id
stringlengths
27
33
source
stringclasses
1 value
format
stringclasses
1 value
text
stringlengths
13
1.81M
warning/0001/hep-ph0001308.html
ar5iv
text
# 1 Introduction ## 1 Introduction Low-energy pion and pion–nucleon physics is described in QCD via chiral perturbation theory (CHPT) . This is an effective field theory designed to solve the Ward identities of the chiral symmetry of QCD order by order. CHPT is based on an effective Lagrangian constructed in a systematic way in terms of the asymptotically observed hadronic fields and consistent with all symmetries of QCD. The effective Lagrangian is organized in the form of a chiral expansion, i.e. an expansion in powers of momenta and light quark masses . At every order, the effective Lagrangian contains free parameters, the so–called low-energy constants (LECs). Various quantities calculated from the effective Lagrangian are related by the same set of LECs entering the results. In addition, there are some rare cases where, to a certain given order in the chiral expansion, predictions are free of LECs. At lowest order, CHPT just reproduces the well–known results of current algebra. However, within CHPT a systematic improvement of current algebra predictions is possible. Furthermore, to restore unitarity perturbatively, one has to calculate (pion) loop corrections because the tree diagrams contributing at lowest order are always real. Every loop increases the chiral order by two, therefore the lowest order results are the tree level ones. The relation between current algebra and the lowest order tree level effective Lagrangian has already been established in the sixties. Loops enter at higher orders, together with tree contributions with insertions from the higher order Lagrangians. In the meson sector, the chiral expansion thus proceeds in steps of two, if one assumes the standard scenario of chiral symmetry breaking with a large quark–antiquark condensate. In the one–nucleon sector of CHPT the situation is different. Here, due to the fermionic nature of the matter fields, couplings with odd powers in momenta are allowed and the lowest order tree graphs stem from a dimension one chiral Lagrangian. Loops start to enter at third order (in a scheme that respects the power counting, see below). However, for various reasons (convergence, completeness, and so on) calculations are now performed at the complete one–loop level, i.e. at fourth chiral order (for a review and status report, see refs.). The full effective Lagrangian is, however, only known up to third order . The most general fourth order $`\pi `$N chiral Lagrangian is still the missing ingredient in a complete one–loop analysis within pion–nucleon CHPT. It is the purpose of this paper to provide this missing component. We refrain here from discussing the three–flavor case, which certainly deserves more study in the future. Most CHPT calculations for the $`\pi `$N system (coupled to external fields) have been performed in the framework of Heavy Baryon CHPT (HBCHPT), a specific non-relativistic projection of the theory. This scheme was developed in ref. motivated by the methods used in heavy quark effective field theory and the observation that a relativistic formulation with baryons leads to complications in the power counting when standard dimensional regularization is employed . In HBCHPT the troublesome mass scale, the nucleon mass $`m`$, appearing in the fermion propagator, is simply transformed in a string of vertex corrections with fixed coefficients and increasing powers in $`1/m`$. However, as was pointed out recently in ref. and ref., a scheme consistent with the power counting, using the relativistic version of the $`\pi `$N Lagrangian, can also be set up if one performs a different method of regularization. Therefore not only the heavy baryon (HB) projection of the Lagrangian is of interest, but also its fully relativistic version. Note also that if one wants to match the HB approach to the relativistic theory, this is most easily and naturally done starting from the relativistic approach (for a detailed discussion, see e.g. ref.). The divergent part of the fourth order HBCHPT $`\pi `$N Lagrangian is already known . Also, the complete but not minimal fourth order Lagrangian with virtual photons has been worked out in ref.. In addition, recently an attempt to work out the complete fourth–order isospin–symmetric Lagrangian in the absence of external fields has been presented in ref..<sup>#4</sup><sup>#4</sup>#4We will comment in more detail on that paper below. However, the complete fourth order heavy baryon as well as relativistic $`\pi `$N Lagrangian (including external fields and in particular strong isospin breaking) is constructed here for the first time. The number of LECs in the resulting Lagrangian turns out to be larger than 100. This is similar to the case of the two–loop $`𝒪(p^6)`$ meson Lagrangian, see refs.. In case of such a large number of free parameters in the theory, one might question the predictive power of the whole approach. However, most physical processes are only sensitive to a small number of LECs. As an example, we mention elastic–pion nucleon scattering. At second, third and fourth order, one has four, four and five independent (combinations of) LECs, respectively, and these can be easily determined by a fit to the vast amount of low–energy pion–nucleon scattering data (this argument is spelled out in more detail in ref.). In photo–nucleon processes, one usually has even fewer LECs, e.g. the HBCHPT expression of the electric dipole amplitude $`E_{0+}`$ for neutral pion scattering off nucleons involves only two combinations of LECs up–to–and–including fourth order. The paper is organized as follows. In Sect. 2, the construction of the Lagrangian is described step by step (choice of fields, construction of invariant monomials, elimination of linearly dependent terms, heavy baryon projection). The result is presented and discussed in Sect. 3. Appendix A contains a more detailed discussion of the elimination of the so-called equation-of-motion terms in the relativistic Lagrangian. Appendix B contains some relevant formulae for the heavy baryon projection. ## 2 Construction of the Lagrangian ### 2.1 Choice of fields Let us briefly collect the basic ingredients for the construction of the effective $`\pi `$N Lagrangian. The underlying Lagrangian is that of QCD with massless $`u`$ and $`d`$ quarks, coupled to external hermitian $`2\times 2`$ matrix–valued fields $`v_\mu ,`$ $`a_\mu ,`$ $`s`$ and $`p`$ (vector, axial–vector, scalar and pseudoscalar, respectively) $$=_{\text{QCD}}^0+\overline{q}\gamma ^\mu \left(v_\mu +\gamma _5a_\mu \right)q\overline{q}(si\gamma _5p)q,q=\left(\genfrac{}{}{0pt}{}{u}{d}\right).$$ (2.1) With suitably transforming external fields, this Lagrangian is locally $`SU(2)_L\times SU(2)_R\times U(1)_V`$ invariant. The chiral $`SU(2)_L\times SU(2)_R`$ symmetry of QCD is spontaneously broken down to its vectorial subgroup, $`SU(2)_V`$. To simplify life further, we disregard the isoscalar axial currents as well as the winding number density.<sup>#5</sup><sup>#5</sup>#5Note that isoscalar axial currents only play a role in the discussion of the so–called spin content of the nucleon. That topic can only be addressed properly in a three flavor scheme. For a lucid discussion of the $`\theta `$ term, we refer to ref.. Explicit chiral symmetry breaking, i.e. the nonvanishing $`u`$ and $`d`$ current quark mass, is taken into account by setting $`s==diag(m_u,m_d).`$ On the level of the effective field theory, the spontaneously broken chiral symmetry is non-linearly realized in terms of the pion and the nucleon fields . The pion fields $`\mathrm{\Phi },`$ being coordinates of the chiral coset space, are naturally represented by elements $`u(\mathrm{\Phi })`$ of this coset space. The most convenient choice of fields for the construction of the effective Lagrangian is given by $`u_\mu `$ $`=`$ $`i\{u^{}(_\mu ir_\mu )uu(_\mu i\mathrm{}_\mu )u^{}\},`$ $`\chi _\pm `$ $`=`$ $`u^{}\chi u^{}\pm u\chi ^{}u,`$ $`F_{\mu \nu }^\pm `$ $`=`$ $`u^{}F_{\mu \nu }^Ru\pm uF_{\mu \nu }^Lu^{},`$ (2.2) where $`\chi `$ $`=`$ $`2B(s+ip),`$ $`F_R^{\mu \nu }`$ $`=`$ $`^\mu r^\nu ^\nu r^\mu i[r^\mu ,r^\nu ],r_\mu =v_\mu +a_\mu ,`$ $`F_L^{\mu \nu }`$ $`=`$ $`^\mu \mathrm{}^\nu ^\nu \mathrm{}^\mu i[\mathrm{}^\mu ,\mathrm{}^\nu ],\mathrm{}_\mu =v_\mu a_\mu ,`$ (2.3) and $`B`$ is the parameter of the meson Lagrangian of $`O(p^2)`$ related to the strength of the quark–antiquark condensate . We work here in the standard framework, $`BF_\pi `$, with $`F_\pi `$ the pion decay constant. For a discussion of the generalized scenario in the presence of matter fields, see e.g. ref.. The reason why the fields in eq.(2.2) are so convenient is that they all transform in the same way under chiral transformations, namely as $$X\stackrel{𝑔}{}h(g,\mathrm{\Phi })Xh^1(g,\mathrm{\Phi }),$$ (2.4) where $`g`$ is an element of $`SU(2)_L\times SU(2)_R`$ and the so-called compensator $`h(g,\mathrm{\Phi })`$ defines a non–linear realization of the chiral symmetry. The compensator $`h(g,\mathrm{\Phi })`$ depends on $`g`$ and $`\mathrm{\Phi }`$ in a complicated way, but since the nucleon field $`\mathrm{\Psi }`$ transforms as $$\mathrm{\Psi }\stackrel{𝑔}{}h(g,\mathrm{\Phi })\mathrm{\Psi },\overline{\mathrm{\Psi }}\stackrel{𝑔}{}\overline{\mathrm{\Psi }}h^1(g,\mathrm{\Phi })$$ (2.5) one can easily construct invariants of the form $`\overline{\mathrm{\Psi }}O\mathrm{\Psi }`$ without explicit knowledge of the compensator. Moreover, for the covariant derivative defined by $$D_\mu =_\mu +\mathrm{\Gamma }_\mu ,\mathrm{\Gamma }_\mu =\frac{1}{2}\{u^{}(_\mu ir_\mu )u+u(_\mu i\mathrm{}_\mu )u^{}\},$$ (2.6) it follows that also $`[D_\mu ,X],`$ $`[D_\mu ,[D_\nu ,X]],`$ etc. transform according to eq.(2.4) and $`D_\mu \mathrm{\Psi },`$ $`D_\mu D_\nu \mathrm{\Psi },`$ etc. according to eq.(2.5). This allows for a simple construction of invariants containing these derivatives. The definition of the covariant derivative eq.(2.6) implies two important relations. The first one is the so-called curvature relation $$[D_\mu ,D_\nu ]=\frac{1}{4}[u_\mu ,u_\nu ]\frac{i}{2}F_{\mu \nu }^+,$$ (2.7) which allows one to consider products of covariant derivatives only in the completely symmetrized form (and ignore all the other possibilities). The second one is (be aware that some authors like e.g. in refs., use a different convention which leads to a minus sign in the definition of $`F_{\mu \nu }^{}`$ as compared to the one used here) $$[D_\mu ,u_\nu ][D_\nu ,u_\mu ]=F_{\mu \nu }^{},$$ (2.8) which in a similar way allows one to consider covariant derivatives of $`u_\mu `$ only in the explicitly symmetrized form in terms of the tensor $`h_{\mu \nu }`$, $$h_{\mu \nu }=[D_\mu ,u_\nu ]+[D_\nu ,u_\mu ].$$ (2.9) For the construction of terms, as well as for further phenomenological applications, it is advantageous to treat isosinglet and isotriplet components of the external fields separately. We therefore define $$\stackrel{~}{X}=X\frac{1}{2}X,$$ (2.10) where $`\mathrm{}`$ stands for the flavor trace, and we work with the following set of fields: $`u_\mu ,`$ $`\stackrel{~}{\chi }_\pm ,`$ $`\chi _\pm ,`$ $`\stackrel{~}{F}_\pm ,`$ $`F_+`$ (the trace $`F_{}`$ is zero because we have omitted the isoscalar axial current.) To construct a hermitian effective Lagrangian, which is not only chiral, but also parity (P) and charge conjugation (C) invariant, one needs to know the transformation properties of the fields under space inversion, charge conjugation and hermitian conjugation. For the fields under consideration (and their covariant derivatives), hermitian conjugation is equal to $`\pm `$itself and charge conjugation amounts to $`\pm `$transposed, where the signs are given (together with the parity) in Table 1. This table also contains the chiral dimensions of the fields. This is necessary for a systematic construction of the chiral effective Lagrangian, order by order. Covariant derivatives acting on pion or external fields count as quantities of first chiral order. In what follows, an analogous information for Clifford algebra elements, the metric $`g_{\mu \nu }`$ and the totally antisymmetric (Levi-Civita) tensor $`\epsilon _{\lambda \mu \nu \rho }`$ (for $`d=4)`$ together with the covariant derivative acting on the nucleon fields will be needed (for a more detailed discussion, see ref.). In this case, hermitian conjugate equals $`\pm \gamma ^0`$(itself)$`\gamma ^0`$ and charge conjugation amounts to $`\pm `$transposed, where the signs are given (together with parity and chiral dimension) in Table 2. The covariant derivative acting on nucleon fields counts as a quantity of zeroth chiral order, since the time component of the derivative gives the nucleon energy, which cannot be considered a small quantity. However, the combination $`\left(i\overline{)𝐷}m\right)\mathrm{\Psi }`$ is of first chiral order. The minus sign for the charge and hermitian conjugation of $`D_\mu \mathrm{\Psi }`$ as well as the chiral dimension of $`\gamma _5`$ in Table 2 is formal and requires some comment, which will be given below. ### 2.2 Invariant monomials We are now in the position to combine these building blocks to form invariant monomials.<sup>#6</sup><sup>#6</sup>#6Systematic methods to construct these invariant monomials can also be found in refs.. Any invariant monomial in the effective $`\pi `$N Lagrangian is of the generic form $$\overline{\mathrm{\Psi }}A^{\mu \nu \mathrm{}}\mathrm{\Theta }_{\mu \nu \mathrm{}}\mathrm{\Psi }+\mathrm{h}.\mathrm{c}..$$ (2.11) Here, the quantity $`A^{\mu \nu \mathrm{}}`$ is a product of pion and/or external fields and their covariant derivatives. $`\mathrm{\Theta }_{\mu \nu \mathrm{}}`$, on the other hand, is a product of a Clifford algebra element $`\mathrm{\Gamma }_{\mu \nu \mathrm{}}`$ and a totally symmetrized product of $`n`$ covariant derivatives acting on nucleon fields, $`D_{\alpha \beta \mathrm{}\omega }^n=\{D_\alpha ,\{D_\beta ,\{\mathrm{},D_\omega \}\}\}`$, $$\mathrm{\Theta }_{\mu \nu \mathrm{}\alpha \beta \mathrm{}}=\mathrm{\Gamma }_{\mu \nu \mathrm{}}D_{\alpha \beta \mathrm{}}^n.$$ (2.12) The Clifford algebra elements are understood to be expanded in the standard basis ($`1`$, $`\gamma _5`$, $`\gamma _\mu `$, $`\gamma _\mu \gamma _5`$, $`\sigma _{\mu \nu }`$) and all the metric and Levi-Civita tensors are included in $`\mathrm{\Gamma }_{\mu \nu \mathrm{}}`$. Note that Levi-Civita tensors may have some upper indices, which are, however, contracted with other indices within $`\mathrm{\Theta }_{\mu \nu \mathrm{}}`$. The structure given in eq.(2.11) is not the most general Lorentz invariant structure, it already obeys some of the restrictions dictated by chiral symmetry. The curvature relation eq.(2.7) manifests itself in the fact that we only consider symmetrized products of covariant derivatives acting on $`\mathrm{\Psi }`$. Another feature of eq.(2.11) is that, except for the $`\epsilon `$-tensors, no two indices of $`\mathrm{\Theta }_{\mu \nu \mathrm{}}`$ are contracted with each other. The reason for this lies in the fact that at a given chiral order, $`\overline{)𝐷}\mathrm{\Psi }`$ can always be replaced by $`im\mathrm{\Psi }`$, since their difference $`\left(\overline{)𝐷}+im\right)\mathrm{\Psi }`$ is of higher order. One can therefore ignore $`D_\mu \mathrm{\Psi }`$ contracted with $`\gamma ^\mu ,`$ $`\gamma ^5\gamma ^\mu `$ and also with $`\sigma ^{\lambda \mu }=i\gamma ^\lambda \gamma ^\mu ig^{\lambda \mu }`$. The last relation explains also why $`g^{\lambda \mu }\{D_\lambda ,D_\mu \}\mathrm{\Psi }`$ can be ignored. Other important restrictions on the structure of $`\mathrm{\Theta }_{\mu \nu \mathrm{}}`$ are discussed in appendix A. To get the complete list of terms contributing to $`A^{\mu \nu \mathrm{}}`$, one writes down all possible products of the pionic and external fields and covariant derivatives thereof. Every index of $`A^{\mu \nu \mathrm{}}`$ increases the chiral order by one, therefore the overall number of indices of $`A^{\mu \nu \mathrm{}}`$ (as well as lower indices of $`\mathrm{\Theta }_{\mu \nu \mathrm{}}`$) is constrained by the chiral order under consideration. Since the matrix fields do not commute, one has to take all the possible orderings. To get $`A^{\mu \nu \mathrm{}}`$ with simple transformation properties under charge and hermitian conjugation, all the products are rewritten in terms of commutators and anticommutators. In the SU(2) case, this is equivalent to decomposing each product of any two terms into isoscalar and isovector parts, since the only nontrivial product is that of two traceless matrices, and the isoscalar (isovector) part of this product is given by an anticommutator (commutator) of the matrices. After such a rearrangement of products has been performed, the following relations hold: $$A^{}=(1)^{h_A}A,A^c=(1)^{c_A}A^T,$$ (2.13) where $`(1)^{h_A}`$ and $`(1)^{c_A}`$ are determined by the signs from Table 1 (as products of factors $`\pm 1`$ for every field and covariant derivative, and an extra factor $`1`$ for every commutator). For the Clifford algebra elements one has similar relations, $$\mathrm{\Gamma }^{}=(1)^{h_\mathrm{\Gamma }}\gamma ^0\mathrm{\Gamma }\gamma ^0,\mathrm{\Gamma }^c=(1)^{c_\mathrm{\Gamma }}\mathrm{\Gamma }^T,$$ (2.14) where $`h_\mathrm{\Gamma }`$ and $`c_\mathrm{\Gamma }`$ are determined by the signs from Table 2. The monomial eq.(2.11) can now be written as $$\overline{\mathrm{\Psi }}A^{\mu \nu \mathrm{}\alpha \beta \mathrm{}}\mathrm{\Gamma }_{\mu \nu \mathrm{}}D_{\alpha \beta \mathrm{}}^n\mathrm{\Psi }+(1)^{h_A+h_\mathrm{\Gamma }}\overline{\mathrm{\Psi }}\stackrel{}{D}_{\alpha \beta \mathrm{}}^n\mathrm{\Gamma }_{\mu \nu \mathrm{}}A^{\mu \nu \mathrm{}\alpha \beta \mathrm{}}\mathrm{\Psi }.$$ (2.15) After elimination of total derivatives ($`\stackrel{}{D}^n(1)^nD^n`$) and subsequent use of the Leibniz rule, one obtains (modulo higher order terms with derivatives acting on $`A^{\mu \nu \mathrm{}}`$) $$\overline{\mathrm{\Psi }}A^{\mu \nu \mathrm{}\alpha \beta \mathrm{}}\mathrm{\Gamma }_{\mu \nu \mathrm{}}D_{\alpha \beta \mathrm{}}^n\mathrm{\Psi }+(1)^{h_A+h_\mathrm{\Gamma }+n}\overline{\mathrm{\Psi }}A^{\mu \nu \mathrm{}\alpha \beta \mathrm{}}\mathrm{\Gamma }_{\mu \nu \mathrm{}}D_{\alpha \beta \mathrm{}}^n\mathrm{\Psi }.$$ (2.16) We see that to a given chiral order, the second term in eq.(2.15) either doubles or cancels the first one. If the later is true, i.e. if $`h_A+h_\mathrm{\Gamma }+n`$ is odd, this term only contributes at higher orders and can thus be ignored. Consider now charge conjugation acting on the remaining terms of the type given in eq.(2.16) (again modulo higher order terms, with derivatives acting on $`A^{\mu \nu \mathrm{}}`$) $$2\overline{\mathrm{\Psi }}A^{\mu \nu \mathrm{}\alpha \beta \mathrm{}}\mathrm{\Gamma }_{\mu \nu \mathrm{}}D_{\alpha \beta \mathrm{}}^n\mathrm{\Psi }+(1)^{c_A+c_\mathrm{\Gamma }+n}2\overline{\mathrm{\Psi }}A^{\mu \nu \mathrm{}\alpha \beta \mathrm{}}\mathrm{\Gamma }_{\mu \nu \mathrm{}}D_{\alpha \beta \mathrm{}}^n\mathrm{\Psi }.$$ (2.17) By the same reasoning as before, the terms with odd $`c_A+c_\mathrm{\Gamma }+n`$ are to be discarded. The formal minus sign for the charge and hermitian conjugation of $`D_\mu \mathrm{\Psi }`$ in Table 2 takes care of this in a simple way. With this convention, for any $`\mathrm{\Theta }_{\mu \nu \mathrm{}}`$, the two numbers $`h_\mathrm{\Theta }`$ and $`c_\mathrm{\Theta }`$ are determined by the entries in Table 2 ($`h_\mathrm{\Theta }=`$ $`h_\mathrm{\Gamma }+n`$, $`c_\mathrm{\Theta }=`$ $`c_\mathrm{\Gamma }+n`$) and invariant monomials eq.(2.11) of a given chiral order are obtained if and only if $$(1)^{h_A+h_\mathrm{\Theta }}=1,(1)^{c_A+c_\mathrm{\Theta }}=1.$$ (2.18) ### 2.3 Elimination of terms The list of invariant monomials generated by the complete lists of $`A^{\mu \nu \mathrm{}}`$ and $`\mathrm{\Theta }_{\mu \nu \mathrm{}}`$ together with the condition (2.18) is still overcomplete. It contains linearly dependent terms, which can be reduced to a minimal set by use of various identities. In this paragraph, we will discuss the different types of such identities which allow to eliminate the linearly dependent terms. First of all, there are general identities, like the cyclic property of the trace, implying $$a[b,c]=c[a,b],$$ (2.19) or Schouten’s identity $$\epsilon ^{\lambda \mu \nu \rho }a^\tau +\epsilon ^{\mu \nu \rho \tau }a^\lambda +\epsilon ^{\nu \rho \tau \lambda }a^\mu +\epsilon ^{\rho \tau \lambda \mu }a^\nu +\epsilon ^{\tau \lambda \mu \nu }a^\rho =0.$$ (2.20) Another general identity, frequently used in the construction of chiral Lagrangians, is provided by the Cayley-Hamilton theorem. For 2$`\times `$2 matrices $`a`$ and $`b`$, this theorem just implies $$\{a,b\}=ab+ab+abab.$$ (2.21) This was already accounted for by the separate treatment of the traces and the traceless matrices. However, there are other nontrivial identities among products of traceless matrices. Let us explicitly mention one such identity, which turns out to reduce the number of independent terms containing four $`u`$ fields. It reads $$\{\stackrel{~}{a},\stackrel{~}{b}\}[\stackrel{~}{a},\stackrel{~}{c}]\{\stackrel{~}{a},\stackrel{~}{c}\}[\stackrel{~}{a},\stackrel{~}{b}]=\{\stackrel{~}{a},\stackrel{~}{a}\}[\stackrel{~}{b},\stackrel{~}{c}]\stackrel{~}{a}\{\stackrel{~}{a},[\stackrel{~}{b},\stackrel{~}{c}]\},$$ (2.22) where $`\stackrel{~}{a},\stackrel{~}{b},\stackrel{~}{c}`$ are traceless 2$`\times `$2 matrices. Another set of identities is provided by the curvature relation eq.(2.7). In connection with the Bianchi identity for covariant derivatives, $$[D_\lambda ,[D_\mu ,D_\nu ]]+\mathrm{cyclic}=0,$$ (2.23) where “$`\mathrm{cyclic}`$” stands for cyclic permutations, it entails $$[D_\lambda ,F_{\mu \nu }^+]+\mathrm{cyclic}=\frac{i}{2}[u_\lambda ,F_{\mu \nu }^{}]+\mathrm{cyclic},$$ (2.24) where we have used the Leibniz rule and eq.(2.8) on the right–hand–side. On the other hand, when combined with eq.(2.8), the curvature relation gives $$[D_\lambda ,F_{\mu \nu }^{}]+\mathrm{cyclic}=\frac{i}{2}[u_\lambda ,F_{\mu \nu }^+]+\mathrm{cyclic},$$ (2.25) where we have used the Jacobi identity $`[[u_\lambda ,u_\mu ],u_\nu ]+\mathrm{cyclic}=0`$ on the right–hand–side. These relations can be used for the elimination of (some) terms which contain $`[D_\lambda ,F_{\mu \nu }^\pm ]`$. Yet another set of identities is based on the equations of motion (EOM) deduced from the lowest order $`\pi \pi `$ $$[D_\mu ,u^\mu ]=\frac{i}{2}\stackrel{~}{\chi }_{},$$ (2.26) and $`\pi `$N Lagrangians (its explicit form is given in paragraph 3.1) $`\left(i\overline{)𝐷}m+{\displaystyle \frac{1}{2}}g_A\overline{)𝑢}\gamma ^5\right)\mathrm{\Psi }`$ $`=`$ $`0,`$ (2.27) $`\overline{\mathrm{\Psi }}\left(i\stackrel{}{\overline{)𝐷}}+m{\displaystyle \frac{1}{2}}g_A\overline{)𝑢}\gamma ^5\right)`$ $`=`$ $`0.`$ (2.28) Strictly speaking, in these equations $`m`$ and $`g_A`$ refer to the nucleon mass and the axial–vector coupling constant in the SU(2) chiral limit ($`m_u=m_d=0`$, $`m_s`$ fixed). We will not further specify this but it should be kept in mind. One can directly use these EOM or (equivalently) perform specific field redefinitions — both techniques yield the same result. The pion EOM is used to get rid of all the terms containing $`h_\mu ^\mu `$, as well as $`[D^\mu ,h_{\mu \nu }]`$, which can be eliminated using eq.(2.26) together with eqs.(2.72.8). The nucleon EOM, the main effect of which is a remarkable restriction of the structure of $`\mathrm{\Theta }_{\mu \nu \mathrm{}}`$, is discussed in full generality in appendix A. Here, we prefer to follow the procedure adopted in ref. and give only the specific relations based on partial integrations and the nucleon EOM used in the reduction from the overcomplete to the minimal set of terms. Some of these relations have already appeared in , but for completeness we prefer to show them all. They read: $`\overline{\mathrm{\Psi }}A^\mu iD_\mu \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.`$ $``$ $`2m\overline{\mathrm{\Psi }}\gamma _\mu A^\mu \mathrm{\Psi },`$ (2.29) $`\overline{\mathrm{\Psi }}A^{\mu \nu }D_\nu D_\mu \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.`$ $``$ $`m(\overline{\mathrm{\Psi }}\gamma _\mu A^{\mu \nu }iD_\nu \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.),`$ (2.30) $`\overline{\mathrm{\Psi }}A^{\mu \nu \lambda }iD_\lambda D_\nu D_\mu \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.`$ $``$ $`m(\overline{\mathrm{\Psi }}\gamma _\mu A^{\mu \nu \lambda }D_\lambda D_\nu \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.),`$ (2.31) $`\overline{\mathrm{\Psi }}\gamma _5\gamma _\lambda A^{\mu \lambda }iD_\mu \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.`$ $``$ $`2im\overline{\mathrm{\Psi }}\gamma _5\sigma _{\mu \lambda }A^{\mu \lambda }\mathrm{\Psi }`$ (2.32) $`+`$ $`(\overline{\mathrm{\Psi }}\gamma _5\gamma _\mu A^{\mu \lambda }iD_\lambda \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.),`$ $`\overline{\mathrm{\Psi }}\gamma _5\gamma _\lambda A^{\mu \lambda \alpha }D_\alpha D_\mu \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.`$ $``$ $`m(\overline{\mathrm{\Psi }}\gamma _5\sigma _{\mu \lambda }A^{\mu \lambda \alpha }D_\alpha \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.)`$ (2.33) $`+`$ $`(\overline{\mathrm{\Psi }}\gamma _5\gamma _\mu A^{\mu \lambda \alpha }D_\alpha D_\lambda \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.),`$ $`\overline{\mathrm{\Psi }}\gamma _5\gamma _\lambda A^{\mu \lambda \alpha \beta }iD_\beta D_\alpha D_\mu \mathrm{\Psi },+\mathrm{h}.\mathrm{c}.`$ $``$ $`m(i\overline{\mathrm{\Psi }}\gamma _5\sigma _{\mu \lambda }A^{\mu \lambda \alpha \beta }D_\beta D_\alpha \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.)`$ (2.34) $`+`$ $`(\overline{\mathrm{\Psi }}\gamma _5\gamma _\mu A^{\mu \lambda \alpha \beta }iD_\beta D_\alpha D_\lambda \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.),`$ $`\overline{\mathrm{\Psi }}\sigma _{\alpha \beta }A^{\alpha \beta \mu }iD_\mu \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.`$ $``$ $`2m\overline{\mathrm{\Psi }}ϵ_{\alpha \beta \mu \nu }\gamma _5\gamma ^\nu A^{\alpha \beta \mu }\mathrm{\Psi }(\overline{\mathrm{\Psi }}\sigma _{\beta \mu }A^{\alpha \beta \mu }iD_\alpha \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.)`$ (2.35) $`+`$ $`(\overline{\mathrm{\Psi }}\sigma _{\alpha \mu }A^{\alpha \beta \mu }iD_\beta \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.),`$ $`\overline{\mathrm{\Psi }}\sigma _{\alpha \beta }A^{\alpha \beta \nu \mu }D_\nu D_\mu \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.`$ $``$ $`m(i\overline{\mathrm{\Psi }}ϵ_{\alpha \beta \mu \lambda }\gamma _5\gamma ^\lambda A^{\alpha \beta \nu \mu }D_\nu \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.)`$ (2.36) $``$ $`(\overline{\mathrm{\Psi }}\sigma _{\beta \mu }A^{\alpha \beta \nu \mu }D_\nu D_\alpha \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.)`$ $`+`$ $`(\overline{\mathrm{\Psi }}\sigma _{\alpha \mu }A^{\alpha \beta \nu \mu }D_\nu D_\beta \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.),`$ $`\overline{\mathrm{\Psi }}\gamma _5\sigma _{\alpha \beta }A^{\alpha \beta \mu }D_\mu \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.`$ $``$ $`(\overline{\mathrm{\Psi }}\gamma _5\sigma _{\beta \mu }A^{\alpha \beta \mu }D_\alpha \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.)`$ (2.37) $`+`$ $`(\overline{\mathrm{\Psi }}\gamma _5\sigma _{\alpha \mu }A^{\alpha \beta \mu }D_\beta \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.),`$ $`i\overline{\mathrm{\Psi }}\gamma _5\sigma _{\alpha \beta }A^{\alpha \beta \nu \mu }D_\nu D_\mu \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.`$ $``$ $`(i\overline{\mathrm{\Psi }}\gamma _5\sigma _{\beta \mu }A^{\alpha \beta \nu \mu }D_\nu D_\alpha \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.)`$ (2.38) $`+`$ $`(i\overline{\mathrm{\Psi }}\gamma _5\sigma _{\alpha \mu }A^{\alpha \beta \nu \mu }D_\nu D_\beta \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.),`$ $`\overline{\mathrm{\Psi }}\gamma _\mu [iD^\mu ,A]\mathrm{\Psi }`$ $``$ $`{\displaystyle \frac{g_A}{2}}\overline{\mathrm{\Psi }}\gamma ^\mu \gamma _5[A,u_\mu ]\mathrm{\Psi },`$ (2.39) $`\overline{\mathrm{\Psi }}\gamma _5\gamma _\mu [iD^\mu ,A]\mathrm{\Psi }`$ $``$ $`2m\overline{\mathrm{\Psi }}\gamma _5A\mathrm{\Psi }{\displaystyle \frac{g_A}{2}}\overline{\mathrm{\Psi }}\gamma ^\mu [A,u_\mu ]\mathrm{\Psi },`$ (2.40) $`\overline{\mathrm{\Psi }}\gamma _5\gamma _\nu [D^\mu ,A^\nu ]iD_\mu \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.`$ $``$ $`0,`$ (2.41) $`\overline{\mathrm{\Psi }}ϵ_{\alpha \beta \mu \lambda }\gamma ^\lambda [D^\nu ,A^{\alpha \beta \mu }]iD_\nu \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.`$ $``$ $`0,`$ (2.42) $`\overline{\mathrm{\Psi }}[D^\mu ,A^\nu ]D_\nu D_\mu \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.`$ $``$ $`0,`$ (2.43) $`\overline{\mathrm{\Psi }}\sigma _{\alpha \beta }[D^\mu ,A^{\alpha \beta \nu }]D_\nu D_\mu \mathrm{\Psi }+\mathrm{h}.\mathrm{c}.`$ $``$ $`0.`$ (2.44) Here, the symbol $``$ means equal up to terms of higher order. ### 2.4 Heavy baryon projection The heavy baryon (HB) projection of the relativistic theory is well documented in the literature, see e.g. refs.. Here, we only collect some basic definitions and formulae which are needed in the following. For a more detailed exposition, we refer to the review . The procedure we adopt follows closely the one in ref., therefore we only give some steps for completeness. Basically, one considers the mass of the nucleon large compared to the typical external momenta transferred by pions or external probes and writes the nucleon four–momentum as $`p_\mu =mv_\mu +\mathrm{}_\mu `$, $`p^2=m^2`$, subject to the condition that $`v\mathrm{}m`$. Here, $`v_\mu `$ is the nucleon four–velocity (in the rest–frame, we have $`v_\mu =(1,\stackrel{}{0})`$). Note again that strictly speaking, the nucleon mass appearing here should be taken at its value in the chiral limit. Consequently, one can decompose the wavefunction $`\mathrm{\Psi }(x)`$ into velocity eigenstates $$\mathrm{\Psi }(x)=\mathrm{exp}[imvx][N_v(x)+h_v(x)]$$ (2.45) with $$\overline{)𝑣}N_v=N_v,\overline{)𝑣}h_v=h_v,$$ (2.46) or in terms of velocity projection operators $`P_v^\pm `$ $$P_v^+N_v=N_v,P_v^{}h_v=h_v,P_v^\pm =\frac{1}{2}(1\pm \overline{)𝑣}),P_v^++P_v^{}=1.$$ (2.47) One now eliminates the small component $`h_v(x)`$ either by using the equations of motion or path–integral methods. The Dirac equation for the velocity–dependent nucleon field $`N_v`$ takes the form $`ivN_v=0`$ to lowest order in $`1/m`$. Apart from the four–velocity, the covariant spin–operator $`S_\mu `$ is of crucial importance in the HB projection. Its definition and basic properties are: $$S_\mu =\frac{i}{2}\gamma _5\sigma _{\mu \nu }v^\nu ,Sv=0,\{S_\mu ,S_\nu \}=\frac{1}{2}\left(v_\mu v_\nu g_{\mu \nu }\right),[S_\mu ,S_\nu ]=iϵ_{\mu \nu \alpha \beta }v^\alpha S^\beta ,$$ (2.48) in the convention $`ϵ^{0123}=1`$. After this projection, the Lagrangian does not contain the nucleon mass term any more and also, all Dirac matrices can be expressed as combinations of $`v_\mu `$ and $`S_\mu `$. Therefore, one can translate all terms of the relativistic Lagrangian into their heavy fermion counterpieces. In addition, there are $`1/m`$ corrections to the various operators. These can be worked out along the lines spelled out in appendix A of ref.. The $`\pi `$N Lagrangian to fourth order for the “light” components $`N_v`$ then becomes: $`_{\pi N}`$ $`=`$ $`\overline{N}_v\{𝒜^{(1)}+𝒜^{(2)}+\gamma _0^{(1)}\gamma _0𝒞^{1(0)}^{(1)}`$ $`+`$ $`𝒜^{(3)}+\gamma _0^{(1)}\gamma _0𝒞^{1(1)}^{(1)}+\gamma _0^{(2)}\gamma _0𝒞^{1(0)}^{(1)}+\gamma _0^{(1)}\gamma _0𝒞^{1(0)}^{(2)}`$ $`+`$ $`𝒜^{(4)}+\gamma _0^{(1)}\gamma _0𝒞^{1(2)}^{(1)}+\gamma _0^{(2)}\gamma _0𝒞^{1(1)}^{(1)}+\gamma _0^{(1)}\gamma _0𝒞^{1(1)}^{(2)}`$ $`+\gamma _0^{(2)}\gamma _0𝒞^{1(0)}^{(2)}+\gamma _0^{(3)}\gamma _0𝒞^{1(0)}^{(1)}+\gamma _0^{(1)}\gamma _0𝒞^{1(0)}^{(3)}\}N_v,`$ in the standard notation. Explicit expressions for the operators $`𝒜^{(i)}`$, $`^{(i)}`$, and $`𝒞^{1(i)}`$ are collected in appendix B. ### 2.5 Checks To arrive at the final form of the Lagrangian, which contains more than 100 independent terms, one has to perform rather lengthy algebraic manipulations. Due to the large number of terms, these calculations are prone to errors. To eliminate the possibility of making simple algebraic errors, we went through the whole calculation in two independent ways. We really calculated everything by hand, but simultaneously we have written codes in *Mathematica* for the construction of the relativistic Lagrangian as well as for the extraction of its HB projection. The *Mathematica* programs were not used just to check the hand-made algebraic manipulations step by step, we have rather used different approaches in the codes wherever possible. The final coherence of the results obtained in these two different ways is a very nontrivial cross–check. ## 3 The Lagrangian The resulting effective Lagrangian is given by a string of terms with increasing chiral dimension, $$_{\pi N}=_{\pi N}^{(1)}+_{\pi N}^{(2)}+_{\pi N}^{(3)}+_{\pi N}^{(4)}+\mathrm{},$$ (3.1) where the ellipsis denotes terms of chiral dimension five (or higher). We will first briefly summarize the first three terms in this expansion and then display the novel complete and minimal fourth order Lagrangian. We always give the Lagrangian in the relativistic as well as in the heavy baryon formulation. ### 3.1 Dimension one, two and three At lowest order, the effective $`\pi `$N Lagrangian is given in terms of two parameters, the nucleon mass and the axial–vector coupling constant (in the chiral limit). In its relativistic form, it reads $$_{\pi N}^{(1)}=\overline{\mathrm{\Psi }}\left(i\overline{)𝐷}m+\frac{g_A}{2}\overline{)𝑢}\gamma _5\right)\mathrm{\Psi }.$$ (3.2) In the HB formulation, the mass term is absent and all Clifford algebra elements can be expressed in terms of the nucleon four–velocity and the spin–vector, $$\widehat{}_{\pi N}^{(1)}=\overline{N}_v\left(ivD+g_ASu\right)N_v.$$ (3.3) At second order, seven independent terms with LECs appear, so that the relativistic Lagrangian reads (the explicit form of the various operators $`O_i^{(2)}`$ is given in Table 3) $$_{\pi N}^{(2)}=\underset{i=1}{\overset{7}{}}c_i\overline{\mathrm{\Psi }}O_i^{(2)}\mathrm{\Psi }.$$ (3.4) The LECs $`c_i`$ are, of course, finite. The HB projection is straightforward. We work here with the standard form (see e.g. refs.) and do not transform away the $`(vD)^2`$ term from the kinetic energy (as it was done e.g. in ref.), $$\widehat{}_{\pi N}^{(2)}=\frac{1}{2m}\overline{N}_v\left(\left(vD\right)^2D^2ig_A\{SD,vu\}\right)N_v+\underset{i=1}{\overset{7}{}}\widehat{c}_i\overline{N}_v\widehat{O}_i^{(2)}N_v.$$ (3.5) Again, the monomials $`\widehat{O}_i^{(2)}`$ are listed in Table 3 together with the $`1/m`$ corrections, which some of these operators receive (this splitting is done mostly to have an easier handle on estimating LECs via resonance saturation ). At third order, one has 23 independent terms. We follow the notation of ref. (see also ), $$_{\pi N}^{(3)}=\underset{i=1}{\overset{23}{}}d_i\overline{\mathrm{\Psi }}O_i^{(3)}\mathrm{\Psi }.$$ (3.6) The LECs $`d_i`$ decompose into a renormalized scale–dependent and an infinite (also scale–dependent) part in the standard manner, $`d_i=d_i^r(\lambda )+\kappa _i/F^2L(\lambda )`$, with $`\lambda `$ the scale of dimensional regularization, $`L(\lambda )`$ as defined in ref., and the $`\beta `$–functions $`\kappa _i`$, first worked out by Ecker . The HB projection including the $`1/m`$ corrections reads $$\widehat{}_{\pi N}^{(3)}=\underset{i=1}{\overset{23}{}}\widehat{d}_i\overline{N}_v\widehat{O}_i^{(3)}N_v+\overline{N}_v\widehat{O}_{\mathrm{fixed}}^{(3)}N_v+\overline{N}_v\widehat{O}_{\mathrm{div}}^{(3)}N_v,$$ (3.7) with the corresponding operators and $`1/m`$ corrections collected in Table 4 and $`𝒪_{\mathrm{fixed}}^{(3)}`$ given by $`\widehat{O}_{\mathrm{fixed}}^{(3)}`$ $`=`$ $`{\displaystyle \frac{g_A}{8m^2}}[D^\mu ,[D_\mu ,Su]]{\displaystyle \frac{g_A^2}{32m^2}}ϵ^{\mu \nu \alpha \beta }v_\alpha S_\beta F_{\mu \nu }^{}vui{\displaystyle \frac{1}{4m^2}}\left(vD\right)^3`$ (3.8) $``$ $`{\displaystyle \frac{g_A}{4m^2}}v\stackrel{}{D}SuvD+{\displaystyle \frac{1}{8m^2}}(iD^2vD+\text{h.c.})`$ $``$ $`{\displaystyle \frac{g_A}{4m^2}}\left(\{SD,vu\}vD+\text{h.c.}\right)+{\displaystyle \frac{3g_A^2}{64m^2}}\left(i(vu)^2vD+\text{h. c.}\right)`$ $`+`$ $`{\displaystyle \frac{1}{32m^2}}\left(ϵ^{\mu \nu \alpha \beta }v_\alpha S_\beta [u_\mu ,u_\nu ]vD+\text{h.c.}\right){\displaystyle \frac{1}{16m^2}}\left(iϵ^{\mu \nu \alpha \beta }v_\alpha S_\beta \stackrel{~}{F}_{\mu \nu }^+vD+\text{h.c.}\right)`$ $``$ $`{\displaystyle \frac{1}{32m^2}}\left(iϵ^{\mu \nu \alpha \beta }v_\alpha S_\beta F_{\mu \nu }^+vD+\text{h.c.}\right)`$ $``$ $`{\displaystyle \frac{g_A}{8m^2}}(SuD^2+\text{h.c.}){\displaystyle \frac{g_A}{4m^2}}(S\stackrel{}{D}uD+\text{h. c.})`$ $``$ $`{\displaystyle \frac{1+2c_6}{8m^2}}\left(iϵ^{\mu \nu \alpha \beta }v_\alpha S_\beta \stackrel{~}{F}_{\mu \sigma }^+v^\sigma D_\nu +\text{h.c.}\right){\displaystyle \frac{1+2c_6+4c_7}{16m^2}}\left(iϵ^{\mu \nu \alpha \beta }v_\alpha S_\beta F_{\mu \sigma }^+v^\sigma D_\nu +\text{h.c.}\right)`$ $`+`$ $`{\displaystyle \frac{1+g_A^2+8mc_4}{16m^2}}\left(ϵ^{\mu \nu \alpha \beta }v_\alpha S_\beta [u_\mu ,vu]D_\nu +\text{h.c.}\right)`$ $`+`$ $`{\displaystyle \frac{g_A}{32m^2}}\left(iϵ^{\mu \nu \alpha \beta }v_\alpha F_{\mu \nu }^{}D_\beta +\text{h.c.}\right){\displaystyle \frac{g_A^2}{16m^2}}\left(ivuuD+\text{h.c.}\right)`$ $`+`$ $`i{\displaystyle \frac{1+8mc_4}{32m^2}}[vu,[D^\mu ,u_\mu ]]+{\displaystyle \frac{c_2}{2m}}\left(ivuu_\mu D^\mu +\text{h.c.}\right).`$ In addition, there are eight terms just needed for the renormalization (we give these here for completeness). Their explicit form is: $`\widehat{O}_{\mathrm{div}}^{(3)}`$ $`=`$ $`\stackrel{~}{d}_{24}(\lambda )i(vD)^3+\stackrel{~}{d}_{25}(\lambda )v\stackrel{}{D}SuvD+\stackrel{~}{d}_{26}(\lambda )(iuuvD+\text{h.c.})`$ (3.9) $`+`$ $`\stackrel{~}{d}_{27}(\lambda )\left(i(vu)^2vD+\text{h.c.}\right)+\stackrel{~}{d}_{28}(\lambda )\left(i\chi _+vD+\text{h.c.}\right)`$ $`+`$ $`\stackrel{~}{d}_{29}(\lambda )\left(S^\mu [vD,u_\mu ]vD+\text{h.c.}\right)+\stackrel{~}{d}_{30}(\lambda )\left(ϵ^{\mu \nu \alpha \beta }v_\alpha S_\beta [u_\mu ,u_\nu ]vD+\text{h.c.}\right)`$ $`+`$ $`\stackrel{~}{d}_{31}(\lambda )\left(ϵ^{\mu \nu \alpha \beta }v_\alpha S_\beta \stackrel{~}{F}_{\mu \nu }^+vD+\text{h.c.}\right).`$ The LECs $`\stackrel{~}{d}_i`$ have no finite piece . ### 3.2 Dimension four This paragraph constitutes the main result of our work. We have found that there are 118 independent dimension four operators, which are in principle measurable.<sup>#7</sup><sup>#7</sup>#7We note that the mastertable in ref. contains more terms, 199 to be precise. In that work, however, no attempt was made to find the minimal basis. This is similar to what was done for the dimension three Lagrangian, compare e.g. refs. and . The last four ($`i=115,\mathrm{},118)`$, however, are special in the sense that they are pure contact interactions of the nucleons with the external sources that have no pion matrix elements. For example, the operators 115 and 116 contribute to the scalar nucleon form factor but not to pion–nucleon scattering. The relativistic dimension four Lagrangian takes the form $$_{\pi N}^{(4)}=\underset{i=1}{\overset{118}{}}e_i\overline{\mathrm{\Psi }}O_i^{(4)}\mathrm{\Psi },$$ (3.10) and the monomials $`O_i^{(4)}`$ are tabulated in Table LABEL:dim4. Also given in that table are the processes and observables with the least number of particles (including nucleons, pions, photons and W bosons) to which each operator can contribute. As pointed out in the introduction, many of the operators contribute only to very exotic processes, like three or four pion production induced by photons or pions. Moreover, for a given process, it often happens that some of these operators appear in certain linear combinations. Note also that operators with a separate $`\chi _+`$ simply amount to a quark mass renormalization of the corresponding dimension two operator. The HB projection leads to a much more complicated Lagrangian, $`\widehat{}_{\pi N}^{(4)}`$ $`=`$ $`\overline{N}_v({\displaystyle \underset{i=1}{\overset{118}{}}}\widehat{e}_i\widehat{O}_i^{(4)}+{\displaystyle \underset{i=1}{\overset{23}{}}}e_i^{}W_i+{\displaystyle \underset{i=1}{\overset{67}{}}}(e_i^{\prime \prime }X_i^\lambda D_\lambda +\mathrm{h}.\mathrm{c}.)`$ (3.11) $`+{\displaystyle \underset{i=1}{\overset{23}{}}}(e_i^{\prime \prime \prime }D_\mu Y_i^{\mu \nu }D_\nu +\mathrm{h}.\mathrm{c}.)+{\displaystyle \underset{i=1}{\overset{4}{}}}(e_i^{iv}Z_i^{\lambda \mu \nu }D_\lambda D_\mu D_\nu +\mathrm{h}.\mathrm{c}.)`$ $`+{\displaystyle \frac{1}{8m^3}}(vDD_\mu D^\mu vD(vD)^4)+\widehat{O}_{\mathrm{div}}^{(4)})N_v,`$ where the first sum contains the 118 low–energy constants in the basis of the heavy nucleon fields. Various additional terms appear: First, there are the leading $`1/m`$ corrections to (most of) the 118 dimension four operators. These contribute to the difference in the LECs $`e_i`$ and $`\widehat{e}_i`$; the $`\widehat{e}_i`$ are tabulated in Table LABEL:eim. Furthermore we have additional terms with fixed coefficients. They can be most compactly represented by counting the number of covariant derivatives acting on the nucleon fields. The corresponding (tensor) structures $`W_i`$, $`X_i^\lambda `$, $`Y_i^{\mu \nu }`$, and $`Z_i^{\lambda \mu \nu }`$ are collected in Tables LABEL:tab:Wi, LABEL:tab:Xi, LABEL:tab:Yi, LABEL:tab:Zi together with the pertinent coefficients $`e_i^{},\mathrm{},e_i^{iv}`$. There are two other fixed coefficient terms which are listed in the last line of eq.(3.11). We briefly comment on the fourth order heavy baryon Lagrangian recently presented in ref.. First, it is entirely based on the HB projection and contains no information concerning the relativistic formulation. Furthermore, strong isospin breaking terms as well as external sources (apart from the quark mass matrix) are entirely omitted. The tables in that work contain considerably more terms than found here, which could be reduced by some of the identities spelled out in this paper. ## 4 Summary In this work, we have constructed the minimal effective chiral pion–nucleon two–flavor Lagrangian at fourth order in the chiral expansion. To arrive at this minimal form, we have studied in detail all strictures arising from the equations of motion, trace relations and other algebraic identities. The so constructed Lagrangian contains 118 in principle measurable terms. These are accompanied by the so–called low–energy constants, which can be obtained by a fit to data, from some models or maybe from lattice gauge theory. Note that four of these operators are special since they have no pion matrix elements. Based on the fact that a consistent power counting can be set up in the relativistic as well as in the heavy baryon formulation of the effective field theory, we have worked out the effective Lagrangian in both schemes. In the heavy baryon case, it is mandatory to calculate the various $`1/m`$ corrections at a given chiral order. Here, these are explicitly spelled out up–to–and–including fourth order. ### Acknowledgment We would like to thank Miro Helbich and Guido Müller for collaboration in the early stages of the work, Thomas Becher for a useful comment, and Peter Stelmachovic for pointing out a typographical error. ## Appendix A EOM eliminations The number of independent terms in the effective $`\pi `$N chiral Lagrangian is considerably reduced by the use of the so-called EOM (equation of motion) eliminations. One can view them simply as replacements of $`\overline{)𝐷}\mathrm{\Psi }`$ by $`im\mathrm{\Psi }`$ (a consequence of the fact that $`\left(\overline{)𝐷}+im\right)\mathrm{\Psi }`$ is of higher order), or as a consequence of the lowest order EOM (defining the classical solution, around which the loop expansion is performed), or as a result of specific field transformations (leaving the S–matrix elements untouched). However, the simple replacements $`\overline{)𝐷}\mathrm{\Psi }im\mathrm{\Psi }`$ do not exhaust the EOM eliminations. The point is that also terms (or combinations of terms) not containing $`\overline{)𝐷}\mathrm{\Psi }`$ explicitly can be brought into a form that includes $`\overline{)𝐷}\mathrm{\Psi }`$. Such terms can then also be eliminated. To systematically obtain all such terms, one has to inspect all the possible terms with $`\mathrm{\Gamma }_{\mathrm{}}\overline{)𝐷}\mathrm{\Psi }`$ and rewrite them as combinations of terms without $`\overline{)𝐷}\mathrm{\Psi }`$. To achieve this, we expand any product $`\mathrm{\Gamma }_{\mathrm{}}\gamma _\lambda `$ in the form (the last $`\gamma `$–matrix originates from $`\overline{)𝐷}`$) $$\mathrm{\Gamma }_{\mathrm{}}\gamma _\lambda =\underset{i}{}\mathrm{\Gamma }_{i,\mathrm{}\lambda }^{}+\underset{j}{}\mathrm{\Gamma }_{j,\mathrm{}\lambda }^{\prime \prime },c_{\mathrm{\Gamma }_j^{\prime \prime }}=c_{\mathrm{\Gamma }_i^{}}+1=c_\mathrm{\Gamma },$$ (A.1) where $`\mathrm{\Gamma }_i^{}`$ and $`\mathrm{\Gamma }_j^{\prime \prime }`$ contain only elements from the standard basis of the Clifford algebra, multiplied by appropriate combinations of $`g`$\- and $`\epsilon `$-tensors. When contracted with $`D_\lambda `$ and inserted into (2.15) instead of $`\mathrm{\Gamma }_{\mathrm{}}`$, this yields (after the replacement $`\overline{)𝐷}\mathrm{\Psi }im\mathrm{\Psi }`$) $`{\displaystyle \underset{i}{}}\overline{\mathrm{\Psi }}A^{\mu \nu \mathrm{}\alpha \beta \mathrm{}}\mathrm{\Gamma }_{i,\mu \nu \mathrm{}}^\lambda D_{\lambda \alpha \beta \mathrm{}}^n\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.`$ $``$ $`im\overline{\mathrm{\Psi }}A^{\mu \nu \mathrm{}\alpha \beta \mathrm{}}\mathrm{\Gamma }_{\mu \nu \mathrm{}}D_{\alpha \beta \mathrm{}}^{n1}\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.,`$ $`{\displaystyle \underset{j}{}}\overline{\mathrm{\Psi }}A^{\mu \nu \mathrm{}\alpha \beta \mathrm{}}\mathrm{\Gamma }_{j,\mu \nu \mathrm{}}^{\prime \prime \lambda }D_{\lambda \alpha \beta \mathrm{}}^n\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.`$ $``$ $`0,`$ (A.2) where the symbol $``$ stands for ”equal up-to higher orders”. For $`\mathrm{\Gamma }=1`$, one obtains the so–called “master relation” $$\overline{\mathrm{\Psi }}A^{\alpha \beta \mathrm{}}\gamma ^\lambda D_{\lambda \alpha \beta \mathrm{}}^n\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.im\overline{\mathrm{\Psi }}A^{\alpha \beta \mathrm{}}D_{\alpha \beta \mathrm{}}^{n1}\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.,$$ (A.3) but letting $`\mathrm{\Gamma }`$ run through the rest of the standard basis of the Clifford algebra, a couple of additional relations is generated $$0im\overline{\mathrm{\Psi }}A^{\alpha \beta \mathrm{}}\gamma _5D_{\alpha \beta \mathrm{}}^{n1}\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.,$$ (A.4) $$\overline{\mathrm{\Psi }}A^{\alpha \beta \mathrm{}}\gamma ^5\gamma ^\lambda D_{\lambda \alpha \beta \mathrm{}}^n\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.0,$$ (A.5) $$\overline{\mathrm{\Psi }}A^{\mu \alpha \beta \mathrm{}}D_{\mu \alpha \beta \mathrm{}}^n\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.im\overline{\mathrm{\Psi }}A^{\mu \alpha \beta \mathrm{}}\gamma _\mu D_{\alpha \beta \mathrm{}}^{n1}\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.,$$ (A.6) $$\overline{\mathrm{\Psi }}A^{\mu \alpha \beta \mathrm{}}\sigma _\mu ^\lambda D_{\lambda \alpha \beta \mathrm{}}^n\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.0,$$ (A.7) $$\overline{\mathrm{\Psi }}A^{\mu \alpha \beta \mathrm{}}\gamma _5\sigma _\mu ^\lambda D_{\lambda \alpha \beta \mathrm{}}^n\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.m\overline{\mathrm{\Psi }}A^{\mu \alpha \beta \mathrm{}}\gamma _5\gamma _\mu D_{\alpha \beta \mathrm{}}^{n1}\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.,$$ (A.8) $$\overline{\mathrm{\Psi }}A^{\mu \alpha \beta \mathrm{}}\gamma _5D_{\mu \alpha \beta \mathrm{}}^n\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.0,$$ (A.9) $$\overline{\mathrm{\Psi }}A^{\mu \nu \alpha \beta \mathrm{}}\epsilon _{\mu \nu }^{\lambda \rho }\gamma _5\gamma _\rho D_{\lambda \alpha \beta \mathrm{}}^n\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.im\overline{\mathrm{\Psi }}A^{\mu \nu \alpha \beta \mathrm{}}\sigma _{\mu \nu }D_{\alpha \beta \mathrm{}}^{n1}\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.,$$ (A.10) $$\overline{\mathrm{\Psi }}A^{\mu \nu \alpha \beta \mathrm{}}(\gamma _\mu D_{\nu \alpha \beta \mathrm{}}^n\gamma _\nu D_{\mu \alpha \beta \mathrm{}}^n)\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.0.$$ (A.11) For the sake of completeness, we have written down both relations (A.2) for every $`\mathrm{\Gamma }`$ $``$ $`(\gamma _5`$, $`\gamma _\mu `$,$`\gamma _5\gamma _\mu `$, $`\sigma _{\mu \nu })`$, in spite of the fact that they are not all independent (e.g. eqs.(A.4) and (A.9) are equivalent, and (A.11) follows from (A.6)). For a general $`\mathrm{\Gamma }`$ one obtains, as a rule, some linear combinations of the above relations. But there are exceptions, namely for $`\epsilon `$–tensors contracted with elements of the standard basis. The first such case is $`\mathrm{\Gamma }=\gamma _5\sigma _{\mu \nu }=(1/2i)\epsilon _{\mu \nu \rho \tau }\sigma ^{\rho \tau }`$, which yields two new relations, $$\overline{\mathrm{\Psi }}A^{\mu \nu \alpha \beta \mathrm{}}(\gamma _5\gamma _\mu D_{\nu \alpha \beta \mathrm{}}^n\gamma _5\gamma _\nu D_{\mu \alpha \beta \mathrm{}}^n)\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.m\overline{\mathrm{\Psi }}A^{\mu \nu \alpha \beta \mathrm{}}\gamma _5\sigma _{\mu \nu }D_{\alpha \beta \mathrm{}}^{n1}\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.,$$ (A.12) $$\overline{\mathrm{\Psi }}A^{\mu \nu \alpha \beta \mathrm{}}\epsilon _{\mu \nu }^{\lambda \rho }\gamma _\rho D_{\lambda \alpha \beta \mathrm{}}^n\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.0.$$ (A.13) For $`\mathrm{\Gamma }=\epsilon _{\mu \nu \rho \tau }\gamma ^\tau `$, one obtains $$\overline{\mathrm{\Psi }}A^{\mu \nu \rho \alpha \mathrm{}}\epsilon _{\mu \nu \rho }^\lambda D_{\lambda \alpha \mathrm{}}^n\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.im\overline{\mathrm{\Psi }}A^{\mu \nu \rho \alpha \mathrm{}}\epsilon _{\mu \nu \rho }^\lambda \gamma _\lambda D_\alpha \mathrm{}^{n1}\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.,$$ (A.14) $$\overline{\mathrm{\Psi }}A^{\mu \nu \rho \alpha \mathrm{}}\gamma _5(g_\mu ^\lambda \sigma _{\nu \rho }+g_\rho ^\lambda \sigma _{\mu \nu }+g_\nu ^\lambda \sigma _{\rho \mu })D_{\lambda \alpha \mathrm{}}^n\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.0,$$ (A.15) and $`\mathrm{\Gamma }=\epsilon _{\mu \nu \rho \tau }\gamma _5\gamma ^\tau `$ gives $`\overline{\mathrm{\Psi }}A^{\mu \nu \rho \alpha \mathrm{}}\left(g_\mu ^\lambda \sigma _{\nu \rho }+g_\rho ^\lambda \sigma _{\mu \nu }+g_\nu ^\lambda \sigma _{\rho \mu }\right)D_{\lambda \alpha \mathrm{}}^n\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.`$ $`im\overline{\mathrm{\Psi }}A^{\mu \nu \rho \alpha \mathrm{}}\epsilon _{\mu \nu \rho }^\tau \gamma _5\gamma _\tau D_\alpha \mathrm{}^{n1}\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.,`$ (A.16) $$\overline{\mathrm{\Psi }}A^{\mu \nu \rho \alpha \mathrm{}}\gamma _5\epsilon _{\mu \nu \rho }^\lambda D_{\lambda \alpha \mathrm{}}^n\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.0.$$ (A.17) The relations (A.4A.17) can be used to eliminate some terms from the list of all the possible invariants, but it is preferable to use them already at the stage of generating such a list.<sup>#8</sup><sup>#8</sup>#8In and implications of the nucleon EOM have been expressed in the form of six identities. The first five of them are equivalent to each other, and they are consequences of our relations (A.4), (A.6A.14) and (A.17). The sixth relation of follows from our relations (A.3) and (A.5), and is not given explicitly (although is accounted for) by . The sixth relation of is neither given, nor used by . The relation follows from our relations (A.15A.16), however, not for a general $`A^{\mu \nu \rho \mathrm{}}`$, but rather only for $`A^{\mu \nu \rho \mathrm{}}`$ antisymmetric in $`\mu \nu `$ (or $`\mu \rho `$ or $`\nu \rho `$). This was exactly what we did, e.g. when counting $`\gamma _5`$ as a quantity of first chiral order in Table 1, or when considering the special structure of $`\mathrm{\Theta }_{\mu \nu \mathrm{}}`$ in eq.(2.11). In this way, the relations (A.3A.5) and (A.7A.9) were taken into account. The relations (A.6), (A.11) and (A.14) are also easily accounted for, namely by ignoring all structures $`\mathrm{\Theta }_{\mu \nu \mathrm{}}`$ with an explicit $`\gamma _\mu `$ (but not with an explicit $`\gamma _5\gamma _\mu `$). The relations (A.10) and (A.12A.13) are simultaneously embodied in the simple requirement that $`\mathrm{\Theta }_{\mu \nu \mathrm{}}`$ can contain an $`\epsilon `$–tensor with at most one index contracted within $`\mathrm{\Theta }_{\mu \nu \mathrm{}}`$. The relation (A.15) is just a simple consequence of eq.(A.12), and the relation (A.16) allows to ignore $`\mathrm{\Theta }_{\mu \nu \rho \mathrm{}}`$ containing $`\epsilon _{\mu \nu \rho \tau }\gamma _5\gamma ^\tau `$. So far we have investigated the implications of the nucleon EOM on the structure of $`\mathrm{\Theta }_{\mu \nu \mathrm{}}`$. There are, however, also consequences for the structure of $`A^{\mu \nu \mathrm{}}`$, namely for $`A^{\mu \nu \mathrm{}}=[D^\mu ,B^\nu \mathrm{}]`$. The point is that $`D^\mu `$ can be reshuffled to act on nucleon fields (by partial integration and exclusion of total derivatives) and, in some cases, eliminated afterwards. Such an elimination is obviously possible if $`D^\mu `$ is contracted with $`\gamma _\mu D_\nu \mathrm{}^n`$ and consequently also if it is contracted with $`\gamma _\nu D_\mu \mathrm{}^n`$ (cf. eq.(A.11)) or $`D_{\mu \nu \mathrm{}}^n`$ (cf. eq.(A.6)). For $`D^\mu `$ contracted with $`\gamma _5\gamma _\mu D_\nu \mathrm{}^n`$ and $`\gamma _5\gamma _\nu D_\mu \mathrm{}^n`$ one obtains in a similar way $$\overline{\mathrm{\Psi }}[D^\mu ,B^\nu \mathrm{}]\gamma _5\gamma _\mu D_\nu \mathrm{}^n\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.2im\overline{\mathrm{\Psi }}B^\nu \mathrm{}\gamma _5D_\nu \mathrm{}^n\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.,$$ (A.18) $$\overline{\mathrm{\Psi }}[D^\mu ,B^\nu \mathrm{}]\gamma _5\gamma _\nu D_\mu \mathrm{}^n\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.0,$$ (A.19) and for $`D^\mu `$ contracted with $`\sigma _{\nu \rho }D_\mu \mathrm{}^n`$ $`\overline{\mathrm{\Psi }}[D^\mu ,B^{\nu \rho \mathrm{}}]\sigma _{\nu \rho }D_\mu \mathrm{}^n\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.`$ $``$ $`i\overline{\mathrm{\Psi }}[D^\mu ,B^{\nu \rho \mathrm{}}]g_{\mu \nu }D_\rho \mathrm{}^n\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.`$ (A.20) $``$ $`i\overline{\mathrm{\Psi }}[D^\mu ,B^{\nu \rho \mathrm{}}]g_{\mu \rho }D_\nu \mathrm{}^n\mathrm{\Psi }+\mathrm{h}.\mathrm{c}.`$ $`+`$ $`im\overline{\mathrm{\Psi }}[D^\mu ,B^{\nu \rho \mathrm{}}]\epsilon _{\mu \nu \rho }^\tau \gamma _5\gamma _\tau D_{\mathrm{}}^{n1}\mathrm{\Psi }+\mathrm{h}.\mathrm{c}..`$ Note that the omitted contraction of $`D^\mu `$ with $`\sigma _{\mu \nu }D_\rho \mathrm{}^n`$ yields nothing but a trivial identity. The relations (A.18A.20) are used to ignore terms of the type written on the left–hand–side. However, in case of eq.(A.18) it is more advantageous to keep the structure on the left–hand–side and eliminate the right–hand–side, which does not contain the special structure $`[D^\mu ,B^\nu \mathrm{}]`$, but rather a general structure $`B^\nu \mathrm{}`$. To summarize: we utilize the nucleon EOM in the form of restrictions on the structure of $`\mathrm{\Gamma }_{\mu \nu \mathrm{}}`$ and $`A^{\mu \nu \mathrm{}}`$: * $`\mathrm{\Gamma }_{\mu \nu \mathrm{}}`$ is a product of $`g`$-tensors, $`\epsilon `$-tensors and one matrix from the following set $`\{1,\gamma _5\gamma _\mu ,\sigma _{\mu \nu }\}`$, * within $`\mathrm{\Gamma }_{\mu \nu \mathrm{}}`$, only $`\epsilon `$-tensors may have an upper index (every $`\epsilon `$ at most one), which has to be contracted with $`D_{\alpha \beta \mathrm{}}^n`$, * for $`A^{\mu \nu \mathrm{}}=[D^\mu ,B^\nu \mathrm{}]`$, derivative $`D^\mu `$ cannot be contracted with $`D_{\mu \alpha \mathrm{}}^n`$ This is not the only way of implementing the nucleon EOM, but it seems to be the most economic one. It results in the following set of $`\mathrm{\Theta }_{\mu \nu \mathrm{}}`$, sufficient for generating the complete chiral Lagrangian up to fourth order: $`1;`$ $`\gamma _5\gamma _\mu ,D_\mu ;`$ $`g_{\mu \nu },\sigma _{\mu \nu },\gamma _5\gamma _\mu D_\nu ,D_{\mu \nu };`$ $`g_{\mu \nu }\gamma _5\gamma _\rho ,g_{\mu \nu }D_\rho ,\sigma _{\mu \nu }D_\rho ,\epsilon _{\mu \nu \rho }^\lambda D_\lambda ,\gamma _5\gamma _\mu D_{\nu \rho },D_{\mu \nu \rho };`$ $`g_{\mu \nu }g_{\rho \tau },\epsilon _{\mu \nu \rho \tau },g_{\mu \nu }\sigma _{\rho \tau },g_{\mu \nu }\gamma _5\gamma _\rho D_\tau ,g_{\mu \nu }D_{\rho \tau },\sigma _{\mu \nu }D_{\rho \tau },`$ $`\epsilon _{\mu \nu \rho }^\lambda D_{\lambda \tau },\gamma _5\gamma _\mu D_{\nu \rho \tau },D_{\mu \nu \rho \tau }.`$ (A.21) ## Appendix B Operators for the heavy baryon Lagrangian In order to construct the $`1/m`$-corrections up-to-and-including fourth order, one needs the expressions for the operators $`𝒜^{(1,2,3,4)}`$, $`^{(1,2,3)}`$, and $`𝒞^{(0,1,2)}`$, the latter allowing one to reconstruct the chiral expansion of the inverse of $`𝒞`$. These are given by: $`𝒞^{(0)}`$ $`=`$ $`2m,`$ (B.1) $`𝒜^{(1)}`$ $`=`$ $`i(vD)+g_A(Su),`$ (B.2) $`^{(1)}`$ $`=`$ $`\gamma _5\left(2i(SD)+{\displaystyle \frac{g_A}{2}}(vu)\right),`$ (B.3) $`𝒞^{(1)}`$ $`=`$ $`i(vD)+g_A(Su),`$ (B.4) $`𝒜^{(2)}`$ $`=`$ $`c_1\chi _++{\displaystyle \frac{c_2}{2}}(vu)^2+{\displaystyle \frac{c_3}{2}}u^2+c_5\stackrel{~}{\chi }_+`$ (B.5) $``$ $`2i[S^\mu ,S^\nu ]\left({\displaystyle \frac{ic_4}{4}}[u_\mu ,u_\nu ]+{\displaystyle \frac{c_6}{8m}}F_{\mu \nu }^++{\displaystyle \frac{c_7}{8m}}F_{\mu \nu }^+\right),`$ $`^{(2)}`$ $`=`$ $`2i\gamma _5(v^\mu S^\nu v^\nu S^\mu )\left({\displaystyle \frac{ic_4}{4}}[u_\mu ,u_\nu ]+{\displaystyle \frac{c_6}{8m}}F_{\mu \nu }^++{\displaystyle \frac{c_7}{8m}}F_{\mu \nu }^+\right),`$ (B.6) $`𝒞^{(2)}`$ $`=`$ $`𝒜^{(2)},`$ (B.7) $`𝒜^{(3)}`$ $`=`$ $`(i{\displaystyle \frac{c_2}{2m}}(vu)u_\mu D^\mu +\mathrm{h}.\mathrm{c}.)+E+2S^\mu G_\mu 2i[S^\mu ,S^\nu ]H_{\mu \nu },`$ (B.8) $`^{(3)}`$ $`=`$ $`\gamma _5\left(v^\mu G_\mu +2i(v^\mu S^\nu v^\nu S^\mu )H_{\mu \nu }\right),`$ (B.9) $`𝒜^{(4)}`$ $`=`$ $`{\displaystyle \frac{c_2}{8m^2}}u_\mu u_\nu \{D^\mu ,D^\nu \}+\mathrm{h}.\mathrm{c}.`$ (B.10) $`{\displaystyle \frac{d_1}{2m}}\left([u_\mu ,[D_\nu ,u^\mu ]]D^\nu +\text{h.c.}\right){\displaystyle \frac{d_2}{2m}}\left([u_\mu ,[D^\mu ,u_\nu ]]D^\nu +\text{h.c.}\right)`$ $`{\displaystyle \frac{d_3}{2m}}\left([vu,[vD,u_\mu ]]D^\mu +[vu,[D_\mu ,vu]]D^\mu +[u_\mu ,[vD,vu]]D^\mu +\text{h.c.}\right)`$ $`{\displaystyle \frac{d_4}{2m}}\left(ϵ^{\mu \nu \alpha \beta }u_\mu u_\nu u_\alpha D_\beta +\text{h.c.}\right)+{\displaystyle \frac{d_5}{2m}}\left(i[\chi _{},u_\mu ]D^\mu +\text{h.c.}\right)`$ $`+{\displaystyle \frac{d_6}{2m}}\left(i[D^\mu ,\stackrel{~}{F}_{\mu \nu }^+]D^\nu +\text{h.c.}\right)+{\displaystyle \frac{d_7}{2m}}\left(i[D^\mu ,F_{\mu \nu }^+]D^\nu +\text{h.c.}\right)`$ $`+{\displaystyle \frac{d_8}{2m}}\left(iϵ^{\mu \nu \alpha \beta }\stackrel{~}{F}_{\mu \nu }^+u_\alpha D_\beta +\text{h.c.}\right)+{\displaystyle \frac{d_9}{2m}}\left(iϵ^{\mu \nu \alpha \beta }F_{\mu \nu }^+u_\alpha D_\beta +\text{h.c.}\right)`$ $`+{\displaystyle \frac{d_{12}}{m}}\left(ivuu_\mu SuD^\mu +\text{h.c.}\right)`$ $`+{\displaystyle \frac{d_{13}}{2m}}\left(iS^\mu u_\mu vuu_\nu D^\nu +iS^\mu u_\mu u_\nu vuD^\nu +\text{h.c.}\right)`$ $`+{\displaystyle \frac{d_{14}}{2m}}\left([S^\mu ,S^\nu ][D_\lambda ,u_\mu ]u_\nu D^\lambda +\text{h.c.}\right)`$ $`+{\displaystyle \frac{d_{15}}{2m}}\left([S^\mu ,S^\nu ]u_\mu [D_\nu ,u_\lambda ]D^\lambda +\text{h.c.}\right)`$ $`{\displaystyle \frac{d_{20}}{2m}}\left(S^\mu v^\nu [\stackrel{~}{F}_{\mu \nu }^+,u_\lambda ]D^\lambda +S^\mu [\stackrel{~}{F}_{\mu \nu }^+,vu]D^\nu +\text{h.c.}\right),`$ with $`E`$ $`=`$ $`id_1[u_\mu ,[vD,u^\mu ]]+id_2[u_\mu ,[D^\mu ,vu]]+id_3[vu,[vD,vu]]`$ (B.11) $`+`$ $`id_4ϵ^{\mu \nu \alpha \beta }v_\beta u_\mu u_\nu u_\alpha +d_5[\chi _{},vu]+d_6v^\nu [D^\mu ,\stackrel{~}{F}_{\mu \nu }^+]`$ $`+`$ $`d_7v^\nu [D^\mu ,F_{\mu \nu }^+]+d_8ϵ^{\mu \nu \alpha \beta }v^\beta \stackrel{~}{F}_{\mu \nu }^+u_\alpha +d_9ϵ^{\mu \nu \alpha \beta }v^\beta F_{\mu \nu }^+u_\alpha ,`$ $`G_\mu `$ $`=`$ $`{\displaystyle \frac{d_{10}}{2}}u^2u_\mu +{\displaystyle \frac{d_{11}}{2}}u_\mu u_\nu u^\nu +{\displaystyle \frac{d_{12}}{2}}(vu)^2u_\mu +{\displaystyle \frac{d_{13}}{2}}u_\mu (vu)(vu)`$ (B.12) $`+{\displaystyle \frac{d_{16}}{2}}\chi _+u_\mu +{\displaystyle \frac{d_{17}}{2}}\chi _+u_\mu +{\displaystyle \frac{id_{18}}{2}}[D_\mu ,\chi _{}]+{\displaystyle \frac{id_{19}}{2}}[D_\mu ,\chi _{}]+{\displaystyle \frac{id_{20}}{2}}v^\nu [\stackrel{~}{F}_{\mu \nu }^+,vu]`$ $`+{\displaystyle \frac{id_{21}}{2}}[\stackrel{~}{F}_{\mu \nu }^+,u^\nu ]+{\displaystyle \frac{d_{22}}{2}}[D^\nu ,F_{\mu \nu }^{}]+{\displaystyle \frac{d_{23}}{2}}ϵ_\mu ^{\nu \alpha \beta }u_\nu F_{\alpha \beta }^{},`$ $`H_{\mu \nu }`$ $`=`$ $`{\displaystyle \frac{d_{14}}{2}}[vD,u_\mu ]u_\nu +{\displaystyle \frac{d_{15}}{2}}u_\mu [D_\nu ,vu],`$ (B.13) and $`𝒞^1`$ $`=`$ $`{\displaystyle \frac{1}{2m}}{\displaystyle \frac{ivD+g_ASu}{(2m)^2}}+{\displaystyle \frac{(ivD+g_ASu)^2}{(2m)^3}}{\displaystyle \frac{𝒞^{(2)}}{(2m)^2}}.`$ (B.14)
warning/0001/math0001182.html
ar5iv
text
# The trace formula for transversally elliptic operators on Riemannian foliations ## 1 Introduction The main goal of the paper is to generalize the Duistermaat-Guillemin trace formula to the case of transversally elliptic operators on a compact foliated manifold. First, let us recall briefly the setting of the classical formula. Let $`P`$ be a positive self-adjoint elliptic pseudodifferential operator of order one on a closed manifold $`M`$ (for example, $`P=\sqrt{\mathrm{\Delta }}`$, where $`\mathrm{\Delta }`$ is the Laplace-Beltrami operator of a Riemannian metric on $`M`$). For any function $`fC_c^{\mathrm{}}()`$, the operator $`U_f=f(t)e^{itP}𝑑t`$ can be shown to be of trace class, and the mapping $`\theta :f\mathrm{tr}U_f`$ is a continuous linear functional on $`C_c^{\mathrm{}}()`$. Otherwise speaking, $`\theta `$ is a distribution on $``$. The principal symbol $`p`$ of the operator $`P`$ is a smooth function on the symplectic manifold $`T^{}M0`$, and, by definition, the bicharacteristic flow $`f_t`$ of the operator $`P`$ is the Hamiltonian flow on $`T^{}M0`$ defined by the function $`p`$. By the theorem due to Colin de Verdiére and Chazarain , the singularities of the distribution $`\theta `$ are contained in the period set of closed trajectories of the bicharacteristic flow $`f_t`$. Moreover, Duistermaat and Guillemin showed that, under the assumption that the bicharacteristic flow is clean, one can write down an asymptotic expansion for the distribution $`\theta `$ near a given period of closed bicharacteristic. A formula for the leading term of this asymptotic expansion is the Duistermaat-Guillemin trace formula mentioned above. It involves the geometry of the bicharacteristic flow in the form of Poincaré map and Maslov indices and provides a far-reaching generalization of the classical Poisson formula and the Selberg trace formula on hyperbolic spaces. In the paper, we prove a trace formula for an operator $`P=\sqrt{A}`$, where $`A`$ is a positive self-adjoint transversally elliptic pseudodifferential operator of order two with the positive, holonomy invariant transverse principal symbol on a compact foliated manifold $`(M,)`$ (see Theorem 6). One can consider such an operator as an elliptic operator on the singular space $`M/`$ of leaves of the foliation $``$ (this statement can be made more precise, using the language of noncommutative geometry, see ) the trace formula stated in this paper as an example of a trace formula for elliptic operators on singular spaces. We hope that this formula will be useful in further study of a general trace formula in noncommutative geometry (see, for instance, for discussion of a noncommutative trace formula). It should be also noted that our trace formula can be viewed as a relative version of the Duistermaat-Guillemin trace formula. ## 2 Preliminaries and main results Let $`(M,)`$ be a compact, connected, oriented foliated manifold. We will use the following notation: $`T`$ is the tangent bundle, $`H=TM/T`$ is the normal bundle and $`N^{}`$ is the conormal bundle to $``$. There is a short exact sequence $$0TTMH0.$$ (2.1) We will consider linear operators, acting on half-densities. Recall that an $`\alpha `$-density ($`\alpha `$) on a real vector space $`V`$ of dimension $`n`$ is a map $`\varphi :\mathrm{\Lambda }^nV`$ such that $`\varphi (\lambda v)=|\lambda |^\alpha \varphi (v),v\mathrm{\Lambda }^nV,\lambda `$. For any real vector bundle $`E`$ over a smooth manifold $`X`$, we will denote by $`|E|^\alpha `$ the $`\alpha `$-density bundle of $`E`$. Given a pseudodifferential operator $`A\mathrm{\Psi }^m(M,|TM|^{1/2})`$, the transversal principal symbol $`\sigma _A`$ of $`A`$ is defined to be the restriction of its principal symbol $`a_m`$ on $`\stackrel{~}{N}^{}=N^{}0`$. An operator $`A\mathrm{\Psi }^m(M,|TM|^{1/2})`$ is said to be transversally elliptic, if $`\sigma _A(\nu )0`$ for any $`\nu \stackrel{~}{N}^{}`$. For any smooth leafwise path $`\gamma `$ from $`xM`$ to $`yM`$, sliding along leaves of the foliation defines the holonomy map $`h_\gamma `$, which associates to every germ of a local transversal to the foliation at the point $`x`$ a germ of a local transversal to the foliation at the point $`y`$ (this map is a natural generalization of the Poincaré first-return map for flows). The differential of this map (the linear holonomy map) is well-defined as a linear map $`dh_\gamma :H_xH_y`$ and the codifferential as a linear map $`dh_\gamma ^{}:N_y^{}N_x^{}`$. The transversal principal symbol $`\sigma _A`$ of an operator $`A\mathrm{\Psi }^m(M,|TM|^{1/2})`$ is said to be holonomy invariant, if $`\sigma _A(dh_\gamma ^{}(\nu ))=\sigma _A(\nu )`$ for any smooth leafwise path $`\gamma `$ from $`x`$ to $`y`$ and for any $`\nu \stackrel{~}{N}_y^{}`$. Throughout in the paper, we will assume that $`A`$ is a linear operator in $`C^{\mathrm{}}(M,|TM|^{1/2})`$, satisfying the following conditions: (A1) $`A\mathrm{\Psi }^2(M,|TM|^{1/2})`$ is a transversally elliptic operator with the positive, holonomy invariant transversal principal symbol; (A2) $`A`$ is an essentially self-adjoint positive operator in $`L^2`$ space of half-densities on $`M`$, $`L^2(M)`$ (with the initial domain $`C^{\mathrm{}}(M,|TM|^{1/2})`$). ###### Example 1. A geometrical example of an operator, satisfying the conditions (A1) and (A2), is given by the operator $`A=I+\mathrm{\Delta }_H`$, where $`\mathrm{\Delta }_H`$ is the transversal Laplacian of a bundle-like metric on a Riemannian foliation. Recall that a foliation $``$ on a smooth Riemannian manifold $`(M,g_M)`$ is Riemannian if it satisfies one of the following equivalent conditions (see, for instance, ): 1. $`(M,)`$ locally has the structure of Riemannian submersion; 2. the transverse part of the Riemannian metric $`g_M`$ (that is, its restriction to $`H=T^{}`$) is holonomy invariant; 3. the horizontal distribution $`H`$ is totally geodesic. In this case, the metric $`g_M`$ is called bundle-like. The Riemannian metric $`g_M`$ defines a decomposition of the cotangent bundle $`T^{}M`$ into a direct sum $`T^{}M=F^{}H^{}`$. With respect to this decomposition, the de Rham differential $`d:C^{\mathrm{}}(M)C^{\mathrm{}}(M,T^{}M)`$ can be written as a sum $`d=d_F+d_H,`$ where $`d_F:C^{\mathrm{}}(M)C^{\mathrm{}}(M,F^{})`$ and $`d_H:C^{\mathrm{}}(M)C^{\mathrm{}}(M,H^{})`$. The transversal Laplacian is a second order transversally elliptic differential operator in the space $`C^{\mathrm{}}(M)`$ defined by the formula $$\mathrm{\Delta }_H=d_H^{}d_H.$$ Its principal symbol $`a_2`$ is given by the formula $$a_2(x,\xi )=g_H(\xi ,\xi ),(x,\xi )\stackrel{~}{T}^{}M,$$ and the holonomy invariance of the transverse principal symbol $`\sigma _{\mathrm{\Delta }_H}`$ is equivalent to the assumption on the Riemannian metric $`g_M`$ to be bundle-like. From now on, we will assume that $`A`$ satisfies the assumptions (A1) and (A2). By the spectral theorem, the operator $`P=\sqrt{A}`$ generates a strongly continuous group $`e^{itP}`$ of bounded operators in $`L^2(M)`$. To define a distributional trace of the operator $`e^{itP}`$, one need an additional regularization. First, let us introduce some notation. Recall that the holonomy groupoid $`G=G_{}`$ of the foliation $``$ is the set of equivalence classes of leafwise paths $`\gamma :[0,1]M`$ with respect to an equivalence relation $`_h`$, setting $`\gamma _1_h\gamma _2`$ if $`\gamma _1`$ and $`\gamma _2`$ have the same initial and final points and the same holonomy maps. $`G`$ is equipped with maps $`s,r:GM`$ given by $`s(\gamma )=\gamma (0)`$ and $`r(\gamma )=\gamma (1)`$ and has a composition law given by the composition of paths. For any $`\gamma _1,\gamma _2G`$, the composition $`\gamma _1\gamma _2`$ makes sense iff $`r(\gamma _2)=s(\gamma _1)`$. We will make use of standard notation: $`G^x=r^1(x)`$, $`G_x=s^1(x)`$, $`G_x^x=s^1(x)r^1(x)`$, $`xM`$. For any $`xM`$, $`G_x^x`$ is the holonomy group of the leaf $`L_x`$ through the point $`x`$ and the maps $`s:G^xL_x`$ and $`r:G_xL_x`$ are covering maps associated with $`G_x^x`$. We will identify a point $`xM`$ with the element in $`G`$ given by the constant path $`\gamma (t)=x,t[0,1]`$. Let $`s^{}(|T|^{1/2})`$ and $`r^{}(|T|^{1/2})`$ be the lifts of the vector bundle of leafwise half-densities $`|T|^{1/2}`$ to vector bundles on $`G`$ via the mappings $`s`$ and $`r`$ respectively, and $`|T𝒢|^{1/2}=r^{}(|T|^{1/2})s^{}(|T|^{1/2}).`$ The line bundle $`|T𝒢|^{1/2}`$ is the bundle of leafwise half-densities on $`G`$ with respect to the natural foliation $`𝒢`$ . The space $`C_c^{\mathrm{}}(G,|T𝒢|^{1/2})`$ has the structure of involutive algebra (see, for instance, ). There is a natural $``$-representation $`R`$ of the involutive algebra $`C_c^{\mathrm{}}(G,|T𝒢|^{1/2})`$ in $`L^2(M)`$. For any $`kC_c^{\mathrm{}}(G,|T𝒢|^{1/2})`$, the operator $`R(k)`$ in $`L^2(M)`$ is defined as follows. According to the short exact sequence (2.1), the half-density vector bundle $`|TM|^{1/2}`$ can be decomposed as $$|TM|^{1/2}|T|^{1/2}|H|^{1/2}.$$ For any $`\gamma G,s(\gamma )=x,r(\gamma )=y`$, the corresponding linear holonomy map defines a map $$dh_\gamma ^{}:|H_y|^{1/2}|H_x|^{1/2}.$$ Given $`uL^2(M)`$ of the form $`u=u_1u_2,u_1L^2(M,|T|^{1/2}),u_2L^2(M,|H|^{1/2}),`$ $`R(k)uL^2(M)`$ is defined by the formula $$R(k)u(x)=_{G^x}k(\gamma )s^{}u_1(\gamma )𝑑h_{\gamma ^1}^{}[u_2(s(\gamma ))],xM.$$ ###### Proposition 2. For any $`kC_c^{\mathrm{}}(G,|T𝒢|^{1/2})`$ and $`fC_c^{\mathrm{}}()`$, the operator $`R(k)f(t)e^{itP}𝑑t`$ is of trace class. Moreover, for any $`kC_c^{\mathrm{}}(G,|T𝒢|^{1/2})`$, the formula $$\theta _k,f=\mathrm{tr}R(k)f(t)e^{itP}𝑑t,fC_c^{\mathrm{}}(),$$ defines a distribution $`\theta _k`$ on the real line $``$, $`\theta _k𝒟^{}()`$. Let $`_N`$ be a foliation in $`\stackrel{~}{N}^{}`$, which is the horizontal foliation for the natural leafwise flat connection in $`\stackrel{~}{N}^{}`$ (the Bott connection). The leaf of the foliation $`_N`$ through a point $`\nu \stackrel{~}{N}^{}`$ is the set of all $`dh_\gamma ^{}(\nu )\stackrel{~}{N}^{}`$ such that $`\gamma G,r(\gamma )=\pi (\nu )`$, where $`\pi :N^{}M`$ is the natural projection. Denote by $`H_N`$ the normal bundle to $`_N`$, $`H_N=T(N^{})/T_N`$. For any $`\nu N^{}`$, the space $`T_\nu N^{}`$ is a coisotropic subspace of the space $`T_\nu T^{}M`$ equipped with the canonical symplectic structure, and $`T_\nu _N`$ its skew-orthogonal complement, therefore, the normal bundle $`H_\nu _N`$ has a natural symplectic structure (see, for instance, ). Given an operator $`A`$ under the conditions (A1) and (A2) with the principal symbol $`a`$, let $`\stackrel{~}{p}`$ be a smooth function on $`\stackrel{~}{T}^{}M`$ homogeneous of degree one such that $`\stackrel{~}{p}(\xi )0`$ for $`\xi \stackrel{~}{T}^{}M`$, which is equal to $`p=a^{1/2}`$ in some conical neighborhood of $`N^{}`$, and $`\stackrel{~}{f}_t`$ the Hamiltonian flow of the function $`\stackrel{~}{p}`$. Define $`\sigma _P`$ to be the restriction of $`p`$ on $`N^{}`$: $`\sigma _P=\sigma _A^{1/2}`$. The function $`\sigma _P`$ coincides with the transverse principal symbol of any operator $`P_1\mathrm{\Psi }^1(M,|TM|^{1/2})`$ such that the principal symbols of $`P_1^2`$ and $`A`$ are equal on $`N^{}`$. The holonomy invariance assumption on $`\sigma _A`$ implies $$d\stackrel{~}{p}(\nu )(X)=0,\nu \stackrel{~}{N}^{},XT_\nu _N.$$ (2.2) Using (2.2) and the fact that $`\stackrel{~}{f}_t`$ preserves the symplectic structure of $`T^{}M`$, one can easily check that the Hamiltonian flow $`\stackrel{~}{f}_t`$ can be restricted on $`N^{}`$. The resulting flow will be denoted by $`f_t`$. By definition, the flow $`f_t`$ depends only on the $`1`$-jet of the principal symbol $`a`$ on $`N^{}`$, therefore, it doesn’t depend on a choice of $`\stackrel{~}{p}`$ and can be naturally called the transverse bicharacteristic flow of the operator $`A`$. Since $`\stackrel{~}{f}_t`$ preserves the symplectic structure of $`T^{}M`$ and $`T_N`$ is the skew-adjoint complement to $`TN^{}`$, $`f_t`$ maps leaves of the foliation $`_N`$ to leaves. In particular, the differential $`df_t`$ defines a map $`T_\nu _NT_{f_t(\nu )}_N`$ and a symplectic map $`H_\nu _NH_{f_t(\nu )}_N`$. We say that a point $`\nu \stackrel{~}{N}^{}`$ is a relative fixed point of the diffeomorphism $`f_t:\stackrel{~}{N}^{}\stackrel{~}{N}^{}`$ (with respect to the foliation $`_N`$), if there exist $`\gamma G`$ such that $`r(\gamma )=\pi (\nu )`$ and $`f_tdh_\gamma ^{}(\nu )=\nu `$. For any $`t`$, denote by $`Z_t`$ the set of relative fixed points of $`f_t`$. We also introduce the corresponding set in the cospherical bundle $`SN^{}=\{\nu N^{}:\sigma _P(\nu )=1\}`$: $`SZ_t=Z_tSN^{}.`$ This set might be not closed, but, for any $`kC_c^{\mathrm{}}(G,|T𝒢|^{1/2}))`$, the corresponding part $$SZ_{t,k}=\{\nu SN^{}:(\gamma \mathrm{supp}k,r(\gamma )=\pi (\nu ))f_tdh_\gamma ^{}(\nu )=\nu \}$$ is closed. By the tranversal ellipticity of $`\sigma _A`$, the flow $`f_t`$ is transverse to $`_N`$, therefore, the relative period set $`𝒯_k=\{t:SZ_{t,k}\mathrm{}\}`$ is a discrete subset of $``$. The following theorem was proved in , but we will give an independent proof. ###### Theorem 3. Given an operator $`A`$ under the conditions (A1) and (A2) and $`kC_c^{\mathrm{}}(G,|T𝒢|^{1/2})`$, the distribution $`\theta _k`$ is smooth outside of the relative period set $`𝒯_k`$ of the transverse bicharacterictic flow $`f_t`$. Let $`G__N`$ denote the holonomy groupoid of the foliation $`_N`$. $`G__N`$ consists of all pairs $`(\gamma ,\nu )G_{}\times \stackrel{~}{N}^{}`$ such that $`r(\gamma )=\pi (\nu )`$ with the source map $`s_N:G__N\stackrel{~}{N}^{},s_N(\gamma ,\nu )=dh_\gamma ^{}(\nu )`$, and the target map $`r_N:G__N\stackrel{~}{N}^{},r_N(\gamma ,\nu )=\nu `$. There is a projection $`\pi _G:G__NG_{}`$ given by the formula $`\pi _G(\gamma ,\nu )=\gamma `$. Put also $`G_{SN^{}}=G__Nr_N^1(SN^{})`$. For any $`(\gamma ,\nu )G__N`$, denote by $`dH_{(\gamma ,\nu )}`$ the associated linear holonomy map: $$dH_{(\gamma ,\nu )}:H_{dh_\gamma ^{}(\nu )}_NH_\nu _N.$$ It is easy to see that $`dH_{(\gamma ,\nu )}`$ preserves the symplectic structure of $`H_N`$. Denote by $`Q:TN^{}H_N`$ the projection map. The differential of the map $`(r_N,s_N):G__N\stackrel{~}{N}^{}\times \stackrel{~}{N}^{}`$ defines an isomorphism of the tangent space $`T_{(\gamma ,\nu )}G__N`$ with the set of all $`(v_1,v_2)T_\nu N^{}T_{dh_\gamma ^{}(\nu )}N^{}`$ such that the normal components of $`v_1`$ and $`v_2`$ are connected by the holonomy map: $`Q(v_1)=dH_{(\gamma ,\nu )}(Q(v_2))`$. The holonomy groupoid $`G__N`$ has the natural foliation $`𝒢__N`$ such that the tangent bundle $`T_{(\gamma ,\nu )}𝒢__N`$ corresponds to $`T_\nu _NT_{dh_\gamma ^{}(\nu )}_N`$ under the isomorphism described above. The normal space $`H_{(\gamma ,\nu )}𝒢__N`$ to $`𝒢__N`$ is isomorphic to the set of all $`(v_1,v_2)H_\nu _NH_{dh_\gamma ^{}(\nu )}_N`$ such that $`v_1=dH_{(\gamma ,\nu )}(v_2)`$, and therefore the maps $`dr_N`$ ($`ds_N`$) define isomorphisms of $`H_{(\gamma ,\nu )}𝒢__N`$ with $`H_\nu _N`$ ($`H_{dh_\gamma ^{}(\nu )}_N`$) respectively. ###### Lemma 4. The set $`Z_t`$ is a saturated subset of $`N^{}`$, that is, it is a union of leaves of the foliation $`_N`$. ###### Proof. By the holonomy invariance of $`p`$, the Hamiltonian vector field $`\mathrm{\Xi }_p`$ with the Hamiltonian $`p`$ satisfies the identity $$dH_{(\gamma ,\nu )}(Q(\mathrm{\Xi }_p(dh_\gamma ^{}(\nu ))))=Q(\mathrm{\Xi }_p(\nu )),\nu \stackrel{~}{N}^{},\gamma G,r(\gamma )=\pi (\nu ),$$ therefore, there exists a vector field $`\widehat{\mathrm{\Xi }}_p`$ on $`G__N`$ such that $$ds_N(\widehat{\mathrm{\Xi }}_p(\gamma ,\nu ))=\mathrm{\Xi }_p(dh_\gamma ^{}(\nu )),dr_N(\widehat{\mathrm{\Xi }}_p(\gamma ,\nu ))=\mathrm{\Xi }_p(\nu ),(\gamma ,\nu )G__N.$$ (2.3) Let $`\widehat{F}_t`$ be a flow on $`G__N`$ generated by $`\widehat{\mathrm{\Xi }}_p`$. By (2.3), we have $$f_tr_N=r_N\widehat{F}_t,f_ts_N=s_N\widehat{F}_t,$$ or, if we write $`\widehat{F}_t:G__NG__N`$ as $`\widehat{F}_t(\gamma ,\nu )=(F_t(\gamma ,\nu ),f_t(\nu ))`$, $$f_t(dh_\gamma ^{}(\nu ))=dh_{F_t(\gamma ,\nu )}^{}(f_t(\nu )).$$ (2.4) Take any $`\nu Z_t`$ with the corresponding $`\gamma G`$ such $`r(\gamma )=\pi (\nu ),f_tdh_\gamma ^{}(\nu )=\nu `$. Let $`(\gamma _1,\nu )G__N`$. Then we have $`f_t(dh_{\gamma _1}^{}(\nu ))`$ $`=dh_{F_t(\gamma _1,\nu )}^{}(f_t(\nu ))\text{by (}\text{2.4}\text{)}`$ $`=dh_{F_t(\gamma _1,\nu )}^{}(dh_\gamma ^{}(\nu ))`$ $`=(dh_{F_t(\gamma _1,\nu )}^{}dh_\gamma ^{}dh_{\gamma _1^1}^{})(dh_{\gamma _1}^{}(\nu ))`$ $`=dh_\gamma ^{}^{}(dh_{\gamma _1}^{}(\nu )),`$ where $`\gamma ^{}=\gamma _1^1\gamma F_t(\gamma _1,\nu )`$ that implies $`dh_{\gamma _1}^{}(\nu )Z_t`$. ∎ The relative fixed point sets $`Z_t`$ can be naturally lifted to the holonomy groupoid $`G__N`$: $$𝒵_t=\{(\gamma ,\nu )G__N:f_tdh_\gamma ^{}(\nu )=\nu \},S𝒵_t=𝒵_tG_{SN^{}},$$ By Lemma 4, $`Z_t=r_N(𝒵_t)=s_N(𝒵_t)`$. Let us assume that $`𝒵_t`$ is a smooth submanifold of $`G__N`$. By Lemma 4, the tangent space to $`𝒵_t`$ at a point $`(\gamma ,\nu )𝒵_t`$ contains a subspace $`F_{(\gamma ,\nu )}𝒵_t`$, which is the graph of the linear map $`df_t(\nu ):T_\nu _NT_{dh_\gamma ^{}(\nu )}_N=T_{f_t(\nu )}_N`$: $$F_{(\gamma ,\nu )}𝒵_t=\{(v_1,v_2)T_\nu _N\times T_{dh_\gamma ^{}(\nu )}_N:v_2=df_t(\nu )(v_1)\}.$$ Let $$H_{(\gamma ,\nu )}𝒵_t=T_{(\gamma ,\nu )}𝒵_t/F_{(\gamma ,\nu )}𝒵_t,H_{(\gamma ,\nu )}S𝒵_t=T_{(\gamma ,\nu )}S𝒵_t/F_{(\gamma ,\nu )}𝒵_t.$$ ###### Definition 5. Let $`t`$ be a relative period of the flow $`f_t`$. We say that the flow $`f_t`$ is clean on $`𝒵_t`$, if: * $`𝒵_t`$ is a smooth submanifold of $`G__N`$; * the normal space $`H_{(\gamma ,\nu )}𝒵_t`$ at any point $`(\gamma ,\nu )𝒵_t`$ coincides with the set of all $`(v_1,v_2)H_{(\gamma ,\nu )}𝒢__N`$ such that $`v_2=df_t(\nu )(v_1).`$ Let $`|T_N|^{1/2}`$ be the vector bundle of leafwise half-densities on $`N^{}`$, and $`s_N^{}(|T_N|^{1/2})`$ and $`r_N^{}(|T_N|^{1/2})`$ are the lifts of this vector bundle to vector bundles on $`G__N`$ via the mappings $`s_N`$ and $`r_N`$ respectively. Let $`|T𝒢__N|^{1/2}`$ be the vector bundle of leafwise half-densities on $`G__N`$: $$|T𝒢__N|^{1/2}=r_N^{}(|T_N|^{1/2})s_N^{}(|T_N|^{1/2}).$$ The projection $`\pi _G:G__NG`$ defines a local diffeomorphism $`\pi _G:𝒢__N𝒢`$, that induces a map $$\pi _G^{}:C_c^{\mathrm{}}(G,|T𝒢|^{1/2})C^{\mathrm{}}(G__N,|T𝒢__N|^{1/2}).$$ Define a restriction map $$R_𝒵:C_c^{\mathrm{}}(G__N,|T𝒢__N|^{1/2})C_c^{\mathrm{}}(𝒵_t,|T_N|^1)$$ as follows. If $`\rho =fr_N^{}\rho _1s_N^{}\rho _2,fC_c^{\mathrm{}}(G__N),\rho _1,\rho _2C_c^{\mathrm{}}(M,|T_N|^{1/2})`$, then $$R_𝒵\rho (\gamma ,\nu )=f(\gamma ,\nu )\rho _1(\gamma ,\nu )df_t^{}(\nu )[\rho _2(dh_\gamma ^{}(\nu ))],(\gamma ,\nu )𝒵_t,$$ where the map $`df_t^{}(\nu ):|T_{f_t(\nu )}_N|^{1/2}|T_\nu _N|^{1/2}`$ is induced by the linear map $`df_t(\nu ):T_\nu _NT_{f_t(\nu )}_N`$. If the flow $`f_t`$ is clean, there is defined a natural density $`d\mu _𝒵`$ on $`H_{(\gamma ,\nu )}𝒵_t`$, being the fixed point set of the symplectic linear map $`dH_{(\gamma ,\nu )}df_t(\nu )`$ of the symplectic space $`H_\nu _N`$ (see, for instance, \[3, Lemma 4.3\]). Dividing $`d\mu _𝒵`$ by $`d\sigma _P`$, we get a density $`d\mu _{S𝒵}`$ on $`H_{(\gamma ,\nu )}S𝒵_t`$. Using the natural isomorphism $$|TS𝒵_t||F𝒵||HS𝒵_t|.$$ one can combine the densities $`R_𝒵\pi _G^{}kC_c^{\mathrm{}}(S𝒵_t,|FS𝒵_t|)`$ and $`d\mu _{S𝒵}C_c^{\mathrm{}}(S𝒵_t,|HS𝒵_t|)`$ to get a smooth density $`R_𝒵\pi _G^{}kd\mu _{S𝒵}`$ on $`S𝒵_t`$. Let $`\sigma _{\mathrm{sub}}(A)`$ denote the subprincipal symbol of $`A`$. Define $`\sigma _{\mathrm{sub}}(P)=\frac{1}{2}a^{\frac{1}{2}}\sigma _{\mathrm{sub}}(A)`$ in some conic neighborhood of $`N^{}`$. The restriction of $`\sigma _{\mathrm{sub}}(P)`$ on $`N^{}`$ is equal to the restriction on $`N^{}`$ of the subprincipal symbol of any operator $`P_1\mathrm{\Psi }^1(M,|TM|^{1/2})`$ such that the complete symbols of $`P_1^2`$ and $`A`$ are equal $`modS^{\mathrm{}}`$ in some neighborhood of $`N^{}`$. ###### Theorem 6. Let $`t`$ be a relative period of the flow $`f_t`$. Assume that the relative fixed point set $`𝒵_t`$ is clean. Then, for any $`kC_c^{\mathrm{}}(G,|T𝒢|^{1/2})`$ and for any $`\tau `$ in some neighborhood of $`t`$, we have $$\theta _k(\tau )=\underset{𝒵_j}{}_{\mathrm{}}^+\mathrm{}\alpha _j(s,k)e^{is(\tau t)}𝑑s,$$ (2.5) where: 1. $`𝒵_j`$ are all connected components of the set $`S𝒵_t`$ in $`G_{SN^{}}`$ of dimensions $`d_j=dim𝒵_j`$; 2. $`\alpha _j`$ has an asymptotic expansion $$\alpha _j(s,k)\left(\frac{s}{2\pi i}\right)^{(d_jp1)/2}i^{\sigma _j}\underset{r=0}{\overset{+\mathrm{}}{}}\alpha _{j,r}(k)s^r,s+\mathrm{}$$ (2.6) with $`\alpha _{j,0}`$ given by the formula $$\alpha _{j,0}(k)=_{𝒵_j}e^{i_0^t\sigma _{\mathrm{sub}}(P)(f_\tau dh_\gamma ^{}(\nu ))𝑑\tau }R_𝒵\pi _G^{}k(\gamma ,\nu )𝑑\mu _{S𝒵_j}(\gamma ,\nu ),$$ (2.7) where $`\sigma _j`$ denotes the Maslov index associated with the connected component $`𝒵_j`$ (see below for the definition). ## 3 Reduction to the case when $`A`$ is elliptic In this section, we will assume that $`A`$ is an operator under the assumptions (A1) and (A2). We will use the classes $`\mathrm{\Psi }^{m,\mathrm{}}(M,,|TM|^{1/2})`$ of transversal pseudodifferential operators (see for the definition) and the Sobolev spaces $`H^s(M)`$ of half-densities on $`M`$. Put also $`\mathrm{\Psi }^,\mathrm{}(M,,|TM|^{1/2})=_m\mathrm{\Psi }^{m,\mathrm{}}(M,,|TM|^{1/2})`$. By , the operator $`P=A^{1/2}`$ satisfies the following conditions: (H1) $`P`$ has the form $$P=P_1+R_1,$$ where: (a) $`P_1\mathrm{\Psi }^1(M,|TM|^{1/2})`$ is a transversally elliptic operator with the positive, holonomy invariant transversal principal symbol such that the complete symbols of $`P_1^2`$ and $`A`$ are equal $`modS^{\mathrm{}}`$ in some neighborhood of $`N^{}`$; (b) $`R_1`$ is a bounded operator from $`L^2(M)`$ to $`H^1(M)`$ and for any $`K\mathrm{\Psi }^,\mathrm{}(M,,|TM|^{1/2})`$ the operator $`KR_1`$ is a smoothing operator in $`L^2(M)`$, that is, it defines a bounded operator from $`L^2(M)`$ to $`C^{\mathrm{}}(M,|TM|^{1/2})`$. (H2) $`P`$ is essentially self-adjoint in $`L^2(M)`$ (with the initial domain $`C^{\mathrm{}}(M,|TM|^{1/2})`$). ###### Lemma 7 (). Any operator $`P`$, satisfying the conditions (H1) and (H2), can be represented in the form $$P=P_2+R_2,$$ (3.1) where: (a) $`P_2\mathrm{\Psi }^1(M,|TM|^{1/2})`$ is an essentially self-adjoint, elliptic operator with the positive principal symbol and the holonomy invariant transversal principal symbol such that the complete symbols of $`P_1`$ and $`P_2`$ are equal $`modS^{\mathrm{}}`$ in some neighborhood of $`N^{}`$; (b) $`R_2`$ is a bounded operator from $`L^2(M)`$ to $`H^1(M)`$ and, for any $`K\mathrm{\Psi }^,\mathrm{}(M,,|TM|^{1/2})`$, the operator $`KR_2`$ is a smoothing operator in $`L^2(M)`$. ###### Proof. Take a foliated coordinate chart $`\mathrm{\Omega }`$ on $`M`$ with coordinates $`(x,y)I^p\times I^q`$ ($`I`$ is the open interval $`(0,1)`$) such that the restriction of $``$ on $`U`$ is given by the sets $`y=\mathrm{const}`$. Let $`p_1S^1(I^n\times ^n)`$ be the complete symbol of the operator $`P_1`$ in this chart. Assume that $`p_1(x,y,\xi ,\eta )`$ is invertible for any $`(x,y,\xi ,\eta )U,|\xi |^2+|\eta |^2>R^2`$, where $`R>0`$, $`U`$ is a conic neighborhood of the set $`\eta =0`$. Take any function $`\varphi C^{\mathrm{}}(I^n\times ^n),\varphi =\varphi (x,y,\xi ,\eta ),xI^p,yI^q,\xi ^p,\eta ^q`$, homogeneous of degree $`0`$ in $`(\xi ,\eta )`$ for $`|\xi |^2+|\eta |^2>1`$, which is supported in some conic neighborhood of $`\eta =0`$ and is equal to $`1`$ in $`U`$, and put $$p_2(x,y,\xi ,\eta )=\varphi p_1(x,y,\xi ,\eta )+(1\varphi )(1+|\xi |^2+|\eta |^2)^{1/2}.$$ Take $`P_2`$ to be the operator $`p_2(x,y,D_x,D_y)`$ with the complete symbol $`p_2`$ (or, more precisely, $`p_2(x,y,D_x,D_y)+p_2(x,y,D_x,D_y)^{}`$ to provide self-adjointness) and put $`R_2=PP_2`$. The operator $`P_1P_2`$ has order $`\mathrm{}`$ in some conic neighbourhood of $`N^{}`$, therefore, for any $`K\mathrm{\Psi }^,\mathrm{}(M,,|TM|^{1/2})`$ the operator $`K(P_1P_2)`$ is a smoothing operator , that completes immediately the proof. ∎ Denote by $`W(t)=e^{itP_2}`$ the wave group generated by the elliptic operator $`P_2`$. It is well-known that $`W(t)`$ is a Fourier integral operator (see below for more details). Put also $`R(t)=e^{itP}W(t)`$. ###### Proposition 8. For any $`K\mathrm{\Psi }^,\mathrm{}(M,,|TM|^{1/2})`$, the family $`KR(t),t,`$ is a smooth family of bounded operators from $`L^2(M)`$ to $`C^{\mathrm{}}(M,|TM|^{1/2})`$. ###### Proof. Since $`P^2=A\mathrm{\Psi }^2(M,|TM|^{1/2})`$, by interpolation and duality, $`P`$ defines a bounded operator from $`H^1(M)`$ to $`L^2(M)`$ and from $`L^2(M)`$ to $`H^1(M)`$ and, for any natural $`N`$, $`P^N`$ defines a bounded operator from $`H^N(M)`$ to $`L^2(M)`$ and from $`L^2(M)`$ to $`H^N(M)`$. Since $`R_2=PP_2`$ and $`P_2\mathrm{\Psi }^1(M,|TM|^{1/2})`$, $`R_2`$ also defines a bounded operator from $`H^1(M)`$ to $`L^2(M)`$ and from $`L^2(M)`$ to $`H^1(M)`$. By assumption, for any $`K\mathrm{\Psi }^,\mathrm{}(M,,|TM|^{1/2})`$, the operator $`KR_2`$ is a smoothing operator, therefore, the operator $`KR_2P^N`$ is defined as an operator from $`H^N(M)`$ to $`C^{\mathrm{}}(M,|TM|^{1/2})`$. ###### Lemma 9. For any $`K\mathrm{\Psi }^,\mathrm{}(M,,|TM|^{1/2})`$ and for any $`N`$, the operator $`KR_2P^N`$ extends to a bounded operator from $`L^2(M)`$ to $`C^{\mathrm{}}(M,|TM|^{1/2})`$. ###### Proof. We will prove the lemma by induction on $`N`$. For $`N=0`$, the statement is true by assumption. Let us assume that it is true for some $`N`$, that is, for any $`K\mathrm{\Psi }^,\mathrm{}(M,,|TM|^{1/2})`$, the operator $`KR_2P^N`$ extends to a bounded operator from $`L^2(M)`$ to $`C^{\mathrm{}}(M,|TM|^{1/2})`$. We have the equality $`P^2P_2^2=R_2P+P_2R_2`$ as operators from $`H^1(M)`$ to $`H^1(M)`$, therefore, for any $`K\mathrm{\Psi }^,\mathrm{}(M,,|TM|^{1/2})`$, $$KR_2P^{N+1}=K(R_2P)P^N=K(P^2P_2^2)P^NKP_2R_2P^N.$$ The operator $`P^2P_2^2\mathrm{\Psi }^2(M,|TM|^{1/2})`$ has order $`\mathrm{}`$ in some conic neighbourhood of $`N^{}`$, therefore, for any $`K\mathrm{\Psi }^,\mathrm{}(M,,|TM|^{1/2})`$ the operator $`K(P^2P_2^2)`$ extends to a bounded operator from $`H^s(M)`$ to $`C^{\mathrm{}}(M,|TM|^{1/2})`$ for any $`s`$ and $`K(P^2P_2^2)P^N`$ extends to a bounded operator from $`L^2(M)`$ to $`C^{\mathrm{}}(M,|TM|^{1/2})`$. Since $`P_2\mathrm{\Psi }^1(M,|TM|^{1/2})`$ and $`K\mathrm{\Psi }^,\mathrm{}(M,,|TM|^{1/2})`$, by the composition theorem , $`KP_2\mathrm{\Psi }^,\mathrm{}(M,,|TM|^{1/2})`$ and, by induction hypothesis, $`KP_2R_2P^N`$ extends to a bounded operator from $`L^2(M)`$ to $`C^{\mathrm{}}(M,|TM|^{1/2})`$. ∎ By the Duhamel formula, we have $$R(t)u=i_0^te^{i\tau P_2}R_2e^{i(t\tau )P}u𝑑\tau ,uH^1(M)D(P),$$ therefore, for any $`K\mathrm{\Psi }^,\mathrm{}(M,,|TM|^{1/2})`$, $$KR(t)=i_0^tKe^{i\tau P_2}R_2e^{i(t\tau )P}𝑑\tau =i_0^te^{i\tau P_2}e^{i\tau P_2}Ke^{i\tau P_2}R_2e^{i(t\tau )P}𝑑\tau .$$ Any operator $`K\mathrm{\Psi }^,\mathrm{}(M,,|TM|^{1/2})`$ is a Fourier integral operator (see below for more details) and, using the composition theorem for Fourier integral operators, one can check that $`e^{i\tau P_2}Ke^{i\tau P_2}\mathrm{\Psi }^,\mathrm{}(M,,|TM|^{1/2})`$. Therefore, the operator $`e^{i\tau P_2}Ke^{i\tau P_2}R_2`$ extends to a bounded operator from $`L^2(M)`$ to $`C^{\mathrm{}}(M,|TM|^{1/2})`$. Since $`e^{i\tau P_2}`$ maps $`C^{\mathrm{}}(M,|TM|^{1/2})`$ to $`C^{\mathrm{}}(M,|TM|^{1/2})`$ and, by the spectral theorem, $`e^{i(t\tau )P}`$ is a bounded operator in $`L^2(M)`$, for any $`K\mathrm{\Psi }^,\mathrm{}(M,,|TM|^{1/2})`$, the operator $`KR(t)`$ extends to a bounded operator from $`L^2(M)`$ to $`C^{\mathrm{}}(M,|TM|^{1/2})`$. Moreover, one can be easily seen from above arguments that the function $`KR(t)`$ is continuous as a function on $``$ with values in the space $`(L^2(M),C^{\mathrm{}}(M,|TM|^{1/2}))`$ of bounded operators from $`L^2(M)`$ to $`C^{\mathrm{}}(M,|TM|^{1/2})`$. For any $`uH^1(M)`$, the function $`H:tKR(t)u`$ is differentiable, and $$\frac{d}{dt}KR(t)u=iK(Pe^{itP}uP_2e^{itP_2}u)=i(KP_2R(t)+KR_2e^{itP})u.$$ The operator $`KP_2R(t)+KR_2e^{itP}`$ extends to a bounded operator from $`L^2(M)`$ to $`C^{\mathrm{}}(M,|TM|^{1/2})`$, and, moreover, the function $`tKP_2R(t)+KR_2e^{itP}`$ is a continuous function on $``$ with values in $`(L^2(M),C^{\mathrm{}}(M,|TM|^{1/2}))`$. Using this, one can be easily seen that the function $`tKR(t)`$ is differentiable as a function on $``$ with values in $`(L^2(M),C^{\mathrm{}}(M,|TM|^{1/2}))`$ and $$\frac{d}{dt}KR(t)=i(KP_2R(t)+KR_2e^{itP}).$$ Let us proceed by induction. Assume that, for any $`K\mathrm{\Psi }^,\mathrm{}(M,,|TM|^{1/2})`$ and for some natural $`n`$, the function $`KR(t)`$ is $`n`$-times differentiable as a function on $``$ with values in $`(L^2(M),C^{\mathrm{}}(M,|TM|^{1/2}))`$ and the derivative $`KR^{(n)}(t),t`$ satisfies the equation $$KR^{(n)}(t)=iKP_2R^{(n1)}(t)+i^nKR_2P^{n1}e^{itP}.$$ (3.2) To prove that the function $`tKR^{(n)}(t)`$ is differentiable as a function on $``$ with values in $`(L^2(M),C^{\mathrm{}}(M,|TM|^{1/2}))`$, as above, it suffices to prove that the derivative $`(d/dt)KR^{(n)}(t)u`$ exists for any $`u`$ from a dense subspace of $`L^2(M)`$, it extends to a bounded operator from $`L^2(M)`$ to $`C^{\mathrm{}}(M,|TM|^{1/2})`$, and its extension is continuous as a function on $``$ with values in $`(L^2(M),C^{\mathrm{}}(M,|TM|^{1/2}))`$. From (3.2), one can easy to see that the derivative $`(d/dt)KR^{(n)}(t)u`$ exists for any $`uH^1(M)`$ and satisfies the equation $$\frac{d}{dt}KR^{(n)}(t)u=iKP_2R^{(n)}(t)u+i^{n+1}KR_2P^ne^{itP}u.$$ (3.3) The first term in the right-hand side of (3.3), $`iKP_2R^{(n)}(t)`$, is a bounded operator from $`L^2(M)`$ to $`C^{\mathrm{}}(M,|TM|^{1/2})`$ by the induction hypothesis. By Lemma 9, the operator $`KR_2P^n`$ extends to a bounded operator from $`L^2(M)`$ to $`C^{\mathrm{}}(M,|TM|^{1/2})`$, and, by the spectral theorem, $`e^{itP}`$ is a bounded operator in $`L^2(M)`$, therefore, the second term in the right-hand side of (3.3), the operator $`i^{n+1}KR_2P^ne^{itP}`$, extends to a bounded operator from $`L^2(M)`$ to $`C^{\mathrm{}}(M,|TM|^{1/2})`$. It is also clear the right-hand side of (3.3) is continuous as a function on $``$ with values in $`(L^2(M),C^{\mathrm{}}(M,|TM|^{1/2}))`$. This completes the proof of the existence of the derivative $`KR^{(n+1)}(t)=(d/dt)KR^{(n)}(t)`$ and the induction arguments. ∎ ###### Proof of Proposition 2. Let $`W(t)`$ and $`R(t)`$ be as in Proposition 8 and $`kC_c^{\mathrm{}}(G,|T𝒢|^{1/2})`$. Define $`\theta _k(t)`$ by the formula $$\theta _k(t)=\mathrm{tr}R(k)W(t)+\mathrm{tr}R(k)R(t).$$ Since $`P_2`$ is an elliptic operator, the operator $`f(t)e^{itP_2}𝑑t`$ is a smoothing operator in $`𝒟^{}(M)`$ and the trace of the operator $`R(k)W(t)`$ is well-defined as a distribution on $``$ . Since any bounded operator $`T`$ in $`L^2(M)`$, which extends to a bounded operator from $`L^2(M)`$ to $`H^s(M)`$ with $`s>n=dimM`$, is a trace class operator, the trace $`\mathrm{tr}R(k)R(t)`$ is a well-defined smooth function on $``$ by Proposition 8. ∎ ###### Corollary 10. For any $`kC_c^{\mathrm{}}(G,|T𝒢|^{1/2})`$, the function $`\mathrm{tr}R(k)R(t)=\theta _k(t)\mathrm{tr}R(k)W(t)`$ is a smooth function on $``$. It should be noted that, without any additional assumption about the operator $`P`$ in question, the corresponding distribution on $`G__N\times `$, $`k\mathrm{tr}R(k)R(t)`$, might be very singular, but the singularities of the distribution $`k\mathrm{tr}R(k)W(t)`$ can be described rather explicitly under the clear intersection assumption. ## 4 The case of an elliptic operator Let $`P_2\mathrm{\Psi }^1(M,|TM|^{1/2})`$ be an essentially self-adjoint, elliptic operator with the positive principal symbol and the holonomy invariant transversal principal symbol and $`W(t)=e^{itP_2}`$. The singularities of the distribution $`t\mathrm{tr}R(k)W(t)`$ can be studied in a standard manner, using microlocal analysis. Fix $`kC_c^{\mathrm{}}(G,|T𝒢|^{1/2})`$. We will consider the operator family $`R(k)W(t)`$ as a single operator $`R(k)W`$ from $`L^2(M)`$ to $`L^2(\times M)`$. We will prove that this operator is a Fourier integral operator. At first, let us recall well-known facts about the structure of the operators $`R(k)`$ and $`W`$. As above, let $`\stackrel{~}{p}`$ be a smooth function on $`\stackrel{~}{T}^{}M`$ homogeneous of degree one such that $`\stackrel{~}{p}(\xi )0`$ for $`\xi \stackrel{~}{T}^{}M`$, which is equal to $`p=a^{1/2}`$ in some conic neighborhood of $`N^{}`$ and $`\stackrel{~}{f}_t`$ the Hamiltonian flow of $`\stackrel{~}{p}`$. Without loss of generality, we may assume that $`\stackrel{~}{p}`$ is the principal symbol of the operator $`P_2`$. Let $`\mathrm{\Lambda }_{\stackrel{~}{p}}`$ be the Lagrangian submanifold in $`\stackrel{~}{T}^{}\times \stackrel{~}{T}^{}M\times \stackrel{~}{T}^{}M`$: $$\mathrm{\Lambda }_{\stackrel{~}{p}}=\{((t,\tau ),(x,\xi ),(y,\eta ))\stackrel{~}{T}^{}\times \stackrel{~}{T}^{}M\times \stackrel{~}{T}^{}M:\tau =\stackrel{~}{p}(x,\xi ),(x,\xi )=\stackrel{~}{f}_t(y,\eta )\}.$$ Then $`W`$ is a Fourier integral operator associated with the canonical relation $`\mathrm{\Lambda }_{\stackrel{~}{p}}^{}`$, $`WI^{1/4}(\times M\times M,\mathrm{\Lambda }_{\stackrel{~}{p}}^{})`$. The operator $`R(k)`$ belongs to $`\mathrm{\Psi }^{0,\mathrm{}}(M,,|TM|^{1/2})`$ and, therefore, is a Fourier integral operator associated with an immersed canonical relation, which is the image of $`G__N`$ under the mapping $$(r_N,s_N):G__NT^{}M\times T^{}M,(\gamma ,\nu )(\nu ,dh_\gamma ^{}(\nu )),$$ given by the source and the target mappings of the groupoid $`G__N`$ (see ). More precisely, $`R(k)I^{p/2}(M\times M,G__N^{})`$. By the tranversal ellipticity of $`\stackrel{~}{p}`$, the intersection of $`\mathrm{\Lambda }_{\stackrel{~}{p}}^{}`$ with $`G__N^{}`$ is transverse, and, by the composition theorem of Fourier integral operators , the operator $`R(k)W`$ is a Fourier integral operator associated with an immersed canonical relation from $`T^{}M`$ to $`T^{}(\times M)`$ given by the map $$\mathrm{\Pi }:\times G__NT^{}\times T^{}M\times T^{}M,(t,\gamma ,\nu )(t,p(\nu ),\nu ,f_tdh_\gamma ^{}(\nu )).$$ More precisely, $`R(k)WI^{p/21/4}(\times M\times M;\times G__N,\mathrm{\Pi }).`$ Recall that the trace functional can be treated from the point of microlocal analysis as follows . Let $`\mathrm{\Delta }:\times M\times M\times M`$ be the diagonal map, $`\mathrm{\Delta }(t,x)=(t,x,x),(t,x)\times M`$, and $`\pi :\times MM`$ the projection map. Then $$\text{tr}R(k)W=\pi _{}\mathrm{\Delta }^{}W_k,$$ (4.1) where $`W_kC^{\mathrm{}}(\times M\times M,|T(\times M\times M)|^{1/2})`$ is the Schwartz kernel of the operator $`R(k)W`$, $`\mathrm{\Delta }^{}:C^{\mathrm{}}(\times M\times M,|T(\times M\times M)|^{1/2})C^{\mathrm{}}(,|T|^{1/2})C^{\mathrm{}}(M,|TM|)`$ is defined by the formula $$\mathrm{\Delta }^{}(s_1s_2s_3)(t,x)=s_1(t)(s_2(x)s_3(x)),t,xM,$$ where $`s_1C^{\mathrm{}}(,|T|^{1/2})`$, $`s_2C^{\mathrm{}}(M,|TM|^{1/2})`$, $`s_3C^{\mathrm{}}(M,|TM|^{1/2})`$, and $$\pi _{}:C^{\mathrm{}}(,|T|^{1/2})C^{\mathrm{}}(M,|TM|)C^{\mathrm{}}(,|T|^{1/2})$$ is given by integration along fibers of the projection $`\pi `$. It is known that $`\pi _{}\mathrm{\Delta }^{}I^0(\times M\times M\times ,\mathrm{\Gamma })`$, where $`\mathrm{\Gamma }`$ is the conormal bundle to the diagonal in $`\times M\times M\times `$: $$\mathrm{\Gamma }=\{(t,\tau _1,\nu _1,\nu _2,t,\tau _2)T^{}\times T^{}M\times T^{}M\times T^{}:\nu _1=\nu _2,\tau _1=\tau _2\}.$$ There is a commutative diagram $$\begin{array}{ccc}\mathrm{\Gamma }& \stackrel{p_1}{}& 𝒵\\ \phi & & p_2& & \\ T^{}(\times M\times M)& \stackrel{\mathrm{\Pi }}{}& \times G__N\end{array}$$ (4.2) where $$p_1(t,\gamma ,\nu )=(\mathrm{\Pi }(t,\gamma ,\nu ),t,p(\nu ))=(t,p(\nu ),\nu ,\nu ,t,p(\nu )),(t,\gamma ,\nu )𝒵,$$ $`p_2`$ is a natural inclusion and $$\phi (t,\tau ,\nu ,\nu ,t,\tau )=(t,\tau ,\nu ,\nu ),(t,\tau ,\nu ,\nu ,t,\tau )\mathrm{\Gamma }.$$ It is easy to see that (4.2) is a fiber product diagram, that is, $$𝒵\{(x,y)(\times G__N)\times \mathrm{\Gamma }:\mathrm{\Pi }(x)=\phi (y)\}.$$ Using this fact and the functoriality properties of the wave-front sets (see, for instance, ), one get immediately the description of the singularities of the distribution $`\theta _k`$, given by Theorem 3. To finish the proof of Theorem 6, we will state under the assumption on the flow $`f_t`$ to be clean in the sense of Definition 5 that $`\pi _{}\mathrm{\Delta }^{}W_k`$ is a Lagrangian distribution and compute its symbol. We begin with computation of the symbol of the operator $`R(k)W`$. Recall first the description of the principal symbol of the operator $`R(k)`$. According to the short exact sequence $$0T𝒢__NTG__NH𝒢__N0,$$ the half-density vector bundle on $`G__N`$ can be decomposed as $$|TG__N|^{1/2}|T𝒢__N|^{1/2}|H𝒢__N|^{1/2},$$ where $`|H𝒢__N|^{1/2}`$ is the transverse half-density bundle on $`G__N`$: $`|H_{(\gamma ,\nu )}𝒢__N|^{1/2}|H_\nu _N|^{1/2}|H_{dh_\gamma ^{}(\nu )}_N|^{1/2}.`$ Let $`|dyd\eta |^{1/2}C^{\mathrm{}}(N^{},|H_N|^{1/2})`$ be given by the Liouville form of the canonical transverse symplectic structure on the foliated manifold $`(N^{},_N)`$, and $`r_N^{}(|dyd\eta |^{1/2})C^{\mathrm{}}(G__N,|H𝒢__N|^{1/2})`$ its pull back via the map $`r_N:G__NN^{}`$. Recall that the space $`S^m(G__N,|TG__N|^{1/2})`$ is defined to be the space of all smooth sections $`s`$ of the vector bundle $`|TG__N|^{1/2}`$ on $`G__N`$ homogeneous of degree $`m`$ such that $`\pi _G(\mathrm{supp}s)`$ is compact in $`G_{}`$. The half-density principal symbol of $`R(k)`$ is an element of $`S^0(G__N,|TG__N|^{1/2})`$ given by the formula $$\sigma (R(k))(\gamma ,\nu )=\pi _G^{}k(\gamma ,\nu )r_N^{}(|dyd\eta |^{1/2}),(\gamma ,\nu )G__N.$$ The Maslov bundle $`M(\times G__N,\mathrm{\Pi })`$ of the immersed canonical relation $`(\times G__N,\mathrm{\Pi })`$ restricted to $`t=0`$ is isomorphic to the Maslov bundle $`M(G__N)`$ of $`G__N`$, therefore, it has a canonical constant section, which extends to a global section $`s`$ of $`M(\times G__N,\mathrm{\Pi })`$ by requiring it to be constant along each bicharacteristic $`(t,\tau ,\nu _1,\nu _2),\nu _1=f_t(\nu _2),t.`$ Using the description of the principal symbol of the operator $`W(t)`$ given, for instance, in and the composition theorem of Fourier integral operators, we get immediately that the principal symbol of the operator $`R(k)W`$ is an element of $`S^0(\times G__N,M(\times G__N,\mathrm{\Pi })|T(\times G__N)|^{1/2})`$, whose value at a point $`(t,\gamma ,\nu )\times G__N`$ is given by $$\sigma (R(k)W)(t,\gamma ,\nu )=e^{i_0^t\sigma _{\mathrm{sub}}(P)(f_sdh_\gamma ^{}(\nu ))𝑑s}s|dt|^{1/2}\pi _G^{}k(\gamma ,\nu )r_N^{}(|dyd\eta |^{1/2}).$$ Now let us turn to the composition (4.1). First, we check the corresponding cleanness assumption. ###### Lemma 11. The assumption on the flow $`f_t`$ to be clean on $`𝒵_t`$ in the sense of Definition 5 guarantees that the composition of $`\times G__N`$ with $`\mathrm{\Gamma }`$ is clean. ###### Proof. By definition, the composition of $`\times G__N`$ with $`\mathrm{\Gamma }`$ is clean iff $`𝒵_t`$ is a submanifold of $`\times G__N`$ and in addition the fiber product diagram (4.2) is clean at any point $`(t,\gamma ,\nu )𝒵`$, that is, the linearized diagram $$\begin{array}{ccc}T_{p_1(t,\gamma ,\nu )}\mathrm{\Gamma }& \stackrel{dp_1}{}& T_{(t,\gamma ,\nu )}𝒵\\ d\phi & & dp_2& & \\ T_{(t,p(\nu ),\nu ,\nu )}(T^{}(\times M\times M))& \stackrel{d\mathrm{\Pi }}{}& T_{(t,\gamma ,\nu )}(\times G__N)\end{array}$$ (4.3) is a fiber product diagram. Since $`T_{(t,\gamma ,\nu )}𝒵`$ is always contained in $`T_\nu N^{}T_{dh_\gamma ^{}(\nu )}N^{}`$, this is true iff the diagram $$\begin{array}{ccc}T_{p_1(t,\gamma ,\nu )}(\mathrm{\Gamma }T^{}\times N^{}\times N^{}\times T^{})& \stackrel{dp_1}{}& T_{(t,\gamma ,\nu )}𝒵\\ d\phi & & dp_2& & \\ T_{(t,p(\nu ),\nu ,\nu )}(T^{}\times N^{}\times N^{})& \stackrel{d\mathrm{\Pi }}{}& T_{(t,\gamma ,\nu )}(\times G__N)\end{array}$$ (4.4) is a fiber product diagram. The diagram (4.4) has a subdiagram $$\begin{array}{ccc}L_{p_1(t,\gamma ,\nu )}\mathrm{\Gamma }& \stackrel{dp_1}{}& L_{(t,\gamma ,\nu )}𝒵\\ d\phi & & dp_2& & \\ 0T_\nu _NT_\nu _N& \stackrel{d\mathrm{\Pi }}{}& L_{(t,\gamma ,\nu )}(\times G__N)\end{array}$$ (4.5) where $`L_{(t,\gamma ,\nu )}(\times G__N)T_{(t,\gamma ,\nu )}(\times G__N)`$ is given by $$\begin{array}{c}L_{(t,\gamma ,\nu )}(\times G__N)\hfill \\ \hfill =\{(U,V_1,V_2,W)T_t()T_\nu _NT_{dh_\gamma ^{}(\nu )}_NH_{(\gamma ,\nu )}𝒢__N:U=0,W=0\}\\ \hfill T_\nu _NT_{dh_\gamma ^{}(\nu )}_N\end{array}$$ $`L_{(t,\gamma ,\nu )}𝒵T_{(t,\gamma ,\nu )}𝒵`$ by $$\begin{array}{c}L_{(t,\gamma ,\nu )}𝒵=\{(U,V_1,V_2,W)T_t()T_\nu _NT_{dh_\gamma ^{}(\nu )}_NH_{(\gamma ,\nu )}𝒢__N\hfill \\ \hfill :U=0,V_1=df_t(dh_\gamma ^{}(\nu ))(V_2),W=0\}T_\nu _N\end{array}$$ and $`L_{p_1(t,\gamma ,\nu )}\mathrm{\Gamma }T_{p_1(t,\gamma ,\nu )}\mathrm{\Gamma }`$ by $$\begin{array}{c}L_{p_1(t,\gamma ,\nu )}\mathrm{\Gamma }\{(U_1,V_1,V_2,U_2)T_{(t,p(\nu ))}(T^{})T_\nu _NT_\nu _NT_{(t,p(\nu ))}(T^{})\hfill \\ \hfill :U_1=U_2=0,V_1=V_2\}\end{array}$$ which can be easily seen to be a fiber diagram. Therefore, the diagram (4.4) is a fiber product diagram iff the quotient diagram is a fiber product diagram: $$\begin{array}{ccc}H_{p_1(t,\gamma ,\nu )}\mathrm{\Gamma }& \stackrel{dp_1}{}& H_{(t,\gamma ,\nu )}𝒵\\ d\phi & & dp_2& & \\ T_{(t,p(\nu ))}(T^{})H_\nu _NH_\nu _N& \stackrel{d\mathrm{\Pi }}{}& H_{(t,\gamma ,\nu )}(\times G__N)\end{array}$$ (4.6) where $$\begin{array}{c}H_{(t,\gamma ,\nu )}(\times G__N)\hfill \\ \hfill =\{(U,V_1,V_2,W)T_t()T_\nu _NT_{dh_\gamma ^{}(\nu )}_NH_{(\gamma ,\nu )}𝒢__N:V_1=0,V_2=0\}\\ \hfill T_t()H_{(\gamma ,\nu )}𝒢__N,\end{array}$$ $$H_{(t,\gamma ,\nu )}𝒵=T_{(t,\gamma ,\nu )}𝒵/L_{(t,\gamma ,\nu )}𝒵T_\nu Z_t/T_\nu _N,$$ $$\begin{array}{c}H_{p_1(t,\gamma ,\nu )}\mathrm{\Gamma }\{(U_1,V_1,V_2,U_2)T_{(t,p(\nu ))}(T^{})H_\nu _NH_\nu _NT_{(t,p(\nu ))}(T^{})\hfill \\ \hfill :U_1=U_2,V_1=V_2\}.\end{array}$$ In its turn, the diagram (4.6) is a fiber product diagram iff the flow $`f_t`$ is clean on $`𝒵_t`$ in the sense of Definition 5. ∎ For any connected component $`𝒵_j`$ of $`𝒵_t`$, the excess of the clean diagram (4.2) equals $`d_j`$, the dimension of the relative fixed point set $`S𝒵_j`$ in $`G_{SN^{}}`$. By the composition theorem of Fourier integral operators, $`\theta _k`$ belongs to $`_jI^{\frac{d_jp}{2}\frac{1}{4}}(\mathrm{\Lambda }_t)`$, where $`\mathrm{\Lambda }_t=\{(t,\tau )T^{}:\tau _{}\}`$, that proves the desired representation of $`\theta _k`$ in the form (2.5) and the existence of the asymptotic expansion (2.6) for $`\alpha _j(s,k)`$.<sup>1</sup><sup>1</sup>1Note that the appearance of the term $`p/2`$ in the exponent is due to the fact $`R(k)WI^{p/21/4}(M\times M;\times G__N,\mathrm{\Pi })`$. To obtain the explicit formula for the leading coefficients $`\alpha _{j,0}`$, we compute the principal symbol of $`\theta _k`$, $`\sigma (\theta _k)`$, following the arguments in . Fix a connected component $`𝒵_j`$ of $`𝒵_t`$ and $`(t,\gamma ,\nu )𝒵_j`$. The fiber product diagram (4.3) defines a composition map $$:|T_{p_1(t,\gamma ,\nu )}\mathrm{\Gamma }|^{1/2}|T_{(t,\gamma ,\nu )}(\times G__N)|^{1/2}|T_{(t,p(\nu ))}(\mathrm{\Lambda }_t)|^{1/2}|T_{(t,\gamma ,\nu )}𝒵_j|$$ (4.7) and, due to (4.1) and the composition theorem for Fourier integral operators, the value of the principal symbol $`\sigma (\theta _k)|T\mathrm{\Lambda }_t|^{1/2}`$ at a point $`(t,\tau )\mathrm{\Lambda }_t`$ is given by integration over $`𝒵_j`$ of $`\sigma (\pi _{}\mathrm{\Delta }^{})\sigma (W_k)C^{\mathrm{}}(\mathrm{\Lambda }_t\times 𝒵_j,|T\mathrm{\Lambda }_t|^{1/2}|T𝒵_j|)`$. Using the functoriality of the $``$ operation on half densities with respect to reduction (cf. \[13, proof of Lemma 4.6\]), the computation of the $``$-product $`\sigma (\pi _{}\mathrm{\Delta }^{})\sigma (W_k)`$ can be reduced to the computation of a $``$-product defined by the transverse fiber product diagram (4.6): $$_t:|H_{p_1(t,\gamma ,\nu )}\mathrm{\Gamma }|^{1/2}|H_{(t,\gamma ,\nu )}(\times G__N)|^{1/2}|T_{(t,p(\nu ))}\mathrm{\Lambda }_t|^{1/2}|H_{(t,\gamma ,\nu )}𝒵_j|.$$ More precisely, we apply the result stated in \[13, proof of Lemma 4.6\] with a symplectic vector space $`𝒱=T_{(t,p(\nu ),\nu ,\nu ,t,p(\nu ))}(T^{}\times T^{}M\times T^{}M\times T^{})`$, two Lagrangian subspaces in $`𝒱`$: $`\mathrm{\Lambda }_1=T_{p_1(t,\gamma ,\nu )}\mathrm{\Gamma }`$ and $`\mathrm{\Lambda }_2`$, which is the image of $`T_{(t,\gamma ,\nu )}(\times G__N)`$ in $`𝒱`$ and the reduction given by the coisotropic subspace $`\mathrm{\Gamma }=T_{(t,p(\nu ),\nu ,\nu ,t,p(\nu ))}(T^{}\times N^{}\times N^{}\times T^{}).`$ By this result, the leafwise component of $`\sigma (\theta _k)`$, $`\sigma _l(\theta _k)|L_{(t,\gamma ,\nu )}𝒵_j|`$, is obtained from the leafwise component of $`\sigma (R(k)W)`$: $$\sigma _l(R(k)W)=e^{i_0^t\sigma _{\mathrm{sub}}(P)(f_sdh_\gamma ^{}(\nu ))𝑑s}\pi _G^{}k$$ by application of the restriction map $`R_𝒵`$: $$\sigma _l(\theta _k)=e^{i_0^t\sigma _{\mathrm{sub}}(P)(f_sdh_\gamma ^{}(\nu ))𝑑s}R_𝒵\pi _G^{}k,$$ and the transversal component of $`\sigma (\theta _k)`$, $`\sigma _t(\theta _k)|T_{(t,p(\nu ))}(\mathrm{\Lambda }_t)|^{1/2}|H_{(t,\gamma ,\nu )}𝒵_j|`$, is equal to the $`_t`$-product of the transverse component of $`\sigma (R(k)W)`$, $$\sigma _t(R(k)W)=|dt|^{1/2}r_N^{}(|dyd\eta |^{1/2}),$$ and the transverse component of $`\sigma (\pi _{}\mathrm{\Delta }^{})`$. By , we get $$\sigma _t(\theta _k)=|dt|^{1/2}d\mu _{𝒵_j}$$ that completes the calculation of the half-density principal symbol of $`\theta _k`$ and implies the formula (2.7) for the leading coefficients $`\alpha _{j,0}`$ as in . ## 5 Maslov indices In this section, we define the Maslov factors $`\sigma `$, corresponding to $`(\gamma ,\nu )𝒵_t`$. For this goal, we will use local coordinates on the holonomy groupoid $`G`$ described, for instance, in . Choose a pair of compatible foliated charts near the points $`\pi (\nu )`$ and $`\pi (f_t(\nu ))`$ with the coordinates $`(x,y)`$ and $`(x^{},y)`$, corresponding to $`\gamma G`$. Then we have the corresponding coordinates in $`H_\nu _N=T_\nu N^{}/T_\nu _N`$ defined as $`(\delta y,\delta \eta )`$, and the vertical and horizontal subspaces $`V_\nu `$ and $`H_\nu `$, given by the equations $`\delta y=0`$ and $`\delta \eta =0`$ accordingly. The linear holonomy map $`dH_{(\gamma ,\nu )}`$ of the foliation $`(N^{},_N)`$ defines an isomorphism of the symplectic spaces $`H_{f_t(\nu )}_N`$ and $`H_\nu _N`$, which preserves the vertical and horizontal subspaces. Due to this isomorphism, we can obtain a closed curve $`\omega _{(\gamma ,\nu )}`$ in the Lagrangian Grassmannian $`𝒢`$ of the symplectic space $`H_\nu _N`$, pulling back, via $`df_t`$, the vertical subspace at $`f_t(\nu )`$ for $`t`$ between $`0`$ and $`T`$. Denote by $`\kappa _{(\gamma ,\nu )}`$ the intersection number of $`\omega `$ with the horizontal subspace $`H_\nu `$: $$\kappa _{(\gamma ,\nu )}=[\omega _{(\gamma ,\nu )}:H_\nu ].$$ Let $`\chi (t,x,y,\xi ,\eta )`$ be the generating function of the canonical transformation $`f_t`$ in the chosen coordinates. Recall that $`\chi `$ is the solution of the Cauchy problem $$d_t\chi =p(x,y,d_x\chi ,d_y\chi ),\chi (0,x,y,\xi ,\eta )=x\xi +y\eta .$$ (5.1) By the holonomy invariance of $`p`$, it can be easily seen that $`\chi (t,x,y,0,\eta )`$ is independent of $`x`$ and $`\xi `$: $`\chi (t,x,y,0,\eta )=\chi (t,y,\eta )`$. Let $$R_{(\gamma ,\nu )}=\left[\begin{array}{ccc}d_{yy}^2\chi & d_{y\eta }^2\chi & 1\\ d_{\eta y}^2\chi & d_{\eta \eta }^2\chi & 0\\ 1& 0& 0\end{array}\right].$$ We define the Maslov factor $`\sigma (\gamma ,\nu )`$ as $$\sigma (\gamma ,\nu )=\mathrm{sgn}R_{(\gamma ,\nu )}+2\kappa _{(\gamma ,\nu )},(\gamma ,\nu )𝒵.$$ It is clear that $`\sigma (\gamma ,\nu )`$ is a locally constant function on $`𝒵`$. To handle with Maslov factors in the proof of Theorem 6, let us write the Schwartz kernel of the operator $`R(k)`$ in a foliated coordinate chart on $`G`$ given by a pair of compatible coordinate systems as $`k(x,x_1,y)\delta (yy_1)`$, and the Schwartz kernel $`W(t)`$ microlocally as an oscillatory integral of the form $$e^{i\alpha (t,x_1,y_1,x_2,y_2,\xi ,\eta )}a(t,x_1,y_1,x_2,y_2,\xi ,\eta )𝑑\xi 𝑑\eta ,$$ where $$\alpha (t,x_1,y_1,x_2,y_2,\xi ,\eta )=\chi (t,x_1,y_1,\xi ,\eta )x_2\xi y_2\eta $$ and $`\chi (t,x_1,y_1,\xi ,\eta )`$ is the generating function of the canonical transformation $`f_t`$ given by (5.1). Then the Schwartz kernel of the operator $`R(k)W`$ is given by the formula $$e^{i\alpha (t,x_1,y_1,x_2,y_2,\xi ,\eta )}k(x,x_1,y_1)a(t,x_1,y_1,x_2,y_2,\xi ,\eta )𝑑x_1𝑑\xi 𝑑\eta ,$$ from where one can easily derive the desired assertion, following the arguments of .
warning/0001/cs0001022.html
ar5iv
text
# Recognition Performance of a Structured Language ModelThis work was funded by the NSF IRI-19618874 grant STIMULATE ## 1. Introduction The main goal of the present work is to develop and evaluate a language model that uses syntactic structure to model long-distance dependencies. The model we present is closely related to the one investigated in , however different in a few important aspects: $``$ our model operates in a left-to-right manner, allowing the decoding of word lattices, as opposed to the one referred to previously, where only whole sentences could be processed, thus reducing its applicability to N-best list re-scoring; the syntactic structure is developed as a model component; $``$ our model is a factored version of the one in , thus enabling the calculation of the joint probability of words and parse structure; this was not possible in the previous case due to the huge computational complexity of that model. The structured language model (SLM), its probabilistic parameterization and performance in a two-pass speech recognizer — we evaluate the model in a lattice decoding framework — are presented. Experiments on the SWITCHBOARD corpus show an improvement in both perplexity (PPL) and word error rate (WER) over conventional trigram models. ## 2. Structured Language Model An extensive presentation of the SLM can be found in . The model assigns a probability $`P(W,T)`$ to every sentence $`W`$ and its every possible binary parse $`T`$. The terminals of $`T`$ are the words of $`W`$ with POStags, and the nodes of $`T`$ are annotated with phrase headwords and non-terminal labels. Let $`W`$ be a sentence of length $`n`$ words to which we have prepended `<s>` and appended `</s>` so that $`w_0=`$`<s>` and $`w_{n+1}=`$`</s>`. Let $`W_k`$ be the word k-prefix $`w_0\mathrm{}w_k`$ of the sentence and $`W_kT_k`$ the *word-parse k-prefix*. Figure 1 shows a word-parse k-prefix; `h_0 .. h_{-m}` are the *exposed heads*, each head being a pair(headword, non-terminal label), or (word, POStag) in the case of a root-only tree. ### 2.1. Probabilistic Model The probability $`P(W,T)`$ of a word sequence $`W`$ and a complete parse $`T`$ can be broken into: $`P(W,T)=`$ (1) $`{\displaystyle \underset{k=1}{\overset{n+1}{}}}[`$ $`P(w_k/W_{k1}T_{k1})P(t_k/W_{k1}T_{k1},w_k)`$ $`{\displaystyle \underset{i=1}{\overset{N_k}{}}}P(p_i^k/W_{k1}T_{k1},w_k,t_k,p_1^k\mathrm{}p_{i1}^k)]`$ where: $``$ $`W_{k1}T_{k1}`$ is the word-parse $`(k1)`$-prefix $``$ $`w_k`$ is the word predicted by WORD-PREDICTOR $``$ $`t_k`$ is the tag assigned to $`w_k`$ by the TAGGER $``$ $`N_k1`$ is the number of operations the PARSER executes at sentence position $`k`$ before passing control to the WORD-PREDICTOR (the $`N_k`$-th operation at position k is the `null` transition); $`N_k`$ is a function of $`T`$ $``$ $`p_i^k`$ denotes the i-th PARSER operation carried out at position k in the word string; the operations performed by the PARSER are illustrated in Figures 2-3 and they ensure that all possible binary branching parses with all possible headword and non-terminal label assignments for the $`w_1\mathrm{}w_k`$ word sequence can be generated. Our model is based on three probabilities, estimated using deleted interpolation , parameterized as follows: $`P(w_k/W_{k1}T_{k1})`$ $`=`$ $`P(w_k/h_0,h_1)`$ (2) $`P(t_k/w_k,W_{k1}T_{k1})`$ $`=`$ $`P(t_k/w_k,h_0.tag,h_1.tag)`$ (3) $`P(p_i^k/W_kT_k)`$ $`=`$ $`P(p_i^k/h_0,h_1)`$ (4) It is worth noting that if the binary branching structure developed by the parser were always right-branching and we mapped the POStag and non-terminal label vocabularies to a single type then our model would be equivalent to a trigram language model. Since the number of parses for a given word prefix $`W_k`$ grows exponentially with $`k`$, $`|\{T_k\}|O(2^k)`$, the state space of our model is huge even for relatively short sentences so we had to use a search strategy that prunes it. Our choice was a synchronous multi-stack search algorithm which is very similar to a beam search. The probability assignment for the word at position $`k+1`$ in the input sentence is made using: $`P_{SLM}(w_{k+1}/W_k)`$ $`=`$ $`{\displaystyle \underset{T_kS_k}{}}P(w_{k+1}/W_kT_k)\rho (W_k,T_k),`$ $`\rho (W_k,T_k)`$ $`=`$ $`P(W_kT_k)/{\displaystyle \underset{T_kS_k}{}}P(W_kT_k)`$ (5) which ensures a proper probability over strings $`W^{}`$, where $`S_k`$ is the set of all parses present in our stacks at the current stage $`k`$. An N-best EM variant is employed to reestimate the model parameters such that the PPL on training data is decreased — the likelihood of the training data under our model is increased. The reduction in PPL is shown experimentally to carry over to the test data. ## 3. $`A^{}`$ Decoder for Lattices ### 3.1. $`A^{}`$ Algorithm The $`A^{}`$ algorithm is a tree search strategy that could be compared to depth-first tree-traversal: pursue the most promising path as deeply as possible. To be more specific, let a set of hypotheses $`L=\{h:x_1,\mathrm{},x_n\},x_i𝒲^{}`$ — to be scored using the function $`f()`$ — be organized as a prefix tree. We wish to obtain the hypothesis $`h^{}=\mathrm{arg}\mathrm{max}_{hL}f(h)`$ without scoring all the hypotheses in $`L`$, if possible with a minimal computational effort. To be able to pursue the most promising path, the algorithm needs to evaluate the possible continuations of a given prefix $`x=w_1,\mathrm{},w_p`$ that reach the end of the lattice. Let $`C_L(x)`$ be the set of complete continuations of $`x`$ in $`L`$ — they all reach the end of the lattice, see Figure 4. Assume we have an overestimate $`g(x.y)=f(x)+h(y|x)f(x.y)`$ for the score of *complete* hypothesis $`x.y`$. denotes concatenation; imposing that $`h(y|x)=0`$ for empty $`y`$, we have $`g(x)=f(x),completexL`$. This means that the quantity defined as: $`g_L(x)`$ $``$ $`\underset{yC_L(x)}{\mathrm{max}}g(x.y)=f(x)+h_L(x),`$ (6) $`h_L(x)`$ $``$ $`\underset{yC_L(x)}{\mathrm{max}}h(y|x)`$ (7) is an overestimate of the most promising complete continuation of $`x`$ in $`L`$: $`g_L(x)f(x.y),yC_L(x)`$ and that $`g_L(x)=f(x),completexL`$. The $`A^{}`$ algorithm uses a potentially infinite stack<sup>1</sup><sup>1</sup>1The stack need not be larger than $`|L|=n`$ in which prefixes $`x`$ are ordered in decreasing order of $`g_L(x)`$; at each extension step the top-most prefix $`x=w_1,\mathrm{},w_p`$ is popped form the stack, expanded with all possible one-symbol continuations of $`x`$ in $`L`$ and then all the resulting expanded prefixes — among which there may be complete hypotheses as well — are inserted back into the stack. The stopping condition is: whenever the popped hypothesis is a complete one, retain that one as the overall best hypothesis $`h^{}`$. ### 3.2. $`A^{}`$ for Lattice Decoding There are a couple of reasons that make $`A^{}`$ appealing for our problem: $``$ the algorithm operates with whole prefixes $`x`$, making it ideal for incorporating language models whose memory is the entire prefix; $``$ a reasonably good overestimate $`h(y|x)`$ and an efficient way to calculate $`h_L(x)`$ (see Eq.6) are readily available using the n-gram model, as we will explain later. The lattices we work with retain the following information after the first pass: $``$ time-alignment of each node; $``$ for each link connecting two nodes in the lattice we retain: word identity, acoustic model score and n-gram language model score. The lattice has a unique starting and ending node, respectively. A lattice can be conceptually organized as a prefix tree of paths. When rescoring the lattice using a different language model than the one that was used in the first pass, we seek to find the complete path $`p=l_0\mathrm{}l_n`$ maximizing: $`f(p)`$ $`=`$ $`{\displaystyle \underset{i=0}{\overset{n}{}}}[logP_{AM}(l_i)`$ (8) $`+`$ $`LMweightlogP_{LM}(w(l_i)|w(l_0)\mathrm{}w(l_{i1}))`$ $``$ $`logP_{IP}]`$ where: $``$ $`logP_{AM}(l_i)`$ is the acoustic model log-likelihood assigned to link $`l_i`$; $``$ $`logP_{LM}(w(l_i)|w(l_0)\mathrm{}w(l_{i1}))`$ is the language model log-probability assigned to link $`l_i`$ given the previous links on the partial path $`l_0\mathrm{}l_i`$; $``$ $`LMweight>0`$ is a constant weight which multiplies the language model score of a link; its theoretical justification is unclear but experiments show its usefulness; $``$ $`logP_{IP}>0`$ is the “insertion penalty”; again, its theoretical justification is unclear but experiments show its usefulness. To be able to apply the $`A^{}`$ algorithm we need to find an appropriate stack entry scoring function $`g_L(x)`$ where $`x`$ is a partial path and $`L`$ is the set of complete paths in the lattice. Going back to the definition (6) of $`g_L()`$ we need an overestimate $`g(x.y)=f(x)+h(y|x)f(x.y)`$ for all possible $`y=l_k\mathrm{}l_n`$ complete continuations of $`x`$ allowed by the lattice. We propose to use the heuristic: $`h(y|x)`$ $`=`$ $`{\displaystyle \underset{i=k}{\overset{n}{}}}[logP_{AM}(l_i)+LMweight(logP_{NG}(l_i)`$ (9) $`+logP_{COMP})logP_{IP}]`$ $`+LMweightlogP_{FINAL}\delta (k<n)`$ A simple calculation shows that if $$logP_{NG}(l_i)+logP_{COMP}logP_{LM}(l_i),l_i$$ is satisfied then $`g_L(x)=f(x)+max_{yC_L(x)}h(y|x)`$ is a an appropriate choice for the $`A^{}`$ search. The justification for the $`logP_{COMP}`$ term is that it is supposed to compensate for the per word difference in log-probability between the n-gram model $`NG`$ and the superior model $`LM`$ with which we rescore the lattice — hence $`logP_{COMP}>0`$. Its expected value can be estimated from the difference in perplexity between the two models $`LM`$ and $`NG`$. The $`logP_{FINAL}>0`$ term is used for practical considerations as explained in the next section. The calculation of $`g_L(x)`$ (6) is made very efficient after realizing that one can use the dynamic programming technique in the Viterbi algorithm . Indeed, for a given lattice $`L`$, the value of $`h_L(x)`$ is completely determined by the identity of the ending node of $`x`$; a Viterbi backward pass over the lattice can store at each node the corresponding value of $`h_L(x)=h_L(ending\mathrm{\_}node(x))`$ such that it is readily available in the $`A^{}`$ search. ### 3.3. Some Practical Considerations In practice one cannot maintain a potentially infinite stack. We chose to control the stack depth using two thresholds: one on the maximum number of entries in the stack, called *stack-depth-threshold* and another one on the maximum log-probability difference between the top most and the bottom most hypotheses in the stack, called *stack-logP-threshold*. A gross overestimate used in connection with a finite stack may lure the search on a cluster of paths which is suboptimal — the desired cluster of paths may fall short of the stack if the overestimate happens to favor a wrong cluster. Also, longer partial paths — thus having shorter suffixes — benefit less from the per word $`logP_{COMP}`$ compensation which means that they may fall out of a stack already full with shorter hypotheses — which have high scores due to compensation. This is the justification for the $`logP_{FINAL}`$ term in the compensation function $`h(y|x)`$: the variance $`var[logP_{LM}(l_i|l_0\mathrm{}l_{i1})logP_{NG}(l_i)]`$ is a finite positive quantity so the compensation is likely to be closer to the expected value $`E[logP_{LM}(l_i|l_0\mathrm{}l_{i1})logP_{NG}(l_i)]`$ for longer $`y`$ continuations than for shorter ones; introducing a constant $`logP_{FINAL}`$ term is equivalent to an adaptive $`logP_{COMP}`$ depending on the length of the $`y`$ suffix — smaller equivalent $`logP_{COMP}`$ for long suffixes $`y`$ for which $`E[logP_{LM}(l_i|l_0\mathrm{}l_{i1})logP_{NG}(l_i)]`$ is a better estimate for $`logP_{COMP}`$ than it is for shorter ones. Because the structured language model is computationally expensive, a strong limitation is being placed on the width of the search — controlled by the `stack-depth` and the `stack-logP-threshold`. For an acceptable search width — runtime — one seeks to tune the compensation parameters to maximize performance measured in terms of WER. However, the correlation between these parameters and the WER is not clear and makes search problems diagnosis extremely difficult. Our method for choosing the search and compensation parameters was to sample a few complete paths $`p_1,\mathrm{},p_N`$ from each lattice, rescore those paths according to the $`f()`$ function (8) and then rank the $`h^{}`$ path output by the $`A^{}`$ search among the sampled paths. A correct $`A^{}`$ search should result in average rank 0. In practice this doesn’t happen but one can trace the topmost path $`p^{}`$ — in the offending cases $`p^{}h^{}`$ and $`f(p^{})>f(h^{})`$ — and check whether the search failed strictly because of insufficient compensation — a prefix of the $`p^{}`$ hypothesis is present in the stack when $`A^{}`$ returns — or because the path $`p^{}`$ fell short of the stack during the search — in which case the compensation and the search-width interact. The method we chose for sampling paths from the lattice was an N-best search using the n-gram language model scores; this is appropriate for pragmatic reasons — one prefers lattice rescoring to N-best list rescoring exactly because of the possibility to extract a path that is not among the candidates proposed in the N-best list — as well as practical reasons — they are among the “better” paths in terms of WER. ## 4. Experiments ### 4.1. Experimental Setup In order to train the structured language model (SLM) as described in we need parse trees from which to initialize the parameters of the model. Fortunately a part of the Switchboard (SWB) data has been manually parsed at UPenn ; let us refer to this corpus as the SWB-Treebank. The SWB training data used for speech recognition — SWB-CSR — is different from the SWB-Treebank in two aspects: $``$ the SWB-Treebank is a subset of the SWB-CSR data; $``$ the SWB-Treebank tokenization is different from that of the SWB-CSR corpus; among other spurious small differences, the most frequent ones are of the type presented in Table 1. Our goal is to train the SLM on the SWB-CSR corpus. #### 4.1.1. Training Setup The training of the SLM model proceeded as follows: $``$ train SLM on SWB-Treebank — using the SWB-Treebank closed vocabulary — as described in ; this is possible because for this data we have parse trees from which we can gather initial statistics; $``$ process the SWB-CSR training data to bring it closer to the SWB-Treebank format. We applied the transformations suggested by Table 1; the resulting corpus will be called SWB-CSR-Treebank, although at this stage we only have words and no parse trees for it; $``$ transfer the SWB-Treebank parse trees onto the SWB-CSR-Treebank training corpus. To do so we parsed the SWB-CSR-Treebank using the SLM trained on the SWB-Treebank; the vocabulary for this step was the union between the SWB-Treebank and the SWB-CSR-Treebank closed vocabularies; at this stage SWB-CSR-Treebank is truly a “treebank”; $``$ retrain the SLM on the SWB-CSR-Treebank training corpus using the parse trees obtained at the previous step for gathering initial statistics; the vocabulary used at this step was the SWB-CSR-Treebank closed vocabulary. #### 4.1.2. Lattice Decoding Setup To be able to run lattice decoding experiments we need to bring the lattices — SWB-CSR tokenization — to the SWB-CSR-Treebank format. The only operation involved in this transformation is splitting certain words into two parts, as suggested by Table 1. Each link whose word needs to be split is cut into two parts and an intermediate node is inserted into the lattice as shown in Figure 5. The acoustic and language model scores of the initial link are copied onto the second new link. For all the decoding experiments we have carried out, the WER is measured after undoing the transformations highlighted above; the reference transcriptions for the test data were not touched and the NIST SCLITE package was used for measuring the WER. ### 4.2. Perplexity Results As a first step we evaluated the perplexity performance of the SLM relative to that of a deleted interpolation 3-gram model trained in the same conditions. We worked on the SWB-CSR-Treebank corpus. The size of the training data was 2.29 Mwds; the size of the test data set aside for perplexity measurements was 28 Kwds — WS97 DevTest . We used a closed vocabulary — test set words included in the vocabulary — of size 22Kwds. Similar to the experiments reported in , we built a deleted interpolation 3-gram model which was used as a baseline; we have also linearly interpolated the SLM with the 3-gram baseline showing a modest reduction in perplexity: $$P(w_i|W_{i1})=\lambda P(w_i|w_{i1},w_{i2})+(1\lambda )P_{SLM}(w_i|W_{i1})$$ The results are presented in Table 2. ### 4.3. Lattice Decoding Results We proceeded to evaluate the WER performance of the SLM using the $`A^{}`$ lattice decoder described previously. Before describing the experiments we need to make clear one point; there are two 3-gram language model scores associated with the each link in the lattice: $``$ the language model score assigned by the model that generated the lattice, referred to as the LAT3-gram; this model operates on text in the SWB-CSR tokenization; $``$ the language model score assigned by rescoring each link in the lattice with the deleted interpolation 3-gram built on the data in the SWB-CSR-Treebank tokenization, referred to simply as the 3-gram — used in the experiments reported in the previous section. The perplexity results show that interpolation with the 3-gram model is beneficial for our model. Note that the interpolation: $$P(l)=\lambda P_{LAT3gram}(l)+(1\lambda )P_{SLM}(l)$$ between the LAT3-gram model and the SLM is illegitimate due to the tokenization mismatch. As explained previously, due to the fact that the SLM’s memory extends over the entire prefix we need to apply the $`A^{}`$ algorithm to find the overall best path in the lattice. The parameters controlling the $`A^{}`$ search were set to: $`logP_{COMP}`$ = 0.5, $`logP_{FINAL}`$ = 2, $`LMweight`$ = 12, $`logP_{IP}`$ = 10, `stack-depth-threshold=30`, `stack-depth-logP-threshold=100` — see (8) and ( 9). The parameters controlling the SLM were the same as in . The results for different interpolation coefficient values are shown in Table 3. The structured language model achieved an absolute improvement of 1% WER over the baseline; the improvement is statistically significant at the 0.002 level according to a sign test. In the 3-gram case, the $`A^{}`$ search looses 0.3% over the Viterbi search due to finite stack and heuristic lookahead. ### 4.4. Search Evaluation Results For tuning the search parameters we have applied the N-best lattice sampling technique described in Section 3.3.. As a by-product, the WER performance of the structured language model on N-best list rescoring — N = 25 — was 40.9%. The average rank of the hypothesis found by the $`A^{}`$ search among the N-best ones — after rescoring them using the structured language model interpolated with the trigram — was 1.07 (minimum achievable value is 0). There were 585 offending sentences — out of a total of 2427 test sentences — in which the $`A^{}`$ search led to a hypothesis whose score was lower than that of the top hypothesis among the N-best (1-best). In 310 cases the prefix of the rescored 1-best was still in the stack when $`A^{}`$ returned — inadequate compensation — and in the other 275 cases the 1-best hypothesis was lost during the search due to the finite stack size. One interesting experimental observation was that even though in the 585 offending cases the score of the 1-best was higher than that of the hypothesis found by $`A^{}`$, the WER of those hypotheses — as a set — was higher than that of the set of $`A^{}`$ hypotheses. ## 5. Conclusions Similar experiments on the Wall Street Journal corpus are reported in showing that the improvement holds even when the WER is much lower. We believe we have presented an original approach to language modeling that takes into account the hierarchical structure in natural language. Our experiments showed improvement in both perplexity and word error rate over current language modeling techniques demonstrating the usefulness of syntactic structure for improved language models. ## 6. Acknowledgments The authors would like to thank to Sanjeev Khudanpur for his insightful suggestions. Also thanks to Bill Byrne for making available the SWB lattices, Vaibhava Goel for making available the N-best decoder and Vaibhava Goel, Harriet Nock and Murat Saraclar for useful discussions about lattice rescoring.
warning/0001/quant-ph0001046.html
ar5iv
text
# Untitled Document Quantum authentication protocol <sup>1</sup>Guihua Zeng, and <sup>2</sup>Guangcan Guo <sup>1</sup>National Key Lab. on ISDN, XiDian University, Xi’an 710071, China <sup>2</sup>Department of physics and Nonlinear science centre, University of Science and technology of China, Heifei, 230026, China ## Abstract In this letter, we proposed a quantum authentication protocol. The authentication process is implemented by the symmetric cryptographic scheme with quantum effects. PACS:03.67.Dd,03.65.Bz Since the first finding that quantum effects may protect privacy information transmitted in an open quantum channel by S.Wiesner , and then by C.H.Bennett and G.Brassard , a remarkable surge of interest in the international scientific and industrial community has propelled quantum cryptography into mainstream computer science and physics \[3-10\]. Furthermore, quantum cryptography is becoming increasingly practical at a fast pace . Many quantum key distribution protocols have been proposed and some have been verified in experimental. However, the previous quantum key distribution (QKD) protocols are based on the legitimate users. In practice, the presented QKD protocols are completely insecure under the men-in-middle attack <sup>1</sup><sup>1</sup>1Men-in-middle attack: When the legitimate communicator Alice communicates the legitimate communicator Bob, Eve intercepts all qubit sent by Alice, and communicates Bob with impersonating Alice. Finally, Eve obtains two keys $`K_{AE},K_{EB}`$, where $`K_{AE}`$ represents the secret key between Alice and Eve, and $`K_{AE}`$ represents the secret key between Bob and Eve. As a result Eve can easily decrypt the ciphertext sent by Alice or Bob. . In addition, the users’ identities need to be verified in the point-to-point communicator and in the quantum cryptographic network , even if one does not need to generate the quantum key distribution. For axample, if the communicator Alice is a member of a network, when she want to enter the network, she first need to input the her password or others characteristics to verify her identities. Alice can enter the network only her password is correct. This process is called the authentication. If the implementation is completed by the quantum effects, we call it as quantum authentication. Unfortunately, there is no known way to initiate quantum authentication. Consider these cases, we propose a quantum authentication protocol in this letter, which is implemented by using EPR pair with the Bell’s theorem (or the EPR correlation ). In classic cryptography, there are three classes protocols: the key distribution protocol, the key verification protocol, and the authentication protocol, the uses of these protocols are different. These conceptions are useful in quantum cryptography, correspondingly, they are quantum key distribution, quantum verification and quantum authentication. It needs to stress that the quantum authentication scheme is different from the quantum key verification scheme . The quantum authentication scheme is always used in the case of the non-QKD like that in classic cryptography, e.g., when a user want to enter a network, he does not need to generate a key but need to input his correct identity or other characteristics. However, the quantum key verification scheme is used for verification of the obtained key in the quantum key distribution. In this paper, We proposed a quantum authentication protocol. It executes the following steps: Step 1. Alice and Bob transfer the sharing key $`K_1`$ <sup>2</sup><sup>2</sup>2Here we assume Alice and Bob have a sharing key before the currently communication into a sequence of measurement basis. While Alice and Bob need to verify their identification, or need to set up a new communication, they secretly transfer the reserved common key into a sequence of measurement basis according to the appointment. For example, if Alice and Bob use the measurement basis of polarization photon which was used in BB84 protocol, they may let the bit ’0’ correspond to rectilinear measurement basis and ’1’ correspond to diagonal measurement basis, or vice versa. We represent rectilinear measurement basis by the symbol $``$, and represent diagonal measurement basis by the symbol $``$. After transferred, Alice and Bob obtain a sequence of measurement basis, respectively. For example, if the common key is $`K_1=001101`$, the sequence of measurement basic is $`M_{K_1}=`$. Step 2. Alice and Bob set up a quantum communication channel. When Alice wants to secretly communicate Bob, Alice and Bob need to set up a quantum channel. The transmitting quantum states in the quantum channel may be arbitrary. For example the polarization photon state or the phase correction states. In this protocol, we use the phase correction states. So the channel consists of a source that emits EPR pairs of spin-$`\frac{1}{2}`$ particle, in a singlet state. The particles fly apart along the $`z`$ axis, towards the two legitimate users of the channel. Alice chooses a random basis for measuring one numbering of each EPR pair of particles. The other particle of each EPR pair is measured by Bob in the next step. Alice’s measurement results in effect determine, through the EPR corrections, a sequence of states for Bob’s particles. Step 3. Bob measures the strings of quantum states. Bob randomly measures the sequence of quantum states by using two measurement basis $`M`$ and $`M_{K_1}`$, where $`M`$ is the measurement basis for obtaining new authentication key and for quantum key distribution, $`M`$ is like the basis used in EPR protocol. $`M_{K_1}`$ is the measurement basis for identity authentication in the current communication. Step 4. Alice and Bob check the eavesdropper. For secure communication, the legitimate communicators Alice and Bob need to firstly detect the eavesdroppers. Bob randomly chooses some measurement results measured by the basis $`M`$ for checking the correction of EPR pair. Then the communicators judge the eavesdropping according to the Bell’s theorem (or EPR correlation). Step 5. Bob encrypts his results measured by $`M_{K_1}`$. Although Bob does not know the qubits measured by Alice, it will not influence the identity verification. Expressing the strings of quantum states for authentication by $$|\mathrm{\Psi }>=\{|\psi _1>,|\psi _2>,\mathrm{},|\psi _n>\}.$$ Where $`|\psi _i>`$ represents a qubit received by Bob. After finished measurement, Bob obtains $$|\mathrm{\Phi }>=M_{K_1}|\mathrm{\Psi }>,$$ where $`|\mathrm{\Phi }>=\{|\varphi _1>,|\varphi _2>,\mathrm{},|\varphi _n>\}`$ represents the measurement results under measurement basis $`M_{K_1}`$, i.e., $`|\varphi _i>=M_{K_1}|\psi _i>,i=1,2\mathrm{},n`$. Transferring $`|\mathrm{\Phi }>`$ into binary bits strings $`m`$, and then using $`K_1`$ to encrypt it, Bob obtains the ciphertext $$y=E_{K_1}\left(m\right).$$ Bob sends Alice the ciphertext $`y`$, and tells Alice the corresponding sequence numbers of quantum states $`|\psi _i>,i=1,2,\mathrm{},n`$. Step 6. Verifying Bob’s identity. Having received Bob’s results, Alice analyzes Bob’s results. Alice decrypts the ciphertext, $$m=E_{K_1}^1\left(y\right),$$ and compare her results with $`m`$, thereby Alice gets the measurement basis $`M_{K_t}`$. If $`K_t=K_1`$, Bob’s identity is true. Step 7 Verifying Alice’ identity. After Alice decrypted the ciphertext, Alice sends Bob the result $`m^{}`$. If $`m^{}=m`$, Alice’s identity is true. Step 8. Alice and Bob discard the authentication keys $`K_1`$, and set up a new authentication keys. After finished authentication, the authentication key $`K_1`$ is no longer use. The legitimate users obtain a new authentication key $`K_2`$. The method is same as the quantum key distribution in the EPR protocol. In practical communication, because of the noise effects errors are evitable in quantum channel. If the errors are produced by the Bob’s measurement, Bob tell Alice the error qubits to overcome the noise effects. If the errors are produced in transmission, Alice and Bob estimate the bound of errors $`e_t`$, and consider it in the identity verification. While the error is more than $`e_t`$, the communicators refuses each other, otherwise, the communicators are legitimate. The proposed scheme need a pre-key $`K_1`$, it means that an initial phase is necessary. Communicators may use quantum method to acquire the authentication key, e.g., the Biham’s technology . In the Ref., Biham et al. proposed a method to distribute the quantum key between Alice and Bob by the center. To prevent the center’s cheating (men-in-middle attack), the center must be legitimate and believable. Communicators can also use classic cryptographic method with quantum key distribution protocol to get the authentication key, e.g., the famous RSA system with QKD protocol. The method is as the following: first the legitimate communicators use the RSA system distributes the key, then in the valid time of RSA system obtain the sharing key $`K_1`$ by the QKD protocol <sup>3</sup><sup>3</sup>3For practical quantum cryptography, the combination of quantum cryptography and classic cryptography perhaps be useful and pontential direction, here the ‘valid time’ is important. The proposed quantum authentication protocol is provably secure. Because: i) Our protocol does not have the conspiracy problem of masquerading. If a forger wants to masquerade user Alice or Bob to communicate with others, he must find the common key. However, it is difficult to obtain the shared common secret because of the follows two reasons. First, the authentication key is obtained by the quantum key distribution protocol which is provably secure, so the authentication key is secure. Second the authentication key is used only one times, eavesdropper does not know any information about the authentication key. ii) The replay-attack will also not succeed in our protocol because the key is used only one times. iii) The quantum attacking strategy is invalid, the reason is the same as the analysis for previous QKD protocols. There is a weakness in our protocol. Although the obtaining of the common key in the last quantum communication is provably secure, the common key reservation has not circumvented possibility of attacking by eavesdroppers like in classic cryptography. In fact, this drawback exists in all symmetric cryptographic system. Of course, we can use the EPR effects or other quantum effect, i.e., quantum memory, to keep the common key, but the reservation time is very short according to current technology. A long time correlation of quantum states is need in the future. We use EPR effects with Bell’ theorem (or EPR correlation) to implement quantum authentication. It can also be implemented by noncommute quantum states or non-orthogonal quantum states with Heisenberg uncertainty principle (including the mixed states) by using a similar procedure. This project was supported by the Natural Science Foundation of China, Grant no: 69803008. References 1. S. Wiesner, Sigact News, vol. 15, no. 1, 78 (1983). 2. C. H. Bennett, G. Brassard, S. Breidbart, and S. Wiesner, Advances in Cryptology: Proceedings of Crypto 82, August 1982, Plenum Press, New York, pp. 267 - 275. 3. C. H. Bennett, and G. Brassard, Advances in Cryptology: Proceedings of Crypto’84, August 1984, Springer - Verlag, pp. 475 - 480. 4. C. H. Bennett, Phys. Rev. Lett. 68, 3121 (1992). 5. A. K. Ekert, Phys. Rev. Lett. 67, 661 (1991). A. K. Ekert, J. G. Rarity, P. R. Tapster, and G. M. Palma, Phys. Rev. Lett. 69, 1293 (1992). 6. C.H.Bennett,F.Bessette, G.Brassard, L.Salvail and J.Smolin, J.Cryptology 5, 3 (1992). 7. J. Breguet, A. Muller, and N. Gisin, J. Modern Optics, 41, 2405 (1994). 8. S. M. Barnett, and S. J. D.Phoenix, Journal of Modern Optics, 40, 1443 (1993). 9. C. A. Fuchs, N. Gisin, R. B. Griffiths, C. S. Niu, and A. Peres, Phys. Rev. A 56, 1163 (1997). 10. B. A. Slutsky, R. Rao, P. C. Sun, and Y. Fainman, Phys. Rev. A 57, 2383 (1998). 11. S. Phoenix, S. Barnett, P. Townsend and K. Blow, Journal of Modern Optics, 42, 1155 (1995). 12. C. Marand and P. D. Townsend, Optics Lett. 20, 1695 (1995). 13. E. Biham, B. Huttner, and T. Mor, Phys. Rev. A, Vol. 54, 2651 (1996). 14. J. S. Bell, Physics (Long Island City, N.Y.) 1, 195 (1965). 15. C. H. Bennett, G. Brassard, and N. D. Mermin, Phys. Rev. Lett. 68, 557 (1992). 16. G. Zeng and W. Zhang, Phys. Rev. A, vol 61, no 1, 2000, (in press).
warning/0001/nucl-th0001006.html
ar5iv
text
# Triplet Pairing in Neutron Matter ## 1 Introduction The study of superfluidity in infinitely extended nuclear systems has a long history \[1–82\], predating the 1967 discovery of pulsars , which were soon identified as rapidly rotating magnetic neutron stars . Interest in nucleonic pairing has intensified in recent years, owing primarily to experimental developments on two different fronts. In the field of astrophysics, a series of $`X`$-ray satellites (including Einstein, EXOSAT, ROSAT, and ASCA) has brought a flow of data on thermal emission from neutron stars, comprising both upper limits and actual flux measurements. The recent launching of the Chandra $`X`$-ray observatory provides further impetus for more incisive theoretical investigations. Realistic ab initio prediction of the microscopic physics of nucleonic superfluid components in the interiors of neutron stars is crucial to a quantitative understanding of neutrino cooling mechanisms \[85–95\] that operate immediately after their birth in supernova events, as well as the magnetic properties, vortex structure, rotational dynamics, and pulse timing irregularities \[96–99\] of these superdense stellar objects. In particular, when nucleonic species enter a superfluid state in one or another region of the star, suppression factors of the form $`\mathrm{exp}(\mathrm{\Delta }_F/k_BT)`$ are introduced into the expression for the emissivity, $`\mathrm{\Delta }_F`$ being an appropriate average measure of the energy gap at the Fermi surface. On the terrestrial front, the expanding capabilities of radioactive-beam and heavy-ion facilities have stimulated a concerted exploration of exotic nuclei and nuclei far from stability, with special focus on neutron-rich species . Pairing plays a prominent role in modeling the structure and behavior of these new nuclei. In this article we will be principally concerned with nucleonic pairing in the astrophysical setting. To a first approximation, a neutron star is described as a neutral system of nucleons (and possibly heavier baryons) and electrons (and possibly muons) in beta equilibrium at zero temperature, with a central density several times the saturation density $`\rho _0`$ of symmetrical nuclear matter . The gross structure of the star (mass, radius, pressure and density profiles) is determined by the Tolman-Oppenheimer-Volkov general relativistic equation of hydrostatic equilibrium, consistently with the continuity equation and the equation of state (which embodies the microscopic physics of the system). The star contains (i) an outer crust made up of bare nuclei arranged in a lattice interpenetrated by relativistic electrons, (ii) an inner crust where a similar Coulomb lattice of neutron-rich nuclei is embedded in Fermi seas of relativistic electrons and neutrons, (iii) a quantum fluid interior of coexisting neutron, proton, and electron fluids, and finally (iv) a core region of uncertain constitution and phase (but possibly containing hyperons, a pion or kaon condensate and/or quark matter). It is generally accepted that the neutrons in the background fluid of the inner crust of a neutron star will pair-condense in the <sup>1</sup>S<sub>0</sub> state. Qualitatively, this phenomenon can be understood as follows. At the relatively large average particle spacing at the “low” densities involved in this region ($`\rho \rho _0/10`$), the delocalized neutrons experience mainly the attractive component the <sup>1</sup>S<sub>0</sub> interaction. However, the pairing effect is quenched at higher densities, $`\rho _0`$ and beyond, due to the strong short-range component of this interaction. By similar reasoning, one expects <sup>1</sup>S<sub>0</sub> proton pairing to occur in the quantum fluid interior, in a density regime where the proton contaminant (necessary for charge balance and chemical equilibrium) reaches a partial density $`\rho _p\rho _0/10`$. The energy dependence of the nucleon-nucleon ($`NN`$) phase shifts in different partial waves offers some guidance in judging what nucleonic pair-condensed states are possible or likely in different regions of a neutron star. A rough correspondence between baryon density and $`NN`$ bombardment energies can be established through the Fermi momenta assigned to the nucleonic components of neutron-star matter , which provide a crude measure of the maximum relative momenta attained in the nuclear medium. In (pure) neutron matter, only $`T=1`$ partial waves are allowed by the Pauli principle. Moreover, one need only consider partial waves with $`L4`$ in the range of baryon density – optimistically, $`\rho <(34)\rho _0`$ – where a nucleonic model of neutron-star material is tenable. The <sup>1</sup>S<sub>0</sub> phase shift is positive at low energy (indicating an attractive in-medium force) but turns negative (repulsive) at around 250 MeV lab energy. Thus, unless the in-medium pairing force is dramatically different from its vacuum counterpart, the situation already suggested above should prevail: S-wave pairs should form at low densities but should be inhibited from forming when the density approaches that of ordinary nuclear matter. The next lowest $`T=1`$ partial waves are the three triplet P waves <sup>3</sup>P<sub>J</sub>, with $`J=0,1,2`$, which are actually coupled by the tensor force to the triplet F waves with the same total angular momentum $`J`$. Of these channels, the first two are not good candidates for pairing. In the <sup>3</sup>P<sub>0</sub> state, the phase shift is only mildly attractive at low energy, turning repulsive at a lab energy of 200 MeV, while the <sup>3</sup>P<sub>1</sub> phase shift is repulsive at all energies. On the other hand, the <sup>3</sup>P<sub>2</sub> phase shift is negative at all energies, and indeed is the most attractive $`T=1`$ phase shift at energies above about 160 MeV. This exceptional attractive strength in the <sup>3</sup>P<sub>2</sub> state is due mainly to the spin-orbit interaction and in part to the coupling to <sup>3</sup>F<sub>2</sub> wave via the tensor component of the $`NN`$ force. A substantial pairing effect in the <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> channel may be expected at densities somewhat in excess of $`\rho _0`$, again assuming that the relevant in-vacuum interaction is not greatly altered within the medium. The remaining $`T=0`$ partial waves with $`L4`$ are both singlets: <sup>1</sup>D<sub>2</sub> and <sup>1</sup>G<sub>4</sub>. However, the <sup>1</sup>D<sub>2</sub> phase shift, though positive over the energy domain of interest, is clearly dominated by the <sup>3</sup>P<sub>2</sub> phase shift, while the <sup>1</sup>G<sub>4</sub> phase shift, again positive, is even less important. Therefore, in the conventional picture of neutron-star matter (see, however, ref. ), singlet and triplet pairing are operationally equivalent to <sup>1</sup>S<sub>0</sub> and <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> pairing, respectively. There have been many illuminating studies of the former problem, notably . Here we shall concentrate on triplet pairing in the pure neutron system, ignoring the relatively minor modifications that arise in beta-stable neutral matter with its dilute admixture of protons. BCS theory furnishes a general framework for microscopic treatments of nucleonic pairing. As is the case for singlet pairing , two rather different problems must be overcome to realize a quantitative description of triplet pairing within this framework and make realistic predictions for the associated energy gap. The first problem to be solved is the construction of an accurate in-medium (or “effective”) particle-particle interaction that provides the pairing force, together with a consistent mass operator that provides the single-particle input for the BCS equation. The second problem is to solve the BCS gap equation (actually a system of gap equations when one deals with $`J0`$), once the inputs for the pairing matrix and normal-state single-particle energies have been determined. This first task places great demands on even the most advanced tools of many-body theory. For <sup>1</sup>S<sub>0</sub> pairing, there exists a “first generation” of results on this facet of the problem, obtained by several different approaches (Landau Fermi-liquid theory , the method of correlated basis functions , the polarization-potential approach and diagrammatic perturbation theory ). However, little progress has been made on this aspect of <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> pairing, and the issue of modifications of the associated input pairing interaction and single-particle energies due to short- and long-range correlations inside the nuclear medium remains open. We shall not attempt to address this issue here. Rather, we shall be concerned with the task of actually solving the triplet BCS system, assuming the particle-particle interaction is identical with the bare or vacuum $`NN`$ interaction, and that the input single-particle energies needed to complete the BCS equation(s) have the same form as for free particles (perhaps with an effective mass). This problem also proves to be nontrivial, since we face essentially the same difficulties as in singlet S-wave pairing, and more: 1. Again, the kernel of the BCS gap equation(s) has a pole at zero gap amplitude, implying an incipient logarithmic divergence. 2. Again, the nonlocality of the particle-particle interaction in momentum space implies slow convergence of the integral over momentum modulus $`k`$ in the BCS equation(s) – precluding application of the standard weak-coupling estimates of BCS theory . This feature of nucleonic pairing implies that one must integrate out to large values of the momentum variable. 3. As already intimated, an additional complication of pairing in higher angular momentum states with $`J0`$ arises from the (possible) angle-dependence of the gap, which can be parametrized in terms of dependences on the quantum numbers $`J,L`$ of the total and relative orbital angular momenta and on the magnetic quantum number $`M`$ associated with $`J`$. One must deal with a set of coupled gap equations for the gap components $`\mathrm{\Delta }_L^{JM}(k)`$ in an expansion of the (generally anisotropic) gap in the spin-angle functions $`\widehat{G}_L^{JM}(𝐧)`$. The combination of difficulties (a) and (b) can make numerical solution of the gap equation(s) a problematic exercise. To overcome these obstacles, a special procedure called the separation approach has been introduced in ref. and applied to <sup>1</sup>S<sub>0</sub> pairing. In this approach, the generic pairing matrix $`V(k,k^{})`$ is decomposed, identically, into a separable part, plus a remainder whose every element vanishes when either momentum variable lies on the Fermi surface. (As is characteristic of BCS theory, we assume that the particle-particle interaction is independent of frequency.) Applied to <sup>1</sup>S<sub>0</sub> pairing, this separation procedure reduces the BCS gap equation to two equivalent coupled equations: a nonsingular, quasilinear integral equation for the shape of the gap function and a nonlinear equation determining its amplitude at the Fermi surface. We hasten to point out that this strategy is by no means limited to the case where the pairing matrix elements are supplied by the bare $`NN`$ interaction – it will retain its practical and analytical value when a realistic in-medium particle-particle interaction is employed. The primary aim of this article is to generalize the separation approach to pairing in an arbitrary angular momentum channel and apply it to <sup>3</sup>P<sub>2</sub> pairing in neutron matter. Since, according to consideration (c) above, we are now confronted – in general – with a set of coupled gap equations rather than a single one, the situation is much more complicated than for <sup>1</sup>S<sub>0</sub> pairing. However, in this case the power of the separation method becomes even more apparent. It enables a decomposition of the problem into (i) the determination of the shape of the gap function in $`k`$-space through an appropriate nonsingular quasilinear integral equation whose kernel is practically independent of the energy gap, and, once this is done, (ii) the determination of a set of numerical coefficients specifying the angular dependence of the gap through a system of nonlinear ‘algebraic’ equations. In both singlet and triplet cases, the effectiveness of the separation scheme is greatly enhanced by the existence of a small parameter, given by the ratio of a suitable measure of the gap amplitude to the Fermi energy. Small-parameter expansions work well in the singlet problem, and even better in the triplet case. Corrections to the leading order in a small-parameter expansion of the <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> energy gap are down by a factor $`10^5`$ or $`10^6`$. This property leads to considerable simplifications in the analytical and numerical treatment of the triplet problem, greatly ameliorating complication (c). The separation approach is not merely a tool for numerical solution of gap equations. Following the lead of refs. , we shall demonstrate its very important role in facilitating incisive analysis of the nature of pairing solutions in different angular momentum channels. In particular, the separation formalism provides a basis for establishing certain universal properties of <sup>3</sup>P<sub>2</sub> pairing, properties that are independent of density and independent of the specifics of the pairing interaction itself, whether furnished by the in-vacuum or in-medium particle-particle interaction. The analytical results obtained here and in refs. remain valid outside the context of neutron matter and carry implications for the still more complex problem of pairing in liquid <sup>3</sup>He . In sect. 2, we display the BCS gap equations that describe pairing in an arbitrary two-body channel of given total spin $`S`$ and isospin $`T`$. Sect. 3 extends the separation method of ref. to this general problem, with special attention to <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> pairing. Sect. 4 is devoted to an analysis of <sup>3</sup>P<sub>2</sub> pairing in the absence of tensor coupling to the <sup>3</sup>F<sub>2</sub> state (“pure” <sup>3</sup>P<sub>2</sub> pairing). We offer a straightforward criterion for determination of the upper critical density at which a solution of the pairing problem ceases to exist in the given two-body channel, in terms of vanishing of a characteristic determinant associated associated with the separation formulation. In sect. 5 we introduce the small-parameter expansion of the pairing problem and use it to clarify the nature of the angle-average approximation proposed by Baldo et al. . In turn, the latter approximation is adopted to formulate an efficient procedure that allows reasonably accurate numerical solution of the coupled-channel, <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> pairing system when the goal is to find the magnitude of the pairing effect and the detailed structure of the pairing solutions is a lesser concern. We go on to recount how the fundamental character of the diverse solutions of the nonlinear pairing problem may be elucidated within the separation method. The derivation of universal features of <sup>3</sup>P<sub>2</sub> pairing is exemplified and discussed. Results from numerical application of the separation approach to triplet pairing in neutron matter are reported in sect. 6. Three representative models of the free-space $`NN`$ interaction are used to construct pairing matrix elements, the Argonne $`v_{18}`$ potential being chosen as an example from the current generation of models yielding high-precision fits of the $`NN`$ scattering data . The results presented explicitly include the critical upper density for gap closure, the momentum dependence of gap components, and the conventional average measure of the gap amplitude at the Fermi surface, variously calculated for “pure” and tensor-coupled <sup>3</sup>P<sub>2</sub> pairing, with and without use of the angle-average approximation. These results are compared with previous findings of other authors. In sect. 7, we indicate some promising future directions for research on pairing in infinitely extended nucleonic systems. An appendix reviews the separation method as developed for <sup>1</sup>S<sub>0</sub> pairing. Small-parameter expansion is employed to devise an approximation for the gap amplitude that may be taken as a generalization of the BCS weak-coupling formula. Additionally, the technique of small-parameter expansion is extended to the triplet pairing problem. ## 2 Any-channel gap equation To describe pairing in an arbitrary two-body state with total spin $`S`$, angular momentum $`J`$, and isospin $`T`$, it is useful to introduce a $`2\times 2`$ gap matrix $`\widehat{\mathrm{\Delta }}(𝐤)`$ which has the expansion (cf. ) $$\widehat{\mathrm{\Delta }}(𝐤)=\underset{LJM}{}\mathrm{\Delta }_L^{JM}(k)\widehat{G}_L^{JM}(𝐧)$$ (1) in the spin-angle matrices $$\widehat{G}_L^{JM}(𝐧)\underset{M_SM_L}{}\frac{1}{2}\frac{1}{2}\sigma _1\sigma _2|SM_SSLM_SM_L|JMY_{LM_L}(𝐧),$$ (2) with $`𝐧𝐤/k`$. The total spin and isospin quantum numbers $`S,T`$ ($`=(1,1)`$ for triplet states and neutron matter) are fixed throughout and generally suppressed. Setting $`S_{L^{}L_1}^{JMJ_1M_1}(𝐧)=\mathrm{Tr}\left[\widehat{G}_L^{}^{JM}(𝐧)\widehat{G}_{L_1}^{J_1M_1}(𝐧)\right]`$, the spin-angle matrices (2) may be shown to obey the orthogonality condition $$S_{L^{}L_1}^{JMJ_1M_1}(𝐧)d𝐧=\delta _{L^{}L_1}\delta _{JJ_1}\delta _{MM_1}.$$ (3) The particle-particle interaction has the corresponding expansion $$V(𝐤,𝐤^{})=\underset{LL^{}M}{}k|V_{LL^{}}^J|k^{}\widehat{G}_L^{JM}(𝐧)\widehat{G}_L^{}^{JM}(𝐧^{}),$$ (4) with $`|LL^{}|2`$ in the case of tensor forces. The coupled set of generalized gap equations for the components $`\mathrm{\Delta }_L^{JM}(k)`$ then reads $$\mathrm{\Delta }_L^{JM}(k)=\underset{L^{}L_1J_1M_1}{}(1)^\mathrm{\Lambda }k|V_{LJ}^{L^{}J}|k^{}S_{L^{}L_1}^{JMJ_1M_1}(𝐧^{})\frac{\mathrm{\Delta }_{L_1}^{J_1M_1}(k^{})}{2E(𝐤^{})}d𝐧^{}d\tau ^{},$$ (5) with $`\mathrm{\Lambda }S+(LL^{})/2`$ and $`\mathrm{d}\tau ^{}(2/\pi )(k^{})^2\mathrm{d}k^{}`$. The energy denominator $`E(𝐤)=[\xi ^2(k)+D^2(𝐤)]^{1/2}`$ involves the square of the single-particle excitation energy $`\xi (k)=\epsilon (k)\mu `$ of the normal system, measured relative the Fermi energy $`\mu `$, and a term $$D^2(𝐤)=\frac{1}{2}\underset{LJML_1J_1M_1}{}\mathrm{\Delta }_L^{JM}(k)\mathrm{\Delta }_{L_1}^{J_1M_1}(k)S_{LL_1}^{JMJ_1M_1}(𝐧)$$ (6) giving rise to an energy gap in the single-particle excitation spectrum of the pair-condensed system. Examination of the the explicit formula for the angular factor $`S_{LL_1}^{JMJ_1M_1}(𝐧)`$ reveals that a summation over total momentum $`J_1`$ must be present in (5) and (6). (This fact is well known in the theory of superfluid <sup>3</sup>He : the anisotropic ABM phase is built of states with total momentum $`J=1`$ and 2.) The sum over orbital momenta $`L^{}`$ on the l.h.s. of (5) appears only when tensor forces are present and is then restricted by $`|LL^{}|2`$ and parity conservation. On the other hand, the sum over $`L_1`$ resulting from the angular dependence of the gap contains, in principle, an infinite number of terms. We identify the energy gap as $`|D(𝐤)|`$ and note that it is in general angle dependent and in general has nodes. It is convenient to define an angle-averaged gap function $`\mathrm{\Delta }(k)`$ through the positive root of $$\mathrm{\Delta }^2(k)\overline{D^2}(k)=\frac{1}{4\pi }D^2(𝐤)𝑑𝐧=\frac{1}{4\pi }\frac{1}{2}\underset{LJM}{}\left|\mathrm{\Delta }_L^{JM}(k)\right|^2.$$ (7) (The last equality follows from the orthogonality property (3).) Analogously to the <sup>1</sup>S<sub>0</sub> pairing problem, the quantity (7), evaluated at the Fermi momentum $`k_F`$ and denoted $`\mathrm{\Delta }_F^2\overline{D^2}(k_F)`$ furnishes an overall measure of the strength of pairing correction to the ground-state energy in the preferred state. The notation $`\overline{D^2}(k_F)`$ is included here since it has been used earlier in the literature . We take this opportunity to caution the reader that the literature is inconsistent and sometimes vague in the choice of energy gap measure. Care must be taken in interpreting statements about the magnitude of “the gap.” Moreover, the choice made in normalizing or defining the pairing matrix elements is not uniform. We have attempted to be precise, and for the most part we follow the conventions established by Takatsuka and Tamagaki . ## 3 Separation formulation of the triplet pairing problem Generalizing the separation technique developed in ref. for the <sup>1</sup>S<sub>0</sub> problem, we again split the pairing matrix elements into a separable part and a remainder $`W_{LL^{}}^J(k,k^{})`$ that vanishes when either of the momentum variables $`k`$, $`k^{}`$ lies on the Fermi surface. Thus $$k^{}|V_{LL^{}}^J|kV_{LL^{}}^J(k,k^{})=v_{LL^{}}^J\varphi _{LL^{}}^J(k)\varphi _{LL^{}}^J(k^{})+W_{LL^{}}^J(k,k^{}),$$ (8) where $`v_{LL^{}}^JV_{LL^{}}^J(k_F,k_F)k_F|V_{L^{}L}^J|k_F`$ and $`\varphi _{LL^{}}^J=k_F|V_{LL^{}}^J|k_F/v_{LL^{}}^J`$. Substitution of this identity into the generic-channel gap equation (5) yields $`\mathrm{\Delta }_L^{JM}(k)`$ $``$ $`{\displaystyle \underset{L^{}}{}}(1)^\mathrm{\Lambda }{\displaystyle W_{LL^{}}^J(k,k^{})\underset{L_1J_1M_1}{}S_{L^{}L_1}^{JMJ_1M_1}(𝐧^{})\frac{\mathrm{\Delta }_{L_1}^{J_1M_1}(k^{})}{2\sqrt{\xi ^2(k^{})+\delta ^2}}\mathrm{d}𝐧^{}\mathrm{d}\tau ^{}}`$ (9) $`=`$ $`{\displaystyle \underset{L^{}}{}}D_{LL^{}}^{JM}\varphi _{LL^{}}^J(k)`$ where the coefficients $$D_{LL^{}}^{JM}=(1)^\mathrm{\Lambda }v_{LL^{}}^J\varphi _{LL^{}}^J(k)\underset{L_1J_1M_1}{}S_{L^{}L_1}^{JMJ_1M_1}(𝐧)\frac{\mathrm{\Delta }_{L_1}^{J_1M_1}(k)}{2\sqrt{\xi ^2(k)+D^2(𝐤)}}\mathrm{d}𝐧\mathrm{d}\tau $$ (10) (with $`\mathrm{d}\tau (2/\pi )k^2dk)`$ are just numbers. As in ref. (see also the appendix), we have introduced a typical scale $`\delta `$ in place of the gap in single-particle spectrum, in the term of (9) involving the portion $`W_{LL^{}}^J(k,k^{})`$ of the pairing matrix that vanishes on the Fermi surface. Such terms are quite insensitive to the choice of this scale, since the vanishing of $`W_{LL^{}}^J(k,k^{})`$ at $`k^{}=k_F`$ ensures that the integral over $`k^{}`$ receives its overwhelming contributions some distance from the Fermi surface where $`D^2(k^{})`$ is negligible compared to $`\xi ^2(k^{})`$. In the present context where the gap functions are angle-dependent, it is important to note that the $`\delta `$ quantity may be regarded as angle-independent. Numerical calculations show that the final result is practically independent of the choice of $`\delta `$ and we shall normally take it to be zero. Employing the orthogonality condition (3), one finds that only diagonal terms (meaning $`L_1=L^{}`$, $`J_1=J`$, $`M^{}=M`$) contribute to the sum on the l.h.s. of eq. (9), which becomes $$\mathrm{\Delta }_L^{JM}(k)\underset{L^{}}{}(1)^\mathrm{\Lambda }W_{LL^{}}^J(k,k^{})\frac{\mathrm{\Delta }_L^{}^{JM}(k^{})}{2|\xi (k^{})|}d\tau ^{}=\underset{L^{}}{}D_{LL^{}}^{JM}\varphi _{LL^{}}^J(k).$$ (11) Henceforth we shall frequently suppress the fixed index $`J`$. The functions $`\varphi _{LL_1}^J(k)\varphi ^{LL_1}(k)`$ superposed on the r.h.s. of (11) do not depend on $`M`$. Accordingly, the gap components $`\mathrm{\Delta }_L^M(k)=\mathrm{\Delta }_L^{JM}(k)`$ can be constructed as linear combinations $$\mathrm{\Delta }_L^M(k)=\underset{L_1L_2}{}D_{L_1L_2}^M\chi _L^{L_1L_2}(k)$$ (12) of $`M`$independent functions $`\chi _L^{L_1L_2}(k)=\chi _L^{JL_1L_2}(k)`$ multiplied by coefficients $`D_{L_1L_2}^M`$ determined by eq. (10). It is helpful to regard the quantities $`\mathrm{\Delta }_L^M(k)`$ as forming a vector $`\stackrel{}{\mathrm{\Delta }}^M`$ with components indexed by $`L`$, and similarly the functions $`\chi _{L_1L_2}^L(k)`$ as composing a vector $`\stackrel{}{\chi }^{L_1L_2}`$. The equations $$\chi _L^{L_1L_2}(k)\underset{L^{}}{}(1)^\mathrm{\Lambda }W_{LL^{}}(k,k^{})\frac{\chi _L^{}^{L_1L_2}(k^{})}{2|\xi (k^{})|}d\tau ^{}=\delta _{LL_1}\varphi ^{L_1L_2}(k)$$ (13) for the shape factors $`\chi _L^{L_1L_2}(k)`$ are readily obtained from eq. (11). From this result we see that $`\chi _L^{L_1L_2}(k_F)=\delta _{LL_1}`$ for any $`L_2`$, since the potentials $`W_{LL^{}}(k,k^{})`$ vanish for $`k`$ on the Fermi surface while $`\varphi ^{L_1L_2}(k_F)=1`$ for any $`L_1,L_2`$. It is important to observe that the equations (13) are free from any log singularities and form a coupled system of linear integral equations. We further remark that these equations can be solved without regard to the values of the coefficients $`D_{LL^{}}^{JM}=D_{LL^{}}^M`$ given by the set of equations (10) with $`\mathrm{\Delta }_{L_1}^{J_1M_1}(k)=\mathrm{\Delta }_{L_1}^{M_1}`$ expanded as in (12). It is also important to take note of an essential difference in the reformulated equations (11)–(13) as compared to the original gap equations (5). The sum over orbital angular momenta $`L_1`$ in (5) is in principle infinite, while in (11) only the states with $`L^{}=L`$ and $`L^{}=L\pm 2`$ can contribute, due to the conservation of the total angular momentum $`J`$ and parity. Therefore the number of components of the vectors $`\stackrel{}{\mathrm{\Delta }}^M`$ and $`\stackrel{}{\chi }^{L_1L_2}`$ does not exceed 3, since $`|LL^{}|2`$ and in eqs. (11)–(13) the total angular momentum $`J`$ has a fixed value. We may infer that, in the case of a <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> mixture where $`J=2`$, orbital angular momentum quantum numbers beyond 5 contribute only to the numerical factors $`D_{L_1L_2}^M`$. It is instructive to write out eqs. (11) and (13) in extended form for the <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> problem. Using a symbolic notation in terms of the operator $`T=W/2[\xi ^2+\delta ^2]^{1/2}`$, the system (11) reads $`\mathrm{\Delta }_1^M+T_{11}\mathrm{\Delta }_1^M+T_{13}\mathrm{\Delta }_3^M`$ $`=`$ $`D_{11}^M\varphi ^{11}+D_{13}^M\varphi ^{13},`$ $`\mathrm{\Delta }_3^M+T_{31}\mathrm{\Delta }_1^M+T_{33}\mathrm{\Delta }_3^M`$ $`=`$ $`D_{31}^M\varphi ^{31}+D_{33}^M\varphi ^{33}.`$ (14) Suppose all the coefficients $`D_{LL^{}}^M`$ are zero except $`D_{11}^M`$; then the solution $`\stackrel{}{\mathrm{\Delta }}^M`$ of (14), a two-component vector, reduces to $$\mathrm{\Delta }_L^M(k)=D_{11}^M\chi _L^{11}(k)(L=1,3),$$ (15) where the components $`\chi _1^{11}`$ and $`\chi _3^{11}`$ of $`\stackrel{}{\chi }_{11}`$ obey the pair of equations $`\chi _1^{11}+T_{11}\chi _1^{11}+T_{13}\chi _3^{11}`$ $`=`$ $`\varphi ^{11},`$ $`\chi _3^{11}+T_{31}\chi _1^{11}+T_{33}\chi _3^{11}`$ $`=`$ $`0.`$ (16) The vector $`\stackrel{}{\chi }_{13}`$ is given by a second pair of equations, $`\chi _1^{13}+T_{11}\chi _1^{13}+T_{13}\chi _3^{13}`$ $`=`$ $`\varphi ^{13},`$ $`\chi _3^{13}+T_{31}\chi _1^{13}+T_{33}\chi _3^{13}`$ $`=`$ $`0,`$ (17) derived by supposing that $`D_{13}^M`$ is present but all other $`D_{LL^{}}^M`$ are zero. The remaining two pairs of equations for the vector functions $`\stackrel{}{\chi }^{31}`$ and $`\stackrel{}{\chi }^{33}`$, namely $`\chi _1^{31}+T_{11}\chi _1^{31}+T_{13}\chi _3^{31}`$ $`=`$ $`0,`$ $`\chi _3^{31}+T_{31}\chi _1^{31}+T_{33}\chi _3^{31}`$ $`=`$ $`\varphi ^{31},`$ (18) and $`\chi _1^{33}+T_{11}\chi _1^{33}+T_{13}\chi _3^{33}`$ $`=`$ $`0,`$ $`\chi _3^{33}+T_{31}\chi _1^{33}+T_{33}\chi _3^{33}`$ $`=`$ $`\varphi ^{33}`$ (19) are constructed in the same fashion. The general solution of the <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> pairing problem is given by eq. (12) with the indices $`L_1`$ and $`L_2`$ taking on the values 1 or 3. It is worth noting that the same system of equations arises when one deals with $`SD`$ pairing in the deuteron channel, except that the indices $`L_1`$ and $`L_2`$ then take on the values 0 and 2. Let us return now to the problem of pairing in an arbitrary two-body channel. Having learned that the shape factors $`\chi _L^{L_1L_2}(k)`$ in the representation (12) may be found by solving a set of nonsingular linear integral equations independently of the coefficients $`D_{L_1L_2}^M`$, we must next consider the determination of these coefficients, assuming knowledge of the $`\chi `$ functions. Substitution of (12) into (10) provides a set of nonlinear ‘algebraic’ equations to be solved for the desired set of numbers: $`D_{LL^{}}^{JM}`$ $`=`$ $`(1)^\mathrm{\Lambda }v_{LL^{}}^J{\displaystyle \varphi ^{LL^{}}(k)\underset{L_1J_1M_1L_2L_3}{}S_{L^{}L_1}^{JMJ_1M_1}(𝐧)D_{L_2L_3}^{J_1M_1}}`$ (20) $`\times {\displaystyle \frac{\chi _{L_1}^{JL_2L_3}(k)}{2\sqrt{\xi ^2(k)+D^2(𝐤)}}}\mathrm{d}𝐧\mathrm{d}\tau .`$ In the vicinity of the (upper) critical value $`\rho _c`$ of the density $`\rho `$ where the gap value $`\mathrm{\Delta }_F`$ vanishes, the term $`D^2(𝐤)`$ in the denominator of (20) can be omitted, and in this case we arrive at a system of linear equations for the coefficients $`D_{LL^{}}^{JM}`$. The determinant of the linearized system should vanish at $`\rho =\rho _c`$, since by definition there is no solution at that point. Thus we obtain a new condition for gap closure. For example, in the case of <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> pairing (where channel coupling is included), the zero of the determinant corresponding to the dominant $`V_{11}^{J=2}`$ matrix elements of the Argonne $`v_{18}`$ potential gives a value $`k_c=3.6`$ fm<sup>-1</sup> for the critical Fermi momentum when the input function $`\xi (k)=\epsilon (k)\mu `$ in eqs. (20) is constructed from free single-particle energies $`\epsilon (k)`$. In the standard numerical approach to BCS pairing problems, the original system (5) for the gap components $`\mathrm{\Delta }_L^{JM}(k)`$ is solved by iteration. One starts with suitable “guesses” for these functions (ten in number, for <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> pairing) and inserts them into the r.h.s. of (5), generating a new set of component functions; and so on until convergence is hopefully achieved. Such a procedure encounters obstacles even in the case of <sup>1</sup>S<sub>0</sub> pairing . A conventional iterative approach is even more problematic for pairing in higher angular momentum states, where the dimensionality of the nonlinear problem is much larger. From this point of view, Eqs. (20) are far more appropriate as basis for numerical computation, since shape factors $`\chi _L^{L_1L_2}(k)`$ are determined separately and iterations are needed only to find the set of coefficients $`D_{L_1L_2}^{JM}`$ entering the representation (12). In principle, instead of trying to solve the system (20) directly, we could use the variables $`D_{LL^{}}^{JM}`$ (twenty in number, for <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> pairing) to parametrize the condensation energy and apply the variational principle to determine an appropriate set of these factors and a corresponding set of gap components $`\mathrm{\Delta }_L^{JM}(k)`$. However, one must remember that the performance of optimization algorithms, as well as the convergence of iterative procedures, is very sensitive to the degeneracies of the functional under consideration. Accordingly, any attempt at implementing such an approach should be prefaced by a thorough study of the symmetries of the gap problem and the degeneracies of its solutions. Such a study is of considerable fundamental interest in any case . ## 4 Triplet-P pairing without channel coupling To gain a better understanding of the triplet pairing problem in neutron matter, we revisit the pure (i.e., uncoupled) <sup>3</sup>P<sub>2</sub> case analyzed in some detail in . If tensor coupling to the <sup>3</sup>F<sub>2</sub> state is neglected, the system of gap equations (5) reduces to a set of five equations, corresponding to $`M=0,\pm 1,\pm 2`$. Taking into account the property of time-reversal invariance, which implies the relations $$\left(\mathrm{\Delta }_L^{JM}(k)\right)^{}=(1)^{J+M}\mathrm{\Delta }_L^{J,M}(k),$$ (21) it is seen that only three of the generally complex components $`\mathrm{\Delta }_1^{2M}`$ are independent. The independent components have customarily been written as $`\mathrm{\Delta }_1^{20}(k)`$ $`=`$ $`\delta _0(k),`$ $`\mathrm{\Delta }_1^{21}(k)`$ $`=`$ $`\left(\mathrm{\Delta }_1^{2,1}(k)\right)^{}=\delta _1(k)+i\eta _1(k),`$ $`\mathrm{\Delta }_1^{22}(k)`$ $`=`$ $`\left(\mathrm{\Delta }_1^{2,2}(k)\right)^{}=\delta _2(k)+i\eta _2(k),`$ (22) where the $`\delta _i(k)`$ and $`\eta _i(k)`$ are real functions of $`k`$. Takatsuka and Tamagaki , as well as Amundsen and Østgaard , have carried out numerical studies of the pure <sup>3</sup>P<sub>2</sub> problem that demonstrate an interesting feature of triplet pairing in neutron matter (and in fact of triplet-P pairing more generally). Solutions of different types were found to be nearly degenerate, i.e., they yield very nearly the same condensation energy. The spread of condensation energies reported by Takatsuka and Tamagaki at a typical density is only 3%, and Østgaard and Amundsen obtained very similar results. The calculations of these authors were based on an older nucleon-nucleon potential called OPEG (see sec. 6 for details). Although this potential model is no longer competitive, its essential aspects are semi-quantitatively correct. At any rate, it will be argued that the nearly degenerate character of the level structure in the triplet pairing problem and the energy ordering of different solution types are very insensitive to the interaction assumed. Important insights into the uncoupled problem can be achieved within the separation method . For pure <sup>3</sup>P<sub>2</sub> pairing, eq. (15) becomes simply $$\mathrm{\Delta }_1^{2M}(k)=D_{11}^{2M}\chi _1^{11}(k),$$ (23) while the shape factor $`\chi _1^{11}(k)`$ obeys the uncoupled equation $$\chi _1^{11}(k)+\frac{W_{11}(k,k^{})\chi _1^{11}(k^{})}{2|\xi (k)|}d\tau ^{}=\varphi ^{11}(k).$$ (24) We now make the decomposition $`D_{112}^{M0}=(\lambda _M+i\kappa _M)\delta _0(k_F)/\sqrt{6}`$, where $`\delta _0(k_F)=D_{112}^{M=0}`$ serves as a scale factor. Relations between the parameters $`\lambda _M`$ and $`\kappa _M`$ and the quantities $`\delta _i(k_F)`$ and $`\eta _i(k_F)`$ ($`i=1,2`$) stem from the definitions contained in eqs. (22). Evaluating the spin sums $`S_{11}^{2M2M^{}}`$ and separating real and imaginary parts in eq. (20), we extract the following set of equations determining the coefficients $`D_{11}^{JM}`$ (cf. refs. ): $`\lambda _2`$ $`=`$ $`v^{}[\lambda _2(J_0+J_5)\lambda _1J_1\kappa _1J_2J_3],`$ $`\kappa _2`$ $`=`$ $`v^{}[\kappa _2(J_0+J_5)\kappa _1J_1+\lambda _1J_2+J_4],`$ $`\lambda _1`$ $`=`$ $`v^{}[\lambda _1J_6(\lambda _2+1)J_1+\kappa _2J_2\kappa _1J_4/2],`$ $`\kappa _1`$ $`=`$ $`v^{}[\kappa _1J_7\kappa _2J_1(\lambda _21)J_2\lambda _1J_4/2],`$ $`1`$ $`=`$ $`v^{}[(\lambda _1J_1\kappa _1J_2+\lambda _2J_3\kappa _2J_4)/3+J_5],`$ (25) where $`v^{}=(\pi /2)v_{11}^{J=2}`$. The integrals $`J_i`$ are given by $$J_i=f_i(\theta ,\phi )\frac{\varphi ^{11}(k)\chi _1^{11}(k)}{2E(𝐤)}k^2dk\frac{\mathrm{d}𝐧}{4\pi },$$ (26) with $`f_0=13z^2`$, $`f_1=3xz/2`$, $`f_2=3yz/2`$, $`f_3=3(2x^2+z^21)/2`$, $`f_4=3xy`$, and $`f_5=(1+3z^2)/2`$, where $`x=\mathrm{sin}\theta \mathrm{cos}\phi `$, $`y=\mathrm{sin}\theta \mathrm{sin}\phi `$, and $`z=\mathrm{cos}\theta `$. Of these integrals, only $`J_5`$ contains a principal term behaving like $`\mathrm{ln}\mathrm{\Delta }_F`$, since only $`f_5`$ among the $`f_i`$ does not vanish under the angular integration when multiplied by an angle-independent factor. The normal-state single-particle energy $`\xi (k)`$ present in the energy denominator $`E(𝐤)=[\xi ^2(k)+D^2(𝐤)]^{1/2}`$ is conventionally assumed to have finite slope $`d\xi (k)/dk`$, while the gap term takes the explicit form $`D^2(𝐤)`$ $`=`$ $`{\displaystyle \frac{\delta _0^2}{16\pi }}[(1+\lambda _2)^2+\kappa _1^2+\kappa _2^2+(\lambda _1^24\lambda _2\kappa _1^2)x^2`$ (27) $``$ $`2(\lambda _1+\lambda _1\lambda _2+\kappa _1\kappa _2)xz+(3+\lambda _1^2\lambda _2^22\lambda _2)z^2`$ $`+`$ $`2(2\kappa _2\kappa _1\lambda _1)xy+2(\kappa _1+\lambda _1\kappa _2\lambda _2\kappa _1)yz][\chi _1^{11}(k)]^2.`$ In their angular content, the system (25) and the trigonometric decomposition (27) are consistent with the formulations of the uncoupled <sup>3</sup>P<sub>2</sub> problem given in refs. , and specifically with eqs. (4.7)–(4.8) of Amundsen and Østgaard. However, our separation formulation, as expressed in eqs. (23)–(27), has distinct advantages as a basis for numerical solution of this problem. To begin with, there is only one shape function $`\chi (k)\chi _1^{11}(k)`$, which satisfies a linear integral equation of exactly the same form as for singlet-S pairing (cf. eq. (60). This property of shape invariance follows naturally from the formal development presented in sect. 3. Given the key physical fact that the angle dependence of the problem only comes into play significantly in the close vicinity of the Fermi surface, we observe that this part of momentum space is almost entirely eliminated from the $`k^{}`$ integration in eq. (24) by the vanishing of $`W(k,k^{})`$ at $`k^{}=k_F`$. Thus, to an excellent approximation, all solutions $`\mathrm{\Delta }_1^{2M}`$ of the uncoupled triplet problem have identical shapes. In particular, the positions of their zeros and maxima coincide, as is clearly seen in the figures of previous works on triplet pairing in neutron matter . To establish the angular dependences of the allowed solutions of the <sup>3</sup>P<sub>2</sub> problem, the shape equation (24) must of course be supplemented by a set of equations that determine the coefficients or amplitudes corresponding to the different $`M`$ values. An important benefit of the separation strategy is that the nonlinear aspects of the problem are now concentrated entirely in the latter system of equations, given by (25) – equations for a set of numbers $`D_{112}^M`$ rather than for a set of functions. By contrast, earlier authors have worked directly with a set of singular linear integral equations for the five functions $`\delta _0(k)`$, $`\delta _1(k)`$, $`\eta _1(k)`$, $`\delta _2(k)`$, and $`\eta _2(k)`$ defined in eq. (22). In this approach, the dependences of gap quantities on the modulus of the momentum and on angles (or magnetic quantum numbers) are confronted simultaneously, with attendant opportunities for numerical difficulties. Obviously, the process of numerical solution is facilitated by the separation of “shape” and “angular” aspects of the problem that has been accomplished in eqs. (23)–(27). Less obviously, this transformation opens the way to an incisive analysis of the nature of these solutions. Such an analysis has been carried out in refs. , extending the formalism to finite temperature $`T`$ via introduction of the factor $`\mathrm{tanh}\left[E(𝐤)/k_BT\right]`$ in the integrands of the basic gap equations eq. (5) and the $`J_i`$ integrals (26). Appealing to the orthogonality of the spherical functions for different $`M`$, it is possible to find special solutions of the system (25) (or of the system (4.7) of ref. ) that satisfy most or some of its equations identically. The simplest types of consistent solutions are characterized as follows : 1. Only $`M=\pm 2`$ components are present: $`\delta _0=\delta _1=\eta _1=0`$. 2. Only $`M=0`$ is present: $`\delta _1=\eta _1=\delta _2=\eta _2=0`$. 3. Only $`M=\pm 1`$ are present: $`\delta _0=\delta _2=\eta _2=0`$. 4. Coupled case containing $`M=0,\pm 2`$: $`\delta _1=\eta _1=0`$. In addition, there is naturally the general case involving all five $`M`$ values. For purposes of illustration, let us examine the situation for solution Types 1 and 2 in some detail. Type 1. Under the assumption that only the component with $`M=2`$ is present (its $`M=2`$ partner being included trivially), we need only deal with $`\mathrm{\Delta }_1^{22}(k)=\delta _2(k_F)\chi (k)`$, noting that a free choice of phase allows us to set $`\eta _2=0`$. In this case only the first of eqs. (25) is relevant and it becomes $$\frac{1}{v}=\varphi (k)\chi (k)\frac{\frac{3}{2}\mathrm{sin}^2\theta }{2\sqrt{\xi ^2(k)+\frac{3}{2}\mathrm{\Delta }_F^2\chi ^2(k)\mathrm{sin}^2\theta }}d𝐧d\tau ,$$ (28) where $`v=v_{11}^{J=2}`$. All the other equations of the set (25) are met identically. (We remind the reader that $`\mathrm{\Delta }_F^2`$ is the value, on the Fermi surface, of the angle average defined in eq. (7).) Apart from the angular dependences and angular integration, eq. (28) has the same form as the equation (61) for the gap amplitude in the <sup>1</sup>S<sub>0</sub> problem (see the appendix). The angular integration can actually be performed analytically, with the result $$\frac{1}{v}=\frac{\varphi (k)\chi (k)}{2\sqrt{\xi ^2(k)+\frac{3}{2}\mathrm{\Delta }_F^2\chi ^2(k)}}\mathrm{\Phi }_2(\beta _2)d\tau ,$$ (29) where $`\mathrm{\Phi }_2(\beta _2)`$ $`=`$ $`{\displaystyle \frac{3}{2\beta _2}}\mathrm{arcsin}\beta _2+{\displaystyle \frac{3}{4\beta _2^2}}\left[\sqrt{1\beta _2^2}{\displaystyle \frac{\mathrm{arcsin}\beta _2}{\beta _2}}\right],`$ $`(\beta _2(k))^2`$ $`=`$ $`{\displaystyle \frac{\frac{3}{2}\mathrm{\Delta }_F^2\chi ^2(k)}{\xi ^2(k)+\frac{3}{2}\mathrm{\Delta }_F^2\chi ^2(k)}}.`$ (30) The root of eq. (29) can be found numerically without difficulty, after finding the universal shape function $`\chi (k)=\chi _1^{11}(k)`$ by matrix inversion of the discretized form of eq. (24). This has been done for pairing matrix elements calculated from the OPEG potential employed by Takatsuka and Tamagaki and Amundsen and Østgaard ; some results for this case are reported in sect. 6. Type 2. Here there is the single component $`\mathrm{\Delta }_1^{20}(k)=\delta _0(k_F)\chi (k)=\mathrm{\Delta }_F\chi (k)`$. The dispersion equation determining $`\mathrm{\Delta }_F`$ is given by the last of eqs. (25), which becomes $$\frac{1}{v}=\varphi (k)\chi (k)\frac{\frac{1}{2}(1+3\mathrm{cos}^2\theta ))}{2\sqrt{\xi ^2(k)+\frac{1}{2}\mathrm{\Delta }_F^2(1+3\mathrm{cos}^2\theta )\chi ^2(k)}}d𝐧d\tau .$$ (31) Analytic integration over angles is again possible, yielding $$\frac{1}{v}=\frac{\varphi (k)\chi (k)}{2\sqrt{\xi ^2(k)+\frac{1}{4}\mathrm{\Delta }_F^2\chi ^2(k)}}\mathrm{\Phi }_0(\beta _0)d\tau ,$$ (32) where $`\mathrm{\Phi }_0(\beta _0)`$ $`=`$ $`{\displaystyle \frac{1}{2\beta _0}}\mathrm{ln}\left(\beta _0+\sqrt{1+\beta _0^2}\right)+{\displaystyle \frac{3}{4\beta _{0}^{}{}_{}{}^{2}}}\left[\sqrt{1+\beta _0^2}\mathrm{ln}\left(\beta _0+{\displaystyle \frac{1}{\beta _0}}\sqrt{1+\beta _0^2}\right)\right],`$ $`(\beta _0(k))^2`$ $`=`$ $`{\displaystyle \frac{\frac{3}{4}\mathrm{\Delta }_F^2\chi ^2(k)}{\xi ^2(k)+\frac{1}{4}\mathrm{\Delta }_F^2\chi ^2(k)}}.`$ (33) Some numerical results for this case are also reported in sect. 6, again for the OPEG potential. ## 5 Small-parameter expansion and angle-average approximation Quantitative description of pairing in nuclear systems is beset with special difficulties because the strong inner repulsion (or strong momentum dependence) present in some partial waves (notably S states) implies that the momentum integral entering the gap equation must be extended to large momenta . This behavior precludes the application of simple estimates such as the BCS weak-coupling formula, which assumes sharp localization of the pairing interaction around the Fermi surface. However, important simplifications are permitted by the existence of a small dimensionless parameter $`d_F`$, given essentially by the scale of the energy gap at the Fermi surface, relative to the value of the Fermi energy $`\epsilon _F`$. In the case of <sup>1</sup>S<sub>0</sub> pairing, the small parameter is just $`d_F=\mathrm{\Delta }_F/\epsilon _F`$, where $`\mathrm{\Delta }_F\mathrm{\Delta }(k_F)`$. Framing this problem within the separation method, Taylor-series expansion of the energy denominator of the equation for the gap amplitude $`\mathrm{\Delta }_F`$ has been employed in ref. to derive a sequence of approximants for this quantity expressed as series of terms in $`(d_F^2)^n\mathrm{ln}d_F`$ and $`(d_F^2)^n`$ with integral $`n`$. The derivations involved are outlined in appendix A. Even the leading (zeroth) approximant yields reasonable results: the exact solution for $`\mathrm{\Delta }_F`$ within the relevant density range is reproduced within 5%, based on pairing matrix elements formed from the <sup>1</sup>S<sub>0</sub> component of the Argonne $`v_{18}`$ interaction. In the triplet case considered here, the small parameter is appropriately defined as $`d_F=[\overline{D^2}(k_F)]^{1/2}/\epsilon _F\mathrm{\Delta }_F/\epsilon _F`$. A characteristic value of this quantity in the density region of interest is 0.5 MeV/80 MeV $`6\times 10^3`$, which is at least an order of magnitude smaller than typical values for singlet-S pairing. On this basis, one may expect to achieve good accuracy through an approximation that retains only the principal logarithmic terms $`\mathrm{ln}D^2(𝐤)`$ together with those independent of $`D^2(𝐤)`$, while omitting corrections $`d_F^2\mathrm{ln}d_F`$. Small-parameter approximations are examined in more detail in the appendix. Here we restrict ourselves only to brief remarks. Consider the integral on the r.h.s. of eqs. (20) for the coefficients $`D_{LL^{}}^{JM}`$. We may carry out an integration by parts with respect to the variable $`\xi `$, to derive a principal term (within the angular integration) that goes like $`\mathrm{ln}D^2(k_F,𝐧)`$ and a remainder that involves $`D^2(𝐤)`$ only in the factor $$\mathrm{ln}\left(|\xi |+\left[\xi ^2+D^2(𝐤)\right]^{1/2}\right).$$ Since latter quantity is quite smooth in the vicinity of the Fermi surface, its replacement by $`\mathrm{ln}(2|\xi |)`$ is valid within corrections of order of $`d_F^2\mathrm{ln}d_F`$ that belong to the next stage of approximation. Writing $`D^2(k_F,𝐧)=\mathrm{\Delta }_F^2𝒟^2(𝐧)`$, we extract the scale $`\mathrm{\Delta }_F^2`$ and express the angular dependence by $`𝒟^2(𝐧)`$. The dominant log term then decomposes as $`\mathrm{ln}D^2(k_F,𝐧)=\mathrm{ln}\mathrm{\Delta }_F^2+\mathrm{ln}𝒟^2(𝐧)`$. By virtue of the orthogonality condition (3), the angle-independent portion $`\mathrm{ln}\mathrm{\Delta }_F^2`$ contributes only to terms on the r.h.s. of eq. (20) that are diagonal in the magnetic quantum number, i.e., only for $`M_1=M`$. Since $`\mathrm{\Delta }_F`$ is a small quantity, the contribution from the scale term $`\mathrm{ln}\mathrm{\Delta }_F^2`$ greatly exceeds that from the angle-dependent part $`\mathrm{ln}𝒟^2(𝐧)`$, which remains of order of 1 even in the limit $`\mathrm{\Delta }_F0`$. It is worth noting that the $`\mathrm{\Delta }`$-independent contributions coming from terms in $`\mathrm{ln}(2|\xi |)`$ are also of order of 1. However, these terms are angle-independent and, like the scale term, contribute only to the diagonal part of the r.h.s. of (20). In fact, to within corrections of order $`d_F^2\mathrm{ln}d_F`$, the result for the $`D`$ coefficients calculated from the angle-independent terms alone (i.e., dropping $`\mathrm{ln}𝒟^2(𝐧)`$), is the same as the result we obtain if we simply replace the quantity $`D^2(𝐤)`$ in the energy denominator of the original equation by its angle average (7), $$\overline{D^2}(k)=\frac{1}{8\pi }\underset{L}{}|\mathrm{\Delta }_L^{JM=0}(k)|^2+\frac{1}{4\pi }\underset{LM>0}{}|\mathrm{\Delta }_L^{JM}(k)|^2.$$ (34) We reiterate that justification of this replacement, which corresponds to the angle-average approximation first introduced by Baldo et al. , rests on (i) the unimportance of the angle-dependent quantity $`\mathrm{ln}𝒟^2(𝐧)`$ relative to $`\mathrm{ln}\mathrm{\Delta }_F^2`$ and (ii) the unimportance of correction terms of order $`d_F^2\mathrm{ln}d_F`$ or higher in the small parameter expansion. The angle-average approximation generally overestimates the pairing gap and the magnitude of the pairing energy (i.e. the pair condensation energy correction to the energy of the normal ground state). However, it allows a considerable simplification of the triplet pairing problem with very modest sacrifice in accuracy, if our primary interest lies in the magnitude of the pairing effect rather than the detailed structure of pairing solutions. Making the replacement (34) in the general coupled-channel equations (20) for the gap amplitudes, we arrive at the quasi-uniform system $$D_{LL^{}}^M=(1)^\mathrm{\Lambda }v_{LL^{}}\varphi ^{LL^{}}(k)\underset{L_1L_2}{}D_{L_1L_2}^M\frac{\chi _L^{}^{L_1L_2}(k)}{2\sqrt{\xi ^2(k)+\overline{D^2}(k)}}\mathrm{d}\tau .$$ (35) Since the index $`M`$ appears explicitly only on the quantities being sought, there exists a degeneracy with respect to this quantum number: All sets of $`D_{LL^{}}^{JM}`$ yielding the same result for the gap quantity $`\overline{D^2}(k)`$ of eq. (34) are equivalent. We first apply the angle-averaging prescription to the pure <sup>3</sup>P<sub>2</sub> problem. In this case eq. (35) becomes simply $$1=v\frac{\left(\varphi ^{11}(k)\right)^2\chi _1^{11}(k)}{2\sqrt{\xi ^2(k)+\overline{D^2}(k)}}d\tau $$ (36) with $`\overline{D^2}(k)=\mathrm{\Delta }_F^2[\chi _1^{11}(k)]^2=\mathrm{\Delta }_F^2\chi ^2(k)`$ and $`v=v_{11}^{J=2}`$. This equation, having no explicit reference to magnetic substates, is identical in form to the singlet-channel equation (61) and is solved by the same method. Numerical results from solution of the uncoupled gap problem in this approximation are presented and discussed in sect. 6. If we do not care about the detailed angular behavior of solutions, but only about the size of the gap (as measured by $`\mathrm{\Delta }_F`$) and the condensation energy, the angular-averaging prescription gives very good predictions. This statement is justified by the close agreement with the results shown in fig. 1 for solution Types 1 and 2 and by the near degeneracy of the $`\mathrm{\Delta }_F`$ values and condensation energies found numerically and analytically for the different solution types. Turning to the general coupled-channel case, observe that the terms with $`L_1L^{}`$ on the r.h.s. of (35) contain integrals $$A_{LL^{}}^{L_1L_2}=\frac{\varphi ^{LL^{}}(k)\chi _L^{}^{L_1L_2}(k)}{2\sqrt{\xi ^2(k)+\overline{D^2}(k)}}d\tau \frac{\varphi ^{LL^{}}(k)\chi _L^{}^{L_1L_2}(k)}{2|\xi (k)|}d\tau $$ (37) which are almost independent of the gap value, since the property $`\chi _L^{}^{L_1L_2}=\delta _{LL_1}`$ implies that the numerators of the integrands vanish at the Fermi surface. In the diagonal case, $`L_1=L^{}`$, we introduce an unknown positive number $$Z=\frac{[\chi _1^{11}(k)]^2}{2\sqrt{\xi ^2(k)+\overline{D^2}(k)}}d\tau $$ (38) and write $$\frac{\varphi ^{LL^{}}(k)\chi _L^{}^{L^{}L_2}(k)}{2\sqrt{\xi ^2(k)+\overline{D^2}(k)}}d\tau Z+B_{LL^{}}^{L^{}L_2},$$ (39) where $$B_{LL^{}}^{L^{}L_2}=\frac{\varphi ^{LL^{}}(k)\chi _L^{}^{L^{}L_2}(k)[\chi _1^{11}(k)]^2}{2\sqrt{\xi ^2(k)+\overline{D^2}(k)}}d\tau \frac{\varphi ^{LL^{}}(k)\chi _L^{}^{L^{}L_2}(k)[\chi _1^{11}(k)]^2}{2|\xi (k)|}d\tau .$$ (40) The simplification on the right in eq. (40) is permitted since $`\varphi ^{LL^{}}(k_F)=\chi _L^{}^{L^{}L_2}(k_F)=1`$ and the numerator in the integrands again vanishes at the Fermi surface. The integrals $`B_{LL^{}}^{L^{}L_2}`$ are effectively already known: with the indicated simplifications, they only depend on the $`\chi `$ functions, which may be found from eqs. (13) without reference to the gap quantities $`D_{LL^{}}^M`$ being sought. We now specialize to coupled-channel, <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> pairing where the system (35) for the coefficients $`D_{LL^{}}^M`$ contains four equations. For a solution of this set of equations to exist, the determinant $$C(Z)=\left|\begin{array}{cccc}(1/v_{11})+B_{11}^{11}+Z& B_{11}^{13}+Z& A_{11}^{31}& A_{11}^{33}\\ A_{13}^{11}& (1/v_{13})+A_{13}^{13}& B_{13}^{31}+Z& B_{13}^{33}+Z\\ B_{31}^{11}+Z& B_{31}^{13}+Z& (1/v_{31})+A_{31}^{31}& A_{31}^{33}\\ A_{33}^{11}& A_{33}^{31}& B_{33}^{31}+Z& (1/v_{33})+B_{33}^{33}+Z\end{array}\right|,$$ (41) which depends on $`\mathrm{\Delta }_F`$ (or the $`D_{LL^{}}^M`$ quantities) only through the parameter $`Z`$, must vanish. Evaluation of this determinant may be simplified by subtracting the third row from the second row and the second row from the fourth row, and then subtracting the second column from the first column and the third column from the fourth column. This leads to a quadratic equation for the determination of the unknown parameter $`Z`$: $$C(Z)=c_2Z^2+c_1Z+c_0=0,$$ (42) where $$c_0=C(0)=\left|\begin{array}{cccc}C_{00}& C_{01}& C_{02}& C_{03}\\ C_{10}& C_{11}& C_{12}& C_{13}\\ C_{20}& C_{21}& C_{22}& C_{23}\\ C_{30}& C_{31}& C_{32}& C_{33}\end{array}\right|,$$ (43) $$c_2=\left|\begin{array}{cc}C_{30}C_{10}[C_{31}C_{11}]& C_{33}C_{13}[C_{32}C_{12}]\\ C_{00}C_{20}[C_{01}C_{21}]& C_{03}C_{23}[C_{02}C_{22}]\end{array}\right|,$$ (44) $$c_1=C(1)c_0c_2.$$ (45) Equation (42) can have two roots $`\{Z_0^\pm \}`$, and only positive solutions are of interest. Having found the appropriate root of (42), one may solve any three of the four equations of the system (35) for the ratios $`D_{LL^{}}^M/D_{11}^M`$. These ratios may then be used to obtain an equation for $`D_{11}^M`$ that is analogous to the algebraic equation for the gap amplitude $`\mathrm{\Delta }_F`$ in the <sup>1</sup>S<sub>0</sub> problem, namely $$Z=\frac{[\chi _1^{11}(k)]^2}{2\sqrt{\xi ^2(k)+\overline{D^2}(k)}}d\tau .$$ (46) Solution of this equation yields the desired gap function in angle-average approximation. The angle-average approximation serves well if the goal is to calculate the gap amplitude $`\mathrm{\Delta }_F=[\overline{D^2}(k_F)]^{1/2}`$. However, if information is required on the spectrum of pairing solutions and the energy splittings between these solutions, it is essential to include the angle-dependent effects from the terms in eq. (20) going like $`\mathrm{ln}𝒟^2(𝐧)`$. By so doing, accurate results and valuable insights can be gained even if we stop at leading order in the small-parameter approximation. The situation is made more tangible by considering two simple solutions involving only one $`M`$ value (and its negative), specifically $`M=2`$ or $`M=0`$ (corresponding to Types 1 and 2, respectively). Working at lowest order in the small-parameter expansion, partial integration of the definition (26) of the integrals $`J_i`$ over the variable $`\xi `$ leads to the simplified form $$J_i=\frac{1}{2}_0^1_0^{2\pi }f_i(z,\phi )\mathrm{ln}D^2(z,\phi )\frac{\mathrm{d}z\mathrm{d}\phi }{2\pi },$$ (47) having made use of the equalities $`\varphi (\xi =0)=\chi (\xi =0)=1`$. With this result, eqs. (28) and (31) are recast, respectively, as $$1=vJ_5\left(\mathrm{\Delta }_F^{(M=2)}\right)\mathrm{and}1=v\left[J_0\left(\mathrm{\Delta }_F^{(M=0)}\right)+J_5\left(\mathrm{\Delta }_F^{(M=0)}\right)\right].$$ (48) The splitting between the states in question is determined by subtracting one equation from the other. Straightforward manipulations give $$\mathrm{ln}\left(\frac{\mathrm{\Delta }_F^{(2)}}{\mathrm{\Delta }_F^{(0)}}\right)^2=\frac{2}{3}\mathrm{ln}3+\frac{2\pi }{9\sqrt{3}}0.028.$$ (49) When the angle-average approximation is applied, eq. (28) or (31) is replaced by $$\frac{8\pi ^2ϵ_F}{vk_F^3}+_1^{\mathrm{}}dϵ\frac{\sqrt{ϵ+1}\varphi \chi }{2ϵ}+_1^1dϵ\frac{\sqrt{ϵ+1}\varphi \chi 1}{2|ϵ|}+\mathrm{ln}\frac{4}{\mathrm{\Delta }_F^2}\mathrm{ln}\frac{1}{8\pi }=0,$$ (50) which is to be solved for $`\mathrm{\Delta }_F=\mathrm{\Delta }_F^{\mathrm{aa}}`$. (Here, $`\varphi =\varphi (ϵ)`$ and $`\chi =\chi (ϵ)`$ after a change of integration variable from $`k`$ to $`ϵ=(k/k_F)^21`$.) Evaluating the ratios of $`\mathrm{\Delta }_F^{(2)}`$ and $`\mathrm{\Delta }_F^{(1)}`$ to $`\mathrm{\Delta }_F^{\mathrm{aa}}`$ along the lines of the above derivation of $`\mathrm{\Delta }_F^{(2)}/\mathrm{\Delta }_F^{(0)}`$, the ordering of the three gap amplitudes is seen to be $`\mathrm{\Delta }_F^{(2)}<\mathrm{\Delta }_F^{(0)}<\mathrm{\Delta }_F^{\mathrm{aa}}`$. This order agrees with the numerical results appearing in fig. 1 and with the findings of other authors . It is interesting and highly significant that the result (49) for the ratio of pairing energies for the Type 1 ($`|M|=0`$) and Type 2 ($`M=0`$) solutions – obtained to leading order in the small parameter of the theory yet very accurate—is completely independent of the parameters of the actual interaction and independent of density. In other words, it is universal. Proceeding to the next layer of complexity, let us consider the Type 4 solution, again studied in leading order. In this case, $`M=0`$ and $`|M|=2`$ substates coexist, and the system of equations that determines the coefficients $`D_{11}^{JM}`$ becomes $$\lambda _2=v\left[\lambda _2(J_0+J_5)J_3\right],$$ (51) $$1=v\left[\lambda _2J_3/3+J_5\right],$$ (52) where we have introduced the more compact notation $`\lambda _2=\delta _2\sqrt{6}/\delta _0`$ and the $`J_i`$ are given by (26). The first of these equations can be conveniently rewritten as $$0=\lambda _2J_0+(\lambda _2^2/31)J_3$$ (53) by subtracting $`\lambda _2`$ times eq. (52) from eq. (51). The alternative form (53) is distinctly advantageous: unlike $`J_5`$, the integrals $`J_0`$ and $`J_3`$ do not contain a term proportional to $`\mathrm{ln}\mathrm{\Delta }_F^2`$ and hence are independent of any details of interaction. Consequently, eq. (53) provides a closed equation for the parameter $`\lambda _2`$ whose solution is universal – i.e., the same for all pairing interactions. Once the universal parameter $`\lambda _2`$ is determined, the scale parameter $`\delta _0`$ is calculated from (52); its value does, of course, depend on the interaction. The formulas for $`J_0`$ and $`J_3`$ can be further simplified by partial integration of $`J_0`$ over $`z`$ and $`J_3`$ over $`\phi `$. The remaining integration over $`\phi `$ is performed analytically. After simple operations, we arrive at $$J_0(0,\lambda _2)=\frac{1}{3}+4_0^1\frac{z^2\mathrm{d}z}{f(z)},J_3(0,\lambda _2)=\frac{3+\lambda _2^2}{4\lambda _2}\frac{3}{8\lambda _2}_0^1f(z)dz,$$ (54) where $`f(z)=\left[(1+\lambda _2^2+(3\lambda _2^2)z^2)^224\lambda _2^2(1z^2)^2\right]^{1/2}`$. In general, these integrals are expressible in elliptic functions, but in exceptional cases they reduce to elementary integrals. This is exactly what happens when we seek roots of the system (51–(53). In addition to the root $`\lambda _2=0`$ corresponding to the $`M=0`$ state (the Type 2 solution), we find the roots $`\lambda _2=\pm 3`$ and $`\lambda _2=\pm 1`$. It should not be surprising that the roots are degenerate with respect to signs, since eq. (53) is invariant under reversal of the sign of $`\lambda _2`$. What is most surprising is that these roots are rational numbers. Rationality of the roots $`\lambda _2`$ is a consequence of a hidden symmetry of the problem that transcends the approximation we have employed. To substantiate this assertion, consider first the particular solution $`\lambda _2=3`$, for which the squared gap function (27) becomes $`D^2(k,x,z;\lambda _2=3)[\chi _1^{11}(k)]^2[43(x^2+z^2)]`$. The symmetry of this function with respect to $`x`$ and $`z`$ implies the relation $`3J_0(0,3)+2J_3(0,3)=0`$, since the combination $`3f_0+2f_3=6(x^2z^2)`$ changes sign on interchange of $`x`$ and $`z`$ whereas $`d^2`$ and other factors within the integrand of (26) are left unchanged. The above relation between $`J_0`$ and $`J_3`$ satisfies eq. (53). The essential point is that this relation derives from a simple symmetry of the problem; it is not predicated on any approximate scheme for the evaluation of the $`J_i`$. Now consider the root $`\lambda _1=0`$, for which $`D^2(k,x,y,z)[\chi _1^{11}(k)]^2(1x^2)`$. Since $`D^2(k,x,y,z)`$ is independent of $`y`$ and $`z`$, one may trivially integrate (26) over these variables. Simple algebra then gives $`3J_0(0,1)=2J_3(0,1)`$, and this relation also satisfies Eq. (53). We may therefore conclude that the ratio of $`M=2`$ and $`M=0`$ components in the Type 4 solution is independent of any input parameters including density, temperature, and the parameters specifying the pairing interaction. A general analysis along similar lines, again exploiting the simplifications produced by the separation method, reveals the fundamental structure of the full set of solutions of the BCS <sup>3</sup>P<sub>2</sub> pairing problem in neutron matter . Considering general multicomponent solutions, the allowed ratios of coefficients of different spin-angle matrices $`\widehat{G}_L^{JM}(𝐧)`$ in the expansion (1) of the gap matrix turn out to be universal. As in the above derivation, any reference to interaction, density, or temperature cancels out to very high accuracy in forming the ratios. The complete set of solutions for these ratios is obtained explicitly by solving a set of five coupled ‘algebraic’ equations, one of which serves to determine the energetic scale of the pairing effect. The solution set and the spectrum of pairing energies are found to be highly degenerate in character. The small splittings between different solutions can be found by manipulations similar to those performed above for the Type 1 and Type 2 solutions. In principle, analogous analytic arguments may be used to establish universalities of pairing in the much more difficult problem of liquid <sup>3</sup>He, where states with $`L=S=1`$ and $`J=0,1,2`$ contribute on an equal footing and the number of equations to be solved rises from five to nine. ## 6 Numerical applications: uncoupled and coupled channels The separation approach to BCS-type pairing problems which has been further developed and studied in this paper leads to considerable simplifications of the process of solving the gap equations in a generic pairing channel. Aided by the existence of a small parameter, this approach allows one to find the interval of existence of solutions without having to perform any iterations, and it allows one actually to find solutions without having to iterate a set of integral equations. Some iterations may be required to construct solutions, but they need be performed only on a set of algebraic equations for the amplitudes of the components of the gap matrix (quantities that are momentum-independent but dependent on the magnetic projection $`M`$). The “division of labor” that underlies the separation method thus facilitates the process of finding solutions; moreover, because of the reduced number of numerical operations, it permits higher accuracy to be achieved. In addition to these numerical advantages, the new method offers an efficient framework for fundamental analytic investigations of pairing, as has been demonstrated in sect. 4 and more systematically in refs. . A necessary step in testing our procedures for numerical solution of the triplet pairing problem in neutron matter is comparison with results obtained by other authors. As indicated in sect. 4, Takatsuka and Tamagaki (see also their review ) have studied the pure <sup>3</sup>P<sub>2</sub> pairing problem at some depth, classifying the types of solutions and generating numerical solutions by iteration. Subsequently, Takatsuka (see also ) implemented an iterative procedure for solution of the gap equations in the <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> coupled-channel case with fixed $`M=0`$. This early numerical work was based on several older potential models with soft (Gaussian) cores. The Takatsuka-Tamagaki calculations were repeated and supplemented by Amundsen and Østgaard . Baldo et al. introduced the idea of angular averaging in considering $`M=0`$ and $`|M|=2`$ solutions (i.e., Type 2 and Type 1 solutions) of the pure <sup>3</sup>P<sub>2</sub> case and reported preliminary numerical results for a separable approximation to the Argonne $`v_{14}`$ potential . More recently, Elgarøy et al. have performed uncoupled <sup>3</sup>P<sub>2</sub> and coupled <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> calculations with fixed $`|M|`$ (0 or 2) in $`\beta `$stable matter for the Bonn potential models designated A, B, and C . In the most recent microscopic study of triplet pairing, Baldo et al. obtained results for several realistic models of the nucleon-nucleon interaction, including four that may legitimately be regarded as phase-shift equivalent. The latter potentials – Argonne $`v_{18}`$ , CD-Bonn , and Nijmegen I and II – belong to a new generation of models that yield high-precision fits of the sanitized world supply of $`pp`$ and $`np`$ scattering data up to a bombarding energy of 350 MeV. All four produce fits with $`\chi ^2`$ per data point close to unity. In addition, gap calculations were performed for three “older” potentials – Argonne $`v_{14}`$, Paris , and Bonn B. The authors of this extensive study conclude that reliable predictions of the <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> pairing gap at densities above about 1.7 times the saturation density of symmetrical nuclear matter will require potential models that fit the nucleon-nucleon phase shifts up to 1 GeV. For the purpose of comparison with the pre-1990 results on the triplet pairing problem, we have chosen the OPEG (“one-pion-exchange Gaussian”) potential used in refs. , and specifically the version denoted OPEG $`{}_{}{}^{3}O1`$. This potential has been fitted to scattering data collected before 1970, and it cannot be regarded as a realistic model of the nucleon-nucleon interaction. However, it possesses all of the essential qualitative features of the interaction and describes them at a semi-quantitative level. Of local character, it contains central, tensor, spin-orbit ($`𝐋𝐒`$), and quadratic spin-orbit (e.g. $`(𝐋𝐒)^2`$) components, as written out in detail by Amundsen and Østgaard . To compare with the latest calculations , we have generated results for the Argonne $`v_{14}`$ and Argonne $`v_{18}`$ models. The diagonal pairing matrix elements of these two potentials are plotted in fig. 2. In presenting the numerical results, we employ two measures of the gap matrix, namely (i) the average quantity $`\mathrm{\Delta }_F\left[\overline{D^2}(k_F)\right]^{1/2}`$ defined via eq. (7) or eq. (34) and (ii) the gap component $`\mathrm{\Delta }_L^{JM}(k)`$, either as a function of the momentum modulus $`k`$ or evaluated at the Fermi momentum $`k_F`$. We note that when the pairing problem is solved in the angle-average approximation specified in sect. 5, the $`M`$ dependence of the latter measure becomes moot. Thus the index $`M`$ will often be suppressed, along with the fixed index $`J=2`$. The zeros of the characteristic determinant of the linearized version of eq. (20) provide us with candidate values for the upper critical Fermi momentum $`k_c`$. This is true whether we restrict ourselves to pure <sup>3</sup>P<sub>2</sub> pairing or address the coupled-channel <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> case. Examining first the uncoupled case, in fig. 3 we superimpose plots of the diagonal matrix elements $`v_{11}^{J=2}=V_{11}^{J=2}(k_F,k_F)`$ of the Argonne $`v_{18}`$ potential, the associated characteristic determinant, and the gap component $`\mathrm{\Delta }_1^{2M}(k_F)=\mathrm{\Delta }_1(k_F)`$ obtained in the angle-average approximation, all considered as functions of the Fermi momentum $`k_F`$. From the preceding formal developments it is evident that a zero of the potential implies a pole of $`\varphi ^{11}(k)`$ and hence a simple pole of the determinant. Such a pole is very prominent in fig. 3 and is naturally followed closely by a zero of the determinant. The position of this zero gives a value $`k_c3.6`$ fm<sup>-1</sup> for the upper critical $`k_F`$ value at which the gap closes, the corresponding upper critical density being $`\rho _c=k_c^3/3\pi ^2`$. On the plot, the gap amplitude is clearly negligible at this point. Because of the exponential falloff, it appears to be zero even below this point, but is in fact still finite over an interval in $`k_F`$ where $`v_{11}^{J=2}`$ has positive values. Fig. 4 displays the characteristic determinant in the coupled-channel case, again with the Argonne $`v_{18}`$ potential as input for the pairing force. The determinant is evaluated for three different effective-mass ratios $`m^{}=M_n^{}/M_n`$, the free normal-state single-particle spectrum $`\epsilon (k)=\mathrm{}^2k^2/2M_n`$ being replaced by $`\mathrm{}^2k^2/2M_n^{}`$ to simulate dispersive effects of the medium. For any of the chosen values of $`m^{}`$, the determinant has a pole near $`k_F=1.9`$ fm<sup>-1</sup> followed closely by a zero near $`k_F=2.2`$ fm<sup>-1</sup>, and a second pole near $`k_F=3.3`$ fm<sup>-1</sup> followed closely by a zero near $`k_F=3.5`$ fm<sup>-1</sup>. It is the second zero that should be identified with the upper critical density, since the adjacent pole corresponds to a zero of $`v_{11}^{J=2}`$. The point is that we are seeking to describe $`T=1`$ triplet-P-wave pairing as modified by the tensor coupling to the $`T=1`$ triplet-F channel. The first pole of the determinant corresponds to the zero of $`v_{33}^{J=2}`$. Thus the accompanying determinantal zero is not relevant to the task at hand, which is the identification of the upper critical point for triplet-P rather than triplet-F pairing. Generally, a lower bound on $`k_c`$ for pairing predominantly in a state with orbital angular momentum quantum number $`L`$ is given by the location of the pole that is produced by the vanishing of $`v_{LL}^J`$ (we assume that this occurs only at one density). Conventional wisdom gives $`v_{LL}^J<0`$ as a sufficient (but not necessary!) condition for a pairing instability in partial wave $`L`$. The indicated lower bound property is of course consistent with this statement, but we are able to provide a more incisive condition for the extent of the pairing instability in terms of the behavior of the characteristic determinant associated with the integral equation (20) (or alternatively, with the characteristic determinant of the integral equation (13) for the shape function $`\chi (k)`$ ). Concerning the dependence of $`k_c`$ on effective mass, it is seen in fig. 4 that a smaller $`m^{}`$ leads to a lower value of the upper critical Fermi momentum (cf. further discussion below). Numerical results for the OPEG interaction that may be compared with those of Takatsuka and Tamagaki and Amundsen and Østgaard (and with general findings of other authors) are displayed in fig. 1 and fig. 5. Fig. 1 refers to the uncoupled <sup>3</sup>P<sub>2</sub> case with $`m^{}=1`$ and shows results for the $`k_F`$ dependence of values of the energy-gap parameter $`\mathrm{\Delta }_F`$ corresponding, respectively, to the angle-average approximation (uppermost curve), the pure $`M=0`$ solution (middle curve), and the $`|M|=2`$ solution (lowest curve). The latter two curves are in reasonable agreement with results reported in refs. , having maxima near 2 MeV in the vicinity of $`k_F=2.2`$ fm<sup>-1</sup>. It is worth mentioning again that our calculations are in accord with the energy ordering of the $`M=0`$ and $`|M|=2`$ solutions found numerically in the early work ; moreover, they confirm the generic results that have been derived in sect. 5 using the separation method and the small parameter expansion. At $`k_F=2.2`$ fm<sup>-1</sup> the ratio predicted by eq. (49) is reproduced with better than 1% accuracy, and the energy order $`\mathrm{\Delta }_F^{(2)}<\mathrm{\Delta }_F^{(0)}<\mathrm{\Delta }_F^{\mathrm{aa}}`$ implied by eqs. (49) and (50) is preserved. Considering the near-degeneracies in the condensation energies for the different solutions, documented originally by Takatsuka and Tamagaki , these favorable comparisons warrant some confidence in the application of the angle-average approximation to the coupled-channel case. Introducing the coupling to the <sup>3</sup>F<sub>2</sub> channel, we have solved the triplet pairing problem for the OPEG and Argonne potentials in the angle-average approximation, using the procedure outlined in sect. 5. Fig. 5 shows the calculated momentum dependence of the gap components $`\mathrm{\Delta }_1^{2M}(k)\mathrm{\Delta }_1(k)`$ (solid curve) and $`\mathrm{\Delta }_3^{2M}(k)\mathrm{\Delta }_3(k)`$ (dashed curve) for the case of the OPEG potential and $`m^{}=1`$, at a Fermi momentum $`k_F=2.2`$ fm<sup>-1</sup> near the value yielding the maximum strength of triplet pairing. The locations of the maxima, minima, and nodes of these functions are in close agreement with results given in refs. , and the magnitudes are in general agreement. (It is to be noted that in his coupled-channel calculation, Takatsuka included only the $`M=0`$ magnetic substate.) Figs. 6, 7, and 8 depict the principal results of our coupled-channel <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> gap calculations for the more realistic potential models, $`v_{18}`$ and $`v_{14}`$. Fig. 6 illustrates the momentum dependence of the $`L=1`$ and $`L=3`$ components $`\mathrm{\Delta }_L^{2M}(k)`$ for the Argonne $`v_{18}`$ interaction (abbreviated as $`\mathrm{\Delta }_1(k)`$ and $`\mathrm{\Delta }_2(k)`$). As expected, the shapes of these gap functions are rather similar to those found for the OPEG interaction at the same density (cf. fig. 5). However, the overall scale of the pairing effect in the OPEG case is almost 50% larger than for the $`v_{18}`$ model. In fig. 7, we view the dependence of the gap component values $`\mathrm{\Delta }_2(k_F)`$ and $`\mathrm{\Delta }_3(k_F)`$ on the Fermi momentum $`k_F`$, as calculated in angle-average approximation for the $`v_{18}`$ interaction. Contrasting the results for $`\mathrm{\Delta }_1(k_F)`$ with and without coupling to the $`F`$ wave, it is seen that the tensor force acts to substantially increase the pairing effect, as has been demonstrated in greater or lesser measure by earlier investigators . Fig. 8 compares the $`k_F`$ dependence of the <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> energy-gap parameter $`\mathrm{\Delta }_F`$ for the three potentials investigated, the angle-average approximation being invoked in all cases. In all three cases, the gap measure $`\mathrm{\Delta }_F`$ peaks in the range between $`k_F=2`$ and 2.5 fm<sup>-1</sup>. With the free single-particle spectrum, i.e., $`m^{}=1`$, the maximum value reached by this quantity for the $`v_{14}`$ interaction is twice that attained for the more refined $`v_{18}`$ model. As is clear from its definition, $`\mathrm{\Delta }_F`$ contains direct contributions both from P-wave and F-wave pairing. Inspecting fig. 2, we observe that the greater pairing effect for the $`v_{14}`$ potential relative to $`v_{18}`$ can be traced to the greater attraction present for $`v_{14}`$ in both P- and F-wave $`J=2`$ channels over the relevant range of the momentum variable $`k`$. (The coupling matrix elements $`V_{13}(kk)`$ for the two potentials are essentially equivalent over the latter range.) Fig. 8 also demonstrates the greater pairing effect of the OPEG potential relative to $`v_{18}`$. Such discrepancies between the predictions obtained for different interaction models are not unexpected: due to the log singularity intrinsic to the BCS pairing phenomenon, the overall strength $`\mathrm{\Delta }_F`$ of the effect in the <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> channel is extremely sensitive to the behavior of the input pairing matrix elements at momenta in the vicinity of the Fermi surface. In turn, since the pairing matrix elements are constructed directly from the given free-space $`NN`$ potential model and since $`k_F(23)`$ fm<sup>-1</sup>, one must expect a sensitive dependence on properties of the bare interaction that are not determined, or only poorly determined, by fits of the $`NN`$ scattering data up to 350 MeV. This point has been developed thoroughly and convincingly in the recent work of Baldo et al. . In table 1, we report some explicit numerical values of $`\mathrm{\Delta }_F`$ obtained in our calculations based on the two “realistic” potential models, Argonne $`v_{14}`$ and $`v_{18}`$. A comparison of our numerical results for these potentials with the corresponding results of Baldo et al. is in order, restricting attention perforce to the case of free single-particle energies $`\epsilon (k)`$. First, we point out that the pairing matrix elements entering our formulas are larger than those appearing in ref. by a numerical factor $`\pi /2`$, due to different choices of normalization for the spherical Bessel functions. Upon accounting for this (immaterial) factor, the diagonal matrix elements plotted for the $`v_{18}`$ interaction in fig. 3 of ref. are apparently in excellent agreement with their counterparts in fig. 2. Second, it must be recognized that the quantity $`D^2(k)`$ defined in eq. (6) of ref. is actually twice the function $`D^2(k)`$ appearing in our treatment (see our eqs. (6) and (34)). Thus the gap measure $`\mathrm{\Delta }_F[\overline{D^2}(k_F)]^{1/2}`$ we have calculated is smaller than the $`\mathrm{\Delta }_F`$ quantity of Baldo et al. by a factor $`1/\sqrt{2}`$. The definition we have adopted is motivated by the desideratum of consistency with the earlier literature . In any case, the two calculations should agree once this scale factor is applied, since both calculations employ the angle-average approximation. The results are generally in good agreement; however, there are some modest but significant numerical discrepancies. For the $`v_{14}`$ interaction, the current results for $`\mathrm{\Delta }_F`$ versus $`k_F`$ are somewhat higher than those of ref. below the peak, and somewhat lower above. The position of the peak, occurring near $`k_F=2.3`$ fm<sup>-1</sup> in the calculation of Baldo et al., is shifted to a slightly smaller $`k_F`$ value, near 2.2 fm<sup>-1</sup>. The value of $`\mathrm{\Delta }_F`$ at the peak is higher in our calculation by roughly 5%. In the case of the $`v_{18}`$ interaction, the numerical agreement is excellent below the peak, but beyond the maximum the current results lie below those of ref. by an amount that can reach about 0.1 MeV (in the $`\mathrm{\Delta }_F`$ measure used by Baldo et al.). The origin of these relatively minor numerical disparities is not clear. We would, however, like to reiterate the concern expressed in refs. about the sensitivity of the numerical integration over $`k`$ in the triplet gap equations to the choice of mesh points near the Fermi momentum. The difficulty arises because the functions $`k^2\mathrm{\Delta }_1^{2M}(k)/E(k)`$ and $`k^2\mathrm{\Delta }_3^{2M}(k)/E(k)`$ are very strongly peaked at $`k_F`$. To handle this problem within the present treatment, we have integrated out the troublesome log-like part by hand and performed a numerical integration only for the remainder, which is much smoother and well-behaved. For the latter integration, we employ 288 points, dividing the $`k`$ domain into three intervals: $`[0,k_F]`$, $`[k_F,2k_F]`$, and $`[k_F,k_{\mathrm{max}}]`$, with $`k_{\mathrm{max}}=50`$ fm<sup>-1</sup>. Each interval contains 96 mesh points distributed in Gaussian fashion – very dense around the ends of the interval and sparse in the middle. In fig. 8 and table 1, we have included $`v_{18}`$ results not only for the baseline choice $`m^{}=1`$, but also for a value $`m^{}=0.78`$ of the effective-mass parameter conventionally thought to provide a sensible representation of the in-medium single-particle energy. This brings us to the last point that we would like to address in the current treatment of triplet pairing in neutron matter. Relative to the bare-mass case, the propagator renormalization implied by $`m^{}=0.78`$ yields a substantially smaller gap as well as a lower value for the upper critical density (see fig. 4). These effects can be readily understood in terms of a reduction of the density of states $`N(0)`$ near the Fermi surface, which is related to the effective mass by the familiar expression $`N(0)=k_FM_n^{}/\pi ^2\mathrm{}^2`$. In the singlet-S-wave problem, the approximate formula (67) derived in the appendix shows clearly the inverse exponential dependence of the leading component of the gap on the density of states in the case of a negative pairing interaction on the Fermi surface, as does the BCS weak-coupling formula in a more limited context. The same physical reasoning and similar mathematical connections apply in the triplet case. It is a common belief that at the higher densities where triplet pairing becomes important, the interactions with particles in the background neutron sea produce a considerably greater reduction from the bare mass than is the case at inner-crust densities where singlet pairing dominates . This conclusion rests on the results of the Brueckner-Hartree-Fock and correlated-basis calculations. However, in these calculations, the contribution of pion-exchange effects to the renormalization of the nucleon effective mass is ignored. The role of the pion degrees of freedom grows rapidly with increasing density, eventually giving rise to a rearrangement of the ground state (associated, in particular, with pion condensation ). The latest variational prediction of the density $`\rho _c`$ at which this phase transition sets in is based on a nuclear Hamiltonian containing the Argonne $`v_{18}`$ two-nucleon interaction (with boost corrections to account for the leading relativistic effects) and phenomenological three-nucleon interactions. The result $`\rho _c0.2`$ fm<sup>-3</sup> actually lies below the density at which triplet pairing is expected to reach its maximum strength. Straightforward considerations demonstrate that in the vicinity of this transition point the contribution of the pion-exchange diagram to the effective mass $`M_n^{}`$ diverges and $`m^{}`$ goes to infinity. Therefore it is not unlikely that the operative value of the effective mass in the domain of triplet pairing considerably exceeds the bare mass. From this standpoint, conventional treatments that assume an $`m^{}`$ value somewhat below unity may lose their relevance. ## 7 Prospectus Exploiting the separation approach and the existence of a small parameter, we have been able to clarify important general properties of triplet pairing in neuron matter and to establish a framework and explicit procedures for reliable numerical solution of the relevant BCS gap equations. This effort complements a recent exhaustive analysis of the full set of generic solutions of the pure (uncoupled) <sup>3</sup>P<sub>2</sub> problem, also performed with the aid of the separation strategy . In view of the corresponding advances that have been made in the <sup>1</sup>S<sub>0</sub> pairing problem , it can be reasonably claimed that the mathematical and analytical aspects of BCS pairing of identical fermions in <sup>1</sup>S<sub>0</sub> and <sup>3</sup>P<sub>2</sub> states are now well understood, both within the relevant nucleonic contexts and in a more universal, context-independent sense. The same cannot be said of the specific physical and mechanistic aspects of triplet or singlet pairing of strongly-interacting fermions, which are associated with the realistic in-medium particle-particle interaction and renormalized single-particle energies that should provide the raw material for the gap equations. A complete and realistic treatment of pairing in a given strongly-coupled Fermi system demands ab initio calculation of these essential inputs. The success of the method of correlated-basis (or CBF) theory across a broad range of many-body systems and phenomena suggests the following approach to the quantitative physics of pairing in extended nucleonic systems: 1. Dressing of the pairing interaction by Jastrow correlations within CBF theory 2. Dressing of the pairing interaction by dynamical collective effects within CBF theory (including polarization effects arising from exchange of density and spin-density fluctuations, etc.) 3. Consistent renormalization of single-particle energies by short- and long-range correlations within CBF theory (cf. ref. ) This approach has already been explored in the <sup>1</sup>S<sub>0</sub> neutron pairing problem , although the assumed Jastrow correlations have not been optimized and only a second-order CBF perturbation treatment is available for step (b). Application of this scheme to <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> pairing in neutron-star matter is a challenging but potentially rewarding prospect. The inclusion of other physical effects will also be important in achieving a realistic description of pairing in the neutron-star medium. Modifications due to three-nucleon interactions should be considered, the calculations should be performed in $`\beta `$-stable matter (cf. ref. ), and relativistic effects should be examined (see, for example, ref. ). As pointed out in sect. 6, the impact on pairing of mesonic degrees of freedom, especially through an enhancement of the nucleon effective mass in the vicinity of pion condensation, deserves close investigation. In conclusion, it is important to emphasize that although the specific calculations performed in the present work were based on pairing matrix elements computed directly from bare, in-vacuum $`NN`$ interactions and on quadratic single-particle energies, many of our findings are unaffected by these simplifications. Indeed, the analyses that have been performed (leading to the universal properties discussed here and in ref. ), as well as the calculational procedures that have been developed, will retain their applicability and essential validity when realistic in-medium inputs are employed in the gap equations. Within this in mind, further studies in the same spirit – aimed at developing accurate computational procedures and uncovering universal properties – are of considerable interest. As a continuation of the work of sect. 6, the separation-transformed BCS gap equations for pure neutron matter may be solved in the <sup>3</sup>P<sub>2</sub><sup>3</sup>F<sub>2</sub> coupled channel without resorting to the angle-average approximation, but still within the setting of the rapidly convergent small-parameter expansion, and with bare pairing matrix elements and unrenormalized single-particle energies. Another project that awaits conclusion is a perturbative treatment of the lifting of the universalities and degeneracies that were established for the uncoupled <sup>3</sup>P<sub>2</sub> pairing by the analysis carried out in ref. . Also, the separation analysis performed in that paper may be extended to explore the superfluid phase diagram of liquid <sup>3</sup>He – although this task is considerably more demanding than in the case of <sup>3</sup>P<sub>2</sub> pairing in neutron matter. Finally, application of the separation approach to studies of <sup>3</sup>S<sub>1</sub><sup>3</sup>D<sub>1</sub> pairing in symmetrical nuclear matter and in the expanding phase of heavy-ion collisions may produce new insights into these intriguing problems. ## 8 Acknowledgements This research was supported in part by the US National Science Foundation under Grant Nos. PHY-9602127 and PHY-9900713 and by the McDonnell Center for the Space Sciences. We thank M. Hjorth-Jensen, Ø. Elgarøy, and R. B. Wiringa for stimulating discussions and helpful communications. ## 9 Appendix In this appendix we review the application of the separation method to <sup>1</sup>S<sub>0</sub> pairing and display the first two approximants to the gap amplitude $`\mathrm{\Delta }_F\mathrm{\Delta }(k_F)`$ within a small-parameter expansion of the separated problem . Having set the pattern, we then sketch the extension of the small-parameter expansion to <sup>3</sup>P<sub>2</sub> pairing. To provide a baseline for the separation method, it is useful to revisit the problem of a separable pairing matrix $`V(k,k^{})=v_0\varphi (k)\varphi (k^{})`$ with $`\varphi (k_F)=1`$ and $`v_0V^{J=0}(k_F,k_F)0`$. The standard gap equation in the <sup>1</sup>S<sub>0</sub> channel, $$\mathrm{\Delta }(k)=\frac{V(k,k^{})\mathrm{\Delta }(k^{})}{2\sqrt{\xi ^2(k^{})+\mathrm{\Delta }^2(k^{})}}d\tau _0$$ (55) with $`\mathrm{d}\tau _0=k^2dk/2\pi ^2`$, becomes $$\mathrm{\Delta }(k)=v_0\varphi (k)\frac{\varphi (k^{})\mathrm{\Delta }(k^{})}{2\sqrt{\xi ^2(k^{})+\mathrm{\Delta }^2(k^{})}}d\tau _0.$$ (56) (One may note that the differential volume element $`\mathrm{d}\tau _0`$ entering eqs. (55)–(56) and other formulas below differs from the volume element $`\mathrm{d}\tau =(2/\pi )k^2dk`$ occurring in integrals elsewhere in this paper. The difference arises from a different choice of normalization in forming the pairing matrix elements. In this appendix, the S-wave pairing matrix elements are defined as in refs. , whereas the convention employed by Takatsuka and Tamagaki for general pairing matrix elements is adopted in the rest of the paper.) Since the integral on the r.h.s. of eq. (56) is just a number, we see that $`\varphi (k)`$ determines the shape of the gap function, while the integral determines its scale. Introducing a dimensionless shape function $`\chi (k)`$ such that $`\mathrm{\Delta }(k)\mathrm{\Delta }_F\chi (k)`$, the gap equation for this problem is equivalent to $`\chi (k)`$ $`=`$ $`\varphi (k),`$ (57) $`1`$ $`=`$ $`v_0{\displaystyle \frac{\varphi ^2(k)}{2\sqrt{\xi ^2(k)+\mathrm{\Delta }_F^2\varphi ^2(k)}}d\tau _0}.`$ (58) Thus, one has only to solve a nonlinear equation for the gap amplitude rather a nonlinear singular integral equation for the gap function. We note that a solution of this equation exists only when $`v_F`$ is negative, in which case the monotonicity of the r.h.s. implies that there exists one and only one solution. Now we turn to the general situation where $`V(k,k^{})`$ need not be separable. To gain some benefits of the separable case, let us split the pairing interaction into a separable part and a remainder that vanishes when either momentum variable is on the Fermi surface: $$V(k,k^{})=v_0\varphi (k)\varphi (k^{})+W(k,k^{}),$$ (59) where $`W(k_F,k)=W(k,k_F)0`$ and, as before, $`\varphi (k)=V(k,k_F)/v_F`$. Again writing $`\mathrm{\Delta }(k)=\mathrm{\Delta }_F\chi (k)`$, the gap equation (55) is readily seen to be equivalent to the integral equation for the shape function $`\chi (k)`$ $$\chi (k)+\frac{W(k,k^{})\chi (k^{})}{2\sqrt{\xi ^2(k)+\mathrm{\Delta }_F^2\chi ^2(k)}}d\tau _0=\varphi (k),$$ (60) together with the ‘algebraic’ equation $$1+v_0\frac{\varphi (k)\chi (k)}{2\sqrt{\xi ^2(k)+\mathrm{\Delta }_F^2\chi ^2(k)}}d\tau _0=0$$ (61) for the gap amplitude $`\mathrm{\Delta }_F`$ (assumed nonzero). It is important to note that since $`W(k,k_F)`$ is zero by construction, the integral equation (60) has a nonsingular kernel, the log-singularity of the original BCS formulation having been isolated in the amplitude equation (61). An iterative solution of this set of equations converges very rapidly. Indeed, for neutron matter at densities corresponding to the inner-crust region of a neutron star, a good approximation to the gap is already obtained by replacing $`\mathrm{\Delta }_F\chi (k)`$ in the denominator of (60) by a small constant scale factor $`\delta `$ (e.g. $`0.01\epsilon _F`$) solving (60) for $`\chi (k)`$ by matrix inversion, and substituting the result into (61) to obtain the scale factor $`\mathrm{\Delta }_F`$ by (say) Newton’s method. The existence of a small parameter $`d_F=\mathrm{\Delta }_F/\epsilon _F`$ for the problem plays a key role in the success of the separation method based on eqs. (60)–(61). Consider the equation (60) for the shape function $`\chi (k)`$. Since the Taylor expansion of the remainder function $`W(k,k^{})`$ around $`k^{}=k_F`$ starts with the first-order term $`\mathrm{d}W(k,k^{})/\mathrm{d}k^{}|_{k^{}=k_F}(k^{}k_F)`$, the integral over $`k^{}`$ in the symmetric vicinity $`[k_F\delta k,k_F+\delta k]`$ of this point vanishes for sufficiently smooth functions $`W(k,k^{})`$ and $`\chi (k)`$, $$\frac{^2W(k,k^{})}{k_{}^{}{}_{}{}^{2}}|_{k_F}\delta k\frac{W(k,k^{})}{k^{}}|_{k_F}\text{and}\frac{\mathrm{d}\chi (k)}{\mathrm{d}k}\delta k1.$$ (62) Also, if beyond this interval we have $`\mathrm{\Delta }_F\chi (k)\xi (k)`$, or in other words if $`\delta kd_Fk_F`$, it is a safe approximation to replace the $`\mathrm{\Delta }_F`$ term in the denominator of (60) by a rather arbitrary scale factor $`\delta `$. We thereby obtain a practically equivalent linear integral equation $$\chi (k)+\frac{W(k,k^{})\chi (k^{})}{2\sqrt{\xi ^2(k^{})+\delta ^2}}d\tau _0=\varphi (k),$$ (63) which is always soluble for well-behaved interactions. For most purposes, it is permissible simply to set $`\delta `$ equal to zero. The nature of this “$`\delta `$ approximation,” and more generally of expansions in the small parameter $`d_F`$, is developed more fully below. For realistic nuclear potentials, the accuracy of this approximation in the singlet-S problem is in fact much higher than indicated by the simple error estimate $`d_F^2\mathrm{ln}1/d_F`$ that accompanies the above preceding. The reason for this heightened accuracy is that the main contribution to the integral in (60) comes from a region well separated from the Fermi surface. The <sup>1</sup>S<sub>0</sub> pairing matrix elements of realistic nuclear potentials show maxima at $`k3`$ fm<sup>-1</sup>. Hence the contribution from the Fermi surface region, already small because of the structure of $`W(k,k^{})`$, becomes quite negligible in comparison. Based on Taylor-series expansion of the energy denominator of eq. (61), the gap amplitude $`\mathrm{\Delta }_F`$ may be systematically expressed as a combination of series of terms in $`(d_F^2)^n\mathrm{ln}1/d_F`$ and $`(d_F^2)^n`$ with integral $`n`$. Introducing the dimensionless variable $`ϵ(k/k_F)^21`$, eq. (61) is first recast as $$\frac{2}{N(0)v_F}+_1^1\frac{\sqrt{ϵ+1}\varphi (ϵ)\chi (ϵ)\mathrm{d}ϵ}{2\sqrt{ϵ^2+d_F^2\chi ^2(ϵ)}}+_1^{\mathrm{}}\frac{\sqrt{ϵ+1}\varphi (ϵ)\chi (ϵ)}{2\sqrt{ϵ^2+d_F^2\chi ^2(ϵ)}}dϵ=0.$$ (64) The integration over $`ϵ`$ has been divided two intervals, namely (i) the vicinity of Fermi surface $`ϵ[1,1]`$, where terms of both types indicated above will appear, and (ii) the tail $`ϵ[1,\mathrm{}]`$, where a simple series in $`d_F^2`$ is generated, $$_1^{\mathrm{}}\frac{\sqrt{ϵ+1}\varphi (ϵ)\chi (ϵ)\mathrm{d}ϵ}{2\sqrt{ϵ^2+d_F^2\chi ^2(ϵ)}}_1^{\mathrm{}}\frac{\sqrt{ϵ+1}\varphi (ϵ)\chi (ϵ)}{2|ϵ|}dϵ\frac{d_F^2}{2}\epsilon _F^2_1^{\mathrm{}}\frac{\sqrt{ϵ+1}\varphi (ϵ)\chi ^3(ϵ)}{2|ϵ|^3}dϵ+\mathrm{}.$$ (65) Extracting the logarithmic contribution from the second term of equation (64), we have $`{\displaystyle \frac{2}{N(0)v_F}}`$ $`+`$ $`\mathrm{ln}{\displaystyle \frac{2}{d_F}}+\mathrm{ln}{\displaystyle \frac{1}{2}}\left(1+\sqrt{1+d_F^2}\right)`$ (66) $`+`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _1^1}{\displaystyle \frac{\sqrt{ϵ+1}\varphi (ϵ)\chi (ϵ)1}{\sqrt{ϵ^2+d_F^2\chi (ϵ)^2}}}+{\displaystyle \frac{1}{2}}{\displaystyle _1^{\mathrm{}}}dϵ{\displaystyle \frac{\sqrt{ϵ+1}\varphi (ϵ)\chi (ϵ)}{\sqrt{ϵ^2+d_F^2\chi (ϵ)^2}}}`$ $`+`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _1^1}dϵ{\displaystyle \frac{\sqrt{ϵ^2+d_F^2}\sqrt{ϵ^2+d_F^2\chi ^2(ϵ)}}{\sqrt{ϵ^2+d_F^2}\sqrt{ϵ^2+d_F^2\chi ^2(ϵ)}}}=0.`$ The leading approximation to $`d_F`$, $$d_F^0=2\mathrm{exp}\left[\frac{2}{v_FN(0)}+\frac{1}{2}_1^1\frac{\mathrm{d}ϵ}{|ϵ|}\left(\sqrt{ϵ+1}\varphi (ϵ)\chi (ϵ)1\right)+\frac{1}{2}_1^{\mathrm{}}\frac{\mathrm{d}ϵ}{ϵ}\sqrt{ϵ+1}\varphi (ϵ)\chi (ϵ)\right],$$ (67) is obtained by setting $`d_F0`$ everywhere except within the first logarithmic term of eq. (66). The first term inside the square brackets is the usual BCS contribution and is clearly dominant if the interaction is weak and localized in momentum space around the Fermi surface. If – as is the case for nuclear forces – the localization requirement is not fulfilled, the contribution from the second term of eq. (67) must also be taken into account. One can then derive meaningful results even for potentials with positive $`v_0`$. The approximant (67) is offered as a useful generalization of the BCS weak-coupling formula to a broad class of pairing interactions. The first correction and higher corrections to the formula (67) may be developed as follows. Near the Fermi surface, the denominators appearing in eq. (66) are simplified by noting that $$ϵ^2+d_F^2\chi ^2(ϵ)=ϵ^2+d_F^2+d_F^2\left[\chi ^2(ϵ)1\right]$$ (68) can be expanded in powers of $`d_F^2\left[\chi ^2(ϵ)1\right]/(ϵ^2+d_F^2)`$. This is a small quantity: for $`ϵ>d_F`$ it is of order $`d_F^2/ϵ^2`$, while for $`ϵ<d_F`$ we can make the expansion $$\chi (ϵ)=1+\chi _0^{}ϵ+\frac{1}{2}\chi _0^{\prime \prime }ϵ^2+\mathrm{},$$ (69) where the subscript 0 stands for $`ϵ=0`$, and use $$\frac{2\epsilon d_F}{ϵ^2+d_F^2}1.$$ (70) (The prime indicates the derivative with respect to $`ϵ`$, while the subscript 0 stands for $`ϵ=0`$.) Accordingly, we invoke the expansion $$\frac{1}{\sqrt{ϵ^2+d_F^2\chi ^2}}=\frac{1}{\sqrt{ϵ^2+d_F^2}}\left[1\frac{1}{2}d_F^2\frac{\chi ^21}{ϵ^2+d_F^2}+\frac{3}{8}d_F^4\frac{(\chi ^21)^2}{(ϵ^2+d_F^2)^2}\mathrm{}\right].$$ (71) At this point, any smooth functions in the numerators of the integral terms in eq. (66) may be expanded in Taylor series. Since (i) the integrations in the fourth and sixth addends on the l.h.s. of (66) are performed on a symmetric interval and (ii) the denominators contain only even powers of $`ϵ`$, we need only consider terms with even powers of $`ϵ`$ in the corresponding numerators. For example, $$_1^1dϵ\frac{\sqrt{ϵ^2+d_F^2}\sqrt{ϵ^2+d_F^2\chi ^2(ϵ)}}{\sqrt{ϵ^2+d_F^2}\sqrt{ϵ^2+d_F^2\chi ^2(ϵ}}=$$ (72) $``$ $`{\displaystyle \frac{1}{2}}d_F^2{\displaystyle _1^1}dϵ{\displaystyle \frac{\chi ^21}{(ϵ^2+d_F^2)^{\frac{3}{2}}}}+{\displaystyle \frac{3}{8}}d_F^4{\displaystyle _1^1}dϵ{\displaystyle \frac{(\chi ^21)^2}{(ϵ^2+d_F^2)^{\frac{5}{2}}}}\mathrm{}`$ $`=`$ $`{\displaystyle \frac{1}{2}}d_F^2\left[(\chi _0^{})^2+\chi _0^{\prime \prime }\right]\left[2+2\mathrm{ln}{\displaystyle \frac{2}{d_F}}\right]+{\displaystyle \frac{3}{8}}d_F^44(\chi _0^{})^2\left[{\displaystyle \frac{2}{3}}{\displaystyle \frac{1}{d_F^2}}1\right]\mathrm{}.`$ This step reveals a slight complication in the analysis. The integral of the first term of the Taylor expansion of $`\chi ^21`$ vanishes, and consequently the integral of the second term of this Taylor expansion is of the same order as the integral of the term of the expansion (71) containing $`(\chi ^21)^2`$ are of the same order in $`d_F`$. In evaluating corrections to the result (67), it is convenient to define $$I_n^m_1^1dϵ\frac{ϵ^{2n}}{(ϵ^2+d_F^2)^{m+\frac{1}{2}}}$$ (73) and $$A_n(\sqrt{ϵ+1}\varphi \chi ^n)_0^{\prime \prime }.$$ (74) The first correction to $`d_F^0`$ is then determined by applying recursion relations involving the $`I_n^m`$ (see ref. ). Eq. (64) is written as $$0\mathrm{ln}\frac{2}{d_F^0}+\frac{1}{4}d_F^2B+\left(1\frac{1}{4}d_F^2A_3\right)\mathrm{ln}\frac{2}{d_F},$$ (75) where $`B`$ $`=`$ $`{\displaystyle _1^{\mathrm{}}}{\displaystyle \frac{\sqrt{ϵ+1}}{ϵ^3}}\varphi (ϵ)\chi ^3(ϵ)dϵ{\displaystyle _1^1}{\displaystyle \frac{1}{|ϵ|^3}}\left[\sqrt{ϵ+1}\varphi \chi ^31{\displaystyle \frac{1}{2}}ϵ^2A_3\right]dϵ`$ (76) $`{\displaystyle \frac{1}{4}}A_1+{\displaystyle \frac{1}{2}}A_3+{\displaystyle \frac{1}{4}}A_5+1,`$ and the improved estimate is $$d_Fd_F^0\left[1+\frac{1}{4}\left(d_F^0\right)^2\left(BA_3\mathrm{ln}\frac{2}{d_F^0}\right)\right]$$ (77) follows. A calculation for the Argonne $`v_{14}`$ interaction illustrates the utility and accuracy of the leading approximation (67) to the energy gap $`\mathrm{\Delta }_F=\epsilon _Fd_F`$ within the small-parameter expansion. In fig. (9), the $`k_F`$ dependence of $`\mathrm{\Delta }_F`$ as given by the zeroth-order approximation in $`d_F`$ is compared with the essentially exact result obtained by iteration of eqs. (60) and (61). Free single-particle energies ($`m^{}=1`$) are assumed. The estimate (67) produces excellent results, especially at densities beyond the peak of the $`\mathrm{\Delta }_F`$ curve. The same strategy of constructing approximants to the gap problem through small-parameter expansion is applicable to pairing in higher angular momentum states and in particular to <sup>3</sup>P<sub>2</sub> pairing. By way of illustration, consider the Type 1 ($`M=\pm 2`$) solution of the uncoupled-channel problem. The transformed gap equation (28) associated with this case has a form similar to that for <sup>1</sup>S<sub>0</sub> pairing, with differences arising from the anisotropy of the P-wave problem. A weight factor $`3(\mathrm{sin}^2\theta )/2=3(1\mathrm{cos}^2\theta )/2`$ appears in the integration over angles in eq. (28), and $`3\mathrm{\Delta }_F^2(1\mathrm{cos}^2\theta )/2`$ appears in place of $`\mathrm{\Delta }_F^2`$ inside the square-root energy denominator. Doing the $`k`$ integral first, we can directly exploit results and relations of sects. 4 and 5 (specifically, the first of eqs. (25) along with eqs. (26) and (36)). Ignoring corrections proportional to $`d_F^2=\mathrm{\Delta }_F/ϵ_F`$, the only additional term that needs to be considered, relative to the angle-average approximation, is $$d𝐧\frac{3}{2}(1\mathrm{cos}^2\theta )/\mathrm{ln}\left[\frac{3}{2}(1\mathrm{cos}^2\theta )\right].$$ (78) For the Type 2 solution (with $`M=0`$), the additional term has the slightly different expression (see the last of eqs.(25)) $$d𝐧\frac{1}{2}(1+3\mathrm{cos}^2\theta )/\mathrm{ln}\left[\frac{1}{2}(1+3\mathrm{cos}^2\theta )\right].$$ (79) The most salient points are that both these contributions are of order unity relative to the dominant logarithm in the gap and both are negative. We may thus conclude that the true gap will be smaller than that given by the angle-average approximation and that the correction to the gap value calculated in this approximation will generally be small.
warning/0001/astro-ph0001137.html
ar5iv
text
# 1 First Encounter ## 1 First Encounter Particle Physics today is a zoo with plenty of wild species very difficult to trace or capture. They are in general known under phyla WIMPs, LSP, Axions, Higgs, Heavy Neutrinos, etc. Most of these species are potential candidates of the Dark Matter. Recently a new phylum called WIMPZILLA has been added to this zoo. The only common characteristic of the members of this family is their enormous mass, close to GUT scale $`10^{16}GeV`$ and their presumed long lifetime, much larger than the age of the Universe. Theoretical motivations for existence of these particles is not very new. Since early 90s, some compactification scenarios in string theory have predicted composite particles (e.g cryptons) with large symmetry groups and $`M10^{14}GeV`$. M-theory provides better candidates if compactification scale is much larger than Standard Model weak interaction scale . Messenger bosons in soft supersymmetry breaking models (see for review) also can have close to GUT scale masses. Being composite and decaying only through non-renormalizable interactions or having discrete gauge symmetry can make these ultra heavy particles meta-stable. Parametric resonance or vacuum fluctuation at the end of inflation can produce a large amount of UHDM and unitarity constraint on their mass can be overcome if they have never been thermalized . If there was not other motivation for existence of an ultra heavy meta-stable particle, it was just one of many predictions of Particle Physics models waiting for detection. However, the discovery of Ultra High Energy Cosmic Rays (UHECRs) by large Air Shower detectors (For a review of UHECRs detection and observed properties see ) is an observational motivation. The predicted GZK cutoff in the spectrum of CRs at energies around $`10^{18}eV`$ due to interaction with CMBR and IR photons restricts the distance to sources to less than $`2030Mpc`$. At present there are about $`600`$ events with $`E10^{19}eV`$, $`50`$ with $`E>4\times 10^{19}eV`$ and 14 events with $`E>10^{20}eV`$ including one with $`E10^{21}eV`$ . The UHECR spectrum has a local minimum around $`E10^{19}eV`$ but it rises again at higher energies. Composition of the primary particles can be estimated from the shower and its muon content maximum position and their elongation rate in the atmosphere. In spite of uncertainties and dependence on hadrons interaction models at high energies , all analyses of the data are compatible with a composition change from iron nuclei to proton at $`E>10^{18}eV`$ . Based on theoretical arguments however, some authors suggest that events with highest energies can be produced by heavy nuclei (But see also for maximum fly distance of $`Fe`$ nuclei). Some of ideas about the origin of UHECRs has been reviewed in and references therein. For the sake of completeness here we briefly review arguments for and against conventional and exotic sources. Classical Candidates: In a recent review of conventional candidates of UHECR sources , R. Blandford rules out practically all of them. It is expected that charged particles are accelerated in the shock waves of AGNs, SNs, or in-falling gas in rich galaxy clusters. Their maximum energy is proportional to the magnetic field strength and the size of the accelerator. Using this simple estimation, remnants of old SNs are expected to accelerate protons up to $`10^{15}eV`$ and iron to $`10^{18}eV`$ . A $`100TeV`$ $`\gamma `$-ray is expected from synchrotron radiation of these protons but nothing has been observed . Shock front of AGNs relativistic jets can accelerate protons to $`E10^{20}eV`$ . Central black hole in Fannaroff-Riley galaxies or the remnant of QSO can produce UHECRs up to $`E10^{20}eV`$ with a marginal maximum energy of $`4\times 10^{21}eV`$ . Adiabatic expansion and in situ interaction with radiation and matter fields however reduce the achievable maximum energy by orders of magnitude. Abrupt termination of the acceleration zone or composition change from $`p`$ to $`n`$ increases the chance of particles to keep their energies. The former however needs a fine tuning of the source structure and in the latter case particles lose part of their energy anyway. Gamma Ray Bursts (GBR) also have been proposed as a candidate source of UHECRs . A simulation of cosmological distribution of sources with a power law flux of UHECRs shows however that the expected flux on Earth is much lower than observed value. In it is argued that Poisson noise in the process of proton interaction with background photons leaves a non-interacting tail in the flux of UHECRs and increases the probability of detecting UHECRs from further distances. Optical depth of protons around GZK cutoff can be roughly estimated by $`\tau _{opt}\sigma n_{cmb}`$. For $`\sigma 0.45mb`$ close to the resonance, the probability of non-interacting in a distance of $`30Mpc`$ is at most $`10^8`$. Correlation with Astronomical Objects: At present no serious astronomical candidate source has been observed. A correlation between Super Galactic Plane and UHECR events direction has been claimed , but ruled out by other analyses . It is plausible that the apparent clustering of events reported by AGASA collaboration originates from caustics generated by the galactic magnetic field . Most of UHECRs burst models predict only one important nearby source . M87 in Virgo Cluster is practically the only conventional candidate that respects the distance constraint and is probably able to accelerate protons to ultra high energies. Recently, it has been shown that galactic wind and its induced magnetic field can deflect protons with $`E>10^{20}eV`$ if the magnetic field is as high as $`7\mu G`$. In this case, most of UHECR events point to Virgo Cluster . Even with such strong magnetic field, this model can correlate most energetic events with M87/Virgo only if the primaries are $`He`$ nuclei. There are also claims of correlation between the direction of UHECRs and pulsars or QSOs . The latter is reliable only if ultra high energy $`\nu `$s produce protons by interacting with relic or a halo of $`\nu `$ . Recent works however rule out a large part of the parameter space. Close to uniform distribution of UHECRs events has been concluded to be the evidence that UHECRs originate from some extra-galactic astronomical objects and not from a decaying UHDM in the Galactic Halo . MACHOs observation however shows that the Halo has a heterogeneous composition and a precise modeling of the anisotropy must take into account the distribution of various components. Exotic Sources: Ultra Heavy particles can be either long life DM or short life particles produced by the decay of topological defects . In the latter case, the heavy particles decay in their turn to ordinary species and make the observed UHECRs. The result of the decay of a heavy short life particle and a heavy relic can be quite the same, but their production rate and its cosmological evolution are very different and depend on the defect type and model . The possibility that topological defects be the source of UHECRs has been studied extensively . Nonetheless, the observed power spectrum of LSS and CMB anisotropies rule out the existence of large amount of defects in the early Universe and consequently the chance that UHECRs be produced by their decay (there are also limits from high energy $`\nu `$ ). Another proposed source of UHECRs is the evaporation of primordial black holes (PBH). The Hawking temperature at the end of their life is enough high to produce extremely energetic elementary particles like quarks and gluons and thus UHECRs. Most of models for production of PBH however needs fine tuning. Moreover, they are mainly produced when a large over density region crosses the horizon . PBHs with present temperature of the same order as UHECRs energy must have an initial mass of $`10^{14}10^{15}gr`$ and thus formed when the temperature of the Universe was $`10^9GeV`$. A thermal inflation at EW scale would have reduced their density $`10^{12}`$ times. In models with low reheating temperature, at these scales the Universe is not yet thermalized and parametric resonance and fluctuation production happens only at superhorizon scales . Summarizing the discussion of this section, it seems that conventional sources of Cosmic Rays are not able to explain the observed rate of UHECRs . Between exotic sources, the decay of a meta-stable ultra heavy particle seems to be the most plausible one. ## 2 WIMPZILLA Decay and Energy Dissipation of Remnants The decay of UHDM can have important implications for the evolution of high energy backgrounds. This can also be used for verifying this hypothesis and constraining the mass and lifetime of these particles. A number of authors have already tried to estimate the possible range of parameters as well as the flux of remnants on Earth (see and for defects, and and for UHDM). In the present work more interactions have been included in the energy dissipation of UHDM remnants. Hadronization is implemented using PYTHIA Monte Carlo program . We consider two distributions for UHDM. First we study the evolution of the spectrum of stable particles i.e. $`e^\pm ,p^\pm ,\nu ,\overline{\nu }`$ and $`\gamma `$ in a homogeneous universe from photon decoupling to today. Then to show the effect of matter clumpiness, we simulate the Galactic Halo by simply considering a uniform over-density. A complete treatment of a clumpy universe will be reported elsewhere. We also estimate the effect of very slow decay of DM on the equation of state of the Universe and on the baryon and lepton asymmetries. In this section we describe the decay model of UHDM and interactions which are included in the simulation. ### 2.1 Decay Model Theoretical predictions for mass and lifetime of UH particles cover a large range of values $`m_{UH}=10^{22}10^{26}eV`$ and $`\tau _{UH}=10^710^{20}yr`$ . Nevertheless, at the end of inflation it is more difficult to produce the highest range of the masses and special types of inflationary models are needed. For UH particles make a substantial part of the Dark Matter today, their lifetime must be at least comparable to the age of the Universe. In this work we perform the simulation for $`m_{UH}=10^{22}eV`$ and $`m_{UH}=10^{24}eV`$. For lifetime, $`\tau _{UH}=5\tau _0`$ and $`\tau _{UH}=50\tau _0`$, where $`\tau _0`$ is the age of the Universe, are studied. These values are smaller than what have been used by other groups . We show below that taking into account a realistic model for energy dissipation of remnants, even these relatively short lifetime can not explain the flux of UHECRs in a homogeneous universe. For some halo and IR background models, these lifetimes or slightly larger ones are compatible with observations. The decay modes of UHDM are very model dependent. It is very likely that they don’t decay directly to known particles and their decay has a number of intermediate unstable states that decay in their turn. It is also very probable that remnants include stable WIMPs which are not easily observable. To study the maximal effects of the decay on high energy backgrounds, we assume that at the end, the whole decayed energy goes to stable visible particles. Most of WIMPZILLA models consider them to be neutral bosons. Due to lack of precise information about their decay, we assume that it looks like the decay of $`Z^{}`$. Theoretical and experimental arguments show that leptonic and hadronic decay channels of $`Z^{}`$ have a branching ratio of $`1/32/3`$ . As hadronic channel is dominant, here we only consider this mode. It maximizes the flux of nucleons which at present are the dominant observable at ultra high energies. To mimic the softening of energy spectrum due to multiple decay level, we assume that decay is similar to hadronization of a pair of gluon jets. Experimental data as well as MLLA (Modified Leading Logarithm Approximation) , LPHD (Local Parton-Hadron Duality) ( and references therein) and string hadronization model predict a softer spectrum with higher multiplicity for gluon jets than for quarks. We use PYTHIA program for jet hadronization. This program, like many other available ones, can not properly simulate ultra high energy events, not only because we don’t know the exact physics at $`10^{16}GeV`$ scale, but also because of programming limits. For this reason, we had to extrapolate simulation results for $`E_{CM}10^{20}eV`$ up to $`E_{CM}=10^{24}eV`$. Fig.1 shows as an example, proton and photon multiplicity in hadronization of a pair of gluon jets. At middle energies, the multiplicity per $`\mathrm{log}(E)`$ is roughly constant. The same behavior exists for other species. This is a known shortcoming of present fragmentation simulations (See also Appendix 1) and makes spectrum harder at middle energies. As we would like to study the maximal flux of UHECRs and their effects on high energy backgrounds, this problem can not change our conclusions. In the simulation, all particles except $`e^\pm ,p^\pm ,\nu ,\overline{\nu }`$ and $`\gamma `$ decay. We neglect neutrinos mass and for simplicity we assume only one family of neutrinos i.e. $`\nu _e`$. Contribution of the stable species in the total multiplicity and the total decay energy is summarized in Table 1. For all species, more than $`99\%`$ of the total energy belongs to the particles with energies higher than $`10^{20}eV`$ and $`10^{18}eV`$ respectively for two masses considered here. Apparently the mass of UHDM has little effect on the composition of remnants. However, one has to admit an uncertainty about this conclusion which is a direct consequence of the uncertain behavior of the multiplicity spectrum as mentioned above and in the Appendix 1. ### 2.2 Interactions We have included roughly all relevant interactions between remnants (except $`\nu \nu `$ and $`\overline{\nu }\overline{\nu }`$ elastic scattering) to the simulation either analytically or by using the results of PYTHIA Monte Carlo. Previous works either don’t consider the energy dissipation or take into account only the first order perturbative interactions (except for $`p\gamma `$ where a fitting is used for $`N\gamma N\pi `$ cross-section). An early version of the present simulation studied only energy dissipation of UHECRs using PYTHIA without considering low energy processes and interactions of secondary particles. It found a higher lifetime for UHECRs than present work. The main reason for considering only first order interactions is that it is usually assumed that interaction with CMB is dominated by the minimal process i.e $`N\gamma N\pi `$. However, the CMB spectrum even on its peak spans over a wide range of energies where radiation correction and hadronization become important. For instance, at $`E_{CM}=4GeV`$ the mean multiplicity is $`15`$ in place of 5 (after pion decay) in the minimal interaction. Moreover, in the galactic medium, the IR and visible radiations are comparable with CMB and play an important role in the energy dissipation of protons of $`E10^{18}10^{19}eV`$. In extragalactic medium, the number density of background high energy photons with ($`E>1eV`$) is larger than visible and near IR. Another factor which accelerates energy dissipation of protons in the interaction with high energy photons is energy loss of leading proton in $`p\gamma `$ interaction. It increases with energy (See Fig.2) and results a higher dissipation rate. PYTHIA can not simulate processes with invariant CM energies $`E_{CM}\sqrt{s}`$, $`E_{CM}<2GeV`$ to $`E_{CM}<4GeV`$ (the lower limit depends on the interaction). Consequently, for smaller energies we have included only perturbative, first order interactions using analytical expressions. Table 3 summarizes processes which are included in the simulation, the energy range of analytical and/or Monte Carlo calculation of the cross-sections, and the cuts used for removing singularities at small energies or angles in the case of analytical calculation (results depend somehow on these cuts specially in the case of moderate energy and angular resolution of our program). For energy ranges that PYTHIA has been used, in general we use default value of various parameters of the program as defined in PYTHIA manual. A number of parameters have been changed for all processes e.g. to make unstable particles like mesons decay etc. They are listed in Table 2. For some interactions, the value of a few other parameters are changed. They are also listed in the Table 3. From now on $`s`$, $`t`$ and $`u`$ are Mandelstam variables. At very high energies, $`E_{CM}>10^{14}eV`$ for nucleon-nucleon and $`E_{CM}>10^{15}eV`$ for other interactions, PYTHIA becomes very slow and the number of rejected events increases rapidly. For these energies we perform a linear extrapolation from lower energies as explained below. ## 3 Evolution We assume that non-baryonic Dark Matter is totally composed of slowly decaying UH particles. From (2) and (4) below it is evident that width and fraction of UHDM in DM are degenerate and in the evolution equations, decreasing contribution is equivalent to increasing lifetime. Boltzmann equation for space-time and energy-momentum distribution of a particle $`i`$ is (We use units with $`c=\mathrm{}=1`$): $`p^\mu _\mu f^{(i)}(x,p)(\mathrm{\Gamma }_{\nu \rho }^\mu p^\nu p^\rho e_iF_\nu ^\mu p^\nu ){\displaystyle \frac{f^{(i)}}{p^\mu }}`$ $`=`$ $`(𝒜(x,p)+(x,p))f^{(i)}(x,p)+𝒞(x,p)+`$ (1) $`𝒟(x,p)+(x,p).`$ $`𝒜(x,p)`$ $`=`$ $`\mathrm{\Gamma }_im_i.`$ (2) $`(x,p)`$ $`=`$ $`{\displaystyle \underset{j}{}}{\displaystyle \frac{1}{(2\pi )^3g_i}}{\displaystyle 𝑑\overline{p}_jf^{(j)}(x,p_j)A(s)\sigma _{ij}(s)}.`$ (3) $`𝒞(x,p)`$ $`=`$ $`{\displaystyle \underset{j}{}}\mathrm{\Gamma }_jm_j{\displaystyle \frac{1}{(2\pi )^3g_i}}{\displaystyle 𝑑\overline{p}_jf^{(j)}(x,p_j)\frac{d_{}^{(i)}{}_{j}{}^{}}{d\overline{p}}}.`$ (4) $`𝒟(x,p)`$ $`=`$ $`{\displaystyle \underset{j,k}{}}{\displaystyle \frac{1}{(2\pi )^6g_i}}{\displaystyle 𝑑\overline{p}_j𝑑\overline{p}_kf^{(j)}(x,p_j)f^{(k)}(x,p_k)A(s)\frac{d\sigma _{j+ki+\mathrm{}}}{d\overline{p}}}.`$ (5) $`x`$ and $`p`$ are coordinate and momentum 4-vectors; $`f^{(i)}(x,p)`$ is the distribution of species $`i`$; $`m_i`$, $`e_i`$ and $`\mathrm{\Gamma }_i`$, are its mass, electric charge and width $`=1/\tau _i`$, $`\tau _i`$ is the lifetime; $`\sigma _{ij}`$ is the total interaction cross-section of species $`i`$ and species $`j`$ at a fixed $`s`$; $`\frac{d\sigma _{j+ki+\mathrm{}}}{d\overline{p}}=\frac{(2\pi )^3Ed\sigma }{g_ip^2dpd\mathrm{\Omega }}`$ is the Lorantz invariant differential cross-section of production of $`i`$ in the interaction of $`j`$ and $`k`$; $`g_i`$ is the number of internal degrees of freedom (e.g. spin, color); $`d\overline{p}=\frac{d^3p}{E}`$. We treat interactions classically, i.e. we consider only two-body interactions and we neglect the interference between outgoing particles. It is a good approximation when the plasma is not degenerate. It is assumed that cross-sections include summation over internal degrees of freedom like spin; $`\frac{d_{}^{(i)}{}_{j}{}^{}}{d\overline{p}}`$ is the differential multiplicity of species $`i`$ in the decay of $`j`$; $`\mathrm{\Gamma }_{\nu \rho }^\mu `$ is the connection; $`F_\nu ^\mu `$ an external electromagnetic field; and finally $`(x,p)`$ presents all other external sources. $`A(s)`$ is a kinematic factor : $$A(p_i,p_j)=((p_i.p_j)^2m_i^2m_j^2)^{\frac{1}{2}}=\frac{1}{2}((sm_i^2m_j^2)^24m_i^2m_j^2)^{\frac{1}{2}}.$$ (6) The quantity $`A\sigma `$ presents the probability of an interaction. In a homogeneous universe $`f(x,p)=f(t,|p|)`$ and in (1) the term corresponding to interaction with external electromagnetic field is zero. Therefore, to have a consistent formalism for evolution of distribution of all species, we don’t include the synchrotron radiation of high energy electrons in a magnetic field. We only consider the evolution of stable particles and slowly decaying UHDM. The term (2) concerns only UHDM. In (4), the only non-zero term in the sum is the decay of UHDM. We assume that stable species don’t have any interaction with UHDM and corresponding interaction integrals in (3) and (5) are zero. Due to the very large mass of UHDM, its momentum is negligible and we can assume that in comoving frame it is at rest. This permits to use its number density $`n_{dm}`$ which is more convenient for numerical calculation. In a homogeneous universe the metric in comoving frame is: $$ds^2=dt^2a^2(t)\delta _{ij}dx^idx^j.$$ (7) and with respect to local Lorantz frame (1) to (5) take the following form (in the following the species index indicates one of the stable species): $`{\displaystyle \frac{f^{(i)}(t,p)}{t}}{\displaystyle \frac{\dot{a}}{a}}p{\displaystyle \frac{f^{(i)}}{p}}`$ $`=`$ $`{\displaystyle \frac{1}{E}}((t,p)f^{(i)}(t,p)+𝒞(t,p)+𝒟(t,p)).`$ (8) $`(t,p)`$ $`=`$ $`{\displaystyle \underset{j}{}}{\displaystyle \frac{1}{(2\pi )^2g_i}}{\displaystyle 𝑑p_j\frac{p_j^2}{E_j}f^{(j)}(t,p_j)d(\mathrm{cos}\theta _{ij})A(s)\sigma _{ij}(s)}.`$ (9) $`𝒞(t,p)`$ $`=`$ $`{\displaystyle \frac{E}{4\pi g_ip^2}}\mathrm{\Gamma }_{dm}n_{dm}{\displaystyle \frac{d^{(i)}}{dp}}.`$ (10) $`𝒟(t,p)`$ $`=`$ $`{\displaystyle \underset{j,k}{}}{\displaystyle \frac{1}{(2\pi )^5g_i}}{\displaystyle 𝑑p_j𝑑p_k\frac{p_j^2}{E_j}\frac{p_k^2}{E_k}f^{(j)}(t,p_j)f^{(k)}(t,p_k)d(\mathrm{cos}\theta _{jk})d(\mathrm{cos}\theta _{ji})}`$ (11) $`d\varphi _iA(s){\displaystyle \frac{d\sigma _{j+ki+\mathrm{}}}{d\overline{p}}}.`$ $`{\displaystyle \frac{dn_{dm}}{dt}}+{\displaystyle \frac{3\dot{a}}{a}}n_{dm}`$ $`=`$ $`\mathrm{\Gamma }_{dm}n_{dm}.`$ (12) In (9) and (11) $`s`$ depends on the angle between species $`j`$ and $`k`$, and $`j`$ and $`i`$. Consequently, it is not possible to use cross-sections integrated over angular variables. Evolution of $`a(t)`$ is ruled by Einstein equation: $`{\displaystyle \frac{\dot{a}^2}{a^2}}`$ $`=`$ $`{\displaystyle \frac{8\pi G}{3}}T_{00}+{\displaystyle \frac{\mathrm{\Lambda }}{3}}.`$ (13) $`T^{00}(t)`$ $`=`$ $`{\displaystyle \underset{i}{}}{\displaystyle \frac{g_i}{2\pi ^2}}{\displaystyle 𝑑pp^2Ef^{(i)}(t,p)}.`$ (14) In a homogeneous cosmology, $`T^{00}`$ in Local Lorantz frame is the same as comoving frame and $`T_{comov}^{ii}=a^2T_{Loc.Lor.}^{ii}`$. Equations (8) and (13) determine the cosmological evolution of species. Due to interaction terms, even in a homogeneous universe, these equations are non-linear and coupled. It is not therefore possible to solve them analytically. Evolution equation for DM can be solved analytically for a short period of time. For other species if in (8) we consider absorption and production integrals as $`t`$ and $`p`$ dependent coefficients of a linear partial differential equation, (8) can be solved analytically . Giving the value of $`f^{(i)}(t,p)`$, $`n_{dm}(t)`$ and $`a(t)`$, at time $`t`$, we can then determine $`a(t+\mathrm{\Delta }t)`$, $`n_{dm}(t+\mathrm{\Delta }t)`$, $`(t,p)`$, $`𝒞(t,p)`$ and $`𝒟(t,p)`$ for a short time interval $`\mathrm{\Delta }t`$. $`f^{(i)}(t+\mathrm{\Delta }t,p)`$ would be obtain from solution of partial differential equation (8) using difference method. The solution of metric and distributions in one step of numerical calculation are the followings: $`a(t+\mathrm{\Delta }t)`$ $`=`$ $`a(t)\mathrm{exp}(\mathrm{\Delta }t({\displaystyle \frac{8\pi G}{3}}T_{00}+{\displaystyle \frac{\mathrm{\Lambda }}{3}})^{\frac{1}{2}}).`$ (15) $`n_{dm}(t+\mathrm{\Delta }t)`$ $`=`$ $`n_{dm}(t){\displaystyle \frac{a^3(t_0)}{a^3(t)}}\mathrm{exp}({\displaystyle \frac{tt_0}{\tau }}).`$ (16) $`f^{(i)}(t+\mathrm{\Delta }t,p)`$ $`=`$ $`(f^{(i)}(t,p^{})+\mathrm{\Delta }t(𝒞(t,p^{})+𝒟(t,p^{})))\mathrm{exp}((t,p^{})\mathrm{\Delta }t)`$ (17) $`p^{}`$ $`=`$ $`{\displaystyle \frac{a(t+\mathrm{\Delta }t)p}{a(t)}}.`$ (18) This prescription is more precise than a pure numerical calculation using e.g. difference method. ## 4 Numerical Simulation What makes numerical calculation of (15) to (17) difficult is the extension on roughly $`34`$ orders of magnitude of energy from $`10^9eV`$ (radio background) to $`10^{24}eV`$ (mass of UHDM) (from now on we call this energy range $`_E`$). Physical processes in this vast energy range have varieties of behavior, resonances, etc. Moreover, species have distributions which are orders of magnitude different from each others. In other term, these equations are very stiff. Semi-analytic method explained above helps to increase the precision of the numerical calculation. However, the 5-dimensional integration in (11) is extremely time and memory consuming and it is impossible to determine it with the same precision. In the following we describe in detail the numerical calculation of (15) to (17), as well as cosmological model, initial conditions and backgrounds which have been used. ### 4.1 Multiplicity and Cross-section For processes simulated by PYTHIA, we need to calculate total and differential cross-sections (see (9) and (11)). The former is given by the program itself. To determine the latter, we divide $`_E`$ to logarithmic bins (one per order of magnitude) and classify particles according to their momentum. The angular distribution of produced particles with respect to the axis of incoming particles in CM also is divided linearly to 90 bins. At a given $`s`$, the cross section in each bin $`\mathrm{\Delta }\sigma _{ij}=\sigma _{tot}N_{ij}/N`$. $`N_{ij}`$ is the number of particles of a given species in the bin $`ij`$. $`N`$ is the total number of simulated events. The same procedure is used for determination of $`\frac{d}{dp}`$. In this case it is not necessary to consider the angular distribution because in the rest frame of WIMPZILLA the decay has a spherical symmetry. Because PYTHIA can not cover the totality of the energy range, for high energy bins we use a linear extrapolation in $`\mathrm{log}p`$. The contribution of these energies i.e. $`E_{CM}10^6GeV`$ on the evolution of species is nevertheless small because the density of concerning particles is very low. The reason for adding them is not to have an artificial cut in the calculation. As mentioned above, for most processes including $`p\gamma `$, PYTHIA can not simulate the interaction with $`E_{CM}<4GeV`$. $`p\gamma `$ is the most important process for the energy dissipation of protons specially in this uncovered energy range where interaction with CMB photon is concentrated. In these energies we use directly the total cross-section obtained from experience . For differential cross-section, we extrapolate angular distribution from higher energies and normalize it to the exact total cross-section. ### 4.2 Evolution Equations In (9) and (11), the integrals over angular degrees of freedom can be separated from energy integrals and they don’t depend on any cosmological or DM parameter. It is therefore very convenient to calculate them separately. We divide each $`180^{}`$ interval to $`9`$ bins and use trapezoid method for integration. For the single integral in (9) a better resolution with $`90`$ bins has been used. However, our tests show that even the moderate resolution of $`9`$ bins gives, up to a few percents, the same results as the more precise integration. This is a reassuring results and means that the triple integrals in production term also must be enough correct even with a low resolution. Calculation of production term for first-order interactions is more complicate. They are processes of type $`2part.2part`$. In the CM frame where all cross-sections are determined, analytically or numerically, the incoming and outgoing particles have the same momentum. This is equivalent to having a Delta function in the integrand. The numerical realization of this function specially with a moderate resolution is very difficult. Consequently, one has to analytically absorb this function into integrand. Details of the calculation can be found in Appendix 2. Numerical solution of evolution equation itself needs much better energy resolution due to stiffness of distributions. The resolution must be at least comparable to smallest quantities. Our tests show that a division to 680 logarithmic bins of the energy range $`_E`$ (i.e $`20`$ bins per one order of magnitude) gives an acceptable compromise between precision and calculation time. We divide the interval $`z=[z_{dec}0.001]`$ to 30 logarithmic bins and the last step is from $`z=0.001`$ to $`z=0`$. The program is written in $`C^{++}`$ language and is highly modulable. It can be requested from the author. To test the precision of our numerical calculation, we have run the program without interaction terms. The error on total $`T^{00}`$ is $`0.7\%`$ and on non-baryonic DM is practically zero. For other species it is $`5.5\%`$ to $`7.5\%`$. Including interaction terms but not the decay of DM gives the same answer. This test is crucial for correct simulation of thermal equilibrium of the Universe. ### 4.3 Cosmology Model and Initial Conditions We consider a flat universe with present value of parameters as the followings: $`\mathrm{\Omega }_M=0.3`$, $`\mathrm{\Omega }_\mathrm{\Lambda }=0.7`$, $`h=H_0/100km\mathrm{sec}^1Mpc^1=0.7`$ and $`\mathrm{\Omega }_b=0.02h^2`$. We fix the decoupling redshift at $`z_{dec}=1100`$. The distribution of species at that time was thermal with a temperature $`T_{dec}=T_{cmb}(z_{dec}+1)=0.26eV`$, $`T_{cmb}=2.728K`$ for $`e^\pm `$, $`p^\pm `$ and $`\gamma `$ and $`\frac{4}{11}T_{dec}`$ for $`\nu `$ and $`\overline{\nu }`$. In the same way, one can determine the temperature of Dark Matter <sup>2</sup><sup>2</sup>2One should consider $`T_{dm}`$ as an estimation of kinetic energy scale rather than a real temperature because if ultra heavy particles exist, they could never be thermalized. at decoupling : $$T_{dm}(z_{dec})=\frac{g_{s}^{dec}{}_{}{}^{\frac{2}{3}}T_{dec}^2}{g_{s}^{dm}{}_{}{}^{\frac{2}{3}}T_{dmdec}}$$ (19) $`T_{dmdec}`$ is the decoupling temperature of the Dark Matter. If we assume that $`T_{dmdec}10^{16}eV`$, $`T_{dm}(z_{dec})<10^{18}eV`$. Therefore, our approximation $`T_{dm}=0`$ is quite justified. We know the density of species at present. Their initial value at decoupling depends on the equation of state of the Universe. However, it is exactly what we want to calculate! Consequently, we have to determine their value at decoupling approximately by neglecting the effect of UHDM decay. As the lifetime of WIMPZILLA is assumed to be much longer than the age of the Universe, the present value of densities after evolution must stay very close to our initial assumption. We define the initial densities as the followings: $`n_\gamma =n_{COBE}(z_{dec}+1)^3`$ $`n_\nu =n_{\overline{\nu }}=3\times {\displaystyle \frac{4}{11}}\times n_\gamma `$ (20) $`n_p={\displaystyle \frac{\mathrm{\Omega }_b\rho _c}{m_p}}(z_{dec}+1)^3`$ $`n_e^{}=n_pn_{\overline{p}}=n_{e^+}=0`$ (21) $`n_{dm}={\displaystyle \frac{\rho _c(\mathrm{\Omega }_M\mathrm{\Omega }_{hot}\mathrm{\Omega }_b(1+\frac{m_e}{m_p}))}{m_{dm}}}(z_{dec}+1)^3`$ $`\mathrm{\Omega }_{hot}={\displaystyle \frac{\pi ^2}{30}}g_{}T_{cmb}^4.`$ (22) Knowing initial density and temperature, other quantities like chemical potential and distributions can be determined . We assume that the age of the Universe $`\tau _0=14.8Gyr`$. This quantity also depends on the equation of state and we use it only for fixing the lifetime of WIMPZILLA. ### 4.4 Backgrounds Apart from CMB and relic neutrinos which are included in the initial conditions, we don’t include any other background to high redshift distributions. For $`z3`$, we add near-IR to UV emissivity of stars to equation (17). Far-IR which is very important for energy dissipation of UHECRs, is not added because there is very little information about its evolution with redshift. High energy backgrounds are not added because we want to be able to distinguish the contribution of remnants from other sources. Adding backgrounds by ”hand” evidently violates the energy conservation of the model, but there is not any other simple alternative method. Moreover, the violation is very small and comparable to numerical errors. Here a comment is in order: why have we included star backgrounds and not the synchrotron radiation ? Star background is considered as an external source. By contrast, synchrotron radiation concerns high energy electrons which are involved in the evolution. In (1), the elimination of interaction with an external field in a homogeneous universe means that the probability for production and absorption of synchrotron photons is the same. It is not therefore possible to consider their production without their absorption i.e. interaction of electrons with external magnetic field. ## 5 Results Figure 3 shows the energy flux of high energy protons and photons in a homogeneous universe. Fig.4 shows the same quantity for all species. The GZK cutoff is very transparent. With our background model it begins at $`E10^{18.2}eV`$ for protons due to $`p\gamma `$ interaction (see optical depth in Fig.5) and at $`E10^{13}eV`$ for photons due to $`e^\pm `$ production. For purely kinematic reasons, proton cutoff is much shallower than photon one. The resonance of $`p\gamma `$ interaction is very close to $`p`$ rest mass where $`A(s)`$ in (3) is very small. Moreover, division of (8) by $`E`$ reduces the effect of absorption on the distribution. According to Fig.3, even the shortest lifetime we have considered, can not explain the observed flux of protons (in contrast to that assumes a 2-particle decay mode). The same figure shows also the flux without energy dissipation. It is compatible with Ref. which assumes a hadronic decay but does not consider the energy dissipation of secondary particles. In this latter case, the lifetime must be $`46`$ orders of magnitude larger than the age of the Universe. On the one hand, this result proves the rôle of a realistic model of energy dissipation in the estimation of mass and lifetime of UHDM. On the other hand, it shows the importance of clumping of Dark Matter, i.e. most of observed UHECRs must come from nearby sources. This conclusion is independent of the source of UHECRs. A similar conclusion has been obtained in by fitting the expected flux from extragalactic sources and from the Halo on the data. However they don’t consider the dissipation. In the next section we show that even in a halo, the dissipation of photons energy is significant. In the case of decaying UHDM hypothesis, the most important source is the Galactic Halo. Before trying to make a simple model of halo in the next section, we discuss some of other conclusions one can make from this study. From Fig.3 and by taking into account the fact that $`p\gamma `$ cross-section is $`10^4`$ times smaller than $`pp`$, with present statistics of UHECRs, less than one photon shower could be observed. The comparison of EGRET data for $`10^8eV<E<10^{11}eV`$ with our calculation shows that it is compatible with a short lifetime UHDM. Future observations of GLAST at $`E>10^{11}eV`$ is crucial for understanding the source of UHECRs because one expects an additional contribution from synchrotron radiation of ultra high energy electrons at $`E10^{12}10^{14}eV`$ . Fig.5 illustrates the total and partial optical depth, i.e. $`(t,p)`$ in equation (9) at $`z=0`$. It shows that for protons with $`E10^{20}eV`$, even a source at $`20Mpc`$ must be very strong to be able to provide the observed flux because its flux is reduced by $`11`$ orders of magnitude! The optical depth of protons must be even larger than what we have obtained here because we didn’t take into account the far-IR and radio backgrounds. This makes difficulties for the recent suggestion by Ahn E.J., et al. that Virgo cluster can be the only source of UHECRs, even if the magnetic field of Galactic wind is as strong as what is considered in that work. The main challenge is finding a conventional source with enough emissivity of CRs at ultra high energies. ### 5.1 CMB Distortion and Entropy Excess To see if a relatively short living UHDM can distort CMB, we reduced the spectrum of photons for a decaying DM from the spectrum with a stable DM. Fig.6 shows the result for $`z=0`$. There is no distortion at least up to $`1`$ to $`10^8`$ parts for $`E3eV`$. However, one should note that this conclusion depends somehow on the cross-section cuts at low energies. Nevertheless, giving the fact that CMB flux at its maximum is $`10^{20}`$ times larger than the rest of the spectrum, it seems very unlikely that it can be disturbed, otherwise SZ effect due to stars had to be observed everywhere. The distortion of CMB anisotropy will be studied elsewhere. We have also examined the entropy excess separately for each species. There is no entropy enhancement except for $`e^+`$ and $`p^{}`$ which in our model are absent from the initial conditions. Comparing to other species, their contribution is very small and negligible. ### 5.2 Baryon and Lepton Asymmetry Generation It has been suggested that the decay of UHDM may be able to generate additional baryon and lepton asymmetry. At GUT scale, i.e. the mass scale of WIMPZILLA, we expect such processes and the fact that a late time decay is out of thermal equilibrium and satisfies Sakharov conditions for baryogenesis make growing asymmetry plausible. The rate of baryonic (or leptonic) number production by decay of UHDM in comoving frame can be expressed as: $$\frac{d(n_bn_{\overline{b}})}{dt}+3\frac{\dot{a}(t)}{a(t)}(n_bn_{\overline{b}})=\frac{\epsilon n_{dm}}{\tau }.$$ (23) $`\epsilon `$ is the total baryon number violation per decay. The solution of this equation is: $`\mathrm{\Delta }(n_bn_{\overline{b}})`$ $`=`$ $`\epsilon n_{dm}(t_0)(1\mathrm{exp}({\displaystyle \frac{tt_0}{\tau }})){\displaystyle \frac{(1+z_0)^3}{(1+z)^3}}.`$ (24) $`\mathrm{\Delta }B`$ $``$ $`{\displaystyle \frac{\mathrm{\Delta }(n_bn_{\overline{b}})}{2g_{}n_\gamma }}={\displaystyle \frac{\epsilon n_{dm}(t_0)}{2g_{}n_\gamma (t_0)}}(1\mathrm{exp}({\displaystyle \frac{tt_0}{\tau }})){\displaystyle \frac{(1+z)}{(1+z_0)}}.`$ (25) If $`t_0=t_{dec}`$, $`\frac{n_{dm}(t_0)}{n_\gamma (t_0)}10^{22}`$ (for $`m_{dm}=10^{24}eV`$). Therefore $`\mathrm{\Delta }B10^{22}\epsilon `$ at $`z=0`$. As $`\epsilon `$ can not be larger than total multiplicity, $`1000`$, $`\mathrm{\Delta }B10^{19}`$, i.e. much smaller than primordial value $`10^{10}`$. We tested this argument by assuming $`\epsilon =0.1`$ at all energies, i.e. $`\epsilon _{tot}=0.1_{tot}`$. Evidently $`n_{\overline{p}}`$ is smaller, but the change of $`n_p`$ is too small to be measured. The same is true for number density of leptons, but energy density of leptons with respect to anti-leptons increases by an amount comparable to $`\epsilon `$. ### 5.3 Equation of State of the Universe Decay of UHDM gradually changes part of CDM to HDM and thus changes the equation of state of the Universe. Fig.7 shows the variation of equation of state. For $`\tau 50\tau _0`$ or larger, it would be too small to be measurable. For smaller $`\tau `$, DM decay plays the rôle of a running cosmological constant. A complete study of this issue and comparison with data is under preparation. ## 6 Halo To see the effect of clumping of a decaying UHDM on the flux of UHECRs, here we try to make a very simple model. A complete treatment of halos will be reported elsewhere. We consider a halo as a uniform over-density with a limited size at $`z=0`$. This is simulated by following the decay of UHDM and evolution of remnants for a time comparable to the propagation time in the halo. Evolution equation and energy binning is taken to be the same as in the homogeneous universe case with $`a(t)=cte`$. Because we want to study the propagation of remnants in a volume comparable to the Galactic Halo, we consider time steps equivalent to $`10kpc`$. Our tests show that after a few steps ($`7`$), the accumulation rate of ultra high energy particles becomes very slow. We consider two cases. In the first case we simply evolve distributions for a number of steps (up to 30). In the second case, after some steps, we stop the decay of the Dark Matter to simulate an inner halo of MACHOs. Then, the evolution is continued for more $`5`$ steps (i.e. $`50kpc`$ to simulate propagation through MACHOs. Evidently this model is very approximative. We use it only to make a crude estimate of production and absorption of UHECRs. ### 6.1 Initial Conditions and Galactic Backgrounds Galactic baryonic matter is taken to have a thermal distribution with $`T_b=10^4K`$. We assume that baryonic over-density is biased with respect to DM i.e. the fraction of baryons to DM is larger than its mean value in the Universe. With these assumptions the initial number densities can be expressed as: $`n_p={\displaystyle \frac{b\delta \rho _c}{m_p}}n_e^{}=n_p`$ $`n_{\overline{p}}=n_{e^+}=0`$ $`n_{dm}={\displaystyle \frac{(1b)\delta \rho _c}{m_{dm}}}`$ (26) $`b`$ is the fraction of baryons in the halo. In the following $`b=0.3`$, i.e. $`2`$ times primordial value in the cosmological model explained above. $`\delta `$ is the mean over-density of the Halo. Inspired by universal halo density distribution of NFW , we consider a halo with characteristic radius (i.e. virial radius) $`r_{200}=0.12Mpc`$. According to NFW distribution and by definition this means $`\delta =200`$. These parameters defines a halo of mass $`M_H=6\times 10^{12}M_{}`$. Neutrino density is assumed to be the same as relic at $`z=0`$. For photons, in addition to CMB, we consider a galactic background as the following: Galactic IR and visible backgrounds are not very well known. We use the results of the model developed by DIRBE group for detection of extragalactic component of the IRB . We consider the observed value of IRB after elimination of Inter-Planetary Dust (IDP) contribution as the galactic background. It is just an estimation of average galactic IRB. It is not clear if we can extend the local value of IRB to whole galaxy or take it as a representative average. For this reason we also increase it $`10`$ times (probably an extremely high value) to see the effect on the energy dissipation of UHECRs (see below for conclusions). Our simulation does not include radio background. For soft and hard X-Ray galactic backgrounds, we use the model developed for extraction of extragalactic component from ROSAT and ASCA observations . It considers GXB as two thermal components, a soft component with $`T_{sx}=70eV`$ from Local Bubble, and a hard component with $`T_{hx}=145eV`$ from hot gas, probably in the Halo. We add also the extragalactic component for $`0.25keV<E<10keV`$. ## 7 Results Fig.8 shows the distribution of high energy protons and photons for a uniform halo and for a halo that its inner part is composed of MACHOs. Only the result for $`m_{dm}=10^{24}eV`$ with $`\tau =5\tau _0`$ and $`\tau =50\tau _0`$ is shown. Because of importance of the spectrum close to observed trough in the interpretation of the results, Table 4 summarizes the numerical value of simulated and observed spectrum. For photons the trough of GZK cutoff is shallower than in a homogeneous universe and for protons it is practically absent. Consequently, the calculated flux at $`E10^{19.5}eV`$ is somehow higher than observation. However, at higher energies simulation results specially one with a MACHO halo is in the $`1\sigma `$ error range of the observations. The simplest explanation of having a smaller observed flux close to the minimum of the spectrum can be the need for increasing the lifetime somehow (but not by many orders of magnitude as suggested by previous works ). However, the lack of a minimum in the simulated spectrum<sup>3</sup><sup>3</sup>3In fact optical depth has a maximum at $`E3.3\times 10^{19}eV`$. means that in some way our simulation is not exact. It can be due to a too simple halo model. Other possibilities are a different (probably harder) decay spectrum and/or energy loss in the Galactic magnetic field. If these suggestions are true, fluxes will be smaller and there is no need for increasing the lifetime. A more complete simulation of the Halo and magnetic field and a better understanding of non-baryonic DM distribution is necessary for making any definitive conclusion. The minimum in the spectrum can also be interpreted as a wider distribution of sources. In this case, it is hardly probable that sources responsible for CRs at lower energies can explain this behavior because even if the spectrum of these sources can be extrapolated with the same slop to higher energies, it can not explain the rising slop of the spectrum. As the IR background is crucial for energy dissipation of UHECRs, we have also increased it to $`10`$ times of the DIRBE model to see the effect. Proton flux at high energies slightly decreases, but it can not explain the observed minimum. We have also tested the distortion of the CMB by remnants as described for a homogeneous universe. There is no distortion up to at least $`1`$ to $`10^8`$ for $`E10eV`$. Summarizing this section, it seems that a decaying UHDM with a lifetime as short as $`\tau 10100\tau _0`$ is not ruled out by present observations. ## 8 Conclusions The main purpose of this work was showing the importance of a realistic physical model for finding the answer to the mystery of UHECRs and introducing readers to the program that has been developed for achieving this goal. We showed that $`\xi \tau _0/\tau 0.10.01`$ where $`\xi `$ is the contribution of UHDM in the Dark Matter. This value is larger (or equivalently the lifetime is shorter for the same contribution) than what has been suggested in previous works which don’t consider the dissipation . If a more realistic halo model confirms this conclusion, the cosmological implication of a UHDM can be important. We have studied a special decay mode. Any other mode that produces invisible WIMPs or leptonic or semi-leptonic modes decreases the lifetime of UHDM. If more nucleon are produced, the lifetime must be longer but it seems less probable than other cases. We showed the reciprocal influence of UHECRs and backgrounds on each others. Consequently, whatever the source of UHECRs, it is extremely important to correlate their observations to the observation of high energy photon and neutrino backgrounds. This work is the first step to a comprehensive study of the effects of a decaying Dark Matter. Other issues like a realistic model for halos, effects on the determination of cosmological parameters and equation of state and comparison with more data from cosmic rays and high energy backgrounds remain for future works. Note: Shortly after completion of this work the Lake Baikal Experiment collaboration have published their upper limit on the flux of high energy neutrinos. It is well above what is obtained in our simulation (see Fig.5). ## Appendix 1: Fragmentation in MLLA The MLLA treats fragmentation as a Markov process. Consequently, the differential multiplicity is proportional to splitting function $`P(x)`$(See e.g. ): $$\frac{d(x_E)}{dx_E}P(x)x_E=\frac{E}{E_j}.$$ (27) For a gluon jet and at $`x_E1`$, $`P(x_E)\frac{1}{x_E}`$, and one expects that: $$\frac{d(x_E)}{d(ln(x_E))}\text{Total Num. of splitting}E_j.$$ (28) In PYTHIA however, at high energies $`E100GeV`$, increasing CM energy just increases the probability of having more high energy fragments which simply escape fragmentation. A consequence of this behavior is the narrowing of the multiplicity distribution (See Fig.9) in contrast to theoretical prediction i.e. KNO scaling for $`E_j\mathrm{}`$. This limit is obtained at parton level. However, if LPHD is valid, one expects the same type of behavior at hadron level. Another factor that can explain, at least partially, the deviation from theory is the decay of hadrons in our simulation. It increases total multiplicity more than its statistical variation and makes the distribution narrower, but it can not explain the absence of a tail of large multiplicity events. Even after decay, these events should keep their difference with average. As both MLLA and Monte Carlos fail to reproduce observations in relatively low energies , the exact behavior at high energies is not clear. ## Appendix 2: Production Integral for $`22`$ Processes Cross-sections are Lorantz invariant. However, to have a unique expression for what are simulated and what are calculated analytically, we determine all of them in their CM. The triple integral in (11) depends on three momentum variables corresponding to the momentum of two incoming particles and one of the outgoing particles that its evolution is under calculation. In the case of a $`22`$ process, the amplitude of the momentum of final particles depends only on s. If: $$𝐩=(E,p\mathrm{cos}\varphi \mathrm{sin}\theta ,p\mathrm{sin}\varphi \mathrm{sin}\theta ,p\mathrm{cos}\theta ).$$ (29) is the 4-momentum of outgoing particle, and $`𝐩^{}`$ is its counterpart in the CM: $`𝐩_{}^{}{}_{}{}^{2}`$ $`=`$ $`{\displaystyle \frac{(sm^2m^2)^24m^2m^2}{4s}}.`$ (30) $`𝐩^{}`$ $`=`$ $`\mathrm{\Gamma }𝐩`$ (31) $`m`$ and $`m^{}`$ are the mass of out-going particles, $`\mathrm{\Gamma }`$ is the boost matrix. The equality of two expressions for $`𝐩^{}`$ leads to an equation that can be solved for one of the angular variables in (29). The calculation is tedious but strait forward. With respect to $`\varphi `$, the equation is $`4^{th}`$ order but in the case of a homogeneous cosmology where the boost matrix depends only on the relative angle between incoming particles, it depends only on $`\mathrm{tan}^2(\frac{\varphi }{2})`$ and is analytically solvable. In this way the integration over $`\varphi `$ in (11) reduces to sum of integrand evaluated at the roots of the equation. This method provides a general way to deal with this problem and is more convenient than doing calculation for each cross-section separately. Acknowledgement I would like to thanks people of Strasburg Observatory for their kindness.
warning/0001/nucl-th0001057.html
ar5iv
text
# The reaction ²𝐻⁢(𝛾,𝜋⁰)⁢𝑛⁢𝑝 in the threshold region ## I Introduction Nowadays, the threshold $`s`$\- and $`p`$\- amplitudes of $`\pi ^0`$ photoproduction off the deuteron are of considerable theoretical and experimental interest. This is motivated by two reasons. Firstly, the deuteron is a natural ‘source’ of neutrons. If one wants to minimize uncertainties stemming from nuclear forces when extracting neutron threshold parameters one should use a deuteron target. At present, there is a significant disagreement for values of the threshold electric dipole amplitude $`E_{0+}^{n\pi ^0}`$ calculated in chiral perturbation theory (ChPT) and dispersion theory (DR), respectively, 2.13 and 1.19 (in units of $`10^3/\mu _{\pi ^+}`$ which are suppressed from here on). At the same time the predicted values for $`E_{0+}^{p\pi ^0}`$ agree reasonably with each other ($`1.16`$ and $`1.22`$, respectively) and with experimental values $`1.32\pm 0.08`$ and $`1.31\pm 0.08`$ . Resolving the problem mentioned in favour of the one or the other theory requires experimental measurements of which in the reactions $`\gamma d\pi ^0d`$ and $`ede^{}\pi ^0d`$ seem to be most promising. However, even in these simplest reactions many theoretical problems in the extraction of neutron data are anticipated. On the other hand, the deuteron cannot be considered as a direct sum of the proton and neutron (impulse approximation). Photons may also interact with potential (virtual) mesons. This effect is known as ‘Meson Exchange Currents’ (MEC). It was found to be very important in the case of coherent neutral pion photoproduction on the deuteron . Accepting the ChPT predictions for free nucleons we could expect in the impulse approximation a value of about $`+0.5`$ ($`=\frac{1}{2}[E_{0+}^{p\pi ^0}+E_{0+}^{n\pi ^0}]`$) for the deuteron electric dipole amplitude $`E_d`$. In fact, however, the ChPT result is quite different being about $`1.8\pm 0.2`$ and the difference is due to the meson exchange contributions. A very recent measurement of $`E_d`$ performed at SAL in coherent $`\pi ^0`$ photoproduction on the deuteron within 20 MeV of threshold has given a value of $`1.45\pm 0.09`$ which is about 20% lower than the ChPT prediction. There is also a significant deviation from the ChPT prediction for the free-nucleon $`p_1`$ amplitude. A comparison of the differential cross sections calculated in a theoretical model with the data revealed a noticeable disagreement as well. It was claimed in Ref. that a possible reason for the disagreements might consist in inadequate treatment of the inelastic channel in Ref. . Since the $`{}_{}{}^{2}H(\gamma ,\pi ^0)np`$ reaction cannot be resolved one has to estimate its contribution to the coherent channel making the use of theoretical predictions. Indeed, a model developed in Ref. seems to be oversimplified. For example, the authors employed a square-well deuteron wave function and used the effective range approximation to describe the final state $`n`$-$`p`$ interaction. The deuteron $`d`$-wave was ignored. Explicit calculations of very important diagrams with $`\pi `$-$`N`$ rescattering were not carried out. Effectively these latter where included into the model by accepting the extracted value $`1.45`$ for $`E_d`$ which, as it has been noted, contains contributions from MEC or, in other words, from $`\pi `$-$`N`$ rescattering. Although a theoretical uncertainty of $`\pm 25\%`$, based on input parameters, was assigned in Ref. to the calculated cross sections it would be desirable to have more realistic calculations for the inelastic channel. The present paper is aimed at fixing the defects of the model . We are going to study whether these defects may be responsible for the deviations mentioned. In our previous paper on the reaction $$\gamma d\pi ^0np$$ (1) we restricted ourselves to the energy region 200 to 400 MeV. Now we extend the model to the threshold region. The extension includes: i) a more realistic treatment of the $`\pi ^0`$ photoproduction amplitudes at threshold energies; ii) in Ref. we considered $`\pi `$-$`N`$ rescattering mechanism but without taking into account the $`n`$-$`p`$ rescattering process which may follow $`\pi `$-$`N`$ rescattering. In the present paper a corresponding diagram has been taken into account. We would like to emphasize that the study of the process (1) in the threshold region is hardly of theoretical interest in its own. Rather it is needed for experimentalists to estimate the relative contribution of the inelastic channel to the cross section of coherent pion photoproduction on the deuteron. Furthermore, these studies may be useful for extracting the elementary $`E_{0+}^{n\pi ^{}}`$ amplitude from exclusive inelastic data where the recoil neutron is detected. ## II Kinematics Let us denote by $`k=(\omega ,𝐤),p_d=(\epsilon _d,𝐤),q=(\epsilon _\pi ,𝐪),p_n=(\epsilon _n,𝐩_n)`$ and $`p_p=(\epsilon _p,𝐩_p)`$ the 4-momenta of the initial photon and deuteron and the final pion, neutron and proton, respectively in the $`\gamma `$-$`d`$ c.m. frame. A symbol $`E_\gamma `$ we reserve for the lab photon energy $`E_\gamma =W_{\gamma d}\omega /m_d`$ with $`W_{\gamma d}=\omega +\epsilon _d=\omega +\sqrt{\omega ^2+m_d^2}`$ and $`m_d`$ being the deuteron mass. It is convenient to take as independent kinematical variables the photon energy and pion angle, $`\mathrm{\Theta }_\pi `$, in the c.m. frame and momentum $`𝐩`$ of one of the nucleons - say, the neutron - in the final $`n`$-$`p`$ c.m. frame (see Ref. ). Once these variables have been specified the pion momentum can be easily found: $`|𝐪|={\displaystyle \frac{1}{2W_{\gamma d}}}\sqrt{[W_{\gamma d}^2(W_{np}+\mu )^2][W_{\gamma d}^2(W_{np}\mu )^2]},`$ (2) where $`W_{np}=2\sqrt{𝐩^2+m^2}`$ ($`m`$ is the nucleon mass) and $`\mu `$ is the $`\pi ^0`$ mass. Then making a boost with the velocity $`𝐪/(W_{\gamma d}\epsilon _\pi )`$ we obtain the neutron momentum $`p_n`$ and, therefore, totally restore the kinematics in the $`\gamma `$-$`d`$ c.m. frame. Note that in kinematic calculations a distinction is not made between the proton and neutron masses. It is, however, taken into account in the threshold $`\pi ^0`$ photoproduction amplitude when parametrizing the $`s`$\- and $`p`$-waves multipoles (see Sect. III, Eq. (7)). The differential cross section is $`{\displaystyle \frac{d^4\sigma }{d𝐩d\mathrm{\Omega }_\pi }}={\displaystyle \frac{1}{(2\pi )^5}}{\displaystyle \frac{2m^2\epsilon _d|𝐪|^3}{4\omega W_{\gamma d}W_{np}(\epsilon _\pi qp_p\epsilon _p\mu ^2)}}{\displaystyle \frac{1}{6}}{\displaystyle \underset{m_pm_n\lambda m_d}{}}|m_pm_n|T|\lambda m_d|^2,`$ (3) where $`m_p`$, $`m_n`$, $`\lambda `$, and $`m_d`$ are spin states of the proton, neutron, photon, and deuteron, respectively. Assuming that near threshold energies the nucleons are not detected we should integrate the l.h.s. of Eq. (3) over the momentum $`𝐩`$ to obtain $`{\displaystyle \frac{d\sigma }{d\mathrm{\Omega }_\pi }}={\displaystyle \underset{0}{\overset{p^{max}}{}}}{\displaystyle \frac{d^4\sigma }{d𝐩d\mathrm{\Omega }_\pi }}p^2𝑑p𝑑\mathrm{\Omega }_𝐩,`$ (4) where the maximum value $`p^{max}`$ is found from energy-momentum conservation and reads $`p^{max}={\displaystyle \frac{1}{2}}\sqrt{(W_{\gamma d}\mu )^24m^2}.`$ (5) ## III Theoretical model for the $`\gamma d\pi ^0np`$ process The scattering amplitude of the reaction (1) is described as a sum of contributions from diagrams which are expected to be most important in the energy region under consideration. The pole diagrams a and b in Fig. 1 and one-loop diagram 1c correspond to the impulse approximation without and with $`n`$-$`p`$ interaction in the final state, respectively. The necessity to consider the latter mechanism follows from the very strong nucleon-nucleon interaction at small energies. It might seem that diagrams 1d to f with $`\pi `$-$`N`$ rescattering could be disregarded in the threshold region since the $`\pi `$-$`N`$ scattering lengths are about two orders smaller than those for $`n`$-$`p`$ scattering. In fact, however, there is actually no suppression of the diagrams 1d to f in comparison with diagram 1c. Indeed, keeping in mind that the threshold electric dipole amplitudes $`E_{0+}`$ for charged channels are about 30 times larger in absolute numbers than those for the neutral channels one can expect the corresponding contributions from $`\pi `$-$`N`$ rescattering to be of the same order as the ones from final state $`n`$-$`p`$ interaction. This expectation will be confirmed below by numerical calculations. A very big effect of $`\pi `$-$`N`$ rescattering in the case of coherent $`\pi ^0`$ photoproduction on the deuteron in the threshold region was also found in Refs. . Note that our treatment of $`\pi `$-$`N`$ contribution is similar to the one of ‘three-body interactions to order $`q^3`$’ from Ref. in the case of coherent $`\pi ^0`$ photoproduction off the deuteron. It was found in that paper that three-body contributions at order $`q^4`$ are less important. Based on this finding we do not take into account the corresponding diagrams for the reaction (1) though they of course exist. We begin with the pole diagrams. The matrix element corresponding to the diagram in Fig. 1a reads $$m_pm_n|T^{1a}(𝐤,𝐪,𝐩_p)|\lambda m_d=\underset{m_{\stackrel{~}{n}}}{}\mathrm{\Psi }_{m_pm_{\stackrel{~}{n}}}^{m_d}\left(𝐩_p+\frac{𝐤}{2}\right)m_n|T_{\gamma \stackrel{~}{n}\pi ^0n}(𝐤_{\pi n},𝐪_{\pi n})|\lambda m_{\stackrel{~}{n}},$$ (6) where $`\mathrm{\Psi }_{m_pm_{\stackrel{~}{n}}}^{m_d}(𝐩_p+𝐤/2)`$ is the deuteron wave function (DWF) and $`m_n|T_{\gamma n\pi ^0n}|\lambda m_{\stackrel{~}{n}}`$ is the amplitude of the elementary process $`\gamma n\pi ^0n`$. The amplitude depends on photon ($`𝐤_{\pi n}`$) and pion ($`𝐪_{\pi n}`$) momenta taken in the c.m. frame of the $`\pi `$-$`n`$ pair. These momenta can be obtained from the corresponding momenta in the $`\gamma `$-$`d`$ c.m. frame through a boost with the velocity $`𝐩_p/(W_{\gamma d}\epsilon _p)`$. In calculations we used DWF for non-relativistic versions of the Bonn OBE potential (the same models were used when solving the Lippmann-Schwinger equation for the $`n`$-$`p`$ scattering amplitude needed for calculations of diagrams 1c and f, see below). Analytical parametrizations of the $`s`$\- and $`d`$-wave amplitudes were taken from Ref. . We would like to note just here that our results are practically independent of the choice of the potentials. In our previous paper we used the well-known Blomqvist-Laget parametrization of the pion photoproduction amplitudes. This model is rather simple and, nevertheless, fits reasonably to the experimental data. Unfortunately, it is obsolete in the sense that it predicts the threshold electric dipole amplitudes for the neutral channels in accordance with the classical low-energy theorem, $`E_{0+}^{p\pi ^0}=2.4`$ and $`E_{0+}^{n\pi ^0}=0.4`$, but in disagreement with the modern values (see Sect. I). To be consistent with these latter, we use in the energy region $`\omega 150`$ MeV a parametrization for the neutral pion photoproduction amplitude given in Ref. . The threshold amplitude in the $`\gamma `$-$`N`$ c.m. frame can be written in the following form $`T_{\gamma N\pi ^0N}={\displaystyle \frac{4\pi W_{\gamma N}}{m}}\left[i𝝈\mathit{ϵ}(E_{0+}+𝐤𝐪p_1)+i𝝈𝐤\mathit{ϵ}𝐪p_2+𝐪(𝐤\times \mathit{ϵ})p_3\right],`$ (7) where $`W_{\gamma N}`$ is the total energy of the $`\gamma `$-$`N`$ system, and $`p_1`$, $`p_2`$ and $`p_3`$ are the $`p`$-waves multipoles. The $`E_{0+}`$ amplitude was taken in the form of Eq. (25) of Ref. with parameters from Table I of the same work. The parametrization takes into account the energy dependence of the $`E_{0+}`$ multipole as well as the so-called cusp effect. Just at threshold it reproduces the ChPT and DR values (see Sect. I). These parametrizations were used in the present paper to study uncertainties due to the choice of a theoretical model for pion photoproduction. As to the $`p`$-wave amplitudes, they are also taken for two models and listed in Table I. As was shown in Ref. , the parametrizations corresponding to ChPT and DR reproduce very well the available data on the energy dependence of the real part of $`E_{0+}`$ as well as the total and differential cross sections of the reaction $`\gamma p\pi ^0p`$ up to $`E_\gamma 180`$ MeV. One should emphasize that, although we are going to work in the energy region below 160 MeV, we need parametrizations for the photoproduction amplitudes at higher energies too. Such energies emerge because of two kinds of boosts. The first one occurs at the integration over $`𝐩`$ in Eq. (4). Below 160 MeV the value of $`p^{max}`$ does not exceed 125 MeV/c so that the corresponding boosts lead to only small energy shifts. But at integrations over loops when calculating $`n`$-$`p`$ final state interaction and $`\pi `$-$`N`$ rescattering (see below), energy shifts can be much bigger. In addition, when considering $`\pi `$-$`N`$ rescattering the pion photoproduction amplitudes for the charged channels are also needed. Both these latter and the $`\pi ^0`$ photoproduction amplitudes at photon c.m. energies above 150 MeV were taken from the Blomqvist-Laget model . In full analogy with Eq. (6) we write the matrix element corresponding to diagram 1b in the form $$m_pm_n|T^{1b}(𝐤,𝐪,𝐩_n)|\lambda m_d=\underset{m_{\stackrel{~}{p}}}{}\mathrm{\Psi }_{m_nm_{\stackrel{~}{p}}}^{m_d}\left(𝐩_n+\frac{𝐤}{2}\right)m_p|T_{\gamma \stackrel{~}{p}\pi ^0p}(𝐤_{\pi p},𝐪_{\pi p})|\lambda m_{\stackrel{~}{p}}.$$ (8) The photon ($`𝐤_{\pi p}`$) and pion ($`𝐪_{\pi p}`$) momenta in the c.m. frame of the $`\pi `$-$`p`$ pair can be obtained by boosting the corresponding momenta in the $`\gamma `$-$`d`$ c.m. frame with the velocity $`𝐩_n/(W_{\gamma d}\epsilon _n)`$. The one-loop diagram in Fig. 1c with $`n`$-$`p`$ rescattering in the final state is expected to be very important in the energy region under consideration. The smallness of the relative $`n`$-$`p`$ momenta leads to strong $`{}_{}{}^{1}S_{0}^{}`$\- and $`{}_{}{}^{3}S_{1}^{}`$-wave interactions which significantly modify the $`n`$-$`p`$ plane wave states appearing in the pole diagrams 1a and b. The matrix element corresponding to diagram 1c is $`m_pm_n|T^{1c}(𝐤,𝐪,𝐩_n)|\lambda m_d=m{\displaystyle \frac{d^3𝐩_s}{(2\pi )^3}\frac{1}{p_{in}^2p_{out}^2i0}}`$ (9) $`\times {\displaystyle \underset{m_{\stackrel{~}{p}}m_{\stackrel{~}{n}}}{}}[𝐩_{out},m_pm_n|T_{np}|𝐩_{in},m_{\stackrel{~}{p}}m_{\stackrel{~}{n}}m_{\stackrel{~}{p}}m_{\stackrel{~}{n}}|T^{1a}(𝐤,𝐪,𝐩_s)|\lambda m_d`$ (10) $`+𝐩_{out},m_pm_n|T_{np}|𝐩_{in},m_{\stackrel{~}{p}}m_{\stackrel{~}{n}}m_{\stackrel{~}{p}}m_{\stackrel{~}{n}}|T^{1b}(𝐤,𝐪,𝐩_s)|\lambda m_d],`$ (11) where $`𝐩_{out}=(𝐩_p𝐩_n)/2`$ and $`𝐩_{in}=𝐩_s+𝐪/2`$ are the relative momenta of the $`n`$-$`p`$ pair after and before scattering, respectively, and $`𝐩_{out},m_pm_n|T_{np}|𝐩_{in},m_{\stackrel{~}{p}}m_{\stackrel{~}{n}}`$ being the half off-shell $`n`$-$`p`$ scattering amplitude. We will not discuss here details of computations of the amplitude (11) because they are given in Ref. . Note only that all partial waves with the total angular momentum $`J1`$ were retained in the $`n`$-$`p`$ scattering amplitude. We cut off the integration in Eq. (11) at $`|𝐩_s|^{max}=500`$ MeV/c since it is difficult to control the pion photoproduction amplitudes at higher momenta. Such a cutting off has no impact on our final results because the integral is mainly saturated at smaller momenta. For example, a variation of $`|𝐩_s|^{max}`$ from 400 to 500 MeV changes the calculated differential cross sections by less than 2%. One further remark is concerned with the choice of particle energies when integrating over the momenta $`𝐩_s`$. In accordance with a prescription of Ref. we will suppose a spectator particle (it is denoted in Fig. 1c by a subscript $`s`$) to be on its mass shell. The amplitude for the diagram in Fig. 1d reads $`m_pm_n|T^{1d}(𝐤,𝐪,𝐩_n)|\lambda m_d=`$ (12) $`{\displaystyle }{\displaystyle \frac{d^3𝐩_s^{}}{(2\pi )^3}}{\displaystyle \underset{m_1m_2}{}}\mathrm{\Psi }_{m_1m_2}^{m_d}(𝐩_s+{\displaystyle \frac{𝐤}{2}})\{{\displaystyle \frac{\epsilon _n^{}+\epsilon _s^{}}{2W_{\pi n}}}{\displaystyle \frac{1}{p_{s}^{}{}_{}{}^{2}p_{\pi n}^{}{}_{}{}^{2}i0}}`$ (13) $`\times [m_n|T_{\stackrel{~}{\pi }^0n\pi ^0n}(𝐩_{\pi n}^{},𝐩_s^{})|m_1m_p|T_{\gamma p\stackrel{~}{\pi }^0p}(𝐤_{\stackrel{~}{\pi }p},𝐪_{\stackrel{~}{\pi }p})|m_2`$ (14) $`m_n|T_{\stackrel{~}{\pi }^{}p\pi ^0n}(𝐩_{\pi n}^{},𝐩_s^{})|m_1m_p|T_{\gamma n\stackrel{~}{\pi }^{}p}(𝐤_{\stackrel{~}{\pi }p},𝐪_{\stackrel{~}{\pi }p})|m_2]\}.`$ (15) In Eq. (15) the asterisk denotes kinematical variables in the $`\pi ^0`$-$`n`$ c.m. frame. $`W_{\pi n}=\epsilon _\pi ^{}+\epsilon _n^{}=\sqrt{p_{\pi n}^{}{}_{}{}^{2}+\mu ^2}+\sqrt{p_{\pi n}^{}{}_{}{}^{2}+m^2}`$ is the total energy in this frame and $`p_{\pi n}^{}{}_{}{}^{2}={\displaystyle \frac{1}{4W_{\pi n}^2}}[W_{\pi n}^2(m+\mu )^2][W_{\pi n}^2(m\mu )^2]>0.`$ (16) A boost with the velocity $`𝐩_p/(W_{\gamma d}\epsilon _p)`$ is needed to transform the momentum $`𝐩_s^{}`$ to $`𝐩_s`$ in the $`\gamma `$-$`d`$ c.m. frame. As in the case of Eq. (11) we cut off the integration in Eq. (15) at $`|𝐩_s^{}|^{max}=500`$ MeV/c without any noticeable changes in the calculated differential cross sections. Spectator particles in diagrams 1d and e are again supposed to be on their mass shells. The $`\pi `$-$`N`$ scattering amplitude has the following form $`m^{}|T_{\pi N\pi ^{}N^{}}^I(𝐪^{},𝐪)|m`$ $`=`$ $`(2\pi )^3\sqrt{{\displaystyle \frac{4\epsilon _\pi ^{}\epsilon _{\pi }^{}{}_{}{}^{}\epsilon _N^{}\epsilon _{N}^{}{}_{}{}^{}}{m^2}}}`$ (18) $`\times {\displaystyle \underset{JLm_J}{}}C_{\frac{1}{2}m^{}Lm_L^{}}^{Jm_J}Y_L^{m_L^{}}(\widehat{𝐪}^{})C_{\frac{1}{2}mLm_L}^{Jm_J}Y_{L}^{m_L}{}_{}{}^{}(\widehat{𝐪})T_{IJ}^L(q^{},q),`$ where $`𝐪^{}`$ and $`𝐪`$ are the final and initial momenta of (generally, off-shell) particles in the $`\pi `$-$`N`$ c.m. frame, $`Y_L^{m_L}(\widehat{𝐪})`$ are the spherical harmonics, $`C_{J_1M_1J_2M_2}^{JM}`$ are the Clebsch-Gordan coefficients. Partial wave amplitudes $`T_{IJ}^L(q^{},q)`$ for all $`s`$\- and $`p`$-waves channels were obtained by solving the Lippmann-Schwinger equation for a separable energy-dependent $`\pi `$-$`N`$ potential built in Ref. . Only for the $`p_{33}`$-wave we preferred using a model from Ref. with explicit inclusion of a $`s`$-channel pole diagram with an intermediate $`\mathrm{\Delta }`$ isobar. Contributions from the $`d`$-waves were found to be negligible. Two $`\pi `$-$`N`$ scattering amplitudes in isospin space are numbered by a symbol $`I=\frac{1}{2}`$ or $`\frac{3}{2}`$. Needed in Eq. (15) physical amplitudes are linear combinations of these isospin amplitudes: $`T_{\pi ^0p\pi ^0p}`$ $`=T_{\pi ^0n\pi ^0n}={\displaystyle \frac{1}{3}}\left(T^{\frac{1}{2}}+2T^{\frac{3}{2}}\right),`$ (19) $`T_{\pi ^+n\pi ^0p}`$ $`=T_{\pi ^{}p\pi ^0n}={\displaystyle \frac{\sqrt{2}}{3}}\left(T^{\frac{1}{2}}T^{\frac{3}{2}}\right).`$ (20) In full analogy with Eq. (15) one has for diagram 1e $`m_pm_n|T^{1e}(𝐤,𝐪,𝐩_p)|\lambda m_d=`$ (21) $`{\displaystyle }{\displaystyle \frac{d^3𝐩_s^{}}{(2\pi )^3}}{\displaystyle \underset{m_1m_2}{}}\mathrm{\Psi }_{m_1m_2}^{m_d}(𝐩_s+{\displaystyle \frac{𝐤}{2}})\{{\displaystyle \frac{\epsilon _p^{}+\epsilon _s^{}}{2W_{\pi p}}}{\displaystyle \frac{1}{p_{s}^{}{}_{}{}^{2}p_{\pi p}^{}{}_{}{}^{2}i0}}`$ (22) $`\times [m_p|T_{\stackrel{~}{\pi }^0p\pi ^0p}(𝐩_{\pi p}^{},𝐩_s^{})|m_1m_n|T_{\gamma n\stackrel{~}{\pi }^0n}(𝐤_{\stackrel{~}{\pi }n},𝐪_{\stackrel{~}{\pi }n})|m_2`$ (23) $`m_p|T_{\stackrel{~}{\pi }^+n\pi ^0p}(𝐩_{\pi p}^{},𝐩_s^{})|m_1m_n|T_{\gamma p\stackrel{~}{\pi }^+n}(𝐤_{\stackrel{~}{\pi }n},𝐪_{\stackrel{~}{\pi }n})|m_2]\},`$ (24) where the meaning of all variables is easily understood from the corresponding ones in Eq. (15). Finally, let us consider diagram 1f which includes simultaneously both $`\pi `$-$`N`$ and $`n`$-$`p`$ rescattering. One has for the corresponding matrix element $`m_pm_n|T^{1f}(𝐤,𝐪,𝐩_n)|\lambda m_d=m{\displaystyle \frac{d^3𝐩_s}{(2\pi )^3}\frac{1}{p_{in}^2p_{out}^2i0}}`$ (25) $`\times {\displaystyle \underset{m_{\stackrel{~}{p}}m_{\stackrel{~}{n}}}{}}[𝐩_{out},m_pm_n|T_{np}|𝐩_{in},m_{\stackrel{~}{p}}m_{\stackrel{~}{n}}m_{\stackrel{~}{p}}m_{\stackrel{~}{n}}|T^{1d}(𝐤,𝐪,𝐩_s)|\lambda m_d`$ (26) $`+𝐩_{out},m_pm_n|T_{np}|𝐩_{in},m_{\stackrel{~}{p}}m_{\stackrel{~}{n}}m_{\stackrel{~}{p}}m_{\stackrel{~}{n}}|T^{1e}(𝐤,𝐪,𝐩_s)|\lambda m_d].`$ (27) The amplitude (27) includes as subblocks the amplitudes (15) and (24) but with off-shell outgoing nucleons so that $`p_{\pi N}^{}{}_{}{}^{2}`$ in Eq. (16) may be both positive and negative. One point which has to be emphasized is how to specify the energies in the loops. It is reasonable to suppose that the nucleon N$`{}_{\mathrm{S}}{}^{}{}_{}{}^{}`$ in the left loop is on its mass shell since such a choice precisely corresponds to that used when evaluating diagrams 1d and e. As to the right loop, a prescription for the nucleon energies used in the case of coherent $`\pi ^0`$ photoproduction on the deuteron consisted in putting the recoil nucleon on its mass shell in the pion photoproduction vertex . This prescription (which is referred to from here on as Case 1) was motivated by the location of the leading singularities of elementary amplitudes (the $`\mathrm{\Delta }`$ pole singularity) and internal particle propagators. It was found in Ref. that putting the recoil nucleon on its mass shell at the $`\pi `$-$`N`$ vertex led to a one and a half times larger rescattering amplitude. We think, however, that the prescription above can hardly be considered as a hard-and-fast rule. Even a glance on Fig. 1f enables one to see that there are no reasons why such a choice of nucleon energies should be preferred. In fact, we have checked other realistic possibilities to specify the nucleon energies in the right loop. One of them was to suppose that each of the nucleons carried equal parts of the total energy of the final $`n`$-$`p`$ pair, namely $`(\epsilon _n+\epsilon _p)/2`$ (Case 2). One more choice rejected in Ref. was to put the recoil nucleon on its mass shell at the $`\pi `$-$`N`$ vertex (Case 3). Summations over polarizations of the particles in Eqs. (6)-(27) as well as the three-dimensional integrations in Eqs. (11), (15) and (24) and six-dimensional one in Eq. (27) have been carried out numerically making use of the methods described in Refs. . Note, since Eq. (4) involves a three dimensional integration we were forced to carry out integrations with dimensions up to and including nine. Additional one-dimensional integrations were needed to solve the Lippmann-Schwinger equations for the $`N`$-$`N`$ and $`\pi `$-$`N`$ scattering amplitudes. The computing work was very hard but we were in a position with reasonable computer resources to evaluate the integrals with a precision providing an accuracy of calculated cross sections better than 10%. ## IV Results and Discussion Based on the approach described in the previous section, we calculated the differential cross section (4) and total cross section of the reaction (1) in the energy region from the threshold ($`E_\gamma ^{thr}=142.2`$ MeV) to 160 MeV. In Fig. 2 we present our predictions for the differential cross section with the ChPT set for the multipoles in Eq. (7) and with the choice Case 1 for nucleon energies in the right loop of diagram 1f. One can see that the contributions of all the diagrams in Fig. 1 are of the same order so that none of the diagrams can be ignored in the calculation. The two pole diagrams give a forward-peaked angular distribution. This reflects the fact that the ChPT values of the isospin averaged threshold electric dipole amplitude $`E_{0+}^{(+)}=\frac{1}{2}[E_{0+}^{p\pi ^0}+E_{0+}^{n\pi ^0}]=0.5`$ and the amplitude $`p_1^{(+)}=\frac{1}{2}[p_1^{p\pi ^0}+p_1^{n\pi ^0}]=8.9`$ are positive. Indeed, it is seen from Eq. (7) that the amplitudes $`E_{0+}^{(+)}`$ and $`p_1^{(+)}`$ are responsible for the forward-backward asymmetry in the differential cross section. An effect of $`n`$-$`p`$ final state interaction (diagram 1c) is strongly dependent on the photon energy. Just above threshold (e.g., at $`E_\gamma `$=145 MeV) it leads to increasing the differential cross sections over the full angular region and this increase is attributed to the strong attractive $`n`$-$`p`$ interaction in the singlet $`{}_{}{}^{1}S_{0}^{}`$-state. At higher photon energies the repulsive triplet $`{}_{}{}^{3}S_{1}^{}`$-interaction grows in importance. At $`E150`$ MeV the total effect of diagram 1c consists in a noticeable lowering of the cross sections except in the backward angle region. As to diagrams 1d and e with $`\pi `$-$`N`$ rescattering in the final state, they increase the differential cross section at all angles. At $`E155`$ MeV these diagrams contribute mainly at backward angles. Finally, diagram 1f gives a noticeable reduction of the cross section at forward angles. It is seen in Fig. 2 that after inclusion of diagrams 1d to f the differential cross section becomes to be backward peaked at $`E150`$ MeV. This means that due to $`\pi `$-$`N`$ rescattering the amplitude $`E_{0+}^{(+)}`$ effectively acquires a negative contribution altering its sign. When making such a statement we suppose (and with some justice, see below) that although rescattering can also modify the $`p_1^{(+)}`$ amplitude the latter remains to be positive. This effect was first discovered in the framework of the ChPT in the study of coherent $`\pi ^0`$ photoproduction on the deuteron. With the ChPT set for the multipoles in Eq. (7) we have calculated and presented in Fig. 3 the differential cross sections for various choices of nucleon energies in the right loop of diagram 1f. One can see that the results at $`\mathrm{\Theta }_\pi 90^{}`$ are practically independent of the choice of the energies. This is not the case for backward angles. Case 1 leads to minimum cross sections and Case 3 to maximum ones. Results with Case 2 lie between. Simultaneously, with increasing photon energy the deviation is declining. It is worth mentioning that analogous observations were made in Ref. in consideration of coherent $`\pi ^0`$ photoproduction in the threshold region. It was found in that paper that various approximations for the nucleon kinetic energy at integrations over loops gave noticeable changes in the calculated differential cross sections especially near threshold and backward angles. In Fig. 4 we compare our predictions for the angular distribution of the differential cross section with the ChPT and DR sets for the multipoles. It is seen that the ChPT and DR sets produce rather close cross sections at backward angles for $`E_\gamma 150`$ MeV though there exist noticeable deviations at forward angles. In the energy region under consideration the DR sets produce only half of the ChPT cross sections at forward angles. This reflects the significant difference between the ChPT and DR predictions for the threshold electric dipole amplitude $`E_{0+}^{n\pi ^0}`$ (see Sect. I) and the close consistency of the two for the $`p_1^{(+)}`$ amplitude. The previous consideration shows that $`\pi `$-$`N`$ rescattering (diagrams 1d-f) markedly changes the results calculated without this effect (diagrams 1a-c). It is instructive to study the question: what modification of the low energy parameters in the model without $`\pi `$-$`N`$ rescattering is required in order to reproduce the predictions of the full model. We will be interested in the amplitudes $`E_{0+}^{(+)}`$ and $`p_1^{(+)}`$ only. Let us suppose that due to $`\pi `$-$`N`$ rescattering these amplitudes acquire additional contributions $`\mathrm{\Delta }E_{0+}^{(+)}`$ and $`\mathrm{\Delta }p_1^{(+)}`$. Numerical estimates for them are obtained by switching off diagrams 1d-f and adjusting these parameters to reproduce the differential cross sections at $`\mathrm{\Theta }_\pi =0^{}`$ and $`180^{}`$ for the full model. Our results are given in Table II. One comment should be made here. In a rigorous consideration the values $`\mathrm{\Delta }E_{0+}^{(+)}`$ and $`\mathrm{\Delta }p_1^{(+)}`$ must be independent of the parametrization for the low energy amplitudes in Eq. (7), i.e. the former should be the same for ChPT and DR amplitudes. Indeed, the difference between ChPT and DR manifests itself in the $`\pi ^0`$ production amplitudes but the main contribution from $`\pi `$-$`N`$ rescattering stems from charged pion exchanges. The slight difference for ChPT and DR results in Table II is because of a rather rough numerical method of extracting $`\mathrm{\Delta }E_{0+}^{(+)}`$ and $`\mathrm{\Delta }p_1^{(+)}`$. Nevertheless, we believe that such a method properly reproduces the tendency of variation of $`E_{0+}^{(+)}`$ and $`p_1^{(+)}`$ due to $`\pi `$-$`N`$ rescattering. The dependence of $`\mathrm{\Delta }E_{0+}^{(+)}`$ and $`\mathrm{\Delta }p_1^{(+)}`$ on the choice of nucleon energies as seen in Table II is expected since the Cases 1 to 3 effectively lead to different pion propagators in Eq. (27). Averaging over all values in Table II we obtain $`\mathrm{\Delta }E_{0+}^{(+)}1.1`$. One can see that $`\mathrm{\Delta }E_{0+}^{(+)}`$ is negative in accordance with the ChPT result from Ref. being, however, nearly two times smaller in the absolute value. The $`p_1^{(+)}`$ amplitude acquires a positive contribution an averaged value for which is $`4.1\pm 0.6`$. This result being put together with the free-nucleon value of about 9 gives $`p_1=13.1\pm 0.6`$ and reasonably reproduces an experimental result of Ref. , $`p_1^{exp}=12.88\pm 0.28`$. In other words, our estimates confirm an assumption from that paper that a mechanism responsible for renormalizing the $`p_1`$ amplitude is the two-body process corresponding to $`\pi `$-$`N`$ rescattering. To make it much easier using our results in practice, we have parametrized the differential cross sections at four photon energies 145, 150, 155, and 160 MeV with the help of the following formula $`{\displaystyle \frac{d\sigma }{d\mathrm{\Omega }_\pi }}=A_1+A_2\mathrm{cos}\mathrm{\Theta }_\pi ,`$ (28) with parameters given in Table III. The values for $`A_1`$ and $`A_2`$ have been obtained by a fit to the average of the minimum and the maximum cross sections (see above) and the errors in $`A_1`$ account for variations of the cross sections both due to various choices of sets for multipoles and nucleon energies in the right loop of diagram 1f. The corresponding cross sections at arbitrary energies up to 160 MeV can be obtained by making use of a suitable interpolation procedure. We think that such a simple parametrization provides reasonable accuracy to be used in estimates of inelastic channel contributions to coherent $`\pi ^0`$ production. Performing in Eq. (4) an integration over the pion solid angle $`\mathrm{\Omega }_\pi `$ we can calculate the total cross section $`\sigma _{\mathrm{tot}}`$ of the reaction (1) (note that $`\sigma _{\mathrm{tot}}=4\pi A_1`$ where $`\sigma _{\mathrm{tot}}`$ is in units of $`\mu `$b). In Fig. 5 we show contributions of different diagrams to $`\sigma _{\mathrm{tot}}`$ with the ChPT set for the multipoles in Eq. (7). The two pole diagrams produce a total cross section which is rapidly increasing with photon energy. After inclusion of the $`n`$-$`p`$ final state interaction the energy distribution of $`\sigma _{\mathrm{tot}}`$ becomes to be more flat. Just above threshold this interaction in the $`{}_{}{}^{1}S_{0}^{}`$-wave leads to an increase of $`\sigma _{\mathrm{tot}}`$ but at $`E_\gamma 155`$ MeV the cross section diminishes by a factor of about 2. Diagrams 1d and e with $`\pi `$-$`N`$ rescattering give noticeable positive contribution to the total cross section. Of all the diagrams 1f is the least important one, decreasing $`\sigma _{\mathrm{tot}}`$ by about 10%. The total cross sections for various choices of nucleon energies in the right loop of diagram 1f are presented in the left panel of Fig. 6. It is seen that they are practically independent of a prescription for the energy. This reflects the smallness of the diagram 1f contribution to the total cross section. Our predictions for $`\sigma _{\mathrm{tot}}`$ with the ChPT and DR sets of the multipoles are also shown in Fig. 6. There is a deviation of about 15% in the full energy region. In Fig. 7 we compare our results with the ones of the model for the differential and total cross sections. One can see that the predictions of the two calculations are quite consistent. There is some disagreement at $`E_\gamma 155`$ MeV and forward angles for the differential cross section. It should be emphasized, however, that this disagreement cannot make any noticeable changes in the cross sections of coherent $`\pi ^0`$ photoproduction measured in Ref. because of smallness of the inelastic channel cross sections in comparison with the coherent channel ones. At the same time, at backward angles the inelastic channel is growing in importance but there the deviation is not very significant. One can see that the maximum deviation between the present calculation and the one of Ref. does not exceed 20% at 160 MeV and backward angles, being even smaller when the energy is decreasing. (Note here that the authors of that paper ascribed to their calculated cross sections of the inelastic channel uncertainties of about $`\pm 25\%`$ which were taken into account in their data analysis). In the case of the total cross section we observe good agreement with Ref. . An analysis performed by Bergstrom has shown that the $`p_1`$ amplitude is insensitive to the variation in the inelastic cross section from the present work and that from Ref. . Likewise, the extrapolation of the $`E_d`$ amplitude to threshold changes very little, remaining within the error in Eq. (12) of Ref. . In summary, we have performed an investigation of the reaction $`{}_{}{}^{2}H(\gamma ,\pi ^0)np`$ in the threshold region and found a good agreement with the only previous calculation for the total cross section. A reasonable agreement also exists for the differential cross sections. Our predictions coincide within of 20%. If one takes into account that the authors of that work ascribed to their results an uncertainty of about $`\pm 25\%`$ we conclude that the present calculation cannot lead to any changes in the total and differential cross section of coherent $`\pi ^0`$ photoproduction on the deuteron measured in Ref. . This means that the deviation of about 20% from theoretical predictions found in that paper for the $`E_d`$ threshold amplitude still remains to be resolved. Our calculations have shown that the difference of about $`410^3/\mu _{\pi ^+}^3`$ between the free-nucleon value for the $`p_1`$ threshold amplitude and the experimental value can be attributed to two-body effects due to $`\pi `$-$`N`$ rescattering. ###### Acknowledgements. We would like to thank J.C. Bergstrom for supplying us with results of his numerical calculations and for fruitful discussions. We are pleased to acknowledge helpful comments by A.I. L’vov. One of the authors (M.L.) is indebted to M.V. Galynsky for the use of his computer. This work was supported by Advance Research Foundation of Belarus and Deutsche Forschungsgemeinschaft under contract 436 RUS 113/510.
warning/0001/astro-ph0001084.html
ar5iv
text
# Signatures of galactic magnetic lensing upon ultra high energy cosmic rays ## 1 Introduction Galactic and intergalactic magnetic fields deflect extragalactic charged cosmic ray trajectories in their journey from their sources to the Earth, causing several effects upon the observed properties of ultra high energy cosmic rays (UHECRs). For a review see for instance . In a previous paper (in what follows paper I), we have shown that the regular component of the galactic magnetic field acts as a giant lens upon charged CRs, and this can sizeably amplify (or demagnify) the flux arriving from any given source, modifying its spectrum. The magnification of the CR flux by the galactic magnetic field becomes divergent for directions along critical curves in the sky seen from the Earth, corresponding to caustic curves in the “source plane”, i.e. in the corresponding directions outside the Galaxy. The location of the caustics move with energy and, as a caustic crosses a given source direction, pairs of additional images of the source appear or disappear. Multiple image formation is a rather common phenomenon, at least within the galactic magnetic field models considered in I. Indeed, the caustics sweep a rather significant fraction of the sky as the ratio $`E/Z`$ between the CR energy and charge steps down to values of the order of a few EeV ($`1\mathrm{EeV}=10^{18}\mathrm{eV}`$), before the drift and diffusive regimes turn on. At values of $`E/Z`$ larger than around 50 EeV caustics are present but sweep out a relatively small fraction of the sky. The effects under discussion are thus relevant even for the highest energy events so far detected if the CRs have a component which is not light. In the present paper we analyse in detail the energy dependence of the magnification of image pairs near a caustic, and discuss several implications of the existence of caustics upon the observed properties of UHECRs. In section 2 and in appendix A we show that if a source lies along a caustic at energy $`E_0`$, the magnification factor $`\mu `$ of each image in the pair that becomes visible at energies below $`E_0`$ diverges as $`\mu A/\sqrt{1E/E_0}`$. $`A`$ is a dimensionless constant, fixed by the structure of the magnetic field along the CR trajectory to the source. It must be determined numerically, and we do so for some examples of source locations. We discuss several implications of this result, such as the expected enhancement in the observed event rate at energies near a caustic, which can lead to an amplification bias, making it more likely to detect images of UHECR sources that lie along caustics than sources in ordinary regions. In appendix B we provide a simple proof of the relation between the magnification of the three images appearing when there are two nearby folds. In section 3 we show that the angular separation between images in a pair increases near the caustic as $`\mathrm{\Delta }\theta \sqrt{1E/E_0}`$. We analyse the expected event rate from the original source and its multiple images not only as a function of energy but also in terms of the observed arrival direction. We show that magnetic lensing near caustics is a source of clustering in the arrival directions of events with comparable energy. In section 4 we stress the fact that at fixed energy the magnetic lensing effects depend upon the CR electric charge, and show that this dependence can significantly alter the observed CR composition at the highest energies. In section 5 we discuss features of the time delay between events from multiple images, relevant in the case of bursting or highly variable sources of UHECRs . We show that the delay between the arrival of events of equal energy from the two images in a pair increases as $`\mathrm{\Delta }t\sqrt{1E/E_0}`$ around a caustic. We also show that the time delay between events at different energy from one image in a pair may monotonically decrease with decreasing energy near the caustic, opposite to the behaviour of time delays in normal regions. Section 6 rounds up our conclusions. The galactic magnetic field model that will be used throughout this paper to illustrate magnetic lensing effects is the bisymmetric spiral model with even symmetry (BSS-S) described in paper I, which is a smoothed version of one of the models used in and to study the effects of CR deflections in our galaxy. We refer the reader to paper I for details about the field configuration and about the numerical methods implemented to determine CR trajectories and flux magnifications. ## 2 Flux enhancement by magnetic lensing near critical points As shown in paper I, the galactic magnetic field can act as a giant lens that amplifies or demagnifies extragalactic sources of UHECRs, much alike the gravitational lensing effect upon distant quasars by intervening matter along the line of sight . CRs from a distant extragalactic source that enter the galactic halo from a direction $`(\mathrm{},b)_H`$ (in galactic coordinates) are deflected by the magnetic field and thus observed on Earth as if coming from a different direction $`(\mathrm{},b)_E`$, and their flux is amplified (or demagnified) by a factor $`\mu `$. The magnitude of the effect depends upon the direction of observation and upon the ratio $`E/Z`$ between energy and charge of the CR. The effect is most dramatic at the critical curves of the lens mapping, the directions along which the observed flux diverges for a fixed value of $`E/Z`$. The corresponding lines in the source coordinates (the direction from which CRs enter the galactic halo) are the caustics of the lens mapping. The location of the caustics changes with energy, for fixed $`Z`$. If a source position lies along a caustic of the magnetic lens mapping at energy $`E_0`$, a pair of images of the source either becomes visible or dissapears at energies below $`E_0`$. Their magnification diverges at $`E=E_0`$. This behaviour was illustrated in paper I for sources at galactic coordinates $`(\mathrm{},b)=(282.5^{},74.4^{})`$ (M87 in the Virgo cluster) and $`(\mathrm{},b)=(320^{},30^{})`$ (visible from the southern hemisphere). In appendix A we give a geometrical interpretation of the magnification of a pair of images near a caustic. Similar to the gravitational lens case , the magnification diverges as $`1/\sqrt{x}`$, where $`x`$ measures the distance of the source to the caustic. In the magnetic lensing case the location of the caustics “move” with energy, and thus the magnification of a pair of images diverges as the energy approaches the energy at which the source lies along the caustic as $`1/\sqrt{E_0E}`$. In appendix A we also find the first two corrections to this leading order behaviour, and show that the magnification of the two images behaves as: $$\mu _i(E)\frac{A}{\sqrt{1E/E_0}}\pm B+C_i\sqrt{1E/E_0}.$$ (1) The dimensionless coefficient $`A`$ is the same for the two images. The constant term $`\pm B`$ has the same value with opposite sign for each image. The third term has a different coefficient $`C_i`$ for each image due to the truncation of the expansion up to this order. The value of the coefficients $`A,B`$ and $`C_i`$ are fixed by the properties of the magnetic field along the CR trajectories. We determine them through a fit to the numerical output for the energy dependence of the amplification, evaluated through the method described in paper I. Figure 1 displays the numerical result for the amplification near the caustics along with the fit to the analytic expression in eq. (1), for the same examples of source locations as in paper I. The top panel corresponds to M87 and the bottom panel to the source in the southern hemisphere. The figures display the magnification of the principal image (the one that is also visible at the highest energies) and of the images $`A`$ and $`B`$, visible at energies below the caustic only. The analytic expression (1) fits with very high accuracy the numerical result for the magnification of the secondary images near the caustic, at least down to energies 10% below the energy $`E_0`$ of the caustic. The fit to M87 determines $`E_0/Z=20.41`$ EeV, $`A=1.3`$, $`B=4.0`$, $`C_A=3.1`$, $`C_B=5.0`$. The fit to the southern source fixes $`E_0/Z=15.425`$ EeV, $`A=0.44`$, $`B=0.37`$, $`C_A=0.19`$, $`C_B=0.28`$. The divergence of the flux magnification of a CR source at $`E=E_0`$ is softened down to finite values when integrated across an extended source. This still allows for extremely large magnifications. The limiting factor to the maximum attainable magnification in a realistic situation is not the extended nature of the sources, but the fact that the divergence in the magnification, even in the case of a point source, arises only at a fixed energy $`E_0`$. Since realistic sources are not monoenergetic, the integrated flux of a magnified source around $`E_0`$ is thus always finite, even in the point source approximation. Large magnification of image pairs around caustics leads to a significant enhancement of the detection probability of a source in a flux-limited sample. In gravitational lensing this effect is termed “amplification bias”, and may be responsible for the observation of quasars that would otherwise be too dim to detect. Let us now consider the strength of this effect due to galactic magnetic lensing of UHECRs. Consider a differential flux of CRs injected in the galactic halo at energies beyond the “ankle” given by $`dF=F_0(E_0/E)^{2.7}dE`$. The flux observed on Earth coming from two images formed at a caustic at energy $`E_0`$, in the energy interval between $`0.9E_0`$ and $`E_0`$, is $$F_{A+B}(0.9E_0<E<E_0)2_{0.9E_0}^{E_0}dE\frac{\mathrm{d}F}{\mathrm{d}E}\frac{A}{\sqrt{1E/E_0}}12A_{0.9E_0}^{E_0}dE\frac{\mathrm{d}F}{\mathrm{d}E}.$$ (2) We have neglected a small $`O(\sqrt{1E/E_0})`$ correction (the term proportional to $`C_A+C_B`$). The flux observed from the two images in this energy interval near the caustic is $`12A`$ times larger than the flux of the principal image of the source in the same energy range in the absence of magnification. This is also $`2.4A`$ times the flux that would arrive from the principal image at all energies above $`E_0`$ if there were no magnetic lensing. Detection of an UHECR source in a narrow energy range around a caustic may thus be more likely than its detection at any higher energy. It may also be more likely to detect the CR source at energies around the caustic than at significantly lower energies (say half the energy of the caustic or even less) in cases where both the principal image as well as the secondary images are not magnified at energies below the caustic. Large enhancements of the observed flux in a narrow energy range also occur in the interesting case when the source position lies along two caustics at nearby energies so that the principal image of the source (the only one visible from Earth at the highest energies) has also divergent magnification. Such a situation is exemplified in figure 2, for a source located at $`(\mathrm{}=90^{},b=10^{})`$. A pair of images becomes visible at energies below $`E_0^{III}=29.31`$ EeV, one of which disappears in a caustic along with the principal image at an energy $`E_0^I=22.31`$ EeV. The secondary image that merges with the principal image is that with opposite parity. In appendix B we develop the geometrical interpretation and analytic approximation to the magnification of images for two nearby folds. The analog result in the gravitational lens case was obtained in . As shown in appendix B, and verified numerically with high accuracy, between the folds the sum of the magnifications of the two images with equal parity coincides with the magnification of the image with opposite parity, up to a constant term. The analytic fits in figure 2 to the divergent magnification of the principal image $`(I)`$ at $`E_0^I`$ and to the magnification at $`E_0^{III}`$ of the image that survives at low energies $`(III)`$, lead to $`A_I=0.275`$, $`B_I=0.15`$, $`C_I=0.63`$; $`A_{III}=.096`$, $`B_{III}=0.11`$, $`C_{III}=0.03`$. The magnification of the image that is visible at energies between $`E_0^I`$ and $`E_0^{II}`$ is fit to $`\mu _{II}=\mu _I+\mu _{III}0.7`$. ## 3 Angular distribution of events: clustering around caustics The arrival directions of the so far detected CRs in the highest energy range is compatible (within the limited statistics available) with an isotropic distribution, except for some small angle clustering of events (eight doublets and two triplets within a total of 92 events with energies above 40 EeV). The observed relatively uniform distribution does not preclude the possibility that UHECRs originate from very few nearby sources, if their trajectories underwent significant energy-dependent deflections in their journey to the Earth. Trajectories of UHECRs with $`E/Z`$ below approximately 50 EeV are sensibly deflected in the magnetic field model considered in this paper. Several quite separated events may actually originate from the same source if CRs have a component which is not light (see paper I and references therein). It has even been speculated that all the events so far detected at energies above $`10^{20}`$ eV may come from M87 in the Virgo cluster, if the Galaxy has a rather strong and extended magnetic wind. In paper I we illustrated the angular displacement in the arrival directions of CRs from the principal and secondary images of a magnetically lensed source. Here we analyse in more detail the angular displacement of image pairs around caustics. We argue that caustics can produce a significant clustering of arrival directions of events with comparable energies. It may well be the case (depending on the sources flux and composition, and on the exact nature of the galactic magnetic field) that a large fraction of the trajectories of the UHECRs so far detected have been significantly bent and do not point to their sources. In this scenario clustering of events would be infrequent, at relatively low energies due to large deflections and at high energies due to the smaller flux. However, when a source is near a caustic, since the significant enhancement of its flux occurs within a narrow energy range, the CRs arrive from relatively nearby directions, leading to an angular concentration of events. We illustrate the clustering effect in figure 3. We consider the same examples of source locations as in the previous section. We assume that the differential flux injected by the source scales as $`E^{2.7}`$, and that the detecting system has the same efficiency at all energies within the range considered. The energy range was divided in 50 bins of equal detection probability. Figure 3 displays the predicted arrival directions of the events detected from each source. In the case of M87 (top panel) 11 events out of a total of 50 with $`E/Z`$ larger than 10 EeV (half the energy of the caustic) fall in the narrow energy range between the energy of the caustic and just 10% below (’+’ signs). 3 of those 11 events are in the principal image and 8 in the secondary images. Notice that only 5% of the events would fall in the same energy range if the source were not magnified. Notice also that just 8 events correspond to the principal image at all energies higher than the energy of the caustic. The angular clustering effect for M87 is significant but not extremely large (in this magnetic field model), partly due to the fact that the principal image is also largely magnified around the energy of the caustic, and partly because deflections are quite large. The middle panel in figure 3 displays the effect for the source located at $`(\mathrm{},b)=(320^{},30^{})`$. In this case 9 events fall within 10% of the energy of the caustic in the secondary images, separated by no more than $`5^{}`$. Only 4 events are seen for all energies higher than the energy of the caustic, scattered over almost $`10^{}`$. The events at energies below the caustic, down to half its energy, are scattered over more than $`20^{}`$. The bottom panel in figure 3 corresponds to the source located at $`(\mathrm{},b)=(90^{},10^{})`$, which presents two nearby folds, as depicted in figure 2. In this case nearly half of the events (23) fall within the energy range between the two caustics $`(22.3\mathrm{EeV}<E/Z<29.3\mathrm{EeV})`$, while 15 events occur for all higher energies, and just 12 events appear at lower energies down to $`E/Z=7\mathrm{EeV}`$. The relatively small number of events in the lower energy range is due to the large deamplification of the image flux. An unlensed source would have instead 86% of the events in the lower energy range considered, 9% in the higher energy range, and just 5% at the intermediate energies. Notice that while the arrival directions of the events considered are spread over more than $`15^{}`$, nearly half of the events fall within just $`2^{}`$ around $`(l=88^{},b=11^{})`$. We stress the fact that six out of the eight doublets in the UHECR data listed in table 6 in are such that the energies of the events in a pair differ by less than 10%. The same happens with two events in one of the triplets. This may be an indication that at least a fraction of the observed clustering of events may be due to magnetic lensing around caustics. Figure 3 (and figure 6 in paper I) illustrate the angular displacement of the images as a function of energy. As discussed in appendix A, the observed angular separation between images scales as $`\sqrt{x}`$, where $`x`$ measures the distance from the source to the caustic. Consequently, the angular separation between the pair of images created at the caustic crossing scales with energy near a caustic as $$\mathrm{\Delta }\theta \mathrm{\Theta }\sqrt{1E/E_0}.$$ (3) The proportionality constant $`\mathrm{\Theta }`$ is fitted from the numerical output for the angular displacement of the images. In the case of M87, $`\mathrm{\Theta }56^{}`$, and in the case of the source at $`(\mathrm{},b)=(320^{},20^{})`$ it is $`\mathrm{\Theta }15.1^{}`$. ## 4 Effects upon composition Another interesting effect of magnetic lensing is that it can strongly alter the measured composition of CRs at high energies. This is because at each source location the magnification is a function of $`E/Z`$. Thus, for a given energy the CR flux of components with different $`Z`$ suffers different (de)magnification as the CRs travel through the Galactic magnetic field. To give an idea of the strength of this effect, let us consider a simple example: suppose that from a given source the CR flux arriving to the halo is composed by a fraction $`f`$ of protons and a fraction $`1f`$ of heavier nuclei with charge $`Z_h`$. In the absence of magnetic lensing the mean value of $`Z`$ of the arriving particles would be $`Z=f+(1f)Z_h`$. Due to the effect of the magnetic field, the flux of protons on Earth will be (de)magnified by $`\mu (E)`$ and the flux of heavier nuclei by $`\mu (E/Z)`$. Thus, the mean value of $`Z`$ of the CRs arriving on Earth will be $$Z=\frac{f\mu (E)+(1f)Z_h\mu (E/Z_h)}{f\mu (E)+(1f)\mu (E/Z_h)}.$$ (4) As the magnifications can be large, $`Z`$ can be strongly modified by the magnetic field. The effect is specially important for regions of the sky for which the total magnification $`\mu `$ (adding the contribution from all the images) has noticeable changes as $`E/Z`$ varies. We show as an example the source located at $`(\mathrm{},b)=(90^{},10^{})`$, which undergoes the magnification depicted in figure 2. We took for reference a flux composed by a fraction $`f=0.5`$ of protons and a fraction 0.5 of Nitrogen ($`Z_h=7`$). In the absence of lensing the mean value of $`Z`$ would be $`Z=4`$. We see in figure 4 that a clear change from a light composition in the smaller energy region to a heavier composition at the larger energies appears due to magnetic lensing. ## 5 Time delays A charged UHECR that traverses a distance $`L`$ within a homogeneous magnetic field $`B`$ perpendicular to its trajectory is deflected by an angle of the order of $`\eta 5^{}(10\mathrm{EeV}Z/E)(B/\mu \mathrm{G})(L/\mathrm{kpc})`$, in the limit of small deflections. Consider two initially parallel CRs emitted simultaneously, one with a much lower ratio $`E/Z`$ than the other. If they enter a region permeated by a homogeneous magnetic field and converge to the same point after a distance $`L`$, the CR with lower ratio $`E/Z`$ arrives later, with a relative time delay $`\delta t\eta ^2L/210\mathrm{yrs}(10\mathrm{EeV}Z/E)^2(B/\mu \mathrm{G})^2(L/\mathrm{kpc})^3`$. Higher energy events arrive earlier. Time delays induced by intergalactic magnetic fields are an essential ingredient in bursting models for the origin of UHECRs . There are very definite observational signatures of such a scenario. For instance, individual bursting CR sources would have very narrow observed spectra, since only CRs with a fixed time delay would be observed at any given time. The energy at the peak in the differential CR flux received on Earth should shift with time as $`t^{1/2}`$. Here we analyse the implications of the regular component of the galactic magnetic field upon time delays between events from a UHECR source at different energies, and between events from different images of a single magnetically lensed source. Since UHECRs are extremely relativistic, their time delay compared to straight propagation at the speed of light from the source to the observer is simply determined by the excess path length. We assume that the extragalactic UHECR sources are sufficiently distant that we can approximate the CR flux incident upon the galactic halo as a beam of parallel trajectories, all emitted simultaneously. We then numerically determine the difference in path length between the trajectory of a charged CR and the path length along the parallel straight trajectory in the beam that reaches the Earth. In other words, we measure the delay in the arrival time of a charged CR with respect to the arrival time of photons (or charged CRs with much higher energy) emitted simultaneously from a distant source. Figure 5 displays the time delay with respect to straight propagation for the principal and secondary images of the source locations discussed in the previous sections. The time delay in normal regions (far from critical points) monotonically increases with decreasing energy. The spiraled galactic magnetic field is far from homogeneous, and thus departures from the $`E^2`$ dependence are certainly expected. We have checked that the time delay averaged over a regular grid of arrival directions scales, in the BSS-S galactic magnetic field model considered, as $`<\delta t>1000\mathrm{yrs}(10\mathrm{EeV}Z/E)^2`$ from very high energies down to $`E/Z`$ of order 5 EeV. At lower $`E/Z`$ values the increase in the time delay with decreasing energy is faster, since a sharp transition from quasirectilinear to drift motion occurs at values of $`E/Z`$ between 3 and 1 EeV. What strikes the eye in figure 5 is that the time delay of one of the images in a pair can have an energy dependence opposite to that in normal regions. Indeed, the time delay of one of the images that are visible at energies below the energy of the caustic increases with decreasing energy, while the time delay of the other member of the pair decreases. Thus, the relative arrival time of events from a single image of a CR source does not necessarily increase with decreasing energy. It is often argued that the doublets in which the highest energy event arrived later than the other member in the pair can not arise from bursting sources. As we have seen, this is not necessarily true near a caustic. The relative arrival time delay between equal energy events from different images $`A`$ and $`B`$ in a pair behaves near the caustic as $$\mathrm{\Delta }t=\delta t_A\delta t_BT\sqrt{1E/E_0}.$$ (5) This can be understood as follows, in the limit of small deflections. The trajectory of each of the images in the pair is deflected by $`\eta `$ and $`\eta +\mathrm{\Delta }\theta `$ respectively, with $`\mathrm{\Delta }\theta `$ given by eq. (3). The time delays with respect to straight propagation are thus proportional to $`\eta ^2`$ and $`\eta ^2+2\eta \mathrm{\Delta }\theta `$ respectively, as long as $`\mathrm{\Delta }\theta \eta `$. Thus the relative time delay between events at the same energy from the two images scales as $`\mathrm{\Delta }\theta `$ at energies sufficiently close to that of the caustic. The fit of eq. (5) to the numerical output is highly accurate, with $`T3200`$ yrs in the case of M87 and $`T1030`$ yrs in the case of the source in the southern sky at $`(\mathrm{}=320^{},b=30^{})`$. ## 6 Conclusions UHECRs beyond the ankle in the spectrum are most probably nuclei of extragalactic origin. In their way to the Earth they feel the magnetic field structure permeating the Galaxy and hence their trajectories become deflected and their fluxes are lensed. The implications of this magnetic lensing are manyfold and they have to be taken into account in the analysis of the observations. Since the lensing effect depends on the energy it can sizeably affect the observed spectrum of the sources. Moreover, for sources located in a large fraction of the sky CRs can reach the Earth following different paths, and hence multiple images of those sources will appear. A useful way to visualise these effects is to display the mapping of the arrival directions in the Earth into the incoming directions outside the Galaxy, as was introduced in paper I with what was dubbed a ‘sky sheet’. In the locations where this surface develops folds, pairs of additional images of the source are present. These folds, corresponding to the caustics, move with energy, and as they cross a given source location pairs of additional images appear or disappear. Near the fold the magnification of each image in the pair diverges as $`\mu 1/\sqrt{|EE_0|}`$. Thus the probability to detect events from a given source is noticeably enhanced for energies close to $`E_0`$, at which the caustic crosses the source location. This also leads to an expected concentration of events near the location at which the new pair of images appears. This is relevant in the analysis of the small scale clustering present in the UHECR distribution and in this respect it is remarkable that the observed events in doublets and triplets in most cases are very close in energies. With the increased statistics expected with the new detectors, such as Auger or High Res , these features may become testable through a careful analysis of the clustering of the events. Also the observed CR composition can be affected by magnetic lensing due to the dependence of the flux amplification on $`E/Z`$. Nuclei with different charges are magnified by different amounts for a given energy. This effect is sizeable for sources whose magnification has a strong energy dependence (in particular when there are caustics) and which have a mixed composition. Another feature which we have discussed is the time delay due to the galactic field between the different images of a lensed source. These are typically larger than the lifetimes of the CR observatories, and are strongly dependent on the energy. If burst sources exist, narrow spectra will then result at any given time, and the different images will be observed simultaneously with different $`E/Z`$ values. We have to stress that in addition to the effects related to the magnetic field of the Galaxy discussed in this paper there may be also similar effects associated to the magnetic field in the source galaxy or even to the intergalactic fields if these ones are strong. Also the magnetic field model adopted here is plausible but the real one may differ from it, changing the quantitative details but not the general qualitative results. ###### Acknowledgments. Work partially supported by ANPCyT, CONICET and Fundación Antorchas, Argentina. ## Appendix A Magnification near caustics The relation between the source position and the image positions is given by a mapping from the source coordinates $`(\mathrm{},b)_H`$ to the observer’s ones $`(\mathrm{},b)_E`$. When this mapping is multiple valued there will be additional image pairs appearing (see paper I). The magnification of a given image will be $`\mu =\mathrm{d}\mathrm{\Omega }_H/`$d$`\mathrm{\Omega }_E`$, i.e. the ratio of differential solid angles associated to the mapping. To visualise these facts it is useful to consider the inverse map $`(\mathrm{},b)_E(\mathrm{},b)_H`$ and look at it as the mapping of a surface (the observer’s sky) into another surface (dubbed the ‘sky sheet’ in I), which will be folded when multiple images appear, and whose stretching is related to the magnification of the images. The location of the folds correspond to the caustics along which image pairs with divergent magnification appear and the magnification varies very rapidly with the distance to the fold. Since the folds move as the energy is decreased, the knowledge of the magnification as a function of the angular distance to the fold for a given energy can be used to obtain the magnification, for a given source, as a function of the energy near a caustic, which is the quantity of interest to us here. The angular dependence of the magnification near a fold has a very simple geometrical interpretation in terms of the folded sky sheet. Let us take local angular coordinates $`(x,y)`$ in the source sky such that $`x=0`$ describes locally the location of the fold (and hence the fold is along the $`y`$ axis while the $`x`$ axis is orthogonal to it). We can also adopt (non–orthogonal) local coordinates on Earth $`(X,Y)`$ such that $`X=0`$ is mapped into $`x=0`$ and similarly $`Y=0`$ is mapped into $`y=0`$. Furthermore, we will assign a third coordinate to the mapping, i.e. $`(X,Y)(x,y,z)`$, giving a depth to the sky sheet so that the fold can be visualised. The simplest choice for this (arbitrary) third coordinate is to take $`z=X`$, as we will do in the following. The results will be easier to obtain with this choice, but are independent of it. In this way the mapping for a fixed value of $`Y`$ in a neighbourhood of $`X=0`$ will look as shown in figure 6. The magnification is given by $`\mu =|K\mathrm{d}X/\mathrm{d}x|`$, where $`K[\mathrm{cos}b_H|(\mathrm{},b)_H/(x,y)|]/[\mathrm{cos}b_E|(\mathrm{},b)_E/(X,Y)|]\mathrm{d}y/\mathrm{d}YK_0(Y)+K_1(Y)X+O(X^2)`$. The next step is to relate the factor d$`X`$/d$`x=\mathrm{d}z/`$d$`x`$ with the slope of the curve in figure 6, and approximate the fold near the caustic by its Taylor expansion $`x=az^2+bz^3+O(z^4)`$. Hence we have $$\frac{\mathrm{d}x}{\mathrm{d}z}2az+3bz^2$$ (6) and hence, since $`z=\pm \sqrt{x/a}(1+O(x))`$, with the sign indicating to which side of the fold the image belongs, we have $$\mu =\frac{A^{}}{\sqrt{x}}\pm B^{}+O(\sqrt{x}),$$ (7) $`A^{}`$ and $`B^{}`$ being $`x`$-independent parameters related to the parameters $`a`$, $`b`$, $`K_0`$ and $`K_1`$. The final step is to relate this with the magnification for one given source as a function of energy. Suppose that at energy $`E_0`$ a source located at $`(\mathrm{}_0,b_0)_H`$ would be just on top of the fold. If we define the angular ‘velocity’ of the fold as the energy is changed as $`V(E)\mathrm{d}x_f/\mathrm{d}E`$, with $`x_f`$ being the distance from the source to the fold ($`x_f=0`$ for $`E=E_0`$), one has $`x_fV(E_0)|E_0E|`$. Hence we finally get $$\mu (E)=\frac{A}{\sqrt{1E/E_0}}\pm B+O(\sqrt{1E/E_0}).$$ (8) ## Appendix B Magnification for two nearby folds The previous Appendix dealt with the magnification for the two images which appear when a fold crosses the source location. The third image (e.g. the one present originally) was assumed to be far from the fold so that its magnification was supposed to have a non-singular behaviour. Another situation of interest is when two folds are nearby (e.g. when the sky–sheet has a narrow fold which moves with energy across the source, rather than becoming wider and remaining on top of the source, or when the source is near a cusp where two folds merge) so that the three images can have large magnifications. When a couple of nearby folds crosses a source, we see that a pair of images appear at some energy and at a somewhat smaller energy one of the new images merges with the original one present at high energies and they disappear (figure 2). In this case there is a relation between the magnification of the three images, as we now show. A similar relation holds in the gravitational lens case . Following similar lines as in the previous Appendix, the folded surface can be described as in figure 7. Without loss of generality we can choose the origin $`x=0`$ so that the vertical coordinates of the other two images at $`z_1`$ and $`z_2`$ in figure 7 satisfy $`z_2=z_1`$. In this case a Taylor expansion of the fold will be $`x=az^3cz`$ (with $`a,c>0`$ and no quadratic term due to our choice of origin for $`x`$). The location of the folds are at $`x_I=2a(c/3a)^{3/2}`$ and $`x_{II}=x_I`$. The images in region I ($`z>z_I=\sqrt{c/3a}`$) and III ($`z<z_{II}=z_I`$) will have positive parities, while the one in region II ($`z_{II}<z<z_I`$) will have negative parity. The vertical coordinates of the three images for a given $`x`$ are simply obtained as $$z_k=2z_I\mathrm{cos}\frac{\alpha +2k\pi }{3}(k=0,1,2)$$ (9) where $`\mathrm{cos}\alpha =x/x_{II}`$. Here $`k=2`$ corresponds to the negative parity image (region II) while $`k=0`$ and 1 are the images in region I and III respectively. The magnification of the images will be $`\mu _k=K|\mathrm{d}z_k/\mathrm{d}x|`$, where $$\frac{\mathrm{d}z_k}{\mathrm{d}x}=\frac{2z_I}{3x_{II}\sqrt{1(x/x_{II})^2}}\mathrm{sin}\left[\frac{\mathrm{cos}^1(x/x_{II})}{3}+\frac{2k\pi }{3}\right].$$ (10) From this we find that $$\left|\frac{\mathrm{d}z_0}{\mathrm{d}x}\right|+\left|\frac{\mathrm{d}z_1}{\mathrm{d}x}\right|=\left|\frac{\mathrm{d}z_2}{\mathrm{d}x}\right|,$$ (11) i.e. that the sum of the magnifications of the two positive parity images coincides with the magnification of the negative parity image. Next to leading terms (i.e. corrections of $`O(\sqrt{x})`$ will add constant terms to the magnifications, so that the final result up to corrections of $`O(x)`$ is $$\mu _I+\mu _{III}=\mu _{II}+\mathrm{const}.$$ (12) This theorem is illustrated in figure 2.
warning/0001/gr-qc0001021.html
ar5iv
text
# Second order perturbations of a Schwarzschild black hole: inclusion of odd parity perturbations ## I Introduction There has been a recent surge of interest in black hole perturbation theory, stemming from the successful applicability of such a formalism in the case of black hole collisions (see for a review). In particular, the use of second order perturbative computations has proven quite useful to endow the perturbative formalism with “error bars” and also to increase its accuracy. The computations associated with the second order formulation are quite laborious. In particular, the second order perturbative equations are linear differential equations that contain “source” terms that are quadratic in the first order perturbations. This forbids us from doing a “generic” second order computation, since it would involve an infinite number of source terms. Computations have to be restricted to certain multipolar orders and parities to achieve a computable expression. The expressions we have presented in the past are all for even parity axisymmetric $`\mathrm{}=2`$ first order perturbations. This is a significant case, since for first order perturbations such modes are the ones that dominate the initial data for black hole collisions (at least in the head-on case, in the non-head on case one also has non-axisymmetric $`m=\pm 2`$ modes). If the black holes individually have spin, however, odd parity terms appear as well. If one wishes to study such collisions with the second order formulation one needs to lay out the second order equations for odd and even parity perturbations. Since the “source” terms that appear in the second order equations are quadratic, one has quadratic contributions from odd parity first order modes in the even parity second order equations. The intention of this paper is to present a comprehensive account of the perturbative equations involving odd terms, up to second order. In a separate publication we will apply the formalism to the collision of counterrotating black holes. ## II Perturbative formalism: review of some first order results ### A The Regge–Wheeler decomposition and the Regge–Wheeler gauge In order to fix notation and to set the computational philosophy for the rest of the paper, we recollect here several results of first order perturbation theory that are known, but that it is useful to have laid out in a compact fashion for even and odd perturbations together. We consider the axially symmetric perturbations of a spherically symmetric background, up to and including second order, using the Regge–Wheeler framework. In first order perturbations there is no loss of generality in the consideration of axisymmetric perturbations since modes with different $`m`$ value decouple. This is not the case, however, in second order theory and our treatment is therefore restricted to the axisymmetric case. In terms of the usual Schwarzschild-like coordinates $`t,r,\theta ,\varphi `$, any $`\mathrm{}=2`$ axially symmetric perturbation may be written in the “Regge–Wheeler form” $`\stackrel{~}{g}_{tt}`$ $`=`$ $`(12M/r)\left\{1\left[ϵ\stackrel{~}{H}_0^{(1)}(r,t)+ϵ^2\stackrel{~}{H}_0^{(2)}(r,t)\right]P_2(\theta )\right\}`$ (1) $`\stackrel{~}{g}_{rr}`$ $`=`$ $`(12M/r)^1\left\{1+\left[ϵ\stackrel{~}{H}_2^{(1)}(r,t)+ϵ^2\stackrel{~}{H}_2^{(2)}(r,t)\right]P_2(\theta )\right\}`$ (2) $`\stackrel{~}{g}_{rt}`$ $`=`$ $`\left[ϵ\stackrel{~}{H}_1^{(1)}(r,t)+ϵ^2\stackrel{~}{H}_1^{(2)}(r,t)\right]P_2(\theta )`$ (3) $`\stackrel{~}{g}_{t\theta }`$ $`=`$ $`[ϵ\stackrel{~}{h}_0^{(1)}(r,t)+ϵ^2\stackrel{~}{h}_0^{(2)}(r,t)]P_2(\theta )/\theta )`$ (4) $`\stackrel{~}{g}_{t\varphi }`$ $`=`$ $`[ϵ\stackrel{~}{h}_{0o}^{(1)}(r,t)+ϵ^2\stackrel{~}{h}_{0o}^{(2)}(r,t)]\mathrm{sin}(\theta )P_2(\theta )/\theta )`$ (5) $`\stackrel{~}{g}_{r\varphi }`$ $`=`$ $`[ϵ\stackrel{~}{h}_{1o}^{(1)}(r,t)+ϵ^2\stackrel{~}{h}_{1o}^{(2)}(r,t)]\mathrm{sin}(\theta )P_2(\theta )/\theta )`$ (6) $`\stackrel{~}{g}_{r\theta }`$ $`=`$ $`[ϵ\stackrel{~}{h}_1^{(1)}(r,t)+ϵ^2\stackrel{~}{h}_1^{(2)}(r,t)]P_2(\theta )/\theta )`$ (7) $`\stackrel{~}{g}_{\theta \theta }`$ $`=`$ $`r^2\{1+[ϵ\stackrel{~}{K}^{(1)}(r,t)+ϵ^2\stackrel{~}{K}^{(2)}(r,t)]P_2(\theta )`$ (9) $`+[ϵ\stackrel{~}{G}^{(1)}(r,t)+ϵ^2\stackrel{~}{G}^{(2)}(r,t)]^2P_2(\theta )/\theta ^2)\}`$ $`\stackrel{~}{g}_{\varphi \varphi }`$ $`=`$ $`r^2\{\mathrm{sin}^2\theta +\mathrm{sin}^2\theta [ϵ\stackrel{~}{K}^{(1)}(r,t)+ϵ^2\stackrel{~}{K}^{(2)}(r,t)]P_2(\theta )`$ (11) $`+\mathrm{sin}(\theta )\mathrm{cos}(\theta )[ϵ\stackrel{~}{G}^{(1)}(r,t)+ϵ^2\stackrel{~}{G}^{(2)}(r,t)]P_2(\theta )/\theta )\}`$ $`\stackrel{~}{g}_{\theta \varphi }`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left[ϵ\stackrel{~}{h}_{2o}^{(1)}(r,t)+ϵ^2\stackrel{~}{h}_{2o}^{(2)}(r,t)\right]\left[\mathrm{cos}(\theta )P_2(\theta )/\theta \mathrm{sin}(\theta )^2P_2(\theta )/\theta ^2\right]`$ (12) where $`P_2(\theta )=3\mathrm{cos}^2(\theta )/21/2`$. We introduce the expansion parameter $`ϵ`$, which defines the perturbation order, and identify the corresponding metric coefficients with the superscripts $`(1)`$, $`(2)`$ and use the subscript “o” to refer to the odd parity quantities. As we stated above, any axisymmetric perturbation can be put into the Regge–Wheeler form. However, for calculational simplicity, it is usually convenient to take advantage of the coordinate freedom to restrict somewhat the form of the metric. An example of such choice is the so called Regge - Wheeler gauge (RWG), where the nonvanishing metric coefficients are $`g_{tt}`$ $`=`$ $`(12M/r)\left\{1[ϵH_0^{(1)}(r,t)+ϵ^2H_0^{(2)}(r,t)]P_2(\theta )\right\}`$ (13) $`g_{rr}`$ $`=`$ $`(12M/r)^1\left\{1+[ϵH_2^{(1)}(r,t)+ϵ^2H_2^{(2)}(r,t)]P_2(\theta )\right\}`$ (14) $`g_{rt}`$ $`=`$ $`[ϵH_1^{(1)}(r,t)+ϵ^2H_1^{(2)}(r,t)]P_2(\theta )`$ (15) $`g_{t\theta }`$ $`=`$ $`[ϵh_{0o}^{(1)}(r,t)+ϵ^2h_{0o}^{(2)}(r,t)]\mathrm{sin}(\theta )P_2(\theta )/\theta `$ (16) $`g_{r\theta }`$ $`=`$ $`[ϵh_{1o}^{(1)}(r,t)+ϵ^2h_{1o}^{(2)}(r,t)]\mathrm{sin}(\theta )P_2(\theta )/\theta `$ (17) $`g_{\theta \theta }`$ $`=`$ $`r^2\left\{1+[ϵK^{(1)}(r,t)+ϵ^2K^{(2)}(r,t)]P_2(\theta )\right\}`$ (18) $`g_{\varphi \varphi }`$ $`=`$ $`r^2\left\{\mathrm{sin}^2\theta +\mathrm{sin}^2\theta [ϵK^{(1)}(r,t)+ϵ^2K^{(2)}(r,t)]P_2(\theta )\right\},`$ (19) that is, we are demanding that $`h_0^{(i)}=h_1^{(i)}=G^{(i)}=h_{2o}^{(i)}=0`$, $`i=1,2`$. A remarkable aspect of the Regge–Wheeler gauge (RWG) is that it is unique, in the sense that the metric perturbation coefficients in the RWG can be uniquely recovered from those in an arbitrary gauge. That is, given an arbitrary metric, in an arbitrary gauge, there is a well defined procedure that brings it to the RWG. Let us exhibit this property for first order gauge transformations. One can uniquely choose a gauge transformation vector that brings the metric to the RWG form . The resulting transformation formulae are, $`K^{(1)}`$ $`=`$ $`\stackrel{~}{K}^{(1)}+(r2M)(\stackrel{~}{G}^{(1)},_r{\displaystyle \frac{2}{r^2}}\stackrel{~}{h}_1^{(1)})`$ (20) $`H_2^{(1)}`$ $`=`$ $`\stackrel{~}{H}_2^{(1)}+(2r3M)(\stackrel{~}{G}^{(1)},_r{\displaystyle \frac{2}{r^2}}\stackrel{~}{h}_1^{(1)})+r(r2M)(\stackrel{~}{G}^{(1)},_r{\displaystyle \frac{2}{r^2}}\stackrel{~}{h}_1^{(1)})_{,r}`$ (21) $`H_1^{(1)}`$ $`=`$ $`\stackrel{~}{H}_1^{(1)}+r^2\stackrel{~}{G}^{(1)},_{tr}\stackrel{~}{h}_1^{(1)},_t{\displaystyle \frac{2M}{r(r2M)}}\stackrel{~}{h}_0^{(1)}+\stackrel{~}{h}_0^{(1)},_r+{\displaystyle \frac{r(r3M)}{r2M}}\stackrel{~}{G}^{(1)},_t`$ (22) $`H_0^{(1)}`$ $`=`$ $`\stackrel{~}{H}_0^{(1)}M(\stackrel{~}{G}^{(1)},_r{\displaystyle \frac{2}{r^2}}\stackrel{~}{h}_1^{(1)}){\displaystyle \frac{2r}{r2M}}\stackrel{~}{h}_0^{(1)},_t+{\displaystyle \frac{r^3}{(r2M)}}\stackrel{~}{G}^{(1)},_{tt}`$ (23) $`h_{0o}^{(1)}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\stackrel{~}{h}_{2o,t}^{(1)}+\stackrel{~}{h}_{0o}^{(1)}`$ (24) $`h_{1o}^{(1)}`$ $`=`$ $`{\displaystyle \frac{1}{2r}}\left(r\stackrel{~}{h}_{2o,r}^{(1)}2\stackrel{~}{h}_{2o}^{(1)}\right)+\stackrel{~}{h}_{1o}^{(1)}.`$ (25) Something that should be stressed is that given the uniqueness of the RWG, quantities computed in such a gauge (like, for instance, metric components), are gauge invariant. That is, substituting the above formulae in the definition of any quantity in the RWG one obtains expressions for the quantity in any gauge. The procedure can in principle be repeated at every order in perturbation theory. To achieve this, one first performs a first order transformation that yields the metric (at first order) in the RWG. This transformation will induce transformations at all higher orders as well. One then makes a purely second order gauge transformation that brings the second order portion of the metric to the RWG. This will leave the first order piece in RWG, and will modify the third and higher orders as well. One can continue this procedure up to an arbitrarily high order and the metric will be, up to that order, in the RWG. Unfortunately, the procedure is not quite unique at higher orders. The reason for this is that if before performing the second order gauge transformation, one performs an $`\mathrm{}=0`$ first order gauge transformation, the procedure we outlined yields at the end of the day a different second order metric, but still in the RWG. That is, the metric one obtains at the end of the procedure is not unique, in spite of being in the RWG. Fortunately, one can isolate second order quantities that are invariant under these $`\mathrm{}=0`$ transformations. This point has been discussed in reference in detail, so we will not repeat it here. From the expressions for waveforms given in this paper, one can apply the techniques of reference in order to compare with numerical results without the $`\mathrm{}=0`$ ambiguity. From now on, we will do computations in the RWG. Because of what we just discussed, there is no issue of gauge invariance in our formalism, since all quantities in the RWG (appropriately modified to take into account the $`\mathrm{}=0`$ issue) will be “gauge invariant” in the sense described above. ### B The first order Zerilli equation As shown by Zerilli, we may introduce a function $`\psi ^{(1)}(r,t)`$, such that if we write for the metric components in the RWG, $`K^{(1)}(r,t)`$ $`=`$ $`6{\displaystyle \frac{r^2+rM+M^2}{r^2(2r+3M)}}\psi ^{(1)}(r,t)+\left(12{\displaystyle \frac{M}{r}}\right){\displaystyle \frac{\psi ^{(1)}(r,t)}{r}}`$ (26) $`H_0^{(1)}(r,t)`$ $`=`$ $`{\displaystyle \frac{}{r}}\left[{\displaystyle \frac{2r^26rM3M^2}{r(2r+3M)}}\psi ^{(1)}(r,t)+(r2M){\displaystyle \frac{\psi ^{(1)}(r,t)}{r}}\right]K^{(1)}(r,t)`$ (27) $`H_1^{(1)}(r,t)`$ $`=`$ $`{\displaystyle \frac{2r^26rM3M^2}{(r2M)(2r+3M)}}{\displaystyle \frac{\psi ^{(1)}(r,t)}{t}}+r{\displaystyle \frac{^2\psi ^{(1)}(r,t)}{rt}}`$ (28) $`H_2^{(1)}(r,t)`$ $`=`$ $`H_0^{(1)}(r,t),`$ (29) then the linearized Einstein equations are satisfied, provided only that $`\psi ^{(1)}(r,t)`$ is a solution of the Zerilli equation (30). $$\frac{^2\psi ^{(1)}(r,t)}{r_{}^{}{}_{}{}^{2}}\frac{^2\psi ^{(1)}(r,t)}{t^2}V_Z(r^{})\psi ^{(1)}(r,t)=0$$ (30) where $$r^{}=r+2M\mathrm{ln}[r/(2M)1]$$ (31) and $$V_Z(r)=6\left(12\frac{M}{r}\right)\frac{4r^3+4r^2M+6rM^2+3M^3}{r^3(2r+3M)^2}$$ (32) It is straightforward to obtain “inversion” formulas for $`\psi ^{(1)}(r,t)`$ in terms of the metric coefficients. From (26) and (28), we have $$\frac{\psi ^{(1)}(r,t)}{t}=\frac{r}{2r+3M}\left[r\frac{K^{(1)}}{t}\left(12\frac{M}{r}\right)H_1^{(1)}\right]$$ (33) while from (26) and (27) we find $$\psi ^{(1)}(r,t)=\frac{r(r2M)}{3(2r+3M)}\left(H_0^{(1)}r\frac{K^{(1)}}{r}\right)+\frac{r}{3}K^{(1)}$$ (34) Equations (33) and (34) are equivalent, as far as first order perturbations are concerned, up to an additive function of $`r`$. This ambiguity is of no concern since the relevant physical quantities, like radiated energies and waveforms are all computed in terms of the time derivative of $`\psi ^{(1)}`$, so it is conventionally ignored, we will assume we set the additive function to zero. Equation (34) can be used to obtain a gauge invariant form for $`\psi ^{(1)}(r,t)`$, i.e., an expression for $`\psi ^{(1)}(r,t)`$, valid in an arbitrary gauge. This is an especially important result, because it allows us to obtain the initial data for the Zerilli equation, directly in an appropriate gauge. It should be noted, however, that although $`\psi ^{(1)}(r,t)`$ is uniquely defined in a general gauge by (34), we do not have, in a general gauge, unique expressions for the metric perturbation functions in terms of $`\psi ^{(1)}(r,t)`$. The general statement of Zerilli’s results is, in this case, that if the metric perturbations satisfy the first order Einstein equations, then $`\psi ^{(1)}(r,t)`$ satisfies the Zerilli equation (30). The situation, as we shall see, is somewhat different for the second order perturbations, because of the presence of “source terms” in the Zerilli equation. ### C The Odd parity equation One can also write a “Zerilli”-like equation (known as the Regge–Wheeler equation, historically it was discovered earlier than the even parity one ) for the odd part of first order metric perturbations and the equations they satisfy. If we define $`h_{0o}^{(1)}(r,t)`$ $`=`$ $`\left(1{\displaystyle \frac{2M}{r}}\right)\left(r{\displaystyle \frac{}{r}}Q^{(1)}(r,t)+Q^{(1)}(r,t)\right)`$ (35) $`h_{1o}^{(1)}(r,t)`$ $`=`$ $`{\displaystyle \frac{r}{12M/r}}{\displaystyle \frac{}{t}}Q^{(1)}(r,t)`$ (36) the Einstein equations are satisfied, provided that $`Q^{(1)}(r,t)`$ is a solution of the equation $$\frac{^2Q^{(1)}(r,t)}{r_{}^{}{}_{}{}^{2}}\frac{^2Q^{(1)}(r,t)}{t^2}V_Q(r^{})Q^{(1)}(r,t)=0$$ (37) where $$r^{}=r+2M\mathrm{ln}[r/(2M)1]$$ (38) and $$V_Q(r)=6\frac{(rM)(r2M)}{r^4}.$$ (39) It is easier to obtain inversion formulas in this case that in the even parity case. We can write, $`Q^{(1)}(r,t)`$ $`=`$ $`{\displaystyle \frac{1}{4}}\left(r{\displaystyle \frac{h_{1o}^{(1)}(r,t)}{t}}r{\displaystyle \frac{h_{0o}^{(1)}(r,t)}{r}}+2h_{0o}^{(1)}(r,t)\right)`$ (40) $`{\displaystyle \frac{Q^{(1)}(r,t)}{t}}`$ $`=`$ $`{\displaystyle \frac{(12M/r)}{r}}h_{1o}^{(1)}(r,t)`$ (41) We can make the same comments as in the even case. Equation (40) can be used to give a gauge invariant form for the odd parity Zerilli function. Both equations allow us to give initial conditions for $`Q^{(1)}(r,t)`$ in an any gauge. As in the even case, there is no unique way to write the metric perturbations in terms of $`Q^{(1)}(r,t)`$ in an arbitrary gauge. ## III Second order perturbations We define the second order Regge–Wheeler gauge by imposing that $`h_0^{(i)}=h_1^{(i)}=G^{(i)}=h_{2o}^{(i)}=0`$, $`i=1,2`$. It is easy to check that this gauge is uniquely defined, as long as we restrict to $`\mathrm{}=2`$ axisymmetric first and second order gauge transformations, but the general expressions for the second order gauge transformations are rather lengthy. It should be clear from the perturbation expansion, that the second order Einstein equations imply, in turn, that the second order perturbation functions satisfy the same linear set of equations as the first order perturbations, but with additional terms quadratic in the first order perturbations, that may be thought of as representing “sources” for these equations. Therefore if motivated by the corresponding expression for the time derivatives $`\psi ^{(1)}/t`$, $`Q^{(1)}/t`$, of the Zerilli functions in first order perturbation theory, we introduce gauge invariant functions, given by $`\chi ^{(2)}(r,t)`$ $`=`$ $`{\displaystyle \frac{r}{2r+3M}}\left[r{\displaystyle \frac{K^{(2)}(r,t)}{t}}\left(1{\displaystyle \frac{2M}{r}}\right)H_1^{(2)}(r,t)\right]`$ (42) $`\mathrm{\Theta }^{(2)}(r,t)`$ $`=`$ $`{\displaystyle \frac{(12M/r)}{r}}h_{1o}^{(2)}(r,t)`$ (43) it follows that if the second order Einstein equations are satisfied, then $`\chi ^{(2)}(r,t)`$, $`\mathrm{\Theta }^{(2)}`$ satisfy equations of the form $`{\displaystyle \frac{^2\chi ^{(2)}(r,t)}{r_{}^{}{}_{}{}^{2}}}{\displaystyle \frac{^2\chi ^{(2)}(r,t)}{t^2}}V_Z(r^{})\chi ^{(2)}(r,t)+𝒮_Z=0`$ (44) $`{\displaystyle \frac{^2\mathrm{\Theta }^{(2)}(r,t)}{r_{}^{}{}_{}{}^{2}}}{\displaystyle \frac{^2\mathrm{\Theta }^{(2)}(r,t)}{t^2}}V_Q(r^{})\mathrm{\Theta }^{(2)}(r,t)+𝒮_Q=0`$ (45) where $`𝒮_Z`$, $`𝒮_Q`$ are quadratic polynomials in the first order functions and their $`t`$ and $`r`$ derivatives, with $`r`$ dependent coefficients. $`𝒮_Z`$,$`𝒮_Q`$ may be considered as a kind of ”source terms” for the otherwise homogeneous equations satisfied by $`\chi ^{(2)}`$, $`\mathrm{\Theta }^{(2)}`$. But, precisely because of the presence of these “source terms”, if we redefine $`\chi ^{(2)}(r,t)`$, $`\mathrm{\Theta }^{(2)}(r,t)`$ by the addition of a quadratic polynomial in the first order functions and their derivatives, the new, redefined $`\chi ^{(2)}(r,t)`$, $`\mathrm{\Theta }^{(2)}(r,t)`$ will still satisfy an equation of the form (44), provided only that the first and second order Einstein equations are satisfied. This arbitrariness in the definition of the Zerilli functions associated to the second order perturbations may be used to simplify the form, and asymptotic behavior of the “source terms” $`𝒮_Z`$, $`𝒮_Q`$. As was discussed in , the source terms are delicate in the sense that they can be divergent for large values of $`t`$ and $`r`$. The divergence of the source does not imply a physical divergence in the problem, but it poses serious difficulties for numeric computations. If we make the following choice, based on the study of the solutions of the Zerilli equations for large values of $`r`$, the source simplifies and also is well behaved at spatial infinity, $$\chi ^{(2)}(r,t)=\frac{r}{2r+3M}\left[r\frac{K^{(2)}}{t}\left(1\frac{2M}{r}\right)H_1^{(2)}\right]\frac{2}{7}\left[\frac{r^2}{2r+3M}K^{(1)}\frac{K^{(1)}}{t}+(K^{(1)})^2\right]$$ (46) as our definition for $`\chi ^{(2)}(r,t)`$, while $`\mathrm{\Theta }^{(2)}(r,t)`$ is defined by (43). Making the appropriate replacements, we then find for the sources, $`𝒮_Z`$ $`=`$ $`{\displaystyle \frac{12}{7}}{\displaystyle \frac{(r2M)^3}{r^3(2r+3M)}}[{\displaystyle \frac{\left(24r^5108Mr^472M^2r^3+24M^3r^2+180M^4r+180M^5\right)\left(\psi \right)^2}{r^5\left(2r+3M\right)\left(r2M\right)^2}}`$ (47) $`+`$ $`{\displaystyle \frac{\left(64r^5+176Mr^4+592M^2r^3+1020M^3r^2+1122M^4r+540M^5\right)\psi _r\psi }{r^4\left(2r+3M\right)^2\left(r2M\right)}}`$ (48) $``$ $`{\displaystyle \frac{\left(112r^5+480Mr^4+692M^2r^3+762M^3r^2+441M^4r+144M^5\right)\psi _t\psi }{r^2\left(2r+3M\right)^3\left(r2M\right)^2}}`$ (49) $`+`$ $`{\displaystyle \frac{\left(8r^3+16r^2M+36rM^2+24M^3\right)\psi _{rr}\psi }{r^3\left(2r+3M\right)}}+{\displaystyle \frac{\left(8r^2+12rM+7M^2\right)\psi _{rt}\psi }{r\left(2r+3M\right)\left(r+2M\right)}}+{\displaystyle \frac{M\psi \psi _{rrt}}{2r+3M}}+{\displaystyle \frac{r\psi _t\psi _{rrr}}{3}}`$ (50) $``$ $`{\displaystyle \frac{\left(12r^3+36r^2M+59rM^2+90M^3\right)\left(\psi _r\right)^2}{3r^3\left(r2M\right)}}+{\displaystyle \frac{\left(18r^3+4r^2M+33rM^2+48M^3\right)\psi _r\psi _t}{3\left(2r+3M\right)r\left(r2M\right)^2}}`$ (51) $``$ $`{\displaystyle \frac{\left(12r^2+20rM+24M^2\right)\psi _r\psi _{rr}}{3r^2}}{\displaystyle \frac{\left(3r+7M\right)\psi _r\psi _{rt}}{3r+6M}}{\displaystyle \frac{r\psi _r\psi _{rrt}}{3}}`$ (52) $`+`$ $`{\displaystyle \frac{12\left(r^2+rM+M^2\right)^2\left(\psi _t\right)^2}{\left(2r+3M\right)r\left(r2M\right)^3}}{\displaystyle \frac{\left(2r^2+M^2\right)\psi _{rr}\psi _t}{\left(2r+3M\right)\left(r2M\right)}}+{\displaystyle \frac{\left(4r^2+4rM+4M^2\right)\psi _t\psi _{rt}}{\left(r2M\right)^2}}`$ (53) $``$ $`{\displaystyle \frac{\left(2r+3M\right)\left(r2M\right)\left(\psi _{rr}\right)^2}{3r}}{\displaystyle \frac{\left(2r+3M\right)r\left(\psi _{rt}\right)^2}{3r+6M}}{\displaystyle \frac{\left(2r+10M\right)\left(3r+8M\right)Q_rQ_t}{r\left(r2M\right)^2}}`$ (54) $`+`$ $`{\displaystyle \frac{\left(22r^2+70rM+66M^2\right)Q_rQ_{rt}}{\left(2r+3M\right)\left(r2M\right)}}{\displaystyle \frac{\left(136r^3462r^2M+50rM^2+456M^3\right)QQ_t}{r^2\left(2r+3M\right)\left(r2M\right)^2}}`$ (55) $`+`$ $`{\displaystyle \frac{\left(14r^2+58rM+66M^2\right)QQ_{rt}}{r\left(2r+3M\right)\left(r2M\right)}}{\displaystyle \frac{\left(22r+54M\right)\left(r+M\right)Q_tQ_{rr}}{\left(2r+3M\right)\left(r2M\right)}}]`$ (56) $`𝒮_Q`$ $`=`$ $`{\displaystyle \frac{12}{7}}{\displaystyle \frac{(r2M)^3}{r^3(2r+3M)}}[{\displaystyle \frac{\left(40r^4150Mr^3138r^2M^2+162M^3r+81M^4\right)Q\psi _t}{3r^3\left(2r+3M\right)\left(r2M\right)^2}}`$ (57) $`+`$ $`{\displaystyle \frac{\left(2r+3M\right)\left(r+M\right)Q\psi _{rrt}}{r\left(r2M\right)}}+{\displaystyle \frac{\left(4r^317r^2M33rM^2+81M^3\right)Q\psi _{rt}}{3r^2\left(r2M\right)^2}}`$ (58) $`+`$ $`{\displaystyle \frac{\left(160M^2r^3+216M^5+102M^3r^2160Mr^4+387M^4r128r^5\right)Q_t\psi }{2r^3\left(2r+3M\right)^2\left(r2M\right)^2}}`$ (59) $`+`$ $`{\displaystyle \frac{\left(75M^3r^212r^5+81M^510M^2r^3+14Mr^4+150M^4r\right)\psi Q_{rt}}{r^2\left(2r+3M\right)^2\left(r2M\right)^2}}{\displaystyle \frac{\left(2r+3M\right)r\psi _{rr}Q_{rrt}}{3}}`$ (60) $`+`$ $`{\displaystyle \frac{\left(4r^3+4r^2M+6rM^2+3M^3\right)\psi Q_{rrt}}{r\left(2r+3M\right)\left(r2M\right)}}{\displaystyle \frac{\left(8r^4+56Mr^340r^2M^239M^3r+72M^4\right)Q_t\psi _r}{2r^2\left(2r+3M\right)\left(r+2M\right)^2}}`$ (61) $``$ $`{\displaystyle \frac{\left(18r^4+19Mr^313r^2M^2+78M^3r+81M^4\right)\psi _rQ_{rt}}{3r\left(2r+3M\right)\left(r2M\right)^2}}+{\displaystyle \frac{\left(2r^22rM+3M^2\right)\psi _rQ_{rrt}}{3r+6M}}`$ (62) $``$ $`{\displaystyle \frac{\left(28r^472Mr^396r^2M^2+114M^3r+81M^4\right)\psi _tQ_r}{3r^2\left(2r+3M\right)\left(r2M\right)^2}}+{\displaystyle \frac{\left(2r^2+6rM+3M^2\right)\psi _tQ_{rrr}}{3r+6M}}`$ (63) $``$ $`{\displaystyle \frac{\left(2r^49Mr^3+9r^2M^2+22M^3r+9M^4\right)\psi _tQ_{rr}}{r\left(2r+3M\right)\left(r2M\right)^2}}+{\displaystyle \frac{\left(4r^311r^2M15rM^2+63M^3\right)Q_t\psi _{rr}}{3r\left(r2M\right)^2}}`$ (64) $`+`$ $`{\displaystyle \frac{\left(6r^2+7rM+18M^2\right)\psi _{rr}Q_{rt}}{6r+12M}}{\displaystyle \frac{\left(10r^310r^2M37rM^2+27M^3\right)\psi _{rt}Q_r}{3r\left(r2M\right)^2}}`$ (65) $``$ $`{\displaystyle \frac{\left(6r^2+7rM+18M^2\right)\psi _{rt}Q_{rr}}{6r+12M}}+{\displaystyle \frac{\left(2r+3M\right)rQ_{rrr}\psi _{rt}}{3}}{\displaystyle \frac{\left(2r+3M\right)\left(r+3M\right)Q_t\psi _{rrr}}{3r+6M}}`$ (66) $``$ $`{\displaystyle \frac{\left(2r+3M\right)r\psi _{rrr}Q_{rt}}{3}}+{\displaystyle \frac{\left(2r+3M\right)\left(r+M\right)\psi _{rrt}Q_r}{3r+6M}}+{\displaystyle \frac{\left(2r+3M\right)rQ_{rr}\psi _{rrt}}{3}}]`$ (67) where $`\psi =\psi ^{(1)}`$, $`Q=Q^{(1)}`$ and the subindices indicate partial derivatives. ## IV Gravitational wave amplitude A straightforward computational way to define a gravitational wave amplitude, is to consider a gauge in which the space-time is manifestly asymptotically flat. For instance, consider the typical gauge discussed in reference . In this gauge the metric has components that behave in the following way, $`H_2`$ $``$ $`O(1/r^3)`$ (68) $`h_1`$ $``$ $`O(1/r)`$ (69) $`K`$ $``$ $`O(1/r)`$ (70) $`G`$ $``$ $`O(1/r)`$ (71) $`h_{1o}`$ $``$ $`O(1/r)`$ (72) $`h_{2o}`$ $``$ $`O(r)`$ (73) and, of course, $`H_0=H_1=h_0=h_{0o}=0`$. The gravitational wave amplitude can then simply be read off from the components $`K`$, $`G`$ and $`h_{2o}`$. We therefore need to relate the behavior of these components of the metric to the Zerilli function in an asymptotic region far away from the source. In order to study the asymptotic behavior, one considers a solution of the Zerilli equation in terms of an expansion in inverse powers of $`r`$ with coefficients that are functions of the retarded time $`tr^{}`$, as discussed in detail in . Substituting in the Zerilli equations we obtain, $`\psi ^{(1)}`$ $`=`$ $`F_\psi ^{(iii)}(tr^{})+{\displaystyle \frac{3}{r}}F_\psi ^{(ii)}(tr^{}){\displaystyle \frac{3}{4r^2}}\left[7F_\psi ^{(ii)}(tr^{})M+4F_\psi ^{(i)}(tr^{})\right]+O(1/r^3)`$ (74) $`Q^{(1)}`$ $`=`$ $`F_Q^{(iii)}(tr^{})+{\displaystyle \frac{3}{r}}F_Q^{(ii)}(tr^{})`$ (76) $`+{\displaystyle \frac{1}{r^2}}\left[{\displaystyle \frac{3}{2}}MF_Q^{(ii)}(tr^{})+3F_Q^{(i)}(tr^{})\right]+O(1/r^3)`$ For this behavior of first order Zerilli functions, the sources $`S_Z`$, $`S_Q`$ for second order Zerilli functions $`\chi ^{(2)}`$, $`\mathrm{\Theta }^{(2)}`$ decay as $`O(1/r^2)`$ asymptotically. Therefore, $`\chi ^{(2)}`$,$`\mathrm{\Theta }^{(2)}`$ will behave in a similar way to $`\psi ^{(1)}`$, $`Q^{(1)}`$ for large $`r`$: $`\chi ^{(2)}`$ $`=`$ $`F_\chi ^{(20)}(tr^{})+{\displaystyle \frac{1}{r}}F_\chi ^{(21)}(tr^{})+O(1/r^2)`$ (77) $`\mathrm{\Theta }^{(2)}`$ $`=`$ $`F_Q^{(20)}(tr^{})+{\displaystyle \frac{1}{r}}F_Q^{(21)}(tr^{})+O(1/r^2)`$ (78) We can now compute the asymptotic form for the metric perturbations, to first order and second order, in Regge–Wheeler gauge, using (76) and (78) in (27) and (29). Unfortunately, in the Regge–Wheeler gauge the metric is not in the manifestly asymptotically flat form (73). We therefore need to perform a gauge transformation to bring it to such a form, to first and second order. In order to do this, we start by recalling that for any $`\mathrm{}=2`$ axisymmetric gauge transformation, we can write the gauge transformation vectors as, $$\begin{array}{cccccc}\hfill \xi _{(even)}^{(1)t}& =& _0^{(1)}(r,t)P_2(\mathrm{cos}(\theta ))\hfill & \hfill \xi _{(even)}^{(1)r}& =& _1^{(1)}(r,t)P_2(\mathrm{cos}(\theta ))\hfill \\ \hfill \xi _{(even)}^{(1)\theta }& =& ^{(1)}(r,t)P_2(\mathrm{cos}(\theta ))/\theta \hfill & \hfill \xi _{(even)}^{(1)\varphi }& =& 0\hfill \end{array}$$ (79) for even parity, and $$\begin{array}{cccccc}\hfill \xi _{(odd)}^{(1)t}& =& 0\hfill & \hfill \xi _{(odd)}^{(1)r}& =& 0\hfill \\ \hfill \xi _{(odd)}^{(1)\theta }& =& 0\hfill & \hfill \xi _{(odd)}^{(1)\varphi }& =& _2^{(1)}(r,t)\mathrm{sin}(\theta )^1P_2(\mathrm{cos}(\theta ))/\theta \hfill \end{array}$$ (80) for the odd parity case. To achieve the asymptotically flat gauge (73) we demand that $`H_0^{(1)}(r,t)=H_1^{(1)}(r,t)=h_0^{(1)}(r,t)=h_{0o}^{(1)}=0`$. From the general form of the gauge transformation equations we have, $`0`$ $`=`$ $`\stackrel{~}{H}_0^{(1)}+{\displaystyle \frac{2M}{r(r2M)}}_1^{(1)}+2{\displaystyle \frac{_0^{(1)}}{t}}`$ (81) $`0`$ $`=`$ $`\stackrel{~}{H}_1^{(1)}{\displaystyle \frac{r}{r2M}}{\displaystyle \frac{_1^{(1)}}{t}}+{\displaystyle \frac{r2M}{r}}{\displaystyle \frac{_0^{(1)}}{r}}`$ (82) $`0`$ $`=`$ $`r^2{\displaystyle \frac{^{(1)}}{t}}+{\displaystyle \frac{r2M}{r}}_0^{(1)}`$ (83) $`H_2^{(1)}`$ $`=`$ $`\stackrel{~}{H}_2^{(1)}+{\displaystyle \frac{2M}{r(r2M)}}_1^{(1)}2{\displaystyle \frac{_1^{(1)}}{r}}`$ (84) $`h_1^{(1)}`$ $`=`$ $`{\displaystyle \frac{r}{r2M}}_1^{(1)}r^2{\displaystyle \frac{^{(1)}}{r}}`$ (85) $`G^{(1)}`$ $`=`$ $`2^{(1)}`$ (86) $`K^{(1)}`$ $`=`$ $`\stackrel{~}{K}^{(1)}{\displaystyle \frac{2}{r}}_1^{(1)}`$ (87) $`0`$ $`=`$ $`\stackrel{~}{h}_{0o}^{(1)}+r^2{\displaystyle \frac{_2^{(1)}(r,t)}{t}}`$ (88) $`h_{1o}^{(1)}`$ $`=`$ $`\stackrel{~}{h}_{1o}^{(1)}+r^2{\displaystyle \frac{_2^{(1)}(r,t)}{r}}`$ (89) $`h_{2o}^{(1)}`$ $`=`$ $`2r^2_2^{(1)}(r,t)`$ (90) where the quantities with tildes are given in the Regge–Wheeler gauge, $`\{^{(1)},_0^{(1)},_1^{(1)}\}`$, are the components of the even parity gauge transformation vector, and $`_2^{(1)}`$ that of the odd parity gauge transformation vector. We can see that, given $`\stackrel{~}{H}_0^{(1)}`$ and $`\stackrel{~}{H}_1^{(1)}`$, the first and second equation above may be used to solve for $`_0^{(1)}`$ and $`_1^{(1)}`$. Then, the third equation leads to a partial differential equation for $`^{(1)}`$ while the equation for $`h_{0o}^{(1)}`$ fixes $`_2^{(1)}`$. Therefore, the set of equations, and therefore the gauge choice, is consistent. The solution, assuming only that $`\stackrel{~}{H}_0^{(1)}`$, $`\stackrel{~}{H}_1^{(1)}`$, $`\stackrel{~}{H}_2^{(1)}`$, and $`\stackrel{~}{K}^{(1)}`$ are given, is non unique since one has to integrate partial differential equations, so the solution in general involves free functions. One may choose such functions in a judicious way to ensure that the non-vanishing metric components behave in a manner consistent with asymptotic flatness. Similarly, we define the second order asymptotically flat gauge by imposing the first order asymptotically flat gauge conditions and $`H_0^{(2)}(r,t)=0`$, $`H_1^{(2)}(r,t)=0`$, $`h_0^{(2)}(r,t)=0`$ and $`h_{0o}^{(2)}(r,t)=0`$. Clearly, the same reasoning as above, applied to the second order gauge transformations, shows that this is a consistent choice. The end result of these calculations is that one may write the first order gauge transformation vectors in the form, $`^{(1)}(r,t)`$ $`=`$ $`{\displaystyle \frac{1}{2r}}F_\psi ^{(iii)}(tr^{})+{\displaystyle \frac{1}{r^2}}F_\psi ^{(ii)}(tr^{})+{\displaystyle \frac{1}{r^3}}\left[{\displaystyle \frac{1}{4}}MF_\psi ^{(ii)}(tr^{})+{\displaystyle \frac{3}{2}}F_\psi ^{(i)}(tr^{})\right]+O(1/r^4)`$ (91) $`_0^{(1)}(r,t)`$ $`=`$ $`{\displaystyle \frac{1}{2}}rF_\psi ^{(iv)}(tr^{})+\left[F_\psi ^{(iii)}(tr^{})+MF_\psi ^{(iv)}(tr^{})\right]`$ (93) $`+{\displaystyle \frac{1}{r}}\left[{\displaystyle \frac{9}{4}}MF_\psi ^{(iii)}(tr^{})+{\displaystyle \frac{3}{2}}F_\psi ^{(ii)}(tr^{})+2M^2F_\psi ^{(iv)}(tr^{})\right]+O(1/r^2)`$ $`_1^{(1)}(r,t)`$ $`=`$ $`{\displaystyle \frac{1}{2}}rF_\psi ^{(iv)}(tr^{})+{\displaystyle \frac{3}{2}}F_\psi ^{(iii)}(tr^{})`$ (95) $`+{\displaystyle \frac{1}{r}}\left[{\displaystyle \frac{3}{2}}F_\psi ^{(ii)}(tr^{}){\displaystyle \frac{3}{4}}MF_\psi ^{(iii)}(tr^{})\right]+O(1/r^2)`$ $`_2^{(1)}(r,t)`$ $`=`$ $`{\displaystyle \frac{3}{r}}F_Q^{(iii)}(tr^{})+{\displaystyle \frac{2}{r^2}}F_Q^{(ii)}(tr^{})`$ (97) $`+{\displaystyle \frac{1}{r^3}}\left[3F_Q^{(i)}(tr^{})+{\displaystyle \frac{1}{2}}MF_Q^{(ii)}(tr^{})\right]+O(1/r^4)`$ The second order gauge equations are more complex then the first order ones. It can be seen that the equations can also be solved iteratively for all orders in $`r`$, but for simplicity we only compute the terms that are going to be relevant in the gravitational waveforms. The meaningful terms are given by $`^{(2)}`$ $`=`$ $`{\displaystyle \frac{1}{r}}^{(20)}(tr^{})+O(1/r^2)`$ (98) $`_0^{(2)}`$ $`=`$ $`r_0^{(20)}(tr^{})+O(r^0)`$ (99) $`_1^{(2)}`$ $`=`$ $`r_1^{(20)}(tr^{})+_1^{(21)}(tr^{})+O(1/r)`$ (100) $`_2^{(2)}`$ $`=`$ $`{\displaystyle \frac{1}{r}}_2^{(20)}(tr^{})+O(1/r^3)`$ (101) and if we compute them in terms of $`F_\psi `$,$`F_Q`$, we finally get $`{\displaystyle \frac{^{(20)}(tr^{})}{t}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}F_\chi ^{(20)}(tr^{}){\displaystyle \frac{1}{28}}F_\psi ^{(v)}(tr^{})F_\psi ^{(iii)}(tr^{}){\displaystyle \frac{1}{28}}F_\psi ^{(iv)}(tr^{})^2{\displaystyle \frac{3}{14}}F_Q^{(iv)}(tr^{})^2`$ (102) $`_0^{(20)}(tr^{})`$ $`=`$ $`{\displaystyle \frac{1}{2}}F_\chi ^{(20)}(tr^{}){\displaystyle \frac{1}{14}}F_\psi ^{(v)}(tr^{})F_\psi ^{(iii)}(tr^{}){\displaystyle \frac{3}{14}}F_Q^{(iv)}(tr^{})^2`$ (103) $`_1^{(20)}(tr^{})`$ $`=`$ $`{\displaystyle \frac{1}{2}}F_\chi ^{(20)}(tr^{}){\displaystyle \frac{1}{14}}F_\psi ^{(v)}(tr^{})F_\psi ^{(iii)}(tr^{}){\displaystyle \frac{3}{14}}F_Q^{(iv)}(tr^{})^2`$ (104) $`{\displaystyle \frac{^2_1^{(21)}(tr^{})}{t^2}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}F_\chi ^{(21)(ii)}(tr^{})+{\displaystyle \frac{3}{14}}MF_\psi ^{(v)}(tr^{})^2{\displaystyle \frac{12}{7}}F_\psi ^{(iv)}(tr^{})F_\psi ^{(v)}(tr^{})`$ (109) $`{\displaystyle \frac{3}{14}}F_\psi ^{(ii)}(tr^{})F_\psi ^{(vii)}(tr^{}){\displaystyle \frac{9}{14}}F_\psi ^{(vi)}(tr^{})F_\psi ^{(iii)}(tr^{})`$ $`{\displaystyle \frac{69}{14}}F_Q^{(v)}(tr^{})F_Q^{(iv)}(tr^{})+{\displaystyle \frac{9}{14}}F_Q^{(vi)}(tr^{})F_Q^{(iv)}(tr^{})`$ $`+{\displaystyle \frac{3}{28}}MF_\psi ^{(v)}(tr^{})F_\psi ^{(v)}(tr^{}){\displaystyle \frac{3}{28}}MF_\psi ^{(iii)}(tr^{})F_\psi ^{(vii)}(tr^{})`$ $`{\displaystyle \frac{15}{14}}F_Q^{(vi)}(tr^{})F_Q^{(iii)}(tr^{})+{\displaystyle \frac{9}{14}}MF_Q^{(v)}(tr^{})^2`$ $`{\displaystyle \frac{_2^{(20)}(tr^{})}{t}}`$ $`=`$ $`F_Q^{(20)}(tr^{}){\displaystyle \frac{1}{14}}F_Q^{(v)}(tr^{})F_\psi ^{(iii)}(tr^{}){\displaystyle \frac{1}{14}}F_Q^{(iii)}(tr^{})F_\psi ^{(v)}(tr^{})`$ (110) We can now compute the gravitational waveforms by reading off the appropriate metric components. For first order we find, $`{\displaystyle \frac{G^{(1)}(r,t)}{t}}`$ $`=`$ $`{\displaystyle \frac{1}{r}}F_\psi ^{(iv)}(tr^{})+O(1/r^2)`$ (111) $`{\displaystyle \frac{K^{(1)}(r,t)}{t}}`$ $`=`$ $`{\displaystyle \frac{3}{r}}F_\psi ^{(iv)}(tr^{})+O(1/r^2)`$ (112) $`{\displaystyle \frac{h_{2o}^{(1)}(r,t)}{t}}`$ $`=`$ $`2rF_Q^{(iv)}(tr^{})+O(r^0)`$ (113) and the second order components are $`{\displaystyle \frac{G^{(2)}(r,t)}{t}}`$ $`=`$ $`\left\{F_\chi ^{20}(tr^{})+{\displaystyle \frac{1}{7}}{\displaystyle \frac{}{t}}\left[F_\psi ^{(iv)}(tr^{})F_\psi ^{(iii)}(tr^{})\right]{\displaystyle \frac{3}{7}}F_Q^{(iv)}(tr^{})^2\right\}{\displaystyle \frac{1}{r}}+O(1/r^2)`$ (114) $`{\displaystyle \frac{K^{(2)}(r,t)}{t}}`$ $`=`$ $`3{\displaystyle \frac{G^{(2)}(r,t)}{t}}+O(1/r^2)`$ (115) $`{\displaystyle \frac{h_{2o}^{(2)}(r,t)}{t}}`$ $`=`$ $`2r\{F_Q^{20}(tr^{})+{\displaystyle \frac{1}{7}}{\displaystyle \frac{^2}{t^2}}\left[F_Q^{(iii)}(tr^{})F_\psi ^{(iii)}(tr^{})\right]`$ (117) $`+{\displaystyle \frac{1}{7}}F_Q^{(iv)}(tr^{})F_\psi ^{(iv)}(tr^{})\}+O(r^0)`$ and in terms of first and second order Zerilli functions, we get $`{\displaystyle \frac{G^{(1)}(r,t)}{t}}`$ $`=`$ $`{\displaystyle \frac{1}{r}}{\displaystyle \frac{\psi ^{(1)}}{t}}+O(1/r^2)`$ (118) $`{\displaystyle \frac{K^{(1)}(r,t)}{t}}`$ $`=`$ $`3{\displaystyle \frac{G^{(1)}(r,t)}{t}}`$ (119) $`{\displaystyle \frac{h_{2o}^{(1)}(r,t)}{t}}`$ $`=`$ $`2r{\displaystyle \frac{Q^{(1)}(r,t)}{t}}+O(r^0)`$ (120) and $`{\displaystyle \frac{G^{(2)}(r,t)}{t}}`$ $`=`$ $`\left\{\chi ^{(2)}+{\displaystyle \frac{1}{7}}{\displaystyle \frac{}{t}}\left[{\displaystyle \frac{\psi ^{(1)}}{t}}\psi ^{(1)}\right]{\displaystyle \frac{3}{7}}\left[{\displaystyle \frac{Q^{(1)}}{t}}\right]^2\right\}{\displaystyle \frac{1}{r}}+O(1/r^2)`$ (121) $`{\displaystyle \frac{K^{(2)}(r,t)}{t}}`$ $`=`$ $`3{\displaystyle \frac{G^{(2)}(r,t)}{t}}+O(1/r^2)`$ (122) $`{\displaystyle \frac{h_{2o}^{(2)}(r,t)}{t}}`$ $`=`$ $`2r\left\{\mathrm{\Theta }^{(2)}+{\displaystyle \frac{1}{7}}{\displaystyle \frac{^2}{t^2}}\left[\psi ^{(1)}Q^{(1)}\right]+{\displaystyle \frac{1}{7}}{\displaystyle \frac{}{t}}\psi ^{(1)}{\displaystyle \frac{}{t}}Q^{(1)}\right\}+O(r^0)`$ (123) It is straightforward to compute the radiated energy, since we are in an explicitly asymptotically flat gauge. One can therefore simply apply the formulae stemming from Landau and Lifshitz , $$\frac{d\mathrm{Power}}{d\mathrm{\Omega }}=\frac{1}{16\pi r^2}\left[\left(\frac{}{t}h_{\theta \varphi }\right)^2+\frac{1}{4}\left(\frac{}{t}h_{\theta \theta }\frac{1}{\mathrm{sin}^2\theta }\frac{}{t}h_{\varphi \varphi }\right)^2\right].$$ (124) For the perturbations we are considering, $`h_{\theta \theta }`$ $`=`$ $`r^2\{[ϵ\stackrel{~}{K}^{(1)}(r,t)+ϵ^2\stackrel{~}{K}^{(2)}(r,t)]P_2(\theta )`$ (126) $`+[ϵ\stackrel{~}{G}^{(1)}(r,t)+ϵ^2\stackrel{~}{G}^{(2)}(r,t)]^2P_2(\theta )/\theta ^2)\}`$ $`h_{\varphi \varphi }`$ $`=`$ $`r^2\{\mathrm{sin}^2\theta [ϵ\stackrel{~}{K}^{(1)}(r,t)+ϵ^2\stackrel{~}{K}^{(2)}(r,t)]P_2(\theta )`$ (128) $`+\mathrm{sin}(\theta )\mathrm{cos}(\theta )[ϵ\stackrel{~}{G}^{(1)}(r,t)+ϵ^2\stackrel{~}{G}^{(2)}(r,t)]P_2(\theta )/\theta )\}`$ $`h_{\theta \varphi }`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left[ϵ\stackrel{~}{h}_{2o}^{(1)}(r,t)+ϵ^2\stackrel{~}{h}_{2o}^{(2)}(r,t)\right]\left[\mathrm{cos}(\theta )P_2(\theta )/\theta \mathrm{sin}(\theta )^2P_2(\theta )/\theta ^2\right]`$ (129) and substituting, we get, $`\mathrm{Power}`$ $`=`$ $`{\displaystyle \frac{3}{10}}\left[ϵ{\displaystyle \frac{\psi ^{(1)}}{t}}+ϵ^2\left(\chi ^{(2)}+{\displaystyle \frac{1}{7}}{\displaystyle \frac{}{t}}\left(\psi ^{(1)}{\displaystyle \frac{\psi ^{(1)}}{t}}\right){\displaystyle \frac{3}{7}}\left({\displaystyle \frac{Q^{(1)}}{t}}\right)^2\right)\right]^2`$ (130) $`+`$ $`{\displaystyle \frac{36}{35}}\left[ϵ{\displaystyle \frac{Q^{(1)}}{t}}+ϵ^2\left(\mathrm{\Theta }^{(2)}+{\displaystyle \frac{1}{7}}\left({\displaystyle \frac{^2Q^{(1)}\psi ^{(1)}}{t^2}}+{\displaystyle \frac{Q^{(1)}}{t}}{\displaystyle \frac{\psi ^{(1)}}{t}}\right)\right)\right]^2.`$ (131) As can be seen, the energy has two quadratic contributions, one coming from the $`h_{\theta \varphi }`$ component of the metric and one from the diagonal elements of it. They correspond to the “$`\times `$” and “$`+`$” polarization modes of the gravitational field and therefore we can consider the expressions in the square brackets as the“waveforms” of the gravitational waves emitted. Contrary to the first order case, in which the waveform is directly given by the Zerilli (or Regge–Wheeler) function, for second order corrections the relationship is more involved. ## V Initial data Initial data for solutions of Einstein equations is usually given in terms of the initial value of the three-metric and the initial value of the extrinsic curvature of the initial Cauchy surface. For the case of initial data for colliding black holes, the popular families of initial data of Bowen and York and punctures naturally come cast in a gauge in which<sup>*</sup><sup>*</sup>* This happens to coincide exactly with the same gauge condition of the asymptotically flat gauge. $`H_0=0`$, $`H_1=0`$, $`h_0=0`$ and $`h_{0o}=0`$. We will therefore present the appropriate formulae that would allow us, given a metric and extrinsic curvature cast in such a gauge, to provide the initial data for the first and second order Zerilli functions and their time derivatives. The extrinsic curvature $`K_{ab}`$ of a spatial slice is defined by $$K_{ab}=n_{(c;d)}\left(\delta _{}^{c}{}_{a}{}^{}+n^cn_a\right)\left(\delta _{}^{d}{}_{b}{}^{}+n^dn_b\right)$$ (132) where $`n_a`$ is the unit normal to the Cauchy surface where the data are specified. One can relate the extrinsic curvature to the time derivative of the three-metric through, $`ds^2`$ $`=`$ $`g_{ij}dx^idx^jN^2(dt)^2`$ (133) $`{\displaystyle \frac{g_{ij}}{t}}`$ $`=`$ $`2NK_{ij}.`$ (134) That is, if through some procedure we are given the extrinsic curvature and the three metric, we can compute the time derivative of the three metric. Re-expressing this in terms of components of the tensor spherical harmonic decomposition due to Regge and Wheeler we get, $$\begin{array}{cccccc}\hfill h_1^{(1)}/t& =& 2\sqrt{12M/r}K_{(h_1)}^{(1)}\hfill & \hfill H_2^{(1)}/t& =& 2\sqrt{12M/r}K_{(H_2)}^{(1)}\hfill \\ \hfill K^{(1)}/t& =& 2\sqrt{12M/r}K_{(K)}^{(1)}\hfill & \hfill G^{(1)}/t& =& 2\sqrt{12M/r}K_{(G)}^{(1)}\hfill \\ \hfill h_{1o}^{(1)}/t& =& 2\sqrt{12M/r}K_{(h_{1o})}^{(1)}\hfill & \hfill h_{2o}^{(1)}/t& =& 2\sqrt{12M/r}K_{(h_{2o})}^{(1)}\hfill \end{array}$$ (135) for first order components. $`K_{(xx)}^{(1)}`$ are spherical tensor harmonic components of the first order extrinsic curvature tensor, written in the Regge–Wheeler notation that we used in formulae (1-12). For second order we get, $$\begin{array}{cccccc}\hfill h_1^{(2)}/t& =& 2\sqrt{12M/r}K_{(h_1)}^{(2)}\hfill & \hfill H_2^{(2)}/t& =& 2\sqrt{12M/r}K_{(H_2)}^{(2)}\hfill \\ \hfill K^{(2)}/t& =& 2\sqrt{12M/r}K_{(K)}^{(2)}\hfill & \hfill G^{(2)}/t& =& 2\sqrt{12M/r}K_{(G)}^{(2)}\hfill \\ \hfill h_{1o}^{(2)}/t& =& 2\sqrt{12M/r}K_{(h_{1o})}^{(2)}\hfill & \hfill h_{2o}^{(2)}/t& =& 2\sqrt{12M/r}K_{(h_{2o})}^{(2)}\hfill \end{array}$$ (136) Here $`K_{(xx)}^{(2)}`$ are spherical harmonic components of the second order extrinsic curvature tensor. Note that the expressions are similar to the first order ones only in the gauge we are considering. In other gauges (in which the perturbations of lapse and shift may not vanish), there will be extra terms, quadratic in first order elements. A case in which the formalism just developed will be of use (and we will discuss in the forthcoming publication) is that of the collision of counterrotating Bowen–York holes (the cosmic screw case). In such a case, to first order the metric has even portions only (similar to those in the Misner initial data) and the extrinsic curvature odd portions. So putting together the above formulae, for this case one gets, $`\chi ^{(2)}`$ $`=`$ $`{\displaystyle \frac{2}{7}}{\displaystyle \frac{\sqrt{r2M}}{(2r+3M)\sqrt{r}}}[7r^2K_{(K)}^{(2)}7r^2K_{(G)}^{(2)}7rK_{(h_1)}^{(2)}+2\left(K^{(1)}\right)^2r\sqrt{{\displaystyle \frac{r}{r+2M}}}+`$ (137) $`+`$ $`3\left(K^{(1)}\right)^2M\sqrt{{\displaystyle \frac{r}{r+2M}}}+14K_{(h_1)}^{(2)}M+21rK_{(G)}^{(2)}M]`$ (138) $`{\displaystyle \frac{\chi ^{(2)}}{t}}`$ $`=`$ $`{\displaystyle \frac{1}{14}}{\displaystyle \frac{r2M}{r^4(2r+3M)}}[140Mrh_1^{(2)}+84M^2r^2G_r^{(2)}+r^5\left(K_r^{(1)}\right)^215r^3\left(K^{(1)}\right)^22Mr^4\left(K_r^{(1)}\right)^2+`$ (139) $`+`$ $`2Mr^2\left(K^{(1)}\right)^2+2r^4K^{(1)}K_r^{(1)}+36r^3\left(K_{(h_{1o})}^{(1)}\right)^2+96r\left(K_{(h_{2o})}^{(1)}\right)^2+276M\left(K_{(h_{2o})}^{(1)}\right)^2`$ (140) $``$ $`144M^2r\left(K_{(h_{1o})}^{(1)}\right)^2+72r^2K_{(h_{1o})}^{(1)}K_{(h_{2o})}^{(1)}+12r^2K_{(h_{2o})}^{(1)}K_{(h_{2o})r}^{(1)}14Mr^3K_r^{(2)}+28r^3K^{(2)}`$ (141) $``$ $`28r^5G_{rr}^{(2)}28r^4G_r^{(2)}144MrK_{(h_{1o})}^{(1)}K_{(h_{2o})}^{(1)}24MrK_{(h_{2o})}^{(1)}K_{(h_{2o})r}^{(1)}+`$ (142) $`+`$ $`84M^2r^3G_{rr}^{(2)}+14Mr^4G_{rr}^{(2)}42Mr^3G_r^{(2)}168M^2rh_{1r}^{(2)}28Mr^2h_{1r}^{(2)}+56r^3h_{1r}^{(2)}+`$ (143) $`+`$ $`168M^2h_1^{(2)}56r^2h_1^{(2)}28r^3H_2^{(2)}28Mr^2H_2^{(2)}]`$ (144) $`\mathrm{\Theta }^{(2)}`$ $`=`$ $`{\displaystyle \frac{\left(r+2M\right)\left(2h_{2o}^{(2)}+rh_{2or}^{(2)}+2rh_{1o}^{(2)}\right)}{2r^3}}`$ (145) $`{\displaystyle \frac{\mathrm{\Theta }^{(2)}}{t}}`$ $`=`$ $`{\displaystyle \frac{1}{7}}{\displaystyle \frac{\sqrt{r2M}}{r^3\sqrt{r}}}[r^2K_{(h_{1o})}^{(1)}K^{(1)}14r^2K_{(h_{1o})}^{(2)}7r^2K_{(h_{2o})r}^{(2)}35K_{(h_{2o})}^{(2)}M+14rK_{(h_{2o})}^{(2)}+`$ (146) $`+`$ $`6rK_r^{(1)}K_{(h_{2o})}^{(1)}M+8rK^{(1)}K_{(h_{2o})r}^{(1)}M+14rK_{(h_{2o})r}^{(2)}M+28rK_{(h_{1o})}^{(2)}M4r^2K^{(1)}K_{(h_{2o})r}^{(1)}`$ (147) $``$ $`3r^2K_r^{(1)}K_{(h_{2o})}^{(1)}+8rK^{(1)}K_{(h_{2o})}^{(1)}20K^{(1)}K_{(h_{2o})}^{(1)}M2rK_{(h_{1o})}^{(1)}K^{(1)}M]`$ (148) ## VI Conclusions We have written explicitly the second order perturbative equations for axisymmetric $`\mathrm{}=2`$ perturbations including odd parity modes. We set up explicitly the evolution equations and the formulae for generating initial data starting from a metric and an extrinsic curvature. We also performed an asymptotic analysis to obtain formulas for radiated energies and waveforms. These formulae will be useful for considering perturbatively the evolution of spacetimes arising from the collision of nearby black holes. They might also be useful for other comparison with nonlinear evolutions, as the ones considered in . ###### Acknowledgements. This work was supported in part by grants of the National Science Foundation of the US INT-9512894, PHY-9423950, PHY-9800973, PHY-9407194, by funds of the University of Córdoba, the Pennsylvania State University and the Eberly Family Research Fund at Penn State. We also acknowledge support of CONICET and CONICOR (Argentina). JP also acknowledges support from the Alfred P. Sloan and John S. Guggenheim Foundations. RJG is a member of CONICET.
warning/0001/cond-mat0001150.html
ar5iv
text
# Microscopic theory of the coupling of intrinsic Josephson oscillations and phonons ## 1 INTRODUCTION The main features of the electronic $`c`$-axis transport in the high-$`T_c`$-superconductors Tl<sub>2</sub>Ba<sub>2</sub>Ca<sub>2</sub>Cu<sub>3</sub>O<sub>10+δ</sub> and Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+δ</sub> below the critical temperature $`T_c`$ can be well understood in the model of a stack of superconducting CuO<sub>2</sub>-layers, which form a onedimensional stack of intrinsic Josephson junctions . One of the few phenomena, which are specific to the intrinsic Josephsoneffect and cannot be seen in arrays of conventional superconductors, is the interaction of the Josephson oscillations and phonons. This has been observed experimentally in the form of resonances in the $`I_{\mathrm{dc}}`$-$`V_{\mathrm{dc}}`$-characteristics at low voltages . Their main features could already be understood theoretically in the framework of a simple lattice dynamical model of damped local oscillators in the barrier . Despite the clear evidence of the effect a general multi-band theory, in which parameters like the phonon frequencies or the oscillator strength in the dielectric function are specified microscopically, is still missing and shall be presented in the following. This more general formalism is in principle appropriate to include the eigenvalues and eigenvectors of the dynamical matrix as obtained in any microscopic lattice dynamical calculation into the Josephson theory. Qualitative effects of the phonon dispersion are discussed in a simple rigid ion model, e.g. the van-Hove singularity at the upper edge of the acoustical band, which can be detected in the $`I_{\mathrm{dc}}`$-$`V_{\mathrm{dc}}`$-curve. Details of the derivation and further results can be found elsewhere . ## 2 THEORY OF PHONON COUPLING The Josephson effect between the layers $`n`$ and $`n+1`$ is described by a RSJ-like equation $$j=j_c\mathrm{sin}\gamma _n(t)+j_{qp}(E_n(t))+ϵ_0\dot{E}_n^\rho (t),$$ where $`j`$ is the bias current (density), $`j_c\mathrm{sin}\gamma _n(t)`$ the supercurrent, $`j_{qp}(E_n)`$ the quasiparticle current and $`\gamma _n(t)`$ the gauge-invariant phase difference, which is related to the average electric field $`E_n(t)`$ in $`c`$-direction in the barrier: $`\mathrm{}\dot{\gamma }_n(t)=2edE_n(t)`$ ($`d`$: thickness of the barrier). In contrast to this, the displacement current density $`ϵ_0\dot{E}_n^\rho (t)`$ does not depend on $`E_n`$, but the field $`E_n^\rho (t)`$ set-up by the conduction-electron charge-fluctuations $`\delta \rho _n`$ on the layers, which obey the Poisson equation $`\delta \rho _n(t)=ϵ_0(E^\rho (t)E_{n1}^\rho (t)`$ and the continuity equation $`j_n(t)j_{n1}(t)=\delta \dot{\rho }_n(t)`$. All quantities are constant along the layers ($`q_{}=0`$) as in short junctions. The crucial point is to determine the (linear) relation between the field $$E^\rho (q_z,\omega )=EE^{\mathrm{ion}}=ϵ_{\mathrm{ph}}(q_z,\omega )E(q_z,\omega )$$ (1) created by the charge fluctuations alone and the total electric field $`E(q_z,\omega )`$ due to both electrons and ions. This can be described by the (longitudinal) phonon dielectric function $$ϵ_{\mathrm{ph}}^L(q_z,\omega )=\frac{ϵ_{\mathrm{}}}{1\chi (q_z,\omega )},$$ (2) where the suszeptibility ($`\lambda `$: phonon branch) $$\chi (q_z,\omega )=\underset{\lambda }{}\frac{|\mathrm{\Omega }(q_z\lambda )|^2}{\omega ^2(q_z\lambda )\omega ^2}$$ (3) and the oscillator strength $$|\mathrm{\Omega }(q_z\lambda )|^2=\underset{\kappa \kappa ^{}}{}\stackrel{~}{Z}_\kappa \frac{e_z(\kappa |q_z\lambda )e_z^{}(\kappa ^{}|q_z\lambda )}{v_cϵ_0\sqrt{M_\kappa M_\kappa ^{}}}\stackrel{~}{Z}_\kappa ^{}^{}.$$ (4) depend on the eigenvalues $`\omega (q_z,\lambda )`$ and eigenvectors $`\stackrel{}{e}(\kappa |q_z\lambda )`$ of the dynamical matrix $`D_{\alpha \beta }(\genfrac{}{}{0pt}{}{\stackrel{}{q}}{\kappa \kappa ^{}})`$, which contains all the short and long range interaction between the ionic cores in the insulating state. The eigenvalues $`\omega (q_z,\lambda )`$ therefore represent the bare phononfrequencies in the absence of the charge fluctuations $`\delta \rho _n`$. The appearence of special $`q_z`$-dependent effective charges $`\stackrel{~}{Z}_\kappa (q_z)`$ in equ. 4 reflects the different contribution of ions on and between the superconducting layers and is one of the distinctions to the conventional dielectric function. The function $`ϵ_{\mathrm{ph}}^L(q_z,\omega )`$ has zeros at the longitudinal eigenfrequencies $`\omega (q_z,\lambda )`$ of modes with polarization in $`c`$-direction. A similar consideration shows the general form $$ϵ_{\mathrm{ph}}(\stackrel{}{q},\omega )=1+\frac{\chi (\stackrel{}{q},\omega )}{1f(\stackrel{}{q})\chi (\stackrel{}{q},\omega )}$$ (5) of the phononic dielectric constant, which reduces in special cases to the longitudinal $`ϵ_{\mathrm{ph}}^L(q_z)=ϵ_{\mathrm{ph}}(q_{}=0,q_z)`$ and transversal dielectric constant $`ϵ_{\mathrm{ph}}^T(q_{})=ϵ_{\mathrm{ph}}(q_{},q_z=0)=ϵ_{\mathrm{}}^T+\chi (q_{},\omega )`$, which is relevant in optical experiments. With the knowledge of the phononic dielectric function $`ϵ_{\mathrm{ph}}`$ the $`I_{\mathrm{dc}}`$-$`V_{\mathrm{dc}}`$-curve can be calculated with the ansatz $$\gamma _n(t)=\{\begin{array}{cc}\gamma _0+\omega t+\delta \gamma _n(t)\hfill & n\mathrm{resistive}\hfill \\ \gamma _0+\delta \gamma _n(t)\hfill & n\mathrm{else}\hfill \end{array},$$ (6) by linearizing in the small oscillating parts $`\delta \gamma _n(t)1`$. The final result for the first branch ($`V_{\mathrm{dc}}=\mathrm{}\omega /(2e)`$, $`\omega _p^2:=2edj_c/(\mathrm{}ϵ_0)`$) $`j(V)`$ $`=`$ $`j_{\mathrm{qp}}(V){\displaystyle \frac{j_c}{2}}{\displaystyle \frac{\omega _p^2}{\omega ^2}}\mathrm{Im}{\displaystyle \frac{1}{\stackrel{~}{ϵ}(\omega )}}`$ (7) exhibits a phonon resonance exactly at the zeros of the modified dielectric function $$\stackrel{~}{ϵ}(\omega )=\left[\frac{1}{N_z}\underset{q_z}{}\frac{1}{ϵ_{\mathrm{ph}}^L(q_z,\omega )\frac{\omega _p^2}{\omega ^2}+\frac{i\sigma }{ϵ_0\omega }}\right]^1+\frac{\omega _p^2}{\omega ^2},$$ (8) i.e. in the unrenormalized phonon bands $`\omega (\stackrel{}{q},\lambda )`$. More specifically peaks develop at the van-Hove Singularities of the phonon density of states due to the averaging over $`q_z`$ in equ. 8. As a reliable microscopic lattice dynamical theory for the dynamical matrix is not yet available and as the intrinsic Josepshon effect has to be treated in a phenomenological (one-band) picture until the details of the pairing mechanism are understood, the above results will be illustrated qualitatively in a simple model of a two-atomic chain of masses $`M_\kappa `$ and charges $`Z_1=Z_2`$. If the mass on the layer ($`M_l`$) is larger than the one in the barrier ($`M_b`$), the acoustical van-Hove singularity is suppressed, but the optical band shows a double peak at the band edges (cf. fig. 1), the distance being the bandwidth. The (more realistic) case of $`M_l<M_b`$ exhibits single peaks at the upper edges of both acoustical and optical bands and is compared with experiment in fig. 3. If several junctions are in the resistive state, their dynamics is coherent due to the coupling with phonons, even if no other (e.g. inductive) interaction is present. This is important for the application of short junctions as effective high-frequency devices. Analytical expressions similar to equ. 7 for the second branch can in principle distinguish different phase locked solutions in the $`I_{\mathrm{dc}}`$-$`V_{\mathrm{dc}}`$-characteristic . ## 3 COLLECTIVE MODES The zeros of the real part of the total dielectric function $$ϵ_{\mathrm{tot}}(\stackrel{}{q},\omega )=ϵ_{\mathrm{ph}}(\stackrel{}{q},\omega )\frac{\omega _p^2(\stackrel{}{q})}{\omega ^2}+\frac{i\sigma }{ϵ_0\omega },$$ (9) define the frequencies $`\omega _{\mathrm{el}\mathrm{ph}}(\stackrel{}{q},\omega )`$ of the collective modes of the coupled conduction electrons and phonons, where the Josephson plasma frequency is given by ($`\lambda _{ab}`$: coherence length): $$\omega _p(\stackrel{}{q})=\omega _p^2(1+\frac{c_0^2q_{}^2/\omega _p^2}{12\frac{\lambda _{ab}^2}{d^2}(\mathrm{cos}(q_zd)1)}).$$ (10) These modes are different from the bare phonon bands $`\omega (\stackrel{}{q},\lambda )`$ and can in principle be observed in neutron scattering. In fig. 2 the plasma band $`\omega _{\mathrm{pl}}(\stackrel{}{q})`$ for $`q_xq_z`$ mixes with the longitudinal acoustical (LA) and optical (LO) phonon modes $`\omega (\stackrel{}{q},\lambda )`$, which are almost dispersionless on this scale. Note especially that a frequency gap of width $`\sqrt{2\omega _{\mathrm{pl}}\mathrm{\Omega }}`$ appears in the spectrum, which is formally (not physically) similar to the polariton dispersion of coupled phonons and photons . ## 4 EXPERIMENTAL RESULTS Recently the explanation of the subgap resonances in Refs. with the phonon coupling mechanism presented here could be well confirmed by Raman measurements on the same samples and infrared reflectivity experiments with grazing incidence (see table 1). Note that in our theory also Raman-active or silent modes may couple to intrinsic Josephson oscillations for $`q_z0`$ and that in contrast to optical experiments (at the $`\mathrm{\Gamma }`$-point) the whole Brioullin zone contributes in an average way (cf. equ. 8). This provides a reason for slight discrepancies of these experimental data in table 1 For more experimental and theoretical references see . The qualitative features of the subgap resonances can already be understood in the local oscillator model of : The position of the resonance is completely independent on temperature, magnetic field or the geometry of the probe, while the intensity of the structure varies $`j_c^2`$ with the critical current density $`j_c(T,B)`$. Also the behaviour in external pressure is consistent with the phonon interpretation . Beyond this, in the more general model developed above resonances at van-Hove singularities, e.g. at the upper band edge of the acoustical phonon band, can be described. This might be an explanation for a peak seen in at $`3.2`$ mV ($`\widehat{=}1.54`$ THz) in the $`I_{\mathrm{dc}}`$-$`V_{\mathrm{dc}}`$-characteristic of Tl<sub>2</sub>Ba<sub>2</sub>Ca<sub>2</sub>Cu<sub>3</sub>O<sub>10</sub>, as the same frequency is predicted in lattice dynamical calculations for the upper edge of the acoustical band and because there are no optical phonon bands at that low frequency (cf. Fig. 3). Also the double peaks shown in fig. 1 might have been seen in the satellite structures at 5.17 THz (10.7 mV) and 5.6 (11.6 mV) in the $`I_{\mathrm{dc}}`$-$`V_{\mathrm{dc}}`$-curve of Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+δ</sub> , which is consistent with the theoretical width $`0.3`$ THz of this phonon band . ## 5 CONCLUSIONS In this contribution the microscopic theory for the coupling between Josephson oscillations and longitudinal phonons in intrinsic Josephson systems like the highly anisotropic cuprate superconductors has been developed. The analytical result equ. (7) for the $`I_{\mathrm{dc}}`$-$`V_{\mathrm{dc}}`$-curve for one resistive junction and the detailed form of the longitudinal dielectric function equ. (2) describing the coupling has been obtained. In paritcular, not only optical but also acoustical phonons at the edge of the Brillouin zone couple to Josephson oscillations explaining a structure observed in in the $`I_{\mathrm{dc}}`$-$`V_{\mathrm{dc}}`$-curve occurring at low voltage. The discussion of the dispersion of collective phonon-electron modes is an important first step to understand the effect of phonons in long junctions . The authors thank A. Yurgens, A. Tsvetkov, A. Mayer, D. Strauch, R. Kleiner, P. Müller, L. Bulaevskii and A. Bishop for fruitful discussions and DFG, FORSUPRA and DOE under contract W-7405-ENG-36 (C.H.) for financial support.
warning/0001/hep-ph0001092.html
ar5iv
text
# Diffraction: A New Approach ## Abstract A phenomenological model of hard diffraction is presented, in which the structure of the Pomeron is derived from the structure of the parent hadron. Predictions for diffractive deep inelastic scattering are compared with data. The inclusive and diffractive deep inelastic scattering (DIS) cross sections are proportional to the corresponding $`F_2`$ structure functions of the proton, | Inclusive DIS | $`\frac{d^2\sigma }{dxdQ^2}\frac{F_2^h(x,Q^2)}{x}`$ | | --- | --- | | Diffractive DIS | $`\frac{d^3\sigma }{d\xi dxdQ^2}\frac{F_2^{D(3)}(\xi ,x,Q^2)}{x}`$ | where the superscripts $`h`$ and $`D(3)`$ indicate, respectively, a hard structure function (at scale $`Q^2`$) and a 3-variable diffractive structure function (integrated over $`t`$). The latter depends not only on the hard scale $`Q^2`$, but also on the soft scale, $`M_T1`$ GeV, which is the relevant scale for the formation of the diffractive rapidity gap. The only marker of the rapidity gap is the variable $`\xi `$. We therefore postulate that the rapidity gap probability is proportional to the soft parton density at $`\xi `$ and write the DDIS (diffractive DIS) cross section as $$\frac{d^3\sigma }{d\xi dxdQ^2}\frac{F_2^h(x,Q^2)}{x}\times \frac{F_2^s(\xi )}{\xi }\xi \mathrm{norm}$$ where the symbolic notation “$`\xi \mathrm{norm}`$” is used to indicate that the $`\xi `$ probability is normalized. Since $`x=\beta \xi `$, the normalization over all available $`\xi `$ values involves not only $`F_2^s`$ but also $`F_2^h`$, breaking down factorization. It is therefore prudent to write the DDIS cross section in terms of $`\beta `$ instead of $`x`$, so that the dependence of $`F_2^h`$ on $`\xi `$ be shown explicitly: $$\frac{d^3\sigma }{d\xi d\beta dQ^2}\frac{1}{\beta }\left[F_2^h(\beta \xi ,Q^2)\times \frac{F_2^s(\xi )}{\xi }\xi \mathrm{norm}\right]$$ The term in the brackets represents the DDIS structure function $`F_2^{D(3)}(\xi ,\beta ,Q^2)`$. In the next step, we seek guidance from the scaling behavior of the soft single-diffractive (sd) differential cross section , $$\frac{d\sigma _{sd}}{dM^2}\frac{1}{(M^2)^{1+ϵ}}\text{(no }s\text{-dependence!)}$$ which in terms of $`\xi `$ takes the form $$\frac{d\sigma _{sd}}{d\xi }\underset{\text{gap probability}}{\underset{}{\frac{1}{s^{2ϵ}}\frac{1}{\xi ^{1+2ϵ}}}}\times (s^{})^ϵ$$ where $`s^{}M^2`$ is the s-value of the diffractive sub-system. Noting that $`\xi `$ is related to the associated rapidity gap by $`\mathrm{\Delta }Y=\mathrm{ln}\frac{1}{\xi }`$, and that the integral $`_{s_{}/s}^1\frac{1}{s^{2ϵ}}\frac{d\xi }{\xi ^{1+2ϵ}}`$ is equal to a constant, the above equation may be viewed as representing the product of the total cross section at the sub-system energy multiplied by a normalized rapidity gap probability. In analogy with this experimentally established behavior, we factorize $`F_2^{D(3)}(\xi ,\beta ,Q^2)`$ into $`F_2^h(\beta ,Q^2)`$, the sub-energy DIS cross section, times a normalized gap probability: $$F_2^{D(3)}(\xi ,\beta ,Q^2)=P_{gap}(\xi ,\beta ,Q^2)\times F_2^h(\beta ,Q^2)$$ The gap probability is therefore given by<sup>2</sup><sup>2</sup>2In hep-ph/9911210, a factor of $`\frac{1}{\beta }`$ was erroneously included in Eq. (13) of the gap probability; however, this factor is carried over into the normalization factor $`N(s,\beta ,Q^2)`$ through Eqs. (14) and (15) and cancels out in the final result for $`F_2^{D(3)}`$ in Eq. (16). $$P_{gap}(\xi ,\beta ,Q^2)=F_2^h(\beta \xi ,Q^2)\times \frac{F_2^s(\xi )}{\xi }\times N(s,\beta ,Q^2)$$ The normalization factor, $`N(s,\beta ,Q^2)`$, is obtained from the following equation, using $`\xi _{min}=Q^2/s`$, $$N^1(s,\beta ,Q^2)=\frac{1}{f_q}_{\xi _{min}}^1F_2^h(\beta \xi ,Q^2)\times \frac{F_2^s(\xi )}{\xi }𝑑\xi $$ where $`f_q`$ is the quark fraction of the hard structure and is used here since only quarks participate in DIS. At small $`x`$ ($`0.1`$), the structure functions $`F_2^h`$ and $`F_2^s`$ are represented well by the power law expressions $`F_2^h(x,Q^2)=A^h/x^{\lambda _h(Q^2)}`$ and $`F_2^s(\xi )=A^s/\xi ^{\lambda _s}`$. Using these forms we obtain $$N^1(s,\beta ,Q^2)=\frac{1}{f_q}\left[\frac{A^h}{\beta ^{\lambda _h}}\frac{A^s}{\lambda _h+\lambda _s}\left(\frac{\beta s}{Q^2}\right)^{\lambda _h+\lambda _s}\right]$$ $$F_2^{D(3)}(\xi ,\beta ,Q^2)=\frac{1}{\xi ^{1+\lambda _h+\lambda _s}}\times f_q(\lambda _h+\lambda _s)\left(\frac{Q^2}{\beta s}\right)^{\lambda _h+\lambda _s}\times \frac{A^h}{\beta ^{\lambda _h}}$$ Since in DDIS $`x`$ is always smaller than $`\xi `$, the above form of $`F_2^{D(3)}`$, derived for small $`x`$, should be valid for all $`x`$ when $`\xi `$ is small; it should also be valid for all $`\beta (=x/\xi )`$. We therefore expect $`F_2^{D(3)}`$ to have the following $`\xi `$ and $`\beta `$ dependence at small $`\xi `$: $$F_2^{D(3)}(\xi ,\beta ,Q^2)|_{\beta ,Q^2}\frac{1}{\xi ^{1+n}}n=\lambda _h(Q^2)+\lambda _s$$ $$F_2^{D(3)}(\xi ,\beta ,Q^2)|_{\xi ,Q^2}\frac{1}{\beta ^m}m=2\lambda _h(Q^2)+\lambda _s$$ The HERA (non-diffractive) DIS measurements yield $`\lambda _s0.1`$, which is in agreement with the value of $`ϵ=\alpha (0)1=0.104`$ , where $`\alpha (0)`$ is the intercept of the Pomeron trajectory at t=0. In the $`Q^2`$ range of 10-50 GeV<sup>2</sup>, where the DDIS data are concentrated, these measurements yield $`\lambda _h0.3`$. Using these values we obtain $`n=0.4`$ and $`m=0.7`$. We therefore expect $$\text{Prediction:}F_2^{D(3)}\frac{1}{\xi ^{1.4}}\times \frac{1}{\beta ^{0.7}}$$ We observe the following features: Factorization Our prediction exhibits factorization between $`\xi `$ and $`\beta `$, in agreement with HERA results at small $`\xi `$. $`\xi `$-dependence In the Regge framework, the $`\xi `$-dependence of $`F_2^{D(3)}`$ is expected to have the “Pomeron flux” form of $`1/\xi ^{1+n}`$ with $`n=2ϵ=0.2`$, independent of $`Q^2`$. In the $`Q^2`$ range of 10-50 GeV<sup>2</sup>, the HERA experiments find that $`n`$ is $`0.4`$ and has a small $`Q^2`$ dependence, in agreement with our prediction of $`n=\lambda _h(Q^2)+\lambda _s`$. $`\beta `$-dependence The predicted form $`1/\beta ^m`$ for $`F_2^{D(3)}`$ is valid in the region of (fixed) small $`\xi `$ and high $`Q^2`$, where the $`x`$-distribution of $`F_2(x,Q^2)`$ has the form $`A^h/x^{\lambda _h(Q^2)}`$. As there are no data points at strictly fixed $`\xi `$, we have selected the following set of five points at $`\xi 0.01`$ and $`Q^2=45`$ GeV<sup>2</sup> from Ref. : | $`\beta `$ | $`x`$ | $`\xi =x/\beta `$ | $`\xi F_2^{D(3)}\pm stat\pm syst`$ | | --- | --- | --- | --- | | 0.10 | 0.00133 | 0.0133 | $`0.0384\pm 0.0066\pm 0.0030`$ | | 0.20 | 0.00237 | 0.0118 | $`0.0406\pm 0.0061\pm 0.0026`$ | | 0.40 | 0.00421 | 0.0105 | $`0.0215\pm 0.0046\pm 0.0016`$ | | 0.65 | 0.00750 | 0.0115 | $`0.0240\pm 0.0054\pm 0.0026`$ | | 0.90 | 0.00750 | 0.0083 | $`0.0088\pm 0.0041\pm 0.0005`$ | Figure 1 shows the above values of $`\xi F_2^{D(3)}(\beta )`$ versus $`\beta `$ along with our prediction (solid line). The following parameters were used in the calculation of $`F_2^{D(3)}`$: $`\sqrt{s}=280`$ GeV, $`\xi =0.01`$, $`Q^2=45`$ GeV<sup>2</sup>, $`\lambda _s=0.1`$, $`\lambda _h=0.3`$, $`f_q=0.4`$ , and $`A^h=0.2`$; the latter was evaluated from $`F_2(Q^2=50,x=0.00133)=1.46`$ assuming a $`\frac{A^h}{x^{0.3}}`$ dependence. The observed agreement between data and prediction, both in shape and normalization, is considered satisfactory, particularly since no free parameters are used in the calculation.
warning/0001/cond-mat0001295.html
ar5iv
text
# Asymmetric exclusion process with next-nearest-neighbor interaction: some comments on traffic flow and a nonequilibrium reentrance transition ## I Introduction One of the basic properties of many driven, interacting many-body systems is the occurrence of shocks. A shock in a system of classical flowing particles marks the sudden transition from a region of low density to a region of high density. A well-known example for a shock is the beginning of a traffic jam on a motorway where incoming cars (almost freely flowing particles in the low density regime) have to slow down very quickly over a short distance and then form part of the (high-density) congested region. A remarkable feature of such shocks is their stability over long periods of time, i.e., they remain localized over distances comparable to the size of particles. In some sense one may regard shocks as soliton-like collective excitations of the particle system. Lattice gas models have proven to be excellent systems for the theoretical investigation of shocks and of the consequences of shocks for the collective behavior of the particle system. An interesting situation is the coupling of such a driven lattice gas system to external particle reservoirs . It is intuitively clear that unlike in equilibrium systems, here the boundaries will play a decisive part in determining the bulk behavior of the system: Since the system is open at the boundaries, particles will flow in and out and the current will carry boundary effects into the bulk. Among the fundamental questions to ask in such a set-up is the stationary (i.e. long-time) behavior of the system as a function of the boundary densities. Numerical observations and mean-field based arguments show that varying boundary densities leads to boundary-induced phase transitions . To understand why this happens it is clearly necessary to get insight into the collective behavior of the lattice gas and to investigate the role of the shocks. The best-studied example is the one-dimensional asymmetric simple exclusion process (ASEP) , a hard-core lattice gas where each lattice site can be occupied by at most one particle and particles hop stochastically with constant bias to vacant nearest-neighbor sites. For this model not only the structure and motion of shocks is largely understood, but also the stationary phase diagram of the open system and the exact particle-density profiles are explicitly known as function of the external reservoir densities. Based on the exact solution of the ASEP and of a related lattice gas model the role of shocks for the stationary phase of more generic open driven diffusive systems was elucidated . This has led to a theory of boundary-induced phase transitions for one-component driven particle systems in one dimension , reviewed below. Of particular interest are systems where the stationary current $`j(\rho )`$ as a function of particle density $`\rho `$ has a single maximum, an example being traffic flow on single-lane or multi-lane highways . The theory predicts a phase diagram with a first-order (non-equilibrium) phase transition between a low-density phase (LD) and a high density phase (HD) and a continuous (second-order) order transition from both phases to a maximal-current phase (MC), see Fig. 2 below. In the context of traffic flow the properties of the second-order transition suggest more efficient control mechanisms for avoiding jams . The strength of the theory lies in the small number and generality of its assumptions. Hence it is tempting to test its validity in real traffic flow. The openness of the system models in- and outflow of cars on a road between two junctions. Indeed, the main features of the predicted first-order transition have been observed using data collected on a German motorway . The second-order transition has been confirmed by Monte-Carlo simulations of a suitably modified Nagel-Schreckenberg model for traffic flow, originally introduced only for periodic boundary conditions . Since to our knowledge the data necessary to test the existence of the second-order transition in real traffic are not available at present, independent investigations of other traffic flow models are required to establish the presence of such a transition. We consider here a continuous-time exclusion model with additional short-range interaction on top of the pure hard-core repulsion of the usual ASEP. The model is designed to be exactly solvable like the ASEP (to some degree) on the one hand and to be somewhat more realistic in describing real traffic than the ASEP on the other hand. However, given the wide range of experimental applications of hard-core lattice gases which comprise phenomena as diverse as the kinetics of protein synthesis , diffusion in thin channels , or polymer reptation , broader relevance of our model is to be expected. Indeed, in order to obtain a more complete picture we also analyse a similar model with attractive next-nearest-neighbour interaction. This model is not related to traffic flow, but describes driven hard-core particles with attractive short-range interaction. To our knowledge, this is the first investigation of the steady-state selection of a driven diffusive system with this type of interaction. The paper is organized as follows. In Sec. II we comment on some of the requirements of traffic flow modeling and we describe our model. This model is a special case of the driven diffusive systems studied some while ago by Katz, Lebowitz and Spohn . We explain in which respect this model is more realistic for traffic flow than the standard ASEP which is a limiting case of our lattice gas. This phenomenological explanation is then confirmed by the calculation of the exact flow diagram, i.e. the stationary bulk current as a function of the density. In this section the emphasis is on bulk properties and hence we investigate the system with periodic boundary conditions in the thermodynamic limit. In Sec. III we review some details of the theory of boundary-induced phase transitions of Ref. and we introduce the boundary dynamics for modeling the coupling of a finite system to boundary reservoirs of constant density. We give an exact solution for the stationary distribution along the line in the phase diagram corresponding to equal boundary densities (with proof postponed to the appendices) and discuss a mean-field analysis of the full phase diagram. In Sec. IV we present a mean-field analysis of the full phase diagram and discuss Monte Carlo data for the phase transition lines while in Sec. V we perform a similar analysis for the model with attractive interaction. Finally we summarize our findings and present some concluding remarks (Sec. VI). Technical details of the derivation of exact results are presented in the appendices. ## II ASEP with next-nearest-neighbor interaction - some comments on modelling traffic flow The ASEP is a stochastic lattice gas of hard-core particles with biased particle hopping. Particles hop with exponential waiting-time distribution with parameter 1 to their nearest-neighbor site to the right, provided this site is empty. If the site is occupied, the hopping attempt is rejected. Symbolically one may represent these stochastic dynamics as follows $`A\mathrm{}`$ $``$ $`\mathrm{}A\text{ with rate }1.`$ (1) Here $`A`$ represent a particle and $`\mathrm{}`$ represents the vacant neighboring site. Even though this stochastic process is too simple to be a realistic model for traffic flow, some qualitative features of real traffic can already be seen: Shocks exist and the stationary current $`j(\rho )=\rho (1\rho )`$ as a function of particle density $`\rho `$ has a single maximum. An apparently unrealistic feature of the particle distribution is the absence of correlations in the steady state which are seen in more sophisticated traffic flow models . An unrealistic feature of the current-density relation (known in traffic engineering as flow diagram) is the reflection symmetry w.r.t. the maximal-current density $`\rho ^{}=1/2`$ and its rounded shape close to the maximum. The basic mechanisms which determine traffic flow appear to be firstly the competition between the desire to reach an optimal speed while keeping a (velocity-dependent) safety distance. When the traffic density becomes sufficiently large this distance cannot be kept anymore and the speed has to be reduced. As a consequence the current drops at some density $`\rho ^{}`$. Secondly, there is a certain amount of randomness due to variations in individual drivers behavior. This “noise” necessitates a statistical description of traffic flow phenomena. Various one-dimensional lattice gas models which incorporate these mechanisms in different manners have been proposed . The following is part of the picture that emerges: (i) The existence of a shock is generic and appears to be the consequence of the non-linear current-density relation . (ii) The symmetric shape of the current-density relation results from particle-hole symmetry and is a feature of models in which cars move with constant probability or rate, independently of the environment beyond the nearest neighbor site to which they move. This is an unrealistic assumption since clearly car drivers slow down when they see a slowly moving car already some distance ahead. They do not just perform an emergency stop when the car is immediately in front of them. Numerical and analytical results for models which allow for a reduction of speed that depends on the occupation of sites further ahead show an asymmetric current-density relation resembling the shape of the current-density relation of real traffic. In these models speed is implemented by jumps over a variable number of lattice sites. (iii) The unrealistically round shape of the current-density relation at $`\rho ^{}`$ is specific for the ASEP. Deterministic exclusion processes with parallel update also show a symmetric current-density relation with one maximum, but the derivative of the current is discontinuous at the maximal-current density $`\rho ^{}`$, in this respect resembling the shape of the current in real traffic and of more realistic traffic flow models . Increasing the hopping probability in a discrete-time process towards deterministic hopping, leads to an increasingly sharp jump in the current derivative at $`\rho ^{}`$ . Hence the rather broad exponential waiting-time distribution of the standard ASEP seems to be responsible for the round shape of the current-density relation at $`\rho ^{}`$. In terms of the motion of a single particle these dynamics correspond to an overestimated single-particle diffusion coefficient. (iv) For parallel update, but not for sublattice parallel update, increasing the hopping probability strengthens antiferromagnetic particle correlations , i.e., cars are less likely to be found close to each other than some distance apart. In order to further investigate this picture and to disentangle the various effects associated with a small diffusion coefficient and with speed-reduction respectively it would be interesting to study both ingredients separately for models in which the stationary distribution can be calculated exactly. A small single-particle diffusion coefficient can be implemented in discrete-time models choosing the hopping probability close to one (low noise). In the single-speed ASEP this has been shown to lead to the (almost) discontinuous behavior of the current derivative discussed above. The numerical results on the effect of a small diffusion coefficient on correlations are inconclusive. In our toy model we keep the exponential waiting-time distribution (large diffusion coefficient), but introduce a next-nearest-neighbor interaction which in the repulsive case models slowing down of a car if the next-nearest-neighbor site is occupied as well. A particle hops to the right with rate $`r`$ if the next-nearest-neighbor site is empty and with rate $`q`$ if it is occupied: $`A\mathrm{}\mathrm{}`$ $``$ $`\mathrm{}A\mathrm{}\text{ with rate }r`$ (2) $`A\mathrm{}A`$ $``$ $`\mathrm{}AA\text{ with rate }q.`$ (3) The condition $`q<r`$ models slowing-down, in the limiting case $`r=q`$ one recovers the usual ASEP. For $`q>r`$ this model has not an interpretation as traffic model, but may be regarded as describing hard-core particles with attractive short-range interaction which are driven by an external field. This model is a special case in the class of driven diffusive systems investigated in Ref. . On a ring with $`N`$ sites with periodic boundary conditions the stationary distribution turns out to be given by the equilibrium distribution of the one-dimensional Ising model. Each state of the system is defined by the set of occupation numbers $`\underset{¯}{n}=\{n_1,\mathrm{},n_N\}`$ with $`n_i=0,1`$. The stationary probability of finding a state $`\underset{¯}{n}`$ is given by $$P^{}(\underset{¯}{n})=\frac{1}{Z_N}\left(\frac{q}{r}\right)^{_{i=1}^N(n_in_{i+1}+hn_i)}.$$ (4) Here $`Z_N`$ is the partition function and the “chemical potential” $`h`$ parametrizes the conserved bulk density $`\rho `$. This grand-canonical distribution is a non-equilibrium stationary state, i.e., it is invariant under the stochastic time evolution, but it does not satisfy detailed balance with respect to the dynamics. It is interesting to notice that correlations are non-vanishing and, in the repulsive case $`q<r`$ which corresponds to speed reduction, become antiferromagnetic. In fact, the stationary state is identical to that of the discrete-time ASEP with parallel update for hopping probability $`p=1q/r`$. According to the dynamics described above the local density satisfies the continuity equation $$\frac{d}{dt}n_i=j_{i1}j_i$$ (5) with the current $$j_i=n_i(1n_{i+1})[qn_{i+2}+r(1n_{i+2})]$$ (6) between sites $`(i,i+1)`$. The stationary particle current $`j=j_i`$ is constant and is readily calculated using standard transfer matrix techniques for the one-dimensional Ising model (see Appendix A). In the thermodynamic limit $`N\mathrm{}`$ one finds the exact current density relation $$j=r\rho \left[1+\frac{\sqrt{14\rho (1\rho )(1q/r)}1}{2(1\rho )(1q/r)}\right]$$ (7) shown in (Fig. 1) in the repulsive case. ## III Open boundaries The analysis of the previous section yields the bulk properties of the lattice gas at a given density. Now we address the question of the steady-state selection of the system with open boundaries, i.e., we investigate the bulk density of an open system coupled at both ends to reservoirs of different fixed boundary densities. Translated into traffic language, open boundaries correspond to traffic junctions at the ends of a road where particles (cars) enter or leave the road respectively with certain fixed attempt rates, a situation envisaged with different aims already in for the ASEP with parallel update. As a result of the coupling to reservoirs a non-trivial stationary density profile in the vicinity of the boundaries will emerge and the (spatially constant) bulk-density will be a function of the two reservoir densities. There are two distinct mechanisms at work. Firstly, because of the particle interaction, coupling of a semi-infinite system to a reservoir will generically lead to some discontinuous behavior of the stationary distribution close to the boundary. The boundary represents an inhomogeneity of the system since the interaction of the particles with the fixed boundary leads to different dynamics than that which results from the interaction of particles among themselves. This is a non-universal phenomenon which depends on the precise nature of the coupling mechanism and on the nature of the particle interaction. For short-range hopping and systems with short-range stationary bulk correlations one expects the following picture: Coupling of a semi-infinite system at site 1 to a reservoir of constant density $`\rho _L`$ will give rise to a non-universal boundary density profile starting at $`\rho _L`$ and approaching (on the scale of lattice units) some bulk density $`\rho _{}`$ which is a non-universal function of $`\rho _L`$ (Fig. 2). A similar picture holds for coupling of a semi-infinite system at the right boundary where particles flow out of the system into the reservoir. Here the bulk density $`\rho _+`$ may change close to the boundary to the reservoir density $`\rho _R`$. Superimposed on this non-universal boundary structure is a universal behavior which depends only the effective boundary densities $`\rho _{},\rho _+`$ close to, but not at the boundaries. The theory of boundary induced phase transition describes the stationary phase diagram in terms of these effective boundary densities. ### A Theory of boundary-induced phase transitions We review only the principal ideas of this theory which is based on an interplay of the collective velocity $$v_c=\frac{j}{\rho }$$ (8) of the lattice gas and the shock velocity $$v_s=\frac{j_+j_{}}{\rho _+\rho _{}}$$ (9) of a shock with limiting densities $`\rho _+`$ and $`\rho _{}`$ and with limiting currents $`j_+`$ and $`j_{}`$ to the right and to the left respectively. The collective velocity is the velocity of the center of mass of a local perturbation in a homogeneous stationary background (Fig. 3a). It is positive for background density $`\rho <\rho ^{}`$, but becomes negative for $`\rho >\rho ^{}`$. In this case the perturbation creates a back-moving traffic jam, which leads to a negative center-of-mass velocity, even though all individual particles move with positive velocity. In terms of traffic flow one might think of such a perturbation as being a car which has just entered a major throughway from some side road. The shock velocity describes the motion of the shock which performs, due to fluctuations, a biased random walk with velocity $`v_s`$ (Fig. 3b). If the incoming current $`j_{}`$ exceeds the outgoing current $`j_+`$, the shock velocity is negative. This is analogous to the back-moving shock of a traffic jam for sufficiently high incoming traffic flow. To get an intuitive understanding how these velocities determine the stationary phase diagram of a driven system coupled to boundary reservoirs let us assume $`\rho _\pm =\rho _{R,L}`$. We consider first $`\rho _+>\rho ^{}`$ where $`\rho ^{}`$ is the density where the current takes its maximal value $`j^{}`$. To make the argument more transparent we also assume that particles hop only to the right and that initially the lattice is empty. Because of ergodicity the stationary distribution is independent of the initial state and hence this assumption involves no loss of generality. Consider now the time evolution of the averaged density profile in a large system, starting from the empty lattice. We start the discussion by assuming the left boundary density to be very low. (i) As time proceeds, particles from the left reservoir will enter the system and (possibly after some distance describing the non-universal boundary layer) create a region of constant density $`\rho _{}`$. This region decays to the right to zero, because after a finite time the rightmost particles will have traveled only a finite distance. Eventually however, after a time which is of the order of system size, the rightmost particles will hit the right boundary with the reservoir of density $`\rho _+`$. This reservoirs makes it more difficult for particles to travel further and hence creates a little traffic jam. The result is a shock profile, with shock densities $`\rho _{}`$ on the left and $`\rho _+`$ to the right resp. like in Fig. 3b. \[For the sake of the argument one could have chosen such an initial state. The reason for choosing an empty initial state becomes clear below.\] The decisive question is now how this shock profile evolves in time. According to (9) the shock velocity under the circumstances described here is positive, simply because the incoming current of particles $`j_{}`$ is less than the outgoing current $`j_+`$ for sufficiently small left boundary density. Hence, even though a shock forms by fluctuations, it has an average drift towards the right boundary. Hence the system remains in the low density (LD) regime with bulk density $`\rho =\rho _{}<\rho ^{}`$. (ii) The situation changes when the left boundary density takes a value such that the incoming current $`j_{}`$ equals the outgoing current $`j_+`$. In this case the shock velocity vanishes and the shock performs an unbiased random walk over the lattice. Hence the density profile may be regarded as being composed of two stationary domains with densities $`\rho _{}`$ and $`\rho _+`$, separated by a “domain wall” which is the sharp transitional region of the shock. Since the shock motion is unbiased, the stationary probability of finding the shock is constant in space. This leads to an equal superposition of shock profiles and hence to a linearly increasing stationary density profile. In analogy to first order equilibrium phase transitions where a domain wall separates regions of coexisting equilibrium regimes, we call the line defined by $`j_+=j_{}`$ and $`\rho _{}<\rho ^{}`$, $`\rho _+>\rho ^{}`$, a first order phase transition line. (iii) This line marks the transition to a high-density phase (HD) with bulk density $`\rho =\rho _+>\rho ^{}`$ since for even higher left boundary density the incoming current into the shock exceeds the outgoing current, and according to (9) the shock moves to the left. This leaves the bulk in the high-density regime determined by the coupling to the right boundary reservoir. At the first-order transition line the stationary bulk density is discontinuous, it jumps from $`\rho _{}`$ to $`\rho _+`$. Next we consider the case of low right boundary density $`\rho _+<\rho ^{}`$. For definiteness we choose $`\rho _+=0`$. The essential part of the following discussion, the explanation of the occurrence of a continuous phase transition, is unaffected by this choice. Again we start the discussion with the empty lattice. (i) As argued above, after some initial time the system will have filled up to a bulk density $`\rho =\rho _{}`$. Since $`\rho _+=0`$, particles hitting the right boundary can leave the system without creating a shock. As a result, the system is in the low-density phase. \[For $`\rho _+>0`$ a shock could form, but would have positive velocity under the circumstances considered here. Hence also in this case the system is in the low-density phase.\] Now we examine the seemingly trivial reason why an increase in the left boundary density leads to an increase in the bulk density. Above we simply claimed this to be true on the basis of plausibility. Here we support this claim with an argument that becomes important below. Suppose we create a little perturbation at the left boundary by injecting an extra particle (on top of those particles that are injected anyway from the reservoir). This creates a perturbation which, by definition of the collective velocity, travels with $`v_c>0`$ into the bulk and leads to a local increase of the density, moving away from the boundary and spreading out as time goes on. The collective velocity is positive, since by assumption we have a background density $`\rho _{}<\rho ^{}`$ and hence the current as a function of the density has positive slope. The point is that maintaining such a perturbation (which corresponds to increasing the left reservoir density permanently) leads to a permanent additional flow of particles into the bulk and hence to the anticipated increase of the stationary bulk density. (ii) It is now clear that this argument holds only as long as $`\rho _{}<\rho ^{}`$. Assume now $`\rho _{}=\rho ^{}`$. Following the reasoning above this results in a maximal-current bulk density $`\rho ^{}`$. However, if the left reservoir density increases beyond the maximal current density, the collective velocity becomes negative. No extra particles flow into the system, which therefore remains in its bulk at the maximal current density $`\rho ^{}`$. The system is now in the maximal current phase for all $`\rho _{}>\rho ^{}`$ and $`\rho _+<\rho ^{}`$. This transition is continuous, the bulk density approaches $`\rho ^{}`$ smoothly from below. Intuitively this phenomenon may be understood as “overfeeding” . The injected particles act as blockages for further incoming particles, leading to a back-moving traffic jam at the origin. This increased density at the origin blocks further injection attempts and prevents the actual increase of the current. Finally, using similar arguments, one can show that the transition from the high-density phase to the maximal current phase is also continuous. To summarize, the theory predicts a first-order transition along the line defined by $`j_+=j_{}`$ and $`\rho _{}<\rho ^{}`$, $`\rho _+>\rho ^{}`$ where the stationary bulk density jumps from $`\rho _{}`$ (low density phase LD) to $`\rho _+`$ (high density phase HD). On the phase transition line the stationary density is linearly increasing from $`\rho _{}`$ to $`\rho _+`$. From both phases there is a continuous phase transition to the maximal current phase MC defined by $`\rho _{}>\rho ^{}`$ and $`\rho _+<\rho ^{}`$. In this phase the bulk density takes the maximal current value $`\rho =\rho ^{}`$. These rules are encoded in an extremal principle for the current $$j=\{\begin{array}{c}\underset{\rho [\rho _R,\rho _L]}{\mathrm{max}}j(\rho )\text{ for }\rho _L>\rho _R\hfill \\ \underset{\rho [\rho _L,\rho _R]}{\mathrm{min}}j(\rho )\text{ for }\rho _L<\rho _R.\hfill \end{array}$$ (10) derived in . It is worthwhile pointing out that this dynamical theory explains in mesoscopic terms the predictions one would obtain by viewing the system from a coarse-grained hydrodynamic view-point . At the same time, verification of this scenario suggests the validity of the hydrodynamic approach to the description of the large-scale dynamics of the particle system by using the exact current-density relation. ### B Coupling to boundary reservoirs To verify this scenario which correctly describes the exactly solvable ASEP with open boundaries one has to investigate our model in terms of the effective boundary densities $`\rho _{},\rho _+`$. Given two reservoir densities there is no general recipe how to eliminate the non-universal boundary effects that result in the effective boundary densities $`\rho _{},\rho _+`$. Hence these quantities are not easy to control. Ideally, for purposes of theoretical investigation, one would like to construct an injection and absorption mechanism which leads to a constant density profile for a semi-infinite system so that $`\rho _L=\rho _{}`$ and $`\rho _R=\rho _+`$, i.e. the effective boundary densities are identical to the actual control parameters of the model. Here we choose an injection mechanism where the particles on the lattice interact with the reservoir particles in the same way as among each other. We define two injection rates from the reservoir at the left boundary: * injection at site 1 if site 2 is occupied $`|\mathrm{}A|AA`$ with rate $`\alpha _1`$ * injection at site 1 if site 2 is empty $`|\mathrm{}\mathrm{}|A\mathrm{}`$ with rate $`\alpha _2`$ and two new hopping rates at the right boundary: * hopping from site $`N1`$ to site $`N`$ $`A\mathrm{}|\mathrm{}A|`$ with rate $`\beta _1`$ * hopping out from site $`N`$ (absorption) $`A|\mathrm{}|`$ with rate $`\beta _2`$. These four hopping rates are those that would be affected by the interaction with particles in the reservoirs. The rates $`\alpha _i`$ ($`\beta _i`$ resp.) have now to be determined as functions of the left (right) reservoir density. This can be illustrated e.g. for the injection process with rate $`\alpha _1`$. We imagine the reservoir to include a site 0 of the chain. The injection rate into the first site is defined by the (stationary) average occupation $`\rho _L`$ of the imaginary site 0, but with the condition that the first site is empty and the second site is occupied. Considering the zeroth, the first and the second site as three neighboring sites of an infinite chain, this conditional probability can be expressed readily as correlations in the stationary state of an infinite chain. Thus we find $`\alpha _1=q\mathrm{\hspace{0.17em}101}/\mathrm{\hspace{0.17em}01}`$ where expectation values like $`\mathrm{\hspace{0.17em}101}=n_i(1n_{i+1})n_{i+2}`$ are calculated in the thermodynamic limit of the distribution (4) with density $`\rho =\rho _L`$. The case of $`\alpha _2`$ is entirely similar and one finds $`\alpha _2=r\mathrm{\hspace{0.17em}100}/\mathrm{\hspace{0.17em}00}`$. For the calculation of the right boundary rates we note that the probability of jumping from the site $`N1`$ to the site $`N`$ is affected by the average occupation of the imaginary reservoir site $`N+1`$. With the jump condition that site $`N1`$ is occupied and site $`N`$ is empty one finds $`\beta _1=(r\mathrm{\hspace{0.17em}100}+q\mathrm{\hspace{0.17em}101})/\mathrm{\hspace{0.17em}10}`$. The case of $`\beta _2`$ is similar but one has to take into account the conditional probability of the occupation of two imaginary reservoir sites $`N+1`$ and $`N+2`$. This gives $`\beta _2=(r\mathrm{\hspace{0.17em}100}+q\mathrm{\hspace{0.17em}101})/\mathrm{\hspace{0.17em}1}`$. One has to determine these correlations in an infinite chain with density $`\rho _R`$. These rates can be expressed as a function of the density through the form of the current and we find: $`\alpha _1`$ $`=`$ $`q{\displaystyle \frac{\mathrm{\hspace{0.17em}101}}{\mathrm{\hspace{0.17em}01}}}=q{\displaystyle \frac{\mathrm{\hspace{0.17em}10}_{\rho _L}}{1\rho _L}}`$ (11) $`\alpha _2`$ $`=`$ $`r{\displaystyle \frac{\mathrm{\hspace{0.17em}100}}{\mathrm{\hspace{0.17em}00}}}=r{\displaystyle \frac{\mathrm{\hspace{0.17em}10}_{\rho _L}}{1\rho _L}}`$ (12) $`\beta _1`$ $`=`$ $`{\displaystyle \frac{r\mathrm{\hspace{0.17em}100}+q\mathrm{\hspace{0.17em}101}}{\mathrm{\hspace{0.17em}10}}}={\displaystyle \frac{j(\rho _R)}{\mathrm{\hspace{0.17em}10}_{\rho _R}}}`$ (13) $`\beta _2`$ $`=`$ $`{\displaystyle \frac{r\mathrm{\hspace{0.17em}100}+q\mathrm{\hspace{0.17em}101}}{\mathrm{\hspace{0.17em}1}}}={\displaystyle \frac{j(\rho _R)}{\rho _R}}`$ (14) with $$\mathrm{\hspace{0.17em}10}_\rho =\mathrm{\hspace{0.17em}01}_\rho =(1\rho )\left(1\frac{j(\rho )}{r\rho }\right)$$ (15) For $`\rho _L=1`$ we use the limiting values $`\alpha _1=q`$, $`\alpha _2=r`$, and for $`\rho _R=0`$ we use $`\beta _1=\beta _2=r`$. Together with the bulk hopping rates $`r,q`$ the dynamics of the model is now completely defined. Before discussing the full phase diagram we consider the case of equal boundary densities $`\rho =\rho _R=\rho _L`$. Somewhat surprisingly, it turns out that we can actually obtain the full stationary distribution of the process: $$P^{}(\{n\})=\frac{1\rho }{\lambda _1^{N1}}\left(\frac{q}{r}\right)^{_{i=1}^{N1}n_in_{i+1}}z^{_{i=2}^{N1}n_i}(\lambda _11)^{n_1+n_N}$$ (16) with the eigenvalue $`\lambda _1`$ of the transfer matrix of the one-dimensional Ising model and the “fugacity” $`z=e^{\beta h}`$ (Appendix A). Stationarity of this distribution can be proved by writing the time evolution operator of the process in a quantum Hamiltonian formalism (see appendix B). Like in the periodic case, the exact stationary non-equilibrium distribution of the open system with equal boundary densities is the equilibrium distribution of an Ising chain of length $`N`$, but with boundary fields $`g(n_1+n_N)`$ rather than with the Ising coupling $`n_Nn_1`$. Moreover, with the transfer matrix formulation of the Ising model it is straightforward to show that the density profile is constant. The theoretical scenario described above then suggests the identification $`\rho _{}=\rho _L`$ for $`\rho _L<\rho ^{}`$ and $`\rho _+=\rho _R`$ for $`\rho _R>\rho ^{}`$. For $`\rho _L>\rho ^{}`$ or $`\rho _R<\rho ^{}`$ respectively constant profiles are not stable with respect to fluctuations and the relationship between the reservoir densities and the effective boundary densities may be more subtle (see below). ## IV Phase diagram for repulsive interaction Having defined the model and given the theoretical background we are now set to investigate the phase diagram of the system. We consider first the traffic flow scenario (repulsive interaction). ### A Mean field Unfortunately the exact stationary distribution of the open system for different boundary densities does not have a simple form. However, also the density in the open system satisfies a continuity equation of the form (5), with the boundary currents $`j_0=\alpha _1(1n_1)n_2+\alpha _2(1n_1)(1n_2)`$ (17) $`j_{N1}=\beta _1n_{N1}(1n_N)`$ (18) $`j_N=\beta _2n_N`$ (19) Using also the expression (6) for the bulk current one can obtain the stationary density profile in a mean field approximation by neglecting the correlations in the expectation values. Stationarity implies equal current everywhere in the lattice, $`j_i=j`$. This gives rise to the bulk recursion relation $$\rho _i=\frac{j}{(1\rho _{i+1})\left(q\rho _{i+2}+r(1\rho _{i+2})\right)}$$ (20) for the density profile. The boundary values of the recursion are determined by the boundary currents (17) - (19). Because of the interaction with the boundary there are two possibilities to perform a mean-field analysis. The most straightforward choice would be to choose the boundaries rates (11) - (14) and analyze the mean field phase diagram in terms of the reservoir densities $`\rho _{R,L}`$. For a given choice of current one can draw a density profile and read off the various phases. However, it turns out that within such an approximation scheme a constant density profile cannot be achieved even if $`\rho _R=\rho _L`$ and that the boundary densities are not equal to the reservoir densities. As a result, the mean field phase diagram obtained in this way differs considerably from the theoretical expectation. Within this approximation the MC phase disappears for $`r/q`$ smaller than $`0.6`$. From a theoretical point of view it is more natural to neglect the correlations even in the expressions of the boundary rates in terms of the Ising expectation values. This gives rise to simple expressions of the boundary rates in terms of the reservoir densities. Using the boundary conditions $`\rho _0=\rho _L`$ and $`\rho _{N+1}=\rho _{N+2}=\rho _R`$ extends the validity of the recursion (20) to the range $`0iN`$. In this way $`\rho _N`$ can be expressed as a function of $`\rho _R`$ and $`j`$ and this was the reason of choosing the direction of the recursion from the right to the left. It is easy to see that for a semi-infinite system there are the constant solutions $`\rho _i=\rho _L`$ and $`\rho _i=\rho _R`$ resp. with the mean field current $$j^{MF}(\rho )=\rho (1\rho )\left(q\rho +r(1\rho )\right)$$ (21) Like in the exact solution this is also a solution of the finite system with equal boundary densities. Therefore we shall discuss this mean-field approach in some more detail. For general boundary densities we did not find a solution of the recursion (20) in closed form, but it is easily solved numerically on a computer. It is sufficient to choose a density $`\rho _R`$ and a current $`j`$ and apply the recursion relation for drawing the density profile. If any of $`\rho _i`$’s is smaller than 0 or greater than 1 then there is no solution corresponding to these values of $`\rho _R`$ and $`j`$ at this system size since any physical solution has to satisfy $`0\rho _i1`$ for all $`1iN`$. For fixed $`\rho _R,\rho _L`$ and length $`N`$ this requirement fixes the current $`j`$. In this way one can map out the whole phase diagram, finding the corresponding current for all values of $`\rho _L`$ and $`\rho _R`$. For sufficiently large $`N`$ the size dependence is negligibly small. We have chosen $`r=1.0`$ and $`q=0.1`$. The low- (LD) and the high-density (HD) phases resp. are easy to distinguish as the bulk density is equal to one of the end densities. This allows us to identify $`\rho _{}=\rho _L`$ and $`\rho _+=\rho _R`$. At the other end oscillatory behavior is observed. This does not happen in the usual ASEP with open boundaries , but has been observed in an ASEP with particles covering more than one lattice site . The maximal mean field current accessible at given parameters $`r`$ and $`q`$ and the corresponding bulk density define the maximal current phase MC. The location of these phases in the parameter space differs from those obtained from theoretical scenario reviewed above (Fig. 4) by using the exact value (7) of the current. Further analysis shows that the discrepancy increases with increasing repulsion. ### B Monte Carlo results In order to check the phase diagram based on the theory of boundary induced phase transition described above we have performed Monte Carlo simulations of the model. The boundary rates are determined as explained above to simulate given left and right densities. During one MC step one of $`N+1`$ sites is chosen randomly ($`N`$ is the size of the system and the +1 is for the jumping into the system) and if there is a particle then it can hop with a probability given by the hopping rates. The initial configuration was an empty lattice. The required time to reach the stationary state for given system size can be determined by investigating the time dependence of the current and the bulk density (the density in the middle of the system). The order parameter which is the stationary value of the bulk density obtained as an average over $`10^7`$ MC steps is the best indicator of the phase transition lines. At the first order phase transition between the high and the low density phases it has a pronounced jump from the left to the right density. A system size of 1000 is sufficiently large to localize the phase transition line. The transition into the maximal current phase is continuous, only the derivative of the order parameter has a (jump) discontinuity. In order to localize this phase transition line one needs larger systems of size 5000. In a given phase the simulated bulk density depends on the boundary densities the same way as it is predicted by the theory, namely it equals to the left (right) density in the LD (HD) phase, and it is a well defined constant value in the MC phase. At the phase boundary the crossover between two kinds of behavior of the bulk density can be localized within an error shown the figures (). The Monte Carlo results are seen to be in agreement with the phase diagram derived from the theory of boundary-induced phase transitions (Fig. 5). The mean-field phase transition lines are significantly outside the numerical error bars except on part of the transition line between LD and MC phase. ## V Attractive interaction Attractive (ferromagnetic) interaction leads to bulk particle domains as in the one-dimensional Ising model. Dynamically this is qualitatively understandable since because of $`q>r`$ particles tend to form clusters. One expects a shift of the maximal-current density $`\rho ^{}`$ to higher densities. This can indeed be shown by calculating the derivative of the exact current (7). The full current-density relation for $`q=1`$, $`r=0.1`$ is shown in Fig. 6. Analysis of the model with open boundaries proceeds in the same way as in the repulsive case. With $`r=0.1`$ and $`q=1.0`$ the mean field approximation yields the three phases discussed above, with the phase transition lines differing substantially from those expected theoretically if one identifies $`\rho _{}=\rho _L`$ and $`\rho _+=\rho _R`$ as suggested by the density profiles in the low and high density phases respectively. As additional feature the mean field theory predicts a further phase transition located at the large $`\rho _L`$ and small $`\rho _R`$ corner of the phase diagram (Fig. 7). In this “fourth phase” the bulk density is larger than the density corresponding to the maximal current. The area of this phase increases with increasing attraction, and the area of MC phase decreases but does not disappear. The phase diagram obtained from Monte-Carlo simulation has the phase transitions expected from the theoretical prediction, but also shows that the fourth phase exists in a small neighborhood of the ($`\rho _+=0`$, $`\rho _{}>\rho ^{}`$) line and hence is not an artefact of the mean-field approximation. The location of the phase transition lines is rather different compared to mean field (Fig. 8). The transition to this phase from the LD phase is first order and from the maximal current phase it is second order. In this sense the fourth phase is like the high-density phase, however, the bulk density differs from both the boundary densities $`\rho _{R,L}`$. Within the theory of boundary-induced phase transitions this high density phase can be understood by recalling the non-universal relationship between real boundary density and effective boundary density (in terms of which the theory is formulated). While for repulsive interaction and the choice of coupling mechanism used here these two quantities may be identified, the connection is apparently more complicated for attractive interaction if the (real) boundary densities become sufficiently high (left boundary) and low (right boundary) respectively. An adequate explanation of this boundary phenomenon which leads to a reentrance transition into the high-density phase is not available. In order to rule out the possibility of a finite-size effect we performed simulations with different system sizes. Using N=500, 1000 and 5000 resp. we found no indication that the reentrance transition would disappear for large enough systems. ## VI Conclusions Our analysis of the totally asymmetric exclusion model with next-nearest-neighbor interaction consists of two parts. First we considered the periodic system in order study the bulk properties of the steady state. With a view on traffic flow modelling we remark that the observations (i) - (iv) listed above are consistent with the results obtained here. The current-density relation becomes asymmetric (Fig. 1) in a way which is closer to real traffic data as the symmetric relation $`j=\rho (1\rho )`$ for the ASEP with $`r=q=1`$. There is no discontinuity in the derivative at the maximal-current density $`\rho ^{}`$, in agreement with the arguments given above. As more far-reaching conclusion we note that the exact result (4) sheds light on the so far somewhat unclear relationship between the updating mechanism and the occurrence of correlations. In order to obtain correlations in traffic flow models it is not essential to use a discrete-time updating mechanisms. Since other discrete-time models have uncorrelated stationary states it appears that correlations have their physical origin in speed-reduction rather than in the nature of the updating scheme. In the second part we investigated the steady-state selection in the open system. It turns out that the theoretical scenario based on the interplay of shocks and overfeeding correctly describes the phase diagram in terms of effective boundary densities. For repulsive interactions this strengthens the case for a maximal current phase in real traffic. For attractive interaction there is, however, a surprising reentrance phase transition to the high density phase which originates in the so far poorly understood relationship between actual and effective boundary densities. Apparently these two are not always monotic functions of each other as one would naively expect. For a deeper understanding the next step to be done is the analysis of the density profiles close the reentrance phase transition lines in order analyze whether universal properties of the usual transition lines can be observed or not. T.A. thanks for partial support by the Hungarian Academy of Sciences (Grant OTKA T 029792). G.M.S. thanks K. Klauck, V. Popkov, L. Santen and A. Schadschneider for illuminating discussions on traffic flow. ## A Transfer matrix for the one-dimensional Ising model The energy of the one-dimensional Ising model with the classical spin variables $`s_i=\pm 1`$ can be written in term of the variables $`n_i=(1s_i)/2`$ in the form $`E=J_in_in_{i+1}`$. In this interpretation the Ising model is a classical lattice gas of $`M=_in_i`$ hard-core particles with nearest neighbor interaction. Using standard techniques one can express the grand canonical partition function $`Z=_{config}z^Me^{\beta E}`$ at inverse temperature $`\beta =1/kT`$ in terms of the eigenvalues of the transfer matrix $$T=\left(\begin{array}{cc}1& \sqrt{z}\\ \sqrt{z}& ze^\beta \end{array}\right)$$ (A.1) where, for our model, $`e^{\beta J}=q/r`$ and we define $`z=e^{\beta h}`$. In the spin interpretation of the model $`h`$ plays the role of a magnetic field. The eigenvalues of the transfer matrix are $$\lambda _{1,2}=\frac{1}{2}(1+\frac{q}{r}z)\pm \sqrt{\frac{1}{4}(1+\frac{q}{r}z)^2+z(1\frac{q}{r})}$$ (A.2) where we choose $`\lambda _1`$ to the positive sign. For a system with $`N`$ sites and periodic boundaries one has $`Z=\text{Tr}T^N=\lambda _1^N+\lambda _2^N`$. The equilibrium distribution of the Ising model is the stationary distribution of the particle hopping model. This does not mean that the particle hopping model reaches thermal equilibrium at long times, since the stationary distribution does not satisfy detailed balance with respect to the dynamics of the model. The non-equilibrium nature of the steady state results in a non-vanishing stationary particle current. We stress that completely different dynamical models may have the same stationary distribution, see e.g. the list of models given in . Another non-equilibrium example with the same Ising distribution is the discrete-time ASEP with parallel update . Even though the distribution is the same and hence all particle correlations are the same, the stationary current of this model (which is defined via the dynamics by the continuity equation) is different from the current (6) in our model. In order to calculate expectation values in the transfer matrix formalism we define the diagonal matrix $$n=\left(\begin{array}{cc}0& 0\\ 0& 1\end{array}\right).$$ (A.3) The density is then given by $`\rho =\text{Tr}(nT^N)/Z_N`$ and a two-point correlation function is given by $`n_kn_l=\text{Tr}(nT^{kl}nT^{Nk+l})`$. Higher order correlators are calculated analogously. By diagonalizing $`T`$ one obtains the expression $$\rho =\frac{1\lambda _1}{\lambda _2\lambda _1}.$$ (A.4) for the particle density in terms of the eigenvalues of $`T`$. This relation may be used to express the current (6) as a function of the particle density. In the thermodynamic limit $`N\mathrm{}`$ one obtains (7). We note that expectation values for the distribution (16) are calculated with the same transfer matrix as in the periodic case. However, instead of taking a trace, one calculates a scalar product with suitably chosen vectors which are determined by the boundary fields. ## B Proof of stationarity for equal boundary densities The proof of stationarity of the distribution (4) for the periodic system is given in Ref. . An alternative, constructive proof can be obtained by using translational invariance and taking the distance between neighboring particles as stochastic variables. In this way the particle hopping process turns into a zero-range process for which the stationary distribution is known. Reexpressing the stationary zero-range distribution in terms of particle occupation numbers yields (4). Here we prove stationarity of the Ising distribution (16) for the open system coupled to reservoirs of equal density. We use a convenient standard approach for the stochastic description of classical interacting particle systems, known as “quantum Hamiltonian formalism” . The basic idea is to formulate the generator of the Markov process in terms of a many-body quantum operator. For processes without exclusion one obtains in this way a Fock space representation of the generator in terms of bosonic creation and annilation operators . For exclusion processes with at most one particle per site the same strategy yields an operator expressed in terms of Pauli-spin matrices . We define the exclusion process with state space $`X=\{0,1\}^N`$ and transition rates $`w_{\underset{¯}{n}\underset{¯}{n}^{}}`$ from state $`\underset{¯}{n}`$ to $`\underset{¯}{n}^{}`$ in terms of a master equation $$\frac{d}{dt}P(\underset{¯}{n};t)=\underset{\underset{¯}{n}^{}X}{}\left[w_{\underset{¯}{n}^{}\underset{¯}{n}}P(\underset{¯}{n}^{};t)w_{\underset{¯}{n}\underset{¯}{n}^{}}P(\underset{¯}{n};t)\right]$$ (B.1) for the probability $`P(\underset{¯}{n};t)`$ of finding, at time $`t`$, a configuration $`\underset{¯}{n}`$ of particles on a lattice of $`N`$ sites. Here $`\underset{¯}{n}=\{n_1,n_2,\mathrm{},n_N\}`$ where $`n_i=0,1`$ are the integer-valued particle occupation numbers at site $`i`$. The master equation is a linear first order DGL in the time-variable and therefore it is natural to write it in a vector notation with the probabilities $`P(\underset{¯}{n};t)`$ as vector components. We represent each of the possible particle configurations $`\underset{¯}{n}`$ by a column vector $`|\underset{¯}{n}`$ which form a basis of a vector space $`𝐗=(\text{ }\mathrm{C}^2)^N`$. The transposed vectors $`\underset{¯}{n}|`$ form a basis of the dual space and we define the usual scalar product $`\underset{¯}{n}|\underset{¯}{n}^{}=\delta _{\underset{¯}{n},\underset{¯}{n}^{}}`$. The probability distribution is now represented by a state vector $`|P(t)=_{\underset{¯}{n}X}P(\underset{¯}{n};t)|\underset{¯}{n}`$ and one can write the master equation in the form $$\frac{d}{dt}P(\underset{¯}{n};t)=\underset{¯}{n}|H|P(t)$$ (B.2) where the off-diagonal matrix elements of $`H`$ are the (negative) transition rates between states and the diagonal entries are the inverse of the exponentially distributed life times of the states. In formal analogy to the quantum mechanical Schrödinger equation we shall refer to $`H`$ as quantum Hamiltonian. A state at time $`t^{}=t_0+t`$ is given in terms of an initial state at time $`t_0`$ by $$|P(t_0+t)=\text{e}^{Ht}|P(t_0).$$ (B.3) We stress that the physicists notion “quantum Hamiltonian” for the matrix $`H`$ is somewhat misleading in so far as $`H`$ is, in fact, the generator of the Markov semigroup of the process, rather than the Hamiltonian of an actual quantum system. This by now well-established notion has its origin in the fact that for various stochastic processes the generator $`H`$ is identical to the quantum Hamiltonian of some well-known spin system. In this context we would also like to point out that quantum mechanical expectation values $`A\mathrm{\Psi }|A|\mathrm{\Psi }`$ for an observable $`A`$ are calculated in a different way than probabilistic expectation values for a function $`F(\underset{¯}{n})`$ of the stochastic variables $`\eta `$. In the quantum Hamiltonian formalism one writes $`F_{\underset{¯}{n}X}F(\underset{¯}{n})P(\underset{¯}{n};t)=s|F|P(t)`$ with the matrix $`F=_{\underset{¯}{n}X}F(\underset{¯}{n})|\underset{¯}{n}\underset{¯}{n}|`$ and the summation vector $`s|=_{\underset{¯}{n}X}\underset{¯}{n}|`$ which performs the average over all possible final states of the stochastic time evolution. For our considerations the expectation value $`\rho _k(t)=s|n_k|P(t)`$ for the density at site $`k`$ is of special interest. It is given by the projection operator $`n_k`$ which has value 1 if there is a particle at site $`k`$ and 0 otherwise. $`m`$-point density correlations are then given by the expression $`s|n_{k_1}\mathrm{}n_{k_m}|P(t)`$. In this paper we study only stationary expectation values. For the formal description of a stationary probability distribution we use the transposed summation vector $`|s=_{\underset{¯}{n}X}|\underset{¯}{n}`$. A general stationary measure $`P^{}(\underset{¯}{n})`$ may then be written in vector notation in the form $`|P^{}=e^{\beta E(\underset{¯}{n})}|s/Z_N`$ with the configuration-dependent “energy” matrix $`E(\underset{¯}{n})`$ and the “partition function” $`Z_N=s|e^{\beta E(\underset{¯}{n})}|s`$ which normalizes the measure $`|\stackrel{~}{P}^{}=e^{\beta E(\underset{¯}{n})}|s`$. Notice that in vector notation the expression $`E(\underset{¯}{n})`$ represents a diagonal matrix with the energies as diagonal elements. Below we shall also use the invertible diagonal matrix $`P^{}=e^{\beta E(\underset{¯}{n})}`$ which has the (unnormalized) stationary probabilities $`P^{}(\underset{¯}{n})`$ as diagonal elements. To obtain the quantum Hamiltonian for the time evolution of the interacting ASEP with open boundaries we note that one can represent any two-state particle system as a spin system by identifying a particle (vacancy) on site $`k`$ with a spin-down (up) state on this site. This allows for a representation of $`H`$ in terms of Pauli matrices where $`n_k=(1\sigma _k^z)/2`$ projects on states with a particle on site $`k`$ and $`v_k=1n_k`$ is the projector on vacancies. The off-diagonal matrices $`s_k^\pm =(\sigma _k^x\pm i\sigma _k^y)/2`$ create ($`s_k^{}`$) and annihilate ($`s_k^+`$) particles. We stress that in the present context the “spins” are just convenient labels for particle occupancies. Using this pseudospin formalism one finds $`H`$ $`=`$ $`{\displaystyle \underset{k=1}{\overset{N2}{}}}(n_kv_{k+1}s_k^+s_{k+1}^{})(rn_{k+2}+qv_{k+2})`$ (B.6) $`+(1n_1s_1^{})(\alpha _1n_2+\alpha _2v_2)`$ $`+\beta _1(n_{N1}v_Ns_{N1}^+s_N^{})+\beta _2(n_Ns_N^+).`$ Within this formalism proof of stationarity is now a straightforward calculation, using simple expressions for the jumping rates at the ends: $`\alpha _1`$ $`=`$ $`{\displaystyle \frac{qz}{\lambda _1(\lambda _11)}}`$ (B.7) $`\alpha _2`$ $`=`$ $`{\displaystyle \frac{rz}{\lambda _1(\lambda _11)}}`$ (B.8) $`\beta _1`$ $`=`$ $`{\displaystyle \frac{qz}{\lambda _11}}`$ (B.9) $`\beta _2`$ $`=`$ $`{\displaystyle \frac{qz}{\lambda _1}}.`$ (B.10) and assuming that the boundary densities are equal at the two end. According to (B.3) stationarity is equivalent to the eigenvector relation $`H|P^{}=0`$ which in turn is equivalent to $$(P^{})^1H|\stackrel{~}{P}^{}=(P^{})^1HP^{}|s=0.$$ (B.11) The diagonal similarity transformation of $`H`$ with $`P^{}`$ leads to a sum of transition matrices which act non-trivially on at most four neighboring sites. To calculate the action of the non-diagonal parts on $`|s`$ we use $`s_k^+|s=v_k|s`$ and $`s_k^{}|s=n_k|s`$. This leaves only diagonal terms the sum of which vanishes identically. This proves stationarity of the measure (16).
warning/0001/cond-mat0001146.html
ar5iv
text
# Untitled Document Probing quantum statistical mechanics with Bose gases: Non-trivial order parameter topology from a Bose-Einstein quench J.R. Anglin<sup>a,b</sup> and W.H. Zurek<sup>b</sup> <sup>a</sup>Institut für Theoretische Physik, Universität Innsbruck, Austria <sup>b</sup>T-6, MS B288, Los Alamos National Laboratory, Los Alamos, New Mexico 87545 A rapid second order phase transition can have a significant probability of producing a metastable state instead of the equilibrium state. We consider the case of rapid Bose-Einstein condensation in a toroidal trap resulting in a spontaneous superfluid current, and compare the phenomenological time-dependent Ginzburg-Landau theory with quantum kinetic theory. A simple model suggests the effect should be observable. 1. THE ORDER PARAMETER AND THE EFFECTIVE POTENTIAL A qualitative understanding of a second order phase transition may be had by considering the system to be described by a two-component order parameter, consisting of a modulus $`R`$ and an angle $`\theta `$. The order parameter is taken to behave as a particle in a two-dimensional effective potential $`V(R)`$, with some form of dissipation dragging the system towards the bottom of this potential (possibly opposed by some random noise). Figure 1: Effective potential theory of a second order transition. In the disordered phase of the system, the minimum of $`V(R)`$ is at $`R=0`$, so that $`\theta `$ is indefinite (or purely random) (Fig. 1, left picture). The phase transition then occurs through a modification of the effective potential, such that $`R=0`$ becomes a local maximum, and the new minimum is a circular trough at some finite radius. The system rolls down the central hill to the trough, (Fig. 1, right picture) and $`\theta `$ acquires a definite value (with only small fluctuations due to noise). Which definite value $`\theta `$ acquires, however, is random, determined by small fluctuations while the system is near $`R=0`$. One can also allow $`R,\theta `$ to be functions of position (with $`V(R)`$ assumed to have the same form everywhere). In the ordered phase, the definite angle $`\theta (x)`$ may then vary spatially. A qualitative question thus arises, namely the topology of $`\theta (x)`$. Suppose we let $`x`$ lie on a circle, for example: what winding number $`𝑑x_x\theta /(2\pi )`$ should we typically observe after a transition? If the transition is slow, $`\theta (x)`$ will have time to organize itself all around the circle into some lowest energy configuration (such as $`\theta (x)`$ constant); but if it is fast, the random choice of $`\theta `$ around the circle will have some finite correlation length, and non-zero winding number will sometimes result. If it does, it will tend to be stable, since there is a high effective potential barrier for $`R=0`$, and without making $`R(x)0`$ somewhere, it is impossible to alter the winding number. See Fig. 2. Figure 2: Winding number two; distance of ribbon from torus represents $`R(x)`$, angle of winding around torus represents $`\theta (x)`$. The emergence of non-trivial topology from a rapid ( quench-like ) phase transition is common to a wide range of physical systems, real and postulated, from vortex loops in superfluid helium to cosmic strings in the early universe. With the advent of dilute alkali gas Bose condensates, we can now investigate this basic statistical mechanical problem in a new arena. The condensate mean field wave function $`\psi =Re^{i\theta }`$ provides our order parameter; and the current experimental technique of evaporative cooling automatically provides a rapid phase transition (on the Boltzmann scattering time scale). Our example of $`x`$ lying on a circle can realistically be achieved with a toroidal trap, sufficiently tight for the system to be approximately one-dimensional. (We will consider such a trap of circumference $`100\mu `$m, which we understand is a credible prospect for the relatively near future.) In this context, non-zero winding number implies circulation around the trap, since $`\theta `$ is the superfluid velocity. (Angular momentum conservation is an issue we will discuss briefly below, only noting for now that the condensate is not isolated.) 2. TIME-DEPENDENT GINZBURG-LANDAU THEORY To proceed to a quantitative description of our subject, we first consider the phenomenological theory obtained by letting the time evolution of $`\psi `$ be governed by a first order equation involving a potential of the Ginzburg-Landau form we have sketched above: $$\tau _0\dot{\psi }=\beta \left(\frac{\mathrm{}^2}{2M}^2+\mu \mathrm{\Lambda }|\psi |^2\right)\psi ,$$ (1) where $`\beta =(k_BT)^1`$, and $`\tau _0`$ and $`\mathrm{\Lambda }>0`$ are phenomenological parameters. The thermodynamical variable $`\mu `$ behaves near the critical point, in the case we consider, as $$\mu =\frac{3}{2}(T_cT)+𝒪(T_cT)^2,$$ (2) where $`T_c`$ is the critical temperature. The equilibration time for long wavelengths is $`\tau =\tau _0k_BT/|\mu |`$. The system’s disordered phase is described by $`\mu <0`$; the ordered phase appears when $`\mu >0`$. Note that this time-dependent Ginzburg-Landau (TDGL) theory also typically assumes some small stochastic forces acting on $`\psi `$; we will leave these implicit. A quench occurs if $`\mu `$ changes with time from negative to positive values. The divergence of the equilibration time $`\tau `$ at the critical point $`\mu =0`$ is associated with critical slowing down. Because of this critical slowing down, $`\frac{d\mu }{dt}/\mu `$ must exceed $`1/\tau `$ in some neighbourhood of the critical point, and so there must be an epoch in which the system is out of equilibrium. What are at the beginning of this epoch mere fluctuations in the disordered phase, in which higher energy modes happen momentarily to be more populated than the lowest mode, can thus pass unsuppressed by equilibration into the ordered phase, to become topologically non-trivial configurations of $`\psi (x)`$. The interval within which equilibration is negligible can be identified as the period wherein $`|t|/\tau <1`$. If we define the quench time scale $`\tau _Q`$ by letting $`\beta \mu =t/\tau _Q`$ (choosing $`t=0`$ as the moment the system crosses the critical point), this implies that the crucial interval is $`\widehat{t}<t<\widehat{t}`$, for $`\widehat{t}=\sqrt{\tau _Q\tau _0}`$. The correlation length $`\widehat{\xi }`$ for fluctuations at time $`t=\widehat{t}`$ is then given by $`\mathrm{}/(2M\widehat{\xi }^2)=\mu (\widehat{t})`$, which (assuming $`T(\widehat{t})T_c`$) implies that $`\widehat{\xi }=\lambda _{T_c}(\tau _Q/\tau _0)^{1/4}`$, for $`\lambda _T=\mathrm{}(2Mk_BT)^{\frac{1}{2}}`$ the thermal de Broglie wavelength. Assuming one independently chosen phase within each correlation length $`\widehat{\xi }`$, then modeling the phase distribution around the torus as a random walk suggests that the net winding number should be of order $`\sqrt{L/\widehat{\xi }}`$ for $`L`$ the trap circumference. Since evaporative cooling may be expected to yield $`(\tau _Q/\tau _0)^{1/4}`$ of order one, and $`T_c`$ is several hundred nK, we can estimate $`\widehat{\xi }100`$ nm, leading us to expect winding numbers in our $`100\mu `$m torus of order ten. At current experimental densities, this implies a current approaching the Landau critical velocity, in the range of mm/s. Considering this intriguing possibility raises an obvious question: is TDGL actually relevant to finite samples of dilute gas, far from equilibrium? 3. QUANTUM KINETIC THEORY Because the condensates now available are dilute enough to be weakly interacting, there are good prospects for answering this question theoretically, by constructing from first principles a quantum kinetic theory to describe the whole process accurately. Using a second-quantized description of the trapped gas, one treats all modes above some judiciously chosen energy level as a reservoir of particles, coupled to the lower modes by two-particle scattering. Tracing over the reservoir modes leads to a master equation for the low energy modes, very similar to the equation for a multimode laser. Scattering of particles from the reservoir modes into the low modes, and vice versa, provides gain and loss terms. If the reservoir is described at all times by a grand canonical ensemble (of time-dependent temperature and chemical potential, in general), then the gain and loss processes are related by a type of fluctuation-dissipation relation, in which the reservoir chemical potential and the repulsive self-interaction of the gas combine in precisely the Ginzburg-Landau form. As a result, as long as these gain and loss processes are the only significant scattering channels, and the condensate self-Hamiltonian can be linearized to a good approximation, the system is indeed described quite well by time-dependent Ginzburg-Landau theory. Since these two conditions can be shown to hold near the critical point, all the results of section 2 can be recovered from quantum kinetic theory. We can also clarify how the onset of a runaway process like Bose-Einstein condensation by evaporation can be consistent with critical slowing down: though the particle numbers in the lowest modes of the trapped gas are growing very fast, the required particle numbers to be in equilibrium with the higher energy modes grow much faster still as the temperature passes through $`T_c`$. So even with Bose-enhanced scattering at ever increasing rates, the rate at which the system is able to maintain equilibrium actually decreases. The problem is that circulating states only become metastable above a threshold density, precisely because nonlinearity is required. So understanding the probability for a circulating condensate to grow into metastability requires extending quantum kinetic theory into the fully nonlinear regime, and without begging the basic question by assuming from the start that some particular mode will end up with the condensate. 3.1. A two-mode model As a first step towards this goal, we consider a toy model of just two competing modes, such as the torus modes of different winding numbers $`k_0`$, $`k_1`$. The self-Hamiltonian of this system is taken as $`\widehat{H}=E[\widehat{n}_1+{\displaystyle \frac{1}{2N_c}}(\widehat{n}_0^2+\widehat{n}_1^2+4\widehat{n}_1\widehat{n}_0)].`$ (3) Since the self-Hamiltonian must conserve both particle number and angular momentum, it can only be a function of $`\widehat{n}_0`$ and $`\widehat{n}_1`$ (which greatly simplifies the derivation of the master equation). Because we have incorporated the Bose enhancement of inter-mode repulsion (the factor of 4 instead of 2 in front of the $`\widehat{n}_0\widehat{n}_1`$ term, which is of course the best case value, obtained when the two orbitals overlap completely), we make the state with all particles in the 1 mode a local minimum of the energy for fixed $`n_0+n_1>(N_c+1)`$. We have thus obtained an simple model which exhibits metastability above a threshold. To make our results more meaningful, we estimate the experimental ranges of our parameters $`E`$ and $`N_c`$. For our toroidal trap $`\beta E`$ would be on the order of $`10^4`$ at current experimental temperatures, and proportional to $`k_1^2k_0^2`$. $`N_c`$ would be of order one for $`(k_0,k_1)=(0,1)`$, but higher for higher modes (since $`E/N_c`$ is actually constant). Figure 3: Trajectories from QKT (solid) and TDGL (dotted); heavy dashed line is threshold for metastability of mode 1. Initial times are $`\widehat{t}`$; quench is $`\beta \mu =\mathrm{tanh}(t/\tau _Q)`$, $`\beta =\beta _ce^{\mathrm{tanh}(t/\tau _Q)}`$. Parameters are $`N_c=100`$, and (a) $`\mathrm{\Gamma }\tau _Q=10`$, $`\beta _cE=0.01`$; (b) $`\mathrm{\Gamma }\tau _Q=100`$, $`\beta _cE=0.05`$. All trajectories shown initially have $`n_1>n_0`$. Tracing out a reservoir of higher modes as described above leads to a master equation for the two modes 0 and 1, which even for this highly simplified model is only directly tractable numerically. Here we will simply consider it as a kind of Fokker-Plank equation for the particle numbers $`n_0,n_1`$, and by neglecting the diffusion terms (valid for fast quenches after the early stage, which can be solved separately by assuming linearity), we extract an equation of motion for $`n_0`$ and $`n_1`$ which we can compare to TDGL: $`\dot{n}_0`$ $`=`$ $`\mathrm{\Gamma }n_0\left[e^{\beta \mu }e^{\frac{\beta E}{N_c}(n_0+2n_1)}+2\beta En_1e^{\frac{\beta E}{2N_c}|N_c+n_0n_1|}\mathrm{sinh}{\displaystyle \frac{\beta E}{2N_c}}(N_c+n_0n_1)\right]`$ $`\dot{n}_1`$ $`=`$ $`\mathrm{\Gamma }n_1\left[e^{\beta \mu }e^{\frac{\beta E}{N_c}(N_c+n_1+2n_0)}2\beta En_0e^{\frac{\beta E}{2N_c}|N_c+n_0n_1|}\mathrm{sinh}{\displaystyle \frac{\beta E}{2N_c}}(N_c+n_0n_1)\right].`$ When $`n_0+2n_1`$ and $`N_c+n_1+2n_0`$ are both close to $`N_c\mu /E`$, or for low enough particle numbers, we may replace $`n_j|\psi _j|^2`$ in the first two terms in each equation of (S0.Ex1) to obtain a TDGL equation, in the sense that $`\dot{\psi }_j`$ is set equal to the variation of a Ginzburg-Landau effective potential with respect to $`\psi _j^{}`$. But the last term in each equation is not of Ginzburg-Landau form (it does not even involve $`\mu `$). These non-GL terms conserve $`n_0+n_1`$, and so are neither gain nor loss but another scattering channel hitherto neglected: collisions with reservoir particles shifting particles between our two low modes. These processes are insignificant at low particle numbers, but grow much stronger as the condensate grows larger, because they are doubly Bose-enhanced. And they turn out to have a simple but potentially drastic effect: they allow they system to equilibrate in energy faster than in particle number. Representative solutions to (S0.Ex1) are shown in Fig. 3, together with the $`|\psi _j|^2`$ given by the TDGL theory formed by keeping only the first two terms of each equation in (S0.Ex1), expanding the exponentials to first order. It is clear that for sufficiently fast quenches the two theories accord quite well, but that for slower quenches TDGL significantly overestimates the probability of reaching the metastable state, because more rapid equilibration in energy than particle number strongly favours the lowest mode. This is the main new feature we have identified in dilute Bose gases as a GL-like system. 3.2. Angular momentum Finally, we return to the question of angular momentum. Spontaneous appearance of a circulating state, condensing out of a cold atomic cloud with no net rotation, obviously implies angular momentum concentration in the condensate. Compensating ejection of fast atoms with opposite angular momentum during the evaporation process, in which after all most of the initial atoms are expelled, is a plausible mechanism by which this can proceed. Our assumption that the non-condensate atoms remain in a grand canonical ensemble relies implicitly on some such process. In our two-mode model, constraining the reservoir’s mean angular momentum to be always opposite to that of the condensate will effectively just raise $`E`$ and $`N_c`$. Angular momentum transport during evaporative cooling has as yet received no detailed study, however. 4. CONCLUSIONS Despite the shortcomings of TDGL revealed by our toy model, we emphasize that kinetic theory does show that TDGL is indeed relevant to trapped dilute gases: what TDGL requires is not outright rejection, but corrections. And although the corrections may be substantial, the qualitative features predicted by TDGL are still recovered. While the extension of quantum kinetic theory beyond toy models, to realistic descriptions of spontaneous currents, will obviously require further study, the prospects for experimental realization of spontaneous currents are quite encouraging. Athough in this brief treatment we have neglected noise, diffusive nucleation actually gives a lower bound of typical winding number one for condensation in our hundred micron toroidal trap; with $`10^6`$ atoms, even this has an equilibrium probability of order $`e^{10}`$. REFERENCES * V.L. Ginzburg and L.D. Landau, Zh. Eksperim. i. Teor. Fiz. 20, 1064 (1950) . * W.H. Zurek, Nature 317, 505 (1985); Acta Physica Polonica B 24, 1301 (1993); Phys. Rep. 276, 177 (1996). * P.C. Hendry et al., Nature 368, 315 (1994); V.M.H. Ruutu et al., Nature 382, 332 (1996); C. Baüerle et al., Nature 382, 334 (1996). * T.W.B. Kibble, J. Phys. A 9, 1387 (1976). * M. Anderson et al., Science 269, 198 (1995); K.B. Davis et al., Phys. Rev. Lett. 75, 3969 (1995); C.C. Bradley et al., Phys. Rev. Lett. 75, 1687 (1995). * C.W. Gardiner and P. Zoller, Phys. Rev. A55, 2901 (1997); cond-mat/9712002; D. Jaksch, C.W. Gardiner, and P. Zoller, Phys. Rev. A56, 575 (1997); J.R. Anglin, Phys. Rev. Lett. 79, 6 (1997).
warning/0001/gr-qc0001079.html
ar5iv
text
# Generalised hyperbolicity in space-times with conical singularities ## 1. Introduction A desirable property of any space-time used to model a physically plausible scenario is that the evolution of the Einstein’s equations is well posed i.e. the initial value problem has a unique solution. Space-times whose metrics are at least $`C^2`$, which guarantees the existence of unique geodesics, fall within the context of the Cosmic Censorship Hypothesis of Penrose (1979). The hypothesis states that that the space-time will be globally hyperbolic, i.e. strong causality is satisfied and $`J^+(p)J^{}(q)`$ is compact $`p,qM`$, and hence the evolution of Einstein’s equations is well defined. There are however a number of space-times with weak singularities which model physically plausible scenarios such as thin cosmic strings (Vickers, 1987), impulsive gravitational waves (Penrose, 1972) and dust caustic space-times (Clarke and O’Donnell, 1992). Typically such a space-time has a locally bounded metric whose differentiability level is lower than $`C^2`$, but whose curvature is well defined as a distribution, often with its support on a proper submanifold. Although cosmic censorship may be violated for such space-times, it does not rule out the possibility that the evolution of some fields is well posed. A concept of hyperbolicity for such space-times was proposed by Clarke (1998). This was based on the extent to which singularities disrupted the local evolution of the initial value problem for the scalar wave equation. $`\mathrm{}\varphi `$ $`=f`$ $`\varphi _{|S}`$ $`=\varphi _0`$ $`n^a\varphi _{,a}^{}{}_{|S}{}^{}`$ $`=\varphi _1`$ Clarke reformulated the initial value problem, on an open region $`\mathrm{\Omega }`$ with a compact closure admitting a space-like hypersurface $`S`$, which partitions $`\mathrm{\Omega }`$ into two disjoint sets $`\mathrm{\Omega }^+`$ and $`\mathrm{\Omega }^{}`$, in a distributional form, obtained by multiplying a test field $`\omega `$ and integrating by parts once to give $`{\displaystyle _{\mathrm{\Omega }^+}}\varphi _{,a}\omega _{,b}g^{ab}(g)^{1/2}d^4x={\displaystyle _{\mathrm{\Omega }^+}}f\omega (g)^{1/2}d^4x{\displaystyle _S}\varphi _1\omega 𝑑S\omega 𝒟(\mathrm{\Omega })`$ $`\varphi _{|S}=\varphi _0`$ and then defined a point $`pM`$ as being $`\mathrm{}`$-regular if it admitted such a neighbourhood $`\mathrm{\Omega }`$ for which the above equation had a unique solution for each set of Cauchy data $`(\varphi _0,\varphi _1)H^1(S)\times H^0(S)`$. A space-time which was $`\mathrm{}`$-regular everywhere could then be said to be $`\mathrm{}`$-globally hyperbolic. It was shown if a space-time satisfied the following curve integrability conditions at a given point $`pM`$, then that point was $`\mathrm{}`$-regular. The proof involved the construction of a congruence of time-like geodesics, whose tangent admitted an essentially bounded weak derivative, and a suitable energy inequality from which uniqueness and existence could be deduced. In particular it was shown that these results were applicable to the dust caustic space-times. Not all space-times with weak singularities satisfy these integrability conditions. One such example is a space-time with a conical singularity representing a thin string. Such a space-time may be constructed by cutting out a sector of angle $`2\pi (1A)`$ from Minkowski space-time and identifying the resulting edges. In a suitable Cartesian coordinate system $`(t,x,y,z)^4`$ the metric $`g_{ab}`$ may be written as the following line element $$ds^2=dt^2+dx^2+dy^2+dz^2\frac{(1A)^2}{x^2+y^2}(xdyydx)^2$$ $`1`$ and the non-zero energy-momentum tensor components may be shown to be the well defined distribution (Balasin and Nachbagauer, 1993; Clarke et al., 1996) $$T^t{}_{t}{}^{}(g)_{}^{1/2}=T^z{}_{z}{}^{}(g)_{}^{1/2}=2\pi (1A)\delta (x)\delta (y)$$ $`2`$ which justifies the interpretation of this space-time as representing a thin cosmic string. Curve integrability is violated at the axis because $`g_{ab,c}`$ is not locally square integrable there, however there does exist a natural congruence of time-like geodesics; namely the integral curves of the vector $`/t`$ whose covariant derivative vanishes. In this paper it will be shown that $`\mathrm{}`$-regularity is also applicable to such space-times. ### A remark on notation The lower case indices $`a`$, $`b`$ etc. denote coordinates $`t`$, $`x`$, $`y`$ and $`z`$ where as the upper case indices $`A`$, $`B`$ etc. denote coordinates $`x`$ and $`y`$. The Sobolev spaces $`H^k(\mathrm{\Omega })`$, where $`\mathrm{\Omega }`$ is an open subset of $`^n`$, are defined as the spaces of functions $`fL^2(\mathrm{\Omega })`$ such that all partial derivatives up to order $`k`$ are defined as distributions and such that the Sobolev norm $$f_\mathrm{\Omega }^k=\left(\underset{\alpha _1+\mathrm{}+\alpha _n=0}{\overset{k}{}}_\mathrm{\Omega }\left|\frac{^{\alpha _1+\mathrm{}+\alpha _n}f}{(x^1)^{\alpha _1}\mathrm{}(x^n)^{\alpha ^n}}\right|^2d^nx\right)^{1/2}$$ is finite. For an $`m`$-dimensional subspace $`S\mathrm{\Omega }`$ we may also define the embedding norm $$f_{S\mathrm{\Omega }}^k=\left(\underset{\alpha _1+\mathrm{}+\alpha _n=0}{\overset{k}{}}_S\left|\frac{^{\alpha _1+\mathrm{}+\alpha _k}f}{(x^1)^{\alpha _1}\mathrm{}(x^n)^{\alpha ^n}}\right|^2d^mx\right)^{1/2}$$ It should be noted that derivatives in all tangential directions of $`\mathrm{\Omega }`$ are summed over, not only those tangential to $`S`$. Clearly $`f_S^kf_{S\mathrm{\Omega }}^k`$. ## 2. The initial value problem In order to prove hyperbolicity, we must show that in each open region $`\mathrm{\Omega }`$ with a compact boundary $`\mathrm{\Omega }`$, that a unique solution $`\varphi H^1(\mathrm{\Omega }^+)`$ exists to the initial value problem; $`\mathrm{}\varphi `$ $`=f`$ $`3`$ $`\varphi _{|S}`$ $`=\varphi _0`$ $`n^a\varphi _{,a}^{}{}_{|S}{}^{}`$ $`=\varphi _1`$ where $`fH^0(\mathrm{\Omega })`$, $`\varphi _0H^1(S)`$, $`\varphi _1H^1(S)`$. We shall assume that without loss of generality that $`\mathrm{\Omega }`$ is generated by a foliation of space-like hypersurfaces $`(S_\tau )_{\tau _1<\tau <\tau _2}`$ (with $`\tau _1<0<\tau _2`$) having a common boundary, with $`S_0`$ coinciding with the initial hypersurface $`S`$ described by $`t=0`$. The region of $`\mathrm{\Omega }`$ to the future of $`S`$, $`\mathrm{\Omega }^+`$ will be denoted as $`U`$. Figure 1. The foliation of region $`\mathrm{\Omega }`$ There will now be two cases to consider, according to whether $`U`$ intersects the axis $`\mathrm{\Lambda }`$ given by $`x=y=0`$. In the case where $`U\mathrm{\Lambda }=\mathrm{}`$, the metric is equivalent to that of Minkowski space and so one can write down a unique solution in terms of the Kirchoff formula (see e.g. Choquet-Bruhat et al., 1977). We shall therefore consider the more problematic case $`U\mathrm{\Lambda }\mathrm{}`$, where the metric is locally bounded but admits directional dependent limits as one approaches the axis. By writing $`\varphi =\psi +q`$ where $`qH^1(U)`$ is an arbitrary function satisfying $`q_{|S}`$ $`=\varphi _0`$ $`4`$ $`n^aq_{,a}^{}{}_{|S}{}^{}`$ $`=\varphi _1`$ we can express (3) as an initial value problem for $`\psi `$ satisfying zero initial conditions $`\mathrm{}\psi `$ $`=f\mathrm{}q`$ $`5`$ $`\psi _{|S}`$ $`=0`$ $`n^a\psi _{,a}^{}{}_{|S}{}^{}`$ $`=0`$ One possible way of defining a function $`qL^2(U)`$ which satisfies the initial conditions (4) is $`q(t,x,y,z)=\varphi _0(x,y,z)+t\varphi _1(x,y,z){\displaystyle \frac{t^3}{t^2+x^2+y^2}}\left(x\varphi _{1,x}(0,0,z)+y\varphi _{1,y}(0,0,z)\right)`$ $`(\varphi _0,\varphi _1)H^1(S)\times H^0(S)`$ By imposing stronger initial conditions, we can also achieve $`qH^1(U)`$. On differentiating $`q_{,t}`$ $`=\varphi _1(x,y,z){\displaystyle \frac{t^2(t^2+3x^2+3y^2)}{(t^2+x^2+y^2)^2}}\left(x\varphi _{1,x}(0,0,z)+y\varphi _{1,y}(0,0,z)\right)`$ $`q_{,z}`$ $`=\varphi _{0,z}(x,y,z)+t\varphi _{1,z}(x,y,z)`$ $`{\displaystyle \frac{t^3}{t^2+x^2+y^2}}\left(x\varphi _{1,xz}(0,0,z)+y\varphi _{1,yz}(0,0,z)\right)`$ $`q_{,A}`$ $`=\varphi _{0,A}(x,y,z)+t\varphi _{1,A}(x,y,z)`$ $`{\displaystyle \frac{t^3}{t^2+x^2+y^2}}\varphi _{1,A}(0,0,z)`$ $`+2{\displaystyle \frac{t^3x_A}{(t^2+x^2+y^2)^2}}\left(x\varphi _{1,x}(0,0,z)+y\varphi _{1,x}(0,0,z)\right)`$ we can see that sufficient conditions on the initial data are $$\varphi _0,\varphi _1H^1(S),\varphi _{1,A}^{}{}_{|\mathrm{\Lambda }}{}^{}H^1(S\mathrm{\Lambda })$$ Since we shall be working exclusively with the Sobolev spaces $`H^k`$; it would be very desirable for $`\mathrm{}qH^0(U)`$ so that $`\mathrm{}\psi H^0(U)`$ in (5). This will require a further strengthening of the conditions on the initial data $`(\varphi _0,\varphi _1)`$. Now $`\mathrm{}q`$ may be expressed in the form $$\mathrm{}q=g^{ab}q_{,ab}\frac{(1/A^21)}{(x^2+y^2)^{1/2}}\alpha ^aq_{,a}$$ where $$\alpha ^a=(0,x/(x^2+y^2)^{1/2},y/(x^2+y^2)^{1/2},0)$$ The non-zero components $`g^{tt}`$, $`g^{AB}`$, $`g^{zz}`$ and $`\alpha ^A`$ are essentially bounded. Therefore it is sufficient to require that $`q_{,tt}`$, $`q_{,AB}`$, $`q_{,zz}`$, $`q_{,A}/(x^2+y^2)^{1/2}`$ are all square integrable. On differentiating we find that $`q_{,tt}`$ $`=2{\displaystyle \frac{t(x^2+y^2)(3x^2+3y^2t^2)}{(t^2+x^2+y^2)^{5/2}}}`$ $`\times \left({\displaystyle \frac{x}{(x^2+y^2)^{1/2}}}\varphi _{1,x}(0,0,z)+{\displaystyle \frac{y}{(x^2+y^2)^{1/2}}}\varphi _{1,y}(0,0,z)\right)`$ $`q_{,zz}`$ $`=\varphi _{0,zz}(x,y,z)+t\varphi _{1,zz}(x,y,z)`$ $`{\displaystyle \frac{t^3}{t^2+x^2+y^2}}\left(x\varphi _{1,xzz}(0,0,z)+y\varphi _{1,yzz}(0,0,z)\right)`$ $`q_{,AB}`$ $`=\varphi _{0,AB}(x,y,z)+t\varphi _{1,AB}(x,y,z)`$ $`+2{\displaystyle \frac{t^3}{(t^2+x^2+y^2)^2}}\left(x_B\varphi _{1,A}(0,0,z)+x_B\varphi _{1,A}(0,0,z)\right)`$ $`+2{\displaystyle \frac{t^3}{(t^2+x^2+y^2)^{3/2}}}\left(\delta _{AB}4{\displaystyle \frac{x_Ax_B}{t^2+x^2+y^2}}\right)`$ $`\times \left({\displaystyle \frac{x}{(x^2+y^2)^{1/2}}}\varphi _{1,x}(0,0,z)+{\displaystyle \frac{y}{(x^2+y^2)^{1/2}}}\varphi _{1,y}(0,0,z)\right)`$ Also we apply the mean value theorem (with $`\xi ,\eta [0,1]`$) to $`q_{,A}`$ to obtain $$\begin{array}{cc}\hfill \frac{q_{,A}}{(x^2+y^2)^{1/2}}& =\frac{1}{(x^2+y^2)^{1/2}}\varphi _{0,A}(0,0,z)+\frac{t(x^2+y^2)^{1/2}}{t^2+x^2+y^2}\varphi _{1,A}(0,0,z)\hfill \\ & +\frac{x}{(x^2+y^2)^{1/2}}\left(\varphi _{0,Ax}(\xi x,0,z)+t\varphi _{1,Ax}(\xi x,0,z)\right)\hfill \\ & +\frac{y}{(x^2+y^2)^{1/2}}\left(\varphi _{0,Ay}(0,\eta y,z)+t\varphi _{1,Ay}(0,\eta y,z)\right)\hfill \\ & +2\frac{x_At^3}{(t^2+x^2+y^2)}\left(\frac{x}{(x^2+y^2)^{1/2}}\varphi _{1,x}(0,0,z)+\frac{y}{(x^2+y^2)^{1/2}}\varphi _{1,y}(0,0,z)\right)\hfill \end{array}$$ Sufficient conditions on the initial data for the integrability of each of these components are We therefore have sufficient conditions on the initial data $`(\varphi _0,\varphi _1)`$, for $`\mathrm{}qL^2(U)`$, of ## 3. Existence and uniqueness On obtaining a function $`qH^1(U)`$ which makes the right hand side of (5) square integrable, we may proceed to prove existence of a unique solution $`\psi H^1(U)`$ to (5), and hence a unique solution $`\varphi H^1(U)`$ to (3). The first step in establishing such a result is to construct an energy inequality (Clarke, 1998; Hawking and Ellis, 1973). We define the following energy integral $$E(\tau )=_{S_\tau }S^{ab}t_an_b(g)^{1/2}d^3x$$ where $`t^a`$ is the tangent to a time-like congruence of geodesics (here we can take $`t_a=\delta _a^t`$), $`n_a`$ is the normal to the surface $`S_\tau `$ and $$S^{ab}=(g^{ac}g^{bd}\frac{1}{2}g^{ab}g^{cd})\psi _{,a}\psi _{,b}\frac{1}{2}g^{ab}\psi ^2$$ It will turn out that estimating this energy integral easier than working directly with the classical Sobolev norm $`\psi _{S_\tau \mathrm{\Omega }}^1`$. However $`E(\tau )`$ and $`(\psi _{S_\tau \mathrm{\Omega }}^1)^2`$ are equivalent in the sense that one can find global positive constants $`B_1`$ and $`B_2`$ such that $$B_1(\psi _{S_\tau \mathrm{\Omega }}^1)^2E(\tau )B_2(\psi _{S_\tau \mathrm{\Omega }}^1)^2$$ $`6`$ The crucial step in obtaining an energy inequality is to apply Stokes’ theorem to the vector $`S^{ab}t_b`$ on the region $$U_\tau =\underset{0<\tau ^{}<\tau }{}S_\tau ^{}$$ however caution must be exercised because of the possible lack of differentiability of $`g_{ab}`$ and $`\psi `$ in a neighbourhood of the axis $`\mathrm{\Lambda }`$. We therefore apply Stokes’ theorem to the region $$U_\tau ^\epsilon =\{xU_\tau |x^2+y^2>\epsilon ^2\}$$ which has a boundary $`U_\tau ^\epsilon `$ consisting of $`S_\tau ^\epsilon `$ $`=S_\tau U^\epsilon `$ $`S^\epsilon `$ $`=SU^\epsilon `$ $`W_\tau ^\epsilon `$ $`=\{(t,x,y,z)U_\tau |x^2+y^2=\epsilon ^2\}`$ and will consider the limit $`\epsilon 0`$. On applying Stokes’ theorem to this region we obtain $$_{U_\tau ^\epsilon }(S^{ab}t_a)_{;b}(g)^{1/2}d^4x=_{S_\tau ^\epsilon }S^{ab}t_an_b𝑑S_{S^\epsilon }S^{ab}t_an_b𝑑S_{W_\tau ^\epsilon }S^{ab}t_an_b𝑑W$$ Taking the limit as $`\epsilon 0`$ and using the fact that $`t_{a;b}=0`$ $$E(\tau )=E(0)+F(\tau )+_{U_\tau }S^{ab}{}_{;b}{}^{}t_{a}^{}(g)^{1/2}d^4x$$ where $$F(\tau )=\underset{\epsilon 0}{lim}_{W_\tau ^\epsilon }S^{ab}t_an_b𝑑W$$ The quantity $`F(\tau )`$ is the limiting flux integral, whose significance we shall discus later; for the time being we shall assume that this quantity exists. We have $$E(\tau )=E(0)+F(\tau )+_{U_\tau }\psi _{,t}(\psi \mathrm{}\psi )(g)^{1/2}d^4x$$ which implies that $$E(\tau )E(0)+|F(\tau )|+K(\mathrm{}\psi _{U_\tau }^0)^2+K(\psi _{U_\tau }^1)^2$$ Using the fact that $$(\psi _{U_\tau }^1)^2=_0^\tau (\psi _{S_\tau ^{}\mathrm{\Omega }}^1)^2𝑑\tau ^{}$$ and on applying (6) we have $$E(\tau )E(0)+|F(\tau )|+K\mathrm{}\psi _{U_\tau }^0+(K/B_1)_0^\tau E(\tau ^{})𝑑\tau ^{}$$ This inequality may be solved by Gronwall’s Lemma (See e.g. Abraham et al., 1988) to give $$E(\tau )\left(E(0)+|F(\tau )|+K(\mathrm{}\psi _{U_\tau }^0)^2\right)e^{K/B_1\tau }$$ or equivalently $$(\psi _{S_\tau }^1)^2\frac{1}{B_1}\left(B_2(\psi _{S\mathrm{\Omega }}^1)^2+|F(\tau )|+K(\mathrm{}\psi _{U_\tau }^0)^2\right)e^{K/B_1\tau }$$ $`7`$ The first term in the right hand side is determined by the initial data in (5) and will be zero in our case. We first consider the issue of whether any such solution to (5) is unique; suppose that $`\gamma `$ is the difference of two such solutions of (5), then it must be a solution of the initial value problem $`\mathrm{}\gamma `$ $`=0`$ $`\gamma _{|S}`$ $`=0`$ $`n^a\gamma _{,a}^{}{}_{|S}{}^{}`$ $`=0`$ and so the corresponding energy inequality is $$(\gamma _{S_\tau \mathrm{\Omega }}^1)^2\frac{1}{B_1}|F(\tau )|e^{K/B_1\tau }$$ A vanishing limiting flux term $`F(\tau )`$ will imply that $`\gamma =0`$ and therefore any such solution to (5) would be unique. In the case of a non-vanishing $`F(\tau )`$, it could be possible for uniqueness to be violated. However $`F(\tau )`$ may be expressed as $$F(\tau )=\underset{\epsilon 0}{lim}_{W_\tau ^\epsilon }\psi _{,t}(n^x\psi _{,x}+n^y\psi _{,y})𝑑W$$ $`8`$ Since the metric is locally bounded and the surface element can be expressed as $`A\epsilon dtd\phi dz`$ in suitable cylindrical coordinates, a sufficient condition for $`F(\tau )`$ to vanish is that $`\psi `$ has a locally bounded derivative. We finally consider the question of whether a solution exists to (5) using a method following those of Egorov and Shubin (1992) and Clarke (1998). We shall show that a $`C^1`$ solution exists to (5), whose differentiability is sufficient for the flux term (8) to vanish and therefore guarantees that the solution is unique. We first define the solution function space $`V_0`$ and its dual $`V_1`$ $`V_0`$ $`=\{\psi C^1(U)|\psi _{|S}=n^a\psi _{,a}^{}{}_{|S}{}^{}=0\}`$ $`V_1`$ $`=\{\omega C^1(U)|\omega _{|S_{\tau _2}}=n^a\omega _{,a}^{}{}_{|S_{\tau _2}}{}^{}=0\}`$ We may apply the energy inequality to $`\psi V_0`$. It should be noted, as in the previous section, the imposed differentiability level will force the flux term $`F(\tau )`$ to vanish and so we are left with $$(\psi _{S_\tau \mathrm{\Omega }}^1)^2B_2Ke^{K/B_1\tau _2}(\mathrm{}\psi _{U_\tau }^0)^2$$ In particular we apply this to the region $`U`$ and integrate to obtain $$\psi c_1\mathrm{}\psi \psi V_0$$ where we use $`\psi `$ to denote the $`L^2`$ norm of $`\psi `$ over $`U`$. We may obtain a similar inequality for $`\omega V_1`$, by regarding $`S_{\tau _2}`$ as an initial surface, with zero initial data, evolving back in time and constructing an analogous energy inequality; thus we obtain $$\omega c_2\mathrm{}\omega \omega V_1$$ $`9`$ We now apply Stokes’ theorem to $`\psi \omega ^{;a}`$ and $`\omega \psi ^{;a}`$ for $`\psi V_0`$ and $`\omega V_1`$ noting that the boundary contributions on $`S`$ and $`S_{\tau _2}`$ vanish. $`{\displaystyle _{U^\epsilon }}\psi ^{;a}\omega _{;a}(g)^{1/2}d^4x+{\displaystyle _{U^\epsilon }}\psi \mathrm{}\omega (g)^{1/2}d^4x`$ $`={\displaystyle _{W_{\tau _2}^\epsilon }}\psi \omega _{,a}n^a𝑑W`$ $`{\displaystyle _{U^\epsilon }}\psi ^{;a}\omega _{;a}(g)^{1/2}d^4x+{\displaystyle _{U^\epsilon }}\mathrm{}\psi \omega (g)^{1/2}d^4x`$ $`={\displaystyle _{W_{\tau _2}^\epsilon }}\omega \psi _{,a}n^a𝑑W`$ On taking the limit $`\epsilon 0`$, the flux integrals on the right hand side vanish because $`\psi `$ and $`\omega `$ are $`C^1`$. Subtracting the resulting equations gives $$_U\mathrm{}\psi \omega (g)^{1/2}d^4x=_U\psi \mathrm{}\omega (g)^{1/2}d^4x$$ which, because $`(g)^{1/2}`$ is constant may be written as $$_U\mathrm{}\psi \omega d^4x=_U\psi \mathrm{}\omega d^4x$$ $`10`$ As a consequence, we may define a linear functional $`k:\mathrm{}V_1`$ by $$k(\mathrm{}\omega )=_U(f\mathrm{}q)\omega d^4x$$ and by (9) $$\begin{array}{cc}\hfill \left|k(\mathrm{}\omega )\right|& f\mathrm{}q\omega \hfill \\ & c_2f\mathrm{}q\mathrm{}\omega \hfill \end{array}$$ implying that this linear functional is also bounded. In order for $`k`$ to correspond to an element of $`V_0`$ (a solution of (5)) we must show that $`\mathrm{}V_1`$ is a dense subspace of $`L^2(U)`$. It is sufficient to show that any function in the space $`(\mathrm{}V_1)^{}`$ is necessarily zero. Suppose that $`\lambda (\mathrm{}V_1)^{}`$ then, because $`V_0`$ is dense, there exists a sequence $`(\psi _n)`$ in $`V_0`$ converging to $`\lambda `$. By (10), we have for all $`\omega V_1`$ $$\begin{array}{cc}\hfill \underset{n\mathrm{}}{lim}_U\mathrm{}\psi _n\omega d^4x& =\underset{n\mathrm{}}{lim}_U\mathrm{}\omega \psi _nd^4x\hfill \\ & =_U\mathrm{}\omega \lambda d^4x\hfill \\ & =0\hfill \end{array}$$ Since $`V_1`$ is dense, this implies that $`\mathrm{}\psi _n0`$ a.e. That is we have $`\lambda =0`$ ## 4. Conclusion We have shown that by using the functional analytic methods of Clarke (1998), that for initial data satisfying the following conditions; a unique solution $`\varphi H^1(U)`$ to (3) exists in every open region $`U`$ of the conical space-time (1), and therefore such a space-time can be regarded as hyperbolic in this sense. The integrability conditions on the initial data resulted from the choice of the arbitrary function $`qH^1(U)`$ which satisfied the initial data. It may be possible to weaken them by a more careful choice of such a function. The actual proof of existence and uniqueness depended on two properties of the metric; namely that both its covariant and contravariant components are locally essentially bounded, and that there existed a natural congruence of time-like geodesics whose tangent has an essentially bounded covariant derivative. It is therefore possible to modify the proof to show that a similar evolution of the wave equation is possible in other conical space-times whose metric differs from (1) by a $`C^2`$ perturbation. In such a space-time the conical singularity will have a constant deficit angle and the axis may be regarded as a totally geodesic two dimensional time-like submanifold, and therefore represents a thin string on a curved background. More recently the same problem was approached by Vickers and Wilson (1999) using Colombeau’s generalised functions (See e.g. Colombeau, 1984). Colombeau’s theory enables a distributional interpretation to be given to products of distributions that would otherwise not be defined in classical distribution theory, and therefore has a natural application to non-linear theories such as General Relativity. One important recent application was to rigorously establish the form of the energy momentum tensor (2). In that paper, the initial value problem (3) was formulated as an initial value problem in the Colombeau Algebra and existence and uniqueness was proved within the algebra. It should be noted that, by working with regularised functions in a Colombeau Algebra, it was possible to establish existence and uniqueness given any initial data $`(\varphi _0,\varphi _1)H^1(S)\times H^0(S)`$, where as in this paper we needed stronger differentiability conditions on the initial data. ## Acknowledgements The author wishes to thank Professor Chris Clarke and Dr James Vickers for many helpful discussions.
warning/0001/hep-th0001077.html
ar5iv
text
# ROM2F-2000/2hep-th/0001077 Open-string models with broken supersymmetry ## 1. Broken supersymmetry and type-0 models In this talk I would like to review the key features of some open-string models with broken supersymmetry constructed in . These models may be derived in a systematic fashion from corresponding models of oriented closed strings , and once more display a surprising richness compared to them. Since the relevant techniques have been discussed at length in the original papers, I will not present any explicit derivations. Rather, referring to some of the resulting vacuum amplitudes, I will try to illustrate how supersymmetry can be broken at tree level in the bulk, on some branes or everywhere. Closed-string models with broken supersymmetry were among the first new examples considered in the last decade. In particular, the type-0 models provided the first non-trivial instances of modified GSO projections compatible with modular invariance. In order to describe their partition functions, I will begin by introducing some notation that will be used repeatedly in the following, defining the four level-one SO(8) characters $`O_8={\displaystyle \frac{\vartheta _3^4+\vartheta _4^4}{2\eta ^4}},V_8={\displaystyle \frac{\vartheta _3^4\vartheta _4^4}{2\eta ^4}},`$ $`S_8={\displaystyle \frac{\vartheta _2^4\vartheta _1^4}{2\eta ^4}},C_8={\displaystyle \frac{\vartheta _2^4+\vartheta _1^4}{2\eta ^4}},`$ (1.1) where the $`\vartheta _i`$ are Jacobi theta functions and $`\eta `$ is the Dedekind function. In terms of these characters, and leaving aside the contribution of the eight transverse bosonic coordinates, the type II models are described by $`𝒯_{IIA}`$ $`=`$ $`(V_8S_8)(\overline{V}_8\overline{C}_8),`$ (1.2) $`𝒯_{IIB}`$ $`=`$ $`|V_8S_8|^2,`$ (1.3) while the type-0A and type-0B models are described by $`𝒯_{0A}=|O_8|^2+|V_8|^2+S_8\overline{C}_8+C_8\overline{S}_8,`$ (1.4) $`𝒯_{0B}=|O_8|^2+|V_8|^2+|S_8|^2+|C_8|^2.`$ (1.5) In these expressions, the characters $`(O_8,V_8,S_8,C_8)`$ depend on $`q=exp(i2\pi \tau )`$, with $`\tau `$ the modulus of the torus, while their conjugates depend of $`\overline{q}`$. All these characters have power series expansions of the type $$\chi (q)=q^{hc/24}\underset{n=0}{\overset{\mathrm{}}{}}d_nq^n,$$ (1.6) where the $`d_n`$ are integers. The low-lying spectra, essentially manifest in this notation, include in all cases the universal triple ($`g_{\mu \nu }`$, $`B_{\mu \nu }`$, $`\varphi `$). The corresponding states fill a generic transverse matrix, the direct product of the ground states of the $`V_8`$ module and of its conjugate $`\overline{V}_8`$. In addition, the type-IIA superstring has a Majorana gravitino and a Majorana spinor from the NS-R and R-NS sectors and a vector and a three-form from the R-R sector. The fermions result from pairs of Majorana-Weyl spinors of opposite chiralities, ground states of $`V_8\overline{C}_8`$ and $`S_8\overline{V}_8`$, while the nature of the R-R bosons is determined by the direct product of the two inequivalent spinor representations of SO(8). These are the ground states of $`S_8`$ and $`\overline{C}_8`$, and the product $`8_s\times 8_c`$ indeed decomposes into a vector and a three-form. A similar reasoning shows that the type-IIB superstring has a complex Majorana-Weyl gravitino and a complex Weyl fermion from the NS-R and R-NS sectors, and a scalar, a two-form and a self-dual four-form from the R-R sector. The spectra of the type-0 models are purely bosonic, and can be essentially deduced from these. Aside from the universal triple ($`g_{\mu \nu }`$, $`B_{\mu \nu }`$, $`\varphi `$), their low-lying excitations include a tachyon, the ground state of the $`O_8\overline{O}_8`$ sector, while their R-R sectors are two copies of the previous ones, and include a pair of vectors and a pair of three-forms for the 0A model, and a pair of scalars, a pair of two-forms and an unconstrained four-form for the 0B model. Both type-0 models are clearly non-chiral, and are thus free of gravitational anomalies. Let us now turn to the open descendants of the type-0 models. Their structure is essentially determined by the Klein bottle projection. Leaving aside the contributions of the transverse bosons, the conventional choice, originally discussed in , corresponds to $`𝒦_{0A}`$ $`=`$ $`{\displaystyle \frac{1}{2}}(O_8+V_8),`$ (1.7) $`𝒦_{0B}`$ $`=`$ $`{\displaystyle \frac{1}{2}}(O_8+V_8S_8C_8).`$ (1.8) It eliminates the NS-NS two-form $`B_{\mu \nu }`$, but does not affect the tachyon. The effect of $`𝒦`$ can be simply summarized recalling that a positive (negative) sign implies a symmetrization (antisymmetrization) of the sectors fixed under left-right interchange. This unoriented projection is frequently called $`\mathrm{\Omega }`$. For instance, $`𝒦_{0B}`$ symmetrizes the two NS-NS sectors, described by $`|O_8|^2`$ and $`|V_8|^2`$, and antisymmetrizes the two R-R sectors, eliminating the NS-NS $`B_{\mu \nu }`$ and leaving only a pair of R-R two-forms. In addition, the projected spectrum generally includes invariant combinations of all pairs of sectors interchanged by $`\mathrm{\Omega }`$, that do not contribute to $`𝒦`$. Thus, the low-lying spectrum of the projected 0A model includes also a R-R vector and a R-R three-form. As usual, the characters in these direct-channel Klein-bottle amplitudes depend on $`q\overline{q}=exp(4\pi \tau _2)`$. The open sector of the 0A model, described by $`𝒜_{0A}`$ $`=`$ $`{\displaystyle \frac{n_B^2+n_F^2}{2}}(O_8+V_8)n_Bn_F(S_8+C_8),`$ (1.9) $`_{0A}`$ $`=`$ $`{\displaystyle \frac{n_B+n_F}{2}}\widehat{V}_8{\displaystyle \frac{n_Bn_F}{2}}\widehat{O}_8,`$ (1.10) is not chiral and involves two different “real” charges, corresponding to orthogonal or symplectic groups. These enter the partition functions via the dimensions, here $`n_B`$ and $`n_F`$, of the corresponding fundamental representations, and in general are subject to linear relations originating from (massless) tadpole conditions. The low-lying modes are simply identified from the contributions of $`𝒜`$ and $``$. For instance, the model contains two sets of $`n_B(n_B1)/2`$ and $`n_F(n_F1)/2`$ vectors (corresponding to $`V_8`$), enough to fill the adjoint representations of a pair of orthogonal groups, tachyons (corresponding to $`O_8`$) in doubly (anti)symmetric representations and fermions (corresponding to the R characters $`S_8`$ and $`C_8`$) in bi-fundamental representations. This description of open-string spectra is also useful in Conformal Field Theory, where it provides a convenient encoding of the spectrum of boundary operators in a generating function of their multiplicities. In this case, if one insists on demanding the cancellation of all NS-NS tadpoles, the result is the family of gauge groups SO$`(n_B)\times `$ SO$`(n_F)`$, with $`n_B+n_F=32`$. Here $`𝒜`$ depends on $`(q\overline{q})^{1/4}`$ = $`exp(\pi \tau _2)`$. On the other hand, $``$ depends on $`(q\overline{q})^{1/4}`$ = $`exp(\pi \tau _2+i\pi )`$, but the “hatted” characters are redefined by suitable phases, and are thus real. The open sector of the 0B model, described by $`𝒜_{0B}`$ $`=`$ $`{\displaystyle \frac{n_o^2+n_v^2+n_s^2+n_c^2}{2}}V_8+(n_on_v+n_sn_c)O_8`$ (1.11) $``$ $`(n_vn_s+n_on_c)S_8(n_vn_c+n_on_s)C_8,`$ $`_{0B}`$ $`=`$ $`{\displaystyle \frac{n_o+n_v+n_s+n_c}{2}}\widehat{V}_8,`$ (1.12) involves four different “real” charges, and is chiral but free of anomalies, as a result of the R-R tadpole conditions $`n_o=n_v`$ and $`n_s=n_c`$. All irreducible gauge and gravitational anomalies cancel as a result of the R-R tadpole conditions, while the residual anomaly polynomial requires a generalized Green-Schwarz mechanism . If one insists on demanding the cancellation of all NS-NS tadpoles, not related to anomalies as the previous ones, one obtains the family of gauge groups SO$`(n_o)\times `$ SO$`(n_v)\times `$SO$`(n_s)\times `$ SO$`(n_c)`$, with $`n_o+n_v+n_s+n_c=64`$. Despite their apparent complication, these open-string models are actually simpler than the previous ones, since the 0B torus amplitude corresponds to the “charge-conjugation” modular invariant. This circumstance implies a one-to-one correspondence between types of boundaries and types of bulk sectors typical of the “Cardy case” of boundary CFT . In equivalent terms, this model has four types of boundary states, in one-to-one correspondence with the chiral sectors of the bulk spectrum. The boundary states of the 0A model are a bit subtler, since they are proper combinations of these that do not couple to the R-R states, that cannot flow in the transverse channel compatibly with 10D Lorentz invariance. Indeed, the product of the two spinor representations $`8_s`$ and $`8_c`$ does not contain the identity, and consequently a right-moving $`S`$ state cannot reflect into a left-moving $`C`$ state at a Lorentz-invariant boundary. The modified Klein bottle projection $$𝒦_{}^{}{}_{0B}{}^{}=\frac{1}{2}(O_8+V_8+S_8C_8),$$ (1.13) first proposed in as an amusing application of the results of , removes the tachyon and the NS-NS two-form $`B_{\mu \nu }`$ from the closed spectrum, and leaves a chiral unoriented closed spectrum that comprises the $`(g_{\mu \nu },\varphi )`$ NS-NS pair, together with a two-form, an additional scalar and a self-dual four-form from the R-R sectors. The resulting open spectrum, described by $`𝒜_{0B}^{}`$ $`=`$ $`{\displaystyle \frac{n^2+\overline{n}^2+m^2+\overline{m}^2}{2}}C_8+(n\overline{n}+m\overline{m})V_8`$ (1.14) $`+`$ $`(n\overline{m}+m\overline{n})O_8(mn+\overline{m}\overline{n})S_8,`$ $`_{0B}^{}`$ $`=`$ $`{\displaystyle \frac{m+\overline{m}n\overline{n}}{2}}C_8,`$ (1.15) involves the “complex” charges of a pair of unitary groups, subject to the R-R tadpole constraint $`mn=32`$, that eliminates all (non-Abelian) gauge and gravitational anomalies. The notation resorts to pairs of multiplicities , say $`m`$ and $`\overline{m}`$, to emphasize the different roles of the fundamental and conjugate fundamental representations of a unitary group U(m). The tadpole conditions identify the numerical values of $`m`$ and $`\overline{m}`$, but once again one can read the low-lying spectrum directly and conveniently from the amplitudes written in this form. The choice $`n=0`$ selects a U(32) gauge group, with a spectrum that is free of tachyons both in the closed and in the open sectors. All irreducible (non-Abelian) gauge and gravitational anomalies cancel, while the residual anomaly polynomial requires a generalized Green-Schwarz mechanism . On the other hand, the U(1) factor is anomalous, and is thus lifted by a ten-dimensional generalization of the mechanism of , so that the effective gauge group of this model is SU(32). Here one does not have the option of eliminating all the NS-NS tadpoles, and as a result a dilaton potential is generated. These models have also been studied in some detail in , first with the aim of connecting them to the bosonic string, and more recently with the aim of relating them to (non-supersymmetric) reductions of M theory. This last approach goes beyond the perturbative analysis, and therefore has the potential of discriminating between the various options. According to , both types of $`0B`$ descendants admit a non-perturbative definition, while the $`0A`$ descendants do not. It would be interesting to take a closer look at this relatively simple model and try to elicit some manifestation of this phenomenon. Let us now spend a few words to summarize the key features of these descendants, where supersymmetry is broken both in the closed and in the open sectors. Whereas the first two models have tachyons both in the closed and in the open sectors, the last results from a non-tachyonic brane configuration of impressive simplicity. This feature actually extends to lower-dimensional compactifications, as first shown by Angelantonj . These type 0 models, and in particular the non-tachyonic one, have interesting applications in the framework of the AdS/CFT correspondence . A simple generalisation of this setting allows one to describe the branes allowed in these ten-dimensional models and in their “parent” oriented closed models. These results, originally obtained by a number of authors, can be efficiently described in this formalism as in . ## 2. Scherk-Schwarz deformations and brane supersymmetry We may now turn to the second class of models. These rest on elegant extensions of the Kaluza-Klein reduction, known as Scherk-Schwarz deformations , that allow one to induce the breaking of supersymmetry from the different behaviors of fermionic and bosonic modes in the internal space. This setting, as adapted to the entire perturbative spectra of models of oriented closed strings in , is the starting point for the constructions in . I will confine my attention to particularly simple examples, related to the reduction of the type IIB superstring on a circle of radius $`R`$ where the momenta or the windings are subjected to 1/2-shifts, compatibly with modular invariance, in such a way that all massless fermions are lifted in mass. In these models, supersymmetry is completely broken, but several more complicated open-string models, with partial breaking of supersymmetry are discussed in . This Scherk-Schwarz deformation generically introduces tachyons, in the first case (momentum shifts) for $`R<\sqrt{\alpha ^{}}`$, and in the second case (winding shifts) for $`R>\sqrt{\alpha ^{}}`$. The former choice is essentially a Scherk-Schwarz deformation of the low-energy field theory, here lifted to the entire string spectrum. On the other hand, the latter is a bit more subtle to interpret from a field theory perspective, and indeed the resulting deformation of the spectrum is removed in the limit of small radius $`R`$, that strictly speaking is inaccessible to the field theory description. Naively, in the first case the open descendants should not present new subtleties. One would expect that the momentum deformations be somehow inherited by the open spectrum, and this is indeed what happens. On the other hand, naively the open spectrum should be insensitive to winding deformations, simply because the available Neumann strings have only momentum excitations. Here the detailed analysis settles the issue in an interesting way. The open spectrum is indeed affected, although in a rather subtle way, and supersymmetry is effectively broken again, at the compactification scale $`1/R`$, but is exact for the massless modes. In order to appreciate this result, let us present the closed-string amplitudes for the two cases, here written in the Scherk-Schwarz basis, $`𝒯_1`$ $`=`$ $`Z_{m,2n}(V_8\overline{V}_8+S_8\overline{S}_8)+Z_{m,2n+1}(O_8\overline{O}_8+C_8\overline{C}_8)`$ (2.1) $``$ $`Z_{m+1/2,2n}(V_8\overline{S}_8+S_8\overline{V}_8)Z_{m+1/2,2n+1}(O_8\overline{C}_8+C_8\overline{O}_8)`$ and $`𝒯_2`$ $`=`$ $`Z_{2m,n}(V_8\overline{V}_8+S_8\overline{S}_8)+Z_{2m+1,n}(O_8\overline{O}_8+C_8\overline{C}_8)`$ (2.2) $``$ $`Z_{2m,n+1/2}(V_8\overline{S}_8+S_8\overline{V}_8)Z_{2m+1,n+1/2}(O_8\overline{C}_8+C_8\overline{O}_8),`$ and the corresponding Klein bottle projections $`𝒦_1`$ $`=`$ $`{\displaystyle \frac{1}{2}}(V_8S_8)Z_m,`$ (2.3) $`𝒦_2`$ $`=`$ $`{\displaystyle \frac{1}{2}}(V_8S_8)Z_{2m}+{\displaystyle \frac{1}{2}}(O_8C_8)Z_{2m+1}.`$ (2.4) In these expressions, $`Z_{m,n}`$ denotes the usual Narain lattice sum for the circle $$Z_{m,n}=\underset{m,n}{}q^{\frac{\alpha ^{}}{4}(\frac{m}{R}+\frac{nR}{\alpha ^{}})^2}\overline{q}^{\frac{\alpha ^{}}{4}(\frac{m}{R}\frac{nR}{\alpha ^{}})^2},$$ (2.5) while, for instance, $`Z_{2m,n}`$ denotes the sum restricted to even momenta. In a similar fashion, $`Z_m`$ in (2.4) denotes the restriction of the sum to the momentum lattice. In writing the corresponding open sectors, I will now eliminate several contributions, restricting the charge configurations in such a way that no tachyons are introduced. This is, to some extent, in the spirit of the previous discussion of the 10D U(32) model, but here one can also cancel all NS-NS tadpoles. Moreover, I will take into account the infrared subtlety discussed in , that in the model with winding shifts leads the emergence of additional tadpoles in the singular limit $`R0`$, where whole towers of massive excitations collapse to zero mass. With this proviso, the corresponding open spectra are described by $`𝒜_1`$ $`=`$ $`{\displaystyle \frac{n_1^2+n_2^2}{2}}(V_8Z_mS_8Z_{m+1/2})+n_1n_2(V_8Z_{m+1/2}S_8Z_m),`$ $`_1`$ $`=`$ $`{\displaystyle \frac{n_1+n_2}{2}}(\widehat{V}_8Z_m\widehat{S}_8Z_{m+1/2}),`$ (2.6) and $`𝒜_2`$ $`=`$ $`{\displaystyle \frac{n_1^2+n_2^2}{2}}(V_8S_8)Z_m+n_1n_2(O_8C_8)Z_{m+1/2},`$ $`_2`$ $`=`$ $`{\displaystyle \frac{n_1+n_2}{2}}\left(\widehat{V}_8(1)^m\widehat{S}_8\right)Z_m,`$ (2.7) As anticipated, the first model, with $`n_1+n_2=32`$, is essentially a conventional Scherk-Schwarz deformation of the type-I superstring. It can also describe the type I spectrum at a finite temperature related to the internal radius $`R`$. On the other hand, the second model, where $`n_1=n_2=16`$, is more interesting, and displays the first novel phenomenon reviewed here, “brane supersymmetry”: although supersymmetry is broken at the compactification scale by the Scherk-Schwarz deformation, the massless modes of the open sector fill complete supersymmetry multiplets. I would like to stress that the breaking of supersymmetry in the massive spectrum of the second model can also be regarded as a deformation, now resulting from the unpairing of the Chan-Paton representations for bosonic and fermionic modes with lattice excitations at alternate massive levels. This is the simplest instance of the phenomenon that, following , we can call “brane supersymmetry”. Here the residual supersymmetry is present only for the massless modes, but in more complicated models it extends to entire sectors of the open spectrum, as first shown in . T-dualities turn these descriptions into equivalent ones, that can often have more intuitive appeal . This is particularly rewarding in the second case: a T-duality along the circle can turn the winding deformation into a momentum deformation orthogonal to the brane responsible for the open-string excitations. It is then perhaps simpler to accept the previous result that the deformation, now a momentum shift orthogonal to the brane, does not affect its massless excitations. A little more work results in a duality argument that associates the (now momentum) shift to the eleventh dimension of M theory, thus realizing the proposal of . Thus, as is often the case, a simple perturbative type I phenomenon has a non-perturbative origin in the heterotic string (and vice versa). ## 3. Brane supersymmetry breaking I will now conclude by reviewing the third possibility afforded by these constructions. Here I will follow , concocting a six-dimensional analogue of “discrete torsion” . The construction of the resulting closed string model is another application of the methods of , in the same spirit as the construction of the 10D U(32) model. Starting from the $`T^4/Z_2`$ U(16) $`\times `$ U(16) model , one can revert the Klein-bottle projection for all twisted states. This results in an unoriented closed spectrum with (1,0) supersymmetry, whose massless excitations, aside from the gravitational multiplet, comprise 17 tensor multiplets and 4 hypermultiplets. In it is shown how this choice, described by $`𝒯`$ $`=`$ $`{\displaystyle \frac{1}{2}}|Q_o+Q_v|^2\mathrm{\Lambda }+{\displaystyle \frac{1}{2}}|Q_oQ_v|^2\left|{\displaystyle \frac{2\eta }{\theta _2}}\right|^4`$ $`+`$ $`{\displaystyle \frac{1}{2}}|Q_s+Q_c|^2\left|{\displaystyle \frac{2\eta }{\theta _4}}\right|^4+{\displaystyle \frac{1}{2}}|Q_s+Q_c|^2\left|{\displaystyle \frac{2\eta }{\theta _3}}\right|^4,`$ $`𝒦`$ $`=`$ $`{\displaystyle \frac{1}{4}}\left\{(Q_o+Q_v)(P+W)2\times 16(Q_s+Q_c)\right\},`$ (3.2) does not allow a consistent supersymmetric solution of the tadpole conditions. A consistent solution does exist , but requires the introduction of anti-branes, with the end result that supersymmetry, exact to lowest order in the bulk, is necessarily broken on their world volume. Hence the name “brane supersymmetry breaking” for this peculiar phenomenon, that has the attractive feature of confining the breaking of supersymmetry, and the resulting contributions to the vacuum energy, to a brane, or to a collection of branes, that float in a bath of supersymmetric gravity. In writing these expressions, I have introduced the $`(1,0)`$ supersymmetric characters $`Q_o`$ $`=`$ $`V_4O_4C_4C_4,Q_v=O_4V_4S_4S_4,`$ $`Q_s`$ $`=`$ $`O_4C_4S_4O_4,Q_c=V_4S_4C_4V_4.`$ (3.3) The two untwisted ones, $`Q_o`$ and $`Q_v`$, start with a vector multiplet and a hypermultiplet, and are $`Z_2`$ orbifold breakings of $`(V_8S_8)`$. Out of the two twisted ones $`Q_s`$ and $`Q_c`$, only $`Q_s`$ describes massless modes, that in this case correspond to a half-hypermultiplet. The breaking of supersymmetry is demanded by the consistency of String Theory. This can be seen rather neatly from the dependence of the transverse-channel Klein bottle amplitude at the origin of the lattices on the sign $`ϵ`$ associated to the twisted states $$\stackrel{~}{𝒦}_0=\frac{2^5}{4}\left\{Q_o\left(\sqrt{v}\pm \frac{1}{\sqrt{v}}\right)^2+Q_v\left(\sqrt{v}\frac{1}{\sqrt{v}}\right)^2\right\},$$ (3.4) where the upper signs would correspond to the conventional case of the U(16) $`\times `$ U(16) model, while the lower signs correspond to the model of eq. (3.2). Since the terms with different powers of $`\sqrt{v}`$ are related by tadpole conditions to the multiplicities of the $`N`$ and $`D`$ charge spaces, a naive solution of the model corresponding to the upper signs would require a negative multiplicity, $`D=32`$. The open sector is described by $`𝒜`$ $`=`$ $`{\displaystyle \frac{1}{4}}\{(Q_o+Q_v)(N^2P+D^2W)+2ND(Q_s^{}+Q_c^{})\left({\displaystyle \frac{\eta }{\theta _4}}\right)^2`$ $`+`$ $`(R_N^2+R_D^2)(Q_oQ_v)\left({\displaystyle \frac{2\eta }{\theta _2}}\right)^2+2R_NR_D(O_4S_4C_4O_4+V_4C_4+S_4V_4)\left({\displaystyle \frac{\eta }{\theta _3}}\right)^2\}`$ $``$ $`=`$ $`{\displaystyle \frac{1}{4}}\{NP(\widehat{O}_4\widehat{V}_4+\widehat{V}_4\widehat{O}_4\widehat{S}_4\widehat{S}_4\widehat{C}_4\widehat{C}_4)DW(\widehat{O}_4\widehat{V}_4+\widehat{V}_4\widehat{O}_4+\widehat{S}_4\widehat{S}_4+\widehat{C}_4\widehat{C}_4)`$ $``$ $`N(\widehat{O}_4\widehat{V}_4\widehat{V}_4\widehat{O}_4\widehat{S}_4\widehat{S}_4+\widehat{C}_4\widehat{C}_4)\left({\displaystyle \frac{2\widehat{\eta }}{\widehat{\theta }_2}}\right)^2+D(\widehat{O}_4\widehat{V}_4\widehat{V}_4\widehat{O}_4+\widehat{S}_4\widehat{S}_4\widehat{C}_4\widehat{C}_4)\left({\displaystyle \frac{2\widehat{\eta }}{\widehat{\theta }_2}}\right)^2\}.`$ Supersymmetry is broken on the antibranes, and indeed the amplitudes involve new characters $`Q_s^{}`$ and $`Q_c^{}`$, that describe supermultiplets of a chirally flipped supercharge and may be obtained from eq. (3.3) upon the interchange of $`S_4`$ and $`C_4`$, as well as other non-supersymmetric combinations. The tadpole conditions determine the gauge group $`[`$ SO(16) $`\times `$ SO(16) $`]_9\times [`$ USp(16) $`\times `$ USp(16) $`]_{\overline{5}}`$, and the $`99`$ spectrum is supersymmetric, with (1,0) vector multiplets for the SO(16) $`\times `$ SO(16) gauge group and a hypermultiplet in the $`(\mathrm{𝟏𝟔},\mathrm{𝟏𝟔},\mathrm{𝟏},\mathrm{𝟏})`$. On the other hand, the $`\overline{5}\overline{5}`$ spectrum is not supersymmetric and, aside from the $`[`$ USp(16) $`\times `$ USp(16) $`]`$ gauge vectors, contains quartets of scalars in the $`(\mathrm{𝟏},\mathrm{𝟏},\mathrm{𝟏𝟔},\mathrm{𝟏𝟔})`$, right-handed Weyl fermions in the $`(\mathrm{𝟏},\mathrm{𝟏},\mathrm{𝟏𝟐𝟎},\mathrm{𝟏})`$ and $`(\mathrm{𝟏},\mathrm{𝟏},\mathrm{𝟏},\mathrm{𝟏𝟐𝟎})`$ and left-handed Weyl fermions in the $`(\mathrm{𝟏},\mathrm{𝟏},\mathrm{𝟏𝟔},\mathrm{𝟏𝟔})`$. Finally, the ND sector, also not supersymmetric, comprises doublets of scalars in the $`(\mathrm{𝟏𝟔},\mathrm{𝟏},\mathrm{𝟏},\mathrm{𝟏𝟔})`$ and in the $`(\mathrm{𝟏},\mathrm{𝟏𝟔},\mathrm{𝟏𝟔},\mathrm{𝟏})`$, and additional (symplectic) Majorana-Weyl fermions in the $`(\mathrm{𝟏𝟔},\mathrm{𝟏},\mathrm{𝟏𝟔},\mathrm{𝟏})`$ and $`(\mathrm{𝟏},\mathrm{𝟏𝟔},\mathrm{𝟏},\mathrm{𝟏𝟔})`$. These fields are a peculiar feature of six-dimensional space time, where one can define Majorana-Weyl fermions, if the Majorana condition is supplemented by the conjugation in a pseudo-real representation. All irreducible gauge and gravitational anomalies cancel also in this model, while the residual anomaly polynomial requires a generalized Green-Schwarz mechanism . The radiative corrections in this model are quite interesting, since they convey the (soft) breaking to the gravitational sector. At any rate, the situation in which a model requires the simultaneous presence of branes and presents itself in at least two other instances , the four dimensional $`Z_2\times Z_2`$ model with discrete torsion and the four-dimensional $`Z_4`$ model. In both cases, brane supersymmetry breaking allows a solution of all tadpole conditions and a consistent definition of the open descendants. One can actually enrich these constructions, allowing for the simultaneous presence of branes and antibranes . These configurations are generically unstable, and their instability reflects itself in the presence of tachyonic excitations, a feature that we have already confronted in our analysis of the ten-dimensional type-0 models. The internal lattice can be used to lift in mass the tachyons, at least within certain ranges of parameters for the internal geometry, that is actually partly stabilized, as a result of the different scaling behavior ($`O(\sqrt{v})`$ and $`O(1/\sqrt{v})`$) of the contributions of the different D<sub>p</sub> branes. Thus, for instance, starting from the type-IIB superstring, one can introduce both branes and anti-branes at the price of having tachyonic excitations in the open spectrum . In addition, even with special tachyon-free configurations, simply waiving the restriction to configurations free of NS-NS tadpoles often gives new interesting models with broken supersymmetry. The simplest setting is provided again by the type-IIB superstring that, aside from the type-I superstring, has an additional chiral tachyon-free descendant, free of gauge and gravitational anomalies, but with broken supersymmetry and a USp(32) gauge group. In lower-dimensional models, more possibilities are afforded by the internal lattice, that may be used to lift in mass some tachyons, leading to stable vacuum configurations including both branes and antibranes. Some of these , related to the four-dimensional $`Z_3`$ orientifold of , appear particularly interesting. I would like to conclude by mentioning that brane configurations similar to these have also been studied by several other authors over the last couple of years, from a different vantage point, following Sen . Brane studies are reminiscent of monopole studies in gauge theories of a classical Electrodynamics of charge probes in a given external field, and are an interesting enterprise in their own right. Open-string vacua are particular brane configurations that are also vacuum configurations for a perturbative construction, and thus become exact solutions in the limit of vanishing string coupling. This, in retrospect, makes them particularly attractive and instructive, and makes their study particularly rewarding. In addition, they have very amusing applications, in particular to issues related to the AdS/CFT correspondence, that we are only starting to appreciate. Acknowledgments I am grateful to the Organizers for their kind invitations and to C. Angelantonj, I. Antoniadis, M. Bianchi, G. D’ Appollonio, E. Dudas, G. Pradisi, Ya.S. Stanev for stimulating discussions and very enjoyable collaborations on the topics reviewed in this talk, originally prepared for the QFTHEP Meeting, held in Moscow in June 1999 and for the QG99 Meeting, held in Villasimius in September 1999. With slight corrections in the text and in the references, it is now being contributed to the Proceedings of the 2001 Como Conference on “Statistical Field Theories”, of the 2001 Les Houches Summer Institute on “Gravity, gauge theory and strings” and of the 2001 Johns Hopkins Meeting, since my presentations in all cases are within the spirit of this short review. A forthcoming review article will provide a comprehensive derivation of these world-sheet constructions more easily accessible to interested readers. This work was partly supported by the EEC contracts HPRN-CT-2000-001222, HPRN-CT-2001-00148 and by the INTAS contract 99-1-590.
warning/0001/hep-ph0001203.html
ar5iv
text
# The Nucleon Sigma Term from Threshold Parameters ## I Introduction The sigma term is a fundamental scale of chiral physics, a measure of chiral symmetry breaking. For decades there has been a lively discussion between “large” sigma term advocates and “small” ones. We propose here a new evaluation method which is independent of global phase shift analyses. Our expression depends only on threshold parameters. Recent experiments allow phase shifts to be determined using only data near threshold. When combined with the present sum rule, a truly local sigma term value is obtained together with a realistic error. There are two basic methods for determining the sigma term. In the framework of QCD the expected chiral symmetry-breaking term is proportional to the bare quark mass and transforms as the $`(3\overline{3})+(\overline{3}3)`$ representation of $`\mathrm{SU}(3)\times \mathrm{SU}(3)`$. This is equivalent to the original (pre-QCD) GMOR model. Using a conventional ratio for quark masses, $`\frac{m_u+m_d}{2m_s}\frac{1}{25}`$, and assuming the nucleon has no strangeness content, the “theoretical” sigma term comes out to be about 25 MeV. With higher order corrections and strangeness the predicted sigma term might rise to about 40 MeV. The second method employs the “on-shell theorem”, where the sigma term is related to the $`\pi N`$ scattering amplitude at the “Cheng-Dasher” point. This method historically yields a larger value of about 65 MeV. This method is used here and is described in some detail below. It is the apparent discrepancy between the “theoretical” first method and the “experimental” second method that makes the $`\pi N`$ sigma term so interesting. The soft pion chiral constraints on the crossing even $`\pi N\pi N`$ amplitude $`A^{(+)}(\nu ,t,q^2,q^2)`$ are most succinctly expressed in terms of $$\overline{D}(\nu _B,q^2,q^2)A^{(+)}(\nu =0,t,q^2,q^2)\frac{g(q^2)g(q^2)}{M},$$ (1) where $`t=(qq^{})^2`$, $`4M\nu _B=t+q^2+q^2`$, $`4M\nu =su`$, $`g`$ is the renormalized $`\pi NN`$ coupling constant $`\frac{g^2(\mu ^2)}{4\pi }\frac{g^2}{4\pi }14`$, and $`M`$ and $`\mu `$ are the nucleon and charged pion masses. The Adler PCAC consistency condition requires that if one pion mass is soft, $$\overline{D}(\nu _B=0,q^2=\mu ^2,q^2=0)=\overline{D}(\nu _B=0,q^2=0,q^2=\mu ^2)=0.$$ (2) When both pions are soft, the chiral limit predicts that $`\overline{D}(0,0,0)=0`$ and the sigma term is defined as the chiral symmetry breaking in this limit $$\frac{\sigma (0)}{F_\pi ^2}\overline{D}(\nu _B=0,q^2=0,q^2=0),$$ (3) where $`F_\pi 93`$ MeV. The sigma term can be evaluated accurately in terms of the physical (on-shell) amplitude using the well-known “on-shell theorem”. Since this result plays a central role in this, and other, determinations of the sigma term, a brief review of its origin is warranted. The amplitude $`\overline{D}(\nu _B=0,q^2,q^2)`$ is assumed to be slowly varying in $`q^2`$ and $`q^2`$ and hence $$\overline{D}(0,q^2,q^2)=\frac{\sigma (t)}{F_\pi ^2}\left[1+\frac{q^2+q^2}{\mu ^2}\right]+R(0,q^2,q^2),$$ (4) where $`R(\nu _B,q^2,q^2)`$ contains all non-linear dependence and this remainder function satisfies $$R(0,0,0)=R(0,\mu ^2,0)=R(0,0,\mu ^2)=0.$$ (5) The amplitude $`\overline{D}(0,q^2,q^2)`$ in (4) clearly satisfies the the Adler condition (2) and the sigma term definition (3). The remainder amplitude $`R`$ is small and will be evaluated by considering the nearby singular contributions primarily due to the $`\mathrm{\Delta }(1232)`$. The on-shell theorem follows by taking $`q^2=q^2=\mu ^2`$ (and hence $`t=2\mu ^2`$), $$F_\pi ^2\overline{D}(t=2\mu ^2)=\sigma (2\mu ^2)\mathrm{\Delta }_R$$ (6) where $$\mathrm{\Delta }_R=F_\pi ^2R(0,\mu ^2,\mu ^2)$$ (7) and by definition $`R`$ is of order $`\mu ^4`$. A simple model of the $`\mathrm{\Delta }(1230)`$ contribution to $`A^{(+)}(\nu =0,\nu _B=0,q^2,q^2)`$ is obtained through the successful effective Lagrangian $`\mathrm{\Delta }N\pi `$ coupling $$L=g_\mathrm{\Delta }\overline{\mathrm{\Delta }}_\mu \mathrm{\Theta }_{\mu \nu }N_\nu \varphi +\mathrm{h}.\mathrm{c}.,\mathrm{\Theta }_{\mu \nu }=\delta _{\mu \nu }(\frac{1}{2}+\mathrm{Z})\gamma _\mu \gamma _\nu ,$$ (8) yielding $$R(\nu _B=0,q^2,q^2)=\frac{2g_\mathrm{\Delta }^2(2M_\mathrm{\Delta }+M)q^2q^2}{9M_\mathrm{\Delta }^2(M_\mathrm{\Delta }^2M^2)},$$ (9) which automatically satisfies the conditions (5). Using $`g_\mathrm{\Delta }^2/4\pi 0.28\mu ^2`$ we obtain $$R(\nu _B=0,\mu ^2,\mu ^2)=7.5\times 10^3\mu ^1,$$ (10) which by (6) is equivalent to $$\sigma (2\mu ^2)=\left[F_\pi ^2\overline{D}(2\mu ^2)\right](0.46\mathrm{MeV}).$$ (11) The only uncertainty arises from the model dependence of $`R`$. Several evaluations agree on the result given in (10) and under very weak assumptions that $`|\mathrm{\Delta }_R|<2`$ MeV and is negligible compared to the sigma term. In this paper a new threshold method for finding $`\sigma (2\mu ^2)`$ is developed. The important question of determining $`\sigma (0)`$ is not addressed. In the literature there is some variation in results for the quantity $$\mathrm{\Delta }_\sigma =\sigma (2\mu ^2)\sigma (0).$$ (12) A recent one-loop ChPT estimate of the nucleon scalar form factor implies $`\mathrm{\Delta }_\sigma 5`$–7 MeV whereas a dispersive estimate gives $`\mathrm{\Delta }_\sigma 15`$ MeV. ## II The dispersive sum rule As outlined in the introductory section, the key quantity to evaluate is the on-shell amplitude $`\overline{D}(t=2\mu ^2)=A^{(+)}(\nu =0,t=2\mu ^2)g^2/M`$. Our evaluation involves the difference between two subtracted fixed-$`t`$ dispersion relations. Both start from the usual fixed momentum transfer dispersion relation for $`D^{(+)}(\nu ,t)=A^{(+)}(\nu ,t)+\nu B^{(+)}(\nu ,t)`$, $$eD^{(+)}(\nu ,t)=\frac{g^2}{M}\frac{\nu ^2}{\nu ^2\nu _B^2}+\frac{2}{\pi }P_{\nu _T}^{\mathrm{}}\frac{\nu ^{}d\nu ^{}}{\nu ^2\nu ^2}mD^{(+)}(\nu ^{},t),$$ (13) where $`\nu _T=\mu +t/4M`$. ### A Initial sigma term sum rule From (13) we find by evaluating at $`\nu =0`$ and $`\nu _T`$ and subtracting, $$\overline{D}(t)=D^{(+)}(\nu _T,t)\frac{g^2}{M}\frac{\nu _B^2}{\nu _T^2\nu _B^2}\frac{2}{\pi }_{\nu _T}^{\mathrm{}}\frac{d\nu }{\nu (\nu ^2\nu _T^2)}mD^{(+)}(\nu ,t).$$ (14) The PV singularity is not present since $`mD`$ vanishes at $`\nu =\nu _T`$. We now fix $`t=2\mu ^2`$ and change variable to $`d\nu =d\omega =kdk/\omega `$, where $`k`$ is the lab pion momentum $`k^2=\omega ^2\mu ^2`$, to obtain $$\overline{D}(2\mu ^2)=D^{(+)}(\nu _T,2\mu ^2)\frac{2\nu _T^2}{\pi }_0^{\mathrm{}}\frac{dk(\omega +\mu )mD^{(+)}(k,2\mu ^2)/k}{\omega (\omega +\mu r)(\omega +\mu +2\mu r)},$$ (15) where now $`\nu _T=\mu (1+r)`$ and $$r=\mu /2M0.0744.$$ (16) The Born term vanishes since $`\nu _B=0`$ at $`t=2\mu ^2`$. Although the integral converges at both limits we explicitly remove the threshold parameters by adding and subtracting the quantity $$\underset{k0}{lim}\frac{mD^{(+)}(\omega ,2\mu ^2)}{k}\lambda _2,$$ (17) yielding $`\mu \overline{D}(2\mu ^2)=\mu D^{(+)}(\nu _T,2\mu ^2)1.047\mu ^2\lambda _2I_0,`$ (18) $`I_0={\displaystyle \frac{2\mu ^3(1+r)^2}{\pi }}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dk(\omega +\mu )}{\omega (\omega +\mu r)(\omega +\mu +2\mu r)}}\left[{\displaystyle \frac{mD^{(+)}(\omega ,2\mu ^2)}{k}}\lambda _2\right],`$ (19) where we note that each term in (18) is dimensionless. ### B Forward dispersive identity Returning to (13) but now with momentum transfer $`t`$ fixed at $`t=0`$ we evaluate at threshold $`(\nu =\omega =\mu )`$ and above threshold ($`\omega >\mu `$) and take the difference, $$\frac{1}{k^2}\left[eD^{(+)}(\omega ,0)D^{(+)}(\mu ,0)\right]=\frac{g^2}{M}\frac{\omega _B^2}{(\omega ^2\omega _B^2)(\mu ^2\omega _B^2)}+\frac{2}{\pi }P_0^{\mathrm{}}\frac{dk^{}}{k^2k^2}\frac{mD^{(+)}(\omega ^{},0)}{k^{}},$$ (20) where $`\omega _B=\mu r`$. To remove the PV singularity we use the identity $$P_0^{\mathrm{}}\frac{dk^{}}{k^2k^2}=0;k^2>0$$ (21) and the definitions $`\underset{k0+}{lim}{\displaystyle \frac{mD^{(+)}(\omega ,0)}{k}}\lambda _0,`$ (22) $`\underset{k0+}{lim}{\displaystyle \frac{eD^{(+)}(\omega ,0)D^{(+)}(\mu ,0)}{k^2}}K.`$ (23) With the above definitions, and for $`k\mu `$, Eq. (20) becomes $$K\frac{g^2}{\mu ^2M}\left(\frac{r}{1r^2}\right)^2\frac{2}{\pi }_0^{\mathrm{}}\frac{dk}{k^2}\left[\frac{mD^{(+)}(\omega ,0)}{k}\lambda _0\right]=0.$$ (24) The integrand vanishes at threshold. ### C The final sum rule It might be noted that in the relativistic regime $`(\omega k)`$ the $`\mu \overline{D}(2\mu ^2)`$ integrand in (19) and $`\mu \nu _T^2K`$ in (24) are similar. To obtain our final sum rule we subtract $`\mu \nu _T^2=\mu ^3(1+r)^2`$ times the identity (24) from the $`\mu \overline{D}(2\mu ^2)`$ sum rule (18), giving $`\mu \overline{D}(2\mu ^2)`$ $`=`$ $`\mu \left[D^{(+)}(\nu _T,2\mu ^2)\nu _T^2K\right]1.047\mu ^2\lambda _2+{\displaystyle \frac{\mu g^2}{M}}\left({\displaystyle \frac{r}{1r^2}}\right)^2I,`$ (25) where $`I`$ $`=`$ $`{\displaystyle \frac{2\mu ^3}{\pi }}(1+r)^2{\displaystyle _0^{\mathrm{}}}dk\{(\omega +\mu ){\displaystyle \frac{\left[mD^{(+)}(\omega ,2\mu ^2)/k\lambda _2\right]}{\omega (\omega +\mu r)(\omega +\mu +2\mu r)}}`$ (27) $`{\displaystyle \frac{1}{k^2}}[mD^{(+)}(\omega ,0)/k\lambda _0]\}.`$ The various quantities $`D^{(+)}(\nu _T,2\mu ^2)`$, $`K`$, $`\lambda _2`$ and $`\lambda _0`$ can be evaluated in terms of threshold parameters. ## III Threshold Expansions The relation of the invariant amplitude to the partial-wave scattering amplitudes and phase shifts is standard. Using we have $$\frac{1}{4\pi }D^{(+)}(\nu ,t)=\frac{W}{M}(f_1+f_2)+\frac{t}{4MQ^2}\left[E(f_1+f_2)M(f_1f_2)\right],$$ (28) where $`2M\omega =W^2M^2\mu ^2`$, the c.m. momentum is $`Q=Mk/W`$ and the nucleon c.m. energy $`E=\sqrt{M^2+Q^2}`$. The quantities $`f_1`$ and $`f_2`$ are defined by $`f_1`$ $`=`$ $`{\displaystyle \underset{\mathrm{}=0}{\overset{\mathrm{}}{}}}f_\mathrm{}+^{(+)}P_{\mathrm{}+1}^{}(x)f_{\mathrm{}}^{(+)}P_\mathrm{}1^{}(x).`$ (29) $`f_2`$ $`=`$ $`{\displaystyle \underset{\mathrm{}=1}{\overset{\mathrm{}}{}}}(f_{\mathrm{}}^{(+)}f_\mathrm{}+^{(+)})P_{\mathrm{}}^{}(x).`$ (30) Here $`P_{\mathrm{}}^{}`$ are derivatives of Legendre polynomials with $`t=2Q^2(1x)`$. The above partial wave amplitudes $`f_\mathrm{}\pm ^{(+)}`$ are the crossing-even combination of the isospin amplitudes $`f_\mathrm{}\pm ^{(I)}`$, $$f_\mathrm{}\pm ^{(+)}=\frac{1}{3}f_\mathrm{}\pm ^{(1/2)}+\frac{2}{3}f_\mathrm{}\pm ^{(3/2)}$$ (31) and the isospin partial-wave amplitudes are $$f_\mathrm{}\pm ^{(I)}=\frac{1}{2iQ}\left(\eta e^{2i\delta }1\right)_{I,\mathrm{}\pm }\underset{\mathrm{elastic}}{}\left(\frac{e^{i\delta }\mathrm{sin}\delta }{Q}\right)_{I,\mathrm{}\pm }.$$ (32) Close to threshold we have $`\delta _{I,\mathrm{}\pm }`$ $``$ $`(a_\mathrm{}\pm ^{(I)})Q^{2\mathrm{}+1},`$ (33) $`{\displaystyle \frac{W}{M+\mu }}ef_{0+}^{(+)}`$ $``$ $`a_{0+}^{(+)}+{\displaystyle \frac{1}{3}}C^{(+)}k^2.`$ (34) The latter ($`s`$-wave expansion) is motivated by dispersion relations and the the static model where $`C^{(+)}`$ is energy independent. Using the above threshold expansion, the quantities needed in the sum rule are $$\frac{1}{4\pi }D^{(+)}(\nu _T,2\mu ^2)=\underset{\mathrm{}=0}{\overset{\mathrm{}}{}}\left(\beta _{\mathrm{}}a_{\mathrm{}}^{(+)}+\beta _\mathrm{}+a_\mathrm{}+^{(+)}\right)\mu ^2\mathrm{},$$ (35) where the first few factors $`\beta _\mathrm{}\pm `$ are given in Table I and the crossing-even scattering lengths are expressed, as in (31), in terms of the isospin scattering lengths. Corresponding expansions for the remaining quantities appearing in the sum rule (18) are $$\frac{\mu ^2}{4\pi }K=(1+2r)^1(a_1^{(+)}+2a_{1+}^{(+)})+\frac{1}{3}(1+2r)\mu C^{(+)}.$$ (36) Finally, we evaluate the two imaginary threshold quantities $`\lambda _0`$ and $`\lambda _2`$, $$\lambda _{0,2}=\underset{k0}{lim}\frac{mD^{(+)}(\nu _r,t=0\mathrm{or}2\mu ^2)}{k}.$$ (37) In each case only the $`\mathrm{}=0`$ partial wave enters and hence only $`f_1`$ contributes to $`D^{(+)}`$. Using (32) and $`QW=kM`$ we find $$\frac{1}{4\pi }\lambda _0=\frac{1}{3}\left(a_{0+}^{(1/2)}\right)^2+\frac{2}{3}\left(a_{0+}^{(3/2)}\right)^2.$$ (38) and $$\lambda _2=\frac{(1+r)^2}{1+2r}\lambda _0=1.036\lambda _0.$$ (39) We note that $`\lambda _2`$ is the same as $`\lambda _0`$ to better than 4%. It should be noted that the amplitude combination $`D=A+\nu B`$ is unique in giving a threshold expansion without $`p`$-wave effective ranges in $`K`$. As one sees from (28) the $`1/Q^2`$ term vanishes at $`t=0`$. ## IV Evaluation In order to evaluate the final sum rule (25) we first consider the integral (27). The integration is straightforward in terms of the partial wave amplitude (28) once the phase shifts and absorption factors are given. We use the very convenient VPI/GW partial wave analysis On Line. In Fig. 1 we show the integrand of (27) with thirteen crossing-even partial waves $`\mathrm{}\pm =0+,1\pm ,2\pm ,3\pm ,4\pm ,5\pm ,6\pm `$. Large cancellation between the two parts of the integrand reduce it by nearly an order of magnitude. It is evident that the $`\mathrm{\Delta }(1230)`$ resonance dominates and the integrated result for $`I`$ is $$I0.21\pm 0.02.$$ (40) The error is less than 10% and is negligible compared to other errors. Since the partial wave analysis extends from threshold to about $`k=2`$ GeV some estimate must be made for the contribution above this limit. Two approaches were tried which gave similar results. The integrand in (27) for $`\omega \mu `$ becomes $$\frac{2}{\pi }(1+r)^2\left[\frac{\mathrm{\Delta }D^{(+)}}{k^2}\frac{(\lambda _2\lambda _0)}{k^2}\right],$$ (41) where $$\mathrm{\Delta }D^{(+)}=mD^{(+)}(\omega ,2\mu ^2)mD^{(+)}(\omega ,0).$$ (42) The $`(\lambda _2\lambda _0)`$ part is very small because of (38) and (39) and will be ignored. We make two estimates for $`\mathrm{\Delta }D^{(+)}`$ in the asymptotic region. In the first we use the phase shift analysis to give $`\mathrm{\Delta }D^{(+)}(\omega _0)`$ at $`\omega _02`$ GeV. Then the asymptotic contribution is assume to be $`\mathrm{\Delta }D^{(+)}(\omega )=\frac{\omega }{\omega _0}\mathrm{\Delta }D^{(+)}(\omega _0)`$. This would be expected for diffractive (Pomeron) dominance. Alternatively we used a comprehensive Regge pole analysis to provide $`\mathrm{\Delta }D^{(+)}`$. This approach incorporates $`P`$, $`P^{}`$ and $`P^{\prime \prime }`$ Regge exchanges. Both estimates give $$I_{\mathrm{ASY}}0.01,$$ (43) which has been incorporated into the result (40). We next consider the nucleon Born term in (25) which, with $`g^2/4\pi 14`$, largely cancels the continuum part (40) above to give $$\frac{\mu g^2}{M}\left(\frac{r}{1r^2}\right)^2I0.06\pm 0.02.$$ (44) As a final step we separate out the small contributions of the $`\mathrm{}2`$ scattering lengths. We use the Koch dispersive analysis for these higher scattering lengths to write (35) as $$D^{(+)}(\nu _T,2\mu ^2)=4\pi \left((1+r)^2a_{0+}^{(+)}+a_1^{(+)}+\left[3(1+r)^21\right]a_{1+}^{(+)}\right)(0.02\pm 0.02).$$ (45) Here and subsequently we assume threshold parameters are in charged pion mass units. Combining the results (36), (44) and (45) with our general sum rule (25) yields $`{\displaystyle \frac{1}{4\pi }}\overline{D}(2\mu ^2)`$ $`=`$ $`(1+r)^2\left[a_{0+}^{(+)}{\displaystyle \frac{1.047}{3}}\left(a_{0+}^{(1/2)}\right)^2{\displaystyle \frac{2}{3}}(1.047)\left(a_{0+}^{(3/2)}\right)^2{\displaystyle \frac{1}{3}}(1+2r)C^{(+)}\right]`$ (47) $`{\displaystyle \frac{r^2}{1+2r}}a_1^{(+)}+{\displaystyle \frac{r}{1+2r}}(6+13r+6r^2)a_{1+}^{(+)}{\displaystyle \frac{(0.08\pm 0.03)}{4\pi }}.`$ Using the value $`r=0.0744`$ from (16) we find (in units of $`\mu `$) $`\overline{D}(2\mu ^2)`$ $`=`$ $`14.5a_{0+}^{(+)}5.06\left(a_{0+}^{(1/2)}\right)^210.13\left(a_{0+}^{(3/2)}\right)^25.55C^{(+)}`$ (49) $`0.06a_1^{(+)}+5.70a_{1+}^{(+)}(0.08\pm 0.03).`$ We see that if the $`s`$-wave and $`p`$-wave scattering lengths and the effective range $`C^{(+)}`$ are known the sigma term $`\sigma (2\mu )`$ then follows by (11). We note that the last term in (49), which represents the net contributions of Born, continuum, and higher partial wave scattering lengths, is small. The near cancellation of the coefficient of $`a_1^{(+)}`$ also makes it irrelevant. It turns out that $`a_{1+}^{(+)}`$ and $`C^{(+)}`$ are the dominant terms. Although $`a_{0+}^{(+)}`$ does not make a large contribution due to its small magnitude, it must be accurately determined because of its large weighting. A recent analysis parametrizes the very low energy $`\pi ^\pm p`$ data with a $`K`$-matrix expansion. Applying this method yields the $`s`$\- and $`p`$-wave scattering lengths as well as the $`s`$-wave effective range parameters $`C^{(\pm )}`$. The threshold parameters which are needed in our sum rule (49) are given (private communication): $`a_{0+}^{(+)}`$ $`=`$ $`0.0035\pm 0.0026,`$ (50) $`a_{0+}^{(1/2)}`$ $`=`$ $`0.179\pm 0.006,`$ (51) $`a_{0+}^{(3/2)}`$ $`=`$ $`0.079\pm 0.003,`$ (52) $`C^{(+)}`$ $`=`$ $`0.116\pm 0.026,`$ (53) $`a_1^{(+)}`$ $`=`$ $`0.063\pm 0.003,`$ (54) $`a_{1+}^{(+)}`$ $`=`$ $`0.135\pm 0.002.`$ (55) Inserting these into (49) and using (11) we obtain $$\sigma (2\mu ^2)=71\pm 9\mathrm{MeV}.$$ (56) The dominant error is from the effective range $`C^{(+)}`$. To show the results of earlier sets of threshold parameters, we first consider the very old set of Hamilton and Woolcock. This set gives $`\sigma _{\mathrm{HW}}(2\mu ^2)=64\pm 9`$ MeV. A slightly more recent set due to Koch yields $`\sigma _\mathrm{K}(2\mu ^2)=55\pm 6`$ MeV. The slightly smaller values of $`\sigma _{\mathrm{HW}}`$ and $`\sigma _\mathrm{K}`$ can be ascribed to smaller values of $`C^{(+)}`$ and the change in sign of $`a_{0+}^{(+)}`$. ## V Conclusions Combining subtracted $`t=2\mu ^2`$ and $`t=0`$ dispersion relations allows the amplitude $`\overline{D}(2\mu ^2)`$ to be determined by threshold parameters alone. By the “on-shell theorem” (11) the sigma term $`\sigma (2\mu )`$ then follows. The method presented here is an improvement on that used in . The central results obtained here are: 1. We demonstrate the relation of the sigma term $`\sigma (2\mu )`$ to threshold parameters. 2. Recent data and analyses determine the scattering lengths and crossing-even effective range using only data near to threshold $`(T_\pi <100`$ MeV). We use these results to compute $`\sigma (2\mu ^2)=71\pm 9`$ MeV. The assignment of error is straightforward using (49). This is in contrast to previous determinations using a global phase shift analysis with dispersive constraints where the error is difficult to determine. 3. The value of $`\sigma (2\mu ^2)`$ is not sensitive to $`g_{\pi NN}`$ but only depends on quantities close to the observables. 4. Our threshold method finds $`\sigma (2\mu ^2)`$ in agreement with recent global calcuations. The discrepancy between experiment and conventional ChPT theory is reinforced. ## Acknowledgments This research was supported in part by the U.S. Department of Energy under Grant No. DE-FG02-95ER40896 and in part by the University of Wisconsin Research Committee with funds granted by the Wisconsin Alumni Research Foundation.
warning/0001/gr-qc0001083.html
ar5iv
text
# Exact 𝑈⁢(1) symmetric cosmologies with local Mixmaster dynamics e-mail: berger@oakland.edu, vincent.moncrief@yale.edu ## I Introduction Recent numerical studies have provided strong evidence that $`U(1)`$-symmetric, vacuum spacetimes on $`T^3\times 𝐑`$ generically develop Mixmaster-like, oscillatory singularities of the type predicted long ago by Belinskii, Khalatnikov and Lifschitz (BKL) . These results confirm numerically some of the most surprising features of the BKL prediction, namely that nearby spatial points are effectively decoupled in their asymptotic metric evolution and that the metric variables at each of these points evolve, at least qualitatively, like those of a Mixmaster spacetime. Several years ago B. Grubišić and one of us (V.M.) made an analytical effort to generate some exact vacuum spacetimes which were spatially inhomogeneous and which were expected to exhibit the sort of oscillatory singularities which have since been seen in the numerical studies . That effort was not completed at the time since it was not realized that several seemingly intractable integrals actually cancel in the course of the calculations leaving only elementary computations to be done. We shall therefore complete that project here and use the results to compare, in a more quantitative way, the numerical results with some exact oscillatory singularities. To generate new solutions having Mixmaster-like oscillations, we begin with the actual Mixmaster solutions and apply a standard solution generating technique. We choose one of the Killing fields shared by the Mixmaster family and treat it as the generator of a spacelike $`U(1)`$ action on $`S^3\times 𝐑`$, ignoring the presence of the other Killing symmetries. We compute the twist potential associated with the chosen Killing field and reexpress the field equations, in a well-known way , as a Kaluza-Klein reduced system on the base manifold $`S^2\times 𝐑`$ of the $`S^1`$-bundle $`S^3\times 𝐑S^2\times 𝐑`$. The field equations on the base take the form of $`2+1`$ Einstein gravity coupled to a wave map whose target space is the hyperbolic plane, $`𝐇^2`$. The isometry group of this latter space, $`SL(2,𝐑)`$, acts on the base fields in a natural way so as to transform the given solution to a family of potentially inequivalent solutions. By a careful choice of the applied group element one can arrange that the transformed solution either lifts to the same bundle defined for the original spacetime or perhaps to a different one (e.g., the trivial bundle,$`S^2\times S^1\times 𝐑S^2\times 𝐑`$, or a “squashed sphere”, $`S^3/𝐙_k\times 𝐑S^2\times 𝐑`$). Typically, the new solutions will preserve only the Killing field that generates the common $`U(1)`$ action and not preserve those Killing fields of the seed solutions which fail to commute with the chosen $`U(1)`$ generator. Thus the new solutions are expected to be spatially inhomogeneous and yet to exhibit Mixmaster-like oscillations inasmuch as their metrics are parametrized by the same functions appearing in the Mixmaster seed metrics themselves. A previous application of this technique involved transforming an infinite dimensional family of “generalized Taub-NUT” spacetimes defined on $`S^3\times 𝐑`$, which have smooth Cauchy horizons at their “singular” boundaries, to a new family of curvature singular spacetimes defined on $`S^2\times S^1\times 𝐑`$ . Because of the special nature of the seed solutions in this case, the transformed solutions developed only velocity dominated singularities and never exhibited Mixmaster-like oscillations. A new technique based upon expressing the Einstein evolution equations in a so-called Fuchsian form seems capable of significantly enlarging this set of rigorous, $`U(1)`$-symmetric, curvature singular cosmological spacetimes but, so far, is also only capable of yielding velocity dominated singularities . So far as we know the solutions presented for the first time here are the only known exact inhomogeneous vacuum spacetimes which exhibit Mixmaster oscillations. Though only a finite dimensional family they presumably display behavior representative of more general, $`U(1)`$-symmetric vacuum spacetimes and thus warrant comparison with numerically produced $`U(1)`$-solutions. Making such a comparison is the second main aim, after producing the solutions themselves, of this paper. As a byproduct of this work, we also resolve a potential paradox that was pointed out in Ref. . There it was shown that every $`U(1)`$-symmetric vacuum spacetime admits a certain gauge invariant conserved quantity which is expressible purely locally in terms of the instantaneous Cauchy data for that solution and serves as a Casimir invariant for the $`SL(2,𝐑)`$ action. For generic $`U(1)`$ solutions this quantity is known to be non-trivial but, if non-trivial for the Mixmaster subfamily, would seem to contradict the anticipated “chaos” of the Mixmaster dynamics . The only sensible resolution, as was discussed in Ref. , is that the quantity actually vanishes on the Mixmaster subfamily. This we find to be the case by explicit calculation. The inhomogeneity in our transformed solutions is produced, roughly speaking, by the fact that we choose to reduce with respect to a Killing field which fails to commute with the remaining Killing fields of the seed metric. This is unavoidable with the generic Mixmaster solution but special cases such as the Taub-NUT metrics allow for different possibilities. The additional Killing field admitted by Taub space commutes with all the generators and is preserved upon reduction with respect to one of these (non-abelian) generators. The resulting spacetime has therefore (at least) two commuting Killing fields and is thus a special case of the so-called Gowdy family of spacetimes. By contrast one could instead choose to reduce with respect to the additional, commutative Killling field but, in this case, all the symmetries are preserved and one arrives, as was first shown by Geroch , at only the Kantowski-Sachs (i.,e. locally interior Schwarzschild) spacetime. One might wonder if the “new” solutions we produce are really inhomogeneous at all or perhaps because of their expression in an unusual gauge, are merely homogeneous solutions in disguise. We shall use the Gowdy transform of Taub space mentioned above, to show that this is not the case—the new solutions are not in general globally homogeneous. ## II Mixmaster spacetimes The Mixmaster spacetimes are spatially homogeneous vacuum metrics on $`S^3\times 𝐑`$ whose line elements can be written $$ds^2=N^2(t)dt^2+A^2(t)(\widehat{\sigma }^1)^2+B^2(t)(\widehat{\sigma }^2)^2+C^2(t)(\widehat{\sigma }^3)^2.$$ (1) Here the $`\{\widehat{\sigma }^i\}`$ are a global, analytic basis of one-forms on $`S^3`$ expressible in terms of the usual Euler angle coordinates $`\{x^1,x^2,x^3\}=\{\theta ,\phi ,\psi \}\{[0,\pi ),[0,2\pi ),[0,4\pi )\}`$ by $`\widehat{\sigma }^1`$ $`=`$ $`\mathrm{cos}\phi d\theta +\mathrm{sin}\theta \mathrm{sin}\phi d\psi ,`$ (2) $`\widehat{\sigma }^2`$ $`=`$ $`\mathrm{sin}\phi d\theta +\mathrm{sin}\theta \mathrm{cos}\phi d\psi ,`$ (3) $`\widehat{\sigma }^3`$ $`=`$ $`d\phi +\mathrm{cos}\theta d\psi .`$ (4) These forms, and therefore the above line element, are invariant with respect to the $`U(1)`$ action on $`S^3`$ generated by the Killing field $`\widehat{X}_3=\frac{}{\psi }`$ whose orbits yield a Hopf fibration of $`S^3`$, i.e. make $`S^3`$ into a principal fiber bundle over $`S^2`$ with bundle projection given by $$\pi _\psi :S^3S^2,(\theta ,\phi ,\psi )(\theta ,\phi ).$$ (5) Of course the Mixmaster metrics are invariant with respect to a full $`SU(2)`$ action generated by Killing fields $`\widehat{X}_1`$ $`=`$ $`\mathrm{cos}\psi {\displaystyle \frac{}{\theta }}+\text{csc}\theta \mathrm{sin}\psi {\displaystyle \frac{}{\phi }}\text{cot}\theta \mathrm{sin}\psi {\displaystyle \frac{}{\psi }},`$ (6) $`\widehat{X}_2`$ $`=`$ $`\mathrm{sin}\psi {\displaystyle \frac{}{\theta }}+\text{csc}\theta \mathrm{cos}\psi {\displaystyle \frac{}{\phi }}\text{cot}\theta \mathrm{cos}\psi {\displaystyle \frac{}{\psi }},`$ (7) $`\widehat{X}_3`$ $`=`$ $`{\displaystyle \frac{}{\psi }}`$ (8) but, for the transformations we shall consider, only invariance with respect to $`\widehat{X}_3=\frac{}{\psi }`$ will in general be preserved. The equations of motion for the Mixmaster solutions are most simply expressed in a gauge for which $`N=ABC`$ where they take the form $`(\mathrm{ln}A^2),_{tt}`$ $`=`$ $`(B^2C^2)^2A^4,`$ (9) $`(\mathrm{ln}B^2),_{tt}`$ $`=`$ $`(C^2A^2)^2B^4,`$ (10) $`(\mathrm{ln}C^2),_{tt}`$ $`=`$ $`(A^2B^2)^2C^4,`$ (11) and are to be supplemented by the Hamiltonian constraint $`{\displaystyle \frac{A_{,t}}{A}}{\displaystyle \frac{B_{,t}}{B}}+{\displaystyle \frac{A_{,t}}{A}}{\displaystyle \frac{C_{,t}}{C}}+{\displaystyle \frac{B_{,t}}{B}}{\displaystyle \frac{C_{,t}}{C}}+`$ (12) $`{\displaystyle \frac{1}{4}}[A^4+B^4+C^42(A^2B^2+B^2C^2+A^2C^2)]0.`$ (13) In terms of the Misner anistropy variables $`\alpha ,\beta _+,\beta _{}`$, $`A`$ $`=`$ $`e^{\alpha +\beta _++\sqrt{3}\beta _{}},`$ (14) $`B`$ $`=`$ $`e^{\alpha +\beta _+\sqrt{3}\beta _{}},`$ (15) $`C`$ $`=`$ $`e^{\alpha 2\beta _+},`$ (16) and the chosen gauge condition is $`N=e^{3\alpha }`$. We now rewrite the line element in the $`U(1)`$-symmetric form developed in Refs. and . Taking $`\{x^a\}=\{\theta ,\phi \}`$ and noting that the shift vector vanishes we express $`ds^2`$ in the form $`ds^2=e^{2\gamma }\{\stackrel{~}{N}^2dt^2+\stackrel{~}{g}_{ab}dx^adx^b\}`$ (17) $`+e^{2\gamma }\{d\psi +\mathrm{cos}\theta d\phi +\beta _adx^a\}^2`$ (18) where $`e^{2\gamma }`$ is the scalar field $`\frac{}{\psi }\frac{}{\psi }`$ given explicitly by $`e^{2\gamma }`$ $`=`$ $`A^2\mathrm{sin}^2\theta \mathrm{sin}^2\phi +B^2\mathrm{sin}^2\theta \mathrm{cos}^2\phi `$ (20) $`+C^2\mathrm{cos}^2\theta .`$ Since $`\gamma `$ is invariant with respect to the $`U(1)`$ action generated by $`\widehat{X}_3=\frac{}{\psi }`$ it induces a function on the quotient manifold $`S^3\times 𝐑/U(1)S^2\times 𝐑`$ which (with a slight abuse of notation) we shall also designate by $`\gamma `$. In a similar way one finds induced upon the quotient manifold $`S^2\times 𝐑`$ a Lorentzian metric $$d\sigma ^2=\stackrel{~}{N}^2dt^2+\stackrel{~}{g}_{ab}dx^adx^b$$ (21) and a one-form field $$\underset{}{𝛽}=\beta _adx^a$$ (22) where $`\{a,b,\mathrm{}\}=\{1,2\}`$. These forms are slightly specialized because of the vanishing of the shift vector field in $`ds^2`$. The most general $`U(1)`$-symmetric line element would yield $`d\sigma ^2`$ $`=`$ $`\stackrel{~}{N}^2dt^2+\stackrel{~}{g}_{ab}(dx^a+\stackrel{~}{N}^adt)(dx^b+\stackrel{~}{N}^bdt)`$ (23) $`\underset{}{𝛽}`$ $`=`$ $`\beta _adx^a+\beta _0dt.`$ (24) The explicit formulas for the $`2+1`$ Lorentzian metric $`d\sigma ^2`$ and the one-form potential $`\underset{}{𝛽}`$ may be read off upon expressing $`ds^2`$ in the form of Eq. (1). One finds that $`\stackrel{~}{N}`$ $`=`$ $`Ne^\gamma ,`$ (25) $`\stackrel{~}{g}_{\theta \theta }`$ $`=`$ $`C^2\mathrm{cos}^2\theta (A^2\mathrm{cos}^2\phi +B^2\mathrm{sin}^2\phi )`$ (27) $`+A^2B^2\mathrm{sin}^2\theta ,`$ $`\stackrel{~}{g}_{\phi \phi }`$ $`=`$ $`C^2\mathrm{sin}^2\theta (A^2\mathrm{sin}^2\phi +B^2\mathrm{cos}^2\phi ),`$ (28) $`\stackrel{~}{g}_{\theta \phi }`$ $`=`$ $`C^2(A^2B^2)\mathrm{cos}\phi \mathrm{sin}\phi \mathrm{cos}\theta \mathrm{sin}\theta ,`$ (29) $`\beta _\theta `$ $`=`$ $`{\displaystyle \frac{(A^2B^2)\mathrm{cos}\phi \mathrm{sin}\phi \mathrm{sin}\theta }{e^{2\gamma }}},`$ (30) $`\beta _\phi `$ $`=`$ $`({\displaystyle \frac{C^2\mathrm{cos}\theta }{e^{2\gamma }}}\mathrm{cos}\theta ),`$ (31) and computes, for example, that $$\sqrt{det{}_{}{}^{(2)}\stackrel{~}{g}}=ABC\mathrm{sin}\theta e^\gamma .$$ (32) As in Refs. and we introduce the momenta $`\{\stackrel{~}{p},\stackrel{~}{e}^a,\stackrel{~}{\pi }^{ab}\}`$ conjugate to $`\{\gamma ,\beta _a,\stackrel{~}{g}_{ab}\}`$, which, taken together, parameterize the full $`3+1`$ spatial metric $`g_{ij}`$ and its conjugate momentum $`\pi ^{ij}`$. For the case of vanishing shift the formulas relating the momentum variables $`\{\stackrel{~}{p},\stackrel{~}{e}^a,\stackrel{~}{\pi }^{ab}\}`$ to the metric variables $`\{\gamma ,\beta _a,\stackrel{~}{g}_{ab}\}`$ are given by $`\stackrel{~}{p}`$ $`=`$ $`\left({\displaystyle \frac{\sqrt{{}_{}{}^{(2)}\stackrel{~}{g}}}{\stackrel{~}{N}}}\right)4\gamma _{,t},`$ (33) $`\stackrel{~}{e}^a`$ $`=`$ $`\left({\displaystyle \frac{\sqrt{{}_{}{}^{(2)}\stackrel{~}{g}}}{\stackrel{~}{N}}}\right)e^{4\gamma }\stackrel{~}{g}^{ab}\beta _{b,t},`$ (34) $`\stackrel{~}{\pi }^{ab}`$ $`=`$ $`\left({\displaystyle \frac{\sqrt{{}_{}{}^{(2)}\stackrel{~}{g}}}{\stackrel{~}{N}}}\right){\displaystyle \frac{1}{2}}(\stackrel{~}{g}^{ac}\stackrel{~}{g}^{bd}\stackrel{~}{g}^{ab}\stackrel{~}{g}^{cd})\stackrel{~}{g}_{cd,t}.`$ (35) For the Mixmaster metrics one computes that $`\stackrel{~}{e}^\theta ={\displaystyle \frac{2\mathrm{sin}^2\theta }{N}}\{BCA_{,t}\mathrm{sin}\phi \mathrm{cos}\phi ACB_{,t}\mathrm{sin}\phi \mathrm{cos}\phi \},`$ (36) $`\stackrel{~}{e}^\phi ={\displaystyle \frac{2\mathrm{sin}\theta \mathrm{cos}\theta }{N}}\{ABC_{,t}BCA_{,t}\mathrm{sin}^2\phi ACB_{,t}\mathrm{cos}^2\phi \},`$ (37) $`\stackrel{~}{p}={\displaystyle \frac{4ABC\mathrm{sin}\theta }{Ne^{2\gamma }}}[AA_{,t}\mathrm{sin}^2\theta \mathrm{sin}^2\phi +BB_{,t}\mathrm{sin}^2\theta \mathrm{cos}^2\phi +CC_{,t}\mathrm{cos}^2\theta ].`$ (38) These momenta (along with $`\stackrel{~}{\pi }^{ab}`$ which we shall not need explicitly) project to yield smooth tensor densities on the base manifold $`S^2\times 𝐑`$ and one easily verifies that $`\stackrel{~}{e}_{,a}^a=0`$ which is one of the components of the $`3+1`$ momentum constraint. Note that the $`U(1)`$ connection one-form on $`S^3\times 𝐑`$ given by $$\underset{}{𝜆}:=\widehat{\omega }^3+\underset{}{𝛽}=d\psi +\mathrm{cos}\theta d\phi +\beta _adx^a$$ (39) does not project to yield a one-form on the base but that the difference between this connection one form and the reference one form $`\widehat{\omega }^3:=d\psi +\mathrm{cos}\theta d\phi `$ does yield a one-form (namely $`\underset{}{𝛽}=\beta _adx^a`$) which projects to the base. Even though $`\underset{}{𝜆}`$ itself does not project to the base, its exterior derivative $`\underset{}{d\lambda }`$ (i.e., the curvature of the connection $`\underset{}{𝜆}`$) does project. Pulling back the induced two-form to a $`t`$ = constant slice of the base manifold and computing its dual, one gets a scalar density $`\stackrel{~}{r}`$ defined by $`\stackrel{~}{r}`$ $`=`$ $`^{ab}\lambda _{a,b}`$ (40) $`=`$ $`^{ab}\beta _{a,b}+\mathrm{sin}\theta `$ (41) whose explicit form is $`\stackrel{~}{r}`$ $`=`$ $`{\displaystyle \frac{(A^2B^2)}{(e^{2\gamma })^2}}\mathrm{sin}^3\theta [B^2\mathrm{cos}^2\phi A^2\mathrm{sin}^2\phi ]`$ (43) $`+{\displaystyle \frac{C^2\mathrm{sin}\theta }{(e^{2\gamma })^2}}[A^2+B^2C^2+\mathrm{sin}^2\theta (C^2A^2\mathrm{cos}^2\phi B^2\mathrm{sin}^2\phi )].`$ One computes on an arbitrary $`t=`$ constant slice of the base manifold, that $$\widehat{B}:=_{S^2}\stackrel{~}{r}=4\pi .$$ (44) The value $`4\pi `$ reflects the particular bundle $`S^3\times 𝐑S^2\times 𝐑`$ under study and would be the same for any $`U(1)`$-symmeteric metric defined on this bundle. Taking into account the equation $`\stackrel{~}{e}_{,a}^a=0`$ satisfied by $`\stackrel{~}{e}^a`$ and the fact that $`S^2\times 𝐑`$ admits no non-trivial harmonic one forms, we now introduce the “twist potential” function $`\omega `$ (a salar field on $`S^2\times 𝐑`$) by imposing $`\stackrel{~}{e}^a`$ $`=`$ $`^{ab}\omega _{,b},`$ (45) $`\stackrel{~}{r}`$ $`=`$ $`{\displaystyle \frac{\sqrt{{}_{}{}^{(2)}\stackrel{~}{g}}}{\stackrel{~}{N}}}e^{4\gamma }\omega _{,t}.`$ (46) These equations are self consistent and yield the solution $`\omega `$ $`=`$ $`{\displaystyle \frac{\mathrm{sin}^2\theta }{N}}\{ABC_{,t}+BCA_{,t}\mathrm{sin}^2\phi +ACB_{,t}\mathrm{cos}^2\phi \}`$ (48) $`+k(t)`$ which is unique up to the additive constant $`k_0:=k(t_0)`$ where $`k(t)`$ is the function defined by $$k(t)=k(t_0)+_{t_0}^t𝑑t^{}\left(\frac{N}{ABC}\right)C^2(A^2+B^2C^2).$$ (49) As discussed in Refs. and the fields $`\{\gamma ,\omega ,d\sigma ^2=\stackrel{~}{N}^2dt^2+\stackrel{~}{g}_{ab}dx^adx^b\}`$ induced upon the base manifold $`S^2\times 𝐑`$ satisfy a $`2+1`$ dimensional system of Einstein-wave map equations for which the target space of the wave map is hyperbolic two-space (endowed with global coordinates $`\{\gamma ,\omega \}`$ and the natural metric $`dh^2=4d\gamma ^2+e^{4\gamma }d\omega ^2`$). As a consequence of the $`SL(2,𝐑)`$ isometry group of this target space the Einstein-wave map system admits three independent constants of the motion which serve as the Hamiltonian generators of the action of $`SL(2,𝐑)`$ on the phase space of fields $`\{\gamma ,\stackrel{~}{p},\omega ,\stackrel{~}{r},\stackrel{~}{g}_{ab},\stackrel{~}{\pi }^{ab}\}`$. These conserved quantities are given explicitly by the integrals $`\widehat{A}`$ $`:=`$ $`{\displaystyle _{S^2}}(2\omega \stackrel{~}{r}+\stackrel{~}{p}),`$ (50) $`\widehat{B}`$ $`:=`$ $`{\displaystyle _{S^2}}\stackrel{~}{r},`$ (51) $`\widehat{C}`$ $`:=`$ $`{\displaystyle _{S^2}}(\stackrel{~}{r}(e^{4\gamma }\omega ^2)\stackrel{~}{p}\omega ),`$ (52) and we have already noticed that $`\widehat{B}=4\pi `$ for the Mixmaster spacetimes in particular. In fact $`\widehat{B}`$ would take this same value for any $`U(1)`$-symmetric vacuum metric on $`S^3\times 𝐑`$ but for other $`S^1`$ bundles over the same base the value would (as discussed in Refs. and ) be modified to $`\widehat{B}=4\pi n`$ where $`n`$ is an integer determining the Chern class of the bundle. In particular, for solutions on the trivial bundle $`S^2\times S^1\times 𝐑`$, $`n`$ would vanish whereas if $`n=2,3,\mathrm{},`$ the bundle would correspond to various “squashed spheres” rather that a true $`S^3`$. Note that, in view of the integral expression for $`k(t)`$ arising in the formula for $`\omega `$, both $`\widehat{A}`$ and $`\widehat{C}`$ are non-local in time. The same feature occurs in more general $`U(1)`$ symmetric solutions but this non-locality cancels from the Casimir invariant $$\widehat{K}:=\widehat{A}^2+4\widehat{B}\widehat{C}$$ (53) which, however, vanishes identically for the Mixmaster family of solutions (though not in general). The vanishing of $`\widehat{K}`$ resolves a potential mystery pointed out in Ref. whereby a non-vanishing, local, constant of the motion for Mixmaster metrics would seem to contradict their empirically observed “chaotic” properties. More specifically one finds, for the Mixmaster metrics, that $`\widehat{A}`$ $`=`$ $`8\pi k(t)+8\pi \left({\displaystyle \frac{ABC}{N}}\right)\left({\displaystyle \frac{A_{,t}}{A}}+{\displaystyle \frac{B_{,t}}{B}}\right),`$ (54) $`\widehat{B}`$ $`=`$ $`4\pi ,`$ (55) $`\widehat{C}`$ $`=`$ $`{\displaystyle \frac{\widehat{A}^2}{16\pi }},`$ (56) so that $`\widehat{K}=0`$. The non-locality of $`\widehat{A}`$ and $`\widehat{C}`$ sidesteps any conflict with the observed “chaos” in Mixmaster solutions since, in fact, any Hamiltonian system will admit such non-local constants of the motion. To see this (even for a chaotic system) simply time integrate Hamilton’s equations and express the initial values of the canonical variables in terms of time integrals of their driving “forces.” ## III The new solutions To generate new solutions of Einstein’s equations from a given one (such as a Mixmaster solution) we choose an element $`gSL(2,𝐑)`$, $$g=\left(\begin{array}{cc}a& b\\ c& d\end{array}\right),adbc=1$$ (57) and transform the fields $`\{\gamma ,\omega ,\stackrel{~}{p},\stackrel{~}{r}\}`$ according to $`e^{2\gamma _g}`$ $`=`$ $`{\displaystyle \frac{e^{2\gamma }}{[c^2(\omega ^2+e^{4\gamma })+2cd\omega +d^2]}},`$ (58) $`\omega _g`$ $`=`$ $`{\displaystyle \frac{ac(\omega ^2+e^{4\gamma })+(ad+bc)\omega +bd}{[c^2(\omega ^2+e^{4\gamma })+2cd\omega +d^2]}},`$ (59) $`\stackrel{~}{p}_g`$ $`=`$ $`{\displaystyle \frac{\{\stackrel{~}{p}[c^2(\omega ^2e^{4\gamma })+2cd\omega +d^2]\stackrel{~}{r}[4e^{4\gamma }(cd+\omega c^2)]\}}{[c^2(\omega ^2+e^{4\gamma })+2cd\omega +d^2]}},`$ (60) $`\stackrel{~}{r}_g`$ $`=`$ $`\stackrel{~}{p}(c^2\omega +cd)+\stackrel{~}{r}[d^2+c^2(\omega ^2e^{4\gamma })+2cd\omega ],`$ (61) while leaving $`\{\stackrel{~}{g}_{ab},\stackrel{~}{\pi }^{ab},\stackrel{~}{N},\stackrel{~}{N}^a\}`$ invariant. The induced transformation of the conserved quantities $`\widehat{A},\widehat{B},\widehat{C}`$ (by the so-called co-adjoint action of $`SL(2,𝐑)`$)is found to be $`\widehat{A}_g`$ $`=`$ $`(ad+bc)\widehat{A}+2bd\widehat{B}2ac\widehat{C},`$ (62) $`\widehat{B}_g`$ $`=`$ $`d^2\widehat{B}c^2\widehat{C}+cd\widehat{A},`$ (63) $`\widehat{C}_g`$ $`=`$ $`a^2\widehat{C}b^2\widehat{B}ab\widehat{A},`$ (64) $`\widehat{K}_g`$ $`=`$ $`\widehat{K}=(\widehat{A}_g)^2+4\widehat{B}_g\widehat{C}_g.`$ (65) To avoid a trivial transformation we shall require that $`c`$ be non-zero and, to ensure that the transformed solution lifts to an $`S^1`$ bundle over$`S^2\times 𝐑`$, we shall demand that $$\widehat{B}_g=4\pi n,n=0,1,2,\mathrm{}.$$ (66) Defining $$\mathrm{}(t):=\frac{ABC}{N}\left(\frac{A,_t}{A}+\frac{B,_t}{B}\right)$$ (67) we see from Eq. (57) that $`\widehat{A}`$ $`=`$ $`8\pi k(t)+8\pi \mathrm{}(t)=8\pi k(t_0)+8\pi \mathrm{}(t_0)=8\pi (k_0+\mathrm{}_0),`$ (68) $`\widehat{B}`$ $`=`$ $`4\pi ,`$ (69) $`\widehat{C}`$ $`=`$ $`4\pi (k_0+\mathrm{}_0)^2.f`$ (70) Setting $`\widehat{B}_g=4\pi n,n=0,1,2,\mathrm{}`$ gives the restriction $$[d+c(k_0+\mathrm{}_0)]^2=n0$$ (71) or, equivalently, $$d+c(k_0+\mathrm{}_0)=\pm n^{1/2}$$ (72) which can always be solved for $`k_0`$ since $`c0`$. Exploiting the fact that $`\widehat{A}`$, hence also $`k(t)+\mathrm{}(t)`$, is conserved one finds that the integral occurring in the formula for $`k(t)`$ can be expressed as $$_{t_0}^t𝑑t^{}\left(\frac{N}{ABC}\right)C^2(A^2+B^2C^2)=(\mathrm{}(t)\mathrm{}_0)$$ (73) (this is also easily verified upon differentiation by using the equations of motion (9)). Using this result, one can easily show that $`(c\omega +d)=\pm n^{1/2}`$ (74) $`+c{\displaystyle \frac{\mathrm{sin}^2\theta }{N}}[ABC_{,t}+BCA_{,t}\mathrm{sin}^2\phi +ACB_{,t}\mathrm{cos}^2\phi ]`$ (75) $`c{\displaystyle \frac{ABC}{N}}\left({\displaystyle \frac{A_{,t}}{A}}+{\displaystyle \frac{B_{,t}}{B}}\right).`$ (76) With this and Eq. (20) for $`e^{2\gamma }`$ one easily evaluates the transformed field variables $`\{e^{2\gamma _g},\omega _g,\stackrel{~}{p}_g,\stackrel{~}{r}_g\}`$ using Eq. (58). The new spacetime metric thus takes the form $`ds_g^2`$ $`=`$ $`e^{2\gamma _g}\{\stackrel{~}{N}^2dt^2+\stackrel{~}{g}_{ab}dx^adx^b\}`$ (78) $`+e^{2\gamma _g}\{d\psi +n\mathrm{cos}\theta d\phi +\beta _{(g)a}dx^a\}^2`$ where however, $`\beta _{(g)a}`$ remains to be computed. As discussed in Ref. , $`\beta _{(g)a}`$ can be expanded (via the Hodge decomposition for a one-form on $`S^2`$) as $$\beta _{(g)a}=\left(\stackrel{~}{g}_{ac}\frac{^{cd}}{\stackrel{~}{\mu }_g}\right)\eta _{,d}+\delta _{,a}$$ (79) where $`\stackrel{~}{\mu }_g=\sqrt{{}_{}{}^{(2)}\stackrel{~}{g}}`$ and $`\eta `$ and $`\delta `$ are suitable functions defined on $`S^2`$. The equation for $`\beta _{(g)a}dx^a`$ is $$\stackrel{~}{r}_g=^{ab}\beta _{(g)a,b}+n\mathrm{sin}\theta $$ (80) which, upon substitution of the decomposition (79), becomes $$\stackrel{~}{r}_gn\mathrm{sin}\theta =(\stackrel{~}{\mu }_g\stackrel{~}{g}^{ab}\eta _{,a})_{,b}$$ (81) a Poisson equation for $`\eta `$ for which the necessary and sufficient integrability condition is ensured by Eq. (66). This uniquely determines $`\eta `$, at fixed $`t`$, up to an arbitrary additive constant and leaves $`\delta `$ arbitrary. The presence of $`\delta `$ reflects the freedom to make an arbitrary coordinate transformation of the form $`\psi \psi +\delta `$ without affecting the $`U(1)`$-form of the spacetime metric. The time development of $`\beta _{(g)a}`$ can now be obtained by integrating the (zero shift) evolution equation $`\beta _{(g)a,t}`$ $`=`$ $`\left({\displaystyle \frac{\stackrel{~}{N}}{\stackrel{~}{\mu }_g}}\right)e^{4\gamma _g}\stackrel{~}{g}_{ab}e_{(g)}^b`$ (82) $`=`$ $`\left({\displaystyle \frac{\stackrel{~}{N}}{\stackrel{~}{\mu }_g}}\right)e^{4\gamma _g}\stackrel{~}{g}_{ab}^{bc}\omega _{g,c}`$ (83) with $`\gamma _g,\omega _g`$ determined as above. Equations (80) and (82) are consistent with each other by virtue of the Hamilton equations satisfied by $`\stackrel{~}{r}_g,\omega _g`$. Note that whereas we have used the actual metric $`\stackrel{~}{g}_{ab}`$ in defining a Hodge decomposition of $`\beta _{(g)a}`$, any smooth metric on $`S^2`$ could have been used instead. Furthermore one could have used Eq. (81) to determine $`\eta `$ at an arbitrary time and then adjusted the time dependence of $`\delta `$ to impose the zero shift condition which is implicit in Eq. (82). In either case the new metric (78) will satisfy the vacuum field equations on the chosen $`S^1`$ bundle over $`S^2\times 𝐑`$. One might still wonder how we know that the transformed solutions are genuinely inhomogeneous. Could they not be merely homogeneous solutions disguised through the choice of a time slicing that is not adapted to the (hypothetical) homogeneity? To show that this is not the case, in general,we shall examine a special case for which the transformed solution has a hypersurface of time symmetry at $`t=t_0`$, i.e., has $`K_{ij}^{(g)}_{t=t_0}=0`$. To arrange this, we choose the seed solution to have this property and make a careful choice of transformation parameters so that the desired feature is not destroyed by the $`SL(2,𝐑)`$ transformation. We then show that the transformed spatial metric $`g_{ij}^{(g)}_{t=t_0}`$ is not homogeneous as it would have to be for the resulting spacetime to have this property. The key point here is the fact that on any compact slice, having constant mean curvature the first and second fundamental forms $`\{g_{ij}^{(g)},K_{ij}^{(g)}\}_{t=t_0}`$ would both have to be homogeneous in order that the spacetime have this property. Consider a Mixmaster solution for which $`\dot{A}(t_0)=\dot{B}(t_0)=\dot{C}(t_0)=0`$. This spacetime has the $`t=t_0`$ slice as a surface of time symmetry and, because of the Hamiltonian constraint, must satisfy $`C(t_0)=\pm (A(t_0)\pm B(t_0)),`$ (84) $`A(t_0),B(t_0),C(t_0)>0.`$ (85) To maintain this property we choose the trivial target bundle $`S^2\times S^1\times 𝐑`$ by taking $`n=0`$. We further simplify the computations by choosing $`A(t_0)=B(t_0)>0`$ and $`C(t_0)=2A(t_0)`$ and find that the transformed metric at $`t=t_0`$ satisfies $`\stackrel{~}{g}_{\theta \theta }_{t=t_0}`$ $`=`$ $`A^4(t_0)(\mathrm{sin}^2\theta +4\mathrm{cos}^2\theta ),`$ (86) $`\stackrel{~}{g}_{\theta \phi }_{t=t_0}`$ $`=`$ $`0,`$ (87) $`\stackrel{~}{g}_{\phi \phi }_{t=t_0}`$ $`=`$ $`4A^4(t_0)\mathrm{sin}^2\theta ,`$ (88) $`\underset{}{𝛽}{}_{(g)}{}^{}_{t=t_0}^{}`$ $`=`$ $`\beta _{(g)a}dx^a_{t=t_0}=c^24A^4(t_0)\mathrm{cos}\theta \mathrm{sin}^2\theta d\phi ,`$ (89) $`e^{2\gamma _g}_{t=t_0}`$ $`=`$ $`{\displaystyle \frac{1}{c^2A^2(t_0)(\mathrm{sin}^2\theta +4\mathrm{cos}^2\theta )}}.`$ (90) Thus the new spatial metric induced at $`t=t_0`$ on $`S^2\times S^1`$ is $`d\mathrm{}_{(g)}^2_{t=t_0}`$ $`=`$ $`c^2A^2(t_0)(\mathrm{sin}^2\theta +4\mathrm{cos}^2\theta )\{4A^4(t_0)\mathrm{sin}^2\theta d\phi ^2`$ (93) $`+A^4(t_0)(\mathrm{sin}^2\theta +4\mathrm{cos}^2\theta )d\theta ^2\}`$ $`+{\displaystyle \frac{1}{c^2A^2(t_0)(\mathrm{sin}^2\theta +4\mathrm{cos}^2\theta )}}\{d\psi 4c^2A^4(t_0)\mathrm{cos}\theta \mathrm{sin}^2\theta d\phi \}^2.`$ A straightforward computation of $`R_{ij}R^{ij}`$ (the square of the Ricci tensor of this metric) proves that the resulting spacetime is not homogeneous. Indeed, the only vacuum homogeneous solution on $`S^2\times S^1\times 𝐑`$ is known to be the Kantowski-Sachs universe which does not have a hypersurface of time symmetry. It is possible to get the Kantowski-Sachs metric upon transformation of a Taub metric (i.e., a Mixmaster solution having $`A(t)=B(t)`$) but to do so one must reduce with respect to the “extra” Killing field the Taub metric possesses, $`\frac{}{\phi }`$, rather than with respect to the common Killing field $`\frac{}{\psi }`$ of the Mixmaster family as we have done. The extra Killing field $`\frac{}{\phi }`$ commutes with all the Killing symmetries of the Taub solution and allows all of these symmetries to be preserved upon reduction . The Taub metric used in the example above, is known explicitly but exhibits no $`BKL`$ type oscillations. To see such oscillations in an inhomogeneous setting, we combine a previously developed code for solving the Mixmaster equations of motion with the transformations discussed above. Our results are discussed in the following section and compared with results derived from a general $`U(1)`$-syummetric, vacuum Einstein code. ## IV Numerical results Elsewhere we have shown that even the homogeneous Mixmaster model reproduces the local behavior seen in generic $`U(1)`$-symmetric cosmologies . From Eq. (20), it is clear that $`\gamma `$ is dominated by the largest of the Mixmaster scale factors $`A`$, $`B`$, or $`C`$. The local oscillations seen in $`\gamma `$ in the $`U(1)`$-symmetric models are interpreted as follows: Assume that the BKL approximate description of a homogeneous Mixmaster model as a sequence of Kasner epochs is valid. In a given approximate Kasner epoch assume $`A>B>C`$ and that $`A`$ is increasing. Then $`\gamma ,_t>0`$ for $`A,_t>0`$ while $`B,_t`$ and $`C,_t`$ are less than zero. The usual Mixmaster bounce changes the sign of $`A,_t`$ and thus of $`\gamma ,_t`$. However, after the bounce, either $`B,_t`$ (within an era) or $`C,_t`$ (at the end of an era) becomes positive. When the growing scale factor surpasses the decreasing $`A`$, $`\gamma ,_t`$ will start to grow again since it will now track the new dominant scale factor. A similar analysis indicates that the remaining “dynamical” variables, $`\omega `$, $`\stackrel{~}{p}`$, and $`\stackrel{~}{r}`$, depend on an order unity ratio of scale factors and thus do not oscillate, as $`\gamma `$ does, between order unity and exponentially small values. (Here we shall use “order unity” to mean some finite value which is not exponentially small.) In our previous numerical simulations of generic $`U(1)`$-symmetric cosmologies on $`T^3\times 𝐑`$, we noted that the oscillations in $`\gamma `$ could be interpreted as bounces off the potentials $`V_1=\frac{1}{2}\stackrel{~}{r}^2e^{4\gamma }`$ and $`V_2=\frac{1}{2}\stackrel{~}{g}\stackrel{~}{g}^{ab}e^{4\gamma }\omega ,_a\omega ,_b`$. For a Mixmaster solution, $`V_1`$ is exponentially small unless $`e^{2\gamma }`$ is of order unity while $`V_2`$ is exponentially small unless the two largest scale factors are approximately equal to each other . This is clearly consistent with a presumption that the generic models exhibit local Mixmaster dynamics. To explore the nature of the new inhomogeneous $`U(1)`$-symmetric models, we note that the transformed variables $`\omega _g`$, $`\stackrel{~}{p}_g`$, and $`\stackrel{~}{r}_g`$ will remain of order unity (i.e. they will not oscillate between exponentially small and order unity values) because the right hand sides of Eqs. (58b)-(58d) are always of order unity. On the other hand, $`\gamma _g`$ is dominated by the behavior of the oscillatory $`\gamma `$ since the denominator on the right hand side of Eq. (58a) is always order unity while the numerator oscillates. To explore the differences between our new solution and the Mixmaster seed solution, we construct the new solutions as follows: First use the algorithm of Berger et al to obtain a numericallly generated Mixmaster model. This code is known to solve the Mixmaster ODE’s with machine-level precision and can follow hundreds of bounces. The presumed stochastic properties of such a model imply that almost any Mixmaster initial conditions will yield generic Mixmaster behavior. Thus, we need only consider a single Mixmaster trajectory. Next, Eqs. (20), (36), (43), (48), (49), (67), and (73) are used to numerically evaluate $`\gamma `$, $`\omega `$, $`\stackrel{~}{p}`$, and $`\stackrel{~}{r}`$ from the numerically generated sequence of values of the BKL scale factors and their time derivatives. Finally, for a representative choice of the $`SL(2,𝐑)`$ parameters and, e.g., $`n=1`$, the transformed variables $`\gamma _g`$, $`\omega _g`$, $`\stackrel{~}{p}_g`$, and $`\stackrel{~}{r}_g`$ are computed using Eqs. (58). In Figures 1–3, we compare the Mixmaster and transformed $`\gamma `$ and $`\omega `$ at a representative spatial point for typical Mixmaster seed and set of $`SL(2,𝐑)`$ parameters. Note that, in Fig. 1, the original and transformed $`\gamma `$’s become indistinguishable after only a small number of Mixmaster epochs. It is clear that this will be so from Eq. (20) for $`\gamma `$ and Eq. (58a) for the transformation. Since $`\gamma `$ and $`\gamma _g`$ are found from the logarithm of Eqs. (20) and (58a), both $`\gamma `$ and $`\gamma _g`$ will be approximately equal to the logarithm of the largest scale factor and depend only logarithmically on the spatially dependent function associated with it. On a finer scale, in Fig. 2, the difference between the solutions (especially near the “bounce” where $`\gamma \gamma _g0`$) may be seen. On the other hand, as is seen in Fig. 3, $`\omega `$ and $`\omega _g`$ are always of order unity and may easily be distinguished. Figure 4 demonstrates the close link between Mixmaster dynamics and the oscillatory behavior observed in our studies of generic $`U(1)`$-symmetric models and should be compared to Figs. 2–6 in . It shows the oscillations of $`\gamma _g`$ (or essentially equivalently $`\gamma `$), $`V_1`$, and $`V_2`$ at a representative spatial point—reproducing the behavior seen in our simulations of generic $`U(1)`$-symmetric cosmologies. Since we know that these oscillations indicate local Mixmaster behavior in the new solutions, we can infer that the observed oscillations in the generic models also indicate local Mixmaster dynamics. Since $`\gamma `$ is the key variable in the $`U(1)`$-symmetric models and $`\gamma \gamma _g`$, one may then ask where these new $`U(1)`$-symmetric models differ from both Mixmaster and generic $`U(1)`$-symmetric models. First, we emphasize that, except at special values of the spatial coordinate angles, there are no qualitative differences attributable to spatial topology. The Mixmaster spatial dependence of course represents a realization of the Bianchi type IX symmetry. From Eq. (20), it is clear that three distinct spatial patterns will appear in $`\gamma `$ (in the logarithm) depending on which scale factor dominates. In Fig. 5, we compare the spatial dependence of $`\gamma `$ and $`\gamma _g`$ for 12 epochs of the seed Mixmaster solution. The epochs are arranged according to the dominant scale factor. The numerical scale in each frame is chosen so that the average value of $`\gamma `$ or $`\gamma _g`$ is the centroid. (If this were not done, no spatial dependence would be visible.) From Eqs. (20) and (48), $`\gamma `$ and $`\omega `$ have three possible spatial dependences. The $`SL(2,𝐑)`$ transformation of Eqs. (58) clearly mixes the spatial dependence of $`\gamma `$ and $`\omega `$ to form $`\gamma _g`$ and $`\omega _g`$. In Fig. 5, we see the evolving spatial dependence of $`\gamma _g`$. This is additional evidence that the new solutions are spatially inhomogeneous. In generic $`U(1)`$-symmetric models, one could qualitatively interpret the asymptotic approach to the singularity as the evolution of a different Mixmaster model at every spatial point. In particular, the Mixmaster epochs have spatially dependent durations—bounces at different spatial points occur at different times. In contrast, our new solution is characterized as is Mixmaster itself by spatially independent epoch durations since the new solution bounces only when the seed solution does so. While one could modify the spacetime slicing to yield spatially dependent epoch durations, one would expect to be able to detect the difference between between a single underlying Mixmaster seed in the new solutions and a continuum of approximate Mixmaster solutions in the generic case. ## Acknowledgments We would like to thank the Institute for Theoretical Physics at the University of California / Santa Barbara for hospitality. BKB would like to thank the Institute for Geophysics and Planetary Physics of Lawrence Livermore National Laboratory for hospitality. This work was supported in part by National Science Foundation Grants PHY9732629, PHY9800103, PHY9973666, and PHY9407194. ## Figure Captions Figure 1. Comparison of $`\gamma `$ and $`\gamma _g`$ at a typical spatial point. The Mixmaster seed solution has initial values $`\beta _+=0.9847899998176387`$, $`\beta _{}=0.09987655443789`$, $`\mathrm{\Omega }=8.00000000000000`$, $`\dot{\beta }_+=3.632980009876544`$, $`\dot{\beta }_{}=4.58987654433567878`$, and the Hamiltonian constraint (12) solved for $`\dot{\mathrm{\Omega }}`$. The $`SL(2,𝐑)`$ parameters are $`a=1`$, $`b=1`$, $`c=10000`$, and $`d=10001`$. Figure 2. Detail of the comparison of $`\gamma `$ and $`\gamma _g`$. To emphasize the approach of $`\gamma _g`$ to $`\gamma `$, data from later in the simulation of Fig. 1 are shown. The actual, saved data values are indicated by the $`\times `$ and $`+`$ symbols. Figure 3. Comparison of $`\omega `$ and $`\omega _g`$ for the same models as in Fig. 1. Note that $`\omega _g`$ appears to decrease to zero. This is due to the fact that choice of $`SL(2,𝐑)`$ parameters causes $`|\omega _g|10^4`$ if $`|\omega |1`$. Figure 4. New solution as an inhomogeneous $`U(1)`$-symmetric cosmology. As in previous studies of generic $`U(1)`$-symmetric cosmologies, $`\gamma _g`$, $`V_1`$, and $`V_2`$ are shown at a typical spatial point. Figure 5. Evidence for the spatial inhomogeneity of the new solutions. The spatial dependence of $`\gamma `$ and $`\gamma _g`$ is shown for the $`(\mathrm{cos}\theta ,\phi )`$ plane in a series of side-by-side frames arranged in three separate panels. Each pair of frames shows the spatial dependence of $`\gamma `$ and $`\gamma _g`$ respectively during an approximate Kasner epoch of the seed Mixmaster solution. The panels are grouped according to the identity of the dominant scale factor in the spatially homogeneous solution rather than sequentially. According to Eq. (20), $`\gamma `$ will have the spatial dependence $`\mathrm{ln}(\mathrm{sin}\theta \mathrm{sin}\phi )`$, $`\mathrm{ln}(\mathrm{sin}\theta \mathrm{cos}\phi )`$, or $`\mathrm{ln}(\mathrm{cos}\theta )`$ depending on whether $`A`$, $`B`$, or $`C`$ respectively is dominant. In each of the three panels, the 4 left-hand frames reproduce one of these three spatial dependences with, reading from left to right, $`B`$, $`C`$, or $`A`$ dominant. In each case, the accompanying right-hand frame represents the spatial dependence of the corresponding $`\gamma _g`$ for that epoch. In every case, the numerical scales for $`\gamma `$ and $`\gamma _g`$ have been centered on their average values to enhance the visibility of the spatial dependence.
warning/0001/gr-qc0001103.html
ar5iv
text
# Boson stars in massive dilatonic gravity ## I Introduction Boson stars are gravitationally bound macroscopic quantum states made up of scalar bosons. They differ from the fermionic stars in that they are only prevented from collapsing gravitationally by the Heisenberg uncertainty principle. Boson stars were first considered by Kaup and then by Ruffini and Bonazola . They found that the boson stars as described by non-interacting massive, complex scalar field had masses of the order of $`\frac{_{Pl}^2}{m_B}`$, where $`_{Pl}`$ is the Planck mass and $`m_B`$ is the boson mass. In a latter work Colpi, Shapiro and Wasserman analyzed the consequence of switching on a quartic self-interaction for the boson field. This results in a drastic increase of the mass of the boson star, which even for small values of the coupling constant turns out to be of the order of the Chandrasekhar’s mass when the boson mass is similar to a proton mass. Thus, the boson stars arise as possible candidates for non-baryonic dark matter in the universe. Although we still have no astrophysical evidences for their existence, the boson stars are good model to learn more about the nature of strong gravitational fields not only in general relativity but also in the other theories of gravity. The most natural and promising generalizations of Einstein’s general relativity are scalar-tensor theories of gravity. The most studied class of such theories are scalar-tensor theories with massless gravitational scalar field including the Brans-Dicke theory as a special case. It should be noted that an interesting specific class of scalar-tensor theories with dilaton field arise from the low energy limit of string theory, which confirms their importance for current physics. Boson stars in the framework of scalar-tensor theories with massless gravitational scalar have been studied widely, too. The first scalar-tensor model of a boson star was studied by Gundersen and Jensen , who considered Brans-Dicke theory with $`\omega _{BD}=6.`$ Their work was generalized by Torres , who studied boson stars in scalar-tensor theories with nonconstant $`\omega _{BD}(\mathrm{\Phi }).`$ The conclusion is that boson stars can exist in any scalar-tensor theory of gravity with massless gravitational scalar, as the masses of the boson stars are always smaller than in the general relativistic case irrespective of the coupling functions $`\omega _{BD}(\mathrm{\Phi })`$. We should mention also the interesting paper by Z. Tao and X. Xue where boson stars are studied in the framework of a scalar-tensor theory with a coupling between the dilaton and the mass term of the boson field. More recently boson stars in scalar-tensor theories have been studied in the papers by Torres et al , and in the paper by Comer and Shinkai . In boson stars have been studied in connection with so-called gravitational memory , while their stability through cosmic history has examined in and . The dynamical evolution of boson stars has been investigated in the paper by Balakrishna and Shinkai . Charged boson stars in scalar-tensor gravity with massless gravitational scalar have been studied in . As we already mentioned boson stars in scalar-tensor theories have been studied only in the case of scalar-tensor theories of gravity with massless gravitational scalar. From a field-theoretical point of view it is to be expected that the scalar part of the gravitational field is massive and leads to a force with a finite range. It also should be noted that there is no symmetry which forbids a mass term, or a more general potential term for the gravitational scalar. Moreover, if string theory and it’s low energy limit are relevant to the real world then the dilaton must be massive. Unfortunately, our current understanding of how the dilaton acquires mass is rather primitive and it’s tied to our lack of understanding of supersymmetry breaking. Since we don’t have a definite model for dilaton mass generation the mass of the dilaton and the form of its potential are completely unknown from theoretical point of view. The purpose of the present paper is to study boson stars in the framework of some class of scalar-tensor theories of gravity with massive gravitational scalar. In particular we study the possible influence of the mass and more generally of the potential of gravitational scalar on the equilibrium configurations and stability of the boson stars. The obtained results, however, are typical for wider classes of scalar-tensor theories with massive dilaton not only for the class we consider. ## II A scalar-tensor theory of gravity with massive gravitational scalar The most general scalar-tensor theory of gravity including such a term is described by the following action $`\stackrel{~}{𝒜}={\displaystyle \frac{1}{16\pi G_{}}}{\displaystyle }d^4x\sqrt{\stackrel{~}{g}}(\mathrm{\Phi }\stackrel{~}{R}h(\mathrm{\Phi })\stackrel{~}{g}^{\mu \nu }_\mu \mathrm{\Phi }_\nu \mathrm{\Phi }`$ (1) $`+\stackrel{~}{U}(\mathrm{\Phi }))+𝒜_{matter}(\mathrm{\Psi }_{matter},\stackrel{~}{g}_{\mu \nu })`$ (2) where $`\stackrel{~}{g}_{\mu \nu }`$ is the space-time metric, $`\stackrel{~}{R}`$ is the Ricci scalar curvature with respect to the space-time metric. The functions $`h(\mathrm{\Phi })`$ and $`\stackrel{~}{U}(\mathrm{\Phi })`$ are, in general, arbitrary differentiable functions of the gravitational scalar $`\mathrm{\Phi }`$. A new method for determination of these functions using astrophysical observations is developed in the recent articles , but at present we have no good enough data to make use of this method. The last term in (1) denotes the action of the matter, which is a functional of the matter variables, collectively denoted by $`\mathrm{\Psi }_{matter}`$, and of the space-time metric $`\stackrel{~}{g}_{\mu \nu }`$. For our purpose in this article the action (1) is too general. That’s why we restrict ourselves to a subclass of scalar-tensor theories which don’t include explicitly a kinetic term for the gravitational scalar in the action (i.e. with function $`h(\mathrm{\Phi })0`$). In particular we consider scalar-tensor theories described by the action $`\stackrel{~}{A}={\displaystyle \frac{1}{16\pi G_{}}}{\displaystyle d^4x\sqrt{\stackrel{~}{g}}\left(\mathrm{\Phi }\stackrel{~}{R}+\stackrel{~}{U}(\mathrm{\Phi })\right)}`$ (3) $`+𝒜_{matter}(\mathrm{\Psi }_{matter},\stackrel{~}{g}_{\mu \nu }).`$ (4) Such scalar-tensor theory with only one unknown function $`\stackrel{~}{U}(\mathrm{\Phi })`$ is much more definite. It has been considered at first by O’Hanlon in connection with Fujii’s theory of massive dilaton . It may be considered also as a part of more general theories of gravity like (4+1)-dimensional Kaluza-Klein model described by Fujii in , or model of gravity with violated local conformal symmetry in affine-connected spaces which probably may be related with string theory . A recent development of model of dilatonic gravity with action (3) related with newest astrophysical data on cosmological constant problem may be found in . In addition there it was shown that this model is consistent with extremely high precision with all known data in solar system gravitational experiments if one chooses properly the potential $`\stackrel{~}{U}(\mathrm{\Phi })`$. Under proper normalization of this potential solar system end Earth surface experiments actually give restrictions only on the mass term in the potential $`\stackrel{~}{U}(\mathrm{\Phi })`$. The corresponding field equations are $`\stackrel{~}{G}_{\mu \nu }={\displaystyle \frac{\kappa _{}}{\mathrm{\Phi }}}\stackrel{~}{T}_{\mu \nu }+{\displaystyle \frac{1}{\mathrm{\Phi }}}\left(\stackrel{~}{}_\mu \stackrel{~}{}_\nu \mathrm{\Phi }\stackrel{~}{g}_{\mu \nu }\stackrel{~}{}_\rho \stackrel{~}{}^\rho \mathrm{\Phi }\right)+{\displaystyle \frac{1}{2}}`$ $`{\displaystyle \frac{\stackrel{~}{U}(\mathrm{\Phi })}{\mathrm{\Phi }}}\stackrel{~}{g}_{\mu \nu },`$ (5) $`\stackrel{~}{}_\rho \stackrel{~}{}^\rho \mathrm{\Phi }+{\displaystyle \frac{1}{3}}\left(\mathrm{\Phi }{\displaystyle \frac{d\stackrel{~}{U}(\mathrm{\Phi })}{d\mathrm{\Phi }}}2\stackrel{~}{U}(\mathrm{\Phi })\right)`$ $`=`$ $`{\displaystyle \frac{\kappa _{}}{3}}\stackrel{~}{T}.`$ (7) For the treatment of the local Earth surface, astrophysical and astronomical problems in star systems (but not for cosmological problems in the scales of universe) we demand that a weak-field approximation should be possible assuming in addition that the space-time is asymptotically flat. Then in weak-field approximation we may write $`\stackrel{~}{g}_{\mu \nu }=\eta _{\mu \nu }+\delta \stackrel{~}{h}_{\mu \nu },\mathrm{\Phi }=\mathrm{\Phi }_{\mathrm{}}+\delta \mathrm{\Phi }`$ (8) where $`\eta _{\mu \nu }`$ is the flat space-time metric, $`\mathrm{\Phi }_{\mathrm{}}`$ is the background value of the $`\mathrm{\Phi }`$, and $`\delta h_{\mu \nu }`$ and $`\delta \mathrm{\Phi }`$ are small perturbations due to the local masses. Consistency of the field equations then requires that $`\stackrel{~}{U}(\mathrm{\Phi }_{\mathrm{}})={\displaystyle \frac{d\stackrel{~}{U}}{d\mathrm{\Phi }}}(\mathrm{\Phi }_{\mathrm{}})=0.`$ (9) Using now the consistency conditions (9) we obtain that the mass of the gravitational scalar $`\mathrm{\Phi }`$ is $`m^2={\displaystyle \frac{\mathrm{\Phi }_{\mathrm{}}}{3}}{\displaystyle \frac{d^2\stackrel{~}{U}}{d\mathrm{\Phi }^2}}(\mathrm{\Phi }_{\mathrm{}}).`$ (10) Further on, without loss of generality we take $`\mathrm{\Phi }_{\mathrm{}}=1.`$ In this case, taking into account that the range of the gravitational scalar is finite we obtain that the bare gravitational constant $`G_{}`$ coincides with the background physical gravitational constant. The PPN formalism for scalar-tensor theories of gravity with a massive gravitational scalar has been developed by Helbig , who has also calculated the basic gravitational effects as perihelion shift and so on. Using Helbig’s results and the experimental data published in and we obtain that the scalar-tensor theory we consider here is consistent with the experiments when the Compton length of the gravitational scalar $`\mathrm{\Phi }`$ is less than $`2`$ mm, or equivalently when the mass of the gravitational scalar satisfies the constraint $`m10^4eV`$ . So far our considerations have been made in the Jordan frame. For some purposes, in particular for numerical calculations, the scalar-tensor theories are better formulated in the conformal Einstein frame. The metric in the Einstein frame is given by the conformal transformation $`g_{\mu \nu }=\mathrm{\Phi }\stackrel{~}{g}_{\mu \nu }.`$ (11) It should be stressed that it’s the Jordan frame which is the physical frame while the Einstein frame is a convenient mathematical tool. Including also the dilaton field $`\phi `$ via the formula $`\mathrm{\Phi }=\mathrm{exp}(2\alpha \phi )`$, where $`\alpha =\frac{1}{\sqrt{3}}`$, the action (3) written in terms of $`g_{\mu \nu }`$ and $`\phi `$ reads $`𝒜={\displaystyle \frac{1}{16\pi G_{}}}{\displaystyle d^4x\sqrt{g}\left(R2g^{\mu \nu }_\mu \phi _\nu \phi +U(\phi )\right)}`$ (12) $`+𝒜_{matter}(\mathrm{\Psi }_{matter},A^2(\phi )g_{\mu \nu }).`$ (13) Here $`R`$ is the Ricci scalar curvature with respect to the metric $`g_{\mu \nu }`$, $`U(\phi )=A^4(\phi )\stackrel{~}{U}(\mathrm{\Phi }(\phi ))`$ is the potential of the dilaton field $`\phi `$ and $`A^2(\phi )=\mathrm{\Phi }^1(\phi )`$. It should be noted that the convention $`\mathrm{\Phi }_{\mathrm{}}=1`$ in terms of the dilaton field writes $`\phi _{\mathrm{}}=0`$. As must be expected the dilaton potential $`U(\phi )`$ satisfies the conditions $`U(\phi _{\mathrm{}})=\frac{dU(\phi )}{d\phi }(\phi _{\mathrm{}})=0`$ and the mass of the dilaton coincides with the mass of the gravitational scalar $`\mathrm{\Phi }`$ which gives another equivalent representation of the same physical object. ## III Boson star model We take the matter action to be the action of complex, massive and self-interacting scalar field $`\mathrm{\Psi }`$ which in the physical Jordan frame has the form $`𝒜_{matter}={\displaystyle d^4x\sqrt{\stackrel{~}{g}}\left(\frac{1}{2}\stackrel{~}{g}^{\mu \nu }_\mu \mathrm{\Psi }^+_\nu \mathrm{\Psi }W(\mathrm{\Psi }^+\mathrm{\Psi })\right)}`$ (14) where $`W(\mathrm{\Psi }^+\mathrm{\Psi })=\frac{1}{2}m_B^2\mathrm{\Psi }^+\mathrm{\Psi }+\frac{1}{4}\widehat{\mathrm{\Lambda }}(\mathrm{\Psi }^+\mathrm{\Psi })^2`$ is the potential of the boson field. Then the action of the gravity and matter in Einstein frame is $`𝒜={\displaystyle \frac{1}{16\pi G_{}}}{\displaystyle }`$ $`d^4x\sqrt{g}\left(R2g^{\mu \nu }_\mu \phi _\nu \phi +U(\phi )\right)`$ (15) $`+{\displaystyle d^4x\sqrt{g}}`$ $`({\displaystyle \frac{1}{2}}A^2(\phi )g^{\mu \nu }_\mu \mathrm{\Psi }^+_\nu \mathrm{\Psi }`$ (16) . $`A^4(\phi )W(\mathrm{\Psi }^+\mathrm{\Psi })).`$ (17) Varying the action (17) with respect to $`g_{\mu \nu }`$, $`\phi `$, $`\mathrm{\Psi }`$ and $`\mathrm{\Psi }^+`$ we obtain the field equations $`G_{\mu \nu }=\kappa _{}T_{\mu \nu }+2_\mu \phi _\nu \phi g_{\mu \nu }_\rho \phi ^\rho \phi +{\displaystyle \frac{1}{2}}U(\phi )g_{\mu \nu },`$ (18) $`_\rho ^\rho \phi +{\displaystyle \frac{1}{4}}U^{}(\phi )={\displaystyle \frac{\alpha }{2}}\kappa _{}T,`$ (19) $`_\rho ^\rho \mathrm{\Psi }+2\alpha _\rho \phi ^\rho \mathrm{\Psi }=2A^2(\phi ){\displaystyle \frac{dW(\mathrm{\Psi }^+\mathrm{\Psi })}{d\mathrm{\Psi }^+}},`$ (20) $`_\rho ^\rho \mathrm{\Psi }^++2\alpha _\rho \phi ^\rho \mathrm{\Psi }^+=2A^2(\phi ){\displaystyle \frac{dW(\mathrm{\Psi }^+\mathrm{\Psi })}{d\mathrm{\Psi }}}`$ (21) where $`_\rho `$ is the Levi-Civita connection with respect to the metric $`g_{\mu \nu }`$, $`T_{\mu \nu }`$ is the Einstein frame energy-momentum tensor of the boson filed $`\mathrm{\Psi }`$ and $`T`$ is its trace. The Einstein frame energy-momentum tensor $`T_{\mu \nu }`$ is given by $`T_{\mu \nu }={\displaystyle \frac{1}{2}}A^2(\phi )\left(_\mu \mathrm{\Psi }^+_\nu \mathrm{\Psi }+_\mu \mathrm{\Psi }_\nu \mathrm{\Psi }^+\right)`$ (22) $`{\displaystyle \frac{1}{2}}A^2(\phi )\left(_\rho \mathrm{\Psi }^+^\rho \mathrm{\Psi }2A^2(\phi )W(\mathrm{\Psi }^+\mathrm{\Psi })\right)g_{\mu \nu }.`$ (23) We consider a static and spherically symmetric boson star. Then the metric $`g_{\mu \nu }`$ can be written in the standard form $`ds^2=e^{\nu (\rho )}dt^2e^{\lambda (\rho )}d\rho ^2\rho ^2\left(d\theta ^2+\mathrm{sin}^2(\theta )d\varphi ^2\right).`$ (24) We also demand a spherically-symmetric form for the boson field and we adopt a form consistent with the static metric: $`\mathrm{\Psi }=\widehat{\sigma }(\rho )e^{i\omega t}.`$ (25) Using the metric (24) and the equation defining the form of the boson field together with the Einstein frame energy-momentum tensor (22) in the field equations (18) we get the equations for the structure of the boson star. Before we explicitly write them we are going to introduce a dimensionless radial coordinate by $`r=m_B\rho .`$ (26) From now on, a prime will denote a differentiation with respect to the dimensionless coordinate $`r`$. We also define other dimensionless quantities by $`\mathrm{\Omega }={\displaystyle \frac{\omega }{m_B}},\sigma =\sqrt{\kappa _{}}\widehat{\sigma },\mathrm{\Lambda }={\displaystyle \frac{\widehat{\mathrm{\Lambda }}}{\kappa _{}m_B}},\gamma ={\displaystyle \frac{m}{m_B}}`$ (27) and dimensionless potential $`V(\phi )`$ given by $`U(\phi )=m^2V(\phi ).`$ (28) With all these definitions, the equations of the structure of the boson star in Einstein frame reduce to the following $`\lambda ^{}={\displaystyle \frac{1e^\lambda }{r}}+re^\lambda 𝒯_0^0+r\phi _{}^{}{}_{}{}^{2}+{\displaystyle \frac{1}{2}}\gamma ^2re^\lambda V(\phi ),`$ (29) $`\nu ^{}={\displaystyle \frac{e^\lambda 1}{r}}re^\lambda 𝒯_1^1+r\phi _{}^{}{}_{}{}^{2}{\displaystyle \frac{1}{2}}\gamma ^2re^\lambda V(\phi ),`$ (30) $`\nu ^{\prime \prime }={\displaystyle \frac{\nu _{}^{}{}_{}{}^{2}+\lambda ^{}\nu ^{}+\lambda ^{}\nu ^{}}{2}}2e^\lambda 𝒯_2^22\phi _{}^{}{}_{}{}^{2}\gamma ^2e^\lambda V(\phi ),`$ (31) $`\phi ^{\prime \prime }=\left({\displaystyle \frac{\nu ^{}\lambda ^{}}{2}}+{\displaystyle \frac{2}{r}}\right)\phi ^{}+{\displaystyle \frac{1}{4}}\gamma ^2e^\lambda {\displaystyle \frac{dV(\phi )}{d\phi }}+{\displaystyle \frac{\alpha }{2}}𝒯e^\lambda ,`$ (32) $`\sigma ^{\prime \prime }=\left({\displaystyle \frac{\nu ^{}\lambda ^{}}{2}}+{\displaystyle \frac{2}{r}}\right)\sigma ^{}\mathrm{\Omega }^2e^{\lambda \nu }\sigma 2\alpha \phi ^{}\sigma ^{}+`$ (33) $`2A^2(\phi )e^\lambda {\displaystyle \frac{dW(\sigma ^2)}{d\sigma ^2}}\sigma .`$ (34) Here $`𝒯_\nu ^\mu =\frac{\kappa _{}}{m_B^2}T_\nu ^\mu `$ is the dimensionless Einstein frame energy-momentum tensor and $`𝒯`$ is its trace. In explicit form we have $`𝒯_0^0={\displaystyle \frac{1}{2}}\mathrm{\Omega }^2A^2(\phi )e^\nu \sigma ^2+{\displaystyle \frac{1}{2}}A^2(\phi )e^\lambda \sigma _{}^{}{}_{}{}^{2}+A^4(\phi )W(\sigma ^2),`$ (35) $`𝒯_1^1={\displaystyle \frac{1}{2}}\mathrm{\Omega }^2A^2(\phi )e^\nu \sigma ^2{\displaystyle \frac{1}{2}}A^2(\phi )e^\lambda \sigma _{}^{}{}_{}{}^{2}+A^4(\phi )W(\sigma ^2),`$ (36) $`𝒯_2^2={\displaystyle \frac{1}{2}}\mathrm{\Omega }^2A^2(\phi )e^\nu \sigma ^2+{\displaystyle \frac{1}{2}}A^2(\phi )e^\lambda \sigma _{}^{}{}_{}{}^{2}+A^4(\phi )W(\sigma ^2),`$ (37) $`𝒯=\mathrm{\Omega }^2A^2(\phi )e^\nu \sigma ^2+A^2(\phi )e^\lambda \sigma _{}^{}{}_{}{}^{2}+4A^4(\phi )W(\sigma ^2).`$ (38) By reasons imposed by the numerical method used in the present paper, instead to investigate the system (29) we solve numerically an equivalent system obtained as follows. First making use of the second equation of (29) we express $`e^\lambda `$ as a function of $`\nu ,\nu ^{},\sigma ,\sigma ^{},\phi ,\phi ^{}`$ and then substitute in the other equations. In this way we obtain the system: $`\nu ^{\prime \prime }={\displaystyle \frac{\nu ^{}}{r}}+({\displaystyle \frac{\nu ^{}}{r}}+𝒯_0^0𝒯_1^1+2𝒯_2^2\gamma ^2V(\phi )+`$ (39) $`{\displaystyle \frac{\nu ^{}r}{2}}(𝒯_0^0+𝒯_1^1+\gamma ^2V(\phi )))e^\lambda ,`$ (40) $`\phi ^{\prime \prime }={\displaystyle \frac{\phi ^{}}{r}}+({\displaystyle \frac{\phi ^{}}{r}}+{\displaystyle \frac{\alpha }{2}}𝒯+{\displaystyle \frac{1}{4}}\gamma ^2{\displaystyle \frac{dV(\phi )}{d\phi }}+`$ (41) $`{\displaystyle \frac{\phi ^{}r}{2}}(𝒯_0^0+𝒯_1^1+\gamma ^2V(\phi )))e^\lambda ,`$ (42) $`\sigma ^{\prime \prime }={\displaystyle \frac{\sigma ^{}}{r}}2\alpha \phi ^{}\sigma ^{}+({\displaystyle \frac{\sigma ^{}}{r}}+2A^2(\phi ){\displaystyle \frac{dW(\sigma ^2)}{d\sigma ^2}}\sigma `$ (43) $`\mathrm{\Omega }^2e^\nu \sigma +{\displaystyle \frac{\sigma ^{}r}{2}}(𝒯_0^0+𝒯_1^1+\gamma ^2V(\phi )))e^\lambda `$ (44) where $`e^\lambda `$ is given by $`e^\lambda ={\displaystyle \frac{1+r\nu ^{}r^2\phi _{}^{}{}_{}{}^{2}\frac{1}{2}A^2(\phi )r^2\sigma _{}^{}{}_{}{}^{2}}{1r^2\left(\frac{1}{2}\gamma ^2V(\phi )\frac{1}{2}\mathrm{\Omega }^2A^2(\phi )e^\nu \sigma ^2+A^4(\phi )W(\sigma ^2)\right)}}.`$ (45) We solve numerically the system (39). The boundary conditions for the system are the following. We demand asymptotic flatness which means that $`\nu (\mathrm{})=0`$. The non-singularity of the $`\nu `$ at the center of the star requires that $`\nu ^{}(0)=0`$. Concerning the dilaton $`\phi `$, the non-singularity at the center implies $`\phi ^{}(0)=0`$, while at infinity we must have $`\phi (\mathrm{})=0`$ as required by the asymptotic flatness. Non-singularity of the boson field at the center implies $`\sigma ^{}(0)=0`$ and the asymptotic flatness requires that $`\sigma (\mathrm{})=0`$. To complete the nonlinear eigenvalue problem for $`\mathrm{\Omega }`$ we also must give the central value of the boson field $`\sigma (0)=\sigma _c`$. Here we must make some comments. We have imposed the boundary conditions in Einstein frame. But what is important is the non-singularity and asymptotic flatness in the physical Jordan frame. However, it’s not difficult to see that all Einstein frame boundary conditions phrased in Jordan frame remain the same. Indeed, the only quantity which changes under transition to Jordan frame is $`\nu `$. The non-singularity at the center and the asymptotic flatness in Jordan frame require correspondingly that $`\stackrel{~}{\nu }^{}(0)=0`$ and $`\stackrel{~}{\nu }(\mathrm{})=0`$. Taking into account that $`\stackrel{~}{\nu }=\nu +2\alpha \phi `$ and Einstein frame boundary conditions for the dilaton field $`\phi `$ we obtain that Jordan frame boundary conditions for $`\stackrel{~}{\nu }`$ are satisfied if and only if $`\nu ^{}(0)=\nu (\mathrm{})=0`$. ## IV Conserved quantities The action of the boson field is $`U(1)`$-invariant under the global gauge transformations $`\mathrm{\Psi }e^{ia}\mathrm{\Psi }`$, where $`a`$ is a constant. This global $`U(1)`$-symmetry gives rise to the following conserved current in Jordan frame $`\stackrel{~}{J}^\mu ={\displaystyle \frac{i}{2}}\sqrt{\stackrel{~}{g}}\stackrel{~}{g}^{\mu \nu }\left(\mathrm{\Psi }_\nu \mathrm{\Psi }^+\mathrm{\Psi }^+_\nu \mathrm{\Psi }\right).`$ (46) The same current written in terms of the Einstein frame metric is $`J^\mu ={\displaystyle \frac{i}{2}}A^2(\phi )\sqrt{g}g^{\mu \nu }\left(\mathrm{\Psi }_\nu \mathrm{\Psi }^+\mathrm{\Psi }^+_\nu \mathrm{\Psi }\right).`$ (47) The conserved current leads to conserved charge - the total number of particles making up the star: $`N={\displaystyle \stackrel{~}{J}^0d^3x}.`$ (48) It should be stressed that $`N`$ is the total number of particles in the physical Jordan frame. Taking into account the explicit form of the metric and of the boson field we obtain $`N=\omega {\displaystyle d^3x\sqrt{\stackrel{~}{g}}\stackrel{~}{g}^{00}\widehat{\sigma }^2}`$ (49) In order to calculate numerically the total number of particles we must present the above integral in terms of Einstein frame metric. This can be done easily by presenting the Jordan frame metric as conformally transformed Einstein frame metric. This way we obtain $`_R=m_BN=\left({\displaystyle \frac{_{}^{2}{}_{Pl}{}^{}}{2m_B}}\right)\mathrm{\Omega }{\displaystyle _0^{\mathrm{}}}𝑑rr^2A^2(\phi )e^{\frac{\lambda \nu }{2}}\sigma ^2`$ (50) where $`_R`$ is the rest mass of the star in the Jordan frame. The binding energy in Jordan frame is then defined by $`_b=_R`$ (51) where $``$ is the total mass of the star. In contrast to the general relativity the definition of mass in scalar-tensor theories is subtle . There are three possible mass definitions in Jordan frame, notably the Schwarzschild mass $`_S`$ (i.e. ADM mass in Jordan frame), the Kepler mass $`_K`$ and the tensor mass $`_T`$ which is the ADM mass in Einstein frame . These mass definitions are equivalent in general relativity, but they may give different results in scalar-tensor theories of general type. It’s the tensor mass which has got physically acceptable properties: The tensor mass has the important property to peak, as a function of the central density at the same location where the particle number (respectively the rest mass) takes its maximum. This property has been proved in the most general scalar-tensor theory including a potential term for the gravitational scalar (dilaton), see for details . That the tensor mass and particle number peak at the same location results in a cusp in the bifurcation diagram $``$ (respectively $`_b`$) versus $`_R`$. All masses we have mentioned are asymptotically measured quantities. Taking into account that in the scalar-tensor theory we consider the gravitational scalar (dilaton) varies essentially in a finite space-time domain , it’s not difficult to see that all masses in our case are identical. That’s why all properties proven for the tensor mass are shared also by the others and the mass of the star may be expressed in three different ways as one wishes. We take the expression for the mass of the star in the form of the Schwarzschild mass in Einstein frame, namely $`=\left({\displaystyle \frac{_{}^{2}{}_{Pl}{}^{}}{2m_B}}\right){\displaystyle _0^{\mathrm{}}}𝑑rr^2\left(𝒯_0^0+e^\lambda \phi _{}^{}{}_{}{}^{2}+{\displaystyle \frac{1}{2}}U(\phi )\right).`$ (52) For numerical purposes we also introduce the following dimensionless masses $`M=/{\displaystyle \frac{_{Pl}^2}{2m_B}},M_R=_R/{\displaystyle \frac{_{Pl}^2}{2m_B}}.`$ (53) Therefore the dimensionless binding energy is $`E_b=MM_R.`$ (54) We also define an effective dimensionless radius of the boson star as: $`R={\displaystyle \frac{m_B}{N}}{\displaystyle rJ^0d^3x}.`$ (55) ## V Numerical results and discussions To solve the nonlinear ODE’s (39) with the corresponding boundary conditions and spectral parameter $`\mathrm{\Omega }`$ we apply the continuous analog of the Newton method , . In our case this method leads to a linear systems ODE’s with respect to the increments of unknown functions coupling with an algebraic equation for the increment of $`\mathrm{\Omega }`$. The appropriate linear boundary value problems we solve numerically using the spline collocation at the Gaussian points on the irregular mesh condensing to the centre $`r=0`$ of the star. For details see our work . As we have already mentioned at present we have only some weak experimental constrains on the mass of the dilaton and on the form of its potential. A definite theoretical suggestions for them do not exist. That’s why we consider first the simplest form for the potential of the gravitational scalar $`\stackrel{~}{U}(\mathrm{\Phi })={\displaystyle \frac{3}{2}}m^2\left(\mathrm{\Phi }1\right)^2.`$ (56) The corresponding dilaton potential is $`U(\phi )=m^2V(\phi )={\displaystyle \frac{3}{2}}m^2\left(1\mathrm{exp}({\displaystyle \frac{2\phi }{\sqrt{3}}})\right)^2.`$ (57) The numerical results for the chosen potential are presented in the figures. Fig.1 presents a configuration diagram which is typical for large variations of the parameters not only for those values presented on the figure. The dependence of the boson star mass on the central value $`\sigma _c`$ is presented in Fig.2 for three different values of the ratio $`\gamma =\frac{m}{m_B}`$ at fixed value of the parameter $`\lambda =10.`$ It’s seen that the curves $`M(\sigma _c)`$ differ slightly from each other. More detailed investigation for much more values of $`\gamma `$ shows that the curves tend to the same curve $`M(\sigma _c)`$ from the general relativistic case for large values of $`\gamma `$ ($`\gamma 10`$). This behaviour is confirmed also by Fig.3 where the dependence of the boson star mass on the parameter $`\gamma `$ is shown for $`\sigma _c=0.1`$ and $`\mathrm{\Lambda }=10`$. For small values of $`\gamma `$ the boson star mass increases when the parameter $`\gamma `$ increases. Then for larger values of $`\gamma `$ the mass $`M`$ tends to a fixed value which is just the general relativistic value. The cause for this behaviour is that the increasing of the parameter $`\gamma `$ is actually an increasing of the dilaton mass with respect to the boson mass which leads to smaller dilaton range and therefore to a suppressing of the dilaton field which means that the general relativity is recovered. The Fig. 3 also shows that the boson stars in the model under consideration always have masses smaller than those in general relativity. The deviations from general relativity are within a typical few percent. The influence of the dilaton mass, respectively $`\gamma `$, on the dependence of the boson field on the radial coordinate $`r`$ is very slightly as one may see from Fig. 4. The dependence $`\sigma (r)`$ is presented there for three different values of $`\gamma `$. The three curves are extremely close to each other. The dependence $`\stackrel{~}{\nu }(r)`$ is shown in Fig. 5 for three different values of $`\gamma `$. As it is seen the influence of $`\gamma `$ on $`\stackrel{~}{\nu }(r)`$ is slight. The most sensitive to the value of the parameter $`\gamma `$ is the dilaton field $`\phi `$. The dependence $`\phi (r)`$ is presented in Fig. 6 for three different values of $`\gamma `$. One sees that the increasing of $`\gamma `$ leads to suppressing the dilaton field as one must expect. The dependence $`\mathrm{\Phi }(r)`$ is shown in Fig. 7. At the center of the star one has $`\mathrm{\Phi }(0)>1`$ which means that the gravitational constant at the center satisfies $`G(0)<1`$. Therefore the boson stars are less gravitationally bounded than in general relativistic case. We have also examined the cases of different values of $`\mathrm{\Lambda }`$. The picture is qualitatively the same. It is interesting to know how the exact form of the dilaton potential $`V(\phi )`$ influences the boson star structure. We have examined several potentials $`V(\phi )`$ with the same dilaton mass (for example $`V(\phi )=2\phi ^2`$, $`V(\phi )=2\mathrm{sin}^2(\phi )`$, $`V(\phi )=4(1\mathrm{cos}(\phi ))`$, $`V(\phi )=2(1\mathrm{exp}(\phi ^2))`$). Our numerical results show that boson star structure in practice does not depend on the exact form of the dilaton potential in the class of potential we have examined. This gives strong evidences that the boson star structure is sensitive mainly to the second derivative of the dilaton potential $`\frac{d^2V(\phi )}{d\phi ^2}(0)`$ determining the dilaton mass rather to the exact form of the potential. Finally in Fig. 8 one can see that the usual dependence of the mass $`M(R)`$ of boson star on its effective radius $`R`$ take place in the scalar-tensor model with massive dilaton, too. As seen, there exist a domain of stability of the boson star, followed by domains of unstable states. ## VI Boson star stability Here we briefly discuss stability property of the boson stars we consider. Our analysis is based on catastrophe theory . The basis for our analysis is Fig. 9 where the binding energy $`E_b`$ is plotted against the rest mass $`M_R`$ (i.e. particle number). This figure, as our numerical results show, is typical for a large number of values of $`\gamma `$ and $`\mathrm{\Lambda }`$. Therefore the forthcoming conclusions concern the general case not only the special case presented by the figure. Fig. 9 is actually a bifurcation diagram. For small central densities the binding energy is negative which shows that the boson star is potentially stable and one assumes that the boson star is stable against small radial perturbations. As the central density is increased one meets a cusp. The second branch as a whole has higher mass and therefore it is unstable. At the cusp, the boson star stability changes - one radial perturbation mode develops instability. ## VII Conclusion In this paper we have analyzed static spherically symmetric boson stars in a scalar tensor theory of gravity with massive gravitational scalar (dilaton). We have studied their equilibrium properties for different values of the ratio $`\gamma =\frac{m}{m_B}`$. The conclusion is that stable boson stars may exist for any value of $`\gamma `$. When $`\gamma `$ increases the equilibrium configurations tends to those from general relativity and for sufficiently large $`\gamma `$ the general relativistic case is in practice recovered. The masses of the boson stars are turned out to be always smaller than those in general relativity. The typical deviations of the masses in our model of dilatonic gravity from those in general relativity is a few percent. The small deviations of bosonic star structure in the considered model of dilatonic gravity from general relativistic case are significantly greater then the corresponding deviations in solar system and Earth surface experiments studied in . It turns out also that the boson star structure is sensitive mainly to the mass term of the dilaton potential rather to the exact form of the potential. The boson stars in our case are less gravitationally bound as a consequence of the fact that the gravitational constant $`G`$ within the stars is smaller than one. For completeness we have examined also the case of scalar-tensor theories with a massive gravitational scalar with kinetic term $`\omega _{BD}\mathrm{\Phi }^2\stackrel{~}{g}^{\mu \nu }_\mu \mathrm{\Phi }_\nu \mathrm{\Phi }`$. In these cases the obtained results for different values of the parameter $`\omega _{BD}`$ are qualitatively the same as in the case $`\omega _{BD}0`$ and for large $`\omega _{BD}`$ the differences between bosonic stars in dilatonic gravity and in general relativity are even smaller. It is worth noting that, in the domain of stability of boson stars, our results are qualitatively close to the results obtained for boson stars in Brans-Dicke theory , , and for boson stars considered in a scalar-tensor theory with a kinetic term and simple coupling between the dilaton and the mass term of the boson filed in Jordan frame which in our notations reads $`m_B^2\mathrm{\Phi }\mathrm{\Psi }^+\mathrm{\Psi }`$. In the domain of instability of boson stars the results of the last reference differ essentially from ours. As in general relativity and Brans-Dicke theory beyond the point of stability for large central densities the boson star mass in our model of massive dilatonic gravity drops, oscillates a bit and approaches a constant value while in the model considered in it increases rapidly. Acknowledgments: We are grateful to the unknown referee for useful suggestions and for pointing out the references and . This work was supported by Bulgarian National Scientific Fund, Contr. NoNo F610/99, MM602/96 and by Sofia University Research Fund, Contr. No. 245/99.
warning/0001/math0001161.html
ar5iv
text
# 1 The irreducible orthogonal representations with free algebra of skew-invariants The exterior algebra and ‘Spin’ of an orthogonal $`g`$-module Dmitri I. Panyushev<sup>1</sup><sup>1</sup>1supported in part by R.F.F.I. Grant N0 98–01–00598 Introduction Let $`g`$ be a reductive algebraic Lie algebra over an algebraically closed field 𝕜 of characteristic zero and $`G`$ is the corresponding connected and simply connected group. The symmetric algebra of a (finite-dimensional) $`g`$-module $`V`$ is the algebra of polynomial functions on the dual space $`V^{}`$. Therefore one can study the algebra of symmetric invariants using geometry of $`G`$-orbits in $`V^{}`$. In case of the exterior algebra, $`^{}V`$, lack of such geometric picture results by now in absence of general structure theorems for the algebra of skew-invariants $`(^{}V)^g`$. One may find in the literature several interesting results related to skew-symmetric invariants. We only mention Kostant’s computation for cohomology of the nilradical of a parabolic subalgebra in $`g`$ \[Ko61\] and R. Howe’s classification of “skew-multiplicity-free” $`g`$-modules \[Ho95, ch. IV\]. But the general situation still remains unsatisfactory, and developing of Invariant Theory in the skew-symmetric setting represents an attractive problem. In this paper, we begin with describing all irreducible orthogonal $`g`$-modules such that $`(^{}V)^g`$ is again an exterior algebra. It is shown that in this case either $`Vg`$ and hence $`g`$ is simple or $`gV`$ has a structure of simple $`Z_2`$-graded Lie algebra, which quickly leads to a short classification, see Table 1. Obviously, none of the symplectic representations (with $`dimV>2`$) can have an exterior algebra of skew-invariants. But the situation for the representations of “general type” is not yet clear. In case $`V`$ is orthogonal, a better understanding of the $`g`$-module structure of $`^{}V`$ can be achieved through the notion of ‘Spin’ of $`V`$. This goes as follows. Let $`\pi :gso(V)`$ be the corresponding representation. Restricting the spinor representation of $`so(V)`$ to $`g`$ gives us a $`g`$-module, which is denoted by $`Spin(V)`$. The motivation came from Kostant’s result that $`Spin(g)`$ is a primary $`g`$-module; namely, $`Spin(g)=2^{[\mathrm{rk}g/2]}V_\rho `$, the highest weight $`\rho `$ being the half-sum of the positive roots \[Ko61, p. 358\]. The main property of $`Spin(V)`$ is that, depending on parity of $`dimV`$, $`^{}V`$ is isomorphic to either $`Spin(V)^2`$ or $`2Spin(V)^2`$. It is thus interesting to find the orthogonal $`g`$-modules, where $`Spin(V)`$ has a simple structure. In general, $`Spin(V)`$, as element of the representation ring, has a numerical factor depending on the zero-weight multiplicity. Omitting this factor yields a $`g`$-module, which is called the reduced ‘Spin’ of $`V`$ and denoted by $`Spin_0(V)`$; e.g. $`Spin_0(g)=V_\rho `$. In a sense, $`Spin_0(V)`$ behaves better than $`Spin(V)`$. For, regardless of parity of $`dimV`$, we have $`^{}V2^{m(0)}Spin_0(V)^2`$, where $`m(0)`$ is the zero-weight multiplicity, and $`Spin_0(V_1V_2)=Spin_0(V_1)Spin_0(V_2)`$. An orthogonal $`g`$-module $`V`$ is said to be co-primary, if $`Spin_0(V)`$ is irreducible. In sections 1 and 1, a classification of the co-primary modules is obtained. To this end, we give a geometric description of some highest weights of $`Spin_0(V)`$. These weights are called extreme. The assumption that $`Spin_0(V)`$ has a unique extreme weight imposes strong constraints on the weight structure of $`V`$. Using this, one shows that $`g`$ must be simple whenever $`V`$ is an irreducible faithful co-primary $`g`$-module, and that any reducible co-primary module is being obtained by iterating the “direct sum” procedure: $`(g_i,V_i),i=1,2(g_1g_2,V_1V_2)`$. It is thus sufficient to classify the irreducible co-primary modules. The resulting list appears to be rather short: 1. $`g`$ is simple and $`Vg`$; 2. $`g`$ is of type $`𝐁_n`$ or $`𝐂_n`$ or $`𝐅_4`$, and $`V`$ is the little adjoint module; 3. $`g=so(W)`$ and $`V=\{\text{the Cartan component of }𝒮^2W\}`$; $`dimW=3,\mathrm{\hspace{0.17em}5},\mathrm{\hspace{0.17em}7},\mathrm{}`$. In case 2, $`g`$ has roots of two lengths and the $`g`$-module whose highest weight is the short dominant root is called little adjoint (l.a.). Actually, we give a unified proof for the fact that the l. a. module is co-primary whenever the ratio of root lengths is $`\sqrt{2}`$. Note that, for $`𝐆_2`$, where this ratio is $`\sqrt{3}`$, the l. a. module is not co-primary. A true reason why this is so is that the l. a. module for $`𝐆_2`$ is not the isotropy representation of a symmetric space, whereas this is the case for B, C, and F. Furthermore, all representations listed above are the isotropy representations of symmetric spaces. A curious coincidence in this regard is the following. Let $`\stackrel{~}{G}/G`$ be a symmetric space and $`\stackrel{~}{g}=gV`$ the corresponding Lie algebra decomposition. Then the $`g`$-module $`V`$ is co-primary if and only if $`g`$ is non-homologous to zero in $`\stackrel{~}{g}`$. Another by-product of our classifications is that any co-primary module has an exterior algebra of skew-invariants. Having observed that any co-primary representation is a very specific isotropy representation, one may suggest that $`Spin_0(V)`$ admits a nice description for all symmetric spaces. This is really the case, and a transparent formulation can be given for the inner involutory automorphisms. Let $`g=g_0g_1`$ be a $`Z_2`$-grading of inner type, i.e., $`\mathrm{rk}g_0=\mathrm{rk}g`$. As the $`g_0`$-module $`g_1`$ has no zero weight, $`Spin(g_1)=Spin_0(g_1)`$. Choose a common Cartan subalgebra $`t`$ for $`g`$ and $`g_0`$, and consider the natural inclusion of the Weyl groups $`W_0W`$. Although $`W_0`$ is not necessary a parabolic subgroup of $`W`$, each coset $`wW_0`$ contains a unique element of minimal length (see 1). Let $`W^0W`$ be the set of such elements. Then the irreducible constituents of the $`g_0`$-module $`Spin(g_1)`$ are parameterized by $`W^0`$. Namely, $`Spin(g_1)={\displaystyle \underset{wW^0}{}}V_{\lambda _w}`$, where $`\lambda _w=w^1\rho \rho _0`$ is the highest weight, see section 1. Moreover, the weights $`\lambda _w`$ ($`wW^0`$) are distinct and hence $`Spin(g_1)`$ is a multiplicity free $`g_0`$-module. It is worth noting that the above expression for $`Spin(g_1)`$ is equivalent to an identity for root systems that seem to have not been observed before. Let $`\mathrm{\Delta }`$ be the root system of $`(g,t)`$ and let $`\mathrm{\Delta }^+=\mathrm{\Delta }_0^+\mathrm{\Delta }_1^+`$ be the partition of the set of positive roots corresponding to the sum $`g=g_0g_1`$. In this situation, one can introduce the “cunning” parity $`\tau :W\{1,1\}`$, determined by $`\mathrm{\Delta }_0^+`$. If $`wW_0`$, then $`\tau (w)=(1)^{l_0(w)}`$, where $`l_0()`$ is the length in $`W_0`$ relative to the set of positive roots $`\mathrm{\Delta }_0^+`$. To extend $`\tau `$ to $`W`$, one uses the aforementioned subset $`W^0`$ (see section 1). Then the identity reads $$\underset{wW}{}\tau (w)e^{w\rho }=\underset{\alpha \mathrm{\Delta }_0^+}{}(e^{\alpha /2}e^{\alpha /2})\underset{\mu \mathrm{\Delta }_1^+}{}(e^{\mu /2}+e^{\mu /2}).$$ For the outer involutory automorphisms, the final description of $`Spin_0(g_1)`$ is almost identical to the previous one, see section 1. However, it requires much more preparations and its proof uses the classification of involutory automorphisms. Our arguments suggest that there should exist interesting connections between cohomology of symmetric spaces, twisted affine Kac–Moody algebras, and $`Spin(g_1)`$. The description of the highest weights of $`Spin(g_1)`$ (for all involutions!) shows that these weights are extreme. This also implies the following claim (see sect. 7): Let $`\mathrm{\Phi }(,)`$ be an invariant bilinear form on $`g`$ and $`\mathrm{\Phi }(,)_0`$ its restriction to $`g_0`$. Let $`c_0U(g_0)`$ be the Casimir element with respect to $`\mathrm{\Phi }(,)_0`$. Then $`c_0`$ acts scalarly on $`Spin(g_1)`$; the value is $`(\rho ,\rho )(\rho _0,\rho _0)`$, where $`(,)`$ is the $`W`$-invariant bilinear form on $`t^{}`$ induced by $`\mathrm{\Phi }(,)`$. A similar result holds for the isotropy representation $`hso(m)`$ of non-symmetric space $`G/H`$, if $`\mathrm{rk}h=\mathrm{rk}g`$ and one considers the submodule of $`Spin(m)`$ generated by the extreme weight vectors. Recently, B. Kostant obtained a series of nice results for $`Spin(g)`$ \[Ko97\]. Since the adjoint representation is one of the isotropy representations of symmetric spaces, our results for $`Spin(g_1)`$ suggest that many parts of Kostant’s theory can be generalized to the setting of arbitrary symmetric spaces. Main notation. $`g`$ is a reductive Lie algebra with a fixed triangular decomposition: $`g=u^+tu^{}`$. All $`g`$-modules are assumed to be finite-dimensional. $`\mathrm{\Delta }`$ (resp. $`\mathrm{\Delta }^+`$) is the set of roots (resp. positive roots); $`\mathrm{\Pi }\mathrm{\Delta }^+`$ is the set of simple roots; $`\mathrm{\Pi }=\{\alpha _i\}_{iI}`$ and $`\phi _i`$ is the fundamental weight corresponding to $`\alpha _i`$. For simple Lie algebras, we follow the numeration of the simple roots from \[VO88\] and \[On95\]. $`𝒫`$ – the lattice of integral weights, $`𝒫_+`$ – the monoid of dominant integral weights. $`W=N_G(t)/Z_G(t)=N_G(t)/T`$ – the Weyl group; for $`\beta \mathrm{\Delta }`$, $`s_\beta `$ is the reflection in $`W`$. $`𝒫_Q=𝒫_ZQt^{}`$ and $`(,)`$ is the $`W`$-invariant positive-definite scalar product in $`𝒫_Q`$ determined by a non-degenerate invariant bilinear form $`\mathrm{\Phi }(,)`$ on $`g`$. If $`M𝒫`$ is any finite set of weights, then $`|M|=_{mM}m`$; $`\rho :=\frac{1}{2}|\mathrm{\Delta }^+|`$. If $`\lambda 𝒫_+`$, then $`V_\lambda `$ stands for the irreducible $`g`$-module with highest weight $`\lambda `$. Acknowledgements. I would like to thank A.L. Onishchik for conversations about cohomology of compact homogeneous spaces. I am indebted to B. Kostant for drawing my attention to Conlon’s paper \[Co72\]. Part of this work was done while I was visiting Université de Poitiers. I am grateful to this institution for the great hospitality I enjoyed there. 1 Orthogonal $`g`$-modules with an exterior algebra of skew-invariants Let $`V`$ be a $`g`$-module. Study of the algebra $`(𝒮^{}V)^g`$ of symmetric (or polynomial) invariants is the subject of a rich and well-developed theory. In contrast, little is known about the algebra $`(^{}V)^g`$ of skew-invariants. The skew-symmetric theory has some parallels to the symmetric case, and many interesting differences. We begin with two observations. (1.1) Put $`n=dimV`$. Suppose $`gsl(V)`$, e.g. $`g`$ is semisimple. Then $`^nV`$ is a trivial $`g`$-module. Therefore $`dim(^{}V)^g2`$, $`^iV`$ and $`^{ni}V`$ are isomorphic $`g`$-modules, and the Poincaré polynomial of $`(^{}V)^g`$ is symmetric. (1.2) Let $`𝒫(V)`$ be the set of all weights of $`V`$ relative to $`tg`$ and $`V=_{\mu 𝒫(V)}V^\mu `$ the weight decomposition. Set $`m(\mu )=dimV^\mu `$. Recall that the character of $`V`$ is the element of the group algebra $`Z[𝒫]`$ defined by $`\mathrm{ch}V=_{\mu 𝒫(V)}m(\mu )e^\mu `$. Then, $`t`$ being an indeterminate, we have $$\underset{i=1}{\overset{n}{}}(\mathrm{ch}^iV)t^i=\underset{\mu 𝒫(V)}{}(1+te^\mu )^{m(\mu )}.$$ In particular, $`\mathrm{ch}^{}V=2^{m(0)}_{\mu 0}(1+e^\mu )^{m(\mu )}`$. Obviously, $`_{\mu 0}(1+e^\mu )^{m(\mu )}`$ must be the character of a $`g`$-module, say $`W`$. Hence $`1dim(^{}V)^g=2^{m(0)}dimW^g`$ and therefore $`dim(^{}V)^g2^{m(0)}`$. It is natural to first describe $`g`$-modules, where the algebra of skew-invariants has a simple structure. Definition. The algebra $`(^{}V)^g`$ is said to be free (or an exterior algebra), if there exists a graded subspace $`P(^{}V)^g`$ such that $`(^{}V)^g`$ is the exterior algebra over $`P`$. Suppose $`gsl(V)`$ and $`(^{}V)^g`$ is an exterior algebra. Let $`P=p_1,\mathrm{},p_l`$ with $`\mathrm{deg}p_i=d_i`$. Then $`0p_1\mathrm{}p_l`$ must be an element of $`^nV`$. Hence $`_id_i=n`$. It follows from the definition that all the $`d_i`$’s must be odd whenever $`l>1`$. However, if $`l=1`$, then $`d_1=n`$ is allowed to be even. In other words, all 2-dimensional algebras of skew-invariants are proclaimed to be exterior. Example. Let $`V=g`$. Then $`(^{}g)^g`$ is free. Here $`l=\mathrm{rk}g`$ and $`d_i=2m_i+1`$, where $`m_1,\mathrm{},m_l`$ are the exponents of $`g`$. A purely algebraic proof of this result was given by Koszul \[Kos50\]. From now on, $`V`$ is an orthogonal $`g`$-module, i.e., we are given a representation $`\pi :gso(V)`$. In particular, $`gsl(V)`$. 1.3 Lemma. Let $`V`$ be an irreducible orthogonal $`g`$-module with $`V^g=0`$. Suppose $`(^{}V)^g`$ is free. Then either $`(^4V)^g=0`$ or $`dimV=4`$. Proof. Since $`(^1V)^g=(^2V)^g=0`$, any nonzero element of $`(^4V)^g`$ is a generator of $`(^{}V)^g`$. Let $`\mu :V\times V^{}g^{}g`$ be the moment mapping associated with the standard symplectic structure on $`V\times V^{}T^{}(V)`$. Identifying $`V`$ and $`V^{}`$, one obtains an anti-commutative bilinear mapping $`\overline{\mu }:V\times Vg`$. Using the $`g`$-invariant symmetric bilinear forms $`\mathrm{\Phi }(,)`$ and $`(,)_V`$, one may explicitly define $`\overline{\mu }`$ by $$\mathrm{\Phi }(\overline{\mu }(v_1,v_2),g):=(v_2,gv_1)_V,$$ (1.4) where $`v_1,v_2V`$ and $`gv_1`$ is a shorthand for $`\pi (g)v_1`$. This $`\overline{\mu }`$ yields an anti-commutative multiplication, denoted by $`[,]\stackrel{~}{}`$, in $`gV`$: $$[(g_1,v_1),(g_2,v_2)]\stackrel{~}{}:=([g_1,g_2]+\overline{\mu }(v_1,v_2),g_1v_2g_2v_1).$$ The following assertion is stated in \[Co72, p. 152\], in the context of compact group representations, as the “Cartan-Kostant theorem”. It is an easy part of Kostant’s characterization of the isotropy representation of compact homogeneous spaces \[loc. cit\]. 1.5 Proposition. The multiplication $`[,]\stackrel{~}{}`$ satisfies the Jacobi identity if and only if the skew-symmetric $`g`$-invariant 4-form on $`V`$ $$(v_1,v_2,v_3,v_4)\stackrel{\kappa }{}\mathrm{\Phi }(\overline{\mu }(v_1,v_2),\overline{\mu }(v_3,v_4))+\mathrm{\Phi }(\overline{\mu }(v_2,v_3),\overline{\mu }(v_1,v_4))+\mathrm{\Phi }(\overline{\mu }(v_3,v_1),\overline{\mu }(v_2,v_4))$$ is identically equal to zero. Proof. By bilinearity of $`[,]\stackrel{~}{}`$, it suffices to verify the Jacobi identity for 4 sorts of triples: (i) $`(g_1,g_2,g_3)`$, (ii) $`(g_1,g_2,v_1)`$, (iii) $`(g_1,v_1,v_2)`$, (iv) $`(v_1,v_2,v_3)`$, where $`g_ig`$ and $`v_iV`$. The Jacobi identity is always satisfied for cases (i)–(iii), because, respectively, $`g`$ is a Lie algebra, $`V`$ is a $`g`$-module, and $`\overline{\mu }`$ is a homomorphism of $`g`$-modules. For $`v_1,v_2,v_3V`$, the identity means that $`\overline{\mu }(v_1,v_2)v_3+\overline{\mu }(v_2,v_3)v_1+\overline{\mu }(v_3,v_1)v_2=0V`$ or $$(\overline{\mu }(v_1,v_2)v_3+\overline{\mu }(v_2,v_3)v_1+\overline{\mu }(v_3,v_1)v_2,v_4)_V=0$$ for any $`v_4V`$. Using Eq. (1.4), one rewrites the last equality as the condition that the mapping $`\kappa :V^4\text{𝕜}`$ is zero. It is also easily seen that $`\kappa `$ is skew-symmetric and $`g`$-invariant. 1.6 Corollary. If $`(^4V)^g=0`$, then $`gV`$, endowed with multiplication $`[,]\stackrel{~}{}`$, is a $`Z_2`$-graded Lie algebra. Notice that the condition $`(^4V)^g=0`$ is not necessary for $`gV`$ to be a $`Z_2`$-graded Lie algebra. We are going to list all irreducible orthogonal $`g`$-modules $`V`$ such that $`(^{}V)^g`$ is free. 1.7 Theorem. Let $`g`$ be semisimple and $`V`$ a faithful orthogonal irreducible $`g`$-module. Suppose $`(^{}V)^g`$ is free. Then either $`g`$ is simple and $`Vg`$ or $`\stackrel{~}{g}:=gV`$ is a simple $`Z_2`$-graded Lie algebra. Proof. 1. Assume that $`dimV4`$. By Lemma 1 and Corollary 1, it follows that $`[,]\stackrel{~}{}`$ makes $`\stackrel{~}{g}`$ a $`Z_2`$-graded Lie algebra. Let $`a\stackrel{~}{g}`$ be an ideal. Then $`aV`$ and $`ag`$ are $`g`$-stable spaces. (i) If $`aV=V`$, then $`a`$ also contains $`\overline{\mu }(V,V)=[V,V]\stackrel{~}{}`$. Since $`V`$ is faithful, $`\overline{\mu }(V,V)`$ meets all the simple components of $`g`$. Therefore $`a=\stackrel{~}{g}`$. (ii) If $`ag0`$, then, since $`V`$ is faithful, $`(ag)V0`$. That is, $`aV0`$ and we are back in part (i). (iii) If $`aV=0`$ and $`ag=0`$, then the $`g`$-module $`a`$ is isomorphic to its projections to both $`V`$ and $`g`$. Hence $`pr_g(a)apr_V(a)V`$. Therefore $`pr_g(a)`$ is a a simple component of $`g`$. As $`V`$ is a faithful $`g`$-module, we conclude that $`gV`$ and therefore $`g`$ is simple in this case. Here $`\stackrel{~}{g}`$ is the sum of two isomorphic ideals, $`\stackrel{~}{g}aa`$. The subalgebra $`g`$, which is isomorphic to $`a`$, is the diagonal in $`\stackrel{~}{g}`$, and $`V=\{(x,x)xa\}`$. 2. Assume that $`dimV=4`$. Then $`gso_4=so(V)`$. Obviously, $`so_4Vso_5`$, and one easily verifies that $`(^{}V)^g`$ is not free for any proper reductive subalgebra $`g`$ of $`so_4`$. As is mentioned above, $`(^{}g)^g`$ is free. Thus, all other irreducible orthogonal modules with free algebra of skew-invariants arise in connection with $`Z_2`$-gradings of simple Lie algebras. If $`\stackrel{~}{g}=gV`$ is a simple $`Z_2`$-graded Lie algebra, then $`(^{}V)^g`$ is isomorphic to $`H^{}(\stackrel{~}{G}/G)`$, the cohomology ring of the symmetric space $`\stackrel{~}{G}/G`$ \[On95, § 9, n.11\]. The cases, where $`H^{}(\stackrel{~}{G}/G)`$ is an exterior algebra, are well known, see \[On95, § 13, Th. 1\]. Note however that our interpretation of “exterior algebras” is a bit wider. In case $`dimH^{}(\stackrel{~}{G}/G)=2`$, the generator is allowed to be of even degree. The resulting classification is presented in Table 1. If $`V`$ is a symplectic $`g`$-module, then $`(^2V)^g0`$. Thus, $`(^{}V)^g`$ cannot be free unless $`dimV=2`$. If $`V`$ is neither orthogonal nor symplectic, then all known instances of free algebras of skew-invariants are those with $`dim(^{}V)^g=2`$. 2 ‘Spin’ of an orthogonal $`g`$-module and its properties Let $`V`$ be a 𝕜-vector space endowed with a non-degenerate quadratic form $`Q`$. Denote by $`so(V)=so_Q(V)`$ the respective orthogonal Lie algebra and by $`𝒞_Q(V)`$ the Clifford algebra of $`Q`$. Let $`W,W^{}`$ be maximal $`Q`$-isotropic subspaces of $`V`$ and $`WW^{}=0`$. The following relations are well-known in the theory of Clifford algebras (see e.g. \[FH96, § 20.1\]) : (i) $`𝒞_Q(V)\text{End }(^{}W)`$, if $`dimV`$ is even, (ii) $`𝒞_Q(V)\text{End }(^{}W)\text{End }(^{}W^{})`$, if $`dimV`$ is odd. As $`^{}W`$ (or $`^{}W^{}`$) is the underlying space of the spin representation of $`so(V)`$ (in case (i) this representation is the sum of two half-spin representations), we shall write $`Spin(V)`$ in place of $`^{}W`$. The above relations are thought of as isomorphisms of $`so(V)`$-modules. It is well-known (and easily seen) that $`𝒞_Q(V)`$ has an $`so(V)`$-stable filtration such that the associated graded algebra is isomorphic to the exterior algebra of $`V`$. Since in both cases $`Spin(V)`$ is a self-dual module, we obtain the following isomorphisms of $`so(V)`$-modules: $$\begin{array}{c}^{}VSpin(V)Spin(V),\text{ if }dimV\text{ is even},\hfill \\ ^{}V2(Spin(V)Spin(V)),\text{ if }dimV\text{ is odd.}\hfill \end{array}$$ (2.1) Let $`g`$ be a reductive Lie algebra and $`\pi :gso(V)`$ an orthogonal representation. Using $`\pi `$, one may regard $`Spin(V)`$ as $`g`$-module. In this way, we obtain a mapping from the set of orthogonal $`g`$-modules to a set of $`g`$-modules: $`VSpin(V)`$. Of course, the $`g`$-modules of the form $`Spin(V)`$ must satisfy some constraints; e.g. $`dimSpin(V)`$ is a power of 2. Equations (2.1), which can be treated as isomorphisms of $`g`$-modules, suggest that ‘$`Spin`$’ could be used for better understanding of $`g`$-module structure of the exterior algebra of an orthogonal module. The point of departure for our considerations is a simple formula for the character of the $`g`$-module $`Spin(V)`$. Fix some notation, which applies to arbitrary $`g`$-modules (i.e. not necessarily orthogonal ones). Let $`𝒫(V)`$ (resp. $`\mathrm{\Delta }(V)`$) denote the set of all (resp. all nonzero) weights of $`V`$. For instance, $`\mathrm{\Delta }(g)=\mathrm{\Delta }`$. For $`\mu 𝒫(V)`$, $`V^\mu `$ is the corresponding weight space and $`m(\mu )=dimV^\mu `$. If $`V=V_\lambda `$ is irreducible, then the multiplicity is denoted by $`m_\lambda (\mu )`$. Recall that $`V`$ is self-dual if and only if $`\mathrm{\Delta }(V)=\mathrm{\Delta }(V)`$ and $`m(\mu )=m(\mu )`$ for all $`\mu \mathrm{\Delta }(V)`$. Given an orthogonal $`g`$-module $`V`$, let $`\mathrm{\Delta }(V)^+`$ denote an arbitrary subset such that $`\mathrm{\Delta }(V)=\mathrm{\Delta }(V)^+(\mathrm{\Delta }(V)^+)`$. 2.2 Lemma. $`\mathrm{ch}Spin(V)=2^{[m(0)/2]}{\displaystyle \underset{\mu \mathrm{\Delta }(V)^+}{}}(e^{\mu /2}+e^{\mu /2})^{m(\mu )}`$. Proof. Using (1), one obtains $`\mathrm{ch}(^{}V)={\displaystyle \underset{\mu 𝒫(V)}{}}(1+e^\mu )^{m(\mu )}=`$ $$2^{m(0)}\underset{\mu \mathrm{\Delta }(V)^+}{}[(1+e^\mu )(1+e^\mu )]^{m(\mu )}=2^{m(0)}\underset{\mu \mathrm{\Delta }(V)^+}{}(e^{\mu /2}+e^{\mu /2})^{2m(\mu )}.$$ Since $`dimVm(0)`$ is even, comparing with Eq. (2.1) completes the proof. Roughly speaking, Eq. (2.1) asserts that a “square root” of $`^{}V`$ is again a $`g`$-module whenever $`V`$ is orthogonal. Lemma 1 gives a precise form for this. Notice that the transformation from the proof of Lemma can be performed for any self-dual $`g`$-module $`V`$. But the respective “square root” does not yield in general the character of a $`g`$-module. It is convenient to omit the numerical factor in $`\mathrm{ch}Spin(V)`$. The remaining expression is still the character of a $`g`$-module. This module is said to be the reduced Spin of $`V`$ and we write $`Spin_0(V)`$ for it: $$\mathrm{ch}Spin_0(V)=\underset{\mu \mathrm{\Delta }(V)^+}{}(e^{\mu /2}+e^{\mu /2})^{m(\mu )}.$$ (2.3) Several easy properties of $`Spin_0`$ are summarized below. 2.4 Proposition. Let $`V=V^{(1)}`$ and $`V^{(2)}`$ be orthogonal $`g`$-modules. Then (i) $`dimSpin_0(V)=2^{(dimVm(0))/2}`$; (ii) $`^{}V2^{m(0)}Spin_0(V)^2`$; (iii) $`Spin(V)=Spin_0(V)`$ if and only if $`m(0)1`$; (iv) $`Spin_0(V^{(1)}V^{(2)})Spin_0(V^{(1)})Spin_0(V^{(2)})`$; (v) $`Spin_0(V)`$ is a self-dual $`g`$-module. Proof. This immediately follows from (2.1), (1), and (2.3). 2.5 Examples. 1. Our consideration of $`Spin(V)`$ was motivated by the following observation of Kostant, see \[Ko61, p. 358\] and \[Ko97\]. Suppose $`V=g`$ and $`\pi =\mathrm{ad}`$ is the adjoint representation. Then $`^{}g2^{\mathrm{rk}g}(V_\rho V_\rho )`$. This means that $`Spin(g)=2^{[\mathrm{rk}g/2]}V_\rho `$ and $`Spin_0(g)=V_\rho `$. 2. $`g=sl_2`$. We shall write $`R_d`$ in place of $`V_{d\phi _1}`$. Recall that $`R_2=g`$, $`R_d=𝒮^dR_1`$, and $`R_d`$ is orthogonal if and only if $`d`$ is even. Let $`𝒫(R_1)=\{\epsilon ,\epsilon \}`$. Applying Prop. 1, we obtain $`Spin_0(R_{2d})=SpinR_{2d}`$ and $`\mathrm{ch}SpinR_{2d}=_{k=1}^d(e^{k\epsilon }+e^{k\epsilon })`$. It is not hard to compute this character for small values of $`d`$. Here are first few formulas: $`SpinR_2=R_1`$, $`SpinR_4=R_3`$, $`SpinR_6=R_6+R_0`$, $`SpinR_8=R_{10}+R_4`$, $`SpinR_{10}=R_{15}+R_9+R_5`$. It is easily seen that if $`SpinR_{2d}=_{iI}R_{m_i}`$, then $`SpinR_{2(d+1)}_{iI}R_{m_i+d+1}`$. Therefore the number of summands is a nondecreasing function of $`d`$. 3. Let $`W`$ be an arbitrary $`g`$-module. We may regard $`V:=WW^{}`$ as orthogonal $`g`$-module equipped with the quadratic form $`Q((w,w^{})):=w|w^{}`$, where $`(w,w^{})V`$ and $`|`$ is the canonical pairing of $`W`$ and $`W^{}`$. Assuming for simplicity that the weights of $`W`$ and $`W^{}`$ are distinct, we see that $`𝒫(W)`$ can be taken as $`\mathrm{\Delta }(V)^+`$. Therefore $`\mathrm{ch}Spin(V)={\displaystyle \underset{\mu 𝒫(W)}{}}(e^{\mu /2}+e^{\mu /2})^{m(\mu )}=e^\nu {\displaystyle \underset{\mu 𝒫(W)}{}}(1+e^\mu )^{m(\mu )}`$, where $`\nu =\frac{1}{2}_{\mu 𝒫(W)}m(\mu )\mu `$. Whence $$Spin(WW^{})\text{𝕜}_\nu ^{}W\text{𝕜}_\nu ^{}W^{},$$ where $`\text{𝕜}_\nu `$ is 1-dimensional $`g`$-module with character $`\nu `$. Obviously, $`\nu =0`$ if and only if $`gsl(W)`$, e.g. $`g`$ is semisimple. It is not hard to verify that the above formula for $`Spin(WW^{})`$ remains true for all $`W`$. Definition. An orthogonal $`g`$-module $`V`$ is said to be co-primary, if $`Spin_0(V)`$ is irreducible. In this case $`Spin(V)`$ is a primary $`g`$-module. We are going to list all co-primary modules for the semisimple Lie algebras. At the moment, the following examples of such modules are known: $`V=g`$, $`g`$ simple; $`V=R_4`$, $`g=sl_2`$. As a step towards a classification, we describe another series of co-primary modules. Let $`g`$ be a simple Lie algebra having two root lengths. We use subscripts ‘s‘ and ‘l‘ to mark objects related to short and long roots, respectively. For instance, $`\mathrm{\Delta }_s`$ is the set of short roots, $`\mathrm{\Delta }=\mathrm{\Delta }_s\mathrm{\Delta }_l`$, and $`\mathrm{\Pi }_s=\mathrm{\Pi }\mathrm{\Delta }_s`$. Set $`\rho _s=\frac{1}{2}|\mathrm{\Delta }_s^+|`$ and $`\rho _l=\frac{1}{2}|\mathrm{\Delta }_l^+|`$. As usual, $`s_\alpha W`$ is the reflection corresponding to $`\alpha \mathrm{\Delta }`$ and $`s_i:=s_{\alpha _i}`$. 2.6 Lemma. $`\rho _s=_{\alpha _i\mathrm{\Pi }_s}\phi _i`$. Proof. It is easily seen that $`s_i(\rho _s)=\{\begin{array}{cc}\rho _s,& \text{if }\alpha _i\mathrm{\Pi }_l\hfill \\ \rho _s\alpha _i,& \text{if }\alpha _i\mathrm{\Pi }_s\hfill \end{array}`$. Let $`\theta \mathrm{\Delta }^+`$ be the highest root and $`\theta _s`$ the short dominant root. Recall that $`\mathrm{\Delta }_l=W\theta `$, $`\mathrm{\Delta }_s=W\theta _s`$, and $`\theta ^2/\theta _s^2=2`$ or 3. If $`\mu \mathrm{\Delta }`$, then $`\mu ^{}:=2\mu /\mu ^2`$. 2.7 Lemma. Suppose $`\theta ^2/\theta _s^2=2`$ and $`\mu \mathrm{\Delta }_s`$. Then $`(\rho +\rho _s,\mu ^{})`$ is even. Proof. Let $`\mu =_{\alpha _i\mathrm{\Pi }_s}n_i\alpha _i+_{\alpha _j\mathrm{\Pi }_l}m_j\alpha _j`$. Then $`\mu ^{}=_{\alpha _i\mathrm{\Pi }_s}n_i\alpha _i^{}+2_{\alpha _j\mathrm{\Pi }_l}m_j\alpha _j^{}`$. Therefore $`(\rho +\rho _s,\mu ^{})=(2\rho _s+\rho _l,\mu ^{})=(2\rho _s,_{\alpha _i\mathrm{\Pi }_s}n_i\alpha _i^{})+(\rho _l,2_{\alpha _j\mathrm{\Pi }_l}m_j\alpha _j^{})=2(_in_i+_jm_j)`$. The following assertion can be proved using classification, but we give a unified proof. 2.8 Proposition. (i) $`dimV_{\theta _s}=(h+1)m_{\theta _s}(0)`$, where $`h`$ is the Coxeter number of $`g`$; (ii) $`m_{\theta _s}(0)=\mathrm{\#}\mathrm{\Pi }_s`$. Proof. (i) It is clear that $`𝒫(V_{\theta _s})=\{0\}\mathrm{\Delta }_s`$. Moreover, $`m_{\theta _s}(\alpha )=1`$ for all $`\alpha \mathrm{\Delta }_s`$. Applying Freudenthal’s multiplicity formula to $`m_{\theta _s}(0)`$, we obtain $$(\theta _s+2\rho ,\theta _s)m_{\theta _s}(0)=2\underset{\alpha \mathrm{\Delta }^+}{}\underset{t1}{}m_{\theta _s}(t\alpha )(t\alpha ,\alpha )=2\underset{\alpha \mathrm{\Delta }_s^+}{}m_{\theta _s}(\alpha )(\alpha ,\alpha )=2\underset{\alpha \mathrm{\Delta }_s^+}{}(\alpha ,\alpha ).$$ Whence $$(1+(\rho ,\theta _s^{}))m_{\theta _s}(0)=\mathrm{\#}\mathrm{\Delta }_s=dimV_{\theta _s}m_{\theta _s}(0).$$ As $`\theta _s^{}`$ is the highest root in the dual root system $`\mathrm{\Delta }^{}`$, we have $`(\rho ,\theta _s^{})=h1`$. (ii) By part (i), we have $`m_{\theta _s}(0)={\displaystyle \frac{dimV_{\theta _s}m_{\theta _s}(0)}{h}}={\displaystyle \frac{\mathrm{\#}\mathrm{\Delta }_s}{h}}`$. Let $`cW`$ be a Coxeter element associated with $`\mathrm{\Pi }`$. It is known that each orbit of $`c`$ in $`\mathrm{\Delta }`$ has cardinality $`h`$ and contains a unique simple root, see \[BOU, ch.VI, § 1, Prop. 33\]. Hence $`\mathrm{\#}\mathrm{\Delta }_s=h(\mathrm{\#}\mathrm{\Pi }_s)`$. Some authors call $`V_{\theta _s}`$ the little adjoint module. To a great extent, properties of $`V_{\theta _s}`$ are similar with properties of $`g`$. 2.9 Theorem. Suppose $`\theta ^2/\theta _s^2=2`$. Then $`^{}V_{\theta _s}2^{\mathrm{\#}\mathrm{\Pi }_s}(V_{\rho _s}V_{\rho _s})`$. Proof. By Proposition 1, we have $`\mathrm{ch}V_{\theta _s}=\mathrm{\#}\mathrm{\Pi }_s+_{\alpha \mathrm{\Delta }_s}e^\alpha `$. Therefore $$\mathrm{ch}^{}V_{\theta _s}=2^{\mathrm{\#}\mathrm{\Pi }_s}\underset{\alpha \mathrm{\Delta }_s}{}(1+e^\alpha )=2^{\mathrm{\#}\mathrm{\Pi }_s}\underset{\alpha \mathrm{\Delta }_s^+}{}(e^{\alpha /2}+e^{\alpha /2})^2.$$ Thus, the statement of theorem is equivalent to that $$\mathrm{ch}V_{\rho _s}=\underset{\alpha \mathrm{\Delta }_s^+}{}(e^{\alpha /2}+e^{\alpha /2})=e^{\rho _s}\underset{\alpha \mathrm{\Delta }_s^+}{}(1+e^\alpha ).$$ (2.10) By Weyl’s character formula $$\mathrm{ch}V_{\rho _s}=\frac{_{wW}\epsilon (w)e^{w(\rho +\rho _s)}}{_{wW}\epsilon (w)e^{w\rho }}=\frac{_{wW}\epsilon (w)e^{w(\rho +\rho _s)}}{_{\alpha \mathrm{\Delta }^+}(e^{\alpha /2}e^{\alpha /2})}.$$ (2.11) Here $`\epsilon (w)=(1)^{l(w)}`$, where $`l(w)`$ is the length of $`w`$ with respect to $`\mathrm{\Delta }^+`$. Take $`\alpha \mathrm{\Delta }_s^+`$. We are going to prove that $`1+e^\alpha `$ divides $`\mathrm{ch}V_{\rho _s}`$ in $`Z[𝒫]`$. Since $`\mathrm{ch}V_{\rho _s}`$ is $`W`$-invariant, it is enough to consider the case in which $`\alpha `$ is simple, i.e., $`\alpha \mathrm{\Pi }_s`$. Actually, we shall prove that $`1+e^\alpha `$ divides the numerator in Eq. (2.11). For this, we show how to group together the summands of the numerator. Let $`W^\alpha =\{wWw^1\alpha \mathrm{\Delta }^+\}`$. Then $`W`$ is the disjoint union of pairs $`\{s_\alpha w,w\}`$ ($`wW^\alpha `$). Consider the corresponding pairs of summands in the numerator of (2.11). Since $`\alpha \mathrm{\Pi }_s`$, we have $`\epsilon (s_\alpha w)=\epsilon (w)`$ and $$\epsilon (w)e^{w(\rho +\rho _s)}+\epsilon (s_\alpha w)e^{s_\alpha w(\rho +\rho _s)}=\epsilon (w)e^{w(\rho +\rho _s)}(1e^{n\alpha }),$$ where $`n=(w(\rho +\rho _s),\alpha ^{})=(\rho +\rho _s,(w^1\alpha )^{})`$. By the definition of $`W^\alpha `$, $`n`$ is positive. The divisibility will follow from the fact that $`n`$ is even. But this is just Lemma 1. Since $`Z[𝒫]`$ is factorial and the factors $`1+e^\alpha `$ ($`\alpha \mathrm{\Delta }_s^+`$) are coprime (see \[BOU, ch. VI, § 3, Lemma 1\]), $`e^{\rho _s}_{\alpha \mathrm{\Delta }_s^+}(1+e^\alpha )`$ divides $`\mathrm{ch}V_{\rho _s}`$. The quotient is a $`W`$-invariant element of $`Z[𝒫]`$. Comparing the maximal terms in both expressions, we see that the quotient must be equal to 1. 2.12 Corollary. 1. $`Spin_0(V_{\theta _s})=V_{\rho _s}`$; 2. $`dim(^{}V_{\theta _s})^g=2^{\mathrm{\#}\mathrm{\Pi }_s}`$. 2.13 Examples. To realize the scope of Theorem 1, we look at all simple Lie algebras with two root lengths. 1. $`g=sp_{2n}`$. Here $`\theta =2\phi _1`$, $`\theta _s=\phi _2`$, and $`\rho _s=\phi _1+\mathrm{}+\phi _{n1}`$. Thus $$^{}V_{\phi _2}=2^{n1}(V_{\phi _1+\mathrm{}+\phi _{n1}})^2\text{and}Spin_0(V_{\phi _2})=V_{\phi _1+\mathrm{}+\phi _{n1}}.$$ 2. $`g=f_4`$. Here $`\theta =\phi _4`$, $`\theta _s=\phi _1`$, and $`\rho _s=\phi _1+\phi _2`$. Thus $$^{}V_{\phi _1}=4(V_{\phi _1+\phi _2})^2\text{and}Spin_0(V_{\phi _1})=V_{\phi _1+\phi _2}.$$ 3. $`g=so_{2n+1}`$. Here $`\theta =\phi _2`$, $`\theta _s=\phi _1`$, and $`\rho _s=\phi _n`$. In this case $`Spin_0(V_{\phi _1})=Spin(V_{\phi _1})=V_{\phi _n}`$ and the formula of Theorem 1 is nothing but the second equality in Eq. (2.1). Hence the theorem also yields another approach to defining ‘Spin’ of an orthogonal representation. 4. $`g=g_2`$. Here $`\theta ^2/\theta _s^2=3`$ and Theorem 1 does not apply. In this case $`\rho _s=\theta _s=\phi _1`$ and $`\theta =\phi _2`$. An explicit (easy) computation with characters shows that $`^{}V_{\phi _1}=2(V_{\phi _1}\mathrm{I}1)^2`$, i.e., $`Spin(V_{\phi _1})=V_{\phi _1}\mathrm{I}1`$. Hence $`V_{\theta _s}`$ is not co-primary. Here $`\mathrm{I}1`$ stands for the trivial 1-dimensional module. (2.14) Another proof of Theorem 1. Making use of Weyl’s character formula, we interpret Eq. (2.10) as Weyl’s denominator identity for the dual root system. Recall that $`\mathrm{\Delta }=\mathrm{\Delta }_l\mathrm{\Delta }_s`$ and we assume that $`\theta ^2/\theta _s^2=2`$. The dual root system is therefore isomorphic to $`\stackrel{~}{\mathrm{\Delta }}:=\mathrm{\Delta }_l2\mathrm{\Delta }_s`$. Here $`(\stackrel{~}{\mathrm{\Delta }})_l=2\mathrm{\Delta }_s`$ and $`(\stackrel{~}{\mathrm{\Delta }})_s=\mathrm{\Delta }_l`$. Since $`\stackrel{~}{W}W`$, Weyl’s denominator identity for $`\stackrel{~}{\mathrm{\Delta }}`$ reads $$\underset{wW}{}\epsilon (w)e^{w\stackrel{~}{\rho }}=\underset{\alpha \stackrel{~}{\mathrm{\Delta }}^+}{}(e^{\alpha /2}e^{\alpha /2}).$$ We have $`\stackrel{~}{\rho }=\rho +\rho _s`$ on the left hand side and $$\underset{\alpha \mathrm{\Delta }_l^+}{}(e^{\alpha /2}e^{\alpha /2})\underset{\mu \mathrm{\Delta }_s^+}{}(e^\mu e^\mu )=\underset{\alpha \mathrm{\Delta }^+}{}(e^{\alpha /2}e^{\alpha /2})\underset{\alpha \mathrm{\Delta }_s^+}{}(e^{\mu /2}+e^{\mu /2})$$ on the right hand side. Hence dividing Weyl’s identity by $`_{\alpha \mathrm{\Delta }^+}(e^{\alpha /2}e^{\alpha /2})`$ yields $`\mathrm{ch}V_{\rho _s}=_{\mu \mathrm{\Delta }_s^+}(e^{\mu /2}+e^{\mu /2})`$. Remark. The previous argument suggests a proper analogue of (2.10) for the exceptional Lie algebra $`g_2`$. Here the dual root system is isomorphic to $`\stackrel{~}{\mathrm{\Delta }}:=\mathrm{\Delta }_l3\mathrm{\Delta }_s`$ and a similar transformation proves that $`\mathrm{ch}V_{2\rho _s}=_{\mu \mathrm{\Delta }_s^+}(e^\mu +1+e^\mu )`$. 3 Classification of co-primary $`g`$-modules In this section, $`g`$ is a semisimple Lie algebra and $`V`$ an orthogonal $`g`$-module. From Eq. (2.3) it is clear that $`V^g`$ has no affect on $`Spin_0(V)`$. We may therefore assume that $`V^g=0`$. 3.1 Proposition. Suppose $`V`$ is co-primary. Then there exist decompositions $`g=g_1\mathrm{}g_s`$, $`V=V_1\mathrm{}V_s`$ such that (i) Each $`g_i`$ is a (semisimple) ideal of $`g`$, (ii) $`g_i`$ acts trivially on $`V_j`$ ($`ij`$), (iii) $`V_i`$ is an irreducible orthogonal co-primary $`g_i`$-module. Proof. Assume that $`V=V_1V_2`$, where $`V_1`$ and $`V_2`$ are orthogonal $`g`$-modules. It follows from the assumptions and Proposition 1(iv) that the $`g`$-module $`Spin_0(V_1)Spin_0(V_2)`$ is irreducible. Since both factors are non-trivial, the only possibility for this is that $`g=g_1g_2`$, where $`g_i`$ acts trivially on $`V_j`$ ($`ij`$) and $`V_i`$ is a co-primary $`g_i`$-module ($`i=1,2`$). Repeating this procedure, we obtain a decomposition satisfying (i) and (ii), where each $`V_i`$ is orthogonal co-primary and is not a sum of two proper orthogonal $`g_i`$-submodules. Then either $`V_i`$ is irreducible or $`V_i=W_iW_i^{}`$, where $`W_i`$ is already irreducible. In the second case, we have $`Spin(V_i)^{}W_i`$ (see Example 1(3)). It is easily seen that the $`g_i`$-module $`^{}W_i`$ is never primary, i.e., $`Spin_0(V_i)`$ can not be irreducible here. Whenever $`(g,V)`$ admits a decomposition satisfying conditions (i) and (ii) of the Proposition, this will be denoted by $`(g,V)=(g_1,V_1)\mathrm{}(g_s,V_s)`$. Notice that if each $`V_i`$ is irreducible, then all the summands in the above decomposition are uniquely determined. 3.2 Lemma. If $`V_\lambda `$ is an irreducible co-primary $`g`$-module, then $`m_\lambda (0)0`$. Proof. If $`m_\lambda (0)=0`$, then $`^{}V_\lambda Spin_0(V_\lambda )^2`$, see 1(ii). Since $`dim(^{}V_\lambda )^g2`$, the Schur lemma shows that $`Spin_0(V_\lambda )`$ cannot be irreducible. It follows from the above two assertions that $`𝒫(V)`$ lies in the root lattice whenever $`V`$ is co-primary. Let us present an explicit way for finding some irreducible constituents of $`Spin_0(V)`$. To write an expression for $`\mathrm{ch}Spin_0(V)`$ in (2.3), we exploited an arbitrary ‘half’ $`\mathrm{\Delta }(V)^+`$ of $`\mathrm{\Delta }(V)`$. However a clever choice of $`\mathrm{\Delta }(V)^+`$ will provide us with a maximal term in $`\mathrm{ch}Spin_0(V)`$ and hence with a highest weight. Take $`\nu 𝒫_+`$ such that $`(\nu ,\mu )0`$ for all $`\mu \mathrm{\Delta }(V)`$. Put $`\mathrm{\Delta }(V)_\nu ^+=\{\mu \mathrm{\Delta }(V)(\mu ,\nu )>0\}`$. A subset of such form is said to be a dominant half of $`\mathrm{\Delta }(V)`$. Set $`\mathrm{\Lambda }_\nu :=\frac{1}{2}_\mu m(\mu )\mu `$, where $`\mu `$ ranges over $`\mathrm{\Delta }(V)_\nu ^+`$. 3.3 Lemma. $`\mathrm{\Lambda }_\nu `$ is a highest weight of $`Spin_0(V)`$. Proof. We show that $`\mathrm{\Lambda }_\nu `$ is dominant and it is a maximal element in $`𝒫(Spin_0(V))`$. Note that the first part is not tautological. We exploit formula 2.3 with $`\mathrm{\Delta }(V)_\nu ^+`$: $$\mathrm{ch}Spin_0(V)=\underset{\mu \mathrm{\Delta }(V)_\nu ^+}{}(e^{\mu /2}+e^{\mu /2})^{m(\mu )}.$$ This shows that $`e^{\mathrm{\Lambda }_\nu }`$ occurs in $`\mathrm{ch}Spin_0(V)`$ with coefficient 1, $`(\nu ,\mathrm{\Lambda }_\nu )=\underset{\mu 𝒫(Spin_0(V))}{\mathrm{max}}(\nu ,\mu )`$, and $`\mathrm{\Lambda }_\nu `$ is the unique element of $`𝒫(Spin_0(V))`$, where the maximal value is attained. Let $`\mathrm{\Lambda }_\nu ^{}`$ be the dominant representative in $`W\mathrm{\Lambda }_\nu `$. Then $`\mathrm{\Lambda }_\nu ^{}𝒫(Spin_0(V))`$ and $`\mathrm{\Lambda }_\nu ^{}\mathrm{\Lambda }_\nu =_{\alpha _i\mathrm{\Pi }}n_i\alpha _i`$ with $`n_i0`$. Therefore $`(\nu ,\mathrm{\Lambda }_\nu ^{})(\nu ,\mathrm{\Lambda }_\nu )`$ and hence $`\mathrm{\Lambda }_\nu ^{}=\mathrm{\Lambda }_\nu `$. (3.4) The highest weights of $`Spin_0(V)`$ of the form $`\mathrm{\Lambda }_\nu `$ are said to be extreme. It is easy to describe all dominant halfs of $`\mathrm{\Delta }(V)`$ and hence all extreme weights of $`Spin_0(V)`$. Consider the Weyl chamber $`C:=Q_+𝒫_+𝒫_Q`$ and its interior $`C^o`$. Let $`H_\mu `$ denote the hyperplane in $`𝒫_Q`$ orthogonal to $`\mu 𝒫`$. Recall that $`C^o`$ is the connected component<sup>2</sup><sup>2</sup>2Strictly speaking, use of the term “connected component” is correct only for the real vector space $`𝒫_R`$. of $`𝒫_Q_{\gamma \mathrm{\Delta }}H_\gamma `$, containing dominant weights. Then the hyperplanes $`H_\mu `$ ($`\mu \mathrm{\Delta }(V)`$) cut $`C`$ in smaller chambers. When $`\nu `$ varies inside of such a ‘small’ chamber the corresponding extreme weight does not change. We thus obtain a bijection $$\{\text{extreme weights of }Spin(V)\}\{\text{connected components of }C^o\underset{\mu \mathrm{\Delta }(V)}{}H_\mu \}.$$ In particular, $`Spin_0(V)`$ has a unique extreme weight if and only if $`\mathrm{\Delta }(V)`$ has a unique dominant half if and only if none of the hyperplanes $`H_\mu `$ cuts $`C^o`$. 3.5 Lemma. Suppose $`\mathrm{\Delta }(V)`$ lies in the root lattice. Then: none of the hyperplanes $`H_\mu `$ ($`\mu \mathrm{\Delta }(V)`$) cuts $`C^o`$ $``$ $`\mathrm{\Delta }(V)_{\alpha \mathrm{\Delta }}Z\alpha `$. Proof. “$``$” This is obvious. $``$” Assume that $`M:=\mathrm{\Delta }(V)_{\alpha \mathrm{\Delta }}Z\alpha \text{}`$. Let $`\mu M𝒫_+`$ be an element closest to 0. Write $`\mu `$ as sum of positive roots with positive integral coefficients $`\mu =_{i=1}^dk_i\gamma _i`$ ($`\gamma _i\gamma _j`$) and so that $`_ik_i`$ is minimal over all such presentations. Then $`\gamma _i+\gamma _j`$ is not a root, i.e., $`(\gamma _i,\gamma _j)0`$. Therefore $`(\mu ,\gamma _1)>0`$ and hence $`\mu \gamma _1\mathrm{\Delta }(V)`$. As $`\mu \gamma _1<\mu `$, we obtain $`\mu \gamma _1_{\alpha \mathrm{\Delta }^+}N\alpha `$. Thus, $`k_1=1`$, $`d=2`$ and, by symmetry, $`\mu =\gamma _1+\gamma _2`$. Since $`(\gamma _1,\gamma _2)0`$ and $`\gamma _1+\gamma _2`$ is not a multiple of a root, it is easily seen that $`(\gamma _1,\gamma _2)`$ is a basis of the root system $`\mathrm{\Delta }(Q\gamma _1+Q\gamma _2)`$. Therefore $`(\gamma _1,\gamma _2)`$ is $`W`$-conjugate to a pair of simple roots $`(\alpha _i,\alpha _j)`$ (see \[BOU, ch. VI, § 1,Prop. 24\]). Thus, $`\alpha _i\alpha _j\mathrm{\Delta }(V)`$ and $`H_{\alpha _i\alpha _j}`$ cuts $`C^o`$. 3.6 Proposition. Let $`V`$ be a co-primary faithful irreducible $`g`$-module. Then $`\mathrm{\Delta }(V)_{\alpha \mathrm{\Delta }}Z\alpha `$ and $`g`$ is simple. Proof. By Lemma 1, $`\mathrm{\Delta }(V)`$ lies in the root lattice. Therefore the first claim readily follows from (S3.Ex16) and Lemma 1. Assume that $`g=g_1g_2`$ is a sum of two ideals. Then $`V=V_1V_2`$, where $`V_i`$ is a non-trivial $`g_i`$-module. Obviously, if $`\mu _i\mathrm{\Delta }(V_i)`$ ($`i=1,2`$), then $`\mu _1+\mu _2\mathrm{\Delta }(V)`$ and it is not a multiple of a root of $`g`$. Now, we are ready to state a classification. 3.7 Theorem. (i) Let $`g`$ be semisimple and $`V`$ a faithful orthogonal $`g`$-module with $`V^g=0`$. Suppose $`V`$ is co-primary. Then $$(g,V)=(g_1,V_1)\mathrm{}(g_s,V_s),$$ where each $`g_i`$ is simple and $`V_i`$ is irreducible and co-primary. Each weight of $`V`$ is a multiple of a root of $`g`$. (ii) If $`g`$ is simple and $`V=V_\lambda `$ is irreducible and co-primary, then the pair $`(g,\lambda )`$ is one of the following: (a) $`g`$ is any and $`\lambda =\theta `$. (b) $`g\{so_{2n+1},sp_{2n},f_4\}`$ and $`\lambda =\theta _s`$. (c) $`g=so_{2n+1}`$, $`\lambda =2\theta _s=2\phi _1`$ ($`n2`$). (d) $`g=sl_2`$, $`\lambda =4\phi _1`$. Proof. (i) By Proposition 1, such a decomposition with irreducible and co-primary summands $`V_i`$ exists. The other assertions are proved in Proposition 1. (ii) By part (i), we have $`\lambda \{k\theta ,k\theta _skN\}`$. Let $`\mathrm{rk}g=1`$. It follows from Example 1(2) that the only co-primary $`sl_2`$-modules are $`sl_2=R_2`$ and $`R_4`$. Let $`\mathrm{rk}g2`$. Consider the following possibilities. $``$ $`\lambda =2\theta `$. Take $`\alpha _i\mathrm{\Pi }`$ such that $`(\alpha _i,\theta )0`$. Then $`2\theta \alpha _i`$ is a weight of $`V_{2\theta }`$, which is not a multiple of a root. Thus, $`V_{2\theta }`$ is not co-primary. $``$ $`g=sp_{2n}`$ or $`f_4`$ and $`\lambda =2\theta _s`$. If $`\alpha _i`$ is the unique simple root such that $`(\alpha _i,\theta _s)0`$, then $`2\theta _s\alpha _i\mathrm{\Delta }(V_{2\theta _s})`$ is not a multiple of a root. $``$ $`g=so_{2n+1}`$ and $`\lambda =3\theta _s=3\phi _1`$. Here $`3\phi _1\alpha _1`$ is not proportional to a root. $``$ $`g=g_2`$. We have already shown in Example S2.Ex11(4) that $`V_{\theta _s}`$ is not co-primary. Obviously, if $`V_{k\lambda }`$ is not co-primary, then the same holds for any $`mk`$. Thus, comparing with results of section 1, we see that the only unclear case is ii(c). Our proof that this module is co-primary is similar to the first proof of Theorem 1. It will be given in the next proposition, where we also compute the reduced $`Spin`$ of $`V_{2\theta _s}`$. 3.8 Proposition. Let $`g=so_{2n+1}`$. Then 1. $`Spin_0(V_{2\phi _1})=V_{\rho +2\phi _n}`$; 2. $`^{}V_{2\phi _1}=2^n(V_{\rho +2\phi _n})^2`$. Proof. First, we describe the weight structure of the $`g`$-module $`V_{2\phi _1}`$. This is easy, since $`V_{2\phi _1}`$ is the Cartan (highest) component in $`𝒮^2V_{\phi _1}`$. Here $`𝒫(V_{2\phi _1})=\{0\}\mathrm{\Delta }2\mathrm{\Delta }_s`$. Hence $`\mathrm{\Delta }(V)^+=\mathrm{\Delta }^+2\mathrm{\Delta }_s^+`$. The non-zero weights are of multiplicity 1, and $`m_{2\phi _1}(0)=n`$. Therefore, making use of Eq. (2.3), we obtain $$\mathrm{ch}Spin_0(V_{2\phi _1})=\underset{\alpha \mathrm{\Delta }^+}{}(e^{\alpha /2}+e^{\alpha /2})\underset{\alpha \mathrm{\Delta }_s^+}{}(e^\alpha +e^\alpha )=e^{\rho +2\phi _n}\underset{\alpha \mathrm{\Delta }^+}{}(1+e^\alpha )\underset{\alpha \mathrm{\Delta }_s^+}{}(1+e^{2\alpha })=$$ $$=e^{\rho +2\phi _n}\underset{\alpha \mathrm{\Delta }_l^+}{}(1+e^\alpha )\underset{\alpha \mathrm{\Delta }_s^+}{}(1+e^\alpha +e^{2\alpha }+e^{3\alpha }).$$ On the other hand, $$\mathrm{ch}V_{\rho +2\phi _n}=\frac{_{wW}\epsilon (w)e^{w(\rho +\rho +2\phi _n)}}{_{\alpha \mathrm{\Delta }^+}(e^{\alpha /2}e^{\alpha /2})}.$$ (3.9) Since $`\mathrm{ch}Spin_0(V_{2\phi _1})`$ and $`\mathrm{ch}V_{\rho +2\phi _n}`$ have the same maximal term $`e^{\rho +2\phi _n}`$, it suffices to prove that each factor in the last expression for $`\mathrm{ch}Spin_0(V_{2\phi _1})`$ divides $`\mathrm{ch}V_{\rho +2\phi _n}`$, i.e., the numerator in Eq. (3.9). The same procedure, as in the proof of Theorem 1, reduces the problem to proving that, for any $`wW^\alpha `$, $$(w(2\rho +2\phi _n),\alpha ^{})\{\begin{array}{ccc}\text{is even}\hfill & ,& \text{ if }\alpha \mathrm{\Pi }_l\hfill \\ \text{is divisible by 4}\hfill & ,& \text{ if }\alpha \mathrm{\Pi }_s\hfill \end{array}.$$ That is, we need actually to verify that $`(w(\rho +\phi _n),\alpha ^{})`$ is even whenever $`\alpha `$ is short. As $`\phi _n=\rho _s`$ for our $`g`$, this is just Lemma 1. 2. This is a formal consequence of part 1, see Proposition 1. Having obtained the list of all irreducible co-primary modules in Theorem 1(ii), it is worth looking it through again in order to find out common features and latent regularities for the representations in question. First, item (ii)d in (1) can be thought of as starting point for the series in (ii)c. Indeed, $`V_{2\phi _1}`$ is the Cartan component in $`𝒮^2V_{\phi _1}`$ and $`V_{\phi _1}`$ is the tautological module for $`so_{2n+1}`$ ($`n2`$), whereas $`sl_2`$-module $`R_4`$ is the Cartan component in $`𝒮^2R_2`$ and $`R_2`$ is the tautological module for $`so_3`$. Thus, the list consists of three groups of representations: 1. $`(g,V_\theta =g)`$; 2. $`(g,V_{\theta _s})`$, where $`g`$ is of type $`𝐁`$, $`𝐂`$, or $`𝐅`$; 3. $`g=so(W)`$ and $`V=𝒮_0^2(W)`$, where $`dimW=3,5,7,\mathrm{}`$ The second (more interesting) observation is that, for all items $`(g,V)`$ in the list, $`gso(V)`$ is the isotropy representation of an irreducible symmetric space. In other words, $`\stackrel{~}{g}:=gV`$ has a structure of irreducible $`Z_2`$-graded semisimple Lie algebra. More precisely, $`\stackrel{~}{g}`$ is simple for items 2 and 3, and $`\stackrel{~}{g}gg`$ for item 1. Furthermore, it follows from the well-known classification of symmetric spaces that items 1–3 correspond exactly to the cases, where $`g`$ is non-homologous to zero<sup>3</sup><sup>3</sup>3this means that the canonical map of homology spaces $`H_{}(g)H_{}(\stackrel{~}{g})`$ is injective. in $`\stackrel{~}{g}`$. The class of homogeneous spaces $`\stackrel{~}{G}/G`$ (not necessarily symmetric ones) such that $`\stackrel{~}{G},G`$ are connected and $`g`$ is non-homologous to zero in $`\stackrel{~}{g}`$ has many nice descriptions. We refer the reader to \[On95, § 13, n.2\] for a thorough treatment in the context of homogeneous spaces of compact Lie groups. In the symmetric case, yet another characterization is that this happens if and only if $`g`$ is determined by a diagram involutory automorphism of $`\stackrel{~}{g}`$. An explicit description of the diagram automorphisms of simple Lie algebras is found in \[Ka90, § 7.9, 7.10\]. The third observation is that any irreducible co-primary module occurs in Table 1 in section 1, i.e., it has a free algebra of skew-invariants. These observations give us some hope that the reduced $`Spin`$ of the isotropy representation of an arbitrary symmetric spaces might have some interesting properties. This is really the case and we turn to such considerations in the following sections. 4 Some auxiliary results In this section, we prove an auxiliary result on Weyl groups and recall some standard facts on involutions of simple Lie algebras. Let $`W`$ be the Weyl corresponding to a reduced root system $`\mathrm{\Delta }`$ with a set of positive roots $`\mathrm{\Delta }^+`$. Let $`wl(w)`$ be the length function on $`W`$ determined by $`\mathrm{\Delta }^+`$. Recall that $`l`$ can be defined as $`l(w)=\mathrm{\#}\{\alpha \mathrm{\Delta }^+w(\alpha )\mathrm{\Delta }^{}\}`$. Consider an arbitrary subset $`\mathrm{\Delta }_0\mathrm{\Delta }`$ which is a root system in its own right, but is not necessarily closed in $`\mathrm{\Delta }`$. That is, it is allowed that $`\alpha +\beta \mathrm{\Delta }\mathrm{\Delta }_0`$ for some $`\alpha ,\beta \mathrm{\Delta }_0`$. It is easily seen that such a phenomenon can only occur if $`\mathrm{\Delta }`$ has roots of different length. As a sample of such non-closed subset, we mention $`\mathrm{\Delta }_0=\mathrm{\Delta }_s`$. Nevertheless, $`W_0`$, the Weyl group of $`\mathrm{\Delta }_0`$, is always identified with a subgroup of $`W`$. Clearly, $`\mathrm{\Delta }_0^+:=\mathrm{\Delta }_0\mathrm{\Delta }^+`$ can be taken as set of positive roots for $`\mathrm{\Delta }_0`$. 4.1 Proposition. 1. Any coset $`wW_0W`$ contains a unique representative of minimal length. Denoting by $`W^0`$ the set of minimal length representatives, we have $`W^0=\{wWw(\mathrm{\Delta }_0^+)\mathrm{\Delta }^+\}`$. 2. The mapping $`W^0\times W_0W`$ ($`(w^o,w_o)w_o(w^o)^1`$) is a bijection. Proof. 1. Set $`W^{}=\{wWw(\mathrm{\Delta }_0^+)\mathrm{\Delta }^+\}`$. Then $`W^0W^{}`$. Indeed, assume that $`wW^0`$ and $`w(\beta )\mathrm{\Delta }^{}`$ for some $`\beta \mathrm{\Delta }_0^+`$. Then $`ws_\beta (\beta )\mathrm{\Delta }^+`$ and it follows from \[BGG, 2.3\] that $`l(ws_\beta )<l(w)`$. But this contradicts the fact that $`wW_0`$. Obviously, each coset contains elements of minimal length and hence elements from $`W^{}`$. Assume that $`u,vW^{}vW_0`$. Then $`u=vw`$ for some $`wW_0`$. If $`we`$, then $`w(\beta )\mathrm{\Delta }_0^{}`$ for some $`\beta \mathrm{\Delta }_0^+`$. Whence $`u(\beta )=v(w(\beta ))\mathrm{\Delta }^{}`$, which contradicts the assumption. Thus, each coset contains a unique element of $`W^{}`$, $`W^0=W^{}`$, and we are done. 2. Obvious. Remark. If $`\mathrm{\Delta }_0`$ is generated by a part of the basis $`\mathrm{\Pi }\mathrm{\Delta }^+`$ (i.e., $`\mathrm{\Delta }_0\mathrm{\Pi }`$ is a basis of $`\mathrm{\Delta }_0`$), then $`W_0`$ is a parabolic subgroup of $`W`$. In this case the Proposition is well known and, moreover, the relation $`l(w^ow_0)=l(w^o)+l(w_o)`$ holds, see e.g. \[Hu95, 1.10\]. However this relation does not hold in general. For $`w_oW_0`$, let $`l_0(w_o)`$ denote the length of $`w_o`$ in $`W_0`$. That is, $`l_0(w_o)=\mathrm{\#}\{\mu \mathrm{\Delta }_0^+w_o(\mu )\mathrm{\Delta }_0^{}\}`$. If $`W_0`$ is a parabolic subgroup, then $`l_0(w_o)=l(w_o)`$, but in general we have only “$``$”. The usual determinant or parity for the elements of $`W`$ is defined by $`\epsilon (w)=(1)^{l(w)}`$. Making use of the above bijection, one may introduce a parity depending on $`\mathrm{\Delta }_0`$. By Prop. 1(2), each element $`wW`$ has a unique presentation $`w=w_o(w^o)^1`$, where $`w_oW_0`$ and $`w^oW^0`$. Set $`l_0(w):=l_0(w_o)`$ and $`\tau (w):=(1)^{l_0(w)}`$. So, if $`w=w_o`$, then $`\tau (w_o)`$ is nothing but the usual parity on $`W_0`$, which will be denoted by $`\epsilon _0(w_o)`$. Therefore one may say that $`\tau `$ is the extension of the parity $`\epsilon _0`$ to $`W`$ determined by the ‘section’ $`W^0`$. The function $`wW\tau (w)\{1,1\}`$ is said to be the cunning parity on $`W`$, determined by $`\mathrm{\Delta }_0^+`$ (or $`W_0`$). It is convenient to give an expression for $`l_0(w)`$, and hence for $`\tau (w)`$, where $`w_o`$ is not explicitly mentioned. 4.2 Lemma. $`l_0(w)=\mathrm{\#}\{\alpha \mathrm{\Delta }^{}w(\alpha )\mathrm{\Delta }_0^+\}`$. Proof. Let $`w=w_o(w^o)^1`$, as above. Consider the subsets $`M_1=\{\alpha \mathrm{\Delta }^{}w(\alpha )\mathrm{\Delta }_0^+\}`$ and $`M_2=\{\mu \mathrm{\Delta }^{}w_o(\mu )\mathrm{\Delta }_0^+\}`$. Since $`\mathrm{\Delta }_0`$ is $`W_0`$-stable, $`M_2\mathrm{\Delta }_0^{}`$ and therefore $`l_0(w_o)=\mathrm{\#}M_2`$. By Prop. 1(1), we have $`(w^o)^1M_1\mathrm{\Delta }_0^+=\text{}`$. Since $`w_o((w^o)^1M_1)\mathrm{\Delta }_0^+`$, we see that $`(w_o)^1M_1\mathrm{\Delta }_0^{}`$. Thus $`(w^o)^1M_1M_2`$. Similarly, one proves the opposite containment. Thus, $`l_0(w)=\mathrm{\#}M_2=\mathrm{\#}M_1`$, and we are done. (4.3) Classes of involutory automorphisms. Here $`g`$ is a simple Lie algebra. Given an involutory automorphism $`\mathrm{\Theta }`$ of $`g`$, consider the $`Z_2`$-grading $`g=g_0g_1`$, where $`g_i=\{xg\mathrm{\Theta }(x)=(1)^ix\}`$. The reductive subalgebra $`g_0`$ is called symmetric. The involutory automorphisms fall into three classes: a) $`\mathrm{rk}g=\mathrm{rk}g_0`$ and $`g_0`$ is semisimple; b) $`\mathrm{rk}g=\mathrm{rk}g_0`$ and $`g_0`$ has 1-dimensional centre; c) $`\mathrm{rk}g>\mathrm{rk}g_0`$. In cases a) and b), $`\mathrm{\Theta }`$ is inner and, accordingly, both $`g_0`$ and the $`Z_2`$-grading are said to be of inner type. It is well known that the $`g_0`$-module $`g_1`$ is irreducible in cases a) and c), and is the sum of two dual submodules in case b). However, $`g_1`$ is orthogonal in all three cases and one may consider the $`g_0`$-module $`Spin(g_1)`$. An important feature of this situation is that all nonzero weights of $`g_1`$ are of multiplicity 1. This is clear in the equal rank cases, and can also be proved for c). An invariant theoretic proof of this uses Lemma 3.4 in \[Ka80\] and the fact that the linear group $`G_0GL(g_1)`$ is visible. 5 $`\mathrm{𝐒𝐩𝐢𝐧}(g_1)`$ for the inner involutory automorphisms In this section, $`g`$ is simple and $`g_0`$ is a symmetric subalgebra of inner type. Retain for $`g`$ the previous notation such as $`t`$, $`\mathrm{\Delta }`$, $`\mathrm{\Delta }^+`$, $`\rho `$, $`C`$, etc. Since $`\mathrm{rk}g=\mathrm{rk}g_0`$, we may assume that $`t`$ is a Cartan subalgebra in both $`g`$ and $`g_0`$. Let $`\mathrm{\Delta }_0`$ be the root system of $`(g_0,t)`$ and $`\mathrm{\Delta }_1`$ the set of weights of the $`g_0`$-module $`g_1`$. Then $`\mathrm{\Delta }=\mathrm{\Delta }_0\mathrm{\Delta }_1`$ and $`\mathrm{\Delta }_0`$ is a closed subset of $`\mathrm{\Delta }`$. We regard $`\mathrm{\Delta }_0^+:=\mathrm{\Delta }^+\mathrm{\Delta }_0`$ as set of positive roots for $`g_0`$. Note also that $`\mathrm{\Delta }_1`$ contains a distinguished ‘half’ $`\mathrm{\Delta }_1^+=\mathrm{\Delta }^+\mathrm{\Delta }_1`$. Then Prop. 1 applies to the Weyl groups $`W_0W`$ and one obtains the “minimal length” subset $`W^0W`$. Our aim is to describe the $`g_0`$-module $`Spin_0(g_1)`$. As $`g_1`$ has no zero weight, we have $`Spin(g_1)=Spin_0(g_1)`$. As a first step, we find all extreme weights of $`Spin(g_1)`$. Recall from (1)(S3.Ex16) that each dominant half of $`\mathrm{\Delta }_1`$ determines an extreme weight for $`Spin(g_1)`$. According to that discussion, one has to take the dominant chamber $`C_0`$ for $`g_0`$ and cut it up by the hyperplanes orthogonal to the roots of $`\mathrm{\Delta }_1`$. Clearly, each small chamber is isomorphic to $`C`$. Since there are $`\mathrm{\#}W`$ chambers for $`g`$ and $`\mathrm{\#}W_0`$ chambers for $`g_0`$, we obtain the partition of $`C_0`$ in $`\mathrm{\#}(W/W_0)`$ small chambers. Then any weight inside of a small chamber determines a dominant half of $`\mathrm{\Delta }_1`$ and an extreme weight. In the following proposition we give a formula for these extreme weights. 5.1 Proposition. 1. The set of hyperplanes $`H_\mu `$ ($`\mu \mathrm{\Delta }_1`$) cuts $`C_0`$ in $`\mathrm{\#}(W/W_0)`$ small chambers; 2. The collection of ($`g_0`$-dominant) weights $`w^1\rho `$ ($`wW^0`$) contains representatives of all small chambers in $`C_0`$. 3. The extreme weight of $`Spin(g_1)`$ corresponding to $`w^1\rho `$ is $`\lambda _w:=w^1\rho \rho _0`$. Proof. 1. This is proved in the previous paragraph. 2 & 3. If $`\alpha \mathrm{\Delta }_0^+`$ and $`wW^0`$, then $`w\alpha \mathrm{\Delta }^+`$, see Prop. 1(1). Therefore $`w^1\rho `$ is $`g_0`$-dominant. Since the number of these weights is $`\mathrm{\#}(W/W_0)`$, as required, it suffices to verify that the corresponding dominant halfs are different. By definition, the dominant half of $`\mathrm{\Delta }_1`$ associated with $`w^1\rho `$ is $$(\mathrm{\Delta }_1)_w^+=\{\mu \mathrm{\Delta }_1(w^1\rho ,\mu )>0\}=\{\mu \mathrm{\Delta }_1w\mu \mathrm{\Delta }^+\}.$$ Because all weight multiplicities in $`g_1`$ are equal to 1, the corresponding extreme weight is $`\lambda _w:=\frac{1}{2}|(\mathrm{\Delta }_1)_{w^1\rho }^+|`$. Set $`M_w=\{\mu \mathrm{\Delta }_1^+w\mu \mathrm{\Delta }^+\}`$ and $`\overline{M}_w=\mathrm{\Delta }_1^+M_w`$. Then $`\mathrm{\Delta }^+=\mathrm{\Delta }_0^+M_w\overline{M}_w`$ and $$\rho =\rho _0+\frac{1}{2}|M_w|+\frac{1}{2}|\overline{M}_w|.$$ Since $`wW^0`$, we obtain $$w^1\rho =\rho _0+\frac{1}{2}|M_w|\frac{1}{2}|\overline{M}_w|.$$ Note also that $`|(\mathrm{\Delta }_1)_w^+|=|M_w||\overline{M}_w|`$. Whence $`\lambda _w=w^1\rho \rho _0`$. Thus, we have obtained the required number of different extreme weights. In the remainder of the section, notation $`V_\lambda `$ refers to a $`g_0`$-module. 5.2 Theorem. Let $`g=g_0g_1`$ be a $`Z_2`$-grading of inner type. Then $`Spin_0(g_1)=Spin(g_1)={\displaystyle \underset{wW^0}{}}V_{\lambda _w}`$ . Proof. It follows from the preceding exposition that $$\underset{wW^0}{}V_{\lambda _w}Spin_0(g_1)=Spin(g_1).$$ Since $`Spin(g_1)`$ is self-dual, $`dim(Spin(g_1)^2)^{g_0}`$ is greater than or equal to the number of irreducible summands of $`Spin(g_1)`$. Therefore the desired equality is equivalent to that $`dim(Spin(g_1)^2)^{g_0}=\mathrm{\#}W^0`$. Recall the main property of ‘Spin’ in our situation: $$^{}g_1Spin(g_1)^2.$$ Hence the question about $`g_0`$-invariants is being translated in the setting of exterior algebras. Assuming that $`\text{𝕜}=C`$, we can exploit de Rham cohomology with complex coefficients. It is well known that $`(^{}g_1)^{g_0}`$ is isomorphic to $`H^{}(G/G_0)`$, the cohomology ring of the symmetric space $`G/G_0`$ \[On95, §9 n.11\], and that $`dimH^{}(G/G_0)=\mathrm{\#}(W/W_0)`$ \[On95, §13 n.3\]. This completes the proof. 5.3 Example. Let $`\mathrm{\Theta }`$ be a ‘Hermitian’ involutory automorphism, i.e., $`g_0`$ has a 1-dimensional centre and $`g_1WW^{}`$, where $`W`$ is a faithful irreducible $`g_0`$-module. This is just case 1(b). Then $`\mathrm{\Delta }(W)=𝒫(W)=\mathrm{\Delta }_1^+`$ and, according to Example 1(3), $`Spin(g_1)\text{𝕜}_{\rho _1}^{}W^{}`$, where $`\rho _1=\frac{1}{2}|\mathrm{\Delta }_1^+|`$. It then follows from Theorem 1 that $$^{}W^{}=\text{𝕜}_{\rho _1}Spin(g_1)=\underset{wW^0}{}V_{w\rho \rho }.$$ Or, equivalently, $`\{\rho w\rho wW^0\}`$ is the set of all highest weights for the $`g_0`$-module $`^{}W`$. This result was obtained by Kostant (see \[Ko61, 8.2\]) as application of his results on the cohomology of the nilpotent radical of a parabolic subalgebra of $`g`$. In this situation, $`W`$ is the Abelian nilpotent radical of the parabolic subalgebra $`g_0W`$. So, the concept of ‘Spin’ and Theorem 1 yield another generalization of Kostant’s result. Purists may condemn the above proof of Theorem 1, since it invokes the cohomology theory of compact Lie groups over $`C`$. Fortunately, there exists also a rather simple and purely algebraic proof. We shall show that the equality in 1 is equivalent to an identity in $`Z[𝒫]`$, which is a variation of the Weyl denominator formula. Recall from section 1 the cunning parity $`\tau (w)`$ for $`wW`$, determined by $`W_0`$. 5.4 Theorem. Let $`\mathrm{\Delta }=\mathrm{\Delta }_0\mathrm{\Delta }_1`$ be the partition corresponding to a $`Z_2`$-grading of $`g`$ of inner type. Then $`{\displaystyle \underset{wW}{}}\tau (w)e^{w\rho }={\displaystyle \underset{\alpha \mathrm{\Delta }_0^+}{}}(e^{\alpha /2}e^{\alpha /2}){\displaystyle \underset{\mu \mathrm{\Delta }_1^+}{}}(e^{\mu /2}+e^{\mu /2})`$ . Proof. The fact that $`\mathrm{\Delta }_0`$ and $`\mathrm{\Delta }_1`$ originate from an inner involutory automorphism can alternatively be stated as follows: $`()`$ if $`\alpha \mathrm{\Delta }_i`$, $`\beta \mathrm{\Delta }_j`$, and $`\alpha +\beta \mathrm{\Delta }`$, then $`\alpha +\beta \mathrm{\Delta }_{i+j}`$, where, of course, $`i,jZ/2Z`$. Let $`𝒬𝒫`$ be the root lattice. For $`Z[𝒬]`$, with basis $`e^\alpha `$ ($`\alpha 𝒬`$), one has a version of Weyl’s denominator identity: $$\underset{wW}{}\epsilon (w)e^{w\rho \rho }=\underset{\alpha \mathrm{\Delta }^+}{}(1e^\alpha ).$$ Consider the second copy of $`Z[𝒬]`$, with basis $`q^\alpha `$ ($`\alpha 𝒬`$), and the equality in $`Z[𝒬]Z[𝒬]`$: $$\underset{wW}{}\epsilon (w)q^{w\rho \rho }e^{w\rho \rho }=\underset{\alpha \mathrm{\Delta }^+}{}(1q^\alpha e^\alpha ).$$ (5.5) Take the specialization of this identity such that $`q^\alpha \{\begin{array}{cc}\hfill 1,& \alpha \mathrm{\Delta }_0\\ \hfill 1,& \alpha \mathrm{\Delta }_1\end{array}`$. It has to be verified that one obtains a well-defined homomorphism $`(𝒬,+)\{1,1\}Z/2Z`$. In other words, if $`\nu =_{iI}\mu _i`$ is a sum of roots then the number of summands lying in $`\mathrm{\Delta }_1`$ should have the same parity for all such presentations. Indeed, assume that $`_{iI}\mu _i=_{jJ}\beta _j`$. We argue by induction on $`\mathrm{\#}I+\mathrm{\#}J`$. Since $`(_{iI}\mu _i,_{jJ}\beta _j)>0`$, there exist $`i_0,j_0`$ such that $`(\mu _{i_0},\beta _{j_0})>0`$. Hence $`\mu _{i_0}\beta _{j_0}`$ is a root and $`_{iI\{i_0\}}\mu _i+(\mu _{i_0}\beta _{j_0})=_{jJ\{j_0\}}\beta _j`$. We conclude by applying the inductive hypothesis to this equality and using $`()`$. Thus, the specialization is well-defined and we obtain $`{\displaystyle \underset{\alpha \mathrm{\Delta }_0^+}{}}(1e^\alpha ){\displaystyle \underset{\mu \mathrm{\Delta }_1^+}{}}(1+e^\mu )`$ at the right hand side of Eq. (5.5). It is easily seen that $`w\rho \rho =|\mathrm{\Delta }(w)|`$, where $`\mathrm{\Delta }(w)=\{\alpha \mathrm{\Delta }^+w^1\alpha \mathrm{\Delta }^{}\}=\mathrm{\Delta }^+w(\mathrm{\Delta }^{})`$. Therefore $`q^{w\rho \rho }`$ specializes to $`(1)^n`$, where $`n=\mathrm{\#}(\mathrm{\Delta }_1^+w\mathrm{\Delta }^{})`$. Recall that $`\epsilon (w)=(1)^{l(w)}`$ and $`l(w)=\mathrm{\#}(\mathrm{\Delta }^+w\mathrm{\Delta }^{})`$. Thus the resulting sign on the left hand side is $`(1)^{\mathrm{\#}(\mathrm{\Delta }_0w\mathrm{\Delta }^{})}`$, which is just $`\tau (w)`$ by Lemma 1. This completes the proof of the theorem. (5.6) Another proof of theorem 1. By Weyl’s character formula for $`g_0`$-modules and Prop. 1(3), $$\mathrm{ch}V_{\lambda _w}=\frac{_{\stackrel{~}{w}W_0}\epsilon _0(\stackrel{~}{w})e^{\stackrel{~}{w}(\rho _0+\lambda _w)}}{_{\alpha \mathrm{\Delta }_0^+}(e^{\alpha /2}e^{\alpha /2})}=\frac{_{\stackrel{~}{w}W_0}\epsilon _0(\stackrel{~}{w})e^{\stackrel{~}{w}w^1\rho }}{_{\alpha \mathrm{\Delta }_0^+}(e^{\alpha /2}e^{\alpha /2})}.$$ Hence $$\mathrm{ch}\left(\underset{wW^0}{}V_{\lambda _w}\right)=\frac{_{wW^0}_{\stackrel{~}{w}W_0}\epsilon _0(\stackrel{~}{w})e^{\stackrel{~}{w}w^1\rho }}{_{\alpha \mathrm{\Delta }_0^+}(e^{\alpha /2}e^{\alpha /2})}.$$ By the very definition of $`\tau (w)`$ (see section 1) and Prop. 1(2), it follows that the numerator is equal to $`_{wW}\tau (w)e^{w\rho }`$. Whence, by Theorem 1, $$\mathrm{ch}\left(\underset{wW^0}{}V_{\lambda _w}\right)=\underset{\mu \mathrm{\Delta }_1^+}{}(e^{\mu /2}+e^{\mu /2})=\mathrm{ch}Spin(g_1).$$ 5.7 Examples.1. $`g=so_{2n+1}`$, $`g_0=so_{2n}`$. Here $`g_1V_{\phi _1}`$ is the tautological $`so_{2n}`$-module and $`\mathrm{\#}(W/W_0)=2`$. Let $`\{\epsilon _1,\mathrm{},\epsilon _n\}`$ be the standard basis of $`t^{}`$ so that $`\mathrm{\Delta }=\{\pm \epsilon _i\pm \epsilon _j,\pm \epsilon _i1i,jn,ij\}`$. Here $`\mathrm{\Delta }_0^+=\{\epsilon _i\pm \epsilon _j(i<j),\epsilon _i\}`$ and $`\mathrm{\Delta }_0=\mathrm{\Delta }_l`$. Then $`W^0=\{id,w_n\}`$, where $`w_n(\epsilon _i)=\epsilon _i`$ ($`in1`$) and $`w_n(\epsilon _n)=\epsilon _n`$. Since $`\mathrm{\Delta }_1=\mathrm{\Delta }(V_{\phi _1})=\{\pm \epsilon _1,\mathrm{},\pm \epsilon _n\}`$, the corresponding dominant halfs are $`(\mathrm{\Delta }_1)_{id}^+=\{\epsilon _1,\mathrm{},\epsilon _n\}`$ and $`(\mathrm{\Delta }_1)_{w_n}^+=\{\epsilon _1,\mathrm{},\epsilon _{n1},\epsilon _n\}`$, and the corresponding extreme weights are $`\phi _n`$ and $`\phi _{n1}`$. Thus, $`Spin(V_{\phi _1})=V_{\phi _{n1}}V_{\phi _n}`$ and $`^{}V_{\phi _1}=(V_{\phi _{n1}}V_{\phi _n})^2`$. Notice that the last equality is nothing but the first equality in Eq. (2.1). 2. $`g=f_4`$, $`g_0=so_9`$. Here $`g_1V_{\phi _4}`$ and $`\mathrm{\#}(W/W_0)=3`$. In the standard notation for $`f_4`$, we have $`\mathrm{\Delta }^+=\{\epsilon _i\pm \epsilon _j(i<j),\epsilon _i,\frac{1}{2}(\epsilon _1\pm \epsilon _2\pm \epsilon _3\pm \epsilon _4)\}`$. Then $`\mathrm{\Delta }_0^+=\mathrm{\Delta }_l^+\{\epsilon _1,\epsilon _2,\epsilon _3,\epsilon _4\}`$. An explicit computation shows that $`W^0=\{id,w^{},w^{\prime \prime }\}`$, where $`w^{}:\{\begin{array}{c}\epsilon _1\frac{1}{2}(\epsilon _1+\epsilon _2+\epsilon _3+\epsilon _4)\\ \epsilon _2\frac{1}{2}(\epsilon _1+\epsilon _2\epsilon _3\epsilon _4)\\ \epsilon _3\frac{1}{2}(\epsilon _1\epsilon _2+\epsilon _3\epsilon _4)\\ \epsilon _4\frac{1}{2}(\epsilon _1\epsilon _2\epsilon _3+\epsilon _4)\end{array}`$ and $`w^{\prime \prime }:\{\begin{array}{c}\epsilon _1\frac{1}{2}(\epsilon _1+\epsilon _2+\epsilon _3\epsilon _4)\\ \epsilon _2\frac{1}{2}(\epsilon _1+\epsilon _2\epsilon _3+\epsilon _4)\\ \epsilon _3\frac{1}{2}(\epsilon _1\epsilon _2+\epsilon _3+\epsilon _4)\\ \epsilon _4\frac{1}{2}(\epsilon _1\epsilon _2\epsilon _3\epsilon _4)\end{array}`$. (One may notice that any $`wW^0`$ must preserve $`(\mathrm{\Delta }_0)_l^+=\mathrm{\Delta }_l^+`$, $`\mathrm{\Delta }_l`$ being the root system of type $`𝐃_4`$. Hence $`w`$ takes $`\epsilon _2\epsilon _3`$ to itself and permutes somehow $`\epsilon _1\epsilon _2,\epsilon _3\epsilon _4`$, and $`\epsilon _3+\epsilon _4`$.) Whence $`\lambda _{id}=\rho \rho _0=2\epsilon _1`$, $`\lambda _w^{}=(w^{})^1\rho \rho _0=\epsilon _1+\epsilon _2+\epsilon _3`$, $`\lambda _{w^{\prime \prime }}=(w^{\prime \prime })^1\rho \rho _0=(3\epsilon _1+\epsilon _2+\epsilon _3+\epsilon _4)/2`$. Thus $`Spin(V_{\phi _4})=V_{2\phi _1}V_{\phi _3}V_{\phi _1+\phi _4}`$. 6 $`\mathrm{𝐒𝐩𝐢𝐧}(g_1)`$ for the outer involutory automorphisms In this section $`g`$ is simple and $`\mathrm{\Theta }`$ is outer. Since $`\mathrm{rk}g_0<\mathrm{rk}g`$, there is no clear relation between roots and Weyl groups of the two algebras, and the approach of section 1 seems to fail completely. Yet, it appears to be possible to describe $`Spin(g_1)`$ in a similar fashion, but with some complications. Another price is that we have to exploit case-by-case arguments several times. (6.1) Associated diagram involutory automorphism of $`g`$. By a result of Steinberg, $`\mathrm{\Theta }`$ keeps stable a Borel subalgebra and a Cartan subalgebra in it. Therefore we may (and shall) assume that $`\mathrm{\Theta }t=t`$ and $`\mathrm{\Theta }u^+=u^+`$. Then $`\mathrm{\Theta }`$ also preserves $`\mathrm{\Delta }^+`$ and $`\mathrm{\Pi }`$, as subsets of $`t^{}`$. In particular, $`\mathrm{\Theta }`$ induces an involution of the Dynkin diagram. Associated with this involution, one has the specific involutory automorphism of $`g`$, which is called the diagram involutory automorphism and denoted by $`\overline{\mathrm{\Theta }}`$. Roughly speaking, $`\overline{\mathrm{\Theta }}`$ performs the same involution on $`\mathrm{\Pi }`$, as $`\mathrm{\Theta }`$, and transforms ‘well’ the Chevalley generators of $`g`$<sup>4</sup><sup>4</sup>4Explicit formulas for the diagram automorphisms of all simple Lie algebras are written in \[Ka90, § 7.9,7.10\].. We are going to compare properties of the $`Z_2`$-gradings $$g=g_0g_1\text{and}g=g_{\overline{0}}g_{\overline{1}}$$ arising from $`\mathrm{\Theta }`$ and $`\overline{\mathrm{\Theta }}`$. By construction, $`\mathrm{\Theta }|_t=\overline{\mathrm{\Theta }}|_t`$. Therefore $`\mathrm{\Theta }`$ and $`\overline{\mathrm{\Theta }}`$ act identically on $`\mathrm{\Delta }^+`$ and $`t_0:=t^\mathrm{\Theta }`$ is a Cartan subalgebra for both $`g_0`$ and $`g_{\overline{0}}`$. Let us organize notation for roots and weights of the symmetric subalgebras in question: $``$ $`\mathrm{\Delta }_0`$ (resp. $`\mathrm{\Delta }_{\overline{0}}`$) is the root system of $`g_0`$ (resp. $`g_{\overline{0}}`$) relative to $`t_0`$; $``$ $`\mathrm{\Delta }_1`$ (resp. $`\mathrm{\Delta }_{\overline{1}}`$) is the set of non-zero weights of the $`g_0`$-module $`g_1`$ (resp. $`g_{\overline{0}}`$-module $`g_{\overline{1}}`$) relative to $`t_0`$. Since all these sets are defined with respect to a common Cartan subalgebra, $`\mathrm{\Delta }_0\mathrm{\Delta }_1=\mathrm{\Delta }_{\overline{0}}\mathrm{\Delta }_{\overline{1}}`$ and, more precisely, the totality of weights occurring in $`\{\mathrm{\Delta }_0,\mathrm{\Delta }_1\}`$ is the same as in $`\{\mathrm{\Delta }_{\overline{0}},\mathrm{\Delta }_{\overline{1}}\}`$. Because $`t_0`$ contains regular elements of $`g`$ (see e.g. \[Ka90, 8.1(b)\]), none of the roots of $`g`$ vanishes on $`t_0`$. Therefore the above totality of weights consists of all restricted roots. Moreover, since the non-zero weights of $`t_0`$ in $`g_1`$ (or $`g_{\overline{1}}`$) are of multiplicity 1, $$\mathrm{\#}\mathrm{\Delta }=\mathrm{\#}\mathrm{\Delta }_0+\mathrm{\#}\mathrm{\Delta }_1=\mathrm{\#}\mathrm{\Delta }_{\overline{0}}+\mathrm{\#}\mathrm{\Delta }_{\overline{1}}.$$ Warning. Unlike section 1, elements of $`\mathrm{\Delta }_0`$ and $`\mathrm{\Delta }_1`$ have not much in common with roots of $`g`$. Actually, we do not need $`\mathrm{\Delta }`$ in this section. The next assertion follows from the classification. Fact. The fixed point subalgebra of an outer involutory automorphism always has roots of different length, with $`long^2/short^2=2`$. This applies to both $`g_0`$ and $`g_{\overline{0}}`$ and, as in section 1, we use the subscripts ‘s’ and ‘l’ to denote the objects related to short and long roots in $`\mathrm{\Delta }_0`$ and $`\mathrm{\Delta }_{\overline{0}}`$. A close look to the classification list reveals important features of this situation. (6.2) Suppose $`\mathrm{\Theta }\overline{\mathrm{\Theta }}`$. Then $`\{\begin{array}{c}g_{\overline{0}}\text{ is simple and }g_{\overline{1}}\text{ is the little adjoint module for }g_{\overline{0}}\text{;}\hfill \\ \mathrm{\Delta }_{\overline{1}}=(\mathrm{\Delta }_{\overline{0}})_s\text{;}\hfill \\ \mathrm{\Delta }_0\mathrm{\Delta }_{\overline{0}}\text{ and }(\mathrm{\Delta }_0)_s=(\mathrm{\Delta }_{\overline{0}})_s\text{}\hfill \end{array}`$ In fact, there are 7 series of outer involutory automorphisms of simple Lie algebras. They form three pairs $`(\mathrm{\Theta },\overline{\mathrm{\Theta }})`$ and one “isolated” diagram involutory automorphism, where (S6.Ex33) is not satisfied. The relevant data for all these series are presented in Table 2. It follows from (S6.Ex33) that $`\mathrm{\Delta }_{\overline{0}}\mathrm{\Delta }_{\overline{1}}=\mathrm{\Delta }_{\overline{0}}`$. Thus, everything lies in $`\mathrm{\Delta }_{\overline{0}}`$. Therefore a choice of the set of positive roots $`\mathrm{\Delta }_{\overline{0}}^+`$ determines $`\mathrm{\Delta }_{\overline{1}}^+`$, $`\mathrm{\Delta }_0^+`$, and $`\mathrm{\Delta }_1^+`$ as well. Of course, we choose $`\mathrm{\Delta }_{\overline{0}}^+`$ so that it is the image of $`\mathrm{\Delta }^+`$ under the projection $`t^{}(t_0)^{}`$. Then $`\{\mathrm{\Delta }_0^+,\mathrm{\Delta }_1^+\}`$ and $`\{\mathrm{\Delta }_{\overline{0}}^+,\mathrm{\Delta }_{\overline{1}}^+\}`$ are two presentations for the totality of all restricted positive roots of $`g`$. Set $`\rho _i=\frac{1}{2}|\mathrm{\Delta }_i^+|`$ and $`\rho _{\overline{ı}}=\frac{1}{2}|\mathrm{\Delta }_{\overline{ı}}^+|`$ ($`i=0,1`$). Recall that $`\rho =\frac{1}{2}|\mathrm{\Delta }^+|`$ and therefore $`\mathrm{\Theta }\rho =\rho `$. That is, $`\rho (t^{})^\mathrm{\Theta }t_0^{}`$. It then follows from the above discussion that $$\rho =\rho _0+\rho _1=\rho _{\overline{0}}+\rho _{\overline{1}}=\rho _{\overline{0}}+(\rho _{\overline{0}})_s.$$ (6.3) Let $`W_0`$ and $`W_{\overline{0}}`$ be the Weyl groups of $`g_0`$ and $`g_{\overline{0}}`$, respectively. Although $`\mathrm{\Delta }_0\mathrm{\Delta }_{\overline{0}}`$, $`g_0`$ is not a subalgebra of $`g_{\overline{0}}`$ (if $`\mathrm{\Theta }\overline{\mathrm{\Theta }}`$). In other words, $`\mathrm{\Delta }_0`$ is a non-closed subset of $`\mathrm{\Delta }_{\overline{0}}`$. Nevertheless, Prop. 1 applies to $`W_0W_{\overline{0}}`$ and one obtains the subset $`W^{}W_{\overline{0}}`$ consisting of the elements of minimal length in the cosets $`\{wW_0\}`$. Equivalently, $`W^{}=\{wW_{\overline{0}}w(\mathrm{\Delta }_0^+)\mathrm{\Delta }_{\overline{0}}^+\}`$. Below, we consider the Weyl chambers $`C_0`$ and $`C_{\overline{0}}`$, and the hyperplanes $`H_\mu `$ ($`\mu \mathrm{\Delta }_1`$). They are regarded as subsets of the rational span of $`\mathrm{\Delta }_{\overline{0}}`$ in $`t_0^{}`$. 6.4 Proposition. 1. The set of hyperplanes $`H_\mu `$ ($`\mu \mathrm{\Delta }_1`$) cuts $`C_0`$ in $`\mathrm{\#}(W_{\overline{0}}/W_0)`$ small chambers; 2. The collection of ($`g_0`$-dominant) weights $`w^1\rho _{\overline{0}}`$ ($`wW^{}`$) contains representatives of all small chambers in $`C_0`$. 3. The extreme weight of $`Spin_0(g_1)`$ corresponding to $`w^1\rho _{\overline{0}}`$ is $`\lambda _w:=w^1(\rho _0+\rho _1)\rho _0=w^1\rho \rho _0`$. Proof. To a great extent, the proof is parallel to the proof of Prop. 1. 1. The union $`\mathrm{\Delta }_0\mathrm{\Delta }_1`$ coincides with $`\mathrm{\Delta }_{\overline{0}}`$. Therefore each small chamber is isomorphic to $`C_{\overline{0}}`$. Comparing the total number of chambers, we see that $`C_0`$ splits into $`\mathrm{\#}(W_{\overline{0}}/W_0)`$ small chambers. 2 & 3. By the definition of $`W^{}`$, it follows that $`w^1\rho _{\overline{0}}`$ is $`g_0`$-dominant. So, we have the required number of dominant weights and it suffices to verify that the corresponding extreme weights of $`Spin_0(g_1)`$ are different. Given $`wW^{}`$, the dominant half of $`\mathrm{\Delta }_1`$ associated with $`w^1\rho _{\overline{0}}`$ is $$(\mathrm{\Delta }_1)_w^+:=\{\mu \mathrm{\Delta }_1(w^1\rho _{\overline{0}},\mu )>0\}=\{\mu \mathrm{\Delta }_1w\mu \mathrm{\Delta }_{\overline{0}}^+\}$$ and the corresponding extreme weight is $`\lambda _w:=\frac{1}{2}|(\mathrm{\Delta }_1)_w^+|`$. Set $`M_w=\{\mu \mathrm{\Delta }_1^+w\mu \mathrm{\Delta }_{\overline{0}}^+\}`$ and $`\overline{M}_w=\mathrm{\Delta }_1^+M_w`$. Then $$\rho _{\overline{0}}+\rho _{\overline{1}}=\rho _0+\rho _1=\rho _0+\frac{1}{2}|M_w|+\frac{1}{2}|\overline{M}_w|.$$ Since $`wW^{}`$, we have $$w^1(\rho _{\overline{0}}+\rho _{\overline{1}})=\rho _0+\frac{1}{2}|M_w|\frac{1}{2}|\overline{M}_w|.$$ Noting that $`|(\mathrm{\Delta }_1)_w^+|=|M_w||\overline{M}_w|`$, we obtain $`\lambda _w=w^1(\rho _{\overline{0}}+\rho _{\overline{1}})\rho _0`$. Obviously, these weights are different, and we are done. In the next theorem, $`V_\lambda `$ denotes a $`g_0`$-module. 6.5 Theorem. Let $`g=g_0g_1`$ be a $`Z_2`$-grading of outer type and $`g=g_{\overline{0}}g_{\overline{1}}`$ the associated diagram $`Z_2`$-grading. Let $`W^{}`$ be the set of representatives of minimal length for $`W_{\overline{0}}/W_0`$. Then $`Spin_0(g_1)=_{wW^{}}V_{\lambda _w}`$ . Proof. First, note that if $`\mathrm{\Theta }`$ is a diagram involutory automorphism, then $`W_{\overline{0}}=W_0`$. Here the theorem claims that $`Spin_0(g_1)`$ is irreducible, with highest weight $`(\rho _{\overline{0}})_s`$. This was already demonstrated in Theorem 1 and Prop. 1. In the general case, we proceed as follows. By Prop. 1, $`_{wW^{}}V_{\lambda _w}Spin_0(g_1)`$, and the equality will follow from the fact that $`dim(Spin_0(g_1)^2)^{g_0}=\mathrm{\#}W^{}`$. As in the proof of Theorem 1, a crucial step in the next argument is of “cohomological” nature. Since $`t_0`$ contains regular elements, $`dim(g_1)^{t_0}=dimtdimt_0`$, i.e., the multiplicity of the zero weight in $`g_1`$ is equal to $`\mathrm{rk}g\mathrm{rk}g_0`$. By Prop. 1(ii), $$^{}g_12^{\text{rk}g\text{rk}g_0}Spin_0(g_1)^2$$ and hence $$dim(^{}g_1)^{g_0}=2^{\text{rk}g\text{rk}g_0}dim(Spin_0(g_1)^2)^{g_0}.$$ At the rest of the proof, $`\text{𝕜}=C`$. Inspecting the list of the symmetric spaces of outer type and their cohomology rings over $`C`$ (see e.g. \[Ta62, § 4\]) yields the equality $$dimH^{}(G/G_0)=2^{\text{rk}g\text{rk}g_0}\mathrm{\#}(W_{\overline{0}}/W_0).$$ Since $`(^{}g_1)^{g_0}H^{}(G/G_0)`$, we are done. The following proof, although also being not free of case-by-case arguments, does not appeal to $`C`$. (6.6) Another proof of Theorem 1. Arguing as in (5.5) and using Prop. 1(3), we obtain $$\mathrm{ch}\left(\underset{wW^{}}{}V_{\lambda _w}\right)=\underset{wW^{}}{}\frac{_{\stackrel{~}{w}W_0}\epsilon _0(\stackrel{~}{w})e^{\stackrel{~}{w}(\rho _0+\lambda _w)}}{_{\alpha \mathrm{\Delta }_0^+}(e^{\alpha /2}e^{\alpha /2})}=$$ (6.7) $$=\frac{_{wW^{}}_{\stackrel{~}{w}W_0}\epsilon _0(\stackrel{~}{w})e^{\stackrel{~}{w}w^1\rho }}{_{\alpha \mathrm{\Delta }_0^+}(e^{\alpha /2}e^{\alpha /2})}=\frac{_{wW_{\overline{0}}}\tau (w)e^{w\rho }}{_{\alpha \mathrm{\Delta }_0^+}(e^{\alpha /2}e^{\alpha /2})}.$$ Here $`\tau (w)`$ is the cunning parity for $`wW_{\overline{0}}`$, relative to the subgroup $`W_0`$. To get another expression for the numerator, we exploit the following observation concerning the pairs $`(g_0,g_{\overline{0}})`$ in Table 2. Although $`\mathrm{\Delta }_0`$ is not closed in $`\mathrm{\Delta }_{\overline{0}}`$, the dual root system $`\stackrel{~}{\mathrm{\Delta }}_0`$ is closed in $`\stackrel{~}{\mathrm{\Delta }}_{\overline{0}}`$ and, moreover, it is a “symmetric” subset. That is, the partition $`\stackrel{~}{\mathrm{\Delta }}_{\overline{0}}=\stackrel{~}{\mathrm{\Delta }}_0(\stackrel{~}{\mathrm{\Delta }}_{\overline{0}}\stackrel{~}{\mathrm{\Delta }}_0)`$ arises from an inner involutory automorphism of the “dual” Lie algebra. (E.g. the pair $`(𝐂_4,𝐅_4)`$ inverts in $`(𝐁_4,𝐅_4`$).) Here $`\stackrel{~}{\mathrm{\Delta }}_{\overline{0}}=(\mathrm{\Delta }_{\overline{0}})_l2(\mathrm{\Delta }_{\overline{0}})_s`$. Recall from (S6.Ex33) that $`\mathrm{\Delta }_{\overline{1}}=(\mathrm{\Delta }_{\overline{0}})_s=(\mathrm{\Delta }_0)_s`$. This means in particular that $`\mathrm{\Delta }_{\overline{0}}\mathrm{\Delta }_0`$ consists of long roots and these are exactly the roots constituting $`(\mathrm{\Delta }_1)_l`$. Hence $`\stackrel{~}{\mathrm{\Delta }}_{\overline{0}}\stackrel{~}{\mathrm{\Delta }}_0=(\mathrm{\Delta }_1)_l`$. After these preparations, write out the identity from Theorem 1 for the partition $`\stackrel{~}{\mathrm{\Delta }}_{\overline{0}}=\stackrel{~}{\mathrm{\Delta }}_0(\mathrm{\Delta }_1)_l`$: $$\underset{wW_{\overline{0}}}{}\tau (w)e^{w\stackrel{~}{\rho }}=\underset{\alpha \stackrel{~}{\mathrm{\Delta }}_0^+}{}(e^{\alpha /2}e^{\alpha /2})\underset{\mu (\mathrm{\Delta }_1^+)_l}{}(e^{\mu /2}+e^{\mu /2}).$$ (6.8) Here $`\stackrel{~}{\rho }:=\frac{1}{2}|\stackrel{~}{\mathrm{\Delta }}_{\overline{0}}^+|=\frac{1}{2}|(\mathrm{\Delta }_{\overline{0}}^+)_l|+|(\mathrm{\Delta }_{\overline{0}}^+)_s|=(\rho _{\overline{0}})_l+2(\rho _{\overline{0}})_s=\rho _{\overline{0}}+(\rho _{\overline{0}})_s=\rho `$ (see Eq. 6.3). Transforming the first factor on the right hand side of Eq. (6.8) yields $$\underset{\alpha \stackrel{~}{\mathrm{\Delta }}_0^+}{}(e^{\alpha /2}e^{\alpha /2})=\underset{\alpha (\mathrm{\Delta }_0^+)_s}{}(e^\alpha e^\alpha )\underset{\beta (\mathrm{\Delta }_0^+)_l}{}(e^{\beta /2}e^{\beta /2})=$$ $$=\underset{\alpha (\mathrm{\Delta }_0^+)_s}{}(e^{\alpha /2}e^{\alpha /2})(e^{\alpha /2}+e^{\alpha /2})\underset{\beta (\mathrm{\Delta }_0^+)_l}{}(e^{\beta /2}e^{\beta /2})=$$ $$=\underset{\alpha \mathrm{\Delta }_0^+}{}(e^{\alpha /2}e^{\alpha /2})\underset{\mu (\mathrm{\Delta }_0^+)_s}{}(e^{\mu /2}+e^{\mu /2}).$$ Since $`(\mathrm{\Delta }_0^+)_s=(\mathrm{\Delta }_1^+)_s`$, the whole expression on the right hand side of Eq. (6.8) is equal to $$\underset{\alpha \mathrm{\Delta }_0^+}{}(e^{\alpha /2}e^{\alpha /2})\underset{\alpha \mathrm{\Delta }_1^+}{}(e^{\mu /2}+e^{\mu /2}).$$ Substituting this expression for $`_{wW_{\overline{0}}}\tau (w)e^{w\rho }`$ in Eq. (6.7), we obtain $$\mathrm{ch}\left(\underset{wW^{}}{}V_{\lambda _w}\right)=\underset{\alpha \mathrm{\Delta }_1^+}{}(e^{\mu /2}+e^{\mu /2})=\mathrm{ch}Spin(g_1).\text{q.e.d.}$$ (6.9) Connexion with cohomology of symmetric spaces. Using classical structure results on $`H^{}(G/G_0)`$ (see e.g. \[On95, ch.3\]), one may notice some interesting coincidences. 1. Recall that the image of the canonical map $`\eta :H^{}(G/G_0)H^{}(G)`$ is generated by primitive elements. Actually, there exists a subalgebra of $`H^{}(G/G_0)`$ that is mapped isomorphically onto $`\mathrm{Im}(\eta )`$. It is called Samelson and denoted by $`Sam(G/G_0)`$. There exists also another subalgebra of $`H^{}(G/G_0)`$, which is called characteristic and denoted by $`{}_{0}{}^{}H_{}^{}(G/G_0)`$. It is the zero component for a natural grading in $`H^{}(G/G_0)`$. Then $`H^{}(G/G_0)Sam(G/G_0)_0H^{}(G/G_0)`$. This holds not only for the symmetric spaces but for a wider class of formal homogeneous spaces, see \[loc. cit., § 12, Th.2\]. A general fact for the formal homogeneous spaces is that $`dimSam(G/G_0)=2^{\text{rk}g\text{rk}g_0}`$. Hence $`dimSpin_0(g_1)=dim_0H^{}(G/G_0)`$, which suggests that $`Spin_0(g_1)`$ might somehow be related to the characteristic subalgebra of $`H^{}(G/G_0)`$. 2. Note that $`H^{}(G/G_0)=_0H^{}(G/G_0)`$ if and only if $`\mathrm{\Theta }`$ is inner, and $`H^{}(G/G_0)=Sam(G/G_0)`$ if and only if $`\mathrm{\Theta }`$ is a diagram involutory automorphism. In the mixed case, the associated diagram involutory automorphism $`\overline{\mathrm{\Theta }}`$ seems to yield a splitting for $`H^{}(G/G_0)`$. Namely, one has $`dimSam(G/G_0)=dimH^{}(G/G_{\overline{0}})`$ and $`dim_0H^{}(G/G_0)=dimH^{}(\stackrel{~}{G}_{\overline{0}}/\stackrel{~}{G}_0)`$, where $`\stackrel{~}{G}_{\overline{0}}`$ and $`\stackrel{~}{G}_0`$ are the groups corresponding to the dual root systems $`\stackrel{~}{\mathrm{\Delta }}_{\overline{0}}`$ and $`\stackrel{~}{\mathrm{\Delta }}_0`$. Moreover, these two pairs of graded algebras have equal Poincaré polynomials and it is likely that they are naturally isomorphic. I think that a better understanding of this situation as well as elimination of the case-by-case arguments in section 1 can be achieved through application of the theory of twisted affine Kac–Moody algebras. 3. In the above exposition, $`\overline{\mathrm{\Theta }}`$ has appeared as deus ex machina. But the Kac-Moody theory provides some explanation for this. Namely, the outer automorphism $`\mathrm{\Theta }`$ determines the twisted affine Kac-Moody algebra $`\widehat{}(g,\mathrm{\Theta },2)=\widehat{}(g)`$ \[Ka90, ch. 8\]. Next, $`\widehat{}(g)`$ has the standard $`Z`$-grading associated with the special vertex of the Dynkin diagram of $`\widehat{}(g)`$. If $`gso_{2n+1}`$, this $`Z`$-grading determines another outer automorphism of $`g`$, which is just $`\overline{\mathrm{\Theta }}`$. 6.10 Examples.1. $`g=sl_{2n}`$, $`g_0=so_{2n}`$. Then $`g_1V_{2\phi _1}`$. As indicated in Table 2, $`g_{\overline{0}}=sp_{2n}`$ and therefore $`\mathrm{\#}(W_{\overline{0}}/W_0)=2`$. Here $`\mathrm{\Delta }_0=\{\pm \epsilon _i\pm \epsilon _j1i,jn,ij\}`$ and $`\mathrm{\Delta }_{\overline{0}}=\{\pm \epsilon _i\pm \epsilon _j,\pm 2\epsilon _i\}`$. This example is a kind of outer version of Example S5.Ex31(1). Indeed, taking the “dual” Lie algebras for $`(g_0,g_{\overline{0}})`$ yields the symmetric pair considered there. In our case, $`\mathrm{\Delta }_1=\mathrm{\Delta }_{\overline{0}}`$ and $`W^{}=\{id,w_n\}`$, where $`w_n(\epsilon _i)=\epsilon _i`$ ($`in1`$) and $`w_n(\epsilon _n)=\epsilon _n`$. Therefore $`(\mathrm{\Delta }_1)_{id}^+=\mathrm{\Delta }_{\overline{0}}^+=\mathrm{\Delta }_0^+\{2\epsilon _1,\mathrm{},2\epsilon _n\}`$ and $`(\mathrm{\Delta }_1)_{w_n}^+=\mathrm{\Delta }_0^+\{2\epsilon _1,\mathrm{},2\epsilon _{n1},2\epsilon _n\}`$. Hence the extreme weights are $`\rho _0+2\phi _n`$ and $`\rho _0+2\phi _{n1}`$. Thus, $`Spin_0(V_{2\phi _1})=V_{\rho _0+2\phi _{n1}}V_{\rho +2\phi _n}`$ and, because $`m_{2\phi _1}(0)=n1`$, $`^{}V_{2\phi _1}=2^{n1}(V_{\rho +2\phi _{n1}}V_{\rho +2\phi _n})^2`$. 2. $`g=e_6`$, $`g_0=sp_8`$. Then $`g_1V_{\phi _4}`$. As indicated in Table 2, $`g_{\overline{0}}=f_4`$ and hence $`\mathrm{\#}(W_{\overline{0}}/W_0)=3`$. This is the outer version of Example S5.Ex31(2). Here $`\mathrm{\Delta }_{\overline{0}}=\{\pm \epsilon _i\pm \epsilon _j,\pm \epsilon _i,(\pm \epsilon _1\pm \epsilon _2\pm \epsilon _3\pm \epsilon _4)/21i,j4,ij\}`$ and $`\mathrm{\Delta }_0=(\mathrm{\Delta }_{\overline{0}})_s\{\pm \epsilon _1\pm \epsilon _2,\pm \epsilon _3\pm \epsilon _4\}`$. The standard set of simple roots for $`\mathrm{\Delta }_{\overline{0}}`$ is $`\alpha _1^{}=\frac{1}{2}(\epsilon _1\epsilon _2\epsilon _3\epsilon _4)`$, $`\alpha _2^{}=\epsilon _4`$, $`\alpha _3^{}=\epsilon _3\epsilon _4`$, $`\alpha _4^{}=\epsilon _2\epsilon _4`$. The roots for $`sp_8`$ have non standard presentation, but it is not hard to find that the simple roots in $`\mathrm{\Delta }_{\overline{0}}^+\mathrm{\Delta }_0`$ are $`\alpha _1=\epsilon _2`$, $`\alpha _2=\frac{1}{2}(\epsilon _1\epsilon _2\epsilon _3\epsilon _4)`$, $`\alpha _3=\epsilon _4`$, $`\alpha _4=\epsilon _3\epsilon _4`$. Therefore the fundamental weights for $`sp_8`$ are $`\phi _1=\frac{1}{2}(\epsilon _1+\epsilon _2)`$, $`\phi _2=\epsilon _1`$, $`\phi _3=\epsilon _1+\frac{1}{2}(\epsilon _3+\epsilon _4)`$, $`\phi _4=\epsilon _1+\epsilon _3`$. Then an explicit verification shows that $`W^{}=\{id,w^{},w^{\prime \prime }\}`$, where $`w^{},w^{\prime \prime }`$ are permutations of $`\{\epsilon _1,\epsilon _2,\epsilon _3,\epsilon _4\}`$ determined by the cycles (23) and (432), respectively. (E.g. $`w^{\prime \prime }(\epsilon _4)=\epsilon _3`$.) The direct computation of the extreme weights gives: $`\lambda _{id}=\rho _{\overline{0}}+\rho _{\overline{1}}\rho _0=\frac{1}{2}(9\epsilon _1+5\epsilon _2+\epsilon _3+\epsilon _4)=5\phi _1+\phi _2+\phi _3`$, $`\lambda _w^{}=(w^{})^1(\rho _{\overline{0}}+\rho _{\overline{1}})\rho _0=\frac{1}{2}(9\epsilon _1+3\epsilon _2+3\epsilon _3+\epsilon _4)=3\phi _1+\phi _2+\phi _3+\phi _4=\rho _0+2\phi _1`$, $`\lambda _{w^{\prime \prime }}=(w^{\prime \prime })^1(\rho _{\overline{0}}+\rho _{\overline{1}})\rho _0=\frac{1}{2}(9\epsilon _1+\epsilon _2+3\epsilon _3+3\epsilon _4)=\phi _1+\phi _2+3\phi _3`$. Whence $`Spin_0(V_{\phi _4})=V_{5\phi _1+\phi _2+\phi _3}V_{\phi _1+\phi _2+3\phi _3}V_{\rho _0+2\phi _1}`$ and $$^{}V_{\phi _4}=4(V_{5\phi _1+\phi _2+\phi _3}V_{\phi _1+\phi _2+3\phi _3}V_{\rho _0+2\phi _1})^2.$$ 7 Decomposably-generated ‘Spin’ modules and a Casimir element Recall that we have given a geometric description of the extreme weights of Spin-representations in (S3.Ex16). Definition. Given an orthogonal $`g`$-module $`V`$, the $`g`$-submodule of $`Spin_0(V)`$ generated by the extreme weight vectors is denoted by $`Spin_0^{dg}(V)`$; $`Spin_0(V)`$ is called decomposably-generated, if it is equal to $`Spin_0^{dg}(V)`$, i.e., if all its highest weights are extreme. Since the extreme weights are of multiplicity 1, “decomposably-generated” implies “multiplicity free”. As a consequence of previous development, we have 7.1 Proposition. Let $`g=g_0g_1`$ be a $`Z_2`$-graded semisimple Lie algebra. Then the $`g_0`$-module $`Spin_0(g_1)`$ is decomposably-generated (and multiplicity free). Proof. The problem immediately reduces to the case in which $`g`$ is an irreducible $`Z_2`$-graded algebra. Then either $`g`$ is simple or $`ghh`$, where $`h`$ is simple and $`\mathrm{\Theta }(h_1,h_2)=(h_2,h_1)`$. In the second case, $`g_0h`$ is the diagonal in $`g`$, and $`g_1g_0`$ as $`g_0`$-module. Here the conclusion follows by Kostant’s result, see Example 1(1). In the first case, for $`g_0`$ of inner type, use Prop. 1(3) and Theorem 1; for $`g_0`$ of outer type, use Prop. 1(3) and Theorem 1. An explanation of the term “decomposably-generated” comes from Example 1(3). If $`V=WW^{}`$, then $`Spin(V)\text{𝕜}_\nu ^{}W`$ and each extreme weight vector is represented by a decomposable vector in the exterior algebra. I think that the property of being “decomposably-generated” characterizes the representations of the form $`Spin_0(g_1)`$, i.e., 7.2 Conjecture. Let $`V`$ be an orthogonal $`g`$-module. Then $`Spin_0(V)`$ is decomposably-generated if and only if $`\stackrel{~}{g}:=gV`$ is a $`Z_2`$-graded semisimple Lie algebra. The conjecture will be proved in a particular case. Until the end of the section, the following situation is being considered: $`g`$ is semisimple, $`h`$ is a reductive subalgebra of $`g`$, and $`m:=h^{}g`$. Then $`m`$ is an orthogonal $`h`$-module and $`g=hm`$ is a vector space sum. Clearly, this decomposition is a $`Z_2`$-grading if and only if $`[m,m]h`$. The representation $`hso(m)`$ is the isotropy representation of the affine homogeneous space $`G/H`$. Our aim is to study $`Spin_0(m)`$ and $`Spin_0^{dg}(m)`$ in the equal rank case. That is, it is assumed from now on that $`\mathrm{rk}h=\mathrm{rk}g`$ and, more precisely, $`th`$. Then $`\mathrm{\Delta }^+=\mathrm{\Delta }_h^+\mathrm{\Delta }_m^+`$, $`m`$ has no zero weight and $`^{}m(Spin_0(m))^2`$. Denoting by $`W_h`$ the Weyl group of $`(h,t)`$, one may consider the minimal length “section” $`W^h`$ for $`WW/W_h`$ and the cunning parity $`\tau :W\{1,1\}`$, determined by $`\mathrm{\Delta }_h^+`$. The proof of Prop. 1 applies in the present situation as well. This yields exactly $`\mathrm{\#}W^h`$ extreme weights of $`Spin_0(m)`$. Hence $$dim(^{}m)^h=dim(Spin_0(m)^2)^h\mathrm{\#}W^h.$$ For $`wW^h`$, the corresponding extreme weight is $`\lambda _w=w^1\rho \rho _h`$, where $`\rho _h=\frac{1}{2}|\mathrm{\Delta }_h^+|`$. Therefore, arguing as in (5.5), we obtain $$\mathrm{ch}(Spin_0^{dg}(m))=\mathrm{ch}\left(\underset{wW^h}{}V_{\lambda _w}\right)=\frac{_{wW}\tau (w)e^{w\rho }}{_{\alpha \mathrm{\Delta }_h^+}(e^{\alpha /2}e^{\alpha /2})}.$$ Recall that $`\mathrm{ch}Spin_0(m)=_{\mu \mathrm{\Delta }_m^+}(e^{\mu /2}+e^{\mu /2})`$. On the other hand, $`dimH^{}(G/H)=\mathrm{\#}W^h`$ \[On95, § 13, Th. 2\] and $`H^{}(G/H)`$ can be computed via the complex of $`G`$-invariant exterior forms on $`G/H`$, i.e., the complex $`((^{}(g/h)^{})^h,d)`$, where $`d`$ is the usual Lie algebra coboundary operator. We shall identify the $`h`$-modules $`(g/h)^{}`$ and $`m`$. Having compared the previous expressions, we obtain 7.3 Proposition. Let $`hg`$ be a reductive subalgebra of maximal rank and $`hso(m)`$ the isotropy representation. Then the following conditions are equivalent: (i) $`d`$ is trivial on $`(^{}m)^h`$; (ii) $`dim(^{}m)^h=\mathrm{\#}W^h`$; (iii) $`Spin_0(m)=Spin_0^{dg}(m)`$; (iv) $`_{wW}\tau (w)e^{w\rho }=_{\alpha \mathrm{\Delta }_h^+}(e^{\alpha /2}e^{\alpha /2})_{\mu \mathrm{\Delta }_m^+}(e^{\mu /2}+e^{\mu /2})`$. Thus, conjecture 1 claims that neither of these conditions holds unless $`G/H`$ is symmetric. 7.4 Theorem. Let $`h`$ be a reductive subalgebra of $`g`$, with $`\mathrm{rk}h=\mathrm{rk}g`$, and $`m:=h^{}`$. Suppose $`[m,m]h`$; then $`d`$ is non-trivial on $`(^{}m)^h`$. More precisely, $`d((^3m)^h)0`$. Proof. For any $`xg`$, let $`x_h`$ and $`x_m`$ denote its components in $`h`$ and $`m`$, respectively. Given $`x,ym`$, consider the decomposition $`[x,y]=[x,y]_h+[x,y]_m`$. We regard $`[,]_m`$ as mapping from $`m\times m`$ to $`m`$, and likewise for $`[,]_h`$. By assumption, $`[,]_m0`$. On the other hand, applying construction from section 1 (see (1.4) and around) to $`h`$ and $`m`$ in place of $`g`$ and $`V`$, we see that $`[x,y]_h=\overline{\mu }(x,y)`$ for any $`x,ym`$. Define the 3-form $`\mathrm{\Psi }:^3m\text{𝕜}`$ by $`\mathrm{\Psi }(x,y,z)=\mathrm{\Phi }([x,y],z)`$. Obviously, $`\mathrm{\Psi }`$ is $`h`$-invariant. The assumption $`[m,m]h`$ precisely means that $`\mathrm{\Psi }0`$. We shall prove $`d\mathrm{\Psi }0`$. To compute $`d\mathrm{\Psi }`$, we regard $`\mathrm{\Psi }`$ as $`h`$-invariant 3-form on $`g`$, orthogonal to $`h`$, and use the standard formula for $`d`$. The resulting expression is $$d\mathrm{\Psi }(x,y,z,u)=2\left(\mathrm{\Phi }([x,y]_m,[z,u]_m)+\mathrm{\Phi }([y,z]_m,[x,u]_m)+\mathrm{\Phi }([z,x]_m,[y,u]_m)\right).$$ Since $`[x,y]=\overline{\mu }(x,y)+[x,y]_m`$, $`h`$ is orthogonal to $`m`$, and $`([x,y],[z,u])+([y,z],[x,u])+([z,x],[y,u])=0`$ (because of the Jacobi identity), we have $$d\mathrm{\Psi }(x,y,z,u)=2\left(\mathrm{\Phi }(\overline{\mu }(x,y),\overline{\mu }(z,u))+\mathrm{\Phi }(\overline{\mu }(y,z),\overline{\mu }(x,u))+\mathrm{\Phi }(\overline{\mu }(z,x),\overline{\mu }(y,u))\right).$$ In the notation of Prop. 1, for $`h`$ and $`m`$ in place of $`g`$ and $`V`$, this means that $`d\mathrm{\Psi }=2\kappa `$. Assume that $`\kappa =0`$. Then $`hm`$ equipped with the modified multiplication $`[,]\stackrel{~}{}`$ becomes a $`Z_2`$-graded Lie algebra (see Prop. 1). Clearly, the multiplication does change only for pairs of elements in $`m`$: $`[m_1,m_2]\stackrel{~}{}:=[m_1,m_2]_h`$, whereas the structure of Lie algebra on $`h`$ and the $`h`$-module structure on $`m`$ remain undisturbed. Let $`\stackrel{~}{g}`$ denote the Lie algebra with modified multiplication and $`\mathrm{\Theta }`$ the corresponding involutory automorphism of $`\stackrel{~}{g}`$. It is easily seen that $`\stackrel{~}{g}`$ is semisimple (use the proof of Theorem 1) and $`\mathrm{\Theta }`$ is inner (because $`th`$ remains a Cartan subalgebra in $`\stackrel{~}{g}`$). For the symmetric space $`\stackrel{~}{G}/H`$, we have $`H^3(\stackrel{~}{G}/H)=(^3m)^h0`$. But $`H^{odd}()=0`$ for the symmetric spaces of inner type \[On95, § 13, n.3\]. This contradiction proves $`\kappa =d\mathrm{\Psi }0`$. 7.5 Corollary. Conjecture 1 is true for the isotropy representations of affine homogeneous spaces $`G/H`$ with $`\mathrm{rk}g=\mathrm{rk}h`$. Remark. Theorem 1 is true even if $`\mathrm{rk}h<\mathrm{rk}g`$ and some mild conditions are satisfied (e.g. $`g`$ is simple). However this has no immediate relation to Conjecture 1. For a reductive Lie algebra $`h`$, the Casimir element in $`U(h)`$ is determined by the choice of an invariant bilinear form on $`h`$. If $`h`$ is not simple, then the choice is essentially non unique. But for the isotropy representations one has a preferred choice of the bilinear form. In the above setting, let $`\mathrm{\Phi }(,)_h`$ be the restriction of $`\mathrm{\Phi }(,)`$ to $`h`$. Notice that even if $`h`$ is semisimple and we begin with the Killing form on $`g`$, then$`\mathrm{\Phi }(,)_h`$ is not necessarily proportional to the Killing form on $`h`$. Let $`c_hU(h)`$ be the Casimir element with respect to $`\mathrm{\Phi }(,)_h`$. Recall that the $`W`$-invariant scalar product on $`𝒫_Q`$ is determined by $`\mathrm{\Phi }(,)`$. 7.6 Proposition. Suppose $`\mathrm{rk}h=\mathrm{rk}g`$. Then the Casimir element $`c_h`$ acts scalarly on $`Spin_0^{dg}(m)`$. Its eigenvalue is equal to $`(\rho ,\rho )(\rho _h,\rho _h)`$. Proof. As is indicated above, $`Spin_0^{dg}(m)=_{wW^h}V_{\lambda _w}`$ and $`\lambda _w=w^1\rho \rho _h`$. Therefore the eigenvalue of $`c_h`$ on $`V_{\lambda _w}`$ is $`(\lambda _w+2\rho _h,\lambda _w)=(w^1\rho ,w^1\rho )(\rho _h,\rho _h)=(\rho ,\rho )(\rho _h,\rho _h)`$. 7.7 Theorem. Let $`g=g_0g_1`$ be a $`Z_2`$-graded semisimple Lie algebra. Define the Casimir element $`c_0`$ for $`g_0`$ using the restriction of $`\mathrm{\Phi }(,)`$ to $`g_0`$. Then $`c_0`$ acts on $`Spin_0(g_1)`$ scalarly, with value $`(\rho ,\rho )(\rho _0,\rho _0)`$. Proof. 1. If $`\mathrm{\Theta }`$ is inner, then $`\mathrm{rk}g_0=\mathrm{rk}g`$; we conclude by Propositions 1, 1. 2. If $`\mathrm{\Theta }`$ is outer, some accuracy is needed, since $`\mathrm{rk}g_0<\mathrm{rk}g`$. We use notation and information from section 1. Since $`Spin_0(g_1)=Spin_0^{dg}(g_1)=_{wW^{}}V_{\lambda _w}`$ and $`\lambda _w=w^1\rho \rho _0`$, the value of $`c_0`$ on $`V_{\lambda _w}`$ is equal to $`(w^1\rho ,w^1\rho )(\rho _0,\rho _0)`$. Here $`W^{}W_{\overline{0}}`$, where $`W_{\overline{0}}`$ is the Weyl group of $`g_{\overline{0}}`$. Recall that $`t_0=t^\mathrm{\Theta }`$ is a Cartan subalgebra for both $`g_{\overline{0}}`$ and $`g_0`$. As $`t_0`$ contains regular elements of $`t`$, we have $`N_{G_{\overline{0}}}(t_0)N_G(t)`$. Furthermore, since $`G_{\overline{0}}`$ is connected and $`t_0`$ is Cartan, we have $`G_{\overline{0}}T=T_0`$. Therefore $`W_{\overline{0}}=N_{G_{\overline{0}}}(t_0)/T_0`$ can be identified with a subgroup of $`W=N_G(t)/T`$. Hence $`(w^1\rho ,w^1\rho )=(\rho ,\rho )`$. ul. Akad. Anokhina d.30, kor.1, kv.7 Moscow 117602 Russia E-mail: dmitri@panyushev.mccme.ru Current address (until February 29, 2000): Max-Planck-Institut für Mathematik Vivatsgasse 7 D-53111 Bonn, Deutschland e-mail: panyush@mpim-bonn.mpg.de
warning/0001/astro-ph0001181.html
ar5iv
text
# On the optical pulsations from the Geminga pulsar ## 1 Introduction An optical counterpart of the Geminga source was found by Bignami et al. (1987), based upon color considerations. Recent spectral studies (Martin et al. 1998) show a continuous power-law $`\left(\nu ^{0.8\pm 0.5}\right)`$ from 3700 to 8000 Å with a broad absorption feature over $`6300÷6500`$ Å band. Such a spectrum indicates existence of the dominant nonthermal synchrotron component in the optical emission of the Geminga pulsar. Recently Shearer et al. (1998), based on their deep integrated images in $`B`$-band, claimed that the optical counterpart of Geminga pulses with a period $`P=237`$ ms, reported also at $`X`$ (Halpern & Holt 1992) and $`\gamma `$-ray (Bertch et al. 1992) bands. They derived the magnitude of the pulsed emission $`m__B=26.0\pm 0.4`$ mag, and found that the signal shows two peaks with a phase separation of $`0.5`$. It appeared that there is a phase agreement of the optical data with $`\gamma `$ and hard $`X`$-ray curves. Moreover, the form of the optical light curve resembles that of $`\gamma `$ and hard $`X`$-rays rather than the soft $`X`$-rays signature. According to Shearer et al. (1998), this is a further evidence of the predominantly magnetospheric origin of Geminga’s pulsed optical emission (Halpern et al. 1996) over a thermal one (Bignami et al. 1996). We suggest that Geminga’s pulsed optical emission is a synchrotron radiation of the plasma particles gyrating about relatively weak magnetic field of the distant magnetosphere of this pulsar. As we demonstrate below, the plasma particles acquire perpendicular momenta, necessary for the synchrotron emission, due to cyclotron absorption of radio waves on these particles. There is an evidence that the Geminga pulsar is either radio quiet or, more likely, visible only at low radio frequencies below about $`100`$ MHz (e.g. Malofeev & Malov 1997; Vats et al. 1999; see also Paper I). ## 2 Estimation of pitch-angles Mikhailovskii et al. (1982) were first to suggest that the energy lost by radio waves due to cyclotron damping should lead to ’heating’ of the ambient plasma, i.e. increase of its perpendicular temperature. Lyubarskii & Petrova (1998) treated thoroughly spontaneous re-emission of the absorbed energy, and concluded that in short-period pulsars a significant fraction of this energy is re-emitted in the far infrared band. They found that an electron (or positron), moving initially with the relativistic velocity strictly along the local magnetic field, obtains a pitch-angle<sup>1</sup><sup>1</sup>1The pitch-angle is defined as $`\psi \mathrm{arctan}\left(p_{}/p_{}\right)`$, where $`p_{}`$ and $`p_{}`$ are the components of the particle momentum along and across the local magnetic field, respectively. $$\psi \theta \eta ^{1/2}\mathrm{rad}$$ (1) in the course of cyclotron absorption of radio emission. Here $`\theta `$ is an angle between the wave vector $`𝒌_\mathrm{r}`$ and the local magnetic field $`𝑩`$, and $$\eta =\frac{L_\mathrm{r}}{L_{p1}}$$ (2) is the ratio of the radio luminosity to the particle luminosity, that is, the ratio of the power of radio waves (damped due to cyclotron resonance) to the total kinetic energy of the resonant particles produced per second above a fraction of the polar cap with the area $`S_{p1}`$ (see equation 4). Note that equation (1) is valid provided that $`\eta 1`$. Estimation of $`\eta `$ is presented below. The radio luminosity of the Geminga pulsar, by what we mean the integrated radio power of Geminga as it would be in the absence of cyclotron damping on the magnetospheric plasma (Paper I), can be estimated as $$L_\mathrm{r}F\left(100\mathrm{MHz}\right)\mathrm{\Delta }\nu d^2[\mathrm{erg}\mathrm{s}^1],$$ (3) where $`F\left(100\mathrm{MHz}\right)1\mathrm{Jy}=10^{23}\mathrm{erg}\mathrm{cm}^2\mathrm{s}^1\mathrm{Hz}^1`$ (Vats et al. 1999), $`\mathrm{\Delta }\nu 1`$ GHz is a frequency range where the bulk of pulsar radio emission is usually radiated, and $`d160\mathrm{pc}5\times 10^{20}`$ cm is a distance to the Geminga pulsar. Inserting all these values into equation (3) we find that $`L_\mathrm{r}3\times 10^{27}\mathrm{erg}\mathrm{s}^1`$. Let us now evaluate the denominator in equation (2). According to Paper I (see also Malov 1998), the Geminga pulsar is an almost aligned rotator, and for the inclination angle $`\alpha =5^{}`$ and the impact angle $`\beta =20^{}`$, an observer’s line of sight in fact grazes the emission cone radiated from the bundle of the last open field lines, with the opening angle at about 100 MHz being $`\rho _130^{}`$ (Khechinashvili, Melikidze & Gil 1999). The latter defines a group of the radial lines constituting an angle $`\rho _1`$ with the magnetic axis $`𝝁`$ (see Fig. 1). The simulation of radio waves damping along these lines (Paper I; Khechinashvili et al. 1999) shows that $`\nu 100`$ MHz radio waves are damped at the distance $`_22.4`$, whereas $`\nu 1`$ GHz waves are damped at $`_11.1`$. Here $`R/R_{_{\mathrm{LC}}}`$ is a distance from the star’s centre measured in the units of the light cylinder radius ($`R_{_{\mathrm{LC}}}=Pc/2\pi `$). One can speculate that the major part of Geminga’s radio emission is effectively damped within a cut-cone-shaped volume represented in Fig. 1. Then, let us consider the dipolar field line which crosses the radial line, constituting the angle $`\rho _1=30^{}`$ with $`𝝁`$, at the distance $`_11.1`$. The footprint of this field line on the stellar surface has a polar angle $`\vartheta _{p1}\left(R_0/R_1\right)^{0.5}\rho _10.014`$ rad. The footprints of all inner field lines, on which $`0.1÷1`$ GHz radio frequencies are damped at high altitudes (see above), mark a circle with the angular radius $`\vartheta _{p1}`$ within the pulsar polar cap through which the resonant leave the neutron star. Let us notice that $`\vartheta _{p1}0.5\vartheta _p`$, where $`\vartheta _p(R_0/R_{_{\mathrm{LC}}})^{1/2}`$ is an angular radius of the pulsar polar cap, and $`R_010^6`$ cm is a stellar radius. Hence, only a fraction of the polar cap provide resonant particles. The area of this region is $$S_1\pi R_0^2\vartheta _{p1}^26\times 10^8\mathrm{cm}^2,$$ (4) which yields $$L_{p1}=\kappa n_{_{\mathrm{GJ}}}S\gamma __pmc^38\times 10^{27}\kappa \left[\mathrm{erg}\mathrm{s}^1\right].$$ (5) Here the surface value of the corotational charge number density (Goldereich & Julian 1969) $`n_{_{\mathrm{GJ}}}2.2\times 10^{18}\left(\dot{P}/P\right)^{1/2}5.4\times 10^{12}\mathrm{cm}^3`$ was substituted, which was evaluated for the dynamical parameters of the Geminga pulsar ($`P=0.237`$ s and $`\dot{P}=1.1\times 10^{14}`$); $`\kappa `$ is the Sturrock multiplication factor (Sturrock 1971). Evaluating equation (5) we assumed that the mean Lorentz factor of the ambient plasma particles $`\gamma __p100`$. From equation (5) it follows that $$F_{p1}=\frac{L_{p1}}{\gamma __pmc^2}10^{32}\kappa $$ (6) particles are generated per second over the area of $`S_1`$ (equation 4). Substituting the derived values $`L_\mathrm{r}`$ and $`L_{p1}`$ in equation (2) we obtain that $`\eta 0.4\kappa ^1`$. Then, taking into account the average value of the angle $`\theta 0.6`$ rad (Khechinashvili et al. 1999) in equation (1), we find that the characteristic pitch-angle of plasma particles is $`\psi 0.4\kappa ^{0.5}`$ rad. Therefore, a ’kick’ that each electron gets due to the cyclotron absorption of radio waves, depends on the number of particles (determined by the Sturrock multiplication factor $`\kappa `$) among which the energy of these waves is distributed. For example, inside the spark-associated columns of plasma (e.g. Ruderman & Sutherland 1975; Gil & Sendyk 1999), where $`\kappa 10^4`$, we have $`\psi 10^3`$. At the same time, in the regions of reduced number density between the spark-associated columns of plasma, where $`\kappa 10÷100,`$ we obtain $`\psi 0.1.`$ For the particle to emit the synchrotron radiation the condition $$\psi \gamma __p1$$ (7) should be satisfied. Apparently, the latter condition is not met inside the dense plasma ($`\psi \gamma __p1`$ for $`\gamma __p100`$), whereas in the space of reduced number density it is satisfied well. In other words, only the particles, being in the space between the spark-associated plasma columns, obtain (due to cyclotron absorption of radio waves) pitch-angles big enough to start gyrating relativistically, hence, emitting synchrotron radiation. ## 3 The synchrotron model The total radiated power of a single electron synchrotron radiation is defined as (Ginzburg 1979; Lang 1980) $$L_01.6\times 10^{15}B^2\gamma _p^2\mathrm{sin}^2\psi \left[\mathrm{erg}\mathrm{s}^1\right],$$ (8) and the critical frequency of the synchrotron radiation is $$\nu _c4.2\times 10^6B\gamma __p^2\mathrm{sin}\psi \left[\mathrm{Hz}\right].$$ (9) The spectral density of a single electron radiation power near $`\nu _c`$ is $$I_0\left(\nu _c\right)=\frac{L_0}{\nu _c}3.8\times 10^{22}B\mathrm{sin}\psi \left[\mathrm{erg}\mathrm{s}^1\mathrm{Hz}^1\right].$$ (10) Let us notice that, if a source as a whole is moving towards an observer, the power given by equation (8) should be divided by $`\mathrm{sin}^2\psi `$ (see equation (5.12) in Ginzburg (1979) and the discussion on page 75). However, this is not the case in our current application, as the emitting volume represented in Fig. 1 is quasi-stationary with respect to an observer on Earth. Indeed, the particles are permanently flowing in and out but the volume as the whole does not approach an observer with the relativistic velocity. That is why the multiplier $`1/\mathrm{sin}^2\psi `$ should be omitted in this consideration. The average magnetic field in the region where synchrotron emission is generated (Fig. 1) can be estimated from the equation $$B=3.2\times 10^{19}\frac{\left(P\dot{P}\right)^{1/2}}{\mathrm{sin}\alpha }\left(\frac{R_0}{R}\right)^3=\frac{9.4P^{2.5}\dot{P}_{15}^{0.5}}{^3\mathrm{sin}\alpha },$$ (11) as $`B3.3\times 10^3`$ G. This provides the critical frequency $`\nu _c1.4\times 10^{13}`$ Hz (equation 9), which falls into infrared band (the frequency corresponding to the maximum in the single electron synchrotron spectrum yields $`\nu _m0.29\nu _c4\times 10^{12}`$ Hz). From equation (10) we find that $`I_0\left(\nu _c\right)1.3\times 10^{19}\mathrm{erg}\mathrm{s}^1\mathrm{Hz}^1.`$ It is natural to assume that the spectral density of total power radiated by $`N`$ electrons is $$I\left(\nu \right)=NI_0\left(\nu \right).$$ (12) Let us then estimate $`N`$, i.e., the total number of emitting particles. From Fig. 1, assuming the conservation of the particle flux along the dipolar fields lines, we find that in the cut-cone with the height $`l=\left(_2_1\right)R_{_{\mathrm{LC}}}\mathrm{cos}\rho _1=1.3R_{_{\mathrm{LC}}}\mathrm{cos}\rho _11.3\times 10^9`$ cm there are $`NF_{p1}l4\times 10^{30}\kappa `$ particles (where $`F_{p1}`$ is defined from equation 6) at any given moment of time. Then, from equation (12), for $`\kappa =10`$, we obtain $$I\left(\nu _c\right)=6\times 10^{12}\mathrm{erg}\mathrm{s}^1\mathrm{Hz}^1,$$ (13) which is a theoretical value of spectral density near the critical infrared frequency $`\nu _c`$. The spectral density of power in the $`B`$-band can be deduced using the power-law spectrum of Geminga’s optical emission (Martin et al. 1998), as $$I\left(\nu __B\right)=I\left(\nu _c\right)\left(\frac{\nu __B}{\nu _c}\right)^{0.8}3\times 10^{11}\mathrm{erg}\mathrm{s}^1\mathrm{Hz}^1,$$ (14) where $`\nu __B=6.9\times 10^{14}`$ Hz. Let us compare the theoretical value of $`I\left(\nu __B\right)`$ (equation 14) with that provided by the observations. The latter can be derived from the spectral density of flux $`f\left(\nu __B\right)=2\times 10^{30}\mathrm{erg}\mathrm{s}^1\mathrm{Hz}^1`$ cm<sup>-2</sup> (see Fig. 3 in Martin et al. 1998), which gives the spectral density of total power $$I^{\mathrm{obs}}(\nu __B)=\varsigma f\left(\nu __B\right)d^2\mathrm{erg}\mathrm{s}^1\mathrm{Hz}^1,$$ (15) where $`\varsigma `$ is a beaming factor and $`d5\times 10^{20}`$ cm. According to Fig. 1, using the features of the dipolar geometry, we find that $`\varphi 1.5\rho _10.8`$ rad, which leads to the beaming factor $`\varsigma \varphi ^20.6`$ sr (instead of $`4\pi `$ sr for the isotropic radiation). Therefore, equation (15) yields $`I^{\mathrm{obs}}(\nu __B)3\times 10^{11}\mathrm{erg}\mathrm{s}^1\mathrm{Hz}^1`$, which is consistent with our theoretical prediction (equation 14). Obviously, the fraction of the synchrotron optical radiation of Geminga observed on Earth is emitted in the same direction as the low-frequency radio waves (Paper I). Using the feature of dipolar field lines that all of them intersect any given radial line with the same angle, we find a geometrical place of all the points in the magnetosphere where the tangents to the field lines point to the same direction<sup>2</sup><sup>2</sup>2$`\beta =20^{}`$ at the nearest approach of the line of sight to the magnetic axis $`𝝁`$, and $`\beta +2\alpha =30^{}`$ $``$ at the furthest one. as a wavevector $`𝒌_\mathrm{r}`$ of the 100 MHz radio waves. As it is seen in Fig. 2, these points are placed at the radial straight lines inclined to $`𝝁`$ at an angle $`\vartheta 13^{}`$ from the side of the nearest approach to $`𝝁`$ and $`\vartheta 20^{}`$ from the furthest one. The ’optically bright’ region is restricted in the longitudinal direction to the altitudes along these lines, where the cyclotron damping of $`0.1÷1`$ GHz radio waves occurs. Calculations show (Khechinashvili et al. 1999) that the longitudinal size of this region $`l_{}R_{_{\mathrm{LC}}}`$. The dimension of this region in the transverse direction can be estimated as $`l_{}R_c/\gamma _p10^2R_{_{\mathrm{LC}}}`$, where $`R_c`$ is a magnetic field line curvature at corresponding altitude. Such an estimation of $`l_{}`$ assumes that this is a characteristic distance at which a particle moving along the curved magnetic field line keeps emitting towards the observer (i.e., an observer stays in the emission diagram of a singel particle, within the angle range $`1/\gamma __p`$). The corresponding characteristic time is $`\tau _{}l_{}/c4\times 10^4\mathrm{s}`$, during which a particle radiates $`E_0=L_0\tau _{}`$ energy. Obviously, the latter should be much less than the kinetic energy of an electron $`E_k=\gamma __pmc^2`$. Let us notice that in such a consideration the multiplier $`1/\mathrm{sin}^2\psi `$ (see the discussion below equation 10) should be taken into account in equation (8), as the source of the synchrotron radiation (a particle) is moving towards the observer in this case. Using equation (8) with the latter correction we find that $`\chi E_0/E_k10^31,`$ so that only a tiny fraction of the particle’s kinetic energy is radiated away as a synchrotron emission. Assuming that this is valid for all the radiating particles, we can estimate the total radiated power of the synchrotron radiation as $`L=\chi L_{p1}8\times 10^{25}\mathrm{erg}\mathrm{s}^1`$ (where $`L_{p1}`$ is found from equation 5). Then, using equation (10), we obtain $`I\left(\nu _c\right)6\times 10^{12}\mathrm{erg}\mathrm{s}^1\mathrm{Hz}^1`$. We see that this value is consistent with our previous theoretical estimation (equation 13), and hence with the observationally derived value (equation 15). The consideration based on the energy conservation provides an independent test of our model. In the calculations above we assume that both pitch-angles $`\psi `$ and Lorentz factors $`\gamma __p`$ of emitting particles remain constant while the particles cross the ’optically bright’ region. Obviously, this assumption is valid for Lorentz factors because $`\chi 1`$. As for pitch-angles, it requires that a single particle goes through at least few acts of radio-absorption during the time $`\tau _\mathrm{m}=l_{}/c0.04`$ s. This is equivalent to fulfillment of the following condition $$\tau _d/\tau _\mathrm{m}1.$$ (16) Here $`\tau _d=1/\mathrm{\Gamma }`$ is a characteristic time-scale of the cyclotron damping, and $`\mathrm{\Gamma }`$ is the damping decrement. We found (Paper I; Khechinashvili et al. 1999) that $`\mathrm{\Gamma }R/c100`$ at the distances where the bulk of radio band ($`0.1÷1`$ GHz) is typically damped. Therefore, $`\mathrm{\Gamma }2\times 10^3\mathrm{s}^1`$ and $`\tau _d/\tau _\mathrm{m}10^21`$. Thus, the parameters $`\gamma __p`$ and $`\psi `$ used in our model can be considered quasi-stationary. We believe that the apparent optical spectral index $`\alpha 0.8`$ (Martin et al. 1998) indicates a power-law distribution of emitting particles over energy $`n\left(\gamma \right)\gamma ^q`$, where $`q=1+2\alpha 2.6`$ (Ginzburg 1979; Lang 1980). Such a distribution function can be expected in the region where the synchrotron optical radiation of the Geminga pulsar is believed to originate. ## 4 Conclusions The model for the optical radiation of the Geminga pulsar, presented in this paper, is a direct consequence of our model for its erratic radio emission (Paper I). Let us summarise our main results: (i) Relativistic charged particles of the pulsar magnetosphere obtain non-zero pitch-angles due to cyclotron damping of radio waves; (ii) The particles re-emit the absorbed radio energy in the form of synchrotron radiation, with a maximum power in the infrared band; (iii) The observed optical emission of the Geminga pulsar is just a short-wavelength continuation of its power-law synchrotron spectrum; (iv) The optical power estimated from our model, using two different approaches, agrees with the observed value; (v) We predict a significant pulsed infrared emission which should be observed from the Geminga pulsar; (vi) As a by-result, we support a model of the non-uniform plasma outflow in pulsar magnetospheres, originating in the form of spark discharges above the polar cap. ## 5 Acknowledgments The work is supported in parts by the KBN Grants 2 P03D 015 12 and 2 P03D 003 15 of the Polish State Committee for Scientific Research. G. M. and D. K. acknowledge also support from the INTAS Grant 96-0154.
warning/0001/hep-ph0001167.html
ar5iv
text
# ASYMPTOTIC GLUON SHADOWING ## 1 Introduction Nuclear shadowing or the depletion at low Bjorken $`x`$ of the nuclear Deep Inelastic (DI) structure function, $`F_2^A`$, with respect to the nucleon one, $`F_2^N`$, has been observed in a number of experiments (see and references therein). Assuming the universality of parton distributions in nuclei, one expects nuclear shadowing to be present in other high energy processes as well, such as Drell-Yan pair, $`J/\psi `$ and $`\mathrm{{\rm Y}}`$ production in lepton-nucleus, hadron-nucleus and nucleus-nucleus collisions. In particular nuclear shadowing might be concurring with the other mechanisms among which quark-gluon plasma formation, that result in a depletion of the observed cross sections for these processes . A quantitative understanding of both the $`x`$ and $`Q^2`$ dependences of the nuclear parton distributions at low and very low $`x`$ is therefore a necessary step for interpreting the outcome of future experiments at RHIC and at the LHC. Recent calculations rely on non-perturbative models for the nuclear parton distributions at a given (low) scale, $`Q_o^2`$, combined with DGLAP perturbative evolution. They are all therefore affected by the uncertainty in the initial parton distributions and, in particular, in the gluon distribution which governs evolution at low $`x`$, and which is poorly known experimentally. Moreover, evolution to moderate values of $`Q^2`$ ($`Q^210\mathrm{GeV}^2`$) depends even more strongly, by the value of $`Q_o^2`$ itself which, if varied within the range, $`Q_o^2=0.8few\mathrm{GeV}^2`$ can lead to completely different evolution patterns for the shadowing of both the structure functions and the gluon distribution. These simple facts hamper the possibility of predicting the $`Q^2`$ dependence of nuclear shadowing although its behavior within pQCD is quite well understood and it can be predicted accurately for each choice of initial conditions. In this paper we suggest an avenue for dealing with the model dependence associated with the initial conditions of perturbative evolution, which is based on a study of the low $`x`$ and high $`Q^2`$ asymptotic behavior of the shadowing ratios $`R_G=G_A/G_N`$ and $`R_F=F_2^A/F_2^N`$, $`G_{N(A)}`$ and $`F_2^{N(A)}`$ being the gluon distributions and the Deep Inelastic (DI) structure function in a nucleon, $`(N)`$, and in an isoscalar nucleus, $`(A)`$, respectively. In the asymptotic regime defined by $`Q^2Q_o^2`$ and $`xx_o`$, $`x_o0.1`$, the Double Logarithmic Approximation (DLA) to pQCD evolution applies . The proton structure function data analyzed recently at HERA have been proven after the suggestion of Ref., to evolve according to DLA. The key test is to show that the data obey Double Asymtptotic Scaling (DAS) in the variables $`\rho =\gamma ((YY_o)/\xi )^{1/2}`$ and $`\sigma =\gamma ^1((YY_o)\xi )^{1/2}`$, $`\gamma =6/(332N_f)^{1/2}`$, $`Y=\mathrm{ln}1/x`$, $`\xi =\gamma ^2\mathrm{ln}(\mathrm{ln}Q^2/\mathrm{\Lambda }_{QCD}^2/\mathrm{ln}Q_o^2/\mathrm{\Lambda }_{QCD}^2`$). Violations from DAS (other than due to the fact that the data lie in a pre-asymptotic region or to NLO DGLAP corrections b) would signal either the onset of other approximations such as leading-(next-to-leading)-log in $`1/x`$, $`LL_x`$ ($`NL_x`$) resummation, or the beginning of parton saturation. Here we wish to apply a DAS type of analysis to nuclear DI scattering. As a first step we show that if evolution proceeds through ordinary asymptotic pQCD evolution equations in nuclei as well as in a free nucleon, the ratios $`R_F`$ and $`R_G`$ become a function of $`\rho `$ only, which is per se a model independent result. We use this result as a basis for exploring the origin of scaling violations in nuclei. In nuclei in fact, the asymptotic regime is entered in principle at different values of $`x_o^A`$ and $`Q_{o,A}^2`$ than in the proton. In particular Unitarity Shadowing Corrections (USC) are expected to affect evolution at $`x_o^A>x_o`$ because of the increase of the gluon density in a nucleus due to the overlapping of nucleons in the longitudinal direction. On a more speculative basis one might also expect the transition to the $`\mathrm{ln}(1/x)`$ resummation to appear in a different regime. In our approach such questions can be addressed systematically as they introduce specific scaling violations from the DLA result, appearing as a $`\sigma `$ dependence in the ratios $`R_G`$ and $`R_F`$. In summary, although it is technically predictable that in a proton at very low values of $`x`$ and sufficiently large $`Q^2`$, i.e. deeply in the asymptotic region, DGLAP evolution and the DLA should break down and give way to $`\mathrm{ln}(1/x)`$ summation and to USC, it is still a major task to be able to pinpoint where and if the transition from the different regimes is going to take place in the kinematical regimes currently under exploration. Our goal is to obtain some new insight by using nuclear targets where the asymptotic regime can in principle be reached at larger $`x`$. As a by-product we obtain quantitative predictions for RHIC and the LHC. ## 2 DGLAP Evolution in Nuclei We first summarize results for ordinary DGLAP evolution applied to the nuclear ratios at low $`x`$, assuming that the proton and the nuclear distributions evolve along similar paths. As it is well known evolution is driven by the gluon distribution which dominates over the sea quarks one and one can predict the behavior of the shadowing ratios, $`R_G`$ and $`R_F`$ with $`Q^2`$: $`{\displaystyle \frac{R_G}{\mathrm{ln}Q^2}}{\displaystyle _x^1}P_{GG}({\displaystyle \frac{x}{y}},\alpha _S(Q^2)){\displaystyle \frac{G_N(y,Q^2)}{G_N(x,Q^2)}}`$ $`\times \left[R_G(y,Q^2)R_G(x,Q^2)\right]{\displaystyle \frac{dy}{y}}`$ $`{\displaystyle \frac{G_A(x,Q^2)/\mathrm{ln}Q^2}{G_N(x,Q^2)/\mathrm{ln}Q^2}}R_G(x,Q^2)`$ , (1) $`{\displaystyle \frac{R_F}{\mathrm{ln}Q^2}}{\displaystyle _x^1}P_{qG}({\displaystyle \frac{x}{y}},\alpha _S(Q^2)){\displaystyle \frac{G_N(y,Q^2)}{\mathrm{\Sigma }_N(x,Q^2)}}`$ $`\times \left[R_G(y,Q^2)R_F(x,Q^2)\right]{\displaystyle \frac{dy}{y}}`$ , (2) where we have disregarded the sea quarks distribution on the r.h.s. of the coupled DGLAP evolution equations; $`P_{qG}`$ and $`P_{GG}`$ are the splitting functions evaluated at NLO; and we used the approximation $`F_2^{N(A)}5/18\mathrm{\Sigma }_{N(A)}`$. Eqs.(1) and (2) show that the $`Q^2`$ dependence of $`R_G`$ and $`R_F`$ is regulated by a subtle balance involving the parton distributions and the ratios themselves . The following predictions can be made for the ratios $`R_G`$ and $`R_F`$: If, as predicted by current non-perturbative shadowing models $`R_G`$ is a growing function of $`x`$, then it also grows with $`Q^2`$, the r.h.s. of Eq.(1) being positive ($`yx`$). On the other side, defining $`R_G(x,Q_o^2)R_G^o`$ and $`R_F(x,Q_o^2)R_F^o`$, one obtains two opposite behaviors for $`R_F`$ namely: i) if $`R_F^o<R_G^o`$, then $`R_F`$ grows with $`Q^2`$; ii) if instead $`R_F^o>R_G^o`$, then $`R_F`$ initially decreases with $`Q^2`$ until it reaches the value of $`R_G^o`$ and it subsequently starts increasing along with $`R_G`$. <sup>1</sup><sup>1</sup>1Although the form of Eq.(1) might seem suggestive of a fixed point behavior , this is a priori not the case, since the quantity $`G_A(x,Q^2)/\mathrm{ln}Q^2/G_N(x,Q^2)/\mathrm{ln}Q^2`$ depends on $`Q^2`$. The “rate” of change with $`Q^2`$ is governed by the ratios $`G_N(y)/G_N(x)`$ and $`G_N(y)/\mathrm{\Sigma }_N(x)`$, at $`Q^2=Q_o^2`$, respectively. If $`Q_o^21\mathrm{GeV}^2`$, then a rapid evolution strongly reduces the shadowing in both $`R_G`$ and $`R_F`$, between $`Q_o^2`$ and $`Q^223\mathrm{GeV}^2`$. If on the contrary $`Q_o^2`$ ranges from $`2`$ to $`5\mathrm{GeV}^2`$ then the evolution is slower and shadowing remains large even at $`Q^210100\mathrm{GeV}^2`$ (these results affect quarkonia production where $`Q^2M_{J/\psi }^210\mathrm{GeV}^2`$ and $`Q^2M_\mathrm{{\rm Y}}^2100\mathrm{GeV}^2`$). ## 3 Double Asymptotic Scaling We now examine the nuclear DI structure function and gluon distribution in the asymptotic regime. The derivation of the equations of the DLA in a nucleus parallels the one for the proton, namely one first writes the LO DGLAP evolution equation for the gluon distribution in the limit $`x0`$, in moments space (we have omitted the subscripts $`N(A)`$ unless necessary): $$\frac{g(n,Q^2)}{\mathrm{ln}Q^2}=\frac{\alpha _s(Q^2)}{2\pi }\gamma _{GG}^0(n)g(n,Q^2),$$ (3) $`\gamma _{gg}^0(n)2C_A/(n1)`$ being the anomalous dimension in the limit $`n1`$. Solutions in $`(x,Q^2)`$ are found by evaluating the anti-Mellin transform, $`G(x,Q^2)`$ $`=`$ $`{\displaystyle \frac{1}{2\pi i}}{\displaystyle _C}𝑑ng(n,Q_0^2)`$ (4) $`\mathrm{exp}\left\{(n1)Y+\xi /(n1)\right\},`$ with the saddle point method. In Eq.(4) $`g(n,Q_0^2)=_0^1𝑑xx^{(n1)}g(x,Q_0^2)`$, $`g(x,Q_o^2)`$ being the inital gluon distribution, and $`G(x,Q^2)=xg(x,Q^2)`$. If one takes a “soft” initial condition such as, $`G(x,Q_0^2)A_Nx^\lambda `$, $`\lambda 0`$, and $`Y`$ and $`\xi `$ are both similarly large, then $`g(n_0,Q_0^2)A_N/(n1)`$, and the saddle point is: $`n_0=1/(2\mathrm{\Delta }Y)+\sqrt{1/(4\mathrm{\Delta }Y)^2+\xi /\mathrm{\Delta }Y}\sqrt{1+\xi /\mathrm{\Delta }Y}`$, $`\mathrm{\Delta }Y=YY_o`$, $`Y_o=\mathrm{ln}(1/x_o)`$, yielding: $$G(x,Q^2)=\sqrt{2\pi }\left(\frac{\stackrel{~}{g}(n_o,Q_o^2)}{\stackrel{~}{g}^{\prime \prime }(n_o,Q_o^2)}\right)^{1/2}\stackrel{~}{g}(n_o,Q^2),$$ (5) with $`\stackrel{~}{g}(n,Q^2)=g(n,Q^2)\mathrm{exp}[(n1)Y]`$. <sup>2</sup><sup>2</sup>2We have omitted sub-leading corrections for simplicity. In terms of the DAS variables, $`\rho `$ and $`\sigma `$, $`n_0=1+\rho /\gamma ^2`$ and: $$GG^{DLA}(x,Q^2)=f(\rho ,\sigma )\mathrm{exp}(2\gamma \sigma ),$$ (6) where $`f(\rho ,\sigma )`$ is a function that depends on the explicit form of the initial conditions. The main result is that $`\mathrm{ln}(G^{DLA})/f(\rho ,\sigma ))`$ is predicted to be linear in $`\sigma `$, the slope being fixed by $`\gamma `$. A similar behavior is found for $`F_2^N`$ by solving the equation: $$\frac{F_2^N(x,Q^2)}{\mathrm{ln}Q^2}=\frac{2}{9}\frac{\alpha _S}{\pi }G(x,Q^2).$$ (7) Deviations from DAS due to hard initial conditions i.e. when $`\lambda 0.2`$ are technically obtained as follows: $`g(n_0,Q_0^2)A_N/(n(\lambda +1))`$ and a physical solution in the limit $`Y\xi `$ is $$G(x,Q^2)=C/\lambda e^{\lambda (YY_o)+\xi /\lambda },$$ (8) corresponding to the saddle point value, $`n_0(\lambda )=1/\mathrm{\Delta }Y+\lambda +1`$ (see also ). Physically this behavior supports the appearance of USC which become more important just because of the steep rise at small $`x`$. The data of Ref. favor the DAS scenario, whereas the more recent data at lower $`x`$ and $`Q^2`$ are best interpreted in terms of USC . In this context it is therefore important to study nuclear shadowing because, due to the $`A`$-dependent increase in gluon density, the physics we are interested in is expected to show up at larger $`x`$ values and because one can tune the variable $`A`$ in order to discriminate among models. In a nucleus we make the two following assumptions for the asymptotic regime: i) the initial distributions are shadowed because of some non-perturbative mechanism, the evolution equations are not affected by the medium; ii) besides the non-perturbative shadowing the evolution equations have screening corrections at a larger $`x_A`$. The derivation of the ratio $`R_G`$ in the hypothesis i) yields: $`R_G^{Asym}=\left[{\displaystyle \frac{G_A(n_o,Q^2)}{G_N(n_o,Q^2)}}\right]^{3/2}\left[{\displaystyle \frac{G_N^{\prime \prime }(n_o,Q^2)}{G_A^{\prime \prime }(n_o,Q^2)}}\right]^{1/2}`$ $`=f_A(\rho ){\displaystyle \frac{\rho h_N^{(1)}(\rho )+\gamma ^2\sigma h_N^{(2)}(\rho )}{\rho h_A^{(1)}(\rho )+\gamma ^2\sigma h_A^{(2)}(\rho )}}`$ . (9) where $`f_A(\rho )=G_A(n_o,Q_o^2)/G_N(n_o,Q_o^2)`$, and the functions $`h_{N(A)}^{(1,2)}`$ are listed explicitely elsewhere. The main point is that the exponential terms appearing in $`G^{DLA}`$, Eq.(6) calculated for a nucleus and for a nucleon respectively, cancel exactly and the remaining dependence on $`\sigma `$ is small asymptotically. This scaling result (shown in Fig.1) will be our reference point. The onset of a different evolution mechanism in the nucleus will appear as a $`\sigma `$-scaling violation modifying the exponential behavior of Eq.(5). In general one can write $$R_G=R_G^{Asym}(\rho )\times \mathrm{exp}\left\{2\gamma (\sigma _A\sigma )\right\},$$ (10) where $`R_G^{Asym}`$ is the large $`\sigma `$ limit of Eq.(9). The form of $`\sigma _A\sigma _A(x,Q^2)`$ describes the approximation to pQCD evolution in a nucleus. In the case of a “hard pomeron” boundary conditions plus screening corrections it is $$\sigma _A=D_A(x,Q^2)\left(\lambda \sigma \rho +\frac{\gamma ^2}{\lambda }\frac{\sigma }{\rho }\right),$$ (11) $`D_A`$ being a damping correction (more details are going to be found in ). In conclusion, we have shown model independent predictions for gluon shadowing in nuclei in the asymptotic region. We have also outlined an approach for more detailed studies of the violations from DAS in nuclei as a signal of pQCD evolution mechanisms other than DGLAP or the DLA. Our calculations will be relevant in the regime accessible at future experiments at RHIC, LHC and possibly at the $`eA`$ project at DESY . ## Acknowledgments The authors wish to thank the Institute of Nuclear and Particle Physics at the University of Virginia for financial support. ## References
warning/0001/gr-qc0001025.html
ar5iv
text
# The self-force on a static scalar test-charge outside a Schwarzschild black hole ## I INTRODUCTION In order to gain deeper quantitative understanding of highly relativistic binary star systems spiraling toward coalescence, a number of authors have used perturbation theory to study the motion of test-particles orbiting black holes. (See e.g. .) The basic idea is to solve the linearized field equations with out-going radiation boundary conditions on a Schwarzschild or Kerr black-hole background. The source of the perturbing field is generally chosen to be a test-particle (endowed with a small scalar charge, electric charge, or mass charge) moving on a bound geodesic orbit, e.g., a circle. Once the perturbing field (scalar, electromagnetic, or gravitational) is calculated, the energy radiated per orbit (i.e., the time-averaged energy flux) can be computed by a performing a surface integral over a distant sphere surrounding the system. The rate energy is carried away by the radiation is then equated to a loss of orbital energy of the particle; thus one can compute the rate of orbital decay.<sup>*</sup><sup>*</sup>* Finding a gravitational-wave signal in noisy detector data requires an accurate prediction of the orbital decay rate. See Thorne . In effect, this energy-balance argument gives the time-averaged radiation-reaction force associated with imposing out-going radiation boundary conditions . This method has been used successfully to compute the inspiral rate of coalescing binaries to very high relativistic order, and it has been used to check analogous post-Newtonian calculation of the inspiral . These perturbation calculations also add insight into effects such as wave “tails”. (See and for post-Newtonian and perturbation-theory discussions of tails.) However, in spite of the success of the perturbation approach, the method, as it has been applied, has a drawback: it has only been used to compute time-averaged, dissipative forces on the particle. The time-averaging is problematic for two reasons: (1) During the late stages of inspiral the orbit will be decaying swiftly, and there will not be many orbits left before the final splat to average over. (2) In the curved spacetime near a black hole, the self-forces experienced by the particle are not instantaneous forces. Rather these forces arise because the fields produced by the particle at one instant travel away from the particle, encounter the curvature of the spacetime, and then scatter back and interact with the particle at a later time. (See DeWitt and DeWitt for an illuminating discussion of this point.) Thus the fields produced when the particle moves through, say, periastron, and suffers the maximum coordinate acceleration, will not come back to interact with the particle until later in the orbit when the coordinate acceleration will be more gentle. In other words, the state of motion that produced the fields, will not be the state of motion affected by the fields. Although it is unlikely that taking into account the non-local nature of the self-forces will qualitatively change any of the results found by the “averaging” technique, the issue has never been rigorously addressed with perturbation theory. In addition to the dissipative forces, the perturbing fields will produce other non-dissipative — conservative — self-forces on the particle, and these forces have not been studied by the conventional perturbation-theory approach . \[As DeWitt and DeWitt have demonstrated, these forces can be computed using a different perturbative approach.\] For example, suppose we have an electric charge in a circular orbit around a Schwarzschild black hole. Even if we neglect the dissipative effects of the radiation reaction, which cause a secular decay of the orbit, the charged particle still does not travel on an exact geodesic of the spacetime. The particle will feel an additional conservative force — proportional to square of the charge — that pushes it slightly off the geodesic. As this force is not radiative in nature, the portion of the field that produces it is explicitly discarded in the conventional perturbation approach because it falls off faster than $`O[1/r]`$, and therefore any effect it might have on the motion is missed by a technique that uses a surface integral on a distant sphere to determine the back-reaction effect. (2) Another perturbation calculation by Gal’tsov has used the “half-retarded–minus–half-advanced” Green’s function to find the radiation reaction forces. But these calculations also miss the conservative (i.e., time symmetric) parts of the force, because they simply cancel when the two parts of the Green’s function are subtracted. In other words, the Gal’tsov result differs from the DeWitt-DeWitt result in that it misses the conservative part of the force. We can try (but as we will see, fail) to understand the origin of the conservative forces by considering a static electric test-charge held fixed by some non-conducting mechanical struts outside a Schwarzschild black hole. The charge experiences a self-force (which is clearly not a radiation-reaction force in any conventional sense) proportional to the square of the electric charge. A physical origin of such a force might be naively described as follows: the surface of the uncharged black hole will act as a conductor, and therefore the presence of the external charge will induce a dipolar charge distribution on the horizon . The magnitude of the induced dipole moment will depend on the size of the dipole (in this case, the mass $`M`$ of the Schwarzschild black hole) and the magnitude $`e`$ of the external charge. This leads one to believe that the charge will experience an attractive dipolar force scaling as $`b_s^5`$, where $`b_s`$ is the Schwarzschild radial coordinate of the charge. This force will be in addition to the natural attractive force (the weight of the particle) which, according to Newton, should scale as $`b_s^2`$. Thus our naive suspicion is that, in addition to the weight, the strut holding the charge fixed will have to counteract an attractive self-force of the form $$F_{(\mathrm{naive})}\frac{e^2M}{b_s^5}.$$ (1) Smith and Will (SW) have calculated the force required to hold an electric charge fixed outside a Schwarzschild black hole. They find the total force exerted by the mechanical strut holding the charge fixed must be $`F_{(\mathrm{strut})}={\displaystyle \frac{M\mu _{\mathrm{ren}}}{b_s^2}}(1{\displaystyle \frac{2M}{b_s}})^{1/2}{\displaystyle \frac{e^2M}{b_s^3}}.`$ (2) Here $`\mu _{\mathrm{ren}}`$ is the renormalized mass of the test-charge. We use units in which $`G=c=1`$. As expected, the first term scales as $`O[b_s^2]`$, and shows that the strut must support the weight of the particle; such a term would be present whether or not the particle is electrically charged. The second term shows — contrary to the “physical intuition” presented above — that the black hole tries to repel the charged particle and the repulsion scales as $`O[b_s^3]`$. Although the second term in Eq. (2) appears only to be accurate to leading order in $`M`$, the result is exact for static charges; the appearance is only a beautiful artifact of Schwarzschild coordinates. Switching to harmonic or isotropic coordinates makes the second term appear as an infinite series in $`M`$. See Eq. (IVb). SW show for almost any choice of variables in Eq. (2) the first term dominates, and the second term only lightens the load a small amount. The fact that our naive physical intuition led to both the wrong direction and the wrong scaling of the force suggests such calculations should be carried out in detail before conclusions are drawn. In this paper we carry out the analogous calculation for a scalar charge in the presence of a Schwarzschild black hole, i.e. we compute the force required to hold a scalar charge fixed. The primary result of this paper is to show that, although the scalar field does contribute to the renormalized mass, there is no force analogous to the second term in Eq. (2) for a scalar charged particle. The primary motivation for this calculation is much broader: to begin to develop techniques and formalism for tackling a sequence of problems, namely, calculating all the forces, linear in the fields (dissipative and non-dissipative), that act on test-charges and masses moving in the proximity of a black hole. As the implicit underlying motivation for this work is calculating the forces on moving charges, we can ask: does our no–self-force–on–a–static–scalar–charge result carry over to the case of a moving charge? Not directly. However, we can gain some intuition about extending our static results to dynamic results for scalar charges by examining the static and dynamic cases for electric charges. Although the SW force calculation is only valid in the static limit, DeWitt and DeWitt show that when the electric charge is in slow motion around the black hole, in addition to the conventional radiation reaction force, there is a conservative repulsive force acting on the charge. This force is independent of the velocity, and clearly corresponds to the force found in the static calculation, i.e., the second term in Eq. (2). DeWitt and DeWitt only compute the force to leading order in the mass of the hole and their result appears to agree exactly with the second term in Eq. (2). However, the agreement is really only at leading order in $`M`$. See the previous footnote. Therefore, it seems reasonable to expect a correspondence between the static case and the slow-motion case for scalar charges. For scalar charges, this means the absence of a self-force on a static charge should imply there is no conservative (repulsive or attractive) force on a slowly orbiting charge. This, however, does not rule out the presence of a higher-order, conservative force that scales as, say, the square of the orbital velocity. (This issue is under vigorous investigation .) As a consequence, although both electric and scalar charges will spiral inward due to radiation reaction forces, scalar charges will not suffer the same persistent, velocity-independent, conservative force trying to hold them off the geodesic. As a matter of principle, many of the delicate issues of radiation reaction forces in curved spacetime we will be dealing with in this paper have already been addressed in quite a general context by DeWitt and Brehme (using a world-tube calculation), by Quinn and Wald (using an axiomatic approach), and Mino, Sasaki and Tanaka (using matched asymptotic expansions). However, in practice, the results presented in these papers have never been applied to give even simple results like Eq. (2), nor did they work out the general expressions for forces on scalar charges. We begin in Sec. II by calculating the scalar field produced by a static scalar charge in Schwarzschild spacetime. The standard method for solving the field equations on a black-hole background is to use separation of variables and decompose the solution into Fourier, radial and angular modes (spherical harmonics). (See Appendix B for an example.) The result is an infinite-series solution for the field. This method works well for computing the far-zone properties of the field (e.g., the energy flux) where the field is weak and can be accurately described by the first few terms in the series. However, when computing a force, such as Eq. (2) (or its scalar analogue, as we are doing here), the behavior of the field on a distant sphere will not suffice; we need to know the detailed behavior of the field up close to the particle. For this it will not be sufficient to describe the field by a few multipoles. Describing the singular nature of the field near the charge requires an infinite number of multipoles ; therefore it is not surprising that SW did not use the standard multipolar expansion of the electrostatic field (e.g., ) in their force calculation. Rather, when embarking on their electrostatic force calculation SW comment: “… an exact calculation is made possible by the fortuitous existence of a previously discovered analytic solution to the \[static\] curved-space Maxwell’s equations”. The exact solution to which they refer is an old result due to Copson and modified by Linet which gives a closed-form expression for the electrostatic potential of a fixed charge residing in Schwarzschild spacetime. \[See Eq. (B).\] The Copson-Linet formula has the advantage that the singular nature of the field near the electric charge is manifest, thus allowing simple calculations in the close proximity of the particle. We begin by constructing the scalar-field analogue of the Copson-Linet result: a closed-form expression for the field of a fixed scalar charge in Schwarzschild spacetime. Our derivation is similar to Copson’s, and is based on constructing the Hadamard “elementary solution” of the scalar-static field equation. As in the Copson-Linet electrostatic solution, our solution will clearly show the divergent behavior of the field near the particle. In Appendix A, we outline Copson’s derivation of the closed-form expression for the electrostatic field with the Hadamard formalism. In Appendix B, we generate some interesting summation formulas by equating the closed-form solutions with the infinite series solutions . As a historical note, we mention that Copson obtained his solution for a static electric charge in a Schwarzschild spacetime in 1928. This is approximately 40 years before the term “black hole” was coined and researchers began formulating the no-hair theorem. However, Copson was probably the first to note one of the important features of the no-hair theorem: “… the potential of an electron on the boundary sphere $`R=\alpha `$ \[on the horizon\] is independent of its position on the sphere, a rather curious result.” Our force calculation in Sec. III closely follows the SW calculation of the electrostatic force. We use the exact expression for the scalar-static field found in Sec. II to compute the stress-energy tensor and the force density in a local, freely-falling frame near the fixed scalar charge. By integrating the force density over a small spherical volume of radius $`\overline{ϵ}`$ centered on the charge (spherical in the freely-falling frame) we compute the total force the strut must supply to hold the charge fixed. In the limit where the radius of the spherical region of integration shrinks to zero, there remains a formally divergent piece of the force scaling as $`(\mathrm{charge})^2/\overline{ϵ}`$. This factor multiplies an acceleration that is identical to the acceleration in the first term in Eq. (2); therefore the infinite term is simply absorbed into the definition of the mass of the particle, i.e., in Eq. (106) we define $$\mu _{\mathrm{ren}}\mu _{\mathrm{bare}}+\frac{1}{2}\underset{\overline{ϵ}0}{lim}\frac{q^2}{\overline{ϵ}},$$ (3) where $`q`$ is the scalar charge of the particle, and $`\mu _{\mathrm{bare}}`$ is the bare mass of the particle. In this way, the scalar charge does contribute to the renormalized mass of the particle in exactly the same way as the electric charge contributes to the renormalized mass in Eq. (2). The form of classical mass renormalization depicted in Eq. (3) is seen in all calculations of this type (e.g.,). The main conclusion of this paper is that, after the formally infinite piece is absorbed into the renormalized mass, there remains no scalar counterpart to the second term in Eq. (2). Also in Sec. III, we verify this answer using conservation of energy. Neither the charge’s contribution to the renormalized-mass term or to the repulsive term in Eq. (2) should be confused with the contribution to the gravitational force on the particle due the stress energy of the electric field perturbing the metric of the spacetime. Such a force would scale as $`\mu e^2`$. In this calculation we are explicitly ignoring corrections to the metric. See the discussion following Eq. (IVe). In Sec. IV, we discuss a number of alternative methods for solving similar problems, as well as similarities and differences in the scalar-static and electrostatic results. We also collect what is known about the forces on static charges (mass, electric and scalar) in Schwarzschild spacetime. This gives a clear indication of what future research is needed. ## II Solution of the scalar field for a fixed charge in Schwarzschild spacetime In order to find the self-force acting on the scalar charge, we first solve the massless scalar field equation $$\mathrm{}V(1/\sqrt{g})(\sqrt{g}g^{\alpha \beta }V_{,\alpha })_{,\beta }=4\pi \rho $$ (4) in Schwarzschild spacetime.<sup>§</sup><sup>§</sup>§Since the Ricci scalar curvature $`R`$ vanishes in the Schwarzschild spacetime, it would seem that including coupling to the curvature (e.g. a conformally invariant term $`\frac{1}{6}RV`$) in Eq. (4) would have no effect on our results. However, if we include coupling to the curvature, the stress-energy that enters the force calculation \[Eq. (73)\] would have to be modified also. Therefore, we make no claim that our results hold for non minimally coupled fields. Here commas denote partial differentiation, Greek indices run $`0`$ to $`3`$, and Latin indices will run $`1`$ to $`3`$. The source $`\rho `$ for our field will be a point-like scalar charge which can be described by $`\rho (t,𝐱)`$ $`=`$ $`q{\displaystyle _{\mathrm{}}^{\mathrm{}}}(1/\sqrt{g})\delta ^4(x^\alpha b^\alpha (\tau ))𝑑\tau `$ (5) $`=`$ $`{\displaystyle \frac{q}{u^t(t)}}{\displaystyle \frac{\delta ^3(𝐱𝐛(t))}{\sqrt{g}}},`$ (6) where $`b^\alpha (\tau )`$ is the spacetime trajectory of the charged body, $`\tau `$ is the proper time measured along the path, and $`u^t=db^0/d\tau =dt/d\tau `$. \[Notice, $`\rho d(\mathrm{proper}\mathrm{volume})=\rho u^t\sqrt{g}d^3x=q`$ in a frame comoving with the charge.\] We will later restrict our attention to a static field and a fixed source charge at $`𝐛=b\widehat{𝐳}`$, but, for the present, we will leave the time-dependence in the equations. We use isotropic coordinates to describe the Schwarzschild geometry. The line element is $$ds^2=\frac{(2rM)^2}{(2r+M)^2}dt^2+(1+\frac{M}{2r})^4(dx^2+dy^2+dz^2).$$ (7) In these coordinates Eq. (4) can be written $$C^{ij}\stackrel{~}{V}_{,ij}(\omega ,𝐱)+C^j\stackrel{~}{V}_{,j}(\omega ,𝐱)+C\stackrel{~}{V}(\omega ,𝐱)=4\pi (1+M/2r)^4\stackrel{~}{\rho }(\omega ,𝐱),$$ (8) where $`C^{ij}`$ $`=`$ $`\mathrm{diag}(1,1,1),`$ (10) $`C^j`$ $`=`$ $`h(r){\displaystyle \frac{x^j}{r}}={\displaystyle \frac{d}{dr}}\left[\mathrm{ln}(1(M/2r)^2)\right]{\displaystyle \frac{x^j}{r}},`$ (11) $`C`$ $`=`$ $`\omega ^2{\displaystyle \frac{(1+M/2r)^6}{(1M/2r)^2}},`$ (12) and we have used the Fourier transform $$V(t,𝐱)=_{\mathrm{}}^{\mathrm{}}\stackrel{~}{V}(\omega ,𝐱)e^{i\omega t}𝑑\omega $$ (13) to eliminate the time variable. We can also write Eq. (8) in the more compact form $`^2\stackrel{~}{V}+h(r)\stackrel{~}{V}_{,r}+C(r,\omega )\stackrel{~}{V}=4\pi (1+M/2r)^4\stackrel{~}{\rho }(\omega ,𝐱),(\text{8}^{})`$ (14) where $`^2`$ is the flat-space Laplacian. Our primary attention will be focused on the static — zero frequency — case ($`\omega =C=0`$), but the formalism we are developing is valid for the “Helmholtz”-type equation in Eqs. (8) and (8). Following Hadamard (p. 92-107), the elementary solution (i.e. the Green’s function ) for Eq. (8) takes the form $$U_{elem}=\mathrm{\Gamma }^{1/2}(U_0+U_1\mathrm{\Gamma }+U_2\mathrm{\Gamma }^2+\mathrm{}),$$ (15) where $`\mathrm{\Gamma }`$ is the square of the geodesic distance (in the sense of the “metric” $`C^{ij}=\delta ^{ij}`$) from the source point $`𝐱^{}`$ to the field point $`𝐱`$, i.e., $$\mathrm{\Gamma }=(xx^{})^2+(yy^{})^2+(zz^{})^2.$$ (16) Since we are working in three dimensions (an odd number of dimensions), there is no natural-log term $`(\mathrm{ln}\mathrm{\Gamma })`$ in the elementary solution. The $`U_n`$’s are non-singular functions everywhere outside the horizon. Recall, in isotropic coordinates the horizon is at $`r=M/2`$. The simple form of $`\mathrm{\Gamma }`$ in Eq. (16) is a consequence of the isotropic coordinates we are using; therefore the technique we are developing cannot be easily extended to geometries that do not have spatially isotropic coordinates, e.g. Kerr geometry. The formula for the leading order behavior of the series is given by Hadamard (p. 94) $$U_0(𝐱,𝐱^{})=\frac{1}{\sqrt{detC^{ij}}}\mathrm{exp}\left\{_0^\lambda \left[C^{ij}\mathrm{\Gamma }_{,ij}+C^j\mathrm{\Gamma }_{,j}6\right]\frac{d\lambda }{4\lambda }\right\},$$ (17) where $`\lambda `$ is the arc length measured along the geodesic connecting the source and field points. Letting $`\theta `$ denote the angle between the two spatial vectors $`𝐱`$ and $`𝐱^{}`$ and using Eqs .(II) and (16) we have $`U_0(𝐱,𝐱^{})`$ $`=`$ $`\mathrm{exp}\left\{{\displaystyle _0^\lambda }\left[h(r)(rr^{}\mathrm{cos}\theta )\right]{\displaystyle \frac{d\lambda }{2\lambda }}\right\},`$ (19) $`=`$ $`\mathrm{exp}\left\{{\displaystyle _r^{}^r}h(r^{\prime \prime }){\displaystyle \frac{dr^{\prime \prime }}{2}}\right\},`$ (20) $`=`$ $`\sqrt{{\displaystyle \frac{1(M/2r^{})^2}{1(M/2r)^2}}},`$ (21) where $`r=|𝐱|`$, $`r^{}=|𝐱^{}|`$, and $`r^{\prime \prime }`$ is the dummy integration variable over $`r`$. We have also used the geometric relationship $`(rr^{}\mathrm{cos}\theta )d\lambda =\lambda dr`$. Because we are looking for an axially symmetric solution, we may assume the $`U_n`$’s are functions of only the radial variables \[e.g., notice $`U_0=U_0(r,r^{})`$\]. We now substitute Eq. (15) into Eq. (8), use relations such as $`\delta ^{ij}\mathrm{\Gamma }_{,i}\mathrm{\Gamma }_{,j}=4\mathrm{\Gamma }`$, and collect powers of $`\mathrm{\Gamma }`$. The result is $`(2U_{0,r}+hU_0)\left({\displaystyle \frac{r^2r_{}^{}{}_{}{}^{2}}{2r}}\right)\mathrm{\Gamma }^{3/2}+{\displaystyle \underset{n=0}{}}[(2n+1)(2U_{n+1,r}+hU_{n+1})\left({\displaystyle \frac{r^2r_{}^{}{}_{}{}^{2}}{2r}}\right)`$ (22) $`+2(2n^2+n1)U_{n+1}+{\displaystyle \frac{1}{r}}(rU_n)_{,rr}+hU_{n,r}+{\displaystyle \frac{2n1}{2r}}(2U_{n,r}+hU_n)+C(r,\omega )U_n]\mathrm{\Gamma }^{n1/2}=0.`$ (23) (24) Since $`U_{elem}`$ is a solution to the homogeneous equation everywhere (except at the source point) the coefficients of each power of $`\mathrm{\Gamma }`$ must vanish independently. The first term gives an equation for $`U_0`$ $$2U_{0,r}+hU_0=0.$$ (25) Notice that our Eq. (17) has already given us a particular solution (Eq. (21)) to this equation. Setting the coefficient of $`\mathrm{\Gamma }^{n1/2}`$ to zero and using an integrating factor, we obtain a recursion relation for the $`U_n`$’s $`{\displaystyle \frac{U_{n+1}(r,r^{})}{U_0}}={\displaystyle \frac{1}{(r^2r_{}^{}{}_{}{}^{2})^{n+1}}}{\displaystyle _r^{}^r}`$ $`{\displaystyle \frac{(r_{}^{\prime \prime }{}_{}{}^{2}r_{}^{}{}_{}{}^{2})^n}{2n+1}}[{\displaystyle \frac{(r^{\prime \prime }U_n)_{,r^{\prime \prime }r^{\prime \prime }}}{U_0}}{\displaystyle \frac{2r^{\prime \prime }U_{0,r^{\prime \prime }}U_{n,r^{\prime \prime }r^{\prime \prime }}}{U_0^2}}`$ (27) $`+(2n1)(U_n/U_0)_{,r^{\prime \prime }}+r^{\prime \prime }C(r^{\prime \prime },\omega )(U_n/U_0)]dr^{\prime \prime }.`$ In the integrand the $`U_n`$’s are functions of the dummy integration variable $`r^{\prime \prime }`$ and the source point $`r^{}`$, i.e. $`U_n=U_n(r^{\prime \prime },r^{})`$. This recursion relation allows us to construct a series solution for the Green’s function for the Schwarzschild-Helmholtz equation Eq. (8) or (8). We can gain confidence in the recursion relation Eq. (27) by applying it to the true Helmholtz equation in flat spacetime (Eq. (8) with $`M=0`$, but $`C\omega ^20`$). In this case, Eq. (17) gives $`U_0=1`$, and the recursion relation Eq. (27) gives $`U_1=\omega ^2/2`$, $`U_2=\omega ^4/24`$, $`\mathrm{}`$, $`U_n=(1)^n\omega ^{2n}/(2n!)`$. Summing the series gives the Green’s function for the Helmholtz equation $`U_{elem}={\displaystyle \frac{\mathrm{cos}(\omega |𝐱𝐱^{}|)}{|𝐱𝐱^{}|}}.`$ (28) Now compute the inverse Fourier transform and we have $`G(x,x^{})={\displaystyle \frac{1}{8\pi }}\left\{{\displaystyle \frac{\delta [t^{}(t|𝐱𝐱^{}|)]}{|𝐱𝐱^{}|}}+{\displaystyle \frac{\delta [t^{}(t+|𝐱𝐱^{}|)]}{|𝐱𝐱^{}|}}\right\}.`$ (29) This is the half-advanced plus half-retarded Green’s function. See , Eq. (6.61). We now explicitly assume that the charge and the field are static, i.e. we assume that $`\omega =C(r,\omega )=0`$ in Eq. (27). Beginning with $`U_0`$ from Eq. (21) we can construct $`{\displaystyle \frac{U_1}{U_0}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{(2M)^2}{(4r_{}^{}{}_{}{}^{2}M^2)(4r^2M^2)}}{\displaystyle \frac{1}{2}}\gamma ,`$ (31) $`{\displaystyle \frac{U_2}{U_0}}`$ $`=`$ $`{\displaystyle \frac{3}{8}}{\displaystyle \frac{(2M)^4}{(4r_{}^{}{}_{}{}^{2}M^2)^2(4r^2M^2)^2}}={\displaystyle \frac{3}{8}}\gamma ,`$ (32) $`{\displaystyle \frac{U_3}{U_0}}`$ $`=`$ $`{\displaystyle \frac{5}{16}}{\displaystyle \frac{(2M)^6}{(4r_{}^{}{}_{}{}^{2}M^2)^3(4r^2M^2)^3}}={\displaystyle \frac{5}{16}}\gamma .`$ (33) Substituting these into Eq. (15), the pattern is immediately clear and the summation is elementary $`U_{elem}(𝐱,𝐱^{})`$ $`=`$ $`{\displaystyle \frac{U_0}{\sqrt{\mathrm{\Gamma }}}}\left(1{\displaystyle \frac{1}{2}}\gamma \mathrm{\Gamma }+{\displaystyle \frac{3}{8}}(\gamma \mathrm{\Gamma })^2{\displaystyle \frac{5}{16}}(\gamma \mathrm{\Gamma })^3+\mathrm{}\right)`$ (35) $`=`$ $`{\displaystyle \frac{U_0}{\sqrt{\mathrm{\Gamma }}}}{\displaystyle \frac{1}{\sqrt{1+\mathrm{\Gamma }\gamma }}}`$ (36) $`=`$ $`{\displaystyle \frac{1}{\sqrt{\mathrm{\Gamma }}}}{\displaystyle \frac{r(4r_{}^{}{}_{}{}^{2}M^2)}{r^{}\sqrt{(4r_{}^{}{}_{}{}^{2}M^2)(4r^2M^2)+4M^2\mathrm{\Gamma }}}}.`$ (37) Equation (37) is a closed-form expression for the Green’s function for Eq. (8) (with C=0). In the limit $`M=0`$ it reduces to $`|𝐱𝐱^{}|^1`$, the Green’s function for Poisson’s equation in flat space. We use Eq. (37) in precisely the same manner we would use the Green’s function for Poisson’s equation: we integrate it against the source to obtain a particular solution to the inhomogeneous equation. In our case the source is the static scalar test-charge held fixed at the spatial position $`𝐛`$, and we have $`V_{part}(𝐱,𝐛)`$ $`=`$ $`{\displaystyle \left[\frac{q}{u^t(b)\sqrt{g(b)}}\left(1+\frac{M}{2b}\right)^4\delta ^3(𝐱^{}𝐛)\right]U_{elem}(𝐱,𝐱^{})d^3𝐱^{}}`$ (39) $`=`$ $`q{\displaystyle \frac{2bM}{2b+M}}{\displaystyle \frac{1}{\sqrt{\mathrm{\Gamma }(𝐱,𝐛)}}}{\displaystyle \frac{4rb}{\sqrt{(4b^2M^2)(4r^2M^2)+4M^2\mathrm{\Gamma }(𝐱,𝐛)}}}`$ (40) $`=`$ $`q\sqrt{{\displaystyle \frac{b_hM}{b_h+M}}}{\displaystyle \frac{1}{\sqrt{r_h^22r_hb_h\mathrm{cos}\theta +b_h^2M^2\mathrm{sin}^2\theta }}}.`$ (41) In the last two steps we have explicitly included the factor $`1/u^t(b)=\sqrt{g_{00}(b)}=(2bM)/(2b+M)=\sqrt{(b_hM)/(b_h+M)}`$; in the last step we have converted to harmonic coordinates . The leading minus sign in Eq. (II) is a consequence of the source term in Eq. (4); our source is ‘$`+`$$`4\pi \rho `$ and not the familiar ‘$``$$`4\pi \rho `$ of electrostatics. (Like scalar charges attract.) As we are now dealing strictly with a static solution, we have also dropped the twiddle denoting the Fourier transform in Eq. (8). Equation (II) is a particular solution to the scalar-static field equation, but is it the desired solution to the equation? In other words, does it satisfy all the boundary conditions? First, the field and its derivatives are well behaved outside — and on — the horizon; thus our solution has no unphysical regions of infinite energy (save, of course, at the location of the charge). For $`rb>M/2`$ we see from Eq. (IIb) $$V_{part}(r\mathrm{})\frac{q}{r}\left(\frac{2bM}{2b+M}\right).$$ (42) As the charge is lowered toward the horizon, the factor $`(2bM)/(2b+M)`$ extinguishes the field measured by a distant observer. Notice if the charge is lowered to the horizon ($`b=M/2`$) the field completely disappears. (See, e.g., for discussion.) Thus, it is the extinction factor — which had its origin in the factor of $`1/u^t(t)`$ in Eq. (6) and has survived throughout the calculation — that enforces no scalar hair on the black hole. The factor $`1/u^t`$ is present even in a special relativity. Notice it is just the inverse of the Lorentz factor $`\gamma =dt/d\tau =1/\sqrt{1v^2}`$. As a mnemonic, the strength of the “charge” depends on the spin of the field: for masses (spin 2) we have $`m=\gamma ^1m_{\mathrm{rest}}`$, for electric charges (spin 1) we have $`e=\gamma ^0e_{\mathrm{rest}}`$, and clearly, by induction, for scalar charges (spin 0) we have $`q=\gamma ^1q_{\mathrm{rest}}`$. Equation (42) is the appropriate asymptotic form of the scalar field; therefore the particular solution Eq. (II) is the desired solution which satisfies all the boundary conditions $$V(𝐱,𝐛)=V_{part}(𝐱,𝐛).$$ (43) Before embarking on the force calculation, we note a remarkable feature of the closed-form solution we have found. Despite the fact that every term in the Hadamard series is divergent at the horizon \[Eq. (II)\], the closed-form expression for the elementary solution is well behaved on the horizon. The easiest way to see this is with the harmonic-coordinate expression Eq. (IIc) evaluated on the horizon ($`r_h=M`$) $$V_{Horizon}=\frac{q}{b_hM\mathrm{cos}\theta }\sqrt{\frac{b_hM}{b_h+M}}.$$ (44) It is also interesting to note that the horizon is not a surface of constant “potential”. In Sec. I our (flawed) intuition about the self-force on a static electric charge was predicated on the horizon acting as conducting surface; therefore we should not be surprised that the force on the static scalar charge will be different. ## III Self-force on a static scalar charge ### A local method We begin our calculation of the self-force with a picturesque description (Gedankenexperiment) of how such a measurement could be made. We imagine a test-charge with bare mass $`\mu _{bare}`$ and scalar charge $`q`$ held fixed by a non-conducting system of mechanical struts outside the horizon of a Schwarzschild black hole. A non-conducting experimenter at the apex of a vertical, ballistic trajectory momentarily comes to rest with respect to the fixed charge. At this moment, she reaches out and measures the force required to hold the charge fixed, i.e. she measures the force needed to just lift the charge off the strut. The spacetime event $``$ where/when the force is measured will be taken as the origin of the free-falling observer’s coordinates $`x^{\overline{\alpha }}`$. (The over-bar denotes coordinates in the local, freely-falling frame of the observer .) Clearly, by symmetry, we can choose our coordinate system such that the particle is located on the $`z`$-axis, and the freely-falling coordinate system is aligned so that the $`\overline{z}`$-axis coincides with the $`z`$-axis. Thus the only component of the force for the freely-falling observer to measure will be $`F^{\overline{z}}`$. Although this is an elaborate scheme to define the force measurement, working in the freely-falling frame where the charge is momentarily at rest is the surest way to establish unambiguously how the scalar field of the particle contributes to the renormalized mass. As no scalar charges have ever been observed, nor scalar fields measured, our assumption that the scalar field does not interact with the experimental apparatus (the struts) and the experimenter seems to be quite plausible. In order to compute the force in the free-falling frame of the observer, we need to be able to convert quantities from the isotropic coordinates of Eq. (II) to the coordinates of the observer. The defining feature of the free-falling frame is that spacetime is locally flat: i.e., $`(g_{\overline{\alpha }\overline{\beta }})_{}=\mathrm{diag}(1,1,1,1)\eta _{\overline{\alpha }\overline{\beta }}`$ (46) $`(g_{\overline{\alpha }\overline{\beta },\overline{\gamma }})_{}=0`$ (47) $`g_{\overline{\alpha }\overline{\beta }}(x^{\overline{\gamma }})=\eta _{\overline{\alpha }\overline{\beta }}+O[(x^{\overline{\gamma }})^2].`$ (48) In particular, the Christoffel symbols, $`\mathrm{\Gamma }_{\overline{\beta }\overline{\gamma }}^{\overline{\alpha }}=(1/2)g^{\overline{\alpha }\overline{\sigma }}[g_{\overline{\sigma }\overline{\gamma },\overline{\beta }}+g_{\overline{\sigma }\overline{\beta },\overline{\gamma }}g_{\overline{\beta }\overline{\gamma },\overline{\sigma }}]O[\overline{x}^{\overline{\sigma }}]`$ near the event $``$ . Fortunately, SW have done the dirty work in finding the transformations from isotropic coordinates (un-barred) to the free-falling coordinates (barred). They are given by $`\overline{t}`$ $`=`$ $`{\displaystyle \frac{1M/2b}{1+M/2b}}t+{\displaystyle \frac{M}{b^2}}{\displaystyle \frac{1}{(1+m/2b)^2}}t(zb)+O[(x^\mu b^\mu )^3]`$ (50) $`x^{\overline{j}}`$ $`=`$ $`(1+M/2b)^2(x^jb^j)+{\displaystyle \frac{M}{2b^2}}{\displaystyle \frac{1M/2b}{(1+M/2b)^5}}\delta ^{\overline{j}\overline{3}}t^2`$ (52) $`{\displaystyle \frac{M}{2b^2}}(1+m/2b)[2(x^jb^j)(zb)\delta ^{\overline{j}\overline{3}}|𝐱𝐛|^2]+O[(x^\mu b^\mu )^3].`$ These can be used to find $`{\displaystyle \frac{1}{\sqrt{\mathrm{\Gamma }(𝐱,𝐛)}}}={\displaystyle \frac{1}{|𝐱𝐛|}}={\displaystyle \frac{1}{|\overline{𝐱}\overline{𝐗}(\overline{t})|}}(1+M/2b)^2\left[1{\displaystyle \frac{M}{2b^2}}{\displaystyle \frac{1}{(1+M/2b)^3}}\overline{z}+O[(x^{\overline{\alpha }})^2]\right],`$ (53) where $`\overline{𝐗}(\overline{t})`$ is the position of the charge as viewed in the freely-falling frame $$X^{\overline{j}}=\frac{1}{2}a_g\delta ^{\overline{3}\overline{j}}\overline{t}^2+O[\overline{t}^3],$$ (54) and $`a_g`$ is the acceleration of the fixed charge as measured in the freely falling frame at the moment the experimenter comes to rest at the apex of her geodesic trajectory $`a_g`$ $`{\displaystyle \frac{M}{b^2}}{\displaystyle \frac{1}{(1+M/2b)^3(1M/2b)}}`$ (56) $`={\displaystyle \frac{M}{b_s^2}}(1{\displaystyle \frac{2M}{b_s}})^{1/2}.`$ In the second line we have converted to Schwarzschild coordinates. Notice this is the same as the acceleration appearing in Eq. (2). We can use Eq. (III A) to evaluate the scalar field and its derivatives in the free-falling frame $`V(\overline{t},\overline{𝐱})={\displaystyle \frac{q}{|\overline{𝐱}\overline{𝐗}(\overline{t})|}}\left[1{\displaystyle \frac{1}{2}}a_g\overline{z}+O[(x^{\overline{\alpha }})^2]\right],`$ (58) $`V(\overline{t}=0,\overline{𝐱})=q\left[{\displaystyle \frac{1}{\overline{r}}}{\displaystyle \frac{a_g}{2}}n^{\overline{3}}+O[\overline{r}\times \mathrm{even}\mathrm{number}\mathrm{of}(n^{\overline{l}}){}_{}{}^{}\mathrm{s}]\right],`$ (59) $`V_{,\overline{t}}(\overline{t}=0,\overline{𝐱})=qO[\overline{r}^0\times \mathrm{odd}\mathrm{number}\mathrm{of}(n^{\overline{l}}){}_{}{}^{}\mathrm{s}],`$ (60) $`V_{,\overline{k}}(\overline{t}=0,\overline{𝐱})=q\left[{\displaystyle \frac{n^{\overline{k}}}{\overline{r}^2}}+{\displaystyle \frac{a_g}{2\overline{r}}}(n^{\overline{k}}n^{\overline{3}}\delta ^{\overline{3}\overline{k}})+n^{\overline{k}}O[\overline{r}^0\times \mathrm{even}\mathrm{number}\mathrm{of}(n^{\overline{l}}){}_{}{}^{}\mathrm{s}]\right],`$ (61) $`V_{,\overline{t}\overline{t}}(\overline{t}=0,\overline{𝐱})=q\left[{\displaystyle \frac{a_g}{\overline{r}^2}}{\displaystyle \frac{a_g^2}{2}}{\displaystyle \frac{n^{\overline{3}}}{\overline{r}}}+O[\overline{r}^1\times \mathrm{even}\mathrm{number}\mathrm{of}(n^{\overline{l}}){}_{}{}^{}\mathrm{s}]\right],`$ (62) $`V_{,\overline{k}\overline{t}}(\overline{t}=0,\overline{𝐱})={\displaystyle \frac{q}{\overline{r}}}\left[n^{\overline{k}}O[\overline{r}^0\times \mathrm{odd}\mathrm{number}\mathrm{of}(n^{\overline{l}}){}_{}{}^{}\mathrm{s}]\right],`$ (63) where $`\overline{r}=|\overline{𝐱}|`$ and $`n^{\overline{k}}=x^{\overline{k}}/\overline{r}`$. The spatial indices can be freely raised and lowered with $`\delta ^{ij}`$. Notice the extinction factor present in Eq. (II) does not extinguish the charge in the freely-falling frame Eq. (III Aa), that is, in the freely-falling frame the dominant behavior of the field is simply $`(\mathrm{charge}/\mathrm{distance})`$, independent of how deep the charge is in the Schwarzschild potential. We will compute the force required to hold the charge fixed by integrating the force density $$f^{\overline{z}}=T_{;\overline{\beta }}^{\overline{z}\overline{\beta }}$$ (64) over the physical extent of the charged body at the instant of time ($`\overline{t}=0`$) when the measurement is made. More precisely, since we are describing the particle as a Dirac $`\delta `$-function, we will integrate over an infinitesimal sphere of radius $`\overline{ϵ}`$ centered on the particle and take the limit as $`\overline{ϵ}0`$. The stress-energy tensor $`T^{\overline{z}\overline{\beta }}`$ will have contributions from the bare mass of the particle and the scalar field; thus we have $`F_{(\mathrm{strut})}^{\overline{z}}`$ $`=`$ $`\underset{\overline{ϵ}0}{lim}{\displaystyle _{\overline{r}\overline{ϵ}}}[f^{\overline{z}}]_{\overline{t}=0}d^3\overline{x}=\underset{\overline{ϵ}0}{lim}{\displaystyle _{\overline{r}\overline{ϵ}}}\left[T_{;\overline{\beta }}^{\overline{z}\overline{\beta }}\right]_{\overline{t}=0}d^3\overline{x}`$ (66) $`=`$ $`\underset{\overline{ϵ}0}{lim}{\displaystyle _{\overline{r}\overline{ϵ}}}\left[T_{(\mathrm{bare});\overline{\beta }}^{\overline{z}\overline{\beta }}\right]_{\overline{t}=0}d^3\overline{x}+\underset{\overline{ϵ}0}{lim}{\displaystyle _{\overline{r}\overline{ϵ}}}\left[T_{(\mathrm{SF});\overline{\beta }}^{\overline{z}\overline{\beta }}\right]_{\overline{t}=0}d^3\overline{x}.`$ (67) The first term in the Eq. (67) shows that the strut must support the bare weight of the particle. This term is present whether or not the particle is charged. Using the stress-energy tensor for the bare mass of a point particle located at $`𝐛=b\widehat{𝐳}`$ $$T_{(\mathrm{bare})}^{\alpha \beta }=\mu _{\mathrm{bare}}\frac{\dot{b}^\alpha (b)\dot{b}^\beta (b)}{\sqrt{g(b)}u^t(b)}\delta ^3(𝐱𝐛),$$ (68) SW found the necessary force the strut must supply to support the bare weight of the particle to be $$F_{(\mathrm{bare})}^{\overline{z}}=\underset{\overline{ϵ}0}{lim}_{\overline{r}\overline{ϵ}}\left[T_{(\mathrm{bare});\overline{\beta }}^{\overline{z}\overline{\beta }}\right]_{\overline{t}=0}d^3\overline{x}=\frac{M\mu _{\mathrm{bare}}}{b^2}\frac{1}{(1+M/2b)^3}\frac{1}{1M/2b}=\mu _{\mathrm{bare}}a_g.$$ (69) The second term in Eq. (67) involves the stress-energy tensor of the scalar field. In evaluating this integral we make use of Eq. (III A) and note that the connection coefficients in this frame are $`O[x^{\overline{\alpha }}]`$; thus we can write $`F_{(\mathrm{SF})}^{\overline{z}}`$ $`=\underset{\overline{ϵ}0}{lim}{\displaystyle _{\overline{r}\overline{ϵ}}}\left[T_{(\mathrm{SF});\overline{\beta }}^{\overline{z}\overline{\beta }}\right]_{\overline{t}=0}d^3\overline{x}`$ (71) $`=\underset{\overline{ϵ}0}{lim}\left\{{\displaystyle _{\overline{r}\overline{ϵ}}}\left[T_{(\mathrm{SF}),\overline{k}}^{\overline{z}\overline{k}}+T_{(\mathrm{SF}),\overline{t}}^{\overline{z}\overline{t}}+O[x^{\overline{\alpha }}T^{\overline{\beta }\overline{\gamma }}]\right]_{\overline{t}=0}d^3\overline{x}\right\}.`$ We will denote the three contributions to Eq. (71) as $`F_{(\mathrm{SF1})}^{\overline{z}}`$, $`F_{(\mathrm{SF2})}^{\overline{z}}`$ and $`F_{(\mathrm{SF3})}^{\overline{z}}`$ respectively. We also make use of Eq. (III A) in writing the stress-energy tensor for the scalar field $`T^{\overline{\beta }\overline{\gamma }}`$ $``$ $`{\displaystyle \frac{1}{4\pi }}\left[g^{\overline{\alpha }\overline{\sigma }}g^{\overline{\beta }\overline{\tau }}V_{,\overline{\sigma }}V_{,\overline{\tau }}{\displaystyle \frac{1}{2}}g^{\overline{\alpha }\overline{\beta }}g^{\overline{\sigma }\overline{\tau }}V_{,\overline{\sigma }}V_{,\overline{\tau }}\right]`$ (73) $`=`$ $`{\displaystyle \frac{1}{4\pi }}\left[\eta ^{\overline{\alpha }\overline{\sigma }}\eta ^{\overline{\beta }\overline{\tau }}V_{,\overline{\sigma }}V_{,\overline{\tau }}{\displaystyle \frac{1}{2}}\eta ^{\overline{\alpha }\overline{\beta }}\eta ^{\overline{\sigma }\overline{\tau }}V_{,\overline{\sigma }}V_{,\overline{\tau }}+O[(x^{\overline{\alpha }})^2V_{,\overline{\sigma }}V_{,\overline{\tau }}]\right].`$ (74) We now substitute Eq. (74) into Eq. (71) and treat the terms in reverse order. The third term in Eq. (71) gives a contribution of the form $`F_{(\mathrm{SF3})}^{\overline{z}}`$ $`={\displaystyle \frac{1}{4\pi }}\underset{\overline{ϵ}0}{lim}{\displaystyle _{\overline{r}\overline{ϵ}}}\left[O[x^{\overline{\alpha }}T_{(\mathrm{SF})}^{\overline{z}\overline{t}}]\right]_{\overline{t}=0}d^3\overline{x}`$ (76) $`={\displaystyle \frac{1}{4\pi }}\underset{\overline{ϵ}0}{lim}{\displaystyle _{\overline{r}\overline{ϵ}}}\left[O[x^{\overline{\alpha }}V_{,\overline{\beta }}V_{,\overline{\gamma }}]+O[(x^{\overline{\alpha }})^3V_{,\overline{\beta }}V_{,\overline{\gamma }}]\right]_{\overline{t}=0}d^3\overline{x}.`$ Using Eq. (III A), we see the most singular terms come from $`(V_{,\overline{k}}V_{,\overline{l}})`$; therefore we have $`F_{(\mathrm{SF3})}^{\overline{z}}\underset{\overline{ϵ}0}{lim}{\displaystyle _{\overline{r}\overline{ϵ}}}[`$ $`{\displaystyle \frac{1}{\overline{r}}}\times [\mathrm{term}\mathrm{with}\mathrm{odd}\mathrm{number}\mathrm{of}(n^{\overline{l}}){}_{}{}^{}\mathrm{s}]`$ (78) $`+\overline{r}^0\times [\mathrm{term}\mathrm{with}\mathrm{even}\mathrm{number}\mathrm{of}(n^{\overline{l}}){}_{}{}^{}\mathrm{s}]]d\overline{r}d\overline{\mathrm{\Omega }}.`$ The first term contains an odd number of unit vectors, and therefore will vanish when we integrate over the solid angle. The second term vanishes as $`\overline{ϵ}0`$, and similarly for higher powers of $`\overline{r}`$. These two tricks are used repeatedly in evaluating the remaining integrals in Eq. (71). Thus we have $$F_{(\mathrm{SF3})}=0.$$ (79) We now evaluate the second term in Eq. (71) using Eq. (III A) and ruthlessly discarding terms that do not survive the limit or the angular integration: $`F_{(\mathrm{SF2})}^{\overline{z}}`$ $`=\underset{\overline{ϵ}0}{lim}{\displaystyle \frac{1}{4\pi }}{\displaystyle _{\overline{r}\overline{ϵ}}}\left[V_{,\overline{z}}V_{,\overline{z}\overline{t}}\right]\overline{r}^2𝑑\overline{r}𝑑\overline{\mathrm{\Omega }}`$ (83) $`={\displaystyle \frac{1}{4\pi }}q^2a_g\underset{\overline{ϵ}0}{lim}{\displaystyle _0^{\overline{ϵ}}}{\displaystyle \frac{1}{\overline{r}^2}}𝑑\overline{r}{\displaystyle n^{\overline{z}}n^{\overline{z}}𝑑\overline{\mathrm{\Omega }}}`$ $`={\displaystyle \frac{1}{3}}q^2a_g\underset{\overline{ϵ}0}{lim}{\displaystyle _0^{\overline{ϵ}}}{\displaystyle \frac{1}{\overline{r}^2}}𝑑\overline{r}.`$ Using the divergence theorem and Eq. (74), the first term in Eq. (71) can be written $`F_{(\mathrm{SF1})}^{\overline{z}}`$ $`={\displaystyle \frac{1}{4\pi }}\underset{\overline{ϵ}0}{lim}{\displaystyle _{\overline{r}\overline{ϵ}}}\left[T_{(\mathrm{SF}),\overline{k}}^{\overline{z}\overline{k}}\right]_{\overline{t}=0}d^3\overline{x}=\underset{\overline{ϵ}0}{lim}{\displaystyle _{\overline{r}=\overline{ϵ}}}\left[T_{(\mathrm{SF})}^{\overline{z}\overline{k}}\right]_{\overline{t}=0}n^{\overline{k}}\overline{r}^2𝑑\overline{\mathrm{\Omega }}`$ (85) $`={\displaystyle \frac{1}{4\pi }}\underset{\overline{ϵ}0}{lim}{\displaystyle _{\overline{r}=\overline{ϵ}}}\left[V_{,\overline{z}}V_{,\overline{k}}{\displaystyle \frac{1}{2}}\delta ^{\overline{z}\overline{k}}(V_{,\overline{t}}V_{,\overline{t}}+\delta ^{\overline{l}\overline{m}}V_{,\overline{l}}V_{,\overline{m}})+O[(x^{\overline{\alpha }})^2V_{,\overline{\beta }}V_{,\overline{\gamma }}]\right]_{\overline{t}=0}\overline{r}^2n^{\overline{k}}𝑑\overline{\mathrm{\Omega }}.`$ All the terms except the first vanish either by $`\mathrm{}𝑑\overline{\mathrm{\Omega }}=0`$, or $`lim_{\overline{ϵ}0}\mathrm{}=0`$. Using Eq. (III A) and integrating the first term in Eq. (85) over the solid angle, we are left with $$F_{(\mathrm{SF1})}^{\overline{z}}=\frac{1}{3}a_g\underset{\overline{ϵ}0}{lim}\frac{q^2}{\overline{ϵ}}.$$ (86) Combining the results in Eqs. (67), (69), (79), (83) and (86) we have $`F_{(\mathrm{strut})}^{\overline{z}}`$ $`=\left[\mu _{\mathrm{bare}}+{\displaystyle \frac{q^2}{3}}\underset{\overline{ϵ}0}{lim}\left({\displaystyle \frac{1}{\overline{ϵ}}}+{\displaystyle _0^{\overline{ϵ}}}{\displaystyle \frac{1}{\overline{r}^2}}𝑑\overline{r}\right)\right]a_g`$ (90) $`=\left[\mu _{\mathrm{bare}}+{\displaystyle \frac{q^2}{3}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{1}{\overline{r}^2}}𝑑\overline{r}\right]a_g`$ $`=\mu _{\mathrm{ren}}a_g,`$ where we have defined the leading factor in the bracket as the “renormalized” mass $`\mu _{ren}`$. The renormalized mass has the same functional form as the renormalized mass for the electric charge in the SW calculation. Converting to Schwarzschild coordinates , we get $`F_{(\mathrm{strut})}={\displaystyle \frac{M\mu _{\mathrm{ren}}}{b_s^2}}(1{\displaystyle \frac{2M}{b_s}})^{1/2}+\left\{\mathrm{nothing}\mathrm{depending}\mathrm{on}q\right\}.`$ (91) There is no self-force of the form seen in the second term of Eq. (2). ### B global method We verify our no-self-force result by means of a global, energy-conservation calculation. Suppose, instead of measuring the force on the charge while the charge is in place (as we did in the last subsection), the free-falling observer lowers the charge a small amount $`\delta \overline{b}`$. The work done on the experimenter will be $$\delta \overline{W}=F^{\overline{z}}\delta \overline{b}=F^{\overline{z}}(1+M/2b)^2\delta b,$$ (92) where we have used Eq. (7) to convert the free-falling displacement $`\delta \overline{b}`$ to an isotropic coordinate displacement $`\delta b`$. The experimenter then converts this energy into a photon and fires the photon to asymptotic infinity. The energy received at infinity will be red-shifted $$\delta E_{\mathrm{received}}=\sqrt{g_{00}(b)}\delta \overline{W}.$$ (93) By conservation of energy this change in the system will be manifested by a change in asymptotic mass $``$ of the system $`\delta `$ $`=\delta E_{\mathrm{received}}=[1(M/2b)^2]F^{\overline{z}}\delta b.`$ (94) Thus we have $$F^{\overline{z}}=\frac{1}{1(M/2b)^2}\frac{\delta }{\delta b}.$$ (95) We now use the total mass variation law of Carter , which shows how the asymptotic mass will differ between two situations where the gravitational and matter status of the spacetime is slightly altered. In our case we compute the difference in asymptotic mass before and after we make the small displacement of the charge. The relationship is given by $$\delta \frac{\kappa }{8\pi }\delta 𝒜=\frac{1}{8\pi }\delta G_{\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}}^0\sqrt{g}d^3x+\frac{1}{16\pi }G^{\mu \nu }h_{\mu \nu }\sqrt{g}d^3x.$$ (96) Here $`\kappa `$ is the surface gravity of the black hole, $`𝒜`$ is the area of black hole, $`G^{\mu \nu }`$ is the Einstein tensor, and $`h_{\mu \nu }`$ is the difference in the metric between the two configurations. In Eq. (96) we have neglected terms involving the spin of the black hole. Throughout this paper we have assumed that the metric is unperturbed by the presence of the charge; therefore the last term in Eq. (96) vanishes. For a Schwarzschild black hole the area term in Eq. (96) is just the change in the mass of the black hole. During our slow displacement we will assume that no matter or radiation goes down the hole, and therefore this term will vanish. (We revisit this point at the end of the section.) Using Einstein’s equation to write $`G_{\mathrm{\hspace{0.33em}\hspace{0.33em}0}}^0=8\pi T_{\mathrm{\hspace{0.33em}\hspace{0.33em}0}}^0`$ all that remains of Eq. (96) is $`\delta `$ $`=\delta {\displaystyle T_{\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}}^0\sqrt{g}d^3x}`$ (98) $`=\delta {\displaystyle T_{(\mathrm{bare})0}^0\sqrt{g}d^3x}+\delta {\displaystyle T_{(\mathrm{SF})0}^0\sqrt{g}d^3x}.`$ The first integral can be computed by Eq. (68). SW give the result $$_{\mathrm{bare}}T_{(\mathrm{bare})0}^0\sqrt{g}d^3x=\mu _{\mathrm{bare}}\left(\frac{1M/2b}{1+M/2b}\right).$$ (99) Noting that the metric is diagonal, the scalar field is static, and employing the definition of the stress tensor for a scalar field Eq. (73), we can write the second integral in Eq. (98) as $`_{\mathrm{SF}}`$ $`{\displaystyle T_{(\mathrm{SF})0}^0\sqrt{g}d^3x}={\displaystyle \frac{1}{8\pi }}{\displaystyle \left[g^{jk}V_{,j}V_{,k}\right]\sqrt{g}d^3x}`$ (101) $`={\displaystyle \frac{1}{8\pi }}{\displaystyle \left[(g^{jk}V_{,j}V\sqrt{g})_{,k}V(g^{jk}V_{,j}\sqrt{g})_{,k}\right]d^3x}.`$ The first term in Eq. (101) can be converted to two surface integrals: one over the horizon, the other over a sphere at $`r\mathrm{}`$. The field and its derivatives are well behaved on the horizon, but $`g^{jk}\sqrt{g}`$ vanishes there; therefore the integral on the horizon vanishes. The other surface integral vanishes in the limit $`r\mathrm{}`$. Using the original field equation Eq. (4), we have $$(\sqrt{g}g^{jk}V_{,j})_{,k}=4\pi \sqrt{g}\rho .$$ (102) Thus we can write $$_{\mathrm{SF}}=\frac{1}{2}\rho V\sqrt{g}d^3x=\frac{q}{2u^t(b)}V\delta ^3(𝐱𝐛)d^3x.$$ (103) Unfortunately, the scalar field is divergent at the source point b; therefore we must renormalize. We do so by modeling our source as a charged, spherical shell of radius $`ϵ`$, i.e. $$\frac{\delta ^3(𝐱𝐛)}{\sqrt{g}}\underset{ϵ0}{lim}\frac{\delta ^1(|𝐱𝐛|ϵ)}{4\pi ϵ^2}.$$ (104) Performing the integration and using the metric to convert the radius of the ball to free-falling coordinates $`\overline{ϵ}=(1+M/2b)^2ϵ`$, we have $$_{\mathrm{SF}}=\frac{1}{2}\left(\frac{1M/2b}{1+M/2b}\right)\underset{\overline{ϵ}0}{lim}\frac{q^2}{\overline{ϵ}}.$$ (105) As expected the functional form is the same here as in Eq. (99). Combining Eq. (99) and Eq. (105) defining the renormalized mass $$\overline{\mu }_{\mathrm{ren}}=\mu _{\mathrm{bare}}+\frac{1}{2}\underset{\overline{ϵ}0}{lim}\frac{q^2}{\overline{ϵ}},$$ (106) and using Eq. (95) and Eq. (56), we have $$F^{\overline{z}}=\overline{\mu }_{\mathrm{ren}}a_g.$$ (107) This agrees exactly with our previous calculation Eq. (90): no finite part of the self-force. We close this section with a pedagogical comment on the global energy conservation method for computing the force. We note that the calculation was predicated on the assumption that the area term in Eq. (96) vanished, i.e., $$\frac{\kappa }{8\pi }\delta 𝒜=\delta (\mathrm{mass}\mathrm{of}\mathrm{the}\mathrm{hole})=0.$$ (108) In terms of modern black hole theory, this is a valid assumption: No particles were dropped into the hole. The charge was displaced slowly, so no transverse fields (i.e. radiation) heated the horizon. Therefore, the area remains unchanged. However, in Sec. III A we gave a primitive derivation of the force which did not appear to explicitly rely on any sophisticated properties of black holes. (Implicitly, we did assume that the mass of the hole remained constant when the observer wiggled the charge to measure the force.) An interesting interpretation of the two force calculations is to accept the primitive derivation in Sec. III A as the correct force. Then, when we evaluate the right hand side of Eq. (98) and show that it is the same as our local (primitive) force calculation, we have verified that the area of the black hole did not change when we lowered the charge. This tells us, in spite of the fact that energy of the scalar-static field (or the electric field for an electric charge) extends clear down to the horizon, that none of the energy near the horizon is pushed across the horizon when we lower the charge. ## IV Conclusions and Discussion We have developed a formalism for constructing the Hadamard elementary solution for the Schwarzschild-Helmholtz equation (8). This formalism was chosen because it expresses the singular nature of field in the near proximity of the charge. In the case of a static charge, we are able to find a closed-form expression for the field. We have used this expression to show (after mass renormalization) there is no self-force on a static scalar charge outside a Schwarzschild black hole Although we have patterned our discussion after Copson and SW , there are alternative methods for computing the scalar field and the self-force. For example, in the static limit, Linet has used generalized axially symmetric potential theory (GASP) to derive the Green’s function for the scalar field. However, this type of construction has not been extended to the Helmholtz-type equation depicted in Eq. (8). Lohiya has demonstrated a concise method for determining the force on a static electric charge. Lohiya’s method also uses the Copson-Linet closed-form expression for the electrostatic potential; however it is unclear how to extend Lohiya’s method to moving particles. The formalism we have developed can be extended to moving charges. The recursion relation, Eq. (27), can be integrated with $`\omega 0`$. Although it may be hard to find a simple summation of the results as in Eq. (36), it is possible to obtain the Green’s function to the first several orders in $`\mathrm{\Gamma }`$. Now that we have computed the (absence of) forces acting on a static scalar charge, let us assemble what is known about all the forces on a static charge outside a Schwarzschild black hole. In order to express this, let us slightly change the thought experiment. We will give the test-charge a mass $`\mu `$, an electric charge $`e`$, and a scalar charge $`q`$. We will support the charge on a strut as before, but, instead of measuring the force supplied by the strut at some moment $`\overline{t}=0`$, we kick the strut out from under the charge and find the instantaneous acceleration of the falling charge in harmonic coordinates. After the particle begins to move there will also be a radiation-reaction force, so we must make the measurement at the moment we remove the strut. The metric can be used to convert quantities from free-falling (proper) coordinates to harmonic coordinates . The result is $$\left[\mu \frac{d^2r_h}{dt_h^2}\right]_{\overline{t}=0}=\left(\frac{r_hm}{r_h+m}\right)^{3/2}F_{(\mathrm{strut})}^{\overline{z}}.$$ (109) Using this to convert Eqs. (2) and (91), and expanding in the post-Newtonian quantity $`M/r_h`$, we have $`\left[\mu {\displaystyle \frac{d^2r_h}{dt_h^2}}\right]_{\overline{t}=0}={\displaystyle \frac{M}{r_h^2}}\{`$ $`\mu \left[1+4\left({\displaystyle \frac{M}{r_h}}\right)9\left({\displaystyle \frac{M}{r_h}}\right)^2+16\left({\displaystyle \frac{M}{r_h}}\right)^3+\mathrm{known}\mathrm{terms}\right]`$ (115) $`+{\displaystyle \frac{e^2}{M}}\left[\right({\displaystyle \frac{M}{r_h}})6\left({\displaystyle \frac{M}{r_h}}\right)^2+{\displaystyle \frac{39}{2}}\left({\displaystyle \frac{M}{r_h}}\right)^3+\mathrm{known}\mathrm{terms}]`$ $`+{\displaystyle \frac{q^2}{M}}\left[\mathrm{all}\mathrm{terms}\mathrm{are}\mathrm{known}\mathrm{to}\mathrm{be}\mathrm{zero}\right]`$ $`+{\displaystyle \frac{\mu ^2}{M}}\left[2\right({\displaystyle \frac{M}{r_h}}){\displaystyle \frac{87}{4}}\left({\displaystyle \frac{M}{r_h}}\right)^2+\mathrm{unknown}\mathrm{terms}]`$ $`+O[e^2\mu ]+O[q^2\mu ]+O[\mu ^3]\}.`$ Here $`r_h`$ denotes the radial position of the particle in harmonic coordinates. In line (IVa) we have recovered the velocity independent terms of the geodesic equation of motion expressed in harmonic coordinates. Since this is just a Taylor expansion of the first term in Eq. (2), we know these terms to all orders in $`M/r_h`$. Thankfully, the first three terms are in agreement with the second post-Newtonian equations of motions. (See e.g. .) Line (IVb) is just the expansion of the second term in Eq. (2), and thus we know these terms to all orders in $`M/r_h`$. Line (IVc) is the scalar-charge part, which we have shown to vanish for all orders in $`M/r_h`$. For moving charges, there will very likely be non-zero terms. In line (IVd), only the first two terms are known from second post-Newtonian calculations. (See e.g. .) The question remains, can the unknown terms in line (IVd) be obtained by methods similar to those used to find the electric and scalar forces, that is, by looking at the field (metric) perturbations produced by the mass of the test-particle? Obviously there are a number of conceptual issues to tackle in answering this question. For example, when the metric itself is the perturbed field, can we define a freely falling observer in the same way as we did in the scalar-charge force calculation? When solving for the metric perturbation, how do the stresses in the strut affect the solution? This is currently under vigorous investigation . Line (IVe) represents higher order effects, such as additional forces on the particle due to the change in the metric produced by the electric field of the particle. Such terms will also arise from a gauge change, say, $`rr+O[\mu ]`$. These terms are clearly second order in the perturbations. ###### Acknowledgements. It is a pleasure to thank Patrick Brady, Ted Quinn and Bob Wald for many useful conversations. This work was supported in part by NASA Grant NAGW-4268 and NSF Grant AST-9417371 while the author was at Caltech, and by NSF Grant PHY95-14726 while the author was at the University of Chicago. ## A Recap of electrostatic charge in Schwarzschild spacetime In this section we summarize the results of Copson and obtain the closed-form solution for the electrostatic potential using our Hadamard construction. Using isotropic coordinates and assuming the field is strictly static, Maxwell’s equations for the electrostatic potential can be written $$C^{ij}A_{0,ij}(𝐱)+C^jA_{0,j}(𝐱)=16\pi e\frac{b^2(2aM)}{(2b+M)^3}\delta ^3(𝐱𝐛),$$ (A1) where $$C^j=h(r)\frac{x^j}{r}=\frac{d}{dr}\left\{\mathrm{ln}\right[\frac{(1+M/2r)^3}{1M/2r}]\}\frac{x^j}{r}.$$ (A2) Equation (A1) is in the same form as Eq. (8); therefore we proceed using Eq. (17) and we get $$U_0^{(\mathrm{el})}=\frac{r}{r^{}}\left(\frac{2r^{}+M}{2r+M}\right)^{3/2}\left(\frac{2rM}{2r^{}M}\right)^{1/2}.$$ (A3) The superscript $`(\mathrm{el})`$ denotes that these are parts of the electrostatic solution. Using the recursion relation Eq. (27), we have $`U_1^{(\mathrm{el})}(r,r^{})`$ $`={\displaystyle \frac{3}{2}}U_0^{(\mathrm{el})}\gamma ,`$ (A5) $`U_2^{(\mathrm{el})}(r,r^{})`$ $`={\displaystyle \frac{5}{8}}U_0^{(\mathrm{el})}\gamma ^2,`$ (A6) $`U_3^{(\mathrm{el})}(r,r^{})`$ $`={\displaystyle \frac{7}{16}}U_0^{(\mathrm{el})}\gamma ^3,`$ (A7) $`U_4^{(\mathrm{el})}(r,r^{})`$ $`={\displaystyle \frac{45}{128}}U_0^{(\mathrm{el})}\gamma ^4,`$ (A8) where $`\gamma `$ is the same as in Eq. (IIa). Once again the summation is elementary: $$U_{\mathrm{elem}}^{(\mathrm{el})}(r,r^{})=\frac{U_0^{(\mathrm{el})}}{\sqrt{\mathrm{\Gamma }}}\frac{(1+2\gamma \mathrm{\Gamma })}{\sqrt{1+\mathrm{\Gamma }\gamma }}.$$ (A9) As in the scalar case, we integrate the elementary solution against the source and obtain a particular solution to electrostatic field Eq. (A1) $$A_0^{\mathrm{part}}(𝐱,𝐛)=\frac{e}{(b_h+M)(r_h+M)}\frac{b_hr_hM^2\mathrm{cos}\theta }{\sqrt{r_h^22r_hb_h\mathrm{cos}\theta +b_h^2M^2\mathrm{sin}^2\theta }},$$ (A10) where we have switched to harmonic coordinates . This is Copson’s 1928 solution “for the potential of an electron in the Schwarzschild field”. As in the scalar case we must ask: does the particular solution satisfy the boundary conditions? It is well behaved at the horizon, so there is no problem there. However, as $`r_h\mathrm{}`$ the potential does not give the correct value $$A_0^{\mathrm{part}}(r_h\mathrm{})\frac{e}{r_h}\frac{b_h}{b_h+M}\frac{e}{r_h}.$$ (A11) The fact that the field does not behave as $`e/r_h`$ for large $`r_h`$ suggests by Gauss’s law that we have found a solution with some additional charge lying around. However, our solution satisfies the homogeneous (source-free) equation everywhere outside the horizon except at the source point where there is a charge $`e`$. We must conclude that we found a solution with some charge on the horizon. In order to fix the boundary condition, we need to add a monopolar solution of the homogeneous equation, that is, we need to add an image charge. It is easy to see what is needed, $$A_0^{\mathrm{homog}}=\frac{eM}{(r_h+M)(b_h+M)},$$ (A12) and check that this satisfies the homogeneous equation outside the horizon. Linet noticed the discrepancy in Eq. (A11) and added this piece to Copson’s result. Combining the two pieces gives the final result: the electrostatic potential for a fixed charge outside a Schwarzschild black hole $$A_0(𝐱,𝐛)=\frac{e}{(b_h+M)(r_h+M)}\left\{\frac{b_hr_hM^2\mathrm{cos}\theta }{\sqrt{r_h^22r_hb_h\mathrm{cos}\theta +b_h^2M^2\mathrm{sin}^2\theta }}+M\right\}.$$ (A13) The potential $`A_0^{\mathrm{part}}`$ in Eq. (A10) is the particular solution constructed directly from the Hadamard elementary solution. If the Hadamard potential $`A_0^{\mathrm{part}}`$ is taken to be the actual potential and the force calculation is carried out (ı.e., repeat the SW calculation using the method similar to Sec. III A), the resulting force is zero. This means that although our construction of the Hadamard solution was strictly a local calculation, when we summed the series we found a solution with just enough charge on the horizon to cancel the repulsive force in Eq. (2). This also means that the repulsive force that SW found for the electric charge is due solely to the part of the potential $`A_0^{\mathrm{homog}}`$ which is tacked on to satisfy the boundary conditions. In other words, the second term in Eq. (2) is simply the force produced by the image charge on the horizon, and the force can be computed from Eq. (A12) directly $`F_{\mathrm{self}}`$ $`=e\left[{\displaystyle \frac{d}{dr_h}}A_0^{\mathrm{homog}}(r_h,b_h)\right]_{r_h=b_h}`$ (A15) $`={\displaystyle \frac{e^2M}{(b_h+M)^3}}={\displaystyle \frac{e^2M}{b_s^3}}.`$ This gives a physical interpretation of the repulsive force; the charge outside the black hole is repelled by an image charge inside the horizon. ## B Comparing closed-form solutions with series solutions Equating our closed-form solutions for the scalar-static and electrostatic fields with the conventional infinite series solutions, we can obtain some interesting summation formulas that do not appear in the standard references . For a fixed point source Eq. (4) is easily solved by separation of variables. The angular dependence is expressed by Legendre polynomials, and the resulting radial equation is also Legendre’s equation. Equating the series solution to our close-form solution Eq. (IIc) we have $`V(𝐱_𝐡,`$ $`𝐛_𝐡)=q\sqrt{{\displaystyle \frac{b_hM}{b_h+M}}}{\displaystyle \frac{1}{\sqrt{r_h^22r_hb_h\mathrm{cos}\theta +b_h^2M^2\mathrm{sin}^2\theta }}}`$ (B5) $`={\displaystyle \frac{q}{M}}\sqrt{{\displaystyle \frac{b_hM}{b_h+M}}}{\displaystyle \underset{l=0}{\overset{\mathrm{}}{}}}(2l+1)P_l(\mathrm{cos}\theta )\left\{\begin{array}{cc}P_l(r_h/M)Q_l(b_h/M),\hfill & \mathrm{if}\mathrm{r}_\mathrm{h}<\mathrm{b}_\mathrm{h}\hfill \\ P_l(b_h/M)Q_l(r_h/M),\hfill & \mathrm{if}\mathrm{r}_\mathrm{h}>\mathrm{b}_\mathrm{h}\hfill \end{array}\right\}.`$ Here the $`P_l`$ and $`Q_l`$ are the Legendre functions. This summation is a special case of Equation 28 in MacRobert . If the field point is located on the horizon ($`M=r_h<b`$) or on the axis ($`\theta =0`$) we can verify this formula using the standard summation formula $$\underset{n=0}{\overset{\mathrm{}}{}}(2n+1)Q_n(x)P_n(y)=\frac{1}{xy}|x|>1\mathrm{and}|\mathrm{x}|>|\mathrm{y}|.$$ (B6) Applying this summation formula to Eq. (B), we get $$V(\mathrm{horizon})=\frac{\mathrm{q}}{\mathrm{b}_\mathrm{h}\mathrm{M}\mathrm{cos}\theta }\sqrt{\frac{\mathrm{b}_\mathrm{h}\mathrm{M}}{\mathrm{b}_\mathrm{h}+\mathrm{M}}}.$$ (B7) Clearly the horizon is not a surface of constant “potential”. This is in contrast with the electrostatic case where the horizon is a surface of constant potential. \[See Eq. (A13).\] On the axis of symmetry (i.e., $`\theta =0`$, so $`P_l(cos\theta )=1`$ for all values of $`l`$) we can also use Eq. (B6) to sum Eq. (B) $$V(\mathrm{axis})=\frac{\mathrm{q}}{|\mathrm{r}_\mathrm{h}\mathrm{b}_\mathrm{h}|}\sqrt{\frac{\mathrm{b}_\mathrm{h}\mathrm{M}}{\mathrm{b}_\mathrm{h}+\mathrm{M}}}.$$ (B8) In the electrostatic case, the series solution can be similarly obtained by separation of variables. The radial functions are derivatives of Legendre functions. Equating the analytic expression with the series solution gives an interesting summation formula $`A_0(𝐱,𝐛)={\displaystyle \frac{e}{(b_h+M)(r_h+M)}}\left\{{\displaystyle \frac{b_hr_hM^2\mathrm{cos}\theta }{\sqrt{r_h^22r_hb_h\mathrm{cos}\theta +b_h^2M^2\mathrm{sin}^2\theta }}}+M\right\}`$ (B10) $`={\displaystyle \frac{e}{M^3}}{\displaystyle \underset{l=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{2l+1}{l(l+1)}}P_l(\mathrm{cos}\theta )\left\{\begin{array}{cc}(r_hM)(b_hM)P_l^{}(r_h/M)Q_l^{}(b_h/M),\hfill & \mathrm{if}\mathrm{r}_\mathrm{h}<\mathrm{b}_\mathrm{h}\hfill \\ (r_hM)(b_hM)P_l^{}(b_h/M)Q_l^{}(r_h/M),\hfill & \mathrm{if}\mathrm{r}_\mathrm{h}>\mathrm{b}_\mathrm{h}\hfill \end{array}\right\}.`$ (B13) Here it is understood when $`l=0`$ we make the replacement $$\frac{(r_hM)}{l}P_l^{}(r_h/M)M(\mathrm{when}l=0).$$ (B15) Notice that the horizon ($`r_h=M`$) is a surface of constant potential. The closed-form expression \[modulo the homogeneous piece in Eq. (A12)\] was computed in 1928 by Copson , who expanded the result in terms of radial functions and discovered the summation formula. The series result was rederived by Cohen and Wald , and Hanni and Ruffini . The asymptotic form of the solution can be seen immediately from the closed-form result, or from the series by noting that $`Q_0(x\mathrm{})1/x`$. We see that $$A_0(r\mathrm{})\frac{e}{r},$$ (B16) which is the correct behavior. Equation (B) can also be obtained from Eq. (B) by differentiating and using Legendre’s equation.
warning/0001/nucl-th0001051.html
ar5iv
text
# Mass modification of 𝐷-meson at finite density in QCD sum rule ## I Introduction Changes of hadron properties in the nuclear medium have recently attracted great interests in theoretical studies. They also have induced the on-going experiments and forthcoming experimental plans in heavy nuclei in GSI and in relativistic heavy-ion collisions in SPS at CERN and AGS, RHIC at BNL. In particular the spectral changes of vector mesons are expected to be a possible signal to investigate such medium effect, because their leptonic decay in the medium can supply us information of the medium without disturbance of the strong interaction. Motivated by this experimental advantage, the in-medium effect of the light vector-mesons ($`\rho ^0,\omega `$ and $`\varphi `$) has been mainly studied in effective hadronic models and QCD sum rules (QSR’s). Recently we applied the QSR method to low-lying heavy-quarkonium ($`J/\psi `$) in order to investigate its mass shift in nuclear matter and the $`J/\psi `$-nucleon interaction at low energy. It was shown that the $`J/\psi `$ mass at normal matter density drops by $`0.10.2\%`$ of its vacuum value. This negative mass-shift of $`J/\psi `$ corresponds to a small decrease of the mass, $`47`$ MeV. This is much smaller than the similar effect in the light mesons such as $`\rho ^0`$ and $`\omega `$ . This difference can be understood as follows: In QSR the magnitude of the spectral change is related to the in-medium change of quark and gluon condensates . In the linear density approximation, the quark operators give a dominant contribution for $`\rho ^0`$ and $`\omega `$, and have significant change in the matter. On the other hand, the dominant in-medium gluon condensates at the normal matter density for $`J/\psi `$ are only $`510\%`$ smaller than its vacuum value. In this paper, we generalize the above calculations and apply the QSR analysis in Ref.6 to light and heavy quark systems with unequal mass. We focus on in-medium properties of pseudoscalar mesons $`D`$’s. Studying the mass modification of the $`D`$-mesons in the nuclear medium is important by the following physical reasons: 1. A part of total cross section of the $`D`$-meson production attributes a conversion of the $`c\overline{c}`$-states such as $`J/\psi `$ and $`\psi ^{}`$ into open-charm pairs. The component is produced through the reaction, $`\psi +ND+\overline{D}`$, in high energy proton-nucleus collisions and relativistic high energy heavy-ion collisions. It contains a heavy charm-quark and a light quark. The existence of a light quark in the $`D`$-meson causes much difference of the properties in the nuclear medium between the $`D`$-meson and $`J/\psi `$. The latter is dominated by the gluon condensates as discussed above, while the former has a large contribution from the light-quark condensates, multiplied by a charm-quark mass. Furthermore $`J/\psi `$ predominantly interacts with light hadrons solely through gluonic content effects in matter, while the $`D`$-mesons will couple strongly through inelastic channels such as $`DN`$ $``$ $`\mathrm{\Lambda }_c`$ or $`\mathrm{\Sigma }_c`$ $`(+\pi )`$. Thus we can expect a large modification of the $`D`$-meson in the nuclear medium as well as $`\rho ^0`$ and $`\omega `$. 2. The existence of novel states, so-called the $`D`$-mesic nuclei, has been predicted by quark meson coupling (QMC) model . The QMC model suggests that $`D^{}(\overline{c}d)`$ may be bound in heavy atoms such as <sup>208</sup>Pb by an attractive scalar-meson exchange and an attractive Coulomb force. It also suggests that $`D^0(c\overline{u})`$ is deeply bound in the nuclei by an attractive $`\omega `$-meson exchange. In view of QSR, it is of importance to investigate the $`D`$ meson-nucleon ($`N`$) interaction. 3. The nuclear absorption of the $`D`$-mesons created in $`\pi `$-$`A`$ and $`p`$-$`A`$ collisions ($`A`$ denotes targets such as Be, Cu, Al, W and Au) has been measured via the decay $`DK\pi `$ at Fermilab. The result indicates that $`D`$-mesons are not completely absorbed in the nuclear targets irrespective of the charge of the produced $`D`$-mesons. The theoretical analysis for the $`D`$-$`N`$ interaction will give an important suggestion for the experimental data in this case too. Motivated by these points, we investigate the properties of $`D`$-mesons in the nuclear matter through the $`D`$-$`N`$ interactions, using an application of Borel-transformed QSR to the $`D`$-$`N`$ forward scattering amplitude. Here we deal with isospin-averaged $`D`$-meson current for simplicity. After a brief explanation of the QSR formulation, in the next section, we calculate the mass shift at finite density using the Borel QSR and compare the results with the mass shift of $`J/\psi `$ calculated in Ref.10. On the basis of these results, we will discuss a possible mechanism of $`J/\psi `$ suppression through the level crossings between the charmonium states and the $`D\overline{D}`$ threshold. ## II QCD sum rule analysis for $`D`$-meson mass shift We start with a two-point in-medium correlation function $`\mathrm{\Pi }_{PS}^{\mathrm{NM}}`$ to discuss hadron properties in the nuclear matter. In the Fermi gas approximation for the matter, $`\mathrm{\Pi }_{PS}^{\mathrm{NM}}`$ is divided into two parts by applying the operator product expansion (OPE) to the correlators in the deep Euclidean region ($`Q^2=q^2>0`$). One is a vacuum part, $`\mathrm{\Pi }_{PS}^0`$, and another is a static one-nucleon part, $`T_{PS}`$. This decomposition is expected to be valid at relatively low density. Namely, in the framework of QSR, $`\mathrm{\Pi }_{PS}^{\mathrm{NM}}`$ can be approximated reasonably well in the linear density of the nuclear matter that all nucleons are at rest : $`\mathrm{\Pi }_{PS}^{\mathrm{NM}}(q)=i{\displaystyle d^4xe^{iqx}\text{T}J_5(x)J_5^{}(0)_{\mathrm{NM}(\rho _N)}}\mathrm{\Pi }_{PS}^0(q)+{\displaystyle \frac{\rho _N}{2M_N}}T_{PS}(q),`$ (1) where $`\rho _N`$ denotes the nuclear matter density, and the forward scattering amplitude $`T_{PS}`$ of the pseudoscalar current-nucleon is $`T_{PS}(\omega ,𝒒)=i{\displaystyle d^4xe^{iqx}N(p)|\text{T}J_5(x)J_5^{}(0)|N(p)}.`$ (2) Here $`q^\mu =(\omega ,𝒒)`$ is the four-momentum carried by the $`D`$-meson current, $`J_5(x)`$ $`=`$ $`J_5^{}(x)`$ $`=`$ $`(\overline{c}i\gamma _5q(x)+\overline{q}\gamma _5c(x))/2`$, where $`q`$ denotes $`u`$ or $`d`$ quark. $`|N(p)`$ represents the isospin, spin-averaged static nucleon-state with the four-momentum $`p=(M_N,𝒑=\mathrm{𝟎})`$, where $`M_N`$ is the nucleon mass, $`0.94`$ GeV. The state is normalized covariantly as $`N(𝒑)|N(𝒑^{})=(2\pi )^32p^0\delta ^3(𝒑𝒑^{})`$. The QSR analysis on the forward scattering amplitude enables us to obtain the information for the $`D`$-$`N`$ interaction. In Eq. (1), the second term means a slight deviation from in-vacuum properties of the $`D`$-meson determined by $`\mathrm{\Pi }_{PS}^0`$. By applying QSR to $`T_{PS}`$, we get the $`D`$-$`N`$ scattering length $`a_D`$ in the limit of $`𝒒\mathrm{𝟎}`$. In this limit, $`T_{PS}`$ can be related to the $`T`$-matrix, $`𝒯_{𝒟𝒩}(m_D,𝒒=\mathrm{𝟎})`$ $`=`$ $`8\pi (M_N+m_D)a_D`$. Near the pole position of the $`D`$-meson, the spectral function $`\rho (\omega ,𝒒=\mathrm{𝟎})`$ is given with three unknown phenomenological parameters $`a,b`$ and $`c`$ in terms of the $`T`$-matrix: $`\rho (\omega ,𝒒=\mathrm{𝟎})`$ $`=`$ $`{\displaystyle \frac{1}{\pi }}{\displaystyle \frac{f_D^2m_D^4}{m_c^2}}\text{Im}\left[{\displaystyle \frac{𝒯_{𝒟𝒩}(\omega ,\mathrm{𝟎})}{(\omega ^2m_D^2+i\epsilon )^2}}\right]+\mathrm{}`$ (3) $`=`$ $`a\delta ^{}(\omega ^2m_D^2)+b\delta (\omega ^2m_D^2)+c\delta (\omega ^2s_0).`$ (4) Here the leptonic decay constant $`f_D`$ is defined by the relation $`0|J_5|D(k)=f_Dm_D^2/m_c`$, where $`|D(k)`$ is the $`D`$-meson state with the four-momentum $`k`$. $`m_D=1.87`$ GeV and $`m_c=1.35`$ GeV are masses of the $`D`$-meson and charm quark respectively. The terms denoted by $`\mathrm{}`$ in Eq.(3) represent the continuum contribution and $`\delta ^{}`$ in Eq.(4) is the first derivative of $`\delta `$ function with respect to $`\omega ^2`$. The first term proportional to $`a`$ is the double-pole term corresponding to the on-shell effect of the $`T`$-matrix and $`a`$ is related to the scattering length $`a_D`$ as $`a=8\pi (M_N+m_D)a_Df_D^2m_D^4/m_c^2`$. The second term proportional to $`b`$ is the single-pole term corresponding to the off-shell effect of the $`T`$-matrix. The third term proportional to $`c`$ is the continuum term corresponding to other remaining effects, where $`s_0`$ is the continuum threshold in the vacuum. Combining the single-pole term of $`\mathrm{\Pi }_{PS}^0`$ in Eq. (1) with Eq. (4), we can relate the scattering length extracted from the QSR of $`T_{PS}`$ with the mass shift of the $`D`$-meson, $`\delta m_D=2\pi {\displaystyle \frac{M_N+m_D}{M_Nm_D}}\rho _Na_D.`$ (5) We may determine these unknown parameters by using a dispersion relation and matching the phenomenological (ph) side with the OPE side. Before such analysis, we can impose a constraint from the low energy theorem among these parameters: In the low energy limit ($`\omega 0`$), $`T_{PS}(\omega ,\mathrm{𝟎})`$ becomes equivalent to the Born term $`T_{PS}^{\mathrm{Born}}(\omega ,\mathrm{𝟎})`$, $`T_{PS}^{\mathrm{ph}}(0)=T_{PS}^{\mathrm{Born}}(0)`$. We assume two cases to take the Born term into the “ph” side. The case (i) and the case (ii) are defined as follows. At $`q_\mu 0`$, we require the $`\omega ^2`$-dependence of the Born term explicitly in the case (i): $`T_{PS}^{\mathrm{ph}}(\omega ^2)=T_{PS}^{\mathrm{Born}}(\omega ^2)+{\displaystyle \frac{a}{(m_D^2\omega ^2)^2}}+{\displaystyle \frac{b}{m_D^2\omega ^2}}+{\displaystyle \frac{c}{s_0\omega ^2}},`$ (6) with the condition $`{\displaystyle \frac{a}{m_D^4}}+{\displaystyle \frac{b}{m_D^2}}+{\displaystyle \frac{c}{s_0}}=0.`$ (7) In the case (ii), we do not require the $`\omega ^2`$-dependence of the Born term explicitly: $`T_{PS}^{\mathrm{ph}}(\omega ^2)={\displaystyle \frac{a}{(m_D^2\omega ^2)^2}}+{\displaystyle \frac{b}{m_D^2\omega ^2}}+{\displaystyle \frac{c}{s_0\omega ^2}},`$ (8) with the condition $`{\displaystyle \frac{a}{m_D^4}}+{\displaystyle \frac{b}{m_D^2}}+{\displaystyle \frac{c}{s_0}}=T_{PS}^{\mathrm{Born}}(0).`$ (9) If $`T_{PS}^{\mathrm{Born}}(0)=0`$, two cases coincide. We determine two unknown phenomenological parameters $`a`$ and $`b`$ in the QSR after the parameter $`c`$ is removed by the condition of Eq. (7) or (9). The analysis of the Born term is easily performed through a calculation of the Born diagrams at the tree level. The isospin states of the $`D`$-meson determine the contribution to the Born term as follows: For the $`D^0(c\overline{u})`$-$`N`$ and $`D^+(c\overline{d})`$-$`N`$ interactions, we need two reactions, $`D^0(c\overline{u})+p(uud)\text{or}n(udd)`$ $``$ $`\mathrm{\Lambda }_c^+,\mathrm{\Sigma }_c^+(cud)\text{or}\mathrm{\Sigma }_c^0(cdd)`$ (10) and $`D^+(c\overline{d})+p(uud)\text{or}n(udd)`$ $``$ $`\mathrm{\Sigma }_c^{++}(cuu)\text{or}\mathrm{\Lambda }_c^+,\mathrm{\Sigma }_c^+(cud),`$ (11) where the static masses of charmed baryons are $`2.28`$ GeV for $`\mathrm{\Lambda }_c^+`$ and $`2.45`$ GeV for $`\mathrm{\Sigma }_c^0`$, $`\mathrm{\Sigma }_c^+`$ and $`\mathrm{\Sigma }_c^{++}`$. Note that we use $`M_B2.4`$ GeV as the average value, where $`B`$ means either $`\mathrm{\Lambda }_c^+`$, $`\mathrm{\Sigma }_c^+`$, $`\mathrm{\Sigma }_c^{++}`$ or $`\mathrm{\Sigma }_c^0`$. After averaging over the nucleon spin, we obtain $`T_{PS}^{\mathrm{Born}}(\omega ,\mathrm{𝟎})={\displaystyle \frac{1}{2}}{\displaystyle \frac{M_N(M_N+M_B)}{\omega ^2(M_N+M_B)^2}}g_p(\omega )^2.`$ (12) The pseudoscalar form-factor $`g_p(q^2)`$ is introduced through the following relation , $`B(p^{})|\overline{Q}i\gamma _5q|N(p)=g_p(q^2)\overline{u}(p^{})i\gamma _5u(p),`$ (13) where $`u(p)`$ is a Dirac spinor and $`Q(q)`$ in the pseudoscalar current is the heavy(light)-quark. Furthermore, $`g_p(q^2)`$ is related to a coupling constant $`g_{NDB}`$ such as $`g_p(q^2)={\displaystyle \frac{2m_D^2f_D}{m_q+m_c}}{\displaystyle \frac{g_{NDB}}{q^2m_D^2}},`$ (14) where $`m_q`$ denotes a light-quark mass. $`g_{ND\mathrm{\Lambda }_c}`$ has been estimated in Ref.17 which gives $`g_{DN\mathrm{\Lambda }_c}6.74`$. However since $`g_{DN\mathrm{\Sigma }_c}`$ has not been evaluated, we take an approximation $`g_{DN\mathrm{\Sigma }_c}`$ $``$ $`g_{ND\mathrm{\Lambda }_c}`$. On the other hand, there are no inelastic channels like the above for $`\overline{D}^0(\overline{c}u)`$-$`N`$ and $`D^{}(\overline{c}d)`$-$`N`$ interactions, i.e. $`T_{PS}^{\mathrm{Born}}(0)=0`$ in both case (i) and case (ii). Eventually we determine parameters $`a`$ and $`b`$ simultaneously after applying the Borel transform to both the OPE side and the “ph” side. ## III Numerical results in the Borel sum rule After performing the Borel transform $`\widehat{B}`$ of the $`T_{PS}`$ calculated in the OPE up to dimension-4 , we obtain as a function of the Borel mass $`M`$ $`\widehat{B}\left[T_{PS}^{\mathrm{OPE}}\right]`$ $`=`$ $`{\displaystyle \frac{1}{2}}e^{m_c^2/M^2}[m_c\overline{q}q_N+2q^{}iD_0q_N(1+{\displaystyle \frac{m_c^2}{M^2}})`$ (15) $`+`$ $`{\displaystyle \frac{1}{2}}m_q\overline{q}q_N\left(1+{\displaystyle \frac{m_c^2}{M^2}}\right)+{\displaystyle \frac{1}{24}}{\displaystyle \frac{\alpha _s}{\pi }}G^2_N\left(2{\displaystyle \frac{m_c^2}{M^2}}\right)`$ $`+`$ $`{\displaystyle \frac{\alpha _s}{\pi }}(uG)^2{\displaystyle \frac{1}{4}}G^2_N{\displaystyle \frac{1}{3}}\{{\displaystyle \frac{4}{3}}{\displaystyle \frac{1}{6}}{\displaystyle \frac{m_c^2}{M^2}}+{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{m_c^2}{M^2}}\right)^3`$ $`+e^{m_c^2/M^2}\left(2\gamma _E\text{ln}\left({\displaystyle \frac{m_c^2}{M^2}}\right)+{\displaystyle _0^{\frac{m_c^2}{M^2}}}𝑑t{\displaystyle \frac{1e^t}{t}}\right)`$ $`+(1{\displaystyle \frac{m_c^2}{M^2}})\text{ln}\left({\displaystyle \frac{m_c^2}{4\pi \mu ^2}}\right)\}],`$ where $`_N`$ denotes the nucleon matrix element. The renormarization scale is taken to be $`\mu ^2=1`$ GeV<sup>2</sup> and the Euler constant is $`\gamma _E=0.5772\mathrm{}`$. Corresponding formula for the “ph” side, in the case (i) reads $`\widehat{B}\left[T_{PS}^{\text{ph}}\right]`$ $`=`$ $`a\left({\displaystyle \frac{1}{M^2}}e^{m_D^2/M^2}{\displaystyle \frac{s_0}{m_D^4}}e^{s_0/M^2}\right)+b\left(e^{m_D^2/M^2}{\displaystyle \frac{s_0}{m_D^2}}e^{s_0/M^2}\right)`$ (16) $`+`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{M_N(M_N+M_B)}{(M_N+M_B)^2m_D^2}}{\displaystyle \frac{4f_D^2m_D^4}{(m_q+m_c)^2}}g_{NDB}^2`$ $`\times `$ $`\left[{\displaystyle \frac{e^{(M_N+M_B)^2/M^2}}{(M_N+M_B)^2m_D^2}}+\left({\displaystyle \frac{1}{(M_N+M_B)^2m_D^2}}{\displaystyle \frac{1}{M^2}}\right)e^{m_D^2/M^2}\right],`$ and that in the case (ii) reads $`\widehat{B}\left[T_{PS}^{\text{ph}}\right]`$ $`=`$ $`a\left({\displaystyle \frac{1}{M^2}}e^{m_D^2/M^2}{\displaystyle \frac{s_0}{m_D^4}}e^{s_0/M^2}\right)+b\left(e^{m_D^2/M^2}{\displaystyle \frac{s_0}{m_D^2}}e^{s_0/M^2}\right)`$ (17) $`+s_0T_{PS}^{\mathrm{Born}}(0)e^{s_0/M^2}.`$ In Eq. (15), a convention, $`(uG)^2G_{\kappa \lambda }^aG_\rho ^{a\lambda }u^\kappa u^\rho `$ is used, where $`u=(1,\mathrm{𝟎})`$ for the static nucleon. We equate Eq. (15) with Eq. (16) or Eq. (17) in the case (i) or the case (ii) respectively. Furthermore we take a first derivative of its equation with respect to $`M^2`$ in each case. We can derive the $`D`$-meson mass shift as a function of $`M^2`$ by removing the parameter $`b`$ from two equations obtained thus. The Borel curve for the $`D`$-meson mass shift, $`\delta m_D`$, is shown in Fig.1. Here we adopt $`f_D0.18`$ GeV for the decay constant and $`s_06.0`$ GeV<sup>2</sup> for the continuum threshold. The values of these parameters were estimated by the analysis of the in-vacuum correlation function in Ref.19, 20. This $`f_D`$ value is very close to the result of the lattice QCD calculation ($`0.194\pm 0.01`$ GeV) and is consistent with the upper bound of experimental data ($``$ $`0.31`$ GeV) . Note that the parameters are fixed in our calculation. We use the nucleon matrix elements for quark fields such as $`\overline{q}q_N5.3`$ GeV and $`q^{}iD_0q_N0.34`$ GeV<sup>2</sup>, and for gluon fields such as $`\frac{\alpha _s}{\pi }G^2_N`$ $``$ $`1.2`$ GeV<sup>2</sup> and $`\frac{\alpha _s}{\pi }(uG)^2\frac{1}{4}G^2_N`$ $``$ $`0.1`$ GeV<sup>2</sup>. The light-quark mass is taken to be $`m_q0.008`$ GeV. In Fig.1, the solid line (“$`\mathrm{Born}=0`$”\[case (i)\]) are calculated without the contribution of inelastic channels. In the doted line (“With Born term”\[case (ii)\]) and the dashed line (“Without Born term”), we allow for the contribution of the sub-threshold resonances $`\mathrm{\Lambda }_c`$ and $`\mathrm{\Sigma }_c`$, lying very close to the $`D`$-$`N`$ threshold. As shown in Fig.1, the sub-threshold effect is repulsive both in the case (i) and the case (ii) and its contribution in the case (ii) is stronger than that in the case (i). This is analogous to the case for the $`K^{}`$-$`p`$ interaction . In Fig.1, the remaining three lines are evaluated by extracting the contribution of $`m_c\overline{q}q_N`$ term in the OPE from the solid, dashed and doted lines respectively. We find that the contribution of $`m_c\overline{q}q_N`$ term is more than $`95\%`$ of total contribution corresponding to the solid, dashed and doted lines within the plateau regions. The analysis of this graph is summarized in Table 1. We cannot determine th Borel windows, since there is only lowest-dimension term in the OPE side. Therefore we take the following procedure to determine a window of $`M^2`$: We focus on a plateau region of each line shown in Fig.1. First, we take the minimum point in the line as the smallest value of $`\delta m_D`$. Next, we determine two points in $`M^2`$ as the deviation from the minimum value of $`\delta m_D`$ becomes less than $`10\%`$ of the minimum value. We take the region between the two points as a window of $`M^2`$. The window of $`M^2`$ in each line is given in Table 1. As shown in Table 1, all the Borel curves are rather stable within the windows. Note that the windows discussed above are also close to that obtained by scaling up typical Borel-windows for the light-vector mesons. An estimate from the light-vector mesons is $`(\frac{m_D}{m_V})^2\times 0.8(=4.7)<M^2<(\frac{m_D}{m_V})^2\times 1.3(=7.7)`$ for $`V=\rho ,\omega `$ and $`(\frac{m_D}{m_V})^2\times 1.3(=4.4)<M^2<(\frac{m_D}{m_V})^2\times 1.8(=6.1)`$ for $`V=\varphi `$. It should be stressed that the QSR for vacuum correlation function $`\mathrm{\Pi }_{PS}^0`$ up to dimension-6 operators cannot reproduce well the $`D`$-meson mass ($`m_D`$) in free space. The Borel curve of $`m_D`$ at $`s_0=6.0`$ GeV<sup>2</sup> does not have any stability and seems to give larger values than the experimental data ($`m_D=1.87`$ GeV) in the plateau region ($`M^2=38`$ GeV<sup>2</sup>) of $`\delta m_D`$ discussed above. Furthermore, the curve of $`m_D`$ has rather larger change than that of $`\delta m_D`$ in the region. Therefore, if we perform the QSR analysis for the effective mass $`m_D^{}=m_D+\delta m_D`$, the stability in the Borel curve of $`m_D^{}`$ will become obscure. This implies that the QSR analysis for $`m_D^{}`$ is not so valid in this case. However the application of the QSR to $`\delta m_D`$ gives us a possible indication for the $`D`$-meson mass-shift as discussed above. From the above analysis we obtain $`\delta m_D=48\pm 8`$ MeV by considering both cases ((i) and (ii)). Then the $`D`$-$`N`$ scattering length $`a_D`$ is $`0.72\pm 0.12`$ fm. This result suggests that the $`D`$-$`N`$ interaction is more attractive than the $`J/\psi `$-$`N`$ interaction, where $`\delta m_{J/\psi }`$ is about $`5`$ MeV and $`a_{J/\psi }`$ is about $`0.1`$ fm . It is noted the above results were performed at the nuclear matter density, $`\rho _N=\rho _02\rho _0`$, where $`\rho _0`$ is the normal matter density. Assuming that the linear density approximation in the QSR is valid at the nuclear matter density, we find these results lead to a larger decrease of the $`D`$-meson mass than that of the charmonium. We also find that the large mass-shift of the $`D`$-meson originates from the contribution of $`m_c\overline{q}q_N`$ term in the OPE. In the QMC model, the contribution from the $`m_c\overline{q}q_N`$ term may correspond to a quark-$`\sigma `$ meson coupling. In fact, the model predicts the mass shift of the $`D`$-meson becomes $`60`$ MeV for the scalar potential at the normal matter density . Their results are very close to our results. ## IV Concluding remarks We present an analysis of isospin-averaged $`D`$-meson mass-shift through a direct application of the Borel QSR to the forward scattering amplitude of the pseudoscalar current and nucleon. Here we perform the operator product expansion with all the terms up to dimension-4. The result predicts an attractive mass-shift of the isospin-averaged $`D`$-meson about $`50`$ MeV (about $`3\%`$ of the bare mass) at the normal matter density. The contribution of $`m_c\overline{q}q_N`$ term in the OPE is more than $`95\%`$ of the results evaluated up to all the dimension-4 operators. Our result is very close to that reported by the QMC model. Then the $`D`$-$`N`$ scattering length $`a_D`$ is about $`0.7`$ fm and indicates a strongly attractive force between the $`D`$-meson and nucleons. The mass modification of the $`D`$-meson in the nuclear matter corresponds to $`10`$ times larger than that of $`J/\psi `$. Recent NA50 Collaboration has reported a strong (so called “anomalous”) $`J/\psi `$ suppression in $`Pb`$-$`Pb`$ collision at $`158`$ GeV per nucleon. We suggest a following possibility to cause such a strong $`J/\psi `$ suppression using the above large difference of the mass shifts between the $`D`$-meson and $`J/\psi `$: Suppose that the other $`c\overline{c}`$-states such as $`\psi ^{}`$ and $`\chi _c`$’s do not also have a large mass-shift in the nuclear matter as discussed above for $`J/\psi `$. Then on the basis of our calculations, the $`D\overline{D}`$ threshold ($`3.74`$ GeV in free space) decreases by $`100`$ MeV at $`\rho _N=\rho _0`$ and comes down between $`\chi _{c2}`$ ($`3.55`$ GeV) and $`\psi ^{}`$ ($`3.68`$ GeV). At $`\rho _N=2\rho _0`$, it decreases by $`200`$ MeV and comes down between $`\chi _{c1}`$ ($`3.50`$ GeV) and $`\chi _{c2}`$ ($`3.55`$ GeV). Such level crossings between the $`D\overline{D}`$ threshold and the charmonium spectrum are shown in Fig.2. It is well known from the $`p`$-$`A`$ collision data that only $`60\%`$ of $`J/\psi `$ observed are directly produced and the remainder comes from excited states ($`\chi _c`$($`1P`$), $`\psi ^{}`$($`2S`$)) with a ratio of 3 to 1 . So if the disappearance of $`\psi ^{}`$ and $`\chi _c`$ due to their decays into $`D\overline{D}`$ takes place, it could lead to a decrease of the $`J/\psi `$ yield. This $`J/\psi `$ suppression will have stair-shaped form as a function of the energy density of the system. At low energy density, only $`\psi ^{}`$ $`D\overline{D}`$ takes place, which causes a slight suppression of $`J/\psi `$. At intermediate density, subsequent suppression occurs by the level crossing of ($`\chi _{c2},\chi _{c1}`$) states with the $`D\overline{D}`$ threshold. At higher density, direct suppression of $`J/\psi `$ by the decay $`J/\psi `$ $`D\overline{D}`$ occurs. This is an alternative or supplementary effect to the deconfinement scenario in Ref.29 and may have implications to the recent data of NA50 Collaboration. Strictly speaking, as mentioned before, the formulation of QSR used here is applicable to dilute nuclear matter in comparison to highly dense nuclear matter produced in high energy heavy-ion collision. If we try to discuss the mass shifts at higher density, it will need an extension or some improvements in the above formulation. This is one of important future works for us. However we may say that our calculation in this paper can suggest a possibility of the level crossings even in only 1 or 2 times normal matter density and lets one expect the direct $`J/\psi `$ suppression at sufficiently higher density. We may also suggest another phenomenon caused by the level crossings discussed above. It is a change of the decay width of the charmonium. In the vacuum the resonances above $`D\overline{D}`$ threshold, for example the $`\psi ^{\prime \prime }`$ state, have a width of order MeV because of the strong open-charm channel. On the other hand, the resonances below the threshold have a very sharp width of a few hundreds keV. So after the level crossings, the decay modes of the $`\psi ^{}`$ and $`\chi _c`$ states will change drastically at least one order of magnitude. Needless to say, in the high energy heavy-ion collisions we must also perform the theoretical investigation of finite temperature effect to the mass modification. Therefore, a future task will be to understand the medium effect in terms of the QSR when both temperature and density are finite. As for investigation of only the $`D`$-$`N`$ interaction, an inverse kinematics experiment will give useful information. Since the projectile (target) is a heavy-ion beam (light-nuclei), the decays of the $`D`$-meson and the charmonium inside a nucleus will be possible in such an experiment. Before closing, it is stressed that we should take the above analysis for the isospin-averaged $`D`$-$`N`$ interaction as a qualitative estimate. If we wish more quantitative discussion, the isospin-decomposition of the $`D`$-meson and higher correction terms beyond dimension-4 operators must be taken into account. Particularly the odd-components in the OPE play important roles for a difference of the mass shift between $`D`$ and $`\overline{D}`$. ## Acknowledgements I would like to thank T. Hatsuda for careful reading of the manuscript and many useful discussion. I also thank S.H. Lee and D. Jido for helpful discussion and suggestion.
warning/0001/hep-th0001184.html
ar5iv
text
# Phases of dual superconductivity and confinement in softly broken 𝒩=2 supersymmetric Yang–Mills theories ## I Introduction Certainly, one of the most beautiful ideas in the context of quantum chromodynamics (QCD) is the confinement mechanism envisaged by ’t Hooft and Mandelstam through the condensation of light monopoles. In essence it states that the QCD vacuum should behave as a dual superconductor where magnetic order takes place, and electric flux tubes form thus producing color confinement. In the context of QCD it stands for a kind of descriptive scheme, as long as it is not known how magnetically charged quanta can arise and condense in the effective low energy theory. In this respect, the idea of Abelian projection proposed by ’t Hooft has received increasing support from numerical simulations on the lattice in the last few years . Even in the continuum, recent work using a novel parametrization of QCD , points in the direction of the above scenario for color confinement . From the analytical side, the understanding of non-perturbative phenomena in four dimensional quantum field theory has been put several steps forward since Seiberg and Witten constructed an exact solution for the low energy dynamics of $`SU(2)`$ $`𝒩=2`$ supersymmetric Yang-Mills theory . In particular, it was possible for them to show that the mechanism of color confinement devised by ’t Hooft and Mandelstam takes place when supersymmetry is broken down to $`𝒩=1`$. These results were soon extended to the case of $`SU(N)`$ . Furthermore, when $`𝒩=2`$ supersymmetry is softly broken down to $`𝒩=0`$, the same mechanism has been shown to persist . In spite of the fact that these results are well known, not much attention has been paid to the actual solutions in the strong coupling limit corresponding to electric flux lines that would undertake quark confinement. In Ref., it was shown that this sort of vortices should have a spectrum of string tensions that distinguishes among different factors in the magnetic $`U(1)^{N1}`$ theory arising in the infrared. The same result was found in the framework of the M-theory fivebrane version of QCD, also named MQCD . The string tension of the $`N1`$ electric flux tubes $`T_k,k=1,\mathrm{},N1,`$ is given –up to a dimensionful factor which is different for each theory but independent of $`k`$–, by a dimensionless function $`f_N(k)=\mathrm{sin}(\pi k/N)`$. This function is somehow universal as long as the soft breaking perturbation is carried by a single Casimir operator . Even in that case, the problem of finding such solutions in the particular model that emerges in this context has not yet been addressed in detail, <sup>*</sup><sup>*</sup>* Except for $`SU(2)`$, both at the maximal singularity of the Coulomb branch for pure gauge , and on the Higgs branch in the theory with massive fundamental matter . probably due to the naive expectation that the effective theory consists of $`N1`$ copies of the standard $`U(1)`$ Abelian Higgs model. It was already shown by Douglas and Shenker that the magnetic $`U(1)`$ factors of the infrared quantum theory describing the neighborhood of the monopole singularities are coupled . The existence of these off-diagonal couplings, $`\tau _{ij}^{\mathrm{off}}`$, was confirmed in two different frameworks. First, they appear in the expression of the Donaldson–Witten functional for gauge group $`SU(N)`$ . Moreover, these couplings were shown to satisfy a stringent constraint coming from the Whitham hierarchy formulation of the Seiberg–Witten solution in Ref. where, in addition, a general ansatz for $`\tau _{ij}^{\mathrm{off}}`$ is given. In this paper, we extend the work to the case of $`SU(N)`$, $`𝒩=2`$ supersymmetric Yang-Mills theory softly broken to $`𝒩=1`$. The analysis is performed in a “peculiar” scaling limit (named “ultrastrong” in ). We show that, even in that limit, generically there are no BPS electric flux tubes. A perturbative analysis leads to the conclusion that the phase of dual superconductivity is of type II i.e. there is a short range repulsive force between different vortices. This fact supports the expectation that indeed electric flux lines are safely confined into stable flux tubes, a feature of the confinement mechanism which is not a priori granted. It could happen, for example, that the electric vortices result to be unstable, and their core grows and smears in such a way that they do not lead to confinement of electric charges . It is worth mentioning in this respect that numerical simulations in lattice QCD seem to point out that the type of Abelian Higgs model behind the picture of dual superconductivity is a critical one between type I and type II . The plan of the paper is as follows. The setup of the problem is given in Section II where some aspects of the low-energy dynamics of $`𝒩=2`$ supersymmetric gauge theories softly broken down to $`𝒩=1`$ are reviewed. We emphasize the existence of non-vanishing couplings between the different $`U(1)`$ factors –even at the maximal singularity–, which play an essential rôle in our results. In Section III, we show that the string tension of vortex-like configurations obeys a Bogomol’nyi bound in the ultrastrong scaling limit. However, there are no BPS electric vortices in the system unless the complex phases of the soft breaking parameters corresponding to different Casimir operators are aligned. Even in this case, we show that the string tensions of the resulting BPS vortices are governed by a dimensionless function $`f_N(k)`$ which is different from the one obtained in , the latter being recovered as a particular limit of our system corresponding to a single quadratic $`𝒩=1`$ perturbation. In Section IV, we focus for convenience on the group $`SU(3)`$ and analyze the critical vortex solutions in certain simplified cases. We speculate about the full spectrum of such configurations. A perturbative analysis of the dynamics expected for nearly critical vortices is performed in sections V and VI by means of energetic arguments. This analysis reveals the existence of repulsive forces among vortices corresponding to different magnetic $`U(1)`$ factors. Thereafter we refer to this phase as an “hybrid” type II phase. In Section VII, the half Higgsed vacua are considered. The similarities and differences with the model proposed by Witten to describe cosmic superconducting strings are discussed. We find solutions to the Bogomol’nyi equations with the behavior of superconducting strings. Finally, Section VIII is devoted to our conclusions and further remarks. ## II Infrared dynamics at maximal singularities The quantum moduli space of vacua $`_\mathrm{\Lambda }`$ of $`SU(N)`$, $`𝒩=2`$ supersymmetric gauge theory has a singular locus given by hypersurfaces of complex codimension one which may intersect with each other . Along each of these hypersurfaces, an extra massless degree of freedom –whose quantum numbers can be read off from the monodromy matrix corresponding to a closed path encircling the singularity–, must be included into the effective action. At the intersections, many states become simultaneously massless. Of special interest are those singularities where $`N1`$, i.e. the maximum allowed number of mutually local states, become massless. They are accordingly called maximal singularities. In $`SU(3)`$, for example, the singular locus is given by the complex curves $$4u^327(v\pm 2\mathrm{\Lambda }^3)^2=0,$$ where $`u=1/2\mathrm{Tr}\varphi ^2`$ and $`v=1/3\mathrm{Tr}\varphi ^3`$ are the gauge invariant order parameters constructed out of the scalar field belonging to the $`𝒩=2`$ vector supermultiplet, and $`\mathrm{\Lambda }`$ is the quantum dynamical scale. Higher intersections of these curves lead to the so-called $`Z_2`$ and $`Z_3`$ singularities, given respectively by the points $`\{u^3=27\mathrm{\Lambda }^6,v=0\}`$ and $`\{u=0,v^2=4\mathrm{\Lambda }^6\}`$. The addition of a microscopic superpotential breaks supersymmetry and leads to an $`𝒩=1`$ theory $$W_{𝒩=1}=\underset{k=2}{\overset{N}{}}\frac{1}{k}\lambda _k\mathrm{Tr}\mathrm{\Phi }^k.$$ (1) Notice that those contributions in (1) with $`k>3`$ are non-renormalizable. However, this does not necessarily mean that they do not affect the low-energy dynamics. They could be dangerously irrelevant operators . We will not discuss these subtleties here, and shall restrict ourselves to the case of up to cubic perturbations, $$W_{𝒩=1}=\frac{1}{2}\mu \mathrm{Tr}\mathrm{\Phi }^2+\frac{1}{3}\nu \mathrm{Tr}\mathrm{\Phi }^3.$$ (2) This breaking is soft, in fact renormalizable. The continuum vacuum degeneracy is lifted except for a given set of points which depend on the actual values of the parameters $`\mu `$ and $`\nu `$. <sup>§</sup><sup>§</sup>§ For instance, in the case of $`SU(3)`$, the theory has generically five $`𝒩=1`$ vacua, three of which are the maximal $`Z_2`$ points. In the limit $`\mu 0`$, the remaining vacua approach the $`Z_3`$ points . Let us focus on the low-energy effective field theory near a maximal point that we choose, for simplicity, to be that with real quadratic Casimir, $`u=N\mathrm{\Lambda }^2`$. This is a dual $`𝒩=2`$ supersymmetric gauge system with gauge group $`U(1)^{N1}`$ which includes both chiral multiplets $`\mathrm{\Psi }_i^D=(\chi _i^D,V_i^D)`$, as well as hypermultiplets $`H_i=(M_i,\stackrel{~}{M}_i)`$ that correspond to the monopoles that become light in that patch of the moduli space. One can choose a homology basis for the cycles on the auxiliary curve such that each monopole has a unit of charge with respect to each dual gauge field. The quantities $`\chi _i^D`$, $`M_i`$ and $`\stackrel{~}{M}_i`$ are chiral $`𝒩=1`$ superfields, while $`V_i^D`$ are $`𝒩=1`$ vector superfields (and $`W_\alpha ^{Di}`$ their corresponding superfield strengths). For completeness, we give also the $`𝒩=0`$ content of these superfields $`\chi _i^D=(\varphi _i^D,\psi _i,F_i)`$ $`V_i^D=((A_\mu ^D)_i,\lambda _i,D_i)`$ $`M_i=(\varphi _{m_i},\psi _{m_i},F_{m_i})`$ $`\stackrel{~}{M}_i=(\stackrel{~}{\varphi }_{m_i},\stackrel{~}{\psi }_{m_i},\stackrel{~}{F}_{m_i})`$ where the notation for fermionic, bosonic and auxiliary components is the standard one. Setting $`a_i^D\varphi _i^D`$, the coordinates at the point of maximal singularity that we are focusing on are $`a_i^D=0`$. The dominant piece of the $`𝒩=2`$ low energy effective Lagrangian is given in terms of a holomorphic function $``$, called the effective prepotential $`_{eff}^{N=2}`$ $`=`$ $`{\displaystyle \frac{1}{4\pi }}\mathrm{Im}\left[{\displaystyle d^4\theta \frac{(\chi ^D)}{\chi _i^D}\chi _i^D}+{\displaystyle \frac{1}{2}}{\displaystyle d^2\theta \frac{^2(\chi ^D)}{\chi _i^D\chi _j^D}W_\alpha ^{Di}W^{D\alpha j}}\right]`$ (3) $`+`$ $`{\displaystyle d^4\theta \{M_i^{}e^{2V_i^D}M_i+\stackrel{~}{M}_i^{}e^{2V_i^D}\stackrel{~}{M}_i\}}+\mathrm{Re}{\displaystyle d^2\theta W(\chi ^D,M,\stackrel{~}{M})}.`$ (4) The monopole fields have been “integrated in” in order to soak up the singularity of the effective action when $`a_i^D=\varphi _i^D0`$ where $`M_i`$ becomes massless. The effective superpotential at low energies is $$W(\chi ^D,M,\stackrel{~}{M})=\sqrt{2}M_i\chi _i^D\stackrel{~}{M}_i+\mu 𝒰(\chi ^D)+\nu 𝒱(\chi ^D),$$ (5) the last two terms being the effective contribution of the supersymmetry breaking superpotential (2). In fact, $`𝒰`$ and $`𝒱`$ are the Abelian superfields arising respectively from the quadratic and cubic Casimir operators in the low-energy theory. The vacuum expectation value of their lowest components, $`U`$ and $`V`$, are the holomorphic coordinates in $`_\mathrm{\Lambda }`$, $`U=u`$ and $`V=v`$. Written in component fields, the bosonic sector of the system is described by the Lagrangian $`_{eff,B}^{𝒩=1}`$ $`=`$ $`{\displaystyle \frac{1}{4}}b_{ij}(F_{\mu \nu })_i(F^{\mu \nu })_j+(D_\mu \varphi _{m_i})^{}D^\mu \varphi _{m_i}+D_\mu \stackrel{~}{\varphi }_{m_i}(D^\mu \stackrel{~}{\varphi }_{m_i})^{}+b_{ij}_\mu \varphi _{i}^{D}{}_{}{}^{}^\mu \varphi _j^D`$ (6) $``$ $`\left[{\displaystyle \frac{1}{2}}b_{ij}D_iD_j+b_{ij}F_i^{}F_j+F_{m_i}^{}F_{m_i}+\stackrel{~}{F}_{m_i}^{}\stackrel{~}{F}_{m_i}\right],`$ (7) where the auxiliary fields are solved as $`D_i=b_{ij}^1(|\varphi _{m_j}|^2|\stackrel{~}{\varphi }_{m_j}|^2)`$ $`F_i=b_{ij}^1\left(\sqrt{2}\varphi _{m_j}\stackrel{~}{\varphi }_{m_j}+C_j\right)`$ (8) $`F_{m_i}=\sqrt{2}\varphi _{i}^{D}{}_{}{}^{}\stackrel{~}{\varphi }_{m_i}^{}`$ $`\stackrel{~}{F}_{m_i}=\sqrt{2}\varphi _{i}^{D}{}_{}{}^{}\varphi _{m_i}^{},`$ (9) whereas field strengths and covariant derivatives are given by $`(F_{\mu \nu })_i`$ $`=`$ $`_\mu (A_\nu ^D)_i_\nu (A_\mu ^D)_i,`$ (10) $`D_\mu \varphi _{m_i}=_\mu \varphi _{m_i}+i(A_\mu ^D)_i\varphi _{m_i}`$ $`D_\mu \stackrel{~}{\varphi }_{m_i}=_\mu \stackrel{~}{\varphi }_{m_i}i(A_\mu ^D)_i\stackrel{~}{\varphi }_{m_i}.`$ (11) Concerning $`C_j`$ in (8), it stands for $$C_j(\varphi ^D)=\mu U_j(\varphi ^D)+\nu V_j(\varphi ^D)|C_j|e^{i\beta _j},$$ (12) where $`U_j`$ and $`V_j`$ are the derivatives of $`U`$ and $`V`$ with respect to $`\varphi _j^D`$ , $`U_j(\varphi ^D)=u_j^{(0)}\mathrm{\Lambda }+{\displaystyle \underset{p1}{}}u_j^{(p)}(\varphi ^D)\mathrm{\Lambda }^{1p},`$ $`u_j^{(0)}=2j\mathrm{sin}\widehat{\theta }_j`$ (13) $`V_j(\varphi ^D)=v_j^{(0)}\mathrm{\Lambda }^2+{\displaystyle \underset{p1}{}}v_j^{(p)}(\varphi ^D)\mathrm{\Lambda }^{2p},`$ $`v_j^{(0)}=2j\mathrm{sin}2\widehat{\theta }_j,`$ (14) while $`u_j^{(p)}(\varphi ^D)`$ and $`v_j^{(p)}(\varphi ^D)`$ are homogeneous polynomials in $`\varphi _i^D`$ of degree $`p`$, so that $`C_j`$ are regular functions in the vicinity of the maximal singularity. Finally, $`b_{ij}`$ is ($`\frac{1}{4\pi }`$ times) the imaginary part of the period matrix $`\tau _{ij}^D`$, $$\tau _{ij}^D(\varphi ^D)=\frac{^2}{\varphi _i^D\varphi _j^D}=\frac{1}{2\pi i}\mathrm{log}\left(\frac{\varphi _i^D}{\mathrm{\Lambda }_i}\right)\delta _{ij}+\tau _{ij}^{\mathrm{off}}+𝒪\left(\frac{\varphi ^D}{\mathrm{\Lambda }}\right),$$ (15) where $`\mathrm{\Lambda }_j=\mathrm{\Lambda }\mathrm{sin}\widehat{\theta }_j`$ and $`\widehat{\theta }_j=j\pi /N`$. When expanding around the vacuum expectation value $`a_i^D=\varphi _i^D`$, $`\tau _{ij}^D(\varphi ^D)`$ yields the effective coupling constant matrix. The logarithmic singularity when $`a_i^D=0`$ corresponds to the perturbative running of the dual coupling constant up to the maximal point, displaying the asymptotic freedom of the dual description. The coupling flows to zero due to the fact that the quantum fluctuations of massless monopoles have been integrated out. This is fine as long as one is interested only in searching for vacuum solutions. Then $`M`$ and $`M^{}`$ in (4)–(5) stand for the zero modes of the monopole field (see the discussion in ). Here, however, in order not to run into double counting of degrees of freedom we should introduce, on physical grounds, an infrared cut off for the monopole loop integrals. In each $`U(1)`$ factor the natural energy scale is set by the soft breaking parameters $`a_i^D|C_i^{(0)}|^{1/2}`$ with $$C_i^{(0)}=\mu u_i^{(0)}\mathrm{\Lambda }+\nu v_i^{(0)}\mathrm{\Lambda }^2=2i\mathrm{\Lambda }(\mu \mathrm{sin}\widehat{\theta }_i+\nu \mathrm{\Lambda }\mathrm{sin}2\widehat{\theta }_i),$$ (16) and the perturbative couplings of each monopole to its corresponding dual vector field, $$\frac{4\pi }{g_{D}^{2}{}_{i}{}^{}}\frac{1}{4\pi }\mathrm{log}\left(\frac{|C_i^{(0)}|}{\mathrm{\Lambda }_i^2}\right),$$ (17) show logarithmic variations among different $`U(1)`$ factors. In other words, we are dealing here with a macroscopic (classical) theory of the Ginzburg–Landau type, and we should consider the coupling constant of the $`M_i`$ and $`\stackrel{~}{M}_i^{}`$ classical fields to $`V_i^D`$: wave-particle duality connects $`g_D`$ with the running coupling constant of the quantum theory through the formula $`\mathrm{}g_{Di}=g_{Di}(a_i^D|C_i^{(0)}|^{1/2})`$, the strong coupling limit becoming the classical limit for the magnetically charged quanta . Even in the close vicinity of the singularity, different magnetic $`U(1)`$ factors are coupled through $`\tau _{ij}^{\mathrm{off}}`$ . Exactly at the singularity, i.e. at $`a_i^D=0`$, the generic expression proposed in for these off-diagonal couplings is $$\tau _{ij}^{\mathrm{off}}=\frac{2i}{N^2\pi }\underset{k=1}{\overset{N1}{}}\mathrm{sin}k\widehat{\theta }_i\mathrm{sin}k\widehat{\theta }_j\underset{p,q=1}{\overset{N}{}}\tau _{pq}^{(0)}\mathrm{cos}k\theta _p\mathrm{cos}k\theta _q,$$ (18) where $`\tau _{pq}^{(0)}`$ is given by $$\tau _{pq}^{(0)}=\delta _{pq}\underset{kp}{}\mathrm{log}(2\mathrm{cos}\theta _p2\mathrm{cos}\theta _k)^2(1\delta _{pq})\mathrm{log}(2\mathrm{cos}\theta _p2\mathrm{cos}\theta _q)^2,$$ (19) with $`\theta _p=(p1/2)\pi /N`$ and $`p,q=1,\mathrm{},N`$. In the case of $`SU(3)`$, for example, $`\tau _{12}^{\mathrm{off}}=i/\pi \mathrm{log}2`$ . These interactions are also present in the effective potential obtained from the terms in square brackets of (7), $`V_{eff}`$ $`=`$ $`{\displaystyle \frac{1}{2}}b_{ij}^1(\varphi ^D)(|\varphi _{m_i}|^2|\stackrel{~}{\varphi }_{m_i}|^2)(|\varphi _{m_j}|^2|\stackrel{~}{\varphi }_{m_j}|^2)+2|\varphi _i^D|^2(|\varphi _{m_i}|^2+|\stackrel{~}{\varphi }_{m_i}|^2)`$ (20) $`+`$ $`b_{ij}^1(\varphi ^D)(\sqrt{2}\varphi _{m_i}\stackrel{~}{\varphi }_{m_i}+C_i(\varphi ^D))(\sqrt{2}\varphi _{m_j}\stackrel{~}{\varphi }_{m_j}+C_j(\varphi ^D))^{}.`$ (21) Notice that, $`b_{ij}^1`$ being positive definite, the potential is either positive or zero. Given the expectation values of the complex scalars, $$\varphi _i^D=a_i^D\varphi _{m_i}=m_i\stackrel{~}{\varphi }_{m_i}=\stackrel{~}{m}_i,$$ (22) $`𝒩=1`$ supersymmetric vacua are in one to one correspondence with zeroes of $`V_{eff}`$: $`\sqrt{2}m_i\stackrel{~}{m}_i`$ $`\text{}=`$ $`C_i(a^D),`$ (23) $`m_ia_i^D`$ $`\text{}=`$ $`\stackrel{~}{m}_ia_i^D=0,`$ (24) $`|m_i|`$ $`\text{}=`$ $`|\stackrel{~}{m}_i|.`$ (25) $`i=1,2,\mathrm{}N1`$. From (24) we learn that monopole condensation can only occur at hypersurfaces where $`a_i^D=0`$ for some $`i`$. At the maximal singularity, every $`a_i^D`$ vanishes, and it is clear from (23)–(25) that $`N1`$ monopoles have a chance to condense. While soft breaking is parametrized by $`\mu `$ and $`\nu `$, monopole condensation is controlled by $`C_i`$. If for some $`j`$ we have $`a_j^D=0`$ and adjust $`C_j^{(0)}=0`$, the corresponding $`U(1)`$ remains unbroken ($`m_j=\stackrel{~}{m}_j=0`$), and the vacuum is said to be “partially Higgsed”. Summarizing, the Higgs vacuum $``$ at the maximal point is given by $$=\{m_i,\stackrel{~}{m}_i/|m_i|^2=|\stackrel{~}{m}_i|^2=|C_i^{(0)}|/\sqrt{2},\stackrel{~}{m}_j=e^{i\beta _j^{(0)}}m_j^{}\},$$ (26) with $`C_i^{(0)}=|C_i^{(0)}|e^{i\beta _i^{(0)}}`$ given in equation (16). Since the absolute phases of $`m_i`$ are not fixed, it has the topology of a torus of genus $`g=N1`$. Equation (26) shows that the scalar components of the monopole superfields condense in the vacua placed at the maximal points. Although the presence of condensation suggests that confinement indeed takes place, some further analysis is required before this can be definitively established. An important question to be answered is whether the collimation of the electric (or dual magnetic) flux lines is energetically favored or not. This is a dynamical issue that goes beyond the simple vacuum analysis. ## III Bogomol’nyi Bound in the Ultrastrong Scaling Limit The resulting effective theory we have arrived at, in the bosonic sector, is an Abelian $`(N1)`$–Higgs model with coupled $`U(1)`$ factors and a quite non-standard Higgs potential. The search for stable vortex solutions in the complete system is a hard problem. On general grounds, one should not expect to have BPS string solutions in spite of the fact that $`𝒩=1`$ supersymmetry is enough, generically, to have BPS vortices in four dimensions . At least, this is the case of $`𝒩=1`$ QCD, where the strings are conserved modulo $`N`$ so they cannot carry an additive conserved quantity such as a central charge . There is a limit, however, in which the system simplifies and admits BPS vortices . It happens whenever the condensation parameters (12) are independent of $`\varphi ^D`$, something that corresponds to linear perturbations in the superpotential (5), i.e. Fayet–Iliopoulos terms. This kind of terms together with properly normalized quartic potentials are known to lead to Abelian Higgs models that admit BPS vortices . Taking into account (13)–(14), one should consider $`\mathrm{\Lambda }\mathrm{}`$ and small values of the soft breaking parameters $`\mu 0`$ and $`\nu 0`$, such that $`\mu \mathrm{\Lambda }`$ and $`\nu \mathrm{\Lambda }^2`$ remain finite. In this “ultrastrong” limit, $`C_i(\varphi _D)C_i^{(0)}`$ are constants, and one can easily check that setting $`a_i^D=\varphi _i^D=0`$ is a consistent constraint. One may then study the existence of extended solutions in the remaining fields. The (bosonic part of the) effective Lagrangian adopts the following form: $`_{eff,B}^{𝒩=1}`$ $`=`$ $`{\displaystyle \frac{1}{4}}b_{ij}^{(0)}F_{\mu \nu }^iF^{j\mu \nu }+(D_\mu \varphi _{m_i})^{}D^\mu \varphi _{m_i}+D_\mu \stackrel{~}{\varphi }_{m_i}(D^\mu \stackrel{~}{\varphi }_{m_i})^{}`$ (27) $``$ $`\left[{\displaystyle \frac{1}{2}}b_{ij}^{(0)}D_i^{(0)}D_j^{(0)}+b_{ij}^{(0)}F_{i}^{(0)}{}_{}{}^{}F_j^{(0)}\right],`$ (28) where $`b_{ij}^{(0)}`$ stands for the actual value of $`b_{ij}`$ at the maximal singularity and $`D_i^{(0)},F_i^{(0)}`$ are obtained from (8) by replacing $`b_{ij}`$ with $`b_{ij}^{(0)}`$. It is now feasible to give an expression à la Bogomol’nyi for the energy per unit length corresponding to static and magnetically neutral ($`(A_0^D)_i=0`$) vortex-like configurations (i.e. configurations with translational symmetry along one axis) by means of the remainder $`𝒩=1`$ supersymmetry (see also where a multi Higgs system has been treated). Indeed, the energy density can be rearranged as follows: $`_{eff}`$ $`=`$ $`{\displaystyle \frac{1}{2}}b_{ij}^{(0)}\left(F_{12}^i\pm D_i^{(0)}\right)\left(F_{12}^j\pm D_j^{(0)}\right)+\left|(D_1\pm iD_2)\varphi _{m_i}\right|^2+\left|(D_1\pm iD_2)\stackrel{~}{\varphi }_{m_i}\right|^2`$ (29) $`+`$ $`b_{ij}^{(0)}F_i^{(0)}F_{j}^{(0)}{}_{}{}^{}ϵ_{ab}_a𝒥_b,`$ (30) where the last term, corresponding to the current $`𝒥_b=i(\varphi _{m_i}^{}D_b\varphi _{m_i}+\stackrel{~}{\varphi }_{m_i}^{}D_b\stackrel{~}{\varphi }_{m_i})`$, does not contribute to the string tension for finite energy configurations. It is easier to analyze this system in a different set of variables, obtained from the above ones by means of an $`SU(2)_R`$ transformation yielding $`D_i^{(0)}`$ $``$ $`\widehat{D}_i^{(0)}=\sqrt{2}\text{Re}(e^{i\alpha }F_i^{(0)}),`$ (31) $`\text{}\sqrt{2}F_i^{(0)}`$ $``$ $`\sqrt{2}\widehat{F}_i^{(0)}=e^{i\alpha }\left(D_i^{(0)}+i\sqrt{2}\text{Im}(e^{i\alpha }F_i^{(0)})\right),`$ (32) $`\varphi _{m_i}`$ $``$ $`\widehat{\varphi }_{m_i}={\displaystyle \frac{i}{\sqrt{2}}}(\varphi _{m_i}e^{i\alpha }\stackrel{~}{\varphi }_{m_i}^{}),`$ (33) $`\stackrel{~}{\varphi }_{m_i}^{}`$ $``$ $`\widehat{\stackrel{~}{\varphi }}_{m_i}^{}={\displaystyle \frac{i}{\sqrt{2}}}e^{i\alpha }(\varphi _{m_i}+e^{i\alpha }\stackrel{~}{\varphi }_{m_i}^{}).`$ (34) The tension $`\sigma _{eff}=d^2x_{eff}`$ now reads $`\sigma _{eff}`$ $`=`$ $`{\displaystyle }d^2x[{\displaystyle \frac{1}{2}}b_{ij}^{(0)}(F_{12}^i\pm \widehat{D}_i^{(0)})(F_{12}^j\pm \widehat{D}_j^{(0)})+|(D_1\pm iD_2)\widehat{\varphi }_{m_i}|^2`$ (35) $`+`$ $`|(D_1\pm iD_2)\widehat{\stackrel{~}{\varphi }}_{m_i}|^2+b_{ij}^{(0)}\widehat{F}_i^{(0)}\widehat{F}_j^{(0)}\sqrt{2}F_{12}^i\text{Re}(e^{i\alpha }C_i^{(0)})].`$ (36) The last term breaks explicitely $`SU(2)_R`$ symmetry. Finiteness of the string tension demands regularity of the fields on $`^2`$, and vanishing of the potential energy, field strenghts and covariant derivatives at infinity. Altogether, these requirements make the space of solutions to split into $`^{N1}`$ disconnected pieces that differ by the winding numbers of each $`\varphi _{m_i}`$ over the border of the plane. The electric-fluxes label these sectors. In particular, in the $`(n_1,n_2,\mathrm{},n_{N1})`$–sector they are $$\mathrm{\Phi }_j=d^2xF_{12}^j=2\pi n_j,j=1,2,\mathrm{},N1.$$ (37) The string tension of possible vortex configurations with topologically quantized $`(n_1,n_2)`$ electric flux, exhibits a Bogomol’nyi bound $$\sigma _{eff}4\sqrt{2}\pi \mathrm{\Lambda }\underset{i}{}|(\mu \mathrm{sin}\widehat{\theta }_i+\nu \mathrm{\Lambda }\mathrm{sin}2\widehat{\theta }_i)\mathrm{cos}(\alpha +\beta _i^{(0)})n_i|,$$ (38) which is saturated for configurations solving the following set of first order equations: $`F_{12}^i=\pm \sqrt{2}\text{Re}(e^{i\alpha }F_i^{(0)}),`$ $`D_i^{(0)}+i\sqrt{2}\text{Im}(e^{i\alpha }F_i^{(0)})=0,`$ (39) $`(D_1\pm iD_2)\widehat{\varphi }_{m_i}=0,`$ $`(D_1\pm iD_2)\widehat{\stackrel{~}{\varphi }}_{m_i}=0.`$ (40) The second equation in (39) implies $$|\varphi _{m_j}|=|\stackrel{~}{\varphi }_{m_j}|\text{Im}\left(e^{i\alpha }\left[\sqrt{2}\varphi _{m_j}\stackrel{~}{\varphi }_{m_j}+C_j^{(0)}\right]\right)=0.$$ (41) These constraints should hold at any point, in particular, at zeroes of the Higgs field. Thus $$\varphi _{m_i}=e^{i\beta _i^{(0)}}\stackrel{~}{\varphi }_{m_i}^{}.$$ (42) with $`\alpha +\beta _i^{(0)}=0`$ or $`\pi `$. Consequently, for $`ij`$, $`\beta _{ij}^{(0)}\beta _i^{(0)}\beta _j^{(0)}=0`$ or $`\pi `$. Summarizing, there are no BPS electric vortices in the system unless the complex numbers $`C_i^{(0)}`$ are aligned or anti–aligned. This alignment, in turn, requires supersymmetry breaking parameters to have no relative complex phases. Notice that this corresponds to having a CP invariant bare Lagrangian. For definiteness, in the case of $`SU(3)`$ one easily sees that $$C_1^{(0)}=\sqrt{3}\mathrm{\Lambda }(\mu +\nu \mathrm{\Lambda });C_2^{(0)}=\sqrt{3}\mathrm{\Lambda }(\mu \nu \mathrm{\Lambda }),$$ (43) so that $`\beta _{21}^{(0)}=0`$ or $`\pi `$ if and only if arg$`(\nu \mathrm{\Lambda })=`$ arg$`\mu +n\pi `$ and $`|\nu \mathrm{\Lambda }|<|\mu |`$ or $`|\nu \mathrm{\Lambda }|>|\mu |`$ respectively. A comment is in order at this point regarding the string tensions of unit vortices, whose existence will be discussed below. It is immediate to read, from (38), the string tension of electric vortices carrying a single flux quantum $`n_k=1,n_{ik}=0`$. Up to a common factor, it is given by $$T_k\mathrm{\Lambda }f_N(k)f_N(k)=|\mu \mathrm{sin}\widehat{\theta }_k+\nu \mathrm{\Lambda }\mathrm{sin}2\widehat{\theta }_k|.$$ (44) This result makes clear the dependence of $`f_N(k)`$ on the supersymmetry breaking deformation entering the superpotential. It generalizes previous results in and, in particular, it shows that for perturbations other than the quadratic one, the string tensions are modified with respect to those in the above mentioned results. In particular, notice that when $`\mu `$ and $`\nu `$ do not vanish it is possible to have different string tensions even in the case of $`SU(3)`$ and, in general, $`T_kT_{Nk}`$. ## IV Aligned vacua: critical vortices We will focus hereafter on the case of $`SU(3)`$. When the constants $`\mu `$ and $`\nu `$ are fine tuned in such a way that the phases of the two complex energy scales $`C_1^{(0)}`$ and $`C_2^{(0)}`$ are either aligned or antialigned, i.e. $`\beta _{21}^{(0)}=0`$ or $`\pi `$ respectively, we are at the self dual point. The Bogomol’nyi equations (39)–(40), after (42), read $`F_{12}^i`$ $`=`$ $`\pm {\displaystyle \frac{1}{2}}b_{ij}^{(0)1}ϵ_j(|\phi _j|^2v_j^2),`$ (45) $`(D_1`$ $`\pm `$ $`iϵ_jD_2)\phi _j=0,`$ (46) where $`ϵ_j=e^{i(\alpha +\beta _j^{(0)})}=\pm 1`$ and $`b_{ij}^{(0)}`$ is $$b_{ij}^{(0)}=\left(\begin{array}{cc}g_{D,1}^2& \frac{1}{4\pi ^2}\mathrm{log}2\\ \frac{1}{4\pi ^2}\mathrm{log}2& g_{D,2}^2\end{array}\right),$$ (47) Also, we have performed, for convenience, some redefinitions of the fields, $`\phi _j=2\varphi _{m_j}`$, and parameters, $`v_j^2=2\sqrt{2}|C_j^{(0)}|`$. Let us further remark that Eq.(45) gives an unusual contribution to the electric field of each dual $`U(1)`$ factor from zeroes of both Higgs fields. This is a straight consequence of the presence of off-diagonal couplings and leads to interesting results. It is clear that solutions to (45)–(46) also satisfy the Euler-Lagrange equations. Without loss of generality, we can adjust $`\alpha `$ so that $`ϵ_1=+1,ϵ_2ϵ=e^{i\beta _{21}^{(0)}}=\pm 1`$. Let us focus on the BPS solutions with upper sign. The first order system can be written as $`F_{12}^1`$ $`=`$ $`\lambda _1(|\phi _1|^2v_1^2)ϵ\gamma (|\phi _2|^2v_2^2),`$ (48) $`F_{12}^2`$ $`=`$ $`\gamma (|\phi _1|^2v_1^2)+ϵ\lambda _2(|\phi _2|^2v_2^2),`$ (49) $`(D_1`$ $`+`$ $`iD_2)\phi _1=0,`$ (50) $`(D_1`$ $`+`$ $`iϵD_2)\phi _2=0,`$ (51) with $`\lambda _i`$ $`=`$ $`b_{ii}^{(0)1}=\left(1{\displaystyle \frac{g_{D,1}^2g_{D,2}^2}{16\pi ^2}}\mathrm{log}^22\right)^1{\displaystyle \frac{g_{D,i}^2}{2}}`$ (52) $`\gamma `$ $`=`$ $`b_{12}^{(0)1}={\displaystyle \frac{\mathrm{log}2}{8\pi ^2}}(g_{D,1}^2\lambda _2+g_{D,2}^2\lambda _1).`$ (53) Note that, as we are in the weak $`g_D`$-coupling regime, $`\gamma <\lambda _i`$. Naively one would suspect that in the scaling limit we are interested, the system diagonalizes. Notice however the important fact that the relative factor between $`\lambda _i`$ and $`\gamma `$ vanishes only logarithmically. Hence, for example, setting $`|C_i|/\mathrm{\Lambda }_i^210^{10}`$ in (17) yields $`\gamma (\mathrm{log}2/5)\lambda _i0.13\lambda _i`$. The topology of the configuration space determines global properties of the solutions in two ways: the quantization of the fluxes is due either to the asymptotics of the $`A_j`$ fields or to the existence of a prescribed number of zeroes of the $`\phi _j`$. These global inputs should be made compatible with the differential equations, as it happens in the Abelian Higgs model. In the present situation things are less clear; from Eqs.(50)–(51), where no mixing between both $`U(1)`$s shows up, one reads the electric fluxes using Stokes theorem and the asymptotics of $`A_j`$. On the other hand, Eqs.(48)–(49) mix the factors and both $`\phi _1`$ and $`\phi _2`$ contribute together to each $`F_{12}^i`$. In this respect, our system is quite awkward as compared with other non-diagonal models as, for example, non-relativistic non-abelian Chern-Simons theories , in which the same mixing appears in the field strength and covariant derivative equations. Here, there is mixing in the former but not in the latter, and given such an asymmetry, it is much more difficult to show whether the local equations and the global conditions reconcile or not. On general grounds, it is reasonable to expect that the equations (48)–(51) will exhibit solutions in the topological sector $`(n_1,n_2)`$ with $`n_1,n_2`$ representing the integrated flux of an “ensemble” of noninteracting vortices located at different (maybe coincident) positions. Indeed, the smallness of the ratio $`\gamma /\lambda _i`$ suggests to consider this system as a pertubation of the diagonal situation, so that the above solutions would come out from continuous deformations of the standard critical Abrikosov vortices. Only in some simple cases, the question about the existence of solutions can be answered by taking advantage of known results from the standard Abelian Higgs model. This will be done in the following two situations $``$ Solutions of type $`(n,0)`$ and $`(0,n)`$. Clearly it will be enough to prove existence of one type, say $`(n,0)`$. Assume therefore that $`\phi _2=|\phi _2|e^{i\xi _2}`$ is nowhere vanishing on the finite transverse plane. As usual, (51) couples $`\xi _2`$ and $`A_2`$. So, if $`|\phi _2|`$ has nowhere a zero, regularity of the phase enforces $`A_2`$ to have vanishing circulation around any loop. By Stokes theorem $`F_{12}^2=0`$ everywhere, an inserting this back into (49) yields a constraint that correlates the profiles of $`|\phi _1|`$ and $`|\phi _2|`$, $$|\phi _2|^2=ϵ\frac{\gamma }{\lambda _2}(|\phi _1|^2v_1^2)+v_2^2.$$ (54) Existence of the required vortex profile for $`|\phi _1|`$ can be proved by inserting (54) into $`(\text{48})`$, which leads to the standard Bogomol’nyi equations for the critical Abelian Higgs model (after a suitable re-normalization of the Higgs field) $`F_{12}^1`$ $`=`$ $`\lambda _1\left(1{\displaystyle \frac{\gamma ^2}{\lambda _2}}\right)(|\phi _1|^2v_1^2),`$ (55) $`(D_1`$ $`+`$ $`iD_2)\phi _1=0.`$ (56) We learn from (54) that if $`|\phi _1|^2`$ ranges from $`0`$ (at the origin) up to $`v_1^2`$ (at infinity), $`|\phi _2|^2`$ will correspondingly interpolate between $`ϵ\frac{\gamma }{\lambda _2}v_1^2+v_2^2`$ and $`v_2^2`$. To remain consistent with our initial asumption that $`|\phi _2|`$ vanished nowhere we must set either $`\beta _{21}^{(0)}=0`$ with $`v_2^2>\frac{\gamma }{\lambda _2}v_1^2`$, or else $`ϵ=1`$, i.e. , $`\beta _{21}^{(0)}=\pi `$. We observe that the latter possibility is less contrived. $``$ Solutions of type $`(n,n)`$ for a single perturbation. Let us briefly consider the case of $`SU(3)`$ $`𝒩=2`$ supersymmetric Yang–Mills theory softly broken to $`𝒩=1`$ only by means of a single Casimir operator, i.e. $`\mu =0`$ or $`\nu =0`$. In both cases, $`\beta _{21}^{(0)}=0`$ or $`\pi `$, and the theory is critical. Moreover, $`\lambda _1=\lambda _2\lambda `$, $`C_1^{(0)}=C_2^{(0)}`$ and hence $`v_1=v_2v`$, so that the Bogomol’nyi equations have an almost trivial solution of vorticity $`(n,n)`$ (or $`(n,n)`$), by imposing the ansatz $`\phi _j\phi `$, $`A_jA`$ (or $`\phi _2^{}=\phi _1\phi `$, $`A_2=A_1A`$) in the case $`\beta _{21}^{(0)}=0`$ (or $`\pi `$). The system is again reduced, after a suitable normalization of the Higgs field, to the critical Abelian Higgs model $`F_{12}`$ $`=`$ $`\pm (\lambda ϵ\gamma )(|\phi |^2v^2),`$ (57) $`(D_1`$ $`\pm `$ $`iD_2)\phi =0,ϵ=e^{i\beta _{21}^{(0)}}.`$ (58) It is crucial, for the system to admit regular solutions, that $`\gamma <\lambda `$ as it indeed happens. As it is well known, the general solution to this sytem represents an assembly of $`n`$ separated vortices centered at the zeroes of $`\phi `$. In our case, every such zero is doubled and we have assemblies of $`n`$ couples of superimposed vortices of both $`U(1)`$ fields. Also, self-dual configurations in which the center of the vortices of different types split apart, can be easily constructed along the lines in . To see this, we perturb one of the solutions just described for $`\beta _{21}^{(0)}=0`$ $$\phi _j^{}=\phi _j+\delta \phi _j,A_j^{}=A_j+\delta A_j,$$ (59) and linearize the self-duality equations to get $`4i_z\delta A_12\lambda \phi ^{}\delta \phi _1+2\gamma \phi ^{}\delta \phi _2`$ $`=`$ $`0,`$ (60) $`4i_z\delta A_2+2\gamma \phi ^{}\delta \phi _12\lambda \phi ^{}\delta \phi _2`$ $`=`$ $`0,`$ (61) $`ig_D\phi \delta A_j+(_{\overline{z}}+ig_DA_j)\delta \phi _j`$ $`=`$ $`0,`$ (62) where we use the notation $`_z=\frac{1}{2}(_1i_2)`$, $`A_j=\frac{1}{2}[(A_1)_j+i(A_2)_j],j=1,2`$, and fix the gauge conditions as $`_c(\delta A_c)_1`$ $`=`$ $`\lambda |\phi |^2\delta \mathrm{\Omega }_1+\gamma |\phi |^2\delta \mathrm{\Omega }_2,`$ (63) $`_c(\delta A_c)_2`$ $`=`$ $`\gamma |\phi |^2\delta \mathrm{\Omega }_1\lambda |\phi |^2\delta \mathrm{\Omega }_2.`$ (64) By writing $`\delta \phi _j=\phi \xi _j`$ and using (62), the vector perturbations are found to be $`\delta A_j=\frac{i}{g_D}_{\overline{z}}\xi _j`$ and the system of linearized equations reduces to $$^2W_\pm =2(\lambda \gamma )g_D|\phi |^2W_\pm ,$$ (65) with $`W_\pm =\xi _1\pm \xi _2`$. Notice that in both equations $`(\lambda \gamma )g_D>0`$. Although they have not regular square-integrable solutions, we can admit singular ones provided the singularities of $`\xi _j`$ fit with the zeroes of $`\phi `$ in such a way that $`\delta \phi _j`$ is well-behaved. Take for instance the case of a radially symmetric solution of vorticity $`n`$ centered at the origin of the complex plane. Then, for small $`z`$ $$\phi (z,\overline{z})z^n,$$ (66) and a singularity of $`W_\pm `$ at the origin is harmless if its order is lower or equal than $`n`$. Equation (65) has indeed solutions with such a behaviour . To be exact, two sets of linearly independent self-dual perturbations $`W_\pm ^m(z,\overline{z}),m=1,2,3,\mathrm{}..,n`$ with $$W_\pm ^m(z,\overline{z})z^m,z0.$$ (67) In particular, if we consider $`W_\pm =aW_\pm ^m`$, we get, near the origin, $$\xi _1az^m\xi _20,$$ (68) so that $$\phi _1^{}z^{nm}(z^ma)\phi _2^{}z^n.$$ (69) This perturbation realizes the splitting of a $`(n,n)`$ vortex at the origin into a $`(nm,n)`$ at that point and $`m`$ $`(1,0)`$ vortices located at the $`m`$ roots of the coefficient $`a`$. The analysis for $`\beta _{21}^{(0)}=\pi `$ (i.e. $`ϵ=1`$) is totally equivalent and yields nothing but vortices of type 1 and anti-vortices of type 2 or viceversa, moving freely with respect to each other. For the general analysis, following Jaffe and Taubes , the Higgs fields should be “couched” as $$\phi _jv_je^{\frac{1}{2}(u_j+i\mathrm{\Omega }_j)},$$ (70) to recast the Higgs system in the following form $`^2u_1`$ $`=`$ $`2\lambda _1v_1^2(e^{u_1}1)2ϵ\gamma v_2^2(e^{u_2}1)+\epsilon _{bc}_b_c\mathrm{\Omega }_1,`$ (71) $`^2u_2`$ $`=`$ $`2\gamma v_1^2(e^{u_1}1)+2ϵ\lambda _2v_2^2(e^{u_2}1)+\epsilon _{bc}_b_c\mathrm{\Omega }_2.`$ (72) The gauge fields are determined by $`(A_c)_1`$ $`=`$ $`{\displaystyle \frac{1}{2}}(_c\mathrm{\Omega }_1+\epsilon _{ca}_au_1),`$ (73) $`(A_c)_2`$ $`=`$ $`{\displaystyle \frac{ϵ}{2}}(_c\mathrm{\Omega }_2+\epsilon _{ca}_au_2).`$ (74) At each $`(n_1,n_2)`$ sector, regularity implies that $`\phi _j`$ has exactly $`n_j`$ zeroes on $``$, say $`z_1^j,z_2^j,\mathrm{},z_{n_j}^j`$. Also, these are the only points at which the singularities of the phases can occur. We can then choose the particular gauge $$\mathrm{\Omega }_j(z,\overline{z})=2\underset{l=1}{\overset{n_j}{}}\mathrm{arg}(zz_l^j),$$ (75) in which the problem reduces to $`^2u_1`$ $`=`$ $`2\lambda _1v_1^2(e^{u_1}1)2ϵ\gamma v_2^2(e^{u_2}1)+4\pi {\displaystyle \underset{l=1}{\overset{n_1}{}}}\delta (zz_l^1),`$ (76) $`^2u_2`$ $`=`$ $`2\gamma v_1^2(e^{u_1}1)+2ϵ\lambda _2v_2^2(e^{u_2}1)+4\pi {\displaystyle \underset{l=1}{\overset{n_2}{}}}\delta (zz_l^2),`$ (77) where both $`u_j`$ should vanish at space infinity. The general analysis is involved, and usually goes through by numerical relaxation techniques or hard Sovolev estimates. ## V Hybrid type II vortices By itself, the Abelian Higgs model we are dealing with is worth a detailed analysis. For the moment, and awaiting a sounder analytical or numerical study of its solutions, aside from the two simplified samples considered above little can be said about the generic $`(n_1,n_2)`$ vortex solution. An interesting peculiarity comes from the fact that there are only two overall choices of signs available in equations (45) and (46): either upper or lower sign have to be taken simultaneously on all the equations or, else, the bound (38) will not be saturated. This should be contrasted with the situation in the standard diagonal Abelian Higgs model, where each $`U(1)`$ can be conjugated independently. To better grasp what is going on let us consider the Bogomol’nyi equations (48)-(51) with $`\beta _{21}^{(0)}=0`$ $`F_{12}^1`$ $`=`$ $`\pm (\lambda _1W_1\gamma W_2),`$ (78) $`(D_1`$ $`\pm `$ $`iD_2)\phi _1=0,`$ (79) $`F_{12}^2`$ $`=\text{}`$ $`\pm (\lambda _2W_2\gamma W_1),`$ (80) $`(D_1`$ $`\pm `$ $`iD_2)\phi _2=0,`$ (81) with $`W_i=(|\phi _i|^2v_i^2)`$ . If $`\gamma <<\lambda _1,\lambda _2`$, $`(\pm n_1,\pm n_2)`$ vortex with $`n_1,n_2>0`$ come from solutions to the previous equations with the upper (lower) sign which should correspond to deformations of analogous configurations in the case $`\gamma =0`$. In the diagonal limit $`\gamma =0`$ the vortex-antivortex solutions $`(\pm n_1,n_2)`$ would also solve the previous equations but with a choice of sign for (78)–(79) and the opposite one for (80)–(81). If $`\gamma 0`$, as is now the case, solutions with this second choice of sign do not saturate the bound (38) and, indeed, there is an energy remnant coming from the off-diagonal piece $`=\pi |n_1v_1^2+n_2v_2^2|+\delta `$ $$\delta =d^2x\delta \sigma _{eff}=d^2x\mathrm{\hspace{0.17em}2}b_{12}^{(0)}F_{12}^1F_{12}^2=\frac{\mathrm{log}2}{2\pi ^2}d^2xF_{12}^1F_{12}^2$$ (82) For anti-aligned magnetic fields, this extra term is negative and tends to increase the overlap by attracting the cores of vortices of different kind. A similar reasoning can be carried out of $`\beta _{21}^{(0)}=\pi `$. In this case, the equations read $`F_{12}^1`$ $`=`$ $`\pm (\lambda _1W_1+\gamma W_2),`$ (83) $`(D_1`$ $`\pm `$ $`iD_2)\phi _1=0,`$ (84) $`F_{12}^2`$ $`=\text{}`$ $`(\lambda _2W_2+\gamma W_1),`$ (85) $`(D_1`$ $``$ $`iD_2)\phi _2=0,`$ (86) and critical configurations are naturally of the form $`(\pm n_1,n_2),n_1,n_20`$ saturating the bound $`=\pi |n_1v_1^2n_2v_2^2|`$. Here, in contrast, vortex-vortex solutions of the form $`(\pm n_1,\pm n_2)`$ would lead to the same energy surplus as in (82). But now $`\delta 0`$ for aligned magnetic fields, and this term decreases by minimizing the overlap, hence by taking the cores far appart. In summary, to a first approximation, we see that, if not neutral, vortex-vortex (vortex-antivortex) configurations behave repulsively (attractively) as in type II superconductors. Since this interaction involves vortices of different $`U(1)`$’s, we speak of an “hybrid type II” phase. Let us discuss the peculiarities that arise whenever one tries to model confinement in the present scenario. First we fix some notation for convenience: the chromoelectric fluxes $`(n_1,n_2)`$ of the basic vortices arising in the dual Meissner effect are $`(1,0)`$ (“vortex 1”) and $`(0,1)`$ (“vortex 2”). In turn, quarks enter the system as external probes with chromoelectric charges $`(Q_1,Q_2)`$ equal to $`(1,0)`$ (“red quark”), $`(0,1)`$ (“blue quark”) and $`(1,1)`$ (“yellow quark”). $`(h_1,h_2)`$ is the “monopole” basis of the Cartan algebra of the dual $`S\stackrel{ˇ}{U}(3)`$ group and the fundamental BPS monopoles correspond to the simple co-roots of $`SU(3)`$. In other words, the chromomagnetic charges of the $`\phi _i`$–field quanta is $`h_i=1,h_{ji}=0`$. Consider now, for example, the case $`\beta _{21}^{(0)}=0`$. According to our previous analysis, chromoelectric flux tubes of both $`(1,0)`$ and $`(0,1)`$ type form in response to parallel external electric fields $`\stackrel{}{E}_1`$ and $`\stackrel{}{E}_2`$. Vortices of type 1 end at pairs of red quark-antiquark and vortices of type 2 finish at pairs of blue antiquark-quark. There is therefore confinement of red and blue quarks in a critical phase between Type I and Type II superconductivity, whereas the yellow quark confinement occurs in a hybrid Type II phase. The weak repulsion between the vortex 1/antivortex 2 pair pull slightly apart the flux lines from each other. Thus, the quark/antiquark potential energy would increase slower than linearly with the distance, and one is allowed to expect deviations from the area law, but the force is still confining. If, instead, $`\beta _{21}^{(0)}=\pi `$, a pair of yellow quark-antiquark will now be joined by a stable and non-interacting vortex 1/antivortex 2 pair of flux tubes. In conclusion, the cases $`\beta _{21}^{(0)}=0`$ or $`\pi `$ can be physically distinguished by the behaviour of the yellow quark-antiquark force. At large separation W-pair production leads to instability of the string and the lowest string tension governs the large distance regime . In the framework of condensed matter it is well known the fact that, in standard type II superconductivity on a finite piece of material, though mutually repelling, vortices tend to form a regular pattern by lying at the sites of a triangular lattice. This fact can be reproduced analytically by variational methods . We expect a similar situation here, the difference being that now repulsion involves vortex cores of distinct Higgs fields. Upon substitution of (42) into (28), the exact second order equations with $`\beta _{21}^{(0)}=\pi `$, corresponding to vortices of type 1 and 2, in a finite piece of material $`b_{11}^{(0)}_a(F_{ab})_1+b_{12}^{(0)}_a(F_{ab})_2`$ $`=`$ $`{\displaystyle \frac{i}{2}}(\phi _1^{}D_b\phi _1\phi _1D_b\phi _1^{}),`$ (87) $`b_{22}^{(0)}_a(F_{ab})_2+b_{21}^{(0)}_a(F_{ab})_1`$ $`=`$ $`{\displaystyle \frac{i}{2}}(\phi _2^{}D_b\phi _2\phi _2D_b\phi _2^{}),`$ (88) $`D_cD_c\phi _1`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\phi _1^{}b_{1j}^{(0)1}(|\phi _j|^2v_j^2)(1)^{\beta _{1j}},`$ (89) $`D_cD_c\phi _2`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\phi _2^{}b_{2j}^{(0)1}(|\phi _j|^2v_j^2)(1)^{\beta _{2j}},`$ (90) should now be supplemented with periodic boundary conditions. Thus, the system of differential equations is defined in a torus of modular parameter $`\tau =L_2/L_1e^{i\theta }`$. We have chosen the $`x_1`$-axis as the direction of the first $`L_1`$ periodicity; the length and direction of the second periodicity is determined by $`L_2e^{i\theta }`$. Application of the Rayleigh-Ritz variational method as in plus previous work on the rôle of Riemann Theta functions in magnetic systems , suggest the field configurations $`\phi _1`$ $`=`$ $`{\displaystyle \underset{m_1𝐙}{}}C_{m_1}\mathrm{exp}[in_1m_1\mathrm{Im}z{\displaystyle \frac{1}{2}}(\mathrm{Re}zn_1m_1)^2],`$ (91) $`\phi _2`$ $`=`$ $`{\displaystyle \underset{m_2𝐙}{}}C_{m_2}\mathrm{exp}[in_2m_2\mathrm{Im}z{\displaystyle \frac{1}{2}}(\mathrm{Re}zn_2m_2)^2],`$ (92) where $`n_1`$, $`n_2`$ are integers and $`z=\sqrt{g_D(\lambda \gamma )}\left({\displaystyle \frac{x_1+ix_2}{L_1}}\right)`$, as trial functions to model extremals of the energy. In fact, the choice of the coefficients $`C_{m_1}`$ and $`C_{m_2}`$ in such a way that $`\phi _1^{n_1}(z)`$ $`=`$ $`\mathrm{exp}\{\pi n_1{\displaystyle \frac{(\mathrm{Im}z)^2}{\mathrm{Im}\tau }}\}{\displaystyle \underset{l_1=1}{\overset{n_1}{}}}\mathrm{\Theta }\left[\begin{array}{c}0\\ \frac{l_1}{n_1}\end{array}\right](z|{\displaystyle \frac{\tau }{n_1}}),`$ (95) $`\phi _2^{n_2}(z)`$ $`=`$ $`\mathrm{exp}\{\pi n_2{\displaystyle \frac{(\mathrm{Im}z)^2}{\mathrm{Im}\tau }}\}{\displaystyle \underset{l_2=1}{\overset{n_2}{}}}\mathrm{\Theta }\left[\begin{array}{c}\frac{1}{2}\\ \frac{l_2}{n_2}+\frac{1}{2}\end{array}\right](z|{\displaystyle \frac{\tau }{n_2}}),`$ (98) leads to (meta)-stable solutions to the field equations. Here $`l_i=1,\mathrm{},n_i`$, and $`\mathrm{\Theta }[_b^a](z|\tau )`$ are the Riemann Theta functions with characteristics, see and references quoted therein. Notice that the solution describes $`n_1`$ chromoelectric vortices, located at the zeroes of $`\phi _1^{n_1}`$, and $`n_2`$ vortices of the other kind centered around the zeroes of $`\phi _2^{n_2}`$. It corresponds therefore to a hybrid static triangular lattice of vortices; see the Figure. One can check, from a dynamical point of view that a configuration like this where a vortex of type 1 is at the center of a square with vortices of type 2 at the vertices and viceversa is stable against small fluctuations. ## VI Misaligned vacua As discussed earlier, there are no BPS vortices in the generic case where soft breaking parameters are not aligned. We would however be interested in the response of the BPS configuration when an infinitesimal misalignment $`\beta _{21}^{(0)}\epsilon `$ or $`\beta _{21}^{(0)}\pi +\epsilon `$ is turned on. The dynamics of the system drives the configuration off the constraint (42) which, therefore, can no longer be imposed consistently. In fact, though the Higgs mechanism yields a critical mass spectra for any value of $`\epsilon `$ (an obvious consequence of supersymmetry), the eigenvectors do depend on this phase difference in such a way that when it is different from $`0`$ or $`\pi `$, massive excitations do not respect the constraint surface (42). In the same vein as above, for small values of $`\epsilon `$ we will treat the system as a perturbation of the critical situation in which the net effect of the misalignment reflects itself in a force between the former noninteracting vortices. The shortcut to obtain the sign of this force is to split the energy (36) of the configuration as a BPS contribution plus an additional perturbation. Namely, after inserting the ansatz (42) into (28), solutions to (45)–(46) exhibit a string tension $`\sigma _{eff}=\sigma _{eff}^{SD}+\delta \sigma _{eff}`$ where $`\sigma _{eff}^{SD}`$ is given in (38) and $$\delta \sigma _{eff}=ϵ\frac{\gamma \epsilon ^2}{8}d^2x(|\phi _1|^2v_1^2)(|\phi _2|^2v_2^2).(\epsilon <<1)$$ (99) with $`ϵ=1`$ for $`\beta _{21}^{(0)}=0+\epsilon `$ and $`ϵ=1`$ for $`\beta _{21}^{(0)}=\pi +\epsilon `$. Consider a vortex configuration of type $`(1,1)`$ where the zeroes of each Higgs field are well separated. Then, the above surplus of energy is positive for $`ϵ=1`$ and decreases as the cores are taken further appart and the overlap diminishes hence the interaction in this case is repulsive. When perturbing around the anti-aligned case, $`\beta _{21}^{(0)}=\pi +\epsilon `$, the energy increment (99) reverses sign. Previously non-interacting, $`(1,1)`$ antiparallel vortex configurations tend to increase the overlap in order to lower the perturbation, and hence the force is attractive. In summary, when perturbing around the aligned or misaligned scenarios, the vortex configurations are not neutral anymore, and the interactions follow the pattern that was previously named “hybrid type II” where, if made of distinct U(1)’s, parallel vortices repel and antiparallel vortices attract. ## VII Half Higgsed Vacua As pointed out in , for particular values of the soft breaking parameters $`\mu `$ and $`\nu `$ we have four instead of five vacua. This happens whenever one of the two half Higgsed vacua $`\{a_1^D0,a_2^D=0\}`$ with $`C_1(\mu ,\nu )=0`$, or ($`12`$), meets and replaces one of the normal vacua at $`\{a_1^D=0,a_2^D=0\}`$. This possibility is actually achieved by turning off $`C_i^{(0)}`$ for $`i=1`$ or $`2`$. Since precisely at the $`Z_2`$ point we have (43), this amounts to $`\mu `$ and $`\nu `$ fulfilling $`\mu =\nu \mathrm{\Lambda }`$. Let us choose for definiteness, $`C_2^{(0)}=0`$. Inserting this back into (21), the effective potential at the maximal point reads $$V=\frac{1}{8}\lambda _1(|\phi _1|^2v_1^2)^2+\frac{1}{8}\lambda _2|\phi _2|^4\frac{1}{4}\gamma \mathrm{cos}\beta _1|\phi _2|^2(|\phi _1|^2v_1^2).$$ (100) Observe that the phase of $`\phi _2`$ is free. When $`\mathrm{cos}\beta _1<0`$ this is precisely the type of situation that was studied by Witten and shown to lead to superconducting strings for specific ranges of parameters. Let us briefly recall the essence of the mechanism. As the vacuum equations (23) exhibit, only the first $`U(1)`$ is broken by the v.e.v. $`\phi _1=v_1`$, whereas the second $`U(1)`$ remains intact since $`\phi _2=0`$. This is fine for vacuum solutions, but suppose now that $`\phi _1`$ developes a vortex line. At the core of the vortex $`\phi _1=0`$ and, in turn, it may become favorable that $`\phi _20`$ there. Actually the model considered in is slightly more general than ours involving the potential $$V=\frac{1}{8}g(|\phi _1|^2v^2)^2+\frac{1}{4}\stackrel{~}{g}|\phi _2|^4+f|\phi _1|^2|\phi _2|^2m^2|\phi _2|^2.$$ (101) The detailed analysis of the dynamics showed that for parameters in the range $`fv^2m^2`$, instability actually takes over and the superconducting string indeed forms. We see easily that the present situation lies precisely at the boundary of the region of validity, since in our case $`fv^2m^2=0`$, and the induced mass term for $`\phi _2`$ exactly vanishes. In , this situation was also studied and seen to yield a power law decay of the profile of $`\phi _2`$ which leads to a long range scalar attractive interaction among vortices. At this point we would not like to put forward too strong a claim, but simply point out the ocurrence of this coincidence among models. The possible existence and relevance of structures like superconducting strings in the microscopic context of confinement models should be handled with care. For example the question of quantum tunnelling will be certainly much more relevant here than for cosmic strings. Incidentally this question was also addressed in where it was seen that these power law solutions are more stable than the usual ones. As compared with Witten’s model, the one here involves the additional feature of the non-diagonal kinetic term for the (dual) vector particles (cf. eq. (28)). But precisely the fact that the quadratic forms of kinetic term and potential are related paves the way to the possibility of rewritting the energy as a sum of squares (36). We may therefore expect vortex solutions of the superconducting type with dynamical properties of BPS configurations. We can check that this is indeed the case by looking at the smooth deformation of a generic (anti-)aligned scenario. As we approach the situation when $`C_20`$, the parameters that enter (100) are such that $`\gamma ,\lambda _2<<\lambda _1`$ (see eqns.(17) and (52)–(53)). Hence at very low energy the second $`U(1)`$ seemingly decouples. This is suggested by the $`𝒩=2`$ exact effective solution, although it is reasonable to expect modifications of the renormalization group flow in the $`𝒩=1`$ theory. Let us follow a continuous line of anti-aligned ($`\beta _{21}^{(0)}=\pi `$) vacua $`C_1^{(0)}0,C_2^{(0)}0`$. Precisely in this situation, (54) presents no obstruction to a smooth deformation of the $`(n,0)`$ solutions down to the situation where $`v_2=0`$. In this limit the profiles of $`|\phi _1|`$ and $`|\phi _2|`$ are correlated in such a way that both vanish at opposite ends. In fact, as $`|\phi _1|^2`$ varies from zero up to $`v_1^2`$ far away, $`|\phi _2|`$ interpolates between $`\frac{\gamma }{\lambda _2}v_1^2=(\mathrm{log}2/8\pi ^2)g_{D,1}^2v_1^2`$ at the origin (which need not be small !), and $`0`$ at infinity. Moreover, since the phase of $`\phi _2`$ is free, the same arguments of ref. can be used to show that a persistent current occurs. We would call this a BPS superconducting string solution. ## VIII Concluding Remarks The present paper is devoted to the low energy dynamics of $`𝒩=2`$ supersymmetric gauge theories softly broken to $`𝒩=1`$ by a superpotential containing up to cubic perturbations. The effective lagrangian in the neighborhood of maximal singularities of the quantum moduli space corresponds to an Abelian $`U(1)^{N1}`$ multi Higgs system with couplings among different dual $`U(1)`$ factors. The case of $`SU(3)`$ has been analized in some detail. There are generically no BPS electric vortices in the system unless the soft breaking parameters have coincident complex phases (or they differ by $`\pi `$) and the ultrastrong scaling limit is taken. We have seen that the effect over a BPS configuration of turning on an infinitesimal misalignment among these parameters is the appearance of a net repulsive force between parallel vortices corresponding to (zeroes of) different Higgs fields. In a finite piece of material, metastable solutions take place and vortices develope static triangular lattice. We call this phase “hybrid Type II” dual superconductivity. When the theory is perturbed with a cubic superpotential, the ratio of string tensions differs from that computed in the quadratic case both when the $`\mathrm{Tr}\mathrm{\Phi }^2`$ perturbation is present or not. In the former case, we found that these ratios even depend on the supersymmetry breaking parameters. These results were obtained after imposing the ultrastrong scaling limit. It would be certainly interesting to know if similar results emerge in the context of MQCD. This is intriguing in the sense that string tensions in MQCD are given by the distance of D4-branes which, for a single Casimir perturbation, are stretched at the roots of unity over a circle of radius of order $`\mathrm{\Lambda }`$ , so one would not expect them to be modified (except, possibly, for a global factor due to an induced change in $`\mathrm{\Lambda }`$) as compared to the purely quadratic case. A natural extension of the present work involves the case of $`𝒩=2`$ supersymmetric theories softly broken down to $`𝒩=0`$, and possible soft breaking by higher than the two first Casimir operators. This program can be addressed within the Whitham approach to the Seiberg–Witten solution, where the slow-times of the hierarchy can be used as spurionic sources of soft supersymmetry breaking . ###### Acknowledgements. We are pleased to thank José F. Barbón, A. González-Arroyo, Michael Douglas, Amihay Hanany and Marcos Mariño for interesting discussions. J.M. wants to thank J.J. Blanco Pillado for pointing out reference . The work of J.D.E. has been supported by the National Research Council (CONICET) of Argentina and the Ministry of Education and Culture of Spain. The work of J.M. was partially supported by DGCIYT under contract PB96-0960 and European Union TMR grant ERBFM-RXCT960012.
warning/0001/gr-qc0001047.html
ar5iv
text
# Quiescent cosmological singularities ## 1. Introduction The singularity theorems of Penrose and Hawking are among the best known theoretical results in general relativity. They guarantee the existence of spacetime singularities under rather general circumstances but say little about the structure of the singularities they predict. In the literature there are heuristic approaches to describing the structure of singularities, notably that of Belinskii, Khalatnikov and Lifshitz (BKL), described in , , and their references. The BKL work indicates that generic singularities are oscillatory and therefore, in a certain sense, complicated. This complexity may explain why it has not been possible to determine the structure of the singularities by rigorous mathematical arguments. According to the BKL analysis, the presence of oscillatory behaviour in solutions of the Einstein equations coupled to some matter fields is to a great extent independent of the details of the matter content. There are, however, exceptions. It was pointed out by Belinskii and Khalatnikov that a massless scalar field can change the situation dramatically, producing singularities without oscillations. A massless scalar field is closely related to a stiff fluid, i.e. a perfect fluid with pressure equal to energy density, as will be explained in more detail below. Barrow exploited the singularity structure of solutions of the Einstein equations coupled to a stiff fluid for a description of the early universe he called ‘quiescent cosmology’. We will refer to singularities where oscillatory behaviour is absent due to the matter content of spacetime as quiescent singularities. Recently the Einstein equations coupled to a scalar field have once again been a source of interest, this time in the context of string cosmology. A formal low energy limit of string theory gives rise to the Einstein equations coupled to various matter fields. Under simplifying assumptions the collection of matter fields can be reduced to a single scalar field, the dilaton. The field equations are then equivalent to the standard Einstein-scalar field equations. (Note, however, that the metric occurring in this formulation of the equations is not the physical metric.) The structure of the singularity in these models plays a role in the so-called pre-big bang scenario. For more information on these matters the reader is referred to the work of Buonanno, Damour and Veneziano. In view of the above facts, the Einstein-scalar field equations and, more generally, the Einstein-stiff fluid equations represent an opportunity to prove something about the structure of spacetime singularities in a context simpler than that encountered in the case of the vacuum Einstein equations or the Einstein equations coupled to a perfect fluid with a softer equation of state. In this paper we take this opportunity and prove the existence of a family of solutions of the Einstein-scalar field equations whose singularities can be described in detail and are quiescent. These spacetimes are very general in the sense that no symmetry is assumed and they depend on as many free functions as the general solution of the Einstein-scalar field equations. They have an initial singularity near which they can be approximated by solutions of a simpler system of differential equations, the velocity dominated system. Like the full system, it consists of constraints and evolution equations. The evolution equations contain no spatial derivatives and are thus a system of ordinary differential equations. This is an expression of the idea of BKL that the evolution at different spatial points decouples near the singularity. The structure of the paper is as follows. In the second section we recall the Einstein-scalar field and Einstein-stiff fluid equations and define the corresponding velocity dominated systems. This allows the main theorems to be stated. They assert the existence of a unique solution of the Einstein-scalar field equations or Einstein-stiff fluid equations asymptotic to a given solution of the velocity dominated system. The proofs of the existence and uniqueness theorems are described in the third section. In the fourth section the main analytical tool used in these proofs, the theory of Fuchsian systems, is presented. The algebraic machinery needed for the application of the Fuchsian theory is set up in the fifth section. This provides the basis for the estimates of spatial curvature and other important quantities in the section which follows. The seventh section treats relevant aspects of the constraints. The paper concludes with a discussion of what can be learned from the results of the paper and what generalizations are desirable. Throughout the paper the scalar field and stiff fluid cases are treated in parallel. These are independent except in section 7 where the propagation of constraints for the scalar field is deduced from the corresponding statement for the stiff fluid. Hence concentrating on the scalar field case on a first reading would give a good idea of the main features of the proofs. ## 2. The main results Let $`{}_{}{}^{(4)}g_{\alpha \beta }^{}`$ be a Lorentz metric on a four-dimensional manifold $`M`$ which is diffeomorphic to $`(0,T)\times S`$ for a three-dimensional manifold $`S`$. Let a point of $`M`$ be denoted by $`(t,x)`$, where $`t(0,T)`$ and $`xS`$. It will be important in the following to express the geometrical quantities of interest in terms of a local frame $`\{e_a\}`$ on $`S`$. Let $`\{\theta ^a\}`$ denote the coframe dual to $`\{e_a\}`$. Throughout the paper lower case Latin indices refer to components in this frame, except where other conventions are introduced explicitly. Suppose that the metric takes the form: (2.1) $$dt^2+g_{ab}(t)\theta ^a\theta ^b$$ where $`g_{ab}(t)`$ denotes the one-parameter family of Riemannian metrics on $`S`$ defined by the metrics induced on the hypersurfaces $`t=`$constant by the metric $`{}_{}{}^{(4)}g_{\alpha \beta }^{}`$. A function $`t`$ such that the metric takes the form (2.1) is called a Gaussian time coordinate. In this case, the second fundamental form of a hypersurface $`t=`$constant is given by $`k_{ab}=\frac{1}{2}_tg_{ab}`$. ### 2.1. The Einstein-matter equations The Einstein field equations coupled to matter can be written in the following equivalent $`3+1`$ form. The constraints are: (2.2a) $`Rk_{ab}k^{ab}+(\text{tr}k)^2`$ $`=16\pi \rho `$ (2.2b) $`^ak_{ab}e_b(\text{tr}k)`$ $`=8\pi j_b`$ The evolution equations are: (2.3a) $`_tg_{ab}`$ $`=2k_{ab}`$ (2.3b) $`_tk^a_b`$ $`=R^a{}_{b}{}^{}+(\text{tr}k)k^a{}_{b}{}^{}8\pi (S_b^a{\displaystyle \frac{1}{2}}\delta _b^a\text{tr}S)4\pi \rho \delta _b^a`$ Here $`R`$ is the scalar curvature of $`g_{ab}`$ and $`R_{ab}`$ its Ricci tensor. The quantities $`\rho `$, $`j_a`$ and $`S_{ab}`$ are projections of the energy-momentum tensor. Their explicit forms in the cases of interest in this paper will be given below. The energy-momentum tensor of a scalar field is given by (2.4) $$T_{\alpha \beta }=_\alpha \varphi _\beta \varphi \frac{1}{2}(_\gamma \varphi ^\gamma \varphi ){}_{}{}^{(4)}g_{\alpha \beta }^{}$$ The Einstein equations can be written in the equivalent form $`{}_{}{}^{(4)}R_{\alpha \beta }^{}=8\pi _\alpha \varphi _\beta \varphi `$, where $`{}_{}{}^{(4)}R_{\alpha \beta }^{}`$ is the Ricci tensor of $`{}_{}{}^{(4)}g_{\alpha \beta }^{}`$. We have $`\rho =T_{00}`$, $`j_a=T_{0a}`$ and $`S_{ab}=T_{ab}`$ so that in the case of a scalar field it follows from (2.4) that: (2.5a) $`\rho `$ $`={\displaystyle \frac{1}{2}}[(_t\varphi )^2+g^{ab}e_a(\varphi )e_b(\varphi )]`$ (2.5b) $`j_b`$ $`=_t\varphi e_b(\varphi )`$ (2.5c) $`S_{ab}`$ $`=e_a(\varphi )e_b(\varphi )+{\displaystyle \frac{1}{2}}[(_t\varphi )^2g^{cd}e_c(\varphi )e_d(\varphi )]g_{ab}`$ Note that it follows from the Einstein-scalar field equations as a consequence of the Bianchi identity that $`\varphi `$ satisfies the wave equation $`{}_{}{}^{(4)}g_{}^{\alpha \beta }_\alpha _\beta \varphi =0`$. This has the $`3+1`$ form: (2.6) $$_t^2\varphi +(\text{tr}k)_t\varphi +\mathrm{\Delta }\varphi =0$$ The constraints and evolution equations are together equivalent to the full Einstein-scalar field equations. In the following we will work with the $`3+1`$ formulation of the equations rather than the four-dimensional formulation. A stiff fluid is a perfect fluid with pressure equal to energy density. As we will see, it is closely related to the scalar field. The energy-momentum tensor of a stiff fluid is (2.7) $$T_{\alpha \beta }=\mu (2u_\alpha u_\beta +{}_{}{}^{(4)}g_{\alpha \beta }^{})$$ where $`\mu `$ is the energy density of the fluid in a comoving frame and $`u^\alpha `$ is the four-velocity. The Euler equations are obtained by substituting this expression into the equation $`_\alpha T^{\alpha \beta }=0`$. The relation between the scalar field and the stiff fluid is as follows. Given a solution of the Einstein equations coupled to a scalar field, where the gradient of the scalar field $`\varphi `$ is everywhere timelike, define $`\mu =(1/2)_\alpha \varphi ^\alpha \varphi `$ and $`u^\alpha =\pm (_\beta \varphi ^\beta \varphi )^{1/2}^\alpha \varphi `$. Here the sign is chosen so that $`u^\alpha `$ is future pointing, and so can be interpreted as the four-velocity of a fluid. Then the energy-momentum tensor defined by (2.7) is equal to the energy-momentum tensor of the scalar field. Since the latter is divergence-free we see that the fluid variables just defined together with the original metric define a solution of the Einstein equations coupled to a stiff fluid. In the spacetimes of interest in the following, the gradient of $`\varphi `$ is always timelike near the singularity, so that the condition on the gradient is not a restriction in that situation. The matter terms needed for the Einstein equations are given in the stiff fluid case by (2.8a) $`\rho `$ $`=\mu (1+2|u|^2)`$ (2.8b) $`j_b`$ $`=2\mu (1+|u|^2)^{1/2}u_b`$ (2.8c) $`S_{ab}`$ $`=\mu (2u_au_b+g_{ab})`$ Here $`|u|^2=g_{ab}u^au^b`$. The Euler equations can be written in the following 3+1 form: $`_t\mu 2(\text{tr}k)\mu =2|u|^2_t\mu 4\mu u^a_tu_a2\mu k_{ab}u^au^b`$ $`+2(\text{tr}k)\mu |u|^22e_a(\mu )(1+|u|^2)^{1/2}u^a\mu (1+|u|^2)^{1/2}u^a_au_bu^b`$ (2.9a) $`2\mu (1+|u|^2)^{1/2}_au^a`$ $`_tu_a+(\text{tr}k)u_a=\mu ^1[_t\mu 2(\text{tr}k)\mu ]u_a+(1+|u|^2)^1u^bu_a_tu_b`$ (2.9b) $`(1+|u|^2)^{1/2}[(u_au^b+(1/2)\delta _a^b)\mu ^1e_b(\mu )+(_cu^cu_a+u^c_cu_a)]`$ ### 2.2. The velocity dominated system We will prove the existence of a large class of solutions of the Einstein-scalar field equations and Einstein-stiff fluid equations whose singularities we can describe in great detail. These singularities are of the type known as velocity dominated. (Cf. and .) This means that near the singularity the solution can be approximated by a solution of a simpler system, the velocity dominated system. Like the 3+1 version of the full equations it consists of constraints and evolution equations. Solutions of the velocity dominated system will always be written with a left superscript zero and the convention is adopted that the indices of all quantities with this superscript are moved with the velocity dominated metric $`{}_{}{}^{0}g_{cd}^{}`$. The velocity dominated equations will now be written out explicitly. The constraints are: (2.10a) $`{}_{}{}^{0}k_{ab}^{}{}_{}{}^{0}k_{}^{ab}+(\text{tr}{}_{}{}^{0}k)^2`$ $`=16\pi {}_{}{}^{0}\rho `$ (2.10b) $`^a({}_{}{}^{0}k_{ab}^{})e_b(\text{tr}{}_{}{}^{0}k)`$ $`=8\pi {}_{}{}^{0}j_{b}^{}`$ The evolution equations are: (2.11a) $`_t{}_{}{}^{0}g_{ab}^{}`$ $`=2{}_{}{}^{0}k_{ab}^{}`$ (2.11b) $`_t{}_{}{}^{0}k_{b}^{a}`$ $`=(\text{tr}{}_{}{}^{0}k){}_{}{}^{0}k_{b}^{a}8\pi ({}_{}{}^{0}S_{b}^{a}{\displaystyle \frac{1}{2}}\delta _b^a\text{tr}{}_{}{}^{0}S)4\pi ^0\rho \delta _b^a`$ In these equations the matter terms are not identical to those in the full equations but have been obtained from those by discarding certain terms. In the case of a scalar field the (truncated) energy-momentum tensor components are given by (2.12a) $`{}_{}{}^{0}\rho `$ $`={\displaystyle \frac{1}{2}}(_t{}_{}{}^{0}\varphi )^2`$ (2.12b) $`{}_{}{}^{0}j_{b}^{}`$ $`=_t{}_{}{}^{0}\varphi e_b({}_{}{}^{0}\varphi )`$ (2.12c) $`{}_{}{}^{0}S_{ab}^{}`$ $`={\displaystyle \frac{1}{2}}(_t{}_{}{}^{0}\varphi )^2({}_{}{}^{0}g_{ab}^{})`$ The scalar field satisfies the equation $$_t^2({}_{}{}^{0}\varphi )+(\text{tr}{}_{}{}^{0}k)_t{}_{}{}^{0}\varphi =0$$ It is important to note that the velocity dominated evolution equations are ordinary differential equations. However the constraints still include partial differential equations. In the stiff fluid case the (truncated) energy-momentum tensor components are (2.13a) $`{}_{}{}^{0}\rho `$ $`={}_{}{}^{0}\mu `$ (2.13b) $`{}_{}{}^{0}j_{b}^{}`$ $`=2{}_{}{}^{0}\mu {}_{}{}^{0}u_{b}^{}`$ (2.13c) $`{}_{}{}^{0}S_{ab}^{}`$ $`={}_{}{}^{0}\mu {}_{}{}^{0}g_{ab}^{}`$ The velocity dominated Euler equations are (2.14a) $`_t{}_{}{}^{0}\mu 2(\text{tr}{}_{}{}^{0}k){}_{}{}^{0}\mu `$ $`=0`$ (2.14b) $`_t{}_{}{}^{0}u_{a}^{}+(\text{tr}{}_{}{}^{0}k){}_{}{}^{0}u_{a}^{}`$ $`=(1/2){}_{}{}^{0}\mu _{}^{1}e_a({}_{}{}^{0}\mu )`$ In contrast to the scalar field case, this is not a system of ordinary differential equations. However it has a hierarchical ODE structure in the sense that if the ODE for $`{}_{}{}^{0}\mu `$ is solved and the result substituted into the other equations an ODE system for the $`{}_{}{}^{0}u_{a}^{}`$ results. Substituting the expressions for the truncated energy-momentum tensor into the velocity dominated system shows that the matter terms in the velocity dominated evolution equation for $`{}_{}{}^{0}k_{ab}^{}`$ cancel both for the scalar field and the stiff fluid, leaving $$_t{}_{}{}^{0}k_{b}^{a}=(\text{tr}{}_{}{}^{0}k){}_{}{}^{0}k_{b}^{a}$$ Taking the trace of this equation gives $`_t(\text{tr}{}_{}{}^{0}k)=(\text{tr}{}_{}{}^{0}k)^2`$. This has the general solution $`\text{tr}{}_{}{}^{0}k=(Ct)^1`$. If we wish an initial singularity, as signalled by the blow-up of $`\text{tr}{}_{}{}^{0}k`$, to occur at $`t=0`$ then $`\text{tr}{}_{}{}^{0}k=t^1`$. Going back to the equation for $`{}_{}{}^{0}k_{b}^{a}`$ we see that $`t({}_{}{}^{0}k_{b}^{a})`$ is independent of time. Thus all components of the mixed form of the second fundamental form are proportional to $`t^1`$. At any given spatial point we can simultaneously diagonalize $`{}_{}{}^{0}g_{ab}^{}`$ and $`{}_{}{}^{0}k_{ab}^{}`$ by a suitable choice of frame. The matrix of components of the metric in this frame is diagonal with the diagonal elements being proportional to powers of $`t`$. This form of the metric is that originally used by BKL. Its disadvantage is that in general this frame cannot be chosen to depend smoothly on the spatial point. (There are difficulties when there are changes in the multiplicity of the eigenvalues of $`{}_{}{}^{0}k_{b}^{a}`$.) This is one reason why a different formulation is used in this paper. The velocity dominated matter equations can be solved exactly to give (2.15) $${}_{}{}^{0}\varphi (t,x)=A(x)\mathrm{log}t+B(x)$$ for given functions $`A`$ and $`B`$ on $`S`$ and (2.16) $`{}_{}{}^{0}\mu (t,x)`$ $`=A^2(x)t^2`$ (2.17) $`{}_{}{}^{0}u_{a}^{}(t,x)`$ $`=t\mathrm{log}t(A(x))^1e_a(A(x))+tB_a(x)`$ for given quantities $`A(x)`$ and $`B_a(x)`$ on $`S`$. ### 2.3. Statement of the main theorems The main theorems can now be stated. ###### Theorem 2.1. Let $`S`$ be a three-dimensional analytic manifold and let $`({}_{}{}^{0}g_{ab}^{}(t),{}_{}{}^{0}k_{ab}^{}(t),{}_{}{}^{0}\varphi (t))`$ be a $`C^\omega `$ solution of the velocity dominated Einstein-scalar field equations on $`S\times (0,\mathrm{})`$ such that $`t\text{tr}{}_{}{}^{0}k=1`$ and each eigenvalue $`\lambda `$ of $`t{}_{}{}^{0}k_{b}^{a}`$ is positive. Then there exists an open neighbourhood $`U`$ of $`S\times \{0\}`$ in $`S\times [0,\mathrm{})`$ and a unique $`C^\omega `$ solution $`(g_{ab}(t),k_{ab}(t),\varphi (t))`$ of the Einstein-scalar field equations on $`U(S\times (0,\mathrm{}))`$ such that for each compact subset $`KS`$ there are positive real numbers $`\zeta ,\beta ,\alpha _b^a`$, with $`\zeta <\beta <\alpha _b^a`$, for which the following estimates hold uniformly on $`K`$: 1. $`{}_{}{}^{0}g_{}^{ac}g_{cb}=\delta _b^a+o(t^{\alpha _b^a})`$ 2. $`k_b^a={}_{}{}^{0}k_{b}^{a}+o(t^{1+\alpha _b^a})`$ 3. $`\varphi ={}_{}{}^{0}\varphi +o(t^\beta )`$ 4. $`_t\varphi =_t{}_{}{}^{0}\varphi +o(t^{1+\beta })`$ 5. $`{}_{}{}^{0}g_{}^{ac}e_f(g_{cb})=o(t^{\alpha _b^a\zeta })`$ 6. $`e_a(\varphi )=e_a({}_{}{}^{0}\varphi )+o(t^{\beta \zeta })`$ Note that the condition on $`\text{tr}{}_{}{}^{0}k`$ can always be arranged by means of a time translation and that the condition on the eigenvalues of $`{}_{}{}^{0}k_{b}^{a}`$ is satisfied provided it holds for a single value of $`t>0`$. The positivity condition on the eigenvalues together with the velocity dominated Hamiltonian constraint imply that $`A^2`$ must be strictly positive in the velocity dominated solution. Thus vacuum solutions are ruled out by the hypotheses of this theorem. If an analogous analysis were done for the Einstein equations coupled to other matter models, for instance a perfect fluid with equation of state $`p=k\rho `$, $`k<1`$, then in many cases, including that of the fluid just mentioned, the matter would make no contribution to the velocity dominated Hamiltonian constraint, so that it would not be possible to prove an analogous theorem. This reflects the fact that for those matter models an oscillatory approach to the singularity is predicted by the BKL analysis. The scalar field is an exception, as is the stiff fluid which will be discussed next. ###### Theorem 2.2. Let $`S`$ be a three-dimensional analytic manifold and let $`({}_{}{}^{0}g_{ab}^{}(t),{}_{}{}^{0}k_{ab}^{}(t),{}_{}{}^{0}\mu (t),{}_{}{}^{0}u_{a}^{})`$ be a $`C^\omega `$ solution of the velocity dominated Einstein-stiff fluid equations on $`S\times (0,\mathrm{})`$ such that $`t\text{tr}{}_{}{}^{0}k=1`$ and each eigenvalue $`\lambda `$ of $`t{}_{}{}^{0}k_{b}^{a}`$ is positive. Then there exists an open neighbourhood $`U`$ of $`S\times \{0\}`$ in $`S\times [0,\mathrm{})`$ and and a unique $`C^\omega `$ solution $`(g_{ab}(t),k_{ab}(t),\mu (t),u_a(t))`$ of the Einstein-stiff fluid equations on $`U(S\times (0,\mathrm{}))`$ such that for each compact subset $`KS`$ there are positive real numbers $`\zeta ,\beta _1,\beta _2,\alpha _b^a`$, with $`\zeta <\beta _2<\beta _1<\alpha _b^a`$, for which the following estimates hold uniformly on $`K`$: 1. $`{}_{}{}^{0}g_{}^{ac}g_{cb}=\delta _b^a+o(t^{\alpha _b^a})`$ 2. $`k_b^a={}_{}{}^{0}k_{b}^{a}+o(t^{1+\alpha _b^a})`$ 3. $`\mu ={}_{}{}^{0}\mu +o(t^{2+\beta _1})`$ 4. $`u_a={}_{}{}^{0}u_{a}^{}+o(t^{1+\beta _2})`$ 5. $`{}_{}{}^{0}g_{}^{ac}e_f(g_{cb})=o(t^{\alpha _b^a\zeta })`$ The interest of these theorems depends very much on what information is available on constructing solutions of the velocity dominated system. Suppose that a solution of the velocity dominated constraints is given for some $`t=t_0>0`$. The velocity dominated evolution equations constitute a system of ordinary differential equations which can be solved with these initial data. It follows from the remarks above that the solution exists globally on the interval $`(0,\mathrm{})`$. If we define (2.18) $`{}_{}{}^{0}C`$ $`={}_{}{}^{0}k_{ab}^{}{}_{}{}^{0}k_{}^{ab}+(\text{tr}{}_{}{}^{0}k)^216\pi {}_{}{}^{0}\rho `$ (2.19) $`{}_{}{}^{0}C_{b}^{}`$ $`=^a({}_{}{}^{0}k_{ab}^{})e_b(\text{tr}{}_{}{}^{0}k)8\pi {}_{}{}^{0}j_{b}^{}`$ then the velocity dominated evolution equations imply that: (2.20) $`_t{}_{}{}^{0}C+2t^1({}_{}{}^{0}C)`$ $`=0`$ (2.21) $`_t{}_{}{}^{0}C_{a}^{}+t^1({}_{}{}^{0}C_{a}^{})`$ $`={\displaystyle \frac{1}{2}}e_a({}_{}{}^{0}C)`$ To prove this it is necessary to use the following equations for the matter quantities, which can be derived from the velocity dominated matter equations in both the scalar field and stiff fluid cases. (2.22) $`_t{}_{}{}^{0}\rho `$ $`=2(\text{tr}{}_{}{}^{0}k){}_{}{}^{0}\rho `$ (2.23) $`_t{}_{}{}^{0}j_{a}^{}`$ $`=(\text{tr}{}_{}{}^{0}k){}_{}{}^{0}j_{a}^{}e_a({}_{}{}^{0}\rho )`$ Since $`{}_{}{}^{0}C`$ vanishes at $`t=t_0`$ the evolution equation for $`{}_{}{}^{0}C`$ implies that it vanishes everywhere. Then the evolution equation for $`{}_{}{}^{0}C_{a}^{}`$, together with the fact that it vanishes for $`t=t_0`$ implies that $`{}_{}{}^{0}C_{a}^{}`$ vanishes everywhere. To sum up, if the velocity dominated constraints are satisfied at some time and the velocity dominated evolution equations are satisfied everywhere then the velocity dominated constraints are satisfied everywhere. Thus in order to have a parametrization of the general solution of the velocity dominated system, it is enough to obtain a parametrization of solutions of the velocity dominated constraints. The latter question will be treated in section 7. ## 3. Framework of the proofs In this section the proofs of Theorems 2.1 and 2.2 are outlined. Only the general logical stucture of the proof is explained here and the hard technical parts of the argument are left to later sections. These results will be referred to as required in this section. The first step is to make a suitable ansatz for the desired solution. This essentially means giving names to the remainder terms occurring in the statements of the main theorems. Assume that a velocity dominated solution is given as in those statements. Then a solution is sought in the form: (3.1a) $`g_{ab}`$ $`={}_{}{}^{0}g_{ab}^{}+{}_{}{}^{0}g_{ac}^{}t^{\alpha _b^c}\gamma _b^c`$ (3.1b) $`k_{ab}`$ $`=g_{ac}({}_{}{}^{0}k_{b}^{c}+t^{1+\alpha _b^c}\kappa _b^c)`$ In the following the summation convention applies only to repeated tensor indices and not to non-tensorial quantities like $`\alpha _b^a`$. Thus in the above equations there is a summation on the index $`c`$ but none on the index $`b`$. Matter fields are sought in the form: (3.2) $$\varphi ={}_{}{}^{0}\varphi +t^\beta \psi $$ and (3.3a) $`\mu `$ $`={}_{}{}^{0}\mu +t^{2+\beta _1}\nu `$ (3.3b) $`u_a`$ $`={}_{}{}^{0}u_{a}^{}+t^{1+\beta _2}v_a`$ respectively. The Einstein-scalar field equations (2.2), (2.3), (2.5) and (2.6) can be rewritten as equations for $`\gamma _b^a`$, $`\kappa _b^a`$ and $`\psi `$. Similarly the Einstein-stiff fluid equations can be written as equations for $`\gamma _b^a`$, $`\kappa _b^a`$, $`\nu `$ and $`v_a`$. This system of equations (for either choice of matter model) will be called the first reduced system. Since $`\gamma _b^a`$ and $`\kappa _b^a`$ are mixed tensors, there is no direct way to express the fact that they originated from symmetric tensors. Instead it must be shown that when the first reduced system is solved with suitable asymptotic conditions as $`t0`$ then the quantities $`g_{ab}`$ and $`k_{ab}`$ defined by the above equations are in fact symmetric as a consequence of the differential equations and the initial conditions. When allowing non-symmetric tensors $`g_{ab}`$ and $`k_{ab}`$, we need to establish some conventions in order to make the definition of the first reduced system unambiguous. Firstly, define $`g^{bc}`$ as the unique tensor which satifies $`g_{ab}g^{bc}=\delta _a^c`$. Next, use the convention that indices on tensors are lowered by contraction with the second index of $`g_{ab}`$ and raised with the first index of $`g^{bc}`$. This maintains the usual properties of index manipulations in the case of symmetric $`g_{ab}`$ as far as possible. The covariant derivatives in the equations are expressed in terms of the connection coefficients in the frame $`\{e_a\}`$ in an unambiguous way. The definition of the connection coefficients is extended to the case of a non-symmetric tensor $`g_{ab}`$ by fixing the order of indices according to (3.4) $$g_{cd}\mathrm{\Gamma }_{ab}^d=\frac{1}{2}(e_a(g_{bc})+e_b(g_{ac})e_c(g_{ab})+\gamma _{ab}^dg_{cd}\gamma _{ac}^dg_{bd}\gamma _{bc}^dg_{ad})$$ where $`\gamma _{ab}^c=\theta ^c([e_a,e_b])`$ are the structure functions of the frame. Finally, in order to define the Ricci tensor in the evolution equation for $`k_{ab}`$ we define $`R_{ab}`$ to be the Ricci tensor of the symmetric part $`{}_{}{}^{S}g_{ab}^{}=(1/2)(g_{ab}+g_{ba})`$ of $`g_{ab}`$. In Lemma 5.2 it is shown that given a solution of the velocity dominated system as in the statement of one of the main theorems 2.12.2, any solution of the first reduced system which satisfies points 16 of Theorem 2.1, in the scalar field case, and points 15 of Theorem 2.2, in the stiff fluid case, gives rise to symmetric tensors $`g_{ab}`$ and $`k_{ab}`$, and thus to a solution of the Einstein-matter evolution equations. It then follows from Lemma 7.1 and the remarks following it that the Einstein-matter constraints are also satisfied. It follows that to prove the main theorems it is enough to prove the existence and uniqueness of solutions of the first reduced system of the form given in the main theorems. The existence theorem for the first reduced system will be proved using the theory of Fuchsian systems. Since the form of this theory we use in the following concerns a system of first order equations it is not immediately applicable, due to the occurrence of second order derivatives of the metric and scalar field. It is necessary to introduce some suitable new variables representing spatial derivatives of the basic variables. Define $`\lambda _{bc}^a=t^\zeta e_c(\gamma _b^a)`$ and, in the scalar field case, $`\omega _a=t^\zeta e_a(\psi )`$ and $`\chi =t_t\psi +\beta \psi `$, where $`\zeta `$ and $`\beta `$ are positive constants. The evolution equations satisfied by these quantities are given explicitly in (5.13b) and (5.15). With the help of the new variables these equations together with the first reduced system can be written as a first order system. Call the result the second reduced system. It is easy to show, using the evolution equations for differences like $`\gamma _{bc}^at^\zeta e_c(\gamma _b^a)`$ which follow from the second reduced system, that the first and second reduced systems give rise to the same sets of solutions under the assumptions of the main theorems, together with corresponding assumptions on the new variables. Thus it suffices to solve the second reduced system, which is of first order. If it can be shown that the second reduced system is Fuchsian, then the main theorems follow from Theorem 4.2. In fact it is enough to show that the restriction of the system to a neighbourhood of an arbitrary point of $`S`$ is Fuchsian. For the local solutions thus obtained can be pieced together to get a global solution. Moreover asymptotic estimates as in the statements of the main theorems follow from corresponding local statements, since a compact subset of $`S`$ can be covered by finitely many of the local neighbourhoods. In sections 5 and 6 it is shown that the second reduced system is Fuchsian on some neighbourhood of each point of $`S`$ for a suitable choice of the constants $`\alpha _b^a`$, $`\beta `$, $`\beta _1`$, $`\beta _2`$ and $`\zeta `$ depending on the given velocity dominated solution. This requires a detailed analysis of the degree of singularity of all terms in the second reduced system, in particular that of the Ricci tensor. The result of the latter is Lemma 6.3. This is the hardest part of the proof. Due the above considerations, it is clear that the main theorems follow directly from Theorem 4.2. ## 4. Fuchsian systems The proofs of Theorems 2.1 and 2.2 rely on a result of Kichenassamy and Rendall on Fuchsian systems which uses a method going back to Baouendi and Goulaouic. The result of will now be recalled. It concerns a system of the form: (4.1) $$t\frac{u}{t}+A(x)u=f(t,x,u,u_x)$$ Here $`u(t,x)`$ is a function on an open subset of $`𝐑\times 𝐑^n`$ with values in $`𝐑^k`$ and $`A(x)`$ is a $`C^\omega `$ matrix-valued function. The derivatives of $`u`$ with respect to the $`x`$ variables are denoted by $`u_x`$. The function $`f`$ is defined on $`(0,T_0]\times U_1\times U_2`$, where $`U_1`$ is an open subset of $`𝐑^n`$ and $`U_2`$ is an open subset of $`𝐑^{k+nk}`$, and takes values in $`𝐑^k`$. We assume that $`A(x)`$ is defined on $`U_1`$. In this and later sections it will be useful to have some terminology for comparing the sizes of certain expressions. ###### Definition 4.1. Let $`F(t,x,p)`$, $`G(t,x,p)`$ be functions on $`(0,T_0]\times U_1\times U_2`$, where $`U_1,U_2`$ are open subsets of $`𝐑^n`$ and $`𝐑^N`$ respectively. Then we will say that $$FG$$ if for every compact $`KU_1\times U_2`$, there is a constant $`C`$ such that $$|F(t,x,p)|C|G(t,x,p)|\text{for }t(0,t_0]\text{}(x,p)K.$$ In the particular case that $`G`$ is just a function of $`t`$ we will often use the familiar notation $`F=O(G(t))`$ to replace $`FG`$. The notation $`F=o(G(t))`$ will also be used to indicate that $`F/G`$ tends to zero uniformly on compact subsets of $`U_1\times U_2`$ as $`t0`$. In the theorem on Fuchsian systems the function $`f`$ is supposed to be regular in a sense which will now be explained. To do this we need the notion of a function which is continuous in $`t`$ and analytic in other (complex) variables. This means by definition that it should be a continuous function of all variables, that the first order partial derivatives with respect to all variables other than $`t`$ should exist and be continuous, and that the Cauchy-Riemann equations should be satisfied in these variables. For further remarks on this concept see . Assume that there is an open subset $`\stackrel{~}{U}`$ of $`𝐂^{n+k+nk}`$ whose intersection with the real section is equal to $`U_1\times U_2`$ and a function $`\stackrel{~}{f}`$ on $`(0,T_0]\times \stackrel{~}{U}`$ continuous in $`t`$ and analytic in the remaining arguments whose restriction to $`(0,T_0]\times 𝐑^{n+k+nk}`$ is equal to $`f`$. The function $`f`$ is called regular if it has an analytic continuation $`\stackrel{~}{f}`$ of the kind just described, and if there is some $`\theta >0`$ such that $`\stackrel{~}{f}`$ and its first derivatives with respect to the arguments $`u`$ and $`u_x`$ are $`O(t^\theta )`$ as $`t0`$, in the sense introduced above. For a matrix $`A`$ with entries $`A_b^a`$ let $`A=sup\{Ax:x=1\}`$ (operator norm) and $`A_{\mathrm{}}=\mathrm{max}_{a,b}|A_b^a|`$ (maximum norm). Since these two norms are equivalent the operator norm could be replaced by the maximum norm in the statement of the theorem which follows and in Lemma 4.3 below. However in the proof of that lemma below the use of the operator norm is important. ###### Theorem 4.2. Suppose that the function $`f`$ is regular, $`A(x)`$ has an analytic continuation to an open set $`\stackrel{~}{U}_1`$ whose intersection with the real section is $`U_1`$, and there is a constant $`C`$ such that $`\sigma ^{A(x)}C`$ for $`x\stackrel{~}{U}_1`$ and $`0<\sigma <1`$. Then the equation (4.1) has a unique solution $`u`$ defined near $`t=0`$ which is continuous in $`t`$ and analytic in $`x`$ and tends to zero as $`t0`$. If $`\stackrel{~}{f}`$ is analytic for $`t>0`$ then this solution is also analytic in $`t`$ for $`t>0`$. ###### Remark 4.1. Under the hypotheses of the theorem spatial derivatives of any order of $`u`$ are also $`o(1)`$ as $`t0`$. This follows directly from the proof of the theorem in . ###### Remark 4.2. If the coefficients of the equation depend analytically on a parameter and are suitably regular, then the solution depends analytically on the parameter. It suffices to treat the parameter as an additional spatial variable. The statement of the theorem is not identical to that given in but the proof is just the same. We can write $`f(t,x,u,u_x)=t^\theta g(t,x,u,u_x)`$ for some bounded function $`g`$. By replacing $`t`$ as time variable by $`t^\theta `$, it can be assumed without loss of generality that $`\theta =1`$. Then the iteration used in the proof given in converges to the desired solution under the regularity hypothesis we have made on $`f`$. For the applications in this paper an extension of this result which applies to equations slightly more general than (4.1) will be required. These have the form (4.2) $$t\frac{u}{t}+A(x)u=f(t,x,u,u_x)+g(t,x,u)t\frac{u}{t}$$ If we have an equation of this form where $`f`$ and $`g`$ are regular then an analogous existence and uniqueness result holds. For we can rewrite (4.2) in the form (4.3) $$t\frac{u}{t}+A(x)u=[I(Ig(t,x,u))^1]A(x)+(Ig(t,x,u))^1f(t,x,u,u_x)$$ If we call the right hand side of this equation $`h(t,x,u,u_x)`$ then it satisfies the conditions required of $`f(t,x,u,u_x)`$ in Theorem 4.2. In other words $`h`$ is regular. For if $`g(t,x,u)`$ is $`O(t^\theta )`$ then $`(Ig(t,x,u))^1`$ is $`O(1)`$ and $`[I(Ig(t,x,u))^1]`$ is $`O(t^\theta )`$. In the following it will be necessary to verify the regularity hypothesis for some particular systems, and some general remarks which can be used to simplify this task will now be made. In these systems we always have (4.4) $$f(t,x,u,u_x)=\underset{i=1}{\overset{m}{}}t_iF_i(x,v(t,x),u,u_x)$$ where the $`F_i`$ are analytic functions on an open subset $`V`$ of $`𝐑^{n+l+k+nk}`$ which includes $`U_1\times \{0\}\times U_2`$, $`t_1,\mathrm{},t_m`$ are some functions of $`t`$ and $`v(t,x)`$ is a given function with values in $`𝐑^l`$. The $`t_i`$ are continuous functions on $`(0,T_0]`$ which tend to zero as $`t0`$ as least as fast as some positive power of $`t`$ and are analytic for $`t>0`$. The function $`v(t,x)`$ is the restriction of an analytic function on $`(0,T_0]\times \stackrel{~}{U}_1`$ to real values of its arguments. Each component of $`v`$ tends to zero as $`t0`$, uniformly on $`\stackrel{~}{U}_1`$. These properties ensure that $`f`$ is regular. For the functions $`F_i`$ have analytic continuations to some open neighbourhood $`\stackrel{~}{V}`$ of $`V`$ in $`𝐂^{n+l+k+nk}`$. In the examples we will meet the functions $`t_i`$ are either positive powers of $`t`$ or positive powers of $`t`$ times positive powers of $`\mathrm{log}t`$. The role of $`v(t,x)`$ is played by the functions $`t_j`$, the components of the velocity dominated metric $`{}_{}{}^{0}g_{ab}^{}`$, their spatial derivatives of first and second order, and the components of the inverse metric multiplied by suitable powers of $`t`$ so that the product vanishes in the limit $`t0`$. We saw in the last section that $`{}_{}{}^{0}g_{ab}^{}`$ can be written in the form $`t^{2K}`$ which, for each fixed $`t`$, is an entire function of $`K`$. Thus $`t^{2K(x)}`$ is analytic on any region where $`K(x)`$ is analytic. The velocity dominated metric and its derivatives all tend to zero uniformly on compact sets as $`t0`$ while the same is true of the inverse metric multiplied by a suitable power of $`t`$. Next a criterion will be given which allows the hypothesis on the matrix $`A`$ in Theorem 4.2 to be checked in many cases. ###### Lemma 4.3. Let $`A(x)`$ be a $`k\times k`$ matrix-valued continuous function defined on a compact subset of $`𝐑^n`$. If there is a constant $`\alpha `$ such that, for each eigenvalue $`\lambda `$ of $`A(x)`$ at any point of the given compact set, $`\mathrm{Re}\lambda >\alpha `$ then there is a constant $`C`$ such that the estimate $`t^{A(x)}Ct^\alpha `$ holds for $`t`$ small and positive and $`x`$ in the compact set. Proof The general case can be reduced to the case $`\alpha =0`$ by the following computation: (4.5) $$t^{A\alpha I}t^At^\alpha $$ In the rest of the proof only the case $`\alpha =0`$ will be considered. Without the parameter dependence the result could easily be proved by reducing the matrix to Jordan canonical form. The difficulty with a parameter is that the reduction to canonical form is in general not a continuous process. At least it can be concluded from the continuity properties of the eigenvalues that there is a $`\beta >0`$ such that all eigenvalues satisfy $`\mathrm{Re}\lambda >\beta `$. Let $`s=\mathrm{log}t`$. Then the problem is to show that for a fixed $`s_0`$ there is a constant $`C>0`$ such that $`e^{sA}C`$ for all $`s>s_0`$. By scaling $`t`$ we may suppose without loss of generality that $`s_0=0`$. For each $`x`$ we can conclude by reduction to canonical form that exists a value $`s_x`$ of $`s`$ such that the inequality $`e^{s_xA(x)}<e^{\beta s_x/2}`$ holds. By continuity of the exponential function there is an open neighbourhood $`U_x`$ of $`x`$ where this continues to hold for the given value of $`s_x`$. Let $`C_x=sup\{e^{sA(y)}:s[0,s_x],yU_x\}`$. It follows that for any $`s[0,\mathrm{})`$ and any $`yU_x`$ we have (4.6) $$e^{sA(y)}C_xe^{s_xA(y)}^{[s/s_x]}C_xe^{\beta s_x[s/s_x]/2}C_x$$ By compactness, it is possible to pass to a subcover consisting of a finite number of the sets $`U_x`$ and letting $`C`$ be the maximum of the corresponding $`C_x`$ we obtain the required estimate. ###### Remark 4.3. More generally, an analogous estimate is obtained if $`A(x)`$ is the direct sum of a matrix $`B(x)`$ whose eigenvalues have positive real parts and the zero matrix. This is obvious since in that case $`t^{A(x)}`$ is the direct sum of $`t^{B(x)}`$ and the identity. ## 5. Setting up the reduced equations In this section we introduce the adapted frame $`\{e_a\}`$ and the auxiliary exponents $`\{q_a\}`$ which will be used in the curvature estimate. Let $`x_0`$ be given. Let $`{}_{}{}^{0}g_{ab}^{}`$, $`{}_{}{}^{0}k_{b}^{a}`$, be solutions of the velocity dominated evolution and constraint equations (2.11) and (2.10). Let $`p_a`$ be the eigenvalues of $`K_b^a=t{}_{}{}^{0}k_{b}^{a}`$. Assume that $`\{p_a\}`$ are such that $`p_a(x_0)>0`$, $`a=1,2,3`$, $`_ap_a=1`$ (Kasner condition) and $`p_a`$ are ordered so that $`p_ap_b`$, for $`ab`$. Fix an initial time $`t_0(0,1)`$. We will in the following restrict our considerations to $`t(0,1)`$. If $`K`$ has a double eigenvalue at $`x_0`$, then in general the eigenvalues and eigenvectors of $`K`$ are not analytic in a neighbourhood of $`x_0`$, and therefore in general it is not possible to introduce an analytic frame diagonalizing $`K`$ in a given neighbourhood. We will avoid this problem by using the well known fact that if $`p_a^{}`$ is a double eigenvalue of $`K(x_0)`$, the eigenspace of the pair of eigenvalues corresponding to $`p_a^{}`$ is analytic in a neighbourhood of $`x_0`$. This means in particular that it is possible to choose an analytic frame which is adapted to the eigenspace of the pair of eigenvalues. This will play a central role in what follows. Choose numbers $`\alpha _0,ϵ>0`$ so that $`ϵ=\alpha _0/4<\mathrm{min}\{p_a(x_0)\}/40`$. 1. Cases I, II, III: We will distinguish between the following cases: 1. (near Friedmann) $`\mathrm{max}_{a,b}|p_ap_b|<ϵ/2,a=1,2,3`$. 2. (near double eigenvalue) $`\mathrm{max}_{a,b}|p_ap_b|>ϵ/2`$, and $`|p_{}_{}{}^{}ap_b^{}|<ϵ/2`$ for some pair $`a^{},b^{}`$, $`a^{}b^{}`$. Denote by $`p_{}`$ the distinguished exponent not equal to $`p_a^{},p_b^{}`$. 3. (diagonalizable) $`\mathrm{min}_{\begin{array}{c}a,b\\ ab\end{array}}|p_ap_b|>ϵ/2`$ By reducing $`ϵ`$ if necessary we can make sure that condition I, II or III holds at $`x_0`$ if the maximum multiplicity of an eigenvalue at $`x_0`$ is three, two or one respectively. The conditions I, II, III are open, and hence there is an open neighbourhood $`U_0x_0`$ such that, if condition I, II or III holds at $`x_0`$, the condition holds in $`U_0`$ and further, for $`xU_0`$, $`\mathrm{min}_a\{p_a(x)\}>20ϵ`$. 2. Auxiliary exponents $`\{q_a\}`$: Let $`U_0x_0`$ be as in point 1. We will define analytic functions $`q_a,a=1,2,3`$, called auxiliary exponents, in $`U_0`$, with the properties 1. $`q_a>0`$ (positivity) 2. $`q_aq_b`$ if $`ab`$ (ordering) 3. $`_aq_a=1`$ (Kasner) The auxiliary exponents $`q_a`$ will be defined in terms of the eigenvalues $`p_a`$ of $`K`$ depending on whether at $`x_0`$ we are in case I, II, or III. Let $`q_a(x_0)=p_a(x_0)`$, $`a=1,2,3`$, and choose $`q_a`$ on $`U_0`$ satisfying the positivity, ordering and Kasner conditions such that in cases I, II, III, the following holds. 1. $`q_a=1/3`$, $`a=1,2,3`$ 2. $`q_{}=p_{}`$, $`q_a^{}=q_b^{}=\frac{1}{2}(1q_{})`$ 3. $`q_a=p_a`$ Then $`q_a`$ are analytic on $`U_0`$. Note that it follows from the definition of $`q_a`$ that $`q_1\mathrm{min}_i\{p_i\}`$ and that $`\mathrm{max}_a|q_ap_a|<ϵ/2`$. 3. The frame $`\{e_a\}`$: In each case I, II, III define an analytic frame $`\{e_a\}`$ with dual frame $`\{\theta ^a\}`$, by the following prescription: $`\{e_a\}`$ is an ON frame w.r.t. $`{}_{}{}^{0}g(t_0)`$, and in case II, III the following additional conditions hold. 1. $`e_{}`$ is the eigenvector of $`K`$ corresponding to $`q_{}`$ and $`e_a^{},e_b^{}`$ span the eigenspace of $`K`$ corresponding to the eigenvalues $`p_a^{},p_b^{}`$. 2. $`e_a`$ are eigenvectors of $`K`$ corresponding to the eigenvalues $`q_a`$. We will call $`\{q_a\}`$, $`\{e_a\}`$, $`\{\theta ^a\}`$, satisfying the above conditions adapted. In the following we will work in a neighbourhood $`U_0`$ defined as above and assume that we are given adapted $`\{q_a\}`$, $`\{e_a\}`$, $`\{\theta ^a\}`$, on $`U_0`$. The role of $`\alpha _b^a`$ will be to shift the spectrum of the system matrix to be positive. We need to choose $`\alpha _0`$ larger than $`ϵ`$ to compensate for the fact that the $`q_a`$ are not the exact eigenvalues of $`K`$. In the following $`T_{ab}`$ will denote frame components of the tensor $`T`$ w.r.t. the frame $`\{e_a\}`$. Define the rescaled frame $`\stackrel{~}{e}_a=t^{q_a}e_a`$ with dual frame $`\stackrel{~}{\theta }^a=t^{q_a}\theta ^a`$ and denote by $`\stackrel{~}{T}_{ab}`$ the $`\stackrel{~}{e}_a`$ frame components of the tensor $`T`$. Then we have (5.1) $$\stackrel{~}{T}_{ab}=t^{q_aq_b}T_{ab},\stackrel{~}{T}_b^a=t^{q_aq_b}T_b^a.$$ It follows from the definitions, that in case III, (5.2) $$K_b^a=\delta _b^aq_b,{}_{}{}^{0}k_{b}^{a}=t^1\delta _b^aq_b$$ while in case II, the tensors $`K`$, $`{}_{}{}^{0}k`$ and $`{}_{}{}^{0}g`$ are block diagonal in the frame $`\{e_a\}`$, (5.3) $$K_{}^a^{}=0,{}_{}{}^{0}k_{}^{a^{}}=0,{}_{}{}^{0}g_{a^{}}^{}=0$$ For $`s𝐑`$, let $`(s)_+=\mathrm{max}(s,0)`$, for $`a,b\{1,2,3\}`$, let (5.4) $$\alpha _b^a=2(q_bq_a)_++\alpha _0,$$ and $`\stackrel{~}{\alpha }_b^a=|q_bq_a|+\alpha _0`$. In view of the relation $`2(s)_+s=|s|`$ we have (5.5) $$\alpha _b^a+q_aq_b=\stackrel{~}{\alpha }_b^a.$$ The following identities are an immediate consequence of equations (5.2) and (5.3), together with the fact that in case I, $`q_a=1/3`$, $`a=1,2,3`$. (5.6a) $`{}_{}{}^{0}g_{ab}^{}t^{\alpha _c^b}`$ $`={}_{}{}^{0}g_{ab}^{}t^{\alpha _c^a}`$ (5.6b) $`{}_{}{}^{0}g_{ab}^{}t^{q_b}`$ $`={}_{}{}^{0}g_{ab}^{}t^{q_a}`$ Note that in (5.6) no summation over indices is implied. Lemma 4.3 implies the following estimate for the rescaled frame components of $`{}_{}{}^{0}\stackrel{~}{g}`$. ###### Lemma 5.1. (5.7) $${}_{}{}^{0}\stackrel{~}{g}_{ab}^{}_{\mathrm{}}Ct^ϵ,{}_{}{}^{0}\stackrel{~}{g}_{}^{ab}_{\mathrm{}}Ct^ϵ$$ ###### Proof. Let $`Q_b^a=\delta _b^aq_b`$. A direct computation starting from the matrix form of the velocity dominated evolution equation gives $$t_t\stackrel{~}{G}_{ab}=2\stackrel{~}{G}_{ac}(K_b^cQ_b^c)$$ which has the solution $$\stackrel{~}{G}_{ab}(t)=\stackrel{~}{G}_{ac}(t_0)\left(\frac{t}{t_0}\right)^{2(K_b^cQ_b^c)}$$ From the definition of $`q_a`$ and the frame $`e_a`$ the spectrum of $`KQ`$ is of the form $`p_aq_a`$. Therefore since $`|p_aq_a|<ϵ/2`$, we find that the spectrum of $`2(KQ)`$ is contained in the interval $`(ϵ,ϵ)`$. Therefore it follows from Lemma 4.3 that $`\stackrel{~}{G}_{\mathrm{}}Ct^ϵ`$. Similarly, using the fact that $`\stackrel{~}{G}^1`$ satisfies the equation $$t_t\stackrel{~}{G}_{ab}^1=2(Q_a^cK_a^c)\stackrel{~}{G}_{cb}^1$$ an application of Lemma 4.3 yields the estimate $`\stackrel{~}{G}^1_{\mathrm{}}Ct^ϵ`$. ∎ We are now ready to describe the ansatz which will be used to write the Einstein–scalar field and Einstein–stiff fluid systems in Fuchsian form. Assume that a solution of the velocity dominated constraint and evolution equations, $`{}_{}{}^{0}g_{ab}^{}`$, $`{}_{}{}^{0}k_{ab}^{}`$, with adapted frame, coframe, and auxiliary exponents $`\{e_a\}`$, $`\{\theta ^a\}`$, $`\{q_a\}`$, is given. Let $`\alpha _b^a`$ be defined by (5.4). Let $`\zeta =ϵ/200`$. We will consider metrics and second fundamental forms $`g,k`$ of the form (5.8a) $`g_{ab}`$ $`={}_{}{}^{0}g_{ab}^{}+{}_{}{}^{0}g_{ac}^{}t^{\alpha _b^c}\gamma _b^c`$ $`\gamma _b^c=o(1)`$ (5.8b) $`g^{ab}`$ $`={}_{}{}^{0}g_{}^{ab}+t^{\alpha _b^a}\overline{\gamma }_c^a{}_{}{}^{0}g_{}^{cb}`$ $`\overline{\gamma }_c^a=o(1)`$ (5.8c) $`e_c(\gamma _b^a)`$ $`=t^\zeta \lambda _{bc}^a`$ $`\lambda _{bc}^a=o(1)`$ (5.8d) $`k_{ab}`$ $`=g_{ac}({}_{}{}^{0}k_{b}^{c}+t^{1+\alpha _b^c}\kappa _b^c)`$ $`\kappa _b^c=o(1)`$ The form (5.8b) is a consequence of (5.8a). To see this, note that $`{}_{}{}^{0}g_{}^{ac}g_{cb}=\delta _b^a+t^{\alpha _b^a}\gamma _b^a`$ and that $`g^{ac}{}_{}{}^{0}g_{cb}^{}=({}_{}{}^{0}g_{}^{ac}g_{cb})^1.`$ Thus the desired result follows from the matrix identity $`(I+A)^1=IA+(I+A)^1A^2`$ and the fact that, using $`2(x)_+x=|x|`$, it can be concluded that (5.9) $$\alpha _e^a+\alpha _b^e\alpha _b^a=|q_eq_a|+|q_bq_e||q_bq_a|+\alpha _0\alpha _0.$$ The latter relation shows that each component of the square of $`\gamma ^a_b`$ vanishes faster than the corresponding component of $`\gamma ^a_b`$ itself. Let $`\beta =ϵ/100`$. In addition to (5.8) we will use the following ansatz for the scalar field (5.10a) $`\varphi `$ $`={}_{}{}^{0}\varphi +t^\beta \psi `$ $`\psi =o(1)`$ (5.10b) $`e_a(\psi )`$ $`=t^\zeta \omega _a`$ $`\omega _a=o(1)`$ (5.10c) $`t_t\psi +\beta \psi `$ $`=\chi `$ $`\chi =o(1)`$ and for the stiff fluid case, (5.11a) $`\mu `$ $`={}_{}{}^{0}\mu +t^{2+\beta _1}\nu `$ $`\nu =o(1)`$ (5.11b) $`u_a`$ $`={}_{}{}^{0}u_{a}^{}+t^{1+\beta _2}v_a`$ $`v_a=o(1)`$ Equations (5.8), (5.10) and (5.11) with the exception of (5.8c) and (5.10b) will be used to derive the first reduced form of the field equations, and equations (5.8c) and (5.10b) for the spatial derivatives of $`\gamma _b^a`$ and $`\psi `$, will be used to derive the second reduced system. Note that in view of (5.1), we have (5.12a) $`\stackrel{~}{g}_{ab}`$ $`={}_{}{}^{0}\stackrel{~}{g}_{ab}^{}+{}_{}{}^{0}\stackrel{~}{g}_{ac}^{}t^{\stackrel{~}{\alpha }_b^c}\gamma _b^c`$ (5.12b) $`\stackrel{~}{g}^{ab}`$ $`={}_{}{}^{0}\stackrel{~}{g}_{}^{ab}+t^{\stackrel{~}{\alpha }_b^a}\overline{\gamma }_c^a{}_{}{}^{0}\stackrel{~}{g}_{}^{cb}`$ (5.12c) $`\stackrel{~}{k}_{ab}`$ $`=\stackrel{~}{g}_{ac}({}_{}{}^{0}\stackrel{~}{k}_{b}^{c}+t^{1+\stackrel{~}{\alpha }_b^c}\kappa _b^c)`$ We use the following conventions throughout: * indices on velocity dominated fields $`{}_{}{}^{0}g_{ab}^{},{}_{}{}^{0}k_{ab}^{},{}_{}{}^{0}u_{a}^{}`$ are raised and lowered with $`{}_{}{}^{0}g_{ab}^{}`$, while indices on other tensors are raised and lowered with $`g_{ab}`$. * the dynamic tensor fields $`\gamma _b^a,\kappa _b^a`$ in $`g_{ab},k_{ab}`$ are always used in mixed form and only in $`\{e_a\}`$ frame components. * the dynamic 1–form $`v_a`$ in the velocity field $`u_a`$ is always used with lower index. ### 5.1. The reduced Einstein–matter system In this section, we describe the first reduced system for the Einstein–scalar field evolution equations, derived from (2.3) using the ansatz given by equations (5.8) and (5.10) for $`g_{ab},k_{ab},\varphi `$ in terms of the velocity dominated solution $`{}_{}{}^{0}g_{ab}^{},{}_{}{}^{0}k_{ab}^{},{}_{}{}^{0}\varphi `$, the auxiliary exponents $`\{q_a\}`$ and the dynamical fields $`\gamma _b^a,\kappa _b^a,\psi `$. Similarly, we describe the first reduced system for the Einstein–stiff fluid evolution equations obtained using the equations (5.11). For convenience we use the term ‘Einstein–matter system’ to describe the Einstein–scalar field and Einstein–stiff fluid system collectively. The tensor $`g_{ab}`$ of the form (5.8a) is not a priori symmetric, but it will follow from Lemma 5.2 that the solution to the Fuchsian form of the Einstein–matter evolution equations will be symmetric. It is convenient to introduce the symmetrized tensor $${}_{}{}^{S}g_{ab}^{}=\frac{1}{2}(g_{ab}+g_{ba}).$$ Let $`{}_{}{}^{S}R_{ab}^{}`$ be the Ricci tensor computed w.r.t. the symmetrized metric $`{}_{}{}^{S}g_{ab}^{}`$, see section 6 for details. By substituting into the evolution equations (2.3) with $`R_{ab}`$ replaced by $`{}_{}{}^{S}R_{ab}^{}`$, defined in terms of $`\gamma _b^a`$ and $`\lambda _{bc}^a`$, we get the following system for $`\gamma _b^a,\kappa _b^a,\lambda _{bc}^a`$. (5.13a) $`t_t\gamma ^a_b`$ $`+\alpha _b^a\gamma _b^a+2\kappa _b^a+2\gamma _e^a(t{}_{}{}^{0}k_{b}^{e})2(t{}_{}{}^{0}k_{e}^{a})\gamma _b^e=2t^{\alpha _e^a+\alpha _b^e\alpha _b^a}\gamma _e^a\kappa _b^e`$ (5.13b) $`t_t\lambda _{bc}^a`$ $`=t^\zeta e_c(t_t\gamma _b^a)+\zeta t^\zeta e_c(\gamma _b^a)`$ (5.13c) $`t_t\kappa _b^a`$ $`+\alpha _b^a\kappa _b^a(t{}_{}{}^{0}k_{b}^{a})(\text{tr}\kappa )=t^{\alpha _0}(\text{tr}\kappa )\kappa _b^a`$ (5.13d) $`+t^{2\alpha _b^a}({}_{}{}^{S}R_{b}^{a}M_b^a)`$ where $`M_{ab}`$ is is given by (5.14a) $`M_{ab}`$ $`=8\pi e_a(\varphi )e_b(\varphi )`$ for the Einstein–scalar field system (5.14b) $`M_{ab}`$ $`=16\pi \mu u_au_b`$ for the Einstein–stiff fluid system $`M_b^a`$ will be estimated in section 6. Note that the power of $`t`$ occurring on the right hand side of equation (5.13a) is positive due to (5.9). The wave equation (2.6) becomes the following system of equations for $`\psi ,\omega _a,\chi `$. (5.15a) $`t_t\psi +\beta \psi \chi `$ $`=0`$ (5.15b) $`t_t\omega _a`$ $`=t^\zeta [e_a(\chi )+(\zeta \beta )e_a(\psi )]`$ (5.15c) $`t_t\chi +\beta \chi `$ $`=t^{\alpha _0\beta }\text{tr}\kappa (A+t^\beta \chi )+t^{2\beta }\mathrm{\Delta }{}_{}{}^{0}\varphi +t^{2\zeta }^a\omega _a`$ Let $`𝒰=(\gamma _b^a,\kappa _b^a,\psi ,\chi ,\lambda _{bc}^a,\omega _a)`$. Then we can write the second reduced system in the Einstein–scalar field case, which consists of equations (5.13) and (5.15) in the form $$t_t𝒰+𝒜𝒰=(t,x,𝒰,𝒰_x).$$ for a matrix $`𝒜`$ and a function $``$. We will prove that this system is in Fuchsian form. The Einstein-stiff fluid equations will be treated in a similar way. However, due to the complexity of the equations in that case, they will not be written out more explicitly than is absolutely necessary to understand the essential features of their structure. The second reduced system in the stiff fluid case can be brought into the generalized Fuchsian form (5.16) $$t_t𝒰+𝒜𝒰=(t,x,𝒰,𝒰_x)+𝒢(t,x,𝒰)_t𝒰$$ already introduced in section 4. In order to do this it is useful to introduce some abbreviations for certain terms in (2.9) so that the equations become (5.17a) $`_t\mu 2(\text{tr}k)\mu =2|u|^2_t\mu 4\mu u^a_tu_a+F_1`$ (5.17b) $`_tu_a+(\text{tr}k)u_a+(1/2)\mu ^1e_a(\mu )=\mu ^1[_t\mu 2(\text{tr}k)\mu ]u_a`$ $`+(1+|u|^2)^1u^bu_a_tu_b[(1+|u|^2)^{1/2}1]\mu ^1e_a(\mu )+F_2`$ The expressions $`F_1`$ and $`F_2`$ contain only terms which can be incorporated into $``$ in (5.16). Next the ansatz (5.11) must be substituted into these equations. The result is: (5.18a) $`t_t\nu +\beta _1\nu `$ $`=2t^{3\beta _1}(1+t\text{tr}k){}_{}{}^{0}\mu +2(1+t\text{tr}k)\nu +t^{3\beta _1}[_t\mu 2(\text{tr}k)\mu ]`$ (5.18b) $`t_tv_a+\beta _2v_a`$ $`=(1+t\text{tr}k)v_a+t^{\beta _2}[_tu_a+(\text{tr}k)u_a+(1/2){}_{}{}^{0}\mu _{}^{1}e_a({}_{}{}^{0}\mu )]`$ The expressions on the left hand side of the above form of the Euler equations written in terms of the basic variables $`\mu `$ and $`u_a`$ occur on the right hand sides of the above evolution equations for $`\nu `$ and $`v_a`$. In order to get a fully explicit form it would be necessary to substitute for these expressions and then express the final result in terms of $`\nu `$ and $`v_a`$. This is, however, neither necessary nor even helpful for the analysis to be done here. Next we will consider the matrix $`𝒜`$ and prove that $`𝒜`$ is a direct sum of a matrix with spectrum bounded from below by a positive number, with a zero matrix. (The arguments in the scalar field and stiff fluid cases are very similar.) It is therefore of a form such that the theory presented in section 4 applies. In addition we must show that $`(t,x,𝒰,𝒰_x)=O(t^\delta )`$ for some $`\delta >0`$ and, in the stiff fluid case, that $`𝒢(t,x,𝒰)`$ satisfies a similar estimate. This will be done in the next section. The matrix $`𝒜`$ is block diagonal and therefore it is enough to consider each block separately. The rows and columns of $`𝒜`$ corresponding to $`\lambda _{bc}^a,\omega _a`$ are zero, and therefore this $`𝒜`$ is the direct sum of a matrix corresponding to $`\gamma ,\kappa ,\psi ,\chi `$, with a zero matrix in the scalar field case and the direct sum of a matrix corresponding to $`\gamma ,\kappa ,\nu ,v_a`$ with a zero matrix in the stiff fluid case. We now consider the spectrum of this matrix. The submatrix corresponding to $`\gamma ,\kappa `$ is upper block triangular. The $`\gamma ,\gamma `$ block is given by $$\gamma _b^a\alpha _b^a\gamma _b^a+2[\gamma ,t{}_{}{}^{0}k]_b^a$$ To estimate the spectrum of this, it is necessary to consider the cases I,II,III separately. Working in a frame which diagonalizes $`{}_{}{}^{0}k`$, $`t{}_{}{}^{0}k_{b}^{a}=\delta _b^ap_b`$, and hence in this case $$2[\gamma ,t{}_{}{}^{0}k]_b^a=2(p_bp_a)\gamma _b^a$$ Therefore, in case III, we get using the definition of $`\alpha _b^a`$, $$\alpha _b^a\gamma _b^a+2[\gamma ,t{}_{}{}^{0}k]_b^a=2((p_bp_a)_+(p_bp_a)+\alpha _0)\gamma _b^a$$ and hence using $`(x)_+x0`$ for all $`x𝐑`$, the spectrum of the $`\gamma ,\gamma `$ block is bounded from below by $`\alpha _0`$ in case III. Next consider case I. In this case, $`\alpha _b^a=\alpha _0=4ϵ`$ and the spectrum of $`\gamma _b^a2[\gamma ,t{}_{}{}^{0}k]_b^a`$ is bounded from below by $`ϵ`$, which shows that the spectrum of the $`\gamma ,\gamma `$ block is bounded from below by $`3ϵ`$ in case II. Finally, in case II, $`{}_{}{}^{0}k_{b}^{a}`$ is block diagonal in the adapted frame. The spectrum of $`\gamma _b^a2[\gamma ,t{}_{}{}^{0}k]_b^a`$ consists of $`2(p_a^{}p_b^{})`$, $`2(p_a^{}p_{})`$, and $`0`$. Now using the definition of $`\alpha _b^a`$ for case II and arguing as above, we get the lower bound $`3ϵ`$ for the spectrum of the $`\gamma ,\gamma `$ block in case II. Therefore the spectrum of the $`\gamma ,\gamma `$ of $`𝒜`$ is bounded from below by $`3ϵ`$. Next we consider the $`\kappa ,\kappa `$ block. This is of the form $$\kappa _b^a\alpha _b^a\kappa _b^a(t{}_{}{}^{0}k_{b}^{a})\text{tr}\kappa $$ First consider the action on the trace–free part of $`\kappa _b^a`$. Then the spectrum is given by $`\alpha _b^a>\alpha _0`$. On the other hand, restricting to the trace part of $`\kappa _b^a`$, which is diagonal, we see that the spectrum is $`\alpha _0+1`$. Therefore the spectrum of the $`\kappa ,\kappa `$ block is bounded from below by $`\alpha _0`$. The $`\psi ,\chi `$ block is of the form $$\left(\begin{array}{cc}\beta & 1\\ 0& \beta \end{array}\right)$$ which has spectrum $`\beta >0`$. The $`\nu ,v_a`$ block is diagonal with eigenvalues $`\beta _1`$ and $`\beta _2`$. Therefore, in view of the facts that $`3ϵ>\beta >0`$, $`\beta _1>0`$ and $`\beta _2>0`$, the desired properties of the spectrum of $`𝒜`$ have been verified. Given a solution $`𝒰=(\gamma _b^a,\kappa _b^a,\psi ,\chi ,\lambda _{bc}^a,\omega _a)`$ of the reduced system for the Einstein–scalar field equations, define $`g_{ab}`$, $`k_{ab}`$ and $`\varphi `$ by (5.8) and (5.10). Similarly, given a solution $`𝒰=(\gamma _b^a,\kappa _b^a,\nu ,v_a\lambda _{bc}^a)`$ of the reduced system for the Einstein–stiff fluid equations, define $`(g_{ab},k_{ab},\mu ,u_a)`$ by (5.8) and (5.11). If it can be shown that $`g_{ab}`$ and $`k_{ab}`$ are symmetric then a solution of the Einstein-scalar field equations is obtained. The next lemma gives sufficient conditions for this to be true. ###### Lemma 5.2. Let a solution of the velocity dominated Einstein-matter system be given on $`S\times (0,\mathrm{})`$ with all eigenvalues of $`t{}_{}{}^{0}k_{}^{a}_b`$ positive and $`t\text{tr}{}_{}{}^{0}k=1`$. Let $`𝒰`$ be a solution of the reduced system for the Einstein–scalar field or Einstein–stiff fluid system corresponding to the given velocity dominated solution, with $`𝒰=o(1)`$. Define $`(g_{ab},k_{ab},\varphi )`$ by (5.8) and (5.10) in the scalar field case, and define $`(g_{ab},k_{ab},\mu ,u_a)`$ by (5.8) and (5.11) in the stiff fluid case. Then $`g_{ab}`$ and $`k_{ab}`$ are symmetric. ###### Proof. From the evolution equation for $`\gamma _b^a`$ and the definitions of $`g_{ab}`$ and $`k_{ab}`$ it follows that $`_tg_{ab}=2k_{ab}`$. Similarly an equation close to the usual evolution equation for $`k_b^a`$ can be recovered from (5.13d). It differs from the usual one only in the fact that $`R_b^a`$ is replaced by $`{}_{}{}^{S}R_{b}^{a}`$. From these equations we can derive the equations: (5.19a) $`_t(g_{ab}g_{ba})`$ $`=2(k_{ab}k_{ba})`$ (5.19b) $`_t(k_{ab}k_{ba})`$ $`=(\text{tr}k)(k_{ab}k_{ba})`$ It follows from the assumptions on $`\gamma _b^a`$ and $`\kappa _b^a`$ together with the definition of $`k_{ab}`$, that the components of $`k_{ab}k_{ba}`$ are $`o(t^{1+\eta })`$ for some $`\eta >0`$. Hence the quantity $`\mathrm{\Omega }_{ab}=t^{1\eta }(k_{ab}k_{ba})`$ tends to zero as $`t0`$. It satisfies the equation: (5.20) $$t_t\mathrm{\Omega }_{ab}+\eta \mathrm{\Omega }_{ab}=(t\text{tr}k+1)\mathrm{\Omega }_{ab}$$ From Theorem 4.2 we conclude that $`\mathrm{\Omega }_{ab}=0`$. Thus $`k_{ab}`$ is symmetric. It then follows immediately from (5.19a) and the fact that $`g_{ab}=o(1)`$ that $`g_{ab}`$ is also symmetric. ∎ ## 6. Curvature estimates Let $`{}_{}{}^{S}R_{ab}^{}`$ be the Ricci tensor computed w.r.t. the symmetrized metric $`{}_{}{}^{S}g_{ab}^{}=\frac{1}{2}(g_{ab}+g_{ba})`$. In order to get a Fuchsian form for the Einstein–matter evolution equations, we need the following estimate for the frame components of $`{}_{}{}^{S}R_{b}^{a}`$, (6.1) $$t^{2\alpha _b^a}{}_{}{}^{S}R_{b}^{a}=O(t^\delta ),\text{ for some }\delta (0,ϵ).$$ In doing the estimates we will use the notion of comparing the size of functions introduced in Definition 4.1. In proving that the second reduced system $$t_tu+A(x)u=f(t,x,u,u_x)$$ is in Fuchsian form, one essential step is to prove an estimate of the form $$ft^\delta $$ for some $`\delta >0`$. In the present section, we accomplish this task for the expression $`t^{2\alpha _b^a}{}_{}{}^{S}R_{b}^{a}`$, which is now considered as a function $`r(t,x,v(x,t),u,u_x)`$, where $`v(x,t)`$ is defined in terms of the solution $`{}_{}{}^{0}g_{ab}^{},{}_{}{}^{0}k_{ab}^{}`$ to the velocity dominated system and the data $`\{e_a\},\{\theta ^a\},\{q_a\}`$, etc. defined in section 5, and $`u`$ consists of the variables $`\gamma _b^a,\lambda _{bc}^a`$. In terms of the relation $``$, the goal is to prove (6.2) $$t^{2\alpha _b^a}{}_{}{}^{S}R_{b}^{a}t^\delta ,\text{ for some }\delta (0,ϵ).$$ By assumption, $`\alpha _0=4ϵ`$, so $`\alpha _b^a2ϵ=2(q_bq_a)_++\alpha _0/2`$. Therefore, the arguments that apply to $`\alpha _b^a`$ also apply to $`\alpha _b^a2ϵ`$. The symmetrized metric tensor satisfies (6.3a) $`{}_{}{}^{S}\stackrel{~}{g}_{ab}^{}`$ $`={}_{}{}^{0}\stackrel{~}{g}_{ab}^{}+t^{\stackrel{~}{\alpha }_b^a2ϵ}{}_{}{}^{0}\stackrel{~}{g}_{ac}^{}{}_{}{}^{S}\gamma _{b}^{c},{}_{}{}^{S}\gamma _{b}^{c}=o(1),`$ (6.3b) $`{}_{}{}^{S}\stackrel{~}{g}_{}^{ab}`$ $`={}_{}{}^{0}\stackrel{~}{g}_{}^{ab}+t^{\stackrel{~}{\alpha }_b^a2ϵ}{}_{}{}^{S}\overline{\gamma }_{c}^{a}{}_{}{}^{0}\stackrel{~}{g}_{}^{cb},{}_{}{}^{S}\overline{\gamma }_{c}^{a}=o(1)`$ To see this, note the identity $${}_{}{}^{S}\gamma _{b}^{c}=\frac{1}{2}t^{2ϵ}\left(\gamma _b^c+{}_{}{}^{0}\stackrel{~}{g}_{}^{cd}{}_{}{}^{0}\stackrel{~}{g}_{bf}^{}\gamma _d^f\right),$$ which in view of Lemma 5.1 shows that $`{}_{}{}^{S}\gamma _{b}^{c}=o(1)`$. The argument that $`{}_{}{}^{S}\overline{\gamma }_{c}^{a}=o(1)`$ is the same as for $`\overline{\gamma }_c^a=o(1)`$. It is convenient to estimate the rescaled frame components. Note $`R_b^a=t^{q_a+q_b}\stackrel{~}{R}_b^a`$. Hence in view of (5.5) we need to consider $$t^{2\stackrel{~}{\alpha }_b^a}\stackrel{~}{R}_b^a$$ Using Lemma 5.1 and (5.8) gives $$t^{2\stackrel{~}{\alpha }_b^a}\stackrel{~}{R}_b^a_{\mathrm{}}Ct^ϵt^{2\stackrel{~}{\alpha }_b^a}\stackrel{~}{R}_{ab}_{\mathrm{}}$$ To see this, we compute using (5.6a) and (5.12b) $`t^{2\stackrel{~}{\alpha }_b^a}\stackrel{~}{R}_b^a_{\mathrm{}}`$ $`=t^{2\stackrel{~}{\alpha }_b^a}\stackrel{~}{g}^{ac}\stackrel{~}{R}_{cb}_{\mathrm{}}`$ $`t^{2\stackrel{~}{\alpha }_b^a}{}_{}{}^{0}\stackrel{~}{g}_{}^{ac}\stackrel{~}{R}_{cb}_{\mathrm{}}+t^{2\stackrel{~}{\alpha }_b^a+\stackrel{~}{\alpha }_c^a}\overline{\gamma }_d^a{}_{}{}^{0}\stackrel{~}{g}_{}^{dc}\stackrel{~}{R}_{cb}_{\mathrm{}}`$ $`Ct^{2ϵ\stackrel{~}{\alpha }_b^c}\stackrel{~}{R}_{cb}_{\mathrm{}}`$ where we used the triangle inequality in the form $`\stackrel{~}{\alpha }_b^a+\stackrel{~}{\alpha }_c^a\stackrel{~}{\alpha }_b^c`$. Let $`\gamma _{ab}^c=\theta ^c([e_a,e_b])`$ be the structure coefficients of the frame $`\{e_a\}`$. The structure coefficients $`\stackrel{~}{\gamma }_{ab}^c=\stackrel{~}{\theta }^c([\stackrel{~}{e}_a,\stackrel{~}{e}_b])`$ of the frame $`\stackrel{~}{e}_a`$ are given by (6.4) $$\stackrel{~}{\gamma }_{ab}^c=t^{q_cq_aq_b}\gamma _{ab}^c\mathrm{log}(t)(t^{q_a}e_a(q_b)\delta _b^ct^{q_b}e_b(q_a)\delta _a^c)$$ It is convenient to define $`\stackrel{~}{\mathrm{\Gamma }}_{abc}=_{\stackrel{~}{e}_a}\stackrel{~}{e}_b,\stackrel{~}{e}_c.`$ Then $`\stackrel{~}{\mathrm{\Gamma }}_{abc}`$ is given in terms of $`\stackrel{~}{g}_{ab}`$ by (6.5) $$\begin{array}{cc}\hfill 2\stackrel{~}{\mathrm{\Gamma }}_{abc}& =\stackrel{~}{e}_a(\stackrel{~}{g}_{bc})+\stackrel{~}{e}_b(\stackrel{~}{g}_{ac})\stackrel{~}{e}_c(\stackrel{~}{g}_{ab})\hfill \\ & +\stackrel{~}{\gamma }_{ab}^d\stackrel{~}{g}_{cd}\stackrel{~}{\gamma }_{ac}^d\stackrel{~}{g}_{bd}\stackrel{~}{\gamma }_{bc}^d\stackrel{~}{g}_{ad}\hfill \end{array}$$ and $`\stackrel{~}{R}_{dcab}`$ is given in terms of $`\stackrel{~}{\mathrm{\Gamma }}_{abc}`$ and $`\stackrel{~}{\gamma }_{bc}^a`$ by $`\stackrel{~}{R}_{dcab}`$ $`=\stackrel{~}{e}_a\stackrel{~}{\mathrm{\Gamma }}_{bcd}\stackrel{~}{e}_b\stackrel{~}{\mathrm{\Gamma }}_{acd}\stackrel{~}{\gamma }_{ab}^f\stackrel{~}{\mathrm{\Gamma }}_{fcd}`$ $`\stackrel{~}{g}^{fg}\stackrel{~}{\mathrm{\Gamma }}_{bcf}\stackrel{~}{\mathrm{\Gamma }}_{adg}+\stackrel{~}{g}^{fg}\stackrel{~}{\mathrm{\Gamma }}_{acf}\stackrel{~}{\mathrm{\Gamma }}_{bdg}`$ (6.6) $`=(R1)_{dcab}(R2)_{dcab}(R3)_{dcab}(R4)_{dcab}+(R5)_{dcab}`$ Let $$z_{ab}=\{\begin{array}{cc}0,& \text{ if }a=b\hfill \\ 1,& \text{ if }ab\hfill \end{array}$$ Define (6.7) $$\begin{array}{cc}\hfill Z_{abc}(t)& =t^{q_a}+t^{q_b}+t^{q_c}\hfill \\ & +t^{q_aq_bq_c}z_{bc}+t^{q_bq_cq_a}z_{ca}+t^{q_cq_aq_b}z_{ab}\hfill \end{array}$$ Note that $`(R1)_{dcab},\mathrm{},(R5)_{dcab},z_{ab},Z_{abc}`$ are not tensors. In the rest of this section, the frame $`\{e_a\}`$ is fixed and the estimates will be done for tensor components in this frame. We will use the following lemma as the starting point of the estimates in this section. ###### Lemma 6.1. (6.8a) $`\stackrel{~}{g}_{ab}`$ $`t^ϵ`$ (6.8b) $`\stackrel{~}{g}^{ab}`$ $`t^ϵ`$ (6.8c) $`\stackrel{~}{e}_a\stackrel{~}{g}_{bc}`$ $`t^{q_a2ϵ}`$ (6.8d) $`\stackrel{~}{e}_a\stackrel{~}{e}_b\stackrel{~}{g}_{cd}`$ $`t^{q_aq_b3ϵ}`$ (6.8e) $`\stackrel{~}{g}^{ab}t^{q_b}`$ $`t^{ϵq_a}`$ (6.8f) $`\stackrel{~}{e}_c(t^{q_a}\stackrel{~}{g}_{ab})`$ $`t^{q_c+q_b2ϵ}`$ (6.8g) $`\stackrel{~}{g}^{ab}Z_{abc}`$ $`t^ϵ(t^{q_h}+t^{q_c}+t^{q_c2q_h})`$ $`h=\mathrm{min}(a,b)`$ (6.8h) $`\stackrel{~}{\gamma }_{ab}^c\stackrel{~}{g}^{bd}`$ $`t^{2ϵ}(t^{q_cq_aq_d}+t^{q_a}+t^{q_d})`$ ###### Proof. The inequalities (6.8a) and (6.8b) are immediate from Lemma 5.1, (5.12) and definition 4.1. Recalling that the variables $`u,u_x`$ occurring in the second reduced system contain $`\gamma _b^a`$, $`\lambda _{bc}^a`$ and its first order derivatives gives (6.8c) and (6.8d). (Here the inequality $`\zeta <ϵ`$ has been used.) The estimate (6.8e) follows from Lemma 5.1, (5.6) and (5.12) together with the triangle inequality in the form $`q_bq_a+|q_aq_b|`$. The estimate (6.8f) follows from (5.6) and (5.12) together with the observation that since $`{}_{}{}^{0}g_{ab}^{}`$ is block diagonal for $`xU_0`$, $`e_c{}_{}{}^{0}g_{ab}^{}`$ is also block diagonal. The estimates (6.8g) and (6.8h) follow in a similar way starting from (6.7) and (6.4). In (6.8h) a $`\mathrm{log}(t)`$ term is dominated by $`t^ϵ`$. ∎ The following lemma gives estimates of $`\stackrel{~}{\mathrm{\Gamma }}_{abc}`$ in terms of $`Z_{abc}`$ ###### Lemma 6.2. (6.9a) $`\stackrel{~}{\mathrm{\Gamma }}_{abc}`$ $`t^{2ϵ}Z_{abc}`$ (6.9b) $`\stackrel{~}{\mathrm{\Gamma }}_{abb}`$ $`t^{2ϵq_a}`$ (6.9c) $`\stackrel{~}{\mathrm{\Gamma }}_{aab}`$ $`t^{2ϵ}(t^{q_a}+t^{q_b})`$ (6.9d) $`\stackrel{~}{\mathrm{\Gamma }}_{aba}`$ $`t^{2ϵ}(t^{q_a}+t^{q_b})`$ (6.9e) $`\stackrel{~}{e}_a\stackrel{~}{\mathrm{\Gamma }}_{bcd}`$ $`t^{3ϵq_a}Z_{bcd}`$ ###### Proof. First observe that $`\gamma _{bc}^fz_{bc}`$. From (6.4) and (6.8) we have $$\stackrel{~}{\gamma }_{bc}^f\stackrel{~}{g}_{fa}t^{2ϵ}(t^{q_aq_bq_c}z_{bc}+t^{q_b}+t^{q_c})$$ Now noting $`\stackrel{~}{e}_a(\stackrel{~}{g}_{bc})t^{q_a2ϵ}`$, (6.9a) follows. To estimate $`\stackrel{~}{\mathrm{\Gamma }}_{abb}`$ we compute $$\stackrel{~}{\mathrm{\Gamma }}_{abb}=\frac{1}{2}t^{q_a}e_a\stackrel{~}{g}_{bb}t^{2ϵq_a}$$ The estimates for $`\stackrel{~}{\mathrm{\Gamma }}_{aab},\stackrel{~}{\mathrm{\Gamma }}_{aba}`$ follow directly from (6.9a) and the definition of $`Z_{abc}`$. Finally we consider (6.9e). Expanding out $`\stackrel{~}{e}_a\stackrel{~}{\mathrm{\Gamma }}_{bcd}`$, we see that it contains (up to permutations of the indices) terms of the form $`\stackrel{~}{e}_a\stackrel{~}{e}_b(\stackrel{~}{g}_{cd})`$, $`(\stackrel{~}{e}_a\stackrel{~}{\gamma }_{cd}^f)\stackrel{~}{g}_{fb}`$ and $`\stackrel{~}{\gamma }_{cd}^f\stackrel{~}{e}_a\stackrel{~}{g}_{fb}`$. The first and second type of terms are estimated using (6.8d) and (6.8e), using the form of $`\stackrel{~}{\gamma }_{cd}^f`$. Finally, the third type of term is estimated using (6.8f). ∎ An important consequence of the Kasner relation $`_aq_a=1`$, is (6.10) $$2+2(q_1q_2q_3)=4q_1$$ which implies (6.11) $$t^{2+2(q_jq_kq_l)}t^{4q_1}\text{if at least one of }k,l\text{ is different from 3},$$ The strategy will be to eliminate as much as possible the occurence of repeated negative exponents, in order to be able to use this relation. We make note of the following useful relations. (6.12a) $`Z_{abc}`$ $`t^{q_1q_2q_3}`$ (6.12b) $`Z_{hhc}`$ $`t^{q_h}+t^{q_c}`$ (6.12c) $`e_aZ_{bcd}`$ $`t^ϵZ_{bcd}`$ The estimates (6.12a) and (6.12b) are immediate from (6.7). For the rest of this section, we will assume that $`g_{ab}`$ is symmetric and is of the form given by (5.8a). Let $`R_{ab}`$ be the Ricci tensor defined with respect to $`g_{ab}`$. Under these assumptions we will estimate $`R_b^a`$. The estimate then applies after a small modification to $`{}_{}{}^{S}R_{b}^{a}`$. We now proceed to estimate the rescaled components of the Ricci tensor $`\stackrel{~}{R}_{ad}=\stackrel{~}{g}^{bc}\stackrel{~}{R}_{dcab}`$. Corresponding to the terms $`(R1),\mathrm{},(R5)`$ we have $$\stackrel{~}{R}_{ad}=(\text{Ric}1)_{ad}(\text{Ric}2)_{ad}(\text{Ric}3)_{ad}(\text{Ric}4)_{ad}+(\text{Ric}5)_{ad}$$ We make the following simplifying observations. * By the symmetry $`\stackrel{~}{R}_{ad}=\stackrel{~}{R}_{da}`$ we can assume without loss of generality that $`ad`$. * $`\stackrel{~}{R}_{dcab}`$ is skew symmetric in the first and second pair of indices, therefore we can assume without loss of generality that $`cd`$, $`ba`$. Therefore, in the following, we can without loss of generality use the following convention: The indices $`a,b,c,d`$ satisfy the relations (6.13) $$ad,cd,ab$$ We will now estimate $`\stackrel{~}{R}_{ad}`$ by considering each term $`(\text{Ric}1)_{ad},\mathrm{},(\text{Ric}5)_{ad}`$ in turn. The estimate we will actually prove is of the form $$t^{2+q_aq_d}\stackrel{~}{R}_{ad}t^{4q_16ϵ}$$ which will imply the needed estimate for $`{}_{}{}^{S}R_{d}^{a}`$. ### 6.1. $`(\text{Ric}1)`$ $$(\text{Ric}1)_{ad}=\stackrel{~}{g}^{bc}\stackrel{~}{e}_a\stackrel{~}{\mathrm{\Gamma }}_{bcd}$$ where we are summing over repeated indices. Let $`F=t^{2+q_aq_d}(\text{Ric}1)_{ad}`$. Using Lemma 6.2 and (6.8g), we have $`F`$ $`t^{24ϵ}t^{q_aq_d}t^{q_a}(t^{q_c}+t^{q_d}+t^{q_d2q_c})`$ $`t^{24ϵ}(t^{q_dq_c}+t^{2q_d}+t^{2q_c})`$ $`t^{4q_14ϵ}t^{4q_16ϵ}`$ ### 6.2. $`(\text{Ric}2)`$ $$(\text{Ric}2)_{ad}=\stackrel{~}{g}^{bc}\stackrel{~}{e}_b\stackrel{~}{\mathrm{\Gamma }}_{acd}$$ Let $`F=t^{2+q_aq_d}(\text{Ric}2)_{ad}`$. By Lemma 6.2 and (6.8e), $$Ft^{24ϵ}t^{q_aq_dq_c}Z_{acd}$$ By (6.13), $`dc`$ and using (6.12a) and (6.11) this gives $$Ft^{4q_14ϵ}t^{4q_16ϵ}$$ which is the required estimate. ### 6.3. $`(\text{Ric}3)`$ $$(\text{Ric}3)_{ad}=\stackrel{~}{g}^{bc}\stackrel{~}{\gamma }_{ab}^f\stackrel{~}{\mathrm{\Gamma }}_{fcd}$$ Let $`F=t^{2+q_aq_d}(\text{Ric}3)_{ad}`$. We estimate using (6.8h) and Lemma 6.2, $`F`$ $`t^{24ϵ}t^{q_aq_d}(t^{q_fq_aq_c}+t^{q_a}+t^{q_c})Z_{fcd}`$ $`t^{24ϵ}(t^{q_fq_dq_c}+t^{q_d}+t^{q_aq_dq_c})Z_{fcd}`$ use $`cd`$ by (6.13), (6.12a) and (6.11) $`t^{4q_14ϵ}t^{4q_16ϵ}`$ ### 6.4. $`(\text{Ric}4)`$ $$(\text{Ric}4)_{ad}=\stackrel{~}{g}^{bc}\stackrel{~}{g}^{fg}\stackrel{~}{\mathrm{\Gamma }}_{bcf}\stackrel{~}{\mathrm{\Gamma }}_{adg}$$ Let $`F=t^{2+q_aq_d}(\text{Ric}4)_{ad}`$. We have by Lemma 6.2, $$Ft^{2+q_aq_d4ϵ}\stackrel{~}{g}^{bc}\stackrel{~}{g}^{fg}Z_{bcf}Z_{adg}$$ In case $`a=d`$ this gives, with $`h=\mathrm{min}(b,c)`$, $`m\{f,g\}`$, using (6.12) and (6.8), $`t^2(\text{Ric}4)_{aa}`$ $`t^{26ϵ}(t^{q_h}+t^{q_m}+t^{q_m2q_h})(t^{q_a}+t^{q_m})`$ $`t^{4q_16ϵ}`$ Next we consider the case $`a<d`$. In case $`g=d`$, Lemma 6.2 together with (6.8e) and (6.8g) gives $`F`$ $`t^{2+q_aq_d}\stackrel{~}{g}^{bc}\stackrel{~}{g}^{fd}\stackrel{~}{\mathrm{\Gamma }}_{bcf}t^{2ϵq_a}`$ $`t^{26ϵ}(t^{q_fq_h}+t^{2q_f}+t^{2q_h})h=\mathrm{min}(b,c)`$ $`t^{4q_16ϵ}`$ In case $`a=g`$ we have arguing as above $`F`$ $`t^{2+q_aq_d}\stackrel{~}{g}^{bc}\stackrel{~}{g}^{fa}\stackrel{~}{\mathrm{\Gamma }}_{bcf}t^{2ϵ}(t^{q_a}+t^{q_d})`$ $`t^{25ϵ}\stackrel{~}{g}^{fa}(t^{q_h}+t^{q_f}+t^{q_f2q_h})(t^{q_d}+t^{q_a2q_d}),h=\mathrm{min}(b,c)`$ $`t^{26ϵ}(t^{q_h}+t^{q_m}+t^{q_m2q_h})(t^{q_d}+t^{q_m2q_d}),m=\mathrm{min}(a,f)`$ From $`h=\mathrm{min}(b,c)`$ and $`cd`$ which holds by (6.13), we find that either $`h<3`$ or $`d<3`$ must hold. Using this it follows using (6.11) that in case $`a=g<d`$, $`Ft^{4q_16ϵ}`$. It remains to consider the case when $`a<d`$ and $`a,d,g`$ are distinct. In this case, the estimates used above give $`F`$ $`t^{2+q_aq_d6ϵ}(t^{q_h}+t^{q_g}+t^{q_g2q_h})Z_{adg},h=\mathrm{min}(b,c)`$ $`t^{26ϵ}(t^{q_h}+t^{q_g}+t^{q_g2q_h})`$ $`(t^{q_d}+t^{q_a2q_d}+t^{q_aq_dq_g}+t^{2q_a2q_dq_g}+t^{q_g}+t^{q_g2q_d})`$ By construction, $`h<3`$ or $`d<3`$ must hold, which in conjunction with the fact that in the present case, $`gd`$ gives using (6.11), $$Ft^{4q_16ϵ}$$ The above proves that the required estimate $`t^{2+q_aq_d}(\text{Ric}4)_{ad}t^{4q_16ϵ}`$ holds. ### 6.5. $`(\text{Ric}5)`$ $$(\text{Ric}5)_{ad}=\stackrel{~}{g}^{bc}\stackrel{~}{g}^{fg}\stackrel{~}{\mathrm{\Gamma }}_{acf}\stackrel{~}{\mathrm{\Gamma }}_{bdg}$$ The estimate for $`(\text{Ric}5)_{ad}`$ is the most complicated, and will be done in several steps. We review the steps which will be used here. In each step the conditions on the indices $`a,c,f,b,d,g`$ which leads to the required estimate may be excluded from our considerations. Recall that $`ad`$ may be assumed and also note by (6.13) we may assume without loss of generality that $`ab`$, $`cd`$. The steps we will use are: 1. $`a=d`$ can be excluded, so $`a<d`$ may be assumed. 2. $`g=d`$ can be excluded, so $`gd`$ may be assumed. 3. $`g=a`$ can be excluded, so $`ga`$ may be assumed. 4. $`b=d`$ can be excluded, so $`bd`$ may be assumed. 5. $`a=c`$ can be excluded, so $`ac`$ may be assumed. When all the above claims are verified, we may restrict our considerations to the indices satisfying the conditions (6.14) $$a<d,ab,cd,gd,ga,bd,ac$$ These conditions imply that the indices $`\{a,d,g\}`$, $`\{a,d,c\}`$ and $`\{a,d,b\}`$ are distinct, so (6.14) implies $`g=b=c`$, as all indices take values in $`\{1,2,3\}`$. Therefore the required estimate for $`(\text{Ric}5)_{ad}`$ will hold if we can verify that it holds under (6.14) in conjunction with the condition $`g=b=c`$, which is the final step. Let $$F=t^{2+q_aq_d}(\text{Ric}5)_{ad}.$$ Case $`a=d`$: In case $`a=d`$, Lemma 6.2 and (6.12a) give $`F`$ $`t^{26ϵ}t^{2(q_1q_2q_3)}`$ $`t^{4q_16ϵ}`$ Therefore we may assume $`a<d`$ in the following. Further by (6.13), $`ab`$ and $`cd`$. Case $`g=d`$: Next consider the case $`g=d`$. Then using Lemma 6.2, (6.7) and (6.8e) we have $`F`$ $`=t^{2+q_aq_d}\stackrel{~}{g}^{bc}\stackrel{~}{g}^{fd}\stackrel{~}{\mathrm{\Gamma }}_{acf}\stackrel{~}{\mathrm{\Gamma }}_{bdd}`$ $`t^{2+q_aq_d}\stackrel{~}{g}^{bc}\stackrel{~}{g}^{fd}\stackrel{~}{\mathrm{\Gamma }}_{acf}t^{2ϵq_b}`$ $`t^{26ϵ+q_aq_dq_c}(t^{q_a}+t^{q_c}+t^{q_d}+t^{q_aq_cq_d}+t^{q_cq_dq_a}+t^{q_dq_aq_c})`$ $`t^{26ϵ}(t^{q_dq_c}+t^{q_aq_d2q_c}`$ $`+t^{q_a2q_dq_c}+t^{2(q_aq_dq_c)}+t^{2q_d}+t^{2q_c}).`$ By (6.13), $`dc`$, this gives $`Ft^{4q_16ϵ}`$ in the case $`g=d`$. Case $`g=a`$: Next consider the case $`g=a`$. Then we have $`F`$ $`=t^{2+q_aq_d}\stackrel{~}{g}^{bc}\stackrel{~}{g}^{fa}\stackrel{~}{\mathrm{\Gamma }}_{acf}\stackrel{~}{\mathrm{\Gamma }}_{bda}`$ use Lemma 6.2 and (6.8g) $`t^{25ϵ}t^{q_aq_d}\stackrel{~}{g}^{bc}(t^{q_a}+t^{q_c}+t^{q_c2q_a})Z_{bda}`$ use Lemma 5.1 $`t^{26ϵ}(t^{q_d}+t^{q_aq_dq_c}+t^{q_cq_aq_d})Z_{bda}`$ use (6.12a) and $`a<d`$, $`dc`$ $`t^{4q_16ϵ}`$ At this stage we may assume (6.15) $$a<d,ab,cd,gd,ga$$ Case $`b=d`$: Next consider the case $`b=d`$. In this case $`F`$ $`=t^{2+q_aq_d}\stackrel{~}{g}^{dc}\stackrel{~}{g}^{fg}\stackrel{~}{\mathrm{\Gamma }}_{acf}\stackrel{~}{\mathrm{\Gamma }}_{ddg}`$ $`t^{2+q_aq_d}\stackrel{~}{g}^{dc}\stackrel{~}{g}^{fg}\stackrel{~}{\mathrm{\Gamma }}_{acf}t^{2ϵ}(t^{q_d}+t^{q_g})`$ $`t^{26ϵ}(t^{q_aq_dq_c}+t^{q_aq_dq_g})Z_{acf}`$ use $`dc`$ and $`dg`$ from (6.15) $`t^{4q_16ϵ}`$ Case $`a=c`$: Next consider the case $`a=c`$. In this case we have using Lemma 6.2 $`F`$ $`t^{2+q_aq_d}\stackrel{~}{g}^{ba}\stackrel{~}{g}^{fg}t^{2ϵ}(t^{q_a}+t^{q_f})t^{2ϵ}Z_{bdg}`$ use (6.8e) $`t^{26ϵ}(t^{q_d}+t^{q_aq_dq_g})Z_{bdg}`$ use $`gd`$ from (6.15) and (6.12a) $`t^{4q_16ϵ}`$ At this stage we may assume (6.16) $$a<d,ab,cd,gd,ga,bd,ac$$ As discussed above, if we can prove that the required estimate holds under the condition $`g=c=b`$ we are done. Case $`g=b=c`$: Next consider the case $`g=b=c`$. In this case, after making the substitutions $`g=c`$ and $`b=c`$, $`F`$ $`=t^{2+q_aq_d}\stackrel{~}{g}^{cc}\stackrel{~}{g}^{fc}\stackrel{~}{\mathrm{\Gamma }}_{acf}\stackrel{~}{\mathrm{\Gamma }}_{cdc}`$ use Lemma 6.2 (6.17) $`t^{23ϵ}t^{q_aq_d}\stackrel{~}{g}^{fc}\stackrel{~}{\mathrm{\Gamma }}_{acf}(t^{q_d}+t^{q_c})`$ To estimate $`F`$ we must now consider the cases $`f=d`$, $`f=c`$, $`f=a`$ separately. Case $`g=b=c`$ and $`f=d`$: In case $`g=b=c`$ and $`f=d`$, we have from (6.17) $`F`$ $`t^{23ϵ}t^{q_aq_d}\stackrel{~}{g}^{dc}\stackrel{~}{\mathrm{\Gamma }}_{acd}(t^{q_d}+t^{q_c})`$ use (6.8e) and Lemma 6.2 $`t^{26ϵ}t^{q_aq_dq_c}Z_{acd}`$ use $`cd`$ from (6.16) and (6.12a) $`t^{4q_16ϵ}`$ Therefore we may exclude condition (6.16) in conjunction with $`f=d`$ from our considerations. Case $`g=b=c`$ and $`f=c`$: In case $`g=b=c`$ and $`f=c`$ we have from (6.17) $`F`$ $`t^{23ϵ}t^{q_aq_d}\stackrel{~}{g}^{cc}\stackrel{~}{\mathrm{\Gamma }}_{acc}(t^{q_d}+t^{q_c})`$ use Lemma 6.2 $`t^{24ϵ}t^{q_aq_d}t^{2ϵq_a}(t^{q_d}+t^{q_c})`$ $`=t^{26ϵ}(t^{2q_d}+t^{q_dq_c})`$ $`t^{4q_16ϵ}`$ Case $`g=b=c`$ and $`f=a`$: The only remaining case is $`g=b=c`$ and $`f=a`$. In this case we get from (6.17) using Lemma 6.2 $`F`$ $`t^{25ϵ}t^{q_aq_d}\stackrel{~}{g}^{ac}Z_{aca}(t^{q_d}+t^{q_c})`$ use (6.12b) and (6.8g) $`t^{26ϵ}(t^{q_a2q_d}+t^{q_aq_dq_c})t^{q_a}`$ $`t^{26ϵ}(t^{2q_d}+t^{q_dq_c})`$ $`t^{4q_16ϵ}`$ Therefore it now follows that under (6.16), the required estimate $$Ft^{4q_16ϵ}$$ holds and hence by the above argument it follows that this estimate holds under (6.13). This proves for a symmetric metric satisfying (5.8) the estimate (6.18) $$t^{2\alpha _b^a}R_b^at^{4q_16ϵ}$$ We wish to apply this to the symmetrized metric $`{}_{}{}^{S}g_{ab}^{}`$, which has the property that the rescaled symmetrized metric $`{}_{}{}^{S}\stackrel{~}{g}_{ab}^{}`$ satisfies (6.3). The estimate (6.18) translates to an estimate for a metric satisfying (6.3) after replacing $`\alpha _0`$ by $`\alpha _02ϵ`$, which by the definition of $`\alpha _0`$ satisfies $`\alpha _02ϵ>ϵ>0`$. Therefore we get in view of the discussion at the beginning of this section $$t^{2\stackrel{~}{\alpha }_d^a}{}_{}{}^{S}\stackrel{~}{R}_{d}^{a}t^{4p_19ϵ\alpha _0}$$ or $$t^{2\alpha _d^a}{}_{}{}^{S}R_{d}^{a}t^{4p_19ϵ\alpha _0}$$ Now recall $`ϵ=\alpha _0/4=\mathrm{min}_a\{p_a(x_0)\}/40`$ and $`q_1>20ϵ`$ by construction. This gives $`t^{2\alpha _d^a}{}_{}{}^{S}R_{d}^{a}`$ $`t^{4q_113ϵ}`$ $`t^{3q_1}`$ $`t^{3p_1}`$ where we used that fact that $`q_1p_1`$ by construction. This finishes the proof of ###### Lemma 6.3 (Curvature estimate). $$t^{2\alpha _b^a}{}_{}{}^{S}R_{b}^{a}t^{3p_1}$$ Now some estimates will be obtained for matter variables. These will be used to check that the right hand side of the second reduced system has the properties required for a Fuchsian system. Let $`w_a`$ be a one-form with the property that $`w_at^{q_a}`$. We have $`t^2\stackrel{~}{g}^{ab}_aw_b`$ $`=t^2\stackrel{~}{g}^{ab}\stackrel{~}{e}_aw_at^2\stackrel{~}{g}^{ab}\stackrel{~}{g}^{fg}\stackrel{~}{\mathrm{\Gamma }}_{abf}w_g`$ $`t^{2ϵ2q_a}+t^{24ϵ}t^{q_1+q_2q_3}t^{q_g}`$ $`t^{4q_14ϵ}t^{4p_14ϵ}`$ Here it has been assumed that $`w_a`$ behaves in a suitable way upon taking derivatives. In the context of the matter variables this will be the case for the relevant choices of $`w_a`$, namely $`e_a(\varphi )`$ and $`t^1u_a`$. Note that this estimate requires no use of cancellations, since it only uses the relation (6.12a) and not (6.8g). In this situation the estimate for a given quantity is never more difficult than that for the corresponding velocity dominated part, since the difference between the two is always of higher order. This gives the estimates required for the matter equations in the scalar field case. For the stiff fluid some more work is needed. The aim now is to estimate the terms on the right hand side of equations (5.18) by a positive power of $`t`$. For most of these terms no cancellations are required to get the desired estimate. There are only two exceptions to this and they will be discussed explicitly now. The first is the following combination which arises if the evolution equation for $`v_a`$ is written out explicitly: (6.19) $${}_{}{}^{0}\mu _{}^{1}e_a({}_{}{}^{0}\mu )\mu ^1e_a(\mu )$$ This expression is equal to (6.20) $$t^{\beta _1}[A^1\nu (1+A^1\nu t^{\beta _1})^1A^1(_aA+_a\nu )+A^1t^{\beta _1}_a\nu ]$$ which is $`O(t^{\beta _1})`$. The contribution of this expression to the right hand side of the evolution equation for $`v_a`$ is as a consequence $`O(t^{\beta _1\beta _2})`$ which shows that it is necessary to choose $`\beta _2<\beta _1`$. The second expression where a cancellation is necessary is $`1+t\text{tr}k`$. Now $`\text{tr}k=t^1+\text{tr}\kappa t^{1+\alpha _0}`$ and hence $`1+t\text{tr}k=\text{tr}\kappa t^{\alpha _0}`$. It follows that this expression is $`O(t^{\alpha _0})`$. The analysis of the other terms is rather straightforward, although lengthy, and will not be carried out explicitly here. However some comments may be useful. The terms which are a priori most difficult to estimate are those involving covariant derivatives of $`u_a`$. For those it is convenient to use the components in the rescaled frame $`\stackrel{~}{e}_a`$. For all other terms the original frame $`e_a`$ can be used straightforwardly. In order that all terms can be estimated by a positive power of $`t`$ it suffices to choose $`\beta _1`$ and $`\beta _2`$ small enough. One possible choice is $`\beta _2<\beta _1<q_15ϵ`$. In order to show that the reduced systems for the Einstein–scalar field system and the Einstein–stiff fluid systems, are in Fuchsian form, we need to show that $`t^{2\alpha _b^a}M_b^a=o(t^\delta )`$ for some $`\delta >0`$. We consider first the Einstein–scalar field case. In this case, $$M_b^a=g^{ac}e_c(\varphi )e_b(\varphi )$$ Arguing as above for the estimate of $`t^{2\alpha _b^a}{}_{}{}^{S}R_{b}^{a}`$, we have $$t^{2\alpha _b^a}M_b^at^{2\stackrel{~}{\alpha }_b^aϵ}\stackrel{~}{M}_{ab}$$ Therefore it is enough to show $$t^{2\stackrel{~}{\alpha }_b^a}\stackrel{~}{M}_{ab}t^{2ϵ}$$ By definition $`\stackrel{~}{M}_{ab}=t^{q_aq_b}e_a(\varphi )e_b(\varphi )`$ and hence $$\stackrel{~}{M}_{ab}t^{q_aq_b}$$ This give using $`\stackrel{~}{\alpha }_b^a=|q_aq_b|`$, $$t^{2\stackrel{~}{\alpha }_b^a}M_b^at^{3q_1}$$ For the scalar field case this gives, together with the above, $$t^{2\alpha _b^a}({}_{}{}^{S}R_{b}^{a}M_b^a)t^\delta ,\text{ for some }\delta >0\text{ }$$ which is the estimate required for proving that the second reduced system for the Einstein–scalar field system is in Fuchsian form. The argument for the Einstein–stiff fluid system is very similar. ## 7. The constraints The main aim of this section is to show that if a solution of the evolution equations is given which corresponds to a solution of the velocity dominated equations as in Theorem 2.1 or 2.2 then it satisfies the full constraints. It will also be shown how the existence of a large class of solutions of the velocity dominated constraints can be demonstrated. The first result on the propagation of the constraints relies on rough computations which prove the result in the case where all $`p_a`$ are close to $`1/3`$. An analytic continuation argument then gives the general case. ###### Lemma 7.1. Let $`({}_{}{}^{0}g_{ab}^{},{}_{}{}^{0}k_{ab}^{},{}_{}{}^{0}\varphi )`$ be a solution of the velocity dominated system as in the hypotheses of Theorem 2.1 with $`|p_a1/3|<\alpha _0/10`$. Let $`(\gamma ^a{}_{b}{}^{},\kappa ^a{}_{b}{}^{},\psi )`$ be a solution of (5.13) and (5.15) modelled on this velocity dominated solution and define $`g_{ab}`$, $`k_{ab}`$ and $`\varphi `$ by (3.1). Suppose that this solution satisfies the properties 1.-6. of the conclusions of the Theorem 2.1. Then the Einstein constraints are also satisfied. Proof Define: (7.1) $`C`$ $`=k_{ab}k^{ab}+(\text{tr}k)^2R16\pi \rho `$ (7.2) $`C_b`$ $`=_ak^a{}_{b}{}^{}_b(\text{tr}k)8\pi j_b`$ These quantities satisfy the evolution equations (7.3) $`_tC2(\text{tr}k)C`$ $`=^aC_a`$ (7.4) $`_tC_a(\text{tr}k)C_a`$ $`={\displaystyle \frac{1}{2}}_aC`$ Define rescaled quantities by $`\overline{C}=t^{2\eta _1}C`$ and $`\overline{C}=t^{1\eta _2}C_a`$ for some positive real numbers $`\eta _1`$ and $`\eta _2`$. Then the above equations can be written in the form: (7.5) $`t_t\overline{C}+\eta _1\overline{C}`$ $`=2[1+t\text{tr}k]\overline{C}t^{2\eta _1+\eta _2}^a\overline{C}_a`$ (7.6) $`t_t\overline{C}_a+\eta _2\overline{C}_a`$ $`=[1+t\text{tr}k]\overline{C}_a(1/2)t^{\eta _1\eta _2}_a\overline{C}`$ Choose $`\eta _1`$ and $`\eta _2`$ so that $`\eta _1\eta _2>0`$. The aim is to apply Theorem 4.2 to show that $`\overline{C}`$ and $`\overline{C}_a`$ vanish. In order to do this we should show that these two quantities vanish as fast as a positive power of $`t`$ as $`t0`$, that $`1+t\text{tr}k`$ vanishes like a positive power of $`t`$ and that the term $`^a\overline{C}_a`$ is not too singular. In obtaining these estimates it is necessary to use the behaviour of the derivatives of the solution mentioned in the remark following Theorem 4.2. Note that since the velocity dominated constraints are satisfied by assumption, it is enough to estimate the differences of the constraint quantities corresponding to the velocity dominated and full solutions, since these are in fact equal to $`C`$ and $`C_a`$. By property 2. of the conclusions of Theorem 2.1 it follows that $`1+t\text{tr}k=O(t^{\alpha _0})`$ which gives one of the desired statements. Similarly it follows that (7.7) $$k^{ab}k_{ab}+(\text{tr}k)^2=({}_{}{}^{0}k_{}^{ab})({}_{}{}^{0}k_{ab}^{})+(\text{tr}{}_{}{}^{0}k)^2+O(t^{2+\alpha _0})$$ It follows from the curvature estimates done in section 6 that the scalar curvature is also $`O(t^{2+\alpha _0})`$. Now consider the expression $`\rho {}_{}{}^{0}\rho `$. The components of the inverse metric can be estimated by $`t^{2+40\alpha _0}`$, so that the terms in $`\rho `$ involving spatial derivatives can be estimated by $`t^{2+\alpha _0}`$ as well. The difference of the time derivatives can be estimated by $`t^{2+\beta }`$. These are the estimates for the Hamiltonian constraint that will be needed. The estimates just carried out were independent of the restriction on the $`p_a`$ in the hypotheses of the lemma. The following estimates for the momentum constraint are of a cruder type and do use the restriction. First note that the metric and its inverse can be estimated by the powers of $`t`$ equal to $`2p_1`$ and $`2p_3`$ respectively. It follows from the assumption on the $`p_a`$ in the hypotheses of the theorem that $`2p_1>2/3\alpha _0/5`$ and $`2p_3>2/3\alpha _0/5`$. Using (3.4) then shows that the connection coefficients can be estimated in terms of the power $`2\alpha _0/5`$. The effect on the order of a term of taking a divergence can be estimated by the powers $`2\alpha _0/5`$ and $`2/33\alpha _0/5`$ for upper and lower indices respectively. The gradient of the mean curvature is $`O(t^{1+\alpha _0}\mathrm{log}t)`$ while the difference of $`j_a`$ is $`O(t^{1+\beta }\mathrm{log}t)`$. The difference of the divergence of the second fundamental form produces the power $`1+3\alpha _0/5`$. Evidently the last power and that containing $`\beta `$ are the limiting ones and determine the estimate for $`C_a`$. Similarly the divergence of $`C_a`$ can be estimated by the powers $`5/3`$ and $`5/3+\beta 3\alpha _0/5`$. Note that it follows from the definition of $`\alpha _0`$ that $`\alpha _0<1/30`$. Thus given the hypothesis of the lemma it can be concluded that $`\eta _1`$ and $`\eta _2`$ can be chosen in such a way that all terms on the right hand side of the propagation equations for the constraint quantities vanish like positive powers of $`t`$. Thus these equations are Fuchsian and the conclusion follows from Theorem 4.2. The analogue of this lemma with the scalar field replaced by a stiff fluid is also true and can be proved in the same way. Next the restriction on the exponents $`p_a`$ will be lifted. Consider a solution $`({}_{}{}^{0}g_{ab}^{},{}_{}{}^{0}k_{ab}^{},{}_{}{}^{0}\mu ,{}_{}{}^{0}v_{a}^{})`$ of the velocity dominated constraints for the Einstein-stiff fluid system. Let $`{}_{}{}^{0}\widehat{k}_{ab}^{}`$ be the trace-free part of $`{}_{}{}^{0}k_{ab}^{}`$. The velocity dominated constraints become: (7.8) $`{}_{}{}^{0}\widehat{k}_{}^{ab}{}_{}{}^{0}\widehat{k}_{ab}^{}+(2/3)(\text{tr}{}_{}{}^{0}k)^2`$ $`=16\pi \mu `$ (7.9) $`_a{}_{}{}^{0}\widehat{k}_{b}^{a}`$ $`=8\pi \mu u_b`$ Now let $`{}_{}{}^{\lambda }k_{ab}^{}=(1\lambda )\widehat{}^0k_{ab}+(1/3)(\text{tr}{}_{}{}^{0}k){}_{}{}^{0}g_{ab}^{}`$ and (7.10) $`{}_{}{}^{\lambda }\mu `$ $`=(1/16\pi )[(1\lambda )^2({}_{}{}^{0}\widehat{k}_{}^{ab})({}_{}{}^{0}\widehat{k}_{ab}^{})+(2/3)(\text{tr}{}_{}{}^{0}k)^2]`$ (7.11) $`{}_{}{}^{\lambda }\mu _{b}^{}`$ $`=2(1\lambda )^a{}_{}{}^{0}\widehat{k}_{ab}^{}[(1\lambda )^2({}_{}{}^{0}\widehat{k}_{}^{ab})({}_{}{}^{0}\widehat{k}_{ab}^{})+(2/3)(\text{tr}{}_{}{}^{0}k)^2]^1`$ Then $`{}_{}{}^{0}g_{ab}^{},{}_{}{}^{\lambda }k_{ab}^{},{}_{}{}^{\lambda }\mu ,{}_{}{}^{\lambda }u_{a}^{}`$ is a one parameter family of solutions of the velocity dominated constraints which depends analytically on the parameter $`\lambda `$. There exists a corresponding family of solutions of the velocity dominated evolution equations which also depends analytically on $`\lambda `$. Next, Theorem 4.2 provides a corresponding analytic family of solutions of the full evolution equations. (Cf. the second remark following that theorem.) These define constraint quantities depending analytically on $`\lambda `$. For $`\lambda `$ close to one Lemma 7.1 shows that these quantities are zero. Hence by analyticity they are zero for all values of $`\lambda `$, including $`\lambda =0`$. This means that the conclusion of Lemma 7.1 holds for all positive $`p_a`$ and a stiff fluid. Since any solution of the second reduced system for a scalar field defines a solution of the second reduced system for a stiff fluid, this extension also holds for the scalar field. A variant of the conformal method for solving the full Einstein constraints can be used to analyse the velocity dominated constraints. Consider the following set of free data: a Riemannian metric $`\overline{g}_{ab}`$, a symmetric trace-free tensor $`\sigma _{ab}`$ on $`S`$ and two scalar functions $`\overline{\varphi }`$ and $`\overline{\varphi }_t`$ on $`S`$. Next consider the following ansatz: (7.12) $`g_{ab}`$ $`=\omega ^4\overline{g}_{ab}`$ (7.13) $`k_{ab}`$ $`=(1/3)t_0^1g_{ab}+\omega ^2l_{ab}`$ (7.14) $`\varphi `$ $`=\omega ^2\overline{\varphi }`$ (7.15) $`\varphi _t`$ $`=\omega ^4\overline{\varphi }_t`$ where (7.16) $$l_{ab}=\sigma _{ab}+_aW_b+_bW_a(2/3)\overline{g}_{ab}\overline{g}^{cd}_cW_d$$ Putting this into (2.12a) and defining $`\overline{\rho }=\frac{1}{2}(\overline{\varphi }_t)^2`$ gives $`\rho =\omega ^8\overline{\rho }`$, a relation well known from the usual conformal method. As a result of the Hamiltonian constraint the function $`\varphi `$ satisfies the following algebraic analogue of the Lichnerowicz equation: (7.17) $$\omega ^{12}l_{ab}l_{cd}\overline{g}^{ac}\overline{g}^{bd}+\frac{2}{3}t_0^216\pi \omega ^8\overline{\rho }=0$$ Solving this comes down to looking for positive roots of the equation $`a\zeta ^3+b\zeta ^2c=0`$ where $`a`$ and $`b`$ are non-negative and $`c`$ is positive. The derivative of the function on the left hand side of this equation is $`\zeta (3a\zeta +2b)`$. Thus unless $`a`$ and $`b`$ are both zero the derivative has no positive roots. Moreover the function tends to plus infinity for large $`\zeta `$ and is negative at $`\zeta =0`$. Hence the equation has a unique solution for each $`a`$ and $`b`$ not both zero and if $`a`$ and $`b`$ depend analytically on some parameter then the solution does so too. If $`a`$ and $`b`$ are both zero then of course there is no positive solution. In the case of interest here $`a`$ and $`b`$ are both positive. The function $`\omega `$ is given by $`\omega =\mathrm{\Omega }(l_{ab}l_{cd}\overline{g}^{ac}\overline{g}^{bd},t_0,\overline{\rho })`$ where the analytic function $`\mathrm{\Omega }`$ is defined as the solution of the algebraic Lichnerowicz equation. The momentum constraint implies the elliptic equation (7.18) $$\overline{g}^{as}_a[_sW_b+_bW_s(2/3)\overline{g}_{sb}\overline{g}^{cd}_cW_d]=8\pi [\overline{j}_b2\overline{\varphi }_t\overline{\varphi }\omega _b\omega ]_a\sigma ^a_b$$ for $`W_a`$. Here $`\overline{j}_a=\overline{\varphi }_t_a\overline{\varphi }`$. Note that, when $`\omega `$ is expressed in terms of the function $`\mathrm{\Omega }`$ of the basic variables, it depends on the first derivatives of $`W_a`$. Thus the expression $`_a\omega `$ involves second derivatives of $`W_a`$ and is not simply a lower order term. Consider now the linearization of (7.18),with respect to $`W_a`$, where $`\omega `$ has been reexpressed using $`\mathrm{\Omega }`$. In particular, consider the linearization in the particular case where $`\overline{g}_{ab}`$ is the metric of constant negative curvature on a compact hyperbolic manifold, the tensor $`\sigma _{ab}`$ is zero, $`\overline{\varphi }`$ and $`\overline{\varphi }_t`$ are constant and the background value of $`W_a`$ is zero. Because $`\omega `$ is a function of an expression quadratic in $`W_a`$, the right hand side of (7.18) makes no contribution to the linearization. Since $`\overline{g}_{ab}`$ has no non-trivial conformal Killing vectors it follows from the standard theory of the York operator that the operator obtained by linearization of the equation (7.18) is invertible as a map between appropriate Sobolev spaces. Then an application of the implicit function theorem gives solutions of (7.18) for arbitrary choices of the free data sufficiently close (with respect to a Sobolev norm) to the particular free data at which the linearization was carried out. This shows the existence of solutions of the velocity dominated constraints which are as general as the solutions of the full Einstein constraints (at least in the crude sense of function counting). The conformal method can be applied in a similar way in the stiff fluid case and it turns out to be easier than in the scalar field case. This might seem paradoxical, since the scalar field problem can be identified with a subcase of the stiff fluid problem. The explanation is that it is difficult to identify which free data in the procedure for constructing stiff fluid data which will be presented correspond to data for a scalar field. The ansatz used is $`\mu =\omega ^8\overline{\mu }`$ and $`u_a=\omega ^2\overline{u}_a`$. This gives the scaling $`\rho =\omega ^8\overline{\omega }`$ and $`j_a=\omega ^6\overline{j}_a`$ which is often used in the conformal method. The quantities describing the geometry are scaled as in the case of the scalar field. The equations for $`\omega `$ and $`W_a`$ are very similar in both cases, with the notable difference that in the stiff fluid case the term involving the derivative of $`\omega `$ is missing from the equation for $`W_a`$. This means that the equation for $`W_a`$ is independent of $`\omega `$ and can be solved by standard theory, as long as the metric $`\overline{g}_{ab}`$ has no conformal Killing vectors. Once this has been done the algebraic equation for $`\omega `$ can be solved straightforwardly. ## 8. Discussion We have shown the existence of a family of solutions of the Einstein equations coupled to a scalar field or a stiff fluid whose singularity structure we can analyse. No symmetry assumptions are made and the solutions are general in the sense that they depend on the same number of free functions as general initial data for the same system on a regular Cauchy surface. These solutions agree with the picture of general spacetime singularities proposed by Belinskii, Khalatnikov and Lifshitz in two important ways. Firstly, the evolution at different spatial points decouples, in the sense that the solutions of the full equations are approximated near the singularity by a solution of a system of ordinary differential equations. Secondly there exists a Gaussian coordinate system which covers a neighbourhood of the singularity in which the singularity is situated at $`t=0`$. It is easily seen that the curvature invariant $`R_{\alpha \beta }R^{\alpha \beta }=64\pi ^2(_\alpha \varphi ^\alpha \varphi )^2`$ blows up uniformly for $`t0`$. In fact the leading term is proportional to $`A^4(x)t^4`$ and in the solutions we consider $`A`$ can never vanish, as a consequence of the Hamiltonian constraint. Thus these singularities are all consistent with the strong cosmic censorship hypothesis. The mean curvature of the hypersurfaces of constant Gaussian time tends uniformly to infinity as $`t0`$ so that the the singularity in crushing in the sense of . It then follows from well-known results that a neighbourhood of the singularity can be covered by a foliation consisting of constant mean curvature hypersurfaces. This is the most general class of spacetimes in which all these suggested properties of general spacetimes have been demonstrated. A subclass of these spacetimes is covered by the results of Anguige and Tod. The connection between their results and those of the present paper deserves to be examined more closely but intuitively their spacetimes should correspond to the case where, in our notation, the $`p_i`$ are everywhere equal to $`1/3`$. The spacetimes constructed have been shown to be general in the sense of function counting. It would, however, be desirable to prove that the assumption of analyticity of the data can be replaced by smoothness and that, this having been done, the spacetimes constructed include all those arising from a non-empty open set of initial data on a regular Cauchy surface which, in particular, contains the initial data for a Friedmann model. This would be a statement on the stability of the Friedmann singularity. A model for this kind of generalization is provided by the work of Kichenassamy on nonlinear wave equations. It was indicated in the introduction that the results on the Einstein-scalar equations can be interpreted in more than one way. The interpretation which has been emphasized here is that where the metric occurring in this system is considered to be the physical metric. In the interpretation in terms of string cosmology the physical metric is (up to a multiplicative constant) $`e^\varphi g_{\mu \nu }`$. This means that for $`A(x)`$ sufficiently negative the limit $`t0`$ does not correspond to a singularity at all, but rather to a phase which lasts for an infinite proper time. It is the time reverse of this situation which plays a role in the pre-big bang model. Another interpretation is in terms of the vacuum field equations in Brans-Dicke theory. This is very similar to the string cosmology case, with the difference that the conformal factor $`e^\varphi `$ is replaced by $`e^{C\varphi }`$ where $`C`$ is a constant which depends on the Brans-Dicke coupling constant. All the results in this paper have concerned the case of three space dimensions. There are reasons to believe that if the space dimension is at least ten then the vacuum Einstein equations allow stable quiescent singularities, similar in some ways to those of the Einstein-scalar field equations in three space dimensions. The techniques developed in this paper might allow this to be proved rigorously. It would also to be interesting to know what happens to the picture when further matter fields are added. There are several possibilities here. One is to add some other field, not directly coupled to the scalar field, to the Einstein-scalar field system. A second is to reinstate some of the extra fields (axion, moduli) which have been discarded in passing from the low energy limit of string theory to the Einstein-dilaton theory. A third is to add extra matter fields to the Brans-Dicke theory. Another direction in which the results on the Einstein-scalar field and Einstein-stiff fluid equations could be generalized is to start with situations where the solution has one Kasner exponent negative and investigate whether it moves (in the direction towards the singularity) towards the region where all Kasner exponents are non-negative. If this were true, then the singularities in generic solutions of these equations could be quiescent. The set of initial data concerned would be not just open, but also dense. This question is sufficiently difficult that it would seem advisable to first try and investigate it rigorously in the spatially homogeneous case. Acknowledgements This research was supported in part by the Swedish Natural Sciences Research Council (SNSRC), contract no. F-FU 4873-307, and the US National Science Foundation under Grant No. PHY94-07194. Part of the work was done while the authors were enjoying the hospitality of the Institute for Theoretical Physics, Santa Barbara. We gratefully acknowledge stimulating discussions with V. Moncrief which had an important influence on the development of the strategy used in this work.
warning/0001/gr-qc0001075.html
ar5iv
text
# GYROSCOPE PRECESSION IN CYLINDRICALLY SYMMETRIC SPACETIMES ## 1 Introduction As stressed by Bonnor in his review on the physical interpretation of vacuum solutions of Einstein’s field equations, relativists have not been diligent in interpreting such solutions. One way to palliate this deficiency of the physical content of the theory consists in providing link between the characteristic parameters of the solutions and quantities measured from well defined and physically reasonable experiments, providing thereby physical interpretation for those parameters. It is the purpose of this work to establish such a link for cylindrically symmetric spacetimes. The general form of the metric in this case was given by Lewis and describes a stationary spacetime.This metric can be split into two families called Weyl class and Lewis class.Here we shall restrict our study to the Weyl class where all parameters appearing in the metric are real. For the Lewis class these parameters can be complex. The corresponding static limit was obtained by Levi-Civita . The motivation for this choice is provided, on one hand, by the fact that the physical interpretation of the parametrers of these metrics is still a matter of discussion (see ), and on the other, by the fact that some of the parameters of these metrics are related to topological defects not entering into the expression of the physical components of curvature tensor. The physical(gedanken) experiment proposed here consists in observing the precession of a gyroscope under different conditions in such spacetimes. Specifically, we calculate the rate of rotation of a gyroscope at rest in the frame in which the metric is presented, and also, the total precession per revolution of a gyroscope circumventing the symmetry axis, along a circular path (geodesic or not). By doing so, the four parameters of the Weyl class, from the Lewis metric, become measurable, in the sense that they are expressed through quantities obtained from well defined and physically reasonable experiments (we are of course not discussing about the actual technical feasibility of such experiments). All calculations are carried out using the method proposed by Rindler and Perlick , and a very brief resume of which is given in the next section together with the notation and the specification of the spacetime under consideration.In section 3 we calculate the rate of precession of a gyroscope at rest in the original lattice, and in section 4 we obtain the precession per revolution relative to the original frame, of a gyroscope rotating round the axis of symmetry. Finally the results are discussed in the last section. ## 2 The spacetime and the Rindler-Perlick <br>method ### 2.1 The Lewis metric The Lewis metric can be written as $$ds^2=fdt^2+2kdtd\varphi +e^\mu (dr^2+dz^2)+ld\varphi ^2,$$ (1) where $`f=ar^{1n}{\displaystyle \frac{c^2}{n^2a}}r^{1+n},`$ (2) $`k=Af,`$ (3) $`l={\displaystyle \frac{r^2}{f}}A^2f,`$ (4) $`e^\mu =r^{(n^21)/2},`$ (5) with $$A=\frac{cr^{1+n}}{naf}+b.$$ (6) Observe that taking $`ds`$ in (1) to have dimension of length $`L`$ then $$[t]=[L]^{2n/(1+n)}.$$ (7) $$[r]=[L]^{2/(1+n)}.$$ (8) and $$[b]=[L]^{2n/(1+n)}.$$ (9) $$[c]=[L]^{2n/(1+n)}$$ (10) whereas $`n`$ and $`a`$ are dimensionless and $`e^\mu `$ is multiplied by a unit constant with dimensions $$[L]^{(2nn^21)/(n+1)}.$$ (11) The four parameters $`n,a,b`$ and $`c`$ can be either real or complex, and the corresponding solutions belong to the Weyl or Lewis classes respectively. Here we restrict our study to the Weyl class (not to confound with Weyl metrics representing static and axially symmetric spacetimes). The transformation $`d\tau =\sqrt{a}(dt+bd\varphi ),`$ (12) $`d\overline{\varphi }={\displaystyle \frac{1}{n}}[cdt+(nbc)d\varphi ],`$ (13) casts the Weyl class of the Lewis metric into the Levi-Civita metric (for recent discussions on these metrics see -, and references therein). However the transformation above is not valid globally, and therefore both metrics are equivalent only locally, a fact that can be verified by calculating the corresponding Cartan scalars . In order to globally transform the Weyl class of the Lewis metric into the static Levi-Civita metric, we have to make $`b=0`$. Indeed, if $`b=0`$ and $`c`$ is different from zero, (12) gives an admissible transformation for the time coordinate and (13) represents a transformation to a rotating frame. However, since rotating frames (as in special relativity) are not expected to cover the whole space-time and furthermore since the new angle coordinate ranges from $`\mathrm{}`$ to $`\mathrm{}`$, it has been argued in the past that both $`b`$ and $`c`$ ,have to vanish for (12) and (13) to be globally valid. This point of view is also reinforced by the fact that, assuming that only $`b`$ has to vanish in order to globally cast (1) into Levi-Civita, we are lead to the intriguing result that there is not dragging outside rotating cylinders (see comments at the end of section 3).We shall recall this question later. ### 2.2 The Rindler-Perlick method This method consists in transforming the angular coordinate $`\varphi `$ by $$\varphi =\varphi ^{}+\omega t,$$ (14) where $`\omega `$ is a constant. Then the original frame is replaced by a rotating frame. The transformed metric is written in a canonical form, $$ds^2=e^{2\mathrm{\Psi }}(dt\omega _idx^i)^2+h_{ij}dx^idx^j,$$ (15) with latin indexes running from 1 to 3 and $`\mathrm{\Psi },\omega _i`$ and $`h_{ij}`$ depend on the spatial coordinate $`x^i`$ only (we are omitting primes). Then, it may be shown that the four acceleration $`A_\mu `$ and the rotation three vector $`\mathrm{\Omega }^i`$ of the congruence of world lines $`x^i=`$constant are given by , $`A_\mu =(0,\mathrm{\Psi }_{,i}),`$ (16) $`\mathrm{\Omega }^i={\displaystyle \frac{1}{2}}e^\mathrm{\Psi }(deth_{mn})^{1/2}ϵ^{ijk}\omega _{k,j},`$ (17) where the comma denotes partial derivative. It is clear from the above that if $`\mathrm{\Psi }_{,i}=0`$, then particles at rest in the rotating frame follow a circular geodesic. On the other hand, since $`\mathrm{\Omega }^i`$ describes the rate of rotation with respect to the proper time at any point at rest in the rotating frame, relative to the local compass of inertia, then $`\mathrm{\Omega }^i`$ describes the rotation of the compass of inertia (the gyroscope) with respect to the rotating frame. Applying (14) to the original frame of (1), with $`t=t^{},r=r^{}`$ and $`z=z^{}`$, we cast (1) into the canonical form (15), where $`e^{2\mathrm{\Psi }}=f\omega ^2l2\omega k,`$ (18) $`\omega _i=(0,0,\omega _\varphi ),`$ (19) $`\omega _\varphi =e^{2\mathrm{\Psi }}(\omega l+k),`$ (20) $`h_{rr}=h_{zz}=e^\mu ,`$ (21) $`h_{\varphi \varphi }=l+e^{2\mathrm{\Psi }}\omega _\varphi ^2.`$ (22) From (12)-(17) and $$\mathrm{\Omega }(h_{ij}\mathrm{\Omega }^i\mathrm{\Omega }^j)^{1/2},$$ (23) we obtain that $$\mathrm{\Omega }=[h_{zz}(\mathrm{\Omega }^z)^2]^{1/2},$$ (24) where $$(\mathrm{\Omega }^z)^2=\frac{e^{4\mathrm{\Psi }}\omega _{\varphi ,r}^2}{4e^{2\mu }(fl+k^2)}.$$ (25) ### 2.3 Circular geodesics From (18) we obtain with the condition $`\mathrm{\Psi }_{,i}=0`$ that $$\omega =\frac{k__r\pm (f_{,r}l_{,r}+k_{,r}^2)^{1/2}}{l_{,r}},$$ (26) which yields the expression for the angular velocity of a particle on a circular geodesic in Lewis metric (see (49) in ). Now substituting (2)-(5) into (26) we obtain $$\omega =\left(\pm \omega _0+\frac{c}{n}\right)\left[1b\left(\pm \omega _0+\frac{c}{n}\right)\right]^1,$$ (27) where $`\omega _0`$ is the angular velocity when the spacetime is static, $`b=c=0`$, given by Levi-Civita’s metric, $$\omega _0^2=\frac{1n}{1+n}a^2r^{2n}.$$ (28) We can calculate the tangential velocity $`W`$ of the circular geodesic particles (see (53) in ), $$W=\frac{\omega (fl+k^2)^{1/2}}{f\omega k}.$$ (29) Substituting (2)-(5) and (28) into (29), we obtain $$W=\left(\pm \omega _0+\frac{c}{n}\right)\left(ar^n\pm \frac{c}{na}\omega _0r^n\right)^1.$$ (30) It is worth noticing the fact, which follows from (27) and (30), that $`b`$ and $`c`$ affect the angular velocity $`\omega `$, while for the tangential velocity $`W`$ only $`c`$ plays a role. ## 3 Precession of a gyroscope at rest in the original latice To calculate the precession in this case, we only have to put $`\omega =0`$ in (17)-(24) (see for a similar case). Then, we obtain after simple calculations, $$\mathrm{\Omega }=\frac{e^{\mu /2}}{2f}\frac{|kf_{,r}fk_{,r}|}{(fl+k^2)^{1/2}},$$ (31) or, using (2)-(5) in (31), $$\mathrm{\Omega }=cr^{(1n^2)/4}\left(ar^{1n}\frac{c^2}{n^2a}r^{1+n}\right)^1.$$ (32) Thus the parameter $`c`$, appears to be essential in the precession of a gyroscope at rest in the frame of (1), whereas $`n`$ and $`a`$ just modify its absolute value. This fact reinforces the interpretation of $`c`$, already given in and , in the sense that it represents the vorticity of the source, when descibed by a rigidly rotating anisotropic cylinder and that it provides the dragging correction to the angular velocity of a particle in circular orbits in Lewis spacetime . However, here there was no need to specify the source that produces the field and, on the other hand, although some kind of frame dragging effect may also be related to $`b`$ (see (60) in ), this last parameter does not play any role in the precession under consideration. Another interesting point about (32) is that if $`c`$ is small, it becomes $$\mathrm{\Omega }\frac{c}{a}r^{(1n)(n3)/4},$$ (33) and we observe that $`n=3`$ produces a constant $`\mathrm{\Omega }`$, independent of $`r`$. It is known that when $`b=c=0`$ and $`n=3`$ the metric becomes locally Taub’s plane metric . This fact suggests that the gravitational potential becomes constant, not modifying the precession with respect to its distance. Next, introducing $$\beta \frac{c}{na},$$ (34) we can write (32) as $$\mathrm{\Omega }=\frac{\beta nr^{(1n)(n3)/4}}{1\beta ^2r^{2n}}.$$ (35) Then performing a series of measurements of $`\mathrm{\Omega }`$ for different values of $`r`$, we can in principle obtain $`n`$ and $`\beta `$ by adjusting these parameters to the obtained curve $`\mathrm{\Omega }=\mathrm{\Omega }(r)`$, which in turn allows for obtaining the value of $`c/a`$. In the Levi-Civita metric, $`b=c=0`$, the gyroscope at rest will not precess, as expected for a vacuum static spacetime (for the electrovac case however, this may change ). All these comments above (after equation (32)), are valid as long as we adopt the point of view that eqs.(12)-(13) globally transform (1) into the Levi-Civita spacetime , if and only if $`b=c=0`$. However if we adopt the point of view that only $`b`$ has to vanish for that transformation to be globally valid, then the precession given by (32) is just a coordinate effect,implying that there is not frame dragging in the Lewis space-time (Weyl class), since the precession of the gyroscope is due to the fact that the original frame of (1) is rotating itself. This is quite a surprising result, if we recall that material sources for (1) consist in steadly rotating fluids (see and ).Furthermore the vorticity of the source given in is, at the boundary surface, proportional to $`c`$ (not $`b`$). ## 4 Precession of gyroscope moving in a circle around the axis of symmetry According to the meaning of $`\mathrm{\Omega }`$ given above, it is clear that the orientation of the gyroscope, moving around the axis of symmetry, after one revolution, changes by $$\mathrm{\Delta }\varphi ^{}=\mathrm{\Omega }\mathrm{\Delta }\tau ,$$ (36) where $`\mathrm{\Delta }\tau `$ is the proper time interval corresponding to one period. Then from (15), $$\mathrm{\Delta }\varphi ^{}=2\pi \frac{\mathrm{\Omega }e^\mathrm{\Psi }}{\omega },$$ (37) as measured in the rotating frame. In the original system, we have $$\mathrm{\Delta }\varphi =2\pi \left(1\frac{\mathrm{\Omega }e^\mathrm{\Psi }}{\omega }\right).$$ (38) To calculate (24) and (38) for the metric (1) we first obtain from (18) and (19) using (2)-(5), $`e^{2\mathrm{\Psi }}=aM^2r^{1n}{\displaystyle \frac{N^2}{n^2a}}r^{1+n},`$ (39) $`\omega _{\varphi ,r}=2MNe^{4\mathrm{\Psi }}r,`$ (40) where $`M=1+b\omega ,`$ (41) $`N=n\omega c(1+b\omega ).`$ (42) Now substituting (39), (40) and (25) into (24) and (38), we obtain, $`\mathrm{\Omega }=MNr^{(1n^2)/4}\left(M^2ar^{1n}{\displaystyle \frac{N^2}{n^2a}}r^{1+n}\right)^1,`$ (43) $`\mathrm{\Delta }\varphi =2\pi \left[1{\displaystyle \frac{MN}{\omega }}r^{(1n^2)/4}\left(M^2ar^{1n}{\displaystyle \frac{N^2}{n^2a}}r^{1+n}\right)^{1/2}\right].`$ (44) When particle follows a circular geodesic around the axis, then the angular velocity is (27) and then (43) and (44) become, $`\mathrm{\Omega }={\displaystyle \frac{1}{2}}(1n^2)^{1/2}r^{(3+n^2)/4},`$ (45) $`\mathrm{\Delta }\varphi =2\pi \left\{1{\displaystyle \frac{n(1n)^{1/2}ar^n}{n(1n)^{1/2}ar^n+(1+n)^{1/2}c}}\left[{\displaystyle \frac{n(1+n)}{2a}}\right]^{1/2}r^{(1n)^2/4}\right\}.`$ (46) Surprisingly neither $`b`$ nor $`c`$ enter into the expression (45) for $`\mathrm{\Omega }`$. Also, the expression (46) is unaffected by $`b`$. If $`\omega _0r1`$ and $`c\omega _0`$, we have from (46), $$\mathrm{\Delta }\varphi \delta +3\pi \frac{\omega _0^2r^2}{a^{5/2}}+2\pi \frac{c}{\sqrt{a}\omega _0},$$ (47) and if $`a=1`$ we have $`\mathrm{\Delta }\varphi 3\pi \omega _0^2r^2+2\pi c/\omega _0`$, which coincides with the Schiff precession in the Kerr spacetime, if we identify the Kerr parameter with $`c`$. We shall now apply (43)-(46) to some specific cases. ### 4.1 Levi-Civita spacetime case When in (1) $`b=c=0`$ we have the static Levi-Civita spacetime then (43) and (44) reduce to $`\mathrm{\Omega }={\displaystyle \frac{na\omega r^{(1n)(n3)/4}}{a^2\omega ^2r^{2n}}},`$ (48) $`\mathrm{\Delta }\varphi =2\pi \left[1{\displaystyle \frac{n\sqrt{a}r^{(1n)^2/4}}{(a^2\omega ^2r^{2n})^{1/2}}}\right].`$ (49) Let us now assume that the trajectory of the gyroscope is a geodesic, then $`\omega =\omega _0`$ given by (28), we have from (48) and (49), $`\mathrm{\Omega }={\displaystyle \frac{1}{2}}(1n^2)^{1/2}r^{(3+n^2)/4},`$ (50) $`\mathrm{\Delta }\varphi =2\pi \left\{1\left[{\displaystyle \frac{n(1+n)}{2a}}\right]^{1/2}r^{(1n)^2/4}\right\}.`$ (51) In the case $`\omega _0r1`$, (51) becomes $$\mathrm{\Delta }\varphi \delta +3\pi \frac{\omega _0^2r^2}{a^{5/2}},$$ (52) and if $`a=1`$, we have $`\mathrm{\Delta }\varphi 3\pi \omega _0^2r^2`$, which coincides with the Fokker-de Sitter precession in the Schwarzschild spacetime. It is interesting to note that if $`n=0`$, which corresponds to the null circular geodesics as can be seen from (30) when $`W=1`$, we have from (49) that $`\mathrm{\Delta }\varphi =2\pi `$. This behaviour means that the precession becomes so large, that independently of $`\omega `$, the orientation of the gyroscope is locked to the lattice of the rotating frame. Exactly the same behaviour appears in the Schwarzschild spacetime. ### 4.2 Flat spacetime case From the Cartan scalars we see that only $`n`$ helps to curve the spacetime . When $`n=1`$ the spacetime becomes flat and (43) and (44) become, $`\mathrm{\Omega }={\displaystyle \frac{a\stackrel{~}{\omega }}{a^2\stackrel{~}{\omega }^2r^2}},`$ (53) $`\mathrm{\Delta }\varphi =2\pi \left[1{\displaystyle \frac{(1+bc)\sqrt{a}\stackrel{~}{\omega }}{(\stackrel{~}{\omega }+c)(a^2\stackrel{~}{\omega }^2r^2)^{1/2}}}\right],`$ (54) where $`\stackrel{~}{\omega }`$ is the effective coordinate rate of rotation, $$\stackrel{~}{\omega }=\frac{\omega }{1+b\omega }c.$$ (55) From (54) we see that $`b`$ and $`c`$ affect $`\stackrel{~}{\omega }`$ by increasing (decreasing) it when they are in the same (opposite) sense of rotation with respect to $`\omega `$. Let us first consider the case $`b=c=0`$, then (48) and (49) become $`\mathrm{\Omega }={\displaystyle \frac{a\omega }{a^2\omega ^2r^2}},`$ (56) $`\mathrm{\Delta }\varphi =2\pi \left[1{\displaystyle \frac{\sqrt{a}}{(a^2\omega ^2r^2)^{1/2}}}\right].`$ (57) These expressions, (56) and (57), put in evidence the influence of $`a`$ on the Thomas precession of a gyroscope moving around a string with linear energy density $`\lambda `$ given by $$\lambda =\frac{1}{4}\left(1\frac{1}{\sqrt{a}}\right).$$ (58) We recall that $`a`$ changes the topological structure of the spacetime, giving rise to an angular deficit $`\delta `$ equal to $$\delta =2\pi \left(1\frac{1}{\sqrt{a}}\right).$$ (59) In the case $`\omega r1`$, (57) becomes $$\mathrm{\Delta }\varphi \delta \pi \frac{\omega ^2r^2}{a^{5/2}},$$ (60) and if $`a=1`$ we have the usual Thomas precession $`\mathrm{\Delta }\varphi \pi \omega ^2r^2`$. If $`b=0`$ and $`c\omega `$ (54) becomes $$\mathrm{\Delta }\varphi 2\pi \left[1\frac{\sqrt{a}}{(a^2\omega ^2r^2)^{1/2}}+\frac{c}{\omega }\frac{a^2+\omega ^2r^2}{(a^2\omega ^2r^2)^{3/2}}\right],$$ (61) and if $`c=0`$ and $`b\omega 1`$ we have from (54) $$\mathrm{\Delta }\varphi 2\pi \left[1\frac{\sqrt{a}}{(a^2\omega ^2r^2)^{1/2}}+b\omega \frac{\omega ^2r^2}{(a^2\omega ^2r^2)^{3/2}}\right].$$ (62) We see from (60) and (62) that if the order of magnitude of $`c/\omega `$ and $`b\omega `$ are equal, $`O(c/\omega )=O(b\omega )`$, then the contribution of $`c`$ is larger than $`b`$ to the precession. The expression (62) exhibits the modifications on the Thomas precession, associated with the topological defect created by $`b`$. It is worth noticing that a quantum scalar particle moving around a spinning cosmic string, exhibits a phase factor proportional to $`b\sqrt{a}`$, an evident reminiscence of the Aharonov-Bohm effect . ## 5 Conclusions We have been able to establish a set of expressions linking the parameters of the Lewis metric to quantities obtained from the observation of gyroscope’s precession. In the particular case of a gyroscope at rest in the original lattice, the relevance of parameter $`c`$ is clearly illustrated. Curiously enough, the parameter $`b`$ does not enter into the expression of the angular velocity of precession (see discussion above on this point). In the case of the gyroscope rotating around the axis of symmetry, we obtain that in the Levi-Civita case the precession vanishes at the photon orbit. A similar result is known to happen for the Schwarzschild and the Ernst spacetimes. However in the general case, $`b0`$ and $`c0`$, the same result is observed, which is different from previous results found in stationary spacetimes . This happens probably due to the fact already mentioned, that the Weyl class of the Lewis metric, is a rather sui generis class of stationary metrics, since it is locally static. For the special cases with $`n=1`$, we found how different parameters affect the Thomas precession, providing at the same time a tool for their measurement. We would like to conclude with the following comment. Except for $`n`$, neither of the parameters $`a,b`$ and $`c`$ of the Lewis metric enter into the expressions for the physical components of the Riemann tensor . This implies that they cannot be measured by means of tidal forces observations. Therefore gyroscope precession experiments (i.e. experiments leading to the measuring of the rate of precession of a gyroscope at rest in the frame of a given metric and /or the total precession per revolution of a gyroscope circumventing the source of such metric) provide a good alternative for observing those aspects of gravitation not directly related to the curvature. We have in mind not only topological deffects, as is the case here, but other issues appering in the study of gravity and which are not directly related with the value of the physical components of the Riemann tensor (see for example , , , and references therein).In the case of real experiments which are now being contemplated as the GP-B in the solar system, it might in principle (we are completely ignorant about the accuracy of such experiment) help to determine what is, among all stationary solutions, the spacetime associated to a rotating source (which we expect to be the Kerr metric).
warning/0001/quant-ph0001086.html
ar5iv
text
# Decoherence in QED at finite temperature ## 1 Introduction It has been known for a long time that thermal photons can substantially disturb electron beams in an accelerator . It also has been suggested that the problem of the classical limit of quantum mechanics (at least the problem of decoherence) can be solved in the presence of a reservoir of photons . Some models of an interaction of a quantum system with a reservoir confirming the decoherent behaviour have been discussed in refs.. These models are treated as an approximation to the quantum electrodynamics (QED). We discussed QED and some approximations to QED in ref.. We pointed out a difficulty (resulting from the ultraviolet singularity) with some approximations to QED when applied in a discussion of the decoherence. The singular part of the QED propagator does not contribute to the decoherence in formal semiclassical calculations. In this paper we discuss the regular part. We calculate the time evolution of the density matrix of wave packets in an approximation of a small charge $`e`$ treating the particle dynamics in a semiclassical approximation. We neglect the vacuum fluctuations considering the effect of thermal photons on the evolution of the density matrix. We show a decoherence effect of the thermal photons. The vacuum fluctuations lead to divergencies at short distances. After a renormalization the contribution of vacuum fluctuations to decoherence is negligible for a large time. The effect of photons on decoherence has been discussed first by Joos and Zeh . The general arrangement for interference experiments in an environment has been considered in . L. Ford discussed a change of the vacuum fluctuations with a change of the environment’s geometry (the Casimir effect) and its role in the interference experiments. In such a case the relative strength of the vacuum fluctuations rather than its intrinsic value (which is infinite) is relevant. The decoherent effect of the black body radiation has been studied by Stapp but we disagree with his results and approximation methods. We are interested in quantum electrodynamics at finite temperature T determined by the density matrix (the Gibbs state) $$\rho =\left(Tr\left(\mathrm{exp}\left(\beta H_R\right)\right)\right)^1\mathrm{exp}(\beta H_R)$$ where $`\frac{1}{\beta }=KT`$, $`K`$ is the Boltzmann constant and $`H_R`$ is the Hamiltonian for the quantum free electromagnetic field. We compute the correlation functions explicitely $$\begin{array}{c}G_\beta (0,𝐱;t,𝐱^{})_{jl}Tr\left(A_j(0,𝐱)A_l(t,𝐱^{})\rho \right)=\hfill \\ \frac{\mathrm{}c}{\pi ^2}𝑑𝐤|𝐤|^1\mathrm{cos}\left(\left(𝐱𝐱^{}\right)𝐤\right)\delta _{jl}^{tr}(𝐤)\hfill \\ \left(cos(c|𝐤|t)\left(\frac{1}{2}+\left(\mathrm{exp}(\beta \mathrm{}c|𝐤|)1\right)^1\right)\frac{i}{2}sin(c|𝐤|t)\right)\hfill \end{array}$$ (1) where $$\delta _{jl}^{tr}(𝐤)=\delta _{jl}k_jk_l|𝐤|^2$$ The term $`\frac{1}{2}\mathrm{}c|𝐤|`$ corresponds to the zero point energy (of vacuum fluctuations) whereas $`\mathrm{}c|𝐤|(\mathrm{exp}(\mathrm{}c|𝐤|\beta )1)^1`$ is the average energy of thermal photons with the wave number $`𝐤`$. The vacuum fluctuation (noise) is a measurable effect and in general cannot be neglected. The virtual photons corresponding to the vacuum fluctuations are not directly observable. Then, the thermal photons are described by the Green’s function $`G_{th}=G_\beta G_{\mathrm{}}`$ (note that $`G_{th}`$ is real whereas $`G_\beta `$ and $`G_{\mathrm{}}`$ are complex; in quantum field theory the imaginary part of the Green’s function is related to the pair creation and annihilation, so subtracting $`G_{\mathrm{}}`$ means that the processes of pair creation and annihilation are neglected) $$\begin{array}{c}G_{th}(𝐱,𝐱^{},t)_{jl}=\frac{\mathrm{}c}{2\pi ^2}𝑑𝐤|𝐤|^1\delta _{jl}^{tr}(𝐤)\mathrm{cos}\left(\left(𝐱𝐱^{}\right)𝐤\right)\mathrm{cos}\left(ct|𝐤|\right)\left(\mathrm{exp}\left(\beta \mathrm{}c|𝐤|\right)1\right)^1\hfill \end{array}$$ (2) Let us note that $`G_{th}`$ determines the real Gaussian random field. Subsequent computations can be performed either in the Fock space or by means of the functional integration . We do not explain the equivalence of both methods here but refer to our earlier paper . For a small $`|𝐱𝐱^{}|`$ $$\begin{array}{c}G_{th}(𝐱,𝐱^{},t)_{jl}G_{th}(0,0,t)_{jl}=\delta _{jl}\frac{4}{3}\mathrm{}c\pi ^1_0^{\mathrm{}}𝑑kkcos(ckt)\left(\mathrm{exp}(\beta \mathrm{}ck)1\right)^1\hfill \end{array}$$ (3) is approximately $`𝐱`$-independent. ## 2 The semiclassical approximation We approach the semiclassical limit of the wave function in the standard way treating the electromagnetic field $`𝐀`$ as a classical field . Then, the quantum electromagnetic field is realized as a random field with the covariance (2) . A solution of the Schrödinger equation $$i\mathrm{}_t\psi (t,𝐱)=\frac{1}{2m}(i\mathrm{}+\frac{e}{c}𝐀_t)^2\psi (t,𝐱)$$ (4) with the initial condition $`\psi =exp(\frac{i}{\mathrm{}}W)\varphi `$ can be related to the solution of the Hamilton-Jacobi equation $`W_t`$ $$_tW_t+\frac{1}{2m}(W_t+\frac{e}{c}𝐀_t)^2=0$$ (5) with the initial condition $`W_{t=0}(𝐱)=W(𝐱)`$. We express $`\psi _t`$ in the form $$\psi _t\chi _t\varphi _t=exp(\frac{i}{\mathrm{}}W_t)\varphi _t$$ Then, $`\psi _t`$ is the solution of eq.(4) if and only if $`\varphi _t`$ is the solution of the equation $$_t\varphi _t=\frac{i\mathrm{}}{2m}\mathrm{}\varphi _t\frac{1}{m}(W_t+\frac{e}{c}𝐀_t)\varphi _t\frac{1}{2m}(\mathrm{}W_t+\frac{e}{c}div𝐀_t)\varphi _t$$ (6) with the initial condition $`\varphi `$. In a formal limit $`\mathrm{}0`$ the term $`\mathrm{}\varphi `$ can be neglected. In such an approximation the solution of the Schrödinger equation (4) is expressed by the classical flow starting from $`𝐱`$ (here $`0st`$) $$\frac{d𝐲_s}{ds}=\frac{1}{m}\left(W_{ts}(𝐲_s)+\frac{e}{c}𝐀_{ts}(𝐲_s)\right)$$ (7) Till $`o(\mathrm{})`$ terms we have $$\begin{array}{c}\psi (t,𝐱)=exp\left(\frac{i}{\mathrm{}}W_t(𝐱)\right)exp\left(_0^t\frac{1}{2m}\left(\mathrm{}W_{ts}(𝐲_s)+\frac{e}{c}div𝐀_{ts}(𝐲_s)\right)𝑑s\right)\varphi \left(𝐲_t(𝐱)\right)\hfill \end{array}$$ If we know the trajectory (e.g., from the Hamilton equations) then we can compute $`W_t`$ $$W_t(𝐱)=W(𝐲_t(𝐱))+_0^t\left(\frac{m}{2}(\frac{d𝐲}{ds})^2+\frac{e}{c}𝐀_s(𝐲_s)\frac{d𝐲}{ds}\right)𝑑s$$ (8) From the correlation functions (2) of $`𝐀`$ it follows that the assumption that $`𝐀(t,𝐱)`$ is $`𝐱`$-independent is a good approximation for a non-relativistic motion. For a wave packet of momentum $`𝐩`$ we have $`W=\mathrm{𝐩𝐱}`$. Hence, approximately $$\frac{d𝐲_s}{ds}=\frac{1}{m}\left(𝐩+\frac{e}{c}𝐀_{ts}\right)$$ (9) This is a consistent approximation because for the approximate solution of the Hamilton-Jacobi equation $`W_s(𝐱)`$ is space independent . Then, for $`𝐀`$ which is space independent we have the exact solution of eq.(5) $$W_t(𝐱)=\mathrm{𝐩𝐱}\frac{t}{2m}𝐩^2\frac{e}{2mc}𝐩_0^t𝐀_\tau 𝑑\tau \frac{e^2}{2mc^2}_0^t𝐀_\tau ^2𝑑\tau $$ (10) As a result of the evolution in an environment of photons (which are not under an observation) the pure state $$\psi =\mathrm{exp}(iW/\mathrm{})\varphi $$ after an average over the states of the quantum electromagnetic field is transformed into a mixed state with the density matrix $$\rho _t(𝐱,𝐱^{})=\overline{\psi }(𝐱)\psi (𝐱^{})$$ (11) where the average is over the electromagnetic field. We combine the approximation (9)-(10) of a space independent $`𝐀`$ with the exact $`W_t`$ (8) in the following approximate expression for $`W_t`$ $$W_t(𝐱)=\mathrm{𝐩𝐱}\frac{t}{2m}𝐩^2\frac{e}{2mc}𝐩_0^t𝐀_\tau (𝐱\frac{\tau }{m}𝐩)𝑑\tau $$ (12) This expression results from a solution of the equations of motion (7) to the lowest (zeroth) order in $`e`$ . Subsequently, $`W_t`$ in eq.(8) is calculated to the first order in $`e`$. Inserting $`W_t`$ (12) into eq.(11) we obtain $$\begin{array}{c}\rho _t(𝐱,𝐱^{})\mathrm{exp}\left((i\mathrm{𝐩𝐱}^{}i\mathrm{𝐩𝐱})/\mathrm{}\right)\overline{\varphi (𝐱\frac{t}{m}𝐩)}\varphi (𝐱^{}\frac{t}{m}𝐩)\mathrm{exp}(S)=\hfill \\ \mathrm{exp}\left((i\mathrm{𝐩𝐱}^{}i\mathrm{𝐩𝐱})/\mathrm{}\right)\overline{\varphi (𝐱\frac{t}{m}𝐩)}\varphi (𝐱^{}\frac{t}{m}𝐩)\hfill \\ \mathrm{exp}(\frac{e^2}{m^2c^2\mathrm{}^2}_0^t𝐩G_{th}((s\tau )𝐩/m,s\tau )𝐩dsd\tau \hfill \\ +\frac{e^2}{2m^2c^2\mathrm{}^2}_0^t𝐩G_{th}(𝐱𝐱^{}+(s\tau )𝐩/m,s\tau )𝐩𝑑s𝑑\tau \hfill \\ +\frac{e^2}{2m^2c^2\mathrm{}^2}_0^t𝐩G_{th}(𝐱^{}𝐱+(s\tau )𝐩/m,s\tau )𝐩dsd\tau )\hfill \end{array}$$ (13) where $`\mathrm{exp}(S)`$ denotes the last factor in eq.(13). If the vacuum fluctuations were to be taken into account then we would need to make the replacement $`G_{th}G_\beta =G_{th}+i\mathrm{}_F`$, where $`\mathrm{}_F`$ is the Feynman causal propagator (in the notation of Bjorken and Drell ). ## 3 The estimates of the evolution of the density matrix We perform first the integral over time in eq.(13). We denote $`𝐲=𝐱𝐱^{}`$, introduce the spherical coordinates $`d𝐤=dkk^2d\theta \mathrm{sin}\theta d\varphi `$ and write $`S`$ in the form $$\begin{array}{c}S=\frac{e^2}{2\pi m^2c\mathrm{}}_0^{\mathrm{}}𝑑kk_0^\pi 𝑑\theta \mathrm{sin}\theta I(\theta ,k,p)\left(\mathrm{exp}\left(\beta \mathrm{}ck\right)1\right)^1\hfill \end{array}$$ (14) We restrict ourselves to $`|𝐩|mc`$ then $$\begin{array}{c}I(\theta ,k,p)=c^2k^2|𝐩|^2(1\mathrm{cos}^2\theta )\hfill \\ \left(2\left(1\mathrm{cos}\left(tck\right)\right)2\mathrm{cos}\mathrm{𝐤𝐲}+\mathrm{cos}\left(\mathrm{𝐤𝐲}+ckt\right)+\mathrm{cos}\left(\mathrm{𝐤𝐲}ckt\right)\right)\hfill \end{array}$$ (15) If $`𝐲`$ is large then on the basis of eq.(15) the $`𝐲`$-dependent terms are small as a function of $`t`$ in comparison with other terms, because we have an additional oscillation in $`𝐲`$ which makes such a term negligible. The main contribution to $`S`$ for a large $`𝐲`$ is (we write $`\alpha =\mathrm{cos}\theta `$) $$\begin{array}{c}S=\frac{2e^2}{m^2c^3\mathrm{}\pi }𝐩^2_1^1𝑑\alpha (1\alpha ^2)_0^{\mathrm{}}𝑑kk^1\left(1\mathrm{cos}\left(tck\right)\right)\left(\mathrm{exp}\left(\beta \mathrm{}ck\right)1\right)^1\hfill \end{array}$$ (16) For a small time $`t`$ ( and a large $`𝐲`$) we obtain $$\begin{array}{c}|\rho _t|\mathrm{exp}\left(\frac{2}{3}\frac{e^2}{m^2\mathrm{}c\pi }𝐩^2t^2_0^{\mathrm{}}𝑑kk\left(\mathrm{exp}\left(\beta \mathrm{}ck\right)1\right)^1\right)\hfill \\ \mathrm{exp}\left(B\frac{e^2}{\mathrm{}c}(𝐩t/m)^2l_{dB}^2\right)\hfill \end{array}$$ (17) where $`B`$ is a constant of order 1 and $$l_{dB}=c\mathrm{}\beta $$ denotes the thermal de Broglie wave length at temperature T . The result (17) means that the decoherence (or simply the effect of the electromagnetic environment) is visible after a particle makes a path comparable with de Broglie wave length. Let us consider now a large $`t0`$ in eq.(16). We apply the formula $$1\mathrm{cos}w=w_0^1𝑑\gamma \mathrm{sin}(\gamma w)$$ and the formula 3.911 of Gradshtein and Ryzhik $$_0^{\mathrm{}}𝑑u\mathrm{sin}(au)\left(\mathrm{exp}(\beta u)1\right)^1=\frac{\pi }{2\beta }\mathrm{coth}(\frac{\pi a}{\beta })\frac{1}{2a}$$ Then, we obtain $$\begin{array}{c}S=\frac{4}{3}\frac{te^2}{m^2c^2\mathrm{}\pi }𝐩^2_0^1𝑑\gamma _0^{\mathrm{}}𝑑k\mathrm{sin}\left(tck\gamma \right)\left(\mathrm{exp}\left(\beta \mathrm{}ck\right)1\right)^1\hfill \\ =\frac{4}{3}\frac{e^2}{m^2c^3\mathrm{}\pi }𝐩^2_0^t𝑑\gamma \left(\frac{\pi }{\beta \mathrm{}}\mathrm{coth}(\frac{\pi \gamma }{\beta \mathrm{}})\frac{1}{\gamma }\right)\hfill \\ =\frac{4}{3}\frac{e^2}{m^2c^3\mathrm{}\pi }𝐩^2\mathrm{ln}\left(\frac{\beta \mathrm{}}{\pi t}\mathrm{sinh}(\frac{\pi t}{\beta \mathrm{}})\right)\frac{|𝐩|^2}{m^2c^2}\frac{e^2}{c\mathrm{}}\frac{ct}{l_{dB}}\hfill \end{array}$$ (18) for a large $`t`$ (and a large $`|𝐲|`$). Let us consider next small $`t`$ together with a small $`𝐲`$. Then, we can set $`s=\tau =0`$ in the argument of $`G_{th}`$ in eq.(13). In such a case $$\begin{array}{c}S\frac{e^2}{m^2c\mathrm{}\pi }𝐩^2t^2_0^{\mathrm{}}𝑑kk_0^\pi 𝑑\theta \mathrm{sin}\theta \hfill \\ \left(1\mathrm{cos}^2\theta \right)\left(1\mathrm{cos}\left(k\mathrm{cos}\theta |𝐲|\right)\right)\left(\mathrm{exp}\left(\beta \mathrm{}ck\right)1\right)^1\hfill \end{array}$$ (19) The formula (19) is relevant only for a small $`𝐲`$ (for a large $`𝐲`$ we return to eq.(17)). From eq.(19) for a small $`𝐲`$ and small $`t`$ we obtain $$\begin{array}{c}\rho _t(𝐱,𝐱^{})\mathrm{exp}\left(\frac{2e^2}{15\pi m^2\mathrm{}c}𝐩^2t^2𝐲^2_0^{\mathrm{}}𝑑kk^3(\mathrm{exp}(\mathrm{}c\beta k)1)^1\right)\hfill \\ \mathrm{exp}\left(B\frac{e^2}{\mathrm{}c}\left(\frac{𝐩t}{m}\right)^2𝐲^2l_{dB}^4\right)\hfill \end{array}$$ (20) where $`B`$ is a constant of order 1. Again we can conclude that the decoherence is visible on distances comparable to the thermal de Broglie length. For a large $`t`$ we neglect the quickly oscillating $`t`$-dependent terms in eq.(15). Then, similarly as in the computations in eq.(18) (with $`\alpha =cos\theta `$) $$\begin{array}{c}S=\frac{e^2}{\pi m^2c^3\mathrm{}}𝐩^2|𝐲|_0^1𝑑\gamma _1^1𝑑\alpha \alpha (1\alpha ^2)_0^{\mathrm{}}𝑑k\left(\mathrm{exp}\left(\beta \mathrm{}ck\right)1\right)^1\mathrm{sin}\left(k\gamma \alpha |𝐲|\right)\hfill \\ =\frac{2e^2}{\pi m^2c^3\mathrm{}}𝐩^2_0^1𝑑\alpha (1\alpha ^2)\mathrm{ln}\left(\frac{\beta \mathrm{}c}{\pi \alpha |𝐲|}\mathrm{sinh}(\frac{\pi \alpha |𝐲|}{\beta \mathrm{}c})\right)\frac{e^2}{\mathrm{}c}\frac{𝐩^2}{m^2c^2}\frac{|𝐲|}{l_{dB}}\hfill \end{array}$$ (21) for a large $`𝐲`$ (and a large $`t`$). For a small $`|𝐲|`$ (and a large $`t`$ ) the $`t`$-dependent terms in eq.(15) can be neglected. Then, the integral (14) gives $$S=\frac{2}{15}\frac{e^2}{\pi m^2c^3\mathrm{}}|𝐩|^2|𝐲|^2l_{dB}^2_0^{\mathrm{}}𝑑kk(\mathrm{exp}k1)^1$$ Hence, $$|\rho _t|\mathrm{exp}\left(B\frac{e^2}{\mathrm{}c}|𝐩|^2(mc)^2|𝐲|^2l_{dB}^2\right)$$ At this point it is useful to recall the definition of the Wigner function $$𝒲(𝐪,𝐤)=(2\pi \mathrm{})^3𝑑𝐮\mathrm{exp}(i\mathrm{𝐤𝐮}/\mathrm{})\rho (𝐪+𝐮/2,𝐪𝐮/2)$$ If $`\rho \mathrm{exp}(\frac{a}{2}𝐲^2i\mathrm{𝐩𝐲}/\mathrm{})`$ then $$𝒲(𝐪,𝐤)\mathrm{exp}(\frac{1}{2a\mathrm{}^2}(𝐩𝐤)^2)w(𝐪,𝐤)$$ Such a behaviour means a localization on the classical momentum $`𝐩`$. If (as in eq.(18) for a large time) $`\rho _t=\mathrm{exp}(i\mathrm{𝐩𝐲}S)\mathrm{exp}(i\mathrm{𝐩𝐲}b𝐩^2t)`$ then approximately $$_t\rho _tb[𝐏,[𝐏,\rho _t]]$$ (22) where $`𝐏=i\mathrm{}`$ is the quantum momentum operator. Then $$_t𝒲_t(𝐪,𝐤)b𝐤^2𝒲_t(𝐪,𝐤)$$ For N-particles with momenta $`𝐩_j`$ we consider the initial state of the form $$\psi =\mathrm{exp}(\frac{i}{\mathrm{}}\underset{j=1}{\overset{N}{}}𝐩_j𝐱_j)\varphi $$ Then, $`W_t`$ in eq.(12) is a sum of the Hamilton-Jacobi functions for each particle. The subsequent expectation value over the electromagnetic field gives an exponential with a sum of pairings between the $`N`$ particles. Hence, it can be expected that $`\rho _t`$ behaves as $`\mathrm{exp}(RN^2t^2)`$ for a small t and as $`\mathrm{exp}(RN^2|t|)`$ for a large t with a certain constant $`R`$ (note that the behaviour $`\mathrm{exp}(RN|t|)`$ for another model has been obtained by Unruh ). Such a result is valid for N particles of equal charges under the assumption that the terms in the exponential of the form (13) have equal signs. The contributions add if the momenta have a distinguished direction. If the directions are random then the contributions from different particles cancel one another. A similar cancellation takes place if the system is neutral , i.e., if the charges $`e_j`$ can have different signs. Then, in the sum in an exponential of the form (13) we shall have $`e_je_k𝐩_j𝐩_k`$ multiplying $`G_{th}`$. Hence, there can be cancellations from different charges $`e_j`$ even if there is a distinguished direction of the momenta. Under an assumption that the contributions in the exponential do not cancel one another we obtain for a large distance between any pair of coordinates and a large $`t`$ $$|\rho _t|\mathrm{exp}\left(B\frac{e^2}{c\mathrm{}}N^2(\frac{|𝐩|}{mc})^2\frac{ct}{l_{dB}}\right)$$ where B is a constant of order 1. Such a decay rate means that the decay is visible if the distance achieved by a particle (with the mean momentum $`𝐩`$ ) is comparable with the de Broglie wave length. The time needed for this purpose can be short if $`N`$ is large $$\frac{e^2}{c\mathrm{}}N^2>(\frac{mc}{|𝐩|})^2\frac{l_{dB}}{ct}$$ ## 4 Disappearance of the interference So far we have discussed the decay of $`\rho _t(𝐱,𝐱^{})`$ for varying $`t`$ and $`𝐲=𝐱𝐱^{}`$. The disappearance of the off-diagonal matrix elements indicates a classical behaviour of quantum probabilities in the state $`\rho _t`$. We investigate next what happens with the interference describing a typical quantum phenomenon. We consider a superposition of two wave packets $$\psi (𝐱)=\mathrm{exp}(i𝐩^{(1)}𝐱/\mathrm{})\varphi (𝐱)+\mathrm{exp}(i𝐩^{(2)}𝐱/\mathrm{})\varphi (𝐱)$$ Then, at the point $`𝐱`$ (on the screen) after an evolution through a cavity filled with thermal photons the probability density $`|\psi _t(𝐱)|^2`$ is equal to the diagonal part of the density matrix $$\rho _t=|\psi _t\psi _t|$$ (23) For each packet we have the Hamilton-Jacobi function $$W_t(𝐱)=\mathrm{𝐩𝐱}t\frac{𝐩^2}{2m}\frac{𝐩e}{mc}_0^t𝐀(𝐲_s(𝐱),s)𝑑s$$ (24) Under the time evolution $$\psi \psi _t=\mathrm{exp}(iW_t^{(1)}/\mathrm{})\varphi _t^{(1)}+\mathrm{exp}(iW_t^{(2)}/\mathrm{})\varphi _t^{(2)}$$ In our semiclassical approximation $`\varphi (𝐱)\varphi (𝐲_t(𝐱))`$). Then, for weak fields we neglect the dependence of the paths on the electromagnetic field, i.e., we consider straight lines $$𝐲_s^{(1)}=𝐱\frac{s}{m}𝐩^{(1)}$$ and $$𝐲_s^{(2)}=𝐱\frac{s}{m}𝐩^{(2)}$$ In this approximation the expectation value (23) is $$\begin{array}{c}\psi _t(𝐱)^2=|\varphi (𝐱\frac{t}{m}𝐩^{(1)})|^2+|\varphi (𝐱\frac{t}{m}𝐩^{(2)})|^2+\hfill \\ (\overline{\varphi (𝐱\frac{t}{m}𝐩^{(2)})}\varphi (𝐱\frac{t}{m}𝐩^{(1)})\hfill \\ \mathrm{exp}\left(\frac{i}{\mathrm{}}(𝐩^{(2)}𝐩^{(1)})𝐱+\frac{it}{2m\mathrm{}}((𝐩^{(2)})^2(𝐩^{(1)})^2)\right)+\hfill \\ \mathrm{exp}(\frac{i}{\mathrm{}}(𝐩^{(2)}𝐩^{(1)})𝐱\frac{it}{2m\mathrm{}}((𝐩^{(2)})^2(𝐩^{(1)})^2))\overline{\varphi (𝐱\frac{t}{m}𝐩^{(1)})}\varphi (𝐱\frac{t}{m}𝐩^{(2)}))\hfill \\ \mathrm{exp}(\frac{e^2}{2m^2c^2\mathrm{}^2}_0^t𝐩^{(1)}G_{th}(\frac{s}{m}𝐩^{(1)}\frac{\tau }{m}𝐩^{(2)},s\tau )𝐩^{(2)}dsd\tau \hfill \\ +\frac{e^2}{2m^2c^2\mathrm{}^2}_0^t𝐩^{(1)}G_{th}(\frac{s}{m}𝐩^{(2)}\frac{\tau }{m}𝐩^{(1)},s\tau )𝐩^{(2)}𝑑s𝑑\tau \hfill \\ \frac{e^2}{2m^2c^2\mathrm{}^2}_0^t𝐩^{(1)}G_{th}(\frac{s}{m}𝐩^{(1)}\frac{\tau }{m}𝐩^{(1)},s\tau )𝐩^{(1)}𝑑s𝑑\tau \hfill \\ \frac{e^2}{2m^2c^2\mathrm{}^2}_0^t𝐩^{(2)}G_{th}(\frac{s}{m}𝐩^{(2)}\frac{\tau }{m}𝐩^{(2)},s\tau )𝐩^{(2)}dsd\tau )\rho _t^{(1)}+\rho _t^{(2)}+\rho _t^{(12)}\hfill \end{array}$$ (25) Without detailed calculations we can obtain the behaviour for a small t. Let $`s=\tau =0`$ in $`G_{th}`$ in eq.(25) then from eq.(3) $`G_{th}(0,0)B\delta _{jl}\beta ^2\mathrm{}^1c^1`$. It follows that $$\begin{array}{c}|\rho _t^{(12)}|\mathrm{exp}\left(B\frac{e^2}{\mathrm{}c}l_{dB}^2(t|𝐩^{(2)}𝐩^{(1)}|/m)^2\right)\hfill \end{array}$$ (26) Hence, if the decoherence is to be visible the distance between the particles after time $`t`$ must be of the order of the de Broglie length. The calculations for a large time are more involved. Let us denote $$|\rho _t^{(12)}|\mathrm{exp}(S_{12})$$ We obtain for $`S_{12}`$ $$\begin{array}{c}S_{12}=\frac{2}{3}\frac{e^2}{\pi m^2c^3\mathrm{}}|𝐩^{(2)}𝐩^{(1)}|^2_0^{\mathrm{}}\frac{dk}{k}(\mathrm{exp}(\beta \mathrm{}ck)1)^1\left(1\mathrm{cos}\left(tck\right)\right)\hfill \\ =\frac{2}{3}\frac{e^2}{\pi m^2c^3\mathrm{}}|𝐩^{(2)}𝐩^{(1)}|^2ct_0^1𝑑\gamma _0^{\mathrm{}}𝑑k(\mathrm{exp}(\beta \mathrm{}ck)1)^1\mathrm{sin}(\gamma ckt)\hfill \\ =\frac{2}{3}\frac{e^2}{\pi m^2c^2\mathrm{}}|𝐩^{(2)}𝐩^{(1)}|^2_0^1𝑑\gamma (\frac{t\pi }{2\beta \mathrm{}c}\mathrm{coth}(\frac{\pi \gamma t}{\beta \mathrm{}})\frac{1}{2\gamma c})\hfill \\ =\frac{1}{3}\frac{e^2}{\pi m^2c^3\mathrm{}}|𝐩^{(2)}𝐩^{(1)}|^2\mathrm{ln}\left(\frac{\beta \mathrm{}}{\pi t}\mathrm{sinh}(\frac{t\pi }{\beta \mathrm{}})\right)\hfill \\ \frac{e^2}{\mathrm{}c}\frac{ct}{l_{dB}}|𝐩^{(2)}𝐩^{(1)}|^2m^2c^2\hfill \end{array}$$ (27) for a large t. Hence, the mixed term in eq.(25) is multiplied by $`\mathrm{exp}(S_{12})`$ which decays as $`\mathrm{exp}(Rt)`$. Such a behaviour of the probability density proves that the thermal photons lead to the classical addition of probabilities instead of the quantum addition of amplitudes showing the decoherence phenomenon in a physical model of an electron-photon interaction. ## 5 Conclusions We have discussed a model of a quantum charged particle interacting with a quantum electromagnetic field at finite temperature. We have calculated the time evolution of the density matrix in a semiclassical approximation for the wave function and in a weak coupling approximation for the particle-photon interaction. In contradistinction to refs. we do not apply the approximation of a linear coordinate coupling to the environment. For a small time and small space separations we obtain an exponential in time and space decay $`\rho _t\mathrm{exp}(bt^2|𝐱𝐱^{}|^2)`$ of off-diagonal matrix elements (decoherence). For a large time the decay achieves its stationary value (18) and (21). The time and space scale is determined by the thermal de Broglie wave length. If we have a large number $`N`$ of charged particles then the decoherence rate can increase as $`N^2`$ . The density matrix elements decay as $`\rho _t\mathrm{exp}(bt)`$ for a large time. Such a behaviour is in agreement with the Lindblad dynamics if the dissipative part of the dynamics has the form (22). Such Lindblad dynamics has been discussed in refs.. The Lindblad dynamics resulting from a linear coordinate coupling to the environment (studied in ) is of a different type. It could be described by a replacement of the momentum operator by a position operator in eq.(22) (a coupling to the environment linear in the momentum as well as in the coordinate has been discussed by Leggett in ref.). We have studied the interference as another typical aspect of a quantum behaviour. We have shown that in an environment of photons the interference disappears with an exponential speed (eqs.(26)-(27)). Our results suggest an arrangement for an experiment. Such experiments could verify the QED beyond the usual perturbative approximation as well as the principle of the wave function reduction in a non-selective measurement.
warning/0001/hep-ph0001234.html
ar5iv
text
# ON THE QUARK-QUARK SUPERCONDUCTING PAIRING INDUCED BY INSTANTONS IN QCD ## Abstract The superconducting pairing of quarks induced by instantons in QCD is studied by means of the functional integral method. The integral equation determining the critical temperature of the superconducting phase transition is established. It is shown that the Bose condensate of diquarks consists of color antitriplet $`\left(\overline{3}\right)`$ scalar diquarks. The superconducting pairing of quarks in QCD was investigated since more than twenty years<sup>\[1-5\]</sup>. At the present time a strong interest to this problem is motivated by recent works of Alford, Rajagopal and Wilczeck<sup></sup>, Schäfer and Wilczek<sup></sup>, Rapp, Schäfer, Shuryak and Velkovsky<sup></sup> and other authors<sup>\[9-16\]</sup>. In this work the functional integral method<sup></sup> is applied to the study of the superconducting pairing of quarks induced by instantons<sup></sup>, and the integral equation determining the critical temperature of phase transition is established. We work in the imaginary time formalism and use the notations $$x=(𝐱,\tau ),𝑑x=_0^\beta 𝑑\tau 𝑑𝐱,\beta =\frac{1}{kT},$$ $`k`$ and $`T`$ being the Boltzmann constant and the temperature. Denote $`\psi _A\left(x\right)=\psi _A(𝐱,\tau )`$ the quark field, $`A`$ being the set consisting of the index $`\alpha `$ of the Dirac spinor, the color and flavor indices $`a`$ and $`i`$, $`A=\left(\alpha ai\right)`$, $`a=1,2,\mathrm{},N_c`$, $`i=1,2,\mathrm{},N_f`$. Consider the system of massless quarks (and antiquarks) with the parity conserving and chirally symmetric interaction Lagrangian of the form $`L_{int}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\overline{\psi }^A\left(x\right)\overline{\psi }^B\left(x\right)U_{BA}^{CD}\psi _D\left(x\right)\psi _C\left(x\right),`$ (1) $`U_{BA}^{CD}`$ $`=`$ $`U_{AB}^{CD}=U_{BA}^{DC}=U_{AB}^{DC}.`$ (2) The partition function of the system equals $`Z`$ $`=`$ $`{\displaystyle \left[D\psi \right]\left[D\overline{\psi }\right]\mathrm{exp}\left\{𝑑x\overline{\psi }^A\left(x\right)D_A^B\psi _B\left(x\right)\right\}}`$ (3) $`\mathrm{exp}\left\{{\displaystyle \frac{1}{2}}{\displaystyle 𝑑x\overline{\psi }^A\left(x\right)\overline{\psi }^B\left(x\right)U_{BA}^{CD}\psi _D\left(x\right)\psi _C\left(x\right)}\right\}`$ with $$D_A^B=\left[\gamma _4\left(\frac{}{\tau }\mu \right)+\gamma \right]_\alpha ^\beta \delta _a^b\delta _i^j,$$ (4) $`\mu `$ being the chemical potential. Introducing the antisymmetric bi-spinor fields $`\mathrm{\Phi }_{DC}\left(x\right),\overline{\mathrm{\Phi }}^{AB}\left(x\right),`$ $$\mathrm{\Phi }_{CD}\left(x\right)=\mathrm{\Phi }_{DC}\left(x\right),\overline{\mathrm{\Phi }}^{BA}\left(x\right)=\overline{\mathrm{\Phi }}^{AB}\left(x\right),$$ (5) and the functional integral over these fields $$Z_0^\mathrm{\Phi }=\left[D\mathrm{\Phi }\right]\left[D\overline{\mathrm{\Phi }}\right]\mathrm{exp}\left\{\frac{1}{2}\overline{\mathrm{\Phi }}^{AB}\left(x\right)U_{BA}^{CD}\mathrm{\Phi }_{DC}\left(x\right)\right\},$$ (6) we have the Hubbard-Stratonovich transformation<sup></sup> $`\mathrm{exp}\left\{{\displaystyle \frac{1}{2}}{\displaystyle 𝑑x\overline{\psi }^A\left(x\right)\overline{\psi }^B\left(x\right)U_{BA}^{CD}\psi _D\left(x\right)\psi _C\left(x\right)}\right\}`$ $`={\displaystyle \frac{1}{Z_0^\mathrm{\Phi }}}{\displaystyle \left[D\mathrm{\Phi }\right]\left[D\overline{\mathrm{\Phi }}\right]\mathrm{exp}\left\{\frac{1}{2}\overline{\mathrm{\Phi }}^{AB}\left(x\right)U_{BA}^{CD}\mathrm{\Phi }_{DC}\left(x\right)\right\}}`$ (7) $`\mathrm{exp}\left\{{\displaystyle \frac{1}{2}}{\displaystyle 𝑑x\left[\overline{\mathrm{\Delta }}^{CD}\left(x\right)\psi _D\left(x\right)\psi _C\left(x\right)+\overline{\psi }^A\left(x\right)\overline{\psi }^B\left(x\right)\mathrm{\Delta }_{BA}\left(x\right)\right]}\right\},`$ where $$\mathrm{\Delta }_{BA}\left(x\right)=U_{BA}^{CD}\mathrm{\Phi }_{DC}\left(x\right),\overline{\mathrm{\Delta }}^{CD}\left(x\right)=\overline{\mathrm{\Phi }}^{AB}\left(x\right)U_{BA}^{CD}.$$ (8) Applying this transformation in the r.h.s. of the relation (3) and performing the functional integration over fermionic variables, we rewrite the partition function $`Z`$ in the form of a functional integral over the bosonic diquark fields $`\mathrm{\Phi }_{DC}\left(x\right)`$ and $`\overline{\mathrm{\Phi }}^{AB}\left(x\right)`$ : $$Z=\frac{Z_0}{Z_0^\mathrm{\Phi }}\left[D\mathrm{\Phi }\right]\left[D\overline{\mathrm{\Phi }}\right]\mathrm{exp}\left\{I_{eff}[\mathrm{\Phi },\overline{\mathrm{\Phi }}]\right\},$$ (9) $`Z_0`$ being the partition function of the system of the free quarks, $$Z_0=\left[D\psi \right]\left[D\overline{\psi }\right]\mathrm{exp}\left\{𝑑x\overline{\psi }^A\left(x\right)D_A^B\psi _B\left(x\right)\right\}.$$ (10) The effective action of the diquark system equals $`I_{eff}[\mathrm{\Phi },\overline{\mathrm{\Phi }}]`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \overline{\mathrm{\Phi }}^{AB}\left(x\right)U_{BA}^{CD}\mathrm{\Phi }_{DC}\left(x\right)}+W[\mathrm{\Delta },\overline{\mathrm{\Delta }}],`$ (11) $`W[\mathrm{\Delta },\overline{\mathrm{\Delta }}]`$ $`=`$ $`{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}W^{\left(2n\right)}[\mathrm{\Delta },\overline{\mathrm{\Delta }}],`$ (12) $$W^{\left(2\right)}[\mathrm{\Delta },\overline{\mathrm{\Delta }}]=\frac{1}{2}𝑑x_1𝑑x_2\overline{\mathrm{\Delta }}^{D_1C_1}\left(x_1\right)S_{C_1}^{B_2}\left(x_1x_2\right)\mathrm{\Delta }_{B_2A_2}\left(x_2\right)S_{D_1}^{TA_2}\left(x_2x_1\right),$$ (13) $`W^{\left(4\right)}[\mathrm{\Delta },\overline{\mathrm{\Delta }}]`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle 𝑑x_1𝑑x_2𝑑x_3𝑑x_4\overline{\mathrm{\Delta }}^{D_1C_1}\left(x_1\right)S_{C_1}^{B_2}\left(x_1x_2\right)\mathrm{\Delta }_{B_2A_2}\left(x_2\right)}`$ (14) $`S_{D_3}^{TA_2}\left(x_2x_3\right)\overline{\mathrm{\Delta }}^{D_3C_3}\left(x_3\right)S_{C_5}^{B_4}\left(x_3x_4\right)\mathrm{\Delta }_{B_4A_4}\left(x_4\right)S_{D_1}^{TA_4}\left(x_4x_1\right),`$ $`\mathrm{}\mathrm{}..`$ where $`S_A^B\left(xy\right)`$ is the two-point Green function of the free quark, $$D_A^BS_B^C\left(xy\right)=\delta _A^C\left(xy\right),$$ (15) and $$S_B^{TA}\left(xy\right)=S_B^A\left(yx\right).$$ (16) From the variational principle $$\frac{\delta I_{eff}[\mathrm{\Phi },\overline{\mathrm{\Phi }}]}{\delta \overline{\mathrm{\Phi }}^{AB}\left(x\right)}=0$$ (17) we derive the system of field equations $`\mathrm{\Delta }_{A_1B_1}\left(x_1\right)`$ $`=`$ $`U_{A_1B_1}^{D_1C_1}\{{\displaystyle }dx_2S_{C_1}^{A_2}(x_1x_2)\mathrm{\Delta }_{A_2B_2}\left(x_2\right)S_{D_1}^{TB_2}(x_2x_1)`$ $`{\displaystyle 𝑑x_2𝑑x_3𝑑x_4S_{C_1}^{A_2}\left(x_1x_2\right)\mathrm{\Delta }_{A_2B_2}\left(x_2\right)S_{D_3}^{TB_2}\left(x_2x_3\right)}`$ $`\overline{\mathrm{\Delta }}^{D_3C_3}\left(x_3\right)S_{C_3}^{A_4}(x_3x_4)\mathrm{\Delta }_{A_4B_4}\left(x_4\right)S_{D_1}^{TB_4}(x_4x_1)+\mathrm{}\}.`$ The constant solutions $`\mathrm{\Delta }_{AB}\left(x\right)`$ $`=`$ $`\mathrm{\Delta }_{AB}^0=\mathrm{const},`$ $`\overline{\mathrm{\Delta }}^{DC}\left(x\right)`$ $`=`$ $`\overline{\mathrm{\Delta }}^{0DC}=\mathrm{const}`$ of this system of equations are the order parameters of the superconducting phase transition with the formation of the Bose condensate of diquarks - the Cooper pairs of quarks. At the critical temperature $`TT_c`$, $`\beta \beta _c`$, these order parameters tend to vanishing limits and in the r.h.s. of the equations (18) we can neglect the high order terms. Then we obtain the system of linear equations $$\mathrm{\Delta }_{AB}^0=U_{AB}^{DC}𝑑xS_C^A^{}\left(x\right)\mathrm{\Delta }_{A^{}B^{}}^0S_D^{TB^{}}\left(x\right).$$ (19) Using the expression of the two - point Green function of the free quarks at $`\beta \beta _c`$ $$S_A^B\left(x\right)=\frac{1}{\beta _c}\underset{m}{}\mathrm{exp}\left\{i\epsilon _m\tau \right\}\frac{1}{\left(2\pi \right)^3}𝑑𝐩\mathrm{exp}\left\{i\mathrm{𝐩𝐱}\right\}\frac{\left(i\right)\left[\left(\epsilon _m+i\mu \right)\gamma _4+𝐩\gamma \right]_{\alpha \beta }}{\left(\epsilon _m+i\mu \right)^2+𝐩^2},$$ (20) $$\epsilon _m=\left(2m+1\right)\frac{\pi }{\beta _c},$$ $`m`$ being the integers, and introducing the $`4\times 4`$ matrices $`\widehat{\mathrm{\Delta }}_{\left(ai\right)\left(bj\right)}^0`$ with the elements $$\left[\widehat{\mathrm{\Delta }}_{\left(ai\right)\left(bj\right)}^0\right]_{\alpha \beta }=\mathrm{\Delta }_{\left(\alpha ai\right)\left(\beta bj\right)}^0,$$ we rewrite the equations (19) in the matrix notations $`\left[\widehat{\mathrm{\Delta }}_{\left(ai\right)\left(bj\right)}^0\right]_{\alpha \beta }`$ $`=`$ $`U_{\left(\alpha ai\right)\left(\beta bj\right)}^{\left(\delta d\mathrm{}\right)\left(\gamma ck\right)}{\displaystyle \frac{1}{\beta _c}}{\displaystyle \underset{m}{}}{\displaystyle \frac{1}{\left(2\pi \right)^3}}{\displaystyle 𝑑𝐩}`$ (21) $`{\displaystyle \frac{\left[\left\{\left(\epsilon _m+i\mu \right)\gamma _4+𝐩\gamma \right\}\widehat{\mathrm{\Delta }}_{\left(ak\right)\left(d\mathrm{}\right)}^0\left\{\left(\epsilon _mi\mu \right)\gamma _4+𝐩\gamma \right\}^T\right]_{\gamma \delta }}{\left[\left(\epsilon _m+i\mu \right)^2+𝐩^2\right]\left[\left(\epsilon _mi\mu \right)^2+𝐩^2\right]}}.`$ The order parameter matrix with the elements $`\mathrm{\Delta }_{\left(\alpha ai\right)\left(\beta bj\right)}^0`$ and the coupling constant matrix with the elements $`U_{\left(\alpha ai\right)\left(\beta bj\right)}^{\left(\delta d\mathrm{}\right)\left(\gamma ck\right)}`$ have the general forms<sup></sup> $$\mathrm{\Delta }_{\left(\alpha ai\right)\left(\beta bj\right)}^0=\mathrm{\Delta }_{\left(ai\right)\left(bj\right)}^S\left(\gamma _5C\right)_{\alpha \beta }+\mathrm{\Delta }_{\left(ai\right)\left(bj\right)}^P\left(C\right)_{\alpha \beta },$$ (22) $`U_{\left(\alpha ai\right)\left(\beta bj\right)}^{\left(\delta d\mathrm{}\right)\left(\gamma ck\right)}`$ $`=`$ $`{\displaystyle \frac{1}{4}}S_{\left(ai\right)\left(bj\right)}^{\left(d\mathrm{}\right)\left(ck\right)}\left(\gamma _5C\right)_{\alpha \beta }\left(C^1\gamma _5\right)^{\delta \gamma }+{\displaystyle \frac{1}{4}}P_{\left(ai\right)\left(bj\right)}^{\left(d\mathrm{}\right)\left(ck\right)}\left(C\right)_{\alpha \beta }\left(C^1\right)^{\delta \gamma }`$ (23) $`+{\displaystyle \frac{1}{16}}V_{\left(ai\right)\left(bj\right)}^{\left(d\mathrm{}\right)\left(ck\right)}\left(\gamma _\mu \gamma _5C\right)_{\alpha \beta }\left(C^1\gamma _5\gamma _\mu \right)^{\delta \gamma }+{\displaystyle \frac{1}{16}}A_{\left(ai\right)\left(bj\right)}^{\left(d\mathrm{}\right)\left(ck\right)}\left(\gamma _\mu C\right)_{\alpha \beta }\left(C^1\gamma _\mu \right)^{\delta \gamma }`$ $`+{\displaystyle \frac{1}{24}}T_{\left(ai\right)\left(bj\right)}^{\left(d\mathrm{}\right)\left(ck\right)}\left(\sigma _{\mu \nu }\gamma _5C\right)_{\alpha \beta }\left(C^1\gamma _5\sigma _{\mu \nu }\right)^{\delta \gamma }.`$ Substituting these expressions into the r.h.s of the equations (21), we derive two equations for $`\mathrm{\Delta }_{\left(ai\right)\left(bj\right)}^S`$ and $`\mathrm{\Delta }_{\left(ai\right)\left(bj\right)}^P`$ : $$\mathrm{\Delta }_{\left(ai\right)\left(bj\right)}^S=JS_{\left(ai\right)\left(bj\right)}^{\left(d\mathrm{}\right)\left(ck\right)}\mathrm{\Delta }_{\left(ck\right)\left(d\mathrm{}\right)}^S,$$ $`(24.a)`$ $$\mathrm{\Delta }_{\left(ai\right)\left(bj\right)}^P=JP_{\left(ai\right)\left(bj\right)}^{\left(d\mathrm{}\right)\left(ck\right)}\mathrm{\Delta }_{\left(ck\right)\left(d\mathrm{}\right)}^P,$$ $`(24.b)`$ where $$J=\frac{1}{\beta _c}\underset{m}{}\frac{1}{\left(2\pi \right)^3}𝑑𝐩\frac{\epsilon _m^2+𝐩^2+\mu ^2}{\left[\left(\epsilon _m+i\mu \right)^2+𝐩^2\right]\left[\left(\epsilon _mi\mu \right)^2+𝐩^2\right]}.$$ (25) It can be shown that $$J=\frac{1}{8\pi ^2}_0^\lambda \left[\frac{1}{p\mu }\mathrm{tanh}\frac{\beta _c\left(p\mu \right)}{2}+\frac{1}{p+\mu }\mathrm{tanh}\frac{\beta _c\left(p+\mu \right)}{2}\right]p^2𝑑p.$$ (26) The ultraviolet divergence in the integral (25) may be avoided by introducing the momentum cut-off $`\lambda `$. In the case of the effective four-quark direct coupling induced by instantons<sup></sup> we have $`U_{\left(\beta bj\right)\left(\alpha ai\right)}^{\left(\gamma ck\right)\left(\delta d\mathrm{}\right)}={\displaystyle \frac{1}{2}}\delta _a^c\delta _b^d\left[\delta _i^k\delta _j^{\mathrm{}}\left(\tau \right)_i^k\left(\tau \right)_j^{\mathrm{}}\right]\left\{G_1\left[\delta _\alpha ^\gamma \delta _\beta ^\delta +\left(\gamma _5\right)_\alpha ^\gamma \left(\gamma _5\right)_\beta ^\delta \right]+G_2\left(\sigma _{\mu \nu }\right)_\alpha ^\gamma \left(\sigma _{\mu \nu }\right)_\beta ^\delta \right\}`$ $`{\displaystyle \frac{1}{2}}\delta _a^d\delta _b^c\left[\delta _i^{\mathrm{}}\delta _j^k\left(\tau \right)_i^{\mathrm{}}\left(\tau \right)_j^k\right]\left\{G_1\left[\delta _\alpha ^\delta \delta _\beta ^\gamma +\left(\gamma _5\right)_\alpha ^\delta \left(\gamma _5\right)_\beta ^\gamma \right]+G_2\left(\sigma _{\mu \nu }\right)_\alpha ^\delta \left(\sigma _{\mu \nu }\right)_\beta ^\gamma \right\}`$ (27) with $$G_1=g\frac{1}{N_c^21}\frac{2N_c1}{2N_c},G_2=g\frac{1}{N_c^21}\frac{1}{4N_c},$$ (28) where $`g`$ is some positive constant. It follows that $$S_{\left(ai\right)\left(bj\right)}^{\left(d\mathrm{}\right)\left(ck\right)}=P_{\left(ai\right)\left(bj\right)}^{\left(d\mathrm{}\right)\left(ck\right)}=\frac{1}{4}G\epsilon _{ij}\epsilon ^k\mathrm{}\left(\delta _a^c\delta _b^d\delta _a^d\delta _b^c\right)$$ (29) with $$G=g.\frac{1}{2N_c\left(N_c1\right)}.$$ (30) Substituting the expression (29) of $`S_{\left(ai\right)\left(bj\right)}^{\left(d\mathrm{}\right)\left(ck\right)}`$ and $`P_{\left(ai\right)\left(bj\right)}^{\left(d\mathrm{}\right)\left(ck\right)}`$ into the r.h.s. of the equations (24.a) and (24.b), we derive the formulae $`\mathrm{\Delta }_{\left(ai\right)\left(bj\right)}^S`$ $`=`$ $`\epsilon _{ij}\mathrm{\Delta }_{\left[ab\right]}^S,`$ $`\mathrm{\Delta }_{\left(ai\right)\left(bj\right)}^P`$ $`=`$ $`\epsilon _{ij}\mathrm{\Delta }_{\left[ab\right]}^P,`$ (31) where $`\mathrm{\Delta }_{\left[ab\right]}^S`$ and $`\mathrm{\Delta }_{\left[ab\right]}^P`$ are antisymmetric under the permutation of the color indices and satisfy the equations $$\mathrm{\Delta }_{\left[ab\right]}^S\left(1GJ\right)=0,$$ $`(32.a)`$ $$\mathrm{\Delta }_{\left[ab\right]}^P\left(1+GJ\right)=0.$$ $`(32.b)`$ Because both constants $`G`$ and $`J`$ are positive, $`\mathrm{\Delta }_{\left[ab\right]}^P`$ must be zero $$\mathrm{\Delta }_{\left[ab\right]}^P=0.$$ (33) For the existence of a non-vanishing antisymmetric rank 2 spinor $`\mathrm{\Delta }_{\left[ab\right]}^S`$ the equation $$GJ=1$$ (34) must be satisfied. In another form this equation was derived by Pisarski and Rischke<sup></sup>. In some approximation it was discussed in ref<sup></sup>. Together with the expression (26) for $`J`$ the equation (34) determines the critical temperature $`T_c`$: $$T_c\mathrm{const}\lambda \mathrm{exp}\left\{\frac{8\pi ^2}{\mu ^2G}\right\}.$$ (35) The existence of $`\mathrm{\Delta }_{\left[ab\right]}^S0`$ is a consequence of the attractive interaction between two quarks inside a scalar antitriplet $`\left(\overline{3}\right)`$ diquark. The relation (33) means that the pseudoscalar diquark does not exist.ACKNOWLEDGMENT The authors express their sincere thank to the Natural Science Council for the financial support to this work.
warning/0001/math0001009.html
ar5iv
text
# Solutions to congruences using sets with the property of Baire ## 1. Introduction and definitions The basic form of the Banach-Tarski paradox can be stated as follows: The two-dimensional sphere $`S^2`$ can be partitioned into finitely many pieces $`A_1,A_2,\mathrm{},A_n,B_1,B_2,\mathrm{},B_m`$ with the property that the sets $`A_i`$ can be rearranged by rigid motions (rotations) so as to cover the entire sphere, and so can the sets $`B_j`$. This contradicts standard intuitions concerning measure or area; the sets $`A_i`$ and $`B_j`$ cannot all be measurable with respect to the standard rotation-invariant probability measure on $`S^2`$. This result of Banach and Tarski was based on earlier work of Hausdorff \[7, p.~469\] who proved that the free product of cyclic groups $`_2`$ and $`_3`$ can be embedded in the rotation group $`SO_3`$ of $`S^2`$. Using this, Hausdorff showed that there is a countable set $`D`$ such that $`S^2D`$ can be partitioned into three sets $`A,B,C`$ such that $`A`$ is congruent to $`B`$ (i.e., there is a rotation $`\rho `$ such that $`\rho (A)=B`$), $`B`$ is congruent to $`C`$, and $`C`$ is congruent to $`AB`$. This also is counterintuitive, and the sets $`A`$$`B`$, and $`C`$ cannot be measurable with respect to the standard isometry-invariant probability measure on $`S^2`$. (In fact, there is no rotation-invariant finitely additive probability measure on $`S^2`$ which assigns a measure to these sets.) Later, R. Robinson refined the Banach-Tarski construction, and characterized the systems of congruences for which one can partition $`S^2`$ (without a countable exceptional set) into pieces satisfying the congruences. In order to state Robinson’s results precisely, we need some definitions. Fix a positive integer $`r`$. A congruence is specified by two subsets $`L`$ and $`R`$ of $`\{1,2,\mathrm{},r\}`$, and is written formally as $`_{kL}A_k_{kR}A_k`$, where $`A_1,A_2,\mathrm{},A_r`$ are variables. The congruence is proper if both $`L`$ and $`R`$ are nonempty proper subsets of $`\{1,\mathrm{},r\}`$. Now suppose $`G`$ is a group acting on a set $`X`$, and a system of congruences is given by pairs $`L_i,R_i\{1,\mathrm{},r\}`$ for $`im`$; a solution to the system of congruences in $`X`$ is a sequence of sets $`A_kX`$ ($`kr`$) which are pairwise disjoint and have union $`X`$, such that, for each $`im`$, there is $`\sigma _iG`$ such that $`\sigma _i(_{kL_i}A_k)=_{kR_i}A_k`$ (i.e., $`\sigma _i`$ witnesses congruence number $`i`$). It is clear that only proper congruences are useful here, and that, if $`\sigma _i`$ witnesses the congruence given by $`L_i`$ and $`R_i`$, then $`\sigma _i`$ also witnesses the complementary congruence given by $`L_i^c`$ and $`R_i^c`$, where $`S^c=\{1,\mathrm{},r\}S`$. Also, congruence is transitive: if $`\sigma `$ witnesses $`AB`$ and $`\tau `$ witnesses $`BC`$, then $`\tau \sigma `$ witnesses $`AC`$. A system of congruences is called weak if, among all congruences which can be deduced from the system by taking complements and applying transitivity, there is no congruence of the form $`_{kL}A_k_{kL^c}A_k`$ which requires some set to be congruent to its complement. It is easy to see that, if a system of congruences has a solution in $`S^2`$ with rotations witnessing the congruences, then the system must be weak: any rotation has fixed points, and hence cannot witness that a set is congruent to its complement. Robinson showed that the converse is true: any weak system of congruences has a solution in $`S^2`$ with rotations witnessing the congruences. Dekker gave the following abstract form of Robinson’s result: If $`G`$ is a free group on more than one generator which acts locally commutatively on a set $`X`$, then any weak system of congruences has a solution in $`X`$ with elements of $`G`$ witnessing the congruences. (An action of a group $`G`$ is called locally commutative if any two elements of $`G`$ with a common fixed point commute; clearly the rotation group on $`S^2`$ has this property. Variants or corollaries of Hausdorff’s embedding of a free product of $`_2`$ and $`_3`$ into $`SO_3`$ show that free groups on any finite or countable number of generators can be embedded into $`SO_3`$.) Also, Adams showed that, if one allows arbitrary isometries of $`S^2`$ rather than just rotations to witness the congruences, then any system of proper congruences has a solution in $`S^2`$. The preceding information is from Wagon , which is an excellent reference on the Banach-Tarski paradox and related work. In general, a weak system of congruences need not have a solution using measurable subsets of $`S^2`$. For example, consider the system $`A_2A_2A_3A_4`$, $`A_4A_1A_2A_4`$, which Robinson used to get a minimal Banach-Tarski decomposition of $`S^2`$. If these congruences are satisfied by measurable subsets of $`S^2`$, then, since $`S^2`$ has finite measure, the first congruence forces $`A_3`$ and $`A_4`$ to have measure $`0`$, and the second forces $`A_1`$ and $`A_2`$ to have measure $`0`$, so the four sets together cannot cover $`S^2`$. (The same argument applies to Hausdorff’s system of congruences, but this is not weak.) Another example is the system $`A_1A_2A_3A_4A_5`$, $`A_1A_2A_1A_3A_4`$; any measurable solution to the first part of this would have to give measure $`1/5`$ to each of the sets, making the final congruence impossible. If one considers solutions using sets with the property of Baire instead of measurable sets, then the situation is quite different; the arguments of the preceding paragraph do not apply. It was shown in Dougherty and Foreman that the Banach-Tarski paradox can be carried out using pieces with the property of Baire. In the present paper, the methods of Dougherty and Foreman will be used to characterize those systems of congruences which have solutions in $`S^2`$ under rotations using sets with the property of Baire. The result applies more generally, to show that suitable systems of congruences have solutions using sets with the property of Baire in any Polish space (complete separable metric space) $`𝒳`$ with a nonabelian free group of homeomorphisms of $`𝒳`$ which acts locally commutatively on $`𝒳`$ and freely (without fixed points) on a comeager subset of $`𝒳`$. To specify which systems are ‘suitable’ requires further definitions. We will call a system of congruences nonredundant if no congruence in the system can be deduced from the other congruences in the system by complementation and transitivity as above, and there is no identity congruence $`AA`$ in the system. Next, say that $`A`$ is subcongruent to $`B`$ ($`AB`$) if $`A`$ is congruent to a subset of $`B`$. From a given system of congruences, one can deduce subcongruences by the following rules: if $`AB`$, then $`AB`$; if $`AB`$ and $`BC`$, then $`AC`$; and, if $`AB`$ is in the given system, then $`AB`$, $`BA`$, $`A^cB^c`$, and $`B^cA^c`$ (where $`A^c`$ is the complement of $`A`$). We will call the system of congruences consistent if there do not exist sets $`L,R\{1,2,\mathrm{},r\}`$ with $`R`$ a proper subset of $`L`$ such that one can deduce $`_{kL}A_k_{kR}A_k`$ from the system. For example, the systems used by Hausdorff and Robinson as above are not consistent, but the other example system above is consistent (the only subcongruences deducible from it where the left side is a union of more sets than the right side are $`A_1A_3A_4A_1A_2`$ and $`A_3A_4A_5A_2A_5`$). The main result of this paper is that, if a given system of $`m`$ congruences is weak and consistent, and if $`𝒳`$ is a Polish space on which a free group $`G`$ of homeomorphisms with $`m`$ generators acts locally commutatively everywhere and freely on a comeager set, then the system of congruences has a solution on $`𝒳`$ using sets with the property of Baire; furthermore, if the system is nonredundant, then one can use a specified list of $`m`$ free generators of $`G`$ to serve as the witnesses for the congruences. (This latter condition holds for the Robinson-Dekker construction, without any extra assumption.) The conditions of weakness and consistency are necessary for the case of $`S^2`$ with a free group of rotations, at least if we require the sets to be nonmeager; without this requirement, a system has a solution in $`S^2`$ using free rotations if and only if it has a subsystem (obtained by deleting zero or more of the sets $`A_1,\mathrm{},A_r`$ from all congruences) which is weak and consistent. Also, requiring that a redundant congruence be witnessed by a free rotation can make a system unsolvable on $`S^2`$. As in Dougherty and Foreman , the results here concerning sets with the property of Baire are obtained by combining known results about arbitrary sets with new results about open sets. In most cases, one cannot expect to get actual solutions to systems of congruences using open sets; in particular, a connected space cannot be nontrivially partitioned into open sets at all. We will therefore allow meager exceptional sets when trying to satisfy congruences using open sets. This leads to the following definitions: Suppose $`G`$ is a group of homeomorphisms of a space $`𝒳`$. Two sets $`A,B𝒳`$ will be called quasi-disjoint if their intersection is meager. (Of course, quasi-disjoint open sets in a Polish space are actually disjoint.) Sets $`A`$ and $`B`$ are quasi-congruent, as witnessed by $`\sigma G`$, if $`\sigma (A)`$ differs from $`B`$ by a meager set. A quasi-solution to a system of congruences $`_{kL_i}A_k_{kR_i}A_k`$ is a sequence of sets $`A_k𝒳`$ ($`kr`$) which are pairwise quasi-disjoint and whose union is a comeager subset of $`𝒳`$, such that, for each $`im`$, there is $`\sigma _iG`$ which witnesses that $`_{kL_i}A_k`$ is quasi-congruent to $`_{kR_i}A_k`$. The remainder of this paper is as follows. In section 2, it will be shown that, if $`G`$ is a suitable free group of homeomorphisms of a Polish space $`𝒳`$, then any weak consistent system of congruences has a quasi-solution in $`𝒳`$ using nonempty open sets (with specified free generators of $`G`$ witnessing the congruences, if the system is nonredundant). This result is then combined with the results of Robinson and Dekker to produce solutions (not just quasi-solutions) to any weak consistent system using sets with the property of Baire. Section 3 shows the necessity of weakness, consistency, and nonredundancy. Section 4 gives the proof that, if one allows arbitrary isometries of $`S^2`$ as witnesses, then any consistent system of congruences has a quasi-solution using nonempty open sets, and a solution using nonmeager sets with the property of Baire. In a later paper , we consider the problem of finding open sets which actually satisfy congruences rather than quasi-congruences (but still are only required to cover a dense subset of the space, rather than all of it). We will use the symbol $``$ or simple juxtaposition to denote a group operation, interchangeably. All group actions will be written on the left. For standard basic facts about free groups, such as the unique expression of any element as a reduced word in the generators and the fact that any nonidentity element has infinite order, see any text on combinatorial group theory, such as Magnus, Karrass, and Solitar . More advanced facts will be referred to specifically as needed. ## 2. Positive results ###### Theorem 2.1 Suppose $`𝒳`$ is a Polish space and $`G`$ is a countable group of homeomorphisms of $`𝒳`$ which acts freely on a comeager subset of $`𝒳`$, and which has a subgroup which is free on $`m`$ generators ($`m1`$). Suppose that a system of $`m`$ congruences is specified by pairs $`(L_i,R_i)`$ ($`1im`$) of subsets of $`\{1,2,\mathrm{},r\}`$; also suppose that this system is weak and consistent. Then there is a sequence of nonempty open sets $`A_k𝒳`$ ($`kr`$) which is a quasi-solution to the system. Furthermore, if the system is nonredundant, and elements $`f_i`$ ($`1im`$) of $`G`$ are free generators for a free subgroup of $`G`$, then there is a sequence of nonempty open sets $`A_k`$ as above such that, for each $`im`$, $`f_i`$ witnesses that $`_{kL_i}A_k`$ is quasi-congruent to $`_{kR_i}A_k`$. ###### Demonstration Proof First note that, if one congruence in a system is deducible from the other congruences, then one can delete that one congruence to get a smaller system, and any quasi-solution to the smaller system will be a quasi-solution to the original system. (The same holds for solutions.) By iterating this, one can reduce the original system to a nonredundant system with the same quasi-solutions. Therefore, it will suffice to prove only the second part of the theorem. We may assume that $`G`$ is the free group generated by the elements $`f_i`$. We will follow the method of Dougherty and Foreman , but with a few differences. One difference is that we will concentrate on the points to be excluded from the sets $`A_k`$, and not construct the sets $`A_k`$ themselves until the excluded sets are complete. (The reason for this is that it is easier to work with congruences between intersections than with congruences between unions; we can actually make intersections congruent, rather than quasi-congruent.) We will construct sets $`B_k`$ ($`1kr`$) with the following properties: $`_{k=1}^rB_k=\mathrm{}`$; the sets $`_{k^{}k}B_k^{}`$ for $`kr`$ are all nonempty, and their union is dense in $`𝒳`$; and, for each $`im`$, $`f_i(_{kL_i}B_k)=_{kR_i}B_k`$. Once we have these sets, we can define $`A_k`$ to be $`_{k^{}k}B_k^{}`$; then the sets $`A_k`$ will be as desired. (The intersection of any two sets $`A_k`$ will be $`_{k^{}=1}^rB_k^{}=\mathrm{}`$. For any $`L\{1,2,\mathrm{},r\}`$, the set $`_{kL}B_k`$ includes $`A_k^{}`$ for $`k^{}L`$ and is disjoint from $`A_k^{}`$ for $`k^{}L`$, so $`_{kL}A_k`$ differs from the complement of $`_{kL}B_k`$ by a meager set; hence, a congruence between $`_{kL}B_k`$ and $`_{kR}B_k`$ yields a quasi-congruence between $`_{kL}A_k`$ and $`_{kR}A_k`$.) The open sets $`B_k`$ will be built in stages: we will construct open sets $`B_k^0B_k^1B_k^2\mathrm{}`$ for $`kr`$ and then let $`B_k=_{n=0}^{\mathrm{}}B_k^n`$. The sets $`B_k^n`$ will satisfy the following properties, to be maintained as induction hypotheses: (There is no property rm (1); this numbering is used for compatibility with Dougherty and Foreman .) Of course, we must define the terms used in rm (4): Definition. Two points $`x`$ and $`x^{}`$ are linked, or there is a link from $`x`$ to $`x^{}`$, if $`x^{}=f_i(x)`$ or $`x=f_i(x^{})`$ for some $`im`$. Points $`x`$ and $`x^{}`$ are connected by a chain of links if there are points $`x_0,x_1,\mathrm{},x_J`$ with $`x_0=x`$ and $`x_J=x^{}`$ such that there is a link from $`x_{j1}`$ to $`x_j`$ for each $`jJ`$. A link from $`x`$ to $`x^{}`$ is active (for the sets $`B_k^n`$) if there is a point in one or more of the sets $`B_k^n`$ which is connected to $`x`$ or to $`x^{}`$ by a chain of at most $`2^r`$ links. Note that adding one new point to a set $`B_k^{n+1}`$ activates only a finite number of new links, although the finite number is very large. Let $`B_k^0=\mathrm{}`$ for all $`k`$; clearly this makes rm (2)–rm (4) true for $`n=0`$. Let $`Z_n:n=0,1,2,\mathrm{}`$ be a listing of the nonempty sets in some base for the topology of $`𝒳`$; we may assume that $`𝒳`$ itself occurs at least $`r`$ times in the list. We must show how to get from $`B_k^n`$ to $`B_k^{n+1}`$, preserving properties rm (2)–rm (4), so that, for a given nonempty open set $`Z=Z_n`$, one of the sets $`_{k^{}\overline{k}}B_k^{}^{n+1}`$ for $`\overline{k}r`$ will meet $`Z`$. The $`t`$’th time that $`Z=𝒳`$ ($`tr`$), we will set $`\overline{k}=t`$ in order to ensure that $`_{k^{}t}B_k^{}`$ will be nonempty. Once this is accomplished for all $`n`$, the resulting sets $`B_k`$ will have the desired properties. So suppose we are given $`B_k^n`$ ($`kr`$) and $`Z=Z_n`$. The first step is to find a point $`x_0Z`$ to be put into all but one of the sets $`B_k^{n+1}`$. In fact, we will find $`x_0`$ in $`Z^{}`$, where $`Z^{}`$ is $`Z`$ unless $`Z=𝒳`$ and this is one of the first $`r`$ occurrences of $`𝒳`$ in the list of open sets; in the latter case, if this is the $`t`$’th occurrence of $`𝒳`$, then let $`Z^{}`$ be the interior of the complement of $`B_t^n`$. (To see that this set is nonempty, look at a $`G`$-orbit on which $`G`$ acts freely and which does not meet the boundary of $`B_t^n`$; such orbits form a comeager subset of $`𝒳`$. By rm (4), some point in this orbit is not in any of the sets $`B_k^n`$, and hence must be in the interior of the complement of $`B_t^n`$.) So $`x_0`$ will be in $`Z`$ in any case. Let $`D`$ be the complement of a ($`G`$-invariant) comeager set on which $`G`$ acts freely, and let $`D^{}`$ be the union of the images under the elements of $`G`$ of the boundaries of the sets $`B_k^n`$; then $`DD^{}`$ is meager. Let $`x_0`$ be any point in $`Z^{}(DD^{})`$. By rm (2), we can find $`\overline{k}r`$ such that $`x_0B_{\overline{k}}^n`$. In the case where $`Z=𝒳`$ for the $`t`$’th time ($`tr`$), we have $`x_0B_t^n`$ by the definition of $`Z^{}`$, so we may set $`\overline{k}=t`$. We will ensure that $`x_0B_k^{n+1}`$ for all $`k\overline{k}`$; this will take care of the current density requirement (or the current nonemptiness requirement). As in Dougherty and Foreman , we will construct sets $`\widehat{B}_k`$ by adding finitely many points to the sets $`B_k^n`$; $`\widehat{B}_k`$ will be defined to be $`B_k^n\{g(x_0):gT_k\}`$ for some $`T_kG`$. However, we will describe the construction a little differently; instead of giving inductive clauses to define the sets $`T_k`$, we will define a set $`M_g\{1,2,\mathrm{},r\}`$ for each $`gG`$ and then let $`T_k=\{gG:kM_g\}`$. We define $`M_g`$ recursively, based on the reduced form of $`gG`$ in terms of the generators $`f_i`$. If $`g`$ is the identity of $`G`$, then let $`M_g=\{kr:k\overline{k}\}`$. Otherwise, we can write $`g`$ uniquely as $`f_ig^{}`$ or $`f_i^1g^{}`$ where $`g^{}`$ has a shorter reduced form than $`g`$ does, and hence $`M_g^{}`$ is already defined. Let $`M_g^{}^+=M_g^{}\{k:g^{}(x_0)B_k^n\}`$. If $`M_g^{}=\mathrm{}`$, let $`M_g=\mathrm{}`$. If $`M_g^{}\mathrm{}`$ and $`g=f_ig^{}`$, then define $`M_g`$ as follows: if $`L_iM_g^{}^+`$, let $`M_g=R_i`$; if $`L_i^cM_g^{}^+`$, let $`M_g=R_i^c`$; otherwise, let $`M_g=\mathrm{}`$. (If both $`L_i`$ and $`L_i^c`$ are subsets of $`M_g^{}^+`$, make some arbitrary definition such as $`M_g=\{1,2,\mathrm{},r\}`$; we will see in the next paragraph that this case cannot occur.) If $`M_g^{}\mathrm{}`$ and $`g=f_i^1g^{}`$, then define $`M_g`$ in the same way, but with $`L_i`$ and $`R_i`$ interchanged. First, we show by induction on $`gG`$ that $`M_g^+\{1,2,\mathrm{},r\}`$. If $`g`$ is the identity, then $`\overline{k}M_g^+`$ by the definition of $`x_0`$. Otherwise, we have $`g=f_ig^{}`$ or $`g=f_i^1g^{}`$ for some simpler $`g^{}`$. If $`M_g=\mathrm{}`$, then $`M_g^+=\{k:g(x_0)B_k^n\}\{1,2,\mathrm{},r\}`$ by rm (2). If $`M_g\mathrm{}`$, $`g=f_ig^{}`$, and $`L_iM_g^{}^+`$, then $`L_i^cM_g^{}^+`$ by the induction hypothesis, so $`L_i^c\{k:g^{}(x_0)B_k^n\}`$, so $`R_i^c\{k:g(x_0)B_k^n\}`$ by rm (3), so $`M_g^+=R_i\{k:g(x_0)B_k^n\}\{1,2,\mathrm{},r\}`$. The remaining cases are handled the same way. We can now check that, for any $`g`$ and $`g^{}`$ in $`G`$ and $`ir`$, if $`g=f_ig^{}`$, then $`L_iM_g^{}^+`$ iff $`R_iM_g^+`$, and $`L_i^cM_g^{}^+`$ iff $`R_i^cM_g^+`$. First, suppose the reduced form of $`g^{}`$ does not have $`f_i^1`$ as its leftmost term; then $`M_g`$ is defined from $`M_g^{}`$ as above. If $`M_g^{}=\mathrm{}`$ and hence $`M_g=\mathrm{}`$, then these two equivalences follow directly from rm (3); so suppose $`M_g^{}\mathrm{}`$. Now the two left-to-right implications are immediate. For the first right-to-left implication, if $`L_iM_g^{}^+`$, then $`R_iM_g=\mathrm{}`$ by definition of $`M_g`$, while $`R_i\{k:g(x_0)B_k^n\}`$ because otherwise rm (3) would give $`L_i\{k:g^{}(x_0)B_k^n\}M_g^{}^+`$, so $`R_iM_g^+`$. The other implication is proved in the same way. This completes the case where $`g^{}`$ does not have $`f_i^1`$ as its leftmost term. If $`g^{}`$ does have $`f_i^1`$ as its leftmost term, then $`g`$ does not have $`f_i`$ as its leftmost term, so we can write $`g^{}=f_i^1g`$ and proceed as above. We are now ready to prove rm (2)–rm (4) for the sets $`\widehat{B}_k`$. The definitions of $`T_k`$ and $`\widehat{B}_k`$ (and the fact that $`G`$ acts freely on the orbit of $`x_0`$) easily imply that $`\{k:g(x_0)\widehat{B}_k\}=M_g^+`$ for all $`gG`$, while $`\{k:x\widehat{B}_k\}=\{k:xB_k^n\}`$ if $`x`$ is not in the $`G`$-orbit of $`x_0`$. Therefore, properties rm (2) and rm (3) for $`\widehat{B}_k`$ follow from the same properties for $`B_k^n`$ and the above facts about $`M_g^+`$. To prove property rm (4) for the sets $`\widehat{B}_k`$, we will need the following claims, which are the part of this proof where all of the restrictions on the system of congruences are needed. Define a labeled directed graph $`𝒢`$ from the system of congruences as follows. The vertices of $`𝒢`$ are the nonempty proper subsets of $`\{1,2,\mathrm{},r\}`$. If $`S`$ is such a subset, and $`L_iS`$, then $`𝒢`$ has an edge from $`S`$ to $`R_i`$ labeled $`f_i`$. If $`L_i^cS`$, then $`𝒢`$ has an edge from $`S`$ to $`R_i^c`$ labeled $`f_i`$. Similarly, if $`R_iS`$, then $`𝒢`$ has an edge from $`S`$ to $`L_i`$ labeled $`f_i^1`$; if $`R_i^cS`$, then $`𝒢`$ has an edge from $`S`$ to $`L_i^c`$ labeled $`f_i^1`$. The digraph $`𝒢`$ has cycles of length $`2`$ connecting pairs $`(L_i,R_i)`$ or $`(L_i^c,R_i^c)`$; each such cycle consists of an $`f_i`$-edge and an $`f_i^1`$-edge. Call the edges in these $`2`$-cycles (the edges which come from actual congruences rather than subcongruences) good edges, and call all other edges (e.g., an $`f_i`$-edge from a proper superset of $`L_i`$ to $`R_i`$) bad edges. Claim 1. No cycle in $`𝒢`$ contains a bad edge. Proof. Suppose the edges $`e_1,e_2,\mathrm{},e_J`$ form a cycle. Let $`N_0,N_1,\mathrm{},N_J`$ be the vertices of this cycle (with $`N_J=N_0`$), so that $`e_j`$ is an edge from $`N_{j1}`$ to $`N_j`$. For each $`j`$, $`e_j`$ has a label, which is either $`f_{i_j}`$ or $`f_{i_j}^1`$ for some $`i_j`$. We will abuse notation by writing $`MN`$ for $`M,N\{1,2,\mathrm{},r\}`$ to mean that the subcongruence $`_{kM}A_k_{kN}A_k`$ is deducible from the given system of congruences by the usual rules. For each $`j`$ such that $`0<jJ`$, define a set $`N_j^{}`$ as follows. First, suppose $`e_j`$ is labeled $`f_{i_j}`$. Then either $`L_{i_j}N_{j1}`$ and $`N_j=R_{i_j}`$, or $`L_{i_j}^cN_{j1}`$ and $`N_j=R_{i_j}^c`$; let $`N_j^{}`$ be $`L_{i_j}`$ in the former case and $`L_{i_j}^c`$ in the latter. Similarly, if $`e_j`$ is labeled $`f_{i_j}^1`$, let $`N_j^{}`$ be $`R_{i_j}`$ or $`R_{i_j}^c`$, depending on whether $`N_j`$ is $`L_{i_j}`$ or $`L_{i_j}^c`$. In any case, we have $`N_{j1}N_j^{}`$, and congruence number $`i_j`$ relates either the sets $`N_j^{}`$ and $`N_j`$ or their complements. Therefore, $`N_{j1}N_j^{}N_j`$ for all $`j`$. Since $``$ is transitive and $`N_J=N_0`$, we have $`N_j^{}N_{j1}`$ for $`0<jJ`$. Since the system of congruences is consistent, $`N_j^{}`$ cannot be a proper subset of $`N_{j1}`$, so we must have $`N_j^{}=N_{j1}`$ for $`0<jJ`$; this means that all of the edges are good. $`\mathrm{}`$ Now construct the undirected graph $`𝒢_0`$ by treating each pair of oppositely-directed good edges in $`𝒢`$ as a single undirected edge between $`L_i`$ and $`R_i`$ or between $`L_i^c`$ and $`R_i^c`$. Claim 2. The undirected graph $`𝒢_0`$ is acyclic (i.e., its connected components are trees). Proof. Note that sets $`N,N^{}`$ are in the same component of $`𝒢_0`$ if and only if the congruence $`\{A_k:kN\}\{A_k:kN^{}\}`$ follows from the given system of congruences. In particular, $`N`$ and $`N^c`$ cannot be in the same component of $`𝒢_0`$, because the system is weak. Also, note that congruence number $`i`$ gives rise to two edges of $`𝒢_0`$, one between $`L_i`$ and $`R_i`$ and one between $`L_i^c`$ and $`R_i^c`$; these edges must be in different components of $`𝒢_0`$. Now suppose we have a nontrivial cycle in $`𝒢_0`$; by taking a minimal such cycle, we may ensure that there are no repeated edges in the cycle. Let one of the edges in the cycle be an edge from $`L`$ to $`R`$, coming from congruence number $`i`$. Then the rest of this cycle cannot use this edge and cannot use the other edge coming from congruence number $`i`$ (since this is not even in the same component), so it consists entirely of edges coming from other congruences. But the rest of the cycle gives a path from $`L`$ to $`R`$, so $`\{A_k:kL\}\{A_k:kR\}`$ is deducible from the system without using congruence number $`i`$. So congruence number $`i`$ is deducible from the others, contradicting the assumption that the system is nonredundant. $`\mathrm{}`$ Using these two claims, we get: Claim 3. Every path of length $`2^r`$ in the digraph $`𝒢`$ contains a pair of consecutive edges with labels $`f_i`$ and $`f_i^1`$, or vice versa, for some $`i`$. Proof. Suppose we have a path of length $`2^r`$ in $`𝒢`$. Since there are fewer than $`2^r`$ vertices in $`𝒢`$, some vertex must be visited more than once, so we get a nontrivial subpath which starts and ends at the same vertex (i.e., a cycle). By Claim 1, this subpath consists entirely of good edges, so it induces a corresponding path in the graph $`𝒢_0`$ which also starts and ends at the same place. By Claim 2, this latter path cannot be a nontrivial cycle, so it must double back on itself (use the same edge twice in succession); hence, the original path uses both edges of a pair of oppositely-directed good edges successively, which gives the desired conclusion. $`\mathrm{}`$ Now, for any $`gG`$, $`x_0`$ is connected to $`g(x_0)`$ by a chain of links, and this chain can be read off from the reduced form of $`g`$. In order to prove rm (4) for the sets $`\widehat{B}_k`$, it will suffice to show that, if $`M_g\mathrm{}`$, then either all of the links in this chain are active for the sets $`B_k^n`$, or the chain has fewer than $`2^r`$ links; once we know this, rm (4) for $`B_k^n`$ implies that there are only finitely many points $`g(x_0)`$ such that $`M_g\mathrm{}`$ (equivalently, since $`G`$ acts freely on the orbit of $`x_0`$, the set of $`g`$ such that $`M_g\mathrm{}`$ is finite), so only finitely many new links are activated when $`B_k^n`$ is enlarged to $`\widehat{B}_k`$, so rm (4) for $`B_k^n`$ implies rm (4) for $`\widehat{B}_k`$. So suppose $`M_g\mathrm{}`$ and the above chain has at least $`2^r`$ links. Then $`M_h\mathrm{}`$ for all of the intermediate points $`h(x_0)`$ on the chain. It must now be true that, given any $`2^r`$ consecutive links in the chain, at least one of the $`2^r+1`$ endpoints of these links is in one of the sets $`B_k^n`$, because otherwise the sets $`M_h`$ at these $`2^r+1`$ endpoints would give a counterexample to Claim 3. (If none of these points $`h(x_0)`$ is in any of the sets $`B_k^n`$, then we always have $`M_h^+=M_h`$. Now, if $`h`$ and $`h^{}=\rho h`$ are final subwords of the reduced word for $`g`$, where $`\rho `$ is $`f_i`$ or $`f_i^1`$, then the way in which $`M_h^{}`$ is computed from $`M_h`$ shows that there is an edge in $`𝒢`$ from $`M_h`$ to $`M_h^{}`$ labeled $`\rho `$. The resulting path of length $`2^r`$ cannot include consecutive edges labeled $`f_i`$ and $`f_i^1`$ or vice versa because we are working with the reduced form of $`g`$.) It follows that all $`2^r`$ of the links are active for $`B_k^n`$; since this was an arbitrary subchain of the chain, all of the links in the chain are active for $`B_k^n`$. This completes the proof of rm (4) for $`\widehat{B}_k`$. Now that we have rm (2)–rm (4) for $`\widehat{B}_k`$, we can enlarge these sets to get open sets. Let $`S`$ be the set of $`gG`$ such that $`x_0`$ is connected to $`g(x_0)`$ by a chain of links which are active for the sets $`\widehat{B}_k`$, and let $`S^{}`$ be the set of $`g^{}G`$ such that $`g^{}(x_0)`$ is connected to $`g(x_0)`$ for some $`gS`$ by a chain of at most $`2^r+1`$ links. Then $`T_kS`$ for all $`k`$, $`SS^{}`$, and $`S`$ and $`S^{}`$ are finite by rm (4). Let $`U_0`$ be an open neighborhood of $`x_0`$ so small that the images $`g(U_0)`$ for $`gS^{}`$ are pairwise disjoint and each of them is either included in or disjoint from each of the sets $`B_k^n`$. (This is possible because, by the choice of $`x_0`$, no point in $`S^{}`$ is on the boundary of any of the sets $`B_k^n`$.) Now let $`B_k^{n+1}=B_k^n\{g(U_0):gT_k\}`$ for each $`k`$; we must see that these sets satisfy properties rm (2)–rm (4). From the definition of $`B_k^{n+1}`$ and the disjointness of the sets $`g(U_0)`$ for $`gS^{}`$, the following two statements follow easily: If $`xg(U_0)`$ for some $`gS^{}`$, then $`xB_k^{n+1}`$ if and only if $`g(x_0)\widehat{B}_k`$. If $`x𝒳`$ is not in any of the sets $`g(U_0)`$ for $`gS`$, then $`xB_k^{n+1}`$ if and only if $`xB_k^n`$. We can now prove rm (2)–rm (4) for $`B_k^{n+1}`$. rm (2): If a point $`x`$ is in one of the neighborhoods $`g(U_0)`$ where $`gS`$, then $`g(x_0)_{k=1}^r\widehat{B}_k`$ by rm (2) for $`\widehat{B}_k`$, so $`x_{k=1}^rB_k^{n+1}`$; if $`x`$ is not in one of these neighborhoods, then $`x_{k=1}^rB_k^n`$ by rm (2) for $`B_k^n`$, so $`x_{k=1}^rB_k^{n+1}`$. rm (3): We prove $`f_i(_{kL_i}B_k^{n+1})_{kR_i}B_k^{n+1}`$; the other parts are similar. Suppose $`x_{kL_i}B_k^{n+1}`$. If $`xg(U_0)`$ for some $`gS`$, then $`g(x_0)_{kL_i}\widehat{B}_k`$, so $`f_i(g(x_0))_{kR_i}\widehat{B}_k`$ by rm (3) for $`\widehat{B}_k`$; but $`f_igS^{}`$ and $`f_i(x)f_i(g(U_0))`$, so $`f_i(x)_{kR_i}B_k^{n+1}`$. If $`x`$ is not in $`g(U_0)`$ for any $`gS`$, then $`x_{kL_i}B_k^n`$, so $`f_i(x)_{kR_i}B_k^n`$ by rm (3) for $`B_k^n`$. rm (4): Let $`w`$ be any point of $`𝒳`$, and consider the set of all points connected to $`w`$ by a path of links which are active for the sets $`B_k^{n+1}`$. If this set contains no point which is in $`g(U_0)`$ for any $`gS`$, then all of the links connecting the set were in fact active for $`B_k^n`$. (Note: If the link from $`x`$ to $`x^{}`$ is activated by $`x^{\prime \prime }`$, because there is a chain of at most $`2^r`$ links connecting $`x^{\prime \prime }`$ to $`x`$ or to $`x^{}`$, then all of the links in this chain are also activated by $`x^{\prime \prime }`$.) Hence, the set is finite by rm (4) for $`B_k^n`$. So suppose $`yg(U_0)`$ is connected by active links to $`w`$, and $`gS`$. A point is connected to $`w`$ if and only if it is connected to $`y`$, so it will suffice to show that only finitely many points are connected to $`y`$. Suppose $`y`$ is actively linked to $`y^{}`$, say $`y^{}=f_i(y)`$ (the case $`y^{}=f_i^1(y)`$ is similar). Let $`y^{\prime \prime }`$ be a point in one of the sets $`B_k^{n+1}`$ such that $`y^{\prime \prime }`$ is connected to either $`y`$ or $`y^{}`$ by a chain of at most $`2^r`$ links. Then there is an element $`h`$ of $`G`$ such that $`h(y)=y^{\prime \prime }`$, and the reduced form of $`h`$ in terms of the generators $`f_I`$ has length at most $`2^r+1`$ (and, if it has length $`2^r+1`$, then the rightmost component is $`f_i`$). Therefore, $`hgS^{}`$. We now have $`y^{\prime \prime }h(g(U_0))`$, so, since $`y^{\prime \prime }B_k^{n+1}`$, we must have $`h(g(x_0))\widehat{B}_k`$. This means that the link from $`g(x_0)`$ to $`f_i(g(x_0))`$ is active for the sets $`\widehat{B}_k`$, so $`f_i(y)f_i(g(x_0))`$ and $`f_igS`$. Now this argument can be repeated starting at $`y^{}`$, and so on; the result is that, for any chain of active (for the sets $`B_k^{n+1}`$) links starting at $`y`$, all of the links in the corresponding chain starting at $`g(x_0)`$ are also active (for the sets $`\widehat{B}_k`$). Furthermore, if $`y`$ is connected to two different points $`y^{}`$ and $`y^{\prime \prime }`$ by such chains of links, this will give $`y^{}=h^{}(y)`$ and $`y^{\prime \prime }=h^{\prime \prime }(y)`$ for some distinct elements $`h^{},h^{\prime \prime }`$ of $`G`$, and the corresponding points reached from $`g(x_0)`$ will be $`h^{}(g(x_0))`$ and $`h^{\prime \prime }(g(x_0))`$; since $`G`$ acts freely on the orbit of $`x_0`$, these two points will also be different. Therefore, since $`g(x_0)`$ is connected to only finitely many points, $`y`$ (and hence $`w`$) must be connected to only finitely many points. This completes the proof of rm (4) for the sets $`B_k^{n+1}`$. This completes the induction. $`\mathrm{}`$ One can use this result to give a new proof of Theorem 4.8 from Dougherty and Foreman : ###### Corollary 2.2 Suppose $`𝒳`$ is a Polish space and $`G`$ is a countable group of homeomorphisms of $`𝒳`$ which acts freely on a comeager subset of $`𝒳`$. Then, for any $`N3`$, if elements $`f_i`$ ($`1iN`$) of $`G`$ are free generators for a free subgroup of $`G`$, then there is an open subset $`A`$ of $`𝒳`$ such that the sets $`f_i(A)`$ are disjoint and their union is dense in $`𝒳`$. In fact, if $`f_{ij}G`$ for $`1ij`$ and $`3jN`$ are free generators for a free subgroup of $`G`$, then there is an open set $`A`$ such that, for each $`j`$, the sets $`f_{ij}(A)`$ for $`ij`$ are disjoint and have dense union. ###### Demonstration Proof We will prove the second part; the proof of the first part can be obtained from this by omitting most of the congruences (in fact, the first part is essentially a special case of the second). Let $`R`$ be the set of sequences $`s=s_j:3jN`$ such that $`1s_jj`$ for each $`j`$; we will construct a system of congruences between sets $`A_s`$ for $`sR`$. (Of course, one can relabel the sets to make the index set $`\{1,2,\mathrm{},r\}`$, where $`r=|R|=N!/2`$.) The congruences are: $`\{A_s:s(N)=1\}\{A_s:s(j)=i\}`$ for each pair $`(i,j)(1,N)`$ such that $`3jN`$ and $`1ij`$. The only proper congruences that can be deduced from this system are those of the form $`\{A_s:s(j)=i\}\{A_s:s(j^{})=i^{}\}`$ and their complementary versions; it follows easily that the system is weak and consistent. It is also easy to check that the system is nonredundant. Therefore, by Theorem 2.1, one can find a quasi-solution to the system using open sets $`A_s`$ for $`sR`$, where the congruence for $`(i,j)`$ is witnessed by the element $`f_{ij}f_{1N}^1`$ of $`G`$. (Since the elements $`f_{ij}`$ are free generators for their subgroup, the elements $`f_{ij}f_{1N}^1`$ for $`(i,j)(1,N)`$ are free generators for their subgroup.) Now let $`A=f_{1N}^1(\{A_s:s(N)=1\})`$. Then, for each $`(i,j)`$, $`f_{ij}(A)`$ differs from $`\{A_s:s(j)=i\}`$ by a meager set; it follows that, for each $`j`$, the sets $`f_{ij}(A)`$ for $`ij`$ are quasi-disjoint and their union is a comeager (hence dense) subset of $`𝒳`$. Since quasi-disjoint open sets must actually be disjoint, we are done. $`\mathrm{}`$ The trick used here to get congruent rather than quasi-congruent sets is quite specific; many weak consistent systems of congruences do not have quasi-solutions in open sets if one actually requires congruences instead of quasi-congruences. This will be explored further in a later paper . In order to get results concerning sets with the property of Baire, we will combine the preceding results about open sets with the Robinson-Dekker results on arbitrary sets, using the following lemma, which is a variant of Lemma 2.4 from Dougherty and Foreman : ###### Lemma 2.3 Suppose $`𝒳`$ is a Polish space, and $`f_1,\mathrm{},f_m`$ are homeomorphisms from $`𝒳`$ to $`𝒳`$. Also suppose that we have a system of $`m`$ congruences such that: there is a solution to the system in $`𝒳`$ with $`f_i`$ witnessing congruence number $`i`$ for $`i=1,2,\mathrm{},m`$; and there is a quasi-solution to the system in $`𝒳`$ using nonmeager sets with the property of Baire so that $`f_i`$ witnesses congruence number $`i`$. Then there is a solution to the system in $`𝒳`$ using nonmeager sets with the property of Baire so that $`f_i`$ witnesses congruence number $`i`$. ###### Demonstration Proof Let $`G`$ be the countable group of homeomorphisms generated by $`f_1,\mathrm{},f_m`$. Suppose the quasi-solution consists of sets $`A_k^{}`$ with the property of Baire for $`1kr`$, while the solution is given by sets $`A_k^{\prime \prime }`$, $`1kr`$. Let $`D`$ be the union of $`𝒳_{k=1}^rA_k^{}`$, the intersections $`A_k^{}A_k^{}^{}`$ for $`kk^{}`$, and the differences $`f_i(_{kL_i}A_k^{})\mathrm{}_{kR_i}A_k^{}`$ for $`im`$. Then $`D`$ is meager, and so is the union $`D^{}`$ of all of the images of $`D`$ under the elements of $`G`$. Now let $`A_k=(A_k^{}D^{})(A_k^{\prime \prime }D^{})`$. These sets have the property of Baire (since $`A_k^{}`$ has the property of Baire and $`D^{}`$ and $`A_k^{\prime \prime }D^{}`$ are meager), and they are easily seen to be disjoint; using the $`G`$-invariance of $`D^{}`$, it is easy to verify that $`f_i(_{kL_i}A_k)=_{kR_i}A_k`$ for each $`i`$. Also, since $`A_k^{}`$ is nonmeager, $`A_k`$ is nonmeager. Therefore, the sets $`A_k`$ are as desired. $`\mathrm{}`$ ###### Theorem 2.4 Suppose $`𝒳`$ is a Polish space and $`G`$ is a countable group of homeomorphisms of $`𝒳`$ which acts freely on a comeager subset of $`𝒳`$ and locally commutatively on all of $`𝒳`$, and which has a subgroup which is free on $`m`$ generators ($`m1`$). Suppose that a system of $`m`$ congruences is specified by pairs $`(L_i,R_i)`$ ($`1im)`$ of subsets of $`\{1,2,\mathrm{},r\}`$; also suppose that this system is weak and consistent. Then there is a sequence of nonmeager sets $`A_k𝒳`$ ($`kr`$) with the property of Baire which is a solution to the system. Furthermore, if the system is nonredundant, and elements $`f_i`$ ($`1im`$) of $`G`$ are free generators for a free subgroup of $`G`$, then there is a sequence of sets $`A_k`$ as above such that, for each $`im`$, $`f_i`$ witnesses that $`_{kL_i}A_k`$ is congruent to $`_{kR_i}A_k`$. ###### Demonstration Proof The second part follows immediately from Lemma 2.3, Theorem 2.1, and the results of Robinson and Dekker, while the first follows from the second as in the proof of Theorem 2.1. $`\mathrm{}`$ We now recall that a free group on two generators has subgroups which are free on any finite number of generators \[9, Problem~1.4.12\]. This allows us to simplify the statement of the following corollary, which follows from Theorem 2.4 just as Corollary 2.2 follows from Theorem 2.1: ###### Corollary 2.5 \rm\cite{\DoughertyForeman, Theorem~5.4} Suppose $`𝒳`$ is a Polish space and $`G`$ is a countable group of homeomorphisms of $`𝒳`$ which acts freely on a comeager subset of $`𝒳`$ and locally commutatively on all of $`𝒳`$, and which has a subgroup which is free on more than one generator. Then, for any $`N3`$, $`𝒳`$ can be partitioned into $`N`$ $`G`$-congruent pieces with the property of Baire; in fact, there is a set $`A𝒳`$ with the property of Baire such that, for $`3jN`$, $`𝒳`$ can be partitioned into $`j`$ pieces congruent to $`A`$. $`\mathrm{}`$ ## 3. Negative results We will now see why the systems of congruences used in Theorems 2.1 and 2.4 must be weak and consistent in order to get the desired conclusions for all suitable $`𝒳`$ and $`G`$. In fact, it suffices to look at 2.1 only; if one has a solution (or even a quasi-solution) to a system of congruences using sets $`A_k`$ with the property of Baire, and if $`A_k^{}`$ is an open set differing from $`A_k`$ by a meager set, then the sets $`A_k^{}`$ are a quasi-solution to the same system of congruences. We will look at the case of the sphere $`S^2`$ acted on by a free group of rotations. (For other spaces or groups, more systems of congruences might be solvable. For example, if $`G`$ is a free group on countably infinitely many generators and we put the discrete metric on $`G`$, then we get a Polish space acted on by $`G`$ in which every proper system of congruences has a solution \[12, Cor.~4.12\], and this solution will automatically consist of open sets because the space is discrete.) We first see why consistency is necessary, at least if we want quasi-solutions involving nonempty open sets. It is easy to verify that, if the sets $`A_k`$ are a quasi-solution to the system, and the subcongruence $`_{kR}A_k_{kR^{}}A_k`$ is deducible from the system, then $`_{kR}A_k`$ actually is quasi-congruent to a subset of $`_{kR^{}}A_k`$. We now use the following fact, a slight variant of Proposition 5.5 from Dougherty and Foreman : ###### Proposition 3.1 If $`B`$ and $`C`$ are quasi-disjoint subsets of $`S^2`$ with the property of Baire such that $`BC`$ is quasi-congruent to a subset of $`B`$, then $`C`$ is meager. ###### Demonstration Proof Let $`\sigma `$ be an isometry witnessing the quasi-congruence, and let $`B^{}`$ and $`C^{}`$ be the unique regular-open sets such that $`B\mathrm{}B^{}`$ and $`C\mathrm{}C^{}`$ are meager. (For the definition and properties of regular-open sets, see Oxtoby \[10, ~Ch.~4\]. The relevant fact is that $`B^{}`$ is the largest open set such that $`B\mathrm{}B^{}`$ is meager, and similarly for $`C^{}`$.) Since $`\sigma (BC)B`$ is meager, $`\sigma (B^{}C^{})B^{}`$ must be meager; but $`B^{}`$ is regular-open, so we have $`\sigma (B^{}C^{})B^{}`$. Also, since $`B`$ and $`C`$ are quasi-disjoint, $`B^{}C^{}`$ is a meager open set and hence empty. If $`\lambda `$ is the standard rotation-invariant probability measure on $`S^2`$, then $`\lambda (B^{})\lambda (\sigma (B^{}C^{}))=\lambda (B^{}C^{})=\lambda (B^{})+\lambda (C^{})`$, so $`\lambda (C^{})0`$, so $`C^{}`$ must be empty, so $`C`$ is meager. $`\mathrm{}`$ Therefore, if the open sets $`A_kS^2`$ are a quasi-solution for a system of congruences, and $`_{kR}A_k_{kR^{}}A_k`$ is deducible from the system where $`R^{}`$ is a proper subset of $`R`$, then $`A_k`$ must be meager and hence empty for each $`kRR^{}`$. Hence, if one insists on a solution using nonempty open sets, then the system of congruences must be consistent. If it does not matter that some of the sets are empty, then, given a system of congruences, one should proceed as follows: Find all inconsistencies $`_{kR}A_k_{kR^{}}A_k`$, $`R^{}R`$ deducible from the system, list all indices $`k`$ occurring in $`RR^{}`$, and delete the corresponding sets $`A_k`$ from the congruences (i.e., delete these indices from $`\{1,2,\mathrm{},r\}`$ and from all sets $`L_i,R_i`$ to get new sets $`L_i^{},R_i^{}`$ defining a new system of congruences). This may produce new inconsistencies; if so, repeat the process, and continue until no inconsistencies remain. If nothing is left (all sets $`A_k`$ have been declared empty), then the original system had no quasi-solutions using open subsets of $`S^2`$. Otherwise, the final system is consistent. If it is also weak, then the final system, and hence the original system, has solutions for any $`𝒳`$ and $`G`$ as in Theorem 2.4; if the final system is not weak, then we will see below that the final system has no quasi-solutions using open subsets of $`S^2`$ with a free group $`G`$ of rotations, and it follows that the original system also has no quasi-solutions in this case. Suppose $`G`$ is a free group of rotations of $`S^2`$, and we restrict ourselves to congruences which are witnessed in $`G`$; we will now see that only weak systems can have quasi-solutions using open sets in this case. (The corresponding statement about solutions using sets with the property of Baire, or even using arbitrary sets, is trivial because any rotation has fixed points and hence cannot map a set to its complement.) To see this, we use a lemma about open subsets of $`S^2`$ which are quasi-invariant under a rotation of infinite order. (A set $`A`$ is said to be quasi-invariant under a homeomorphism $`f`$ if the symmetric difference $`f(A)\mathrm{}A`$ is meager.) ###### Lemma 3.2 If an open subset $`A`$ of $`S^2`$ is invariant under a rotation of infinite order around an axis $`\mathrm{}`$, then $`A`$ is invariant under all rotations around $`\mathrm{}`$. The same applies to quasi-invariance. ###### Demonstration Proof Let $`\sigma `$ be a rotation of infinite order around $`\mathrm{}`$ under which $`A`$ is invariant. Fix $`xA`$, and let $`C`$ be the circle generated by rotating $`x`$ continuously around $`\mathrm{}`$; we must show that $`CA`$. Let $`2\pi \theta `$ be the angle through which $`\sigma `$ rotates $`S^2`$; since $`\sigma `$ does not have finite order, $`\theta `$ is irrational. It follows that the fractional parts of $`j\theta `$ for positive integers $`j`$ are dense in the interval $`(0,1)`$; equivalently, the rotations through positive integer multiples of $`2\pi \theta `$ arbitrarily well approximate any rotation around $`\mathrm{}`$. In particular, for any $`yC`$, the rotation around $`\mathrm{}`$ which takes $`y`$ to $`x`$ can be approximated by a rotation $`\sigma ^j`$ so well that $`\sigma ^j(y)`$ is in the open set $`A`$; this means that $`y\sigma ^j(A)=A`$. Since $`y`$ was arbitrary, we have $`CA`$, as desired. Now suppose the set $`A`$ is just quasi-invariant under $`\sigma `$. Let $`A^{}`$ be the regular-open set that differs from $`A`$ by a meager set; then $`\sigma (A^{})`$ is a regular-open set that differs from $`\sigma (A)`$ by a meager set. But $`\sigma (A)`$ differs from $`A`$ by a meager set, so $`\sigma (A^{})`$ is a regular-open set differing from $`A`$ by a meager set, so it must be equal to $`A^{}`$. Therefore, $`A^{}`$ is invariant under $`\sigma `$, so it is invariant under any rotation $`\tau `$ around $`\mathrm{}`$; since $`A\mathrm{}A^{}`$ and $`\tau (A)\mathrm{}\tau (A^{})`$ are meager, $`A`$ must be quasi-invariant under $`\tau `$. $`\mathrm{}`$ Now, suppose a given system has a quasi-solution in $`S^2`$ using a free group $`G`$ of rotations, but is not weak; fix a set $`L\{1,\mathrm{},r\}`$ such that the congruence $`_{kL}A_k_{kL^c}A_k`$ can be deduced from the system. Then this quasi-congruence is witnessed by some $`\sigma G`$, which clearly is not the identity and therefore must be a rotation of infinite order. But then $`\sigma ^2`$ is also a rotation of infinite order, and $`_{kL}A_k`$ is quasi-invariant under $`\sigma ^2`$, so it is quasi-invariant under all rotations around the axis of $`\sigma ^2`$. In particular, $`_{kL}A_k`$ is quasi-invariant under $`\sigma `$, so $`_{kL}A_k`$ differs from $`_{kL^c}A_k`$ by a meager set, which is impossible because $`_{kL}A_k`$ differs from the complement of $`_{kL^c}A_k`$ by a meager set, and $`S^2`$ is not meager. This contradiction shows that the non-weak system had no quasi-solution after all. This shows why weakness and consistency are required in Theorem 2.1. Next, we consider the requirement of nonredundancy for a system of congruences to be satisfied with specified witnesses to the congruences. Even for a simple redundant system such as $`A_1A_1`$, $`A_1A_1`$, $`A_1A_2`$, $`A_1A_3`$, one cannot arbitrarily specify the witnesses for the congruences: if the first two congruences are witnessed by rotations of infinite order around different axes, then Lemma 3.2 implies that $`A_1`$ must be quasi-invariant under any rotation around either of these axes, so $`A_1`$ must be either empty or comeager. (By considering the regular-open set which differs from $`A_1`$ by a meager set, we can reduce this claim to the corresponding claim about invariant sets: if $`A`$ is a nonempty open proper subset of $`S^2`$, then $`A`$ cannot be invariant under all rotations around either of two different axes. To see this, note that a connected component of $`A`$ must have nonempty boundary. If $`A`$ is invariant around an axis, then this boundary consists of one or two parts, each of which is a point on the axis or a circle obtained by revolving a point around the axis, so the axis can be reconstructed given the boundary.) Either of these makes the rest of the system impossible to satisfy. With a little more work, one can show that requiring even a single redundant congruence to be witnessed by a rotation of infinite order, such as in the system $`A_1A_1`$, $`A_1A_2`$, $`A_1A_3`$, can make a system unsatisfiable. (It would require $`A_1`$ to be a union of spherical disks and annuli with a common axis, and $`A_2`$ and $`A_3`$ would also have to be such unions but around different axes; such sets cannot fit together closely enough to cover a dense subset of $`S^2`$.) Note that, if $`G`$ is a countable free group of rotations of $`S^2`$, then, since each element of $`G`$ other than the identity has only two fixed points, $`G`$ acts freely on $`S^2D`$ for some countable set $`D`$. But $`S^2D`$ is a $`G_\delta `$ set in $`S^2`$ and is therefore a Polish space itself \[8, ~§33~VI\], and the preceding paragraphs hold for this new space as well. Hence, even in a Polish space on which a free group of homeomorphisms acts freely, one cannot guarantee that a system of congruences has a quasi-solution using open sets unless the reduced form of that system (after deleting inconsistencies as above) is weak and consistent. Another way to modify the space $`S^2`$ is as follows: Let $`G`$ be a free group of rotations of $`S^2`$ on $`\mathrm{}_0`$ generators, and let $`z`$ be a point of $`S^2`$ such that $`G`$ acts freely on the $`G`$-orbit of $`z`$; fix another point $`z^{}`$ of $`S^2`$ which is neither $`z`$ nor the point opposite $`z`$. For each $`gG`$, consider the space $`𝒳_g`$ which is the union of $`S^2`$ and its tangent ray at $`g(z)`$ in the direction of the (shortest) great-circle arc from $`g(z)`$ to $`g(z^{})`$, with the standard Euclidean metric from $`^3`$. Note that $`h(𝒳_g)=𝒳_{hg}`$ if we view the rotations as acting on all of $`^3`$. Now take a copy of $`𝒳_g`$ for each $`g`$, and identify corresponding points of $`S^2`$ to get a space $`𝒳`$. (Tangent rays that happen to intersect will not have their common points identified. If $`x,y𝒳`$ are in different tangent rays, then the distance from $`x`$ to $`y`$ in $`𝒳`$ is the length of the shortest path in $`^3`$ from $`x`$ to $`y`$ via a point of $`S^2`$.) Then $`𝒳`$ is a Polish space, and its group of isometries is precisely $`G`$; $`G`$ acts locally commutatively on $`𝒳`$ and freely on $`𝒳D`$ for a countable meager set $`D`$, and the negative results given above for $`S^2`$ also apply to $`𝒳`$, so we can get such results even when using the entire isometry group of a suitable Polish space. ## 4. Congruences on the sphere using all isometries We now know what congruences have solutions using subsets of $`S^2`$ with the property of Baire and using free rotations; it is natural to ask what can be done if arbitrary isometries of $`S^2`$ are allowed. The results in the preceding section concerning consistency apply just as well for arbitrary isometries, so even here a system must be consistent (or at least reducible to a consistent system by deletion of some sets) in order to have such a solution. However, it turns out that the restriction of weakness can be removed if we allow arbitrary isometries to witness the congruences. As usual, one of the ingredients needed for the proof is a corresponding result for arbitrary subsets of $`S^2`$; this result is due to Adams (see also Wagon \[12, ~Theorem~4.16\]). Unfortunately, Adams’ proof cannot be used here; the particular isometries he uses to witness the congruences cannot be used to get a corresponding result concerning open sets. (Adams’ construction causes a complementary congruence $`AA^c`$ to be witnessed by an isometry $`\tau `$ such that $`\tau ^2`$ is a rotation of infinite order; then $`A`$ is invariant under $`\tau ^2`$, and we saw in the preceding section why this cannot work for open sets.) We therefore give a revised proof of this result. ###### Theorem 4.1 \rm(Adams) Any system of proper congruences has a solution in $`S^2`$, if arbitrary isometries can be used as witnesses for the congruences. ###### Demonstration Proof We first transform the system into an equivalent system having a useful form. Call two systems of congruences (on the same index set $`\{1,2,\mathrm{},r\}`$) equivalent if any congruence in one can be deduced from the other, and vice versa; clearly equivalent systems have the same solutions. By moving to an equivalent system if necessary, we may assume that the system is presented with as few congruences as possible (i.e., there is no equivalent system with fewer congruences). Now, suppose the system (call it $`S_0`$) is not weak. Let $`M_0`$ be a subset of $`\{1,2,\mathrm{},r\}`$ such that the congruence $`_{kM_0}A_k_{kM_0^c}A_k`$ is deducible from $`S_0`$; choose $`M_0`$ so that this deduction requires as few steps as possible. Such a deduction gives a sequence $`M_0,M_1,\mathrm{},M_n`$ of subsets of $`\{1,2,\mathrm{},r\}`$ such that $`M_n=M_0^c`$ and each pair $`(M_i,M_{i+1})`$ appears as one of the congruences in $`S_0`$, perhaps in the complemented form $`(M_i^c,M_{i+1}^c)`$. Because the deduction is minimal, no set appears more than once in the list $`M_0,M_1,\mathrm{},M_n`$, and the only case where both a set and its complement appear in the list is $`M_n=M_0^c`$. But this easily implies that no congruence in $`S_0`$ is used more than once during the deduction: if it were used twice in the same form, this would require a duplication in the list, while if it were used once in the given form and once in the complemented form (assuming these are different), there would be more than one instance of a set and its complement appearing in the list. Let $`S_1`$ be $`S_0`$ with the last congruence used in the above deduction (the one between $`M_{n1}`$ and $`M_n`$, or maybe their complements) deleted, and let $`S_0^{}`$ be $`S_1`$ together with the congruence $`_{kM_0}A_k_{kM_0^c}A_k`$. Since the congruence between $`M_0`$ and $`M_{n1}`$ is deducible from $`S_1`$, the congruence between $`M_n`$ and $`M_{n1}`$ is deducible from $`S_0^{}`$, so $`S_0^{}`$ is equivalent to $`S_0`$. Now look at $`S_1`$, ignoring the new self-complement congruence. If $`S_1`$ is not weak, one can repeat the above process to change $`S_1`$ into an equivalent system $`S_1^{}`$ with the same number of congruences, where $`S_1^{}`$ is $`S_2`$ together with another self-complement congruence. Repeat this process as many times as possible, until we reach a system $`S_j`$ which is weak. Let the congruences in $`S_j`$ be given by pairs $`(L_i,R_i)`$ for $`1r\overline{m}`$, and let the self-complement congruences be given as $`(L_i,R_i)`$ (with $`R_i=L_i^c`$) for $`\overline{m}+1im`$. (If the original system was weak, then $`\overline{m}=m`$.) We have now found a system equivalent to the original system, with a minimal number $`m`$ of congruences (the same number as in $`S_0`$), such that the first $`\overline{m}`$ congruences form a weak system and the remaining $`m\overline{m}`$ congruences are between a set and its complement. Since $`m`$ is minimal, this system is nonredundant. The main result of Dekker states that any reasonable-sized (continuum or smaller) free product of cyclic groups can be embedded in the rotation group of $`S^2`$. Therefore, we can choose rotations $`\sigma _i`$ ($`1i\overline{m}`$) and $`\tau _i^{}`$ ($`\overline{m}<im`$) such that each $`\sigma _i`$ has infinite order, each $`\tau _i^{}`$ has order $`4`$, and the group $`G^{}`$ generated by all of these rotations is the free product of the cyclic groups generated by the rotations individually. Let $`\zeta `$ be the antipodal isometry which maps each point of $`S^2`$ to the point opposite it. Let $`\tau _i=\zeta \tau _i^{}`$ for each $`i`$, and let $`G`$ be the group generated by the isometries $`\sigma _i`$ and $`\tau _i`$. We will construct a solution to the revised system of congruences so that $`\sigma _i`$ ($`i\overline{m}`$) or $`\tau _i`$ ($`i>\overline{m}`$) witnesses congruence number $`i`$. Clearly $`\zeta ^2`$ is the identity on $`S^2`$. Since isometries of $`S^2`$ preserve oppositeness of points, $`\zeta `$ commutes with all isometries of $`S^2`$. Using this, we see that $`\tau _i`$ has order $`4`$ in $`G`$, and the homomorphism from $`G^{}`$ to $`G`$ which sends $`\sigma _i`$ to $`\sigma _i`$ and $`\tau _i^{}`$ to $`\tau _i`$ (which exists and is unique by the definition of free products) is in fact an isomorphism. Therefore, $`G`$ is also a free product of $`\overline{m}`$ copies of $``$ and $`m\overline{m}`$ copies of $`_4`$. Also, $`GG^{}`$ is a group of index $`2`$ in $`G`$ and in $`G^{}`$, consisting of those words in $`G^{}`$ such that the total number of occurrences of the generators $`\tau _i^{}`$ is even. We will now show that much of Robinson’s work on free groups, as presented in Chapter 4 of Wagon , can be carried out as well for free products of $``$’s and $`_4`$’s. The rest of this proof will follow the relevant parts of that chapter rather closely. First, we look at the structure of the group $`G`$ (of course, the same results will apply to $`G^{}`$, which is isomorphic to $`G`$). Any element of $`G`$ has a unique expression as a reduced word. Here a ‘word’ is a product (possibly of length $`0`$) of elements $`\sigma _i^{\pm 1}`$ and $`\tau _i^{\pm 1}`$; a word is reduced if there is no occurrence of $`\sigma _i\sigma _i^1`$, $`\sigma _i^1\sigma _i`$, $`\tau _i^4`$, or $`\tau _i^1`$ (which is equal to $`\tau _i^3`$). Given such a reduced word $`g`$, we can express it in the form $`h_1h_2h_3`$ with $`h_3=h_1^1`$ where $`h_1`$ and $`h_3`$ include as much of the word $`g`$ as possible. (For a reduced word $`h_1`$, the inverse reduced word $`h_1^1`$ is obtained by reversing the word and then replacing $`\sigma _i`$ with $`\sigma _i^1`$, $`\sigma _i^1`$ with $`\sigma _i`$, and maximal consecutive blocks $`\tau _i^j`$ with $`\tau _i^{4j}`$.) To do this, start by setting $`h_1`$ and $`h_3`$ to be the identity and $`h_2`$ to be $`g`$. If the current $`h_2`$ starts with $`\sigma _i`$ and ends with $`\sigma _i^1`$, transfer the $`\sigma _i`$ to $`h_1`$ and the $`\sigma _i^1`$ to $`h_3`$; similarly if $`h_2`$ starts with $`\sigma _i^1`$ and ends with $`\sigma _i`$. If the current $`h_2`$ both starts and ends with one or more copies of $`\tau _i`$, with a total of at least $`4`$ such copies (but $`h_2`$ is not just a power of $`\tau _i`$), then transfer $`j`$ copies from the start of $`h_2`$ to $`h_1`$ and $`4j`$ copies from the end of $`h_2`$ to $`h_3`$, where $`j`$ is $`1`$$`2`$, or $`3`$, as appropriate. (If there are more than $`4`$ such copies at the ends of $`h`$, so that more than one choice of $`j`$ is possible, then use $`j=2`$, for a reason to be seen below.) Repeat until $`h_2`$ cannot be reduced further. When $`g`$ is expressed as above, it is easy to see that the reduced form of $`g^n`$ is $`h_1h_2^nh_3`$ for any positive $`n`$, unless $`h_2`$ is of the form $`\tau _i^k`$, in which case the reduced form of $`g^n`$ is the null word (the identity) if $`kn`$ is divisible by $`4`$, $`h_1\tau _i^{kn\text{ mod }4}h_3`$ otherwise. (If $`h_2`$ is not of the form $`\tau _i^k`$, then the fact that $`h_2`$ cannot be reduced further as above shows that $`h_2^n`$ is a reduced word.) In fact, the same statement also holds for negative $`n`$, because we used $`j=2`$ whenever possible in the preceding paragraph. Furthermore, if $`g^n`$ is broken into three pieces as above, then the three pieces are precisely $`h_1`$$`h_2^n`$, and $`h_3`$. From these facts, it follows immediately that the only elements of $`G`$ of finite order are conjugates of powers $`\tau _i^k`$. In particular, the only elements of $`G`$ of order $`2`$ have the form $`g\tau _i^2g^1`$ for some $`g`$ and $`i`$. (A similar statement holds for $`G^{}`$, of course.) One more fact we will need is that the only abelian subgroups of $`G`$ are the cyclic subgroups (so, if two elements of $`G`$ commute, then they are powers of a single element of $`G`$). This follows from the Kurosh Subgroup Theorem \[9, ~Cor.~4.9.1\]. We now start to work out the analogues for $`G`$ of Robinson’s results for free groups. ###### Lemma 4.2 The group $`G`$ above can be partitioned into subsets $`A_1,A_2,\mathrm{},A_r`$ satisfying the given system of congruences, with $`\sigma _i`$ ($`i\overline{m}`$) or $`\tau _i`$ ($`i>\overline{m}`$) witnessing congruence number $`i`$ for each $`im`$. In fact, for any word $`w`$ from $`G`$ in which the total number of occurrences of the generators $`\tau _i`$ is even, there is such a partition of $`G`$ which puts $`w`$ in the same set $`A_k`$ as the identity element $`1`$ of $`G`$. ###### Demonstration Proof First, we show that the subsets of $`\{1,2,\mathrm{},r\}`$ can be colored with two colors so that: for any set $`L`$, $`L`$ and $`L^c`$ have opposite colors; for any $`i\overline{m}`$, $`L_i`$ and $`R_i`$ have the same color. To do this, define an equivalence relation on subsets of $`\{1,2,\mathrm{},r\}`$ as follows: $`L`$ is equivalent to $`L^{}`$ iff the congruence $`LL^{}`$ can be deduced from the first $`\overline{m}`$ congruences of the given system. Clearly, if $`L`$ is equivalent to $`L^{}`$, then $`L^c`$ is equivalent to $`L^c`$. Also, $`L`$ is never equivalent to $`L^c`$, since the first $`\overline{m}`$ of the given congruences form a weak system. Therefore, the equivalence classes under this relation come in complementary pairs; if we assign colors to the equivalence classes so that complementary classes get opposite colors, then the induced coloring of the subsets of $`\{1,2,\mathrm{},r\}`$ will be as desired. We can view the given $`m`$ congruences as $`2m`$ formal inclusions: the equation $`\sigma _i(_{kL_i}A_k)=_{kR_i}A_k`$ can be expressed as the two inclusions $`\sigma _i(_{kL_i}A_k)_{kR_i}A_k`$ and $`\sigma _i^1(_{kR_i}A_k)_{kL_i}A_k`$, and similarly for $`\tau _i`$. We will therefore use the terms ‘domain of $`\sigma _i`$’ and ‘range of $`\sigma _i`$’ for the sets $`_{kL_i}A_k`$ and $`_{kR_i}A_k`$, respectively, and define ‘domain of $`\tau _i`$,’ ‘range of $`\tau _i^1`$,’ and so on similarly. Suppose $`w=\rho _n\rho _{n1}\mathrm{}\rho _1`$, where each $`\rho _k`$ is $`\sigma _i^{\pm 1}`$ or $`\tau _i`$ for some $`i`$ (and this is the reduced form of $`w`$). We will first assign the end segments $`1,\rho _1,\rho _2\rho _1,\mathrm{},w`$ to suitable sets $`A_k`$, and then handle the remaining elements of $`G`$. First, suppose that, for some $`jn`$, the range of $`\rho _j`$ is neither the domain of $`\rho _{j+1}`$ nor the complement of the domain of $`\rho _{j+1}`$ (here we let $`\rho _{n+1}=\rho _1`$). Then either the range of $`\rho _j`$ meets both the domain of $`\rho _{j+1}`$ and its complement, or the complement of the range of $`\rho _j`$ meets both the domain of $`\rho _{j+1}`$ and its complement; let $`S`$ be the domain of $`\rho _j`$ in the former case, the complement of this domain in the latter case. Assign $`\rho _{j1}\mathrm{}\rho _1`$ to one of the sets in $`S`$. Next, assign $`\rho _{j2}\mathrm{}\rho _1`$ to an appropriate set $`A_k`$; this will be a set in the domain of $`\rho _{j1}`$ if $`\rho _{j1}\mathrm{}\rho _1`$ is in the range of $`\rho _{j1}`$, and a set not in this domain if $`\rho _{j1}\mathrm{}\rho _1`$ is not in this range. Repeat this process to assign all of the shorter end segments of $`w`$, down to $`1`$, to sets $`A_k`$. Put $`w`$ in the same set as $`1`$, and then assign $`\rho _{n1}\mathrm{}\rho _1`$ and so on; continue until only $`\rho _j\mathrm{}\rho _1`$ remains unassigned. This word must be assigned to the range of $`\rho _j`$ if $`S`$ is the domain of $`\rho _j`$, the complement of this range otherwise; it also must be placed in the domain of $`\rho _{j+1}`$ if $`\rho _{j+1}\mathrm{}\rho _1`$ is in the range of $`\rho _{j+1}`$, the complement of this domain otherwise. By the definition of $`S`$, these two requirements can both be met. (This must be reworded slightly in the case $`j=n`$, but the basic argument remains the same.) Now, suppose that the preceding case does not hold; for every $`j`$, the range of $`\rho _j`$ is either the domain of $`\rho _{j+1}`$ or its complement. Let $`S_0`$ be the domain of $`\rho _1`$. Given a set $`S_{j1}`$ which is either the domain of $`\rho _j`$ or the complement of the domain of $`\rho _j`$, let $`S_j`$ be the range of $`\rho _j`$ in the former case, the complement of the range of $`\rho _j`$ in the latter. Then $`S_n`$ must be either $`S_0`$ or the complement of $`S_0`$. Note that each $`S_j`$ is a union of sets $`A_k`$, say $`_{kN_j}A_k`$, and therefore $`S_j`$ (actually, $`N_j`$) has had a color assigned earlier in the proof of the Lemma. Furthermore, if $`\rho _j`$ is $`\sigma _i^{\pm 1}`$, then $`S_{j1}`$ and $`S_j`$ have the same color; if $`\rho _j`$ is $`\tau _i`$, then $`S_{j1}`$ and $`S_j`$ have opposite colors. Since the number of $`j`$’s for which $`\rho _j`$ is a generator $`\tau _i`$ is even (by hypothesis on $`w`$), $`S_n`$ must have the same color as $`S_0`$, so $`S_n`$ must be $`S_0`$ rather than the complement of $`S_0`$. We now easily assign each end segment $`\rho _j\mathrm{}\rho _1`$ to one of the sets $`A_k`$ included in $`S_j`$, making sure to put $`1`$ and $`w`$ in the same set included in $`S_0`$; these assignments are compatible with the required inclusions. Now that we have assigned the end segments of $`w`$ to sets $`A_k`$, the remaining elements $`g`$ of $`G`$ can be assigned by an easy recursion on the reduced form of $`g`$. Suppose this reduced form starts with $`\rho `$, where $`\rho `$ is $`\sigma _i^{\pm 1}`$ or $`\tau _i`$, and $`g=\rho g^{}`$ where $`g^{}`$ has already been assigned to a set $`A_k`$. Then, if $`g^{}`$ is in the domain of $`\rho `$, assign $`g`$ to the range of $`\rho `$; if $`g^{}`$ is in the complement of the domain of $`\rho `$, assign $`g`$ to the complement of the range of $`\rho `$. We must verify that, if $`g=\rho g^{}`$ where $`\rho `$ is $`\sigma _i`$ or $`\tau _i`$, then $`g`$ is in the range of $`\rho `$ if and only if $`g^{}`$ is in the domain of $`\rho `$. Let $`v`$ and $`v^{}`$ be the reduced words for $`g`$ and $`g^{}`$. If $`v=\rho v^{}`$ and $`v`$ is an end segment of $`w`$, then the way in which the end segments of $`w`$ were assigned to sets $`A_k`$ gives the desired result here; the same applies if $`v^{}=\rho ^1v`$ and $`v^{}`$ is an end segment of $`w`$. If $`v=\rho v^{}`$ and $`v`$ is not an end segment of $`w`$, then we get the desired result from the recursive definition of the preceding paragraph; this also holds if $`v^{}=\rho ^1v`$ and $`v^{}`$ is not an end segment of $`w`$. The only remaining case is when neither $`\rho v^{}`$ nor $`\rho ^1v`$ is reduced. This can happen only when $`\rho `$ is $`\tau _i`$ and $`v^{}=\tau _i^3v`$. But then, by the preceding cases, we have $`vL_i`$ iff $`\tau _ivL_i^c`$ iff $`\tau _i^2vL_i`$ iff $`\tau _i^3vL_i^c`$; hence, we have the desired result in this case as well. Therefore, the sets $`A_k`$ satisfy the system of congruences. $`\mathrm{}`$ Another useful fact is that, if $`w`$ is a word in the generators of $`G`$ which has an odd number of occurrences of the generators $`\tau _i`$ (including as inverses), then the corresponding isometry of $`S^2`$ has no fixed points. Let $`w^{}`$ be the element of $`G^{}`$ corresponding to $`w`$ (i.e., replace all generators $`\tau _i`$ with $`\tau _i^{}`$). Since $`\tau _i=\zeta \tau _i^{}`$, $`\zeta `$ commutes with all elements of $`G^{}`$, $`\zeta ^2`$ is the identity, and the number of occurrences of the generators $`\tau _i`$ in $`w`$ is odd, we can compute that $`w=\zeta w^{}`$. Now $`w^{}`$ is a rotation which cannot be of order $`2`$, since the only elements of $`G^{}`$ of order $`2`$ are the conjugates of $`\tau _I^2`$, which all have even numbers of generators $`\tau _i^{}`$. But it is easy to see that the only rotations of $`S^2`$ which map some point to its antipodal point are rotations of order $`2`$. Therefore, $`w^{}`$ does not map any point to its antipodal point, so $`w=\zeta w^{}`$ has no fixed points. We now resume the proof of Theorem 4.1. In order to get the desired partition of $`S^2`$, it will suffice to get such a partition for each $`G`$-orbit in $`S^2`$ and put them together (using the axiom of choice to choose one such partition for each orbit). So consider one such orbit $`𝒪`$. If $`G`$ acts freely on $`𝒪`$, then fixing any element $`x`$ of $`𝒪`$ determines a bijection $`gg(x)`$ from $`G`$ to $`𝒪`$ which preserves the action of $`G`$, so any partition of $`G`$ as in Lemma 4.2 can be transferred to $`𝒪`$, giving a partition of $`𝒪`$ with the desired properties. So suppose $`G`$ does not act freely on $`𝒪`$. Let $`w`$ be a non-identity reduced word of $`G`$, as short as possible, such that $`w`$ has a fixed point in $`𝒪`$; let $`x`$ be such a fixed point. Then $`w`$ cannot start with $`\sigma _i^1`$ and end with $`\sigma _i`$, because, if it did, then the reduced form of $`\sigma _iw\sigma _i^1`$ would be shorter than $`w`$ and would have a fixed point $`\sigma _i(x)𝒪`$. Similarly, $`w`$ cannot start with $`\sigma _i`$ and end with $`\sigma _i^1`$; and $`w`$ cannot start with $`\tau _i^j`$ and end with $`\tau _i^j^{}`$ where $`j+j^{}4`$, except in the case that $`w`$ is just a power of $`\tau _i`$. In fact, if $`w`$ is not a power of $`\tau _i`$, then we may assume that $`w`$ does not both start and end with $`\tau _i`$; if it ends with $`\tau _i^j`$, then we can replace $`w`$ with the reduced form of $`\tau _i^jw\tau _i^j`$, which still starts with $`\tau _i`$ but ends with something else, and has the fixed point $`\tau _i^j(x)`$. We also know that $`w`$ has an even number of occurrences of generators $`\tau _i`$ (and hence represents a rotation of $`S^2`$); in particular, if $`w`$ is a power of some $`\tau _i`$, then $`w`$ must be $`\tau _i^2`$. Let $`\rho `$ be the leftmost term in the reduced word $`w`$ (either $`\sigma _i`$, $`\sigma _i^1`$, or $`\tau _i`$ for some $`i`$). Define $`\rho ^{}`$ to be $`\sigma _i^1`$ if $`\rho =\sigma _i`$, $`\sigma _i`$ if $`\rho =\sigma _i^1`$, and $`\tau _i`$ if $`\rho =\tau _i`$; we have ensured that $`w`$ does not end with $`\rho ^{}`$, unless $`w=\tau _i^2`$ for some $`i`$. Therefore, the word $`w^n`$ is already in reduced form for positive $`n`$, and the reduced form of $`w^n`$ for negative $`n`$ does not begin with $`\rho `$, unless $`w=\tau _i^2`$. The next thing to show is that the only elements of $`G`$ which fix $`x`$ are the powers of $`w`$. Suppose $`v`$ is a nonidentity member of $`G`$ such that $`v(x)=x`$. Then $`v`$ must also have an even number of occurrences of generators $`\tau _i`$, and is therefore a rotation. Since the rotation group acts locally commutatively on $`S^2`$, $`v`$ and $`w`$ must commute, so together they generate an abelian subgroup of $`G`$. We noted earlier that any abelian subgroup of $`G`$ is cyclic, so there must be an element $`u`$ of $`G`$ such that $`v`$ and $`w`$ are both powers of $`u`$. We may assume that $`w`$ is a positive power of $`u`$ (replace $`u`$ with $`u^1`$ if necessary). If $`u`$ has finite order, then $`u=g^1\tau _i^jg`$ for some $`g`$$`i`$, and $`j`$; since $`v`$ and $`w`$ are nonidentity rotations and are powers of $`u`$, we must have $`v=w=g^1\tau _i^2g`$, so $`v`$ is a power of $`w`$. Now suppose $`u`$ is of infinite order. From the general arguments about the structure of $`G`$ given earlier (specifically, the expression of $`u`$ in the form $`h_1h_2h_3`$ so that the reduced form of $`u^n`$ is $`h_1h_2^nh_3`$ for any $`n>0`$), we see that, if $`n>k>0`$, then the reduced form of $`u^n`$ is longer than the reduced form of $`u^k`$. Suppose $`w=u^n`$ and $`v=u^j`$, and let $`k`$ be the greatest common divisor of $`n`$ and $`j`$. Then $`u^k`$ can be expressed as a power of $`v`$ times a power of $`w`$ (by applying the extended Euclidean algorithm to $`n`$ and $`j`$), so $`u^k(x)=x`$. We clearly have $`kn`$, but we cannot have $`k<n`$, since otherwise $`u^k`$ would be shorter than $`w`$, contradicting the choice of $`w`$ as the shortest possible word with a fixed point in $`𝒪`$. Therefore, $`k=n`$, so $`j`$ is divisible by $`n`$, so $`v`$ is a power of $`w`$, as desired. Using the above, we now show that every element of the orbit $`𝒪`$ has a unique expression of the form $`g(x)`$, where $`g`$ is an element of $`G`$ whose reduced form does not end in $`w`$ and does not end in $`\rho ^{}`$. (Exception: if $`w=\tau _i^2`$, then the reduced form of $`g`$ is allowed to end in $`\rho ^{}=\tau _i`$, but not in $`\tau _i^2`$.) Let $`y`$ be any point in this orbit, and let $`v`$ be a shortest reduced word in $`G`$ such that $`v(x)=y`$. Clearly $`v`$ cannot end in $`w`$ (otherwise the reduced form of $`vw^1`$ is shorter). If $`w`$ is not of the form $`\tau _i^2`$, and $`v`$ ends in $`\rho ^{}`$, then $`vw`$ does not end in $`w`$, and it does not end in $`\rho ^{}`$ either, since $`w`$ does not end in $`\rho ^{}`$. (If the entire $`w`$ cancels out when $`vw`$ is transformed to reduced form, then $`vw`$ has a shorter reduced form than $`v`$, contradicting the choice of $`v`$.) Therefore, we can take $`g`$ to be either $`v`$ or $`vw`$. To see that $`g`$ is unique, suppose $`u`$ and $`v`$ are distinct and $`u(x)=v(x)=y`$. Then $`(u^1v)(x)=x`$, so $`u^1v`$ is a power of $`w`$; by interchanging $`u`$ and $`v`$ if necessary, we may ensure that $`u^1v`$ is a positive power of $`w`$, say $`w^j`$. Then either $`v=uw^j`$ ends in $`w`$, or there is some cancellation when $`u`$ is multiplied by $`w^j`$; in the latter case, $`u`$ must end in $`\rho ^{}`$ (in $`\tau _i^2`$ if $`w=\tau _i^2`$, since in this case $`w^j`$ must be $`w`$). This completes the proof that $`g`$ is unique. Now, apply Lemma 4.2 to partition $`G`$ into sets $`A_1,A_2,\mathrm{},A_r`$ satisfying the congruences, so that $`w`$ is in the same set $`A_k`$ as $`1`$. This lets us partition $`𝒪`$ into sets $`B_1,B_2,\mathrm{},B_r`$ as follows: for any point $`y𝒪`$, find the unique expression $`g(x)`$ for $`y`$ as above, and put $`yB_k`$ iff $`gA_k`$. We must see that the sets $`B_k`$ satisfy the system of congruences. First, suppose $`y𝒪`$, $`i\overline{m}`$, and $`z=\sigma _i(y)`$; we must see that $`y_{kL_i}B_k`$ if and only if $`z_{kR_i}B_k`$. Express $`y`$ and $`z`$ as $`g(x)`$ and $`h(x)`$, where $`g`$ and $`h`$ do not end in $`w`$ and (if $`w`$ is not of the form $`\tau _I^2`$) do not end in $`\rho ^{}`$. Then $`y_{kL_i}B_k`$ iff $`g_{kL_i}A_k`$, and $`z_{kR_i}B_k`$ iff $`h_{kR_i}A_k`$. Also, since the sets $`A_k`$ satisfy the congruences, we have $`g_{kL_i}A_k`$ iff $`\sigma _ig_{kR_i}A_k`$, and $`\sigma _i^1h_{kL_i}A_k`$ iff $`h_{kR_i}A_k`$. Therefore, we are done if $`h=\sigma _ig`$. So suppose $`h\sigma _ig`$. Then the reduced form of $`\sigma _ig`$ must end in $`w`$ or in $`\rho ^{}`$, while the reduced form of $`g`$ does not. There are only two cases in which this can happen: either $`\rho =\sigma _i`$ and $`\sigma _ig=w`$, or $`\rho ^{}=\sigma _i`$ and $`g=1`$. In the first of these cases we have $`h=1`$, and since $`1`$ and $`w`$ lie in the same set $`A_k`$, we have $`h_{kR_i}A_k`$ iff $`\sigma _ig_{kR_i}A_k`$, and this gives the desired result. In the second case, we have $`\sigma _i^1h=w`$, so $`g_{kL_i}A_k`$ iff $`\sigma _i^1h_{kL_i}A_k`$, and again we are done. Now suppose $`y𝒪`$, $`i>\overline{m}`$, and $`z=\tau _i(y)`$. Define $`g`$ and $`h`$ as above. Repeating this argument, we again see that we are done unless $`h\tau _ig`$. Again this happens in only two cases: either $`\rho =\tau _i`$ and $`\tau _ig=w`$, or $`\rho ^{}=\tau _i`$ and $`g=1`$ (and $`w\tau _i^2`$). These two cases are handled just as before. This completes the construction of the desired partition for an arbitrary orbit of $`S^2`$ under $`G`$, so we are done. $`\mathrm{}`$ The corresponding result for open sets is: ###### Theorem 4.3 Any consistent system of proper congruences has a quasi-solution in $`S^2`$ using nonempty open sets (and arbitrary isometries). ###### Demonstration Proof Revise the system as in the first three paragraphs of the proof of Theorem 4.1, and define isometries $`\sigma _i`$$`\tau _i^{}`$, $`\zeta `$, and $`\tau _i`$ and groups $`G`$ and $`G^{}`$ as in the fourth paragraph of that proof; we will use the same isometries as witnesses here. The proof will follow that of (the second part of) Theorem 2.1 quite closely, so we will just give the differences here. Let $`f_i`$ be $`\sigma _i`$ if $`i\overline{m}`$, $`\tau _i`$ if $`i>\overline{m}`$. The definition of ‘active link’ is changed slightly: a link from $`x`$ to $`x^{}`$ will be considered active for the sets $`B_k^n`$ if there is a point in one or more of these sets which is connected to $`x`$ or to $`x^{}`$ by a chain of at most $`2^{r+1}`$ (rather than $`2^r`$) links. The next change is at the proof that, if $`g=f_ig^{}`$, then $`L_iM_g^{}^+`$ iff $`R_iM_g^+`$, and $`L_i^cM_g^{}^+`$ iff $`R_i^cM_g^+`$. If $`f_i`$ is $`\sigma _i`$, then the argument is unchanged, but if $`f_i`$ is $`\tau _i`$, then the cases are slightly different. If the reduced form of $`g^{}`$ does not start with $`\tau _i^3`$, then $`M_g`$ is defined from $`M_g^{}`$, and we get the desired result as before. If the reduced form of $`g^{}`$ does start with $`\tau _i^3`$, so that $`g^{}`$ is $`\tau _i^3g`$, then $`M_g^{}`$ is defined from $`M_{\tau _i^2g}`$, which is defined from $`M_{\tau _ig}`$, which is defined from $`M_g`$. Now, using the preceding case and the fact that $`R_i=L_i^c`$, we get $`L_iM_g^{}^+`$ iff $`L_i^cM_{\tau _i^2g}^+`$ iff $`L_iM_{\tau _ig}^+`$ iff $`L_i^cM_g^+`$; similarly, $`L_i^cM_g^{}^+`$ iff $`L_iM_g^+`$, as desired. Next, we must give revised forms of the Claims. Define a labeled directed graph $`𝒢`$ from the system of congruences as before, except that the edges labeled $`f_i^1`$ for $`i>\overline{m}`$ are omitted. Again the digraph $`𝒢`$ has cycles of length $`2`$ connecting pairs $`(L_i,R_i)`$ or $`(L_i^c,R_i^c)`$, for $`i\overline{m}`$; each such cycle consists of an $`\sigma _i`$-edge and an $`\sigma _i^1`$-edge. For $`i>\overline{m}`$ we also get $`2`$-cycles between $`L_i`$ and $`R_i=L_i^c`$; in this case both edges in the cycle will be labeled $`\tau _i`$. Call the edges in all of these $`2`$-cycles good edges, and call all other edges bad edges. Since the system is still consistent, the same argument as before gives: Claim 1. No cycle in $`𝒢`$ contains a bad edge. $`\mathrm{}`$ Again construct the undirected graph $`𝒢_0`$ by treating each pair of oppositely-directed good edges in $`𝒢`$ as a single undirected edge. Claim 2. The undirected graph $`𝒢_0`$ is acyclic; furthermore, each component of $`𝒢_0`$ contains at most one edge coming from the self-complement congruences. Proof. Suppose we have a nontrivial cycle in $`𝒢_0`$; as before, we may assume that this cycle does not use an edge more than once. If this cycle includes an edge coming from congruence number $`i`$ for $`i>\overline{m}`$, then, since congruence number $`i`$ produces only one edge of $`𝒢_0`$, the rest of the cycle must come from the other congruences; as before, this implies that congruence number $`i`$ is deducible from the remaining congruences, contradicting nonredundancy. So all of the edges in the cycle come from the first $`\overline{m}`$ congruences; since these congruences form a weak system, we get a contradiction as in the old Claim 2. Now, suppose there are two self-complement edges in the same component. Find a shortest possible path connecting endpoints of two such edges; this path (possibly of length 0) consists of distinct edges from the first $`\overline{m}`$ congruences. Say this path connects $`L`$ to $`R`$, where $`L\{L_i,L_i^c\}`$, $`R\{L_i^{},L_i^{}^c\}`$, and $`i`$ and $`i^{}`$ are distinct numbers greater than $`\overline{m}`$. Then there is a nontrivial cycle in $`𝒢_0`$ from $`L`$ to $`R`$ (the given path) to $`R^c`$ (the $`\tau _i^{}`$-edge) to $`L^c`$ (the complemented and reversed form of the given path) to $`L`$ (the $`\tau _i`$-edge), contradicting the preceding paragraph. $`\mathrm{}`$ Using these two claims, we can now get: Claim 3. Every path of length $`2^{r+1}`$ in the digraph $`𝒢`$ contains either four consecutive edges with the same label $`\tau _i`$ for some $`i>\overline{m}`$ or a pair of consecutive edges with labels $`\sigma _i`$ and $`\sigma _i^1`$, or vice versa, for some $`i\overline{m}`$. Proof. Suppose we have a path of length $`2^{r+1}`$ in $`𝒢`$. Since there are fewer than $`2^r`$ vertices in $`𝒢`$, some vertex, say $`L`$, must be visited at least three times. Let $`p`$ be the subpath from $`L`$ to $`L`$ to $`L`$. By Claim 1, this subpath consists entirely of good edges, so it induces a corresponding path $`p_0`$ in the graph $`𝒢_0`$ which also goes from $`L`$ to $`L`$ to $`L`$. By Claim 2, $`p_0`$ cannot include a nontrivial cycle, so each of its two $`L`$-to-$`L`$ parts must double back on itself. If either doubles back on itself at a $`\sigma _i`$-edge, then $`p`$ has a pair of consecutive edges with labels $`\sigma _i`$ and $`\sigma _i^1`$, or vice versa, so we are done. If neither part of $`p_0`$ doubles back on itself at a $`\sigma _i`$-edge, then they both must double back at a $`\tau _i`$-edge. By Claim 2, there is only one such edge $`e`$ in the component of $`𝒢_0`$ containing $`p_0`$, and there is a unique path $`q`$ in $`𝒢_0`$ from $`L`$ to the nearest endpoint of $`e`$ which does not double back. Hence, $`p_0`$ must consist of $`q`$, an even number of traversals of $`e`$, $`q^{}`$ (the reversal of $`q`$), $`q`$ again, another even number of traversals of $`e`$, and $`q^{}`$ again. If $`q`$ is non-null, then $`p_0`$ doubles back on itself at a $`\sigma _j`$-edge (the last edge of $`q^{}`$ followed by the first edge of $`q`$), so we are done as before; if $`q`$ is null, then $`p_0`$ consists of at least four consecutive occurrences of $`e`$, so $`p`$ contains four consecutive $`\tau _i`$-edges, as desired. $`\mathrm{}`$ The next step is to show that, for any $`gG`$, if $`M_g\mathrm{}`$, then either all of the links in the canonical chain from $`x_0`$ to $`g(x_0)`$ (i.e., the chain read off from the reduced form of $`g`$; note that there might be other chains from $`x_0`$ to $`g(x_0)`$, since $`G`$ is no longer free) are active for the sets $`B_k^n`$, or this chain has fewer than $`2^{r+1}`$ links. The proof of this works as before (with $`2^r`$ replaced by $`2^{r+1}`$), so properties rm (2)–rm (4) hold for the sets $`\widehat{B}_k`$. The construction of the sets $`B_k^{n+1}`$ goes through as before, with two minor changes: one must replace $`2^r`$ with $`2^{r+1}`$ throughout, and one must not assume that there is a unique chain of links connecting two points in the $`G`$-orbit of $`x_0`$. The wording of the definition of the set $`S`$ does not need to be changed, but one must note that it refers to arbitrary chains from $`x_0`$ to $`g(x_0)`$, rather than just the canonical chain. Also, in the second paragraph of the proof of rm (4) for the sets $`B_k^{n+1}`$, one does not necessarily use the reduced form of the element of $`h`$; instead, one just uses the fact that there is some expression of $`h`$ as a product of elements $`f_I`$ and their inverses (one is allowed to use $`f_I^1`$ even if $`I>\overline{m}`$) such that the product has length at most $`2^{r+1}+1`$, and if its length is equal to $`2^{r+1}+1`$, then the rightmost component is $`f_i`$. Everything else goes through as before. This completes the induction. $`\mathrm{}`$ Since the same isometries were used in the preceding two proofs to witness the congruences, Lemma 2.3 now gives: ###### Theorem 4.4 Any consistent system of proper congruences has a solution in $`S^2`$ using nonmeager sets with the property of Baire (and arbitrary isometries). $`\mathrm{}`$ Actually, the proof of Theorem 4.3 goes through without change if the involution $`\zeta `$ is deleted, so that the group $`G^{}`$ is used instead of $`G`$. This gives: ###### Proposition 4.5 Any consistent system of proper congruences has a quasi-solution in $`S^2`$ using nonempty open sets, with rotations witnessing the congruences. $`\mathrm{}`$ However, this result does not lead to a result about sets with the property of Baire, because there is no corresponding result giving solutions using arbitrary sets (unless the system is weak).
warning/0001/nucl-th0001026.html
ar5iv
text
# REFERENCES Solar Proton Burning Process Revisited within a Covariant Model Based on the Bethe-Salpeter Formalism L.P. Kaptari<sup>a,b</sup>, B. Kämpfer$`^a,`$ E. Grosse<sup>a</sup> <sup>a</sup>Forschungszentrum Rossendorf, PF 510119, 01314 Dresden, Germany <sup>b</sup>Bogoliubov Laboratory of Theoretical Physics, JINR Dubna, 141980, Moscow reg., Russia ## Abstract A covariant model based on the Bethe-Salpeter formalism is proposed for investigating the solar proton burning process $`ppDe^+\nu _e`$ and the near-threshold deuteron disintegration via electromagnetic and weak interactions. Results of numerical calculations of the energy dependence of relevant cross sections and the astrophysical low-energy cross section factor $`S_{pp}`$ of the proton burning process are presented. Our results confirm previous canonical values, and the energy dependence of the $`S_{pp}`$ factor is rather close to phenomenological extrapolations commonly adopted in computations of solar nuclear reaction rates. I. Introduction: The proton-proton fusion reaction $`ppDe^+\nu _e`$ plays an important role in understanding processes occurring in the Sun and in investigating the solar structure. This reaction is the first step in the chain of nuclear reactions producing the solar energy and neutrino fluxes. Unfortunately the corresponding cross section is far too low for a direct laboratory study. In this context reliable theoretical calculations and estimates of the reaction rates are of a great importance. Nowadays, the most precise estimates have been obtained within the framework of low-energy potential models . The accuracy of the obtained results is believed to be of the order of few per cents or even better . For instance, the low-energy cross section factor $`S_{pp}`$ that determines the rates for the proton-proton fusion is estimated as $`S_{pp}=4.00(1\pm 0.007)\times 10^{25}`$ MeV b . While this number is considered as a canonical value, recent estimates , based on an effective relativistic field theory, claimed a considerably larger value being inconsistent with helioseismic data . This has triggered a series of recalculations of $`S_{pp}`$ within various approaches (cf. and further references therein), which recover values of $`S_{pp}`$ centered around the canonical one. However, there are some other indications that $`S_{pp}`$ may be $`4.2\times 10^{25}`$ MeV b or even larger. Calculations of the $`pp`$ rates require the evaluation of two main ingredients: (i) the weak-interaction matrix element, and (ii) the overlap of the $`pp`$ and deuteron wave functions. In most relativistic effective field theories a self-consistent treatment of the deuteron bound state is difficult, so one usually adopts the short-range effective theory to avoid an explicit use of the deuteron wave function. Within such models one may achieve good fits of data in the zero energy limit, whereas a computation of matrix elements at finite values of the energy is not yet available. In this paper we calculate the relevant matrix elements within a fully covariant model of the deuteron, based on the Bethe-Salpeter (BS) formalism. Within this model a variety of quantities, including the total and differential cross sections, energy dependences, angular distributions etc. are accessible. In the first part of this paper we introduce our formalism and consider as examples the processes of near-threshold deuteron disintegration by electromagnetic and weak interactions. The second part extends the formalism and is entirely dedicated to the solar $`pp`$ burning processes. Within the relativistic impulse approximation a fully covariant expression for the $`S_{pp}`$ factor is obtained. II. Deuteron disintegration near thresholds: In this section we consider processes of the type $`lDl^{}N_1N_2`$, i.e. the deuteron disintegration by leptons in electromagnetic ($`eDe^{}pn`$) and weak ($`e^+D\nu pp`$) reactions near threshold. There the amplitudes fall down rapidly and the cross sections are governed by the phase space of final products. The threshold energies are $`E_{thr}=2.7375`$ MeV ($`eD`$ reactions) and $`E_{thr}=0.93139`$ MeV (weak interaction), where $`E_{thr}`$ is the total energy of the incident lepton in the deuteron rest frame. We consider the electromagnetic disintegration process within the one-photon exchange approximation (cf. fig. 1a) and the weak positron-nucleon interaction (cf. fig. 1b) within the effective Fermi model. The deuteron vertex $`DNN`$ is treated in a covariant way within the BS formalism. The BS equation has been solved numerically , with realistic one-boson-exchange interactions, and suitably parameterized . The solution describes fairly accurate the main low-energy and static characteristics of the deuteron, such as the deuteron binding energy, and its quadrupole and magnetic moments . The invariant cross section of the process $`D(l,l^{})NN`$, depicted in fig. 1, has the form $`d\sigma ={\displaystyle \frac{(2\pi )^4\delta ^{(4)}\left(k_l+p_Dk_l^{}p_{N_1}p_{N_2}\right)}{2\sqrt{\lambda (s,M_d^2,\mu ^2)}}}\left|_{lDl^{}NN}\right|^2{\displaystyle \frac{d𝐤_𝐥^{}}{(2\pi )^32^{}}}{\displaystyle \frac{d𝐩_\mathrm{𝟏}^{}}{(2\pi )^32E_1^{}}}{\displaystyle \frac{d𝐩_\mathrm{𝟐}^{}}{(2\pi )^32E_2^{}}},`$ (1) where $`\lambda (s,M_d^2,\mu ^2)[s(M_d+\mu )^2][s(M_d\mu )^2]`$ is a kinematical factor with $`s`$ as usual Mandelstam variable, $`M_d`$ and $`\mu `$ are the deuteron and the electron (positron) masses, respectively. The 4-momenta of leptons are denoted as $`k=(,𝐤)`$, while for hadron momenta the notation $`p=(E,𝐩)`$ is used. The prime superscript specifies the momenta of the final particles. The subscript “2” denotes the spectator nucleon, i.e. the neutron in $`eD`$ reactions and the non-interacting proton in $`e^+D`$ processes (see fig. 1). The two nucleons in the final state are treated in the plane wave approximation, i.e. the effects of final state interaction are here disregarded and all calculations are performed within the relativistic impulse approximation. The square of the invariant amplitude $`_{lDl^{}NN}`$ may be cast in the form $`|_{lDl^{}NN}|^2=S^2(Q^2)l_{\mu \mu }W^{\mu \nu },`$ where $`S(Q^2=(k_lk_l^{})^2)`$ is the Feynman propagator of the intermediate exchanged boson (photon or $`W^+`$ boson), and the leptonic ($`l_{\mu \nu }`$) and hadronic ($`W_{\mu \nu }`$) tensors are defined by $`l_{\mu \nu }={\displaystyle \frac{1}{2}}{\displaystyle \underset{\mathrm{spins}}{}}l^{}|\widehat{j}_\mu |ll|\widehat{j}_\nu ^+|l^{}={\displaystyle \frac{1}{2}}\mathrm{Tr}\left[(\widehat{\mathrm{k}}_\mathrm{l}+\mu )\mathrm{\Gamma }_\mu (\widehat{\mathrm{k}}_\mathrm{l}^{}+\mu ^{})\mathrm{\Gamma }_\nu \right],`$ (2) $`W_{\mu \nu }={\displaystyle \frac{1}{3}}{\displaystyle \underset{\mathrm{spins}}{}}N_1^{}N_2^{}|\widehat{J}_\mu |DD|\widehat{J}_\nu ^+|N_1^{}N_2^{},`$ (3) where $`\widehat{j}_\mu `$ and $`\widehat{J}_\mu `$ denote the leptonic and hadronic current operators, $`\mathrm{\Gamma }_\mu `$ is the corresponding vertex function, related to the choice of interacting currents; $`\widehat{k}`$ is used as a short hand notation for $`\gamma ^\mu k_\mu `$. The electron-nucleon electromagnetic vertex in the on-mass-shell form, $`\mathrm{\Gamma }_\mu ^{eN}(Q^2)=\gamma _\mu F_1(Q^2)+i\frac{\sigma _{\mu \alpha }Q^\alpha }{2m}\kappa F_2(Q^2),`$ contains the electromagnetic form factors of the nucleon $`F_{1,2}`$, and is its anomalous magnetic moment $`\kappa `$. In view of the small transferred momenta and low initial energies, considered here, the second term may be safely disregarded, and with high accuracy also the nucleon formfactor $`F_1(Q^2)`$ is replaced by unity. Then the electromagnetic part of the leptonic tensor is computed straightforwardly. For the weak vertices one may employ the Weinberg-Salam theory for weak interactions, however, in the considered energy range the effective Fermi model is appropriate. Thus $`\widehat{j}_\mu (x)=\overline{\psi }_e(x)\gamma _\mu (1\gamma _5)\psi _\nu (x),`$ (4) $`\widehat{J}_\nu (x)^{(\mathrm{\Delta }S=0)}=\mathrm{cos}\theta _c\overline{\psi }_N(x){\displaystyle \frac{(\tau _1i\tau _2)}{\sqrt{2}}}\gamma _\mu (1R\gamma _5)\psi _N(x),`$ (5) and $`S(Q^2)={\displaystyle \frac{G_F}{\sqrt{2}}}`$, and $`R=1.2681`$ is the ratio of axial to vector coupling constants, $`G_F=1.16639\times 10^5\mathrm{GeV}^2`$ is the effective Fermi constant for the weak nucleon decay, $`\theta _c=0.24`$ is the Cabibbo strangeness mixing angle ($`\mathrm{\Delta }S=0`$ means the strangeness conserving part of the current), $`\psi _{l,N}(x)`$ are the corresponding leptonic and nucleonic field operators, $`\tau `$ stands for the isospin Pauli matrices. With virtue of the Mandelstam technique for computing matrix elements within the BS formalism, the deuteron tensor $`W_{\mu \nu }`$ in eq. (3) receives the form $`W_{\mu \nu }={\displaystyle \frac{1}{3}}(p_2^{\mathrm{\hspace{0.17em}2}}m_2^2)\mathrm{Tr}\left[\overline{\mathrm{\Psi }}^{BS}(p_1,p_2)\mathrm{\Gamma }_\mu (\widehat{p}_1^{}+m_1)\mathrm{\Gamma }_\nu \mathrm{\Psi }^{BS}(p_1,p_2)(\widehat{p}_2^{}+m_2)\right],`$ (6) where $`\mathrm{\Psi }^{BS}`$ is the deuteron amplitude as defined in refs. and $`p_1`$ and $`p_2=p_2^{}`$ are the momenta of the deuteron constituents with a mass $`m`$, i.e. the proton ($`m_p`$) or neutron ($`m_n`$) masses. By inspecting eq. (6) one observes that, due to the adopted mechanism in fig. 1, the spectator nucleon is on mass shell, and the first factor in eq. (6) is always zero. Nevertheless, the BS amplitudes are themselves singular when one nucleon is on mass shell with the singularity of each amplitude being of the order $`(p_2^2m_2^2)^1`$. Consequently, the r.h.s. of eq. (6) is finite. In the general case the BS amplitude $`\mathrm{\Psi }^{BS}`$, which is a $`4\times 4`$ matrix, may be projected on a set of basis matrices and such an operation determines eight independent partial amplitudes . However not all these amplitudes equally contribute to the deuteron matrix elements. One may use a basis set of spin-angular matrices which allows to classify the partial amplitudes corresponding to their contributions in a full analogy with the non-relativistic case as $`{}_{}{}^{3}S_{1}^{++()}`$, $`{}_{}{}^{3}D_{1}^{++()}`$, $`{}_{}{}^{3}P_{1}^{+(+)}`$, $`{}_{}{}^{1}P_{1}^{+(+)}`$ waves in the spectroscopical notation (the superscripts denote the $`\rho `$ spin of each constituent, cf. for details). Then the $`{}_{}{}^{3}S_{1}^{++}`$ and $`{}_{}{}^{3}D_{1}^{++}`$ waves turn out to be a relativistic generalization of the non-relativistic $`S`$ and $`D`$ waves in the deuteron. In the considered reactions the contribution of $`D`$ waves vanishes near the threshold (as do the other amplitudes with at least one negative $`\rho `$ spin), leaving us with contributions only from the $`{}_{}{}^{3}S_{1}^{++}`$ waves. A direct evaluation of traces in eqs. (2, 6) results in analytical, covariant expressions for the cross section. For instance, the electromagnetic tensors read $`l_{\mu \nu }W^{\mu \nu }=`$ $`16M_d(2\pi )^3\mathrm{\Phi }^2(p_0,|𝐩_2|)\times `$ (8) $`\left[(k_ep_1^{})(k^{}p_1)\mu ^2(p_1p_1^{})(k_e^{}k_e)m_p^2+(k_e^{}p_1^{})(k_ep_1)+2\mu ^2m_p^2\right],`$ where $`\mathrm{\Phi }(p_0,|𝐩_2|)`$ is the covariant generalization of the deuteron wave function within the BS formalism, which has the following form in the deuteron rest frame $$\mathrm{\Phi }(|𝐩_2|)=\frac{G_S(p_0,|𝐩_2|)}{(2\pi )^2\sqrt{2M_d}(M_d2E_p)^2},p_0=M_d/2E_p,E_p=\sqrt{m^2+𝐩_2^2},$$ (9) where $`G_S(p_0,|𝐩_2|)`$ is the BS partial vertex corresponding to the $`{}_{}{}^{3}S_{1}^{++}`$ amplitude and $`m`$ is an average value of the nucleon mass. Inserting eq. (8) into eq. (1) yields $`d\sigma ^{eD}={\displaystyle \frac{m_n}{E_2}}\mathrm{\Phi }^2(|𝐩_2|)d𝐩_2\delta (p_1^{{}_{}{}^{}\mathrm{\hspace{0.17em}2}}m_p^2)\times `$ (10) $`\left\{{\displaystyle \frac{4\alpha ^2}{Q^4}}{\displaystyle \frac{4M_d}{m_p\sqrt{\lambda (s,M_d^2,\mu ^2)}}}\left((k_ep_1^{})(k^{}p_1)\mu ^2(p_1p_1^{})(k_e^{}k_e)m_p^2+(k_e^{}p_1^{})(k_ep_1)+2\mu ^2m_p^2\right){\displaystyle \frac{d𝐤_e^{}}{2^{}}}\right\},`$ (11) where the expression enclosed in the curly brackets together with the $`\delta `$ function is merely the cross section $`d\sigma ^{ep}`$ of an electron scattering off a moving proton with the momentum $`p_1`$. Hence, for the electromagnetic $`eD`$ disintegration processes the cross section may be written as $$\sigma ^{eD}=𝑑𝐩_2\frac{m_n}{E_2}\mathrm{\Phi }^2(|𝐩_2|)𝑑\sigma ^{ep}(k,p_1).$$ (12) By processing in an analogous way with the weak disintegration diagram displayed in fig. 1b, one obtains $$\sigma _w^{e^+D}=𝑑𝐩_2\frac{m_n}{E_2}\mathrm{\Phi }^2(|𝐩_2|)𝑑\sigma _w^{en}(k,p_1),$$ (13) where, for convenience, the “elementary” cross section $`d\sigma _w^{en}(k,p_1)`$ for a hypothetical $`e^+n\nu p`$ reaction has been introduced by $`d\sigma _w^{en}(k,p_1)`$ $`=`$ $`{\displaystyle \frac{G_F^2}{(2\pi )^2}}{\displaystyle \frac{4M_d\delta (p_1^{{}_{}{}^{}\mathrm{\hspace{0.17em}2}}m_p^2)}{m_p\sqrt{\lambda (s,M_d^2,\mu ^2)}}}\left\{(R^2+1)\right)[(k_ep_1)(k_\nu ^{}p_1^{})+(k_ep_1^{})(k_\nu ^{}p_1)]`$ (15) $`+(R^21))m_pm_n(k_ek_\nu ^{})2R[(k_ep_1^{})(k_\nu ^{}p_1)(k_ep_1)(k_\nu ^{}p_1^{})]\}{\displaystyle \frac{d𝐤_\nu ^{}}{^{}}}.`$ In computing eq. (13) the angular integration on $`𝐩_2`$ should be performed in only one-half of the hemisphere due to the two identical particles in the final state. In fig. 2 results of the numerical calculation of the total cross section as a function of the initial energy of the electron (positron) in the laboratory frame are displayed. Above the electromagnetic disintegration threshold there is probably no way to experimentally separate the weak processes from the electromagnetic ones. Below the electromagnetic and above the weak thresholds, the cross section of positron weak disintegration of the deuteron seems far too low to be measured under laboratory conditions. So one may conclude that at present an experimental investigation of the matrix element, relevant for the solar $`pp`$ fusion, is still impossible also when using the cross channels. (This is in contrast to other reactions, cf. .) Only reliable theoretical calculations may be applied to estimate the nuclear reaction rates and to interpret the corresponding experimental neutrino data. III. Solar proton burning process: Let us now consider the process of the two-proton fusion $`p_1p_2\nu e^+D`$ at low relative energies. While this reaction looks rather similar to the processes in the previous section, here appears a peculiarity. Due to the relation $`2m_p>M_d+\mu `$ (with $`m_p=938.27231`$ MeV, $`M_d=1875.61339`$ MeV, deuteron binding energy $`\epsilon _D=2.22455`$ MeV, and $`\mu =0.510999`$ MeV) there is no energy threshold for the $`pp`$ fusion and, in principle, the fusion process may occur at arbitrarily small relative energies of the two protons. In this case, in the cross section (1) two factors will play the major role: (i) the two-proton flux ($`\sqrt{\lambda (s,m_p^2,m_p^2)}`$) approaches rather rapidly zero and the computation of the cross section becomes problematic, (ii) a strong Coloumb repulsion between protons makes the amplitude $``$ sharply decreasing and compensates singularities appearing from the vanishing flux. In order to avoid uncertainties connected to these circumstances one usually separates the contribution of Coulomb repulsion by factorizing the Coulomb part of the two-proton final state, which at low relative energies may be taken as $`\psi _{pp}={\displaystyle \frac{\sqrt{2\pi \eta }}{\mathrm{e}^{\pi \eta }1}}\sqrt{2\pi \eta }\mathrm{e}^{\pi \eta },`$ where $`\eta =\alpha /v`$ is the Sommerfeld penetration parameter, $`v`$ denotes the relative velocity of the protons, and $`T_{pp}=m_pv^2/4`$ is the corresponding relative kinetic energy; the exponent is known as the Gamov penetration factor . Then one introduces the notion of the $`S_{pp}`$ factor, which is determined only by the nuclear part of the interaction and is connected with the cross section by $`\sigma {\displaystyle \frac{S_{pp}(T_{pp})}{T_{pp}}}\mathrm{e}^{2\pi \eta }.`$ Then the rate of $`pp`$ fusion reactions can be written in the form $`\sigma v_{pp}=1.3005\times 10^{18}\left(2/T_6^2\right)^{1/3}fS_{pp}(T_{pp})\mathrm{e}^\tau \mathrm{cm}^3\mathrm{s}^1;`$ here $`T_6`$ is the temperature in units of $`10^6`$ K, $`\tau 1540`$ dominates the temperature dependence of the reaction rate, $`f`$ is a screening factor calculated first by Salpeter , and $`S_{pp}`$ is in MeV b. The $`S_{pp}`$ factor is needed at the most probable interaction energy $`T_{pp}^{(0)}`$, which is in a range of 5 keV $`\mathrm{}`$ 30 keV. However, in most analyzes in the literature, the value of $`S_{pp}`$ is quoted at vanishing energy, obtained , e.g. within the low-energy effective theory . Hence, for the sake of extrapolation to the finite value of $`T_{pp}^{(0)}`$, a good estimate of the associated derivatives is also required. We now apply the same formalism as in the previous section, i.e. the relativistic impulse approximation, to estimate the low-energy behavior of $`S_{pp}`$. The effect of initial state interaction is accounted for only in the Coulomb part of the interaction by adopting the corresponding part of the above two-proton wave function $`\psi _{pp}`$. A treatment of the nuclear part of the proton wave function as plane waves means that in the matrix element all transitions are allowed. Nevertheless, at low energies only the lowest partial waves contribute to the proton wave function and one may expect that the total cross section will be determined by the axial part of the current operator (5) and correspondingly by the super-allowed $`0^+0^+`$ Gamov-Teller transition. In the limit $`T_{pp}0`$, apart from the phase space volume, the value of the deuteron wave function $`\psi (p_0,|𝐩|)`$ at vanishing values of its arguments becomes important. As can be seen from eq. (9), at $`|𝐩|0,\mathrm{\hspace{0.17em}2}E_p2m`$ the wave function $`\psi (p_0,|𝐩|)`$ displays a sharp maximum. Our solution of the BS equation has been obtained by solving a system of coupled integral equations with Gaussian meshes with a minimum value of $`|𝐩|=0.310848493733`$ MeV, so that an extrapolation to $`|𝐩|=0`$ needs to be employed. This procedure causes some numerical uncertainty which we estimate to be about 10% in the final results. By observing that in the center of mass of colliding protons $`2\sqrt{\lambda (s,m_p^2,m_p^2)}=4E_1E_2v=sv`$ and adopting the Gamow penetration factor, eq. (1) leads to $`S_{pp}(T_{pp})={\displaystyle \frac{G_F^2}{8s}}{\displaystyle \frac{\pi m_p}{(2\pi )^5}}{\displaystyle \frac{d𝐏_𝐝}{E_d}\frac{d𝐤_\nu ^{}}{_\nu ^{}}\left|_{ppe^+\nu D}\right|^2\delta (k_{e^+}^2\mu ^2)},`$ (16) where, due to the antisymmetrization of the initial two-proton state, the matrix element $`_{ppe^+\nu D}`$ is now determined by two Feynman diagrams, as depicted in fig. 3, with a relative minus sign. Correspondingly, the hadron tensor $`W_{\mu \nu }`$ consists of a direct and an interference term, $`W_{\mu \nu }^{(dir)}`$ $`=`$ $`(p_2^2m_p^2)\mathrm{Tr}\left[\overline{\mathrm{\Psi }}(p_1,p_n)\mathrm{\Gamma }_\mu (\widehat{p}_2+m_p)\mathrm{\Gamma }_\nu \mathrm{\Psi }(p_1,p_n)(\widehat{p}_2+m_p)\right]`$ (18) $`+(p_1^2m_p^2)\mathrm{Tr}\left[\mathrm{\Psi }(p_1,p_n)\mathrm{\Gamma }_\nu (\widehat{p}_1m_p)\mathrm{\Gamma }_\mu \overline{\mathrm{\Psi }}(p_2,p_n)(\widehat{p}_1m_p)\right],`$ $`W_{\mu \nu }^{(int)}`$ $`=`$ $`(p_2^2m_p^2)(p_1^2m_p^2)\mathrm{Tr}\left[\overline{\mathrm{\Psi }}(p_2,p_n)\mathrm{\Gamma }_\nu \mathrm{\Psi }(p_1,p_n)\mathrm{\Gamma }_\mu +\overline{\mathrm{\Psi }}(p_1,p_n)\mathrm{\Gamma }_\mu \mathrm{\Psi }(p_2,p_n)\mathrm{\Gamma }_\nu \right],`$ (19) where $`\mathrm{\Gamma }_\mu `$ is the vertex corresponding to the effective current operator (5). The direct part of the hadron tensor has exactly the same form as for the reactions $`e^+D\nu pp`$ (cf. eq. (6)), resulting in an “elementary” cross section like that in eq. (15). The interference term is different and much more cumbersome, and for the sake of brevity we do not display it here. In fig. 4 the energy dependence of $`S_{pp}`$ computed by eq. (16) is displayed. The $`S_{pp}`$ factor, starting from $`T_{pp}+0`$ up to solar energies $`T_{pp}^0`$, displays a rather weak energy dependence. At very low kinetic energies the value of $`S_{pp}`$ remains fairly constant, having a value of $`3.960\times 10^{25}`$ MeV b in agreement with previous results . At larger energies, however, the $`S_{pp}`$ factor increases rather rapidly. The two curves in fig. 4 correspond to two different methods of numerical calculations. In one case the deuteron wave function is replaced by its value at zero transferred momenta. To some extent this corresponds to calculations within low-energy effective potential models. The full line reflects results of calculations without such an approximation. Only at relatively high values of the kinetic energy, $`T_{pp}>20`$ KeV, the effect of the deuteron wave function becomes sizeable. Within low-energy effective theories the associated derivatives of the $`S_{pp}`$ factor play an important role. The best estimates of the first derivative seem to be those quoted in ref., $`S_{pp}^{}=4.48\times 10^{24}`$ b, and the logarithmic one $`S_{pp}^{}/S=11.2`$ MeV<sup>-1</sup>. In figs. 5 and 6 we display our estimates for the energy dependence of these derivatives. The derivatives are computed at each value of $`T_{pp}`$ by interpolating $`S_{pp}`$ with a quadratic interpolation formula. From these pictures it is seen that within our approach both the derivative and the logarithmic derivative are below the estimates in by $`1520\%`$. However, at energies around $`T_{pp}^{(0)}`$ our results for the $`S_{pp}`$ factor and the results of an interpolation with parameters from are rather close to each other. IV. Summary: We present a covariant model, based on the Bethe-Salpeter formalism, to investigate the energy dependence of the matrix elements relevant to the solar proton-proton burning process. Explicit covariant expressions for the cross section and the low-energy cross section factor $`S_{pp}`$ are presented. Our numerical estimates of $`S_{pp}`$ and its derivatives confirm previous results. Acknowledgments: We thank H.-W. Barz, R. Wünsch and A.I. Titov for useful discussions. L.P.K. would like to thank for the warm hospitality in the Research Center Rossendorf. This work has been partially supported by the Heisenberg-Landau JINR-FRG collaboration project, and by BMBF grant 06DR829/1 and WTZ RUS-678-98.
warning/0001/cond-mat0001139.html
ar5iv
text
# Lattice parameters from direct-space images at two tilts ## I Introduction If 10 nm crystals presented themselves to our perceptions as 10 cm hand specimens, new rules for direct-space crystallography might have emerged. For example, if by tilting the specimen one could locate a set of 4-fold cross-fringes, from which a tilt (at $`45^{}`$ to those fringes) by $`35.3^{}`$ just brings one to a new set of fringes $`15.5`$ percent larger in spacing, then the hand specimen is likely face-centered-cubic. It’s reciprocal lattice definitely includes a body-centered-cubic array like that characteristic of face centered-cubic-crystals. Here we discuss the geometry of crystals from the perspective of lattice imaging in direct space. The required instrumentation consists of a TEM able to deliver phase or Z contrast lattice images of desired periodicities (e.g. spacings down to half the unit cell side for cubic crystals), and a specimen stage with adequate tilt (e.g. two-axes with a combined tilt range of $`\pm 18^{}`$). For crystals with lattice spacings of $`0.25`$nm and larger, many analytical TEM’s will work, while a high resolution microscope with continuous contrast transfer to spatial frequencies beyond $`1/(0.2`$nm$`)`$ can do this for most crystals. We demonstrate the process experimentally by determining the lattice parameters of a tungsten carbide nano-crystal using a Philips EM430ST TEM. Appropriately orienting the crystal, so as to reveal its three-dimensional structure in images, is a key part of the experimental design and will be discussed in detail. In the process, strategies for supporting on-line electron-crystallography, for three-dimensional lattice-correlation darkfield studies of nanocrystalline and paracrystalline materials, and for stereo-diffraction analyses, are suggested as well. ## II Calculations ### II.1 Overview of a “tilt protocol” in action For the stereo lattice-imaging strategy discussed here, low Miller index (hence large) lattice spacings are both easier to see, and more diagnostic of the lattice. Orientation changes directed toward the detection of such spacings are needed. In this section, tilt protocols optimized for getting 3D data from one abundant class of lattice types (namely cubic crystals) are surveyed. We begin with a list of candidate lattice types (e.g. f.c.c. or h.c.p.) based on prior compositional, diffraction, or imaging data. Three non-coplanar reciprocal lattice vectors seen along 2 different zone axes are sufficient for inferring a subset of the 3D reciprocal lattice of a single crystal. Often these are adequate to infer the whole reciprocal lattice. Hence the goal of our experimental design is to look for at least 3 non-coplanar lattice spatial frequencies, in two or more images. We prefer images with “aberration limits” $`r_a`$ smaller than the analyzed spacings, to lessen chances of missing other comparable (or larger) spacings in the exit-surface wavefield. So as to tilt from one zone to another, the crystal must also be oriented so that the desired beam orientations are accessible within the tilt limits of the microscope. To illustrate, we exploit the fact (considered more fully below) that all cubic crystals will provide data on their three dimensional lattice parameters if imaged down selected zones separated by $`35.3^{}`$, provided that spacings at least as large as half the cell side $`a`$ are reported in the images. With the images discussed here we expect to “cast a net” for 3D data on any cubic crystals whose cell side $`a`$ is larger than $`2\times 0.19nm=0.38nm`$. More than 85% of the cubic close packed crystals and nearly 40% of the elemental b.c.c. crystals tabulated in Wyckoff Wyckoff (1982), for example, meet this requirement, as of course would most cubic crystals with asymmetric units comprised of more than one atom. Although $`35.3^{}`$ is too far for the eucentric tilt axis in our microscope, combining two tilts gives us a range of $`35.6^{}`$. Two images therefore can be taken at orientations $`35.3^{}`$ apart, namely at ($`\vartheta _1=15.0^{}`$, $`\vartheta _2=9.7^{}`$) and ($`\vartheta _1=15.0^{}`$, $`\vartheta _2=9.7^{}`$), respectively, where $`\vartheta _1`$ and $`\vartheta _2`$ are goniometer readings on our double tilt holder. This yields an “effective” tilt axis that runs perpendicular to the electron beam. Its azimuth is $`123.5^{}`$ in the $`xy`$ plane of our images. The coordinate system used will be discussed in more detail later. For the special case of fcc crystals, the zones of nterest are \[$`001`$\] and \[$`112`$\]. Since the ($`2\overline{2}0`$) lattice planes are parallel to both desired zones, the tilt must be along these planes. That is, ($`200`$) fringes seen down a 4-fold symmetric \[$`001`$\] zone must make an angle approaching $`45^{}`$ with respect to the effective tilt axis. Nearly a third of the randomly-oriented crystals showing \[$`001`$\]-zone fringes will be sufficiently close Qin (2000). The experimental results were unambiguous. We found many four-fold symmetric images having spacings consistent with WC<sub>1-x</sub>. This has an f.c.c. lattice with $`a=0.4248`$nm Krainer and Robitsch (1967); JCP (1988). When such a zone was found with fringes making $`45^{}`$ to the effective tilt axis in the first image, a new spacing was seen in the second image making a 3D lattice parameter measurement possible. The experiment is illustrated in Figure 3. ### II.2 Experimental designs Here we seek three non-coplanar periodicities from two images, although the analysis also works if they are discovered singly in three images. Given Miller indices $`(h_1k_1l_1)`$ and $`(h_2k_2l_2)`$ for any two periodicities (i.e. vectors $`𝐠_1`$ and $`𝐠_2`$, respectively, in the reciprocal lattice) of any crystal, first find zone indices $`[u_Av_Aw_A]`$ of the beam direction $`𝐫_A`$ $`𝐠_1\times 𝐠_2`$ needed to view both spacings in one image. The axis for the smallest tilt that will make the beam orthogonal to a third periodicity with indices $`(h_3k_3l_3)`$, and reciprocal lattice vector $`𝐠_3`$, may then be defined by the vector $`𝐯_t𝐠_3\times 𝐫_A`$. Lastly, zone indices $`[u_Bv_Bw_B]`$ for the beam after the specimen has been tilted around this axis so as to image the third periodicity, may be obtained from the expression $`𝐫_B𝐯_t\times 𝐠_3`$. Note here that we treat the Bragg angle for electrons as small (i.e. less than one degree). Thus the actual tilt required will be a fraction of a degree less. Although these cross product calculations can be done by first converting for example to “c-axis” cartesian coordinates Fraundorf (1981a), the simplest determination of needed parameters is perhaps done using the metric matrix $`G`$ of a prospective lattice Boisen and Gibbs (1985): $$G\left[\begin{array}{ccc}𝐚𝐚& 𝐚𝐛& 𝐚𝐜\\ 𝐛𝐚& 𝐛𝐛& 𝐛𝐜\\ 𝐜𝐚& 𝐜𝐛& 𝐜𝐜\end{array}\right]\text{.}$$ (1) If we denote row vectors formed from Miller (or lattice) indices as $`ijk|`$, and column vectors as $`|ijk`$, then the zone $`A`$ indices obey $`𝐠_1𝐫_A=h_1k_1l_1|u_Av_Aw_A=0`$ and $`𝐠_2𝐫_A=h_2k_2l_2|u_Av_Aw_A=0`$. From these two equations, $`[u_Av_Aw_A]`$ follows except for a multiplicative constant which is not important. Similarly, the (possibly irrational) Miller indices of the tilt axis $`(h_t,k_t,l_t)`$ may be determined from $`𝐠_t𝐫_A=h_tk_tl_t|u_Av_Aw_A=0`$ and $`𝐯_t𝐠_3=0=h_tk_tl_t|G^1|h_3k_3l_3`$ Spence and Zuo (1992). Only in this fourth equality does $`G`$ affect the calculation, and for cubic crystals it then simply offers a multiplying constant. Finally, the zone $`B`$ indices follow (to within a factor) simply from $`𝐠_t𝐫_B=h_tk_tl_t|u_Bv_Bw_B=0`$ and $`𝐠_3𝐫_B=h_3k_3l_3|u_Bv_Bw_B=0`$. Two parameters which determine the attractiveness and feasibility of a given experiment are the spatial resolution, and range of specimen tilts, that the microscope is able to provide. For a given lattice type, it is useful to: (i) go through the list of all pairs of periodicities calculating the tilt between the zone associated with that pair and any third spacing of possible interest, and (ii) rank the findings according to the minimum-spacing that must be resolved, and the range-of-tilt that the specimen undergoes. This has been done for face-centered, body-centered, and simple-cubic lattices, and the results illustrated in Figure 1, with examples for fcc and bcc spacings no less than half of the cell side illustrated in Figure 4. Note that this calculation needs to be done only once for each unit cell shape. Factors like the multiplicity of a given zone type might also figure into the design of experiments with randomly-oriented crystals, although we have not considered them here. High tilts can be used to lower measured spacing uncertainties, in directions perpendicular to the electron microscope specimen plane Fraundorf (1987). Hence the protocols of interest for a given experiment may be those which approach goniometer tilt limits, or at least the limits of specimen tiltability. Concerning the resolution limit to use, we suggest the spacing associated with the end of the first transfer function passband in the micrograph of interest, sometimes inferrable from regions in the image showing disordered material. Even in this case, possible thickness and misorientation effects warrant caution Spence (1988); Malm and O’Keefe (1997). One may of course explore spatial frequencies in the specimen up to the microscope information limit. However, spherical aberration zeros in the transfer function introduce the possibility that the microscope will suppress some spatial frequencies present in the subspecimen electron wavefield. To lessen this problem, HREM images taken at different focus settings could be compared, if the foci were chosen so that spatial frequencies lost at one defocus are likely recorded at another. This would allow measurement of three non-coplanar reciprocal lattice vectors, without missing any whose (reciprocal) length is shorter than the longest among those three. Lastly, of course, the protocol chosen may depend also on the specimen. For an image field containing hundreds of non-overlapping but randomly-oriented nano-crystals, only two micrographs could allow one to measure the three-dimensional lattice parameters of all cubic crystal types present with cell sides $`a`$ greater than $`2r_a`$. On the other hand, for a single crystal specimen of unknown structure, both a great deal of tilt range, and considerable trial and error tilting (or guess work based on lattice models) might be required before a single set of three indexable non-coplanar spacings is found. ### II.3 Inferring 3D reciprocal lattice vectors from micrographs Consider a specimen stage with two orthogonal tilt axes (and associated rotation matrices) $`T_1`$ and $`T_2`$, both perpendicular to the electron beam (the second only so when the first axis is set at zero tilt). When the specimen is untilted ($`\vartheta _1=0^{}`$, $`\vartheta _2=0^{}`$), vectors in the reciprocal lattice of the specimen may be described, in coordinates referenced to the microscope, as column vectors $`|𝐠`$. When this reciprocal lattice vector is tilted to intersect the Ewald sphere by some double tilt in the sequence $`T_2(\vartheta _2)`$ then $`T_1(\vartheta _1)`$, it’s presence may be inferred from diffraction patterns or micrograph power spectra. In our fixed coordinate system, $`|g`$ has become $`|g_m`$, where the $`m`$ means that $`|g_m`$ is associated with the lattice periodicity $`d_m=1/g_m`$ recorded on a micrograph. Using matrix notation, we might then write: $$|𝐠_m=T_1(\vartheta _1)T_2(\vartheta _2)|𝐠$$ (2) Components of $`g_m`$ may be determined from the polar coordinates ($`g`$, $`\phi `$) of a spot in the power spectrum of a recorded image, following: $$|𝐠_m\left(\begin{array}{c}g_{mx}\\ g_{my}\\ g_{mz}\end{array}\right)=g\left(\begin{array}{c}\mathrm{cos}(\phi )\\ \mathrm{sin}(\phi )\\ 0\end{array}\right)$$ (3) where $`g`$ is the length of the diffraction vector (e.g. in reciprocal nm) and $`\phi `$ is its azimuth corrected for lens rotation. Hence we can calculate the “untilted-coordinates” $`|g`$, of reciprocal lattice objects at $`|g_m`$ inferred experimentally from micrographs, using $$|g=T_2^1(\vartheta _2)T_1^1(\vartheta _1)|𝐠_m=A(\vartheta _1,\vartheta _2)|𝐠_m\text{,}$$ (4) where we’ve defined: $$A(\vartheta _1,\vartheta _2)T_2^1(\vartheta _2)T_1^1(\vartheta _1)\text{.}$$ (5) The resulting $`xyz`$ coordinates of reciprocal lattice features $`g`$, associated with the crystal while in the untilted goniometer specimen orientation, but referenceable from micrographs taken at any orientation, provide the language we use for speaking of our measurements in three dimensions. ### II.4 Calculating lattice parameters Given 3D cartesian coordinates of “points” in the reciprocal lattice of a crystal, we are in much the same situation as if we had diffraction patterns of the crystal from two directions containing three (or more) non-coplanar spots. Hence methods for stereo-analysis of diffraction data Fraundorf (1981a, b) can be used at this point. We summarize in this context briefly. Given measured reciprocal lattice vector coordinates, a natural next step is to infer a basis triplet for the crystal’s reciprocal lattice. Three alternate paths to this basis triplet might be referred to as “matching”, “building” Fraundorf (1981b), and “presumed” Fraundorf (1981a). Given an experimental basis triplet from any of these sources, lattice parameters ($`a`$, $`b`$, $`c`$, $`\alpha `$, $`\beta `$, $`\gamma `$), goniometer settings for other zones, and many other things follow simply from the oriented triplet matrix defined below: $$𝐖\left[\begin{array}{ccc}a_x& a_y& a_z\\ b_x& b_y& b_z\\ c_x& c_y& c_z\end{array}\right]=\left[\begin{array}{ccc}a_x^{}& b_x^{}& c_x^{}\\ a_y^{}& b_y^{}& c_y^{}\\ a_z^{}& b_z^{}& c_z^{}\end{array}\right]^1$$ (6) Given $`W`$, for example, cartesian coordinates in the microscope for any direct lattice vector with indices $`[uvw]`$ follow from $`|r=W|uvw`$, while cartesian coordinates for the reciprocal lattice vector with indices $`(hkl)`$ may be predicted from $`g|=hkl|W^1`$. These rules of course include instructions for calculating basis vectors of the lattice, such as $`𝐚[100]`$, and reciprocal lattice, such as $`𝐛^{}(010)`$, and the angles between. Moreover, the oriented cartesian triplet $`W`$ is simply related to the metric matrix for the lattice in equation 1 by $`G=WW^T`$. From $`G`$, of course, all the familiar orientation-independent properties of the lattice follow Spence and Zuo (1992), including cell volume $`V_{cell}=\sqrt{\left|G\right|}`$, Miller/lattice vector dot products $`g_{hkl}|r_{uvw}=hkl|uvw`$, reciprocal lattice vector and interplanar spacing magnitudes $`g_{hkl}^2=1/d_{hkl}^2=hkl|G^1|hkl`$, lattice vector magnitudes $`r_{uvw}^2=uvw|G|uvw`$, reciprocal lattice dot products $`g_1|g_2=h_1k_1l_1|G^1|h_2k_2l_2`$, interspot angles $`\theta _{12}=\mathrm{cos}^1\left[g_1|g_2/(g_1g_2)\right]`$, etc. Even before a basis triplet is selected, indexing of observed reciprocal lattice vectors $`g`$ can be attempted by matching spacings and interspot angles to candidate lattices. Because of the uniqueness of non-coplanar triplets in three dimensions (providing at least a significant subset of the whole reciprocal lattice), the matches are very discriminating (even for low-symmetry lattices) relative to similar analyses from 2D data, i.e. from a only a single image or diffraction pattern. After a basis triplet is selected, the indices $`(hkl)`$ of any observed spot $`g|`$ in a diffraction pattern follows from $`hkl|=g|W`$. Similarly, indices $`[uvw]`$ of any observed lattice periodicity $`|r`$ in an image follow from $`|uvw=W^1|r`$, once a basis triplet is in hand. Given the triplet one can similarly calculate goniometer settings to align the beam with any other crystallographic zone of interest. ## III The Experimental Setup ### III.1 Instruments The Philips EM430ST TEM used is housed in a triple-bi/story building designed for low vibration, and provides continuous contrast transfer to $`1/(0.19nm)`$ at Scherzer defocus. It is equipped with a $`\pm 15^{}`$ side-entry goniometer specimen stage. A Gatan double tilt holder enables $`\pm 10^{}`$ tilt around an orthogonal tilt axis. The largest orientation difference which can be achieved using this double tilt holder in the microscope is therefore $`35.6^{}`$ Selby (1972). ### III.2 Determining the angle of effective tilt projected onto an image In order to establish the spatial relationship between reciprocal lattice vectors inferred from images taken at different specimen tilts, the direction of the tilt axis relative to those images must be known. The tilt axis direction is defined via the right hand rule, as orthogonal to the relative motion of parts of the specimen as the goniometer reading is increased. In a single tilt, the axis is perpendicular to the electron beam and parallel to the micrographs. This is true also of the effective tilt axis in a double tilt, provided the two specimen orientations are symmetric about the zero tilt position. We limit our discussion of double tilts to this case. We determined the projection of both tilt axes of a Gatan double tilt holder onto 700K HREM images by examining Kikuchi line shifts during tilt in the 1200mm diffraction pattern of single crystal silicon, and then correcting for the rotation between that diffraction pattern and the image Qin (2000). To be specific, with a micrograph placed in front of the operator with emulsion side up as in the microscope, with zero azimuth defined as a vector from left to right, and with counterclockwise defined as the direction of increasing azimuth, the projection of the $`T_1`$ axis on 1200mm camera-length diffraction patterns in our microscope is along $`114.0^{}`$. The rotation angle between electron diffraction patterns at the camera length of 1200mm and 700K HREM images is $`42.9^{}`$. Therefore the direction of the projection of $`T_1`$ on 700K HREM images is along $`156.9^{}`$. The direction of the projection of the second tilt axis, $`T_2`$, on 700K HREM images is along $`113.1^{}`$. ### III.3 A reference coordinate system We then consider a coordinate system fixed to the microscope, for measuring reciprocal lattice vectors from the power spectra of 700K HREM images. The $`y`$ and $`z`$ directions are defined to be along $`T_1`$ and the electron beam direction, respectively, as shown in Figure 2. The projection of these tilts on the power spectrum of a 700K HREM image is shown in the 2nd inset of Figure 3. Azimuths in the remainder of this paper are all measured in the $`xy`$ plane of this coordinate system, with the $`x`$ or $`T_2`$ direction set to zero. Because the $`T_1`$ direction is defined in our coordinate system as the negative $`y`$-direction, azimuthal angles are measured on micrographs from a direction $`90^{}`$ clockwise from the $`T_1`$ direction, when the emulsion side is up. ### III.4 Double tilting The specimen was first tilted about $`T_2`$ to $`\vartheta _2=9.7^{}`$ while $`\vartheta _1`$ remained at $`0^{}`$, made eucentric, then tilted about $`T_1`$ to $`\vartheta _1=15.0^{}`$. The first HREM image was taken at this specimen orientation of ($`\vartheta _1=15^{}`$,$`\vartheta _2=9.7^{}`$). A similar sequence was applied to take a second HREM image at the second specimen orientation of ($`\vartheta _1=15^{}`$, $`\vartheta _2=9.7^{}`$). The process can be modeled with a simple matrix calculation Liu et al. (1989); Tambuyser (1985). Because of the importance of repeatable quantitative tilts, effects of “mechanical hysterisis” were minimized by inferring all relative changes in tilt from goniometer readings taken with a common direction of goniometer rotation. The rotation sequences were “initialized” by first tilting past the starting line, and then returning to it in the direction of subsequent motion. Nonetheless both the precision of angle measurement, and our inability to observe lattice fringes during rotation, were shortcomings that microscopes designed to apply these strategies routinely must address. ### III.5 Specimen preparation The tungsten carbide thin film was deposited by PECVD on glass substrates by introducing a gaseous mixture of tungsten hexacarbonyl and hydrogen into a RF-induced plasma reactor at a substrate temperature of $`330^{}`$C Qin et al. (1998). The specimen was disk-cut, abraded from the glass substrate side and dimpled by a Gatan Model 601 Disk Cutter, a South Bay Technology Model 900 Grinder and a Gatan Model 656 Precision Dimpler, respectively. The specimen was then argon ion-milled by a Gatan DuoMill for about 5 hours to perforation prior to the TEM study, at an incidence angle of $`3^{}`$. ## IV Experimental Results ### IV.1 Diffraction from a known to check column geometry Calibration of these algorithms with geometry in our microscope was first done with diffraction data from a Si crystal. Diffraction patterns of $`100`$ Si along the \[$`1\overline{1}\overline{6}`$\] and \[$`1\overline{1}6`$\] zone axes were obtained by tilting about $`T_1`$ and $`T_2`$. The lattice parameters determined are ($`a=0.383`$nm, $`b=0.387`$nm, $`c=0.386`$nm, $`\alpha =60.0^{}`$, $`\beta =119.6^{}`$, $`\gamma =119.1^{}`$). This set of chosen basis defines the rhombohedral primitive cell of the Si f.c.c lattice. Compared with the literature values of Si ($`a=0.384`$nm, $`b=0.384`$nm, $`c=0.384`$nm, $`\alpha =60^{}`$, $`\beta =120^{}`$, $`\gamma =120^{}`$), the angular disagreements are less than $`1^{}`$ and spatial disagreements are less than $`1\%`$. These uncertainties are comparable to those obtained by other techniques of submicron crystal analysis Liu (1990); Liu et al. (1992, 1989); Tambuyser (1985); Fraundorf (1981b). ### IV.2 Nanocrystal images to infer lattice parameters of an unknown The micrographs in Figure 3 show a nanocrystal in a film rich in tungsten carbide, at the orientations of ($`\vartheta _1=15^{}`$, $`\vartheta _2=9.7^{}`$) and ($`\vartheta _1=15^{}`$, $`\vartheta _2=9.7^{}`$) respectively. The coordinates of periodicities in micrograph power spectra, as well as in the common reference coordinate system, are listed in Table 1, in much the same format as is diffraction data for stereo analysis Fraundorf (1981b). #### IV.2.1 Matching the lattice: The lattice spacings and inter-spot angles of periodicities in image power spectra were used to look for consistent indexing alternatives from a set of 36 tungsten carbide and oxide candidate lattices including WC<sub>1-x</sub>. When an angular tolerance of $`2^{}`$ and a spatial tolerance of $`2\%`$ are imposed, only WC<sub>1-x</sub> provides a consistent indexing alternative. As summarized in Table 2, the Miller indices of the three observed spots then become ($`200`$), ($`020`$) and ($`11\overline{1}`$). The images of Figure 3 thus represent WC<sub>1-x</sub> and zones, respectively. The azimuth of the reciprocal lattice vector ($`2\overline{2}0`$), $$\phi _{(2\overline{2}0)}=\frac{\left(\phi _{(200)}+\left[180^{}+\phi _{(020)}\right]\right)}{2}=123.8^{}\text{,}$$ (7) deviates from the projection of the effective tilt axis by only $`0.3^{}`$. Therefore the ($`2\overline{2}0`$) lattice planes are perpendicular to the effective tilt axis as per Figure 4, and the data acquired are consistent with the expectation for fcc crystals outlined in Figure 3. The actual tilting path in the Kikuchi map of crystal A is shown there as well. From the indexing suggested by this match, the $`x`$, $`y`$, and $`z`$ coordinates of the reciprocal lattice basis vectors $`a^{}`$, $`b^{}`$, and $`c^{}`$ may be inferred. This is shown in Table 3, along with the resulting lattice parameters and a comparison with literature values. The resulting errors in $`a`$ and $`b`$ are less than $`1.3\%`$, while the error in $`c`$ (which is orthogonal to the plane of the first image) is larger (around $`2.3\%`$). Both because of tilt uncertainties and reciprocal lattice broadening in the beam direction, uncertainties orthogonal to the plane of the specimen are expected to be larger than in-plane errors Fraundorf (1987). #### IV.2.2 Building a triplet from scratch: By generating linear integral combinations of the measured periodicities in reciprocal space (i.e. vector triplets of the form $`n_1g_1+n_2g_2+n_3g_3`$ where the $`n_i`$ are integers) until a minimal volume unit cell is obtained (there will be more than one way to achieve the minimum), a primitive triplet for the measured lattice can be inferred quite independent of any knowledge of candidate lattices. The primitive cell parameters determined are also listed in Table 3. With respect to literature values, these show spatial disagreements of less than $`1.6\%`$, and angular disagreements of less than $`1.5^{}`$. Although inference of the “conventional cell” from the primitive cell alone is possible, the process has not been attempted here because of complications attendant to measurement error. #### IV.2.3 Phase Identification Determining a reciprocal lattice triplet, and inferring lattice parameters therefrom, are of course not equivalent to confirming the existence of a particular phase. In order to draw a more robust conclusion about the makeup of crystal A, we extended our analysis of the structure to other lattices capable of indexing the observed spots, albeit with larger errors in spacing and interspot angle. When the spatial and angular tolerances of our candidate match analysis are increased to $`3^{}`$ and $`3\%`$, there are many tungsten oxide and carbide candidates in addition to WC<sub>1-x</sub> which show consistent lattice spacings and inter-planar angles Qin et al. (1998). In order to eliminate these candidates, it was necessary to confirm, using power spectra of amorphous regions in each image, that the spatial frequencies in Figure 3images were continuously transferred within the first passband Spence (1988); Williams and Carter (1996). By then assuming that projected reciprocal lattice frequencies make their way into the exit surface wavefield (at least at the thin edges of the particle), all the candidates except WC<sub>1-x</sub> are eliminated. Specifically, it was found that for each of the candidates except WC<sub>1-x</sub>, along one or more of the suggested zone axes at least one reciprocal lattice vector shorter than the experimental one(s) is missing in a power spectrum Qin et al. (1998). An example of this is the match with hexagonal WC<sub>x</sub> ($`a=1.058`$nm, $`c=1.335`$nm). In this case the Miller indices suggested for spot 3 ($`\overline{4}2\overline{2}`$) were inconsistent with the fact that the ($`\overline{2}1\overline{1}`$) is absent from the power spectrum of the image on the right side of Figure 3. Our conclusion that “this crystal is WC<sub>1-x</sub>” (and as we see later most of the other crystals in the specimen) is consistent with knowledge of the formation conditions, as well as with X-ray powder and EDS analysis of the same film. ### IV.3 The effective tilt direction, and recurring fringes In addition to serving as a guide for correctly choosing the azimuth of the crystal before tilting between desired zones, knowledge of the tilt axis direction plays another role: that of highlighting lattice fringes present in both specimen orientations, but caused by one and the same set of lattice planes. In single tilt experiments, the tilt axis is simply $`T_1`$. This is always perpendicular to the electron beam and hence parallel to the micrographs. Any reciprocal lattice vector parallel or antiparallel to $`T_1`$ remains in Bragg condition throughout the whole tilting process, regardless of the amount of tilt $`\vartheta _1`$. If the spacing is large enough to be recorded in the images, the same lattice fringes are seen perpendicular to the projection of $`T_1`$ in any HREM image as well. For double tilt experiments, it is convenient to introduce the concept of an effective tilt axis. The effective tilt axis is analogous to the tilt axis in a single tilt experiment, in that the double tilt can be characterized by a single tilt around the effective tilt axis of angular size equal to that in the double tilt. This effective axis is perpendicular to the electron beam and hence parallel to the micrographs only if the two specimen orientations are symmetric about the untilted position. Considering only double tilts falling into this category, let ($`\vartheta _1`$, $`\vartheta _2`$) and ($`\vartheta _1`$, $`\vartheta _2`$) denote the specimen orientations before and after. The effective tilt axis direction has an azimuth (with respect to our reference $`x`$-direction) of $$\phi _{eff}=\mathrm{tan}^1\left[\frac{\mathrm{sin}(\vartheta _1)}{\mathrm{tan}(\vartheta _2)}\right]\text{.}$$ (8) A proof of equation 8 is given in Appendix A. There exists a 180 ambiguity in the direction of the effective tilt axis using equation 8. This ambiguity can be resolved through the knowledge of the actual tilting sequence. In our experiment $`\vartheta _1=15^{}`$, $`\vartheta _2=9.7^{}`$, $`\phi _{eff}=123.5^{}`$. This is the effective tilt axis direction mentioned in previous sections. Lattice planes perpendicular to the effective tilt axis, in the double tilt case, diffract and are visible at initial and final, but not intermediate, specimen orientations. This result inspired further experimental work on, and modeling of, fringe visibility loss during tilt Qin (2000). One result of this exercise was a prediction that $`0.213`$nm fringes deviating by as much as $`4^{}`$ from the effective tilt direction in a 10nm $`WC_{1x}`$ specimen will remain visible after a $`35.6^{}`$ tilt. This was confirmed by experiment on these specimens Qin (2000). The result in turn serves to constrain the probability and error analyses below. ### IV.4 Tilt limitations and chances for success This section addresses the chances for successful 3D cell determination from images, depending on properties of both microscope and specimen. Such matters are considered in more detail elsewhere Qin (2000). In a microscope with a single-axis tilt of at least $`\pm 35.3^{}`$ and a stage capable also of $`180^{}`$ rotation, any cubic crystal with an \[$`001`$\] zone in the beam direction at zero tilt can be re-aligned by azimuthal rotation until its ($`2\overline{2}0`$) reciprocal lattice vector is parallel to T<sub>1</sub>. With an untilted tilt-rotate stage, this would allow a crystal’s \[$`001`$\] zone to remain aligned with the beam throughout the rotation. Subsequent tilting by $`35.3^{}`$ will lead to the \[$`112`$\] zone, and the lattice structure in three dimensions confirmed ala Figure 3. Under these conditions, any cubic crystal showing \[$`001`$\] zone cross fringes can be tilted so as to reveal a third spacing. Hence the probability of success with any given crystal is that of finding a randomly-oriented crystal oriented with \[$`001`$\]-zone fringes visible. Fortunately for this method, the spreading of reciprocal lattice points due to finite crystal thickness $`t`$ allows one to visualize fringes within a half angle $`\mathrm{\Theta }_t`$ of order $`1/t`$ surrounding the exact Bragg condition. Otherwise, cross-fringes would be rare indeed! The solid angle subtended by this visibility range for lattice planes intersecting along the \[$`001`$\]-zone allows us to calculate the probability that a randomly-oriented crystal will show cross-fringes of specified type. For example if we approximate the cross-fringe region with a conical bundle of directions about each zone, then for the special case of spherical particles the fraction of crystals showing the fringes of zone $`x`$ is: $$p_x=n(1\mathrm{cos}[\mathrm{\Theta }_t])\text{, where }\mathrm{\Theta }_t=\mathrm{arcsin}[\frac{g_t^2g_x^2+2g_\lambda g_t}{\sqrt{2}g_xg_\lambda }]\text{,}$$ $`n`$ is multiplicity of zone $`x`$ (e.g. $`n=3`$ for cubic $`x=`$\[$`001`$\]), $`g_x=1/d`$, $`g_\lambda =1/\lambda `$, and $`g_t=f/t`$, where $`t`$ is the thickness of the crystal in the direction of the beam and $`f`$ is a parameter of order one that empirically accounts for signal-to-noise in the method used to “visualize” fringes. For example, we expect $`f`$ to decrease if an amorphous film is superimposed on the crystals being imaged. The half-angle $`\mathrm{\Theta }_t`$ is a pivotal quantity in both the probability and accuracy of fringe measurement. A plot of the probability for seeing \[$`001`$\] cross-fringes of spacing $`d=0.202`$nm, as a function of specimen thickness $`t`$ for both spherical and laterally-infinite (rel-rod) particles, is shown in Fig. 5. Here we’ve used fit parameter $`f=0.79`$ based on data points (also plotted) that were obtained experimentally for particles of varying size from HREM images of Au/Pd evaporated onto a carbon film. As you can see, the probability of encountering cross-fringes improves greatly as crystallite size decreases toward a nanometer. Of course, as discussed in the next section, this “reciprocal lattice broadening” is accompanied by a decrease in the precision of measurements for individual lattices. Due to the tilt limits of the specimen holder in our microscope, the first HREM image along the \[$`001`$\] zone of a WC<sub>1-x</sub> nano-crystal had to be taken at a nonzero $`\vartheta _1`$ orientation. Azimuthal symmetry is thus broken. Our solution was to find a \[$`001`$\] nano-crystal whose ($`2\overline{2}0`$) reciprocal lattice vector was by chance parallel to the effective tilt axis, then tilting to the 2nd orientation. Thus nano-crystal $`A`$ was identified (by coincidence) to have an appropriate azimuth during real time study of the ($`200`$) and ($`020`$) lattice fringes. Tilting by $`35.3^{}`$ was done thereafter. For the probability of success in our case, we must multiply $`p_x`$ by the probability of viewing ($`111`$) fringes after tilting a \[$`001`$\] crystal with random aziumth by $`35.26^{}`$. This probability of finding a 3rd spacing takes the form $`p_3=m\delta /\pi `$, where $`m`$ is the multiplicity of target spacings (e.g. $`m=4`$ for a four-fold symmetric \[$`001`$\] starting zone), and the “azimuthal tolerance half-angle” $`\delta `$ (again in the spherical particle case) obeys the implicit relation: $$\theta _o=\mathrm{arctan}[\frac{\mathrm{tan}\theta _o}{\mathrm{cos}\delta }]+\mathrm{arctan}[\frac{\mathrm{cos}\gamma }{\mathrm{sin}^2\gamma (\mathrm{cos}\theta _o\mathrm{sin}\delta )^2}]\text{,}$$ where $`\theta _o`$ is the required tilt (in our case $`35.26^{}`$) and $$\gamma =\mathrm{arccos}[\frac{g_d^2g_l^22g_\lambda g_l}{2g_dg_\lambda }]\text{.}$$ The probability $`p_3`$ is also plotted as a function of specimen thickness in Fig. 5, along with the product $`p_xp_3`$. These models predict a probability of success with the strategy adopted in our experiment, for the “large” $`10`$nm WC<sub>1-x</sub> crystals in our specimen, of $`p_xp_3=7.3\times 10^4\times 0.371=2.4\times 10^4`$. Hence only one in every $`1500`$ crystals will show \[$`001`$\] cross fringes, and one in every $`4000`$ will be suitably oriented for 3D lattice parameter determination. This is consistent with our experience: The image of crystal A was recorded in one negative out of 22, each of which provided an unobstructed view of approximately $`100`$ crystals. As mentioned above, using a microscope capable of side-entry goniometer tilting by $`\pm 35.3^{}`$ with a tilt-rotate stage, the 3D parameters of all cubic crystals, when untilted showing \[$`001`$\] zone cross-fringes, could have been determined. According to Fig. 14a, the fraction of particles $`2`$ nm in thickness that are oriented suitably for such analysis approaches $`1`$ in $`100`$. Moreover, with a goniometer capable of tilting by $`\pm 45^{}`$ plus computer support for automated tilt/rotation from any starting point, each unobstructed nano-crystal in the specimen could have been subjected to this same analysis after a trial-and-error search for accessible \[$`001`$\] zones. Thus a significant fraction of crystals in a specimen become accessible to these techniques, with either a sufficient range of precise computer-supported tilts, or if the crystals are sufficiently thin. Subsequent work Fraundorf and Qin (2001) has shown that information on tilt protocols and fringe visibility for crystals of a given thickness can be elegantly sumarized with spherical maps like those in Figure 6. These are a direct-space analog to reciprocal-space Kikuchi maps, in which band thickness is proportional to d (rather than 1/d). If the sphere used has a diameter equal to specimen thickness, then the first order effect of changing thickness simply increases the separation between zones while holding the width of the bands fixed. ## V Pitfalls and uncertainties ### V.1 Cautions involving specimens and contrast transfer In this section, we discuss effects warranting caution. In the next section, models of lattice parameter uncertainty are discussed. High electron beam intensities can cause lattice rearrangement in sufficiently small nanocrystals, as well as changes in the orientation of a thin film (e.g. due to differential expansion). Sequential images of the same region at fixed tilt might allow one to check for such specimen alterations. Loss of periodicities in the recorded image, due to damping and spherical aberration zeros, were discussed in the section above on experimental design. Nonetheless, careful observations of more than one crystal, and image simulation as well, may be useful adjuncts whenever this technique is applied. We illustrate this below, with a “two-dimensional” experiment done to assess the size of errors due to finite crystal size and random orientation. The result is of help in the section on modeling uncertainties that follows. A recent paper on HREM image simulations Malm and O’Keefe (1997) indicated that deviations in orientation of a 2.8nm spherical palladium nano-crystal from the zone axes may result in fringes unrelated to the structure of the particle. Variability in measured lattice spacings was also reported to be as high as several percent, with the highest reaching 10%. To compare such results with our experimental data, 23 single crystals free of overlap with other crystals, and each showing cross-fringes, were examined. The projected sizes of these crystals range from 3.7nm$`\times `$3.8nm to 10.8nm$`\times `$7.8nm. The spacings and angles between fringes are plotted in Fig. 7. Observed cross-fringes in the HREM images fall into two categories, according to their spacings and angles. The first category is characterized by a $`90^{}`$ inter-planar angle between two $`2.12\AA `$ lattice spacings. The second one by two inter-planar angles of $`55^{}`$, $`70^{}`$ and three lattice spacings of $`2.12\AA `$, $`2.12\AA `$, $`2.44\AA `$. Only the spacings of $`2.44\AA `$ and $`2.12\AA `$ and the corresponding angle of $`55^{}`$ have been shown in Figure 7. Two conclusions can be drawn. First, since the two categories of cross-fringes match those along the \[$`001`$\] and \[$`011`$\] zone axes of WC<sub>1-x</sub>, the only two zones which show cross-lattice fringes in our HREM images, the thin film consists mainly of WC<sub>1-x</sub>. X-ray powder diffraction work on the film supports this conclusion Qin et al. (1998). Secondly, for nano-crystals free of overlap with other crystals, the observed lattice spacings and interplanar angles in HREM images have a standard deviation from the mean of less than $`1.5\%`$, and a standard deviation of less than $`1.3^{}`$, respectively. We have not observed any seriously shortened or bent fringes. Nonetheless we recommend that such fringe abundance analyses go hand in hand with stereo lattice studies of nano-crystal specimens, and that comparative image simulation studies be done where possible as well. ### V.2 Uncertainty forecasts Earlier estimates Fraundorf (1987), as well as the typical size of diffraction broadening in the TEM, suggest that lattice parameter spacing errors may in favorable circumstances be on the order $`1\%`$, and angle errors on the order of $`1^{}`$. Experiment, and a more detailed look at the theory Qin (2000), now support this impression. We will focus the discussion here on equant or spherical nanocrystals. The results should be correct within $`10\%`$ for other (e.g. thin-foil) geometries of the same thickness. The three sources contributing to the lattice spacing measurement uncertainty in images are: expansion of the reciprocal lattice spot in the image plane, uncertainty in the camera constant, and expansion of the reciprocal lattice spot along the electron beam direction, in order of decreasing relative effect Qin (2000). The uncertainty in our camera constant is measured to be about $`0.5\%`$. Uncertainties from the first and third sources above are on the order of $`1\%`$ and $`0.01\%`$, respectively, for a typical lattice spacing of 0.2 nm. The in-plane/out-of-plane error ratio is on the order of 10. Sources contributing to the measurement uncertainty of lattice parameters along the electron beam direction, when the specimen is un-tilted, include uncertainty in goniometer tilt as well as sources analogous to those above. Observation of reciprocal lattice vectors further out of the specimen plane (i.e. of fringes at high tilt) reduces the measurement uncertainty of “out-of-plane” parameters. The measurement uncertainty of inter-planar angles in images is due to lateral uncertainty in their associated reciprocal lattice spots in the image plane. Using a mathematical model of these errors Qin (2000), we predict spacing uncertainties in a 10 nm nanocrystal, tilted by $`\pm 18^{}`$, of $`2.1\%`$ for an imaged spacing and of $`8.6\%`$ for a lattice parameter perpendicular to the plane of the untilted specimen. This large “out-of-plane” uncertainty is a result of the small tilt range available with our high resolution pole piece. The estimated interplanar angle uncertainty is about $`2.3^{}`$. These predicted uncertainties Qin (2000) are between 2 and 3 times the errors observed here, and hence of the right order of magnitude. The model suggests that lattice parameter uncertainties will decrease as camera constant and tilt uncertainties decrease, and will also decrease as the tilt range used for the measurement increases. It suggests that the lattice parameter errors will increase as crystal thickness goes down. The ease of locating spacings, however, goes up as crystal thickness decreases. Hence the best candidates for application of the protocols here may be crystals in the 1 to 20 nm range. Improved tilt accuracy (hopefully with computer guidance), and low-vibration tilting so that fringes may be detected as orientation changes, would make these strategies more accurate and widely applicable as well. ## VI Conclusions When considered from the perspective of direct space imaging, crystals offer a discrete set of opportunities for measuring their lattice parameters in three dimensions. Ennumerating those opportunities for candidate lattices, or lattice classes, opens doors to the direct experimental determination of nano-crystal lattice parameters in 3D. A method for doing this, and lists of those opportunities for the special case of cubic crystals, are presented here. We apply this insight to inferring the 3D lattice of a single crystal from electron phase or Z-contrast images taken at two different orientations. For nano-crystals in particular, a double-axis tilt range of $`\pm 18^{}`$ degrees allows one to get such data from all correctly-oriented cubic crystals with appropriate spacings resolvable in a pair of images taken from directions separated by $`35.3^{}`$. In the experimental example presented, we find less than $`1.5\%`$ spatial and $`1.6^{}`$ angular disagreements between the inferred primitive cell lattice parameters of a 10 nm WC<sub>1-x</sub> nano-crystal, and literature values. We further present data on the variability of lattice fringe spacings measured from images of such randomly-oriented 10nm WC<sub>1-x</sub> crystals in electron phase contrast images. The results suggest that measurement accuracies of $`2\%`$ in spacing and $`2^{}`$ in angle may be attainable routinely from particles in this size range. Smaller size crystals may be easier to obtain data from, but show larger uncertainties, while larger or non-randomly oriented crystals (especially if guesses as to their structure are unavailable) may be more challenging to characterize in three dimensions. Precise knowledge of the tilt axes, as projected on the plane of a micrograph, is crucial to implementation. This information, if coupled with on-line guidance on how to tilt from an arbitrary two-axis goniometer orientation in any desired direction with respect to the plane of an image or diffraction pattern, could make this strategy and related diffraction strategies Fraundorf (1981b) for lattice parameter measurement more routine. Future microscopists might then be able to interface to individual nanocrystals much as the nano-geologist in the first paragraph of this paper examined her “hand specimen”. Lastly, diffraction can also be used in this “stereo analysis” mode (with crystals large enough to provide diffraction patterns), although the easier accessibility of high spatial frequencies via diffraction often makes the large tilts used here unnecessary. They can also be put to use in darkfield imaging applications, by forming images of the specimen using “beams” diffracted by the periodicities which serve as diagnostic of a given lattice (for example those associated with each of the protocols of Figure 1). Images so taken of nano-crystalline specimens, for example at two tilts with three different darkfield conditions, would be expected to show correlations among that subset of the crystals correctly oriented for diffraction with all three reflections. Although this strategy may never allow precise lattice parameter determinations given limits on objective aperature angular size, it may be a very efficient way to search for crystals correctly-oriented and of correct type for one of the imaging protocols described here. Moreover, because such lattice-correlations in three dimensions contain information beyond the pair-correlation function, they may be able to support the new technique of fluctuation microscopy Treacy and Gibson (1993, 1996); Voyles et al. (2000) in the study of paracrystalline specimens like evaporated silicon and germanium Gibson and Treacy (1997, 1998) whose order-range is too small for detection by other techniques. ## Appendix A Effective tilt axis azimuth Let ($`\vartheta _1`$, $`\vartheta _2`$) and ($`\vartheta _1^{}`$, $`\vartheta _2^{}`$) denote orthogonal tilt values for two specimen orientations, and $`\varphi _{eff}`$ the azimuth of the effective tilt axis between these orientations. Any reciprocal lattice vector with untilted Cartesian coordinates $`|g`$, and with identical micrograph coordinates $`|g_m`$ at the two tilted orientations, will (following equation 4) obey: $$A(\vartheta _1,\vartheta _2)|g_m=|g=A(\vartheta _1^{},\vartheta _2^{})|g_m\text{,}$$ (9) where $$|g_m=\left(\begin{array}{c}g\mathrm{cos}\left(\phi _{eff}\right)\\ g\mathrm{sin}\left(\phi _{eff}\right)\\ 0\end{array}\right)\text{.}$$ (10) Expanding, this gives three equations which, combined with equation 10 $`g_{mx}=g_{my}\mathrm{tan}(\phi _{eff})`$, can be solved for the three unknowns $`\vartheta _1^{}`$, $`\vartheta _2^{}`$, $`\phi _{eff}`$, to get: $$\vartheta _1^{}=\vartheta _1\text{}\vartheta _2^{}=\vartheta _2\text{, and}$$ (11) $$\phi _{eff}=\mathrm{tan}^1\left[\frac{\mathrm{sin}(\vartheta _1)}{\mathrm{tan}(\vartheta _2)}\right]\text{.}$$ (12) This provides an equation for the azimuth of the effective tilt, and confirms that symmetry about the zero tilt position is a necessary and sufficient condition for the reciprocal lattice vector $`|g`$, and it’s associated lattice fringe, to show a common direction in micrographs at both tilts. ###### Acknowledgements. Thanks to W. Shi, J. Li and W. James at University of Missouri - Rolla for the tungsten carbide specimens, and to MEMC and Monsanto for regional facility support.
warning/0001/astro-ph0001375.html
ar5iv
text
# 1 Introduction ## 1 Introduction The Luminous Infrared Galaxies (LIGs) are a class of objects characterized by high infrared luminosities (L$`{}_{IR}{}^{}>10^{11}`$ erg s<sup>-1</sup>). Both AGNs and starbursts can power LIGs, but the relative contribution of these two energy sources to the bolometric luminosity is still unclear. X-ray observations in the hard (2-10 keV) band can be a powerful tool to unveil the AGN emission in the LIGs and to estimate its contribution to the total luminosity. It is at present impossible to analyze a representative LIGs sample in the 2-10 keV band, because the hard X-ray observations performed up to now are strongly biased in favor of AGN–dominated sources. Nevertheless, a significant number of IR-selected LIGs with X-ray observations is now available both in the literature and in public archives, and therefore a comparison between the X-ray properties of AGN–dominated and IR–selected LIGs is possible. We collected the data of all the luminous infrared galaxies observed so far in hard X-rays and studied their X-ray emission and the correlation between their X-ray and infrared properties. ## 2 X–ray and IR properties of the sample The X-ray properties of our sample of objects are very heterogeneous both in terms of brightness and spectral shape. Most of the objects optically classified as type 1 Sys or QSOs are characterized by bright X-ray emission (relative to the IR luminosity) and their X-ray spectrum does not show indications for significant cold absorption. A significant fraction of the narrow line AGNs are also relatively bright in the X-rays and their spectrum is characterized by a photoleletric cutoff ascribed to Compton thin absorbing gas ($`\mathrm{N}_\mathrm{H}<10^{24}\mathrm{cm}^2`$). The remaining objects optically classified as AGNs are very weak in the X-rays. This can be due to the intrinsic weakness of the AGN component or to an absorbing column density higher than $`10^{24}`$cm<sup>-2</sup> (Compton thick AGNs). Finally, a few objects are optically classified as starbursts and are all very weak in the X rays. In Fig. 1 we plot the X/IR flux ratio versus the infrared colour defined as C$`{}_{IR}{}^{}=2\times \frac{f_{25}}{f_{60}}`$, where f<sub>25</sub> and f<sub>60</sub> are the flux densities at 25 $`\mu `$m and at 60$`\mu `$m. A clear correlation is apparent in Fig. 1: type 1 AGNs are preferentially in the high X/IR ratio and warm infrared colour part of the diagram. Moving towards lower 25/60$`\mu `$m ratios we find lower X/IR ratios and an increasing fraction of obscured AGNs at first, and of starbursts afterwards. A simple model in agreement with this correlation is shown in Fig. 2: starting from a “pure Seyfert 1” point, at the top -right corner in the plot, the oblique rightmost (light) line gives the expected location of AGN–dominated objects, with an increasing X-ray absorbing column density moving towards the bottom. The dotted (dark) curves give different degrees of mixing of starburst and AGN, with the AGN contribution lowering moving to the left. In this model the absorption suffered by the IR radiation is only a small fraction of that inferred from the X-ray column density by assuming a standard dust–to–gas ratio. Indeed, we find that there is no way to explain the correlation if we assume that (1) the same amount of material obscures both the X-rays and the infrared and that (2) the dust–to–gas ratio is galactic. This is because if none out of the two hypothesis above are relaxed, the absorbed model gives a pure–AGN curve that is almost horizontal, because the IR colour decreases too fast with respect to the X/IR flux ratio. A simple explanation for the difference between the absorption in the IR and in the X-rays can be the following: if the density in the circumnuclear torus decreases radially, the X–rays, emitted in the very central region, are much more absorbed than the 25$`\mu `$m radiation, emitted by the warm dust located along the inner face of the torus, far from the plane of the accretion disk. Besides the optical classification, the simple picture depicted above is supported by additional pieces of evidence: 1) broad lines in the polarized spectrum or in the near IR are found only in type 2 objects with warm IR colour, that, according to our model, are the AGN-dominated LIGs; 2) the large majority of the IR-cold sources have steep (starburst-like) X-ray indices while IR-warm sources have flatter (AGN-like) indices. For a more detailed discussion about these issues, we remind to a forthcoming paper . ## 3 The 20-200 keV emission of LIGs Our sample includes a subsample of 13 sources optically classified as type 2 AGNs and observed by BeppoSAX up to 200 keV. 6 out of the 11 sources that are Compton–thick in the 2-10 keV range are completely absorbed also in the 10 to 200 keV band, therefore implying a column density N$`{}_{H}{}^{}>10^{25}`$ cm<sup>-2</sup>. Only two of the 11 Compton thick sources have an excess in the 15-100 keV range, while for the remaining three the hard 15-100 keV X-ray emission is unconstrained. The shortage of objects with 10<sup>24</sup>cm$`{}_{}{}^{2}<`$N$`{}_{H}{}^{}<10^{25}`$ cm<sup>-2</sup> already pointed out in a sample of optically selected Seyfert 2s , would have important consequences in the synthesis models of the X-ray background, since this class of Sy2 should contribute significantly to the XRB in the 10 keV to 200 keV band, where most of the XRB energy is emitted. These models are by and large successful in synthesizing the XRB from the contributions of individual AGNs; they must however include a dominant contribution from absorbed, type 2 AGNs, which up to now have been observed only at low redshifts and low luminosities. Assuming that type 1 and type 2 AGNs evolve in the same way (as predicted by unified models), the zero-th order extrapolation has to face several discrepancies, and could be cured only by adding extra type 2 sources at intermediate or high redshifts . It has been proposed that LIGs could be the required additional sources, thanks to their high luminosity and high spatial density. However, our work indicates that even those sources in which the presence of a luminous AGN is strongly supported by several optical and IR indicators, are on average very dim in the hard X-rays up to 100 keV, because of heavy obscuration. Therefore, the contribution of this class of sources to the XRB is negligible. Acknowledgements. Much of this work has been don in collaboration with R. Gilli, R. Maiolino and M. Salvati. The author acknowledges the partial financial support from the Italian Space Agency (ASI) through the grant ARS–99–15.
warning/0001/cond-mat0001065.html
ar5iv
text
# Soft spin waves in the low temperature thermodynamics of Pr0.7Ca0.3MnO3 ## Abstract We present a detailed magnetothermal study of Pr<sub>0.7</sub>Ca<sub>0.3</sub>MnO<sub>3</sub>, a perovskite manganite in which an insulator-metal transition can be driven by magnetic field, but also by pressure, visible light, x-rays, or high currents. We find that the field-induced transition is associated with an enormous release of energy which accounts for its strong irreversibility. In the ferromagnetic metallic state, specific heat and magnetization measurements indicate a much smaller spin wave stiffness than that seen in any other manganite, which we attribute to spin waves among the ferromagnetically ordered Pr moments. The coupling between the Pr and Mn spins may also provide a basis for understanding the low temperature phase diagram of this most unusual manganite. The rare earth perovskite manganites (R<sub>1-x</sub>A<sub>x</sub>MnO<sub>3</sub>) are associated with a wide variety of fascinating physics due to the strong coupling between their electronic, magnetic and lattice degrees of freedom. Phenomena observed in these materials include $`\mathrm{`}\mathrm{`}`$colossal$`\mathrm{"}`$ magnetoresistance, real-space charge ordering, electronic phase separation, and a diverse variety of magnetoelectronic ground states . Although this entire class of materials displays unusual behavior, one material, Pr<sub>0.7</sub>Ca<sub>0.3</sub>MnO<sub>3</sub> (PCMO), has been shown to display a particularly rich set of phenomena. Like several other manganites, the resistivity of PCMO is reduced by orders of magnitude in a magnetic field due to an irreversible transition from an insulating antiferromagnetic to a metallic ferromagnetic state . What sets PCMO apart from the other manganites is that the metastable insulating state also can be driven metallic by the application of light , pressure , x-rays , or a high current . Despite the strong recent interest in this wide range of unique phenomena in PCMO, there has been no clear understanding of why this particular manganite is so different from the others. We have performed a detailed study of the low temperature magnetothermal properties of PCMO. We find that there is an enormous release of heat at the field-induced insulator to metal transition at low temperatures, which explains the complete irreversibility of the transition. In the ferromagnetic (FM) state at low temperatures, our specific heat and magnetization measurements indicate a ferromagnetic spin wave stiffness which is far below that seen in other conducting manganites. The data can be explained most easily as a result of ferromagnetism among the moments associated with the Pr ions, suggesting that the Pr magnetism is crucial to understanding the unusual low temperature properties of this material. We studied both a ceramic sample of PCMO, synthesized by a standard solid state technique, and a single crystal grown in a floating zone mirror furnace. Both samples were judged to be single phase from x-ray diffraction studies. The cation concentration ratio, as measured by plasma atomic emission spectroscopy, was also consistent with the nominal concentration. While we show only results from the single crystal, data from the two samples were qualitatively and quantitatively consistent in all of the studied properties, and both had low temperature phase diagrams (including a regime of field-induced ferromagnetic metallic phase) consistent with previously published studies of PCMO . Magnetization was measured in a Quantum Design SQUID magnetometer, and specific heat was measured by a semiadiabatic heat pulse technique calibrated against a copper standard. Magnetocaloric measurements were made by temperature-controlling the calorimeter at a few degrees above the surrounding cryostat temperature while the field was swept. We then measured the input power required to maintain constant temperature during the field sweep and attributed changes in the input power to magnetocaloric effects. The field-temperature phase diagram of PCMO is shown in the inset of figure 1 based on earlier measurements . Upon cooling in zero field, PCMO first undergoes a charge-ordering transition at T<sub>co</sub> $``$ 230 K. Upon further cooling, the Mn spins order first into a pseudo-CE type antiferromagnetic state at T<sub>AF</sub> $``$ 150 K, and then into a canted antiferromagnetic (CAFM) state at a lower temperature T<sub>CAFM</sub> . This low field phase is quite complex due to the incommensurability of the charge order with the 30$`\%`$ Ca doping, and the resultant disorder leads to both spin-glass-like behavior and electronic phase separation . Application of a sufficiently high magnetic field at low temperatures induces a first order transition from the CAFM phase to a conducting FM state, and there is a strong hysteresis associated with this transition (the hatched region in the inset indicates where the state is history-dependent). For T $``$ 60 K, the transition is completely irreversible so that if the material undergoes the field-induced FM transition it remains in a FM conducting state until the temperature is raised above 60 K. In figure 1 we demonstrate this irreversibility by plotting the magnetization vs. field after zero-field-cooling to T = 10 K and then sweeping the field up and down 0 $``$ 7 T $``$ -7 T $``$ 7 T. The reduced moment during the initial sweep up in field for H $`<`$ 4 T corresponds to the magnetization of the canted antiferromagnetic phase, and the rise in magnetization during that sweep at $``$ 4 T corresponds to the phase transition. As seen in the figure, the magnetization maintains its ferromagnetic nature during all subsequent field sweeps. Since the field-induced phase transition is highly irreversible and therefore first order, one expects an associated release of heat which should be observed in our magnetocaloric measurements. In figure 2 we show the raw data of a magnetocaloric measurement, plotting the power input from the temperature controller (P) vs. magnetic field after zero-field-cooling as the field is changed from 0 $``$ 9 T, 9 T $``$ -9 T, and -9 T $``$ 9 T. For H $``$ 1 T we see small rises and falls in P(H) during the initial sweep up in field. We attribute these reproducible features to heat release associated with the spin-glass-like character of the zero-field-cooled CAFM state , i.e. irreversible relaxation of the spin configuration during the initial field sweep (a more detailed treatment of this behavior will be included in a future paper ). The most prominent feature in the magnetocaloric data, however, is the large negative peak in P(H) during the initial field sweep. This peak corresponds to a reduction in the heat required from the temperature controller to maintain constant temperature, and we attribute this reduction to the heat released at the CAFM-FM transition. Note that there is no similar peak observed in subsequent sweeps of the magnetic field, which is consistent with the irreversibility of the transition. The integrated magnitude of the peak in P(H), $`Q=(P/(\frac{dH}{dt}))𝑑H`$, is extraordinarily large at low temperatures - about 15 $`\pm `$ 1 J/mole. Within our resolution, $`Q`$ is independent of both the sweep rate between 6 G/sec and 24 G/sec and the temperature for T $``$ 40 K, but Q gradually decreases for higher temperatures . Comparing the magnitude of Q to $`C(T)𝑑T`$, one finds that the heat release at the transition is sufficient to raise the temperature of a perfectly thermally isolated sample from 5 to 17.5 K, i.e. by more than a factor of 3. Indeed, as can be seen in the inset to figure 2, when we simply sweep the field on our 0.22 gram sample on the calorimeter, the temperature of the calorimeter rises from 5 to 11 K. The magnitude of Q can be understood as a release of the Zeeman energy associated with the CAFM phase relative to that in the FM phase. This can be calculated from our measured value of M(H) as $`H𝑑M`$ over the field range of the transition which is 10 $`\pm `$ 1 J/mole for T $``$ 40 K, in reasonable agreement with the measured values of Q . The large magnitude of Q relative to the specific heat explains the complete irreversibility of the transition at low temperatures, since there is no corresponding energetic advantage to the CAFM phase at low fields which would drive the reversal of the transition. We measured the specific heat as a function of temperature, C(T), at 0, 3, 6, and 9 tesla after cooling at those fields and at 0 and 3 tesla in the FM state (induced by raising the field temporarily to 9 T after reaching low temperatures). As shown in figure 3 we find that we can fit the temperature dependence in the FM state with the simple form C(T) = $`\beta `$ T<sup>3</sup> \+ $`\delta `$ T<sup>3/2</sup> where the two terms correspond to the lattice specific heat and the specific heat associated with ferromagnetic spin waves respectively. While we cannot rule out the possibility of a contribution to the specific heat which is linear in temperature (such as would arise from either free electrons or a spin glass state), we find that $`C_{mag}`$ = C(T) - $`\beta `$ T<sup>3</sup> has a power law temperature dependence with an exponent of almost exactly 1.5 as shown in the log-log plot in the inset to figure 3. Our fitted values of $`\beta `$ vary with field by only a few percent, while $`\delta (H)`$ varies by 65$`\%`$ implying that the field dependence of the specific heat originates almost entirely in the magnetic component. We directly measured the field-dependence of the specific heat, C(H) by zero-field cooling the sample and then measuring the specific heat every 0.25 T while sweeping the field from 0 $``$ 9 T $``$ -9 T $``$ 9 T as shown in figure 4. During the initial sweep up in field, there is a sharp drop in C(H) at about 4 T corresponding to the CAFM $``$ FM phase transition. On all subsequent sweeps, C(H) is quite smooth and monotonically decreases as the magnitude of the applied field increases. Since the fits to C(T) indicate that the lattice contribution to the specific heat ($`\beta T^3`$) is almost field-independent, we fit C(H) in the FM state to a Heisenberg spin wave model as shown by the solid line in figure 4 : $$C(H)=A+\frac{V_{mole}k_B^{5/2}T^{3/2}}{4\pi ^2D^{3/2}}\underset{g\mu _BH/k_BT}{\overset{\mathrm{}}{}}\frac{x^2e^x}{(e^x1)^2}\sqrt{x\frac{g\mu _BH}{k_BT}}𝑑x$$ (1) where $`V_{mole}`$ is the molar volume and the only two fitting constants A and D are an offset to account for the lattice contributions and the stiffness constant of the spin wave spectrum respectively . While such a form can also fit C(H) in other ferromagnetic manganites, the magnitude of the field-induced suppression of C(H) is 15 times larger in PCMO than in the other ferromagnetic manganites such as La<sub>0.7</sub>Sr<sub>0.3</sub>MnO<sub>3</sub> (shown as dashed line in figure 4) and La<sub>0.7</sub>Ca<sub>0.3</sub>MnO<sub>3</sub> . This larger suppression of C in a field implies that the magnetic specific heat is much larger in PCMO than in the other materials and consequently that the spin wave spectrum is much softer. The fit to C(H) of PCMO yields a value of D = 28.0 $`\pm `$ 0.3 meV-Å<sup>2</sup>. This is a factor of 5-6 smaller than that obtained from neutron scattering studies of other ferromagnetic metallic manganites including La<sub>0.70</sub>Sr<sub>0.30</sub>MnO<sub>3</sub> , La<sub>0.67</sub>Ca<sub>0.33</sub>MnO<sub>3</sub> , Pr<sub>0.63</sub>Sr<sub>0.37</sub>MnO<sub>3</sub> , and La<sub>0.70</sub>Pb<sub>0.30</sub>MnO<sub>3</sub> the first two of which are in good agreement with the measured field-dependence of C(H) . The low value of D from the fits to C(H) can be tested in two ways. First, the parameter $`\delta `$ obtained from our fit to C(T) at H = 0 yields an estimate of D, since $`\delta =0.113k_BV_{mole}\left(\frac{k_B}{D}\right)^{3/2}`$ within the Heisenberg spin wave model . Our value of $`\delta `$ corresponds to D = 25.3 $`\pm `$ 0.5 meV-Å<sup>2</sup>, in reasonable agreement with the fit to C(H). Second, the very soft spin wave stiffness in the ferromagnetic state can also be tested by fitting our magnetization data, since the magnetization per unit volume at low temperatures within the Heisenberg model is given by , $$M(0,H)M(T,H)=g\mu _B\left(\frac{k_BT}{4\pi D}\right)^{3/2}f_{3/2}(g\mu _B(HNM)/k_BT)$$ (2) where $`f_p(y)=_{n=1}^{\mathrm{}}\frac{e^{ny}}{n^p}`$, and NM is the demagnetization field and M is the magnetization. Such a fit to M(T) (shown in Figure 5 as the solid line) reproduces our data with the exception of the very lowest temperature points which are suppressed by $``$ 0.3$`\%`$ possibly due to a minute presence of a second phase (this feature is also seen in the polycrystalline sample). A fit to the data using equation 2 yields D = 25.1$`\pm `$ 2.1 meV-Å<sup>2</sup> which is in agreement with our other two values. We now discuss a possible origin of the very low ferromagnetic spin wave stiffness constant in PCMO. There is no reason to expect that the Mn-Mn ferromagnetic double exchange interaction should be a factor of 5 weaker in this compound than in all of the other ferromagnetic metallic manganites (in fact studies have shown that D at low temperatures is remarkably independent of ordering temperature in the manganites ), and recent neutron scattering measurements of D for the Mn spins in PCMO yield D $``$ 150 meV-Å<sup>2</sup> . While one might invoke electronic phase separation to explain the thermodynamic data, it is difficult to imagine that phase separation could reproduce quantitative agreement with the Heisenberg model consistently for the three different data sets. Furthermore, increasing the field should increase the proportion of the ferromagnetic phase which already includes $``$ 90 $`\%`$ of the sample based on the magnetization data. Thus, for $`C_{mag}`$ to decrease as strongly with field as we observe, the residual phase would need to have an associated specific heat about 100 times larger than is typical for antiferromagnetic manganites . A more plausible explanation is suggested by the data of Cox $`\mathrm{𝑒𝑡}`$ $`\mathrm{𝑎𝑙}.`$ who reported that there is a ferromagnetic moment associated with the Pr ions for temperatures below T<sub>c</sub> = 60 K . We propose that the spin waves on this ferromagnetic sublattice are responsible for our data, and we suggest that the Pr magnetism could be critical to the low temperature thermodynamics of PCMO. Softer spin waves associated with the Pr moment would have a much larger associated specific heat than the Mn spins and therefore would dominate both C<sub>mag</sub>(T) and C(H). A careful examination of M(T) for our samples at low fields shows a distinct rise at T $``$ 50-60 K (Figure 5 inset), and C(T) also shows a peak there, confirming the existence of the ordering transition . Furthermore, in all studies of the magnetization of PCMO, the saturation moment at high fields and low temperature is at least 10$`\%`$ higher than that of other ferromagnets, such as La<sub>0.70</sub>Sr<sub>0.30</sub>MnO<sub>3</sub> and La<sub>0.70</sub>Ca<sub>0.30</sub>MnO<sub>3</sub> . An additional ferromagnetic moment such as that from the Pr is required to explain this excess magnetization since the Mn spins in PCMO display considerable canting even at high field . The ferromagnetism among the Pr ions also appears to affect the remarkable phase behavior of this material, since the ferromagnetic T<sub>c</sub> $``$ 60 K for the Pr corresponds to the temperature below which the field-induced and pressure-induced CAFM-FM transitions become irreversible. Furthermore, this is also the temperature above which the x-ray induced metallicity is quenched . It thus appears that the Pr ferromagnetism destabilizes the CAFM phase relative to the FM phase, so that in zero field the CAFM phase is stable for T $`>`$ 60 K but then becomes unstable to the FM phase below T $``$ 60 K when the Pr spins order. The free energy difference between the CAFM and FM state in zero field could be small enough that the FM state does not nucleate spontaneously at such low temperatures, but an applied magnetic field would make the FM state energetically more favorable and induce the first order phase transition. In this scenario, both the Pr ferromagnetism and its coupling to the Mn moments are crucial to understanding the physics and are therefore inseparable from the numerous unique phenomena observed in PCMO . The importance of rare-earth magnetism to the explanation of the unusual properties of the manganites has been largely ignored in previous studies, and at least in the case of PCMO it appears that this assumption is not justified. We gratefully acknowledge helpful discussions with P. Dai and J. A. Fernandez-Baca, and financial support from NSF grant DMR 97-01548, the Alfred P. Sloan Foundation and the Dept. of Energy, Basic Energy Sciences-Materials Sciences under contract #W-31-109-ENG-38.
warning/0001/astro-ph0001149.html
ar5iv
text
# Looking for GRB progenitors ## Introduction It is surprising how little we know of GRB progenitors when we consider how much work is devoted to the subject. Different objects and models of progenitors were proposed, some already forgotten while others still being intensively studied. Considering the diversity of gamma ray bursts it is probable that they come from different types of astronomical objects, and this should promote the work on different types of proposed progenitors. Recently much of the weight was placed on collapsars, and connection between supernovae and GRBs, although other models, as compact object mergers, are still on the stage. Compact object binaries have recently drawn much attention in the astronomical community. Several models of compact object binaries were proposed as GRBs progenitors, namely: double neutron star mergers meszrees97 ; ruffert97 , black hole – neutron star mergers lee95 , black hole – white dwarf mergers fryer99b and helium star mergers fryer98 . Mergers of compact object binaries are also most often considered sources of gravitational waves. As gravitational wave detectors, LIGO and VIRGO, will soon be operational, the question of these merger rates arises allen99 . To predict compact object merger rates we use Monte Carlo simulations to produce populations of different compact object binaries. In this approach one generates massive binaries at ZAMS and evolves them through consecutive stages of single and binary star evolution, which may eventually lead to formation of a compact object binary. Final synthesis of large ensembles of compact object binaries allows then statistical studies and calculation of expected merger rates. Several compact object merger rates estimates narayan91 ; phinney91 ; tutukov93 ; zwart96b ; bb1 ; bbz ; fryer99a have been already published. However, calculations of this rates are based on many assumptions and use parameters with very uncertain values. Whereas the evolution of single stars is reasonably well known eggleton89 , the distributions of initial binary conditions as well as some aspects of binary star evolution are uncertain. Finally the population synthesis codes must deal with uncertainties in supernova explosion mechanisms, and in particular with the value of the kick a neutron star (or a black hole) receives at birth, and the mass of formed in the explosion compact object bbz . Previous population synthesis calculations concerned double neutron stars and only few dealt with neutron star black hole binaries. So far only in one case fryer99a all kinds of the proposed GRBs compact object binary progenitors were considered. Our present contribution follows the same line of work, although we use a different population synthesis code and we would like to communicate our first results here. Preliminary comparison shows striking similarities of the results, which taking into account different codes, points toward robustness of the population synthesis method in spite of many uncertainties involved in the calculations. ## The Model Our evolutionary code is primarily based on the prescriptions given by tout97 and eggleton89 , but we use the number of many revised or newly developed specific evolutionary prescriptions from bethe98 ; zwart96a ; pols94 ; podsiadlowski92 . In our calculation of single and binary evolution we include stellar winds (normal and LBV wind), magnetic breaking, quasi-dynamic mass transfer, common envelope evolution, hyper-accretion onto compact objects, detailed supernova explosion treatment and gravitational wave energy loss in compact object binaries. We start the evolution of a given binary when both components are at ZAMS, then each star evolves through different stages of its life depending on its mass: main sequence, Hertzsprung gap, red giant branch, horizontal branch, asymptotic giant branch, and either supernova explosion which leads to formation of a neutron star (or a black hole) or a phase of enhanced mass loss and formation of a white dwarf. During each stage of the binary evolution the components may interact, which changes the consecutive component’s evolution (either through rejuvenation, stripping the component off its outer layers, or even by swallowing one component by the other). Interaction of binary components may lead to one of the proposed GRB progenitors, the helium star merger, when the compact object is engulfed in giant’s envelope. If there is not enough orbital energy to eject the giant’s envelope then the compact object spirals in through the giant’s envelope finally merging with the giant’s core, which can be torn out in the process, and form an accretion disc around a compact object which is then already a black hole, due to the accretion during spiral in. Other systems follow their evolutionary paths, and we collect the informations on the types of compact object binaries that are proposed for GRBs progenitors: double neutron stars, black hole neutron stars and black hole white dwarf systems. These systems interact only due to gravitational energy wave loss, finally merging and forming a black hole with massive thick accretion disc which presumably may lead to a gamma ray burst. ## Results In Fig. 1 we show the relative numbers of 4 different GRB progenitor types that merge within the Hubble time (15 Gyrs) as a function of the width of the distribution from which we draw kick velocity a compact object receives in a supernova explosion. Two things are clearly seen; first the number of WD-BH binaries and Helium mergers (He-BH) is about the same and is more then an order of magnitude greater then the number of NS-NS and BH-NS binaries. Second, the number of a given progenitor type falls off approximately exponentially with the kick velocity. Relative production rates may be calibrated (eg. see eq. 14 in bb1 ). For example, assuming the width of kick velocity, let say, v$`{}_{kick}{}^{}=200`$ km s<sup>-1</sup> yields: 1 merging event per Milky Way like galaxy per Myr for BH-NS systems, 3 events for NS-NS binaries and 60 for WD-BH and Helium mergers. In Fig. 2 we show the cumulative rate of different merging events. We have combined our relative numbers for different progenitors with the star formation rate function madau96 ; totani97 , and taking into account the evolutionary time delay of a given merging event we integrated our production rates to get the merger rates as a function of redshift. In this example calculation we assumed $`\mathrm{\Omega }_M=1.0,\mathrm{\Omega }_\mathrm{\Lambda }=0.0`$, and the Hubble constant $`H_0=65`$ km s$`{}_{}{}^{1}\mathrm{Mpc}_{}^{1}`$. We used the kicks drawn from the distribution which is a weighted sum of two Gaussians: 80 percent with the width of 200 km s<sup>-1</sup> and 20 percent with the width 800 km s<sup>-1</sup>. In Fig. 2 we show also BATSE gamma ray burst detection rate corrected for the sky exposure. Comparison of the cumulative distributions for different progenitor types with he BATSE rate shows that if any of the progenitor types included here was to reproduce the BATSE rate, then we should not see GRBs from redshifts greater then unity! Of course this is not the case, as GRBs with higher redshifts were observed. However, we haven’t yet introduced the collimation factor into our results, presented in Fig. 2, which certainly will lower down our predicted rates. To lower down our calculated rates to the BATSE rate, for average GRBs redshift of about 2, we would need the collimation of about 4 for BH-WD and Helium mergers and about 12 for BH-NS and NS-NS mergers. As seen in Fig. 2 the curves flatten out for high redshifts (z $``$ 5). In other words we do not expect binary mergers at high redshifts. This is a combined effect of the star formation rate function we have used, which falls down for high redshifts and of the non zero lifetimes of progenitors, prior to the final merging event. Each of our binaries needs a specific time to evolve into a compact object binary (t<sub>evol</sub>) and then needs time to merge due to gravitational wave energy losses (t<sub>merger</sub>). This times are non negligible and are specific for each group of proposed binary GRBs progenitors. For our sample of binaries we found that $`t_{life}`$ ($`t_{evol}+t_{merger}`$) are, for NS-NS: $`10^7`$$`10^{12}`$ yrs, for BH-NS: $`10^7`$$`10^{10}`$ yrs, for BH-He: $`10^6`$$`10^9`$ yrs, for BH-WD: $`10^7`$$`10^{12}`$ yrs. ## Conclusions * GRB binary progenitor production rates fall off exponentially with width of natal kick velocity distribution. * Evolutionary times can not be neglected in computations of GRB’s progenitor rates. * Assuming that all GRB’s are connected to NS-NS or BH-NS binaries, the collimation must be of order $`3\times 10^2`$ (12). If we assume that all GRB’s result from binary mergers, then the population is dominated by BH-WD and BH-He star mergers, and the collimation must be of order $`3\times 10^3`$ (4). * We do not expect binary mergers at high redshifts (z $``$ 5). Acknowledgments. We acknowledge the support of the following grants KBN-2P03D01616, KBN-2P03D00415, KBN2P03D02117.
warning/0001/astro-ph0001052.html
ar5iv
text
# Near infrared imaging spectroscopy of NGC1275 ## 1 Introduction NGC 1275 (Perseus A, 3C 84) at the center of the Perseus cluster has been variously classified as an active galaxy, a merger, a cooling flow galaxy, a BL Lac object, and an FR I radio jet source. However, many of its physical characteristics — from X-ray to radio — remain baffling. NGC 1275 exhibits the characteristics of several source types: Seyfert (Seyfert:43 (1943)) included it in his original list, Minkowski (Minkowski:57 (1957)) called it a galaxy “collision”. The dominant low velocity (LV) component ($`5200\text{ km\hspace{0.17em}s}\text{-1}`$) is the giant cD elliptical NGC 1275, the other component (HV, $`8300`$ km s<sup>-1</sup>) is probably a nearby spiral galaxy (at least partially) in front of the main galaxy (Haschick et al. Haschick:82 (1982)). Veron (Veron:78 (1978)) suggested that NGC 1275 was a BL Lac object, while other authors have suggested that it was more compatible with LINERs (e.g. Norgaard-Nielsen et al. Norgaard-Nielsen:90 (1990)). NGC 1275 has a powerful, nearly point like X-ray emitting nucleus with a flux of L$`_\text{X}`$$`2\times `$10<sup>43</sup> erg s<sup>-1</sup> (Prieto Prieto:96 (1996)), which is embedded in an extended halo of diffuse X-ray emitting intracluster gas. The nucleus of NGC 1275 is displaced by about 25″ from the center of symmetry of the large scale ($`2`$′) cluster gas halo, perhaps due to its peculiar motion through the gas (Böhringer et al. Bohringer:93 (1993)). Temperature profiles of the X-ray halo suggest that a “cooling flow” of $`200`$ Myr<sup>-1</sup> is condensing out of the surrounding intracluster medium onto the galaxy (Fabian et al. Fabian:81 (1981), Cardiel et al. Cardiel:95 (1995)). As in all cooling flow galaxies, the ultimate fate of this purported inflowing material is unclear, but visible band colors and Balmer absorption lines spectra characteristic of A0 stars show that at most 10% of the gas can form stars with a standard Salpeter IMF (Romanishin Romanishin:86 (1986)). NGC 1275 is unique among cooling flow galaxies in that it in fact shows evidence of ongoing star formation. Even by the standards of infrared luminous mergers, NGC 1275 has very powerful, extended H<sub>2</sub> emission (Fischer et al. Fischer:87 (1987)) matched only by NGC 6240 (van der Werf et al. vanderWerf:93 (1993)). In NGC 1275, this emission has been speculated to have its origin in the cooling flow, although fast J-shocks, UV pumping, and X-ray heating via supernovae and/or AGN also produce strong H<sub>2</sub> emission. Merging events are a likely mechanism for supplying NGC 1275 with substantial (M$`\stackrel{>}{}`$ 3$`\times 10^9`$ M) quantities of H<sub>2</sub> in giant molecular clouds which are detected through mm wave CO observations within the central regions (Inoue et al. Inoue:96 (1996)). CO data also show two kinds of motion: a large ($`\stackrel{>}{}`$ 5 kpc) scale rotation which has been attributed to merging/cannibalism, and smaller scale turbulence, attributed to the inflow of cooling flow gas (Reuter et al. Reuter:93 (1993)). Recent 2.6 mm CO interferometric observations reveal the majority of the gas to be in a ring structure of radius $`4`$″ oriented in the East–West direction. The ring has a westerly extension up to 30″ from the nucleus, and nearly $`80\%`$ of the total CO is on the western side (Inoue et al. Inoue:96 (1996)). Very deep interferometric observations of the nucleus itself show no CO absorption from very cold gas, which puts very tight limits on any mass deposition from a cooling flow if the gas cools to nearly 3 K (Braine et al. Braine-co1-0:95 (1995)). NGC 1275 is a Fanaroff-Riley type I (F–R I) radio source with powerful ($`0.4`$ Jy at 22 cm) radio lobes $`10`$ kpc long and oriented at position angle (PA)160 which emanate from a central core (Pedlar et al. Pedlar:90 (1990)). The asymmetry of the radio lobes and their spectral indices suggest that we are looking nearly pole-on at the central F–R I radio engine (Pedlar et al. Pedlar:90 (1990), Levinson et al. Levinson:95 (1995)). Within the central 30″ region, relativistic particles and the magnetic field from the bipolar radio lobes produce a pressure, which exceeds the thermal gas pressure and so must disturb any cooling flow within that region (Pedlar et al. Pedlar:90 (1990), Böhringer et al. Bohringer:93 (1993)). At even smaller scales, VLBI imaging shows that the central core ($`1`$″) breaks up into many smaller components and a southern/northern jet pair (Venturi et al. Venturi:93 (1993), Vermeulen et al. Vermeulen-counterjet:94 (1994)). Recently, a new component has been discovered only 55 mas south of the central peak (Taylor & Vermeulen Taylor:96 (1996)). Near infrared (NIR) annular photometry shows that the nuclear K-band light is dominated by non stellar emission which could be either thermal emission from hot dust, or optically thin synchrotron radiation (Longmore et al. Longmore:84 (1984)). Recent HST observations in the UV show that any nuclear engine/star cluster must be $`<`$17 pc in size (Maoz et al. Maoz:95 (1995)). Our goal in this work is to better constrain the many existing models of NGC 1275 with near infrared (NIR) imaging spectroscopy and spectral synthesis modeling. In dusty environments, NIR imaging has optical depth penetration 10 times deeper than visible wavelengths, and NIR spectroscopy is sensitive to a wealth of diagnostic lines which describe the physical state of the gas and stars in the nucleus. Throughout this paper we assume a scale of 344 pc ″<sup>-1</sup> (H<sub>0</sub> = 75 km s<sup>-1</sup> Mpc<sup>-1</sup>, z = 0.0172). ## 2 Observations and data reduction We observed NGC1275 using the Max-Planck-Institut für extraterrestrische Physik (MPE) imaging spectrometer 3D and the Max-Planck-Institut für Astronomie (MPIA) tip-tilt image tracker “CHARM” on 1995 January 15, 16 and 21 at the 3.5m telescope of the Max-Planck-Institut für Astronomie at Calar Alto, Spain. The CHARM instrument (McCaughrean et al. McCaughrean:94 (1994)) removes atmospherically induced image motion and produces stable images of increased spatial resolution. We locked CHARM on the bright point-like nucleus of NGC1275, and corrected at 5 Hz. Under seeing conditions varying from 1.0 to 1$`\stackrel{}{.}`$7, we achieve a final overall image quality of FWHM 1$`\stackrel{}{.}`$5 as determined by fitting to the point-like nucleus (Maoz et al. Maoz:95 (1995)). The “3D” instrument (Krabbe et al. Krabbe:95 (1995), Weitzel et al. Weitzel:96 (1996)) is an field imaging spectrograph which provides, during in a single exposure, H or K band data cubes with spectral resolution of R $`=\lambda /\mathrm{\Delta }\lambda =1000`$and$`2000`$. The two spatial cube dimensions are 16 pixels on a side, with pixel sizes selected to be 0.5″ (giving an 8″$`\times `$8″ field of view), while the third dimension of 256 pixels is spectral. The intrinsic spectral sampling of the camera is with pixels of size $`\lambda /`$R, so by dithering the spectral sampling by 1/2 pixel on alternate data sets, we achieve fully Nyquist sampled spectra. The spectral resolution, achieved with the current reduction procedure, is 670, or 450 km s<sup>-1</sup>; future software will produce full R = 1000 spectra after co-addition. For extended objects, the efficiency gain over a long slit spectrograph is substantial, as at any point within our 8″$`\times `$8″ field, it is possible to obtain the full band object spectrum. The data (Table 1) frames were interleaved with identical sky exposures 60″ away to the West. The source observations were preceded and/or followed immediately by a set of spectral calibration observations of nearly featureless stars of known brightness. The few known spectral features in the reference star spectra (e.g. Br$`\gamma `$ in K band and Br9 – Br11 in H band) were later removed by interpolation. A temporally variable deep CO<sub>2</sub> absorption near 2.01 $`\mu `$m made accurate sky subtraction in this regime difficult. In addition, the temporal variability of OH lines in the H band produces spurious sky noise which dominates the background across the band. A run-down on typical 3D data reduction with the Groningen GIPSY software is given in Weitzel et al. (Weitzel:96 (1996)). Spectral baselines were calibrated using the color temperature of the calibrator stars and a featureless Nernst glower. The absolute wavelength calibration is accurate to better than 1/4 pixel. Individual data sets were properly weighted, registered onto the central peak of NGC 1275, and co-added. The K and H band spectra of the nucleus within a 3″ aperture are presented in Figs. 1 and 2, which present the enormous variety of emission and absorption lines. The signal to noise in the K-band spectrum is $``$ 70 in the continuum. For the H band, the signal to noise is $``$ 25 in the continuum. ## 3 Nature of the continuum emission Fig. 3 represents the true, line emission free, K band continuum structure of NGC 1275. Its peaking at the nucleus confirms that the bulk of the emission arises in the central 300 pc. The slight south-east north-west extension, almost along the radio PA of 160 (Pedlar et al. Pedlar:90 (1990)), is 2$`\sigma `$ significant. We analyzed the the continuum spectrum with respect to contribution from stars, the active galactic nucleus, hot dust from the circumnuclear environment, and free-free emission. K and M stars (T$`{}_{\text{eff}}{}^{}4000`$), dominating the K band stellar emission, show strong CO absorption features in their more or less BB spectrum which can be used to constrain the fraction of total light originating in stars. NIR spectra from AGN show a power law spectrum of the form I$`{}_{\lambda }{}^{}\lambda ^\alpha `$ where $`\alpha `$ varies between $`0.5`$ and $`1.5`$ (Laor & Draine Laor:93 (1993), Acosta-Pulido et al. Acosta-Pulido:90 (1990)). Hot dust in this context is mainly referred to as being reprocessed radiation from AGN and, to a minor extent, from starburst activity. It’s spectrum is that of a black body multiplied by a power-law emissivity of the form $`ϵ_{\text{dust}}\lambda ^1`$. From these contributions a synthetic galaxy spectrum including extinction effects has been modelled and (least square) fitted to the observed continuum spectrum. Details of the procedure are given in the appendix, contrains on the parameters are being discussed below, and the results are presented in Sect. 3.2. ### 3.1 Physical constrains on the continuum emission The size of the search space was reduced by placing constrains on the total extinction, Galactic extinction, free-free contribution, dust temperature, and stellar populations. *i*) The Galactic extinction toward NGC 1275 is A$`_\text{B}`$ = 0.7, or A$`_\text{V}`$ = 0.53, and there is an additional component of extinction intrinsic to NGC 1275 itself. From UV measurements we have E(B-V) = 0.2 – 0.3 (Maoz et al. Maoz:95 (1995), Levinson et al. Levinson:95 (1995)), which both suggest that A$`{}_{\text{V}}{}^{}1.0`$ at the nucleus. Integrated over the central stellar area, H–K is 0.15 redder than typical ellipticals (Longmore et al. Longmore:84 (1984)), which requires A$`{}_{\text{V}}{}^{}2.4`$. Hence we constrain the V band extinction to A$`{}_{\text{V}}{}^{}<2.0`$, or A$`{}_{\text{K}}{}^{}<0.2`$ (Rieke & Lebofsky Rieke:85 (1985)). Since we know that the central light is likely to be dominated by the AGN (Longmore et al. Longmore:84 (1984)), we choose to use a screen model of extinction there, while farther out at distances $`\stackrel{>}{}`$ 500 pc we will use a mixed stars/gas model. *ii*) The free-free continuum can be obtained from the Br$`\gamma `$ line flux using case B recombination and assuming T = 10000 K. (Joy & Lester Joy:88 (1988)) have shown that this method gives consistent results compared to a NIR multicolor continuum decomposition (Joy & Harvey Joy:87 (1987)) and radio measurements (Wynn-Williams & Becklin Wynn:86 (1986)). Using the Br$`\gamma `$ line flux from Table 2 we can thus constrain the free-free contribution to be about $`2\%`$ of the underlying continuum. There is some suggestion that the free-free emission could be higher in NGC 1275 because the central plasma is constantly being replenished with free electrons via the cooling flow gas. However, if there were a substantial population of such ions, they would contribute to the scattering of the central source at the 1% level (Sarazin & Wise Sarazin:93 (1993)), which would in turn produce a blue excess. If the 1% scattering is responsible for the observed nuclear blue excess of $`10^{44}`$ erg s<sup>-1</sup> (Crawford & Fabian Crawford:93 (1993)), then the central source would require an X-ray/UV luminosity of roughly 10<sup>46</sup> erg s<sup>-1</sup>, which is $`600`$ times the point X-ray luminosity. Hence we set the free-free emission to a typical value of $`<3\%`$ of the total continuum, or sometimes ignore it completely, as our fits constrain the relative contributions to only $`\pm 10\%`$. Had we not done so, the free-free and AGN emission would have traded off influence with each other, as they are both power laws of similar slopes. *iii*) Hot dust has turned out to be very common in the nuclear regions of AGNs (Granato et al. Granato:97 (1997), Maiolino et al. Maiolino:95 (1995)) and, to some extent, in starburst galaxies as well (see e.g. Böker et al. Boeker:98 (1998)). It surrounds the nucleus forming a torus which absorbs radiation from the AGN and reemits it in the mid-infrared. Since hot dust too close to the AGN will start to sublime, the maximum dust temperature is set by the sublimation temperature of the dust material. Silicate and graphite grains, which are the main constituents of interstellar dust (Laor & Draine Laor:93 (1993)), will sublime at temperatures of about 1400K and 1750K resp., setting the limit for the the dust temperature at about 1500K at the inner edge of the dust torus. The minimum radius for dust grains to survive is $`\stackrel{>}{}`$ 0.2L$`{}_{}{}^{1/2}{}_{46}{}^{}`$ pc (Laor & Draine Laor:93 (1993)), where L<sub>46</sub> is the central source bolometric luminosity in units of $`10^{46}`$ erg s<sup>-1</sup>. For NGC 1275 with L<sub>46</sub> = 0.04 (Levinson et al. Levinson:95 (1995)), this is $`\stackrel{>}{}`$ 0.05 pc. The dust temperature observed will be a mixture of dust at different temperatures along the line of sight weighted with their column density distribution. Thus the effective temperature will be lower and, e.g. for NGC 1068, in the range of 700 K (Thatte et al. Thatte:97 (1997)). As NGC 1275 is a F–R I type radio source, we are probably looking nearly pole on at the central source, and hence see the largest possible fraction of the hot inner surfaces of the clouds which make up an obscuring torus (Pier & Krolik Pier:92 (1992)), suggesting that we limit the temperature of the dust in our simulations to a relatively high value of 1000 K. *iv*) The NIR flux of essentially all integrated galaxy populations is dominated by giants and, to a lesser extent, by supergiants (Oliva et al. Oliva:95 (1995)). We rule out K5 V stars because for all reasonable stellar populations they are too faint to contribute substantially to the overall H and K band fluxes. As an underlying stellar population we choose K5 III stars because they dominate, to first order, the light of nearly all galaxies at H and K bands (Frogel et al. Frogel:78 (1978), Oliva et al. Oliva:95 (1995)). This is due to the fact that their effective temperatures of $`4000`$ K have a black body emission peak in those bands, and because they are typically 200 times brighter than K5 V stars with similar effective temperatures. Digitized libraries of NIR stellar spectra are available in the K band Kleinmann:86 (1986), so we used the spectrum of the K5 III star $`\gamma `$ Dra as our template I$`{}_{}{}^{\text{STAR}}{}_{\lambda }{}^{}`$ in the K band. In the H band, no such digitized stellar library was publicly available at the time of this data reduction, so we constructed a continuum spectrum appropriate to the stellar population of NGC 1275 by summing the contributions of many black body curves of different temperature scaled by the fractional number of stars having that temperature. This produces a continuum spectrum without any lines. The H band CO absorption features are relatively weak, so they would have contributed little to the fitting anyway. The galaxy population, necessary for this procedure, was modeled with the stellar population evolution code STARS (Sternberg & Kovo, in preparation). Given a set of input parameters it determines the binned number density function of stars in the HR-diagram. Summing the series of appropriately scaled black body curves at the various temperatures results in a H-band continuum that approximates the underlying H band continuum of NGC 1275, if the input parameters to STARS are varied adequately. The composite spectrum is then normalized to the same wavelength as the K band stellar library spectrum in order to produce the final H and K band underlying stellar spectrum, I$`{}_{}{}^{\text{STAR}}{}_{\lambda }{}^{}`$. All K spectra are normalized to a wavelength of 2.27 $`\mu `$m, and all H spectra are normalized to 1.57 $`\mu `$m, which we chose because the spectra are free of lines there. ### 3.2 Decomposing the continuum emission In the central 1″ (0.3 kpc) region we expect that the contribution of the AGN overwhelms the the stellar population. This is indeed the case, as the continuum emission from the center of NGC 1275 is well modeled by the light of K5 III or K5 Ib stars heavily diluted by emission from hot dust and AGN typical power-law emission (Fig. 4). The best fitting values are $`c_{\text{STAR}}=0.0\pm 0.1`$, $`c_{\text{PL}}=0.5\pm 0.1`$, $`c_{\text{HD}}=0.5\pm 0.1`$, $`T_\text{d}=700\pm 200`$, $`\tau _\text{K}=0.05\pm 0.1`$, where the $`c_\text{i}`$ represent fractional contributions and were the errors are those derived from the estimated systematic errors in the spectra and from multiple trials with slightly varied weights and initial guess parameters. The relative insignificance of spectral features in the underlying stellar population shows that the fitting process at the nucleus is dominated by the power laws, and that we are unable to distinguish between the various possibilities of underlying populations. The relative contribution of power-law emission is not particularly sensitive to the adopted AGN spectral index (-0.5 through -1.5, see appendix), because its main function is to dilute the very steep slope of the stellar continuum (very nearly a black body). In contrast, the same fitting process applied to a 1″ wide annulus 3″ (0.9 kpc) from the nucleus yields completely different result. Because the CO absorption bandheads are clearly observed in the K band spectrum (they are much less diluted), we fit this region only using the K band data. The fit is excellent (Fig. 5), and yields the following contributions: $`c_{\text{STAR}}=0.6\pm 0.1`$, $`c_{\text{PL}}=0.4\pm 0.1`$, $`c_{\text{HD}}=0.0\pm 0.1`$, $`\tau _\text{K}=0.01\pm 0.1`$, where we have chosen to apply a mixed stars/gas model of extinction rather than a simple screen. Again, we assume a K5 III star dominated underlying stellar population. Note that the AGN contribution could also be interpreted as free-free emission at this distance from the nucleus, where we do not expect any direct or scattered contribution from the AGN. By the time we reach 4″ (1.2 kpc) from the nucleus, the stellar component dominates completely, and we find $`c_{\text{STAR}}=0.90\pm 0.1`$, $`c_{\text{PL}}=0.1\pm 0.1`$, $`c_{\text{HD}}=0.0\pm 0.1`$, $`\tau _\text{K}=0.01\pm 0.1`$. Hence the behavior of the the various components is as expected from a central AGN dominated source. ## 4 The H<sub>2</sub> emission The H<sub>2</sub> emission in NGC 1275 is exceptionally luminous (Fischer et al. Fischer:87 (1987)) even if compared with other powerful H<sub>2</sub> emitting galaxies. The excitation of H<sub>2</sub> in AGNs is usually explained to be a mixture of shocks (Brand et al. Brand:89 (1989)), UV photons (Sternberg & Dalgarno Sternberg:89 (1989)), and X-rays (Draine & Woods Draine:91 (1991)). The shocks can be caused by galaxy merging or by supernovae, the UV radiation could come from OB star associations (Puxley et al. Puxley:90 (1990)), while only an AGN with a black hole can produce the quantities of X-ray radiation needed to excite the H<sub>2</sub> seen in Seyfert nuclei. ### 4.1 Spatial distribution of H<sub>2</sub> emission The H<sub>2</sub> line emission (Fig. 6) is strongly dominated by an unresolved source centered on the nucleus surrounded by weak extended emission extending a few arcseconds west. The fact that the primary exciting source is not distributed over scales $`\stackrel{>}{}`$ 300 pc suggests that most of the H<sub>2</sub> emission is associated with the circumnuclear environment and probably the active galactic nucleus and not, as speculated before (Inoue et al. Inoue:96 (1996)) with the cooling flow. The low level extended emission reach $`2\%`$ of the peak flux at our limiting distance of 5″ or 3.5 kpc (Fig. 7). The total H and K band H<sub>2</sub> emission amounts to $`1.3\times 10^{13}`$ erg s<sup>-1</sup> cm<sup>-2</sup> (see Table 2) or, for a distance of 70 Mpc, a total luminosity of L$`{}_{\text{H}_2}{}^{}=8\times 10^{40}`$ erg s<sup>-1</sup>. ### 4.2 Excitation of the H<sub>2</sub> emission The high signal-to-noise in our spectra allow us to check the effective excitation temperature from the column density of molecules in each quantum state (see e.g. Beckwith et al. Beckwith:78 (1978)): $$N_{col}=4\pi I/A_{ul}h\nu $$ (1) where $`I`$ is the surface brightness in the line and $`A_{\text{ul}}`$ is the Einstein A coefficient. The surface brightnesses have been dereddened using A$`{}_{\text{V}}{}^{}=2.0`$ (Sect. 3.1) as a realistic value. Screen and mixed extinction models (McLeod et al. McLeod\_al:93 (1993)) for the dereddening gave flux correction between 10% and 20%. Since the appropriate model was uncertain, the correction was averaged between screen and mixed model and the uncertainties propagated in the errors. In thermal equilibrium, the line ratios should be given solely by the number of atoms in each quantum state and their spontaneous emission coefficients if the H<sub>2</sub> is optically thin. From Table 3 we conclude that this is indeed the case. The column densities are always lower than N$`_{\text{col}}`$(H<sub>2</sub> 1–0 S(1))$`=9.6\times 10^{15}`$ cm<sup>-2</sup>. Summing over all the observed H<sub>2</sub> lines gives N$`_{\text{col}}`$(H<sub>2</sub>$`{}_{}{}^{\text{hot}})3\times 10^{16}`$ cm<sup>-2</sup>. If we assume that this gas is uniformly distributed within the emitting region, then the total mass of contributing H<sub>2</sub> would be $`1000`$ M. In Fig. 8 we plot the natural log of N$`_{\text{col}}`$ per mode against the energy of the upper level given as a temperature (compare with Kawara & Taniguchi Kawara:93 (1993)). The error bars include the errors given in Table 1 and the dereddening. The inverse slope is the effective excitation temperature T$`_{\text{ex}}`$ of the gas and corresponds to the kinetic temperature if local thermal equilibrium (LTE) prevails. Depending on the conditions within a molecular cloud, the initial ortho/para ratio of 3 of molecular hydrogen at formation may evolve and eventually drop to lower values (Martini et al. Martini:97 (1997)). Exchange reactions with H and H<sup>+</sup> (Flower & Watts Flower:84 (1984)) or a low temperature environment (T $`\stackrel{<}{}`$ 200 K; Burton et al. Burton:92 (1992)) can even lower the ratio to below 2.0 (see e.g. Hora & Latter Hora:96 (1996)). A high temperature environment in turn will keep the ortho/para ratio at 3. In a photodissociation region (PDR), however, freshly exposed H<sub>2</sub> may keep the ortho/para ration down (Chrysostomou et al. Chrysostomou:93 (1993)) although the environment is hot. In Fig. 8 the column densities have been divided by the statistical weights corresponding to their states which include the spin degeneracies g$`{}_{\text{s}}{}^{}=1`$ for even (para) and g$`{}_{\text{s}}{}^{}=3`$ for odd (ortho) J. If the ortho/para ratio in the molecular hydrogen is 3:1 then the ortho and para column densities should lie along a line. For the 1–0 transitions, were we have enough data points, this is indeed the case. This demonstrates that the either the molecular hydrogen has just been formed or that the H<sub>2</sub> has been kept in a high temperature environment. The distribution of data in Fig. 8 shows that the H<sub>2</sub> gas is nearly isothermal. Below 8000 K upper level energy, the excitation temperature is T$`{}_{\text{ex}}{}^{}1500`$ K, but above 8000 K, T$`_{\text{ex}}`$ does only vary between $`2600`$ K and $`2900`$ K. If no extinction correction is applied, the temperatures come out to be 1450 K, 2400 K, and 2650 K resp., which is still in agreement with the 1$`\sigma `$ errors given in Fig. 8. All date (except for 2–1 S(3), see below) out to level energies of $`20000`$ K lie along an only slightly curved line. In particular, the column density per mode of 1–0 S(7) is close to those of 2–1 S(1) and 2–1 S(2), indicating that the rotational and vibrational temperatures are about the same and the fluorescent excitation of the gas is rather small (Draine & Bertoldi Draine-Bertoldi:96 (1996)). The proposed likely mechanism to populate the levels of this hot H<sub>2</sub> gas with T$`{}_{\text{ex}}{}^{}2700`$ K is collisional excitation in a post shock environment, which has been observed in many galactic sources (see e.g. Fischer et al. Fischer:87 (1987), Gredel et al. Gredel:94 (1994), Wright et al. Wright:96 (1996)). However, at column densities prevailing in the central 300 pc of NGC 1275, a PDR scenario cannot be completely excluded. In dense PDRs, the vibrational ladder is thermalized while the rotational ladder in a given vibrational state may be out of thermal equilibrium (see e.g. Timmermann et al. Timmermann:96 (1996)). The fact that we do not see gas of temperature above 5000 K (Wright et al. Wright:96 (1996)) may, however, indicate that the contribution of nonthermal processes (UV pumping) to the H<sub>2</sub> excitation is rather moderate. The excitation conditions in NGC 1275 can be investigated in more detail by examining the ratio of two ortho or para lines and put them in context. The remaining uncertainty introduced by the intrinsic ortho-para ratio of the gas clouds can thus be avoided (Mouri Mouri:94 (1994)). Locating these ratios in a ortho-para line ratio plane provides an independent guide for distinguishing between the relative contribution of X-ray heating, shocks, and UV pumping excitation of H<sub>2</sub> gas in starburst and AGN galaxies (Fig. 9). As non thermal processes (e.g. UV pumping in low density gas) excite high vibrational states more easily than the thermal process, the ratio of two ortho transitions such as R$`_\text{o}`$ = H<sub>2</sub> 2–1 S(1)/H<sub>2</sub> 1–0 S(1) gives an ortho-para balance independent measure of the relative contribution of thermal and non-thermal processes. The same holds for R$`_\text{p}`$ = H<sub>2</sub> 1–0 S(2)/H<sub>2</sub> 1–0 S(0). The expected value of R$`_\text{o}`$ in the case of pure non thermal excitation under nearly all realistic conditions is R$`{}_{\text{o}}{}^{}=0.56`$ (Black & van Dishoeck Black:87 (1987)), while thermal UV excitation gives a value of R$`_\text{o}`$ near 0 (Sternberg et al. Sternberg:87 (1987)). From Table 2, we find R$`{}_{\text{o}}{}^{}=7.7/41.1=0.19`$ and R$`{}_{\text{p}}{}^{}=14.4/10.0=1.4`$, which again, as we noted earlier, puts NGC 1275 in the R$`_\text{o}`$ – R$`_\text{p}`$ plane close to the location of NGC 6240 (Fig. 9). Starburst galaxies like NGC 253 and galactic photodissociation regions generally show R$`{}_{\text{p}}{}^{}1.0`$. Following this method one can also define a ratio, R$`_\text{o}`$ = H<sub>2</sub> 1–0 S(3)/H<sub>2</sub> 1–0 S(1), of two ortho lines which is again independent of the initial ortho-para ratio in the molecular clouds. When placed in this diagram (Fig. 9), NGC 1275 falls near other powerful AGNs like NGC 1068. The interpretation of these diagrams tells us that the excitation is mostly “thermal”, with a small component appearing as “nonthermal emission”. The gas must then be quite dense, with n$`{}_{\text{T}}{}^{}\stackrel{>}{}\mathrm{\hspace{0.17em}10}^5`$ cm<sup>-3</sup>. The placement of NGC 1275 near other powerful AGN and merger galaxies suggests that both X-ray and shock mechanisms may be at work creating the powerful H<sub>2</sub> emission. Hence, based on the spatial extent of the H<sub>2</sub> emission and its temperature, we propose that there are two distinct sources of H<sub>2</sub> excitation: the AGN nucleus, which excites the vast majority of the radiation, and which produces the compact emission core seen in our maps, and the extended region of moderate star formation which excites the highly extended, weak H<sub>2</sub> envelope. ### 4.3 H<sub>2</sub> 2–1S(3) line strength The column density of the H<sub>2</sub> 2–1 S(3) line is lower by a factor of 3 compared to what could be expected from weighted extrapolation of the 2–1 S(1) and 2–1 S(2) transitions. Even if the ortho/para ratio within the 2–1 transitions is not close to 3, which it was in the 1–0 transitions, N$`_{\text{col}}`$(2–1 S(3)) is still a factor of $`>2.5`$ lower compared to N$`_{\text{col}}`$(2–1 S(1)) and using T$`{}_{\text{ex}}{}^{}=2600`$ K for the two ortho states. Inspecting the original data and the spectrum in Fig. 1 did not show any abnormalities. Since the chance for blending the line with a strong absorption can be excluded for this part of the spectrum and the atmosphere, we argue that this weakness of the H<sub>2</sub> 2–1 S(3) line is real. Inspecting Fig. 5 reveals that the 2–1 S(1) line is also weak outside the nucleus at a radius of $`1`$ kpc. It would be visible if it was 2.5 time stronger or about half the strength of the 1–0 S(2) line, which is clearly present. The only other galaxy, were a weak H<sub>2</sub> 2–1 S(3) line has been reported so far, is NGC 6240 (Lester et al. Lester:88 (1988)), although only as an upper limit. From their Table 3 one can estimate that the column density of 2–1 S(3) line in NGC 6240 is at least 3.5 times weaker than expected from extrapolating other data, which is in good agreement with our finding and suggest that there may be a common mechanism at work. The column densities in NGC 6240 are only a factor of 2 higher compared with NGC 1275 an their derived temperature range is almost identical with Fig. 8. It should be noted, however, that the excitation conditions in both galaxies are not identical. Draine & Woods (Draine:90 (1990)) have shown that in addition to X-ray heating, UV pumping plays a major role in NGC 6240. This is already indicated in Fig. 9 as well as and in Fig. 4 of Lester et al. (Lester:88 (1988)): The large offset between the line through the 1–0 transitions and the 2–1 data is different from NGC 1275 (Fig. 8). Recently, Sugai et al. (Sugai:97 (1997)) reported that they were unable to confirm the weakness of the H<sub>2</sub> 2–1 S(3) line in NGC 6240 with R = 580 grism spectroscopy. Better quality data at higher spectral resolution is certainly required to clarify the situation. The most likely mechanism for the weakness of the H<sub>2</sub> 2–1 S(3) line is resonant fluorescent excitation of the upper level ($`\nu `$ = 2, J= 5) by Ly$`\alpha `$ photons (Black & van Dishoeck Black:87 (1987)). The Ly$`\alpha `$ photons needed for this pumping process can easily be provided in an environment where a notable fraction of the H<sub>2</sub> is excited by X–ray heating (Black & van Dishoeck Black:87 (1987)). Although detailed models of resonant fluorescent excitation of the $`\nu =2,J=5`$ level does not exist, the weakness of the H<sub>2</sub> 2–1 S(3) line by itself already demonstrates that at least in the nuclear region the excitation of the molecular hydrogen in NGC 1275 is dominated by the central AGN. This is in agreement with our previous finding that the excitation environment in the nuclear region of NGC 1275 is much more dominated by X-ray heating than in NGC 6240 (Draine & Woods Draine:90 (1990)). NGC 1275 may be the first galaxy for which a more complete model on resonant fluorescent excitation can be computed. Finding other galaxies with such a weak H<sub>2</sub> 2–1 S(3) line (e.g. in NGC 1068) would be important for constraining the process, which may provide an additional tool for understanding the radiative interaction between AGN and circumnuclear molecular material. ## 5 Br$`\gamma `$ emission The Br$`\gamma `$ emission from NGC 1275 is unusual in that it is highly asymmetric both in its line profile (Fig. 1), and in its spatial distribution (Fig. 10). The Br$`\gamma `$ emission is extended along roughly the same line (PA =160) as the powerful radio jets (Pedlar et al. Pedlar:90 (1990)), which suggests that the mechanism of its ionization is more strongly influenced by the central engine than by the surrounding stellar population. This result correlates well with H$`\alpha `$ and \[NII\] maps with visible band imaging spectroscopy (Ferruit & Pecontal Ferruit-tiger:94 (1994)). In particular, the position of the extension in our maps correlates with an H$`\alpha `$ and \[NII\] emission peak located $`1\stackrel{}{.}2`$ N, 0$`\stackrel{}{.}`$7 W of the nucleus. From the published H$`\alpha `$ maps, we estimate the H$`\alpha `$ flux in the hot spot to be $`1.3\times 10^{17}`$ W m<sup>-2</sup> arcsec<sup>-2</sup>. Based on the observed \[SII\] line ratios and assuming a temperature of T$`{}_{\text{e}}{}^{}=10^4`$ K, (Ferruit & Pecontal Ferruit-tiger:94 (1994)) derive a mean electron density of n$`{}_{\text{e}}{}^{}=40`$ for this region, and a typical linewidth of 340 km s<sup>-1</sup>. The intrinsic Br$`\gamma `$ linewidth at the nucleus of $`980\pm 30`$ km s<sup>-1</sup> is unusually large for a pure star forming source. However, the width is not unexpected for emission from NLRs associated with AGNs and we therfore safely assume here that the NLR contributes a substatial fraction to the Br$`\gamma `$ emission. The larger linewidth of Br$`\gamma `$ compared to H$`\alpha `$ is probably a differential extinction effect. The emission was too faint to derive accurate Br$`\gamma `$ line profiles in regions 1″ to the NW and SE of the nucleus in order to test whether the excess width and shape of the Br$`\gamma `$ line is due to expansion. The total Br$`\gamma `$ emission within the central 1 kpc is $`9.7\times 10^{15}`$ erg s<sup>-1</sup> cm<sup>-2</sup>, which is a total luminosity of $`1.5\times 10^6`$ L. This corresponds to an ionization rate of $`3.2\times 10^{53}`$ s<sup>-1</sup> or $`2.0\times 10^9`$ L for T$`{}_{\text{e}}{}^{}=7500`$ K, and can be compared with the value of $`1.1\times 10^{54}`$ s<sup>-1</sup> inferred from M82 over a similar region from H53$`\alpha `$ recombination lines (Puxley Puxley:91 (1991)). From the ionization rate one can infer the total Lyman continuum luminosity L$`_{\text{Lyc}}`$ if one assumes OB type stars as its source. L$`{}_{\text{Lyc}}{}^{}=2.0\times 10^9`$ L yields about 1.5% of the L$`_{\text{FIR}}`$ of NGC 1275 ($`1.5\times 10^{11}`$ L) or L$`_{\text{FIR}}`$/ L$`{}_{\text{Lyc}}{}^{}75`$. L$`{}_{\text{Lyc}}{}^{}=2\times 10^9`$ L is a lower limit to the Lyc-luminosity in the IRAS beam. The total bolometric luminosity of such an evolving stellar cluster is typically about 10 - 20 times higher than the Lyman continuum luminosity for typical assumptions on the properties of such a region if it were an active star forming region powered by massive stars (age $``$ a few 10<sup>7</sup> y, M$`{}_{\text{lower}}{}^{}=0.1`$ M, M$`{}_{\text{upper}}{}^{}=120`$ M, see Fig. 10 in Krabbe et al. Krabbe:94 (1994)). With that ratio L$`{}_{\text{Bol}}{}^{\text{ionizing}}3.8\times 10^{10}`$L, which is about 25% of L$`_{\text{FIR}}`$. Taken as face values these numbers are not very consistent with star formation but much more with what one gets empirically, using above formulas for AGNs (Genzel et al. Genzel:98 (1998)). ## 6 The iron emission ### 6.1 Strength and morphology of the \[Fe ii\] emission The \[Fe ii\] 1.644 $`\mu `$m emission from NGC 1275 is very strong (Fig. 2) and centrally concentrated (Fig. 11). Other prominent H and K band iron lines include the \[Fe ii\] 1.534 $`\mu `$m, \[Fe iii\] 2.2178 $`\mu `$m, and \[Fe iii\] 2.2420 $`\mu `$m, in a near blend with H<sub>2</sub> 2–1 S(1). The flux of the \[Fe ii\] 1.644 $`\mu `$m (Table 2) is nearly as large as the \[Fe ii\] 1.257 $`\mu `$m in an 8″ aperture by (Rudy et al. Rudy:93 (1993)). The \[Fe ii\] emission is much stronger compared to other starburst/Seyfert indicators than it is in most active galaxies. For example, in the central 1 kpc of NGC 1275, we find \[Fe ii\]($`\lambda `$1.644)/Br$`\gamma `$ = 6.3, whereas in roughly similar sized regions in the merger driven starburst NGC 3256 \[Fe ii\]/Br$`\gamma `$ = 1.55 and in the starburst/Seyfert systems NGC 4945 \[Fe ii\]/Br$`\gamma `$ = 1.4 (Moorwood & Oliva Moorwood:94 (1994)) and NGC 1068 \[Fe ii\]/Br$`\gamma `$ = 2.5 (Thatte et al., in preparation; Blietz et al. Blietz:94 (1994)). While Br$`\gamma `$ originates predominantly from photoionization of H ii regions by OB stars, \[Fe ii\] has an additional source. Most of the gas phase Fe in the local ISM condenses onto grains, so that the free Fe is often depleted by factors of 100 – 1000. If we presume that the ionization characteristics of the central source in NGC 1275 are similar to those of NGC 3256, NGC 4945, and NGC 1068, then the unusual \[Fe ii\]/Br$`\gamma `$ and \[Fe ii\]/Pa$`\beta `$ ratios (e.g. Rudy et al. Rudy:93 (1993)) would suggest that the iron depletion is $`\stackrel{<}{}`$ 10 in NGC 1275, less severe than in most starburst and active galaxies. This could be due to an ambient radiation field which heats the dust grains and tends to evaporate the Fe off of them. Such a scenario is supported by the relatively high derived dust temperature within the central 300 pc (e.g., Sect. 3.2). Alternatively, the fact that the ionization state of the gas must allow for plentiful Fe<sup>+</sup>, which can then be further ionized by photons with $`E>16.2`$ eV, requires that the ionizing source of radiation must be relatively cool, in order to produce low ionization radiation which would produce the high \[Fe ii\]/Br$`\gamma `$ line ratios. However, fast J-shocks and X-ray heating can also produce copious \[Fe ii\] emission (Blietz et al. Blietz:94 (1994)). In the latter case, the X-rays penetrate deep into molecular clouds where they create large reservoirs of partially ionized gas. #### 6.1.1 Profile and excitation of the \[Fe ii\] emission The profile of the \[Fe ii\] 1.644 $`\mu `$m line cannot be well modeled as a single Gaussian. This is typical of Seyfert galaxies, which often show narrow and broad components of the same emission feature. We therefore fit the observed profile using two Gaussians, a broad component of width $`1340\pm 120`$ km s<sup>-1</sup> and relative amplitude of $`0.26`$, and a narrow component of of width $`280\pm 20`$ km s<sup>-1</sup>. This two component fit (Fig. 12), is excellent, and suggests that $`50\%`$ of the \[Fe ii\] emission arises in the same region as the Br$`\gamma `$, showing a similar linewidth, while the other half of the \[Fe ii\] arises in the same region as the H<sub>2</sub>, where the linewidths are again similar. Hence we propose — as in the case of the H<sub>2</sub> lines (Sect. 4.2) — that the active nucleus of NGC 1275 is responsible for much of the excitation leading to \[Fe ii\] emission. Comparing this result with Blietz et al. (1994), who have found an excellent correlation between the \[Fe ii\] emission and the NLR in NGC 1068 suggests, that we might have a similar case here but unresolved and at much greater distance. The detailed line ratios expected from excitation via a central AGN have been modeled by Hollenbach & Maloney (Hollenbach:96 (1996)). In their model, the X-ray energy absorbed per unit time per hydrogen nucleus in a cloud at distance 100r<sub>100</sub> pc is H$`{}_{\text{X}}{}^{}7\times 10^{22}`$ L<sub>44</sub>r$`{}_{}{}^{2}{}_{100}{}^{}`$N$`{}_{}{}^{1}{}_{22}{}^{}`$ erg s<sup>-1</sup>, where $`10^{44}`$L<sub>44</sub> erg s<sup>-1</sup> is the AGN luminosity and $`10^{22}`$N<sub>22</sub> cm<sup>-2</sup> is the total column density. The heating rate is $`0.3`$ – 0.4 H$`_\text{X}`$, and the cooling is due to line emission from many species, but of particular diagnostic interest are the H<sub>2</sub> 1–0 S(1) and \[Fe ii\] 1.64 $`\mu `$m lines. Using our value of N$`{}_{\text{H}_2^{\text{hot}}}{}^{}=3\times 10^{16}`$ cm<sup>-2</sup>, a molecular abundance of H/H<sub>2</sub> $`10^5`$, and L$`{}_{\text{X}}{}^{}2\times 10^{43}`$ erg s<sup>-1</sup> (Prieto Prieto:96 (1996)), and integrating the Hollenbach & Maloney results over our resolution beam of 300 pc, we can account for the observed \[Fe ii\] to H<sub>2</sub> 1–0 S(1) line ratio. For all radii larger than our beam, the predicted ratio is $`1:1`$. This is close to our measured value of 1.5:1 in a 3″aperture (Table 2) since the predicted ratio depends on the presumed depletion of Fe, which we already know to be $`\stackrel{<}{}\mathrm{\hspace{0.17em}10}`$, and which the model presumes to be $`30`$ (Hollenbach, private communication). We thus conclude that the \[Fe ii\] emission as well as \[Fe ii\]/H<sub>2</sub> 1–0 S(1) can be explained by an X–radiation field typically emitted by AGN. ## 7 The CO$`_{\text{sp}}`$ index The CO$`_{\text{sp}}`$ index of the underlying stellar population can be diluted by continuum emission not originating in stars. If the non-stellar continuum is $`\chi `$ times stronger than the stellar continuum, the CO$`_{\text{sp}}`$ index will be diminished according to $$CO_{sp}(\chi )=2.5log\frac{S_0+\chi }{1+\chi }$$ (2) where S$`{}_{0}{}^{}=10^{0.4\text{CO}_{\text{sp}}(0)}`$ is the average of the undiluted rectified spectrum between 2.31 and 2.40 $`\mu `$m and CO$`_{\text{sp}}`$ is the spectroscopic index. Details are given in Doyon et al. (doyon-co:94 (1994)). According to our model fitting, the non-stellar continuum in the central 1″ contributes $`\stackrel{>}{}\mathrm{\hspace{0.17em}90}\%`$ of the emission, reducing the observed CO$`_{\text{sp}}`$ from its intrinsic value of $`0.3`$ to $`0.03`$, and and giving rise to the CO$`_{\text{sp}}`$ hole at the nucleus of NGC 1275 (Fig. 13). The range for the allowed values of the intrinsic CO$`_{\text{sp}}`$ is indicated in Fig. 14. The nuclear stellar cluster may have a higher CO index than initially assumed, implying that the K supergiants with CO$`{}_{\text{sp}}{}^{}0.3`$ instead of giants may be the dominating population (Doyon et al. doyon-co:94 (1994)). Including this finding into a reiteration on the continuum decomposition will not have any effect on the results, since the temperature of these stellar types are not very different. Outside of the hole, the CO$`_{\text{sp}}`$ of 0.17 is typical of E/S0 galaxies (Frogel et al. Frogel:78 (1978)), confirming that the stellar population of NGC 1275 is quite normal. Furthermore CO$`{}_{\text{sp}}{}^{}=0.17`$ corresponds to a K5 III star (Doyon et al.doyon-co:94 (1994), appendix A), which corroborates our spectral synthesis results for regions at R $`\stackrel{>}{}`$ 3 ″. Given that the light is dominated by K giants or supergiants, we try to estimate the total nuclear stellar mass. Using our result that the emission enclosed in the central 3″ (1 kpc) of NGC 1275 is $`\stackrel{>}{}\mathrm{\hspace{0.17em}90}\%`$ due to non-stellar sources, we correct the observed K magnitude of the *stellar component only* of m$`{}_{\text{K}}{}^{}=11.8`$ \[in agreement with other recent values m$`{}_{\text{K}}{}^{}=11.9`$ (Forbes et al. Forbes:92 (1992))\] to m$`{}_{\text{K}}{}^{}=14.3`$. Converting to absolute visual magnitudes yields M$`{}_{}{}^{\text{K5\hspace{0.17em}III}}{}_{\text{V}}{}^{}`$ (NGC 1275) = (V-K) + m$`{}_{\text{K}}{}^{}25`$A$`{}_{\text{K}}{}^{}5`$ log D$`{}_{\text{Mpc}}{}^{}=16.8`$ using V-K= 3.24, A$`{}_{\text{K}}{}^{}0.2`$, and D = 68.8 Mpc. The mass of a K giant is 1.2 M, and its M$`{}_{\text{V}}{}^{\text{K5\hspace{0.17em}III}}=0.2`$ (Lang Lang:92 (1992)), which implies $`4.6\times 10^6`$ M of K5 III stars in the central 3″. Assuming a Salpeter IMF with P(M)$``$ M<sup>-1.35</sup> dM, the mass fraction of the stars in the range 1.0 to 1.5 M (roughly the range of stellar masses which will become K5 III stars), for a population bounded by 0.1 M and 100 M stars is $`(1.0^{0.35}1.5^{0.35})/(0.1^{0.35}100^{0.35})=0.065`$, so that K5 III stars make up about 7% of the mass of a star cluster. Hence the total mass of stars in the central kpc is thus $`7.1\times 10^7`$ M. If the population is dominated by K supergiants, the total stellar mass comes out to be the same value (mass range = 12 – 14 M M$`{}_{\text{V}}{}^{\text{K5\hspace{0.17em}I}}=5.0`$). We estimate the core radius r<sub>0</sub> of the nuclear stellar cluster to be smaller than 500 pc. Using this value as an upper limit and knowing the stellar mass of the cluster, we can use $`\rho _0=9\sigma _0^2/4\pi Gr_0^2`$ (Eckart et al. Eckart:93 (1993)) to estimate the expected velocity dispersion: $`\sigma =\sqrt{M_{\text{cluster}}G/9r_0}\stackrel{>}{}\mathrm{\hspace{0.17em}8.4}`$ km/s. Even if the CO absorption had been detected on the nucleus, such a small dispersion had been beyond the capabilities of our instrument. We conclude that the absence of CO bandheads in the nuclear emission indicates that those bandheads are heavily diluted by the emission of hot dust and power-law emission. ## 8 Discussion Our spectral synthesis (Sect. 3) shows that inside of the central $`300pc`$, the NIR continuum emission is dominated by the AGN and hot dust components. Together these are $`\stackrel{>}{}`$ 10 times more powerful than the stellar emission from the same region. This result is confirmed by the observed factor of $`\stackrel{>}{}`$ 10 continuum CO$`_{\text{sp}}`$ dilution (Sect. 7) at the nucleus. The nature of the nuclear power source likely is a massive black hole (MBH) with a surrounding molecular gas cloud torus or disk. The lack of strong CO absorption bands indicates that massive star formation probably does not play a very important role in the nucleus. The intensity and morphology of the line emission suggest that more than one mechanism is at work exciting the line emission. In addition to the AGN, shocks are also likely to play a large role. Since the J band \[Fe ii\] lines are unusually strong, even for an AGN, it requires a substantial boost of the H<sub>2</sub> emission to keep the \[Fe ii\]/H<sub>2</sub> 1–0 S(1) ratio near unity. This is indeed observed in NGC 1275, which has unusually strong H<sub>2</sub> emission compared to other Seyferts. Furthermore, the spatial extension of the H<sub>2</sub> emission (Sect. 4.1) suggests that it may be driven by shocks from a recent or ongoing merger event. Such a picture is supported by not only the observed isophotal shell structure (Hernquist & Weil Hernquist:92 (1992)), but also by the extinction analysis of Norgaard-Nielsen et al. (Norgaard-Nielsen:93 (1993)), which shows that the HV system is at least partially *moving through* the main galaxy rather than completely in front of it. Hence the interaction of the HV and LV systems could give rise to the required shocks which boost the H<sub>2</sub> emission. This process is postulated to occur in the merging galaxy NGC 6240 (van der Werf et al. vanderWerf:93 (1993)). Among AGNs, strong \[Fe ii\] emission is not unique to NGC 1275. In fact there are a large number of AGNs with elevated \[Fe ii\] emission (among them many IRAS luminous and starburst galaxies such as MRK 231, MRK 507, and PHL 1092), which can be attributed to violent star formation in a metal rich environment (Fillipenko & Terlevich Fillipenko:93 (1992), Lipari et al. Lipari:93 (1993)). In the merger induced H<sub>2</sub> boosting scenario, the unusually high level of \[Fe ii\] emission observed from NGC 1275 would likely be due to a different effect, namely the dissociation of Fe from interstellar grains via shock heating. This could produce an iron rich gas phase (Sect. 6.1.1), which would then give rise to the very strong observed \[Fe ii\] emission. The fact that the Br$`\gamma `$ emission is generally extended suggests the presence of young stars and thus extended recent star formation. If the slight extension of the continuum is significant, it suggests increased star formation along the jets, perhaps due to shocks, caused by interactions with the surrounding medium. Part of the Br$`\gamma `$ emission along the radio jet may be influenced by the radiation from the nucleus and thus not created by star formation. The extended component of hot molecular hydrogen is probably shock excited as a result of the star formation. The existence (or non-existence) of “cooling flows” — an inflow of cool gas into a galaxy from its halo of hot x-ray emitting gas — is currently a topic of intense debate. The evidence *for* them is that some galaxies which are surrounded by a halo of hot, X-ray emitting gas show a dramatic central drop in the gas temperature. As the primary cooling mechanism of this gas is bremsstrahlung (emission $``$ n<sup>2</sup>), and as the central temperature drop is accompanied by a brightness increase, the gas must be both cooler and denser (see, e.g. Sarazin et al. Sarazin:92 (1992) for a physical introduction and review) . If the gas is cooler and denser, then it must fall toward the center of the galaxy. In the case of NGC 1275, the predicted infall amounts to $`200`$ M yr<sup>-1</sup> (Fabian et al. Fabian:81 (1981)), or $`2\times 10^{12}`$ M in $`10^{10}`$ years. The evidence *against* cooling flows is that if such enormous quantities of gas really are falling in, they must be detectable in some way, yet careful searches have not revealed the gas in any of its expected states. We know that the gas cannot form stars with a normal IMF, because 200 M of star formation would be 20 times more powerful than the prototypical starburst galaxy, M 82, and easily detected via color variations (Cardiel et al. Cardiel:95 (1995)). Our H<sub>2</sub> data provide several evidences against the cooling flow hypothosis. *i*) The spatial distribution of the hot molecular hydrogen strongly peaks at the nucleus of NGC 1275 on a scale $`<`$300pc. If the H<sub>2</sub> emission was due to the cooling flow, one would have expexted a much more extended distribution since the gas has to fall through all of the galaxy. The weakness of the H<sub>2</sub> 2–1 S(3) line (see Sect. 4.3) emphasises the close relation of the hot H<sub>2</sub> with the nucleus. *ii*) The temperature variation within the hot H<sub>2</sub> is small. Between 0 K and about 20000 K upper level energy the temperature only varies by about 1500 K. A cooling hot gas cloud is likely to show a much larger span of temperature compared to what our data show. *iii*) The temperature of the hot H<sub>2</sub> is rather low: between 1500 K and 3000 K. There is no evidence of a very hot ($`\stackrel{>}{}`$ 3000 K) H<sub>2</sub> component which can be expected for a cooling hot gas mass. *iv* The line emission of the hot H<sub>2</sub> can be explained consistently with mostly thermal excitation (e.g. in a postshock environment) with a smaller component of X–ray heating. A “cooling flow” scenario does not need to be invoked to explain the line emission of molecular hydrogen in NGC 1275. Several groups have argued that the gas cannot have collected in the form of molecular hydrogen (H<sub>2</sub>), because searches for CO emission associated with H<sub>2</sub> have revealed only very limited quantities of H<sub>2</sub> gas, although NGC 1275 is unusual in that it shows more CO than most cooling flow galaxies (Lazareff et al. Lazareff:89 (1989), Inoue et al Inoue:96 (1996), O’Dea et al. Odea:94 (1994)). However, since the temperature regime is very different from the cold molecular clouds, the CO $``$ H<sub>2</sub> conversion factor is certainly different and may even not be applicable in this context. The gas could have been collected as atomic hydrogen. However, only very limited quantities have been detected, certainly much less than would be expected from an infall of several hundred solar masses per year over cosmic timescales (Haschick et al. Haschick:82 (1982)). One of the last places in which the purported inflow could be hiding is in stars with a very unusual kind of IMF, one that is biased toward producing low mass stars, so that the star formation rate as determined by the total number of ionizing photons (these are produced only by massive stars) does not reflect the total amount of mass involved in the star formation process. A population of such low mass stars should have a somewhat lower CO$`_{\text{sp}}`$ index than a normal population, because it will have fewer luminous giants and supergiants which are the largest contributors to the CO absorption lines. Our NIR spectroscopy allows us to analyze the stellar population of regions 1 kpc from the nucleus via the CO$`_{\text{sp}}`$ index, where we find a CO$`{}_{\text{sp}}{}^{}=0.17`$. This can be compared with theoretical predictions for a population of very low mass stars (see Fig. 20 in Kroupa & Gilmore Kroupa:94 (1994)), whose expected value of CO$`_{\text{sp}}`$ is between 0.07 and 0.16 for cooling flow rates of 158 M yr<sup>-1</sup>. For higher mass deposition rates, the expected CO$`_{\text{sp}}`$ decreases, so that for a rate of 316 M yr<sup>-1</sup> the expected CO$`_{\text{sp}}`$ lies between 0.0 and 0.1. Hence we cannot completely eliminate the existence of a low mass star population for a cooling flow of 200 M yr<sup>-1</sup> as is postulated for NGC 1275, but the observed CO$`_{\text{sp}}`$ index is at the extreme limit of theoretically acceptable values. ## 9 Summary We observed the central 1 kpc of NGC 1275 with the MPE 3D imaging spectrograph. We find that the central $`300`$ pc are strongly dominated by emission from an AGN and from hot dust: There is no evidence of a nuclear stellar continuum. In the regions at a distance of $`1`$ kpc from the nucleus, the situation is completely reversed, and the emission is totally dominated by a stellar population whose spectra is consistent with a normal, old ($`>10^810^9`$ y) population. In this region, we find no evidence to support the thesis that substantial quantities of infalling “cooling flow” gas has been transformed into stars with an IMF biased toward low masses, although we cannot strictly rule out a low mass population because the purported flow may be disturbed in the central 10 kpc of the galaxy. Within the central $`500`$pc, the emission line ratios of many important diagnostic lines from the NIR regime are consistent with excitation from a combination of AGN emission and shock excitation. The shocks may be due to previous mergers, or to the early phases of a merger between the HV and LV systems. ###### Acknowledgements. We thank the 3D team for their enthusiastic support during the observations. ## Appendix A Spectral Fitting of the continuum emission The total intrinsic continuum emission at each wavelength $`\lambda `$ is $$I_\lambda ^{\text{INT}}=\underset{\text{i}}{}c_\text{i}I_\lambda ^\text{i}with\underset{\text{i}}{}c_\text{i}=1\text{.}$$ (3) In active and starburst galaxies, we identify four sources of continuum emission: stellar ($`I_\lambda ^{\text{STAR}}`$), AGN power-law($`I_\lambda ^\text{P}`$) and hot dust ($`I_\lambda ^{\text{HD}}`$), and free-free ($`I_\lambda ^{\text{FF}}`$). The intrinsic emission of $$I_\lambda ^{\text{INT}}=c_\text{S}I_\lambda ^{\text{STAR}}+c_{\text{PL}}I_\lambda ^\text{P}+c_{\text{HD}}I_\lambda ^{\text{HD}}+c_{\text{FF}}I_\lambda ^{\text{FF}}$$ (4) is then extincted according to $$I_\lambda ^{\text{SYN}}=\{\begin{array}{cc}I_\lambda ^{\text{INT}}e^{g\tau _\lambda }e^{\tau _\lambda },\hfill & \text{Screen}\hfill \\ I_\lambda ^{INT}e^{g\tau _\lambda }(1e^{\tau _\lambda })/\tau _\lambda ,\hfill & \text{Mix}\text{.}\hfill \end{array}$$ (5) to obtain the synthetic spectrum. The extinction has two parts, one due to dust within our Galaxy ($`g\tau _\lambda `$), and one due to dust within the source galaxy ($`\tau _\lambda `$). In our model, the latter can have either a simple screen geometry or a mixed stars/gas geometry. In all cases, the extinction is taken to vary as $`\lambda ^{1.85}`$ (Landini et al. Landini:84 (1984)), and is fixed to be A$`{}_{\text{K}}{}^{}=0.112`$ A$`_\text{V}`$ at $`\lambda =2.2\mu `$m (Rieke & Lebofsky Rieke:85 (1985)). When comparing $`I_\lambda ^{\text{SYN}}`$ with the observed spectrum, $`I_\lambda ^{\text{OBS}}`$, we avoid emission lines due to non-continuum sources (e.g. shocked ISM gas, photoionized gas clouds), and define our goodness of fit parameter, $`\chi ^2`$, such that $$\chi ^2=\frac{1}{N1}\underset{\lambda line}{}\frac{(I_\lambda ^{\text{SYN}}I_\lambda ^{\text{OBS}})^2}{\sigma _\lambda ^2}$$ (6) where $`N`$ is the number of measured spectral points which contain no line emission or absorption from non-continuum sources and $`\sigma _\lambda `$ is the local noise. To model the stellar component, we have resampled the R$`3000`$ normalized K band spectra of Kleinmann & Hall (Kleinmann:86 (1986)) at the 3D resolution of R = 670 onto our $`0.001\mu `$m pix<sup>-1</sup> data format (Förster et al. Forster:96 (1996)), and then multiplied them by a Rayleigh-Jeans power law of $`\lambda ^{3.94}`$ (Doyon et al. doyon-co:94 (1994)) corresponding to an A0V star to reconstruct the true stellar spectrum. The intrinsic spectrum of the hot dust of temperature T$`_\text{d}`$, is taken to be a black body multiplied by a power-law emissivity of the form $`ϵ_{\text{dust}}\lambda ^1`$. Hence, $`I_\lambda ^{\text{HD}}=B_\lambda (T_\text{d})\lambda ^1`$, where B$`{}_{\lambda }{}^{}(T_\text{d})`$ is the black body function. The behavior of free-free emission is given by $`I_\nu e^{\text{-h}\nu \text{/kT}}`$ where T is the plasma temperature (normally $`>5000`$ K). We convert to wavelength units, and find $`I_\lambda \lambda ^2e^{\text{-hc/}\lambda \text{kT}}`$ which we adopt as our functional form for free-free emission.
warning/0001/astro-ph0001282.html
ar5iv
text
# Analysis of the Hipparcos Measurements of HD 10697— A Mass Determination of a Brown-Dwarf Secondary ## 1 INTRODUCTION More than twenty candidates for extrasolar planets have been announced over the past four years (e.g., Mayor & Queloz 1995; Noyes et al. 1997; Marcy & Butler 1998; Marcy, Cochran & Mayor 2000). In each case, very precise stellar radial-velocity measurements, with a precision of 10 m s<sup>-1</sup> or better, indicated the presence of a low-mass unseen companion orbiting a nearby solar-type star. These high-precision discoveries came almost a decade after the first ’planet candidate’ around HD 114762 was discovered (Latham et al. 1989; Mazeh, Latham & Stefanik 1996) with much lower precision. In most cases, the individual masses of the companions are not known, because the inclination angles of their orbital planes relative to our line of sight cannot be derived from the spectroscopic data. The minimum masses for all candidates, attained for an inclination angle of $`90^{}`$, are in the range 0.5 to 10 Jupiter masses ($`M_J`$). To derive the actual mass we need additional information, like precise astrometry of the orbit, from which we can derive the inclination, and therefore the secondary mass. This can be done at least for the cases where the primary mass can be estimated from its spectral type. At present, the astronomical community has at hand the accurate astrometric Hipparcos data, which have already yielded numerous orbits with small semi-major axes (ESA 1997; Söderhjelm 1999), down to a few milli-arc-sec (mas). Perryman et al. (1996) indeed used the Hipparcos data to constrain the orbits of some planet candidates. The Hipparcos data, which span about three years, can be most effective for systems with periods comparable to the lifetime of the satellite mission. As the time base of the precise radial-velocity surveys is getting longer, we expect more such planets to be discovered. Such an example was the outermost planet of $`\upsilon `$ And whose period was found to be close to 3 years. Mazeh et al. (1999) used the Hipparcos data together with the spectroscopic data of $`\upsilon `$ And to get a mass of $`10.1_{4.6}^{+4.7}`$ $`M_J`$ for the outermost known planet of that system. That work has shown the great potential of the combination of spectroscopic data and Hipparcos measurements to derive masses for planet candidates. Very recently Vogt et al. (1999) detected radial-velocity evidence of new six planet candidates. One of these stars is HD 10697, with a period of $`1072\pm 10`$ days and a companion lower mass estimate of 6.35 $`M_J`$. The minimum semi-major axis of the stellar orbit was found to be 0.36 mas. In this Letter we follow Mazeh et al. (1999) and combine the spectroscopic elements with the Hipparcos data to find that the unseen companion has a mass of about 40 $`M_J`$. We therefore suggest that the secondary is probably a brown dwarf in close orbit, of about 2 AU, around its parent star. ## 2 The Hipparcos Astrometry HD 10697 (109 Psc, HIP8159; $`\alpha `$=01:44:55.82, $`\delta `$=+20:04:59.34 \[J2000\]; V = 6.3), is a G5IV star, whose mass is estimated by Vogt et al. (1999) to be 1.10 $`M_{}`$. The Hipparcos data of HD 10697 included 33 independent abscissa measurements, which are the results of work of FAST and NDAC — the two consortia involved in reduction of the raw Hipparcos data. These 33 measurements represent data from 17 orbits of the satellite, analyzed separately by the two consortia and then de-correlated to produce 2 independent measurements out of each orbit (van Leeuwen 1997; van Leeuwen & Evans 1998; Perryman et al. 1997). In orbit no. 614 only the NDAC consortium produced a measurement of HD 10697. The present analysis used the spectroscopic elements of HD 10697 as given by Vogt et al. (1999). These included the spectroscopic period, $`P=1072\pm 10`$ day, the periastron passage, $`T_0=2451482\pm 39`$, the radial-velocity amplitude, $`K=119\pm 3`$ m s<sup>-1</sup>, the eccentricity $`e=0.12\pm 0.02`$, and the longitude of the periastron, $`\omega =113\pm 14`$. The orbital astrometric elements include $`P,T_0,e,\omega `$ and three additional elements — the semi-major axis, $`a_1`$, the inclination, $`i`$, and the longitude of the nodes, $`\mathrm{\Omega }`$. In addition, the astrometric solution includes the five regular astrometric parameters — the parallax, the position (in right ascension and declination) and the proper motion (in right ascension and declination). All together we had 12 parameter model to fit to the astrometric data, four of which are common with the spectroscopic orbit. To find the best astrometric orbit we used the values of $`P,T_0,e,\omega `$ as given by the spectroscopic orbit, and solved for the other parameters. To do that we considered a dense grid on the ($`a_1,i`$) plane, and found the values of the five regular astrometric parameters and $`\mathrm{\Omega }`$ that minimized the $`\chi ^2`$ statistics for each value of ($`a_1,i`$). The result of this search is a $`\chi ^2`$ function, which depends on $`a_1`$ and $`i`$. The square-root normalized $`\chi ^2`$ is plotted in Figure 1 as a two-dimensional function. The figure shows a very pronounced “valley” at about $`a_1=2`$ mas and $`i=160^{}`$, indicating a detection of an astrometric motion. To derive the best semi-major axis and inclination for the system we used another spectroscopic element — the radial-velocity amplitude $`K`$, which has not been used so far in the analysis. This element induces a constraint on the product of $`a_1`$ and $`\mathrm{sin}i`$, which for the HD 10697 case results in $$a_1\mathrm{sin}i=0.36\pm 0.01\left(\frac{P}{1072\mathrm{day}}\right)\left(\frac{K}{119\mathrm{m}\mathrm{s}^1}\right)\left(\frac{\sqrt{1e^2}}{0.993}\right)\left(\frac{\pi }{30.71\mathrm{mas}}\right)\mathrm{mas},$$ (1) where we have used here the Hipparcos catalog parallax, $`\pi =30.71\pm 0.81`$ mas. This constraint is plotted in Figure 1 as a continuous line both on the ($`a_1,i`$) plane and on the square-root normalized $`\chi ^2`$ surface. The product was calculated assuming the best-fit value of $`\pi `$ — 30.3 mas. Using the catalog value yielded very similar results. In Figure 2 we collapsed the two-dimensional function onto the line of Eq. (1). We can see a clear minimum at $`170^{}`$. This corresponds to a semi-major axis of 2.1 mas and a mass of 38 $`M_J`$. Detailed $`\chi ^2`$ analysis resulted in $$a_1=2.1\pm 0.7\mathrm{mas},M_{sec}=38\pm 13M_J,$$ (2) where $`M_{sec}`$ is the mass of the unseen companion. ## 3 Discussion All the 29 planet candidates discovered so far have $`M\mathrm{sin}i`$ in the range of 0.5–10 $`M_J`$. No report has been published yet about any precise radial-velocity detection of a secondary with minimum masses in the range of 10–80 $`M_J`$. The large radial-velocity survey of K and G stars with lower precision of Mayor et al. (1997) also yielded only few binaries with minimum secondary mass in the range of 10–80 $`M_J`$, most of which are actually main-sequence secondaries with low inclination (Halbwachs et al. 1999; Udry et al. 1999). These studies attest to the paucity of secondaries in the range of 10–80 $`M_J`$, and suggest a distinction between the population of stellar secondaries and planets (Vogt et el. 1999; Mazeh 1999). This distinction might also help to find the mass upper limit for planetary companions. An upper limit of the planetary mass at about 10–20 $`M_J`$ is consistent with the accumulated distribution of planetary masses (Mazeh 1999). It is also consistent with the mass of the outermost known companion of $`\upsilon `$ And — $`10\pm 5`$ $`M_J`$ (Mazeh et al. 1999). It therefore seems that the secondary of HD 10697 mass is probably too large for a planetary mass. We therefore suggest that the secondary of HD 10697 is a brown dwarf orbiting around its parent star at a distance of 2 AU. The paradigm behind the distinction between brown dwarfs and planets assumes that binaries, even with small-mass secondaries, are formed by a different mechanism than that of planets (e.g., Boss 1996). Brown-dwarf secondaries are therefore at the low-mass end of the distribution of secondaries that were formed as binaries and not as planets. The study of the lower end of the mass distribution has great importance to our understanding of binary formation, specially if we compare them to the emerging population of brown-dwarf field stars (e.g., Bèjar, Zapatero Osorio & Rebolo 1999; Binney 1999; Tinney 1999 and see references therein). Until now the paucity of brown dwarf secondaries in short-period binaries could give the impression that such objects do not exist. This work suggests that there are such secondaries, even though rare, in this range of masses. As usual, objects in binaries with short-enough periods supply the precious opportunity of dynamical mass determination. The combination of astrometry and radial-velocity work yields here a first mass determination of a brown dwarf which is model independent. The separation between the two components of HD 10697 is about 70 mas (Vogt et al. 1999). Ground-based interferometry might be able to resolve the two objects despite the huge brightness difference between them and supply for the first time brightness measurements of a brown dwarf with dynamical mass determination. In addition, time dependent spectroscopy might help to obtain a spectrum of the faint companion, enabling us to confront the theory of brown-dwarf evolutionary models (e.g., Burrows & Sharp 1999) with the observations. This work was supported by the US-Israel Binational Science Foundation through grant 97-00460 and the Israeli Science Foundation. ## REFERENCES Bèjar, V. J. S., Zapatero Osorio, M. R., & Rebolo, R. 1999, ApJ, 521, 671 Binney, J. 1999, MNRAS, 307, L27 Boss, A. P. 1996, Nature, 379, 397 Burrows, A., & Sharp, C. M. 1999, ApJ, 512, 843 Halbwachs, J. L., Arenou, F., Mayor, M., Udry, S., & Queloz, D. 1999 A&A, submitted Latham, D. W., Mazeh, T., Stefanik, R. P., Mayor, M., & Burki, G. 1989, Nature, 339, 38 ESA 1997, The Hipparcos and Tycho Catalogues, ESA SP-1200 van Leeuwen, F. 1997, Space Sc. Rev., 81, 201 van Leeuwen, F., & Evans, D. W. 1998, A&AS, 130, 157 Marcy, G. W., & Butler R. P. 1998, ARAA, 36, 57 Marcy, G. W., Cochran, W. D., & Mayor, M. 2000, in Protostars and Planets IV, ed. V. Mannings, A. P. Boss & S. S. Russell (Tucson: University of Arizona Press), in press Mayor, M., & Queloz, D. 1995, Nature, 378, 355 Mayor, M., Queloz, D., Udry, S., & Halbwachs, J.-L. 1997, in IAU Coll. 161, Astronomical and Biochemical Origins and Search for Life in the Universe ed. C. B. Cosmovici, S. Boyer, & D. Werthimer (Bolognia: Editrice Compositori) 313 Mazeh, T. 1999, in IAU Coll. 170, Precise Stellar Radial Velocities, eds. J.B. Hearnshaw & C.D. Scarfe, in press Mazeh, T., Latham, D. W., & Stefanik R. P. 1996, ApJ, 466, 415 Mazeh, T., Zucker, S., Dalla Torre, A., & van Leeuwen, F., 1999, ApJL, 522, L149 Noyes, R. W., Jha, S., Korzennik, S. G., Krockenberger, M., Nisenson, P., Brown, T. M., Kennelly, E. J., & Horner, S. D. 1997, ApJL, 483, L111 Perryman, M. A. C., et al. 1996, A&A, 310, L21 Perryman, M. A. C., et al. 1997, A&A, 323, L49 Söderhjelm, S. 1999, A&A, 341, 121 Tinney, C. G. 1999, Nature, 397, 37 Vogt, S. S., Marcy, G. W., Butler, R. P., Apps, K. 1999, ApJ, submitted Udry, S., Mayor, M., Naef, D., Pepe, F., Queloz, D., Santos, N., Burnet, M. 1999, A&A, submitted FIGURE LEGENDS
warning/0001/gr-qc0001050.html
ar5iv
text
# 1. Introduction ## 1. Introduction In the early nineties linearised gravity in terms of connection variables and free Maxwell theory on flat spacetime, were treated as useful toy models on which to test techniques being developed for loop quantum gravity . Significant progress has been made in the field of loop quantum gravity since then. Hence, it is useful to revisit these systems using current techniques to clarify certain questions which arise in the context of those pioneering but necessarily non-rigorous efforts. Two important (and related) questions are: (I) How did similar techniques for the quantization of general relativity and for its linearization about flat space, result in a non Fock representation for the (kinematic sector) of the former and a Fock representation for the latter? In particular, what is the role of Poincare invariance in obtaining the Fock representation? (This last point was a puzzle to the authors themselves ). (II)What is the role of “smeared” loops in in obtaining a Fock representation? In this work, we use the abelian $`C^{}`$ algebra techniques which constitute the mathematically rigorous framework of the loop quantum gravity program today, to investigate (I) and (II) above. It is also our aim to clarify the role of the different mathematical structures in the quantization procedure which determine whether a Fock or non Fock representation results. Although we restrict attention to $`U(1)`$ theory on a flat spacetime, we believe that our results should be of some relevance to the case of linearised gravity. This work is motivated by the following question in loop quantum gravity: how do Fock space gravitons on flat spacetime arise from the non-Fock structure of the Hilbert space which serves as the kinematical arena for loop quantum gravity? Admittedely, the answer to this question must await the construction of the full physical state space (i.e. the kernel of all the constraints) of quantum gravity. Nevertheless, this work may illuminate some facets of the issues involved. The starting point for our analysis is the abelian Poisson bracket algebra of $`U(1)`$ holonomies around loops on a spatial slice. This algebra is completed to the abelian $`C^{}`$ algebra, $`\overline{𝒜}`$ of . Hilbert space representations of $`\overline{𝒜}`$ are in determined by continuous positive linear functions (PLFs) on $`\overline{𝒜}`$. We review the construction of $`\overline{𝒜}`$ and of the PLF introduced in (which we shall call the Haar PLF) in section 2. The resulting representation is a non Fock representation in which the Electric flux is quantized . In section 3 we construct an abelian $`C^{}`$ algebra $`\overline{𝒜}_r`$, based on the Poisson bracket algebra of holonomies around the “Gaussian smeared ” loops of .<sup>1</sup><sup>1</sup>1 $`r`$ is a small length which characterises the width of the Gaussian smearing function in . Next, we derive the key result of this work, namely that there exists a natural $`C^{}`$ algebraic isomorphism, $`I_r:\overline{𝒜}\overline{𝒜}_r`$ with the property that $`I_r(𝒜)=𝒜_r`$. The standard flat spacetime Fock vacuum expectation value restricts to a positive linear function on $`𝒜_r`$. We are unable to show the continuity or lack thereof, of this Fock PLF on $`𝒜_r`$. Nevertheless, since the GNS construction needs only a * algebra (as opposed to a $`C^{}`$ algebra), we can use the Fock PLF to construct a representation of the * algebra $`𝒜_r`$. In section 4 we show that this representation is indeed the standard Fock representation even though $`𝒜_r`$ is a proper subalgebra of the standard Weyl algebra for $`U(1)`$ theory. Using the map, $`I_r`$, we can define a Haar PLF on $`\overline{𝒜}_r`$. We construct the resulting representation in section 5a. Finally, we use $`I_r`$ to define a Fock PLF on $`𝒜`$ . The resulting representation is, in a precise sense, an approximation to the standard Fock representation. We study it in section 5b. Section 6 is devoted to a discussion of our results in the context of the questions (I) and (II). Some useful lemmas are proved in Appendices A1 and A2. In this work the spacetime of interest is flat $`R^4`$ and we use global cartesian coordinates $`(t,x^i),i=1,2,3`$. The spatial slice of interest is the initial $`t=0`$ slice and all calculations are done in the spatial cartesian coordinate chart $`(x^i)`$. We use units in which both the velocity of light and Plancks constant, $`\mathrm{}`$, are equal to 1. We freely raise and lower indices with the flat spatial metric. The Poisson bracket between the $`U(1)`$ connection $`A_a(\stackrel{}{x}),a=1,2,3`$ and its conjugate electric field $`E^b(\stackrel{}{y})`$ is $`\{A_a(\stackrel{}{x}),E^b(\stackrel{}{y})\}=e\delta _a^b\delta (\stackrel{}{x},\stackrel{}{y})`$ where $`e`$ is a constant with units of electric charge. ## 2. Review of the construction and representation theory of $`\overline{𝒜}`$. We quickly review the relevant contents of . We refer the reader to , especially appendix A2 of for details. The mathematical structures of interest are as follows. $`𝒜`$ is the space of smooth $`U(1)`$ connections on the trivial $`U(1)`$ bundle on $`R^3`$.<sup>2</sup><sup>2</sup>2Thus a minor change of notation from A2 of is that we denote $`𝒜_0`$ of that reference by by $`𝒜`$. We restrict attention to connections $`A_a(x)`$ whose cartesian components are functions of rapid decrease at infinity. $`_{x_0}`$ is the space of unparametrized, oriented, piecwise analytic loops <sup>3</sup><sup>3</sup>3This is in contrast to the $`C^1`$ loops of A2 of . on $`R^3`$ with basepoint $`\stackrel{}{x_0}`$. Composition of a loop $`\alpha `$ with a loop $`\beta `$ is denoted by $`\alpha \beta `$. Given a loop $`\alpha _{x_0}`$, the holonomy of $`A_a(x)`$ around $`\alpha `$ is $`H_\alpha (A):=\mathrm{exp}(i_\alpha A_a𝑑x^a)`$. $`\stackrel{~}{\alpha }`$ is the holonomy equivalence class (hoop class) of $`\alpha `$ i.e. $`\alpha ,\beta `$ define the same hoop iff $`H_\alpha (A)=H_\beta (A)`$ for every $`A_a(x)𝒜`$. $`𝒢`$ is the group generated by all hoops $`\stackrel{~}{\alpha }`$, where group multiplication is hoop composition i.e. $`\stackrel{~}{\alpha }\stackrel{~}{\beta }:=\stackrel{~}{\alpha \beta }`$. $`𝒜`$ is the abelian Poisson bracket algebra of $`U(1)`$ holonomies. $`_{x_0}`$ is the free algebra generated by elements of $`_{x_0}`$, with product law $`\alpha \beta :=\alpha \beta `$. With this product, all elements of $`_{x_0}`$ are expressible as complex linear combinations of elements of $`_{x_0}`$. $`K`$ is a 2 sided ideal of $`_{x_0}`$, such that $$\underset{i=1}{\overset{N}{}}a_i\alpha _iK\mathrm{iff}\underset{i=1}{\overset{N}{}}a_iH_{\alpha _i}(A)=0\mathrm{for}\mathrm{every}A_a(x)𝒜,$$ (1) where $`a_i`$ are complex numbers. $`_{x_0}`$ is quotiented by $`K`$ to give the algebra $`_{x_0}/K`$. The $`K`$ equivalence class of $`\alpha `$ is denoted by $`[\alpha ]`$. As abstract algebras, $`𝒜`$ and $`_{x_0}/K`$ are isomorphic. $$(\underset{i=1}{\overset{N}{}}a_i[\alpha _i])^{}:=\underset{i=1}{\overset{N}{}}a_i^{}[\alpha _i^1]$$ (2) defines a $``$ relation on $`𝒜`$. $$\underset{i=1}{\overset{N}{}}a_i[\alpha _i]:=\underset{A𝒜}{sup}|\underset{i=1}{\overset{N}{}}a_iH_{\alpha _i}(A)|$$ (3) defines a norm on $`𝒜`$. $`\overline{𝒜}`$ is the abelian $`C^{}`$ algebra obtained by defining $``$ on $`𝒜`$ and completing the resulting $``$ algebra with respect to $`||||`$. $`\mathrm{\Delta }`$ is the spectrum of $`\overline{𝒜}`$. $`\mathrm{\Delta }`$ is also denoted by $`\overline{𝒜/𝒢}`$ where $`𝒢`$ denotes the $`U(1)`$ gauge group and is a suitable completion of the space of connections modulo gauge, $`𝒜/𝒢`$. From Gel’fand theory, $`\mathrm{\Delta }`$ is the space of continuous, linear, multiplicative $``$ homeomorphisms, $`h`$, from $`\overline{𝒜}`$ to the ($`C^{}`$ algebra of) complex numbers $`𝐂`$. From the elements of $`\mathrm{\Delta }`$ are also in 1-1 correspondence with homeomorphisms from $`𝒢`$ to $`U(1)`$. Given $`X\overline{𝒜}`$, $`h(X)`$ is a complex function on $`\mathrm{\Delta }`$. $`\mathrm{\Delta }`$ is endowed with the weakest topology in which $`h(X)`$ for all $`X\overline{𝒜}`$ are continuous functions on $`\mathrm{\Delta }`$. In this topology, $`\mathrm{\Delta }`$ is a compact, Hausdorff space and the functions $`h([\alpha ]),\alpha _{x_0}`$ are dense in the $`C^{}`$ algebra, $`C(\mathrm{\Delta })`$, of continuous functions on $`\mathrm{\Delta }`$. Further, $`C(\mathrm{\Delta })`$ is isomorphic to $`\overline{𝒜}`$. Every continuous cyclic representation of $`\overline{𝒜}`$ is in 1-1 correspondence with a continuous positive linear functional (PLF) on $`\overline{𝒜}`$. Since $`\overline{𝒜}C(\mathrm{\Delta })`$, every continuous PLF so defined on $`C(\mathrm{\Delta })`$ is in correspondence, by the Riesz lemma, with some regular measure $`d\mu `$ on $`\mathrm{\Delta }`$ and $`\widehat{H}_\alpha `$ is represented on $`\psi L^2(\mathrm{\Delta },d\mu )`$ as unitary operator through $`(\widehat{H}_\alpha \psi )(h)=h([\alpha ])\psi (h)`$. In particular, the continuous ‘Haar’ PLF $`\mathrm{\Gamma }(\alpha )`$ $`=`$ $`1\mathrm{if}\stackrel{~}{\alpha }=\stackrel{~}{o}`$ (4) $`=`$ $`0\mathrm{otherwise}`$ (where $`o`$ is the trivial loop), corresponds to the Haar measure on $`\mathrm{\Delta }`$. $`\mathrm{\Delta }=\overline{𝒜/𝒢}`$ can also be constructed as the projective limit space of certain finite dimensional spaces. Each of these spaces is isomorphic to $`n`$ copies of $`U(1)`$ and is labelled by $`n`$ strongly independent hoops. Recall from that $`\stackrel{~}{\alpha }_ii=1..n`$ are strongly independent hoops iff $`\alpha _i_{x_0}`$ are strongly independent loops; $`\alpha _i,i=1..n`$ are strongly independent loops iff each $`\alpha _i`$ has at least one segment which intersects $`\alpha _{ji}`$ at most at a finite number of points. The Haar measure on $`\mathrm{\Delta }`$ is the projective limit measure of the Haar measures on each of the finite dimensional spaces. <sup>4</sup><sup>4</sup>4 Note that the proof of continuity of the Haar PLF in is incomplete in that it applies only if the loops $`\alpha _j`$ of A.7 of are holonomically independent. Nevertheless, if as in this work, we restrict attention to piecewise analytic loops, continuity of the Haar PLF can immediately be inferred from its definition through the Haar measure. Then the considerations of show that the electric flux $`_SE^a𝑑s_a`$ through a surface $`S`$ can be realised as an essentially self adjoint operator on the dense domain of cylindrical functions <sup>5</sup><sup>5</sup>5Cylindrical functions on $`\mathrm{\Delta }`$ are of the form $`\psi _{\{[\alpha _i]\}}:=\psi (h([\alpha _1])..h([\alpha _n]))`$, where $`\alpha _i,i=1..n`$, are a finite number of strongly indendent loops and $`\psi `$ is a complex function on $`U(1)^n`$. as $$_S\widehat{E}^a𝑑s_a\psi _{\{[\alpha _i]\}}=e\underset{i}{}N(S,\alpha _i)h([\alpha _i])\frac{\psi _{\{[\alpha _i]\}}}{h([\alpha _i])}$$ (5) where $`N(S,\alpha _i)`$ is the number of intersections between $`\alpha _i`$ and $`S`$. ## 3. $`\overline{𝒜}_r`$ and the isomorphism $`I_r`$ In section 3a we recall the definition of ‘smeared’ loops and their holonomies from and construct the ‘smeared’ loop related structures $`\stackrel{~}{\alpha }_r`$, $`K_r`$, $`𝒜_r`$, $`\overline{𝒜}_r`$ and $`\mathrm{\Delta }_r`$. In section 3b, using Appendix A2, we show that an isomorphism exists between the structures $`\stackrel{~}{\alpha }`$, $`K`$, $`𝒜`$, $`\overline{𝒜},\mathrm{\Delta }`$ and their ‘smeared’ versions. ### 3a. The construction of $`\overline{𝒜}_r`$ In the notation of , $`H_\alpha (A)`$ $`=`$ $`\mathrm{exp}i{\displaystyle _{R^3}}X_\gamma ^a(\stackrel{}{x})A_a(\stackrel{}{x})d^3x,`$ (6) $`X_\gamma ^a(\stackrel{}{x})`$ $`:=`$ $`{\displaystyle _\gamma }𝑑s\delta ^3(\stackrel{}{\gamma }(s),\stackrel{}{x})\dot{\gamma }^a,`$ (7) where $`s`$ is a parametrization of the loop $`\gamma `$, $`s[0,2\pi ]`$. $`X_\gamma ^a(\stackrel{}{x})`$ is called the form factor of $`\gamma `$. Its Fourier transform is $`X_\gamma ^a(\stackrel{}{k})`$ $`:=`$ $`{\displaystyle \frac{1}{2\pi ^{\frac{3}{2}}}}{\displaystyle _{R^3}}d^3xX_\gamma ^a(\stackrel{}{x})e^{i\stackrel{}{k}\stackrel{}{x}}`$ (8) $`=`$ $`{\displaystyle \frac{1}{2\pi ^{\frac{3}{2}}}}{\displaystyle _\gamma }𝑑s\dot{\gamma }^a(s)e^{i\stackrel{}{k}\stackrel{}{\gamma }(s)}.`$ The Gaussian smeared form factor is defined as $$X_{\gamma _{(r)}}^a(\stackrel{}{x}):=_{R^3}d^3yf_r(\stackrel{}{y}\stackrel{}{x})X_\gamma ^a(\stackrel{}{y})=_\gamma 𝑑sf_r(\stackrel{}{\gamma }(s)\stackrel{}{x})\dot{\gamma }^a(s)$$ (9) where $$f_r(\stackrel{}{x})=\frac{1}{2\pi ^{\frac{3}{2}}r^3}e^{\frac{x^2}{2r^2}}x:=|\stackrel{}{x}|$$ (10) approximates the Dirac delta function for small $`r`$. The Fourier transform of the smeared form factor is $$X_{\gamma _{(r)}}^a(\stackrel{}{k})=e^{\frac{k^2r^2}{2}}X_\gamma ^a(\stackrel{}{k})$$ (11) and the smeared holonomy is defined as $`H_{\gamma _{(r)}}(A)`$ $`=`$ $`\mathrm{exp}i{\displaystyle _{R^3}}X_{\gamma _{(r)}}^a(\stackrel{}{x})A_a(\stackrel{}{x})d^3x`$ (12) $`=`$ $`\mathrm{exp}i{\displaystyle _{R^3}}X_{\gamma _{(r)}}^a(\stackrel{}{k})A_a(\stackrel{}{k})d^3k.`$ where $`A_a(\stackrel{}{k})`$ is the Fourier transform of $`A_a(\stackrel{}{x})`$. We define $`\stackrel{~}{\alpha }_r`$, $`K_r`$, $`𝒜_r`$, $`\overline{𝒜}_r,\mathrm{\Delta }_r`$ as follows. $`\stackrel{~}{\alpha }_r`$ is the $`r`$-hoop class of $`\alpha `$ i.e. $`\alpha ,\beta `$ define the same $`r`$-hoop iff $`H_{\alpha _{(r)}}(A)=H_{\beta _{(r)}}(A)`$ for every $`A_a(x)𝒜`$. $`𝒢_r`$ is the group generated by all $`r`$-hoops $`\stackrel{~}{\alpha }_r`$ where group multiplication is $`r`$-hoop composition i.e. $$\stackrel{~}{\alpha }_r\stackrel{~}{\beta }_r:=(\stackrel{~}{\alpha \beta })_r.$$ (13) Note that the above definition is consistent because, from (12) and the definition of $`r`$-hoop equivalence, it follows that $$H_{\alpha _{(r)}}(A)H_{\beta _{(r)}}(A)=H_{(\alpha \beta )_{(r)}}(A)$$ (14) Note that from (13), it follows that the identity element of $`𝒢_r`$ is $`\stackrel{~}{o}_r`$ and that $`(\stackrel{~}{\alpha }_r)^1=\stackrel{~}{\alpha ^1}_r`$. $`𝒜_r`$ is the abelian Poisson bracket algebra of the $`r`$-holonomies, $`H_{\alpha _{(r)}}(A)`$, $`A_a𝒜`$, $`\alpha _{x_0}`$. Recall that with the product law defined in section 2, all elements of $`_{x_0}`$ are expressible as complex linear combinations of elements of $`_{x_0}`$. We define the 2 sided ideal of $`K_r_{x_0}`$, through $$\underset{i=1}{\overset{N}{}}a_i\alpha _iK_r\mathrm{iff}\underset{i=1}{\overset{N}{}}a_iH_{\alpha _{i}^{}{}_{(r)}{}^{}}(A)=0\mathrm{for}\mathrm{every}A_a(x)𝒜,$$ (15) where $`a_i`$ are complex numbers. The $`K_r`$ equivalence class of $`\alpha `$ is denoted by $`[\alpha ]_r`$. It can be seen that, as abstract algebras, $`𝒜_r`$ and $`_{x_0}/K_r`$ are isomorphic. It can be checked that the relation $`_r`$ defined on $`𝒜_r`$ by $$(\underset{i=1}{\overset{N}{}}a_i[\alpha _i]_r)^_r:=\underset{i=1}{\overset{N}{}}a_i^{}[\alpha _i^1]_r$$ (16) is a $``$ relation. Note that from (12), the complex conjugate of $`H_{\alpha _{(r)}}(A)`$ is $`H_{\alpha _{}^{1}{}_{(r)}{}^{}}(A)`$ and hence the abstract $`_r`$ relation just encodes the operation of complex conjugation on the algebra $`𝒜_r`$. Next we define the norm $`||||_r`$ as $$\underset{i=1}{\overset{N}{}}a_i[\alpha _i]_r_r:=\underset{A𝒜}{sup}|\underset{i=1}{\overset{N}{}}a_iH_{\alpha _{i}^{}{}_{(r)}{}^{}}(A)|.$$ (17) It is easily verified that $`||||_r`$ is indeed a norm on the $``$ algebra $`𝒜_r`$ with $``$ relation defined by (16). Completion of $`𝒜_r`$ with respect to $`||||_r`$ gives the abelian $`C^{}`$ algebra $`\overline{𝒜}_r`$. Next, we characterize the spectrum $`\mathrm{\Delta }_r`$ of $`\overline{𝒜}_r`$ as the space of all homomorphisms from $`𝒢_r`$ to $`U(1)`$. Let $`h\mathrm{\Delta }_r`$. Thus $`h`$ is a linear, multiplicative, continuous * homorphism from $`\overline{𝒜}_r`$ to $`𝐂`$. $$h([\alpha ]_r)h([\alpha ^1]_r)=h([o]_r).$$ $$\mathrm{Choosing}\alpha =o,h([o]_r)^2=h([o]_r)h([o]_r)=1.$$ (18) $$h([\alpha ^1]_r)=\frac{1}{h([\alpha ]_r)}=h^{}([\alpha ]_r).$$ (19) (19) implies that $`|h([\alpha ]_r)|=1`$ and this, coupled with the fact that $`𝒢_r`$ is commutative, shows that every $`h\mathrm{\Delta }_r`$ defines a homomorphism from $`𝒢_r`$ to $`U(1)`$. Conversely, let $`h`$ be a homomorphism from $`𝒢_r`$ to $`U(1)`$. Its action can be extended by linearity to elements of $`𝒜_r`$ so that $`h(_{i=1}^Na_i[\alpha _i]_r):=_{i=1}^Na_ih([\alpha _i]_r)`$. <sup>6</sup><sup>6</sup>6$`h`$ can be defined on $`[\alpha ]_r`$ because $`K_r`$ equivalence subsumes $`r`$-hoop equivalence. It is also easy to see that $`h([\alpha ^1]_r)=h^{}([\alpha ]_r)`$. These properties and the fact that $`h`$ is a homomorphism from $`𝒢_r`$ to $`U(1)𝐂`$, imply that $`h`$ is a linear, multiplicative, * homomorphism from $`𝒜_r`$ to $`𝐂`$. Finally we show that $`h`$ extends to a continuous homomorphism on $`\overline{𝒜}_r`$. From it follows that for $`\alpha _i_{x_0},i=1..n`$, there exist strongly independent $`\beta _j,j=1..m`$ such that each $`\alpha _i`$ is the composition of some of the $`\{\beta _j\}`$. From this fact and Lemma 2 of Appendix A1, it can be shown that, for a given $`_{i=1}^Na_i[\alpha _i]_r𝒜_r`$ and any $`\delta >0`$, there exists $`A_a^{(a_i,\delta ,r)}𝒜`$ such that $$|\underset{i=1}{\overset{N}{}}a_i\left(h([\alpha _i]_r)H_{\alpha _{i(r)}}(A^{(a_i,\delta ,r)})\right)|<\delta .$$ (20) From (20), it is straightforward to show that $$|\underset{i=1}{\overset{N}{}}a_i(h([\alpha _i]_r)|\underset{A𝒜}{sup}|\underset{i=1}{\overset{N}{}}a_iH_{\alpha _{i}^{}{}_{(r)}{}^{}}(A)|.=||\underset{i=1}{\overset{N}{}}a_i[\alpha _i]_r||_r.$$ (21) Since $`𝒜_r`$ is dense in $`\overline{𝒜}_r`$, (21) implies that $`h`$ can be extended to a continuous (linear, multiplicative) homorphism from $`\overline{𝒜}_r`$ to $`𝐂`$. Thus $`\mathrm{\Delta }_r`$ can be identified with the set of all homomorphisms from $`𝒢_r`$ to $`U(1)`$. ### 3b. The isomorphism $`I_r`$ We show that (i) $`K=K_r`$ : Let $$\underset{i=1}{\overset{N}{}}a_iH_{\alpha _i}(A)=0\mathrm{for}\mathrm{every}A_a(x)𝒜.$$ (22) From Lemma 3 of Appendix A1, given $`A_a𝒜`$, there exists $`A_{a(r)}𝒜`$ such that $$\underset{i=1}{\overset{N}{}}a_iH_{\alpha _{i(r)}}(A)=\underset{i=1}{\overset{N}{}}a_iH_{\alpha _i}(A_{(r)}).$$ (23) (22) and (23) imply that $`KK_r`$. Let $$\underset{i=1}{\overset{N}{}}a_iH_{\alpha _{i}^{}{}_{(r)}{}^{}}(A)=0\mathrm{for}\mathrm{every}A_a(x)𝒜.$$ (24) $``$ Given $`A_a,B_a𝒜`$, $$|\underset{i=1}{\overset{N}{}}a_iH_{\alpha _i}(A)|=|\underset{i=1}{\overset{N}{}}a_iH_{\alpha _i}(A)\underset{i=1}{\overset{N}{}}a_iH_{\alpha _{i(r)}}(B)|.$$ (25) Choose, $`B_a=A_a^ϵ`$ where $`A_a^ϵ`$ is defined in Lemma 1, A1. Then $`|_{i=1}^Na_iH_{\alpha _i}(A)|_{i=1}^N|a_i|ϵ`$ for every $`ϵ>0`$. $`_{i=1}^Na_iH_{\alpha _i}(A)=0`$ and hence, $`K_rK`$. Thus $`K=K_r`$, $`[\alpha ]=[\alpha ]_r`$ and $`\stackrel{~}{\alpha }=\stackrel{~}{\alpha }_r`$ (ii)$`_{i=1}^Na_i[\alpha _i]=_{i=1}^Na_i[\alpha _i]_r_r`$: Let $`_{i=1}^Na_i[\alpha _i]_r_r=c_r`$. Then $`c_r|_{i=1}^Na_iH_{\alpha _{i(r)}}(A)|`$ for every $`A_a𝒜`$. Further, for every $`\tau >0`$ there exists $`{}_{}{}^{(\tau )}A_{a}^{}𝒜`$ such that $`c_r|_{i=1}^Na_iH_{\alpha _{i(r)}}({}_{}{}^{(\tau )}A)|\tau `$.Then, from Lemma 3, A1, there exists $`{}_{}{}^{(\tau )}A_{a(r)}^{}𝒜`$ such that $$0c_r\underset{i=1}{\overset{N}{}}a_iH_{\alpha _i}({}_{}{}^{(\tau )}A_{(r)}^{})\tau .$$ (26) $$\underset{A𝒜}{sup}|\underset{i=1}{\overset{N}{}}a_iH_{\alpha _i}(A)|c_r\underset{i=1}{\overset{N}{}}a_i[\alpha _i]\underset{i=1}{\overset{N}{}}a_i[\alpha _i]_r_r$$ (27) Let $`_{i=1}^Na_i[\alpha _i]=c`$. Then for every $`\tau >0`$ there exists $`{}_{}{}^{(\frac{\tau }{2})}A_{a}^{}𝒜`$ such that $$c|\underset{i=1}{\overset{N}{}}a_iH_{\alpha _i}({}_{}{}^{(\frac{\tau }{2})}A)|\frac{\tau }{2}.$$ (28) From Lemma 1 A1, there exists $`{}_{}{}^{(\frac{\tau }{2})}A_{a}^{ϵ}𝒜`$ such that $`|H_{\alpha _{i(r)}}({}_{}{}^{(\frac{\tau }{2})}A_{}^{ϵ})H_{\alpha _i}({}_{}{}^{(\frac{\tau }{2})}A)|`$ $``$ $`ϵ.`$ $`|{\displaystyle \underset{i=1}{\overset{N}{}}}a_i\left(H_{\alpha _{i(r)}}({}_{}{}^{(\frac{\tau }{2})}A_{}^{ϵ})H_{\alpha _i}({}_{}{}^{(\frac{\tau }{2})}A)\right)|`$ $``$ $`{\displaystyle \underset{i=1}{\overset{N}{}}}|a_i|ϵ.`$ (29) Choose $`0<ϵ<\frac{\tau }{2_{i=1}^N|a_i|}`$. From (28) and (29) it follows that, for every $`\tau >0`$, $$c|\underset{i=1}{\overset{N}{}}a_iH_{\alpha _{i(r)}}({}_{}{}^{(\frac{\tau }{2})}A_{}^{ϵ})|<\tau .$$ (30) $$c\underset{A𝒜}{sup}|\underset{i=1}{\overset{N}{}}a_iH_{\alpha _{i(r)}}(A)|\underset{i=1}{\overset{N}{}}a_i[\alpha _i]\underset{i=1}{\overset{N}{}}a_i[\alpha _i]_r_r.$$ (31) Thus, for any finite $`N`$, $`_{i=1}^Na_i[\alpha _i]=_{i=1}^Na_i[\alpha _i]_r_r`$ From (i) and (ii) it follows that the structures $`K_r`$, $`[\alpha _r]`$, $`\stackrel{~}{\alpha }_r`$, $`𝒢_r`$, $`_r`$, $`𝒜_r`$, $`\overline{𝒜}_r,\mathrm{\Delta }_r`$ are isomorphic to $`K`$, $`[\alpha ]`$, $`\stackrel{~}{\alpha }`$, $`𝒢`$, \*, $`𝒜`$, $`\overline{𝒜},\mathrm{\Delta }`$. Thus a $`C^{}`$ isomorphism $`I_r:\overline{𝒜}\overline{𝒜}_r`$ exists such that $$I_r(\underset{i=1}{\overset{N}{}}a_i[\alpha _i])=\underset{i=1}{\overset{N}{}}a_i[\alpha _i]_r.$$ (32) $`I_r`$ defines a natural 1-1 map from $`\mathrm{\Delta }`$ to $`\mathrm{\Delta }_r`$ (which we shall also call $`I_r`$). Given the * isomorphism $`h\mathrm{\Delta }`$, from $`\overline{𝒜}`$ to $`𝐂`$, its image is $`h_r\mathrm{\Delta }_r`$ where $$h_r(\underset{i=1}{\overset{N}{}}a_i[\alpha _i]_r):=h(\underset{i=1}{\overset{N}{}}a_i[\alpha _i]_r).$$ (33) Note that if $`h_r`$ is defined by some smooth, non-flat $`A_a𝒜`$ then it is not true that $`h`$ is associated with (the gauge equivalence class of) the same connection. ## 4. Fock representation from $`\overline{𝒜}_r`$ The standard Fock space vacuum expectation value restricted to $`𝒜_r`$ defines the Fock PLF on $`𝒜_r`$ as $$\mathrm{\Gamma }_F(\underset{i=1}{\overset{N}{}}a_i[\alpha _i]_r):=\underset{i=1}{\overset{N}{}}a_i\mathrm{exp}(\frac{d^3k}{k}|X_{\alpha _{i(r)}}^a(\stackrel{}{k})|^2).$$ (34) Since $`𝒜_r`$ is a proper subalgebra of the standard Weyl algebra for $`U(1)`$ theory, it is not clear that its quantization (through the GNS construction based on the Fock PLF) reproduces the full Fock space. We prove that the full Fock space is indeed obtained. Let the GNS Hilbert space (based on $`\mathrm{\Gamma }_F`$) be $``$. Let $`𝒟`$ be the linear subspace of $``$ spanned by elements of the form $`\widehat{H}_{\alpha _{(r)}}\mathrm{\Omega },\alpha _{x_0}`$ where $`\mathrm{\Omega }`$ is the GNS vacuum. It can be seen that $`𝒟`$ is dense in $``$. $`𝒟`$ is naturally embedded in the Fock space, $``$, through the map $`U:𝒟`$ defined by $$U(\mathrm{\Omega })=|0>\mathrm{and}U(\underset{i=1}{\overset{N}{}}a_i\widehat{H}_{\alpha _{i(r)}}\mathrm{\Omega })=\underset{i=1}{\overset{N}{}}a_i\mathrm{exp}i_{R^3}X_{\alpha _{i(r)}}^a(\stackrel{}{x})\widehat{A}_a(\stackrel{}{x})d^3x|0>.$$ (35) Here $`\widehat{A}_a`$ is the standard Fock space operator valued distribution at $`t=0`$ $$\widehat{A}_a(\stackrel{}{x})=\frac{1}{2\pi ^{\frac{3}{2}}}\frac{d^3k}{\sqrt{k}}(e^{i\stackrel{}{k}\stackrel{}{x}}\widehat{a}_a(\stackrel{}{k})+e^{i\stackrel{}{k}\stackrel{}{x}}\widehat{a}_a^{}(\stackrel{}{k})$$ (36) where $$\widehat{a}_a(\stackrel{}{k})k^a=0,[\widehat{a}_a(\stackrel{}{k}),\widehat{a}_b^{}(\stackrel{}{l})]=\delta _{ab}\delta (\stackrel{}{k},\stackrel{}{l}).$$ (37) By construction, $`U`$ is a unitary map and can be uniquely extended to $``$ so that it embeds $``$ in $``$. We show that Cauchy limits of states in $`U(𝒟)`$ span a dense set in $``$ \- this suffices to show that the entire Fock space is indeed obtained, i.e. that $`U()=`$. Define the ‘occupation number’ states $$|\varphi ,p>:=d^3k_1..d^3k_p\varphi ^{a_1..a_p}(\stackrel{}{k}_1..\stackrel{}{k}_p)\widehat{a}_{a_1}^{}(\stackrel{}{k}_1)..\widehat{a}_{a_p}^{}(\stackrel{}{k}_p)|0>.$$ (38) $`\varphi ^{a_1..a_p}(\stackrel{}{k}_1..\stackrel{}{k}_p)`$ (with $`p`$ a positive integer) is such that (a) $`d^3k_i|\varphi ^{a_1..a_p}(\stackrel{}{k}_1,..,\stackrel{}{k}_i,..,\stackrel{}{k}_p)|^2<\mathrm{}`$ and $`\varphi ^{a_1..a_p}(\stackrel{}{k}_1..\stackrel{}{k}_p)`$ falls of faster than any inverse power of $`k_i`$ as $`k_i\mathrm{},\stackrel{}{k}_{ji}`$ fixed. (b) $`\varphi ^{a_1..a_i..a_p}(\stackrel{}{k}_1,..,\stackrel{}{k}_i,..,\stackrel{}{k}_p)(k_i)_{a_i}=0`$ i.e. it is transverse. (c) it is symmetric under interchange of $`(a_i,\stackrel{}{k}_i)`$ with $`(a_j,\stackrel{}{k}_j)`$ for all $`i,j=1..p`$. $`|\varphi ,p>`$ for all $`p`$ together with $`|0>`$, span a dense set, $`𝒟_0`$. Given 2 vectors $`\stackrel{}{x},\stackrel{}{v}`$, define the operator $$\widehat{O}_{\stackrel{}{x},\stackrel{}{v}}:=\frac{i}{2\pi ^{\frac{3}{2}}}\frac{d^3k}{\sqrt{2}k}e^{i\stackrel{}{k}\stackrel{}{x}}e^{\frac{k^2r^2}{2}}(\stackrel{}{v}\times \stackrel{}{k})^a(\widehat{a}_a(\stackrel{}{k})+\widehat{a}_a^{}(\stackrel{}{k})).$$ (39) As argued in Appendix A2, states of the form $`|\psi _{\{\stackrel{}{x}_i,\stackrel{}{v}_i\}}>:=_{i=1}^p\widehat{O}_{(\stackrel{}{x}_i,\stackrel{}{v}_i)}|0>,p=1,2..`$ together with $`|0>`$ span $`𝒟_0`$. Our proof that $`\psi _{\{\stackrel{}{x}_i,\stackrel{}{v}_i\}}U()`$ is as follows. (i) Note that $`\psi _{\{\stackrel{}{x}_i,\stackrel{}{v}_i\}}𝒟_0`$. (ii) Let $`\gamma ^{\{m,\stackrel{}{x},\stackrel{}{n}\}}`$ be a circular loop of radius $`ϵ_m:=\frac{1}{2^m}`$ ($`m`$ is a positive integer), centred at $`\stackrel{}{x}`$ and let its plane have unit normal $`\stackrel{}{n}`$ <sup>7</sup><sup>7</sup>7Although $`\gamma ^{\{m,\stackrel{}{x},\stackrel{}{n}\}}`$ is not in $`_{x_0}`$, the loop formed by joining $`\gamma ^{\{m,\stackrel{}{x},\stackrel{}{n}\}}`$ to the base point $`x_0`$ and retracing, is. We shall continue to denote this loop, which represents the same hoop, by $`\gamma ^{\{m,\stackrel{}{x},\stackrel{}{n}\}}`$.. The image of $`\widehat{H}_{\gamma _{(r)}^{\{m,\stackrel{}{x},\stackrel{}{n}\}}}`$ on $`U(𝒟)`$ is $`\mathrm{exp}(iX_{\gamma _{(r)}^{\{m,\stackrel{}{x},\stackrel{}{n}\}}}^a(\stackrel{}{y})\widehat{A}_a(\stackrel{}{y})d^3y)`$. Define $$\widehat{O}_{\stackrel{}{x},\stackrel{}{n},m}:=\frac{e^{iX_{\gamma _{(r)}^{\{m,\stackrel{}{x},\stackrel{}{n}\}}}^a(\stackrel{}{y})\widehat{A}_a(\stackrel{}{y})d^3y}1}{i\pi ϵ_m^2}.$$ (40) The formal limit of $`\widehat{O}_{\stackrel{}{x},\stackrel{}{n},m}`$ as $`m\mathrm{}`$ is $`\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}`$. We show below that $`\widehat{O}_{\stackrel{}{x},\stackrel{}{n},m}|\psi >`$, $`|\psi >𝒟_0`$, <sup>8</sup><sup>8</sup>8Note that since $`\widehat{O}_{\stackrel{}{x},\stackrel{}{n},m}`$ are bounded operators defined on the entire Fock space, $`\widehat{O}_{\stackrel{}{x},\stackrel{}{n},m}`$ are well defined on $`𝒟_0`$. form a Cauchy sequence with limit $`\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}|\psi >`$. Then, choosing $`|\psi >=|0>`$, we see that $`\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}|0>`$ is in the completion of $`U(𝒟)`$. (iii) From (i) above, $`\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}|0>𝒟_0`$. We can repeat the argument in (ii) above to conclude that $`\widehat{O}_{\stackrel{}{x}_2,\stackrel{}{n}_2}\widehat{O}_{\stackrel{}{x}_1,\stackrel{}{n}_1}|0>`$ is obtained as the Cauchy limit of the states $`\widehat{O}_{\stackrel{}{x}_2,\stackrel{}{n}_2,m}\widehat{O}_{\stackrel{}{x}_1,\stackrel{}{n}_1}|0>`$. Iterating this argument we see that $`|\psi _{\{\stackrel{}{x}_i,\stackrel{}{n}_i\}}>,|\stackrel{}{n}_i|=1`$ is in the completion of $`U(𝒟)`$. Finally, set $`\stackrel{}{v}_i:=v_i\stackrel{}{n}_i`$, where $`v_i`$ are real numbers. Then it follows that $`|\psi _{\{\stackrel{}{x}_i,\stackrel{}{v}_i\}}>=(_{i=1}^pv_i)|\psi _{\{\stackrel{}{x}_i,\stackrel{}{n}_i\}}>`$ and hence that $`|\psi _{\{\stackrel{}{x}_i,\stackrel{}{v}_i\}}>U()`$. Thus, it remains to show (see (ii) above) that: Given $`\psi 𝒟_0`$, $`\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}`$ as defined in (39) and $`\widehat{O}_{\stackrel{}{x},\stackrel{}{n},m}`$ as defined in (40), $$\underset{m\mathrm{}}{lim}\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}\widehat{O}_{\stackrel{}{x},\stackrel{}{n},m}|\psi >=0$$ (41) Proof: Let $$\widehat{D}:=X_{\gamma _{(r)}^{\{m,\stackrel{}{x},\stackrel{}{n}\}}}^a(\stackrel{}{y})\widehat{A}_a(\stackrel{}{y})d^3y\pi ϵ_m^2\widehat{O}_{(\stackrel{}{x},\stackrel{}{n})}.$$ (42) Thus $$\widehat{O}_{\stackrel{}{x},\stackrel{}{n},m}=\frac{e^{i\pi ϵ_m^2\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}+i\widehat{D}}1}{i\pi ϵ_m^2}.$$ (43) Since both $`e^{i\pi ϵ_m^2\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}}`$ and $`e^{i\widehat{D}}`$ are commuting elements of the standard Weyl algebra, $$\widehat{O}_{\stackrel{}{x},\stackrel{}{n},m}=\frac{e^{i\pi ϵ_m^2\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}}e^{i\widehat{D}}1}{i\pi ϵ_m^2}$$ (44) $`\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}\widehat{O}_{\stackrel{}{x},\stackrel{}{n},m}|\psi >`$ $`={\displaystyle \frac{e^{i\pi ϵ_m^2\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}}(e^{i\widehat{D}}1)}{i\pi ϵ_m^2}}|\psi >+\left({\displaystyle \frac{e^{i\pi ϵ_m^2\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}}1}{i\pi ϵ_m^2}}\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}\right)|\psi >`$ $`\left({\displaystyle \frac{e^{i\pi ϵ_m^2\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}}1}{i\pi ϵ_m^2}}\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}\right)|\psi >+{\displaystyle \frac{e^{i\pi ϵ_m^2\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}}(e^{i\widehat{D}}1)}{i\pi ϵ_m^2}}|\psi >`$ (45) From Lemma 2 of Appendix A2, $`\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}`$ is a densely defined symmetric operator on $`𝒟_0`$ and admits self adjoint extensions. Hence, from , the first term in (45) vanishes in the $`ϵ_m0`$ limit. Further, since $`e^{i\pi ϵ_m^2\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}}`$ is a unitary operator, we have $$\frac{e^{i\pi ϵ_m^2\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}}(e^{i\widehat{D}}1)}{i\pi ϵ_m^2}|\psi >=\frac{(e^{i\widehat{D}}1)}{i\pi ϵ_m^2}|\psi >.$$ (46) But $$\frac{(e^{i\widehat{D}}1)}{i\pi ϵ_m^2}|\psi >^2=\left(<\psi |\frac{(e^{i\widehat{D}}1)}{\pi ^2ϵ_m^4}|\psi >+<\psi |\frac{(e^{i\widehat{D}}1)}{\pi ^2ϵ_m^4}|\psi >\right).$$ (47) From Lemma 3, A2 and (47), $`\frac{(e^{i\widehat{D}}1)}{i\pi ϵ_m^2}|\psi >0`$ as $`ϵ_m0`$ and then (45) implies (41). Thus we have shown above that the GNS representation of $`𝒜_r`$ on the GNS Hilbert space $``$, is unitarily equivalent to the standard Fock representation on $`=L^2(𝒮^{},d\mu _G)`$ ($`𝒮^{}`$ denotes the appropriate space of tempered distributions and $`\mu _G`$ is the standard Gaussian measure with covariance $`\frac{1}{2}(^2)^{\frac{1}{2}}`$ ) via the unitary map $`U`$. The action of the smeared electric field operator, $`\widehat{E}(\stackrel{}{f}):=d^3xf_a(\stackrel{}{x})\widehat{E}^a(\stackrel{}{x})`$, on $`\psi 𝒞_FL^2(𝒮^{},d\mu _G)`$ is written in the standard way as $$\widehat{E}(\stackrel{}{f})\psi =ed^3x\left(if_a(\stackrel{}{x})\frac{\delta }{\delta A_a(\stackrel{}{x})}+i((^2)^{\frac{1}{2}}f^a(\stackrel{}{x}))A_a(\stackrel{}{x})\right)\psi .$$ (48) Here $`f_a(\stackrel{}{x})`$ is real, divergence free, smooth and of rapid decrease, and $`𝒞_FL^2(𝒮^{},d\mu _G)`$ is the standard dense domain of cylindrical functions appropriate to Fock space. The smeared electric field operator on $``$ is defined as the unitary image of $`\widehat{E}(\stackrel{}{f})`$ by $`U^1`$ i.e. for $`\psi U^1(𝒞_F)`$ $$\widehat{E}(\stackrel{}{f})\psi =U^1ed^3x\left(if_a(\stackrel{}{x})\frac{\delta }{\delta A_a(\stackrel{}{x})}+i((^2)^{\frac{1}{2}}f^a)(\stackrel{}{x})A_a(\stackrel{}{x})\right)U\psi .$$ (49) With this action, $`\widehat{E}(\stackrel{}{f})`$ is densely defined on the dense domain $`U^1(𝒞_F)`$, and just like its unitary image on $`𝒞_F`$, admits a unique self adjoint extension. ## 5. Induced representations through $`I_r`$. It can be verified that $`I_r`$ is a topological homorphism from $`\mathrm{\Delta }`$ to $`\mathrm{\Delta }_r`$ (where $`\mathrm{\Delta }`$ and $`\mathrm{\Delta }_r`$ are equipped with their Gel’fand topologies). Hence, $`I_r`$ defines a measurable isomorphism $`I_r:_r`$ where $``$ and $`_r`$ are the Borel sigma algebras associated with $`\mathrm{\Delta }`$ and $`\mathrm{\Delta }_r`$ respectively. Any regular Borel measure $`\mu `$ on $`\mathrm{\Delta }`$ induces a regular Borel measure $`\mu _r`$ on $`\mathrm{\Delta }_r`$, with $`\mu _r:=\mu I_r^1`$. It follows that $`I_r`$ defines a unitary map $`U_r`$ from $`L^2(\mathrm{\Delta },d\mu )`$ to $`L^2(\mathrm{\Delta }_r,d\mu _r)`$. $`U_r`$ can be explicitly defined through its action on the dense set $`𝒞L^2(\mathrm{\Delta },d\mu )`$, of cylindrical functions (cylindrical functions in the context of $`\overline{𝒜}`$ have been defined in section 2). Denote the dense set of cylindrical functions in $`L^2(\mathrm{\Delta }_r,d\mu _r)`$ by $`𝒞_r`$ <sup>9</sup><sup>9</sup>9Cylindrical functions are of the form $`\psi _{\{[\alpha _i]_r\}}(h):=\psi (h([\alpha _1]_r)..h([\alpha _n]_r))`$, for $`\alpha _i_{x_0},i=1..n`$, $`h\mathrm{\Delta }_r`$, and they span $`𝒞_r`$.. $`U_r`$ maps $`𝒞`$ to $`𝒞_r`$ through $$U_r(\psi _{\{[\alpha _i]\}})=\psi _{\{[\alpha _i]_r\}}.$$ (50) It also follows that $$U_r\widehat{H}_\alpha U_r^1=\widehat{H}_{\alpha _{(r)}}.$$ (51) Thus $`I_r`$ induces a representation of $`\overline{𝒜}_r`$ from a representation of $`\overline{𝒜}`$. In section 5a, we induce a Haar like representation of $`\overline{𝒜}_r`$ from the Haar representation of $`\overline{𝒜}`$. Since the image of $`I_r`$ restricted to $`𝒜`$ is $`𝒜_r`$, $`I_r`$ (or $`I_r^1`$) can also be used to induce representations of $`𝒜`$ from those of $`𝒜_r`$ and vice versa. In section 5b, we induce a Fock like representation of $`𝒜`$ from the Fock representation of $`𝒜_r`$. The elements of $`𝒜_r`$ define a dense subspace of $``$ through the GNS construction and a map $`U_r`$ is defined through (50) and (51). $`U_r^1`$ induces a Fock like representation of $`𝒜`$. ### 5a. Haar representation of $`\overline{𝒜}_r`$ We denote both the Haar measure on $`\mathrm{\Delta }`$ as well as its image on $`\mathrm{\Delta }_r`$ by $`d\mu _H`$. The induced PLF (corresponding to $`d\mu _H`$) on $`\overline{𝒜}_r`$ is defined by $`\mathrm{\Gamma }(\alpha )`$ $`=`$ $`1\mathrm{if}\stackrel{~}{\alpha _r}=\stackrel{~}{o_r}`$ (52) $`=`$ $`0\mathrm{otherwise}.`$ From (50) and (51) it follows that $`\widehat{H}_{\alpha _{(r)}}`$ are represented by unitary operators on $`\psi L^2(\mathrm{\Delta }_r,d\mu _H)`$ by $$(\widehat{H}_{\alpha _{(r)}}\psi )(h)=h([\alpha ]_r)\psi (h),h\mathrm{\Delta }_r.$$ (53) We construct electric field operators on $`L^2(\mathrm{\Delta }_r,d\mu _H)`$ as unitary images of appropriate electric field operators on $`L^2(\mathrm{\Delta }_,d\mu _H)`$ as follows. Define the classical Gaussian smeared electric field as $$E_r^a(\stackrel{}{x}):=d^3yf_r(\stackrel{}{y}\stackrel{}{x})E^a(\stackrel{}{y})$$ (54) where $`f_r`$ has been defined in section 3. Given $`\psi _{\{[\alpha _i]\}}𝒞L^2(\mathrm{\Delta },d\mu _H)`$ it can be checked that $$(\widehat{E}_r^a(\stackrel{}{x})\psi _{\{[\alpha _i]\}})(h)=e\underset{i=1}{\overset{n}{}}X_{\alpha _{i(r)}}^a(\stackrel{}{x})h([\alpha _i])\frac{\psi _{\{[\alpha _i]\}}}{h([\alpha _i])}.$$ (55) The methods of can be used to show that $`\widehat{E}_r^a(\stackrel{}{x})`$ is essentially self adjoint on $`𝒞`$. Note that (55) implies that $$[\widehat{E}_r^a(\stackrel{}{x}),\widehat{H}_\alpha ]=eX_{\alpha _{(r)}}^a(\stackrel{}{x})\widehat{H}_\alpha .$$ (56) The unitary image of (56) is $$[U_r\widehat{E}_r^a(\stackrel{}{x})U_r^1,\widehat{H}_{\alpha _{(r)}}]=eX_{\alpha _{(r)}}^a(\stackrel{}{x})\widehat{H}_{\alpha _{(r)}}.$$ (57) Denote the classical counterpart of $`U_r\widehat{E}_r^a(\stackrel{}{x})U_r^1`$ by $`F^a(\stackrel{}{E})`$. Then (57) provides a quantum representation of the classical Poisson bracket, $$\{F^a(\stackrel{}{E}),H_{\alpha _{(r)}}(A)\}=ieX_{\alpha _{(r)}}^a(\stackrel{}{x})H_{\alpha _{(r)}}(A).$$ (58) Note that $`\{E^a(\stackrel{}{x}),H_{\alpha _{(r)}}(A)\}=ieX_{\alpha _{(r)}}^a(\stackrel{}{x})H_{\alpha _{(r)}}(A)`$. Hence, we can consistently identify $`F^a(\stackrel{}{E})`$ with $`E^a(\stackrel{}{x})`$. Thus, $`U_r\widehat{E}_r^a(\stackrel{}{x})U_r^1=\widehat{E}^a(\stackrel{}{x})`$ and from (55), $$(\widehat{E}^a(\stackrel{}{x})\psi _{\{[\alpha _i]_r\}})(h)=e\underset{i=1}{\overset{n}{}}X_{\alpha _{i(r)}}^a(\stackrel{}{x})h([\alpha _i]_r)\frac{\psi _{\{[\alpha _i]_r\}}(h)}{h([\alpha _i]_r)}.$$ (59) Since $`U_r`$ is unitary, $`\widehat{E}^a(\stackrel{}{x})`$ is essentially self adjoint on $`𝒞_rL^2(\mathrm{\Delta }_r,d\mu _H)`$. To summarize: The induced Haar representation of $`\overline{𝒜}_r`$ provides a quantum representation of the classical Poisson bracket algebra of smeared holonomies $`H_{\alpha _{(r)}}(A)`$ and (divergence free) electric field $`E^a(\stackrel{}{x})`$. $`\widehat{H}_{\alpha _r}`$ are represented by unitary operators through (53) and the unsmeared electric field operator, $`\widehat{E}^a(\stackrel{}{x})`$, is represented through (59) as an essentially self adjoint operator on the dense domain of cylindrical functions, $`𝒞_rL^2(\mathrm{\Delta }_r,d\mu _H)`$. Note that $`\widehat{E}^a(\stackrel{}{x})`$ is a genuine operator as opposed to an operator valued distribution! ### 5b. Fock representation of $`𝒜`$ We denote the ‘Fock’ PLF on $`𝒜_r`$ as well as its image on $`𝒜`$ by $`\mathrm{\Gamma }_F`$. Note that the induced PLF on $`𝒜`$ is defined by $$\mathrm{\Gamma }_F(\underset{i=1}{\overset{N}{}}a_i[\alpha _i]):=\underset{i=1}{\overset{N}{}}a_i\mathrm{exp}(\frac{d^3k}{k}|X_{\alpha _{i(r)}}^a(\stackrel{}{k})|^2).$$ (60) Since $`\widehat{H}_{\alpha _{(r)}}`$ are represented as unitary operators on $``$, it follows that $$\widehat{H}_\alpha :=U_r^1\widehat{H}_{\alpha _{(r)}}U_r$$ (61) are represented as unitary operators on $`U_r^1()`$. It remains to construct, following the strategy of section 5a, electric field operators on $`U_r^1()`$ as unitary images of appropriate electric field operators on $``$. On $`U^1(𝒞_F)`$, $$[\widehat{E}(\stackrel{}{f}),\widehat{H}_{\alpha _{(r)}}]=ed^3xf_a(\stackrel{}{x})X_{\alpha _{(r)}}^a(\stackrel{}{x})\widehat{H}_{\alpha _{(r)}}.$$ (62) $$[U_r^1\widehat{E}(\stackrel{}{f})U_r,\widehat{H}_\alpha ]=ed^3xf_a(\stackrel{}{x})X_{\alpha _{(r)}}^a(\stackrel{}{x})\widehat{H}_\alpha .$$ (63) Define the classical function $$E_r(\stackrel{}{f}):=d^3xf_a(\stackrel{}{x})E_r^a(\stackrel{}{x}),$$ (64) where $`E_r^a(\stackrel{}{x})`$ is defined by (54). Since $$\{E_r(\stackrel{}{f}),H_\alpha (A)\}=ied^3xf_a(\stackrel{}{x})X_{\alpha _{(r)}}^a(\stackrel{}{x})H_\alpha (A),$$ (65) we identify $$\widehat{E}_r^a(\stackrel{}{f}):=U_r^1\widehat{E}(\stackrel{}{f})U_r.$$ (66) To summarize: The induced Fock representation of $`𝒜`$ provides a quantum representation of the classical Poisson bracket algebra of holonomies $`H_\alpha (A)`$ and “Gaussian-smeared, smeared” electric fields $`E_r(\stackrel{}{f})`$. The ‘unsmeared’ holonomy operators $`\widehat{H}_\alpha `$ are represented by unitary operators through (61) and $`\widehat{E}(\stackrel{}{f})`$ is represented as a self adjoint operator through (66). Note that the “Gaussian smeared object” $`\widehat{E}_r(\stackrel{}{x})`$ is represented as an operator valued distribution (as opposed to a genuine operator) on $`U_r^1()`$. ## 6. Discussion Preliminary remarks: In this paper, representations of the Poisson algebra of $`U(1)`$ theory were constructed in 2 steps. First, Hilbert space representations of the abelian Poisson algebra of configuration functions (i.e. functions of $`A_a`$) were constructed by specifying a PLF. Second, real functions of the conjugate electric field were represented by self adjoint operators on this Hilbert space. The Haar representation of $`\overline{𝒜}`$ and its image on $`\overline{𝒜}_r`$ support a representation of the electric field wherein, formally, $$\widehat{E}^a(\stackrel{}{x})=i\frac{\delta }{\delta A_a(\stackrel{}{x})}.$$ (67) This action is not connected with Poincare invariance and it is not surprising that the resulting representations of section 2 and 5a, are non Fock representations. On the Fock representation of $`𝒜_r`$ <sup>10</sup><sup>10</sup>10We remind the reader that we displayed a fairly rigorous argument that the entire Fock space is obtained in such a representation through the constructions of section 4 and Appendix A2. We reiterate our belief that the formal equation (88) can be rendered mathematically well defined in a more careful treatement. and its image on $`𝒜`$, equation (67) is incompatible with the requirement of self adjointness of the electric field operators. Their action necessarily contains a term dependent on the Gaussian measure (see (49)) to ensure self adjointness. The choice of Gaussian measure is intimately associated with the properties of the de Alembertian, $`\frac{^2}{t^2}^2`$, and hence with Poincare invariance. A rephrasing of the above remarks which brings them closer to the strategy of is as follows. Given a representation in which (smeared or unsmeared) holonomies are represented by multiplication by unitary operators and the electric field acts, as in (67), purely by functional differentiation, the requirement of self adjointness of the electric field operator determines the Hilbert space measure to be the Haar measure. The self adjointness of electric field operators results in the Gaussian measure only if their action has a contribution dependent on the Gaussian measure. Thus, to obtain the standard Fock representation or the induced one of section 5b, the Gaussian measure and hence, Poincare invariance, plays an essential and explicit role. Note that this work concerns the ‘connection’ representation of a theory of a real $`U(1)`$ connection. In contrast constructs the loop representation of a description of linearised gravity based on a self dual connection <sup>11</sup><sup>11</sup>11Note, however that the descripton reduces to one in terms of a triplet of abelian connections.. Despite these differences, there is also a certain amount of shared mathematical structure in our work and . Therefore, the delineation of the structures involved in the construction of $`U(1)`$ theory as spelt out in this paper, allows us to identify the role of the key structures in . Discussion of (I) and (II): We use the notation of and when discussing those papers. We first discuss (I). In the action of the linearised metric variable in the loop representation is deduced, ultimately, from its action of the form ‘$`i\frac{\delta }{\delta A_a^i}`$’ in the connection representation. The loop representation then becomes an ‘electric field’ type representation in which the magnetic field operator acts purely by functional differentiation with respect to the loop form factor. Yet a Fock representation (of the positive and negative helicity gravitons) results in apparent contradiction to our claims that such a representation cannot result without using Poincare invariance explicitly. The resolution of this apparent contradiction for the positive helicity graviton sector seems to lie, in what appears at first sight, to be a mere mathematical nicety. In the Gaussian measure contribution to the $`\widehat{B}^+`$ operator is absorbed (and hidden) in the rescaling of the wave function. Such a rescaling is permissible for finite dimensional systems but results in a mathematically ill defined ‘measure’ for the field theory in question. In spite of the fact that for most applications this formal treatment suffices, it is crucial to realise, in the context of (I), that it hides the role of Poincare invariance in constructing the Fock representation. To obtain a well defined (Gaussian) measure, the wave functions ((101) of ) need to be rescaled and a Gaussian measure term needs to be added to the action of the $`\widehat{B}^+`$ \- this, of course, feeds explicit Poincare invariance back into the construction. Note that this argument does not apply to the negative helicity sector. There, the choice of self dual connection results in the negative helicity magnetic field operator, $`\widehat{B}^{}`$ being the same as the negative helicity annihilation operator. $`\widehat{B}^{}`$ is naturally represented as a functional derivative ((70) of ) and, in this aspect, matches the standard Fock representation of the annihilation operator as a pure functional derivative term. The resulting representation is the Fock representation for negative helicity gravitons and indeed, for this sector, it seems that explicit Poincare invariance is not invoked. Thus the Fock representation of linearised gravity seems to result partly due to explicit Poincare invariance (which is suppressed in by a mathematically ill defined operation) and partly due to the use of self dual connections. Considerations similar to those for the positive helicity gravitons also apply to the treatment of free Maxwell theory in . There, it is shown that the extended loop representation coincides, formally, with the electric field representation. Again, an ill defined measure is used and a proper mathematical treatment restores the explicit role of Poincare invariance. We turn now, to a discussion of (II). In the loop representation of the two sets of important operators are the magnetic field, $`\widehat{B}^\pm `$, and the linearised metric, $`\widehat{h}^\pm `$. They are represented by functional differentiation and multiplication, on the representation space of functionals of loop form factors. This representation space supports the holonomies as operators. Indeed, the action of the $`\widehat{B}^\pm `$ operators is deduced from the fact that the classical magnetic flux is the lowest non trivial term in the expansion of the holonomy of a small loop ((57) of ). However, all these constructions are rendered formal because of the distributional nature of the loop form factor and the resulting divergence of the ground state functional. Therefore a regularization procedure is adopted wherein the loop form factors are replaced by their Gaussian smeared, $`r`$\- versions (see (9)) and $`\widehat{B}^\pm `$, $`\widehat{h}^\pm `$ are represented as functional differentiation and multiplication operators on the space of functionals of $`r`$-loop form factors. An important question is: Are the holonomy operators or some regularized version thereof, represented on this space? One may choose to ignore this question and simply postulate the action of $`\widehat{B}^\pm `$, $`\widehat{h}^\pm `$ in terms of $`r`$-loop form factors. Then the primary configuration variables of the theory are $`B^\pm `$ and the construction does not seem to have much to do with loops and holonomies. Since holonomies and loops are the primary objects in the loop approach to full-blown quantum gravity, an interpretation of the regularization which allows for the representation of holonomy operators is of interest. It seems to us that such an interpretation must regard the $`r`$\- form factor representation as an approximation to the standard Fock representation, which becomes better as $`r0`$. A precise formulation of such an interpretation in the context of $`U(1)`$ theory is provided by the induced Fock representation of section 5b. There, the holonomy operators are represented on the Hilbert space and the magnetic field operators can be constructed by a “shrinking of loop” limit, as the image of $`U^1\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}U`$ via $`U_r^1`$. That representation, although not the standard Fock representation, is a good approximation to it for small $`r`$. The nature of the approximation is as follows. For sufficiently small $`r`$, the holonomies $`H_\gamma (A)`$, the electric field $`E^a(\stackrel{}{x})`$ and their Gaussian smeared counterparts, $`H_{\gamma _{(r)}}(A)`$, $`E_r^a(\stackrel{}{x})`$ approximate each other well. An approximate Fock representation can be constructed in which the operators corresponding to $`H_\gamma (A)`$, $`E_r^a(\stackrel{}{x})`$ act in the same way as the operators corresponding to $`H_{\gamma _{(r)}}(A)`$, $`E^a(\stackrel{}{x})`$ in the standard Fock representation. This approximate Fock representation is the induced Fock representation of section 5b. To summarize: The standard Fock representation for $`U(1)`$ theory is obtained only when the algebra of smeared holonomies is used and explict Poincare invariance is invoked. However, the role of Poincare invariance (or equivalently, the choice of PLF) seems to be more important than that of smeared loops. If the requirement of smeared loops is dropped, it is still possible to construct an approximate Fock representation; but dropping Poincare invariance results in the non Fock representations of section 2 and 5a. Comments: (i) The ‘area derivative’ plays an important role in some approaches to loop quantum gravity . Our construction of $`\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}|\psi >`$ (or its image in the induced Fock representation of 5b) as a Cauchy limit is a rigorous realization of the area derivative in the context of Fock like representations. Note that the required limits do not exist in the Haar representation and hence the area derivative is ill defined there. (ii) As noted above, self duality of the connection plays a key role in obtaining the (negative helicity) graviton Fock representation without explicit recourse to Poincare invariance (see for a detailed examination of the relation between self duality and helicity). However, recent efforts in loop quantum gravity use real (as opposed to self dual) connections. It would be useful to reformulate linearised gravity in terms of real connections and construct its quantization. (iii) Note that we have mainly been concerned with the kinematics of $`U(1)`$ theory. The Fock representation, of course, supports the Maxwell Hamiltonian as an operator. Note that the normal ordering prescription adopted in is, of course, connected with Poincare invariance. It is an open question as to how to express (presumably an approximation of) the Hamiltonian as an operator in the Haar representation. (iv) We have not been able to show continuity of the Fock PLF on $`𝒜_r`$ or lack thereof. If the Fock PLF is continuous, a ‘Fock’ measure, $`d\mu _F`$, can be constructed on $`\mathrm{\Delta }_r`$ and $``$ can be identified with $`L^2(\mathrm{\Delta }_r,d\mu _F)`$. The considerations of section 5b can then be extended to the $`C^{}`$ algebras $`\overline{𝒜}_r`$ and $`\overline{𝒜}`$. If, however, the Fock PLF on $`𝒜_r`$ turns out not to be continuous, then a corresponding Fock measure on $`\mathrm{\Delta }_r`$ does not exist and it is incorrect to identify $`\mathrm{\Delta }_r`$ with the ‘quantum configuration space’. If this is indeed the case, then the emphasis on continuous cyclic representations of $`\overline{𝒜}`$ in loop quantum gravity would seem unduly restrictive. (v) The representation of kinematic loop quantum gravity is the $`SU(2)`$ counterpart of the Haar representation for $`U(1)`$ theory. An important question is how the Fock space-graviton description of linearised gravity arises out of loop quantum gravity. It is possible that some insight into this issue may be obtained by considering the following (simpler) question in the context of $`U(1)`$ theory. Is there any way in which an approximate Fock structure can be obtained from the Haar representation of $`U(1)`$ theory? Since the PLFs play a key role in determining the type of representation, this work suggests that to get an approximate Fock structure, it may be a good strategy to try to approximate (in some, yet unknown way) the Fock PLF by the Haar PLF. Acknowledgements: I am very grateful to Jose Zapata for useful discussions and encouragement. ## Appendix ### A1 Lemma 1: Given (i) $`\gamma _i_{x_0}`$, $`i=1..n`$, $`n`$ finite, (ii) $`A_a(\stackrel{}{x})𝒜`$ and (iii) $`ϵ>0`$, there exists a connection $`A_a^ϵ(\stackrel{}{x})𝒜`$ such that $$|H_{\gamma _{i(r)}}(A^ϵ)H_{\gamma _i}(A)|<ϵ$$ (68) for $`i=1..n`$. Proof: For a single loop $`\gamma `$, from (8) $$|X_\gamma ^a(\stackrel{}{k})|<C_\gamma C_\gamma :=\frac{3}{(2\pi )^{\frac{3}{2}}}L_\gamma $$ (69) where $`L_\gamma `$ is the length of the loop as measured by the flat metric. Since $`A_a(\stackrel{}{x})`$ is Schwartz, we have, for arbitrarily large $`N>0`$, $$|A_a(\stackrel{}{k})|<\frac{C_N}{k^N}\mathrm{for}\mathrm{some}C_N>0.$$ (70) From (69) and (70) $$_{k>\mathrm{\Lambda }}d^3k|X_\gamma ^a(\stackrel{}{k})A_a(\stackrel{}{k})|<\frac{C_{N,\gamma }}{\mathrm{\Lambda }^{N3}},C_{N,\gamma }=\frac{4\pi C_\gamma C_N}{N1}.$$ (71) Thus, given $`\delta >0`$, there exists $`\mathrm{\Lambda }(\gamma ,\delta )`$ such that $$_{k>\mathrm{\Lambda }(\gamma ,\delta )}d^3k|X_\gamma ^a(\stackrel{}{k})A_a(\stackrel{}{k})|<\delta .$$ (72) Let $`f(k)>0`$ be a smooth function such that $`f(k)`$ $`=`$ $`e^{\frac{k^2r^2}{2}}\mathrm{for}k<\mathrm{\Lambda }(\gamma ,\delta )`$ (73) $`<`$ $`e^{\frac{k^2r^2}{2}}\mathrm{for}\mathrm{\Lambda }(\gamma ,\delta )<k<2\mathrm{\Lambda }(\gamma ,\delta )`$ $`=`$ $`1\mathrm{for}k>2\mathrm{\Lambda }(\gamma ,\delta )`$ Define $`A_{a(r)}^\delta (\stackrel{}{x})`$ through its Fourier transform, $$A_{a(r)}^\delta (\stackrel{}{k}):=f(k)A_a(\stackrel{}{k}).$$ (74) Note that $`A_{a(r)}^\delta (\stackrel{}{x})𝒜`$. From (11), (74), and (72) it follows that $$|_{k>\mathrm{\Lambda }(\gamma ,\delta )}d^3kX_{\gamma _{(r)}}^a(\stackrel{}{k})A_{a(r)}^\delta (\stackrel{}{k})|<\delta .$$ (75) From (11) and (74) $$_{k<\mathrm{\Lambda }(\gamma ,\delta )}d^3kX_{\gamma _{(r)}}^a(\stackrel{}{k})A_{a(r)}^\delta (\stackrel{}{k})=_{k<\mathrm{\Lambda }(\gamma ,\delta )}d^3kX_\gamma ^a(\stackrel{}{k})A_a(\stackrel{}{k}).$$ (76) Using (76) $$|H_{\gamma _{i(r)}}(A^ϵ)H_{\gamma _i}(A)|=|\mathrm{exp}i(_{k>\mathrm{\Lambda }(\gamma ,\delta )}d^3kX_{\gamma _{(r)}}^a(\stackrel{}{k})A_{a(r)}^\delta (\stackrel{}{k})X_\gamma ^a(\stackrel{}{k})A^a(\stackrel{}{k})).$$ (77) From (72), (75) and (77), for small enough $`\delta >0`$, it can be seen that $$|H_{\gamma _{i(r)}}(A^ϵ)H_{\gamma _i}(A)|<4\delta $$ (78) For the loops $`\gamma _i,i=1..n`$, $`\overline{\mathrm{\Lambda }}(\delta )`$ $`:=`$ $`\underset{i}{\mathrm{max}}\mathrm{\Lambda }(\delta ,\gamma _i)`$ $`\overline{A}_{a(r)}^\delta (\stackrel{}{k})`$ $`:=`$ $`\overline{f}A_a(\stackrel{}{k})`$ (79) with $`\overline{f}(k)>0`$ a smooth function such that $`\overline{f}(k)`$ $`=`$ $`e^{\frac{k^2r^2}{2}},\mathrm{for}k<\overline{\mathrm{\Lambda }}(\delta )`$ (80) $`<`$ $`e^{\frac{k^2r^2}{2}}\mathrm{for}\overline{\mathrm{\Lambda }}(\delta )<k<2\overline{\mathrm{\Lambda }}(\delta )`$ $`=`$ $`1\mathrm{for}k>2\overline{\mathrm{\Lambda }}(\delta )`$ Then, given $`ϵ>0`$, choose some $`\delta \frac{ϵ}{4}`$ and set $$A_a^ϵ(\stackrel{}{k}):=\overline{A}_{a(r)}^{4\delta }(\stackrel{}{k}).$$ (81) Then (68) holds. Lemma 2: Given (i) strongly independent loops $`\gamma _i`$, $`i=1..n`$, $`n`$ finite, (ii) $`g_iU(1)`$, $`i=1..n`$ and (iii) $`ϵ>0`$, there exists a connection $`A_a^ϵ(\stackrel{}{x})𝒜`$ such that $$|H_{\gamma _{i(r)}}(A^ϵ)g_i|<ϵ$$ (82) for $`i=1..n`$. Proof: From , $`A_a𝒜`$ exists such that $`H_{\gamma _i}(A)=g_i`$, $`i=1..n`$. Therefore it suffices to construct $`A_a^ϵ`$ such that $`|H_{\gamma _{i(r)}}(A^ϵ)H_{\gamma _i}(A)|<ϵ`$. But this is exactly the content of Lemma 1. Lemma 3:Given $`A_a(\stackrel{}{x})𝒜`$, there exists $`A_{a(r)}(\stackrel{}{x})𝒜`$ such that $$H_{\gamma _{(r)}}(A)=H_\gamma (A_{(r)}).$$ (83) for every $`\gamma _{x_0}`$. Proof: From (11) and (12) it immediately follows that the required $`A_{a(r)}(\stackrel{}{x})`$ is determined by its Fourier transform via $`A_{a(r)}(\stackrel{}{k})=e^{\frac{k^2r^2}{2}}A_a(\stackrel{}{k})`$. ### A2 Proposition: The states $`|\psi _{\{\stackrel{}{x}_i,\stackrel{}{v}_i\}}>:=_{i=1}^p\widehat{O}_{\stackrel{}{x}_i,\stackrel{}{v}_i}|0>,(p=1,2..)`$ together with $`|0>`$, span $`𝒟_0`$. Heuristic Proof: The argument below is a bit formal, but we expect that it can be converted to a rigorous proof. Define $$|\psi ,p>:=d^3k_1..d^3k_p\psi ^{a_1..a_p}(\stackrel{}{k}_1..\stackrel{}{k}_p)(\underset{i=1}{\overset{p}{}}\widehat{a}_{a_i}(\stackrel{}{k}_i)+\widehat{a}_{a_i}^{}(\stackrel{}{k}_i))|0>,$$ (84) where $`\psi ^{a_1..a_p}(\stackrel{}{k}_1..\stackrel{}{k}_p)`$ has the same properties (a)-(c) (see section 4) as $`\varphi ^{a_1..a_p}(\stackrel{}{k}_1..\stackrel{}{k}_p)`$. $`|\psi ,p>`$ along with $`|0>`$ span $`𝒟_0`$. $`|\psi ,p>`$ can be generated from $`|\psi _{\{\stackrel{}{x}_i,\stackrel{}{n}_i\}}>`$ as follows. Note that from (39), $$\frac{1}{2\pi ^{\frac{3}{2}}}d^3xe^{i\stackrel{}{k}\stackrel{}{x}}\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}=\frac{i(\stackrel{}{n}\times \stackrel{}{k})^a}{\sqrt{2}k}e^{\frac{k^2r^2}{2}}(\widehat{a}_a(\stackrel{}{k})+\widehat{a}_a^{}(\stackrel{}{k})).$$ (85) Define $$g^{a_1..a_p}(\stackrel{}{k}_1..\stackrel{}{k}_p)=(\underset{i=1}{\overset{p}{}}e^{k_i^2r^2}\sqrt{2}k_i)\psi ^{a_1..a_p}(\stackrel{}{k}_1..\stackrel{}{k}_p).$$ (86) Given $`\stackrel{}{k}_ii=1..p`$, it is possible to construct a triplet of vectors $`\stackrel{}{u}_{a_i}^i(\stackrel{}{k}_i)`$ ($`a_i=1,2,3`$ for each $`i`$), such that <sup>12</sup><sup>12</sup>12 An explicit choice is as follows. Fix Cartesian coordinates $`(x,y,z)`$ and the corresponding unit vectors $`(\widehat{x},\widehat{y},\widehat{z})`$. Then for $`i=1..p`$, $`\stackrel{}{u}_1^i=\frac{\stackrel{}{k}_i\times \widehat{x}}{k}`$, $`\stackrel{}{u}_2^i=\frac{\stackrel{}{k}_i\times \widehat{y}}{k}`$, $`\stackrel{}{u}_3^i=\frac{\stackrel{}{k}_i\times \widehat{z}}{k}`$. $$\frac{(\stackrel{}{u}_{b_i}^i\times \stackrel{}{k}_i)^{a_i}}{k}=\delta _{b_i}^{a_i}\frac{k^{a_i}k_{b_i}}{k^2}.$$ (87) Then from (85), (86) and (87), $$|\psi ,p>=(\underset{l=1}{\overset{p}{}}d^3k_l\underset{m=1}{\overset{p}{}}d^3x_m)g^{a_1..a_p}(\stackrel{}{k}_1..\stackrel{}{k}_p)(\underset{i=1}{\overset{p}{}}\frac{e^{i\stackrel{}{k}_i\stackrel{}{x}_i}}{2\pi ^{\frac{3}{2}}})|\psi _{\{\stackrel{}{x}_i,\stackrel{}{u}_{a_i}^i\}}>$$ (88) It is in this formal sense that states of the type $`|\psi _{\{\stackrel{}{x}_i,\stackrel{}{v}_i\}}>`$ together with $`|0>`$ span $`𝒟_0`$. Lemma 2: $`\widehat{O}_{\stackrel{}{x},\stackrel{}{n}}`$ is a densely defined, symmetric operator on the dense domain $`𝒟_0`$, which admits self adjoint extensions. Proof: It is straightforward to check that $`\widehat{O}_{(\stackrel{}{x},\stackrel{}{n})}|\varphi ,p>`$ $`=`$ $`{\displaystyle }({\displaystyle \underset{i=1}{\overset{p+1}{}}}d^3k_i)f^{a_{p+1}}(\stackrel{}{k}_{p+1})\varphi ^{a_1..a_p}(\stackrel{}{k}_1..\stackrel{}{k}_p)({\displaystyle \underset{j=1}{\overset{p+1}{}}}\widehat{a}_{a_j}^{}(\stackrel{}{k}_j))|0>`$ (89) $`+`$ $`p{\displaystyle }({\displaystyle \underset{i=1}{\overset{p}{}}}d^3k_i)f_{a_1}(\stackrel{}{k}_1)\varphi ^{a_1..a_p}(\stackrel{}{k}_1..\stackrel{}{k}_p)({\displaystyle \underset{j=2}{\overset{p+1}{}}}\widehat{a}_{a_j}^{}(\stackrel{}{k}_j))|0>`$ where $$f^a(\stackrel{}{k}):=\frac{i}{2\pi ^{\frac{3}{2}}}e^{i\stackrel{}{k}\stackrel{}{x}}\frac{(\stackrel{}{n}\times \stackrel{}{k})^a}{\sqrt{2}k}e^{\frac{k^2r^2}{2}}.$$ (90) The ultraviolet behaviour of $`\varphi ^{a_1..a_p}(\stackrel{}{k}_1..\stackrel{}{k}_p)`$, $`f^a(\stackrel{}{k})`$ ensures that $`\widehat{O}_{(\stackrel{}{x},\stackrel{}{n})}|\varphi ,p>`$ is finite. Thus $`\widehat{O}_{(\stackrel{}{x},\stackrel{}{n})}`$ is densely defined on $`𝒟_0`$. By inspection $`\widehat{O}_{(\stackrel{}{x},\stackrel{}{n})}`$ is also symmetric on $`𝒟_0`$. To show existence of its self adjoint extensions, it is sufficient to exhibit an antilinear operator $`\widehat{C}`$ on $``$ with $`\widehat{C}^2=\mathrm{𝟏}`$ which leaves $`𝒟_0`$ invariant and commutes with $`\widehat{O}_{(\stackrel{}{x},\stackrel{}{n})}`$ . As in , take $`\widehat{C}`$ to be the complex conjugation operator (in the standard Schrodinger representation) on $`=L^2(𝒮^{},d\mu )`$ where $`𝒮^{}`$ is the appropriate space of tempered distributions and $`d\mu `$ is the standard Gaussian measure, for free Maxwell theory. It can be seen that $`\widehat{C}a_a(\stackrel{}{k})=a_a(\stackrel{}{k})\widehat{C}`$ and $`\widehat{C}a_a^{}(\stackrel{}{k})=a_a^{}(\stackrel{}{k})\widehat{C}`$, and hence $`\widehat{O}_{(\stackrel{}{x},\stackrel{}{n})}\widehat{C}=\widehat{C}\widehat{O}_{(\stackrel{}{x},\stackrel{}{n})}`$. Lemma 3: $$\frac{(e^{i\widehat{D}}1)}{i\pi ϵ_m^2}|\psi >0.$$ (91) as $`ϵ_m0`$. Proof: $$i\widehat{D}=id^3kh^a(\stackrel{}{k})(\widehat{a}_a(\stackrel{}{k})+\widehat{a}_a^{}(\stackrel{}{k}))$$ (92) with $$h^a(\stackrel{}{k})=\frac{e^{i\stackrel{}{k}\stackrel{}{x}}e^{\frac{k^2r^2}{2}}}{\sqrt{2}k2\pi ^{\frac{3}{2}}}\left(𝑑se^{i\stackrel{}{k}(\stackrel{}{\gamma }^{\{m,\stackrel{}{x},\stackrel{}{n}\}}\stackrel{}{x})}\dot{\gamma }^{a\{m,\stackrel{}{x},\stackrel{}{n}\}}i\pi ϵ_m^2(\stackrel{}{n}\times \stackrel{}{k})^a\right)$$ (93) A straightforward calculation, using , shows that $$h^a(\stackrel{}{k})=\frac{ie^{i\stackrel{}{k}\stackrel{}{x}}e^{\frac{k^2r^2}{2}}}{\sqrt{2}k2\pi ^{\frac{3}{2}}}(\stackrel{}{n}\times \stackrel{}{k})^a\pi ϵ_m^2\left(\frac{2J_1(\alpha _kϵ_m)}{\alpha _kϵ_m}1\right)$$ (94) with $`\alpha _k:=|\stackrel{}{n}\times \stackrel{}{k}|`$. Now, from (92) and (37), $$e^{i\widehat{D}}=e^{{\scriptscriptstyle d^3k|h^a(\stackrel{}{k})h_a(\stackrel{}{k})|}}e^{i{\scriptscriptstyle d^3kh^a(\stackrel{}{k})\widehat{a}_a^{}(\stackrel{}{k})}}e^{i{\scriptscriptstyle d^3kh^a(\stackrel{}{k})\widehat{a}_a(\stackrel{}{k})}}.$$ (95) $`<\varphi ,p|e^{i\widehat{D}}|\varphi ,p>`$ $`=`$ $`e^{{\scriptscriptstyle d^3k|h^a(\stackrel{}{k})h_a(\stackrel{}{k})|}}{\displaystyle }({\displaystyle \underset{i=1}{\overset{p}{}}}d^3k_i{\displaystyle \underset{j=1}{\overset{p}{}}}d^3l_j)\varphi ^{a_1..a_p}(\stackrel{}{k}_1..\stackrel{}{k}_p)\varphi ^{b_1..b_p}(\stackrel{}{l}_1..\stackrel{}{l}_p)`$ (96) $`<0|({\displaystyle \underset{j=1}{\overset{p}{}}}\widehat{a}_{b_j}(\stackrel{}{l}_j)ih_{b_j}^{}(\stackrel{}{l}_j))({\displaystyle \underset{i=1}{\overset{p}{}}}\widehat{a}_{a_i}^{}(\stackrel{}{k}_i)+ih_{a_i}(\stackrel{}{k}_i))|0>`$ $`<\varphi ,p|e^{i\widehat{D}}1|\varphi ,p>`$ $`=`$ $`n!(e^{{\scriptscriptstyle d^3k|h^a(\stackrel{}{k})h_a(\stackrel{}{k})|}}1){\displaystyle }{\displaystyle \underset{i=1}{\overset{p}{}}}d^3k_i|\varphi ^{a_1..a_p}(\stackrel{}{k}_1..\stackrel{}{k}_p)|^2`$ $`+n!ne^{{\scriptscriptstyle d^3k|h^a(\stackrel{}{k})h_a(\stackrel{}{k})|}}`$ $`{\displaystyle }({\displaystyle \underset{i=2}{\overset{p}{}}}d^3k_i)d^3kd^3lh_{a_1}(\stackrel{}{k})h^{b_1}(\stackrel{}{l})\varphi ^{a_1a_2..a_p}(\stackrel{}{k},\stackrel{}{k}_2..,\stackrel{}{k}_p)\varphi _{b_1a_2..a_p}(\stackrel{}{l},\stackrel{}{k}_2..,\stackrel{}{k}_p)`$ (97) $`+`$ $`O(h^4).`$ Since $`\left(\frac{2J_1(\alpha _kϵ_m)}{\alpha _kϵ_m}1\right)`$ is a bounded function, (94) implies that the $`O(h^4)`$ terms do not contribute to (91) in the $`ϵ_m0`$ limit. From Lemma 4 below and (94) the first term of (97) is of order $`ϵ_m^{5\frac{1}{2}}`$ and the second is of order $`ϵ_m^5`$. From this is it is clear that $`\frac{(e^{i\widehat{D}}1)}{i\pi ϵ_m^2}|\psi >0`$ as $`ϵ_m0`$. Lemma 4: Let $`n`$ be a positive integer and $`g(\stackrel{}{k})`$ be a bounded function of rapid decrease (i.e. it falls to zero as $`k\mathrm{}`$, faster than any inverse power of $`k`$.). Then, as $`ϵ0`$, $$I:=|d^3kg(\stackrel{}{k})\left(\frac{2J_1(\alpha _kϵ)}{\alpha _kϵ}1\right)^n|<Cϵ^{n\frac{1}{2}}$$ (98) for some positive constant $`C`$ which depends on $`n`$ and $`g`$. Proof: $$I_{kϵ^{\frac{1}{2}}}d^3k|g(\stackrel{}{k})\left(\frac{2J_1(\alpha _kϵ)}{\alpha _kϵ}1\right)^n|+_{k>ϵ^{\frac{1}{2}}}d^3k|g(\stackrel{}{k})\left(\frac{2J_1(\alpha _kϵ)}{\alpha _kϵ}1\right)^n|.$$ (99) In the first term the range of integration is such that $`\alpha _kϵ<ϵ^{\frac{1}{2}}`$. A straightforward calculation shows that the small argument expansion of $`J_1(\alpha _kϵ)`$ coupled with the rapid fall off property of $`g(\stackrel{}{k})`$ gives the bound $$_{kϵ^{\frac{1}{2}}}d^3k|g(\stackrel{}{k})\left(\frac{2J_1(\alpha _kϵ)}{\alpha _kϵ}1\right)^n|C_1(g,n)ϵ^{n\frac{1}{2}}.$$ (100) where $`C_1(g,n)`$ is a positive constant dependent on both $`n`$ and the properties of $`g`$. The rapid decrease property of $`g(\stackrel{}{k})`$ ensures that, for small enough $`ϵ`$, the second term of (99) falls off much faster than the first term. Hence, $`I<Cϵ^{n\frac{1}{2}}`$ where we have set $`C:=2C_1(g,n)`$.
warning/0001/cond-mat0001269.html
ar5iv
text
# Microscopic self-consistent theory of Josephson junctions including dynamical electron correlations. ## I Introduction Some of the most successful and most promising electronics applications of superconductors involve Josephson junctions, where two superconducting regions are coupled through a barrier region, made from a non-superconducting material . The drive to make electronic devices using Josephson junctions has been motivated by their naturally high operating frequencies, far in excess of the speeds obtainable in standard silicon technology. The operational frequency of an ideal Josephson junction using rapid single flux quantum (RSFQ) logic is $`I_cR_Ne/\mathrm{}`$, where $`I_c`$ is the critical current, and $`R_N`$ is the normal state resistance of the junction . Hence a maximum value of the product, $`I_cR_N`$ leads to optimal performance. At present, low temperature superconductors have been used to produce junctions with $`I_cR_N`$ products up to 1mV, and operational speeds reaching $`770`$ GHz , while junctions of high-temperature superconductors have achieved $`I_cR_N`$ products reaching from $`1`$ mV to $`20`$ mV . The choice of barrier material strongly affects both $`I_c`$ and $`R_N`$ (typically poor conductors increase $`R_N`$ but reduce $`I_c`$) so it is appropriate to study in detail what properties of the junction modify $`I_c`$, $`R_N`$, and $`I_cR_N`$. Indeed, it has been suggested , that the maximum value of the product would be reached when the barrier material is close to a metal-insulator transition, so that the system is near the cross-over from an SNS to an SIS junction. In order to investigate such non-trivial barrier materials, we develop a microscopic model of a Josephson junction, which self-consistently incorporates the dynamical correlation effects of the electrons in calculating the conductance of the junction and the critical current. Our model is also appropriate for ballistic junctions that have relatively pure barrier regions so that the extent of the barrier is smaller than its mean-free-path but larger than its proximity-effect-induced superconducting correlation length. Much work has progressed on these systems recently, with a concentration on Niobium-Indium Arsenide-Niobium junctions or Niobium-Silicon-Niobium junctions. Recent work includes an examination of subgap structures and current deficits , quasiparticle reflection effects and spikes in the conductance , effects of multiple Andreev reflection and noise in stacked junctions , and an investigation of a tunable junction that can be altered from an ordinary junction to a $`\pi `$-junction by driving the barrier into a nonequilibrium state via a transverse electrical current . The above-mentioned experimental and theoretical work has concentrated on nonequilibrium effects that occur at finite voltages. This contribution will address only equilibrium and linear-response properties of the microscopic model for Josephson junctions, but our work can be extended to examine such nonequilibrium effects as well. Josephson junctions consist of two superconducting regions coupled through an intermediate region which is not naturally superconducting. The main characteristics of a Josephson junction can be understood by just considering two superconductors coupled together with a single, energy independent transmissivity parameter. The dc Josephson effect, of a direct supercurrent through the junction at zero external voltage, and the ac Josephson effect of an oscillating, dissipative current at finite voltage were both predicted from such a simple model . Further details of the I-V characteristics, such as excess current and sub-harmonic gap structure due to Andreev bound states were explained , by matching boundary conditions across a step-function in the superconducting potential, and including an adjustable scattering potential at the interfaces . Analytic calculations have been performed to provide current-phase relationships , representing the barrier as a single scattering potential with no spatial extent. In this contribution we compare these traditional approaches with a more microscopic model. Originally the term Josephson junction referred to superconductor - insulator - superconductor ($`SIS`$) tunnel junctions, but here we use the term in its common usage to refer to all “weak links”, including superconductor - normal metal - superconductor ($`SNS`$) junctions. In the latter case, the barrier region can occupy tens to hundreds of nanometres, as the proximity effect ensures that the superconductivity extends into the barrier region, decaying on the scale of the coherence length, $`\xi _N=\mathrm{}v_F/k_BT`$ when in the ballistic regime (where the coherence length is longer than the mean free path in the barrier). In this paper, we present a method for studying the effects of the barrier region on the strength of superconductivity it can support, and hence on the supercurrent it can maintain. Effects of electron correlations are incorporated within the dynamic mean field theory , which leads to a local, frequency-dependent self-energy, which we allow to vary from one plane to the next, while assuming it to remain constant within individual planes. Hence the three-dimensional system becomes inhomogeneous in the one dimension where the current flows (which we label the z-direction). We model the system with two sets of $`N_{SC}`$ planes of superconducting material coupled each on one side to the bulk superconductor, and on the other side sandwiching $`N_b`$ planes of barrier material as depicted in Fig. 1. The total number of planes modeled self-consistently is $`2N_{SC}+N_b=N`$. We note that while in this paper we concentrate on applying the method to Josephson junctions, our method would also be applicable to theoretical studies of correlated electrons at surfaces, single interfaces, or multiple interfaces, as first demonstrated by Potthoff and Nolting . An important advantage of our scheme is that the different (material specific) microscopic models that best describe any particular material can be coupled together across the planes. Hence a superconductor with electron-phonon coupling described by a Holstein model (or even using the appropriate $`\alpha ^2F`$ in Migdal-Eliahberg theory) can be connected to a metallic region with impurities described by the Falicov-Kimball model (or a correlated Hubbard model), and so on. The local self energy for each plane is calculated independently, according to the model best-suited to the particular material. Once a set of local self-energies are evaluated for each plane, the Green’s functions are calculated by finding the inverse of an infinite matrix which includes planes of bulk superconductor extending in the positive and negative z-direction. The inversion process, which is made tractable by a continued fraction representation, couples all of the different planes together, such that a change in the self energy on one plane affects the local interacting Green’s functions on all other planes, particularly those nearby in real space. In order for a supercurrent to flow, a phase gradient must be applied to the superconductor and across the barrier region. The critical current, $`I_c`$, is reached, when the planes with the lowest superconducting order, typically at the center of the barrier region, can no longer support the necessary phase gradient to maintain current continuity. In our model we find the supercurrent as a function of phase variation across the barrier, by solving the system self-consistently at each set of phases. The self-consistency is crucial , as the existence of a current flow affects the value of the superconducting order parameter, both inside and outside the barrier region. Section II contains a detailed description of our method, including the physical approximations used, and the general computer algorithm. In Section III we analyze the current-phase relationships of our results in terms of the transmissivity of a barrier. The results presented in Section IV demonstrate the efficacy of the method to solve some simple models of Josephson junctions, using the Bogoliubov-de Gennes equations, where we demonstrate the effects of self-consistency, in particular on the superconducting order and electron density. Section V includes the results for barriers with impurity scattering, with a description of the calculations of normal-state resistance, and results of resistance and $`I_cR_N`$ products. We conclude in Section VI with some comments on the results, and suggestions of other situations well-suited to our model. ## II Method ### A Model Our method consists of two stages. First, we determine the properties of the bulk boundary regions for a uniform system. When no current flows in the bulk (homogeneous) superconductor, dynamical mean field theory is employed to determine the local self energy, using the local approximation. When a uniform current is flowing, one must include a uniform phase variation in the superconducting order parameter, $`\mathrm{\Delta }(z)`$, for the bulk system. Next, we solve the inhomogeneous problem self-consistently by iteration, for a number of planes, $`N`$, coupled on either side to the uniform bulk solution. That is, giving the plane an index, $`\alpha `$, and defining the central link of the junction to be between planes $`\alpha =0`$ and $`\alpha =1`$, we include $`N_b`$ barrier planes, surrounded by $`N_{sc}`$ self-consistently calculated superconductor planes on each side, such that $`N=N_b+2N_{sc}`$ and planes with index, $`\alpha <1N/2`$ or $`\alpha >N/2`$ are invariant homogeneous bulk planes (see Fig. 1). Typically, we solve systems with $`N_{sc}=30`$, which is significantly greater than the coherence length for our bulk superconductor ($`\xi =\mathrm{}v_F/\mathrm{\Delta }10a`$, where $`a`$ is the lattice spacing). We observe that, except when very close to $`T_c`$, the superconducting order has completely healed from its disruption at the interface, by the time we reach the planes at the bulk superconductor boundary (planes with $`\alpha \pm N/2`$). We describe the system with the following tight-binding model: $$\widehat{H}=\underset{i,j,\sigma }{}t_{ij}\widehat{c}_{i\sigma }^{}\widehat{c}_{j\sigma }+\underset{i}{}U_i\left(\widehat{c}_i^{}\widehat{c}_i\frac{1}{2}\right)\left(\widehat{c}_i^{}\widehat{c}_i\frac{1}{2}\right)+\underset{i,\sigma }{}U_i^{FK}\widehat{c}_{i\sigma }^{}\widehat{c}_{i\sigma }w_i$$ (1) where $`\widehat{c}_{i\sigma }^{}`$ and $`\widehat{c}_{i\sigma }`$ are fermionic operators which respectively create and destroy an electron of spin-$`\sigma `$ in a single Wannier (tight-binding) state on the lattice site $`i`$; $`t_{ij}=\{\begin{array}{cc}ϵ_\alpha & \text{if }i=j\text{ on plane }\alpha \text{,}\hfill \\ t_\alpha & \text{if }i\text{ and }j\text{ are neighboring sites on the same plane, }\alpha \text{,}\hfill \\ \sqrt{t_\alpha t_\alpha ^{}}& \text{if }i\text{ and }j\text{ are neighboring sites on consecutive planes, }\alpha \text{ and }\alpha ^{}\text{,}\hfill \\ 0& \text{otherwise,}\hfill \end{array}`$ with $`t_\alpha `$ the overlap or hopping integral for the $`\alpha `$-th plane, $`ϵ_i`$ the local site energy, $`U_i`$ the renormalized on-site, Hubbard interaction energy, $`U_i^{FK}`$ the impurity potential, and $`w_i=1`$ if there is an impurity on site $`i`$, and $`w_i=0`$ otherwise. The magnitude of the hopping integral in the superconducting region, $`t`$, is constant, and defines our energy scale for the entire system ($`t=1`$). In all results that we present here, the superconducting region has an attractive Hubbard interaction, $`U_i=2`$ and no impurities ($`w_i=0`$ for all sites on planes $`\alpha <1N_b/2`$ and $`\alpha >N_b/2`$). We utilize the spinless Falicov-Kimball model to describe non-magnetic charge impurities within the barrier region. The interaction between dopant atoms and conduction electrons is attractive, and represented by a negative $`U_i^{FK}`$. The average impurity concentration within the barrier is given by, $`\rho _{imp}=\frac{1}{N_b}\left(\frac{a}{L}\right)^2w_i`$, where $`N_b\left(\frac{L}{a}\right)^2`$ is the total number of barrier sites. We also consider systems where there are no impurities within the barrier, but there is a Hubbard interaction for sites within the barrier, $`U_i=U_b`$, that differs from $`U_i=2`$ in the superconductor. In addition, we have included systems where the hopping integral, $`t_\alpha `$, differs in the barrier region, and where the interfacial planes ($`\alpha =1N_b/2`$ and $`\alpha =N_b/2`$ only) have a non-zero local on-site potential, $`ϵ_\alpha =V_{Int}`$. ### B Green’s Function Calculations We use the matrix formulation of Nambu for the Green’s function, $`\underset{¯}{\underset{¯}{G}}(𝐫_i,𝐫_j,i\omega _n)`$, between two lattice sites $`𝐫_i`$ and $`𝐫_j`$ at the Matsubara frequency, $`i\omega _n=i\pi (2n+1)k_BT`$, $$\underset{¯}{\underset{¯}{G}}(𝐫_i,𝐫_j,i\omega _n)=\left(\begin{array}{cc}G(𝐫_i,𝐫_j,i\omega _n)& F(𝐫_i,𝐫_j,i\omega _n)\\ \overline{F}(𝐫_i,𝐫_j,i\omega _n)& G^{}(𝐫_i,𝐫_j,i\omega _n)\end{array}\right),$$ (2) and the corresponding local self energy, $$\underset{¯}{\underset{¯}{\mathrm{\Sigma }}}(𝐫_i,i\omega _n)=\left(\begin{array}{cc}\mathrm{\Sigma }(𝐫_i,i\omega _n)& \varphi (𝐫_i,i\omega _n)\\ \varphi ^{}(𝐫_i,i\omega _n)& \mathrm{\Sigma }^{}(𝐫_i,i\omega _n))\end{array}\right).$$ (3) The diagonal and off-diagonal Green’s functions are defined respectively as: $`G(𝐫_i,𝐫_j,i\omega _n)`$ $`=`$ $`{\displaystyle _0^\beta }𝑑\tau \mathrm{exp}(i\omega _n\tau )\mathrm{T}_\tau \widehat{c}_{j\sigma }(\tau )\widehat{c}_{i\sigma }^{}(0),`$ (4) $`F(𝐫_i,𝐫_j,i\omega _n)`$ $`=`$ $`{\displaystyle _0^\beta }𝑑\tau \mathrm{exp}(i\omega _n\tau )\mathrm{T}_\tau \widehat{c}_j(\tau )\widehat{c}_i(0),`$ (5) where $`\mathrm{T}_\tau `$ denotes time-ordering in $`\tau `$ and $`\beta =1/(k_BT)`$. The self energies and Green’s functions are coupled together through Dyson’s equation, $$\underset{¯}{\underset{¯}{G}}(𝐫_i,𝐫_j,i\omega _n)=\underset{¯}{\underset{¯}{G}}^{(0)}(𝐫_i,𝐫_j,i\omega _n)+\underset{l}{}\underset{¯}{\underset{¯}{G}}^{(0)}(𝐫_i,𝐫_l,i\omega _n)\underset{¯}{\underset{¯}{\mathrm{\Sigma }}}(𝐫_l,i\omega _n)\underset{¯}{\underset{¯}{G}}(𝐫_l,𝐫_j,i\omega _n),$$ (6) where we have included the local approximation for the self energy, $`\underset{¯}{\underset{¯}{\mathrm{\Sigma }}}(𝐫_i,𝐫_j,i\omega _n)=\underset{¯}{\underset{¯}{\mathrm{\Sigma }}}(𝐫_i,i\omega _n)\delta _{ij}`$, \[which can be relaxed if we use the Dynamical Cluster Approximation (DCA) \]. The non-interacting Green’s function, $`\underset{¯}{\underset{¯}{G}}^{(0)}(𝐫_i,𝐫_j,i\omega _n)`$ is diagonal in Nambu space, with upper diagonal component given by: $$G^{(0)}(𝐫_i,𝐫_j,i\omega _n)=d^3𝐤\frac{\text{e}^{i𝐤(𝐫_i𝐫_j)}}{i\omega _n\epsilon _𝐤+\mu }.$$ (7) We emphasize that $`\underset{¯}{\underset{¯}{G}}^{(0)}`$ is the non-interacting Green’s function and is not the effective medium of an equivalent atomic problem (see below for athe detailed algorithm used to solve the dynamical mean field theory). A major innovation in our work is to utilize an efficient hybrid real-space—momentum-space method for calculating the Green’s functions from the set of local self-energies. We find this method to be much more powerful in solving systems with spatial variations or inhomogeneity, and it is also faster in bulk systems with current flow than a more conventional $`𝐤`$-space integral technique. Since the stacked planes have translational symmetry within the plane, the systems that we study are inhomogeneous in one direction only. We choose that direction to be labeled the z-axis, which is also the direction of current flow through the Josephson junction. The first stage of our method is to convert the problem from a three-dimensional system to a one-dimensional system following the algorithm of Potthoff and Nolting . We perform a Fourier transform within the planes to determine the planar indexed Green’s functions, $$\underset{¯}{\underset{¯}{G}}_{\alpha \beta }(i\omega _n,k_x,k_y)=\left(\frac{a}{L}\right)^2\underset{x_j,y_j}{}\underset{¯}{\underset{¯}{G}}(𝐫_i,𝐫_j,i\omega _n)\mathrm{exp}\left[k_x(x_jx_i)+k_y(y_jy_i)\right],$$ (8) where $`\alpha `$ and $`\beta `$ denote distinct planes, defined by $`\alpha =z_i/a`$, $`\beta =z_j/a`$ and the summation is over all lattice sites, $`(x_j,y_j)`$, within the $`\beta `$-th plane. The self energy, $`\underset{¯}{\underset{¯}{\mathrm{\Sigma }}}_\alpha (i\omega _n)=\underset{¯}{\underset{¯}{\mathrm{\Sigma }}}(z_i,i\omega _n)=\underset{¯}{\underset{¯}{\mathrm{\Sigma }}}(𝐫_i,i\omega _n)`$, is independent of the planar coordinates $`x_i`$ and $`y_i`$, so that Dyson’s equation \[Eq. (6)\] becomes $$\underset{¯}{\underset{¯}{G}}_{\alpha \beta }(i\omega _n,k_x,k_y)=\underset{¯}{\underset{¯}{G}}_{\alpha \beta }^{(0)}(i\omega _n,k_x,k_y)+\underset{\gamma }{}\underset{¯}{\underset{¯}{G}}_{\alpha \gamma }^{(0)}(i\omega _n,k_x,k_y)\underset{¯}{\underset{¯}{\mathrm{\Sigma }}}_\gamma (i\omega _n)\underset{¯}{\underset{¯}{G}}_{\gamma \beta }(i\omega _n,k_x,k_y),$$ (9) with the summation over all planes, $`\gamma `$. The non-interacting planar Green’s function, is similarly found by the Fourier transform $$\underset{¯}{\underset{¯}{G}}_{\alpha \beta }^{(0)}(i\omega _n,k_x,k_y)=\left(\frac{a}{L}\right)^2\underset{x_j,y_j}{}\underset{¯}{\underset{¯}{G}}^{(0)}(𝐫_i,𝐫_j,i\omega _n)\mathrm{exp}\left[k_x(x_jx_i)+k_y(y_jy_i)\right].$$ (10) The local Green’s function, $`\underset{¯}{\underset{¯}{G}}(𝐫_i,𝐫_i,i\omega _n)`$, is required for calculating the self-consistent potentials, and the Green’s function $`\underset{¯}{\underset{¯}{G}}(𝐫_i,𝐫_j,i\omega _n)`$ between two neighboring sites, $`𝐫_i=(x_i,y_i,z_i)`$ and $`𝐫_j=(x_i,y_i,z_i\pm a)`$ is required for current calculations. These are given by the simple planar momentum integrals: $$\underset{¯}{\underset{¯}{G}}(𝐫_i,𝐫_j,i\omega _n)=\left(\frac{\pi }{a}\right)^2_{\pi /a}^{\pi /a}_{\pi /a}^{\pi /a}\underset{¯}{\underset{¯}{G}}_{\alpha ,\beta }(i\omega _n,k_x,k_y)𝑑k_x𝑑k_y$$ (11) where again, $`\alpha =z_i/a`$ and $`\beta =z_j/a`$, and the phase factors in the integral have canceled as $`x_i=x_j`$ and $`y_i=y_j`$. Hence our goal is to find the interacting Green’s functions, $`\underset{¯}{\underset{¯}{G}}_{\alpha \beta }(i\omega _n,k_x,k_y)`$. We make a huge improvement (by one to two orders of magnitude) in the computational efficiency by transforming the two-dimensional planar momentum integral into a single integral over in-plane kinetic energy. In the case of nearest-neighbor hopping on a square lattice, the kinetic energy within the $`\alpha `$-th plane is given by $$\epsilon _\alpha ^{xy}=2t_\alpha \left[\mathrm{cos}(k_xa)+\mathrm{cos}(k_ya)\right]=\frac{t_\alpha }{t}\overline{\epsilon ^{xy}}$$ (12) where $`t_\alpha `$ is the hopping integral between two nearest-neighbor sites within the $`\alpha `$-th plane and $`a`$ is the lattice spacing. The effect of the in-plane kinetic energy is equivalent to an increase in the on-site energy, $`ϵ_i^{(0)}ϵ_i^{(0)}+\epsilon _\alpha ^{(xy)}`$ which can vary between the different planes. The planar Green’s functions only depend on the planar momentum via the normalized kinetic energy, $`\overline{\epsilon ^{xy}}=2t\left[\mathrm{cos}(k_xa)+\mathrm{cos}(k_ya)\right]`$, such that $`\underset{¯}{\underset{¯}{G}}_{\alpha \beta }(i\omega _n,k_x,k_y)=\underset{¯}{\underset{¯}{G}}_{\alpha \beta }(i\omega _n,\overline{\epsilon ^{xy}})`$. Hence, by using the two-dimensional density of states, $`\rho ^{2D}(\epsilon )`$, for a square lattice, the momentum integral is transformed into $$\left(\frac{a}{2\pi }\right)^2_{\pi /a}^{\pi /a}_{\pi /a}^{\pi /a}G_{\alpha \beta }(i\omega _n,k_x,k_y)𝑑k_x𝑑k_y=_{\mathrm{}}^{\mathrm{}}G_{\alpha \beta }(i\omega _n,\overline{\epsilon ^{xy}})\rho ^{2D}(\overline{\epsilon ^{xy}})𝑑\overline{\epsilon ^{xy}}.$$ (13) The specific hopping integral, $`t_\alpha `$, for each plane affects the on-site potential of a given plane in the continued fraction method \[see Eq. (15) below\], but does not contribute to the change of variables in the momentum integral. Once the system is converted to a one-dimensional model, with nearest-neighbor hopping, the Green’s functions can be solved rapidly by a continued fraction expansion, without recourse to another $`k`$-space integral for the $`z`$-direction. The continued fraction expansion is similar to the recursion method, modified to include superconductivity , but with three important differences. First, the method is much faster, as there is no need to expand about a site to obtain a new basis — the system is already one-dimensional in form. Second, there is no inaccuracy in the termination process, as the hopping integrals are given exactly in the model. The third point is an alteration, because the sites of interest are not at the end of a chain, but in the middle. This leads to a different set of continued fractions that must be calculated compared with the standard recursion method. In test runs, our method proves to be $`4\times 10^5`$ times faster, and with machine precision accuracy, compared to a standard recursion method expansion which is terminated (due to memory limits) at an accuracy of one part in $`10^3`$! An alternate approach is to solve Dyson’s matrix equation directly for a finite system, where the infinitely extended bulk boundaries can be mimicked by appropriate choice of potentials for the end planes. We have carried out such an approach as a comparison, but find it to be much slower (by a factor of 4000 for 60 planes than our method, and it grows like $`N^3`$ for large systems with $`N`$ planes) so we only describe the continued-fraction method below. The equivalence of our method to the recursion method is that we calculate the Green’s functions directly from a continued-fraction representation of the inverse of the Hamiltonian matrix in real space. That is, we find $$\underset{¯}{\underset{¯}{G}}_{\alpha \beta }(i\omega _n,\overline{\epsilon ^{xy}})=\left(\begin{array}{ccccccc}\mathrm{}& \mathrm{}& 0& \mathrm{}& \mathrm{}& \mathrm{}& \\ \mathrm{}& i\omega _n\underset{¯}{\underset{¯}{1}}\underset{¯}{\underset{¯}{a}}_{\alpha 2}& \underset{¯}{\underset{¯}{b}}_{\alpha 1}& 0& 0& 0& \mathrm{}\\ \mathrm{}& \underset{¯}{\underset{¯}{b}}_{\alpha 1}^{}& i\omega _n\underset{¯}{\underset{¯}{1}}\underset{¯}{\underset{¯}{a}}_{\alpha 1}& \underset{¯}{\underset{¯}{b}}_\alpha & 0& 0& \mathrm{}\\ \mathrm{}& 0& \underset{¯}{\underset{¯}{b}}_\alpha ^{}& i\omega _n\underset{¯}{\underset{¯}{1}}\underset{¯}{\underset{¯}{a}}_\alpha & \underset{¯}{\underset{¯}{b}}_{n+1}& 0& \mathrm{}\\ \mathrm{}& 0& 0& \underset{¯}{\underset{¯}{b}}_{\alpha +1}^{}& i\omega _n\underset{¯}{\underset{¯}{1}}\underset{¯}{\underset{¯}{a}}_{\alpha +1}& \underset{¯}{\underset{¯}{b}}_{\alpha +2}& 0\\ & \mathrm{}& \mathrm{}& 0& \mathrm{}& \text{ }\mathrm{}\text{ }& \text{ }\mathrm{}\text{ }\end{array}\right)_{\alpha \beta }^{\text{-1}}$$ (14) where the matrices $`\left\{\underset{¯}{\underset{¯}{a}}_\alpha \right\}`$ are the total in-plane energies for a particular plane, given by $$\underset{¯}{\underset{¯}{a}}_\alpha =\left(\begin{array}{cc}ϵ_\alpha +\epsilon _\alpha ^{(xy)}+\mathrm{\Sigma }_\alpha (i\omega _n)\mu & \varphi _\alpha (i\omega _n)\\ \varphi _\alpha ^{}(i\omega _n)& ϵ_\alpha \epsilon _\alpha ^{(xy)}\mathrm{\Sigma }_\alpha ^{}(i\omega _n)+\mu \end{array}\right)$$ (15) and $`\left\{\underset{¯}{\underset{¯}{b}}_\alpha \right\}`$ couple the $`\alpha 1`$th and $`\alpha `$th planes, $$\underset{¯}{\underset{¯}{b}}_\alpha =\left(\begin{array}{cc}t_{\alpha 1,\alpha }& 0\\ 0& t_{\alpha 1,\alpha }^{}\end{array}\right).$$ (16) The local planar Green’s function, $`\underset{¯}{\underset{¯}{G}}_{\alpha \alpha }(i\omega _n,\overline{\epsilon ^{xy}})`$, is readily evaluated as a combination of continued fractions (as in the renormalized perturbation expansion ). We define the right-directed, $`\underset{¯}{\underset{¯}{R}}_\alpha (i\omega _n)`$, and left-directed $`\underset{¯}{\underset{¯}{L}}_\alpha (i\omega _n)`$ continued fractions from a plane, $`\alpha `$, recursively as $`\underset{¯}{\underset{¯}{R}}_\alpha (i\omega _n,\overline{\epsilon ^{xy}})`$ $`=`$ $`i\omega _n\underset{¯}{\underset{¯}{1}}\underset{¯}{\underset{¯}{a}}_\alpha \underset{¯}{\underset{¯}{b}}_{\alpha +1}\underset{¯}{\underset{¯}{R}}_{\alpha +1}^1(i\omega _n,\overline{\epsilon ^{xy}})\underset{¯}{\underset{¯}{b}}_{\alpha +1}^{}`$ (17) $`\underset{¯}{\underset{¯}{L}}_\alpha (i\omega _n,\overline{\epsilon ^{xy}})`$ $`=`$ $`i\omega _n\underset{¯}{\underset{¯}{1}}\underset{¯}{\underset{¯}{a}}_\alpha \underset{¯}{\underset{¯}{b}}_\alpha ^{}\underset{¯}{\underset{¯}{L}}_{\alpha 1}^1(i\omega _n,\overline{\epsilon ^{xy}})\underset{¯}{\underset{¯}{b}}_\alpha `$ (18) The recursive calculation continues to infinity, but once it has been extended to planes in the uniform bulk medium, where $`\alpha <1N/2`$ or $`\alpha >N/2`$, the coefficients at each level become constant. The effect of a constant phase gradient in $`\varphi `$ is equivalent to a constant phase factor in the hopping integral, $`t_{\alpha ,\alpha +1}`$, that does not change between planes in the bulk. Hence, by equating all $`\underset{¯}{\underset{¯}{R}}_\alpha (i\omega _n,\overline{\epsilon ^{xy}})`$ as $`\underset{¯}{\underset{¯}{R}}_{\mathrm{}}(i\omega _n,\overline{\epsilon ^{xy}})`$ for $`\alpha >N/2`$ and $`\underset{¯}{\underset{¯}{L}}_\alpha (i\omega _n,\overline{\epsilon ^{xy}})`$ as $`\underset{¯}{\underset{¯}{L}}_{\mathrm{}}(i\omega _n,\overline{\epsilon ^{xy}})`$ for $`\alpha <1N/2`$ in the bulk limit, an exact terminator function can be calculated as the solution of a complex quadratic matrix equation: $`\underset{¯}{\underset{¯}{R}}_{\mathrm{}}(i\omega _n,\overline{\epsilon ^{xy}})\underset{¯}{\underset{¯}{b}}_{\mathrm{}}^1\underset{¯}{\underset{¯}{R}}_{\mathrm{}}(i\omega _n,\overline{\epsilon ^{xy}})+\left[\underset{¯}{\underset{¯}{a}}_{\mathrm{}}i\omega _n\underset{¯}{\underset{¯}{1}}\right]\underset{¯}{\underset{¯}{b}}_{\mathrm{}}^1\underset{¯}{\underset{¯}{R}}_{\mathrm{}}(i\omega _n,\overline{\epsilon ^{xy}})+\underset{¯}{\underset{¯}{b}}_{\mathrm{}}`$ $`=`$ $`\underset{¯}{\underset{¯}{0}}`$ (19) $`\underset{¯}{\underset{¯}{L}}_{\mathrm{}}(i\omega _n,\overline{\epsilon ^{xy}})\underset{¯}{\underset{¯}{b}}_{\mathrm{}}^1\underset{¯}{\underset{¯}{L}}_{\mathrm{}}(i\omega _n,\overline{\epsilon ^{xy}})+\left[\underset{¯}{\underset{¯}{a}}_{\mathrm{}}i\omega _n\underset{¯}{\underset{¯}{1}}\right]\underset{¯}{\underset{¯}{b}}_{\mathrm{}}^1\underset{¯}{\underset{¯}{L}}_{\mathrm{}}(i\omega _n,\overline{\epsilon ^{xy}})+\underset{¯}{\underset{¯}{b}}_{\mathrm{}}^{}`$ $`=`$ $`\underset{¯}{\underset{¯}{0}}.`$ (20) Note that the same terminator function is used for all sites in the intermediate layers, and the functions $`\underset{¯}{\underset{¯}{R}}_\alpha `$ and $`\underset{¯}{\underset{¯}{L}}_\alpha `$ calculated for one site are used in the calculation for the next site — so the number of computations required to find solutions for all sites can be less than $`O(N)`$. There are two ways to solve this matrix quadratic equation. When we perform calculations on the real axis, without including a supercurrent, the matrix equation becomes analytically tractable to solve. On the imaginary axis, we find that it is numerically faster to simply find an iterative solution to these quadratic equations within the bulk medium. In most cases, accuracies of one part in $`10^{10}`$ can be achieved in ten iterations or less. The continued fractions form the local planar Green’s functions, according to $$\underset{¯}{\underset{¯}{G}}_{\alpha \alpha }(i\omega _n,\overline{\epsilon ^{xy}})=\left\{i\omega _n\underset{¯}{\underset{¯}{1}}\underset{¯}{\underset{¯}{a}}_\alpha (\overline{\epsilon ^{xy}})\underset{¯}{\underset{¯}{b}}_\alpha ^{}\underset{¯}{\underset{¯}{L}}_{\alpha 1}^1(i\omega _n,\overline{\epsilon ^{xy}})\underset{¯}{\underset{¯}{b}}_\alpha \underset{¯}{\underset{¯}{b}}_{\alpha +1}\underset{¯}{\underset{¯}{R}}_{\alpha +1}^1(i\omega _n,\overline{\epsilon ^{xy}})\underset{¯}{\underset{¯}{b}}_{\alpha +1}^{}\right\}^1$$ (21) which, using Eqs.(17-18), can be simplified to $$\underset{¯}{\underset{¯}{G}}_{\alpha \alpha }(i\omega _n,\overline{\epsilon ^{xy}})=\left[\underset{¯}{\underset{¯}{R}}_\alpha +\underset{¯}{\underset{¯}{L}}_\alpha i\omega _n\underset{¯}{\underset{¯}{1}}+\underset{¯}{\underset{¯}{a}}_\alpha \right]^1.$$ (22) The Green’s functions connecting neighboring planes, $`\alpha `$ and $`\alpha \pm 1`$, which are required to calculate the current flow, are given in two equivalent forms $`\underset{¯}{\underset{¯}{G}}_{\alpha ,\alpha +1}(i\omega _n,\overline{\epsilon ^{xy}})`$ $`=`$ $`\underset{¯}{\underset{¯}{G}}_{\alpha \alpha }(i\omega _n,\overline{\epsilon ^{xy}})\underset{¯}{\underset{¯}{b}}_{\alpha +1}\underset{¯}{\underset{¯}{R}}_{\alpha +1}^1(i\omega _n,\overline{\epsilon ^{xy}})=\underset{¯}{\underset{¯}{L}}_\alpha ^1(i\omega _n,\overline{\epsilon ^{xy}})\underset{¯}{\underset{¯}{b}}_{\alpha +1}\underset{¯}{\underset{¯}{G}}_{\alpha +1,\alpha +1}(i\omega _n,\overline{\epsilon ^{xy}})`$ (23) $`\underset{¯}{\underset{¯}{G}}_{\alpha ,\alpha 1}(i\omega _n,\overline{\epsilon ^{xy}})`$ $`=`$ $`\underset{¯}{\underset{¯}{G}}_{\alpha \alpha }(i\omega _n,\overline{\epsilon ^{xy}})\underset{¯}{\underset{¯}{b}}_\alpha ^{}\underset{¯}{\underset{¯}{L}}_{\alpha 1}^1(i\omega _n,\overline{\epsilon ^{xy}})=\underset{¯}{\underset{¯}{R}}_{\alpha +1}^1(i\omega _n,\overline{\epsilon ^{xy}})\underset{¯}{\underset{¯}{b}}_\alpha ^{}\underset{¯}{\underset{¯}{G}}_{\alpha 1,\alpha 1}(i\omega _n,\overline{\epsilon ^{xy}})`$ (24) The local planar Green’s functions enable us to calculate the self-energy and electron density, $`n_i`$, on each site in a given plane, $`\alpha =z_i/a`$, with the latter given by $$n_i=k_BT\underset{\omega _n}{}_{\mathrm{}}^{\mathrm{}}\rho ^{2D}(\overline{\epsilon ^{xy}})\text{Im}\left[G_{\alpha \alpha }(i\omega _n,\overline{\epsilon ^{xy}})\right]𝑑\overline{\epsilon ^{xy}}.$$ (25) The current, $`I_{\alpha ,\alpha +1}`$, which flows along each link between two neighboring planes, $`\alpha `$ and $`\alpha +1`$, in the $`z`$-direction is given by: $$I_{\alpha ,\alpha +1}=\frac{2te}{\mathrm{}}k_BT\underset{\omega _n}{}_{\mathrm{}}^{\mathrm{}}\rho ^{2D}(\overline{\epsilon ^{xy}})\text{Im}\left[G_{\alpha ,\alpha +1}(i\omega _n,\overline{\epsilon ^{xy}})\right]𝑑\overline{\epsilon ^{xy}}.$$ (26) A stringent convergence check for self-consistency, when there is a phase difference between the bulk superconductors, is that the current flow is constant from one plane to the next. For the bulk boundary regions, the uniform variation of phase in the off-diagonal self energy, $`\varphi (𝐫_i)`$ has the form, $`\varphi (𝐫_i)=\varphi _0\mathrm{exp}\left[i𝐪𝐫_i\right]`$, where the net superfluid momentum depends on $`𝐪=(0,0,q_z)`$ through $`mv_s=\frac{\mathrm{}}{a}\mathrm{sin}(q_za)`$ and leads to the following solution of Dyson’s equation: $$G(𝐫_i,𝐫_j,i\omega _n)=d^3𝐤\frac{\left(i\omega _n+\epsilon _{𝐤𝐪}\mu +\mathrm{\Sigma }^{}(i\omega _n)\right)\text{e}^{i𝐤(𝐫_i𝐫_j)}}{\left(i\omega _n\epsilon _𝐤+\mu \mathrm{\Sigma }(i\omega _n)\right)\left(i\omega _n+\epsilon _{𝐤𝐪}\mu +\mathrm{\Sigma }^{}(i\omega _n)\right)\left|\varphi _0(i\omega _n)\right|^2}$$ (27) and $$F(𝐫_i,𝐫_j,i\omega _n)=d^3𝐤\frac{\varphi _0(i\omega _n)\text{e}^{i(𝐤𝐪)(𝐫_i𝐫_j)}}{\left(i\omega _n\epsilon _𝐤+\mu \mathrm{\Sigma }(i\omega _n)\right)\left(i\omega _n+\epsilon _{𝐤𝐪}\mu +\mathrm{\Sigma }^{}(i\omega _n)\right)\left|\varphi _0(i\omega _n)\right|^2}$$ (28) where the diagonal self energy, $`\mathrm{\Sigma }(i\omega _n)`$, is independent of site index, $`𝐫_i`$, in the bulk. We will need only the local, and nearest-neighbor bulk Green’s functions in our calculations. ### C Local self energy calculations In this contribution, the superconducting region is modeled by the negative-U Hubbard model within the Hartree-Fock (static mean-field) approximation. In this case, the local self energy is found from the local Green’s functions by: $$\mathrm{\Sigma }(𝐫_i,i\omega _n)=UT\underset{\omega _n}{}G(𝐫_i,𝐫_i,i\omega _n)$$ (29) and $$\varphi (𝐫_i,i\omega _n)=UT\underset{\omega _n}{}F(𝐫_i,𝐫_i,i\omega _n),$$ (30) where the instantaneous electron-electron interaction energy, $`U`$, leads to a time-independent self energy. This procedure is identical to the conventional Bogoliubov-de Gennes approach , which neglects retardation effects in the superconductor. As all sites within a plane are identical, the self energy need only be calculated once for each of the $`N`$ planes. For planes within the barrier which include impurities, the dynamical mean field approximation says that the local (site) Green’s function, $`\underset{¯}{\underset{¯}{G}}(𝐫_i,𝐫_i,i\omega _n)`$ is related to a local host Green’s function , $`\underset{¯}{\underset{¯}{𝒢}}(𝐫_i,i\omega _n)`$, via $$\underset{¯}{\underset{¯}{𝒢}}(𝐫_i,i\omega _n)=\left[\underset{¯}{\underset{¯}{G}}^1(𝐫_i,𝐫_i,i\omega _n)+\underset{¯}{\underset{¯}{\mathrm{\Sigma }}}(𝐫_i,i\omega _n)\right]^1.$$ (31) The atomic Green’s function, $`\underset{¯}{\underset{¯}{G^{at}}}(𝐫_i,i\omega _n)`$, which will be equated to the local Green’s function, $`\underset{¯}{\underset{¯}{G}}(𝐫_i,𝐫_i,i\omega _n)`$, in the dynamical mean field approximation, then satisfies $$\underset{¯}{\underset{¯}{G^{at}}}(𝐫_i,i\omega _n)=(1\rho _{imp})\underset{¯}{\underset{¯}{𝒢}}(𝐫_i,i\omega _n)+\rho _{imp}\left[\underset{¯}{\underset{¯}{𝒢}}^1(𝐫_i,i\omega _n)U^{FK}\underset{¯}{\underset{¯}{1}}\right]^1.$$ (32) and the local self energy becomes $$\underset{¯}{\underset{¯}{\mathrm{\Sigma }}}(𝐫_i,i\omega _n)=\underset{¯}{\underset{¯}{𝒢}}^1(𝐫_i,i\omega _n)\underset{¯}{\underset{¯}{G^{at}}}^1(𝐫_i,i\omega _n).$$ (33) Starting from a local self energy, $`\underset{¯}{\underset{¯}{\mathrm{\Sigma }}}(𝐫_i,i\omega _n)`$, and a local Green’s function, $`\underset{¯}{\underset{¯}{G}}(𝐫_i,𝐫_i,i\omega _n)`$, Eqs.(31,32,33) can be employed to iteratively determine a new self energy $`\underset{¯}{\underset{¯}{\mathrm{\Sigma }}}(𝐫_i,i\omega _n)`$ when the plane is described by the Falicov-Kimball model. The method, which is solved for a fixed concentration of impurities, $`\rho _{imp}`$, is equivalent to the coherent potential approximation. The algorithm is summarized in Fig. 2. ## III Phase Variation Standard theory of Josephson junctions predicts the phase variation of the current in the weak-coupling limit to be $`I(\theta )=I_c\mathrm{sin}(\theta )`$ where $`\theta `$ is the phase difference across the barrier, and $`I_c`$ is the temperature-dependent critical current. Such a current variation arises from consideration of two superconductors with different phases being coupled by a single energy-independent transmission coefficient, corresponding to the tunneling of Cooper pairs across a barrier. A more general consideration includes the Bogoliubov-de Gennes equations for a one-dimensional system of two superconductors coupled across a potential barrier, $`V`$. The strength of the barrier is measured by $`Z=mV/\mathrm{}^2k_F`$ which determines the transmission coefficient, $`\tau `$, according to $`\tau =1/(1+Z^2)`$. It is found that Andreev bound states carry the current, whose variation with phase difference across the barrier leads to $$I=G_N\frac{\pi \mathrm{\Delta }_0}{2e}\frac{\mathrm{sin}(\theta )}{\sqrt{1\tau \mathrm{sin}^2(\theta /2)}}\mathrm{tanh}\left(\frac{\mathrm{\Delta }_0\mathrm{cos}(\theta /2)}{2k_BT}\right).$$ (34) This reproduces the weak-coupling result in the limit of low transparency, $`\tau 1`$, with $$I=G_N\frac{\pi \mathrm{\Delta }_0}{2e}\mathrm{sin}(\theta )\mathrm{tanh}\left(\frac{\mathrm{\Delta }_0}{2k_BT}\right),$$ (35) and reproduces the formula for a point contact Josephson junction with perfect transparency , $`\tau =1`$: $$I=G_N\frac{\pi \mathrm{\Delta }_0}{e}\mathrm{sin}(\theta /2)\mathrm{tanh}\left(\frac{\mathrm{\Delta }_0\mathrm{cos}(\theta /2)}{2k_BT}\right).$$ (36) Interestingly, the maximum current is at a phase difference which approaches $`\theta =\pi `$ at low temperatures in the latter case, whereas traditional tunnel junctions have their maximum at $`\theta =\pi /2`$. In the more general case of a three-dimensional geometry , the transmissivity depends on the angle of approach of the electron. The full current is obtained by an angular integral, $`d\alpha `$ over the Fermi surface. The complete calculation leads to: $$I\left(\theta \right)=G_N\frac{\pi \mathrm{\Delta }_0(T)}{2e}\overline{R}_N\mathrm{sin}(\theta )_{\pi /2}^{\pi /2}\frac{\sigma _N(\alpha )\mathrm{cos}(\alpha )}{\sqrt{1\sigma _N(\alpha )\mathrm{sin}^2(\theta /2)}}\mathrm{tanh}\left(\frac{\mathrm{\Delta }_0(T)\sqrt{1\sigma _N(\alpha )\mathrm{sin}^2(\theta /2)}}{2k_BT}\right)𝑑\alpha $$ (37) where $$\overline{R}_N^1=_{\pi /2}^{\pi /2}\sigma _N(\alpha )\mathrm{cos}(\alpha )𝑑\alpha .$$ (38) The transmission probability for a single electron, which is proportional to its contribution to the conductivity, $`\sigma _N`$, (from a Landauer formula) depends on its angle of incidence, as well as the strength of scattering at the barrier. The linear response of current due to a small phase difference, $`I^{}`$, is given by: $$I^{}=\left(\frac{I}{\theta }\right)_{lim\theta 0}=G_N\frac{\pi \mathrm{\Delta }_0(T)}{2e}\mathrm{tanh}\left(\frac{\mathrm{\Delta }_0(T)}{2k_BT}\right)$$ (39) for both the one-dimensional and three-dimensional cases above. So if the barrier is parameterized by a single scattering potential, then the linear response supercurrent only depends on the normal state conductance according to Eq. (39) and the product of $`I^{}R_N`$ is a constant, independent of the microscopic details of the barrier. More realistic treatments of the barrier region are expected to result in deviations from Eq. (39) and allow $`I^{}R_N`$ to vary with the properties of the barrier. Further differences arise between methods which treat the barrier as a single scattering potential and our self-consistent treatment of the order parameter and off-diagonal Green’s functions within the barrier. Our results indicate that the effective scattering strength of the barrier, $`Z`$, changes with temperature and current flow. In general, the proximity effect, which enhances coupling between superconductors, is weakened as the current flow approaches the critical current. Hence at large current flow, the effective barrier is increased compared to its value in the linear response regime at close to zero current flow. The effect of self-consistency on the current-phase relationship appears to be more marked for weak barriers, where the current flow becomes relatively large. We quantify the current response by making use of the linear response $`I^{}=\left(dI/d\theta \right)_{lim\theta 0}`$, as well as the maximum current flow, $`I_c`$, through the barrier. In the weak-coupling limit $`I_c`$ is exactly equal to $`I^{}`$, and our results show that in general the two are within 20% of each other, and scale almost identically with external parameters. As $`I^{}`$ requires much less computational time to calculate, than $`I_c`$, we report values of $`I^{}`$ for many of our results. ## IV Bogoliubov-de Gennes Results To begin, we demonstrate how our model reproduces standard results, by using the Hartree-Fock approximation to calculate the self energy within the barrier region. As such, we are effectively self-consistently solving the full Bogoliubov-de Gennes equations for the system. In all the results that we present in this paper, the superconducting region is modeled with an attractive Hubbard interaction of $`U=2`$ at half filling. The homogeneous bulk superconductor then has a critical temperature, $`T_c=0.11`$ and a zero temperature order parameter, $`\mathrm{\Delta }_0=0.198`$. ### A Barrier thickness We begin by solving for systems with a small attractive interaction within the barrier, $`U_b=0.5`$, over a range of barrier thicknesses. This attraction is small enough that the bulk superconducting transition temperature of the barrier material is always less than any temperature we consider. Fig. 3 demonstrates the proximity effect, with the decay of the anomalous average at the center of the barrier as its width is increased. Fig. 4 is a log plot of the linear-response current and critical current against barrier thickness. As expected, both the linear-response current, $`I^{}`$, and the critical current, $`I_c`$, drop rapidly when the number of planes within the barrier region is increased from 5 to 30. The nearly constant slopes indicate that the decays are close to exponential. The exponent is approximately twice the bulk correlation length, given by $`2\xi \mathrm{}v_f/\mathrm{\Delta }2ta/\mathrm{\Delta }10a`$, where $`a`$ is the lattice spacing. Fig. 5 shows the normalized current as a function of phase difference across the barrier region, for two different barrier thicknesses. The phase variation is compared to the simple form $`I(\theta )/I_c=\mathrm{sin}(\theta )`$, which is appropriate in the weak-coupling limit, and $`I(\theta )/I_c=(I_0/I_c)\mathrm{sin}(\theta /2)\mathrm{tanh}(\mathrm{\Delta }\mathrm{cos}(\theta /2)/2k_BT)`$, the result for a point-contact junction, appropriate in the limit of high transmissivity. The curve for a thin junction falls outside these two limits, indicating that the effects of self-consistency and finite junction width are important in determining the current-phase relation for a Josephson junction. In particular, the calculated difference between a weakly coupled and strongly coupled junction lies in the opposite direction to the analytic formula on such a normalized curve. The actual magnitude of the critical current, is of course, greatly enhanced for the thinner barrier, as shown in Figs. 3-4. ### B Barrier interaction strength Fig. 6 demonstrates the proximity effect as a decay in the anomalous average, $`F_{ii}=F(𝐫_i,𝐫_i,\tau =0^+)`$, within a barrier region of twenty planes. The proximity effect is seen to depend on the Hubbard interaction, $`U_b`$, for sites within the barrier. The corresponding critical currents, $`I_c`$ and linear response currents, $`I^{}`$, are shown in Fig. 7. Note that in the cases where $`U_b`$ is negative, the barrier region is actually a superconductor in its normal state, above its transition temperature. In the case where $`U_b=0`$, the order parameter, $`\mathrm{\Delta }_i`$, is exactly zero within the barrier. In the example where $`U_b`$ is positive, the order parameter actually switches sign within the barrier region, even when there is no external phase variation. The results of Fig. 7 indicate that such a switching in sign of the pairing potential, $`\mathrm{\Delta }_i`$ has no marked effect on the transport properties, which depend on the continuously varying Green’s functions of the system. Once again we find a systematic tracking of $`I^{}`$ and $`I_c`$ with the strength of the Coulomb interaction, $`U_b`$, in the barrier. ### C Barrier hopping integral Our method allows us to consider different hopping integrals in different regions (within planes or between planes). This would be appropriate when the barrier region has a density of states at the Fermi surface, or a Fermi velocity that differs from that found in the normal state of the superconducting regions. When modeling such systems, the hopping integral between successive planes that have differing intraplanar hopping integrals, is taken as the geometric mean of the two planar values ($`t_{\alpha ,\alpha +1}=\sqrt{t_\alpha .t_{\alpha +1}}`$). Fig. 8 indicates that the critical current approximately scales with the hopping integral in the barrier region at low temperature. This effect dominates over any increased scattering at the interfaces due to Fermi velocity mismatch. In fact, when $`t_b`$ is large, the critical current is enhanced by a factor of $`t_b/t`$ over that found in the uniform bulk system ($`t_b=1`$). The proximity effect results in a minimum of the anomalous average, $`F_{ii}=F(𝐫_i,𝐫_i,\tau =0^+)=<c_i(0^+)c_i(0)>`$, at the center of the barrier region. At low temperatures, Fig. 9 indicates that the system with a smaller hopping integral in the barrier, $`t_b=1/2`$, has a larger anomalous average than the uniform system (with $`t_b=1`$), due to its increase in density of states at the Fermi surface. As the temperature is increased, there is a crossover, with the system that has largest hopping integral in the barrier ($`t_b=2`$), having the largest anomalous average just below the critical temperature, $`T_c`$. Such a crossover is due to the differing natural energy scales, $`t_b`$, in the barrier region. For the system with $`t_b=1/2`$, an actual temperature of $`T=0.1`$ corresponds to a temperature of $`0.2`$ in the natural energy units of the barrier, $`t_b`$. When the temperatures are given in units of $`t_b`$, as shown in Fig. 10, a series of approximately parallel curves is seen, ordered according to the differing densities of states at the Fermi surface. The anomalous average, $`F_{ii}`$, plotted on the $`y`$-axis, does not require such scaling, as it is dimensionless. ### D Local density of states It is interesting to observe the density of states, in particular the presence of states within the gap as shown in Fig. 11(a). The two major peaks correspond to Andreev bound states, which carry current, when there is a phase difference across the barrier region. The bound states can be seen to split in two, so that two inner peaks move towards each other, while two outer peaks move to the gap edge as the current flow increases. A careful examination of the curve corresponding to zero current reveals a great deal of structure, due to states trapped within the barrier region. The exact energies of the bound states depend on the width of the barrier, and the states arise in part from normal reflection of quasiparticles with non-zero momenta parallel to the planes. The intraplanar momenta of the quasiparticles may lead to the states being unobservable as current peaks at the corresponding voltages in an I-V curve, where only quasiparticles travelling perpendicular to the planes are measured. In Fig. 12, we examine the bound states in more detail, for a system with ten barrier planes ($`N_b=10`$). It is worthwhile noting that as the energy of the states within the gap approaches the gap edge, so they extend further away from the barrier, into the superconducting region \[which begins after the fifth plane from the center ($`\alpha >5`$)\]. The figure also demonstrates an alternating parity between states. The states with energy closest to zero have even parity, as demonstrated by a maximum amplitude at the central planes ($`\alpha =0,1`$). The next higher energy bound states have a node at the central planes, so exhibit odd parity. We can define a current-carrying local density of states such that the total current, $`I_{\alpha ,\alpha +1}`$, between two successive planes, $`\alpha `$ and $`\alpha +1`$, is given by: $$I_{\alpha ,\alpha +1}=i_{\alpha ,\alpha +1}(E)𝑑E.$$ (40) A plot of the function $`i_{\alpha ,\alpha +1}(E)`$ between the two planes, $`\alpha =0`$ and $`\alpha +1=1`$, at the center of the barrier is shown in Fig. 13. It can be seen that the majority of the current is carried by the Andreev bound states, and that the two peaks that have separated from a single Andreev peak at zero phase difference, carry current in opposite directions. The states at positive energy carry equal and opposite current to the states at negative energy, but their occupation is much lower for low temperatures, $`T<T_c`$. ### E Charge Depletion or Accumulation Regions By including an extra repulsive or attractive potential on the interfacial planes which connect the barrier to the superconductor, we are able to model in a very simple manner, some of the effects of a charge accumulation region or a Schottky Barrier. In our simple model there is a mirror symmetry for a half-filled band : a repulsive potential depletes the electron density in one layer, to reproduce some effects of a Schottky Barrier, while an attractive potential results in a charge accumulation region. The effect on the pairing potential and current density are the same, for equivalent potential strengths at half-filling. That is, if a repulsive interfacial potential, $`V_1`$, results in a reduction in electron density on a specific plane, then the attractive potential, $`V_1`$, at the interface, causes an equal in magnitude increase of electron density at that plane, and results in a system with the same variation in order parameter, and equal current response. Fig. 14(a) indicates the oscillations in the anomalous average resulting from a single layer barrier with $`V_0=\pm 2t`$ and $`V_0=\pm 4t`$. Notice that the proximity effect is reduced by the additional scattering potential, so the order parameter exhibits variation more like a step-function, with increasing barrier strength. We can think of the interfacial scattering as modifying the junction from $`SNS`$ to $`SIS`$ characteristics. Fig. 15 depicts the effect on the electron density of a positive interfacial potential, which is like a Schottky barrier, since the electronic charge is depleted at the interface. The charge density also exhibits Friedel oscillations away from the barrier. The transformation $`n_i2n_i`$ maps results for positive potential barrier to negative potential, $`V_0V_0`$, and will change the system to that of a charge accumulation region. ## V Charge Impurity Scattering We model a barrier region with impurities by using the Falicov-Kimball model, as described in Section II, using Eq. (1) with $`U^{FK}<0`$ and using the self-consistency procedure of Eqs.(31-33), as shown in Fig. 2. We carry out the calculations with an impurity concentration, $`\rho _{imp}`$ that ranges from 0.01 to 0.2. In the limit of $`\rho _{imp}=0`$ there is no scattering in the barrier, and the results correspond to the Hubbard model with $`U_b=0`$. Impurities in the barrier region lead to an imaginary part for the frequency-dependent self energy and, since the lifetime of the quasiparticles at the Fermi surface becomes finite, to a non-Fermi liquid, characteristic of annealed disorder scattering. ### A Current Responses The addition of a small number of impurities is seen in Fig. 16 to decrease the anomalous average in the barrier. The accompanying decrease in both critical current and linear response current, shown in Fig. 17 by the curves labeled $`U_{FK}=2`$, is more severe. Addition of 10% impurities ($`\rho _{imp}=0.10`$) leads to a reduction in both current responses to approximately $`1/3`$ of their initial values, while the anomalous average remains at approximately $`3/4`$ its original amplitude. Anderson’s theorem, which states that non-magnetic impurities do not detract from the superconducting properties of a system only holds for a spatially homogeneous system. A current flow breaks the symmetry, and a Josephson junction is inhomogeneous in one dimension, so the effects we observe do not violate Anderson’s theorem. General considerations of Green’s functions in a homogeneous system show that increasing the imaginary part of the electronic self-energy (hence reducing the quasi-particle lifetime) leads to a reduction in supercurrent for a given phase gradient. Hence it is expected that impurities would have a more deleterious effect on current responses than other superconducting properties, which is in the spirit of Anderson’s theorem. We look at the linear response current, $`I^{}`$, as a function of temperature for a barrier with impurity scattering, in Fig. 18. The critical current remains slightly above zero until the bulk critical temperature, $`T_c`$. As suggested by Eq. (39), we can extract an effective transmissivity, $`\tau =G_N.h/2e^2`$ of the junction from $`I^{}`$ using: $$G_N=I^{}/\left\{\frac{\pi \mathrm{\Delta }_0(T)}{2e}\mathrm{tanh}\left(\frac{\mathrm{\Delta }_0(T)}{2k_BT}\right)\right\}$$ (41) where $`\mathrm{\Delta }_0(T)`$ is the value of the order parameter on the last superconducting plane before the barrier region. Fig. 19 depicts the results, which demonstrate that a real barrier has lower transmissivity with increasing temperature. This can be understood from the variation of the self-consistent order parameter, which is reduced to zero within the barrier as temperature is increased. These results are impossible to predict using more conventional approaches, and only arise from a microscopic model that includes self-consistency. ### B Resistance Calculations We find self-consistent solutions of a system in the normal state, with no current flow, by setting the order parameter to zero on all planes. The self energy of planes outside the barrier contains only a constant real part, as initially we carry out the calculations within the Hartree-Fock approximation. Given the set of local self energies, the Green’s functions coupling any two planes are readily found, for any momentum parallel to the planes. We are interested in the longitudinal components in the z-direction (perpendicular to the uniform planes) of the conductivity matrix. We define the conductivity tensor for our effectively one-dimensional system, from the linear current response $`I_{\alpha ,\alpha +1}`$ across a link between two planes, $`\alpha `$ and $`\alpha +1`$, due to an electric field, $`E_{\beta ,\beta +1}`$ across all links between planes $`\beta `$ and $`\beta +1`$: $$\sigma _{\alpha ,\beta }=\frac{I_{\alpha ,\alpha +1}}{E_{\beta ,\beta +1}}$$ (42) We find the conductivity matrix with frequency component, $`\nu `$, neglecting vertex corrections (which is valid for homogeneous systems in the large dimensional limit) to be: $`\sigma _{\alpha ,\beta }(\nu )`$ $`=`$ $`{\displaystyle \frac{2}{\nu }}\left({\displaystyle \frac{eat}{\mathrm{}}}\right)^2{\displaystyle _{\mathrm{}}^{\mathrm{}}}\rho ^{2D}(\epsilon _{xy})d\epsilon _{xy}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{d\omega }{2\pi }}\{\text{Im}\left[G_{\alpha ,\beta +1}(\omega ,\epsilon _{xy})\right]\text{Im}\left[G_{\beta ,\alpha +1}(\omega +\nu ,\epsilon _{xy})\right]+`$ (45) $`\text{Im}\left[G_{\alpha +1,\beta }(\omega ,\epsilon _{xy})\right]\text{Im}\left[G_{\beta +1,\alpha }(\omega +\nu ,\epsilon _{xy})\right]\text{Im}\left[G_{\alpha ,\beta }(\omega ,\epsilon _{xy})\right]\text{Im}\left[G_{\beta +1,\alpha +1}(\omega +\nu ,\epsilon _{xy})\right]`$ $`\text{Im}\left[G_{\alpha +1,\beta +1}(\omega ,\epsilon _{xy})\right]\text{Im}\left[G_{\beta ,\alpha }(\omega +\nu ,\epsilon _{xy})\right]\left\}\right[f(\omega )f(\omega +\nu )]`$ where $`f(\omega )`$ is the Fermi-Dirac distribution function and $`\rho ^{2D}(\epsilon _{xy})`$ is the two-dimensional tight-binding density of states, used for the sum over momenta parallel to the planes. We are interested in the zero-frequency response, which is found from the appropriate limit of Eq. (45): $`\sigma _{\alpha ,\beta }(0)`$ $`=`$ $`{\displaystyle \frac{1}{k_BT}}\left({\displaystyle \frac{eat}{\mathrm{}}}\right)^2{\displaystyle _{\mathrm{}}^{\mathrm{}}}\rho ^{2D}(\epsilon _{xy})𝑑\epsilon _{xy}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{d\omega }{2\pi }}`$ (47) $`{\displaystyle \frac{\text{Im}\left[G_{\alpha ,\beta +1}(\omega ,\epsilon _{xy})\right]\text{Im}\left[G_{\beta ,\alpha +1}(\omega ,\epsilon _{xy})\right]\text{Im}\left[G_{\alpha ,\beta }(\omega ,\epsilon _{xy})\right]\text{Im}\left[G_{\beta +1,\alpha +1}(\omega ,\epsilon _{xy})\right]}{\mathrm{cosh}^2\left(\omega /(2k_BT)\right)}}`$ When calculating the resistance of the junction, it is important to be aware that for an inhomogeneous 1D system the current flow must be uniform but the electric field is not. The relationship $$I_{\alpha ,\alpha +1}=I_0=\underset{\beta }{}\sigma _{\alpha ,\beta }E_{\beta ,\beta +1}$$ (48) leads to $$E_{\beta ,\beta +1}=\underset{\alpha }{}\left(\sigma ^1\right)_{\beta ,\alpha }I_0,$$ (49) by multiplying on the left by the inverse of the conductivity tensor. The voltage across the junction is the sum of the electric fields across each link, multiplied by the lattice spacing, $`a`$, so we obtain the resistance, $$R_N=\frac{V}{I_0}=a\underset{\alpha ,\beta }{}\left(\sigma ^1\right)_{\beta ,\alpha }$$ (50) given by the sum of components of the inverse conductivity tensor. It is worthwhile pointing out, that where there is no imaginary part to the self energy, so there is no quasiparticle decay, the conductivity tensor consists of constant elements, $`\sigma _{\alpha ,\beta }=\sigma _0`$. In such a region, the electric field required to produce a current flow approaches zero as the inverse of the system size, so the local conductivity becomes infinite. (That is, the sum of elements in one row of the conductivity tensor increases with system size.) However, the voltage drop across the region, given by the product of the electric field and length of perfect lead, remains constant (equal to $`I_0a/\sigma _0`$, so the resistance is non-zero (equal to $`a/\sigma _0`$) while the local resistivity vanishes with large system size. In our calculations, where we neglect any lifetime effects of electrons outside the barrier region (in the Hartree-Fock approximation), there is still a contribution to the resistance of the junction from the ‘perfect’ leads, but the value of the contribution does not depend on the length of the leads, so it can be thought of as a contact resistance. Fig. 20 indicates the variation of junction resistance with impurity concentration in the barrier. The resistance increases linearly with number of scattering sites for the small concentrations calculated, with the slope increasing with the strength of scatterers. The intercept is at a finite resistance, which corresponds to the resistance of an infinitely long, perfectly conducting lead, with conductivity tensor given by $`\sigma _{\alpha ,\beta }=\sigma _01.25`$. The resistance calculated for junctions with impurity scattering within the barrier region, does not change when the number of perfectly conducting planes on either side of the barrier is increased from one to twenty-five. Thereafter, numerical instabilities in the matrix inversion process make the calculations unreliable, but we can be confident that the answer already arrived at is the appropriate one for the infinite system. The product $`I_c.R_N`$ decreases with increasing concentration of impurities in the barrier for the examples shown. That is, the reduction in critical current is greater than the increase in resistance due to impurity scattering. A system which increases the $`I_c.R_N`$ product is found by incorporating an extra (coherent) superconducting region within the barrier . Such an $`SNS^{}NS`$ or $`SIS^{}IS`$ structure has been examined, where an $`SNS`$ structure with 20 barrier planes has its central 6 barrier planes replaced with superconducting material, creating a barrier sandwich of 7 normal, 6 superconducting then 7 more normal planes. Both the critical current and linear response current increase by a factor of greater than two, while the normal state resistance is only reduced by 15% from its value for the SNS junction. ## VI Conclusions We have developed an effective method for calculating the equilibrium properties of Josephson junctions. We are able to examine the microscopic details of self-consistently solved systems through local and non-local Green’s functions. The importance of self-consistency has previously been shown , so we go beyond standard ‘potential barrier’ models of junctions to include the effects of spatial correlation and local fluctuations on the self-consistent potentials. Our results are interpreted in terms of a linear-response current, due to a small phase difference across the junction, as well as the critical current, being the maximum current that a junction can sustain. We find that a self-consistent microscopic determination of the potentials in the system for different phase differences results in current-phase behavior not predictable by standard fixed potential, semiclassical approaches. We study the effects of a number of properties of the barrier material, including its electron-electron interaction potential, its hopping integral and its charge impurity concentration. We also study interfacial scattering potentials, which mimic Schottky barriers and charge accumulation regions, and include barriers with a range of thicknesses. As well as quantifying the change in current response due to such modifications in the barrier, we also depict the alterations in the proximity effect for all cases, and charge variation in the case of an interfacial potential. In future work, we plan to model the charge redistribution at interfaces more realistically, by incorporating a non-local Coulomb potential, which will self-consistently determine an effective potential with the charge density for each plane. By such a model, we expect to discover if there is any significant charge redistribution as a junction passes through its superconducting transition, as suggested by Greene’s group . We have also plotted the local density of states, to observe the bound states which occur at energies less than the bulk gap, within the barrier region. These states can be both current-carrying Andreev states, or states arising from normal reflections. Their evolution with position and phase variation is seen within our model, as is their contribution to the total current flow when a phase difference is applied. We plan in later work to show how the detailed structure of electronic states, apparent in these equilibrium results, will affect the I-V characteristics of a junction. We have carried out resistance calculations for junctions with impurities in the barrier, and found that in general the reduction in critical current outweighs the increase in resistance due to charge impurities. We suggest that a junction of the form $`SNSNS`$, where a thin layer of superconductor is placed within an normal metal barrier, can increase the critical current of a junction dramatically, without markedly reducing its resistance. We are in a position to study how more subtle effects involving electronic correlation close to the metal-insulator can affect the $`I_c.R_N`$ product. We plan to progress by making contact with specific materials, such as a Nb-InAs-Nb junction. Our microscopic formulation will be particularly necessary when we proceed to study junctions created with high-temperature superconductor materials. The d-wave symmetry of the order parameter, with its directional dependence, results in behavior that is not addressed by models utilizing a single transmissivity function. While an appropriate non-local version of the Bogoliubov-de Gennes equations can provide some insight into the properties of d-wave junctions , use of the DCA is necessary to tackle the problem in any realistic manner. ## VII Acknowledgements We are grateful to the Office of Naval Research for funding under grant number N00014-99-1-0328.
warning/0001/hep-ph0001083.html
ar5iv
text
# 1 Introduction ## 1 Introduction The success of coupling unification extrapolations based on the massless spectrum of the MSSM has greatly conditioned the search for a realistic string vacuum. This search has been shaped, to a great extent, by the fact that SM couplings seem to join at a scale $`M_X=2\times 10^{16}`$ GeV, not far from the Planck mass $`M_p`$ and by the necessity of identifying the string scale $`M_s`$ essentially with the Planck scale in heterotic model building. In this situation, it appears natural to look for perturbative heterotic vacua in which, below a scale of order $`M_X`$, essentially only the MSSM remains. Recent p-brane developments have changed our view of the possible ways to embed SM physics into string theory. To start with, it has been realized that the string/M-theory scale may be much below the Planck and unification scales . This is because in the presence of p-branes (like D-branes in Type II and Type I string theory) gauge interactions can be localized in the world-volume of D-branes (e.g., a 3-brane), whereas gravitational interactions in general live in the full ten (or eleven) dimensions. Then the largeness of the Planck mass may be obtained even if $`M_s<<M_p`$ if there are large compact dimensions. Now, if $`M_s<<M_X`$, gauge coupling unification should in principle take place at the string scale $`M_s`$ and thus the nice unification of MSSM couplings is lost. Of course, this unification problem appearing for string models if $`M_s<<M_X`$, could be taken as an argument against them. However, we think that we should first try to answer the following question: Is there any simple alternative framework which is consistent with unification at a string scale $`M_s<<M_X`$ ? After all, the MSSM+big desert orthodoxy is not free of problems. In fact, some unattractive features of the standard scenario are the following: i) The quark/lepton generations come in three chiral copies whereas the Higgs fields come only in one copy and are non-chiral. The fact that there is a single Higgs set is crucial to obtain correct unification predictions. This asymmetry among quarks/leptons on one side and Higgs fields on the other looks quite ad hoc. ii) The MSSM needs to be supplemented by additional symmetries like R-parity in order to ensure proton stability from dimension four operators. It also needs additional symmetries beyond R-parity to get stability against dimension five operators. iii) To obtain a viable heterotic string unification, whose massless sector is just the MSSM and includes the above symmetries, turns out to be a very difficult task, if not impossible. All the models studied up to now require a complicated study of possible scalar flat directions and only very particular ones lead to something of that sort . The reason why dynamics should prefer such vacua with only one set of Higgsses and built-in discrete symmetries to suppress too fast proton decay is unclear. Given the above limitations of the orthodox approach, it seems sensible to look for (if possible, elegant) alternatives. But to be really competitive with the MSSM scenario such alternatives need to 1) improve some of the above problematic aspects of the standard scenario and 2) have a nice and consistent unification of coupling constants. By the latter we mean that couplings unify at the string scale without forcing the structure of the model by adding, for instance, unjustified extra mass scales or ad-hoc extra massless particles. In the present paper we want to propose such an alternative to the standard MSSM+ big desert scenario. We propose that above a scale of order 1 TeV or so, the gauge group is that of the minimal left-right symmetric extension of the supersymmetric standard model, i.e.,$`SU(3)\times SU(2)_L\times SU(2)_R\times U(1)_{BL}`$ . In addition all quarks, leptons and Higgs fields come in three generations. Interestingly enough, this simple structure leads to very precise unification of gauge coupling constants at an intermediate scale of order $`10^{12}`$ GeV as long as the normalization of the $`U(1)_{BL}`$ coupling is the one expected if such gauge group is associated to a collection of D-branes. It is important to remark that this normalization differs from the one predicted by standard GUT left-right symmetric scenarios like $`SO(10)`$. We also construct an explicit Type IIB orientifold string compactification leading to the desired massless spectrum and normalization of coupling constants. In this model there is a simple explanation for the family replication: there are three quark-lepton generations because there are three complex compact dimensions and an underlying $`Z_3`$ orbifold. These features resemble the first three-generation perturbative heterotic string models built, those of ref.. Furthermore, this model has the interesting property of having natural discrete symmetries including $`R`$-parity and $`Z_2`$ lepton numbers. Thus guaranteeing in a natural way the stability of the proton, although allowing for other baryon number violation processes, such as neutron-antineutron oscillations, sufficiantly suppressed to be consistent with current experimental bounds. Finally, the structure of the model allows for candidate axion fields with the right couplings to gauge fields needed to solve the strong CP problem. The structure of this paper is as follows. In chapter 2 we present this alternative scenario which we call D-brane left-right symmetric model. We also discuss the unification of coupling constants and show how, if this model is correct, new $`Z`$’ and $`W`$’ gauge bosons corresponding to left-right symmetry should be found at future or present colliders. In chapter 3 we present a particular Type IIB orientifold model realizing the above scenario and study the cancellation of $`U(1)`$ anomalies and generation of Fayet-Iliopoulos terms. In this realization the unification scale is identified with a string scale $`M_s10^{12}`$ GeV. The model has also some anti-branes in the bulk which provide for hidden-sector supersymmetry breaking at the same scale of order $`M_s`$. We study a number of phenomenological issues of this particular orientifold model in chapter 4. This includes some aspects of the structure of Yukawa couplings, $`SU(2)_R\times U(1)_{BL}`$ symmetry breaking and the presence of natural candidates for invisible axions. In chapter 5 we present an outlook and some final comments. ## 2 The D-brane left-right symmetric model and coupling unification As we discussed above, it has become recently clear that the string scale could well be much below the Planck mass. But, is there any indication or advantage from a lowered string scale? A particularly interesting alternative to the unification at $`M_X`$ close to the Planck scale is getting unification close to the geometric intermediate scale $`M_I=\sqrt{M_WM_p}`$. Indeed, if the string scale is of order $`M_s=M_I`$, gauge couplings should unify at that scale. Now, as argued in ref. , if there are non supersymmetric brane configurations, the scale of supersymmetry breaking would also be of order of $`M_s=M_I`$. This is interesting because hidden sector supersymmetry breaking models also need to have SUSY-breaking at the intermediate scale. Thus in this case the string, unification and SUSY-braking scales would be one and the same. We would like to argue in what follows that the intermediate scale idea is equally good than the standard one in what concerns coupling unification, at least for a model with the following structure: i) The gauge group above a L-R symmetric scale $`M_R`$ slightly above the weak scale is the minimal left-right symmetric extension of the SM: $`SU(3)\times SU(2)_L\times SU(2)_R\times U(1)_{BL}`$. ii) All quarks, leptons and Higgs fields come in three generations. Thus the chiral multiplet content is three copies of $`(3,2,1,1/3)+(\overline{3},1,2,1/3)`$ $`+(1,2,1,1)+(1,1,2,+1)`$ $`+(1,2,2,0)`$. iii) The boundary conditions at the unification (i.e,.string) scale are $`g_3^2=g_L^2=g_R^2=32/3g_{BL}^2`$. This corresponds to a weak angle with $`sin^2\theta (M_s)=3/14=0.215`$. In a model with the above characteristics one finds that gauge couplings naturally unify at a scale of order the intermediate scale $`M_s10^{12}`$ GeV as long as the left-right scale $`M_R`$ is not far from the weak scale $`M_W`$. An important point to remark is that the unification boundary conditions are different from those found in GUT schemes. Indeed in $`SO(10)`$-like schemes the boundary conditions at unification are $`g_3^2=g_L^2=g_R^2=8/3g_{BL}^2`$ yielding the canonical $`sin^2\theta _W=3/8`$. A remarkable point we find is that the new boundary conditions we are proposing are precisely the ones which are natural from the point of view of the embedding of the gauge group in a D-brane scheme. Let us discuss in some more detail how coupling unification takes place. The above mentioned boundary conditions $`g_3^2=g_L^2=g_R^2=32/3g_{BL}^2`$ have a simple group theoretical interpretation. They correspond to the embedding of the left-right symmetric gauge group into a non-semisimple structure: $$U(3)\times U(2)_L\times U(2)_R$$ (2.1) with unified coupling constants $`g_3^2=g_L^2=g_R^2`$ at some mass scale (to be identified later on with the string scale $`M_s`$ ). This is in fact the structure one gets in models with gauge groups living on D-branes, as we will discuss in the specific string model below. As we said, the model contain three identical generations under $`U(3)\times U(2)_L\times U(2)_R`$ with quantum numbers $`(3,\overline{2},1)_{(1,1,0)}+(\overline{3},1,2)_{(1,0,1)}`$ $`+(1,\overline{2},1)_{(0,1,0)}+(1,1,2)_{(0,0,1)}`$ $`+(1,2,\overline{2})_{(0,1,1)}`$, where the subindices denote the charges with respect to the three $`U(1)`$’s. We denote the $`U(1)`$ generators by $`Q_3`$, $`Q_L`$ and $`Q_R`$ respectively. It is easy to check that two of them are anomalous and only one of them, the linear combination $$Q_{BL}=\frac{2}{3}Q_3Q_LQ_R$$ (2.2) is anomaly free <sup>1</sup><sup>1</sup>1In string theory the other two (anomalous) $`U(1)`$’s become massive and decouple due to a generalized Green-Schwarz mechanism. See the discussion in chapter 3. . This is just the familiar $`(BL)`$ of left-right symmetric models which is related to weak hypercharge by $`Y=T_R^3+Q_{BL}/2`$. Now, notice that, if we normalize the original $`U(n)`$ generators in the fundamental representation $`T_a`$ to $`TrT_a^2=1`$, the normalization of the $`U(1)`$’s are $`TrQ_3^2=3`$, $`TrQ_L^2=TrQ_R^2=2`$. Then, the normalization of $`U(1)_{BL}`$ compared to that of the non-Abelian generators is $`k_{BL}=2TrQ_{BL}^2=32/3`$, as remarked above. Notice this implies a hypercharge normalization $`k_1=k_R+1/4k_{BL}=11/3`$, and hence a tree level weak angle $`sin^2\theta _W=3/14=0.214`$. Let us study now the one-loop corrections to the couplings. In between the scales $`M_R`$ and $`M_s`$ the gauge group is $`SU(3)\times SU(2)_L\times SU(2)_R\times U(1)_{BL}`$ and the above chiral field content gives rise to the following one-loop $`\beta `$-function coefficients $`B_a`$ : $$B_3=3;B_L=+3;B_R=+3;B_{BL}=+16$$ (2.3) In between the weak scale $`M_W`$ and $`M_R`$ the gauge group will be that of the SM with $`\beta `$-function coefficients $`b_i`$. Then the one loop running yields: $`\mathrm{sin}^2\theta _W(M_Z)=`$ $`{\displaystyle \frac{1}{1+k_1}}(1+k_1{\displaystyle \frac{\alpha _e(M_Z)}{2\pi }}[(B_L{\displaystyle \frac{1}{k}}_1B_1^{})\mathrm{log}({\displaystyle \frac{M_s}{M_R}})`$ (2.5) $`+(b_2{\displaystyle \frac{1}{k}}_1b_1)\mathrm{log}({\displaystyle \frac{M_R}{M_Z}})]`$ $`{\displaystyle \frac{1}{\alpha _e(M_Z)}}{\displaystyle \frac{1+k_1}{\alpha _3(M_Z)}}=`$ $`{\displaystyle \frac{1}{2\pi }}[(b_1+b_2(1+k_1)b_3)\mathrm{log}({\displaystyle \frac{M_R}{M_Z}})`$ $`+(B_1^{}+B_L(1+k_1)B_3)\mathrm{log}({\displaystyle \frac{M_s}{M_R}})]`$ where one defines $$B_1^{}=B_R+\frac{1}{4}B_{BL}$$ (2.6) and $`k_1=k_R+1/4k_{BL}`$. With the minimal particle content described above one has $`B_1^{}=7`$. Let us suppose for the moment that, below the $`M_R`$ scale down to $`M_Z`$, we were left just with the content of the MSSM. We would then have (for $`k_1=11/3`$, corresponding to $`k_{BL}=32/3`$): $`\mathrm{sin}^2\theta _W(M_Z)=`$ $`{\displaystyle \frac{3}{14}}(1+{\displaystyle \frac{\alpha _e(M_Z)}{2\pi }}[4\mathrm{log}({\displaystyle \frac{M_s}{M_R}}){\displaystyle \frac{22}{3}}\mathrm{log}({\displaystyle \frac{M_R}{M_Z}})])`$ (2.7) $`{\displaystyle \frac{1}{\alpha _e(M_Z)}}{\displaystyle \frac{14}{3\alpha _3(M_Z)}}=`$ $`{\displaystyle \frac{1}{2\pi }}[26\mathrm{log}({\displaystyle \frac{M_R}{M_Z}})+24\mathrm{log}({\displaystyle \frac{M_s}{M_R}})]`$ (2.8) Now, using as input $`\alpha _e(M_Z)^1=127.934\pm 0.027`$ in ref. (in particular its central value), one can plot the predicted $`sin^2\theta _W(M_Z)`$ versus $`\alpha _3(M_Z)`$ and compare it to the experimental data for those two quantities. This is done for several values of $`k_{BL}`$ in fig.1, where the data plotted correspond to the world-average in ref. with two standard deviation errors. It may be observed that, for the particular normalization corresponding to an embedding of the left-right symmetric interactions into D-branes ($`k_{BL}=32/3`$), a very nice agreement with the data is found. Very slight departures from this value are ruled out. The unification scale corresponding to the successful results is of order $`M_s=9\times 10^{11}`$ GeV. We show for comparison a similar plot for the MSSM obtained from the one-loop formulae in fig.2. Again, for the MSSM standard normalization $`k_1=5/3`$ the prediction nicely goes through the data points, in this case for a unification mass of order $`M_X=2\times 10^{16}`$ GeV. Thus we may conclude that, within the approximations made, the D-brane left-right symmetric model is remarkably successful in obtaining appropriate gauge coupling unification. It can also be checked that, within this scheme, both the SUSY-breaking scales and the left-right scale $`M_R`$ cannot be much above the 1 TeV scale. Indeed, let us now use also as input the world average central value for the weak angle, $`sin^2\theta _W(M_Z)=0.23117\pm 0.00016`$ as well as $`\alpha _e(M_Z)`$ and let us plot the predicted $`\alpha _3(M_Z)`$ as a function of the left-right scale $`M_R`$ for various values of the SUSY-breaking mass $`M_{sb}`$. This is shown in figs. 3 and 4, where we have assumed that below the $`M_R`$ (and above the $`M_{sb}`$) scale one is left with the particle content of the MSSM. In fig. 3 we consider a universal SUSY-breaking threshold $`M_{sb}`$ below which the non-SUSY SM is obtained. Fig. 4 plots the same quantities but now with two SUSY-breaking thresholds one ($`M_{sb1}`$) for the coloured SUSY-particles (squarks and gluinos) and a second lower one $`M_{sb2}`$ for non-coloured ones. We observe that for large values of $`M_R`$, a low value for the SUSY threshold seems to be required to get a consistent value for $`\alpha _3(M_Z)`$. Thus either the SUSY threshold or the left-right symmetric threshold (or both) should be below the 1 TeV scale. Although we have only used one-loop formulae and have taken step functions for the different thresholds we expect that those refinements will not substantially change the main conclusion that couplings nicely unify as long as the $`M_R`$ scale is not much above 1 TeV. Several comments are in order: i) Notice that we have not played around with the addition of extra mass scales and /or extra particles beyond the three quark/lepton/Higgs generations in order to get satisfactory unification. So this is not a mere adjustment of the model to get nice coupling unification, it appears naturally as long as $`M_R`$ is not much higher than 1 TeV. ii) In order to obtain the above interesting unification results it is crucial to use the boundary conditions $`g_3^2=g_L^2=g_R^2=32/3g_{BL}^2`$. Thus if we would have used the standard GUT conditions, with $`k_{BL}=8/3`$ instead of $`32/3`$, we would obtain $`M_s10^{13}`$ GeV and $`sin^2\theta _W(M_Z)=0.34`$ (for input $`\alpha _3(M_Z)=0.119`$ and $`M_R1`$ TeV). As we have mentioned, the boundary conditions which work correspond to those expected when the gauge group is embedded inside a collection of Type IIB D-branes. iii) The tree level result $`sin^2\theta _W=3/14=0.214`$ is quite close already to the experimental result 0.231. One loop effects should then be small and positive (i.e., of order 8% ). This contrasts with the standard MSSM/GUT case, where the tree level result $`sin^2\theta _W=3/8=0.375`$ departs from the experimental number. In this case the loop corrections must be large and negative (of order 62% ). In this connection notice that the $`sin^2\theta _W`$ loop corrections have two pieces in our case, one positive and proportional to $`log(M_s/M_R)`$ and one negative and proportional to $`log(M_R/M_Z)`$. Since a positive correction is needed in order to get agreement for $`sin^2\theta _W`$, this is the hidden reason why gauge coupling unification requires in our case large $`log(M_s/M_R)`$ but small $`log(M_R/M_Z)`$. iv) Coupling constant unification works only for three generations of quarks, leptons and Higgs fields. Thus there is a connection between generation number and unification (unlike the MSSM which is insensitive, at one loop, to the number of quark-lepton generations). Let us denote the number of quark/lepton/Higgs generations as $`n_g`$. Thus, $`\beta `$-function coefficients become now $`B_3=9+2n_g`$, $`B_L=B_R=6+3n_g`$, $`B_{BL}=(16n_g)/3`$ and $`B_1^{}=6+(13/3)n_g`$. The combinations relevant for the running are $$[B_L\frac{3}{11}B_1^{}]=\frac{4}{11}(5n_g12);[B_1^{}+B_L\frac{14}{3}B_3]=302n_g$$ (2.9) Now, notice that in order to have a positive correction for $`sin^2\theta _W`$ as required, we need $`n_g3`$. On the other hand, for $`n_g4`$ either $`sin^2\theta _W`$ becomes too large or $`log(M_s/M_R)`$ becomes too small to be compatible with unification. v) We have assumed that at a scale $`M_R`$ the gauge group is broken to that of the SM but we have not specified what the fields, giving rise to such breaking, are. As we will show in the specific string construction below, there are simple additions to the model (e.g., from the presence of some D-branes in the bulk, which lead to non-chiral particle content) which contain the required fields for this breaking without modifying the runnings at one loop, thus preserving the interesting results described above. The above new scheme may probably be obtained in different classes of string models involving D-branes with a gauge group $`U(3)\times U(2)\times U(2)`$. One of the most interesting points we find is that to construct a specific four-dimensional Type I string model with precisely that massless spectrum and with the necessary gauge coupling boundary conditions is very simple. In fact, such a model was briefly discussed in section (4.1) of ref.. We do not claim however that this is the only possible realization of the D-brane left-right symmetric scenario here introduced. Nevertheless, we think it is worth studying such a model, since it may provide clues of more general features of the scheme. We now describe the construction of the mentioned Type IIB orientifold realization. ## 3 A Type IIB orientifold with left-right symmetry ### 3.1 The LR orientifold model The model we are interested in is a $`Z_3`$ Type IIB orientifold with both D-branes and anti -D-branes. A general discussion of such kind of models is given in Ref. where we refer the reader for notation and details of the construction <sup>2</sup><sup>2</sup>2For other constructions involving anti-branes see . Here we only present a brief description in order to settle the general framework. A $`Z_3`$ Type IIB orbifold, in four dimensions , is obtained by dividing closed Type IIB string theory compactified on a six dimensional torus $`T^6`$, by the discrete symmetry group $`Z_3`$. The orientifold model is obtained by further dividing the orbifoldized string by world sheet orientation reversal symmetry . The twist eigenvalues, associated to complex coordinates $`Y_aa=0,1,2`$ are chosen as $`v=\frac{1}{3}(1,1,2)`$ in order to leave $`N=1`$ supersymmetry in four dimensions. The general picture is that the above procedure leads to a Klein-Bottle unoriented world sheet. Amplitudes computed on such a surface contain unphysical tadpole like divergences which can be interpreted as unbalanced charges carried by RR form potentials. Thus, in order to cancel such divergences, D9-branes, carrying opposite charges must be introduced. Moreover, D5-branes and anti-D5-branes can be consistently included. Even if they are not required (in this $`Z_3`$ case) for tadpole cancellation, they open the way for achieving interesting supersymmetry breaking patterns and at the same time provide new possibilities for model building. Let us be more explicit. An open string state is denoted by $`|\mathrm{\Psi },ab\lambda _{ab}^{pq}`$ where $`\mathrm{\Psi }`$ refers to world-sheet degrees of freedom whereas $`a,b`$ are Chan-Paton indices associated to the open string endpoints lying on D$`p`$-branes and D$`q`$-branes respectively . $`\lambda ^{pq}`$ is the Chan-Paton, hermitian matrix, containing the gauge group structure information. Analogously, $`\lambda ^{\overline{p}\overline{q}}`$ ($`\lambda ^{\overline{p}q}`$) is introduced for open strings ending at D$`\overline{p}`$, D$`\overline{q}`$-antibranes ( D$`\overline{p}`$ antibrane, D$`q`$-brane, etc.). The $`Z_3`$ action (denoted by $`\theta `$) that twists the internal complex coordinates has a corresponding action on Chan Paton matrices represented by unitary matrix $`\gamma _{\theta ,p}`$, namely $`\theta :\lambda ^{pq}\gamma _{\theta ,p}\lambda ^{pq}\gamma _{\theta ,q}^1`$. Moreover, Wilson lines, wrapping along internal tori directions can also be included and also have a matrix representation when acting on Chan-Paton factors. Consistency under group algebra operations and the requirement of cancellation of RR tadpoles leads to constraints on the possible twist matrices. Tadpole cancellation, in the $`Z_3`$ case we are discussing, imposes the number of nine branes to be 32 and the requirement that the number of D5-branes and anti-D5-branes must be equal. Moreover, cancellation of twisted tadpoles requires $$\mathrm{Tr}(𝒲)^a\gamma _{\theta ,9}+3(\mathrm{Tr}\gamma _{\theta ,5,a,i}\mathrm{Tr}\gamma _{\theta ,\overline{5},a,i})=4$$ (3.1) for $`a=0,1,2`$. Here we have allowed for the possibility of having a Wilson line, represented by the matrix $`𝒲`$ on Chan -Paton matrices, wrapping along the direction $`e_1`$. We denote with $`a,i`$ with $`a,i=0,1,2`$ the nine orbifold fixed points in the first and second complex planes. For a given $`a`$, $`i=0,1,2`$ label the subset of fixed points that feels the twist $`(𝒲)^a\gamma _{\theta ,9}`$. Generic solutions to these equations are discussed in . Here we consider the specific model characterized by $`\gamma _{\theta ,9}=(\stackrel{~}{\gamma }_{\theta ,9},\stackrel{~}{\gamma }_{\theta ,9}^{}{}_{}{}^{})`$, and $`𝒲=(\stackrel{~}{𝒲},\stackrel{~}{𝒲}^{})`$ where $``$ denotes the complex conjugate and $`\stackrel{~}{\gamma }_{\theta ,9}`$ $`=`$ $`\mathrm{diag}(\alpha I_3,\alpha ^2I_2,I_2,I_2,\alpha I_7)`$ (3.2) $`\stackrel{~}{𝒲}`$ $`=`$ $`\mathrm{diag}(I_3,I_2,I_2,I_2,I_7)`$ (3.3) Also, $$\gamma _{\theta ,5,2,i}=\mathrm{diag}(\alpha ,\alpha ^2)$$ (3.4) where $`\alpha =e^{2i\pi /3}`$. It can be easily checked that such a choice satisfies the tadpole cancellation constraint 3.1. Notice that there are two 5-branes ($`\mathrm{Tr}\gamma _{\theta ,5,2,i}=1`$) stuck at each of the three fixed points of the type $`(2,i)(1,i)`$. The effective twist is thus $`𝒲^1\gamma _{\theta ,9}`$. There are no 5-branes at the other six fixed points ($`\mathrm{Tr}(\gamma _{\theta ,9})=\mathrm{Tr}(𝒲)\gamma _{\theta ,9}=4`$). Since we must have the same number of branes and antibranes, six anti-5-branes must be present in the bulk. It is always possible to add an extra Wilson line in the third complex plane in such a way that anti-brane sector is gauge decoupled from the 9, 5-branes sectors. In this way, anti-branes in the bulk, which lead to a non-supersymmetric spectrum, provide a“hidden sector” which will transmit supersymmetry breaking through gravitational interactions to the “observable” brane sector. An alternative description of such a decoupling situation can be achieved by performing a T-duality transformation in the third complex dimension. Thus, 9-branes become 7-branes located at the origin ($`Y_3=0`$) in the third complex plane with their world volume including the first two complex planes. 5-branes (antibranes) turn into 3-branes (antibranes) which can be located anywhere on the compact space. This is pictorically described in figure 5 <sup>3</sup><sup>3</sup>3Notice that in this figure extra ‘bulk’ branes are added (see below) and only two antibranes are shown, the other antibranes are located at the image of these two under the combination of the $`Z_3`$ defining the orbifold and the orientifold $`\mathrm{\Omega }`$ (or $`(1)^{F_L}\mathrm{\Omega }R_3`$ in the 7,3-brane picture) which is a further $`Z_2`$ symmetry, making for a total of $`2\times 3\times 2=12`$ antibranes which equals the number of 3-branes.. Hence, if bulky anti-3-branes are placed away from $`Y_3=0`$ their worldvolume will not overlap that of the supersymmetric 7,3-brane sectors. Thus bulky, non-supersymmetric fields, will decouple from the supersymmetric sector and will only couple to it through the exchange of closed string interactions. This is the T-dual description of the Wilson line accounted for in the preceding paragraph. We adopt this point of view in what follows. Twist matrices in the T-dual description are obtained just by exchanging $`79`$, $`3(\overline{3})5(\overline{5})`$ indices. Thus, spectrum and interactions are the same in both descriptions. Twist matrices can be described in terms of associated shift vectors (). Such a description, which naturally appears when a Cartan-Weyl basis is chosen for the group algebra, is especially adapted for computing the spectrum . Thus, twist matrices above (now for 7,3-branes) correspond to $`V_7={\displaystyle \frac{1}{3}}(1,1,1;1,1;0,0;0,0;1,1,1,1,1,1,1)`$ (3.5) $`W={\displaystyle \frac{1}{3}}(1,1,1;1,1;1,1;0,0;0,0,0,0,0,0,0)`$ (3.6) and simply $`V_{3,(2,i)}=\frac{1}{3}`$ for matrices in 3.4. We then have for $`(𝒲)\gamma _{\theta ,9}`$ and $`𝒲^1\gamma _{\theta ,9}`$ $`V_7+W={\displaystyle \frac{1}{3}}(2,2,2;0,0;1,1;0,0;1,1,1,1,1,1,1)`$ (3.7) $`V_7W={\displaystyle \frac{1}{3}}(0,0,0;1,1;2,2;0,0;1,1,1,1,1,1,1)`$ (3.8) respectively. The total observable gauge group is $`G=G_7\times G_3`$ with the 77-sector brane group being $`G_7=U(3)\times U(2)_L\times U(2)_R\times SO(4)\times U(7)`$, while the $`\mathrm{𝟑𝟑}_{2,i}`$ sectors, with branes trapped at $`(2,i)`$ ($`i=0,1,2`$) fixed points leads to $`G_3=U(1)_{2,0}\times U(1)_{2,1}\times U(1)_{2,2}`$. The six anti-3-branes in the bulk give rise to a single $`Sp(2)`$: two of them give rise to a $`Sp(2)`$ and the other four are just $`Z_3`$ mirrors of the first ones. Let us analyze the supersymmetric part of the spectrum (see ). We find the chiral content $`\mathrm{𝟕𝟕}sector:`$ (3.9) $`3[(3,2,1,1/3)+(\overline{3},1,2,1/3)+(1,2,2,0)]+3[(1,21)^{}+(4,\overline{7})^{}]`$ where we have indicated the representations under the Left-Right group $`G_{LR}=SU(3)\times SU(2)_L\times SU(2)_R\times U(1)_{BL}`$ and $`G^{}=SO(4)^{}\times U(7)^{}`$. The Abelian factor $`U(1)_{BL}`$ is generated by $`Q_{BL}=\frac{2}{3}Q_3Q_LQ_R`$ introduced in 2.2 where $`Q_3,Q_L,Q_R`$ are the generators of the Abelian factor in the corresponding unitary groups. As mentioned, $`Q_{BL}`$ is identified with $`BL`$ symmetry generator and it can be shown (see next section) to be non anomalous. The Standard Model hypercharge $`Y`$ and electromagnetic charge are thus obtained as $`Y`$ $`=`$ $`{\displaystyle \frac{Q_{BL}}{2}}T_R^3`$ (3.10) $`Q_{em}`$ $`=`$ $`Y+T_L^3`$ (3.11) where $`T_{R,L}^3`$ are the diagonal generators of $`SU(2)_{R,L}`$. We observe that the $`\mathrm{𝟕𝟕}`$ sector contains the standard three quark generations plus a set of three chiral Higgs fields $`(1,2,2,0)`$. The factor three here is associated to the three compact complex dimensions $`Y_a`$ $`(a=0,1,2)`$ each one of them feeling the same orbifold twist. $`\mathrm{𝟑𝟕}_{2,i}sector:`$ (3.12) $`(3,1,1,2/3)_1+(\overline{3},1,1,2/3)_1+(1,2,1,1)_1+(1,1,2,+1)_1+[(1,7)_1^{}+(4,1)_1^{}]`$ A subindex indicates the charge with respect to the $`U(1)_{3,(2,i)}`$ 3-brane group. Since $`i=0,1,2`$, we will have three identical copies. Hence, these sectors provide three generations of standard leptons. We display the summary of the massless spectrum of the model in Table 1 <sup>4</sup><sup>4</sup>4 In Tables 1,2,3 we also display the extra massless fields which might appear if, in addition to the branes discussed above, there are further 3-branes living in the bulk (see fig. 5) of the first two complex dimensions (but at $`Y_3=0`$). These 3-branes may be used to break the left-right symmetry down to the SM, as we discuss in next chapter. They do not modify one-loop coupling unification, though.. Each of the three $`\mathrm{𝟑𝟑}`$ sectors contain a singlet chiral field $`(1)_2`$. As we will see in the next chapter, these three singlets generically get vacuum expectation values of the order of the string scale, giving masses to the extra colour triplets in the $`\mathrm{𝟑𝟕}`$ sector of the model. In this way, the massless spectrum of the model coupling to the $`SU(3)\times SU(2)_L\times SU(2)_R\times U(1)_{BL}`$ group is indeed the one proposed in chapter 2. Thus, the results for gauge coupling unification obtained there directly apply to the present model. ### 3.2 Anomalous $`U(1)`$’s and Fayet-Iliopoulos terms The LR model contains seven Abelian $`U(1)`$ factors. Four of these terms, namely $`Q_3,Q_L,Q_R`$ and $`Q_7`$, appear in the unitary groups of the $`\mathrm{𝟕𝟕}`$ sector, whereas the other three $`Q_{(2,i)}`$ come from each of the $`\mathrm{𝟑𝟑}_{2,i}`$ sectors. As we know, some of these $`U(1)`$’s are anomalous. The spectrum of the model with the corresponding $`U(1)`$ charges is given in Table 2. ¿From there the matrix of mixed $`U(1)`$-non-Abelian anomalies can be computed to be $$T_{IJ}^{\alpha \beta }=\left(\begin{array}{ccccc}0& 9& 9& 0& 0\\ 6& 0& 6& 0& 0\\ 6& 6& 0& 0& 0\\ 0& 0& 0& 21& 21\\ 2& 1& 1& 1& 1\end{array}\right)$$ (3.13) The rows correspond to the seven factors $`Q_3,Q_L,Q_R,Q_7`$ and $`Q_{(2,i)}`$ (same structure repeats for $`i=0,1,2`$). The columns correspond to the nonabelian groups: $`SU(3)`$, $`SU(2)_L`$, $`SU(2)_R`$, $`SU(7)`$ and $`SO(4)`$ respectively. There are two linear independent combinations of above generators which are free of anomalies whereas the other five have triangle anomalies. <sup>5</sup><sup>5</sup>5This is a generic feature of this type of models with 5-branes at just one $`(a,i)`$ ($`a`$ fixed) set of fixed points. If branes are stuck at two different $`a`$’s then there are ten $`U(1)`$’s and three of them are non anomalous. If there are branes at the three sets $`a=0,1,2`$ then four of the thirteen Abelian factors are anomaly free. A possible choice for the non anomalous generators is <sup>6</sup><sup>6</sup>6It is amusing that if one looks at the $`Q_X`$ charges of quarks, leptons and Higgs fields, they are identical to the ones such fields have under the $`U(1)`$ contained in the branching $`E_6SO(10)\times U(1)`$. Here, though, there is no $`E_6`$ nor $`SO(10)`$ symmetry present. $`Q_{BL}`$ $`=`$ $`{\displaystyle \frac{2}{3}}Q_3Q_LQ_R`$ $`Q_X`$ $`=`$ $`Q_RQ_L{\displaystyle \frac{2}{7}}Q_7+2{\displaystyle \underset{i}{}}Q_{n_2^i}`$ while anomalous ones can be chosen as $`Q_{A1}`$ $`=`$ $`Q_{n_2^0}Q_{n_2^1}`$ $`Q_{A2}`$ $`=`$ $`Q_{n_2^0}+Q_{n_2^1}2Q_{n_2^2}`$ $`Q_{A3}`$ $`=`$ $`Q_3Q_L+Q_7`$ $`Q_{A4}`$ $`=`$ $`Q_3Q_RQ_7`$ $`Q_{A5}`$ $`=`$ $`{\displaystyle \underset{i}{}}Q_{n_2^i}+3Q_7`$ (3.15) Notice that even though $`Q_{A1}`$ and $`Q_{A2}`$ present no mixed $`U(1)`$-non-Abelian anomalies, they have cubic and mixed $`U(1)`$ anomalies. Under these new combinations, the charges of the particles in the spectrum are displayed in Table 3. As usual in Type I theory , $`U(1)`$ anomalies are cancelled by a generalized Green-Schwarz mechanism through the coupling to twisted close string RR fields . Anomalous $`U(1)`$ s become massive . At the same time, because of supersymmetry, a Fayet-Iliopoulos term, associated to each of the anomalous groups, appears . The corresponding D-term potential is $$V_r=\frac{1}{2}\left(\xi _r+\underset{l}{}q_r^l|\varphi _l|^2\right)^2$$ (3.16) where $`\varphi _l`$ is the scalar field with charge $`q_l^r`$ under the anomalous group $`U(1)_{Ar}`$. The $`\xi _r`$ $`r=1,\mathrm{}5`$ terms can be explicitly computed (see eq. 3.20 in Ref. ) in terms of the fields $`M_{(a,i)}`$, the Neveu-Schwarz partners of the RR antisymmetric forms mentioned above. We find $`\xi _1`$ $`=`$ $`{\displaystyle \frac{3\sqrt{3}}{2}}\left(M_{02}M_{12}\right)`$ $`\xi _2`$ $`=`$ $`{\displaystyle \frac{3\sqrt{3}}{2}}\left(M_{02}+M_{12}2M_{22}\right)`$ $`\xi _3`$ $`=`$ $`{\displaystyle \frac{\sqrt{3}}{2}}{\displaystyle \underset{i}{}}\left(12M_{0i}+4M_{1i}+5M_{2i}\right)`$ $`\xi _4`$ $`=`$ $`{\displaystyle \frac{\sqrt{3}}{2}}{\displaystyle \underset{i}{}}\left(4M_{0i}+12M_{1i}+5M_{2i}\right)`$ $`\xi _5`$ $`=`$ $`{\displaystyle \frac{\sqrt{3}}{2}}{\displaystyle \underset{i}{}}(7M_{0i}+7M_{1i}+8M_{2i})`$ (3.17) Notice that all the FI-terms $`\xi _r`$ are linearly independent combinations of the $`NSNS`$ twisted moduli. This means that, unlike what usually happens in the perturbative heterotic vacua , there is no need to check for D-flatness of the scalar potentials eq.(3.16). This is because for any field direction of the scalars $`\varphi _l`$ charged under each anomalous $`U(1)_r`$, there will be vevs for the twisted moduli yielding $`\xi _r`$’s compensating them. Furthermore, this indicates, in general, a departure of the orbifold limit and also that the anomalous $`U(1)`$’s do not remain as effective global symmetries, as it would have happened if all twisted moduli were vanishing . ### 3.3 The structure of mass scales The unification of coupling constants in this model takes place at a scale of order $`9\times 10^{11}`$ GeV which should then be identified with the string scale $`M_s`$. In addition that is also the order of magnitude of the compactification scales $`M_1`$, $`M_2`$ of the radii of the first two complex dimensions. This is desirable for two reasons: 1) Since the worldvolume of 7-branes includes the first two complex dimensions, if $`M_{1,2}`$ were much smaller there would be charged Kaluza-Klein fields which might spoil gauge coupling unification; 2) Some phenomenologically interesting non-renormalizable Yukawa couplings involve (see next chapter) 3-branes living at different locations in the first two compact directions. Such couplings would be considerably suppressed if $`M_{1,2}`$ were much smaller than $`M_s`$. On the other hand, the compactification scale $`M_3`$ along the third complex plane (which is transverse to the 7-branes worldvolume) is unconstrained by these considerations. The Planck mass is related to the string scale $`M_s`$ and the compactification scales $`M_i`$ by (see e.g. ) : $$M_p=\frac{2\sqrt{2}M_s^4}{\lambda M_1M_2M_3}=\frac{\sqrt{2}}{\alpha _7}\frac{M_1M_2}{M_3}$$ (3.18) where $`\alpha _7=\frac{\lambda M_1^2M_2^2}{2M_s^4}`$ is the unified coupling of the group coming from 7-branes, which includes the standard model group. Thus, for $`M_{1,2}M_s=9\times 10^{11}`$ GeV, one can obtain the measured $`M_p`$ for $`M_3(100)/\alpha _7`$ TeV. In this scheme (see fig. 5) the size of the $`Y_3`$ coordinate would be thus very large compared to $`Y_{1,2}`$. The present class of models contain anti-3-branes in the bulk in transverse space. As depicted in fig. 5, their worldvolume does not have overlap with that of the “visible world” of 3-branes and 7-branes once the latter are located at the origin in the third compact dimension. Anti-3-branes are instead in the bulk in that dimension. The global configuration of the model is non-supersymmetric, since the supersymmetries preserved by branes are broken by the anti-branes and viceversa<sup>7</sup><sup>7</sup>7It is worth pointing out that even though the presence of the anti-branes explicitly break supersymmetry, the number of massless bosonic degrees of freedom still matches the number of massless fermionic degrees of freedom as can be easily seen in all models of this type, following the general spectrum of reference .. Closed string states living in the bulk of space will generically communicate supersymmetry breaking from the anti-3-brane sector to the visible sector of 3-branes and 7-branes. We will assume that the presence of SUSY-breaking anti-3-branes in the bulk constitutes a SUSY-breaking hidden sector for this model. Since these anti-3-branes live far away in the bulk of the (very large) third complex dimension, SUSY-breaking effects in the 7-branes and 3-branes where the SM resides will be Planck mass suppressed. Thus one expects SUSY-breaking soft terms of order: $$M_{soft}=ϵ\frac{M_s^2}{M_p}$$ (3.19) where the value of the fudge factor $`ϵ`$ will depend on the details of how SUSY-breaking effects in the antibranes are transmitted to the branes by the massless closed string fields. Since gauge coupling unification predicts $`M_s=9\times 10^{11}`$ GeV, in order to get soft terms of order, say 1 TeV, we need <sup>8</sup><sup>8</sup>8This seems to suggest a one-loop transmission of SUSY-breaking to the observable D-brane sectors, as occurs for example in moduli dominated and/or anomaly mediated scenarios. $`ϵ10^2`$. The above assumption of a very large $`Y_3`$ dimensions is a possible simple explanation for the observed large size of $`M_p`$ compared to our predicted $`M_s=9\times 10^{11}`$ GeV. Recently an alternative explanation has been proposed to obtain such an effect which may occur (in some simple models) even if the extra dimensions are infinite. This occurs due to the presence of warp factors in the space-time metric exponentially depending on the extra dimensions. Furthermore, it has also been argued that a localized set of $`D3`$ branes does indeed induce a warped geometry around its location. It would be interesting to explore whether this kind of arguments extend to configurations like the one discussed here which involve intersections of both 3-branes and 7-branes <sup>9</sup><sup>9</sup>9For recent studies of the Randall-Sundrum scenario in the presence of brane intersections see . . An exponential warp factor depending on the dimension $`Y_3`$ transverse to both 3-branes and 7-branes could in this case be a possible alternative origin for the $`M_p`$/$`M_s`$ hierarchy in a model like the one studied here. ### 3.4 Yukawa couplings and conservation rules The general structure of renormalizable couplings in this class of orientifolds was already discussed in ref. . Let us review the couplings involving the supersymmetric sector for the present model, leaving their phenomenological implications for the next section. i) $`(\mathrm{𝟕𝟕})^\mathrm{𝟑}`$ couplings These have the form: $$\varphi _i^{77}\varphi _j^{77}\varphi _k^{77},ijki$$ (3.20) where $`\varphi _i^{77}`$, $`i=1,2,3`$ are any of the charged chiral fields in the $`\mathrm{𝟕𝟕}`$ sector associated to the complex plane $`i`$. These type of couplings give rise for example to quark Yukawa couplings, as we discuss below. The coupling is proportional to the gauge coupling constant for the $`\mathrm{𝟕𝟕}`$ gauge interactions $`g_7`$, which is the one associated to the physical gauge fields. Recall that the latter is related to the string scale $`M_s`$ and the compactification scales $`M_{1,2}`$ of the first two complex planes by: $$\alpha _7=\frac{g_7^2}{4\pi }=\frac{\lambda M_1^2M_2^2}{2M_s^4}$$ (3.21) where $`\lambda `$ is the Type IIB dilaton coupling. It is this $`\alpha _7`$ which provides the boundary conditions for the running of the gauge couplings of the left-right symmetric model. ii) $`(\mathrm{𝟕𝟑})(\mathrm{𝟕𝟑})(\mathrm{𝟕𝟕})`$ couplings These in principle only involve the $`\mathrm{𝟕𝟕}`$ sector associated to the third complex plane : $$\psi _i^{73}\psi _i^{73}\varphi _3^{77}$$ (3.22) where $`i=0,1,2`$ labels the fixed points where the $`3`$-brane is localized. Notice that these couplings are diagonal in the $`i`$ label, i.e., there are no renormalizable couplings involving different fixed points. These Yukawa couplings are also proportional to the $`\mathrm{𝟕𝟕}`$ gauge coupling constant $`g`$. iii) $`(\mathrm{𝟕𝟑})(\mathrm{𝟕𝟑})(\mathrm{𝟑𝟑})`$ couplings In a similar manner there are superpotential couplings of the form $$\psi _i^{73}\psi _i^{73}\varphi _{3,i}^{33}$$ (3.23) in which again $`i`$ labels the fixed point. Again, only the $`\mathrm{𝟑𝟑}`$ chiral fields in the third complex plane appear in the coupling. For example, we already mentioned that there is a coupling of this type between the singlets $`(1)_2`$ in the $`\mathrm{𝟑𝟑}_𝐢`$ sectors and the coloured triplets in the $`\mathrm{𝟕𝟑}_𝐢`$ sectors. Notice however that the gauge coupling $`\stackrel{~}{g}`$ is now different, with $`\stackrel{~}{\alpha }=\lambda /2`$. Several comments concerning the above couplings are in order. From the string point of view, these couplings are obtained from a disk-shaped worldsheet at which boundaries three open string vertex operators are attached. The boundaries of the disk represent the relevant $`p`$-branes, $`3`$-branes and $`7`$-branes in our case. Thus an insertion of a vertex operator of a particle in a $`\mathrm{𝟕𝟑}_𝐢`$ sector turns a $`7`$-brane boundary into a $`3_i`$-brane boundary (and viceversa). This implies that, for the disk worldsheet to make sense, $`\mathrm{𝟕𝟑}_𝐢`$ vertex insertions (for each different $`i`$) have to come in pairs (see figure 6). Thus there is a $`Z_2\times Z_2\times Z_2`$ symmetry which is respected by all disk couplings. This is obviously respected in the couplings discussed above. These symmetries will have an important phenomenological role in the present model, as discussed in the next chapter. A second question concerns the structure of the couplings $`(\mathrm{𝟕𝟕})^\mathrm{𝟑}`$ above. The reader familiar with heterotic orbifold constructions will realize that the same type of couplings involving necessarily the three different complex planes are present for the untwisted particles in those constructions. In the case of heterotic orbifolds this antisymmetric structure may be understood in terms of the conservation of the so called H-momentum (see e.g., ref. for a discussion of these symmetries). The right-moving vertex operators have factors proportional to the RNS fermions which, when bosonized, can be written as $`exp(i𝐩.𝐇)`$ where $`𝐩`$ is an $`SO(10)`$ (space-time) weight. H-momentum conservation is the statement that the overall momentum $`𝐩`$ in a correlator has to vanish for a coupling to be allowed. Now, in the Type I case something completely analogous may be defined leading to equivalent symmetries. The above $`Z_2`$ symmetries and H-momentum conservation rules are still valid for disk (i.e. tree-level) amplitudes leading to non-renormalizable couplings involving charged open string fields. If a given coupling involves branes living at different points in transverse space, it will be exponentially suppressed by the distance between those branes. Thus, for example, couplings involving the $`3_i`$ branes and the $`3_{bulk}`$ branes will get such a suppression. Notice however that if the compactification scales along the first two complex planes $`M_{1,2}`$ are of order the string scale $`M_s`$, no such a suppression will be present. This is in fact the case considered in the previous subsection in order to understand the hierarchy $`M_p>>M_s`$ : only the third complex dimension is large and only $`3`$ branes distant in the 3-d complex dimension are exponentially suppressed. This is what we will assume in the phenomenological analysis in the next chapter. Another point concerning non-renormalizable couplings is the existence of couplings violating the conservation of anomalous $`U(1)`$ symmetries. Tree level couplings renormalizable or not should respect all non-anomalous $`U(1)`$ symmetries. However they may violate anomalous $`U(1)`$ symmetries since those, as we discussed above, are broken by the vevs of twisted NS-NS fields as long as we are away from the orbifold limit (which is the case in the models discussed). As an example of this we show in appendix B how non-renormalizable couplings violating anomalous $`U(1)`$ symmetries are expected to appear (on the basis of heterotic/Type I duality) in the standard $`Z_3`$ orientifold. ## 4 Phenomenology of the D-brane left-right symmetric orientifold In this chapter we discuss several phenomenological aspects of this model. We will not attempt a detailed description of all possible aspects like fermion masses or spontaneous gauge symmetry breaking, rather we will only discuss some possible avenues enabling to address the gross phenomenological issues in this model. i) A nearby vacuum As we mentioned in the previous chapter, apart from the three left-right symmetric generations and Higgs fields, this particular string model has three copies of vector-like colour triplets form the (37) sectors. However, these states are generically massive. Indeed, looking at Table 3 one sees that all gauge interactions allow for a coupling between the three singlets $`1_4`$ from the (33) sectors to the three pairs $`(3,1,1,2/3)_2+(\overline{3},1,1,+2/3)_2`$, where the subindices denote the $`Q_X`$ charge. These Yukawa couplings do indeed exist, as discussed in previous chapter. Thus, if the three singlets $`1_4`$ get a vev of order $`M_s`$, the colour triplets will disappear from the low-energy spectrum and we will be left at low energies with precisely the massless spectrum discussed in chapter 2, leading to very good predictions for gauge coupling unification <sup>10</sup><sup>10</sup>10Notice that, on the other hand, equivalent couplings to the Higgs doublets $`(1,2,2,0)`$ in the $`(77)`$ sector do not exist. Thus a doublet-triplet splitting mechanism is built-in in the symmetries of the model. If we give a vev $`<1_4>_aM_s`$, $`a=1,2,3`$ to these (33) singlets, we have to ensure D-flatness and F-flatness for this direction. In fact we should not care too much about the D-terms of the anomalous $`U(1)`$’s because they can be easily cancelled for appropriate values of the blowing-up fields $`M`$ discussed in chapter 3. On the other hand, the D-term corresponding to the anomaly-free $`Q_X`$ generator has to cancel, which requires giving a vev to some fields with negative $`Q_X`$ charge. A natural option seems to be giving vevs to the antisymmetric $`A^{ij}`$ of $`SU(7)`$ present in the $`(77)_3`$ sector as follows <sup>11</sup><sup>11</sup>11This also turns out to give rise to the required masses for right-handed neutrinos. : $$A^{12}=A^{23}=A^{34}=A^{45}=A^{56}=A^{67}=A^{71}=v$$ (4.1) with $`vM_s`$. This direction can be easily seen to be D-flat and F-flat. Below $`M_R`$ the only $`U(1)`$ interaction left is now $`Q_{BL}`$ since $`Q_X`$ is broken by the above vevs. ii) The breaking of the left-right symmetry and bulk 3-branes The chiral multiplet content of our model below $`M_s`$ includes just thee generations of quarks/leptons/Higgs fields. Some additional fields (particularly, $`SU(2)_R`$ doublets) are needed if we want to break our theory down to the SM gauge group. Probably there is more than one way to modify the model in such a way that one has additional massless $`SU(2)_R`$ doublets for symmetry breaking while the good coupling unification predictions are not spoiled <sup>12</sup><sup>12</sup>12In particular, the variant left-right symmetric model displayed in table 1 of ref. is another possibility. That model has additional matter but one can check that gauge coupling unification along similar lines to those of the present model takes place. A discussion of this variant model is presented in appendix A. The simplest possibility seems to add some additional 3-branes moving in the bulk in the first two compact directions but at the origin in the third compact direction (so that their worldvolume overlaps with that of 7-branes). The simplest set of 3-branes that one can add in the bulk are 6 of them (one 3-brane and their orbifold and orientifold mirrors). They lead to a $`SU(2)`$ gauge group in the $`(\mathrm{𝟑𝟑})_{\mathrm{𝐛𝐮𝐥𝐤}}`$ sector and massless chiral fields in the $`(\mathrm{𝟕𝟑}_{\mathrm{𝐛𝐮𝐥𝐤}})`$ sector transforming like: $$[(3,1,1,2/3;2)+(1,2,1,1;2)+(1,1,2,+1;2)+h.c.]+(1,7;2)^{}+h.c.+(4,1;2)^{}.$$ (4.2) In addition there are chiral fields in the $`(\mathrm{𝟑𝟑})_{\mathrm{𝐛𝐮𝐥𝐤}}`$ sector transforming like $`2(1)+(3)`$ under the $`SU(2)`$ group on the 3-branes. We will see later on when we discuss neutrino masses that the $`SU(2)`$ group coming from this bulky 3-branes should be broken close to the $`M_s`$ scale by vacuum expectation values of the fields $`(1,7;2)^{}+h.c.`$ above. The chiral fields in eq.(4.2) include $`SU(2)_R`$ doublets which can in principle get a vev and break the symmetry. Thus we will assume that some of the fields in $`(1,1,2,+1;2)+(1,1,2,1;2)`$ will get vacuum expectation values of order $`M_R1TeV`$ and break the symmetry to that of the SM. We will briefly discuss below how that could take place due to a radiative symmetry breaking mechanism. One interesting point of the extra particle content provided by the addition of these “bulky 3-branes” is that they give a net vanishing contribution to the combinations $`(B_1^{}+B_L\frac{14}{3}B_3)`$ and $`(B_L\frac{3}{11}B_1^{})`$ which control the joining of coupling constants. Indeed we can easily check that extra contributions to the $`\beta `$-functions are obtained: $$\mathrm{\Delta }B_L=\mathrm{\Delta }B_R=\mathrm{\Delta }B_3=+2;\mathrm{\Delta }B_1^{}=\mathrm{\Delta }B_R+\frac{1}{4}\mathrm{\Delta }B_L=\frac{22}{3}$$ (4.3) so that unification of couplings is not modified <sup>13</sup><sup>13</sup>13This is analogous to the well known fact that complete $`SU(5)`$ representations do not modify the one-loop conditions for unification in the MSSM.at one loop. iii) Quark and charged lepton masses In this model renormalizable quark Yukawa couplings of type $`(77)^3`$ exist with the structure: $$gϵ_{ijk}(3,2,1,1/3)_i(\overline{3},1,2,1/3)_j(1,2,2,0)_k$$ (4.4) where $`i,j,k=1,2,3`$ label the three complex planes and g is the (77) gauge coupling constant. With this simple structure, there would be a massless quark generation and two degenerate generations with masses of order $`g\sqrt{_i|<H_i>|^2}`$. However, this structure is modified by various effects. To start with, the Kahler metric for the $`(77)`$ matter fields in a model like this one needs not be diagonal. There are Kahler untwisted moduli which mix the different complex planes. Furthermore, other effects mixing different complex planes may come from non-renormalizable D-terms like, e.g. $`<21_i21_j^{}>Q_R^iQ_R^j`$. In addition to these, there are mixing terms with the color triplets from the bulky branes. In particular there are renormalizable Yukawa couplings of the form $`(77)(73_{bulk})^2`$: $`(\overline{3},1,2,1/3)_3\times (3,1,1,2/3;2)\times <(1,1,2,1;2)>`$ (4.5) $`(3,2,1,1/3)_3\times (\overline{3},1,1,2/3;2)\times <(1,2,1,1;2)>`$ and hence the right-handed D-quarks of the (77) sector mix with the colour triplets from the $`(73_{bulk})`$ sector once the $`SU(2)_R`$ doublets in that sector get a vev. This means that generically two physical right-handed D-quarks (the three of them in the variant model of the appendix) will be $`SU(2)_R`$ singlets and the other will be contained in a doublet <sup>14</sup><sup>14</sup>14This is analogous to the alternate left-right models considered in refs. . Those models are interesting from the point of view of supression of FCNC, as we comment below.. The second Yukawa coupling above will give masses to the first two D-quarks and the couplings in (4.4) will give masses to the third. Concerning the possible Yukawa couplings for the leptons, the following couplings in the disk are allowed by all non-anomalous gauge symmetries: $$g(1,2,2,0)_3\times (1,2,1,1)_a\times (1,1,2,+1)_a\times f_a(M_\gamma )a=1,2,3$$ (4.6) i.e., the Higgs fields along the third complex plane in the $`(77)`$ sector couple diagonally to the lepton generations. Looking at Tables 2 and 3 we can observe that these couplings are indeed allowed by all non-anomalous symmetries, including the $`Q_X`$ generator. However they are in principle forbidden by the anomalous $`U(1)`$’s which are spontaneously broken at the string scale. The twisted moduli fields discussed in the previous section are charged (non-linearly) under these anomalous $`U(1)`$’s so one expects that upon the insertion of coherent sets of twisted vertex $`M_\gamma `$ operators the above couplings will be allowed, as discussed in previous chapter and exemplified in appendix B. This we denote by the addition of the factor $`f_a(M_\gamma )`$ in the above expression. Notice that this factor does not necessarily mean an exponential suppression, since the physical gauge couplings of the SM gauge interactions are given by the couplings on the $`(77)`$ sector which are proportional to $`\lambda /(M_s^4R_1^2R_2^2)`$, but not to the dilaton $`\lambda `$ itself. Thus $`\lambda `$ need not be too small a number (see also the discussion in appendix B). The precise size of the obtained lepton masses depends on the size of the vev for $`(1,2,2,0)_3`$ and on the value of $`f_a(M_\gamma )`$ for each $`a`$. iv) Neutrino masses In this model there is no right-handed $`SU(2)_R`$ triplet which might give a large Majorana mass to the right-handed neutrinos. However there are fields $`N_a`$ which are singlets under the $`SU(3)\times SU(2)_L\times SU(2)_R\times U(1)_{BL}`$ group and can combine with the right-handed neutrinos which then get a Dirac mass of order $`M_R`$. Specifically, those singlets are contained in the $`(1,7^{})`$ representations in the three $`(37)`$ sectors. For the relevant couplings to appear we have to give vevs of order the string scale to the $`(1,7;2)^{}+(1,\overline{7};2)`$ chiral fields in the $`(73_{bulk})`$ sector. In particular there is a D-flat and F-flat direction along $`\mathrm{\Psi }_6^2=\overline{\mathrm{\Psi }}_6^1=u`$ and $`\mathrm{\Psi }_7^2=\overline{\mathrm{\Psi }}_7^1=iu`$, where in $`\mathrm{\Psi }_r^s`$ ($`\overline{\mathrm{\Psi }}_r^s`$) the index $`r`$ runs over $`SU(7)`$ and $`s`$ over the $`SU(2)`$. Then an effective renormalizable Yukawa coupling is induced at low energies of the form: $$(1,1,2,1)_a\times (1,1,2,+1;2)\times (1,7)_a^{}<(1,21^{})^6\times (1,7;2)h(M_\gamma )>$$ (4.7) where the $`(1,1,2,+1;2)`$ are $`SU(2)_R`$ doublets from the 3-branes in the bulk. One can check that this coupling is allowed by all anomaly-free gauge interactions of the model. As happened with the masses of charged leptons, insertions of twisted moduli fields will be required, which we parameterize by the factor $`h(M_\gamma )`$. Notice that this coupling, since it is non-renormalizable, is in principle suppressed by powers of $`M_s`$. It is of the general form $`(73)^2(73_{bulk})^2(77)^6`$ and hence involves 3-branes located at different points which will also mean exponential suppression in the distance between the location of the 3-branes at the fixed points and those in the bulk. Notice however that, as we discussed in the previous chapter, we have chosen the first two complex compact directions with sizes of order $`1/M_s`$ and hence there is not necessarily any extra suppression, only the third compact complex dimension is assumed to be very large. Once the fields $`(1,1,2,+1;2)`$ get a vev breaking spontaneously the $`SU(2)_R`$ symmetry, the right handed neutrinos inside the three $`(1,1,2,1)_a`$ fields will get a mass of order $`M_R`$ combining with some singlets inside the $`(1,7)_a^{}`$. Notice in this connection that generically the $`SU(7)`$ gauge symmetry is broken and those fields behave indeed like singlet partners of the right-handed neutrinos. In this situation the left-handed neutrinos remain massless. However there are mixing terms from analogous couplings involving $`SU(2)_L`$ doublets of the form: $$(1,2,1,+1)_a\times (1,2,1,1;2)\times (1,7)_a^{}<(1,21^{})^6\times (1,7;2)h(M_\gamma )>$$ (4.8) Then the left handed neutrinos get induced Majorana masses of order $`m_{\nu _L}m_l\times (<(1,2,1,1;2)>/<(1,1,2,+1;2)>)`$. The particular sizes depend on the vev of $`<(1,2,1,1;2)>`$, since $`<(1,1,2,+1;2)>`$ we know is of order $`M_R1`$ TeV. One thus gets neutrino masses of order: $$m_\nu ^am_l^a\times \frac{<(1,2,1,1;2)>}{M_R}$$ (4.9) For $`<(1,2,1,1;2)>m_l`$ a seesaw-like formula is obtained but the precise sizes depend on the unknown values of the vevs of the $`SU(2)_L`$ doublets $`<(1,2,1,1;2)>`$. Notice however that these mass contributions are flavour diagonal, there is no mixing between different lepton families. Thus oscillations can only take place into some inert sterile massless neutrino contained in the original $`(1,7)_a^{}`$ fields. This is not a generic property of the present scenario. One can check that in the variant model described in the appendix mixing between different neutrino flavors can take place, since there are no $`Z_2`$ lepton parities. v) Discrete symmetries and proton stability The couplings in this orientifold model respect a number of discrete $`Z_2`$ symmetries: i) There is a $`Z_2`$ symmetry associated to each of the three $`(37)`$ sectors. Under it all $`(73)`$ particles are odd and the rest are even. Indeed, if we consider the couplings of $`(37)`$ particles on the boundary of the disk, they have to appear in multiplets of two (see fig.6). Since in these sectors live the leptons (and some singlets coming from the $`(1,7^{})`$’s which, as we saw above behave like neutrino-like fields), this corresponded to a discrete $`Z_2`$ lepton number parity. There is one $`Z_2`$ symmetry for each of the three flavours. ii) The flat direction considered gives vevs to the fields $`(1,7;2)^{}+(1,\overline{7};2)^{}`$ and also to some $`SU(7)`$ antisymmetric fields. Thus this direction respects a $`Z_2`$ symmetry under which $`7`$-plets and $`SU(2)`$ doublets (with respect to the $`(33_{bulk})`$ group) are odd. Under this symmetry all quarks and leptons are even but the $`(1,7^{})`$ fields in the $`(73)`$ sectors are odd. The fields in the $`(73_{bulk})`$ are odd, since all are SU(2) doublets. In fact , after breaking of the $`SU(2)_R`$ symmetry by the $`(1,1,2,\pm 1;2)`$ fields and of the $`SU(2)_L`$ by the $`(1,2,2,0)`$ (or, in addition, the $`(1,2,1,\pm 1;2)`$ fields), the diagonal $`Z_2`$ which is the combination of the original $`Z_2`$ and the center of $`SU(2)_R`$ and $`SU(2)_L`$ remains still unbroken. Thus, even after electroweak breaking a $`Z_2`$ symmetry remains under which * All quarks and leptons are odd. * Higgs fields breaking $`SU(2)_L`$ are even. * Singlets combining with right-handed neutrinos are odd. * $`SU(2)_R`$ and $`SU(2)_L`$ doublets in the $`(73_{bulk})`$ sector are even. * Colour triplets in the $`(73_{bulk})`$ sector are odd. Note that this residual symmetry can be identified with the standard R-parity of supersymmetric models. In summary, the effective lagrangian has a residual R-parity symmetry <sup>15</sup><sup>15</sup>15It is well known that a residual R-parity remains in left-right symmetric models if the $`SU(2)_R\times U(1)_{BL}`$ symmetry is broken by $`SU(2)_R`$ triplets $`(1,1,3,2)`$. Notice that this is not the case here and the origin of the residual R-parity is different. and in addition three lepton parities, one per lepton flavour. It is well known that R-parity may be considered as a discrete $`Z_2`$ subgroup of the B-L symmetry. Thus combining it with the lepton parities we thus have a $`Z_2`$ symmetry associated to baryon number. This means that nucleons are stable, since a $`Z_2`$ baryon parity has to be conserved under which baryons are odd and leptons and mesons are even. Thus protons are absolutely stable. On the other hand baryon number can be violated in two units, since there is a $`Z_2`$ symmetry. This means that in principle there can be neutron-antineutron transitions allowed. However those transitions violate B-L symmetry and hence are suppressed by high powers of $`(M_R/M_s)`$ and the rate in this model is negligible. Notice on the contrary that discrete symmetries allow for the neutrino masses discussed in the previous subsection. vi) Gauge symmetry breaking and the low energy spectrum As we discussed in the previous chapter, due to the presence of anti-3-branes in the bulk, one expects the generation of SUSY-breaking soft terms in the effective action. Once soft SUSY-breaking terms appear, $`SU(2)_L\times SU(2)_R\times U(1)_{BL}`$ gauge symmetry breaking can occur due to loop corrections. We will not perform a complete analysis of the (quite involved) scalar potential, but will just study what scalar fields are likely to get vevs once loop corrections are included. As usual they will be the $`SU(3)`$ colour singlets with Yukawa couplings to coloured fields. These include the $`SU(2)_L`$ and $`SU(2)_R`$ doublets in the model, as well as the fields in the $`(33_{bulk})`$ sector which are triplets $`(1;3)`$ under the $`SU(2)_{bulk}`$ gauge group. The following Yukawa couplings appear at the renormalizable level: $`(3,2,1,1/3)_i\times (\overline{3},1,2,1/3)_j\times (1,2,2,0)_k`$ $`(\overline{3},1,2,1/3)_3\times (3,1,1,2/3;2)\times (1,1,2,+1;2)`$ $`(3,2,1,1/3)_3\times (\overline{3},1,1,2/3;2)\times (1,2,1,1;2)`$ $`(3,1,1,2/3;2)\times (\overline{3},1,1,+2/3;2)\times (1;3)`$ $`.`$ The first of these couplings is the $`(77)^3`$ quark Yukawa coupling that we mentioned above. The second and third couplings are of type $`(77)_3(73_{bulk})^2`$ and the fourth of type $`(33_{bulk})(73_{bulk})^2`$. All these four Yukawa couplings tend to give negative mass<sup>2</sup> to the above colour singlet scalars from one-loop diagrams in which the colour triplets circulate in the loop. In addition one also expects generically the presence of trilinear scalar couplings ( ”A-terms”) involving the colour singlet scalars. They are proportional to the scalar couplings $`(1;3)\times (1,2,1,1;2)\times (1,2,1,+1;2)`$ $`+h.c.`$ $`(1;3)\times (1,1,2,+1;2)\times (1,1,2,1;2)`$ $`+h.c.`$ $`(1,2,2,0)_3\times (1,2,1,1;2)\times (1,1,2,+1;2)`$ $`+h.c.`$ $`(1,2,2,0)_3\times (1,2,1,+1;2)\times (1,1,2,1;2)+h.c.`$ The first two have couplings proportional to the $`(33_{bulk})`$ gauge coupling constant whereas the last two are proportional to the $`(77)`$ gauge coupling. The corresponding A-terms are proportional to $`M_{soft}`$ and only involve the corresponding scalars. These contributions to the scalar potential are not positive definite and will favor all $`SU(2)_R`$ and $`SU(2)_L`$ doublets (and the scalars in $`(1,3)`$) to get a non-vanishing vev at some level. We will assume that a stable minimum of the scalar potential exists for vevs of the order of magnitude: $`<(1,1,2,+1;2)><(1,1,2,1;2)><(1;3)>M_R1TeV`$ (4.12) $`<(1,2,2,0)_i>M_Z<<M_R`$ so that the required hierarchy between the left and right gauge symmetries is obtained. Notice that in principle all interactions respect an explicit parity left$``$right symmetry and it is the vacuum which will explicitly break parity symmetry and decide who is left-handed and who is right-handed. Whatever $`SU(2)`$ survives to lower energies we will call $`SU(2)_L`$ by definition. Another relevant question is what is the mass of the extra Higgs and Higgsino fields that this model has both from the $`(77)`$ and $`(73_{bulk})`$ sectors. This is a complicate issue which will depend on the detailed structure of vevs. Looking at the first, third and fourth couplings in eqs.(4) we see that vevs of order $`M_R`$ for $`(1;3)`$ and $`(1,1,2,\pm 1;2)`$ will make massive some of the $`SU(2)_L`$ doublets in the $`(73_{bulk})`$ sector and also the $`(1,2,2,0)_3`$ fields in the $`(77)`$ third complex plane. In this situation we would be left at low energies with the fields $`(1,2,2,0)_1`$ and $`(1,2,2,0)_2`$ corresponding to the first two complex planes. However, as we mentioned when we discussed quark Yukawa couplings, there are different effects which will generically mix the particles living in different complex planes in $`(77)`$ sectors. Thus one also expects that these other doublets could become massive. We will thus assume that at a scale of order $`M_R`$ only one set of SM doublets remains relatively light, so that they are available for $`SU(2)_L`$ spontaneous symmetry breaking. In addition there are the extra right-handed chiral fields $`(1,1,2,\pm 1;2)`$ from the $`(73_{bulk})`$ sector. Some of these where eaten in the process of $`SU(2)_R`$ breaking. The remaining may acquire a mass of order $`M_R`$ from the second equation in (4) , once the scalars $`(1;3)`$ get a vev. The same applies to the extra colour triplets $`(3,1,1,1/3;2)+h.c.`$ from the $`(73_{bulk})`$ sector. We already mentioned that some combination of them mixes with the right-handed quarks from the $`(77)`$ sector. The orthogonal combination will get a mass of order $`M_R`$ once the scalars $`(1;3)`$ get a vev. All in all, the spectrum below the $`M_R`$ would thus be similar to that of the MSSM: three quark-lepton chiral multiplets and one set of $`H_u+H_d`$ Higgs fields. vii) Ramond-Ramond fields and invisible axions We already mentioned in chapter 3 that in this class of orientifold models there are twisted Ramond-Ramond singlet scalars which couple to $`F\stackrel{~}{F}`$. As already discussed in ref. , they are natural candidates to play the role of invisible axions in a model like this. Notice however that, in the absence of other charged scalar vevs, the combinations of twisted Ramond-Ramond fields coupling to the gauge groups get in fact large masses of order the string scale $`M_s`$ by providing the longitudinal degrees of freedom of anomalous $`U(1)`$’s when the latter become massive. This can be easily seen from eq.(3.16). Now, again from eq.(3.16), since $`\xi _r0`$, in the presence of other charged fields from the open string sector acquiring a vev and contributing to the anomalous $`U(1)`$ breaking, there will be a linear combination of RR-field plus the charged field, which will be swallowed by the $`U(1)`$ to become massive. The orthogonal combination will remain massless. This massless linear combination will in general couple to the gauge fields (and in particular, to QCD) in the standard axionic fashion with the gauge kinetic function taking the general form $`f_\alpha =S+s_\alpha ^{(ai)}M_{(ai)}`$ with $`s_\alpha ^{(ai)}`$ constant computable coefficients and with a decay constant of order the string scale $`M_s10^{12}`$ GeV, well within astrophysical limits. viii) Experimental signatures The most obvious experimental implication of the present scheme is the existence of extra $`W_R`$, $`Z_0^{}`$ gauge bosons corresponding to the left-right symmetric gauge interactions at a scale of order 1 TeV or below. The phenomenology of $`SU(3)\times SU(2)_L\times SU(2)_R\times U(1)_{BL}`$ models has been extensively studied in the past, although most of the studies have tacitly assumed an $`SO(10)`$ embedding of such gauge symmetry leading to the canonical value for the weak angle . In addition, many studies have concentrated on a scheme in which $`SU(2)_R`$ chiral triplets transforming like $`(1,1,3,2)+h.c.`$ break the left right symmetry. At the same time these vevs could give rise to large Majorana masses for the right-handed neutrinos, leading to a see-saw structure for neutrino masses. This kind of Higgs fields do not appear in the class of models that we construct, and right-handed neutrinos are expected to become massive by combining with other singlet chiral fields, as explained above. Thus many previous studies do not directly apply to the present model. There are a number of experimental limits on the masses of the extra gauge bosons . If right-handed neutrinos are lighter than the $`W_R`$ mass, the channel $`W_Rl_R\nu _R`$ is open leading to clean signatures at the Tevatron. From searches in that channel $`D0`$ has set the limit $`M_{W_R}>720`$ GeV and $`CDF`$ $`M_{W_R}>650`$ . If right-handed neutrinos are heavier than the $`W_R`$, this signature disappears and weaker limits coming from dijet production are obtained. $`D0`$ excludes the range $`340<M_{W_R}<680`$ GeV whereas $`CDF`$ excludes $`300<M_{W_R}<420`$ . UA2 had excluded the energy range $`100<M_{W_R}<251`$ also from the dijet signature . There are stronger constraints on the $`W_R`$ mass from the $`K_LK_S`$ mass difference but those are much more model dependent . On the other hand, for left-right symmetric models with $`g_L=g_R`$ like this, one can obtain limits from precision LEP-I measurements and low-energy neutral current data yielding $`M_{Z_0^{}}>900`$ GeV, implying $`M_{W_R}>780`$ GeV in this class of models . In summary, the extra gauge bosons appearing in a left-right symmetric model like this should weight more than around 800 GeV or so <sup>16</sup><sup>16</sup>16If the right-handed D-quarks are mostly $`SU(2)_R`$ singlets as discussed above, $`W_R`$ production is very much supressed and direct limits on the $`W_R`$ mass are much weakened . However that is not the case for the $`Z`$’ and hence mass limits of that order are expected to still apply. Masses of this size or a bit higher are compatible with the gauge coupling unification results in chapter 2. However those coupling unification results seem to prefer not very high masses for $`W_R`$ and $`Z_0^{}`$. <sup>17</sup><sup>17</sup>17 It has been recently pointed out that a small amount of missing invisible width in $`Z`$ decays at LEP I as well as atomic parity violation experiments could perhaps indicate already the existence of some extra $`Z`$’ with mass of order 500-1000 GeV.Thus the extra left-right symmetric degrees of freedom could perhaps soon be discovered. In addition in this class of models there is a triplication of the number of Higgs fields transforming like $`(1,2,2,0)`$. Thus one also expects to find at energies of order 1 TeV, charged and neutral Higgs and Higgsino fields. In fact these fields are potentially dangerous. Indeed, it is well known that in generic models with multiple Higgs $`SU(2)_L`$ doublets, the unitary transformations which diagonalize the quark mass matrices do not necessarily diagonalize the Yukawa interactions and FCNC can in principle appear. This FCNC problem is generically present in left-right symmetric models with a low $`M_R`$ scale like this. Thus to suppress sufficiently such kind of transitions, the extra Higgs fields have to be sufficiently heavy and/or the Yukawa couplings will need to have some symmetries. The question of how to evade the problem of FCNC in supersymmetric left-right models with $`M_R1`$ TeV has been adressed in ref. . There it is shown that this problem can be avoided if there are present some extra $`SU(2)_R`$ singlet D-type quarks mixing with the right-handed doublet quarks in the model in such a way that the physical D-quarks are mostly $`SU(2)_R`$ singlets. In addition extra $`SU(2)_R`$ and $`SU(2)_L`$ doublets with non-vanishing B-L charge are also required . Interestingly enough this type of extra fields and mixings are also present in the D-brane model here discussed. It would be interesting to see whether in a D-brane type of model a similar mechanism as in ref. could be made operative. Finally, there are also extra fields from the sector which is in charge of the breaking of the $`SU(2)_R`$ symmetry. In order not to spoil gauge coupling unification we have seen that the required $`SU(2)_R`$ doublets should come along with the same number of $`SU(2)_L`$ doublets and coloured $`SU(3)`$ triplets. In the D-brane explicit constructions presented in this paper such structure of extra particles is very natural and corresponds to the addition of some extra 3-branes in the bulk of the first two complex dimensions. All these extra particles are expected to have masses also around $`M_R`$ or the SUSY-breaking scale. Thus new extra heavy lepton-like and quark-like extra fermions and scalars are expected to be produced at future accelerators corresponding to this sector. In this chapter we have given an overall discussion of phenomenological aspects of the particular orientifold model introduced in chapter 3. We believe that the particular $`Z_3`$ orientifold here discussed is tantalizingly close to the general new scheme proposed in chapter 2. Analogous models based on the same orientifold can be constructed which will lead to somewhat different phenomenological properties. For example, one can construct similar models by adding a second discrete Wilson line breaking e.g. the original $`SU(7)`$ symmetry further. Another possibility is to consider the variant left-right symmetric discussed in the appendix . ## 5 Final comments and outlook During more than 15 years the minimal supersymmetric standard model has been thoroughly studied as the best motivated extension of the standard model. During the past 10, this study became more intense due to the realization that gauge coupling unification works very well in the MSSM as compared to the standard model. In this article we have proposed an interesting alternative to the MSSM mostly motivated by the structure of D-brane models. This alternative scenario matches the success of gauge coupling unification of the MSSM, with the interesting feature that unification works only if there are three families of quarks, leptons and Higgs fields. In the present scheme several physical mass scales are unified: the string unification, susy-breaking and axion scales are one and the same and of order $`M_s=9\times 10^{11}`$ GeV. As it has been noted in several occasions, , this intermediate scale may have important physical implications regarding neutrino masses, the strong CP problem, ultra high-energy cosmic rays, non thermal dark matter candidates, inflation, etc. We have also presented a concrete D-brane model satisfying most of the general properties of the proposed scenario. The existence of three quark-lepton generations has an elegant explanation in these constructions: there are three generations because there are three complex compact dimensions and a $`Z_3`$ orbifold structure<sup>18</sup><sup>18</sup>18This simple explanation, first encountered in the heterotic models in , is to be compared with that in Calabi-Yau compactifications of perturbative or non-perturbative heterotic vacua in which the net number of generations minus antigenerations is related to the Hodge numbers of the CY. Those tend to be quite large in general and only for very particular manifolds is small.. The proton is stable due to a combination of discrete symmetries including lepton and R-parities and Yukawa couplings are obtained for quarks, charged leptons and neutrinos. To our knowledge this is the first example where R-parity appears so naturally in string theoretical models. This guarantees the existence of an LSP and most of the standard searches for supersymmetry. Being a left-right symmetric model, it shares many of the good properties that have been realized over several years about LR models . Recall, however, that the Higgs structure we find differs from the standard treatment in the sense that the fields responsible for breaking the LR group to the standard model are doublets that belong to the spectrum of the corresponding string model instead of $`SU(2)`$ triplets, as often assumed in the literature . One may ask if going to a left-right symmetric model is unavoidable in this class of theories. Indeed, one can construct explicit three generation $`Z_3`$ orientifold models with the gauge group of the standard model and some examples of this type were presented in ref. . However gauge coupling unification does not appear as naturally as in the left-right symmetric models here described. Furthermore, the absence of the $`BL`$ gauged symmetry makes difficult to find field directions with a sufficiently stable proton. We believe that this is probably a generic property and both coupling unification and proton stability seem to point towards a left-right symmetric extension of the SM above a TeV scale. Since the LR scale, $`M_R`$, has to be relatively close to the TeV scale in order for gauge unification to work, this scenario can be experimentally tested very soon. In the first place, the existence of new massive gauge fields for which the experimental constraints are becoming very strong can be explored. The scenario can also be tested by looking at the existence of new Higgs particles at the TeV scale, which in principle can give rise to flavour changing neutral currents, and of course with the production of supersymmetric particles. There are several aspects that remain to be understood in this scenario. First, the Higgs scalar potential needs to be studied in order to understand the possible patterns of gauge symmetry breaking. This is also important in order to address the issue of possible FCNC transitions coming from the multi-Higgs structure. Then, the detailed structure of soft supersymmetry breaking terms needs to be addressed. This is important in order to have more information about the possible experimental signals of supersymmetry. Other issues, such as the origin of baryogenesis, may depend crucially on this knowledge. A through analysis of the general structure of neutrino masses and oscilations in this kind of scenario would also be important. As for the explicit string model constructed, it would be interesting to study different variations e.g. with further symmetry breaking (Wilson lines) or locations of 3-branes which may lead to different phenomenological details. Also interesting would be to look for similar left-right symmetric models using other constructions like Type IIA orientifolds or non-perturbative heterotic orbifolds leading to similar spectra. In addition the standard issues of moduli stabilization <sup>19</sup><sup>19</sup>19For some recent attempts to understand the stabilization of large radii see e.g. ref.. and the cosmological constant <sup>20</sup><sup>20</sup>20For some recent ideas about the cosmological constant problem in the D-brane context see e.g. refs. . remain open. It is highly remarkable that in this string model, the unification scale coincides with the preferred fundamental string scale determined by supersymmetry breaking, since a priori the two scales did not have to be related. If this turns out to be true, we might say that nature has been misleading us for many years into the belief that the unification scale was much higher. In any case, we believe the present new scenario has many appealing properties and deserves to be considered as a serious alternative to the MSSM. Acknowledgements We acknowledge useful conversations with B. Allanach, M. Klein, R. Rabadán and A. Uranga. This work has been partially supported by CICYT (Spain), the European Commission (grant ERBFMRX-CT96-0045) and PPARC. G.A work is partially supported by APCyT grant 03-03403. ## 6 Appendices ### 6.1 Appendix A A variant left-right symmetric orientifold model In reference , we found two LR models with three families in the study of $`ZZ_3`$ orientifold models with a single Wilson line. The model discussed in the text corresponds to the first of such models. Here, for completeness, we will briefly describe the phenomenological aspects of the second model. Following the notation of chapter 3, the model can be defined by the following shift vector and Wilson lines: $`V_7`$ $`=`$ $`{\displaystyle \frac{1}{3}}(1,1,1;1,1;0,0;0;1,1,1,1,1,1,1,1)`$ $`W`$ $`=`$ $`{\displaystyle \frac{1}{3}}(1,1,1;1,1;1,1;0;0,0,0,0,0,0,0,0)`$ (6.1) which leads to a $`\mathrm{𝟕𝟕}`$ sector group $`U(3)\times U(2)_L\times U(2)_R\times [SO(2)\times U(8)]^{}`$. We must also add four 3-branes at each of the six points at $`(a,i)`$ ($`a=0,1`$ and $`i=0,1,2`$). The gauge group on each of the six 3-branes will then be $`Sp(2)\times U(1)`$. Similar to the previous model, the anti-branes live in the bulk and they total 24 (equals to the total number of 3-branes). Again, in the figure we only show 4 of them, all the other ones are located at the images points under the $`Z_3\times Z_2`$ orientifold symmetry. Since there are no branes at the $`(2,i)`$ points we will have three non-anomalous and seven anomalous $`U(1)`$’s. We choose the linear combinations $`Q_{BL}`$ $`=`$ $`{\displaystyle \frac{2}{3}}Q_3+Q_LQ_R`$ (6.2) $`Q_{non}^L`$ $`=`$ $`{\displaystyle \frac{Q_L}{2}}+{\displaystyle \frac{Q_8}{8}}{\displaystyle \frac{2}{3}}{\displaystyle \underset{i}{}}Q_{n_0^i}{\displaystyle \frac{1}{3}}{\displaystyle \underset{i}{}}Q_{n_1^i}`$ $`Q_{non}^R`$ $`=`$ $`{\displaystyle \frac{Q_R}{2}}+{\displaystyle \frac{Q_8}{8}}{\displaystyle \frac{1}{3}}{\displaystyle \underset{i}{}}Q_{n_0^i}{\displaystyle \frac{2}{3}}{\displaystyle \underset{i}{}}Q_{n_1^i}`$ $`Q_A^8`$ $`=`$ $`Q_8`$ $`Q_A^r`$ $`=`$ $`Q_{n_a^i}`$ (6.3) with $`a=0,1`$ and $`i=0,1,2`$. The first 3 are non anomalous whereas the last seven are anomalous. The spectrum with the corresponding charges under the $`U(1)`$’s is shown in the table 4. We present the quantum numbers under the $`U(1)^{10}`$ groups. $`Q_3,Q_L,Q_R`$ and $`Q_8,Q_{n_0^i}^A,Q_{n_1^i}^A`$. The last seven can be chosen as independent anomalous combinations. $`Q_{BL},Q_L^{non}`$ and $`Q_R^{non}`$ are the non anomalous charges. Mixed anomalies can be computed as we did for the model in the text and they are exactly cancelled. The anomaly matrix is: $$T_{IJ}^{\alpha \beta }=\left(\begin{array}{cccccccc}0& 9& 9& 0& 3& 3& 3& 3\\ 6& 0& 6& 0& 2& 0& 2& 0\\ 6& 6& 0& 0& 0& 2& 0& 2\\ 0& 0& 0& 24& 8& 8& 8& 8\\ 1& 1& 2& 1& 3& 0& 3& 0\\ 1& 2& 1& 1& 0& 3& 0& 3\end{array}\right)$$ (6.4) Where the rows correspond to the 10 $`U(1)`$’s ordered as in table 4. The columns correspond to the groups: $`SU(3)`$, $`SU(2)_L`$, $`SU(2)_R`$, $`SU(8)^{}`$, $`U(1)_0^i,U(1)_1^i`$ and $`Sp(2)_0^i`$, $`Sp(2)_1^i`$ respectively. Looking at the spectrum of this model we can immediately see that contrary to the model in the text, there are enough fields on this model to break the group to the standard model one, without the need to add extra branes in the bulk. For instance, if we give a nonvanishing vev to the $`Sp(2)`$ doublets in the 33 sectors we are left with a spectrum in the visible sector consisting only of the three families of quarks, leptons and Higgs fields plus extra matter with exactly the same quantum numbers as the ‘bulk’ matter fields of the model in the text (equation (4.2)), although in three copies. <sup>21</sup><sup>21</sup>21There is an ambiguity about which doublets are assigned to be the leptons and which would be extra matter fields. In fact we can see that the fields $`(1,2,1;2)`$ in the $`\mathrm{𝟑𝟕}_0`$ sector and one of the $`(1,2,1;1)`$ of the $`\mathrm{𝟑𝟕}_1`$ sector have the same quantum numbers as the leptons. One of the components of the $`(1,2,1;2)`$ field will get a mass after the $`Sp(2)`$ doublet in the $`\mathrm{𝟑𝟑}_0`$ sector gets a vev, the remaining component is what we will identify as the physical leptons. The other choice does not have the appropriate Yukawa couplings.. As we have mentioned before, the quantum numbers of those fields are such that they do not modify the analysis of the gauge coupling unification. Therefore we conclude that, quite remarkably, the present model also shares the good properties about gauge coupling unification as the model presented in the text. In order to cancel the D-terms generated by the vevs of the Sp(2) doublets one has to give vevs to other doublets to compensate. The simplest option is to take flat directions for the hidden sector fields such as $`(2,\overline{8})^{}`$ of the $`\mathrm{𝟑𝟕}_{0,1}`$ sectors as well as to the $`(1,28)^{}`$ of the 77 sector. This allows in addition to generate nonvanishing charged lepton and neutrino masses from couplings such as<sup>22</sup><sup>22</sup>22Yukawa couplings giving rise to quark masses are identical to the model in the text since they only include couplings among fields of the 77 sector. $`(1,2,2)(1,2,1;2)^{}(1,1,2;2)^{}<(1,\overline{8};2)^{}(1,\overline{8};2)^{}(1,28)^{}><(1,28)^4>`$ $`(1,1,2;2)^{}(1,1,2)(1,1)_{0,i}<(1,2)_{0,i}(1,\overline{8};2)^{}(1,\overline{8};2)^{}(1,28)^{}>`$ (6.5) where the on the $`Sp(2)`$ doublets stand for the component of the doublet that remains massless after the doublet in the $`\mathrm{𝟑𝟑}`$ sector got a nonvanishing vev and the subindex $`a,i`$ indicates that these are fields from the $`\mathrm{𝟑𝟕}_a`$ sector. Notice that in this case the right-handed neutrinos become massive by combining with singlets in the $`\mathrm{𝟑𝟑}`$ sectors. These couplings are allowed by all the gauge symmetries including the anomalous $`U(1)`$’s, therefore, contrary to the model in text, there is no need to introduce insertions of twisted vertex operators in order to generate lepton masses. Interestingly enough, this flat direction leaves a remaining $`Z_2`$ symmetry consisting in changing sign to all $`SU(8)`$ octets and simultaneously changing sign to all particles in sectors $`\mathrm{𝟑𝟕}`$. It is easy to see that this discrete symmetry combined with the two $`Z_2`$’s remaining after breaking the LR $`SU(2)`$’s imply a residual discrete baryon $`Z_2`$ parity making the proton absolutely stable, similar to the model in the text. Unlike that model however, lepton number is not conserved since couplings like $`QDL`$ are permitted, although suppressed by powers of $`(M_R/M_s)`$ since they violate $`BL`$. Combining these $`SU(8)`$ flat directions with those of the $`Sp(2)`$ doublets does not preserve gauge coupling unification since couplings such as $$(3,1,1;2)^{}(\overline{3},1,1;2)^{}<(1,\overline{8};2)(1,\overline{8};2)(1,28)>$$ (6.6) will in principle give masses to the extra triplets. Thus an alternative which does preserve gauge coupling unification would be combining the $`SU(8)`$ flat directions with those of singlets in the $`\mathrm{𝟑𝟑}`$ sectors (not giving vevs to the Sp(2) doublets). This maintains gauge coupling unification with the good properties for lepton masses. In this case giving a mass to all the extra triplets needs the insertion of twisted vertex operators. In summary this model is very similar to the one presented in the text (although probably a bit more complicated) and may deserve further exploration. ### 6.2 Appendix B Anomalous $`U(1)`$’s and non-renormalizable couplings in the $`Z_3`$ orientifold In this appendix we will argue that non-renormalizable couplings violating anomalous $`U(1)`$ charges are in general expected in orientifolds similar to the ones considered in the present article. More specifically heterotic/Type I duality seems to indicate that this is the case. Consider in particular the standard $`D=4`$, $`N=1`$, $`Z_3`$ orientifold first constructed in ref.. The underlying orbifold is exactly the same than the one considered in the present paper, the only difference being the absence of anti-branes and Wilson lines. The model has 32 9-branes and a gauge group $`SU(12)\times SO(8)\times U(1)_A`$. In the open string $`(99)`$ sector there are charged fields transforming like: $$3(\overline{66},1)_{+2}+3(12,8)_1$$ (6.7) where the subindex denotes the $`U(1)_A`$ charge. It is easy to check that the $`U(1)_A`$ interaction is anomalous. In addition there are 27 chiral singlets $`(1,1)_0`$ coming from the twisted closed string sector. The latter are singlets under $`U(1)_A`$ but transform non-linearly under that anomalous symmetry, and it is this transformation which cancels the $`U(1)_A`$ anomalies by means of a generalized Green-Schwarz mechanism. At the same time the $`U(1)_A`$ becomes massive. Now, $`SU(12)`$ invariance allows for a non-renormalizable superpotential coupling of the type $`(\overline{66},1)_{+2}^6`$ by contraction with the 12-index antisymmetric tensor. However, such a coupling would violate $`U(1)_A`$ conservation in 12 units. Since this $`U(1)_A`$ symmetry is broken one may suspect that such non-renormalizable coupling may however exist if one moves away from the orbifold limit, i.e., if the twisted NS-NS singlet fields get a vev. Heterotic/Type I duality seems to indicate that is the case. The above orientifold has a heterotic dual which is a $`Z_3`$ orbifold compactification of the $`SO(32)`$ heterotic string. This model has the same gauge group and the charged particles in the untwisted sector are identical to those in the $`(99)`$ sector of the Type I model. There are also 27 twisted chiral fields transforming like $`(1,1)_4`$ under the gauge group, which should be identified as duals of the 27 twisted fields of the Type I dual. Notice however that, unlike their Type I counterparts, the heterotic singlets are charged under the anomalous $`U(1)_A`$. Now, one can convince oneself that in this heterotic $`Z_3`$ orbifold the following non-renormalizable couplings do in general exist: $$(1,1)_4^3\times (\overline{66},1)_{+2}^6$$ (6.8) This coupling preserves the $`U(1)_A`$ symmetry. One can also check that this coupling preserves the necessary selection rules in order to be present. In particular, it respects the $`Z_3`$ point group symmetry (the twisted fields appear to the third power). Also the H-momentum conservation is obeyed. To check this it is enough to consider the vertex operators associated to the $`(\overline{66},1)_{+2}`$ written in the ” 0-picture”. In addition, as pointed out in ref. , the FI-term present in this heterotic model forces the singlets $`(1,1)_4`$ to get non-vanishing vevs. Thus away from the orbifold point the heterotic model will present $`(\overline{66},1)_{+2}^6`$ couplings, effectively violating the anomalous $`U(1)_A`$ symmetry. This strongly suggests that similar couplings will also be present in the dual Type I model as long as one stays away from the orbifold point, i.e., as long as the twisted $`NSNS`$ closed string singlets have non-vanishing vevs. This is indeed the case considered in the models in the present article.
warning/0001/cond-mat0001116.html
ar5iv
text
# Semiclassical description of electronic supershells in simple metal clusters ## Abstract A semiclassical approach for calculating shell effects, that has been used in atomic and plasma physics, is applied to describe electronic supershells in metal clusters. Using the spherical jellium model we give the analytical expression for the oscillating part of the binding energy of electrons as an explicit sum of contributions from supershells with quantum numbers $`2n_r+l`$, $`3n_r+l`$, $`4n_r+l`$, … This expression is written in terms of the classical characteristics of the motion of an electron with the Fermi energy in a self-consistent potential. The conditions under which a new supershell appears and the relative contribution of this shell are studied as a function of the cluster size and form of the potential. Specific calculations are performed for a square well. 1. The electronic structure of atomic clusters has been intensively studied experimentally and theoretically in the last two decades (see, for example, reviews in Refs. and ). One general feature of the experimental mass-abundance spectra of $`N`$-atom clasters is the existence of magic numbers: Clusters with these numbers of atoms are produced more abundantly. As $`N`$ enlarges, the amplitude of these variations lessens, then in still larger clusters increases once again, and so on, i.e., oscillations with beats occur. The theoretical reproduction of the beating patterns one can see, for example, in the self-consistent calculations of sodium clusters in Ref. , where the spherical jellium model in local density approximation was used. There are two kinds of periodicity in the density of states and in the oscillating part $`\delta E_{sh}`$ of the binding energy as a function of a cluster radius $`RN_e^{1/3}`$ ($`N_e`$ is a number of valence electrons in cluster): the primary shell periodic structure and beat mode with a higher-order period. The similar results were performed in Ref. using the Woods-Saxon potential and in calculations for various metals in Ref. using nearly self-consistent potentials. For small clusters the interpretation of the numerical results is obvious: the extremum energy cusps occur at $`l_{max}`$-shell closing and represent the magic numbers ($`l_{max}`$ is a maximum orbital number $`l`$ in the cluster). However this law breaks down for $`N_e>100`$. The theory, explaining quantitatively the beating pattern of the level density for spherical cavity by a superposition of the contributions from closed classical trajectories of electrons, has been developed in the fundumental work of Balian and Bloch . The detailed numerical calculations in Ref. using more complex spherical cluster potentials reproduce the similar oscillations, one-electron levels $`\epsilon (n_r,l)`$ of the higher angular momentum states bunching in supershells: $`\epsilon (n_r,l)\epsilon (n_r+1,lK)`$ with pseudoquantum numbers $`Kn_r+l`$. Here $`n_r`$ is a radial quantum number, $`K=2,3,4,\mathrm{}`$. The integer number $`K`$ under classical treatment (see, for example in Ref.) is equal to a ratio of the frequencies in the radial and angular motion of the corresponding closed orbit. At $`K=2`$ pseudoquantum number is the same as a principal quantum number and characterizes the pendulating orbit going through the origin, $`K=3`$ and $`K=4`$ corresponds to the triangular and square orbits, respectively. One can suppose from here that a periodic-orbit-expansion (or supershell- expansion), obtained for a spherical cavity, is a particular case of the more general expansion for spherical cluster potentials. This conclusion is supported in the recent paper where such a generalization was carried out using Woods-Saxon potential by expanding on a parameter $`a/R`$ ($`a`$ is a surface width) around the known results for a spherical potential well. It is of interest to investigate analytically the origin of supershells and the mechanisms leading to their appearance for a potential of arbitrary form. In the present paper it is shown that this problem can be solved by a semiclassical method for distinguishing shell effects, previously applied successfully in atomic ,, and plasma , physics on the basis of the Thomas-Fermi (TF) model. Although the TF model and its conventional variants ETF with quantum and exchange corrections (for application to clusters, see, for example,Refs. and ) give only the average dependences of all quantities on the number of particles, a refinement of this model makes it possible to account for the shell structure of the electronic spectrum. This refinement is based on the use of the Bohr-Zommerfeld quantization conditions and on the possibillity of performing the sum over quantum numbers analytically, provided that the semiclassity parameter, which for clusters is proportional to $`N_e^{1/3}`$, is small. 2. We furthermore use the expression derived in our future extended publication for the correction to the binding energy of electrons (the atomic units are used): $$\delta E=\underset{\mathrm{}}{\overset{\mu }{}}𝑑\mu ^{}𝑑𝐫\delta n(𝐫,\mu ^{})$$ (1) Here $`\delta n(𝐫)`$ is the correction to the electronic density, which because of other effects goes beyond the initial model approach (for example, TF or ETF model), $`\mu `$ is the chemical potential in the initial model, the correction $`\delta n(𝐫)`$ is assumed to be small and is calculated using the initial self-consistent potential. We are interested in the contribution of the shell correction to the electronic density of states $`\delta n_{sh}(𝐫,\mu )`$ or to the number of states: $$\delta N_{sh}(\mu )=𝑑𝐫\delta n_{sh}(𝐫,\mu )=N(\mu )N_{TF}(\mu ),$$ (2) where for a cluster with the closing ”$`l`$”-shells $$N(\mu )=2\underset{n_r,l}{}(2l+1)\theta (\mu \epsilon _{n_r,l}),$$ (3) and in the semiclassical approach the energy levels $`\epsilon _{n_r,l}`$ are determined from the quantization condition $$S_{\epsilon l}=𝑑rp_{\epsilon l}(r)=\pi \left(n_r+\frac{1}{2}\right)$$ (4) Here $`S_{\epsilon l}S_{\epsilon \lambda }`$ and $`p_{\epsilon l}(r)=\sqrt{2(\epsilon U(r))(l+1/2)^2/r^2}\sqrt{p_\epsilon ^2(r)\lambda ^2/r^2}p_{\epsilon \lambda }`$ are, respectively, the classical radial action and momentum of an electron with energy $`\epsilon `$ and orbital angular momentum $`l`$. Simple calculations using the Poisson formula to replace the sums over quantum numbers $`n_r`$ and $`l`$ by integrals make it possible to rewrite expression (3) as $$N(\mu )=\frac{2}{\pi }\underset{k,s=\mathrm{}}{\overset{\mathrm{}}{}}\frac{(1)^{k+s}}{k}\underset{0}{\overset{\lambda _\mu }{}}𝑑\lambda \lambda \mathrm{sin}(2\pi k\nu _{\mu \lambda })\mathrm{cos}(2\pi s\lambda )$$ (5) Here $`\nu _{\epsilon \lambda }=S_{\epsilon \lambda }/\pi `$, and $`\lambda _\epsilon `$ determines the border of the phase area of the classically allowed motion of an electron with an energy $`\epsilon `$: $`\nu _{\epsilon \lambda _\epsilon }=0`$. In Eq.(5) the term with $`k=s=0`$ gives the TF result $`N_{TF}(\mu )`$ and according to Eq.(2) the sum (5) without this term is equal to the desired quantity $`\delta N_{sh}(\mu )`$. Let’s note now, that the supposed approach is agreed with a general, known from the nuclear physics, concept , of separating the total energy of finite sistem as a function of the system size into a smooth part and a fluctuating correction. But here for calculating the last we base the simple expression (1), using initial smooth self-consistent potential or its Woods-Saxon and another fitting. 3. The integral limits and points $`\overline{\lambda }`$ of stationary phase make the main contribution in the integral over $`\lambda `$ in Eq.(5). These points are determined from the relation: $$\frac{\nu _{\mu \lambda }}{\lambda }|_{\overline{\lambda }}=\frac{s}{k},0\lambda \lambda _\mu $$ (6) A function $`\nu _{\mu \lambda }(\lambda )`$ decreases monotonically and for all potentials $`U(r)`$ that are finite at the origin the slope of the corresponding curve at $`\lambda =0`$ is the same , $$\frac{\nu _{\mu \lambda }}{\lambda }|_0=\frac{1}{2},$$ $`(6a)`$ The value of the derivative at $`\lambda =\lambda _\mu `$ $$\frac{\nu _{\mu \lambda }}{\lambda }|_{\lambda _\mu }\nu _\mu ^{}$$ $`(6b)`$ depends strongly on the form of the potential. For an oscillator $`\nu _\mu ^{}=1/2`$, for a square well $`\nu _\mu ^{}=0`$, and for the Woods-Saxon potential the value of $`\nu _\mu ^{}`$ varies with increasing $`N`$, vanishing in the limit of a very large number of atoms. On this basis it follows that the relation (6) distinguishes in the sum over $`k`$ the leading terms $`k=(2+j)s,{\displaystyle \frac{\nu _{\mu \lambda }}{\lambda }}|_{\overline{\lambda }_j}={\displaystyle \frac{1}{2+j}},`$ $`j=0,1,\mathrm{}j_{max},j_{max}=\left[{\displaystyle \frac{1}{\nu _\mu ^{}}}2\right]`$ (7) The terms $`k=2s(j=0)`$ must be studied separately, since in this case the point of stationary phase $`\overline{\lambda }=0`$ is also the lower limit of integration. As a result we obtain $`\delta N_{sh}={\displaystyle \underset{s=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^s}{(\pi s)^2}}\{{\displaystyle \frac{\mathrm{cos}(2\pi s2\nu _{\mu 0})}{\delta _\mu ^{(0)}}}{\displaystyle \frac{\lambda _\mu \mathrm{cos}(2\pi s\lambda _\mu )}{0.5\nu _\mu ^{}}}`$ $`{\displaystyle \underset{j=1}{\overset{j_{max}}{}}}{\displaystyle \frac{4\sqrt{s}j(1)^{js}}{(\delta _\mu ^{(j)})^{3/2}(2+j)^{5/2}}}\mathrm{cos}[2\pi s((2+j)\nu _{\mu \overline{\lambda }_j}+\overline{\lambda }_j){\displaystyle \frac{\pi }{4}}]\}`$ (8) Here $$\delta _\mu ^{(j)}\frac{^2\nu _{\mu \lambda }}{\lambda ^2}|_{\overline{\lambda }_j}.$$ Substituting expression (8) into Eq.(1) and integrating by parts to separate the terms which are of leading order in the semiclassicity parameter give a semiclassical formula for the shell correction to the binding energy of electrons in a cluster $`\delta E_{sh}={\displaystyle \frac{1}{2}}{\displaystyle \underset{s=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^s}{(\pi s)^3}}\{{\displaystyle \frac{\mathrm{sin}(2\pi s2\nu _{\mu 0})}{\delta _\mu ^{(0)}(2\nu _{\mu 0}/\mu )}}{\displaystyle \frac{\lambda _\mu \mathrm{sin}(2\pi s\lambda _\mu )}{\left(0.5\nu _\mu ^{}\right)(\lambda _\mu /\mu )}}`$ $`{\displaystyle \underset{j=1}{\overset{j_{max}}{}}}{\displaystyle \frac{4\sqrt{s}j(1)^{js}}{(\delta _\mu ^{(j)})^{3/2}(2+j)^{5/2}}}{\displaystyle \frac{\mathrm{sin}\left[2\pi s\left((2+j)\nu _{\mu \overline{\lambda }_j}+\overline{\lambda }_j\right)\frac{\pi }{4}\right]}{\left((2+j)\nu _{\mu \overline{\lambda }_j}+\overline{\lambda }_j\right)/\mu }}\},`$ (9) as an sum of contributions from supershells with quantum numbers $`n_j=Kn_r+l,K=2+j`$, the quantization on the Fermi level being substantial. In Eqs.(8) and (9) the symbol $`j`$ numerates a kind of the electronic orbit: $`j=0`$ corresponds to the linear pendulating orbit, the terms $`j1(K3)`$ are connected with the planar regular polygons, $`K`$ being the number of their vertices. The integer value $`s`$ is equal to a number of periods that the electronic trajectory $`(j,s)`$ includes, so the sum over $`s`$ is the trajectory length-expansion for a $`j`$-orbit (compare with the orbits $`(\lambda ,\nu )`$ in Ref. ). 4. The proposed simple method makes it possible for any spherical potential to determine the period and amplitude of oscillations associated with each supershell and to estimate their relative role in beating occurence. The results of such an analysis for a square well potential: $$U(r)=\{\genfrac{}{}{0pt}{}{2\epsilon _F,rR}{0,r>R},R=r_sN_e^{1/3},\epsilon _F=\frac{1}{2r_s^2}\left(\frac{9\pi }{4}\right)^{2/3},\mu =\epsilon _F$$ are displayed in Fig.1. Let’s note that even though we start with the expression (3) for a cluster with the closing ”$`l`$”-shells our results for $`N_e>100`$ check well with the results of the complete calculations in Ref. . Fig.1 shows, that the first period ($`N_e^{1/3}<7`$) of the beats is determined by terms with $`j=0,1,2`$, the term with $`j=0`$ contributing little. The account for term with $`j=3`$ ($`K=5`$) is needed to describe well the second period ($`N_e^{1/3}13`$). Add- ing a term with $`j=4`$ ($`K=6`$) is sufficient to describe the behavior of $`\delta E_{sh}`$ in the entire range under study. This means that the actual value $`j_{max}`$ is less that determined from Eq.(7) and corresponds to the filled states, for which $$\nu _{\mu \overline{\lambda }_j}\frac{1}{2}.$$ 5. For spherical potential we give the analytical expression for the oscillating part of the binding energy of electrons as an explicit sum of contributions from the leading periodic orbits (supershells). We show that a number of the leading orbits in the expansion depends strongly on the form of the potential and is determined by the value of derivative (6b)(see Eq.(7)). The more soft potentials will be dealt in detail in our extended publication where among other things small clusters and finite temperatures will be analyzed.
warning/0001/math0001037.html
ar5iv
text
# Majorant Series ## 1. Introduction What properties of a holomorphic function can be detected from the *moduli* of its Maclaurin series coefficients? To make this question precise, fix a holomorphic function $`f`$ such that $`f(z)=_kc_kz^k`$, and consider the class $``$ of all power series expansions $`_kb_kz^k`$ with the property that $`|b_k|=|c_k|`$ for all $`k`$. Which properties of $`f`$ are inherited by all functions in the class $``$? Here are some examples of properties that hold for all members of $``$ if they hold for one member of $``$. * The radius of convergence of the Maclaurin series equals $`1`$ (since Hadamard’s formula for the radius of convergence depends only on the moduli of the coefficients). * The function belongs to the Hardy space $`H^2`$ of the unit disk (since the Hardy space norm equals the square root of the sum of the squares of the moduli of the Maclaurin series coefficients). * The function is univalent in a neighborhood of the origin (since this property holds if and only if $`c_10`$.) On the other hand, here are some examples of properties that may hold for some members of $``$ but not for others. * The function is holomorphic and univalent in the whole unit disk. For instance, the linear fractional transformation $`z\frac{z}{1z}`$ is univalent in the unit disk, but changing the sign of the first term of the Maclaurin series yields the function $`z2z+\frac{z}{1z}`$, which maps the points $`0`$ and $`1/2`$ both to $`0`$. * The function has zeroes at prescribed locations. Indeed, composing a member of $``$ with the reflection $`zz`$ yields a new member of $``$ with zeroes at different locations. * The function is holomorphic and bounded in the unit disk. For instance, for almost every choice of plus and minus signs, the series $`_{k=1}^{\mathrm{}}\pm z^k/k`$ is continuous when $`|z|1`$ (see, for example, \[15, Chapter V, Theorem 8.34\]); but if all plus signs are taken, the series is unbounded in the unit disk. ## 2. Bohr’s theorem More generally, it might happen that if $`f`$ has a certain property, then every function in the associated class $``$ has some related property. For example, an old theorem of Harald Bohr implies that if $`f`$ is in the unit ball of $`H^{\mathrm{}}`$ of the unit disk, then every element of the class $``$ is in the unit ball of $`H^{\mathrm{}}`$ of the disk of radius $`1/3`$. ###### Theorem 1 (Bohr, 1914). If $`|_{k=0}^{\mathrm{}}c_kz^k|<1`$ when $`|z|<1`$, then $`_{k=0}^{\mathrm{}}|c_kz^k|<1`$ when $`|z|<1/3`$. Moreover, the radius $`1/3`$ is the best possible. This surprising theorem has been largely forgotten. The following proof, based on a classical inequality of Carathéodory, is due to Edmund Landau . ###### Lemma (Carathéodory’s inequality). If $`g`$ is a holomorphic function with positive real part in the unit disk, and $`g(z)=_{k=0}^{\mathrm{}}b_kz^k`$, then $`|b_k|2\mathrm{Re}b_0`$ when $`k1`$. ###### Proof of Bohr’s theorem. Let $`f(z)`$ denote the series $`_{k=0}^{\mathrm{}}c_kz^k`$. Let $`\phi `$ be an arbitrary real number, and set the function $`g`$ in Carathéodory’s inequality equal to $`1e^{i\phi }f`$ to deduce that $`|c_k|2\mathrm{Re}(1e^{i\phi }c_0)`$ when $`k1`$. Since $`\phi `$ is arbitrary, it follows that $`|c_k|2(1|c_0|)`$ when $`k1`$. If $`f`$ is a constant function, then the conclusion of the theorem is trivial. For nonconstant $`f`$, the preceding inequality shows that if $`|z|<1/3`$, then $$\underset{k=0}{\overset{\mathrm{}}{}}|c_kz^k|<|c_0|+2(1|c_0|)\underset{k=1}{\overset{\mathrm{}}{}}(1/3)^k=1.$$ To see that the radius $`1/3`$ in Bohr’s theorem is the best possible, consider the linear fractional transformation $`f_a`$ defined by $`f_a(z)={\displaystyle \frac{za}{1az}}`$ when $`0<a<1`$. Writing $`f_a(z)=_{k=0}^{\mathrm{}}c_k(a)z^k`$, one easily computes that $`_{k=0}^{\mathrm{}}|c_k(a)z^k|=2a+f_a(|z|)`$. Simple algebra shows that $`2a+f_a(|z|)>1`$ when $`|z|>1/(1+2a)`$. Since $`a`$ can approach $`1`$ from below, it follows that the radius $`1/3`$ in Bohr’s theorem cannot be increased. ∎ ###### Proof of Carathéodory’s inequality. A common proof of Carathéodory’s inequality (see \[5, page 41\], for example) uses the Herglotz representation for positive harmonic functions. I learned the following even simpler proof of the inequality from . By considering $`g(rz)`$ and letting $`r`$ increase toward $`1`$, we may assume without loss of generality that $`g`$ is holomorphic in a neighborhood of the closed unit disk. When $`k1`$, orthogonality implies that $$b_k=\frac{1}{2\pi }_0^{2\pi }e^{ik\theta }g(e^{i\theta })𝑑\theta =\frac{1}{2\pi }_0^{2\pi }e^{ik\theta }\left(g(e^{i\theta })+\overline{g(e^{i\theta })}\right)𝑑\theta .$$ Consequently, $$|b_k|\frac{1}{2\pi }_0^{2\pi }|2\mathrm{Re}g(e^{i\theta })|𝑑\theta =2\mathrm{Re}b_0,$$ where the last step follows because $`2\mathrm{Re}g`$ is a positive function with the mean-value property. ∎ A natural way to study the class $``$ associated to a holomorphic function $`f`$ is to single out a canonical representative of the class. If $`f(z)=_{k=0}^{\mathrm{}}c_kz^k`$, then the *majorant function* $`f`$ is defined by $`f(z)=_{k=0}^{\mathrm{}}|c_k|z^k`$. Evidently $`sup_{|z|<r}|f(z)|sup_{|z|<r}|f(z)|=f(r)`$, which justifies the name “majorant function”. Bohr’s theorem may be restated as the inequality $`sup_{|z|<r}|f(z)|sup_{|z|<3r}|f(z)|`$ for every $`r`$ for which the right-hand side makes sense. ## 3. Wintner’s problem In 1956, Aurel Wintner raised the following question related to Bohr’s theorem on majorant series: > For the class of holomorphic functions $`f`$ in the unit disk with modulus bounded by $`1`$, find the best upper bound on $`\underset{0<|z|<1}{inf}\left|{\displaystyle \frac{f(z)}{z}}\right|`$ or on $`\underset{0<r<1}{inf}{\displaystyle \frac{f(r)}{r}}`$. It follows by taking $`r`$ equal to $`1/3`$ that the value $`3`$ is an upper bound. Wintner claimed—incorrectly—that the bound $`3`$ cannot be improved. In a subsequent paper whose title quotes Wintner’s title, Günther Schlenstedt pointed out that Wintner made a blunder in high-school algebra, and that it is easy to see from the maximum principle that the best bound on $`\underset{0<|z|<1}{inf}\left|{\displaystyle \frac{f(z)}{z}}\right|`$ is actually $`1`$. It is amusing to note that the address printed at the end of Schlenstedt’s paper is “Carl-Friedrich-von-Siemens-Schule, Berlin”, so Schlenstedt was presumably either a student or a teacher at a German high school. Perhaps one should not be too critical of Wintner for this error, for he published 26 papers in 1956. Moreover, even the mistakes of a good mathematician are interesting. In my view, the really interesting mistake in Wintner’s paper is one that Schlenstedt did not address. Namely, Wintner claimed to study $`\underset{0<|z|<1}{inf}\left|{\displaystyle \frac{f(z)}{z}}\right|`$, but he actually studied $`\underset{0<r<1}{inf}{\displaystyle \frac{f(r)}{r}}`$, apparently thinking that these quantities admit the same optimal bound. In fact, the best bound on the second quantity is neither $`3`$ nor $`1`$, but $`2`$, as I shall now demonstrate. ###### Theorem 2. If $`f`$ is a holomorphic function such that $`|f(z)|<1`$ when $`|z|<1`$, then the majorant function $`f`$ satisfies the inequality $$\underset{0<r<1}{inf}\frac{f(r)}{r}2,$$ and the bound $`2`$ cannot be replaced by any smaller number. ###### Proof. I shall prove somewhat more than is stated in the theorem. Namely, I do not need the function $`f`$ to belong to the unit ball of $`H^{\mathrm{}}`$; all I will use is that $`f`$ belongs to the unit ball of the Hardy space $`H^2`$. In other words, a sufficient hypothesis on $`f`$ is that $$\underset{0<r<1}{sup}\frac{1}{2\pi }_0^{2\pi }|f(re^{i\theta })|^2𝑑\theta 1.$$ By Parseval’s theorem, this hypothesis implies that if $`f(z)`$ has the series expansion $`_{k=0}^{\mathrm{}}c_kz^k`$ when $`|z|<1`$, then $`_{k=0}^{\mathrm{}}|c_k|^21`$. Consequently, by the Cauchy-Schwarz inequality, $`{\displaystyle \frac{f(r)}{r}}`$ $`={\displaystyle \frac{|c_0|}{r}}+{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}|c_k|r^{k1}{\displaystyle \frac{|c_0|}{r}}+\left({\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}|c_k|^2\right)^{1/2}{\displaystyle \frac{1}{\sqrt{1r^2}}}`$ $`{\displaystyle \frac{|c_0|}{r}}+{\displaystyle \frac{\sqrt{1|c_0|^2}}{\sqrt{1r^2}}}.`$ Let $`r|c_0|`$ to deduce that $`\underset{0<r<1}{inf}{\displaystyle \frac{f(r)}{r}}2`$. For the particular function $`f`$ such that $`f(z)={\displaystyle \frac{z\frac{1}{\sqrt{2}}}{1\frac{z}{\sqrt{2}}}}`$, a straightforward computation shows that $`\underset{0<r<1}{inf}{\displaystyle \frac{f(r)}{r}}=2`$, so the bound $`2`$ cannot be improved. ∎ ## 4. A multi-dimensional analogue of Bohr’s theorem Some researchers in several complex variables feel that power series are boring, because the really interesting parts of multi-dimensional complex analysis are the parts that differ from the one-variable theory, while the elementary theory of power series superficially appears the same in all dimensions. My goal here is to exhibit some interesting and accessible problems about multi-variable power series in which the dependence on the dimension is the key issue. I shall use multi-index notation to write an $`n`$-variable power series as $`_\alpha c_\alpha z^\alpha `$, where $`\alpha `$ denotes an $`n`$-tuple $`(\alpha _1,\mathrm{},\alpha _n)`$ of non-negative integers, and $`z^\alpha `$ denotes the product $`z_1^{\alpha _1}\mathrm{}z_n^{\alpha _n}`$. It is standard (see \[12, pages 78–80\], for example) that such a power series converges in a logarithmically convex complete Reinhardt domain. To each such domain $`G`$ corresponds a *Bohr radius* $`K(G)`$: the largest $`r`$ such that whenever $`|_\alpha c_\alpha z^\alpha |1`$ for $`z`$ in $`G`$, it follows that $`_\alpha |c_\alpha z^\alpha |1`$ for $`z`$ in the scaled domain $`rG`$. The terminology “Bohr radius” is somewhat whimsical, for physicists consider the Bohr radius $`a_0`$ of the hydrogen atom to be a fundamental constant: its value is $`4\pi ϵ_0\mathrm{}^2/m_ee^2`$, or about $`0.529`$ Å. The physicists’ Bohr radius is named for Niels Bohr, a founder of the quantum theory and the 1922 recipient of the Nobel Prize for physics. Since Niels was the elder brother of Harald Bohr, adopting the term “Bohr radius” for mathematical purposes keeps the honor within the family. In contrast to the situation in one dimension, there is no higher-dimensional bounded domain $`G`$ whose Bohr radius $`K(G)`$ is known *exactly*. However, reasonably good bounds for the Bohr radius are known for special domains like the ball and the polydisc. More generally, let $`B_p^n`$ denote $`\{z^n:_{j=1}^n|z_j|^p<1\}`$, which is the unit ball of the complex Banach space $`\mathrm{}_p^n`$ whose norm is defined by $`z_{\mathrm{}_p^n}:=(_{j=1}^n|z_j|^p)^{1/p}`$. The ball $`B_2^n`$ corresponding to the case that $`p=2`$ is the usual Euclidean unit ball in $`^n`$. The ball $`B_{\mathrm{}}^n`$ is to be interpreted as the unit polydisc in $`^n`$. The following theorem quantifies the rate at which the Bohr radius of $`B_p^n`$ decays as the dimension $`n`$ increases. ###### Theorem 3. When $`n>1`$, the Bohr radius $`K(B_p^n)`$ of the $`\mathrm{}_p^n`$ unit ball in $`^n`$ admits the following bounds. * If $`1p2`$, then $$\frac{1}{3\sqrt[3]{e}}\left(\frac{1}{n}\right)^{1\frac{1}{p}}K(B_p^n)<3\left(\frac{\mathrm{log}n}{n}\right)^{1\frac{1}{p}}.$$ * If $`2p\mathrm{}`$, then $$\frac{1}{3}\sqrt{\frac{1}{n}}K(B_p^n)<2\sqrt{\frac{\mathrm{log}n}{n}}.$$ The theorem does not quite fix the sharp decay rate of the Bohr radius with the dimension, for the upper bounds contain a logarithmic factor not present in the lower bounds. This logarithmic factor, an artifact of the proof, presumably should not really be present. The numerical values of the constants in the theorem can be improved. For example, the constants in the two parts of the theorem could be made to agree when $`p=2`$. I have chosen to write simple constants for clarity, since the main point of the theorem is the dependence of the Bohr radius on the dimension $`n`$. The lower bound in the theorem is due to Lev Aizenberg when $`p=1`$. The lower bound when $`2p\mathrm{}`$ and the upper bound when $`p=\mathrm{}`$ are due jointly to Dmitry Khavinson and myself . The other parts of the theorem are new. ## 5. Proof of the estimates for the Bohr radius ### 5.1. The lower bound Lower bounds on the Bohr radius follow from upper bounds on the Maclaurin series coefficients of a bounded holomorphic function. In the writings of Bohr and Landau , one finds a useful trick employed by F. Wiener in the context of coefficient bounds for the one-dimensional Bohr theorem. ###### Lemma (after F. Wiener). Let $`G`$ be a complete Reinhardt domain, and let $``$ be the set of holomorphic functions on $`G`$ with modulus bounded by $`1`$. Fix a multi-index $`\alpha `$ other than $`(0,\mathrm{},0)`$, and suppose that the positive real number $`b`$ is an upper bound for the modulus of the derivative $`f^{(\alpha )}(0)`$ for every function $`f`$ in $``$. Then $`|f^{(\alpha )}(0)|(1|f(0)|^2)b`$ for every function $`f`$ in $``$. ###### Proof. Pick an index $`j`$ for which $`\alpha _j0`$, and let $`\omega _j`$ denote a primitive $`\alpha _j`$th root of unity. Consider the function obtained by averaging $`f`$: $$(z_1,\mathrm{},z_n)\frac{1}{\alpha _j}\underset{k=1}{\overset{\alpha _j}{}}f(z_1,\mathrm{},z_{j1},\omega _j^kz_j,z_{j+1},\mathrm{},z_n).$$ Iterate this averaging procedure for each non-zero component of the multi-index $`\alpha `$. The resulting function $`h`$ is still in the class $``$, and its Maclaurin series starts out $`f(0)+f^{(\alpha )}(0)z^\alpha /\alpha !+\mathrm{}`$, where $`\alpha !`$ denotes the product $`\alpha _1!\mathrm{}\alpha _n!`$. Composing $`h`$ with a linear fractional transformation of the unit disk that maps $`f(0)`$ to $`0`$ gives a new function in the class $``$ whose Maclaurin series begins $`f^{(\alpha )}(0)z^\alpha /(1|f(0)|^2)\alpha !+\mathrm{}`$, and the conclusion of the lemma follows. ∎ To prove the lower bound in Theorem 3 when $`1p2`$, I will follow the argument used by Aizenberg for the case when $`p=1`$. The method actually yields a bound when $`1p\mathrm{}`$, but the bound is interesting only when $`1p2`$. Let $`|\alpha |`$ denote the sum $`\alpha _1+\mathrm{}+\alpha _n`$, and let $`\alpha ^\alpha `$ denote the product $`\alpha _1^{\alpha _1}\mathrm{}\alpha _n^{\alpha _n}`$ (where $`0^0`$ is interpreted as $`1`$). A straightforward calculation using Lagrange multipliers shows that (1) $$sup\{|z^\alpha |:zB_p^n\}=\left(\frac{\alpha ^\alpha }{|\alpha |^{|\alpha |}}\right)^{1/p}.$$ By Cauchy’s estimates, it follows that if $`\left|_\alpha c_\alpha z^\alpha \right|1`$ when $`zB_p^n`$, then $`|c_\alpha |\left(|\alpha |^{|\alpha |}/\alpha ^\alpha \right)^{1/p}|\alpha |^{|\alpha |}/\alpha ^\alpha `$. Wiener’s lemma improves this estimate by a factor of $`(1|c_0|^2)`$. Hence, we have when $`k>0`$ that $$\underset{|\alpha |=k}{}|c_\alpha z^\alpha |(1|c_0|^2)\underset{|\alpha |=k}{}\frac{k^k}{\alpha ^\alpha }|z^\alpha |.$$ Multiplying and dividing by the multinomial coefficient $`\left(\genfrac{}{}{0pt}{}{k}{\alpha }\right)`$, which equals $`k!/\alpha !`$, observing that $`_{|\alpha |=k}\left(\genfrac{}{}{0pt}{}{k}{\alpha }\right)|z^\alpha |=z_{\mathrm{}_1^n}^k`$, and applying Hölder’s inequality to bound $`z_{\mathrm{}_1^n}`$ above by $`n^{1\frac{1}{p}}z_{\mathrm{}_p^n}`$, we deduce that $$\underset{\alpha }{}|c_\alpha z^\alpha ||c_0|+(1|c_0|^2)\underset{k=1}{\overset{\mathrm{}}{}}\frac{k^k}{k!}\left(n^{1\frac{1}{p}}z_{\mathrm{}_p^n}\right)^k.$$ Because the real quadratic function $`tt+(1t^2)/2`$ never exceeds $`1`$, it follows that if $`x`$ is the unique positive number such that (2) $$\underset{k=1}{\overset{\mathrm{}}{}}\frac{k^k}{k!}x^k=\frac{1}{2},$$ then the Bohr radius $`K(B_p^n)`$ is at least as big as $`x/n^{1\frac{1}{p}}`$. In combinatorics, one encounters the tree function $`T`$ (see, for example, \[7, p. 395\]), which satisfies the functional equation $`T(x)e^{T(x)}=x`$ and has the series expansion $`T(x)=_{k=1}^{\mathrm{}}\frac{k^{k1}}{k!}x^k`$. The equation (2) says that $`xT^{}(x)=1/2`$. Moreover, the functional equation implies that $`xT^{}(x)=T(x)/(1T(x))`$, so (2) yields $`T(x)/(1T(x))=1/2`$, or $`T(x)=1/3`$. The functional equation gives the solution $`x=1/(3\sqrt[3]{e})`$. One can also read off the solution of (2) from \[11, p. 707\], a source brought to my attention by Lev Aizenberg. This completes the proof of the lower bound in Theorem 3 when $`1p2`$. A different argument shows that a better lower bound for the Bohr radius is available when $`2<p`$. I reproduce the proof from . If $`|_\alpha c_\alpha z^\alpha |1`$ when $`zB_p^n`$, then applying the one-dimensional version of Wiener’s lemma to the one-variable function $`\zeta _\alpha c_\alpha \zeta ^{|\alpha |}z^\alpha `$ shows that $`|_{|\alpha |=k}c_\alpha z^\alpha |1|c_0|^2`$ when $`k1`$. Integrating the square of the left-hand side of this inequality over a torus and using the orthogonality of the monomials $`z^\alpha `$ shows that $`(_{|\alpha |=k}|c_\alpha |^2|w^\alpha |^2)^{1/2}1|c_0|^2`$ for every point $`w`$ in the unit ball of $`\mathrm{}_p^n`$. Now if $`z`$ lies in the $`\mathrm{}_p^n`$ ball of radius $`1/(3\sqrt{n})`$, then applying the Cauchy-Schwarz inequality with $`w`$ equal to $`3\sqrt{n}z`$ shows that $`{\displaystyle \underset{|\alpha |=k}{}}|c_\alpha z^\alpha |`$ $`\left({\displaystyle \underset{|\alpha |=k}{}}|c_\alpha |^2|w^\alpha |^2\right)^{1/2}\left({\displaystyle \underset{|\alpha |=k}{}}\left({\displaystyle \frac{1}{3\sqrt{n}}}\right)^{2k}\right)^{1/2}`$ $`(1|c_0|^2)/3^k,`$ since $`_{|\alpha |=k}1n^k`$. Consequently, we have for such points $`z`$ that $$\underset{\alpha }{}|c_\alpha z^\alpha ||c_0|+(1|c_0|^2)\underset{k=1}{\overset{\mathrm{}}{}}3^k=|c_0|+(1|c_0|^2)/21.$$ This proves the lower bound in Theorem 3 when $`2p\mathrm{}`$. ### 5.2. The upper bound The case of the polydisc, which corresponds to $`p`$ equal to $`\mathrm{}`$, is considered in . The idea is to use probabilistic methods to construct a homogeneous polynomial having relatively small supremum but relatively large majorant function. The argument in applies a result from on random trigonometric polynomials. To handle the case of general $`p`$, I will use a related random technique from . Unfortunately, the estimate I require is not stated explicitly in that paper: for their purposes, the authors did not need to keep track of the constants in the proof, and besides their argument has to be modified for the case of random tensors that are symmetric. Consequently, I repeat here the argument from with appropriate modifications and amplifications. Although the goal is to show the existence of a homogeneous polynomial of degree $`d`$ in $`n`$ variables that satisfies suitable estimates, technical considerations suggest proving a more general result for symmetric multi-linear functions mapping $`(^n)^d`$ into $``$. ###### Theorem 4. If $`1p\mathrm{}`$, and if $`n`$ and $`d`$ are integers larger than $`1`$, then there exists a symmetric multi-linear function $`F:(^n)^d`$ of the form $$F(Z_1,\mathrm{},Z_d)=\underset{J_1=1}{\overset{n}{}}\mathrm{}\underset{J_d=1}{\overset{n}{}}\pm Z_{1J_1}\mathrm{}Z_{dJ_d}$$ such that the supremum of $`|F(Z_1,\mathrm{},Z_d)|`$ when every $`n`$-vector $`Z_k`$ lies in the unit ball of $`\mathrm{}_p^n`$ is at most (3) $$\sqrt{32d\mathrm{log}(6d)}\times \{\begin{array}{cc}n^{\frac{1}{2}}(d!)^{1\frac{1}{p}},\hfill & \text{if }1p2;\hfill \\ n^{\frac{1}{2}+(\frac{1}{2}\frac{1}{p})d}(d!)^{\frac{1}{2}},\hfill & \text{if }2p\mathrm{}.\hfill \end{array}$$ In the theorem, the plus and minus signs are chosen independently for each set of indices $`J_1`$, …, $`J_d`$, with the proviso that the same choice is made for every permutation of a given set of indices. In other words, $`F`$ is the sum of the $`n^d`$ monomials of degree $`d`$ that can be formed from the components of the vectors $`Z_1`$, …, $`Z_d`$, with plus and minus signs chosen to enforce symmetry under permutations of the vectors. When all the vectors $`Z_k`$ are equal, the theorem provides a special homogeneous polynomial. Observe that the number of $`d`$-tuples $`J_1`$, …, $`J_d`$ that are permutations of a given list of indices is typically less than $`d!`$ because some indices in the list will be repeated. Indeed, if $`\alpha _k`$ denotes the number of times the integer $`k`$ appears in the list $`J_1`$, …, $`J_d`$, then the number of $`d`$-tuples that are permutations of the list is the multinomial coefficient $`\left(\genfrac{}{}{0pt}{}{d}{\alpha }\right)`$. Consequently, we have the following corollary of the theorem. ###### Corollary. If $`1p\mathrm{}`$, and if $`n`$ and $`d`$ are integers larger than $`1`$, then there exists a homogeneous polynomial of degree $`d`$ in the variable $`z`$ in $`^n`$ of the form (4) $$\underset{|\alpha |=d}{}\pm \left(\genfrac{}{}{0pt}{}{d}{\alpha }\right)z^\alpha $$ such that the supremum of the modulus of the polynomial when $`z`$ lies in the unit ball of $`\mathrm{}_p^n`$ is no greater than the above bound (3). To prove the upper bound in Theorem 3, consider the homogeneous polynomial of the corollary when $`z`$ is the vector $`(1,\mathrm{},1)`$ scaled by the factor $`K(B_p^n)/n^{1/p}`$. To write formulas that work for both ranges of $`p`$ simultaneously, I will write $`m(p)`$ for $`\mathrm{min}(p,2)`$ and $`M(p)`$ for $`\mathrm{max}(p,2)`$. The definition of the Bohr radius implies that (5) $$\underset{|\alpha |=d}{}\left(\genfrac{}{}{0pt}{}{d}{\alpha }\right)(K(B_p^n)/n^{1/p})^dn^{\frac{1}{2}+(\frac{1}{2}\frac{1}{M(p)})d}(d!)^{1\frac{1}{m(p)}}\sqrt{32d\mathrm{log}(6d)}.$$ Since $`_{|\alpha |=d}\left(\genfrac{}{}{0pt}{}{d}{\alpha }\right)=n^d`$, it follows that (6) $$K(B_p^n)\left(\frac{(d!)^{1/d}}{n}\right)^{1\frac{1}{m(p)}}\left(32nd\mathrm{log}(6d)\right)^{\frac{1}{2d}}.$$ Stirling’s formula implies that if $`d\mathrm{log}n`$, then the right-hand side of (6) yields the upper bound in Theorem 3 for sufficiently large $`n`$. Numerical calculations show that if $`d=2+\mathrm{log}n`$, then $`n`$ greater than $`148`$ is sufficiently large. On the other hand, the upper bound in Theorem 3 holds automatically for smaller values of the dimension $`n`$, because the multi-dimensional Bohr radius does not exceed the one-dimensional Bohr radius of $`1/3`$, while $`2\sqrt{\mathrm{log}n}/\sqrt{n}>1/3`$ when $`1<n<189`$. Thus the upper bound in Theorem 3 is a consequence of the corollary of Theorem 4. ###### Proof of Theorem 4. The proof consists of a probabilistic estimate and a covering argument. Three parameters are fixed throughout the proof: the dimension $`n`$, the degree $`d`$, and the exponent $`p`$. I will use the following notation: $`Z`$ denotes a $`d`$-tuple $`Z_1`$, …, $`Z_d`$ of $`n`$-vectors, $`J`$ denotes a $`d`$-tuple $`J_1`$, …, $`J_d`$ of integers between $`1`$ and $`n`$, sums and products run over all such $`d`$-tuples of integers, a prime on a sum or product means that it is restricted to $`d`$-tuples of integers that are arranged in non-decreasing order, and $`JK`$ indicates that the $`d`$-tuples $`J`$ and $`K`$ are permutations of each other. The symbol $`Z_J`$ is shorthand for the monomial $`Z_{1J_1}\mathrm{}Z_{dJ_d}`$. Thus, the function in the statement of Theorem 4 can be written in the form $$F(Z)=\underset{K}{^{}}\left(\pm \underset{JK}{}Z_J\right),$$ and in this expression, all of the plus and minus signs are independent of each other. To begin the probabilistic argument, fix a point $`Z`$ in $`(^n)^d`$ such that every $`Z_k`$ lies in the $`\mathrm{}_p^n`$ unit ball $`B_p^n`$. For each $`d`$-tuple $`K`$ in non-decreasing order, choose a different Rademacher function $`r_K`$. (Recall that the Rademacher functions are independent functions each taking the values $`+1`$ and $`1`$ with probability $`1/2`$.) Consider the random sum (7) $$F(t,Z):=\underset{K}{^{}}\left(r_K(t)\underset{JK}{}Z_J\right),$$ where $`t`$ lies in the interval $`[0,1]`$. The immediate goal is to make an upper estimate on the probability that this sum has large modulus. Let $`\lambda `$ be an arbitrary positive real number; in (13) below, I will specify a value for $`\lambda `$ in terms of $`n`$, $`d`$, and $`p`$. Invoking the independence of the Rademacher functions, we can compute the expectation (that is, the integral with respect to $`t`$) of the exponential of the real part of $`\lambda F(t,Z)`$ by computing the product over non-decreasing $`d`$-tuples $`K`$ of the expectations of $`\mathrm{exp}(\lambda r_K(t)_{JK}\mathrm{Re}Z_J)`$: namely, (8) $$\underset{K}{^{}}\mathrm{cosh}\left(\lambda \underset{JK}{}\mathrm{Re}Z_J\right).$$ In view of the inequality $`\mathrm{cosh}x\mathrm{exp}(x^2/2)`$, this expectation is bounded above by (9) $$\mathrm{exp}\left(\frac{1}{2}\lambda ^2\underset{K}{^{}}\left(\underset{JK}{}\mathrm{Re}Z_J\right)^2\right).$$ I will again write $`m(p)`$ for $`\mathrm{min}(p,2)`$ and $`M(p)`$ for $`\mathrm{max}(p,2)`$ in order to analyze the cases $`1p2`$ and $`2p\mathrm{}`$ simultaneously. Hölder’s inequality implies that $$\left(\underset{JK}{}\mathrm{Re}Z_J\right)^2(d!)^{2(11/m(p))}\left(\underset{JK}{}|Z_J|^{m(p)}\right)^{2/m(p)}.$$ The exponent $`2/m(p)`$ is equal to $`1`$ when $`2p\mathrm{}`$, and replacing it by $`1`$ when $`1p<2`$ can only increase the right-hand side when each $`n`$-vector $`Z_k`$ lies in $`B_p^n`$. Consequently, (9) is bounded above by (10) $$\mathrm{exp}\left(\frac{1}{2}\lambda ^2(d!)^{2(11/m(p))}Z_1_{\mathrm{}_{m(p)}^n}^{m(p)}\mathrm{}Z_d_{\mathrm{}_{m(p)}^n}^{m(p)}\right).$$ Since $`Z_kB_p^n`$, Hölder’s inequality implies that $`Z_k_{\mathrm{}_{m(p)}^n}^{m(p)}`$ is bounded above by $`n^{1\frac{2}{M(p)}}`$. Therefore (10) is bounded above by (11) $$\mathrm{exp}\left(\frac{1}{2}\lambda ^2(d!)^{2(11/m(p))}n^{(12/M(p))d}\right).$$ Let $`R`$ be an arbitrary positive real number; in (13) below, I will specify a value for $`R`$ in terms of $`n`$, $`d`$, and $`p`$. By Chebyshev’s inequality, the upper bound (11) on the expectation of $`\mathrm{exp}(\lambda \mathrm{Re}F(t,Z))`$ implies that the measure of the set of points $`t`$ for which $`\mathrm{Re}F(t,Z)`$ exceeds $`R`$ is at most (12) $$\mathrm{exp}\left(R\lambda +\frac{1}{2}\lambda ^2(d!)^{2(11/m(p))}n^{(12/M(p))d}\right).$$ Symmetric reasoning gives the same estimate for the probability that $`\mathrm{Re}F(t,Z)`$ is less than $`R`$, so the probability that $`|\mathrm{Re}F(t,Z)|`$ exceeds $`R`$ is at most $`2`$ times (12). The same argument applies to the imaginary part of $`F(t,Z)`$. Consequently, the probability that $`|F(t,Z)|`$ exceeds $`\sqrt{2}R`$ is at most $`4`$ times (12). This probabilistic estimate holds for an arbitrary but fixed $`Z`$. The second part of the proof is a covering argument to produce an estimate that is uniform in $`Z`$. The following lemma is well known (see \[10, p. 7\], for example). The exponent differs from the statement in because I am considering the complex $`\mathrm{}_p^n`$ space. ###### Lemma. If $`ϵ`$ is a positive real number, then the unit ball of the complex space $`\mathrm{}_p^n`$ can be covered by a collection of open $`\mathrm{}_p^n`$ balls of radius $`ϵ`$, the number of balls in the collection not exceeding $`(1+2ϵ^1)^{2n}`$ and the centers of the balls lying in the closed unit ball of $`\mathrm{}_p^n`$. ###### Proof. Place arbitrarily in the open $`\mathrm{}_p^n`$ ball of radius $`(1+\frac{ϵ}{2})`$ a collection of disjoint open $`\mathrm{}_p^n`$ balls of radius $`ϵ/2`$. Since the Euclidean volume of a ball in $`\mathrm{}_p^n`$ scales like the $`(2n)`$th power of its radius, an obvious volume comparison shows that the number of disjoint balls cannot exceed the $`(2n)`$th power of the ratio of radii $`(1+\frac{ϵ}{2})/(ϵ/2)`$: namely, $`(1+2ϵ^1)^{2n}`$. If the collection of disjoint balls is made maximal, then every point of the closed $`\mathrm{}_p^n`$ unit ball must lie within $`ϵ/2`$ of some point of one of the balls in the collection. Consequently, the balls with the same centers but with radius $`ϵ`$ must cover the closed unit ball. ∎ Next we need a simple Lipschitz estimate for $`F(t,Z)`$. Suppose that $`Z`$ and $`W`$ are points of $`(^n)^d`$ such that all of the component $`n`$-vectors $`Z_1`$, …, $`Z_d`$ and $`W_1`$, …, $`W_d`$ lie in $`B_p^n`$, and $`Z_kW_k_{\mathrm{}_p^n}ϵ`$ for every $`k`$. The multi-linearity of $`F`$ implies that $$\begin{array}{c}F(t,Z_1,\mathrm{},Z_d)=F(t,Z_1W_1,Z_2,\mathrm{},Z_d)\hfill \\ \hfill +F(t,W_1,Z_2W_2,Z_3,\mathrm{},Z_d)\\ \hfill +F(t,W_1,W_2,Z_3W_3,Z_4,\mathrm{},Z_d)+\mathrm{}\\ \hfill +F(t,W_1,W_2,\mathrm{},W_{d1},Z_dW_d)+F(t,W_1,\mathrm{},W_d).\end{array}$$ Consequently, $`|F(t,Z)F(t,W)|`$ is at most $`d`$ times $`ϵ`$ times the supremum of $`|F(t,)|`$ over $`(B_p^n)^d`$. Taking $`ϵ`$ to be $`1/(2d)`$, we see by the lemma that there is a collection of at most $`(1+4d)^{2nd}`$ points of $`(B_p^n)^d`$ such that the supremum of $`|F(t,)|`$ over $`(B_p^n)^d`$ is no more than twice the supremum of $`|F(t,)|`$ over this finite collection of points. Applying the preceding probabilistic estimate to each member of the finite collection of points shows that the probability that the supremum of $`|F(t,)|`$ over $`(B_p^n)^d`$ exceeds $`2\sqrt{2}R`$ is at most $$4(1+4d)^{2nd}\mathrm{exp}\left(R\lambda +\frac{1}{2}\lambda ^2(d!)^{2(11/m(p))}n^{(12/M(p))d}\right).$$ Now take the following values for the parameters $`R`$ and $`\lambda `$: (13) $`R`$ $`:=\left(2(d!)^{2(11/m(p))}n^{(12/M(p))d}\mathrm{log}(8(1+4d)^{2nd})\right)^{1/2},`$ $`\lambda `$ $`:={\displaystyle \frac{R}{(d!)^{2(11/m(p))}n^{(12/M(p))d}}}.`$ With these choices, we find that the probability that the supremum of $`|F(t,)|`$ over $`(B_p^n)^d`$ exceeds $`2\sqrt{2}R`$ is at most $`1/2`$, so we are sure that there exists a particular value of $`t`$ such that the supremum of $`|F(t,)|`$ over $`(B_p^n)^d`$ is no more than (14) $$\left(16(d!)^{2(11/m(p))}n^{(12/M(p))d}\mathrm{log}(8(1+4d)^{2nd})\right)^{1/2}.$$ The values of the Rademacher functions at this particular value of $`t`$ produce the pattern of plus and minus signs indicated in the statement of Theorem 4. Moreover, $`8(1+4d)^{2nd}<(6d)^{2nd}`$ when $`n`$ and $`d`$ are both at least $`2`$, so the upper bound (14) is even smaller than the bound stated in Theorem 4. ∎ ## 6. Another analogue of Bohr’s theorem In , Aizenberg introduced a second Bohr radius $`B(G)`$: namely, the largest $`r`$ such that whenever $`|_\alpha c_\alpha z^\alpha |1`$ for $`zG`$, it follows that $`_\alpha sup_{zrG}|c_\alpha z^\alpha |1`$. Clearly $`B(G)K(G)`$, because the supremum of a sum is no bigger than the sum of the suprema. Equality holds for polydiscs, because points on the torus have the property that they maximize $`|z^\alpha |`$ for all indices $`\alpha `$ simultaneously. The following statement is an analogue of Theorem 3 for the second Bohr radius. ###### Theorem 5. When $`n>1`$, the second Bohr radius $`B(B_p^n)`$ of the $`\mathrm{}_p^n`$ unit ball in $`^n`$ admits the following bounds. * If $`1p2`$, then $$\frac{1}{3}\frac{1}{n}<1\left(\frac{2}{3}\right)^{\frac{1}{n}}B(B_p^n)<4\frac{\mathrm{log}n}{n}.$$ * If $`2p\mathrm{}`$, then $$\frac{1}{3}\left(\frac{1}{n}\right)^{\frac{1}{2}+\frac{1}{p}}B(B_p^n)<4\left(\frac{\mathrm{log}n}{n}\right)^{\frac{1}{2}+\frac{1}{p}}.$$ ###### Proof. The lower bound is due to Aizenberg when $`1p2`$. Namely, Cauchy’s estimates imply that if $`|_\alpha c_\alpha z^\alpha |1`$ in $`B_p^n`$, then $`|c_\alpha |1/sup\{|z^\alpha |:zB_p^n\}`$, and Wiener’s lemma provides an extra factor of $`1|c_0|^2`$. Therefore $$\underset{\alpha }{}sup\{|c_\alpha z^\alpha |:z_{\mathrm{}_p^n}r\}|c_0|+(1|c_0|^2)\underset{\alpha 0}{}r^{|\alpha |},$$ and this quantity will be bounded above by $`1`$ if $`_{\alpha 0}r^{|\alpha |}1/2`$. Since $`_{\alpha 0}r^{|\alpha |}=1+1/(1r)^n`$, we deduce that the second Bohr radius is no smaller than $`1(2/3)^{1/n}`$. It is easy to see that this quantity exceeds $`1/(3n)`$ when $`n>1`$, because the function $`nn(1(2/3)^{1/n})`$ is increasing for positive $`n`$ and takes the value $`1/3`$ when $`n=1`$. When $`p=\mathrm{}`$, the lower bound is contained in Theorem 3, since the two Bohr radii are the same for the polydisc. When $`2<p<\mathrm{}`$, the lower bound follows by observing that an $`\mathrm{}_{\mathrm{}}^n`$ ball of radius $`1/n^{1/p}`$ nests inside $`B_p^n`$, and so $`B(B_p^n)B(B_{\mathrm{}}^n)/n^{1/p}`$. To prove the upper bound in Theorem 5, apply the definition of the second Bohr radius to the homogeneous polynomial of the corollary of Theorem 4. From (1), it follows that $$\underset{|\alpha |=d}{}\left(\genfrac{}{}{0pt}{}{d}{\alpha }\right)\left(\frac{\alpha ^\alpha }{d^d}\right)^{\frac{1}{p}}B(B_p^n)^dn^{\frac{1}{2}+(\frac{1}{2}\frac{1}{M(p)})d}(d!)^{1\frac{1}{m(p)}}\sqrt{32d\mathrm{log}(6d)}.$$ Since $`\alpha ^\alpha 1`$, and $`_{|\alpha |=d}\left(\genfrac{}{}{0pt}{}{d}{\alpha }\right)=n^d`$, we deduce that $$B(B_p^n)\frac{1}{n^{\frac{1}{2}+\frac{1}{M(p)}}}d^{\frac{1}{p}}(d!)^{\frac{1}{d}(1\frac{1}{m(p)})}(32nd\mathrm{log}(6d))^{\frac{1}{2d}}.$$ When $`d=\mathrm{log}n`$, the right-hand side asymptotically gives the desired power of $`(\mathrm{log}n)/n`$, and numerical calculations show that the right-hand side is smaller than the upper bound stated in Theorem 5 when $`n>28`$. On the other hand, the upper bound in the theorem holds automatically for smaller values of $`n`$, because $`(4\mathrm{log}n)/n>1/3`$ when $`1<n<46`$. ∎ ## 7. Directions for future research My goal in this article has been not only to demonstrate some interesting results and techniques in the theory of multi-variable power series, but also to suggest some avenues for further investigation. Here are some concrete problems. 1. Construct examples to show that the logarithmic term is not needed in the upper bound in Theorems 3 and 5. 2. Find the *exact* value of the Bohr radius for the polydisc in dimension $`2`$. 3. Estimate the Bohr radius for domains of the form $$|z_1|^{p_1}+\mathrm{}+|z_n|^{p_n}<1.$$ 4. Generalize Wintner’s problem to higher dimensions.
warning/0001/astro-ph0001439.html
ar5iv
text
# Hydrodynamic Simulations of the Bardeen–Petterson Effect ## 1 Introduction Accretion discs may occur around stellar mass black holes in X–ray sources such as Cygnus X–1, or around very massive black holes in active galaxies. The Lense-Thirring effect, or dragging of inertial frames, has an important influence on the structure of an accretion disc around a rotating black hole in the situation where the equatorial plane of the outer disc far from the black hole is misaligned with the symmetry plane of the black hole. Bardeen & Petterson (1975) showed that a viscous disc would be expected to relax to a form in which the inner regions become aligned with the equatorial plane of the black hole out to a transition radius, beyond which the disc remains aligned with the outer disc. This phenomenon may be important for understanding absorption and reprocessing of X–rays in X–ray sources, and also for providing stability for jet directions in radio sources (Rees 1978). In their calculation Bardeen & Petterson (1975) assumed that the warps in the misaligned disc diffused on a viscous time scale. The viscosity coefficient was taken to be of the the same magnitude as that producing the mass inflow, typically modelled using the standard Shakura & Sunyaev (1973) $`\alpha `$ prescription. However, Papaloizou and Pringle (1983) showed that for an isotropic viscosity, warp diffusion occurs at a rate faster by $`1/(2\alpha ^2),`$ provided $`\alpha >H/r,`$ much reducing the transition radius. For $`\alpha <H/r`$ it is expected that the dynamics is controlled by the wave modes present in the disc. The location of the transition radius has been investigated in the diffusive linear regime ($`\alpha >H/r`$) (Kumar & Pringle 1985), and in a model including viscous effects but neglecting pressure (Scheuer & Feiler 1996). The Bardeen–Petterson effect has also been studied in the linear regime, but with $`\alpha <H/r`$, by Ivanov & Illarianov (1997), where they also considered the effects of some additional post Newtonian corrections. In this paper we study the Bardeen-Petterson effect using SPH simulations. We focus primarily on the parameter regime where the warp dynamics are controlled by bending wave propagation, although we also examine models in which warps are subject to diffusion and are not transmitted by waves. The properties of the wave and diffusive communication that occurs in the discs modelled here are discussed in detail in a recent paper (Nelson & Papaloizou 1999) hereafter referred to as paper I. In that paper we compare SPH simulations of small amplitude disturbances to results obtained using the linearized fluid equations that are solved using a standard finite difference code, though we comment more generally that by using SPH calculations we are able to consider the fully non linear regime. The calculations presented in paper I are used to calibrate the code viscosity. The warping of a gaseous disc generates vertical shearing motions (i.e. horizontal motions that are a function of the vertical position in the disc). The viscosity that acts on this vertical shear may not be the same as that which acts on the usual horizontal shear associated with a planar, differentially rotating accretion disc. We therefore denote the dimensionless viscosity parameter acting on the vertical shear as $`\alpha _1`$, and that acting on the horizontal shear through the $`r`$$`\varphi `$ component of the stress tensor as $`\alpha `$. The test calculations presented in paper I indicate that the thicker discs that we consider, with midplane Mach numbers of $`12`$, (where $`H/r^1`$) have $`\alpha _10.04`$ so that bending waves are expected to propagate in these discs. We note that calibration runs indicate that $`\alpha 0.02`$$`0.03`$ for disc models of this thickness obtained using SPH (e.g. Larwood et al. 1996; Bryden et al. 1999). It is found that the thinner discs with $`30`$ have values of $`\alpha _10.1`$ and similar values for $`\alpha `$ so that in these cases warping disturbances are expected to evolve diffusively. We study the Bardeen–Petterson effect for a variety of disc Mach numbers, and find that the transition radius is always much smaller than that obtained by Bardeen & Petterson (1975). The rate of black hole – disc alignment is, however, found to be consistent with the ideas of Rees (1978), and the later work of Scheuer & Feiler (1996) when taking the viscous diffusion coefficients acting in and out of the plane to be equal. A discussion of this result is presented in section 8.5. The plan of the paper is as follows. In section 2 we give the basic equations. In section 3 we describe how Lense-Thirring precession acting on a misaligned accretion disc in the neighbourhood of a rotating black hole leads to the Bardeen-Petterson effect, where the inner regions of the disc become aligned with the equatorial plane of the black hole out to the transition radius. The relaxation of an accretion disc for which the angular momentum vector is misaligned with the spin vector of a rotating black hole is dependent on the dynamics of warps or bending waves (see paper I and references therein). Linear bending waves are expected to propagate if the Shakura & Sunyaev (1973) $`\alpha `$ viscosity appropriate to dissipation of vertical shear, $`\alpha _1<H/r.`$ For larger values of $`\alpha _1,`$ the evolution is diffusive. In section 4 we consider the relaxation of a disc subject to gravomagnetic forces when there is a small misalignment. In section 5 we discuss estimates of the Bardeen-Petterson transition radius based on equating the Lense-Thirring precessional shear rate to the viscous diffusion rate. Here we note that it is the diffusion rate of warps that is important and this occurs at a rate faster by $`1/(2\alpha ^2)`$ than the standard viscous rate (Papaloizou & Pringle 1983) if an isotropic standard Shakura & Sunyaev (1973) $`\alpha `$ viscosity is used so $`\alpha =\alpha _1`$. The transition radius then decreases as $`\alpha `$ decreases until $`\alpha =H/r.`$ Then the warp diffusion rate becomes the same as the propagation rate of bending waves and the transition radius cannot decrease further as $`\alpha `$ is reduced. In section 6 we describe the implementation of SPH used to perform the simulations, and in section 7 we describe the calculations of the Bardeen-Petterson effect. In section 8 we present the results of nonlinear simulations of a disc orbiting around a rotating black hole with Kerr parameter $`a=1`$ and which has its spin vector initially misaligned with that of the disc. We use the lowest order post Newtonian approximation in our simulations. We present simulations for Mach numbers ranging between 5 and 30 and inclinations of the outer disc plane of 10 and 30 degrees. For the effective viscosity $`\alpha _1`$ in the disc, in the lower Mach number cases, warps are governed by bending waves, while for the higher Mach number case of 30 they are diffusive. The central portion of the disc becomes aligned with the equatorial plane of the hole out to a transition radius which was found to be much smaller than that given by Bardeen & Petterson (1975), ranging between 15 and 30 gravitational radii. Models were considered both at high and low inclination. In the nonlinear high inclination case the transition between the aligned and non-aligned disc was more abrupt than in the low inclination case, indicative of a tendency for the outer part of the disc to become disconnected from the inner part. A calculation is presented for a disc of substantially larger outer radius than the other disc models we consider. We do not expect that the location of the transition radius will be affected by the location of the outer disc radius provided that it is sufficiently far away from the transition zone. In the majority of cases that we consider, the outer disc radius ($`r2`$) is a factor of between three and four times larger than the final transition radius ($`r0.7`$). We find that a disc that is more than a factor of three larger again (i.e. $`r7`$) has a transition radius that is unchanged, indicating that the smaller disc models are sufficiently large to not affect the location of the transition radius. We go on to investigate the time scale required for black-hole disc alignment. Our results are consistent with the ideas of Rees (1978), and the later work of Scheuer & Feiler (1996) when the viscous diffusion coefficients acting in and out of the plane are taken to be equal, even though the transition radius is much smaller in our case. We also find that the effects of Einstein precession are to increase the alignment rate, as expected from a linear analysis. Finally in section 9 we summarize our results and conclusions. ## 2 Equations of motion In order to describe a compressible fluid, we adopt continuity and momentum equations in the form $$\frac{d\rho }{dt}+\rho .𝐯=0$$ (1) and $$\frac{d𝐯}{dt}=\frac{1}{\rho }P+𝐯\times 𝐡\mathrm{\Phi }+𝐒_{visc}$$ (2) where $$\frac{d}{dt}=\frac{}{t}+𝐯.$$ (3) denotes the convective derivative, $`\rho `$ is the density, $`𝐯`$ is the velocity, $`P`$ is the pressure, $`\mathrm{\Phi }`$ is the gravitational potential and $`𝐒_{visc}`$ represents the viscous force per unit mass. The above equations are the standard Newtonian equations of fluid dynamics apart from the addition of the $`𝐯\times 𝐡`$ term to equation (2) which gives the gravomagnetic force per unit mass experienced by the fluid in the vicinity of a rotating black hole to lowest post Newtonian order (e.g. Blandford 1996). Here $`𝐡`$ is defined by the expression $$𝐡=\frac{2𝐒}{R^3}\frac{6(𝐒.𝐫)𝐫}{R^5}$$ (4) and $$𝐒=\frac{G𝐉}{c^2}.$$ (5) Here $`R`$ is the distance to the central mass and $`𝐫`$ denotes the coordinate vector measured from there. It is the gravomagnetic force that gives rise to Lense–Thirring precession which in turn leads to the alignment of the rotation axis of a viscous disc with that of the black hole. The spin angular momentum of a Kerr black hole is given by the equation $$𝐉=\frac{aGM^2}{c}\widehat{𝐤},$$ (6) where $`\widehat{𝐤}`$ denotes the unit vector in the $`z`$ direction associated with cylindrical coordinates $`(r,\varphi ,z)`$ based on the central mass. Thus $`R=\sqrt{(r^2+z^2)}.`$ For a maximally rotating hole, the Kerr parameter $`a=1`$, with $`M`$ being the mass of the black hole, and $`c`$ the speed of light. Because the effects of interest simulated here take place many gravitational radii from the black hole, relativistic effects are included only in a post Newtonian approximation. In the calculations described in subsequent sections, different prescriptions for the gravitational force were used. In some cases, a softened Keplerian potential was used such that the gravitational force per unit mass is $$\mathrm{\Phi }=\frac{GM}{(R^2+b^2)^{3/2}}𝐫$$ (7) where $`b`$ is the gravitational softening parameter. This was adopted in order to prevent numerical divergences as the disc material approaches the central object. In other cases, the effects of Einstein precession were included, and the central object was treated as a uniform density sphere of finite size to prevent the gravitational force from becoming infinite as $`r0`$. In this case the gravitational force per unit mass outside the sphere was taken to be $$\mathrm{\Phi }=\frac{GM}{R^3}\left(1+\frac{6R_+}{R}\right)𝐫$$ (8) where $`R_+=GM/c^2`$, denotes the gravitational radius. This force differs from the pseudo–Newtonian expression, often adopted in calculations of this kind, that gives the correct radius of the last stable circular orbit (e.g. Paczynski & Wiita 1980), but instead it gives the correct orbital apsidal precession frequency at large distances from the black hole. The calculations that included softening of the gravitational force were performed before we arrived at the idea of treating the central object as a uniform density sphere as a means of preventing numerical divergences. Although there is no astrophysical phenomenon that has a direct analogy with the use of a softening parameter as described in equation (7), it plays a relatively small role in our calculations of the Bardeen–Petterson transition zone. The fact that there is an associated retrograde precession changes the behaviour slightly, and this acts as a convenient way of illustrating the effect of the disc rotation profile on the behavious of warps in accretion discs by comparing the results obtained with softening with those obtained when Einstein precession is included. These equations are supplemented by an equation of state, which in this case is taken to be a polytrope of index $`\gamma =5/3`$: $$P=K\rho ^\gamma .$$ (9) The sound speed is $`c_s=\sqrt{dP/d\rho }.`$ Energy dissipated through the action of artificial viscosity is simply allowed to leave the system, so that a barotropic equation of state is assumed throughout. ## 3 Precession Frequencies and the Bardeen–Petterson Effect The effect of Lense–Thirring precession is to cause the plane of an orbit inclined to the $`(x,y)`$ plane to precess about the angular momentum $`(z)`$ axis. If a viscous disc with negligible inertia is set up with its midplane initially misaligned with the $`(x,y)`$ plane, differential Lense–Thirring precession will lead to the formation of a warped disc. The (prograde) nodal precession frequency that arises due to the inclusion of the gravomagnetic force term in the equation of motion (2) is given by $$\omega _z=\frac{2S}{R^3},$$ (10) where $`S=|𝐒|.`$ Note that this decreases with radius resulting in differential precession. The Bardeen–Petterson effect is caused by the combined effects of Lense–Thirring precession and internal disc viscosity, which, assuming the disc has negligible inertia, acts to align the midplane of the inner region of the accretion disc with the $`(x,y)`$ plane. This is because the damping produced by viscosity, acting on the small scale structure (twisting up) produced by strong differential precession, causes the inner disc to settle into the plane of the black hole equator. However, the outer parts of the disc remain in their original plane because the Lense–Thirring precession rate drops off sharply as $`r`$ increases, such that the internal pressure and viscous stresses acting in the disc are able to limit the effects of differential precession. The radius of transition between the two disc midplanes is expected to occur approximately where the rate at which the disc is twisted up by differential precession is balanced by the rate at which disc warps are diffused or propagated away. The rate at which the disc is twisted up is given by $`1/\tau _{dp}=|R(d\omega _z/dR)|`$, where $`\tau _{dp}`$ is the differential precession time scale. The rate at which warps are propagated or diffused is discussed in the next section. Important dynamical effects can also occur because precession of orbital apsidal lines occurs when the central potential is non Keplerian. The two cases considered here are associated with the softening of the central potential and taking into account Einstein precession. The rate of (retrograde) apsidal precession of a near circular orbit in the softened gravitational potential (7) is given by $$\omega _{ps}=\frac{3}{2}\mathrm{\Omega }\frac{b^2}{R^2},$$ (11) where $`\mathrm{\Omega }`$ is the circular orbit rotation frequency. It can be seen that the importance of this apsidal precession drops off quickly as $`R>>b`$. The rate of (prograde) Einstein apsidal precession associated with the potential (8) is given by $$\omega _{pE}=3\mathrm{\Omega }\frac{R_+}{R}.$$ (12) Comparison of these two rates of precession shows that the effects of Einstein precession remain relatively more important further away from the black hole than the effects of gravitational softening. Our numerical experiments show that use of (8) rather than just a softened potential has a noticeable effect on the results, in particular on the way in which the disc undergoes the transition between the plane occupied by the inner disc and that occupied by the outer disc. This is related to the effects of a non Keplerian potential on the propagation and diffusion of warps. ## 4 Misalignment and The Bardeen–Petterson Transition ### 4.1 Small Misalignments at Large Radii The radius at which the disc midplane undergoes a transition between alignment with the $`(x,y)`$ plane and its initial inclination at large distance can be simply estimated in the situation when the latter inclination is small. To do this we consider small amplitude bending disturbances of the disc out of the $`(x,y)`$ plane. These are required to have an inclination that is asymptotically constant and equal to its value at large radii. The theory and propagation of small amplitude bending waves has been discussed in paper I and we may apply the analysis presented there. To make this application we first work on the $`𝐯\times 𝐡`$ term in equation (2). Using (4) this may be written $$𝐯\times 𝐡=\frac{2S(𝐯\times \widehat{𝐤})}{R^3}+\frac{6Sz(𝐫\times 𝐯)}{R^5}.$$ (13) For linear disturbances at large radii, we may regard $`S`$ as a first order quantity and accordingly replace the velocity by the unperturbed Keplerian value in (13). Thus in cylindrical coordinates $`𝐯=(0,v_\varphi ,0)`$ with $`v_\varphi =\sqrt{GM/r}.`$ Additionally as we are interested in a thin disc \[neglecting corrections of order $`(H/r)^2`$\] we replace $`R`$ by the cylindrical radius $`r.`$ Then (13) becomes $$𝐯\times 𝐡=\frac{2S\sqrt{GM}\widehat{𝐫}}{r^{7/2}}+\frac{6Sz\sqrt{GM}\widehat{𝐤}}{r^{9/2}}.$$ (14) This can be considered \[again neglecting corrections of order $`(H/r)^2`$\] to arise from an effective potential such that $$𝐯\times 𝐡=\mathrm{\Phi }_{gr},$$ (15) with $$\mathrm{\Phi }_{gr}=\frac{4S\sqrt{GM}}{5r^{5/2}}\frac{3Sz^2\sqrt{GM}}{r^{9/2}}.$$ (16) If we now add this potential to the already existing one we obtain the total potential governing the motion of the fluid $$\mathrm{\Phi }_{Tot}\mathrm{\Phi }+\mathrm{\Phi }_{gr}.$$ (17) The governing equations are now of the same general form as those describing bending disturbances considered in paper I. We shall consider the low frequency limit incorporating viscosity as in that paper. Accordingly we adopt equation (19) of that paper. This applies to disturbances for which the $`\varphi `$ and $`t`$ dependence is through a factor $`e^{i(\varphi +\sigma t)}`$, $`\sigma `$ being the mode frequency and is $$4g\mathrm{\Omega }(\omega _z+\sigma )+\frac{d}{dr}\left(\frac{\mu \mathrm{\Omega }}{\sigma i\alpha _1\mathrm{\Omega }+\mathrm{\Omega }\kappa }\frac{dg}{dr}\right)=0$$ (18) where $`g=v_z^{}/(\mathrm{\Omega }r)`$ is the local disc tilt, $`v_z^{}`$ being the vertical component of the perturbed velocity. As in paper I the symbols are defined such that $$\kappa ^2=\frac{2\mathrm{\Omega }}{r}\frac{d(r^2\mathrm{\Omega })}{dr}$$ is the square of the epicyclic frequency, $$=_{\mathrm{}}^{\mathrm{}}\frac{\rho z^2}{c_s^2}𝑑z,$$ (19) $$\mu =_{\mathrm{}}^{\mathrm{}}\rho z^2𝑑z.$$ (20) Here it is supposed that the free particle nodal and apsidal precession frequencies $`\omega _z=2S/r^3,`$ and $`\mathrm{\Omega }\kappa =3\mathrm{\Omega }R_+/r1.5S/r^3`$ are small compared to the unperturbed angular velocity $`\mathrm{\Omega }`$ in magnitude. The relaxation of small amplitude bending disturbances at large radii is described by (18). As in paper I, this occurs through bending waves if $`\alpha _1<H/r,`$ and through diffusion with diffusion coefficient $`𝒟H^2\mathrm{\Omega }/4\alpha _1`$ on the characteristic time $`t_{diff}=r^2/𝒟`$ at radius $`r`$ if $`\alpha _1>H/r.`$ For $`\alpha _1`$ not too different from $`H/r`$, as in the numerical calculations described here, the relaxation is on the time scale required for bending waves to propagate across the relevant length scale. The waves propagate with speed $`c_B=(1/2)\sqrt{\mu /}(1/2)H\mathrm{\Omega },`$ which is one half of a vertically averaged sound speed. The relaxation is towards the solution of (18) corresponding to the least rapid decay in time. For a disc of infinite extent it is possible to look for a solution of (18) with $`\sigma =0.`$ This then takes the form $$g\left(\frac{\mathrm{\Sigma }\omega _z}{\mathrm{\Omega }}\right)+\frac{d}{dr}\left(\frac{c_B^2\mathrm{\Sigma }}{\mathrm{\Omega }(i\alpha _1\mathrm{\Omega }+\mathrm{\Omega }\kappa )}\frac{dg}{dr}\right)=0.$$ (21) When $`\mathrm{\Sigma }`$ and the disc aspect ratio are asymptotically constant for $`r\mathrm{},`$ (21) has solutions for which $`g`$ approaches an asymptotically constant value $`g_{\mathrm{}}`$ such that $$gg_{\mathrm{}}g_1_r^{\mathrm{}}\frac{\mathrm{\Omega }(i\alpha _1\mathrm{\Omega }+\mathrm{\Omega }\kappa )}{c_B^2\mathrm{\Sigma }}𝑑r,$$ (22) where $`g_1`$ is a constant which should be determined according to an appropriate inner boundary condition. As $`r`$ decreases, one eventually reaches a point where $`g`$ must begin to depart significantly from its asymptotic value $`g_{\mathrm{}}`$. Physically this is because the rate at which the disc is being twisted up by the Lense–Thirring precession begins to dominate over the local propagation/diffusion of disc warp, such that the disc is forced to lie in the equatorial plane of the black hole. We identify the radius where $`g`$ has to begin to depart significantly from $`g_{\mathrm{}}`$ with the Bardeen-Petterson transition radius. This radius can be estimated by equating it to the length scale associated with equation (21), or equivalently by the more physical approach of equating the rate at which the disc is twisted up by Lense–Thirring precession with the rate at which warping disturbances are propagated/diffused away. This latter approach is used below. These arguments were also used by Scheuer & Feiler (1996) to derive the transition or warp radius (see their equation (8) and discussion). We comment that the derivation of this transition radius requires the disc to be large in comparison but not necessarily infinite. We emphasise that if the disc is finite but still of significantly greater extent than the transition radius, although some slow long term realignment and precession of the disc will occur due to the its finite inertia, the initial relaxation to an approximate steady state like the one considered here is expected. This is in fact borne out by the simulation results. Scheuer & Feiler (1996) gave a solution of (21) for a constant $`\mathrm{\Sigma }`$ disc model with constant kinematic viscosity or equivalently, assuming $`\alpha _1`$ to be constant, constant $`c_B^2/\mathrm{\Omega }.`$ They adopted $`\mathrm{\Omega }=\kappa ,`$ however their solution can also be extended to the case $`(\mathrm{\Omega }\kappa )/\mathrm{\Omega }=ϵ,`$ with $`ϵ`$ being constant. The solution for $`g`$ (which in this context may be identified with the complex conjugate $`W^{}`$ of their function $`W`$to within a sign) is then $$g=g_{\mathrm{}}\mathrm{exp}\left(\left(\frac{8S\mathrm{\Omega }}{c_B^2r}\right)^{1/2}\left(ϵ+i\alpha _1\right)^{1/2}\right).$$ (23) As long as $`\alpha _10,`$ this solution is such that $`g0`$, for $`r0.`$ However, for $`ϵ>0,`$ corresponding to prograde precession, as in the Einstein case, and small $`|\alpha _1/ϵ|,`$ the solution exhibits many wavelike oscillations which may be interpreted as inward propagating bending waves excited in the transition region. This is in contrast to the situation of retrograde precession, $`ϵ<0,`$ as in the softened case, where the waves are evanescent. This phenomenon has been already noted by Ivanov & Illarianov (1997) and, as we shall see, it has some consequences for the calculation of the torque exerted between the black hole and disc in the limit of small $`\alpha _1.`$ But note that solutions with many oscillations in the inner disc appear in a linear theory. It is likely that non linear effects are likely to make very short wavelength waves disappear. The case of Einstein precession has $`ϵ=3R+/r,`$ which is not constant. However a solution of the Scheuer & Feiler (1996) type may be found in this case for constant $`\mathrm{\Sigma }`$ in the limit of vanishing $`\alpha _1,`$ if we assume the kinematic viscosity or equivalently $`c_B^2/\mathrm{\Omega }1/r.`$ This is $$g=g_{\mathrm{}}\mathrm{exp}\left(i\left(\frac{24SR_+\mathrm{\Omega }}{c_B^2r}\right)^{1/2}r^{1/2}\right).$$ (24) ## 5 The Transition Radius In order to obtain an estimate of the transition radius for the warp, we need to equate the rate at which the disc is twisted up by the rate at which warping disturbances are propagated across the disc. We note that estimates of the transition radius based on this approach should provide results that are accurate to within a factor of order unity. The physical processes of twisting the disc up and propagating/diffusing the warp occur over a range in radii such that the transition between aligned and misaligned planes will also occur over a range in radii, rather than at a single location. In the diffusive limit where $`H/r<\alpha _1,`$ and $`\alpha _1>|\mathrm{\Omega }\kappa |/\mathrm{\Omega },`$ we equate the rate of twisting up by differential precession and the rate of diffusion of warps. The rate at which the disc is twisted up is given by $`1/\tau _{dp}=|r(d\omega _z/dr)|=6S/r^3.`$ In the diffusion limit, the warping of the disc is counteracted by diffusion of the warp, which acts at a rate $`1/t_{diff}`$ where $`t_{diff}=r^2/𝒟`$ and $`𝒟=c_B^2/(4\alpha _1\mathrm{\Omega })`$ is the diffusion coefficient. Equating these two rates leads to an expression from which the transition radius can be found: $$R_{T1}=\left\{\frac{24\alpha _1S}{\sqrt{GM}}\left(\frac{H}{r}\right)^2\right\}^{2/3}$$ (25) This should be contrasted with the expression for the transition radius, $`R_{BP}`$, obtained using thin disc theory in which the effects of pressure forces are neglected and an isotropic viscosity is assumed (Bardeen & Petterson 1975; Hatchet, Begelman, & Sarazin 1981): $$R_{BP}=\left\{\frac{6S}{\alpha \sqrt{GM}}\left(\frac{H}{r}\right)^2\right\}^{2/3}.$$ (26) Here $`\alpha `$ is the standard viscosity parameter associated with the radial accretion. It is apparent that if $`\alpha <<1`$, then $`R_{BP}>>R_{T1}`$ and the transition radius predicted by (26) will lie much further away from the black hole, with the warp extending over a much greater proportion of the accretion disc than when (25) is used. This arises because the viscous damping of the resonantly driven horizontal motions included when (25) is used leads to a larger diffusion coefficient, and thus to a more rapid damping of the warp. It should be noted that $`R_{T1}`$ given by (25) continues to decrease as $`\alpha _1`$ decreases. Here we remark that for a strictly Keplerian disc the diffusive behaviour of warps used in (25) requires for its validity that $`\alpha _1>H/r.`$ When $`\alpha _1H/r,`$ warps are diffused across the disc on essentially the sound crossing time. At the beginning of the transition to the wave communication regime, we replace $`\alpha _1`$ by $`H/r`$ in (25), giving a second estimate of the transition radius $$R_{T2}=\left\{\frac{24S}{\sqrt{GM}}\left(\frac{H}{r}\right)^1\right\}^{2/3}.$$ (27) Note further that the disc is not strictly Keplerian in which case apsidal precession, which inhibits the communication of warps, should be taken into account. This modifies (25) for the diffusive regime to provide a third estimate of the transition radius. $$R_{T3}=\left\{\frac{24S\sqrt{(\alpha _1^2\mathrm{\Omega }^2+(\mathrm{\Omega }\kappa )^2)}}{\mathrm{\Omega }\sqrt{GM}}\left(\frac{H}{r}\right)^2\right\}^{2/3}.$$ (28) Note that if we take account of the fact that $`\mathrm{\Omega }\kappa ,`$ because $`S0,`$ but ignore the effect of Einstein precession, then we recover (27) in the limit of small $`\alpha _1.`$ The above discussion suggests that to estimate the Bardeen-Petterson radius we should take the maximum value found from using (25), (27) or (28). In this way unrealistic efficiency in warp communication is avoided. However, we remark that the transition radius is expected to be many gravitational radii from the black hole justifying the post Newtonian approximations in estimating its location. We further remark that the appropriate values of $`\alpha _1`$ for our disc models were calibrated using calculations similar to those presented in paper I and those in Nelson, Papaloizou, & Terquem (1999). These values are given in the introduction. ## 6 Numerical Method The set of fluid equations described in section (2) are solved using smoothed particle hydrodynamics (Lucy 1977; Gingold & Monaghan 1977). The SPH code used to perform the calculations presented in this paper is identical to that used in paper I, and we refer readers to this paper and references therein for a description of our numerical scheme. ## 7 Calculations of the Bardeen–Petterson Effect We have performed a number of calculations to examine the structure of an accretion disc produced when in orbit about a Kerr black hole, with the effects of Lense–Thirring precession included in the equations of motion. The calculations assume that the angular momentum vectors of the disc material and the spinning black hole are initially misaligned by some amount, with the system being evolved until a quasi steady state is achieved. The main issue that we wish to address in these simulations is the location of the so–called ‘Bardeen–Petterson radius’. This is the region of transition in the disc between the inner part, which in a steady state is expected to lie in the equatorial plane of the black hole due the combined effects of differential precession and viscosity, and the external part which is expected to remain in its original plane if the disc is effectively infinite. Due to the limitations of computational tractability, we are only able to consider discs which are of finite size. We do not expect this to affect the results on the transition radius, provided that the outer disc radius is sufficiently far from the transition zone, as is the case in our disc models. The initial conditions of each of the calculations were varied, such that the effects of changing the disc thickness, the exact form of the central potential, the number of particles, the disc radius, and the inclination angle between the black hole equator and initial disc midplane could be studied. These initial conditions are described in detail in the following section, which is followed by a description of the results from the numerical calculations. ### 7.1 Initial Conditions In each of the Bardeen–Petterson calculations, the central mass $`M=1`$, the gravitational constant $`G=1`$, $`a=1,`$ and $`𝐒=0.008`$ in equation (4). This leads to a value of the gravitational radius of $`R_+=0.04`$ in our computational units. The calculations can be divided essentially into two groups, those in which the central Keplerian potential was softened, and those in which no softening was used but in which the effects of Einstein precession were included. We will describe the former group of calculations first, followed by the latter group. One additional calculation was performed in which the effects of Einstein precession were neglected and no softening of the central potential was used, leading to a disc with a Keplerian rotation profile. All calculations are presented in table 1 , with the predicted value of $`R_{T2}`$ from equation (27) being shown in the eighth column, and that measured from the SPH calculations, $`R_{TS}`$ being presented in the last. In this paper we define $`R_{TS}`$, the transition radius measured from the simulations, to be the radius at which the disc tilt has a value that is half way between that of the inner and outer parts of the relaxed warp disc models, unless otherwise stated. The gravitational softening parameter used in equation (7) was $`b=0.2`$. A number of disc models were employed to examine the Bardeen–Petterson effect. An initial calculation used a disc model of radius $`R=1`$ and employed $`N=20,000`$ particles. This calculation had a midplane Mach number of $`12`$. Two calculations were performed with disc models of radius $`R=2`$ and with $`N=52,000`$. These calculations had $`12`$ and $`30`$ respectively. Two calculations were performed with a disc of radius $`R=\sqrt{2}`$ and $`N=102,000`$ particles, one with $`9`$, and the other with $`12`$. An additional calculation was performed with $`R=\sqrt{2}`$, $`N=60,000`$, and $`12`$. In all cases, the inclination between the black hole equator and the initial disc midplane was $`i=10^o`$. The calculations in which no gravitational softening was employed, and which included the effect of Einstein precession, had the central black hole treated as a uniform density sphere of finite size ($`b=0.2`$) in order to prevent numerical divergences as the disc material extends towards $`r0`$. Six calculations of this type were computed overall. Five of them had $`R=2`$ and $`N=52,000`$, two of which had had $`i=10^o`$ and $`12`$ and $`30`$, two of which had $`i=30^o`$ with $`12`$ and $`30`$, and one had $`i=10^o`$ and $`=5`$. The sixth calculation had $`i=10^o`$, $`R=7`$, $`N=200,000`$, and $`14`$. An additional calculation was performed in which the effects of Einstein precession were neglected and no softening of the gravitational potential was used. This calculation had $`i=30^o`$, $`=12`$, $`R=2`$, and $`N=52,000`$, and was performed to provide a comparison case with the other high inclination runs. All calculations were initiated with the disc orbiting the black hole, and with the inclination between the black hole equator and the original disc plane being constant throughout the calculations. The calculations were all evolved towards a quasi steady state, such that a precessing, but otherwise almost stationary warped structure was obtained in each case. The results of these calculations are described in the following section. ## 8 Results In this section we present and discuss the results of the calculations described in section (7.1). The results of the calculations that were performed using a softened gravitational potential are described first, followed by the calculations in which the effects of Einstein precession were included. In the discussion of the results that follows, we will be describing the evolution of the angle of inclination between angular momentum vector of the black hole, $`𝐉_{bh}`$, and the specific angular momentum vector of the disc material, $`𝐉_d`$. This quantity is defined through the expression $$i(r)=\mathrm{arccos}\left(\frac{𝐉_{bh}.𝐉_d(r)}{|𝐉_{bh}||𝐉_d(r)|}\right),$$ (29) and may be a function of radius through the disc, such that the disc is warped. The calculations are initiated with this quantity being constant throughout the disc. We note that $`g`$, introduced in equation (18), and $`i`$ are equivalent. We will also be describing the precession of the local angular momentum vector in the disc, projected onto the fixed $`x`$$`y`$ plane of the black hole equator. This precession angle is defined by the expression $$\beta (r)=\mathrm{arccos}(\frac{𝐉_{bh}\times 𝐉_d(r)}{|𝐉_{bh}\times 𝐉_d(r)|}.\widehat{𝐲_o}),$$ (30) and may vary as a function of radius in the disc, such that the disc is twisted. Here, $`\widehat{𝐲_o}`$ is the unit vector pointing in the $`y`$ direction located in the fixed equatorial plane of the black hole. The calculations are initiated with $`\beta =\pi /2`$ throughout the disc. The results shown in the particle projection plots are presented in a fixed reference frame centred on the black hole. The initial inclination between the disc midplane and the black hole equator is provided by rigidly rotating the disc about a diameter coincident with the $`x`$ axis in this fixed reference frame. The evolution of all of the disc models in the simulations is as follows. Initially, there is a short period of rapid evolution during which the inner parts of the disc undergo strong differential precession and are forced to lie in the symmetry plane of the black hole out to a transition region, beyond which the discs remain in their original plane. During this time, warping disturbances propagate through the disc either as bending waves or through diffusion, and a warped disc configuration is set up on the time scale required for these warps to propagate across the disc. This time scale for the the warped configuration and the transition zone to be set up also corresponds to the differential precession time scale measured at the transition radius, $`\tau _{dp}`$. The discs were all evolved for a time $`>>\tau _{dp}(R_{TS})`$ to ensure that the position of the transition zone had been firmly established. The run time of each calculation, in terms of the relevant differential precession time $`\tau _{dp}`$, is given in table 1. There then follows a period of evolution that occurs on a much longer time scale involving near–rigid body precession of the disc, and a slow alignment of the outer parts of the disc with the symmetry plane of the central black hole. This period of evolution arises because of viscous coupling between the misaligned inner and outer parts of the disc, and as a result of residual torquing by the black hole beyond the transition radius. It does not arise because warping disturbances are being propagated to large radii from within the transition zone of the disc. The communication of warps by bending waves or diffusion appears to be confined within the transition zone, exactly as expected from the linear theory. This is evidenced by the fact that the outer regions evolve much more slowly than they would if they were being torqued into alignment as a result of warps propagating from the inner regions of the disc. We have examined the outer parts of the discs in our models, and find no evidence of warping disturbances reaching them from their central parts over time scales that are typically $``$ few hundred $`\mathrm{\Omega }(R=1)^1`$. We interpret this to mean that using disc models with larger outer radii will not result in changes to the locations of the transition radii that we find in our models. This is confirmed by run E6. ### 8.1 Calculations Employing Gravitational Softening A number of calculations were performed in which the central gravitational potential was softened and the effects of Einstein precession were neglected. These calculations were performed with $`9`$, 12 or $`30`$ (i.e. with $`H/r0.11`$, 0.08 or 0.03). #### 8.1.1 $`12`$ Calculations A time sequence showing the evolution of calculation S1, listed in table 1, is shown in Fig. (1). Projections of particles contained within a thin slice centred about the $`x`$ axis (left panels) and the $`y`$ axis (right panels) are presented, with the time corresponding to the panels being shown in units of $`\mathrm{\Omega }^1`$ at $`R=1`$ in the top right hand corner of the left hand panels. The top two panels show the disc at the beginning of the calculation, with the initial tilt of the disc relative to the fixed coordinate system being apparent in the right hand panel. As the calculation proceeds, the Lense–Thirring precession twists up the inner parts of the disc, which are forced eventually to lie in the equatorial plane of the black hole, with the outer disc remaining close to its original plane. The middle two panels show the disc at an intermediate state of its evolution, during the time in which the inner parts of the disc are still in the process of flattening out into the equatorial plane of the black hole. The final two panels show the disc after it has evolved towards a quasi–steady state configuration, in which a smoothly varying warped disc has been established which undergoes approximate solid–body precession. We comment that this non zero precession rate is a consequence of having a finite disc. If the disc were to extend to arbitrary large radii, the angular momentum content may tend to arbitrary large values while the total mass remained finite. Then the precession period would approach arbitrary large values. Fig. (2) shows the time evolution of the disc tilt versus radius for the calculation S2 shown in table 1, with the unit of time being $`\mathrm{\Omega }^1`$ evaluated at $`R=1`$. The angular momentum vector of the disc at each radius is initially tilted by $`i=10^o`$ with respect to the angular momentum vector of the black hole. As the calculation proceeds the inner parts of the disc begin to lie in the equatorial plane of the black hole, and eventually a disc configuration is developed in which the disc tilt varies smoothly as a function of radius, and with the radius of transition between the inner and outer planes being at around $`R_{TS}0.6`$, which is equivalent to $`R_{TS}=15R_+`$ in our units. To recap, we define the transition radius measured in the simulations, $`R_{TS}`$, to be the radius at which the disc angle of tilt is half way between that of the outer part of the disc and the inner part of the disc. In the case of calculation S2, the transition zone approximately occupies the range of radii 0.4 – 1.5. The differential precession time measured at $`R=1.5`$ is $`\tau _{dp}(1.5)=70\mathrm{\Omega }(R=1)^1`$. The calculation is evolved for 170 $`\mathrm{\Omega }^1`$ without any apparent change in the transition radius between the times 100 – 170 $`\mathrm{\Omega }^1`$. We find in this calculation, and in the others also, that the position of $`R_{TS}`$ agrees reasonably well with the semi–analytic formulae presented in section 4, such as equation (27), to within a factor of 2 or 3, thus justifying the physical arguments leading to these expressions. We note that the choice of defining $`R_{TS}`$ to be the radius corresponding to the half way point of the warped region is somewhat arbitrary, and that a definition of $`R_{TS}`$ being the position where the inclination approaches its value at the outer disc edge would produce values of $`R_{TS}`$ in table 1 that are in greater agreement with the tabulated values of $`R_{T2}`$. We also comment that the inclination of the outer disc boundary is free, and thus its location should not affect the position of the transition radius if it is sufficiently distant (compare models S1, S2, S3, S4 in Fig., and E6 in Fig. (13). The equations (25), (27), and (28) that predict the transition region were derived under the assumption of linearity of the warp. We note that the warp in the Mach 30 cases is nonlinear, as described in paper I, such that we should expect a poorer agreement between the transition radii measured in the simulations and those predicted using equations such as (27). Fig. (3) shows the time evolution of the nodal precession angle, $`\beta `$, for different annuli of the accretion disc, computed from the results of calculation S2. The definition of $`\beta `$ is given by equation (30), such that at time $`t=0`$, $`\beta =90^o`$ for the whole disc. In Fig. (3), the disc has been subdivided into three annuli which each lie in the range $`0r<0.66`$, $`0.66r<1.33`$, and $`1.33r2`$, respectively. As the calculation proceeds, it may be observed that the disc initially undergoes a period of differential precession, with the inner disc precessing more rapidly than the outer disc, before settling down to a quasi steady state in which the disc precesses approximately as a rigid body. The degree of twist in the disc at the end of the calculation is represented in Fig. (4). This figure indicates that the angle between the line of nodes of disc annuli in the inner part of the disc, say at $`r=0.6`$, and the outer disc edge is $`25^o`$, showing that the disc has developed a $`25^o`$ twist before reaching a quasi equilibrium state. A comparison between each of the runs with midplane Mach numbers $`12`$ is presented in Fig. (5). This figure shows the results from each of the runs S1, S2, S3, and S4 presented in table 1 . It is apparent that the radial variations in tilt, and in particular the positions of the transition radii, are similar in each case, even though the resolution of each calculation differs, indicating that convergence of the results has been attained. We note that the position of outer radius of the disc does not seem to alter the location of the transition region, since the outer disc lies sufficiently beyond the warped region. The curve corresponding to the R=1 disc (calculation S1) shows that the outer parts of this disc are tending towards alignment with the black hole equatorial plane. This is expected for all discs of finite extent, and hence finite angular momentum and moment of inertia, since there is a torque between the inner and outer parts of the disc that attempts to bring them into alignment. The rate at which this occurs obviously depends on the angular momentum content, and hence radius, of the disc, with a smaller disc being aligned more quickly. The larger (R = $`\sqrt{2}`$ and $`R=2`$) discs will also eventually evolve towards alignment, but on a time scale that is longer than the time for which these calculations were run. This issue will be discussed in greater detail in section 8.5. As described in paper I, the result of warping a vertically stratified, gaseous accretion disc is to set up vertical shearing motions, such that the perturbed radial velocities, $`v_r`$, are an odd function of $`z`$. The warped discs generated by the action of Lense–Thirring precession indeed show such a kinematic feature over those portions of the disc where the inclination, $`i`$, varies with radius. This fact is illustrated by Fig. (6), in which the radial velocities of the particles contained within a thin slice of the disc centred about the $`y`$ axis are represented by velocity vectors. The arrows located at the bottom of the figure indicate the magnitude of the midplane sound speed at each position. The data used to generate this figure were obtained from the calculation S4. #### 8.1.2 Changing the Disc Thickness – $`30`$ and $`9`$ Calculations A time sequence showing the evolution of the disc during the calculation S6 is shown in Fig. (7). Projections of particles contained within a slice centred about the $`x`$ axis (left panels) and the $`y`$ axis (right panels) are presented. The time corresponding to each panel is shown in the top right hand corner of each left hand panel in units of $`\mathrm{\Omega }^1`$ evaluated at $`R=1`$. The first two panels show the disc after it has evolved for $`t15\mathrm{\Omega }^1`$, and the existence of apparent ‘wiggles’ may be observed in the inner parts of the disc. These features are transient phenomena and are caused by the differential precession of the inner disc. As time proceeds, the wiggles damp away, as the inner disc flattens into the equatorial plane of the black hole, eventually leaving a disc with a warp that is slowly varying as a function of radius. We notice that the inner region of the disc that becomes aligned with the black hole equator is of much greater radial extent that that occurring in Fig. (1), as expected from the discussion presented in section 4. The variation of disc tilt as a function of radius is plotted for three different runs in Fig. (8) in order to examine the effect of changing the disc thickness. The midplane Mach number of each run is $`9`$ (dot–dashed line), $`12`$ (solid line), and $`30`$ (dashed line), and it is apparent that the transition radius increases with increasing $``$. This result is expected from the analytic estimates of $`R_{T1}`$ and $`R_{T2}`$ presented in section 4, since a thicker disc diffuses disc warp more efficiently. The value of $`R_{TS}`$ changes from $`R_{TS}10R_+`$ to $`16R_+`$, and $`33R_+`$ as $``$ changes from $`9`$, 12, and 30, respectively. It should be noted that these values of $`R_{TS}`$ are very much smaller than those predicted using the thin viscous disc theory neglecting hydrodynamical effects, (Bardeen & Petterson 1975), and confirm the previous assertions of Papaloizou & Pringle (1983), and Kumar & Pringle (1985), that taking into account the full disc hydrodynamics results in a transition radius that lies much closer to the central black hole. The degree of twist in the $`30`$ disc is shown in Fig. (9), and indicates that the twist angle between the line of nodes in the inner ($`R0.9`$) and outer parts ($`R=2`$) of the disc is $`90^o`$, which is significantly larger than was found to be the case for the $`12`$ disc. This larger twist occurs because communication through the disc by bending waves/diffusion is much less efficient in this case, so that the disc needs to acquire a more distorted configuration before it can obtain a quasi steady state structure which undergoes approximate solid body precession. We note that although the numerical resolution in the $`30`$ calculations is reduced relative to the $`12`$ calculations, a plot of the velocity field similar to that shown in Fig. (6) produces a similar picture of a disc in which the vertical shearing motion extends over those regions of the disc in which $`i`$, the disc inclination, varies with radius (i.e. in those regions where the curvature of the disc is non–zero). We also found that these motions were subsonic which is consistent with the idea that damping due to shocks keeps them down to this level. This indicates that the thinner disc models still have their vertical structure modeled to a reasonable degree of accuracy, such that it is able to produce this vertical shear. This is because the smoothing lengths still remain $`<H`$, the disc semi–thickness, beyond that region of the disc which resides in the equatorial plane of the black hole, and where it becomes noticeably curved. We remark that calculation E5, which has the largest disc thickness of the models we considered with $`=5`$, showed no signs of alignment in its inner regions, and represents the limiting disc thickness that allows the formation of a noticeably warped disc. This has relevance to the properties of the thick advection dominated accretion discs (i.e. ADAFs) around rotating black holes, which will presumably show no outward signs of being warped. ### 8.2 Calculations Including the Effects of Einstein Precession The discussion of warp propagation given in paper I indicates that efficient, non dispersive bending wave propagation is expected to occur only for Keplerian accretion discs in which $`\mathrm{\Omega }\kappa `$. In a situation where $`\mathrm{\Omega }\kappa `$, then the communication of warping disturbances is expected to become dispersive, with the effects of this dispersion becoming more important as $`|\mathrm{\Omega }\kappa |`$ increases (i.e. as the apsidal precession frequency increases), since the resonance between the circular and epicyclic motion becomes increasingly detuned. As was noted in section (3), the frequency of Einstein precession (ı.e. advance of periapse) remains relatively high out to quite large radii in our disc models, indicating that it may be an important consideration in determining the structure and evolution of warped accretion discs around black holes. A comparison between the calculations S2 and E1 is presented in Fig. (10). The solid line shows the variation of disc tilt versus radius for calculation S2, whereas calculation E1 is indicated by the dashed line. The primary difference between the cases with Einstein precession included and those with it neglected is that the transition between the inner and outer disc planes occurs much more sharply when Einstein precession is included. The calculations for $`i=10^o`$ and $`=30`$ are also shown in Fig. (10), S6 being shown by the dashed–dotted line and E2 by the dotted line, and illustrate this steeper transition between the inner and outer disc planes more clearly. This arises because the communication of the disc tilt between between the two misaligned planes is more effective when Einstein precession is neglected, so that the transition may be smoothed out more effectively by the enhanced diffusion of disc tilt. One effect of this modification of disc structure by the inclusion of Einstein precession in our calculations is to change the rate at which the outer disc tends to align with the equatorial plane of the black hole. The existence of a more discontinuous transition between the two disc planes leads to a larger frictional/viscous interaction between the outer and inner disc at their points of intersection, and thus to a larger torque that acts to realign the two separate regions of the disc. Since the inner disc is anchored to the black hole equatorial plane by the Bardeen–Petterson effect, the outer disc is also forced to evolve towards this plane. Our calculations indicate that this rate of realignment is enhanced when Einstein precession is included. Although this seems counter intuitive since disc tilt is supposed to diffuse through the disc at a lower rate when the rotation profile is non Keplerian, this result is in agreement with the predictions of linear theory, as discussed in section (8.5). In addition to the transition between the inner and outer disc planes being steeper, we also find that the disc twist is greater when Einstein precession is included. Fig. (11) shows the precession angles of different disc annuli at the end of the calculations E1 and E2. These plots should be compared with Figs. (4) and (9) which show the precession angles for the calculations S2 and S6. We find that calculation E1 results in a twist of the disc of $`90^o`$ between the radii $`R=0.6`$ to 2.0, which should be compared with $`25^o`$ for calculation S2. A similar comparison between E2 and S6 yields twist angles of $`120^o`$ and $`90^o`$, respectively. This is entirely consistent with linear analysis \[see section (8.5)\] which indicates that inward propagating short wavelength bending waves should be present in the inner part of the transition region producing a twist there. Physically the large twist angles produced in the simulations are a result of the diminished efficacy of communication between neighbouring annuli in the disc when the rotation profile becomes significantly non Keplerian. In order for such a disc to settle to a steady state structure which is in a state of near solid body precession, the disc must become more severely distorted and twisted so as to increase the torques acting between neighbouring annuli of gas. We remark that many oscillations in disc tilt as indicated by the linear analysis discussed here and found in the linear calculations of Ivanov & Illarianov (1997), do not appear in our nonlinear simulations. This is presumably because non linear effects lead to the damping of these short wavelength features, and alignment of the inner disc regions in which the tilt amplitude would otherwise change rapidly on small length scales. ### 8.3 Effects of Increasing the Inclination Angle $`i`$ Three calculations were performed with larger degrees of inclination between the black hole equatorial plane and the original disc plane. These were the calculations listed as E3, E4, and K1 in table 1 . The results of these calculations are plotted in Fig. (12), along with those of calculations E1 and E2 which are plotted for the purpose of comparison. As in the previous runs, the calculations result in a disc structure which is essentially composed of two mutually inclined regions, the inner and outer disc, and a transition zone between them. In the low inclination ($`i=10^o`$) runs, the transition zone appears to join the two misaligned inner and outer disc parts more smoothly than is the case with the higher inclination ($`i=30^o`$) runs. This is because the inner disc, which becomes aligned with the black hole equator, and the outer disc, which remains in its original plane, occupy approximately the same intervals in radius irrespective of whether the inclination angle is $`10^o`$ or $`30^o`$. Thus, the transition between these two planes occurs more steeply simply because of the greater degree of inclination when $`i=30^o`$. The steepness of the transition in the $`30^o`$ inclination degree runs, combined with an inspection of the disc structures resulting from the calculations E3 and E4, indicate that the discs are close to breaking into two discrete pieces with only a tenuous bridge of material connecting them. The dominant means by which the two separate pieces of the disc communicate with one another is probably through the viscous coupling at their region of intersection rather than through wave–like or diffusive communication. ### 8.4 A Larger Disc Model A calculation was performed using a disc model with an outer radius at $`r=7`$ in order to ensure that the transition radii obtained in previous sections are not affected by the smaller disc models used. This calculation is listed as run E6 in table 1, and employed $`2\times 10^5`$ particles. The midplane Mach number $`14`$, and the expected value of $`\alpha _10.04`$, so that bending waves are expected to travel in this disc model. The effects of Einstein precession were included so that equation (8) was used to calculate the gravitational force. The evolution of the disc tilt for this run can be observed in Fig. (13) where each line corresponds to the disc tilt versus radius at different times during the run. The times are shown in the figure in units of $`\mathrm{\Omega }^1`$ evaluated at $`R=1`$. We note that the warped disc and the transition radius are expected to be set up after a time corresponding to the differential precession time at the radius of the transition zone. From Fig. (13) we notice that the transition radius, as defined in the first paragraph of section 7.1, is located at $`R0.77=20R_+`$, with the disc tilt approaching that of the original disc beyond a radius of $`r2`$. This result agrees very well with those obtained for the disc models with $`12`$ and outer radii of $`r=2`$ described in previous sections, as seen in table 1. The larger disc of radius $`r=7`$ is slightly thinner with $`14`$, with the result that the transition zone is slightly further out in this case. The differential precession time at a radius of $`r=2`$ is $`\tau _{dp}166`$, which is the time required for the transition zone to become fully relaxed. The calculation ran for a time of $`400`$ $`\mathrm{\Omega }(R=1)^1`$. It is apparent that once the transition zone has been established on the precession time, its position remains fixed when the calculation is continued for a number of precession times. The disc exterior to the transition zone undergoes some longer time scale readjustment as a transient disturbance propagates through the disc from the central regions. As is the case with the calculations presented in previous sections, the outer disc undergoes a process of slow realignment with the black hole equator, with the location of the transition radius remaining fixed throughout this period of the evolution. Particle projection plots are presented in Fig. (14) showing the structure of the warp at a time $`t=313\mathrm{\Omega }(R=1)^1`$, after the warp has settled to a quasi-steady state. The first panel shows a projection in the $`y`$$`z`$ plane of a slice through the disc, and presents a global view of the warp. A closer view of the central regions is shown in the second panel, and illustrates how the warp and transition zone are confined to the central regions of the disc. ### 8.5 Alignment Time Scale for Black Hole and Accretion Disc As the calculations in the previous sections have shown, the effect of the Lense–Thirring precession induced by the dragging of inertial frames by a Kerr black hole is to cause the inner regions of a misaligned accretion disc to become aligned with the black hole angular momentum vector. Newton’s third law, however, implies that equal but opposite torques are exerted on the black hole by the disc, causing it’s angular momentum vector, in turn, to become aligned with the angular momentum vector of the outer disc material. This process was considered by Rees (1978), who estimated the alignment time scale for the black hole and accretion disc by considering the mass flux through the disc. The direction of the angular momentum of this infalling matter is changed as it passes through the transition region between the aligned and misaligned disc, so that a simple estimate of the alignment time scale may be obtained from $$\tau _{align}\frac{J_{BH}}{\dot{M}j_d(R_{BP})\mathrm{sin}(i)},$$ (31) where $`J_{BH}`$ is the black hole angular momentum, $`\dot{M}`$ is the mass accretion rate through the disc, $`j_d(R_{BP})`$ is the specific angular momentum of disc material as it enters the warped region at $`R_{BP}`$, and $`i`$ is the inclination angle between the outer disc and black hole equator. Using this general picture, Rees (1978) calculated the time scale for alignment in AGN and estimated it to be in the region of $`10^8`$ yr, comparable to the ages of jets in AGN whose directions seem to have been constant over these time scales. The above process was considered in more detail by Scheuer & Feiler (1996). These authors derived an analytic expression to describe the steady state warped structure of a disc around a Kerr black hole, using a linearized set of equations derived from those of Pringle (1992), to describe the evolution of a warped disc. By calculating the Lense–Thirring torque due to the black hole on this disc, they were able to calculate the reverse torque experienced by the black hole, and hence to estimate its alignment time scale. They found that the more detailed calculations gave time scales in essential agreement with the ideas proposed by Rees (1978), provided that the viscous diffusion coefficients acting in and out of the disc plane, $`\nu _1`$ and $`\nu _2`$, are equal. These estimates of the alignment time scale were calculated by assuming that disc warp is communicated through a disc on the standard viscous time scale, so that the transition radius $`R_{BP}`$ is then expected to be at a large distance from the black hole. The calculations presented in previous sections indicate that when the full disc hydrodynamics are considered, then the evolution of warped discs changes substantially, such that the transition radius moves closer to the black hole. This effect has recently been considered by Natarajan & Pringle (1998), who have estimated alignment time scales of $`10^5`$ yr for black holes in AGN using the formula suggested by Scheuer & Feiler (1996), but taking into account differences between the viscous diffusion coefficients that act in the plane and perpendicular to the plane of the disc. The accretion disc models that we calculate may be different from those expected to occur in AGN, since they are relatively thick ($`H/r0.08`$$`0.03`$), whereas $`H/r10^3`$ may be more appropriate to the outer discs of AGN. In addition, they are of small radial extent. This prevents us from making detailed calculations of the alignment time scale for black holes in AGN on the basis of our calculations. Nonetheless, we can examine the rate at which the global disc tilt changes in the simulations, and compare it with that expected from the tilt evolution caused by the advection of misaligned angular momentum through the warped region, and with the work of Scheuer & Feiler (1996). If the disc tilt in our simulations changes solely because of advection of material with misaligned $`J`$ through the warped region, where it is torqued into alignment, then the rate of change of misaligned disc angular momentum may be estimated to be $$\frac{dJ}{dt}\dot{M}j(R_{TS})\mathrm{sin}(i),$$ (32) where $`j(R_{TS})`$ is the specific angular momentum of disc material as it enters the warped region located at $`R_{TS}`$. The corresponding expression that applies when the transition zone occurs at the Bardeen–Petterson radius, $`R_{BP}`$ is given by $$\frac{dJ}{dt}\dot{M}j(R_{BP})\mathrm{sin}(i),$$ (33) where $`j(R_{BP})`$ is the specific angular momentum of disc material as it enters the region located at $`R_{BP}`$. By calculating $`\dot{M}`$ from our simulations, we can estimate the effect of these processes, and the associated time scales for them to align the disc with the black hole and thus by Newton’s third law, the time to align the black hole with the disc. We denote these time scales by $`\tau _{TS}`$ and $`\tau _{BP}`$, respectively. In our finite disc models, we expect the global change in the disc tilt to arise because of advection of tilt angular momentum through $`R_{TS}`$, viscous coupling between misaligned disc annuli, and residual torquing by the black hole on disc elements that lie somewhat beyond $`R_{TS}`$, which remains relatively close to the black hole in our simulations. The time scale for the alignment of a black hole’s spin angular momentum vector with that of the outer part of an accretion disc is given by Scheuer & Feiler (1996) to be $$\tau _{align}\frac{1}{\pi \mathrm{\Sigma }}\left(\frac{acM}{\nu _2G}\right)^{1/2}$$ (34) where $`a`$ is the Kerr parameter, $`c`$ is the speed of light, $`M`$ is the black hole mass, $`\mathrm{\Sigma }`$ is the disc surface density, and $`\nu _2`$ is the viscous diffusion coefficient acting on the disc warp. We remark that the expression of Scheuer & Feiler (1996) which gives the aligning torque (their equation ) indicates that most of the torque contribution arises from the region of the disc that is in the vicinity of the warp or transition radius, as was commented upon by these authors themselves, implying that the alignment time scale obtained is not reliant upon the disc being infinite, but just significantly larger than the transition radius. This is also the case if the expressions for the tilt (23) and (24), applicable when $`\alpha _1`$ is small, are used in order to evaluate their expression for the torque. Then $`\nu _2`$ in (34) should be replaced such that $$\nu _2\nu _2\alpha _1\frac{\left(\sqrt{(ϵ^2+\alpha _1^2)}+ϵ\right)}{\left(ϵ^2+\alpha _1^2\right)}.$$ (35) In addition for these cases $`\nu _2=c_B^2/(\alpha _1\mathrm{\Omega }).`$ Note that in the Einstein precession case (24), the limit $`\alpha _10,`$ with both $`\nu _2`$ and $`ϵ`$ $`1/r,`$ is to be taken with the result that $`\nu _2`$ is replaced by a constant value. An interesting feature of the above is that in the case of prograde Einstein precession $`(ϵ>0`$) an alignment torque survives in the limit $`\alpha _10,`$ whereas in the retrograde case $`(ϵ<0)`$, as occurs with softening, it vanishes. This is because of the short wavelength oscillations or bending waves in the former case. In the latter case the solution is evanescent and of the wrong phase, having zero twist, to produce an alignment torque. The above suggests that cases with prograde precession should show significantly larger twists and alignment torques than those with retrograde precession when the viscosity is small. This is found to be the case in our simulations. The black hole alignment time scale may be written as $$\tau _{align}\left|\frac{dJ}{dt}\right|^1\left(\frac{GM^2a\mathrm{sin}(i)}{c}\right)$$ (36) By Newton’s third law, the alignment torque acting on the black hole is equal and opposite to that acting to align the disc plane with the equatorial plane of the black hole. Thus, the alignment torque acting on the disc may be written $$\frac{dJ}{dt}\pi \mathrm{\Sigma }\left(\frac{\nu _2G}{acM}\right)^{1/2}\left(\frac{GM^2a\mathrm{sin}(i)}{c}\right).$$ (37) We note that $`S=G^2M^2a/c^3`$, so that we may write equation (37) as $$\frac{dJ}{dt}=\pi \mathrm{\Sigma }\left(S\nu _2GM\right)^{1/2}\mathrm{sin}(i).$$ (38) In their discussion, Scheuer & Feiler (1996) distinguish between the viscosity coefficient, $`\nu _1,`$ acting in the plane of the disc which is responsible for mass flow through the disc, and that which is responsible for the damping or diffusion of disc warp, $`\nu _2.`$ In the situation where the diffusion coefficients are equal (i.e. $`\nu _2=\nu _1`$ where $`\nu _1=\alpha H^2\mathrm{\Omega }`$) equation (38) may be written as $$\frac{dJ}{dt}\pi \mathrm{\Sigma }(S\nu _1GM)^{1/2}\mathrm{sin}(i).$$ (39) Recent work on the evolution of warped discs, including that presented in paper I, indicates that disc warp diffuses on a time scale much faster than that on which mass diffuses through the disc, and that the appropriate value for $`\nu _2`$ is $`\nu _2=\nu _1/(2\alpha ^2)`$. In order to estimate the alignment time scale for the black hole angular momentum vector, Natarajan & Pringle (1998) used this relation in equation (34). For our purposes of calculating the torque acting on the disc, using this relation in equation (38) leads to the expression $$\frac{dJ}{dt}\frac{\pi \mathrm{\Sigma }}{\alpha }\left(\frac{S\nu _1GM}{2}\right)^{1/2}\mathrm{sin}(i).$$ (40) We are able to compare our simulated disc alignment torques with the values expected from equations (32), (33),(39), and (40). We denote the expected time scales for the discs to align with the black hole equatorial plane calculated from these equations as $`\tau _{TS}`$, $`\tau _{BP}`$, $`\tau _{SF1}`$, and $`\tau _{SF2}`$, respectively. We denote the time scale for disc alignment measured from the simulations by $`\tau _m.`$ We define the time scales for disc alignment to be $$\tau =|𝐉_{}|\left(\frac{dJ}{dt}\right)^1$$ (41) where we take $`𝐉_{}`$ to be the perpendicular component of the angular momentum of the disc models at time $`t=0.`$ This value is used in the estimates of $`\tau _m`$, $`\tau _{TS}`$, $`\tau _{BP}`$, $`\tau _{SF1}`$, and $`\tau _{SF2}`$. The value of $`\tau _m`$, the time scale for disc alignment measured from the simulations, was obtained by measuring the change in $`𝐉_{}`$ over time, and extrapolating forward to the point when $`𝐉_{}=0`$. The values of $`\dot{M}`$ used in equations (32) and (33), to estimate the values of $`\tau _{TS}`$ and $`\tau _{BP}`$, were obtained from the simulations by measuring the mass flux through the disc models. The values of $`R_{TS}`$ were taken from table 1 and the values of $`R_{BP}`$ were calculated using equation equation (26). The values of $`\nu _1`$ and $`\alpha `$ used in equations (39) and (40) were obtained from the measurements of $`\dot{M}`$ and assuming a steady state such that $`\dot{M}=3\pi \nu \mathrm{\Sigma }`$. The values of $`\alpha `$ used in the calculation of the alignment time scales for each model were (S2 – $`\alpha =0.022`$; S6 – $`\alpha =0.12`$; E1 – $`\alpha =0.022`$; E2 – $`\alpha =0.14`$; E6 – $`\alpha =0.04`$). These are very similar to the values of $`\alpha _1`$ obtained from the bending wave calculations presented in paper I, and indicate that SPH produces a reasonably isotropic viscosity when applied to this problem. The alignment time scales calculated from the simulations and the analytical torque estimates are presented in table 2. First of all, we consider the runs that employed softening of the gravitational potential, since in this case it is relevant to compare these calculations with the analytical estimates of the alignment time scales, given that the softening plays only a minor role in the simulations. The physical models that lead to the analytical estimates of the alignment time scales do not include the effects of Einstein precession, and so strictly speaking they should not be compared with the results of calculations E1, E2, and E6 in which Einstein precession is included and plays an important role. From table 2 we see that the results for the runs S2 and S6 indicate that the predicted alignment time scale, $`\tau _{TS}`$, estimated by assuming that material is only torqued into alignment when it accretes through the transition zone, $`R_{TS}`$, is too long, indicating that the viscous torques acting between the aligned and misaligned disc components, and the residual torque due to the black hole beyond $`R_{TS}`$, are having a substantial effect on the evolution of the disc tilt. The alignment time scale $`\tau _{SF2}`$ leads to a prediction that the discs will align with the black hole equatorial plane on too short a time scale, whereas the estimate $`\tau _{SF1}`$ predicts too long a time scale. Interestingly, it appears that the estimate $`\tau _{BP}`$ based on equation (33) provides the best estimate of the alignment time scale in the cases S2 and S6. The physical picture suggested by Rees (1978) is of a warped disc with constant mass flow in which the misaligned component of the disc material’s angular momentum is transferred to the black hole as it passes through the warped region located at $`R_{BP}`$. It may seem surprising that we obtain results for the alignment time scale that are consistent with Rees (1978) in view of the facts that the transition radius is much smaller in our case and (40) predicts a significantly faster rate. But we note that (40) is based on a linearized approximation which does not allow for the torques applied in the warped region of the disc to feedback into the disc structure. One would expect that this feedback would clear material with misaligned angular momentum from this region which would then have to be resupplied by mass accretion. This would result in a dependence of the alignment rate on the mass accretion rate. From the discussion presented in section 4, we would expect that the alignment time scales obtained when Einstein precession is included would be noticeably shorter than when it is neglected. This arises because of the larger twist in these cases, leading to a larger torque. By examining the values of $`\tau _m`$ listed in table 2 we see that this is indeed the case. Comparing the runs S2 and E1, we see that the alignment time scale is shorter by a factor of $`4`$ when Einstein precession is included. Comparing runs S6 and E2, we observe the same trend, but at a reduced level because the transition zone is further from the black hole in these cases and the effects of Einstein precession are reduced accordingly. By comparing the results of runs E1 and E6, we notice that $`\tau _m`$ is larger than in the latter case because of the larger disc size. Although the disc model parameters for these two cases differ slightly, giving slightly different transition radii, we would expect that the ratio of alignment time scales would be similar to the ratio of the initial disc angular momenta if the aligning torques are similar. We find that $`\tau _m(E1)/\tau _m(E6)0.1`$ and $`J_{}(E1)/J_{}(E6)0.12`$, suggesting that we not only calculate the position of the transition zone correctly with our smaller disc models with outer radii $`r=2`$, but that we also accurately calculate the alignment torque between the outer and inner discs since this arises from the transition zone. ## 9 Discussion and Conclusion In this paper we have performed simulations of accretion discs in orbit around a rotating black hole, which had its spin vector initially misaligned with that of the disc, in the lowest order post Newtonian approximation. Discs with midplane Mach numbers between 5 and 30 were considered. For discs with Mach numbers of 5, 9, and 12, the effective viscosity acting through the $`r`$$`z`$ component of the viscous stress, $`\alpha _1`$, is such that the warps are controlled by bending waves (see paper I). For the higher Mach number of 30, the warps evolve diffusively, but with an associated diffusion coefficient that is a factor of $`1/(2\alpha _1^2)`$ larger than that associated with mass flow through the disc (Papaloizou & Pringle 1983). As expected the central portions of the disc models became aligned with the equatorial plane of the black hole out to a transition radius, beyond which the discs remained close to their original plane. This structure was established on the sound crossing, or warp diffusion, time. This period of initial relaxation to a well defined warped disc structure was followed by a period of evolution occurring on a longer time scale, corresponding to solid body precession of the disc and a slow process of alignment of the outer disc with equatorial plane of the black hole. Both of these are caused by the fact that we are forced to consider isolated discs of finite radii for reasons of numerical tractability. Nonetheless the position of the transition between the aligned and misaligned disc components is not affected by this. Because of the much more effective communication of warps than implied by the standard viscous time scale, the transition radius was found to be much smaller than that given by Bardeen & Petterson (1975), ranging between 15 and 30 gravitational radii for a hole with Kerr parameter $`a=1.`$ In the warped regions of the disc with changing inclination, the expected vertical shear (Papaloizou & Pringle 1983) was seen in the simulations. We found that these velocities were limited to be subsonic, presumably by nonlinear effects such as shocks. An interesting issue is the the stability of this shear flow. Parametric instability associated with bending waves is a possibility (Papaloizou & Terquem 1995; Gammie, Goodman, & Ogilvie 1998, private communication). This may produce a level of turbulence in a disc with low viscosity. But it should be noted that the SPH simulations presented here are for discs that are already viscous, such that the growth rate of any instability is exceeded by the viscous damping. Simulations with a viscosity small enough to investigate shear flow instability require too many particles to be practicable at present. Models were considered both at high and low inclination. In the nonlinear high inclination case, although the transition radius remained approximately in the same location, the transition was more abrupt than in the low inclination case, indicative of a tendency for the outer part of the disc to become disconnected from the inner part. The implications of this are that a disc which is forced to maintain a nonlinear warp due to severe misalignment will tend to break into two or more disconnected pieces, rather than maintain a smoothly warped structure in which the perturbed horizontal motions remain transonic. Although the calculations performed here are in a very different physical regime to that expected for discs around AGN, the time scale we obtain for black hole – disc alignment is in essential agreement with the ideas of Rees (1978) and accordingly the later work of Scheuer & Feiler (1996), if the viscous diffusion coefficients acting in and out of the disc plane are taken to be equal. We also found that this alignment time scale could be significantly shortened by the inclusion of Einstein precession, in line with the predictions of a linear analysis.
warning/0001/math0001154.html
ar5iv
text
# I Introduction ## I Introduction The purpose of this paper is two-fold. One is to study irreducible highest weight representations and q-vertex operators of the quantum affine superalgebra $`U_q[\widehat{sl(N|1)}]`$, $`N>2`$. Another one is to apply these results to bosonize the multi-component super $`tJ`$ model on an infinite lattice. We shall adapt the bosonization technique initiated in , which turns out to be very powerful in constructing highest weight representations and q-vertex operators. Recently, free bosonic realizations of the level-one representations and “elementary” q-vertex operators have been obtained for $`U_q[\widehat{sl(M|N)}]`$, $`MN`$ and $`U_q[\widehat{gl(N|N)}]`$ . However, these free boson representations are not irreducible in general. Moreover, the elementary q-vertex operators obtained in were determined solely from their commutation relations with the bosonized Drinfeld generators of the relevant algebras, and thus one can ask on which representations these bosonized q-vertex operators act. To construct irreducible highest weight representations and q-vertex operators acting on them, we need to study in details the structure of the bosonic Fock space generated by the free boson fields. This has been done for $`U_q[\widehat{sl(2|1)}]`$ and $`U_q[\widehat{gl(N|N)}]`$, $`N2`$ . In this paper we treat the $`U_q[\widehat{sl(N|1)}]`$ ($`N>2`$) case. Irreducible highest weight representations and bosonized q-vertex operators acting on them play an essential role in the algebraic analysis method of lattice integrable models, which was invented by the Kyoto group and collaborators . In this approach, the following assumption is the vital key : $`\mathrm{`}\mathrm{`}\mathrm{𝐭𝐡𝐞}\mathrm{𝐩𝐡𝐲𝐬𝐢𝐜𝐚𝐥}\mathrm{𝐬𝐩𝐚𝐜𝐞}\mathrm{𝐨𝐟}\mathrm{𝐬𝐭𝐚𝐭𝐞𝐬}\mathrm{𝐨𝐟}\mathrm{𝐭𝐡𝐞}\mathrm{𝐦𝐨𝐝𝐞𝐥}\mathrm{"}={\displaystyle \underset{\alpha ,\alpha ^{}}{}}V(\lambda _\alpha )V(\lambda _\alpha ^{})^S`$ (I.1) where $`V(\lambda _\alpha )`$ is the level-one irreducible highest weight module of the underlying quantum affine algebras and $`V(\lambda _\alpha )^S`$ is the dual module of $`V(\lambda _\alpha )`$. By this method, various integrable models have been analysed such as the higher spin XXZ chains , the higher rank cases , the twisted $`A_2^{(2)}`$ case , and the face type statistical models . Spin chain models with quantum superalgebra symmetries have been the focus of recent studies in the context of strongly correlated fermion systems . It is natural to generalize the algebraic analysis method to treat super spin chains on an infinite lattice. In , the q-deformed supersymmetric $`tJ`$ model which has $`U_q^{}[\widehat{sl(2|1)}]`$ as its non-abelian symmetry has been analysed. However, the super case is fundamentally different from the non-super case. Unlike the latter, $`U_q[\widehat{sl(2|1)}]`$ has infinite number of level-one irreducible highest weight representations and the bosonized q-vertex operators act in all of them. This leads to the assumption that for the $`q`$-deformed supersymmetric $`t`$-$`J`$ model $`\alpha `$, $`\alpha ^{}`$ in (I.1) take infinite number of integer values. In this paper we extend the work to treat the multi-component $`tJ`$ model with $`U_q^{}[\widehat{sl(N|1)}]`$ ($`N>2`$) symmetry. As we shall see, the level-one irreducible highest weight representations of $`U_q[\widehat{sl(N|1)}]`$ ($`N>2`$) have similar structures as the $`N=2`$ case. So we shall make the assumption that the physical space of states of the multi-component $`tJ`$ model on an infinite lattice is of the form (I.1) with $`\alpha ,\alpha ^{}`$ being any integers. This paper is organized as follows. After presenting some neccessary preliminaries, we in section 3 construct the level-one irreducible highest weight representations of $`U_q[\widehat{sl(N|1)}]`$ and calculate their (super)characters by means of the BRST resolution. In section 4, we compute the exchange relations of the q-vertex operators and show that they form the graded Faddeev-Zamolodchikov algebra. In section 5, we consider the application of these results to the multi-component super $`tJ`$ model on an infinite lattice. Generalizing the Kyoto group’s work , we give the bosonization of this model using the bosonized vertex operators of $`U_q[\widehat{sl(N|1)}]`$. Finally, we compute the one-point correlation functions of the local operators and give an integral expression of the correlation functions. ## II Preliminaries ### II.1 Quantum affine superalgebra $`U_q[\widehat{sl(N|1)}]`$ Let us introduce orthonormal basis $`\{ϵ_i^{}|i=1,2,\mathrm{},N+1\}`$ with the bilinear form $`(ϵ_i^{},ϵ_j^{})=\nu _i\delta _{ij}`$, where $`\nu _i=1`$ for $`iN+1`$ and $`\nu _{N+1}=1`$. The classical fundamental weights are defined by $`\overline{\mathrm{\Lambda }}_i=_{j=1}^iϵ_j`$ ($`i=1,2,\mathrm{},N`$), with $`ϵ_i=ϵ_i^{}\frac{\nu _i}{N1}_{j=1}^{N+1}ϵ_j^{}`$. Introduce the affine weight $`\mathrm{\Lambda }_0`$ and the null root $`\delta `$ having $`(\mathrm{\Lambda }_0,ϵ_i^{})=(\delta ,ϵ_i^{})=0`$ for $`i=1,2,\mathrm{},N+1`$ and $`(\mathrm{\Lambda }_0,\mathrm{\Lambda }_0)=(\delta ,\delta )=0`$, $`(\mathrm{\Lambda }_0,\delta )=1`$. The affine simple roots and fundamental weights are given by $`\alpha _i=\nu _iϵ_i^{}\nu _{i+1}ϵ_{i+1}^{},i=1,2,\mathrm{},N,\alpha _0=\delta {\displaystyle \underset{i=1}{\overset{N}{}}}\alpha _i,`$ $`\mathrm{\Lambda }_0=\mathrm{\Lambda }_0,\mathrm{\Lambda }_i=\mathrm{\Lambda }_0+\overline{\mathrm{\Lambda }}_i,i=1,2,\mathrm{},N.`$ (II.1) The Cartan matrix of the affine superalgebra $`\widehat{sl(N|1)}`$ reads as $`(a_{ij})=\left(\begin{array}{ccccccc}0& 1& & & & 1& \\ 1& 2& 1& & & & \\ & 1& 2& \mathrm{}& & & \\ & & \mathrm{}& \mathrm{}& \mathrm{}& & \\ & & & 1& 2& 1& \\ 1& & & & 1& 0& \end{array}\right)(i,j=0,1,2,\mathrm{},N).`$ (II.8) The Quantum affine superalgebra $`U_q[\widehat{sl(N|1)}]`$ is a $`q`$-analogue of the universal enveloping algebra of $`\widehat{sl(N|1)}`$ generated by the Chevalley generators $`\{e_i,f_i,q^{h_i},d|i=0,1,2,\mathrm{},N\}`$, where $`d`$ is the usual derivation operator. The $`Z_2`$-grading of the generators are $`[e_0]=[f_0]=[e_N]=[f_N]=1`$ and zero otherwise. The defining relations are $`[h_i,h_j]=0,h_id=dh_i,[d,e_i]=\delta _{i,0}e_i,[d,f_i]=\delta _{i,0}f_i,`$ $`q^{h_i}e_jq^{h_i}=q^{a_{ij}}e_j,q^{h_i}f_jq^{h_i}=q^{a_{ij}}f_j,[e_i,f_j]=\delta _{ij}{\displaystyle \frac{q^{h_i}q^{h_i}}{qq^1}},`$ $`[e_i,e_j]=[f_i,f_j]=0,\mathrm{for}a_{ij}=0,`$ $`[e_j,[e_j,e_i]_{q^1}]_q=0,[f_j,[f_j,f_i]_{q^1}]_q=0,\mathrm{for}|a_{ij}|=1,j0,N.`$ Here and throughout, $`[a,b]_xab(1)^{[a][b]}xba`$ and $`[a,b][a,b]_1`$. We do not write down the extra $`q`$-Serre relations which can be obtained by using Yamane’s Dynkin diagram procedure . $`U_q[\widehat{sl(N|1)}]`$ is a $`𝐙_2`$-graded quasi-triangular Hopf algebra endowed with the following coproduct $`\mathrm{\Delta }`$, counit $`ϵ`$ and antipode $`S`$: $`\mathrm{\Delta }(h_i)`$ $`=`$ $`h_i1+1h_i,\mathrm{\Delta }(d)=d1+1d,`$ $`\mathrm{\Delta }(e_i)`$ $`=`$ $`e_i1+q^{h_i}e_i,\mathrm{\Delta }(f_i)=f_iq^{h_i}+1f_i,`$ $`ϵ(e_i)`$ $`=`$ $`ϵ(f_i)=ϵ(h)=0,`$ $`S(e_i)`$ $`=`$ $`q^{h_i}e_i,S(f_i)=f_iq^{h_i},S(h)=h,`$ (II.9) where $`i=0,1,\mathrm{},N`$. Notice that the antipode $`S`$ is a $`𝐙_2`$-graded algebra anti-homomorphism. Namely, for any homogeneous elements $`a,bU_q[\widehat{sl(N|1)}]`$ $`S(ab)=(1)^{[a][b]}S(b)S(a)`$, which extends to inhomogeneous elements through linearity. Moreover, $$S^2(a)=q^{2\rho }aq^{2\rho },aU_q[\widehat{sl(N|1)}],$$ (II.10) where $`\rho `$ is an element in the Cartan subalgebra such that $`(\rho ,\alpha _i)=(\alpha _i,\alpha _i)/2`$ for any simple root $`\alpha _i,i=0,1,2,\mathrm{},N`$. Explicitly, $$\rho =(N1)d+\overline{\rho }=(N1)d+\frac{1}{2}\underset{k=1}{\overset{N}{}}(N2k)ϵ_k^{}\frac{1}{2}Nϵ_{N+1}^{},$$ (II.11) which $`\overline{\rho }`$ is the half-sum of positive roots of $`sl(N|1)`$ . The multiplication rule on the tensor products is $`𝐙_2`$-graded: $`(ab)(a^{}b^{})=(1)^{[b][a^{}]}(aa^{}bb^{})`$ for any homogeneous elements $`a,b,a^{},b^{}U_q[\widehat{sl(N|1)}]`$. $`U_q[\widehat{sl(N|1)}]`$ can also be realized in terms of the Drinfeld generators $`\{X_m^{\pm ,i},H_n^i,q^{\pm H_0^i}`$, $`c,d|m𝐙,n𝐙\{0\},i=1,2,\mathrm{},N\}`$. The $`𝐙_2`$-grading of the Drinfeld generators is given by $`[X_m^{\pm ,N}]=1`$ for $`m𝐙`$ and zero otherwise. The relations satisfied by the Drinfeld generators read $`[c,a]=[d,H_0^i]=[H_0^i,H_n^j]=0,[d,H_n^i]=nH_n^i,aU_q[\widehat{sl(N|1)}]`$ $`[d,X_n^{\pm ,i}]=nX_n^{\pm ,i},q^{H_0^j}X_n^{\pm ,i}q^{H_0^j}=q^{\pm a_{ij}}X_n^{\pm ,i},`$ $`[H_n^i,H_m^j]=\delta _{n+m,0}{\displaystyle \frac{[a_{ij}n]_q[nc]_q}{n}},[H_n^i,X_m^{\pm ,j}]=\pm {\displaystyle \frac{[a_{ij}n]_q}{n}}X_{n+m}^{\pm ,j}q^{\pm |n|c/2},`$ $`[X_n^{+,i},X_m^{,j}]={\displaystyle \frac{\delta _{ij}}{qq^1}}\left(q^{\frac{c}{2}(nm)}\psi _{n+m}^{+,i}q^{\frac{c}{2}(nm)}\psi _{n+m}^{,i}\right),`$ $`[X_n^{\pm ,i},X_m^{\pm ,j}]=0,\mathrm{for}a_{ij}=0,`$ $`[X_{n+1}^{\pm ,i},X_m^{\pm ,j}]_{q^{\pm a_{ij}}}[X_{m+1}^{\pm ,j},X_n^{\pm ,i}]_{q^{\pm a_{ij}}}=0,\mathrm{for}a_{ij}0,`$ $`Sym_{l,m}[X_l^{\pm ,i},[X_m^{\pm ,i},X_n^{\pm ,j}]_{q^1}]_q=0,\mathrm{for}a_{ij}=0,iN,`$ (II.12) where $`_{n𝐙}\psi _n^{\pm ,j}z^n=q^{\pm H_0^j}\mathrm{exp}\left(\pm (qq^1)_{n>0}H_{\pm n}^jz^n\right)`$ , and the symbol $`Sym_{k,l}`$ means symmetrization with respect to $`k`$ and $`l`$. We used the standard notation $`[x]_q=(q^xq^x)/(qq^1)`$. The Chevalley generators are related to the Drinfeld generators by the formulas: $`h_i=H_0^i,e_i=X_0^{+,i},f_i=X_0^{,i},i=1,2,\mathrm{},N,h_0=c{\displaystyle \underset{k=1}{\overset{N}{}}}H_0^k,`$ $`e_0=[X_0^{,N},[X_0^{,N1},\mathrm{},[X_0^{,2},X_1^{,1}]_{q^1}]_{q^1}\mathrm{}]_{q^1}q^{_{k=1}^NH_0^k},`$ $`f_0=q^{_{k=1}^NH_0^k}[[\mathrm{}[[X_1^{+,1},X_0^{+,2}]_q,\mathrm{},X_0^{+,N1}]_q,X_0^{+,N}]_q.`$ (II.13) ### II.2 Free Bosonic realization of the quantum affine superalgebra $`U_q[\widehat{sl(N|1)}]`$ at level one Introduce bosonic oscillators $`\{a_n^i,b_n,c_n`$, $`Q_{a^i},Q_b,Q_c`$ $`|n𝐙,i=1,2,\mathrm{},N,\}`$ which satisfy the commutation relations $`[a_n^i,a_m^j]=\delta _{n+m,0}\delta _{ij}{\displaystyle \frac{[n]_q[n]_q}{n}},[a_0^i,Q_{a^j}]=\delta _{ij},`$ $`[b_n,b_m]=\delta _{n+m,0}{\displaystyle \frac{[n]_q^2}{n}},[b_0,Q_b]=1,`$ $`[c_n,c_m]=\delta _{n+m,0}{\displaystyle \frac{[n]_q^2}{n}},[c_0,Q_c]=1.`$ (II.14) The remaining commutation relations are zero. Define $`\{h_m^i|i=1,2,\mathrm{},N,m𝐙\}`$ : $`h_m^i=a_m^iq^{|m|/2}a^{i+1}q^{|m|/2},Q_{h_i}=Q_{a^i}Q_{a^{i+1}},i=1,2,\mathrm{},N1,`$ $`h_m^N=a_m^Nq^{|m|/2}+b_mq^{|m|/2},Q_{h_N}=Q_{a^N}+Q_b.`$ (II.15) Let us introduce the notation $`h^j(z;\kappa )=Q_{h_j}+h_0^j\mathrm{ln}z_{n0}\frac{h_n^j}{[n]_q}q^{\kappa |n|}z^n`$. The bosonic fields $`c(z;\beta )`$, $`b(z;\beta )`$ and $`h_j^{}(z;\beta )`$ are defined in the same way. Define the Drinfeld currents, $`X^{\pm ,i}(z)=_{n𝐙}X_n^{\pm ,i}z^{n1}`$, $`i=1,2,\mathrm{},N`$, and the q-differential operator $`_zf(z)=\frac{f(qz)f(q^1z)}{(qq^1)z}`$. Then, the Drinfeld generators of $`U_q[\widehat{sl(N|1)}]`$ at level one can be realized by the free boson fields as $`c=1,H_m^i=h_m^i,X^{+,N}(z)=:e^{h^N(z;\frac{1}{2})}e^{c(z;0)}:e^{\sqrt{1}\pi _{i=1}^{N1}a_0^i},`$ $`X^{,N}(z)=:e^{h^N(z;\frac{1}{2})}_z\{e^{c(z;0)}\}:e^{\sqrt{1}\pi _{i=1}^{N1}a_0^i},`$ $`X^{\pm ,i}(z)=\pm :e^{\pm h^i(z;\frac{1}{2})}:e^{\pm \sqrt{1}\pi a_0^i},i=1,2,\mathrm{},N1.`$ (II.16) ### II.3 Bosonization of level-one vertex operators In order to construct the vertex operators of $`U_q[\widehat{sl(N|1)}]`$, we firstly consider the level-zero representations (i.e. the evaluation representations) of $`U_q[\widehat{sl(N|1)}]`$. Let $`E_{i,j}`$ be the $`(N+1)\times (N+1)`$ matrix whose $`(i,j)`$-element is unity and zero elsewhere. Let $`\{v_1,v_2,\mathrm{},v_{N+1}\}`$ be the basis vectors of the (N+1)-dimensional graded vector space $`V`$. The $`𝐙_2`$-grading of these basis vectors is chosen to be $`[v_i]=(\nu _i+1)/2`$. The $`(N+1)`$-dimensional level-zero representation $`V_z`$ of $`U_q[\widehat{sl(N|1)}]`$ is given by $`e_i=E_{i,i+1},f_i=\nu _iE_{i+1,i},t_i=q^{\nu _iE_{i,i}\nu _{i+1}E_{i+1,i+1}},`$ $`e_0=zE_{N+1,1},f_0=z^1E_{1,N+1},t_0=q^{E_{1,1}E_{N+1,N+1}},`$ (II.17) where $`i=1,\mathrm{},N`$. Let $`V_z^S`$ be the left dual module of $`V_z`$, defined by $$\pi _{V_z^S}(a)=\pi _{V_z}(S(a))^{st},aU_q[\widehat{sl(N|1)}],$$ (II.18) where $`st`$ denotes the supertansposition. Now, we study the level-one vertex operators of $`U_q[\widehat{sl(N|1)}]`$. Let $`V(\lambda )`$ be the highest weight $`U_q[\widehat{sl(N|1)}]`$-module with the highest weight $`\lambda `$ and the highest weight vector $`|\lambda >`$. Consider the following intertwiners of $`U_q[\widehat{sl(N|1)}]`$-modules : $`\mathrm{\Phi }_\lambda ^{\mu V}(z):V(\lambda )V(\mu )V_z,\mathrm{\Phi }_\lambda ^{\mu V^{}}(z):V(\lambda )V(\mu )V_z^S,`$ $`\mathrm{\Psi }_\lambda ^{V\mu }(z):V(\lambda )V_zV(\mu ),\mathrm{\Psi }_\lambda ^{V^{}\mu }(z):V(\lambda )V_z^SV(\mu ).`$ (II.19) They are intertwiners in the sense that for any $`xU_q[\widehat{sl(N|1)}]`$ $$\mathrm{\Xi }(z)x=\mathrm{\Delta }(x)\mathrm{\Xi }(z),\mathrm{\Xi }(z)=\mathrm{\Phi }_\lambda ^{\mu V}(z),\mathrm{\Phi }_\lambda ^{\mu V^{}}(z),\mathrm{\Psi }_\lambda ^{V\mu }(z),\mathrm{\Psi }_\lambda ^{V^{}\mu }(z).$$ (II.20) We expand the vertex operators as $`\mathrm{\Phi }_\lambda ^{\mu V}(z)={\displaystyle \underset{j=1}{\overset{N}{}}}\mathrm{\Phi }_{\lambda ,j}^{\mu V}(z)v_j,\mathrm{\Phi }_\lambda ^{\mu V^{}}(z)={\displaystyle \underset{j=1}{\overset{N}{}}}\mathrm{\Phi }_{\lambda ,j}^{\mu V^{}}(z)v_j^{},`$ $`\mathrm{\Psi }_\lambda ^{V\mu }(z)={\displaystyle \underset{j=1}{\overset{N}{}}}v_j\mathrm{\Psi }_{\lambda ,j}^{V\mu }(z),\mathrm{\Psi }_\lambda ^{V^{}\mu }(z)={\displaystyle \underset{j=1}{\overset{N}{}}}v_j^{}\mathrm{\Psi }_{\lambda ,j}^{V^{}\mu }(z).`$ (II.21) The intertwiners are even, which implies $`[\mathrm{\Phi }_{\lambda ,j}^{\mu V}(z)]=[\mathrm{\Phi }_{\lambda ,j}^{\mu V^{}}(z)]=[\mathrm{\Psi }_{\lambda ,j}^{V\mu }(z)]=[\mathrm{\Psi }_{\lambda ,j}^{V^{}\mu }(z)]=[v_j]=\frac{\nu _j+1}{2}`$. According to , $`\mathrm{\Phi }_\lambda ^{\mu V}(z)\left(\mathrm{\Phi }_\lambda ^{\mu V^{}}(z)\right)`$ is called type I (dual) vertex operator and $`\mathrm{\Psi }_\lambda ^{V\mu }(z)\left(\mathrm{\Psi }_\lambda ^{V^{}\mu }(z)\right)`$ type II (dual) vertex operator. Introduce the bosonic operators $`\varphi _j(z)`$, $`\varphi _j^{}(z)`$, $`\psi _j(z)`$ and $`\psi _j^{}(z)`$ : $`\varphi _{N+1}(z)=:e^{h_N^{}(q^Nz;\frac{1}{2})}e^{c(q^Nz;0)}(q^Nz)^{\frac{N2}{2(N1)}}:e^{\sqrt{1}\pi _{i=1}^N\frac{1i}{N1}a_0^i},`$ $`\nu _l\varphi _l(z)(1)^{[f_l]([v_l]+[v_{l+1}])}=[\varphi _{l+1}(z),f_l]_{q^{\nu _{l+1}}},`$ $`\varphi _1^{}(z)=:e^{h_1^{}(qz;\frac{1}{2})}(q^Nz)^{\frac{N2}{2(N1)}}:e^{\sqrt{1}\pi _{i=1}^N\frac{1i}{N1}a_0^i},`$ $`\nu _lq^{\nu _l}\varphi _{l+1}^{}(z)(1)^{[f_l]([v_l]+[v_{l+1}])}=[\varphi _l^{}(z),f_l]_{q^{\nu _l}},`$ $`\psi _1(z)=:e^{h_1^{}(qz;\frac{1}{2})}(q^Nz)^{\frac{N2}{2(N1)}}:e^{\sqrt{1}\pi _{i=1}^N\frac{1i}{N1}a_0^i},`$ $`\psi _{l+1}(z)=[\psi _l(z),e_l]_{q^{\nu _l}},`$ $`\psi _{N+1}^{}(z)=:e^{h_N^{}(q^{2N}z;\frac{1}{2})}_z\{e^{c(q^{2N}z;0)}\}(q^Nz)^{\frac{N2}{2(N1)}}:e^{\sqrt{1}\pi _{i=1}^N\frac{1i}{N1}a_0^i},`$ $`\nu _l\nu _{l+1}q^{\nu _l}\psi _l^{}(z)=[\psi _{l+1}^{}(z),e_l]_{q^{\nu _{l+1}}},`$ (II.22) where $`h_n^i={\displaystyle \underset{j=1}{\overset{N}{}}}{\displaystyle \frac{[\alpha _{ij}m]_q[\beta _{ij}m]_q}{[(N1)m]_q[m]_q}}h_n^j,Q_{h^i}^{}={\displaystyle \underset{j=1}{\overset{N}{}}}{\displaystyle \frac{\alpha _{ij}\beta _{ij}}{N1}}Q_{h^j},h_0^i={\displaystyle \underset{j=1}{\overset{N}{}}}{\displaystyle \frac{\alpha _{ij}\beta _{ij}}{N1}}h^j,`$ with $`\alpha _{ij}=min(i,j)`$, and $`\beta _{ij}=N1max(i,j)`$. Define the even operators $`\varphi (z),\varphi ^{}(z),\psi (z)`$ and $`\psi ^{}(z)`$ by $`\varphi (z)=_{j=1}^{N+1}\varphi _j(z)v_j`$, $`\varphi ^{}(z)=_{j=1}^{N+1}\varphi _j^{}(z)v_j^{}`$, $`\psi (z)=_{j=1}^{N+1}v_j\psi _j(z)`$ and $`\psi ^{}(z)=_{j=1}^{N+1}v_j^{}\psi _j^{}(z)`$. Then the vertex operators $`\mathrm{\Phi }_\lambda ^{\mu V}(z),\mathrm{\Phi }_\lambda ^{\mu V^{}}(z),\mathrm{\Psi }_\lambda ^{V\mu }(z)`$ and $`\mathrm{\Psi }_\lambda ^{V^{}\mu }(z)`$, if they exist, are bosonized by $`\varphi (z),\varphi ^{}(z),\psi (z)`$ and $`\psi ^{}(z)`$, respectively . We remark that our vertex operators differ from those of Kimura et al by a scalar factor $`(q^Nz)^{\frac{N2}{2(N1)}}`$, which is needed in order for the vertex operators also satisfy (II.20) for the element $`x=d`$. $`\varphi (z),\varphi ^{}(z),\psi (z)`$ and $`\psi ^{}(z)`$ are referred to as the “elementary q-vertex operators” of $`U_q[\widehat{sl(N|1)}]`$. ## III Highest weight $`U_q[\widehat{sl(N|1)}]`$-modules We begin by defining the Fock module. Denote by $`F_{\lambda _1,\lambda _2,\mathrm{},\lambda _{N+1};\lambda _{N+2}}`$ the bosonic Fock space generated by $`a_m^i,b_m,c_m(m>0)`$ over the vector $`|\lambda _1,\lambda _2,\mathrm{},\lambda _{N+1};\lambda _{N+2}>`$: $`F_{\lambda _1,\lambda _2,\mathrm{},\lambda _{N+1};\lambda _{N+2}}=𝐂[a_1^i,a_2^i,\mathrm{};b_1,b_2,\mathrm{};c_1,c_2,\mathrm{}]|\lambda _1,\lambda _2,\mathrm{},\lambda _{N+1};\lambda _{N+2}>,`$ where $`|\lambda _1,\lambda _2,\mathrm{},\lambda _{N+1};\lambda _{N+2}>=e^{_{i=1}^N\lambda _iQ_{a^i}+\lambda _{N+1}Q_b+\lambda _{N+2}Q_c}|0>.`$ The vacuum vector $`|0>`$ is defined by $`a_m^i|0>=b_m|0>=c_m|0>=0`$ for $`i=1,2,\mathrm{},N`$, and $`m0`$. Obviously, $`a_m^i|\lambda _1,\lambda _2,\mathrm{},\lambda _{N+1};\lambda _{N+2}>=0,\mathrm{for}i=1,2,\mathrm{},N\mathrm{and}m>0,`$ $`b_m|\lambda _1,\lambda _2,\mathrm{},\lambda _{N+1};\lambda _{N+2}>=c_m|\lambda _1,\lambda _2,\mathrm{},\lambda _{N+1};\lambda _{N+2}>=0,\mathrm{for}m>0.`$ To obtain the highest weight vectors of $`U_q[\widehat{sl(N|1)}]`$, we impose the conditions: $`e_i|\lambda _1,\mathrm{},\lambda _{N+1};\lambda _{N+2}>=0,i=0,1,2,\mathrm{},N,`$ $`h_i|\lambda _1,\mathrm{},\lambda _{N+1};\lambda _{N+2}>=\lambda ^i|\lambda _1,\mathrm{},\lambda _{N+1};\lambda _{N+2}>,i=0,1,2,\mathrm{},N.`$ (III.1) Solving these equations, we obtain two classess of solutions: 1. $`(\lambda _1,\mathrm{},\lambda _i,\lambda _{i+1},\mathrm{},\lambda _{N+1};\lambda _{N+2})=(\beta +1,\mathrm{},\underset{i,i+1}{\underset{}{\beta +1,\beta }},\mathrm{},\beta ;0)`$, where $`i=1,\mathrm{},N`$, and $`\beta `$ is arbitrary. It follows that $`(\lambda ^0,\lambda ^1,\mathrm{},\lambda ^i,\lambda ^{i+1}\mathrm{},\lambda ^N)`$=$`(0,0,\mathrm{},\underset{i1,i,i+1}{\underset{}{0,\mathrm{\hspace{0.33em}1},\mathrm{\hspace{0.33em}0}}},\mathrm{},0)`$ and we have the identification $`|\mathrm{\Lambda }_i>`$=$`|\beta +1,\mathrm{},\underset{i,i+1}{\underset{}{\beta +1,\beta }},\mathrm{},\beta ;0>`$. 2. $`(\lambda _1,\mathrm{},\lambda _N,\lambda _{N+1};\lambda _{N+2})=(\beta ,\mathrm{},\beta ,\beta \alpha ;\alpha )`$, where $`\alpha `$, $`\beta `$ are arbitrary. We have $`(\lambda ^0,\lambda ^1,\mathrm{},\lambda ^{N1},\lambda ^N)`$=$`(1\alpha ,0,\mathrm{},0,\alpha )`$ and $`|(1\alpha )\mathrm{\Lambda }_0+\alpha \mathrm{\Lambda }_N>`$=$`|\beta ,\mathrm{},\beta ,\beta \alpha ;\alpha >`$. Associated to the above two classes of solutions are the following Fock spaces: $`_\beta ^m={\displaystyle \underset{\{i_1,\mathrm{},i_N\}𝐙}{}}F_{\beta +1+i_1,\beta +1i_1+i_2,\mathrm{},\beta +1i_{m1}+i_m,\beta i_m+i_{m+1},\mathrm{},\beta i_{N1}+i_N,\beta +i_N;i_N},`$ $`_{(\alpha ;\beta )}={\displaystyle \underset{\{i_1,\mathrm{},i_N\}𝐙}{}}F_{\beta +i_1,\beta i_1+i_2,\mathrm{},\beta i_{N1}+i_N,\beta \alpha +i_N;\alpha +i_N},`$ where $`m=1,2,\mathrm{},N`$, and it should be understood that $`i_00`$. However, it is easily seen that $`_\beta ^m=F_{(m;\beta )}`$, $`m=1,\mathrm{},N`$. Thus, it is sufficient to study the Fock space $`_{(\alpha ;\beta )}`$. In the following we shall also restrict ourselves to the $`\alpha 𝐙`$ case. It can be shown that the bosonized action of $`U_q[\widehat{sl(N|1)}]`$ (II.16) on $`_{(\alpha ;\beta )}`$ is closed : $`U_q[\widehat{sl(N|1)}]_{(\alpha ;\beta )}=_{(\alpha ;\beta )}.`$ Hence each Fack space $`_{(\alpha ;\beta )}`$ constitutes a $`U_q[\widehat{sl(N|1)}]`$-module. However, these modules are not irreducible in general. To obtain irreducible subspaces, we introduce a pair of ghost fields $`\eta (z)={\displaystyle \underset{n𝐙}{}}\eta _nz^{n1}=:e^{c(z)}:,\xi (z)={\displaystyle \underset{n𝐙}{}}\xi _nz^n=:e^{c(z)}:.`$ The mode expansion of $`\eta (z)`$ and $`\xi (z)`$ is well defined on $`_{(\alpha ;\beta )}`$ for $`\alpha 𝐙`$, and the modes satisfy the relations $`\xi _m\xi _n+\xi _n\xi _m=\eta _m\eta _n+\eta _n\eta _m=0,\xi _m\eta _n+\eta _n\xi _m=\delta _{m+n,0}.`$ (III.2) Since $`\eta _0\xi _0`$ and $`\xi _0\eta _0`$ qualify as projectors, we use them to decompose $`_{(\alpha ;\beta )}`$ into a direct sum $`_{(\alpha ;\beta )}=\eta _0\xi _0_{(\alpha ;\beta )}\xi _0\eta _0_{(\alpha ;\beta )}`$ for $`\alpha 𝐙`$. $`\eta _0\xi _0_{(\alpha ;\beta )}`$ is referred to as $`Ker_{\eta _0}`$ and $`\xi _0\eta _0_{(\alpha ;\beta )}=_{(\alpha ;\beta )}/\eta _0\xi _0_{(\alpha ;\beta )}`$ as $`Coker_{\eta _0}`$. Since $`\eta _0`$ commutes (or anticommutes) with the bosonized action of $`U_q[\widehat{sl(N|1)}]`$, $`Ker_{\eta _0}`$ and $`Coker_{\eta _0}`$ are both $`U_q[\widehat{sl(N|1)}]`$-modules for $`\alpha 𝐙`$. ### III.1 Character and supercharacter We want to determine the character and supercharacter formulae of the $`U_q[\widehat{sl(N|1)}]`$-modules constructed in the bosonic Fock space. We first of all bosonize the derivation operator $`d`$ as $`d={\displaystyle \underset{m1}{}}{\displaystyle \frac{m^2}{[m]_q^2}}\{{\displaystyle \underset{i=1}{\overset{N}{}}}h_m^ih_m^i+c_mc_m\}{\displaystyle \frac{1}{2}}\{{\displaystyle \underset{i=1}{\overset{N}{}}}h_0^ih_0^i+c_0(c_0+1)\}.`$ (III.3) It obeys the commutation relations $`[d,h_i]=0,[d,h_m^i]=mh_m^i,[d,X_m^{\pm ,i}]=mX_m^{\pm ,i},i=1,2,\mathrm{},N,`$ as required. Moreover, $`[d,\xi _0]=[d,\eta _0]=0`$. The character and supercharacter of a $`U_q[\widehat{sl(N|1)}]`$-module $`M`$ are defined by $`Ch_M(q;x_1,x_2,\mathrm{},x_N)`$ $`=`$ $`tr_M(q^dx_1^{h_1}x_2^{h_2}\mathrm{}x_N^{h_N}),`$ $`Sch_M(q;x_1,x_2,\mathrm{},x_N)`$ $`=`$ $`Str_M(q^dx_1^{h_1}x_2^{h_2}\mathrm{}x_N^{h_N})`$ (III.4) $`=`$ $`tr_M((1)^{N_f}q^dx_1^{h_1}x_2^{h_2}\mathrm{}x_N^{h_N}),`$ respectively. The Fermi-number operaor $`N_f`$ can be bosonized as $`N_f=\{\begin{array}{cc}(N1)b_0\hfill & ifNeven,i.e.N=2L\hfill \\ L(_{k=1}^Na_0^ib_0)+c_0\hfill & ifNodd,i.e.N=2L+1\hfill \end{array}.`$ (III.7) Indeed, $`N_f`$ satisfies $`(1)^{N_f}\mathrm{\Theta }(z)=(1)^{[\mathrm{\Theta }(z)]}\mathrm{\Theta }(z)(1)^{N_f},`$ where $`\mathrm{\Theta }(z)=X^{\pm ,i}(z)`$, $`\varphi _i(z)`$, $`\varphi _i^{}(z)`$, $`\psi _i(z)`$ and $`\psi _i^{}(z)`$. We calculate the characters and supercharacters by using the BRST resolution . Let us define the Fock spaces, for $`l𝐙`$ $`_{(\alpha ;\beta )}^{(l)}={\displaystyle \underset{\{i_1,\mathrm{},i_N\}𝐙}{}}F_{\beta +i_1,\beta i_1+i_2,\mathrm{},,\beta i_{N1}+i_N,\beta \alpha +i_N;\alpha +i_N+l}.`$ We have $`_{(\alpha ;\beta )}^{(0)}=_{(\alpha ;\beta )}`$. It can be shown that $`\eta _0`$ and $`\xi _0`$ intertwine these Fock spaces as follows: $`\eta _0:_{(\alpha ;\beta )}^{(l)}_{(\alpha ;\beta )}^{(l+1)},\xi _0:_{(\alpha ;\beta )}^{(l)}_{(\alpha ;\beta )}^{(l1)}.`$ We have the following BRST complexes: $`\begin{array}{ccccccc}\mathrm{}& \stackrel{Q_{l1}=\eta _0}{}& _{(\alpha ;\beta )}^{(l)}& \stackrel{Q_l=\eta _0}{}& _{(\alpha ;\beta )}^{(l+1)}& \stackrel{Q_{l+1}=\eta _0}{}& \mathrm{}\\ & & |𝐎& & |𝐎& & \\ \mathrm{}& \stackrel{Q_{l1}=\eta _0}{}& _{(\alpha ;\beta )}^{(l)}& \stackrel{Q_l=\eta _0}{}& _{(\alpha ;\beta )}^{(l+1)}& \stackrel{Q_{l+1}=\eta _0}{}& \mathrm{}\end{array}`$ (III.11) where $`𝐎`$ is an operator such that $`_{(\alpha ;\beta )}^{(l)}_{(\alpha ;\beta )}^{(l)}`$. Noting the fact that $`\eta _0\xi _0+\xi _0\eta _0=1`$, and $`\eta _0\xi _0`$ ($`\xi _0\eta _0`$) is the projection operator from $`_{(\alpha ;\beta )}^{(l)}`$ to $`Ker_{Q_l}`$ ($`Coker_{Q_l}`$), we get $`Ker_{Q_l}=Im_{Q_{l1}},\mathrm{for}\mathrm{any}l𝐙,`$ $`tr(𝐎)|_{Ker_{Q_l}}=tr(𝐎)|_{Im_{Q_{l1}}}=tr(𝐎)|_{Coker_{Q_{l1}}}.`$ (III.12) By the above results, we can write the trace over $`Ker`$ or $`Coker`$ as the sum of trace over $`_{(\alpha ;\beta )}^{(l)}`$, and compute the latter by using the technique introduced in . The results are $`Ch_{Ker_{_{(\alpha ;\beta )}}}(q;x_1,\mathrm{},x_N)`$ $`=`$ $`{\displaystyle \frac{q^{\frac{1}{2}\alpha (\alpha 1)}}{_{n=1}^{\mathrm{}}(1q^n)^{N+1}}}{\displaystyle \underset{l=1}{\overset{\mathrm{}}{}}}(1)^{l+1}q^{\frac{1}{2}\{l^2+l(2\alpha 1)\}}`$ $`\times {\displaystyle \underset{\{i_1,\mathrm{},i_N\}𝐙}{}}q^{\frac{1}{2}\{i_N^2+i_N(12\alpha 2l)\}}q^{\frac{1}{2}\mathrm{\Delta }(i_1,\mathrm{}.i_N)}`$ $`\times x_1^{2i_1i_2}x_2^{2i_2i_1i_3}\mathrm{}x_{N1}^{2i_{N1}i_Ni_{N2}}x_N^{\alpha i_N},`$ $`Ch_{Coker_{_{(\alpha ;\beta )}}}(q;x_1,\mathrm{},x_N)`$ $`=`$ $`{\displaystyle \frac{q^{\frac{1}{2}\alpha (\alpha 1)}}{_{n=1}^{\mathrm{}}(1q^n)^{N+1}}}{\displaystyle \underset{l=1}{\overset{\mathrm{}}{}}}(1)^{l+1}q^{\frac{1}{2}\{l^2+l(12\alpha )\}}`$ $`\times {\displaystyle \underset{\{i_1,\mathrm{},i_N\}𝐙}{}}q^{\frac{1}{2}\{i_N^2+i_N(12\alpha +2l)\}}q^{\frac{1}{2}\mathrm{\Delta }(i_1,\mathrm{}.i_N)}`$ $`\times x_1^{2i_1i_2}x_2^{2i_2i_1i_3}\mathrm{}x_{N1}^{2i_{N1}i_Ni_{N2}}x_N^{\alpha i_N},`$ where $`\mathrm{\Delta }(i_1,\mathrm{},i_N)=_{l,l^{}=1}^N\frac{\alpha _{ll^{}}\beta _{ll^{}}}{N1}\lambda _{i_1,\mathrm{},i_N}^l\lambda _{i_1,\mathrm{},i_N}^l^{}`$ and $`\{\begin{array}{cc}\lambda _{i_1,\mathrm{},i_N}^l=2i_li_{l1}i_{l+1},\hfill & 2lN1\hfill \\ \lambda _{i_1,\mathrm{},i_N}^1=2i_1i_2,\hfill & \lambda _{i_1,\mathrm{},i_N}^N=\alpha i_N\hfill \end{array}.`$ (III.15) Similary, the supercharacters of $`Ker_{_{(\alpha ;\beta )}}`$ and $`Coker_{_{(\alpha ;\beta )}}`$ are given by 1. For $`N=2L`$: $`Sch_{Ker_{_{(\alpha ;\beta )}}}(q;x_1,\mathrm{},x_N)`$ $`=`$ $`{\displaystyle \frac{(1)^\alpha q^{\frac{1}{2}\alpha (\alpha 1)}}{_{n=1}^{\mathrm{}}(1q^n)^{N+1}}}{\displaystyle \underset{l=1}{\overset{\mathrm{}}{}}}(1)^{l+1}q^{\frac{1}{2}\{l^2+l(2\alpha 1)\}}`$ $`\times {\displaystyle \underset{\{i_1,\mathrm{},i_N\}𝐙}{}}(1)^{i_N}q^{\frac{1}{2}\{i_N^2+i_N(12\alpha 2l)\}}q^{\frac{1}{2}\mathrm{\Delta }(i_1,\mathrm{}.i_N)}`$ $`\times x_1^{2i_1i_2}x_2^{2i_2i_1i_3}\mathrm{}x_{N1}^{2i_{N1}i_Ni_{N2}}x_N^{\alpha i_N},`$ $`Sch_{Coker_{_{(\alpha ;\beta )}}}(q;x_1,\mathrm{},x_N)`$ $`=`$ $`{\displaystyle \frac{(1)^\alpha q^{\frac{1}{2}\alpha (\alpha 1)}}{_{n=1}^{\mathrm{}}(1q^n)^{N+1}}}{\displaystyle \underset{l=1}{\overset{\mathrm{}}{}}}(1)^{l+1}q^{\frac{1}{2}\{l^2+l(12\alpha )\}}`$ $`\times {\displaystyle \underset{\{i_1,\mathrm{},i_N\}𝐙}{}}(1)^{i_N}q^{\frac{1}{2}\{i_N^2+i_N(12\alpha +2l)\}}q^{\frac{1}{2}\mathrm{\Delta }(i_1,\mathrm{}.i_N)}`$ $`\times x_1^{2i_1i_2}x_2^{2i_2i_1i_3}\mathrm{}x_{N1}^{2i_{N1}i_Ni_{N2}}x_N^{\alpha i_N},`$ 2. For $`N=2L+1`$: $`Sch_{Ker_{_{(\alpha ;\beta )}}}(q;x_1,\mathrm{},x_N)`$ $`=`$ $`{\displaystyle \frac{(1)^{(L+1)\alpha }q^{\frac{1}{2}\alpha (\alpha 1)}}{_{n=1}^{\mathrm{}}(1q^n)^{N+1}}}{\displaystyle \underset{l=1}{\overset{\mathrm{}}{}}}q^{\frac{1}{2}\{l^2+l(2\alpha 1)\}}`$ $`\times {\displaystyle \underset{\{i_1,\mathrm{},i_N\}𝐙}{}}(1)^{i_N}q^{\frac{1}{2}\{i_N^2+i_N(12\alpha 2l)\}}q^{\frac{1}{2}\mathrm{\Delta }(i_1,\mathrm{}.i_N)}`$ $`\times x_1^{2i_1i_2}x_2^{2i_2i_1i_3}\mathrm{}x_{N1}^{2i_{N1}i_Ni_{N2}}x_N^{\alpha i_N},`$ $`Sch_{Coker_{_{(\alpha ;\beta )}}}(q;x_1,\mathrm{},x_N)`$ $`=`$ $`{\displaystyle \frac{(1)^{(L+1)\alpha }q^{\frac{1}{2}\alpha (\alpha 1)}}{_{n=1}^{\mathrm{}}(1q^n)^{N+1}}}{\displaystyle \underset{l=1}{\overset{\mathrm{}}{}}}q^{\frac{1}{2}\{l^2+l(12\alpha )\}}`$ $`\times {\displaystyle \underset{\{i_1,\mathrm{},i_N\}𝐙}{}}(1)^{i_N}q^{\frac{1}{2}\{i_N^2+i_N(12\alpha +2l)\}}q^{\frac{1}{2}\mathrm{\Delta }(i_1,\mathrm{}.i_N)}`$ $`\times x_1^{2i_1i_2}x_2^{2i_2i_1i_3}\mathrm{}x_{N1}^{2i_{N1}i_Ni_{N2}}x_N^{\alpha i_N}.`$ Since $`_{(\alpha (N1);\beta +1)}^{(1)}=_{(\alpha ;\beta )}`$ and by (III.12), we have $`Ch_{Coker_{_{(\alpha (N1);\beta +1)}}}=Ch_{Ker_{_{(\alpha ;\beta )}}},Sch_{Coker_{_{(\alpha (N1);\beta +1)}}}=Sch_{Ker_{_{(\alpha ;\beta )}}}.`$ (III.16) Relations (III.16) can also be checked by using the above explicit formulae of the (super)characters. ### III.2 $`U_q[\widehat{sl(N|1)}]`$-module structure of $`_{(\alpha ;\beta \frac{1}{N1}\alpha )}`$ Set $`\lambda _\alpha =(1\alpha )\mathrm{\Lambda }_0+\alpha \mathrm{\Lambda }_N`$ and $`|\lambda _\alpha >=|\beta ,\mathrm{},\beta ,\beta \alpha ;\alpha >_{(\alpha ;\beta )},\alpha 𝐙,`$ $`|\mathrm{\Lambda }_m>=|\beta +1,\mathrm{},\beta +1,\beta ,\mathrm{},\beta ;0>_{(m;\beta )},m=1,\mathrm{},N,`$ The above vectors play the role of the highest weight vectors of $`U_q[\widehat{sl(N|1)}]`$-modules. one can check that $`\{\begin{array}{cc}\eta _0|\lambda _\alpha >=0,\hfill & \mathrm{for}\alpha =0,1,\mathrm{}\hfill \\ \eta _o|\mathrm{\Lambda }_m>=0,\hfill & \mathrm{for}m=1,\mathrm{},N\hfill \\ \eta _0|\lambda _\alpha >0,\hfill & \mathrm{for}\alpha =1,2,\mathrm{}\hfill \end{array}.`$ (III.20) It follows that the modules $`Coker_{_{(\alpha ;\beta )}}(\alpha =1,2,\mathrm{}),Ker_{_{(\alpha ;\beta )}}(\alpha =0,1,2,\mathrm{}),`$ $`Ker_{_{(m;\beta )}}(m=1,2,\mathrm{},N),`$ are highest weight $`U_q[\widehat{sl(N|1)}]`$-modules. Denote them by $`\overline{V}(\lambda _\alpha )`$ and $`\overline{V}(\mathrm{\Lambda }_m)`$, respectively. From (III.20) and (III.16), we have the following identifications of the highest weight $`U_q[\widehat{sl(N|1)}]`$-modules: $`\overline{V}(\lambda _\alpha )`$ $``$ $`Ker_{_{(\alpha ;\beta \frac{1}{N1}\alpha )}}Coker_{_{(\alpha (N1);\beta \frac{1}{N1}\alpha +1)}},\mathrm{for}\alpha =0,1,2,\mathrm{},`$ (III.21) $``$ $`Coker_{_{(\alpha ;\beta \frac{1}{N1}\alpha )}}Ker_{_{(\alpha +(N1);\beta \frac{1}{N1}\alpha 1)}},\mathrm{for}\alpha =1,2,\mathrm{},`$ $`\overline{V}(\mathrm{\Lambda }_m)`$ $``$ $`Ker_{_{(m;\beta \frac{1}{N1}m)}}Coker_{_{(m(N1);\beta \frac{1}{N1}m+1)}},\mathrm{for}m=1,\mathrm{},N.`$ (III.22) It is easy to see that the vertex operators (II.22) also commute (or anti-commute) with $`\eta _0`$. It follows from (III.21)-(III.22) that each Fock space $`_{(\alpha ;\beta \frac{1}{N1}\alpha )}`$ is decomposed into a direct sum of the highest weight $`U_q[\widehat{sl(N|1)}]`$-modules: $`\begin{array}{cccc}& Ker& & Coker\\ .& .& & .\\ .& .& & .\\ F_{(N;\beta +1+\frac{1}{N1})}=& \overline{V}(\lambda _N)& & \overline{V}(\lambda _1)\\ & \varphi (z)\varphi ^{}(z)& & \varphi (z)\varphi ^{}(z)\\ F_{(N+1;\beta +1)}=& \overline{V}(\lambda _{N+1})& & \overline{V}(\mathrm{\Lambda }_0)\\ & \varphi (z)\varphi ^{}(z)& & \varphi (z)\varphi ^{}(z)\\ F_{(N+2;\beta +1\frac{1}{N1})}=& \overline{V}(\lambda _{N+2})& & \overline{V}(\mathrm{\Lambda }_1)\\ & \varphi (z)\varphi ^{}(z)& & \varphi (z)\varphi ^{}(z)\\ .& .& & .\\ .& .& & .\\ F_{(2;\beta +1\frac{N3}{N1})}=& \overline{V}(\lambda _2)& & \overline{V}(\mathrm{\Lambda }_{N3})\\ & \varphi (z)\varphi ^{}(z)& & \varphi (z)\varphi ^{}(z)\\ F_{(1;\beta +1\frac{N2}{N1})}=& \overline{V}(\lambda _1)& & \overline{V}(\mathrm{\Lambda }_{N2})\\ & \varphi (z)\varphi ^{}(z)& & \varphi (z)\varphi ^{}(z)\\ F_{(0;\beta )}=& \overline{V}(\mathrm{\Lambda }_0)& & \overline{V}(\mathrm{\Lambda }_{N1})\\ & \varphi (z)\varphi ^{}(z)& & \varphi (z)\varphi ^{}(z)\\ F_{(1;\beta \frac{1}{N1})}=& \overline{V}(\mathrm{\Lambda }_1)& & \overline{V}(\mathrm{\Lambda }_N)\\ & \varphi (z)\varphi ^{}(z)& & \varphi (z)\varphi ^{}(z)\\ F_{(2;\beta \frac{2}{N1})}=& \overline{V}(\mathrm{\Lambda }_2)& & \overline{V}(\lambda _2)\\ & \varphi (z)\varphi ^{}(z)& & \varphi (z)\varphi ^{}(z)\\ .& .& & .\\ .& .& & .\\ F_{(N2;\beta \frac{N2}{N1})}=& \overline{V}(\mathrm{\Lambda }_{N2})& & \overline{V}(\lambda _{N2})\\ & \varphi (z)\varphi ^{}(z)& & \varphi (z)\varphi ^{}(z)\\ F_{(N1;\beta 1)}=& \overline{V}(\mathrm{\Lambda }_{N1})& & \overline{V}(\mathrm{\Lambda }_{N1})\\ & \varphi (z)\varphi ^{}(z)& & \varphi (z)\varphi ^{}(z)\\ F_{(N;\beta 1\frac{1}{N1})}=& \overline{V}(\mathrm{\Lambda }_N)& & \overline{V}(\lambda _N)\\ & \varphi (z)\varphi ^{}(z)& & \varphi (z)\varphi ^{}(z)\\ F_{(N+1;\beta 1\frac{2}{N1})}=& \overline{V}(\lambda _2)& & \overline{V}(\lambda _{N+1})\\ .& .& & .\\ .& .& & .\end{array}.`$ (III.55) It is expected that $`\overline{V}(\lambda _\alpha )`$ ($`\alpha 𝐙`$) and $`\overline{V}(\mathrm{\Lambda }_m)`$ ($`m=1,2,\mathrm{},N1`$) are irreducible highest weight $`U_q[\widehat{sl(N|1)}]`$-modules with the highest weights $`\lambda _\alpha `$ and $`\mathrm{\Lambda }_m`$, respectively. Thus we conjecture that $`\overline{V}(\lambda _\alpha )=V(\lambda _\alpha ),\overline{V}(\mathrm{\Lambda }_m)=V(\mathrm{\Lambda }_m).`$ (III.56) ## IV Exchange Relations of Vertex Operators In this section, we derive the exchange relations of the type I and type II bosonized vertex operators of $`U_q[\widehat{sl(N|1)}]`$. As expected, these vertex operators satisfy the graded Faddeev-Zamolodchikov algebra. ### IV.1 The R-matrix Throughout, we use the abbreviation $`(z;x_1,\mathrm{},x_m)_{\mathrm{}}={\displaystyle \underset{\{n_1,\mathrm{},n_m\}=0}{\overset{\mathrm{}}{}}}(1zx_1^{n_1}\mathrm{}x_m^{n_m}),`$ $`\{z\}_{\mathrm{}}\stackrel{def}{=}(z;q^{2(N1)},q^{2(N1)})_{\mathrm{}}.`$ (IV.1) Let $`\overline{R}(z)End(VV)`$ be the R-matrix of $`U_q[\widehat{sl(N|1)}]`$, $`\overline{R}(z)(v_iv_j)={\displaystyle \underset{k,l=1}{\overset{2N}{}}}\overline{R}_{kl}^{ij}(z)v_kv_l,v_i,v_j,v_k,v_lV,`$ (IV.2) where the matrix elements of $`\overline{R}(z)`$ are given by $`\overline{R}_{i,i}^{i,i}(z)=1,\overline{R}_{N+1,N+1}^{N+1,N+1}(z)={\displaystyle \frac{zq^1q}{zqq^1}},i=1,2,\mathrm{},N,`$ $`\overline{R}_{ij}^{ij}(z)={\displaystyle \frac{z1}{zqq^1}},ij,`$ $`\overline{R}_{ij}^{ji}(z)={\displaystyle \frac{qq^1}{zqq^1}}(1)^{[i][j]},i<j,`$ $`\overline{R}_{ij}^{ji}(z)={\displaystyle \frac{(qq^1)z}{zqq^1}}(1)^{[i][j]},i>j,`$ $`\overline{R}_{kl}^{ij}(z)=0,\mathrm{otherwise}.`$ Define the R-matrices $`R^{(I)}(z)`$ and $`R^{(II)}(z)`$ by $`R^{(I)}(z)=r(z)\overline{R}(z),R^{(II)}(z)=\overline{r}(z)\overline{R}(z),`$ (IV.3) where $`r(z)=z^{\frac{2N}{N1}}{\displaystyle \frac{(zq^2;q^{2(N1)})_{\mathrm{}}(z^1q^{2N2};q^{2(N1)})_{\mathrm{}}}{(z^1q^2;q^{2(N1)})_{\mathrm{}}(zq^{2N2};q^{2(N1)})_{\mathrm{}}}},`$ $`\overline{r}(z)=z^{\frac{1}{N1}}{\displaystyle \frac{(zq^{2N4};q^{2(N1)})_{\mathrm{}}(z^1q^{2N2};q^{2(N1)})_{\mathrm{}}}{(z^1q^{2N4};q^{2(N1)})_{\mathrm{}}(zq^{2N2};q^{2(N1)})_{\mathrm{}}}}.`$ These R-matrices satisfy the graded Yang-Baxter equation on $`VVV`$: $`R_{12}^{(i)}(z)R_{13}^{(i)}(zw)R_{23}^{(i)}(w)=R_{23}^{(i)}(w)R_{13}^{(i)}(zw)R_{12}^{(i)}(z),i=I,II.`$ Moreover, they enjoy (i) the initial condition $`R^{(i)}(1)=P,i=I,II`$, where $`P`$ is the graded permutation operator; (ii) the unitarity condition $`R_{12}^{(i)}(\frac{z}{w})R_{21}^{(i)}(\frac{w}{z})=1,i=I,II`$, where $`R_{21}^{(i)}(z)=PR_{12}^{(i)}(z)P`$; (iii) the crossing-unitarity $`(R^{(i)})^{1,st_1}(z)\left((q^{2\overline{\rho }}1)R^{(i)}(zq^{2(1N)})(q^{2\overline{\rho }}1)\right)^{st_1}=1,i=I,II,`$ where $`q^{2\overline{\rho }}`$ $``$ $`diag(q^{2\rho _1},q^{2\rho _2},\mathrm{},q^{2\rho _N},q^{2\rho _{N+1}})`$ $`=`$ $`diag(q^{N2},q^{N4},\mathrm{},q^N,q^N).`$ The various supertranspositions of the R-matrix are given by $`(R^{st_1}(z))_{ij}^{kl}=R_{kj}^{il}(z)(1)^{[i]([i]+[k])},(R^{st_2}(z))_{ij}^{kl}=R_{il}^{kj}(z)(1)^{[j]([l]+[j])},`$ $`(R^{st_{12}}(z))_{ij}^{kl}=R_{kl}^{ij}(z)(1)^{([i]+[j])([i]+[j]+[k]+[l])}=R_{kl}^{ij}(z).`$ ### IV.2 The graded Faddeev-Zamolodchikov algebra We now calculate the exchange relations of the type I and type II bosonic vertex operators of $`U_q[\widehat{sl(N|1)}]`$. Define $`{\displaystyle 𝑑zf(z)}=Res(f)=f_1,\mathrm{for}\mathrm{a}\mathrm{formal}\mathrm{series}\mathrm{function}f(z)={\displaystyle \underset{n𝐙}{}}f_nz^n.`$ Then, the Chevalley generators of $`U_q[\widehat{sl(N|1)}]`$ can be expressed by the integrals $`e_i={\displaystyle 𝑑zX^{+,i}(z)},f_i={\displaystyle 𝑑zX^{,i}(z)},i=1,2,\mathrm{},N.`$ One can also get the integral expressions of the bosonic vertex operators $`\varphi (z)`$, $`\varphi ^{}(z)`$, $`\psi (z)`$ and $`\psi ^{}(z)`$ . Using these integral expressions and the relations given in appendices A and B, we find that the bosonic vertex operators defined in (II.22) satisfy the graded Faddeev-Zamolodchikov algebra $`\varphi _j(z_2)\varphi _i(z_1)={\displaystyle \underset{k,l=1}{\overset{N+1}{}}}R^{(I)}({\displaystyle \frac{z_1}{z_2}})_{ij}^{kl}\varphi _k(z_1)\varphi _l(z_2)(1)^{[i][j]},`$ $`\psi _i^{}(z_1)\psi _j^{}(z_2)={\displaystyle \underset{k,l=1}{\overset{N+1}{}}}R^{(II)}({\displaystyle \frac{z_1}{z_2}})_{kl}^{ij}\psi _l^{}(z_2)\psi _k^{}(z_1)(1)^{[i][j]},`$ $`\psi _i^{}(z_1)\varphi _j(z_2)=\tau ({\displaystyle \frac{z_1}{z_2}})\varphi _j(z_2)\psi _i^{}(z_1)(1)^{[i][j]},`$ (IV.4) where $`\tau (z)=z^{\frac{2N}{N1}}{\displaystyle \frac{(zq;q^{2(N1)})_{\mathrm{}}(z^1q^{2N3};q^{2(N1)})_{\mathrm{}}}{(z^1q;q^{2(N1)})_{\mathrm{}}(zq^{2N3};q^{2(N1)})_{\mathrm{}}}}.`$ By $`:e^{h_N^{}(zq^N;\frac{1}{2})+h_1^{}(zq;\frac{1}{2})h^1(zq^2;\frac{1}{2})h^2(zq^3;\frac{1}{2})\mathrm{}h^N(zq^{N+1};\frac{1}{2})}:=1,`$ we obtain the first invertibility relations $`\varphi _i(z)\varphi _j^{}(z)=g^1(1)^{[i]}\delta _{ij},{\displaystyle \underset{k=1}{\overset{N+1}{}}}(1)^{[k]}\varphi _k^{}(z)\varphi _k(z)=g^1,`$ (IV.5) and the second invertibility relations $`\varphi _i^{}(zq^{2(N1)})\varphi _j(z)=g^1q^{2\rho _i}\delta _{ij},{\displaystyle \underset{k=1}{\overset{N+1}{}}}q^{2\rho _k}\varphi _k(z)\varphi _k^{}(zq^{2(N1)})=g^1,`$ (IV.6) where $`g=e^{\frac{\sqrt{1}\pi N}{2(N1)}}\frac{(q^2;q^{2(N1)})_{\mathrm{}}}{(q^{2(N1)};q^{2(N1)})_{\mathrm{}}}`$. Using the fact that $`\eta _0\xi _0`$ is a projection operator, we can make the following identifications: $`\mathrm{\Phi }_i(z)=\eta _0\xi _0\varphi _i(z)\eta _0\xi _0,\mathrm{\Phi }_i^{}(z)=\eta _0\xi _0\varphi _i^{}(z)\eta _0\xi _0,`$ $`\mathrm{\Psi }_i(z)=\eta _0\xi _0\psi _i(z)\eta _0\xi _0,\mathrm{\Psi }_i^{}(z)=\eta _0\xi _0\psi _i^{}(z)\eta _0\xi _0.`$ (IV.7) Set $`\mu _\alpha =\{\begin{array}{cc}\mathrm{\Lambda }_\alpha ,\hfill & \alpha =0,1,\mathrm{},N\hfill \\ \lambda _{\alpha (N1)},\hfill & \mathrm{for}\alpha >N\hfill \\ \lambda _\alpha ,\hfill & \mathrm{for}\alpha <0\hfill \end{array}.`$ (IV.11) It is easy to see that the vertex operators $`\varphi (z)`$, $`\varphi ^{}(z)`$, $`\psi (z)`$ and $`\psi ^{}(z)`$ commute (or anti-commute) with the BRST charge $`\eta _0`$. It follows from (III.55) and (III.56) that the vertex operators (IV.7) intertwine all the level-one irreducible highest weight $`U_q[\widehat{sl(N|1)}]`$-modules $`V(\mu _\alpha )`$ ($`\alpha 𝐙`$) as follows $`\mathrm{\Phi }(z):V(\mu _\alpha )V(\mu _{\alpha 1})V_z,\mathrm{\Phi }^{}(z):V(\mu _\alpha )V(\mu _{\alpha +1})V_z^S,`$ $`\mathrm{\Psi }(z):V(\mu _\alpha )V_zV(\mu _{\alpha 1}),\mathrm{\Psi }^{}(z):V(\mu _\alpha )V_z^SV(\mu _{\alpha +1}).`$ (IV.12) ¿From (IV.4), we have $`\mathrm{\Phi }_j(z_2)\mathrm{\Phi }_i(z_1)={\displaystyle \underset{k,l=1}{\overset{N+1}{}}}R^{(I)}({\displaystyle \frac{z_1}{z_2}})_{ij}^{kl}\mathrm{\Phi }_k(z_1)\mathrm{\Phi }_l(z_2)(1)^{[i][j]},`$ $`\mathrm{\Psi }_i^{}(z_1)\mathrm{\Psi }_j^{}(z_2)={\displaystyle \underset{k,l=1}{\overset{N+1}{}}}R^{(II)}({\displaystyle \frac{z_1}{z_2}})_{kl}^{ij}\mathrm{\Psi }_l^{}(z_2)\mathrm{\Psi }_k^{}(z_1)(1)^{[i][j]},`$ $`\mathrm{\Psi }_i^{}(z_1)\mathrm{\Phi }_j(z_2)=\tau ({\displaystyle \frac{z_1}{z_2}})\mathrm{\Phi }_j(z_2)\mathrm{\Psi }_i^{}(z_1)(1)^{[i][j]}.`$ (IV.13) Moreover, we have the following invertibility relations: $`\mathrm{\Phi }_i(z)\mathrm{\Phi }_j^{}(z)=g^1(1)^{[i]}\delta _{ij}id_{V(\mu _\alpha )},`$ $`{\displaystyle \underset{k=1}{\overset{N+1}{}}}(1)^{[k]}\mathrm{\Phi }_k^{}(z)\mathrm{\Phi }_k(z)=g^1id_{V(\mu _\alpha )},`$ $`\mathrm{\Phi }_i^{}(zq^{2(N1)})\mathrm{\Phi }_j(z)=g^1q^{2\rho _i}\delta _{ij}id_{V(\mu _\alpha )},`$ $`{\displaystyle \underset{k=1}{\overset{N+1}{}}}q^{2\rho _k}\mathrm{\Phi }_k(z)\mathrm{\Phi }_k^{}(zq^{2(N1)})=g^1id_{V(\mu _\alpha )}.`$ (IV.14) ## V Multi-component super $`t`$-$`J`$ model In this section, we give a mathematical definition of the multi-component super $`t`$-$`J`$ model on an infinite lattice. ### V.1 Space of states By means of the R-matrix (IV.2) of $`U_q[\widehat{sl(N|1)}]`$, one defines a spin chain model, referred to as the multi-component super $`tJ`$ model, on the infinte lattice $`\mathrm{}VVV\mathrm{}`$. Let $`h`$ be the operator on $`VV`$ such that $`P\overline{R}({\displaystyle \frac{z_1}{z_2}})=1+uh+\mathrm{},u0,`$ $`P:\mathrm{the}\mathrm{graded}\mathrm{permutation}\mathrm{operator},e^u{\displaystyle \frac{z_1}{z_2}}.`$ The Hamiltonian $`H`$ of this model is given by $`H={\displaystyle \underset{lZ}{}}h_{l+1,l}.`$ (V.1) $`H`$ acts formally on the infinite tensor product, $$\mathrm{}VVV\mathrm{}.$$ (V.2) It can be easily checked that $`[U_q^{}(\widehat{sl}(N|1)),H]=0,`$ where $`U_q^{}[\widehat{sl}(N|1)]`$ is the subalgebra of $`U_q[\widehat{sl(N|1)}]`$ with the derivation operator $`d`$ being dropped. So $`U_q^{}[\widehat{sl}(N|1)]`$ plays the role of infinite dimensional non-abelian symmetry of the multi-component super $`t`$-$`J`$ model on the infinite lattice. ¿From the intertwining relation (IV.12), one have the following composition of the type I vertex operators: $`V(\mu _\alpha )\stackrel{\mathrm{\Phi }(1)}{}V(\mu _{\alpha 1})V\stackrel{\mathrm{\Phi }(1)id}{}V(\mu _{\alpha 1})VV\stackrel{\mathrm{\Phi }(1)idid}{}\mathrm{}W_l,`$ (V.3) where $`W_l\stackrel{def}{=}\mathrm{}VV`$, i.e the left half-infinite tensor product. We conjecture that such a composition converges to a map : $`i:V(\mu _\alpha )W_l.`$ Such a map $`i`$ satisfies $`i(xv)=\mathrm{\Delta }^{(\mathrm{})}(x)i(v)`$, $`xU_q[\widehat{sl(N|1)}]`$ and $`vV(\mu _\alpha )`$. Following , we could replace the infinite tensor product (V.2) by the level-zero $`U_q[\widehat{sl(N|1)}]`$-module, $`F_{\alpha \alpha ^{}}=\mathrm{Hom}(V(\mu _\alpha ),V(\mu _\alpha ^{}))V(\mu _\alpha )V(\mu _\alpha ^{})^{},`$ where $`V(\mu _\alpha )`$ is level-one irreducible highest weight $`U_q[\widehat{sl(N|1)}]`$-module and $`V(\mu _\alpha ^{})^{}`$ is the dual module of $`V(\mu _\alpha ^{})`$. By (III.55), this homomorphism can be realized by applying the type I vertex operators repeatedly. So we shall make the (hypothetical) identification: $`\mathrm{`}\mathrm{`}\mathrm{𝐭𝐡𝐞}\mathrm{𝐬𝐩𝐚𝐜𝐞}\mathrm{𝐨𝐟}\mathrm{𝐩𝐡𝐲𝐬𝐢𝐜𝐚𝐥}\mathrm{𝐬𝐭𝐚𝐭𝐞𝐬}\mathrm{"}={\displaystyle \underset{\alpha ,\alpha ^{}𝐙}{}}V(\mu _\alpha )V(\mu _\alpha ^{})^{}.`$ Namely, we take $`FEnd({\displaystyle \underset{\alpha 𝐙}{}}V(\mu _\alpha )){\displaystyle \underset{\alpha ,\alpha ^{}𝐙}{}}F_{\alpha \alpha ^{}}`$ as the space of states of the multi-component super $`t`$-$`J`$ model on the infinite lattice. The left action of $`U_q[\widehat{sl(N|1)}]`$ on $`F`$ is defined by $`x.f={\displaystyle x_{(1)}fS(x_{(2)})(1)^{[f][x_{(2)}]}},xU_q[\widehat{sl(N|1)}],fF,`$ where we have used notation $`\mathrm{\Delta }(x)=x_{(1)}x_{(2)}`$. Note that $`F_{\alpha \alpha }`$ has the unique canonical element $`id_{V(\mu _\alpha )}`$. We call it the vacuum and denote it by $`|vac>_\alpha `$. ### V.2 Local structure and local operators Following Jimbo et al , we use the type I vertex operators and their variants to incorporate the local structure into the space of physical states $`F`$, that is to formulate the action of local operators of the multi-component super $`t`$-$`J`$ model on the infinite tensor product (V.2) in terms of their actions on $`F_{\alpha \alpha ^{}}`$. Using the isomorphisms $`\mathrm{\Phi }(1)`$ $`:`$ $`V(\mu _\alpha )V(\mu _{\alpha 1})V,`$ $`\mathrm{\Phi }^{,st}(q^{2(N1)})`$ $`:`$ $`VV(\mu _\alpha )^{}V(\mu _{\alpha 1})^{},`$ (V.4) were $`st`$ is the supertransposition on the quantum space, we have the following identification: $`V(\mu _\alpha )V(\mu _\alpha ^{})^{}\stackrel{}{}V(\mu _{\alpha 1})VV(\mu _\alpha ^{})^{}\stackrel{}{}V(\mu _{\alpha 1})V(\mu _{\alpha ^{}1})^{}.`$ The resulting isomorphism can be identified with the super translation (or shift) operator defined by $`T=g{\displaystyle \underset{i}{}}\mathrm{\Phi }_i(1)\mathrm{\Phi }_i^{,st}(q^{2(N1)})(1)^{[i]}q^{2\rho _i}.`$ Its inverse is given by $`T^1=g{\displaystyle \underset{i}{}}\mathrm{\Phi }_i^{}(1)\mathrm{\Phi }_i^{st}(1).`$ Thus we can define the local operators on $`V`$ as operators on $`F_{\alpha \alpha ^{}}`$ . Let us label the tensor components from the middle as $`1,2,\mathrm{}`$ for the left half and as $`0,1,2,\mathrm{}`$ for the right half. The operators acting on the site 1 are defined by $`E_{ij}\stackrel{def}{=}E_{ij}^{(1)}=g\mathrm{\Phi }_i^{}(1)\mathrm{\Phi }_j(1)(1)^{[j]}id.`$ (V.5) More generally we set $`E_{ij}^{(n)}=T^{(n1)}E_{ij}T^{n1}(nZ).`$ (V.6) Then, from the invertibility relations of the type I vertex operators of $`U_q[\widehat{sl(N|1)}]`$, we can show that the local operators $`E_{ij}^{(n)}`$ acting on $`F_{\alpha \alpha ^{}}`$ satisfy the following relations: $`E_{ij}^{(m)}E_{kl}^{(n)}=\{\begin{array}{cc}\delta _{jk}E_{il}^{(n)}\hfill & \mathrm{if}m=n\hfill \\ (1)^{([i]+[j])([k]+[l])}E_{kl}^{(n)}E_{ij}^{(m)}\hfill & \mathrm{if}mn\hfill \end{array}.`$ This result implies that the local operators $`E_{ij}^{(n)}`$ are nothing but the $`U_q[sl(N|1)]`$ generators acting on the n-th component of $`\mathrm{}VV\mathrm{}`$. They include all the local operators in the multi-component super $`tJ`$ model . As is expected from the physical point of view, the vacuum vectors $`|vac>_\alpha `$ are super-translationally invariant and singlets (i.e. they belong to the trivial representation of $`U_q[\widehat{sl(N|1)}]`$): $`T|vac>_\alpha =|vac>_{\alpha 1},x.|vac>_\alpha =ϵ(x)|vac>_\alpha ,xU_q[\widehat{sl(N|1)}].`$ This is proved as follow. Let $`u_l^{(\alpha )}`$ ($`u_l^{(\alpha )}`$) be a basis vectors of $`V(\mu _\alpha )(V(\mu _\alpha )^{})`$ and $`|vac>_\alpha \stackrel{def}{=}id_{V(\mu _\alpha )}={\displaystyle \underset{l}{}}u_l^{(\alpha )}u_l^{(\alpha )}.`$ Then $`T|vac>_\alpha =g{\displaystyle \underset{m,l}{}}q^{2\rho _m}\mathrm{\Phi }_m(1)u_l^{(\alpha )}\mathrm{\Phi }_m^{,st}(q^{2(N1)})u_l^{(\alpha )}(1)^{[m]+[l][m]}.`$ We want to show $`T|vac>_\alpha =|vac>_{\alpha 1}`$. This is equivalent to proving $`g{\displaystyle \underset{m,l}{}}q^{2\rho _m}\mathrm{\Phi }_m(1)u_l^{(\alpha )}\mathrm{\Phi }_m^{,st}(q^{2(N1)})u_l^{(\alpha )}(v)(1)^{[m]+[l][m]}=v,vV(\mu _{\alpha 1}).`$ Now $`l.h.s`$ $`=`$ $`g{\displaystyle \underset{m,l}{}}q^{2\rho _m}\mathrm{\Phi }_m(1)u_l^{(\alpha )}u_l^{(\alpha )}(\mathrm{\Phi }_m^{}(q^{2(N1)})^{st})^{st}v)(1)^{[m]}`$ $`=`$ $`g{\displaystyle \underset{m,l}{}}q^{2\rho _m}\mathrm{\Phi }_m(1)u_l^{(\alpha )}u_l^{(\alpha )}(\mathrm{\Phi }_m^{}(q^{2(N1)})v)`$ $`=`$ $`g{\displaystyle \underset{m}{}}q^{2\rho _m}\mathrm{\Phi }_m(1)\mathrm{\Phi }_m^{}(q^{2(N1)})v=v,`$ where we have used $`(\mathrm{\Phi }_m^{}(z)^{st})^{st}=\mathrm{\Phi }_m^{}(z)(1)^{[m]}`$ and (IV.14). As to the second equation, we have $`x|vac>_\alpha `$ $`=`$ $`{\displaystyle x_{(1)}u_l^{(\alpha )}x_{(2)}u_l^{(\alpha )}(1)^{[l][x_{(2)}]}}`$ $`=`$ $`{\displaystyle x_{(1)}u_l^{(\alpha )}\pi _{V(\mu _\alpha )^{}}(x_{(2)})_{ml}u_m^{(\alpha )}(1)^{[l][x_{(2)}]}}`$ $`=`$ $`{\displaystyle x_{(1)}u_l^{(\alpha )}\pi _{V(\mu _\alpha )}(S(x_{(2)}))_{lm}u_m^{(\alpha )}}`$ $`=`$ $`{\displaystyle x_{(1)}\pi _{V(\mu _\alpha )}(S(x_{(2)}))_{lm}u_l^{(\alpha )}u_m^{(\alpha )}}`$ $`=`$ $`{\displaystyle x_{(1)}S(x_{(2)})u_m^{(\alpha )}u_m^{(\alpha )}}=ϵ(x)|vac>_\alpha .`$ This completes the proof. For any local operator $`OF`$, its vacuum expectation value is defined by $`{}_{\alpha }{}^{}<vac|O|vac>_\alpha \stackrel{def}{=}{\displaystyle \frac{tr_{V(\mu _\alpha )}(q^{2\rho }O)}{tr_{V(\mu _\alpha )}(q^{2\rho })}}={\displaystyle \frac{tr_{V(\mu _\alpha )}(q^{2(N1)d2h_{\overline{\rho }}}O)}{tr_{V(\mu _\alpha )}(q^{2(N1)d2h_{\overline{\rho }}})}},`$ (V.8) where $`2h_{\overline{\rho }}={\displaystyle \underset{l=1}{\overset{N}{}}}l(N1l)h_l.`$ We shall denote the correlator $`{}_{\alpha }{}^{}<vac|O|vac>_\alpha `$ by $`<O>_\alpha `$. ## VI Correlation functions The aim of this section is to calculate $`<E_{mn}>_\alpha `$. The generalization to the calculation of the multi-point functions is straightforward. Set $`P_n^m(z_1,z_2|q|\alpha )={\displaystyle \frac{tr_{V(\mu _\alpha )}(q^{2(N1)d2h_{\overline{\rho }}}\mathrm{\Phi }_m^{}(z_1)\mathrm{\Phi }_n(z_2))}{tr_{V(\mu _\alpha )}(q^{2(N1)d2h_{\overline{\rho }}})}},`$ then $`<E_{mn}>_\alpha =P_n^m(z,z|q|\alpha )`$. By (IV.11), it is sufficient to calculate $`F_{mn}^{(\alpha )}(z_1,z_2)={\displaystyle \frac{tr_{F_{(\alpha ;\beta \alpha )}}(q^{2(N1)d2h_{\overline{\rho }}}\varphi _m^{}(z_1)\varphi _n(z_2)\eta _0\xi _0)}{tr_{F_{(\alpha ;\beta \alpha )}}(q^{2(N1)d2h_{\overline{\rho }}}\eta _0\xi _0)}}.`$ (VI.1) Using the Clavelli-Shapiro technique , we get $`F_{mn}^{(\alpha )}(z_1,z_2)={\displaystyle \frac{\delta _{mn}}{\chi _\alpha }}F_m^{(\alpha )}(z_1,z_2){\displaystyle \frac{\delta _{mn}}{\chi _\alpha }}{\displaystyle \underset{l=1}{\overset{\mathrm{}}{}}}(1)^{l+1}F_{m,l}^{(\alpha )}(z_1,z_2),`$ where $`\chi _\alpha `$ $`=`$ $`Ch_{Ker_{_{(\alpha ;\beta )}}}(q^{2(N1)};q^{(N2)},\mathrm{},q^{l(N1l)},\mathrm{},q^N),`$ $`F_{m,l}^{(\alpha )}(z_1,z_2)`$ $`=`$ $`e^{\frac{\sqrt{1}\pi N}{2(N1)}}C_1^{}C_N^{}(C_1)^{N1}(C_{N+1})^2`$ $`(z_1q)^{\frac{1}{N1}}{\displaystyle \frac{\{\frac{z_1}{z_2}q^{2(N1)}\}_{\mathrm{}}\{\frac{z_2}{z_1}q^{2(N1)}\}_{\mathrm{}}}{\{\frac{z_1}{z_2}q^{2N}\}_{\mathrm{}}\{\frac{z_2}{z_1}q^{2N}\}_{\mathrm{}}}}{\displaystyle 𝑑w_1\mathrm{}𝑑w_N}`$ $`\times \{{\displaystyle \underset{k=1}{\overset{m1}{}}}{\displaystyle \frac{(1q^2)}{qw_{k1}(\frac{w_k}{w_{k1}}q;q^{2(N1)})_{\mathrm{}}(\frac{w_{k1}}{w_k}q;q^{2(N1)})_{\mathrm{}}}}\}`$ $`\times {\displaystyle \frac{1}{w_{m1}(\frac{w_m}{w_{m1}}q;q^{2(N1)})_{\mathrm{}}(\frac{w_{m1}}{w_m}q^{2N1};q^{2(N1)})_{\mathrm{}}}}`$ $`\times \{{\displaystyle \underset{k=m+1}{\overset{N}{}}}{\displaystyle \frac{(1q^2)}{w_k(\frac{w_k}{w_{k1}}q;q^{2(N1)})_{\mathrm{}}(\frac{w_{k1}}{w_k}q;q^{2(N1)})_{\mathrm{}}}}\}`$ $`\times {\displaystyle \underset{\{i_1,\mathrm{},i_N\}𝐙}{}}I_{i_1,\mathrm{},i_N}^{(a,l)}(z_1,z_2|w_1,\mathrm{},w_N)`$ $`\times \{{\displaystyle \frac{(\frac{z_2}{w_N}q^{N1})^{l\alpha +i_N}}{w_Nq(\frac{z_2}{w_N}q^{N1};q^{2(N1)})_{\mathrm{}}(\frac{w_N}{z_2}q^{N1};q^{2(N1)})_{\mathrm{}}}}`$ $`+{\displaystyle \frac{(\frac{z_2}{w_N}q^{N+1})^{l\alpha +i_N}}{z_2q^N(\frac{z_2}{w_N}q^{3N1};q^{2(N1)})_{\mathrm{}}(\frac{w_N}{z_2}q^{N1};q^{2(N1)})_{\mathrm{}}}}\},`$ $`\mathrm{for}m=1,\mathrm{},N,`$ $`F_{N+1,l}^{(\alpha )}(z_1,z_2)`$ $`=`$ $`e^{\frac{\sqrt{1}\pi N}{2(N1)}}C_1^{}C_N^{}(C_1)^N(C_{N+1})^2(z_1q)^{\frac{1}{N1}}{\displaystyle \frac{\{\frac{z_1}{z_2}q^{2(N1)}\}_{\mathrm{}}\{\frac{z_2}{z_1}q^{2(N1)}\}_{\mathrm{}}}{\{\frac{z_1}{z_2}q^{2N}\}_{\mathrm{}}\{\frac{z_2}{z_1}q^{2N}\}_{\mathrm{}}}}`$ $`\times {\displaystyle }dw_1\mathrm{}{\displaystyle }dw_N\{{\displaystyle \underset{k=1}{\overset{N}{}}}{\displaystyle \frac{(1q^2)}{qw_{k1}(\frac{w_k}{w_{k1}}q;q^{2(N1)})_{\mathrm{}}(\frac{w_{k1}}{w_k}q;q^{2(N1)})_{\mathrm{}}}}\}`$ $`\times {\displaystyle \frac{1}{w_N(\frac{z_2}{w_N}q^{N+1};q^{2(N1)})_{\mathrm{}}(\frac{w_N}{z_2}q^{N1};q^{2(N1)})_{\mathrm{}}}}`$ $`\times {\displaystyle \underset{\{i_1,\mathrm{},i_N\}𝐙}{}}I_{i_1,\mathrm{},i_N}^{(a,l)}(z_1,z_2|w_1,\mathrm{},w_N)`$ $`\times _{w_N}\{{\displaystyle \frac{(\frac{z_2}{w_N}q^N)^{l\alpha +i_N}}{w_N(\frac{z_2}{w_N}q^N;q^{2(N1)})_{\mathrm{}}(\frac{w_N}{z_2}q^{N2};q^{2(N1)})_{\mathrm{}}}}\}.`$ In the above equations, $`w_0z_1q`$, and $`I_{i_1,\mathrm{},i_N}^{(a,l)}(z_1,z_2|w_1,\mathrm{},w_N)`$ $`=`$ $`q^{(N1)\alpha (\alpha 1)}(z_1q)^{i_1\frac{\alpha }{N1}}(z_2q^N)^{\frac{N}{N1}\alpha i_N}`$ $`\times q^{(N1)\{l^2+l(12\alpha )+i_N^2+i_N(12\alpha +2l)+\mathrm{\Delta }(i_1,\mathrm{}.i_N)\}}`$ $`\times {\displaystyle \underset{k=1}{\overset{N}{}}}(w_kq^{k(N1k)})^{\lambda _{i_1,\mathrm{},i_N}^k},`$ $`C_1^{}={\displaystyle \frac{\{q^{2N}\}_{\mathrm{}}}{\{q^{4N4}\}_{\mathrm{}}}},C_N^{}={\displaystyle \frac{\{q^{4N2}\}_{\mathrm{}}}{\{q^{2(N1)}\}_{\mathrm{}}}},`$ $`C_1=(q^{2(N1)};q^{2(N1)})_{\mathrm{}}(q^{2N};q^{2(N1)})_{\mathrm{}},C_{N+1}=(q^{2(N1)};q^{2(N1)})_{\mathrm{}}.`$ We now derive the difference equations satisfied by these one-point functions. Noticing that $`x^d\varphi _i(z)x^d=\varphi _i(zx^1),x^d\varphi _i^{}(z)x^d=\varphi _i^{}(zx^1),`$ $`x^d\psi _i(z)x^d=\psi _i(zx^1),x^d\psi _i^{}(z)x^d=\psi _i^{}(zx^1),`$ $`x^d\eta _0x^d=\eta _0,x^d\xi _0x^d=\xi _0,`$ we get the difference equations $`F_m^{(\alpha )}(z_1,z_2q^{2(N1)})=q^{2\rho _m}{\displaystyle \underset{k}{}}R(z_2,z_1)_{mk}^{km}F_k^{(\alpha 1)}(z_1,z_2)(1)^{[m]+[k]+[m][k]}.`$ Since $`\alpha 𝐙`$, it is easily seen that this is a set of infinite number of difference equations. ### Acknowledgements This work has been financially supported by the Australian Research Council Large, Small and QEII Fellowship grants. W.-L. Yang would like to thank Y.-Z. Zhang, and the department of Mathematics, the University of Queensland, for their kind hospitality. W.-L. Yang has also been partially supported by the National Natural Science Foundation of China. ### Appendix A In this appendix, we give the normal ordered relations of the fundamental bosonic fields: $`:e^{h^i(z;\beta _1)}::e^{h^j(w;\beta _2)}:=z^{a_{ij}}(1{\displaystyle \frac{w}{z}}q^{\beta _1+\beta _2})^{a_{ij}}:e^{h^i(z;\beta _1)+h^j(w;\beta _2)}:,ij,`$ $`:e^{h^i(z;\beta _1)}::e^{h^i(w;\beta _2)}:=z^2(1{\displaystyle \frac{w}{z}}q^{\beta _1+\beta _21})(1{\displaystyle \frac{w}{z}}q^{\beta _1+\beta _2+1}):e^{h^i(z;\beta _1)+h^i(w;\beta _2)}:,iN,`$ $`:e^{h^N(z;\beta _1)}::e^{h^N(w;\beta _2)}:=:e^{h^N(z;\beta _1)+h^N(w;\beta _2)}:,`$ $`:e^{h^i(z;\beta _1)}::e^{h_j^{}(w;\beta _2)}:=z^{\delta _{ij}}(1{\displaystyle \frac{w}{z}}q^{\beta _1+\beta _2})^{\delta _{ij}}:e^{h^i(z;\beta _1)+h_j^{}(w;\beta _2)}:,`$ $`:e^{h_i^{}(z;\beta _1)}::e^{h^j(w;\beta _2)}:=z^{\delta _{ij}}(1{\displaystyle \frac{w}{z}}q^{\beta _1+\beta _2})^{\delta _{ij}}:e^{h_i^{}(z;\beta _1)+h^j(w;\beta _2)}:,`$ $`:e^{h_N^{}(z;\beta _1)}::e^{h_N^{}(w;\beta _2)}:=z^{\frac{N}{N1}}{\displaystyle \frac{(\frac{w}{z}q^{\beta _1+\beta _2+2N1};q^{2(N1)})}{(\frac{w}{z}q^{\beta _1+\beta _21};q^{2(N1)})}}:e^{h_N^{}(z;\beta _1)+h_N^{}(w;\beta _2)}:,`$ $`:e^{h_1^{}(z;\beta _1)}::e^{h_1^{}(w;\beta _2)}:=z^{\frac{N2}{N1}}{\displaystyle \frac{(\frac{w}{z}q^{\beta _1+\beta _2+1};q^{2(N1)})}{(\frac{w}{z}q^{\beta _1+\beta _2+2N3};q^{2(N1)})}}:e^{h_1^{}(z;\beta _1)+h_1^{}(w;\beta _2)}:,`$ $`:e^{h_1^{}(z;\beta _1)}::e^{h_N^{}(w;\beta _2)}:=z^{\frac{1}{N1}}{\displaystyle \frac{(\frac{w}{z}q^{\beta _1+\beta _2+N};q^{2(N1)})}{(\frac{w}{z}q^{\beta _1+\beta _2+N2};q^{2(N1)})}}:e^{h_1^{}(z;\beta _1)+h_N^{}(w;\beta _2)}:,`$ $`:e^{h_N^{}(z;\beta _1)}::e^{h_1^{}(w;\beta _2)}:=z^{\frac{1}{N1}}{\displaystyle \frac{(\frac{w}{z}q^{\beta _1+\beta _2+N};q^{2(N1)})}{(\frac{w}{z}q^{\beta _1+\beta _2+N2};q^{2(N1)})}}:e^{h_N^{}(z;\beta _1)+h_1^{}(w;\beta _2)}:,`$ $`:e^{c(z;\beta _1)}::e^{c(w;\beta _2)}:=z(1{\displaystyle \frac{w}{z}}q^{\beta _1+\beta _2}):e^{c(z;\beta _1)+c(w;\beta _2)}:,`$ where $`a_{ij}`$ is the Cartan-matrix of $`\widehat{sl(N|1)}`$ and $`i,j=1,2,\mathrm{},N`$. ### Appendix B By means of the bosonic realization (II.16) of $`U_q[\widehat{sl(N|1)}]`$, the integral expressions of the bosonized vertex operators (II.22) and the technique given in , one can check the following relations * For the type I vertex operators: $`[\varphi _k(z),f_l]=0\mathrm{if}kl,l+1,[\varphi _{l+1}(z),f_l]_{q^{\nu _{l+1}}}=\nu _l\varphi _l(z)(1)^{[f_l]([v_l]+[v_{l+1}])},`$ $`[\varphi _l(z),f_l]_{q^{\nu _l}}=0,[\varphi _l(z),e_l]=q^{h_l}\varphi _{l+1}(z)(1)^{[e_l]([v_l]+[v_{l+1}])},`$ $`[\varphi _k(z),e_l]=0\mathrm{if}kl,q^{h_l}\varphi _l(z)q^{h_l}=q^{\nu _l}\varphi _l(z),`$ $`q^{h_l}\varphi _k(z)q^{h_l}=\varphi _k(z)\mathrm{if}kl,l+1,q^{h_l}\varphi _{l+1}(z)q^{h_l}=q^{\nu _{l+1}}\varphi _{l+1}(z),`$ $`[\varphi _k^{}(z),f_l]=0\mathrm{if}kl,l+1,[\varphi _{l+1}^{}(z),f_l]_{q^{\nu _{l+1}}}=0,`$ $`[\varphi _k^{}(z),e_l]=0\mathrm{if}kl+1,[\varphi _{l+1}^{}(z),e_l]=\nu _l\nu _{l+1}q^{h_l\nu _l}\varphi _l^{}(z)(1)^{[e_l]([v_l]+[v_{l+1}])},`$ $`[\varphi _l^{}(z),f_l]_{q^{\nu _l}}=\nu _lq^{\nu _l}\varphi _{l+1}^{}(z)(1)^{[f_l]([v_l]+[v_{l+1}])},q^{h_l}\varphi _l^{}(z)q^{h_l}=q^{\nu _l}\varphi _l^{}(z),`$ $`q^{h_l}\varphi _k^{}(z)q^{h_l}=\varphi _k^{}(z)\mathrm{if}kl,l+1,q^{h_l}\varphi _{l+1}^{}(z)q^{h_l}=q^{\nu _{l+1}}\varphi _{l+1}^{}(z).`$ * For the type II vertex operators: $`[\psi _k(z),e_l]=0\mathrm{if}kl,l+1,[\psi _{l+1}(z),e_l]_{q^{\nu _{l+1}}}=0,[\psi _l(z),e_l]_{q^{\nu _l}}=\psi _{l+1}(z),`$ $`[\psi _k(z),f_l]=0\mathrm{if}kl+1,[\psi _{l+1}(z),f_l]=\nu _lq^{h_l}\psi _l(z),`$ $`q^{h_l}\psi _l(z)q^{h_l}=q^{\nu _l}\psi _l(z),q^{h_l}\psi _{l+1}(z)q^{h_l}=q^{\nu _{l+1}}\psi _{l+1}(z),`$ $`q^{h_l}\psi _k(z)q^{h_l}=\psi _k(z)\mathrm{if}kl,l+1,`$ $`[\psi _k^{}(z),e_l]=0\mathrm{if}kl,l+1,[\psi _l^{}(z),e_l]_{q^{\nu _l}}=0,`$ $`[\psi _k^{}(z),f_l]=0\mathrm{if}kl,[\psi _l^{}(z),f_l]=\nu _lq^{h_l+\nu _l}\psi _{l+1}^{}(z),`$ $`[\psi _{l+1}^{}(z),e_l]_{q^{\nu _{l+1}}}=\nu _l\nu _{l+1}q^{\nu _l}\psi _l^{}(z),q^{h_l}\psi _l^{}(z)q^{h_l}=q^{\nu _l}\psi _l^{}(z),`$ $`q^{h_l}\psi _k^{}(z)q^{h_l}=\psi _k^{}(z)\mathrm{if}kl,l+1,q^{h_l}\psi _{l+1}^{}(z)q^{h_l}=q^{\nu _{l+1}}\psi _{l+1}^{}(z).`$
warning/0001/hep-ph0001104.html
ar5iv
text
# 1 Bethe-Salpeter equation for the generalized mesons in the QCD superconductor. ## Acknowledgments We thank Youngman Kim for help with the figures. MR thanks Deog Ki Hong, Hyun Kyu Lee and Maciek Nowak for discussions. This work was supported in part by US-DOE DE-FG-88ER40388 and DE-FG02-86ER40251. ## Appendix In this appendix, we give some of the missing steps in deriving the formulae of the main text. 1. Direct calculation of the propagator (LABEL:PropPart) including mass corrections (5): According to Ref. , the general entries for $`𝐒(q)`$ read $$\begin{array}{ccccc}S_{11}(q)\hfill & =& i\psi (q)\overline{\psi }(q)\hfill & =& \left\{\left(G_0^+(q)\right)^1\gamma ^0\mathrm{\Delta }^{}(q)\gamma ^0G_0^{}(q)\mathrm{\Delta }(q)\right\}^1,\hfill \\ S_{12}(q)\hfill & =& i\psi (q)\overline{\psi }_C(q)\hfill & =& G_0^+(q)\gamma ^0\mathrm{\Delta }^{}(q)\gamma ^0S_{22}(q),\hfill \\ S_{21}(q)\hfill & =& i\psi _C(q)\overline{\psi }(q)\hfill & =& G_0^{}(q)\mathrm{\Delta }(q)S_{11}(q),\hfill \\ S_{22}(q)\hfill & =& i\psi _C(q)\overline{\psi }_C(q)\hfill & =& \left\{\left(G_0^{}(q)\right)^1\mathrm{\Delta }(q)G_0^+(q)\gamma ^0\mathrm{\Delta }^{}(q)\gamma ^0\right\}^1\hfill \end{array}$$ (A1) with $`\left(G_0^\pm (q)\right)^1`$ $`=`$ $`\gamma _\mu q^\mu \pm \mu \gamma ^0m`$ (A2) $`=`$ $`\gamma ^0\left(q_0\pm \mu 𝜶𝐪\right)m,`$ where, in general $`m=\mathrm{diag}(m_u,m_d,m_s)`$. Furthermore, $$\mathrm{\Delta }(q)=𝐌G(q)\mathrm{\Lambda }^+(𝐪)+𝐌\overline{G}(q)\mathrm{\Lambda }^{}(𝐪),$$ (A3) where $`𝐌=ϵ_f^aϵ_c^a\gamma _5=𝐌^{}`$ with $`(ϵ^a)^{bc}=ϵ^{abc}`$. Note that $$\gamma ^0\mathrm{\Delta }^{}(q)\gamma ^0=𝐌^{}G^{}(q)\mathrm{\Lambda }^{}(𝐪)𝐌^{}\overline{G}^{}(q)\mathrm{\Lambda }^+(𝐪)$$ (A4) and $$\left(𝐌^{}𝐌\right)_{ij}^{\alpha \beta }=\delta _{\alpha \beta }\delta _{ij}+\delta _{\alpha i}\delta _{\beta j},$$ (A5) where $`\alpha ,\beta =1,2,3`$ and $`i,j=1,2,3`$ are color and flavor indices, respectively. Inserting (A2) and (A3) and (A4) into $`S_{11}(q)`$, we get up to second order in $`m`$: $`S_{11}(q)`$ $``$ $`\{{\displaystyle \frac{}{}}(\gamma q+\mu \gamma ^0)m𝐌^{}{\displaystyle \frac{\gamma q\mu \gamma ^0}{(q_0\mu )^2|𝐪|^2m^2}}𝐌(|G(q)|^2\mathrm{\Lambda }^+(𝐪)+|\overline{G}(q)|^2\mathrm{\Lambda }^{}(𝐪))`$ (A6) $`+𝐌^{}m{\displaystyle \frac{G^{}(q)\overline{G}(q)\mathrm{\Lambda }^{}(𝐪)+\overline{G}^{}(q)G(q)\mathrm{\Lambda }^+(𝐪)}{(q_0\mu )^2|𝐪|^2m^2}}𝐌\}^1`$ $`=`$ $`\{(\gamma q\mu \gamma ^0)(\gamma q+\mu \gamma ^0)`$ $`\left(𝐌^{}𝐌+{\displaystyle \frac{𝐌^{}m^2𝐌}{(q_0\mu )^2|𝐪|^2}}\right)\left(|G(q)|^2\mathrm{\Lambda }^+(𝐪)+|\overline{G}(q)|^2\mathrm{\Lambda }^{}(𝐪)\right)`$ $`(\gamma q\mu \gamma ^0)(m𝐌^{}m𝐌{\displaystyle \frac{G^{}(q)\overline{G}(q)\mathrm{\Lambda }^{}(𝐪)+\overline{G}^{}(q)G(q)\mathrm{\Lambda }^+(𝐪)}{(q_0\mu )^2|𝐪|^2}})\}^1`$ $`\times (\gamma q\mu \gamma ^0).`$ Using $$(\gamma q\mu \gamma ^0)(\gamma q+\mu \gamma ^0)=\{q_0^2(\mu |𝐪|)^2\}\mathrm{\Lambda }^+(𝐪)+\{q_0^2(\mu +|𝐪|)^2\}\mathrm{\Lambda }^{}(𝐪),$$ we can transform (A6) into $`S_{11}(q)`$ $`=`$ $`\{\mathrm{\Lambda }^+(𝐪)[q_0^2(\mu |𝐪|)^2(𝐌^{}𝐌+{\displaystyle \frac{𝐌^{}m^2𝐌}{(q_0\mu )^2|𝐪|^2}})|G(q)|^2]\mathrm{\Lambda }^+(𝐪)`$ (A7) $`+\mathrm{\Lambda }^{}(𝐪)\left[q_0^2(\mu +|𝐪|)^2\left(𝐌^{}𝐌+{\displaystyle \frac{𝐌^{}m^2𝐌}{(q_0\mu )^2|𝐪|^2}}\right)|\overline{G}(q)|^2\right]\mathrm{\Lambda }^{}(𝐪)`$ $`\mathrm{\Lambda }^+(𝐪)\gamma ^0\left[q^0\mu +|𝐪|\right]\left(m𝐌^{}m𝐌{\displaystyle \frac{G^{}(q)\overline{G}(q)}{(q_0\mu )^2|𝐪|^2}}\right)\mathrm{\Lambda }^{}(𝐪)`$ $`\mathrm{\Lambda }^{}(𝐪)\gamma ^0[q^0\mu |𝐪|](m𝐌^{}m𝐌{\displaystyle \frac{\overline{G}^{}(q)G(q)}{(q_0\mu )^2|𝐪|^2}})\mathrm{\Lambda }^+(𝐪)\}^1`$ $`\times \gamma ^0(q_0\mu 𝜶𝐪).`$ Note that (A7) has the structure $`S_{11}(q)`$ $`=`$ $`\left\{\mathrm{\Lambda }^+A\mathrm{\Lambda }^++\mathrm{\Lambda }^{}D\mathrm{\Lambda }^{}+\mathrm{\Lambda }^+B\mathrm{\Lambda }^{}+\mathrm{\Lambda }^{}C\mathrm{\Lambda }^+\right\}^1\gamma ^0(q_0\mu 𝜶𝐪)`$ (A8) $``$ $`\left\{F\right\}^1\gamma ^0(q_0\mu 𝜶𝐪)`$ with $`A`$ $`=`$ $`q_0^2\mathit{ϵ}_q^2𝐌^{}m^2𝐌{\displaystyle \frac{|G(q)|^2}{(q_0\mu )^2|𝐪|^2}},`$ $`D`$ $`=`$ $`q_0^2\overline{\mathit{ϵ}}_q^2𝐌^{}m^2𝐌{\displaystyle \frac{|\overline{G}(q)|^2}{(q_0\mu )^2|𝐪|^2}},`$ $`B`$ $`=`$ $`\gamma ^0\left[q_0\mu +|𝐪|\right]\left(m𝐌^{}m𝐌{\displaystyle \frac{G^{}(q)\overline{G}(q)}{(q_0\mu )^2|𝐪|^2}}\right),`$ $`C`$ $`=`$ $`\gamma ^0\left[q_0\mu |𝐪|\right]\left(m𝐌^{}m𝐌{\displaystyle \frac{\overline{G}^{}(q)G(q)}{(q_0\mu )^2|𝐪|^2}}\right),`$ where $`\mathit{ϵ}_q`$ $``$ $`(\mu |𝐪|)^2+𝐌^{}𝐌|G(q)|^2,`$ (A10) $`\overline{\mathit{ϵ}}_q`$ $``$ $`(\mu +|𝐪|)^2+𝐌^{}𝐌|\overline{G}(q)|^2.`$ (A11) Here the projectors $`\mathrm{\Lambda }^\pm `$ are short for $`\mathrm{\Lambda }^\pm (𝐪)`$, and satisfy $`\mathrm{\Lambda }^\pm \mathrm{\Lambda }^\pm =\mathrm{\Lambda }^\pm `$ and $`\mathrm{\Lambda }^++\mathrm{\Lambda }^{}=\mathrm{𝟏}`$, where the unit matrix refers to the Dirac space. We will use now that the inverse of $$F\mathrm{\Lambda }^+A\mathrm{\Lambda }^++\mathrm{\Lambda }^+B\mathrm{\Lambda }^{}+\mathrm{\Lambda }^{}C\mathrm{\Lambda }^++\mathrm{\Lambda }^{}D\mathrm{\Lambda }^{}$$ reads $`F^1`$ $`=`$ $`\mathrm{\Lambda }^+A^1\mathrm{\Lambda }^++\mathrm{\Lambda }^{}D^1\mathrm{\Lambda }^{}`$ $`\mathrm{\Lambda }^+A^1BD^1\mathrm{\Lambda }^{}\mathrm{\Lambda }^{}D^1CA^1\mathrm{\Lambda }^+`$ $`+\mathrm{\Lambda }^+A^1BD^1CA^1\mathrm{\Lambda }^++\mathrm{\Lambda }^{}D^1CA^1BD^1\mathrm{\Lambda }^{}`$ $`+𝒪(m^3)`$ and furthermore that $`\gamma ^0\mathrm{\Lambda }^\pm =\mathrm{\Lambda }^{}\gamma ^0`$ and $`\mathrm{\Lambda }^\pm (𝐪)𝜶𝐪=\pm |𝐪|\mathrm{\Lambda }^\pm (𝐪)`$. Then, we have $`S_{11}(q)`$ $``$ $`{\displaystyle \frac{\gamma ^0\mathrm{\Lambda }^{}(𝐪)(q_0\mu +|𝐪|)}{q_0^2\mathit{ϵ}_q^2}}+{\displaystyle \frac{\gamma ^0\mathrm{\Lambda }^+(𝐪)(q_0\mu |𝐪|)}{q_0^2\overline{\mathit{ϵ}}_q^2}}`$ (A12) $`+\left((q_0\mu )^2|𝐪|^2\right)\left\{{\displaystyle \frac{1}{q_0^2\mathit{ϵ}_q^2}}m{\displaystyle \frac{1}{q_0^2\overline{\mathit{ϵ}}_q^2}}\mathrm{\Lambda }^+(𝐪)+{\displaystyle \frac{1}{q_0^2\overline{\mathit{ϵ}}_q^2}}m{\displaystyle \frac{1}{q_0^2\mathit{ϵ}_q^2}}\mathrm{\Lambda }^{}(𝐪)\right\}`$ $`G^{}(q)\overline{G}(q){\displaystyle \frac{1}{q_0^2\mathit{ϵ}_q^2}}𝐌^{}m𝐌{\displaystyle \frac{1}{q_0^2\overline{\mathit{ϵ}}_q^2}}\mathrm{\Lambda }^+(𝐪)`$ $`\overline{G}^{}(q)G(q){\displaystyle \frac{1}{q_0^2\overline{\mathit{ϵ}}_q^2}}𝐌^{}m𝐌{\displaystyle \frac{1}{q_0^2\mathit{ϵ}_q^2}}\mathrm{\Lambda }^{}(𝐪)`$ $`+{\displaystyle \frac{1}{q_0^2\mathit{ϵ}_q^2}}𝐌^{}m^2𝐌{\displaystyle \frac{|G(q)|^2}{(q_0\mu )^2|𝐪|^2}}{\displaystyle \frac{\gamma ^0\mathrm{\Lambda }^{}(𝐪)(q_0\mu +|𝐪|)}{q_0^2\mathit{ϵ}_q^2}}`$ $`+{\displaystyle \frac{1}{q_0^2\overline{\mathit{ϵ}}_q^2}}𝐌^{}m^2𝐌{\displaystyle \frac{|\overline{G}(q)|^2}{(q_0\mu )^2|𝐪|^2}}{\displaystyle \frac{\gamma ^0\mathrm{\Lambda }^+(𝐪)(q_0\mu |𝐪|)}{q_0^2\overline{\mathit{ϵ}}_q^2}}`$ $`+{\displaystyle \frac{(q_0\mu )^2|𝐪|^2}{q_0^2\mathit{ϵ}_q^2}}\left(m𝐌^{}m𝐌{\displaystyle \frac{G^{}(q)\overline{G}(q)}{(q_0\mu )^2|𝐪|^2}}\right){\displaystyle \frac{1}{q_0^2\overline{\mathit{ϵ}}_q^2}}`$ $`\times \left(m𝐌^{}m𝐌{\displaystyle \frac{\overline{G}^{}(q)G(q)}{(q_0\mu )^2|𝐪|^2}}\right){\displaystyle \frac{\gamma ^0\mathrm{\Lambda }^{}(𝐪)(q_0\mu +|𝐪|)}{q_0^2\mathit{ϵ}_q^2}}`$ $`+{\displaystyle \frac{(q_0\mu )^2|𝐪|^2}{q_0^2\overline{\mathit{ϵ}}_q^2}}\left(m𝐌^{}m𝐌{\displaystyle \frac{\overline{G}^{}(q)G(q)}{(q_0\mu )^2|𝐪|^2}}\right){\displaystyle \frac{1}{q_0^2\mathit{ϵ}_q^2}}`$ $`\times \left(m𝐌^{}m𝐌{\displaystyle \frac{G^{}(q)\overline{G}(q)}{(q_0\mu )^2|𝐪|^2}}\right){\displaystyle \frac{\gamma ^0\mathrm{\Lambda }^+(𝐪)(q_0\mu |𝐪|)}{q_0^2\overline{\mathit{ϵ}}_{q}^{}{}_{}{}^{2}}}`$ $`+𝒪(m^3).`$ We apply now the following approximations $`\mathit{ϵ}_q^2`$ $`=`$ $`(\mu |𝐪|)^2+𝐌^{}𝐌|G(q)|^2(\mu |𝐪|)^2+|G(q)|^2ϵ_q^2,`$ $`\overline{\mathit{ϵ}}_q^2`$ $`=`$ $`(\mu +|𝐪|)^2+𝐌^{}𝐌|\overline{G}(q)|^2(\mu +|𝐪|)^2+|\overline{G}(q)|^2\overline{ϵ}_q^2,`$ (A13) and, since $`|𝐪|\mu +q_{||}`$, $`ϵ_q^2q_{||}^2+|G(q)|^2,`$ (A14) $`\overline{ϵ}_q^24\mu ^2,`$ (A15) $`(q_0\mu )^2|𝐪|^22\mu (q_0+q_{||}).`$ (A16) Then equation (A12) simplifies to $`S_{11}(q)`$ $`=`$ $`\gamma ^0{\displaystyle \frac{q_0+q_{||}}{q_0^2ϵ_q^2}}\mathrm{\Lambda }^{}(𝐪)+{\displaystyle \frac{\gamma ^0\mathrm{\Lambda }^+(𝐪)}{2\mu }}+{\displaystyle \frac{m}{2\mu }}{\displaystyle \frac{q_0+q_{||}}{q_0^2ϵ_q^2}}`$ (A17) $`+\gamma ^0{\displaystyle \frac{m^2}{2\mu }}\left({\displaystyle \frac{q_0+q_{||}}{q_0^2ϵ_q^2}}\right)^2\mathrm{\Lambda }^{}(𝐪)\gamma ^0{\displaystyle \frac{𝐌^{}m^2𝐌}{2\mu }}\left({\displaystyle \frac{|G(q)|}{q_0^2ϵ_q^2}}\right)^2\mathrm{\Lambda }^{}(𝐪)`$ $`+𝒪(\mu ^2,m^3,|G(q)|^2\frac{𝐌^{}𝐌\mathrm{𝟏}}{\left(q_0^2ϵ_q^2\right)^2}).`$ Note that $`S_{22}(q)`$ follows from (A12) under the substitutions $$\mathrm{\Lambda }^\pm (𝐪)\mathrm{\Lambda }^{}(𝐪),\pm \mu \mu ,\pm |𝐪||𝐪|,\pm G^{}G,\pm \overline{G}^{}\overline{G},\text{and}𝐌^{}𝐌,$$ (A18) which also imply $`\pm q_{||}q_{||}`$. In fact, these rules can be traced back to the replacements $`G_0^+(q)G_0^{}(q)`$ and $`\mathrm{\Delta }(q)\gamma ^0\mathrm{\Delta }^{}(q)\gamma ^0`$ which link the various Nambu-Gorkov components in (A1) to each other. Using (A15) and $$(q_0+\mu )^2|𝐪|^22\mu (q_0q_{||}),$$ (A19) we get $`S_{22}(q)`$ $`=`$ $`{\displaystyle \frac{\gamma ^0\mathrm{\Lambda }^+(𝐪)(q_0q_{||})}{q_0^2ϵ_q^2}}{\displaystyle \frac{\gamma ^0\mathrm{\Lambda }^{}(𝐪)}{2\mu }}{\displaystyle \frac{m}{2\mu }}{\displaystyle \frac{q_0q_{||}}{q_0^2ϵ_q^2}}`$ (A20) $`\gamma ^0{\displaystyle \frac{m^2}{2\mu }}\left({\displaystyle \frac{q_0q_{||}}{q_0^2ϵ_q^2}}\right)^2\mathrm{\Lambda }^+(𝐪)+\gamma ^0{\displaystyle \frac{𝐌m^2𝐌^{}}{2\mu }}\left({\displaystyle \frac{|G(q)|}{q_0^2ϵ_q^2}}\right)^2\mathrm{\Lambda }^+(𝐪)`$ $`+𝒪(\mu ^2,m^3,|G(q)|^2\frac{𝐌^{}𝐌\mathrm{𝟏}}{\left(q_0^2ϵ_q^2\right)^2}).`$ Finally, $`S_{21}(q)`$ $`=`$ $`G_0^{}(q)\mathrm{\Delta }(q)S_{11}(q)`$ $`=`$ $`{\displaystyle \frac{\gamma ^0(q_0\mu 𝜶𝐪)+m}{(q_0\mu )^2|𝐪|^2m^2}}\times 𝐌\left[G(q)\mathrm{\Lambda }^+(𝐪)+\overline{G}(q)\mathrm{\Lambda }^{}(𝐪)\right]\times S_{11}(q)`$ $`=`$ $`+\left[{\displaystyle \frac{q_0\mu |𝐪|}{(q_0\mu )^2|𝐪|^2}}𝐌G(q)\mathrm{\Lambda }^{}(𝐪)+{\displaystyle \frac{q_0\mu +|𝐪|}{(q_0\mu )^2|𝐪|^2}}𝐌\overline{G}(q)\mathrm{\Lambda }^+(𝐪)\right]\gamma ^0S_{11}(q)`$ $`{\displaystyle \frac{m}{(q_0\mu )^2|𝐪|^2}}𝐌\left[G(q)\mathrm{\Lambda }^+(𝐪)+\overline{G}(q)\mathrm{\Lambda }^{}(𝐪)\right]S_{11}(q)`$ $`+m^2\left[{\displaystyle \frac{q_0\mu |𝐪|}{\left((q_0\mu )^2|𝐪|^2\right)^2}}𝐌G(q)\mathrm{\Lambda }^{}(𝐪)+{\displaystyle \frac{q_0\mu +|𝐪|}{(q_0\mu )^2|𝐪|^2}}𝐌\overline{G}(q)\mathrm{\Lambda }^+(𝐪)\right]`$ $`\times \gamma ^0S_{11}(q)+𝒪\left(m^3\right)`$ $`=`$ $`{\displaystyle \frac{𝐌G(q)\mathrm{\Lambda }^{}(𝐪)}{q_0^2ϵ_q^2}}\gamma ^0{\displaystyle \frac{𝐌m}{2\mu }}{\displaystyle \frac{G(q)\mathrm{\Lambda }^+(𝐪)}{q_0^2ϵ_q^2}}\gamma ^0{\displaystyle \frac{m𝐌}{2\mu }}{\displaystyle \frac{G(q)\mathrm{\Lambda }^{}(𝐪)}{q_0^2ϵ_q^2}}`$ $`+{\displaystyle \frac{𝐌m^2}{2\mu }}{\displaystyle \frac{G(q)\left(q_0+q_{||}\right)\mathrm{\Lambda }^{}(𝐪)}{\left(q_0^2ϵ_q^2\right)^2}}{\displaystyle \frac{\mathrm{𝐌𝐌}^{}m^2𝐌}{2\mu }}{\displaystyle \frac{G(q)|G(q)|^2\mathrm{\Lambda }^{}(𝐪)}{\left(q_0+q_{||}\right)\left(q_0^2ϵ_q^2\right)^2}}`$ $`{\displaystyle \frac{m^2}{2\mu }}{\displaystyle \frac{𝐌G(q)\mathrm{\Lambda }^{}(𝐪)}{\left(q_0+q_{||}\right)\left(q_0^2ϵ_q^2\right)}}+𝒪(\mu ^2,m^3,|G(q)|^2\frac{𝐌^{}𝐌\mathrm{𝟏}}{\left(q_0^2ϵ_q^2\right)^2}).`$ The latter equation can be simplified to $`S_{21}(q)`$ $`=`$ $`{\displaystyle \frac{𝐌G(q)\mathrm{\Lambda }^{}(𝐪)}{q_0^2ϵ_q^2}}\gamma ^0{\displaystyle \frac{𝐌m}{2\mu }}{\displaystyle \frac{G(q)\mathrm{\Lambda }^+(𝐪)}{q_0^2ϵ_q^2}}\gamma ^0{\displaystyle \frac{m𝐌}{2\mu }}{\displaystyle \frac{G(q)\mathrm{\Lambda }^{}(𝐪)}{q_0^2ϵ_q^2}}`$ (A22) $`+{\displaystyle \frac{1}{2\mu }}\left\{𝐌m^2\left(q_0+q_{||}\right)m^2𝐌\left(q_0q_{||}\right)\right\}{\displaystyle \frac{G(q)\mathrm{\Lambda }^{}(𝐪)}{\left(q_0^2ϵ_q^2\right)^2}}`$ $`+𝒪(\mu ^2,m^3,|G(q)|^2\frac{𝐌^{}𝐌\mathrm{𝟏}}{\left(q_0^2ϵ_q^2\right)^2}).`$ Again, $`S_{12}`$ follows under the substitutions (A18) from $`S_{21}`$. Using (A15) and(A19), we get $`S_{12}(q)`$ $`=`$ $`{\displaystyle \frac{𝐌^{}G^{}(q)\mathrm{\Lambda }^+(𝐪)}{q_0^2ϵ_q^2}}\gamma ^0{\displaystyle \frac{𝐌^{}m}{2\mu }}{\displaystyle \frac{G^{}(q)\mathrm{\Lambda }^{}(𝐪)}{q_0^2ϵ_q^2}}\gamma ^0{\displaystyle \frac{m𝐌^{}}{2\mu }}{\displaystyle \frac{G^{}(q)\mathrm{\Lambda }^+(𝐪)}{q_0^2ϵ_q^2}}`$ (A23) $`+{\displaystyle \frac{1}{2\mu }}\left\{𝐌^{}m^2\left(q_0q_{||}\right)m^2𝐌^{}\left(q_0+q_{||}\right)\right\}{\displaystyle \frac{G^{}(q)\mathrm{\Lambda }^+(𝐪)}{\left(q_0^2ϵ_q^2\right)^2}}`$ $`+𝒪(\mu ^2,m^3,|G(q)|^2\frac{𝐌^{}𝐌\mathrm{𝟏}}{\left(q_0^2ϵ_q^2\right)^2}).`$ Note that the above determined quark propagators (A17), (A20), (A22) and (A23) show neither a $`\overline{G}(q)`$ nor a $`\overline{G}^{}(q)`$ dependence. In fact, such terms first arise at order $`𝒪(\mu ^2)`$. 2. Perturbative calculation of the mass corrections (5) to the propagator (LABEL:PropPart): As a check on the precedent analysis, we now carry a mass perturbation analysis of the quark propagator in the CFL phase. We recall that the massless propagator reads <sup>#12</sup><sup>#12</sup>#12These expressions can easily be derived with the methods of the former section. Especially, they are consistent with (A17), (A20), (A22) and (A23) to order $`𝒪(m^0,\mu ^1)`$ and under the approximations (A14), (A15), (A16) and (A19). $$\begin{array}{ccccc}S_{11}(q)\hfill & =& \left[\mathrm{\Lambda }^+(𝐪)\gamma ^0\frac{q_0\mu +|𝐪|}{q_0^2\mathit{ϵ}_q^2}\mathrm{\Lambda }^{}(𝐪)+\mathrm{\Lambda }^{}(𝐪)\gamma ^0\frac{q_0\mu |𝐪|}{q_0^2\overline{\mathit{ϵ}}_q^2}\mathrm{\Lambda }^+(𝐪)\right],\hfill & & \\ S_{12}(q)\hfill & =& \left[\mathrm{\Lambda }^+(𝐪)𝐌^{}\frac{G^{}(q)}{q_0^2\mathit{ϵ}_q^2}\mathrm{\Lambda }^+(𝐪)+\mathrm{\Lambda }^{}(𝐪)𝐌^{}\frac{\overline{G}^{}(q)}{q_0^2\overline{\mathit{ϵ}}_q^2}\mathrm{\Lambda }^{}(𝐪)\right],\hfill & & \\ S_{21}(q)\hfill & =& \left[\mathrm{\Lambda }^{}(𝐪)𝐌\frac{G(q)}{q_0^2\mathit{ϵ}_q^2}\mathrm{\Lambda }^{}(𝐪)+\mathrm{\Lambda }^+(𝐪)𝐌\frac{\overline{G}(q)}{q_0^2\overline{\mathit{ϵ}}_q^2}\mathrm{\Lambda }^+(𝐪)\right],\hfill & & \\ S_{22}(q)\hfill & =& \left[\mathrm{\Lambda }^{}(𝐪)\gamma ^0\frac{q_0+\mu |𝐪|}{q_0^2\mathit{ϵ}_q^2}\mathrm{\Lambda }^+(𝐪)+\mathrm{\Lambda }^+(𝐪)\gamma ^0\frac{q_0+\mu +|𝐪|}{q_0^2\overline{\mathit{ϵ}}_q^2}\mathrm{\Lambda }^{}(𝐪)\right]\hfill & & \end{array}$$ (A24) with $`\mathit{ϵ}_q^2=(\mu |𝐪|)^2+𝐌^{}𝐌|G(q)|^2`$ and $`\overline{\mathit{ϵ}}_q^2=(\mu +|𝐪|)^2+𝐌^{}𝐌|\overline{G}(q)|^2`$. Using (A24), we can calculate the mass corrections in perturbation theory, i.e. $`\mathrm{\Delta }_m𝐒(q)`$ $`=`$ $`𝐒(q)\left(\begin{array}{cc}m& 0\\ 0& m\end{array}\right)𝐒(q),`$ (A27) $`\mathrm{\Delta }_{m^2}𝐒(q)`$ $`=`$ $`𝐒(q)\left(\begin{array}{cc}m& 0\\ 0& m\end{array}\right)𝐒(q)\left(\begin{array}{cc}m& 0\\ 0& m\end{array}\right)𝐒(q),`$ (A32) etc., where, in general, $`m=\mathrm{diag}(m_u,m_d,m_s)`$. The remaining task is just to insert the terms (A24) into (A27) and (A32). One can easily check that the $`m`$ and $`m^2`$ terms of (A17), (A20), (A22) and (A23) are recovered in this way. Hence, the direct and perturbative arguments give the same result to the order quoted. In retrospect, this is not surprising. Both approaches use the Dyson expansion of the propagator. In the perturbative argument, the full massive propagator is expanded in terms of the full massless propagator. In the direct approach, the same expansion is performed on the level of the free propagators. The gap functions $`𝐌G(q)`$ and $`𝐌\overline{G}(q)`$ are completely passive with respect to these expansions. Therefore the results are the same. The neglect of the color-flavor non-diagonal terms in (A5) through the approximation (A13) which has also been used in to simplify the denominators, can be justified as follows. The eigenvalues of $`𝐌`$ (=$`𝐌^{}`$) read $$\mathrm{eig}\left(𝐌\right)=+2,+1,+1,+1,1,1,1,1,1.$$ Thus the eigenvalues of $`𝐌^{}𝐌=𝐌^2`$ are <sup>#13</sup><sup>#13</sup>#13For the general case $`N_c=N_f`$, the “+2” and “4” have to be replaced by $`N_c1`$ and $`(N_c1)^2`$, respectively. Furthermore, there are $`\frac{1}{2}N_c(N_c1)`$ eigenvalues $`+1`$ and $`\frac{1}{2}N_c(N_c+1)1`$ eigenvalues $`1`$. $$\mathrm{eig}\left(𝐌^{}𝐌\right)=4,1,1,1,1,1,1,1,1.$$ Note that eight eigenvalues are equal to unit and only the ninth deviates from this value . This is related to an explicit U(1) degree of freedom in the U(3) color-flavor phase, whereas the agreement of the other eight eigenvalues corresponds to the SU(3) sector in the color-flavor phase. Throughout, we have specialized to the SU(3) phase as indicated in the introduction, leaving the issue of the additional U(1) in the presence of the triangle anomaly for a future discussion. 3. Derivation of the mass formula of the generalized Goldstone meson: Following Ref. , we consider the chiral Ward identity implied by the underlying flavor symmetry in the CFL phase. Indeed, when chiral symmetry is softly broken by massive quarks $`m=\mathrm{diag}(m_u,m_d,m_s)`$, then the pions are expected to be massive. Hence $$0d^4x_x^\mu \mathrm{BCS}\left|T^{}𝐀_\mu ^\alpha (x)𝝅_B^\beta (0)\right|\mathrm{BCS},$$ (A33) where the axial-vector current $`𝐀_\mu ^a`$ is given in (A82) and the pion field $`𝝅_B(x)`$ in the CFL phase is defined as (see Appendix-7) $`𝝅_B^\beta (x)=\left(\begin{array}{cc}0& \overline{\psi }\gamma ^0\left(𝐌i\tau ^\beta \gamma _5\right)^{}\gamma ^0\psi _C(x)\\ \overline{\psi }_C𝐌i\tau ^\beta \gamma _5\psi (x)& 0\end{array}\right),`$ (A36) which is consistent with (19). The flavor axial-vector current in the CFL phase (see Appendix-6) obeys the local divergence equation $`𝐀^\alpha (x)=\left(\begin{array}{cc}\overline{\psi }i[m,\frac{1}{2}\tau ^\alpha ]_+\gamma _5\psi (x)& 0\\ 0& \overline{\psi }_Ci[m,\frac{1}{2}\tau _{}^{\alpha }{}_{}{}^{}]_+\gamma _5\psi _C(x)\end{array}\right).`$ (A39) For massless quarks, the hermitean axial-isovector charge $`𝐐_5^\alpha 𝐐_5^\alpha (x^0)={\displaystyle d^3x\left(\begin{array}{cc}\overline{\psi }\frac{1}{2}\tau ^\alpha \gamma ^0\gamma _5\psi (x)& 0\\ 0& \overline{\psi }_C\frac{1}{2}\tau _{}^{\alpha }{}_{}{}^{}\gamma ^0\gamma _5\psi _C(x)\end{array}\right)}`$ (A42) is conserved and generates axial-vector rotations, e.g. $$[𝐐_5^\alpha ,𝚿(x)]=\gamma _5\frac{1}{2}𝐓^\alpha 𝚿(x).$$ (A43) In terms of (A39-A43), the identity (A36) yields the axial Ward-identity $$d^4x\mathrm{BCS}\left|T^{}[m,\frac{1}{2}𝝅^\alpha (x)]_+𝝅_B^\beta (0)\right|\mathrm{BCS}=\mathrm{BCS}\left|𝚺_B^{\alpha \beta }(0)\right|\mathrm{BCS},$$ (A44) where the diquark field $`𝚺_B^{\alpha \beta }(x)`$ is defined as $`𝚺^{\alpha \beta }=\overline{𝚿}(x)[\frac{1}{2}𝐓^\alpha ,\left(\begin{array}{cc}0& i\tau ^\beta 𝐌^{}\\ i𝐌\tau ^\beta & 0\end{array}\right)]_+𝚿(x)`$ (A47) and $`𝝅(x)`$ is the diagonal pion field $`𝝅^\alpha (x)=\left(\begin{array}{cc}\overline{\psi }i\tau ^\alpha \gamma _5\psi (x)& 0\\ 0& \overline{\psi }_Ci\tau _{}^{\alpha }{}_{}{}^{}\gamma _5\psi _C(x)\end{array}\right).`$ (A50) The nonconfining character of the weak coupling description allows for the occurrence of the gapped $`qq`$ and/or $`\overline{q}q`$ exchange. Hence, $`\mathrm{BCS}\left|𝚺_B^{\alpha \beta }(0)\right|\mathrm{BCS}{\displaystyle \frac{d^4q}{(2\pi )^4}\mathrm{Tr}\left[i\gamma _5\frac{1}{2}[m,𝐓^\alpha ]_+i𝐒(q)𝚷_B^\beta i𝐒(q)\right]}`$ $`+\left\{{\displaystyle \frac{d^4q}{(2\pi )^4}\mathrm{Tr}\left[i\gamma _5\frac{1}{2}[m,𝐓^\alpha ]_+i𝐒(q)i𝚪_\pi ^\xi i𝐒(q)\right]}\right\}\left({\displaystyle \frac{i}{M^2}}\right)^{\xi \xi ^{}}\left\{{\displaystyle \frac{d^4q}{(2\pi )^4}\mathrm{Tr}\left[i𝚪_\pi ^\xi ^{}i𝐒(q)𝚷_B^\beta i𝐒(q)\right]}\right\}`$ (A51) with $`𝚷_B^\beta \left(\begin{array}{cc}0& \gamma ^0\left(i𝐌\tau ^\beta \gamma _5\right)^{}\gamma ^0\\ i𝐌\tau ^\beta \gamma _5& 0\end{array}\right).`$ (A54) In the chiral limit $`m_i0`$, $`i\{u,d,s\}`$, the first term in (A51) drops out and the identity is fulfilled if $`1/M^2`$ is sufficiently singular in $`m_i`$ to match the numerator. The traces can be evaluated in weak coupling. The result is <sup>#14</sup><sup>#14</sup>#14The use of $`F_T`$ instead of $`F_S`$ in the pion vertex follows from the fact that the intermediate BCS pion is generated by a chiral rotation of the BCS ground state. A similar interpretation in matter is made in . $`{\displaystyle \frac{d^4q}{(2\pi )^4}\mathrm{Tr}\left[i\gamma _5\frac{1}{2}[m,𝐓^\alpha ]_+i𝐒(q)i𝚪_\pi ^\xi i𝐒(q)\right]}=𝒪(m^2),`$ (A55) $`{\displaystyle \frac{d^4q}{(2\pi )^4}\mathrm{Tr}\left[i𝚪_\pi ^\xi ^{}i𝐒(q)𝚷_B^\beta i𝐒(q)\right]}=\delta ^{\xi ^{}\beta }{\displaystyle \frac{16i}{F_T}}{\displaystyle \frac{d^4q}{(2\pi )^4}\frac{G(q)}{q_0^2ϵ_q^2}},`$ (A56) which shows that $`M^2=𝒪(m^2)`$. To determine the coefficient, we need to expand the vertices and the propagators in (A51) to leading order in $`m`$. The $`𝒪(m)`$ corrections to both $`G(p)`$ and $`\mathrm{\Gamma }(p)`$ do not contribute. They trace to zero because of a poor spin structure. Therefore, only the $`𝒪(m)`$ correction to the propagator (LABEL:PropPart) is needed, i.e. (5). Inserting (LABEL:PropPart) together with the mass correction (5) into (A55) yields $`{\displaystyle \frac{d^4q}{(2\pi )^4}\mathrm{Tr}\left[i\gamma _5\frac{1}{2}[m,𝐓^\alpha ]_+i𝐒(q)i𝚪_\pi ^\xi i𝐒(q)\right]}`$ $`={\displaystyle \frac{\mu G_0}{8\pi ^2F_T}}\mathrm{Tr}_{cf}\left([m^2,\tau ^\alpha ]\left(𝐌^{}𝐌^\beta 𝐌_{}^{\beta }{}_{}{}^{}𝐌\right)+[m^2,\tau _{}^{\alpha }{}_{}{}^{}]\left(\mathrm{𝐌𝐌}_{}^{\beta }{}_{}{}^{}𝐌^\beta 𝐌^{}\right)\right).`$ (A57) Here, the resulting integral simplifies under a contour integration in the following way: $`i{\displaystyle \frac{d^4q}{(2\pi )^4}\frac{q_0\pm q_{||}}{\left(q_0^2ϵ_q^2\right)^2}\frac{|G(q_{||})|^2}{2\mu F_T}}`$ $`=`$ $`i2\pi {\displaystyle _0^{2\mu }}{\displaystyle \frac{dq_{}q_{}}{(2\pi )^2}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{dq_{||}}{2\pi }}{\displaystyle \frac{|G(q_{||})|^2}{2\mu F_T}}{\displaystyle \frac{idq_4}{2\pi }\frac{iq_4\pm q_{||}}{\left(q_4^2+ϵ_q^2\right)^2}}`$ $`={\displaystyle \frac{\mu ^2}{\pi ^2}}{\displaystyle _0^{\mathrm{}}}𝑑q_{||}{\displaystyle \frac{|G(q_{||})|^2}{2\mu F_T}}{\displaystyle \frac{q_{||}}{4ϵ_q^3}}`$ $``$ $`{\displaystyle \frac{\mu }{8\pi ^2F_T}}{\displaystyle _{G_0}^\mathrm{\Lambda }_{}}𝑑q_{||}{\displaystyle \frac{|G(q_{||})|^2}{q_{||}^2}}`$ $``$ $`{\displaystyle \frac{\mu }{8\pi ^2F_T}}{\displaystyle _0^{x_0}}𝑑x{\displaystyle \frac{e^x}{\mathrm{\Lambda }_{}}}G_0^2\mathrm{sin}^2\left({\displaystyle \frac{\pi x}{2x_0}}\right)`$ $``$ $`{\displaystyle \frac{\mu }{8\pi ^2F_T}}{\displaystyle \frac{G_0^2}{\mathrm{\Lambda }_{}}}e^{x_0}={\displaystyle \frac{\mu }{8\pi ^2F_T}}G_0,`$ where the gap solution (9) and the logarithmic scales $`x=\mathrm{ln}(\mathrm{\Lambda }_{}/q_{||})`$ and $`x_0=\mathrm{ln}(\mathrm{\Lambda }_{}/G_0)`$ were inserted in the second to last line. Furthermore, $`[m,[m,\tau ^\alpha ]_+]=[m^2,\tau ^\alpha ]`$ was used. Inserting (A56) and (A57) in (A51) and noting that $$\mathrm{\Sigma }_B^{\alpha \beta }\mathrm{Tr}\left([\frac{1}{2}𝐓^\alpha ,\left(\begin{array}{cc}0& i\tau ^\beta 𝐌^{}\\ i𝐌\tau ^\beta & 0\end{array}\right)]_+i𝐒\right)=8\delta ^{\alpha \beta }\frac{d^4q}{(2\pi )^4}\frac{G(q)}{q_0^2ϵ_q^2},$$ (A58) we obtain for the mass of the Goldstone modes $$\left(M^2\right)^{\alpha \beta }\frac{\mu G_0}{4\pi ^2F_T^2}\mathrm{Tr}_{cf}\left([m^2,\tau ^\alpha ]\left(𝐌^{}𝐌^\beta 𝐌_{}^{\beta }{}_{}{}^{}𝐌\right)+[m^2,\tau _{}^{\alpha }{}_{}{}^{}]\left(\mathrm{𝐌𝐌}_{}^{\beta }{}_{}{}^{}𝐌^\beta 𝐌^{}\right)\right),$$ (A59) which is (28), see also Ref. . Note that the color-flavor traces, which first appeared in the transition from (A55) to (A57), yield zero. In order to get a non-zero result for the mass matrix of the generalized pion, we have to insert in (A55) the next-to-leading order, $`𝒪(1/\mu )`$, even for the massless terms of the propagator, i.e. the second terms on the right hand sides of (A17) and (A20) which can be traced back to the leading terms of the antiparticle propagator, see (A24<sup>#15</sup><sup>#15</sup>#15The $`1/\mu `$ expansion of the numerators and denominators of the particle propagators only modifies the leading result (A57) by overall factors that do not prevent the vanishing of the color-flavor traces.. These terms are in fact antiparticle-gap independent. The first antiparticle-gap dependent piece appears at order $`𝒪(1/\mu ^2)`$ and is therefore subleading. This is fortunate, since according to the antiparticle-gap is gauge-fixing-term dependent. Inserting the above mentioned terms in (A55), we get $`(A37)`$ $`=`$ $`(A39)+{\displaystyle \frac{i}{4\mu ^2F_T}}{\displaystyle \frac{d^4q}{(2\pi )^4}\frac{|G(q)|^2}{q_0^2ϵ_q^2}}`$ (A60) $`\times \{\mathrm{Tr}_{cf}\left([m,\tau ^\alpha ]_+(𝐌_{}^{\xi }{}_{}{}^{}m𝐌+𝐌^{}m𝐌^\xi )\right)`$ $`+\mathrm{Tr}_{cf}\left([m,\tau _{}^{\alpha }{}_{}{}^{}]_+(𝐌^\xi m𝐌^{}+𝐌m𝐌_{}^{\xi }{}_{}{}^{})\right)\}.`$ Note that $`i{\displaystyle \frac{d^4q}{(2\pi )^4}\frac{|G(q)|^2}{q_0^2ϵ_q^2}}`$ $`=`$ $`i2\pi {\displaystyle _0^{2\mu }}{\displaystyle \frac{dq_{}q_{}}{(2\pi )^2}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{dq_{||}}{2\pi }}|G(q_{||})|^2{\displaystyle \frac{idq_4}{2\pi }\frac{1}{q_4^2+ϵ_q^2}}`$ (A61) $`=`$ $`{\displaystyle \frac{\mu ^2}{\pi ^2}}{\displaystyle _0^{\mathrm{}}}𝑑q_{||}{\displaystyle \frac{|G(q_{||})|^2}{2ϵ_q}}`$ $``$ $`{\displaystyle \frac{\mu ^2}{2\pi ^2}}{\displaystyle _{G_0}^\mathrm{\Lambda }_{}}𝑑q_{||}{\displaystyle \frac{|G(q_{||})|^2}{q_{||}}}`$ $`=`$ $`{\displaystyle \frac{\mu ^2}{2\pi ^2}}G_0^2{\displaystyle _0^{x_0}}𝑑x\mathrm{sin}^2\left({\displaystyle \frac{\pi x}{2x_0}}\right)`$ $`=`$ $`{\displaystyle \frac{\mu ^2}{2\pi ^2}}G_0^2{\displaystyle \frac{x_0}{2}}={\displaystyle \frac{\mu ^2G_0^2x_0}{4\pi ^2}},`$ where logarithmic scales $`x=\mathrm{ln}(\mathrm{\Lambda }_{}/q_{||})`$ and $`x_0=\mathrm{ln}(\mathrm{\Lambda }_{}/G_0)`$ were used in the fourth line. Thus we have for $`\mathrm{\Delta }(A37)(A37)(A39)`$: $`\mathrm{\Delta }(A37)`$ $``$ $`{\displaystyle \frac{G_0^2x_0}{16\pi ^2F_T}}\{\mathrm{Tr}_{cf}\left([m,\tau ^\alpha ]_+(𝐌_{}^{\xi }{}_{}{}^{}m𝐌+𝐌^{}m𝐌^\xi )\right)`$ (A62) $`+\mathrm{Tr}_{cf}\left([m,\tau _{}^{\alpha }{}_{}{}^{}]_+(𝐌^\xi m𝐌^{}+𝐌m𝐌_{}^{\xi }{}_{}{}^{})\right)\}.`$ Since $`G_0`$ and $`F_T`$ are of order $`𝒪(\mu )`$, we find that $`\mathrm{\Delta }(A37)`$ is of order $`𝒪(\mu )`$. This means that the corresponding $`(M^2)^{\alpha \beta }`$ is of order $`𝒪(\mu ^0)`$, since there is an additional $`1/F_T`$ factor from (A56), namely $`\left(M^2\right)^{\alpha \beta }`$ $``$ $`{\displaystyle \frac{G_0^2x_0}{8\pi ^2F_T^2}}\{\mathrm{Tr}_{cf}\left([m,\tau ^\alpha ]_+(𝐌_{}^{\beta }{}_{}{}^{}m𝐌+𝐌^{}m𝐌^\beta )\right)`$ (A63) $`+\mathrm{Tr}_{cf}\left([m,\tau _{}^{\alpha }{}_{}{}^{}]_+(𝐌^\beta m𝐌^{}+𝐌m𝐌_{}^{\beta }{}_{}{}^{})\right)\}.`$ In the flavor-symmetric case $`m_u=m_d=m_s`$ ($`m_q`$), the color-flavor traces simplify to $$2m_q^2\mathrm{Tr}_{cf}\left(\tau ^\alpha \left\{𝐌_{}^{\xi }{}_{}{}^{}𝐌+𝐌^{}𝐌^\xi \right\}\right)=2m_q^2\mathrm{Tr}_{cf}\left(\tau _{}^{\alpha }{}_{}{}^{}\left\{𝐌^\xi 𝐌^{}+\mathrm{𝐌𝐌}_{}^{\xi }{}_{}{}^{}\right\}\right)=16m_q^2\delta ^{\alpha \xi }.$$ The mass of the generalized pion at next-to-leading order now reads $`M^2`$ $``$ $`{\displaystyle \frac{4G_0^2x_0}{\pi ^2F_T^2}}m_q^2={\displaystyle \frac{4m_q^2}{\mu ^2}}\left\{{\displaystyle \frac{4\mathrm{\Lambda }_{}^6}{\pi m_E^5}}\mathrm{exp}\left({\displaystyle \frac{3\pi ^2}{\sqrt{2}g}}\right)\right\}^2{\displaystyle \frac{3\pi ^2}{\sqrt{2}g}}`$ (A64) $`=`$ $`{\displaystyle \frac{2^{23}\pi ^{10}m_q^2}{3^4\sqrt{2}g^{11}}}\mathrm{exp}\left({\displaystyle \frac{3\sqrt{2}\pi ^2}{g}}\right),`$ where we have used that $`F_T=\mu /\pi `$, eq. (10), $`x_0=\frac{\sqrt{3}\pi }{2h_{}}`$, eq. (11), $`\mathrm{\Lambda }_{}=2\mu `$ and $`m_E=\sqrt{\frac{N_f}{2}}\frac{g\mu }{\pi }`$ with $`N_f=3`$. 4. Proof of the color-identity (6): Apply the standard identity for $`SU(N_c)`$ Gell-Mann matrices, $$\underset{a=1}{\overset{N_c^21}{}}\frac{1}{4}\lambda _{\alpha \beta }^a\lambda _{\gamma \delta }^a=\frac{1}{2}\delta _{\alpha \delta }\delta _{\gamma \beta }\frac{1}{2N}\delta _{\alpha \beta }\delta _{\gamma \delta },$$ (A65) to the expression $`_{a=1}^{N_c^21}\left(\frac{\lambda ^{aT}}{2}ϵ^A\frac{\lambda ^a}{2}\right)_{\alpha \delta }`$, i.e. $`{\displaystyle \underset{a=1}{\overset{N_c^21}{}}}\frac{1}{4}\lambda _{\alpha \beta }^{aT}ϵ^{A\beta \gamma }\lambda _{\gamma \delta }^a`$ $`=`$ $`{\displaystyle \underset{a=1}{\overset{N_c^21}{}}}\frac{1}{4}\lambda _{\beta \alpha }^a\lambda _{\gamma \delta }^aϵ^{A\beta \gamma }=\left(\frac{1}{2}\delta _{\beta \delta }\delta _{\gamma \alpha }\frac{1}{2N_c}\delta _{\beta \alpha }\delta _{\gamma \delta }\right)ϵ^{A\beta \gamma }`$ (A66) $`=`$ $`\frac{1}{2}\left(ϵ^{A\delta \alpha }\frac{1}{N_c}ϵ^{A\alpha \delta }\right)={\displaystyle \frac{N_c+1}{2N_c}}ϵ^{A\alpha \delta }`$ which is identical to (6) for the case $`N_c=3`$. It is easy to check that also $$\underset{a}{}\frac{\lambda ^a}{2}ϵ_c^A\frac{\lambda ^{aT}}{2}=\frac{N_c+1}{2N_c}ϵ_c^A$$ (A67) holds. Finally, by replacing $`\beta \alpha `$ in the bracket of the third term of (A66) one can easily derive $$\underset{a}{}\frac{\lambda ^a}{2}ϵ_c^A\frac{\lambda ^a}{2}=\frac{1}{2N_c}ϵ_c^A.$$ (A68) 5. Proof of the relations (21): The first identity follows immediately from (6) or (A66), if one inserts $`{\displaystyle \underset{a}{}}{\displaystyle \frac{\lambda ^{aT}}{2}}\left(𝐌^A\right){\displaystyle \frac{\lambda ^a}{2}}`$ $`=`$ $`{\displaystyle \underset{a}{}}{\displaystyle \frac{\lambda ^{aT}}{2}}\left(ϵ_f^Iϵ_c^J\gamma _5\left(\tau ^A\right)^{IJ}\right){\displaystyle \frac{\lambda ^a}{2}}`$ (A69) $`=`$ $`{\displaystyle \frac{N_c+1}{2N_c}}\left(ϵ_f^Iϵ_c^J\gamma _5\left(\tau ^A\right)^{IJ}\right)={\displaystyle \frac{N_c+1}{2N_c}}𝐌^A.`$ It is easy to see that the same relation holds, if $`𝐌^A`$ is replaced by $`𝐌_{}^{A}{}_{}{}^{}`$. In order to show the second relation of (21), we use that $$\mathrm{𝐌𝐌}^A𝐌=\gamma _5ϵ_f^Iϵ_c^Iϵ_f^J\left(\tau ^A\right)^{JK}ϵ_c^Kϵ_f^Lϵ_c^L$$ and that $`\left(ϵ_c^Iϵ_c^Kϵ_c^L\right)^{\alpha \beta }`$ $`=`$ $`ϵ_c^{I\alpha \beta }\delta ^{KL}+ϵ_c^{I\alpha L}\delta ^{K\beta },`$ $`\left(ϵ_f^Iϵ_f^Jϵ_f^L\right)^{ij}`$ $`=`$ $`ϵ_f^{Lij}\delta ^{IJ}+ϵ_f^{LIj}\delta ^{iJ}.`$ Therefore $`\left(\mathrm{𝐌𝐌}^A𝐌\right)^{\alpha i,\beta j}`$ $`=`$ $`\gamma _5ϵ_f^{Kij}ϵ_c^{J\alpha \beta }\left(\tau ^A\right)^{JK}+\mathrm{}`$ $`=`$ $`\gamma _5ϵ_f^{Kij}ϵ_c^{J\alpha \beta }\left(\tau _{}^{A}{}_{}{}^{T}\right)^{KJ}+\mathrm{}=\left(𝐌_{}^{A}{}_{}{}^{}\right)^{\alpha i,\beta j}+\mathrm{},`$ where the dots refer to terms which are symmetric in color-flavor and finally subleading to leading logarithm order. Naturally, also $`\left(\mathrm{𝐌𝐌}_{}^{A}{}_{}{}^{}𝐌\right)=𝐌^A+\mathrm{}`$ holds. Therefore the second relation of (21) approximately follows from the first one, if the above mentioned subleading terms are neglected. 6. The structure of the vector and axial-vector currents: The two-component Nambu-Gorkov field $$𝚿=\left(\begin{array}{c}\psi \\ C\overline{\psi }^T\end{array}\right)$$ transforms under vector and axial-vector transformations as follows $$𝐔_V𝚿=\left(\begin{array}{cc}e^{i{\scriptscriptstyle \frac{1}{2}}\tau ^A\alpha ^A}& 0\\ 0& e^{i{\scriptscriptstyle \frac{1}{2}}\tau _{}^{A}{}_{}{}^{}\alpha ^A}\end{array}\right)𝚿\text{and}𝐔_A𝚿=\left(\begin{array}{cc}e^{i{\scriptscriptstyle \frac{1}{2}}\gamma _5\tau ^A\beta ^A}& 0\\ 0& e^{i{\scriptscriptstyle \frac{1}{2}}\gamma _5\tau _{}^{A}{}_{}{}^{}\beta ^A}\end{array}\right)𝚿.$$ (A70) The vector and axial-vector currents are diagonal in the Nambu-Gorkov formalism, since they result as Noether currents from the diagonal kinetic term $$\overline{𝚿}i\gamma ^\mu _\mu 𝝆_0𝚿=\left(\begin{array}{cc}\overline{\psi }\gamma ^\mu _\mu \psi & 0\\ 0& \psi ^T\gamma _{}^{\mu }{}_{}{}^{T}_\mu \overline{\psi }^T\end{array}\right)=\overline{𝚿}\left(\begin{array}{cc}\gamma ^\mu _\mu & 0\\ 0& C\gamma _{}^{\mu }{}_{}{}^{T}C^1_\mu \end{array}\right)𝚿,$$ with $`𝝆_0`$ the unit matrix in the Nambu-Gorkov space. Alternatively, they can be derived by a prescription from which generalizes the standard current structure $`\overline{\psi }\mathrm{\Gamma }\psi `$ to the charge conjugated sector as $`\overline{\psi }_CC\mathrm{\Gamma }^TC^1\psi _C`$. Because of $`\gamma _\mu ^T=C^1\gamma _\mu C`$ and $`C^1\gamma _5C=\gamma _5^T=\gamma _5`$, we have $`𝐕_\mu ^A`$ $``$ $`\overline{𝚿}\left(\begin{array}{cc}\gamma _\mu \frac{1}{2}\tau ^A& 0\\ 0& C\left(\gamma _\mu \frac{1}{2}\tau ^A\right)^TC^1\end{array}\right)𝚿=\overline{𝚿}\left(\begin{array}{cc}\gamma _\mu \frac{1}{2}\tau ^A& 0\\ 0& \gamma _\mu \frac{1}{2}\tau _{}^{A}{}_{}{}^{}\end{array}\right)𝚿`$ (A75) $`=`$ $`\overline{𝚿}\gamma _\mu \frac{1}{2}𝐓^A𝝆_3𝚿,`$ (A76) and $`𝐀_\mu ^a`$ $``$ $`\overline{𝚿}\left(\begin{array}{cc}\gamma _\mu \gamma _5\frac{1}{2}\tau ^A& 0\\ 0& C\left(\gamma _\mu \gamma _5\frac{1}{2}\tau ^A\right)^TC^1\end{array}\right)𝚿=\overline{𝚿}\left(\begin{array}{cc}\gamma _\mu \gamma _5\frac{1}{2}\tau ^A& 0\\ 0& \gamma _\mu \gamma _5\frac{1}{2}\tau _{}^{A}{}_{}{}^{}\end{array}\right)𝚿`$ (A81) $`=`$ $`\overline{𝚿}\gamma _\mu \gamma _5\frac{1}{2}𝐓^A𝚿,`$ (A82) with $`𝐓^A=\mathrm{diag}(\tau ^A,\tau _{}^{A}{}_{}{}^{})`$ and $`𝝆_3`$ the standard Pauli matrix. In (A82) it was used that $`C\gamma _5^T\gamma _\mu ^TC^1=CC^1\gamma _5C(C^1\gamma _\mu C)C^1=\gamma _5\gamma _\mu =\gamma _\mu \gamma _5`$. Furthermore, note that the derivation of the gluon-vertex (13) is totally analogous to (A76). 7. The structure of the vertices for the generalized mesons: Because of the particle-particle (or hole-hole) substructure, all generalized mesons have the vertex structure $$\overline{𝚿}𝚪_M𝚿=\overline{𝚿}\left(\begin{array}{cc}0& \left(\mathrm{\Gamma }_M\right)_{12}\\ \left(\mathrm{\Gamma }_M\right)_{21}& 0\end{array}\right)𝚿.$$ (A83) By conjugating the 21-component $`\overline{\psi }_C\left(\mathrm{\Gamma }_M\right)_{21}\psi =\psi ^TC\left(\mathrm{\Gamma }_M\right)_{21}\psi `$, namely <sup>#16</sup><sup>#16</sup>#16The minus sign results from the Grassman property of the fermion spinors. Note that $`C=C^{}=C^1`$ and $`\gamma _0C^1=C\gamma _{}^{0}{}_{}{}^{T}`$. $$\psi ^{}\left\{\left(\mathrm{\Gamma }_M\right)_{21}\right\}^{}C^{}\psi _{}^{T}{}_{}{}^{}=\overline{\psi }\gamma ^0\left\{\left(\mathrm{\Gamma }_M\right)_{21}\right\}^{}\gamma ^0\gamma ^0C^1\psi _{}^{}{}_{}{}^{T}=\overline{\psi }\gamma ^0\left\{\left(\mathrm{\Gamma }_M\right)_{21}\right\}^{}\gamma ^0C(\overline{\psi })^T,$$ (A84) one can derive a general rule (see ) which links the 12 and 21 components of $`𝚪_M`$ $$\left(\mathrm{\Gamma }_M\right)_{12}=\gamma ^0\left\{\left(\mathrm{\Gamma }_M\right)_{21}\right\}^{}\gamma ^0.$$ (A85) Thus, in order to determine the structure of the generalized vertices, we only have to determine the structure of the 21-component. As in the standard case, the structure of the (generalized) meson vertices follows from the transformation properties of the (hermitean) bilinears $`\overline{𝚿}𝚪_M𝚿`$ under proper (ordinary continuous) Lorentz transformations $`\psi (x)\psi ^{}(x^{})=S(\mathrm{\Lambda })\psi (\mathrm{\Lambda }x)`$ with $`S(\mathrm{\Lambda })=\mathrm{exp}(\frac{i}{4}\sigma _{\alpha \beta }\omega ^{\alpha \beta })`$ and under the (discrete) parity transformation $`\psi (x)\psi ^{}(x^{})=\gamma _0\psi (t,𝐱)`$. Using that $`S(\mathrm{\Lambda })^T=CS(\mathrm{\Lambda })^1C^1`$ and $`\gamma _{}^{0}{}_{}{}^{T}C=\gamma ^0C=C\gamma ^0`$ , we can easily derive the following transformation properties of the bilinears $`\overline{\psi }_C\left(\mathrm{\Gamma }_M\right)_{21}\psi `$ under proper Lorentz and parity transformations (which generalize the transformation properties of the standard bilinears $`\overline{\psi }\mathrm{\Gamma }\psi `$, see e.g. ): $$\begin{array}{cccccc}\hfill \overline{\psi }_C^{}(x^{})\gamma _5\psi ^{}(x^{})& =& \overline{\psi }_C(x)\gamma _5\psi (x)\hfill & & \text{scalar}\hfill & 0^+,\hfill \\ \hfill \overline{\psi }_C^{}(x^{})\psi ^{}(x^{})& =& det(\mathrm{\Lambda })\overline{\psi }_C(x)\psi (x)\hfill & & \text{pseudoscalar}\hfill & 0^{},\hfill \\ \hfill \overline{\psi }_C^{}(x^{})\gamma ^\mu \gamma _5\psi ^{}(x^{})& =& \mathrm{\Lambda }_\nu ^\mu \overline{\psi }_C(x)\gamma ^\nu \gamma _5\psi (x)\hfill & & \text{vector}\hfill & 1^{},\hfill \\ \hfill \overline{\psi }_C^{}(x^{})\gamma _\mu \psi ^{}(x^{})& =& det(\mathrm{\Lambda })\mathrm{\Lambda }_\nu ^\mu \overline{\psi }_C(x)\gamma ^\nu \psi (x)\hfill & & \text{axial-vector}\hfill & 1^+.\hfill \end{array}$$ (A86) Note the appearance of the extra $`\gamma _5`$ relative to the standard rules of e.g. . Taking the flavor matrix in the color-flavor-locked way into account, we have the following 21 components of the generalized meson vertices: $$\begin{array}{ccccc}\hfill \left(\mathrm{\Gamma }_\sigma (p,P)\right)_{21}& =& i𝐌^A\mathrm{\Gamma }_S(p,P)/F_S\hfill & & \text{generalized sigma},\hfill \\ \hfill \left(\mathrm{\Gamma }_\pi (p,P)\right)_{21}& =& i\gamma _5𝐌^A\mathrm{\Gamma }_{PS}(p,P)/F_{PS}\hfill & & \text{generalized pion},\hfill \\ \hfill \left(\mathrm{\Gamma }_\mu (p,P)\right)_{21}& =& \gamma _\mu 𝐌^A\mathrm{\Gamma }_V(p,P)/F_V\hfill & & \text{generalized vector meson},\hfill \\ \hfill \left(\mathrm{\Gamma }_{5\mu }(p,P)\right)_{21}& =& i\gamma _\mu \gamma _5𝐌^A\mathrm{\Gamma }_{AV}(p,P)/F_{AV}\hfill & & \text{generalized axial-vector meson},\hfill \end{array}$$ (A87) where $`𝐌^A\gamma _5ϵ_f^aϵ_c^\alpha (\tau ^A)^{a\alpha }`$ and $`(ϵ^a)^{bc}=ϵ^{abc}`$. The form factors and decay constants have been introduced in for the pionic case and in (52) for the vector case. Note thay after inserting unity $`\mathrm{𝟏}=\mathrm{\Lambda }^+(𝐩)+\mathrm{\Lambda }^{}(𝐩)`$ to the left and right of the $`\left(\mathrm{\Gamma }_M(p,P)\right)_{21}`$’s, we can project these vertices onto the particle-particle, particle-antiparticle, antiparticle-particle and antiparticle-antiparticle sectors in analogy to (A3): $`\left(\mathrm{\Gamma }_M(p,P)\right)_{21}`$ $`=`$ $`\mathrm{\Lambda }^+(𝐩)\left(\mathrm{\Gamma }_M(p,P)\right)_{pp}\mathrm{\Lambda }^+(𝐩)+\mathrm{\Lambda }^+(𝐩)\left(\mathrm{\Gamma }_M(p,P)\right)_{pa}\mathrm{\Lambda }^{}(𝐩)`$ (A88) $`+\mathrm{\Lambda }^{}(𝐩)\left(\mathrm{\Gamma }_M(p,P)\right)_{ap}\mathrm{\Lambda }^+(𝐩)+\mathrm{\Lambda }^{}(𝐩)\left(\mathrm{\Gamma }_M(p,P)\right)_{aa}\mathrm{\Lambda }^{}(𝐩).`$ In the scalar and pseudoscalar case, the mixed (particle-antiparticle and antiparticle-particle) vertices vanish identically since $`\mathrm{\Lambda }^\pm (𝐩)\gamma _5\mathrm{\Lambda }^{}(𝐩)=0=\mathrm{\Lambda }^\pm (𝐩)\mathrm{\Lambda }^{}(𝐩)`$. The sum of the remaining particle-particle and antiparticle-antiparticle has exactly the structure of (A3). In the vector and axialvector case, the mixed terms $`\left(\mathrm{\Gamma }_M(p,P)\right)_{pa}`$ and $`\left(\mathrm{\Gamma }_M(p,P)\right)_{ap}`$ survive. However, in the leading-logarithm approximation only the particle-particle parts $`\left(\mathrm{\Gamma }_M(p,P)\right)_{pp}`$ of the vertices are needed. Contrary to the standard case, the phases cannot be determined from the hermiticity property of the quark bilinears, since the hermiticity is automatically satisfied under the condition (A85). In general, the form factors are complex-valued, such that the phases can be chosen at will. Our phase choice corresponds to real-valued form factors with attractive Bethe-Salpeter kernels (see (49) and (A98)). Taking the (A85) rule into account, we have $$\begin{array}{ccccc}\hfill 𝚪_\sigma ^A(p,P)& =& i𝝆_1𝐌_𝐓^A\mathrm{\Gamma }_S(p,P)/F_S\hfill & & \text{generalized sigma},\hfill \\ \hfill 𝚪_\pi ^A(p,P)& =& \gamma _5𝝆_2𝐌_𝐓^A\mathrm{\Gamma }_{PS}(p,P)/F_{PS}\hfill & & \text{generalized pion},\hfill \\ \hfill 𝚪_\mu ^A(p,P)& =& \gamma _\mu 𝝆_1𝐌_𝐓^A\mathrm{\Gamma }_V(p,P)/F_V\hfill & & \text{generalized vector meson},\hfill \\ \hfill 𝚪_{5\mu }^A(p,P)& =& \gamma _\mu \gamma _5𝝆_2𝐌_𝐓^A\mathrm{\Gamma }_{AV}(p,P)/F_{AV}\hfill & & \text{generalized axial-vector meson},\hfill \end{array}$$ (A89) where $`𝐌_𝐓^A\left(\begin{array}{cc}𝐌^A& 0\\ 0& 𝐌_{}^{A}{}_{}{}^{}\end{array}\right)=\left(\begin{array}{cc}𝐌^{i\alpha }\left(\tau ^A\right)^{i\alpha }& 0\\ 0& 𝐌^{i\alpha }\left(\tau _{}^{A}{}_{}{}^{}\right)^{i\alpha }\end{array}\right)=𝐌^{i\alpha }\left(𝐓^A\right)^{i\alpha }`$ (A94) and $`𝝆_1`$ and $`𝝆_2`$ are the standard Pauli matrices acting on the Nambu-Gorkov indices. As mentioned above, the form factors $`\mathrm{\Gamma }_S(p,P)`$, $`\mathrm{\Gamma }_{PS}(p,P)`$, $`\mathrm{\Gamma }_V(p,P)`$ and $`\mathrm{\Gamma }_{AV}(p,P)`$ are now assumed to be real, where $`\mathrm{\Gamma }_{PS}(p,0)=G(p)`$. The vertex structure of (A89) is in agreement with (16) and (19) as well as (52) as used in (45). If $`\left(\mathrm{\Gamma }_M(p,P)\right)_{21}`$ is projected onto the particle-particle, particle-antiparticle, antiparticle-particle and antiparticle-antiparticle sectors as in (A88), it follows from the (A85)-rule and $`\gamma ^0\mathrm{\Lambda }^\pm (𝐩)\gamma ^0=\mathrm{\Lambda }^{}(𝐩)`$ that the corresponding contributions of the 1-2 component read $`\left(\mathrm{\Gamma }_M(p,P)\right)_{12}`$ $`=`$ $`\mathrm{\Lambda }^{}(𝐩)\gamma ^0\left(\mathrm{\Gamma }_M(p,P)\right)_{pp}^{}\gamma ^0\mathrm{\Lambda }^{}(𝐩)+\mathrm{\Lambda }^{}(𝐩)\gamma ^0\left(\mathrm{\Gamma }_M(p,P)\right)_{pa}^{}\gamma ^0\mathrm{\Lambda }^+(𝐩)`$ $`+\mathrm{\Lambda }^+(𝐩)\gamma ^0\left(\mathrm{\Gamma }_M(p,P)\right)_{ap}^{}\gamma ^0\mathrm{\Lambda }^{}(𝐩)+\mathrm{\Lambda }^+(𝐩)\gamma ^0\left(\mathrm{\Gamma }_M(p,P)\right)_{aa}^{}\gamma ^0\mathrm{\Lambda }^+(𝐩)`$ in analogy to (A4). In the leading logarithm approximation, only the first term, i.e. the particle-particle term, is needed. In this case the vertex structure (A89) has to be modified by the “sandwich-rule” (35) which is compatible with (A88) and (LABEL:pp-12-projections). As mentioned in section 3 (see also Appendix-12), the simplified vertex structure (A89) can be used, if only leading propagators (LABEL:PropPart<sup>#17</sup><sup>#17</sup>#17In the scalar and pseudoscalar case, it already is sufficient that only one leading propagator is coupled per vertex. are coupled to the vertex and if the simplified form (8) of the gluon propagator is used. 8. Derivation of equation (20) from (12) and equation (53) from (49): Defining $`Qq+P/2`$ and $`KqP/2`$, the 12 component of (49) reads: $`\left(\mathrm{\Gamma }_j^A(p,P)\right)_{12}`$ $`=`$ $`g^2{\displaystyle }{\displaystyle \frac{d^4q}{(2\pi )^4}}i𝒟(pq)\gamma _\mu {\displaystyle \frac{\lambda ^a}{2}}[S_{11}(Q)\left(\mathrm{\Gamma }_j^A(q,P)\right)_{12}S_{22}(K)`$ (A96) $`+S_{12}(Q)\left(\mathrm{\Gamma }_j^A(q,P)\right)_{21}S_{12}(K)](\gamma ^\mu ){\displaystyle \frac{\lambda _{}^{a}{}_{}{}^{T}}{2}}.`$ The expression for $`\left(\mathrm{\Gamma }_j^A(p,P)\right)_{21}`$ follows from (A96) with the replacements $`12`$ and $`\lambda ^a\lambda _{}^{a}{}_{}{}^{T}`$. We now write $`\left(\mathrm{\Gamma }_j^A\right)_{12}=\frac{1}{F_V}\gamma _j\mathrm{\Gamma }_V𝐌_{}^{A}{}_{}{}^{}`$ where we assumed $`\mathrm{\Gamma }_V`$ to be real. Inserting then (LABEL:PropPart) for the propagators, we can transform (A96) to <sup>#18</sup><sup>#18</sup>#18In the following, we will use the simplified form (A89) of the generalized-meson vertex. The results for the sandwiched form (35) will be discussed at the end of this section. $`\gamma _j\mathrm{\Gamma }_V(p,P)𝐌_{}^{A}{}_{}{}^{}`$ $`=`$ $`g^2{\displaystyle \frac{d^4q}{(2\pi )^4}\frac{i𝒟(pq)\mathrm{\Gamma }_V(q,P)}{(Q_0^2ϵ_Q^2)(K_0^2ϵ_K^2)}}`$ (A97) $`\times \{Q_+K_{}\left[\gamma _\mu \gamma ^0\mathrm{\Lambda }^{}(𝐐)\gamma _j\gamma ^0\mathrm{\Lambda }^+(𝐊)\gamma ^\mu \right]\left[{\displaystyle \frac{\lambda ^a}{2}}𝐌_{}^{A}{}_{}{}^{}{\displaystyle \frac{\lambda _{}^{a}{}_{}{}^{T}}{2}}\right]`$ $`+G(Q)G(K)\gamma _\mu \mathrm{\Lambda }^+(𝐐)\gamma _j\mathrm{\Lambda }^+(𝐊)\gamma ^\mu \left[{\displaystyle \frac{\lambda ^a}{2}}\mathrm{𝐌𝐌}^A𝐌{\displaystyle \frac{\lambda _{}^{a}{}_{}{}^{T}}{2}}\right]\},`$ where it was used that both $`𝐌`$ (=$`𝐌^{}`$) and $`𝐌^A`$ contain a $`\gamma _5`$ matrix. Note that the corresponding expression for the composite axial-vector meson differs from (A97) by the replacement $`𝐌^Ai\gamma _5𝐌^A`$ and by an additional minus sign in front of the second term on the right hand side. Using (21) for the flavor contractions and moving the $`\gamma ^0`$ matrices through, we finally get $`\gamma _j\mathrm{\Gamma }_V(p,P)`$ $`=`$ $`{\displaystyle \frac{2g^2}{3}}{\displaystyle \frac{d^4q}{(2\pi )^4}\frac{i𝒟(pq)\mathrm{\Gamma }_V(q,P)}{(Q_0^2ϵ_Q^2)(K_0^2ϵ_K^2)}}`$ (A98) $`\times \left\{Q_+K_{}G(Q)G(K)\right\}\left[\gamma _\mu \mathrm{\Lambda }^+(𝐐)\gamma _j\mathrm{\Lambda }^+(𝐊)\gamma ^\mu \right],`$ where we ignored a symmetric contribution in color-flavor which is subleading to logarithmic accuracy. This equation will be simplified by applying the longitudinal and transverse projection (46) and (47), respectively, and then taking the Dirac trace. The left hand side becomes $`4\mathrm{\Gamma }_{L,T}(p,P)`$, whereas on the right hand side the Dirac structure becomes $`\frac{1}{3}\mathrm{Tr}\left[\gamma ^j\gamma _\mu \mathrm{\Lambda }^+(𝐐)\gamma _j\mathrm{\Lambda }^+(𝐊)\gamma ^\mu \right]`$ $`=`$ $`\frac{2}{3}\mathrm{Tr}\left[\gamma ^j\mathrm{\Lambda }^+(𝐐)\gamma _j\mathrm{\Lambda }^+(𝐊)\right]`$ (A99) $`=`$ $`{\displaystyle \frac{2}{3}}\mathrm{Tr}\left[\left(3\mathrm{\Lambda }^+(𝐐)\widehat{𝐐}^k\alpha ^k\right)\mathrm{\Lambda }^+(𝐊)\right],`$ where we summed over the index $`j`$ and used the Dirac matrix identity $`\gamma _\mu \gamma _j\gamma ^\mu =2\gamma _j`$. In the rest frame, the trace (A99) is just $`8/3`$. Inserting this back into (A98) and dividing by four, we get the quoted result (53) which identically holds in the axial-vector case. For deriving the corresponding expression of the generalized pion from (12), the $`\gamma _j`$ matrices in (A97) and (A98) have to be replaced by $`i\gamma _5`$. After multiplying both sides with $`i\gamma _5`$ and then taking the Dirac trace, we still get $`4\mathrm{\Gamma }_{PS}(p,P)`$ on the left hand side, whereas on the right hand the Dirac trace $$\mathrm{Tr}[\gamma _5\gamma _\mu \mathrm{\Lambda }^+(𝐐)\gamma _5\mathrm{\Lambda }^+(𝐊)\gamma ^\mu ]=\mathrm{Tr}[\gamma ^\mu \gamma _\mu \mathrm{\Lambda }^+(𝐐)\mathrm{\Lambda }^+(𝐊)]=4\mathrm{T}\mathrm{r}[\mathrm{\Lambda }^+(𝐐)\mathrm{\Lambda }^+(𝐊)]$$ reduces to a factor $`8`$ instead of $`8/3`$ in the rest frame. This is the reason why the prefactor in the Bethe-Salpeter kernel of the generalized pion (see (20) or (7) for the gap itself) is three times bigger than the one of the generalized vector and axial-vector (see (53)). For the generalized sigma, the $`\gamma _j`$ in (A97) and (A98) has to be replace by the i times the unit matrix. Furthermore, there is an additional minus sign in front of the second term on the right hand side of (A97) and an additional overall minus sign on the right hand side of (A98). After multiplying both sides with $`i`$ and taking the Dirac trace, there is still $`4\mathrm{\Gamma }_S(p,P)`$ on the left hand side, whereas on the right hand side the Dirac trace $$\mathrm{Tr}[\gamma _\mu \mathrm{\Lambda }^+(𝐐)\mathrm{\Lambda }^+(𝐊)\gamma ^\mu ]=4\mathrm{T}\mathrm{r}[\mathrm{\Lambda }^+(𝐐)\mathrm{\Lambda }^+(𝐊)]$$ reduces to $`+8`$ in the rest frame. This opposite sign, relative to the pion case, cancels against the above mentioned opposite sign on the right hand side of (A98). Thus the Bethe-Salpeter equation of the generalized sigma and pion are the same, see (20). If the sandwiched form of the generalized mesons is used, equation (A98) must be first projected from the left and right with $`\mathrm{\Lambda }^{}(p)`$, before we can multiply with $`\gamma _j`$ (or correspondingly with $`i\gamma _5`$ in the pion or $`i\mathrm{𝟏}`$ in the scalar case) and take the Dirac trace. This modification induces a factor $`\frac{1}{2}`$ on the left and right hand side of the projected (A98) and therefore does not change the final answer (53) and (20) for the gap equations for the (axial-)vector and (pseudo-)scalar case, respectively. The result of (A98) is valid for the phase choice of (A87), i.e. for real-valued form factors. If the opposite phase-choice had been made, i.e. if the form factors were assumed to be purely imaginary-valued, the $`Q_+K_{}G(Q)G(K)`$ term in (A98) would have to read $`Q_+K_{}+G(Q)G(K)`$ instead. The error which this choice will induce can be estimated by perturbation theory by inserting (9) and (58) or an analogous form factor into $$\frac{h_{}^2}{18}_{G_M}^\mathrm{\Lambda }_{}\frac{dq_{||}}{q_{||}}\frac{2G^2(q_{||})}{q_{||}^2}\mathrm{ln}\left(\frac{\mathrm{\Lambda }_{}^2}{(p_{||}q_{||})^2}\right)\mathrm{\Gamma }(q_{||},M_V)$$ (A100) as very small, i.e. $`𝒪(h_{}^2)`$ relatively to $`\mathrm{\Gamma }(q_{||},M_V)`$. 9. The structure of the vertices for the standard mesons: The Bethe-Salpeter kernels of the diagonal standard$`\overline{q}q`$” -type mesons in the CFL phase are $$\begin{array}{ccccc}\hfill \stackrel{~}{𝚪}_\sigma ^A(p,P)& =& 𝝆_0𝐍_{\stackrel{~}{T}}^A\stackrel{~}{\mathrm{\Gamma }}_S(p,P)/\stackrel{~}{F}_\mathrm{\Sigma }\hfill & & \text{standard sigma},\hfill \\ \hfill \stackrel{~}{𝚪}_\pi ^A(p,P)& =& i\gamma _5𝝆_0𝐍_{\stackrel{~}{T}}^A\stackrel{~}{\mathrm{\Gamma }}_{PS}(p,P)/\stackrel{~}{F}\hfill & & \text{standard pion},\hfill \\ \hfill \stackrel{~}{𝚪}_\mu ^A(p,P)& =& \gamma _\mu 𝝆_3𝐍_{\stackrel{~}{T}}^A\stackrel{~}{\mathrm{\Gamma }}_V(p,P)/\stackrel{~}{F}_V\hfill & & \text{standard vector meson},\hfill \\ \hfill \stackrel{~}{𝚪}_{5\mu }^A(p,P)& =& \gamma _\mu \gamma _5𝝆_0𝐍_{\stackrel{~}{T}}^A\stackrel{~}{\mathrm{\Gamma }}_{AV}(p,P)/\stackrel{~}{F}_{AV}\hfill & & \text{standard axial-vector meson},\hfill \end{array}$$ (A101) where $`𝐍_{\stackrel{~}{T}}^A=ϵ_f^aϵ_c^\alpha (\stackrel{~}{T}^A)^{a\alpha }`$. Thus $`𝐍_{\stackrel{~}{T}}^A`$ is of the same form as $`𝐌_𝐓^A`$, without the $`\gamma _5`$, however, and with $`𝐓^A=\mathrm{diag}(\tau ^A,\tau _{}^{A}{}_{}{}^{})`$ replaced by $`\stackrel{~}{T}^A=\mathrm{diag}(\stackrel{~}{\tau }^A,\stackrel{~}{\tau }^A)`$, where $`\stackrel{~}{\tau }_{1,3}=\tau _{1,3}`$ and $`\stackrel{~}{\tau }_2=i\tau _2`$. Furthermore $`𝝆_3`$ is the usual Pauli matrix, whereas $`𝝆_0`$ is the corresponding unit matrix. We have checked that the Bethe-Salpeter equations resulting from (A101) vanish identically. Standard scalar, pseudoscalar, vector and axial-vector excitations are not supported by the QCD superconductor in leading logarithm approximation. 10. From equation (22) to equation (27) and other variational approximations: After multiplying (22) with $`3/4g^2`$ and using the Fourier-transformations $`\mathrm{\Gamma }(p,M)`$ $`=`$ $`{\displaystyle d^4xe^{ipx}\mathrm{\Gamma }(x)},`$ $`𝒟(pq)`$ $`=`$ $`{\displaystyle d^4xe^{i(pq)x}𝒟(x)},`$ we can transform (22) into $$\frac{3}{4g^2}\mathrm{\Gamma }(x)=\frac{d^4q}{(2\pi )^4}e^{iqx}i𝒟(x)\frac{q_0^2ϵ_q^2M^2/4}{(q_0^2ϵ_q^2+M^2/4)^2M^2q_0^2}\mathrm{\Gamma }(q),$$ where $`\mathrm{\Gamma }(q)\mathrm{\Gamma }(q,M)`$. Multiplying both sides with $`\mathrm{\Gamma }(x)/i𝒟(x)`$ and integrating over $`x`$, we get $`{\displaystyle \frac{3}{4g^2}}{\displaystyle d^4x\frac{\mathrm{\Gamma }^2(x)}{i𝒟(x)}}`$ $`=`$ $`{\displaystyle \frac{d^4q}{(2\pi )^4}\left\{d^4xe^{iqx}\mathrm{\Gamma }(x)\right\}\frac{q_0^2ϵ_q^2M^2/4}{(q_0^2ϵ_q^2+M^2/4)^2M^2q_0^2}\mathrm{\Gamma }(q)}`$ $`=`$ $`{\displaystyle \frac{d^4q}{(2\pi )^4}\mathrm{\Gamma }(q)\frac{q_0^2ϵ_q^2M^2/4}{(q_0^2ϵ_q^2+M^2/4)^2M^2q_0^2}\mathrm{\Gamma }(q)}.`$ Assuming that $`\mathrm{\Gamma }(q)`$ is an even function in $`q`$ in analogy to $`G(q)`$ and Taylor-expanding the right-hand side in $`M^2`$, we get (24), i.e. $`{\displaystyle \frac{3}{4g^2}}{\displaystyle d^4x\frac{\mathrm{\Gamma }^2(x)}{i𝒟(x)}}`$ $``$ $`{\displaystyle \frac{d^4q}{(2\pi )^4}\frac{\mathrm{\Gamma }^2(q)}{q_0^2ϵ_q^2}}+{\displaystyle \frac{M^2}{4}}{\displaystyle \frac{d^4q}{(2\pi )^4}\frac{q_0^2+3ϵ_q^2}{(q_0^2ϵ_q^2)^3}\mathrm{\Gamma }^2(q)}`$ $`=`$ $`{\displaystyle \frac{d^4q}{(2\pi )^4}\frac{\mathrm{\Gamma }^2(q)}{q_0^2ϵ_q^2}}iM^2F^2,`$ where definition (25) was used. By assuming that $`\mathrm{\Gamma }(q)=\kappa G(q)`$ is an even real-valued function of $`q_{||}`$, by Wick-rotating to Euclidean space and evaluating the resulting $`q_4`$ integration as a contour integration, we can calculate $`F^2`$ defined in the last equation as follows: $`F^2`$ $``$ $`{\displaystyle \frac{i}{4}}{\displaystyle \frac{d^4q}{(2\pi )^4}\frac{q_0^2+3ϵ_q^2}{(q_0^2ϵ_q^2)^3}\mathrm{\Gamma }^2(q)}`$ $`=`$ $`{\displaystyle \frac{1}{4}}\mathrm{\hspace{0.17em}2}\pi {\displaystyle _0^{2\mu }}{\displaystyle \frac{q_{}dq_{}}{(2\pi )^2}}2{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dq_{||}}{2\pi }}\mathrm{\Gamma }^2(q_{||}){\displaystyle _{\mathrm{}}^+\mathrm{}}{\displaystyle \frac{dq_4}{2\pi }}{\displaystyle \frac{q_4^2+ϵ_q^2}{(q_4iϵ_q)^3(q_4+iϵ_q)^3}}`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle \frac{\mu ^2}{\pi ^2}}{\displaystyle _0^{\mathrm{}}}𝑑q_{||}\mathrm{\Gamma }^2(q_{||}){\displaystyle \frac{2\pi i}{2\pi }}{\displaystyle \frac{1}{i2ϵ_q^3}}`$ $`=`$ $`{\displaystyle \frac{\mu ^2}{8\pi ^2}}{\displaystyle _0^{\mathrm{}}}𝑑q_{||}{\displaystyle \frac{\mathrm{\Gamma }^2(q_{||})}{ϵ_q^3}}={\displaystyle \frac{\kappa ^2\mu ^2}{8\pi ^2}}{\displaystyle _0^{\mathrm{}}}𝑑q_{||}{\displaystyle \frac{G^2(q_{||})}{ϵ_q^3}}{\displaystyle \frac{\kappa ^2\mu ^2}{8\pi ^2}}{\displaystyle _{G_0}^\mathrm{\Lambda }_{}}𝑑q_{||}{\displaystyle \frac{G^2(q_{||})}{q_{||}^3}}.`$ After inserting (9) into the last equation and shifting to the logarithmic scales $`x=\mathrm{ln}(\mathrm{\Lambda }_{}/q_{||})`$ and $`x_0=\mathrm{ln}(\mathrm{\Lambda }_{}/G_0)`$, one finally arrives at (27), i.e. $$F^2\frac{\kappa ^2\mu ^2}{8\pi ^2}_0^{x_0}𝑑xe^{2(xx_0)}\mathrm{sin}^2(\pi x/2x_0)=\frac{\kappa ^2\mu ^2}{8\pi ^2}\frac{8x_0^2+\pi ^2e^{2x_0}\pi ^2}{16x_0^2+4\pi ^2}\frac{\kappa ^2\mu ^2}{16\pi ^2},$$ (A102) since $`x_01`$, see . Finally note the variational approximation $`{\displaystyle \frac{d^4q}{(2\pi )^4}\frac{\mathrm{\Gamma }_{T,L}^2(q)}{q_0^2ϵ_q^2}}`$ $`=`$ $`2\pi {\displaystyle _0^{2\mu }}{\displaystyle \frac{q_{}dq_{}}{(2\pi )^2}}2{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dq_{||}}{2\pi }}\mathrm{\Gamma }_{T,L}^2(q_{||}){\displaystyle \frac{idq_4}{2\pi }\frac{1}{q_4^2ϵ_q^2}}`$ $`=`$ $`i{\displaystyle \frac{\mu ^2}{\pi ^2}}{\displaystyle _0^{\mathrm{}}}𝑑q_{||}{\displaystyle \frac{\mathrm{\Gamma }_{T,L}^2(q_{||})}{2ϵ_q}}`$ $``$ $`i{\displaystyle \frac{\kappa ^2\mu ^2}{2\pi ^2}}{\displaystyle _{G_0}^\mathrm{\Lambda }_{}}𝑑q_{||}{\displaystyle \frac{G_M^2\mathrm{sin}^2\left(\frac{h_{}}{3}\mathrm{ln}\left(\frac{\mathrm{\Lambda }_{}}{q_{||}}\right)\right)}{q_{||}}}`$ $`=`$ $`i{\displaystyle \frac{\kappa ^2\mu ^2G_M^2}{2\pi ^2}}{\displaystyle _0^{x_0}}𝑑x\mathrm{sin}^2\left({\displaystyle \frac{\pi x}{2x_M}}\right)`$ $`=`$ $`i{\displaystyle \frac{\kappa ^2\mu ^2G_M^2}{4\pi ^2}}x_0\left(1{\displaystyle \frac{x_M}{\pi x_0}}\mathrm{sin}\left({\displaystyle \frac{\pi x_0}{x_M}}\right)\right),`$ which is used to derive (66) from the vector-meson analog of (24). 11. Gluon polarization function in the CFL phase: In the CFL phase, the gluon polarization function for the bare gluon ($`AA`$) fields is $$\mathrm{\Pi }_{\mu \nu }^{ab}(Q)=g^2\frac{d^4q}{(2\pi )^4}\mathrm{Tr}\left(i𝒱_\mu ^ai𝐒(q+\frac{Q}{2})i𝒱_\nu ^bi𝐒(q\frac{Q}{2})\right).$$ (A103) Using $`\mathrm{Tr}_{cf}(\lambda ^a\lambda ^b)=6\delta ^{ab}`$ and $$\mathrm{Tr}_{cf}(\lambda ^a𝐌\lambda _{}^{b}{}_{}{}^{T}𝐌)=2\mathrm{T}\mathrm{r}_c(\lambda ^a\lambda ^b)=4\delta ^{ab},$$ we may write (A103) in the following form $`\mathrm{\Pi }_{\mu \nu }^{ab}(Q)`$ $`=`$ $`g^2\delta ^{ab}{\displaystyle \frac{d^4q}{(2\pi )^4}\frac{1}{(K_0^2ϵ_K^2)(P_0^2ϵ_P^2)}}`$ (A104) $`\times \{{\displaystyle \frac{}{}}3(K_0P_0+K_{||}P_{||})\mathrm{Tr}\left[\gamma _\mu \gamma ^0\mathrm{\Lambda }^{}(K)\gamma _\nu \gamma ^0\mathrm{\Lambda }^{}(P)\right]`$ $`+2G(K)G(P)\mathrm{Tr}\left[\gamma _\mu \mathrm{\Lambda }^+(K)\gamma _\nu \mathrm{\Lambda }^{}(P)\right]{\displaystyle \frac{}{}}\}`$ with $`K=q+Q/2`$ and $`P=qQ/2`$. For the temporal polarization, we have $$\mathrm{\Pi }_{00}^{ab}(Q)=g^2\delta ^{ab}\frac{d^4q}{(2\pi )^4}\frac{3(K_0P_0+K_{||}P_{||})+2G(K)G(P)}{(K_0^2ϵ_K^2)(P_0^2ϵ_P^2)}\left(1+\widehat{K}\widehat{P}\right).$$ (A105) For $`Q=0`$, this simplifies to $`\mathrm{\Pi }_{00}^{ab}(0)`$ $`=`$ $`g^2\delta ^{ab}{\displaystyle \frac{d^4q}{(2\pi )^4}\frac{6(q_0^2+q_{||}^2)+4G^2(q)}{(q_0^2ϵ_q^2)^2}}`$ (A106) $`=`$ $`\delta ^{ab}{\displaystyle \frac{g^2\mu ^2}{\pi ^2}}{\displaystyle _0^{\mathrm{}}}𝑑q_{||}{\displaystyle _{\mathrm{}}^+\mathrm{}}{\displaystyle \frac{idq_4}{2\pi }}{\displaystyle \frac{6q_4^2+6ϵ_q^22G^2(q)}{(q_4^2+ϵ_q^2)^2}}`$ $`=`$ $`\delta ^{ab}{\displaystyle \frac{g^2\mu ^2}{\pi ^2}}{\displaystyle _0^{\mathrm{}}}𝑑q_{||}{\displaystyle \frac{iG^2(q)}{2ϵ_q^3}}=i\delta ^{ab}\left({\displaystyle \frac{g\mu }{2\pi }}\right)^2.`$ Here we have used that the angle and $`q_{}`$ integrations contribute a factor $`2\pi 2\mu ^2`$ and, after a Wick rotation to Euclidean space and contour integration, that $`_0^{\mathrm{}}𝑑q_{||}G^2(q)/ϵ_q^3=1/2`$ (see above). For the spatial polarization, we obtain $$\mathrm{\Pi }_{ij}^{ab}(Q)=g^2\delta ^{ab}\frac{d^4q}{(2\pi )^4}\frac{3(K_0P_0+K_{||}P_{||})2G(K)G(P)}{(K_0^2ϵ_K^2)(P_0^2ϵ_P^2)}\left(g_{ij}(1\widehat{K}\widehat{P})\widehat{K}_i\widehat{P}_j\widehat{K}_j\widehat{P}_i\right)$$ (A107) after using the spin trace $$\mathrm{Tr}[\gamma ^i\mathrm{\Lambda }^\pm (𝐊)\gamma ^j\mathrm{\Lambda }^{}(𝐏)]=g^{ij}(g^{im}g^{jn}g^{ij}g^{mn}+g^{in}g^{jm})\widehat{𝐊}^m\widehat{𝐏}^n.$$ (A108) For $`Q=0`$, this simplifies $`\mathrm{\Pi }_{ij}^{ab}(0)`$ $`=`$ $`g^2\delta ^{ab}{\displaystyle \frac{d^4q}{(2\pi )^4}\frac{6(q_0^2+q_{||}^2)4G^2(q)}{(q_0^2ϵ_q^2)^2}\widehat{𝐪}_i\widehat{𝐪}_j}`$ (A109) $`=`$ $`\frac{1}{3}g^2\delta ^{ab}\delta _{ij}{\displaystyle \frac{d^4q}{(2\pi )^4}\frac{6q_0^2+6ϵ_q^210G^2(q)}{(q_0^2ϵ_q^2)^2}}`$ $`=`$ $`\frac{5}{6}\delta ^{ab}\delta _{ij}{\displaystyle \frac{g^2\mu ^2}{\pi ^2}}{\displaystyle _0^{\mathrm{}}}𝑑q_{||}{\displaystyle \frac{iG^2(q)}{ϵ_q^3}}=i\delta ^{ab}\delta _{ij}{\displaystyle \frac{5g^2\mu ^2}{12\pi ^2}}.`$ The lack of transversality in the $`AA`$ polarization, which is manifest in (A106) and (A109), is fixed by the mixing with the scalars (Higgs mechanism) and the additional contribution from the modes within the Fermi surface (nonsurface modes). 12. The comparison of the exact and simplified in-medium gluon propagator: According to eq. (6.51) of and eq. (10) of , the in-medium retarded (Minkowski-space) gluon-propagator in a general covariant gauge reads (modulo an overall phase factor) $$𝒟_{\mu \nu }(q)=i\frac{P_{\mu \nu }^T}{q^2G}+i\frac{P_{\mu \nu }^L}{q^2F}i\xi \frac{P_{\mu \nu }^{GF}}{q^2},$$ (A110) where $`Fm_D^2=\frac{N_f}{2\pi ^2}g^2\mu ^2`$ ($`m_E^2`$) and $`G=\frac{\pi }{4}\frac{iq_0}{|𝐪|}m_D^2`$ ($`m_M^2`$). The propagator contains the gauge parameter $`\xi `$, which must not appear in physical results. The projectors appearing in (A110) read $`P_{\mu \nu }^T`$ $`=`$ $`(1g_{\mu 0})(1g_{\nu 0})\left(g_{\mu \nu }{\displaystyle \frac{q_\mu q_\nu }{𝐪^2}}\right),`$ (A111) $`P_{\mu \nu }^L`$ $`=`$ $`g_{\mu \nu }+{\displaystyle \frac{q_\mu q_\nu }{q^2}}P_{\mu \nu }^T,`$ (A112) $`P_{\mu \nu }^{GF}`$ $`=`$ $`{\displaystyle \frac{q_\mu q_\nu }{q^2}},`$ (A113) where $`q^2=(q^0)^2𝐪^2`$ and $`g_{\mu \nu }=g^{\mu \nu }=\mathrm{diag}(1,1,1,1)`$ for $`\mu ,\nu =0,1,2,3`$. The transverse projector can be written as $`P_{\mu 0}^T`$ $`=`$ $`P_{0\nu }^T=0,`$ (A114) $`P_{ij}^T`$ $`=`$ $`P_{ji}^T=\delta _{ij}\widehat{𝐪}_i\widehat{𝐪}_j,`$ (A115) where $`\widehat{𝐪}_iq_i/|𝐪|`$. Eq. (A110) should be compared with the simplified form $$𝒟_{\mu \nu }(q)=i\frac{1}{2}\frac{g_{\mu \nu }}{q^2G}+i\frac{1}{2}\frac{g_{\mu \nu }}{q^2F},$$ (A116) which is the analog in Minkowski space of the screened perturbative gluon propagator in Euclidean space used here in (8), see also . We now proceed to show that (A110-A116) yield equivalent results in the context of our analysis. The three-momentum can be split into the Fermi-momentum $`𝐪_F`$ and a momentum $`\stackrel{}{l}`$ measured relative to the Fermi surface, $`|𝐪|=|𝐪_F+\stackrel{}{l}|=\mu +l_{||}+{\displaystyle \frac{l_{}^2}{2\mu }}+𝒪(1/\mu ^2),`$ (A117) where $`l_{||}`$ and $`l_{}`$ are the projections of the relative momentum in the direction of and orthogonal to the Fermi-momentum $`𝐏`$, respectively. Because of the decomposition (A117), we have, modulo $`1/\mu ^2`$ corrections, $`q^2=q_0^2𝐪^2𝐪^2`$ and we can simplify the longitudinal projector as follows (see )) $`P_{\mu \nu }^L`$ $``$ $`g_{\mu 0}g_{\nu 0}.`$ (A118) Finally we will use that $`\delta ^{ij}P_{ij}^T`$ $`=`$ $`2,`$ (A119) $`\widehat{𝐪}\widehat{𝐐}\widehat{𝐤}\widehat{𝐐}`$ $`=`$ $`\frac{1}{2}\left(1\widehat{𝐪}\widehat{𝐤}\right)+𝒪(1/\mu ),`$ (A120) where $`Qkq`$. The second formula can be derived with the help of eq. (A117) applied to $`|𝐤|`$, $`|𝐪|`$ and $`|𝐐|=|𝐤𝐪|`$, i.e., $`|𝐐|^2=|𝐤𝐪|^22\mu ^2\left(1\widehat{𝐪}\widehat{𝐤}\right)`$ and $`\widehat{𝐪}𝐐\widehat{𝐤}𝐐=(|𝐤|\widehat{𝐪}\widehat{𝐤}|𝐪|)(|𝐤||𝐪|\widehat{𝐪}\widehat{𝐤})`$ $`\mu ^2(1\widehat{𝐪}\widehat{𝐤})^2`$. The Dirac structure of the gap equation of (and of the Bethe-Salpeter equation (12) in case the generalized-pion-vertex follows the sandwich rule (35), i.e., the projectors $`\frac{1}{2}(1\pm 𝜶\widehat{𝐪})`$ are kept in the generalized-pion-vertex (19)) reads: $`A^{\mu \nu }`$ $``$ $`\frac{1}{2}\mathrm{Tr}\left[\gamma ^\mu \frac{1}{2}(1s_R\gamma ^0\gamma ^m\widehat{𝐪}^m)\gamma ^\nu \frac{1}{2}(1+s_L\gamma ^0\gamma ^n\widehat{𝐤}^n)\right]`$ (A121) $`=\frac{1}{2}g^{\mu \nu }+{\displaystyle \frac{s_L}{2}}\left(g^{\mu n}g^{\nu 0}g^{\mu 0}g^{\nu n}\right)\widehat{𝐤}^n+{\displaystyle \frac{s_R}{2}}\left(g^{\mu m}g^{\nu 0}g^{\mu 0}g^{\nu m}\right)\widehat{𝐪}^m`$ $`+{\displaystyle \frac{s_Rs_L}{2}}\left(2g^{\mu 0}g^{\nu 0}\delta ^{mn}g^{\mu \nu }\delta ^{mn}g^{\mu m}g^{\nu n}g^{\mu n}g^{\nu m}\right)\widehat{𝐪}^m\widehat{𝐤}^n,`$ where $`s_L=s_R=\pm 1`$ for the gap and $`s_L=s_R=\pm 1`$ for the anti-gap, see eq. (9) of . The Dirac structure of the Bethe-Salpeter equation (12) without the projectors $`\frac{1}{2}(1\pm 𝜶\widehat{𝐪})`$ in the generalized-pion-vertex (19) reads $`B^{\mu \nu }`$ $`=`$ $`\frac{1}{4}\mathrm{Tr}[\gamma ^\mu \frac{1}{2}(1\gamma ^0\gamma ^m\widehat{𝐪}^m)\gamma ^\nu ]`$ (A122) $`=`$ $`\frac{1}{2}g^{\mu \nu }\frac{1}{2}\left(g^{\mu 0}g^{m\nu }g^{\mu m}g^{\nu 0}\right)\widehat{𝐪}^m.`$ Here, the projector $`\frac{1}{2}(1\gamma ^0\gamma ^m\widehat{𝐪}^m)`$ results solely from the leading parts of the quark propagators (LABEL:PropPart) and not from the generalized-pion vertex. The prefactor $`\frac{1}{4}`$ in (A122) versus the prefactor $`\frac{1}{2}`$ in (A121) can be traced back to the division by the Dirac trace on the l.h.s. of the Bethe-Salpeter equation, namely to the division by $`\mathrm{Tr}[1]=4`$ in (A122) versus $`\mathrm{Tr}[\frac{1}{2}(1+s_L\gamma ^0\gamma ^n\widehat{𝐤}^n)]=2`$ in (A121). Using the projectors of $`𝒟_{\mu \nu }(kq)`$ as given in (A110), we get $`A^{\mu \nu }P_{\mu \nu }^L=A^{00}=\frac{1}{2}\left(1+s_Rs_L\widehat{𝐪}\widehat{𝐤}\right).`$ (A123) Assuming as in that $`\widehat{𝐪}\widehat{𝐤}\mathrm{cos}(\theta )1`$ in the numerators of the gap-equation, we get the weight $`1`$ for the longitudinal contribution to the gap and 0 for the longitudinal contribution to the anti-gap. This should be compared with $$B^{\mu \nu }P_{\mu \nu }^L=\frac{1}{2}$$ (A124) for both the gap and the anti-gap. Note that (A124) is just the average of (A123). Furthermore, we have using (A120) $`A^{\mu \nu }P_{\mu \nu }^T`$ $`=`$ $`\frac{1}{2}g^{ij}P_{ij}^T+{\displaystyle \frac{s_Rs_L}{2}}\left(2P_{mn}^T+\delta ^{ij}P_{ij}^T\delta ^{mn}\right)\widehat{𝐤}^n\widehat{𝐪}^m`$ (A125) $`=`$ $`1+s_Rs_L\left(\widehat{𝐤}\widehat{𝐪}\widehat{𝐤}\widehat{𝐪}+\widehat{𝐤}\widehat{𝐐}\widehat{𝐪}\widehat{𝐐}\right)`$ $``$ $`1\frac{1}{2}s_Rs_L+\frac{1}{2}s_Rs_L\widehat{𝐤}\widehat{𝐪}.`$ Under the approximation $`\widehat{𝐤}\widehat{𝐪}1`$, we get for the gap-case $`\frac{3}{2}+\frac{1}{2}\widehat{𝐤}\widehat{𝐪}1`$ and for the anti-gap case $`\frac{1}{2}\frac{1}{2}\widehat{𝐤}\widehat{𝐪}1`$. This agrees with the unprojected case $$B^{\mu \nu }P_{\mu \nu }^T=\frac{1}{2}\delta ^{ij}P_{ij}^T=1$$ (A126) for both the gap and the anti-gap. Finally, $`A^{\mu \nu }P_{\mu \nu }^{GF}`$ $`=`$ $`\frac{1}{2}{\displaystyle \frac{g^{\mu \nu }Q_\mu Q_\nu }{Q^2}}+{\displaystyle \frac{s_Rs_L}{2}}\left(2{\displaystyle \frac{Q^0Q^0}{Q^2}}\widehat{𝐤}\widehat{𝐪}2{\displaystyle \frac{𝐐\widehat{𝐤}𝐐\widehat{𝐪}}{Q^2}}{\displaystyle \frac{Q^2}{Q^2}}\widehat{𝐤}\widehat{𝐪}\right)`$ (A127) $`=`$ $`\frac{1}{2}{\displaystyle \frac{s_Rs_L}{2}}\widehat{𝐤}\widehat{𝐪}+s_Rs_L\left(\widehat{𝐤}\widehat{𝐪}{\displaystyle \frac{Q_{}^{0}{}_{}{}^{2}}{Q^2}}{\displaystyle \frac{𝐐\widehat{𝐤}𝐐\widehat{𝐪}}{Q^2}}\right)`$ $``$ $`\frac{1}{2}{\displaystyle \frac{s_Rs_L}{2}}\widehat{𝐤}\widehat{𝐪}+s_Rs_L\left({\displaystyle \frac{𝐐\widehat{𝐤}𝐐\widehat{𝐪}}{𝐐^2}}\right)`$ $``$ $`\frac{1}{2}{\displaystyle \frac{s_Rs_L}{2}}\widehat{𝐤}\widehat{𝐪}+s_Rs_L\left(\frac{1}{2}\widehat{𝐤}\widehat{𝐪}\frac{1}{2}\right)`$ $`=`$ $`\frac{1}{2}\left(1s_Rs_L\right),`$ where $`Q^2𝐐^2`$ and (A120) was used. Note that the gauge-fixing dependence vanishes for the gap and gives a weight factor $`+1`$ for the anti-gap. This should be compared with $$B^{\mu \nu }P_{\mu \nu }^{GF}=+\frac{1}{2}$$ (A128) for both the gap and the anti-gap. Again, the result (A128) of the unprojected case is the average of the sandwiched one, (A127). If $`A^{\mu \nu }`$ is contracted with $`\frac{1}{2}g_{\mu \nu }`$ as in the gluon-propagator (A116) and in the Euclidean analog (8), we get $`A^{\mu \nu }{\displaystyle \frac{1}{2}}g_{\mu \nu }`$ $`=`$ $`1+{\displaystyle \frac{s_Rs_L}{2}}\left(g^{00}\widehat{𝐪}\widehat{𝐤}+2\widehat{𝐪}\widehat{𝐤}+\frac{1}{2}g^{mn}\widehat{𝐪}^m\widehat{𝐤}^n+\frac{1}{2}g^{mn}\widehat{𝐪}^m\widehat{𝐤}^n\right)`$ (A129) $`=`$ $`1+{\displaystyle \frac{s_Rs_L}{2}}\left(\widehat{𝐪}\widehat{𝐤}+2\widehat{𝐪}\widehat{𝐤}\widehat{𝐪}\widehat{𝐤}\right)`$ $`=`$ $`1`$ for both the gap and the anti-gap. If $`B^{\mu \nu }`$ is contracted with $`\frac{1}{2}g_{\mu \nu }`$, then $$B^{\mu \nu }\frac{1}{2}g_{\mu \nu }=\frac{1}{4}g^{\mu \nu }g_{\mu \nu }=1$$ (A130) for both the gap and the anti-gap. In summary: the use of the simplified propagator (8) together with the unprojected meson vertices, i.e. (35) without the additional projectors, yields the same results as the use of the exact propagator (A110) (or even the simplified propagator) together with the sandwiched meson vertices (35), to leading order. At higher order, the simplifications require amendments. To leading order, the differences resulting from (A121) and (A122) can be traced back to the presence or absence of the projector $`\mathrm{\Lambda }^\pm (𝐤)`$ on the l.h.s. of the Bethe-Salpeter equation, i.e. at the amputated vertex.
warning/0001/quant-ph0001077.html
ar5iv
text
# 1 Introduction ## 1 Introduction Feynman \[Fy\] described how, in principle, a finite dimensional quantum mechanical system could be made to evolve so as to execute a sequence of $`\mathrm{`}\mathrm{`}`$quantum gates.” By a quantum gate, we mean one of the several rather standard unitary operators applied to a small (usually 1, 2, or 3) number of tensor factors (and the identity on the remaining factors) within a larger tensor product of factors of bounded dimension, often taken to be $`(^2)^n`$. For example: NOT, Controlled NOT, and Controlled $`i`$phase is known to be a complete basis \[Ki\]. Lloyd \[L\] made the converse explicit. Let $`H`$ be a poly-local Hamiltonian, $`H=\underset{\mathrm{}=1}{\overset{𝐿}{}}H_{\mathrm{}}`$ is a sum of $`L`$ polynomial(problem size) terms $`H_{\mathrm{}}`$, each the identity except for a constant number of tensor factors. The Hamiltonian $`H`$ exponentiates to $`𝐔=e^{2\pi itH}`$ which is well approximated by $`𝐔^{}=\underset{i=1}{\overset{𝑀}{}}(\text{gate})_i`$, where $`M\text{constant}t\frac{1}{ϵ}L^2`$ and $`ϵ>0`$ is the allowed (operator norm) discrepancy between $`𝐔`$ and $`𝐔^{}`$. The poly-local assumption on $`H`$ is very natural for most real world quantum mechanical systems and is often summarized by the axiom: $`\mathrm{`}\mathrm{`}`$physics is local.” For example, in quantum field theory interactions tensors with more than four indices are generally not renormalizable. In more exotic contexts, such as, fractional quantum Hall systems \[W\] topological models have been proposed for which the Hamiltonian acting on internal degrees of freedom vanishes identically $`H0`$, yet after a topological event, such as a braiding $`b`$ is completed, the internal state is transformed by $`𝐔(b)`$. For systems of this type (called: a unitary topological modular functor) a distinct argument was provided \[FKW\] that $`𝐔(b)`$ is poly-local. There seem to be few remaining candidates - perhaps quantum gravity - for a physical system having a finite dimensional state space whose evolution cannot be well approximated by a poly-local transformation $`𝐔`$. This reinforces confidence in the quantum circuit model QCM \[Ki\], and motivates the examination of (families of) transformations having a poly-local description. Any poly-local unitary transform is available as a $`\mathrm{`}\mathrm{`}`$subroutine” on a quantum computer, but it will not generally be a simple matter to recognize a transformation as lying in (or efficiently approximated by) that class. While a simple volume argument shows that only a tiny fraction of the unitary group $`𝐔(2^n)`$ can be approximated (up to phase) by poly-local transformations, it is not known how to give concrete examples which cannot be thus approximated. The situation is reminiscent of the fact that transcendental numbers are easily seen to predominate, but less easy to identify. To add to the puzzle, in most contexts, we will be free to add ancilli to create an inclusion $`vv0\mathrm{}0>`$, $`HH^+`$ of one Hilbert space into a larger one. It seems likely that there may be an operator $`T:HH`$ which is not well approximated by poly-local operators but its extension $`T\text{id}:H^+H^+`$ is. In this regard, Michael Nielsen has pointed out (private communication and Chpt. 6 \[CN\]) that the exponential of any Hamiltonian which is a sum of polynomially many product terms can, with the help of ancilli, be efficiently simulated on a quantum computer. There is no requirement on the individual product terms except that they be products, i.e. a tensors of operators each acting on one (or up to log many) qubit(s). To appreciate the importance of a poly-local description consider the FFT. The reader familiar with the Fast Fourier Transform may be surprised that it has a virtue in the quantum context since it is already fast, $`𝒪(N\mathrm{log}N)`$, classically. But on this scale, we should think of the quantum version as poly$`(\mathrm{log}N)`$ in the sense that we may Fourier transform in poly$`(\mathrm{log}N)`$ time a data vector $`(f(1>),\mathrm{},f(N>))`$ (held in superposition and previously computed in poly$`(\mathrm{log}N)`$ time) into its discrete Fourier transformation $`(\widehat{f}(\widehat{1}>),\mathrm{},\widehat{f}(\widehat{N}>))`$. The measurement phase of quantum computation samples the transform according to the density of its $`L^2`$norm squared. In a sense, this is the best we could hope for in poly-log time. We could never output the entire transform in less than $`𝒪(N)`$ time, instead sampling shows us its sharpest spikes. Similarly with other transforms, a poly-local description allows us to sample, according to the $`L^2`$ norm, that transform applied to any rapidly computable function. A key step in the Shor factoring algorithm is his description of the discrete Fourier transform $`F:C_m\widehat{C}_m`$ for the cyclic group of order $`m`$, where $`m`$ is a large smooth number, as a poly-local transform (The size parameter is $`\mathrm{log}m`$ so $`F`$ must be written as a composition of poly$`(\mathrm{log}m)`$ many gates). We construct another class of examples by showing that cyclic periodic near diagonal matrices are poly-local (Thm 1). The construction is direct - no ancilla qubits are used. As an application, we prove (Corollary 1) that all the standard orthogonal discrete wavelet transforms, in particular the Daubechies transforms $`D_{2n}`$ \[D\], are poly-local. I thank C. Williams for pointing out that he and A. Fijany previously obtained an explicit factoring of $`D_2`$( = Haar) and $`D_4`$ into quantum circuits \[AW\]. Our alternative approach has the usual virtues and demerits of abstraction. Some perspective into the limitations poly-locality imposes on the poly-time computational class of the QCM, BQP, can be achieved by adjoining oracles. Let $`𝒪_n:(^2)^n(^2)^n`$ be a $`\mathrm{`}\mathrm{`}`$super-gate” which can be called upon in addition to gates to transform the computational state $`\mathrm{\Psi }`$. It is clear that if $`𝒪_n`$ encodes the answer to a difficult (or even undecidable) problem $`P_n`$, $`\text{QCM}^𝒪`$ will be improbable strong: If $`f_n:\{0,1\}^{n1}\{0,1\}`$ is a function and $`(f_n):\{0,1\}^n\{0,1\}^n`$ is defined by $`f_n(x_1,\mathrm{},x_{n1},x_n)=\{x_1,\mathrm{},x_{n1},f_n(x_1,\mathrm{},x_{n1})+x_n\}`$ and $`𝒪_n`$ is the linear extension of $`f_n`$ operating on $`(^2)^n`$, then we can compute $`f_n(x_1,\mathrm{},x_{n1})`$ by measuring the last qubit of $`𝒪_n(x_1,\mathrm{},x_{n1},0)`$. However, if we stipulate that $`𝒪_n`$ is $`\mathrm{`}\mathrm{`}`$locally-poly” meaning that the $`(i,j)`$ entry of $`𝒪_n`$ be computable in poly$`(n)`$ time, can the nonlocal structure of such $`𝒪_n`$ by virtue of their size alone, extend QCM? The following observation is evidence that the answer is $`\mathrm{`}\mathrm{`}`$yes.” ###### Observation 1. For each family of bijections $`f_n:\{0,1\}^n\{0,1\}^n`$ which is computable in polynomial(n) time there is a family of oracles $`𝒪_n`$ with entries $`(𝒪_n)_{i,j}`$ computable in poly$`(n)`$time so that $`f^1`$ can be computed by $`\text{BQP}^{𝒪_n}`$. Simply set $`(𝒪_n)_{i,j}=\delta _{f(i),j}`$. Write the bit string $`(y_1^j,\mathrm{},y_n^j)=\stackrel{}{y}=y^j,1j2^n`$. Then $`𝒪_n\stackrel{}{y}=\underset{j=1}{\overset{2^n}{\mathrm{\Sigma }}}\delta _{f(i),j}y^i=y^{f^1(j)}=f^1(\stackrel{}{y})`$, so applying $`𝒪_n`$ to $`(y_1,\mathrm{},y_n)`$ and then measuring each qubit separately solves the inversion problem $`𝒪_n(y_1^j,\mathrm{},y_n^j)=(y_1^{f^1(j)},\mathrm{},y_n^{f^1(j)})`$. Thus, one-way functions disappear if we are permitted to adjoin locally-poly yet non-poly-local oracles. We have not succeeded in solving an $`NP`$ complete problem with such an oracle. ## 2 Periodic Band-Diagonal Transforms It is common in signal processing to encounter smoothing operators which when written as matrices are band diagonal. Often mathematical reasons suggest wrapping the interval into a circle and considering cyclically banded matrices $`M`$, $`M_{i,j}=0`$ unless $`ij<bmodN,1i,jN`$. Two additional conditions natural in signal processing are unitary, $`(M^1=M^{})`$ and cyclicity $`(k<<N\text{s.t.}M_{i,j}=M_{i+k,j+k}`$, addition of indices taken $`(\text{mod}N))`$. All these conditions occur in quadrature mirror filters \[P\]. We fix $`b`$ and $`k`$ and considering $`N`$ as parameter approaching infinity for studying the family of $`N\times N`$ matrices $`M_N`$. To summarize the $`M_N`$ are $`N\times N`$ unitary matrices, $`b`$band diagonal in the cyclic sense, cyclic with period $`k`$, $`k|N`$, and $`4b<N`$. Furthermore even among different $`M_N`$, $`M_N^{}`$, we assume $`(M_N)_{i,j}=(M_N^{})_{i^{},j^{}}`$. If $`ii^{}`$ and $`jj^{}modk`$, so the matrices look locally identical. Purely for convenience, we also assume all $`N`$ are powers of $`2`$, $`N=2^n`$, $`n>n_{}`$ some fixed constant and $`k=2^{\mathrm{}}`$, this makes the fit with qubits effortless. A final technical assumption is necessary to efficiently approximate even a small $`M_N_{}`$, $`N_{}=2^n_{}`$, from gates available in the QCM. It is sufficient here \[Ki\] to require the entries $`M_{i,j}`$ are algebraic numbers. This prevents some difficult to obtain information being encoded in the entries themselves. The theorem below asserts that the family $`\{M_N\}`$ is poly-local in the strictest sense: No ancilla are introduced to stabilize the problem. ###### Theorem 2.1. Let $`\{M_N\}`$ be as above, there is a polynomial $`p`$ so that for any $`ϵ>0`$ there is an $`M_N^{}`$, $`ϵ`$ close to $`M_N`$ in the operator norm, so that $`M_N^{}`$ is a composition of $`p(n,\mathrm{log}\left(\frac{1}{ϵ}\right))`$ gates in the QCM. ###### Proof 2.2. It is sufficient to consider $`N=2^n`$, for $`n`$ greater than any fixed number, so the reader may image $`M_N`$, $`N`$ comfortably large. For some $`K=2^jk`$, some $`j>0`$, we will modify $`M_N`$ to $`\overline{M}_N`$ to a unitary $`K`$cyclic matrix which agrees with $`M_N`$ except $`K`$periodically for a small (order $`b`$) slab of rows where $`\overline{M}_N`$ agrees with the identity matrix and a small (order $`b`$) transition region between the two types of rows. The construction of $`\overline{M}_N`$ is very similar to the truncation process given in \[F,P\] so we will not give formulas here, but simple principles from which $`\overline{M}_N`$ may be constructed. Let $`V_I`$ be the vector subspace spaned by the rows of $`M_N`$ whose row index lies in $`I`$ (We think of $`I`$ as an interval with integer end points on the circle consisting of $`R/NZ`$). Let $`W_I`$ be the subspace spanned by the basis vectors with indexes in $`I`$equivalently those rows of the identity $`N\times N`$matrix. From the nonsingularity of unitary matrices we obtain: ###### Lemma 2.3. Whenever $`J`$ contains the $`b`$neighborhood of $`I`$, $`W_IV_J`$ and $`V_IW_J.\mathrm{}`$ We apply the lemma several times below: consider an interval $`I`$ of $`\mathrm{}>2b`$ rows and a larger interval $`J`$ of $`\mathrm{}\text{rows}+2b`$ rows containing $`I`$ symmetrically. Let $`J^+`$ and $`J^{}`$ denote the intervals of length$`=2b`$ at containing the $`+`$ and $``$ end points of $`J`$ respectively. The rows $`\{q\}`$ of $`M_N`$ in the complement of $`I`$ together with the rows $`\{s\}`$ of the identity matrix in the interval $`J`$ span the entire space. Let $`\{r\}`$ denote the rows of $`M_N`$ with indices in $`J^+`$ and $`J^{}`$. The Gram-Schmidt process applied to $`\{r\}\{s\}`$ (The Gram-Schmidt process requires an ordering; work from the middle of the $`J`$ interval outward) produces a Hermitian- orthogonal frame $`\{t\}`$ for the space spanned by $`\{r\}\{s\}`$ (Note: $`\{q\}\{r\}\{s\}`$ span the entire space). This process naturally leaves the central $`\mathrm{}2b`$ rows of $`\{s\}`$ unchanged, i.e. they remain standard basis vectors. The geometry of this process is pictured below. Figure 2 The frame $`\{t\}`$ consists of $`\mathrm{}`$ standard basis vectors together with $`2b`$ vectors of length $`2b`$. Think of these $`2b`$vector as mediating a transition from rows $`\{q\}`$ of $`M_N`$ to rows $`\{s\}`$ of the identity matrix. The truncation process preserves unitarity and band diagonality and the band width. $`\overline{M}_N`$ is block diagonal consisting of $`2^{nj\mathrm{}}`$ identical blocks $`\overline{M}`$ (independent of N) down the diagonal. Thus, $`\overline{M}_N`$ has a tensor decomposition $`\overline{M}_N=\overline{M}\text{id}_{(^2)^{nj\mathrm{}}}`$. Define $`\overline{\overline{M}}=(\overline{M}_N)^1M_N`$. Clearly the rows of $`\overline{\overline{M}}_N`$ agree with the identity matrix outside the $`J`$intervals and $`\overline{\overline{M}}_N`$ is $`b`$band diagonal. Thus, $`\overline{\overline{M}}_N`$ can be written in a tensor product form as well: $`\overline{\overline{M}}_N`$ consists of identical blocks down the diagonal of size $`K`$ provided $`K2b`$. Thus, we may write $`\overline{\overline{M}}_N=P(\overline{\overline{M}}\text{id}_{(^2)^{nj\mathrm{}}})P^1`$ where $`P`$ is a cyclic permutation matrix of $`\{1,2,\mathrm{},N\}`$. The permutation must arise since the blocks of $`\overline{\overline{M}}`$ are off-set from the blocks $`\overline{M}_N`$ by a translation of $`\frac{K}{2}`$. Because we assumed the entries of $`M_N`$ to be algebraic so are the entries of $`\overline{M}`$ and $`\overline{\overline{M}}`$. It is well-known that algebraic numbers can be efficiently computed using Newton’s method so that refining the numerical accuracy of an algebraic $`\alpha `$ to up to a factor of $`(1+ϵ)`$ requires only constant$`(\alpha )`$ poly $`\mathrm{log}\left(\frac{1}{ϵ}\right)`$ computational steps. Now by a theorem of \[Ki\] for approximating a unitary transformation in a fixed dimension as a product of gates, we find matrices $`\overline{M}^{}`$ and $`\overline{\overline{M}}^{}`$, $`|\overline{M}\overline{M}^{}|<\frac{ϵ}{2}`$, and $`|\overline{\overline{M}}\overline{\overline{M}}^{}|<\frac{ϵ}{2}`$ using the operator norms, with $`\overline{M}^{}`$ and $`\overline{\overline{M}}^{}`$ a product of poly $`\mathrm{log}\left(\frac{1}{ϵ}\right)`$ elementary gates. The cyclic permutation $`P`$ and its inverse $`P^1`$ are, of course, poly$`(n)=\text{poly}(\mathrm{log}N)`$ time computable classically. Writing $`|x>\{1,2,\mathrm{},N\}`$ in an initial register and $`|0>`$ in an equal large (size $`=n`$, $`N=2^n`$) register of ancilli a well-known sequence of operations effects the transformation: $$\begin{array}{c}P_i(|x>|0>)\stackrel{}{\text{compute}P}\left(|x>|\mathrm{\hspace{0.17em}0}+P(x)>\right)\stackrel{}{\text{flip}}\left(|P(x)>|x>\right)\\ \stackrel{}{\text{compute}P^1}\left(|P(x)>|x+P^1P(x)>\right)\\ =\left(|P(x)>|0>\right)\end{array}$$ (where $`+`$ denotes bit wise mod $`2`$ sum in the register). More directly, it is possible to build size $`𝒪(n^2)`$ circuits from $`\mathrm{`}\mathrm{`}`$NOT, $`\mathrm{`}\mathrm{`}`$Controlled NOT,” and $`\mathrm{`}\mathrm{`}`$swap” gates (See Section 2.4 \[AW\].), which use no ancilli and implement addition of $`\frac{K}{2}`$ modulo $`N`$. For example, a circuit implementing $`+2mod(8)`$ is shown below. Figure 3 Thus, all the 4 terms of an $`ϵ`$approximation $`M_N^{}`$ may be written as a product of poly $`(n,\mathrm{log}\frac{1}{ϵ})`$ many gates: $`M_N\stackrel{\mathit{ϵ}}{}M_N^{}=(\overline{M}^{}\text{id})P(\overline{\overline{M}^{}}\text{id})P^1`$. $`\mathrm{}`$ ## 3 Wavelets We define a discrete wavelet transform as the result of the pyramid algorithm \[M\] - applied to a band diagonal quadrature mirror filter. Daubachies’ \[D\] original basis is recovered by choosing a quadrature mirror filter whose differencing rows have a maximal number of vanishing moments (consistent with orthogonality relations). For example, the simplest Daubachies wavelet Daub<sub>4</sub> is built from an orthogonal matrix(and mirror filter): $$M_N=\stackrel{2^n=N}{\stackrel{}{\left|\begin{array}{cccccccc}C_1& C_2& C_3& & & & & C_0\\ C_2& C_1& C_0& & & & & C_3\\ & C_0& C_1& C_2& C_3\\ & C_3& C_2& C_1& C_0\\ \mathrm{}& & \mathrm{}& & & & \mathrm{}\\ & & & C_0& C_1& C_2& C_3\\ & & & C_3& C_2& C_1& C_0\\ C_3& & & & & C_0& C_1& C_2\\ C_0& & & & & C_3& C_2& C_1\end{array}\right|}}\}2^n=N$$ satisfying moment conditions: $$C_3C_2+C_1C_0=0\text{and}\mathrm{\hspace{0.33em}\hspace{0.33em}0}C_31C_2+2C_13C_0=0.$$ First let us explain in algorithmic terms how the wavelet coefficients of an data (column) vector $`X=(x_1,\mathrm{},x_N)`$ are computed starting from a filter $`M_N`$ as above (compare \[P\]). The even components of $`M_NX`$ are regarded as $`\mathrm{`}\mathrm{`}`$smoothed” and these are set aside for further processing. The odd components of $`M_NX`$ are the finest scale wavelet coefficients. To continue processing apply the matrix identical in description to $`M_N`$, but of half the size to the even output of $`M_NX:M_{N/2}\left(\text{even}(M_NX)\right)`$. Again set the even output aside: the odd output are the wavelet coefficients at the next-to-smallest scale. Now apply $`M_{N/4}\left(\text{even}\left(M_{N/2}(\text{even}M_NX)\right)\right)`$, etc… The process must terminate when $`M`$ becomes so small that wrapping around of rows threatens unitarity. For Daub<sub>4</sub> the last $`M`$ will be a $`4\times 4`$ matrix. The wavelet coefficients of $`X`$, usually written as $`X`$ integrated against a function $`\psi `$, are the output of the final differencing (proceeded, possible, by several smoothings). The $`\mathrm{`}\mathrm{`}`$mother” or $`\mathrm{`}\mathrm{`}`$scaling” coefficients will be the output of successive smoothings (i.e. even rows). These are traditionally written as $`X`$ integrated against a function $`\varphi `$; there will be two of these for Daub<sub>4</sub>. The pyramid algorithm consist of only $`\mathrm{log}(N)`$ many applications of a filter matrix $`M`$, obeying the hypotheses of the theorem, and some easily computed permutations needed to sort out the even from odd rows between applications of $`M`$. Thus from the theorem, we have an evident corollary, which should be compared with the constructions of \[AW\] and \[Kl\]. ###### Corollary 1. Consider a poly-time computable function of $`f`$ from the integers $`[1,\mathrm{},N]`$ into a fixed basis of a vector space $`V`$ of dimension $`=d`$. The function $`f`$ is recorded by a vector $`v_fϵV^N`$. For any discrete (algebraic) wavelet transform $`T`$ applied to $`f`$, there is a quantum circuit, without ancilla, of length poly$`(\mathrm{log}N,\mathrm{log}\frac{1}{ϵ})`$ which yields a state space vector $`T^{}(v_f)ϵV^N`$ which is within $`ϵ>0`$ of the exact transform $`T(v_f)`$. The components of $`T^{}(V_f)`$ can be sampled by von Neumann measurement according to the density $`|\text{component}_{i,j}|^2`$, $`1iN`$, $`1jd`$. The proceeding recipe, in so far as repeated smoothing is applied, sends a $`\mathrm{`}\mathrm{`}`$Dirac function” $`\delta `$ toward a fixed point of the iteration $`\varphi (x)\mathrm{\Sigma }c_n\varphi (2xn)`$. Such a fixed function $`\varphi `$ is called the scaling function of the filter $`M`$. Wavelet coefficients (in the limit) correspond to integrations against $`\psi (2^kxn)`$ where $`\psi (x)=\mathrm{\Sigma }(1)^kc_{1k}\varphi (2xk)`$. Whereas, the Fourier transform efficiently extracts information about periodicity (hence its relevance to divisibility and factoring questions) the wavelet transforms (discrete or continuous) extract qualitatively different information. It is less easy to say precisely what combinatorial structures wavelets are best suited to detecting, but the answer must relate to phenomenon with a certain amount of both spacial and frequency localization. Some studies suggest fractal behaviors, Hausdorff dimension, and the like are readily observed in wavelet bases. Thinking democratically, one should not prefer either Fourier or wavelet bases, but seek different problems which can be solved by sampling the density of one transform or the other. So far on the Fourier side there is discrete log, factoring, and the abelian stabilizer problem \[S\]\[Ki\] - essentially similar applications of that transform, and no interesting quantum applications, yet known, for the wavelet transforms. ## 4 Continuum Versus Vectorized Transforms Both Shors’ \[S\] use of the Fourier transform and our treatment of wavelet transforms concerns the $`\mathrm{`}\mathrm{`}`$vectorized” transform: The function values of $`x^amod(n)`$ in Shor’s work and the values of $`f`$ mod$`(d)`$ in our corollary 1 are treated as independent basis vectors in an abstract vector space. The transforms do not treat similar function values similarly, but rather place each possible outcome of $`f(t)`$ into a separate bin. Thus, there is no notion of the continuum limit of these transforms as presently constituted and they should not be confused with their continuous relatives in signal processing. The property of a function $`f`$ on the integers having period $`=d`$ is invariant under arbitrary permutation (or vectorization) of the target. This fact is crucial to the efficacy of the Fourier transform in Shor’s algorithm. It is more difficult to see what portion of the information captured by a wavelet transform is invariant under target permutation or vectorization. Although it seems that a continuous - as opposed to vectorized - Fourier transform is not as useful for factoring, their may be other quantum applications for the continuous version. In the case of wavelets, it is more likely that continuous rather than vectorized transforms will be important. In the following paragraph, we explain how to use phase coding of states to build efficient continuous versions of any transform $`T`$ (such as Fourier or wavelet) for which there is an efficient quantum circuit implementation of the vectorized version. Let $`F:[1,\mathrm{},N][1,\mathrm{},d]`$ be a funtion and denote $`t[1,\mathrm{},N]`$ as the variable. For $`1xd`$, we may define $`f_x(t)=0`$ if $`f(t)x`$, and $`f_x(t)=1`$ if $`f(t)=x;f(t)=\underset{x=1}{\overset{𝑑}{\mathrm{\Sigma }}}f_x(t)`$. A $`\mathrm{`}\mathrm{`}`$vectorized” unitary transform obeys: $$T\left(\underset{t=1}{\overset{𝑁}{\mathrm{\Sigma }}}|t>|f(t)>\right)=\underset{x=1}{\overset{𝑑}{\mathrm{\Sigma }}}\underset{\widehat{t}=1}{\overset{𝑁}{\mathrm{\Sigma }}}\left(T(f_x)\widehat{t}\right)|\widehat{t}>|x>$$ where $`1\widehat{t}N`$ is the transform variable, and states are written without numerical normalization factors. If we pick $`D>>d`$ and let $`w=e^{2\pi i/D}`$, the set of complex numbers $`S=\{w^i|1id\}`$ is a short discretized arc of a circle and is a rather good affine model for $`\mathrm{}=[1,\mathrm{},d]`$ under the bijection $`\delta \stackrel{𝑏}{}e^{2\pi i\delta /D}`$. After linear rescaling the quasi-isometric distortion of b is $`𝒪\left(\left(\frac{d}{D}\right)^2\right)`$ (A map $`g:XY`$ between metric spaces is said to have quasi-isometric distortion $`K`$ for some $`K0`$ iff: $$\frac{1}{1+K}d(g(x_1),g(x_2))d(x_1,x_2)(1+K)d\left(g(x_1)g(x_2)\right).$$ We will neglect this distortion and code the values of a bounded real function $`h:[1,\mathrm{},N][0,\frac{2\pi d}{D}]`$ into $`\stackrel{~}{h}:[1,\mathrm{},N]`$, $`\stackrel{~}{h}=b^1r\text{exp}(ih)`$ where $`r`$ rounds to the nearest element of the form $`e^{2\pi i\delta /D}`$. Now a continuous version $`\stackrel{~}{T}`$ of $`T`$ applied to $`\stackrel{~}{h}`$ is given by the formula: $$\stackrel{~}{T}\left(\underset{t=1}{\overset{𝑁}{\mathrm{\Sigma }}}|t>|\stackrel{~}{h}(t)>\right)=\underset{x=1}{\overset{𝑑}{\mathrm{\Sigma }}}\underset{t=1}{\overset{𝑁}{\mathrm{\Sigma }}}w^x\left(\stackrel{~}{T}(f_x)\widehat{t}\right)|\widehat{t}>.$$ Note that there is a simple modification of a quantum circuit computing $`T`$ to one computing $`\stackrel{~}{T}`$. Instead of simply recording the qubit $`|x>`$ by adding $`x`$ to a $`D`$state register $`R`$ holding zero, the trick is to initialize $`R`$ to the superposition: $$w^0|1>+w^1|2>+\mathrm{}w^{D1}|D>.$$ Now adding $`x`$ to the above state simply changes its phase by a factor of $`w^x`$. Thus, the transform $`\stackrel{~}{T}`$, possessing a natural continuum limit, can be sampled in the QCM with no more difficulty than the discrete version $`T`$. If quantum computers are built, such continuous transforms would find application in signal processing. Beyond these obvious applications, it is interesting to wonder what algorithmic potential lies in these continuum transformations.
warning/0001/math0001045.html
ar5iv
text
# Derived categories for the working mathematician ## 1. Introduction I should begin by apologising for the title of this talk \[Ma\]; while the title is intended for mathematicians the talk itself is aimed at least as much at physicists. Kontsevich’s mirror symmetry programme \[K\] and the mathematical description of D-branes have brought triangulated categories into mainstream string theory and geometry (there are now even papers in which the reference \[Ha\] means \[RD\] rather than \[AG\]). Since they have such a fearsome reputation (probably mainly due to the references being in French), and since I will need them for my second talk, as presumably will other speakers, I wanted to motivate this beautiful piece of homological algebra. This motivation is of course what guided the creators of derived categories (principally Verdier, or as he is traditionally known in this context, Grothendieck’s student Verdier) but, equally of course, is not written down. For the technical details of the theory the working mathematician should consult the excellent \[GM\]. The main idea of derived categories is simple: work with complexes rather than their (co)homology. We will take simple examples from algebraic geometry to demonstrate why one might want to do this, then examples from algebraic topology to show that the ideas and structure are already familiar. (The link between the two subjects is sheaves, but we shall not pursue this.) This also gives us a way of using pictures and techniques from topology in algebraic geometry. Section 1 begins with chain complexes in algebraic geometry and topology, and why they are preferable to (co)homology. Thus we would like to consider the natural invariant of a space/sheaf/etc. to be a complex rather than its homology, so to do this we must identify which complexes should be considered isomorphic – this leads to the notion of quasi-isomorphism that is the subject of Section 2. Setting up the general (abstract) theory is where categories creep in, in Section 3, leading to cones, triangles, and triangulated categories in Section 4. Finally Section 5 shows how more classical homological algebra, in particular derived functors, fit into this framework; this makes them more transparent and leads to easier proofs and explanations of standard results. Acknowledgements. I learnt a lot of homological algebra from Brian Conrad, Mikhail Khovanov and Paul Seidel. I would like to thank members of the CALF (junior Cambridge-Oxford-Warwick) algebraic geometry seminar for enduring a trial run of a preliminary version of this talk, and the organisers of the 1999 Harvard Winter School on Mirror Symmetry for inviting me to take part. ## 2. Chain complexes ### Chain complexes in topology Chain complexes are familiar in algebraic topology. A simplicial complex $`X`$ (such as a triangulation of a manifold) gives rise to two chain complexes. Letting $`C_i`$ be the free abelian group generated by the $`i`$-simplices in the space, and $`C^i`$ its dual, (2.1) $$C_i=\{i\text{-simplices}\},C^i=\mathrm{Hom}(C_i,),$$ we have the complexes $$C_{}=\mathrm{}C_i\stackrel{_i}{}C_{i1}\mathrm{},C^{}=\mathrm{}C^{i1}\stackrel{d^i}{}C^i\mathrm{}.$$ Here $``$ is the boundary operator taking a simplex to its boundary, $`d`$ is its adjoint, and $`^2=0=d^2`$. Thus we can form the homology and cohomology groups $$H_i(X)=\frac{\mathrm{ker}_i}{\mathrm{im}_{i+1}},H^i(X)=\frac{\mathrm{ker}d^{i+1}}{\mathrm{im}d^i}.$$ Simplicial maps $`XY`$ induce chain maps (that is, they commute with the (co)boundary operators) on the corresponding chain complexes, inducing maps on the (co)homology groups. However, the (co)homology $`H_{}(H^{})`$ contains less information than the complexes $`C_{},C^{}`$, for instance Massey products. There exist spaces with the same homology $`H_{}`$ but different homotopy type, making $`H_{}`$ a limited invariant of homotopy type. But $`C_{}`$ is a very powerful “invariant”, at least for simply connected spaces (to which we shall confine ourselves in this talk), due to the Whitehead theorem. This states that the underlying topological spaces $`|X|,|Y|`$ of simplicial complexes $`X`$ and $`Y`$ (both simply connected) are homotopy equivalent if and only if there are maps of simplicial complexes (2.2) *inducing isomorphisms on homology* $`H_{}(X)\stackrel{}{}H_{}(Z)\stackrel{}{}H_{}(Y)`$. (The reason for the appearance of $`Z`$ is that we may need to subdivide $`X`$ – thus giving a $`Z`$ with a simplicial approximation $`ZX`$ to the identity map $`|Z|\stackrel{}{}|X|`$ – before we can get a simplicial approximation $`ZY`$ to a given homotopy equivalence $`|X||Y|`$.) Thus $`C_{}`$ contains as much information as we could hope for, and we would like to think of it as an invariant of the topological space $`|X|`$. We shall see how to do this later; for now we can think of it as an invariant of the triangulation (simplicial complex). Another place in topology where the advantage of using complexes rather than their homology is already familiar is the construction of the dual, cohomology, theory as above. Namely we *do not* define $`H^{}`$ to be the dual of $`H_{}`$ in the sense that $$H^{}\mathrm{Hom}(H_{},).$$ This is because $`\mathrm{Hom}(.,)`$ destroys torsion information and its square is not the identity (the double dual in this sense is not the original $`H_{}`$). Instead we take $`\mathrm{Hom}(.,)`$ at the level of chain complexes as in (2.1). Then no information is lost, and the double dual is the original complex. For instance applying $`\mathrm{Hom}(.,)`$ to the complex $$\stackrel{2}{}H_1=0,H_0=/2,$$ gives the dual complex $$\stackrel{2}{}H^1=/2,H^0=0.$$ Thus the torsion $`/2`$ has not been lost on dualising (as it would have been using $`\mathrm{Hom}(H_{},)`$), it has just moved in degree. So we have seen two examples of what will be the theme of this talk, namely ### Chain complexes in algebraic geometry It might seem strange to consider chain complexes in algebraic geometry at all. The main place they arise is in resolutions of sheaves, which we describe now. Coherent sheaves often arise naturally as kernels or cokernels of maps, i.e. as the cohomology of the 2-term complex made from the map. For instance if $`DX`$ is a divisor in a smooth algebraic variety $`X`$, it corresponds to a line bundle $`L`$, and a section $`sH^0(L)`$ vanishing on $`D`$ (we shall confuse $`L`$ and its sheaf of sections $`L`$). This gives us the standard exact sequence (2.3) $$0L^1\stackrel{s}{}𝒪_X𝒪_D0,$$ where $`𝒪_X`$ is the structure sheaf of $`X`$ (sections of the trivial line bundle on $`X`$) and $`𝒪_D`$ is the structure sheaf of $`D`$ pushed forward to $`X`$ (extending by zero – this is a torsion sheaf concentrated on $`D`$). Thus $`𝒪_D`$ is the cohomology of the complex $`\{L^1\stackrel{s}{}𝒪_X\}`$. (Notice this is the cohomology of a complex of sheaves, and as such is a sheaf, and should not be confused with sheaf cohomology, which is vector-space-valued.) Similarly if $`ZX`$ is a codimension $`r`$ subvariety, the transverse zero locus of a regular section $`sH^0(E)`$ of a rank $`r`$ vector bundle $`E`$, the exact sequence (Koszul complex) $$0\mathrm{\Lambda }^rE^{}\mathrm{\Lambda }^{r1}E^{}\mathrm{}E^{}𝒪_X,$$ where each arrow is given by interior product with $`s`$, has cokernel $`𝒪_Z`$ by inspection. Thus $`𝒪_Z`$ is the cohomology of this complex. We have actually gained something here: we have replaced nasty torsion sheaves $`𝒪_D,𝒪_Z`$ by nicer, locally free sheaves (i.e. vector bundles) on $`X`$. In general one can consider such *resolutions*, replacing arbitrary sheaves $``$ by complexes $`F^{}`$ of sheaves that are “nicer” in some way, $$F^{}\mathrm{or}F^{},$$ and now use the complex $`F^{}`$ instead of its less manageable cohomology $``$. We think of the nasty looking curly $``$ being made up from the nicer straight $`F`$ s as $`F^1F^2+F^3\mathrm{}`$, where the sense in which we subtract the $`F^i`$s is given by the maps between them in the resolution, and the $`F^i`$ are the building blocks: the generators of $``$ form $`F^1`$, the relations $`F^2`$, relations amongst the relations $`F^3`$, and so on. Of course the resolution may not arise naturally in general and we must pick one; this is directly analogous to picking a (non-canonical) triangulation of a topological space as we shall see in Section 3, and is the problem that derived categories resolve. The sense in which the sheaves $`F^{}`$ are “nicer” will be tackled in general in Section 6, but the following is a simple example. ### Intersection theory via sheaves We concentrate on the easiest case of intersecting two divisors $`D_1,D_2`$ in a smooth complex surface $`X`$. These correspond to line bundles $`L_i`$ and sections $`s_iH^0(L_i)`$ with zero locus $`D_i`$. The $`D_i`$ have homology classes in $`H_2(X)`$ that we may intersect, or dually we can consider $`c_1(L_i)H^2(X)`$. In terms of sheaf theory the first corresponds to tensoring structure sheaves: if $`D_1,D_2`$ intersect transversely then we have $`𝒪_{D_1D_2}=𝒪_{D_1}𝒪_{D_2}`$. The second corresponds to using the resolution (2.3) for $`𝒪_{D_1}`$ and tensoring *that* with $`𝒪_{D_2}`$, as the following table shows: $$\begin{array}{ccc}& & \\ \mathrm{Data}& \mathrm{Divisor}D_i& s_iH^0(L_i)\mathrm{such}\mathrm{that}s_i^1(0)=D_i\hfill \\ & & \\ \mathrm{Alg}\mathrm{geom}& 𝒪_{D_i}& \{L_i^1\stackrel{s_i}{}𝒪_X\}\hfill \\ & & \\ \mathrm{Topology}& [D_i]H_2(X)& c_1(L_i)H^2(X)\hfill \\ & & \\ \mathrm{Intersection}& [D_1].[D_2]& c_1(L_1)c_1(L_2)=c_1(L_1),D_2=c_1(L_1|_{D_2})\hfill \\ & & \\ & 𝒪_{D_1}𝒪_{D_2}& c_1(L_1|_{D_2})=(s_1|_{D_2})^1(0),\mathrm{i}.\mathrm{e}.\mathrm{restrict}L_1\hfill \\ \mathrm{Transverse}\mathrm{case}& =𝒪_{D_1D_2}& \mathrm{to}D_2\mathrm{and}\mathrm{take}\mathrm{zeros}\mathrm{of}\mathrm{its}\mathrm{section}s_1|_{D_2},\hfill \\ & & \mathrm{i}.\mathrm{e}.\mathrm{take}\mathrm{cokernel}\mathrm{of}\{L_1^1|_{D_2}\stackrel{s_1|_{D_2}}{}𝒪_{D_2}\}\hfill \\ & & \\ \mathrm{Non}\mathrm{transverse}& 𝒪_D𝒪_D& \{L^1|_D\stackrel{s|_D=0}{}𝒪_D\}.\mathrm{So}\mathrm{we}\mathrm{still}\mathrm{see}\hfill \\ \mathrm{case};\mathrm{e}.\mathrm{g}.& =𝒪_D& L\mathrm{on}D,\mathrm{just}\mathrm{with}\mathrm{the}\mathrm{section}=0.\hfill \\ D_1=D_2=D& & \end{array}$$ So restricting $`L_1`$ to $`D_2`$ corresponds to tensoring $`\{L_1^1𝒪_X\}`$ with $`𝒪_{D_2}`$, and in the transverse case the cokernel of this is just $`𝒪_{D_1D_2}`$. Thus the way to pass from the right hand column to the left is to take cokernels; this corresponds to the fact that tensoring with $`𝒪_{D_2}`$ is *right exact*: tensoring the exact sequence (2.3) with $`𝒪_{D_2}`$ gives a sequence in which the final three arrows are still exact. What the table shows is that while tensoring $`𝒪_{D_1}`$ with $`𝒪_{D_2}`$ gives the correct answer for $`D_1`$ and $`D_2`$ transverse, in the non-transverse case it does not ($`𝒪_D`$ is the structure sheaf of $`DD`$, but not of the correct topological intersection of $`D`$ with itself). However tensoring instead $`\{L_1^1𝒪_X\}`$ with $`𝒪_{D_2}`$ always gives the right answer: taking the cokernel of the resulting complex gives the same $`𝒪_{D_1}𝒪_{D_2}`$ as before, but in the non-transverse case we get more – we still have the line bundle $`L_1|_{D_2}`$ and so the intersection information (we need only take its first Chern class); by changing the section $`L_1^1|_{D_2}\stackrel{s_1|_{D_2}}{}𝒪_{D_2}`$ from one which may be identically zero on (components of) $`D_2`$ to a transverse section we get the correct intersection. More generally we take the divisor of $`L_1|_{D_2}`$ on $`D_2`$. If $`D_1`$ can be moved to be transverse to $`D_2`$ this gives the same answer; if $`L_1|_{D_2}`$ has no sections this cannot be achieved but our method still gives the correct intersection product, *inside* the scheme-theoretic intersection $`D_1D_2`$ as well. The moral is that $$𝒪_D𝒪_D\text{“should be”}\{L^1\stackrel{s}{}𝒪_X\}𝒪_D=\{L^1|_D\stackrel{0}{}𝒪_D\}.$$ This still has cokernel $`𝒪_D𝒪_D=𝒪_D`$, but now (because the intersection was not transverse) has kernel too, namely $`L^1|_D`$, containing all the intersection information. The difference of the first Chern classes of these sheaves on $`D`$ is precisely the self intersection $`D.D`$. So just as the dual of $`H_{}`$, in the case of simplicial complexes in Section 2, “should be” given by applying $`\mathrm{Hom}(.,)`$ not to $`H_{}`$ but to the chain complex, here we apply $`𝒪_D`$ to the complex of locally free (and so better behaved) sheaves rather than $`𝒪_D`$. This is the prototype of a derived functor which will be dealt with systematically in Section 6. The kernel $`L^1|_D`$ above will be the first derived functor Tor$`{}_{1}{}^{}(𝒪_D,𝒪_D)`$ of $`𝒪_D`$. Thus we see an example where it is beneficial to consider complexes of sheaves rather than just single (cohomology) sheaves. Having applied $`𝒪_{D_2}`$ we get a genuine complex, potentially with cohomology in more than one degree, i.e. it is not simply the resolution of a single sheaf. But again we should not now pass to its cohomology, as we may want to intersect with further cycles, by tensoring the complex with some $`𝒪_{D_3}`$ if $`X`$ is higher dimensional, for instance. Complexes good, cohomology bad, after all. So having motivated considering *all* complexes (rather than just resolutions of sheaves) we will now set about working with them, forgetting all about derived functors until Section 6. Replacing sheaves by complexes of which they are the cohomology, i.e. by resolutions, we come across the problem mentioned earlier: how do we pick a resolution functorially ? Again there is an analogous issue in topology. ## 3. Quasi-isomorphisms ### Quasi-isomorphisms in topology We saw in Section 2 that a good invariant of a (simply connected) topological space $`|X|`$ underlying a simplicial complex $`X`$ is the simplicial chain complex $`C_{}^X`$ which determines the homotopy type of $`|X|`$ by the Whitehead theorem. The problem is functoriality: how to pick a triangulation of $`|X|`$ canonically, to pass from the topological space to a complex. The standard mathematical trick is to consider all at once on an equal footing, for instance by making them all isomorphic. As described in Section 2, different triangulations of a space (yielding different chain complexes $`C_{},D_{}`$) may have no simplicial map between them, but taking finer subdivisions and using simplicial approximations we can find a third chain complex $`E_{}`$ fitting into the diagram where both maps are *quasi-isomorphisms* – they are chain maps inducing isomorphisms on homology. We extend this to be an equivalence relation; thus two chain complexes are quasi-isomorphic if they can be related by a sequence of quasi-isomorphisms of the above type. So quasi-isomorphic complexes have the same homology, but the converse does *not* hold; for instance the complexes $$[x,y]^2\stackrel{(x,y)}{}[x,y]\mathrm{and}[x,y]\stackrel{0}{}$$ have the same homology but are not quasi-isomorphic. Quasi-isomorphism is exactly the equivalence relation we want on complexes: by the Whitehead theorem (2.2), $`|X|`$ and $`|Y|`$ are homotopy equivalent if and only if there is a $`Z`$ inducing a quasi-isomorphism (3.1) Thus we would like to think of quasi-isomorphic complexes as isomorphic. Though there may not be a map between them (as there may be no simplicial map between two different triangulations of the same space) we pretend there is by putting one in by hand, putting in the dotted arrow in (3.1) even if it does not exist as a genuine map. We consider the complexes to be isomorphic since the underlying spaces are, after all. In particular homotopy equivalences are quasi-isomorphisms. Algebraically this means that if there is an $`s:C_{}D_{+1}`$ such that $`fg=_Ds+s_D`$ for two chains maps $`f,g:C_{}D_{}`$, then $`f`$ and $`g`$ induce the same map on homology. ### Quasi-isomorphisms in algebraic geometry The cokernel map (2.3) $$\{L^1𝒪_X\}𝒪_D$$ is of course a quasi-isomorphism (it induces an isomorphism on cohomology). Similarly any resolution $`F^{}`$ (or $`F^{}`$) gives a quasi-isomorphism $`\{F^{}\}`$ (or $`\{F^{}\}`$). As in the topological case we want to consider these as equalities (more strictly, isomorphisms) $`\{F^{}\}`$, to make choice of a resolution functorial. So again we need to introduce inverse arrows $`\{F^{}\}`$ (or $`\{F^{}\}`$), which is what we turn to now. ## 4. Category theory In general we would like to replace *objects* (e.g. sheaves, vector bundles, abelian groups, modules, etc.) by complexes of objects that are *“better behaved”* for some particular *operation* such as $`A,\mathrm{Hom}(.,A),\mathrm{Hom}(A,.),\mathrm{\Gamma }(.),\mathrm{}`$ (better behaved in the sense that *no information is lost* when the operation is applied to such objects). Sometimes there is a canonical such complex, in general we want to make its choice *natural*. The words in italics are meant to be suggestive of category theory; abelian categories are the natural setting for the general theory, with the above words being replaced by, respectively, objects of a category, acyclic objects, functor, exact, functorial. We shall now define such things; those scared of the word category should think of the (category of) sheaves on a complex manifold. ###### Definition 4.1. An additive category is a category $`𝒜`$ such that * Each set of morphisms $`\mathrm{Hom}(A,B)`$ forms an abelian group. * Composition of morphisms distributes over the addition of morphisms given by the abelian group structure, i.e. $`f(g+h)=fg+fh`$ and $`(f+g)h=fh+gh`$. * There exist products (direct sums) $`A\times B`$ of any two objects $`A,B`$ satisfying the usual universal properties (see e.g. \[Ma\]). * There exists a zero object $`0`$ such that $`\mathrm{Hom}(0,0)`$ is the zero group (i.e. just the identity morphism). Thus $`\mathrm{Hom}(0,A)=0=\mathrm{Hom}(A,0)`$ for all $`A`$, and the unique zero morphism between any two objects is the one that factors through the zero object. An abelian category is an additive category that also satisfies * All morphisms have kernels and cokernels such that monics (morphisms with zero kernel) are the kernels of their cokernel maps and epis (zero cokernel) are the cokernels of their kernels. The kernels and cokernels mentioned above are defined in a general category via obvious universal properties (e.g. a kernel of $`f:AB`$ is a $`KA`$ through which any $`g:CA`$ factors uniquely if and only if $`fg=0`$). We refer to \[Ma\] for such things; we shall only be concerned with concrete categories (those whose objects are sets) and then kernels and cokernels will be the usual ones, and the final axiom will be immediate. So in an abelian category we can talk about exact sequences and *chain complexes*, and cohomology of complexes. Additive functors between abelian categories are *exact* (respectively left or right exact) if they preserve exact sequences (respectively short exact sequences $`0ABC`$ or $`ABC0`$). ###### Definition 4.2. The bounded derived category $`D^b(𝒜)`$ of an abelian category $`𝒜`$ has as objects bounded (i.e. finite length) $`𝒜`$-chain complexes, and morphisms given by chain maps with quasi-isomorphisms inverted as follows (\[GM\] III 2.2). We introduce morphisms $`f`$ for every chain map between complexes $`f:X_fY_f`$, and $`g^1:Y_gX_g`$ for every quasi-isomorphism $`g:X_g\stackrel{}{}Y_g`$. Then form all products of these morphisms such that the range of one is the domain of the next. Finally identify any combination $`f_1f_2`$ with the composition $`f_1f_2`$, and $`gg^1`$ and $`g^1g`$ with the relevant identity maps id$`_{Y_g}`$ and id$`_{X_g}`$. ###### Remark 4.3. Similarly one can define the unbounded derived category, and the categories $`D^+(𝒜),D^{}(𝒜)`$ of bounded below and above complexes respectively. We shall use $`D(𝒜)`$ to mean one of the four such derived categories. Morphisms in $`D(𝒜)`$ are represented by roofs where $`s`$ is a quasi-isomorphism, and we set $`f=gs^1`$ (it is clear any morphism is a composition of such roofs; that one will suffice is proved in \[GM\] III 2.8). This “localisation” procedure of inverting quasi-isomorphisms has some remarkable properties (for instance we shall see that homotopic maps are identified with each other to give the same morphism in the derived category). Although they give $`D(𝒜)`$ the structure of an additive category, we will see it does not have kernels or cokernels. As ever we go back to the topology to see why not, what they are replaced by, and what the structure of $`D(𝒜)`$ is (since it is not that of an abelian category). ## 5. Cones and triangles When working with topological spaces (or simplicial or cell complexes) up to homotopy there is no notion of kernel or cokernel. In fact the standard cylinder construction shows that any map $`f:XY`$ is homotopic to an inclusion $`X`$cyl$`(f)=Y(X\times [0,1])/f(x)(x,1)`$, while the path space construction shows it is also homotopic to a fibration. For some *fixed* maps $`f:XY`$, rather then equivalence classes of homotopic maps, we can make sense of the kernel (the fibre of $`f`$ if it is a fibration) or cokernel ($`Y/X`$, the space with the image of $`X`$ collapsed to a point, if $`f`$ is a cofibration, which means an inclusion for our purposes). In general of course neither makes sense, but there is something which acts as both, namely cones and the Dold-Puppe construction. The cone $`C_f`$ on a map $`f:XY`$ is the space formed from $`Y(X\times [0,1])`$ by identifying $`X\times \{1\}`$ with its image $`f(X)Y`$, and collapsing $`X\times \{0\}`$ to a point. It fits into the sequence of maps It is clear that this can act as a cokernel, in that if $`X\stackrel{f}{}Y`$ is an inclusion, then $`C_f`$ above is clearly homotopy equivalent to $`Y/X`$. In fact we can now iterate this process, forming the cone on the natural inclusion $`i:YC_f`$ to give the sequence: By retracting $`Y`$ to the right hand apex of $`C_i`$ in the last term of the above diagram, we see that $`C_i`$ is homotopic to $`\mathrm{\Sigma }X`$, the suspension of $`X`$. Thus, up to homotopy, we get a sequence (5.1) $$XYY/X\mathrm{\Sigma }X\mathrm{}$$ Taking the $`i`$th cohomology $`H_i`$ of each term, and using the suspension isomorphism $`H_i(\mathrm{\Sigma }X)H_{i1}(X)`$ gives a sequence (5.2) $$H_i(X)H_i(Y)H_i(Y,X)H_{i1}(X)H_{i1}(Y)\mathrm{}$$ which is just the long exact sequence associated to the pair $`XY`$ (it is an interesting exercise to check that the map $`H_i(Y/X)H_i(\mathrm{\Sigma }X)`$ induced above is indeed the boundary map $`H_i(Y,X)\stackrel{_{}}{}H_{i1}(X)`$). Up to homotopy we can make (5.1) into a sequence of simplicial maps, so that taking the associated chain complexes we get a lifting of the long exact sequence of homology (5.2) to the level of complexes. Of course it exists for all maps $`f`$, not just inclusions, with $`Y/X`$ replaced by $`C_f`$. For instance if $`f`$ is a fibration, $`C_f`$ acts as the “kernel” or fibre of the map. The most extreme case is $`f:X`$ point, which gives $`C_f=\mathrm{\Sigma }X`$, the suspension of the fibre $`X`$, which to homology is $`X`$ shifted in degree by 1. So $`C_f`$ acts as a combination of both cokernel and kernel, and as such gives an easy proof of the Whitehead theorem we have been quoting: i.e. if $`f:XY`$ is a map inducing an isomorphism of homology groups of simply connected spaces then the sequence $$H_i(X)H_i(Y)H_i(C_f)H_{i1}(X)H_{i1}(Y)\mathrm{}$$ shows that $`H_{}(C_f)`$=0. Thus, by the Hurewicz theorem, the homotopy groups $`\pi _{}(C_f)`$ are zero too, making $`C_f`$ homotopy equivalent to a point by the more famous Whitehead theorem. It is an easy consequence of this that $`f`$ is a homotopy equivalence. The final thing we note from the topology is that if $`X`$ and $`Y`$ are simplicial complexes, with $`f:XY`$ a simplicial map, then the cone $`C_f`$ is naturally a simplicial complex with $`i`$-simplices being those in $`Y`$, plus the cones on $`(i1)`$-simplices in $`X`$. A little thought shows that this makes the corresponding (cohomology, for convenience) complex $$C_X^{}[\mathrm{\hspace{0.17em}1}]C_Y^{}\mathrm{with}\mathrm{differential}d_{C_f}=(\begin{array}{cc}d_X[\mathrm{\hspace{0.17em}1}]& 0\\ f& d_Y\end{array}),$$ where $`[n]`$ means shift a complex $`n`$ places left. Thus we can define the cone $`C_f`$ on any map of chain complexes $`f:A^{}B^{}`$ in an abelian category $`𝒜`$ by the above formula, replacing $`C_X^{}`$ by $`A^{}`$ and $`C_Y^{}`$ by $`B^{}`$. If $`A^{}=A`$ and $`B^{}=B`$ are chain complexes concentrated in degree zero then $`C_f`$ is the complex $`\{A\stackrel{f}{}B\}`$. This has zeroth cohomology $`h^0(C_f)=`$ ker$`f`$, and $`h^1(C_f)=`$ coker $`f`$, so combines the two (in different degrees). In general it is just the total complex of $`A^{}B^{}`$. There is an obvious map $`i:B^{}C_f`$, and $`C_i`$ is, as could be guessed from above, quasi-isomorphic to $`A^{}[\mathrm{\hspace{0.17em}1}]`$. So what we get in a derived category is not kernels or cokernels, but “exact triangles” $$A^{}B^{}C^{}A^{}[\mathrm{\hspace{0.17em}1}].$$ (For a proof see (\[GM\] III 3.5) – we should really modify $`B`$, replacing it by the quasi-isomorphic cyl $`(f)`$, but once quasi-isomorphisms are inverted it does not matter.) Thus we have long exact sequences rather than short exact ones; taking $`i`$th cohomology $`h^i`$ of the above gives the standard long exact sequence $$h^i(A^{})h^i(B^{})h^i(C^{})h^{i+1}(A^{})\mathrm{}$$ Thus $`D(A)`$ is not an abelian category, it is an example of a *triangulated category*. This is an additive category with a functor $`T`$ (often denoted $`[\mathrm{\hspace{0.17em}1}]`$) and a set of *distinguished triangles* satisfying a list of axioms. We refer to (\[GM\] IV 1.1) for the precise details, but the triangles include, for all objects $`X`$ of the category, $$X\stackrel{\mathrm{id}}{}X0X[\mathrm{\hspace{0.17em}1}],$$ and any morphism $`f:XY`$ can be completed to a distinguished triangle $$XYCX[\mathrm{\hspace{0.17em}1}].$$ There is also a derived analogue of the 5-lemma, and a compatibility of triangles known as the octahedral lemma, which is pretty unable, as you might imagine. Why are we abstracting again ? It is because this structure that is present in $`D(𝒜)`$, for an abelian category $`𝒜`$, is also present in topology (as we have been hinting all along) without any underlying abelian category. I.e. the topology described throughout this talk might lead one to suspect that some category of nice topological spaces should have just this structure, and we have certainly not constructed it as a derived category of any abelian category. In fact we are not quite there yet, the reason being that the translation functor of suspension $`T=\mathrm{\Sigma }`$ is not invertible on spaces. If it did exist it is clear it would be the loop space functor $`\mathrm{\Omega }`$, as $`\mathrm{\Omega }`$ and $`\mathrm{\Sigma }`$ are adjoints by a standard argument<sup>1</sup><sup>1</sup>1If we denote homotopy classes of maps from $`X`$ to $`Y`$ by $`[X,Y]`$, then applying the invertible $`\mathrm{\Sigma }`$ to $`[X,\mathrm{\Sigma }^1Y]`$ gives an isomorphism $`[X,\mathrm{\Sigma }^1Y][\mathrm{\Sigma }X,Y][X,\mathrm{\Omega }Y]`$. Choosing $`X=\mathrm{\Sigma }^1Y`$ and $`X=\mathrm{\Omega }Y`$ gives canonical maps $`\mathrm{\Sigma }^1Y\mathrm{\Omega }Y`$ that induce isomorphisms on homotopy groups (as is seen by choosing $`X=S^n`$). Thus $`\mathrm{\Sigma }^1Y`$ is, up to weak homotopy equivalence, $`\mathrm{\Omega }Y`$.. In particular $`\mathrm{\Sigma }X\mathrm{\Omega }^1X`$ and we would be able to deloop spaces, making them infinite loop spaces. So it is not surprising to find that *spectra* (essentially infinite loop spaces, plus all their loopings and deloopings to keep track of homotopy in negative degrees) form a triangulated category in the way we have been describing (though there are many technicalities we have bypassed, due to the spaces involved not being simplicial complexes). The final thing to note from the topology is that homotopic maps $`XY`$ get identified when we invert quasi-isomorphisms. By composing with the homotopy $`X\times [0,1]Y`$ it is clear it is enough to show the two inclusions $`\iota _0,\iota _1`$ of $`X`$ into $`X\times [0,1]`$ (as $`X\times \{0\}`$ and $`X\times \{1\}`$ respectively) are identified. But they are both right inverses for the projection $`p:X\times [0,1]X`$: $`p\iota _i=`$ id. As they are also isomorphisms on homology, when we invert quasi-isomorphisms they both become identified with $`p^1`$. Algebraically, in $`D(𝒜)`$, we can mimic the (dual, cohomology) proof by defining the cylinder of a complex in the way dictated by the obvious product cell complex structure on the cylinder of a simplicial complex. Namely consider, for any chain complex $`A^{}`$, $$\mathrm{cyl}(A^{})=A^{}A^{}[\mathrm{\hspace{0.17em}1}]A^{},\mathrm{with}\mathrm{differential}d=(\begin{array}{ccc}d_A& 0& 0\\ \mathrm{id}& d_A[\mathrm{\hspace{0.17em}1}]& \mathrm{id}\\ 0& 0& d_A\end{array}).$$ Set $`\iota _1^{},\iota _2^{}`$ to be the projections to the first and third factors respectively, with $`p^{}`$ the sum of the corresponding two inclusions. These are all chain maps, and in fact quasi-isomorphisms, with $`\iota _i^{}p^{}=`$id. Thus $`\iota _1^{},\iota _2^{}`$ are identified with $`(p^{})^1`$ in $`D(𝒜)`$ and so also with each other. ## 6. Derived functors Now we can go back to (derived) functors and replace arbitrary complexes by resolutions of objects which are suited to the functor concerned. We deal with left exact functors; right exact functors are similar. ###### Definition 6.1. Let $`𝒜`$ and $``$ be abelian categories. A class of objects $`𝒜`$ is *adapted* (\[GM\] III 6.3) to a left exact functor $`F:𝒜`$ if * $``$ is stable under direct sums, * $`F`$ applied to an acyclic complex in $``$ (that is, a complex with vanishing cohomology) is acyclic, and * any $`A𝒜`$ injects $`0AR`$ into some $`R`$. Let $`K^+()`$ be the category of bounded below chain complexes in $``$ with morphisms homotopy equivalence classes of chain maps. Then \[GM\] inverting quasi-isomorphisms in $`K^+()`$ gives a category equivalent to $`D^+(𝒜)`$. The final statement is a fancy way of saying we have finally reached our goal – we can functorially replace (i.e. resolve) any $`𝒜`$-complex by a quasi-isomorphic $``$-complex using the conditions of the definition. Examples of such $`𝒜`$ include $`\{`$injective sheaves$`\}\{`$quasi-coherent sheaves$`\}`$, or for right exact functors, $`\{`$projectives modules over a ring$`\}\{`$all modules$`\},\{`$locally free sheaves over an affine variety or scheme$`\}\{`$coherent sheaves$`\},\{`$flat sheaves$`\}\{`$quasi-coherent sheaves$`\}`$ if $`F`$ is tensoring with a sheaf, etc. So instead of applying $`F`$ to arbitrary complexes, we apply it only to $``$-complexes. Thus we define the *right derived functor* $`𝐑F`$ of $`F`$ to be the composition $`D^+(𝒜)K^+()/\mathrm{q}.\mathrm{i}.\stackrel{F}{}D^+()`$. (The more classical right derived functors $`R^iF`$ are the cohomology $`R^iF=h^i(𝐑F)`$, but taking cohomology is bad, we now know.) We are glossing over some details here; really $`𝐑F`$ should be defined by some universal property (\[GM\] III 6.11) to make it independent of $``$ up to canonical isomorphism. Similarly we can define the left derived functor $`𝐋F`$ of a right exact $`F`$. This gives an *exact functor* $`𝐑F:D^+(𝒜)D^+()`$, i.e. it takes exact triangles to exact triangles. In particular taking cohomology gives the classical long exact sequence of derived functors of the form $$\mathrm{}R^iF(A)R^iF(B)R^iF(C)R^{i+1}F(A)\mathrm{}$$ For instance the right derived functor of the left exact global sections functor $`\mathrm{\Gamma }:\{`$ Sheaves $`\}\{`$ Vector spaces $`\}`$ is just sheaf cohomology $`𝐑\mathrm{\Gamma }=𝐑H^{}`$, with cohomology (of the complex of vector spaces) the standard sheaf cohomology $`H^i`$. Then the above sequence becomes the long exact sequence in cohomology. There are two main advantages of this approach. Firstly that we have managed to make the *complex* $`𝐑F(A)`$, rather than its less powerful cohomology $`R^iF(A)`$, into an invariant of $`A`$, unique up to quasi-isomorphism. Secondly, the derived functor has simply become the original functor applied to complexes (though not arbitrary ones, they have to be in $``$). This gives easier and more conceptual proofs for results about derived functors that usually require complicated double complex, spectral-sequence type arguments. We give some examples in sheaf theory, skating over a few technical conditions (issues about boundedness of complexes and resolutions that are certainly not a problem on a smooth projective variety; for precise statements see \[HRD\]): ###### Example 6.2. Tensor product of sheaves is symmetric, thus its derived functor $`\stackrel{𝐋}{}`$, and its homology Tor<sub>i</sub>, are symmetric: Tor$`{}_{i}{}^{}(A,B)`$ Tor$`{}_{i}{}^{}(B,A)`$. Here we simply tensor complexes of flat sheaves (for instance, locally free sheaves). ###### Example 6.3. $`𝐑om`$ of sheaves, being just local $`om`$ on complexes, can be defined by resolving either the first variable by locally frees or the second by injectives. ###### Example 6.4. Under some mild conditions, $`𝐑(FG)𝐑F𝐑G`$. If we take cohomology before applying $`𝐑F`$ we get an approximation to $`𝐑(FG)`$, and the Grothendieck spectral sequence $`R^iF(R^jG)R^{i+j}(FG)`$. For instance for a morphism $`p:XY`$ the equality $`\mathrm{\Gamma }_X=\mathrm{\Gamma }_Yp_{}`$ yields $`𝐑\mathrm{\Gamma }_X𝐑\mathrm{\Gamma }_Y𝐑p_{}`$, and so the Leray spectral sequence $`H_X^{i+j}H_Y^i(R^jp_{})`$. Similarly $`\mathrm{Hom}=\mathrm{\Gamma }om`$ yields $`𝐑\mathrm{Hom}𝐑H^{}(𝐑om)`$ and the local-to-global spectral sequence Ext$`{}_{}{}^{i+j}H^i(xt^j)`$. Hypercohomology $``$ is given by the derived functor of global sections of complexes (so it is nothing but normal sheaf cohomology for us). Applying the above to $`F=\mathrm{\Gamma }`$ and $`G=`$ id we get the hypercohomology spectral sequence $`H^i(h^j(A^{}))^{i+j}(A^{})`$. ###### Example 6.5. For $`A`$ a sheaf or complex of sheaves, denote by $`A^{}`$ the dual complex $`𝐑om(A,𝒪)`$. Then $`𝐑om(A,B)B\stackrel{𝐋}{}A^{}`$. ###### Example 6.6. We now consider an example on a real $`n`$-manifold, namely the DeRham theorem. Let $`A^i()`$ denote the sheaf of $`C^{\mathrm{}}`$ $`i`$-forms on a manifold $`X`$, with $`d`$ the exterior derivative. Then the Poincaré lemma gives a quasi-isomorphism (resolution of the constant sheaf $``$) $$\{A^0()\stackrel{d}{}A^1()\stackrel{d}{}\mathrm{}\stackrel{d}{}A^n()\}.$$ Because of the existence of partitions of unity the $`A^i()`$ sheaves are acyclic for the global sections functor $`\mathrm{\Gamma }`$, i.e. they have no higher sheaf cohomology (they are what is known as *fine* sheaves). Thus it is easy to see that we may apply $`\mathrm{\Gamma }`$ to the resolution to obtain a complex isomorphic to the derived functor $`𝐑\mathrm{\Gamma }()`$ of the constant sheaf $``$, i.e. its sheaf cohomology. This yields $$\mathrm{\Omega }^0()\stackrel{d}{}\mathrm{\Omega }^1()\stackrel{d}{}\mathrm{}\stackrel{d}{}\mathrm{\Omega }^n(),$$ which means that the DeRham complex computes the real cohomology $`H^{}(X;)`$ of the manifold $`X`$. Similarly taking cohomology of the Dolbeault complex or resolution $$\{𝒪\stackrel{}{}\mathrm{\Omega }^{1,0}\stackrel{}{}\mathrm{\Omega }^{2,0}\stackrel{}{}\mathrm{}\stackrel{}{}\mathrm{\Omega }^{n,0}\},$$ on a complex manifold $`X`$, gives a spectral sequence relating $`H^{}(X;)`$ to the Hodge groups $`_{i,j}H^i(\mathrm{\Omega }^{j,0})=_{i,j}H^{i,j}(X)`$ (the spectral sequence famously degenerates for a Kähler manifold). Hopefully this talk has shown derived and triangulated categories to be natural objects, just categories of complexes with quasi-isomorphisms made into isomorphisms, which we can think of via topological pictures (modulo some technical details). Unfortunately this is a bit of a disservice to anyone who comes to work in some $`D(𝒜)`$, where it is important to chase all quasi-isomorphisms, compatibilities, etc. The main problem is that while cones are defined up to isomorphism in triangulated categories, choosing them functorially is not possible. This leads many to think that there should be some more refined concept still to be worked out.
warning/0001/cond-mat0001321.html
ar5iv
text
# Counter-ion release and electrostatic adsorption ## Abstract The effective charge of a rigid polyelectrolyte (PE) approaching an oppositely charged surface is studied. The cases of a weak (annealed) and strongly charged PE with condensed counterions (such as DNA) are discussed. In the most interesting case of the adsorption onto a substrate of low dielectric constant (such as a lipid membrane or a mica sheet) the condensed counterions are not always released as the PE approaches the substrate, because of the major importance of the image charge effect. For the adsorption onto a surface with freely moving charges, the image charge effect becomes less important and full release is often expected. A deep understanding of the adsorption of DNA or other charged biomolecules onto oppositely charged membranes is of fundamental importance to understand many key physiological processes, and many experimental studies have approached this problem from very different viewpoints, strongly motivated by applications to gene therapy. On the theoretical side, the adsorption and the interaction of charged rods on a charged surface have been studied within the Debye-Huckel approximation in different situations. In this work, we want to focus on the effective charge of the adsorbed macromolecule. We calculate the attraction energy between an infinitely long, charged cylinder (rigid polyelectrolyte or PE) parallel to an oppositely charged plane, as a function of their distance $`h`$. The energy variation leads to the determination of the equilibrium charge density of the rod as a function of $`h`$. This (effective) charge density can be interpreted in terms of the release of condensed counterions for highly charged PE such as DNA, which are beyond the Manning condensation threshold, or in terms of the recombination of ionized charges on the rod for weak (annealed) PE. We show that in contrast to what would be naively expected, the full release of the counterions condensed onto an highly charged rod is not always observed in the vicinity of an oppositely charged surface. The adsorption energy is derived by perturbating the Gouy-Chapmann solution of the Poisson-Boltzmann equation in a planar geometry. The perturbative treatment is strictly speaking valid only for low linear charge densities $`\tau `$ of the rod $`ł_B\tau 1`$, where $`l_B`$ is the Bjerrum length $`l_B=e^2/(4\pi ϵ)`$ (all energies are in $`k_BT`$ units). However, the physical picture which emerges from this calculation leads to qualitative statements concerning highly charged PE as well. We discuss the adsorption free energy on a substrate of low dielectric constant with respect to water ($`ϵ_w=80`$), which is of most practical importance in biology related problem (adsorption of DNA on a lipid membrane for which $`ϵ_{lp}2`$) and other situations (adsorption onto the mica surfaces of an SFA: $`ϵ_{mc}6`$). We have checked that the case of a membrane of thickness $`l=50A`$ and dielectric constant $`ϵ_{lp}=2`$ does not show quantitative differences with the present situation ($`l=\mathrm{}`$ and $`ϵ_{lp}=0`$). Finally, we also study the case where the charges on the plane are free to adjust to the field created by the rod, a situation of great interest for fluid interfaces such as biological lipid membranes. The electrostatic potential $`\varphi ^{(0)}`$ near a charged wall of density $`\sigma >0`$ (or equivalently with a Gouy-Chapmann length $`\lambda 1/(2\pi l_B\sigma )`$), in a salt solution of average concentration $`n_0`$ (or Debye length $`\kappa ^1`$ with $`\kappa ^28\pi l_Bn_0`$) satisfies the Poisson-Boltzmann (PB) equation. In the low salt limit $`\kappa \lambda 1`$, which is the case discussed in this paper, the potential near the wall ($`\kappa z1`$) follows the Gouy-Chapmann (G-C) law: $$\varphi ^{(0)}=2\mathrm{log}\frac{\kappa (\lambda +z)}{2}n^{(0)}(z)=\frac{1}{2\pi l_B(\lambda +z)^2}$$ (1) where $`z`$ is the coordinate normal to the wall ($`z=0`$ at the interface), $`n^{(0)}`$ is the density of (negative) counterions (or c-i) near the wall. The Gouy-Chapmann solution predicts a dense counterion layer (the G-C layer) of thickness $`\lambda `$ containing a finite fraction of the c-i, followed by a diffuse c-i region. The local screening length in the G-C layer is small: $`L_\kappa \lambda /\sqrt{2}`$, while the screening length of the diffuse region is self-similar: $`L_\kappa z/\sqrt{2}`$. For distances larger than the Debye length: $`z>\kappa ^1`$, the electrostatic potential decreases exponentially. The potential variation $`\delta \varphi `$ due to a negatively charge rod (of linear charge density $`\tau `$) located at an altitude $`z=h`$ above the plane is calculated by a linear expansion of the G-C theory, provided that the perturbation is small. A small perturbation of Eq.(1): $`\delta nn^{(0)}`$ (note that $`\delta n`$ is an algebraic average taking into account the sign of the free charges) supposes that the perturbation potential $`\delta \varphi `$ is smaller than unity. The calculation of the potential is carried out in Fourier space for the coordinate $`x`$ parallel to the wall and perpendicular to the rod: $`\stackrel{~}{f}_q=𝑑xe^{iqx}f(x)`$. The linearized PB equation and the boundary conditions are: $`_z^2\delta \stackrel{~}{\varphi }=\left(q^2+{\displaystyle \frac{2}{(z+\lambda )^2}}\right)\delta \stackrel{~}{\varphi }_z\delta \stackrel{~}{\varphi }|_{z=0}=0`$ (2) $`\mathrm{and}_z\delta \stackrel{~}{\varphi }|_{z=h_+}_z\delta \stackrel{~}{\varphi }|_{z=h_{}}=4\pi l_B\tau `$ (3) with natural boundary conditions for $`z\mathrm{}`$ From the perturbed potential $`\delta \varphi (x,z)`$, the free energy due to the presence of the rod can be calculated by a charging process of the rod: $`\delta =_0^\tau \left(\varphi _{z=h}^{(0)}+\delta \varphi _{z=h}\right)𝑑\tau `$. The first part of the integral gives the interaction energy between the rod and the charged plane and the unperturbed G-C counterion layer, while the second part represents the self-energy of the rod, and its interaction with the perturbed c-i cloud. The direct interaction energy is: $$\delta _{int}=2\tau \mathrm{log}\frac{\kappa (\lambda +h)}{2}$$ (4) The solution of the perturbed PB equation Eq.(3) leads to the “self-energy”: $`\delta _{self}=1/2l_B\tau ^2_p`$, with $`_p={\displaystyle _0^{p_M}}dp{\displaystyle \frac{\left(1+p(1+\overline{h})\right)^2}{p^3(1+\overline{h})^2}}\times `$ (5) $`\left(e^{2p\overline{h}}{\displaystyle \frac{1p+p^2}{1+p+p^2}}{\displaystyle \frac{1p(1+\overline{h})}{1+p(1+\overline{h})}}\right)`$ (6) with $`pq\lambda `$ and $`\overline{h}h/\lambda `$. The integral cutoff is $`p_M2\pi \lambda /a`$ where $`a`$ is a “microscopic” size of order the rod radius ($`a=20A`$ for DNA). This complicated expression can be approximated in the two important limits: for $`h\lambda `$ $$\delta _{self}\frac{1}{2}l_B\tau ^2\left(\frac{2}{3}+\mathrm{log}\left(\frac{4\pi }{3}\frac{h}{a}\right)\right)$$ (7) and for $`h\lambda `$. $`\delta _{self}{\displaystyle \frac{1}{2}}l_B\tau ^2\left(\mathrm{\Gamma }+{\displaystyle \frac{2\pi }{3\sqrt{3}}}+\mathrm{log}{\displaystyle \frac{\pi \lambda ^2}{ah}}\right)`$ (8) The self-energy behaves very differently for large and small distances from the wall. It shows an attraction between the rod and the plane at large distances (the energy increases with the distance), which superimposes to the bare attraction between the two oppositely charged macroions. At short distances, there is a strong (logarithmically divergent) repulsion between the rod and the plane (Eq.(8)). The self-energy shows a deep minimum for a position of the rod which only depends upon the G-C length: $`h_{min}0.8\lambda `$. This minimum results from a balance between the repulsive image charge effect and the attraction due to the screening of the c-i in the Gouy-Chapman layer. The self-energy of a cylinder in a bath of mobile charges is of order $`l_B\tau ^2\mathrm{log}L_\kappa /a`$, where $`L_\kappa `$ is the characteristic (“screening”) length of the bath. This energy can be interpreted as the interaction free energy of the charges on the cylinder (the interaction having a range $`L_\kappa `$). Note that although the free ions are mostly of the same sign as the rod, one can speak of a “screening” effect, as the interaction between the rod and the unperturbed ion cloud is taken into account in $`\delta _{int}`$ (Eq.(4)). In addition to the self-interaction, one should consider the interaction with the image charge, which in the case $`ϵ_{z<0}ϵ_{z>0}`$, is a virtual rod of same charge located at $`z=h`$. This gives an extra contribution $`l_B\tau ^2\mathrm{log}L_\kappa /h`$ per unit length. The large and short distance behaviors Eq.(7,8) can now be explained from the counterion profile Eq.(1), since $`L_\kappa h`$ for $`h>\lambda `$ and $`L_\kappa \lambda `$ for for $`h<\lambda `$. The adsorption of a polyelectrolyte onto a biomembrane or other fluid membranes, which are generally a mixture of charged and neutral lipid molecules, involves the movement of surface charges as a response to the field created by the rod. We address this case, disregarding the fact that the lipid bilayer is a flexible object which would, to a certain extent, wrap around the PE \- this interesting phenomenon will be studied in future works. We also assume that the charge reorganization at the interface is not limited by the availability of moving charges on the plane. The charge on the plane follows a Boltzmann law: $`\sigma =\sigma _0e^{\delta \varphi }`$. As a result, the boundary condition for the perturbed field on the plane (corresponding to Eq.(3)) has to be modified: $`_z\delta \stackrel{~}{\varphi }|_{z=0}=4\pi l_B\sigma _0\delta \stackrel{~}{\varphi }|_{z=0}`$. This affects the self-energy term only, which can be expressed similarly to Eq.(6): $`\delta _{self}=1/2l_B\tau {}_{}{}^{2}_{p}^{bis}`$ with $`_p^{bis}={\displaystyle _0^{p_M}}dp{\displaystyle \frac{\left(1+p(1+\overline{h})\right)^2}{p^3(1+\overline{h})^2}}\times `$ (9) $`\left(e^{2p\overline{h}}{\displaystyle \frac{33p+p^2}{3+3p+p^2}}{\displaystyle \frac{1p(1+\overline{h})}{1+p(1+\overline{h})}}\right)`$ (10) For $`\overline{h}>1`$ this self-energy is equivalent to the constant surface charge case (Eq.(7)), while for $`\overline{h}<1`$, the approximation gives: $$\delta _{self}\frac{1}{2}l_B\tau ^2\left(\mathrm{\Gamma }\frac{\pi }{3\sqrt{3}}+\mathrm{log}\frac{\pi \lambda ^2}{6ah}\right)$$ (11) The mobile (positive) charges of the surface are attracted toward the rod ($`x=0`$); this effectively decreases the “Gouy-Chapman” screening length around the PE and reduces the image charge effect. The case of an annealed surface charge is qualitatively similar to (see scaling in Eq.(11)), but quantitatively different from the case of a quenched surface charge. Since the repulsion of the wall is weakened, the minimum of the self-energy is much deeper, and closer to the wall in the case of moving surface charges. We now discuss the variation of the line charge of an annealed, or weak, polyelectrolyte near a charged wall. The charges on the chain result from a partial ionization of specific chemical groups. The ionization occurs at chemical equilibrium with the free ions in solution, and is governed by quantities such as the pH of the solution. Formally, the PE charge density can be determined by equating the chemical potential of the charges on the chain, to a (given) chemical potential $`\mu _0`$ for the free charges. For an infinitely long rigid PE in a salt solution, the free energy (per unit length) of the charges on the chain is the sum of the translational entropy of the charges along the rod, the electrostatic energy, and the chemical potential: $`=\tau \mathrm{log}\tau a/e+l_B\tau ^2\mathrm{log}\kappa a\mu _0\tau `$. The equilibrium charge density of the rod for a given chemical potential is obtained by differentiation of the free energy: $`\mu _0=\mathrm{log}\tau a+2l_B\tau \mathrm{log}\kappa a`$. The equivalent expression for a charged rod near an oppositely charged plane can be computed from the electrostatic free energy Eq.(4,6): $`\mu _0=\mathrm{log}\tau a+2\pi l_B\tau _p+2\mathrm{log}\kappa (\lambda +h)/2`$. Because of the minimum in the self-energy, the equilibrium charge density $`\tau `$ is maximum for a finite height of order the G-C length, and decreases sharply near the wall. This non-trivial behavior of the charge density of an annealed PE near a wall is mostly due to the importance of the image charge effect in the vicinity of the wall. It is extended further below in the case of highly charge PE with Manning condensation. Note that the optimum charge of the weak PE can reach higher values if the charges on the surfaces are mobile, but still decreases at shorter distances. The perturbative treatment is expected to fail in the important case of the release of the counterions condensed onto a highly charged rod, as the rod approaches an oppositely charged plane. However, the qualitative argument which translates the concentration of free charges into a local screening length should still hold in this case. A charged cylinder surrounded by its counterions undergoes the so-called Manning condensation. Solutions of the poisson-Boltzmann equation in this geometry predict that if the rod is highly charged (namely $`l_B\tau >1`$), a finite fraction $`1\beta `$ of the counterions are confined in the close vicinity of the rod. The electrostatic properties at large distances from the rod are the same as those of a rod with an effective charge $`l_B\tau ^{}1`$. The (over)simplified Oosawa picture of counterion condensation gives a qualitative account of this phenomenon; it is based on a chemical equilibrium between two types of counterions: condensed c-i with a reduced entropy and subjected to a large electrostatic attraction from the rod on the one hand, and free c-i in solution far from the rod on the other hand. This picture can be adapted to the case of a cylinder of charge $`\tau `$ in a salt solution of density $`n_0`$ (or screening length $`\kappa ^1`$). The effective charge of the cylinder, $`\tau ^{}=\beta \tau `$, is obtained by balancing the chemical potentials of the condensed c-i $`\mu _{cond}=\mathrm{log}\left[\frac{(1\beta )\tau v}{\pi a^2}\right]+2l_B\tau \beta \mathrm{log}\left[\kappa a\right]`$ and of the c-i dispersed among the salt molecules $`\mu _{free}=\mathrm{log}\left[n_0v\right]`$ ($`v`$ is the volume of a c-i molecule). The resulting fraction of free c-i $`\beta `$ is: $`\mathrm{log}\left[8l_B\tau (1\beta )\right]=\left(1l_B\tau \beta \right)\mathrm{log}\left[(\kappa a)^2\right]`$. This result is very similar to the Oosawa relationship for a rod which would occupy a (small) volume fraction $`\mathrm{\Phi }=(\kappa a)^2`$ in a salt-free solution, namely $`\beta 1`$ for $`l_B\tau <1`$ and $`\beta l_B\tau 1`$ for $`l_B\tau >1`$. The counterion condensation on a rod near a charged plane can be derived in the same way. The chemical potential of the condensed counterions depends upon the electrostatic potential on the rod, which can be determined (at the level of the scaling laws) using the calculation of the previous sections. Equating the chemical potentials of the free and condensed c-i, we obtain the fraction of free c-i $`\beta `$: $`\mathrm{log}\left[8l_B\tau (1\beta )\right]=\mathrm{log}\left[(\kappa a)^2\right]`$ (12) $`2l_B\tau \beta \mathrm{log}\left[{\displaystyle \frac{ah}{L_\kappa (h)^2}}\right]+2\mathrm{log}\left[{\displaystyle \frac{\kappa (h+\lambda )}{2}}\right]`$ (13) where the local screening length follows the asymptotic behaviors: $`L_\kappa =\lambda `$ for $`0<h<\lambda `$ and $`L_\kappa =h`$ for $`\lambda <h<\kappa ^1`$. Examples of the counterion release as a function of the distance to the wall is shown on Fig.1 (for which proper screening due to the salt has been taken into account for $`\kappa h>1`$). The Manning parameter $`l_B\tau =4`$ and the radius $`a=20A`$ are of order those for DNA, three surface charge densities ranging from $`\lambda =200A`$ to $`50A`$ as been considered, corresponding to one charge every $`(100A)^2`$ to $`(50A)^2`$. The salt concentration $`\kappa ^1=1000A`$ corresponds to a concentration of $`10^5mol/l`$. While the c-i are released as the rod penetrates the Gouy-Chapmann layer, full release ($`\beta =1`$) is only reached for $`\lambda =50A`$ \- the maximum being $`60\%`$ for $`\lambda =200A`$ (it is of the order of $`30\%`$ for the free rod). Furthermore, the free charges recondense at short distance if $`\lambda `$ is large enough; the effective charge reaches $`50\%`$ of the bare charge for $`\lambda =200A`$. It can be shown that for large salt concentration or weak surface charge of the wall: $`\kappa ^2\lambda ^3<2a`$, the PE in contact with the wall can have a lower effective charge than the free PE. It should be noted that the short range repulsion due to the image charge can prevent real adsorption with an equilibrium contact between the rod and the plane. In many cases, however, a short range attraction of non electrostatic origin (such as a hydrophobic force) dominates the adsorption. This interaction must be added to the electrostatic free energy calculated here in order to study the equilibrium adsorption. Different conclusions are reached in the case of moving surface charges. A highly charge rod has a strong effect on the surface charge distribution, for the potential it creates is likely to dominate over the Gouy-Chapman potential, even near the wall. A quantitative description of the phenomenon would require the solution of the full non-linear Poisson-Boltzmann equation, with complex (non-linear as well) boundary conditions. In the following, we merely try to give a feeling of the way moving charges can influence the counterion release of an adsorbed polyelectrolyte. We assume that the surface charge distribution obeys Boltzmann statistics: $`\sigma _{x=0}=\sigma _0e^{\varphi _h}`$, where the potential $`\varphi _h`$ created by the rod at the surface reflects the screening due to the Gouy-Chapmann layer, the image charge, and the fraction $`(1\beta )`$ of condensed counterions: $`\varphi _h2l_B\tau \beta \mathrm{log}(L_\kappa /h)^2`$. Since the local screening length $`L_\kappa (\lambda +h)/\sqrt{2}`$ is influenced by the surface charge $`\sigma `$, which in turn, depends upon the screening length, we obtain a self-consistent relationship for the local Gouy-Chapman length near the rod: $$\lambda (h)\left(\frac{h+\lambda (h)}{\sqrt{2}h}\right)^{4l_B\tau \beta }=\lambda _0$$ (14) This expression shows that moving surface charges strongly reduce the screening length near the rod, which becomes of order $`h`$ when the rod is close to the wall (recall that $`l_B\tau \beta 1`$). As a consequence, the interaction with the image charge and the self-interaction along the rod are both strongly screened. The dominant interaction between the cylinder and the wall is the attractive part given by Eq.(4), and we are likely to observe a full release of the condensed counterions in this case. To summarize, we have studied the evolution of the effective charge of a polyelectrolyte near an oppositely charged plane. The case of a weak PE is studied fairly rigorously, via a perturbative treatment of the non-linear Poisson-Boltzmann equation. We show that at large distance, the charge of the PE increases as the distance to the wall $`h`$ decreases, as expected. However, the charge decreases “strongly” as the PE enters the Gouy-Chapman layer ($`h\lambda `$) because of the combined effect of image charge (for the most common case of a wall with a low dielectric constant) and self-interaction along the PE. These two effects are very much influenced by the value of the Gouy-Chapman length $`\lambda `$, and are partly suppressed in the case of a fluid interface with moving surface charges (for a fluid lipid bilayer for instance), where the charge of the adsorbed PE can reach higher values. The most interesting case of a strongly charge PE beyond the Manning condensation threshold (such as DNA) is discussed qualitatively, using “scaling” arguments inferred from the perturbation theory. We predict that in the case of fixed surface charges, and in contrary to a widely spread idea, most of the condensed counterions are not released if the Gouy-Chapman length is larger than the radius of the rod: $`\lambda a`$. In the case of freely moving surface charges, a full release of the condensed counterions is expected, as the effective Gouy-Chapman length near the rod is of order the rod radius. In all the discussion, we have assumed that the Gouy Chapman length is larger than a molecular size. In many real cases, the two lengths are of the same order of magnitude and the finite size of the ions must be taken into account in order to obtain quantitative results. In this case our results can at best be considered as qualitative. We would like to thank A. Johner and R. Netz for many stimulating discussions.
warning/0001/math-ph0001031.html
ar5iv
text
# An inversion theorem in Fermi surface theory ## 1 Introduction The Fermi surface is an important feature of the quantum field theory of solid state models. Besides being central to the theoretical analysis of such models it is also important from a conceptual point of view. In experiments, one observes and measures the Fermi surface of an interacting system (for brevity, we call this the interacting Fermi surface) – or more precisely, an approximation to it due to positive temperature effects, because the electrons interact with each other (say via a screened Coulomb interaction, phonons and so on). On the other hand, the theoretical analysis usually starts from a model of noninteracting electrons, moving in a crystal background, which exhibits the noninteracting Fermi surface. The effects of the electron–electron interaction are taken into account by ‘turning on a coupling constant’. Thus, while the model of independent electrons exists only theoretically, important notions of solid state physics, for instance Fermi liquid theory, start from it and then incorporate the changes in the system caused by the interaction. One of these is a change in the dispersion relation, that gives the energy of a particle as a function of momentum. This results in the transformation of the Fermi surface from the noninteracting to the interacting one. In this paper, we complete our perturbative analysis of the regularity properties of interacting nonspherical Fermi surfaces by proving an inversion theorem for the map between the interacting and the free dispersion relation that we used in the renormalization of these models. The main ingredients in the inversion theorem are an abstract iteration theorem that generalizes the usual contraction mapping theorem (which is not sufficient here) and a number of regularity estimates. The estimates are used to verify the hypotheses of this iteration theorem. The regularity estimates are an application of the methods and the results of , , and , referred to as I, II, and III in the following. By ‘perturbative analysis’ we mean that the perturbation series is truncated at any finite order $`R`$ (which may be arbitrarily large) in the coupling constant $`\lambda `$. There are situations where this expansion can be proven to converge, so that the limit $`R\mathrm{}`$ exists, but we do not give such bounds here. In the remainder of this introduction, we define our class of models and state the inversion theorem. For a more detailed motivation, see the introductory sections of I and II. ### 1.1 The models Let $`\mathrm{\Gamma }`$ be a nondegenerate lattice in $`^d`$ and $$\mathrm{\Gamma }^\mathrm{\#}=\{𝐛^d:𝐛\gamma 2\pi \text{ for all }\gamma \mathrm{\Gamma }\}$$ (1) its dual lattice. We denote the first Brillouin zone by $``$ and choose it to be the $`d`$-dimensional torus $`=^d/\mathrm{\Gamma }^\mathrm{\#}`$. It is compact. For example, if $`\mathrm{\Gamma }=^d`$, then $`\mathrm{\Gamma }^\mathrm{\#}=2\pi ^d`$ and $`=^d/2\pi ^d`$. We are interested in a class of models characterized by an action $`𝒜(\psi ,\overline{\psi })`$ that is a function of two variables $`\psi =\left(\psi _{k,\sigma }\right)_{k\times ,\sigma \{,\}}`$ and $`\overline{\psi }=\left(\overline{\psi }_{k,\sigma }\right)_{k\times ,\sigma \{,\}}`$. Note that $`\overline{\psi }`$ is not the complex conjugate of $`\psi `$. It is just another vector that is totally independent of $`\psi `$. The zero component $`k_0`$ of $`k`$ is usually thought as an energy, the final $`d`$ components $`𝐤`$ as (crystal) momenta and $`\sigma `$ as a spin. There really should also be a sum over a band index $`n`$, but it will not play a role here and has been suppressed. In these models, the quantities one measures are represented by other functions $`f(\psi ,\overline{\psi })`$ of the same two vectors and the value of the observable $`f(\psi ,\overline{\psi })`$ in the model with action $`𝒜(\psi ,\overline{\psi })`$ is given formally by the ratio of integrals $$f(\psi ,\overline{\psi })_𝒜=\frac{f(\psi ,\overline{\psi })e^{𝒜(\psi ,\overline{\psi })}_{k,\sigma }d\psi _{k,\sigma }d\overline{\psi }_{k,\sigma }}{e^{𝒜(\psi ,\overline{\psi })}_{k,\sigma }d\psi _{k,\sigma }d\overline{\psi }_{k,\sigma }}$$ (2) The integrals are fermionic functional integrals. That is, linear functionals on a Grassmann algebra. A typical action of interest is that corresponding to a gas of electrons, of strictly positive density, interacting through a two–body potential $`u(𝐱𝐲)`$. It is $`𝒜_{\mu ,\lambda }`$ $`=`$ $`_{\sigma \{,\}}{\displaystyle _\times }\frac{d^{d+1}k}{(2\pi )^{d+1}}\left(ik_0\left(\frac{𝐤^2}{2m}\mu \right)\right)\overline{\psi }_{k,\sigma }\psi _{k,\sigma }`$ $`\frac{\lambda }{2}_{\sigma ,\tau \{,\}}{\displaystyle _\times }_{i=1}^4\frac{d^{d+1}k_i}{(2\pi )^{d+1}}(2\pi )^{d+1}\delta (k_1+k_2k_3k_4)`$ $`\overline{\psi }_{k_1,\sigma }\psi _{k_3,\sigma }\widehat{u}(𝐤_1𝐤_3)\overline{\psi }_{k_2,\tau }\psi _{k_4,\tau }`$ Here $`\frac{𝐤^2}{2m}`$ is the kinetic energy of an electron, $`\mu `$ is the chemical potential, which controls the density of the gas, and $`\widehat{u}`$ is the Fourier transform of the two–body interaction. The coupling constant $`\lambda `$ is assumed to be small, so that the interaction is weak. More generally, when the electron gas is subject to a periodic potential due to the crystal lattice, $`\mathrm{\Gamma }`$, and when the electrons are interacting with the motion of the crystal lattice through the mediation of harmonic phonons, the action is of the form $`𝒜_\lambda `$ $`=`$ $`_{\sigma \{,\}}{\displaystyle _\times }\frac{d^{d+1}k}{(2\pi )^{d+1}}\left(ik_0E(𝐤)\right)\overline{\psi }_{k,\sigma }\psi _{k,\sigma }`$ $`\frac{\lambda }{2}_{\sigma ,\tau \{,\}}{\displaystyle _\times }_{i=1}^4\frac{d^{d+1}k_i}{(2\pi )^{d+1}}(2\pi )^{d+1}\delta (k_1+k_2k_3k_4)`$ $`\overline{\psi }_{k_1,\sigma }\psi _{k_3,\sigma }\widehat{v}(k_{1,0}k_{3,0},𝐤_1𝐤_3)\overline{\psi }_{k_2,\tau }\psi _{k_4,\tau }`$ where $`E(𝐤)`$ is the dispersion relation minus the chemical potential $`\mu `$. ### 1.2 The class of dispersion relations Let $``$ be a fundamental cell for the action of the translation group $`\mathrm{\Gamma }^\mathrm{\#}`$. In other words, $``$ is an open set in $`^d`$ with the property that it together with its translates under $`\mathrm{\Gamma }^\mathrm{\#}`$ are dense in $`^d`$. For example, if $`\mathrm{\Gamma }=^d`$, then $`\mathrm{\Gamma }^\mathrm{\#}=2\pi ^d`$ and we may choose $`=(\pi ,\pi )^d`$. Let $`_2=\{𝐩:2𝐩\}`$. For a continuous function $`E`$ from $``$ to $``$ let $$S_E=\{𝐩:E(𝐩)=0\}$$ (5) be the corresponding Fermi surface and $`_E=\{𝐩:E(𝐩)<0\}`$ the corresponding Fermi sea. For $`k2`$ let $$C_s^k(,)=\{EC^k(,):E(𝐩)=E(𝐩)\text{ for all }𝐩\}$$ (6) With the norm $`|f|_k=_{|\alpha |k}D^\alpha f_{\mathrm{}}`$, it is a Banach space. For $`EC_s^k(,)`$, let $`B_\epsilon ^{(k)}(E)=\{eC_s^k(,):\left|eE\right|_k<\epsilon \}`$. For positive constants $`\delta _0,g_0,G_0,\omega _0`$, let $`_s=_s(\delta _0,g_0,G_0,\omega _0)`$ be the set of all $`EC_s^2(,)`$ that satisfy the following conditions $`(i)`$ $`S_E_2,_E\mathrm{},_E,d(S_E,_2)>\delta _0,`$ $`(ii)`$ $`\left|E(𝐩)\right|>g_0\text{ for all }𝐩S_E:`$ $`(iii)`$ $`\left|E\right|_2<G_0`$ $`(iv)`$ $`(𝐭(𝐩),E^{\prime \prime }(𝐩)𝐭(𝐩))>\omega _0\text{ for all }𝐩S_E\text{ and all unit vectors }𝐭(𝐩)`$ $`\text{ tangent to }S_E\text{ at }𝐩`$ Since $`E`$ is $`C^2`$, the condition that $`E0`$ on $`S_E`$ implies that the Fermi surface $`S_E`$ is a $`(d1)`$–dimensional $`C^2`$-submanifold of $``$, (in $`d=2`$, the ‘surface’ is a curve). The condition $`(𝐭(𝐩),E^{\prime \prime }(𝐩)𝐭(𝐩))>\omega _0`$ implies that $`S_E`$ has strictly positive curvature everywhere. The set $`_s`$ is open in $`(C_s^2(,),||_2)`$. In this paper, we fix any $`\delta _0>0,g_0>0,\omega _0>0`$ and $`G_0>\mathrm{max}\{g_0,\omega _0\}`$. ### 1.3 The class of interactions We also define the class $`𝒱`$ of allowed interactions to be the set of all functions $`V`$, whose Fourier transforms $`\widehat{v}(p_0,𝐩)`$ obey $`(i)`$ $`\left|\widehat{v}\right|_21`$ $`(ii)`$ $`\widehat{v}(p_0,𝐩)=\overline{\widehat{v}(p_0,𝐩)}`$ $`(iii)`$ $`\widehat{v}(p_0,𝐩)=\widehat{v}(p_0,𝐩)`$ $`(iv)`$ There is a bounded function $`\stackrel{~}{v}C^2(,)`$ and an $`\alpha >0`$ such that $`\underset{p_0\mathrm{}}{lim\; sup}|p_0|^\alpha \underset{𝐩}{sup}\left|\widehat{v}(p_0,𝐩)\stackrel{~}{v}(𝐩)\right|<\mathrm{}`$ Condition (iv) is used only in the large $`k_0`$ regime. If an ultraviolet cutoff is placed on $`k_0`$, it may be omitted. Condition (i) implies that the interaction in momentum space, $`\widehat{v}`$, is in $`C^2(^{d+1})`$. This is the case if the position space integral kernel $`V(xy)`$ is bounded by $`\frac{\text{ const }}{1+\left|xy\right|^{d+3+\epsilon }}`$ for some $`\epsilon >0`$. The $`1`$ in the condition $`\left|\widehat{v}\right|_21`$ is not a restriction, since $`V`$ and $`\lambda `$ appear only in the combination $`\lambda V`$ in the definition of the model, so a rescaling of $`V`$ can be absorbed by a rescaling of $`\lambda `$. ### 1.4 The counterterm function In I, we constructed a counterterm function $`K`$ as a formal power series in $`\lambda `$, $$K(e,\lambda V,𝐩)=\underset{r=1}{\overset{\mathrm{}}{}}\lambda ^rK_r(e,V,𝐩)$$ (7) where $`K_r:𝒟\times 𝒱\times `$ is defined for a set $`𝒟`$ of dispersion relations $`e`$ with $`_s𝒟`$. The conditions required for having a finite $`K_r`$ for all $`r`$ are much weaker than the conditions we impose here (see I and Section 4). $`K`$ is constructed such that, for a model with action $`𝒜_\lambda `$ $`=`$ $`_\sigma {\displaystyle _\times }\frac{d^{d+1}k}{(2\pi )^{d+1}}\left(ik_0e(𝐤)K(e,\lambda V,𝐤)\right)\overline{\psi }_{k,\sigma }\psi _{k,\sigma }`$ $`\frac{\lambda }{2}_{\sigma ,\tau }{\displaystyle _\times }_{i=1}^4\frac{d^{d+1}k_i}{(2\pi )^{d+1}}(2\pi )^{d+1}\delta (k_1+k_2k_3k_4)`$ $`\overline{\psi }_{k_1,\sigma }\psi _{k_3,\sigma }\widehat{v}(k_{1,0}k_{3,0},𝐤_1𝐤_3)\overline{\psi }_{k_2,\tau }\psi _{k_4,\tau }`$ the Fermi surface of the interacting model is fixed to $`S_e`$, independently of $`\lambda `$. The function $`K(𝐩)`$ is real–valued, and under the symmetry hypotheses made here, $`K(𝐩)=K(𝐩)`$. By introducing the counterterm function, we removed the infrared divergences to all orders in the perturbation expansion in powers of $`\lambda `$. That is, when the expansion is truncated at any finite order $`R`$, all Green functions are finite almost everywhere. We showed in I that the counterterm function to any order $`R`$ in $`\lambda `$, $$K^{(R)}(e,\lambda V)=\underset{r=1}{\overset{R}{}}\lambda ^rK_r(e,V),$$ (9) is differentiable in $`𝐩`$ and $`e`$ (and, of course, $`C^{\mathrm{}}`$ in $`\lambda `$ since it is a polynomial for any finite $`R`$). Thus a model that has an action whose quartic part (in the fields) is that corresponding to $`V`$ and whose quadratic part is that corresponding to a dispersion relation $`E`$ will have an interacting Fermi surface that is the zero set of a dispersion relation $`e`$ if $$e+K(e,\lambda V)=E.$$ (10) In this paper we take a given $`E`$ and $`V`$ and solve $$e+K^{(R)}(e,\lambda V)=E.$$ (11) for $`e=e^{(R)}(E,\lambda V)`$. The dispersion relation $`e`$ that appears in the propagator is only an auxiliary quantity, which is to be determined by $`(\text{11})`$. We shall solve $`(\text{11})`$ by iteration, starting from the given $`E`$. Clearly this requires having bounds with uniform constants on a set of dispersion relations that is mapped to itself by the function $`1\mathrm{l}+K^{(R)}`$. We proved in IIII that the following estimate holds (Theorem III.3.13). For all $`r1`$, there are constants $`\kappa _r>0`$ such that, for all $`e_s(\delta _0,g_0,G_0,\omega _0)`$ and $`V𝒱`$, the contribution $`K_r(e,V)`$ is in $`C_s^2(,)`$ and obeys $$\left|K_r(e,V)\right|_2\kappa _r.$$ (12) The constant $`\kappa _r`$ depends only on $`(\delta _0,g_0,G_0,\omega _0)`$ and $`r`$. Consequently, $`K^{(R)}`$ satisfies $$\left|K^{(R)}(e,\lambda V)\right|_2\underset{r=1}{\overset{R}{}}|\lambda |^r\kappa _r$$ (13) so $`|K^{(R)}(e,\lambda V)|_2`$ can be made arbitrarily small by decreasing $`\lambda `$. Because $`_s`$ is open in $`||_2`$, $`e+K^{(R)}(e,\lambda V)_s`$ if $`e_s`$ and $`\lambda `$ is sufficiently small. ### 1.5 The inversion theorem To show that an iteration scheme for the solution converges, we need to have bounds for the distance between successive elements of the iteration sequence. For technical reasons that have nothing to do with the analysis of IIII and that will be explained later, we have to restrict to dispersion relations that have certain third order derivatives bounded, in order to control the distance between successive iterates. This is the reason why, in the following theorem, the starting $`E_0`$ is required to be in $`C^3`$. ###### Theorem 1 Let $`\delta _0,g_0,\omega _0>0`$, $`G_0>\mathrm{max}\{g_0,\omega _0\}`$ and $`R`$. Then there is a $`\lambda _R>0`$ such that for each $`\left|\lambda \right|\lambda _R`$, each $`E_s(\delta _0,g_0,G_0,\omega _0)C^3(,)`$ and each $`V𝒱`$, there is a unique $`e^{(R)}_s(\delta _0/2,g_0/2,2G_0,\omega _0/2)`$ solving $`(\text{11})`$. Moreover, there is a constant $`A_R>0`$ such that $$\left|e^{(R)}E\right|_2A_R\left|\lambda \right|.$$ (14) Theorem 1 follows from the more detailed Theorem 2 below. We shall discuss the more detailed theorems about inversion in Section 5. In this paper, we do not prove optimal bounds about the $`R`$–dependence of $`\lambda _R`$. For the models at hand, in particular because of the symmetry $`E(𝐩)=E(𝐩)`$, one expects that convergence does not hold at zero temperature. That is, one expects $`\lambda _R0`$ as $`R\mathrm{}`$. The reason for this is that at temperatures below a critical temperature, the ground state of the system is superconducting, in which case the above perturbation expansion cannot converge. As noted in , at a positive temperature $`T=\frac{1}{\beta }>0`$, one can expect convergence of the expansion for coupling constants $`\lambda `$ in the region where $`\lambda \mathrm{log}\beta `$ is small enough, that is, for $`TT_0\mathrm{e}^{\lambda _0/|\lambda |}`$ where $`\lambda _0`$ and $`T_0`$ are fixed constants (see for a Fermi liquid criterion based on this convergence). For $`d=2`$, a proof of this may be possible using the techniques of . The bounds derived here do not change in an essential way at positive temperature. So a variant of our theorems can be expected to hold in this convergent positive temperature regime. Note, however that convergence of the expansion for $`K`$ does not imply that the solution of the inversion equation can be expanded in $`\lambda `$. In fact, it can’t. See for an informal explanation. ## 2 Preliminaries ### 2.1 Coordinates Since $`e`$ is going to change under the iteration, it is convenient to use momentum space coordinates that are independent of $`e`$. Under our assumptions, we can simply use polar coordinates in addition to the Fermi surface coordinates that we used in IIII. We shall review the latter shortly. It will be important that the angular variables $`\theta `$ are the same in both coordinate systems. Only the radial coordinate is different. Polar coordinates: Consider a small ball $`B`$ around an $`E_0_s`$. Regard a small neighbourhood of the Fermi surface $`S_{E_0}`$ as a subset of $`^d`$ instead of the torus $``$ and introduce polar coordinates $`(r,\theta )_0^+\times S^{d1}`$, $`𝐩=𝐩(r,\theta )`$. For $`d=2`$, $`\theta S^1`$. In polar coordinates, the Fermi surface can be parametrized, for $`eB`$, as $$S_e=\{𝐩(r_F(e,\theta ),\theta ):\theta S^{d1}\}$$ (15) with $`r_F:B\times S^{d1}^+`$. If $`eC^k(,)`$, then $`r_FC^k(S^{d1},^+)`$. ###### Lemma 1 Let $`0<kK`$. Let $`S`$ be a $`(d1)`$–dimensional $`C^2`$ convex surface in $`^d`$ all of whose principal curvatures are between $`k`$ and $`K`$. Let $`𝐜_1,𝐜_2`$ be any two maximally separated points of $`S`$. That is, $`𝐜_1,𝐜_2S`$ with $$𝐜_1𝐜_2=\mathrm{max}\left\{𝐩_1𝐩_2:𝐩_1,𝐩_2S\right\}$$ (16) Set $`𝐜=\frac{1}{2}\left(𝐜_1+𝐜_2\right)`$. Then, for every $`𝐩S`$, $$\frac{1}{K}𝐩𝐜\frac{1}{k}$$ (17) and the angle $`\theta (𝐩)`$ between $`𝐩𝐜`$ and the outward pointing normal vector $`𝐧(𝐩)`$ to $`S`$ at $`𝐩`$ obeys $$\mathrm{cos}(\theta (𝐩))\frac{k}{K}$$ (18) If, in addition, $`𝐩S`$ for every $`𝐩S`$, then $`𝐜`$ is the origin. Proof: See Appendix A ###### Lemma 2 Let $`\delta _0,g_0,\omega _0>0`$ and $`G_0>\mathrm{max}\{g_0,\omega _0\}`$. There are $`\epsilon ,r_0,g_1>0`$ such that, for every $`E_0_s(\delta _0,g_0,G_0,\omega _0)`$ and every $`eB_\epsilon ^{(2)}(E_0)`$ $$e_s(\delta _0/2,g_0/2,2G_0,\omega _0/2)$$ (19) and $`\left|r_F(e,\theta )r_F(E_0,\theta )\right|r_0`$ $`\text{for all }\theta S^{d1}`$ (20) $`{\displaystyle \frac{}{r}}e(𝐩(r,\theta ))>g_1`$ $`\text{for all }\left|rr_F(E_0,\theta )\right|2r_0,\theta S^{d1}`$ (21) Note that the constants $`\epsilon ,r_0`$ and $`g_1`$ are independent of $`E_0`$. Proof: See Appendix B Let $`E_0,r_0`$ and $`\epsilon `$ be as in Lemma 2. Set $`\underset{¯}{R}(\theta )`$ $`=`$ $`r_F(E_0,\theta )2r_0`$ $`\overline{R}(\theta )`$ $`=`$ $`r_F(E_0,\theta )+2r_0`$ (22) and $`A`$ $`=`$ $`\{(r,\theta ):\theta S^{d1},\underset{¯}{R}(\theta )<r<\overline{R}(\theta )\}`$ $`\stackrel{~}{A}`$ $`=`$ $`\{𝐩(r,\theta ):(r,\theta )A\}`$ (23) Then, $`S_e=\{𝐩:e(𝐩)=0\}\stackrel{~}{A}`$ for all $`eB_\epsilon ^{(2)}(E_0)`$. We use the notation $`\stackrel{~}{F}(r,\theta )=F(𝐩(r,\theta ))`$ for functions in terms of the variables $`r`$ and $`\theta `$, for instance, $`\stackrel{~}{e}(r,\theta )=e(𝐩(r,\theta ))`$. The above Lemma then states that for all $`(r,\theta )A`$, and all $`eB_\epsilon ^{(2)}(E_0)`$, $`_r\stackrel{~}{e}(r,\theta )>g_1>0`$. We could have introduced coordinates in the annulus $`\stackrel{~}{A}`$, based on any vector field that is transversal to $`S_{E_0}`$. This would only have changed the constant in the lower bound for $`_re`$. Fermi surface coordinates: These are the coordinates used in IIII. They are the polar coordinate $`\theta `$ and $`\rho =e(𝐩)`$, and thus obviously depend on $`e`$. We denote the corresponding inverse map, whose range is a neighbourhood of $`E_0`$’s Fermi surface, by $`\pi \pi `$: $$\pi \pi :(\rho _0,\rho _0)\times S^{d1},(\rho ,\theta )\pi \pi (\rho ,\theta )$$ (24) Clearly $`e(\pi \pi (\rho ,\theta ))=\rho `$. The projection to the Fermi surface is obtained by setting $`\rho =0`$. In terms of the polar coordinates, it is constructed as follows. If $`\stackrel{~}{F}:B_\epsilon ^{(2)}(E_0)\times A`$ maps $`(e,r,\theta )\stackrel{~}{F}(e,r,\theta )`$, then $`\mathrm{}_e\stackrel{~}{F}(e,r,\theta )=\stackrel{~}{F}(e,r_F(e,\theta ),\theta )`$. Obviously, $`_r(\mathrm{}_eF)=0`$. Observe that $`\pi \pi (0,\theta )=𝐩(r_F(e,\theta ),\theta )`$. ### 2.2 Norms Let $`_k`$ be the seminorm $`F_k=\underset{|\alpha |=k}{}\underset{p}{sup}\left|^\alpha F(p)\right|`$ and $$\left|F\right|_k=\underset{l=0}{\overset{k}{}}F_l.$$ (25) It does not matter whether we use the norm in Cartesian or polar coordinates since the two are equivalent. We define the radial norms for $`p1`$ as $$\left|F\right|_{p,r}=\left|F\right|_{p1}+_r\stackrel{~}{F}_{p1}$$ (26) and denote the angular norms for $`p0`$ as $`\left|F\right|_{p,\theta }`$. In the latter norms, all derivatives are taken in the $`\theta `$–directions. margin: ###### Lemma 3 1. $`\left|F\right|_{p,r}\left|F\right|_{p+1}`$. 2. For all $`eB_\epsilon ^{(2)}(E_0)`$, $`_r\mathrm{}_eF=0`$, and $$\left|\mathrm{}_eF\right|_{p,r}=\left|\mathrm{}_eF\right|_{p1}=\left|\mathrm{}_eF\right|_{p1,\theta }.$$ (27) 3. $`\left|FG\right|_p`$ $``$ $`2^p\left|F\right|_p\left|G\right|_p,`$ (28) $`\left|FG\right|_p`$ $``$ $`F_0G_p+F_pG_0+2^{p+1}\left|F\right|_{p1}\left|G\right|_{p1},`$ 4. $`\left|FG\right|_{p+1,r}`$ $``$ $`2^{p+2}\left|F\right|_p\left|G\right|_p+_rF_pG_0+F_0_rG_p`$ (29) $``$ $`2^{p+2}\left(\left|F\right|_{p+1,r}\left|G\right|_p+\left|F\right|_p\left|G\right|_{p+1,r}\right).`$ Proof: The first statement is an immediate consequence of $`_rF_pF_{p+1}`$. The second statement is an immediate consequence of the observation that the localization map $`\mathrm{}_e`$ does not depend on $`r`$. For the third and fourth statements, use the Leibniz rule and that $`\left(\genfrac{}{}{0pt}{}{\alpha _k}{\beta _k}\right)\left(\genfrac{}{}{0pt}{}{p}{q}\right)`$ for all $`\alpha _1+\mathrm{}+\alpha _n=p`$ and $`\beta _1+\mathrm{}+\beta _n=q`$ (all nonnegative), to prove that $$FG_p\underset{q=0}{\overset{p}{}}\left(\genfrac{}{}{0pt}{}{p}{q}\right)F_qG_{pq}$$ (30) and $$_r(FG)_p\underset{q=0}{\overset{p}{}}\left(\genfrac{}{}{0pt}{}{p}{q}\right)\left(_rF_qG_{pq}+F_{pq}_rG_q\right).$$ (31) ## 3 The iteration Given $`e_0`$, $`e_1`$, and $`t[0,1]`$, denote $`e_t=(1t)e_0+te_1`$. ###### Theorem 2 Let $`\delta _0,g_0,\omega _0>0`$ and $`G_0>\mathrm{max}\{g_0,\omega _0\}`$. Let $`\epsilon >0`$ be as in Lemma 2. 1. Regularity. For each $`R`$, there is a constant $`D1`$ such that for all $`\left|\lambda \right|1`$, all $`V𝒱`$ and all $`e_s(\delta _0/2,g_0/2,2G_0,\omega _0/2)`$ $$\left|K^{(R)}(e)\right|_{3,r}=\left|K^{(R)}(e)\right|_2<D\left|\lambda \right|.$$ (32) 2. Norm bounds for the iteration. There is $`0<\delta <1`$ (independent of $`\delta _0,g_0,G_0,\omega _0`$) and, for each $`R`$, there are constants $`Q_0,Q_11`$ such that for all $`\left|\lambda \right|1`$, all $`V𝒱`$, all $`E_s(\delta _0,g_0,G_0,\omega _0)`$ and all $`e_0`$ and $`e_1B_\epsilon ^{(2)}(E)C^3`$ $`\left|K^{(R)}(e_1)K^{(R)}(e_0)\right|_0`$ $``$ $`Q_0\left|\lambda \right|\left|e_1e_0\right|_0`$ (33) $`\left|K^{(R)}(e_1)K^{(R)}(e_0)\right|_1`$ $``$ $`Q_0\left|\lambda \right|\left[\left|e_1e_0\right|_0^\delta +\left|e_1e_0\right|_1\right]`$ (34) $`\left|K^{(R)}(e_1)K^{(R)}(e_0)\right|_{3,r}`$ $``$ $`Q_0\left|\lambda \right|\left[\left|e_1e_0\right|_1^\delta +\left|e_1e_0\right|_2\right]`$ (35) $`+`$ $`Q_1\left|\lambda \right|\underset{0t1}{sup}\left|e_t\right|_{3,r}\left|e_1e_0\right|_0.`$ In addition, if $`V_1,V_2𝒱`$ with $`\left|V_1V_2\right|_21`$ and $`e_s`$ then $$\left|K^{(R)}(e,\lambda V_1)K^{(R)}(e,\lambda V_2)\right|_2Q_0\left|\lambda \right|\left|V_1V_2\right|_2.$$ (36) 3. Existence of a unique solution to the inversion equation. Let $`E_s(\delta _0,g_0,G_0,\omega _0)`$ with $`\left|E\right|_{3,r}=G_3<\mathrm{}`$. (This is the case if, for example, $`E_sC^3`$). Set $`Q=\mathrm{max}\{Q_0+Q_1(1+G_3),D\}`$. Let $$B_{\mathrm{rad}}=\{e_s(\delta _0/2,g_0/2,2G_0,\omega _0/2):\left|eE\right|_2<\epsilon ,\left|eE\right|_{3,r}<1\}$$ (37) and let $`\lambda _R>0`$ be such that $`Q\lambda _R<\mathrm{min}\{1,\epsilon \}`$. Then for all $`\left|\lambda \right|\lambda _R`$ and all $`V𝒱`$, there is a unique $`e^{(R)}B_{\mathrm{rad}}`$ such that $`E=e^{(R)}+K^{(R)}(e^{(R)},\lambda V)`$. Moreover $$\left|e^{(R)}E\right|_{3,r}D\left|\lambda \right|.$$ (38) 4. Continuity in $`E`$ and $`V`$. Let $`E,E^{}_s(\delta _0,g_0,G_0,\omega _0)`$ satisfy $`\left|E\right|_{3,r}`$, $`\left|E^{}\right|_{3,r}G_3`$ and $`\left|EE^{}\right|_{3,r}<\epsilon /2`$. Then, for all $`\left|\lambda \right|\lambda _R/2`$ and all $`V,V^{}𝒱`$ with $`\left|VV^{}\right|_21`$, $`\left|e^{(R)}(E,\lambda V)e^{(R)}(E^{},\lambda V^{})\right|_2`$ $`4\left(\left|EE^{}\right|_2+\left|EE^{}\right|_1^\delta +\left|EE^{}\right|_0^{\delta ^2}+\left|VV^{}\right|_2^{\delta ^2}\right).`$ (39) Proof: Part 1 was proven in IIII: equation $`(\text{32})`$ follows directly from $`(\text{13})`$. The bound $`(\text{33})`$ follows from Theorem I.3.5 by summation over $`r\{1,\mathrm{}R\}`$. We reexplain that argument briefly in the proof of Theorem 4 (Section 4.3). We shall shortly prove the remaining statements of part 2 from the more detailed estimates given in Theorem 3. To prove part 3, fix $`R`$, let $`V𝒱`$, and denote for brevity $`K(E)=K^{(R)}(E,\lambda V)`$ and $`B=B_{\mathrm{rad}}`$. Define $`\mathrm{\Phi }:BC_s^2(,)`$ by $$\mathrm{\Phi }(e)=EK(e).$$ (40) By $`(\text{32})`$ and the hypothesis on $`\lambda _R`$, $$|\mathrm{\Phi }(e)E|_{3,r}=|K(e)|_{3,r}D\left|\lambda \right|Q\lambda _R<\mathrm{min}\{\epsilon ,1\}<\epsilon $$ (41) so $`\mathrm{\Phi }(B)B`$. Thus the sequence $`(e_n)_{n0}`$ given by $`e_0=E`$ and $`e_{n+1}=\mathrm{\Phi }(e_n)`$ is well-defined. For $`n1`$, let $`f_n=e_ne_{n1}`$. Then $`f_1=K(E)`$, $`e_n=E+_{k=1}^nf_k`$, and $$f_{n+1}=\mathrm{\Phi }(e_n)\mathrm{\Phi }(e_{n1})=K(e_{n1})K(e_n).$$ (42) Let $`|\lambda |\lambda _R`$. We show that, for all $`n1`$, $`\left|f_n\right|_0`$ $``$ $`(Q\left|\lambda \right|)^n`$ (43) $`\left|f_n\right|_1`$ $``$ $`B_R(\lambda )(Q\left|\lambda \right|)^{n\delta }`$ (44) $`\left|f_n\right|_{3,r}`$ $``$ $`C_R(\lambda )\mathrm{max}\{B_R(\lambda )^\delta ,1\}(Q\left|\lambda \right|)^{n\delta ^2}`$ (45) with $$B_R(\lambda )=\frac{(Q\left|\lambda \right|)^{1\delta }}{1(Q\left|\lambda \right|)^{1\delta }},C_R(\lambda )=\frac{(Q\left|\lambda \right|)^{1\delta ^2}}{1(Q\left|\lambda \right|)^{1\delta ^2}}.$$ (46) Once this is done, $`(\text{45})`$ implies that $`f_n`$ converges in $`||_{3,r}`$. Thus $`e^{(R)}=lim_n\mathrm{}e_n`$ exists. By $`(\text{35})`$, $`\mathrm{\Phi }`$ is continuous in $`||_{3,r}`$, so $`\mathrm{\Phi }(e^{(R)})=e^{(R)}`$ and hence, by $`(\text{40})`$, $`E=e^{(R)}+K(e^{(R)})`$. By $`(\text{41})`$, every $`e_n`$ obeys $`|e_nE|_{3,r}D\left|\lambda \right|`$, so $`e^{(R)}`$ satisfies $`(\text{38})`$. Since $`Q_0\left|\lambda \right|<1`$, uniqueness follows from $`(\text{33})`$. We prove $`(\text{43})`$$`(\text{46})`$ by induction on $`n`$. The statements are true for $`n=1`$ because $$\left|f_1\right|_0\left|f_1\right|_1\left|f_1\right|_{3,r}=\left|K(E)\right|_{3,r}D\left|\lambda \right|Q\left|\lambda \right|$$ (47) and $`(Q\left|\lambda \right|)^\delta B_R(\lambda )`$ $`=`$ $`{\displaystyle \frac{Q\left|\lambda \right|}{1(Q\left|\lambda \right|)^{1\delta }}}>Q\left|\lambda \right|,`$ $`(Q\left|\lambda \right|)^{\delta ^2}C_R(\lambda )`$ $`=`$ $`{\displaystyle \frac{Q\left|\lambda \right|}{1(Q\left|\lambda \right|)^{1\delta ^2}}}>Q\left|\lambda \right|.`$ (48) Assume $`(\text{43})`$$`(\text{46})`$ to hold for $`n`$. By $`(\text{42})`$, $`(\text{33})`$, and the inductive hypothesis $`(\text{43})`$, $$\left|f_{n+1}\right|_0=\left|K(e_n)K(e_{n1})\right|_0Q_0\left|\lambda \right||f_n|_0Q_0\left|\lambda \right|(Q\left|\lambda \right|)^n(Q\left|\lambda \right|)^{n+1}$$ (49) which proves $`(\text{43})`$ for $`n+1`$. By $`(\text{42})`$, $`(\text{34})`$, $`(\text{43})`$ and the inductive hypothesis $`(\text{44})`$, $`\left|f_{n+1}\right|_1`$ $`=`$ $`\left|K(e_n)K(e_{n1})\right|_1`$ (50) $``$ $`Q_0\left|\lambda \right|\left[|f_n|_0^\delta +|f_n|_1\right]`$ $``$ $`Q\left|\lambda \right|\left[(Q\left|\lambda \right|)^{n\delta }+B_R(\lambda )(Q\left|\lambda \right|)^{n\delta }\right]`$ $`=`$ $`\left[(Q\left|\lambda \right|)^{1\delta }+B_R(\lambda )(Q\left|\lambda \right|)^{1\delta }\right](Q\left|\lambda \right|)^{(n+1)\delta }`$ Thus the induction goes through for $`(\text{44})`$ if $$(Q\left|\lambda \right|)^{1\delta }+B_R(\lambda )(Q\left|\lambda \right|)^{1\delta }B_R(\lambda )$$ (51) With the definition $`(\text{46})`$, equality holds in $`(\text{51})`$. By $`(\text{42})`$, $`(\text{33})`$, $`(\text{44})`$ and the inductive hypothesis $`(\text{45})`$, $`\left|f_{n+1}\right|_{3,r}`$ $`=`$ $`\left|K(e_n)K(e_{n1})\right|_{3,r}=\left|K(e_n)K(e_{n1})\right|_2`$ $``$ $`Q_0\left|\lambda \right||f_n|_1^\delta +Q_0\left|\lambda \right||f_n|_2+Q_1(1+G_3)\left|\lambda \right||f_n|_0`$ $``$ $`Q\left|\lambda \right|\left[|f_n|_1^\delta +|f_n|_2\right]`$ $``$ $`Q\left|\lambda \right|\left[B_R(\lambda )^\delta (Q\left|\lambda \right|)^{n\delta ^2}+C_R(\lambda )\mathrm{max}\{B_R(\lambda )^\delta ,1\}(Q\left|\lambda \right|)^{n\delta ^2}\right]`$ $``$ $`\mathrm{max}\{B_R(\lambda )^\delta ,1\}\left[(Q\left|\lambda \right|)^{1\delta ^2}+(Q\left|\lambda \right|)^{1\delta ^2}C_R(\lambda )\right](Q\left|\lambda \right|)^{(n+1)\delta ^2}`$ Here we used that $`e_nB`$ implies $`|e_n|_{3,r}1+|E|_{3,r}1+G_3`$. Thus the induction goes through for $`(\text{45})`$ if $$(Q\left|\lambda \right|)^{1\delta ^2}+(Q\left|\lambda \right|)^{1\delta ^2}C_R(\lambda )C_R(\lambda )$$ (53) With the definition $`(\text{46})`$, equality holds in $`(\text{53})`$. This completes the proof of part 3. We now prove part 4. Denote for brevity $`e=e^{(R)}(E,\lambda V),e^{}=e^{(R)}(E^{},\lambda V^{})`$ and $`K=K^{(R)}`$. First, observe that both $`e,e^{}B_{\mathrm{rad}}B_\epsilon ^{(2)}(E)`$ because, by part 3, $`\left|eE\right|_2`$ $``$ $`D\left|\lambda \right|D\lambda _R/2<\epsilon /2`$ $`\left|e^{}E\right|_2`$ $``$ $`\left|e^{}E^{}\right|_2+\left|EE^{}\right|_2D\lambda _R/2+\epsilon /2<\epsilon ,`$ (54) and because $`\left|eE\right|_{3,r}D|\lambda |<1`$ and $`\left|e^{}E\right|_{3,r}D|\lambda |+\epsilon /2<1`$ hold by $`(\text{38})`$. Thus $`\mathrm{max}\{|e|_{3,r},|e^{}|_{3,r}\}1+G_3`$. By definition, $`e`$ and $`e^{}`$ obey $`E=e+K(e,\lambda V)`$ and $`E^{}=e^{}+K(e^{},\lambda V^{})`$. Hence $`EE^{}`$ $`=`$ $`ee^{}+K(e,\lambda V)K(e^{},\lambda V^{})`$ $`=`$ $`ee^{}+K(e,\lambda V)K(e^{},\lambda V)+K(e^{},\lambda V)K(e^{},\lambda V^{})`$ so that, by $`(\text{33})`$ and $`(\text{36})`$, $$\left|ee^{}\right|_0\left|EE^{}\right|_0+Q_0\left|\lambda \right|\left|ee^{}\right|_0+Q_0\left|\lambda \right|\left|VV^{}\right|_2$$ (56) Recalling that $`Q_0\left|\lambda \right|\frac{1}{2}Q\lambda _R<\frac{1}{2}`$, $$\left|ee^{}\right|_02\left(\left|EE^{}\right|_0+\frac{1}{2}\left|VV^{}\right|_2\right)2\left|EE^{}\right|_0+\left|VV^{}\right|_2$$ (57) Similarly, by $`(\text{34})`$ and $`(\text{36})`$, $$\left|ee^{}\right|_1\left|EE^{}\right|_1+Q_0\left|\lambda \right|\left[\left|ee^{}\right|_0^\delta +\left|ee^{}\right|_1\right]+Q_0\left|\lambda \right|\left|VV^{}\right|_2$$ (58) and $`\left|ee^{}\right|_1`$ $``$ $`2\left(\left|EE^{}\right|_1+\frac{1}{2}\left|ee^{}\right|_0^\delta +\frac{1}{2}\left|VV^{}\right|_2\right)`$ (59) $``$ $`2\left(\left|EE^{}\right|_1+\frac{2^\delta }{2}\left|EE^{}\right|_0^\delta +\frac{1}{2}\left|VV^{}\right|_2^\delta +\frac{1}{2}\left|VV^{}\right|_2\right)`$ $``$ $`2\left(\left|EE^{}\right|_1+\left|EE^{}\right|_0^\delta +\left|VV^{}\right|_2^\delta \right)`$ Similarly, by $`(\text{33})`$ and $`(\text{36})`$, $`\left|ee^{}\right|_2`$ $``$ $`\left|EE^{}\right|_2+Q_0\left|\lambda \right|\left[\left|ee^{}\right|_1^\delta +\left|ee^{}\right|_2\right]`$ (60) $`+`$ $`Q_1\left|\lambda \right|(1+G_3)\left|ee^{}\right|_0+Q_0\left|\lambda \right|\left|VV^{}\right|_2`$ and $`\left|ee^{}\right|_2`$ $``$ $`2\left(\left|EE^{}\right|_2+\frac{1}{2}\left|ee^{}\right|_1^\delta +\frac{1}{2}\left|ee^{}\right|_0+\frac{1}{2}\left|VV^{}\right|_2\right)`$ (61) $``$ $`2\left(\right|EE^{}|_2+\frac{2^\delta }{2}\left[\right|EE^{}|_1^\delta +|EE^{}|_0^{\delta ^2}+|VV^{}|_2^{\delta ^2}]`$ $`+\frac{1}{2}\left[2\right|EE^{}|_0+|VV^{}|_2]+\frac{1}{2}|VV^{}|_2)`$ $``$ $`2\left(\left|EE^{}\right|_2+\left|EE^{}\right|_1^\delta +2\left|EE^{}\right|_0^{\delta ^2}+2\left|VV^{}\right|_2^{\delta ^2}\right)`$ Part 2 of Theorem 2 is proven by a multiscale analysis in which the function $`K^{(R)}(E,\lambda V)`$ is represented as an infinite series $$K^{(R)}(E,\lambda V)=\underset{j<0}{}K_j^{(R)}(E,\lambda V),$$ (62) where, very roughly speaking, $`K_j^{(R)}`$ is the contribution from integrating out those fermions that have an energy in the interval $`[M^{j1},M^j]`$. Here $`M>1`$ and $`j<0`$, so the limit $`j\mathrm{}`$ corresponds to momenta on the Fermi surface. ###### Theorem 3 Let $`\delta _0,g_0,\omega _0>0`$ and $`G_0>\mathrm{max}\{g_0,\omega _0\}`$. Let $`\epsilon >0`$ be as in Lemma 2. Let $`E_0_s(\delta _0,g_0,G_0,\omega _0)C^3`$ and let $$B_{\mathrm{rad}}=\{e_s(\delta _0/2,g_0/2,2G_0,\omega _0/2):\left|eE_0\right|_2<\epsilon ,\left|eE_0\right|_{3,r}<1\}.$$ (63) There is a $`0<\gamma <1`$ such that, for each $`R`$, there is $`Q_2>0`$ ($`Q_2`$ is uniform on $`_s`$!) such that for all $`e_0,e_1B_{\mathrm{rad}}`$ and all $`j<0`$, $`\left|K_j^{(R)}(e)\right|_{3,r}`$ $`=`$ $`\left|K_j^{(R)}(e)\right|_2Q_2\left|\lambda \right|M^{\gamma j}`$ (64) $`\left|K_j^{(R)}(e_1)K_j^{(R)}(e_0)\right|_1`$ $``$ $`Q_2\left|\lambda \right|\left(M^{1.1j}\left|e_1e_0\right|_0+M^{\gamma j}\left|e_1e_0\right|_1\right)`$ (65) and $`\left|K_j^{(R)}(e_1)K_j^{(R)}(e_0)\right|_{3,r}=\left|K_j^{(R)}(e_1)K_j^{(R)}(e_0)\right|_2`$ $``$ $`Q_2\left|\lambda \right|\left(M^{2.1j}\left|e_1e_0\right|_1+M^{\gamma j}\underset{t[0,1]}{sup}\left|e_t\right|_{3,r}\left|e_1e_0\right|_0+M^{\gamma j}\left|e_1e_0\right|_2\right)`$ Moreover, for all $`e_s(\delta _0/2,g_0/2,2G_0,\omega _0/2)`$ and all $`V_1,V_2𝒱`$, $$\left|K_j^{(R)}(e,\lambda V_1)K_j^{(R)}(e,\lambda V_2)\right|_2Q_2\left|\lambda \right|M^{\gamma j}\left|V_1V_2\right|_2$$ (67) The proof of Theorem 3 is given in the next section. The factors $`M^{1.1j}`$ and $`M^{2.1j}`$ come from bounds of the type $`M^j|j|^\alpha \text{ const }(\alpha )M^{1.1j}`$. Proof of parts 1 and 2 of Theorem 2: Eq. $`(\text{64})`$ implies $`(\text{32})`$ when summed over $`j`$, with $`D=Q_2\frac{M^\gamma }{1M^\gamma }`$. Eq. $`(\text{33})`$ was proven in I (Theorem I.3.5). Again by summation, $`(\text{67})`$ implies continuity in the interaction $`V`$. Denote, for brevity, $`K_j(e)=K_j^{(R)}(e,\lambda V)`$. To prove $`(\text{34})`$, with $`\delta =\frac{\gamma }{3}`$, we split the sum over $`j`$ in two parts. If $`j`$ is such that $`|e_1e_0|_0M^{2j}`$, then the inequality $$|e_1e_0|_0\left(M^{2j}\right)^{1\gamma /3}|e_1e_0|_{0}^{}{}_{}{}^{\gamma /3}$$ (68) implies, by $`(\text{65})`$, $`\left|K_j(e_1)K_j(e_0)\right|_1`$ $``$ $`Q_2\left|\lambda \right|\left(M^{(0.92\gamma /3)j}|e_1e_0|_0^{\gamma /3}+M^{\gamma j}\left|e_1e_0\right|_1\right)`$ (69) $``$ $`Q_2\left|\lambda \right|\left(M^{0.2j}|e_1e_0|_0^{\gamma /3}+M^{\gamma j}\left|e_1e_0\right|_1\right)`$ and hence $$\underset{\genfrac{}{}{0pt}{}{j0}{|e_1e_0|_0M^{2j}}}{}\left|K_j(e_1)K_j(e_0)\right|_1Q_3\left|\lambda \right|\left(\left|e_1e_0\right|_0^{\gamma /3}+\left|e_1e_0\right|_1\right)$$ (70) with $`Q_3=Q_2\frac{1}{1M^\gamma ^{}}`$, where $`\gamma ^{}=\mathrm{min}\{0.2,\gamma /3\}`$. If $`j`$ is such that $`|e_1e_0|_0>M^{2j}`$, then $`|e_1e_0|_0^{\gamma /3}M^{2\gamma j/3}`$ and therefore, by $`(\text{64})`$, $$\frac{\left|K_j(e_1)K_j(e_0)\right|_1}{|e_1e_0|_0^{\gamma /3}}2M^{2\gamma j/3}\underset{p=1,2}{\mathrm{max}}\{|K_j(e_p)|_1\}2Q_2\left|\lambda \right|M^{\gamma j/3}$$ (71) so $$\underset{\genfrac{}{}{0pt}{}{j0}{|e_1e_0|_0>M^{2j}}}{}\left|K_j(e_1)K_j(e_0)\right|_12Q_3\left|\lambda \right||e_1e_0|_0^{\gamma /3}.$$ (72) To prove $`(\text{35})`$, with $`\delta =\frac{\gamma }{4}`$, we split the sum over $`j`$ at $`|e_1e_0|_1=M^{3j}`$. This time, writing $`S_3=\underset{t[0,1]}{sup}\left|e_t\right|_{3,r}`$, and using $$|e_1e_0|_1\left(M^{3j}\right)^{(1\gamma /4)}|e_1e_0|_1^{\gamma /4}$$ (73) when $`|e_1e_0|_1M^{3j}`$ gives, by $`(\text{3})`$, for the $`j`$ with $`|e_1e_0|_1M^{3j}`$, $`\left|K_j(e_1)K_j(e_0)\right|_{3,r}`$ (74) $``$ $`Q_2\left|\lambda \right|\left(M^{2.1j}\left|e_1e_0\right|_1+M^{\gamma j}S_3\left|e_1e_0\right|_0+M^{\gamma j}\left|e_1e_0\right|_2\right)`$ $``$ $`Q_2\left|\lambda \right|\left(M^{(0.93\gamma /4)j}|e_1e_0|_1^{\gamma /4}+M^{\gamma j}S_3\left|e_1e_0\right|_0+M^{\gamma j}\left|e_1e_0\right|_2\right)`$ $``$ $`Q_2\left|\lambda \right|\left(M^{0.15j}|e_1e_0|_1^{\gamma /4}+M^{\gamma j}S_3\left|e_1e_0\right|_0+M^{\gamma j}\left|e_1e_0\right|_2\right)`$ so $`{\displaystyle \underset{\genfrac{}{}{0pt}{}{j0}{|e_1e_0|_1M^{3j}}}{}}`$ $`\left|K_j(e_1)K_j(e_0)\right|_{3,r}`$ (75) $``$ $`Q_4\left|\lambda \right|\left(|e_1e_0|_1^{\gamma /4}+S_3\left|e_1e_0\right|_0+\left|e_1e_0\right|_2\right)`$ with $`Q_4=Q_2\frac{1}{1M^\gamma ^{}}`$, where $`\gamma ^{}=\mathrm{min}\{0.15,\gamma /4\}`$. If $`j`$ is such that $`|e_1e_0|_1>M^{3j}`$, then $`|e_1e_0|_{1}^{}{}_{}{}^{\gamma /4}M^{3\gamma j/4}`$ and therefore, by $`(\text{64})`$, $$\frac{\left|K_j(e_1)K_j(e_0)\right|_{3,r}}{|e_1e_0|_1^{\gamma /4}}2M^{3\gamma j/4}\underset{p=1,2}{\mathrm{max}}\{|K_j(e_p)|_{3,r}\}2Q_2\left|\lambda \right|M^{\gamma j/4}$$ (76) so $$\underset{\genfrac{}{}{0pt}{}{j0}{|e_1e_0|_1>M^{3j}}}{}\left|K_j(e_1)K_j(e_0)\right|_{3,r}2Q_4\left|\lambda \right||e_1e_0|_1^{\gamma /4}.$$ (77) ## 4 Bounds with scales – proof of Theorem 3 The counterterm is the localization of a selfenergy function, $$K_j^{(R)}(e,\lambda V,𝐩)=\mathrm{}_eY_j^{(R)}(e,\lambda V,p_0,𝐩).$$ (78) The renormalized tree expansion gives $`Y_j^{(R)}`$ explicitly as $$Y_j^{(R)}(e,\lambda V,p)=\underset{r=1}{\overset{R}{}}\lambda ^r\underset{G}{}\underset{TG}{}\underset{fT}{}\frac{1}{n_f!}\underset{J𝒥(T,j,G)}{}\text{ Val}(G^J)(p)$$ (79) where $`G`$ is summed over all one–particle irreducible (1PI) Feynman graphs with two external legs and $`r`$ interaction vertices. We now briefly describe the genesis of this formula as well as the meaning of $`T`$, $`J`$, $`𝒥(T,j,G)`$ and $`\text{ Val}(G^J)`$. For the details, see, e.g., . The formula is generated by successive applications of renormalization group maps, as follows (for details, see Section 2.3 of I). The covariance corresponding to the quadratic part of the action is expressed as an infinite sum $`C=_{j<0}C_j`$, where the single–scale covariance, $`C_j`$, is supported in the subset of $`\times `$ where $`M^{j2}|\mathrm{i}p_0e(𝐩)|M^j`$ (see Section 2.1 of I). An infrared cutoff $`I<0`$ is introduced by restricting the sum to $`jI`$. Correspondingly, the Gaussian integral with the cutoff covariance is expressed as an $`|I|`$–fold integral $$f(\phi )𝑑\mu _{\mathrm{\Sigma }_{0<jI}C_j}(\phi )=\mathrm{}f\left(_{j=1}^I\phi _j\right)_{j=1}^Id\mu _{C_j}(\phi _j)$$ (80) with respect to the Gaussian measures of covariance $`C_1,\mathrm{},C_I`$. Fields with lower and lower energy scales are integrated out one scale after the other. The Gaussian integral with covariance $`C_j`$ generates an effective interaction on scale $`j`$. The integral kernels of the effective action on scale $`j`$ are given by a sum of values of Feynman graphs whose vertex functions are the integral kernels of the effective action on scale $`j`$ and whose propagators are $`C_j`$. The kernel of the part of the effective interaction on scale $`j`$ that is quadratic in the fields is renormalized by subtracting from it the part of the counterterm whose value is $`\mathrm{}_e`$ applied to the kernel. The renormalized two–legged kernel is called an $`r`$–fork of scale $`j`$. The remaining part of the counterterm is the sum of all $`c`$–forks of scale $`j`$. See Section 2.3 of I. The structure of the iteration is represented by GN (Gallavotti–Nicolò) trees in a natural way. Each graph $`G`$ contributing to the effective interaction at scale $`j`$ has associated to it a GN tree, $`T`$. Each fork, $`f`$, in the tree represents a connected subgraph $`G_f`$ of $`G`$. The subgraph was introduced as a vertex contributing to the effective interaction of some scale $`j_f`$. Hence each fork of $`T`$ carries a label, $`j_f`$, giving its scale and, if $`G_f`$ is two–legged, a label specifying it as an $`r`$–fork or a $`c`$–fork. The fork of $`T`$ corresponding to the entire graph $`G`$ is called the root of $`T`$ and its scale, $`j`$, the root scale of $`T`$. The lines of $`T`$ give the partial ordering of the forks of $`T`$ induced by the partial ordering of subgraphs of $`G`$ by inclusion. If $`\pi (f)`$ is the fork immediately below $`f`$ in the partial ordering of $`T`$, then $`Ij_fj_{\pi (f)}`$ if $`\pi (f)`$ is a c–fork (81) $`1j_f>j_{\pi (f)}`$ otherwise The labelling $`J`$ of $`G`$ assigns a scale $`0<j_lI`$ to every line $`l`$ of $`G`$ and a scale $`0<j_fI`$ to every fork $`f`$ of $`T`$. The set $`𝒥(T,j,G)`$ is the set of labellings determined by the requirements that (a) the root scale is $`j`$, (b) (81) is satisfied and (c) if $`G_f`$ is the smallest of the subgraphs $`G_f^{},f^{}T`$ that contain the line $`l`$, then $`j_l=j_f`$. The value $`\text{ Val}(G^J)(p)`$ of a Feynman graph is the integral over momenta of the integrand which is a product of propagators associated to the lines and vertex functions associated to the vertices (see (I.2.54)). For now, the propagators are given by the covariances $`C_j`$. Later we shall combine strings of two–legged graphs into single lines, and thereby get more general propagators on the lines. For each $`r`$, the coefficient of $`\lambda ^r`$ is a sum of only finitely many terms. Thus most perturbative questions can be reduced to bounding values of individual graphs. In some of our estimates in I, however, we also needed to avoid termwise bounds; this will also play a role in this paper. It was shown in I that under general conditions, the limit $$K^{(R)}(e,\lambda V,𝐩)=\underset{I\mathrm{}}{lim}\underset{Ij<0}{}\mathrm{}_eY_j^{(R)}(e,\lambda V,𝐩)$$ (82) exists and is $`C^1`$ in $`𝐩`$ and Frèchet differentiable in $`e`$. ### 4.1 Proof of $`(\text{64})`$ and $`(\text{67})`$ Eq. $`(\text{64})`$ is just a restatement of (III.3.110) in Theorem III.3.11. Because the function $`\lambda _n(j,\epsilon )`$ in (III.3.110) is bounded by a constant times a power of $`|j|`$ by Lemma I.2.44 $`(v)`$, any $`\gamma <1/3`$ will do. To see $`(\text{67})`$, we note that the value of any graph $`G`$ contributing to $`K_j^{(R)}`$ in $`(\text{79})`$ contains a product of factors $`V`$ associated to the vertices. The localization operator $`\mathrm{}_e`$ does not depend on $`V`$, and the expression $`(\text{79})`$ is linear in $`\text{ Val}(G)`$. Let $`G`$ be a graph contributing to $`K_j^{(R)}`$. By the discrete product rule (II.3.126), the corresponding graph contributing to the difference on the left hand side of $`(\text{67})`$ has a difference $`V_1V_2`$ instead of $`V`$ in one factor. Because all that happens to the vertex functions in the proofs is that they get differentiated (at most twice), and because the estimate is linear in each vertex function, $`(\text{67})`$ follows trivially from the proofs in IIII. ### 4.2 Weaker hypotheses for the proof of $`(\text{33})`$, $`(\text{65})`$, and $`(\text{3})`$ The bounds $`(\text{33})`$, $`(\text{65})`$, and $`(\text{3})`$ hold under much weaker hypotheses than those stated in Theorem 3. In this section, we prove them under hypotheses that are only slightly stronger than those of I. In particular, we shall need neither convexity nor symmetry under $`𝐩𝐩`$ nor the requirement that the Fermi surface be small in the sense that $`S_E_2`$. In fact, it need not even be connected. Let $`𝒩`$ be an open set whose boundary has finitely many connected components, each of which is a $`C^{\mathrm{}}`$ $`(d1)`$–dimensional submanifold of $``$ $`u`$ be a unit $`C^{\mathrm{}}`$ vector field on a neighbourhood of the closure of $`𝒩`$ that is transverse to the boundary of $`𝒩`$ $`e_0,e_1C^0(,)C^2(𝒩,)`$ We assume that there are constants $`\delta _0,u_0,Q_{\mathrm{vol}},\gamma >0`$ such that, for all $`s[0,1]`$, $`e_s=(1s)e_0+se_1`$ has the following properties. The set $`S_{e_s}=\{𝐩:e_s(𝐩)=0\}`$ satisfies $`S_{e_s}𝒩`$ and the distance of $`S_{e_s}`$ to $`𝒩`$ is bounded below by $`\delta _0`$. For all $`𝐩𝒩`$, $$𝒟_ue_s(𝐩)=u(𝐩)e_s(𝐩)>u_0.$$ (83) For $`\epsilon >0`$, let $`𝒰(e,\epsilon )=\{𝐩:|e(𝐩)|\epsilon \}`$. For all $`0<\epsilon _1\epsilon _2\epsilon _3`$, and all $`𝐪`$, $$\underset{𝒰(e_s,\epsilon _1)}{}d𝐩_1\underset{𝒰(e_s,\epsilon _2)}{}d𝐩_2\mathrm{\hspace{0.33em}1}\mathrm{l}\left(|e_s(\pm 𝐩_1\pm 𝐩_2+𝐪)|\epsilon _3\right)Q_{\mathrm{vol}}\epsilon _1\epsilon _2\epsilon _{3}^{}{}_{}{}^{2\gamma }.$$ (84) These hypotheses imply those imposed in I (the volume improvement exponent $`ϵ`$ of I equals $`2\gamma `$), so the results of I apply. Moreover, the stronger hypotheses stated in Section 1.2 imply F1F3 by the following Lemma. ###### Lemma 4 Let $`B=B_\epsilon ^{(2)}(E_0)`$ be the ball of Lemma 2, $`𝒩`$ be the annulus $`\stackrel{~}{A}`$ defined in $`(\text{2.1})`$ and $`u=\widehat{r}`$, the radial vector field of polar coordinates. Then there are constants $`\delta _0,u_0,Q_{\mathrm{vol}},\gamma >0`$ such that F1F3 hold for all $`e_0,e_1B`$. Proof: $`B`$ is convex, so for all $`s[0,1]`$, $`e_s=(1s)e_0+se_1B_s(\delta _0/2,g_0/2,2G_0,\omega _0/2)`$. F1 is obvious by the definition of $`_s`$. F2 follows directly from Lemma 2, with $`u_0=g_1`$. F3 follows from Theorem II.1.1 by the usual Taylor expansion which is described in (I.A.2)–(I.A.6). F2 implies that there is $`g_0>0`$ such that for all $`s[0,1]`$ and all $`𝐩𝒩`$, $`|e_s(𝐩)|>g_0`$. For a fixed $`e`$, the converse is proven in Lemma I.2.1. ###### Lemma 5 Let $`𝒩_c`$ be a connected component of $`𝒩`$ which has a nonempty intersection with $`S_{e_s}`$ for some $`0s1`$. 1. The boundary of $`𝒩_c`$ has precisely two connected components. These two components are diffeomorphic. 2. Denote by $`S`$ one of the two components of the boundary of $`𝒩_c`$. There is, for each $`0s1`$, a $`C^2`$ bijection $`\pi \pi _s`$ from a neighbourhood of $`\{0\}\times S`$ in $`\times S`$ to $`𝒩_c`$ such that $`e_s(\pi \pi _s(\rho ,\theta ))=\rho `$, $`\frac{\pi \pi _s}{\rho }(\rho ,\theta )`$ is parallel to $`u(\pi \pi _s(\rho ,\theta ))`$ and $$\frac{1}{sup_{s,𝐩}|e_s|}\left|\frac{\pi \pi _s}{\rho }\right|\frac{1}{u_0}.$$ (85) Proof: Denote by $`B_1,\mathrm{},B_n`$, the connected components of the boundary of $`𝒩`$. Since $`e_0C^0(,)`$ and $``$ is compact, $`e_0`$ is bounded above and below on $`𝒩`$. By F2, the value of $`e_0`$ changes at a rate of at least $`u_0`$ per unit time along each trajectory of the vector field $`u`$. Hence each trajectory must start on some $`B_i`$ and end on some $`B_j`$. Because $`u`$ is transverse to the boundary of $`𝒩`$ and $`B_i`$ and $`B_j`$ do not themselves have boundaries, each trajectory starting on $`B_i`$ and ending on $`B_j`$ has an open neighbourhood in $`𝒩`$ that is a union of trajectories starting on $`B_i`$ and ending on $`B_j`$. Let, for each $`1i,jn`$, $`𝒩_{i,j}`$ be the set of all points of $`𝒩`$ that lie on a trajectory which starts on $`B_i`$ and ends on $`B_j`$. Then the $`𝒩_{i,j}`$’s are all open and mutually disjoint and their union is $`𝒩`$. Hence each $`𝒩_{i,j}`$ is either empty or a connected component of $`𝒩`$. We claim that if $`𝒩_{i,j}`$ has a nonempty intersection with $`S_{e_s}`$, then $`ij`$. By F1, $`e_s`$ may not vanish in a neighbourhood of the boundary of $`𝒩`$ and hence must be of uniform sign near each $`B_k`$. If $`e_s`$ has the same sign, say positive, near both $`B_i`$ and $`B_j`$ (as will certainly be the case if $`i=j`$) then, as it vanishes somewhere in $`𝒩_{i,j}`$, $`e_s`$ must have a local minimum somewhere in $`𝒩_{i,j}`$. This violates F2. Suppose that $`𝒩_c=𝒩_{i,j}`$. Then $`ij`$ and the components of the boundary of $`𝒩_c`$ are $`B_i`$ and $`B_j`$. The map which associates to each $`𝐩B_i`$ the unique point of $`B_j`$ that is on the same trajectory as $`𝐩`$ is a diffeomorphism, so we have completed the proof of part 1. For each $`𝐩𝒩_c`$, denote by $`\mathrm{\Theta }(𝐩)`$ the unique point of $`B_i`$ that is on the same trajectory of $`u`$ as $`𝐩`$. As $`B_i`$ is a $`C^{\mathrm{}}`$ manifold, $`u`$ is transverse to $`B_i`$ and the trajectories are $`C^{\mathrm{}}`$ in their dependence on time and initial conditions, $`\mathrm{\Theta }(𝐩)`$ is $`C^{\mathrm{}}`$. The map $`𝐩(e_s(𝐩),\mathrm{\Theta }(𝐩))`$ is defined and $`C^2`$ on $`𝒩_c`$, injective (as $`e_s`$ is strictly monotone on each trajectory and each trajectory hits a different point of $`B_i`$) onto a neighbourhood of $`\{0\}\times S`$ (since $`e_s`$ is of opposite sign near $`B_i`$ and $`B_j`$ it must vanish once on each trajectory). Furthermore the Jacobian of this map is nonsingular at each $`𝐩𝒩_c`$ by F2 and the transversality of $`u`$ at $`B_i`$. We may thus take $`\pi \pi _s`$ to be the inverse of this map. Let $`𝐏(e_s,𝐩)=\pi \pi _s(0,\mathrm{\Theta }(𝐩))`$ be the projection on $`S_{e_s}`$, and let $`\mathrm{}_{e_s}`$ denote the localization operator for $`e_s`$, as given by Definition I.2.6. Then $`𝒟_u\mathrm{}_{e_s}=0`$ for all $`s[0,1]`$. Under the hypotheses of Section 1, and if $`u`$ is chosen to be the radial field $`u=\widehat{r}`$, $`𝐏`$ agrees with the projection $`𝐩(r,\theta )𝐩(r_F(e_s,\theta ),\theta )`$ in a neighbourhood of the Fermi surface. We now take a fixed $`V𝒱`$ and prove bounds that are uniform on $`𝒱`$. Thus we again drop the $`\lambda V`$ from the notation. ###### Theorem 4 Under the hypotheses F1–F3, there are constants $`\stackrel{~}{Q}_0`$ and $`\stackrel{~}{Q}_1`$, depending on $`G=sup_s|e_s|_2`$, $`Q_{\mathrm{vol}}`$, $`\gamma `$, $`R`$, $`r_0`$, and $`u_0`$, such that $`\left|K^{(R)}(e_1)K^{(R)}(e_0)\right|_0`$ $``$ $`\stackrel{~}{Q}_0\left|\lambda \right|\left|e_1e_0\right|_0`$ (86) $`\left|K_j^{(R)}(e_1)K_j^{(R)}(e_0)\right|_1`$ $``$ $`\stackrel{~}{Q}_1\left|\lambda \right|\left(M^{1.1j}\left|e_1e_0\right|_0+M^{\gamma j}\left|e_1e_0\right|_1\right).`$ (87) If for all $`s[0,1]`$ the norm $`|e_s|_{3,r}=|e_s|_2+𝒟_ue_s_2`$ is finite, then $`\left|K_j^{(R)}(e_1)K_j^{(R)}(e_0)\right|_{3,r}=\left|K_j^{(R)}(e_1)K_j^{(R)}(e_0)\right|_2`$ $``$ $`\stackrel{~}{Q}_1\left|\lambda \right|\left(M^{2.1j}\left|e_1e_0\right|_1+M^{\gamma j}\underset{t[0,1]}{sup}\left|e_t\right|_{3,r}\left|e_1e_0\right|_0+M^{\gamma j}\left|e_1e_0\right|_2\right).`$ By Lemma 4, Theorem 4 implies $`(\text{33})`$, $`(\text{65})`$, and $`(\text{3})`$, with $`Q_0=\stackrel{~}{Q}_0`$ and $`Q_2=\stackrel{~}{Q}_1`$. ### 4.3 Proof of Theorem 4 Dropping uniform constants in the notation: We introduce the notation $`AB`$ meaning that $`A\text{ const }B`$ where the constant depends only on $`G`$, $`Q_{\mathrm{vol}}`$, $`\gamma `$, $`R`$, $`r_0`$, and $`u_0`$ (thus in particular the constant is uniform on $`_s`$). For instance, we have, for $`p3`$, $`\left|FG\right|_p\left|F\right|_p\left|G\right|_p`$, and $`\left|e\right|_21`$ if $`e_s`$. For a function $`F`$ that depends on $`e`$, let $`D_hF`$ denote the directional derivative of $`F`$ with respect to $`e`$, $`D_hF=\frac{}{\alpha }F(e+\alpha h)_{\alpha =0}`$. We proved in I that $`K`$ is Fréchet differentiable in $`e`$, so these derivatives exist. Moreover, Fréchet differentiability holds for all quantities in which there is an infrared cutoff. #### Proof of $`(\text{86})`$ By $`(\text{79})`$, for any $`s[0,1]`$, $`\left|D_h\left(\mathrm{}_{e_s}{\displaystyle \underset{Ij<0}{}}Y_j^{(R)}(e_s)\right)\right|_0`$ (89) $``$ $`{\displaystyle \underset{r=1}{\overset{R}{}}}|\lambda |^r{\displaystyle \underset{G}{}}\left|{\displaystyle \underset{Ij<0}{}}{\displaystyle \underset{TG}{}}{\displaystyle \underset{fT}{}}{\displaystyle \frac{1}{n_f!}}{\displaystyle \underset{J𝒥(T,j,G)}{}}D_h\left(\mathrm{}_{e_s}\text{ Val}(G^J)(e_s)\right)\right|_0`$ By (I.3.35), there is a constant, depending only on $`G`$ and on the constants given in the Lemma, such that $$\left|D_h\left(\mathrm{}_{e_s}\underset{Ij<0}{}Y_j^{(R)}(e_s)\right)\right|_0\underset{r=1}{\overset{R}{}}|\lambda |^r\underset{G}{}\text{ const }(G)|h|_0.$$ (90) For fixed $`R`$, the sum over graphs $`G`$ contains finitely many terms, so $$\left|D_{e_1e_0}\left(\mathrm{}_{e_s}\underset{Ij<0}{}Y_j^{(R)}(e_s)\right)\right|_0|\lambda ||e_1e_0|_0$$ (91) uniformly in $`I`$ and $`s`$. Thus $`(\text{86})`$ follows by $`{\displaystyle \underset{j<0}{}}(K_j^{(R)}(e_1)K_j^{(R)}(e_0))`$ $`=`$ $`{\displaystyle \underset{0}{\overset{1}{}}}ds{\displaystyle \frac{}{s}}\mathrm{}_{e_s}{\displaystyle \underset{j<0}{}}Y_j^{(R)}(e_s)`$ (92) $`=`$ $`{\displaystyle \underset{0}{\overset{1}{}}}dsD_{e_1e_0}\mathrm{}_{e_s}{\displaystyle \underset{j<0}{}}Y_j^{(R)}(e_s).`$ #### Preliminaries for the proof of $`(\text{87})`$ and $`(\text{4})`$ To prove the single–scale bounds $`(\text{87})`$ and $`(\text{4})`$, we show that for $`k2`$, the seimnorms $`K_j^{(R)}(e_1)K_j^{(R)}(e_0)_k`$ obey bounds with the same right hand side as in $`(\text{87})`$ and $`(\text{4})`$. Note that even the bound for $`k=0`$ does not follow from $`(\text{91})`$ because we are now considering a fixed scale $`j`$, not a sum over scales, and the summation over scales provided a cancellation that was important in the proof of Theorem I.3.5. However, the proof does not require very detailed estimates because the coefficient of $`|e_1e_0|_{k1}`$ is (up to factors $`|j|`$, which we bound by $`M^{0.1j}`$) a factor $`M^{kj}`$ larger than the undifferentiated power counting behaviour $`M^j`$ of a single–scale selfenergy contribution like $`Y_j^{(R)}`$. This is naive power counting behaviour. The estimates will again follow by applying bounds already proven in I. We now interpolate the difference of the two $`K`$ functions. The derivative of $`\mathrm{}_e`$ with respect to $`e`$ was calculated in Lemma I.3.1. The interpolation gives $$K_j^{(R)}(e_1)K_j^{(R)}(e_0)=\underset{0}{\overset{1}{}}ds\left(𝒴_1(s)𝒴_2(s)\right)$$ (93) with $`𝒴_1(s)`$ $`=`$ $`\mathrm{}_{e_s}\left(D_{e_1e_0}Y_j^{(R)}(e_s)\right)`$ (94) $`𝒴_2(s)`$ $`=`$ $`\mathrm{}_{e_s}\left[(e_1e_0){\displaystyle \frac{1}{𝒟_ue_s}}𝒟_uY_j^{(R)}(e_s)\right],`$ (95) with $`𝒟_u`$ defined in $`(\text{83})`$. Because $`K_j`$ is the localization of $`Y_j`$, $`𝒟_uK_j=0`$, so the first equality in $`(\text{4})`$ holds. Thus we have to bound $`𝒴_i_k`$ for $`k\{0,1,2\}`$. In the following, we drop the superscript $`R`$ from $`Y_j^{(R)}`$. #### Estimates for $`𝒴_2_k`$ Let $`k=0`$. The bound $`|𝒴_2|_0\left|e_1e_0\right|_0\frac{1}{u_0}|𝒟_uY_j|_0`$ and Theorem I.2.46 (i) imply that $$\left|𝒴_2\right|_0|j|^RM^{2\gamma j}\left|e_1e_0\right|_0M^{\gamma j}\left|e_1e_0\right|_0.$$ (96) Let $`k=1`$. Because $$\frac{}{p_\alpha }(\mathrm{}_{e_s}F)(p)=\frac{}{p_\alpha }F(0,𝐏(e_s,𝐩))=\underset{\beta }{}\frac{𝐏_\beta }{p_\alpha }(e_s,𝐩)\left[\frac{}{q_\beta }F(0,𝐪)\right]_{𝐪=𝐏(e_s,𝐩)},$$ (97) we have $$\frac{}{p_\alpha }𝒴_2(s)(p)=\underset{\beta }{}\frac{𝐏_\beta }{p_\alpha }(e_s,𝐩)𝒳_\beta (𝐏(e_s,𝐩))$$ (98) with $$𝒳_\beta (𝐪)=\frac{}{q_\beta }\left[(e_1e_0)(𝐪)\frac{1}{𝒟_ue_s(𝐪)}𝒟_uY_j(0,𝐪)\right].$$ (99) Thus $`𝒴_2(s)_1`$ $``$ $`d𝐏(e_s)_1(e_1e_0_1\frac{1}{u_0}|𝒟_uY_j|_0`$ (100) $`+\left|e_1e_0\right|_0\frac{1}{u_0^2}𝒟_ue_s_1\left|𝒟_uY_j\right|_0`$ $`+|e_1e_0|_0\frac{1}{u_0}𝒟_uY_j_1)`$ $``$ $`\left|e_1e_0\right|_1\left|𝒟_uY_j\right|_0+\left|e_1e_0\right|_0𝒟_uY_j_1`$ because $`𝐏(e_s)_11`$ and $`𝒟_ue_s_1\left|e_s\right|_21`$. Let $`k=2`$. Because $`\frac{^2}{p_\gamma p_\alpha }𝒴_2(p)`$ $`=`$ $`{\displaystyle \underset{\beta }{}}[𝒳_\beta (𝐏(e_s,𝐩))\frac{^2𝐏_\beta }{𝐩_\gamma 𝐩_\alpha }(e_s,𝐩)`$ (101) $`+`$ $`{\displaystyle \underset{\rho }{}}(_\rho 𝒳_\beta )(𝐏(e_s,𝐩))_\gamma 𝐏_\rho (e_s,𝐩)_\alpha 𝐏_\beta (e_s,𝐩)],`$ we have $$𝒴_2(s)_2d𝐏(e_s)_2\underset{\beta }{\mathrm{max}}\left|𝒳_\beta \right|_0+𝐏(e_s)_{1}^{}{}_{}{}^{2}\underset{\beta ,\rho }{}\left|_\rho 𝒳_\beta \right|_0.$$ (102) Because $`𝐏(e_s)_21`$ and $`\left|\frac{}{q_\rho }𝒳_\beta (𝐪)\right|`$ $``$ $`\left|e_1e_0\right|_2\left|𝒟_uY_j\right|_0+\left|e_1e_0\right|_0\left|𝒟_uY_j\right|_0𝒟_ue_s_2`$ (103) $`+`$ $`\left|e_1e_0\right|_1𝒟_uY_j_1+\left|e_1e_0\right|_0𝒟_uY_j_2,`$ we have $`𝒴_2(s)_2`$ $``$ $`\left|e_1e_0\right|_2\left|𝒟_uY_j\right|_0+\left|e_1e_0\right|_1𝒟_uY_j_1`$ (104) $`+`$ $`\left|e_1e_0\right|_0\left(\left|𝒟_uY_j\right|_0𝒟_ue_s_2+𝒟_uY_j_2\right).`$ The term $`𝒟_ue_s_2`$ is the reason why we have to deal with functions that have bounded radial derivatives. Because it arises only from the derivative of the localization operator, it has got nothing to do with the scale dependence of $`Y_j`$. By $`(\text{79})`$, it suffices to bound the contribution from every 1PI two–legged graph $`G`$ separately. That is, we may replace $`Y_j`$ by $`W=_{J𝒥(T,j,g)}\text{ Val}(G^J)`$ in $`(\text{96})`$, $`(\text{100})`$, and $`(\text{104})`$ if we take a maximum over $`G`$ and $`T`$ and multiply by the number of graphs and the number of possible $`T`$’s. By Theorem I.2.46 (i), and using $`\lambda _n(j,\gamma )M^{\gamma j}1`$, we have $`𝒴_2(s)_1`$ $``$ $`\left|e_1e_0\right|_1M^{\gamma j}+\left|e_1e_0\right|_0M^{(\gamma 1)j}`$ (105) $`𝒴_2(s)_2`$ $``$ $`\left|e_1e_0\right|_2M^{\gamma j}+\left|e_1e_0\right|_1M^{(\gamma 1)j}`$ (106) $`+`$ $`\left|e_1e_0\right|_0\left(M^{\gamma j}𝒟_ue_s_2+𝒟_uY_j_2\right).`$ Thus $`sup_s𝒴_2(s)_k`$ obey bounds that imply $`(\text{87})`$ and $`(\text{4})`$ if we can prove that $$𝒟_uY_j_2M^{2.1j}$$ (107) and that $$𝒴_1(s)_1M^{1.1j}\left|e_1e_0\right|_0,𝒴_1(s)_2M^{2.1j}\left|e_1e_0\right|_0.$$ (108) To do this, we need to exhibit the structure of the graphs $`G`$ that contribute to $`Y_j`$ in a little bit more detail. #### Graphical tools Let $`G`$ be a graph contributing to $`(\text{79})`$, $`T`$ a rooted tree compatible to $`G`$, with an $`r`$ and $`c`$ labelling assigned to the forks, and $`𝒥(T,j,G)`$ the set of labellings of $`G`$ compatible with $`T`$ and root scale $`j`$. Let $`\varphi `$ be the root of $`T`$. To every fork $`fT`$ there corresponds a connected subgraph $`G_f`$ of $`G`$, which is a proper subgraph of $`G`$ for $`f>\varphi `$. We call $`f`$ an $`m`$–legged fork if $`G_f`$ has $`m`$ external legs. In the following we construct a graph $`\mathrm{\Gamma }`$, a tree $`T^{}`$ compatible with $`\mathrm{\Gamma }`$, and a set of labellings $`𝒥^{}`$ with the following properties. * $`\mathrm{\Gamma }`$ is two–legged and 1PI, and $`\mathrm{\Gamma }`$ has only four–legged vertices with vertex functions $`\widehat{v}`$. * The associated tree $`T^{}`$ has no 2–legged forks. * The scale assignments in $`𝒥^{}`$ are $`j_f>j_{\pi (f)}`$ for all $`fT^{}`$. With propagators associated to $`\mathrm{\Gamma }`$ in the way given below, $$\underset{J𝒥(T,j,G)}{}\text{ Val}(G^J)=\underset{J^{}𝒥^{}(T^{},j,\mathrm{\Gamma })}{}\text{ Val}(\mathrm{\Gamma }^J^{})$$ (109) Summation over the trees gives $$\underset{TG}{}\underset{fT}{}\frac{1}{n_f!}\underset{J𝒥(T,j,G)}{}\text{ Val}(G^J)=\underset{T^{}\mathrm{\Gamma }}{}\underset{f^{}T^{}}{}\frac{1}{n_f^{}!}\underset{J^{}𝒥^{}(T^{},j,\mathrm{\Gamma })}{}\text{ Val}(\mathrm{\Gamma }^J^{}).$$ (110) This construction is similar to that of Remark I.2.45, only simpler, because here we do not aim at tight bounds for the powers of $`|j|`$ generated by scale sums of four–legged subdiagrams. If no $`f>\varphi `$ is two–legged, then $`\mathrm{\Gamma }=G`$, $`T^{}=T`$, $`𝒥^{}=𝒥`$. Otherwise, let $`f_1,\mathrm{}f_n>\varphi `$ be all minimal two–legged forks of $`T`$. That is, there is no two–legged fork $`f^{}`$ with $`\varphi <f^{}<f_i`$. Let $`\stackrel{~}{T}`$ be the tree where the subtrees $`T_i`$ rooted at the forks $`f_i`$ are replaced by leaves $`\lambda _i`$. To obtain the corresponding graph $`\stackrel{~}{G}`$, replace $`G_{f_i}`$ by a two–legged vertex $`v_i`$ with ($`j_{\pi (f_i)}`$–dependent) vertex function $$A_i=𝒫_i\underset{j_i}{}\underset{J_i𝒥(T_i,j_{\pi (f_i)},G_{f_i})}{}\text{ Val}(G_{f_i}^{J_i}).$$ (111) The projection $`𝒫_i`$ is $`\mathrm{}_{e_s}`$ if $`f_i`$ is a $`c`$–fork and $`1\mathrm{}_{e_s}`$ if $`F_i`$ is an $`r`$–fork of $`T`$. The summation range is $`j_i>j_{\pi (f_i)}`$ if $`f_i`$ is an $`r`$–fork and $`j_ij_{\pi (f_i)}`$ if $`f_i`$ is a $`c`$–fork. Because all $`c`$–forks have now been replaced by vertices (or hidden inside two–legged vertices), $`\stackrel{~}{𝒥}=\{J|_{\stackrel{~}{T}}:J𝒥(T,j,G)\}`$ consists only of labellings with $`j_f>j_{\pi (f)}`$ for all $`f\stackrel{~}{T}`$. With the standard definition of the value of a labelled graph (see, e.g., (I.2.54)), $$\underset{J𝒥(T,j,G)}{}\text{ Val}(G^J)=\underset{\stackrel{~}{J}\stackrel{~}{𝒥}(\stackrel{~}{T},j,\stackrel{~}{G})}{}\text{ Val}\left(\stackrel{~}{G}^{\stackrel{~}{J}}\right).$$ (112) The graph $`\stackrel{~}{G}`$ is not yet what we want because the graph $`G_{f_i}`$ whose value appears in $`(\text{111})`$ is not necessarily 1PI and because $`\stackrel{~}{G}`$ may contain two–legged vertices. In order to apply Theorem I.2.46, we want to reduce all vertex functions of two–legged vertices to sums over values of 1PI graphs. If $`f_i`$ is a $`c`$–fork, $`G_{f_i}`$ is 1PI because otherwise $`\mathrm{}_{e_s}`$ of its value would vanish. If $`f_i`$ is an $`r`$–fork, $`G_{f_i}`$ may be 1PR; then $`𝒫_i\text{ Val}(G_{f_i}^J)=\text{ Val}(G_{f_i}^J)`$ and it is a string of two–legged subgraphs, some of which may be single–scale insertions (SSI’s) defined in Remark I.2.45. Momentum conservation, the scale structure on $`T`$, and the support properties of the cutoff function fix the scale of the lines connecting the 1PI pieces to $`j_{\pi (f_i)}+1`$. When every $`r`$–fork corresponding to an 1PR graph is replaced by its string as above, the only changes to $`\stackrel{~}{G}`$ are that additional two–legged vertices may appear and that, besides the cases $`𝒫_i=\mathrm{}_{e_s},1\mathrm{}_{e_s}`$, there is the third case $`𝒫_i=1`$ for SSI’s, with the scale sum for a SSI consisting only of the one term where all scales are $`j_{\pi (f_i)}+1`$ (see Remark I.2.45 for details). Let $`\mathrm{\Gamma }`$ be the graph where all strings of two–legged subgraphs are replaced by single lines, and $`T^{}`$ be the tree in which all leaves of $`\stackrel{~}{T}`$ that correspond to two–legged vertices of $`\stackrel{~}{G}`$ are removed. For a line $`\mathrm{}`$ of $`\mathrm{\Gamma }`$, let $`j_{\mathrm{}}`$ be the minimum over all $`j_\stackrel{~}{\mathrm{}}`$, where $`\stackrel{~}{\mathrm{}}`$ runs over the lines of $`\stackrel{~}{G}`$ on the string $`\sigma _{\mathrm{}}`$ in $`\stackrel{~}{G}`$ replaced by $`\mathrm{}`$. The propagator associated to $`\mathrm{}`$ is $$S_{\mathrm{},j_{\mathrm{}}}(p)=\underset{(j_\stackrel{~}{\mathrm{}})_{\stackrel{~}{\mathrm{}}\mathrm{on}\sigma _{\mathrm{}}}}{}\underset{\stackrel{~}{\mathrm{}}\mathrm{on}\sigma _{\mathrm{}}}{}C_{j_\stackrel{~}{\mathrm{}}}(p)\underset{v\mathrm{on}\sigma _{\mathrm{}}}{}A_v$$ (113) where the summation is over all scale assignments $`j_\stackrel{~}{\mathrm{}}\{j_{\mathrm{}},j_{\mathrm{}}+1\}`$ that are compatible with $`\stackrel{~}{T}`$, and, if $`n`$ propagators appear in the product, $`n1`$ factors $`A_v`$ appear. By construction, $`(\text{109})`$ and $`(\text{110})`$ hold. ###### Lemma 6 Let $`\alpha `$ be a multiindex with $`w=|\alpha |1`$. Then the propagators $`S_{\mathrm{},j_{\mathrm{}}}`$ given by $`(\text{113})`$ satisfy $$\left|D^\alpha S_{\mathrm{},j_{\mathrm{}}}(p)\right|M^{j_{\mathrm{}}(1+w)+j_{\mathrm{}}\gamma g}\mathrm{\hspace{0.33em}1}\mathrm{l}\left(|\mathrm{i}p_0e_s(𝐩)|M^j_{\mathrm{}}\right)$$ (114) where $`g`$ is the number of $`c`$–forks plus the number of SSI on the string $`\sigma _{\mathrm{}}`$ corresponding to $`S_{\mathrm{},j_{\mathrm{}}}`$. Proof: The support condition follows directly from that of $`C_j_{\mathrm{}}`$. We now bound the functions $`A_v`$ and their first derivatives. This is a direct application of Theorem I.2.46 (i), which states (with $`\epsilon =2\gamma `$) that if $`G`$ is two–legged and 1PI, then for all $`r\{0,1,2\}`$, $$\underset{J𝒥(T,j,G)}{}\left|\text{ Val}G^J\right|_r|j|^{n_G}M^{j(1+2\gamma r)}M^{j(1+\gamma r)}.$$ (115) Let $`w\{0,1\}`$ and $`\alpha `$ be a multiindex with $`|\alpha |=w`$. For $`v`$ corresponding to an $`r`$–fork and for $`p`$ such that $`\left|\mathrm{i}p_0e_s(𝐩)\right|M^j_{\mathrm{}}`$, $$\left|D^\alpha A_v(p)\right|\underset{j>j_{\mathrm{}}}{}\left|\underset{J𝒥(T_i,j,G_{f_i})}{}D^\alpha (1\mathrm{}_{e_s})\text{ Val}(G_{f_i}^J)(p)\right|$$ (116) For $`w=0`$, Taylor expansion gives the renormalization gain $`M^j_{\mathrm{}}`$ and one derivative acting on $`\text{ Val}(G_{f_i}^J)`$. By $`(\text{115})`$, with $`r=1+w=1`$, $$\left|D^\alpha A_v(p)\right|M^j_{\mathrm{}}\underset{j>j_{\mathrm{}}}{}M^{j(1+\gamma 1)}M^j_{\mathrm{}}.$$ (117) For $`w=1`$, we estimate the $`1`$ and $`\mathrm{}_{e_s}`$ terms separately. By $`(\text{115})`$, $$\left|D^\alpha A_v(p)\right|2\underset{j>j_{\mathrm{}}}{}M^{j(1+\gamma 1)}1.$$ (118) For $`v`$ corresponding to a $`c`$–fork, $$\left|A_v\right|_w\underset{jj_{\mathrm{}}}{}\left|\underset{J𝒥(T_i,j,G_{f_i})}{}\mathrm{}_{e_s}\text{ Val}(G_{f_i}^J)\right|_w,$$ (119) so $`(\text{115})`$ implies $$\left|A_v\right|_w\underset{jj_{\mathrm{}}}{}M^{j(1+\gamma w)}M^{j_{\mathrm{}}(1+\gamma w)}.$$ (120) The estimate for $`v`$ corresponding to an SSI is similar to that of a $`c`$–fork, except that there is not even a scale sum to do because the scales are all fixed in an SSI. Using the product rule for derivatives acting on $`(\text{113})`$ and using that $$\left|D^\alpha C_j(p)\right|M^{j(1+|\alpha |)}\mathrm{\hspace{0.33em}1}\mathrm{l}\left(|\mathrm{i}p_0e_s(𝐩)|M^j\right)$$ (121) we get the statement of the Lemma. Lemma 6 gives us control over first order derivatives of the propagators $`S_{\mathrm{},j_{\mathrm{}}}`$ with respect to momentum. The next lemma will imply that we can always arrange the integral for the value of a graph contributing to $`Y_j`$ such that every line of the graph gets differentiated at most once, even if we take three derivatives with respect to the external momentum. In I, Definition 2.19, we introduced the notion of overlapping graphs. A graph is overlapping if there is a line $`\mathrm{}`$ of $`G`$ which is part of two independent (non self–intersecting) loops. We say that the line $`\mathrm{}`$ is part of the two overlapping loops. ###### Lemma 7 Let $`G`$ be a two–legged 1PI graph with two external vertices $`v_1`$ and $`v_2`$. Let all vertices of $`G`$ have an even incidence number. Let $`T`$ be any spanning tree of $`G`$, and let $`\theta `$ be the linear subtree of $`T`$ corresponding to the unique path from $`v_1`$ to $`v_2`$ over lines of $`T`$. Then every line $`\mathrm{}\theta `$ is part of two overlapping loops generated by lines $`\mathrm{}_1T`$ and $`\mathrm{}_2T`$. For $`i\{1,2\}`$, the graph $`T_i`$, obtained from $`T`$ by removing $`\mathrm{}`$ and adding $`\mathrm{}_i`$, is a spanning tree for $`G`$. Proof: Let $`\mathrm{}`$ be a line of $`\theta `$. Cut $`\mathrm{}`$ to get a four–legged graph $`F=G\mathrm{}`$. Because $`G`$ is 1PI, $`F`$ is connected, so there is a (nonselfintersecting) path $`\pi `$ in $`F`$ that joins the endpoints of $`\mathrm{}`$. Because $`T`$ is a tree, $`T\mathrm{}`$ has two connected components, $`T_1`$ and $`T_2`$. As $`T_1T_2\pi `$ is connected, one of the lines on $`\pi `$, say $`\mathrm{}_1`$, joins $`T_1`$ and $`T_2`$, but is not in $`T`$. Thus $`\mathrm{}`$ is on the loop generated by $`\mathrm{}_1`$. Go back to $`G`$ and cut $`\mathrm{}_1`$. The result is a four–legged graph $`F^{}=G\mathrm{}_1`$. Because $`\mathrm{}_1T`$, $`T`$ is still a spanning tree for $`F^{}`$. Cutting $`\mathrm{}`$ does not disconnect $`F^{}`$ because if it did, each of the connected components would have to have three external lines – one of $`G`$’s original external lines, one end of $`\mathrm{}_1`$ and one end of $`\mathrm{}`$ (as all vertices of $`G`$ have even incidence number, all connected graphs must have an even number of external lines). Let $`\mathrm{}_2`$ be a line on the shortest path in $`F^{}\mathrm{}`$ connecting the endpoints of $`\mathrm{}`$ with $`\mathrm{}_2`$ joining $`T_1`$ and $`T_2`$ but not in $`T`$. Then $`\mathrm{}`$ is in the loop generated by $`\mathrm{}_2`$. Thus the loops generated by $`\mathrm{}_1`$ and $`\mathrm{}_2`$ overlap on $`\mathrm{}`$. It would not have been a loss of generality to assume that $`G`$ has no proper two–legged subgraphs. In that case, Remark I.2.23 implies that $`F^{}`$ is also 1PI. If $`T`$ is chosen such that $`\theta `$ is a shortest path from $`v_1`$ to $`v_2`$ in $`G`$, the statement of the Lemma is an obvious consequence of Lemma III.2.5 (see Figures III.2.3–III.2.6; note that the lines from $`v_r`$ to $`v_{r+1}`$ and from $`v_s`$ to $`v_{s+1}`$ can be any pair of lines on $`\theta `$). #### The bound for $`𝒟_uY_j_2`$ Because $`𝒟_uY_j_2\left|Y_j\right|_3`$, it suffices to prove that $$\left|Y_j\right|_3M^{2.1j}.$$ (122) By $`(\text{79})`$, it suffices to prove the same bound for $$𝒲=\underset{J𝒥(T,j,G)}{}\text{ Val}(G^J).$$ (123) All graphs that contribute are two–legged and 1PI, so by $`(\text{115})`$, $$\left|𝒲\right|_2M^{j(1+2\gamma 2)}(1+|j|^R)M^jM^{\gamma j}M^{2j},$$ (124) so it suffices to bound $`𝒲_3`$. Let $`\mathrm{\Gamma }`$ be the graph associated to $`G`$ with the properties $`(\text{109})`$ and $`(\text{110})`$, then $$𝒲_3\underset{J𝒥(T^{},j,\mathrm{\Gamma })}{}\text{ Val}(\mathrm{\Gamma }^J)_3.$$ (125) Let $`T`$ be a spanning tree for $`\mathrm{\Gamma }`$. The only factors in the integrand for $`\text{ Val}\mathrm{\Gamma }^J`$ that can depend on the external momentum $`q`$ are * vertex functions $`\widehat{v}`$; the dependence is of the form $`\widehat{v}(qp)`$ where $`p`$ is a loop momentum or a sum of loop momenta because $`G`$ is 1PI and two–legged (it can happen that $`\widehat{v}`$ does not depend on any loop momentum; this is, however, only the case for tadpoles, in which case only $`\widehat{v}(0)`$ appears). * propagators $`S_{\mathrm{},j_{\mathrm{}}}`$ for those $`\mathrm{}`$ that are in the path on $`T`$ connecting the external vertices (if there is only one external vertex, no propagator depends on $`q`$). We now take three derivatives of $`\text{ Val}(\mathrm{\Gamma }^J)`$ and use the above lemmas to avoid having two derivatives acting on any propagator and three on any vertex function, as follows. If $`\mathrm{\Gamma }`$ has only one external vertex and is not a tadpole, we first route $`q`$ through the $`\widehat{v}`$ of the external vertex. We let two derivatives act and then change variables from $`p`$ to $`qp`$ in the loop integral in which $`\widehat{v}(qp)`$ appears. The third derivative can then not act on this vertex function any more. It can act on another vertex function or on a propagator. If $`\mathrm{\Gamma }`$ has two external vertices, there are two cases, depending on where the first derivative acted. 1. The first derivative acts on a vertex function. Take another derivative. If it acts on the same vertex function, change variables from $`p`$ to $`qp`$ in the loop integral in which $`\widehat{v}(qp)`$ appears. The third derivative can then not act on this vertex function any more. If the second derivative acts on the propagator $`S_{\mathrm{},j_{\mathrm{}}}`$, we change the spanning tree using Lemma 7. The third derivative can then not act on $`S_{\mathrm{},j_{\mathrm{}}}`$ any more. 2. The first derivative acts on the propagator $`S_{\mathrm{},j_{\mathrm{}}}`$. We change the spanning tree to $`T_1`$ by replacing $`\mathrm{}`$ with another line $`\mathrm{}_1`$ (this is possible by Lemma 7) and take another derivative. It can act on a propagator on a line $`\mathrm{}^{}`$ on the path in $`T_1`$ that connects the external vertices ($`\mathrm{}^{}=\mathrm{}_1`$ is possible). By Lemma 7, there are two lines, $`\mathrm{}_1^{}`$ and $`\mathrm{}_2^{}`$, such that for $`i\{1,2\}`$, $`T_i^{}`$, obtained by replacing $`\mathrm{}^{}`$ by $`\mathrm{}_i^{}`$ in $`T_1`$, is still a spanning tree for $`\mathrm{\Gamma }`$. At most one of $`\mathrm{}_1^{}`$ and $`\mathrm{}_2^{}`$ may be $`\mathrm{}`$, so we may change to a spanning tree that contains neither $`\mathrm{}`$ nor $`\mathrm{}^{}`$. Once this is done, the third derivative cannot act on the propagators associated to the lines $`\mathrm{}`$ and $`\mathrm{}^{}`$. In summary, the net effect of taking three derivatives in the way just described is, by Lemma 6, at most a factor $`M^{3j}`$, as compared to standard power counting (a factor $`M^{3j}`$ arises only if all three derivatives act on propagators; when vertex functions get differentiated, no factor $`M^j`$ is produced). Because the GN tree $`T^{}`$ associated to $`\mathrm{\Gamma }`$ has no 2–legged forks, the scale sum converges by standard arguments (see Lemma I.2.4 and Remark I.2.5), and is bounded by $`|j|^RM^j`$. Thus $$𝒲_3|j|^RM^jM^{3j}M^{2.1j}.$$ (126) #### The bound for $`𝒴_1(s)_2`$ In the following bounds we keep the tree sums inside of the norms. By $`(\text{110})`$, we thus need to estimate $$\underset{T^{}\mathrm{\Gamma }}{}\underset{fT^{}}{}\frac{1}{n_f!}\underset{J𝒥(T^{},j,\mathrm{\Gamma }^{})}{}D_h\text{ Val}(\mathrm{\Gamma }^J)_k$$ (127) for $`k\{0,1,2\}`$, with $`h=e_1e_0`$. By construction of $`\mathrm{\Gamma }`$, $`D_h`$ acts only on the propagators $`S_{\mathrm{},j_{\mathrm{}}}`$. ###### Lemma 8 For all $`s[0,1]`$ and all lines $`\mathrm{}`$ of $`\mathrm{\Gamma }`$ $$\left|D_hS_{\mathrm{},j_{\mathrm{}}}(p)\right||h|_0M^{2j_{\mathrm{}}}\mathrm{\hspace{0.33em}1}\mathrm{l}\left(|ip_0e_s(𝐩)|M^j_{\mathrm{}}\right).$$ (128) Proof: By definition $`(\text{113})`$, $`D_h`$ can act on factors (a) $`C_j`$, (b) $`A_v`$ coming from an $`r`$–fork, (c) $`A_v`$ coming from a $`c`$–fork, (d) $`A_v`$ coming from an SSI. In the last three cases, by $`(\text{111})`$, we have to estimate the norms of $$\stackrel{~}{𝒲}_i=𝒫_i\underset{j_i}{}\underset{T_iG_{f_i}}{}\underset{fT_i}{}\frac{1}{n_f!}\underset{J_i𝒥(T_i,j_{\pi (f_i)},G_{f_i})}{}\text{ Val}(G_{f_i}^{J_i}).$$ (129) by (I.3.44), $$\left|D_hC_{j_\stackrel{~}{\mathrm{}}}(p)\right||h|_0M^{2j_{\mathrm{}}}\mathrm{\hspace{0.33em}1}\mathrm{l}\left(|ip_0e_s(𝐩)|M^{j_\stackrel{~}{\mathrm{}}}\right).$$ (130) By Lemma I.3.1, $$D_h(\mathrm{}_{e_s}𝒲_i)(e_s)=\mathrm{}_{e_s}(D_h𝒲_i)\mathrm{}_{e_s}\left(\frac{h}{𝒟_ue_s}𝒟_u𝒲_i\right)(e_s)$$ (131) so $$D_h(1\mathrm{}_{e_s})𝒲_i=(1\mathrm{}_{e_s})D_h𝒲_i+\mathrm{}_{e_s}\left(\frac{h}{𝒟_ue_s}𝒟_u𝒲_i\right).$$ (132) If $`p`$ is such that $`|ip_0e_s(𝐩)|M^j_{\mathrm{}}`$, then by Taylor expansion $$\left|(1\mathrm{}_{e_s})D_h𝒲_i(p)\right|M^j_{\mathrm{}}\left|D_h𝒲_i\right|_1.$$ (133) By (I.3.42), this is $$\left|h\right|_0M^j_{\mathrm{}}\underset{j>j_{\mathrm{}}}{}M^{j(2\gamma 1)}|j_{\mathrm{}}|^{R_i}\left|h\right|_0|j_{\mathrm{}}|^{R_i}M^{2\gamma j_{\mathrm{}}}\left|h\right|_0$$ (134) with $`R_i`$ the number of vertices of $`G_{f_i}`$. The second term in $`(\text{132})`$ is bounded by $$\left|\frac{h}{𝒟_ue_s}𝒟_u𝒲_i\right|_0\left|h\right|_0\left|𝒟_u𝒲_i\right|_0\left|h\right|_0\left|𝒲_i\right|_1\left|h\right|_0$$ (135) (in the last step, we used $`(\text{115})`$). Eq. (I.3.41) (with depth $`PR`$) implies that $$\left|D_hA_v\right|_0\left|h\right|_0.$$ (136) Eq. (I.3.42) again implies $`(\text{136})`$. Thus in all cases, the derivative produces at most an additional factor $`M^j_{\mathrm{}}`$ in the bounds. Applying $`(\text{117})`$, $`(\text{120})`$ with $`w=0`$, and $`(\text{121})`$ with $`|\alpha |=0`$, counting up factors $`M^j_{\mathrm{}}`$, now implies the bound. Thus the effect of a derivative with respect to the dispersion relation acting on the propagator $`S_{\mathrm{},j_{\mathrm{}}}`$ can be bounded in exactly the same way as a derivative with respect to momentum (see Lemma 6), except that $`\gamma `$ (which was never actually used) has been replaced by zero. By Lemma 7 we can again prevent the at most two derivatives that appear in the norms from acting on $`D_hS_{\mathrm{},j_{\mathrm{}}}`$. Thus, repeating the argument from $`(\text{125})`$ to $`(\text{126})`$, again using Lemma I.2.4 and Remark I.2.5, and using $`|j|^RM^{0.1j}`$, we have $`𝒴_1(s)_1`$ $``$ $`M^{1.1j}\left|e_1e_0\right|_0,`$ (137) $`𝒴_1(s)_2`$ $``$ $`M^{2.1j}\left|e_1e_0\right|_0.`$ (138) Summing the seminorms $`_k`$, we get $`(\text{87})`$ and $`(\text{4})`$. ## 5 Discussion In this section, we briefly discuss the role of the various hypotheses we used in our proofs, to summarize which parts of our argument extend easily to general Fermi surface geometries and where more work is needed. We also discuss the role of the symmetry condition $`e(𝐩)=e(𝐩)`$ because cases where this symmetry does not hold are interesting from a physical point of view. The two main ingredients for the iteration by which we construct the solution to $`(\text{11})`$ are 1. the existence of an invariant set for the map $`ee+K^{(R)}`$, 2. the contraction–like bounds $`(\text{33})`$, $`(\text{34})`$, and $`(\text{35})`$. To prove item 2, we needed only rather weak hypotheses on the Fermi surface geometry. In particular, we neither used a symmetry $`e(𝐩)=e(𝐩)`$ in that part of the proof, nor any assumption about strict convexity, nor that $`S_e_2`$. With a different localization operator, defined as in , one can even drop F3 in the proof of $`(\text{87})`$ and $`(\text{4})`$ (recall that these bounds imply $`(\text{34})`$ and $`(\text{35})`$ by Theorem 3 and Lemma 4). However, F3 is also necessary for the Lipschitz continuity, eq. $`(\text{33})`$, in $`||_0`$, proven in I, which is essential for our iteration estimates. One should also keep in mind that if F3 does not hold, the selfenergy $`\mathrm{\Sigma }`$ and the function $`K`$ will in general not even be $`C^1`$ (in one dimension, where there are no curvature effects, $`\mathrm{\Sigma }`$ is not $`C^1`$; this is the source of anomalous decay exponents of the two–point function). The result that requires the most restrictive hypotheses is that, for $`e_s`$, the bound $`(\text{32})`$ for $`\left|K^{(R)}\right|_2`$ holds. This provides an invariant set for the iteration. The proof of $`(\text{32})`$, contained in II and III, uses very detailed geometric estimates which require convexity and positive curvature, as well as the condition $`S_e_2`$. The conditions stated in Section 1.2 (including, in particular, the symmetry (Sy): $`e(𝐩)=e(𝐩)`$ for all $`𝐩`$) imply hypotheses (H2)<sub>2,0</sub>, (H3), (H4), and (H5) of II and thus imply $`(\text{32})`$. In the asymmetric case, where the condition (Sy) is dropped, the regularity proof of II and III requires an additional hypothesis, stated as (H4’) in II, which imposes a minimal rate of change of the curvature of the Fermi surface at those points where the curvature coincides with that at the antipode. This condition (H4’) is not stable under an iteration in $`||_{3,r}`$. It is, however, only needed to estimate the contributions to $`K`$ of a very special class of graphs (the so–called wicked ladders; see Section II.4). We shall analyze these contributions in a further paper, to extend our regularity proof, and thus the inversion theorem, to the asymmetric case. The asymmetry plays a critical role in the proof of the existence of a two–dimensional Fermi liquid at zero temperature that was announced in . The set $`_s(\delta _0,g_0,G_0,\omega _0)C^3(,)`$ of starting $`E`$ allowed in Theorem 1 is not an open subset of $`C_s^2(,)`$. However, a look at the more detailed Theorem 2 shows that the inversion map really maps the ball $`B_{\mathrm{rad}}`$, defined in $`(\text{37})`$, which is open in $`||_{3,r}`$, to itself (see $`(\text{38})`$). Thus in the space of functions with bounded radial derivatives, there is an open set for which the inversion equation has a solution. Observe that, for our inversion theorem, in contrast to the KAM theorem, there is no diophantine condition for irrationality of frequencies. As mentioned above, we needed the norm $`||_{3,r}`$ instead of $`||_2`$ merely for apparently rather technical reasons. A superficial look at part 4 of Theorem 2 even seems to suggest that one can extend the inversion map to balls in $`_s`$ that are open in $`||_2`$ However, this is not the case because $`\lambda _R`$ depends on $`G_3`$, so $`(\text{4})`$ does not imply that the inverse map is defined on a dense subset of $`B_{\epsilon /2}^{(2)}`$. ## Appendix A Proof of Lemma 1 We first show that $`(\text{18})`$ follows from $`(\text{17})`$. Fix any $`𝐩S`$. Let $`T`$ be the tangent plane to $`S`$ at $`𝐩`$ and let $`𝐱`$ be the point of $`T`$ nearest $`𝐜`$. Since $`S`$ is convex it lies on one side of $`T`$. So the sphere of radius $`\frac{1}{K}`$ centered on $`𝐜`$, which by $`(\text{17})`$ is inside $`S`$, also lies on one side of $`T`$. Hence $`𝐱𝐜\frac{1}{K}`$. The vector $`𝐱𝐜`$ is normal to $`T`$ and hence parallel to $`𝐧(𝐩)`$. So $`\theta (𝐩)`$ is the angle between $`𝐱𝐜`$ and $`𝐩𝐜`$ and $$\mathrm{cos}\theta (𝐩)=\frac{𝐱𝐜}{𝐩𝐜}\frac{1/K}{1/k}=\frac{k}{K}$$ (139) We now prove $`(\text{17})`$, starting with $`𝐩𝐜\frac{1}{K}`$. This is a variant of a classical result. See, for example, §24 of . Let $`L>K`$ and define, for each $`𝐩S`$, $$\stackrel{~}{𝐩}(𝐩)=𝐩\frac{1}{L}𝐧(𝐩)$$ (140) Set $$\stackrel{~}{S}=\{\stackrel{~}{𝐩}(𝐩):𝐩S\}$$ (141) Then $`\stackrel{~}{S}`$ is a $`C^1`$ surface. We claim further that $`𝐧(𝐩)`$ is normal to $`\stackrel{~}{S}`$ at $`\stackrel{~}{𝐩}(𝐩)`$. To see this, let $`𝐭`$ be a unit vector that is a principal direction for $`S`$ at $`𝐩`$. Call the corresponding principal curvature $`\kappa `$. Let $`𝐪(s)`$ be a curve on $`S`$ that is parametrized by arc length, passes through $`𝐩`$ at $`s=0`$ and has tangent vector $`𝐭`$ there. Then $`s\stackrel{~}{𝐩}\left(𝐪(s)\right)=𝐪(s)\frac{1}{L}𝐧\left(𝐪(s)\right)`$ is a curve on $`\stackrel{~}{S}`$ that passes through $`\stackrel{~}{𝐩}(𝐩)`$ at $`s=0`$ and has tangent vector $$\frac{d}{ds}\stackrel{~}{𝐩}\left(𝐪(s)\right)|_{s=0}=𝐭\frac{1}{L}\frac{d}{ds}𝐧\left(𝐪(s)\right)|_{s=0}=𝐭\frac{\kappa }{L}𝐭$$ (142) there. Since $`\kappa <L`$, $`𝐭`$ is also a tangent vector to $`\stackrel{~}{S}`$ at $`\stackrel{~}{𝐩}(𝐩)`$. As this is the case for all principal directions $`𝐭`$, the tangent plane to $`\stackrel{~}{S}`$ at $`\stackrel{~}{𝐩}(𝐩)`$ is parallel to the tangent plane to $`S`$ at $`𝐩`$. Since $`S`$ is strictly convex, with principal curvatures bounded away from zero, the Gauss map $`𝐩S𝐧(𝐩)`$ is bijective and has a $`C^1`$ inverse $`𝐧S^{d1}𝐩(𝐧)S`$. The map $`𝐧S^{d1}\stackrel{~}{𝐩}\left(𝐩(𝐧)\right)`$ is then $`C^1`$ and surjective. Furthermore, the normal to $`\stackrel{~}{S}`$ at $`\stackrel{~}{𝐩}\left(𝐩(𝐧)\right)`$ is the same as the normal to $`S`$ at $`𝐩(𝐧)`$, which is $`𝐧`$. Consequently, $`\stackrel{~}{S}`$ is convex. As the chord $`𝐜_1𝐜_2`$ is of maximal length, it must be parallel to both $`𝐧(𝐜_1)`$ and $`𝐧(𝐜_2)`$. Thus $$𝐧(𝐜_1)=\frac{𝐜_1𝐜_2}{𝐜_1𝐜_2}=𝐧(𝐜_2)$$ (143) so that $$𝐜=\frac{1}{2}\left(𝐜_1+𝐜_2\right)=\frac{1}{2}\left(𝐜_1\frac{1}{L}𝐧(𝐜_1)\right)+\frac{1}{2}\left(𝐜_2\frac{1}{L}𝐧(𝐜_2)\right)$$ (144) is also the midpoint of a line joining two points of $`\stackrel{~}{S}`$. By convexity, $`𝐜`$ is inside $`\stackrel{~}{S}`$. The convexity of $`\stackrel{~}{S}`$ also implies that $`\stackrel{~}{S}`$ lies on one side of the tangent plane at $`\stackrel{~}{𝐩}(𝐩)`$, the side opposite $`𝐧(𝐩)`$. Hence $`𝐜`$, which is inside $`\stackrel{~}{S}`$ and $`𝐩S`$ are on opposite sides of the tangent plane to $`\stackrel{~}{𝐩}(𝐩)`$. In particular, the straight line from $`𝐜`$ to the nearest point, say $`𝐩_0`$, of $`S`$ is parallel to $`𝐧(𝐩_0)`$ and coincides, in part, with the line from $`\stackrel{~}{𝐩}(𝐩_0)`$ to $`𝐩_0`$, which is of length $`\frac{1}{L}`$. We conclude that $`𝐩𝐜\frac{1}{L}`$ for every $`L>K`$ and every $`𝐩S`$. The proof that $`𝐩\frac{1}{k}`$ is similar. This time, one lets $`\mathrm{}<k`$ and defines $$\stackrel{~}{𝐩}(𝐩)=𝐩\frac{1}{\mathrm{}}𝐧(𝐩)$$ (145) and sets $$\stackrel{~}{S}=\{\stackrel{~}{𝐩}(𝐩):𝐩S\}$$ (146) This time, $`\stackrel{~}{S}`$, and hence $`𝐜`$, lies on the same side of the tangent plane at $`\stackrel{~}{𝐩}(𝐩)`$ as $`𝐧(𝐩)`$. So the straight line from $`𝐜`$ to the farthest point, say $`𝐩_0`$, of $`S`$ is contained in the line from $`\stackrel{~}{𝐩}(𝐩_0)`$ to $`𝐩_0`$, which is of length $`\frac{1}{\mathrm{}}`$. When $`S`$ is invariant under inversion in the origin, $`𝐧(𝐜_1)=𝐧(𝐜_2)`$ implies that $`𝐜_1=𝐜_2`$ so that $`𝐜=\frac{1}{2}\left(𝐜_1+𝐜_2\right)=\mathrm{𝟎}`$. ## Appendix B Proof of Lemma 2 Let $`𝐩`$ be any point of $`S_{E_0}`$ and let $`𝐭`$ be any principal direction for $`S_{E_0}`$ at $`𝐩`$. Let $`𝐪(s)`$ be a curve on $`S_{E_0}`$ that is parametrized by arc length, passes through $`𝐩`$ at $`s=0`$ and has tangent vector $`𝐭`$ there. The principal curvature $`\kappa `$ corresponding to $`𝐭`$ obeys $$\kappa 𝐭=\frac{d}{ds}\frac{E_0\left(𝐪(s)\right)}{E_0\left(𝐪(s)\right)}|_{s=0}=\frac{E_{0}^{}{}_{}{}^{\prime \prime }(𝐩)𝐭}{E_0(𝐩)}+E_0\left(𝐩\right)\frac{d}{ds}\frac{1}{E_0\left(𝐪(s)\right)}|_{s=0}$$ (147) and hence $$\kappa =\frac{(𝐭,E_{0}^{}{}_{}{}^{\prime \prime }(𝐩)𝐭)}{E_0(𝐩)}$$ (148) Consequently, $`S_{E_0}`$ is a convex surface that is invariant under inversion in the origin and has all principal curvatures between $`\frac{\omega _0}{G_0}`$ and $`\frac{G_0}{g_0}`$. By Lemma 1, $$\frac{}{r}E_0(𝐩(r,\theta ))=E_0(𝐩(r,\theta ))\frac{𝐩}{r}(r,\theta )E_0(𝐩(r,\theta ))\frac{\omega _0/G_0}{G_0/g_0}\frac{\omega _0g_{0}^{}{}_{}{}^{2}}{G_{0}^{}{}_{}{}^{2}}$$ (149) for all $`r=r_F(E_0,\theta )`$. Choose $`g_1=\frac{\omega _0g_{0}^{}{}_{}{}^{2}}{4G_{0}^{}{}_{}{}^{2}}`$ and $`r_0=\mathrm{min}\{\frac{g_1}{G_0},\delta _0\}`$. Then $`{\displaystyle \frac{}{r}}E_0(𝐩(r,\theta ))`$ $`=`$ $`E_0\left(𝐩(r_F(E_0,\theta ),\theta )\right)\frac{𝐩}{r}(r,\theta )`$ $`+\left[E_0\left(𝐩(r,\theta )\right)E_0\left(𝐩(r_F(E_0,\theta ),\theta )\right)\right]\frac{𝐩}{r}(r,\theta )`$ and $$\left|\left[E_0\left(𝐩(r,\theta )\right)E_0\left(𝐩(r_F(E_0,\theta ),\theta )\right)\right]\frac{𝐩}{r}(r,\theta )\right|G_0\left|rr_F(E_0,\theta )\right|$$ (151) so $$\frac{}{r}E_0(𝐩(r,\theta ))2g_1\text{ for all }\left|rr_F(E_0,\theta )\right|2r_0,\theta S^{d1}$$ (152) Similarly, if $`\left|eE_0\right|_1g_1`$, $$\frac{}{r}e(𝐩(r,\theta ))g_1\text{ for all }\left|rr_F(E_0,\theta )\right|2r_0,\theta S^{d1}$$ (153) This verifies $`(\text{21})`$. We merely need to choose $`\epsilon <g_1`$. To verify $`(\text{20})`$, observe that if $`\left|eE_0\right|_0r_0g_1`$, then $`\left|e(r_F(E_0,\theta ),\theta )\right|r_0g_1`$ and hence $$\left|r_F(e,\theta )r_F(E_0,\theta )\right|r_0$$ (154) by $`(\text{21})`$. The same argument that shows that $`_s`$ is open in $`(C_s^2(,),||_2)`$ also yields $`e_s(\delta _0/2,g_0/2,2G_0,\omega _0/2)`$, if we choose $`\epsilon `$ small enough, depending only on $`\delta _0,g_0,G_0`$ and $`\omega _0`$. ## Acknowledgements We thank H. Knörrer for suggesting the proof of Lemma 1.
warning/0001/astro-ph0001118.html
ar5iv
text
# Accepted for publication in Monthly Notices Evolution of the cosmological density distribution function ## 1 introduction The standard scenario for the formation of structure in the Universe is via the gravitational amplification of primordial random Gaussian fluctuations generated in the Early Universe during an Inflationary phase. An attractive feature of this scenario is its predictive power in determining the history of mass perturbations from initial times up until the present day. In principle this should allow one to compare observation with theory. But while detailed predictions at high redshift for the Cosmic Microwave Background (CMB) are possible due to their linearity and dependence on well tested laboratory physics, predictions at lower redshift are complicated by nonlinear gravitational evolution and the physics of galaxy formation. One of the goals of cosmology is accounting for the statistical evolution of mass and galaxies up to the present day. In this paper we focus on the properties of the one–point density distribution function of matter. Interest in the distribution functions of cosmological fields has grown as they encode a great deal of information on initial conditions, gravitational evolution and galaxy bias. However the challenge is to separate out each of these effects. The evolution of the one–point distribution function has been computed by a number of authors using a variety of techniques here we present a new approach, based on propagating the distribution function by the Chapman–Kolmogorov equation, incorporating nonlinear evolution to second order. In a further paper we shall develop the formalism to include the effects of galaxy biasing and redshift–space distortions (Watts & Taylor 2000). The 1-pt PDF is a useful quantity in cosmology. While reasonably straightforward to measure (see Hamilton 1985, Alimi et al 1990, Szapudi, Szalay & Boschán 1992, Gaztañaga 1992, 1994, Bouchet et al 1993, Szapudi, Meiksin & Nichol 1996, Kim & Strauss 1998) the 1-pt PDF embodies the full hierarchy of correlation functions and its measurement ensures that physical constraints such as the Lyapunov inequalities for the moments ($`x^{a+b}x^ax^b`$, for any random variable $`x`$) are automatically satisfied. This is not necessarily the case for direct measurement of the moments. In addition the PDF offers a convenient way of dealing with Poisson sampling in the galaxy distribution. From a theoretical perspective, the 1-pt PDF is a convenient quantity to calculate if it is initially Gaussian. A number of different methods have been developed to calculate its evolution. Fry (1985) first suggested calculating the probability function by applying a hierarchical series suggested by the BBGKY equations to solve for the moment generating function. Bernardeau (1992) derived an exact expression for the evolution of the moment generating function. The generalisation to top-hat filtered density and velocity fields is found in Bernardeau (1994), while Bernardeau (1996) studied the 2-point cumulants and the PDF. However so far a generalisation of these results to include bias and redshift–space distortions in the PDF has not been presented. An approximation for the PDF can also be constructed from the first few cumulants, in the limit of small variance, by the Edgeworth expansion (Juszkiewicz et al., 1995 and Bernardeau & Kofman, 1995), where the cumulants have been derived directly from Eulerian perturbation theory (Bouchet et al., 1992) or Lagrangian perturbation theory (Bouchet et al. 1995). This can be extended to redshift–space using the Lagrangian perturbation calculations of Hivon et al. (1995) for the skewness. Colombi et al. (1997) introduced an extended perturbation theory, where the results of perturbation theory are allowed extra freedom and extrapolated to the nonlinear regime. As well as Eulerian perturbation theory other approximations have also been applied. Kofman et al (1994) and Bernardeau & Kofman (1995) derived the PDF in the Zel’dovich approximation, while Hui, Kofman & Shandarin (1999) extended this to include the effects of redshift–space distortions. A more phenomenological approach was taken by Coles & Jones (1991) who approximated the properties of the PDF by a lognormal distribution while Colombi (1994) suggested an Edgeworth expansion about the lognormal distribution. In this letter we apply second–order perturbation theory to an initially Gaussian distributed density field to calculate the exact second–order characteristic function. We then numerically inverse Fourier transform this to yield the 1-pt pdf. Our method therefore treats the propagation of probabilities exactly. The letter is organised as follows. In Section 2 we calculate the evolution of the cosmological probability distribution function and discuss the transformation to a discrete count distribution. In Section 3 we compare our results with the results of N-body simulations and with other distributions in Section 4. Conclusions are presented in Section 5. We begin by describing second–order perturbation theory and calculating the probability distribution function. ## 2 The cosmological probability distribution function ### 2.1 Second–order perturbation theory Before shell–crossing, and in a spatially flat universe, the Eulerian density field, $`\delta (𝒙,t)`$, can be expanded in a series of separable functions; $$\delta (𝒙,t)=\underset{n=1}{\overset{\mathrm{}}{}}\delta _n(𝒙,t)=\underset{n=1}{\overset{\mathrm{}}{}}D_n(t)\epsilon _n(𝒙),$$ (1) where $`\epsilon _n`$ is an $`n^{th}`$-order time-independent density field and $`D_n`$ is an $`n^{th}`$-order universal growth function. It is a special property of this expansion that the spatial and time components are separable. In a universe with spatial curvature this expansion is only possible to second–order. The growth of perturbations in the nonlinear regime can be calculated by solving the continuity, Euler and Poisson equations sequentially for each order in the perturbation expansion. To second–order the density field can be derived from linear quantities by the relation (Peebles 1980, Bouchet et al 1992) $$\delta =\delta _1+\frac{1}{3}(2\kappa )\delta _1^2𝜼.𝒈+\frac{1}{2}(1+\kappa )E^2,$$ (2) where $$𝜼(𝒙,t)=\mathbf{}\delta _1(𝒙,t)$$ (3) is the gradient of the linear density field and $$𝒈(𝒙,t)=\mathbf{}^2\delta _1(𝒙,t)$$ (4) is the linear peculiar gravity field<sup>1</sup><sup>1</sup>1In this paper we define $`4\pi G\rho _0=3/2\mathrm{\Omega }H^2=1`$ and the expansion parameter $`a(t)=1`$, since our final distribution function will be dimensionless. where $`^2`$ is the inverse Laplacian. The trace-free tidal tensor is given by $$E_{ij}(𝒙,t)=_i_j^2\delta _1(𝒙,t)\frac{1}{3}\delta _1(𝒙,t)\delta _{ij}.$$ (5) Equation (2) is the general result for an arbitrary cosmology where<sup>2</sup><sup>2</sup>2We define $`\kappa `$ following Bouchet et al. (1995) and Hivon et al. (1995), but note that this differs from the definition of Bouchet et al. (1992) who define $`\kappa 2/14\mathrm{\Omega }^{2/63}`$ $`\kappa =D_2/D_1^23/7\mathrm{\Omega }^{2/63}`$ (Bouchet et al 1992, see also Catelan et al. 1995), and $`\delta _1(𝒙,t)=D_1(t)\epsilon _1(𝒙)`$, where $`D_1(t)(1+z)^{\mathrm{\Omega }^{0.6}}`$ is the linear growth function (Peebles 1980). ### 2.2 The distribution function of initial fields In order to calculate the second–order density distribution function, $`P(\delta )`$, we must first calculate the joint probability of each of the fields in equation (2), $`P(\delta _1,𝜼,𝒈,𝑬)`$. Defining the parameter vector $`𝒚=(\delta _1,𝜼,𝒈,𝑬)`$ the joint distribution function is given by the multivariate Gaussian $$P(𝒚)=\frac{1}{((2\pi )^ndet𝑪)^{1/2}}\mathrm{exp}\left(\frac{1}{2}𝒚^t𝑪^1𝒚\right),$$ (6) where $$𝑪=𝒚𝒚^t$$ (7) is the $`12\times 12`$ covariance matrix that gives the degree of correlation between each of the components of $`𝒚`$. The elements of the covariance matrix are $`\delta ^2`$ $`=`$ $`\sigma _0^2,\eta _i\eta _j={\displaystyle \frac{1}{3}}\sigma _1^2\delta _{ij},`$ $`g_ig_j`$ $`=`$ $`{\displaystyle \frac{1}{3}}\sigma _1^2\delta _{ij},\eta _ig_j={\displaystyle \frac{1}{3}}\sigma _0^2\delta _{ij},`$ $`E_{ij}E_{kl}`$ $`=`$ $`{\displaystyle \frac{1}{15}}\sigma _0^2\left(\delta _{ik}\delta _{jl}+\delta _{il}\delta _{jk}{\displaystyle \frac{2}{3}}\delta _{ij}\delta _{kl}\right),`$ (8) with the remaining entries in the covariance matrix zero. These variances are defined as $$\sigma _n^2=D^2(t)_0^{\mathrm{}}\frac{dk}{2\pi ^2}k^{2+2n}P(k),$$ (9) where $`P(k)`$ is the initial power spectrum of density perturbations. The components of this distribution are $$𝒚^t𝑪^1𝒚=\frac{\delta ^2}{\sigma _0^2}+\frac{3}{(1\gamma _\nu ^2)}\left(\frac{\eta ^2}{\sigma _1^2}+\frac{g^2}{\sigma _1^2}2\gamma _\nu \frac{𝜼.𝒈}{\sigma _1\sigma _1}\right)+15\frac{E^2}{\sigma _0^2}$$ (10) and $$det𝑪=\frac{20}{3105^3\pi ^4}\sigma _0^{12}\sigma _1^6\sigma _1^6(1\gamma _\nu ^2)^3.$$ (11) The correlation parameter, $`\gamma _\nu `$, is defined as $$\gamma _\nu =\frac{\sigma _0^2}{\sigma _1\sigma _1},$$ (12) providing a measure of the correlation between the linear velocity and density gradient fields. If we assume a power-law power spectra with a Gaussian cut-off, $`P(k)k^ne^{k^2R^2}`$, where n is the spectral index and $`R`$ some arbitrary length scale, then $$\gamma _\nu =\frac{n+1}{n+3},$$ (13) with the constraint $`n>1`$ to insure convergence of the velocity field. The correlation parameter must be positive definite since under gravity matter will be displaced from low– to high–density regions. The special case of $`n1`$ for a power-law initial spectra results in $`\gamma _\nu =0`$ because the velocity field diverges on small scales. In fact the diverging velocities form an incoherent Gaussian random field which is uncorrelated with the density field (Taylor & Hamilton 1996). ### 2.3 Propagation of the density distribution function The distribution function can be propagated to later times by the Chapman–Kolmogorov equation, $$P(𝒙)=𝑑𝒚W(𝒙|𝒚)P(𝒚)$$ (14) where $`W(𝒙|𝒚)`$ is the transition probability from $`𝒙`$ to $`𝒚`$. In the case of a deterministic process, such as the one considered here, the transition probability reduces to a delta function restricting the number of possible paths of evolution to one. This transition probability is given by $$W(\delta |𝒚)=\delta _D[\delta \delta _1\frac{1}{3}(2\kappa )\delta _1^2+𝜼.𝒈\frac{1}{2}(1+\kappa )E^2]$$ (15) Inserting this into equation (14) we find $$P(\delta )=\delta _D(\delta \delta (𝒚))_y.$$ (16) where $`\delta (𝒚)`$ is the right hand side of equation (2). Hence for deterministic transitions the Chapman–Kolmogorov equation becomes the expectation value of the delta function. This is a well–known result from probability theory (eg van Kampen 1992). Fourier transforming the delta function we find $$P(\delta )\frac{1}{2\pi }_{\mathrm{}}^{\mathrm{}}𝑑J𝒢(J)\mathrm{exp}(iJ\delta )$$ (17) where $`𝒢(J)`$ $``$ $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑\delta P(\delta )\mathrm{exp}\left(iJ\delta \right)`$ (18) $`=`$ $`\mathrm{exp}\left(iJ\delta (𝒚)\right)_y`$ is the characteristic function. In the second line we have used equation (16) to write the characteristic function as an expectation over all the linear fields, $`𝒚`$. This expression reduces to a set of multivariate Gaussian–type integrals that we can easily evaluate yielding $$𝒢(J)=\mathrm{\Theta }(J)\mathrm{exp}\left[\frac{J^2\sigma _0^2}{2(1+i\alpha _1J)}\right]$$ (19) where $$\mathrm{\Theta }(J)=(1+i\alpha _1J)^{1/2}(1+i\alpha _2J)^{5/2}(1i\alpha _3J+\alpha _4J^2)^{3/2}$$ (20) and where the $`\alpha `$ coefficients are given by $$\alpha _1=\frac{2}{3}(2\kappa )\sigma _0^2,\alpha _2=\frac{2}{15}(1+\kappa )\sigma _0^2,$$ $$\alpha _3=\frac{2}{3}\sigma _0^2,\alpha _4=\frac{1}{9}\frac{(1\gamma _\nu ^2)}{\gamma _\nu ^2}\sigma _0^4$$ (21) The probability distribution can then be found by numerically integrating equation (17). Equations (19), (20) and (21) are the central results of this letter. The characteristic function for $`\delta `$ can be separated into characteristic functions for each term in the summation in equation (2). This is to be expected, since the distribution of a summation of random variables is equal to the convolution of their individual distributions, and so the characteristic functions just multiply. We expect this to be true for all orders of the perturbation series. But rather than converging to a Gaussian distribution, as suggested by a naive application of the Central Limit Theorem, the final distribution is driven away from it by the correlations induced by gravity. Figure 1 shows the evolved density distribution function for a range of $`\sigma _0`$. We use linear axis for the top plot emphasising the peak and logarithmic axis in the lower plot emphasising the tail of the distribution. As expected, the shape of the distribution is very nearly Gaussian when the variance is small, becoming very rapidly non-Gaussian for higher values. For high variances the probability density does not drop to zero at $`\delta =1`$, since in second–order perturbation theory the density field can be negative, generating non-vanishing regions with $`\delta 1`$. This is also true of linear theory where there are always negative density regions for Gaussian initial conditions. This is a feature of Eulerian distribution functions. Those calculated in Lagrangian perturbation theory, or the lognormal, have positive definite densities at all times. In Figure 2 we demonstrate the effects of the correlation parameter, $`\gamma _\nu `$, on the PDF. The main effects are an increase in volume of underdense regions with a corresponding decrease in extremely underdense regions when $`\gamma _\nu `$ is high. For low $`\gamma _\nu `$, the underdensities are smaller and deeper. The physical reason for this is that the $`𝜼.𝒈`$ term in second–order perturbation theory deals with the evacuation of the voids, rather than the amplification of peaks. When $`\gamma _\nu `$ is low this term is weakened and voids tend to be smaller and deeper, as they would be if the linear field were extrapolated. Increasing this correlation widens the voids, but makes them shallower to help satisfy the $`\delta 1`$ constraint. This effect is small for CDM-type initial power spectra where the correlation parameter is in the range $`0.55<\gamma _\nu <0.65`$ over a wide range of scales. ### 2.4 Skewness from the characteristic function Since our characteristic function is correct to second–order a useful check is to calculate the variance and skewness to compare with previous estimates. Taking the derivatives of the characteristic function with respect to $`J`$ and setting $`J`$ equal to zero yields the moments of the evolved density distribution function, $$\frac{^n}{[iJ]^n}𝒢(J=0)=\delta ^n.$$ (22) In the literature it is common to express the moments in the form of the moment parameters $$S_n=\delta ^n_c/\delta ^2^{n2}$$ (23) where $`\delta ^n_c`$ is the connected or irreducible part of $`\delta ^n`$. The irreducible moments, or cumulants, can be generated by $$\frac{^n}{[iJ]^n}\mathrm{ln}𝒢(J=0)=\delta ^n_c$$ (24) From this we can calculate the second and third order connected moments of our distribution function; $$S_2=1$$ (25) and $$S_3=2(2\kappa )\stackrel{\mathrm{\Omega }1}{}\frac{34}{7}$$ (26) to lowest order, reproducing the results of Peebles (1980). Hence our PDF leads to the correct second–order skewness. The intrinsic effects of the shape of the power spectrum via the correlation coefficient $`\gamma _v`$ are to higher order. ### 2.5 The discrete distribution function In reality the density field of galaxies is not a continuous function, but a discrete distribution. To account for this it is usual to assume Poisson sampling of the continuous density field as a crude approximation to galaxy formation processes. Any variation is attributed to biased galaxy formation, which modulates the underlying function. We shall consider this elsewhere (Watts & Taylor, 2000). The continuous density distribution function can be transformed to a discrete form by the expectation $$P(n)=\frac{\overline{n}^n}{n!}(1+\delta )^ne^{\overline{n}(1+\delta )}_\delta ,$$ (27) where $`\overline{n}`$ is the mean galaxy count and the expectation is taken with respect to the nonlinear density distribution. Expanding this in terms of the characteristic function and taking the expectation we find $$P(n)=_{\mathrm{}}^{\mathrm{}}𝑑JG(J)\left(1\frac{iJ}{\overline{n}}\right)^{n1}\mathrm{exp}(iJ)$$ (28) In the limit $`\overline{n}\mathrm{}`$ this returns the continuum distribution where $`\delta =n/\overline{n}1`$. We note that again that the discrete distribution has a non-vanishing value at $`n=0`$, $`P(n=0)`$, since the probability of finding a no galaxies within a cell is finite. This is the Void Probability Function (White 1979). It is also useful to have the generating function for the discrete moments of this distribution, $`G_n(k)`$. Following some straightforward calculation we find that $$G_n(k)=G[i\overline{n}(e^{ik}1)]\mathrm{exp}\overline{n}(e^{ik}1).$$ (29) The discrete moments can then be found by differentiating $`G_n(k)`$; $$\delta ^m=\frac{^mG_n(k=0)}{(ik)^m}$$ (30) Again the connected moments can be found by differentiating $`\mathrm{ln}G(k)`$ with respect to $`ik`$. ## 3 Comparison of results with n-body simulations In this section we show a comparison between our theoretical PDF and the counts in cells PDF found from cosmological n-body simulations. We used a version of Hugh Couchman’s (1991) Adaptive P<sup>3</sup>M code altered by Peacock & Dodds (1994) to allow simulations of low density open and flat universes. The simulation volume was a periodic cube of comoving side $`200h^1\mathrm{Mpc}`$ containing $`100^3`$ particles. We chose a CDM–type linear power spectrum of Gaussian initial perturbations (Bardeen et al. 1986), normalised to match the observed abundance of clusters with linear variance given by $`\sigma _8=0.6\mathrm{\Omega }^{0.53}`$ (Viana & Liddle 1996). The simulation was carried out on a $`128^3`$ Fourier mesh. To measure the PDF from the numerical simulations we smoothed the discrete particle distribution with a Gaussian filter of radius $`R`$. The PDF was then found from the smoothed density field evaluated on a $`128^3`$ grid. Since the effective radius of the binning grid was much smaller than that of the filter radius we expected negligible contribution to the smoothing from it. The choice of variance to use in our model is slightly ambiguous, given that we do not treat smoothing exactly. Hence it is better to match variances rather than smoothing scales. However the variance in second–order perturbation theory is the same as in linear theory so we could use either the linear input variance from the simulations, or take the measured nonlinear variance when making a comparison. In practice we found that all of the models provided a better fit if the nonlinear variances where used. This is a well know effect and is the basis of the extended perturbation theory approach of Colombi et al (1997). Figures (3, top) and (3, bottom) show the results of numerical simulations (points) alongside our second–order PDF (solid line), the lognormal model (Coles & Jones 1991, dotted) and the Edgeworth expansion (Juszkiewicz et al. 1995, Bernardeau & Kofman 1995; dashed). Each plot shows the PDF for two different variances, $`\sigma _0=0.2`$, smoothed on a scale of $`R=20h^1\mathrm{Mpc}`$ (open circles) and $`0.3`$ (filled squares) smoothed on $`13.5h^1\mathrm{Mpc}`$. The effects of shot noise were taken into account as shown in Section 2.3 with the mean particle density $`\overline{n}`$ chosen to coincide with the mean particle count in a boxes with side $`l=\sqrt{5}R`$. Since $`\overline{n}`$ was a very large number in all cases, shot noise effects were extremely small. This would not be the case in a real galaxy survey with lower number counts. The remaining factor controlling the shape of the theoretical PDF was $`\gamma _\nu `$. For linear CDM power spectra the calculated $`\gamma _\nu `$ was approximately 0.55 for a wide range of smoothing radii and it was this value that was used to prepare all the plots in this section. Error bars on the N-body data points are the standard deviation over 5 independent simulations with identical cosmological parameters. The linear axes of figures (3, top) shows the accuracy around the peak of the distributions while the logarithmic scale of figure (3, bottom) shows the tails of the distributions. At low variance ($`\sigma _0=0.2`$) we find very good agreement between all of the models and the simulations. Overall we found that the lognormal model fitted the simulations extremely well for all values of the variance and around both the peak and the tails. As has been remarked elsewhere we regard this as something of a fluke (Bernardeau & Kofman 1995), and that the lognormal makes a very useful fitting function. Our second–order PDF fits the peak fairly accurately, but tends to be slightly too skewed. This effect becomes greater at higher variance. The tails of the distribution match the N-body simulations rather well, although at high variances and low densities the distribution is too high. This is due to the real distribution being constrained to go to zero at zero density, whereas, as we have already remarked the second–order PDF is not. In comparison the Edgeworth expansion, taken to second order (Juszkiewicz et al. 1995, Bernardeau & Kofman 1995) $$P_E(\delta )=\left[1+\frac{1}{6}S_3\sigma _0H_3(\delta /\sigma _0)\right]\frac{e^{\delta ^2/2\sigma _0^2}}{\sqrt{2\pi }},$$ (31) where $`H_n(x)`$ is a Hermite polynomial, is also slightly too skewed and tends to undershoot the high density tail compared to the simulations. At large variance the expansion breaks down and the Edgeworth distribution develops a wiggle on the high density tail. In fact the Edgeworth expansion fares badly even when taken to third order because the unphysical wiggles are exaggerated and the low density tail becomes more rapidly negative. ## 4 Comparison with other approximations Finally we compare our PDF with other approximations for the 1-point density distribution. In figure 4 we show three distributions, our second–order PDF calculation (solid line), the Zel’dovich approximation (Kofman et al. 1994; dot-dashed line) and the Edgeworth expansion (dotted line) plotted relative to the lognormal distribution. Our choice of the lognormal distribution as the point of normalisation is based on the agreement we find between it and our N-body simulations. Over the range $`0.4<\delta <1.5`$ all of the models agree to within a few percent, with the exception of the second–order Edgeworth expansion. But for densities below this regime, our second–order PDF overpredicts because the positive density condition is only weakly met. Both the Zel’dovich and Edgeworth models underpredict the probability of low density regions. Above the regime of good agreement our model again overpredicts, again because we over-extrapolate the high density evolution, but not as much as the Zel’dovich approximation, where caustic formation tends towards a too high distribution. In contrast to both models the Edgeworth approximation underpredicts the high density regions and develops a large wiggle in the positive density regime. ## 5 Summary We have presented a new method for calculating the nonlinear evolution of the 1-point density distribution function using the Chapman–Kolmogorov equation to propagate the probability distribution, and second–order perturbation theory to evolve the density field. This has the advantage over other methods that the resultant probability distribution must be positive definite, and can be readily extended to include the effects of Eulerian deterministic or stochastic bias and redshift space distortions. The main disadvantages of our method are that it is not obvious how to include the effects of smoothing in the final distribution and it is difficult to see how one could extend the evolution of the density field to higher order. However, despite these problems we find extremely good agreement with numerical simulations and that our distribution is an improvement on other approximations. Since we derive the characteristic function the moments and cumulants (connected moments) of the field are easily derived, as is the distribution of any local transformations of the density field. The formalism we have introduced can also be used to calculate more complicated 1- and 2-point distributions of cosmologically interesting fields. ## Acknowledgements PIRW thanks the PPARC for a postgraduate grant. ANT thanks the PPARC for a postdoctoral fellowship. ## References Alimi J.-M., Blanchard A., Schaeffer R., 1990, ApJ, 349, L5 Bardeen J.M., Bond J.R., Kaiser N., Szalay A.S., 1986, ApJ, 304, 15 Bouchet F.R., Juszkiewicz R., Colombi S., Pellat R., 1992, ApJ, 394, L5 Bouchet F.R., Strauss M.A., Davis M., Fisher K.B., Yahil A., Huchra J.P., 1993, ApJ, 417, 36 Bouchet F.R., Colombi S., Hivon E., Juszkiewicz R., 1995, AA, 296, 575 Bernardeau F., 1992, ApJ, 392, 1 Bernardeau F., 1994, AA 291, 697 Bernardeau F., 1996, AA, 312, 11 Bernardeau F., Kofman L., 1995, ApJ, 443, 479 Catelan P, Lucchin F., Matarrese S., Moscardini L, 1995, MNRAS, 276, 39 Coles P., Jones B., 1991, MNRAS, 248, 1 Colombi S., 1994, ApJ, 435, 536 Colombi S., Bernardeau F., Bouchet F.R., Hernquist L., 1997, MNRAS, 287, 241 Couchmann 1991, ApJ, 368, L23 Fry J.N., 1985, ApJ, 289, 10 Gaztañaga E., 1992, ApJ, 398, L17 Gaztañaga E., 1994, MNRAS, 286, 913 Hamilton A.J.S., 1985, ApJ, 292, L35 Hivon E., Bouchet F.R., Colombi S., Juszkiewicz R., 1995, AA, 298, 643 Hui L., Kofman L., Shandarin S.F., 1999 (astro-ph/9901104) Juszkeiwicz R., Weinberg D.H., Amsterdamski P., Chodorowski M., Bouchet F., 1995, ApJ, 442, 39 Kim R.S.J., Strauss M.A., 1998, ApJ, 493, 39 Kofman L., Bertschinger E., Gelb J.M., Nusser A., Dekel A., 1994, ApJ, 420, 44 Viana P.T.P., Liddle A.R., 1996, MNRAS, 281, 323 Peacock J.A., Dodds S., 1994, MNRAS, 267, 1020 Peebles P.J., 1980, “Large–Scale Structure in the Universe”, Princeton University Press, Princeton Szapudi I., Szalay A., Boschán P., 1992, ApJ, 390, 350 Szapudi I., Meiksin A., Nichol R., 1996, ApJ, 473, 15 Taylor A.N., Hamilton A.J.S., 1996, MNRAS, 282, 767 van Kampen, N.G., 1992, “Stochastic Processes in Physics and Chemistry”, North–Holland, Amsterdam Watts P.I.R, Taylor A.N., 2000, in Preparation White S.D.M, 1979, MNRAS, 186, 145
warning/0001/hep-ph0001282.html
ar5iv
text
# References Quantum electrodynamics of composite fermions. Serge B. Afanas’ev Sankt–Petersburg, Russia e-mail: serg@ptslab.ioffe.rssi.ru ## Abstract The quantum Dirac–like equation and the QED vertex operator for a composite particle are suggested. The vertex operator and the fermionic propagator are connected by the QED Ward identity. It is shown that all of the Feynman QED–integrals for a three–component particle are finite with some suggestion about large momentum dependence of the vertex operator. This dependence is in agreement with the component counting rules. The ultraviolet divergences appearance in calculations is one of the most difficult problem of the quantum field theory. At the present time many methods of ultraviolet divergences elimination are developed. Thus the ultraviolet divergences are not in the superstring theory, and this fact is connected with the finite sizes of superstrings as the space–time objects. In this work the quantum electrodynamics of a composite particle with the vertex operator and the motion equation different from the Dirac–Feynman–Schwinger pointlike QED ones is considered. Pointlike QED based on the Dirac equation $$\left(\widehat{p}m\right)\psi =0$$ (1) The general form of the vertex operator consequential from requirements of gauge and relativistic invariances is $$\mathrm{\Gamma }^\mu \left(q^2\right)=f\left(q^2\right)\gamma ^\mu \frac{1}{2m}g\left(q^2\right)\sigma ^{\mu \nu }q_\nu ,$$ (2) where $`q_\mu `$ is the photon momentum. The calculations with pointlike form factors in each exact vertex $`f(q^2)=1`$, $`g(q^2)=0`$ yield the ultraviolet divergences caused by momentum dependence of integrand in the large momentum region. It is noted that allowing the momentum dependence of form factors $`f(q^2)`$ and $`g(q^2)`$ results in eliminating ultraviolet divergences of the Feynman integrals in majority of QED–calculations if $`f(q^2)`$ and $`g(q^2)`$ are decreasing functions of momentum in the large momentum region. Asymptotical momentum dependence of composite particle form factors is described with functions $`f(q^2)(1/q^2)^{n1}`$, where $`n`$ is a number of components . In agreement with the components counting rules the suggestion about leptons substructure leads to elimination of all ultraviolet divergences, besides the polarization operator divergence and divergences in other Feynman diagrams in which all of external lines are photon. Leptons substructure and the upper limits of leptons sizes are discussed in . However, $`q^2`$–dependence of the form factors leads to the contradiction with the Ward identity: $$G^1\left(p+q\right)G^1\left(p\right)=q_\mu \mathrm{\Gamma }^\mu (p,p+q;q)$$ (3) because the one puts very rigid conditions on the form of the fermionic propagator. The identity (3) cannot be satisfied if form factors depend on $`q^2`$ only. Suppose that the vertex operator can depend on the lepton momentum as well as the photon momentum. Obviously, $`p`$–dependence of the vertex operator leads to a modification of the particle propagator, as the Ward identity (3) shows. Consider the following system consisting of the composite particle quantum relativistic motion equation: $$\left(f\left(p^2\right)\widehat{p}m\right)\psi =0$$ (4) and the vertex operator in the form $$\mathrm{\Gamma }^\mu (p,p+q;q)=h_1(p^2,q^2)\gamma ^\mu +h_2(p^2,q^2)\gamma _\nu p^\nu q^\mu $$ (5) In (5) the ”spin term” is omitted as it does not take a contribution to the Ward identity due to antisymmetric spin tensor presence. Define the fermionic propagator as the Green function of the equation (4): $$G\left(p\right)=\frac{f\left(p^2\right)\widehat{p}+m}{f^2\left(p^2\right)p^2m^2}$$ (6) Let us suppose that $`f(p^2)`$ is the decreasing function of $`p^2`$ and $$f\left(p^2\right)|_{p^2=0}=1$$ (7) The equation (4) and the propagator (6) under the $`p^20`$ conditions have pointlike QED – forms. If $$f\left(p^2\right)|_{p^2\mathrm{}}\frac{1}{p^{\alpha +1}},\alpha >0$$ (8) then the propagator (6) $$G\left(p\right)|_{p^2\mathrm{}}\frac{1}{m}$$ (9) Poles of the propagator (6) are yielded by the equation $$f^2\left(p^2\right)p^2=m^2$$ (10) As is shown below, the function $`f(p^2)`$ is identical to the elastic form factor in the small momentum region. Therefore it may be noted that $`|f(p^2)|<<1`$ in the region $`p^2<r^2>^1`$ where $`<r^2>`$ is the average composite particle size squared. Thus if $`m^2<r^2>1`$ then the propagator poles are $`p^2m^2`$ and $`p^2<r^2>^1`$. In the case $`m^2<r^2>1`$ both poles are in the region $`p^2m^2`$. If $`m^2<r^2>1`$ then poles of propagator are absent. Solutions of the equation (4) for a particle without external field are not eigenstates of the momentum operator. Thus the solutions of (4) are some linear combinations of the plane waves. Considering the equation (4) and the vertex operator (5) the Ward identity (3) is: $$f\left(\left(p+q\right)^2\right)\left(\widehat{p}+\widehat{q}\right)f\left(p^2\right)\widehat{p}=q_\mu \left(h_1(p^2,q^2)\gamma ^\mu +h_2(p^2,q^2)\gamma _\nu p^\nu q^\mu \right)$$ (11) This equation puts the following conditions on the functions $`f(p^2)`$ , $`h_1(p^2,q^2)`$ and $`h_2(p^2,q^2)`$: $$f\left(\left(p+q\right)^2\right)f\left(p^2\right)=q^2h_2(p^2,q^2)$$ (12) $$f\left(\left(p+q\right)^2\right)=h_1(p^2,q^2)$$ (13) The functions $`h_1(p^2,q^2)`$ and $`h_2(p^2,q^2)`$ are exactly determined by the equations (12), (13) and are connected by these equations. The last equation leads to the condition $$h_1(p_1^2,p_2^2)=h_1(p_2^2,p_1^2)$$ (14) Analyzing the Ward identity (3) with the vertex operator in the form (5), (12), (13) we come to the conclusion that this $`\mathrm{\Gamma }^\mu `$ is orthogonal to $`q_\mu `$. Thus the system of the composite particle quantum equation (4) and the vertex operator (5), (12), (13), is gradient-invariant and satisfies the electric charge conservation law. The vertex operator (5) is analogous to the vertex functions that appear at the investigation of the deep inelastic scattering. There can be suggested that the operator (5) describes the internal processes in a composite system. $`p^2`$–dependence of the function $`h_1(p^2,q^2)=f(p^2)`$ even at the zero momentum transmission describes the internal processes, but not the interaction of a composite particle with the electromagnetic field. This dependence coincides with $`p^2`$–dependence of the function $`G^1(p^2)`$. Consider the ultraviolet divergences in composite particle QED (4), (5). In this theory the photon propagator has the standard QED–form. In the large momentum region the fermionic propagator is finite. A number of integrations over internal momenta of a diagram with $`N`$ vortices, $`N_f`$ external fermionic lines and $`N_{ph}`$ external photon lines, is $`2(NN_fN_{ph}+2)`$. Thus the index of a diagram $`(N,N_f,N_{ph})`$ equals $$\mathrm{\Omega }_{NN_fN_{ph}}=2\left(NN_fN_{ph}+2\right)2\frac{NN_{ph}}{2}N_\mathrm{\Gamma }N$$ (15) where $`N_\mathrm{\Gamma }`$ is the power of momentum in the expression for the vertex operator. Hence $$\mathrm{\Omega }_{NN_fN_{ph}}=N2N_fN_{ph}+4N_\mathrm{\Gamma }N$$ (16) It is known that the most critical divergence is in the polarization operator ($`N_{ph}=2,N_f=0,N=2`$). In this case (16) yields: $$\mathrm{\Omega }_{202}=42N_\mathrm{\Gamma }$$ (17) Thus if $`N_\mathrm{\Gamma }>2`$ then all of Feynman integrals are finite. Consider possible large momentum dependence of the vertex operator (5), (12), (13). In the approximation $`p^20`$ this vertex operator has the following form: $$\mathrm{\Gamma }^\mu h_1(0,q^2)\gamma ^\mu $$ (18) i.e. the vertex operator (2) without ”spin term”. Hence we might suppose that $$h_1(p^2,q^2)|_{\genfrac{}{}{0pt}{}{p^2\mathrm{}}{q^2\mathrm{}}}\frac{1}{\left(p+q\right)^{2\left(n1\right)}}$$ (19) in agreement with the components counting rules. $`h_2(p^2,q^2)`$ has analogical $`p^2,q^2`$–dependence. Thus with the supposition (19) in the case $`n=3`$ the Feynman integrals are finite for all diagrams. In the case $`n=2`$ ultraviolet divergence remain in the polarization operator. This work approach is not consistently one to the Feynman diagram calculations as well as the Feynman rules in pointlike QED where the vertex operator form is consequence of the Dirac equation with the interaction taking into account. It is to be supposed that the equation (4) and the vertex operator (5), (12), (13) could be obtained from multifermionic Green functions consideration. The information about the momentum dependence of the form factors could be found from concrete models of a composite particle with knowledge of the components interactions. The functions $`f(p^2),h_1(p^2,q^2),h_2(p^2,q^2)`$ for leptons could be obtained from preon models . Hence this problem is not a pure QED–problem as the electron structure cannot be determined by electrodynamics. Thus the fact of the vertex operator and the lepton propagator being in principle written in the form that allows for calculating QED–quantities without divergences has a principal character only. The QED–calculations with the account of small corrections, which are proportional to $`<r_{lept}^n>`$, is possible. The author thanks T. Kinoshita for valuable questions, that enabled deeper look on the problem, S.J. Brodsky for the support of this work and M. Tyntarev and M. Golod for their technical help.
warning/0001/physics0001005.html
ar5iv
text
# An asymptotic form of the reciprocity theorem with applications in x-ray scattering ## 1 Introduction The principle of reciprocity can be traced to Helmholtz in the field of acoustics. It states that everything else being equal the amplitude of a wave at a point $`A`$ due to a source at point $`B`$ is equal to the amplitude at $`B`$ due to a source at $`A`$. With its extension to electromagnetic waves by Lorentz and later to quantum mechanical amplitudes , the applicability to all sorts of fields was made manifest. Nowadays the principle is regarded as a symmetry of Green’s functions when the source point and the field point are reversed. This symmetry is actually quite general. As shown in the conditions of time-reversal invariance and hermiticity of the Hamiltonian are sufficient to guarantee reciprocity, but they are not necessary; in fact, reciprocity holds even in the presence of complex absorbing potentials. In the case of electromagnetic waves the only requirement is that the material medium be linear and described by symmetric permittivity and permeability tensors . This excludes plasmas and ferrite media in the presence of magnetic fields. In the field of x-ray optics the principle was used by von Laue to explain the diffraction patterns generated by sources within the crystal, the so-called Kossel lines . More recently there has been a widespread recognition that these interference patterns contain information not just about intensities but also about phases and can be thought of as holographic records from which real space images of the location of the internal sources can be reconstructed. Thus, under the modern name of ‘x-ray holography’ there has been a considerable revival of interest in this subject . However, powerful as it is, the usual formulation of the reciprocity principle suffers from a rather serious drawback: it refers to the exchange of source and field *points*. As a consequence, a careful application of the principle requires one to consider the emission of spherical waves which in crystalline media or even in the mere presence of plane boundaries, can be surprisingly difficult (recall *e.g.* studying the radiation by an antenna in the vicinity of the conducting surface of the Earth or of layered media ). Furthermore, one is typically interested in the asymptotic radiation fields so the relevant exchange should involve a source point *here* with a field point at *infinity*. These technical difficulties have not deterred the users of reciprocity from using the principle to make valuable predictions, but a high price has been paid. The required asymptotic limits are usually taken *verbally* and no accounts are given of where and how spherical waves are replaced by plane waves. Such sleights of hand, because skillfully performed, have not lead to wrong results, but intensities are predicted only up to undetermined proportionality factors and this excludes applications to classes of problems where absolute intensities are needed. Moreover one is left with the uneasy feeling that the validity of the predictions is justified mostly on the purely pragmatic grounds that for the problem at hand they seem to work which, again, limits applications to problems that are already familiar. The main goal of this paper (section 2) is to obtain a modified form of the reciprocity theorem that gives the asymptotic radiation fields directly and that accommodates plane waves and both point and extended sources in a natural way. Remarkably the resulting expressions, which include all the relevant proportionality factors and yield absolute, not just relative intensities, are very simple. For many problems the Asymptotic Reciprocity Theorem (ART) obtained here represents an improvement not only over the usual form of the reciprocity theorem but also over the method of Green’s functions. Computing the Green’s function requires solving a boundary value problems for spherical waves in the presence of plane boundaries and/or periodic media; this may well be an intractably difficult problem. Furthermore, a considerable effort is wasted by first obtaining both near fields and far fields and then discarding the uninteresting near fields. The ART is a shortcut that discards the near fields before, rather than after they are computed. To illustrate the power of the method we consider several applications. The first three (section 3) are brief pedagogical examples of increasing complexity. First the ART is used to calculate the fields radiated by an arbitrary prescribed source in vacuum; next as an application to scattering problems we reproduce the kinematical theory of diffraction by crystals. The third example, the radiation by a current located near a plane dielectric boundary, is straightforward when the ART is used but not if other methods are used. One must emphasize that what is new in these examples are not the results, but the method; the first two are standard textbook material, a special case of the third is treated in . As a more involved application of the ART, in section 4 we combine ideas from the three previous examples to study two other related scattering problems, the specular reflection of polarized x rays by a rough surface and by a continuously graded surface. The technique of the grazing-incidence reflection of x-rays has received considerable attention - from both the theoretical and the experimental sides as a means to obtain structural information about surfaces. The effect of surface roughness on the reflection is taken into account by multiplying the Fresnel reflectivity of an ideal sharp and planar surface by a “static Debye-Waller” factor. The problem is to calculate this corrective factor. The calculation has been carried out in several different approximations. The Rayleigh or Born approximation is satisfactory for rough surfaces with long lateral correlation lengths but for x-rays the situations of interest generally involve short lateral correlation lengths. Here other approximations such as the distorted-wave Born approximation and the Nevot-Croce approximation are used. For variations and interpolations between these two methods see , and for a generalization to surfaces with non-Gaussian roughness and to graded interfaces of arbitrary profile see . In these treatments ( is an exception) the x rays are treated as scalar waves. One expects this approximation to hold at grazing incidence but at higher incidence angles (*e.g.*, for soft x rays) its validity becomes increasingly questionable. Using a modified first Born approximation Dietrich and Haase took the vector character of the x rays into account but they point out that the validity of their approximation is not in general easy to assess and they restrict themselves to studying special interface profiles. In section 4 we study this problem using a different approximation; we use the ART to develop approximations of the Nevot-Croce type . There is, of course, a trivial polarization dependence that is already described by Fresnel formulas for the reflectivity of the ideal flat step surface. The question we address here is whether the “static Debye-Waller” factor shows any additional dependence on polarization. The final result is remarkably simple: the “static Debye-Waller” factor for the specularly reflected vector waves is the same for both polarizations and coincides with that for scalar waves. Finally, some brief concluding remarks are collected in section 5. ## 2 The reciprocity theorem and its asymptotic form We wish to calculate the asymptotic radiation fields $`\stackrel{}{E}`$ and $`\stackrel{}{H}`$ generated by a prescribed current $`\stackrel{}{J}(t,\stackrel{}{r})`$ located near or within a linear medium, $$D_i=\epsilon _{ij}E_j\text{and}B_i=\mu _{ij}H_j\text{ .}$$ We will assume that the tensors $`\epsilon _{ij}\left(\stackrel{}{r}\right)`$ and $`\mu _{ij}\left(\stackrel{}{r}\right)`$ are symmetric, but otherwise the situation remains quite general, the medium may have an irregular shape, or be inhomogeneous, crystalline or amorphous, absorbing, dispersive, etc. As in the usual deduction of the reciprocity theorem (see *e.g.*, ), we consider a second set of fields $`\stackrel{}{E}_c`$ and $`\stackrel{}{H}_c`$, which we will call the “connecting fields”, generated by a source $`\stackrel{}{J}_c`$ (see fig.1a). For simplicity we will also assume that all fields and sources are monochromatic $`\stackrel{}{E}=\stackrel{}{E}(\stackrel{}{r})e^{i\omega t}`$, $`\stackrel{}{J}=\stackrel{}{J}(\stackrel{}{r})e^{i\omega t}`$, etc. For linear media this is not a restriction. From Maxwell’s equations $$\times \stackrel{}{E}=iK\stackrel{}{B}\text{ and }\times \stackrel{}{H}=iK\stackrel{}{D}+\frac{4\pi }{c}\stackrel{}{J},$$ (1) where $`K\omega /c`$, one easily obtains the following identity $$\left(\stackrel{}{E}\times \stackrel{}{H}_c\stackrel{}{E}_c\times \stackrel{}{H}\right)=\frac{4\pi }{c}\left(\stackrel{}{E}_c\stackrel{}{J}\stackrel{}{E}\stackrel{}{J}_c\right),$$ which, on integrating over a large volume $`V`$ bounded by the surface $`S`$, can be rewritten as $$_S\left(\stackrel{}{E}\times \stackrel{}{H}_c\stackrel{}{E}_c\times \stackrel{}{H}\right)𝑑\stackrel{}{s}=\frac{4\pi }{c}_V\left(\stackrel{}{E}_c\stackrel{}{J}\stackrel{}{E}\stackrel{}{J}_c\right)𝑑v.$$ (2) This expression simplifies if one deals with point sources. For example, consider oscillating point dipoles $`\stackrel{}{p}_oe^{i\omega t}`$ and $`\stackrel{}{p}_ce^{i\omega t}`$, located at $`\stackrel{}{r}_o`$ and $`\stackrel{}{r}_c`$ respectively. The current density $`\stackrel{}{J}`$ is given by $`\stackrel{}{J}=i\omega \stackrel{}{p}_o\delta \left(\stackrel{}{r}\stackrel{}{r}_o\right)e^{i\omega t}`$ and $`\stackrel{}{J}_c`$ is given by an analogous expression. Further simplification is achieved if one assumes that the surface $`S`$ is so remote that the surface integral is negligibly small , then $$\stackrel{}{E}_c\left(\stackrel{}{r}_o\right)\stackrel{}{p}_o=\stackrel{}{E}\left(\stackrel{}{r}_c\right)\stackrel{}{p}_c\text{ .}$$ (3) This is the usual form of the reciprocity theorem; it says that if we know $`\stackrel{}{E}_c`$ at the location of $`\stackrel{}{p}_o`$ we can calculate $`\stackrel{}{E}`$ at the location of $`\stackrel{}{p}_c`$. This elegant result takes us a long way toward a final answer for $`\stackrel{}{E}`$, but the remaining problem of calculating $`\stackrel{}{E}_c`$, that is, the calculation of how the spherical wave generated by $`\stackrel{}{p}_c`$ is scattered by the medium, can still be too difficult. A more useful version of the theorem can be obtained once one realizes that the connecting field is merely a tool that codifies information about the influence of the non-trivial medium. Above, the field $`\stackrel{}{E}_c`$ has been introduced by first specifying a source $`\stackrel{}{J}_c`$, but clearly this is an unnecessary additional complication. In fact, since the most convenient $`\stackrel{}{J}_c`$ is that which results in the simplest $`\stackrel{}{E}_c`$ it is best to focus attention directly on the field rather than its source. Thus we move $`\stackrel{}{J}_c`$ outside the surface $`S`$, to infinity (see fig.1b) so that throughout the volume $`V`$ the connecting field $`\stackrel{}{E}_c`$ is a pure radiation field. Furthermore, let the surface $`S`$ itself be so distant that on $`S`$ itself both $`\stackrel{}{E}`$ and $`\stackrel{}{E}_c`$ are *vacuum* radiation fields. Then $$_S\left(\stackrel{}{E}\times (\times \stackrel{}{E}_c)\stackrel{}{E}_c\times (\times \stackrel{}{E})\right)𝑑\stackrel{}{s}=\frac{4\pi iK}{c}_V\stackrel{}{E}_c\stackrel{}{J}𝑑v.$$ (4) At this point it is not yet clear that this form of the reciprocity theorem is simpler than eq. (3) but one remarkable feature can already be seen: eq. (4) relates the field $`\stackrel{}{E}`$ at a distant surface $`S`$ to its source $`\stackrel{}{J}`$ within a nontrivial medium *without* having to calculate $`\stackrel{}{E}`$ in the vicinity of $`\stackrel{}{J}`$. The “connection” between the distant radiation field $`\stackrel{}{E}`$ and its source $`\stackrel{}{J}`$ is achieved through the much simpler (i.e., hopefully calculable) “connecting” field $`\stackrel{}{E}_c`$. To bring eq. (4) into a form that is manifestly simpler than (3) the surface $`S`$ is chosen as a cube with edges of length $`L\mathrm{}`$. In fig.1b the upper face, defined by a constant $`z`$ coordinate, $`z=z_+`$, has been singled out as $`S_+`$, the remaining seven faces are denoted $`S^{^{}}`$. On the upper face $`S_+`$ we write the field $`\stackrel{}{E}`$ as a superposition of outgoing plane waves of wave vector $`\stackrel{}{k}`$ satisfying $`\stackrel{}{k}\stackrel{}{k}=\omega ^2/c^2=K^2`$ and $`k_z>0`$, $$\stackrel{}{E}\left(\stackrel{}{r}\right)=\underset{k_z>0}{}\frac{d^3k}{\left(2\pi \right)^3}\mathrm{\hspace{0.17em}2}\pi \delta \left(kK\right)\stackrel{}{E}(\stackrel{}{k})e^{i\stackrel{}{k}\stackrel{}{r}},$$ (5) where, in a self-explanatory notation, $`\stackrel{}{k}\stackrel{}{r}=\stackrel{}{k}_{}\stackrel{}{r}_{}+k_zz_+`$. It is here, by the very act of writing $`\stackrel{}{E}`$ in this form, that the asymptotic limit of discarding near fields is being taken. For $`z<<z_+`$ additional terms describing the near fields should be included. The integral over $`dk_z`$ is most easily done using $$\delta \left(kK\right)=\frac{K}{k_z}\left[\delta \left(k_z\sqrt{K^2k_{}^2}\right)\delta \left(k_z+\sqrt{K^2k_{}^2}\right)\right].$$ (6) The result is $$\stackrel{}{E}\left(\stackrel{}{r}\right)=\frac{d^2k_{}}{\left(2\pi \right)^2}\frac{K}{k_z}\stackrel{}{E}(\stackrel{}{k})e^{i\stackrel{}{k}\stackrel{}{r}},$$ (7) where $`k_z=+\sqrt{K^2k_{}^2}`$. The choice of the connecting field $`\stackrel{}{E}_c`$ is dictated purely by convenience. A particularly good choice for $`\stackrel{}{E}_c`$ is the superposition of an incoming plane wave of unit amplitude (we use $`\mathrm{^}`$ to denote vectors of unit length) and wave vector $`\stackrel{}{k}_c`$, with $`\stackrel{}{k}_c\stackrel{}{k}_c=K^2`$ and $`k_{cz}<0`$, plus all the waves scattered by the medium, $$\stackrel{}{E}_c\left(\stackrel{}{r}\right)=\widehat{e}_ce^{i\stackrel{}{k}_c\stackrel{}{r}}+\stackrel{}{E}_c^{^{}}\left(\stackrel{}{r}\right).$$ (8) On $`S_+`$, the scattered field $`\stackrel{}{E}_c^{^{}}`$ is a superposition of outgoing plane waves and is given also in a form analogous to eq. (7), $$\stackrel{}{E}_c^{^{}}\left(\stackrel{}{r}\right)=\frac{d^2k_{}^{^{}}}{\left(2\pi \right)^2}\frac{K}{k_z^{^{}}}\stackrel{}{E}_c^{^{}}(\stackrel{}{k}^{^{}})e^{i\stackrel{}{k}^{^{}}\stackrel{}{r}},$$ (9) with $`\stackrel{}{k}^{^{}}\stackrel{}{r}=\stackrel{}{k}_{}^{^{}}\stackrel{}{r}_{}+k_z^{^{}}z_+`$ and $`k_z^{^{}}=+\sqrt{K^2k_{}^{}_{}{}^{}2}`$. Now we are ready to calculate the surface integral on the left hand side of eq. (4). Substituting (8) into (4) the integral over $`S_+`$ separates into two terms, one due to the incoming plane wave $`\widehat{e}_ce^{i\stackrel{}{k}_c\stackrel{}{r}}`$, and the other due to the scattered waves $`\stackrel{}{E}_c^{^{}}\left(\stackrel{}{r}\right)`$. The first term is $$I_1=_{S_+}𝑑x𝑑y\widehat{e}_z\left(\stackrel{}{E}\times (i\stackrel{}{k}_c\times \widehat{e}_c)\widehat{e}_c\times (\times \stackrel{}{E})\right)e^{i\stackrel{}{k}_c\stackrel{}{r}}\text{,}$$ (10) and substituting (7) its evaluation is straightforward. The integral over $`dxdy`$ yields $`(2\pi )^2\delta (\stackrel{}{k}_{}+\stackrel{}{k}_c)`$. Since $`k^2=k_c^2=K^2`$, $`k_{cz}<0`$ and $`k_z>0`$ this implies that the plane waves superposed in (7) yield a vanishing contribution except when $`\stackrel{}{k}=\stackrel{}{k}_c`$. Thus, $$I_1=2iK\widehat{e}_c\stackrel{}{E}(\stackrel{}{k}_c)$$ (11) The contribution of the scattered waves $`\stackrel{}{E}_c^{^{}}\left(\stackrel{}{r}\right)`$, $$I_2=_{S_+}𝑑x𝑑y\widehat{e}_z\left(\stackrel{}{E}\times (\times \stackrel{}{E}_c^{^{}})\stackrel{}{E}_c^{^{}}\times (\times \stackrel{}{E})\right),$$ (12) is calculated in a similar way. Substitute (7) and (9) and integrate over $`dxdy`$ to obtain a delta function. This eliminates all Fourier components except those with $`\stackrel{}{k}_{}=\stackrel{}{k}_{}^{^{}}`$. Since $`k^2=k^{}_{}{}^{}2=K^2`$, and both $`k_z,k_z^{^{}}>0`$ this implies $`k_z=`$ $`k_z^{^{}}>0`$. Thus, $$I_2=\frac{d^2k_{}^{^{}}}{\left(2\pi \right)^2}\frac{K^2}{k_z^^2}\widehat{e}_z\left[\stackrel{}{E}(\stackrel{}{k})\times \left(i\stackrel{}{k}^{^{}}\times \stackrel{}{E}_c^{^{}}(\stackrel{}{k}^{^{}})\right)\stackrel{}{E}_c^{^{}}(\stackrel{}{k}^{^{}})\times \left(i\stackrel{}{k}\times \stackrel{}{E}(\stackrel{}{k})\right)\right],$$ (13) where $`\stackrel{}{k}=\stackrel{}{k}^{^{}}+2k_z^{^{}}\widehat{e}_z`$. Further manipulation using $`\stackrel{}{k}^{^{}}\stackrel{}{E}_c^{^{}}(\stackrel{}{k}^{^{}})=\stackrel{}{k}\stackrel{}{E}(\stackrel{}{k})=0`$ gives $`\widehat{e}_z[\mathrm{}]=0`$, so that $$I_2=0.$$ (14) According to eq. (11) and (14) the only contributions to the surface integral over the distant plane $`S_+`$ come from products of outgoing with incoming waves. Products of two outgoing waves yield vanishing contributions. This result applies also to the remaining seven faces of the cube $`S`$. Since on each of these faces there are only outgoing waves we find that the integral over $`S^{^{}}`$ makes no contribution to the left hand side of (4). Incidentally, this argument completes our previously unfinished deduction of the usual form of the reciprocity theorem, eq. (3): if both sources $`\stackrel{}{J}`$ and $`\stackrel{}{J}_c`$ are internal to the surface $`S`$ the surface integral in eq. (2) vanishes because it only involves products of outgoing waves. Substituting (11) into (4) leads us to the main result of this paper, the asymptotic reciprocity theorem, $$\widehat{e}\stackrel{}{E}(\stackrel{}{k})=\frac{2\pi }{c}_V\stackrel{}{E}_c\stackrel{}{J}𝑑v.$$ (15) In words: *The field* $`\stackrel{}{E}(\stackrel{}{k})`$ *radiated in a direction* $`\stackrel{}{k}`$ *with a certain polarization* $`\widehat{e}`$ *is* $`2\pi /c`$ *times the “component” of the source* $`\stackrel{}{J}(\stackrel{}{r})`$ *“along” a connecting field* $`\stackrel{}{E}_c(\stackrel{}{r})`$ *with incoming wave vector* $`\stackrel{}{k}_c=\stackrel{}{k}`$ *and polarization* $`\widehat{e}_c=\widehat{e}`$. Typically one is interested in the intensity radiated into a solid angle $`d\mathrm{\Omega }`$; since the amplitude $`\stackrel{}{E}(\stackrel{}{k})`$ that appears in (7) and (15) is not quite the Fourier transform of $`\stackrel{}{E}(\stackrel{}{r})`$ it may be useful to derive an explicit expression for $`dW/d\mathrm{\Omega }`$. The total power radiated through the plane $`S_+`$ is given by the flux of the time-averaged Poynting vector, $`\frac{c}{8\pi }\mathrm{R}e[\stackrel{}{E}\times \stackrel{}{B}^{}]`$, $$W=d^2x_{}\frac{c}{8\pi }\mathrm{R}e[\stackrel{}{E}\times \stackrel{}{B}^{}]\widehat{e}_z=𝑑\mathrm{\Omega }\frac{dW}{d\mathrm{\Omega }}$$ (16) Using (7) and $`d^2k_{}=k_{}dk_{}d\varphi =Kk_zd\mathrm{\Omega }`$ (where $`\varphi `$ is the usual azimuthal angle about the $`z`$ axis) we get $$W=\frac{c}{8\pi }\frac{d^2k_{}}{\left(2\pi \right)^2}\frac{K}{k_z}\stackrel{}{E}(\stackrel{}{k})\stackrel{}{E}^{}(\stackrel{}{k}),$$ (17) so that $$\frac{dW}{d\mathrm{\Omega }}=\frac{c}{8\pi }\left(\frac{K}{2\pi }\right)^2\stackrel{}{E}(\stackrel{}{k})\stackrel{}{E}^{}(\stackrel{}{k}).$$ (18) In the next section we offer a few illustrative examples of the ART in action. ## 3 Some simple examples The ART, eq. (15), holds for an arbitrary linear medium. In particular it holds if the medium is vacuum. Our first trivial example is the radiation by a prescribed current in vacuum. Next, to show that the ART can be used to study scattering problems we deal with another equally trivial example, the kinematical theory of diffraction by crystals. The third example, the radiation by currents located near a dielectric boundary, is also straightforward. What is remarkable here is the ease with which the results are obtained compared to conventional methods . ### 3.1 Radiation in vacuum In this case the connecting field is just an incoming plane wave, $`\stackrel{}{E}_c\left(\stackrel{}{r}\right)=\widehat{e}_ce^{i\stackrel{}{k}_c\stackrel{}{r}}`$. The ART, eq. (15), gives the radiated field with polarization $`\widehat{e}=\widehat{e}_c`$ as $$\widehat{e}\stackrel{}{E}(\stackrel{}{k})=\frac{2\pi }{c}_V\stackrel{}{J}(\stackrel{}{r})\widehat{e}e^{i\stackrel{}{k}\stackrel{}{r}}𝑑v=\frac{2\pi }{c}\widehat{e}\stackrel{}{J}(\stackrel{}{k}),$$ (19) so that $$\stackrel{}{E}(\stackrel{}{k})=\frac{2\pi }{c}\widehat{k}\times (\widehat{k}\times \stackrel{}{J}(\stackrel{}{k})).$$ (20) The radiated power, eq. (18), is $$\frac{dW}{d\mathrm{\Omega }}=\frac{K^2}{8\pi c}\left|\widehat{k}\times (\widehat{k}\times \stackrel{}{J}(\stackrel{}{k}))\right|^2,$$ (21) as expected. (For radiation by a point dipole just substitute $`\stackrel{}{J}(\stackrel{}{k})=icK\stackrel{}{p}`$.) ### 3.2 Bragg diffraction Consider a crystal described by its dielectric susceptibility $`\chi (\stackrel{}{r})`$ which for x rays is quite small (typically about $`10^5`$ or less). An incident plane wave $`\stackrel{}{E}_oe^{i\stackrel{}{k}_o\stackrel{}{r}}`$ induces a current $$\stackrel{}{J}(\stackrel{}{r})=i\omega \stackrel{}{P}(\stackrel{}{r})=\frac{i\omega }{4\pi }\chi (\stackrel{}{r})\stackrel{}{E}_oe^{i\stackrel{}{k}_o\stackrel{}{r}},$$ (22) which radiates. The connecting field needed to calculate this radiation is a simple incoming plane wave, $`\stackrel{}{E}_c\left(\stackrel{}{r}\right)=\widehat{e}_ce^{i\stackrel{}{k}_c\stackrel{}{r}}`$, and the ART, eq. (15), gives the radiated field as $$\stackrel{}{E}(\stackrel{}{k})=\frac{i\omega }{2c}\widehat{k}\times (\widehat{k}\times \stackrel{}{E}_o)\chi (\stackrel{}{k}\stackrel{}{k}_o).$$ (23) The scattered field is proportional to the Fourier transform of the susceptibility of the medium; for a periodic medium this is Bragg diffraction. ### 3.3 Radiation in the vicinity of a reflecting surface Consider a current $`\stackrel{}{J}_{in}(\stackrel{}{r})`$ located within a uniform medium with dielectric susceptibility $`\chi _0`$ occupying the region $`z<0`$ (see fig. 2). To calculate the radiation in the direction $`\stackrel{}{k}`$ with polarization $`\widehat{e}`$ we choose as connecting field an incoming plane wave with wave vector $`\stackrel{}{k}_c=\stackrel{}{k}`$ and unit amplitude $`\widehat{e}_c=\widehat{e}`$ plus the corresponding reflected and transmitted waves, $$\stackrel{}{E}_c\left(\stackrel{}{r}\right)=\{\begin{array}{cc}\widehat{e}_ce^{i\stackrel{}{k}_c\stackrel{}{r}}+\stackrel{}{\epsilon }_{cr}e^{i\stackrel{}{k}_{cr}\stackrel{}{r}},\hfill & \text{for }z>0\hfill \\ \stackrel{}{\epsilon }_{ct}e^{i\stackrel{}{k}_{ct}\stackrel{}{r}},\hfill & \text{for }z<0\hfill \end{array}$$ (24) The various wave vectors are given by $$\stackrel{}{k}_c=K\mathrm{cos}\theta \widehat{e}_xq\widehat{e}_z=\stackrel{}{k},$$ (25) $$\stackrel{}{k}_{cr}=K\mathrm{cos}\theta \widehat{e}_x+q\widehat{e}_z,$$ (26) $$\stackrel{}{k}_{ct}=K\mathrm{cos}\theta \widehat{e}_x\overline{q}\widehat{e}_z,$$ (27) where $`K=\omega /c`$, and the normal components $`q`$ and $`\overline{q}`$ are given by $$q=K\mathrm{sin}\theta \text{and }\overline{q}=K\left(\mathrm{sin}^2\theta +\chi _0\right)^{1/2}.$$ (28) The amplitudes $`\stackrel{}{\epsilon }_{cr}`$ and $`\stackrel{}{\epsilon }_{ct}`$ of the reflected and transmitted waves are given by the Fresnel expressions $$\stackrel{}{\epsilon }_{cr}=r_s\widehat{e}_{cr}\text{where }r_s=\frac{q\overline{q}}{q+\overline{q}}\frac{\widehat{e}_{cr}\widehat{e}_{ct}}{\widehat{e}_c\widehat{e}_{ct}},$$ (29) and $$\stackrel{}{\epsilon }_{ct}=t_s\widehat{e}_{ct}\text{where }t_s=\frac{2q}{q+\overline{q}}\frac{1}{\widehat{e}_c\widehat{e}_{ct}}.$$ (30) ($`\widehat{e}_{cr}`$ and $`\widehat{e}_{ct}`$ are unit vectors describing the polarization of the specular reflected and transmitted waves.) Then for a source $`\stackrel{}{J}_{in}(\stackrel{}{r})`$ located within the medium, the ART, eq.(15), gives the radiated field as $$\widehat{e}\stackrel{}{E}(\stackrel{}{k})=\frac{2\pi }{c}_{z<0}\stackrel{}{J}_{in}(\stackrel{}{r})\stackrel{}{\epsilon }_{ct}e^{i\stackrel{}{k}_{ct}\stackrel{}{r}}𝑑v.$$ (31) On the other hand, had the source $`\stackrel{}{J}_{out}(\stackrel{}{r})`$ been located outside the dielectric medium ($`z>0`$) the corresponding radiated field would be $$\widehat{e}\stackrel{}{E}(\stackrel{}{k})=\frac{2\pi }{c}_{z>0}\stackrel{}{J}_{out}(\stackrel{}{r})\left(\widehat{e}_ce^{i\stackrel{}{k}_c\stackrel{}{r}}+\stackrel{}{\epsilon }_{cr}e^{i\stackrel{}{k}_{cr}\stackrel{}{r}}\right)𝑑v.$$ (32) For an oscillating dipole on the $`z`$ axis, $`\stackrel{}{J}(\stackrel{}{r})=i\omega \stackrel{}{p}\delta \left(\stackrel{}{r}z_p\widehat{e}_z\right)`$, eq.(31) and (32) give $$\widehat{e}\stackrel{}{E}(\stackrel{}{k})=\{\begin{array}{cc}2\pi iK\left(\stackrel{}{p}\widehat{e}e^{iqz_p}+\stackrel{}{p}\widehat{e}_{cr}r_se^{iqz_p}\right)\hfill & \text{if }z_p>0\hfill \\ 2\pi iK\stackrel{}{p}\widehat{e}_{ct}t_se^{i\overline{q}z_p}\hfill & \text{if }z_p<0.\hfill \end{array}$$ (33) The power radiated with polarization $`\widehat{e}`$, eq.(18), is $$\frac{dW}{d\mathrm{\Omega }}=\left[\frac{c}{8\pi }K^4\left(\widehat{e}\stackrel{}{p}\right)^2\right]\left|1+\frac{\widehat{e}_{cr}\stackrel{}{p}}{\widehat{e}\stackrel{}{p}}r_se^{2iqz_p}\right|^2\text{for }z_p>0,$$ (34) and $$\frac{dW}{d\mathrm{\Omega }}=\left[\frac{c}{8\pi }K^4\left(\widehat{e}\stackrel{}{p}\right)^2\right]\left|\frac{\widehat{e}_{ct}\stackrel{}{p}}{\widehat{e}\stackrel{}{p}}t_se^{i\overline{q}z_p}\right|^2\text{for }z_p<0.$$ (35) In these two expressions we can recognize the first factor (in square brackets) as the power radiated by a dipole in vacuum. The second factor accounts for the presence of the dielectric medium. ## 4 Specular reflection of polarized x rays In this section ideas from the three previous examples are combined to study two similar and considerably more involved scattering problems, the specular reflection of polarized x rays by a rough surface and by graded interfaces. We show that within approximations of the Nevot-Croce type grading and roughness affect the specular reflectivity in a manner that is independent of the polarization of the incident radiation. ### 4.1 Reflection by rough surfaces The dielectric susceptibility $`\chi (\stackrel{}{r})`$ that describes the rough surface from which we wish to scatter x rays is given by $$\chi (x,y,z)=\{\begin{array}{cc}0\hfill & \text{for }z>\zeta (x,y)\hfill \\ \chi _0\hfill & \text{for }z<\zeta (x,y)\hfill \end{array}$$ (36) where the height $`\zeta (x,y)`$, is a Gaussian random variable with zero mean, $`\zeta =0`$, and variance $`\zeta ^2=\sigma ^2`$ (see fig. 3). To apply the ART it is convenient to rewrite $`\chi \left(\stackrel{}{r}\right)`$ as $$\chi \left(\stackrel{}{r}\right)=\chi _s\left(\stackrel{}{r}\right)+\delta \chi \left(\stackrel{}{r}\right),$$ (37) where $`\chi _s\left(\stackrel{}{r}\right)`$ represents a medium with an ideally flat surface at $`z_0`$, $$\chi _s\left(\stackrel{}{r}\right)=\{\begin{array}{cc}0\hfill & \text{for }z>z_0\hfill \\ \chi _0\hfill & \text{for }z<z_0\hfill \end{array}$$ (38) and $`\delta \chi \left(\stackrel{}{r}\right)`$ represents the roughness. Let $`\widehat{e}_0e^{i\stackrel{}{k}_0\stackrel{}{r}}`$ be the incident field. The total scattered field $`\stackrel{}{\epsilon }(\stackrel{}{r})`$ includes the wave $`\stackrel{}{\epsilon }_s(\stackrel{}{r})`$ specularly reflected by the step $`\chi _s\left(\stackrel{}{r}\right)`$ plus waves $`\delta \stackrel{}{\epsilon }\left(\stackrel{}{r}\right)`$ scattered by $`\delta \chi \left(\stackrel{}{r}\right)`$ $$\stackrel{}{\epsilon }(\stackrel{}{r})=\stackrel{}{\epsilon }_s(\stackrel{}{r})+\delta \stackrel{}{\epsilon }\left(\stackrel{}{r}\right).$$ (39) The first term on the right is $$\stackrel{}{\epsilon }_s(\stackrel{}{r})=\widehat{e}_rr_se^{2iqz_0}e^{i\stackrel{}{k}_r\stackrel{}{r}},$$ (40) where $$r_s=\frac{q\overline{q}}{q+\overline{q}}\frac{\widehat{e}_r\widehat{e}_t}{\widehat{e}_0\widehat{e}_t}$$ (41) ($`\widehat{e}_r`$ and $`\widehat{e}_t`$ are unit vectors describing the polarization of the specular reflected and transmitted waves). The second contribution in eq.(39), the field $`\delta \stackrel{}{\epsilon }\left(\stackrel{}{r}\right)`$ includes a specular component plus diffusely scattered and evanescent waves, $$\delta \stackrel{}{\epsilon }\left(\stackrel{}{r}\right)=\delta \epsilon _r\widehat{e}_re^{i\stackrel{}{k}_r\stackrel{}{r}}+\delta \stackrel{}{\epsilon }_d(\stackrel{}{r}).$$ (42) Using eq.(7) this may be written as $$\delta \stackrel{}{\epsilon }\left(\stackrel{}{r}\right)=\frac{d^2k_{}^{^{}}}{\left(2\pi \right)^2}\frac{K}{k_z^{^{}}}\delta \stackrel{}{\epsilon }(\stackrel{}{k}^{^{}})e^{i\stackrel{}{k}^{^{}}\stackrel{}{r}},$$ (43) where $$\delta \stackrel{}{\epsilon }(\stackrel{}{k}^{^{}})=\frac{k_z^{^{}}}{K}\delta \epsilon _r\widehat{e}_r\left(2\pi \right)^2\delta (\stackrel{}{k}_{}^{^{}}\stackrel{}{k}_0)+\delta \stackrel{}{\epsilon }_d(\stackrel{}{k}^{^{}}).$$ (44) To calculate $`\delta \stackrel{}{\epsilon }\left(\stackrel{}{r}\right)`$ we can proceed exactly as in the previous section (3.3): $`\delta \stackrel{}{\epsilon }\left(\stackrel{}{r}\right)`$ is the field radiated by a current $`\delta \stackrel{}{J}(\stackrel{}{r})`$ in the presence of the medium $`\chi _s\left(\stackrel{}{r}\right)`$. The current $$\delta \stackrel{}{J}(\stackrel{}{r})=\frac{i\omega }{4\pi }\delta \chi (\stackrel{}{r})\stackrel{}{E}\left(\stackrel{}{r}\right),$$ (45) originates in the polarization of the roughness $`\delta \chi (\stackrel{}{r})`$ by the total electric field $`\stackrel{}{E}(\stackrel{}{r})`$ due to the incident and all scattered waves, including those generated by the roughness itself. Thus, the challenge here is that the field $`\stackrel{}{E}\left(\stackrel{}{r}\right)`$ is itself unknown; an approximation for it must be obtained as part of our solution. We can exploit the arbitrariness in the separation of $`\chi \left(\stackrel{}{r}\right)`$ into $`\chi _s\left(\stackrel{}{r}\right)`$ plus $`\delta \chi \left(\stackrel{}{r}\right)`$ to suggest a self-consistent approximation for $`\stackrel{}{E}`$. Suppose we choose $`z_0`$ positive and considerably larger than the roughness $`\sigma `$ (see fig. 3). Then $`\delta \chi \left(\stackrel{}{r}\right)`$ represents a fictitious overlayer that extends well into the vacuum; the sign of $`\delta \chi \left(\stackrel{}{r}\right)`$ is opposite to that of $`\chi _s\left(\stackrel{}{r}\right)`$ and in the vicinity of $`z_0`$ they completely cancel out. The field $`\delta \stackrel{}{\epsilon }(\stackrel{}{k})`$ in a direction $`\stackrel{}{k}`$ with polarization $`\widehat{e}`$ is given by eq.(31) $$\widehat{e}\delta \stackrel{}{\epsilon }(\stackrel{}{k})=\frac{iK}{2}𝑑v\delta \chi (\stackrel{}{r})\stackrel{}{E}\left(\stackrel{}{r}\right)\stackrel{}{\epsilon }_{ct}e^{i\stackrel{}{k}_{ct}\stackrel{}{r}},$$ (46) where the connecting field is precisely as in eqs.(24)-(30) except for phase shifts due to the reflecting surface being at $`z_0`$, $$\stackrel{}{\epsilon }_{cr}=e^{2iqz_0}r_s\widehat{e}_{cr}\text{and}\stackrel{}{\epsilon }_{ct}=e^{i(\overline{q}q)z_0}t_s\widehat{e}_{ct}.$$ (47) The reason behind the somewhat surprising choice for $`z_0`$ will now become clear: slightly above $`z_0`$, in vacuum, the exact field is $$\stackrel{}{E}\left(\stackrel{}{r}\right)=\widehat{e}_0e^{i\stackrel{}{k}_0\stackrel{}{r}}+\stackrel{}{\epsilon }(\stackrel{}{r})=\widehat{e}_0e^{i\stackrel{}{k}_0\stackrel{}{r}}+\stackrel{}{\epsilon }_s(\stackrel{}{r})+\delta \stackrel{}{\epsilon }\left(\stackrel{}{r}\right),$$ (48) but slightly below $`z_0`$ and, in fact, over all of the extension occupied by $`\delta \chi \left(\stackrel{}{r}\right)`$, we are also in vacuum ($`\delta \chi \left(\stackrel{}{r}\right)`$ and $`\chi _s\left(\stackrel{}{r}\right)`$ cancel each other) and therefore $`\stackrel{}{E}\left(\stackrel{}{r}\right)`$ is given by the same expression (48). The last term $`\delta \stackrel{}{\epsilon }\left(\stackrel{}{r}\right)`$, given by (42), includes some weak diffusely scattered and evanescent waves $`\delta \stackrel{}{\epsilon }_d(\stackrel{}{r})`$. Our approximation consists of neglecting them. Therefore, $$\stackrel{}{E}\left(\stackrel{}{r}\right)\widehat{e}_0e^{i\stackrel{}{k}_0\stackrel{}{r}}+r\widehat{e}_re^{i\stackrel{}{k}_r\stackrel{}{r}},$$ (49) where the specular reflections by $`\chi _s\left(\stackrel{}{r}\right)`$ and $`\delta \chi \left(\stackrel{}{r}\right)`$ have been combined into the single, and still unknown, reflection coefficient $`r`$, $$r=r_se^{2iqz_0}+\delta \epsilon _r.$$ (50) Substituting into eq.(46) yields $$\widehat{e}\delta \stackrel{}{\epsilon }(\stackrel{}{k})=\frac{iK\chi _0}{2}𝑑x𝑑y_{\zeta (x,y)}^{z_0}𝑑z\left[\widehat{e}_0e^{i\stackrel{}{k}_0\stackrel{}{r}}+r\widehat{e}_re^{i\stackrel{}{k}_r\stackrel{}{r}}\right]\stackrel{}{\epsilon }_{ct}e^{i\stackrel{}{k}_{ct}\stackrel{}{r}}.$$ (51) From now on we focus our attention on the specularly reflected component; let $`\widehat{e}=\widehat{e}_c=\widehat{e}_r`$, $`\widehat{e}_{cr}=\widehat{e}_0`$, $`\stackrel{}{k}=\stackrel{}{k}_r=\stackrel{}{k}_c`$. Substituting eq.(44) into the left hand side ($`l.h.s.`$), using $`\left(2\pi \right)^2\delta (k_{}k_0)=\left(2\pi \right)^2\delta (0)=𝑑x𝑑y`$ we get $$l.h.s.=\frac{q}{K}\delta \epsilon _r\left(2\pi \right)^2\delta (0)=\frac{q}{K}(rr_se^{2iqz_0})dxdy.$$ (52) This shows that the unknown reflection coefficient $`r`$ we want to calculate appears in both the left and the right hand sides of (51), as part of the radiated field and also as part of the field that induces the source; eq.(51) permits a self-consistent calculation of $`r`$. The integral over $`dz`$ in the right hand side of eq.(51) is elementary and the remaining integral over $`dxdy`$ is performed using the identity $$\frac{𝑑x𝑑ye^{iQ\zeta (x,y)}}{𝑑x𝑑y}=e^{iQ\zeta }=e^{Q^2\sigma ^2/2},$$ (53) where $`\zeta `$ is a Gaussian random variable with zero mean, $`\zeta =0`$, and variance $`\zeta ^2=\sigma ^2`$. The right hand side ($`r.h.s.`$) of eq.(51) becomes $$r.h.s.=\frac{K\chi _0}{2}(dxdy)\{\frac{\widehat{e}_0\stackrel{}{\epsilon }_{ct}}{q+\overline{q}}[e^{i(q+\overline{q})z_0}e^{(q+\overline{q})^2\sigma ^2/2}]$$ $$r\frac{\widehat{e}_r\stackrel{}{\epsilon }_{ct}}{q\overline{q}}[e^{i(q\overline{q})z_0}e^{(q\overline{q})^2\sigma ^2/2}]\},$$ (54) which can be further rewritten by substituting $`\stackrel{}{\epsilon }_{ct}`$ as given by eq.(47), and using $`\overline{q}^2q^2=K^2\chi _0`$, and $$\frac{\widehat{e}_0\widehat{e}_{ct}}{\widehat{e}\widehat{e}_{ct}}=\frac{\widehat{e}_{cr}\widehat{e}_{ct}}{\widehat{e}\widehat{e}_{ct}}=\frac{\widehat{e}_r\widehat{e}_t}{\widehat{e}_0\widehat{e}_t}\text{and}\frac{\widehat{e}_r\widehat{e}_{ct}}{\widehat{e}\widehat{e}_{ct}}=1.$$ (55) Finally, equating eq.(52) to (54) yields a self-consistent approximation to $`r`$, $$r=\frac{q\overline{q}}{q+\overline{q}}\frac{\widehat{e}_r\widehat{e}_t}{\widehat{e}_0\widehat{e}_t}e^{2q\overline{q}\sigma ^2}=r_se^{2q\overline{q}\sigma ^2}.$$ (56) This coincides exactly with the Nevot-Croce result for the polarization $`\widehat{e}_0=\widehat{e}_t=`$ $`\widehat{e}_r`$ for which the ratio $`\widehat{e}_0\widehat{e}_t/`$ $`\widehat{e}_r\widehat{e}_t`$ is unity, and provides the correct generalization to all polarizations. According to this approximation the specular reflection coefficient $`r`$ has no polarization dependence beyond that already implicit in the reflection coefficient $`r_s`$ for the ideal flat step surface; the “static Debye-Waller” factor $`exp(2q\overline{q}\sigma ^2)`$ is polarization independent. Notice that any possible dependence on the arbitrary choice of $`z_0`$ has cancelled out. ### 4.2 Reflection by smoothly graded surfaces The problem of scattering by a smoothly graded interface is similar and somewhat simpler. Here the susceptibility $`\chi (z)`$ depends only on the normal coordinate $`z`$ and not on the transverse coordinates $`x`$ and $`y`$. This implies that the tangential component of momentum is conserved in the scattering; there are no diffuse waves, there is only specular scattering. As before, it is convenient to separate $`\chi \left(z\right)`$ into $$\chi \left(z\right)=\chi _s\left(z\right)+\delta \chi \left(z\right),$$ (57) where $`\chi _s\left(z\right)`$ represents an ideally flat surface at $`z_0`$, $$\chi _s\left(z\right)=\{\begin{array}{cc}0\hfill & \text{for }z>z_0\hfill \\ \chi _0\hfill & \text{for }z<z_0\hfill \end{array}$$ (58) and $`\delta \chi \left(z\right)`$ is an overlayer (see fig.4) describing the smooth transition from bulk to vacuum. Let $`\widehat{e}_0e^{i\stackrel{}{k}_0\stackrel{}{r}}`$ be the incident field. The total scattered field $`\stackrel{}{\epsilon }(\stackrel{}{r})`$, eq.(39), $$\stackrel{}{\epsilon }(\stackrel{}{r})=\stackrel{}{\epsilon }_s(\stackrel{}{r})+\delta \stackrel{}{\epsilon }\left(\stackrel{}{r}\right).$$ (59) includes the wave $`\stackrel{}{\epsilon }_s(\stackrel{}{r})`$ reflected by the step $`\chi _s\left(z\right)`$, eq.(40), plus waves $`\delta \stackrel{}{\epsilon }\left(\stackrel{}{r}\right)`$ scattered by the overlayer $`\delta \chi \left(z\right)`$. While diffusely scattered waves are not present in $`\delta \stackrel{}{\epsilon }\left(\stackrel{}{r}\right)`$, faint evanescent waves could be; these are weak near field effects and we neglect them. Thus $$\delta \stackrel{}{\epsilon }\left(\stackrel{}{r}\right)=\delta \epsilon _r\widehat{e}_re^{i\stackrel{}{k}_r\stackrel{}{r}},$$ (60) and the Fourier expansion, eq.(43), and transform $`\delta \stackrel{}{\epsilon }(\stackrel{}{k})`$, eq.(44), remain otherwise unchanged. Once again, $`\delta \stackrel{}{\epsilon }\left(\stackrel{}{r}\right)`$ is radiated by a current $$\delta \stackrel{}{J}(\stackrel{}{r})=\frac{i\omega }{4\pi }\delta \chi (z)\stackrel{}{E}\left(\stackrel{}{r}\right),$$ (61) where the field $`\stackrel{}{E}(\stackrel{}{r})`$ includes the incident and the unknown reflected waves; $`\stackrel{}{E}\left(\stackrel{}{r}\right)`$ must be self-consistently obtained as part of the solution. Then the ART, in the form of eq.(31), gives the field $`\delta \stackrel{}{\epsilon }(\stackrel{}{k})`$ in a direction $`\stackrel{}{k}`$ with polarization $`\widehat{e}`$ as $$\widehat{e}\delta \stackrel{}{\epsilon }(\stackrel{}{k})=\frac{iK}{2}𝑑v\delta \chi (z)\stackrel{}{E}\left(\stackrel{}{r}\right)\stackrel{}{\epsilon }_{ct}e^{i\stackrel{}{k}_{ct}\stackrel{}{r}},$$ (62) with the same connecting field given back in eq.(47). The approximation we use for $`\stackrel{}{E}\left(\stackrel{}{r}\right)`$ is the same as in last section. The arbitrariness of $`z_0`$ can be exploited by choosing it large enough that the overlayer extends well into the vacuum. Near $`z_0`$ the overlayer and the sharp step $`\chi _s\left(\stackrel{}{r}\right)`$ cancel each other out; slightly above $`z_0`$, in vacuum, the field is $$\stackrel{}{E}\left(\stackrel{}{r}\right)=\widehat{e}_0e^{i\stackrel{}{k}_0\stackrel{}{r}}+r\widehat{e}_re^{i\stackrel{}{k}_r\stackrel{}{r}},$$ (63) where $`r`$ is the unknown reflection coefficient we want to calculate, $$r=r_se^{2iqz_0}+\delta \epsilon _r.$$ (64) Slightly below $`z_0`$ and over most of the extension occupied by $`\delta \chi \left(z\right)`$ we are also in vacuum (provided the bulk to vacuum transition is not too gradual) and we approximate $`\stackrel{}{E}\left(\stackrel{}{r}\right)`$ by the same expression, eq.(63). Substituting into eq.(62) yields an equation for $`r`$, $$\frac{q}{K}\left(rr_se^{2iqz_0}\right)\left(2\pi \right)^2\delta (k_{}k_0)=$$ $$=\frac{iK}{2}_{\mathrm{}}^{z_0}𝑑z\delta \chi (z)𝑑x𝑑y\left[\widehat{e}_0e^{i\stackrel{}{k}_0\stackrel{}{r}}+r\widehat{e}_re^{i\stackrel{}{k}_r\stackrel{}{r}}\right]\stackrel{}{\epsilon }_{ct}e^{i\stackrel{}{k}_{ct}\stackrel{}{r}}.$$ (65) The integral over $`dxdy`$ yields a delta function, $`\left(2\pi \right)^2\delta (k_{}k_0)`$, and we can substitute $`\widehat{e}=\widehat{e}_c=\widehat{e}_r`$, $`\widehat{e}_{cr}=\widehat{e}_0`$, $`\stackrel{}{k}=\stackrel{}{k}_r=\stackrel{}{k}_c`$. The integral over $`z`$ is conveniently expressed as $$_{\mathrm{}}^{z_0}𝑑z\delta \chi (z)e^{iQz}=\frac{\chi _0}{iQ}[e^{iQz_0}+\frac{\chi ^{}(Q)}{\chi _0}],$$ (66) where $`\chi ^{}(Q)`$ is the Fourier transform of $`d\chi (z)/dz`$, $$\chi ^{}(Q)=_{\mathrm{}}^+\mathrm{}𝑑z\frac{d\chi (z)}{dz}e^{iQz}$$ (67) Eq.(66) is proved by integrating the left hand side by parts, using $`\delta \chi (z_0)\chi _0`$, and $`d\delta \chi (z)/dz=d\chi (z)/dz`$. Using $`\overline{q}^2q^2=K^2\chi _0`$ and the identities in eq.(55) the final result is $$r=r_s\frac{\chi ^{}(\overline{q}+q)}{\chi ^{}(\overline{q}q)}.$$ (68) Notice that any possible dependence on the arbitrary choice of $`z_0`$ has cancelled out. This coincides exactly with the scalar wave result and provides the correct generalization to all polarizations. Within these approximations the specular reflection coefficient $`r`$ has no polarization dependence beyond that already implicit in the reflection coefficient $`r_s`$ for the ideal flat step surface; the “static Debye-Waller” factor is polarization independent. To conclude we mention some illustrative examples: (a) The error-function profile $$\chi (z)=\frac{\chi _0}{\sqrt{2\pi \sigma ^2}}_{\mathrm{}}^z𝑑x\mathrm{exp}\left(\frac{x^2}{2\sigma ^2}\right),$$ (69) gives $$\frac{\chi ^{}(\overline{q}+q)}{\chi ^{}(\overline{q}q)}=e^{2q\overline{q}\sigma ^2},$$ (70) the same factor obtained in the previous section for a Gaussian rough surface. This is as expected, the error function is the averaged profile for the Gaussian rough surface. (b) The Epstein (or Fermi distribution) profile $$\chi (z)=\frac{\chi _0}{1+e^{z/\sigma }},$$ (71) gives $$\frac{\chi ^{}(\overline{q}+q)}{\chi ^{}(\overline{q}q)}=\frac{\overline{q}+q}{\overline{q}q}\frac{\mathrm{sinh}[\pi \sigma (q\overline{q})]}{\mathrm{sinh}[\pi \sigma (q+\overline{q})]}.$$ (72) (c) The triangular profile $$\chi (z)=\{\begin{array}{ccc}\chi _0\hfill & \text{for}\hfill & z<\sigma /2\hfill \\ \chi _0\left(12z/\sigma \right)\hfill & \text{for}\hfill & |z|<\sigma /2\hfill \\ 0\hfill & \text{for}\hfill & z>\sigma /2\hfill \end{array},$$ (73) gives $$\frac{\chi ^{}(\overline{q}+q)}{\chi ^{}(\overline{q}q)}=\frac{q\overline{q}}{q+\overline{q}}\frac{\mathrm{sin}[(q+\overline{q})\sigma /2]}{\mathrm{sin}[(q\overline{q})\sigma /2]}.$$ (74) The reliability of these approximations was studied in in the case of scalar waves. There is no reason to expect any difference from the conclusions reached there: the “static Debye-Waller” in eq.(68) provides a remarkably good approximation for the intensities reflected by interfaces of arbitrary grading profile even for transition regions that are quite wide ($`\sigma `$ as large as several nanometers). The phase of the reflected waves is however more sensitive; eq.(68) provides a good approximation for more abrupt transitions ($`\sigma `$ of the order of 1 *nm* or less). ## 5 Conclusion The main result of this work, eq.(15), is an asymptotic form of the reciprocity theorem which can be used as the basis for a practical method for calculations. The theorem states that the field radiated in the presence of a nontrivial medium, in a certain direction and with a given polarization, is a suitable ‘component’ of the radiating source. This ‘component’ is to be extracted by introducing an auxiliary ‘connecting’ field which contains the necessary information about the medium. The practical advantage of the method lies in the simplifications achieved by systematically avoiding unnecessary calculations; it thereby allows one to tackle problems of increasing complexity. In forthcoming papers we will further explore the application of the ART to the study of the dynamical diffraction of radiation generated by sources within a crystal, the so-called Kossel lines. Even this well explored topic has not been exhausted. Of particular interest are situations where the Bragg angle lies close to $`\pi /2`$ and the Kossel cones degenerate into single beams , and situations where the source location is revealed by the oscillatory ‘Pendellösung’ structure of the diffraction pattern . Other applications will include a new approach to thermal diffuse scattering under conditions of dynamical diffraction . Acknowledgments. I am indebted to P. Zambianchi and E. Sutter for valuable discussions.
warning/0001/cond-mat0001250.html
ar5iv
text
# Hydrodynamic theory for granular gases URL: http://www.cec.uchile.cl/cinetica/ ## I Introduction Granular systems subjected to a sufficiently strong excitation may have a fluid-like behavior . From the very beginning several authors have attempted to derive hydrodynamic equations for these systems . If the excitation of the granular system is through a permanent injection of energy, the fluid system may stabilize to a low density stationary gaseous state which necessarily is a nonequilibrium state and it usually is inhomogeneous as well. To develop the basic features of the theory of gaseous granular systems we restrict the analysis to the simplifying inelastic hard sphere model (IHS) . Many authors studying granular gases have put particular attention to studying the spontaneous homogeneous cooling of a granular system using periodic boundary conditions . This time dependent state is called homogeneous cooling state (HCS) and the understanding of its properties has been improving through many articles . A crucial breakthrough was the realization by Goldshtein and Shapiro that the homogeneous cooling distribution function has a scaling property with respect to the instantaneous temperature. Such distribution—which we will be calling $`f_{\mathrm{𝖧𝖢𝖲}}`$—is known in approximate forms . It is known, among other things, that its fourth cumulant $`\kappa `$ does not vanish and that it has a long velocity tail. A nonequilibrium inhomogeneous gaseous system, on the other hand, is described by a distorted distribution function typically obtained from Boltzmann’s equation expanding the distribution either in gradients of the hydrodynamic fields (Chapmann-Enskog’s method) or making a moment expansion (Grad’s method) . For normal gases the expansion is made about the equilibrium Maxwell’s distribution. In this article we will assume that a low density nonequilibrium granular system has a local distribution function which can be obtained expanding about a distribution $`f_0`$ resembling a $`f_{\mathrm{𝖧𝖢𝖲}}`$ in the sense that it has a significantly nonvanishing fourth cumulant. We introduce a reference function (see Eq. (5) below) which is a Maxwellian distorted by a factor which incorporates the next non trivial cumulant, a fourth cumulant, to the distribution function and then we undertake a perturbative moment expansion à la Grad about $`f_0`$ to solve Boltzmann’s equation. Some authors have done some calculations in this direction but using the gradient expansion (Chapman-Enskog) method . The point is that, without the notion of equilibrium, we expect that the reference state, $`f_0`$ in our case, should resemble more the homogeneous cooling state than the simple Maxwellian. We study a two-dimensional system of hard disks, and the moment expansion—in dimension two—is an 8 moment expansion: the number density $`n(\stackrel{}{r},t)`$, the velocity field $`\stackrel{}{v}(\stackrel{}{r},t)`$, the granular temperature field $`T(\stackrel{}{r},t)`$, the pressure tensor $`P_{ij}(\stackrel{}{r},t)`$ and the heat flux vector field $`\stackrel{}{Q}(\stackrel{}{r},t)`$. The dynamic variables are not the components of the pressure tensor $`IP`$ itself but the components of the symmetric traceless part $`p_{ij}`$ where $`P_{ij}=p\delta _{ij}+p_{ij}`$ and $`p`$ is the hydrostatic pressure. As it can be seen in Grad’s article or in the method yields hydrodynamic equations for all the fields mentioned above. In particular, the dynamic equations for $`p_{ij}`$ and $`\stackrel{}{Q}`$ take the place of what would normally be the constitutive (transport) equations of standard hydrodynamics. This last point means that we are not assuming any constitutive equations whatsoever, their present counterparts are dynamic equations. It is well established that, in the case of the IHS model for granular systems, Boltzmann’s equation is modified in that the restitution coefficient $`r`$ enters solely in the gain term of the collision integral and it does so in two forms. First the gain term has an overall factor $`r^2`$ and second, the distribution functions appearing in the gain term depend upon the precollision velocities, and these velocities depend on $`r`$ . When the modified Boltzmann’s equation is used the stemming hydrodynamic equations get factors that depend on the inelasticity coefficient $`q`$, $$q=\frac{1r}{2}$$ (1) ($`q=0`$ in the perfectly elastic case) except that the mass continuity equation and the momentum balance equation remain unchanged since mass and momentum continue being microscopically conserved. In the context of Boltzmann’s equation a dissipative gas satisfies the ideal gas equation of state $$p=nT$$ (2) where the granular temperature $`T`$ is defined in energy units as the average kinetic energy per particle. If we were to consider the Boltzmann-Enskog equation, then the inelasticity coefficient would enter through the Enskog collision factor $`\chi `$, and the equation of state of a normal gas and a dissipative gas would differ, but in the present context Eq. (2) holds. Usual moment expansion methods (as Grad’s is) are appropriate to describe bulk properties. Wall effects are not well described unless higher order momenta are included in the expansion, which are not trivial to handle . Already in normal gases hydrodynamic fields may have discontinuities at walls. In a previous article we gave a kinetic description of a one-dimensional granular system with theoretical tools such that our description was correct and precise up to the walls but we have not generalized yet that type of formalism to higher dimensions, hence, in the present article, we use moment expansions. In consequence the formalism we are going to present suffers too of the weakness of moment expansions: it is unreliable precisely at the points where the boundary conditions should be imposed forcing us to trade the boundary conditions for conditions imposed far from the walls based on the actual behavior of the system according to our molecular dynamic simulations. Our moment expansion method, explained in detail in Sec. II, uses as reference distribution function a distribution $`f_0`$ which differs from a Maxwellian distribution in that it has a nonvanishing fourth cumulant, $`\kappa `$. The method leads to hydrodynamic equations for low density granular systems that depend parametrically on the inelasticity coefficient $`q`$ and $`\kappa `$. We apply this hydrodynamic equations to a stationary and purely conductive case as in . The value of the fourth cumulant $`\kappa `$, or better, the dependence of the fourth cumulant on $`q`$ is determined directly from molecular dynamic simulations, and it turns out to be independent of the size $`N`$ of the system. The predictions that follow from our formalism agree very well with all our simulational data. In Sec. II we briefly present the moment expansion method, in Sec. III the hydrodynamic equations that follow are given and specialized to a purely conductive case and finally in Sec. IV theory and simulational results are compared. Final comments are in Sec. V. ## II The moment expansion method for granular systems Moment expansion methods can summarily be described as follows. Take a velocity distribution function $`f_0(\stackrel{}{r},\stackrel{}{c},t)`$ which is considered to be the reference function about which an expansion is going to be made. For normal gases the natural choice for $`f_0`$ is a local Maxwellian distribution $`f_𝖬`$ written in terms of the peculiar velocity $`\stackrel{}{C}=\stackrel{}{c}\stackrel{}{v}(\stackrel{}{r},t)`$, where $`\stackrel{}{v}(\stackrel{}{r},t)`$ is the hydrodynamic velocity. Next a set of orthonormal polynomials on $`\stackrel{}{C}`$, $`H_a(\stackrel{}{C})`$, are built in the sense that $`H_0=1`$ and $$H_a(\stackrel{}{C})H_b(\stackrel{}{C})f_0(\stackrel{}{r},\stackrel{}{C},t)𝑑\stackrel{}{C}=\delta _{ab}$$ (3) The polynomials $`H_a`$ are obtained simply building a base of orthonormal polynomials starting from $`H_0=1`$ and from first degree upwards. When $`f_0`$ is a Maxwellian the $`H_a`$ are Hermite polynomials but in general they are not. Then a solution of the form $$f(\stackrel{}{r},\stackrel{}{C},t)=\left(1+\underset{a}{}H_a(\stackrel{}{C})R_a(\stackrel{}{r},t)\right)f_0(\stackrel{}{r},\stackrel{}{C},t)$$ (4) is replaced in Boltzmann’s equation. The $`R_a`$ are the moments of $`f`$ with respect to the $`H_a`$ and they can directly be related to the moments associated to $`\stackrel{}{c}`$ (the velocity field $`\stackrel{}{v}`$), to $`\rho C_iC_j`$ (the kinematic pressure tensor $`P_{ij}`$) and to $`\frac{1}{2}\rho C^2\stackrel{}{C}`$ (the heat flux vector $`\stackrel{}{Q}`$). One could go on but we have used polynomials only containing $`C_i`$, $`C_iC_j`$ and $`C^2C_i`$ as Grad did. The following step is to derive integrability conditions multiplying the kinetic equation consecutively by the $`H_n`$ and then integrating the equation over $`\stackrel{}{C}`$. The idea is to do this up to a given order and drop all contributions coming from polynomials of degree higher than a chosen value (up to order 3 in our case). This gives a set of hydrodynamic equations for the different moments. A key point is the choice of the reference function $`f_0`$. The two dimensional Maxwellian $`f_𝖬=n\frac{m}{2\pi T}\mathrm{exp}[mC^2/(2T)]`$ is privileged as the solution describing the equilibrium state of a normal gas. Since in granular systems there is no such thing as equilibrium a next best choice, seems a distorted Maxwellian distribution $$f_0=\left[1+\frac{\kappa }{2}\left(1𝒞^2+\frac{𝒞^4}{8}\right)\right]f_𝖬$$ (5) where the dimensionless peculiar velocity is $$\stackrel{}{𝒞}=\sqrt{\frac{m}{T}}\stackrel{}{C}$$ (6) and $`T`$ is the granular temperature. The coefficient $`\kappa `$ is the fourth cumulant of $`f_0`$ and it depends on the inelasticity coefficient $`q`$ while the coefficients in front of $`𝒞^2`$ and $`𝒞^4`$ are derived from requiring that $`f_0`$ is normalized and that $`\frac{m}{2}C^2_{f_0}=T`$. The fourth cumulant $`\kappa `$ in dimension two is $$\kappa =\frac{𝒞^42𝒞^2^2}{𝒞^2^2},$$ (7) In the case of the homogeneous cooling state, recent articles have justified explicit forms for $`\kappa `$ in the context of distribution functions like $`f_0`$. Their results are well approximated by $$\kappa =\frac{b_1+b_2q}{1+b_3q}q,$$ (8) with $$b_1=2,b_213.619,b_34.5969.$$ (9) The rational expression given in (8) is valid within 2.9% for $`q0.08`$, $`r=12q=0.84`$. Notice that $`\kappa (q=0)=0`$ allowing to recover the elastic case. We describe quite satisfactorily nonhomogeneous, nonequilibrium stationary states with a $`\kappa `$ as in (8), but with numerical factors different from those in Eq. (9). The latter values are valid only for the homogeneous cooling state while our system is kept in a stationary inhomogeneous state with appropriate boundary conditions. For any $`\kappa `$ the resulting distribution obtained with the method summarized above is $`f^{(\kappa )}`$ $`=`$ $`[1+{\displaystyle \frac{p_{xx}}{p(2+\kappa )}}(𝒞_x^2𝒞_y^2)+2{\displaystyle \frac{p_{xy}}{p(2+\kappa )}}𝒞_x𝒞_y`$ (11) $`+{\displaystyle \frac{Q_x}{2Q_0}}{\displaystyle \frac{(𝒞^242\kappa )𝒞_x}{(2+5\kappa \kappa ^2)}}+{\displaystyle \frac{Q_y}{2Q_0}}{\displaystyle \frac{(𝒞^242\kappa )𝒞_y}{(2+5\kappa \kappa ^2)}}]f_0`$ where, $`p_{yy}=p_{xx}`$ and $`Q_0=\sqrt{\frac{T}{m}}p`$. The distribution $`f^{(\kappa )}`$ shares with $`f_0`$ the first scalar moments: density, temperature and fourth cumulant, for any value of $`q`$. If $`\kappa `$ is chosen to be zero then, in Eq. (11), $`f_0f_𝖬`$ and $`f^{(\kappa )}`$ becomes the usual Grad’s distribution. Hence the whole method would be the original method devised by Grad and, if $`\kappa `$ is chosen to be Eq. (8) with coefficient values as in Eq. (9), then $`f_0`$ would be what we are calling $`f_{\mathrm{𝖧𝖢𝖲}}`$. Given our ignorance regarding granular gases one could, in principle, accept $`f_𝖬`$ or $`f_{\mathrm{𝖧𝖢𝖲}}`$ as legitimate reference functions to make the moment expansion. In the following sections we compare the three formalisms (reference functions $`f_𝖬`$, $`f_{\mathrm{𝖧𝖢𝖲}}`$ and $`f_0`$, the latter with a $`\kappa `$ adjusted to the results) concluding that only the one based on $`f_0`$ gives acceptable results for a sufficiently large range of $`q`$. In fact they are very good. ## III The hydrodynamic equations As already mentioned, the inelasticity coefficient $`q`$ enters the kinetic equation in two different forms. It appears as a factor $`\frac{1}{r^2}=\frac{1}{(12q)^2}`$ in the gain term and it appears in the expression for the precollision velocities which are part of the argument of the distribution functions appearing in the gain term. When the solution $`f^{(\kappa )}`$ is inserted in Boltzmann’s equation, $`q`$ enters in a still third form, precisely through the $`\kappa `$ coefficient given by Eqs. (8) and (11). Expanding the collisonal term of Boltzmann’s equation in powers of $`q`$ and $`\kappa `$, the moment method yields the following hydrodynamics equations for a granular gas $`{\displaystyle \frac{Dn}{Dt}}+n\stackrel{}{v}`$ $`=`$ $`0,`$ (12) $`mn{\displaystyle \frac{D\stackrel{}{v}}{Dt}}n\stackrel{}{F}+IP`$ $`=`$ $`0,`$ (13) $`n{\displaystyle \frac{DT}{Dt}}+\stackrel{}{Q}+IP:\stackrel{}{v}`$ $`=`$ $`{\displaystyle \frac{2A(q)nT}{\tau }},`$ (14) $`{\displaystyle \frac{p_{ij}}{t}}+{\displaystyle \frac{}{x_k}}\left(v_kp_{ij}\right)+{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{Q_i}{x_j}}+{\displaystyle \frac{Q_j}{x_i}}\delta _{ij}{\displaystyle \frac{Q_k}{x_k}}\right)+`$ (15) $`p_{rj}{\displaystyle \frac{v_i}{x_r}}+p_{ri}{\displaystyle \frac{v_j}{x_r}}\delta _{ij}p_{rs}{\displaystyle \frac{v_s}{x_r}}+p\left({\displaystyle \frac{v_i}{x_j}}+{\displaystyle \frac{v_j}{x_i}}\delta _{ij}{\displaystyle \frac{v_r}{x_r}}\right)`$ $`=`$ $`{\displaystyle \frac{B(q)}{\tau }}p_{ij},`$ (16) $`{\displaystyle \frac{Q_k}{t}}+{\displaystyle \frac{}{x_r}}\left(v_rQ_k\right)+{\displaystyle \frac{3}{2}}{\displaystyle \frac{v_k}{x_r}}Q_r+{\displaystyle \frac{1}{2}}{\displaystyle \frac{v_r}{x_k}}Q_r+{\displaystyle \frac{1}{2}}{\displaystyle \frac{v_r}{x_r}}Q_k+`$ (17) $`{\displaystyle \frac{T}{m}}{\displaystyle \frac{p_{kr}}{x_r}}+{\displaystyle \frac{3p_{kr}}{m}}{\displaystyle \frac{T}{x_r}}{\displaystyle \frac{p_{kr}}{mn}}{\displaystyle \frac{P_{rs}}{x_s}}+{\displaystyle \frac{3p}{m}}{\displaystyle \frac{T}{x_k}}`$ $`=`$ $`{\displaystyle \frac{C(q)}{2\tau }}Q_k,`$ (18) where $`D/Dt/t+\stackrel{}{v}`$, $$\tau =\frac{1}{2\sigma p}\sqrt{\frac{mT}{\pi }}$$ (19) is a characteristic relaxation time, $`\sigma `$ is the diameter of the particles, and the coefficients $`A`$, $`B`$ and $`C`$ for the generic distribution $`f^{(\kappa )}`$ are $`A(q)`$ $`=`$ $`[q(1q)+𝒪(q^5)](1+{\displaystyle \frac{3}{32}}\kappa +{\displaystyle \frac{9}{4096}}\kappa ^2),`$ (20) $`B(q)`$ $`=`$ $`{\displaystyle \frac{1+\frac{1}{2}q\frac{3}{2}q^2+\frac{23}{32}\kappa (1+\frac{q}{2})+\frac{\kappa ^2}{4096}+𝒪(q^3)+𝒪(\kappa ^3)}{(1+\frac{1}{2}\kappa )}},`$ (21) $`C(q)`$ $`=`$ $`{\displaystyle \frac{1+\frac{13}{2}q\frac{15}{2}q^2+\frac{\kappa }{64}(206+1267q)\frac{2415}{4096}\kappa ^2+𝒪(q^3)+𝒪(\kappa ^3)}{(1+\frac{5}{2}\kappa \frac{\kappa ^2}{2})}}.`$ (22) They depend on $`q`$ and $`\kappa `$, but $`\kappa `$ depends on $`q`$. At least for small $`q`$ the coefficients $`A`$, $`B`$ and $`C`$ are positive ($`q`$ varies between 0 and $`\frac{1}{2}`$). The coefficient $`A`$ in Eq. (14) determines the energy dissipation in the system and consequently it vanishes in the elastic limit while $`B`$ and $`C`$ tend to 1. In the previous equations both $`p_{ij}`$ and $`P_{ij}`$ appear depending on which of them gives a more compact expression. It is our hope that the above hydrodynamics is valid in a wide variety of situations compatible with gaseous states and no clustering but in this article we restrict our study to a hydrostatic case. We are going to consider the purely conductive regime, with no external force, $`\stackrel{}{F}=0`$. The system of $`N`$ disks is in a rectangular box of dimension $`L_x\times L_y`$, with thermal walls, at $`y=\frac{1}{2}L_y`$ and at $`y=\frac{1}{2}L_y`$, both at temperature $`T_0`$, and periodic boundary conditions in the $`X`$ direction. The case with no external force is quite simple because the system is symmetric with respect to $`yy`$ and the pressure is uniform. If the system were conservative, as a normal gas, this would be a homogeneous system at thermal equilibrium, but since the system is dissipative the temperature depends on the coordinate $`y`$ ($`T`$ has a minimum at the symmetry axis) and the problem is much less trivial. Since the system is purely conductive there is no velocity field, $`P_{xy}=0`$, $`Q_x=0`$ and the other fields depend only on the coordinate $`y`$. The set of equations (12-18) becomes $`\begin{array}{cccc}P_{xx}\hfill & ={\displaystyle \frac{BA}{B}}p,\hfill & P_{yy}\hfill & ={\displaystyle \frac{B+A}{B}}p,\hfill \\ & & & \\ Q_y^{}\hfill & =4Ap^2\sigma \sqrt{{\displaystyle \frac{\pi }{mT(y)}}},\hfill & T^{}\hfill & ={\displaystyle \frac{\sigma BC}{3A+2B}}\sqrt{{\displaystyle \frac{\pi m}{T}}}Q_y\hfill \end{array}`$ (26) where the prime indicates derivative with respect to $`y`$. Notice that because there is inelasticity the pressure tensor is anisotropic in the sense that $`P_{xx}P_{yy}`$. In fact $`P_{yy}P_{xx}Aq`$ and they do not depend on the coordinate $`y`$. Next we are going to compare the implications of these equations with our molecular dynamics results. To this end we should, in principle, solve these hydrodynamic equations using the boundary conditions associated to the particular simulations that we have studied. This is not a straightforward task because, as we have mentioned at the end of the introduction, the moment expansion method (behind the previous hydrodynamic equations) does not give a good description near boundaries. For example, if we impose that a wall behaves as a stochastic wall at temperature $`T_0`$, the observed field $`T`$ is not expected to take that value near the wall. Later on it will be shown how we tackle this problem. From Eq. (26) it is direct to derive that the temperature field satisfies the equation $$TT^{\prime \prime }+\frac{1}{2}T_{}^{}{}_{}{}^{2}=k^2\sigma ^2p^2,$$ (27) where $$k^2\frac{4\pi ABC}{3A+2B}.$$ (28) Since the pressure is uniform, and because of Eq. (2), this equation can also be used as an equation for the inverse of the density. Before proceeding to solve the equations we adimensionalize the problem defining a coordinate $`\xi =y/L_y`$, ($`\frac{1}{2}\xi \frac{1}{2}`$) and we also define $$\begin{array}{cccc}\hfill \overline{n}& =\frac{N}{L_xL_y},\hfill & \hfill n(y)& =\overline{n}n^{}(\xi ),\hfill \\ & & & \\ \hfill K& =\frac{k}{2\sqrt{2}}\left(\frac{N\sigma }{L_x}\right),\hfill & \hfill Q_y(y)& =\overline{n}T_0\sqrt{\frac{T_0}{m}}Q_y^{}(\xi ),\hfill \\ & & & \\ \hfill T(y)& =T_0T^{}(\xi ),\hfill & \hfill p& =\overline{n}T_0p^{}.\hfill \end{array}$$ (29) Once Eq. (27) is solved and converted to a solution for the dimensionless number density $`n^{}(\xi )`$ the result is $$4K|\xi |=\frac{1}{n^{}(\xi )}\sqrt{\frac{n_{\mathrm{max}}^{}n^{}(\xi )}{n_{\mathrm{max}}^{}}}+\frac{1}{n_{\mathrm{max}}^{}}\mathrm{𝖺𝗋𝖼𝗍𝖺𝗇𝗁}\sqrt{\frac{n_{\mathrm{max}}^{}n^{}(\xi )}{n_{\mathrm{max}}^{}}},$$ (30) where $`n_{\mathrm{max}}^{}`$ is the value taken by $`n^{}(\xi )`$ in the middle of the system, $`\xi =0`$. In what follows it is shown that $`n_{\mathrm{max}}^{}`$ is a simple function of the parameter $`K`$ defined above. The effect of the boundaries is traded in favor of the observed values of $`n_{\mathrm{max}}^{}`$. The integral condition $`n𝑑x𝑑y=N`$ can be cast as $$_0^{1/2}n^{}𝑑\xi =\frac{1}{2},$$ (31) but since $`d\xi =(d\xi /dn^{})dn^{}`$ the integral condition yields, $$\frac{n_b^{}}{n_{\mathrm{max}}^{}}=1\mathrm{tanh}^2K,$$ (32) where $`n_b^{}`$ is the value that $`n^{}(\xi )`$ takes at the two boundaries, $`n_b^{}=n^{}(\pm \frac{1}{2})`$. Equation (30) evaluated at $`\xi =\frac{1}{2}`$ yields a different condition over $`n_b^{}`$ and it follows that $`n_{\mathrm{max}}^{}`$ $`=`$ $`{\displaystyle \frac{1}{2}}+{\displaystyle \frac{\mathrm{sinh}(2K)}{4K}}.`$ (33) This is a strong result, it says that the dimensionless number density at the center, $`n_{\mathrm{max}}^{}`$, is determined by $`K^2`$ alone, $$K^2=a_0^2\frac{2ABC}{3A+2B},$$ (34) where $`a_0^2=N\rho _A/\alpha `$, $`\alpha =L_x/L_y`$ is the aspect ratio of the box, and $`\rho _A=(\pi /4)(N\sigma ^2/(L_xL_y))`$ is the fraction of area occupied by the disks. Since $`A`$ vanishes in the elastic limit and for small inelasticity coefficient $`q`$, $`Aq`$ while $`B1`$ and $`C1`$ then $$K^2a_0^2q\frac{1}{\alpha }qN\rho _A.$$ (35) In the quasielastic limit then, the control parameter is basically $`qN\rho _A`$. This result, for fixed area density, resembles what we obtained in the one dimensional case , namely, that the relevant control parameter of the one dimensional equations is $`qN`$. The temperature is $`T^{}(\xi )=p^{}/n^{}(\xi )`$ but since the pressure is uniform one may be tempted to use $`p^{}=n_b^{}T_b^{}`$ with the value for $`n_b^{}`$ already derived from the theory and the value $`T_b^{}`$ imposed in the simulation. This would give a bad fit however because, as we have been emphasizing, the formalism is not reliable near the boundaries. Therefore we choose for $`p^{}`$ the value $`p^{}=T_{\mathrm{𝗆𝗂𝗇}}^{}n_{\mathrm{max}}^{}`$. We know $`n_{\mathrm{max}}^{}`$ from Eq. (33), and we take the value for $`T_{\mathrm{𝗆𝗂𝗇}}^{}`$ directly from our simulations. It may be said that $`T_{\mathrm{𝗆𝗂𝗇}}^{}`$ is a parameter to adjust our results. The dimensionless heat flux becomes, $$Q_y^{}(\xi )=\frac{3A+2B}{2BC}\frac{1}{a_0}\sqrt{\frac{p^{}}{n^{}(\xi )}}\frac{p^{}}{n^{}(\xi )^2}\frac{dn^{}}{d\xi }.$$ (36) ## IV Simulation-theory comparison In this section we compare the simulational results with the values given by three formalisms which use as reference function: ($`i`$) $`f_𝖬`$, ($`ii`$) $`f_{\mathrm{𝖧𝖢𝖲}}`$ and ($`iii`$) $`f_0`$. We are calling $`f_{\mathrm{𝖧𝖢𝖲}}`$ the function like $`f_0`$ but with $`\kappa `$ defined with the values given in Eq. (9). Since the formalisms differ by terms of higher order in $`q`$ their predictions are quite similar unless $`qN`$ is large enough. Typically the Maxwellian theory and the one based on $`f_{\mathrm{𝖧𝖢𝖲}}`$ are valid until about $`qN=20`$ and are reasonable until $`qN=40`$ while the theory based on $`f_0`$, with an adjusted $`\kappa `$, is valid for values of $`qN`$ up to 200, and reasonably good until about $`qN300`$. ### A Simulational setup We have performed simulations of a two dimensional system of $`N=2300`$, $`N=3600`$, $`N=10000`$, and $`N=19600`$ inelastic hard disks inside a $`L\times L`$ box with lateral periodic boundary conditions while the upper and lower walls are kept at granular temperature $`T_0=1`$. The area fraction covered by the disks was chosen to be $`\rho _A=0.01`$ (in which case the nonideal corrections to the equation of state are less than 2%) while the $`qN`$ dissipation parameter ranges from $`qN=10`$ up to $`qN=400`$. In the $`N=2300`$ case, the smallest simulated system, the ratio between the mean free path and the linear size of the system (Knudsen number) is 0.065 and it is smaller in the other cases. This value guarantees that not too close to the walls the fluid has a hydrodynamic behavior. The wall temperature $`T_0`$ is imposed sorting the velocity of the bouncing particles as if they were coming from a heat bath at $`T=T_0`$. In every simulation the system was relaxed from an initial condition for a sufficiently long time. After the relaxation we measure local time averages of the main moments of the distribution (i.e., $`n`$, $`\stackrel{}{v}`$, $`T`$, $`p_{ij}`$, $`\stackrel{}{Q}`$) inside each one of a set of square cells. Taking advantage of the translation invariance in the $`X`$ direction, it is natural to take horizontal averages getting, in this way, smooth vertical profiles for the observed hydrodynamic fields. ### B Theory and simulation comparison In order to compare theory and simulation some analysis is needed because size effects are quite noticeable unless the number $`N`$ of particles is above about 3600. We use expression (33) for $`n_{\mathrm{𝗆𝖺𝗑}}^{}`$ as our point of contact between theory and simulation and proceed to adjust the coefficients $`b_k`$ in Eq. (8) so that $`K`$ takes the values that make theory give the observed values for $`n_{\mathrm{𝗆𝖺𝗑}}^{}`$. With this aim we first write $`K`$ as a rational expression $`K=a_0\sqrt{q}(1+a_1q)/(1+a_2q)`$ and find the values of the $`a_k`$ so that Eq. (33) reproduces the observed values of $`n_{\mathrm{𝗆𝖺𝗑}}^{}`$. We have to adjust $`a_0`$ in spite of its definition, given under Eq. (34), because the effective values for the number of particles, the global density and the aspect ratio get distorted since there is a layer near the thermalizing walls which does not behave hydrodynamically. Once this expression for $`K`$ is fixed, we invert (34) to obtain values for $`\kappa `$ as a function of $`q`$. The size effect is in $`a_0`$ alone and $`\kappa (q)`$ is approximately the same for systems with $`N`$ larger than about 3600, as shown in Fig. 1. In this figure there is a solid line which is Eq.(8) with $$b_1=22.541,b_2=6.19187,b_3=79.0362.$$ (37) The discrepancies between this curve and the empirical values of $`\kappa `$ are less than 2% for the whole range of $`q`$ considered. More in detail, Fig. 1 shows the behavior of $`\kappa (q)`$ for systems of different size. It is seen that in the case of $`N=2300`$ (solid circles) $`\kappa (q)`$ is dependent on the system’s size, while the predictions in the cases $`N=3600`$, $`10000`$, and $`19600`$ differ among themselves by less than 2%. Only the smallest simulated system ($`N=2300`$) departs from this otherwise universal shape. In the following we present the results corresponding to the $`N=10000`$ case. #### a The density: As a first step and in order to check the validity of the kinetic description (no clustering, ) we have plotted the final configurational positions of the particles (not shown here) for different values of $`qN`$ up to $`qN=400`$, founding that clustering begins at about $`qN300`$. In fact, we have detected that the number of collisions per unit time increases dramatically shortly after $`qN=300`$. As a second step we check the validity of Eq. (33) (which for fixed geometry depends only on $`qN`$) with simulations. The top Fig. 2 shows both $`n_{\mathrm{max}}^{}`$ (growing curves with $`qN`$) and $`n_b^{}`$ (the decreasing curves). It is seen how well the values of $`n_{\mathrm{𝗆𝖺𝗑}}^{}`$ (solid circles) come from the $`f_0`$ distribution (there should be no surprise as these are the fitted data), compared with the prediction using the other two distributions. This graph shows that $`n_{\mathrm{𝗆𝖺𝗑}}^{}`$ with the other two formalisms is good only up to $`qN40`$. Also at top Fig. 2 shows the values predicted by Eq. (32) and the observed values of $`n_b^{}`$ (see caption), and it is seen that all theoretical schemes give values close to the simulational results in the considered range of $`qN`$. This is the only case where, for large $`qN`$, predictions coming from $`f_𝖬`$ are better that those coming from $`f_0`$, but we do not believe there is anything deep here since our moment expansion method is not reliable near walls. Figure 2, bottom, compares theory and simulational density profiles for $`qN=30`$ and $`qN=200`$. As it has already been explained, the value $`n_{\mathrm{𝗆𝖺𝗑}}^{}`$ in Eq. (30) is fixed by Eq. (33) and there are no extra parameters to adjust. It is seen that theory in all cases ($`f_𝖬`$, $`f_{\mathrm{𝖧𝖢𝖲}}`$ and $`f_0`$) give good agreement for low values of $`qN`$. However predictions for $`qN=200`$, when using $`f_𝖬`$ or $`f_{\mathrm{𝖧𝖢𝖲}}`$ fail. It is amazing how large $`qN`$ can be when $`f_0`$ is used. In the last case the parameter $`K`$ is adjusted only from the knowledge of $`n_{\mathrm{max}}^{}`$ and it accurately predicts the behavior in almost all the volume. From the figure it can be appreciated that near the walls $`(\xi =\pm 0.5)`$, as mentioned in the introduction, the theory does not predict well the behavior of $`n^{}(\xi )`$. #### b The temperature: It has already been mentioned that the temperature profiles exhibit a minimum at the center and there is a temperature jump at the boundaries. In Fig. 3 we show the simulational results for $`T_{\mathrm{𝗆𝗂𝗇}}^{}=T^{}(\xi =0)`$ and temperature $`T_b^{}=T^{}(\xi =\pm 1/2)`$. It is seen that for $`qN=40`$ the temperature of the fluid by the walls is about 40% lower than the imposed value. This effect is due to dissipation and in 1D it has been shown to be a $`𝒪(qN)`$ effect . Because at present we have no theory to describe the temperature jump that takes place near the thermal walls, and we know that Grad’s method does not give good results near boundaries, we have chosen the observed value of the temperature at mid height, $`T_{\mathrm{𝗆𝗂𝗇}}^{}`$, as the value to use in the formalism. The observed values decrease with $`qN`$ and, in the case $`N=10000`$, we have adjusted them with the following expression $`\mathrm{ln}(T_{\mathrm{𝗆𝗂𝗇}})`$ $`=`$ $`0.0163259qN+\mathrm{5.27656\hspace{0.17em}10}^5\left(qN\right)^2`$ (40) $`\mathrm{1.86396\hspace{0.17em}10}^7\left(qN\right)^3+\mathrm{4.241\hspace{0.17em}10}^{10}\left(qN\right)^4`$ $`\mathrm{4.23878\hspace{0.17em}10}^{13}\left(qN\right)^5`$ whose faithfulness is shown in Fig. 3. Since the pressure is uniform the temperature profile can be written directly as $$T^{}(\xi )=\frac{T_{\mathrm{𝗆𝗂𝗇}}^{}n_{\mathrm{max}}^{}}{n^{}(\xi )}.$$ (41) Figure 4 shows the simulational and theoretical temperature profiles for some values of $`qN`$. The three upper curves show the profiles for small $`qN`$ values ($`qN=10`$, 20, 30) and it is seen that the predictions using $`f_0`$ (solid line) give an excellent fit to the simulational data while for the results obtained using $`f_𝖬`$ or $`f_{\mathrm{𝖧𝖢𝖲}}`$ (dashed and light lines respectively) the fit is only fair. The lowest curves shows $`T`$ profiles for larger values of $`qN`$ ($`qN=50`$, 100, 200). In this case only the formalism based on $`f_0`$ give an acceptable description and the agreement is very good up to $`qN=200`$. In the case of $`qN=275`$, not shown, there is still a reasonable agreement when $`f_0`$ is used. #### c The pressure: Now that the values of $`n_{\mathrm{𝗆𝖺𝗑}}^{}`$ and $`T_{\mathrm{𝗆𝗂𝗇}}^{}`$ are known and since the pressure is uniform (both in theory and simulationally) and theory asserts that $`p^{}=T_{\mathrm{𝗆𝗂𝗇}}^{}n_{\mathrm{𝗆𝖺𝗑}}^{}`$, then we have the value of $`p^{}`$. From the value for the pressure and the set of equations (26) one gets the theoretical values for $`P_{yy}`$ in terms of the pressure tensor. These values are compared with our simulational data in Fig. 5. Again the comparison is good when the formalism with $`f_0`$ is used and it is only fairly good with the other two formalisms. As it has already been mentioned after Eq. (26), the two diagonal terms of the pressure tensor are not equal because the system is anisotropic. #### d The heat current: Equation (36) gives the theoretical expression for the heat current $`Q_y^{}`$. At top Fig. 6 shows simulational data and theoretical predictions for small values of $`qN`$. The three formalism predict well the observed values. At bottom in Fig. 6 it is possible to see that for $`qN=200`$ only the formalism using $`f_0`$ fits the observed data, while the other two fail badly. In the case of $`qN=275`$ the agreement is still reasonable within about 3%. ## V Final comments In this article we have studied a bidimensional granular system of $`N`$ inelastic hard disks (normal restitution coefficient $`r`$). The system is placed in a rectangular box and it is kept in a stationary regime with upper and lower walls at granular temperature $`T_0`$. The lateral walls are periodic. The quantity $`qN`$ has been used as control parameter, where $`q=\frac{1r}{2}`$, and simulations were made with different values of $`N`$ ranging from $`N=2300`$ to $`19600`$. If the system were conservative it would remain in a perfectly homogeneous state at temperature $`T_0`$ with a homogeneous area density $`\rho _A=0.01`$. Because there is dissipation the dimensionless number density $`n^{}`$ has a maximum in the middle, $`n_{\mathrm{𝗆𝖺𝗑}}^{}`$, and the dimensionless temperature $`T^{}`$ has a minimum, $`T_{\mathrm{𝗆𝗂𝗇}}^{}`$, also at the center of the system. The pressure is uniform and $`p^{}=n_{\mathrm{max}}^{}T_{\mathrm{𝗆𝗂𝗇}}^{}`$. Hydrodynamic equations were derived using a moment expansion method. This method has the fourth cumulant, $`\kappa `$, of the velocity distribution as a parameter and three possible $`\kappa `$’s where considered. These are: $`\kappa =0`$ which implies that the reference distribution function in a Maxwellian, and two $`\kappa `$’s of the form (8), one with parameters $`b_k`$ as in (9) which corresponds to using $`f_{\mathrm{𝖧𝖢𝖲}}`$ as the reference function and finally using the values $`b_k`$, Eq. (37), numerically determined to ensure that the correct values of the density at the middle of the system come out. The last case gives our reference function $`f_0`$. The empirical rational form of $`\kappa =\kappa (q)`$, Eq. (8), in the case of the successful distribution $`f_0`$ turns out to be independent of the size of the system. When this $`f_0`$ is used, the comparison between the predictions and simulation results for $`n_{\mathrm{𝗆𝖺𝗑}}^{}`$, $`p^{}`$, $`P_{xx}^{}`$, $`P_{yy}^{}`$ against dissipation, and the density, temperature and heat flux profiles are very good. The obvious conclusion is that the formalism based on $`f_0`$ gives an excellent description of the behavior of the system for values of $`qN`$ up to 200 and it gives a reasonable description up to $`qN`$ nearly 300 (slightly above $`qN300`$ clustering begins), while the other formalisms fail beyond about $`qN=40`$. Even though the fourth cumulant $`\kappa `$ of $`f_0`$ is obtained to fit the results of a particular hydrodynamic regime, we expect that the theory with this expression for $`\kappa `$ is still valid for other regimes. ## VI Acknowledgments This work has been partly financed by Fondecyt research grant 296-0021 (R.R.), Fondecyt research grant 1990148 (D.R.), Fondecyt research grant 197-0786 (P.C.) and by FONDAP grant 11980002. One of us (R.S.) acknowledges a grant from MIDEPLAN.
warning/0001/astro-ph0001303.html
ar5iv
text
# Weak Lensing of the CMB: A Harmonic Approach ## I Introduction As the cosmic microwave background (CMB) photons propagate from the last scattering surface through intervening large-scale structure, they are gravitationally lensed. Weak lensing effects on the the temperature and polarization distributions of the cosmic microwave background is already a well-studied field. As in other aspects of the field, early work treating the effects on the temperature correlation function has largely been superceded by harmonic space power spectrum analyses in the post-COBE era . In harmonic space, the physical processes of anisotropy formation are most directly manifest. However for weak lensing in the CMB, correlation function underpinnings have typically remained, forcing transformations between real and Fourier space to define the effect in a small-angle (flat-sky) approximation. Exceptions include recent work on the non-Gaussianity of the lensed temperature field where a direct harmonic space approach has been taken . In this paper, we provide a complete framework for the study of lensing effects in the temperature and polarization fields directly in harmonic space. Not only does this greatly simplify the power spectrum calculations but it also establishes a clear link between weak lensing power spectrum observables in wide-field galaxy surveys and CMB observables for cross-correlation studies. Furthermore, this approach is easily generalized to lensing on the full sky by replacing Fourier harmonics with spherical harmonics. We show that counterintuitively, corrections from employing an exact all-sky treatment are not confined to large angles. The second order nature of the effect brings in large scale power through mode coupling. Since the all-sky expressions are as simple to evaluate as their flat-sky approximations, which themselves are much simpler to evaluate than the correlation function analogues, they should be employed where full accuracy is required, e.g. for the analysis of precise measurements from CMB satellite missions. Beyond the power spectrum, lensing induces three point correlations in the CMB through its correlation with secondary anisotropies , even when the intrinsic distribution at last scattering is Gaussian. Detection of these effects in the temperature maps however are severely limited by cosmic variance. The primary anisotropies themselves act at as Gaussian noise for these purposes. In this case, the low level at which the CMB is polarized can be an asset not a liability. Three point correlations involving the polarization, where orientation plays a role, are most simply considered with their harmonic space analogue, the bispectrum. We introduce polarization and polarization-temperature bispectra and show that they can have signal-to-noise advantages over those involving the temperature alone. The outline of the paper is as follows. In §II, we treat the basic elements of the cosmological framework, CMB temperature and polarization, and weak lensing needed to understand these effects. Detailed derivations are presented in a series of Appendices: A covers the all-sky weak lensing approach, B the evaluation of the all-sky formulae and C the correspondence between the flat and all sky approaches for scalar, vector and tensor fields on the sky. The lensing effects on the power spectrum are treated in the flat-sky approximation in §III and in the exact all-sky approach in §IV. In §V, we study the effects of lensing on the bispectra of the temperature and polarization distributions. We conclude in §VI. ## II Formalism In this section, we review and develop the formalism necessary for calculating lensing effects in the CMB. We review the relevant properties of the adiabatic cold dark matter (CDM) model in §II A. In §II B, we discuss the power spectra and bispectra of the temperature fluctuations, polarization and temperature-polarization cross correlation. Finally in §II C, we review the properties of weak lensing relevant for the CMB calculation. ### A Cosmological Model We work in the context of the adiabatic CDM family of models, where structure forms through the gravitational instability of the CDM in a background Friedmann-Robertson-Walker metric. In units of the critical density $`3H_0^2/8\pi G`$, where $`H_0=100h`$ km s<sup>-1</sup> Mpc<sup>-1</sup> is the Hubble parameter today, the contribution of each component is denoted $`\mathrm{\Omega }_i`$, $`i=c`$ for the CDM, $`b`$ for the baryons, $`\mathrm{\Lambda }`$ for the cosmological constant. It is convenient to define the auxiliary quantities $`\mathrm{\Omega }_m=\mathrm{\Omega }_c+\mathrm{\Omega }_b`$ and $`\mathrm{\Omega }_K=1_i\mathrm{\Omega }_i`$, which represent the matter density and the contribution of spatial curvature to the expansion rate respectively. The expansion rate $$H^2=H_0^2\left[\mathrm{\Omega }_m(1+z)^3+\mathrm{\Omega }_K(1+z)^2+\mathrm{\Omega }_\mathrm{\Lambda }\right].$$ (1) then determines the comoving conformal distance to redshift $`z`$, $$D(z)=_0^z\frac{H_0}{H(z^{})}𝑑z^{},$$ (2) in units of the Hubble distance today $`H_0^1=2997.9h^1`$Mpc. The comoving angular diameter distance $$D_A=\mathrm{\Omega }_K^{1/2}\mathrm{sinh}(\mathrm{\Omega }_K^{1/2}D),$$ (3) plays an important role in lensing. Note that as $`\mathrm{\Omega }_K0`$, $`D_AD`$. The adiabatic CDM model possesses a power spectrum of fluctuations in the gravitational potential $`\mathrm{\Phi }`$ $$\mathrm{\Delta }_\mathrm{\Phi }^2(k,z)=\frac{k^3}{2\pi ^2}P_\mathrm{\Phi }=A(z)\left(\frac{k}{H_0}\right)^{n1}T^2(k),$$ (4) where the the transfer function is normalized to $`T(0)=1`$. We employ the CMBFast code to determine $`T(k)`$ at intermediate scales and extend it to small scales using analytic fits . The cosmological Poisson equation relates the power spectra of the potential and density perturbations $`\delta `$ $$\mathrm{\Delta }_\mathrm{\Phi }^2=\frac{9}{4}\left(\frac{H_0}{k}\right)^4\left(1+3\frac{H_0^2}{k^2}\mathrm{\Omega }_K\right)^2\mathrm{\Omega }_m^2(1+z)^2\mathrm{\Delta }_\delta ^2,$$ (5) and gives the relationship between their relative normalization $$A(z)=\frac{9}{4}\left(1+3\frac{H_0^2}{k^2}\mathrm{\Omega }_K\right)^2\mathrm{\Omega }_m^2F(z)\delta _H^2.$$ (6) Here $`\delta _H`$ is the amplitude of present-day density fluctuations at the Hubble scale; we adopt the COBE normalization for $`\delta _H`$ . $`F(z)/(1+z)`$ is the growth rate of linear density perturbations $`\delta (z)=F(z)\delta (0)/(1+z)`$ $$F(z)(1+z)\frac{H(z)}{H_0}_z^{\mathrm{}}𝑑z^{}(1+z^{})\left(\frac{H_0}{H(z^{})}\right)^3.$$ (7) For the matter dominated regime where $`H(1+z)^{3/2}`$, $`F`$ is independent of redshift. Although we maintain generality in all derivations, we illustrate our results with a $`\mathrm{\Lambda }`$CDM model. The parameters for this model are $`\mathrm{\Omega }_c=0.30`$, $`\mathrm{\Omega }_b=0.05`$, $`\mathrm{\Omega }_\mathrm{\Lambda }=0.65`$, $`h=0.65`$, $`Y_p=0.24`$, $`n=1`$, and $`\delta _H=4.2\times 10^5`$. This model has mass fluctuations on the $`8h`$ Mpc<sup>-1</sup> scale in accord with the abundance of galaxy clusters $`\sigma _8=0.86`$. A reasonable value here is important since the lensing calculation is second order. ### B CMB We decompose the CMB temperature perturbation on the sky $`\mathrm{\Theta }(\widehat{𝐧})=\mathrm{\Delta }T(\widehat{𝐧})/T`$ into its multipole moments $$\mathrm{\Theta }(\widehat{𝐧})=\underset{lm}{}\mathrm{\Theta }_{lm}Y_l^m(\widehat{𝐧}).$$ (8) The polarization on the sky is represented by the trace-free symmetric Stokes matrix on the sky $`𝐏(\widehat{𝐧})`$ $`=`$ $`{}_{+}{}^{}X(\widehat{𝐧})(𝐦_+𝐦_+)+{}_{}{}^{}X(\widehat{𝐧})(𝐦_{}𝐦_{}),`$ (9) where $`{}_{\pm }{}^{}X(\widehat{𝐧})`$ $`=`$ $`Q(\widehat{𝐧})\pm iU(\widehat{𝐧}),`$ (10) $`𝐦_\pm `$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}(\widehat{𝐞}_\theta i\widehat{𝐞}_\varphi ).`$ (11) The complex Stokes parameter $`{}_{\pm }{}^{}X`$ is a spin-2 object which can be decomposed in the spin-spherical harmonics $${}_{\pm }{}^{}X(\widehat{𝐧})=\underset{lm}{}{}_{\pm }{}^{}X_{lm}^{}{}_{\pm 2}{}^{}Y_{l}^{m}(\widehat{𝐧}).$$ (12) We have assumed that the Stokes $`V`$ parameter vanishes as appropriate for cosmological perturbations; for a full set add the term $`Vϵ_{ij}`$ to the polarization matrix, where $`ϵ_{ij}`$ is the Levi-Civita tensor. Due to the parity properties of the spin-spherical harmonics $${}_{s}{}^{}Y_{l}^{m}(1)^l{}_{s}{}^{}Y_{l}^{m},$$ (13) one introduces the parity eigenstates $${}_{\pm }{}^{}X_{lm}^{}=E_{lm}\pm iB_{lm},$$ (14) such that $`E_{lm}`$ just like $`\mathrm{\Theta }_{lm}`$ has parity $`(1)^l`$ (“electric” parity) whereas $`B_{lm}`$ has parity $`(1)^{l+1}`$ (“magnetic” parity). Density (scalar) fluctuations in linear theory only stimulate the $`E`$ component of polarization. The power spectra and cross correlation of these quantities is defined as $$X_{lm}^{}X_{l^{}m^{}}^{}=\delta _{l,l^{}}\delta _{m,m^{}}C_l^{XX^{}},$$ (15) where $`X`$ and $`X^{}`$ can take on the values $`\mathrm{\Theta }`$,$`E`$,$`B`$. Note that the cross power spectra between $`B`$ and $`\mathrm{\Theta }`$ or $`E`$ have odd total parity and thus vanish assuming anisotropy formation is a parity invariant process. The bispectrum is defined as $$X_{lm}X_{l^{}m^{}}^{}X_{l^{\prime \prime }m^{\prime \prime }}^{\prime \prime }=\left(\begin{array}{ccc}l& l^{}& l^{\prime \prime }\\ m& m^{}& m^{\prime \prime }\end{array}\right)B_{ll^{}l^{\prime \prime }}^{XX^{}X^{\prime \prime }},$$ (16) and vanishes if the fluctuations are Gaussian. Even in the presence of non-Gaussianity due to non-linear but parity-conserving sources, bispectra involving an even number of magnetic parity terms (including zero) vanish for $`l+l^{}+l^{\prime \prime }=`$odd and those involving an odd number vanish for $`l+l^{}+l^{\prime \prime }=`$even. For a small section of the sky or high multipole moments, it is sufficient to treat the sky as flat. In the flat-sky approximation, the Fourier moments of the temperature fluctuations are given as $`\mathrm{\Theta }(\widehat{𝐧})`$ $`=`$ $`{\displaystyle \frac{d^2𝐥}{(2\pi )^2}\mathrm{\Theta }(𝐥)e^{i𝐥\widehat{𝐧}}},`$ (17) and the polarization as $${}_{\pm }{}^{}X(\widehat{𝐧})=\frac{d^2𝐥}{(2\pi )^2}{}_{\pm }{}^{}X(𝐥)e^{\pm 2i(\phi _l\phi )}e^{i𝐥\widehat{𝐧}},$$ (18) where $`\phi _l`$ is azimuthal angle of $`𝐥`$. Again one separates the Stokes moments as $${}_{\pm }{}^{}X(𝐥)=E(𝐥)\pm iB(𝐥).$$ (19) As in the all-sky case, the power spectra and cross correlations can be defined as with power spectra $`X^{}(𝐥)X^{}(𝐥^{})`$ $`=`$ $`(2\pi )^2\delta (𝐥𝐥^{})C_{(l)}^{XX^{}},`$ (20) $`X^{}(𝐥)X^{}(𝐥^{})X^{\prime \prime }(𝐥^{\prime \prime })`$ $`=`$ $`(2\pi )^2\delta (𝐥𝐥^{}𝐥^{\prime \prime })B_{(𝐥,𝐥^{},𝐥^{\prime \prime })}^{XX^{}X^{\prime \prime }}.`$ (21) The power spectra for the fiducial $`\mathrm{\Lambda }`$CDM model are shown in Figs. 1 and 2. In Appendix C, we establish the correspondence between the all-sky and flat-sky spectra. For the power spectra and bispectra, $`C_l^{XX^{}}`$ $``$ $`C_{(l)}^{XX^{}},`$ (22) $`B_{ll^{}l^{\prime \prime }}^{XX^{}X^{\prime \prime }}`$ $``$ $`\left(\begin{array}{ccc}l& l^{}& l^{\prime \prime }\\ 0& 0& 0\end{array}\right)\sqrt{{\displaystyle \frac{(2l+1)(2l^{}+1)(2l^{\prime \prime }+1)}{4\pi }}}`$ (26) $`B_{(𝐥,𝐥^{},𝐥^{\prime \prime })}^{XX^{}X^{\prime \prime }},`$ for sufficiently high $`l`$’s. For the bispectra, we have assumed that the triplet is composed of an even number of magnetic parity ($`B`$) objects such that it vanishes for $`l+l^{}+l^{\prime \prime }=`$ odd. For combinations involving an odd number (e.g. $`B\mathrm{\Theta }\mathrm{\Theta }`$), the Wigner-3$`j`$ symbol should be replaced with its algebraic approximation (B4) but with $`l+l^{}+l^{\prime \prime }=`$ even terms set to zero instead. However the overall sign depends on the orientation of the triangle in the flat-sky approximation since the bispectrum is then antisymmetric to reflections about either axis. ### C Weak Lensing In the so-called Born approximation where lensing effects are evaluated on the the null-geodesics of the unlensed photons, all effects can be conveniently encapsulated in the projected potential $$\varphi (\widehat{𝐧})=2𝑑Dg_\varphi (D)\mathrm{\Phi }(𝐱(\widehat{𝐧}),D),$$ (27) where $$g_\varphi (D)=\frac{1}{D_A(D)}_D^{\mathrm{}}𝑑D^{}\frac{D_A(D^{}D)}{D_A(D^{})}g_s(D^{}).$$ (28) For the CMB, the source distribution $`g_s`$ is the Thomson visibility and may be replaced by a delta function at the last scattering surface $`D_s=D(z10^3)`$; for galaxy weak lensing this is the distance distribution of the sources. We explicitly relate this quantity to the more familiar convergence and shear in Appendix A. Note that the deflection angle is given by the angular gradient $`𝜶`$$`(\widehat{𝐧})=\varphi (\widehat{𝐧})`$. As with the temperature perturbations, we can decompose the lensing potential into multipole moments $$\varphi (\widehat{𝐧})=\underset{lm}{}\varphi _{lm}Y_l^m(\widehat{𝐧}),$$ (29) or Fourier moments as $$\varphi (\widehat{𝐧})=\frac{d^2𝐥}{(2\pi )^2}\varphi (𝐥)e^{i𝐥\widehat{𝐧}},$$ (30) The power spectra of the lensing potential in the all-sky and flat-sky cases as $`\varphi _{lm}^{}\varphi _{l^{}m^{}}`$ $`=`$ $`\delta _{l,l^{}}\delta _{m,m^{}}C_l^{\varphi \varphi },`$ (31) $`\varphi ^{}(𝐥)\varphi (𝐥^{})`$ $`=`$ $`(2\pi )^2\delta (𝐥𝐥^{})C_{(l)}^{\varphi \varphi },`$ (32) where again $`C_{(l)}^{\varphi \varphi }=C_l^{\varphi \varphi }`$. The lensing potential also develops a bispectrum in the non-linear density regime, $`\varphi _{lm}\varphi _{l^{}m^{}}\varphi _{l^{\prime \prime }m^{\prime \prime }}`$ $`=`$ $`\left(\begin{array}{ccc}l& l^{}& l^{\prime \prime }\\ m& m^{}& m^{\prime \prime }\end{array}\right)B_{ll^{}l^{\prime \prime }}^{\varphi \varphi \varphi },`$ (35) $`\varphi (𝐥)\varphi (𝐥^{})\varphi (𝐥^{\prime \prime })`$ $`=`$ $`(2\pi )^2\delta (𝐥𝐥^{}𝐥^{\prime \prime })B_{(𝐥,𝐥^{},𝐥^{\prime \prime })}^{XX^{}X^{\prime \prime }},`$ (36) which is responsible for skewness in convergence maps and other higher order effects. Since the lensing potential is not affected by non-linearity until very high multipoles (see Fig. 3), we neglect these terms here. Finally, the lensing potential can also be correlated with secondary temperature and polarization anisotropies , so that one must also consider the cross power spectra $`X_{lm}^{}\varphi _{l^{}m^{}}`$ $`=`$ $`\delta _{l,l^{}}\delta _{m,m^{}}C_l^{X\varphi },`$ (37) $`X^{}(𝐥)\varphi (𝐥^{})`$ $`=`$ $`(2\pi )^2\delta (𝐥𝐥^{})C_{(l)}^{X\varphi },`$ (38) in the all and flat sky limits. To calculate the power spectra of the lensing potential for a given cosmology one expands the gravitational potential in eqn. (27) in plane waves and then expanding the plane waves in spherical harmonics. The result is $`C_l^{\varphi \varphi }`$ $`=`$ $`4\pi {\displaystyle \frac{dk}{k}\mathrm{\Delta }_\mathrm{\Phi }^2(k,z)[I_l^{\mathrm{len}}(k)]^2},`$ (39) where $`I_{\mathrm{}}^{\mathrm{len}}(k)`$ $`=`$ $`{\displaystyle 𝑑DW^{\mathrm{len}}(D)j_l(\frac{k}{H_0}D)},`$ (40) $`W^{\mathrm{len}}(D)`$ $`=`$ $`2F(D){\displaystyle \frac{D_A(D_sD)}{D_A(D)D_A(D_s)}}.`$ (41) For curved universes, replace the spherical Bessel function with the ultra-spherical Bessel function. In the small scale limit, this expression may be replaced by its equivalent Limber approximated integral $`C_{(l)}^{\varphi \varphi }`$ $`=`$ $`{\displaystyle \frac{2\pi ^2}{l^3}}{\displaystyle 𝑑DD_A[W^{\mathrm{len}}(D)]^2\mathrm{\Delta }_\mathrm{\Phi }^2(k,0)}|_{k=l\frac{H_0}{D_A}},`$ (42) This expression also has the useful property that its non-linear analogue can be calculated with the replacement $`F(D)^2\mathrm{\Delta }_\mathrm{\Phi }^2(k,0)\mathrm{\Delta }_\mathrm{\Phi }^2(k,D),`$ (43) where the time-dependent non-linear power spectrum is given by the scaling formula and the Poisson equation (5). Since non-linear effects generally only appear at small angles, the full non-linear all-sky spectrum can be obtained by matching these expressions in the linear regime (see Fig. 3). Similarly, the cross correlation may be calculated for any secondary effect once its relation to the gravitational potential is known. We shall illustrate these results with the integrated Sachs-Wolfe effect. It contributes to temperature fluctuations as $$\mathrm{\Theta }^{\mathrm{ISW}}(\widehat{𝐧})=2𝑑D\dot{\mathrm{\Phi }}(𝐱(\widehat{𝐧}),D).$$ (44) It then follows that the all-sky cross correlation is given by $`C_l^{\mathrm{\Theta }\varphi }`$ $`=`$ $`4\pi {\displaystyle \frac{dk}{k}\mathrm{\Delta }_\mathrm{\Phi }^2(k)I_l^{\mathrm{len}}(k)I_l^{\mathrm{ISW}}(k)},`$ (45) where $`I_l^{\mathrm{ISW}}(k)`$ $`=`$ $`{\displaystyle 𝑑DW^{\mathrm{ISW}}(D)j_l(\frac{k}{H_0}D)},`$ (46) $`W^{\mathrm{ISW}}(D)`$ $`=`$ $`2\dot{F}(D),`$ (47) again with the understanding that one replaces the spherical Bessel function with the ultra-spherical Bessel functions for curved universes. Similarly the flat-sky expression becomes, $`C_{(l)}^{\mathrm{\Theta }\varphi }`$ $`=`$ $`{\displaystyle \frac{2\pi ^2}{l^3}}{\displaystyle 𝑑DD_AW^{\mathrm{ISW}}(D)W^{\mathrm{len}}(D)\mathrm{\Delta }_\mathrm{\Phi }^2(k)}|_{k=l\frac{H_0}{D_A}}.`$ (48) Figure 3 also shows the cross-correlation for the $`\mathrm{\Lambda }`$CDM cosmology. Cross lensing-CMB bispectrum terms can also included but require an external measure of lensing (e.g. a galaxy weak lensing survey) to be observable with three-point correlations. ## III Flat-Sky Power Spectra In this section, we calculate the effects of lensing on the CMB temperature (§III A), polarization and cross (§III B) power spectra. The simplicity of the resulting expressions have calculational and pedagogical advantages over the traditional flat-sky correlation function approach . However we also show why one cannot expect a flat-sky approach to be fully accurate even on small scales. ### A Temperature Weak lensing of the CMB remaps the primary anisotropy according to the deflection angle $`\varphi `$ $`\stackrel{~}{\mathrm{\Theta }}(\widehat{𝐧})`$ $`=`$ $`\mathrm{\Theta }(\widehat{𝐧}+\varphi )`$ (49) $`=`$ $`\mathrm{\Theta }(\widehat{𝐧})+_i\varphi (\widehat{𝐧})^i\mathrm{\Theta }(\widehat{𝐧})`$ (51) $`+{\displaystyle \frac{1}{2}}_i\varphi (\widehat{𝐧})_j\varphi (\widehat{𝐧})^i^j\mathrm{\Theta }(\widehat{𝐧})+\mathrm{}`$ Because surface brightness is conserved in lensing only changes the distribution of the anisotropies and has no effect on the isotropic part of the background. The Fourier coefficients of the lensed field then become $`\stackrel{~}{\mathrm{\Theta }}(𝐥)`$ $`=`$ $`{\displaystyle 𝑑\widehat{𝐧}\stackrel{~}{\mathrm{\Theta }}(\widehat{𝐧})e^{i𝐥\widehat{𝐧}}}`$ (52) $`=`$ $`\mathrm{\Theta }(𝐥){\displaystyle \frac{d^2𝐥_1}{(2\pi )^2}\mathrm{\Theta }(𝐥_1)L(𝐥,𝐥_1)},`$ (53) where $`L(𝐥,𝐥_1)`$ $`=`$ $`\varphi (𝐥𝐥_1)(𝐥𝐥_1)𝐥_1+{\displaystyle \frac{1}{2}}{\displaystyle \frac{d^2𝐥_2}{(2\pi )^2}\varphi (𝐥_2)}`$ (55) $`\times \varphi ^{}(𝐥_2+𝐥_1𝐥)(𝐥_2𝐥_1)(𝐥_2+𝐥_1𝐥)𝐥_1.`$ This determines the lensed power spectrum $`\stackrel{~}{\mathrm{\Theta }}^{}(𝐥)\stackrel{~}{\mathrm{\Theta }}(𝐥^{})`$ $`=`$ $`(2\pi )^2\delta (𝐥𝐥^{})\stackrel{~}{C}_l^{\mathrm{\Theta }\mathrm{\Theta }},`$ (56) as $`\stackrel{~}{C}_l^{\mathrm{\Theta }\mathrm{\Theta }}`$ $`=`$ $`\left(1l^2R\right)C_l^{\mathrm{\Theta }\mathrm{\Theta }}+{\displaystyle \frac{d^2𝐥_1}{(2\pi )^2}C_{|𝐥𝐥_1|}^{\mathrm{\Theta }\mathrm{\Theta }}C_{l_1}^{\varphi \varphi }}`$ (58) $`\times [(𝐥𝐥_1)𝐥_1]^2,`$ where $$R=\frac{1}{4\pi }\frac{dl}{l}l^4C_l^{\varphi \varphi }.$$ (59) The second term in eqn. (58) represents a convolution of the power spectra. Since $`l^4C_l^{\varphi \varphi }`$ peaks at low $`l`$’s compared with the peaks in the CMB (see Fig. 3), it can be considered as a narrow window function on $`C_l^{\mathrm{\Theta }\mathrm{\Theta }}`$ in the acoustic regime $`200l2000`$. It is useful to consider the limit that $`C_l^{\mathrm{\Theta }\mathrm{\Theta }}`$ is slowly varying. It may then be evaluated at $`𝐥𝐥_1𝐥`$ and taken out of the integral $`C_l^{\mathrm{\Theta }\mathrm{\Theta }}{\displaystyle \frac{d^2𝐥_1}{(2\pi )^2}C_l^{\varphi \varphi }(𝐥𝐥_1)^2}`$ $``$ $`l^2RC_l^{\mathrm{\Theta }\mathrm{\Theta }}.`$ (60) Note that the two terms in eqn. (56) cancel in this limit $$\stackrel{~}{C}_l^{\mathrm{\Theta }\mathrm{\Theta }}C_l^{\mathrm{\Theta }\mathrm{\Theta }}.$$ (61) This is the well known result that lensing shifts but does not create power on large scales. Intrinsic features with width $`\mathrm{\Delta }l`$ less than the $`l`$ of the peak in $`l^4C_l^{\varphi \varphi }`$ are washed out by the convolution (see Fig. 3). Note that in the $`\mathrm{\Lambda }`$CDM model this scale is $`l40`$. The implication is that for such a model, the smoothing effect even for high multipoles arises from such low multipoles that the flat-sky approach is suspect. On scales small compared with the damping length $`l2000`$, there is little intrinsic power in the CMB so that the first term in eqn. (58) can be ignored and the second term behaves instead like a smoothing of $`C_l^{\varphi \varphi }`$ of width $`\mathrm{\Delta }l`$ approximately the $`l`$ of the peak in $`l^4C_l^{\mathrm{\Theta }\mathrm{\Theta }}`$. Since $`C_l^{\varphi \varphi }`$ is very smooth itself, the term is approximately, $$\stackrel{~}{C}_l^{\mathrm{\Theta }\mathrm{\Theta }}C_l^{\mathrm{\Theta }\mathrm{\Theta }}+\frac{1}{2}l^2C_l^{\varphi \varphi }\frac{d^2𝐥_1}{(2\pi )^2}l_1^2C_{l_1}^{\mathrm{\Theta }\mathrm{\Theta }},$$ (62) where we have interchanged the roles of $`𝐥_1`$ and $`𝐥_1𝐥`$. The power generated is proportional to the lensing power at the same scale and may be approximated as the lensing of a pure temperature gradient . In this limit the flat-sky approximation should be fully adequate. ### B Polarization The lensing of the polarization field may be obtained by following the same steps as for the temperature field $`{}_{\pm }{}^{}\stackrel{~}{X}(\widehat{𝐧})`$ $`=`$ $`{}_{\pm }{}^{}X(\widehat{𝐧}+\varphi )`$ (63) $``$ $`{}_{\pm }{}^{}X(\widehat{𝐧})+_i\varphi (\widehat{𝐧})^i{}_{\pm }{}^{}X(\widehat{𝐧})`$ (65) $`+{\displaystyle \frac{1}{2}}_i\varphi (\widehat{𝐧})_j\varphi (\widehat{𝐧})^i^j{}_{\pm }{}^{}X(\widehat{𝐧}),`$ where we have used the shorthand notation $`{}_{\pm }{}^{}X=Q\pm iU`$. The Fourier coefficients of the lensed field are then $`{}_{\pm }{}^{}\stackrel{~}{X}(𝐥)`$ $`=`$ $`{}_{\pm }{}^{}X(𝐥){\displaystyle \frac{d^2𝐥_1}{(2\pi )^2}{}_{\pm }{}^{}X(𝐥_1)e^{\pm 2i(\phi _{l_1}\phi _l)}L(𝐥,𝐥_1)},`$ (66) where $`L`$ was defined in eqn. (55). Recalling that $`{}_{\pm }{}^{}X(𝐥)=E(𝐥)\pm iB(𝐥)`$, we obtain the power spectra directly $`\stackrel{~}{C}_l^{EE}`$ $`=`$ $`\left(1l^2R\right)C_l^{EE}+{\displaystyle \frac{1}{2}}{\displaystyle \frac{d^2𝐥_1}{(2\pi )^2}[(𝐥𝐥_1)𝐥_1]^2C_{|𝐥𝐥_1|}^{\varphi \varphi }}`$ (68) $`\times [(C_{l_1}^{EE}+C_{l_1}^{BB})+\mathrm{cos}(4\phi _{l_1})(C_{l_1}^{EE}C_{l_1}^{BB})],`$ $`\stackrel{~}{C}_l^{BB}`$ $`=`$ $`\left(1l^2R\right)C_l^{BB}+{\displaystyle \frac{1}{2}}{\displaystyle \frac{d^2𝐥_1}{(2\pi )^2}[(𝐥𝐥_1)𝐥_1]^2C_{|𝐥𝐥_1|}^{\varphi \varphi }}`$ (70) $`\times [(C_{l_1}^{EE}+C_{l_1}^{BB})\mathrm{cos}(4\phi _{l_1})(C_{l_1}^{EE}C_{l_1}^{BB})],`$ $`\stackrel{~}{C}_l^{\mathrm{\Theta }E}`$ $`=`$ $`\left(1l^2R\right)C_l^{\mathrm{\Theta }E}+{\displaystyle \frac{d^2𝐥_1}{(2\pi )^2}[(𝐥𝐥_1)𝐥_1]^2C_{|𝐥𝐥_1|}^{\varphi \varphi }}`$ (72) $`\times C_{l_1}^{\mathrm{\Theta }E}\mathrm{cos}(2\phi _{l_1}),`$ where recall that $`R`$ was defined in eqn. (59). The cross correlations between $`B`$ and $`\mathrm{\Theta }`$ or $`E`$ still vanish since lensing is parity conserving. Unlike the case of the temperature fluctuations, lensing does not conserve the broadband large scale power of the $`E`$ and $`B`$ , but only the total polarization power. For example, lensing will create a $`B`$ component in a field that originally had only an $`E`$-component. Furthermore, lensing actually destroys temperature-polarization cross correlations due to the lack of correlation with the generated $`B`$ polarization. From Fig. 1, one can see that the largest relative effect of lensing is on the correlation. ## IV All-Sky Power Spectra In this section, we treat lensing effects on the temperature (IV A), polarization and cross (IV B) power spectra in a full all-sky formalism. Corrections to the flat-sky results remain at the 10% even on small scales. Moreover, although the derivation appears more complicated, the end results for the power spectra are simple. They are as readily evaluated their the flat-sky counterparts and should be used in their stead. ### A Temperature In the all-sky case, the Fourier harmonics are replaced with spherical harmonics, and the lensed field becomes $`\stackrel{~}{\mathrm{\Theta }}_{lm}`$ $``$ $`\mathrm{\Theta }_{lm}+{\displaystyle 𝑑\widehat{𝐧}Y_l^m_i\varphi (\widehat{𝐧})^i\mathrm{\Theta }(\widehat{𝐧})}`$ (74) $`+{\displaystyle \frac{1}{2}}{\displaystyle 𝑑\widehat{𝐧}Y_l^m_i\varphi (\widehat{𝐧})_j\varphi (\widehat{𝐧})^i^j\mathrm{\Theta }(\widehat{𝐧})}`$ $`=`$ $`\mathrm{\Theta }_{lm}+{\displaystyle \underset{l_1m_1}{}}{\displaystyle \underset{l_2m_2}{}}\varphi _{l_1m_1}\mathrm{\Theta }_{l_2m_2}`$ (76) $`\times \left[I_{ll_1l_2}^{mm_1m_2}+{\displaystyle \frac{1}{2}}{\displaystyle \underset{l_3m_3}{}}\varphi _{l_3m_3}^{}J_{ll_1l_2l_3}^{mm_1m_2m_3}\right],`$ with the geometrical factors expressed as integrals over the spherical harmonics $`I_{ll_1l_2}^{mm_1m_2}`$ $`=`$ $`{\displaystyle 𝑑\widehat{𝐧}Y_l^m\left(_iY_{l_1}^{m_1}\right)\left(^iY_{l_2}^{m_2}\right)},`$ (77) $`J_{ll_1l_2l_3}^{mm_1m_2m_3}`$ $`=`$ $`{\displaystyle 𝑑\widehat{𝐧}Y_l^m\left(_iY_{l_1}^{m_1}\right)\left(_jY_{l_3}^{m_3}\right)^i^jY_{l_2}^{m_2}}.`$ (78) The lensed power spectrum then becomes $`\stackrel{~}{C}_l`$ $`=`$ $`C_l+{\displaystyle \underset{l_1l_2}{}}C_{l_1}^{\varphi \varphi }C_{l_2}^{\mathrm{\Theta }\mathrm{\Theta }}S_1+C_l^{\mathrm{\Theta }\mathrm{\Theta }}{\displaystyle \underset{l_1}{}}C_{l_1}^{\varphi \varphi }S_2,`$ (79) with $`S_1`$ $`=`$ $`{\displaystyle \underset{m_1m_2}{}}\left(I_{ll_1l_2}^{mm_1m_2}\right)^2,`$ (80) $`S_2`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{m_1}{}}J_{ll_1ll_1}^{mm_1mm_1}+\mathrm{cc},`$ (81) where “cc” denotes the complex conjugate and we have suppressed the $`l`$-indices. These formidable looking expressions simplify considerably. The second term may be rewritten through integration by parts and the identity $`^2Y_l^m=l(l+1)Y_l^m`$ , $`I_{ll_1l_2}^{mm_1m_2}`$ $`=`$ $`{\displaystyle \frac{1}{2}}[l_1(l_1+1)+l_2(l_2+1)l(l+1)]`$ (83) $`\times {\displaystyle }d\widehat{𝐧}Y_l^mY_{l_1}^{m_1}Y_{l_2}^{m_2}.`$ The remaining integral may be expressed in terms of the Wigner-3$`j`$ symbol through the general relation $`{\displaystyle 𝑑\widehat{𝐧}\left({}_{s_1}{}^{}Y_{l_1}^{m_1}\right){}_{s_2}{}^{}Y_{l_2}^{m_2}\left({}_{s_3}{}^{}Y_{l_2}^{m_3}\right)}=`$ (84) $`(1)^{m_1+s_1}\sqrt{{\displaystyle \frac{(2l_1+1)(2l_2+1)(2l_3+1)}{4\pi }}}`$ (85) $`\times \left(\begin{array}{ccc}l_1& l_2& l_3\\ s_1& s_2& s_3\end{array}\right)\left(\begin{array}{ccc}l_1& l_2& l_3\\ m_1& m_2& m_3\end{array}\right),`$ (90) where note that $`{}_{0}{}^{}Y_{l}^{m}=Y_l^m`$. It is therefore convenient to define $`F_{l_1l_2l_3}`$ $`=`$ $`{\displaystyle \frac{1}{2}}[l_2(l_2+1)+l_3(l_3+1)l_1(l_1+1)]`$ (94) $`\times \sqrt{{\displaystyle \frac{(2l_1+1)(2l_2+1)(2l_3+1)}{4\pi }}}\left(\begin{array}{ccc}l_1& l_2& l_3\\ 0& 0& 0\end{array}\right).`$ Finally the Wigner-3$`j`$ symbol obeys $`{\displaystyle \underset{m_1m_2}{}}\left(\begin{array}{ccc}l_1& l_2& l_3\\ m_1& m_2& m_3\end{array}\right)\left(\begin{array}{ccc}l_1& l_2& l_3\\ m_1& m_2& m_3\end{array}\right)={\displaystyle \frac{1}{2l_3+1}}.`$ (99) Putting these relations together, we find that $$S_1=\frac{1}{2l+1}\left(F_{ll_1l_2}\right)^2.$$ (100) An algebraic expression for the relevant Wigner-3$`j`$ symbol is given in the Appendix. The second term in eqn. (79) can be simplified by re-expressing the gradients of the spherical harmonics with spin-1 spherical harmonics. As shown in Appendix A, the spin-1 harmonics are the eigenmodes of vector fields on the sky and naturally appear in expressions for deflection angles. Note that there is a general relation for raising and lowering the spin of a spherical harmonic , $`𝐦_{}{}_{s}{}^{}Y_{l}^{m}`$ $`=`$ $`\sqrt{{\displaystyle \frac{(ls)(l+s+1)}{2}}}{}_{s+1}{}^{}Y_{l}^{m},`$ (101) $`𝐦_+{}_{s}{}^{}Y_{l}^{m}`$ $`=`$ $`\sqrt{{\displaystyle \frac{(l+s)(ls+1)}{2}}}{}_{s1}{}^{}Y_{l}^{m},`$ (102) so that $`Y_l^m=\sqrt{{\displaystyle \frac{l(l+1)}{2}}}\left[{}_{1}{}^{}Y_{l}^{m}𝐦_+{}_{1}{}^{}Y_{l}^{m}𝐦_{}\right].`$ (103) As an aside, we note that equation (100) can alternately be derived from this relation and the integral (90) with $`s=\pm 1`$. Further, we note that spin spherical harmonics also obey a sum rule $`{\displaystyle \underset{m}{}}{}_{s_1}{}^{}Y_{l}^{m}(\widehat{𝐧}){}_{s_2}{}^{}Y_{l}^{m}(\widehat{𝐧})`$ $`=`$ $`\sqrt{{\displaystyle \frac{2l+1}{4\pi }}}{}_{s_2}{}^{}Y_{l}^{s_1}(\mathrm{𝟎}).`$ (104) For the spin-1 harmonics $${}_{1}{}^{}Y_{l}^{1}(\mathrm{𝟎})={}_{1}{}^{}Y_{l}^{1}(\mathrm{𝟎})=\sqrt{\frac{2l+1}{4\pi }},$$ (105) and the others involving $`s_1,s_2=\pm 1`$ vanish. These results imply that $`{\displaystyle \underset{m}{}}_iY_l^m_jY_l^m`$ $`=`$ $`{\displaystyle \frac{1}{2}}l(l+1){\displaystyle \frac{2l+1}{4\pi }}[(𝐦_+)_i(𝐦_{})_j`$ (107) $`+(𝐦_{})_i(𝐦_+)_j].`$ To evaluate the second derivative term in equation (78), we again apply equation (102) to show that $`[(𝐦_+)_i(𝐦_{})_j+(𝐦_{})_i(𝐦_+)_j]^i^j{}_{s}{}^{}Y_{l}^{m}=`$ (108) $`[l(l+1)s^2]{}_{s}{}^{}Y_{l}^{m}.`$ (109) Putting these expressions together we obtain, $`S_2={\displaystyle \frac{1}{2}}l(l+1)l_1(l_1+1){\displaystyle \frac{2l_1+1}{4\pi }}.`$ (110) Finally combining expressions eqns. (79), (100) and (110), we have the following simple result $`\stackrel{~}{C}_l^{\mathrm{\Theta }\mathrm{\Theta }}`$ $`=`$ $`\left[1l(l+1)R\right]C_l^{\mathrm{\Theta }\mathrm{\Theta }}+{\displaystyle \underset{l_1,l_2}{}}C_{l_1}^{\varphi \varphi }C_{l_2}{\displaystyle \frac{\left(F_{ll_1l_2}\right)^2}{2l+1}},`$ (111) where $$R=\frac{1}{2}\underset{l_1}{}l_1(l_1+1)\frac{2l_1+1}{4\pi }C_{l_1}^{\varphi \varphi }.$$ (112) This expression is computationally no more involved than the flat-sky expression eqn. (58) and has the benefit of being exact. Since the lensing effect even at high $`l`$ in the CMB originates from the low order multipoles of $`\varphi `$, corrections due to the curvature of the sky are not confined to low $`l`$. We show in Fig. 1 that the correction causes a 10% difference in the effect. The change in $`\stackrel{~}{C}_l^{\mathrm{\Theta }\mathrm{\Theta }}`$ itself is even smaller (of order 1%). Nonetheless it is larger than the cosmic variance of these high multipoles and thus should be included in calculations for full accuracy. Corrections can be even larger in models with a red tilt $`n<1`$ in the initial spectrum. ### B Polarization The derivation of the all-sky generalization for polarization is superficially more involved but follows the same steps as in the temperature case and results in expressions that are no more difficult to evaluate. The lensed polarization multipoles are given by $`{}_{\pm }{}^{}X_{lm}^{}`$ $`=`$ $`{}_{\pm }{}^{}X_{lm}^{}+{\displaystyle \underset{l_1m_1}{}}{\displaystyle \underset{l_2m_2}{}}\varphi _{l_1m_1}{}_{\pm }{}^{}X_{l_2m_2}^{}`$ (114) $`\times \left[{}_{\pm 2}{}^{}I_{ll_1l_2}^{mm_1m_2}+{\displaystyle \frac{1}{2}}{\displaystyle \underset{l_3m_3}{}}\varphi _{l_3m_3}^{}{}_{\pm 2}{}^{}J_{ll_1l_2l_3}^{mm_1m_2m_3}\right],`$ with the geometrical factors expressed now as integrals over the spin-spherical harmonics $`{}_{\pm 2}{}^{}I_{ll_1l_2}^{mm_1m_2}`$ $`=`$ $`{\displaystyle 𝑑\widehat{𝐧}{}_{\pm 2}{}^{}Y_{l}^{m}\left(_iY_{l_1}^{m_1}\right)\left(^i{}_{\pm 2}{}^{}Y_{l_2}^{m_2}\right)},`$ (115) $`{}_{\pm 2}{}^{}J_{ll_1l_2l_3}^{mm_1m_2m_3}`$ $`=`$ $`{\displaystyle 𝑑\widehat{𝐧}{}_{\pm 2}{}^{}Y_{l}^{m}\left(_iY_{l_1}^{m_1}\right)\left(_iY_{l_3}^{m_3}\right)}`$ (117) $`\times \left(^i^j{}_{\pm 2}{}^{}Y_{l_2}^{m_2}\right).`$ Noting that $`{}_{\pm 2}{}^{}I_{ll_1l_2}^{mm_1m_2}`$ $`=`$ $`(1)^L{}_{2}{}^{}I_{ll_1l_2}^{mm_1m_2},`$ (118) where $`L=l+l_1+l_2`$ and recalling that $`{}_{\pm }{}^{}X_{lm}^{}=E_{lm}\pm iB_{lm}`$, the power spectra then become $`\stackrel{~}{C}_l^{EE}`$ $`=`$ $`C_l^{EE}+{\displaystyle \frac{1}{2}}{\displaystyle \underset{l_1l_2}{}}C_{l_1}^{\varphi \varphi }[(C_{l_2}^{EE}+C_{l_2}^{BB})`$ (121) $`+(1)^L(C_{l_2}^{EE}C_{l_2}^{BB})]_{22}S_1`$ $`+{\displaystyle \frac{1}{2}}C_l^{EE}{\displaystyle \underset{l_1}{}}C_{l_1}^{\varphi \varphi }\left({}_{2}{}^{}S_{2}^{}+{}_{2}{}^{}S_{2}^{}\right),`$ where $`{}_{22}{}^{}S_{1}^{}`$ $`=`$ $`{\displaystyle \underset{m_1m_2}{}}\left({}_{2}{}^{}I_{ll_1l_2}^{mm_1m_2}\right)^2,`$ (122) $`{}_{\pm 2}{}^{}S_{2}^{}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{m_1}{}}{}_{\pm 2}{}^{}J_{ll_1l_1l}^{mm_1m_1m}+\mathrm{cc}.`$ (123) The expression for $`\stackrel{~}{C}_l^{BB}`$ follows by interchanging $`EE`$ and $`BB`$. The cross power spectrum is $`\stackrel{~}{C}_l^{\mathrm{\Theta }E}`$ $`=`$ $`C_l^{\mathrm{\Theta }E}+{\displaystyle \frac{1}{2}}{\displaystyle \underset{l_1l_2}{}}C_{l_1}^{\varphi \varphi }C_{l_2}^{\mathrm{\Theta }E}[1+(1)^L]{}_{02}{}^{}S_{1}^{}`$ (125) $`+{\displaystyle \frac{1}{4}}C_l^{\mathrm{\Theta }E}{\displaystyle \underset{l_1}{}}C_{l_1}^{\varphi \varphi }\left({}_{2}{}^{}S_{2}^{}+{}_{2}{}^{}S_{2}^{}+2S_2\right),`$ with $${}_{02}{}^{}S_{1}^{}=\underset{m_1m_2}{}\left(I_{ll_1l_2}^{mm_1m_2}{}_{2}{}^{}I_{ll_1l_2}^{mm_1m_2}\right).$$ (126) Just as in the case for the temperature field, these expressions simplify considerably. The spin-$`2`$ harmonics are eigenfunctions of the angular Laplacian of a tensor $$^2{}_{\pm 2}{}^{}Y_{l}^{m}=[l(l+1)+4]{}_{\pm 2}{}^{}Y_{l}^{m},$$ (127) which follows from contracting indices in equation (109). It then follows that $`{}_{\pm 2}{}^{}I_{ll_1l_2}^{mm_1m_2}`$ $`=`$ $`{\displaystyle \frac{1}{2}}[l_1(l_1+1)+l_2(l_2+1)l(l+1)]`$ (129) $`{\displaystyle 𝑑\widehat{𝐧}\left({}_{\pm 2}{}^{}Y_{l}^{m}\right)Y_{l_1}^{m_1}\left({}_{\pm 2}{}^{}Y_{l_2}^{m_2}\right)}.`$ Comparison with eqn. (90) implies that it is convenient then to define the quantity $`{}_{2}{}^{}F_{l_1l_2l_3}^{}`$ $`=`$ $`{\displaystyle \frac{1}{2}}[l_2(l_2+1)+l_3(l_3+1)l_1(l_1+1)]`$ (133) $`\sqrt{{\displaystyle \frac{(2l_1+1)(2l_2+1)(2l_3+1)}{4\pi }}}\left(\begin{array}{ccc}l_1& l_2& l_3\\ 2& 0& 2\end{array}\right),`$ so that $`{}_{22}{}^{}S_{1}^{}`$ $`=`$ $`{\displaystyle \frac{1}{2l+1}}({}_{2}{}^{}F_{ll_1l_2}^{})^2,`$ (134) $`{}_{02}{}^{}S_{1}^{}`$ $`=`$ $`{\displaystyle \frac{1}{2l+1}}\left(F_{ll_1l_2}\right)\left({}_{2}{}^{}F_{ll_1l_2}^{}\right).`$ (135) The third term in equation (121) can be simplified by following the same steps for the analogous temperature term except for the replacement of $`s=0`$ with $`s=\pm 2`$ in equation (102). The result is $`{}_{\pm 2}{}^{}S_{2}^{}={\displaystyle \frac{1}{2}}[l(l+1)4]l_1(l_1+1){\displaystyle \frac{2l_1+1}{4\pi }}.`$ (136) Putting these relations together, we obtain the result for the power spectra $`\stackrel{~}{C}_l^{EE}`$ $`=`$ $`\left[1(l^2+l4)R\right]C_l^{EE}+{\displaystyle \frac{1}{2}}{\displaystyle \underset{l_1,l_2}{}}C_{l_1}^{\varphi \varphi }{\displaystyle \frac{\left({}_{2}{}^{}F_{ll_1l_2}^{}\right)^2}{2l+1}}`$ (138) $`\left[C_{l_2}^{EE}+C_{l_2}^{BB}+(1)^L(C_{l_2}^{EE}C_{l_2}^{BB})\right],`$ $`\stackrel{~}{C}_l^{BB}`$ $`=`$ $`\left[1(l^2+l4)R\right]C_l^{BB}+{\displaystyle \frac{1}{2}}{\displaystyle \underset{l_1,l_2}{}}C_{l_1}^{\varphi \varphi }{\displaystyle \frac{\left({}_{2}{}^{}F_{ll_1l_2}^{}\right)^2}{2l+1}}`$ (140) $`\left[C_{l_2}^{EE}+C_{l_2}^{BB}(1)^L(C_{l_2}^{EE}C_{l_2}^{BB})\right],`$ $`\stackrel{~}{C}_l^{\mathrm{\Theta }E}`$ $`=`$ $`\left[1(l^2+l2)R\right]C_l^{\mathrm{\Theta }E}+{\displaystyle \underset{l_1,l_2}{}}C_{l_1}^{\varphi \varphi }`$ (142) $`{\displaystyle \frac{\left(F_{ll_1l_2}{}_{2}{}^{}F_{ll_1l_2}^{}\right)}{2l+1}}C_{l_2}^{\mathrm{\Theta }E}.`$ Recall that $`L=l+l_1+l_2`$ and $`R`$ was defined in eqn. (112). These expressions are plotted for the $`\mathrm{\Lambda }`$CDM model in Fig. 2. ## V Flat and All Sky Bispectra In this section, we consider the lensing contributions to CMB bispectra through the correlation with secondary anisotropies. We begin by reviewing the calculations for the temperature bispectrum as previously treated by . We then introduce the polarization and cross bispectra which in principle have signal-to-noise advantages over the temperature bispectra. We illustrate the formalism with a concrete calculation of the effect due to the ISW secondary anisotropy. ### A Temperature Contributions to the temperature bispectra from the cross power spectrum $`C_l^{\mathrm{\Theta }\varphi }`$ discussed in §II C follow immediately from the first order lensing term, i.e. eqn. (76) for the all-sky bispectrum , $$B_{l_1l_2l_3}^{\mathrm{\Theta }\mathrm{\Theta }\mathrm{\Theta }}=F_{l_1l_2l_3}C_{l_2}^{\mathrm{\Theta }\varphi }C_{l_3}^{\mathrm{\Theta }\mathrm{\Theta }}+5\mathrm{p}\mathrm{e}\mathrm{r}\mathrm{m}.,$$ (143) and eqn. (53) for the flat sky bispectrum $`B_{(𝐥_1,𝐥_2,𝐥_3)}^{\mathrm{\Theta }\mathrm{\Theta }\mathrm{\Theta }}`$ $`=`$ $`(𝐥_2𝐥_3)C_{l_2}^{\mathrm{\Theta }\varphi }C_{l_3}^{\mathrm{\Theta }\mathrm{\Theta }}+5\mathrm{p}\mathrm{e}\mathrm{r}\mathrm{m}.`$ (144) One can show that these relations satisfy the general expression for the correspondence between flat and all sky bispectra eqn. (26) by noting that $$𝐥_2𝐥_3=\frac{1}{2}(l_2^2+l_3^2l_1^2),$$ (145) since the angles of a triangle is fully defined by the length of its sides. Note that there can be strong cancellation between the terms in the permutation in both cases. As we have seen, the spectrum of $`\varphi `$ is generally peaked to low multipoles implying a corresponding weighting of $`C_l^{\mathrm{\Theta }\varphi }`$ to low multipoles for secondary anisotropies that correlate strongly with $`\varphi `$. In this case the triangles $`(l_1,l_2,l_3)`$ that contribute most strongly are highly flattened such that two sides nearly coincide in length $`l_1l_3l_2`$. In this case, contributions $`l_1^2`$ and $`l_3^2`$ in eqn. (145) are cancelled off the permutation $`𝐥_3𝐥_1`$ leaving only a term of order $`l_2^2`$. These considerations also signal problems for the flat-sky expressions. It is important to know what on scales most of the detectable signal is coming from. In the all-sky formalism, the signals from the $`m`$ modes are added together with weights given by the Wigner-3$`j`$ symbol $`B_{l_1l_2l_3}^{\mathrm{\Theta }\mathrm{\Theta }\mathrm{\Theta }}`$ $`=`$ $`{\displaystyle \underset{m_1m_2m_3}{}}\left(\begin{array}{ccc}l_1& l_2& l_3\\ m_1& m_2& m_3\end{array}\right)\mathrm{\Theta }_{l_1m_1}\mathrm{\Theta }_{l_2m_2}\mathrm{\Theta }_{l_3m_3}.`$ (148) For the small effects due to the correlation of secondary anisotropies with lensing, the covariance of the bispectrum estimators is dominated by the Gaussian noise from the power spectrum $`\mathrm{Cov}=C_{l_1}^{\mathrm{\Theta }\mathrm{\Theta }}C_{l_2}^{\mathrm{\Theta }\mathrm{\Theta }}C_{l_3}^{\mathrm{\Theta }\mathrm{\Theta }}\delta _{l_1l_1^{}}\delta _{l_2l_2^{}}\delta _{l_3l_3^{}}+5\mathrm{p}\mathrm{e}\mathrm{r}\mathrm{m}.,`$ (150) where the permutations are in the indices of the $`l^{}`$ triplet. The overall signal-to-noise becomes $`\left({\displaystyle \frac{S}{N}}\right)^2={\displaystyle \underset{l_1l_2l_3}{}}{\displaystyle \underset{l_1^{}l_2^{}l_3^{}}{}}B_{l_1l_2l_3}^{\mathrm{\Theta }\mathrm{\Theta }\mathrm{\Theta }}[\mathrm{Cov}^1]B_{l_1^{}l_2^{}l_3^{}}^{\mathrm{\Theta }\mathrm{\Theta }\mathrm{\Theta }}.`$ (151) The covariance is in general diagonal in the 6$`\times `$6 blocks of permutations of $`(l_1,l_2,l_3)`$ and for this simple case of the temperature bispectrum, the blocks are proportional to the trivial matrix of all ones. The result is one can take a simple sum over all distinct triplets or equivalently divide the full sum by a factor of 6, $$\left(\frac{S}{N}\right)^2=\underset{l_1l_2l_3}{}\frac{(B_{l_1l_2l_3}^{\mathrm{\Theta }\mathrm{\Theta }\mathrm{\Theta }})^2}{6C_{l_1}^{\mathrm{\Theta }\mathrm{\Theta }}C_{l_2}^{\mathrm{\Theta }\mathrm{\Theta }}C_{l_3}^{\mathrm{\Theta }\mathrm{\Theta }}},$$ (152) for a cosmic variance limited experiment. For a realistic experiment with noise from the detectors and residual foregrounds, one simply replaces $$C_l^{XX^{}}C_l^{XX^{}}+C_l^{XX^{}(\mathrm{noise})},$$ (153) here and below. Note that one can also construct the Fisher information matrix of the bispectrum along these lines . Correspondingly, in the flat-sky one constructs the optimal inverse-variance weighted statistic (see also Appendix C) $$\left(\frac{S}{N}\right)^2=\frac{f_{\mathrm{sky}}}{\pi }\frac{1}{(2\pi )^2}d^2l_1d^2l_2\frac{[B_{(l_1,l_2,l_3)}^{\mathrm{\Theta }\mathrm{\Theta }\mathrm{\Theta }}]^2}{6C_{l_1}^{\mathrm{\Theta }\mathrm{\Theta }}C_{l_2}^{\mathrm{\Theta }\mathrm{\Theta }}C_{l_3}^{\mathrm{\Theta }\mathrm{\Theta }}},$$ (154) where $`f_{\mathrm{sky}}`$ is the fraction of the sky covered. We show that these expressions are equivalent in the high $`l`$, $`f_{\mathrm{sky}}=1`$ limit in Appendix C. Thus the extra factor of $`f_{\mathrm{sky}}`$ can be included in the all-sky expression to approximate the effects incomplete sky coverage due to exclusion of regions contaminated by galactic foregrounds. The weighting of the modes is such that the quantity of interest in the lensing-temperature correlation is $`l^3C_l^{\mathrm{\Theta }\varphi }`$ where the extra factor of $`l`$ over the straight bispectrum contribution comes from the square root of the volume factor in $`l`$-space. This quantity is plotted in Fig. 3 for the cross correlation with the ISW effect. The implication is that for this effect, full accuracy requires an all-sky approach and we shall hereafter use this to evaluate the signal-to-noise. The overall signal-to-noise as a function of the largest $`l`$ included in the sum is shown in Fig. 4 for a cosmic variance limited experiment and the Planck satellite (see for the specification of the noise). Note that the Planck satellite is effectively cosmic variance limited to $`l1000`$ and even so the $`S/N`$ is only of order a few . ### B Polarization and Cross Correlation Bispectra involving the $`E`$ and $`B`$ parity polarization will also receive contributions from the correlation induced by lensing. Although these signals are smaller than the temperature bispectrum in an absolute sense, we have seen that the main obstacle in detecting the temperature bispectrum is cosmic variance from the Gaussian contributions. We begin by analyzing terms that do not involve the $`B`$-parity polarization. For these all-sky bispectra, only terms with $`Ll_1+l_2+l_3`$=even are non-vanishing, and we will implicitly assume that only even terms are considered. With the help of eqns. (65) and (114), we can immediately write the all and flat sky results as $`B_{l_1l_2l_3}^{E\mathrm{\Theta }\mathrm{\Theta }}`$ $`=`$ $`{}_{2}{}^{}F_{l_1l_2l_3}^{}C_{l_2}^{\mathrm{\Theta }\varphi }C_{l_3}^{\mathrm{\Theta }E}`$ (156) $`+F_{l_2l_1l_3}(C_{l_1}^{E\varphi }C_{l_3}^{\mathrm{\Theta }\mathrm{\Theta }}+C_{l_1}^{\mathrm{\Theta }E}C_{l_3}^{\mathrm{\Theta }\varphi })+(l_2l_3),`$ $`B_{(𝐥_1,𝐥_2,𝐥_3)}^{E\mathrm{\Theta }\mathrm{\Theta }}`$ $`=`$ $`(𝐥_2𝐥_3)\mathrm{cos}2\phi _{31}C_{l_2}^{\mathrm{\Theta }\varphi }C_{l_3}^{\mathrm{\Theta }E}`$ (158) $`(𝐥_1𝐥_3)(C_{l_1}^{E\varphi }C_{l_3}^{\mathrm{\Theta }\mathrm{\Theta }}+C_{l_1}^{\mathrm{\Theta }E}C_{l_3}^{\mathrm{\Theta }\varphi })+(𝐥_2𝐥_3),`$ where $$\phi _{AB}=\phi _{l_A}\phi _{l_B}.$$ (159) The general correspondence between the flat and all sky expressions in eqn. (26) is established by the use of eqn. (145) the approximation discussed in the Appendix $$\left(\begin{array}{ccc}l_1& l_2& l_3\\ 2& 0& 2\end{array}\right)\mathrm{cos}2\phi _{31}\left(\begin{array}{ccc}l_1& l_2& l_3\\ 0& 0& 0\end{array}\right),$$ (160) for $`L=`$even. The cancellation for flattened triangles discussed in §V A still applies and is easiest to see in the flat-sky limit: the flatness of the triangles implies $`\mathrm{cos}2\phi _{31}1`$. For the $`S/N`$ calculation, note that the covariance is given by $`\mathrm{Cov}`$ $`=`$ $`C_{l_1}^{EE}C_{l_2}^{\mathrm{\Theta }\mathrm{\Theta }}C_{l_3}^{\mathrm{\Theta }\mathrm{\Theta }}\delta _{l_1l_1^{}}\delta _{l_2l_2^{}}\delta _{l_3l_3^{}}`$ (163) $`+C_{l_1}^{\mathrm{\Theta }E}C_{l_2}^{\mathrm{\Theta }\mathrm{\Theta }}C_{l_3}^{\mathrm{\Theta }E}\delta _{l_1l_2^{}}\delta _{l_2l_3^{}}\delta _{l_3l_1^{}}`$ $`+C_{l_1}^{\mathrm{\Theta }E}C_{l_2}^{\mathrm{\Theta }E}C_{l_3}^{\mathrm{\Theta }\mathrm{\Theta }}\delta _{l_1l_3^{}}\delta _{l_2l_1^{}}\delta _{l_3l_2^{}}+(l_2^{}l_3^{}),`$ so that a full calculation requires inverting this matrix for each distinct triplet. Since we are interested mainly in the order of magnitude of $`S/N`$, we can set the lower bound as $$\left(\frac{S}{N}\right)^2\underset{l_1l_2l_3}{}\frac{(B_{l_1l_2l_3}^{E\mathrm{\Theta }\mathrm{\Theta }})^2}{6C_{l_1}^{EE}C_{l_2}^{\mathrm{\Theta }\mathrm{\Theta }}C_{l_3}^{\mathrm{\Theta }\mathrm{\Theta }}},$$ (164) which amounts to ignoring duplicate triplets and replacing the remaining triplet with the average $`S/N`$ of the set. This limit is plotted for the ISW effect in Fig. 4 as a function of the maximal $`l_1`$ included in the sum. As expected, it is comparable to the signal-to-noise in the temperature bispectrum. Of course, it is experimentally more difficult to achieve the cosmic variance limit in the polarization with a realistic experiment containing detector and foreground noise. There is also a qualitatively new effect from the polarization-lensing correlation $`C_l^{E\varphi }`$. However since secondary polarization only arises from Thomson scattering effects, we expect this contribution to be small in $`\mathrm{\Lambda }`$CDM models where the optical depth during reionization is $`\tau <0.3`$ . The $`E\mathrm{\Theta }\mathrm{\Theta }`$ bispectrum term is $`B_{l_1l_2l_3}^{EE\mathrm{\Theta }}`$ $`=`$ $`({}_{2}{}^{}F_{l_1l_2l_3}^{}C_{l_2}^{E\varphi }C_{l_3}^{\mathrm{\Theta }E}+{}_{2}{}^{}F_{l_1l_3l_2}^{}C_{l_2}^{EE}C_{l_3}^{\mathrm{\Theta }\varphi })`$ (166) $`+F_{l_3l_1l_2}C_{l_1}^{E\varphi }C_{l_2}^{\mathrm{\Theta }E}+(l_1l_2),`$ $`B_{(𝐥_1,𝐥_2,𝐥_3)}^{EE\mathrm{\Theta }}`$ $`=`$ $`(𝐥_2𝐥_3)(\mathrm{cos}2\phi _{31}C_{l_2}^{E\varphi }C_{l_3}^{\mathrm{\Theta }E}`$ (169) $`+\mathrm{cos}2\phi _{21}C_{l_2}^{EE}C_{l_3}^{\mathrm{\Theta }\varphi })`$ $`(𝐥_1𝐥_2)C_{l_1}^{E\varphi }C_{l_2}^{\mathrm{\Theta }E}+(𝐥_1𝐥_2),`$ with covariance $`\mathrm{Cov}`$ $`=`$ $`C_{l_1}^{EE}C_{l_2}^{EE}C_{l_3}^{\mathrm{\Theta }\mathrm{\Theta }}\delta _{l_1l_1^{}}\delta _{l_2l_2^{}}\delta _{l_3l_3^{}}`$ (172) $`+C_{l_1}^{EE}C_{l_2}^{E\mathrm{\Theta }}C_{l_3}^{\mathrm{\Theta }E}\delta _{l_1l_2^{}}\delta _{l_2l_3^{}}\delta _{l_3l_1^{}}`$ $`+C_{l_1}^{\mathrm{\Theta }E}C_{l_2}^{EE}C_{l_3}^{\mathrm{\Theta }E}\delta _{l_1l_3^{}}\delta _{l_2l_1^{}}\delta _{l_3l_2^{}}+(l_1^{}l_2^{}),`$ with which we can bound the $`S/N`$ $$\left(\frac{S}{N}\right)^2>\underset{l_1l_2l_3}{}\frac{(B_{l_1l_2l_3}^{EE\mathrm{\Theta }})^2}{6C_{l_1}^{EE}C_{l_2}^{EE}C_{l_3}^{\mathrm{\Theta }\mathrm{\Theta }}}.$$ (173) Again, the ISW example is shown in Fig. 4. Finally the $`EEE`$ bispectrum is given by $`B_{l_1l_2l_3}^{EEE}`$ $`=`$ $`{}_{2}{}^{}F_{l_1l_2l_3}^{}C_{l_2}^{E\varphi }C_{l_3}^{EE}+5\mathrm{p}\mathrm{e}\mathrm{r}\mathrm{m}.`$ (174) $`B_{(𝐥_1,𝐥_2,𝐥_3)}^{EEE}`$ $`=`$ $`(𝐥_2𝐥_3)\mathrm{cos}2\phi _{31}C_{l_2}^{E\varphi }C_{l_3}^{EE}+5\mathrm{p}\mathrm{e}\mathrm{r}\mathrm{m}.`$ (175) with covariance $`\mathrm{Cov}=C_{l_1}^{EE}C_{l_2}^{EE}C_{l_3}^{EE}\delta _{l_1l_1^{}}\delta _{l_2l_2^{}}\delta _{l_3l_3^{}}+5\mathrm{p}\mathrm{e}\mathrm{r}\mathrm{m}.,`$ (176) and signal-to-noise $$\left(\frac{S}{N}\right)^2=\underset{l_1l_2l_3}{}\frac{(B_{l_1l_2l_3}^{EEE})^2}{6C_{l_1}^{EE}C_{l_2}^{EE}C_{l_3}^{EE}}.$$ (177) This bispectrum signal vanishes for the ISW effect. Bispectra involving the $`B`$-parity polarization have distinct properties. For terms involving one B-parity polarization term, only $`l_1+l_2+l_3`$=odd contributes to the all-sky spectrum and we implicitly assume below that even terms vanish. For the $`B\mathrm{\Theta }\mathrm{\Theta }`$ bispectrum, $`B_{l_1l_2l_3}^{B\mathrm{\Theta }\mathrm{\Theta }}`$ $`=`$ $`i\left({}_{2}{}^{}F_{l_1l_2l_3}^{}\right)C_{l_2}^{\mathrm{\Theta }\varphi }C_{l_3}^{\mathrm{\Theta }E}+(l_2l_3),`$ (178) $`B_{(𝐥_1,𝐥_2,𝐥_3)}^{B\mathrm{\Theta }\mathrm{\Theta }}`$ $`=`$ $`(𝐥_2𝐥_3)\mathrm{sin}2\phi _{31}C_{l_2}^{\mathrm{\Theta }\varphi }C_{l_3}^{\mathrm{\Theta }E}(𝐥_2𝐥_3).`$ (179) Again the correspondence between the flat and all sky expressions in eqn. (26) is established by the approximation discussed in the Appendix $$\left(\begin{array}{ccc}l_1& l_2& l_3\\ 2& 0& 2\end{array}\right)\pm i\mathrm{sin}2\phi _{31}\left(\begin{array}{ccc}l_1& l_2& l_3\\ 0& 0& 0\end{array}\right),$$ (180) for $`L=`$odd. The sign ambiguity comes from the fact that a reflection of the triangle $`(𝐥_1,𝐥_2,𝐥_3)`$ across one of the axes corresponds to remappings $`\phi \pi \phi `$ or $`\phi \phi `$ and hence a reversal in sign of the flat-sky bispectrum in equation (179). In this case the cancellation for flattened triangles discussed in §V A does not apply. However since $`\mathrm{sin}2\phi _{31}2\phi _{31}1`$, a suppression still exists. The covariance of the $`B\mathrm{\Theta }\mathrm{\Theta }`$ bispectrum is $`\mathrm{Cov}`$ $`=`$ $`C_{l_1}^{BB}C_{l_2}^{\mathrm{\Theta }\mathrm{\Theta }}C_{l_3}^{\mathrm{\Theta }\mathrm{\Theta }}\delta _{l_1l_1^{}}\delta _{l_2l_2^{}}\delta _{l_3l_3^{}}+(l_2^{}l_3^{}),`$ (181) leading to a signal-to-noise $$\left(\frac{S}{N}\right)^2=\underset{l_1l_2l_3}{}\frac{(B_{l_1l_2l_3}^{B\mathrm{\Theta }\mathrm{\Theta }})^2}{2C_{l_1}^{BB}C_{l_2}^{\mathrm{\Theta }\mathrm{\Theta }}C_{l_3}^{\mathrm{\Theta }\mathrm{\Theta }}}.$$ (182) In a cosmic variance limited experiment (see Fig. 4), the $`B\mathrm{\Theta }\mathrm{\Theta }`$ bispectrum has signal-to-noise advantages over its temperature and $`E`$ polarization counterparts due to the fact that for scalar perturbations $`C_{l_1}^{BB}`$ is dominated by the lensing contributions themselves. Moreover, even if the tensor contributions are near their current limits of $`T/S0.3`$, the signal-to-noise is not much affected for $`l100`$ due to the strong damping of gravity wave contributions under the horizon scale at last scattering. However for the Planck experiment, the detection is severely limited by detector noise and may also suffer further degradation from incomplete foreground subtraction . Next, the $`BE\mathrm{\Theta }`$ bispectrum is given by $`B_{l_1l_2l_3}^{BE\mathrm{\Theta }}`$ $`=`$ $`i({}_{2}{}^{}F_{l_1l_2l_3}^{}C_{l_2}^{E\varphi }C_{l_3}^{\mathrm{\Theta }E}+{}_{2}{}^{}F_{l_1l_3l_2}^{}C_{l_2}^{EE}C_{l_3}^{\mathrm{\Theta }\varphi }),`$ (183) $`B_{(𝐥_1,𝐥_2,𝐥_3)}^{BE\mathrm{\Theta }}`$ $`=`$ $`(𝐥_2𝐥_3)(\mathrm{sin}2\phi _{31}C_{l_2}^{E\varphi }C_{l_3}^{\mathrm{\Theta }E}`$ (185) $`+\mathrm{sin}2\phi _{21}C_{l_2}^{EE}C_{l_3}^{\mathrm{\Theta }\varphi }),`$ with a covariance $`\mathrm{Cov}`$ $`=`$ $`C_{l_1}^{BB}C_{l_2}^{EE}C_{l_3}^{\mathrm{\Theta }\mathrm{\Theta }}\delta _{l_1l_1^{}}\delta _{l_2l_2^{}}\delta _{l_3l_3^{}}`$ (187) $`+C_{l_1}^{BB}C_{l_2}^{\mathrm{\Theta }E}C_{l_3}^{\mathrm{\Theta }E}\delta _{l_1l_1^{}}\delta _{l_2l_3^{}}\delta _{l_3l_2^{}},`$ leading to a signal-to-noise $$\left(\frac{S}{N}\right)^2\underset{l_1l_2l_3}{}\frac{(B_{l_1l_2l_3}^{B\mathrm{\Theta }\mathrm{\Theta }})^2}{2C_{l_1}^{BB}C_{l_2}^{EE}C_{l_3}^{\mathrm{\Theta }\mathrm{\Theta }}}.$$ (188) The signal-to-noise of this term can be greater than that of $`B\mathrm{\Theta }\mathrm{\Theta }`$ due to the fact that the temperature and $`E`$-polarization are only partially correlated in the unlensed sky. Finally, $`B_{l_1l_2l_3}^{BEE}`$ $`=`$ $`i\left({}_{2}{}^{}F_{l_1l_2l_3}^{}\right)C_{l_2}^{E\varphi }C_{l_3}^{EE}+(l_2l_3),`$ (189) $`B_{(𝐥_1,𝐥_2,𝐥_3)}^{BEE}`$ $`=`$ $`(𝐥_2𝐥_3)\mathrm{sin}2\phi _{31}C_{l_2}^{E\varphi }C_{l_3}^{EE}(𝐥_2𝐥_3),`$ (190) with a covariance $`\mathrm{Cov}`$ $`=`$ $`C_{l_1}^{BB}C_2^{EE}C_3^{EE}\delta _{l_1l_1^{}}\delta _{l_2l_2^{}}\delta _{l_3l_3^{}}+(l_2^{}l_3^{}),`$ (191) leading to a signal-to-noise $$\left(\frac{S}{N}\right)^2=\underset{l_1l_2l_3}{}\frac{(B_{l_1l_2l_3}^{BEE})^2}{2C_{l_1}^{BB}C_{l_2}^{EE}C_{l_3}^{EE}}.$$ (192) This signal vanishes for the ISW effect. Terms involving more than one $`B`$ term have no contributions to first order in the correlation power spectrum. ## VI Discussion We have shown that a harmonic approach to weak lensing in the CMB provides a simple and exact means of calculating its effects on the temperature and polarization power spectra, given the power spectrum of the lensing potential or convergence, and on the analogous bispectra given their power spectrum of the cross correlation with secondary anisotropies. Corrections to the flat-sky approximations appear even at high multipoles because even there, lensing effects arises from the large-scale fluctuations in the deflection angles. These corrections correspond to a change in the predictions at the $`\mu `$K level. While this is a negligible change given observations today, it is above the cosmic-variance limit and should be included when interpreting the high-precision results expected from Planck. Unlike the temperature bispectrum, bispectra involving both the temperature and polarization multipoles of the CMB have the potential of producing a high signal-to-noise $`(10)`$ detection of secondary anisotropies such as the ISW effects even with relatively modest angular resolutions $`l<1000`$. Other secondary anisotropies such as the Sunyaev-Zel’dovich effect are expected to contribute even stronger signals, although their exact amplitude is far more uncertain presently . Achieving a cosmic-variance limited detection of the magnetic-parity polarization is a daunting challenge. Even signal-to-noise near unity requires detectors which are a factor of 3 more sensitive to polarization than those planned for the Planck satellite. Also of concern are the residual foreground contamination remaining in the maps after multifrequency subtraction. Our current best models of the foregrounds indicate that with the Planck channels and sensitivities, foregrounds and detector noise may enter into the polarization maps with comparable amplitudes . Thus improving the actual sensitivity to the cosmic signal beyond the specifications of the Planck experiment will not only require better detectors but also a better understanding of the foregrounds, perhaps with increased frequency coverage and sampling. Nonetheless, the polarization of the CMB offers the potential to open a new window on physical processes at low redshifts and the opportunity to learn more from the CMB than can be achieved with the next generation of CMB satellites. Acknowledgements: I would like to thank M. Zaldarriaga for many useful discussions and help during the early stages of this work. This work was supported by the Keck Foundation, a Sloan Fellowship, and NSF-9513835. ## A All-Sky Weak Lensing Observables All weak lensing observables may be defined in terms of the projected potential $`\varphi `$ $$\varphi (\widehat{𝐧})=2𝑑Dg_\varphi (D)\mathrm{\Phi }(𝐱(\widehat{𝐧}),D),$$ (A1) or equivalently its multipole moments $`\varphi _{lm}`$ in the all-sky formalism or Fourier coefficients $`\varphi (𝐥)`$. Recall from eqn. (28) that $`g_\varphi `$ is the lensing efficiency function. The deflection angle that a photon suffers while traveling from the source at $`D_s`$ is given by the angular gradient of the potential $`𝜶`$$`(\widehat{𝐧})=\varphi (\widehat{𝐧})`$. Applying equation (103) to the the spherical harmonic expansion, we obtain $$𝜶=\underset{lm}{}\sqrt{\frac{l(l+1)}{2}}\varphi _{lm}\left[{}_{1}{}^{}Y_{l}^{m}𝐦_+{}_{1}{}^{}Y_{l}^{m}𝐦_{}\right].$$ (A2) This implies that the quantity $`\alpha _1\pm i\alpha _2`$ is a spin $`\pm 1`$ object $`[\alpha _1\pm i\alpha _2](\widehat{𝐧})`$ $``$ $`{\displaystyle \underset{lm}{}}(c\pm ig)_{lm}{}_{\pm 1}{}^{}Y_{l}^{m}(\widehat{𝐧})`$ (A3) $`=`$ $`\pm {\displaystyle \underset{lm}{}}\sqrt{l(l+1)}\varphi _{lm}{}_{\pm 1}{}^{}Y_{l}^{m}(\widehat{𝐧}),`$ (A4) which states that the curl term $`c_{lm}`$ vanishes and the gradient term $$g_{lm}=i\sqrt{l(l+1)}\varphi _{lm}.$$ (A5) The power spectrum of the angular deflection is then $`g_{l^{}m^{}}^{}g_{lm}`$ $``$ $`\delta _{l,l^{}}\delta _{m,m^{}}C_l^{gg}`$ (A6) $`=`$ $`\delta _{l,l^{}}\delta _{m,m^{}}l(l+1)C_l^{\varphi \varphi },`$ (A7) with the curl power vanishing. This accounts for the factors of $`l(l+1)`$ in equations involving the angular deflection \[e.g. eqn. (112)\]. The corresponding flat-sky quantity is given by the decomposition \[see eqn. (C14)\] $`[\alpha _1\pm i\alpha _2](\widehat{𝐧})`$ $``$ $`\pm i{\displaystyle \frac{d^2l}{(2\pi )^2}[c\pm ig](𝐥)e^{\pm i(\phi _l\phi )}e^{i𝐥\widehat{𝐧}}},`$ (A8) with $`c(𝐥)=0`$ and $`g(𝐥)`$ $`=`$ $`il\varphi (𝐥),`$ (A9) $`C_{(l)}^{gg}`$ $`=`$ $`l^2C_{(l)}^{\varphi \varphi }.`$ (A10) These relations also give the bispectrum of the deflection angle in terms of bispectrum of the lensing potential in the obvious manner. The convergence $`(\kappa )`$ and shear $`(\gamma _1,\gamma _2)`$ are familiar weak lensing observables from galaxy weak lensing studies . Although they are not directly needed for CMB studies, they are of interest for cross-correlation of galaxy weak-lensing maps and the CMB. An equivalent all-sky lensing treatment is given by . These quantities are given by the second derivatives $`_i_j\varphi `$ $``$ $`\kappa g_{ij}+(\gamma _1+i\gamma _2)(𝐦_+𝐦_+)_{ij}`$ (A12) $`+(\gamma _1i\gamma _2)(𝐦_{}𝐦_{})_{ij},`$ convergence where $`g_{ij}`$ is the metric on the sphere. For the all-sky harmonics, it is useful to note that equation (102) implies $`_i_jY_l^m`$ $`=`$ $`{\displaystyle \frac{l(l+1)}{2}}Y_l^mg_{ij}+{\displaystyle \frac{1}{2}}\sqrt{{\displaystyle \frac{(l+2)!}{(l2)!}}}`$ (A14) $`\times \left[{}_{2}{}^{}Y_{l}^{m}(𝐦_+𝐦_+)+{}_{2}{}^{}Y_{l}^{m}(𝐦_{}𝐦_{})\right]_{ij},`$ and hence $`\kappa (\widehat{𝐧})`$ $`=`$ $`{\displaystyle \underset{lm}{}}{\displaystyle \frac{1}{2}}l(l+1)\varphi _{lm}Y_l^m(\widehat{𝐧}),`$ (A15) $`\gamma _1(\widehat{𝐧})\pm i\gamma _2(\widehat{𝐧})`$ $`=`$ $`{\displaystyle \underset{lm}{}}{\displaystyle \frac{1}{2}}\sqrt{{\displaystyle \frac{(l+2)!}{(l2)!}}}\varphi _{lm}{}_{\pm 2}{}^{}Y_{l}^{m}(\widehat{𝐧}).`$ (A16) Consequently, the power spectra are related as $`C_l^{\kappa \kappa }`$ $`=`$ $`{\displaystyle \frac{l^2(l+1)^2}{4}}C_l^{\varphi \varphi },`$ (A17) $`C_l^{ϵϵ}`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle \frac{(l+2)!}{(l2)!}}C_l^{\varphi \varphi },`$ (A18) $`C_l^{X\kappa }`$ $`=`$ $`{\displaystyle \frac{1}{2}}l(l+1)C_l^{X\varphi },`$ (A19) $`C_l^{Xϵ}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\sqrt{{\displaystyle \frac{(l+2)!}{(l2)!}}}C_l^{X\varphi },`$ (A20) where the $`ϵ`$ shear power spectra is defined in the same way as that of the $`E`$ polarization and $`X=\mathrm{\Theta },E,B`$. The $`\beta `$ shear power is the analogue of the $`B`$ polarization power and vanishes for weak lensing. In the flat-sky limit, these expressions become $`\kappa (\widehat{𝐧})`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{d^2𝐥}{(2\pi )^2}l^2\varphi (𝐥)e^{i𝐥\widehat{𝐧}}}`$ (A21) $`\gamma _1(\widehat{𝐧})\pm i\gamma _2(\widehat{𝐧})`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{d^2𝐥}{(2\pi )^2}l^2\varphi (𝐥)e^{\pm 2i(\phi _l\phi )}e^{i𝐥\widehat{𝐧}}},`$ (A22) so that $`C_{(l)}^{\kappa \kappa }`$ $`=`$ $`C_{(l)}^{ϵϵ}={\displaystyle \frac{1}{4}}l^4C_{(l)}^{\varphi \varphi },`$ (A23) $`C_{(l)}^{X\kappa }`$ $`=`$ $`C_{(l)}^{Xϵ}={\displaystyle \frac{1}{2}}l^2C_{(l)}^{X\varphi }`$ (A24) These relations also give the bispectrum of the shear and convergence in terms of the bispectrum of the lensing potential $`B_{l_1l_2l_3}^{\kappa \kappa \kappa }`$ $`=`$ $`{\displaystyle \frac{1}{8}}[l_1(l_1+1)l_2(l_2+1)l_3(l_3+1)]B_{l_1l_2l_3}^{\varphi \varphi \varphi },`$ (A25) $`B_{l_1l_2l_3}^{ϵϵϵ}`$ $`=`$ $`{\displaystyle \frac{1}{8}}\sqrt{{\displaystyle \frac{(l_1+2)!}{(l_12)!}}{\displaystyle \frac{(l_2+2)!}{(l_22)!}}{\displaystyle \frac{(l_3+2)!}{(l_32)!}}}B_{l_1l_2l_3}^{\varphi \varphi \varphi },`$ (A26) with a similar relation for the flat-sky bispectra. ## B Wigner-3$`j`$ Evaluation ### 1 Exact Expressions The expressions for the power spectrum of the lensed temperature and polarization distributions involve specific sets of Wigner-3$`j`$ symbols that can be efficiently evaluated. The expression for the temperature involves, a set which has a closed algebraic form: $`\left(\begin{array}{ccc}l_1& l_2& l_3\\ 0& 0& 0\end{array}\right)`$ $`=`$ $`(1)^{L/2}{\displaystyle \frac{(\frac{L}{2})!}{(\frac{L}{2}l_1)!(\frac{L}{2}l_2)!(\frac{L}{2}l_3)!}}`$ (B3) $`\times `$ $`\left[{\displaystyle \frac{(L2l_1)!(L2l_2)!(L2l_3)!}{(L+1)!}}\right]^{1/2},`$ (B4) for even $`L=l_1+l_2+l_3`$ and zero for odd $`L`$. The required set for the polarization does not have an exact closed form expression. However it may be equally efficiently evaluated for our purposes with the realization that in the sums, we require $$\left(\begin{array}{ccc}l_1& l_2& l_3\\ m_1& m_2& m_3\end{array}\right)w_{l_1}$$ (B5) for fixed $`l_2,l_3,m_1,m_2,m_3`$ and all allowed $`l_1`$. The recursion relations for the Wigner-$`3j`$ symbol, $`l_1A_{l_1+1}w_{l_1+1}+B_{l_1}w_{l_1}+(l_1+1)A_{l_1}w_{l_11}=0,`$ (B6) where $`A_{l_1}`$ $`=`$ $`\sqrt{l_1^2(l_2l_3)^2}\sqrt{(l_2+l_3+1)^2l_1^2}\sqrt{l_1^2m_1^2},`$ (B7) $`B_{l_1}`$ $`=`$ $`(2l_1+1)[l_2(l_2+1)m_1l_3(l_3+1)m_1`$ (B9) $`l_1(l_1+1)(m_3m_2)],`$ allow us to generate the whole set at once . For a stable recursion, one begins at the minimum and maximum $`l_1`$ values $`l_{1\mathrm{m}\mathrm{i}\mathrm{n}}`$ $`=`$ $`\mathrm{max}(|l_2l_3|),|m_1|),`$ (B10) $`l_{1\mathrm{m}\mathrm{a}\mathrm{x}}`$ $`=`$ $`l_2+l_3,`$ (B11) with $`w_{l_{1\mathrm{m}\mathrm{i}\mathrm{n}}}=w_{l_{1\mathrm{m}\mathrm{a}\mathrm{x}}}=1`$ and carries the recursion in both directions to the midpoint $`l_{1\mathrm{m}\mathrm{i}\mathrm{d}}`$ in the range (or any non-vanishing entry in the vicinity). One then renormalizes either the left or right recursion to make the $`w_{l_{1\mathrm{m}\mathrm{i}\mathrm{d}}}`$ agree. The remaining overall normalization is fixed by requiring $$\underset{l_1}{}(2l_1+1)w_{l_1}^2=1,$$ (B12) and $$\mathrm{sgn}\left(w_{l_{1\mathrm{m}\mathrm{a}\mathrm{x}}}\right)=(1)^{l_2l_3m_1}.$$ (B13) Putting these relations together, we obtain the full set of symbols as required. ### 2 Approximations We can use the general relation between the all and flat sky bispectra of equation (26) compared with the explicit calculation of the flat sky bispectrum in §V B to develop an high-$`l`$ approximation for the specific symbol in the polarization calculations. The comparison implies that $$\left(\begin{array}{ccc}l_1& l_2& l_3\\ 2& 0& 2\end{array}\right)\mathrm{cos}2\phi _{31}\left(\begin{array}{ccc}l_1& l_2& l_3\\ 0& 0& 0\end{array}\right),$$ (B14) for $`L=l_1+l_2+l_3=`$ even. By the law of cosines, $`\mathrm{cos}2\phi _{31}={\displaystyle \frac{1}{2}}{\displaystyle \frac{(l_2^2l_1^2l_3^2)^2}{l_1^2l_3^2}}1.`$ (B15) Then $`\left(\begin{array}{ccc}l_1& l_2& l_3\\ 2& 0& 2\end{array}\right)`$ $``$ $`(1)^{L/2}\left[{\displaystyle \frac{1}{2}}{\displaystyle \frac{(l_2^2l_1^2l_3^2)^2}{l_1^2l_3^2}}1\right]`$ (B20) $`\times {\displaystyle \frac{(L/2)!}{(L/2l_1)!(L/2l_2)!(L/2l_3)!}}`$ $`\times \left[{\displaystyle \frac{(L2l_1)!(L2l_2)!(L2l_3)!}{(L+1)!}}\right]^{1/2},`$ for $`L=`$ even. For odd values of $`L`$, we use the relation $$\left(\begin{array}{ccc}l_1& l_2& l_3\\ 2& 0& 2\end{array}\right)\pm i\mathrm{sin}2\phi _{31}\left(\begin{array}{ccc}l_1& l_2& l_3\\ 0& 0& 0\end{array}\right),$$ (B21) and fix the overall sign ambiguity by an explicit evaluation. By the triangle relations, $`\mathrm{sin}2\phi _{31}`$ $`=`$ $`{\displaystyle \frac{1}{2}}[L(L2l_1)(L2l_2)(L2l_3)]^{1/2}`$ (B23) $`\times \left({\displaystyle \frac{l_2^2l_1^2l_3^2}{l_1^2l_3^2}}\right).`$ Putting this together with equation (B4) and fixing the sign ambiguity, we obtain $`\left(\begin{array}{ccc}l_1& l_2& l_3\\ 2& 0& 2\end{array}\right)`$ $``$ $`(1)^{\frac{L1}{2}}{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{l_2^2l_1^2l_3^2}{l_1^2l_3^2}}\right)`$ (B29) $`\times {\displaystyle \frac{(L/2)!}{(L/2l_1)!(L/2l_2)!(L/2l_3)!}}`$ $`\times [L(L2l_1)(L2l_2)(L2l_3)`$ $`\times {\displaystyle \frac{(L2l_1)!(L2l_2)!(L2l_3)!}{(L+1)!}}]^{1/2}`$ for $`L=\mathrm{odd}`$. The half integer factorials are defined by the gamma function $`x!=\mathrm{\Gamma }(1+x)`$. By explicit calculation we find that these expressions are valid to better than $`3\%`$ of the rms amplitude of the symbol when averaged over neighboring $`l`$ for all $`l_1|l_2l_3|25`$ and $`l_2+l_3l_125`$, i.e. for triangles that are sufficiently far from being flat. Near zero crossings, the fractional error can be large but the absolute error remains a small fraction of the rms. A typical case is shown in Fig. 5. These relations may be useful in cases where only a single symbol is needed. However for the lensing calculation where the whole set is required, the recursion relations are as efficient as the approximation and are exact. ## C Flat and All Sky Correspondence ### 1 Harmonics We establish here the correspondence between the all and flat sky harmonic coefficients of spin zero (scalar), spin one (vector) and spin two (tensor) quantities on the sky. Following , let us begin by introducing the following weighted sum over the multipole moments of the field $`X=\mathrm{\Theta }`$, $`E`$, $`B`$, or $`\varphi `$ for a given $`l`$ and its inverse relation, $`X(𝐥)`$ $`=`$ $`\sqrt{{\displaystyle \frac{4\pi }{2l+1}}}{\displaystyle \underset{m}{}}i^mX_{lm}e^{im\phi _l},`$ (C1) $`X_{lm}`$ $`=`$ $`\sqrt{{\displaystyle \frac{2l+1}{4\pi }}}i^m{\displaystyle \frac{d\phi _l}{2\pi }e^{im\phi _l}X(𝐥)}.`$ (C2) The goal is then to show that this quantity is the Fourier coefficient of the flat-sky expansion. Spin-0 quantities, such as the temperature flucutations and the lensing potential, are decomposed as $$X(\widehat{𝐧})=\underset{lm}{}X_{lm}Y_l^m(\widehat{𝐧}).$$ (C3) For small angles around the pole, the spherical harmonics may be approximated as<sup>*</sup><sup>*</sup>*Note that our definition of $`Y_l^m`$ differs from the usual one by $`(1)^m`$ to conform with the spin spherical harmonic convention . $$Y_l^mJ_m(l\theta )\sqrt{\frac{l}{2\pi }}e^{im\phi },$$ (C4) and the expansion of the plane wave $`e^{i𝐥\widehat{𝐧}}`$ $`=`$ $`{\displaystyle \underset{m}{}}i^mJ_m(l\theta )e^{im(\phi \phi _l)}`$ (C5) $``$ $`\sqrt{{\displaystyle \frac{2\pi }{l}}}{\displaystyle \underset{m}{}}i^mY_l^me^{im\phi _l},`$ (C6) Thus $`X(\widehat{𝐧})`$ $`=`$ $`{\displaystyle \underset{lm}{}}X_{lm}Y_l^m`$ (C7) $``$ $`{\displaystyle \underset{l}{}}{\displaystyle \frac{l}{2\pi }}{\displaystyle \frac{d\phi _l}{2\pi }X(𝐥)\underset{m}{}J_m(l\theta )i^me^{im(\phi \phi _l)}}`$ (C8) $``$ $`{\displaystyle \frac{d^2l}{(2\pi )^2}X(𝐥)e^{i𝐥\widehat{𝐧}}},`$ (C9) which is the desired correspondence. Spin-1 quantities like the deflection angles are decomposed as $${}_{\pm }{}^{}X(\widehat{𝐧})=\underset{lm}{}{}_{\pm }{}^{}X_{lm}^{}{}_{\pm 1}{}^{}Y_{l}^{m}.$$ (C10) Here one notes that $${}_{\pm 1}{}^{}Y_{l}^{m}\pm \frac{1}{l}e^{i\phi }\left(_x\pm i_y\right)Y_l^m,$$ (C11) and thus $`{}_{\pm }{}^{}X(\widehat{𝐧})`$ $`=`$ $`{\displaystyle \underset{lm}{}}{}_{\pm }{}^{}X_{lm}^{}{}_{\pm 1}{}^{}Y_{l}^{m}`$ (C12) $``$ $`\pm {\displaystyle \underset{l}{}}{\displaystyle \frac{l}{2\pi }}{\displaystyle \frac{d\phi _l}{2\pi }{}_{\pm }{}^{}X(𝐥)e^{i\phi }\frac{1}{l}\left(_x\pm i_y\right)e^{i𝐥\widehat{𝐧}}}`$ (C13) $``$ $`\pm i{\displaystyle \frac{d^2l}{(2\pi )^2}{}_{\pm }{}^{}X(𝐥)e^{\pm i(\phi _l\phi )}e^{i𝐥\widehat{𝐧}}}.`$ (C14) Finally, spin-2 quantities like the polarization are decomposed as $${}_{\pm }{}^{}X(\widehat{𝐧})=\underset{lm}{}{}_{\pm }{}^{}X_{lm}^{}{}_{\pm 2}{}^{}Y_{l}^{m}.$$ (C15) Here one notes that $${}_{\pm 2}{}^{}Y_{l}^{m}\frac{1}{l^2}e^{2i\phi }\left(_x\pm i_y\right)^2Y_l^m,$$ (C16) and thus $`{}_{\pm }{}^{}X(\widehat{𝐧})`$ $`=`$ $`{\displaystyle \underset{lm}{}}{}_{\pm }{}^{}X_{lm}^{}{}_{\pm 2}{}^{}Y_{l}^{m}`$ (C17) $``$ $`{\displaystyle \underset{l}{}}{\displaystyle \frac{l}{2\pi }}{\displaystyle \frac{d\phi _l}{2\pi }{}_{\pm }{}^{}X(𝐥)e^{2i\phi }\frac{1}{l^2}\left(_x\pm i_y\right)^2e^{i𝐥\widehat{𝐧}}}`$ (C18) $``$ $`{\displaystyle \frac{d^2l}{(2\pi )^2}{}_{\pm }{}^{}X(𝐥)e^{\pm 2i(\phi _l\phi )}e^{i𝐥\widehat{𝐧}}},`$ (C19) as desired. ### 2 Power Spectra The correspondence between power spectra then follows from the relationship between the harmonics $`X_{lm}^{}X_{l^{}m^{}}^{}`$ $``$ $`i^{m^{}m}{\displaystyle \frac{\sqrt{ll^{}}}{2\pi }}C_{(l)}^{XX^{}}{\displaystyle 𝑑\phi _le^{im\phi _l}𝑑\phi _l^{}e^{im^{}\phi _l^{}}}`$ (C21) $`\times \delta (𝐥𝐥^{}).`$ We then expand the delta function in plane waves functions $`\delta (𝐥𝐥^{})`$ $`=`$ $`{\displaystyle \frac{d\widehat{𝐧}}{(2\pi )^2}e^{i(𝐥𝐥^{})\widehat{𝐧}}}`$ (C22) $``$ $`{\displaystyle \frac{2\pi }{\sqrt{ll^{}}}}{\displaystyle \frac{d\widehat{𝐧}}{(2\pi )^2}\underset{mm^{}}{}i^{mm^{}}Y_l^{}^m^{}Y_l^me^{im\phi _lim^{}\phi _l^{}}}.`$ (C23) Integrating over the azimuthal angles $`\phi _l,\phi _l^{}`$ collapses the sum to $`X_{lm}^{}X_{l^{}m^{}}^{}`$ $``$ $`\delta _{l,l^{}}\delta _{m,m^{}}C_l^{XX^{}}`$ (C24) $``$ $`C_{(l)}^{XX^{}}{\displaystyle 𝑑\widehat{𝐧}Y_l^mY_l^{}^m^{}}`$ (C25) $`=`$ $`\delta _{l,l^{}}\delta _{m,m^{}}C_{(l)}^{XX^{}}`$ (C26) which proves the desired relation in eqn. (26), $`C_l^{XX^{}}`$ $``$ $`C_{(l)}^{XX^{}}.`$ (C27) ### 3 Bispectra The correspondence between bispectra is established in exactly the same way as with the power spectra. The only difference is that the expansion of $`\delta (𝐥^{}𝐥)`$ in eqn. (C23) is replaced with that of $`\delta (𝐥_1+𝐥_2+𝐥_3)`$ leading to $`X_{lm}X_{l^{}m^{}}^{}X_{l^{\prime \prime }m^{\prime \prime }}^{\prime \prime }`$ $``$ $`\left(\begin{array}{ccc}l& l^{}& l^{\prime \prime }\\ m& m^{}& m^{\prime \prime }\end{array}\right)B_{ll^{}l^{\prime \prime }}^{XX^{}X^{\prime \prime }}`$ (C30) $``$ $`B_{(l,l^{},l^{\prime \prime })}^{XX^{}X^{\prime \prime }}{\displaystyle 𝑑\widehat{𝐧}Y_l^mY_l^{}^m^{}Y_{l^{\prime \prime }}^{m^{\prime \prime }}}`$ (C31) $`=`$ $`B_{(l,l^{},l^{\prime \prime })}^{XX^{}X^{\prime \prime }}\left(\begin{array}{ccc}l& l^{}& l^{\prime \prime }\\ 0& 0& 0\end{array}\right)\left(\begin{array}{ccc}l& l^{}& l^{\prime \prime }\\ m& m^{}& m^{\prime \prime }\end{array}\right)`$ (C38) $`\times \sqrt{{\displaystyle \frac{(2l+1)(2l^{}+1)(2l^{\prime \prime }+1)}{4\pi }}}.`$ This establishes the relation $`B_{ll^{}l^{\prime \prime }}^{XX^{}X^{\prime \prime }}`$ $``$ $`\left(\begin{array}{ccc}l& l^{}& l^{\prime \prime }\\ 0& 0& 0\end{array}\right)\sqrt{{\displaystyle \frac{(2l+1)(2l^{}+1)(2l^{\prime \prime }+1)}{4\pi }}}`$ (C42) $`B_{(l,l^{},l^{\prime \prime })}^{XX^{}X^{\prime \prime }}.`$ Note that we have implicitly assumed that the bispectrum only depends on the the magnitudes ($`l,l^{},l^{\prime \prime }`$) so that it may be removed from the azimuthal integrals. This is not true for terms not involving the magnetic parity. In this case, the sign of the flat-sky bispectrum depends on orientation but we find empirically that a similar relationship holds up to a sign ambiguity as discussed in §II B. ### 4 Signal-to-Noise Here we establish the correspondence between the all and flat sky signal-to-noise statistics for the case of diagonal contributions to the covariance matrix \[$`\mathrm{Cov}`$ = diag($`\mathrm{Var}`$)\], $`\left({\displaystyle \frac{S}{N}}\right)^2`$ $`=`$ $`{\displaystyle \underset{lm}{}}{\displaystyle \frac{(C_l^{XX})^2}{\mathrm{Var}}}={\displaystyle \underset{l}{}}(2l+1){\displaystyle \frac{(C_l^{XX})^2}{\mathrm{Var}}}.`$ (C43) For the flat sky case, one defines a weighted sum of Fourier harmonics $`P={\displaystyle d^2lW(l)X(𝐥)X(𝐥)},`$ (C44) with optimal weights given by $`W(l)=C_{(l)}^{XX}/\mathrm{Var}`$ from which one calculates the signal-to-noise $`P^2/P^2`$ as $`\left({\displaystyle \frac{S}{N}}\right)^2`$ $`=`$ $`{\displaystyle \frac{f_{\mathrm{sky}}}{\pi }}{\displaystyle d^2l\frac{[C_{(l)}^{XX}]^2}{\mathrm{Var}}}`$ (C45) $``$ $`2f_{\mathrm{sky}}{\displaystyle l𝑑l\frac{(C_l^{XX})^2}{\mathrm{Var}}},`$ (C46) where we have used the fact that $`\delta (\mathrm{𝟎})V/(2\pi ^2)=f_{\mathrm{sky}}/\pi `$. These expressions agree in the high $`l`$-limit and imply the familiar result that $`f_{\mathrm{sky}}`$ should multiply the signal-to-noise of angular power spectrum measurements given incomplete sky coverage. The bispectrum signal-to-noise similarly is $`\left({\displaystyle \frac{S}{N}}\right)^2`$ $`=`$ $`{\displaystyle \underset{l_1l_2l_3}{}}{\displaystyle \frac{(B_{l_1l_2l_3}^{XXX})^2}{\mathrm{Var}}}{\displaystyle \underset{m_1m_2m_3}{}}\left(\begin{array}{ccc}l_1& l_2& l_3\\ m_1& m_2& m_3\end{array}\right)^2`$ (C49) $`=`$ $`{\displaystyle \underset{l_1l_2l_3}{}}{\displaystyle \frac{(B_{l_1l_2l_3}^{XXX})^2}{\mathrm{Var}}},`$ (C50) for the all sky bispectrum and $`\left({\displaystyle \frac{S}{N}}\right)^2`$ $`=`$ $`{\displaystyle \frac{f_{\mathrm{sky}}}{\pi }}{\displaystyle \frac{1}{(2\pi )^2}}{\displaystyle d^2l_1d^2l_2\frac{[B_{(l_1,l_2,l_3)}^{XXX}]^2}{\mathrm{Var}}},`$ (C51) for the flat sky bispectrum . The extra factor of $`(2\pi )^2`$ compared with the power spectrum is from the extra delta function in the noise term. One can show that these expressions agree in the high-$`l`$ limit by restoring the integration over $`l_3`$, expanding the delta function into spherical harmonics as in eqn. (C23), and integrating over azimuthal angles, $`{\displaystyle d^2l_1d^2l_2d^2l_3\delta (𝐥_1+𝐥_2+𝐥_3)}`$ (C52) $`{\displaystyle l_1𝑑l_1l_2𝑑l_2l_3𝑑l_3\sqrt{\frac{(2\pi )^5}{l_1l_2l_3}}𝑑\widehat{𝐧}Y_{l_1}^0Y_{l_2}^0Y_{l_3}^0}`$ (C53) $`8\pi ^2{\displaystyle l_1𝑑l_1l_2𝑑l_2l_3𝑑l_3\left(\begin{array}{ccc}l_1& l_2& l_3\\ 0& 0& 0\end{array}\right)^2}.`$ (C56) With the general correspondence of bispectra from eqn. (C42), this becomes $$\left(\frac{S}{N}\right)^2f_{\mathrm{sky}}𝑑l_1𝑑l_2𝑑l_3\frac{(B_{l_1l_2l_3}^{XXX})^2}{\mathrm{Var}},$$ (C57) which proves the equivalence of the signal-to-noise for high-$`l`$ and $`f_{\mathrm{sky}}=1`$.
warning/0001/hep-ph0001215.html
ar5iv
text
# 1 Introduction ## 1 Introduction In a recent paper a new determination of the weak charge of atomic cesium has been reported. The most precise atomic parity violating (APV) experiment compares the mixing among $`S`$ and $`P`$ states due to neutral weak interactions to an induced Stark mixing . The 1.2% uncertainty on the previous measurement of the weak charge $`Q_W`$ was dominated by the theoretical calculations on the amount of Stark mixing and on the electronic parity violating matrix elements. In the Stark mixing was measured and, incorporating new experimental data, the uncertainty in the electronic parity violating matrix elements was reduced. The new result $$Q_W(_{55}^{133}Cs)=72.06\pm (0.28)_{\mathrm{expt}}\pm (0.34)_{\mathrm{theor}}$$ (1) represents a considerable improvement with respect to the previous determination $$Q_W(_{55}^{133}Cs)=71.04\pm (1.58)_{\mathrm{expt}}\pm (0.88)_{\mathrm{theor}}$$ (2) On the theoretical side, $`Q_W`$ can be expressed in terms of the $`S`$ parameter or the $`ϵ_3`$ $$Q_W=72.72\pm 0.13102ϵ_3^{\mathrm{rad}}+\delta _NQ_W$$ (3) including hadronic-loop uncertainty. We use here the variables $`ϵ_i`$ (i=1,2,3) of ref. , which include the radiative corrections, in place of the set of variables $`S`$, $`T`$ and $`U`$ originally introduced in ref. . In the above definition of $`Q_W`$ we have explicitly included only the Standard Model (SM) contribution to the radiative corrections. New physics (that is physics beyond the SM) is represented by the term $`\delta _NQ_W`$ including also contributions to $`ϵ_3`$. Also, we have neglected a correction proportional to $`ϵ_1^{\mathrm{rad}}`$. In fact, as well known , due to the particular values of the number of neutrons ($`N=78`$) and of protons ($`Z=55`$) in cesium, the dependence on $`ϵ_1`$ almost cancels out. From the theoretical expression we see that $`Q_W`$ is particularly sensitive to new physics contributing to the $`ϵ_3`$ parameter. This kind of new physics is severely constrained by the high energy experiments. From a recent fit , the value of $`ϵ_3`$ from the high energy data is $`ϵ_3^{\mathrm{expt}}=(4.19\pm 1.0)\times 10^3`$. To estimate new physics contributions to this parameter one has to subtract the SM radiative corrections, which, for $`m_{top}=175GeV`$ and for $`m_H(GeV)=100,300`$, are given respectively by $$\begin{array}{ccc}m_H=100GeV& & ϵ_3^{\mathrm{rad}}=5.110\times 10^3\\ m_H=300GeV& & ϵ_3^{\mathrm{rad}}=6.115\times 10^3\end{array}$$ (4) Therefore new physics contributing to $`ϵ_3`$ cannot be larger than a few per mill. Since $`ϵ_3`$ appears in $`Q_W`$ multiplied by a factor 102, the kind of new physics which contributes through $`ϵ_3`$ cannot contribute to $`Q_W`$ for more than a few tenth. On the other side the discrepancy between the SM and the experimental data is given by (for a light Higgs) $$Q_W^{\mathrm{expt}}Q_W^{SM}=1.18\pm 0.46$$ (5) where we have added in quadrature the uncertainties. This corresponds to 2.6(2.8)-$`\sigma `$ deviation with respect to the SM for $`m_H=100(300)GeV`$. The 95%CL limits on $`\delta _NQ_W`$ are $$0.28\delta _NQ_W2.08\mathrm{for}m_H=100GeV$$ (6) $$0.38\delta _NQ_W2.18\mathrm{for}m_H=300GeV$$ (7) For increasing $`m_H`$ both bounds increase. One possible contribution to $`Q_W`$ which was neglected is the difference between neutron and proton spatial distributions in the nucleus. With the increasing APV measurement precision, this effect has been reconsidered adding an additional $`\mathrm{\Delta }Q_W^{np}=\pm 0.3`$ to the theoretical error of eq. (1). Including this additional uncertainty, the theoretical error in eq. (1) becomes $`\pm (0.45)`$ and the deviation with respect to SM for $`m_H=100GeV`$ decreases from 2.6-$`\sigma `$ to 2.2-$`\sigma `$. Our analysis will be based on the result given in eq. (1). The lower positive bounds from eqs. (6) and (7) exclude the SM at 99%CL and, a fortiori, all the models leading to negative extra contribution to $`Q_W`$, as for example models with a sequential $`Z^{}`$ . This 2.6(2.8)-$`\sigma `$ deviation with respect to the SM for $`m_H=100(300)GeV`$ could be explained by assuming the existence of an extra $`Z^{}`$ from $`E_6`$ or $`O(10)`$ or from $`Z_{LR}^{}`$ of left-right (LR) models . ## 2 Bounds on extra $`Z^{}`$ from $`Q_W`$ Let us now look at models which, at least in principle, could give rise to a sizeable modification of $`Q_W`$. In ref. it was pointed out that models involving extra neutral vector bosons coupled to ordinary fermions can do the job. The high energy data at the $`Z`$ resonance strongly bound the $`ZZ^{}`$ mixing . For this reason we will assume zero mixing in the following calculations. In this case $`\delta _NQ_W`$ is due to the direct exchange of the $`Z^{}`$ and is completely fixed by the $`Z^{}`$ parameters: $$\delta _NQ_W=16a_e^{}[(2Z+N)v_u^{}+(Z+2N)v_d^{}]\frac{M_Z^2}{M_Z^{}^2}$$ (8) $`a_f^{},v_f^{}`$ are the couplings $`Z^{}`$ to fermions. We will discuss the following classes of models involving an extra $`Z^{}`$: the LR models and the extra-U(1) models. The relevant couplings of the $`Z^{}`$ to the electron and to the up and down quarks are given in the Table 1 of . The different extra-U(1) models are parameterized by the angle $`\theta _6`$. In we used a different definition of the angle. The relation between the angle $`\theta _2`$ of ref. and $`\theta _6`$ is given by $`\theta _6=\theta _2\mathrm{arctan}\sqrt{5/3}`$. In the case of the LR model considered in the extra contribution to the weak charge is $$\delta _NQ_W=\frac{M_Z^2}{M_Z^{}^2}Q_W^{SM}$$ (9) For this model one has a 95%CL lower bound on $`M_{Z_{LR}^{}}`$ from Tevatron given by $`M_{Z_{LR}^{}}630GeV`$. A LR model could then explain the APV data allowing for a mass of the $`Z_{LR}^{}`$ varying between the intersection from the 95%CL bounds $`540M_{Z_{LR}^{}}(GeV)1470`$ deriving from eq. (6) and the lower bound of $`630GeV`$ . In the case of the extra-U(1) models the CDF experimental lower bounds for the masses vary according to the values of the parameter $`\theta _6`$ which parameterizes different extra-U(1) models, but in general they are about $`600GeV`$ at 95%CL . In particular for the model known in the literature as $`\eta `$ (or $`A`$), which corresponds to $`\theta _6=\mathrm{arctan}\sqrt{5/3}`$, the 95%CL lower bound is $`M_{Z_\eta ^{}}620GeV`$, for the model $`\psi `$ (or $`C`$), which corresponds to $`\theta _6=\pi /2`$, the lower bound is $`M_{Z_\psi ^{}}590GeV`$ and for the model $`\chi `$, which corresponds to $`\theta _6=0`$, the lower bound is $`M_{Z_\chi ^{}}595GeV`$. By comparing eqs. (6), (7) with eq. (8) we see that the models $`\eta `$ and $`\psi `$ are excluded. The bounds on $`\delta _NQ_W`$ at 95%CL can be translated into lower and upper bounds on $`M_Z^{}`$. The result is given in Fig. 1, where the bounds are plotted versus $`\theta _6`$. In looking at this figure one should also remember that the direct lower bound from Tevatron is about $`600GeV`$ at 95%CL. We see that the presence of an extra $`Z^{}`$ can explain the discrepancy with the SM prediction for the $`Q_W`$ for a wide range of $`\theta _6`$ angle. In particular the $`\chi `$ (or $`B`$) model, corresponding to $`\theta _6=0`$, is allowed for $`M_{Z_\chi ^{}}`$ less than about 1.2 $`TeV`$. ## 3 $`Z^{}`$ at future colliders The search for a $`Z^{}`$ is one of the tasks of future colliders. The existing bounds for $`E_6`$ models, $`M_Z^{}600GeV`$ from direct search at Tevatron will be upgraded by the future run with $`\sqrt{s}=2TeV`$, $`L=1fb^1`$ to $`M_Z^{}800900GeV`$ and pushed to $`1TeV`$ for $`L=10fb^1`$. The bounds are based on 10 events in the $`e^+e^{}+\mu ^+\mu ^{}`$ channels and decays to SM final states only is assumed . At the LHC with an integrated luminosity of $`100fb^1`$ one can explore a mass range up to $`44.5TeV`$ depending on the $`\theta _6`$ value. Concerning LR models, the 95%CL lower limits from Tevatron run with $`\sqrt{s}=2TeV`$, $`L=1(10)fb^1`$ are $`900(1000)GeV`$ and extend to $`4.5TeV`$ at LHC . Therefore if the deviation for $`Q_W`$ with respect to the SM prediction is not due to a statistical fluctuation but to the presence of new physics like new extra gauge bosons from $`E_6`$ or LR models, LHC can verify or disprove this possible evidence. However little can be learned on the $`Z^{}`$ properties. With $`e^+e^{}`$ colliders the properties of a $`Z^{}`$ can be easily investigated if the center-of-mass energy is large enough to produce it. Anyway, from the measurements of $`e^+e^{}\gamma ,Z,Z^{}\overline{f}f`$ below threshold one can get information about the nature of the $`Z^{}`$. In ref. exclusion limits for $`M_Z^{}`$ have been determined from the analysis of all leptonic and hadronic observables. Here we show the upgrading of this analysis for a collider scenario $`\sqrt{s}=500GeV`$, $`L=500fb^1`$ and polarized beams: $`P_e^{}=0.9`$, $`P_{e^+}=0.6`$ (solid line). The 95%CL lower bounds on $`M_Z^{}`$ are given in Fig. 2 for different values of $`\theta _6`$ for $`E_6`$ models. Also shown is the case of unpolarized positron beam (dash line) and the case of $`\sqrt{s}=300GeV`$, $`L=300fb^1`$ (solid-dot line). The observables which are considered are the total lepton and hadron cross sections, the forward backward asymmetry, the left-right asymmetry (for leptons, hadrons, $`c\overline{c}`$ and $`b\overline{b}`$ final states), $`R_b`$ and $`R_c`$. The assumed identification efficiencies are: for leptons $`ϵ_l=95\%`$, for $`c\overline{c}`$ $`ϵ_c=40\%`$, for $`b\overline{b}`$ $`ϵ_b=60\%`$ and the corresponding systematic errors $`\mathrm{\Delta }ϵ_l/ϵ_l=0.5\%`$ $`\mathrm{\Delta }ϵ_c/ϵ_c=1.5\%`$, and $`\mathrm{\Delta }ϵ_b/ϵ_b=1\%`$. An uncertainty of 0.5% on the luminosity and $`\mathrm{\Delta }P/P=1\%`$ are considered. The same analysis for the LR models leads to a 95%CL lower bound for $`M_{Z_{LR}^{}}`$ shown in Fig. 3. For example the particular LR model considered in corresponds to $`\alpha _{LR}=\sqrt{\mathrm{cot}^2\theta _W1}`$ and for this model the bound extracted from Fig. 3 is around $`7TeV`$. Assuming that the $`Z^{}`$ mass is known, one can study the $`Z^{}`$ couplings to fermions. The 95%CL contours on $`Z^{}\overline{l}l`$ couplings for the $`\chi `$ model for $`M_Z^{}=1,1.5,3TeV`$ are presented in Fig. 4 for $`\sqrt{s}=300GeV`$ and $`L=300fb^1`$ and in Fig. 5 for $`\sqrt{s}=500GeV`$ and $`L=500fb^1`$. In Fig. 6 the 95%CL contours for $`a_b^{}`$ $`v_b^{}`$ for $`M_Z^{}=1,1.5TeV`$, and $`\sqrt{s}=500GeV`$ and $`L=500fb^1`$ are shown. The dash lines correspond to $`P_{e^+}=0`$. In conclusion, once a $`Z_\chi ^{}`$ is seen at LHC, already a LC with $`\sqrt{s}=300GeV`$, $`L=300fb^1`$ and polarized beams is able to measure the $`Z_\chi ^{}`$ couplings to fermions with good precision. For example for a $`Z_\chi ^{}`$ with mass of $`1TeV`$ (which can explain the new APV result) the couplings to leptons can be determined within 10%, unless a sign ambiguity. By increasing the center of mass energy, the precision improves.
warning/0001/hep-th0001092.html
ar5iv
text
# Gravity in the brane-world for two-branes model with stabilized modulus ## I introduction Recently, there has been a growing interest in considering extra-dimensions in non-trivial form. One important suggestion was done by Randall and Sundrum. They realized that if one considers 5-dimensional gravity with a negative cosmological constant bounded by two positive and negative tension branes, the induced gravity on the negative tension brane can possess a very small gravitational constant as compared with the energy scale introduced in the original Lagrangian. This fact potentially gives a solution to the hierarchy problem of explaining the extraordinary weakness of the gravitational coupling. Later, the same authors pointed out that the model with the positive tension brane alone is also possible. In this case, the extension of the extra dimension is infinite. Nevertheless, the induced gravity on the brane can be expected to mimic Einstein gravity. There are many discussions about the cosmology based on these scenarios . The behavior of gravity on these models has been also investigated by many authors. In this direction, an explicit method to deal with the linear order metric perturbations induced by the matter fields confined on the branes was developed in the paper by Garriga and Tanaka(Paper I). In that paper, considering the zero mode truncation approximation, it was shown that the induced gravity on the branes becomes of the Brans-Dicke type in the two-branes model. On the other hand, Einstein gravity is recovered in the single-brane model. However, the analysis of the two-branes model in Paper I was not complete. The stabilization mechanism of the distance between the two branes by means of a bulk scalar field was not taken into account. Moreover, the Kaluza-Klein contribution was not estimated for the two-branes model. In this paper we consider a model with stabilization mechanism, and we confirm that the 4-dimensional Einstein gravity is recovered on both branes under the approximation of zero mode truncation. An approximate formula for the mass of the lowest massive mode in the scalar-type perturbations is also obtained. Furthermore, we present a method to evaluate the metric perturbations taking into account all the KK contributions under a contact interaction approximation. This approximation is valid if the source which excites the KK modes is smoothly distributed compared with the scale corresponding to the mass of the lowest massive mode. This paper is organized as follows. In Sec. II we describe the background model that we consider, and derive the basic equations for the linear perturbations on it. In Sec. III we discuss the mass spectrum of the perturbations. We identify the mode functions for the massless degrees of freedom, and also derive the formula for the mass and the mode function of the lowest massive mode. In Sec. IV the mechanism to recover Einstein gravity is explained. This result is expected from the notion of the mass spectrum but its derivation is not so trivial when we consider the explicit construction of metric perturbations. We shall find that the contribution from the massive modes of the scalar-type perturbations must be also taken into account in some sense. For this purpose, we introduce a contact interaction approximation. In Sec. V we apply the same technique to the KK modes of the tensor-type perturbations to complete the description of the linear perturbations of this system. Section VI is devoted to summary. ## II model and basic equations We consider linear perturbations of the model proposed by Randall and Sundrum with the stabilization mechanism by Goldberger and Wise. The fields existing on the 5-dimensional spacetime (bulk) are the usual 5-dimensional gravity with a negative cosmological constant $`\mathrm{\Lambda }`$, and a 5-dimensional scalar field $`\phi `$ introduced to stabilize the distance between the two branes. We denote the 5-dimensional gravitational constant by $`G_5`$. The unperturbed metric is supposed to take the form, $$ds^2=a^2(y)\eta _{\mu \nu }dx^\mu dx^\nu +dy^2,$$ (1) where $`\eta _{\mu \nu }`$ is the 4-dimensional Minkowski metric with $`(+++)`$ signature. We use $`\gamma _{\mu \nu }=a^2(y)\eta _{\mu \nu }`$ to raise 4-dimensional tensor indices. The $`y`$-direction is bounded by two branes located at $`y=y^{(\pm )}`$. On these two branes, $`𝐙_2`$-symmetry is imposed. The Lagrangian for the bulk scalar field is $$=\frac{1}{2}g^{ab}\phi _{,a}\phi _{,b}V_B(\phi )\underset{\sigma =\pm }{}V^{(\sigma )}(\phi )\delta (yy^{(\sigma )}).$$ (2) For most of the present analysis, we do not need to specify the explicit form of the potentials $`V_B(\phi )`$ and $`V^{(\pm )}(\phi )`$. In the bulk, the background scale factor and the scalar field, $`(a,\phi _0)`$, must satisfy $`\dot{H}(y)`$ $`=`$ $`{\displaystyle \frac{\kappa }{3}}\dot{\phi }_0^2(y),`$ (3) $`H^2(y)`$ $`=`$ $`{\displaystyle \frac{\kappa }{6}}\left({\displaystyle \frac{1}{2}}\dot{\phi }_0^2(y)V_B(\phi _0(y))\kappa ^1\mathrm{\Lambda }\right),`$ (4) $`\ddot{\phi }_0(y)`$ $`+`$ $`4H(y)\dot{\phi }_0(y)V_B^{}(\phi _0(y))=0,`$ (5) where $`H(y):=\dot{a}(y)/a(y)\sqrt{\mathrm{\Lambda }/6}`$ and $`\kappa =8\pi G_5`$. Ordinary matter fields reside on both branes, and the values of the 4-dimensional vacuum energy on both branes are adjusted to realize a static background configuration. This way of model construction is different from the standard approach which assumes the model potential and the values of the vacuum energy on the branes from the very beginning. In the present approach, we obtain a constraint on the model potential by fixing the distance between the two branes. Conversely, in the standard approach, the stabilized distance is determined for each given model (although some tuning of one of the model parameters is necessary to realize a static configuration). We call the brane at $`y=y^{(+)}(y^{()})`$ the positive (negative) tension brane, and adopt the convention $`y^{(+)}<y^{()}`$. We first consider the metric perturbation $`\delta (ds^2)=h_{ab}dx^adx^b`$ and the scalar field perturbation $`\delta \phi `$ in the bulk. We are using the convention that Latin indices run through $`0,1,2,3,5`$, while Greek indices run through $`0,1,2,3`$. In the bulk, we can always impose the “Newton gauge” condition, $`h_{55}`$ $`=`$ $`2\varphi ^N,h_{5\mu }=0,`$ (6) $`h_{\mu \nu }`$ $`=`$ $`h_{\mu \nu }^{(TT)}\varphi ^N\gamma _{\mu \nu },`$ (7) $`\delta \phi `$ $`=`$ $`{\displaystyle \frac{3}{2\kappa \dot{\phi }_0}}\left[_y+2H\right]\varphi ^N,`$ (8) where $`h_{\mu \nu }^{(TT)}`$ satisfies the transverse-traceless condition, and $`\varphi ^N`$ is the 5-dimensional “Newtonian potential”. The equation for $`h_{\mu \nu }^{(TT)}`$ in the bulk is given by $$\left[\frac{1}{a^2}\mathrm{}^{(4)}+\widehat{L}^{(TT)}\right]h_{\mu \nu }^{(TT)}=0,$$ (9) where $`\mathrm{}^{(4)}`$ is the 4-dimensional d’Alembertian operator with respect to $`\eta _{\mu \nu }`$, and we have introduced the operator $$\widehat{L}^{(TT)}:=\frac{1}{a^2}_ya^4_y\frac{1}{a^2}.$$ (10) For the scalar-type perturbations, we obtain $$\left[\mathrm{}^{(4)}+\widehat{L}^{(\varphi ^N)}\right]\varphi ^N=0,$$ (11) where $`\widehat{L}^{(\varphi ^N)}`$ is the operator defined by $$\widehat{L}^{(\varphi ^N)}:=a^2\dot{\phi }_0^2_y\frac{1}{a^2\dot{\phi }_0^2}_ya^2\frac{2\kappa }{3}a^2\dot{\phi }_0^2.$$ (12) As discussed in Paper I, the $`y`$ = constant hypersurface does not correspond to the location of the branes in this gauge. Following it, we introduce two Gaussian normal coordinate systems near both branes, respectively. We denote quantities in these coordinate systems by associating a bar, such as $`\overline{y}`$. Then, the junction condition for the bulk scalar field is given by $`\pm 2\dot{\overline{\phi }}=V^{}{}_{}{}^{(\pm )}(\overline{\phi })`$ at $`\overline{y}=\overline{y}^{(\pm )}\pm `$. Hence, its perturbation become $$\pm 2\delta \dot{\overline{\phi }}=V^{\prime \prime }{}_{}{}^{(\pm )}(\phi _0)\delta \overline{\phi },(\overline{y}=\overline{y}^{(\pm )}\pm ).$$ (13) The junction condition for the metric perturbation is $`\pm (_{\overline{y}}2H)\overline{h}_{\mu \nu }`$ $`=`$ $`\kappa \left[T_{\mu \nu }{\displaystyle \frac{1}{3}}\gamma _{\mu \nu }T\right]^{(\pm )}`$ (15) $`{\displaystyle \frac{2\kappa }{3}}\gamma _{\mu \nu }\dot{\phi }_0\delta \overline{\phi },(\overline{y}=\overline{y}^{(\pm )}\pm ).`$ Here, we have introduced the energy momentum tensor of the matter fields confined on the branes, $`T_{\mu \nu }^{(\pm )}`$, and $`T^{(\pm )}:=\gamma ^{\mu \nu }T_{\mu \nu }^{(\pm )}`$. $`T_{\mu \nu }^{(\pm )}`$ satisfies the 4-dimensional conservation law $`T_{\mu \nu }{}_{}{}^{;\nu }=0`$. Notice that there appears a contribution from the perturbation of the scalar field $`\phi `$ which was not present in the models discussed in Paper I. Whereas the junction conditions have been easily derived using Gaussian normal coordinates, the equations of motion for the perturbations are simpler in the “Newton gauge”. So we need to consider the gauge transformation which relates normal coordinates and the “Newton gauge”. This is given by $`h_{ab}=\overline{h}_{ab}+\xi _{a;b}+\xi _{b;a}`$ with $`\xi _{(\pm )}^5`$ $`=`$ $`{\displaystyle ^y}\varphi ^N(y^{})𝑑y^{}={\displaystyle _{y^{(\pm )}}^y}\varphi ^N(y^{})𝑑y^{}+\widehat{\xi }_{(\pm )}^5,`$ (16) $`\xi _{(\pm )}^\nu `$ $`=`$ $`{\displaystyle ^y}\gamma ^{\mu \nu }(y^{})𝑑y^{}{\displaystyle ^y^{}}\varphi _{,\mu }^N(y^{\prime \prime })𝑑y^{\prime \prime }`$ (17) $`=`$ $`{\displaystyle _{y^{(\pm )}}^y}\gamma ^{\mu \nu }(y^{})𝑑y^{}{\displaystyle ^y^{}}\varphi _{,\mu }^N(y^{\prime \prime })𝑑y^{\prime \prime }+\widehat{\xi }_{(\pm )}^\nu ,`$ (18) where $`\widehat{\xi }_{(\pm )}^5`$ and $`\widehat{\xi }_{(\pm )}^\nu `$ are independent of $`y`$. Then, we have $`\delta \overline{\phi }(y)`$ $`=`$ $`\delta \phi (y)\dot{\phi }_0(y)\left[{\displaystyle _{y^{(\pm )}}^y}\varphi ^N(y^{})𝑑y^{}+\widehat{\xi }_{(\pm )}^5\right],`$ (19) $`\overline{h}_{\mu \nu }(y)`$ $`=`$ $`h_{\mu \nu }(y)+2a^2(y){\displaystyle ^y}{\displaystyle \frac{dy^{}}{a^2(y^{})}}{\displaystyle ^y^{}}\varphi _{,\mu \nu }^N(y^{\prime \prime })𝑑y^{\prime \prime }`$ (21) $`2H\gamma _{\mu \nu }(y){\displaystyle ^y}\varphi ^N(y^{})𝑑y^{}.`$ Here we should stress that the arguments in the l.h.s. are not $`\overline{y}`$ but $`y`$. Combining Eqs.(15) and (21), the junction condition in the “Newton gauge” for the $`TT`$ part is obtained as $`\pm (_y2H)h_{\mu \nu }^{(TT)}=\kappa \mathrm{\Sigma }_{\mu \nu }^{(\pm )},(y=y^{(\pm )}\pm ),`$ (22) where we have defined $`\mathrm{\Sigma }_{\mu \nu }^{(\pm )}`$ $`:=`$ $`\left(T_{\mu \nu }{\displaystyle \frac{1}{4}}\gamma _{\mu \nu }T\right)^{(\pm )}`$ (24) $`\pm {\displaystyle \frac{2}{\kappa }}(\widehat{\xi }_{(\pm ),\mu \nu }^5{\displaystyle \frac{1}{4}}\gamma _{\mu \nu }\widehat{\xi }_{(\pm )}^5{}_{,\rho }{}^{}{}_{}{}^{,\rho }).`$ Combining Eq. (22) with Eq. (9), we obtain the equation for $`h_{\mu \nu }^{(TT)}`$, $`\left[{\displaystyle \frac{\mathrm{}^{(4)}}{a^2}}+\widehat{L}^{(TT)}\right]h_{\mu \nu }^{(TT)}=2\kappa {\displaystyle \underset{\sigma =\pm }{}}\mathrm{\Sigma }_{\mu \nu }^{(\sigma )}\delta (yy^{(\sigma )}).`$ (25) The resulting $`h_{\mu \nu }^{(TT)}`$ is automatically consistent with the traceless condition, but the transverse condition gives us the equation which determines $`\widehat{\xi }_{(\pm )}^5`$, $$\frac{1}{a_{(\pm )}^2}\mathrm{}^{(4)}\widehat{\xi }_{(\pm )}^5=\pm \frac{\kappa }{6}T^{(\pm )},$$ (26) where we have defined $`a_{(\pm )}:=a(y^{(\pm )})`$. The set of Eqs.(25) and (26) is exactly the same that was obtained in Paper I once we specialize the background bulk geometry to pure anti de Sitter. Using Eq. (26), it can be easily seen that the trace part of the metric junction condition is trivially satisfied. The remaining junction condition is the one for the scalar field, Eq. (13). After some computations, it reduces in the “Newton gauge” to $$\frac{2\kappa }{3}(\delta \phi \dot{\phi }_0\widehat{\xi }_{(\pm )}^5)=\frac{ϵ^{(\pm )}}{a^2\dot{\phi }_0}\mathrm{}^{(4)}\varphi ^N,(y=y^{(\pm )}\pm ),$$ (27) where we have defined $$ϵ^{(\pm )}:=\frac{2}{V^{\prime \prime }{}_{}{}^{(\pm )}2(\ddot{\phi }_0/\dot{\phi }_0)}.$$ (28) Combining Eq. (27) with Eq. (11), we obtain the equation for $`\varphi ^N`$, $`\widehat{L}^{(\varphi ^N)}\varphi ^N{\displaystyle \underset{\sigma =\pm }{}}\sigma {\displaystyle \frac{4\kappa a^2\dot{\phi }_0^2}{3}}\widehat{\xi }_{(\sigma )}^5\delta (yy^{(\sigma )})`$ (29) $`=\mathrm{}^{(4)}\varphi ^N\left(1+{\displaystyle \underset{\sigma =\pm }{}}2ϵ^{(\sigma )}\delta (yy^{(\sigma )})\right).`$ (30) ## III mass spectrum ### A Zero modes Let us first consider the solution of the source free equations by setting $`T_{\mu \nu }^{(\pm )}=0`$. We will first consider the zero eigenmodes of $`\mathrm{}^{(4)}`$, which are the most important because they correspond to the 4-dimensional massless field responsible for the propagation of a long range force. If we set $`\mathrm{}^{(4)}=0`$ in Eq. (25), the solution for $`h_{\mu \nu }^{(TT)}`$ is $$h_{\mu \nu }^{(TT)}=\widehat{h}_{\mu \nu }^{(1)}(x^\rho )a^2(y)+\widehat{h}_{\mu \nu }^{(2)}(x^\rho )a^2(y)^y\frac{dy^{}}{a^4(y^{})},$$ (31) where $`\widehat{h}_{\mu \nu }^{(1)}`$ and $`\widehat{h}_{\mu \nu }^{(2)}`$ are 4-dimensional $`TT`$ tensors independent of $`y`$ satisfying $`\mathrm{}^{(4)}\widehat{h}_{\mu \nu }^{(i)}=0`$ . Furthermore, the junction condition (22) gives $$\widehat{h}_{\mu \nu }^{(2)}=2a_{(\pm )}^2\widehat{\xi }_{(\pm ),\mu \nu }^5.$$ (32) Hence, there is no extra constraint on $`\widehat{h}_{\mu \nu }^{(1)}`$, but $`\widehat{h}_{\mu \nu }^{(2)}`$ must be written as the second derivative of a scalar function. The former mode corresponds to 4-dimensional gravitational wave perturbations. Although there seem to exist 5 independent degrees of freedom in this mode, three of them are pure gauge. On the other hand, the latter mode should be classified as scalar-type perturbations. If we forget about the stabilization mechanism, this mode corresponds to the mode called radion in Ref.. Notice that Eq. (32) gives a relation between $`\widehat{\xi }_{(+)}^5`$ and $`\widehat{\xi }_{()}^5`$, $$a_{(+)}^2\widehat{\xi }_{(+)}^5=a_{()}^2\widehat{\xi }_{()}^5.$$ (33) In the present case with stabilization, $`\widehat{\xi }_{(\pm )}^5`$ also appears in the junction condition for the scalar-type perturbations. Hence, the scalar field perturbation must also be chosen to be compatible with this condition. Here we should note that this relation holds only when we restrict our considerations to zero modes. In the general case, which will be considered later, there is no reason for such a relation to be satisfied. Let us now consider the solution of the zero eigenvalue scalar-type perturbation. One solution is $`\varphi ^N={\displaystyle \frac{H}{a^2}}f^{(1)}(x^\rho ),`$ (34) $`\delta \phi ={\displaystyle \frac{\dot{\phi }_0}{2a^2}}f^{(1)}(x^\rho ).`$ (35) This mode satisfies the scalar field junction condition (27) with (33) if $`f^{(1)}=2a_{(\pm )}^2\widehat{\xi }_{(\pm )}^5`$. However, this solution can be transformed to nothing by a gauge transformation with parameters $`\xi ^5=a^2f^{(1)}/2`$, $`\xi ^\mu =f_{,\nu }^{(1)}^ya^2(y^{})\gamma ^{\mu \nu }(y^{})𝑑y^{}`$. The other solution is $`\varphi ^N=\left(1{\displaystyle \frac{2H}{a^2}}{\displaystyle ^y}a^2(y^{})𝑑y^{}\right)f^{(2)}(x^\rho ),`$ (36) $`\delta \phi =f^{(2)}(x^\rho ){\displaystyle \frac{\dot{\phi }_0}{a^2}}{\displaystyle ^y}a^2(y^{})𝑑y^{}.`$ (37) However, this mode is not compatible with condition (33). Hence, if we include the stabilization mechanism, no physical massless mode is present in the scalar-type perturbation spectrum, in contrast with the the case discussed in Paper I. ### B Massive modes Let us now consider the lowest massive mode. They are also important because they give the leading order correction to the zero mode truncation approximation. Furthermore, if tachyonic modes are present, such a model must be rejected. To estimate the lowest mass eigenvalue in the general case, one needs numerical calculations. In order to obtain analytical approximations, here we assume that the effect of the bulk scalar field back reaction to the background geometry is small. Namely, we assume $$\frac{|\dot{H}|}{H^2}=\frac{\kappa \dot{\phi }_0^2}{3H^2}1.$$ (38) For the metric, we can use the pure anti-de Sitter form $$a(y)=e^{y/\mathrm{}}.$$ (39) For simplicity, here we set $$y^{(+)}=0,y^{()}=d.$$ (40) First we consider the tensor-type perturbations. The expression for the mode functions $`u_i(y)`$ for the operator $`\widehat{L}^{(TT)}`$ satisfying the junction condition $`(_y2H)u_i(y)=0`$ on the positive tension brane is found in Refs.. Denoting the eigenvalue corresponding to $`u_i(y)`$ by $`m_i^2`$, the mode function is given in terms of Bessel functions, $`u_i(y)\{J_1(m_i\mathrm{})Y_2(m_i\mathrm{}e^{y/\mathrm{}})Y_1(m_i\mathrm{})J_2(m_i\mathrm{}e^{y/\mathrm{}})\}`$. The boundary condition on the negative tension brane reduces to $$\left\{J_1\left(m_i\mathrm{}\right)Y_1\left(m_i\mathrm{}e^{d/\mathrm{}}\right)Y_1\left(m_i\mathrm{}\right)J_1\left(m_i\mathrm{}e^{d/\mathrm{}}\right)\right\}=0,$$ (41) and determines the discrete eigenvalues $`m_i`$. For small $`m_i`$, this condition becomes $`J_1(m_i\mathrm{}e^{d/\mathrm{}})0`$, and hence the eigenvalues are given by $`m_ie^{d/\mathrm{}}j_i\mathrm{}^1`$, where $`j_i`$ is the $`i`$-th zero-point of $`J_1`$. Hence the physical mass of the lowest massive KK mode on the positive tension brane is given by $`3.8e^{d/\mathrm{}}\mathrm{}^1`$ while that on the negative tension brane by $`3.8\mathrm{}^1`$. Next, let us consider the scalar-type perturbations. When we discuss the scalar-type perturbations, we should not assume an scale factor of the form (39) from the beginning, because the non-vanishing second or higher derivative of $`a(y)`$ can be important. Hence we will derive formula (53) for the lowest mass eigenvalue without assuming (39). Only to obtain an estimate of this expression, we will use the scale factor (39). To discuss scalar-type perturbations, it is convenient to introduce a new variable defined by $$q:=\frac{3a^2}{2\kappa A}\varphi ^N,$$ (42) where a prime denotes derivative with respect to the conformal coordinate $`z`$ defined by $`dz=a(y)^1dy`$, and we have introduced $`A=\pm a^{1/2}\phi _0^{}`$. The signature in the definition of $`A`$ is chosen so that $`A`$ is continuous on the branes. The junction condition for $`q`$ is obtained from (27), $$2[_z+(A^{}/A)]q=\pm 2A\widehat{\xi }_{(\pm )}^5(y=y^{(\pm )}\pm ).$$ (43) The perturbation equation in terms of $`q`$ becomes $`\left[\mathrm{}^{(4)}+_z^2+A\left({\displaystyle \frac{1}{A}}\right)^{\prime \prime }{\displaystyle \frac{2\kappa }{3}}\phi _{0}^{}{}_{}{}^{2}\right]q`$ (44) $`={\displaystyle }\pm 2A\widehat{\xi }_{(\pm )}^5\delta (yy^{(\pm )}).`$ (45) Integrating the above equation once at the vicinity of the branes, we correctly reproduce the junction condition (43). To obtain the solution of Eq. (45), we will construct the Green function for the differential operator appearing in it. It is given by $$G_q=\frac{d^4k}{(2\pi )^4}e^{ik_\mu \mathrm{\Delta }x^\mu }\underset{i}{}\frac{q_i(z)q_i(z^{})}{m_i^2+k^2},$$ (46) where $`m_i`$ is the mass eigenvalue. The mode functions satisfy $$\left[_z^2A\left(\frac{1}{A}\right)^{\prime \prime }\frac{2\kappa }{3}\phi _{0}^{}{}_{}{}^{2}\right]q_i=m_i^2q_i,$$ (47) with the boundary condition $`\left[_z+(A^{}/A)\right]q_i=0`$, at $`y=y^{(\pm )}\pm `$. In the weak back reaction case, the term $`\delta U:=2\kappa \phi _{0}^{}{}_{}{}^{2}/3`$ in Eq. (47) is assumed to be small. If we perturbatively expand the mode function and the mass eigenvalue with respect to powers of $`\delta U`$ like $`q_i=q_i^{(0)}+q_i^{(1)}+\mathrm{}`$ and $`m_i=m_i^{(0)}+m_i^{(1)}+\mathrm{}`$, the lowest eigenmode for the unperturbed system is trivially given by $$q_0^{(0)}=\frac{𝒩}{A},$$ (48) with $`(m_0^{(0)})^2=0`$, where $`𝒩`$ is a normalization constant. It is straightforward to obtain the next order correction. We find $`q_0^{(1)}`$ $`=`$ $`{\displaystyle \frac{𝒩}{A}}{\displaystyle ^z}A^2dz^{}{\displaystyle ^z^{}}{\displaystyle \frac{dz^{\prime \prime }}{A^2}}\left({\displaystyle \frac{2\kappa }{3}}\phi _0^{}{}_{}{}^{2}(m_0^{(1)})^2\right).`$ (49) In terms of $`\delta \phi `$, the above first order correction is given by $$\delta \phi _0^{(1)}=𝒩\dot{\phi }_0(y)_{y^{(+)}}^y𝑑y^{}\left(\frac{2\kappa }{3a^2}\frac{(m_0^{(1)})^2}{a^4(y^{})\dot{\phi }_0^2(y^{})}\right).$$ (50) The junction condition on the positive tension brane is already imposed due to the choice of the integration constant. Here, for simplicity, we have taken the $`ϵ^{(\pm )}0`$ limit. This expression is seen to be regular even when $`\dot{\phi }_0`$ vanishes at some points. Denoting the point of vanishing $`\dot{\phi }_0`$ by $`y_0`$, we can show that $`a^4\dot{\phi }_0^2`$ can be expanded around this point as $`a^4\dot{\phi }_0^2=\left[a^4(V_B^{})^2\right]_{y=y_0}[`$ $`(yy_0)^2`$ (52) $`+\alpha ^2(yy_0)^4+\mathrm{}],`$ where $`\alpha ^2:=\frac{1}{3}(4H^2+V_B^{\prime \prime })|_{y=y_0}`$. From this, it is easy to see that the expression for $`\delta \phi _0^{(1)}`$ is regular at $`y=y_0`$. The junction condition (43) on the negative tension brane gives the formula for the lowest mass eigenvalue, $$m_0^2\frac{2\kappa }{3}\left[_{y^{(+)}}^{y^{()}}\frac{dy}{a^2}\right]/\left[_{y^{(+)}}^{y^{()}}\frac{dy}{a^4\dot{\phi }_0^2}\right].$$ (53) This expression is positive definite if there is no point at which $`\dot{\phi }_0`$ vanishes. However, when $`\dot{\phi }_0`$ vanishes, the integral in the denominator changes its signature at that point. Hence, $`m_0^2`$ can be negative. As seen from Eq. (52), the integrand does not have residue at this point. Hence the integral is independent of the way of modification of the integration path. It is expected that the dominant contribution to this integral comes from the region near the point $`y_0`$, where $`\dot{\phi }`$ vanishes. By substituting equation (52), we can approximately evaluate the denominator in Eq. (53) as $`{\displaystyle 𝑑y\frac{\left[a^4(V_B^{})^2\right]_{y=y_0}^1}{(yy_0)^2+\alpha ^2(yy_0)^4}}=\left[{\displaystyle \frac{\pi \alpha }{a^4(V_B^{})^2}}\right]_{y=y_0},`$ which becomes negative. Although we cannot give a proof here, this seems to be the case in general. Here we simply conjecture it. If tachyonic modes appear, they mean the breakdown of the model. Hence, we do not consider the case in which $`\dot{\phi }_0`$ vanishes at some point, and hereafter we assume that $`\dot{\phi }_0`$ has a definite sign. Without explicitly specifying any model, we can make a crude estimate of the lowest mass eigenvalue. To evaluate the numerator in the r.h.s of Eq. (53), we can use the form (39) for the scale factor $`a`$. To evaluate the denominator, we use the following fact. Besides some exceptional cases, $`a^4\dot{\phi }_0^2`$ will have a minimum (cases with no minima will be discussed later). There $`\ddot{\phi }_0+2H\dot{\phi }_0=V_B^{}2H\dot{\phi }_0=0`$, and accordingly the integrand has a rather sharp peak. We denote this point by $`y_c`$. At this point, the second derivative of $`a^4\dot{\phi }_0^2`$ is evaluated as $`(8H^2+2V_B^{\prime \prime }4\dot{H})a^4\phi _0^2`$. Hence, we can approximate the integral like $`(1/a^4\dot{\phi }_0^2)𝑑y[4H^2/a^4(V_B^{})^2]_{y=y_c}\mathrm{exp}[(4H^2+V_B^{\prime \prime })_{y=y_c}(yy_c)^2]𝑑y`$. Under these approximations, the formula for the mass of the lowest eigenmode is obtained as $`m_0^2{\displaystyle \frac{\kappa e^{2d/\mathrm{}}\mathrm{}^2}{6\sqrt{\pi }}}\left[e^{4y/\mathrm{}}(V_B^{})^2\sqrt{1+{\displaystyle \frac{V_B^{\prime \prime }\mathrm{}^2}{4}}}\right]_{y=y_c}.`$ (54) To proceed further, we consider the specific model for the bulk potential $`V_B(\phi )`$ discussed in Ref., $$V_B(\phi )=\frac{M^2\phi ^2}{2}.$$ (55) For this model, the background solution for the bulk scalar field in the weak back reaction case is already given by $$\phi _0=B_1e^{\nu _1y}+B_2e^{\nu _2y},$$ (56) where $`\nu _1=2\mathrm{}^1+\sqrt{4\mathrm{}^2+M^2}`$, $`\nu _2=2\mathrm{}^1\sqrt{4\mathrm{}^2+M^2}`$ and $`B_1e^{\nu _1d}\left(\phi _{()}\phi _{(+)}e^{\nu _2d}\right),`$ (57) $`B_2\phi _{(+)}\phi _{()}e^{\nu _1d}.`$ (58) $`\phi _{(+)}`$ and $`\phi _{()}`$ are the values of $`\phi _0`$ on the positive and the negative tension branes, respectively. Let us briefly elaborate on the stabilization distance. For simplicity, we will assume that the ratio $`\phi _{(+)}/\phi _{()}`$ is not extremely large, and also that all the input scales are similar, i.e. $`\kappa \mathrm{}^3`$. The condition for weak back reaction is $$\dot{\phi }_0^2\frac{1}{\kappa \mathrm{}^2}.$$ (59) Since $`\dot{\phi }_0^2`$ does not have its maximum in the bulk, it is sufficient to consider the conditions for weak back reaction on the boundaries. Then, we can evaluate the values of $`\dot{\phi }_0`$ on boundaries, $`\dot{\phi }_0(y^{(+)})\nu _2\phi _{(+)},`$ (60) $`\dot{\phi }_0(y^{()})\nu _1\phi _{()}4\mathrm{}^1\sqrt{1+(M^2\mathrm{}^2/4)}e^{\nu _2d}\phi _{(+)}.`$ (61) The conditions for weak back reaction on both branes become $`\left({\displaystyle \frac{\nu _2\phi _{(+)}}{4\mathrm{}^{5/2}}}\right)^2{\displaystyle \frac{\mathrm{}^3}{\kappa }},`$ (62) $`\left({\displaystyle \frac{\nu _1\phi _{()}}{4\mathrm{}^{5/2}}}\right)^2\left(1{\displaystyle \frac{4e^{\nu _2d}}{\nu _1\mathrm{}}}{\displaystyle \frac{\phi _{(+)}}{\phi _{()}}}\sqrt{1+{\displaystyle \frac{M^2\mathrm{}^2}{4}}}\right){\displaystyle \frac{\mathrm{}^3}{\kappa }}.`$ (63) The condition on the positive tension brane is satisfied by choosing (case A) $`\phi _{(\pm )}\mathrm{}^{3/2}`$ with $`M^2\mathrm{}^2`$ or (case B) $`M^2\mathrm{}^2`$ with $`\phi _{(\pm )}\mathrm{}^{3/2}`$. In case A, the condition on the negative tension brane is automatically satisfied. In this case, the stabilization distance is sensitive to the change of the values of the vacuum energy on the branes. Hence, without specifying the complete model, we cannot estimate the stabilization distance. If we consider the case in which $`M^2\mathrm{}^2`$ is not small, we find that it is natural to assume that the signature of $`\phi _{(+)}`$ is different from that of $`\phi _{()}`$. If $`\phi _{(+)}`$ and $`\phi _{()}`$ have the same signature, the ratio between them must be chosen to be extremely large to find a node-less solution of $`\dot{\phi }_0`$ with sufficiently large brane separation $`d`$. In the case that $`\phi _{(+)}`$ and $`\phi _{()}`$ have different signature, the solution of $`\dot{\phi }_0`$ becomes node-less for any choice of parameters. Case B is the case discussed in Ref.. In this case, to satisfy the condition for weak back reaction on the negative tension brane, $$\frac{d}{\mathrm{}}\frac{1}{\nu _2\mathrm{}}\mathrm{ln}\left(\frac{4\mathrm{}^1\phi _{(+)}\sqrt{1+\frac{M^2\mathrm{}^2}{4}}}{\nu _1\phi _{()}}\right)$$ (64) is required. Hence, $`\phi _{(+)}`$ and $`\phi _{()}`$ must have the same signature. To realize a sufficiently large value of $`d/\mathrm{}`$, it is necessary that $`|\phi _{(+)}|`$ is slightly larger than $`|\phi _{()}|`$. If we take the small $`M^2`$ limit, the above expression for the stabilization distance coincides with that obtained in Ref. . Let us now return to the computation of $`m_0^2`$. First we consider the case in which $`a^4\dot{\phi }_0^2`$ has a minimum and the estimate (54) is valid. As we shall see below, case B is exceptional in this sense. From the condition that $`V_B^{}2H\dot{\phi }_0=0`$ at $`y=y_c`$, we have $`e^{(\nu _2\nu _1)y_c}=\nu _1B_1/\nu _2B_2`$. Substituting this estimate of $`y_c`$ into (54), we obtain $$m_0^2\frac{8\kappa M^2e^{2d/\mathrm{}}\sqrt{1+(M^2\mathrm{}^2/4)}}{3\sqrt{\pi }}(B_1B_2).$$ (65) To solve the hierarchy problem on the negative tension brane, we need to set $`e^{d/\mathrm{}}10^{19}`$(GeV)$`/10^3`$(GeV)$`=10^{16}`$. Hence, $`d/\mathrm{}37`$. Therefore, it is rather natural to suppose that $`M^2\mathrm{}d`$ is larger than unity. If $`|\phi _{(+)}|`$ is not much larger than $`|\phi _{()}|`$, as assumed, we can approximate $`B_1B_2|\phi _{(+)}\phi _{()}|e^{\nu _1d}`$. In this case, the mass eigenvalue $`m_0^2`$ becomes $`O(\kappa |\phi _{(+)}\phi _{()}|M^2e^{2\mathrm{}^1d\sqrt{1+(M^2\mathrm{}^2/4)}})`$. Hence the physical squared mass on the negative tension brane becomes $`O(\kappa |\phi _{(+)}\phi _{()}|M^2e^{2\mathrm{}^1d(\sqrt{1+(M^2\mathrm{}^2/4)}1)})`$ while that on the positive tension brane is the same as $`m_0^2`$. As for $`|\phi _{(\pm )}|`$, they must be smaller than $`\mathrm{}^{3/2}`$ for the approximation of weak back reaction to be valid. Also, there is a factor $`M^2e^{2\mathrm{}^1d(\sqrt{1+(M^2\mathrm{}^2/4)}1)}`$, which takes its maximum value $`(0.2/\mathrm{})^2`$ at $`M0.33\mathrm{}^1`$. Hence, the mass scale on the negative tension brane tends to be smaller than the typical background energy scale $`\mathrm{}^1`$. Finally, we consider the special case in which $`a^4\dot{\phi }_0^2`$ in Eq. (53) has no minima. This happens when one of the terms in $`\dot{\phi }_0=\nu _1B_1e^{\nu _1y}+\nu _2B_2e^{\nu _2y}`$ can be totally neglected. This condition becomes $$|\phi _{()}\phi _{(+)}e^{\nu _2d}|\left|\frac{\nu _2}{\nu _1}\right||\phi _{()}|,$$ (66) or $$|\phi _{(+)}\phi _{()}e^{\nu _1d}|\left|\frac{\nu _1}{\nu _2}\right||\phi _{(+)}|.$$ (67) In the former case we can set $`B_10`$, and in the latter case we can set $`B_20`$. Then, for the former case, formula (53) gives $$m_0^2=\frac{4\kappa }{3}\nu _2^2\phi _{()}^2e^{2d/\mathrm{}}\left[\frac{\sqrt{1+(M^2\mathrm{}^2/4)}}{1e^{4\sqrt{1+(M^2\mathrm{}^2/4)}d/\mathrm{}}}\right].$$ (68) It is easy to see that case B corresponds to this case. The model discussed in Sec. 4.1 of Ref. with negative small value of $`b`$ also corresponds to this case (for the definition of $`b`$, see Ref.). The factor in the square brackets is almost unity. Substituting the approximation $`\nu _2M^2\mathrm{}/4`$, we recover the result obtained in . Note that their $`M^3`$ and $`k`$ are our $`1/(4\kappa )`$ and $`\mathrm{}^1`$, respectively. For the latter case, we have $$m_0^2=\frac{4\kappa }{3}\nu _1^2\phi _{()}^2e^{2d/\mathrm{}}\left[\frac{\sqrt{1+(M^2\mathrm{}^2/4)}}{e^{4\sqrt{1+(M^2\mathrm{}^2/4)}d/\mathrm{}}1}\right].$$ (69) The model discussed in Sec. 4.1 of Ref. with positive small value of $`b`$ corresponds to this case. To keep the mass sufficiently large, we have to choose $`|\phi _{()}|`$ extremely large, which is not compatible with the weak back reaction condition. ## IV Recovery of Einstein gravity As we have shown in the preceding section, in two-brane models with stabilization mechanism, a physical massless mode is absent in the scalar-type perturbation spectrum, which is different from what happens in the case without stabilization mechanism discussed in Paper I. This fact indicates that the resulting 4-dimensional effective gravity can resemble Einstein gravity at linear order. Let us show how it can be recovered. To compute the induced metric on the branes, it is convenient to transform back to Gaussian coordinates. Thus, applying “minus” the gauge transformation (18), the induced metric on each brane is given by $`\overline{h}_{\mu \nu }^{(\pm )}`$ $`=`$ $`h_{\mu \nu }^{(0)}(y^{(\pm )})+h_{\mu \nu }^{(KK)}(y^{(\pm )})`$ (71) $`\gamma _{\mu \nu }\left(\varphi ^N(y^{(\pm )})+2H\widehat{\xi }_{(\pm )}^5\right)(\widehat{\xi }_{\mu ,\nu }^{(\pm )}+\widehat{\xi }_{\nu ,\mu }^{(\pm )}).`$ Here we have decomposed $`h_{\mu \nu }^{(TT)}`$ into two parts, the zero mode contribution $`h_{\mu \nu }^{(0)}`$ and the KK contribution $`h_{\mu \nu }^{(KK)}`$. The last term represents the residual 4-dimensional gauge transformation. As we have already noted, Eq. (25) and Eq. (26) determining $`h_{\mu \nu }^{(TT)}(y^{(\pm )})`$ and $`\widehat{\xi }_{(\pm )}^5`$ are essentially the same as the ones obtained in Paper I for the case without stabilization mechanism. There, it was found that the induced gravity approximated by the zero mode truncation becomes of the Brans-Dicke type. Here we only quote the result for $`h_{\mu \nu }^{(0)}(y^{(\pm )})`$ up to 4-dimensional gauge transformation, $`h_{\mu \nu }^{(0)}(y^{(\pm )})`$ $`=`$ $`\left[{\displaystyle \frac{\mathrm{}^{(4)}}{a^2}}\right]^1{\displaystyle \underset{\sigma =\pm }{}}16\pi G^{(\sigma )}\left[T_{\mu \nu }\gamma _{\mu \nu }{\displaystyle \frac{T}{3}}\right]^{(\sigma )},`$ (72) where $`8\pi G^{(\pm )}:=\kappa Na_{(\pm )}^2`$ and $$N:=\left[2_{y^{(+)}}^{y^{()}}a^2𝑑y\right]^1.$$ (74) Hence, the only possible source of an additional contribution is that from $`h_{\mu \nu }^{(KK)}`$ or $`\varphi ^N(y^{(\pm )})`$. Although we have just shown that there are no massless degrees of freedom in the scalar-type perturbations, we will find that the contribution from $`\varphi ^N(y^{(\pm )})`$ gives the correct long range force required to recover Einstein gravity. First we consider the weak back reaction case discussed in Sec. III-B. From the normalization condition of $`q_0^{(0)}`$ we obtain $`𝒩^2=\left[2_0^d𝑑z\left(1/A\right)^2\right]^1`$. Comparing it with the formula for the mass (53), we find a simple relation $$m_0^2\frac{4\kappa 𝒩^2}{3}_0^d\frac{dy}{a^2}\frac{2\kappa \mathrm{}𝒩^2}{3}(e^{2d/\mathrm{}}1).$$ (75) Then, taking into account the contribution from the modes with the lowest mass eigenvalue alone, the Green function for $`q`$, Eq. (46), is approximated by $`G_q`$ $``$ $`{\displaystyle \frac{d^4k}{(2\pi )^4}\frac{3e^{ik_\mu \mathrm{\Delta }x^\mu }}{2\kappa \mathrm{}(e^{2d/\mathrm{}}1)}\frac{m_0^2A^1(y)A^1(y^{})}{m_0^2+k^2}}`$ (76) $`=`$ $`{\displaystyle \frac{3A^1(y)A^1(y^{})}{2\kappa \mathrm{}(e^{2d/\mathrm{}}1)}}`$ (78) $`\times \left[\delta ^4(\mathrm{\Delta }x^\mu ){\displaystyle \frac{d^4k}{(2\pi )^4}\frac{k^2e^{ik_\mu \mathrm{\Delta }x^\mu }}{m_0^2+k^2}}\right].`$ Using this Green function, and with the aid of (42) and (45), we find that $`\varphi ^N`$ is given by $`\varphi ^N`$ $``$ $`{\displaystyle \frac{2\mathrm{}^1}{a^2(e^{2d/\mathrm{}}1)}}[(\widehat{\xi }_{(+)}^5+\widehat{\xi }_{()}^5)`$ (80) $`{\displaystyle \frac{\kappa }{6}}[m_0^2\mathrm{}^{(4)}]^1(a_{(+)}^2T^{(+)}+a_{()}^2T^{()})].`$ Only the first term inside the square brackets gives a long range contribution to the induced metric. The propagation of this force is essentially due to $`\widehat{\xi }_{(\pm )}^5`$. The source for the second term exactly becomes the trace of the energy momentum tensor of the ordinary matter field. Hence, the force due to this term becomes short-ranged with typical length scale $`am_0^1`$. For the first term, using Eq. (26) for $`\widehat{\xi }_{(\pm )}^5`$, we finally obtain the contribution to the long range component of $`\overline{h}_{\mu \nu }^{(\pm )}`$ coming from $`\varphi ^N`$ as $`\gamma _{\mu \nu }`$ $`\varphi ^N`$ $`(y^{(\pm )}){\displaystyle \frac{8\pi G^{()}}{3}}`$ (82) $`\times \left[{\displaystyle \frac{\mathrm{}^{(4)}}{a_{(\pm )}^2}}\right]^1\left(a_{(+)}^2T^{(+)}+a_{()}^2T^{()}\right)\eta _{\mu \nu },`$ where we have used the fact that $`8\pi G^{(\pm )}=\kappa \mathrm{}^1a_{(\pm )}^2/(1e^{2d/\mathrm{}})`$ in the weak back reaction case. Substituting this into Eq. (71), and using Eq. (LABEL:quote) and Eq. (26), the zero mode truncation reproduces the formula for the linearized Einstein gravity, $$\frac{\mathrm{}^{(4)}}{a_{(\pm )}^2}\overline{h}_{\mu \nu }^{(\pm )}=\underset{\sigma =\pm }{}16\pi G^{(\sigma )}\left[T_{\mu \nu }\gamma _{\mu \nu }\frac{T}{2}\right]^{(\sigma )}.$$ (83) In the above derivation of Eq. (82), we have used several approximations. This derivation has the merit of having a result with a rather easy intuitive interpretation. The gravitational field is propagated through the massless field $`\widehat{\xi }_{(\pm )}^{(5)}`$. At a point far from the source $`T_{\mu \nu }`$, $`\widehat{\xi }_{(\pm )}^{(5)}`$ generates a cloud of metric perturbations through the interaction with the massive KK modes. However, the above approximate derivation is not completely satisfactory because some aspects of general relativity are already tested with very high precision. Hence, we present an alternative and more complete treatment below. First we consider to apply the same trick used in the second line of Eq.(78) to the complete Green function containing all massive modes. Then the long range part of the Green function reduces to $`G_q\delta ^4(\mathrm{\Delta }x^\mu )_iq_i(z)q_i(z^{})/m_i^2`$. This replacement is exactly valid when we focus on the long range force. Now one can notice that to use this Green function neglecting the terms corresponding to the short range force is equivalent to solve the equation for $`\varphi ^N`$ (30) by setting $`\mathrm{}^{(4)}=0`$ from the beginning. For the case with $`\mathrm{}^{(4)}=0`$, we have already obtained the general solutions to Eq. (30), i.e., Eqs.(35) and (37). For convenience, here we quote the previous results in a slightly different notation, namely $$\varphi ^N=\underset{\sigma =\pm }{}u_\sigma f^{(\sigma )}(x^\rho ),$$ (84) with $$u_\pm :=1\frac{2H}{a^2}_{y^{()}}^ya^2(y^{})𝑑y^{}.$$ (85) Then the junction condition (27) with $`\mathrm{}^{(4)}=0`$ determines $`f^{(\pm )}`$ as $`f^{(\pm )}=2Na_{(\pm )}^2\widehat{\xi }_{(\pm )}^5`$. Substituting these back into the expression for $`\varphi ^N`$, we obtain $$\varphi ^N(y^{(\pm )})=2H\widehat{\xi }_{(\pm )}^52N\underset{\sigma =\pm }{}\left[\sigma a_{(\sigma )}^2\widehat{\xi }_{(\sigma )}^5\right].$$ (86) Adding this contribution to the contributions coming from the zero mode $`TT`$ part and $`\widehat{\xi }_{(\pm )}^5`$ in Eq. (71) for the induced metric on the branes, Einstein gravity at the linear order is recovered. It will be illustrative to give the next order correction. The source term for the next order correction is $`\mathrm{}^{(4)}\varphi _0^N\left(1+_{\sigma =\pm }2ϵ^{(\sigma )}\delta (yy^{(\sigma )})\right)`$, where we have denoted the solution of the lowest order approximation (86) by $`\varphi _0^N`$. By using Eq. (26), $`\mathrm{}^{(4)}\varphi _0^N`$ is evaluated as $$\mathrm{}^{(4)}\varphi _0^N=\frac{\kappa N}{3}\underset{\sigma =\pm }{}u_\sigma (y)a_{(\sigma )}^4T^{(\sigma )}.$$ (87) Then, by using the standard Green function method, we obtain $`\varphi _1^N(y)`$ $`=`$ $`u_+\left[{\displaystyle _{y^{(+)}}^y}{\displaystyle \frac{u_{}}{\dot{\phi }_0^2}}\left({\displaystyle \frac{3N}{\kappa }}\mathrm{}^{(4)}\varphi _0^N\right)𝑑y^{}+K_+\right]`$ (89) $`u_{}\left[{\displaystyle _{y^{()}}^y}{\displaystyle \frac{u_+}{\dot{\phi }_0^2}}\left({\displaystyle \frac{3N}{\kappa }}\mathrm{}^{(4)}\varphi _0^N\right)𝑑y^{}K_{}\right],`$ where $`K_+`$ and $`K_{}`$ are integration constants determined from the junction condition as $`K_\pm `$ $`=`$ $`\left[{\displaystyle \frac{N^2}{\dot{\phi }^2}}\right]^{(\pm )}ϵ^{(\pm )}`$ (91) $`\times \left[\left\{\left(1\pm {\displaystyle \frac{H}{a^2N}}\right)a^4T\right\}^{(\pm )}+a_{()}^4T^{()}\right].`$ For simplicity, we take again the $`ϵ^{(\pm )}0`$ limit. Then, we have $`K_\pm =0`$, and we obtain $$\varphi _1^{N(\pm )}=N^2\underset{\sigma =\pm }{}I_{\pm ,\sigma }a_{(\sigma )}^4T^{(\sigma )},$$ (92) where we have defined $$I_{i,j}:=_{y^{(+)}}^{y^{()}}\frac{u_i(y)u_j(y)}{\dot{\phi }_0^2}𝑑y.$$ (93) In the weak back reaction case, we can approximately evaluate this integral by using the following facts. First, we note that the combination $`a^2u_\pm `$ can be shown to be a slowly changing function for $`yy_c`$ in the weak back reaction case. As before, $`y_c`$ is the peak location of $`(a^4\dot{\phi }_0^2)^1`$. To show this, we use the fact that $`\dot{\phi }_0`$ is approximated by $`\nu _1B_1e^{\nu _1y}`$ for $`yy_c`$. This constancy of $`a^2u_\pm `$ indicates that the integrand of Eq. (93) is approximately proportional to $`(a^4\dot{\phi }_0^2)^1`$. Then by using the background geometry defined by Eqs. (39) and (40), $`I_{+,+}`$ is evaluated like $`I_{+,+}`$ $``$ $`\left[a^2u_+\right]_{y=d}^2{\displaystyle _0^d}{\displaystyle \frac{dy}{a^4\dot{\phi }_0^2}}{\displaystyle \frac{\kappa \mathrm{}}{4m_0^2}}e^{2d/\mathrm{}},`$ where we have used (53). The other components are also evaluated in a similar way to obtain the relations $`I_{+,+}e^{2d/\mathrm{}}I_{+,}e^{4d/\mathrm{}}I_,`$. The correction obtained by substituting these estimates for $`I_{i,j}`$ into (92) is consistent with the contribution from the second term in the square brackets in Eq. (80). If we set $`\mathrm{}^{(4)}=0`$ there, we recover the same result. It is easy to check that the treatment of taking $`\mathrm{}^{(4)}`$ to be small is consistent if the distribution of the energy momentum tensor is sufficiently smooth. In the treatment presented here, we have taken into account the contribution from all the massive modes simultaneously. The approximation is completely valid as long as we consider smooth matter distribution as compared with the mass scale of the lowest massive mode. However, the lowest mass on the positive tension brane becomes very small if we consider the case in which we are living on the negative tension brane. In this sense, the treatment presented here is rather restrictive when we discuss perturbations caused by the matter fields on the positive tension brane. ## V TT part Revisited Applying the same technique that we have used in the preceding section for the scalar-type perturbations, we can also deal with the KK mode contribution for the $`TT`$ part $`h_{\mu \nu }^{(TT)}`$. By using the Green function method, we can easily calculate the zero mode contribution like $$h_{\mu \nu }^{(0)}=2N\kappa a^2(y)(\mathrm{}^{(4)})^1\underset{\sigma =\pm }{}a_{(\sigma )}^2\mathrm{\Sigma }_{\mu \nu }^{(\sigma )}.$$ (94) After using Eq. (26), we can recover Eq. (LABEL:quote). Substituting $`h_{\mu \nu }^{(TT)}=h_{\mu \nu }^{(0)}+h_{\mu \nu }^{(KK)}`$ into (25), we obtain the equation for the KK contribution, $`\left[a^2\mathrm{}^{(4)}+\widehat{L}^{(TT)}\right]h_{\mu \nu }^{(KK)}`$ (95) $`=2\kappa {\displaystyle \underset{\sigma =\pm }{}}\mathrm{\Sigma }_{\mu \nu }^{(\sigma )}\left[a_{(\sigma )}^2N\delta (yy^{(\sigma )})\right].`$ (96) As before, neglecting the $`\mathrm{}^{(4)}`$-term for the massive KK contribution, we can solve this equation like $`h_{\mu \nu }^{(KK)}`$ $`=`$ $`2N\kappa {\displaystyle \underset{\sigma =\pm }{}}a_{(\sigma )}^2\mathrm{\Sigma }_{\mu \nu }^{(\sigma )}a^2(y)`$ (98) $`\times \left({\displaystyle _{y^{(\sigma )}}^y}{\displaystyle \frac{dy^{}}{a^4(y^{})}}{\displaystyle _{y^{(\sigma )}}^y^{}}𝑑y^{\prime \prime }a^2(y^{\prime \prime })C_{(\sigma )}\right),`$ where $`C_{(+)}`$ and $`C_{()}`$ are constants. We should recall that we have already subtracted the zero mode contribution. Hence $`h_{\mu \nu }^{(KK)}`$ must be orthogonal to the zero mode. This requires $$_{y^{(+)}}^{y^{()}}\frac{dy}{a^2}u_0(y)h_{\mu \nu }^{(KK)}(y)_{y^{(+)}}^{y^{()}}𝑑yh_{\mu \nu }^{(KK)}(y)=0.$$ (99) The constants $`C_{(+)}`$ and $`C_{()}`$ in the solution (98) are determined by imposing this condition. For simplicity we again adopt (39) with (40) as the background geometry. Then, from condition (99), $`C_{(\pm )}`$ are explicitly calculated, and the resulting KK contribution becomes $`h_{\mu \nu }^{(KK)}={\displaystyle \frac{\kappa \mathrm{}}{4(1e^{2d/\mathrm{}})}}{\displaystyle \underset{\sigma =\pm }{}}\mathrm{\Sigma }_{\mu \nu }^{(\sigma )}`$ (100) $`\times \left[e^{(2y2d)/\mathrm{}}2a_{(\sigma )}^2+e^{2y/\mathrm{}}\left({\displaystyle \frac{4a_{(\sigma )}^2d\mathrm{}^1}{1e^{2d/\mathrm{}}}}1\right)\right].`$ (101) Hence, on the respective branes, the contributions from the KK modes become $`h_{\mu \nu (+)}^{(KK)}{\displaystyle \frac{\kappa \mathrm{}}{4}}\left((34d/\mathrm{})\mathrm{\Sigma }_{\mu \nu }^{(+)}+\mathrm{\Sigma }_{\mu \nu }^{()}\right),`$ (102) $`h_{\mu \nu ()}^{(KK)}{\displaystyle \frac{\kappa \mathrm{}}{4}}\left(\mathrm{\Sigma }_{\mu \nu }^{(+)}+\mathrm{\Sigma }_{\mu \nu }^{()}\right),`$ (103) where we have assumed $`d/\mathrm{}1`$. Here we should recall the remaining degrees of freedom for the 4-dimensional gauge transformation. Using these degrees of freedom, $`\mathrm{\Sigma }_{\mu \nu }`$ can be replaced with $`T_{\mu \nu }\gamma _{\mu \nu }T/3`$ with the aid of Eq. (26). Hence, the KK modes, of course, do not give any long range force contribution. If one takes the $`d/\mathrm{}\mathrm{}`$ limit, $`h_{\mu \nu (+)}^{(KK)}`$ seems to diverge. But this is just due to the breakdown of the approximation. In this limit, the mass difference of the KK modes becomes zero. ## VI summary In this paper we have developed a systematic procedure to evaluate the perturbations in the 5-dimensional brane world model proposed by Randall and Sundrum supplemented with the moduli stabilization mechanism by Goldberger and Wise. We have first investigated in detail the mass spectrum of this model. In the case without stabilization mechanism, there was a scalar-type massless mode, which was called radion in Ref.. We have clarified how this massless mode disappears once we switch on the stabilization mechanism. We have also estimated the mass eigenvalue and the mode function corresponding to the lowest mass eigenmodes for both tensor-type and scalar-type perturbations, assuming that the back reaction to the background geometry due to the bulk scalar field introduced for the moduli stabilization is not large. The physical mass of the lowest tensor-type mode becomes $`\mathrm{}^1`$ on the negative tension brane and $`e^{d/\mathrm{}}\mathrm{}^1`$ on the positive tension brane. Here $`\mathrm{}`$ is the curvature scale of the background geometry, and $`d`$ is the proper distance between the two branes. In the original model, $`\mathrm{}^1`$ is supposed to be TeV scale. For the physical mass of the lowest scalar-type mode, we have obtained rather general formulas, (53) and (54). To proceed further in estimating the mass, we have specified a model for the bulk potential of the scalar field. We have considered a simple quadratic potential whose mass is given by $`M`$, and we have assumed that the 5-dimensional gravitational constant is also $`\mathrm{}^3`$. We pointed out that in this model there are two regimes in which the weak back reaction condition holds. The first case is the one in which the vacuum expectation values of the scalar field on the branes, $`\phi _{(\pm )}`$, are sufficiently small compared with the background energy scale, i.e., $`\phi _{(\pm )}\mathrm{}^{3/2}`$. In this case, we found that the mass of the lowest scalar-type mode becomes $`\sqrt{|\phi _{()}\phi _{(+)}|\mathrm{}^3}Me^{(\sqrt{1+(M^2\mathrm{}^2/4)}1)d/\mathrm{}}`$. (For a more precise formula, see Eq. (65).) If we take into account the fact that we need to set $`d/\mathrm{}37`$ to solve the hierarchy problem on the negative tension brane, this factor is at most $`0.2\mathrm{}^1\sqrt{|\phi _{()}\phi _{(+)}|\mathrm{}^3}`$ for $`M\mathrm{}0.33`$. The second case is the one in which $`M`$ is small compared with $`\mathrm{}^1`$, which is the situation discussed in Ref.. In this case, we reproduced the same result found there. Namely, the mass is approximately given by $`(|\phi _{()}|\mathrm{}^{3/2})M^2\mathrm{}`$. As pointed out in Ref., the mass scale on the negative tension brane tends to be smaller than the typical background energy scale $`\mathrm{}^1`$. We confirmed that this is a general feature within the context of weak back reaction. However, looking at these formulas, it seems that we can raise the mass by taking a slightly larger vacuum expectation values of the bulk scalar field on both branes, although a large value of $`|\phi _{(\pm )}|`$ is inconsistent with our approximation. Hence, if we remove the technical limitation of weak back reaction, it is not clear whether the physical mass of the lowest scalar-type mode is always smaller than that of the lowest tensor-type mode. We also mention that the mass scale on the positive tension brane is smaller by a factor of $`e^{d/\mathrm{}}`$. Next, we have developed a method to evaluate the explicit form of the perturbations caused by the matter fields confined on the branes. This is a generalization of the results presented in the previous paper. We found that, as is expected, Einstein gravity is exactly recovered for the long range force at the order of linear perturbations when we impose the stabilization mechanism. The formulas for the leading correction from the scalar-type perturbations (92) and that from the KK modes (103) are also derived without specifying the model for the potential of the bulk scalar field. From these formulas, we can read the coupling of the metric perturbations induced on the branes to the matter fields on both branes. As we have confirmed in Sec.V, there is no pathological behavior in all perturbation modes at the level of linear perturbations. This is a rather expected result because almost all important information at the level of linear perturbations is contained in the mass spectrum except for the strength of coupling to the matter fields. However, it is still unclear what happens once we take into account the non-linearity of gravity. To investigate this issue, the formulas obtained in this paper will be useful. Acknowledgements We thank J. Garriga for useful comments and discussions. T.T. acknowledges support from Monbusho System to Send Japanese Researchers Overseas.
warning/0001/cond-mat0001155.html
ar5iv
text
# Spin-dynamic field coupling in strongly THz driven semiconductors : local inversion symmetry breaking ## I INTRODUCTION Optical properties of semiconductors are sensitive to external conditions such as strong static electric fields being applied to the sample. This was predicted more than 40 years ago by Franz and Keldysh and leads to the Franz-Keldysh effect. The effect manifests itself in finite absorption within the gap, near the band edge, and modulation of the above gap absorption spectrum. The effect can be understood in terms of tunneling assisted absorption. This work was generalized to oscillating external fields, $`\stackrel{}{E}(t)=\stackrel{}{E}\mathrm{cos}(\mathrm{\Omega }t)`$, by Yacoby 30 years ago. Yacoby concluded that similar effects to the Franz-Keldysh effect manifest provided that the ponderomotive energy, $$E_f=\mathrm{}\omega _f=\frac{e^2E^2}{4m_r\mathrm{\Omega }^2},$$ (1) magnitude is similar to the photon energy, $`\mathrm{}\mathrm{\Omega }`$, of the external field, i.e. $$\gamma =\frac{\omega _f}{\mathrm{\Omega }}=\frac{e^2E^2}{4\mathrm{}m_r\mathrm{\Omega }^3}1.$$ (2) Here $`1/m_r=1/m_c+1/m_v`$ is the reduced mass of the effective valence and conduction band masses $`m_v`$ and $`m_c`$ respectively. Then the field induced effects extend a few $`\mathrm{}\mathrm{\Omega }`$ around gap and may be viewed as being due to a combination of quantum mechanical tunneling and multi photon processes. Experimentally it is exceedingly difficult to fulfill the condition (2) such that the effect extends over a range which can be resolved experimentally. The advent of the free electron laser as a source of intense coherent radiation in the THz regime has made it possible to enter the regime $`\gamma 1`$, which now has come to be known as the dynamical Franz-Keldysh (DFK) regime. Works reporting on optical experiments in the DFK regime include Refs. \[\]. This has also lead to renewed theoretical interest of THz electro optics, see e.g. Refs. \[\]. It has been shown that the physical signature of the DFK effect is enhanced with reduced dimensionality , which has lead to experimental investigation of the two dimensional analogue of Yacoby’s prediction in quantum wells at low temperatures . Of particular relevance to the present work is the observation of optical sidebands in the transmission spectrum at frequencies $`\omega _p\pm 2n\mathrm{\Omega }`$, where $`\omega _p`$ is the probe frequency and $`n`$ is an integer. Thus the sidebands only appear at even multiples of the driving frequency $`\mathrm{\Omega }`$ in quantum wells. This observation can be understood as being a direct consequence of the underlying inversion symmetry of the system , e.g. the dispersion has the property $`ϵ_n(\stackrel{}{k})=ϵ_n(\stackrel{}{k})`$, $`n`$ is the band index. We remark from the onset that it is assumed that the THz field does not induce interband coupling. Absence of odd sidebands has also been observed in the presence of strong quantizing magnetic fields . In Ref. \[\] it has now been experimentally demonstrated that odd sidebands appear if the inversion symmetry is broken by driving the system such that the field oscillates in the growth direction of an asymmetric quantum well. These observations support the assumption that the THz field only induces intraband dynamics. In the light of these observations and the theoretical understanding thereof it therefore comes as a curious surprise that odd sideband formation is observed of resonance, e.g. away from the gap, in apparently inversion symmetric bulk samples . In the present work we theoretically demonstrate that inversion symmetry breaking results from spin-dynamic field coupling leading to the formation of odd sidebands. We refer to the well known relativistic effect that a charged particle traveling with velocity $`\stackrel{}{v}`$ in an electric field $`\stackrel{}{E}`$ experiences an effective magnetic field proportional to $`\stackrel{}{v}\times \stackrel{}{E}`$ which couples to the particles spin degree of freedom via a Zeeman term, see e.g. Ref. \[\]. In the present case the electric field is the strong THz field driving the system. The article is organized as follows; in Sec. II we define the theoretical model and calculate the sideband intensities neglecting the spin-THz field coupling and show how only even sidebands appear, in Sec. II we calculate effect of the spin-THz field coupling and finally in Sec. IV we discuss the physical consequences of our results and present numerical calculations for sideband generation in GaAs and InAs. ## II THE MODEL In this section we define the model which we shall study. Our approach is to apply non-equilibrium Green function techniques to include the THz field nonperturbatively. This can be done analytically and the resulting Green functions form the starting point for our subsequent calculations, which is to account for the relativistic term perturbatively in the Born approximation. The resulting Green functions are then applied to determine the interband susceptibility which describes the linear response of the system to a weak optical near infrared probe field. We consider a direct gap undoped bulk semiconductor subject to an intense linearly polarized THz field, $`E_{THz}(t)`$. We shall assume that effects due to the finite wave-vector are negligible, that is we assume that the THz field is uniform. Thus the system considered remains translationally invariant and the wave vector is still a good quantum number. We shall model the system with a four band model which is a generalization of the model introduced in Ref. \[\]. A conduction band and a valence band, each carrying one of two spin states. We start our analysis with the second quantized Hamiltonian $$H^0=\underset{\alpha =\{c,v\},\sigma ,\stackrel{}{k}}{}ϵ_{\alpha \sigma }[\mathrm{}\stackrel{}{k}+e\stackrel{}{A}(t)]c_{\alpha \sigma \stackrel{}{k}}^{}c_{\alpha \sigma \stackrel{}{k}}.$$ (3) Here $`\sigma `$ is the spin state index, $`\alpha `$ the band index and $`k`$ is the wave vector of the state. The THz field is introduced nonperturbatively via the vector potential $`\stackrel{}{A}=\stackrel{}{E}\mathrm{sin}(\mathrm{\Omega }t)/\mathrm{\Omega }`$. We shall assume that the THz field is oriented in the $`\widehat{z}`$-direction. The probe has wave vector $`\stackrel{}{q}`$ and is oriented such that it forms an angle $`\theta `$ to the $`\widehat{z}`$-direction, see Fig. 1. For simplicity, we shall assume that the bands are parabolic, even away from the gap. We assume that $$ϵ_{c\sigma }(\stackrel{}{p})=\frac{p^2}{2m_c}+ϵ_g$$ (4) for the conduction band. Here $`m_c`$ is effective mass for the conduction band and $`ϵ_g`$ is energy gap of the semiconductor. For the valence band we describe the dispersion as $$ϵ_{v\sigma }(\stackrel{}{p})=\frac{p^2}{2m_v},$$ (5) where the $`m_v`$ is the effective mass of valence band electrons. Note that $`ϵ_{\alpha \sigma }(\stackrel{}{p})=ϵ_{\alpha \sigma }(\stackrel{}{p})`$, so the initial Hamiltonian is inversion symmetric. Furthermore, we shall assume that the interband dipole matrix elements $`d_{cv\sigma _c\sigma _v}`$ are locally independent of the wave vector, but that they depend on the general underlying symmetries of the Brillouin zone being probed. For instance, near a point of high symmetry like the $`\mathrm{\Gamma }`$-point we take $`d_{cv\sigma _c\sigma _v}\delta _{\sigma _c\sigma _v}`$, reflecting the near gap selection rules. Away from the $`\mathrm{\Gamma }`$-point these selection rules are not in force, and we shall assume that all the matrix elements are finite but constant. Below we shall thus consider two distinct regimes, the near gap regime characterized by the selection rules and the far regime, where the selection rules are not in force. The relevant quantity to study in order to glean the optical properties of the system, is the interband susceptibility. The non-equilibrium causal interband susceptibility on the Keldysh contour is given by $$\chi _{cv\beta }\stackrel{}{q},t,t^{})=\frac{i}{\mathrm{}}\frac{dk}{(2\pi )^3}g_{c\sigma _c\sigma _c^{}}(\stackrel{}{k}+\stackrel{}{q},t,t^{})g_{v\sigma _v^{}\sigma _v}(\stackrel{}{k},t^{},t),$$ (6) where $`\beta =(\sigma _c\sigma _v\sigma _c^{}\sigma _v^{})`$ and the contour ordered Keldysh Green functions are defined by $$g_{\alpha \sigma \sigma ^{}}(\stackrel{}{k},t,t^{})=i\mathrm{T}_c[c_{\alpha \sigma \stackrel{}{k}}(t)c_{\alpha ^{}\sigma ^{}\stackrel{}{k}}^{}(t^{})].$$ (7) For an overview on non-equilibrium Green functions see . The physically relevant susceptibility is the real time retarded susceptibility. We analytically continue the causal susceptibility using the Langreth rules and find that $$\chi _{cv\beta }^r(\stackrel{}{q},t,t^{})=\frac{i}{\mathrm{}}\frac{dk}{(2\pi )^3}\left[g_{c\sigma _c\sigma _c^{}}^<(\stackrel{}{k}+\stackrel{}{q},t,t^{})g_{v\sigma _v^{}\sigma _v}^a(\stackrel{}{k},t^{},t)+g_{c\sigma _c\sigma _c^{}}^r(\stackrel{}{k}+\stackrel{}{q},t,t^{})g_{v\sigma _v^{}\sigma _v}^<(\stackrel{}{k},t^{},t)\right].$$ (8) In the quasiparticle picture one describes the lesser function, $`g^<`$, as a distribution function $`f`$ times the spectral function via $`g^<=ifa`$. The spectral function is given by $`a=i(g^rg^a)`$ in a non-equilibrium system. In equilibrium this generalization becomes the familiar $`a_{eq}=2\mathrm{I}\mathrm{m}g^r`$. In an undoped semiconductor all the valence band is occupied, hence the distribution function is one, but no states in the conduction band are occupied and the corresponding distribution function is zero. We thus put $`g_v^<=(g_v^rg_v^a)`$ and $`g_c^<=0`$. The interband susceptibility becomes $$\chi _{cv\beta }^r(\stackrel{}{q},t,t^{})=\frac{i}{\mathrm{}}\frac{dk}{(2\pi )^3}g_{c\sigma _c\sigma _c^{}}^r(\stackrel{}{k}+\stackrel{}{q},t,t^{})g_{v\sigma _v^{}\sigma _v}^a(\stackrel{}{k},t^{},t).$$ (9) Probing the system with a weak beam $`E_p(\stackrel{}{q},\omega _pt)E_pe^{i\omega _pt}`$, the linearly induced interband polarization is $$P_{cv}(t)=e^2E_p\underset{\sigma _c^{}\sigma _v^{}\sigma _c\sigma _v}{}d_{cv\sigma _c\sigma _v}d_{cv\sigma _c^{}\sigma _v^{}}𝑑t^{}\chi _{cv\sigma _c\sigma _v\sigma _c^{}\sigma _v^{}}^r(\stackrel{}{q},t,t^{})e^{i\omega _pt^{}}.$$ (10) Driving the system with THz frequency $`\mathrm{\Omega }`$ the induced polarization will be of the form $$P_{cv}(t)=e^2E_p\underset{n}{}\eta _n(\omega _p)e^{i(\omega _p+n\mathrm{\Omega })t},$$ (11) which spectrally is a comb of oscillating dipoles giving rise to sidebands irradiating with intensity proportional to $`I_n(\omega _p)=|\eta _n(\omega _p)|^2`$ and frequency $`\omega _p+n\mathrm{\Omega }`$. The absorption of the probe is proportional to $`\mathrm{Im}\eta _0(\omega _p)`$. Neglecting relativistic effects, the Green functions obeys the Dyson equation $$\left\{i\mathrm{}_tϵ_{\alpha \sigma }(\mathrm{}\stackrel{}{k}+e\stackrel{}{A}(t))\right\}g_{\alpha \sigma \sigma ^{}}^{0r/a}(\stackrel{}{k},t,t^{})=\mathrm{}\delta _{\sigma \sigma ^{}}\delta (tt^{}),$$ (12) with the appropriate boundary conditions. These Green functions which include the THz field to all orders form the starting point of our subsequent calculations. Thus our unperturbed Green functions take the THz field into account from the onset. The equations are readily integrated with the results $$g_{\alpha \sigma \sigma ^{}}^{0r/a}(\stackrel{}{k},t,t^{})=\pm \mathrm{}\delta _{\sigma \sigma ^{}}\theta (\pm tt^{})\mathrm{exp}\left\{i_t^{}^t\frac{ds}{\mathrm{}}ϵ_{\alpha \sigma }(\mathrm{}\stackrel{}{k}+e\stackrel{}{A}(s))\right\}.$$ (13) Using these Green functions in order to evaluate the interband susceptibility leads to a Gaussian integral in $`\stackrel{}{k}`$-space which we evaluate with the result $$\chi _{cv\beta }^{0r}(\stackrel{}{q},\tau ,T)=\frac{e^{i\pi /4}}{(2\pi )^{3/2}}\frac{(m_r\mathrm{\Omega })^{3/2}}{\sqrt{\mathrm{}}}e^{i(\mathrm{}q^2/(2M)+\omega _f+\omega _g)\tau }\underset{n}{}e^{in\pi /2}f_n^0(\mathrm{\Omega }\tau )e^{i2n\mathrm{\Omega }T}.$$ (14) Here $`\tau =tt^{}`$, $`T=(t+t^{})/2`$, $`M=m_v+m_c`$, and we have defined $$f_n^l(x)=\frac{\theta (x)}{x^{3/2}}\mathrm{sin}^l(x/2)\mathrm{exp}\left(4i\gamma \frac{\mathrm{sin}^2(x/2)}{x}\right)J_n\left(\gamma \left(\mathrm{sin}x4\frac{\mathrm{sin}^2(x/2)}{x}\right)\right).$$ (15) Other components of the susceptibility are zero. We thus obtain that $$\eta _{2n}(\omega _p)=2d_{cv}d_{cv}\frac{e^{i(12n)\pi /4}}{(2\pi )^{3/2}}\frac{(m_r\mathrm{\Omega })^{3/2}}{\sqrt{\mathrm{}}}_0^{\mathrm{}}𝑑\tau f_n^0(\mathrm{\Omega }\tau )e^{i\omega _{qn}\tau },$$ (16) where we have used that $`d_{cv}=d_{cv}`$ and we have defined $`\omega _{qn}=\omega _p+n\mathrm{\Omega }\mathrm{}q^2/(2M)\omega _f\omega _g`$. The remaining integral is readily evaluated numerically. ### A PHYSICAL IMPLICATIONS Here several physical results emerge. Sidebands do only appear at frequencies $`\omega _p\pm 2n\mathrm{\Omega }`$, no sidebands involving an odd number of THz photons appear. This is a direct consequence of the inversion symmetry of the Hamiltonian. The response is independent on the relative orientation of the probe and the THz field, which is reflected by that the result only depends on the magnitude of $`q`$. In Fig. 2 we illustrate the sideband intensities for the first 4 sidebands as a function of the probe frequency. These results are valid in the near zone, and are in qualitative agreement with the experimental findings reported in for quantum wells. Both the probe and the irradiating sidebands are close to real states, hence the asymmetry around $`\omega _g`$. The relation $`I_n(\omega )=I_n(\omega +n\mathrm{\Omega })`$ holds. The probe or the sideband is near the main spectral feature, the band edge in the present case. In the quantum well the main feature is the 1s exciton resonance . ## III THE RELATIVISTIC SPIN-THz FIELD COUPLING In this section we shall take into account the relativistic effects. The coupling to the spins comes physically about because a moving electron in the presence of an electric field experiences a local magnetic field, proportional to $`\stackrel{}{v}\times \stackrel{}{E}`$, which couples to its spin via a Zeeman interaction, see e.g. . This is the effect which gives rise to spin orbit coupling in atoms and solids. In the situation considered here the strong THz field provides the electric field giving rise to the effect. We describe the contribution to the Hamiltonian as $$U_i(\stackrel{}{k},t)=\alpha _ie\stackrel{}{\sigma }(\stackrel{}{k}\times \stackrel{}{E}(t)),$$ (17) here $`\alpha _i`$, $`i=\{c,v\}`$, is an effective coupling constant which depends on the material, $`\stackrel{}{\sigma }`$ is a vector of the Pauli matrices and $`\stackrel{}{E}(t)`$ is the intense THz field. With $`\stackrel{}{E}`$ oriented in the $`\widehat{z}`$ direction (17) becomes $$U_i(\stackrel{}{k},t)=\alpha _ie\left(\begin{array}{cc}0& k_y+ik_x\\ k_yik_x& 0\end{array}\right)E_z(t).$$ (18) Physically the same kind of term has been studied by Raman spectroscopy in asymmetric GaAs quantum wells. Then the external field is static and comes about due to the asymmetry of the quantum well confining potential. The term breaks time reversal symmetry and thus influences weak localization leading to weak anti-localization . The important thing for the present study is that $`\alpha _c`$ has been determined both experimentally and theoretically for GaAs leading to $`\alpha _c5`$Å<sup>2</sup>, and for InAs $`\alpha _c110`$Å<sup>2</sup>. From the point of view of an effective mass picture we estimate the coefficient for the valence band by $`\alpha _v=\alpha _c(m_c/m_v)^2`$. These parameters are such that they lead to small corrections compared to the remainder of the Hamiltonian. We thus calculate the correction to the Green functions to lowest order in the Born approximation, with the result $`g_i^{r/a}(\stackrel{}{k},t,t^{})`$ $`=`$ $`g_i^{0r/a}(\stackrel{}{k},t,t^{})+{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{ds}{\mathrm{}^2}}g_i^{0r/a}(\stackrel{}{k},t,s)U_i(\stackrel{}{k},s)g_i^{0r/a}(\stackrel{}{k},s,t^{})`$ (19) $`=`$ $`\left(\begin{array}{cc}1& \pm l_i(t,t^{})(k_y+ik_x)\\ \pm l_i(t,t^{})(k_yik_x)& 1\end{array}\right)g_i^{0r/a}(\stackrel{}{k},t,t^{}),`$ (22) where we have introduced the quantity $`l_i(t,t^{})`$ $`=`$ $`2{\displaystyle \frac{\alpha _ieE_z}{\mathrm{}\mathrm{\Omega }}}\mathrm{cos}\left(\mathrm{\Omega }{\displaystyle \frac{t+t^{}}{2}}\right)\mathrm{sin}(\mathrm{\Omega }(tt^{})).`$ (23) Note that $`l_i(t,t^{})=l_i(t^{},t)`$. Again in this case it is straightforward to perform the $`\stackrel{}{k}`$ integration in order to obtain the interband susceptibility via Eq. (9). The integrals involved are Gaussian times polynomials in $`\stackrel{}{k}`$. We remark that higher order contributions preserve this structure as well and are thus readily determined. Now the relative orientation of the probe with respect to the driving field becomes important. We assume that $`\stackrel{}{q}`$ is perpendicular to the $`\widehat{x}`$ direction. We find that $`\chi _{cv\beta }(\stackrel{}{q},\tau ,T)=\chi _{cv\beta }^{0r}(\stackrel{}{q},\tau ,T)`$ for $`\beta \{,,,\}`$, $`\chi _{cv\beta }(\stackrel{}{q},\tau ,T)`$ $`=`$ $`{\displaystyle \frac{1}{(2\pi )^{3/2}}}{\displaystyle \frac{2e\alpha _vq_yE_z}{\mathrm{}\mathrm{\Omega }}}{\displaystyle \frac{e^{i\pi 5/4}}{\sqrt{\mathrm{}}}}{\displaystyle \frac{m_r^{5/2}\mathrm{\Omega }^{3/2}}{m_c}}e^{i(\mathrm{}q^2/(2M)+\omega _f+\omega _g)\tau }`$ (25) $`\times \mathrm{cos}(\mathrm{\Omega }T){\displaystyle \underset{n}{}}e^{in\pi /2}f_n^1(\mathrm{\Omega }\tau )e^{i2n\mathrm{\Omega }T},`$ for $`\beta \{,,,\}`$, $`\chi _{cv\beta }(\stackrel{}{q},\tau ,T)`$ $`=`$ $`{\displaystyle \frac{1}{(2\pi )^{3/2}}}{\displaystyle \frac{2e\alpha _cq_yE_z}{\mathrm{}\mathrm{\Omega }}}{\displaystyle \frac{e^{i\pi 5/4}}{\sqrt{\mathrm{}}}}{\displaystyle \frac{m_r^{5/2}\mathrm{\Omega }^{3/2}}{m_c}}\left(1i{\displaystyle \frac{m_r}{m_c}}\right)e^{i(\mathrm{}q^2/(2M)+\omega _f+\omega _g)\tau }`$ (27) $`\times \mathrm{cos}(\mathrm{\Omega }T){\displaystyle \underset{n}{}}e^{in\pi /2}f_n^1(\mathrm{\Omega }\tau )e^{i2n\mathrm{\Omega }T},`$ for $`\beta \{,,,\}`$. Other components are of second order in $`\alpha `$ and are negligible. We can now find the sideband intensities from $`\eta _{2n}(\omega _p)`$ $`=`$ $`2(d_{cv}d_{cv}+d_{cv}d_{cv}){\displaystyle \frac{e^{i\pi (12n)/4}}{(2\pi )^{3/2}}}{\displaystyle \frac{(m_r\mathrm{\Omega })^{3/2}}{\sqrt{\mathrm{}}}}{\displaystyle _0^{\mathrm{}}}𝑑\tau f_n^0(\mathrm{\Omega }\tau )e^{i\omega _{qn}\tau }`$ (28) and $`\eta _{2n+1}(\omega _p)`$ $`=`$ $`4d_{cv}d_{cv}{\displaystyle \frac{e^{i\pi (5+2n)/4}}{(2\pi )^{3/2}}}{\displaystyle \frac{(m_r\mathrm{\Omega })^{3/2}}{\sqrt{\mathrm{}}}}{\displaystyle \frac{m_r}{m_c}}{\displaystyle \frac{2e(\alpha _v+\alpha _c(1i\frac{m_r}{m_c}))q_yE_z}{\mathrm{}\mathrm{\Omega }}}{\displaystyle _0^{\mathrm{}}}𝑑\tau (f_n^1(\mathrm{\Omega }\tau )if_{n+1}^1(\mathrm{\Omega }\tau ))e^{i\omega _{qn}\tau }`$ (29) This is the main result of the present work. ## IV RESULTS AND DISCUSSIONS When $`d_{cv}d_{cv}`$, as is the case near the $`\mathrm{\Gamma }`$-point, near resonance, the relativistic effects are suppressed and we obtain the results of the previous section. The relativistic effects, i.e. contributions to the sideband intensity proportional to $`\alpha _i`$, manifest themselves in the far zone, away from resonance. The relativistic term leads to the formation of odd sidebands. Thus the inversion symmetry has been broken and relativistic self induced local symmetry breaking has taken place. The inversion symmetry breaking is local because it is only detectable with finite probe wave vector $`\stackrel{}{q}`$. We say that the symmetry breaking is global if it leads to odd sidebands even when $`q0`$. An example of a global symmetry breaking is for instance when a linear term is present in the dispersion. The nature of the even (28) and the odd (29) sidebands is fundamentally different. Let $`\theta `$ be the angle between the probe beam, $`\widehat{q}`$, and the direction the electric field is polarized in, $`\widehat{z}`$, see Fig. 1. The even sidebands do not depend on $`\theta `$. They can thus be regarded as a result of pure amplitude modulations of the fundamental probe. For the odd sidebands however, we note that $`I_{2n+1}\mathrm{sin}^2\theta `$. The odd sidebands therefore vanish when the probe is aligned parallel with the electric field, but is at a maximum when they are aligned perpendicular to each other. Taking $`d_{cv}=d_{cv}`$ to be constant, we have evaluated the sideband intensities according to (28) and (29) using the material parameters for the heavy hole band and conduction band for GaAs and InAs. We have taken $`\mathrm{}\mathrm{\Omega }=30`$meV for the calculations. In Fig. 3 we show $`I_n`$, $`n=1,2,3,4`$, for THz intensities corresponding to $`\gamma =0.1,0.5,1.0,2.0`$ as a function of the probe frequency for GaAs. Near resonance $`I_2`$ is dominating. Moving away from the resonance $`I_1`$ decays much slower than the other sidebands and eventually becomes dominating. Notice also that $`I_4>I_3`$ within the region shown. The even sidebands are more sensitive to the THz intensity than the odd sidebands. In Fig. 4 we show the results for InAs. Qualitatively the behavior is identical except that for the higher THz intensities we note that $`I_4<I_3<I_2<I_1`$ away from resonance. Quantitatively the odd sidebands are about a factor $`100`$ stronger than for GaAs. In Fig. 5 we show the same sideband intensities as a function of the THz intensity keeping the probe frequency fixed, $`\omega /\mathrm{\Omega }=2.5,5.0,7.5,10.0`$, for GaAs. Note how $`I_2`$ and $`I_3`$ seem to rise in a similar manner. It is noteworthy as well how feature less $`I_1`$ is, in fact $`I_1`$ is mostly linear $`\gamma `$ within the regimes we have investigated. In Fig. 6 we show the same results for InAs. We have considered other possible sources of inversion symmetry breaking. Spin-orbit splitting due to the crystal field leads to a cubic contribution in the effective Hamiltonian for zinc-blend structures , which breaks the inversion symmetry. We find however that the contribution to the sidebands due to this term is three orders of magnitude less than the contribution due to (17) considered above for GaAs and InAs. Another possible source is the band bending at the edge of the sample due to the pinning of the Fermi level associated with charge accumulation of residual charge carriers at the surface of the sample. Thus the relevant inversion symmetry may be broken if the THz field has an electric field component perpendicular to the surface of the sample. We have assumed that this is not the case. In summary, we have studied the nonlinear generation of optical sidebands in strongly THz driven undoped direct gap semiconductors. We have found that relativistic spin-THz field coupling leads to local breaking of inversion symmetry which results in the formation of, otherwise suppressed, odd sidebands in the transmitted wave of a weak near infrared interband probe. We have shown that the even sidebands are independent on the relative orientation of the probe and the linearly polarized THz field, while the odd sidebands depend strongly on the relative orientation of the probe to the THz field. The even sidebands dominate the transmission spectrum for probe frequencies near the gap, but moving the probe away from the gap odd sidebands will eventually dominate. We find that the relativistic effects are orders of magnitude stronger in InAs than in GaAs. ###### Acknowledgements. We thank Dr. Antti-Pekka Jauho and Dr. Simon Pedersen for useful discussions on various inversion symmetry breaking effects in semiconductors, and in particular Dr. Junichiro Kono for discussing his experimental results prior to publication.
warning/0001/hep-ph0001130.html
ar5iv
text
# 1 The lowest anomalous dimensions in the spectrum of twist-3 operators for different 𝑁 calculated by taking into account the 𝒪⁢(1/𝑁_𝑐) correction, Eqs. () and (), in comparison with the corresponding exact numerical results and leading large-𝑁_𝑐 expressions (). 1. Twist-three parton distributions in the nucleon are attracting increasing interest as unique probes of novel quark-gluon correlations in hadrons with clear experimental signature, giving rise to certain asymmetries in experiments with polarized beams and targets. Quantitative studies of such asymmetries are becoming possible with the increasing precision of experimental data at SLAC and RHIC, and can provide for an important part of the future spin physics program on high-luminosity accelerators like ELFE, upgraded CEBAF etc. With this perspective, a detailed theoretical study of twist-three parton distributions in QCD becomes mandatory. Altogether, there exist three twist-3 distribution functions – chiral-odd, $`e(x,Q^2)`$ and $`h_L(x,Q^2)`$, and chiral-even, $`g_2(x,Q^2)`$ – each being a function of parton momentum fraction $`x`$ and the energy scale $`Q^2`$. By virtue of Lorentz invariance the distributions $`h_L`$ and $`g_2`$ contain contributions of twist-2 structure functions $`h_1`$ and $`g_1`$, respectively: $$h_L(x)=2x_x^1\frac{dy}{y^2}h_1(y)+\stackrel{~}{h}_L(x),g_2(x)=g_1(x)+_x^1\frac{dy}{y}g_1(y)+\stackrel{~}{g}_2(x).$$ (1) The QCD description of the remaining genuine twist-3 part of these distributions functions, $`\stackrel{~}{h}_L(x)`$, $`\stackrel{~}{g}_2(x)`$ and of $`e(x)`$, is usually believed to be quite sophisticated. Their moments are related through the QCD equations of motion to the matrix elements of quark-antiquark-gluon operators which have a nontrivial scale dependence and mix with each other under the renormalization.Additional mixture with three-gluon operators is present for flavor-singlet contribution to $`g_2(x,Q^2)`$. In this letter we consider only the flavor-nonsinglet part. As a consequence, the QCD evolution equations for the functions $`\stackrel{~}{h}_L(x)`$, $`\stackrel{~}{g}_2(x)`$, $`e(x)`$ cannot be written in a closed form and require additional nonperturbative input. An important simplification occurs, however, in the large-$`N_c`$ limit. It was shown that to this accuracy the twist-three distributions $`\stackrel{~}{h}_L`$ and $`e`$ as well as the flavor-nonsinglet contribution to $`\stackrel{~}{g}_2`$ satisfy simple DGLAP-type evolution equations (see e.g. ) $`Q^2{\displaystyle \frac{d}{dQ^2}}f(x,Q^2)={\displaystyle \frac{\alpha _s}{4\pi }}{\displaystyle _x^1}{\displaystyle \frac{dz}{z}}P_f^{(0)}(x/z)f(z,Q^2),f=\{\stackrel{~}{h}_L,\stackrel{~}{g}_2^{_{\mathrm{NS}}},e\},`$ $`P_f^{(0)}(z)=2N_c\left\{\left[{\displaystyle \frac{1}{1z}}\right]_++{\displaystyle \frac{1}{4}}\delta (1z)+\sigma _f{\displaystyle \frac{1}{2}}\right\},`$ (2) where $`\sigma _f=1,0`$ and $`1`$ for $`f=\stackrel{~}{h}_L,\stackrel{~}{g}_2`$ and $`e`$, respectively, and $`1/(1z)_+=1/(1z)\delta (1z)_0^1𝑑z^{}/(1z^{})`$. This result implies that the inclusive measurements of twist-three distributions are complete (to the stated accuracy) in the sense that knowledge of the distribution at one value of $`Q^2`$ is enough to predict the distribution at arbitrary $`Q^2`$. From phenomenological point of view it allows to relate the measurements of different experiments to each other and to compare them to the model predictions (including lattice calculations) that typically refer to a low scale. Although the large-$`N_c`$ version of QCD presents a valid theoretical limit, its relevance to the actual $`N_c=3`$ world and its accuracy in predicting the scale dependence of twist-3 distributions is a priory not clear. This work presents the first attempt to go beyond the large-$`N_c`$ approximation in a systematic way. In particular, we calculate the $`1/N_c^2`$ corrections to the evolution kernels in (2) and show that mixing with quark-antiquark-gluon operators remains under control. 2. From the OPE analysis one finds that the scale dependence of the moments of the twist-3 distributions $`_1^1𝑑xx^{N+2}f(x)`$ is governed by renormalization of the set of local composite quark-antiquark-gluon operators ($`k=0`$, $`\mathrm{}`$, $`N`$) $`[S_\mu ^\pm ]_N^k`$ $`=`$ $`\overline{q}(\stackrel{}{D}n)^k\overline{)}nn^\nu [\stackrel{~}{G}_{\mu \nu }\pm iG_{\mu \nu }\gamma _5](\stackrel{}{D}n)^{Nk}q,`$ $`[T_\mathrm{\Gamma }]_N^k`$ $`=`$ $`\overline{q}(\stackrel{}{D}n)^kn_\mu \sigma ^{\mu \rho }\mathrm{\Gamma }n^\nu G_{\nu \rho }(\stackrel{}{D}n)^{Nk}q,\mathrm{\Gamma }=\{\text{1}\text{l},i\gamma _5\}`$ (3) for chiral-even ($`S_\mu ^\pm `$) and chiral-odd distributions ($`T_\mathrm{I}`$ and $`T_{i\gamma _5}`$ for $`e(x)`$ and $`\stackrel{~}{h}_L(x)`$, respectively), see e.g. . Here, $`n_\mu `$ is a light-like vector and $`\stackrel{~}{G}_{\mu \nu }=ϵ_{\mu \nu \rho \lambda }G^{\rho \lambda }/2`$ stands for a dual gluon field strength. To leading order, renormalization of $`T_\mathrm{I}`$ and $`T_{i\gamma _5}`$ is the same and we, therefore, drop the subscript in what follows. Similarly, it is enough to consider the operator $`S^+`$. The operators $`[T]_N^k`$ (and $`[S^+]_N^k`$) with different $`k=0,\mathrm{},N`$ and the same number of covariant derivatives $`N0`$ mix with each other under renormalization. The mixing matrices have been calculated to the leading order (e.g. ) and can be diagonalized numerically for any given $`N`$. The eigenvectors, then, define the multiplicatively renormalizable operators and the eigenvalues give their anomalous dimensions. A disadvantage of this (traditional) approach is that the mixing matrix does not have any obvious structure in this basis and is not symmetric. As the result, the structure of the spectrum remains obscure and the eigenvectors are not mutually orthogonal with any simple weight function. Relative importance of various contributions is also far from being clear.E.g. the hierarchy of entries in the mixing matrix $`𝒪(N)`$, $`𝒪(1)`$, $`𝒪(1/N)`$, etc. is not preserved in the eigenvalues. A systematic $`1/N_c^2`$ expansion of the evolution equations is made possible by going over to the Hamiltonian formulation developed in in . The renormalization group evolution is driven to leading logarithmic accuracy by tree-level counterterms and has the conformal symmetry of the QCD Lagrangian. The evolution kernel can, therefore, be written in an abstract operator form in terms of Casimir operators of the collinear subgroup $`SL(2,R)`$ of the conformal group. In this way, diagonalization of the mixing matrix of the twist-3 operators in (3) can be reformulated as a three-particle quantum mechanical problem $$\mathrm{\Psi }_{N,q}(x_1,x_2,x_3)=_{N,q}\mathrm{\Psi }_{N,q}(x_1,x_2,x_3),$$ (4) defined by the Hamiltonian $$_A=N_c_A^{(0)}\frac{2}{N_c}_A^{(1)},A=\{T,S^+\},$$ (5) with $`_T^{(0)}`$ $`=`$ $`V_{qg}^{(0)}(J_{12})+V_{qg}^{(0)}(J_{23}),_T^{(1)}=V_{qg}^{(1)}(J_{12})+V_{qg}^{(1)}(J_{23})+V_{qq}^{(1)}(J_{13}),`$ $`_{S^+}^{(0)}`$ $`=`$ $`V_{qg}^{(0)}(J_{12})+U_{qg}^{(0)}(J_{23}),_{S^+}^{(1)}=V_{qg}^{(1)}(J_{12})+U_{qg}^{(1)}(J_{23})+U_{qq}^{(1)}(J_{13}).`$ (6) Here, the notation was introduced for two-particle quark-quark and quark-gluon kernels $`V_{qg}^{(0)}(J)`$ $`=`$ $`\psi (J+3/2)+\psi (J3/2)2\psi (1)3/4,`$ (7) $`U_{qg}^{(0)}(J)`$ $`=`$ $`\psi (J+1/2)+\psi (J1/2)2\psi (1)3/4,`$ $`V_{qg}^{(1)}(J)`$ $`=`$ $`{\displaystyle \frac{(1)^{J5/2}}{(J3/2)(J1/2)(J+1/2)}},U_{qg}^{(1)}(J)={\displaystyle \frac{(1)^{J5/2}}{2(J1/2)}},`$ $`V_{qq}^{(1)}(J)`$ $`=`$ $`\psi (J)\psi (1)3/4,U_{qq}^{(1)}(J)={\displaystyle \frac{1}{2}}\left[\psi (J1)+\psi (J+1)\right]\psi (1)3/4.`$ Here and below $`\psi (x)=d\mathrm{ln}\mathrm{\Gamma }(x)/dx`$ stands for the Euler $`\mathrm{\Psi }`$-function; the subscripts ‘1,2,3’ refer to antiquark, gluon and quark fields, respectively. The operators $`J_{ik}`$ are defined as follows $$J_{ik}(J_{ik}1)=L_{ik}^2=(\stackrel{}{L}_i+\stackrel{}{L}_k)^2,$$ (8) where $`\stackrel{}{L}_i`$ are the generators of the $`SL(2,R)`$ group and $`L_{ik}^2`$ are the corresponding two-particle Casimir operators. The generators $`\stackrel{}{L}_i`$ can be realized as differential operators acting on coordinates $`x_i`$ of the wave function $`\mathrm{\Psi }(x_1,x_2,x_3)`$ $$L_{,i}=x_i_{x_i}^2+2j_i_{x_i},L_{+,i}=x_i,L_{0,i}=x_i_{x_i}+j_i,$$ (9) where the conformal spin $`j`$ is equal to $`j_1=j_3=1`$ for the (anti)quark and $`j_2=3/2`$ for the gluon field, respectively. The eigenvalues of the operator $`J_{ik}`$ define the possible values of the conformal spin in the two-parton channel and are given by $`j_i+j_k+n`$ with $`n`$ being nonnegative integer. The Hamiltonians defined above are manifestly $`SL(2,R)`$ invariant: $$[,L_\alpha ]=[,L^2]=[L^2,L_\alpha ]=0,$$ (10) where $`L_\alpha `$ ($`\alpha =0,+,`$) are the total three-particle $`SL(2)`$ generators $$L_\alpha =L_{\alpha ,1}+L_{\alpha ,2}+L_{\alpha ,3},L^2=L_0(L_01)+L_+L_{}.$$ (11) One can, therefore, diagonalize the three operators $`L^2,L_0`$ and $``$ simultaneously. The conditions $$L_0\mathrm{\Psi }_{N,q}(x_1,x_2,x_3)=\left(N+7/2\right)\mathrm{\Psi }_{N,q}(x_1,x_2,x_3),L_{}\mathrm{\Psi }_{N,q}(x_1,x_2,x_3)=0$$ (12) define $`\mathrm{\Psi }_{N,q}(x_1,x_2,x_3)`$ to be a homogeneous polynomial of degree $`N`$ that does not contain factors of $`(x_1+x_2+x_3)`$ and therefore does not vanish as $`_ix_i=0`$. Solving the Schrödinger equation (4) one constructs the basis of multiplicatively renormalizable local operators, $`𝒪_{N,q}`$. Omitting the Lorentz structure, the correspondence is, schematically $$𝒪_{N,q}=\mathrm{\Psi }_{N,q}(D_{\overline{q}},D_g,D_q)\overline{q}Gq$$ (13) with covariant derivatives $`D_{\overline{q}},D_g,D_q`$ acting on the antiquark, gluon and quark, respectively. The corresponding eigenvalues provide the anomalous dimensions $`_{N,q}`$: $$𝒪_{N,q}(Q^2)=\left(\frac{\alpha _s(Q^2)}{\alpha _s(\mu ^2)}\right)^{_{N,q}/b}𝒪_{N,q}(\mu ^2),$$ (14) where $`b=11N_c/32n_f/3`$. Here, parameter $`q`$ enumerates different eigenstates of the mixing matrix. Note that the conformal operators defined in (13) diagonalize the full mixing matrix, including mixing with the operators containing total derivatives. Since at the end only forward matrix elements of these operators enter the parton distributions, taking into account this additional mixing may be seen as an unnecessary complication. The advantage for such formulation is, however, that the Hamiltonian becomes hermitian with respect to the scalar product $$\mathrm{\Psi }_1|\mathrm{\Psi }_2=_0^1𝒟xx_1x_2^2x_3\mathrm{\Psi }_1^{}(x_i)\mathrm{\Psi }_2(x_i)$$ (15) with $`𝒟x=dx_1dx_2dx_3\delta (x_1+x_2+x_31)`$. As well known in quantum mechanics, hermiticity of the Hamiltonian implies that all its eigenvalues (anomalous dimensions) $`_{N,q}`$ are real<sup>§</sup><sup>§</sup>§ Although the conformal symmetry is lost beyond one-loop, it is easy to prove that anomalous dimensions of twist-3 operators remain real to all orders in perturbation theory. Indeed, entries in the mixing matrix are real and complex eigenvalues may only appear in (complex conjugate) pairs. On the other hand, the number of eigenvalues is fixed and in leading order they are non-degenerate (which one can establish by inspection). These two conditions are contradictory since a pair of complex conjugate eigenvalues (i.e. with the same real parts) cannot be obtained from a discrete non-degenerate spectrum at $`\alpha _s0`$ (i.e. with all real parts different) by a perturbative renormalization group flow., and the eigenfunctions are mutually orthogonal: $$\mathrm{\Psi }_{N,q}|\mathrm{\Psi }_{N^{},q^{}}=\delta _{NN^{}}\delta _{qq^{}}\mathrm{\Psi }_{N,q}^2,\mathrm{\Psi }_{N,q}^2=_0^1𝒟xx_1x_2^2x_3|\mathrm{\Psi }_{N,q}(x_i)|^2.$$ (16) Completeness and orthogonality of the eigenstates corresponding to multiplicatively renormalizable operators will be of crucial importance for this work. In particular, using these properties one can expand the moments of the twist-3 parton distributions over the contributions of multiplicatively renormalizable operators as $$_1^1𝑑xx^{N+2}f(x,Q^2)=\underset{q}{}\frac{\mathrm{\Psi }_{N,q}|\mathrm{\Phi }_N}{\mathrm{\Psi }_{N,q}^2}𝒪_{N,q}(Q^2).$$ (17) Here, $`𝒪_{N,q}`$ are reduced (scalar and dimensionless) forward matrix elements of the corresponding local operators over the nucleon state and $`\mathrm{\Phi }_N(x_i)`$ are the coefficients in the re-expansion of twist-3 quark-antiquark operators that define the parton distributions, in terms of quark-antiquark-gluon operators. They are given by QCD equations of motion and for $`x_1+x_2+x_3=0`$ take the simple form $$\mathrm{\Phi }_N^{T,+}=\frac{x_1^{N+1}(x_3)^{N+1}}{x_1+x_3},\mathrm{\Phi }_N^{T,}=(_{x_1}+_{x_3})\frac{\mathrm{\Phi }_{N+1}^{T,+}}{N+2},\mathrm{\Phi }_N^{S^+}=_{x_1}\frac{\mathrm{\Phi }_{N+1}^{T,+}}{N+2}$$ (18) for $`f=e,\stackrel{~}{h}_L`$ and $`\stackrel{~}{g}_2`$, respectively. Note, however, that in order to form the scalar product in (17) one needs to know $`\mathrm{\Phi }_N(x_i)`$ for $`x_1+x_2+x_3=1`$ that corresponds to taking into account contributions to the QCD equations of motion of the operators containing total derivatives. The corresponding expressions are given below in (S0.Ex12). Each term in (17) has an autonomous $`Q^2`$evolution, Eq. (14), and brings a new nonperturbative parameter $`𝒪_{N,q}(\mu ^2)`$. 3. The large-$`N_c`$ Hamiltonians $`^{(0)}`$ in (S0.Ex3) possess an additional ‘hidden’ symmetry, related to the existence of a nontrivial conserved charge $`[_T^{(0)},Q_T]=0,Q_T=\{L_{12}^2,L_{23}^2\}{\displaystyle \frac{9}{2}}L_{12}^2{\displaystyle \frac{9}{2}}L_{23}^2,`$ $`[_{S^+}^{(0)},Q_{S^+}]=0,Q_{S^+}=\{L_{12}^2,L_{23}^2\}{\displaystyle \frac{1}{2}}L_{12}^2{\displaystyle \frac{9}{2}}L_{23}^2,`$ (19) where $`\{,\}`$ stands for the anticommutator. As the result, eigenstates of $`^{(0)}`$ can be labeled by quantized eigenvalues $`q_{\mathrm{}}`$, $`(\mathrm{}=0,1,\mathrm{},N)`$, of the operator $`Q`$: $$Q\mathrm{\Psi }_{N,q}^{(0)}(x_1,x_2,x_3)=q\mathrm{\Psi }_{N,q}^{(0)}(x_1,x_2,x_3).$$ (20) Since the number of degrees of freedom (=3) equals in this case to the number of conserved charges ($`L^2,L_0,Q`$), the quantum mechanical system described by the Hamiltonian $`^{(0)}`$ is completely integrable. This allows to use advanced mathematical methods for the analysis of the spectrum and, in particular, develop the WKB expansion of the eigenvalues at large $`N`$ . The remarkable property of the large-$`N_c`$ evolution kernels that is responsible for the simplicity of the evolution in (2) is that the eigenfunctions corresponding to the states with lowest energy coincide with the coefficient functions $`\mathrm{\Phi }_N`$ entering the expansion in Eq. (17) $$\mathrm{\Phi }_N(x_i)=\mathrm{\Psi }_{N,q_0}^{(0)}(x_i),q_0^S=(N+3)^2+3/8,q_0^{T,\pm }=(N+32)^253/8.$$ (21) To be precise, $`\mathrm{\Phi }_N^{S^+}(x_i)`$ coincides with the ground state of $`_{S^+}^{(0)}`$ and $`\mathrm{\Phi }_N^{T,\pm }(x_i)`$ with the two lowest eigenstates of $`_T^{(0)}`$ with opposite parity with respect to permutations of quarks, $`x_1x_3`$. The corresponding eigenvalues determine the lowest anomalous dimensions for each $`N`$, $`\mathrm{min}_q_{N,q}=_{N,q_0}=N_cE_N+𝒪(1/N_c)`$ and they are equal to $`E_N^{T,\pm }=2\psi (N+3)+{\displaystyle \frac{12}{N+3}}{\displaystyle \frac{1}{2}}+2\gamma __\mathrm{E},`$ $`E_N^S=2\psi (N+3)+{\displaystyle \frac{1}{N+3}}{\displaystyle \frac{1}{2}}+2\gamma __\mathrm{E}.`$ (22) We will often omit the subscript “$`q=q_0`$” for these special states. Note useful relations: $$2E_N^S=E_N^{T,+}+E_N^{T,},2\mathrm{\Phi }_N^{S^\pm }(x_i)=\mathrm{\Phi }_N^{T,+}(x_i)\pm \mathrm{\Phi }_N^{T,}(x_i),\mathrm{\Phi }_N^{T,}|\mathrm{\Phi }_N^{T,+}=0.$$ (23) Since the wave functions of eigenstates with different energies are mutually orthogonal, Eq. (21) implies that the single term $`q=q_0`$ survives in the sum (17) in the leading large $`N_c`$ limit. As a consequence, the moments of twist-3 parton distributions are entirely given by the reduced matrix elements of the corresponding local operators $`𝒪_{N,q_0}`$, Eq. (13) : $`{\displaystyle _1^1}𝑑xx^{N+2}e(x,Q^2)`$ $`=`$ $`𝒪_N^{T,+}(Q^2),`$ $`(N+4){\displaystyle _1^1}𝑑xx^{N+2}\stackrel{~}{h}_L(x,Q^2)`$ $`=`$ $`𝒪_N^{T,}(Q^2),`$ (24) $`{\displaystyle \frac{N+3}{N+2}}{\displaystyle _1^1}𝑑xx^{N+2}\stackrel{~}{g}_2(x,Q^2)`$ $`=`$ $`𝒪_N^{S^+}(Q^2)+𝒪_N^S^{}(Q^2).`$ Combined with the scale dependence of matrix elements, Eq. (14), these relations are equivalent to the DGLAP evolution equations in (2) with $`N_cE_N^f=_0^1𝑑zz^{N+2}P_f^{(0)}(z)`$. Treating the $`1/N_c`$ contribution to the Hamiltonian in (5) as a perturbation, we expand $`_{N,q}`$ $`=`$ $`N_cE_{N,q}+N_c^1\delta E_{N,q}+\mathrm{},`$ (25) $`\mathrm{\Psi }_{N,q}`$ $`=`$ $`\mathrm{\Psi }_{N,q}^{(0)}+N_c^2\delta \mathrm{\Psi }_{N,q}+\mathrm{},`$ with the usual quantum-mechanical expressions $`\delta E_{N,q}`$ $`=`$ $`2\mathrm{\Psi }_{N,q}^{(0)}^2\mathrm{\Psi }_{N,q}^{(0)}|^{(1)}|\mathrm{\Psi }_{N,q}^{(0)},`$ (26) $`\delta \mathrm{\Psi }_{N,q}(x_i)`$ $`=`$ $`2{\displaystyle \underset{q^{}q}{}}{\displaystyle \frac{\mathrm{\Psi }_{N,q}^{(0)}|^{(1)}|\mathrm{\Psi }_{N,q^{}}^{(0)}}{\mathrm{\Psi }_{N,q^{}}^{(0)}^2}}{\displaystyle \frac{\mathrm{\Psi }_{N,q^{}}^{(0)}(x_i)}{E_{N,q}E_{N,q^{}}}}.`$ (27) To this accuracy, moments of the twist-3 distributions are no longer renormalized multiplicatively and have to be expanded in contributions of multiplicatively renormalizable operators, Eq. (17): $$c_f(N)_1^1𝑑xx^{N+2}f(x,Q^2)=𝒪_{N,q_0}(Q^2)\frac{2}{N_c^2}\underset{qq_0}{}\frac{\mathrm{\Psi }_{N,q}^{(0)}|^{(1)}|\mathrm{\Phi }_N}{\mathrm{\Psi }_{N,q}^{(0)}^2}\frac{𝒪_{N,q}(Q^2)}{E_{N,q}E_{N,q_0}},$$ (28) where $`c_e=1`$, $`c_{h_L}=N+4`$, $`c_{g_2}=(N+3)/(N+2)`$, $`\mathrm{\Phi }_N\mathrm{\Psi }_{N,q_0}^{(0)}`$ is one of the functions $`\mathrm{\Phi }_N^{S^+}`$, $`\mathrm{\Phi }_N^{T,+}`$, $`\mathrm{\Phi }_N^{T,}`$. $`𝒪_{N,q_0}(Q^2)`$ is the quark-gluon-antiquark operator with the lowest anomalous dimension corresponding to the wave function $`\mathrm{\Psi }_{N,q_0}=\mathrm{\Psi }_{N,q_0}^{(0)}+N_c^2\delta \mathrm{\Psi }_{N,q_0}`$ and normalized at the scale $`Q^2`$. Thus, the calculation of $`1/N_c^2`$ corrections to the moments of the structure functions consists in two separate tasks. First, one has to calculate the $`𝒪(1/N_c^2)`$ corrections to the anomalous dimension of $`𝒪_{N,q_0}`$, or equivalently to the energies (22) of the lowest eigenstates and, second, calculate the mixing coefficients $`\mathrm{\Psi }_{N,q}^{(0)}|^{(1)}|\mathrm{\Psi }_{N,q_0}^{(0)}`$ with higher levels; evolution of the latter can be taken into account in the leading large-$`N_c`$ approximation of Refs. . We address both questions in what follows. 4. The $`𝒪(1/N_c^2)`$ correction $`\delta E_N`$ to the energy of the lowest eigenstates is given by Eq. (26) for $`q=q_0`$. The calculation is most easily done by going over to the so-called conformal basis . The idea is to define a basis of (orthogonal with respect to (15)) polynomials that satisfy the conformal constraints in (12) and, in addition, diagonalize the two-particle Casimir operator (8) in one of the channels, e.g. $$L_{12}^2Y_{N,n}^{(12)3}(x_1,x_2,x_3)=(n+5/2)(n+3/2)Y_{N,n}^{(12)3}(x_1,x_2,x_3)$$ (29) and, similarly, $`Y_{N,n}^{1(23)}`$ and $`Y_{N,n}^{(13)2}`$. Here, the superscripts indicate the order in which the conformal spins of the three partons sum up to the total conformal spin $`N+7/2`$. The three different sets of $`Y`$-functions are related to each other through the Racah $`6j`$-symbols of the $`SL(2)`$ group. Permutation symmetry between the quarks implies that $`Y_{N,n}^{(23)1}(x_1,x_2,x_3)=(1)^nY_{N,n}^{(12)3}(x_3,x_2,x_1)`$. The explicit expressions for the $`Y`$functions are given in terms of the Jacobi polynomials . The expansion coefficients of the lowest eigenstates (18) in each of the three conformal basis are easily obtained $`\mathrm{\Phi }_N^{T,\pm }`$ $`=`$ $`{\displaystyle \underset{n=0}{\overset{N}{}}}(1)^n(N+n+5)\left[{\displaystyle \frac{N+3}{n+1}}\pm (n+3)\right]Y_{N,n}^{(12)3}`$ $`=`$ $`{\displaystyle \underset{n=0}{\overset{N}{}}}\left[(1)^{Nn}\pm 1\right](2n+3)(n+2){\displaystyle \frac{2(N+n)+9\pm 1}{2(Nn)+31}}Y_{N,n}^{(31)2}`$ and $`\mathrm{\Phi }_N^{S^+}=(\mathrm{\Phi }_N^{T,+}+\mathrm{\Phi }_N^{T,})/2`$. The normalization is $`Y_{N,n}^{(12)3}^2`$ $`=`$ $`{\displaystyle \frac{(n+1)(Nn+1)}{(n+2)(n+3)(N+n+5)}},`$ $`Y_{N,n}^{(31)2}^2`$ $`=`$ $`{\displaystyle \frac{2(n+1)(Nn+1)(Nn+2)}{(n+2)(2n+3)(N+n+4)(N+n+5)}}.`$ (31) Notice that the two-particle kernels defined in (7) become diagonal in the corresponding basis. Expanding each contribution in a suitable basis, one obtains the matrix elements, $`\mathrm{\Psi }_N^{T,\pm }|_T^{(1)}|\mathrm{\Psi }_N^{T,\pm }`$ and $`\mathrm{\Psi }_N^{S^+}|_{S^+}^{(1)}|\mathrm{\Psi }_N^{S^+}`$, as finite sums over two-particle spins $`n`$. Full expressions are rather cumbersome and will be presented elsewhere . Expanding the resulting expressions for $`\delta E_N`$ in powers of $`1/(N+3)`$ we obtain $`\delta E_N^S`$ $`=`$ $`2\left(\mathrm{ln}(N+3)+\gamma __\mathrm{E}+{\displaystyle \frac{3}{4}}{\displaystyle \frac{\pi ^2}{6}}\right)+𝒪\left({\displaystyle \frac{\mathrm{ln}^2(N+3)}{(N+3)^2}}\right),`$ (32) $`\delta E_N^{T,\pm }`$ $`=`$ $`\delta E_N^S\pm {\displaystyle \frac{4}{N+3}}\left[\left(3{\displaystyle \frac{\pi ^2}{3}}\right)\left(\mathrm{ln}(N+3)+\gamma __\mathrm{E}\right){\displaystyle \frac{5}{2}}+{\displaystyle \frac{\pi ^2}{3}}\right]+𝒪\left({\displaystyle \frac{\mathrm{ln}^2(N+3)}{(N+3)^2}}\right).`$ With this correction, Eq. (25) gives an excellent description of the lowest anomalous dimension in the spectrum of twist-3 operators for all integer $`N0`$, see Table 1. Note that to this accuracy $`_N^{T,+}+_N^{T,}=2_N^S`$, cf. Eq. (23). <sup>§</sup>The difference of these values with the exact result is entirely due to the truncation of the $`1/(N+3)`$expansion in (32). Since only one independent operator exists in these three cases, there is no mixing and the $`𝒪(1/N_c)`$ approximation to the energies $`_{N=0}^S`$, $`_{N=0}^{T,+}`$ and $`_{N=1}^{T,}`$ is in fact exact. The following comments are in order. First, we note that $`\delta E_N`$ has the same large-$`N`$ behavior, $`\delta E_N\mathrm{ln}N`$, as the leading large-$`N_c`$ result in (22). The coefficient in front of the $`\mathrm{ln}N`$ term at large $`N`$ is redefined, therefore, from $`2N_c`$ to $`4C_F=2(N_c^21)/N_c`$, in agreement with . It is possible to show that this coefficient is exact and higher order $`1/N_c^2`$ corrections do not grow with $`N`$, $`_{N,q_0}=2C_F\mathrm{ln}N+𝒪(N^0)`$. Second, the constant $`𝒪(N^0)`$ term in the large $`N`$ expansion of the anomalous dimensions (32) does not agree with . The difference is due to the contribution of the operators $`V_{qg}^{(1)}`$ and $`U_{qg}^{(1)}`$ in (S0.Ex3) that was overlooked in the previous works. The result for the $`𝒪(1/N)`$ correction is new. Finally, from the expressions in (32) it is easy to read out the corresponding modification of the DGLAP splitting functions (2), $`P_f(z)=P_f^{(0)}(z)+P_f^{(1)}(z)`$. For example $$P_{g_2}^{(1)}(z)\stackrel{z1}{=}\frac{2}{N_c}\left\{\left[\frac{1}{1z}\right]_++\left(\frac{3}{4}\frac{\pi ^2}{6}\right)\delta (1z)+\frac{1}{2}+𝒪(1z)\right\}.$$ (33) This result is exact to the $`𝒪(1/N_c^4)`$ accuracy and neglecting all terms that vanish at $`z1`$. Expressions for $`P_e^{(1)}(z)`$ and $`P_{h_L}^{(1)}(z)`$ have similar structure. The opposite limit of the small$`z`$ behavior of the DGLAP splitting functions $`P_f(z)`$ is of special interest. It is well known that this behavior is governed by singularities of the anomalous dimensions $`_N`$ in the complex $`N`$-plane. In particular, the leading large-$`N_c`$ anomalous dimensions (22) have simple poles at negative integer $`N3`$. The pole at $`N=3`$ corresponds to $`P^{(0)}(z)\stackrel{z0}{}z^0`$ in (2). With this connection in mind, we have studied the analytic structure of $`1/N_c^2`$ corrections to the anomalous dimensions, $`\delta E_N`$, continued analytically to the complex $`N`$ plane. The calculation is straightforward, albeit tedious . We find that the analytic continuation has to be performed separately for even and odd $`N`$, as familiar from studies of the evolution of twist-2 parton distribution beyond the leading order . Similarly to the latter case, we defineThis corresponds to the decomposition over partial waves with definite signature. $$_1^1𝑑xx^Nf(x)=\frac{1+(1)^N}{2}_0^1𝑑xx^Nf_{\mathrm{even}}(x)+\frac{1(1)^N}{2}_0^1𝑑xx^Nf_{\mathrm{odd}}(x)$$ (34) and consider the evolution of even- and odd-signature component of $`f=e,\stackrel{~}{h}_L,\stackrel{~}{g}_2`$ separately. When the restriction to even (odd) $`N`$ is imposed, the normalized matrix elements $`\mathrm{\Phi }_N|^{(1)}|\mathrm{\Phi }_N/\mathrm{\Phi }_N^2`$ develop singularities at, generically, $`N=0,1,2`$. With one exception, all singularities to the right of $`N=3`$ are simple poles and appear due to vanishing of the norm of the leading large-$`N_c`$ wave functions $`\mathrm{\Phi }_N^2`$. They give rise to small-$`z`$ behavior of the $`1/N_c^2`$ correction to the DGLAP splitting functions of the form $`P_{g_2^{\mathrm{even}}}^{(1)}(z)`$ $`\stackrel{z0}{=}`$ $`{\displaystyle \frac{1}{N_c}}\left\{5{\displaystyle \frac{172\pi ^2}{514\pi ^2}}{\displaystyle \frac{1}{z^2}}+6+𝒪(z)\right\},`$ $`P_{g_2^{\mathrm{odd}}}^{(1)}(z)`$ $`\stackrel{z0}{=}`$ $`{\displaystyle \frac{1}{N_c}}\left\{{\displaystyle \frac{6}{6\pi ^2}}{\displaystyle \frac{\mathrm{ln}z}{z}}+{\displaystyle \frac{\pi ^418\pi ^2+54+72\zeta (3)}{2(6\pi ^2)^2}}{\displaystyle \frac{1}{z}}+𝒪(z)\right\}.`$ (35) This asymptotics is more singular than that of $`P_{g_2}^{(0)}(z)`$ and is in apparent contradiction with the Regge theory expectations. The origin of these “spurious” singularities and their physical significance is not clear and deserves a special study that goes beyond the tasks of this letter. We would like to stress that appearance of singularities of the anomalous dimensions to the right from $`N=3`$ can well be artifact of the $`1/N_c`$ expansion, since close to the singularities the $`1/N_c^2`$ correction dominates over leading order term in (25) and the $`1/N_c^2`$ expansion formally breaks down. Another delicate point is that analytic continuation in $`N`$ has to be done in (28) simultaneously with the analytic continuation of the sum over anomalous dimensions $`_{N,q_{\mathrm{}}}`$ belonging to different “trajectories” parameterized by integer $`\mathrm{}`$ . To the best of our knowledge, the problem of analytic continuation from a set of discrete points $`_{N,q}`$ on a $`(N,q)`$plane has never been addressed in mathematical literature. To illustrate that the problem is nontrivial, consider the trace of the (full) Hamiltonian (5) over the subspace spanned by wave functions with given $`N0`$ $$\mathrm{Tr}_N=\underset{\mathrm{}=0}{\overset{N}{}}_{N,q_{\mathrm{}}}.$$ (36) It is equal to the sum of all $`N+1`$ anomalous dimensions and is easily calculated exactly as a sum of diagonal matrix elements in any suitable basis. We obtain, in particular $`\mathrm{Tr}_N_{S^+}^{(0)}`$ $`=`$ $`2(2N+5)\left[\psi (N+3)+\gamma __\mathrm{E}\right]{\displaystyle \frac{11}{2}}N13+{\displaystyle \frac{1}{N+2}}+{\displaystyle \frac{1}{N+3}},`$ $`\mathrm{Tr}_N_{S^+}^{(1)}`$ $`=`$ $`(N+2)\left[\psi (N+3)+\gamma __\mathrm{E}\right]{\displaystyle \frac{7}{4}}N5+{\displaystyle \frac{1}{2(N+2)}}+{\displaystyle \frac{5}{2}}\mathrm{ln}2`$ $`+(1)^N\left\{{\displaystyle \frac{5}{4}}\left[\psi \left({\displaystyle \frac{N}{2}}+{\displaystyle \frac{3}{2}}\right)\psi \left({\displaystyle \frac{N}{2}}+2\right)\right]+{\displaystyle \frac{1}{2(N+2)}}+{\displaystyle \frac{1}{2(N+3)}}\right\}.`$ Notice now that the trace of already the leading large-$`N_c`$ Hamiltonian $`^{(0)}`$ is singular at $`N=2`$. This implies that either one (or several) trajectories of the anomalous dimensions, $`_{N,q_{\mathrm{}}}`$, becomes singular at this point, or the (analytically continued) sum over the trajectories in (28) diverges. In both cases, analytic continuation of the lowest trajectory (22) beyond $`N=2`$ is questionable and, therefore, asymptotic expressions in (35) (or, equivalently, the DGLAP-evolution for $`z<1/N_c^2`$) have to be taken with caution. 5. According to (28), the moments of twist-3 distributions receive contributions from the whole tower of operators with non-minimal anomalous dimensions at the level of $`𝒪(1/N_c^2)`$ corrections. Aside from the “energy denominators”, the coefficients in front of these operators in (28) are given by (normalized) matrix elements $`\mathrm{}^{}|^{(1)}|\mathrm{}=0`$ of the perturbation $`^{(1)}`$ $$\mathrm{}^{}|^{(1)}|\mathrm{}=\frac{\mathrm{\Psi }_{N,q_{\mathrm{}^{}}}^{(0)}|^{(1)}|\mathrm{\Psi }_{N,q_{\mathrm{}}}^{(0)}}{\mathrm{\Psi }_{N,q_{\mathrm{}^{}}}^{(0)}\mathrm{\Psi }_{N,q_{\mathrm{}}}^{(0)}},$$ (38) where $`\mathrm{},\mathrm{}^{}=0,1,\mathrm{},N`$ numerate the anomalous dimensions from below. Our main observation is that the matrix elements (38) are concentrated close to the diagonal $`\mathrm{}=\mathrm{}^{}`$, see Figs. 1, 2 and Table 2. In addition, the lowest anomalous dimensions $`E_{N,q_0}E_N`$ (22) are separated from the rest of the spectrum by a finite gap $`E_{N,q_1}E_N=0.227`$ as $`N\mathrm{}`$ . These two properties guarantee that the expansion in (28) is rapidly converging and remains well defined in the large-$`N`$ limit (when the number of eigenstates diverges). The general structure for the cases of $`_{S^+}`$ and $`_T`$ is very similar; we present the results for $`S^+`$ as related to the structure function $`g_2(x)`$ and therefore being of more direct phenomenological significance. At large $`N`$ the matrix elements $`\mathrm{}|^{(1)}|0`$ can be calculated semi-classically using the approach developed in . It turns out that the matrix elements $`|\mathrm{}|^{(1)}|0|`$ approach maximal value $`1/\sqrt{\mathrm{ln}N}`$ at $`\mathrm{}\mathrm{ln}N`$, while for $`\mathrm{}\mathrm{ln}N`$ they rapidly decrease as $`1/\mathrm{}^2`$. As the result, the sum in (28) is effectively cut off at $`\mathrm{}\mathrm{ln}N`$ terms. For example, for $`N100`$ only $`5`$ first (lowest) levels give a sizeable contribution to the r.h.s. of (28), cf. Fig. 2. Analytic expressions for the matrix elements $`\mathrm{}|^{(1)}|0`$ in the large $`N`$ limit are complicated and will be given elsewhere . An independent argument in favor of the dominance of the operators with lowest anomalous dimensions comes from the structure of matrix elements $`𝒪_{N,q}`$ in the large-$`N`$ limit. In an analogy with twist-2 distributions, we can write the matrix elements $`𝒪_{N,q}`$ as (generalized) moments of nonperturbative (chiral-odd or chiral-even) distribution function $`D_f(x_1,x_2,x_3)`$ describing quark-gluon-antiquark correlations in the nucleon, schematically<sup>\**</sup><sup>\**</sup>\**see Appendix A in for exact expressions. $$𝒪_{N,q}=_1^1[dx]\mathrm{\Psi }_{N,q}(x_1,x_2,x_3)D_f(x_1,x_2,x_3),$$ (39) where $`[dx]=dx_1dx_2dx_3\delta (x_1+x_2+x_3)`$ and the variables $`x_i`$ have the meaning of the quark, antiquark and gluon momentum fractions, $`1x_i1`$. At large $`N`$, the eigenfunctions $`\mathrm{\Psi }_{N,q}(x_i)`$ are sharply peaked at the boundary of the integration region $`x_1+x_2+x_3=0`$ so that the matrix elements (39) probe the behavior of the quark-gluon distribution functions $`D(x_i)`$ near the kinematical boundaries. The precise position of the peak depends on $`q`$. It turns out that the eigenfunctions of the low-lying states are peaked at $`x_1=x_3=\pm 1,x_2=0`$ corresponding to configurations when the quark and the antiquark carry all the momentum of nucleon and the gluon is soft. For higher states the position of the peak is shifted gradually to $`2x_1=2x_3=x_2=\pm 1`$ corresponding to a hard gluon and (relatively) soft quark and antiquark. If one assumes that the nonperturbative distributions are generated by gluon radiation already at low scales, the configurations with soft gluons should be enhanced and those with hard gluons suppressed. We conclude that the contribution of the conformal operators with large anomalous dimensions to the moments of the structure functions, (28), is suppressed because of smallness of both perturbative mixing coefficients and the corresponding nonperturbative matrix elements as describing rare parton configurations in nucleon involving a hard gluon. To summarize, we have argued that (up to $`𝒪(1/N_c^4)`$ corrections) the contribution of multiplicatively renormalizable operators with higher anomalous dimensions to the moments of twist-3 structure functions $`e(x)`$, $`\stackrel{~}{h}_L`$ and $`\stackrel{~}{g}_2(x)`$ is small as compared with that of the ground state operator. With the increasing precision of the experimental data this contribution can be estimated in a “two-channel” approximation in which one supplements the dominant ground state quark-gluon component with one extra effective partonic component in order to account for the admixture of $`\mathrm{ln}N`$ lowest eigenstates. Acknowledgements. Work by A.M. was partially supported by the DFG.
warning/0001/hep-ph0001179.html
ar5iv
text
# Updated Bounds on Milli-Charged Particles ## 1 Introduction A puzzle of continuing interest in particle physics is the apparent quantization of electric charge. The electric charge of all the particles we see appears to be an integer multiple of $`Q_e/3`$, where $`Q_e`$ is the electron charge. However it is theoretically consistent to have particles with electric charge $`ϵQ_e`$ where $`ϵ`$ is any real number. We consider $`ϵ<1`$ in this paper, and refer to these particles as “milli-charged.” Milli-charged particles can be introduced into the Standard Model in a variety of ways. One is to add an $`SU(3)\times SU(2)`$ singlet Dirac fermion with hypercharge $`Y=2ϵ`$. However, this is not simple if the hypercharge $`U(1)_Y`$ is embedded in a grand unified gauge group. A second way of generating effectively milli-charged particles, which works even if hypercharge is quantised, is due to Holdom . He introduced a second unbroken “mirror” $`U(1)^{}`$, and showed that the photon and the “paraphoton” can mix, so that a particle charged under the $`U(1)^{}`$ appears to have a small coupling to the photon. A third mechanism for introducing milli-charged particles is to remain with the Standard Model particle content and allow the neutrinos to have small electric charges . If the Standard Model hypercharge operator $`Y_{SM}`$ is redefined to be $`Y^{}=Y_{SM}+2_iϵ_i(B/3L_i)`$, neutrinos acquire small electric charges $`ϵ_i`$ and the Standard Model anomaly cancellation is preserved. We assume here that $`U(1)_{em}`$ is an unbroken symmetry; it can be spontaneously broken in models with milli-charged scalars , but the bounds in this case are slightly different from what we discuss here (for instance the photon has a mass). The purpose of this paper is to collect and update experimental and observational bounds on milli-charged particles. Such limits have been previously considered for milli-charged fermions that are not neutrinos and for milli-charged neutrinos . See also for a discussion of astrophysical constraints on all types of milli-charged particles. We recalculate the laboratory bounds using recent data, and include the limits from recent experiments searching for milli-charged particles . We numerically evaluate the constraint from Big Bang Nucleosynthesis on particles with milli-hypercharge, and find a bound similar to previous estimates. We also revisit the astrophysical constraints, resolving a discrepancy in the White Dwarf bounds between and (in favour of the stronger bound of ). Our Supernova limit differs slightly from what is found in Ref. , as will be explained later, and the Red Giant bound remains unchanged. We quote the previous estimated bounds from balloon experiments searching for strongly interacting Dark Matter, and from underground WIMP detectors. For a more detailed discussion of the calculations, see for instance . Our conclusions are presented in Fig. 1 which is an update of Fig. 1 of Ref. . It presents bounds on milli-charged fermions in models with an extra $`U(1)^{}`$, and without. Milli-charged neutrinos will be briefly discussed at the end of the paper. The limits from laboratory experiments, from Red Giants and White Dwarfs are independent of whether there is a paraphoton, and appear as solid lines. We discuss the Supernova and nucleosynthesis bounds with some care for the model without a paraphoton; the limits should be similar in the presence of a paraphoton so these bounds also appear as solid lines. The limit from $`\mathrm{\Omega }<1`$ differs in the two models: in the absence of a paraphoton, the millicharge annihilation cross-section is proportional to $`ϵ^2/m^2`$. So the excluded area, indicated with dotted lines, extends to lower masses for smaller $`ϵ`$. The bound on models with a paraphoton is the horizontal dashed line. ## 2 Laboratory Bounds We plot four lines on Figure 1 from laboratory data: 1) a combined “accelerator” line consisting of the limits from LEP and beam dump experiments <sup>1</sup><sup>1</sup>1 The values of $`m_ϵ`$ excluded by direct searches are quoted in Table II of , but the values given are in MeV, not GeV as claimed in the caption. The limit is placed correctly, however, in Fig. 2 of ., 2) the bound found by the Tokyo group from the non-observation of invisible Ortho-positronium decay, 3) the limit from the dedicated milli-charged particle search experiment at SLAC , and 4) an updated constraint from the Lamb shift using more recent data . LEP has taken many years of data since limits from LEP were previously considered , so we briefly discuss possible bounds, although these are weak because the milli-charge coupling to the $`Z`$ is suppressed by $`\mathrm{sin}^2\theta _W`$. A search for particles with fractional charge $`ϵ=2/3`$ was performed by OPAL using 1991-93 data , which rules out $`ϵ2/3`$ for $`m_ϵ<84`$ GeV. This bound could be extended to the present kinematic limit $`m_ϵ<100`$ GeV, if one assumes that a particle with $`1>ϵ>2/3`$ would be seen as such in the detector. Fractionally charged particles would contribute to the invisible width of the $`Z`$ if they were not seen in the detector <sup>2</sup><sup>2</sup>2This bound assumes that particles with 1/4 $`<ϵ<2/3`$ would look like noise in the detector, and not be mis-identified as $`ϵ=1`$ tracks., in which case the LEP bound can be extended to $$ϵ<0.24m_ϵ>45\mathrm{GeV},$$ (1) from requiring that that milli-charges not contribute more than the $`2\sigma `$ error to the invisible width of the $`Z`$ at LEP1. We calculate an improved bound from the Lamb shift measurements by requiring that the milli-charged particle vacuum polarisation contribution be less than the $`2\sigma `$ error ( = 20 $`\times 10^3`$ MHz) . This is a more recent measurements of the $`2S_{1/2}2P_{1/2}`$ Lamb shift than used in . There are more precise measurements of combinations of Lamb shifts of different principle quantum number $`n`$ ($`n2`$) ; however they do not substantially improve our bound so we use the $`2S_{1/2}2P_{1/2}`$ results because the milli-charged contribution is easier to disentangle. The bound from the Lamb shift is stronger than from $`g2`$, despite the fact that $`g2`$ is measured very precisely , because milli-charged particles contribute to the Lamb shift at one loop but to $`g2`$ at two loop. For a discussion of the effects of milli-charged particles on precision QED measurements, see . ## 3 Big-Bang Nucleosynthesis ### 3.1 Dominant Processes Particles with small electric charge will interact with the plasma in the early universe and in this way get thermally excited. Therefore, their primordial energy density and thus their interaction strength can be constrained by the standard nucleosynthesis limit on $`N_{\mathrm{eff}}`$, the effective number of thermally excited neutrino degrees of freedom. A milli-charged fermion must be a Dirac particle and thus has four intrinsic degrees of freedom. If it were fully excited in the early universe, it would contribute $`\mathrm{\Delta }N_{\mathrm{eff}}=2`$ prior to the annihilation of electrons and positrons <sup>3</sup><sup>3</sup>3This is when the BBN bound on the number of neutrinos applies. Such a particle would be excluded by a large margin. During this annihilation process entropy is transferred to the milli-charged particles, heating them relative to the neutrinos. At the same time, the photon temperature, $`T_\gamma `$, relative to the neutrino temperature, $`T_\nu `$, is less than in the standard model. Normally, $`T_\nu /T_\gamma 0.714`$, whereas here $`T_\nu /T_\gamma 0.849`$. Altogether, neutrinos plus milli-charged particles contribute $`N_{\mathrm{eff}}=3\times (0.849/0.714)^4+2\times 1.401^4=13.7`$, so that $`\mathrm{\Delta }N_{\mathrm{eff}}=10.7`$ at late times. While it is easy to estimate $`\mathrm{\Delta }N_{\mathrm{eff}}`$ for milli-charged particles where full thermal equilibrium obtains, we can derive a more restrictive limit by studying its value as a function of $`ϵ`$ in a regime where full equilibrium is not achieved. To this end we assume that the new particles interact only electromagnetically by their milli-charge. The bounds in the case where there is an extra $`U(1)`$ are at least as restrictive, due to the presence of the paraphoton . We find that the dominant electromagnetic processes which couple the gas of milli-charged particles $`f`$ to the electromagnetic plasma are $`e^+e^{}`$ $``$ $`f\overline{f},`$ $`\gamma `$ $``$ $`f\overline{f},`$ $`ef`$ $``$ $`ef.`$ (2) The rates for these reactions are proportional to $`ϵ^2`$ while those for other conceivable processes such as $`\gamma \gamma f\overline{f}`$ or $`\gamma f\gamma f`$ are proportional to $`ϵ^4`$ and thus much smaller. At low temperatures, the electron-positron number density is exponentially suppressed, making the $`f\overline{f}`$ production rate approach zero quickly when T < me < 𝑇subscript𝑚𝑒T\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}m_{e}. The third process in Eq. (3.1) does not contribute to the production of milli-charged particles but helps to achieve kinetic equilibrium. For a calculation of the $`f\overline{f}`$ production rates we use Boltzmann statistics for all species. Then the rate for the pair process is $`\mathrm{\Gamma }_{e^+e^{}f\overline{f}}=n_e\sigma |v|`$, where $`\sigma |v|`$ is the thermally averaged annihilation cross section. It can be expressed as $$\sigma |v|=\frac{1}{8m_e^4TK_2^2(m_e/T)}_{4m_e^2}^{\mathrm{}}𝑑ss^{1/2}(s4m_e^2)K_1(s^{1/2}/T)\sigma _{\mathrm{CM}},$$ (3) where $`K_n`$ is a modified Bessel function of the second kind and $`\sigma _{\mathrm{CM}}`$ is the center-of-mass cross section. The squared matrix element is $`{\displaystyle \frac{1}{4}}{\displaystyle |M|^2}`$ $`=`$ $`{\displaystyle \frac{8ϵ^2e^4}{(p_e^{}+p_{e^+})^4}}[(p_e^{}p_f)(p_{e^+}p_{\overline{f}})+(p_e^{}p_{\overline{f}})(p_{e^+}p_f)+m_f^2(p_e^{}p_{e^+})`$ (4) $`+m_e^2(p_fp_{\overline{f}})+2m_e^2m_f^2],`$ which translates into the usual CM cross section $`\sigma _{\mathrm{CM}}=(4\pi /3)ϵ^2\alpha ^2s^1`$ if all the particles can be considered massless. In this limit we find $$\mathrm{\Gamma }_{e^+e^{}f\overline{f}}=\frac{\zeta _3}{2\pi }ϵ^2\alpha ^2T,$$ (5) where $`\zeta `$ refers to the Riemann zeta function. The decay rate of a transverse plasmon (photon) with energy $`\omega `$ and momentum $`k`$ into massless milli-charged particles is $`\mathrm{\Gamma }_{\gamma f\overline{f}}=ϵ^2\alpha Z(\omega ^2k^2)/3\omega `$ . In a relativistic plasma, the plasma frequency is given by $`\omega _\mathrm{P}^2=(4\pi /9)\alpha T^2`$. Except for the low-energy part of the blackbody photon spectrum we have $`\omega \omega _\mathrm{P}`$, a limit where the dispersion relation is $`\omega ^2k^2=\frac{3}{2}\omega _\mathrm{P}^2`$ and the wave-function renormalization factor is $`Z=1`$ . With Boltzmann statistics we find $`\omega ^1=(2T)^1`$ so that finally the average plasmon decay rate is $$\mathrm{\Gamma }_{\gamma f\overline{f}}=\frac{\pi }{9}ϵ^2\alpha ^2T.$$ (6) Comparing the rates for the two processes we find $$\frac{\mathrm{\Gamma }_{e^+e^{}f\overline{f}}}{\mathrm{\Gamma }_{\gamma f\overline{f}}}=\frac{9\zeta _3}{2\pi ^2}=0.55,$$ (7) implying that the plasma process is actually somewhat more important than pair annihilation. Notice, however, that the plasma process is only important if m < ωP/2 < 𝑚subscript𝜔𝑃2m\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}\omega_{P}/2. During BBN, the plasma frequency is roughly given by $`\omega _P\sqrt{(4\pi /9)\alpha }T0.1T`$, so that at $`T=1`$ MeV, where BBN begins, the plasma process is only important for m < 50 < 𝑚50m\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}50 keV. Therefore the plasma process is important only for milli-charged particles with masses in the keV region, where astrophysical bounds from stellar evolution are much more stringent. Thus, it seems safe no neglect the plasma process in the BBN calculations. ### 3.2 Solving the Boltzmann Equation The standard procedure for calculating the evolution of number and energy density of a given species is to use the Boltzmann collision equation. It describes the time evolution of the single particle distribution function, $`f_1`$, of any given species. In the expanding universe, formally this equation can be written in our case as $$\frac{f_1}{t}Hp\frac{f_1}{p}=(C_{\mathrm{ann}}+C_{\mathrm{el}})[f_1],$$ (8) where $`H\dot{R}/R`$ is the Hubble expansion parameter. On the right-hand side, $`C_{\mathrm{ann}}`$ is the collision operator describing pair annihilation. $`C_{\mathrm{el}}`$ is the elastic-scattering term which includes the combined effect of scattering on electrons, positrons, and photons. The $`e^+e^{}`$ process is of the form $`1+23+4`$ so that the collision term can be written generically as $$C_{\mathrm{ann}}[f_1]=\frac{1}{2E_1}d^3\stackrel{~}{p}_2d^3\stackrel{~}{p}_3d^3\stackrel{~}{p}_4\mathrm{\Lambda }(f_1,f_2,f_3,f_4)|M|^2\delta ^4(p_1+p_2p_3p_4)(2\pi )^4.$$ (9) Here, $`d^3\stackrel{~}{p}d^3p/[(2\pi )^32E]`$, $`|M|^2`$ is the squared matrix element, and $`p_i`$ is the four-momentum of particle $`i`$. The phase-space factor is $`\mathrm{\Lambda }f_3f_4(1f_1)(1f_2)f_1f_2(1f_3)(1f_4)`$. The dominant elastic scattering term from scattering on electrons and positrons suffers from the usual infrared Coulomb divergence. It can be regulated by including Debye screening effects in the medium. However, instead of treating elastic scattering in any detail we calculate the evolution of the milli-charged particle ensemble for two extreme cases, a) No elastic scattering, and b) Elastic scattering is assumed to be efficient enough to bring the milli-charged species into complete kinetic equilibrium at all times. It will turn out that the difference between these two cases is quite small, justifying our neglect of a detailed treatment of elastic scattering. The dynamical evolution of the cosmic scale factor is governed by the Friedmann equation and the equation of energy conservation , $`H^2`$ $`=`$ $`{\displaystyle \frac{8\pi G\rho }{3}},`$ (10) $`{\displaystyle \frac{d}{dt}}(\rho R^3)`$ $`=`$ $`P{\displaystyle \frac{d}{dt}}(R^3),`$ (11) where $`\rho `$ is the energy density and $`P`$ the pressure, both including the effect of milli-charged particles. We have solved these equations to find the energy density of milli-charged particles. We have always neglected their mass, an approximation which is accurate for $`m_ϵ`$ up to the electron mass. For larger masses the milli-charged particles can pair annihilate into electron-positron pairs at low temperatures. We have always taken a zero initial population of milli-charged particles, an assumption which is valid if they interact only via their electric charge. Adding energy density to the cosmic plasma around the epoch of $`e^+e^{}`$ annihilation perturbs Big Bang Nucleosynthesis . In order to calculate constraints on $`ϵ`$ we have used the nucleosynthesis code of Kawano , modified to include the energy density in milli-charged particles as well as the change in neutrino-to-photon temperature ratio. The effect of the milli-charged particles corresponds to additional neutrino degrees of freedom of $$\mathrm{\Delta }N_{\mathrm{eff}}=0.69\times 10^{17}ϵ^2\times \{\begin{array}{cc}1\hfill & \text{no elastic scattering,}\hfill \\ 1.39\hfill & \text{full equilibrium,}\hfill \end{array}$$ (12) an approximation is valid for ϵ < 6×109 < italic-ϵ6superscript109\epsilon\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}6\times 10^{-9}. This is more than sufficient for our purposes since all values of $`ϵ`$ higher than this produce such a large $`\mathrm{\Delta }N_{\mathrm{eff}}`$ that they are excluded by a huge margin. As expected, the effect of elastic scattering on the energy density in milli-charged particles is quite small as this process does not produce additional particles. ### 3.3 BBN Limits There are still some unresolved issues regarding the observationally determined values of the primordial light-element abundances. For the past few years there have been two favoured solutions, namely the so-called High-Helium/Low-Deuterium and the Low-Helium/High-Deuterium solutions . There seems to be growing consensus that the first is the correct one. Nevertheless any bound derived from nucleosynthesis should be used with some caution since the question of primordial deuterium and helium abundances is not yet fully settled. In the present paper we shall use the data on primordial helium obtained by Izotov and Thuan and the deuterium data from Burles and Tytler $`Y_P`$ $`=`$ $`0.244\pm 0.002,`$ (13) $`D/H`$ $`=`$ $`(3.39\pm 0.25)\times 10^5,`$ (14) where errors are estimated $`1\sigma `$ uncertainties. These data give the High-He/Low-D solution and are completely consistent with standard BBN for a baryon-to-photon ratio of $`\eta =(5.1\pm 0.3)\times 10^{10}`$ . Used together, these data provide the constraint $$N_{\mathrm{eff}}=2.98\pm 0.33(90\%\mathrm{C}.\mathrm{L}.),$$ (15) on the effective number of neutrinos, thus tightly constraining any non-standard nucleosynthesis scenario. Together with the less restrictive case (no elastic scattering) of Eq. (12) BBN then implies a limit $$ϵ<2.1\times 10^9.$$ (16) This bound applies to masses in the regime mϵ < me < subscript𝑚italic-ϵsubscript𝑚𝑒m_{\epsilon}\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}m_{e}. While our result is very similar to what one finds from simple dimensional estimates, our calculation is quantitative and the errors are controlled. ## 4 Stellar Evolution ### 4.1 Globular Clusters New low-mass particles will be produced in the hot and dense medium in the interior of stars and subsequently escape. This new energy-loss channel leads to observational modifications of the standard course of stellar evolution and thus can be used to set limits on the particle’s interaction strength. For milli-charged particles, limits have been set from red giants , horizontal-branch (HB) stars , white dwarfs , and supernova (SN) 1987A . For small masses of the milli-charged particles, the most restrictive limits arise from HB stars and low-mass red giants in globular clusters. At the end of the main-sequence evolution of normal stars, the hydrogen in the inner part has been consumed, leaving the star with a core consisting mainly of helium. In low-mass stars, this helium core reaches degeneracy before it is hot and dense enough to be ignited. This process is very dependent on density and temperature, and even minor changes in these quantities produce observable changes in the brightness at the tip of the red-giant branch in globular clusters. Therefore, the core mass at helium ignition as implied by the color-magnitude diagrams of several globular clusters implies a limit on any new energy-loss channel. After helium ignition, the stars move to the horizontal branch where they burn helium in their core. A new energy-loss mechanism will lead to an accelerated consumption of nuclear fuel, shortening the helium-burning lifetime which can be “measured” by number counts of HB stars in globular clusters. When applied to a new energy-loss channel, usually one of these arguments is more restrictive . For example, a putative neutrino dipole moment will add to the efficiency of the plasmon decay process $`\gamma \nu \overline{\nu }`$ and thus enhance neutrino losses. The helium-ignition argument provides far more restrictive limits than the helium-burning lifetime argument because the plasmon decay is more effective in the degenerate red-giant core. On the other hand, axion losses by the Primakoff process are more effective in the nondegenerate cores of HB stars so that the helium-burning lifetime argument yields more restrictive limits. The emission of milli-charged particles is a special case in that both arguments yield comparable limits of about $$ϵ2\times 10^{14}.$$ (17) The reason is that the rate of the plasma decay process $`\gamma f\overline{f}`$ for milli-charged particles is proportional to $`\omega _\mathrm{P}^2`$, as opposed to the magnetic-dipole case ($`\omega _\mathrm{P}^4`$) or standard-model neutrino case ($`\omega _\mathrm{P}^6`$). The low power of the plasma frequency implies that the emission rate per unit mass is almost independent of density. This, in turn, implies that the core expansion caused by helium ignition leaves the energy-loss rate per unit mass nearly unchanged, while it is “switched off” for the dipole-moment case, or “switched on” for the axion case. An average value for the plasma frequency in the core of a globular-cluster star before helium ignition is $`\omega _\mathrm{P}10\mathrm{keV}`$ (in the center it is about twice that) while it is about 2 keV in the core of a HB star. Therefore, the helium-ignition argument constrains milli-charged particles with masses up to about 5 keV. Of course, this sort of argument applies only if the interaction strength is small enough that the milli-charged particles escape freely once produced in the stellar core; one can check that this is the case for ϵ < 108 < italic-ϵsuperscript108\epsilon\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}10^{-8} . For larger charges, the particles would contribute to the transfer of energy rather than carrying away energy directly. In order to avoid observable consequences, the efficiency of energy transfer would have to be less than that of photons, i.e. the mean free path would have to be very short. It is unlikely that there exists an allowed range of milli-charges ϵ < 1 < italic-ϵ1\epsilon\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}1 where all stars would be left unchanged. However, since laboratory limits and the BBN argument exclude values for $`ϵ`$ above $`10^8`$ anyway, a detailed discussion of the trapping limit is not warranted. ### 4.2 White Dwarfs The observed population of hot young white dwarfs is consistent with cooling by surface emission of photons and by volume emission of neutrinos produced by plasmon decay via Standard Model interactions . Blinnikov and Dunina-Barkovskaya set a bound on the neutrino magnetic moment, $`\mu _\nu <10^{11}\mu _B`$ (where $`\mu _B=e/2m_e`$), by requiring that the additional neutrino emission not cool the white dwarfs faster than observed. We can translate this bound into one on milli-charged particles. For neutrinos or neutrino-like particles coupling to the photon via a dipole moment, the energy loss rate can be written as $$Q_\mu \frac{\mu ^2}{2}\left(\frac{\omega _\mathrm{P}^2}{4\pi }\right)^2Q_2,$$ (18) whereas for milli-charged particles, it is given by $$Q_ϵϵ^2\alpha \frac{\omega _\mathrm{P}^2}{4\pi }Q_1,$$ (19) where $`Q_1/Q_2`$ is a factor of order unity . Here, we set $`Q_1=Q_2=1`$. One then obtains $$\frac{Q_ϵ}{Q_\mu }=2.09\frac{ϵ_{14}^2}{\mu _{12}^2}\left(\frac{\omega _\mathrm{P}}{10\mathrm{keV}}\right)^2,$$ (20) where $`ϵ_{14}10^{14}ϵ`$ and $`\mu _{12}10^{12}\mu /\mu _B`$. In order to calculate an average value of this ratio over the entire star, we need to perform an “emissivity-average” of the form $$\frac{Q_ϵ}{Q_\mu }=\frac{𝑑rr^2\frac{Q_ϵ}{Q_\mu }Q_\mu }{𝑑rr^2Q_\mu }.$$ (21) In order to do these integrals, we assume that the white dwarf is a polytrope of index $`n=3/2`$, so that $$\frac{Q_ϵ}{Q_\mu }=2.40\times 10^5\rho _{c,6}^1\frac{ϵ_{14}^2}{\mu _{12}^2}\frac{𝑑\xi \xi ^2\theta ^n}{𝑑\xi \xi ^2\theta ^{2n}},$$ (22) where $`r=\alpha \xi `$ and $`\rho =\rho _c\theta ^n`$ . For a white-dwarf mass of 0.7 solar masses, this gives $$\frac{Q_ϵ}{Q_\mu }=0.34\frac{ϵ_{14}^2}{\mu _{12}^2}.$$ (23) If we then demand that the emission rate due to milli-charges is not larger than for neutrino dipole moments, we find $$ϵ_{14}17.$$ (24) This bound is consistent with that found in Ref. . One may worry that the way the emissivity-average is performed has consequences for the bound. One could instead use $$\frac{Q_ϵ}{Q_\mu }=\frac{𝑑rr^2\frac{Q_ϵ}{Q_\mu }Q_ϵ}{𝑑rr^2Q_ϵ}.$$ (25) This yields $`ϵ_{14}15`$ so that the way the average is performed matters little for the final result. The emissivity-averaged plasma frequency is $`(10\mathrm{keV}/\omega _\mathrm{P})^2=0.16`$ so that $`\omega _\mathrm{P}=25\mathrm{keV}`$. Therefore, the white-dwarf bound applies to milli-charged particles with mϵ < 10 < subscript𝑚italic-ϵ10m_{\epsilon}\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}10 keV, similar to the red-giant case. ### 4.3 Supernova 1987A Finally, the stellar energy-loss argument can be applied to SN 1987A where the number of neutrinos detected at Earth agree roughly with theoretical expectations. If there are other particles contributing to the cooling of the proto neutron star, this will reduce the neutrino fluxes and the duration of the neutrino signal. Therefore, if we assume that such hypothetical particles freely stream from the core where they are produced, one can put an approximate bound on the allowed loss rate of Q/ρ < 1019ergg1s1, < delimited-⟨⟩𝑄𝜌superscript1019ergsuperscriptg1superscripts1\langle Q/\rho\rangle\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}10^{19}~{}{\rm erg~{}g^{-1}~{}s^{-1}}, (26) to be calculated at average core conditions of about $`3\times 10^{14}\mathrm{g}\mathrm{cm}^3`$ and 30 MeV for density and temperature, respectively. $`10^{19}\mathrm{erg}\mathrm{g}^1\mathrm{s}^1`$ is the proto-neutron star’s average rate of energy loss to neutrinos. The main production process for milli-charged particles in a SN core is the plasmon process. In a degenerate, relativistic electron plasma it is orders of magnitude larger than $`e^+e^{}`$ annihilation. The energy-loss rate is $$Q_\mathrm{P}=\frac{8\zeta _3}{9\pi ^3}ϵ^2\alpha ^2\left(\mu _e^2+\frac{\pi ^2T^2}{3}\right)T^3Q_1,$$ (27) where $`Q_1`$ is again a factor of order unity. Setting $`Q_1=1`$, one finds $$ϵ1\times 10^9$$ (28) This bound matches the one found by Mohapatra and Rothstein . However, they considered nucleon-nucleon bremsstrahlung as a production process and notably an amplitude where the electromagnetic current is coupled to an intermediate charged pion. We believe that a naive perturbative calculation of this process in a nuclear medium can be unreliable . However, since the bounds are so similar, a detailed study of the nucleon process is not warranted. If $`ϵ`$ is much larger than our limit, the milli-charged particles are trapped inside the proto neutron star, and can only escape via diffusion. The main process which keeps milli-charged particles trapped is Coulomb scattering on protons <sup>4</sup><sup>4</sup>4The scattering on electrons was instead considered in , but this is less important than protons.. The differential scattering cross section on nonrelativistic protons is $$\frac{d\sigma }{d\mathrm{\Omega }}=2ϵ^2\alpha ^2\frac{E^2(1+\mathrm{cos}\theta )}{|𝐪|^4}$$ (29) where $`𝐪`$, with $`|𝐪|^2=2E^2(1\mathrm{cos}\theta )`$, is the momentum transfer, $`E`$ the milli-charged particle’s energy, and $`\theta `$ the scattering angle. This cross section is strongly forward peaked and thus not a good measure for particle trapping: a particle which is deflected by a small angle continues its way out of the star essentially as if it had not been scattered at all. Therefore, we rather consider the usual transport cross section which includes an additional weight $`(1\mathrm{cos}\theta )`$. Therefore, $$\frac{d\sigma _\mathrm{T}}{d\mathrm{\Omega }}=ϵ^2\alpha ^2\frac{(1+\mathrm{cos}\theta )}{𝐪^2}$$ (30) is a more adequate measure for the effectiveness of particle trapping. In addition, we need to include proton-proton correlations induced by their mutual Coulomb repulsion, i.e. we need to include screening effects. This is achieved by multiplying the cross section with the static structure function $`S(𝐪)=𝐪^2/(𝐪^2+k_\mathrm{S}^2)`$ so that finally $$\frac{d\sigma _{\mathrm{T},\mathrm{eff}}}{d\mathrm{\Omega }}=ϵ^2\alpha ^2\frac{(1+\mathrm{cos}\theta )}{𝐪^2+k_\mathrm{S}^2}.$$ (31) For nonrelativistic, nondegenerate protons, the screening scale $`k_\mathrm{S}^2`$ is given by the proton Debye scale $`k_\mathrm{S}^2=k_\mathrm{D}^2=4\pi \alpha n_p/T`$ with $`n_p`$ the proton density. It would have been incorrect to use the electron screening scale since the background of degenerate electrons is much “stiffer” than the protons, i.e. most of the polarization of the plasma by a test charge is due to the protons. With these modifications we find a total cross section $$\sigma _{\mathrm{T},\mathrm{eff}}=\frac{2\pi ϵ^2\alpha ^2}{E^2}\left[\frac{(2+z)}{2}\mathrm{ln}\left(\frac{2+z}{z}\right)1\right],$$ (32) where $$z\frac{k_\mathrm{D}^2}{2E^2}=\frac{2\alpha }{3\pi }\eta _e^3\left(\frac{T}{E}\right)^2=0.335\left(\frac{\eta _e}{6}\right)^3\left(\frac{T}{E}\right)^2$$ (33) and $`\eta _e=\mu _e/T`$ is the electron degeneracy parameter. We have used that $`n_p=n_e=\mu _e^3/3\pi ^2`$. The transport mean free path is then found to be $$\lambda _{\mathrm{T},\mathrm{eff}}^1\sigma _{\mathrm{T},\mathrm{eff}}n_p=ϵ^2\alpha Tz\left[\frac{(2+z)}{2}\mathrm{ln}\left(\frac{2+z}{z}\right)1\right],$$ (34) where we have expressed $`n_p`$ in terms of $`k_\mathrm{D}`$. Taking $`T=30\mathrm{MeV}`$ as a typical value, and using 1 for the $`z`$-dependent expression, we find that $`\lambda _{\mathrm{T},\mathrm{eff}}`$ exceeds the proto-neutron star radius of about 10 km for ϵ < 1×108 < italic-ϵ1superscript108\epsilon\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}1\times 10^{-8}. When $`ϵ`$ is larger than this, the particles no longer freely escape. Rather, a “photo-sphere” for the milli-charged particles is created. For them to be less effective at carrying away energy, their transport cross section must be about as large as that for neutrinos or larger. Very crudely, we must compare Eq. (32) with the weak-interaction transport cross section for neutral-current scattering on nucleons, $$\sigma _{\mathrm{T},\mathrm{weak}}=\frac{2G_F^2E^2}{3\pi }(C_V^2+5C_A^2),$$ (35) where $`|C_A|1.26/2`$ for protons and neutrons, while $`|C_V|0`$ for protons and $`1/2`$ for neutrons. If we use $`z=0.01`$ near the neutrino sphere, the expression in square brackets in Eq. (32) is about 4. With this value we find $$\left(\frac{\sigma _{\mathrm{T},\mathrm{eff}}}{\sigma _{\mathrm{T},\mathrm{weak}}}\right)^{1/2}1.2\times 10^7ϵ\left(\frac{20\mathrm{MeV}}{E}\right)^2.$$ (36) Taking an average energy of 20 MeV we conclude that $`ϵ`$ should exceed approximately $`8\times 10^8`$ for emission of milli-charged particles to be less important than neutrinos. While this derivation is somewhat crude, we conclude that milli-charged particles in the range 109 < ϵ < 107 < superscript109italic-ϵ < superscript10710^{-9}\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}\epsilon\mathrel{\vbox{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}}10^{-7} are excluded. The plasma frequency in a proto-neutron star is roughly 10 MeV so that this bound applies to milli-charged particles with mass below about 5 MeV. The upper bound on $`ϵ`$ will also apply to the model with a second $`U(1)`$. The trapping limit $`ϵ>10^7`$ could be slightly modified by the presence of the paraphoton, but this area of parameter space is already ruled out by nucleosynthesis, so we do not discuss this further. In summary, SN 1987A excludes only a very narrow sliver of parameter space in addition to what is excluded by other arguments (Fig. 1). ## 5 Milli-Charged Neutrinos In the Standard Model with massless neutrinos, there are four anomaly-free $`U(1)`$ symmetries, corresponding to the Standard Model hypercharge $`Y_{\mathrm{SM}}`$, and $`\{B/3L_i\}`$. $`B`$ is baryon number, and $`L_i`$ are the lepton numbers of the three lepton families. The hypercharge operator can be redefined to be $$Y^{}=Y_{\mathrm{SM}}+2\underset{i}{}ϵ_i\left(\frac{B}{3}L_i\right)$$ (37) without making the theory anomalous. This gives electric charge $`ϵ_i`$ to $`\nu _i`$ and generates a proton-electron charge difference $`ϵ_e`$. Constraints on this model and variants have been discussed in . It is more difficult to give charge to massive neutrinos in this way, as discussed in . The solar and atmospheric neutrino deficits can be explained by masses, mixing the three Standard Model neutrinos. If this is the case, lepton number and the lepton flavours are not conserved, so the neutrinos cannot acquire electric charge by the redefinition of equation (37). If there are three additional gauge singlet “right-handed” neutrinos without Majorana masses, the neutrino masses can conserve $`BL`$ which would allow the three flavours to have the same charge $`ϵ`$. In this case, the observed neutrality of matter implies $`ϵ<10^{21}`$ . It would be possible for $`\nu _\mu `$ and $`\nu _\tau `$ to have significantly larger charges than this if the observed solar and atmospheric neutrino deficits are not due to a neutrino mass matrix mixing the three Standard Model neutrinos<sup>5</sup><sup>5</sup>5For instance, it has been suggested that the solar neutrino deficit could be caused by the deflection of milli-charged massless neutrinos in the Sun’s magnetic field . Bounds on the electric charge of $`\nu _\tau `$ were calculated in without assuming any relation between the tau neutrino charge and the electric charge of the other two neutrinos. We do not consider this possibility. ## 6 Conclusion We have updated the bounds from astrophysics, cosmology and laboratory experiments on fermions with electric charge $`ϵe`$, where $`ϵ<1`$. These “milli-charged” particles could be neutrinos or new fermions from beyond-the-Standard-Model. If milli-charged neutrinos have mass, and the Standard Model particle content is non-anomalous, the electric charge of neutrinos is constrained to be $`<10^{21}`$. The updated bounds on milli-charged particles from beyond-the-Standard-Model are presented in figure 1, for both the cases where there is, and there is not, a paraphoton. For masses $`<5`$ keV, a fermion with electric charge $`2\times 10^{14}<ϵe<1`$ is ruled out. An electric charge $`ϵ>10^8`$ is ruled out up to masses $``$ MeV. Milli-charged particles could be possible for masses between an MeV and a TeV. ###### Acknowledgments. We are grateful to Andrzej Czarnecki and Alick Macpherson for informative discussions. In Munich, this work was supported, in part, by the Deutsche Forschungsgemeinschaft under grant No. SFB-375. In Copenhagen, it was supported by a grant from the Carlsberg Foundation.
warning/0001/gr-qc0001097.html
ar5iv
text
# On the spontaneous breaking of Lorentz invariance ## 1 INTRODUCTION Lorentz invariance is one of the most fundamental symmetries of physics and is an underlying ingredient of all known physical theories. However, more recently, there has been theoretical evidence that, in the realm of string/M-theory, this symmetry may be violated. This naturally poses the question of verifying this violation experimentally. In this contribution<sup>1</sup><sup>1</sup>1Invited talk delivered at the Third Meeting on Constrained Dynamics and Quantum Gravity, Villasimius, Sardinia, September 1999., we study the implications of a putative violation of Lorentz invariance in the context of a Lorentz-violating extension of the Standard Model (SM) aiming to analyse its impact on the physics of ultra-high energy cosmic rays and the possibility of an astrophysical test of this violation . Actually, in what concerns the latter issue, it is striking that there is convincing evidence that the observed jets of Active Galactic Nuclei (AGN) are efficient cosmic proton accelerators. Furthermore, the photoproduction of neutral pions by accelerated protons is assumed to be the source of the highest-energy photons through which most of the luminosity of the galaxy is emitted. The decay of charged pions with the ensued production of neutrinos is another distinct signature of the proton induced cascades and estimates of the neutrino flux are believed to be fairly model independent . Gamma-Ray Bursts (GRB) have been also suggested as a possible source of high-energy neutrinos . Moreover, a deeper understanding of these sources is expected as large area ($`km^2`$) high-energy neutrino telescopes are under construction (see e.g. ). These telescopes will allow obtaining information that is congenially correlated with gamma-ray flares and bursts emitted by AGN and GRB sources. On the other hand, it has already been pointed out that astrophysical observations of faraway sources of gamma radiation could provide important hints on the nature of gravity-induced effects and hence on physics beyond the Standard Model (SM). We argue that delay measurements in the arrival time of correlated sources of gamma radiation and high-energy neutrinos can, when considered in the context of a Lorentz-violating extension of the SM , help setting relevant limits on the violation of that fundamental symmetry. As we shall see, we can relate our results with the recently discussed limit on the violation of Lorentz symmetry from the observations of high-energy cosmic rays beyond the GKZ cut-off . The idea of dropping the Lorentz symmetry has been sugested long ago. Indeed, a background cosmological vector field has been considered as a way to introduce our velocity with respect to a preferred frame of reference into the physical description . It has also been proposed, based on the behaviour of the renormalization group $`\beta `$function of non-abelian gauge theories, that Lorentz invariance could be actually a low-energy symmetry . In higher dimensional theories of gravity, models that are not locally Lorentz invariant have been studied in order to obtain light fermions in chiral representations . The spontaneous breaking of Lorentz symmetry due to non-trivial solutions of string field theory was first discussed in Refs. . These non-trivial solutions arise in the context of the string field theory of open strings and may have striking implications at low-energy. For example, assuming that the contribution of Lorentz-violating interactions to the vacuum energy is about half of the critical density leads to the conclusion that quite feeble tensor mediated interactions in the range of about $`10^4m`$ should exist . Lorentz violation may lie, with the help of inflation, at the origin of the primordial magnetic fields required to explain the observed galactic magnetic field as it may also imply in the breaking of conformal symmetry of electromagnetism . It is quite natural that violations of the Lorentz invariance may imply in the breaking of CPT symmetry . Interestingly, this possibility can be verified experimentally in neutral-meson experiments, Penning-trap measurements and hydrogen-antihydrogen spectroscopy . Moreover, the breaking of CPT symmetry also allows for an explanation of the baryon asymmetry of the Universe, as tensor-fermion-fermion interactions expected in the low-energy limit of string field theories give rise to a chemical potential that creates in equilibrium a baryon-antibaryon asymmetry in the presence of baryon number violating interactions . Limits on the violation of Lorentz symmetry have been directly investigated through laser interferometric versions of the Michelson-Morley experiment where comparison between the velocity of light, $`c`$, and the maximum attainable velocity of massive particles, $`c_i`$, up to $`\delta |c^2/c_i^21|<10^9`$ . In the so-called Hughes-Drever experiment much more stringent limits can be obtained, searching for a time dependence of the quadrupole splitting of nuclear Zeeman levels along Earth’s orbit, e.g. $`\delta <3\times 10^{22}`$ . Impressively, a more recent assessment of these experiments reveals that more accurate bounds, up to 8 orders of magnitude, can be reached . From the astrophysical side, limits on the violation of momentum conservation and the existence of a preferred reference frame can also be established from bounds on the parametrized post-Newtonian parameter, $`\alpha _3`$. This parameter vanishes in General Relativity and can be extracted from the pulse period of pulsars and millisecond pulsars . The most recent bound, $`|\alpha _3|<2.2\times 10^{20}`$ , indicates that Lorentz symmetry is unbroken up to that level. In the next sections, we shall calculate the corrections to the dispersion relation arising from a Lorentz-violating extension of the SM and confront it with the evidence on the violation of Lorentz invariance arising from the ultra-high energy cosmic ray physics. We shall also see that these corrections induce a time delay in the arrival of signals from faraway sources carried by different particles. ## 2 LORENTZ-VIOLATING EXTENSION OF THE STANDARD MODEL It is widely believed that SM is a low-energy description of a more fundamental theory, where all interactions including gravity are unified and the hierarchy problem is solved. It is quite plausible that, in this most likely higher-dimensional fundamental theory, symmetries such as CPT and Lorentz invariance, may undergo spontaneous symmetry breaking. The fact that within string/M-theory, currently the most promising candidate for a fundamental theory, a mechanism where spontaneous breaking of Lorentz symmetry is known , suggests that the violation of those symmetries might actually occur and that its implications should be investigated. There is no reason, at least in principle, for this breaking not to extend into the four-dimensional spacetime. If this is indeed the case, CPT and Lorentz symmetry violations will be likely to take place within the SM and its effects might be detected. In order to account for the CPT and Lorentz-violating effects an extension to the minimal SM has been developed based on the assumption that CPT and Lorentz-violating terms might arise from interactions of tensor fields to Dirac fields when Lorentz tensors acquire non-vanishing vacuum expectation values. Interactions of this form are expected to arise from the string field trilinear self-interaction, as is in the open string field theory . These interactions may also emerge from the scenario where our world is wrapped in a $`3`$-brane and this is allowed to tilt . In order to preserve the SM power-counting renormalizability only terms involving operators of mass dimension four or less are considered in this extention. In , only the fermionic sector of the extension was considered <sup>2</sup><sup>2</sup>2It was assumed that SM gauge sector is unaltered. Changing this sector has been already considered, but the phenomenological restrictions are quite severe, at least in what concerns the term that gives origin to a cosmological birefringence .. This sector includes both leptons and quarks, since SU(3) symmetry ensures violating extensions to be colour-independent. The extended fermionic sector consists of CPT-odd and CPT-even contributions, which are given by $$_{\mathrm{Fermion}}^{\mathrm{CPT}\mathrm{odd}}=a_\mu \overline{\psi }\gamma ^\mu \psi b_\mu \overline{\psi }\gamma _5\gamma ^\mu \psi ,$$ (1) $`_{\mathrm{Fermion}}^{\mathrm{CPT}\mathrm{even}}`$ $`=`$ $`\frac{1}{2}ic_{\mu \nu }\overline{\psi }\gamma ^\mu \stackrel{}{^\mu }\psi `$ (2) $`+`$ $`\frac{1}{2}id_{\mu \nu }\overline{\psi }\gamma _5\gamma ^\mu \stackrel{}{^\mu }\psi `$ $``$ $`H_{\mu \nu }\overline{\psi }\sigma ^{\mu \nu }\psi ,`$ where the coupling coefficients $`a_\mu `$ end $`b_\mu `$ have dimensions of mass, $`c_{\mu \nu }`$ and $`d_{\mu \nu }`$ are dimensionless and can have both symmetric and anti-symmetric components, and $`H_{\mu \nu }`$ has dimension of mass and is anti-symmetric. All Lorentz violating parameters are hermitian and flavour-dependent. Some of them may induce flavour changing neutral currents when non-diagonal in flavour. The Langrangian density of the fermionic sector including Lorentz-violating terms reads: $``$ $`=`$ $`\frac{1}{2}i\overline{\psi }\gamma _\mu \stackrel{}{^\mu }\psi a_\mu \overline{\psi }\gamma ^\mu \psi b_\mu \overline{\psi }\gamma _5\gamma ^\mu \psi `$ (3) $`+`$ $`\frac{1}{2}ic_{\mu \nu }\overline{\psi }\gamma ^\mu \stackrel{}{^\nu }\psi +\frac{1}{2}id_{\mu \nu }\overline{\psi }\gamma _5\gamma ^\mu \stackrel{}{^\nu }\psi `$ $``$ $`H_{\mu \nu }\overline{\psi }\sigma ^{\mu \nu }\psi m\overline{\psi }\psi ,`$ where only kinetic terms are kept as we are interested in deducing the free particle energy-momentum relation. From the above Lagrangian density, we can get the Dirac-type equation $`[i\gamma ^\mu (_\mu +(c_\mu ^\alpha d_\mu ^\alpha \gamma _5)_\alpha )a_\mu \gamma ^\mu `$ $`b_\mu \gamma _5\gamma ^\mu H_{\mu \nu }\sigma ^{\mu \nu }m]\psi =0.`$ (4) In order to obtain the corresponding Klein-Gordon equation, we multiply eq. (2) from the left by itself with an opposite mass sign yielding: $`[[i(_\mu +c_\mu ^\alpha _\alpha )a_\mu ]^2+(d_\mu ^\alpha _\alpha )^2b^2m^2`$ (5) $``$ $`i\sigma ^{\mu \rho }[i_\mu c_\rho ^\beta _\beta +ic_\mu ^\alpha _\alpha i_\rho +ic_\mu ^\alpha _\alpha ic_\rho ^\beta _\beta `$ $``$ $`i(_\mu +c_\mu ^\alpha _\alpha )id_\rho ^\beta \gamma _5_\beta +id_\mu ^\alpha \gamma _5_\alpha i(_\rho `$ $`+`$ $`c_\rho ^\beta _\beta )2b_\mu \gamma _5[i(_\rho +c_\rho ^\beta _\beta )a_\rho ]]`$ $``$ $`2i(ia_\mu \sigma ^{\mu \rho }b_\mu \gamma _5g^{\mu \rho })d_\rho ^\beta \gamma _5_\beta `$ $`+`$ $`\sigma ^{\mu \nu }\sigma ^{\rho \sigma }H_{\mu \nu }H_{\rho \sigma }H_{\rho \sigma }(\gamma ^\mu \sigma ^{\rho \sigma }+\sigma ^{\rho \sigma }\gamma ^\mu )`$ $`[i(_\mu +(c_\mu ^\alpha d_\mu ^\alpha \gamma _5)_\alpha )`$ $``$ $`a_\mu +b_\mu \gamma _5]]\psi =0.`$ To eliminate the off-diagonal terms, the squaring procedure has to be repeated once again. However, since Lorentz symmetry breaking effects are quite constrained experimentally, violating terms higher than second order will be dropped. After some algebra, we find that off-diagonal terms cannot be fully cancelled, but that these terms are higher order in the Lorentz violating parameters. To simplify further we also drop $`H_{\mu \nu }`$. This is justifiable as only time-like Lorentz-violating parameters are going to be studied. Hence, we obtain, for the Klein-Gordon type equation, up to second order in the new parameters: $`[[(i)^2+2i_\mu ic^{\mu \alpha }_\alpha 2i_\mu a^\mu m^2]^2`$ (6) $`+`$ $`4i_\mu id_\rho ^\beta _\beta i_\eta id_\varphi ^\delta _\delta (g^{\mu \eta }g^{\rho \varphi }g^{\mu \rho }g^{\eta \varphi })`$ $``$ $`8i_\mu id_\rho ^\beta _\beta b_\eta i_\varphi (g^{\mu \rho }g^{\eta \varphi }g^{\mu \varphi }g^{\rho \eta })`$ $`+`$ $`4b_\mu b_\eta i_\rho i_\varphi (g^{\mu \eta }g^{\rho \varphi }g^{\mu \rho }g^{\eta \varphi })]\psi =0.`$ Thus, in momentum space we get at lowest non-trivial order, the following relationship: $`(p_\mu p^\mu +2p_\mu c^{\mu \alpha }p_\alpha +2p_\mu a^\mu m^2)^2`$ (7) $`+`$ $`4[p_\mu p^\mu d_\rho ^\beta p_\beta d^{\rho \delta }p_\delta (p_\mu d^{\mu \beta }p_\beta )^2]`$ $`+`$ $`8(p_\mu p^\mu d_\eta ^\beta p_\beta b^\eta p_\mu d^{\mu \beta }p_\beta b_\eta p^\eta )`$ $`+`$ $`4[b_\mu b^\mu p_\nu p^\nu (b_\mu p^\mu )^2]=0.`$ Hence, the dispersion relation arising from the Lorentz-violating extension of the SM is the following: $`p_\mu p^\mu m^2=2p_\mu c^{\mu \alpha }p_\alpha 2p_\mu a^\mu `$ (8) $`\pm `$ $`2[(p_\mu d^{\mu \beta }p_\beta )^2p_\mu p^\mu d_\eta ^\beta p_\beta d^{\eta \delta }p_\delta `$ $`+`$ $`2(p_\mu d^{\mu \beta }p_\beta b_\rho p^\rho p_\mu p^\mu d_\eta ^\beta p_\beta b^\eta )`$ $``$ $`b_\mu b^\mu p_\nu p^\nu +(b_\mu p^\mu )^2]^{1/2},`$ where the $`\pm `$ sign refers to the fact that the effects of $`b_\mu `$ and $`d_{\mu \nu }`$ depend on chirality. Finally, we consider, for simplicity, the scenario where coefficients $`a_\mu `$, $`b_\mu `$, $`c_{\mu \nu }`$ and $`d_{\mu \nu }`$ have only time-like components, which yields the simplified dispersion relation $$p_\mu p^\mu m^2=2c_{00}E^22aE\pm 2(b+d_{00}E)p,$$ (9) where we have dropped the component indices of coefficients $`a`$ and $`b`$. From now on we drop parameter $`a`$ as it is potentially dangerous from the point of view of giving origin to flavour changing neutral currents when more than one flavour is involved. In the next section, we shall use the dispersion relation (9) to examine how GZK cut-off for ultra-high energy cosmic rays can be relaxed. The ensued discussion is similar to the one described in , where it is assumed that the limiting velocities of particles in different reference frames are ad hoc different. ## 3 ULTRA-HIGH ENERGY COSMIC <br>RAYS The discovery of the cosmic background radiation made inevitable the question of how the most energetic cosmic-ray particles would be affected by the interaction with the microwave photons. Actually, the propagation of the ultra-high energy nucleons is limited by inelastic impacts with photons of the background radiation so that nucleons with energies above $`5\times 10^{19}eV`$ are unable to reach Earth from further than $`50100Mpc`$. This is the well known GZK cut-off . However, events where the estimated energy of the cosmic primaries is beyond the GZK cut-off have been observed by different collaborations . It has been suggested (see also ) that slight violations of Lorentz invariance could be at the origin of energy-dependent effects which would suppress processes, otherwise dynamically inevitable, such as for instance the resonant reaction, $$p+\gamma _{2.73K}\mathrm{\Delta }_{1232},$$ (10) which is central to the GZK cut-off. Were this process untenable, the GZK cut-off would not hold and therefore a cosmological origin for the high-energy cosmic radiation is theoretically acceptable. As discussed in , this can occur through a change in the dispersion relation for free particles. We shall see that this is indeed what happens when process (10) is analysed with dispersion relation (9). Considering a head-on impact of a proton of energy $`E`$ with a cosmic background radiation photon of energy $`\omega `$, the likelihood of the process (10) would be conditioned on satisfying (cf. eq. (9)) $$2\omega +Em_\mathrm{\Delta }(1c_{00}^\mathrm{\Delta }).$$ (11) Thus, we get from (9) after squaring (11) and dropping the $`\omega ^2`$ term $$2\omega +\frac{m_p^2}{2E}(c_{00}^pc_{00}^\mathrm{\Delta })E+\frac{m_\mathrm{\Delta }^2}{2E},$$ (12) which clearly exhibits Lorentz-violating terms. Let us now compare (12) with the results of Ref. and show that this leads to a bound on $`\mathrm{\Delta }c_{00}`$. In order to modify the usual dispersion relation for free particles, Coleman and Glashow suggested assigning a maximal attainable velocity to each particle. Thus, for a given particle $`i`$ moving freely in the preferred frame, which could be thought of as the one in relation to which the cosmic background radiation is isotropic, the dispersion relation would be $$E^2=p^2c_i^2+m_i^2c_i^4.$$ (13) Hence, the likelihood of the process (10) to occur under the conditions described above would depend on satisfying the kinematical requirement $`2\omega +Em_{eff}`$, where the effective mass $`m_{eff}`$ is given by $$m_{eff}^2m_\mathrm{\Delta }^2(c_p^2c_\mathrm{\Delta }^2)p^2,$$ (14) the momentum being with respect to the preferred frame. The likelihood condition takes then the following form $$2\omega +\frac{m_p^2}{2E}(c_pc_\mathrm{\Delta })E+\frac{m_\mathrm{\Delta }^2}{2E},$$ (15) where the term proportional to $`c_pc_\mathrm{\Delta }`$ is clearly Lorentz-violating. If the difference in the maximal velocities exceeds the critical value $$\delta (\omega )=\frac{2\omega ^2}{m_\mathrm{\Delta }^2m_p^2},$$ (16) then reaction (10) would be forbidden and consequently the GZK cut-off relaxed. For photons of the microwave background, $`T=2.73K`$, and $`\omega _0kT=2.35\times 10^4eV`$, this condition would be $$c_pc_\mathrm{\Delta }=\delta (\omega _0)1.7\times 10^{25},$$ (17) which is a quite impressive bound on the violation of the Lorentz symmetry, even though valid only for the process in question. Similar bounds for other particle pairs, although less stringent, were discussed in . Finally, comparison of (15) with (12) yields: $$c_{00}^pc_{00}^\mathrm{\Delta }1.7\times 10^{25}.$$ (18) Thus, we see that the Lorentz-violating extension of the SM can also describe the phenomenology of ultra-high energy cosmic rays and explain the violation of the GZK cut-off. Of course, the situation would be more complex if the Lorentz-violating parameters were allowed to have space-like components which would lead to direction and helicity-dependent effects. ## 4 AN ASTROPHYSICAL TEST OF LORENTZ INVARIANCE We now turn to the discussion of a possible astrophysical test of Lorentz invariance. From eq. (18) we see that $`\mathrm{\Delta }c_{00}ϵ`$, where $`ϵ`$ is a small constant specific of the process involved (cf. eqs. (17) and (18)) and, for instance, $`|ϵ|\text{ }<few\times 10^{22}`$ from the search of neutrinos oscillations . As far as signals simultaneously emitted by faraway sources are concerned, the resulting effect in the propagation velocity of particles with energy, $`E`$, and momentum, $`p`$, is given in the limit where $`m<<p,E`$ or, for massless particles, by $`c_i=c[1(c_{00}\pm d_{00})_i]`$. Therefore, for sources at a distance $`D`$, the delay in the arrival time will be given by: $$\mathrm{\Delta }t\frac{D}{c}[(c_{00}\pm d_{00})_i(c_{00}\pm d_{00})_j]ϵ_{ij}^\pm \frac{D}{c},$$ (19) where a new constant, $`ϵ_{ij}^\pm `$, involving a pair of particles is defined. This time delay may, despite being given by the difference between two fairly small numbers, be measurable for sufficiently far away sources. Furthermore, our result shows that the estimated time delay is energy independent, in opposition to what was obtained from general arguments . We have also found that the time delay has a dependence on the chirality of the particles involved. In the next section we shall discuss how to estimate the observational value of $`c_{00}^i`$ (and $`d_{00}^i`$ if $`d_{00}^ic_{00}^i`$). Therefore if, for instance, the signals from faraway sources were, as suggested previously in the introduction, from AGN $`TeV`$ gamma-ray flares and the intrinsically related neutrino emission, we should expect for the time delay, $$\mathrm{\Delta }t(c_{00}\pm d_{00})_\nu \frac{D}{c},$$ (20) if as argued above the photon propagation is unaltered and assuming that the neutrinos are massless, an issue that will be most probably settled experimentally in the near future. It is worth stressing that even before that, the effect of neutrino masses and other intrinsic effects related with the nature of the neutrino emission can, at least in principle, be extracted from data of several correlated detections of $`TeV`$ gamma-ray flares and neutrinos, if a systematic delay of neutrinos is observed. Moreover, the available knowledge of AGN phenomena and the confidence on the astrophysical methods available to determine their distance from us, make it plausible that the time delay strategy may provide relevant bounds on the violation of Lorentz symmetry. Of course, similar arguments may equally apply to GRB, however, the lack of a deeper understanding of these transient phenomena introduces further undesirable uncertainty. It is also important to point out that bounds involving photon and neutrinos are currently unknown and that a difference in the arrival time between neutrinos and antineutrinos is expected if $`d_{00}^i`$ is non-vanishing. ## 5 CONCLUSIONS In summary, we have shown that parameters of the Lorentz-violating extension of the SM proposed in Ref. can be related with the phenomenology of ultra-high energy cosmic rays with the conclusion that, as in , it may lead to the suppression of processes responsible for the GZK cut-off. This is a crucial argument in favour of a possible extra-galactic origin for the ultra-high energy cosmic rays. We have also found that the relevant Lorentz-violating parameter is, at high energies, $`c_{00}^i`$ so that $`\mathrm{\Delta }c_{00}ϵ`$ with $`|ϵ|\text{ }<few\times 10^{22}`$ from neutrino physics and $`|ϵ|\text{ }<10^{25}`$ from the ultra-high energy cosmic ray physics. It is possible to estimate the typical scales assuming that the source of Lorentz symmetry violation is due to non-trivial solutions in string field theory. Indeed, these solutions imply that Lorentz tensors acquire vacuum expection values as Lorentz symmetry is spontaneously broken due to string induced interactions . A parametrization for these expectation values and, hence, for $`c_{00}^i`$ would be for a fixed energy scale, E, the following : $$c_{00}^i\frac{T}{M_S}=\lambda _i\left(\frac{m_L}{M_S}\right)^l\left(\frac{E}{M_S}\right)^k,$$ (21) where $`T`$ denotes a generic Lorentz tensor, $`\lambda _i`$ is presumably an order one flavour-dependent constant<sup>3</sup><sup>3</sup>3A scenario were $`\lambda _i`$ is of order of the respective Yukawa coupling has been also discussed ., $`m_L`$ is a light mass scale, $`M_S`$ is a string scale presumably close to Planck’s mass or a few orders of magnitude below it, and $`k,l`$ are integers labelling the order of the string corrections at low-energy. Thus, in the lowest non-trivial order, $`k=0,l=1`$ ($`k=l=0`$ being already excluded experimentally), $`c_{00}^i=\lambda _i\left(\frac{m_L}{M_S}\right)`$ and different $`\lambda _i`$ constants lead to $`ϵ\left(\frac{m_L}{M_S}\right)`$. Thus, we see that theoretical estimates are consistent with high-energy cosmic ray data. Moreover, if for example, $`ϵ\text{ }<10^{23}`$, then it follows that $`m_L10^2KeV`$ for $`M_SM_P`$ or $`m_L10^2eV`$ if $`M_Sfew\times 10^{16}GeV`$ . Estimates for $`m_L`$ would clearly change by many orders of magnitude if $`\lambda _i`$ were of order of the Yukawa coupling. In either case, we can conclude that choice $`k=0,l=1`$ implies the time delay in the arrival of signals from faraway sources is energy independent. We have found, however, an interesting dependence on the chirality of particles involved. Of course, another scenario would emerge from a different choice of integers $`k,l`$. For instance, the choice $`k=2`$ and $`l=0`$, the relevant one in the CPT symmetry violating baryogenesis scenario , where the energy should, in this case, be related with the early Universe temperature. This would imply that the time delay in the arrival of signals from faraway sources would be proportional to the square of the energy. This choice would also lead to the conclusion that Lorentz violating effects, whether due to string physics or quantum gravity, are quadratic in the energy. Similar conclusions concerning the order of quantum gravity low-energy effects are obtained from the study of corrections to the Schrödinger equation arising from quantum cosmology in the minisuperspace approximation . In another theoretical setting, the spontaneous breaking of the Lorentz symmetry may occur is the so-called braneworld . In this scenario, SM particles lie on a $`3`$-brane, $`\varphi (x)`$, embedded in spacetime, with possibly large compact extra dimensions, whereas gravity propagates in the bulk. Thus, a tilted brane would induce rotational and Lorentz non-invariant terms in the four-dimensional effective theory as brane-Goldstones couple to all particles on the brane via an induced metric on the brane. This will lead to operators of the form: $$_\mu \varphi _\nu \varphi \overline{\psi }\gamma ^\mu ^\nu \psi +\mathrm{},$$ (22) which clearly resemble some of the Lorentz-violating terms in the SM extention discussed above. Corrections to the kinetic terms of gauge fields are also expected. As before, phenomenology sets tight constraints on this scenario remaining, however, unable to establish whether the breaking of Lorentz invariance, if observed at all, has its origin on the non-perturbative nature of branes or has its roots in the perturbative string field theory scenario described above. The former possibility should more probably be associated with a $`M_S`$ scale that is a few orders of magnitude below Planck scale, while the latter to a value of $`M_S`$ that should be associated with the Planck scale itself. Finally, we have outlined a strategy to establish to what extent Lorentz invariance is violated, from the observation of the time delay in the detection of $`TeV`$ gamma-ray flares and neutrinos from AGN. Our analysis reveals that the time delay has a dependence on the chirality of the particles involved, but is energy independent, contrary to what one could expect from general arguments. In either case, if ever observed, a time delay in the arrival of signals from faraway sources would be a strong evidence of new physics beyond the SM. I would like to thank Carla Carvalho for collaboration in the project that gave origin to this contribution and Alan Kostelecký for the relevant comments and suggestions. I am also extremely grateful to the organizing committee of the Constrained Dynamics and Quantum Gravity 1999 for the superb organizational work and for the most warm hospitality with which myself and my family were received in Sardegna.
warning/0001/quant-ph0001051.html
ar5iv
text
# Spin dynamics of wave packets evolving with the Dirac Hamiltonian in atoms with high Z ## 1 Introduction In the last decade large efforts have been devoted to the detailed understanding of the quantum dynamics of WP in simple systems like H atoms or hydrogenoid atoms. Theoretical investigations resulted in very good understanding of such subtle interference effects like collapse, revivals, fractional revivals and super-revivals of wavepackets created in atomic and molecular systems. Many signatures of these phenomena have been observed in numerous sophisticated experimental studies. Indeed with contemporary lasers and tunable electromagnetic fields one can engineer many different desired initial states and analyze their time evolution . Our aim in this paper is to perform a new step in the analysis of the WP dynamics for hydrogen atom and for ions with one electron by considering spin motion. Until now this effect was not considered for isolated atoms since it requires very long time scales to manifest itself due to the weakness of the spin–orbit coupling. In order to decrease as much as possible this time scale we will consider WP within ions with very large $`Z`$. We intended to make a calculation which include all possible effects for these elements and therefore we have considered WP which evolve under the relativistic Dirac Hamiltonian i.e. which are four component spinors. Spin effects and relativistic effects have been studied many times but mostly for atoms submitted to intense laser fields. A recent review has been given in . Some of the most relevant research works are found in . It is generally agreed in them that spin and relativistic effects are small but that their magnitude increase with the intensity of the field. For fields of the order of $`10^{16}`$ to $`10^{17}`$ W/cm<sup>2</sup> they should be taken into account. In a most recent paper the interest has been focussed on the spin–orbit coupling, essentially for Al<sup>+12</sup> and Ga<sup>+12</sup> ions. The dynamics of the spin degree of freedom was analyzed there during the time of interaction of the laser with the ion. The authors did not use the Dirac equation but the Schrödinger equation corrected up to order $`1/c^3`$ and therefore the spin–orbit potential could be clearly identified by its mere suppression. The most significant signature occur in the radiation spectrum. However little polarization was produced for the ions which were considered. Spin dynamics was previously studied by us in harmonic oscillator potentials in non-relativistic dynamics. In ref. we have found a new effect that we have called the spin–orbit pendulum for a spin–orbit potential with a constant radial part. This phenomenon is simply the periodic collapse and revival of the average spin and an exchange of angular momentum between spin and the orbital angular momentum. The analysis is quite parallel to that of the Jaynes-Cummings model . In order to extend this situation to a relativistic dynamics it was quite natural to study the Dirac oscillator (DO). There, the strong entanglement that gives rise to spin oscillation is still present . Because of nonlinearities of the relativistic energies the dynamics is not periodic. We have also studied relativistic effects like zitterbewegung. Our studies have been also extended to WP motion corresponding to linear trajectories in non-relativistic approach (HO+LS Hamiltonian) as well as in the relativistic Dirac oscillator . Our previous studies stimulated the present work for hydrogen or for heavy ions. It is expected that many features of the DO should take place in this much more realistic situation. Our paper is organized as follows. In order to make it self-contained we in Sec. 2 present basic formula for solutions of Dirac equation with Coulomb potential. In Sec. 3 we show the construction of the circular WP and in Sec. 4 we discuss its time evolution, focusing on the autocorrelation function, spin expectation values and spatial motion. Particular attention is paid to relativistic effects and different time scales. The concluding remarks are contained in Sec. 5. ## 2 Dirac equation with central potential Solution of Dirac equation for a central field, in particular for Coulomb potential is one of standard, textbook problems in quantum mechanics . Therefore we will recall here only few formula that will be relevant for our further considerations. Dirac equation for a spherical potential $`V(r)`$ is $$H_D\mathrm{\Psi }(𝒓)=(c\widehat{𝜶}𝒑+\widehat{\beta }mc^2+V(r))\mathrm{\Psi }(𝒓)=E\mathrm{\Psi }(𝒓).$$ (1) Its eigenfunctions can be found in a representation that diagonalizes simultaneously with $`H_D`$ three other operators, $`𝒋^2,j_z`$ and $`K=\mathrm{}+𝑳𝝈`$. Then we may write for $`\mathrm{\Psi }(𝒓)`$ (using phase convention as in ) $$\mathrm{\Psi }=\mathrm{\Psi }_\kappa ^m=\mathrm{}\left(\begin{array}{c}g(r)\chi _\kappa ^m\\ \mathrm{}f(r)\chi _\kappa ^m\end{array}\right)=\mathrm{}\left(\begin{array}{c}g(r)\mathrm{\Omega }_{ljm}\\ \mathrm{}f(r)\mathrm{\Omega }_{l^{}jm}\end{array}\right)$$ (2) with $$l^{}=2jl=\{\begin{array}{c}2(l+\frac{1}{2})l=l+1\text{for}j=l+\frac{1}{2}=j_+\\ 2(l\frac{1}{2})l=l1\text{for}j=l\frac{1}{2}=j_{}\end{array}.$$ (3) The explicit forms of the spherical spinors $`\mathrm{\Omega }_{ljm}`$ are: $$\mathrm{\Omega }_{l,j=l+{\scriptscriptstyle \frac{1}{2}},m}=\mathrm{\Omega }_{l,j_+,m}=\left(\begin{array}{c}\sqrt{\frac{j+m}{2j}}Y_{l,m{\scriptscriptstyle \frac{1}{2}}}\\ \sqrt{\frac{jm}{2j}}Y_{l,m+{\scriptscriptstyle \frac{1}{2}}}\end{array}\right)$$ (4a) and $$\mathrm{\Omega }_{l,j=l{\scriptscriptstyle \frac{1}{2}},m}=\mathrm{\Omega }_{l,j_{},m}=\left(\begin{array}{c}\sqrt{\frac{jm+1}{2j+2}}Y_{l,m{\scriptscriptstyle \frac{1}{2}}}\\ \sqrt{\frac{j+m+1}{2j+2}}Y_{l,m+{\scriptscriptstyle \frac{1}{2}}}\end{array}\right).$$ (4b) Radial wave functions are given in terms of confluent hypergeometric functions $`F(a,c,x)`$ $$F(a,c,x)=1+\frac{a}{c}x+\frac{a(a+1)}{c(c+1)}\frac{x^2}{2!}+\mathrm{}.$$ (4e) Explicit formula are : $`f(r)`$ $`=`$ $`{\displaystyle \frac{\sqrt{2}\lambda ^{5/2}}{\mathrm{\Gamma }(2\gamma +1)}}\left[{\displaystyle \frac{\mathrm{\Gamma }(2\gamma +n^{}+1)(1E)}{n^{}!\xi (\xi \lambda /\kappa )}}\right]^{\frac{1}{2}}(2\lambda r)^{\gamma 1}\mathrm{}^{\lambda r}`$ $`\times `$ $`[n^{}F(n^{}+1,2\gamma +1,2\lambda r)(\kappa \xi /\lambda )F(n^{},2\gamma +1,2\lambda r)]`$ and $`g(r)`$ $`=`$ $`{\displaystyle \frac{\sqrt{2}\lambda ^{5/2}}{\mathrm{\Gamma }(2\gamma +1)}}\left[{\displaystyle \frac{\mathrm{\Gamma }(2\gamma +n^{}+1)(1+E)}{n^{}!\xi (\xi \lambda /\kappa )}}\right]^{\frac{1}{2}}(2\lambda r)^{\gamma 1}\mathrm{}^{\lambda r}`$ $`\times `$ $`[n^{}F(n^{}+1,2\gamma +1,2\lambda r)(\kappa \xi /\lambda )F(n^{},2\gamma +1,2\lambda r)],`$ where $`\xi =Z\alpha `$, $`\gamma =\sqrt{\kappa ^2\xi ^2}`$, $`\kappa ^2=(j+1/2)^2`$, $`E=[1+(\frac{\xi }{n^{}+\gamma })^2]^{1/2}`$ and $`\lambda =\sqrt{1E^2}`$, $`\alpha =e^2/\mathrm{}c=1/137.036`$ being the fine structure constant. The nonnegative integer number $`n^{}`$ is related to $`n`$ and $`|\kappa |`$ by $`n=n^{}+|\kappa |`$. In the following calculations we use atomic units $`m_e=\mathrm{}=c=1`$. ## 3 Construction of wavepacket In our first paper on relativistic WP in Dirac oscillator (DO) we were considering a circular spinor without small components written as $$\mathrm{\Psi }=\underset{l}{}w_l\left(\begin{array}{c}a|lll\\ b|lll\\ 0\\ 0\end{array}\right)=\underset{l}{}w_lR_{n=l}(r)|ll\left(\begin{array}{c}a\\ b\\ 0\\ 0\end{array}\right).$$ (4fg) Here $`a`$ and $`b`$ are coefficients which determine the initial direction of the spin. There is no loss of generality to choose them real. $`R_{n=l}(r)`$ denotes radial wave function of the harmonic oscillator. The large component for the partial wave $`l`$ is written as $$|ll\left(\begin{array}{c}a\\ b\end{array}\right)=a\mathrm{\Omega }_{l,j_+,j_+}+b\left(\frac{\mathrm{\Omega }_{l,j_+,j_{}}+\sqrt{2l}\mathrm{\Omega }_{l,j_{},j_{}}}{\sqrt{2l+1}}\right).$$ (4fh) Weights $`w_l`$ of the superposition (4fg) were chosen in a way providing $`\mathrm{\Psi }`$ to be the HO coherent state. For DO the bispinors without small components were constructed as a superposition of spinors with positive and negative energies that provide a cancellation of the small components. For the Dirac equation with Coulomb potential we are not able to get rid of small components completely, as we want to build the WP only from bound states. We can, however, construct small components in a way similar to the big ones providing their circular motion. For this purpose we write the angular parts of the large components as in eq. (4fh), then we complete each of the spinors $`\mathrm{\Omega }_{l,j_+,j_+},\mathrm{\Omega }_{l,j_+,j_{}}`$ and $`\mathrm{\Omega }_{l,j_{},j_{}}`$ by their corresponding small components associating each with its radial part different for the states with $`j_+`$ and $`j_{}`$ as in eq. (2). Thus there are four radial parts respectively called $`g_{j_+},f_{j_+},g_j_{}`$ and $`f_j_{}`$ for a given $`l`$. In order to produce circular waves for hydrogenoid atoms the connection between $`n`$ and $`l`$ is $`l=n1`$. A partial wave at time $`t`$ can be written with these notations using $`E_{j_+}`$ and $`E_j_{}`$ the eigenvalues of the Dirac equation: $`|\psi _l(t)`$ $`=`$ $`\left[a\left(\begin{array}{c}\mathrm{}g_{j_+}\mathrm{\Omega }_{l,j_+,j_+}\\ f_{j_+}\mathrm{\Omega }_{l+1,j_+,j_+}\end{array}\right)+{\displaystyle \frac{b}{\sqrt{2l+1}}}\left(\begin{array}{c}\mathrm{}g_{j_+}\mathrm{\Omega }_{l,j_+,j_{}}\\ f_{j_+}\mathrm{\Omega }_{l+1,j_+,j_{}}\end{array}\right)\right]\mathrm{}^{\mathrm{}E_+t}`$ (4fp) $`+b\sqrt{{\displaystyle \frac{2l}{2l+1}}}\left(\begin{array}{c}\mathrm{}g_j_{}\mathrm{\Omega }_{l,j_{},j_{}}\\ f_{j_+}\mathrm{\Omega }_{l1,j_{},j_{}}\end{array}\right)\mathrm{}^{\mathrm{}E_{}t}.`$ After expansion of the $`\mathrm{\Omega }`$ as in (4a,4b) the circular wavepacket at time $`t`$ can be written as $$\mathrm{\Psi }(t)=\left(\begin{array}{c}|c_1(t)\\ |c_2(t)\\ |c_3(t)\\ |c_4(t)\end{array}\right).$$ (4fq) For Coulomb potential weights $`w_n`$ (that can be made real) are chosen in a standard way , i.e. as a Gaussian distribution with respect to the mean value $`N`$: $$|w_n|^2=(2\pi \sigma _G^2)^{\frac{1}{2}}\mathrm{}^{(nN)^2/2\sigma _G^2}.$$ (4fr) Explicit form of the wavepacket (4fq) is obtained in the following form: $`|c_1(t)=\mathrm{}{\displaystyle \underset{l}{}}w_l\{g_{j_+}[a|l,l+b{\displaystyle \frac{\sqrt{2l}}{2l+1}}|l,l1]\mathrm{}^{\mathrm{}E_{j_+}t}`$ (4fsa) $`bg_j_{}{\displaystyle \frac{\sqrt{2l}}{2l+1}}|l,l1\mathrm{}^{\mathrm{}E_j_{}t}\}`$ $$|c_2(t)=\mathrm{}\underset{l}{}w_lb|l,l\left\{g_{j_+}\frac{1}{2l+1}\mathrm{}^{\mathrm{}E_{j_+}t}+g_j_{}\frac{2l}{2l+1}\mathrm{}^{\mathrm{}E_j_{}t}\right\}$$ (4fsb) $`|c_3(t)={\displaystyle \underset{l}{}}w_l\{f_{j_+}[a{\displaystyle \frac{1}{\sqrt{2l+3}}}|l+1,l+b\sqrt{{\displaystyle \frac{2}{(2l+1)(2l+3)}}}|l+1,l1]\mathrm{}^{\mathrm{}E_{j_+}t}`$ $`bf_j_{}\sqrt{{\displaystyle \frac{2l}{2l+1}}}|l1,l1\mathrm{}^{\mathrm{}E_j_{}t}\}`$ (4fsc) $$|c_4(t)=\underset{l}{}w_lf_{j_+}\left\{a\sqrt{\frac{2l+2}{2l+3}}|l+1,l+1+b\sqrt{\frac{1}{2l+3}}|l+1,l\right\}\mathrm{}^{\mathrm{}E_{j_+}t}.$$ (4fsd) For $`t=0`$ spin and orbital angular momentum are only approximately decoupled, they were exactly decoupled in (4fg). In numerical calculations we use the full bispinor (4fsa-4fsd). However, particularly for $`N>10`$, the contributions arising from the small components ($`|c_3`$ and $`|c_4`$) are almost negligible, because the radial $`f(r)`$ functions are always few orders of magnitude smaller than the corresponding $`g(r)`$ functions. This is due to the fact that their prefactors differ mainly by the factor $`\sqrt{1E}`$ (few orders of magnitude smaller than 1) for $`f(r)`$ and $`\sqrt{1+E}`$ (close to $`\sqrt{2}`$) for $`g(r)`$. Using eqs. (4fs-4fsd) one can calculate the contributions from small components of our circular WP exactly. One obtains $$(\mathrm{\Psi }^{}\mathrm{\Psi })_{small}=c_3(t)|c_3(t)+c_4(t)|c_4(t)$$ (4fst) where $$c_3(t)|c_3(t)=\underset{l}{}w_l^2\left\{\left[a^2\frac{1}{2l+3}+b^2\frac{2}{(2l+1)(2l+3)}\right]F_{j_+}+b^2\frac{2l}{2l+1}F_j_{}\right\},$$ (4fsua) $$c_4(t)|c_4(t)=\underset{l}{}w_l^2\left\{a^2\frac{2l+2}{2l+3}+b^2\frac{1}{2l+3}\right\}F_{j_+}.$$ (4fsub) In fact, $`(\mathrm{\Psi }^{}\mathrm{\Psi })_{small}`$ $`c_3|c_3`$, and $`c_4|c_4`$ are constant contributions for our WP. Terms $`F_j_{}`$ and $`F_{j_+}`$ are radial integrals defined in next section. Fig. 1 displays the magnitude of $`(\mathrm{\Psi }^{}\mathrm{\Psi })_{small}`$ as function of $`Z`$ and $`N`$ for $`Z[1,92]`$ and $`N[2,60]`$. Only for very large $`Z`$ and very small $`N`$ this contribution exceeds $`1\%`$. ## 4 Wavepacket evolution ### 4.1 Time scales We can define several characteristic times on the basis of the previous works about recurrent wave packets . Writing the energy in function of a single quantum number $`n`$ we define for $`n=N`$ a hierarchy of times $$\frac{1}{k!}\left(\frac{\text{d}^kE}{\text{d}n^k}\right)_{n=N}=\frac{2\pi \mathrm{}}{T(k)}k=1,2,3,\mathrm{}$$ (4fsuv) For $`k=1`$ we obtain the classical Kepler time, for $`k=2`$ the revival time, for $`k=3`$ the superrevival time etc…For hydrogenoid atoms those terms increase as $`N^{k+2}`$. These times can be compared to a characteristic spin orbit time $`T_{ls}=2\pi /\omega _{ls}`$ with $`\omega _{ls}`$ defined by $$\mathrm{}\omega _{ls}=E_{j_+=N1/2}E_{j_{}=N3/2}.$$ (4fsuw) To first order we obtain $$\omega _{ls}\frac{(\alpha Z)^2}{N^5}.$$ (4fsux) Therefore the spin–orbit frequency follows a variation parallel to the superrevival time with $`k=3`$. Those times are plotted on Fig. 2 for $`N2`$ and for a large range of atoms (ions) including uranium. For hydrogen $`T_{ls}`$ is orders of magnitude larger than all the other characteristic times $`T(1)T(5)`$. If we want to observe spin–orbit effects in the picosecond regime we can concentrate on $`N=20`$ for $`Z=92`$. With such a choice $`T_{ls}1685T(1)`$. The other times are $`T_2=13.33T(1)`$, $`T_3=200T(1)`$, $`T_4=3200T(1)`$. The choice of $`T_{ls}`$ as a convenient time is justified by our plot of the autocorrelation function of Fig. 3 which shows an interesting oscillations up to 12 periods $`T_{ls}`$. ### 4.2 Autocorrelation function The autocorrelation function (often referred as the recurrence probability) contain a lot of information on wavepacket evolution. It is a convenient measure of the degree of recurrences and can be defined as: $$A(t)=\mathrm{\Psi }(0)|\mathrm{\Psi }(t)=\underset{i=1}{\overset{4}{}}c_i(0)|c_i(t).$$ (4fsuy) Using equations (4fsa-4fsd) one obtains $`c_1(0)|c_1(t)={\displaystyle \underset{l}{}}w_l^2\{(a^2+b^2{\displaystyle \frac{2l}{(2l+1)^2}})G_{j_+}\mathrm{}^{\mathrm{}E_{j_+}t}`$ (4fsuza) $`+b^2{\displaystyle \frac{2l}{(2l+1)^2}}[G_j_{}\mathrm{}^{\mathrm{}E_j_{}t}G_{j_\pm }(\mathrm{}^{\mathrm{}E_{j_+}t}+\mathrm{}^{\mathrm{}E_j_{}t})]\},`$ $`c_2(0)|c_2(t)={\displaystyle \underset{l}{}}w_l^2b^2\{{\displaystyle \frac{1}{(2l+1)^2}}G_{j_+}\mathrm{}^{\mathrm{}E_{j_+}t}+{\displaystyle \frac{(2l)^2}{(2l+1)^2}}G_j_{}\mathrm{}^{\mathrm{}E_j_{}t}`$ (4fsuzb) $`+{\displaystyle \frac{2l}{(2l+1)^2}}G_{j_\pm }(\mathrm{}^{\mathrm{}E_{j_+}t}+\mathrm{}^{\mathrm{}E_j_{}t})\},`$ $`c_3(0)|c_3(t)={\displaystyle \underset{l}{}}w_l^2\{[a^2{\displaystyle \frac{1}{2l+3}}+b^2{\displaystyle \frac{2}{(2l+1)(2l+3)}}]F_{j_+}\mathrm{}^{\mathrm{}E_{j_+}t}`$ (4fsuzc) $`+b^2{\displaystyle \frac{2l}{2l+1}}F_j_{}\mathrm{}^{\mathrm{}E_j_{}t}\},`$ and $`c_4(0)|c_4(t)={\displaystyle \underset{l}{}}w_l^2\left\{a^2{\displaystyle \frac{2l+2}{2l+3}}+b^2{\displaystyle \frac{1}{2l+3}}\right\}F_{j_+}\mathrm{}^{\mathrm{}E_{j_+}t}.`$ (4fsuzd) Terms $`G_{j_+},G_j_{},F_{j_+},F_j_{}`$ and $`G_{j_\pm }`$ are radial integrals (for instance $`G_{j_+}=_0^{\mathrm{}}r^2g_{j_+}^2𝑑r`$ and $`G_{j_\pm }=_0^{\mathrm{}}r^2g_{j_+}g_j_{}𝑑r`$ and so on) that can be easily expressed in terms of $`\mathrm{\Gamma }`$ functions. In the presentation of our results below we focus (as in our previous papers) on the most interesting case, i.e. $`a=b=1/\sqrt{2}`$. This means that initial direction of the spin is along the $`Ox`$ axis, in the orbit’s plane. In Fig. 3 we present the square of the autocorrelation function $`|A(t)|^2`$ as function of time for long term evolution ($`T_{ls}`$ units) of the WP with $`N=20`$. Heavy ion system with $`Z=92`$ is considered with the full relativistic approach. Very regular revival structure connected with integer values of time in $`T_{ls}`$ units is immediately recognized. This regularity is somewhat distorted by a strong influence of $`T_4`$ periodicity (see right bottom picture of Fig. 2) but becomes even more evident for larger $`N`$. Both the main part and the insert of Fig. 3 show very complicated shorter scale behaviour of the autocorrelation function related to the three shorter time scales: $`T_{cl},T_{rev},T_{sr}`$, known already from the non-relativistic studies . The case with $`N=40`$ shown in Fig. 4 suggests that for times $`t[0,1.5]T_{ls}`$ and $`t[6,10]T_{ls}`$ the $`T_{ls}`$ time scale is essential for the evolution of the WP. It shows also that there exist a deep connection between the time behaviour of the autocorrelation function and the average values of the spin operators as discussed in the next subsection. ### 4.3 Expectation values of spin operators Expectation values of spin operators are derived directly from (4fsa-4fsd). One obtains: $$\sigma _x_t=\underset{l}{}w_l^2\left\{\frac{2ab}{2l+1}(G_{j_+}+2lG_{j_\pm }\mathrm{cos}\omega _lt)\frac{2ab}{2l+3}F_{j_+}\right\}+\delta \sigma _x,$$ (4fsuzaaa) $$\sigma _y_t=\underset{l}{}w_l^2\left\{2ab\frac{2l}{2l+1}G_{j_\pm }\mathrm{sin}\omega _lt\right\}+\delta \sigma _y,$$ (4fsuzaab) $`\sigma _z_t={\displaystyle \underset{l}{}}w_l^2\{a^2G_{j_+}+b^2{\displaystyle \frac{2l1}{(2l+1)^2}}(G_{j_+}2lG_j_{})b^2{\displaystyle \frac{8l}{(2l+1)^2}}G_{j_\pm }\mathrm{cos}\omega _lt`$ $`+b^2{\displaystyle \frac{2l}{2l+1}}F_j_{}(a^2{\displaystyle \frac{2l+1}{2l+3}}+b^2{\displaystyle \frac{2l1}{(2l+1)(2l+3)}})F_j_{}\},`$ (4fsuzaac) where $`\omega _l=(E_{j_+}E_j_{})/\mathrm{}`$. Terms denoted by $`\delta \sigma _x`$ and $`\delta \sigma _y`$, which are only small contributions, stem from coupling between states with $`l`$ or $`j`$ different by 2 units. Their explicit expressions are: $$\delta \sigma _x=\underset{l}{}K_l\mathrm{cos}\stackrel{~}{\omega }_lt,$$ (4fsuzaaaba) $$\delta \sigma _y=\underset{l}{}K_l\mathrm{sin}\stackrel{~}{\omega }_lt,$$ (4fsuzaaabb) where $$\mathrm{}\stackrel{~}{\omega }_l=E_{j_+}E_{(j+2)_{}},$$ (4fsuzaaabaca) $$F^{}=_0^{\mathrm{}}r^2f_{j+}f_{(j+2)_{}}𝑑r,$$ (4fsuzaaabacb) and $$K_l=2abw_lw_{l+2}F^{}\sqrt{\frac{(2l+2)(2l+4)}{(2l+3)(2l+5)}}.$$ (4fsuzaaabacc) In a non-relativistic limit the terms containing the radial parts $`f_{j_+}`$ and $`f_j_{}`$ cancel and the correction terms $`\delta \sigma _x`$ and $`\delta \sigma _y`$ tend to zero, the same for $`F_{j_+}`$ and $`F_j_{}`$. For the harmonic oscillator with a constant spin–orbit potential the radial integrals verify $$G_{j_+}=G_j_{}=1.$$ (4fsuzaaabacad) The dynamics of the spin motion in such a case has been discussed by us in and . Using the dispersion law appropriate to a constant spin–orbit potential $$\omega _l=(2l+1)\omega _{ls}$$ (4fsuzaaabacae) we have obtained there the periodic oscillations between spin and orbital angular momentum that we have called spin–orbit pendulum. In the present, more general case we see that both $`\sigma _x`$ and $`\sigma _z`$ contain an additional constant shift with respect to the non-relativistic expressions, coming clearly from the small components. In addition there are time dependent shifts due to the $`\delta \sigma _x`$ and $`\delta \sigma _y`$ terms. Numerically these shifts may be safely neglected. Therefore the most interesting dependence comes from the use of the eigenvalues of the Dirac hamiltonian in the time behaviour that we will now discuss in more details. In Figs. 5 and 6 we present the average values of the spin operators and the length of the spin vector as functions of time for cases $`N=20`$ and $`N=40`$ ($`Z=92`$). The latter case the most distinctly reminds both non-relativistic spin–orbit pendulum for the HO+LS Hamiltonian and the relativistic one for the DO . We clearly recognize time ranges for the spin collapse $`(0t2.5)`$, strong spin entanglement $`(2.5t4.5)`$ and spin revivals ($`6t10`$, all times in $`T_{ls}`$ units). For the case $`N=20`$, shown in Fig. 5, we see a similar behaviour of the spin averages, distorted, however, much more due to stronger interferences with $`T_4`$ characteristic time. In cases discussed above the ratio $`T_4/T_{ls}`$ is approximately equal to 1.9 for $`N=20`$ and 3.7 for $`N=40`$. ### 4.4 Spatial motion of circular WP Let us now discuss the motion of WP in space. A complex structure of the autocorrelation function and the existence of several time scales suggest that qualitatively similar spatial evolution can occur in different time scales. Below we focus our presentation of the probability density motion in two particular time scales: the shortest one, corresponding to the Kepler period of classical electron motion on the circular orbit with $`n=N`$, and the spin–orbit motion time scale. The admixture of the components $`|l,l1`$ in $`|c_1(t)`$ (4fsa) \[as shown earlier the small components give negligible contributions to $`\mathrm{\Psi }^{}\mathrm{\Psi }`$\] causes deviations of the WP motion from the equatorial plane, known already from our previous non-relativistic studies . The magnitude of these deviations decreases, however, when $`N`$ increases and for $`N1012`$ the maximum of the probability density moves very close to the equatorial plane, i.e. the plane of the classical orbit. Therefore we limit our presentation of $`[\mathrm{\Psi }^{}\mathrm{\Psi }](t)`$ to this plane only. For the short time scale, $`t[0,9/4]T_{cl}`$, it is shown in Fig. 7. The sequence of pictures exhibits the behaviour known from non-relativistic studies on WP in hydrogen atom , namely circular motion with spreading along the orbit and forming interference patterns that lead to fractional revivals. n Fig. 7 we display the full probability density $`\mathrm{\Psi }^{}\mathrm{\Psi }`$, not the spin up and spin down components separately, because for such short time scale no difference in their motion is visible. In order to see this difference one needs to observe WP in time scale comparable to $`T_{ls}`$. In Fig. 8 we present the sequence of shapes of the same WP ($`N=20,Z=92`$) for times $`t=0,1/2,1,3/2`$ and 2 (in $`T_{ls}`$ units). In these cases the presentation is divided into the two parts: in the left column spin up components of $`\mathrm{\Psi }^{}\mathrm{\Psi }`$ are displayed, whereas in the right column the spin down components are shown. At the presented time instants, that are multiples of $`T_{ls}/2`$ both sub-packets are quite well revived (localized) and move on a circle with different velocities. Fig. 9 shows (in the same convention) the shapes of WP’s components with spin up (left) and spin down (right) for time $`t=10.063545T_{ls}`$ ($`N=20,Z=92`$). As seen from Fig. 3 this time corresponds to the highest value of the autocorrelation function ($`|A(t)|^20.8`$) in the presented range of times. One sees from this figure that the degree of the recurrences (of both sub-packets separately) is indeed very high in this case. ## 5 Conclusions We have discussed the full 3+1–dimensional evolution of the relativistic wave packets in Coulomb potential. The main relativistic effect is the appearance of the new time scale due to the spin–orbit coupling. In general this effect has the same origin as spin–orbit pendulum, introduced by us both in non-relativistic and relativistic studies, where harmonic oscillator potential with a strong spin–orbit coupling was considered. In the Coulomb potential (relativistic) case presented here the time scale of the spin–orbit motion is large, therefore there is practically no chance for observing it in such systems like hydrogen or hydrogenoid atoms. However, this time scale decreases rapidly as function $`1/Z^2`$, when heavy ions with the single electron are considered as systems for experimental studies. At $`Z=82`$, and even better $`Z=92`$ the $`T_{ls}`$ time becomes reasonably short and in principle experimental observations of the spin–orbit effects in electron WP’s motion may become possible with existing techniques. This conclusion is based on the fact that the single electron heavy ions with $`Z=82,92`$ have been already created and some of their properties have been measured . It is also possible to suitably adjust the two parameters $`N,Z`$ in order to fit the energies of electron states and time scales of the motion to ranges of action of available lasers that have to be used for creation of wave packets and analysis of their motion. Comparing the phenomenon of the spin–orbit pendulum discussed in our previous studies with the present Coulomb potential case one sees the following differences. During their motion Coulomb wave packets experience several important time scales (only one is present in the HO case) that make the motion much more complex and obscure the spin–orbit motion. Relativistic (nonlinear) dependence of the eigenenergies on quantum numbers destroy exact periodicity of the motion. P.R. and M.T kindly acknowledge financial support (in part) of Polish Committee for Scientific Research (KBN) under the grant 2 P03B 143 14. ## References
warning/0001/cond-mat0001303.html
ar5iv
text
# Density Wave States of Non-Zero Angular Momentum ## I Introduction In recent years, a number of materials have been uncovered in which the competition between an effective attractive interaction and short-range repulsion appears to lead to the formation of superconducting states in which the Cooper pairs have non-zero relative angular momentum. In this paper, we suggest that such competition can also lead to density-wave states formed by the condensation of particle-hole pairs of non-zero relative angular momentum. These states generalize the familiar charge- and spin-density-wave states, in which $`s`$-wave particle-hole pairs condense. We discuss several different possible ordering schemes, the types of interactions which favor them, their physical properties, and their possible relevance to experiments. Several such states are already commonly known by other names, as we will show below. The singlet $`l=1`$ density-wave state is simply the Peierls state (or bond-ordered wave), while the singlet $`l=2`$ density-wave state is known as the staggered flux state of . However, the triplet analogues of these states have not been discussed. Since the triplet analogues of these states break spin-rotational invariance, they have $`S=1`$ Goldstone boson excitations. However, the ground state does not have a non-zero expectation value for the spin at any wavevector. Hence, as we will see, these Goldstone bosons cannot be detected in experiments which couple simply to the spin density, such as neutron scattering or NMR. Instead, Raman scattering or NQR are necessary to couple to the Goldstone bosons of these more subtle types of ordering. More generaly, $`s`$-wave probes cannot couple directly to the orders discussed here; instead, local probes or those which couple to higher powers of the order parameter are necessary. The $`p`$-wave and $`d`$-wave density wave states are favored by the same types of interactions which favor the $`s`$-wave state – i.e. the CDW. However, they evade interactions which disfavor CDW order. Similarly, they are favored by superconductivity-favoring pair-hopping terms while evading interactions which disfavor superconductivity. Hence, they are rather natural candidates for systems with competing repulsive interactions. As in the case of higher angular momentum superconducting states, there is the possibility of gapless excitations since the order parameter can have nodes on the Fermi surface. To consider one consequence of this, suppose that the shape of the Fermi surface is such that the nodes of the order parameter do not lie on the Fermi surface. Let us distort the shape of the Fermi surface by, say, changing the anisotropy between the hopping parameters, which can be done by applying uniaxial pressure. At the mean-field level, a third-order phase transition can occur at which gapless excitations appear. After this point, the system remains critical as a result of the node. The analogy with supercondutivity can be taken a step further by combining density wave order with superconducting order in a pseudospin $`SU(2)`$ triplet following Yang and Zhang . At a critical point between the two types of order, this pseudospin $`SU(2)`$ could become exact, giving – together with $`SU(2)`$ spin symmetry – an $`O(4)`$-invariant critical point. We discuss the possible relevance of such a critical point to the pseudogap regime of the cuprate superconductors. Particle-hole condensates with non-zero angular momentum were considered in the context of excitonic insulators by Halperin and Rice . They were rediscovered in the context of the mean-field instabilities of of extended Hubbard models by Schulz and Nersesyan and collaborators. At around the same time, Kotliar and Marston and Affleck found the staggered flux state as a mean-field solution of the Hubbard model. However, it was apparently not recognized that the singlet $`d_{x^2y^2}`$ density-wave state is the same as the staggered flux state. More recently, this state and a related variant have been discussed in the context of the cuprate superconductors. A version of this state (see the comments in the concluding section) has appeared in mean field analyses of an $`SU(2)`$ mean field theory of the $`tJ`$ model . The Nodal Liquid state of also bears a family resemblance to the staggered flux state; we will return to the relationship between these states in the concluding section. ## II Order Parameters and Broken Symmetries We define the different possible density-wave orderings by analogy with the more familiar superconducting case. Consider a system of electrons on a square lattice of side $`a`$. A superconductor is defined by a non-vanishing expectation value of $$\psi _\alpha (k,t)\psi _\beta (k,t)$$ (1) A triplet superconductor is characterized by the expectation value $$\psi _\alpha (k,t)\psi _\beta (k,t)=\stackrel{}{\mathrm{\Delta }}(p)\stackrel{}{\sigma }_\alpha ^\gamma ϵ_{\gamma \beta }$$ (2) Fermi statistics requires that $`\stackrel{}{\mathrm{\Delta }}(p)`$ be odd in $`\stackrel{}{p}`$. $`p`$-wave superconductors can have the components of $`\stackrel{}{\mathrm{\Delta }}(p)`$ chosen from $`\mathrm{sin}k_xa`$, $`\mathrm{sin}k_ya`$, or $`\mathrm{sin}k_xa\pm i\mathrm{sin}k_ya`$. For instance, a $`p_x`$ superconductor with all spins polarized along the $`3`$-direction will have $`\mathrm{\Delta }_1+i\mathrm{\Delta }_20`$ and $`\mathrm{\Delta }_3=\mathrm{\Delta }_1i\mathrm{\Delta }_2=0`$: $$\psi _\alpha (k,t)\psi _\beta (k,t)=\mathrm{\Delta }_0\left(\mathrm{sin}k_xa\right)\sigma _\alpha ^{+\gamma }ϵ_{\gamma \beta }$$ (3) Spin-polarized $`p_y`$ and $`p_x+ip_y`$ superconductors have $`\mathrm{sin}k_xa`$ replaced, repectively, by $`\mathrm{sin}k_ya`$ and $`\mathrm{sin}k_xa+i\mathrm{sin}k_ya`$. The analog of the $`A^{}`$ phase of <sup>3</sup>He has equal numbers of $``$ and $``$ pairs: $`\psi _\alpha (k,t)\psi _\beta (k,t)`$ $`=`$ (5) $`\mathrm{\Delta }_0\left(\mathrm{sin}k_xa\sigma _\alpha ^{1\gamma }+\mathrm{sin}k_ya\sigma _\alpha ^{2\gamma }\right)ϵ_{\gamma \beta }`$ An unpolarized $`p_x`$ superconductor of $``$ pairs has $`\mathrm{\Delta }_30`$ and $`\mathrm{\Delta }_1=\mathrm{\Delta }_2=0`$: $$\psi _\alpha (k,t)\psi _\beta (k,t)=\mathrm{\Delta }_0\left(\mathrm{sin}k_xa\right)\sigma _\alpha ^{3\gamma }ϵ_{\gamma \beta }$$ (6) As in the polarized case, unpolarized $`p_y`$ and $`p_x+ip_y`$ superconductors have $`\mathrm{sin}k_xa`$ replaced, repectively, by $`\mathrm{sin}k_ya`$ and $`\mathrm{sin}k_xa+i\mathrm{sin}k_ya`$. In principle, more complicated order parameters are possible, with all components of $`\stackrel{}{\mathrm{\Delta }}`$ taking non-vanishing values. If any component of $`\stackrel{}{\mathrm{\Delta }}(p)`$ is not real, time-reversal symmetry ($`T`$) is broken. A $`d`$-wave superconductor must be a spin-singlet superconductor. A $`d_{x^2y^2}`$ superconductor has $$\psi _\alpha (k,t)\psi _\beta (k,t)=\mathrm{\Delta }_0\left(\mathrm{cos}k_xa\mathrm{cos}k_ya\right)ϵ_{\alpha \beta }$$ (7) while a $`d_{xy}`$ superconductor has $`\mathrm{cos}k_xa\mathrm{cos}k_ya`$ replaced by $`\mathrm{sin}k_xa\mathrm{sin}k_ya`$. A $`d_{x^2y^2}+id_{xy}`$ superconductor breaks $`T`$ with the order parameter: $`\psi _\alpha (k,t)\psi _\beta (k,t)=`$ (8) $`\mathrm{\Delta }_0\left(\mathrm{cos}k_xa\mathrm{cos}k_ya+i\mathrm{sin}k_xa\mathrm{sin}k_ya\right)ϵ_{\alpha \beta }`$ (9) We can define analogous orders for density-wave states. However, the spin structures will no longer be determined by Fermi statistics. Let us first consider the singlet orderings. A singlet $`s`$-wave density wave is simply a charge-density-wave<sup>*</sup><sup>*</sup>*Extended $`s`$-wave is also possible.: $$\psi ^\alpha (k+Q,t)\psi _\beta (k,t)=\mathrm{\Phi }_Q\delta _\beta ^\alpha $$ (10) A singlet $`p_x`$ density-wave state has ordering $$\psi ^\alpha (k+Q,t)\psi _\beta (k,t)=\mathrm{\Phi }_Q\mathrm{sin}k_xa\delta _\beta ^\alpha $$ (11) The singlet $`p_x+ip_y`$ density-wave states are defined by: $$\psi ^\alpha (k+Q,t)\psi _\beta (k,t)=\mathrm{\Phi }_Q\left(\mathrm{sin}k_xa+i\mathrm{sin}k_ya\right)\delta _\beta ^\alpha $$ (12) Similarly, the singlet $`d_{x^2y^2}`$ density-wave states have $$\psi ^\alpha (k+Q,t)\psi _\beta (k,t)=\mathrm{\Phi }_Q\left(\mathrm{cos}k_xa\mathrm{cos}k_ya\right)\delta _\beta ^\alpha $$ (13) while the singlet $`d_{x^2y^2}+id_{xy}`$ density-wave states have $`\psi ^\alpha (k+Q,t)\psi _\beta (k,t)=`$ (14) $`\mathrm{\Phi }_Q\left(\mathrm{cos}k_xa\mathrm{cos}k_ya+i\mathrm{sin}k_xa\mathrm{sin}k_ya\right)\delta _\beta ^\alpha `$ (15) These states belong to a class of states of the form: $$\psi ^\alpha (k+Q,t)\psi _\beta (k,t)=\mathrm{\Phi }_Qf(k)\delta _\beta ^\alpha $$ (16) $`f(k)`$ is an element of some representation of the space group of the vector $`\stackrel{}{Q}`$ in the square lattice. In this paper, we will focus primarily on the cases $`f(k)=\mathrm{sin}k_xa`$ and $`f(k)=\mathrm{cos}k_xa\mathrm{cos}k_ya`$, but $`f(k)`$ could be an element of some larger representation. The $`s`$-wave (or extended $`s`$-wave) cases, $`f(k)=|f(k)|`$, are the usual charge-density wave states. $`Q`$ is the wavevector at which the density-wave ordering takes place. It may be commensurate or incommensurateIn this paper, we will take commensurate to mean the situation in which $`2\stackrel{}{Q}`$ is a reciprocal lattice vector. The term ‘incommensurate’ will actually include higher-order commensurability.. For commensurate ordering such that $`2Q`$ is a reciprocal lattice vector, e.g. $`Q=(\pi /a,0)`$ or $`Q=(\pi /a,\pi /a)`$, we can take the hermitian conjugate of the order parameter: $`\psi ^\beta (k,t)\psi _\alpha (k+Q,t)`$ $`=`$ $`\mathrm{\Phi }_Q^{}f^{}(k)\delta _\beta ^\alpha `$ (17) $`\psi ^\beta (k+Q+Q,t)\psi _\alpha (k+Q,t)`$ $`=`$ $`\mathrm{\Phi }_Q^{}f^{}(k)\delta _\beta ^\alpha `$ (18) $`\mathrm{\Phi }_Qf(k+Q)\delta _\beta ^\alpha `$ $`=`$ $`\mathrm{\Phi }_Q^{}f^{}(k)\delta _\beta ^\alpha `$ (19) Therefore, for $`Q`$ commensurate $$\frac{f(k+Q)}{f^{}(k)}=\frac{\mathrm{\Phi }_Q^{}}{\mathrm{\Phi }_Q}$$ (20) Hence, if $`f(k+Q)=f^{}(k)`$, $`\mathrm{\Phi }_Q`$ must be imaginary. For singlet $`p_x`$ ordering, this will be the case if $`Q=(\pi /a,0)`$ or $`Q=(\pi /a,\pi /a)`$. For singlet $`d_{x^2y^2}`$ ordering, this will be the case if $`Q=(\pi /a,\pi /a)`$. If $`f(k+Q)=f^{}(k)`$, $`\mathrm{\Phi }_Q`$ must be real. For singlet $`p_x`$ ordering, this will be the case if $`Q=(0,\pi /a)`$. For singlet $`d_{xy}`$ ordering, this will be the case if $`Q=(\pi /a,\pi /a)`$. For incommensurate ordering, $`\mathrm{\Phi }_Q`$ can have arbitrary phase: the phase of $`\mathrm{\Phi }_Q`$ is the Goldstone boson of broken translational invariance, i.e. the sliding density-wave mode. Impurities will pin this mode – at second order in the impurity potential, as in the case of a spin-density-wave – so we will not consider it further. All of these states break translational and rotational invariance. To further analyze the symmetries of these states, it is instructive to write these orderings in real space. The singlet $`p_x`$ density-waves have non-vanishing expectation value: $`\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}+a\widehat{x},t)\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}a\widehat{x},t)=`$ (21) $`\mathrm{}{\displaystyle \frac{i}{2}}\left(\mathrm{\Phi }_Qe^{i\stackrel{}{Q}\stackrel{}{x}}+\mathrm{\Phi }_Qe^{i\stackrel{}{Q}\stackrel{}{x}}\right)\delta _\beta ^\alpha `$ (22) We have only written the modulated term; the $`\mathrm{}`$ refers to the uniform contribution coming from the Fourier transform of $`\psi ^{}(k)\psi (k)`$. Let us consider the commensurate and incommensurate cases separately. The incommensurate singlet $`p_x`$ density-wave states completely break the translational and rotational symmetries. If $`\mathrm{\Phi }_Q=\mathrm{\Phi }_Q^{}`$, $`T`$ is preserved; otherwise, it is broken. The singlet $`p_x+ip_y`$ density-wave states always break $`T`$. The commensurate states, on the other hand, break translation by one lattice spacing; translation by two lattice spacings is preserved. From (20), a commensurate singlet $`p_x`$ density-wave state with $`Q=(\pi /a,0)`$ must have imaginary $`\mathrm{\Phi }_Q`$: $`\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}+a\widehat{x},t)\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}a\widehat{x},t)=`$ (23) $`\mathrm{}+\left|\mathrm{\Phi }_Q\right|e^{i\stackrel{}{Q}\stackrel{}{x}}\delta _\beta ^\alpha `$ (24) The singlet state of this type breaks no other symmetries; it is usually called the Peierls state or bond order wave. If $`Q=(0,\pi /a)`$, $`\mathrm{\Phi }_Q`$ must be real. $`\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}+a\widehat{x},t)\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}a\widehat{x},t)=`$ (25) $`\mathrm{}i\left|\mathrm{\Phi }_Q\right|e^{i\stackrel{}{Q}\stackrel{}{x}}\delta _\beta ^\alpha `$ (26) As a result of the $`i`$, the $`Q=(0,\pi /a)`$ singlet $`p_x`$ density-wave states break $`T`$. However, the combination of $`T`$ and translation by an odd number of lattice spacings remains unbroken. The same is true of the commensurate singlet $`p_x+ip_y`$ density-wave states. Examples of commensurate and incommensurate singlet $`p_x`$ and $`p_x+ip_y`$ density-wave states are depicted in figure 1. The singlet $`d_{x^2y^2}`$ density-wave states have non-vanishing expectation value: $`\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}+a\widehat{x},t)+\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}a\widehat{x},t)`$ (27) $`\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}+a\widehat{y},t)+\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}a\widehat{y},t)=`$ (28) $`\mathrm{}+{\displaystyle \frac{1}{2}}\left(\mathrm{\Phi }_Qe^{i\stackrel{}{Q}\stackrel{}{x}}+\mathrm{\Phi }_Qe^{i\stackrel{}{Q}\stackrel{}{x}}\right)\delta _\beta ^\alpha `$ (29) The incommensurate singlet $`d_{x^2y^2}`$ density-wave states will preserve $`T`$ if $`\mathrm{\Phi }_Q=\mathrm{\Phi }_Q^{}`$; otherwise, they break $`T`$. The same is true of the incommensurate singlet $`d_{xy}`$ density-wave states: $`\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}+a\widehat{x}+a\widehat{y},t)+`$ (30) $`\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}a\widehat{x}a\widehat{y},t)`$ (31) $`\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}a\widehat{x}+a\widehat{y},t)`$ (32) $`\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}+a\widehat{x}a\widehat{y},t)=`$ (33) $`\mathrm{}{\displaystyle \frac{1}{4}}\left(\mathrm{\Phi }_Qe^{i\stackrel{}{Q}\stackrel{}{x}}+\mathrm{\Phi }_Qe^{i\stackrel{}{Q}\stackrel{}{x}}\right)\delta _\beta ^\alpha `$ (34) Incommensurate singlet $`d_{x^2y^2}+id_{xy}`$ density-wave states necessarily break $`T`$: $`\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}+a\widehat{x},t)+\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}a\widehat{x},t)`$ (35) $`\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}+a\widehat{y},t)+\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}a\widehat{y},t)`$ (36) $`+i\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}+a\widehat{x}+a\widehat{y},t)`$ (37) $`+i\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}a\widehat{x}a\widehat{y},t)`$ (38) $`i\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}a\widehat{x}+a\widehat{y},t)`$ (39) $`i\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}+a\widehat{x}a\widehat{y},t)=`$ (40) $`\mathrm{}+\left({\displaystyle \frac{1}{2}}{\displaystyle \frac{i}{4}}\right)\left(\mathrm{\Phi }_Qe^{i\stackrel{}{Q}\stackrel{}{x}}+\mathrm{\Phi }_Qe^{i\stackrel{}{Q}\stackrel{}{x}}\right)\delta _\beta ^\alpha `$ (41) The commensurate $`Q=(\pi /a,\pi /a)`$ singlet $`d_{x^2y^2}`$ density-wave states must have imaginary $`\mathrm{\Phi }_Q`$: $`\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}+a\widehat{x},t)+\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}a\widehat{x},t)`$ (42) $`\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}+a\widehat{y},t)+\psi ^\alpha (\stackrel{}{x},t)\psi _\beta (\stackrel{}{x}a\widehat{y},t)=`$ (43) $`\mathrm{}+{\displaystyle \frac{i}{2}}\left|\mathrm{\Phi }_Q\right|e^{i\stackrel{}{Q}\stackrel{}{x}}\delta _\beta ^\alpha `$ (44) As a result of the $`i`$, the singlet $`d_{x^2y^2}`$ density-wave breaks $`T`$ as well as translational and rotational invariance. The combination of time-reversal and a translation by one lattice spacing is preserved by this ordering. The commensurate $`Q=(\pi /a,\pi /a)`$ singlet $`d_{x^2y^2}`$ density-wave state is often called the staggered flux state. There is also a contribution to this correlation function coming from $`\psi ^{}(k)\psi (k)`$ which is uniform in space (the $`\mathrm{}`$); as a result, the phase of the above bond correlation function – and, therefore, the flux through each plaquette – is alternating. The commensurate $`Q=(\pi /a,\pi /a)`$ singlet $`d_{xy}`$ must have real $`\mathrm{\Phi }_Q`$; therefore, it does not break $`T`$. On the other hand, the singlet $`d_{x^2y^2}+id_{xy}`$ state does break $`T`$. Note that the nodeless commensurate singlet $`d_{x^2y^2}+id_{xy}`$ density-wave state does not break more symmetries than the commensurate singlet $`d_{x^2y^2}`$ density-wave state, in contrast to the superconducting case. Examples of singlet $`d_{x^2y^2}`$, $`d_{xy}`$, and $`d_{x^2y^2}+id_{xy}`$ density-wave states are depicted in figure 2. We now consider the triplet density-wave states. Triplet states all break spin-rotational invariance and, therefore, have Goldstone boson excitations. We will discuss the experimental consequences of these Goldstone bosons later. The triplet $`s`$-wave density wave state is simply a spin-density-wave. The triplet $`p`$-wave and $`d`$-wave states are characterized by: $$\psi ^\alpha (k+Q,t)\psi _\beta (k,t)=\stackrel{}{\mathrm{\Phi }}_Q(k)\stackrel{}{\sigma }_\beta ^\alpha $$ (45) with the components of $`\stackrel{}{\mathrm{\Phi }}_Q(k)`$ chosen from, respectively, $`\mathrm{sin}k_xa`$, $`\mathrm{sin}k_ya`$, $`\mathrm{sin}k_xa\pm i\mathrm{sin}k_ya`$; and $`\mathrm{cos}k_xa\mathrm{cos}k_ya`$, $`\mathrm{sin}k_xa\mathrm{sin}k_ya`$, $`\mathrm{cos}k_xa\mathrm{cos}k_ya\pm i\mathrm{sin}k_xa\mathrm{sin}k_ya`$. A state in which the particle-hole pairs are polarized, which is the most direct analogue of a spin-density-wave has $`\stackrel{}{\mathrm{\Phi }}_Q(p)`$ of the form $`\mathrm{\Phi }_Q^30`$, $`\mathrm{\Phi }_Q^1=\mathrm{\Phi }_Q^2=0`$: $$\psi ^\alpha (k+Q,t)\psi _\beta (k,t)=\mathrm{\Phi }_Qf(k)\sigma _\beta ^{3\alpha }$$ (46) where $`f(k)`$ is chosen from the above set. Alternatively, the particle-hole pairs can be unpolarized, e.g. $$\psi ^\alpha (k+Q,t)\psi _\beta (k,t)=\mathrm{\Phi }_Q\left(\mathrm{sin}k_xa\sigma _\beta ^{1\alpha }+i\mathrm{sin}k_ya\sigma _\beta ^{2\alpha }\right)$$ (47) As in the superconducting case, more complicated order parameters are possible, with all components of $`\stackrel{}{\mathrm{\Delta }}`$ taking non-vanishing values. For commensurate ordering, we can follow the same logic as in (17). The phases of the components of $`\stackrel{}{\mathrm{\Phi }}_Q(p)`$ are constrained in the same way as the singlet order parameters, as illustrated by the $`i`$ in front of the second term in (47). The orders discussed here can be generalized to other $`2D`$ lattices and to $`3D`$ lattices. The orbital wavefunctions $`\mathrm{sin}k_xa`$, etc. will be replaced by representations of the point groups of these other lattices. To each $`T`$-preserving singlet ordering, we can associate an ordering of a pure spin model, in the same way that spin-Peierls ordering is related to Peierls ordering: $$\psi ^\alpha (k+Q,t)\psi _\alpha (k,t)\stackrel{}{S}(k+Q,t)\stackrel{}{S}(k,t)$$ (48) These spin orderings are states in which the exchange energies are large along preferred directions. These preferred diractions oscillate from one lattice point to the next with spatial frequency $`\stackrel{}{Q}`$. The simplest case is spin-Peierls ordering, in which the spins form dimers. Another example is $`d_{xy}`$ ordering of a spin model, which takes the form: $`\stackrel{}{S}(\stackrel{}{x},t)\stackrel{}{S}(x+a\widehat{x}+a\widehat{y},t)+\stackrel{}{S}(\stackrel{}{x},t)\stackrel{}{S}(xa\widehat{x}a\widehat{y},t)`$ $`\stackrel{}{S}(\stackrel{}{x},t)\stackrel{}{S}(xa\widehat{x}+a\widehat{y},t)\stackrel{}{S}(\stackrel{}{x},t)\stackrel{}{S}(x+a\widehat{x}a\widehat{y},t)`$ $`={\displaystyle \frac{1}{4}}\left(\mathrm{\Phi }_Qe^{i\stackrel{}{Q}\stackrel{}{x}}+\mathrm{\Phi }_Qe^{i\stackrel{}{Q}\stackrel{}{x}}\right)`$ in analogy with (30). ## III Model Hamiltonians We are primarily concerned in this paper with the universal properties of the states introduced above. We will not attempt to show that particular realistic models of interacting electrons have $`p`$\- or $`d`$-wave density-wave ground states. Rather, we will content ourselves with discussing the types of interactions which favor such orders and showing that they lead to energetically favorable trial variational wavefunctions for some idealized Hamiltonians. The analog of the BCS reduced Hamiltonian for singlet density-wave order is: $`H`$ $`=`$ $`{\displaystyle \frac{d^2k}{(2\pi )^2}ϵ(k)\psi ^\alpha (k)\psi _\alpha (k)}`$ (51) $`g{\displaystyle }{\displaystyle \frac{d^2k}{(2\pi )^2}}{\displaystyle \frac{d^2k^{}}{(2\pi )^2}}[f(k)f(k^{})\times `$ $`\psi ^\alpha (k+Q)\psi _\alpha (k)\psi ^\beta (k^{})\psi _\beta (k^{}+Q)]`$ In the triplet case, we replace the four-fermion operator of (51) by $`\psi ^\alpha (k+Q)\sigma _\alpha ^{a\beta }\psi _\beta (k)\psi ^\gamma (k^{})\sigma _\gamma ^{a\delta }\psi _\delta (k^{}+Q)`$ (52) We now introduce the variational wavefunction $`|\mathrm{\Psi }={\displaystyle \underset{k,\alpha }{}}\left(u_{k,\alpha }\psi ^\alpha (k)+v_{k,\alpha }\psi ^\alpha (k+Q)\right)|0`$ (53) Its energy can be minimized if we take $$\overline{u}_{k,\alpha }v_{k,\alpha }=\frac{g\mathrm{\Phi }_Qf(k)}{\sqrt{\left(ϵ(k)ϵ(k+Q)\right)^2+4g^2\left|\mathrm{\Phi }_Q\right|^2\left(f(k)\right)^2}}$$ (54) in the singlet case and $$\overline{u}_{k,\alpha }\stackrel{}{\sigma }_\alpha ^\beta v_{k,\beta }=\frac{g\stackrel{}{\mathrm{\Phi }}_Qf(k)}{\sqrt{\left(ϵ(k)ϵ(k+Q)\right)^2+4g^2\left|\mathrm{\Phi }_Q\right|^2\left(f(k)\right)^2}}$$ (55) in the triplet case, and require $`\mathrm{\Phi }_Q`$ to satisfy the gap equation: $$g\frac{d^2k}{(2\pi )^2}\frac{\left(f(k)\right)^2}{\sqrt{\left(ϵ(k)ϵ(k+Q)\right)^2+4g^2\left|\mathrm{\Phi }_Q\right|^2\left(f(k)\right)^2}}=1$$ (56) The reduced Hamiltonian has long-ranged interactions, so the variational wavefunction is essentially correct. We will now show that short-ranged Hamiltonians will include terms of the form (51), and that the trial wavefunction (53) is reasonable for these short-ranged Hamiltonians. Consider, then, the following lattice model of interacting electrons: $`H`$ $`=`$ $`t{\displaystyle \underset{<i,j>}{}}(c_{i\sigma }^{}c_{j\sigma }+\mathrm{h}.\mathrm{c}.)+U{\displaystyle \underset{i}{}}n_in_i`$ (62) $`+V{\displaystyle \underset{<i,j>}{}}n_in_j`$ $`t_{c1}{\displaystyle \underset{<i,j>,<i^{},j>,ii^{}}{}}c_{i\sigma }^{}c_{j\sigma }c_{j\sigma }^{}c_{i^{}\sigma }`$ $`t_{c2}{\displaystyle \underset{i}{}}[(c_{i+\widehat{x},\sigma }^{}c_{i\sigma }c_{i\sigma }^{}c_{i+\widehat{x},\sigma })\times `$ $`\left(c_{i+\widehat{x}+\widehat{y},\sigma }^{}c_{i+\widehat{y},\sigma }c_{i+\widehat{y},\sigma }^{}c_{i+\widehat{x}+\widehat{y},\sigma }\right)`$ $`+xy]`$ The first two terms are the usual hopping, $`t`$, and on-site repulsion, $`U`$ of the Hubbard model. The third term is a nearest-neighbor repulsion, $`V`$. The third and fourth terms lead to the correlated motion of pairs of electrons. $`t_{c1}`$ hops an electron from $`i^{}`$ to $`j`$ when $`j`$ is vacated by an electron hopping to $`i`$. $`t_{c2}`$ hops nearest-neighbor pairs in the same direction. Terms of this general form have been discussed in as a mechanism for superconductivity. As we will see below, they not only favor superconductivity, but $`p`$\- and $`d`$-wave density wave order as well. Fourier transforming the interaction terms into momentum space, we see that terms of the form of the reduced interaction (51) are, indeed, present: $`H_{\mathrm{int}}`$ $`=`$ $`{\displaystyle }{\displaystyle \frac{d^2k}{(2\pi )^2}}[U\psi ^{}(k_1)\psi _{}(k_2)\psi ^{}(k_3)\psi _{}(k_4)`$ (68) $`+(2V(\mathrm{cos}(k_3^xk_4^x)a+\mathrm{cos}(k_3^yk_4^y)a)`$ $`\mathrm{\hspace{0.17em}2}t_{c1}(\mathrm{cos}(k_1^xk_4^x)a+\mathrm{cos}(k_1^yk_4^y)a`$ $`+2\mathrm{cos}k_1^xa\mathrm{cos}k_4^xa+2\mathrm{cos}k_1^ya\mathrm{cos}k_4^ya)`$ $`\mathrm{\hspace{0.17em}2}t_{c2}(\mathrm{sin}k_1^xa\mathrm{sin}k_4^xa+\mathrm{sin}k_1^ya\mathrm{sin}k_4^ya))`$ $`\times \psi ^\alpha (k_1)\psi _\alpha (k_2)\psi ^\beta (k_3)\psi _\beta (k_4)]`$ Let us now consider various candidate orderings and the terms which favor or penalize them. Antiferromagnetic order is favored by $`U`$ but penalized by $`V`$. Charge-density-wave order is favored by $`V`$ but penalized by $`U`$. $`p`$ and $`d`$-wave superconductivity are favored by $`t_{c2}`$ and $`t_{c1}`$ respectively and penalized by $`V`$. $`p`$ and $`d`$-density-wave order are favored by $`t_{c2}`$ and $`t_{c1}`$, respectively, and are both favored by $`V`$. The density-wave states can be favored over the others by taking $`V`$ large. The $`p`$\- or $`d`$-wave states can be favored by taking $`t_{c2}`$ or $`t_{c1}`$ large. To be more precise, the mean-field equations for various ordered states read: $$\lambda \frac{d^2k}{(2\pi )^2}\frac{\left(f(k)\right)^2}{\sqrt{\left(ϵ(k)ϵ(k+Q)\right)^2+4\lambda ^2\left|\mathrm{\Phi }_Q\right|^2\left(f(k)\right)^2}}=1$$ (70) where $`\lambda _{dDW}`$ $`=`$ $`8V+96t_{c1}`$ (71) $`\lambda _{pDW}`$ $`=`$ $`4V+16t_{c1}+16t_{c2}`$ (72) $`\lambda _{CDW}`$ $`=`$ $`16V+96t_{c1}+16t_{c2}2U`$ (73) $`\lambda _{AF}`$ $`=`$ $`2U`$ (74) Hence, the singlet $`d_{x^2y^2}`$ density-wave state will be the ground state if $`8t_{c1}<U4V<48t_{c1}`$ (75) $`8t_{c2}<2V+40t_{c1}`$ (76) while the singlet $`p_x`$ density-wave state will be the ground state if $`6V+40t_{c1}<U<2V+8t_{c1}+8t_{c2}`$ (78) $`2V+40t_{c1}<8t_{c2}`$ (79) assuming that the van Hove singularities are at the antinodes of the order parameters. Otherwise, the $`p`$\- and $`d`$-wave density wave states will be favored over somewhat smaller regions of parameter space. By including spin-dependent interactions such as $`J\stackrel{}{S}_i\stackrel{}{S}_i`$, we can favor the triplet $`p`$\- or $`d`$-wave states. Hence, it appears that the orderings discussed in this paper are viable. The detailed energetics at large coupling strengths – which surely hold in physically interesting systems – are beyond the scope of this paper. ## IV Experimental Signatures We now turn to the question of the experimental signatures of such states. Since the order parameter changes sign as the Fermi surface is circled, there is no net CDW or SDW order which could be measured in, for instance, neutron scatteringOne can expect, on general grounds, that incommensurate singlet or triplet $`p`$\- or $`d`$-wave density-wave order at wavevector $`\stackrel{}{Q}`$ will induce CDW order at $`2\stackrel{}{Q}`$ since a term of the form $`\mathrm{\Phi }_Q^2\rho _{2Q}`$ or $`\stackrel{}{\mathrm{\Phi }}_Q\stackrel{}{\mathrm{\Phi }}_Q\rho _{2Q}`$ is allowed by symmetry in the effective action. Nevertheless, we may wish to distinguish such a state from one which has only CDW order.. When another symmetry – in addition to translational invariance – is broken, this is easier. Let’s first consider broken time-reversal symmetry. The commensurate singlet $`d_{x^2y^2}`$ density wave state – or staggered flux state – breaks $`T`$; there is an alternating pattern of currents circulating about each plaquette of the lattice. These currents produce an alternating magnetic field measurable by $`\mu `$SR and, in principle, by neutron scattering . The magnitude of the current along a link of the lattice will be: $$j=\frac{et}{\mathrm{}}\mathrm{\Phi }_Q10^5\mathrm{Ampere}\times \mathrm{\Phi }_Q$$ (81) Now, $`\mathrm{\Phi }_Q`$ is related to the maximum of the gap according to $`\mathrm{\Delta }_0=g\mathrm{\Phi }_Q`$ where $`g`$ is the appropriate coupling constant. Let us suppose, for the purposes of illustration, that the formation of the ordered state is driven by $`\lambda _{dDW}`$. Then, for $`\lambda _{dDW}`$ small, $`\mathrm{\Phi }_Q(t/\lambda _{dDW})e^{(\mathrm{const}.)t/\lambda _{dDW}}`$. Alternatively, we may take the high-$`T_c`$ context as a guideline: observed gaps are $`100300K`$, while interactions such are $`1eV`$. In this case, we expect $`\mathrm{\Phi }_Q10^2`$. This translates to a magnetic field at the center of each plaquette on the order of $`10G`$. The muons in a $`\mu SR`$ experiment might see a lowwer field if they sit at points of high symmetry or away from the plane. The orbital magnetic moments are likely to be dwarfed by local spin moments . Incommensurate ordering may or may not break $`T`$; if it does, the above analysis applies. In the $`\mathrm{\Phi }_Q^30`$, $`\mathrm{\Phi }_Q^1=\mathrm{\Phi }_Q^2=0`$ triplet $`d_{x^2y^2}`$ density wave state, there are counter-circulating currents of up- and down-spin electrons. These currents cancel, so there is no net current circulating about each plaquette, but there is an alternating pattern of spin currents circulating about each plaquette. The checkerboard pattern of spin currents will generate, via the spin-orbit coupling, $`H_{SO}`$ $`=`$ $`{\displaystyle }{\displaystyle \frac{d^2k}{(2\pi )^2}}{\displaystyle \frac{d^2q}{(2\pi )^2}}\stackrel{}{E}(q)(2\stackrel{}{k}+\stackrel{}{q})\times `$ (83) $`\psi ^\alpha (k+q)\stackrel{}{\sigma }_\alpha ^\beta \psi _\beta (k)`$ a quadrupolar electric field which is, in principle, measurable in NQR experiments. With the above estimate of the current, a nucleus with a non-zero quadrupole moment would have an induced splitting of order $`10Hz`$. We now turn to broken spin-rotational invariance, characteristic of the triplet states. Since it transforms non-trivially under the point group of the square lattice, the triplet order parameter $`\stackrel{}{\mathrm{\Phi }}_Q`$ will not couple to photons, neutrons, or nuclear spins according to $`\stackrel{}{\mathrm{\Phi }}_Q\stackrel{}{F}`$, where $`\stackrel{}{F}`$ is, respectively, $`\stackrel{}{B}`$, $`\stackrel{}{S}_N`$, or $`\stackrel{}{I}`$. Said more physically, the triplet ordered states do not have anomalous expectation values for the spin density but, rather, for spin currents; spin currents do not couple simply to these probes. However, the order parameter $`\stackrel{}{\mathrm{\Phi }}_Q`$ will couple to such probes at second order since its square transforms trivially under the point group. Such a coupling will be of the form: $$H_{\mathrm{probe}}=d^2x\left[2\left(\stackrel{}{F}\stackrel{}{\mathrm{\Phi }}_Q\right)\left(\stackrel{}{F}\stackrel{}{\mathrm{\Phi }}_Q^{}\right)\left|\stackrel{}{\mathrm{\Phi }}_Q\right|^2F^2\right]$$ (84) In the case of photons, this will lead to $`2`$-magnon Raman scattering. The coupling to nuclear spins couples directly to the nuclear quadrupole moment, and will lead to a shift in the nuclear quadrupole resonance frequencies. In the presence of disorder, rotational symmetry will be broken. Hence, there will be a small coupling, proportional to the disorder strength, of the Goldstone bosons to s-wave probes such as NMR and neutron scattering. ## V Gapless Fermionic Excitations The mean-field Hamiltonian is: $`H`$ $`=`$ $`{\displaystyle _{\mathrm{B}.\mathrm{Z}.}}{\displaystyle \frac{d^2k}{(2\pi )^2}}[ϵ(k)\psi ^\alpha (k)\psi _\alpha (k)+`$ (86) $`g\mathrm{\Phi }_Qf(k)\psi ^\alpha (k+Q)\psi _\alpha (k)]`$ If we define the four component object $`\chi _{A\alpha }`$ according to $`\left(\begin{array}{c}\chi _{1\alpha }\\ \chi _{2\alpha }\end{array}\right)=\left(\begin{array}{c}\psi _\alpha (k)\\ \psi _\alpha (k+Q)\end{array}\right)`$ (91) then the mean-field Hamiltonian can be written in the form: $`H`$ $`=`$ $`{\displaystyle _{\mathrm{R}.\mathrm{B}.\mathrm{Z}.}}{\displaystyle \frac{d^2k}{(2\pi )^2}}\chi ^\alpha (k)({\displaystyle \frac{1}{2}}(ϵ(k)ϵ(k+Q))\tau _z+`$ (93) $`\mathrm{\Delta }(k)\tau _x+{\displaystyle \frac{1}{2}}(ϵ(k)+ϵ(k+Q)))\chi _\alpha `$ The integral is over the reduced Brillouin zone. The $`\tau `$’s are Pauli matrices; the ‘flavor’ index $`A=1,2`$ on which they act has been suppressed. $`\mathrm{\Delta }(k)`$ is defined by. $$\mathrm{\Delta }(k)\mathrm{\Delta }_0f(k)g\mathrm{\Phi }_Qf(k)$$ (94) The single-quasiparticle energies are: $`E_\pm (k)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(ϵ(k)+ϵ(k+Q)\right)\pm `$ (96) $`{\displaystyle \frac{1}{2}}\sqrt{\left(ϵ(k)ϵ(k+Q)\right)^2+4\mathrm{\Delta }^2(k)}`$ Let’s consider the situation in which there is a node, i.e. when the argument of the square root vanishes (we discuss below the conditions under which this occurs). For simplicity, we will consider the commensurate $`\stackrel{}{Q}=(\pi /a,\pi /a)`$ singlet $`p_x`$ density-wave state in a model with anisotropic nearest-neighbor hopping: $$ϵ(k)=2t\left(r\mathrm{cos}k_xa+\mathrm{cos}k_ya\right)$$ (97) with $`r<1`$. The mean-field quasiparticle energies are: $$E(k)=\pm \sqrt{4t^2\left(r\mathrm{cos}k_xa+\mathrm{cos}k_ya\right)^2+\mathrm{\Delta }_{0}^{}{}_{}{}^{2}\mathrm{sin}^2k_xa}$$ (98) There is a node at $`k_x=0`$, $`k_ya=\mathrm{arccos}(r)`$. Expanding about this node, $$E(q)=\pm \sqrt{v_x^2q_x^2+v_y^2q_y^2}$$ (99) with momenta $`\stackrel{}{q}`$ now measured from the node and $$v_x=\mathrm{\Delta }_0a,v_y=2ta\sqrt{1r^2}$$ (100) The effective Lagrangian for the quasiparticles near the nodes can be written: $$_{\mathrm{eff}}=\chi ^\alpha \left(_\tau \tau _zv_yi_y\tau _xv_xi_x\right)\chi _\alpha $$ (101) Terms which break the nesting of the Fermi surface, such as the chemical potential or next-neighbor hopping, open hole pockets at the nodes: $$_\mu =\mu \chi ^\alpha \chi _\alpha $$ (102) We now turn to the question of when a $`p`$\- or $`d`$-wave density wave will have nodal excitations. Let’s again begin with the commensurate $`\stackrel{}{Q}=(\pi /a,\pi /a)`$ singlet $`p_x`$ density-wave state: $$\psi ^\alpha (k+Q,t)\psi _\beta (k,t)=i\left|\mathrm{\Phi }_Q\right|\mathrm{sin}k_xa\delta _\beta ^\alpha $$ (103) in a system in which the Fermi surface is nested at $`\stackrel{}{Q}`$. This state will have gapless excitations if the nodal line $`k_x=0`$ crosses the Fermi surface. For an open Fermi surface, this need not be the case. In an anisotropic nearest-neighbor tight-binding model, Eq. (97), with $`r>1`$, the Fermi surface at half-filling is an open Fermi surface which does not cross the line $`k_x=0`$. Consequently, there are no gapless excitations. For $`r<1`$, however, the Fermi surface does cross the line $`k_x=0`$, and there are gapless excitations. Are there any thermodynamic singularities at the transition at which gapless excitations occur? To answer this question, let us consider the mean-field ground state energy: $`E_0`$ $`=`$ $`{\displaystyle _{\mathrm{R}.\mathrm{B}.\mathrm{Z}.}}{\displaystyle \frac{d^2k}{(2\pi )^2}}E(k)`$ (104) $`=`$ $`{\displaystyle _{\mathrm{R}.\mathrm{B}.\mathrm{Z}.}}{\displaystyle \frac{d^2k}{(2\pi )^2}}\sqrt{ϵ^2(k)+\mathrm{\Delta }^2(k)}`$ (105) The first and second derivatives of $`E_0`$ are continuous. However, the third derivative of the ground state energy with respect to $`r`$ contains a term of the form: $`{\displaystyle \frac{^3E_0}{r^3}}={\displaystyle _{\mathrm{R}.\mathrm{B}.\mathrm{Z}.}}{\displaystyle \frac{d^2k}{(2\pi )^2}}{\displaystyle \frac{8ϵ(k)t^3\mathrm{cos}^3k_xa}{\left(ϵ^2(k)+\mathrm{\Delta }^2(k)\right)^{3/2}}}+\mathrm{}`$ (106) This term diverges if there is a node on the Fermi surface, but is finite otherwise. Hence, the phase with a node on the Fermi surface is a critical line with a singular third derivative of the ground state energy. We will call this phase the ‘critical phase’ of the $`p_x`$ density wave. Note that the second derivative of the ground state energy is everywhere continuous but nowhere differentiable in the critical phase. It is separated by a third-order phase transition from the phase with no gapless excitations, the non-critical phase of the $`p_x`$ density wave. How does this observation generalize (a) away from half-filling and to non-nested Fermi surfaces; and (b) to $`d`$-wave and/or incommensurate ordering? To answer (a), let’s change the chemical potential in order to move away from half-filling. Now, $`{\displaystyle \frac{^3E_0}{r^3}}={\displaystyle _{E(k)<\mu }}{\displaystyle \frac{d^2k}{(2\pi )^2}}{\displaystyle \frac{8ϵ(k)t^3\mathrm{cos}^3k_xa}{\left(ϵ^2(k)+\mathrm{\Delta }^2(k)\right)^{3/2}}}+\mathrm{}`$ (107) Below half-filling, the denominator never diverges. Hence, the system is always in the non-critical phase, despite the fact that there are gapless excitations. As the chemical potential is increased, the system crosses a third-order phase transition and enters the critical phase. Above half-filling, it is always in the critical phase. Suppose, now, that we allow next-nearest neighbor hopping $`t^{}`$, thereby spoiling nesting. The ground state energy is given by $`E_0`$ $`=`$ $`{\displaystyle _{E(k)<\mu }}{\displaystyle \frac{d^2k}{(2\pi )^2}}E(k)`$ (108) $`=`$ $`{\displaystyle _{E(k)<\mu }}{\displaystyle \frac{d^2k}{(2\pi )^2}}[{\displaystyle \frac{1}{2}}(ϵ(k)+ϵ(k+Q))`$ (110) $`{\displaystyle \frac{1}{2}}\sqrt{\left(ϵ(k)ϵ(k+Q)\right)^2+4\mathrm{\Delta }^2(k)}]`$ and $`{\displaystyle \frac{^3E_0}{r^3}}`$ $`=`$ $`{\displaystyle _{E(k)<\mu }}{\displaystyle \frac{d^2k}{(2\pi )^2}}{\displaystyle \frac{4ϵ(k)t^3\mathrm{cos}^3k_xa}{\left(\left(ϵ(k)ϵ(k+Q)\right)^2+4\mathrm{\Delta }^2(k)\right)^{3/2}}}`$ $`+\mathrm{}`$ This diverges if the nodal line of $`\mathrm{\Delta }(k)`$ crosses the curve $`ϵ(k)=ϵ(k+Q)`$ and this crossing point lies below the chemical potential. In such a case, the system is in the critical phase. Regardless of the details of the band structure, the curve $`ϵ(k)=ϵ(k+Q)`$ is determined by symmetry for commensurate $`\stackrel{}{Q}`$: it is the set of points for which $`\stackrel{}{k}`$ and $`\stackrel{}{k}+\stackrel{}{Q}`$ are related by a symmetry of the square lattice. For $`\stackrel{}{Q}=(\pi /a,0)`$, $`ϵ(k)=ϵ(k+Q)`$ if $`k_x=\pm \pi /2a`$. For $`\stackrel{}{Q}=(\pi /a,\pi /a)`$, $`ϵ(k)=ϵ(k+Q)`$ if $`k_x\pm k_y=\pm \pi /a`$. A $`d`$-wave density-wave will always have nodal lines which cross the curve $`ϵ(k)=ϵ(k+Q)`$; a $`p`$-wave density wave may or may not. If there is no crossing point, or the crossing point is not below the chemical potential, the system is in the non-critical phase. Again, the non-critical phase can have gapless excitations. In the case of incommensurate ordering, similar considerations hold. Let us suppose that the Fermi surface is nested at incommensurate $`\stackrel{}{Q}`$, i.e. if $`\stackrel{}{k}`$ is on the Fermi surface, then $`\stackrel{}{k}+\stackrel{}{Q}`$ is as well, and $`ϵ(k)=ϵ(k+Q)=\mu `$. If the Fermi surface intersects the nodal lines of $`\mathrm{\Delta }(k)`$, then there will be gapless nodal excitations. If the chemical potential is now lowered or the hopping parameters are changed, so that the Fermi surface is no longer perfectly nested, then the nodes will open into hole pockets. Again, as the nesting condition is approached, a third-order phase transition will occur, as in the commensurate case. In summary, the system will be in a ‘critical’ state if nodal points are at or below the Fermi surface. Otherwise, the system will be ‘non-critical’, whether or not there other gapless excitations. The transition between these two states and the entire critical phase is characterized, in mean-field-theory, by a diverent third derivative of the ground state energy. There is no reason to mistrust mean-field theory since there aren’t strong order parameter fluctuations which might destabilize out calculations. ## VI Transitions to Superconducting States As Zhang has recently emphasized, enhanced symmetry can be dynamically generated at a critical point between two different ordered electronic states. The focus of that work was a critical point between an antiferromagnet and a $`d_{x^2y^2}`$ superconductor. In earlier work, Yang identified an $`SU(2)`$ symmetry (which, together with $`SU(2)`$ spin-rotational symmetry, trivially forms an $`O(4)=SU(2)\times SU(2)\times Z_2`$) which is an exact symmetry of the Hubbard model at half-filling with $`\mu =U/2`$. This symmetry would be dynamically generated at a critical point between a CDW and an $`s`$-wave superconductor. We now consider the modification of this idea to $`p`$\- and $`d`$-wave ordering. We first consider a transition at half-filling between a singlet commensurate $`d_{x^2y^2}`$ density-wave and a $`d_{x^2y^2}`$ superconductor. We group the two order parameters into a vector<sup>§</sup><sup>§</sup>§We will use underlined lowercase Roman letters such as $`\underset{¯}{i}=1,2,3`$ to denote pseudospin triplet indices and uppercase Roman letters to denote peudospin doublet indices $`A=1,2`$. Lowercase Roman indices $`a=1,2,3`$ will be vector indices (i.e. real spin triplet indices) and Greek letters $`\alpha =1,2`$ will be used for real spin $`SU(2)`$ spinor indices. Pauli matrices $`\tau ^{\underset{¯}{i}}`$ will be used for pseudospin, while $`\sigma ^a`$ will be reserved for spin.: $`\mathrm{\Phi }_{\underset{¯}{i}}(q)f(k)=\left(\begin{array}{c}\sqrt{2}\mathrm{Re}\left\{\psi _{}^{}(k+\frac{q}{2})\psi _{}^{}(k+\frac{q}{2})\right\}\\ \sqrt{2}\mathrm{Im}\left\{\psi _{}^{}(k+\frac{q}{2})\psi _{}^{}(k+\frac{q}{2})\right\}\\ i\psi ^\alpha (k+Q+\frac{q}{2})\psi _\alpha (k\frac{q}{2})\end{array}\right)`$ (114) If, following Yang , we introduce the following $`SU(2)`$ generators which we will call pseudospin $`SU(2)`$ $`O^3`$ $`=`$ $`{\displaystyle _{\mathrm{R}.\mathrm{B}.\mathrm{Z}.}}{\displaystyle \frac{d^2k}{(2\pi )^2}}(\psi ^\alpha (k)\psi _\alpha (k)+`$ (116) $`\psi ^\alpha (k+Q)\psi _\alpha (k+Q))`$ $`O^+`$ $`=`$ $`{\displaystyle _{\mathrm{R}.\mathrm{B}.\mathrm{Z}.}}{\displaystyle \frac{d^2k}{(2\pi )^2}}i\psi _{}^{}(k)\psi _{}^{}(k+Q)`$ (117) $`O^{}`$ $`=`$ $`{\displaystyle _{\mathrm{R}.\mathrm{B}.\mathrm{Z}.}}{\displaystyle \frac{d^2k}{(2\pi )^2}}i\psi _{}(k)\psi _{}(k+Q)`$ (118) then the order parameters form a triplet under this $`SU(2)`$, $`\left(\begin{array}{c}\mathrm{\Phi }_+(q)f(k)\\ \mathrm{\Phi }_0(q)f(k)\\ \mathrm{\Phi }_{}(q)f(k)\end{array}\right)=\left(\begin{array}{c}\psi _{}^{}(k+\frac{q}{2})\psi _{}^{}(k+\frac{q}{2})\\ i\psi ^\alpha (k+Q+\frac{q}{2})\psi _\alpha (k\frac{q}{2})\\ \psi _{}(k+\frac{q}{2})\psi _{}(k+\frac{q}{2})\end{array}\right)`$ (125) There is a small but important difference between our pseudospin $`SU(2)`$ and Yang’s : the factors of $`i`$ in the definitions of $`O^\pm `$. These are necessary since a commensurate $`d_{x^2y^2}`$ density-wave breaks $`T`$, while a $`d_{x^2y^2}`$ superconductor does not. Consequently, our pseudospin $`SU(2)`$ does not commute with $`T`$, which is an inversion followed by a rotation by $`\pi `$ about the $`\underset{¯}{3}`$-axis. The electron fields transform as a doublet under the pseudospin $`SU(2)`$ as well as the spin $`SU(2)`$. We will group them into $`4`$-component objects $`\mathrm{\Psi }_{A\alpha }`$, where $`A`$ is the pseudospin index, $`A=1,2`$, and $`\alpha `$ is the spin index, $`\alpha =,`$: $`\left(\begin{array}{c}\mathrm{\Psi }_{1\alpha }\\ \mathrm{\Psi }_{2\alpha }\end{array}\right)=\left(\begin{array}{c}\psi _\alpha (k)\\ iϵ_{\alpha \beta }\psi ^\beta (k+Q)\end{array}\right)`$ (130) A ‘microscopic’ Hamiltonian which is $`O(4)`$ invariant can be written down: $$H=H_0+H_{int}$$ (131) $`H_0`$ $`=`$ $`{\displaystyle _{\mathrm{R}.\mathrm{B}.\mathrm{Z}.}}{\displaystyle \frac{d^2k}{(2\pi )^2}}ϵ(k)\mathrm{\Psi }^{A\alpha }\mathrm{\Psi }_{A\alpha }`$ (132) if $`H_{int}`$ is given byWe have only written down the quartic terms; higher-order $`O(4)`$ invariants also exist, but they are irrelevant at weak coupling.: $`H_{int}`$ $`=`$ $`{\displaystyle }{\displaystyle \frac{d^2q}{(2\pi )^2}}[u^{(0,0)}\lambda ^{(0,0)}(q)\lambda ^{(0,0)}(q)+`$ $`u^{(1,0)}\lambda _{\underset{¯}{i}}^{(1,0)}(q)\lambda _{\underset{¯}{i}}^{(1,0)}(q)+u^{(1,0)}\lambda _{\underset{¯}{i}Q}^{(1,0)}(q)\lambda _{\underset{¯}{i}Q}^{(1,0)}(q)+`$ $`u^{(0,1)}\lambda _a^{(0,1)}(q)\lambda _a^{(0,1)}(q)+u_Q^{(0,1)}\lambda _{aQ}^{(0,1)}(q)\lambda _{aQ}^{(0,1)}(q)+`$ $`u^{(1,1)}\lambda _{\underset{¯}{i}a}^{(1,1)}(q)\lambda _{\underset{¯}{i}a}^{(1,1)}(q)+u_Q^{(1,1)}\lambda _{\underset{¯}{i}aQ}^{(1,1)}(q)\lambda _{\underset{¯}{i}aQ}^{(1,1)}(q)]`$ where $`\lambda ^{(0,0)}`$ $`=`$ $`{\displaystyle \frac{d^2k}{(2\pi )^2}f(k)\mathrm{\Psi }^{A\alpha }\left(k+\frac{q}{2}\right)\mathrm{\Psi }_{A\alpha }\left(k\frac{q}{2}\right)}`$ (133) $`\lambda _{\underset{¯}{i}}^{(1,0)}`$ $`=`$ $`{\displaystyle }{\displaystyle \frac{d^2k}{(2\pi )^2}}f(k)\mathrm{\Psi }^{A\alpha }(k+{\displaystyle \frac{q}{2}})\times `$ (135) $`\tau _{\underset{¯}{i}A}^B\mathrm{\Psi }_{B\alpha }\left(k{\displaystyle \frac{q}{2}}\right)`$ $`\lambda _{\underset{¯}{i}Q}^{(1,0)}`$ $`=`$ $`{\displaystyle }{\displaystyle \frac{d^2k}{(2\pi )^2}}f(k)ϵ^{\alpha \beta }\mathrm{\Psi }_{C\alpha }(k+{\displaystyle \frac{q}{2}})ϵ^{CA}\times `$ (137) $`\tau _{\underset{¯}{i}A}^B\mathrm{\Psi }_{B\beta }\left(k+{\displaystyle \frac{q}{2}}\right)`$ $`\lambda _a^{(0,1)}`$ $`=`$ $`{\displaystyle }{\displaystyle \frac{d^2k}{(2\pi )^2}}f(k)\mathrm{\Psi }^{A\alpha }(k+{\displaystyle \frac{q}{2}})\sigma _{\underset{¯}{i}\alpha }^\beta \times `$ (139) $`\mathrm{\Psi }_{A\beta }\left(k{\displaystyle \frac{q}{2}}\right)`$ $`\lambda _{aQ}^{(0,1)}`$ $`=`$ $`{\displaystyle }{\displaystyle \frac{d^2k}{(2\pi )^2}}f(k)ϵ^{AB}\mathrm{\Psi }_{A\gamma }(k+{\displaystyle \frac{q}{2}})ϵ^{\gamma \beta }\times `$ (141) $`\sigma _{\underset{¯}{i}\alpha }^\beta \mathrm{\Psi }_{B\beta }\left(k+{\displaystyle \frac{q}{2}}\right)`$ $`\lambda _{\underset{¯}{i}a}^{(1,1)}`$ $`=`$ $`{\displaystyle }{\displaystyle \frac{d^2k}{(2\pi )^2}}f(k)\mathrm{\Psi }^{A\alpha }(k+{\displaystyle \frac{q}{2}})\tau _{\underset{¯}{i}A}^B\times `$ (143) $`\sigma _{\underset{¯}{i}\alpha }^\beta \mathrm{\Psi }_{B\beta }\left(k{\displaystyle \frac{q}{2}}\right)`$ $`\lambda _{\underset{¯}{i}aQ}^{(1,1)}`$ $`=`$ $`{\displaystyle }{\displaystyle \frac{d^2k}{(2\pi )^2}}f(k)\mathrm{\Psi }_{C\gamma }(k+{\displaystyle \frac{q}{2}})ϵ^{CA}\tau _{\underset{¯}{i}A}^B\times `$ (145) $`ϵ^{\gamma \beta }\sigma _{\underset{¯}{i}\alpha }^\beta \mathrm{\Psi }_{B\beta }\left(k+{\displaystyle \frac{q}{2}}\right)`$ These ‘microscopic’ Hamiltonians describe electrons at half-filling with a nested Fermi surface and interactions which favor density-wave and superconducting order equally. In other words, they describe a critical point at half-filling between a $`d_{x^2y^2}`$ density-wave and a $`d_{x^2y^2}`$ superconductor. Near the critical point, we can focus on the low-energy degrees of freedom: the Goldstone modes and the nodal fermionic excitations. We can write down an $`O(4)`$ invariant action for this: $`S_{\mathrm{eff}}`$ $`=`$ $`{\displaystyle 𝑑\tau \frac{d^2k}{(2\pi )^2}\mathrm{\Psi }^{A\alpha ^{}}(k)\left(_\tau ϵ(k)\right)\mathrm{\Psi }_{A\alpha }(k)}+`$ (150) $`ig{\displaystyle }d\tau {\displaystyle \frac{d^2k}{(2\pi )^2}}{\displaystyle \frac{d^2q}{(2\pi )^2}}\mathrm{\Phi }_{\underset{¯}{i}}(q)f(k)\times `$ $`[ϵ^{\alpha \beta }\mathrm{\Psi }_{C\alpha }(k+{\displaystyle \frac{q}{2}})ϵ^{CA}\tau _A^{\underset{¯}{i}B}\mathrm{\Psi }_{B\beta }(k+{\displaystyle \frac{q}{2}})+`$ $`ϵ_{\alpha \beta }\mathrm{\Psi }^{A\alpha }(k+{\displaystyle \frac{q}{2}})\tau _A^{\underset{¯}{i}B}ϵ^{BC}\mathrm{\Psi }^{B\beta }(k+{\displaystyle \frac{q}{2}})]`$ $`+{\displaystyle 𝑑\tau d^2x\left(\left(_\mu \mathrm{\Phi }_{\underset{¯}{i}}\right)^2+\frac{1}{2}r\mathrm{\Phi }_{\underset{¯}{i}}\mathrm{\Phi }_{\underset{¯}{i}}+\frac{1}{4!}u\left(\mathrm{\Phi }_{\underset{¯}{i}}\mathrm{\Phi }_{\underset{¯}{i}}\right)^2\right)}`$ In this Lagrangian, we have rescaled all of the velocities and stiffnesses to $`1`$. In general, these quantities will be different – breaking the $`O(4)`$ symmetry – and this cannot be done. Symmety-breaking terms will be briefly addressed below. The transition between the $`d_{x^2y^2}`$ density-wave and the $`d_{x^2y^2}`$ superconductor is driven by a pseudospin-$`2`$ symmetry-breaking field, which we will call $`u`$. $`_u`$ $`=`$ $`u\left(\mathrm{\Phi }_0^2+\mathrm{\Phi }_+\mathrm{\Phi }_{}\right)`$ (151) $`=`$ $`u\left(\mathrm{\Phi }_{\underset{¯}{3}}^2\mathrm{\Phi }_{\underset{¯}{1}}^2\mathrm{\Phi }_{\underset{¯}{2}}^2\right)`$ (152) For $`u<0`$, the $`3`$-axis is an easy axis and the $`d_{x^2y^2}`$ density-wave state is favored; for $`u>0`$, the $`12`$-plane is an easy plane and the $`d_{x^2y^2}`$ superconducting state is favored. We can move away from a nested Fermi surface by tuning the chemical potential or a next-neighbor hopping parameter. Such effects are encapsulated by a pseudospin-$`1`$ symmetry-breaking term: $`S_\mu `$ $`=`$ $`\mu O^3`$ (153) $`=`$ $`{\displaystyle 𝑑\tau d^2x\left(ϵ_{\underset{¯}{i}\underset{¯}{j}}\mathrm{\Phi }_{\underset{¯}{i}}_\tau \mathrm{\Phi }_{\underset{¯}{j}}+\mathrm{\Psi }^{}\tau ^{\underset{¯}{3}}\mathrm{\Psi }\right)}`$ (154) where $`O^3`$ is the pseudospin $`SU(2)`$ generator defined above. If $`u=0`$, $`\mu `$ will immediately force the pseudospin into the $`12`$ plane – i.e. the supercondcutor will be favored. If $`u<0`$, the $`d_{x^2y^2}`$ density-wave state will be favored until $`\mu _c(\sqrt{u})`$. At this point, a first-order phase transition – the pseudospin-flop transition – will occur at which the pseudospin switches from an easy-axis phase to an easy-plane phase. If we allow $`\mathrm{\Phi }_0`$ to have a different velocity than $`\mathrm{\Phi }_\pm `$, then this first order phase transition can become two second order phase transitions. Depending on the values of these parameters and the strength of quantum fluctuations, the intervening phase can either have both types of order or neither. The critical point occurs when the jump in $`\mathrm{\Phi }`$ is tuned to zero. Hence, it is a tricritical point. At such a critical point, $`O(4)`$-breaking terms can scale to zero. The critical point and the quantum critical region are described by the physics of the critical fluctuations coupled to nodal fermionic excitations. By arguments similar to those of , the nodal fermions are neutral, spin-$`1/2`$ objects. A more detailed analysis will be given elsewhere . Similar conclusions can be drawn for $`d_{xy}`$ and $`d_{x^2y^2}+id_{xy}`$ transitions; the latter case is particularly simple since there are no fermions. In the case of transitions between $`p_x`$-wave density-wave and superconducting states, the order parameters are both pseudospin and spin-triplets. Hence, the effective field theory for such a transition takes the form: $`S_{\mathrm{eff}}`$ $`=`$ $`{\displaystyle 𝑑\tau \frac{d^2k}{(2\pi )^2}\mathrm{\Psi }^{A\alpha ^{}}\left(_\tau ϵ(k)\right)\mathrm{\Psi }_{A\alpha }}+`$ $`ig{\displaystyle }d\tau {\displaystyle \frac{d^2k}{(2\pi )^2}}{\displaystyle \frac{d^2q}{(2\pi )^2}}\mathrm{\Phi }_{\underset{¯}{i}}^a(q)f(k)\times `$ $`[ϵ^{\gamma \alpha }\sigma _\alpha ^{a\beta }\mathrm{\Psi }_{C\gamma }(k+{\displaystyle \frac{q}{2}})ϵ^{CA}\tau _A^{aB}\mathrm{\Psi }_{B\beta }(k+{\displaystyle \frac{q}{2}})+`$ $`\sigma _\alpha ^{a\beta }ϵ_{\beta \gamma }\mathrm{\Psi }^{A\alpha }(k+{\displaystyle \frac{q}{2}})\tau _A^{aB}ϵ^{BC}\mathrm{\Psi }^{B\gamma }(k+{\displaystyle \frac{q}{2}})]`$ $`+{\displaystyle 𝑑\tau d^2x\left(\left(_\mu \mathrm{\Phi }_{\underset{¯}{i}}\right)^2+\frac{1}{2}r\mathrm{\Phi }_{\underset{¯}{i}}^a\mathrm{\Phi }_{\underset{¯}{i}}^a+\frac{1}{4!}u\left(\mathrm{\Phi }_{\underset{¯}{i}}^a\mathrm{\Phi }_{\underset{¯}{i}}^a\right)^2\right)}`$ where $`\left(\begin{array}{c}\mathrm{\Phi }_{\underset{¯}{1}}^a\\ \mathrm{\Phi }_{\underset{¯}{2}}^a\\ \mathrm{\Phi }_{\underset{¯}{3}}^a\end{array}\right)=\left(\begin{array}{c}\sqrt{2}\mathrm{Re}\left\{\psi _\gamma ^{}(k)ϵ^{\gamma \alpha }\sigma _\alpha ^{a\beta }\psi _\beta ^{}(k)\right\}\\ \sqrt{2}\mathrm{Im}\left\{\psi _\gamma ^{}(k)ϵ^{\gamma \alpha }\sigma _\alpha ^{a\beta }\psi _\beta ^{}(k)\right\}\\ i\psi ^\alpha (k+Q)\sigma _\alpha ^{a\beta }\psi _\beta (k)\end{array}\right)`$ (161) ## VII Discussion In this paper, we have discussed the properties of ordered states in which particle-hole pairs with non-zero angular momentum condense. These states generalize charge- and spin-density wave states in the same way that $`p`$\- and $`d`$-wave superconductors generalize $`s`$-wave superconductivity. However, unlike in the superconducting case – where the Meissner effect follows directly from the broken symmetry, irrespective of the pairing channel – the angular variation of the condensate makes $`p`$\- and $`d`$-wave density-wave ordering difficult to detect. Experiments seeking to uncover such order must (a) be sensitive to spatial variations of kinetic energy or currents or (b) measure higher-order correlations of the charge or spin density. We explained how $`\mu `$SR, neutron scattering, NQR, and Raman scattering can be used in this regard. Impurities, which break rotational invariance, would cause admixture of $`p`$\- or $`d`$-wave ordering with $`s`$-wave ordering. It is natural to wonder whether experiments which appear to detect SDW order should be re-examined to see if they have actually uncovered $`p`$\- or $`d`$-wave order which, as a result of impurities, is masquerading as $`s`$-wave order. As in the superconducting case, the non-trivial pairing symmetry can lead to the existence of nodal excitations. As parameters such as the chemical potential or next-neighbor hopping are varied, nodal excitations appear at a transition which is third-order in mean field theory. The ‘phase’ with nodal excitations is always critical. The analogies between $`p`$\- and $`d`$-wave density-wave ordering and $`p`$\- and $`d`$-wave superconductivity begs the question: what is the nature of a phase transition between such states? In answering this question, we are led to one of the motivations of this work. The pseudo-gap regime of the cuprate superconductors exhibits some properties which can be associated with $`d_{x^2y^2}`$ ordering. One explanation is that some features of the $`d_{x^2y^2}`$ superconducting state have been inherited. However, it is natural to inquire whether the physics of this regime could also be determined in part by proximity to a $`d_{x^2y^2}`$ density-wave state or the transition between the density-wave and superconducting states. In other words, we ask whether the physics of the pseudo-gap regime should be described by a theory which incorporates fluctuations between $`d_{x^2y^2}`$ density-wave and superconducting states. In this connection, we note that the physics of the critical point between $`d_{x^2y^2}`$ density-wave and superconducting states bears a rough resemblance to that of the $`SU(2)`$ mean field theory of the $`tJ`$ model . In that theory, the gauge field parametrizes fluctuations between the $`d_{x^2y^2}`$ density-wave and superconducting states, a role played in our analysis by the Goldstone bosons of the $`O(4)`$ effective theory. We also note that the Nodal Liquid state shares many features of the $`d_{x^2y^2}`$ density-wave and superconducting states. These issues and their possible relevance to the cuprates will be further explored elsewhere. I would like to thank Sudip Chakravarty, Stuart Brown, Subir Sachdev, and, especially, Dror Orgad for discussions.
warning/0001/hep-lat0001024.html
ar5iv
text
# 1 Introduction ## 1 Introduction The regularization is a fundamentally important issue of field theory with an infinite number of degrees of freedom. A closely related issue in field theory is quantum anomaly, though the anomaly itself is perfectly finite. The anomaly is more closely related to “conditional convergence”in a loose sense, the boundary between divergence and convergence. For this reason, the treatment of anomaly becomes rather subtle in a finite theory such as the lattice theory. In this talk, I briefly review some of the fundamental issues related to the regularization and amomalies from my own view point. I will discuss the continuum regularization as well as the lattice regularization, with particular emphasis on the recent exciting development in lattice gauge theory. ## 2 Brief review of continuum path integral We start with a brief summary of the continuum path integral approach to chiral anomaly and a regularization which may be called “mode cut-off”. We study the QCD-type Euclidean path integral with $`\overline{)}D\gamma ^\mu \left(_\mu igA_\mu ^aT^a\right)=\gamma ^\mu \left(_\mu igA_\mu \right)`$, $$𝒟\overline{\psi }𝒟\psi \left[𝒟A_\mu \right]\mathrm{exp}\left[\overline{\psi }\left(i\overline{)}Dm\right)\psi d^4x+S_{YM}\right]$$ (2.1) where $`\gamma ^\mu `$ matrices are anti-hermitian with $`\gamma ^\mu \gamma ^\nu +\gamma ^\nu \gamma ^\mu =2g^{\mu \nu }=2\delta ^{\mu \nu }`$, and $`\gamma _5=\gamma ^1\gamma ^2\gamma ^3\gamma ^4`$ is hermitian. $`S_{YM}`$ stands for the Yang-Mills action and $`\left[𝒟A_\mu \right]`$ contains a suitable gauge fixing. To analyze the chiral Jacobian we expand the fermion variables $`\psi \left(x\right)`$ $`=`$ $`{\displaystyle \underset{n}{}}a_n\phi _n\left(x\right)`$ $`\overline{\psi }\left(x\right)`$ $`=`$ $`{\displaystyle \underset{n}{}}\overline{b}_n\phi _n^{}\left(x\right)`$ (2.2) in terms of the eigen-functions of hermitian $`\overline{)}D`$ $`\overline{)}D\phi _n\left(x\right)`$ $`=`$ $`\lambda _n\phi _n\left(x\right)`$ $`{\displaystyle d^4x\phi _n^{}\left(x\right)\phi _l\left(x\right)}`$ $`=`$ $`\delta _{n,l}`$ (2.3) which diagonalize the fermionic action in (2.1). The fermionic path integral measure is then written as $$𝒟\overline{\psi }𝒟\psi =\underset{N\mathrm{}}{lim}\underset{n=1}{\overset{N}{}}d\overline{b}_nda_n$$ (2.4) Under an infinitesimal global chiral transformation $$\delta \psi =i\alpha \gamma _5\psi ,\delta \overline{\psi }=\overline{\psi }i\alpha \gamma _5$$ (2.5) we obtain the Jacobian factor $`J`$ $`=`$ $`\mathrm{exp}\left[2i\alpha \underset{N\mathrm{}}{lim}{\displaystyle \underset{n=1}{\overset{N}{}}}{\displaystyle d^4x\phi _n^{}\left(x\right)\gamma _5\phi _n\left(x\right)}\right]`$ (2.6) $`=`$ $`\mathrm{exp}\left[2i\alpha \left(n_+n_{}\right)\right]`$ where $`n_\pm `$ stand for the number of eigenfunctions with vanishing eigenvalues and $`\gamma _5\phi _n=\pm \phi _n`$, respectively, in (2.3). We here used the relation $`d^4x\phi _n^{}\left(x\right)\gamma _5\phi _n\left(x\right)=0`$ for $`\lambda _n0`$. The Atiyah-Singer index theorem $`n_+n_{}=\nu `$ with Pontryagin index $`\nu `$, which was confirmed for one- instanton sector in $`R^4`$ space by Jackiw and Rebbi, shows that the chiral Jacobian contains the correct information of chiral anomaly. To extract a local version of the index (i.e., anomaly), we start with the expression $`n_+n_{}`$ $`=`$ $`\underset{N\mathrm{}}{lim}{\displaystyle \underset{n=1}{\overset{N}{}}}{\displaystyle d^4x\phi _n^{}\left(x\right)\gamma _5f\left(\left(\lambda _n\right)^2/M^2\right)\phi _n\left(x\right)}`$ (2.7) $`=`$ $`\underset{N\mathrm{}}{lim}{\displaystyle \underset{n=1}{\overset{N}{}}}{\displaystyle d^4x\phi _n^{}\left(x\right)\gamma _5f\left(\overline{)}D^2/M^2\right)\phi _n\left(x\right)}`$ $``$ $`Tr\gamma _5f\left(\overline{)}DD/M^2\right)`$ for any smooth function $`f\left(x\right)`$ which rapidly goes to zero for $`x=\mathrm{}`$ with $`f\left(0\right)=1`$. Since $`\gamma _5f\left(\overline{)}D^2/M^2\right)`$ is a well-regularized operator, we may now use the plane wave basis of fermionic variables to extract an explicit gauge field dependence, and we define a local version of the index as $`\underset{M\mathrm{}}{lim}tr\gamma _5f\left(\overline{)}D^2/M^2\right)`$ $``$ $`\underset{M\mathrm{}}{lim}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}\phi _n^{}\left(x\right)\gamma _5f\left(\overline{)}D^2/M^2\right)\phi _n\left(x\right)`$ $`=`$ $`\underset{M\mathrm{}}{lim}tr{\displaystyle \frac{d^4k}{\left(2\pi \right)^4}e^{ikx}\gamma _5f\left(\overline{)}D/M^2\right)e^{ikx}}`$ $`=`$ $`\underset{M\mathrm{}}{lim}tr{\displaystyle \frac{d^4k}{\left(2\pi \right)^4}\gamma _5f\left\{\left(ik_\mu +D_\mu \right)^2/M^2\frac{ig}{4}[\gamma ^\mu ,\gamma ^\nu ]F_{\mu \nu }/M^2\right\}}`$ $`=`$ $`\underset{M\mathrm{}}{lim}trM^4{\displaystyle \frac{d^4k}{\left(2\pi \right)^4}\gamma _5f\left\{\left(ik_\mu +D_\mu /M\right)^2\frac{ig}{4}[\gamma ^\mu ,\gamma ^\nu ]F_{\mu \nu }/M^2\right\}}`$ where the remaining trace stands for Dirac and Yang-Mills indices. We also used the relation $$\overline{)}D^2=D_\mu D^\mu \frac{ig}{4}[\gamma ^\mu ,\gamma ^\nu ]F_{\mu \nu }$$ (2.9) and the rescaling of the variable $`k_\mu Mk_\mu `$. By noting $`tr\gamma _5=tr\gamma _5[\gamma ^\mu ,\gamma ^\nu ]=0`$, the above expression ( after expansion in powers of $`1/M`$) is written as ( with $`ϵ^{1234}=1`$) $`\underset{M\mathrm{}}{lim}tr\gamma _5f\left(\overline{)}D^2/M^2\right)`$ $`=`$ $`tr\gamma _5{\displaystyle \frac{1}{2!}}\left\{{\displaystyle \frac{ig}{4}}[\gamma ^\mu ,\gamma ^\nu ]F_{\mu \nu }\right\}^2{\displaystyle \frac{d^4k}{\left(2\pi \right)^4}f^{\prime \prime }\left(k_\mu k^\mu \right)}`$ (2.10) $`={\displaystyle \frac{g^2}{32\pi ^2}}trϵ^{\mu \nu \alpha \beta }F_{\mu \nu }F_{\alpha \beta }`$ where we used $`{\displaystyle \frac{d^4k}{\left(2\pi \right)^4}f^{\prime \prime }\left(k_\mu k^\mu \right)}`$ $`=`$ $`{\displaystyle \frac{1}{16\pi ^2}}{\displaystyle _0^{\mathrm{}}}f^{\prime \prime }\left(x\right)x𝑑x`$ (2.11) $`=`$ $`{\displaystyle \frac{1}{16\pi ^2}}`$ with $`x=k_\mu k^\mu >0`$ in our metric. One can confirm that any finite interval $`Lk_\mu L`$ of momentum variables in (2.8) before the rescaling $`k_\mu Mk_\mu `$ gives rise to a vanishing contribution to (2.10). In this sense, the short distance contribution determines the anomaly. When one combines (2.7) and (2.10), one establishes the Atiyah-Singer index theorem (in $`R^4`$ space). We note that the local version of the index (anomaly) is valid for Abelian theory also. The global index (2.7) as well as a local version of the index (2.10) are both independent of the regulator $`f\left(x\right)`$ provided $$f\left(0\right)=1,f\left(\mathrm{}\right)=0,f^{}\left(x\right)x|_{x=0}=f^{}\left(x\right)x|_{x=\mathrm{}}=0$$ (2.12) If one chooses a smooth function $`f\left(x\right)`$ such that $$f\left(x\right)1,0x1$$ (2.13) and $`f\left(x\right)`$ goes to $`0`$ very rapidly for $`x>1`$, one has $$\underset{N\mathrm{}}{lim}\underset{n=1}{\overset{N}{}}\phi _n^{}\left(x\right)\gamma _5f\left(\left(\lambda _n\right)^2/M^2\right)\phi _n\left(x\right)\underset{\left|\lambda _n\right|=0}{\overset{\left|\lambda _n\right|M}{}}\phi _n^{}\left(x\right)\gamma _5\phi _n\left(x\right)$$ (2.14) The essence of the present regularization may thus be called gauge invariant “mode cut-off”,following the terminology of Zinn-Justin. ## 3 Index theorem on the lattice We now come to the recent intersting development in lattice gauge theory. This development is based on the so-called Ginsparg-Wilson relation $$\gamma _5D+D\gamma _5=aD\gamma _5D.$$ (3.1) where $`a`$ stands for the lattice spacing. If one defines the operator $$\mathrm{\Gamma }_5\gamma _5\left(1\frac{1}{2}aD\right)$$ (3.2) which is hermitian, the above relation is written as $$\mathrm{\Gamma }_5\gamma _5D+\gamma _5D\mathrm{\Gamma }_5=0.$$ (3.3) Namely, $`\mathrm{\Gamma }_5`$ plays a role of $`\gamma _5`$ in continuum theory. An explicit example of the operator $`D`$ which satisfies the Ginsparg-Wilson relation has been constructed by Neuberger and it is known as the overlap operator. All the finite dimensional representations of the Ginsparg-Wilson algebra (3.3)or the eigenstates $`\varphi _n`$ of the hermitian $`\gamma _5D`$ $$\gamma _5D\varphi _n=\lambda _n\varphi _n$$ (3.4) on a finite lattice are categorized into the following 3 classes: (i) $`n_\pm `$ states, $$\gamma _5D\varphi _n=0,\gamma _5\varphi _n=\pm \varphi _n,$$ (3.5) (ii) $`N_\pm `$ states($`\mathrm{\Gamma }_5\varphi _n=0`$), $$\gamma _5D\varphi _n=\pm \frac{2}{a}\varphi _n,\gamma _5\varphi _n=\pm \varphi _n,respectively,$$ (3.6) (iii) Remaining states with $`0<\left|\lambda _n\right|<2/a`$, $$\gamma _5D\varphi _n=\lambda _n\varphi _n,\gamma _5D\left(\mathrm{\Gamma }_5\varphi _n\right)=\lambda _n\left(\mathrm{\Gamma }_5\varphi _n\right),$$ (3.7) and the sum rule $`n_++N_+=n_{}+N_{}`$ holds. All the $`n_\pm `$ and $`N_\pm `$ states are the eigenstates of $`D`$, $`D\varphi _n=0`$ and $`D\varphi _n=\left(2/a\right)\varphi _n`$, respectively. If one denotes the number of states in (iii) by $`2N_0`$, the total number of states (dimension of the representation) $`N`$ is given by $`N=2\left(n_++N_++N_0\right)`$, which is expected to be a constant independent of background gauge field configurations. The index theorem on the lattice formulated by Hasenfratz, Laliena and Niedermayer is stated as the equality $$Tr\mathrm{\Gamma }_5=n_+n_{}=\nu $$ (3.8) in the continuum limit$`a0`$.Here $`n_\pm `$ stand for the number of zero eigenvalue states in (3.5) with $`\gamma _5\varphi _n=\pm \varphi _n`$, respectively, and $`\nu `$ stands for the Pontrygin index or integrated form of chiral anomaly.The proof of this index ralation proceeds as follows: We first evaluate by using the above classification of states $`Tr\mathrm{\Gamma }_5`$ $`=`$ $`{\displaystyle \underset{\lambda _n}{}}\varphi _n^{}\mathrm{\Gamma }_5\varphi _n`$ (3.9) $`=`$ $`{\displaystyle \underset{\lambda _n=0}{}}\varphi _n^{}\mathrm{\Gamma }_5\varphi _n`$ $`=`$ $`{\displaystyle \underset{\lambda _n=0}{}}\varphi _n^{}\gamma _5\varphi _n=n_+n_{}`$ The explicit evaluation of $`Tr\mathrm{\Gamma }_5`$ has been performed by various authors by perturbative calculation. The result confirms the relation for $`a0`$ $$Tr\gamma _5\left(1\frac{a}{2}D\right)\left(x\right)=d^4\frac{g^2}{32\pi ^2}trϵ^{\mu \nu \alpha \beta }F_{\mu \nu }F_{\alpha \beta }.$$ (3.10) The actual calculation is rather involved. We here present a somewhat simpler calculation, which is similar to the continuum calculation in Section 2. We start with $$Tr\left\{\gamma _5\left[1\frac{1}{2}aD\right]f\left(\frac{\left(\gamma _5D\right)^2}{M^2}\right)\right\}=n_+n_{}$$ (3.11) Namely, the index is not modified by any regulator $`f\left(x\right)`$ with $`f\left(0\right)=1`$, as can be confirmed by using the basis in (3.5)-(3.7). The hermitian operator $`\gamma _5D`$ plays a priviledged role in the present analysis of the index theorem. We then consider a local version of the index $$tr\left\{\gamma _5\left[1\frac{1}{2}aD\right]f\left(\frac{\left(\gamma _5D\right)^2}{M^2}\right)\right\}$$ (3.12) where trace stands for Dirac and Yang-Mills indices. A local version of the index is not sensitive to the precise boundary condition , and one may take the infinite volume limit $`L=Na\mathrm{}`$ in the above expression. We now examine the continuum limit $`a0`$ of the above local expression (3.12)<sup>2</sup><sup>2</sup>2This continuum limit corresponds to the so-called “naive” continuum limit in the context of lattice gauge theory.. We first observe that the term $$tr\left\{\frac{1}{2}a\gamma _5Df\left(\frac{\left(\gamma _5D\right)^2}{M^2}\right)\right\}$$ (3.13) goes to zero in this limit. The large eigenvalues of $`\gamma _5D`$ are truncated at the value $`M`$ by the regulator $`f\left(x\right)`$ which rapidly goes to zero for large $`x`$. In other words, the global index of the operator $`Tr\frac{a}{2}\gamma _5Df\left(\frac{\left(\gamma _5D\right)^2}{M^2}\right)O\left(aM\right)`$. We thus examine the small $`a`$ limit of $$tr\left\{\gamma _5f\left(\frac{\left(\gamma _5D\right)^2}{M^2}\right)\right\}$$ (3.14) The operator appearing in this expression is well regularized by the function $`f\left(x\right)`$ , and we evaluate the above trace by using the plane wave basis to extract an explicit gauge field dependence. We consider a square lattice where the momentum is defined in the Brillouin zone $$\frac{\pi }{2a}k_\mu <\frac{3\pi }{2a}$$ (3.15) We assume that the operator $`D`$ is free of species doubling; in other words, the operator $`D`$ blows up rapidly ($`\frac{1}{a}`$) for small $`a`$ in the momentum region corresponding to species doublers. The contributions of doublers are eliminated by the regulator $`f\left(x\right)`$ in the above expression. We thus examine the above trace in the momentum range of the physical species $$\frac{\pi }{2a}k_\mu <\frac{\pi }{2a}$$ (3.16) We now obtain the limiting $`a0`$ expression $`\underset{a0}{lim}tr\left\{\gamma _5f\left({\displaystyle \frac{\left(\gamma _5D\right)^2}{M^2}}\right)\right\}`$ (3.17) $`=`$ $`\underset{a0}{lim}tr{\displaystyle _{\frac{\pi }{2a}}^{\frac{\pi }{2a}}}{\displaystyle \frac{d^4k}{\left(2\pi \right)^4}}e^{ikx}\gamma _5f\left({\displaystyle \frac{\left(\gamma _5D\right)^2}{M^2}}\right)e^{ikx}`$ $`=`$ $`\underset{L\mathrm{}}{lim}\underset{a0}{lim}tr{\displaystyle _L^L}{\displaystyle \frac{d^4k}{\left(2\pi \right)^4}}e^{ikx}\gamma _5f\left({\displaystyle \frac{\left(\gamma _5D\right)^2}{M^2}}\right)e^{ikx}`$ $`=`$ $`\underset{L\mathrm{}}{lim}tr{\displaystyle _L^L}{\displaystyle \frac{d^4k}{\left(2\pi \right)^4}}e^{ikx}\gamma _5f\left({\displaystyle \frac{\left(i\gamma _5\overline{)}D\right)^2}{M^2}}\right)e^{ikx}`$ $`=`$ $`tr\left\{\gamma _5f\left({\displaystyle \frac{\overline{)}D^2}{M^2}}\right)\right\}`$ where we first take the limit $`a0`$ with fixed $`k_\mu `$ in $`Lk_\mu L`$, and then take the limit $`L\mathrm{}`$. This procedure is justified if the integral is well convergent. We also assumed that the operator $`D`$ satisfies the following relation in the limit $`a0`$ $`De^{ikx}g\left(x\right)`$ $``$ $`e^{ikx}\left(\overline{)}ki\overline{)}\overline{)}A\right)g\left(x\right)`$ (3.18) $`=`$ $`i\overline{)}D\left(e^{ikx}g\left(x\right)\right)`$ for any fixed $`k_\mu `$, ($`\frac{\pi }{2a}<k_\mu <\frac{\pi }{2a}`$), and a sufficiently smooth function $`g\left(x\right)`$. The function $`g\left(x\right)`$ corresponds to the gauge potential in our case, which in turn means that the gauge potential $`A_\mu \left(x\right)`$ is assumed to vary very little over the distances of the elementary lattice spacing. It is shown that an explicit example of $`D`$ given by Neuberger satisfies the property (3.18) without species doublers. Our final expression (3.17) in the limit $`M\mathrm{}`$ thus reproduces the index theorem in the continuum formulation, (2.10), by using the quite general properties of the basic operator $`D`$ only: The basic relation (3.1) with hermitian $`\gamma _5D`$ and the continuum limit property (3.18) without species doubling in the limit $`a0`$. ### 3.1 Modified chiral transformation Utilizing the notion of the index on the lattice, Lüscher introduced a new kind of chiral transformation $$\delta \psi =i\alpha \gamma _5\left(1\frac{1}{2}aD\right)\psi ,\delta \overline{\psi }=\overline{\psi }i\alpha \left(1\frac{1}{2}aD\right)\gamma _5$$ (3.19) with an infinitesimal constant parameter $`\alpha `$. This transformation leaves the action in $$𝒟\overline{\psi }𝒟\psi \mathrm{exp}\left[a^4\overline{\psi }D\psi \right]$$ (3.20) invariant due to the property (3.1), and gives rise to the chiral Jacobian factor $$J=\mathrm{exp}\left\{2i\alpha Tr\gamma _5\left(1\frac{1}{2}aD\right)\right\}$$ (3.21) The index theorem (3.8) shows that this Jacobian factor indeed carries the correct chiral anomaly. As a generalization of the vector-like(QCD-type) theory discussed so far,Lüscher showed that a chiral Abelian gauge theory can be consistently defined on a lattice. In particular, the anomaly in the fermion number current, which generally appears in chiral gauge theory, arises as a result of the non-vanishing index of a rectangular $`m\times n`$ matrix $`M`$ $$dimkerMdimkerM^{}=mn$$ (3.22) Namely, the regularized Lagrangian for chiral fermion is characterized by a rectangular matrix instead of a naive square matrix. A further comment on the Abelian chiral theory on the lattice will be given later. The lattice regularization of chiral non-Abelian gauge theory has not been formulated so far. ## 4 Generalized Pauli-Villars regularization A Lgrangian level regularization of chiral non-Abelian gauge theory in continuum has been formulated by Frolov and Slavnov, and Narayanan and Neuberger. This scheme is based on a generalization of the Pauli-Villars regularization. To regularize one chiral fermion, one needs to introduce an infinte number of fermions and unphysical bosonic fermions.This regularization is applicable to anomaly-free gauge theory only. Instead of writing the regularized Lagrangian, we here discuss the generalized Pauli-Villars regularization from a view point of the regularization of currents. The essence of the generalized Pauli-Villars regularization is summarized in terms of regularized currents as follows: $`<\overline{\psi }\left(x\right)T^a\gamma ^\mu \left({\displaystyle \frac{1+\gamma _5}{2}}\right)\psi \left(x\right)>_{PV}`$ $`=`$ $`\underset{yx}{lim}\{{\displaystyle \frac{1}{2}}Tr\left[T^a\gamma ^\mu f(\overline{)}D^2/\mathrm{\Lambda }^2){\displaystyle \frac{1}{i\overline{)}D}}\delta (xy)\right]`$ $`+{\displaystyle \frac{1}{2}}Tr\left[T^a\gamma ^\mu \gamma _5{\displaystyle \frac{1}{i\overline{)}D}}\delta (xy)\right]\}`$ $`<\overline{\psi }\left(x\right)\gamma ^\mu \left({\displaystyle \frac{1+\gamma _5}{2}}\right)\psi \left(x\right)>_{PV}`$ $`=`$ $`\underset{yx}{lim}\{{\displaystyle \frac{1}{2}}Tr\left[\gamma ^\mu f(\overline{)}D^2/\mathrm{\Lambda }^2){\displaystyle \frac{1}{i\overline{)}D}}\delta (xy)\right]`$ $`+{\displaystyle \frac{1}{2}}Tr\left[\gamma ^\mu \gamma _5{\displaystyle \frac{1}{i\overline{)}D}}\delta (xy)\right]\}`$ $`<\overline{\psi }\left(x\right)\gamma ^\mu \gamma _5\left({\displaystyle \frac{1+\gamma _5}{2}}\right)\psi \left(x\right)>_{PV}`$ (4.1) $`=`$ $`\underset{yx}{lim}\{{\displaystyle \frac{1}{2}}Tr\left[\gamma ^\mu \gamma _5f(\overline{)}D^2/\mathrm{\Lambda }^2){\displaystyle \frac{1}{i\overline{)}D}}\delta (xy)\right]`$ $`+{\displaystyle \frac{1}{2}}Tr\left[\gamma ^\mu {\displaystyle \frac{1}{i\overline{)}D}}\delta (xy)\right]\}.`$ where the regularization function is defined by $$f\left(\overline{)}D^2/\mathrm{\Lambda }^2\right)=\underset{n=\mathrm{}}{\overset{\mathrm{}}{}}\left(1\right)^n\overline{)}D^2/\left[\overline{)}D^2+\left(n\mathrm{\Lambda }\right)^2\right]=\frac{\pi \left(\overline{)}D/\mathrm{\Lambda }\right)}{sinh\left(\pi \overline{)}D/\mathrm{\Lambda }\right)}$$ (4.2) Note that $`f\left(x\right)`$ satisfies all the properties in (2.12). In the left-hand sides of these equations (4.1),the currents are defined in terms of the fields appearing in the original chiral Lagrangian, which one wants to regularize,while the right-hand sides of these equations stand for the regularized expressions. The axial-vector and vector $`U\left(1\right)`$ currents in terms of the chiral fermion fields in the original Lagrangian are identical if one note $`\gamma _5^2=1`$, but the regularized versions (i.e. the last two equations in (4.1)) are different. In particular , the vector $`U\left(1\right)`$ current(i.e. , the second equation in (4.1)) is not completely regularized. We emphasize that all the one-loop diagrams are generated from the (partially) regularized currents in (4.1), as will be discussed later in connection with the effective action ; in otherwords,(4.1) retains all the information of the generalized Pauli-Villars regularization. It is interesting that this regularization is implemented in the Lagrangian level. ### 4.1 Covariant regularization A closely related regularization of chiral currents is known as the covarianr regularization, which regularizes all the currents (and consequently all the one-loop fermionic diagrams) and reproduces the so-called covariant anomalies. This covariant regularization is,however, not implemented in the Lagrangian level,in general. The currents in the covariant regularization are written as $`<\overline{\psi }\left(x\right)T^a\gamma ^\mu \left({\displaystyle \frac{1+\gamma _5}{2}}\right)\psi \left(x\right)>_{cov}`$ $`=`$ $`\underset{yx}{lim}\left\{Tr\left[T^a\gamma ^\mu \left({\displaystyle \frac{1+\gamma _5}{2}}\right)f\left(\overline{)}D^2/\mathrm{\Lambda }^2\right){\displaystyle \frac{1}{i\overline{)}D}}\delta \left(xy\right)\right]\right\}`$ $`<\overline{\psi }\left(x\right)\gamma ^\mu \left({\displaystyle \frac{1+\gamma _5}{2}}\right)\psi \left(x\right)>_{cov}`$ $`=`$ $`\underset{yx}{lim}\{Tr\left[\gamma ^\mu \left({\displaystyle \frac{1+\gamma _5}{2}}\right)f(\overline{)}D^2/\mathrm{\Lambda }^2){\displaystyle \frac{1}{i\overline{)}D}}\delta (xy)\right]`$ The difference of this regularization from the generalized Pauli-Villars regularization in (4.1) is that all the components ( either vector or axial-vector) are well-regularized. All the fermionic one-loop diagrams are thus regularized. The price we have to pay for this is that this regularization is not implemented in the Lagrangain level. The anomaly in the gauge current is given by $`D_\mu <\overline{\psi }\left(x\right)T^a\gamma ^\mu \left({\displaystyle \frac{1+\gamma _5}{2}}\right)\psi \left(x\right)>_{cov}`$ (4.4) $`=`$ $`D_\mu {\displaystyle \underset{n}{}}\phi _n\left(x\right)^{}T^a\gamma ^\mu \left({\displaystyle \frac{1+\gamma _5}{2}}\right)f\left(\lambda _n^2/\mathrm{\Lambda }^2\right){\displaystyle \frac{1}{i\lambda _n}}\phi _n\left(x\right)`$ $`=`$ $`{\displaystyle \underset{n}{}}\left(\overline{)}D\phi _n\left(x\right)\right)^{}T^a\left({\displaystyle \frac{1+\gamma _5}{2}}\right)f\left(\lambda _n^2/\mathrm{\Lambda }^2\right){\displaystyle \frac{1}{i\lambda _n}}\phi _n\left(x\right)`$ $`{\displaystyle \underset{n}{}}\phi _n\left(x\right)^{}T^a\left({\displaystyle \frac{1\gamma _5}{2}}\right)f\left(\lambda _n^2/\mathrm{\Lambda }^2\right){\displaystyle \frac{1}{i\lambda _n}}\overline{)}D\phi _n\left(x\right)`$ $`=`$ $`i{\displaystyle \underset{n}{}}\phi _n\left(x\right)^{}T^a\gamma _5f\left(\lambda _n^2/\mathrm{\Lambda }^2\right)\phi _n\left(x\right)`$ where we used the eigenfunctions $$\overline{)}D\phi _n=\lambda _n\phi _n$$ (4.5) We thus recover the Jacibian factor corresponding to the covariant anomaly. As for the fermion number anomaly, we have similarly $`_\mu <\overline{\psi }\left(x\right)\gamma ^\mu \left({\displaystyle \frac{1+\gamma _5}{2}}\right)\psi \left(x\right)>_{cov}`$ (4.6) $`=`$ $`i{\displaystyle \underset{n}{}}\phi _n\left(x\right)^{}\gamma _5f\left(\lambda _n^2/\mathrm{\Lambda }^2\right)\phi _n\left(x\right)`$ $`=`$ $`_\mu <\overline{\psi }\left(x\right)\gamma ^\mu \gamma _5\left({\displaystyle \frac{1+\gamma _5}{2}}\right)\psi \left(x\right)>_{PV}`$ This shows that one can reproduce the correct fermion number anomaly by using the axial $`U\left(1\right)`$ current in (4.1) in the generalized Pauli-Villars regularization. From this analysis, one can see that the generalized Pauli-Villars regularization is closely related to the covariant regularization. Since the covariant regularization is applicable to any chiral gauge theory, it is useful to decide if any theory is anomalous or not.However, the covariant current as it stands does not generate the integrable (or consistent) anomaly. This issue is discussed in the next Section. ## 5 Definition of effective action in terms of covariant currents It is known that the effective action for the fermion is written in terms of the current. By using this fact, it has been proposed by H.Banerjee, R.Banerjee and P.Mitra to write the regularized effective action in terms of the regularized covariant current. As a simplest example, we discuss the Abelian chiral gauge theory defined by $`Z`$ $`=`$ $`{\displaystyle 𝒟\overline{\psi }𝒟\psi e^{{\scriptscriptstyle d^4\overline{\psi }i\overline{)}D\left({\scriptscriptstyle \frac{1+\gamma _5}{2}}\right)\psi }}}`$ $`W`$ $`=`$ $`\mathrm{ln}Z=\mathrm{ln}det\left[i\overline{)}D\left({\displaystyle \frac{1+\gamma _5}{2}}\right)\right]`$ $`\overline{)}D`$ $`=`$ $`\gamma ^\mu \left(_\mu igA_\mu \right)`$ (5.1) We then obtain $$\frac{W}{g}=d^4xA_\mu \left(x\right)\overline{\psi }\left(x\right)\gamma ^\mu \left(\frac{1+\gamma _5}{2}\right)\psi \left(x\right)$$ (5.2) The regularized effective action may be defined in terms of the covariant current by $$W_{reg}_0^gd^4xA_\mu \left(x\right)\overline{\psi }\left(x\right)\gamma ^\mu \left(\frac{1+\gamma _5}{2}\right)\psi \left(x\right)_{cov}$$ (5.3) The consistent current is then derived from this regularized effective action as $`j^\mu \left(x\right)_{cons}`$ $``$ $`{\displaystyle \frac{\delta }{\delta A_\mu \left(x\right)}}W_{reg}`$ $`=`$ $`{\displaystyle _0^g}𝑑gj^\mu \left(x\right)_{cov}+{\displaystyle _0^g}𝑑g{\displaystyle 𝑑yA_\nu \left(y\right)\frac{\delta j^\nu \left(y\right)_{cov}}{\delta A_\mu \left(x\right)}}`$ $`=`$ $`j^\mu \left(x\right)_{cov}{\displaystyle _0^g}𝑑gg{\displaystyle \frac{}{g}}j^\mu \left(x\right)_{cov}+{\displaystyle _0^g}𝑑g{\displaystyle 𝑑yA_\nu \left(y\right)\frac{\delta j^\nu \left(y\right)_{cov}}{\delta A_\mu \left(x\right)}}`$ Also $$\frac{}{g}j^\mu \left(x\right)_{cov}=𝑑yA_\nu \left(y\right)\frac{\delta }{\delta \left(gA_\nu \left(y\right)\right)}j^\mu \left(x\right)_{cov}$$ (5.5) as $`j^\mu \left(x\right)_{cov}`$ depends on $`g`$ only through the combination $`aA_\nu \left(y\right)`$. We thus obtain $$j^\mu \left(x\right)_{cons}=j^\mu \left(x\right)_{cov}+_0^g𝑑g𝑑yA_\nu \left(y\right)\left\{\frac{\delta j^\nu \left(y\right)_{cov}}{\delta A_\mu \left(x\right)}\frac{\delta j^\mu \left(x\right)_{cov}}{\delta A_\nu \left(y\right)}\right\}$$ (5.6) We note that by using (5.6) $$W_{reg}_0^gd^4xA_\mu \left(x\right)j^\mu \left(x\right)_{cov}=_0^gd^4xA_\mu \left(x\right)j^\mu \left(x\right)_{cons}$$ (5.7) namely, the regularized effective action is independent of whether the reguralized covariant current or regularized consistent current is used in its construction. All the naive properties are reproduced by our definition of $`W_{reg}`$. As for the chiral anomaly, we have(by noting that the Abelian covariant current is gauge invariant) $`W\left(A_\mu +_\mu \omega \right)_{reg}`$ $`=`$ $`{\displaystyle _0^g}{\displaystyle d^4x\left(A_\mu \left(x\right)+_\mu \omega \left(x\right)\right)j^\mu \left(x\right)_{cov}}`$ (5.8) $`=`$ $`W_{reg}{\displaystyle _0^g}{\displaystyle d^4x\omega \left(x\right)_\mu j^\mu \left(x\right)_{cov}}`$ and if one lets the cut-off parameter $`\mathrm{\Lambda }\mathrm{}`$ in the last covariant current, we generate the covariant anomaly $`W\left(A_\mu +_\mu \omega \right)_{reg}`$ $`=`$ $`W_{reg}{\displaystyle \frac{1}{16\pi ^2}}{\displaystyle _0^g}{\displaystyle d^4x\omega \left(x\right)F\left(gA_{mu}\right)\stackrel{~}{F}\left(gA_\mu \right)}`$ (5.9) $`=`$ $`W_{reg}{\displaystyle \frac{1}{3}}{\displaystyle \frac{1}{16\pi ^2}}{\displaystyle d^4x\omega \left(x\right)F\left(gA_{mu}\right)\stackrel{~}{F}\left(gA_\mu \right)}`$ and we reproduce the consistent anomaly with the correct Bose symmetrization factor$`1/3`$.It is known that this scheme works for the non-Abelian theory also. ### 5.1 Application to lattice gauge theory It has been pointed out by H. Suzuki that the basic aspect of the above construction of the regularized effective action in terms of the covariant current works for the lattice theory also, and one in fact ontains a formula closely related to the construction of the Abelian chiral theory given by Lüscher. The starting expression is $$W=_0^g𝑑gTr\frac{D}{g}\left(\frac{1+\widehat{\gamma }_5}{2}\right)D^1$$ (5.10) where $`D`$ stands for the lattice Dirac operator which satisfies the Ginsparg-Wilson relation, and $$\widehat{\gamma }_5\gamma _5\left(1aD\right)$$ (5.11) with $`\left(\widehat{\gamma }_5\right)^2=1`$. A naive continuum limit of (5.10) is $$W_{naive}=_0^g𝑑gTr\left[A_\mu \gamma ^\mu \left(\frac{1+\gamma _5}{2}\right)\frac{1}{i\overline{)}D}\right]$$ (5.12) and $`W`$ is gauge invarinat in Abelian theory. Thus $`W`$ in (5.10) is a counter part of $`W_{reg}`$ in continuum theory. It has been shown by Suzuki that $`W`$ in (5.10) for lattice theory gives the first term of the consistent lattice Abelian anomaly $$\frac{1}{3}\frac{1}{16\pi ^2}F\stackrel{~}{F}_{lattice}+_\mu K^\mu $$ (5.13) where the second term is a “lattice artifact” found by Lüscher. Namely, $`K^\mu `$ is gauge invariant and goes to $`0`$ in the naive continuum limit $`a0`$. An improvement of the above $`W`$ by using this $`K^\mu `$ has been shown to be identical to the result in Ref.. ## 6 Conclusion The remarkable development in lattice theory enriched our understanding of the regularization of fermions and the basic aspects of chiral symmetry and anomalies in gauge theory. It is interesting to see that the covariant current and consistent current play mutually complementary roles in these constructions. The interesting notion of index on the lattice deserves further investigation. The lattice formulation of chiral non-Abelian theory ( and eventually supersymmetric theory) remains as a challenging problem.
warning/0001/hep-th0001093.html
ar5iv
text
# Scales of String Theory 22footnote 2Talk given at the Strings 99 Conference, Potsdam, July 1999. To be published in Class. Quantum Grav. 17 (2000) 1 . ## 1 Introduction String/M theory has a single dimensionful parameter and a large number of dynamical moduli. The expectation values of these moduli determine the semiclassical properties of the vacuum, and in particular the masses of its Kaluza-Klein and Regge excitations. It is hoped that some still unknown dynamical mechanism ultimately selects the (unique?) vacuum in which we live, thereby fixing also the scales of the new physics. A more down-to-earth approach is to rely on arguments of theoretical plausibility, on direct experimental limits and on indirect experimental hints. In the wake of the ‘duality revolution’ there has been a revival of interest on the question of scales \[3 – 17\] which I will try to review in the present talk. <sup>3</sup><sup>3</sup>3The audio version and transparencies can be found in . The conventional (and conservative) hypothesis is that the string, compactification and Planck scales lie all to within two or three orders of magnitude from each other, and are hence far beyond direct experimental reach. The non-gravitational physics at lower energies is believed to be described by a renormalizable supersymmetric quantum field theory (SQFT), which must include the Minimal Supersymmetric Standard Model (MSSM). Faith in this hypothesis has been bolstered by the following well-known facts : (i) Softly broken SQFTs can indeed be extrapolated consistently to near-Planckian energies without destabilizing the electroweak scale ; (ii) the minimal (or ‘desert’) unification assumption is in remarkable agreement with some of the measured low-energy parameters of our world, and (iii) the hypothesis is almost automatic within the weakly-coupled heterotic string. The negative side of the coin is that this story is at best incomplete : coupling the broken SQFT to (super)gravity gives rise to vacuum instabilities, including an unacceptably-large cosmological term. As has been (re)appreciated in recent years, our knowledge of the gravitational interaction is in fact also very limited in another way: Einstein’s classical theory has not been tested experimentally at distances shorter than the macroscopic (millimeter or $`10^3`$ eV) regime . Since this is also the observational upper bound on the cosmological constant , it is very tempting to speculate that a resolution of the associated long-standing puzzle will require a drastic modification of gravity at such scales. Recent proposals of a higher-dimensional gravity do not seem to point to a resolution of the problem , while more drastic modifications like millimeter-sized fundamental strings are hard to accomodate in a consistent theoretical framework. The problem of gravitational (in)stability is, in any case, an important motivation for pursuing alternatives to the conventional heterotic compactification scheme. Two other important motivations are (i) that compactification scenaria which confine gauge interactions on a brane arise, as we have learned, very naturally in controllable corners of the moduli space of M-theory, and (ii) that they can bring string and Kaluza-Klein physics closer to experiment. Furthermore, the extreme brane-world scenario with TeV type-I string scale and a (near-millimetric ?) transverse space of dimension two, is singled out as I will explain in this talk. The logarithmic sensitivity on the transverse size sets a stage similar to that of the conventional energy desert : it makes hierarchies of scales very natural, and allows reliable (model-independent) calculations of the effective parameters in the brane theory. The apparent unification of the measured low-energy gauge couplings, on the other hand, has not found a convincing explanation outside the traditional energy-desert scenario yet. Since this is arguably the only clear quantitative hint for physics beyond the Standard Model, I will start and end my discussion from this point. ## 2 Hints of Supersymmetric Unification The well-known observation is that if one extrapolates the three gauge couplings of $`SU(3)\times SU(2)\times U(1)`$ using the $`\beta `$-funcions of the MSSM they meet at a scale $`M_U2\times 10^{16}`$ GeV. This is consistent with the one-loop formulae $$\alpha _{i}^{}{}_{}{}^{1}(\mu )=\alpha _{U}^{}{}_{}{}^{1}+b_i\mathrm{log}\frac{\mu }{M_U}+\mathrm{\Delta }_i$$ (1) where $`b_i`$ are the $`\beta `$-function coefficients, and $`\alpha _U`$ is the fine structure constant at $`M_U`$. The threshold corrections $`\mathrm{\Delta }_i`$ parametrize our ignorance of the details of the theory at the unification scale, and of the details of supersymmetry breaking. Assuming that the $`\mathrm{\Delta }_i`$ are negligible, and treating $`M_U`$ and $`\alpha _U`$ as input parameters, we have one prediction which is verified by LEP data at the level of a few percent. In the minimal heterotic unification there is furthermore one extra relation between the input parameters and the experimental value of $`M_{\mathrm{Planck}}`$. It implies $`M_U5\times 10^{17}`$ GeV, which on the (appropriate) logarithmic scale is a second successful prediction of the theory at the level again of a few percent . Put differently: there was no a priori reason why the extrapolated low-energy couplings should not have met, if at all, at say $`10^{35}`$ GeV ! It is important here to realize that it is the very existence of the perturbative desert which renders the above predictions meaningful and robust. Threshold effects make a few-percent correction to differences of couplings only because the logarithm in equations (1) is very large. It would be impossible to ignore the nitty-gritty details of the model if the unification scale were say instead at $`100`$ TeV. Similar arguments can be given for the mass matrices of quarks and leptons. Minimal assumptions (such as discrete symmetries and Higgs-field content) determine boundary conditions for Yukawa couplings, which can then be evolved with the equations of the renormalization group. The agreement with low-energy data is suggestive , though less compelling than for the gauge-coupling constants. Finally, I should mention that the existence of the superheavy scale $`M_U`$ is indirectly supported by two other independent pieces of data : the very long lifetime of the proton and the extraordinary smallness of neutrino masses . ## 3 Weakly-coupled Heterotic String Let me proceed next to the theoretical arguments that make the SQFT hypothesis quasi-automatic within the weakly-coupled heterotic string. Both the graviton and the perturbative gauge bosons live in this case in the ten-dimensional bulk, and interact through the sphere diagram at tree-level. The four-dimensional Yang-Mills and Einstein actions therefore read $$_{\mathrm{gauge}}\frac{(rM_h)^6}{g_h^2}\mathrm{tr}F^2\mathrm{and}_{\mathrm{grav}}\frac{r^6M_h^8}{g_h^2},$$ (2) where $`M_h`$ and $`g_h`$ are the heterotic string scale and string coupling constant, while $`r`$ is the typical compactification radius. By virtue of T-duality $`r`$ can be always taken greater than, or equal to, the string length. From the coefficients of these actions we can read the four-dimensional gauge coupling and Planck mass with the result $$\alpha _U\frac{g_h^2}{(rM_h)^6},$$ (3) and $$M_h^2\alpha _UM_{\mathrm{Planck}}^2.$$ (4) Factors of $`2`$’s and $`\pi `$’s in these relations are irrelevant for our arguments and have been dropped. I have also dropped the level of the corresponding Kac-Moody algebra which for all practical purposes is an integer of order one. Assume now that (i) the bare gauge coupling is of order one, and (ii) the heterotic theory stays weakly coupled ($`g_h1`$). Then the universal relation (4), which we have already encountered in the previous section, implies that the string scale is tied automatically to the Planck mass. Relation (3) on the other hand also implies that $`rM_h`$ cannot be much larger than one. Since it cannot be smaller than one by T-duality, it is necessarily of order one. Thus there is little leeway for abandoning the conventional scenario within the context of the weakly-coupled heterotic string. We can of course try to relax one of the above two assumptions : either (i) allow $`g_h`$ to be hierarchically large and hope that the gauge sector still stays under control in some special models , or (ii) let $`\alpha _U`$ be hierarchically small and hope that the Standard Model gauge couplings will be driven to their measured values by the large threshold corrections of a higher dimensional field theory .<sup>4</sup><sup>4</sup>4Higher-dimensional thresholds have been discussed also in other contexts, see for example . Such exotic possibilities have been motivated in the past by the search for classical vacua with broken low-energy supersymmetry. Known ‘mechanisms’ of continuous supersymmetry breaking indeed tie the breaking scale to the size of some internal dimensions. Classical supersymmetry restoration can be furthermore argued to be singular, in string theory, on general grounds . It was thus suggested early on that one (or more) radii of inverse size at the TeV would be required if the ‘observed’ supersymmetry breaking in nature were classical. Tree-level breaking, on the other hand, is at best an assumption of convenience – there is no reason in principle why the breaking in nature should not have a non-perturbative origin. Furthermore the classical mechanisms have not so far lead to new insights on the crucial problems of vacuum selection and stability. Thus, there seemed to be little theoretical motivation for abandoning the conventional compactification scheme, and its successful unification predictions, in heterotic string theory. ## 4 Brane World and Type-I Theory One of the important developments of the ‘duality revolution’ has been the realization that various branes – Dirichlet branes , or their dual heterotic fivebranes and Horava-Witten walls – can trap non-abelian gauge interactions in their worldvolumes. This has placed on a firmer basis an old idea according to which we might be living on a brane embedded in a higher-dimensional world. The idea arises naturally in compactifications of type I theory , which typically involve collections of orientifold planes and D-branes. I will from now on restrict my discussion to this context, because in it the ‘brane-world’ scenario admits a fully perturbative string description. All other interesting possibilities (see for example ) involve some type of non-perturbative dynamics and are a priori harder to control. In type I string theory the graviton (a closed-string state) lives in the ten-dimensional bulk, while open-string vector bosons are in general localized on lower-dimensional D-branes. Furthermore while closed strings interact to leading order via the sphere diagram, open strings interact via the disk diagram which is of higher order in the genus expansion. The four-dimensional Planck mass and Yang-Mills couplings therefore take the form $$\alpha _U\frac{g_\mathrm{I}}{(\stackrel{~}{r}M_\mathrm{I})^{6n}},M_{\mathrm{Planck}}^2\frac{r^n\stackrel{~}{r}^{6n}M_\mathrm{I}^8}{g_\mathrm{I}^2},$$ (5) where $`r`$ is the typical radius of the $`n`$ compact dimensions transverse to the brane, $`\stackrel{~}{r}`$ the typical radius of the remaining (6-$`n`$) compact longitudinal dimensions, $`M_\mathrm{I}`$ the type-I string scale and $`g_\mathrm{I}`$ the string coupling constant. By appropriate T-dualities we can again ensure that both $`r`$ and $`\stackrel{~}{r}`$ are greater than or equal to the fundamental string scale. T-dualities change $`n`$ and may take us either to Ia or to Ib theory (also called I or I’, respectively) but I will not make a distinction between these two. It follows from these formulae that (i) there is no universal relation between $`M_{\mathrm{Planck}}`$, $`\alpha _U`$ and $`M_\mathrm{I}`$ anymore, and (ii) tree-level gauge couplings corresponding to different sets of D-branes have radius-dependent ratios and need not unify at all. Thus type-I string theory is much more flexible (and less predictive) than its heterotic counterpart. The fundamental string scale, $`M_\mathrm{I}`$, in particular is a free parameter, even if one insists that $`\alpha _U`$ be kept fixed and of order one, and that the string theory be weakly coupled. This added flexibility can be used to ‘remove’ the order-of magnitude discrepancy between the apparent unification and string scales of the heterotic theory , to lower $`M_\mathrm{I}`$ to an intemediate scale or even all the way down to its experimentally-allowed limit of order the TeV . Keeping for instance $`g_\mathrm{I}`$, $`\alpha _U`$ and $`(\stackrel{~}{r}M_\mathrm{I})`$ fixed and of order one, leads to the condition $$r^nM_{\mathrm{Planck}}^2/M_\mathrm{I}^{2+n}.$$ (6) A TeV string scale would then require from $`n=2`$ millimetric to $`n=6`$ fermi-size dimensions transverse to our brane world. The relative weakness of gravity is in this picture attributed to the transverse spreading of gravitational flux. ## 5 Experimental Bounds What has brought the brane-world idea into focus was the realization that it cannot be a priori ruled out by the existing data, even in the most extreme case of ‘TeV-ish’ string scale and millimmeter-size transverse dimensions. Gravity is hard to test at submillimeter distances because of the large background of residual electromagnetic interactions. The ratio for instance of the Van der Waals to Newtonian force between two hydrogen atoms a distance $`d`$ apart is $$\frac{F_{\mathrm{VdW}}}{F_{\mathrm{grav}}}\left(\frac{1mm}{d}\right)^5.$$ (7) At $`d=10\mu m`$ Newton’s force is thus ten orders of magnitude weaker than Van der Waals ! As a result the present-day data allows practically any modification of Newton’s law, as long as it is of comparable strength at and screened beyond the millimeter range. This has been appreciated in the past in the context of gravitational axions , and similar bounds hold for light string moduli or extra Kaluza-Klein dimensions . Besides mesoscopic gravity experiments, there are two other types of direct experimental limits one should worry about : those coming from precision observables of the Standard Model, and those coming from various exotic processes. Precision tests of the SM and compositeness bounds cannot rule out in a model-independent way any new physics above the TeV-ish scale. Bounds for instance from LEP data on four-fermion operators, or bounds on dimension-five operators contributing to the $`g2`$ of the electron/muon are safe, as long as the characteristic scale of the new physics is a few TeV . Proton decay and other exotic processes could of course rule out large classes of low-scale models. There exist however plausible suppression mechanisms, such as bulk U(1) gauge symmetries which are spontaneously-broken at some distant brane and look like approximate global symmetries in our brane world. One type of model-independent exotic process is graviton emission in the bulk, which could be seen as missing-energy events in collider experiments . The process is however suppressed by the four-dimensional Newton constant at low energies, and only becomes appreciable (as one should expect) near string scale where quantum gravity effects are strong. None of these (or other) phenomenological considerations seems a priori fatal to the brane-world scenario, even in its extreme realization. Put together they will, however, probably make realistic type-I model building a very strenuous exercise indeed. ## 6 The Trasverse Desert Although $`M_\mathrm{I}`$ could lie anywhere between the Planck mass and the TeV , lowering it to the latter scale has two advantages : (i) it brings string physics within the reach of future acceleretor experiments, and (ii) it is a natural starting point for discussing the problem of the gauge hierarchy, which becomes now a question in the infrared . In a certain sense this extreme choice is antipodal to the energy-desert scenario: although the MSSM is a stable renormalizable field theory, we are shrinking its range of validity to one order of magnitude at most! Nevertheless, as I will now argue, these two scenaria share many common features when the number of large dimensions transverse to our brane is exactly two . The key feature of the SQFT hypothesis is that low-energy parameters receive large logarithmic corrections, which are effectively resummed by the equations of the Renormalization Group. The logarithmic sensitivity of parameters also generates naturally hierarchies of scales, and has been the key ingredient in all efforts to understand the origin of the gauge hierarchy in the past . Consider now the brane world scenario. The parameters of the effective brane lagrangian are dynamical open- and closed-string moduli with constant expectation values along the four non-compact space-time dimensions of our world. The closed-string moduli, $`m_a`$, are bulk fields whose expectation values will generically vary as a function of the transverse coordinates $`\xi `$. They include the dilaton, twisted-sector massless scalars, the metric of the transverse space etc. For weak type-I string coupling these variations can be described by a lagrangian of the form $$_{\mathrm{bulk}}+_{\mathrm{source}}d^n\xi \left[\frac{1}{g_\mathrm{I}^2}(_\xi m_a)^2+\frac{1}{g_\mathrm{I}}\underset{s}{}f_s(m_a)\delta (\xi \xi _s)\right].$$ (8) Here $`_{\mathrm{bulk}}`$ is a reduced supergravity Lagrangian while the sources are the D-branes and orientifolds which are localized at positions $`\xi _s`$ in the transverse space. The couplings $`f_s(m_a)`$ may vary from source to source – they can for instance depend on open-string moduli – and are subject to global consistency conditions. What is important is that they are weak in the type-I limit, leading to weak field variations, $$m_a(\xi )=m_a^0+g_\mathrm{I}m_a^1(\xi )+\mathrm{},$$ (9) with $`m_a^0`$ the (constant) average value, $`m_a^1(\xi )`$ given by a sum of Green’s functions, and so on. For $`n=2`$ transverse dimensions the leading variation grows logarithmically with the size, $`r`$, of the transverse space. Since our Standard Model parameters will be a function of the moduli evaluated at the position of our brane world, they will have logarithmic sensitivity on $`M_{\mathrm{Planck}}`$, very much like the (relevant) parameters of a supersymmetric renormalizable QFT. Similar sensitivity may occur even if $`n>2`$ , as long as some of the ‘bulk’ moduli propagate in only two extra large dimensions. The bulk supergravity Lagrangian receives both stringy and higher-genus corrections, but these involve higher derivatives of fields, and should therefore be negligible for moduli varying logarithmically over distances much larger than the string scale. The source functions, $`f_s(m_a)`$, will also be generically modified by such corrections – the D-branes have indeed string-scale thickness when probed by the supergravity fields . Such source modifications can, however, be absorbed into boundary conditions for the supergravity equations, at the special marked points $`\xi _s`$. The situation thus looks (at least superficially) analogous to that prevailing under the SQFT hypothesis : large corrections to effective low-energy couplings can be in both cases resummed by differential equations subject to appropriate boundary conditions. Furthermore ‘threshold corrections’ parametrizing our ignorance of the detailed physics at distant branes – the analog of physics near the unification scale – have a small effect on the relative evolution of parameters, provided the transverse two-dimensional space is sufficiently large. Clearly $`n=2`$ is critical: for exactly one transverse dimension bulk fields vary linearly in space and one expects to hit strong-coupling singularities before $`r`$ can grow very large, while for $`n>2`$ the dynamics on our brane completely decouples from the dynamics elsewhere in the bulk. ## 7 The Puzzle of Unification The logarithmic sensitivity of brane parameters with $`r`$ seems a natural setting for generating scale hierarchies dynamically, as in the case of renormalizable SQFT. Gauge dynamics on a given brane, for example, could become strong as the transverse space expands to an exponentially large size, thereby inducing gaugino condensation and supersymmetry breaking. The problems of vacuum selection and stability are, to be sure, still with us – the situation is, in this respect, neither better nor worse than in the traditional compactification scenario. The apparent unification of the (MS)SM gauge couplings, on the other hand, has not yet found a convincing ‘explanation’ in this (or any alternative) context. The basic ingredients are, nevertheless, present (see also for related ideas): (i) the logarithmic variation of bulk fields in the large transverse space can give equally-robust predictions for brane parameters, as renormalization group running over the energy desert, (ii) some bulk fields (like twisted scalar moduli) have non-universal couplings to gauge fields that live on the same set of D-branes – they can thus split their gauge couplings apart without separating them in transverse space, and (iii) the boundary condition imposing unification at high energy could be replaced by a boundary condition that these non-universally coupled bulk moduli vanish on some far-distant brane (for instance because of a non-perturbative superpotential). The main difficulty of these ideas is to explain why the coefficients of the logarithmic real-space evolutions should have the same (ratio of?) differences as the MSSM beta functions. This is clearly required if we are to ‘understand’ the actual values of the SM gauge couplings, as measured at LEP energies. Aknowledgements I thank the organizers for the invitation to speak, and for organizing a marvellous strings 99 conference. This work was partially supported by the TMR contract ERBFMRX-CT96-0090. ## References
warning/0001/hep-th0001068.html
ar5iv
text
# 1 Introduction ## 1 Introduction The general theory of duality invariance of abelian gauge theory was developed in and further elaborated in a series of publications (see and references therein). In this paper we generalize the duality equation of Gaillard and Zumino , also obtained independently in , to $`𝒩=1,2`$ supersymmetric theories. This duality equation is the condition for a theory with Lagrangian $`L(F_{ab})`$ to be invariant under $`U(1)`$ duality transformations $$\delta F=\lambda G,\delta G=\lambda F,$$ (1.1) where $$\stackrel{~}{G}_{ab}=\frac{1}{2}\epsilon _{abcd}G^{cd}=2\frac{L}{F^{ab}}.$$ (1.2) The equation reads $$G^{ab}\stackrel{~}{G}_{ab}+F^{ab}\stackrel{~}{F}_{ab}=0$$ (1.3) and presents a nontrivial constraint on the Lagrangian. The Born-Infeld (BI) theory is a particular solution of eq. (1.3). The BI action naturally appears in string theory (see for a recent review). Its $`𝒩=1`$ supersymmetric generalization (see also ) turns out to be the action for a Goldstone multiplet associated with partial breaking of $`𝒩=2`$ to $`𝒩=1`$ supersymmetry . It has been conjectured that a $`𝒩=2`$ supersymmetric generalization of the BI action should provide a model for partial breakdown $`𝒩=4𝒩=2`$, with the $`𝒩=2`$ vector multiplet being the corresponding Goldstone field, but the existing mechanisms of partial supersymmetry breaking are very difficult to implement in the $`𝒩=4`$ case. A candidate for $`𝒩=2`$ BI action has been suggested in . It correctly reduces to the Cecotti-Ferrara action once the $`(0,\frac{1}{2})`$ part of the $`𝒩=2`$ vector multiplet is switched off. However, there exist infinitely many $`𝒩=2`$ superfield actions with that property. Therefore, requiring the correct $`𝒩=1`$ reduction does not suffice to fix a proper $`𝒩=2`$ generalization of the BI action. One has to impose additional physical requirements. Since no mechanism for partial $`𝒩=4𝒩=2`$ breaking is currently available, it is natural to look for the $`𝒩=2`$ BI action as a solution of the supersymmetric generalization of the Gaillard-Zumino equation (1.3). In this paper we find $`𝒩=1,2`$ supersymmetric generalizations of the duality equation (1.3). They are presented in eqs. (2.8) and (3.10), respectively. It is not surprising that the Cecotti-Ferrara action is a solution of the $`𝒩=1`$ duality equation. In contrast, the action proposed in does not satisfy the $`𝒩=2`$ duality equation. However, the key to the construction of duality invariant $`𝒩=2`$ BI action was given in where a nonlinear $`𝒩=2`$ superfield constraint was introduced as a minimal extension of that generating the $`𝒩=1`$ BI action . It was asserted that the constrained superfield introduced does generate the $`𝒩=2`$ action given in . While this claim is incorrect, the constrained superfield nevertheless does generate the duality invariant $`𝒩=2`$ action that reduces to the $`𝒩=1`$ BI action after the $`(0,\frac{1}{2})`$ part of the $`𝒩=2`$ vector multiplet is switched off. One application of the $`𝒩=2`$ duality equation may be to compute the duality invariant low-energy effective actions of supersymmetric gauge theories. The $`𝒩=4`$ super Yang-Mills theory is expected to be self-dual . It was proposed in to look for its low-energy action on the Coulomb branch as a solution of the self-duality equation via the $`𝒩=2`$ superfield Legendre transformation, and a few subleading corrections to the low-energy action were determined. For non-supersymmetric theories it was shown in that the Gaillard-Zumino equation (1.3) implies self-duality via Legendre transformation. The Gaillard-Zumino equation is much simpler to solve and this advantage becomes essential in supersymmetric theories, where the procedure of inverting the Legendre transformation appears to be more involved at higher orders of perturbation theory . We have already remarked that (1.3) implies self-duality via Legendre transformation, but it is in fact a stronger condition. With reference to recent interest in the (supersymmetric) BI action within the context of D-branes, this stronger condition is in fact what one would like to impose. As was noted in , the D3-brane world-volume action, which contains, in addition to the gauge field also the axion and the dilaton fields, possesses a non-trivial $`SL(2,𝐑)`$ symmetry. The BI action we are considering corresponds to the CP-even part of this action for the special choice of vanishing axion and dilaton. This background is invariant precisely under the $`U(1)SL(2,𝐑)`$ duality group we are considering. Our paper is organized as follows. In section 2 we derive the $`𝒩=1`$ generalization of the Gaillard-Zumino equation and give a family of duality invariant nonlinear $`𝒩=1`$ models. The $`𝒩=1`$ BI action is a special member of this family. We also introduce a superconformally invariant generalization of the $`𝒩=1`$ BI action by coupling the vector multiplet to a scalar multiplet. In section 3 we present the $`𝒩=2`$ duality equation and derive its nonperturbative solution that reduces to the $`𝒩=1`$ BI action when the $`(0,\frac{1}{2})`$ part of $`𝒩=2`$ vector multiplet is switched off. We also develop a consistent perturbative scheme of computing duality invariant $`𝒩=2`$ superconformal actions. In Appendix A we discuss the general structure of the duality equation in the non-supersymmetric case and we show that any solution of (1.3) admits a supersymmetric extension. In Appendix B we give an explicit proof that the $`𝒩=2`$ BI action is self-dual with respect to Legendre transformation. ## 2 $`𝒩`$ = 1 duality rotations Let $`S[W,\overline{W}]`$ be the action describing the dynamics of a single $`𝒩=1`$ vector multiplet. The (anti) chiral superfield strengths $`\overline{W}_{\dot{\alpha }}`$ and $`W_\alpha `$,<sup>1</sup><sup>1</sup>1Our $`𝒩=1`$ conventions correspond to . $$W_\alpha =\frac{1}{4}\overline{D}^2D_\alpha V,\overline{W}_{\dot{\alpha }}=\frac{1}{4}D^2\overline{D}_{\dot{\alpha }}V,$$ (2.1) are defined in terms of a real unconstrained prepotential $`V`$. As a consequence, the strengths are constrained superfields, that is they satisfy the Bianchi identity $$D^\alpha W_\alpha =\overline{D}_{\dot{\alpha }}\overline{W}^{\dot{\alpha }}.$$ (2.2) Suppose that $`S[W,\overline{W}]`$ can be unambiguously defined<sup>2</sup><sup>2</sup>2This is always possible if $`S[W,\overline{W}]`$ does not involve the combination $`D^\alpha W_\alpha `$ as an independent variable. as a functional of unconstrained (anti) chiral superfields $`\overline{W}_{\dot{\alpha }}`$ and $`W_\alpha `$. Then, one can define (anti) chiral superfields $`\overline{M}_{\dot{\alpha }}`$ and $`M_\alpha `$ as $$\mathrm{i}M_\alpha 2\frac{\delta }{\delta W^\alpha }S[W,\overline{W}],\mathrm{i}\overline{M}^{\dot{\alpha }}2\frac{\delta }{\delta \overline{W}_{\dot{\alpha }}}S[W,\overline{W}].$$ (2.3) The equation of motion following from the action $`S[W,\overline{W}]`$ reads $$D^\alpha M_\alpha =\overline{D}_{\dot{\alpha }}\overline{M}^{\dot{\alpha }}.$$ (2.4) Since the Bianchi identity (2.2) and the equation of motion (2.4) have the same functional form, one may consider infinitesimal $`U(1)`$ duality transformations $$\delta W_\alpha =\lambda M_\alpha ,\delta M_\alpha =\lambda W_\alpha .$$ (2.5) To preserve the definition (2.3) of $`M_\alpha `$ and its conjugate, the action should transform as $$\delta S=\frac{\mathrm{i}}{4}\lambda \mathrm{d}^6z\left\{W^\alpha W_\alpha M^\alpha M_\alpha \right\}+\frac{\mathrm{i}}{4}\lambda \mathrm{d}^6\overline{z}\left\{\overline{W}_{\dot{\alpha }}\overline{W}^{\dot{\alpha }}\overline{M}_{\dot{\alpha }}\overline{M}^{\dot{\alpha }}\right\},$$ (2.6) in complete analogy with the analysis of for the non–supersymmetric case.<sup>3</sup><sup>3</sup>3Note that the action $`S`$ itself is not duality invariant, but rather the combination $`S\frac{\mathrm{i}}{4}\mathrm{d}^6zWM+\frac{\mathrm{i}}{4}\mathrm{d}^6\overline{z}\overline{W}\overline{M}`$. The invariance of this functional under a finite $`U(1)`$ duality rotation by $`\pi /2`$, is equivalent to the self-duality of $`S`$ under Legendre transformation, $`S[W,\overline{W}]\frac{\mathrm{i}}{2}\mathrm{d}^6zWW_\mathrm{D}+\frac{\mathrm{i}}{2}\mathrm{d}^6\overline{z}\overline{W}\overline{W}_\mathrm{D}=S[W_\mathrm{D},\overline{W}_\mathrm{D}]`$, with $`W_\mathrm{D}`$ being the dual chiral field strength. On the other hand, $`S`$ is a functional of $`W_\alpha `$ and $`\overline{W}_{\dot{\alpha }}`$ only, and therefore it variations under (2.5) is $$\delta S=\frac{\mathrm{i}}{2}\lambda \mathrm{d}^6zM^\alpha M_\alpha \frac{\mathrm{i}}{2}\lambda \mathrm{d}^6\overline{z}\overline{M}_{\dot{\alpha }}\overline{M}^{\dot{\alpha }}.$$ (2.7) Since these two variations must coincide, we arrive at the following reality condition $$\mathrm{d}^6z\left(W^\alpha W_\alpha +M^\alpha M_\alpha \right)=\mathrm{d}^6\overline{z}\left(\overline{W}_{\dot{\alpha }}\overline{W}^{\dot{\alpha }}+\overline{M}_{\dot{\alpha }}\overline{M}^{\dot{\alpha }}\right).$$ (2.8) In eq. (2.8), the superfields $`M_\alpha `$ and $`\overline{M}_{\dot{\alpha }}`$ are defined as in (2.3), and $`W_\alpha `$ and $`\overline{W}_{\dot{\alpha }}`$ should be considered as unconstrained chiral and antichiral superfields, respectively. Eq. (2.8) is the condition for the $`𝒩=1`$ supersymmetric theory to be duality invariant. We call it the $`𝒩=1`$ duality equation. A nontrivial solution of eq. (2.8) is the $`𝒩=1`$ supersymmetric Born-Infeld action (see also ) $`S_{\mathrm{BI}}`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle \mathrm{d}^6zW^2}+{\displaystyle \frac{1}{4}}{\displaystyle \mathrm{d}^6\overline{z}\overline{W}^2}+{\displaystyle \frac{1}{g^4}}{\displaystyle \mathrm{d}^8z\frac{W^2\overline{W}^2}{1+\frac{1}{2}A+\sqrt{1+A+\frac{1}{4}B^2}}},`$ (2.9) $`A`$ $`=`$ $`{\displaystyle \frac{1}{2g^4}}\left(D^2W^2+\overline{D}^2\overline{W}^2\right),B={\displaystyle \frac{1}{2g^4}}\left(D^2W^2\overline{D}^2\overline{W}^2\right),`$ where $`g`$ is a coupling constant. This is a model for a Goldstone multiplet associated with partial breaking of $`𝒩=2`$ to $`𝒩=1`$ supersymmetry (see also ), with $`W_\alpha `$ being the Goldstone multiplet. New examples of $`𝒩=1`$ duality invariant models can be obtained by considering a general action of the form (see also Appendix A) $$S=\frac{1}{4}\mathrm{d}^6zW^2+\frac{1}{4}\mathrm{d}^6\overline{z}\overline{W}^2+\frac{1}{2}\mathrm{d}^8zW^2\overline{W}^2L(D^2W^2,\overline{D}^2\overline{W}^2),$$ (2.10) where $`L(u,\overline{u})`$ is a real analytic function of the complex variable $`uD^2W^2`$ and its conjugate. One finds $$\mathrm{i}M_\alpha =W_\alpha \left\{\mathrm{\hspace{0.33em}1}\frac{1}{2}\overline{D}^2\left[\overline{W}^2\left(L+D^2\left(W^2\frac{L}{u}\right)\right)\right]\right\}.$$ (2.11) Then, eq. (2.8) leads to $$4\mathrm{d}^8zW^2\overline{W}^2\left(\mathrm{\Gamma }\overline{\mathrm{\Gamma }}\right)=\mathrm{d}^8zW^2\overline{W}^2\left(\mathrm{\Gamma }^2\overline{D}^2\overline{W}^2\overline{\mathrm{\Gamma }}^2D^2W^2\right),$$ (2.12) where $$\mathrm{\Gamma }L+\frac{L}{u}D^2W^2=\frac{(uL)}{u}.$$ (2.13) Since the latter functional relation must be satisfied for arbitrary (anti) chiral superfields $`\overline{W}_{\dot{\alpha }}`$ and $`W_\alpha `$, we arrive at the following differential equation for $`L(u,\overline{u})`$: $$4\left(\frac{(uL)}{u}\frac{(\overline{u}L)}{\overline{u}}\right)=\overline{u}\left(\frac{(uL)}{u}\right)^2u\left(\frac{(\overline{u}L)}{\overline{u}}\right)^2.$$ (2.14) Similar to the non–supersymmetric case , the general solution of this equation involves an arbitrary real analytic function of a single real argument, $`f(\overline{u}u)`$.<sup>4</sup><sup>4</sup>4Among non–supersymmetric duality invariant models, only the Maxwell action and the BI action satisfy the requirement of shock-free wave propagation . It is an easy exercise to check that the $`𝒩=1`$ BI action (2.9) satisfies eq. (2.14). We conclude this section by giving an extension of the model (2.9), in which the vector multiplet is coupled to an external chiral superfield $`\mathrm{\Phi }`$ in such a way that the system is not only duality invariant but also invariant under the $`𝒩=1`$ superconformal group. The action is $`S`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle \mathrm{d}^6zW^2}+{\displaystyle \frac{1}{4}}{\displaystyle \mathrm{d}^6\overline{z}\overline{W}^2}+{\displaystyle \mathrm{d}^8z\frac{W^2\overline{W}^2(\mathrm{\Phi }\overline{\mathrm{\Phi }})^2}{1+\frac{1}{2}𝐀+\sqrt{1+𝐀+\frac{1}{4}𝐁^2}}},`$ (2.15) $`𝐀`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{D^2}{\overline{\mathrm{\Phi }}^2}}\left({\displaystyle \frac{W^2}{\mathrm{\Phi }^2}}\right)+{\displaystyle \frac{\overline{D}^2}{\mathrm{\Phi }^2}}\left({\displaystyle \frac{\overline{W}^2}{\overline{\mathrm{\Phi }}^2}}\right)\right),𝐁={\displaystyle \frac{1}{2}}\left({\displaystyle \frac{D^2}{\overline{\mathrm{\Phi }}^2}}\left({\displaystyle \frac{W^2}{\mathrm{\Phi }^2}}\right){\displaystyle \frac{\overline{D}^2}{\mathrm{\Phi }^2}}\left({\displaystyle \frac{\overline{W}^2}{\overline{\mathrm{\Phi }}^2}}\right)\right).`$ Superconformal invariance follows from the superconformal transformation properties as given in . The theory is invariant under the duality rotations (2.5) with $`\mathrm{\Phi }`$ being inert. By its very construction, the action is also invariant under global phase transformations of $`\mathrm{\Phi }`$. In a sense, this model is analogous to the BI theory coupled to dilaton and axion fields . Similar to the analysis of , it is possible to show that the action (2.15) can be represented in the form $$S=\frac{1}{4}\mathrm{d}^6z𝐗+\frac{1}{4}\mathrm{d}^6\overline{z}\overline{𝐗},$$ (2.16) where the chiral superfield $`𝐗`$ is a functional of $`W_\alpha `$ and $`\overline{W}_{\dot{\alpha }}`$ such that it satisfies the nonlinear constraint $$𝐗+𝐗\frac{\overline{D}^2}{4\mathrm{\Phi }^2}\left(\frac{\overline{𝐗}}{\overline{\mathrm{\Phi }}^2}\right)=W^2.$$ (2.17) The $`𝒩=1`$ BI theory is obtained from this model by freezing $`\mathrm{\Phi }`$. More generally, for any duality invariant system defined by eqs. (2.10) and (2.14), the replacement $$W^2\overline{W}^2\frac{W^2\overline{W}^2}{\mathrm{\Phi }^2\overline{\mathrm{\Phi }}^2},D^2\frac{1}{\overline{\mathrm{\Phi }}^2}D^2\frac{1}{\mathrm{\Phi }^2}$$ (2.18) in (2.10) preserves the duality invariance but turns the action into a $`𝒩=1`$ superconformal functional. ## 3 $`𝒩`$ = 2 duality rotations We now generalize the results of the previous section to the case of $`𝒩=2`$ supersymmetry. We will work in $`𝒩=2`$ global superspace $`𝐑^{4|8}`$ parametrized by $`𝒵^A=(x^a,\theta _i^\alpha ,\overline{\theta }_{\dot{\alpha }}^i)`$, where $`i=\underset{¯}{1},\underset{¯}{2}`$. The flat covariant derivatives $`𝒟_A=(_a,𝒟_\alpha ^i,\overline{𝒟}_i^{\dot{\alpha }})`$ satisfy the standard algebra $$\{𝒟_\alpha ^i,𝒟_\beta ^j\}=\{\overline{𝒟}_{\dot{\alpha }i},\overline{𝒟}_{\dot{\beta }j}\}=0,\{𝒟_\alpha ^i,\overline{𝒟}_{\dot{\alpha }j}\}=2\mathrm{i}\delta _j^i(\sigma ^a)_{\alpha \dot{\alpha }}_a.$$ (3.1) Throughout this section, we will use the notation: $`𝒟^{ij}𝒟^{\alpha (i}𝒟_\alpha ^{j)}=𝒟^{\alpha i}𝒟_\alpha ^j,`$ $`\overline{𝒟}^{ij}\overline{𝒟}_{\dot{\alpha }}^{(i}\overline{𝒟}^{j)\dot{\alpha }}=\overline{𝒟}_{\dot{\alpha }}^i\overline{𝒟}^{j\dot{\alpha }}`$ $`𝒟^4{\displaystyle \frac{1}{16}}(𝒟^{\underset{¯}{1}})^2(𝒟^{\underset{¯}{2}})^2,`$ $`\overline{𝒟}^4{\displaystyle \frac{1}{16}}(\overline{𝒟}_{\underset{¯}{1}})^2(\overline{𝒟}_{\underset{¯}{2}})^2.`$ (3.2) An integral over the full superspace can be reduce to one over the chiral subspace or over the antichiral subspace as follows: $$\mathrm{d}^{12}𝒵(𝒵)=\mathrm{d}^8𝒵𝒟^4(𝒵)=\mathrm{d}^8\overline{𝒵}\overline{𝒟}^4(𝒵).$$ (3.3) ### 3.1 $`𝒩`$ = 2 duality equation The discussion in this subsection is completely analogous to the one presented in the first part of sect. 2. We will thus be brief. If $`𝒮[𝒲,\overline{𝒲}]`$ is the action describing the dynamics of a single $`𝒩=2`$ vector multiplet, the (anti) chiral superfield strengths $`\overline{𝒲}`$ and $`𝒲`$ are $$𝒲=\overline{𝒟}^4𝒟^{ij}V_{ij},\overline{𝒲}=𝒟^4\overline{𝒟}^{ij}V_{ij}$$ (3.4) in terms of a real unconstrained prepotential $`V_{(ij)}`$. The strengths then satisfy the Bianchi identity $$𝒟^{ij}𝒲=\overline{𝒟}^{ij}\overline{𝒲}.$$ (3.5) Suppose that $`𝒮[𝒲,\overline{𝒲}]`$ can be unambiguously defined as a functional of unconstrained (anti) chiral superfields $`\overline{𝒲}`$ and $`𝒲`$. Then, one can define (anti) chiral superfields $`\overline{}`$ and $``$ as $$\mathrm{i}4\frac{\delta }{\delta 𝒲}𝒮[𝒲,\overline{𝒲}],\mathrm{i}\overline{}4\frac{\delta }{\delta \overline{𝒲}}𝒮[𝒲,\overline{𝒲}]$$ (3.6) in terms of which the equations of motion read $$𝒟^{ij}=\overline{𝒟}^{ij}\overline{}.$$ (3.7) Again, since the Bianchi identity (3.5) and the equation of motion (3.7) have the same functional form, one can consider infinitesimal $`U(1)`$ duality transformations $$\delta 𝒲=\lambda ,\delta =\lambda 𝒲.$$ (3.8) Repeating the analysis of Gaillard and Zumino (see also section 2), we now have to impose $`\delta 𝒮`$ $`=`$ $`{\displaystyle \frac{\mathrm{i}}{8}}\lambda {\displaystyle \mathrm{d}^8𝒵\left(𝒲^2^2\right)}+{\displaystyle \frac{\mathrm{i}}{8}}\lambda {\displaystyle \mathrm{d}^8\overline{𝒵}\left(\overline{𝒲}^2\overline{}^2\right)}`$ $`=`$ $`{\displaystyle \frac{\mathrm{i}}{4}}\lambda {\displaystyle \mathrm{d}^8𝒵^2}{\displaystyle \frac{\mathrm{i}}{4}}\lambda {\displaystyle \mathrm{d}^8\overline{𝒵}\overline{}^2}`$ The theory is thus duality invariant provided the following reality condition is satisfied: $$\mathrm{d}^8𝒵\left(𝒲^2+^2\right)=\mathrm{d}^8\overline{𝒵}\left(\overline{𝒲}^2+\overline{}^2\right).$$ (3.10) Here $``$ and $`\overline{}`$ are defined as in (3.6), and $`𝒲`$ and $`\overline{𝒲}`$ should be considered as unconstrained chiral and antichiral superfields, respectively. Eq. (3.10) serves as our master functional equation to determine duality invariant models of the $`𝒩=2`$ vector multiplet. We remark that, as in the $`𝒩=1`$ case, the action itself is not duality invariant, but $$\delta \left(𝒮\frac{\mathrm{i}}{8}\mathrm{d}^8𝒵𝒲+\frac{\mathrm{i}}{8}\mathrm{d}^8\overline{𝒵}\overline{}\overline{𝒲}\right)=0.$$ (3.11) The invariance of the latter functional under a finite $`U(1)`$ duality rotation by $`\pi /2`$, is equivalent to the self-duality of $`𝒮`$ under Legendre transformation, $$𝒮[𝒲,\overline{𝒲}]\frac{\mathrm{i}}{4}\mathrm{d}^8𝒵𝒲𝒲_\mathrm{D}+\frac{\mathrm{i}}{4}\mathrm{d}^8\overline{𝒵}\overline{𝒲}\overline{𝒲}_\mathrm{D}=𝒮[𝒲_\mathrm{D},\overline{𝒲}_\mathrm{D}],$$ (3.12) where the dual chiral field strength $`𝒲_\mathrm{D}`$ is given by eq. (B.2). ### 3.2 $`𝒩`$ = 2 BI action Recently, Ketov suggested the following action $`𝒮_{\mathrm{BI}}`$ $`=`$ $`{\displaystyle \frac{1}{8}}{\displaystyle \mathrm{d}^8𝒵𝒲^2}+{\displaystyle \frac{1}{8}}{\displaystyle \mathrm{d}^8\overline{𝒵}\overline{𝒲}^2}+{\displaystyle \frac{1}{4}}{\displaystyle \mathrm{d}^{12}𝒵\frac{𝒲^2\overline{𝒲}^2}{1\frac{1}{2}𝒜+\sqrt{1𝒜+\frac{1}{4}^2}}},`$ (3.13) $`𝒜`$ $`=`$ $`𝒟^4𝒲^2+\overline{𝒟}^2\overline{𝒲}^2,=𝒟^4𝒲^2\overline{𝒟}^4\overline{𝒲}^2`$ as the $`𝒩=2`$ supersymmetric generalization of the BI action. We will first demonstrate that it indeed reduces to the $`𝒩=1`$ BI action. We then show that this condition is not strong enough to uniquely fix the $`𝒩=2`$ BI action but this is possible if, in addition, one imposes eq. (3.10). Given a $`𝒩=2`$ superfield $`U`$, its $`𝒩=1`$ projection is defined to be $`U|=U(𝒵)|_{\theta _{\underset{¯}{2}}=\overline{\theta }^{\underset{¯}{2}}=0}`$. The $`𝒩=2`$ vector multiplet contains two independent chiral $`𝒩=1`$ components $$𝒲|=\sqrt{2}\mathrm{\Phi },𝒟_\alpha ^{\underset{¯}{2}}𝒲|=2\mathrm{i}W_\alpha ,(𝒟^{\underset{¯}{2}})^2𝒲|=\sqrt{2}\overline{D}^2\overline{\mathrm{\Phi }}.$$ (3.14) Using in addition that $$\mathrm{d}^8𝒵=\frac{1}{4}\mathrm{d}^6z(𝒟^{\underset{¯}{2}})^2,\mathrm{d}^{12}𝒵=\frac{1}{16}\mathrm{d}^8z(𝒟^{\underset{¯}{2}})^2(\overline{𝒟}_{\underset{¯}{2}})^2,$$ (3.15) the free $`𝒩=2`$ vector multiplet action straightforwardly reduces to $`𝒩=1`$ superfields $$𝒮_{\mathrm{free}}=\frac{1}{8}\mathrm{d}^8𝒵𝒲^2+\frac{1}{8}\mathrm{d}^8\overline{𝒵}\overline{𝒲}^2=\mathrm{d}^8z\overline{\mathrm{\Phi }}\mathrm{\Phi }+\frac{1}{4}\mathrm{d}^6zW^2+\frac{1}{4}\mathrm{d}^6\overline{z}\overline{W}^2.$$ (3.16) If one switches off $`\mathrm{\Phi }`$, $$\mathrm{\Phi }=0(𝒟^{\underset{¯}{2}})^2𝒲|=0,$$ (3.17) the action (3.13) reduces to the $`𝒩=1`$ BI theory (2.9) (with $`g=1`$). However, as we will now demonstrate, there exist infinitely many $`𝒩=2`$ actions with that property.<sup>5</sup><sup>5</sup>5 The property $`W_\alpha W_\beta W_\gamma =0`$ of the $`𝒩=1`$ vector multiplet, which is crucial in the discussion of the $`𝒩=1`$ BI action, has no direct analog for its $`𝒩=2`$ counterpart. To demonstrate why this is possible, consider the following obviously different functionals $`{\displaystyle \mathrm{d}^{12}𝒵𝒲^2\overline{𝒲}^2\left\{(𝒟^4𝒲^2)^2\overline{𝒟}^4\overline{𝒲}^2+(\overline{𝒟}^4\overline{𝒲}^2)^2𝒟^4𝒲^2\right\}},`$ $`{\displaystyle \mathrm{d}^{12}𝒵𝒲^2\overline{𝒲}^2\left\{(𝒟^4𝒲^2)\overline{𝒟}^4\left[\overline{𝒲}^2𝒟^4𝒲^2\right]+(\overline{𝒟}^4\overline{𝒲}^2)𝒟^4\left[𝒲^2\overline{𝒟}^4\overline{𝒲}^2\right]\right\}}.`$ They coincide under (3.17). Therefore, the requirement of correct $`𝒩=1`$ reduction is too weak to fix a proper $`𝒩=2`$ generalization of the BI action<sup>6</sup><sup>6</sup>6It was claimed in that the action (3.13) is self-dual with respect to the $`𝒩=2`$ Legendre transformation. This is, however, not correct.. We suggest to search for a $`𝒩=2`$ generalization of the BI action as a solution of the $`𝒩=2`$ duality equation (3.10) compatible with the requirement to give the correct $`𝒩=1`$ reduction. We have checked to some order in perturbation theory that these two requirements uniquely fix the solution: $`𝒮_{\mathrm{BI}}`$ $`=`$ $`{\displaystyle \frac{1}{8}}{\displaystyle \mathrm{d}^8𝒵𝒲^2}+{\displaystyle \frac{1}{8}}{\displaystyle \mathrm{d}^8\overline{𝒵}\overline{𝒲}^2}+𝒮_{\mathrm{int}},`$ $`𝒮_{\mathrm{int}}`$ $`=`$ $`{\displaystyle \frac{1}{8}}{\displaystyle }\mathrm{d}^{12}𝒵𝒲^2\overline{𝒲}^2\{1+{\displaystyle \frac{1}{2}}(𝒟^4𝒲^2+\overline{𝒟}^4\overline{𝒲}^2)`$ $`+`$ $`{\displaystyle \frac{1}{4}}\left((𝒟^4𝒲^2)^2+(\overline{𝒟}^4\overline{𝒲}^2)^2\right)+{\displaystyle \frac{3}{4}}(𝒟^4𝒲^2)(\overline{𝒟}^4\overline{𝒲}^2)`$ $`+`$ $`{\displaystyle \frac{1}{8}}\left((𝒟^4𝒲^2)^3+(\overline{𝒟}^4\overline{𝒲}^2)^3\right)`$ $`+`$ $`{\displaystyle \frac{1}{2}}\left((𝒟^4𝒲^2)^2(\overline{𝒟}^4\overline{𝒲}^2)+(𝒟^4𝒲^2)(\overline{𝒟}^4\overline{𝒲}^2)^2\right)`$ $`+`$ $`{\displaystyle \frac{1}{4}}((𝒟^4𝒲^2)\overline{𝒟}^4\left[\overline{𝒲}^2𝒟^4𝒲^2\right]+(\overline{𝒟}^4\overline{𝒲}^2)𝒟^4\left[𝒲^2\overline{𝒟}^4\overline{𝒲}^2\right])\}+O(𝒲^{12}).`$ The expression in the last two lines of (3.2) constitutes the leading perturbative corrections where our solution of the duality equation (3.10) differs from the action (3.13). We now present the nonperturbative solution of (3.10) which reduces to the $`𝒩=1`$ BI action (2.9) under the condition (3.17). The action reads $$𝒮_{\mathrm{BI}}=\frac{1}{4}\mathrm{d}^8𝒵𝒳+\frac{1}{4}\mathrm{d}^8\overline{𝒵}\overline{𝒳},$$ (3.19) where the chiral superfield $`𝒳`$ is a functional of $`𝒲`$ and $`\overline{𝒲}`$ defined via the constraint<sup>7</sup><sup>7</sup>7The property $`𝐗^2=0`$ of the $`𝒩=1`$ constraint (2.17) has no direct analog for $`𝒳`$. $$𝒳=𝒳\overline{𝒟}^4\overline{𝒳}+\frac{1}{2}𝒲^2.$$ (3.20) Solving it iteratively for $`𝒳`$ one may verify the equivalence of (3.19) and (3.2) up to the indicated order. The constraint (3.20) was introduced in as a $`𝒩=2`$ generalization of that generating the $`𝒩=1`$ BI action (2.9) (see eq. (2.17)). It was also claimed in that the action (3.13) can be equivalently described by eqs. (3.19) and (3.20). This is clearly incorrect, since they lead to the action (3.2) rather than to (3.13). But the constraint (3.20) has a deep origin: the $`SL(2,𝐑)`$ invariant system introduced in admits a minimal $`𝒩=2`$ extension on the base of the constraint (3.20) such that the original $`SL(2,𝐑)`$ invariance remains intact. Let us prove that the system described by eqs. (3.19) and (3.20) provides a solution of the duality equation (3.10). Under an infinitesimal variation of $`𝒲`$ only, we have $`\delta _𝒲𝒳`$ $`=`$ $`\delta _𝒲𝒳\overline{𝒟}^4\overline{𝒳}+𝒳\overline{𝒟}^4\delta _𝒲\overline{𝒳}+𝒲\delta 𝒲,`$ $`\delta _𝒲\overline{𝒳}`$ $`=`$ $`\delta _𝒲\overline{𝒳}𝒟^4𝒳+\overline{𝒳}𝒟^4\delta _𝒲𝒳.`$ (3.21) From these relations one gets $$\delta _𝒲𝒳=\frac{1}{1𝒬}\left[\frac{𝒲\delta 𝒲}{1\overline{𝒟}^4\overline{𝒳}}\right],\delta _𝒲\overline{𝒳}=\frac{\overline{𝒳}}{1𝒟^4𝒳}𝒟^4\delta _𝒲𝒳,$$ (3.22) where $`𝒬=𝒫\overline{𝒫},`$ $`\overline{𝒬}=\overline{𝒫}𝒫,`$ $`𝒫={\displaystyle \frac{𝒳}{1\overline{𝒟}^4\overline{𝒳}}}\overline{𝒟}^4,`$ $`\overline{𝒫}={\displaystyle \frac{\overline{𝒳}}{1𝒟^4𝒳}}𝒟^4.`$ (3.23) With these results, it is easy to compute $``$: $$\mathrm{i}=\frac{𝒲}{1\overline{𝒟}^4\overline{𝒳}}\left\{1+\overline{𝒟}^4\overline{𝒫}\frac{1}{1𝒬}\frac{𝒳}{1\overline{𝒟}^4\overline{𝒳}}+\overline{𝒟}^4\frac{1}{1\overline{𝒬}}\frac{\overline{𝒳}}{1𝒟^4𝒳}\right\}.$$ (3.24) Now, a short calculation gives $$\mathrm{Im}\mathrm{d}^8𝒵\left\{^2+2\frac{1}{1𝒬}\frac{𝒳}{1\overline{𝒟}^4\overline{𝒳}}\right\}=0.$$ (3.25) On the other hand, the constraint (3.20) implies $$\mathrm{d}^8𝒵𝒳\mathrm{d}^8\overline{𝒵}\overline{𝒳}=\frac{1}{2}\mathrm{d}^8𝒵𝒲^2\frac{1}{2}\mathrm{d}^8\overline{𝒵}\overline{𝒲}^2,$$ (3.26) and hence $$\frac{\delta }{\delta 𝒲}\left\{\mathrm{d}^8𝒵𝒳\mathrm{d}^8\overline{𝒵}\overline{𝒳}\right\}=𝒲.$$ (3.27) The latter relation can be shown to be equivalent to $$\frac{1}{1𝒬}\frac{𝒳}{1\overline{𝒟}^4\overline{𝒳}}=𝒫\frac{1}{1\overline{𝒬}}\frac{\overline{𝒳}}{1𝒟^4𝒳}+𝒳.$$ (3.28) Using this result in eq. (3.25), we arrive at the relation $$\mathrm{d}^8𝒵^2\mathrm{d}^8\overline{𝒵}\overline{}^2=2\mathrm{d}^8𝒵𝒳+2\mathrm{d}^8\overline{𝒵}\overline{𝒳}$$ (3.29) which is equivalent, due to (3.26), to (3.10). In Appendix B we prove the self-duality of the $`𝒩=2`$ BI action under Legendre transformation explicitly, although this property already follows from the general analysis of or our discussion in subsect. 3.1. ### 3.3 Duality invariant $`𝒩`$ = 2 superconformal actions The $`𝒩=4`$ super Yang-Mills theory is believed to be self-dual . It was therefore suggested in to look for its low-energy effective action on the Coulomb branch as a solution to the self-duality equation via the $`𝒩=2`$ Legendre transformation such that the leading (second- and fourth- order) terms in the momentum expansion of the action look like $$𝒮_{\mathrm{lead}}=\frac{1}{8}\mathrm{d}^8𝒵𝒲^2+\frac{1}{8}\mathrm{d}^8\overline{𝒵}\overline{𝒲}^2+\frac{1}{4}c\mathrm{d}^{12}𝒵\mathrm{ln}𝒲\mathrm{ln}\overline{𝒲}+\mathrm{},$$ (3.30) where the third term represents the leading quantum correction computed in . In our opinion, the perturbative scheme of solving the self-duality equation via the $`𝒩=2`$ Legendre transformation is difficult as one has to invert the Legendre transformation. We suggest to look for the low-energy action of $`𝒩=4`$ SYM as a solution of the $`𝒩=2`$ duality equation (3.10). This equation is easy to deal with and it implies self-duality via Legendre transformation. The low-energy effective action we are looking for should be in addition invariant under the $`𝒩=2`$ superconformal group. This means that, along with the structures given in (3.30), the action may involve the following manifestly superconformal functionals $`𝒮_1`$ $`=`$ $`{\displaystyle \mathrm{d}^{12}𝒵\mathrm{ln}𝒲\mathrm{\Lambda }(\mathrm{ln}𝒲)}+\mathrm{c}.\mathrm{c}.,`$ (3.31) $`𝒮_2`$ $`=`$ $`{\displaystyle \mathrm{d}^{12}𝒵\mathrm{{\rm Y}}(\mathrm{ln}𝒲,\overline{}\mathrm{ln}\overline{𝒲})},`$ (3.32) where $$\frac{1}{\overline{𝒲}^2}𝒟^4,\overline{}\frac{1}{𝒲^2}\overline{𝒟}^4,$$ (3.33) and $`\mathrm{\Lambda }`$ and $`\mathrm{{\rm Y}}`$ are arbitrary holomorphic and real analytic functions, respectively. The superfields $`\mathrm{ln}𝒲`$ and $`\overline{}\mathrm{ln}\overline{𝒲}`$ prove to be superconformal scalars . The main property of the operators (3.33) is that, for any superconformal scalar $`\mathrm{\Psi }`$, $`\mathrm{\Psi }`$ and $`\overline{}\mathrm{\Psi }`$ are also superconformal scalars. In components, the functionals (3.30), (3.31) and (3.32) contain all possible structures which involve the physical scalar fields $`\phi =𝒲|_{\theta =0}`$ and the electromagnetic field strength $`F_{ab}`$ (where $`F_{\alpha \beta }𝒟_\alpha {}_{}{}^{i}𝒟_{\beta i}^{}𝒲|_{\theta =0}`$) without derivatives, along with terms containing derivatives and auxiliary fields. Simple power counting determines the necessary number of covariant derivatives in the action in order to produce a given power of $`F`$. Since $`F𝒟^2𝒲`$, there should be $`4n`$ $`𝒟`$’s in the superfield Lagrangian to get $`F^{4+2n}`$ (additional 8 derivatives come from the superspace measure, $`\mathrm{d}^{12}𝒵=\mathrm{d}^4x𝒟^4\overline{𝒟}^4`$). We are looking for a perturbative solution of (3.10) in the framework of the momentum expansion or, equivalently, as a series in powers of $``$ and $`\overline{}`$. But with the Ansatz $`𝒮=𝒮_{\mathrm{lead}}+𝒮_1+𝒮_2`$ it is easy to see that no solution of (3.10) exists. To obtain a consistent perturbation theory, we should allow for higher derivatives. More precisely, we should add new terms such that any number of operators $``$ and $`\overline{}`$ are inserted in the Taylor expansion of $`\mathrm{{\rm Y}}`$ (3.32). In other words, $`𝒮_2`$ should be extended to a more general functional $`\widehat{𝒮}_2`$ which can be symbolically written as <sup>8</sup><sup>8</sup>8There exist more general superconformal invariants of the $`𝒩=2`$ vector multiplet , as compared to the action (3.34), and some of them were determined in from the requirement of scale and $`U(1)_R`$ invariance. It suffices for our purposes that (3.34) provides a consistent Ansatz to solve the $`𝒩=2`$ duality equation (3.10). $$\widehat{𝒮}_2=\mathrm{d}^{12}𝒵\widehat{\mathrm{{\rm Y}}}(\mathrm{ln}𝒲,\overline{}\mathrm{ln}\overline{𝒲},,\overline{}).$$ (3.34) For the action $$𝒮[𝒲,\overline{𝒲}]=𝒮_{\mathrm{lead}}+𝒮_1+\widehat{𝒮}_2$$ (3.35) the equation of motion can be represented in terms of $$\mathrm{i}4\frac{\delta }{\delta 𝒲}𝒮[𝒲,\overline{𝒲}]=𝒲\left\{\mathrm{\hspace{0.17em}1}+\overline{}\mathrm{\Gamma }\right\},$$ (3.36) for some functional $`\mathrm{\Gamma }(\mathrm{ln}𝒲,\mathrm{ln}\overline{𝒲},,\overline{})`$ such that $`\mathrm{\Gamma }=c\mathrm{ln}\overline{𝒲}+O()`$. Then, the duality equation (3.10) is equivalent to $$\mathrm{Im}\mathrm{d}^{12}𝒵\left\{2\mathrm{\Gamma }+\mathrm{\Gamma }\overline{}\mathrm{\Gamma }\right\}=0.$$ (3.37) In the framework of perturbation theory, the procedure of solving of eq. (3.37) amounts to simple algebraic operations. To low order in the perturbation theory, the solution reads $`𝒮`$ $`=`$ $`{\displaystyle \frac{1}{8}}{\displaystyle \mathrm{d}^8𝒵𝒲^2}+{\displaystyle \frac{1}{8}}{\displaystyle \mathrm{d}^8\overline{𝒵}\overline{𝒲}^2}+{\displaystyle \frac{1}{4}}{\displaystyle \mathrm{d}^8\overline{𝒵}},`$ $``$ $`=`$ $`c\mathrm{ln}𝒲\mathrm{ln}\overline{𝒲}+{\displaystyle \frac{1}{4}}c^2(\mathrm{ln}𝒲\mathrm{ln}𝒲+\mathrm{c}.\mathrm{c}.)`$ (3.38) $`+{\displaystyle \frac{1}{4}}c^3d(\mathrm{ln}𝒲)\overline{}\mathrm{ln}\overline{𝒲}{\displaystyle \frac{1}{8}}c^3(\mathrm{ln}𝒲(\mathrm{ln}𝒲)^2+\mathrm{c}.\mathrm{c}.)`$ $`+{\displaystyle \frac{1}{16}}c^4((14d)(\mathrm{ln}𝒲)^2\overline{}\mathrm{ln}\overline{𝒲}+(2d1)(\mathrm{ln}𝒲)\overline{}\mathrm{ln}𝒲`$ $`+{\displaystyle \frac{5}{3}}\mathrm{ln}𝒲(\mathrm{ln}𝒲)^3+\mathrm{c}.\mathrm{c}.)+O(^4).`$ Here $`d`$ is the first parameter in the derivative expansion of $`𝒮`$ which is not fixed by the $`𝒩=2`$ duality equation (3.10). Note that if we had only imposed the condition of self-duality under Legendre transformation, as was done in , we could not have fixed the coefficent $`\frac{1}{8}c^3`$ of the fourth term in $``$. In general, for any self-conjugate monomial in the expansion of $`𝒮`$, like $`(\mathrm{ln}𝒲)\overline{}\mathrm{ln}\overline{𝒲}`$, the corresponding coefficient is not determined by eq. (3.10) in terms of those appearing in the structures in $`𝒮`$ with less derivatives. However, such coefficients can be fixed if one imposes some additional conditions on the solution of eq. (3.10). For example, one can require the solution to reduce to a given $`𝒩=1`$ action under the condition $`𝒲|=\mathrm{const}`$. It should be pointed out that the $`c^3`$–corrections in (3.38) have been determined in by solving the self-duality equation via the $`𝒩=2`$ Legendre transformation. To compute the $`𝒪(c^4)`$ term via the duality equation (3.10) involves only elementary algebraic manipulations. As is seen from (3.38), solutions of the duality equation (3.10) contain higher derivative structures $`\overline{}\mathrm{ln}𝒲`$, $`\overline{}\mathrm{ln}𝒲`$, etc. What is the fate of such terms? The striking result of is the fact that, to the order $`c^3`$, there exists a nonlinear $`𝒩=1`$ superfield redefinition which eliminates all higher derivative (accelerating) component structures (contained already in the first term of $``$ (3.38)). The price for such a redefinition is that the original linear $`𝒩=2`$ supersymmetry turns into a nonlinear one being typical for D3-brane actions . The nonlinear redefinition of eliminates the higher derivative terms to some order of perturbation theory, but it in turn generates new such terms at higher orders in the momentum expansion. Therefore, in order for such a nonlinear redefinition to be consistently defined, the superfield action should involve higher derivatives of arbitrary order. The duality equation (3.10) might guarantee the existence of a consistent redefinition to eliminate acceleration terms. Acknowledgements We are grateful to Gleb Arutyunov, Evgeny Ivanov and Arkady Tseytlin for helpful discussions. Support from DFG, the German National Science Foundation, from GIF, the German-Israeli foundation for Scientific Research and from the EEC under TMR contract ERBFMRX-CT96-0045 is gratefully acknowledged. This work was also supported in part by the NATO collaborative research grant PST.CLG 974965, by the RFBR grant No. 99-02-16617, by the INTAS grant No. 96-0308 and by the DFG-RFBR grant No. 99-02-04022. ## Appendix A $`𝒩=0`$ duality invariant models In this appendix we give several equivalent forms of the Gallard-Zumino equation (1.3) by representing the Lagrangian $`L(F_{ab})`$ as a real function of one complex variable, $`L(F_{ab})=L(𝒰,\overline{𝒰}),𝒰=+\mathrm{i}𝒢,`$ $`={\displaystyle \frac{1}{4}}F^{ab}F_{ab},𝒢={\displaystyle \frac{1}{4}}F^{ab}\stackrel{~}{F}_{ab}.`$ (A.1) The theory is parity invariant iff $`L(𝒰,\overline{𝒰})=L(\overline{𝒰},𝒰)`$. One calculates $`\stackrel{~}{G}`$ (1.2) to be $$\stackrel{~}{G}_{ab}=\left(F_{ab}+\mathrm{i}\stackrel{~}{F}_{ab}\right)\frac{L}{𝒰}+\left(F_{ab}\mathrm{i}\stackrel{~}{F}_{ab}\right)\frac{L}{\overline{𝒰}},$$ (A.2) and the Gallard-Zumino equation (1.3) takes the form $$\mathrm{Im}\left\{𝒰4𝒰\left(\frac{L}{𝒰}\right)^2\right\}=0,$$ (A.3) which is equivalent to the equations obtained in but turns out to be more convenient for supersymmetric generalizations. If one splits $`L`$ into the sum of Maxwell’s part and an interaction, $$L=\frac{1}{2}\left(𝒰+\overline{𝒰}\right)+L_{\mathrm{in}},L_{\mathrm{in}}=𝒪(|𝒰|^2),$$ (A.4) the above equation turns into $$\mathrm{Im}\left\{𝒰\frac{L_{\mathrm{in}}}{𝒰}𝒰\left(\frac{L_{\mathrm{in}}}{𝒰}\right)^2\right\}=0.$$ (A.5) We restrict $`L_{\mathrm{in}}`$ to be a real analytic function of $`𝒰`$ and $`\overline{𝒰}`$. Then, every solution of eq. (A.5) is of the form $$L_{\mathrm{in}}(𝒰,\overline{𝒰})=𝒰\overline{𝒰}\mathrm{\Omega }(𝒰,\overline{𝒰}),\mathrm{\Omega }=𝒪(1),$$ (A.6) where $`\mathrm{\Omega }`$ satisfies $$\mathrm{Im}\left\{\frac{(𝒰\mathrm{\Omega })}{𝒰}\overline{𝒰}\left(\frac{(𝒰\mathrm{\Omega })}{𝒰}\right)^2\right\}=0.$$ (A.7) Note that for any solution $`L_{\mathrm{in}}(𝒰,\overline{𝒰})`$ of (A.5), or any solution $`\mathrm{\Omega }(𝒰,\overline{𝒰})`$ of (A.7), the functions $$\widehat{L}_{\mathrm{in}}(𝒰,\overline{𝒰})=\frac{1}{\kappa ^2}L_{\mathrm{in}}(\kappa ^2𝒰,\kappa ^2\overline{𝒰}),\widehat{\mathrm{\Omega }}(𝒰,\overline{𝒰})=\kappa ^2\mathrm{\Omega }(\kappa ^2𝒰,\kappa ^2\overline{𝒰})$$ (A.8) are also solutions of eqs. (A.5) and (A.7), respectively, for arbitrary real parameter $`\kappa ^2`$. Up to a trivial rescaling, eq. (A.7) coincides with the $`𝒩=1`$ duality equation (2.14). Therefore, any non-supersymmetric duality invariant model admits a $`𝒩=1`$ supersymmetric extension given by eqs. (2.10) and (2.14). It is easy to read off the bosonic sector of the action (2.10). For vanishing fermionic fields, $`W_\alpha |_{\theta =0}=0`$, one finds $$\frac{1}{8}D^2W^2|_{\theta =0}=𝒰2𝒟^2,$$ (A.9) where $`𝒟(x)`$ is the auxiliary field of the $`𝒩=1`$ vector multiplet. If we take the solution $`𝒟=0`$ of the equation of motion for $`𝒟`$, then the action (2.10) reduces to a generic $`𝒩=0`$ duality invariant model. ## Appendix B $`𝒩=2`$ BI action and Legendre transformation To prove that the system defined by eqs. (3.19) and (3.20) is self-dual under Legendre transformation, we replace the action (3.19) by the following one $$𝒮=\frac{1}{4}\mathrm{d}^8𝒵\left\{𝒳[𝒲,\overline{𝒲}]\mathrm{i}𝒲𝒲_\mathrm{D}\right\}+\frac{1}{4}\mathrm{d}^8\overline{𝒵}\left\{\overline{𝒳}[𝒲,\overline{𝒲}]+\mathrm{i}\overline{𝒲}\overline{𝒲}_\mathrm{D}\right\},$$ (B.1) where $`𝒲`$ is now considered to be an unconstrained chiral superfield, and its dual chiral strength $`𝒲_\mathrm{D}`$ reads $$𝒲_\mathrm{D}=\overline{𝒟}^4𝒟^{ij}U_{ij},$$ (B.2) with $`U_{ij}`$ an unconstrained real prepotential. The equation of motion for $`U_{ij}`$ implies the Bianchi identity (3.5), and hence the action reduces to (3.19). On the other hand, varying the action with respect to $`𝒲`$ leads to $$𝒲_\mathrm{D}=,$$ (B.3) where $``$ is given in eq. (3.24). The latter equation can be solved to express $`𝒲`$ in terms of $`𝒲_\mathrm{D}`$ and its conjugate. Instead of doing this explicitly, we note that eqs. (3.28) and (B.3) allow one to rewrite the action (B.1) as $$𝒮=\frac{1}{4}\mathrm{d}^8𝒵𝒳_\mathrm{D}+\frac{1}{4}\mathrm{d}^8\overline{𝒵}\overline{𝒳}_\mathrm{D},$$ (B.4) where $$𝒳_\mathrm{D}\frac{1}{1𝒬}\frac{𝒳}{1\overline{𝒟}^4\overline{𝒳}}𝒫\frac{1}{1\overline{𝒬}}\frac{\overline{𝒳}}{1𝒟^4𝒳}.$$ (B.5) Using eqs. (3.28) and (B.3) once more, one can prove that $`𝒳_\mathrm{D}`$ satisfies the constraint $$𝒳_\mathrm{D}=𝒳_\mathrm{D}\overline{𝒟}^4\overline{𝒳}_\mathrm{D}+\frac{1}{2}𝒲_\mathrm{D}{}_{}{}^{2}.$$ (B.6) This completes the proof.
warning/0001/quant-ph0001043.html
ar5iv
text
# Quantization of Soliton Cellular Automata 11footnote 1Submitted to Journal of Nonlinear Mathematical Physics. Special Issue of Proccedings of NEEDS’99. ## Abstract A method of quantization of classical soliton cellular automata (QSCA) is put forward that provides a description of their time evolution operator by means of quantum circuits that involve quantum gates from which the associated Hamiltonian describing a quantum chain model is constructed. The intrinsic parallelism of QSCA, a phenomenon first known from quantum computers, is also emphasized. Introduction. Soliton cellular automata (SCA), is a class of cellular automata operating on binary sequences with an updating rule function $`f`$ for each cell, that depends on past and present time cells, the number of which determines the radius r of the automaton. Contrary to usual CA ( that evolve their cells by means of past time cells only ), SCA exhibit a variety of evolution patterns that is mainly known to characterize the temporal behavior of solutions of non linear PDE’s, namely: periodic evolution of particles (i.e localized groups of binary cells ), or solitonic type of scattering of digital particles, or even breathing modes of oscillations between particles. All these properties have motivated a number of suggestive applications for a new kind of computational architecture that will utilize these evolution patterns of SCA in order to provide a ”gateless” implementation of logical operations. Towards a physical microscopic realization of these suggestions, envisaged in the context of the new paradigm of Quantum Computing, it is plausible to formulate SCA in terms of Quantum Mechanics and to investigate the possible quantum effects in their time evolution. To this end we put forward here the quantization of classical SCA, and point out their quantum parallelism. Quantization. Let the product Hilbert space $`=_{i\text{Z}\text{Z}}_i`$ where $`_i=span\{|0,|1\}\text{C}|^2`$. Take $`n𝐍`$, and consider the subspace $`H`$ where $`H=_{k=r}^r_{nk}`$, then define the vectors $$|a^t_{i=r}^1|a_{ni}^{t+1}|a_n^t_{j=1}^r|a_{n+j}^tH,$$ (1) and the dual vectors $`a^t|\stackrel{~}{H}`$, with orthogonormality relation $`a^t|b^t`$ $`=`$ $`\mathrm{\Pi }_{i=r}^1a_{ni}^{t+1}|b_{ni}^{t+1}a_n^t|b_n^t\mathrm{\Pi }_{j=1}^ra_{n+j}^t|b_{n+j}^t`$ (2) $`=`$ $`\mathrm{\Pi }_{i=r}^1\delta _{a_{ni}^{t+1},b_{ni}^{t+1}}\delta _{a_n^t,b_n^t}\mathrm{\Pi }_{j=1}^r\delta _{a_{n+j}^t,b_{n+j}^t}.`$ Recall that for $`(x,y)\text{Z}\text{Z}_2^2`$ the Kronecker delta function is defined as $`\delta _{x,y}=1xy`$, where $``$ denotes $`\mathrm{𝚇𝙾𝚁}`$, i.e modulo-$`2`$ addition. Now take $`\text{Z}\text{Z}_2^{2r+1}\text{Z}\text{Z}_2^{2r+1}O^{2r+1}`$, where $`O^{2r+1}`$ is the $`2r+1`$-fold null string, $`\text{Z}\text{Z}_2^{2r+1}\text{Z}\text{Z}_2^{2r+1}\{a_0\}`$ where $`f(\{a_0\})=O^{2r+1}`$, the preimage of the null string, and define the one-to-one function $`f:\text{Z}\text{Z}_2^{2r+1}\text{Z}\text{Z}_2^{2r+1}`$, as $`f(a^t)=a^{t+1}`$, where $`a^t\{a_{nr}^{t+1},\mathrm{},a_{n1}^{t+1},a_n^t,a_{n+1}^t,\mathrm{},a_{n+r}^t\}`$, and $`a^{t+1}\{a_{nr}^{t+1},\mathrm{},a_{n1}^{t+1},a_n^{t+1},a_{n+1}^t,\mathrm{},a_{n+r}^t\}`$, where the updated bit takes the value $`a_n^{t+1}=1_{i=r}^1a_{ni}^{t+1}_{j=0}^ra_{n+j}^t`$. Introduce now the quantization of classical cellular automaton by means of the following quantization diagram. $$\begin{array}{ccccc}& \text{Z}\text{Z}_2^{2r+1}& \stackrel{f}{}& \text{Z}\text{Z}_2^{2r+1}& \\ \hfill \rho & & & & \rho \hfill \\ & H& \begin{array}{c}\\ U_f\end{array}& H& \end{array}$$ This diagram implies that: $$U_f\rho (a^t)=U_f|a^t=|f(a^t)=|a^{t+1}=\rho f(a^t)=\rho (a^{t+1})=|a^{t+1}.$$ (3) The transition operator $`U_f`$ implements in the space of qubits $`H`$ the update rule $`f`$ of the classical bits of the automaton, and acts trivially as the unit operator in the rest space. Since in order to establish the one-to-one property for $`f`$ we have excluded the null string from its domain of values the associated operator $`U_f\mathrm{𝐸𝑛𝑑}H`$, is a partial isometry in $`H_0^{}`$, the orthogonal complement space of $`H_0\mathrm{𝑠𝑝𝑎𝑛}\{|0\}`$, assuming the decomposition $`H=H_0+H_0^{}`$. To be more specific, let $`U_f=_{a^t\text{Z}\text{Z}_2^{2r+1}}\left|f(a^t)a^t\right|`$, then its isometric property is due to the following relations: $`U_fU_{f}^{}{}_{}{}^{}`$ $`=`$ $`{\displaystyle \underset{(a^t,b^t)\text{Z}\text{Z}_2^{2r+1}}{}}|f(a^t)a^t|b^tf(b^t)|={\displaystyle \underset{a^t\text{Z}\text{Z}_2^{2r+1}\{a_0^t\}}{}}|a^ta^t|,`$ $`U_{f}^{}{}_{}{}^{}U_f`$ $`=`$ $`{\displaystyle \underset{(a^t,b^t)\text{Z}\text{Z}_2^{2r+1}}{}}|b^tf(b^t)|f(a^t)a^t|={\displaystyle \underset{a^t\text{Z}\text{Z}_2^{2r+1}\{0\}}{}}|a^ta^t|,`$ (4) where the preimage of the null string for e.g the simplest case of $`r=2`$ is $`a_0^t=00100`$. Choosing a basis of vectors in $`H`$, e.g the computational basis $`\{|x_1\mathrm{}|x_{2r+1}:\{x_i\}_{i=1}^{2r+1}\text{Z}\text{Z}_2^{2r+1}\}`$, we can obtain a $`2^{2r+1}\times 2^{2r+1}`$ matrix representation of $`U_f`$ viz. $`\pi _r(U_f)=`$ $`{\displaystyle \underset{a^t\text{Z}\text{Z}^{2r+1}}{}}_{i=r}^1\left(\begin{array}{cc}\delta _{0,a_{ni}^{t+1}}& 0\\ 0& \delta _{1,a_{ni}^{t+1}}\end{array}\right)\left(\begin{array}{cc}\delta _{0,a_n^{t+1}}\delta _{0,a_n^t}& \delta _{0,a_n^{t+1}}\delta _{1,a_n^t}\\ \delta _{1,a_n^{t+1}}\delta _{0,a_n^t}& \delta _{1,a_n^{t+1}}\delta _{1,a_n^t}\end{array}\right)_{j=1}^r\left(\begin{array}{cc}\delta _{0,a_{n+j}^t}& 0\\ 0& \delta _{1,a_{n+j}^t}\end{array}\right).`$ (11) More explicitly if we partition the basis $``$ into two orthogonal compliments corresponding to invariant and non invariant subspaces of $`U_f`$ we will obtain for e.g the case of radius $`r=2`$, the transition matrix : $$\pi _2(U_f)=\left(\begin{array}{cc}\mathrm{𝟏}& \mathrm{𝟎}\\ \mathrm{𝟎}& \sigma _1^{^{}}\end{array}\right),$$ (12) where $`\mathrm{𝟏}`$,$`\mathrm{𝟎}`$, stand for the square $`16`$-dimensional unit and null matrices respectively, while $`\sigma _1^{^{}}`$, is a $`N=16`$-dimensional matrix with elements along the main antidiagonal $`(a_{1,N},a_{2,N1},\mathrm{},a_{N,1})=(1,\mathrm{},1,0)`$, and all others been zero. Fig. 1. The factorization of $`U_f`$ in terms of quantum gates for radius $`r=2`$. Fig. 2. Quantum circuit implementation of FRT for three quantum BS’s with $`r=2`$. Hamiltonian model. In order to construct a Hamiltonian model associated with the QSCA that generates the total evolution of the automaton, and in view of the fact that the time step evolution operator is determined by unitary $`\mathrm{𝙽𝙾𝚃}`$ viz. $`U_N^i|i=|1i`$ and $`\mathrm{𝙲𝙾𝙽𝚃𝚁𝙾𝙻}\mathrm{𝙽𝙾𝚃}(\mathrm{𝙲𝙽})`$ viz. $`U_{CN}^{ij}|ij=|iij`$ gate operators, cf. its explicit factorization in Fig. 1, we recall here first the Hermitian operators corresponding to those gates \[$`|i|j`$ is abbreviated to $`|ij`$\]. The quantum negation of the $`i`$-th qubit is given by the matrix $`U_N^i=\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right)_i`$ acting non trivially only in the $`i`$-th subspace, with associated Hermitian generator given by $`H_N^i=\frac{1}{2}(\sigma _z^i+\sigma _x^i)`$, in terms of Pauli sigma matrices. For the conditional negation of $`\mathrm{𝙲𝙽}`$ gate with $`i`$-control and $`j`$-target qubits and $`i<j`$ the unitary matrix $`U_{CN}^{ij}=\left|00\right|\mathrm{𝟏}+\left|11\right|\sigma _x=\mathrm{exp}(i\pi H_{CN}^{ij})`$ is generated by the Hamiltonian $`H_{CN}^{ij}=\frac{1}{2}(\mathrm{𝟏}\sigma _z^i)(\sigma _x^j\mathrm{𝟏})`$. Correspondingly for $`i>j`$, $`U_{CN}^{ij}=\mathrm{𝟏}\left|00\right|+\sigma _x\left|11\right|=\mathrm{exp}(i\pi H_{CN}^{ij})`$, and $`H_{CN}^{ij}=\frac{1}{2}(\mathrm{𝟏}\sigma _z^j)(\sigma _x^i\mathrm{𝟏})`$. These gate operators can be utilized to factorize the $`U_f`$ that is used to update each qubit in a given string of qubits of a QSCA. To this end in Fig.1 we construct the quantum circuit that provides such a factorization for the special case $`r=2`$; the generalization to any $`r`$ is straightforward. Specifically the input state vector in Fig. 1 reads $`|a^t|a_{n2}^{t+1},a_{n1}^{t+1},a_n^t,a_{n+1}^t,a_{n+2}^t`$, and is tranformed by $`U_f`$ to $`|a^{t+1}U_f|a^t=|a_{n2}^{t+1},a_{n1}^{t+1},a_n^{t+1},a_{n+1}^t,a_{n+2}^t=U_N^nU_{CN}^{n+2,n}U_{CN}^{n+1,n}U_{CN}^{n1,n}U_{CN}^{n2,n}|a_n^t`$, where $`|a_n^{t+1}=|1a_{n2}^{t+1}a_{n1}^{t+1}a_n^ta_{n+1}^ta_{n+2}^t`$. In Fig. 1 graphically the $`\mathrm{𝙲𝙽}`$ gate is indicated as a vertical line segment with a bullet (control qubit) and a crossed circle (target qubit) at its ends . Then the total unitary evolution operator $`U_f=_i^{\mathrm{}}U_i^f\mathrm{}U_2^fU_1^fU_0^f`$, is the product of update operators for each qubit in the automaton starting with the first one $`U_0^f`$, from which all later ones are determined by shifting i.e $`U_{i+1}^f=\mathrm{𝟏}U_i^f`$. The total Hamiltonian $`H_f=lim_{t0}\frac{1}{i}\frac{dU_f}{dt}=_{i0}H_i^f`$, is the sum of all Hamiltonians generationg each evolution operator $`\{U_i^f,i0\}`$. Since $`U_i^f=U_N^i_{k=1}^rU_{CN}^{i+k,i}_{k=1}^rU_{CN}^{ik,i}`$, the corresponding Hamiltonian at each site $`i`$ turns out to be $`H_i^f=H_N^i+_{k=1}^r(H_{CN}^{i+k,i}+H_{CN}^{ik,i}).`$ In view of the generators of the quantum gates given above, the Hamiltonian model of the QSCA reads, $$H_f=\underset{i0}{}\frac{1}{2}[(\sigma _x^i+\sigma _z^i)+\underset{k=1}{\overset{r}{}}(\mathrm{𝟏}^{i+k}\sigma _z^{i+k})(\sigma _x^i\mathrm{𝟏}^i)+(\mathrm{𝟏}^{ik}\sigma _z^{ik})(\sigma _x^i\mathrm{𝟏}^i)].$$ (13) This can be interpreted as the Hamiltonian operator of a infinite quantum spin chain model with free boundaries and interactions ranged over $`r`$ neighbors. Obvious modifications such that QCSA with periodic boundary conditions, or higher dimensions are in the present formalism appealing and worth of further study. Quantum Fast Rule Theorem. We turn now to study QSCA in terms of the so called Fast Rule Theorem(FRT) that provides analytic tools for the classification and the study of types of dynamical evolution of classical SCA. In the limited space available here we shall confine ourselves to the special case of periodic particles and provide a quantum circuit of operators that governs the time development of periodic qubit-particles, i.e the quantum analogue of periodic particles of classical CA theory. We recall that a particle is an integer multiple of $`r+1`$ consecutive sites and it can be a collection of the so called basic strings (BS), which are $`r+1`$ consecutive sites starting with a boxed site. If a particle consists of a single BS, we call it a simple particle. Then a basic theorem states the following: Let a single particle $`OA^1A^2\mathrm{}A^LO`$ with $`A^1O,A^LO`$, consisting of $`L`$ basic strings $`A^1,A^2,\mathrm{},A^L`$ (here $`O`$ denotes the null BS consisting of $`r+1`$ zeros), and let $``$ denotes sitewise $`\mathrm{𝚇𝙾𝚁}`$ among bits of the BS. Suppose that the BS’s $`A^1,A^1A^2,A^2A^3,\mathrm{},A^{L1}A^L,A^l`$ , contain $`l_0,l_1,\mathrm{},l_L`$ $`1`$’s respectively. Then if such a particle does not split or loose any BS’s from its right end at all times during its evolution, then at times $`t=l_0+l_1+\mathrm{}l_m,mkL`$, it becomes $`A^{m+1}(A^{m+2}A^{m+3}\mathrm{}A^LA^0A^1\mathrm{}A^m)`$. Especially for $`k=L`$, i.e at time $`t=l_0+l_1+\mathrm{}l_L`$ it returns to its initial state $`A^1A^2\mathrm{}A^L`$. To illustrate the workings of the quantum FRT i.e the quantum adaptation of the preceding theorem, let us consider an initial particle of $`L=3`$ classical BS’s $`A^1A^2A^3`$. Such a classical particle by means of the quantization map $`\rho `$ becomes the qubit word $`|A^1A^2A^3OOOO`$. Let us assume that the classical word labelling this state vector evolves under the previously stated conditions so that the classical FRT applies and assures that it has a periodic behaviour. Then the quantization map induces into the quantum state vector the periodic evolution of its classical label word. Namely the quantum state vector evolves periodically, and we seek to determine the evolution operator that implemends this periodic motion. Assuming we have a sufficient number of null quantum BS’s from each side of the initial particle state i.e $`|A^1A^2A^3OOOO`$, let us consider the following two operators. First, the operator $`𝒫_0^I=_{x\text{Z}\text{Z}^{r+1}}\left|0^{r+1}x\right|`$, which projects each $`r+1`$-fold tensor product BS state vector onto the $`r+1`$-fold null state vector. With the upper index $`I`$ denotes the fact that the $`(r+1)`$-fold input vector is placed in the $`I`$-th position of a given chain of tensor product of state vector, and that $`𝒫_0^I`$ acts non trivial only in that $`I`$-th subspace. Second, let us define the collective $`\mathrm{𝙲𝙽}`$ gate operator $`𝒰_{CN}^{IJ}U_{CN}^{I_{r+1}J_{r+1}}\mathrm{}U_{CN}^{I_2J_2}U_{CN}^{I_1J_1}`$, with control BS placed in the $`I`$-th position and target BS placed in the $`J`$-th position. This collective $`\mathrm{𝙲𝙽}`$ gate is actually defined as a succesion of $`(r+1)`$ $`\mathrm{𝙲𝙽}`$ gates for the corresponding qubits of the $`(r+1)`$-fold tensor product control and the target BS states vectors. In Fig. 2 we indicate graphically the sequence of operations that tranform the initial state vector $`|A^1A^2A^3OOOO`$ into the final state $`|OOOOA^1A^2A^3`$, indicating in this way its periodic $`\mathrm{𝑝𝑟𝑜𝑝𝑎𝑔𝑎𝑡𝑖𝑜𝑛}`$. Each parallel wire in the figure represents an $`(r+1)`$-fold tensor product of qubits and the operators $`𝒫_0`$ are indicated by boxed zeros, while operators $`𝒰_{CN}^{IJ}`$ are indicated a double circled $`\mathrm{𝙲𝙽}`$ gate symbol. Explicitly the sequence of operations of the quantum circuit in Fig. 2 reads: $`|A^1A^2A^3OOOO`$ $``$ $`|OA^1A^2A^1A^3A^1OOO`$ $`=`$ $`𝒫_0^1𝒰_{CN}^{14}𝒰_{CN}^{13}𝒰_{CN}^{12}|A^1A^2A^3OOOO`$ $`|OOA^2A^3A^2A^1A^2OO`$ $`=`$ $`𝒫_0^2𝒰_{CN}^{25}𝒰_{CN}^{24}𝒰_{CN}^{23}|OA^1A^2A^1A^3A^1OOO`$ $`|OOOA^3A^1A^3A^2A^3O`$ $`=`$ $`𝒫_0^3𝒰_{CN}^{36}𝒰_{CN}^{35}𝒰_{CN}^{34}|OOA^2A^3A^2A^1A^2OO`$ $`|OOOOA^1A^2A^3`$ $`=`$ $`𝒫_0^4𝒰_{CN}^{47}𝒰_{CN}^{46}𝒰_{CN}^{45}|OOOA^3A^1A^3A^2A^3O.`$ (14) Quantum parallelism. The quantum description of the cells of a CA in terms of qubits allows for having a superposition $`a|0+b|1`$ at each cell. According to the standard interpretation of Quantum Mechanics this combination means that we have the state $`|0`$($`|1`$) with probability $`|a|^2`$($`|b|^2`$). This is the inherent probabilistic character of QCA that entails two important advantages of QCA over classical probabilistic (noisy) CA (see e.g ). First, there is an exponential overhead in storing cell values in QCA over classical CA: an $`r`$-radious probabilistic SCA needs at each discrete time step to store $`N=2^{2r+1}`$ input words in order to process them later on and to update the current cell value. On the contrary a QSCA may form and admit as input a superposition state vector $`|\psi =\frac{1}{\sqrt{N}}_{x\text{Z}\text{Z}_2^{2r+1}}|x\frac{1}{\sqrt{N}}_{x\text{Z}\text{Z}_2^{2r+1}}|x_1\mathrm{}x_{2r+1}`$, which with resources linear in $`r`$ i.e $`2r+1`$ qubits only can constructs an exponential in $`r`$ register of $`2^{2r+1}`$ equiprobable input states. Second, there is an exponential overhead in the updating time of the cells. Once a superposition of all input states has been prepared as the single state vector $`|\psi `$, then by acting the linear evolution operator $`U_f`$ only once on it i.e $`U_f|\psi =\frac{1}{\sqrt{N}}_{x\text{Z}\text{Z}_2^{2r+1}}U_f|x=\frac{1}{\sqrt{N}}_{x\text{Z}\text{Z}_2^{2r+1}}|f(x)`$, we can update all the $`N`$ state vectors simultaneously. This same exponential acceleration in storing space and processing time of quantum information has been known in quantum computation, where it is utilized in a number of tasks such as fast quantum algorithms, information encoding, cryptography etc. Harnessing the space-time resources made available by the quantum parallelism of QCSA in various applications is an open challenge. The research program initiated here will require further studies of QSCA both as a new computational machine and as an implementable physical system. Some of these questions however will be addressed elsewhere. Discussion. We close by discussing some prospects of our work. Towards a realization of QSCA as a physical process that could be implemented experimentally and so would potentially provide a novel quantum computational machine, we may construct an optical circuit made of elementary passive optical elements i.e light beam splitters and phase shifters that are assembled so that they implement the operator $`U_f`$ of a QSCA. For that purpose we may utilize the theorem of embedding the $`SU(2)`$ Lie group into $`SU(N)`$, which recently has been used in the form of expressing every unitary matrix by a sequence of embedded $`2\times 2`$ elementary unitary matrices of only two types : one matrix that realizes a beam splitter with transmission and reflection coefficients determined by the matrix elements and one that similarly realizes a phase shifter of a classical propagating light wave. By virtue of this embedding we may factorize the $`U_f`$ matrix of a $`r=2`$, say QSCA into elementary matrices of beam splitters and phase shifters. Assuming a two-state encoding of the qubits of the QSCA cells (e.g taking the vertical and horizontal polarization states of a laser beam as $`|0>`$ and $`|1>`$), we may construct the quantum optical analog of the $`U_f`$ of the QSCA. Finally, we should mention a number of ramifications of our QSCA formalism that are currently also under investigation, namely: QSCA beyond the binary case ( this would involve higher dimensional representations of the discrete Heisenberg group ); differential versus matrix formulation of QSCA (i.e partial differential operators acting on a space of multivariable complex polynomials realizing the respective action of the $`U_f`$ matrix ); and also the case of noisy QSCA (i.e QSCA evolution that occurs in the presence of of quantum mechanical noise, exemplified by errors due to random bit-flipping and phase shifting in the qubits of the QSCA string ); this would require the description of QSCA in terms of trace-preserving operators instead of the unitary operators as in the noiseless case, and the respective quantization of the classical SCA to be carried out not by Hilbert-space state vectors but by employing the $`\rho `$-density operator formalism of Quantum Mechanics.
warning/0001/quant-ph0001081.html
ar5iv
text
# Probabilistic quantum cloning via Greenberger-Horne-Zeilinger states ## I Introduction Quantum computers can solve problems that classical computers can never solve . However, the practical implementation of such device need careful consideration of the minimum resource requirement and feasibility of quantum operation. The basic operation in quantum computer is unitary evolution, which can be performed using some single-qubit unitary operations and Controlled-NOT (C-NOT) gates . While single-qubit unitary operation can be executed easily , the implementation of C-NOT operation between two particles (for example two photons) encounters great difficulty in experiment . With linear optical devices (beam splitters, phase shifters, etc.), the C-NOT operations between the several quantum qubits (such as location and polarization) of a single photon is within the reach of current quantum optics technology , but nonlinear interactions are required for the construction of practical C-NOT gate of two particles . Those nonlinear interactions are normally very weak, which forecloses the physical implementation of quantum logic gate. To solve this problem, Gottesman and Chuang suggested that a generalization of quantum teleportation $``$using single-qubit operations , Bell-basis measurements and certain entangled quantum states such as Greenberger-Horne-Zeilinger (GHZ) states $``$is sufficient to construct a universal quantum computer and presented systematic constructions for an infinite class of reliable quantum gates (including Tele-C-NOT gate). Experimentally, quantum teleportation has been partially realized and three-photon GHZ entanglement has been observed . Thus their construction of quantum gates offers possibilities for relaxing experimental constraints on realizing quantum computers. Unfortunately, up to now there has been no way to experimentally distinguish all four of the Bell states, although some schemes do work for two of the four required cases $``$yielding at most a $`50\%`$ absolute efficiency . In Gottesman and Chuang’s scheme, two GHZ states and three Bell-basis measurements are needed to perform a C-NOT operation, which yields $`1/8`$ probability of success in experiment. To complete a unitary operator, many C-NOT gates may be needed, which makes the probability of success close to zero. Moreover, the creation efficiency of GHZ states is still not high in experiment now . Therefore a practical experiment protocol requires careful consideration of the minimum resource and the maximum probability of success. In this paper we investigate the problem of probabilistic quantum cloning using GHZ states, Bell basis measurements, single-qubit unitary operations and generalized measurements. The single-qubit generalized measurement can be performed by the unitary transformation on the composite system of that qubit and the auxiliary probe with reduction measurement of the probe . In an optical quantum circuit, the probe qubit can be represented as the location of a photon and such process can be implemented using only linear optical components (such as polarizing beam splitter and polarization rotation) . We mention above that the construction of practical C-NOT between two particles is not within current experimental technology, but it does not prohibit the C-NOT operation between different degrees of freedom of one photon. This kind of C-NOT is allowed in linear optical circuit and is of different type from C-NOT between different particles . So the single-qubit generalized measurement on the polarization can be performed with location as the probe. Consider a sender Alice holds an one-qubit quantum state $`|\varphi `$ and wishes to transmit identical copies to $`N`$ associates (Bob, Claire, etc.). Quantum no-cloning theorem implies that the copies cannot be perfect; but this result does not prohibit cloning strategies with a limited degree of success. Two most important cloning machines $``$universal and state-dependent $``$have been proposed by some authors. However, it is not available <sup>*</sup><sup>*</sup>*However, it is available to clone the states of one qubit of a single photon to two qubits of that photon using optical simulation . for Alice to generate the copies locally using an appropriate quantum network and then teleport each one to its recipient by means of teleportation due to the difficulty of executing C-NOT operation . To avoid such difficulty, recently, Murao et al. presented an optimal $`1`$ to $`N`$ universal quantum telecloning strategy via a $`\left(2N\right)`$-particle entangled state. Such entanglement is difficult to prepare in experiment when $`N`$ is large. Quantum probabilistic (state-dependent) cloning machine is designed to perfectly reproduce linear independent states secretly chosen from a finite set with no-zero probability . The corresponding telecloning process can be executed via the Tele-C-NOT gates according to the cloning strategies provided in ; but such procedure requires too many GHZ states and Bell basis measurements and can succeed with probability close to zero. The scheme we propose in this paper needs only $`\left(N1\right)`$ GHZ states and $`\left(N1\right)`$ Bell-basis measurements to implement $`MN`$ cloning. Although such process cannot reach the optimal probability as that in local situation, it may be used in current experiment to cloning the states of one particle to those of two different particles with higher probability and less GHZ resources. The rest of the paper is organized as follows. In section II we discuss some strategies of probabilistic cloning and present the concept of probability spectrum to describe different strategies. Comparing two most important ones, we show that $`M`$ entries $`1N`$ cloning give more copies at the price of higher probability of failure than one $`MN`$ cloning. In Section III, we present the probabilistic telecloning process via three-particle entangled state and also show how to construct the entangled state from GHZ state by local operations. A summary is given in Section IV. ## II Strategies of probabilistic cloning Generally, the most useful states are $`|\varphi _\pm \left(\theta \right)=\mathrm{cos}\theta |1\pm \mathrm{sin}\theta |0`$ in quantum information theory. Given $`M`$ initial copies, Alice need not to always execute the cloning operation by taking these copies as a whole. Suppose Alice divides the $`M`$ copies into $`m`$ different kinds of shares, each of which includes $`\vartheta _i`$ entries $`k_iN_i`$ cloning processes. For different kinds of shares, one of the two parameters $`k_i`$ and $`N_i`$ should be different. These parameters should satisfy $$\underset{i=1}{\overset{m}{}}k_i\vartheta _i=M\text{.}$$ (1) The probability of obtaining $`x`$ copies for Alice can be represented as $$P\left(x\right)=\underset{\underset{i=1}{\overset{m}{}}g_iN_i=x}{}\underset{i=1}{\overset{m}{}}C_{\vartheta _i}^{g_i}\gamma _{k_iN_i}^{g_i}\left(1\gamma _{k_iN_i}\right)^{\vartheta _ig_i},$$ (2) where $`C_{\vartheta _i}^{g_i}=\vartheta _i!/g_i!(\vartheta _ig_i)!`$, $`g_i`$ denotes successful cloning attempts in $`\vartheta _i`$ same processes, $`\gamma _{k_iN_i}`$ is the success probability of $`k_iN_i`$ cloning, which is $$\gamma _{k_iN_i}=\frac{1\mathrm{cos}^{k_i}2\theta }{1\mathrm{cos}^{N_i}2\theta }.$$ (3) $`P\left(x\right)`$ is the discrete function of $`x`$ and can be represented as a series of discrete lines in the $`P\left(x\right)x`$ plane, which we called as Probability Spectrum. Different probabilistic cloning strategies are corresponding to different Probability Spectrums. Two important parameters can be obtained from Probability Spectrum, that is, the expected value of the output copies number $`E`$ and the probability of failure $`F`$, which are defined as $$E\{k_i,N_i,\vartheta _i\}=\underset{x=0}{\overset{\underset{i=1}{\overset{m}{}}\vartheta _iN_i}{}}xP\left(x\right),$$ (4) $$F\{k_i,N_i,\vartheta _i,K\}=\underset{x=0}{\overset{K1}{}}P\left(x\right).$$ (5) It is regarded as failure if the copies number Alice attains less than the cloning goal $`K`$. When $`M`$ is large, above two parameters can well describe different cloning strategies. In the following, we discuss two most important cloning strategies (the cloning goal $`K=N`$): $`\left(1\right)`$ cloning the $`M`$ copies as a whole ($`MN`$). $`\left(2\right)`$ cloning each copy respectively ($`M\times \left(1N\right)`$). The second is included for it is the strategy we choose in probabilistic telecloning process. Comparing above two strategies with the two parameters $`E`$ and $`F`$, we find the second give more copies at the price of higher probability of failure. In fact, if Alice choose the second strategy, the cloning attempts may succeed for two or more initial copies, thus Alice may have chance to get more than $`N`$ copies. The expected values for the two different strategies can be represented as $$E_1=N\gamma _{MN},$$ (6) $`E_2`$ $`=`$ $`{\displaystyle \underset{k=0}{\overset{M}{}}}kNC_M^k\gamma _{1N}^k\left(1\gamma _{1N}\right)^{Mk}`$ () $`=`$ $`MN\gamma _{1N}{\displaystyle \underset{k=1}{\overset{M}{}}}C_{M1}^{k1}\gamma _{1N}^{k1}\left(1\gamma _{1N}\right)^{\left(M1\right)\left(k1\right)}`$ (7) $`=`$ $`MN\gamma _{1N},`$ (8) where $`2M<N`$. Denote $`t=\mathrm{cos}2\theta `$, we get $`\mathrm{\Delta }E=E_2E_1=N\stackrel{~}{\mathrm{\Delta }E}/\left(1t^N\right)`$, where $`\stackrel{~}{\mathrm{\Delta }E}=MMt1+t^M`$. Obviously $`0t1`$. When $`t=0`$, $`\stackrel{~}{\mathrm{\Delta }E}=M1>0`$. If $`t=1`$, $`\gamma _{MN}=\frac{M}{N}`$ and $`\mathrm{\Delta }E=\stackrel{~}{\mathrm{\Delta }E}=0`$. When $`t1`$, $`\frac{d\stackrel{~}{\mathrm{\Delta }E}}{dt}=M+Mt^{M1}<0`$, thus $`\stackrel{~}{\mathrm{\Delta }E}`$ is monotonously decreasing and always greater than or equal to zero, that is $$E_1E_2,$$ (9) with equality only for $`|\phi _+\left(\theta \right)=|\phi _{}\left(\theta \right)`$ ($`t=1`$). $`\mathrm{\Delta }E`$ is very large when $`M`$ is large. The expected values for different $`M`$, $`N`$ are plotted in Fig. 1. Fig. 1 The failure probabilities of above two strategies are $$F_1=1\gamma _{MN}=t^M\left(1t^{NM}\right)/\left(1t^N\right),$$ (9) $$F_2=(1\gamma _{1N})^M=\left(\left(tt^N\right)/\left(1t^N\right)\right)^M,$$ (10) respectively. Note the fact that for any $`a_i0`$, $`\left(_{i=1}^na_i\right)^{\frac{1}{M}}\frac{1}{M}\left(_{i=1}^na_i\right)`$ with equality only for $`a_1=a_2=\mathrm{}=a_n`$, we derive $`F_1`$ $`=`$ $`{\displaystyle \frac{t^M}{\left(1t^N\right)^M}}\left(1t^N\right)^{M1}\left(1t^{NM}\right)`$ () $``$ $`{\displaystyle \frac{t^M}{\left(1t^N\right)^M}}\left(1{\displaystyle \frac{t^{NM}+(M1)t^N}{M}}\right)^M`$ (11) $``$ $`{\displaystyle \frac{t^M}{\left(1t^N\right)^M}}\left(1t^{N1}\right)^M=F_2`$ (12) with equality only for $`t=0`$ or $`1`$ ($`\theta =\pi /4`$ or $`0`$). The failure probabilities for different $`M`$, $`N`$ are illustrated in Fig. 2. Fig. 2 Now that the two different strategies have both advantage and shortage, Alice should choose one according to her need. If she need more copies, she can adopt $`1N`$ strategy. If she wishes to obtain the copies with greater success probability, she should choose $`MN`$ cloning process. ## III Probabilistic telecloning process Suppose Alice holds $`M`$ copies of one-qubit quantum state $`|\varphi _X`$ that is secretly chosen from the set $`\left\{|\varphi _\pm \left(\theta \right)=\mathrm{cos}\theta |1\pm \mathrm{sin}\theta |0\right\}`$ and wishes to clone it to $`N`$ associates (Bob, Claire, etc.). In local situation, she can do so using the unitary-reduction operation $``$ a combination of unitary evolution together with measurements $``$ on the $`N+1`$ qubit ($`N`$ qubit of the cloning system and a probe to determine whether the cloning is successful) with maximum success probability $`\gamma _{MN}=\left(1\mathrm{cos}^M2\theta \right)/\left(1\mathrm{cos}^N2\theta \right)`$. This unitary-reduction operator can be decomposed into the interaction between two particles using a special unitary gate : $$D(\theta _1,\theta _2)|\varphi _\pm \left(\theta _3\right)|1=|\varphi _\pm \left(\theta _1\right)|\varphi _\pm \left(\theta _2\right).$$ (13) with $`\mathrm{cos}2\theta _3=\mathrm{cos}2\theta _1\mathrm{cos}2\theta _2`$ and $`0\theta _j\pi /4`$, which suffice to determine $`\theta _3`$ uniquely. This operation $`D^{}(\theta _1,\theta _2)`$ transforms the information describing the initial states $`|\varphi _\pm \left(\theta _1\right)|\varphi _\pm \left(\theta _2\right)`$ into one qubit $`|\varphi _\pm \left(\theta _3\right)`$. With such pairwize interaction, the initial states $`|\varphi _\pm \left(\theta \right)^M`$ can be transferred into states $`|\varphi _\pm \left(\theta _M\right)|0^{\left(M1\right)}`$ using corresponding operator $`D_M=D_1(\theta _{M1},\theta _1)D_2(\theta _{M2},\theta _1),\mathrm{}D_{M1}(\theta _1,\theta _1)`$, where $`D_j(\theta _{Mj},\theta _1)`$ is denoted as the operator $`D(\theta _{Mj},\theta _1)`$ acts on particles $`(1,j+1)`$ and $`\theta _j`$ is determined by $`\mathrm{cos}2\theta _j=\mathrm{cos}^j2\theta `$. This operator is unitary and $`D_M^{}`$ can perform the reverse transformation. Thus we only need to transfer the states $`|\varphi _\pm \left(\theta _M\right)`$ to the appropriate form $`|\varphi _\pm \left(\theta _N\right)`$ to obtain $`|\varphi _\pm \left(\theta \right)^N`$ using operation $`D_N^{}`$ (with similar definition as $`D_M^{}`$). This process can be accomplished by a unitary-reduction operation $$U|\varphi _\pm \left(\theta _M\right)_1|P_0=\sqrt{\gamma }|\varphi _\pm \left(\theta _N\right)_1|P_0+\sqrt{1\gamma }|1_1|P_1,$$ (13) where $`|P_0`$ and $`|P_1`$ are the orthogonal bases of the probe system. If a postselective measurement of probe $`P`$ results in $`|P_0`$, the transformation is successful, otherwise the cloning attempt has failed and the result is discarded. $`U`$ is unitary and the transformation probability $`\gamma =\gamma _{MN}`$. $`U`$ is a qubit $`1`$ controlling probe $`P`$ rotation $`R_y\left(2\omega \right)=\left(\begin{array}{cc}\mathrm{cos}\omega & \mathrm{sin}\omega \\ \mathrm{sin}\omega & \mathrm{cos}\omega \end{array}\right)`$ with $`\omega =\mathrm{arccos}\sqrt{\frac{\left(1\mathrm{cos}^M2\theta \right)\left(1+\mathrm{cos}^N2\theta \right)}{\left(1+\mathrm{cos}^M2\theta \right)\left(1\mathrm{cos}^N2\theta \right)}}`$. Operations $`D_M`$ and $`D_N^{}`$ involve the interactions of two particles which is difficult to implement in current experiment. In this paper, we adopt $`M\times \left(1N\right)`$ strategy to substitute $`D_M`$ and transfer $`M`$ copies of the states $`|\varphi _\pm \left(\theta \right)`$ to $`|\varphi _\pm \left(\theta _N\right)`$ respectively using similar unitary-reduction operation as that in Eq. (3.2). To substitute the operation $`D_N^{}`$, we use three-particle entanglement to implement the operator $`D_j(\theta _{Nj},\theta _1)`$, which acts as $$D_j(\theta _{Nj},\theta _1)|\varphi _\pm \left(\theta _{Nj+1}\right)|1=|\varphi _\pm \left(\theta _{Nj}\right)|\varphi _\pm \left(\theta _1\right).$$ (14) Assume Alice and the $`j`$-th associate $`C_j`$ share a three-particle entangled state $`|\psi ^j_{SAC_j}`$ as a starting resource. This state must be chosen so that, after Alice performs local Bell measurements and informs $`C_j`$ of the results, she and $`C_j`$ can obtain the state $`|\varphi _\pm \left(\theta _{Nj}\right)_A|\varphi _\pm \left(\theta _1\right)_{C_j}`$ by using only local operation. Denote $`|\phi _i^j=D_j(\theta _{Nj},\theta _1)|i|1`$, $`i\{0,1\}`$, a choice of $`|\psi ^j_{SAC_j}`$ with these properties may be the three-particle state $$|\psi ^j_{SAC_j}=\frac{1}{\sqrt{2}}\left(|0_S|\phi _1^j_{AC_j}|1_S|\phi _0^j_{AC_j}\right),$$ (15) where $`S`$ represents a single qubit held by Alice, which we should refer to as the “port” qubit. The tensor product of $`|\psi ^j_{SAC_j}`$ with the state $`|\varphi _\pm \left(\theta _{Nj+1}\right)_X=h_j|1\pm t_j|0`$ ($`h_j=\mathrm{cos}\theta _{Nj+1}`$, $`t_j=\mathrm{sin}\theta _{Nj+1}`$) held by Alice is four-qubit state. Rewriting it in a form that singles out the Bell basis of qubit $`X`$ and $`S`$, we get $`|\mathrm{\Omega }^{\pm j}_{XSAC_j}`$ () $`=`$ $`{\displaystyle \frac{1}{2}}|\mathrm{\Psi }^{}_{XS}\left(h_j|\phi _1^j_{AC_j}\pm t_j|\phi _0^j_{AC_j}\right)`$ (19) $`+{\displaystyle \frac{1}{2}}|\mathrm{\Psi }^+_{XS}\left(h_j|\phi _1^j_{AC_j}t_j|\phi _0^j_{AC_j}\right)`$ $`\pm {\displaystyle \frac{1}{2}}|\mathrm{\Phi }^{}_{XS}\left(t_j|\phi _1^j_{AC_j}\pm h_j|\phi _0^j_{AC_j}\right)`$ $`\pm {\displaystyle \frac{1}{2}}|\mathrm{\Phi }^+_{XS}\left(t_j|\phi _1^j_{AC_j}h_j|\phi _0^j_{AC_j}\right),`$ where $`|\mathrm{\Psi }^\pm _{XS}=\frac{1}{\sqrt{2}}\left(|01_{XS}\pm |10_{XS}\right)`$, $`|\mathrm{\Phi }^\pm _{XS}=\frac{1}{\sqrt{2}}\left(|00_{XS}\pm |11_{XS}\right)`$ are the Bell basis of the two-qubit system $`XS`$. The telecloning process can now be accomplished by the following procedure. $`\left(i\right)`$ Alice performs a Bell-basis measurement of qubits $`X`$ and $`S`$, obtaining one of the four results $`|\mathrm{\Psi }^\pm _{XS}`$, $`|\mathrm{\Phi }^\pm _{XS}`$. $`\left(ii\right)`$ Alice use different strategies according to different measurement results. If the result is $`|\mathrm{\Psi }^{}_{XS}`$, the subsystem $`AC_j`$ is projected precisely into the state $`h_j|\phi _1^j_{AC_j}\pm t_j|\phi _0^j_{AC_j}=|\varphi _\pm \left(\theta _{Nj}\right)_A|\varphi _\pm \left(\theta _1\right)_{C_j}`$. If $`|\mathrm{\Psi }^+_{XS}`$ is obtained, $`\sigma _z\sigma _z`$ must be performed on system $`AC_j`$ since $`|\phi _0^j_{AC_j}`$ and $`|\phi _1^j_{AC_j}`$ obey the following simple symmetry: $$\sigma _z\sigma _z|\phi _i^j_{AC_j}=\left(1\right)^{i+1}|\phi _i^j_{AC_j}.$$ (20) With above operations, the states of system $`AC_j`$ are transferred to $`|\varphi _\pm \left(\theta _{Nj}\right)_A|\varphi _\pm \left(\theta _1\right)_{C_j}`$, just as operation $`D_j(\theta _{Nj},\theta _1)`$ functions. $`\left(iii\right)`$ In the case one of the other two Bell states $`|\mathrm{\Phi }^\pm _{XP}`$ is obtained, the corresponding states are entangled states. For example, if measurement result is $`|\mathrm{\Phi }^{}_{XP}`$, the remained states can be written as $`|\alpha _\pm =\frac{\pm 1}{\mathrm{sin}2\theta _{Nj+1}}\left(\right|\varphi _\pm \left(\theta _{Nj}\right)|\varphi _\pm \left(\theta _1\right)`$ $`\mathrm{cos}2\theta _{Nj+1}|\varphi _{}\left(\theta _{Nj}\right)|\varphi _{}\left(\theta _1\right))`$, which lie in the subspace spanned by states $`\left\{\right|\varphi _+\left(\theta _{Nj}\right)|\varphi _+\left(\theta _1\right)`$, $`|\varphi _{}\left(\theta _{Nj}\right)|\varphi _{}\left(\theta _1\right)\}`$. The inner-products show that $`|\alpha _\pm `$ are orthogonal to $`|\varphi _{}\left(\theta _{Nj}\right)|\varphi _{}\left(\theta _1\right)`$. So they are entangled states unless $`|\varphi _+\left(\theta _{Nj}\right)|\varphi _+\left(\theta _1\right)`$ are orthogonal to $`|\varphi _{}\left(\theta _{Nj}\right)|\varphi _{}\left(\theta _1\right)`$, which means $`|\varphi _\pm \left(\theta \right)`$ are orthogonal. When $`|\varphi _\pm \left(\theta _1\right)`$ are not orthogonal, Alice and $`C_j`$ must disentangle the states to the needed states $`|\varphi _\pm \left(\theta _{Nj}\right)|\varphi _\pm \left(\theta _1\right)`$ simultaneously using only local operations and classical communication (LQCC). Unfortunately, this process cannot be deterministic although both transformation $`|\alpha _+|\varphi _+\left(\theta _{Nj}\right)|\varphi _+\left(\theta _1\right)`$ and $`|\alpha _{}|\varphi _{}\left(\theta _{Nj}\right)|\varphi _{}\left(\theta _1\right)`$ can be deterministically executed according to Nielsen theorem . In fact, suppose there exists a process $`H`$ to accomplish so using only LQCC, the evolution equation of the composite system of particles $`A,C_j`$ and the local auxiliary particles $`G^A`$, $`G^{C_j}`$ can be expressed as $`H|\alpha _\pm |G_0^A|G_0^{C_j}`$ () $`=`$ $`{\displaystyle \underset{i=1}{\overset{h}{}}}{\displaystyle \underset{k=1}{\overset{l}{}}}\sqrt{\eta _{ik}}|\varphi _\pm \left(\theta _{Nj}\right)|\varphi _\pm \left(\theta _1\right)|G_i^A|G_k^{C_j}.`$ (20) $`H`$ is a linear operation, thus we get $`H|\varphi _\pm \left(\theta _{Nj}\right)|\varphi _\pm \left(\theta _1\right)|G_0^A|G_0^{C_j}`$ () $`=`$ $`|\alpha _\pm {\displaystyle \underset{i=1}{\overset{h}{}}}{\displaystyle \underset{k=1}{\overset{l}{}}}\sqrt{\eta _{ik}}|G_i^A|G_k^{C_j}.`$ (21) Operation $`H`$ use only local operations and classical communications which cannot enhance the entanglement. Obviously no entanglement exists in the left side of Eq. (3.9), but the right side is an entangled state between particle $`A`$, $`C_j`$. Thus such process $`H`$ does not exist. However, consider current experiment technology, only two Bell basis $`|\mathrm{\Psi }^\pm `$ of the four can be identified by interferometric schemes, with the others $`|\mathrm{\Phi }^\pm `$ giving the same detection signal , so we only need to consider $`|\mathrm{\Psi }^\pm `$ in our protocol. After Alice obtain the state $`|\varphi _\pm \left(\theta _{Nj}\right)_A`$, she take it as the input states $`|\varphi _\pm \left(\theta _{Nj}\right)_X`$ and use another three-particle entangled state $`|\psi ^{j+1}`$ to obtain the states $`|\varphi _\pm \left(\theta _{N\left(j+1\right)}\right)_A|\varphi _\pm \left(\theta _1\right)_{C_{j+1}}`$ between Alice and $`C_{j+1}`$, etc,. In the last process, if Alice wishes to transmit the copies to the associates $`C_{N1}`$ and $`C_N`$, the system $`A`$ should be on the side $`C_N`$. With the series transformations, the associates $`C_1`$, $`C_2`$,…, $`C_N`$ obtain the states $`|\varphi _\pm \left(\theta _1\right)_{C_j}`$ respectively and they finish the telecloning process. In the following, we show how to prepare the three-particle entangled state $`|\psi ^j`$ represented in Eq. (3.5) by LQCC using GHZ state as resource. Consider Alice and $`C_j`$ initially share a GHZ state $`|\xi _{SAC_j}=\frac{1}{\sqrt{2}}\left(|000+|111\right)`$, to implement the telecloning process, they must transfer it to the suitable state using only LQCC. First a local unitary operation $`R_y^S\left(\pi /2\right)R_y^A\left(\pi /2\right)R_y^{C_j}\left(\pi /2\right)`$ is performed to transfer $`|\xi _{SAC_j}`$ to $`|\xi ^{^{}}_{SAC_j}=\frac{1}{4}\left(\left(|0|1\right)_S\left(|1+|0\right)_{AC_j}^2+\left(|0+|1\right)_S\left(|1|0\right)_{AC_j}^2\right)`$. To obtain required states, local generalized measurement (POVM) is needed, which is described by operators $`M_m`$ on corresponding system, satisfying the completeness relation $`_mM_m^{}M_m=I`$. After the measurement, the results (classical communication) are sent to other system, who performs a local quantum operation $`\epsilon _m`$ on its system according to the requirement of the transformation task. The operation $`\epsilon _m`$ is conditional on the result $`m`$ and may be non-unitary. However, it is difficult to perform the operation $`\epsilon _m`$ according to classical communication in experiment. In the following, we introduce a method to prepare the initial state by systems $`S`$, $`A`$ and $`C_j`$ performing local operations respectively without classical communication. In our protocol, there are two possible final states and both of them can be used for telecloning with same Bell states $`|\mathrm{\Psi }^\pm `$ measured. Define operations $`M_{jim}`$ $`\left(i=1,2,3,m=0,1\right)`$ on $`S`$, $`A`$, and $`C_j`$ system with matrix representations $`M_{j10}=\left(\begin{array}{cc}\mathrm{sin}\theta _{Nj+1}& 0\\ 0& \mathrm{cos}\theta _{Nj+1}\end{array}\right)`$, $`M_{j11}=\left(\begin{array}{cc}\mathrm{cos}\theta _{Nj+1}& 0\\ 0& \mathrm{sin}\theta _{Nj+1}\end{array}\right)`$, $`M_{j20}=\left(\begin{array}{cc}\mathrm{sin}\theta _{Nj}& 0\\ 0& \mathrm{cos}\theta _{Nj}\end{array}\right)`$, $`M_{j21}=\left(\begin{array}{cc}\mathrm{cos}\theta _{Nj}& 0\\ 0& \mathrm{sin}\theta _{Nj}\end{array}\right)`$, $`M_{j30}=\left(\begin{array}{cc}\mathrm{sin}\theta _1& 0\\ 0& \mathrm{cos}\theta _1\end{array}\right)`$, $`M_{j31}=\left(\begin{array}{cc}\mathrm{cos}\theta _1& 0\\ 0& \mathrm{sin}\theta _1\end{array}\right)`$ on the basis $`|0`$, $`|1`$ respectively. Note that $`M_{ji0}^{}M_{ji0}+M_{ji1}^{}M_{ji1}=I`$, therefore those define a generalized measurement on each system, which may be implemented using standard techniques involving only projective measurements and unitary transforms . If we consider a probe $`P`$ to assist the generalized measurement $`M_0=\left(\begin{array}{cc}\mathrm{sin}\theta & 0\\ 0& \mathrm{cos}\theta \end{array}\right)`$, $`M_1=\left(\begin{array}{cc}\mathrm{cos}\theta & 0\\ 0& \mathrm{sin}\theta \end{array}\right)`$, the unitary operator acting on the particle and the probe can be represented as $`\left(\begin{array}{cc}R_y\left(\pi +2\theta \right)& 0\\ 0& R_y\left(2\theta \right)\end{array}\right)`$ on the basis $`\{|0P_0,|0P_1,|1P_0,|1P_1\}`$, where $`R_y\left(\theta \right)=\left(\begin{array}{cc}\mathrm{cos}\frac{\theta }{2}& \mathrm{sin}\frac{\theta }{2}\\ \mathrm{sin}\frac{\theta }{2}& \mathrm{cos}\frac{\theta }{2}\end{array}\right)`$ is a rotation by $`\theta `$ around $`\widehat{y}`$. If the measurement result gives $`m=1`$ for a system, then a rotation $`\sigma _x`$ is performed on this system. Let $`|\xi _{\left(1\right)^{k+p+t}}_{SAC_j}`$ denote the state after the measurement and local $`\sigma _x`$, given that outcome $`k`$, $`p`$, $`t`$ occurred for $`A`$, $`C_j`$, $`S`$ system respectively, then $$|\xi _{\left(1\right)^{k+p+t}}_{SAC_j}=\{\begin{array}{c}\frac{1}{\sqrt{2}}\left(|0_S|\phi _1^j_{AC_j}|1_S|\phi _0^j_{AC_j}\right)\text{, when }\left(1\right)^{k+p+t}=1\\ \kappa \left(\frac{t_j}{h_j}|0_S|\phi _0^j_{AC_j}\frac{h_j}{t_j}|1_S|\phi _1^j_{AC_j}\right)\text{, when }\left(1\right)^{k+p+t}=1\end{array}$$ (22) where $`\kappa =\sqrt{\frac{1\mathrm{cos}^22\theta _{Nj+1}}{2(1+\mathrm{cos}^22\theta _{Nj+1})}}`$. The probability to obtain the first state $`|\xi _1_{SAC_j}`$ is $`p_1=\frac{\mathrm{sin}^22\theta _{Nj+1}}{2}`$ and the second $`|\xi _1_{SAC_j}`$ is $`p_1=\frac{1+\mathrm{cos}^22\theta _{Nj+1}}{2}`$. The first state in Eq. (3.10) is exactly the state in Eq. (3.5) and the second state can also be used for telecloning. In fact, the combined states of systems $`XSAC_j`$ can be rewritten in a form that singles out the Bell basis of qubit $`X`$ and $`S`$ as $`|\psi ^{\pm j}_{XSAC_j}^{^{}}`$ () $`=`$ $`{\displaystyle \frac{\kappa }{\sqrt{2}}}|\mathrm{\Psi }^{}_{XS}\left(h_j|\phi _1^j_{AC_j}\pm t_j|\phi _0^j_{AC_j}\right)`$ (25) $`\pm {\displaystyle \frac{\kappa }{\sqrt{2}}}|\mathrm{\Psi }^+_{XS}\left(h_j|\phi _1^j_{AC_j}t_j|\phi _0^j_{AC_j}\right)`$ $`+{\displaystyle \frac{\eta }{\sqrt{2}}}|\mathrm{\Phi }^{}_{XS}\left(h_j^3|\phi _1^j_{AC_j}\pm t_j^3|\phi _0^j_{AC_j}\right)`$ $`{\displaystyle \frac{\eta }{\sqrt{2}}}|\mathrm{\Phi }^+_{XS}\left(h_j^3|\phi _1^j_{AC_j}t_j^3|\phi _0^j_{AC_j}\right)`$ where $`\eta =\frac{2\kappa }{\mathrm{sin}2\theta _{Nj+1}}`$. Obviously the first two terms can be transferred to the target states using same unitary operations as those in Eq. (3.6) and states $`h_j^3|\varphi _0^j_{AC_j}\pm t_j^3|\varphi _1^j_{AC_j}=|\varphi _\pm \left(\theta _{Nj}\right)|\varphi _\pm \left(\theta _1\right)+\mathrm{cos}2\theta _{Nj+1}|\varphi _{}\left(\theta _{Nj}\right)|\varphi _{}\left(\theta _1\right)`$ need not be considered. The probabilistic quantum cloning process via GHZ states is illustrated in Fig. 3(A) and Fig. 3(B) for the case $`M=1`$, $`N=2`$. Fig. 3(A) and Fig. 3(B) The unitary-reduction operation $`U`$ in Eq. (3.2) and the generalized measurements $`M_{jim}`$ can be implemented using linear optical components, i.e., polarizing beam splitter (PBS) and polarization rotation (PR). In Ref. , Cerf et al. constructed the location controlling polarization (LCP) NOT gate using a PR. A general LCP unitary rotation can also be executed similarly. The polarization controlling location (PCL) NOT gate is performed by the use of a PBS. However, a PCL unitary rotation need two PBS and some PR since direct rotation of location qubit is impossible. Generally, a PCL unitary rotation can be represented as $`V=\left(\begin{array}{cc}R_y\left(\xi \right)& 0\\ 0& R_y\left(\chi \right)\end{array}\right)`$ on the orthogonal basis $`\{|0|P_0,|0|P_1,|1|P_0,|1|P_1\}`$, with $`|0,|1`$ denoted as the polarization qubit and $`|P_0,|P_1`$ as the location qubit. $`V`$ can be decomposed into $`V=V_1V_2V_3V_2V_1`$, where $`V_1`$ is a LCP-NOT gate, $`V_2`$ is a PCL-NOT gate and $`V_3`$ represents a LCP unitary operation that performs $`R_y\left(\xi \right)`$ on the polarization qubit if the location qubit is on $`|P_0`$, and $`R_y\left(\chi \right)`$ if the location qubit on $`|P_1`$. So operation $`V`$ can be implemented using linear optical components as that in Fig. 4. Fig. 4 Each generalized measurement $`M`$ gives two output paths $`0`$ and $`1`$ and eight possible results may be output for the three photons while they only represent two possible final states $`|\xi _1_{SAC_j}`$ and $`|\xi _1_{SAC_j}`$. By the use of fiber the two paths for each $`M`$ can be convert into one, which means tracing out over the location qubit, and the final state of the three photons turns into the mixed state $`\rho _{SAC_j}=p_1|\xi _1\xi _1\left|+p_1\right|\xi _1\xi _1|`$. However, after Bell basis measurement of the tensor product state $`|\varphi _\pm \left(\theta _{Nj+1}\right)_X\varphi _\pm \left(\theta _{Nj+1}\right)|\rho _{SAC_j}`$, the final states are still $`h_j|\phi _1^j_{AC_j}\pm t_j|\phi _0^j_{AC_j}`$ and $`h_j|\phi _1^j_{AC_j}t_j|\phi _0^j_{AC_j}`$ corresponding to $`|\mathrm{\Psi }^{}_{XS}`$ and $`|\mathrm{\Psi }^+_{XS}`$ because of Eq. (3.6) and (3.11). Let us compare the efficiency of above telecloning process and that using Tele-C-NOT gates . To complete a Tele-C-NOT operation, two GHZ states and three Bell basis measurement are need, which yields $`1/8`$ probability. Performing a $`D_j(\theta _{Nj},\theta _1)`$ operation needs three C-NOT gates , that is, Alice only has probability of $`\frac{1}{512}`$ to succeed. While our protocol use one GHZ states and yields the probability $`p`$ $`=`$ $`p_1\times {\displaystyle \frac{1}{2}}+p_1\times \kappa ^2`$ () $`=`$ $`{\displaystyle \frac{\mathrm{sin}^22\theta _{Nj+1}}{2}}={\displaystyle \frac{1\mathrm{cos}^{2\left(Nj+1\right)}2\theta }{2}}.`$ (26) When $`\theta `$ is not too small, the success probability is not too low. If we do not consider the preparation of three-particle entanglement states, the efficiency of Tele-$`D_j(\theta _{Nj},\theta _1)`$ is $`50\%`$, which is exactly the efficiency of Bell measurement. If we have enough GHZ states, we can prepare enough required three-particle entangled states. In the initial information compress process, we adopt the $`M\times \left(1N\right)`$ cloning strategy. Using this strategy, more than one $`|\varphi _\pm \left(\theta _N\right)`$ can be obtained. So if the Tele-$`D_j(\theta _{Nj},\theta _1)`$ operation fails to one $`|\varphi _\pm \left(\theta _N\right)`$, we have chance to use another and that increases the success probability. The overall cloning probability of our protocol (not include that in states preparation) can be represented as $$P=\underset{k=1}{\overset{M}{}}C_M^k\gamma _{1N}^k\left(1\gamma _{1N}\right)^{Mk}\left(1\left(1\left(\frac{1}{2}\right)^{N1}\right)^k\right).$$ (27) $`P`$decreases with the increase of $`N`$, therefore we often adopt $`12`$ cloning strategy in practice. Up to this point, our discussion has assumed that the initially shared three-partite entangled states are pure GHZ states. Suppose, however, that $`|\xi _{SAC_j}`$ is corrupted a little by decoherence before it is made available to the systems $`S`$, $`A`$ and $`C_j`$, so they receive a density matrix $`\sigma `$ instead. What can we say about the final states and the probabilities of success? We argue that the final states and the probabilities do not change too much if the windages of initial states are not too large. We discuss this problem using the trace distance, a metric on Hermitian operators defined by $`T(A,B)\text{Tr}(|AB|)`$, where $`|X|`$ denotes the positive square root of the Hermitian matrix $`X^2`$. The trace distance is a quantity with a well-defined operational meaning as the probability of making an error distinguishing two states . In this sense it may reflect the possible physical approximation between the states: the value of the trace distance smaller, the two states more similar. A direct example is that for pure states $`\psi `$ and $`\varphi `$ the trace distance and the fidelity are related by a simple formula, $`T(\psi ,\varphi )=2\sqrt{1F(\psi ,\varphi )}.`$ () Ruskai has shown that the trace distance contracts under physical processes. More precisely, if $`\varpi `$ and $`\sigma `$ are any two density operators, and if $`\varpi ^{}(\varpi )`$ and $`\sigma ^{}(\sigma )`$ denote states after some physical process represented by the (trace-preserving) quantum operation $``$ occurs, then $`T(\varpi ^{},\sigma ^{})T(\varpi ,\sigma ).`$ () So, after the telecloning process, the change of the final states is limited by the trace distance between initial states $`|\xi _{SAC_j}\xi |`$ and $`\sigma `$, and the continuity of probability also promises the less alteration of the successful probabilities represented by Eq. (3.12) and Eq. (3.13). Of course, the final states may not be the pure cloning states we required at this situation. It may be a mixed states resembling the cloning states with the accuracy dependent on the windage of the initial states. Such telecloning process can also be accomplished using a multiparticle entangled state, similar as that has been shown in . The quality of our method is that only three-particle entanglement is used. In this scheme, we use local generalized measurements and Bell basis measurement to avoid the interactions between particles, so it may be feasible in current experiment condition. ## IV Summary In summary, we have presented a probabilistic quantum cloning scheme using GHZ states, Bell basis measurements, single-qubit unitary operations and generalized measurements, all of which are within the reach of current technology. We considered different strategies and propose the concept of Probability Spectrum to describe them. For two most important, we show that $`M`$ entries $`1N`$ cloning process give more copies than one $`MN`$ process at the price of higher probability of failure. Compared to another possible scheme via Tele-C-NOT gate, our scheme may be feasible in experiment to clone the states of one particle to those of two different particles with higher probability and less GHZ resource. ACKNOWLEDGMENT This work was supported by the National Natural Science Foundation of China. Figure Captions: Fig. 1: The expected values of copy number for the two different strategies. Angle $`\theta `$ is corresponding to initial states set $`\left\{\mathrm{cos}\theta |1\pm \mathrm{sin}\theta |0\right\}`$. Here Solid line, Dashed line, Dotted line and Dash-Dotted line denote $`10\times \left(120\right)`$, $`1\times \left(1020\right)`$, $`2\times \left(13\right)`$ and $`1\times \left(23\right)`$ cloning strategies respectively. Fig. 2: The failure probabilities for the two different strategies. The four kinds of lines represent the same strategies as those in Fig. 1. Fig. 3(A): The logic network of $`12`$ probabilistic cloning via GHZ state. Alice and her associate $`C_1`$, $`C_2`$ initially share a GHZ state consisting of the qubit $`S`$ (the port), $`C_1`$ and $`C_2`$ (outputs, or ‘copy qubits’). Alice successfully transforms the initial states $`\mathrm{cos}\theta |1_X\pm \mathrm{sin}\theta |0_X`$ to $`\mathrm{cos}\theta _2|1_X\pm \mathrm{sin}\theta _2|0_X`$ if the probe (the location qubit of the photon $`X`$) results in $`|P_0`$, where the parameters $`\mathrm{cos}^22\theta _2=\mathrm{cos}2\theta `$, $`\omega =\mathrm{arccos}\sqrt{\frac{\left(1+\mathrm{cos}^22\theta \right)}{\left(1+\mathrm{cos}2\theta \right)^2}}`$. Using the unitary rotation $`R_y\left(\varsigma \right)`$ and generalized measurement $`M\left(\theta \right)`$, Alice and $`C_1`$, $`C_2`$ transform GHZ state to the required three-particle entangled state in the form Eq. (3.10). Then Alice performs a Bell measurement of the port $`S`$ along with ‘input’ qubit $`X`$ and has $`25\%`$ probability to obtain $`|\mathrm{\Psi }^{}`$ or $`|\mathrm{\Psi }^+`$ respectively; subsequently, the receivers $`C_1`$ and $`C_2`$ do no operation or $`\sigma _x`$ rotations on the output qubits, obtaining two perfect quantum clones. The implementation of generalized measurement $`M\left(\theta \right)`$ is illustrated in Fig. 3(B). Fig. 3(B): The implementation of generalized measurement $`M\left(\theta \right)`$ in Fig. 3(A). The location qubit of the photon is adopted as the probe $`P`$. Fig. 4: Optical simulation of PCL unitary rotation by the use of two polarizing beam splitters and some polarizing rotators, where PR1 performs operation $`R_y\left(\xi \right)`$ and PR2 executes operation $`R_y\left(\chi \right)`$.
warning/0001/math0001064.html
ar5iv
text
# Polynomial and Rational Solutions of Holonomic Systems ## 1 Introduction Polynomial and rational solutions for linear ordinary differential equations can be obtained by algorithmic methods. For instance, the maple package DEtools provides efficient functions polysols and ratsols to find polynomial and rational solutions for a given linear ordinary differential equation with rational function coefficients. A natural analogue of the notion of linear ordinary differential equation in the several variable case is the notion of holonomic system. A holonomic system is a system of linear partial differential equations whose characteristic variety is middle dimensional. Chyzak gave an algorithm to find the rational solutions of holonomic systems by using elimination in the ring of differential operators with rational function coefficients combined with Abramov’s algorithm for rational solutions of ordinary differential equations with parameters. To the authors, solving holonomic systems is analogous to solving systems of algebraic equations of zero-dimensional ideals. Under this analogy, the method of Chyzak corresponds to the elimination method for solving systems of algebraic equations. The aim of this paper is to give two new algorithms, which are elimination free, to find polynomial and rational solutions for a given holonomic system associated to a set of linear differential operators in the Weyl algebra $$D=𝐤x_1,\mathrm{},x_n,_1,\mathrm{},_n$$ where $`𝐤`$ is a subfield of $`𝐂`$. Polynomial and rational solutions can be obtained, if they exist, by using an exhaustive search. For instance, when $`f=0`$ is the singular locus of a holonomic system $`M=D/I`$, any rational solution has the form $`g/f^r`$. If we have upper bounds for the degree of the polynomial $`g`$ and for $`r`$, then we can construct all rational solutions by solving linear equations satisfied by the coefficients of $`g`$. Alternatively, if we know the dimension of rational solutions, then we can obtain all rational solutions by increasing the degree of $`g`$ and $`r`$. Hence, the problem reduces to finding effective bounds for these numbers. In sections 2 and 3, we give algorithms for upper bounds on the degree of $`g`$ and on $`r`$. The main techniques we use are Gröbner deformations in $`D`$ as introduced in the book and the $`b`$-function for $`D/I`$ and $`f`$. In section 4, we give an algorithm to evaluate the dimension of polynomial and rational solutions. Our approach is an analog in $`D`$ of a question studied by Singer , who gave an algorithm to compute $`\mathrm{Hom}_R(M,N)`$ for left $`R:=𝐤(x_1)_1`$-modules $`M`$ and $`N`$ and studied its relation to factorizations of ordinary differential operators. The theory of $`D`$-modules translates our problem on polynomial and rational solutions to constructions in the ring of differential operators $`D`$. For example, the $`𝐤`$-vector space $$\mathrm{Hom}_D(D/I,𝐤[𝐱])H^n(\mathrm{\Omega }_D^L𝐃(D/I))$$ is the space of the polynomial solutions of the left ideal $`I`$. Here, $`\mathrm{\Omega }`$ is the module of the top dimensional differential forms and $`𝐃`$ is the dualizing functor. See, e.g., the book of Björk on this translation. We evaluate the dimension of the right hand side by recent developments of computational algebra such as construction of free resolutions in the ring $`D`$ and restrictions of $`D`$-modules , , , . Our method also allows us to evaluate the dimension of solutions to a holonomic system inside any holonomic module. For instance, we can find the dimension of the delta function solutions to $`I`$. Throughout the paper, we refer to the book for fundamental facts on the algorithmic treatment of $`D`$. Also, the algorithms which appear in the paper have been implemented in either kan or Macaulay 2 . We deeply thank Dan Grayson, Anton Leykin, and Mike Stillman, who helped to implement $`D`$-modules in Macaulay 2, and Frédéric Chyzak and Michael Singer for discussions on rational solutions. ## 2 Polynomial solutions by Gröbner deformations How can we obtain all polynomial solutions for ordinary differential equations? One method is to compute the indicial polynomial at infinity, find an upper bound on the degrees of polynomial solutions, and determine the coefficients of polynomials. The analogous method works for holonomic systems by using Gröbner deformations. For $`\mathrm{}D`$ and the weight vector $`w𝐑^n`$, we denote by $`\mathrm{in}_{(w,w)}(\mathrm{})`$ the initial term of $`\mathrm{}`$ with respect to the weight $`(w,w)`$ (see, e.g., \[11, §1.1\]). The following proposition follows from the definition of $`\mathrm{in}_{(w,w)}(\mathrm{})`$. ###### Proposition 2.1 Suppose that $`f(x_1,\mathrm{},x_n)`$ is a polynomial solution of $`I=D\{\mathrm{}_1,\mathrm{},\mathrm{}_m\}`$. Take $`w𝐙^n`$. Then $`f(t^{w_1}x_1,\mathrm{},t^{w_n}x_n)`$ can be expanded as a polynomial in $`t`$ as $$f_w(𝐱)t^p+O(t^{p+1}).$$ Then we have $$\mathrm{in}_{(w,w)}(\mathrm{}_i)f_w=0.$$ The initial ideal $`\mathrm{in}_{(w,w)}(I)`$ is sometimes called the Gröbner deformation of $`I`$ with respect to $`(w,w)`$. ###### Theorem 2.2 There exist only finitely many Gröbner deformations. The Newton polytope of a polynomial solution $`f`$ is defined as the convex hull of the exponent vectors of $`f`$. For generic $`w`$, $`f_w`$ is a monomial $`cx^a`$ and the point $`a`$ is a vertex of the Newton polytope of $`f`$. Let $`R=𝐤(x_1,\mathrm{},x_n)_1,\mathrm{},_n`$ and $`\theta _i=x_i_i`$. Since $`a𝐙^n`$ belongs to the zero set of the indicial ideal $$\stackrel{~}{\mathrm{in}_{(w,w)}}(I)=R\mathrm{in}_{(w,w)}(I)𝐤[\theta _1,\mathrm{},\theta _n],$$ we can construct a polytope that contains the Newton polytopes of the polynomial solutions by taking the convex hull of all the non-negative integral roots of all possible indicial ideals. It is not necessary to find all Gröbner deformations to obtain polynomial solutions. Let $`b(s)`$ be the generator of $`\mathrm{in}_{(w,w)}(I)𝐤[s]`$, $`s=_{i=1}^nw_i\theta _i`$. The polynomial $`b(s)`$ is called the $`b`$-function of $`I`$ with respect to $`(w,w)`$. The next proposition follows from the definition of $`b(s)`$. ###### Proposition 2.3 Let $`w`$ be a strictly negative weight vector. In other words, we assume that $`w_i<0`$ for all $`i`$. Consider the b-function $`b(s)`$ of $`I`$ with respect to $`(w,w)`$ and let $`k_1`$ be the smallest integer root of $`b(s)=0`$. The polynomial solutions of $`I`$ have the form $$\underset{p_i0,pwk_1}{}c_px^p.$$ (1) ###### Algorithm 2.4 (Finding the polynomial solutions by a Gröbner deformation) Input: a holonomic left ideal $`I`$. Output: the polynomial solutions of $`I`$. 1. Take a strictly negative weight vector $`w`$, compute the Gröbner deformation $`\mathrm{in}_{(w,w)}(I)`$, and compute the smallest non-positive integer root $`k_1`$ of the $`b`$-function with respect to $`(w,w)`$. See, e.g., \[11, Alg. 5.15\] for these procedures. 2. If we do not have such a root, then there is no polynomial solution other than $`0`$. 3. If there is a minimal integer root, then determine the coefficients $`c_p`$ of (1) by solving linear equations for the coefficients. ###### Example 2.5 The following system of differential equations of two variables is called the Appell differential equation $`F_1(a,b,b^{},c)`$ : $`\theta _x(\theta _x+\theta _y+c1)x(\theta _x+\theta _y+a)(\theta _x+b),`$ $`\theta _y(\theta _x+\theta _y+c1)y(\theta _x+\theta _y+a)(\theta _y+b^{}),`$ $`(xy)_x_yb^{}_x+b_y`$ where $`a,b,b^{},c`$ are complex parameters. Let us demonstrate how Algorithm 2.4 works for the system of parameter values $`(a,b,b^{},c)=(2,3,2,5)`$. First, we choose a strictly negative weight vector $`w=(1,2)`$ and compute the $`b`$-function $`b(s)`$, $`s=\theta _x2\theta _y`$, which is the generator of the principal ideal $`\mathrm{in}_{(w,w)}(I)𝐐[\theta _x2\theta _y]`$. We can use the V-homogenization or the homogenized Weyl algebra to get the generator (see, e.g., \[11, §1.2\]). Second, we need to find the integer roots of the $`b`$-function $`b(s)=0`$. In our example, these are $$7,0,4.$$ From Proposition 2.1, the highest $`(w)`$-degree monomial $`cx^py^q`$ in a polynomial solution gives rise to an integer solution $`w_1p+w_2q=p2q`$ of the $`b`$-function. Hence, the polynomial solutions are of the form $$f=\underset{p,q0,p+2q7}{}c_{pq}x^py^q.$$ Finally, we determine the coefficients $`c_{pq}`$ by applying the differential operators to $`f`$ and putting the results to $`0`$. In our example, we have only one polynomial solution $`({\displaystyle \frac{1}{21}}y^2+{\displaystyle \frac{1}{7}}y{\displaystyle \frac{4}{35}})x^3+({\displaystyle \frac{3}{14}}y^2{\displaystyle \frac{24}{35}}y+{\displaystyle \frac{3}{5}})x^2`$ $`+`$ $`({\displaystyle \frac{12}{35}}y^2+{\displaystyle \frac{6}{5}}y{\displaystyle \frac{6}{5}})x+{\displaystyle \frac{1}{5}}y^2{\displaystyle \frac{4}{5}}y+1.`$ ## 3 Rational solutions by Gröbner deformations The singular locus of a $`D`$-ideal $`I`$ is defined to be the projection of the characteristic variety of $`I`$ minus the zero section from the cotangent bundle to the coordinate base space. In other words, it is the zero set $$\mathrm{Sing}(I)=V(\mathrm{in}_{(0,e)}(I):(\xi _1,\mathrm{},\xi _n)^{\mathrm{}}𝐤[x_1,\mathrm{},x_n]).$$ Any rational solution to $`I`$ has its poles contained inside the singular locus. Thus if $`f(𝐱)`$ defines the codimension 1 component of $`\mathrm{Sing}(I)`$, we may limit our search for rational solutions to $`𝐤[𝐱][\frac{1}{f}]`$. We will present a method to obtain an upper bound of the order of the poles along $`f=0`$ for each rational solution. For this purpose we use the notion of the $`b`$-function for $`f`$ and a section $`u`$ of a holonomic system, which was introduced by Kashiwara : Let $`𝒟`$ be the sheaf of algebraic differential operators on $`X=𝐂^n`$. For a holonomic $`𝒟`$-module $`=𝒟/𝒟I`$ and a polynomial $`f`$, consider the tensor product $$𝒩=𝒪[f^1,s]f^s_{𝒪_X}.$$ (2) This $`𝒩`$ has a structure of a left $`𝒟`$-module via the Leibnitz rule. Let $`u`$ be a section of $``$. Then the $`b`$-function for $`f`$ and $`u`$ (or for $`f^su`$) at $`p𝐂^n`$ is the minimum degree monic polynomial $`0b(s)𝐂[s]`$ such that $$b(s)f^su𝒟[s](f^{s+1}u)$$ (3) holds in $`𝒩`$ at $`p`$ (i.e., as a germ of $`𝒩`$ at $`p`$). This $`b`$-function depends on the point $`p`$. As a function of $`p`$, there is a stratification of $`𝐂^n`$ for which the $`b`$-function does not change on each strata (see e.g. for an algorithmic proof of this fact). In the definitions (2) and (3) for $`b`$-function, if we replace $`𝒪`$ by the polynomial ring $`𝐤[𝐱]`$, $`𝒟`$ by the Weyl algebra $`D`$, and $``$ by a holonomic $`D`$-module $`M=D/I`$, then we obtain the global $`b`$-function for $`f`$ and $`u`$. It is the least common multiple of $`b`$-functions at every point. ###### Theorem 3.1 Let $`u`$ be the residue class of $`1`$ in $`𝒟/𝒟I`$, and let $`b(s)`$ be the $`b`$-function for $`f`$ and $`u`$ at a point $`p𝐂^n`$ where $`f(p)=0`$. Assume that $`I`$ admits an analytic solution of the form $`gf^r`$ around $`p`$, where $`r𝐂`$, $`g`$ is a holomorphic function on a neighborhood of $`p`$, and $`g(p)0`$. Then $`s+r+1`$ divides $`b(s)`$. Proof: Let $`𝒟^{\mathrm{an}}`$ and $`𝒪^{\mathrm{an}}`$ be respectively the sheaf of analytic differential operators and the sheaf of holomorphic functions on $`𝐂^n`$. We may define the analytic $`b`$-function by replacing $`𝒪`$ by $`𝒪^{\mathrm{an}}`$, $`𝒟`$ by $`𝒟^{\mathrm{an}}`$, and $``$ by a $`𝒟^{\mathrm{an}}`$-module $`^{\mathrm{an}}`$ in the definitions (2) and (3). Since the $`b`$-function is an analytic invariant and the analytic and the algebraic $`b`$-functions coincide (see e.g. \[8, §8\]), we may work in the analytic category. We do this to consider solutions $`gf^r`$ where $`g`$ is holomorphic at $`p`$. If we only wish to consider solutions $`gf^r`$ where $`g`$ is a polynomial, then we may work in the algebraic category. In general, given a map of left $`𝒟^{\mathrm{an}}`$-modules $`\varphi :_1^{\mathrm{an}}_2^{\mathrm{an}}`$ and a section $`u`$ of $`_1^{\mathrm{an}}`$, the $`b`$-function for $`f^su`$ at a point $`p`$ is divisible by the $`b`$-function for $`f^s\varphi (u)`$ at $`p`$. We apply this basic fact to the following map $`\phi `$. Let $`J^{\mathrm{an}}`$ be the annihilating ideal of $`gf^r`$ in $`𝒟^{\mathrm{an}}`$. Since $`J^{\mathrm{an}}I^{\mathrm{an}}:=𝒟^{\mathrm{an}}I`$ and $`g(p)0`$, we have a left $`𝒟^{\mathrm{an}}`$-homomorphism $$\phi :𝒟^{\mathrm{an}}/I^{\mathrm{an}}𝒟^{\mathrm{an}}gf^r=𝒟^{\mathrm{an}}f^r𝒪^{\mathrm{an}}[f^1]f^r$$ which sends $`u`$ to $`gf^r`$. This map extends to a left $`𝒟^{\mathrm{an}}[s]`$-homomorphism $$\begin{array}{cc}1\phi :& 𝒪^{\mathrm{an}}[f^1,s]f^s_{𝒪^{\mathrm{an}}}𝒟^{\mathrm{an}}/I^{\mathrm{an}}\hfill \\ & \\ & 𝒪^{\mathrm{an}}[f^1,s]f^s_{𝒪^{\mathrm{an}}}𝒪^{\mathrm{an}}[f^1]f^r=𝒪^{\mathrm{an}}[f^1,s]f^{s+r}\hfill \end{array}$$ which sends $`f^su`$ to $`gf^{s+r}`$. By the definition of $`b(s)`$, there exists a germ $`P(s)`$ of $`𝒟[s]`$ at $`p`$ such that $$(P(s)fb(s))(f^su)=0.$$ Since $`1\phi `$ is a left $`𝒟^{\mathrm{an}}`$-homomorphism, applying it to the above equation gives the equation $`(P(s)fb(s))(gf^{s+r})=0`$, or in other words, $$g^1P(s)gf^{s+r+1}=b(s)f^{s+r}.$$ Thus, we see that the Bernstein-Sato polynomial $`b_f(s)`$ of $`f`$ at $`p`$ divides $`b(sr)`$. Note that $`s+1`$ divides $`b_f(s)`$ since $`f(p)=0`$ (cf. ). In conclusion, we have proved that $`s+1`$ divides $`b(sr)`$. This completes the proof. \[\] By virtue of the above theorem, we can obtain upper bounds by computing the $`b`$-function for $`f^su`$ at a smooth point of each irreducible component of the singular locus of $`I`$. From now on, let us also take $`f𝐤[𝐱]`$ to be a square-free polynomial defining the codimension one component of the singular locus, and let $`f=f_1\mathrm{}f_m`$ be its irreducible decomposition in $`𝐤[𝐱]`$. ###### Theorem 3.2 Let $`b_i(s)`$ be the $`b`$-function for $`f_i^su`$ at a generic point of $`f_i=0`$. Denote by $`r_i`$ the maximum integer root of $`b_i(s)=0`$. Then any rational solution (if any) to $`I`$ can be written in the form $`gf_1^{r_11}\mathrm{}f_m^{r_m1}`$ with a polynomial $`g𝐂[𝐱]`$. If some $`b_i(s)`$ has no integral root, then there exist no rational solutions to $`I`$ other than zero. Proof: An arbitrary rational solution to $`M`$ is written in the form $`gf_1^{\nu _1}\mathrm{}f_m^{\nu _m}`$ with integers $`\nu _1,\mathrm{},\nu _m`$ and $`g𝐂[𝐱]`$. Since the space of the rational solutions with coefficients in $`𝐂`$ is spanned by those with coefficients in $`𝐤`$, we may assume $`g𝐤[x]`$, and $`f`$ and $`g`$ are relatively prime in $`𝐤[𝐱]`$. Let $`p`$ be a generic point of $`f_i=0`$. We may assume that $`f_i`$ is smooth at $`p`$, $`g(p)0`$, and $`f_j(p)0`$ for $`ji`$. It follows from Theorem 3.1 that $`b_i(\nu _i1)=0`$. This implies $`\nu _ir_i+1`$. \[\] Since $`b`$-functions divide the global $`b`$-function, an upper bound can also be obtained from the global $`b`$-function. ###### Corollary 3.3 Let $`b_i(s)`$ be the global $`b`$-function for $`f_i^su`$, and denote by $`r_i`$ the maximum integer root of $`b_i(s)=0`$. Then any rational solution (if any) to $`I`$ can be written in the form $`gf_1^{r_11}\mathrm{}f_m^{r_m1}`$ with a polynomial $`g𝐂[𝐱]`$. We mention the corollary since the algorithm to compute global $`b`$-functions is simpler than the algorithm to compute $`b`$-functions. However, the $`b`$-function offers finer information. For instance, the well-known example $`f=x^2+y^2+z^2+w^2`$ has Bernstein-Sato polynomial $`(s+1)(s+2)`$ coming from the functional equation, $`\frac{1}{4}(_x^2+_y^2+_z^2+_w^2)f^{s+1}=(s+1)(s+2)f^s`$. Now consider the module $`M=Df^1`$ and let $`u`$ be the section of $`f^1`$. The global $`b`$-function for $`f^su`$ is $`s(s+1)`$, and hence Corollary 3.3 implies that rational solutions of $`M`$ all have the form $`gf^1`$ or $`gf^0`$, where $`g`$ is a polynomial not divisible by $`f`$. On the other hand, the Bernstein-Sato polynomial of $`f`$ at any nonsingular point $`p`$ of $`f=0`$ (i.e. except for the origin) is $`s+1`$. It follows that the $`b`$-function for $`f^su`$ equals $`s`$ at the generic point of $`f=0`$ and hence Theorem 3.2 implies that all rational solutions actually have the form $`gf^1`$. An algorithm to compute the $`b`$-function and the global $`b`$-function for $`f^su`$ was first given in based upon tensor product computation, which is slow and memory intensive. Shortly thereafter, Walther introduced in a more efficient method to compute the global $`b`$-function for $`f^su`$. Both methods give the global $`b`$-function exactly, under the condition that $`I`$ is $`f`$-saturated. Otherwise, we get a multiple of the global $`b`$-function. Similarly, the method of gives the $`b`$-function exactly if $`I`$ is $`f`$-saturated and additionally a certain primary decomposition in $`𝐂[𝐱]`$ is known. If primary decomposition is only available in $`𝐤[𝐱]`$, we again get a multiple of the $`b`$-function. Let us now describe an algorithm to compute the $`b`$-function for $`f^su`$ at a generic point of $`f=0`$ by combining the method of and the primary decomposition as was used in . ###### Algorithm 3.4 (Computing an upper bound of the $`b`$-function at a generic point) Input: a finite set $`G_0`$ of generators of a holonomic $`D`$-ideal $`I`$ and an irreducible polynomial $`f𝐤[𝐱]`$. Output: $`b^{}(s)𝐤[s]`$, which is a multiple of the $`b`$-function $`b(s)`$ for $`f^su`$ at a generic point of $`f=0`$, where $`u`$ is the residue class of $`1`$ in $`D/I`$. 1. Introducing a new variable $`t`$, put $`\vartheta _i=_i+(f/x_i)_t`$. Let $`\stackrel{~}{I}`$ be the left ideal of $`D_{n+1}`$, the Weyl algebra on the variables $`x_1,\mathrm{},x_n,t`$, that is generated by $$\{P(𝐱,\vartheta _1,\mathrm{},\vartheta _n)P(𝐱,_1,\mathrm{},_n)G_0\}\{tf(𝐱)\}.$$ 2. Let $`G_1`$ be a finite set of generators of the left ideal $`\mathrm{in}_{(1,0,\mathrm{},0;1,0,\mathrm{},0)}(\stackrel{~}{I})`$ of $`D_{n+1}`$. Here, $`1`$ is the weight for $`t`$ and $`1`$ is the weight for $`_t`$. 3. Rewrite each element $`P`$ of $`G_1`$ in the form $$P=_t^\mu P^{}(t_t,𝐱,_1,\mathrm{},_n)\text{or}P=t^\mu P^{}(t_t,𝐱,_1,\mathrm{},_n)$$ with a non-negative integer $`\mu `$, and define $`\psi (P)`$ by, $$\psi (P):=t^\mu _t^\mu P^{}=t_t\mathrm{}(t_t\mu +1)P^{}(t_t,𝐱,_1,\mathrm{},_n)$$ $$\text{or}\psi (P):=_t^\mu t^\mu P^{}=(t_t+1)\mathrm{}(t_t+\mu )P^{}(t_t,𝐱,_1,\mathrm{},_n).$$ Put $$G_2:=\{\psi (P)(s1,𝐱,_1,\mathrm{},_n)PG_1\}.$$ 4. Compute the elimination ideal $`J:=𝐤[s,𝐱]D[s]G_2`$. (The global $`b`$-function can be obtained at this stage by computing the monic generator of the ideal $`J𝐤[s]`$.) 5. Compute a primary decomposition of $`J`$ in $`𝐤[s,𝐱]`$ as $$J=Q_1\mathrm{}Q_\nu .$$ 6. For each $`i=1,\mathrm{},\nu `$, compute $`Q_{ix}:=Q_i𝐤[𝐱]`$, which is a primary ideal of $`𝐤[𝐱]`$. 7. Let $`b^{}(s)`$ be the monic generator of the ideal $$\{Q_i𝐤[s]\sqrt{Q_{ix}}𝐤[𝐱]f\}$$ of $`𝐤[s]`$. (Note that $`\sqrt{Q_{ix}}𝐤[𝐱]f`$ implies that $`\sqrt{Q_{ix}}`$ equals $`𝐤[𝐱]f`$ or $`\{0\}`$.) ###### Theorem 3.5 In the above algorithm, the polynomial $`b^{}(s)`$ is precisely the $`b`$-function for $`f^su`$ at a generic point of $`f=0`$ if $`I`$ is $`f`$-saturated (i.e., $`I:f^{\mathrm{}}=I`$) and each $`𝐂[s,𝐱]Q_i`$ remains primary in $`𝐂[s,𝐱]`$. Otherwise, the polynomial $`b^{}(s)`$ is a multiple of the $`b`$-function for $`f^su`$ at a generic point of $`f=0`$. Proof: Using essentially the same method as the proof of Lemma 4.1 in , we can prove that $`\stackrel{~}{I}`$ is precisely the annihilator ideal for $`\delta (tf(𝐱))u`$ in $$\stackrel{~}{M}:=(D_{n+1}\delta (tf(𝐱)))_{𝐂[𝐱]}D/I,$$ where $`\delta (tf(𝐱))`$ denotes the residue class of $`(tf(𝐱))^1`$ in $`𝐤[𝐱,(tf(𝐱))^1]/𝐤[𝐱]`$. Let $`b_t(s)`$ be the indicial polynomial for $`\delta (tf(𝐱))u`$ along $`t=0`$ at a point $`(0,p)`$ with $`f(p)=0`$. Then by Theorem 6.14 of , the $`b`$-function $`b(s)`$ for $`f^su`$ at $`p`$ divides out, and if $`I`$ is $`f`$-saturated, coincides with $`b_t(s1)`$. It follows from the definition that $`b_t(s1)`$ is a generator of the ideal $`𝒪_p[s]J𝐂[s]`$ of $`𝐂[s]`$, where $`𝒪_p`$ denotes the stalk of $`𝒪`$ at $`p`$. If $`𝐂[s,𝐱]Q_i`$ are primary in $`𝐂[s,𝐱]`$, $`b^{}(s)`$ generates the above ideal in view of Theorem 4.7 of (cf. also Lemma 4.4 of ). In general, although $`Q_i`$ is primary in $`𝐤[s,𝐱]`$, the extension $`𝐂[s,𝐱]Q_i`$ is no longer primary in $`𝐂[s,𝐱]`$ and admits a primary decomposition $$𝐂[s,𝐱]Q_i=Q_{i1}\mathrm{}Q_{i\mu _i}.$$ In this case, $`b^{}(s)`$ is the least common multiple of the generators of the ideals $`𝒪_p[s]Q_{ij}𝐂[s]`$ for $`j=1,\mathrm{},\mu _i`$, while $`b_t(s1)`$ is the generator of $`𝒪_p[s]Q_{ij}𝐂[s]`$ for some $`j`$ (such that $`p`$ belongs to the zero set of $`Q_{ij}𝐂[𝐱]`$ which is also the zero set of a factor of $`f`$). This completes the proof. \[\] ###### Remark 3.6 In the notation in the above proof, the linear factors of $`b^{}(s)`$ and those of $`b_t(s1)`$ coincide. In particular the set of integer roots of $`b^{}(s)=0`$ is the same as that of $`b_t(s1)=0`$. In fact, this follows from the fact that the linear factors of $`b^{}(s)`$ in $`𝐤[s]`$ are invariant under the action of the Galois group of $`\overline{𝐤}`$ over $`𝐤`$. ###### Remark 3.7 $`I`$ is $`f`$-saturated if and only if the $`1`$-th cohomology group of the restriction of $`\stackrel{~}{M}`$ in the proof of Proposition 3.5 to $`t=0`$ vanishes (see Theorem 6.4 and Proposition 6.13 of ), which is computable by Algorithm 5.10 of or by Algorithm 5.4 of . ###### Remark 3.8 The $`f`$-saturation of $`I`$, which is the ideal $`D[\frac{1}{f}]ID`$, may be computed by using the localization algorithm of (if $`I`$ is specializable along $`f`$) or by using the less efficient algorithm of (if $`I`$ is general). By replacing $`I`$ with its $`f`$-saturation, we may then compute the local $`b`$-function exactly. However, since saturation is often an expensive algorithm, we avoid making this replacement in practice. Once we have determined the integers $`r_1,\mathrm{},r_m`$ of Theorem 3.2, we can use Gröbner deformations to obtain the rational solutions. Put $`k_i=r_i+1`$. Then by virtue of Theorem 3.2, we have only to determine rational solutions of the form $`gf_1^{k_1}\mathrm{}f_m^{k_m}`$ for some polynomial $`g`$. This amounts to computing polynomial solutions of some twisted ideal $`I_{(k_1,\mathrm{},k_m)}`$ of $`I`$. Namely, consider how $`f_i`$ acts on an element $`gf_1^{k_1}\mathrm{}f_m^{k_m}`$: $$f_i(gf_1^{k_1}\mathrm{}f_m^{k_m})=\left(f\frac{g}{x_i}\underset{j=1}{\overset{m}{}}k_j\frac{f}{f_j}\frac{f_j}{x_i}g\right)f_1^{k_1}\mathrm{}f_m^{k_m}.$$ In other words, $`f_i`$ acts on the numerator $`g`$ as the differential operator $$L_i=f_i\underset{j=1}{\overset{m}{}}k_j\frac{f}{f_j}\frac{f_j}{x_i}.$$ (4) Thus, if we multiply a set of generators $`\{g_1,\mathrm{},g_m\}`$ of $`I`$ by sufficiently high powers $`f^{m_i}`$ such that $`f^{m_i}g_i𝐤x_1,\mathrm{},x_n,f_1,\mathrm{},f_n,`$ and if $`I^{}`$ is the ideal obtained from $`\{f^{m_1}g_1,\mathrm{},f^{m_r}g_r\}`$ by substituting $`L_i`$ for $`f_i`$, then the rational solutions of $`I`$ have the polynomial solutions of $`I^{}`$ as numerators. Thus, it only remains to compute the polynomial solutions of $`I^{}`$. However, since $`I^{}`$ might not be holonomic, we cannot apply Algorithm 2.4 just yet. Hence let us define $`I_{(k_1,\mathrm{},k_m)}:=𝐤(𝐱)I^{}D`$, which is the Weyl closure of $`I^{}`$ and whose polynomial solutions are the same as $`I`$. The advantage of $`I_{(k_1,\mathrm{},k_m)}`$ is that it is indeed holonomic, which follows from a theorem of Kashiwara. Namely, note that since $`I`$ is of finite rank, $`I^{}`$ remains of finite rank because an element in $`I`$ of the form $`(g_i(𝐱)_i^{N_i}+\text{ lower order elements})`$ will be sent to an element in $`I^{}`$ of the form $`(f^{M_i}g_i(𝐱)_i^{N_i}+\text{ lower order elements})`$. Now let $`h(𝐱)`$ be any polynomial vanishing on the singular locus of $`I^{}`$. Then the non-holonomic locus of $`I^{}`$ is contained inside the zero set of $`h(𝐱)`$ regarded as a function on the cotangent bundle, and a theorem due to Kashiwara states that the ideal $`D[h^1]I^{}D`$ is holonomic. Furthermore, an argument in shows that the Weyl closure of $`I^{}`$ also equals $`D[h^1]I^{}D`$. Summing up, we arrive at the following algorithm. ###### Algorithm 3.9 (Computing the rational solutions of a holonomic ideal) Input: generators of a holonomic $`D`$-ideal $`I`$. Output: A basis of the rational solutions $`h𝐤(𝐱)`$ of $`Ih=0`$. 1. Compute a polynomial $`f`$ defining the codimension 1 component of $`\mathrm{Sing}(I)`$. 2. Compute the irreducible decomposition $`f=f_1\mathrm{}f_m`$ in $`𝐤[𝐱]`$. 3. For each $`i=1,\mathrm{},m`$, compute the output $`b^{}(s)`$ of Algorithm 3.4 with $`I`$ and $`f_i`$ as input. Let $`r_i`$ be the maximum integer root of $`b^{}(s)=0`$ and put $`k_i=r_i+1`$. If $`b^{}(s)`$ has no integral root for some $`i`$, then there exists no rational solution other than zero. 4. Compute the twisted ideal $`I_{(k_1,\mathrm{},k_m)}`$ as follows. First, form the ideal $`I^{}`$ described in the paragraphs preceding the algorithm. Second, compute any polynomial $`h(𝐱)`$ vanishing on the singular locus of $`I^{}`$. Third, compute the localization $`(D/I^{})[h^1]`$ using the algorithm in . Then the ideal $`I_{(k_1,\mathrm{},k_m)}`$ is the kernel of $`DD/I^{}(D/I^{})[h^1]`$. 5. Compute a basis $`\{g_1,\mathrm{},g_k\}`$ of the polynomial solutions of $`I_{(k_1,\mathrm{},k_m)}`$ using Algorithm 2.4. 6. Output: $`\{g_1f_1^{k_1}\mathrm{}f_m^{k_m},\mathrm{},g_kf_1^{k_1}\mathrm{}f_m^{k_m}\}`$, a basis of the rational solutions of $`I`$. ###### Example 3.10 Let $`I`$ be the left ideal generated by $$L_1=\theta _x(\theta _x+\theta _y)x(\theta _x+\theta _y+3)(\theta _x1)$$ and $$L_2=\theta _y(\theta _x+\theta _y)y(\theta _x+\theta _y+3)(\theta _y+1).$$ The Appell function $`F_1(3,1,1,1;x,y)`$ is a solution of this system. The singular locus of $`I`$ is $`xy(xy)(1x)(1y)=0`$. We can compute the local indicial polynomial of $`u`$, the modulo class of $`1`$ in $`D_2/I`$, along $`x=0`$ directly by the algorithm of \[8, Section 4\]: It is $`s(s1)`$ on $`\{(0,y)y0\}`$, and $`s(s1)^2`$ at $`(0,0)`$. In the same way, the indicial polynomial of $`u`$ along $`y=0`$ is $`s(s+1)`$ on $`\{(x,0)x0\}`$, and $`s(s+1)(s1)`$ at $`(0,0)`$. Now let us compute the $`b`$-function for $`(1y)^su`$. The local indicial polynomial of $`\delta (t+y1)u`$ along $`t=0`$ is $`s(s+3)`$ at any point of $`t=0`$. Hence the $`b`$-function for $`(1y)^su`$ divides $`(s+1)(s2)`$. In the same way, the local indicial polynomial of $`\delta (t+x1)u`$ along $`t=0`$ is $`s(s+1)`$ at any point of $`t=0`$. Finally, the indicial polynomial of $`\delta (tx+y)u`$ is $`s(s1)`$ on $`\{(x,x)x0\}`$, and $`s(s1)(s2)`$ at $`(0,0)`$. Therefore, we conclude that any rational solution to $`I`$, if it exists, can be written in the form $`g(x,y)y^1(1x)^1(1y)^3`$ with a polynomial $`g`$. Now we may compute the twisted ideal $`I_{(0,1,0,1,3)}`$, where $`f_1=x,f_2=y,f_3=xy,f_4=x1,f_5=y1`$, and $`f`$ is the product. Multiplying by $`f^2`$, we get the expressions, $$\begin{array}{ccc}f^2L_1& =& (x^2x^3)(f_x)^2+x((13x)f(1x)y\frac{f}{y}(1x)x\frac{f}{x})(f_x)+\hfill \\ & & x(1x)y(f_y)(f_x)+xyf(f_y)+3xf^2\hfill \\ f^2L_2& =& (y^2y^3)(f_y)^2+y((15y)f(1y)x\frac{f}{x}(1y)y\frac{f}{y})(f_y)+\hfill \\ & & y(1y)x(f_x)(f_y)yxf(f_x)3yf^2,\hfill \end{array}$$ and we set $`T_1`$ and $`T_2`$ to be the operators obtained from the substitution (4). We remark that the ideal generated by $`T_1`$ and $`T_2`$ is neither holonomic nor specializable with respect to the weight vector $`(1,1,1,1)`$, so it is difficult to apply Gröbner deformations to it just yet. The twisted ideal $`I_{(0,1,0,1,3)}`$ is the Weyl closure of the ideal generated by $`T_1`$ and $`T_2`$. At the moment, this Weyl closure is computationally too intensive to compute. However, we are able to compute a partial closure by noting that $`T_1`$ is divisible by $`g=f_1^3f_2f_3^2f_4f_5`$ and $`T_2`$ is divisible by $`h=f_1^2f_2^2f_3^2f_4f_5`$. Now the ideal $`J`$ generated by $$\begin{array}{ccc}\frac{1}{g}T_1& =& (x^3y+x^3+2x^2y2x^2xy+x)_x^2+\hfill \\ & & (x^2y^2+x^2y+2xy^22xyy^2+y)_x_y+\hfill \\ & & (3x^2y6xy+3y)_x+(2xy^22xy2y^2+2y)_y+\hfill \\ & & (4xy2x+4y+2)\hfill \\ \frac{1}{h}T_2& =& (xy^4+2xy^3+y^4xy^22y^3+y^2)_y^2+\hfill \\ & & (x^2y^3+2x^2y^2+xy^3x^2y2xy^2+xy)_x_y+\hfill \\ & & (3x^2y^24x^2y3xy^2+x^2+4xyx)_x+\hfill \\ & & (4xy^36xy^23y^3+2xy+4y^2y)_y+\hfill \\ & & (6xy^2+8xy+3y^22x4y+1),\hfill \end{array}$$ is indeed holonomic, hence we may apply Algorithm 2.4. We find that the $`b`$-function with respect to the weight $`w=(1,1)`$ is $`(s+5)(s+2)^2(s+1)s(s2)(s3)^2`$, which implies that a polynomial solution to $`J`$ must have degree less than or equal to $`5`$. We find that $`J`$ has 2 polynomial solutions, so that $`I`$ has 2 rational solutions, $`(xy^23xy+3x1)/(y1)^3`$ and $`(xy)/y(y1)^3`$. ## 4 Solutions by duality For holonomic $`M`$ and $`N`$, it is well known that $$\text{Ext}_D^i(M,N)H^{in}(\mathrm{\Omega }_D^L(𝐃(M)_{𝐤[x]}^LN)),$$ (5) where $`\mathrm{\Omega }:=(D/\{x_1,\mathrm{},x_n\}D)`$, and $`𝐃(M)`$ is the holonomic dual, $$𝐃(M):=\text{Hom}_{𝐤[x]}(\mathrm{\Omega },\text{Ext}_D^n(M,D)).$$ The spaces $`\text{Ext}_D^i(M,N)`$ are finite-dimensional $`𝐤`$-vector spaces and correspond to the solutions of $`M`$ in $`N`$ when $`i=0`$. For example, if $`N=𝐤[𝐱]`$, then we obtain the polynomial solutions of $`M`$, whereas if $`N=D/D\{x_1,\mathrm{},x_n\}`$, then we obtain the delta function solutions of $`M`$ with support at the origin. In this section, we explain how (5) can be used to compute the dimensions of $`\text{Ext}_D^i(M,N)`$. We first discuss how to compute the holonomic dual, next discuss the special cases $`N=𝐂[𝐱]`$ and $`N=𝐂[𝐱][\frac{1}{f}]`$, and last discuss the general case of holonomic $`N`$. A method to extend these algorithms to compute an explicit basis of $`\text{Hom}_D(M,N)`$ and $`\text{Ext}_D^i(M,N)`$ is the subject of the forthcoming paper . Notation: Let us explain the notation we will use to write maps of left or right $`D`$-modules. As usual, maps between finitely generated modules will be represented by matrices, but some care has to be given to the order in which elements are multiplied due to the noncommutativity of $`D`$. Given an $`r\times s`$ matrix $`A=[a_{ij}]`$ with entries in $`D`$, we get a map of free left $`D`$-modules, $$D^r\stackrel{A}{}D^s[g_1,\mathrm{},g_r][g_1,\mathrm{},g_r]A,$$ where $`D^r`$ and $`D^s`$ are regarded as modules of row vectors, and the map is matrix multiplication. Under this convention, the composition of maps $`D^r\stackrel{A}{}D^s`$ and $`D^s\stackrel{B}{}D^t`$ is the map $`D^r\stackrel{AB}{}D^t`$ where $`AB`$ is usual matrix multiplication. In general, suppose $`M`$ and $`N`$ are left $`D`$-modules with presentations $`D^r/M_0`$ and $`D^s/N_0`$. Then the matrix $`A`$ induces a left $`D`$-module map between $`M`$ and $`N`$, denoted $`(D^r/M_0)\stackrel{A}{}(D^s/N_0)`$, precisely when $`\stackrel{}{g}AN_0`$ for all row vectors $`\stackrel{}{g}M_0`$. Conversely, any map of left $`D`$-modules between $`M`$ and $`N`$ can be represented by some matrix $`A`$ in the manner above. Now let us discuss maps of right $`D`$-modules. The $`r\times s`$ matrix $`A`$ also defines a map of right $`D`$-modules in the opposite direction, $$(D^s)^T\stackrel{A}{}(D^r)^T[h_1,\mathrm{},h_s]^TA[h_1,\mathrm{},h_s]^T,$$ where the superscript-$`T`$ means to regard the free modules $`(D^s)^T`$ and $`(D^r)^T`$ as consisting of column vectors. This map is equivalent to the map obtained by applying $`\text{Hom}_D(,D)`$ to $`D^r\stackrel{A}{}D^s`$, thus $`(D^s)^T`$ may be regarded as the dual module $`\text{Hom}_D(D^s,D)`$. We will suppress the superscript-$`T`$ when the context is clear. As before, the matrix $`A`$ induces a right $`D`$-module map between right $`D`$-modules $`N^{}=(D^s)^T/N_0^{}`$ and $`M^{}=(D^r)^T/M_0^{}`$ when $`A\stackrel{}{g}M_0^{}`$ for all column vectors $`\stackrel{}{g}N_0^{}`$. We denote the map by $`(D^s)^T/N_0^{}\stackrel{A}{}(D^r)^T/M_0^{}`$. Left-right correspondence and $`𝛀`$: As is well-known, a standard use for $`\mathrm{\Omega }`$ is to establish a correspondence between the categories of left and right $`D`$-modules. The correspondence can be expressed through the adjoint operator $`\tau `$, which is the algebra involution $$\tau :DDx^\alpha ^\beta ()^\beta x^\alpha .$$ Namely, given a left $`D`$-module $`MD^r/M_0`$, the corresponding right $`D`$-module is $`\mathrm{\Omega }_{𝐤[𝐱]}MD^r/\tau (M_0)`$. Conversely, given a right $`D`$-module $`ND^s/N_0`$, the corresponding left $`D`$-module is $`\text{Hom}_{𝐤[𝐱]}(\mathrm{\Omega },N)D^s/\tau (N_0)`$. Similarly, given a homomorphism of left $`D`$-modules $`\varphi :(D^r/M_0)(D^s/N_0)`$ defined by left multiplication by the $`r\times s`$ matrix $`A=[a_{ij}]`$, the corresponding homomorphism of right $`D`$-modules $`\tau (\varphi ):(D^r/\tau (M_0))(D^s/\tau (N_0))`$ is defined by right multiplication by the $`s\times r`$ matrix $`\tau (A):=[\tau (a_{ij})]^T`$. Let us explain details of the above correspondence for the non-specialist. Given a left $`D`$-module $`M`$, there is a corresponding right $`D`$-module $`\mathrm{\Omega }_{𝐤[𝐱]}M`$ where the structure is given by extending the actions, $$(wm)f=wfm(wm)\xi =w\xi mw\xi m$$ for $`f𝐤[𝐱]`$ and $`\xi \text{Der}(𝐤[𝐱])`$. Given a presentation $`D^r/M_0`$ for $`M`$ with generators denoted $`\{e_i\}_{i=1}^r`$, then in $`\mathrm{\Omega }_{𝐤[𝐱]}M`$ we have $$(1e_i)x^\alpha ^\beta =(1x^\alpha e_i)^\beta =1()^\beta x^\alpha e_i=1\tau (x^\alpha ^\beta )e_i.$$ It follows that $`\mathrm{\Omega }_{𝐤[𝐱]}M`$ is generated by $`\{1e_i\}_{i=1}^r`$ and gets the presentation $`D_n^r/\tau (M_0)`$. Conversely, given a right $`D`$-module $`N`$, there is a corresponding left $`D`$-module $`\text{Hom}_{𝐤[𝐱]}(\mathrm{\Omega },N)`$ where the structure is given by extending the action, $$(f\phi )(w)=\phi (w)f(\xi \phi )(w)=\phi (w\xi )\phi (w)\xi $$ for $`\phi \text{Hom}_{𝐤[𝐱]}(\mathrm{\Omega },N)`$, $`w\mathrm{\Omega }`$, $`f𝐤[𝐱]`$, and $`\xi \text{Der}(𝐤[𝐱])`$. A morphism $`\phi \text{Hom}_{𝐤[𝐱]}(\mathrm{\Omega },N)`$ can be identified with its image $`\phi (1)N`$. Since $$\begin{array}{cc}(x^\alpha ^\beta \phi )(1)& =(x^\alpha (^\beta \phi ))(1)=(^\beta \phi )(1)x^\alpha =\phi (1)()^\beta x^\alpha \hfill \\ & =\phi (1)\tau (x^\alpha ^\beta ),\hfill \end{array}$$ the morphism $`x^\alpha ^\beta \phi `$ gets identified with $`\phi (1)\tau (x^\alpha ^\beta )`$. In particular, given a presentation $`D^s/N_0`$ of $`N`$, then $`\text{Hom}_{𝐤[𝐱]}(\mathrm{\Omega },N)`$ is generated as a left $`D`$-module by the morphisms $`\{\phi _i\}_{i=1}^s`$ such that $`\phi _i(1)=e_i`$. By the computation above, a relation $`_ie_ig_i=0`$ in $`N`$ corresponds to a relation $`_i\tau (g_i)\phi _i`$ in $`\text{Hom}_{𝐤[𝐱]}(\mathrm{\Omega },N)`$ because $`(_i\tau (g_i)\phi _i)(1)=_ie_i\tau (\tau (g_i))=_ie_ig_i.`$ It follows that $`\text{Hom}_{𝐤[𝐱]}(\mathrm{\Omega },N)`$ is generated by $`\{\phi _i\}_{i=1}^s`$ and gets the presentation $$\mathrm{Hom}_{𝐤[𝐱]}(\mathrm{\Omega },N)D_n^s/\tau (N_0).$$ (6) ### 4.1 Holonomic dual Let us discuss how $`𝐃(M)`$ can be computed. ###### Algorithm 4.1 \[Computing the holonomic dual\] Input: $`D^{r_0}/D\{\stackrel{}{g}_1,\mathrm{},\stackrel{}{g}_{r_1}\}`$, a presentation of a holonomic left $`D`$-module $`M`$. Output: The holonomic dual $`𝐃(M)`$. 1. Compute the first $`n+1`$ steps of any free resolution of $`M`$. Let the $`n`$-th part of the resolution be $`D^p\stackrel{P}{}D^q\stackrel{Q}{}D^r.`$ 2. Dualize and apply the adjoint operator (recall if $`P=[p_{ij}]`$, then $`\tau (P)=[\tau (p_{ij})]^T`$) to get $`D^p\stackrel{\tau (P)}{}D^q\stackrel{\tau (Q)}{}D^r.`$ 3. Return $`\mathrm{ker}(\tau (P))/\text{Im}(\tau (Q))`$. Proof: Let the first $`n+1`$ steps of a free resolution of $`M`$ be denoted, $$F^{}:D^{r_{n+1}}\stackrel{P}{}D^{r_n}\stackrel{Q}{}D^{r_{n1}}\mathrm{}D^{r_0}0.$$ Applying $`\text{Hom}_D(D,)`$ yields a complex of right $`D`$-modules, $$\text{Hom}_D(D,F^{}):(D^{r_{n+1}})^T\stackrel{P}{}(D^{r_n})^T\stackrel{Q}{}(D^{r_{n1}})^T\mathrm{}(D^{r_0})^T0,$$ and by definition, $$\text{Ext}_D^n(M,D)\frac{\mathrm{ker}(D^{r_{n+1}}\stackrel{P}{}D^{r_n})}{\text{Im}(D^{r_n}\stackrel{Q}{}D^{r_{n1}})}.$$ Since $`𝐃(M)=\text{Hom}_{𝐤[𝐱]}(\mathrm{\Omega },\text{Ext}_D^n(M,D))`$, it only remains to determine the effect of applying $`\text{Hom}_{𝐤[𝐱]}(\mathrm{\Omega },)`$. Using the equation (6), if $`\{\stackrel{}{L}_1,\mathrm{},\stackrel{}{L}_k\}`$ are generators of $`K=\mathrm{ker}(D^{r_{n+1}}\stackrel{P}{}D^{r_n})`$, and $`_i\stackrel{}{L}_ig_iI=\text{Im}(D^{r_n}\stackrel{Q}{}D^{r_{n1}})`$ is a relation, then the corresponding relation $`_i\tau (g_i)\phi _i`$ in $`\text{Hom}_{𝐤[𝐱]}(\mathrm{\Omega },\text{Ext}_D^n(M,D))`$ can be realized as the relation $`_i\tau (\stackrel{}{L}_ig_i)=\tau (g_i)\tau (\stackrel{}{L}_i)\tau (I)`$. It follows that $$𝐃(M)\frac{\mathrm{ker}(D^{r_{n+1}}\stackrel{\tau (P)}{}D^{r_n})}{\text{Im}(D^{r_n}\stackrel{\tau (Q)}{}D^{r_{n1}})},$$ which is the output of step 3. \[\] ###### Example 4.2 The Appell differential equation $`F_1(2,3,2,5)`$ of Example 2.5 has the resolution $`0D^1\stackrel{Q_1}{}D^2\stackrel{Q_0}{}D^10,`$ where $$\begin{array}{c}Q_0=\left[\begin{array}{c}(\theta _x3)_y(\theta _y2)_x\hfill \\ (y^2y)(_x_y+_y^2)2(y+x)_x+4y_y+2_x8_y4\hfill \end{array}\right]^T\hfill \\ \\ Q_1=\left[\begin{array}{c}(y^2y)(_x_y+_y^2)2x_x+6y_y+_x9_y\\ (\theta _x3)_y+(\theta _y1)_x\end{array}\right]\hfill \end{array}$$ The holonomic dual $`𝐃(F_1(2,3,2,5))`$ is the cokernel of $`\tau (Q_1)`$ and is the Appell differential equation $`F_1(1,4,2,3)`$. ### 4.2 Polynomial and rational solutions by duality When $`N=𝐤[𝐱]`$, the isomorphism (5) specializes to $$\text{Ext}_D^i(M,𝐤[𝐱])H^{ni}(\mathrm{\Omega }_D^L𝐃(M)).$$ (7) The right hand side is equivalently the $`(ni)`$-th integration of $`𝐃(M)`$ to the origin. An algorithm to compute integration is given in . Using it, we can evaluate the dimensions of $`\text{Ext}_D^i(M,𝐤[𝐱])`$ and in particular $`\text{Hom}_D(M,𝐤[𝐱])`$. ###### Algorithm 4.3 \[Evaluating dimensions of polynomial solution spaces\] Input: a holonomic left $`D`$-module $`M`$. Output: dimensions of $`\text{Ext}_D^i(M,𝐤[𝐱])`$. 1. Compute the dual $`𝐃(M)`$ using Algorithm 4.1 2. Compute the integrations of $`𝐃(M)`$ to the origin using the algorithm in . They are finite dimensional vector spaces. 3. Return the dimensions. The dimensions of rational solution spaces can be evaluated in a similar way. When $`N=𝐤[𝐱][\frac{1}{f}]`$, the isomorphism (5) specializes to $$\text{Ext}_D^i(M,𝐤[𝐱][1/f])H^{ni}(\mathrm{\Omega }_D^L𝐃(M)[1/f]).$$ (8) The right hand side is now equivalently the $`(ni)`$-th integration of $`𝐃(M)[\frac{1}{f}]`$ to the origin. An algorithm to compute localization is given in . Using it and the integration algorithm, we can evaluate the dimensions of $`\text{Ext}_D^i(M,𝐤[𝐱][\frac{1}{f}])`$ and $`\text{Hom}_D(M,𝐤[𝐱][\frac{1}{f}])`$. To get the dimension of all rational solutions, take $`f`$ to be any polynomial vanishing on the singular locus. We summarize how to compute the integration of a module $`N`$ to the origin according to in a slightly more general way. The generalization sometimes gives a more efficient strategy than . We need to recall some definitions. To any strictly positive $`w𝐙_{>0}^n`$, we get an integration filtration $`F_w`$ of $`D`$ defined by $`F_w^i(D)=\text{Span}_k\{x^\alpha ^\beta |w\alpha w\beta i\}`$. More generally, for $`\stackrel{}{m}Z^r`$, we also get a shifted filtration $`F_w[\stackrel{}{m}]`$ of the free module $`D^r`$ defined by $`F_w^i[\stackrel{}{m}](D^r)=\text{Span}_k\{x^\alpha ^\beta e_j|w\alpha w\beta m_ji\}`$. We will often write $`D^r[\stackrel{}{m}]`$ for the free module $`D^r`$ equipped with the shifted filtration $`F_w[\stackrel{}{m}]`$ when the context is clear. The filtrations $`F_w[\stackrel{}{m}]`$ induce filtrations on subquotients of $`D^r`$ in the natural way. Now we may say the steps of the integration algorithm. First, compute a $`(w,w)`$-strict free resolution $`G^{}`$ of $`N`$ of length $`n+1`$. This is a resolution of $`N`$ by free modules $`D^{r_j}[\stackrel{}{m}_j]`$ with the property that the differentials preserve the filtration and moreover induce a resolution on the associated graded level. Second, compute the integration $`b`$-function of $`N`$ with respect to $`(w,w)`$, and find its minimal and maximal integral roots $`k_0`$ and $`k_1`$. The integration $`b`$-function is the monic polynomial $`b(s)`$ of least degree satisfying $`b(_iw_i_ix_i)F^0(N)F^1(N)`$. Third, compute the cohomology of the complex $`F_{k_0}(\mathrm{\Omega }_DG^{})/F_{k_11}(\mathrm{\Omega }_DG^{})`$, which is a complex of finite-dimensional vector spaces. The dimensions of the cohomology groups are equal to the dimensions of the integration modules of $`N`$. ###### Example 4.4 Let us evaluate the dimension of polynomial solutions to the Appell differential equation $`M=F_1(2,3,2,5)`$ of Example 2.5. Choose the weight vector $`w=(1,2)`$. The resolution of Example 4.2, after dualizing, applying the adjoint operator, and shifting, $$0D^1[0]\stackrel{\tau (Q_0)}{}D^2[1,1]\stackrel{\tau (Q_1)}{}D^1[0]0,$$ preserves filtrations but does not induce a resolution on the associated graded level. On the other hand, if we adjust the resolution to $$G^{}:0D^1[1]\stackrel{P_0}{}D^2[0,1]\stackrel{P_1}{}D^1[0]0,$$ where $$P_0=\left[\begin{array}{c}(\theta _x+5)_y+(\theta _y+2)_x\\ (x^2x)(_x^2+_x_y)+4x_x+2(3x+2y)_y+4_x5_y2\end{array}\right]^T$$ $$P_1=\left[\begin{array}{c}(x^2x)(_x^2+_x_y)+2x_x+4(x+y)_y+5_x4_y4\\ (\theta _x+4)_y(\theta _y+2)_x\end{array}\right],$$ then we do obtain a $`(w,w)`$-strict resolution of $`𝐃(M)=F_1(1,4,2,3)`$. The integration $`b`$-function with respect to $`(w,w)`$ is $`(s+5)(s2)(s5)`$, hence the integration complex for $`𝐃(M)`$ is quasi-isomorphic to the truncated complex $`F_w^5(\mathrm{\Omega }_DG^{})/F_w^6(\mathrm{\Omega }_DG^{})`$, which is a complex of finite-dimensional vector spaces with dimensions, $$0𝐐^{16}\stackrel{P_0}{}𝐐^{28}\stackrel{P_1}{}𝐐^{12}0.$$ For instance, $`F_w^5(\mathrm{\Omega }[0])`$ consists of the $`12`$ monomials, $$\{1,x,y,x^2,xy,y^2,x^3,x^2y,xy^2,x^4,xy^3,x^5\},$$ and so on. Note that $`\tau (P_1)`$ is a $`(w,w)`$-Gröbner basis of $`F_1(2,3,2,5)`$ and hence for this case, the duality method essentially coincides with the Gröbner deformation method of Section 2 at the level of $`\mathrm{Hom}_D(M,𝐤[𝐱])`$. The above computations were made in Macaulay 2, where we get the output, $$\begin{array}{ccccccc}\mathrm{𝚒𝟷}& :& \mathrm{𝙿𝚘𝚕𝚢𝙴𝚡𝚝}\left(𝙼\right)\hfill & & & & \\ \mathrm{𝚘𝟷}& =& \mathrm{𝙷𝚊𝚜𝚑𝚃𝚊𝚋𝚕𝚎}\{\hfill & \mathrm{𝟶}\hfill & =>\hfill & \mathrm{𝚀𝚀}^\mathrm{𝟷}\hfill & \}\hfill \\ & & & \mathrm{𝟷}\hfill & =>\hfill & \mathrm{𝚀𝚀}^\mathrm{𝟸}\hfill & \\ & & & \mathrm{𝟸}\hfill & =>\hfill & \mathrm{𝚀𝚀}^\mathrm{𝟷}\hfill & \end{array}$$ Here, the output $`𝚒=>\mathrm{𝚀𝚀}^𝚓`$ means that $`dim\text{Ext}_D^i(M,𝐤[𝐱])=j`$. ###### Example 4.5 Let us now evaluate the dimension of rational solutions to $`M=F_1(2,3,2,5)`$. The singular locus is $`xy(xy)(x1)(y1)`$. We will search for solutions in $`𝐤[x,y][\frac{1}{x}]`$ first. From Example 4.2, $`𝐃(M)`$ has the presentation $`D/\tau (Q_1)`$. Let $`u`$ be the section corresponding to the residue class of $`\overline{1}`$ in this presentation. Then the localization $`𝐃(M)[\frac{1}{x}]`$ is generated by $`u\frac{1}{x^7}`$ and gets the presentation $`D/J`$, where $$J=D\left\{\begin{array}{c}(\theta _x\theta _y+\theta _y^2+8\theta _y+2\theta _x+12)(\theta _x+\theta _y+4)_y\\ (\theta _x\theta _y+2\theta _x+7\theta _y+14)(\theta _x+10)x_y\end{array}\right\}.$$ The natural localization map can be written as $`\phi :D/\tau (Q_1)D/J`$, where $`\phi (1)=x^7`$. Choose the integration weight vector $`w=(1,2)`$. Then $`𝐃(M)[\frac{1}{x}]`$ has a $`(w,w)`$-strict resolution $$G^{}:0D^1[1]\stackrel{[v_1,v_2]}{}D^2[0,1]\stackrel{[u_1,u_2]^T}{}D^1[0]0$$ where $$\begin{array}{ccc}u_1& =& x^2_x_y+xy_x_y+2x_x11x_y+7y_y+14\hfill \\ u_2& =& x^3_x^2+x^3_x_yx^2_x^2x^2_x_y+16x^2_x+11x^2_y+\hfill \\ & & 4xy_y9x_x11x_y+52x7\hfill \\ v_1& =& x^3_x^2+x^3_x_yx^2_x^2x^2_x_y+16x^2_x+12x^2_y+\hfill \\ & & 4xy_y8x_x11x_y+52x6\hfill \\ v_2& =& x^2_x_yxy_x_y2x_x+11x_y6y_y12\hfill \end{array}$$ The integration $`b`$-function is $`(s+12)(s+5)(s+2)`$, hence we want the cohomology of the complex $`F_w^{12}(\mathrm{\Omega }_DG^{})/F_w^1(\mathrm{\Omega }_DG^{})`$, which has the shape, $$0𝐐^{41}𝐐^{86}𝐐^{45}0.$$ By evaluating the dimensions of the cohomology groups in Macaulay 2, we find $$\begin{array}{ccccccc}\mathrm{𝚒𝟸}& :& \mathrm{𝚁𝚊𝚝𝚕𝙴𝚡𝚝}(𝙼,𝚡)\hfill & & & & \\ \mathrm{𝚘𝟸}& =& \mathrm{𝙷𝚊𝚜𝚑𝚃𝚊𝚋𝚕𝚎}\{\hfill & \mathrm{𝟶}\hfill & =>\hfill & \mathrm{𝚀𝚀}^\mathrm{𝟸}\hfill & \}\hfill \\ & & & \mathrm{𝟷}\hfill & =>\hfill & \mathrm{𝚀𝚀}^\mathrm{𝟻}\hfill & \\ & & & \mathrm{𝟸}\hfill & =>\hfill & \mathrm{𝚀𝚀}^\mathrm{𝟹}\hfill & \end{array}$$ Since we already computed a polynomial solution, this means there is one rational solution with pole along $`x`$. Similarly, we get the exact same dimensions for $`\text{Ext}_D^i(M,𝐤[x,y][{\scriptscriptstyle \frac{1}{y}}])`$, which means that there is also one rational solution with pole along $`y`$. The rank of the system is $`3`$, therefore we have found all the solutions. We could also compute, $$\begin{array}{ccccccc}\mathrm{𝚒𝟹}& :& \mathrm{𝚁𝚊𝚝𝚕𝙴𝚡𝚝}(𝙼,𝚏)\hfill & & & & \\ \mathrm{𝚘𝟹}& =& \mathrm{𝙷𝚊𝚜𝚑𝚃𝚊𝚋𝚕𝚎}\{\hfill & \mathrm{𝟶}\hfill & =>\hfill & \mathrm{𝚀𝚀}^\mathrm{𝟷}\hfill & \}\hfill \\ & & & \mathrm{𝟷}\hfill & =>\hfill & \mathrm{𝚀𝚀}^\mathrm{𝟹}\hfill & \\ & & & \mathrm{𝟸}\hfill & =>\hfill & \mathrm{𝚀𝚀}^\mathrm{𝟸}\hfill & \end{array}$$ where $`f`$ is any of the polynomials $`xy`$, $`x1`$, or $`y1`$. As expected, there are no rational solutions with poles along $`xy`$, $`x1`$, or $`y1`$, but in all cases there are new $`\text{Ext}^1`$ and $`\text{Ext}^2`$. We have not computed Ext with respect to any products of poles since it is computationally too intensive for now. Once we have evaluated the dimension of the solution spaces, we can compute the solutions by a brute force method. 1. For a given holonomic system $`M`$, compute its singular locus. Let $`f`$ be a polynomial such that $`f=0`$ contains the singular locus. 2. Evaluate the dimension $`d`$ of the rational solutions by the homological duality method. 3. Try to find rational solutions of the form $`r(x)/f^k`$, $`\mathrm{degree}(r)=p`$. Increase $`p+k`$ until we find $`d`$ linearly independent solutions. ### 4.3 Holonomic solutions by duality The isomorphism (5) can also be expressed (see e.g. ) as $$\text{Ext}_D^i(M,N)\text{Tor}_{ni}^D(\text{Ext}_D^n(M,D),N).$$ To compute the right-hand-side, it is also well known that for all $`M^{}`$, and in particular $`M^{}=\text{Ext}_D^n(M,D)`$, $$\text{Tor}_{ni}^D(M^{},N)H^i(K^{}((M^{}_{𝐤[𝐱]}\mathrm{\Omega }^1)\widehat{}N;\{x_iy_i,_i+\delta _i\}_{i=1}^n)),$$ where $`K^{}`$ denotes the Koszul complex and $`\widehat{}`$ denotes the external tensor product into the category of $`D_{2n}=𝐤x_1,y_1,\mathrm{},x_n,y_n,_1,\delta _1,\mathrm{},_n,\delta _n`$-modules. Combining these isomorphisms leads to $$\text{Ext}_D^i(M,N)H^i(K^{}((𝐃(M)\widehat{}N;\{x_iy_i,_i+\delta _i\}_{i=1}^n))$$ By an automorphism of $`D`$, we can transform $`\{x_iy_i,_i+\delta _i\}_{i=1}^n`$ into $`\{x_i,y_i\}_{i=1}^n`$, for which the Koszul complex computes the derived restriction to the origin. ###### Algorithm 4.6 \[Evaluating dimensions of holonomic solution spaces\] Input: holonomic left $`D`$-modules $`M`$ and $`N`$ Output: dimensions of $`\text{Ext}_D^i(M,N)`$. 1. Compute the dual $`𝐃(M)`$ using Algorithm 4.1 2. Form the $`D_{2n}`$-module $`𝐃(M)_kN`$ and apply the change of coordinates $`\eta :D_{2n}D_{2n}`$ where $`\eta `$ maps, $$\begin{array}{ccc}x_i\frac{1}{2}x_i\delta _i,\hfill & & _i\frac{1}{2}y_i+_i,\hfill \\ y_i\frac{1}{2}x_i\delta _i,\hfill & & \delta _i\frac{1}{2}y_i_i\hfill \end{array}.$$ 3. Compute the restrictions of $`\eta (𝐃(M)_kN)`$ to the origin using the algorithm in . They are finite dimensional vector spaces. 4. Return the dimensions. ###### Example 4.7 Let $`M=F_1(2,3,2,5)`$ be the Appell differential equation of example 2.5, and let $`N=𝐤[x,y][\frac{1}{x}]/𝐤[x,y]`$. It has presentation $`D/D\{x,_y\}`$, where the generator $`1`$ corresponds to $`\frac{1}{x}`$. Using the above algorithm, we compute $$\begin{array}{ccccccc}\mathrm{𝚒𝟺}& :& \mathrm{𝙳𝙴𝚡𝚝}(𝙼,𝙽)\hfill & & & & \\ \mathrm{𝚘𝟺}& =& \mathrm{𝙷𝚊𝚜𝚑𝚃𝚊𝚋𝚕𝚎}\{\hfill & \mathrm{𝟶}\hfill & =>\hfill & \mathrm{𝚀𝚀}^\mathrm{𝟷}\hfill & \}\hfill \\ & & & \mathrm{𝟷}\hfill & =>\hfill & \mathrm{𝚀𝚀}^\mathrm{𝟹}\hfill & \\ & & & \mathrm{𝟸}\hfill & =>\hfill & \mathrm{𝚀𝚀}^\mathrm{𝟸}\hfill & \end{array}$$ Similarly, let $`N=𝐤[_x,_y]D/D\{x,y\}`$, the module of the delta functions with the support $`(0,0)`$. Then we compute $$\begin{array}{ccccccc}\mathrm{𝚒𝟻}& :& \mathrm{𝙳𝙴𝚡𝚝}(𝙼,𝙽)\hfill & & & & \\ \mathrm{𝚘𝟻}& =& \mathrm{𝙷𝚊𝚜𝚑𝚃𝚊𝚋𝚕𝚎}\{\hfill & \mathrm{𝟶}\hfill & =>\hfill & \mathrm{𝚀𝚀}^\mathrm{𝟶}\hfill & \}\hfill \\ & & & \mathrm{𝟷}\hfill & =>\hfill & \mathrm{𝚀𝚀}^\mathrm{𝟷}\hfill & \\ & & & \mathrm{𝟸}\hfill & =>\hfill & \mathrm{𝚀𝚀}^\mathrm{𝟸}\hfill & \end{array}$$ As before, once we know the dimension of $`\text{Hom}_D(M,N)`$, we can compute the solutions of $`M`$ in $`N`$ by a brute force method. 1. For given holonomic systems $`M`$ and $`N`$, evaluate the dimension $`d`$ of $`\text{Hom}_D(M,N)`$ by the homological duality method. 2. Filter $`N`$ by finite-dimensional vector spaces $`F^i(N)`$ and search for solutions in $`F_i(N)`$ for increasing $`i`$ until $`d`$ linearly independent solutions are found. For instance in step 2, if $`N=D/J`$, then we can use the induced Bernstein filtration $`B`$ where $`B^i(D/J)`$ consists of residues of elements $`LD`$ whose total degree is less than or equal to $`i`$. Toshinori Oaku, oaku@twcu.ac.jp Department of Mathematics, Tokyo Woman’s Christian University 2-6-1 Zempukuji, Suginami-ku, Tokyo 167-8585, Japan Nobuki Takayama, takayama@math.kobe-u.ac.jp Department of Mathematics, Kobe University Kobe 657-8501, Japan Harrison Tsai, htsai@math.berkeley.edu Department of Mathematics, University of California, Berkeley Berkeley, CA 94720-3840, USA
warning/0001/hep-th0001136.html
ar5iv
text
# CANCELLATION OF LINEARISED AXION-DILATON SELF ACTION DIVERGENCE IN STRINGS ## 1 Introduction This article is intended as a contribution to the clarification of a question raised by Dabholkar and Harvey who pointed out that the finiteness of superstring theory implied a similar finiteness for the corresponding low energy classical limit theory as constructed in terms of Nambu Goto type strings interacting via coupling to gravitational, dilatonic, and axionic type fields. What this means is that although the gravitational, dilatonic and axionic self interaction contributions will each be separately divergent in generic classical string models, their net effect should cancel out – so that no “infinite” renormalisation is required – in the special case of the particular model obtained from the low energy limit of superstring theory. Although there has always been a general concensus to the effect that this noteworthy prediction is indeed correct, the question of how the expected cancellation actually comes about has been the subject of considerable discord. Relying on the straightforward – but with hindsight obviously flawed – analysis of Dabholkar and Harvey themselves, the opinion that was most widely held until recently is succinctly summarised by the assertion that “Since the quantum answer is zero the classical answer must be zero. This is indeed true. There are three divergent contributions to the classical self energy. The dilaton contribution is the same as the axion but the gravitational contribution is negative and twice as large.” The flaw in this particular analysis is that it takes account only of the external field energy contributions, while neglecting the internal (world sheet supported) contributions, either because the authors assumed (wrongly) that they would be relatively negligible, or else because they simply forgot about them, perhaps due to having started by working out the axion part, for which the previous literature was already well developped and for which it just so happens that there is no internal contribution. There was however a noteworthy dissenting opinion: while agreeing that that total was indeed zero, it was claimed on the basis of a rather obscure calculation by Copeland, Haws and Hindmarsh that – far from being the same as that of the axion – the dilaton contribution cancelled out all by itself, leaving the axion contribution to be cancelled just by the gravitational contribution. It appears with hindsight that this nonconformist conclusion was not quite so wide of the mark as that of Dabholkar and Harvey, but it does not seem to have been taken so seriously, presumably because of the failure to make clear the logical reasonning on which it was supposed to have been based. It has now become evident that both of these competing alternatives were incorrect, essentially for the same underlying reason, which was the use of naive force or energy formulae based on the omission of various important terms that were wrongly assumed to be negligible or simply forgotten about. These overhasty omissions were originally motivated to a large extent by the technical difficulties involved in actually evaluating the terms in question, but such difficulties have recently been alleviated by the introduction of more efficient geometrical methods. Following the derivation by these methods of the correct formula for the complete gravitational force contribution acting on a general string model, it has recently been found that the divergent part of the gravitational self force cancels itself out all by itself in the case of a Nambu-Goto string in 4-dimensions. The implication of this is that – in order for the total to vanish in the low energy classical limit derived from superstring theory – the corresponding dilaton contribution should neither be equal to the axion contribution (as asserted by Dabholkar and Harvey) nor even zero (as asserted by Copeland, Haws and Hindmarsh) but in fact exactly opposite to the axion contribution, at least in a 4-dimensional background. The purpose of the present work is to provide a direct verification of this corollary – i.e. of the mutual cancellation of the relevant axionic and dilatonic string self action divergences for the linearised classical limit derived from superstring theory in 4-dimensions. A less direct confirmation, and and a generalisation to higher dimensions (where the total still vanishes, but not the gravitational part on its own) has already been provided using a new approach based on the use of an effective action by Buonanno and Damour . It is worthwhile to provide further confirmation because the pertinance of this kind of approach was explicitly – but unjustly – cast into doubt by Copeland, Haws and Hindmarsh who alleged that “in general it is not correct” to deduce the divergent part of the self force from the divergent part of the effective action, the reason for their scepticism being the discrepancies that arose in their own rather incoherent approach. The verdict of the present analysis is that use of the effective action approach is inherently correct after all, and that there are no discrepancies, provided all the calculations are carried out properly without omission of relevant terms. The consistency of an approach based on a force analysis with an approach using an effective action had previously been demonstrated for electromagnetic interactions in strings . Moreover the the results of the effective action approach used by Buonanno and Damour were known from the outset to be consistent with the detailed force analysis – as correctly carried out with all relevant terms included – for purely gravitational interactions . In this gravitational case the detailed relationship between the two kinds of approach has since been demonstrated explicitly . The present work provides an analogously explicit demonstration of consistency between the effective action approach used by Buonanno and Damour and a detailed force analysis for the technically simpler cases of dilatonic and axionic interactions. ## 2 Lagrangian for dilatonic and axial current couplings. Before considering the divergences that arise from self interaction it is first necessary to consider the effect of linear interactions with generic fields of the kinds with which we are concerned, namely a dilatonic (scalar) field $`\varphi `$ and an axionic (pseudoscalar) field represented by a antisymmetric Kalb-Ramond type 2-form $`B_{\mu \nu }`$ say. We shall ultimately be concerned with the application to the classical low energy limit of superstring theory which also involves a symmetric gravitational perturbation field $`h_{\mu \nu }`$. However due to the assumption of linearity (which will be physically justified when the fields are sufficiently weak) each of the three pertinant (dilatonic, axionic, and gravitational) parts can be treated independently of the other two. Since the relevant analysis of the gravitational part is already available the present article will be restricted to the corresponding analysis for the technically simple axionic and dilatonic parts. Unlike the dilatonic part, the axionic part seems to have been correctly treated in the earlier work , but we shall work through it again here in order to demonstrate the use of the neater, more efficient, and therefore less error prone mathematical formalism that have since been developed and that are indispensible for avoiding unnecessarily heavy algebra in more complicated applications such the gravitational part, and that are helpful even for relatively simple applications such as to the dilatonic part dealt with here. The action governing the kind of system to be analysed here will consist of a total, $`_{_{\mathrm{to}}}=_{_{\mathrm{ra}}}+_{_{\mathrm{ma}}}`$ in which the first term is a a free radiation contribution of the form $$_{_{\mathrm{ra}}}=\widehat{}_{_{\mathrm{ra}}}g^{1/2}d^4x,$$ (1) where $`g`$ is the modulus of the divergence of the 4-dimensional spacetime background metric $`g_{\mu \nu }`$ as expressed with respect to local coordinates $`x^\mu `$, in which the Lagrangian density scalar $`\widehat{}_{_{\mathrm{ra}}}`$ depends only, in a homogeneous quadratic manner, on the relevant linearised long range field variables, namely $`\varphi `$ and $`B_{\mu \nu }`$ in the present instance. The other part of the action is a material contribution $$_{_{\mathrm{ma}}}=\widehat{}_{_{\mathrm{ma}}}g^{1/2}d^4x,$$ (2) in which the Lagrangian density scalar $`\widehat{}_{_{\mathrm{ma}}}`$ is restricted to have a purely linear dependence on these long range field variables, while also having a generically non-linear dependence on whatever other variables may be needed to characterise the localised material system under consideration, which in the application that follows will be taken to be string. It will be postulated that the material system is unpolarised in the technical sense that its dependence on the linearised long range field variables does not involve derivatives, which means that its Lagrangian density scalar will have the form $$\widehat{}_{_{\mathrm{ma}}}=\widehat{}+\widehat{}_{_{\mathrm{co}}}$$ (3) in which the primary contribution $``$ is entirely independent of the linearised long range field variables $`\varphi `$ and $`B_{\mu \nu }`$ while the coupling term $`\widehat{}_{_{\mathrm{co}}}`$ will have the homogeneous form $$\widehat{}_{_{\mathrm{co}}}=\frac{_1}{^2}\widehat{W}^{\mu \nu }B_{\mu \nu }+\widehat{T}\varphi ,$$ (4) in which the antisymmetric tensor field $`\widehat{W}^{\mu \nu }`$ and the scalar field $`\widehat{T}`$ are each independent of both $`B_{\mu \nu }`$ and $`\varphi `$. Unlike the linear coupling term the homogeneously quadratic free radiation contribution $`_{_{\mathrm{ra}}}`$ will involve field gradients. This contribution will be given in terms of covariant differentiation with respect to the background metric $`g_{\mu \nu }`$, for which the usual symbol $``$ will be used, by an expression of the form $$_{_{\mathrm{ra}}}=\frac{1}{8\pi }\left(m__\mathrm{D}^{\mathrm{\hspace{0.17em}2}}\varphi ^{;\mu }\varphi _{;\mu }+\frac{1}{6m__\mathrm{J}^{\mathrm{\hspace{0.17em}2}}}J^{\mu \nu \rho }J_{\mu \nu \rho }\right),$$ (5) using the notation $$\varphi _{;\mu }=_\mu \varphi J_{\mu \nu \rho }=3_{[\mu }B_{\nu \rho ]},$$ (6) where square brackets indicate index antisymetrisation. The parameters $`m__\mathrm{D}`$ and $`m__\mathrm{J}`$ in this expression are fixed coupling constants having the dimensions of mass on the understanding that we are using units in which the speed of light $`c`$ and the Dirac-Planck constant $`\mathrm{}`$ are set to unity. The quantity $`m__\mathrm{D}`$ is what may conveniently be referred to as the Dicke mass. This dilation mass scale should not be confused with the dilaton mass, $`m__\varphi `$ say, which is usually assumed to be very small, and which is simply taken to be zero in the present work. The Dicke dilation mass scale $`m__\mathrm{D}`$ is usually supposed to be very large, at least comparable with the Planck mass defined by $`m__\mathrm{P}=\text{G}^{1/2}`$ where G is Newton’s constant. In particular, if the theory is to be applied to the modern solar system then there are severe observational limits that can be interpreted as implying that the relevant value of the dimensionless Brans-Dicke parameter $`\omega =2\text{G}m__\mathrm{D}^{\mathrm{\hspace{0.17em}2}}+3/2`$ should be very large compared with unity and hence that $`m__\mathrm{D}`$ is large compared with $`m__\mathrm{P}`$. The other mass scale, $`m__\mathrm{J}`$, is what may conveniently be referred to as the Joukowski mass, since the corresponding Kalb-Ramond coupling – of the axial kind associated in the context of superstring theory with the names of Wess and Zumino – gives rise to a lift force of the type that has long been well known in the context of aerofoil theory, where it was originally derived as a corollary of the magnus effect by the Russian theoretician Joukowski. This mass scale can also usually be supposed to be very large, unlike the associated axion mass, $`m__\mathrm{a}`$ say, with which it should not be confused. Like the dilaton mass $`m__\varphi `$, the axion mass $`m__\mathrm{a}`$ is usually assumed to be very small, and will be taken to be zero for the purposes of the present work. The axial current model obtained in this way is interpretable as the “stiff” (Zel’dovich type) limit (characterised by sound perturbation propagation at the speed of light) within the more general category of ordinary perfect fluid models. The Kalb-Ramond Wess-Zumino Joukowski coupling bivector field $`\widehat{W}^{\mu \nu }`$ is interpretable as a vorticity flux, and must be such as to satisfy a flux conservation law of the form $$_\mu \widehat{W}{}_{}{}^{\mu \nu }=0,$$ (7) in order to ensure invariance under local Kalb-Ramond gauge transformations of the form $`B_{\mu \nu }B_{\mu \nu }+2_{[\mu }\phi _{\nu ]}`$ for an arbitrary covector field $`\phi _\nu `$. In typical applications to continuous media this condition is fullfilled by by a specification of the form $`\widehat{W}^{\mu \nu }=\epsilon ^{\mu \nu \rho \sigma }\chi _{;\rho }^+\chi _{;\sigma }^{}`$ where the scalars $`\chi ^+`$ and $`\chi ^{}`$ are two of the intrinsic field variables characterising the material system and $`\epsilon ^{\mu \nu \rho \sigma }`$ is the usual antisymmetric measure tensor induced on the background space (modulo a choice of signature) by the metric $`g_{\mu \nu }`$. In such a case, as in the case of the string type systems for which the relevant formula for $`\widehat{W}^{\mu \nu }`$ will be described below, the corresponding coupling action has the special feature (which it shares with its electromagnetic analogue ) that its dependence on the metric $`g_{\mu \nu }`$ cancels out. Failure to make proper allowance for the absence of this familiar simplifying property in more general couplings, and in particular in couplings of gravitational and dilatonic type, seems to be one of the main reasons why early evaluations of the latter were so systematically erroneous. The conditions on the other coupling term are more restrictive. In principle the scalar source coefficient $`\widehat{T}`$ might have various forms for diverse scalar coupling theories that might be conceived, but in order for the coupling to be describable as “ dilatonic” in the sense associated most particularly with the work of Dicke , it must be given by the trace $$\widehat{T}=\widehat{T}_\mu {}_{}{}^{\mu },$$ (8) of the material stress energy tensor that is obtained from the primary Lagrangian contribution $``$ of the system according to the usual geometric specification $$\widehat{T}{}_{}{}^{\mu \nu }=2g^{1/2}\frac{\left(\widehat{}g^{1/2}\right)}{g_{\mu \nu }}=2\frac{\widehat{}}{g_{\mu \nu }}+\widehat{\mathrm{\Lambda }}g^{\mu \nu }.$$ (9) The reasonning whereby Dicke and other early workers (notably his predecessor Jordan and his colleague Brans) were lead to a coupling of this kind was based on the postulate that in a fully non linear treatment the primary material action contribution would be given by an expression of the form $$__\mathrm{D}=__\mathrm{D}g__\mathrm{D}^{1/2}d^4x,$$ in which the Lagrangian density $`__\mathrm{D}`$ depends on whatever intrinsic material constituent fields may be necessary as well as on a certain background metric $`g_{{}_{\mathrm{D}}{}^{}\mu \nu }`$ say, that may conveniently be referred to as the Dicke metric, in order to distinguish it from the associated Einsten metric, $`g_{{}_{\mathrm{E}}{}^{}\mu \nu }`$ say. The latter is characterised by the condition that the relevant gravitational field action be proportional to the spacetime volume integral of its Ricci tensor. In a fully non linear treatment the Dicke metric is related to the Einstein metric by a conformal transformation of the form $`g_{{}_{\mathrm{D}}{}^{}\mu \nu }=e^{2\varphi }g_{{}_{\mathrm{E}}{}^{}\mu \nu }`$. In a linearised treatment such as is used here, the relevant Einstein metric can be taken to have the form $`g_{{}_{\mathrm{E}}{}^{}\mu \nu }g_{\mu \nu }+h_{\mu \nu }`$ where the “unperturbed” background metric $`g_{\mu \nu }`$ is strictly flat or at least Ricci flat, and where $`h_{\mu \nu }`$ is the usual gravitational perturbation tensor. The linearisation implies that one can also take $`e^{2\varphi }1+2\varphi `$ and hence that the relevant Dicke metric can be taken to be given in terms of the flat or Ricci flat background by a relation of the form $`g_{{}_{\mathrm{D}}{}^{}\mu \nu }g_{\mu \nu }+h_{\mu \nu }+2\varphi g_{\mu \nu }`$. In the traditional kind of dilaton coupling theory theory (as envisaged by Dicke ), and also in the particular kind of low energy classical limit of superstring theory that was considered by Dabholkar and Harvey , the $`\varphi `$ dependence only comes in indirectly via the dependence on $`g_{{}_{\mathrm{D}}{}^{}\mu \nu }`$. (Rather more complicated couplings occur in some of the more eleborate models derived from superstring theory or M-theory but, as remarked by Cho and Keum , this does not necessarily affect the form of their linearised weak field limits.) Under such conditions it follows that in the linearised limit we shall simply have $$__\mathrm{D}g__\mathrm{D}^{1/2}\left(+T^{\mu \nu }(h_{\mu \nu }+2\varphi g_{\mu \nu })\right)g^{1/2}$$ where $``$ is obtained be substituting the unperturbed background metric $`g_{\mu \nu }`$ in place of the Dicke metric $`g_{{}_{\mathrm{D}}{}^{}\mu \nu }`$ in $`__\mathrm{D}`$ and $`T^{\mu \nu }`$ is the associated stress energy tensor as given by the usual formula (9). The consequences of the purely gravitational part of the coupling that arises in this way have already been treated elsewhere . The present work will be concerned just with the dilatonic part, which will evidently be specified by a coupling term that reduces to the simple form $`\widehat{T}\varphi `$ as presented in (4) with $`\widehat{T}`$ as given by (8) and (9). Not much can be said about the equations of motion for the internal fields characterising the material system until the form of its primary Lagrangian contribution $`\widehat{}`$ has been specified, but but it is evident quite generally that independently of such details, in an empty background the equation of motion for the axion field will be obtainable, using the gauge condition $$^\mu B_{\mu \nu }=0$$ (10) (the analogue of the Lorentz condition whose is familiar in the context of electromagnetism) in the well known Dalembertian form $$_\sigma ^\sigma B_{\mu \nu }=4\pi m__\mathrm{J}^{\mathrm{\hspace{0.17em}2}}\widehat{W}_{\mu \nu }.$$ (11) For the dilaton field no question of gauge arises at this stage: the relevant field equation is thus immediately obtainable in the simple form $$_\sigma ^\sigma \varphi =4\pi m__\mathrm{D}^2\widehat{T}.$$ (12) ## 3 Distributional Sources As in the more familiar case of point particle models, or “zero-branes”, the problem of ultraviolet divergences for higher dimensional “$`p`$-branes”, and in particular for string models, with $`p`$=1, because in these cases the relevant source densities $`\widehat{W}^{\mu \nu }`$, and $`\widehat{T}^{\mu \nu }`$ are not regular functions but Dirac type distributions that vanish outside the relevant one or two dimensional world sheets, except of course in the case of an ordinary medium, with the maximal space dimension, namely $`p=3`$, in an ordinary 4-dimensional spacetime background of the kind to which the peresent analysis is restricted. In the case of general $`p`$-brane, with local $`(p+1)`$-dimensional worldsheet imbedding given by $`x^\mu =\overline{x}{}_{}{}^{\mu }\{\sigma \}`$ in terms of intrinsic coordinates $`\sigma ^i`$ ($`i=0,1,\mathrm{},p`$), so that the induced surface metric will have the form $$\gamma _{ij}=g_{\mu \nu }\overline{x}{}_{}{}^{\mu }{}_{,i}{}^{}\overline{x}{}_{}{}^{\nu }{}_{,j}{}^{},$$ (13) the relevant source distributions will be expressible using the terminology of Dirac delta “functions” in the form exemplified in the case of the vorticity flux by $$\widehat{W}{}_{}{}^{\mu \nu }=g^{1/2}\overline{W}{}_{}{}^{\mu \nu }\delta _{}^{4}[x\overline{x}\{\sigma \}]\gamma ^{1/2}d^{p+1}\sigma ,$$ (14) where $`\gamma `$ is the determinant of the induced metric (13), and where the where the surface vorticity flux bivector $`\overline{W}^{\mu \nu }`$ is a regular antisymmetric tensor field on the worldsheet (but undefined off it). The analogous expression for the stress momentum energy source will be given by $$\widehat{T}{}_{}{}^{\mu \nu }=g^{1/2}\overline{T}{}_{}{}^{\mu \nu }\delta _{}^{4}[x\overline{x}\{\sigma \}]\gamma ^{1/2}d^{p+1}\sigma ,$$ (15) where the surface stress momentum energy density $`\overline{T}^{\mu \nu }`$, is a regular symmetric tensor field on the worldsheet (but undefined off it). The distributional nature of these source terms in cases for which $`p<3`$ is a corollary of the similarly distributional nature of the action density $`\widehat{}_{_{\mathrm{ma}}}`$ as defined, according to (3), to include both the purely internal contribution $`\widehat{}`$ and the cross coupling contribution $`\widehat{}_{_{\mathrm{co}}}`$. This distributional action density will itself be expressible in the form $$\widehat{}_{_{\mathrm{ma}}}=g^{1/2}\overline{}_{_{\mathrm{ma}}}\delta ^4[x\overline{x}\{\sigma \}]\gamma ^{1/2}d^{\mathrm{p}+1}\sigma ,$$ (16) with $$\overline{}_{_{\mathrm{ma}}}=\overline{}+\overline{}_{_{\mathrm{co}}}$$ (17) where the primary contribution or “master function” $`\overline{}`$ and the secondary coupling contribution $`\overline{}_{_{\mathrm{co}}}`$ are both well behaved scalar functions on the string worldsheet but undefined off it. Of these, the master function $`\overline{}`$ will be the intrinsic worldsheet Lagrangian, defined as a function just of the relevant internal, worldsheet confined, fields, such as currents, on the string, and of its induced metric, while the cross coupling contribution will be given in terms of the worldsheet confined fields $`\overline{W}^{\mu \nu }`$ and $`\overline{T}^{\mu \nu }`$ by $$\overline{}_{_{\mathrm{co}}}=\frac{_1}{^2}\overline{W}{}_{}{}^{\mu \nu }B_{\mu \nu }^{}+\overline{T}\varphi ,$$ (18) where $`\overline{T}=\overline{T}_\mu ^\mu `$. In terms of these well behaved worldsheet functions, the corresponding material contribution (2) to the action can be expressed directly, without recourse to heavy distributional machinery, as a simple (p+1)-surface integral in the form $$_{_{\mathrm{ma}}}=\overline{}_{_{\mathrm{ma}}}\gamma ^{1/2}d^{\mathrm{p}+1}\sigma .$$ (19) In particular, the regular surface stress energy tensor $`\overline{T}^{\mu \nu }`$ needed for the purpose of applying (18) is obtainable directly from the world sheet master function, without the use of distributions, using the formula $$\overline{T}{}_{}{}^{\mu \nu }=2\gamma ^{1/2}\frac{(\overline{}\gamma ^{1/2})}{g_{\mu \nu }}.$$ (20) As was remarked above, the errors in the early literature on this subject were largely attributable to failure to take proper account of the fact that unlike what occurs in the historically familiar special cases of electromagnetic and axionic coupling, for more general cases such as those of gravitational and dilatonic coupling the relevant coupling action contribution will be metric dependent. In order to allow for this, it is useful to work out the appropriately constructed hyper-Cauchy tensor (a relativistic generalisation of the Cauchy elasticity tensor of classical mechanics) which is defined by $$\overline{𝒞}{}_{}{}^{\mu \nu \rho \sigma }=\gamma ^{1/2}\frac{}{g_{\mu \nu }}\left(\overline{T}{}_{}{}^{\rho \sigma }\gamma ^{1/2}\right),$$ (21) or equivalently, in manifestly symmetric form by $$\overline{𝒞}{}_{}{}^{\mu \nu \rho \sigma }=\overline{𝒞}{}_{}{}^{\rho \sigma \mu \nu }=2\gamma ^{1/2}\frac{^2(\overline{}\gamma ^{1/2})}{g_{\mu \nu }g_{\rho \sigma }}.$$ (22) ## 4 Brane worldsheet geometry If the relevant radiation fields $`B_{\mu \nu }`$, and $`\varphi `$ are considered to be regular background fields attributable to external souces, the treatment of a brane system of the kind described in the preceeding section will be straightforeward, but it is evident that this will not be the case for the radiation fields produced by the brane itself, since they will be singular just where their evaluation is needed. Even for the treatment just of the regular case in which the relevant radiation fields are of purely external origin, and a fortiori for the treatment of the more delicate problem of self interaction, it is desirable – before proceeding to the derivation of the dynamical equations that ensue from the action (19) to recapitulate the essential geometric concepts that are needed for the kinematic description of the evolving worldsheet. (The unavailability of this machinery at the time of the pioneering work on the Goto-Nambu case is one of the reasons for the use of the misleading shortcut methods responsible for the confusion that beset this subjet before the recent clarification ). Point particle kinematics can conveniently be developed in terms of the timelike worldline tangent vector $`u^\mu `$ that is uniquely fixed by the condition of being future directed with unit normalisation, $`u^\mu u_\nu =1`$, and of the associated acceleration vector that is given in terms of covariant differentiation with respect to the (flat or curved) spacetime background metric $`g_{\mu \nu }`$ by $`a^\mu =u^\nu _\nu u^\mu `$. For higher brane cases, and in particular for the strings with which we shall be concerned here, a less specialised kinematic description must be used. Instead of the unique tangent vector $`u^\mu `$ and the derived vector $`a_\mu `$ that suffice for the “zero brane” case, the kinematic behaviour of higher “branes” starting with the case of strings (i.e. “one branes”) is most conveniently describable in terms of its first and second fundamental tensors. The former is definable as tangential projection tensor $`\eta _\nu ^\mu `$ say, which obtained by index lowering from the spacetime background projection of the inverse of the induced metric as given by the formula $$\eta ^{\mu \nu }=\gamma ^{ij}\overline{x}{}_{}{}^{\mu }{}_{,i}{}^{}\overline{x}{}_{}{}^{\mu }{}_{,j}{}^{},$$ (23) This first fundamental tensor can conveniently be used to rewrite the expressions (20) and (21) in the more practical forms $$\overline{T}{}_{}{}^{\mu \nu }=2\frac{\overline{}}{g_{\mu \nu }}+\overline{\mathrm{\Lambda }}\eta ^{\mu \nu }.$$ (24) and $$\overline{𝒞}{}_{}{}^{\mu \nu \rho \sigma }=\frac{\overline{T}^{\rho \sigma }}{g_{\mu \nu }}+\frac{_1}{^2}\overline{T}{}_{}{}^{\rho \sigma }\eta _{}^{\mu \nu },$$ (25) The corresponding second fundamental tensor $`K_{\mu \nu }{}_{}{}^{\rho }=K_{\nu \mu }^\rho `$, is obtained from the first fundamental tensor using the tangentially projected differentiation operator $$\overline{}_\mu =\eta _\mu ^\nu _\nu ,$$ (26) according to the prescription $$K_{\mu \nu }{}_{}{}^{\rho }=\eta _\nu ^\sigma \overline{}_\mu \eta _\sigma ^\rho .$$ (27) The condition of integrability of the worldsheet is the Weingarten identity, to the effect that this second fundamental tensor should be symmetric on its first two indices, i.e. $$K_{[\mu \nu ]}{}_{}{}^{\rho }=0.$$ (28) This tensor has the noteworthy property of being worldsheet orthogonal on its last index, but tangential on its (by the Weingarten identity interchangable) first pair of indices, $$K_{\mu \nu }{}_{}{}^{\sigma }\eta _{\sigma }^{}{}_{}{}^{\rho }=0=_\mu ^\lambda K_{\lambda \nu }{}_{}{}^{\rho }=0,$$ (29) using the notation $$_\nu ^\mu =g^\mu {}_{\nu }{}^{}\eta ^\mu _\nu $$ (30) for the tensor of projection orthogonal to the worldsheet. Whereas the full second fundamental tensor is needed for dealing with gravitational coupling , the treatment of the simpler cases considered here requires only its trace, namely the curvature vector $$K^\rho =K_\mu {}_{}{}^{\mu \rho }=\overline{}_\nu \eta ^{\nu \rho },$$ (31) which must evidently be worldsheet orthogonal, i.e. $$\eta ^\rho {}_{\sigma }{}^{}K_{}^{\sigma }=0.$$ (32) In terms of the background Riemann Christoffel connection $`\mathrm{\Gamma }_{\mu \rho }^\nu =g^{\nu \sigma }\left(g_{\sigma (\mu ,\rho )}\frac{_1}{^2}g_{\mu \rho ,\sigma }\right)`$ this curvature vector will be expressible in explicit detail as $$K^\nu =\frac{1}{\sqrt{\gamma }}\left(\sqrt{\gamma }\gamma ^{ij}\overline{x}_{,i}^\nu \right){}_{,j}{}^{}+\gamma ^{ij}\overline{x}_{,i}^\mu \overline{x}_{,j}^\rho \mathrm{\Gamma }_{\mu \rho }^\nu .$$ (33) In the particulary simple case a Dirac membrane or a Nambu Goto string (i.e. one for which the master function $`\overline{}`$ is just a constant) that is free, in the sense that it is not subjected to any external force, the “on shell” configurations (i.e. the solutions of the variational dynamical equations) will simply be characterised by the condition that the vector (33) vanishes, $`K^\mu =0`$, but this simple vanishing curvature condition will not be satisfied for more general models (such as those needed for superconducting strings ) nor when dilatonic and axionic forces are involved as in the cases analysed in the present work. It is to be remarked that the orthogonality property (32) of the curvature vector is to be contrasted with the tangentiality property of the stress energy tensor, $$_\mu ^\lambda \overline{T}{}_{}{}^{\mu \nu }=0,$$ (34) and of the hyper-Cauchy tensor, $$_\mu ^\lambda \overline{𝒞}{}_{}{}^{\mu \nu \rho \sigma }=0.$$ (35) Subject to the requirement that the worldsheet supported field $`\overline{W}^{\mu \nu }`$ be constructed from internal worldsheet fields in such a way as to aquire the corresponding tangentiality property $$_\mu ^\lambda \overline{W}{}_{}{}^{\mu \nu }=0,$$ (36) it can be seen that the corresponding distributional conservation law (7) can be equivalently expressed in terms of tangentially projected differentiation as the regular worldsheet flux conservation law $$\overline{}_\mu \overline{W}{}_{}{}^{\mu \nu }=0.$$ (37) ## 5 The force density formula For the purpose of the deriving the equations of motion of the material system from the variation principle, the most general variations to be considered are perturbations of the relevant internal fields, which have not yet been specified, and infinitesimal displacements with respect to the background characterised by the metric $`g_{\mu \nu }`$ and the the linearly coupled axionic and dilatonic fields. The effect of displacements can be conveniently analysed using a Lagrangian treatment in which not just the internal coordinates $`\sigma ^i`$ but also the background coordinates $`x^\mu `$ are considered to be dragged along by the displacement, so that the relevant field variations are given just by the corresponding Lie derivatives with respect to the displacement vector field $`\xi ^\mu `$ under consideration. This leads to the formulae $$\delta B_{\mu \nu }=\xi ^\sigma _\sigma B_{\mu \nu }2B_{\sigma [\mu }_{\nu ]}\xi ^\sigma $$ (38) for the axionic field, and $$\delta \varphi =\xi ^\sigma _\sigma \varphi $$ (39) for the dilatonic field, while finally for the background metric itself one has the well known formula $$\delta g_{\mu \nu }=2_{(\mu }\xi _{\nu )}.$$ (40) The postulate that the internal field equations are satisfied means that perturbations of the relevant internal fields have no effect on the action integral $`_{_{\mathrm{ma}}}`$, with the implication that for the purpose of evaluating the variation $`\delta _{_{\mathrm{ma}}}`$ there will be no loss of generality taking the variations of these so far unspecified internal fields simply to be zero. This means that the only contribution from the first term in (17) will the one provided by the background metric variation, for which we obtain the familiar formula $$\delta \left(\gamma ^{1/2}\overline{}\right)=\frac{_1}{^2}\gamma ^{1/2}\overline{T}{}_{}{}^{\mu \nu }\delta g_{\mu \nu }.$$ (41) The worldsheet flux conservation law (37) is interpretable as meaning that $`\overline{W}^{\mu \nu }`$ is related by Hodge type duality to the exterior derivatives of corresponding worldsheet differential forms (that will generically be of order p-2 respectively). In all the usual applications (including the continuuous medium example mentionned in the preceeding section and the string case developed below) these differential forms will be included among (or depend only on) the relevant independently variable internal fields whose variation can be taken to be zero for the purpose of evaluationg $`\delta _{_{\mathrm{ma}}}`$ when the internal field equations are satisfied. This means that the variation of the bivector surface density will also vanish, i.e. we shall have $$\delta (\gamma ^{1/2}\overline{W}{}_{}{}^{\mu \nu })=0.$$ (42) It follows that the axionic contribution from (18) to the variation of the integrand in (19) will be given simply by $$\delta \left(\gamma ^{1/2}\left(\frac{_1}{^2}\overline{W}{}_{}{}^{\mu \nu }B_{\mu \nu }^{}\right)\right)=\gamma ^{1/2}\left(\frac{_1}{^2}\overline{W}{}_{}{}^{\mu \nu }\delta B_{\mu \nu }\right).$$ (43) The systematic absence of any contribution from the background metric variation $`\delta g_{\mu \nu }`$ to action variation terms such as this, not only in the axionic case considered here but also in its more widely familiar analogue, sets a potentially misleading precedent that encourages a dangerous tendency (one of the main sources of error in earlier work ) to forget to check the possibility of metric variations in more general contexts. Although it does not contribute to the axionic term (43) allowance for the background metric variation (40) turns out to be of paramount importance not only in the gravitational case but also for the evaluation of the dilatonic contribution with which we are concerned here. It can be seen from (25) that we shall have $$\delta (\gamma ^{1/2}\overline{T})=\gamma ^{1/2}(\overline{T}{}_{}{}^{\mu \nu }+\overline{𝒞}{}_{}{}^{\mu \nu })\delta g_{\mu \nu },$$ (44) using the notation $$\overline{𝒞}{}_{}{}^{\mu \nu }=\overline{𝒞}{}_{}{}^{\mu \nu \rho }{}_{\rho }{}^{},$$ (45) for the trace of the hyper Cauchy tensor. Thus despite its deceptively simple scalar nature, the dilatonic coupling gives rise to a corresponding contribution that works out to be given by an expression of the not quite trivially obvious form $$\delta (\gamma ^{1/2}\overline{T}\varphi )=\gamma ^{1/2}(\overline{T}\delta \varphi +(\overline{T}{}_{}{}^{\mu \nu }+\overline{𝒞}{}_{}{}^{\mu \nu })\varphi \delta g_{\mu \nu }).$$ (46) To evaluate the integrated effect of the contributions (41), (43), and (46) the next step is the routine procedure of substitution of the relevant Lie derivative formulae (38), (39), and (40) followed by absorbtion of the terms involving derivatives of the displacement fields into pure worldsheet current divergences. For the primary contribution given by (41) one thereby obtains an expression of the familiar form $$\frac{_1}{^2}\overline{T}{}_{}{}^{\mu \nu }\delta g_{\mu \nu }=\xi ^\mu \overline{}_\nu \overline{T}{}_{}{}^{\nu }{}_{\mu }{}^{}+\overline{}_\mu (\xi ^\nu \overline{T}{}_{}{}^{\mu }{}_{\nu }{}^{}).$$ (47) while the corresponding expression for axionic coupling contribution (43) will simply be given by $$\frac{_1}{^2}\overline{W}{}_{}{}^{\mu \nu }\delta B_{\mu \nu }=\frac{_1}{^2}\xi ^\mu N_{\mu \nu \rho }\overline{W}{}_{}{}^{\nu \rho }+\overline{}_\mu (\xi ^\nu B_{\nu \rho }\overline{W}{}_{}{}^{\mu \rho }).$$ (48) However the dilatonic coupling contribution (46) is not so simple: in addition to the obvious scalar field variation contribution given by $$\overline{T}\delta \varphi =\xi ^\mu \overline{T}_\mu \varphi .$$ (49) there will be another less obvious contribution (the one that tended to be overlooked in earlier work) given by $$(\overline{T}{}_{}{}^{\mu \nu }+\overline{𝒞}{}_{}{}^{\mu \nu })\varphi \delta g_{\mu \nu }=\xi ^\mu \overline{}_\nu \left(2(\overline{T}{}_{}{}^{\nu }{}_{\mu }{}^{}+\overline{𝒞}{}_{}{}^{\nu }{}_{\mu }{}^{})\varphi \right)+\overline{}_\mu \left(2\xi ^\nu (\overline{T}{}_{}{}^{\mu }{}_{\nu }{}^{}+\overline{𝒞}{}_{}{}^{\mu }{}_{\nu }{}^{})\varphi \right).$$ (50) When these expressions are used to evaluate the variation of the action integral (16) due to a displacement confined to a bounded region of the worldsheet, the application of the relevant (p+1) dimensional version of Green’s theorem removes the contributions from the divergence terms, so that one is left with an expression of the standard form $$\delta _{_{\mathrm{ma}}}=\xi ^\mu (\overline{f}_\mu \overline{}_\nu \overline{T}{}_{}{}^{\nu }{}_{\mu }{}^{})\gamma ^{1/2}d^{p+1}\sigma ,$$ (51) so that the application of the variation principle gives the corresponding dynamical equation $$\overline{}_\nu \overline{T}{}_{}{}^{\mu \nu }=\overline{f}^\mu $$ (52) in which the vector $`\overline{f}^\mu `$ represents the total force density exerted by the various radiation fields involved. Using the foregoing expressions, this force density can immediately be read out in the form $$\overline{f}{}_{}{}^{\mu }=\overline{f}__\mathrm{J}{}_{}{}^{\mu }+\overline{f}__\mathrm{D}{}_{}{}^{\mu },$$ (53) in which the axionic contribution can be seen from (48) to be given by the well known formula (the axionic analogue of the Lorentz force formula in electromagnetism) given by $$\overline{f}__\mathrm{J}{}_{}{}^{\mu }=\frac{_1}{^2}N^\mu {}_{\nu \rho }{}^{}\overline{W}{}_{}{}^{\nu \rho },$$ (54) which seemed unfamiliar when first derived in the present context, but which is in fact interpretable just as the immediate relativistic generalisation of the historic formula on which the theory of flying is based, namely the Joukowski force law for the lift (due to the Magnus effect) on a long thin (i.e. string-like) aeroplane wing. What is not so well known is the corresponding formula for the dilatonic force density, which can be seen from (49) and (50) to be given by $$\overline{f}__\mathrm{D}{}_{}{}^{\mu }=\overline{T}^\mu \varphi \overline{}_\nu \left(2(\overline{T}{}_{}{}^{\mu \nu }+\overline{𝒞}{}_{}{}^{\mu \nu })\varphi \right).$$ (55) ### 5.1 The force on a Nambu-Goto string. The proceeding formulae apply to domain wall type membrane models (with $`p=2`$) as well as to simple point particle models (with $`p=0`$), but from this stage onwards we shall restrict our attention to the case of string models, as characterised by $`p=1`$. Before further restricting attention to the very special case of Nambu-Goto type string models, it is worthwhile to recapitulate some relevant features that are shared by more general string models, including the kind needed for describing the effects of the type of supercontivity whose likely occurrence in cosmic strings was originally predicted by Witten . In the higher dimensional branes the vorticity flux $`\overline{W}^{\mu \nu }`$ might depend on internal field variables of the model, while for a point particle model no such source can exist at all. In the intermediate case of a string there is no obstruction to the existence of a vorticity flux but it cannot depend on any internal field variables of the model: the only way the conservation law (37) can be satisfied on a 2-dimensional worldsheet is for the vorticity flux to have the form $$\overline{W}^{\mu \nu }=\overline{\kappa }^{\mu \nu },$$ (56) where $`^{\mu \nu }`$ is the antisymmetric unit surface element tensor and $`\overline{\kappa }`$ is a constant that is interpretable as representing the momentum circulation around the string of the Zel’dovich type fluid representing the axion field – which means that it will be an integral multiple of Planck’s constant, i.e. an integral multiple of $`2\pi `$ in the unit system we are using with $`\mathrm{}`$ set to unity. Using the traditional dot and dash notation $`\dot{x}^\mu =\overline{x}_{,_0}^\mu `$ and $`x{}_{}{}^{\mu }=\overline{x}_{,_1}^\mu `$ for the effect of partial differentiation with respect to worldsheet coordinates $`\sigma ^_0`$ and $`\sigma ^_1`$ the antisymmetric unit surface element tensor will be given by $$^{\mu \nu }=2\left(\gamma \right)^{1/2}x_{,_0}^{[\mu }x_{,_1}^{\nu ]},$$ (57) In the case of a string, the fundamental tensor will be given in terms of this unit tangent bivector by $$\eta _\nu ^\mu =_\rho ^\mu _\nu ^\rho .$$ (58) One of the reasons why the kind tensorial analysis used here was not developed much earlier for the purpose of application to string dynamics is that the heavy algebra involved in the use of coordinate dependent expressions such as that on the right hand side of the curvature formula (33) could be alleviated to some extent by the use of specialised internal coordinate systems of the conformal type characterised by the conditions $$\dot{x}^\mu x_\mu ^{}=0\dot{x}^\mu \dot{x}_\mu +x^\mu x_\mu ^{}=0,$$ (59) which imply that the relation $$\gamma ^{1/2}=x^\mu x_\mu ^{}=\dot{x}^\mu \dot{x}_\mu ,$$ (60) If $`\sigma ^_0`$ and $`\sigma ^_1`$ are restricted to satisfy these conditions (which are frequently incompatible with other compelling desiderata) then the unweildy formula (33) for the curvature vector that plays such an important role will be replaceable by the handier expression $$K^\nu =\gamma ^{1/2}\left(x^{\prime \prime \nu }\ddot{x}^\nu +(x^\mu x^\rho \dot{x}^\mu \dot{x}^\rho )\mathrm{\Gamma }_{\mu \rho }^\nu \right).$$ (61) If the background metric $`g_{\mu \nu }`$ is not just empty but actually flat then this formula will be further simplifiable by elimination of the final term if one is willing to restrict the background cordinates to be of Minkowski type, but of couse such a restriction may not be what is most convenient for exploiting symmetries, such as occur in circular string loop configurations for which spherical or cylindrical coordinates might be preferable. The development of string dynamics has been unnecessarily delayed by over reliance on the special gauge characterised by Minkowski coordinates on the background and conformal coordinates on the worldsheet, rather that using the kind of geometrical approach followed here, which provides more elegant and concise formulae for general purposes. This more powerful geometric approach is of course particularly advantageous for applications in which for various technical reasons the usual (conformal cum Minkowski) kind of gauge may be unsuitable. From this point on, attention will be restricted to the special simple case of string models of Nambu-Goto type, which includes the case that arises in the low energy limit of string theory considered by Dabholkar and Harvey . Such models are characterised by the condition that the relevant master function $`\overline{}`$ is simply a constant, which means that it will be expressible in the form $$\overline{}=m__\mathrm{K}^{\mathrm{\hspace{0.17em}2}},$$ (62) where $`m__\mathrm{K}`$ is a fixed mass scale that will be referred to as the Kibble mass to distinguish it from other mass scales in the theory. In the context of cosmic string theory it is generally expected that it should be of the same order of magnitude as the Higgs mass, $`m__\mathrm{X}`$ say, that is associated with the underlying “spontaneously broken” symmetry of the vacuum. In the context of superstring theory the quantity $`m__\mathrm{K}`$ is usually supposed to be of the order of magnitude of the Planck mass $`m__\mathrm{P}`$. The other parameters needed to complete the specification of the theory are the quantities $`m__\mathrm{J}`$, $`m__\mathrm{D}`$ and $`\overline{\kappa }`$ that have already been introduced. To match the present formulation to the equivalent low energy linearised limit theory in the slightly different notation used by Buonanno and Damour their parameters $`\alpha `$, $`\lambda `$, $`\mu `$ are identifiable as being given by the relations $`\alpha =m__\mathrm{P}/m__\mathrm{D}`$, $`\lambda =\overline{\kappa }m__\mathrm{P}m__\mathrm{J}/2`$ and $`\mu =m__\mathrm{K}^{\mathrm{\hspace{0.17em}2}}`$. The special values corresponding to the low energy superstring theory limit considered by Dabholkar and Harvey are given by $`\alpha =1`$, $`\lambda =\mu `$, which in the formulation used here is equivalent to the conditions $$m__\mathrm{D}=m__\mathrm{P}\mathrm{\hspace{0.17em}2}m__\mathrm{K}^2=\overline{\kappa }m__\mathrm{J}m__\mathrm{P}.$$ (63) Whether or not the particular conditions (63) are satisfied (and of course quite independently of whether the internal coordinate gauge satisfies the conditions (59) on which the specialised formulae (60) and (61) depend) the surface stress momentum energy tensor of the string can be seen from (24) to be proportional to the fundamental tensor, according to the formula $$\overline{T}{}_{}{}^{\mu \nu }=m__\mathrm{K}^{\mathrm{\hspace{0.17em}2}}\eta ^{\mu \nu },$$ (64) and so its trace will be given by $$\overline{T}=2m__\mathrm{K}^{\mathrm{\hspace{0.17em}2}}.$$ (65) The corresponding the hyper-Cauchy tensor is obtainable from (25) in the form $$\overline{𝒞}{}_{}{}^{\mu \nu \rho \sigma }=m__\mathrm{K}^{\mathrm{\hspace{0.17em}2}}(\eta ^{\mu (\rho }\eta ^{\sigma )\nu }\frac{1}{2}\eta ^{\mu \nu }\eta ^{\rho \sigma }).$$ (66) It can be seen from this that in this special Nambu-Goto case the trace tensor that appears in the dilatonic force formula (55) will vanish, i.e. one obtains $$\overline{𝒞}{}_{}{}^{\mu \nu }=0.$$ (67) It is to be emphasised that that this simplification is a special feature of the string case, and that it does not hold in the higher dimensional case of a Dirac membrane, nor even in the trivial lower dimensional case of a point particle with mass $`m`$ and unit 4-velocity vector $`u^\mu `$, for which one obtains $`\overline{𝒞}{}_{}{}^{\mu \nu }=mu^\mu u^\nu /2`$. It follows from (54) and (56) that the Joukowsky force density exerted by the axionic fluid on a string (of any kind) will be given by $$\overline{f}__\mathrm{J}{}_{}{}^{\mu }=\frac{_1}{^2}\overline{\kappa }N^\mu {}_{\nu \rho }{}^{}{}_{}{}^{\nu \rho }.$$ (68) It follows from (55) using (64) and (67) that in the case of a Nambu Goto string the corresponding dilatonic force density contribution will be obtainable – with the aid of the defining formula (31) for the curvature vector $`K^\mu `$ – in the form $$\overline{f}__\mathrm{D}{}_{}{}^{\mu }=2m__\mathrm{K}^{\mathrm{\hspace{0.17em}2}}(\varphi K^\mu ^{\mu \nu }_\nu \varphi ).$$ (69) Simple though it is, this formula does not seem to have been previously made available in the literature. It is to be observed that – as needed to avoid overdetermination in the Goto-Nambu case – the force contributions (68) and (69) are both identically orthogonal to the string worldsheet. It is evident that if the dilatonic field were due only to high frequency radiation from an external source then the first term on the right in (69) would be relatively negligible, but it will be shown below that this term is not at all negligible for a dilatonic field due to self interaction. ## 6 Allowance for regularised self interaction In cases where self interaction is involved, it is commonly convenient to decompose the relevant linear interaction fields – which in the present instance are $`B_{\mu \nu }`$ and $`\varphi `$ – into a short range contribution determined via the relevant Green function by the immediately neighbouring source distribution, and a longer range residual contribution that includes allowance for incoming radiation from external sources. More particularly, in the present instance, it will be useful to consider the relevant fields $`B_{\mu \nu }`$ and $`\varphi `$ the sums of short range contributions that will be indicated by a widehat and long range parts that will be indicated by a widetilde in the form $$B_{\mu \nu }=\stackrel{~}{B}_{\mu \nu }+\widehat{B}_{\mu \nu },\varphi =\stackrel{~}{\varphi }+\widehat{\varphi },$$ (70) This will evidently give rise to corresponding decompositions $$f__\mathrm{J}^\mu =\stackrel{~}{f}__\mathrm{J}^\mu +\widehat{f}__\mathrm{J}^\mu ,f__\mathrm{D}^\mu =\stackrel{~}{f}__\mathrm{D}^\mu +\widehat{f}__\mathrm{D}^\mu $$ (71) for the asociated force densities as specified by the general formulae (54), (55) or their Nambu Goto string specialisations (68), (69). In many contexts the coupling is so weak that the local self force contributions $`\widehat{f}__\mathrm{J}^\mu `$ and $`\widehat{f}__\mathrm{D}^\mu `$ can be neglected. However in cases for which one needs to take account of the self induced contributions $`\widehat{B}_{\mu \nu }`$ and $`\widehat{\varphi }`$, there will be difficulties arising from the fact that the relevant source fields on the right hand sides of the field equations (12) and (12) will not be the regular worldsheet supported tensor fields $`\overline{W}^{\mu \nu }`$, and $`\overline{T}`$, but the corresponding 4-dimensionally supported distributions, $`\widehat{W}^{\mu \nu }`$, and $`\widehat{T}`$ as constructed according to the prescriptions (14) and (15). For sources such as these, the resulting field contributions will diverge in the thin worldsheet limit. As in the familiar point particle case, so also for a string, one can obtain an appropriately regularised result by supposing that (as will be entirely realistic in cases such as that of a cosmic string model for a vortex defect of the vacuum) the underlying that the physical system one wishes to describe is not quite infinitely thin but actually has finite spacial extent that can be used to specify an appropriately microscopic “ultraviolet” cut off length scale $`\delta _{}`$ say. This will be sufficient for regularisation in the point particle case, but in the string case it will also be necessary to introduce a long range “infrared” cut off length scale $`\mathrm{\Delta }`$ say that might represent the macroscopic mean distance between neigbouring strings. In the case of a string an an ordinary 4-dimensional spacetime background, it can be seen (following the example of the electromagnetic prototype considered by Witten in his original discussion of “superconducting” cosmic strings) that the dominant contribution to relevant Green function integrals in the ultraviolet limit will then be proportional to a logarithmic regularisation factor of the form $$\widehat{l}=\mathrm{ln}\left\{\mathrm{\Delta }^2/\delta _{}^{\mathrm{\hspace{0.17em}2}}\right\}.$$ (72) More specifically, the dominant contribution to the axionic self field arising from the Dalembertian source equation (11) will be given by $$\widehat{B}_{\mu \nu }=\widehat{l}m__\mathrm{J}^{\mathrm{\hspace{0.17em}2}}\overline{W}_{\mu \nu },$$ (73) with $`\overline{W}_{\mu \nu }`$ given by (56), while similarly the dominant contribution to the dilatonic self field arising from the Dalembertian source equation (12) will be given by $$\widehat{\varphi }=\widehat{l}m__\mathrm{D}^2\overline{T}.$$ (74) (If the microscopic axial current source distribution were very different from that of the stress energy trace source for the dilatonic field, the natural cut off $`\delta _{}`$ that would be most appropriate for the latter might differ somewhat from what would apply to the former, but the effect of such a difference could be considered as a higher order correction that need not be taken into account so long as we are only concerned with the dominant contribution.) For the purposes of substitution in the force formulae – (54), (55) or their Nambu Goto string specialisations (68), (69) – knowledge just of the regularised self fields $`\widehat{B}_{\mu \nu }`$ and $`\widehat{\varphi }`$ is not immediately sufficient. The problem is that these regularised values are well defined only on the worldsheet, which means that they do not directly provide what is needed for a direct evaluation of the gradients that are required: there is no difficulty for the terms involving just the tangentially projected gradient operator $`\overline{}_\nu `$ but it can be seen that there are also contributions from the unprojected gradient operator $`_\nu `$ which is directly meaningfull only when acting fields whose support extends off the worldsheet. It fortunately turns out that this problem has a very simple general solution, of which particular applications in particular gauges are implicit in much previous work and which I formulated explicitly in conveniently covariant and more generally utilisable form in the specific context of the electromagnetic case . What one finds – by examining the string worldsheet limit behaviour of derivatives of the relevant Dalembertian Green function – is that the appropriate regularisation of the gradients on the string worldsheet is obtainable simply by replacing the ill defined operator $`_\nu `$ by by the corresponding regularised gradient operator $$\widehat{}_\nu =\overline{}_\nu +\frac{_1}{^2}K_\nu ,$$ (75) where $`K^\mu `$ is the worldsheet curvature vector that is defined by the formula (31) and that is expressible, if one is willing to allow oneself to be restricted to the use of conformal worldsheet coordinates, by a more detailed prescription of the form (61). In the explicit application of the formula (75), it is sufficient, in the case of a scalar field $`\phi `$, to use the simple expression $`\overline{}{}_{}{}^{\mu }\phi =\gamma ^{ij}\overline{x}^\mu {}_{,i}{}^{}\phi _{,j}^{}`$ for the tangentially projected gradient, but for a tensorial field there will also be contributions depending on the background Riemann Christoffel connection $`\mathrm{\Gamma }_{\mu \rho }^\nu `$, which is also involved in the detailed expressions (33) and (61), unless one is using Minkowski coordinates in a flat spacetime backckround. Applying this procedure to the axionic Kalb Ramond 2-form, one sees from (6) and (73) that the corresponding regularised local current 3-form contribution will be given by $$\widehat{J}_{\mu \nu \rho }=3\widehat{}_{[\mu }\widehat{B}_{\nu \rho ]}=\frac{_1}{^2}\widehat{l}\left(6\overline{}_{[\mu }\overline{W}_{\nu \rho ]}+3K_{[\mu }\overline{W}_{\nu \rho ]}\right).$$ (76) Taking account of the surface flux conservation law (37), and using the defining relation (31) for the curvature vector $`K^\mu `$, it can be seen that the axionic force density (54) will be expressible as a world sheet divergence in the form $$\widehat{f}__\mathrm{J}^\mu =\overline{}_\nu \widehat{T}__\mathrm{J}{}_{}{}^{\mu \nu },$$ (77) in terms of a regularised local axionic stress energy tensor given by $$\widehat{T}__\mathrm{J}{}_{}{}^{\mu \nu }=\widehat{B}{}_{}{}^{\mu }{}_{\rho }{}^{}\overline{W}{}_{}{}^{\nu \rho }\frac{_1}{^4}\widehat{B}_{\rho \sigma }\overline{W}{}_{}{}^{\rho \sigma }\eta _{}^{\mu \nu },$$ (78) which can be seen from (56) and (73) to reduce with the aid of (58) to the simple explicit form $$\widehat{T}__\mathrm{J}{}_{}{}^{\mu \nu }=\frac{_1}{^2}\widehat{l}\overline{\kappa }^2m__\mathrm{J}^{\mathrm{\hspace{0.17em}2}}\eta ^{\mu \nu },$$ (79) which is evidently isotropic with respect to the 2-dimensional worldsheet geometry, like the intrinsic stress energy tensor in the Nambu-Goto case When one applies same procedure to the dilatonic self force contribution in (55) one finds that it too can be formulated as a world sheet divergence in the analogous form $$\widehat{f}__\mathrm{D}^\mu =\overline{}_\nu \widehat{T}__\mathrm{D}{}_{}{}^{\mu \nu },$$ (80) in terms of a regularised local dilatonic stress energy tensor given by $$\widehat{T}__\mathrm{D}{}_{}{}^{\mu \nu }=\widehat{\varphi }(\overline{T}{}_{}{}^{\mu \nu }\frac{_1}{^4}\overline{T}\eta ^{\mu \nu }+\overline{𝒞}{}_{}{}^{\mu \nu }).$$ (81) More specificly, in the particular case of a Nambu Goto type string, as characterised by (64) and (67) it can be seen that this dilatonic stress energy contribution will reduce to the form $$\widehat{T}__\mathrm{D}{}_{}{}^{\mu \nu }=2\widehat{l}m__\mathrm{K}^{\mathrm{\hspace{0.17em}4}}m__\mathrm{D}^2\eta ^{\mu \nu }.$$ (82) Comparing (82) with (79) it can be seen that (in contradiction with previous assertions to the effect that it would vanish or even that it would augment the corresponding axionic contribution ) this dilatonic contribution must always be oppositely directed to the corresponding axionic contribution. More particularly it can be seen that these dominant local axionic and dilatonic self interaction contributions will exactly cancel each other out if the relevant coupling constants are related by the condition $$2m__\mathrm{K}^{\mathrm{\hspace{0.17em}2}}=\overline{\kappa }m__\mathrm{D}m__\mathrm{J},$$ (83) which will in fact be satisfied automatically in the special case (63) envisaged by Dabholkar and Harvey (whose faulty analysis lead to a condition that was somewhat different from (83) – but that also happenned to be satisfied in the special case (63) they were considering, and that thereby provided a spurious verification of their reasonning.) The mutual cancellation subject to (83) of the dominant axionic and dilatonic self interactions for a Nambu-Goto string in a 4-dimensional background has already been confirmed by Buonanno and Damour using an entirely different approach formulated in terms of an effective action. The fact that – as in the previously investigated electromagnetic and gravitational cases – the result of the present approach based on direct evaluation of the self force is in full agreement with the result of the approach based on the use of an effective action provides a check on the validity of the latter approach, whose credibility had previously been questionned . The complete consistency between the two kinds of approach has already been made clear for the electromagnetic and gravitational cases, and will be made clear for the axionic and dilatonic cases dilatonic in the next section, where the relevant effective action contributions will be explicitly derived. ## 7 Action renormalisation The fact that the dominant local force contributions are expressible as divergences of the form (77) and (80) is what makes it possible to describe the the result of this regularisation as a “renormalisation”: it implies that these self force contributions can be absorbed into the left hand side of the basic force balance equation by a renormalisation whereby the original “bare” stress momentum energy density tensor $`\overline{T}^{\mu \nu }`$ undergoes a replacement $`\overline{T}{}_{}{}^{\mu \nu }\stackrel{~}{T}^{\mu \nu }`$ in which the “dressed” stress momentum energy tensor is given by $$\stackrel{~}{T}{}_{}{}^{\mu \nu }=\overline{T}{}_{}{}^{\mu \nu }+\widehat{T}^{\mu \nu }$$ (84) with $$\widehat{T}{}_{}{}^{\mu \nu }=\widehat{T}__\mathrm{J}{}_{}{}^{\mu \nu }+\widehat{T}__\mathrm{D}{}_{}{}^{\mu \nu }.$$ (85) The basic force balance equation (52) can thereby be rewritten in the equivalent form $$\overline{}_\nu \stackrel{~}{T}{}_{}{}^{\mu \nu }=\stackrel{~}{f}{}_{}{}^{\mu },$$ (86) in which the force density on the right consists just of well behaved long range radiation contributions as given by the sum $$\stackrel{~}{f}{}_{}{}^{\mu }=\stackrel{~}{f}__\mathrm{J}{}_{}{}^{\mu }+\stackrel{~}{f}__\mathrm{D}{}_{}{}^{\mu },$$ (87) in which each of the terms is entirely regular. What will be shown in this final section is that the renormalised stress energy tensor $`\stackrel{~}{T}^{\mu \nu }`$ is derivable, by a prescription of the standard form (20), from a correspondingly remormalised action in which the original Lagrangian master function $`\overline{}_{_{\mathrm{ma}}}`$ is replaced by an appropriately renormalised master function $`\stackrel{~}{}_{_{\mathrm{ma}}}`$. In order to incorporate the effects of self interaction, as described by the renormalised force balance equation (87), it can be verified that all one needs to do is to replace $`_{_{\mathrm{ma}}}`$ by a corresponding renormalised action $$\stackrel{~}{}_{_{\mathrm{ma}}}=\stackrel{~}{}_{_{\mathrm{ma}}}\gamma ^{1/2}d^2\sigma ,$$ (88) with $$\stackrel{~}{}_{_{\mathrm{ma}}}=\stackrel{~}{}+\frac{_1}{^2}\stackrel{~}{B}_{\mu \nu }\overline{W}{}_{}{}^{\mu \nu }+\stackrel{~}{\varphi }\overline{T}.$$ (89) where the renormalised master function is given simply by $$\stackrel{~}{}=+\widehat{\mathrm{\Lambda }}__\mathrm{J}+\widehat{}__\mathrm{D},$$ (90) in which the axionic contribution will be given by $$\widehat{}__\mathrm{J}=\frac{_1}{^4}\widehat{B}_{\mu \nu }\overline{W}{}_{}{}^{\mu \nu },$$ (91) which works out simply to be a constant, $$\widehat{}__\mathrm{J}=\frac{_1}{^2}\widehat{l}\overline{\kappa }^2m__\mathrm{Z}^{\mathrm{\hspace{0.17em}2}}.$$ (92) in which it is to be recalled that $`\widehat{l}`$ is the logarithmic factor given by (72). The corresponding dilatonic contribution (which has not been evaluated for a general string model before) is obtained as $$\widehat{}__\mathrm{D}=\frac{_1}{^2}\widehat{\varphi }\overline{T}=\frac{_1}{^2}\widehat{l}m__\mathrm{D}^2\overline{T}^2.$$ (93) It is to be observed that in terms of the string energy density $`𝒰`$ and tension $`𝒯`$ (as conventionaly defined to be the eigenvalues of $`\overline{T}^{\mu \nu }`$) the trace in the proceeding formula will be given by $`\overline{T}=(𝒰+𝒯)`$, so the dilatonic self interaction contribution will be proportional to the $`(𝒰+𝒯)^2`$. This is to be contrasted with the case of the corresponding gravitational self interaction contribution which has been shown to be proportional to $`(𝒰𝒯)^2`$. For a Nambu-Goto string, as characterised by $`𝒰=𝒯=m__\mathrm{K}^{\mathrm{\hspace{0.17em}2}}`$ this gravitational contribution will simply vanish, while the dilatonic contribution (93) will just have the constant value given by $$\widehat{}__\mathrm{D}=2\widehat{l}m__\mathrm{K}^{\mathrm{\hspace{0.17em}4}}m__\mathrm{D}^2.$$ (94) The self interaction contributions (92) and (94) are in perfect agreement with those already obtained by the effective action approach developed by Buonanno and Damour – whose results were more general than those provided here in so much as they were not limited to a 4-dimensional background, though on the other hand they were more restricted in so much as they considered only strings of Nambu-Goto type. This completes the demonstration that, as has already been shown for electromagnetic and gravitational interactions (and contrary to what was previously alleged ) so also for the axionic and dilatonic contributions, the treatment of the divergent self interaction contribution by an approach developped directly in terms of effective action is entirely consistent with the treatment based on the detailed analysis of the corresponding force contributions, as given in the axionic case by (68) and in the dilatonic case by the general formula (55) or its Nambu-Goto specialisation (69). The discrepancies that arose in earlier work were due to the use of unreliable short cut methods (largely motivated by the unavailability of the more efficient methods of geometrical analysis that have since been developed) which lead to the omission of some of the (less easily calculable) terms in the relevant formulae. In particular the sign error in the original estimate of the net dilatonic contribution by Dabholkar and Harvey can be accounted for as being due to omission of the contribution provided by the first term on the right in (69). As was seen for the corresponding forces in the previous section, the axionic contribution (92) will evidently be exactly cancelled by the dilatonic contribution (94) when the special condition (83) is satisfied. Like the self cancellation of the gravitational contribution in the Nambu=Goto case , this mutual cancellation of the corresponding axionic and dilatonic contributions is a special feature of 4-dimensional space-time: it has been shown by the work of Buonanno and Damour that it does not carry over to Nambu-Goto strings in backgrounds of higher dimension. I wish to thank A. Buonanno, T. Damour, R. Battye, A. Gangui, G. Gibbons, P. Peter and P. Shellard for instructive conversations.
warning/0001/hep-ex0001019.html
ar5iv
text
# 1 Introduction ## 1 Introduction It might appear surprising to include a report from the Fermilab Tevatron, a proton-antiproton collider, in a session about “Brief Reports from the $`B`$ factories”. Does this mean the Tevatron would qualify as a $`B`$ factory? There are two advantages of studying $`B`$ physics at the Tevatron. First, all $`B`$ hadrons are produced; not only charged and neutral $`B`$ mesons as at the $`B`$ factories, but also $`B_\text{S}^0`$ mesons and $`b`$-baryons. The second advantage is the $`b`$ quark production cross section, which is about 1 nb at the $`\mathrm{{\rm Y}}(4S)`$ resonance while it is about 50 $`\mu `$b for $`p\overline{p}`$ collisions at $`\sqrt{s}=1.8`$ TeV. This is an enormous cross section which is about 50,000 times larger at the Tevatron than at the $`B`$ factories. It resulted in about $`5\times 10^9`$ $`b\overline{b}`$ quark pairs being produced during the 1992-1995 data taking period of the Tevatron, called Run I. To illustrate the enormous $`b`$ production rate at the Tevatron, we compare the yield of fully reconstructed $`B`$ mesons between the CLEO experiment and CDF. In a data sample of about 3000 pb<sup>-1</sup>, CLEO reconstructs about 200 $`B`$ mesons decaying into $`J/\psi K^+`$ while CDF finds in a sample of about 100 pb<sup>-1</sup> of data a signal of about 1000 $`J/\psi K^+`$ events with a good signal-to-background ratio . The goal of the $`B`$ factories is to discover $`CP`$ violation in $`B^0J/\psi K_S^0`$ decays. CDF has already presented a measurement of $`CP`$ violation in the $`B`$ meson system , measuring the time-dependent asymmetry in the yield of $`J/\psi K_S^0`$ events coming from a $`B^0`$ versus $`\overline{B}^0`$: $$𝒜_{CP}(t)\frac{N(\overline{B}{}_{}{}^{0}(t))N(B^0(t))}{N(\overline{B}{}_{}{}^{0}(t))+N(B^0(t))}=\mathrm{sin}2\beta \mathrm{sin}\mathrm{\Delta }m_dt.$$ (1) This asymmetry is directly related to the $`CP`$ violation parameter $`\mathrm{sin}2\beta `$. ### 1.1 CDF Measurement of $`\mathrm{𝐬𝐢𝐧}\mathrm{𝟐}𝜷`$ Here, we briefly summarize CDF’s initial measurement of $`\mathrm{sin}2\beta `$. Figure 1(a) shows the $`J/\psi K_S^0`$ yield at CDF, where $`395\pm 31`$ events have been identified. This is currently the world’s largest sample of fully reconstructed $`J/\psi K_S^0`$ events. Measuring a $`CP`$ asymmetry requires knowing whether the $`J/\psi K_S^0`$ originated from a $`B^0`$ or $`\overline{B}^0`$ meson. This is usually referred to as $`B`$ flavor tagging. Several methods of $`B`$ flavor tagging exist. Some of them exploit the other $`B`$ hadron in the event and search for a lepton from the semileptonic decay of the other $`B`$ hadron or determine the net charge of the jet produced by the other $`b`$ quark. These two methods are called lepton tagging and jet charge tagging, respectively. The $`B`$ flavor can also be determined by searching for pions which are produced through fragmentation or $`B^{}`$ mesons in correlation with the $`B`$ meson of interest. This method is known as same side tagging. $`B`$ flavor tagging is the crucial element for a $`CP`$ violation measurement at the Tevatron. The figure of merit quantifying how well a flavor tagging algorithm works is the so-called effective tagging efficiency $`\epsilon 𝒟^2`$. Here, $`\epsilon `$ is the efficiency for obtaining a particular flavor tag, and $`𝒟`$ is the dilution defined by the number of right tags ($`N_R`$) and the number of wrong tags ($`N_W`$): $`𝒟=(N_RN_W)/(N_R+N_W)`$. CDF determined the tagging power of various tagging methods with data measuring the time dependence of $`B^0\overline{B}^0`$ flavor oscillations. Such a measurement serves as a demonstration that a particular flavor tag does work in a hadron collider environment and determines its $`\epsilon 𝒟^2`$. Figure 2(a) shows the measured mixing asymmetries as a function of proper decay length using a same side tag . From this measurement, CDF extracts $`\mathrm{\Delta }m_d=(0.471_{0.068}^{+0.078}\pm 0.034)\mathrm{ps}^1`$ and the effective tagging efficiency for the same side tag to be $`\epsilon 𝒟^2=(1.8\pm 0.4\pm 0.3)\%`$. As another example, Fig. 2(b) shows the fraction of mixed events as a function of proper decay length using a jet charge and lepton flavor tag . This measurement yields $`\mathrm{\Delta }m_d=(0.500\pm 0.052\pm 0.043)\mathrm{ps}^1`$ as well as $`\epsilon 𝒟^2=(0.91\pm 0.10\pm 0.11)\%`$ and $`\epsilon 𝒟^2=(0.78\pm 0.12\pm 0.08)\%`$ for a lepton tag and jet charge tag, respectively. Applying all three tagging methods to CDF’s sample of $`J/\psi K_S^0`$ events gives the $`CP`$ asymmetry distribution shown in Fig. 1(b). The data prefer a positive asymmetry, resulting in a measurement of $`\mathrm{sin}2\beta =0.79\pm 0.39\pm 0.16`$. This can be translated into a limit on $`\mathrm{sin}2\beta `$ being positive ($`0<\mathrm{sin}2\beta <1`$) at 93% confidence level. This is the best direct measurement of $`CP`$ violation in the $`B`$ system to date. With this result, CDF demonstrated that a $`CP`$ violation measurement is feasible at the Tevatron in Run II. Returning to our initial question, we think the Tevatron does qualify as a $`B`$ factory capable of measuring $`CP`$ violation. ## 2 The Tevatron in Run II The most important element of the Tevatron upgrade for Run II is the Main Injector. It is a new 150 GeV accelerator, half the circumference of the Tevatron, which will increase the antiproton intensity into the Tevatron, providing 20 times higher luminosities. The Main Injector project was finished in June 1999 with the first beam circulating at that time. Run II of the Tevatron had originally been defined as 2 fb<sup>-1</sup> being delivered in two years to the collider experiments. Run II has recently been extended beyond the initial two years, to continue until 2006 with no major shutdowns, maximizing the delivered luminosity to a total of up to 15 fb<sup>-1</sup>. The current Fermilab schedule fixes the start of Run II in March 2001. ### 2.1 DØ Detector Upgrade The DØ detector upgrade is built on previous strengths, combining excellent calorimetry with good muon coverage and purity. A cross section of the upgraded DØ detector is shown in Fig. 3. The most important improvement is a superconducting solenoid ($`B=2`$ T) providing significantly improved tracking capabilities (see right hand side of Fig. 3). A central fiber tracker consisting of eight superlayers of scintillating fibers allows a measurement of the charged particle momentum. Together with the silicon micro-strip tracker, a momentum resolution of $`\sigma (p_T)/p_T=0.002p_T`$ will be achieved. The silicon tracker consists of six barrels with four layers each ($`r\varphi `$ and $`rz`$ readout) and 12+4 forward disks reaching out to 1.25 m in $`z`$. In addition, improvements to the muon system will allow enhanced muon triggering for $`p_T>1.5`$ GeV$`/c`$ ($`|\eta |<2`$). Central and forward preshower detectors will improve electron identification and triggering on electrons with $`p_T>1`$ GeV$`/c`$ ($`|\eta |<2.5`$). Finally, an impact parameter trigger detecting tracks from displaced vertices is under development. The current DØ schedule expects the central and forward preshower fabrication as well as the central fiber tracker project to be completed by June 2000. The silicon tracker will be finished by September 2000 and the full tracking system installed and hooked up a few weeks later. The muon system will be in place by November 2000 and the calorimeter electronics by January 2001, allowing DØ to be rolled in and ready for beam by February 2001. ### 2.2 CDF Detector Upgrade The goal for the CDF detector upgrade is to maintain detector occupancies at Run I levels, although many of the detector changes also provide qualitatively improved detector capabilities. A schematic view of the CDF II detector is shown in Fig. 4. One major improvement is to the charged particle tracking system (see Fig. 4), vital for the $`B`$ physics program at CDF. A new silicon vertex detector will consist of five layers of double sided silicon from radii of 2.9 cm to 10 cm. The silicon detector will include three modules covering the entire $`p\overline{p}`$ luminous region. In addition, an intermediate silicon layer consisting of two double-sided silicon sensors at larger radii permits stand-alone silicon tracking out to $`|\eta |=2`$. A new open cell drift chamber (COT) will operate at a beam crossing time of 132 ns with a maximum drift time of $`100`$ ns. The COT consists of 96 layers arranged in four axial and four stereo superlayers. It also provides d$`E`$/d$`x`$ information for particle identification. The upgrades to the muon system almost double the central muon coverage. A new scintillating tile plug calorimeter will give good electron identification up to $`|\eta |=2`$. New front-end electronics will be installed, and a DAQ upgrade will allow the operation of a pipelined trigger system. Finally, two additional upgrade projects significantly enhancing the $`B`$ physics capabilities of the CDF II detector have been approved. These include the installation of a low-mass radiation hard single-sided silicon detector with axial strips at very small radius of $`1.6`$ cm, as well as the installation of a time-of-flight system employing 216 three-meter-long scintillator bars located between the outer radius of the COT and the superconducting solenoid. The current CDF schedule foresees cosmic ray running of the detector at the beginning of 2000 and expects the COT to be installed in April 2000. A commissioning run will take place from August to November of 2000. The silicon upgrade will be complete by September 2000 and installed by January 2001. The full CDF II detector will be ready for collisions by March 2001. ## 3 Run II $`𝑩`$ Physics Prospects When discussing the Run II $`B`$ physics prospects in this section, we will refer to a data sample of 2 fb<sup>-1</sup> delivered in two years. We will focus on the prospects for the CDF experiment. DØ has similar expectations. For a measurement of $`\mathrm{sin}2\beta `$, CDF expects 10,000 $`J/\psi K_S^0`$ events in 2 fb<sup>-1</sup> with the $`J/\psi `$ decaying to muon pairs and $`K_S^0\pi ^+\pi ^{}`$. With the enhanced tracking and vertexing capabilities, extended lepton coverage and better particle identification, CDF will improve the effective flavor tagging efficiencies for the different taggers as detailed in Table 1 to a total $`\epsilon 𝒟^2`$ of approximately 9.1%. With this, CDF expects to measure $`\mathrm{sin}2\beta `$ with an uncertainty of $`0.07`$. Figure 5 shows the current CDF result on $`\mathrm{sin}2\beta `$ in the $`(\rho ,\eta )`$-plane where the light shaded area indicates the present $`1\sigma `$ uncertainty. To illustrate the improvements in Run II, the dark shaded area displays the expected error on $`\mathrm{sin}2\beta `$ in Run II with 2 fb<sup>-1</sup>. With respect to other $`CP`$ modes, CDF plans to measure the time dependence of the $`CP`$ asymmetry in $`B^0\pi ^+\pi ^{}`$ decays, determining $`\mathrm{sin}2\alpha `$. CDF will use a displaced track trigger which will trigger on hadronic tracks from long-lived particles such as $`B`$ hadrons. With a fast track trigger at Level 1, CDF finds track pairs in the COT with $`p_T`$ greater than 1.5 GeV$`/c`$. At Level 2, these tracks are linked into the silicon vertex detector, and cuts on the track impact parameter $`d>100`$ $`\mu `$m ($`\sigma (d)25\mu `$m) are applied. CDF expects to collect 4000-7000 $`B\pi \pi `$ events, assuming $`(B^0\pi ^+\pi ^{})=(4.7_{1.5}^{+1.8}\pm 0.6)10^6`$ as measured by CLEO . In addition, there will be about four times more $`BK\pi `$ decays. The implications of this background for an extraction of $`\mathrm{sin}2\alpha `$ at CDF are still under study. For a measurement of $`\mathrm{sin}\gamma `$ in Run II, CDF can use $`B_\text{S}^0`$ mesons. A signal of about 700 $`B_\text{S}^0/\overline{B}_\text{S}^0D_\text{S}^\pm K^{}`$ events is expected. This might allow for an initial measurement of $`\mathrm{sin}\gamma `$ with 2 fb<sup>-1</sup> in Run II. $`B^0\overline{B}^0`$ and $`B_\text{S}^0\overline{B}_\text{S}^0`$ flavor oscillations measure the Cabibbo-Kobayashi-Maskawa matrix elements $`|V_{td}|/|V_{ts}|`$. The recently approved detector upgrades play an important role in CDF’s prospects for measuring $`B_\text{S}^0`$ mixing. The additional inner layer of silicon improves the time resolution from $`\sigma _t=0.060`$ ps to 0.045 ps. This will be important if $`\mathrm{\Delta }m_\text{S}`$ is unexpectedly large. The time-of-flight system will enhance the effectiveness of $`B`$ flavor tagging, especially through same side tagging with kaons and opposite side kaon tagging, to a total $`\epsilon 𝒟^211.3\%`$. CDF expects a signal of 15,000–23,000 fully reconstructed $`B_\text{S}^0D_\text{S}^{}\pi ^+,D_\text{S}^{}\pi ^+\pi ^{}\pi ^+`$ events from the two-track hadronic trigger in 2 fb<sup>-1</sup>. For 20,000 $`B_\text{S}^0`$ events, a $`5\sigma `$ measurement of $`\mathrm{\Delta }m_\text{S}`$ will be possible at CDF for $`\mathrm{\Delta }m_\text{S}`$ values up to 40 ps<sup>-1</sup>. The current limit on $`\mathrm{\Delta }m_\text{S}`$ is 14.3 ps<sup>-1</sup> at 95% C.L. It is noteworthy that physics with $`B_\text{S}^0`$ mesons will be unique to the Tevatron until the turn-on of the LHC in 2006. ## 4 Conclusions The CDF and DØ detector upgrades are well under way with data taking starting in March 2001. There are excellent prospects for $`B`$ physics in Run II, allowing a measurement of $`\mathrm{sin}2\beta `$ with an uncertainty of 0.07. A discovery of $`B_\text{S}^0`$ mixing is possible for $`\mathrm{\Delta }m_\text{S}`$ values up to 40 ps<sup>-1</sup>. The extension of Run II until 2006 will further increase the sensitivity and the $`B`$ physics potential at the Tevatron. CDF and DØ are looking forward to joining the party with the $`B`$ factories. Discussion Michail Danilov (ITEP, Moscow): What is the sensitivity of the DØ experiment for $`B`$ physics studies in Run II? Paulini: As mentioned in my presentation, the prospects for DØ are similar to the ones at CDF. DØ expects about 8500 $`J/\psi K_S^0`$ events with $`J/\psi \mu ^+\mu ^{}`$ but they plan to also trigger on $`e^+e^{}`$ pairs resulting in additional 6500 $`J/\psi K_S^0`$ events. From a time-dependent analysis, DØ expects to measure $`\mathrm{sin}2\beta `$ with an uncertainty of 0.07. In addition, an impact parameter trigger project has recently been approved allowing DØ to detect $`B\pi \pi `$ events and also to explore $`B_\text{S}^0\overline{B}_\text{S}^0`$ mixing.
warning/0001/hep-ph0001333.html
ar5iv
text
# Non-abelian discrete gauge symmetries and inflation ## 1 Introduction Inflation provides initial conditions in the early universe with attractive and observationally motivated features . When inflation is driven by the almost constant potential energy $`V(\varphi )`$ of a field $`\varphi `$, the equation of motion for $`\varphi (t)`$ ($`\varphi `$ becomes spatially homogeneous quickly after inflation starts) is $$\ddot{\varphi }+3H\dot{\varphi }=\frac{V(\varphi )}{\varphi }$$ (1) where $$(\frac{\dot{a}}{a})^2=H^2=\frac{8\pi G}{3}\rho .$$ (2) Here $`\rho `$ is the energy density, $`a(t)`$ the scale factor, and the curvature scale and cosmological constant have been taken to be zero. Approximately scale invariant perturbations are consistent with observation and can be produced with slow roll inflation. Slow roll inflation not only requires the inflationary acceleration of the scale factor ($`\ddot{a}>0`$ which here becomes $`V\dot{\varphi }^2`$) but also that $$\left(\frac{V^{}}{V}\right)^2\frac{1}{M_{\mathrm{Pl}}^2}$$ (3) and $$\left|\frac{V^{\prime \prime }}{V}\right|\frac{1}{M_{\mathrm{Pl}}^2}.$$ (4) Here $`M_{\mathrm{Pl}}`$ is the Planck constant, $`1/\sqrt{8\pi G}`$. Many models of inflation are built ignoring gravitational strength interactions, and so are implicitly setting $`M_{\mathrm{Pl}}=\mathrm{}`$. In this case, satisfying the conditions above is not possible. Keeping $`M_{\mathrm{Pl}}`$ finite implies one should include gravitational effects. To include gravitational effects, we will work in the context of supergravity. In this case, the first condition suggests we should be near a maximum, or other extremum, of the potential. The second has been found to be non-trivial . In supergravity, the potential is composed of two parts, the $`F`$-term and the $`D`$-term. If the inflationary potential energy is dominated by the $`F`$-term then one can show that $$\frac{V^{\prime \prime }}{V}=\frac{1}{M_{\mathrm{Pl}}^2}+\text{model dependent terms}$$ (5) Unless the model dependent terms cancel the first term, the second slow roll condition, Eq. (4) above, is violated. Thus to build a model of slow-roll inflation one must be able to control the gravitational strength corrections. There have been various attempts at achieving slow-roll inflation naturally, for a discussion see , for a recent review of inflationary models see, e.g. . Discrete non-abelian gauge symmetries provide several ingredients for successful inflation. The symmetries ensure that the potential is flat enough for inflation to occur, and the discreteness of the symmetries provide exit points in the potential where inflation can end. As will be seen below, non-abelian symmetries in particular are useful for hybrid models in low energy effective field theories with supergravity corrections. ## 2 The idea In hybrid and mutated hybrid models, one field slowly rolls and a second is pinned until the first field reaches some critical value. E.g., for concreteness and illustration, take the slowly rolling field to be $`\mathrm{\Phi }`$ and call the second field $`\mathrm{\Psi }`$. The hybrid potential for $`\mathrm{\Psi }`$ in the simplest case has the form $$(\alpha |\mathrm{\Phi }|^\beta m^2)|\mathrm{\Psi }|^2+O(\mathrm{\Psi }^3)$$ (6) where $`m^2,\alpha ,\beta >0`$. The potential for $`\mathrm{\Phi }`$ is not shown, its crucial feature is sufficient flatness to allow for slow roll (the conditions (3,4)). For large $`\mathrm{\Phi }`$, the field $`\mathrm{\Psi }`$ has a large mass and is pinned. During slow roll inflation, $`\mathrm{\Phi }`$ decreases, until near the critical value $`\mathrm{\Phi }_c^\beta m^2/\alpha `$ the field $`\mathrm{\Psi }`$ is freed. The rolling of $`\mathrm{\Psi }`$ eventually ends inflation, often quickly after $`\mathrm{\Phi }`$ reaches the critical value. The specific properties of the exit depend on the other terms in the potential as well. (Slow roll conditions for multiple fields are discussed in ). Supergravity corrections, as mentioned above, will generically give both $`\mathrm{\Phi }`$ and $`\mathrm{\Psi }`$ masses of order the Hubble constant, making slow roll for $`\mathrm{\Phi }`$ hard to achieve. This letter describes a mechanism for providing a flat potential when supergravity corrections are included, and providing a way for inflation to end via special exit points. Although hybrid models are discussed here, a variant also works for mutated hybrid models as well. Hybrid and mutated hybrid inflationary models have the advantage that the energy scale of inflation can sometimes be quite low, which allows for a description in terms of lower energy, and thus often better controlled, theories. (The low energy scale advantage is coupled with the caveat that large fluctuations are likely at the end of inflation in many of these models.) Adding a discrete non-abelian gauged symmetry in the low energy effective field theory, the context for these models, provides for both the flatness and the exit as follows. Instead of taking one field $`\mathrm{\Phi }`$ and one field $`\mathrm{\Psi }`$, take an N-tuple of fields, $`\mathrm{\Phi }_i`$ and $`\mathrm{\Psi }_i`$, with $`i=1,\mathrm{}N`$, and impose a discrete non-abelian gauged symmetry. This symmetry has to be chosen such that the lowest order allowed terms in the superpotential respect a larger continuous symmetry, for example SU(N) in this case. The higher order terms then break this effective continuous symmetry to a discrete subgroup. The other condition on this effective field theory is that the supersymmetry breaking comes from some other sector, and provides a vacuum energy $`V_0`$. As we are using an effective field theory we do not need to say where this term comes from in particular. The vacuum energy in turn generates supersymmetry breaking masses for the fields $`\mathrm{\Phi }_i,\mathrm{\Psi }_i`$ as well as other higher order terms. As the lowest order terms in the superpotential respect the SU(N) symmetry, the mass terms for the fields will be functions of the SU(N) invariants $`|\mathrm{\Phi }|^2=_i|\mathrm{\Phi }_i|^2`$ and $`|\mathrm{\Psi }|^2=_i|\mathrm{\Psi }_i|^2`$. Renormalization group flow will shift the masses $`m_\mathrm{\Phi }^2`$ and $`m_\mathrm{\Psi }^2`$ to have non-trivial minima which to leading order only depend on the SU(N) symmetric field combination $`|\mathrm{\Phi }|`$ and $`|\mathrm{\Psi }|`$, that is, for example, $$V(|\mathrm{\Phi }_i|)m_\mathrm{\Phi }^2(|\mathrm{\Phi }||\mathrm{\Phi }_0|)^2+\mathrm{},$$ (7) This consequently provides for flat directions in the potential where inflation can occur. The field $`|\mathrm{\Phi }|`$ will be locked, $`|\mathrm{\Phi }||\mathrm{\Phi }_0|`$ but the various components of $`\mathrm{\Phi }`$ can vary, causing slow roll inflation. If the SU(N) symmetry were exact, then the corresponding hybrid term above would have to also respect the SU(N) symmetry and be a function of $`|\mathrm{\Phi }|,|\mathrm{\Psi }|`$ alone, just as above: $$V(\mathrm{\Phi }_i,\mathrm{\Psi }_i)(\alpha |\mathrm{\Phi }|^\beta m_\mathrm{\Psi }^2)|\mathrm{\Psi }|^2+\mathrm{}$$ (8) Although the components of $`|\mathrm{\Phi }|`$, the $`\mathrm{\Phi }_i`$, are free to vary, only their combination appears in the hybrid exit, and as $`|\mathrm{\Phi }|`$ remains locked, the exit cannot take place. On the other hand, if the theory only respects a discrete subgroup of SU(N) this can change. For instance if the theory respects say a permutation symmetry, the hybrid term $$V(\mathrm{\Phi }_i,\mathrm{\Psi }_i)\underset{i}{}(\alpha _i|\mathrm{\Phi }_i|^\beta m_\mathrm{\Psi }^2)|\mathrm{\Psi }_i|^2+h.o.t.$$ (9) may be allowed. If $`\beta 4`$, this term will be higher order than the couplings which generate the SU(N) symmetric masses, thus allowing the SU(N) symmetry to consistently pin $`|\mathrm{\Phi }||\mathrm{\Phi }_0|`$. In this case, the individual $`|\mathrm{\Phi }_i|`$ can control the end of inflation, by decreasing to some critical value, although the sum of the $`|\mathrm{\Phi }_i|^2`$ is fixed by the supersymmetry breaking and SU(N) respecting renormalization group effects. In this way, the discrete symmetry allows inflation both to occur (via the continuous effective symmetry) and to end with a hybrid mechanism (via the discrete subgroup). As the norms of the fields, the $`|\mathrm{\Phi }_i|`$ appear in these hybrid potentials, a non-abelian discrete symmetry is needed. An abelian discrete gauge theory would allow for changes in the phase of $`\mathrm{\Phi }_i`$, but as the exit depends on the magnitude of $`\mathrm{\Phi }_i`$ this abelian variation will not trigger the exit. The basic idea of using discrete gauged symmetries generalizes the idea used in Natural Inflation . It is also similar to , where an approximate symmetry is provided by lower terms in the lagrangian (in particular the kinetic terms). Although orbifold constructions, such as also involve a discrete gauge symmetry, the flatness of the potential is obtained differently. ## 3 Example This basic idea can be used to construct a low energy effective model. One chooses a suitable gauge group and representations, ensuring that the gauge symmetries are not anomalous. The symmetries constrain the allowed terms in the superpotential $`W`$ and Kahler potential $`K`$. The couplings and supersymmetry breaking terms correspond to regions in parameter space which have constraints both from observation and underlying physics (for example the supersymmetry breaking terms will naturally have coefficients the scale of supersymmetry breaking). For this effective field theory, as mentioned earlier, we assume a hidden sector breaks supersymmetry. This generates supersymmetry breaking terms in the effective potential, including a vacuum energy $`V_0`$ and masses for the scalars. In addition, as also mentioned earlier, the renormalization group running of the supersymmetry breaking mass term for $`\mathrm{\Phi }`$ to generates a potential for $`\mathrm{\Phi }`$ with non- trivial minimum $`|\mathrm{\Phi }|=\mathrm{\Phi }_0`$. The renormalization is induced (to leading order) by low dimension couplings symmetric under the extended continuous symmetry. Thus the renormalization group masses and the potential will be symmetric under the extended continuous symmetry. A superpotential with a term $`\mathrm{\Phi }^{\frac{\beta }{2}}\mathrm{\Psi }^2`$, will produce a hybrid potential of the form above, eqn.(6), using the same notation for the scalar component of the superfield and the superfield itself. For example, a model based on a discrete subgroup of SU(2) can have the fields and charges shown in table one. The second discrete symmetry must have $`M>4`$, the specific value is not important. The generators of the discrete subgroup of SU(2) are taken to be $$\left(\begin{array}{cc}0\hfill & i\hfill \\ i\hfill & 0\hfill \end{array}\right)\left(\begin{array}{cc}0\hfill & 1\hfill \\ 1\hfill & 0\hfill \end{array}\right)\left(\begin{array}{cc}e^{\frac{2\pi i}{N}}\hfill & 0\hfill \\ 0\hfill & e^{\frac{2\pi i}{N}}\hfill \end{array}\right)$$ (10) with $`N>4`$. The holomorphic terms which respect the symmetries give the superpotential for this model: $$W=\lambda _0(\varphi _1\chi _2\varphi _2\chi _1)\rho +\frac{1}{2}\lambda _1(\varphi _1^2\psi _2^2+\varphi _2^2\psi _1^2)+\lambda _2\varphi _1\varphi _2\psi _1\psi _2+\lambda _\chi \chi _1^2\chi _2^2+h.o.t..$$ (11) The lowest order terms respect the full SU(2), while the higher order terms only have this symmetry for special values of $`\lambda _1,\lambda _2`$. The leading neglected higher terms are of the form $`\rho \varphi \chi ^3`$, $`\varphi ^2\chi ^2\rho ^2`$ and $`\varphi ^2\psi ^2\chi ^2`$. Calculating the bosonic contribution to the potential (again using the same notation for the superfields and their bosonic components) from this superpotential gives $$\begin{array}{cc}V_{susy}=\hfill & |\lambda _0|^2(\varphi _1\chi _2\varphi _2\chi _1|^2+|\varphi _1\rho +\frac{\lambda _\chi }{\lambda _0}\chi _2\chi _1^2|^2+|\varphi _2\rho +\frac{\lambda _\chi }{\lambda _0}\chi _2^2\chi _1|^2)\hfill \\ & +|\lambda _0\chi _2\rho +\lambda _1\varphi _1\psi _2^2+\lambda _2\varphi _2\psi _1\psi _2|^2+|\lambda _0\chi _1\rho +\lambda _1\varphi _2\psi _1^2+\lambda _2\varphi _1\psi _1\psi _2|^2\hfill \\ & +|\lambda _1\varphi _2^2\psi _1+\lambda _2\varphi _1\varphi _2\psi _2|^2+|\lambda _1\varphi _1^2\psi _2+\lambda _2\varphi _1\varphi _2\psi _1|^2+h.o.t.\hfill \end{array}$$ (12) The supersymmetry breaking terms are $$\begin{array}{cc}V_{break}\hfill & =V_0+m_\chi ^2|\chi |^2+\stackrel{~}{m}_\varphi ^2|\varphi |^2m_\psi ^2|\psi |^2+m_\rho ^2|\rho |^2+\mu _0(\varphi _1\chi _2\varphi _2\chi _1)\rho \hfill \\ & +\frac{1}{2}\mu _1(\varphi _1^2\psi _2^2+\varphi _2^2\psi _1^2)+\mu _2\varphi _1\varphi _2\psi _1\psi _2+h.c.+h.o.t.\hfill \end{array}$$ (13) Note that for the hybrid exit the fields $`\psi _i`$ have a negative mass squared. As mentioned earlier, we take the supersymmetry breaking mass term for $`\varphi `$ ($`|\varphi |^2=|\varphi _1|^2+|\varphi _2|^2`$) to be its renormalized form $$\stackrel{~}{m}^2|\varphi |^2=m_\varphi ^2(|\varphi ||\varphi _0|)^2+h.o.t.$$ (14) The mass term for $`\psi `$ can have a similar form, but as it will be considered near the value $`\psi 0`$, this effect is not relevant. An extremum of the combined potential $`V_{susy}+V_{break}`$ occurs when $$\chi _i=\rho =\psi _i=0,|\varphi |\varphi _0.$$ (15) This is the background value around which slow roll inflation is taken to occur. The mass terms for $`\psi _i`$ are $$m_\psi ^2|\psi |^2+\frac{1}{2}\mu _1(\varphi _1^2\psi _2^2+\varphi _2^2\psi _1^2)+\mu _2\varphi _1\varphi _2\psi _1\psi _2+h.c.+|\lambda _1\varphi _2^2\psi _1+\lambda _2\varphi _1\varphi _2\psi _2|^2+|\lambda _1\varphi _1^2\psi _2+\lambda _2\varphi _1\varphi _2\psi _1|^2$$ (16) Taking $`\lambda _2\lambda _1`$ so that the SU(2) is strongly broken, and using that $`\mu _i`$ are small, one disentangles that the leading terms in the mass of $`\psi _i`$ are proportional to $`|\lambda _1ϵ_{ij}\varphi _j^2|^2`$. Thus as mentioned earlier $`|\varphi |`$ can remain fixed while $`|\varphi _j|`$ rolls, and thus inflation can end. ## 4 Summary We have shown that non-abelian discrete gauge symmetries can provide for sufficient flatness for inflation to occur and special exit points where inflation can end. The flatness comes from an approximate continuous symmetry induced by the discrete symmetry, while the special exit points and dynamics are due to the fact that only the discrete symmetry is exact and unbroken. This gives a new way for inflation to happen even when masses from supergravity corrections are taken into account. As a low energy effective field theory is used, one does not need specific information about physics at higher energies, but only the induced field content and symmetries at the scale of interest. Integrating this mechanism into a more complete model requires several steps, which are outside the scope of this letter. The field content must be anomaly free which may require the addition of more fields. Other considerations will constrain the model’s parameter space and background field values. For instance, the stability of the potential around eq.(15) may require specific values of $`\varphi _0`$ and perhaps couplings. The slow roll behavior of $`\varphi _i`$ must lead $`\varphi _i`$ to the hybrid exit, and will be controlled by higher order corrections in the potential such as those mentioned above, corrections to the Kahler potential and supersymmetric loop corrections. The exit from inflation must not introduce large fluctuations in an observable regime, as mentioned earlier. Constraints from the measured COBE normalization and the scale dependence of the fluctuations (tilt) must also be imposed. These requirements fix combinations of the parameters for the model. In a longer forthcoming paper , we present two specific models utilizing discrete non-abelian gauge symmetry as described here. We impose these constraints for a hybrid and mutated hybrid model, including higher order corrections, and calculate detailed properties of the parameter space, the exit from inflation and features of the spectrum. EDS was supported by the DOE and the NASA grant NAG 5-7092 at Fermilab, Grant No. 1999-2-111-002-5 from the interdisciplinary Research Program of the KOSEF and Brain Korea 21 Project. JDC was supported by NSF-PHY-9800978 and NSF-PHY-9896019. For hospitality, EDS thanks M. White and R. Leigh at UIUC, JDC thanks the Aspen Center for Physics, and we both thank the Santa Fe 99 Workshop on Structure Formation and Dark Matter. JDC is grateful to Martin White for numerous discussions, and J. Terning for information about supersymmetry.
warning/0001/cond-mat0001123.html
ar5iv
text
# Enhanced vortex damping by eddy currents in superconductor-semiconductor hybrids ## Abstract An enhancement of vortex-motion damping in thin Pb/In superconducting films is obtained through coupling to an adjacent two-dimensional electron gas formed in a modulation-doped GaAs/AlGaAs heterostructure. This effect is observed by monitoring the power dissipation at the superconductor in the vortex state while increasing the density of the electron gas using a gate voltage. Quantitative agreement is found with calculations based on a viscous model of vortex damping which considers generation of eddy currents in the electron gas by moving flux lines. In the regime of filamentary and channel vortex flow, eddy-current damping leads to striking dissipation breakdown due to stopping of entire vortex channels. Superconductor-semiconductor hybrid structures are emerging as key devices in the search for new physical phenomena resulting from interactions between two systems with dissimilar electronic properties . In particular, Josephson-type junctions with Nb electrodes coupled by a two-dimensional electron gas (2DEG) in InAs layers exhibit phase-sensitive transport due to Andreev reflections of quasi particles at the interfaces between normal metal and superconductor . Other experiments concentrate on commensurability and interference effects on electron ballistic transport in the 2DEG, which occur when a perpendicular magnetic field is spacially modulated by the vortices of an adjacent superconducting film. In this case, a pronounced suppression of the Hall effect was observed and ascribed to electron diffraction by flux quanta . Few investigations are concerned with the influence of a normal metal on the vortex dynamics under a transport current, although the devices were intentionally designed to have no viscous coupling . One can think of our hybrid system as a modified Giaever’s dc transformer , in which one of the superconducting films has been replaced by a 2D electron gas. Under the action of a Lorentz force the vortices move at constant velocity due to viscous damping. For an isolated superconductor, this damping originates from the voltage induced across the normal core of each moving vortex. By bringing a highly mobile electron gas close enough to the superconducting film, i.e. at a distance of the order of the London penetration depth, an additional dissipation mechanism is introduced through magnetic coupling resulting in an increase of viscosity. An interesting issue is to what extent this would affect filamentary and channel vortex flow, for which dissipation jumps are observed in the current-voltage curves . The study of vortex damping in hybrids may provide further insight into vortex-vortex interactions and pinning effects. This Letter reports the first observation of damping enhancement for vortex motion due to the presence of a high-mobility electron gas in superconductor-semiconductor hybrids. The samples used in our experiments consist of thin Pb/In films evaporated on top of modulation-doped GaAs/AlGaAs heterostructures. The evidence is found in the decrease of dissipation voltage measured at the superconducting film due to a higher viscosity for vortex flow in the hybrid system, as compared to the case without the 2DEG beneath. We vary the normal metal conductivity by increasing the carrier density using a gate voltage applied between the 2DEG and a back contact. Our results are in quantitative agreement with the predictions of a model which accounts for the generation of eddy currents in the electron gas by flowing vortices. For an estimated increase in electron density of up to 20% the relative change observed in dissipation voltage lies in the one-percent range. For the currents and magnetic fields at which filamentary vortex flow occurs, however, striking dissipation reductions are readily achieved. The semiconductor component of the hybrid system is either a 25 nm wide GaAs/AlGaAs single quantum well (SQW) or a single heterointerface (SHI) structure. A two-dimensional electron gas is realized by modulation doping and is buried at a distance $`D=`$ 75 nm and 50 nm from the surface for the SQW and SHI structure, respectively . The nominal mobilities and carrier densities of both samples at 4.2 K and under illumination are $`\mu 8\times 10^5\mathrm{cm}^2/\mathrm{Vs}`$ and $`n5.6\times 10^{11}\mathrm{cm}^2`$. The electron gas is contacted from the surface by In alloying in order to apply a gate voltage $`U_g`$ between it and a metallic back contact. The variation of the 2D density was examined previously in photoluminescence experiments . A linear increase in the carrier density $`n_{2\mathrm{D}}`$ between its nominal value and at most $`6.5\times 10^{11}\mathrm{cm}^2`$ can be achieved by applying a gate voltage between 0 and 200 V. Hence the estimated maximum possible increase of density is less than 20%. Superconducting films of Pb with nominally 14 at.% In were evaporated on the semiconductor surface ($`4\times 4`$ mm<sup>2</sup>) with film thicknesses $`d`$ ranging from 60 to 300 nm, as determined using atomic-force microscopy. The superconducting transition in zero field occurs at 7.2 K. For transport experiments, Au leads were pressed against the superconductor film. Current-voltage measurements were performed with standard four-terminal configuration using dc currents up to 1.5 A. Experiments were carried out at 4.2 K and low perpendicular magnetic fields $`B<`$ 0.2 T. To model the coupling between vortex lattice and electron gas in our hybrid samples we consider the effect on the normal metal of the magnetic field $`B`$ of a moving vortex with speed $`v`$. The experimental situation is schematically shown in the inset to Fig. 1. Flowing vortices induce an electric field in the 2DEG that generates eddy currents leading to an additional dissipation which, in turn, forces the fluxoids to slow down. In the limit $`D\lambda ^2/d`$ of small superconductor-2DEG distances as compared to the effective London penetration depth, the magnetic field of a vortex can be approximated as $`𝐁(\mathrm{\Phi }_0/2\pi \lambda ^2)K_0(r/\lambda )\widehat{z}`$, where $`\mathrm{\Phi }_0=h/2e`$ is the flux quantum and $`K_0`$ the zeroth-order Bessel function of imaginary argument . For our samples the effective penetration depth is five to eight times $`D`$ . The vector potential in the plane of the 2DEG can be written as $$𝐀(\rho )=\frac{\mathrm{\Phi }_0}{2\pi \rho }F(\rho /\lambda )\widehat{e}_\phi ,$$ (1) where $`\rho `$ is the polar radius from the vortex core, and $`F(x)=_0^x𝑑yyK_0(y)`$. The time-varying vector potential produced by a flowing vortex induces an electric field $`𝐄=v\frac{}{x}𝐀(xvt,y)`$, which causes joule dissipation in the 2D gas. The energy loss per unit time is calculated according to $$\frac{d\epsilon }{dt}=\sigma _{2\mathrm{D}\mathrm{E}\mathrm{G}}v^2d^2x\left[\frac{}{x}𝐀(\rho /\lambda )\right]^2\frac{\mathrm{\Phi }_0^2}{2\pi \lambda ^2}\sigma _{2\mathrm{D}\mathrm{E}\mathrm{G}}v^2\eta _{2\mathrm{D}\mathrm{E}\mathrm{G}}v^2,$$ (2) where $`\sigma _{2\mathrm{D}\mathrm{E}\mathrm{G}}`$ is the conductivity of the 2D electron gas and the dimensionless integrals are assumed to be of the order of one. We arrive at the general result that eddy current generation in the hybrid system manifests itself in a contribution to the viscosity $`\eta _{2\mathrm{D}\mathrm{E}\mathrm{G}}`$ describing the enhanced damping of vortex motion. This effect can be observed if $`\eta _{2\mathrm{D}\mathrm{E}\mathrm{G}}`$ is comparable to the viscosity of type-II superconducting material, $`\eta _{\mathrm{SC}}=\sigma _nd\mathrm{\Phi }_0^2/2\pi a^2`$ , where $`\sigma _n`$ is the normal state conductivity of the superconductor and $`a`$ the vortex core radius. The electron gas acts as a shunt conductor, thus increasing the system viscosity $`\eta _{\mathrm{tot}}=\eta _{\mathrm{SC}}+\eta _{2\mathrm{D}\mathrm{E}\mathrm{G}}`$ by a factor $$1+\frac{\eta _{2\mathrm{D}\mathrm{E}\mathrm{G}}}{\eta _{\mathrm{SC}}}=1+\left(\frac{a}{\lambda }\right)^2\frac{\sigma _{2\mathrm{D}\mathrm{E}\mathrm{G}}}{\sigma _nd}.$$ (3) Hence the dissipation voltage under a transport current is $`U_dv`$ and the vortex velocity is determined by the balance between the driving force $`jB`$ and the viscous drag $`\eta v`$. Eddy-current damping grows in proportion to $`\sigma _{2\mathrm{D}\mathrm{E}\mathrm{G}}`$, i.e. to the carrier density $`n_{2\mathrm{D}}`$ of the electron gas. The dissipation due to flowing vortices in the PbIn superconducting film is reduced by increasing the charge density in the neighboring electron gas. Figure 1 shows a typical dissipation voltage ($`U_d`$) versus gate bias ($`U_g`$) curve of a PbIn/SHI hybrid sample measured at a constant transport current of 600 mA. Sweeping $`U_g`$ from 0 to 170 V causes a linear decrease in dissipation of $`\mathrm{\Delta }U_d=0.022`$ mV, i.e. $`0.1`$ % of the initial value. Since optical and transport experiments could not be carried out simultaneously, we give here the gate voltage as a measure of the electron density which we assume to vary linearly with $`U_g`$, as inferred from optical measurements . We emphasize that in spite of the relatively small change in 2DEG density the coupling within the hybrid appears to be effective enough to show up in its dissipative behavior. We also notice that during the experiment no leakage current between superconductor and electron gas was ever detected, thus we rule out any spurious bias to be at the origin of the observed effect. We interpret the dissipation change in the superconductor with rising charge density in the 2DEG as the effect of eddy currents in the electron gas which slow down the vortices causing the voltage across the superconductor to decrease. Using Eq. (3) we can now estimate the magnitude of this effect. Taking $`a\lambda `$, $`d100`$ nm, $`\sigma _n1.4\times 10^5\mathrm{\Omega }^1\mathrm{cm}^1`$ as obtained from resistance measurements, and $`\sigma _{2\mathrm{D}\mathrm{E}\mathrm{G}}0.08\mathrm{\Omega }^1`$ with the 2DEG parameters given above, this yields $`\eta _{2\mathrm{D}\mathrm{E}\mathrm{G}}/\eta _{\mathrm{SC}}5\%`$. This change in dissipation voltage corresponds to the difference between having and not having the electron gas next to the superconductor. In our experiments, however, we start from a finite density and produce a variation of about 10%. Thus, the calculated dissipation change is around 0.5%, in very good agreement with the experimental results. Eddy-current damping effects are much more pronounced in the regime of filamentary and channel vortex flow, for they can lead to a striking fall of dissipation voltage by more than one order of magnitude, as shown in Fig. 2. Here, $`U_d`$ was measured at a current close to the repinning transition of a large vortex channel. The inset to Fig. 2 displays the $`IV`$ characteristic of the superconducting film measured for a field of 53 mT and at 4.2 K but without gate bias. The abrupt jumps and large hysteresis apparent in the $`IV`$ curve are the signature of channel vortex flow . As indicated by the arrow, the point where the measurement of dissipation versus gate voltage was carried out is close above the critical current at which the downward jump occurs. In this case, the vortex speed is just high enough for the channel to keep flowing. Energy loss due to eddy currents slows the vortices further down, so that the whole channel will eventually be repinned. When the vortex channel stops, dissipation suddenly drops. This process is irreversible since, as can be seen in Fig. 2, dissipation does not resume to its initial value when the gate voltage is swept back to zero. This effect is highly reproducible even after heating the sample over $`T_c`$ and re-cooling. The study of the dependence on gate voltage of the current values at which dissipation jumps occur provides further information about the nature of filamentary vortex flow and the role played by pinning. Figure 3 shows one example in which the currents for the upward and downward voltage jumps are plotted as a function of gate bias (the corresponding $`IV`$ curve is displayed in the inset to Fig. 3). The jump-up current is independent of the 2D density indicating that the pinning strength is not appreciably affected by the presence of the electron gas. In contrast, the jump-down current increases with $`U_g`$. With increasing viscosity the vortices of a moving channel slow down such that its repinning occurs at larger values of the transport current. This is interpreted as additional evidence of a repinning force which depends on vortex velocity. A similar increase in jump-down current is observed for thin PbIn films on glass by decreasing the external magnetic field . The influence of magnetic field homogeneity on eddy-current damping is revealed by the percentage change in dissipation as a function of magnetic field at constant transport current. As a measure of the damping strength, the maximum dissipation change $`\mathrm{\Delta }U_d`$ for a gate voltage interval of $`\mathrm{\Delta }U_g=170`$ V is normalized by the dissipation voltage $`U_{d0}`$ at zero bias. The corresponding values measured at 500 mA and 1000 mA are shown in Fig. 4 as a function of the external magnetic field $`B`$. At $`B30`$ mT damping causes 0.1–0.2 % change in dissipation. At 40 mT, the data for both 0.5 A and 1 A display a sharp maximum and become very small at fields larger than 50 mT, where the resistance of the film is close to normal but the transport behavior is characterized by massive vortex flow. The weakening of the effect of eddy-current damping on dissipation with increasing magnetic field (solid line in Fig. 4) can be explained as due to the growing homogeneity of the field pattern of the vortices while approaching the upper critical field $`B_{c2}`$. At low magnetic fields, i.e. low vortex density, the field distribution is very inhomogeneous, since $`B`$ has a maximum at the vortex cores dropping to zero between them. Thus, $`\dot{B}`$ is large and damping is efficient. When vortices start to overlap, the lateral modulation of magnetic field in the plane of the 2DEG is continuously reduced. As a consequence, eddy currents as well as damping effects are weak. In contrast, the peak at 40 mT is associated with the enhancement of the effects due to eddy-current damping in the regime dominated by filamentary vortex flow. Here the dissipation change is greatly enhanced by fluctuations in the number of moving vortices contributing to dissipation due to the depinning and repinning of a large number of small vortex filaments in quick succession. In summary, we have observed significant additional damping of vortex motion in superconductor-semiconductor hybrid systems. A theoretical model is used to calculate the damping effect from eddy currents generated in the 2D electron gas showing quantitative agreement with the experiment. Under conditions of filamentary vortex flow, the energy loss due to eddy currents leads to the stopping of entire channels, such that power dissipation in a hybrid device can be switched off by slightly increasing the electron density. At large fields damping is weak due to the vanishing lateral field modulation in the plane of the 2DEG. We point out that, although our observations can be explained within the framework of classical electrodynamics, novel effects due to quantization of the electron gas conductivity are anticipated to occur for the conditions of the experiments. Our results provide further insight into the issue of vortex dynamics with dissipation and open up a new class of devices for the study of correlations between adjacent non-tunneling systems with dissimilar electronic and magnetic properties. A.G.R. acknowledges useful conversations with John Clem and Cagliyan Kurdak and partial support from the National Science Foundation.
warning/0001/cond-mat0001021.html
ar5iv
text
# Non-linear Transport in Quantum-Hall Smectics ## 1 Introduction Recent transport experiments have established that the resistivity of a two-dimensional electron system with weak disorder and valence orbital Landau level filling $`\nu ^{}`$ close to $`1/2`$ is anisotropic when the valence orbital Landau level index $`N2`$. Apparently the ground state spontaneously breaks orientational symmetry, a property believed to be associated with the uni-directional charge-density-wave ground states predicted in this regime by Hartree-Fock mean-field theory . The charge-density-wave (CDW) state consists of electron stripes of width $`a\nu ^{}`$ with local Landau level filling factor $`\nu =[\nu ]+1`$, separated by hole stripes of width $`a(1\nu ^{})`$ with local filling factor $`\nu =[\nu ]`$. Here $`a`$, the CDW period, is comparable to the Landau level’s cyclotron orbit diameter. Because of symmetry properties shared with the smectic state of classical liquid crystals, emphasized by Fradkin and Kivelson , these states have been referred to as quantum Hall smectics, a practice we follow here. The most important transport property of quantum Hall smectics is that dissipation occurs over a wide range of filling factors surrounding $`\nu ^{}=1/2`$ and is not activated at low temperatures. This behavior is not consistent with the properties of a CDW state, which would be pinned by the random disorder potential and have a large gap for mobile quasiparticle excitations, and suggests that, although Hartree-Fock theory hints at the energetic motivation for a ground state with broken orientational symmetry, the description which it provides of the ground state is flawed. Several recent theoretical papers have addressed the properties of quantum Hall smectics and the energetic competition between CDW states, compressible composite-fermion states, and paired incompressible quantum Hall states. In one recent paper we have described a theory of quantum Hall smectics which starts from the Hartree-Fock theory ground state, recognizes the electron stripes as coupled one-dimensional electron systems, and treats residual interaction and disorder terms neglected by Hartree-Fock theory using the convenient bosonization techniques of one-dimensional electron physics. The most important conclusions of this work are the following: i) the quantum Hall smectic state is never the ground state but instead is always unstable, for $`\nu ^{}`$ close to $`1/2`$ likely to an anisotropic electron Wigner crystal state; ii) for $`0.4<\nu ^{}<0.6`$ the interaction terms responsible for the Wigner crystal can be neglected at temperatures available in a dilution fridge; and iii) weak disorder which scatters electrons from stripe to stripe, enabling hard-direction transport, leads to non-linear transport. In this article we emphasize and expand on an experimentally important prediction of this work, namely that the strength of the transport non-linearity is sensitive to the nature of the electron-electron interaction. In particular we predict that the transport non-linearity can be enhanced by placing a screening plane close to the two-dimensional electron system. In Section II we explain our theory of transport in quantum Hall smectics and discuss how the coefficients which govern the power-law behavior of the differential resistivity for weak disorder are related to correlations in the coupled one-dimensional electron stripes. In Section III we briefly review our theory of these correlations and explain why long-range electron-electron interactions weaken transport non-linearities. In Section IV we present numerical results for the non-linear transport coefficients for a model in which interactions in the two-dimensional electron layer are screened by a metallic layer co-planar with the electron system but separated from it by a distance $`d`$. We conclude in Section V with a brief summary. ## 2 Anisotropic Transport Properties Our transport theory is built on a semiclassical Boltzmann-like approach; microscopic physics enters only in the calculation of scattering rates. We choose a coordinate system where the $`\widehat{x}`$ (horizontal) direction runs along the stripes which are separated in the $`\widehat{y}`$ (vertical) direction. We assume that the charge density wave itself is pinned and immobilized by both the edges of the sample and weak impurities which couple to the electrons within the stripes. In this case, collective sliding motion of the charge-density will be absent, and electrical transport will be dominated by single-electron scattering across and between electron stripes. An important property of electronic states in the quantum Hall regime, is the spatial separation of states which carry current in opposite directions. In the case of the electron stripes in quantum Hall smectics, the Fermi edge states which carry oppositely directed currents along the stripes (left-going and right-going) are located on opposite sides of the stripe. Translational invariance along the $`\widehat{x}`$ direction, allows us to use a Landau gauge where this component of wavevector $`\mathrm{}k_x`$ is a good quantum number in the absence of interactions and disorder. The single-particle states at the stripe edges have velocity magnitude $`v_F`$, the Fermi velocity. In the Landau gauge, $`\widehat{x}`$ direction momenta are related to $`\widehat{y}`$ direction positions by $`k_x=y/\mathrm{}^2`$ where $`\mathrm{}(\mathrm{}c/eB)^{1/2}`$ is the magnetic length, so that states on opposite sides of the same stripe differ in momenta by a $`\nu ^{}a/\mathrm{}^2`$ and the adjacent sides of neighboring stripes differ in momenta by $`(1\nu ^{})a/\mathrm{}^2`$. We now summarize the basic assumptions on which our semiclassical transport theory is based and quote the expressions implied by these assumptions. For further details see Ref. . We assume that in the steady state, each edge of each stripe is characterized by a local chemical potential. Translational invariance in the $`\widehat{y}`$ direction implies that the chemical potential drops across each stripe and between any two adjacent stripes are the same, mf $`\mu `$ is the chemical potential drop across an electron stripe, it follows that the potential drop between stripes is $`eE_ya\mu `$, where $`E_y`$ is the hard-direction electric field, the field which supports a steady state transport across the stripes. The electric field in the $`\widehat{x}`$ direction produces a semiclassical drift in momentum space which drives the system from equilibrium. We assume that disorder scattering across and between stripes, then attempts to reestablish equilibrium and that the drift and scattering processes are in balance in the steady state. The scattering currents in the hard-direction are characterized by relaxation times $`\tau _e`$ and $`\tau _h`$ respectively. The current along a stripe is analogous to the quantized current in a long narrow Hall bar and is proportional to the chemical potential difference across that stripe. Combining these ingredients leads to the following expressions for the resistivities: $`\rho _{\mathrm{easy}}`$ $`=`$ $`{\displaystyle \frac{h}{e^2}}{\displaystyle \frac{1}{\tau _e([\nu ]+1)^2+\tau _h[\nu ]^2}}{\displaystyle \frac{a}{v_F}}`$ $`\rho _{\mathrm{hard}}`$ $`=`$ $`{\displaystyle \frac{h}{e^2}}{\displaystyle \frac{1}{\tau _e([\nu ]+1)^2+\tau _h[\nu ]^2}}{\displaystyle \frac{v_F\tau _e\tau _h}{a}}`$ $`\rho _{\mathrm{hall}}`$ $`=`$ $`{\displaystyle \frac{h}{e^2}}{\displaystyle \frac{1}{\tau _e([\nu ]+1)^2+\tau _h[\nu ]^2}}([\nu ]+1)\tau _e+[\nu ]\tau _h,`$ (1) where $`\rho _{\mathrm{easy}}=\rho _{xx}`$, $`\rho _{\mathrm{hard}}=\rho _{yy}`$, and $`\rho _{\mathrm{hall}}=\rho _{xy}`$. For $`\nu ^{}=1/2`$, this theory makes a parameter free prediction for the product $`\rho _{easy}\rho _{hard}`$ which has been confirmed experimentally . In fact, as emphasized by van Oppen et al., this feature of our results has a greater validity than would be suggested by our assumption of largely intact electron stripes. Our main interest here however, is in expanding on our predictions for non-linearities in the easy and hard direction differential resistivities. These predictions were made on the basis of a simple lowest order renormalization group scheme for handling the infrared divergence which appear when disorder terms which scatter electrons either across or between stripes are treated perturbatively. This analysis leads to $`{\displaystyle \frac{1}{\tau _e}}\mathrm{\Gamma }_e`$ $``$ $`\mathrm{\Gamma }_e^{(0)}(V_y/E_c)^{2\mathrm{\Delta }_e2}`$ $`{\displaystyle \frac{1}{\tau _h}}\mathrm{\Gamma }_h`$ $``$ $`\mathrm{\Gamma }_h^{(0)}(V_y/E_c)^{2\mathrm{\Delta }_h2}`$ (2) where $`\mathrm{\Gamma }_e^{(0)}`$ and $`\mathrm{\Gamma }_h^{(0)}`$ are Golden-rule scattering rates at the characteristic microscopic energy scale $`E_c`$, Here $`\mathrm{\Delta }_e`$ is the scaling dimension of the operator which scatters an electron across a stripe, which we discuss in the next section, and $`\mathrm{\Delta }_h`$ is the scaling dimension of the operator which scatters an electron between neighboring stripes. The values of $`\mathrm{\Delta }_e`$ and $`\mathrm{\Delta }_h`$ depend on correlations induced by electron-electron interaction between stripes, and are sensitive in particular to the range of the microscopic electron-electron interaction. At $`\nu ^{}=1/2`$, the case on which we will concentrate, $`\mathrm{\Delta }_e=\mathrm{\Delta }_h`$. Given these expressions, it follows that the non-linear differential resistivity in the hard direction $$\frac{V_y}{I_y}I_y^\alpha ,$$ (3) with an exponent $`\alpha =2(1\mathrm{\Delta }_e)/(2\mathrm{\Delta }_e1)`$. Similar considerations apply for the easy direction current: $$\frac{V_x}{I_x}I_x^\beta $$ (4) with an exponent $`\beta =2(\mathrm{\Delta }_e1)`$. In the next section we show that both $`\alpha `$ and $`\beta `$ increase when the distance to the screening plane is comparable to or smaller than the CDW period. ## 3 Quantum Smectic Model The CDW state of Hartree-Fock theory is a single-Slater-determinant. In the valence Landau level, groups of Landau-gauge single-particle states with adjacent $`k_x`$ (adjacent $`\widehat{y}`$) are occupied to form stripes and separated by groups which are unoccupied. Small fluctuations in the positions and shapes of the stripes can be described in terms of particle-hole excitations near the stripe edges. The residual electron-electron interaction terms which scatter into these low energy states are ignored in Hartree-Fock theory and fall into two classes: “forward” scattering interactions which conserve the number of electrons on each edge of every stripe, and “backward” scattering processes which do not. The latter processes involve large momentum transfer and are unimportant at accessible temperatures for $`\nu ^{}`$ near $`1/2`$. The quantum smectic model , briefly described in this section includes forward scattering only. The interactions are bilinear in the 1D electron density contributions from a particular edge of a particular stripe: $`\rho _{n\alpha }(x)`$, with $`\alpha =\pm `$. Since the density of a single filled Landau level is $`(2\pi \mathrm{}^2)^1`$, the displacement of an edge is related to its associated charge density contribution by $`u_{n\alpha }(x)=\alpha 2\pi \mathrm{}^2\rho _{n\alpha }(x)`$. The quadratic Hamiltonian which describes the classical energetics for small fluctuations has the following general form: $`H_0`$ $`=`$ $`{\displaystyle \frac{1}{2\mathrm{}^2}}{\displaystyle _{x,x^{}}}{\displaystyle \underset{n,n^{}}{}}u_{n\alpha }(x)D_{\alpha \beta }(xx^{};nn^{})u_{n^{}\beta }(x^{})`$ (5) $`=`$ $`{\displaystyle \frac{1}{2\mathrm{}^2}}{\displaystyle _𝐪}u_\alpha (𝐪)D_{\alpha \beta }(𝐪)u_\beta (𝐪),`$ where $`_𝐪d^2𝐪/(2\pi )^2`$. Here the $`q_y`$ integral is over the interval $`(\pi /a,\pi /a)`$ and a high momentum cutoff $`\mathrm{\Lambda }1/\mathrm{}`$ is implicit for $`q_x`$. Symmetry considerations constrain the form of the elastic kernel. In position space, the kernel must be real and symmetric so that $`D_{\alpha \beta }(𝐪)=D_{\alpha \beta }^{}(𝐪)=D_{\beta \alpha }^{}(𝐪)`$. This implies $`D_+(𝐪)=D_+^{}(𝐪)`$ and $`\mathrm{Im}D_{\alpha \alpha }(𝐪)=0`$. Parity invariance (under $`x,n,+x,n,`$), implies moreover $`D_{++}(𝐪)=D_{}(𝐪)`$. Thus, the elastic kernel is fully specified by one real function, $`D_{++}(𝐪)`$, and one complex function, $`D_+(𝐪)`$. It will be important for our present interest that the Hamiltonian must be invariant under: $`u_{n\alpha }(x)u_{n\alpha }(x)+const`$. For short-range interactions this implies that at long wavelengths $$D(q_x=0,q_y)=K_yq_y^2+\mathrm{}.,$$ (6) characteristic of classical smectic elasticity. As we will discuss, this conclusion must be modified in the case of long-range interactions. A quantum theory of the Quantum-Hall smectic is obtained by imposing Kac-Moody commutation relations on the chiral densities: $$[\rho _{n\alpha }(x),\rho _{n^{}\beta }(x^{})]=\frac{i}{2\pi }\alpha \delta _{\alpha ,\beta }\delta _{n,n^{}}_x\delta (xx^{}).$$ (7) Together with Eq.( 5), this relationship fully specifies the quantum dynamics. Electron operators in the chiral edge modes are related to the 1D densities via the usual bosonic phase fields: $`\psi _{n\alpha }e^{i\varphi _{n\alpha }}`$ with $`\rho _{n\alpha }=\alpha _x\varphi _{n\alpha }/2\pi `$. Quantum properties of the smectic can be computed from the imaginary-time action, $`S_0`$ $`=`$ $`{\displaystyle _{x,\tau }}{\displaystyle \frac{1}{4\pi }}{\displaystyle \underset{n,\alpha }{}}i\alpha _\tau \varphi _{n,\alpha }_x\varphi _{n,\alpha }+{\displaystyle _\tau }H_0`$ (8) $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _{𝐪,\omega }}\varphi _\alpha (𝐪,\omega )M_{\alpha ,\beta }(𝐪,\omega )\varphi _\beta (𝐪,\omega ),`$ where in an obvious matrix notation, $$𝐌(𝐪,\omega )=(i\omega q_x/2\pi )\sigma ^z+(q_x\mathrm{})^2𝐃(𝐪).$$ (9) Correlation functions follow from Wick’s theorem and the momentum space correlator $`\varphi _\alpha \varphi _\beta =𝐌^1`$ with $$𝐌^1(𝐪,\omega )=\sigma _z𝐌(𝐪,\omega )\sigma _z/\mathrm{det}𝐌(𝐪,\omega ).$$ (10) The effect of weak disorder on transport in quantum Hall smectics depends sensitively on the elastic constants at $`q_x=0`$. In this limit the relevant excited states are simply Slater determinants with straight stripe edges displaced from those of the Hartree-Fock theory ground state. By evaluating the expectation value of the microscopic Hamiltonian in a state with arbitrary stripe edge locations we find that $$D_{\alpha \beta }(q_x=0,q_y)=\delta _{\alpha \beta }D_0+\alpha \beta \frac{a}{4\pi ^2\mathrm{}^2}\underset{n}{}e^{iq_yan}\mathrm{\Gamma }(y_{n\alpha }^0y_{0\beta }^0),$$ (11) where the constant $`D_0`$ is such that the condition $`_{\alpha \beta }D_{\alpha \beta }(𝐪=0)=0`$, and the positions $`y_{n\pm }^0=a(n\pm \nu ^{}/2)`$ are the ground state stripe edge locations. Here, $`\mathrm{\Gamma }(y)`$ is the interaction potential between two electrons located in guiding center states a distance $`y`$ apart: $$\mathrm{\Gamma }(y)=U(0,y/\mathrm{}^2)U(y/\mathrm{}^2,0),$$ (12) $$U(q,k)=\frac{dp}{2\pi }e^{(q^2+p^2)\mathrm{}^2/2}V_{\mathrm{eff}}^N(q,p)e^{ipk\mathrm{}^2}.$$ (13) The two terms in Eq. (12) are direct and exchange contributions. In Eq. (13), $`V_{\mathrm{eff}}^N(q,p)`$ is the Fourier transform of the effective 2D electron interaction which incorporates form-factors dependent on the Landau level index. $`N`$ and the ground subband wavefunction of the host semiconductor heterojunction or quantum well. The smectic states have relatively long periods proportional to the cyclotron orbit radii. Explicit calculations show that $`a>6\mathrm{}`$ for $`N=2`$. It follows that the exchange contribution to $`\mathrm{\Gamma }(y)`$ is small and that $`\mathrm{\Gamma }(y)`$ decreases with stripe separation in the relevant range. In this paper we address the influence of a metallic screening plane which cuts off this interaction at large distances. For $`y<d`$, $`\mathrm{\Gamma }(y)2e^2\mathrm{ln}(d/y)`$, decreasing extremely slowly with $`y`$. For separation $`y`$ larger than the distance $`d`$ to the screening plane, $`\mathrm{\Gamma }(y)y^2`$, making the sum over $`n`$ in Eq. 11 convergent. ## 4 Screening Dependence of Scaling Dimensions The scaling dimension, $`\mathrm{\Delta }_e`$, of the operator $`e^{i(\varphi _{n,+}\varphi _{n,})}`$ which scatters an electron across the $`n`$-th stripe is readily evaluated from Eq.( 9). We find that $$\mathrm{\Delta }_e=_\pi ^\pi \frac{d(qa)}{2\pi }W(q_x=0,q).$$ (14) Here, $`W`$ is the weight function, $$W(𝐪)=\frac{[D_{++}(𝐪)+\mathrm{Re}D_+(𝐪)]}{[D_{++}^2(𝐪)|D_+(𝐪)|^2]^{1/2}}.$$ (15) For 1D non-interacting electrons $`\mathrm{\Delta }_e=1`$, so that disorder is relevant and eventually leads to localization. As discussed below, $`\mathrm{\Delta }_e<1`$ for quantum Hall smectics. Disorder is even more relevant than in the non-interacting electron case. Nevertheless, since the samples in which the quantum Hall smectic is observed are of extremely high quality, there should be a wide range of temperature over which its effects can be treated perturbatively. If $`\mathrm{\Delta }_e=1`$ both hard direction ($`\alpha `$) and easy direction ($`\beta `$) non-linear transport exponents vanish. We see from Eq.( 15) that $`\mathrm{\Delta }_e=1`$ if the average value of $`W(𝐪_𝐲)`$ is one. To understand the dependence of $`\mathrm{\Delta }_e`$ on screening, we have to understand the dependence of $`W(q_y)`$ on both wavevector and $`d`$. Note that $`W`$ is smaller than one, increasing the relevance of disorder, when $`D_{++}`$ and $`\mathrm{Re}D_+`$ are opposite in sign and similar in magnitude. For each $`q_y`$ in Eq. 14 the weighting factors are like those which enter in the calculation of the scaling dimension of the operator which describes backscattering from disorder in an isolated one-dimensional electron system. In continuum 1D models, $`D_{++}`$ has a contribution, proportional to the Fermi velocity, from the band energy and a contribution proportional to the interaction between electrons traveling in the same direction, while $`D_+`$ has only an interaction contribution. For a continuum model, the effective interactions between electrons traveling in different directions is the same as that between electrons traveling in the same directions. When the interaction term is much larger than the band term $`D_{++}`$ and $`D_+`$ are opposite in sign and nearly equal in magnitude and $`W`$ is very small. This is what happens, for example, for a 1D electron system in which long-range makes the Coulomb interaction very strong at long wavelengths. When $`W`$ is small, the 1D electron system is very close to an electron Wigner crystal, and disorder is very strongly relevant. On the other hand when $`D_+`$ is much smaller than $`D_{++}`$ we have a situation analogous to that in a very weakly interacting Fermion system, in which disorder is relevant but the resistivity is linear when disorder can be treated perturbatively. With this in mind we turn to a discussion of the quantum smectic, limiting our attention to the case $`\nu ^{}=1/2`$. Useful insight comes from examining the value of $`W`$ at the end points of the integration interval, $`q_ya=0`$ and $`q_ya=\pi `$ where both $`D_{++}`$ and $`D_+`$ are real. For $`q_y=0`$, invariance under a uniform translation of the smectic implies that $`D_{++}+D_+=0`$, so that $`W(q_y=0)=0`$. When all the electron stripes move together, the energetics is precisely like that of a single 1D system. In the quantum Hall regime, there is no band energy, only interaction contributions from electrons traveling in the same direction, which appear in $`D_{++}`$, and interaction contributions from electrons traveling in opposite directions, which appear in $`D_+`$. The absence of a band contribution means that, for $`q_y=0`$, $`W`$ vanishes independent of the interaction’s strength or range. When $`q_ya=\pi `$, on the other hand, one has $$D_+(q_ya=\pi )=\underset{n}{}\frac{(1)^na}{4\pi ^2\mathrm{}^2}[\mathrm{\Gamma }(an+a(1\nu ^{}))\mathrm{\Gamma }(an+a\nu ^{})].$$ (16) We see that $`D_+(q_ya=\pi )`$ vanishes because the interaction between the top of one stripe and the bottom of the same stripe is identical to its interaction with the bottom of the next stripe up. The interactions between oppositely directed electrons effectively cancel out, and we obtain $`W=1`$, just as we would for a non-interacting 1D system. Thus $`\mathrm{\Delta }_e`$ is determined by the average over $`q_y`$ of a weighting factor which interpolates between that of a 1D electron Wigner crystal at $`q_ya=0`$ and that of a non-interacting 1D electron system at $`q_ya=\pi `$. The average value of $`W`$ is determined by the rate at which $`W`$ goes from its $`q_y=0`$ limit to its $`q_y=\pi `$ limit. To obtain insight into what controls this, we consider first the limit of short-range interactions. Since $`\mathrm{\Gamma }(y=0)`$ vanishes due to the cancellations of its direct and exchange contributions, the short-range limit is obtained by taking only $`\mathrm{\Gamma }(a/2)\mathrm{\Gamma }^{}0`$. In this case, it follows from Eq.( 11) that $`(4\pi ^2\mathrm{}^2)D_+(q)/a=\mathrm{\Gamma }^{}(1+\mathrm{exp}(iqa))`$ and that $`(4\pi ^2\mathrm{}^2D_{++}(q)/a=2\mathrm{\Gamma }^{}`$, and therefore that $`W(q)=|sin(qa/2)|`$. The numerator of Eq.( 15) vanishes like $`q^2`$ for $`q0`$, and the denominator, which is proportional to the collective mode velocity, vanishes like $`|q|`$. Note that the expression for $`W(q)`$ is independent of $`\mathrm{\Gamma }^{}`$. The average of $`W(q)`$ may be evaluated analytically in this case, and we obtain for short-range interactions $`\mathrm{\Delta }_e=2/\pi 0.6366`$. For the realistic case, analytic calculations are no longer possible, but the behavior of $`W`$ can be simplified, at least at small $`q`$ if the exchange contribution to $`\mathrm{\Gamma }(y)`$ is neglected in constructing $`D_{\alpha ,\beta }(q)`$. In this case $`\mathrm{\Gamma }(y)`$ is the simply the 1D transform to coordinate space of the reciprocal space interaction $`U(p)\mathrm{exp}(p^2\mathrm{}^2/2)V_{eff}^N(0,p)`$, i.e. it is the Coulomb interaction between lines of charged smeared by $`N`$-dependent cyclotron orbit form factors. The components of $`D_{\alpha ,\beta }(q)`$, are then Fourier transforms back to reciprocal space, but with additional ‘umklapp’ terms because this transform is discrete. We find that $`D_+(q)`$ $`=`$ $`{\displaystyle \frac{1}{4\pi ^2\mathrm{}^2}}\mathrm{exp}(iqa/2){\displaystyle \underset{j=\mathrm{}}{\overset{\mathrm{}}{}}}()^jU(q+2\pi j/a)`$ $`D_{++}(q)`$ $`=`$ $`D_0+{\displaystyle \frac{1}{4\pi ^2\mathrm{}^2}}{\displaystyle \underset{j=\mathrm{}}{\overset{\mathrm{}}{}}}U(q+2\pi j/a)`$ (17) and $$D_0=\frac{1}{4\pi ^2\mathrm{}^2}\underset{j=\mathrm{}}{\overset{\mathrm{}}{}}U(2\pi j/a)[()^j1].$$ (18) For the case of a Coulomb interaction, $`U(q)=[2\pi e^2(1\mathrm{exp}(2qd))]/q`$ where $`d`$ is the distance to a screening plane described by an image charge model. For large or infinite $`d`$, the $`j=0`$ terms dominate the sums in Eqs. (17). The above expression for $`U(q)`$ applies when $`q\mathrm{}<1`$, and therefore is always valid for the $`j=0`$ terms in Eqs.( 17). Note that both $`D_{++}(q)`$ and $`D_\pm (q)`$ are proportional to $`d`$ for large $`d`$ and diverge for $`d\mathrm{}`$. On the other hand $`D_0`$ remains finite for $`d\mathrm{}`$ because the $`j=0`$ term is excluded from this sum. We emphasize that the numerical calculations whose results are shown below include the exchange contributions neglected in deriving Eqs.( 17), and important at any value of $`a`$ when $`d`$ is not large. We now examine the large $`d`$, small $`q`$ behavior of the numerator and denominator of Eq.( 15). For the numerator $$D_{++}(q)+\mathrm{Re}D_+(q)=\frac{1}{4\pi ^2\mathrm{}^2}U(q)(1cos(qa/2))\frac{e^2}{2\pi \mathrm{}^2}q^2da^2/2,$$ (19) and for the denominator $$D_{++}(q)^2|D_+(q)|^2U(q)[d][q^2a^2].$$ (20) The first factor is square brackets in Eq.( 20) comes from $`D_{++}(q)+|D_+(q)|`$ for which the $`j=0`$ term dominates. The second factor is proportional to $`D_{++}(q)|D_+(q)|`$ for which only odd $`j`$ terms survive, implying no dependence on $`d`$ and analytic dependence on $`q`$. These formulas apply for $`qd<1`$; for $`d\mathrm{}`$, the small $`q`$ behavior is obtained by replacing $`U(q=0)=4\pi e^2d`$ by $`U(q0)=2\pi e^2/q`$, i.e., by replacing $`d`$ by $`1/2q`$. We plot the square of the denominator, proportional to the square of the collective excitation velocity for several values of the screening length $`d`$ in Fig. . The velocity increases as $`d`$ increases as expected. For $`d=100\mathrm{}`$, the quadratic small $`q`$ behavior predicted in Eq.( 20), applies only for $`qa<0.02`$, and is not apparent in the plot. Instead we see the long range interaction behavior, in which the velocity is proportional to $`q^{1/2}`$. As is apparent from Eqs.( 17), screening is irrelevant except very close to $`q=0`$ when $`d`$ is large. Fig. shows that once $`d>a`$, the large $`d`$ no-screening limit is approached. In Fig. we show the weight functions for $`d=\mathrm{}`$, $`d=10\mathrm{}`$ and $`d=100\mathrm{}`$. At each value of $`d`$, the denominator of the weight function at small $`q`$ is proportional to $`d^{1/2}|q|`$ and the numerator proportional to $`dq^2`$. The weight function is therefore proportional to $`|q|`$ with a coefficient which varies as $`d^{1/2}`$. A larger value of $`d`$ (less screening), leads to a weight function which increases more rapidly with $`q`$ and a scaling dimension for the scattering vertex which is closer to one. In the limit of unscreened interactions, which applies down to small $`q`$ for $`d=100\mathrm{}`$, $`W(q)|q|^{1/2}`$ In Fig. we show the dependence of the scaling dimension $`\mathrm{\Delta }_e`$ and the non-linear transport exponents $`\alpha `$ and $`\beta `$ on the distance $`d`$ to the screening plan. For $`d0`$ the numerical result is very close to that from the analytic short-range interaction model described above which leads to $`\mathrm{\Delta }_e=2/\pi `$. For $`d0`$, the scaling dimension approaches $`\mathrm{\Delta }_e=0.772`$, a value we have been able to obtain only numerically. These relatively modest changes in the scaling dimension translate into relatively large changes in the transport exponents, particularly in $`\alpha `$ which characterizes the hard-axis non-linearity. We predict that these non-linearities will be much stronger if a screening placed in close proximity to the electron layer. ## 5 Summary Recent experiments have established a consistent set of transport properties for high-mobility two-dimensional electron systems with high orbital index ($`N2`$) partially filled Landau levels. These properties differ qualitatively from those which occur in the low orbital index ($`N1`$) fractional quantum Hall effect regime. At large $`N`$, the dissipative resistivities are large, strongly anisotropic, and non-linear for $`0.4<\nu [\nu ]<0.6`$ within each Landau level. This anisotropic transport regime is bracketed by regions of reentrant integer quantum Hall plateaus. We have recently developed a theory which is able to account for most features of these experiments. An important prediction of the theory is that the easy and hard direction resistivities should have non-integral power-law temperature dependences. In this article we have briefly summarized the theory and elaborated on its predictions for the dependence of these exponents on the distance $`d`$ between the two-dimensional electron layer and a remote screening plane, predictions which are summarized in Fig. . We find that $`\mathrm{\Delta }_e`$ approaches two different values, both smaller than one, for $`d0`$ and $`d\mathrm{}`$, and interpolates smoothly between these limits at finite values of $`d`$. Verification of our prediction that non-linear transport can be enhanced by introducing a screening plane and reducing $`d`$, would help substantiate our theory of quantum Hall smectics. ## 6 Acknowledgements We would like to acknowledge insightful conversations with Herb Fertig, Michel Fogler, Tomas Jungwirth, Jim Eisenstein, Eduardo Fradkin, Steve Girvin, and Philip Phillips. AHM acknowledges the hospitality of the ITP at UC Santa Barbara, where this work was initiated. M.P.A.F. is grateful to the NSF for generous support under grants DMR-97-04005, DMR95-28578 and PHY94-07194. A.H.M. is grateful for support under grant DMR-97-14055.
warning/0001/quant-ph0001004.html
ar5iv
text
# Bibliographische Beschreibung ##### Bibliographische Beschreibung Kilin, Dmitri: ”Die Rolle der Umgebung in molekularen Systemen.” Dissertation, Technische Universität Chemnitz, Chemnitz 1999 98 Seiten, 26 Abbildungen, 4 Tabellen. ##### Referat Die Dissipation von Energie von einem molekularen System in die Umgebung und die damit verbundene Zerstörung der Phasenkohärenz hat einen Einfluss auf mehrere physikalische Prozesse wie Bewegung der Schwingungsmoden eines Moleküls, eines Ions in einer Falle oder einer Strahlungsfeldmode, sowie auf Exzitonen- und Elektronentransfer. Elektronentransfer spielt eine wichtige Rolle in vielen Bereichen der Physik und Chemie. In dieser Arbeit wird die Elektronentransferdynamik mit Bewegungsgleichungen für die reduzierte Dichtematrix beschrieben, deren Herleitung ausgehend von der Liouville–von Neumann-Gleichung über die Kumulanten-Entwicklung führt. Durch Ankopplung an ein Wärmebad werden dissipative Effekte berücksichtigt. Zunächst wird diese Theorie auf Modellsysteme angewendet, um die verschiedenen Einflüsse der Umgebung auf Depopulation, Dephasierung und Dekohärenz besser zu verstehen. Dann wird die Dynamik von konkreten intramolekularen Transferreaktionen in realen Molekülen berechnet und die Ergebnisse mit denen von Experimenten und anderer Theorien verglichen. Zu den untersuchten Systemen zählen die Komplexe $`\mathrm{H}_2\mathrm{P}`$ $`\mathrm{ZnP}`$ $`\mathrm{Q}`$ und $`\mathrm{ZnPD}`$ $`\mathrm{H}_2\mathrm{P}`$. ##### Schlagwörter Elektronentransfer, Moleküle, Transferraten, Dichtematrixtheorie, Wärmebad, Dissipation, Relaxation, thermisch aktivierter Transfer, vibronische Zustände, Marcus-Theorie, Superaustausch. Dmitri S. Kilin The Role of the Environment in Molecular Systems ## List of Abbreviations | A | acceptor | | --- | --- | | B | bridge | | CYCLO | cyclohexane | | D | donor | | DGME | differential generalized master equation | | DM | density matrix | | ET | electron transfer | | GME | generalized master equation | | GSLE | generalized stochastic Liouville equation | | $`\mathrm{H}_2\mathrm{P}`$ | free-base porphyrin | | HO | harmonic oscillator | | HOMO | highest occupied molecular orbital | | HSR | Haken, Reineker, Strobl | | IGME | integrodifferential generalized master equation | | JCM | Jaynes-Cummings model | | LUMO | lowest unoccupied molecular orbital | | MTHF | methyltetrahydrofuran | | Q | quinone | | RDM | reduced density matrix | | RDMEM | reduced density matrix equation of motion | | RWA | rotating wave approximation | | SLE | stochastic Liouville equation | | TB | tight-binding | | TLS | two level system | ## Chapter 1 Introduction The behavior of many quantum systems strongly depends on their interaction with the environment. The dissipative processes induced by interaction with the environment have a broad area of applications from isolated molecules to biomolecules. In order to achieve a realistic description of a molecular process, it is important to take the dissipation into account for systems like, e.g., vibrational levels in a big molecule, the quantized mode of an electromagnetic field, or a trapped ion. The interaction of a system with an environment also plays the main role in the modeling of electron transfer (ET) processes as it ensures their irreversibility. ET is a very important process in biology, chemistry, and physics . It constitutes a landmark example for intramolecular, condensed-phase, and biophysical dissipative dynamics. ET plays a significant role in nature in connection with conversion of energy. In the photosynthetic reaction center, ET creates charge imbalance across the membrane, which drives the proton pumping mechanism to produce adenosine triphosphate. In chemical systems, surface ET between metals and oxygen is responsible for corrosion processes. In organic chemistry, mechanisms involving bond fracture or bond making often proceed by ET mechanism. In inorganic chemistry, mixed-valence systems are characterized by ET between linked metal sites. Finally, the nascent area of molecular electronics depends, first and foremost, on understanding and controlling the ET in specially designed chemical structures. That is exactly why the ET problem is the main topic of this work. Of special interest is the ET in configurations where a bridge between donor and acceptor mediates the transfer. The primary step of the charge transfer in the bacterial photosynthetic reaction centers is of this type , and a lot of work in this direction has been done after the structure of the protein-pigment complex of the photosynthetic reaction center of purple bacteria was clarified in 1984 . Many artificial systems, especially self-organized porphyrin complexes, have been developed to model this bacterial photosynthetic reaction centers . The bridge-mediated ET reactions can occur via different mechanisms : incoherent sequential transfer when the mediating bridge level is populated or coherent superexchange when the mediating bridge level is not populated but nevertheless is necessary for the transfer. In the case of the sequential transfer the influence of environment has to be taken into account. Apart from these aspects, one of the fundamental questions of quantum physics has attracted a lot of interest: why does the general principle of superposition work very well in microscopic physics but leads to paradox situations in macroscopic physics as for instance the Schrödinger cat paradox . One possible explanation of the paradox and the non-observability of the macroscopic superposition is that systems are never completely isolated but interact with their environment . Interactions with the environment lead to continuous loss of coherence and drive the system from superposition into a classical statistical mixture. The question about the border between classical and quantum effects and systems, which model this problem, are also under considerations in this work. The interest in the decoherence problem is explained not only by its relation to the fundamental question: ”Where is the borderline between the macroscopic world of classical physics and microscopic phenomena ruled by quantum mechanics?”, but also by the increasing significance of potential practical applications of quantum mechanics, such as quantum computation and cryptography . The rapid development of experimental techniques in the above-mentioned and other branches of physics and chemistry requires to describe, to model, and to analyse possible experiments by numerical and analytical calculations. The mathematical description of the influence of environment for all these examples has attracted a lot of interest but remains a quite complicate problem nowadays. The theory has been developed in recent years and a brief review of its progress is represented in the next chapter of this work. Despite of the intensive attention and investigations of this problem there is a necessity to consider basic model concepts in more detail in order to apply the mathematical techniques to real physical systems, which are studied experimentally in an appropriate way. In the present work we throw a glance on the known principles of the relaxation theory and we are mainly interested in the application of the relaxation theory to some concrete systems, such as a single vibrational mode modelled by a harmonic oscillator (HO), an artificial photosynthetic molecular aggregate, and a porphyrin triad, using simulations and numerical calculations as well as some analytical methods. The questions of the influence of environment on these systems are discussed in detail in this work. Our theoretical arsenal is based on the relaxation theory of dissipative processes containing calculations of coherent effects for electronic states and wave-packet dynamics, which provide the conceptual framework for the study of ET and decoherence. The basic concepts of the relaxation theory, based on the density matrix formalism, are reviewed in chapter 2. This technique will be used throughout this work. Chapter 3 deals with the question of the border between classical and quantum effects and reports on a study of the environmental influence on the time evolution of a coherent state or the superposition of two coherent states of a HO as a simple system displaying the peculiarities of the transition from quantum to classical regime. Chapters 4 and 5 concern the ET problem, namely the mathematical description of the ET in molecular zinc-porphyrin-quinone complexes modeling artificial photosynthesis (chapter 4) and photoinduced processes in the porphyrin triad (chapter 5). Each chapter starts with an introduction and ends with a brief summary. The main achievements of the present work are summarized in the Conclusions. ## Chapter 2 Reduced Density Matrix Method The goal of this chapter is to introduce the reader to the main mathematical tools for calculating the system dynamics induced by the interaction with an environment, which are used in all parts of this work. It should be noted that usually an environment is modelled by a heat bath, thus here and below we use environment and bath as synonyms for each other. The chapter starts with a review of historical developments of the theoretical models for dissipation processes and a brief consideration of various types of master equations, their characteristics, and different techniques used to describe these models (section 2.1). Using the Hamiltonian for system plus bath in the common form (section 2.2), the Green’s matrix technique (section 2.3), and the cumulant expansion method (section 2.4) we arrive at a differential form of the generalized master equation (GME) (secion 2.5), which is applied in the later chapters for the description of particular systems. Although some steps of this derivation are known in the literature we present them here in order to be rigorous and to reach completeness of notation in this work. ### 2.1 Theoretical models for dissipation In the first quantum consideration of the system “atom + field” Landau has introduced an analog of the density matrix (DM), its averaging over the field states and its equation of motion. The rigorous introduction of the statistical operator and its equation of motion has been done by von Neumann in the early 30s . The first derivation of the master equations based on the Liouville equation for the system “atom + field + environment” were closely connected to the radio frequency range and to the saturation of signals in nuclear magnetic resonance . For those systems the projector operator technique has been used in order to derive an exact integro-differential master equation . One started to describe optical processes with the GME a bit later because in the 50s and earlier 60s there were still no experimental hints to the non-Markovian nature of the relaxation for optical processes. Nowadays a quantum dissipation theory has been much sought after as the goal of five communities: quantum optics , condensed matter physicists , mathematical physicists , astrophysicists , and condensed phase chemical physicists . Theories of quantum dissipation can be divided into three main classes. The first class begins with a full system + bath Hamiltonian and then projects the dynamics onto a reduced subspace. Notable examples of this approach are the path integral approach of Feynman and Vernon , Redfield theory , and the projection operator technique of Nakajima and Zwanzig . As also mentioned by Pollard and Friesner this theory can be divided into two subclasses: quantum and classical bath. At the opposite extreme the second class begins with linear equations of motion for the reduced density matrix (RDM) and then deduces a form for the equations of motion compatible with relevant characteristics of the relaxation theory. Examples of this approach are the semigroup approach of Lindblad and the Gaussian ansatz of Yan and Mukamel . An intermediate procedure is to describe the bath as exerting a fluctuating force on the system. This approach is often used in the laser physics community, for example by Agarwal , Louisell , Gardiner and others. This class of theory was also used for the exciton transfer description . Pollard and Friesner denote this class as “stochastic bath theory”. In all mentioned theories the system is described with the RDM which evolves in time under master equations. It is known that the GME can be obtained by the following methods: (i) Decoupling of the relaxation perturbations and the DM of a full system , (ii) averaging over the fast motions taking into account the hierarchy of characterisctic times , (iii) coupled multiparticle Bogolyubov equations , (iv) Nakajima-Zwanzig projection operators technique , (v) the diagrammatic technique , (vi) cumulant expansions , (vii) the method of Green’s functions for the full system, averaged over the realizations of the environment , and (viii) stochastic models . The GMEs can be divided into two groups: integrodifferential GMEs (IGME), e.g., derived using Nakajima-Zwanzig projection operators techniques or differential GME (DGME) often derived with the help of the cumulant expansion technique . In most cases the IGMEs are transformed into the DGME at the outset, based on the assumption that the DM varies slowly over the bath correlation times . In linear optics the question of the relationship between IGMEs and DGMEs was discussed in . In the theory of relaxation it was shown (neglecting the effect of radiation on the relaxation) that IGME can also be constructed with the help of the cumulant expansion . It is generally assumed that a fully satisfactory theory of quantum dissipation should have the following characteristics: (1) The RDM should remain positive semidefinite for all time (i. e., no negative eigenvalues, which would in turn imply negative probabilities). Below we use the word “positivity” in order to point out this property. (2) The RDM should approach an appropriate equilibrium state at long times. (3) The RDM should satisfy the principle of translational invariance, if a coordinate and translation are defined. This condition requires that the frictional force be independent of the coordinate as is generally the case in the classical theory of Brownian motion. In the Markov approximation any GME can be reduced to one of the following types: Redfield (R), Agarwal (A), Caldeira-Leggett (C), Louisell-Lax (Lou), and Lindblad (Li). Now we simply list these types of master equations together with their characteristics. R. In the derivation of Redfield one uses a quantum bath theory to model the environment. In the appropriate basis this master equation coincide with the equation of Agarwal’s type. This equation is not of Lindblad form , while in the Pollard and Friesner parametrisation the two types of master equations seem to be similar. The positivity of the DM which evolves under Redfield equation is violated for large system-bath coupling. At infinite time the DM reaches the thermal state of the bare system. The Redfield equation satisfies the requirement of the translational invariance. A. In the derivation of Agarwal’s master equation one models the environment with a stochastic bath. In the relevant basis this master equation coincides with the equation of Redfield type and does not coincides with Lindblad form. The positivity of the DM which evolves under Agarwal’s equation is violated for large system-bath coupling. At infinite time the DM reaches the thermal state for the bare system. The Agarwal equation satisfies the requirement of translational invariance. System frequencies are modulated because of momentum and coordinate relax in different ways. The approach of Yan and Mukamel ensures the same properties as theories by Redfield and Agarwal. C. In the derivation of the master equation of Caldeira-Leggett type one models the environment with a quantum bath. One derives such a master equation with the path integral technique. A master equation of Caldeira-Leggett type is compatible with the equations of Redfield and Agarwal only in the high temperature limit. This equation is not of Lindblad form. The positivity of the DM is violated in this equation. The DM arrives at the equilibrium state only in the high temperature limit. The Caldeira-Legget master equation satisfies translational invariance requirement. Lou. In the Louisell-Lax approach one accounts for the environment with a quantum bath. It differs from Redfield and Agarwal equations by performing the rotating wave approximation (RWA) and agrees with Lindblad form. So the positivity of DM is maintained in this approach. At infinite time the DM reaches the thermal state of the bare system. In the equations of the Louisell-Lax type translational invariance is violated, because there is a coordinate dependent friction force. The system frequencies are constant in time. Li. The Lindblad master equation are constructed in a special form to conserve the positivity of the RDM. In the same basis this master equation coincides with the equation of Louisell-Lax type. At infinite time the DM reaches the thermal state, but the translational variance is violated. In order to preserve the translational invariance Lindblad has included additional terms into the Hamiltonian and into the master equation. In that case the RDM at large times does not approach the equilibrium state of the bare system but some other state. This state is expected to be a projection of the equilibrium state of system plus bath onto the system subspace . In our contribution we start with the demonstration of the derivation of the GME in differential form using the cumulant expansion technique. Afterwards, in chapter 3 we arrive at a master equation of Agarwal type. In chapter 4 we use the GME in the Louisell-Lax form. The derived master equation with or without RWA is extensively used in all parts of this work. ### 2.2 Hamiltonian and density matrix In the common form the Hamiltonian “system + environment” can be written as $$H=H^\mathrm{S}+H^\mathrm{E}+H^{\mathrm{SE}}$$ (2.1) where $`H^\mathrm{S}=\underset{\mu }{}E_\mu V_{\mu \mu }+\mathrm{}\underset{\mu \nu }{}v_{\mu \nu }V_{\mu \nu }`$ represents a quantum system in the diabatic representation, $`E_\mu `$ the energy of the diabatic state $`\mu `$, $`H^\mathrm{E}=\underset{\xi }{}\mathrm{}\omega _\xi \left(b_\xi ^+b_\xi +\frac{1}{2}\right)`$ a bath of HOs, and $`H^{\mathrm{SE}}=\mathrm{}\underset{\mu \nu }{}\left(r_{\mu \nu }+r_{\mu \nu }^+\right)`$ $`\left(V_{\mu \nu }^++V_{\mu \nu }\right)`$ the linear interaction between them. Here $`V_{\mu \nu }=|\mu \nu |`$ is the transition operator of the system, $`v_{\mu \nu }`$ the coupling of the diabatic states $`\mu `$ and $`\nu `$, $`r_{\mu \nu }=\underset{\xi }{}𝒦_{\mu \nu }^\xi b_\xi `$ the generalized annihilation operator of the bath, $`b_\xi `$ the annihilation operator of the bath mode $`\xi `$ having the frequency $`\omega _\xi `$, $`𝒦_{\mu \nu }^\xi `$ the frequency-dependent interaction constant. Note, that rather often one factorizes the interaction constant on system and bath contributions. The DM of system plus bath $`\rho `$ evolves under the von Neumann equation $`\dot{\rho }=i/\mathrm{}[H,\rho ]`$. The RDM $`\sigma =\mathrm{Tr}^\mathrm{E}\rho `$ is obtained by tracing out the environmental degrees of freedom . The coherent and dissipative dynamics of the system is described by the following equation $$\dot{\sigma }=i/\mathrm{}[H^\mathrm{S},\sigma ]i/\mathrm{}\mathrm{Tr}^\mathrm{E}\left([H^{\mathrm{SE}},\overline{\rho }]\right)=i/\mathrm{}[H^\mathrm{S},\sigma ]+\mathrm{Tr}^\mathrm{E}\left(\dot{\overline{\rho }}\right),$$ (2.2) where $`\overline{\rho }`$ denotes the DM $`\rho `$ in the Heisenberg picture $`\overline{\rho }=\mathrm{exp}\left(i/\mathrm{}H^\mathrm{E}t\right)`$ $`\rho `$ $`\mathrm{exp}\left(i/\mathrm{}H^\mathrm{E}t\right)`$ with respect to environment degrees of freedom. Then we substitute the unit operator of the form $`\mathrm{exp}(i/\mathrm{}H^\mathrm{S}t)\mathrm{exp}(i/\mathrm{}H^\mathrm{S}t)`$ before and after $`\mathrm{Tr}^\mathrm{E}\left(\dot{\overline{\rho }}\right)`$ $$\dot{\sigma }=i/\mathrm{}[H^\mathrm{S},\sigma ]+\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right)\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right)\mathrm{Tr}^\mathrm{E}\left(\dot{\overline{\rho }}\right)\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right)\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right).$$ Thus reduced density matrix equation of motion (RDMEM) takes the following form $$\dot{\sigma }=i/\mathrm{}[H^\mathrm{S},\sigma ]+\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right)\mathrm{Tr}^\mathrm{E}\left(\dot{\stackrel{~}{\rho }}\right)\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right),$$ (2.3) where tilde denotes the interaction representation $`\stackrel{~}{\rho }(t)=\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right)\overline{\rho }\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right)`$. Equation (2.3) is in Schrödinger picture with respect to $`H^\mathrm{S}`$ and in Heisenberg picture with respect to $`H^\mathrm{E}`$; nevertheless it contains the interaction dynamics in the interaction picture. This representation is convenient because the relevant von Neumann equation for system plus environment $`\dot{\stackrel{~}{\rho }}=i/\mathrm{}[\stackrel{~}{H}^{\mathrm{SE}}(t),\stackrel{~}{\rho }]`$ contains neither $`H^\mathrm{S}`$ nor $`H^\mathrm{E}`$. ### 2.3 Green’s matrix The commutator of an operator with the Hamiltonian can be represented symbolically as follows $`[\stackrel{~}{H}^{\mathrm{SE}}(t),\stackrel{~}{\rho }]\stackrel{~}{L}(t)\stackrel{~}{\rho }`$ where $`\stackrel{~}{L}(t)=\stackrel{~}{L}^{\mathrm{SE}}(t)`$ is Liouville’s operator in the interaction picture. Here we assume, that system and bath are disentangled at the initial moment of time $$\stackrel{~}{\rho }(0)=\rho (0)=\rho ^\mathrm{E}(0)\sigma (0).$$ (2.4) In all calculations below we suppose that the initial states of the bath oscillators are thermalized $`\rho _\xi (0)\mathrm{exp}(\mathrm{}\omega _\xi b_\xi ^+b_\xi /k_\mathrm{B}T)`$. In accordance with the evolution of the system is described by a Green’s matrix $$\stackrel{~}{\rho }=D^{\mathrm{SE}}(t,0)\rho (0),$$ (2.5) where $`D^{\mathrm{SE}}(t,0)\rho \left(0\right)=\mathrm{T}\mathrm{exp}\left[i/\mathrm{}\underset{0}{\overset{t}{}}𝑑\tau L(\tau )\right]`$, $`\mathrm{T}`$ stands for the time ordering operator. The ansatz of the disentanglement Eq. (2.4) allows us to decouple the evolution of the system from the evolution of the environment on the basis of the reduced Green’s matrix $`𝐃^{\mathrm{SE}}(t,0)=\mathrm{Tr}^\mathrm{E}\left[D^{\mathrm{SE}}(t,0)\rho ^\mathrm{E}(0)\right]`$ so that $$\mathrm{Tr}^\mathrm{E}\stackrel{~}{\rho }(t)=\mathrm{Tr}^\mathrm{E}\left[D^{\mathrm{SE}}(t,0)\rho ^\mathrm{E}(0)\sigma (0)\right]=𝐃^{\mathrm{SE}}(t,0)\sigma (0)$$ (2.6) The connection between $`\sigma `$ and $`\stackrel{~}{\rho }`$ reads $`\sigma =\mathrm{Tr}^\mathrm{E}\rho =\mathrm{Tr}^\mathrm{E}\stackrel{~}{\rho }`$ $`=`$ $`\mathrm{Tr}^\mathrm{E}\left[\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right)\stackrel{~}{\rho }\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right)\right]`$ (2.7) $`=`$ $`\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right)\mathrm{Tr}^\mathrm{E}\stackrel{~}{\rho }\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right).`$ Formula (2.6) ensures the following property $$(𝐃^{\mathrm{SE}})^1(t,0)\mathrm{Tr}^\mathrm{E}\stackrel{~}{\rho }=\sigma \left(0\right).$$ (2.8) Substituting Eq. (2.6) into Eq. (2.3) gives: $$\dot{\sigma }=i/\mathrm{}[H^\mathrm{S},\sigma ]+\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right)\dot{𝐃}^{\mathrm{SE}}(t,0)\sigma \left(0\right)\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right).$$ (2.9) In the last equation the whole influence of the bath is included in $`\dot{𝐃}^{\mathrm{SE}}(t,0)`$. In order to obtain a differential equation for $`\sigma `$ which is local in time we substitute the property Eq. (2.8) into Eq. (2.9) so that $$\dot{\sigma }=i/\mathrm{}[H^\mathrm{S},\sigma ]+\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right)\dot{𝐃}^{\mathrm{SE}}(t,0)\left[(𝐃^{\mathrm{SE}})^1(t,0)\mathrm{Tr}^\mathrm{E}\stackrel{~}{\rho }\right]\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right).$$ (2.10) Factorizing the operator and the DM term one obtains $`\dot{\sigma }`$ $`=`$ $`i/\mathrm{}[H^\mathrm{S},\sigma ]+\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right)\dot{𝐃}^{\mathrm{SE}}(t,0)(𝐃^{\mathrm{SE}})^1(t,0)\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right)`$ (2.11) $`\times \left[\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right)\mathrm{Tr}^\mathrm{E}\stackrel{~}{\rho }\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right)\right].`$ Substituting Eq. (2.7) in Eq. (2.11) gives: $$\dot{\sigma }=i/\mathrm{}[H^\mathrm{S},\sigma ]+\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right)\dot{𝐃}^{\mathrm{SE}}(t,0)(𝐃^{\mathrm{SE}})^1(t,0)\mathrm{exp}\left(i/\mathrm{}H^\mathrm{S}t\right)\sigma .$$ (2.12) So we have obtained a differential RDMEM instead of the integral Eq. (2.5). ### 2.4 Cumulant expansion Equation (2.12) is found to be exact for the initially disentangled system and bath if the bath does not change in time. Such a precise description of the bath influence is possible with the path inetgral technique but found to be numerically expensive. Here we perform some approximations to consider the influence of the bath in leading order. Taking into account Eq. (2.5) and Eq. (2.6) the reduced Green’s matrix for Eq. (2.12) reads: $$𝐃^{\mathrm{SE}}(t,0)=\mathrm{Tr}^\mathrm{E}\left\{\mathrm{T}\mathrm{exp}\underset{0}{\overset{t}{}}𝑑\tau \left[i/\mathrm{}L\left(\tau \right)\right]\rho ^\mathrm{E}\left(0\right)\right\}$$ (2.13) Below we define a function $`F\left(\lambda \right)`$ as follows: $$\mathrm{Tr}^\mathrm{E}\left[F\left(\lambda \right)\right]=F\left(\lambda \right)=\mathrm{Tr}^\mathrm{E}\left\{\mathrm{T}\mathrm{exp}\lambda \underset{0}{\overset{t}{}}𝑑\tau \left[i/\mathrm{}L\left(\tau \right)\right]\rho ^\mathrm{E}\left(0\right)\right\}$$ (2.14) To expand this expression in the cumulant form $`F=\mathrm{exp}\left\{\underset{n}{}\lambda ^nK_n\right\}`$ we solve the following differential equation: $`\frac{d}{dt}F=i/\mathrm{}\lambda L\left(t\right)F`$. The solutions of this equation in zeroth, first, and second orders of perturbation theory are, respectively: $`F^{\left(0\right)}`$ $`=`$ $`1`$ $`F^{\left(1\right)}`$ $`=`$ $`1\lambda i/\mathrm{}{\displaystyle \underset{0}{\overset{t}{}}}L\left(\tau \right)𝑑\tau `$ $`F^{\left(2\right)}`$ $`=`$ $`1\lambda i/\mathrm{}{\displaystyle \underset{0}{\overset{t}{}}}L\left(\tau \right)𝑑\tau \lambda i/\mathrm{}{\displaystyle \underset{0}{\overset{t}{}}}L\left(\tau \right)𝑑\tau \left[\lambda i/\mathrm{}{\displaystyle \underset{0}{\overset{\tau }{}}}L\left(\tau ^{}\right)𝑑\tau ^{}\right]`$ (2.15) The expression for $`𝐃^{\mathrm{SE}}(t,0)`$ in second order of the perturbative and the cumulant expansions read, respectively $$F(\lambda ,t)1\lambda i/\mathrm{}\underset{0}{\overset{t}{}}L\left(\tau \right)𝑑\tau +\left(\frac{i\lambda }{\mathrm{}}\right)^2\underset{0}{\overset{t}{}}\underset{0}{\overset{\tau }{}}L\left(\tau \right)d\tau L\left(\tau ^{}\right)𝑑\tau ^{}$$ (2.16) $$F(\lambda ,t)=\mathrm{exp}\left\{\underset{n}{}\lambda ^nK_n\right\}1+\underset{n}{}\lambda ^nK_n+\left(\underset{n}{}\lambda ^nK_n\right)^2+\mathrm{}$$ (2.17) As Eq. (2.16) is equal to Eq. (2.17) so $$K_1=i/\mathrm{}\underset{0}{\overset{t}{}}𝑑\tau L\left(\tau \right)$$ (2.18) $$K_2+\frac{1}{2}\left(K_1\right)^2=\left(\frac{i}{\mathrm{}}\right)^2\underset{0}{\overset{t}{}}\underset{0}{\overset{\tau }{}}L\left(\tau \right)d\tau L\left(\tau ^{}\right)𝑑\tau ^{}$$ $$K_2=\left(\frac{i}{\mathrm{}}\right)^2\underset{0}{\overset{t}{}}\underset{0}{\overset{\tau }{}}L\left(\tau \right)L\left(\tau ^{}\right)𝑑\tau 𝑑\tau ^{}\frac{1}{2}\left(i/\mathrm{}\underset{0}{\overset{t}{}}L\left(\tau \right)𝑑\tau \right)^2$$ (2.19) Now we shall use the well known fact that $`\left[\underset{0}{\overset{t}{}}f\left(\tau \right)𝑑\tau \right]^2=2\underset{0}{\overset{t}{}}f\left(\tau \right)𝑑\tau \underset{0}{\overset{\tau }{}}f\left(\tau ^{}\right)𝑑\tau ^{}`$. This fact helps to rewrite the second cumulant Eq. (2.19) in the following form: $$K_2=\left(\frac{i}{\mathrm{}}\right)^2\underset{0}{\overset{t}{}}𝑑\tau \underset{0}{\overset{\tau }{}}𝑑\tau ^{}\left\{L\left(\tau \right)L\left(\tau ^{}\right)L\left(\tau \right)L\left(\tau ^{}\right)\right\}.$$ (2.20) If we know $`K_1`$ and $`K_2`$ from Eq. (2.18) and Eq. (2.20) then the trace of $`F`$ given by Eq. (2.14) yields: $$𝐃^{\mathrm{SE}}(t,0)=F\left(\lambda \right)|_{\lambda =1}=\mathrm{exp}𝐊.$$ (2.21) where $`𝐊K_1+K_2`$. Equation (2.21) allows to express the kernel of the differential Eq. (2.12) as $`\dot{𝐃}𝐃^1=\dot{𝐊}`$, where $$\dot{𝐊}=\frac{d}{dt}\left(K_1+K_2\right)=i/\mathrm{}L\left(t\right)+\left(i/\mathrm{}\right)^2\underset{0}{\overset{t}{}}𝑑\tau \left\{L\left(t\right)L\left(\tau \right)L\left(t\right)L\left(\tau \right)\right\}$$ (2.22) Transforming back from Liouvillian to Hamiltonian form yields $`\dot{𝐊}\sigma `$ $`=`$ $`i/\mathrm{}[\stackrel{~}{H}^{\mathrm{SE}}\left(t\right),\sigma ]\mathrm{}^2{\displaystyle \underset{0}{\overset{t}{}}}d\tau \{[\stackrel{~}{H}^{\mathrm{SE}}\left(t\right),[\stackrel{~}{H}^{\mathrm{SE}}\left(\tau \right),\sigma ]]`$ (2.23) $`[\stackrel{~}{H}^{\mathrm{SE}}\left(t\right),[\stackrel{~}{H}^{\mathrm{SE}}\left(\tau \right),\sigma ]]\}`$ It is easy to show that $`\stackrel{~}{H}^{\mathrm{SE}}\left(t\right)=0`$ since $$\dot{𝐊}\sigma =\mathrm{}^2\underset{0}{\overset{t}{}}𝑑\tau [\stackrel{~}{H}^{\mathrm{SE}}\left(t\right),[\stackrel{~}{H}^{\mathrm{SE}}\left(\tau \right),\sigma ]]$$ (2.24) This expression is found to be precise up to the second order cumulant expansion. Taking some approximation this equation can be transformed into either Agarwal-type master equation used for the analysis of the HO in chapter 3 or Louisell-Lax-type master equation used for the calculation of the ET ansfer dynamics in artificial photosynthetic molecular aggregates in chapter 4. Below we show some steps of derivation for the Louisell-Lax-type master equation. ### 2.5 Master Equation Making the RWA in Eq. (2.24) and performing averaging over the bath degrees of freedom one gets bath correlation functions $`b_\lambda ^{}(\tau )b_\lambda ^+(t)=\delta _{\lambda ^{}\lambda }\left[n(\omega _\lambda )+1\right]\mathrm{exp}\left[i\omega _\lambda (t\tau )\right]`$, where $`n(\omega )=[\mathrm{exp}(\mathrm{}\omega /k_BT)1]^1`$ denotes the Bose-Einstein distribution. Back in the Schrödinger picture Eq. (2.3) reads $`\dot{\sigma }=`$ $``$ $`i/\mathrm{}[H^\mathrm{S},\sigma ]1/\mathrm{}^2{\displaystyle \underset{\mu \nu }{}}{\displaystyle _0^t}𝑑\tau `$ $`\times `$ $`\{[R_{n+1}(\omega _{\mu \nu },t\tau )(V_{\mu \nu }^+V_{\mu \nu }\sigma V_{\mu \nu }\sigma V_{\mu \nu }^+)R_{n+1}(\omega _{\mu \nu },\tau t)(V_{\mu \nu }\sigma V_{\mu \nu }^+\sigma V_{\mu \nu }^+V_{\mu \nu })]`$ $`+`$ $`[R_n(\omega _{\mu \nu },\tau t)(V_{\mu \nu }V_{\mu \nu }^+\sigma V_{\mu \nu }^+\sigma V_{\mu \nu })R_n(\omega _{\mu \nu },t\tau )(V_{\mu \nu }^+\sigma V_{\mu \nu }\sigma V_{\mu \nu }V_{\mu \nu }^+)]\},`$ where $$R_n(\omega ,\tau )=\underset{\lambda }{}n_\lambda K_\lambda ^2\mathrm{exp}[i(\omega _\lambda \omega )\tau ]$$ (2.26) are the correlation functions of the environment perturbations. The subscript $`n`$ in Eq. (2.5) refers to the factor $`n_\lambda `$ in Eq. (2.26). To obtain $`R_{n+1}`$ one replaces $`n_\lambda `$ by $`n_\lambda +1`$. The integral in Eq. (2.5) has different behavior on short and long time scales. On the time scale comparable to the bath correlation time the function $`R`$ allows that non-resonant bath modes $`\omega _\lambda \omega _{\mu \nu }`$ give a contribution to the system dynamics. Here we apply the Markov approximation, i.e., we restrict ourselves to the limit of long times and then the above mentioned integral is an approximation of the delta function $`lim_t\mathrm{}_0^tR_n(\omega ,t\tau )/n_\lambda 𝑑\tau =\pi _\lambda K_\lambda ^2\delta (\omega \omega _\lambda )`$. Furthermore, we replace the discrete set of bath modes with a continuous one. To do so one has to introduce the spectral density of bath modes $`J(\omega )=\pi _\lambda K_\lambda ^2\delta (\omega \omega _\lambda )`$ and to replace the summation by an integration. Finally one obtains the following master equation $`\dot{\sigma }={\displaystyle \frac{i}{\mathrm{}}}[\widehat{H}^\mathrm{S},\sigma ]+L\sigma ,`$ (2.27) with $`L\sigma `$ $`=`$ $`{\displaystyle \underset{\mu \nu }{}}\mathrm{\Gamma }_{\mu \nu }\{[n\left(\omega _{\mu \nu }\right)+1]([\widehat{V}_{\mu \nu }\sigma ,\widehat{V}_{\mu \nu }^+]+[\widehat{V}_{\mu \nu },\sigma \widehat{V}_{\mu \nu }^+])`$ $`+n(\omega _{\mu \nu })([\widehat{V}_{\mu \nu }^+\sigma ,\widehat{V}_{\mu \nu }]+[\widehat{V}_{\mu \nu }^+,\sigma \widehat{V}_{\mu \nu }])\},`$ where the damping constant $$\mathrm{\Gamma }_{\mu \nu }=\mathrm{}^2J(\omega _{\mu \nu })$$ (2.29) depends on the coupling of the transition $`|\mu |\nu `$ to the bath mode of the same frequency. Formally, the damping constant depends on the density of bath modes $`J`$ at the transition frequency $`\omega _{\mu \nu }`$. Equation (2.5) belongs to Louisell-Lax type and maintains the Lindblad form. We apply this equation to the system of discrete levels to describe the ET process in chapter 4. A version of this equation without RWA is carefully investigated in chapter 3 in application to a single HO. ## Chapter 3 First Application: Harmonic Oscillator In this chapter we develop and adopt the theoretical method, introduced in the previous chapter to the HO. This approach is well suited for systems with negligible electronic coupling between different diabatic states and a single reaction coordinate modelled by a HO. Although it is a quite simple system, the study of its dynamics allows to answer the question about the border between classical and quantum effects. This question deals with the superposition of coherent states of the HO. We begin this chapter with the introduction of the superpositional states and a brief review of decoherence problem (section 3.1). In sections 3.2 and 3.3 we briefly rederive the methods of investigation of the HO coupled to a thermal bath. In section 3.4 we discuss the behavior of the superpositional states either using the analytical method derived in section 3.3 or by numerical simulation. One of the goals of this contribution is to present a consistent analysis of the decoherence on the basis of a DM approach starting from von Neumann’s equation for the DM of the whole system, i.e. the microscopic quantum system and the ”macroscopic” environment. ### 3.1 Introduction to the decoherence problem There is a number of propositions how to create the superposition states in mesoscopic systems, or systems that have both macroscopic and microscopic features. A representative example is the superposition of two coherent states of the HO $$|\alpha ,\varphi =N^1\left(|\alpha +e^{i\varphi }|\alpha \right)$$ (3.1) for a relatively large amplitude ($`\alpha 3÷5`$). Here, $`|\alpha `$ is a coherent state and $`N=\left[2+2\mathrm{cos}\varphi \mathrm{exp}\left(2\left|\alpha \right|^2\right)\right]^{1/2}`$ is a normalization constant. These states have been observed recently for the intracavity microwave field and for motional states of a trapped ion . Additionally, it has been predicted that superpositions of coherent states of molecular vibrations could be prepared by appropriately exciting a molecule with two short laser pulses and the practical possibilities of realizing such an experiment have been discussed . In this scheme the quantum interference would survive on a picosecond time scale, which is characteristic for molecular vibrations. From the theoretical point of view, quantum decoherence has been studied extensively . Most efforts focused on the decoherence of the HO states due to the coupling to the heat bath, consisting of a large number of oscillators representing the environment. The energetic spectrum of the bath is usually taken to be broad and dense to provide the transfer of excitation energy from the system to the bath. The system is usually described on the basis of the master equation for the reduced density operator. There are few general approaches for this method. In most approaches listed in chapter 2 one adopts the Markov approximation for real calculations. It means that all details of the complex system-environment interaction are neglected and relaxation is described by the characteristic decay constants. The physical analysis of the system behavior beyond the Markov approximation have been proposed by Zurek : the coupling with the environment singles out a preferred set of states, called ”the pointer basis”. Only vectors of this basis survive the quantum dynamics. The vectors of the pointer basis are the eigenvectors of operators, which commute with the (full) interaction Hamiltonian of the system. This basis depends on the form of the coupling. Very often this pointer basis consists of the eigenstates of the coordinate operator. The density operator describing the system evolves to diagonal form in the pointer basis, which is usually connected to the disappearance of quantum interference. The two approaches give different pictures of the same decoherence processes. ### 3.2 Generalized master equation Let us consider a single molecule vibrational mode as a one-dimensional harmonic potential. For this case the common Hamiltonian Eq. (2.1) contains the molecular system $`H_\mathrm{S}=\mathrm{}\omega \left(a^+a+1/2\right)`$. The molecule interacts with a number of harmonic oscillators modeling the environment. In the interaction Hamiltonian $$H_{\mathrm{SE}}=\mathrm{}\underset{\xi }{}K_\xi \left(b_\xi ^++b_\xi \right)\left(a+a^+\right).$$ (3.2) $`a`$ ($`a^+`$) are annihilation (creation) operators of molecular vibrations with frequency $`\omega `$, $`b_\xi `$ ($`b_\xi ^+`$) operators for the environmental vibrations having the frequencies $`\omega _\xi `$. $`K_\xi `$ is the coupling between them. Performing the same steps of derivation as for Eq. (2.5) but without RWA we arrive at the non-Markovian master equation for the HO $$\dot{\sigma }=i\omega [a^+a,\sigma ]+L\sigma ,$$ (3.3) where the action of the relaxation operator $`L`$ is defined by $`L\sigma `$ $`=`$ $`[\left(A_{n+1}+A_n^+\right)\sigma ,a^++a]+[a^++a,\sigma \left(A_n+A_{n+1}^+\right)].`$ (3.4) Here, the operators $`A_n`$ and $`A_{n+1}`$ are defined by the linear combinations of the operators $`a`$ and $`a^+`$ as $`A_n`$ $`=`$ $`\gamma _n\left(t\right)a+\stackrel{~}{\gamma }_n\left(t\right)a^+,`$ $`A_{n+1}`$ $`=`$ $`\gamma _{n+1}\left(t\right)a+\stackrel{~}{\gamma }_{n+1}\left(t\right)a^+,`$ (3.5) with the functions (see Fig. 3.1) $$\gamma _{n+1}(t)=_0^tR_{n+1}(\tau )𝑑\tau =\underset{\xi }{}K_\xi ^2(n_\xi +1)\frac{e^{i(\omega _\xi \omega )t}1}{i(\omega _\xi \omega )},$$ (3.6) $$\gamma _n(t)=_0^tR_n(\tau )𝑑\tau =\underset{\xi }{}K_\xi ^2n_\xi \frac{e^{i(\omega _\xi \omega )t}1}{i(\omega _\xi \omega )},$$ (3.7) $$\stackrel{~}{\gamma }_{n+1}(t)=_0^tR_{n+1}(\tau )e^{2i\omega \tau }𝑑\tau =\underset{\xi }{}K_\xi ^2(n_\xi +1)\frac{e^{i(\omega _\xi +\omega )t}1}{i(\omega _\xi +\omega )},$$ (3.8) $$\stackrel{~}{\gamma }_n(t)=_0^tR_n(\tau )e^{2i\omega \tau }𝑑\tau =\underset{\xi }{}K_\xi ^2n_\xi \frac{e^{i(\omega _\xi +\omega )t}1}{i(\omega _\xi +\omega )}.$$ (3.9) Note, that $`R_n`$ and $`R_{n+1}`$ are the correlation functions of the environmental perturbations (2.26), where $`n_\xi =[\mathrm{exp}(\mathrm{}\omega _\xi /k_BT)1]^1`$ denotes the number of quanta in the bath mode with the frequency $`\omega _\xi `$. Subscripts $`n`$ and $`n+1`$ indicate the character of the temperature dependence, refering to the factors $`n_\xi `$ and $`n_\xi +1`$ in Eqs. (3.6)-(3.9). The functions $`\gamma `$ correspond to the friction coefficient in the classical limit. The first two coefficients $`\gamma _n`$ and $`\gamma _{n+1}`$ strongly depend on the coupling constant $`K_\xi `$ for frequencies $`\omega _\xi \omega `$ and on the number of quanta in the bath modes with the same frequencies, whilst the coefficients $`\stackrel{~}{\gamma }_n`$ and $`\stackrel{~}{\gamma }_{n+1}`$ are very small for all frequencies. $`\gamma _{n+1}`$ and $`\gamma _n`$ describe the situation when an emission of a quantum from the system with rate $`\gamma _{n+1}`$ occurs more probably than an absorption of a quantum with rate $`\gamma _n`$. The terms of the master equation associated with $`\stackrel{~}{\gamma }_{n+1}`$ and $`\stackrel{~}{\gamma }_n`$ originate from the non-RWA terms $`ab_\xi `$, $`a^+b_\xi ^+`$ of the Hamiltonian (3.2) and correspond to the reverse situation: an absorption $`\stackrel{~}{\gamma }_{n+1}`$ is more probable than an emission $`\stackrel{~}{\gamma }_n`$. As shown in Fig. 3.1 the last two types of terms are essential only for the first stage of relaxation $`t\tau _\mathrm{c}`$, where $`\tau _\mathrm{c}`$ denotes the correlation time of environmental perturbations. The obtained master equation (3.3) describes different stages of vibrational relaxation. The initial stage is defined by a period of time smaller than the correlation time $`\tau _c`$. This time can roughly be estimated as $`\tau _c2\pi /\mathrm{\Delta }\omega `$ from the width $`\mathrm{\Delta }\omega `$ of the perturbation spectrum $`K_\xi ^2`$. For such small times one can approximate the master equation (3.3)-(3.9) in the following form $$\dot{\sigma }=i\omega [a^+a,\sigma ]+\mathrm{\Gamma }\left\{[\left(a^++a\right)\sigma ,a^++a]+[a^++a,\sigma \left(a^++a\right)]\right\},$$ (3.10) where $`\mathrm{\Gamma }=_\xi K_\xi ^2\left(2n_\xi +1\right)`$ is a real constant. As follows from Eq. (3.10), the pointer basis for this step of relaxation is defined by the eigenstates of the position operator $`\widehat{Q}a^++a`$. Another period of time, for which the form of the relaxation operator $`R`$ according to Eq. (3.4) is universal, is the kinetic stage, where $`t\tau _c`$ and the Markov approximation becomes applicable. In this stage the master equation has the form $`\dot{\sigma }`$ $`=`$ $`i\omega [a^+a,\sigma ]`$ $`+\gamma \left\{[\left[(n+1)a+na^+\right]\sigma ,a^++a]+[a^++a,\sigma \left[(n+1)a+na^+\right]]\right\},`$ where $`\gamma =\pi K^2g`$ is the decay rate of the vibrational amplitude. Here $`n=n_\xi `$, $`K=K_\xi `$ and the density of the bath states $`g=g(\omega _\xi )`$ are evaluated at the frequency $`\omega =\omega _\xi `$ of the selected oscillator. It should be stressed that Eq. (3.2) differs from the usual master equation for a damped HO for derivation of which the RWA is applied . Still Eq. (3.2) is only a particular case of the more general Eqs. (3.3)-(3.9). To the best of our knowledge Agarwal was the first who derived this equation. The phase-sensitive relaxation leads to the new effect of classical squeezing and to a decrease of the effective HO frequency . Inbetween there is a time interval, where relaxation is specific and depends on the particular spectrum of $`K_\xi ^2`$. ### 3.3 Analytical solution for wave packet dynamics The solution of the equation of motion of the RDM can be conveniently found using the characteristic function formalism . This formalism enables us to use the differential operators $`\frac{}{\lambda }`$ and $`\frac{}{\lambda ^{}}`$ instead of $`a^+`$ and $`a`$. From the set of normally ordered $`F=\mathrm{Tr}\left(\sigma e^{\lambda a^+}e^{\lambda ^{}a}\right)`$, abnormally ordered $`F=\mathrm{Tr}\left(\sigma e^{\lambda ^{}a}e^{\lambda a^+}\right)`$ and Wigner $`F=\mathrm{Tr}\left(\sigma e^{\lambda ^{}a+\lambda a^+}\right)`$ characteristic functions we use here only the first. Multiplying both sides of Eq. (3.3) with the factor $`f=\mathrm{exp}(\lambda a^+)\mathrm{exp}(\lambda ^{}a)`$ and taking the trace one can rearrange all terms into such a form that $`a`$ and $`a^+`$ precede the appropriate exponent. For this operation we change the order of operators using the expression $`a\mathrm{exp}(\lambda a^+)=\mathrm{exp}(\lambda a^+)\left(a+\lambda \right)`$ to make the transformation $`a^+\mathrm{exp}(\lambda a^+)=\frac{}{\lambda }\mathrm{exp}(\lambda a^+)`$. After that every term can be represented by the normally ordered characteristic function $$F=\mathrm{Tr}(f\sigma ),$$ (3.12) upon which one of the differential operators acts. We obtain, e.g., $`\mathrm{Tr}\left([a^+a,\sigma ]f\right)=\left(\lambda ^{}\frac{}{\lambda ^{}}\lambda \frac{}{\lambda }\right)F`$, $`\mathrm{Tr}\left([a^+\sigma ,a]f\right)=\lambda ^{}\left(\frac{}{\lambda }\lambda ^{}\right)F`$, and $`\mathrm{Tr}\left([a\sigma ,a^+]f\right)=\lambda ^{}\frac{}{\lambda ^{}}F`$. Such manipulations lead us finally to the complex-valued partial differential equation: $`\dot{F}`$ $`=`$ $`\left[i\omega \lambda ^{}+\mu \left(\lambda +\lambda ^{}\right)\right]{\displaystyle \frac{}{\lambda ^{}}}F+\left[i\omega \lambda \mu ^{}\left(\lambda +\lambda ^{}\right)\right]{\displaystyle \frac{}{\lambda }}F`$ (3.13) $`\left(\lambda +\lambda ^{}\right)\left(\nu \lambda ^{}+\nu ^{}\lambda \right)F,`$ where $`\mu (t)`$ $`=`$ $`\gamma _{n+1}(t)+\stackrel{~}{\gamma }_n^{}(t)\nu ^{}(t),`$ $`\nu (t)`$ $`=`$ $`\gamma _n^{}(t)+\stackrel{~}{\gamma }_{n+1}(t)`$ (3.14) are relaxation functions. An analogous method was used by Puri and Lawande , who also treated an HO coupled to the heat bath with the help of the normally ordered characteristic function. They obtained a general expression for time evolution of the characteristic function for an arbitrary initial state of the oscillator, see Eq. (12) in Ref. . This expression is valid for the non-Markovian regime but performed under RWA. We can solve Eq. (3.13) by using the integral representation for the characteristic function $`F(\lambda ,\lambda ^{},0)=\mathrm{Tr}\left(e^{a^+\lambda }e^{a\lambda ^{}}\right)`$ which formally allows us to describe the nondiagonal DM. Below the notation $`𝑑\alpha 𝑑\beta c(\alpha ,\beta )=\alpha |\beta `$ is adopted. An initial characteristic function $$F(\lambda ,\lambda ^{},0)=𝑑\alpha 𝑑\beta e^{\alpha \lambda \beta \lambda ^{}}c(\alpha ,\beta )$$ (3.15) will evolve in accordance with Eq. (3.13) as $$F(\lambda ,\lambda ^{},t)=𝑑\alpha 𝑑\beta \mathrm{exp}\left[\underset{m,n}{}K_{mn}^{(\alpha ,\beta )}(t)\lambda ^m\left(\lambda ^{}\right)^n\right]c(\alpha ,\beta ).$$ (3.16) Taking the derivatives of the characteristic function (3.12) one obtains the mean values of observables, in particular the mean number of quanta is given by $`a^+a`$ $`=`$ $`{\displaystyle \frac{^2F}{\lambda \lambda ^{}}}|_{\lambda =\lambda ^{}=0}{\displaystyle \frac{F}{\lambda }}{\displaystyle \frac{F}{\lambda ^{}}}|_{\lambda =\lambda ^{}=0}.`$ (3.17) Here we restrict the cumulant expansion to the second order, i.e. $`m+n2`$ . For a wide class of initial states (coherent, thermal, squeezed, etc.) higher order cumulants vanish and our approximation becomes exact. The cumulants could hold nondiagonal information, such as the DM, in relevant cases we stress it with the upper index $`(\alpha )`$ or $`(\beta )`$. The functions $`K_{mn}(t)`$ in Eq. (3.16) are given by the solutions of the sets of equations $`\dot{K}_{01}^{(\beta )}`$ $`=`$ $`\left(i\omega +\mu \right)K_{01}^{(\beta )}+\mu ^{}K_{10}^{(\alpha )},`$ $`\dot{K}_{10}^{(\alpha )}`$ $`=`$ $`\left(i\omega \mu ^{}\right)K_{10}^{(\alpha )}+\mu K_{01}^{(\beta )},`$ (3.18) $`\dot{K}_{11}`$ $`=`$ $`2\mathrm{}\nu 2\left(\mathrm{}\mu \right)K_{11}+2\mu K_{02}+2\mu ^{}K_{20},`$ $`\dot{K}_{20}`$ $`=`$ $`\nu ^{}+\mu K_{11}+2\left(i\omega \mu ^{}\right)K_{20},`$ (3.19) $`\dot{K}_{02}`$ $`=`$ $`\nu +\mu ^{}K_{11}2\left(i\omega \mu \right)K_{02},`$ with the initial values $$\begin{array}{c}K_{10}^{(\alpha )}(t=0)=\alpha ,\\ K_{01}^{(\beta )}(t=0)=\beta ,\\ K_{11}\left(t=0\right)=K_{20}\left(t=0\right)=K_{02}\left(t=0\right)=0.\end{array}$$ (3.20) $`\mathrm{}`$ and $`\mathrm{}`$ stands for real and imaginary part of a complex variable, respectively. For the special case $`\beta =\alpha ^{}`$, these initial conditions represent the coherent state with amplitude $`\alpha `$. This solution can be used for the construction of wave packets in different representations. Here, we will discuss the coordinate representation, in particular the dependence of the probability density $`P`$ on the vibrational coordinate $`Q`$ and on time $$P(Q,t)=\frac{1}{2\pi }_{\mathrm{}}^{\mathrm{}}𝑑\lambda e^{i\lambda Q}\chi (\lambda ,t),$$ (3.21) where $$\chi (\lambda ,t)=\mathrm{Tr}\left[e^{i\lambda \left(a^++a\right)}\sigma (t)\right]=e^{\lambda ^2/2}F(i\lambda ,i\lambda ,t)$$ (3.22) is a characteristic function for the position operator, which is nothing but the diagonal of the Wigner characteristic function . Evaluating the integral (3.21) we finally obtain $$P(Q,t)=𝑑\alpha 𝑑\beta P^{(\alpha ,\beta )}(Q,t)c(\alpha ,\beta ),$$ (3.23) where $$P^{(\alpha ,\beta )}(Q,t)=\frac{1}{2\sqrt{\pi V\left(t\right)}}\mathrm{exp}\left\{\frac{\left[QQ^{(\alpha ,\beta )}\left(t\right)\right]^2}{4V\left(t\right)}\right\}$$ (3.24) and $`Q^{(\alpha ,\beta )}\left(t\right)=K_{10}^{(\alpha )}\left(t\right)+K_{01}^{(\beta )}\left(t\right),`$ (3.25) $`V\left(t\right)={\displaystyle \frac{1}{2}}+K_{11}\left(t\right)+K_{20}\left(t\right)+K_{02}\left(t\right).`$ (3.26) For the case $`\beta =\alpha ^{}`$ the function $`Q^{(\alpha ,\alpha ^{})}`$ denotes the expectation values of the coordinate operator of the coherent state, $`V`$ is the broadening of the Gaussian packet Eq. (3.24). The distribution $`P`$ can be used for the investigation of relaxation dynamics for any initial state of molecular vibration, but it is best suited for studying the evolution of the states prepared as a superposition of coherent states. Below, we will discuss the relaxation dynamics of initially coherent states and the superposition Eq. (3.1) of two coherent states. ### 3.4 Coherent states and their superpositions #### 3.4.1 Coherent states Here we investigate the relaxation dynamics using the master equation (3.3). For the coherently exited states $`|\alpha _0`$ the initial characteristic function is $`F(\lambda ,\lambda ^{},0)=\mathrm{exp}\left(\alpha _0^{}\lambda \alpha _0\lambda ^{}\right)`$. Thus, the initial values for $`K_{10}^{(\alpha )}\left(t\right)`$ and $`K_{01}^{(\beta )}\left(t\right)`$ are $`\alpha _0^{}`$ and $`\alpha _0`$. In the first stage of relaxation, when $`t\tau _c`$, the relaxation functions are $`\mu \left(t\right)=0`$, $`\nu \left(t\right)=\mathrm{\Gamma }t`$. Therefore, the solution of the system of Eqs. (3.18)-(3.19) gives $`Q^{(\alpha _0^{},\alpha _0)}\left(t\right)`$ $`=`$ $`2\mathrm{}\left(\alpha _0e^{i\omega t}\right)`$ $`V\left(t\right)`$ $`=`$ $`{\displaystyle \frac{1}{2}}+\mathrm{\Gamma }t^2.`$ (3.27) Even in this early stage there is a small quadratic broadening of the wave packet $`P^{(\alpha _0^{},\alpha _0)}`$ $`(Q,t)`$ without changing its mean amplitude. After the intermediate stage of relaxation the solution of the system goes into the Markovian stage of relaxation, where the master equation (3.2) works. For this stage the solution of Eqs. (3.18)-(3.19) reads $`Q^{(\alpha _0^{},\alpha _0)}\left(t\right)`$ $`=`$ $`2\mathrm{}\left[\alpha _0z\left(t\right)\right]e^{\gamma t},`$ $`V\left(t\right)`$ $`=`$ $`{\displaystyle \frac{1}{2}}+nne^{2\gamma t}\left[1+\left({\displaystyle \frac{\gamma }{\stackrel{~}{\omega }}}\right)^2\left(1\mathrm{cos}2\stackrel{~}{\omega }t\right)+{\displaystyle \frac{\gamma }{\stackrel{~}{\omega }}}\mathrm{sin}2\stackrel{~}{\omega }t\right],`$ (3.28) where $`z\left(t\right)`$ $`=`$ $`\mathrm{cos}\stackrel{~}{\omega }t+(\gamma /\stackrel{~}{\omega })\mathrm{sin}\stackrel{~}{\omega }t+i(\omega /\stackrel{~}{\omega })\mathrm{sin}\stackrel{~}{\omega }t,`$ $`\stackrel{~}{\omega }`$ $`=`$ $`\sqrt{\omega ^2\gamma ^2},`$ (3.29) $`n`$ $`=`$ $`\left[\mathrm{exp}(\mathrm{}\omega /kT)1\right]^1.`$ Equation (3.29) demonstrates the decrease of the effective harmonic oscillator frequency due to the phase-dependent interaction with the bath, well-known in classical mechanics, but absent in the majority of quantum considerations performed with RWA . Neglecting the terms $`a^+b_\xi ^+`$, $`ab_\xi `$ in Hamiltonian (2.1), and repeating the derivation for the first-order cumulant moments one obtains the analogs of Eqs. (3.18) $`\dot{K}_{01}^{(\beta )}`$ $`=`$ $`\left(i\omega +\mu \right)K_{01}^{(\beta )},`$ $`\dot{K}_{10}^{(\alpha )}`$ $`=`$ $`\left(i\omega \mu ^{}\right)K_{10}^{(\alpha )}.`$ (3.30) The solutions $`K_{01}^{(\beta )}=\mathrm{exp}(i\omega \gamma )t`$, $`K_{10}^{(\alpha )}=\mathrm{exp}(i\omega \gamma )t`$ do not present any change of the HO frequency. So we conclude that this reduction of the frequency is due to the phase-sensitivity of the Hamiltonian (3.2). The broadening of the wave packet $`V(t)`$ is displayed in Fig. 3.2 for the damping $`\gamma =0.1\omega `$ chosen to present the prediction of Eq. (3.28) most clearly. An extremely rapid relaxation of the same order was assumed for the exciton motion simulation in a photosynthetic complex . However, for the majority of real systems the damping rate is a few orders of magnitude lower. With methods of so-called quantum tomography even the rapid relaxation of the wave packet like in Fig. 3.2 can be observed experimentally. The increase of $`V(t)`$ appears due to absorption of quanta from the heat bath and can be obtained already using RWA. The oscillation of the broadening which we directly derived in coordinate space is in accordance with the oscillations of second moments of the Green’s function . Ref. assumes initially a squeezed state. Our prediction (3.28) goes further, because the oscillation appears even for the usual coherent state as the initial one. This oscillation does not present quantum squeezing, because the width is never smaller than the ground state width. The oscillations with frequency $`2\stackrel{~}{\omega }`$ induced by the phase sensitivity of Hamiltonian (2.1) also occur for the mean number of quanta (3.17). For the Markovian stage of relaxation, $$a^+a=n+e^{2\gamma t}\left[|\alpha _0|^2n\left(\frac{\omega }{\stackrel{~}{\omega }}\right)^2+2|\alpha _0|^2\frac{\gamma }{\stackrel{~}{\omega }}\mathrm{sin}2\stackrel{~}{\omega }t+n\left(\frac{\gamma }{\stackrel{~}{\omega }}\right)^2\mathrm{cos}2\stackrel{~}{\omega }t\right]$$ (3.31) starts from the number of quanta for the coherent state $`|\alpha _0|^2`$ and relaxes to the number of quanta in the resonant bath mode $`n`$, see Fig. 3.2. Technically the oscillations are induced by the non-RWA relaxation functions $`\stackrel{~}{\gamma }_{n+1}`$ and $`\stackrel{~}{\gamma }_n`$, Eqs. (3.8)-(3.9), while Kohen, Marston, and Tannor treat the features of a non-RWA approach with the concept of a dissipation rate oscillating in time. In Fig. 3.2 we have shown the dynamics of the mean number of quanta $`a^+a`$. Its distribution has been presented by Milburn and Walls for a squeezed state of the HO. #### 3.4.2 Superposition of two coherent states. Creation The methods derived in section 3.3 are easy applicable to the evolution of the superposition of two coherent states. Superpositional states attract attention due to specific quantum effects they are involved in. Such effects of quantum nature which can be realized experimentally are discussed for quantum teleportation , quantum cryptography , and quantum computation . All these effects are possible as long as the states remain pure and keep their superpositional nature. Such states are created experimentally for the motional states of a trapped atom coupled to its hyperfine transition and for the microresonator mode interacting with a Rydberg atom . The dynamics of such type of systems is successfully predicted by the so-called Jaynes-Cummings Model (JCM) including a HO $`\mathrm{}\omega _0a^+a`$ coupled linearly to a two level system (TLS) with $`\mathrm{}\omega _{\mathrm{TLS}}`$ via (in RWA) $$H_{\mathrm{SB}}=g(a^+\sigma _{}+a\sigma _+),$$ (3.32) where $`\sigma _{}=|12|`$ denotes the TLS lowering operator, and allows the analytical solution . There are investigations considering the JCM coupling without RWA . Another extension of JCM is the introduction of dissipation processes associated with spontaneous emission from a TLS and the loss of energy from the cavity through mirrors . In the simplest generalization of the linear coupling the interaction is proportional to the number of quanta . An intensity dependent interaction was considered by Buzek . Gerry has considered a multiquanta interaction. A particular case, namely two-quanta transitions attracts interest up to nowadays . To illustrate the creation of superpositional cat-like states (3.1) we have performed a calculation with the HO-TLS interaction Eq. (3.32). In our calculation the coupling strength $`g=0.2\omega _{\mathrm{TLS}}`$ is taken a few hundred times larger than in experiment to make the wave packet dynamics more pronounced. We have calculated the time evolution of the DM $`\rho _{MmNn}^{\mathrm{JCM}}(t)`$ of the TLS ($`M`$, $`N`$ denoting, e. g., the Rydberg states of a Rb atom in ) and a HO ($`m`$, $`n`$ labeling the exitations of, e. g., the electromagnetic field mode in the microresonator which is tuned to $`\omega _{\mathrm{TLS}}`$). Initially the field is prepared in a coherent state, the atom in the excited state: $$\rho _{MmNn}^{\mathrm{JCM}}(t=0)=\delta _{2M}\delta _{N2}\mathrm{exp}|\alpha |^2\frac{\alpha ^m\left(\alpha ^{}\right)^n}{\sqrt{m!n!}}.$$ (3.33) The trace over the field degree of freedom yields the state of the atom $`\sigma _{MN}^{\mathrm{TLS}}`$ $`=`$ $`\left[\mathrm{Tr}_{\mathrm{HO}}\rho ^{\mathrm{JCM}}\right]_{MN}={\displaystyle \underset{m}{}}\rho _{MmNm}^{\mathrm{JCM}}`$ (3.34) and vice versa $`\sigma _{mn}^{\mathrm{HO}}`$ $`=`$ $`\left[\mathrm{Tr}_{\mathrm{TLS}}\rho ^{\mathrm{JCM}}\right]_{mn}={\displaystyle \underset{M}{}}\rho _{MmMn}^{\mathrm{JCM}}.`$ (3.35) The field mode is characterized by the wave packet $$P^{\mathrm{HO}}(Q,t)=\underset{m}{}\sigma _{mm}^{\mathrm{HO}}(t)\psi _m(Q),$$ (3.36) where $`\psi _m(Q)`$ denotes the HO eigenfunctions, and the mean value of the coordinate is $$\overline{Q}(t)=\underset{\mathrm{}}{\overset{\mathrm{}}{}}P^{\mathrm{HO}}(Q,t)Q𝑑Q.$$ (3.37) The state of the atom is characterized by its population $$P^{\mathrm{TLS}}(t)=\sigma _{22}^{\mathrm{TLS}}(t).$$ (3.38) The wave packet (3.36) shown in Fig. 3.3(a) evolves from the coherent state with one peak to the superpositional state with two peaks (at $`\omega _0t27`$) or interference structure (at $`\omega _0t25`$). The mean value (3.37) of the coordinate, see Fig. 3.3(b), reflects this transition from coherent to superpositional state: it oscillates initially and then relaxes to zero, because the wave packet of a superpositional state (3.1) is symmetric with respect to $`Q=0`$. The TLS population (3.38) in Fig. 3.3(c) also reduces to zero, because the interaction with each pair of HO modes $`n`$, $`n+1`$ yields an oscillation of the TLS population (3.38) with frequency given by the effective coupling $`g\sqrt{n+1}`$, so that the many HO levels initially populated, see Eq. (3.33), lead to destructive interference of these oscillations. When the damping rate is small enough a revival occurs . To prepare the superposition of coherent states we eliminate the revival by restricting the time of the atom-field interaction. Gerry and Knight mention applying two $`\pi /2`$ pulses to the Rydberg atom before and after its interaction with a field mode. In this subsection we illustrated a possible method to create a superpositional state (3.1) of a resonator field mode. The preparation of a motional state of a trapped ion in the state (3.1) is reported in Ref. , while the excitation of the molecular vibration in this state by two short laser pulses is discussed in Refs. . In the next subsection we calculate how the superpositional states evolve in time. #### 3.4.3 Superposition of two coherent states. Decoherence For the initial superposition of two coherent states (3.1) the normally ordered characteristic function Eq. (3.12) consists of four terms: $$F(\lambda ,\lambda ^{},0)=N^2\left[F_{\alpha ^{},\alpha }+F_{\alpha ^{},\alpha }+e^{2\left|\alpha \right|^2}\left(e^{i\varphi }F_{\alpha ^{},\alpha }+e^{i\varphi }F_{\alpha ^{},\alpha }\right)\right]$$ (3.39) with $`F_{\alpha ,\beta }=e^{\alpha \lambda \beta \lambda ^{}}`$. The first terms describe the mixture of two coherent states, $`|\alpha \alpha |+|\alpha \alpha |`$. The last two terms correspond to a quantum interference, i.e., they reflect the coherent properties of the superposition. In accordance with Eqs. (3.18)-(2.20) and the equation of motion (3.2) of the RDM this initial state evolves as $$P(Q,t)=\frac{1}{N^2}\left[P^{(\alpha ^{},\alpha )}(Q,t)+P^{(\alpha ^{},\alpha )}(Q,t)\right]+P_{\mathrm{int}}(Q,t)$$ (3.40) with the interference term $$P_{\mathrm{int}}(Q,t)=\frac{1}{N^2}e^{2\left|\alpha \right|^2}\left[e^{i\varphi }P^{(\alpha ^{},\alpha )}(Q,t)+e^{i\varphi }P^{(\alpha ^{},\alpha )}(Q,t)\right]$$ (3.41) where $`P^{(\alpha ,\beta )}(Q,t)`$ is given by Eq. (3.24) with $$Q^{(\alpha ^{},\alpha )}\left(t\right)=Q^{(\alpha ^{},\alpha )}\left(t\right)=2\mathrm{}\left[\alpha z(t)\right]e^{\gamma t},$$ (3.42) $$Q^{(\alpha ^{},\alpha )}\left(t\right)=Q^{(\alpha ^{},\alpha )}\left(t\right)=2i\mathrm{}\left[\alpha z\left(t\right)\right]e^{\gamma t}.$$ (3.43) $`V\left(t\right)`$ and $`z\left(t\right)`$ are defined by Eqs. (3.28)-(3.29). Rewriting the real part of Eq. (3.41) we finally obtain $`P_{\mathrm{int}}(Q,t)`$ $`=`$ $`{\displaystyle \frac{1}{N^2}}{\displaystyle \frac{1}{\sqrt{\pi V\left(t\right)}}}\mathrm{exp}\left(2\left|\alpha \right|^2+{\displaystyle \frac{4\left\{\mathrm{}\left[\alpha z(t)\right]\right\}^2e^{2\gamma t}Q^2}{4V(t)}}\right)`$ $`\times .\mathrm{cos}\{\varphi Q{\displaystyle \frac{\mathrm{}[\alpha z(t)]}{V(t)}}e^{\gamma t}\}.`$ Figure 3.4 illustrates, how the superpositional state (3.1) evolves in time in accordance with Eq. (3.40). It follows from Eq. (3.4.3), that the interference term describing quantum coherence in the system is only significant when $$\frac{\left\{\mathrm{}\left[\alpha z\left(t\right)\right]\right\}^2}{V\left(t\right)}e^{2\gamma t}2\left|\alpha \right|^2.$$ (3.45) This is true for $`\gamma t1`$, and moreover, when the two wave packets of the state (3.1) come close together (i.e., at the moment when $`z\left(t_i\right)i`$). Expanding Eq. (3.45) into a series at these points and taking into account $`\gamma t_i1`$, it is easy to see that $$P_{\mathrm{int}}\mathrm{exp}\left(2\left|\alpha \right|^2+\frac{\left\{\mathrm{}\left[\alpha z\left(t_i\right)\right]\right\}^2}{V\left(t_i\right)}\left(12\gamma t_i\right)\right).$$ (3.46) The decoherence is due to two reasons: The spreading $`V\left(t_i\right)=1/2+\mathrm{\Delta }V\left(t_i\right)n`$ of the wave packet due to thermal excitations by the bath, and amplitude decoherence. The latter means, that even for the case $`n=0`$ and $`\mathrm{}\left(\alpha z\left(t_i\right)\right)=\left|\alpha \right|`$ quantum interference disappears exponentially with the rate $$t_{\mathrm{dec}}^12\left|\alpha \right|^2\gamma $$ (3.47) This result obtained by Zurek is the main reason why quantum interference is difficult to observe in the mesoscopic and macroscopic world. For example, a physical system with mass 1g in a superposition state with a separation of 1cm shows a ratio of relaxation and decoherence time scales of $`10^{40}`$. Even if our measuring device is able to reflect the quantum properties of the microsystem, nevertheless objectification occurs due to the coupling between the meter and the environment. The fundamental result of Eq. (3.47) is obtained no matter which approach is used, e.g. the RWA or Zurek’s pointer basis approach or the self-consistent description of the present work. However, the time dependence of the superposition terms of the distribution of Eq. (3.4.3) differs a little bit, which can be seen in Fig. 3.5. As quantum interference is more sensitive, we have used it for comparison of three different approaches to the present problem. Figure 3.5 shows the difference between the time evolutions, which result from Eq. (3.27), from Eq. (3.28), and from the corresponding result of the RWA approach. These differences arise due to the fact that during the relaxation there is no constant pointer basis for all steps of the evolution. As follows from Eqs. (3.10) and (3.2), this basis changes from the position eigenstate basis in Eq. (3.10) to the more complicated basis in Eq. (3.2). #### 3.4.4 Partial conservation of superposition As the interference term (3.4.3) of the wave packet decays inevitably, it is difficult to observe the effects mentioned in subsection 3.4.2. The decoherence determines, e. g., the main requirement to the potential elements of quantum computers: the decoherence time should be smaller than the computation time . One should find a system maximally isolated from an environment or increase the decoherence time artificially. One could prolong the coherent interval of the evolution by increasing the number of information transfer channels , by feedback methods , or by so-called passive methods . Here we discuss in detail one more method of coherence preserving, namely organization of interaction with the environment . The idea of this method is to choose the system or the regime of system evolution which ensures a specific type of system-bath coupling, sometimes a nonlinear one. The linear coupling (3.2) to $`A=a^++a`$ corresponds to the resonant exchange of quanta between system and bath. A coupling to the number of phonons used in description of a trapped ion and of exciton evolution in molecular crystals describes the bath-induced modulation of the system transition frequency. A coupling of the form $`A=(a+\alpha )(a\alpha )`$ is discussed in Ref. and Glauber’s generalized annihilation operator $$A=a\mathrm{exp}\left(i\pi a^+a\right)$$ (3.48) in Ref. . An exotic coupling operator $`a\left(a^+aa^+a\right)`$ is discussed in Ref. . In Zurek’s ansatz the eigenstates $`|\psi _A`$ of the coupling operator $`A`$ (pointer states) are not perturbed by the interaction with the reservoir. For the linear coupling (3.2) and high temperature limit $`nn+1`$ in Eq. (3.2) such “pointer states” are represented by eigenstates of the coordinate operator , namely coherent states. In our approach they evolve in time. For the operator (3.48) the eigen- and, respectively, pointer states coincide with the superpositional states $`|\alpha ,\pi `$ given by Eq. (3.1). This coupling preserves the superposition from decoherence, but it is rather difficult to find an experimental system that provides such type of coupling to the environment. We consider a less exotic one, when the majority of the bath modes are not equal in frequency with the selected system, so that loss of amplitude and phase must be delayed. As a simplest example we take the quantum system surrounded by HOs with frequencies doubled to that of the system $`\omega _\xi =2\omega `$ as shown in Fig. 3.6. Although this is still a resonance situation, it leads to unusual behavior compared to the usual case $`\omega _\xi =\omega `$, discussed above. Describing it in RWA we rewrite the interaction (3.2) $$H_{\mathrm{SE}}=\mathrm{}\underset{\xi }{}K_\xi \left[b_\xi ^+a^2+b_\xi (a^+)^2\right],$$ (3.49) as discussed in Ref. . Applying the evolution operator (2.6) we obtain again Eq. (3.3) for the RDM $`\sigma `$ of the selected system, but with some changes of the relaxational part (3.4) $`L\sigma `$ $`=`$ $`\mathrm{\Gamma }(n_\xi +1)\left\{[a^2\sigma ,(a^+)^2]+[(a^+)^2,\sigma a^2]\right\}`$ (3.50) $`+`$ $`\mathrm{\Gamma }n_\xi \left\{[(a^+)^2\sigma ,a^2]+[a^2,\sigma (a^+)^2]\right\},`$ where $`\mathrm{\Gamma }=\pi K_\xi ^2g_\xi `$ is the decay rate of the vibrational amplitude. Here, the number of quanta in the bath mode $`n_\xi `$, the coupling function $`K_\xi `$, and the density of bath states $`g_\xi `$ are evaluated at $`\omega _\xi =2\omega `$. An analogous equation was derived in Ref. . The zero temperature limit was used in Ref. . In the basis of eigenstates $`|n`$ of the unperturbed oscillator the master eq. (3.50) contains only linear combinations of terms as $`\sigma _{m,n}=m\left|\sigma \right|n`$, $`\sigma _{m+2,n+2}`$, and $`\sigma _{m2,n2}`$. It distinguishes even and odd initial states of the system. The odd state $`|1`$ cannot relax to the ground state $`|0`$, but the even state $`|2`$ can. The evolution of the system from different initial conditions was simulated numerically. The equations of motion of the DM elements are integrated using a fourth-order Runge-Kutta algorithm with stepsize control. To make the set of differential equations a finite one we restrict the number of levels by $`m,n20`$. We show the time dependence of the mean value of the coordinate in Fig. 3.7. The mean value in the usual case $`\omega _\xi =\omega `$ decreases with a constant rate. The same initial value of the system coupled to a bath with $`\omega _\xi =2\omega `$ shows a fast decrement in the first stage and almost no decrement afterwards. At temperature $`k_BT=2\mathrm{}\omega /\mathrm{ln3}`$, corresponding to $`n(2\omega )=0.5`$, we have simulated the evolution of the same superpositional states for $`\omega _\xi =\omega `$ and for $`\omega _\xi =2\omega `$, presented in Fig. 3.8(a) and 3.8(b). For $`\omega _\xi =\omega `$ the quantum interference disappears already during the first period, while the amplitude decreases only slightly. The bath with $`\omega _\xi =2\omega `$ leaves the interference almost unchanged, although a fast decrease of the amplitude occurs. Therefore, it partially conserves the quantum superpositional state. ### 3.5 Summary Starting from von Neumann’s equation for a HO interacting with the environment modeled by a set of independent HOs we derived a non-Markovian master equation, which has been solved analytically. For two types of the bath with maximum of the spectral density near the system frequency and near the double of the system frequency and for two different initial states, namely a coherent state and a superposition of coherent states, the wave packet dynamics in coordinate representation have been analysed. It has been shown that wave packet dynamics demonstrates either ”classical squeezing” and the decrease of the effective vibrational oscillator frequency due to the phase-dependent interaction with the bath, or a time-dependent relaxation rate, distinct for even and odd states, and partial conservation of quantum superposition due to the quadratic interaction with the bath. The decoherence also shows differences compared to the usual damping processes adopting RWA and to the description using the pointer basis for decoherence processes. We conclude that there is no permanent pointer basis for the decay. There are two universal stages of relaxation: the coherence stage and the Markovian stage of relaxation, both having different pointer bases. We believe that the proposed method can be applied for other initial states and different couplings with the environment in real existing quantum systems, which is important in the light of recent achievements in single molecule spectroscopy, trapped ion states engineering, and quantum computation. ## Chapter 4 Electron Transfer via Bridges The present chapter deals with an important and quite difficult problem - the mathematical modeling of ET. The purpose of our investigation is, at the one hand, to present a simple, analytically solvable model based on the RDM formalism and to apply it to a porphyrin-quinone complex which is taken as a model system for the reaction centers in bacterial photosynthesis, and on the other hand, to compare this model with another one, which below we call the vibronic model. Before the description of these models and mathematical approaches some brief review of the ET problem in experimental and theoretical contexts is represented in section 4.1. In section 4.2 we introduce the model of a molecular aggregate where only electronic states are taken into account because it is assumed that the vibrational relaxation is much faster than the ET. This model is referred to as the tight-binding (TB) model or model without vibrations below. The properties of an isolated aggregate are modeled in subsection 4.2.1, as well as the static influence of the environment. The dynamical influence of bath fluctuations is discussed and modeled by a heat bath of HOs in subsection 4.2.2. The RDMEM describing the excited state dynamics of the porphyrin aggregate is described in subsection 4.2.3 and compared with an analogous equation of Haken, Strobl, and Reineker (HSR) in appendix A. In subsection 4.2.4 the system parameter dependence on the solvent dielectric constant is discussed for different models of solute-solvent interaction. In Subsection 4.3 system parameters are determined. The methods and results of the numerical and analytical solutions of the RDMEM are presented in subsection 4.4. The dependencies of the transfer rate and final acceptor population on the system parameters are given for the numerical and analytical solutions in subsection 4.5.1. The analysis of the physical processes in the system is also performed there. In subsection 4.5.2 we discuss the dependence of the transfer rate on the solvent dielectric constant for different models of solute-solvent interaction and compare the calculated transfer rates with the experimentally measured ones. The advantages and disadvantages of the presented method in comparison with the method of Davis et al. are analysed in subsection 4.5.3. The vibronic model is described in section 4.6. In this case one pays attention to the fact that experiments in systems similar to the one discussed here show vibrational coherence . Therefore a vibrational substructure is introduced for each electronic level within a multi-level Redfield theory . The comparison of this model with the first one is done in section 4.7. At the end of the chapter the achievements and possible extensions of this consideration are discussed. Unless otherwise stated SI units are used. ### 4.1 Introduction to the electron transfer problem Long-range ET is a very actively studied area in chemistry, biology, and physics; both in biological and synthetic systems. Of special interest are systems with a bridging molecule between donor and acceptor. For example the primary step of charge separation in the bacterial photosynthesis takes place in such a system . #### 4.1.1 Mechanism of the electron transfer in bridge systems It is known that the electronic structure of the bridge component in donor-bridge-acceptor systems plays a critical role . Change of a building block of the complex or change of the environment can modify which mechanism is mainly at work: coherent superexchange or incoherent sequential transfer. Here the bridge energy stands for the energy of the state with electron localized on the bridge, not the locally excited electronic state of the bridge. When the bridge energy is much higher than the donor and acceptor energies, the bridge population is close to zero for all times and the bridge site just mediates the coupling between donor and acceptor. This mechanism is called superexchange and was originally proposed by Kramers to describe the exchange interaction between two paramagnetic atoms spatially separated by a nonmagnetic atom. In the case when donor and acceptor as well as bridge energies are closer than $`k_\mathrm{B}T`$ or the levels are arranged in the form of a cascade, the bridge site is actually populated and the transfer is called sequential. The interplay between these two types of transfer has been investigated theoretically in various publications . Actually, there is still a discussion in the literature whether sequential transfer and superexchange are limiting cases of one process or whether they are two processes which can coexist . To clarify which mechanism is present in an artificial system one can systematically vary both energetics of donor and acceptor and electronic structure of the bridge. In experiments this is done by substituting parts of the complexes or by changing the polarity of the solvent . Also the geometry and size of the bridging block can be varied, and in this way the length of the subsystem through which the electron has to be transfered can be changed. Superexchange occurs due to coherent mixing of the three or more states of the system . The transfer rate in this channel depends algebraically on the differences between the energy levels and decreases exponentially with increasing length of the bridge . When incoherent effects such as dephasing dominate, the transfer is mainly sequential , i. e., the levels are occupied mainly in sequential order . The dependence on the differences between the energy levels is exponential . An increase of the bridge length induces only a small reduction in the transfer rate . This is why sequential transfer is the desired process in molecular wires . #### 4.1.2 Known mathematical theories In the case of coherent superexchange the dynamics is mainly Hamiltonian and can be described on the basis of the Schrödinger equation. The physically important results can be obtained by perturbation theory and, most successfully, by the semiclassical Marcus theory . The complete system dynamics can directly be extracted by numerical diagonalisation of the Hamiltonian . In case of sequential transfer the environmental influence has to be taken into account. There are quite a few different ways how to include the influence of an environment modeled by a heat bath. The simplest phenomenological descriptions of the environmental influence are based on the Einstein coefficients or on the imaginary terms in the Hamiltonian , as well as on the Fokker-Planck or Langevin equations . The most accurate but also numerically most expensive way is the path integral method . This has been applied to bridge-mediated ET especially in the case of bacterial photosynthesis . Bridge-mediated ET has also been investigated using Redfield theory , by propagating a DM in Liouville space and other methods (e. g. ). In most of these methods vibrations are taken into account. The master equation which governs the DM evolution as well as the appropriate relaxation coefficients can be derived from such basic information as system-environment coupling strength and spectral density of the environment . In the model without vibrations the relaxation is introduced in a way similar to Redfield theory but in site representation instead of eigenstate representation. A discussion of advantages and disadvantages of site versus eigenstate representation has been given elsewhere . The equations obtained are similar to those of Ref. where relaxation is introduced in a phenomenological fashion but only a steady-state solution is found in contrast to the model used in this chapter. In addition, the present model is applied to a concrete system. A comparison of the ET time with the bath correlation time allows us to regard three time intervals of system dynamics: the interval of memory effects, the dynamical interval, and the kinetic, long-time interval . In the framework of DM theory one can describe the ET dynamics in all three time intervals. However, often it is enough to find the solution in the kinetic interval for the explanation of experiments within the time resolution of most experimental setups, as has been done in Ref. . The master equation is analytically solvable only for simple models, for example . Most investigations are based on the numerical solution of this equation . However, an estimation can be obtained within the steady-state approximation . We should underline here that the RDMEM contains the coherent dynamics term and, therefore, is able to describe the SE mechanism. This fact has been recognized and successfully used to describe the reactions with dominance of superexchange in a various publications . Davis et al. have explicitly rederived the McConnel analytical expression for superexchange using the RDMEM technique. based on the ability of the diabatic RDMEM to account for the coherent effects we apply the RDMEM formalism to describe the ET in a real system where superexchange most probably takes place. #### 4.1.3 Porphyrin-quinone complex as model for electron transfer The photoinduced ET in the supermolecule consisting of three sequentially connected molecular blocks, i. e., donor (D), bridge (B), and acceptor (A), is under consideration throughout this chapter. D is not able to transfer its charge directly to A because of their spatial separation. D and A can exchange their charges only through B. In the present investigation, the supermolecular system consists of free-base of tetraphenylporphyrin ($`\mathrm{H}_2\mathrm{P}`$) as D, zinc-tetraphenylporphyrin ($`\mathrm{ZnP}`$) as B, and p-benzoquinone as A as shown in Fig. 4.1. ### 4.2 Model without vibrations In each of those molecular blocks shown in Fig. 4.1 we consider only two molecular orbitals, the lowest unoccupied molecular orbital (LUMO) and the highest occupied molecular orbital (HOMO) . Each of the above-mentioned orbitals can be occupied by an electron or not, denoted by $`|1`$ or $`|0`$, respectively. This model allows to describe four states of the molecular block (e. g. D), the neutral ground state $`|1_{\mathrm{HOMO}}|0_{\mathrm{LUMO}}`$ ($`\mathrm{D}`$), the neutral excited state $`|0_{\mathrm{HOMO}}|1_{\mathrm{LUMO}}`$ ($`\mathrm{D}^{}`$),, the positively charged ionic state $`|0_{\mathrm{HOMO}}|0_{\mathrm{LUMO}}`$ ($`\mathrm{D}^+`$),, and the negatively charged ionic state $`|1_{\mathrm{HOMO}}|1_{\mathrm{LUMO}}`$ ($`\mathrm{D}^{}`$),. Below a small roman index denotes the molecular orbital ($`m=0`$ \- HOMO, $`m=1`$ \- LUMO), while a capital index denotes the molecular block ($`M=1`$ \- D, $`M=2`$ \- B, $`M=3`$ \- A). A state of the supermolecule can be described as the direct product of the molecular block states. $`c_{Mm}^+=|1_{Mm}0|_{Mm}`$, $`c_{Mm}=|0_{Mm}1|_{Mm}`$, and $`\widehat{n}_{Mm}=c_{Mm}^+c_{Mm}`$ describe the creation, annihilation, and number of electrons in orbital $`Mm`$, respectively, while $`\widehat{n}_M=_m\widehat{n}_{Mm}`$ gives the number of electrons in a molecular block. The number of particles in the whole supermolecule is conserved $`_M\widehat{n}_M=const`$. Some of the electronic states of the molecular aggregate are shown in Fig. 4.2. Each of the electronic states has its own vibrational substructure. As a rule for the porphyrin containing systems the time of vibrational relaxation is found to be two orders of magnitude faster than the characteristic time of the ET . Because of this we assume that only the vibrational ground states play a role in ET, and we do not include the vibrational structure. A comparison of the models with and without vibrational substructure will be given in section 4.6. #### 4.2.1 System part of the Hamiltonian For the description the charge transfer and other dynamical processes in the system placed in a dissipative environment we use the common form of the Hamiltonian (2.1) where $`\widehat{H}^\mathrm{S}`$ is the Hamiltonian of the supermolecule, $`\widehat{H}^\mathrm{E}`$ the Hamiltonian of the dissipative bath, and $`\widehat{H}^{\mathrm{SE}}`$ describes their interaction. As mentioned in the introduction we are mainly interested in the kinetic limit of the excited state dynamics here. For this limit we assume that the relaxation of the solvent takes only a very short time compared to the timescale of interest for the system. Here the effect of the solvent is considered to be twofold. On the one hand the system states are shifted in energy. This static effect is state-specific and discussed below. On the other hand the system dynamics is perturbed by the solvent state fluctuations, which are independent of the system states. The interaction Hamiltonian shall only reflect the dynamical influence of the fluctuations leading to dissipative processes as discussed in the next subsection. The static influence of the solvent is determined by the relaxed value of the solvent polarization and in general also includes the non-electrostatic contributions such as van-der-Waals attraction and short-range repulsion . It is included into the system state energies and modeled as a function of the dielectric constant of the solvent. The static influence induces a change in the energy levels , $$\widehat{H}^\mathrm{S}=\widehat{H}_0+\widehat{H}_{\mathrm{es}}+\widehat{V},$$ (4.1) where the energy of free and noninteracting blocks $`\widehat{H}_0`$ corresponds to the energy of independent electrons in the field of the ionic nuclei. The term $`\widehat{H}_{\mathrm{es}}`$ denotes the state-selective electrostatic interaction within a molecular aggregate depending on the static dielectric constant $`ϵ_\mathrm{s}`$ of the solvent and $`\widehat{V}`$ the inter-block hopping. It is assumed that the hopping $`\widehat{V}`$ within the supermolecule is affected by the surroundings as discussed in subsection 4.2.4. The energies of the independent electrons can be calculated by $`\widehat{H}_0=_{Mm}E_{Mm}\widehat{n}_{Mm}`$, where $`E_{Mm}`$ denotes the energy of orbital $`Mm`$ in the independent particle approximation . We introduce such a simplified model to be able to calculate the energies of ionic molecular blocks e. g. $`\mathrm{D}^{}`$. For each molecular block a fitting parameter $`A_M`$ is introduced which is used to reproduce the ground state-excited state transition e. g. $`\mathrm{D}\mathrm{D}^{}`$. Because these transitions change only a little for different solvents , the parameters $`A_M`$ are assumed to be solvent-independent. This is why we do not scale $`\widehat{H}_0`$. In order to determine $`E_{Mm}`$ one starts from fully ionized double bonds in each molecular block , calculates the one-particle states in the field of the ions in site representation and fills each of these orbitals with two electrons starting from the lowest orbital; $`E_{Mm}`$ $`=`$ $`{\displaystyle \underset{s=\frac{N_M}{2}d_M+\mathrm{int}\left(\frac{d_M}{2}\right)}{\overset{\frac{N_M}{2}+\mathrm{int}\left(\frac{d_M}{2}\right)}{}}}\left(\stackrel{~}{E}_M2A_M\mathrm{cos}{\displaystyle \frac{2\pi s}{N_M}}\right)`$ $`+`$ $`m\left(\stackrel{~}{E}_M2A_M\mathrm{cos}\left\{{\displaystyle \frac{2\pi }{N_M}}\left[{\displaystyle \frac{N_M}{2}}+\mathrm{int}\left({\displaystyle \frac{d_M}{2}}\right)+1\right]\right\}\right).`$ Here $`N_M`$ and $`d_M`$ are the total number of bonds and number of double bonds, respectively, in the porphyrin rings ($`M=1`$, $`2`$. For $`M=3`$, i. e., the quinone (Q), see Sect. 4.3.). The energy shift $`\stackrel{~}{E}_M`$ is chosen such that the neutral complex has zero energy. $`A_M`$ denotes the energy of hopping between two neighboring sites of the $`M`$th molecular block. The function $`\mathrm{int}()`$ in Eq. (4.2.1) denotes the integer part of a real number. In a similar way, by exciting, removing, or adding the last electron to the model system, one obtains the energy of the excited, oxidized, or reduced molecular block in the independent particle approximation. Below we apply Eqs. (2.27)-(2.5) to the problem of evolution of a single charge-transfer exciton states in the system. In this case the number of states coincides with the number of sites in system: $$\{Mm\}\mu .$$ (4.3) The next contribution to the system Hamiltonian is the inter-block hopping term $$\widehat{V}=\underset{MN}{}v_{\mu \nu }(\widehat{V}_{\mu \nu }^++\widehat{V}_{\mu \nu })\left[(\widehat{n}_M1)^2+(\widehat{n}_N1)^2\right].$$ It includes the hopping operator between two LUMO states $`\widehat{V}_{\mu \nu }=c_{M1}^+c_{N1},`$ (4.4) as well as the corresponding intensities $`v_{\mu \nu }`$, i.e., the coherent coupling between different states of the system. We assume $`v_{13}=0`$ because there is no direct connection between donor and acceptor. The scaling of $`v_{\mu \nu }`$ for different solvents is discussed in subsection 4.2.4. The electrostatic interaction $`\widehat{H}_{\mathrm{es}}`$ scales like energies of a system of charges in a single or in multiple cavities surrounded by a medium with dielectric constant $`ϵ_\mathrm{s}`$ according to the classical reaction field theory . Here we consider two models of scaling. In the first model each molecular block of the aggregate is in the individual cavity as shown in Fig. 4.3. For this case the electrostatic energy reads $$\widehat{H}_{\mathrm{es}}=S^H(ϵ_\mathrm{s})\left(\widehat{H}_{\mathrm{el}}+\widehat{H}_{\mathrm{ion}}\right).$$ (4.5) Here the function $`S^H(ϵ_\mathrm{s})`$ describes the scaling of the electrostatic energy with the static dielectric constant $`ϵ_\mathrm{s}`$ of the solvent. The term $`\widehat{H}_{\mathrm{el}}={\displaystyle \underset{\mu }{}}(\widehat{n}_\mu 1){\displaystyle \frac{e^2}{4\pi ϵ_0}}{\displaystyle \frac{1}{r_\mu }}`$ (4.6) takes the electron interaction into account while bringing an additional charge onto the block $`\mu `$ and thus describes the energy to create an isolated ion. This term depends on the characteristic radius $`r_\mu `$ of the molecular block. The interaction between the ions $`\widehat{H}_{\mathrm{ion}}={\displaystyle \underset{\mu }{}}{\displaystyle \underset{\nu }{}}(\widehat{n}_\mu 1)(\widehat{n}_\nu 1){\displaystyle \frac{e^2}{4\pi ϵ_0}}{\displaystyle \frac{1}{r_{\mu \nu }}}`$ (4.7) depends on the distance between the molecular blocks $`r_{\mu \nu }`$. Both distances $`r_\mu `$ and $`r_{\mu \nu }`$ are also used in the Marcus theory . The term $`H_{\mathrm{el}}+H_{\mathrm{ion}}`$ reflects the interaction of charges inside the aggregate which are compensated by the reaction field according to the Born formula $$S^H=1+\frac{1ϵ_\mathrm{s}}{2ϵ_\mathrm{s}}=\frac{1}{2ϵ_\mathrm{s}}+\frac{1}{2}.$$ (4.8) In the second model sketched in Fig. 4.4, considering the aggregate as an single object placed in a cavity of constant radius one has to use the Onsager term . This term is state selective, i.e., it gives a contribution only for the states with nonzero dipole moment, i.e., charge separation. Defining the static dipole moment operator as $`\widehat{\stackrel{}{p}}=\underset{\mu \nu }{}(\widehat{n}_\mu 1)(\widehat{n}_\nu 1)\stackrel{}{r}_{\mu \nu }e`$ we obtain the Onsager term: $`\widehat{H}_{\mathrm{es}}=S^H\widehat{H}_{\mathrm{dip}}`$, where $`\widehat{H}_{\mathrm{dip}}=\frac{\widehat{\stackrel{}{p}}^2}{r_{13}}`$, $`S^H`$ $`=`$ $`{\displaystyle \frac{1ϵ_\mathrm{s}}{2ϵ_\mathrm{s}+1}}.`$ (4.9) So, the physical meaning of both scalings is briefly summarized as follows: Born-Marcus scaling corresponds to three cavities in the dielectric, each containing one molecular block of the aggregate. The energy scales with the Born formula (4.8). Onsager scaling Eq. (4.9) reflects the naive idea that the whole supermolecule is placed in a single cavity. #### 4.2.2 Microscopic motivation of the system-bath interaction and the thermal bath One can express the dynamic part of the system-bath interaction as $`\widehat{H}^{\mathrm{SE}}={\displaystyle d^3\stackrel{}{r}\underset{\mu \nu }{}\widehat{\stackrel{}{D}}_{\mu \nu }(\stackrel{}{r})\mathrm{\Delta }\widehat{\stackrel{}{P}}(\stackrel{}{r})}.`$ (4.10) Here $`\widehat{\stackrel{}{D}}_{\mu \nu }(\stackrel{}{r})`$ denotes the field of the electrostatic displacement at point $`\stackrel{}{r}`$ induced by the system transition dipole moment $`\widehat{\stackrel{}{p}}_{\mu \nu }=\stackrel{}{p}_{\mu \nu }(\widehat{V}_{\mu \nu }^++\widehat{V}_{\mu \nu })`$ $`\widehat{\stackrel{}{D}}_{\mu \nu }(\stackrel{}{r})={\displaystyle \frac{1}{4\pi ϵ_0}}\left[{\displaystyle \frac{3(\widehat{\stackrel{}{p}}_{\mu \nu }\stackrel{}{r})\stackrel{}{r}}{r^5}}{\displaystyle \frac{\widehat{\stackrel{}{p}}_{\mu \nu }}{r^3}}\right].`$ (4.11) The field of the environmental polarization s denoted as $`\widehat{\stackrel{}{P}}(\stackrel{}{r})=_n\delta (\stackrel{}{r}\stackrel{}{r}_n)\widehat{\stackrel{}{d}}_n`$, where $`\widehat{\stackrel{}{d}}_n`$ is the $`n`$th dipole of the environment and $`\stackrel{}{r}_n`$ its position. Only fluctuations of the environment polarization $`\mathrm{\Delta }\widehat{\stackrel{}{P}}(\stackrel{}{r})`$ influence the system dynamics. Averaged over the angular dependence the interaction reads $`\widehat{H}^{\mathrm{SE}}={\displaystyle \underset{\mu \nu n}{}}{\displaystyle \frac{1}{4\pi ϵ_0}}\left({\displaystyle \frac{2}{3}}\right)^{\frac{1}{2}}{\displaystyle \frac{|\widehat{\stackrel{}{p}}_{\mu \nu }|\mathrm{\Delta }|\widehat{\stackrel{}{d}_n}|}{|\stackrel{}{r}_n|^3}}.`$ (4.12) The dynamical influence of the solvent is described with the thermal bath model. The deviation $`\mathrm{\Delta }\left|\widehat{\stackrel{}{d}}_n\right|`$ of $`d_n`$ from its mean value is determined by temperature induced fluctuations. One could couple $`\widehat{V}_{\mu \nu }`$ to $`\widehat{\stackrel{}{d_n}}`$ or $`\mathrm{\Delta }\left|\widehat{\stackrel{}{d}}_n\right|`$. But, solvent dipoles interact with each other even in the absence of a supermolecular aggregate. For unpolar solvents described by a set of HOs the diagonalisation of their interaction yields the bath of HOs with different frequencies $`\omega _\lambda `$ and effective masses $`m_\lambda `$. In the case of a polar solvent the dipoles are interacting rotators as, e. g. used to describe magnetic phenomena . The elementary excitation of each frequency can again be characterised by an appropriate HO. So, we use generalized coordinates of solvent oscillators modes $`\widehat{Q}_\lambda =\sqrt{{\displaystyle \frac{\mathrm{}}{2m_\lambda \omega _\lambda }}}(\widehat{b}_\lambda +\widehat{b}_\lambda ^+)`$ (4.13) for polar as well as unpolar solvents. The occupation of the $`i`$th state of the $`\lambda `$th oscillator is defined by the equilibrium DM $`\rho _{\lambda ,ij}=\mathrm{exp}\left(\frac{\mathrm{}\omega _\lambda i}{k_BT}\right)\delta _{ij}`$. All mutual orientations and distances of solvent molecules have equal probability. An average over all spatial configurations is performed. The interaction Hamiltonian (4.12) is written in a form which is bilinear in system and bath operators: $`\widehat{H}^{\mathrm{SE}}=\left[{\displaystyle \underset{\mu \nu }{}}p_{\mu \nu }(\widehat{V}_{\mu \nu }+\widehat{V}_{\mu \nu }^+)\right]\left[{\displaystyle \underset{\lambda }{}}K_\lambda (\widehat{b}_\lambda ^++\widehat{b}_\lambda )\right]`$ (4.14) The coefficients $`K_\lambda ={\displaystyle \frac{1}{4\pi ϵ_0}}\left({\displaystyle \frac{2}{3}}\right)^{\frac{1}{2}}{\displaystyle \frac{d^3\stackrel{}{r}}{|\stackrel{}{r}|^3}e\sqrt{\frac{\mathrm{}}{2m_\lambda \omega _\lambda }}S^{\mathrm{SE}}(ϵ_\mathrm{s})}`$ (4.15) depend on properties of the solvent, in particular, the frequencies $`\omega _\lambda `$. The precise determination of these coefficients needs special consideration. Expression (4.12) includes the dipole moment values corresponding to environmental modes and system transitions. The electric field of a dipole in medium is equivalent to the field of an imaginary dipole with a moment depending on the properties of the medium . This influence of the medium is reflected here by the scaling function $`S^{\mathrm{SE}}`$. Explicit expressions for the solvent influence are still under discussion in the literature . #### 4.2.3 Reduced density matrix approach As usual the bath is given by HOs which describe the irradiative relaxation properties of the system. For the full description of the system one also should include photon modes to describe for example the fluorescence from the LUMO to the HOMO in each molecular block transferring an excitation to the electro-magnetic field with rate $`G_\mu `$ . The rate of the radiative processes is small in comparison to other processes in system. That is why only the irradiative contribution is treated below. The treatment is similar to Redfield theory . For sake of notation and completeness we repeat the most important steps here for our model Hamiltonian. The irradiative contribution of the system-bath interaction corresponds to energy transfer to the solvent and spreading of energy over vibrational modes of the supermolecule. Applying the RWA to Eq. (4.14) one gets $`\widehat{H}^{\mathrm{SE}}={\displaystyle \underset{\lambda }{}}{\displaystyle \underset{\mu \nu }{}}K_\lambda p_{\mu \nu }\widehat{b}_\lambda ^+\widehat{V}_{\mu \nu }+h.c.,`$ (4.16) where $`K_\lambda p_{\mu \nu }`$ denotes the interaction intensity between the bath mode $`b_\lambda `$ of frequency $`\omega _\lambda `$ and the quantum transition $`\mu \nu `$ between the LUMOs of molecules $`\mu `$ and $`\nu `$ of frequency $`\omega _{\mu \nu }=\frac{1}{\mathrm{}}\left(E_\mu E_\nu \right)`$. The dynamics of the system plus bath cannot be calculated due to the huge number of degrees of freedom. Therefore one uses the RDMEM technique. Below we assume that the coherent and dissipative dynamics can be represented by the independent terms of the RDMEM. This assumption is applicable if $`v_{\mu \nu }\omega _{\mu \nu }`$. Here we use the RDMEM (2.5) derived in section 2.5. Here we treat the system $`p_{\mu \nu }`$ and bath $`K_\lambda `$ contributions to the system-bath coupling separately. That is why we define the spectral density of bath modes as $`J(\omega )=\pi _\lambda K_\lambda ^2\delta (\omega \omega _\lambda )`$ and the relaxation constant for Eq. (2.5) $$\mathrm{\Gamma }_{\mu \nu }=\mathrm{}^2J(\omega _{\mu 1\nu 1})p_{\mu \nu }^2$$ (4.17) depends on the coupling of the transition $`|\mu |\nu `$ to the bath mode of the same frequency. Formally, the damping constant depends on the density of bath modes $`J`$ at the transition frequency $`\omega _{\mu \nu }`$ and on the transition dipole moments between the system states $`p_{\mu \nu }`$. For the sake of concrete calculations we write the RDMEM in matrix form. Substituting the expressions for the exciton density $`\widehat{\sigma }=\underset{\mu \nu }{}\sigma _{\mu \nu }|\nu \nu |`$, and operators Eq. (4.4) into the relaxation term Eq. (2.5) yields: $`(L_{\mathrm{RWA}}\sigma )_{\kappa \lambda }=`$ $`2\delta _{\kappa \lambda }{\displaystyle \underset{\mu }{}}\left\{\mathrm{\Gamma }_{\mu \kappa }\left[n(\omega _{\mu \kappa })+1\right]+\mathrm{\Gamma }_{\kappa \mu }n(\omega _{\kappa \mu })\right\}\sigma _{\mu \mu }`$ (4.18) $`{\displaystyle \underset{\mu }{}}\{\mathrm{\Gamma }_{\mu \kappa }[n(\omega _{\mu \kappa })+1]+\mathrm{\Gamma }_{\kappa \mu }n(\omega _{\kappa \mu })`$ $`+\mathrm{\Gamma }_{\mu \lambda }[n(\omega _{\mu \lambda })+1]+\mathrm{\Gamma }_{\lambda \mu }n(\omega _{\lambda \mu })\}\sigma _{\kappa \lambda }.`$ A RDMEM of similar structure was used for the description of exciton transfer by Haken, Strobl, and Reineker in . We present the comparison of these RDMEMs with Eq. (4.18) in appendix A. For the description of a concrete system the introduced RDMEM (2.27)-(2.5) is used in the matrix form Eq. (4.18) with the use of an index simplification Eq. (4.3). For the sake of convenience of analytical and numerical calculations we replace the relaxation constant $`\mathrm{\Gamma }_{\mu \nu }`$ and the population of the corresponding bath mode $`n(\omega _{\mu \nu })`$ with the intensity of dissipative transitions $`d_{\mu \nu }=\mathrm{\Gamma }_{\mu \nu }|n(\omega _{\mu \nu })|`$ between two states, as well as the corresponding dephasings intensity $`\gamma _{\mu \nu }={\displaystyle \underset{\kappa }{}}\left(d_{\mu \kappa }+d_{\kappa \nu }\right).`$ (4.19) With this one can express the RDMEM (4.18) in the form $`\dot{\sigma }_{\mu \mu }`$ $`=`$ $`i/\mathrm{}{\displaystyle \underset{\nu }{}}(v_{\mu \nu }\sigma _{\nu \mu }\sigma _{\mu \nu }v_{\nu \mu })+(L\sigma )_{\mu \mu }`$ (4.20) $`\dot{\sigma }_{\mu \nu }`$ $`=`$ $`(i\omega _{\mu \nu }\gamma _{\mu \nu })\sigma _{\mu \nu }i/\mathrm{}v_{\mu \nu }(\sigma _{\nu \nu }\sigma _{\mu \mu }),`$ (4.21) where $$(L\sigma )_{\mu \mu }=\underset{\nu }{}d_{\mu \nu }\sigma _{\mu \mu }+\underset{\nu }{}d_{\nu \mu }\sigma _{\nu \nu }.$$ (4.22) From the manifold of system states we choose the ones which play the essential role in the electron transfer from the donor to acceptor Fig. 4.2. They can be described in terms of single charge transfer exciton and correspond to the index simplification Eq. (4.3). The parameters controlling the transitions between the selected states are discussed in the next section 4.3 and shown explicitly in Fig. 4.5. #### 4.2.4 Scaling of the relaxation coefficients The relaxation coefficients $`\mathrm{\Gamma }_{\mu \nu }`$ include the second power of the scaling function $`\mathrm{\Gamma }_{\mu \nu }(ϵ)=\mathrm{\Gamma }_{\mu \nu }^0S_{}^{\mathrm{SE}}{}_{}{}^{2}(ϵ)`$ because one constructs the relaxation term Eq. (2.24) with the second power of the interaction Hamiltonian. The physical meaning of $`H^{\mathrm{SE}}`$ is similar to the interaction of the system dipole with a surrounding media. That is why it is reasonable to use the relevant expressions from the relaxation theory of Onsager and Kirkwood $`S^{\mathrm{SB}}=\frac{1ϵ_\mathrm{s}}{2ϵ_\mathrm{s}+1}`$. In the work of Mataga, Kaifu, and Koizumi the interaction energy between the system dipole and the media scales in leading order as $$S^{\mathrm{SE}}=\left[\frac{2(ϵ_\mathrm{s}1)}{2ϵ_\mathrm{s}+1}\frac{2(ϵ_{\mathrm{}}1)}{2ϵ_{\mathrm{}}+1}\right],$$ (4.23) where $`ϵ_{\mathrm{}}`$ denotes the optical dielectric constant. In a recent paper of Georgievskii, Hsu, and Marcus a connection $`J(\omega )=\mathrm{}\alpha (\omega )`$ between $`J(\omega )`$ and the generalized susceptibility $`\alpha (\omega )`$ is introduced and the explicit dependence $`\alpha (\omega )`$ on the dielectric parameters of the solvent is given. In application to the present work this means that the relaxation coefficient scales like $`\mathrm{\Gamma }_{\mu \nu }=S_{}^{\mathrm{SE}}{}_{}{}^{2}\mathrm{}^2J(\omega _{\mu \nu })p_{\mu \nu }^2`$. For example the approximation of spherical molecules gives $`\alpha (\omega )\frac{1}{ϵ_\mathrm{s}}\frac{1}{ϵ_\omega }`$. We approximate $`ϵ_\omega =ϵ_{\mathrm{}}`$ here, so $`\mathrm{\Gamma }\frac{1}{ϵ_\mathrm{s}}\frac{1}{ϵ_{\mathrm{}}}`$. In terms of scaling function it can be expressed as $`S^{\mathrm{SE}}=\left({\displaystyle \frac{1}{ϵ_\mathrm{s}}}{\displaystyle \frac{1}{ϵ_{\mathrm{}}}}\right)^{\frac{1}{2}}.`$ (4.24) As an alternative possibility one can test the case of solvent-independent relaxation coefficient $`\mathrm{\Gamma }_{\mu \nu }=const`$, $`S^{\mathrm{SE}}=1`$. The coherent coupling $`v_{\mu \nu }`$ between two electronic states scales with $`ϵ_\mathrm{s}`$ and $`ϵ_{\mathrm{}}`$ too, because a coherent transition in the system is accompanied by a transition of the environment state which is larger for the solvents with larger polarity. Here we involve the concept of reorganization energy from the model with vibrational substructure to account for this effect. For the reorganization energy we take the static part of the system-bath interaction calculated in frames of either individual or multiple cavity model. Here we take the static part $`H_{}^{\mathrm{SE}}{}_{}{}^{\mathrm{s}}=S^{\mathrm{SE}}\left(\widehat{H}_{\mathrm{el}}+\widehat{H}_{\mathrm{ion}}\right)`$, note that the electronic part of $`H_{}^{\mathrm{SE}}{}_{}{}^{\mathrm{s}}`$ does not scale for the single cavity model. $`H_{}^{\mathrm{SE}}{}_{}{}^{\mathrm{s}}`$ is associated with the so-called reorganization energy. The coherent couplings decrease with increase of $`H_{}^{\mathrm{SE}}{}_{}{}^{\mathrm{s}}`$. For the bath of relatively high frequency HOs (like the $`CC`$ stretching vibrations) this scaling can be taken as $$v_{\mu \nu }=v_{\mu \nu }^0\mathrm{exp}\left[\left|\mu \nu |H_{}^{\mathrm{SE}}{}_{}{}^{\mathrm{s}}|\mu \nu \left(2\mathrm{}\omega _{\mathrm{vib}}\right)^1\right|\right],$$ (4.25) where $`v_{\mu \nu }^0`$ is the coupling of electronic states of the isolated molecule, $`\omega _{\mathrm{vib}}`$ the leading (mean) environment oscillator frequency. Unless otherwise stated $`\omega _{\mathrm{vib}}=1500\mathrm{cm}^1`$ is used. ### 4.3 Model parameters The dynamics of the system is controlled by the following parameters: energies of system levels $`E_\mu `$, coherent couplings $`v_{\mu \nu }`$, and dissipation intensities $`\mathrm{\Gamma }_{\mu \nu }`$. The simplification is that we do not calculate these parameters, exept B state energy, rather take the corresponding experimental values. The absorption spectra of porphyrins consist of a high frequency Soret band and a low frequency $`Q`$ band. In the case of ZnP the $`Q`$ band has two subbands, corresponding to pure electronic $`Q(0,0)`$ and vibronic $`Q(1,0)`$ transitions. In the free-base porphyrin $`\mathrm{H}_2\mathrm{P}`$ the reduction of symmetry $`D_{4h}D_{2h}`$ due to the substitution of the central Zn ion by two inner hydrogens induces a splitting of each subband into two, namely $`Q^x(0,0)`$, $`Q^y(1,0)`$ and $`Q^x(0,0)`$, $`Q^y(1,0)`$. So the absorption spectra of $`\mathrm{ZnP}`$ and $`\mathrm{H}_2\mathrm{P}`$ consist of two and four bands respectively. In the emission spectra one sees only two bands for each molecule because of cascade intramolecular relaxations: pure electronic and vibronic one. Each of the above-mentioned spectra can be represented as a sum of Lorentzians with a good accuracy. The absorption spectrum of the complex $`\mathrm{H}_2\mathrm{P}\mathrm{ZnP}\mathrm{Q}`$ consists of the spectra of the individual $`\mathrm{H}_2\mathrm{P}`$, ZnP, and Q compounds without essential changes. It means that intermolecular interactions in this case do not change the structure of electronic states of the subunits . We consider only the states corresponding to the lowest band of each spectrum to find the parameters of ET. The respective frequencies $`\nu `$ and widths $`\gamma `$ are shown in Table 4.1 for $`\mathrm{CH}_2\mathrm{Cl}_2`$ as solvent. The width of each band is formed by transitions to different sublevels of the electronic excited states of the aggregate and by the lifetime of these sublevels. In the present model consideration we do not have an excited level substructure, so that we cannot determine the linewidths correctly. Nevertheless we can consider the widths from Table 4.1 as upper limits for the relaxation constants. On the basis of the given spectral data and taking as reference energy $`E_{\mathrm{DBA}}=0`$ we determine $`E_{\mathrm{D}^{}\mathrm{BA}}=1/2(\nu _{00}^x+\nu _{01}^x)=1.82`$ eV and $`E_{\mathrm{DB}^{}\mathrm{A}}=2.03\mathrm{eV}`$ (in $`\mathrm{CH}_2\mathrm{Cl}_2`$) which allows us to extract the hopping energies $`A_\mathrm{D}=3.11\mathrm{eV}`$, $`A^\mathrm{B}=3.45\mathrm{eV}`$ introduced in Eq. (4.2.1). The excitation spectrum of the quinone lies far from the visible range, in the ultraviolet range. We do not have a precise spectral information about $`\mathrm{Q}`$ and, therefore, we do not calculate $`A_\mathrm{A}`$. We simply take the energy of the state with charge transfer to $`\mathrm{Q}`$ from reference : $`E_{\mathrm{D}^+\mathrm{BA}^{}}=1.42\mathrm{eV}`$. . Further Rempel and coauthors estimate the electron coupling of the initially excited and the charged bridge states as $`\mathrm{D}^{}\mathrm{BA}|H|\mathrm{D}^+\mathrm{B}^{}\mathrm{A}=v_{12}^0=65\mathrm{meV}=9.8\times 10^{13}\mathrm{s}^1`$ and they give the matrix element of two states with charge separation $`\mathrm{D}^+\mathrm{B}^{}\mathrm{A}|H|\mathrm{D}^+\mathrm{BA}^{}=v_{23}^0=2.2\mathrm{meV}=3.3\times 10^{12}\mathrm{s}^1`$. These values of the couplings are essentially lower than the energy differences between these corresponding system states $$\mathrm{}\omega _{ij}v_{ij}^0.$$ (4.26) This is the reason why it is useful to remain in site representation instead of eigenstate representation . The relaxation constants are found with the help of the derived analytical formula to be $`\mathrm{\Gamma }_{21}=\mathrm{\Gamma }_{23}=2.25\times 10^{12}\mathrm{s}^1`$. We discuss it in a more detailed form at the end of the next section. The typical radius of the porphyrin ring is about $`r_1=r_2=5.5`$ $`\mathrm{\AA }`$, $`r_3=3.2`$ $`\mathrm{\AA }`$ , while the distance $`r_{\mu \nu }`$ between the blocks of the aggregate $`\mathrm{H}_2\mathrm{P}\mathrm{ZnP}\mathrm{Q}`$ reaches $`r_{12}=12.5\pm 1\mathrm{\AA }`$ , $`r_{23}=7\pm 1\AA `$, $`r_{13}=14.4\pm 1\mathrm{\AA }`$ . The main parameter which controls ET in the triad is the energy of the state $`E_{\mathrm{D}^+\mathrm{B}^{}\mathrm{A}}`$. This state has a big dipole moment because of its charge separation and is therefore strongly influenced by the solvent. Because of the special importance of this value we calculate it for the different solvents as a matrix element of the unperturbed system Hamiltonian $`\widehat{H}_0+\widehat{H}_{\mathrm{es}}`$ in Eq. (4.1). The calculated values of the bridge state $`\mathrm{D}^+\mathrm{B}^{}\mathrm{A}`$ for some solutions are shown in Table 4.2. ### 4.4 Results The time evolution of the ET in the supermolecule is described by Eqs. (4.20)-(4.21) with the initial condition which corresponds to the excitation of the donor with a $`\pi `$-pulse of appropriate frequency, i.e., the population of the donor is set to one. The system of equations was solved numerically and analytically. For the numerical simulation we express the system of Eqs. (4.20)-(4.21) in the form $`\dot{\overline{\sigma }}=\overline{L}\overline{\sigma }`$, where $`\overline{\sigma }`$ is a vector of dimension $`3^2`$ for the model with $`3`$ system states and the super-operator $`\overline{L}`$ is a matrix of dimension $`3^2\times 3^2`$. The time evolution of an element of the DM can be determined by $`\sigma _{\mu \nu }(t)=\sigma _{\mu \nu }(0)_{\kappa \xi }W_{\mu \nu \kappa \xi }\mathrm{exp}(\lambda _{\kappa \xi }t)`$, where $`\lambda _{\kappa \xi }`$ and $`W_{\mu \nu \kappa \xi }`$ are the eigenvalues and eigenvectors of the super-operator $`\overline{L}`$, respectively. These are obtained numerically. When fluorescence does not have to be taken into account, i.e., in the time interval $`tG_\mu ^1`$ (cp. subsect. 4.2.3) all states except $`|\mathrm{D}^{}\mathrm{BA}`$ ($`\mu =1`$), $`|\mathrm{D}^+\mathrm{B}^{}\mathrm{A}`$ ($`\mu =2`$), and $`|\mathrm{D}^+\mathrm{BA}^{}`$ ($`\mu =3`$) remain essentially unoccupied, while those three take part in the intermolecular transport process. At later times $`tG_\mu ^1`$ excitations decay into the ground state. So the physics of the ET processes in the molecular complex can be understood with the dynamics of the three above-mentioned states. The numerical simulation of the system dynamics with the parameters given in the previous section shows an exponential growth of the acceptor population. Such a behavior can be nicely fitted to a single exponential $$P_3(t)=P_3(\mathrm{})\left[1\mathrm{exp}(k_{\mathrm{ET}}t)\right],$$ (4.27) where for the solvent MTHF $`k_{\mathrm{ET}}3.59\times 10^8\mathrm{s}^1`$ and $`P_3(\mathrm{})0.9994`$. The population of the intermediate state $`\mu =2`$ which corresponds to charge localization on the bridge does not reach a value of more than $`0.005`$. This that means in this case the superexchange mechanism dominates over the sequential transfer mechanism. Besides it ensures the validity of characterizing the system dynamics with $`P_3(\mathrm{})`$ and $`k_{\mathrm{ET}}=P_3(\mathrm{})\left\{{\displaystyle _0^{\mathrm{}}}\left[1P_3(t)\right]𝑑t\right\}^1.`$ (4.28) The alternative analytical approach is performed in the kinetic limit $$t1/\mathrm{min}(\gamma _{\mu \nu }).$$ (4.29) In Laplace space the inequality (4.29) reads $`s\mathrm{min}(\gamma _{\mu \nu })`$, where $`s`$ denotes the Laplace variable. It is equivalent to replacing the factor $`1/(i\omega _{\mu \nu }+\gamma _{\mu \nu }+s)`$ in the Laplace transform of Eqs. (4.20)-(4.21) with $`1/(i\omega _{\mu \nu }+\gamma _{\mu \nu })`$. This trick allows to substitute the expressions for non-diagonal elements of the DM Eqs. (4.21) into Eqs. (4.20), so we do not use them explicitly anymore. After this elimination we describe the coherent transitions which occur with the participance of the non-diagonal elements by the following redefinition of the RDMEM (4.20)-(4.22): $$\dot{\sigma }_{\mu \mu }=\underset{\nu }{}g_{\mu \nu }\sigma _{\mu \mu }+\underset{\nu }{}g_{\nu \mu }\sigma _{\nu \nu }.$$ (4.30) In the given expression the transition coefficients $`g_{\mu \nu }`$ contain both, dissipative and coherent contributions $$g_{\mu \nu }=d_{\mu \nu }+v_{\mu \nu }v_{\nu \mu }\gamma _{\mu \nu }\left[\mathrm{}^2\left(\omega _{\mu \nu }^2+\gamma _{\mu \nu }^2\right)\right]^1.$$ (4.31) Now it is assumed that the bridge state is not occupied. This allows us to find the dynamics of the acceptor state in the form of Eq. (4.27), where $`k_{\mathrm{ET}}=g_{23}+{\displaystyle \frac{g_{23}(g_{12}g_{32})}{g_{21}+g_{23}}},`$ (4.32) $`P_3(\mathrm{})={\displaystyle \frac{g_{12}g_{23}}{g_{21}+g_{23}}}(k_{\mathrm{ET}})^1.`$ (4.33) The value of the dissipative coupling $`\mathrm{\Gamma }_{\mu \nu }=\mathrm{}^2J(\omega _{\mu \nu })p_{\mu \nu }^2`$ can be found by comparison of the experimentally determined ET rate and the derived analytical formula Eq. (4.32). To calculate $`J(\omega _{\mu \nu })`$ would require a microscopic model. We want to avoid a microscopic consideration and this is why we simply take the same $`\mathrm{\Gamma }_{\mu \nu }`$ for all transitions between excited states. The value of ET for $`\mathrm{H}_2\mathrm{P}\mathrm{ZnP}\mathrm{Q}`$ in MTHF is found by Rempel et al. to be $`k_{\mathrm{ET}}=3.6\pm .5\times 10^8\mathrm{s}^1`$. If the bridging state has a rather high energy one can neglect thermally activated processes. $`v_{12}`$ is negligible small with respect to $`v_{23}`$. In this case our result Eq. (4.32) reads $`k_{\mathrm{ET}}={\displaystyle \frac{v_{12}^2\mathrm{\Gamma }_{21}}{\mathrm{}^2\omega _{21}^2+\mathrm{\Gamma }_{21}^2}}{\displaystyle \frac{\mathrm{\Gamma }_{23}}{\mathrm{\Gamma }_{21}+\mathrm{\Gamma }_{23}}}.`$ (4.34) Taking this formula, the relation between the dissipation intensities $`\mathrm{\Gamma }_{21}=\mathrm{\Gamma }_{23}`$, and the experimental value for the transfer rate one obtains $`\mathrm{\Gamma }_{21}=\mathrm{\Gamma }_{23}2.25\times 10^{12}\mathrm{s}^1`$. The fit of the numerical solution of the system of Eqs. (4.20)-(4.21) to the experimental value of the transfer rate in MTHF gives the same value. So the dissipative coupling constants are fixed for a specific solvent and for other solvents they are calculated with the scaling functions. With the method presented the ET in the supermolecule was found to occur with dominance of the superexchange mechanism with rates $`4.6\times 10^6\mathrm{s}^1`$ and $`3.3\times 10^8\mathrm{s}^1`$ for CYCLO and $`\mathrm{CH}_2\mathrm{Cl}_2`$, respectively. ### 4.5 Discussion #### 4.5.1 Sequential versus superexchange Here we discuss how the transfer mechanism depends on the change of parameters. Namely which parameter have to be changed in order to alter not only the transfer rate quantitatively, but the dominant mechanism of transfer and the qualitative behavior of the system. In order to answer this question we calculate the system dynamic by varying one parameter at a time, while all other parameters are kept unchanged. The dependencies of transfer rate $`k_{\mathrm{ET}}`$ and final population $`P_3`$ on such parameters as coherent couplings $`v_{12}`$, $`v_{23}`$ and dissipation intensities $`\mathrm{\Gamma }_{21}`$, $`\mathrm{\Gamma }_{23}`$ are shown in the Figs. 4.6-4.7. The change of each parameter influences the transfer in a different way. In particular, $`k_{\mathrm{ET}}`$ depends quadratically on the coherent coupling $`v_{12}`$ from $`10^{15}\mathrm{s}^1`$ to $`10^{12}\mathrm{s}^1`$ in Fig. 4.6. Below it saturates at the lower bound $`k_{\mathrm{ET}}3\times 10^5\mathrm{s}^1`$. This corresponds to a crossover of the transfer mechanism from superexchange to sequential transfer. But, due to the big energy difference between donor and bridging state the efficiency of this sequential transfer is extremely low. This is displayed by $`P_3(\mathrm{})0`$. In the region $`v_{12}v_{23}`$ both mechanisms contribute to the transfer rate. The transfer rate depends on $`v_{23}`$ in a similar way. The decrease of final population in this region corresponds to coherent back transfer. At rather high values of $`v_{12}`$, $`v_{23}10^{15}\mathrm{s}^1`$ the relation (4.26) is no more valid. For this regime one has to use eigenstate instead of site representation because the wavefunctions are no more localized . This variation of the coherent coupling can be performed experimentally by exchanging building blocks of the supermolecule. The variation of the dissipation intensities $`\mathrm{\Gamma }_{21}`$, $`\mathrm{\Gamma }_{23}`$ in the region near the experimental values shows similar behavior of $`k_{\mathrm{ET}}(\mathrm{\Gamma }_{21})`$ and $`k_{\mathrm{ET}}(\mathrm{\Gamma }_{23})`$ (see Fig. 4.7). Here we assume an independent variation of $`\mathrm{\Gamma }_{21}`$ and $`\mathrm{\Gamma }_{23}`$. Both, $`k_{\mathrm{ET}}(\mathrm{\Gamma }_{21})`$ and $`k_{\mathrm{ET}}(\mathrm{\Gamma }_{23})`$ increase linear until the saturation value $`7\times 10^8\mathrm{s}^1`$ at $`\mathrm{\Gamma }>10^{12}\mathrm{s}^1`$ is reached. There is qualitative agreement between the numerical and analytical values. In both dependencies one observes saturation at $`10^{12}\mathrm{s}^1<\mathrm{\Gamma }<10^{13}\mathrm{s}^1`$. In Eq. (4.28) infinite time is approximated by $`10^5\mathrm{s}`$. It is found that approximating this time with larger values we do not obtain essential changes in the results. But from the formal point of view we note that one cannot obtain transfer rates lower than $`10^5\mathrm{s}^1`$. Here we note that taking larger values for the infinite time does not change the transfer rates with the good precision. The physical meaning of the transfer rate dependence on the dissipation intensities seems to be transparent. At small values of $`\mathrm{\Gamma }`$ a part of the population coherently oscillates back and forth between the states. The increase of the dephasing Eq. (4.19) quenches the coherence and makes the transfer irreversible. So transfer becomes faster up to a maximal value. For the whole range of $`\mathrm{\Gamma }`$, depopulations $`d_{21}`$, $`d_{23}`$ and thermally activated transitions $`d_{12}`$, $`d_{32}`$ always remain smaller than the coherent couplings, therefore they do not play an essential role. Next, the similarity of the dependencies on $`\mathrm{\Gamma }_{21}`$ and $`\mathrm{\Gamma }_{23}`$ will be discussed on the basis of Eq. (4.32). Firstly, in the limit $`k_BT/\mathrm{}\omega _{\mu \nu }0`$ the terms corresponding to thermally activated processes $`\omega _{\mu \nu }<0`$ vanish and so $`|n(\omega _{\mu \nu })|=0`$, while depopulations $`\omega _{\mu \nu }>0`$ remain constant $`|n(\omega _{\mu \nu })|=1`$. Secondly the condition $`\omega _{\mu \nu }\gamma _{\mu \nu }`$ allows to neglect $`\gamma _{\mu \nu }^2`$ in comparison with $`\omega _{\mu \nu }^2`$. With the above-mentioned simplifications the analytical expression for the transfer rate Eq. (4.32) becomes $`k_{\mathrm{ET}}\mathrm{\Gamma }_{21}\mathrm{\Gamma }_{23}\left(\mathrm{\Gamma }_{21}+\mathrm{\Gamma }_{23}\right)^1\left(v_{12}^2/\omega _{21}^2+v_{23}^2/\omega _{23}^2\right).`$ (4.35) This equation is symmetric with respect to the relaxation intensities $`\mathrm{\Gamma }_{21}`$ and $`\mathrm{\Gamma }_{23}`$. This is why the transfer rate depends on each of them in the same way. The most crucial change of the transfer dynamics can be induced by changing the energies of the system levels. As was mentioned in section 4.3 these are the only parameters which can be varied continuously by use of composed solvents. To the largest extent the mechanism of transfer depends on the energy of the bridging state $`|\mathrm{D}^+\mathrm{B}^{}\mathrm{A}`$. The results of the corresponding calculations are presented in Fig. 4.8. In different regions one observes different types of dynamics. For large energy of the bridging state $`E_{\mathrm{D}^+\mathrm{B}^{}\mathrm{A}}E_{\mathrm{D}^{}\mathrm{BA}}`$ and, respectively, $`E_{21}=E_{\mathrm{D}^+\mathrm{B}^{}\mathrm{A}}E_{\mathrm{D}^{}\mathrm{BA}}0`$ the numerical and analytical results do not differ from each other. The transfer occurs with the superexchange mechanism. The transfer rate reaches a maximal value of $`10^{11}\mathrm{s}^1`$ for low energies of the bridge. While the bridge energy approaches the energy of the donor state closer than thermal energy and goes even lower than this, the sequential mechanism of transfer starts to contribute to the reaction process. The traditional scheme of sequential transfer is obtained when donor, bridge, and acceptor levels are arranged in the form of a cascade. In this region the analytical solution need not coincide with the numerical result because the used approximations are no more valid. For equal energies of bridging and acceptor states $`k_{\mathrm{ET}}`$ displays a small resonance peak at $`E_{21}=.4`$ as it is seen in Fig. 4.8(a). In the extremal case when the energy of the bridging state is even lower than the energy of the acceptor state a transfer does not take place anymore because the population gets trapped at the bridge state. The finite transfer rate for $$E_{21}<E_{31}$$ (4.36) does not mean the actual ET because $`P_3(\mathrm{})0`$. For the dynamic time interval $`t<\gamma _{\mu \nu }^1`$ a part of the population tunnels force and back to the acceptor state with the rate $`k_{\mathrm{ET}}`$. The analytical expression Eq. (4.32) gives constant rate for the regime (4.36), while the numerical solution of Eqs. (4.20)-(4.21) displays instability, because such coherent oscillations of population cannot be described by Eq. (4.27) and $`k_{\mathrm{ET}}`$ cannot be fitted with Eq. (4.28). In Fig. 4.8 the regime Eq. (4.36) is displayed two times: for small $`E_{21}`$ while $`E_{31}`$ is kept constant and for large $`E_{31}`$ while $`E_{21}`$ remains constant. The energy dependence of the final population has a transparent physical meaning for the whole range of energy. A large value of the bridging level ensures the transition of the whole population to the acceptor state with charge separation $`|\mathrm{D}^+\mathrm{BA}^{}`$ which has the lowest energy of the excited states. In the intermediate case, when the bridging state has the same energy as the acceptor state, final population spreads itself over these two states $`P_3(\mathrm{})=.5`$. #### 4.5.2 Different solvents Lowering the bridging state even more one arrives at the situation, where the whole population remains on the bridge as the lowest state of the system (taken into account here) and does not move to the acceptor. The dependence of the transfer rate on the relative energy of the acceptor $`E_{31}=E_{\mathrm{D}^+\mathrm{BA}^{}}E_{\mathrm{D}^{}\mathrm{BA}}`$ in Fig. 4.8 remains constant while the acceptor state energy lies below the bridge state energy. Increase of $`E_{31}`$ up to $`E_{21}=1.36\mathrm{eV}`$ gives the maximal value of rate $`k_{\mathrm{ET}}\mathrm{\Gamma }_{21}`$. While the value of $`E_{31}`$ increases further the acceptor level becomes the highest in the system and therefore the population cannot remain on it. For the application of the results to various solvents and comparison with experiment one should use the scaling for energy, coherent coupling, and dissipation as discussed above. The combinations of the energy scaling mentioned in subsection 4.2.1 and relaxation intensities scalings mentioned in subsection 4.2.4 are represented in Fig. 4.9. An increase in the static dielectric constant $`ϵ_\mathrm{s}`$ from $`2`$ to $`4`$ leads to an increase of the transfer rate, no matter which scaling is used. Further increase of $`ϵ_\mathrm{s}`$ induces saturation for the Onsager-Mataga scaling and even a small decrease of the transfer rate for the Born-Marcus scaling. In all those cases the solvent is approximated as the continuous medium. Thus, the transfer rate depends on the interplay of the two mechanisms. Within the used approximations an increase in the solvent polarizability and, hence, of its dielectric constant lowers the energy of the bridging and acceptor states and increases the system-bath interaction and, hence, the relaxation coefficients. It induces a smooth rise of the transfer rate in the whole interval of $`ϵ_\mathrm{s}`$ for the Onsager-Mataga scaling. On the other hand the large values of dielectric constant lead to essentially different polarisational states of the environment for the aggregate states with different dipole moment. The difference of the environmental polarizations reduces the values of the coherent couplings, see Eq. (4.25). This effect is reflected for the Born-Marcus scaling in the small decrease of $`k_{\mathrm{ET}}`$ for large values of $`ϵ_\mathrm{s}`$. The values for the transfer rate, obtained with this scaling come closer to the experimental value $`k_{\mathrm{ET}}(ϵ_\mathrm{s}^{\mathrm{CH}_2\mathrm{Cl}_2})`$. This gives a hint that the model of individual cavities for each molecular block is closer to the reality than the model with a single cavity for the whole supermolecule. Below we consider Born scaling Eq. (4.8) for the system energies and Marcus scaling Eq. (4.24) for the dissipation parameter to compare the calculated transfer rates with the measured ones. For the solvents CYCLO, MTHF, and $`\mathrm{CH}_2\mathrm{Cl}_2`$ one obtains the following relative energies of the bridging level $`E_{21}=1.77\mathrm{eV}`$, $`1.36\mathrm{eV}`$, and $`1.30\mathrm{eV}`$, respectively, i.e., a decrease in the energy of the bridging state. The calculated transfer rate coincides with the experimental value for $`\mathrm{H}_2\mathrm{P}`$ $`\mathrm{ZnP}`$ $`\mathrm{Q}`$ in CYCLO, see table 4.2. For $`\mathrm{CH}_2\mathrm{Cl}_2`$ the numerical transfer rate diverges from the experimentally determined one. The calculated numerical value is found to be approximately thirty percent faster. The following reasons could be responsible for this difference: (i) absence of vibrionic substructure of the electronic states in the present model; (ii) incorrect dependence of system state energies on the solvent properties; (iii) appearing of additional transfer channels not mentioned in the scheme shown in Fig. 4.2. Each of these possibilities requires some comments. ad (i): The incorporation of the vibrational substructure will result in a complication of the model with parameters such as the mass of each vibrational mode . It will also give a more complicated transfer rate dependence on the energy of the electronic states and dielectric constant. In the model with vibrational substructure the interplay between the difference of the energies of each pair of electronic states and corresponding reorganization energy determines if the pair belongs to the normal, activationless, or inverted region according to Marcus . In contrast to the present consideration the model with vibrational substructure should yield the maximal transfer rate for nonequal energies of electronic state, namely for the activationless case: the energy difference equals to the reorganization energy. For a detailed comparison of TB model and the model with vibrational substructure see the next section. ad (ii): In particular such solvent effect as solvation shell, see for example Ref. , do need a molecular dynamics simulation. The total influence of solvent is, probably, reflected in solvent-induced energy shift between the spectroscopically observable states without charge separation $`E_{\mathrm{D}^{}\mathrm{BA}}`$ and $`E_{\mathrm{DB}^{}\mathrm{A}}`$ which is neglected in our consideration. ad (iii): Some states of our three-site system have not been included in the model schematically presented in Fig. 4.2. Using the solvent with strong dielectric constant can bring high-lying states closer to the ones involved in the transfer. The state which might play a role is $`|\mathrm{D}^{}\mathrm{B}^+\mathrm{A}`$ because of its larger dipole moment. So it is strongly influenced by the solvent. The state $`|\mathrm{DBA}^{}`$ which has such a high energy that one can excite it only with ultraviolet radiation, most probably, does not play a role in the discussed transfer. #### 4.5.3 Comparison with the steady-state solution In the work of Davis et al. the vibrational substructure of the electronic states is also not taken into account, the relaxation is incorporated phenomenologically. We use a similar approach in the present consideration derived within a Redfield-like theory. Also we consider the relevant processes of dephasing and depopulation between each pair of levels. In contrast Davis et al. apply relaxation only to selected levels. In their paper dephasing $`\gamma `$ occurs between excited levels, see Eq. (7) in Ref. , while depopulation $`k`$ takes place only for the transition from acceptor to ground state. The advantage of the approach of Davis et al. is the possibility to investigate the transfer rate dependence for more than one bridge state. This was not the goal of the present work but it can be extended into this direction. We are interested in the ET in a concrete molecular complex with realistic parameter values and realistic possibilities to modify those parameter. Our results as well as the results of Davis show that ET can occur as coherent (with the superexchange mechanism) or dissipative process (with the sequential mechanism). We have considered various ways of switching between these two mechanisms including the one suggested by Davis et al., i. e., the change of the relaxation intensities meaning change of the solvent. The numerical steady state method used by Davis et al. is an attractive one due to its simplicity, but unlike our method it is not able to give information about the time evolution of the system. On the other hand the method of Davis et al. does not allow to calculate the most widespread stationary characteristics that clarify the presence of ET process, namely the fluorescence of the donor and its quenching. We suppose this can be included into the approach by Davis et al. by introducing the depopulation coefficient describing the transfer from donor to ground state. ### 4.6 Vibronic model In the model with vibrational substructure (VSS) the influence of the interaction with an environment onto ET and other dynamical processes is also described by the common Hamiltonial of the form Eq. (2.1). The bath again is modeled by a distribution of HOs and characterized by its spectral density $`J(\omega )`$. Starting with a DM of the full system, the RDM of the relevant (sub)system is obtained by tracing out the bath degrees of freedom . While doing so a second-order perturbation expansion in the system-bath coupling and the Markov approximation have been applied . So one arrives at the RDMEM of the already discussed structure. The bridge ET system $`\mathrm{H}_2\mathrm{P}\mathrm{ZnP}\mathrm{Q}`$ is modeled by three diabatic electronic potentials, corresponding to the states $`|1=|\mathrm{D}^{}\mathrm{BA}`$, $`|2=|\mathrm{D}^+\mathrm{B}^{}\mathrm{A}`$, and $`|3=|\mathrm{D}^+\mathrm{BA}^{}`$ (see Fig. 4.10). Each of these electronic potentials has a vibrational substructure. The vibrational frequency is assumed to be 1500 cm<sup>-1</sup> as a typical frequency within carbon structures. The potentials are displaced along a common reaction coordinate which represents the solvent polarization . Following the reasoning of Marcus the free energy differences $`\mathrm{\Delta }G_{\mu \nu }`$ corresponding to the ET from molecular block $`\nu `$ to $`\mu `$ ($`\nu =1`$, $`\mu =2,3`$) are estimated to be $$\mathrm{\Delta }G_{mn}=E_\mu ^{\mathrm{ox}}E_\nu ^{\mathrm{red}}E^{\mathrm{ex}}\frac{e^2}{4\pi ϵ_0ϵ_\mathrm{s}}\frac{1}{r_{\mu \nu }}+\mathrm{\Delta }G_{\mu \nu }(ϵ_\mathrm{s})$$ (4.37) with the term $`\mathrm{\Delta }G_{\mu \nu }(ϵ_\mathrm{s})`$ correcting for the fact that the redox energies $`E_\mu ^{\mathrm{ox}}`$ and $`E_\nu ^{\mathrm{red}}`$ are measured in the reference solvent with dielectric constant $`ϵ_\mathrm{s}^{\mathrm{ref}}`$: $$\mathrm{\Delta }G_{mn}(ϵ_\mathrm{s})=\frac{e^2}{4\pi ϵ_0}\left(\frac{1}{2r_\mu }+\frac{1}{2r_\nu }\right)\left(\frac{1}{ϵ_\mathrm{s}}\frac{1}{ϵ_\mathrm{s}^{\mathrm{ref}}}\right).$$ (4.38) The excitation energy of the donor $`\mathrm{H}_2\mathrm{P}\mathrm{H}_2\mathrm{P}^{}`$ is denoted by $`E^{\mathrm{ex}}`$. $`r_\nu `$ denotes the radius of either donor (1), bridge (2), or acceptor (3) and $`r_{\mu \nu }`$ the distance between two of them as presented in section 4.3. Also sketched in Fig. 4.10 are the reorganization energies $`\lambda _{\mu \nu }=\lambda _{\mu \nu }^\mathrm{i}+\lambda _{\mu \nu }^\mathrm{s}`$. These consist of the internal reorganization energy $`\lambda _{\mu \nu }^\mathrm{i}`$, which is estimated to be 0.3 eV , and the solvent reorganization energy $$\lambda _{\mu \nu }^\mathrm{s}=\frac{e^2}{4\pi ϵ_0}\left(\frac{1}{2r_\mu }+\frac{1}{2r_\nu }\frac{1}{r_{\mu \nu }}\right)\left(\frac{1}{ϵ_{\mathrm{}}}\frac{1}{ϵ_\mathrm{s}}\right).$$ (4.39) Further parameters are the electronic couplings between the potentials. They are the same as described in section 4.3. The damping is described by the spectral density $`J(\omega )`$ of the bath. This is only needed at the frequency of the vibrational transition and is determined $`J(\omega _{\mathrm{vib}})/\omega _{\mathrm{vib}}=0.372`$ by fitting the ET rate for the solvent methyltetrahydrofuran (MTHF). In the vibronic model the spectral density is taken as a constant with respect to $`ϵ_\mathrm{s}`$. Next the calculation of the dynamics is sketched. Starting from the Liouville equation, performing the abovementioned approximations the equation of motion for the RDM $`\rho _𝒩`$ can be obtained $$\frac{}{t}\rho _𝒩=\frac{i}{\mathrm{}}(E_{}E_𝒩)\rho _𝒩i\underset{𝒦}{}\{v_{𝒩𝒦}\rho _𝒦v_𝒦\rho _{𝒦𝒩}\}+R_𝒩.$$ (4.40) The index $``$ combines the electronic quantum number $`\mu `$ and the vibrational quantum number $`M`$ of the diabatic levels $`E_{}`$. $`v_𝒩=V_{\mu \nu }F_{\mathrm{FC}}(\mu ,M,\nu ,N)`$ comprises Franck-Condon factors $`F_{\mathrm{FC}}`$ and the electronic matrix elements $`V_{\mu \nu }`$. The third term $`R_𝒩`$ describes the interaction between the relevant system and the heat bath. In principle one can eliminate the coherent terms containing $`v_𝒦`$ diagonalising the Hamiltonian of the isolated system. In such approach one couples the environment transitions to the transitions between the eigenstates of the system. This is adiabatic approach, used in e.g. . It is rather expensive numerically. We do not apply it here, so we use the diabatic approach. There are still a discussion in the literature wether the diabatic method is able to provide precise results . Nevertheless one uses the diabatic approach rather often because it is less expensive numerically . Equation (4.40) is solved numerically with the initial condition that only the donor state is occupied in the beginning. The population of the acceptor state $$P_3(t)=\underset{M}{}\rho _{3M3M}(t)$$ (4.41) and the ET rate given by Eq. (4.28) are calculated by tracing out the vibrational modes. ### 4.7 Comparison of models with and without vibrations Here we compare the following models: (i) the model where only electronic states without vibrational substructure are taken into account (see Fig. 4.5) and relaxation processes take place between the electronic states and (ii) the model where the relaxation takes place between vibrational states within one electronic state potential surface introduced in the previous section. In this section the energies of the electronic states $`E_m`$ of model without vibrational substructure are chosen to be the ground states of the harmonic potentials of the vibronic model shown in Fig. 4.10. So they vary with the dielectric constant. The electronic couplings are chosen to scale as Eq. (4.25). Note, that in the model with vibrational substructure it corresponds to the scaling provided by the Franck-Condon overlap elements between the vibrational ground states of each pair of electronic surfaces $$v_𝒩=V_{\mu \nu }F_{\mathrm{FC}}(\mu ,0,\nu ,0)=V_{\mu \nu }\mathrm{exp}\frac{|\lambda _{\mu \nu }|}{2\mathrm{}\omega _{\mathrm{vib}}}.$$ (4.42) Here we have used the expression for the Franck-Condon factor for the parabolas with equal effective mass $`m`$, and equal curvature $`F_{\mathrm{FC}}(\mu ,0,\nu ,0)=\mathrm{exp}(\frac{m\omega _{\mathrm{vib}}}{2\mathrm{}}\mathrm{\Delta }q_{\mu \nu }^2)`$ and definition of reorganization energy $`\lambda _{\mu \nu }=m\omega _{\mathrm{vib}}^2\mathrm{\Delta }q_{\mu \nu }^2`$. $`q_{\mu \nu }`$ stands for the difference of the reaction coordinate equilibrium points of these parabolas. In the vibronic model not only the free energy differences $`\mathrm{\Delta }G`$ but also the reorganization energies $`\lambda `$ scale with the dielectric constant $`ϵ_\mathrm{s}`$. Due to this scaling of $`\lambda `$ the system-bath interaction is scaled with the dielectric constant $`ϵ_\mathrm{s}`$. In the high temperature limit the reorganization energy is given by $$\lambda =\mathrm{}_0^{\mathrm{}}𝑑\omega \frac{J(\omega )}{\omega }.$$ (4.43) This relation can be taken as motivation to scale the TB spectral density with $`ϵ_\mathrm{s}`$ like the reorganization energies $`\lambda `$ in the model with vibrational substructure as discussed in the subsection 4.2.4. In the present calculations $`\mathrm{\Gamma }_{21}=\mathrm{\Gamma }_{23}=\mathrm{\Gamma }`$ is assumed. The absolute value of the damping rate $`\mathrm{\Gamma }`$ between the electronic states (see Fig. 4.5) in this section is determined by fitting the ET rate for the solvent MTHF to be $`\mathrm{\Gamma }=2.9\times 10^{11}`$ s<sup>-1</sup>. It differs from the value given in section 4.5 because the different energies of states have been used here. The advantage of the TB model in comparison to the model with vibrational substructure is the possibility to determine the transfer rate $`k_{\mathrm{ET}}`$ and the final population of the acceptor state either numerically or analytically. For all situations described in this paper the differences between analytic and numerical results without the extra assumptions are negligible as shown in section 4.5. In Fig. 4.11 it is shown how the minima of the potential curves change with varying the solvent due to the changes in Eqs. (4.37) to (4.39). The solvents are listed in Table 4.3 together with their parameters and the results for the ET rates in both models. For larger $`ϵ_\mathrm{s}`$ the coordinates of the potential minima of bridge and acceptor increase while their energies decrease with respect to the energy of the donor. The energy difference between donor and bridge decreases with increasing $`ϵ_\mathrm{s}`$. This makes a charge transfer more probable. For small $`ϵ_\mathrm{s}`$ the acceptor state is higher in energy than the donor state; nevertheless there is a small ET rate due to coherent mixing. For fixed $`ϵ_{\mathrm{}}`$ the ET rate is plotted as a function of the dielectric constant $`ϵ_\mathrm{s}`$ in Fig. 4.12. The ET rate in the vibronic model increases strongly for small values of $`ϵ_\mathrm{s}`$ while the increase is very small for $`ϵ_\mathrm{s}`$ in the range between 5 and 8. The increase for small values of $`ϵ_\mathrm{s}`$ is due to the fact that with increasing $`ϵ_\mathrm{s}`$ the minimum of the acceptor potential moves from the position higher than the minimum of the donor level to the position lower than the donor level. So the transfer becomes energy favorable. This can also be seen when looking at the results for the TB model without scaling the electronic coupling with the Franck-Condon factor. In this case the ET rate increases almost linearly with increasing $`ϵ_\mathrm{s}`$. The effect missing in this model is the overlap between the vibrational states. If one corrects the electronic coupling in the TB model by the Franck-Condon factor of the vibrational ground states as described in Eq. (4.42), good agreement is observed between the vibronic and the TB model. The ET rate for the vibronic model shows some oscillations as a function of $`ϵ_\mathrm{s}`$. This is due to the small density of vibrational levels in this model with one reaction coordinate. All three electronic potential curves are harmonic and have the same frequency. So there are small maxima in the rate when two vibrational levels are in resonance and minima when they are far off resonance. Models with more reaction coordinates do not have this problem nor does the simple TB model. If these artificial oscillations would be absent, the agreement between the results for the TB and the vibronic model would be even better, because the rate for the vibronic model happens to have a maximum just at the reference point $`ϵ_\mathrm{s}=6.24`$ which we have chosen to fix the spectral density, i. e. for MTHF. The comparison of the two models has been made assuming that the scaling of energies as a function of the dielectric function is correct in the Marcus theory. There have been a lot of changes to Marcus theory proposed in the last years. Marcus theory assumes excess charges within cavities surrounded by a polarizable medium and there one only takes the leading order into account. Higher order terms are included in the so called reaction field theory (see for example ). But to compare different solvation models is out of the range of the present investigation. Some more details on this issue for the TB model are already given in section 4.2. Here we just want to note that the effect of scaling the system-bath interaction with $`ϵ_\mathrm{s}`$, as assumed in the present work for the TB model, has no big effect on the ET rates. ### 4.8 Summary We have performed a study of the ET in the supermolecular complex $`\mathrm{H}_2\mathrm{P}\mathrm{ZnP}\mathrm{Q}`$ within the TB model treated with the DM formalism. The determined analytical and numerical transfer rates are in a reasonable agreement with the experimental data. The superexchange mechanism of ET dominates over the sequential one. We have investigated the stability of the model varying one parameter at a time. The qualitative character of the transfer is stable with respect to a local change of system parameters. It is determined that the change of the dominating transfer mechanisms can be induced by lowering the bridge state energy. The physical reasons of system parameters scaling as well as the relation of the theory presented here to other theoretical approaches to ET have been discussed. The validity of the TB model and its applicability to $`\mathrm{H}_2\mathrm{P}\mathrm{ZnP}\mathrm{Q}`$ is confirmed because one gets good agreement for the ET rates of the models with and without vibrational substructure, i. e. the vibronic and the TB model, if one scales the electronic coupling with the Franck-Condon overlap matrix elements between the vibrational ground states. The advantage of the model with electronic relaxation only is the possibility to derive analytic expressions for the ET rate and the final population of the acceptor state. But of course for a more realistic description of the ET process in such complicated systems as discussed here, more than one reaction coordinate should be taken into account. The calculations performed in the framework of the present formalism can be extended in the following directions: (i) Considerations beyond the kinetic limit. The solvent dynamics has to be included into the model as well as, probably, non-Markovian RDM equations. (ii) Enlargement of the number of molecular blocks in the complex. (iii) Initial excitation of states with rather high energy should open additional transfer channels. ## Chapter 5 Mixture of Solvents In this chapter we apply the RDM method which was described in chapter 2 to another experimental system i. e. porphyrin triad complexes. This method provides a quantitative description of time-resolved and steady-state properties such as fluorescence quenching of the porphyrins. It will be shown that our calculations do agree with already performed experiments. We start this chapter with a brief review of an important role of the supramolecular porphyrin array in electronic energy transfer, charge transfer, and photoinduced structural changes in biological systems since such processes constitute the basis for the design of molecular devices of applicative interest and for the understanding of the most important photobiological processes. First, in section 5.1 we introduce the chemical structure of the self-assembled porphyrin triad $`\mathrm{ZnPD}\mathrm{H}_2\mathrm{P}`$. Next, in section 5.2 we qualitatively describe the photophysics of the porphyrin triad in different solvents with various dielectric constants. In section 5.3 the parameters of the RDM formalism are discussed and determined for both the porphyrin complexes and the quantitative description of the fluorescence quenching. The explanation of the fluorescence quenching is done in subsection 5.4.1. The analysis of our calculations and the experimental results performed in subsection 5.4.2 yield valuable information about the reaction mechanisms. Finally, in section 5.5 we summarize the achievements of this chapter. Some generalizations are transferred to the Appendix B and require more detailed consideration in the future. ### 5.1 Self-assembled triad of porphyrins Because of their widespread occurrence in photosynthetic reaction centres and other electron-transfer systems supramolecular porphyrin arrays have played a leading role in the study of energy and charge transfer processes in biological systems or to gain insight into the principal possibilities of molecular electronics. Among supramolecular porphyrin arrays three-component covalently-linked donor-acceptor systems have attracted a lot of interest, especially, in the context of photochemical molecular devices . Besides, it is possible to form donor-acceptor systems using both covalent bond and non-covalent binding self-assembling based on coordination interactions of metallo-porphyrins with appropriate extra ligands. These systems are formed from chemical dimer of Zn-octaethylporphyrin with a phynyl spacer (ZnPD) and dipyridyl-diphenylporphyrin with nitrogens in meta-positions of pyridil rings ($`\mathrm{H}_2\mathrm{P}`$) via two-fold coordination of two central Zn ions of the dimer with pyridil rings of the extra-ligand. The chemical structure of a self-assembled porphyrin triad is represented in Fig. 5.1. It was experimentally found in these systems that the population of the excited state of $`\mathrm{H}_2\mathrm{P}`$ is dependent on the solvent dielectric constant as well as on temperature. This is attributed to a close lying charge separated state which exchanges rapidly with the exited state of $`\mathrm{H}_2\mathrm{P}`$. ### 5.2 Photophysics of a porphyrin triad As determined by fluorescence spectroscopic measurements , where one excites $`\mathrm{ZnPD}`$, the triadic aggregate in non-polar solvents (toluene, methylcyclohexane) demonstrates fluorescence quenching for $`\mathrm{ZnPD}`$ and quite intensive fluorescence for $`\mathrm{H}_2\mathrm{P}`$. It is reasonable to assume that energy transfer processes may be involved in this quenching. The fluorescence $`\mathrm{ZnPD}\mathrm{H}_2\mathrm{P}^{}\mathrm{ZnPD}\mathrm{H}_2\mathrm{P}`$ (characterized by the fluorescence time $`\tau _\mathrm{F}=7.7`$ ns) is less intensive with respect to that for pure $`\mathrm{H}_2\mathrm{P}^{}\mathrm{H}_2\mathrm{P}`$ ($`\tau _\mathrm{F}=9.3`$ ns) at the same conditions . The quenching of $`\mathrm{H}_2\mathrm{P}`$ fluorescence increases with the polarity of the solvent. To check whether this is the case an experiment has been performed . In this experiment the triadic complexes were formed in a solution of pure toluene (low dielectric constant) and then in an admixture of acetone (up to 20%) added to toluene (relatively high dielectric constant). In the first case the aggregate shows fluorescence with the mean band maximum of $`716`$ nm which is attributed to $`\mathrm{H}_2\mathrm{P}`$ and fluorescence excitation spectra clearly demonstrate the presence of an energy transfer. While in the second case $`\mathrm{H}_2\mathrm{P}`$ shows noticable fluorescence quenching. Thus the excitation is lost without radiation and ET must take place. The physics of this reaction is qualitatively shown in Fig. 5.2, where D and A represent the electron donor $`\mathrm{ZnPD}`$ and the electron acceptor $`\mathrm{H}_2\mathrm{P}`$, respectively. Note that here D and A denote other compounds than in chapter 4 while the symbols of plus, minus, and star are defined in the same way as in section 4.2. Light excitation of $`\mathrm{ZnPD}`$ to its singlet state $`|1`$ is followed by weak dissipative energy transfer to the exited state $`|3`$ (process a). Inaddition to this process, an ET (process b) takes place, leading to the charge-separated state $`|2`$. The evolution of the population of this state $`|2`$ is distributed between charge recombination (does not enter into the scheme) and further weak coherent and dissipative charge transfer (process c) to give the state $`|3`$. Because the decay of the charge-separated state occurs quite slowly for porphyrin-type systems , we do not consider this process and do not include it in Fig. 5.2. The energy of state $`|2`$ depends on the dielectric constant of the solvent. ### 5.3 Density matrix model parameters The RDM model is able to describe coherent dynamics, dissipative dynamics, and thermally activated processes, thus the RDM formalism can be used for a quantitative explanation of the reaction which was described in subsection 5.2. We apply the system of equations (4.20)-(4.21) to describe the processes in the porphyrin triad. To calculate the population of the excited states physically reasonable values of the model parameters, such as coherent and dissipative couplings and energy of states have been chosen. The physical behavior of the system is determined in leading order by the energies. remains rather stable in respect to change of other parameters. Nevertheless it is reasonable to give some physical foundation in which region the parameters values should lie. In accordance with the consideration given in section 4.7 the parameters for the model with only electronic states can be extracted using some properties of the appropriate model with vibrational substructure. In such model one describes the porphyrin triad with the relevant potential energy surfaces in the space of a single reaction coordinate which reflect the extent of the solvent polarization induced by the field of the triad. In order to allow the analysis of absorption and emission spectra one should add to the model a ground state potential $`|0`$. For the neutral excited states $`|1`$ and $`|3`$ the difference in absorption and emission spectra of the transition ground-excited determines the shift $`q_{i0}`$ of the equilibrium point of the excited state $`|i`$ in respect to the equilibrium point of the ground state $`|0`$. The shift of the equilibrium point of the state $`|2`$ with charge separation can be calculated using the reorganization energy Eq. (4.39). We assume that the sign of the shift for the state $`|3`$ should be negative in respect to the signs of the shifts of $`|1`$ and $`|2`$. To check whether it is really so one should analyse the dependence of $`|1|3`$ transition rate on the solvent polarity. Supposing that the vibrational excitations again do not play a role and in accordance with the explanation after the Eq. (4.42) the coherent couplings can be expressed as $`v_{12}=V_{12}\mathrm{exp}\left[(2\mathrm{}\omega _{\mathrm{vib}})^1(q_{20}q_{10})^2\right]`$, $`v_{23}=V_{23}\mathrm{exp}\left[(2\mathrm{}\omega _{\mathrm{vib}})^1(q_{20}q_{30})^2\right]`$ etc.. Therefore the relation of the coherent couplings reads $$v_{12}/v_{23}=V_{12}/V_{23}\mathrm{exp}\left\{(2\mathrm{}\omega _{\mathrm{vib}})^1[q_{10}^2q_{30}^22q_{20}(q_{10}q_{30})]\right\}.$$ For $`q_{10}q_{30}`$ and $`q_{20}q_{30}`$ the exponential reads $`\mathrm{exp}[(2\mathrm{}\omega _{\mathrm{vib}})^1`$ $`4q_{20}q_{10}]`$. This is an exponential of a positive argument so the $`v_{12}`$ should be at least $`e`$ times larger as $`v_{23}`$. The electronic matrix elements $`V_{\mu \nu }`$ should also have a difference because the $`|3|2`$ process imply the ET from one inner orbital $`\mathrm{HOMO}_\mathrm{D}`$ to another inner orbital $`\mathrm{HOMO}_\mathrm{A}`$. So for the first attempt we take $`v_{12}=0.06\mathrm{eV}`$ as typical value of the coupling between porphyrins (same as the coupling $`\mathrm{H}_2\mathrm{P}^{}\mathrm{ZnP}\mathrm{Q}|H|\mathrm{F}_2\mathrm{P}^+\mathrm{ZnP}^{}\mathrm{Q}`$ for zinc-porphyrin-quinone complex in chapter 4 with the precision up to the first significant figure) and $`v_{32}=0.003\mathrm{eV}`$ is taken to be 20 times smaller. The range of reasonable values of relaxation constants $`\mathrm{\Gamma }_{\mu \nu }`$ can be estimated using Eq. (4.43) and reorganisanization energies from the model with vibrations. The values of the couplings are represented in Tab. 5.1. The precise determination of the parameters needs a special investigation within quantum chemical calculations. This is the next step in the theoretical description of this porphyrin triad. The quantum chemical calculation of the couplings should be done in the future. The excited state energies $`E_{\mathrm{D}^{}\mathrm{A}}=2.10`$ eV and $`E_{\mathrm{DA}^{}}=1.91`$ eV are found spectroscopically . The energy of the state $`|2`$ needs more attention. It depends on the solvent and can be calculated in any solvent with the help of Weller’s formula : $`E_{\mathrm{D}^+\mathrm{A}^{}}(ϵ)=E_{\mathrm{D}^+\mathrm{A}^{}}(ϵ_\mathrm{t})+\left({\displaystyle \frac{1}{ϵ}}{\displaystyle \frac{1}{ϵ_\mathrm{t}}}\right)\times {\displaystyle \frac{e^2}{4\pi ϵ_0}}\left({\displaystyle \frac{1}{2r_\mathrm{D}}}+{\displaystyle \frac{1}{2r_\mathrm{A}}}{\displaystyle \frac{1}{r_{\mathrm{DA}}}}\right),`$ (5.1) where $`ϵ`$ denotes the solvent static dielectric constant, $`r_\mathrm{D}=r_\mathrm{A}=5.5\mathrm{\AA }`$ are the donor and acceptor radius, respectively, $`r_{\mathrm{DA}}=8.8\mathrm{\AA }`$ is the distance between them. In our case the solvent consists of the main compound toluene with a dielectric constant $`ϵ_\mathrm{t}`$ and a small concentration $`c`$ of the additional compound acetone with a dielectric constant $`ϵ_\mathrm{a}`$. In accordance with the effective dielectric constant of the mixture reads $$ϵ=ϵ_\mathrm{t}+c\frac{3(ϵ_\mathrm{a}ϵ_\mathrm{t})ϵ_\mathrm{t}}{ϵ_\mathrm{a}+2ϵ_\mathrm{t}}.$$ (5.2) The change of $`ϵ`$ induces an energy shift of the state $`|2`$. We present the energy difference $`\mathrm{\Delta }E=E_3E_2`$ between states $`|3`$ and $`|2`$ as a function of $`c`$ in Fig. 5.3. The values $`E_{\mathrm{D}^+\mathrm{A}^{}}=1.90`$ eV, $`ϵ_\mathrm{t}=2.38`$ for the pure toluene solution and $`ϵ_\mathrm{a}=10`$ for the acetone were defined in . ### 5.4 Physical processes #### 5.4.1 Fluorescence quenching: simulations and experiment In the porphyrin triad the competition between charge transfer and energy transfer (process a and b in Fig. 5.2, respectively) cause a rather complex dynamics. Mathematically it is easy to calculate the system state at the infinite time $`t=\mathrm{}`$. For our real system the electron transfer time $`\tau _{\mathrm{ET}}`$ is much shorter than the time of fluorescence $`\tau _\mathrm{F}`$. Thus we approximate $`t=\mathrm{}`$ with some time moment when the ET has finished and the fluorescence has not occured yet. On this time $`\tau _{\mathrm{ET}}<t<\tau _\mathrm{F}`$ the system reaches the quasi thermal equilibrium between the excited state $`|3`$ and the close lying charge separated state $`|2`$. We have calculated the equilibrium population of the state $`|3`$ numerically with the RDM-method. This population corresponds to the presence of the fluorescence into the ground state. We denote the population of this state with $`P_{\mathrm{DA}^{}}=\rho _{33}`$. It has been found that $`\rho _{33}(\mathrm{})`$ decreases in two cases: (i) lowering of the energy $`E_{\mathrm{D}^+\mathrm{A}^{}}`$ induced by increase of aceton concentration, see Fig. 5.4 (left) and (ii) lowering of the temperature, see Fig. 5.4 (right) This decrease corresponds to the experimentally observed fluorescence quenching . #### 5.4.2 Reaction mechanisms The variation of the acetone admixture concentration and temperature changes the fluorescence intensity $`\rho _{33}`$ as well as a character of the excited states dynamics. Our simulations display that the time dependence of the population $`\rho _{33}`$ (Fig. 5.5) as well as its temperature dependence (Fig. 5.6) at low acetone concentration qualitatively differs from the behavior at high acetone concentration. This is the most important result of the RDM-calculations. It demonstrates the qualitatively different reaction mechanisms under various experimental conditions. A low concentration of acetone induces a low energy detuning between the states $`|2`$ and $`|3`$. In this case (see Fig. 5.6 for small $`c_\mathrm{a}`$ ), $`\rho _{33}`$ starts to increase with time due to energy transfer (process a) and then does not change as $`|3`$ reaches the quasi-thermal equilibrium. Thus the equilibrium population $`\rho _{33}(\mathrm{})`$ is reached in one-step and a reaction rate $`k`$ can be found with an one-exponential fit Eq. (4.28). As shown in Fig. 5.3 in the case of high aceton concentration the energy detuning between states $`|2`$ and $`|3`$ becomes larger and in addition to the energy transfer (process a), the hole transfer (process c) takes place, thus the equilibrium population $`\rho _{33}`$ is reached in two steps. At first, the energy transfer creates a time-dependent maximum of the population of $`|3`$ (see Fig. 5.6 for large $`c_\mathrm{a}`$) and then hole transfer $`|3`$ $``$ $`|2`$ slowly induces depopulation of $`|3`$ down to the equilibrium population. The reaction occurs with the help of a sequential transfer, which is described by two rates (increase and decrease) and the one-exponential fit for $`k`$ cannot be used in this case. The temperature dependence of $`\rho _{33}(\mathrm{})`$ (Fig. 5.7) also reflects the change of the reaction mechanism which depends on the acetone concentration and temperature. For a low acetone concentration the equilibrium population $`\rho _{33}(\mathrm{})`$ increases with temperature because $`\mathrm{\Delta }E`$ between states $`|3`$ and $`|2`$ is lower than the thermal energy. This case corresponds to a one-step reaction. In the case of a high acetone concentration $`\mathrm{\Delta }E`$ becomes larger. Thus we have two regimes $`k_\mathrm{B}T>\mathrm{\Delta }E`$ and $`k_\mathrm{B}T<\mathrm{\Delta }E`$. The first case correspond to the one-step reaction, the second to the two-step reaction. For high acetone concentration the increase of temperature induces the crossover from the second to the first type of behavior: While the temperature is quite low, its increase does not change $`\rho _{33}(\mathrm{})`$. It occurs because in the abscence of the thermally induced transitions $`\rho _{33}(\mathrm{})`$ is determined in leading order by the coherent mixing $`v_{23}`$ and does not depend on the temperature. Then, when the temperature is quite high in comparison to $`\mathrm{\Delta }E`$, so that $`k_\mathrm{B}T\mathrm{\Delta }E`$, the reaction crossover into the one-step regime is detected and $`\rho _{33}`$ starts to grow with the temperature. For a high acetone concentration (20%) there is no essential temperature dependence of $`\rho _{33}`$ in the considered interval of temperatures. ### 5.5 Summary The RDM method explains the physical properties of a real system, e. g., a self-assembled porphyrin triadic aggregate both qualitatively and quantitatively and gives good agreement with already performed experiments. However, in order to describe the experiments completely it is necessary to take into account a pump-probe laser field. This generalization is quite complicate and requires more detailed consideration in the future. We represent some mathematics involved in this generalization in Appendix B. Our calculations predict a solvent variation induced crossover between two regimes of the time evolution that most probably follows from the energy level dependence. There is no experimental data about this phenomena, that is why it is interesting to prove the existence of the described crossover in experiments. ## ## Chapter 6 Conclusions It is shown that the vibrational wave packet relaxation of initially coherent (displaced) states as well as the quantum superposition of coherent states in heat baths with different spectral densities exhibit a number of peculiarities compared with the cases of linear and quadratic system-bath interactions. A strong dependence of the relaxation rate on the position of the spectral density maximum of the bath is found. The difference discriminates the mechanisms of the molecule-environment interaction. Based on the RDM method, we calculate the dynamics of ET for systems consisting of donor, bridge and acceptor in different solvents. In our first approach it is assumed that vibrational relaxation is much faster than the ET. Transfer rates and final populations of the acceptor state are calculated numerically and in an approximate fashion analytically. In wide parameter ranges these solutions are in very good agreement. The theory is applied to the ET in $`\mathrm{H}_2\mathrm{P}\mathrm{ZnP}\mathrm{Q}`$ with free-base porphyrin ($`\mathrm{H}_2\mathrm{P}`$) being the donor, zinc porphyrin ($`\mathrm{ZnP}`$) the bridge, and quinone ($`\mathrm{Q}`$) the acceptor. It is shown that the transfer rate can be controlled efficiently by changing the energy of the bridge level that can be done by changing the solvent. The effect of the solvent is determined for the models of single and multiple cavities. This approach has been compared to the second approach, where a vibrational substructure is taken into account for each electronic state and the corresponding states are displaced along a common reaction coordinate. In both approaches the system is coupled to the bath of HOs but the way of relaxation is quite different. For the comparison of these two models the parameters are chosen as similar as possible for both approaches and the quality of the agreement of the approaches is discussed. Applying the RDM theory to the photoinduced processes in $`\mathrm{ZnPD}\mathrm{H}_2\mathrm{P}`$ it has been found that the population of the state $`\mathrm{ZnPD}\mathrm{H}_2\mathrm{P}^{}`$ which controls the intensity of fluorescence of this complex is strongly influenced by the temperature and dielectric constant (polarity) of the solvent. The change of the last two parameters alters the character of the dynamics of the state $`\mathrm{ZnPD}\mathrm{H}_2\mathrm{P}^{}`$. ## ## Chapter 7 Outlook The work is devoted to the investigation of the influence of a heat bath on the physical processes in a quantum system. We use the density matrix theory as one of the most powerfool tool for investigation of quantum relaxation. In the beginning of the work we mention and recall the most important steps of derivation of the equation of motion for the RDM (master equation) for an arbitrary quantum system in diabatic representation interacting with the environment modeled by a set of independent HOs. At first we apply the theory to a single state in the diabatic representation decoupled from other states but having vibrational substructure presented by a single HO. We have performed a thorough investigation of this model with the help of the master equation, which has been solved analytically and numerically. For this system the wave packet dynamics in coordinate representation has been analysed for two models of the bath and two initial states. The different models of the bath have their maxima of the spectral density near the system frequency and near the double of the system frequency. The considered initial states are a coherent state and a superposition of coherent states. It has been shown that the wave packet dynamics demonstrates either ”classical squeezing” and the decrease of the effective vibrational oscillator frequency due to the phase-dependent interaction with the bath, or a time-dependent relaxation rate, distinct for even and odd states, and partial conservation of quantum superposition due to the quadratic interaction with the bath. The decoherence also shows differences compared to the usual damping processes. There are two universal stages of relaxation which allows analytical solution: the coherence stage and the Markovian stage of relaxation. The density matrix theory for a quantum system in a diabatic representation has been applied twice to a study of the ET in the supermolecular complex $`\mathrm{H}_2\mathrm{P}\mathrm{ZnP}\mathrm{Q}`$, namely with and without account for vibrations in the complex. With help of the model without vibrations we have determined analytical and numerical ET rates which are in a reasonable agreement with the experimental data. The superexchange mechanism of ET dominates over the sequential one. We have investigated the stability of the model varying one parameter at a time. The qualitative character of the transfer is stable with respect to a local change of system parameters. It is determined that the change of the dominating transfer mechanisms can be induced by lowering the bridge state energy. The physical reasons of system parameters scaling as well as the relation of the theory presented here to other theoretical approaches to ET which do not accounts for the vibrations have been discussed. The validity of the model without vibrations and its applicability to $`\mathrm{H}_2\mathrm{P}\mathrm{ZnP}\mathrm{Q}`$ is confirmed because one gets good agreement for the ET rates of the models with and without vibrational substructure, if one scales the electronic coupling with the Franck-Condon overlap matrix elements between the vibrational ground states. The advantage of the model with electronic relaxation only is the possibility to derive analytic expressions for the ET rate and the final population of the acceptor state. We have also applied the RDM theory to explain the physical properties of a self-assembled non-fluorinated triadic porphyrin aggregate. We have simulated the processes in this complex placed in a mixture of solvents with different dielectric constants. Our simulations reproduce the intensity of the fluorescence of the aggregate and its dependence on temperature and mutual concentration of constituents in the mixture both qualitatively and quantitatively with reasonable agreement with already performed experiments. Our calculations also predict a solvent variation induced crossover between two regimes of time evolution. There is no experimental data about this phenomena, that is why it is interesting to proof the existence of the described crossover in experiments. The calculations performed in the framework of the present formalism can be extended in the following directions: 1. Considerations beyond the kinetic limit. The solvent dynamics has to be included into the model as well as, probably, non-Markovian RDM equations. 2. Enlargement of the number of molecular blocks in the complex. 3. Initial excitation of states with rather high energy should open additional transfer channels. 4. For a more realistic description of the ET process in such complicated systems as discussed here, more than one reaction coordinate should be taken into account. 5. In order to describe the experiments completely it is necessary to take into account a pump-probe laser field. This generalization is quite complicated and requires a more detailed consideration in the future. In this work we have represented some mathematics involved in this generalization. 6. Quantum chemical calculation of coupling matrix elements. The future development of the extension (1) lies in the correct treatment of non-Markovian effects in the description of the photoinduced charge- and exciton transfer of a single molecule in a solvent by the density matrix theory. The model should remain the same as in this work although the solvent correlation functions should enter into the theory in explicite form without approximating them as a delta functions. For the calculation of the correlation functions of the bath of the polar molecules one should take the model of the dipole-dipole interacting spins, similar to the well known Heisenberg model. The response function of this spin lattice should be calculated for the time dependent charge separation in a single molecule embedded in this lattice. The proposed extension of this work would have the following advantages in respect with the traditional treatment of the bath: (a) It accounts for non-Markovian effects. (b) The state of the environment depends on the state of the system. (c) The solvent dipole value is the realistic parameter which enters into this model. As a possible outcome of the investigation of the bath of polar molecules modelled by the lattice of the rotators which interact as dipole-dipole one could expect some phase transitions with temperature as it often occurrs for interacting spins. As a general conclusion we could mention that coupling of a quantum system to the heat bath leads to the loss of energy, to disappearence of phase information, and ensures the irreversibility of the processes in the system. The influence of the environment should be in most cases included in the description of a real physical system. We believe that the developed methods and obtained results can be applied for other initial states and different couplings with the environment in real existing quantum systems. ## ## Appendix A Comparison with the Haken-Strobl-Reineker formalism We compare the RDMEM (2.5) with an analogous equation within the HSR model . There is a term in the RDMEM of the HSR model, which is absent in our calculations. Here we show that this term is nothing but the difference between the full relaxation operator and the relaxation operator in RWA. We neglect this term corresponding to $`\overline{\gamma }_{\mu \nu }`$ both in the equation of motion and in the expression for the transfer rate $`k_{\mathrm{ET}}`$ due to the RWA. The symbol $`\overline{\gamma }_{\mu \nu }`$ is used in the HSR for the rate of changes in the system state induced by this term. First we mention that the Eq. (2.12) from is derived under assumption $`\overline{\gamma }_{\mu \nu }=0`$. The RWA relaxation term within our formalism Eq. (2.5) yields the same RDMEM as Eq. (2.12) of Ref. . This equation is solved in section 4.4 of this work without $`\overline{\gamma }_{\mu \nu }`$. On the other hand the transformation of non-RWA term into matrix form gives the expression associated with $`\overline{\gamma }_{\mu \nu }`$ in the stochastic Liouville equations (SLE) formalism as we show here. The relaxation operator obtained within RWA Eq. (2.5) is assumed to describe the major contribution to the system dynamics. The importance of the non-RWA counterpart $`\widehat{L}_{\mathrm{non}\mathrm{RWA}}\sigma ={\displaystyle \underset{\mu \nu }{}}`$ $`\{\mathrm{\Gamma }_{\mu \nu }[n(\omega _{\mu \nu })+1]([\widehat{V}_{\mu \nu }\widehat{\sigma },\widehat{V}_{\mu \nu }]+[\widehat{V}_{\mu \nu }^+,\widehat{\sigma }\widehat{V}_{\mu \nu }^+])`$ $`+`$ $`\mathrm{\Gamma }_{\mu \nu }n(\omega _{\mu \nu })([\widehat{V}_{\mu \nu }^+\widehat{\sigma },\widehat{V}_{\mu \nu }^+]+[\widehat{V}_{\mu \nu },\widehat{\sigma }\widehat{V}_{\mu \nu }])\}`$ (A.1) obtained using the interaction Hamiltonian Eq. (4.14) instead of Eq. (4.16) remains questionable. The matrix form of Eq. (A) reads $`(L_{\mathrm{non}\mathrm{RWA}}\sigma )_{\kappa \lambda }=`$ $`\left\{\mathrm{\Gamma }_{\lambda \kappa }\left[2n(\omega _{\lambda \kappa })+1\right]+\mathrm{\Gamma }_{\kappa \lambda }\left[2n(\omega _{\kappa \lambda })+1\right]\right\}\sigma _{\lambda \kappa }.`$ (A.2) Thus, the full relaxation dynamics for the system-bath coupling Eq. (4.14) is described as the sum of two terms: Eq. (4.18) and Eq. (A.2). To compare the present approach with the SLE formalism used by HSR we recall the application of both methods to the simplest system, namely the TLS. The SLE method provides the following structure of the incoherent term for the RDMEM (Eqs. (3.1a)-(3.1d) in ): $$(L_{\mathrm{SLE}}\sigma )=\left(\begin{array}{cccc}2\gamma _1& 0& 0& 2\gamma _1\\ 0& 2(\gamma _0+\gamma _1)& 2\overline{\gamma }_1& 0\\ 0& 2\overline{\gamma }_1& 2(\gamma _0+\gamma _1)& 0\\ 2\gamma _1& 0& 0& 2\gamma _1\end{array}\right)\left(\begin{array}{c}\sigma _{11}\\ \sigma _{12}\\ \sigma _{21}\\ \sigma _{22}\end{array}\right).$$ (A.3) Applying the present model to the TLS we obtain $$\left(L_{\mathrm{RWA}}\sigma \right)=\left(\begin{array}{cccc}2\mathrm{\Gamma }n\left(\omega _{21}\right)& 0& 0& 2\mathrm{\Gamma }\left[n\left(\omega _{21}\right)+1\right]\\ 0& \mathrm{\Gamma }\left[2n\left(\omega _{21}\right)+1\right]& 0& 0\\ 0& 0& \mathrm{\Gamma }\left[2n\left(\omega _{21}\right)+1\right]& 0\\ 2\mathrm{\Gamma }n\left(\omega _{21}\right)& 0& 0& 2\mathrm{\Gamma }\left[n\left(\omega _{21}\right)+1\right]\end{array}\right)\left(\begin{array}{c}\sigma _{11}\\ \sigma _{12}\\ \sigma _{21}\\ \sigma _{22}\end{array}\right),$$ (A.4) $$(L_{\mathrm{non}\mathrm{RWA}}\sigma )=\left(\begin{array}{cccc}0& 0& 0& 0\\ 0& 0& \mathrm{\Gamma }[2n(\omega _{21})+1]& 0\\ 0& \mathrm{\Gamma }[2n(\omega _{21})+1]& 0& 0\\ 0& 0& 0& 0\end{array}\right)\left(\begin{array}{c}\sigma _{11}\\ \sigma _{12}\\ \sigma _{21}\\ \sigma _{22}\end{array}\right).$$ (A.5) The relaxation coefficients $`\overline{\gamma }_1`$ and $`\overline{\gamma }_1`$ from Eq. (A.3) that are, evidently, the difference, mix the non-diagonal elements of DM $`\sigma _{21}`$ and $`\sigma _{12}`$. The derivation within the formalism of present paper without RWA $`L_{\mathrm{RWA}}+L_{\mathrm{non}\mathrm{RWA}}`$ ensures the same structure of relaxational dynamics as Eq. (A.3). In details: the non-RWA terms $`\mathrm{\Gamma }[2n(\omega _{21})+1]`$ from Eq. (A.5) ensures the same effect as coefficients $`\overline{\gamma }_1`$ and $`\overline{\gamma }_1`$ from Eq. (A.3). But such kind of terms is really absent in the RWA relaxation term Eq. (A.4) which we use for the material-oriented calculations. At the present stage it is possible to estimate that consideration of non-RWA terms makes a smooth change of the characteristics of the process, i.e., $`k_{\mathrm{ET}}`$ while the expressions lose their simplicity. In our opinion the desired comparison of the more (without RWA) or less (with RWA) precise description of the TB model goes beyond the goals of this work. Another difference of equations (A.3) and (A.4) is that the first of them (SLE method) describes the same dissipative transition probability from $`\sigma _{22}`$ to $`\sigma _{11}`$ and back. It leads finally to the equal population of both levels. In our approach it is possible only if the states $`|1`$ and $`|2`$ are isoenergetic. In any other case the transition from upper level to the lower one will be more intensive as an inverse transition to construct the Boltzmann distribution of populations at the infinite time. The third, and, perhaps, the main difference of SLE and our methods is that SLE method assumes the modulation of system frequency $$\widehat{H}_1=\underset{\mu \nu }{}h_{\mu \nu }(t)b_\mu ^+b_\nu ,$$ (A.6) (Eq. (2.10) in Ref. , Eq. (2.5) in Ref. , Eq. (2.2) in Ref. ), where $`b_\mu `$ denotes exciton annihilation operator, $`h_{\mu \nu }(t)`$ stochastic function $`h_{\mu \nu }(t)=0`$, while in our approach it is assumed that system performs exchanges of quanta with the quantized modes of the thermal bath. So we conclude that the used RDMEM within TB model coincides with well-known HSR equation for exciton motion under certain approximations: i) energy levels are isoenergetic for the first RDMEM and ii) RWA for the second one. The similarity of the equations appears although different models are used for the environment. The generalization of the SLE method appeals to the quantum bath model with SB coupling of the form $`\widehat{H}^{\mathrm{SE}}\widehat{V}^+\widehat{V}\left(\widehat{a}_\lambda ^++\widehat{a}_\lambda \right)`$, which modulates the system transition frequency. In Ref. the equations for exciton motion are using projection operator technique without RWA leading to the presence of $`\overline{\gamma }`$ in the generalized stochastic Liouville equation (GSLE). So, taking the different SB coupling we have rederived a RDMEM which coincides with the GSLE after applying RWA. Both GSLE and our RDMEM are able to describe finite temperatures and non-periodic systems. ## ## Appendix B Full Model: Vibrations, Optics, Memory Effects In the full model the molecule is irradiated by the electromagnetic field as sketched in Fig. B.1. The common Hamiltonian (2.1) is extended by the field term $`H^{\mathrm{SF}}`$ and is written as follows $`H=H^\mathrm{S}+H^\mathrm{E}+H^{\mathrm{SE}}+H^{\mathrm{SF}}`$. Here the diabatic system Hamiltonian $$H^\mathrm{S}=\mathrm{}\underset{I}{}H_{I,n}^\mathrm{S}|I,nI,n|+\underset{I,n,J,m}{}V_{I,n,J,m}|I,nJ,m|$$ is characterised by the electronic-vibrational matrix elements $`H_{I,n}^\mathrm{S}=\mathrm{\Omega }_I+\omega _I(n+1/2)`$ and coherent mixing matrix elements $`V_{I,n,J,m}=V_{IJ}F_{\mathrm{FC}}(I,n,J,m)(1\delta _{IJ})`$. Here $`\mathrm{\Omega }_I`$ and $`\omega _I`$ stand for electronic and vibration transition frequencies, $`V_{IJ}`$ for electronic coupling, $`F_{\mathrm{FC}}(I,n,J,m)=I,n|J,m`$ for Franck-Condon factors. The system couples to the bath of HOs $`b_\xi `$ as follows $`H^{\mathrm{SE}}=[\underset{I}{}K_I(c_I^++c_I)][\underset{\xi }{}k_\xi (b_\xi ^++b_\xi )]`$, where $`c_I=\underset{n}{}\sqrt{n}|I,n1I,n|`$ stands for annihilation of a vibronic quantum in the $`I`$th electronic state. The electro-magnetic field $`\stackrel{}{E}(t)=\stackrel{}{E}_0^+e^{iWt}+\stackrel{}{E}_0^{}e^{iWt}`$ is described by frequency $`W`$ and strength $`\stackrel{}{E}_0^+=\stackrel{}{E}_0^{}=\stackrel{}{E}_0`$. The system-field interaction is written as $`H^{\mathrm{SF}}=\underset{I,n,J,m}{}|\stackrel{}{E}_0|p_{I,m,J,n}|I,nJ,m|`$ with the dipole $`p_{I,n,J,m}=eI,n|\underset{I}{}\sqrt{\frac{m}{2\mathrm{}\omega _I}}(c_I^++c_I)|J,m`$. Taking the notation $`\stackrel{~}{W}_{I,n,J,m}=W\mathrm{sign}(H_{I,n}^\mathrm{S}H_{J,m}^\mathrm{S})`$ we obtain the relevant RDMEM: $`\dot{\rho }_{I,m,I,m}=`$ $``$ $`i{\displaystyle \underset{K}{}}{\displaystyle \underset{l}{}}(E_0p_{I,m,K,l}+V_{I,m,K,l})(\rho _{K,l,I,m}\rho _{I,m,K,l})+L_{I,m,I,m}`$ $`\dot{\rho }_{I,m,J,n}=`$ $`[i(H_{I,m}^\mathrm{S}H_{J,n}^\mathrm{S})+i\stackrel{~}{W}_{I,m,J,n}]\rho _{I,m,J,n}`$ $``$ $`i(E_0p_{I,m,J,n}+V_{I,m,J,n})(\rho _{J,n,J,n}\rho _{I,m,I,m})+L_{I,m,J,n}`$ $`L_{I,m,J,n}=`$ $``$ $`(m[\gamma _{II}+2\mathrm{}\gamma _{II}^N]+n[\gamma _{JJ}^{}+2\mathrm{}\gamma _{JJ}^N]+2\mathrm{}\gamma _{JJ}^N)\rho _{I,m,J,n}`$ $`+`$ $`(2\mathrm{}\gamma _{IJ}^N+2\mathrm{}\gamma _{IJ})\sqrt{(m+1)(n+1)}\rho _{I,m+1,J,n+1}`$ $`+`$ $`2\mathrm{}\gamma _{IJ}^N\sqrt{mn}\rho _{I,m1,J,n1}.`$ The relaxation functions $`\gamma _{IJ}`$ are derived from the bath correlation functions: $$\gamma _{IJ}^N=_0^t𝑑\tau \left[\underset{\xi }{}K_Ik_\xi K_Jk_\xi N(\omega _\xi )\mathrm{exp}(i\omega _\xi t+i\omega _\xi \tau +i\omega _Iti\omega _J\tau )\right],$$ in the same way as introduced in chapter 3. ## Bibliography Ich erkläre, dass ich die vorliegende Arbeit selbständig und nur unter Verwendung der angegebenen Literatur und Hilfsmittel angefertigt habe. Die Passagen der Arbeit, die in Wortlaut oder Sinn anderen Werken entnommen wurden, habe ich entsprechend gekennzeichnet. Ich erkläre, nicht bereits früher oder gleichzeitig bei anderen Hochschulen oder an der Universität Chemnitz ein Promotionsverfahren beantragt zu haben. Ich erkenne die Promotionsordnung der Technischen Universität Chemnitz-Zwickau vom 15.März 1995 an. ## Thesen zur Dissertation * Die analytische Lösung der Bewegungsgleichung für die reduzierte Dichtematrix wurde für einen harmonisches Oszillator und für die Kopplung an ein thermisches Bad für Zeiten kürzer als die Badkohärenzzeit und für lange Zeiten berechnet. * Die gefundene Wellenpaketdynamik in Koordinatendarstellung für ein Bad mit Spektraldichtemaximum bei der Systemfrequenz zeigt eine Oszillation des Koordinatenmittelwerts und eine oszillierende Verringerung der Breite des Wellenpakets und der Energie des anfänglich kohärenten Zustands, wobei die Kohärenz der Superposition von kohärente Anfangszuständen schnell abklingt. * Die Simulation für Modelle mit Badfrequenz, die doppelt so hoch wie die Systemfrequenz ist, zeigt eine zeitabhängige Rate des Energieverlustes, die sich für gerade und ungerade Oszillatorzustände unterscheidet, mit einer relativ langen Dekohärenzzeit eines Superpositionszustandes. * Für die Lösung einer Dichtematrixbewegungsgleichung in diabatischer Darstellung und unter Vernachlässigung der vibronischen Struktur für angeregte Zustände des $`\mathrm{H}_2\mathrm{P}`$ $`\mathrm{ZnP}`$ $`\mathrm{Q}`$ -Komplexes wurde gefunden, daß die Besetzung des ladungsgetrennten Zustandes $`\mathrm{H}_2\mathrm{P}^+\mathrm{ZnP}\mathrm{Q}^{}`$ exponentiell bis zum Gleichgewichtsverteilungswert wächst. * Der Superaustausch-Transfermechanismus ist verglichen mit dem sequentiellen Transfer im Komplex $`\mathrm{H}_2\mathrm{P}`$ $`\mathrm{ZnP}`$ $`\mathrm{Q}`$ vorherrschend. Dieser Schlußfolgerung bleibt auch bei lokaler Änderung eines Systemparameters gültig. * Die Verringerung der $`\mathrm{H}_2\mathrm{P}^+\mathrm{ZnP}^{}\mathrm{Q}`$ — Zustandsenergie ruft die Änderung des Transfermechanismus von Superaustausch zu sequentiellem Transfer hervor. * Bei der Behandlung der dielektrischen Umgebung in einem Modell, das entweder Donator, Akzeptor und Brücke gemeinsam im einem Hohlraum des Dielektrikums enthält oder jeden Baustein in einen einzelnen Hohlraum, zeigte sich, daß die Transferraten im zweiten Fall genauer beschrieben wurden. * Die Simulation des Elektronstranfers unter Berücksichtigung der Schwingungen ergibt Transferraten, die mit den analytischen Transferraten vom Modell ohne vibronische Unterstruktur gut übereinstimmen, woraus wir schließen, daß schwingungslose Modelle geignet sind, Elektronentransfer in $`\mathrm{H}_2\mathrm{P}\mathrm{ZnP}\mathrm{Q}`$ zu beschreiben. * Die Fluoreszenzstärke des selbst-aggregierten molekularen Komplex $`\mathrm{ZnPD}\mathrm{H}_2\mathrm{P}`$ in einer Mischung von Lösungsmitteln mit verschiedenen dielektrischen Konstanten wächst mit der Temperatur und sinkt mit steigender Azetonkonzentration. * Die Simulationen sagen einen von Lösungsmittel induzierten Übergang des Reaktionsmechanismus im Komplex $`\mathrm{ZnPD}\mathrm{H}_2\mathrm{P}`$ voraus. | Curriculum Vitæ | | | | --- | --- | --- | | Name | | Dmitri Sergeevich Kilin | | Date of Birth | | 26.07.1974 | | Place of Birth | | Minsk, Republic of Belarus | | Citizenship | | Republic of Belarus | | Education | 11.1996-12.1999 | Ph. D. student in the research group Theoretical Physics III of the Institute of Physics, Chemnitz University of Technology | | | 09.1991-06.1996 | student at Theoretical Physics Department of Belorussian State University | | | 11.1986-06.1991 | Misk Yanka Kupala School N19 with intensive study of mathematics | | | 09.1988-06.1989 | Laboratory Courses for Young Physicists at the Institute of Heat and Mass Transfer of the Belorussian Acad. of Science | | | 09.1987-05.1988 | Laboratory Courses for Young Chemists at the Institute of Physical and Organic Chemistry of the Belorussian Acad. of Science | | | 09.1980-10.1986 | Minsk Secondary School N20 | | Employments | 11.1996-10.1999 | Scientific Employee at Chemnitz University of Technology | | | 08.1996-10.1996 | Scientific Employee at Belorussian State University | | | 08.1996-08.1996 | Instructor of Summer School for Outstanding Young Physicists “Lujesno-96” | | | 06.1996-07.1996 | Visiting Student Researcher at Chemnitz University of Technology | | | 12.1994-06.1996 | Student Researcher at Laboratory of Physics and Computing Teaching of Belarussian State University | | | 08.1994-11.1994 | Teacher of Physics in Minsk Secondary School N123 | | Awards | 1994 | Soros Foundation Grant | | | 1991 | Winner of the National Physics Olympiad of the Republic of Belarus | ## Acknowledgements (Danksagung) * Zunächst möchte ich mich bei Prof. Dr. Michael Schreiber für die freundliche Aufnahme in seinen Arbeitskreis, die interessante Themenstellung und seine kontinuierliche Unterstützung bedanken. * Viele Probleme und Fragen zur theoretischen Seite dieser Arbeit konnten in Diskussionen mit Dr. Volkhard May, Dr. Ulrich Rempel, Prof. Dr. Alexei Sherman und Prof. Dr. Edward Zenkevich geklärt werden, bei denen ich mich für die gute Zusammenarbeit bedanken möchte. * Ein besonderer Dank gilt den Doktoren Ulrich Kleinekathöfer, Reinhard Scholz und Thomas Vojta, die meine Wegbegleiter sowie bei den verschiedensten Problemen hilfsbereit waren. * Weiterhin danke ich der gesamten Arbeitsgruppe für das angenehme Arbeitsklima. * Last but not least möchte ich mich bei meinen Eltern und Grosseltern für ihre kontinuierliche Unterstützung und bei meine Frau Sveta für das wieder einmal ausgezeichnete Lektorat dieser Arbeit bedanken.
warning/0001/astro-ph0001408.html
ar5iv
text
# Transformations between the theoretical and observational planes in the HST–NICMOS and WFPC2 photometric systems ## 1 Introduction Transformations between the theoretical and the observational planes are fundamental tools to compare stellar evolution models with observed color–magnitude diagrams. In order to calibrate suitable relations among colors, temperatures and bolometric corrections (BCs) several basic ingredients are needed: * homogeneous and complete grids of stellar spectra (observed and/or theoretical); * accurate filter profiles; * reference spectra (the Sun, Vega etc.) to set the zero points of the relations; * suitable routines to interpolate within the grids and to integrate along the spectra. These calibrations are certainly model dependent and systematic shifts between different scales are common features. The goal of this paper is to obtain color–temperature relations and BCs in the HST–NICMOS photometric system. We also derive analogous transformations for a few selected filters in the WFPC2 system to provide homogeneous, self–consistent color–temperature relations and BCs when combinations of optical–infrared colors are used. A more complete calibration of the color–temperature transformations in the WFPC2 system can be found in the paper by Holtzman et al. (1995, hereafter H95). In Sect. 2 we describe the code we used to derive the transformations. In Sect. 3 and 4 we discuss the transformations in the HST–NICMOS and the WFPC2 systems, respectively. In Sect. 5 we derive suitable relations to transform a few selected colors and BCs from the HST to the ground–based photometric system. In Sect. 6 we draw our conclusions. ## 2 The code The synthetic colors and BCs were computed using a modified version of the evolutionary synthesis code by Leitherer et al. (1999). We implemented two different set of model atmospheres: i) the compilation by Lejeune, Cuisinier & Buser (1997, hereafter LCB97), based on the ATLAS code by Kurucz and corrected to match empirical color–temperature calibrations; ii) the compilation by Bessell, Castelli & Plez (1998, hereafter BCP98) using only the homogeneous set of models without overshooting computed by Castelli (1997) and based on Kurucz’s ATLAS9 models. The grid of stellar parameters explored by our computations is: * Metallicities in the range 0.01 – 1.00 solar. * Effective temperatures $`T_{\mathrm{eff}}`$ in the range 3500 – 50000 K. * Gravities log$`g`$ in the range 0.0 – 5.0. The ground–based BVIJHK filter profiles are in the Johnson’s (1966) photometric system and were taken from Buser & Kurucz (1978) (BV filters) and from Bruzual (1983) (IJHK filters) (see also Sect. 5 in Leitherer & Heckman 1995 for more details). The selected broadband filters in the HST–WFPC2 and NICMOS photometric systems are: F439W, F555W, F814W and F110W, F160W, F222M, respectively. The filter profiles have been multiplied by the wavelength dependent detector quantum efficiency of the four WFPC2 and the three NICMOS cameras. We used the most updated post–launch throughput curves and CCD quantum efficiencies, according to the WFPC2 and NICMOS documentation on the WEB. The quoted NICMOS throughput curves are already empirically adjusted to match standard star measurements (the correction factors are less than 10% in the case of the F110W and F160W filters and about 25% for the F222M one). We do not apply any further empirical adjustment to these response curves. The colors have been normalized assuming 0.00 values in all passbands for a Vega–like star with $`T_{\mathrm{eff}}`$=9500K, log$`g`$=4.0 and \[Fe/H\]=–0.5 (cf. e.g. BCP98). The BCs have been normalized assuming a value of –0.07 for a Sun–like star with $`T_{\mathrm{eff}}`$=5750 K, log$`g`$=4.5 and \[Fe/H\]=0.0 (cf. e.g. Montegriffo et al. 1998). Different assumptions for the colors and bolometric corrections of Vega and Sun –like stars can be accounted for by simply scaling the inferred quantities by the corresponding amounts. In order to quantify the influence of the adopted filter response curves on the inferred colors we also compare our B–V and V–K values with those in Table 2 of BCP98 for solar metallicity. The adopted model atmospheres are exactly the same and also similar within 0.01 mag are the reference colors for the Vega–like star. The only difference between the two sets of transformations are the adopted filter response curves. Within 0.01 mag our V–K color is fully consistent with BCP98 values, while only below 5000K our B–V becomes progressively bluer (at most 0.03 mag at 3500K). For each model atmosphere at a given $`T_{\mathrm{eff}}`$ and log$`g`$ we tabulate the corresponding colors in the form (V-F), where F is the selected filter in the ground–based or in the HST system, and the bolometric correction BC<sub>V</sub> in the V passband. We also provided analogous colors and BCs using simple blackbodies at a given temperature. All the tables with the computed transformations can be retrieved from http://www.stsci.edu/science/starburst/. In the following we analyze the behavior of the inferred quantities varying the stellar parameters and the adopted model atmospheres. ## 3 The HST–NICMOS infrared plane In Fig. 1 we plot the (F110W–F160W) and (F110W–F222M) color–temperature transformations in the HST–NIC2 system, using the BCP98 models for three gravities (log$`g`$ of 0.0, 2.0 and 4.0) and two metallicities ($`Z=Z_{}`$ and $`Z=0.1Z_{}`$). For temperatures hotter than 4000 K these infrared colors show a scatter within 0.1 mag with varying log$`g`$ and $`Z`$. At lower temperatures the scatter among models with different gravities increases up to a few tenths of mag at $`T_{\mathrm{eff}}`$=3500 K, while the metallicity dependence is less critical. Pure blackbodies have systematically bluer colors than model atmospheres, especially at the coolest temperatures, as expected from the omission of molecular opacities. This means that for a given color, blackbodies are cooler than the model atmospheres. Analogous color–temperature relations using the LCB97 models have been obtained. In Fig. 2 we report the difference of the color–temperature transformations in the NIC2 system, by adopting the BCP98 or the LCB97 model atmospheres. At temperatures below $``$5000 K the infrared colors using the LCB97 models get progressively bluer (up to 0.2–0.3 mag) than those using the BCP98 models. This behavior mainly reflects the difference between the original and corrected grids of model atmospheres by LCB97. Comparing their Figs. 6,7 (the original synthetic colors for giants and dwarfs, respectively) with their Figs. 14,15 (the corresponding corrected models to match empirical color–temperature relations) one can see that in the range of temperatures between 4000 and 3500 K the corrected (J–H) ground based color is bluer (from 0.1 up to 0.3 mag) than the corresponding original value, regardless the stellar gravity. A similar comparison can be done between the original and the corrected (J–K) ground based color: the latter becomes bluer with decreasing temperature by about 0.2 and up to 0.3 mag for giants and dwarfs at T<sub>eff</sub>=3500 K, respectively. The un–corrected LCB97 models, from the original grids of atmosphere spectra by Kurucz, should provide colors more similar to those obtained from the BCP98 models. Color–temperature transformations in the NIC1 and NIC3 systems are also provided, as shown in Fig. 3. Regardess the model atmosphere used and the adopted stellar parameters, for a given temperature the (F110W–F160W) color in the NIC1 system is only $``$0.01 mag redder than in the NIC2 system, while both the (F110W–F160W) and the (F110W–F222M) colors in the NIC3 system are slightly redder ($``$0.03 mag below $``$4000 K). In Fig. 4 we plot the BC in the NIC2 F110W, F160W and F222M passbands as a function of the temperature using the BCP98 models. Increasing the temperature of the stellar atmospheres, progressively larger corrections (that is smaller values of the BC) to the infrared fluxes have to be applied in order to get the bolometric luminosity. The scatter in the inferred quantities for different gravities and metallicities is generally small. At low temperatures the use of pure blackbodies requires larger corrections than using model atmospheres, as expected since the former have bluer spectra than the latter. In Fig. 5 we plot the difference between the values of the BC in the NIC2 photometric system, adopting the BCP98 or the LCB97 models, as we did for the colors in Fig. 2. At temperatures below 5000 K, larger BC values in the F110W and F222M and smaller ones in the F160W passbands (up to $``$0.2 mag at $`T_{\mathrm{eff}}`$=3500 K) are required when LCB97 models are used compared to the BCP98 ones. The above trends are almost independent of the adopted metallicity. For a given set of model atmospheres, very similar BCs within $`<`$0.01 mag are also obtained in the NIC1 and NIC3 systems, as shown in Fig. 6. Only in the F110W passband we find that the BC values in the NIC3 system are slightly smaller ($``$0.02 mag) than in the NIC2 system. Our discussion on the behavior of the inferred colors and bolometric corrections with varying the stellar parameters has been limited to the low temperature domain (T$`{}_{\mathrm{eff}}{}^{}<`$5000 K) where they are more sensitive. At higher temperatures these infrared quantities become progressively less dependent on the adopted model atmospheres. ## 4 The HST–WFPC2 visual plane In Fig. 7 we plot the (F439W–F555W) and (F555W–F814W) color–temperature transformations in the HST–PC1 system, using the BCP98 models for the three gravities and two metallicities of Fig. 1. The scatter among models with different log$`g`$ and $`Z`$ is $``$0.1 mag at all temperatures in the case of the (F555W–F814W) color, while at low temperatures ($`T_{\mathrm{eff}}`$5000 K) the scatter among models with different gravities in the (F439W–F555W) color increases up to about 0.6 mag at $`T_{\mathrm{eff}}`$=3500 K, while the metallicity dependence is less critical. As we found for the infrared colors, pure blackbodies show bluer colors than model atmospheres. Analogous color–temperature relations using the LCB97 models are also obtained. In Fig. 8 we report the difference in the color–temperature transformations by adopting the BCP98 or the LCB97 model atmospheres. Using the LCB97 models, at low temperatures the (F439W–F555W) color becomes progressively bluer (up to 0.1–0.2 mag) at low gravities and only slightly redder at larger ones, compared to the values obtained from the BCP98 models. The (F555W–F814W) color becomes rapidly redder (particularly at low gravities) than the corresponding quantities using the BCP98 models for T$`{}_{\mathrm{eff}}{}^{}`$4000 K. This behavior has little metallicity dependence. As for the infrared colors, this behavior can be mainly ascribed to the corrections applied by LCB97 to the original model atmospheres. Comparing again their Figs. 6,7 (the original synthetic colors) with their Figs. 14,15 (the corresponding corrected models) one can see that below 4000 K the corrected ground–based (B–V) color of giants becomes bluer and slightly redder for dwarfs, while the ground–based (V–I) color becomes rapidly redder (a few tenths of mag, even larger values for giants). In Fig. 9 we plot the BC in the F439W, F555W and F814W passbands as a function of the temperature using the BCP98 models. The required BC values using pure blackbodies are on average slightly larger than using model atmospheres, as expected since the former have bluer spectra than the latter and, contrary to the infrared case, in the visual range they tend to be brighter than the model atmospheres. In Fig. 10 we plot the difference between the values of the BC, adopting the BCP98 or the LCB97 models, as we did in Fig. 8 for colors. At temperatures below 5000 K, smaller BC values in the F439W and F555W and larger (up to $``$0.2 mag at $`T_{\mathrm{eff}}`$=3500 K) in the F814W passbands are inferred if the LCB97 models are used. As for the infrared plane, the above trends are sensitive to the adopted gravity and almost independent of metallicity. For a given set of model atmospheres, very similar color–temperature transformations and BC (within 0.01 mag) are obtained in all four WFPC2 cameras. ## 5 Discussion Model atmospheres are a powerful tool to calibrate suitable color–temperature transformations since homogeneous and complete grids for a wide range of stellar parameters are available. Nevertheless, for some specific applications empirical scales even if they are less complete in terms of stellar parameters are preferred. Using the BCP98 model atmospheres, we calibrated average relations to transform several representative colors and BCs from the HST to the ground–based photometric system where most of the empirical scales are calibrated (cf. Montegriffo et al. 1998 and references therein). All the BCP98 models included in our grid of stellar parameters for all the three NICMOS and four WFPC2 cameras have been used to compute the best fits. Very similar best fit relations can be obtained using the LCB97 model atmospheres. In Fig. 11 the best model fits to the difference between the ground–based (J–H) and (J–K) colors and the NICMOS (F110W–F160W) and (F110W–F222M) ones, respectively, as a function of the NICMOS quantities are shown. The derived transformations are practically metallicity and gravity independent and very similar relations can be obtained adopting a particular metallicity or gravity. A cubic polynomial relation is required to transform the (F110W–F160W) into the ground–based (J–H) color, while a simple linear relation allows one to transform (F110W–F222M) into the ground–based (J–K) color. Very small ($`<`$0.01 mag) global r.m.s values have been obtained. The maximum scatter between the best fit and the model atmospheres occurs at the lowest temperatures ($`T_{\mathrm{eff}}`$=3500–4000 K) and is $``$0.06 mag in both colors. For comparison, we also plot the colors of the standard stars with measured F110W, F160W and F222M and ground–based JHK magnitudes, according to the NICMOS Photometry Update WEB page (November 25, 1998), even though the quoted values are still in the process of being updated, and of a set of red stars in the Baade’s window measured by Stephens et al. (1999). Unfortunately, most of these stars are cooler than 3500K, that is out of the temperature range covered by the selected set of model atmospheres. Nevertheless, the observed quantities are reasonably reproduced (within $``$0.1 mag) by the best model fits in the temperature range covered by the models, that is $`T_{\mathrm{eff}}`$3500 K, while at lower temperatures the scatter is larger and also depends on the adopted extrapolation. In Fig. 12 we show the best model fits to the difference between the ground–based BC<sub>J</sub> and BC<sub>K</sub> and the corresponding NICMOS BC<sub>F110W</sub> and BC<sub>F222M</sub> values, as a function of the (F110W–F160W) and (F110W–F222M) NICMOS colors, respectively. Cubic polynomial relations with even smaller scatters than for colors are required to transform the BC from the NICMOS into the ground–based infrared photometric system. The numerical relations to transform the selected colors and BCs from the NICMOS to the ground–based photometric system are listed below: $`(JH)=0.063(F110WF160W)^3+0.172(F110WF160W)^2+0.563(F110WF160W)+0.007`$ $`(JK)=0.803(F110WF222M)+0.003`$ $`BC_J=BC_{F110W}+0.069(F110WF160W)^30.181(F110WF160W)^2+0.443(F110WF160W)0.008`$ $`BC_K=BC_{F222M}+0.023(F110WF222M)^30.110(F110WF222M)^2+0.204(F110WF222M)0.001`$ In Fig. 13 the best model fits to the difference between the ground–based (B–V) and (V–I) colors and the corresponding WFPC2 (F439W–F555W) and (F555W–F814W) values as a function of the corresponding WFPC2 quantities are shown. Cubic polynomial relations are required to transform the (F439W–F555W) and the (F555W–F814W) colors into the corresponding ground–based (B–V) and (V–I) quantities. The global r.m.s values are still very small ($``$0.02 mag) as for infrared colors, while the maximum scatter (which occurs at the lowest temperatures and reflects a gravity dependence in the case of the (B–V) and a metallicity dependence in the case of the (V–I)) between the best fit and model atmospheres with selected metallicity or gravity is somewhat larger ($``$0.10–0.16 mag) than for the infrared colors. A direct comparison between our color transformations in the (B–V) planes and those proposed by H95 using their Table 7 indicates an excellent agreement, with only a minor, systematic difference of 0.01 mag (our (B–V)–(F39W–F555W) are bluer than the corresponding H95 values). In Fig. 14 we report the best model fits to the difference between the ground–based BC<sub>V</sub> and BC<sub>I</sub> and the corresponding WFPC2 BC<sub>F555W</sub> and BC<sub>F814M</sub> values, as a function of the (F439W–F555W) and (F555W–F814W) WFPC2 colors, respectively. As for the corresponding infrared quantities, cubic polynomial relations with very small scatters are required to transform the BC from the WFPC2 into the ground–based visual photometric system. The numerical relations to transform the selected colors and BCs from the WFPC2 to the ground–based photometric system are listed below: $`(BV)=0.024(F439WF555W)^30.136(F439WF555W)^2+1.060(F439WF555W)0.014`$ $`(VI)=0.049(F555WF814W)^30.077(F555WF814W)^2+1.120(F555WF814W)0.005`$ $`BC_V=BC_{F555W}+0.016(F439WF555W)^30.070(F439WF555W)^2+2.092(F439WF555W)0.001`$ $`BC_I=BC_{F814W}+0.052(F555WF814W)^30.110(F555WF814W)^2+2.189(F555WF814W)0.007`$ ## 6 Conclusions Using our synthesis code we derive homogeneous, self consistent color–temperature relations and BCs in the HST infrared and visual photometric systems. We investigated the behavior of the derived quantities for different stellar parameters and adopted set of model atmospheres. Major results are: * For a given set of model atmospheres, scatters larger than 0.01 mag in the inferred colors are only observed at low temperatures ($``$6000 K) when models with different gravities and, to a lower degree, with different metallicities are compared. * For given stellar parameters, the BCP98 and LCB97 sets of model atmospheres provide significantly different colors (up to a few tenths of a magnitudes) below 5000 K. The inferred discrepancies can be mainly ascribed to the corrections applied by LCB97 to the original grids of Kurucz’s model atmosphere spectra in order to match empirical color–temperature relations. * The inferred BCs are less dependent on the adopted stellar parameters and model atmospheres than the colors, even at low temperatures. * Very similar colors and BCs have been obtained for the different NICMOS and WFPC2 cameras. * Average relations, with a negligible dependence on the stellar parameters and the selected NICMOS or WFPC2 cameras, can be adopted to provide useful transformations between the HST and ground–based photometric systems. In the low temperature domain the ground–based (B–V), (J–H) and (J–K) colors are bluer than the corresponding ones in the HST photometric system, while the ground–based (V–I) color is redder than the corresponding (F555W–F814W) one in the WFPC2 system. We acknowledge Jon Holtzman for his careful Referee Report and comments and Laura Greggio for the helpful discussions and suggestions. We thank the STScI NICMOS Staff for providing the information on the NICMOS throughput curves. L.O. acknowledges the financial support of the “Ministero della Università e della Ricerca Scientifica e Tecnologica” (MURST) to the project Stellar Evolution. ### REFERENCES Bessel, M. S., Castelli, F., & Plez, B. 1998, A&A, 333, 231 (BCP98) Castelli, F., 1997, private communication Holtzman, J. A., Burrows, J., Casertano, S., Hester J., Trauger, J. T., Watson, A. M., & Worthey, G. 1995, PASP, 107, 1065 (H95) Johnson, H. L. 1966, ARA&A, 4, 193 Leitherer, C., & Heckman, T. M. 1995, ApJS, 96, 9 Leitherer, C., et al. 1999, ApJS, in press Lejeune, T., Cuisinier, F., & Buser, R. 1997, A&AS, 125, 229 (LCB97) Montegriffo, P., Ferraro, F.R., Origlia, L., & Fusi Pecci, F. 1998, MNRAS, 297, 872 Stephens, W., Frogel, J.A., Ortolani, S., Davies, R., Jablonka, P., & Renzini, A., 1999, astro-ph/9909001 <sup>0</sup><sup>0</sup>footnotetext: This preprint was prepared with an adapted version of the AAS L$`A`$ macros v4.0.
warning/0001/quant-ph0001037.html
ar5iv
text
# Hybrid exciton-polaritons in a bad microcavity containing the organic and inorganic quantum wells \[ ## Abstract We study the hybrid exciton-polaritons in a bad microcavity containing the organic and inorganic quantum wells. The corresponding polariton states are given. The analytical solution and the numerical result of the stationary spectrum for the cavity field are finished. \] 1. Introduction Quantum wells (QWs) embedded in semiconductor microcavity structures have been the subject of extensive theoretical and experimental investigation. We know that the excitons play a fundamental role in the optical properties of the QWs. The photodevices of excitons may have small size, low power dissipation, rapidness and high efficiency. All of above these are required to the integrated photoelectric circuits. The excitons are classified as Wannier excitons (they have large radius and weak oscillator strength) and Frenkel excitons (they have smaller radius and strong oscillator strength) by the size of the exciton model radius. Recently, a new excitionic state–hybrid exciton state in the composite organic and inorganic semiconductor heterostructure has been described by pioneering work . Since then from quantum well (QW) to quantum dot, the hybrid exciton states due to resonant mixing of Frenkel and Wannier-Mott excitons have been demonstrate. The reference shows that the hybrid excitons possess a strong oscillator strength and a small saturation density (or large radius). The reference proposes to couple Frenkel excitons and Wannier-Mott excitons through a ideal microcavity. There are two different coupling regime into which interaction between the cavity field and the optical transition of the excitons can be classified. One is the weak coupling regime which the exciton-photon coupling is very small and can be treated as a perturbation to the eigenstates of the uncoupled exciton-photon system. Another is the strong coupling regime which the exciton-photon coupling is so strong that it no longer be treated as perturbation. In the strong coupling regime, the QWs excitons emit photons into the cavity. The photon are bounced back by the mirror and reabsorbed by the QWs to create excitons again. So the Rabi oscillation are formed. The exciton-polaritons mode splitting in a semiconductor microcavity is also observed by many experimental group, such as In this paper, we will deal with the hybrid exciton-polaritons states for the organic and in organic QWs in a bad cavity. In the section 2, we use the motion equations included damping effect to give the mixed hybrid-exciton and cavity field modes, that is, hybrid exciton-polaritons. In the section 3, we will give the emission spectrum of the system in the case of the stationary state and the corresponding numerical results also are given. In section 4, a simple conclusion is given. 2. Model and exciton-polariton states We begin with the Hamiltonian of the organic and inorganic quantum wells in a ideal microcavity $`H`$ $`=`$ $`{\displaystyle \underset{k}{}}[\mathrm{}\omega _WA_k^+A_k+\mathrm{}\omega _FB_k^+B_k+\mathrm{}\mathrm{\Omega }a_k^+a_k]`$ (1) $`+`$ $`{\displaystyle \underset{k}{}}[\mathrm{}\mathrm{\Gamma }_{13}(A_k^+a_k+A_ka_k^+)+\mathrm{}\mathrm{\Gamma }_{23}(B_k^+a_k+B_ka_k^+)].`$ (2) where $`A_k`$($`A_k^+`$), $`B_k`$($`B_k^+`$) and $`a_k`$($`a_k^+`$) are usual boson operators for Wannier, Frenkel and cavity field. $`\mathrm{\Gamma }_{13}=\frac{1}{\mathrm{}}P_W^{01}E_0`$ and $`\mathrm{\Gamma }_{23}=\frac{1}{\mathrm{}}P_F^{01}E_0`$, $`P_{W,F}^{01}`$ is the moment matrix element for Wannier and Frenkel exciton from the ground state. $`E_0`$ is the amplitude of the vacuum electric field at the center of the cavity. This Hamiltonian is linear. For a bad microcavity, we could solve it by the motion equation included the damping coefficient and obtain any mixed solutions of the hybrid exciton and cavity field. But here, we only deal with the case of a single mode cavity field. That is, the above equation is simplified into: $`H`$ $`=`$ $`\mathrm{}\omega _WA^+A+\mathrm{}\omega _FB^+B+\mathrm{}\mathrm{\Omega }a^+a`$ (3) $`+`$ $`\mathrm{}\mathrm{\Gamma }_{13}(A^+a+Aa^+)+\mathrm{}\mathrm{\Gamma }_{23}(B^+a+Ba^+).`$ (4) So we have the motion equation $`{\displaystyle \frac{a}{t}}`$ $`=`$ $`i\mathrm{\Omega }ai\mathrm{\Gamma }_{13}Ai\mathrm{\Gamma }_{23}B\gamma _1a`$ (6) $`{\displaystyle \frac{A}{t}}`$ $`=`$ $`i\omega _WAi\mathrm{\Gamma }_{13}a\gamma _2A`$ (7) $`{\displaystyle \frac{B}{t}}`$ $`=`$ $`i\omega _FBi\mathrm{\Gamma }_{23}a\gamma _3B`$ (8) In order to describe the properties of the bad cavity, The damping coefficient $`\gamma _i`$ are added phenomenologically to the above equation (3.a-3.c). In fact, when we write out the interaction between the system and the reservoir, we could give a motion equation which includes the fluctuation terms and dissipative terms by the Markov approximation. However, because we want to discuss the exciton-polaritons in the strong coupling regime. We aren’t interested in the noise properties of the system. So the fluctuation terms may be neglected. By use of the Fourier transformation, we have: $`(i\gamma _1+\omega \mathrm{\Omega })a(\omega )=ia(0)+\mathrm{\Gamma }_{13}A(\omega )+\mathrm{\Gamma }_{23}B(\omega )`$ (10) $`(i\gamma _2+\omega \omega _W)A(\omega )=iA(0)+\mathrm{\Gamma }_{13}a(\omega )`$ (11) $`(i\gamma _3+\omega \omega _F)B(\omega )=iB(0)+\mathrm{\Gamma }_{23}a(\omega )`$ (12) Where, $`a(0)`$, $`A(0)`$ and $`B(0)`$ are initial operators for the cavity field, Wannier excitons and Frenkel excitons respectively. In order to obtain $`a(t)`$, we need solve the pole equation $`(i\gamma _1+\omega \mathrm{\Omega })(i\gamma _2+\omega \omega _W)(i\gamma _3+\omega \omega _F)`$ (13) $`(i\gamma _2+\omega \omega _W)\mathrm{\Gamma }_{23}^2(i\gamma _3+\omega \omega _F)\mathrm{\Gamma }_{13}^2=0`$ (14) The solutions of this cubic equation could be obtained analytically by using any mathematics handbook, such as . But usually, a cubic equation can be solved more quickly with numerical methods than with analytical procedures. So, we set the form of the analytical solutions of the equation (5) are $`\omega _1=\omega _1^{}i\mathrm{\Gamma }_1`$, $`\omega _2=\omega _2^{}i\mathrm{\Gamma }_2`$ and $`\omega _2=\omega _3^{}i\mathrm{\Gamma }_3`$ respectively. If we make $`F(\omega )=i(i\gamma _2+\omega \omega _W)(i\gamma _3+\omega \omega _F)a(0)+`$ (15) $`+i\mathrm{\Gamma }_{23}(i\gamma _2+\omega \omega _W)B(0)+i\mathrm{\Gamma }_{13}(i\gamma _3+\omega \omega _F)A(0)`$ (16) We have $`a(t)`$ as following: $$a(t)=\frac{F(\omega _1)}{\mathrm{\Delta }_1\mathrm{\Delta }_2}e^{i\omega _1t}\frac{F(\omega _2)}{\mathrm{\Delta }_2\mathrm{\Delta }_3}e^{i\omega _2t}+\frac{F(\omega _3)}{\mathrm{\Delta }_1\mathrm{\Delta }_3}e^{i\omega _3t}$$ (17) with $`\mathrm{\Delta }_1=\omega _1\omega _2`$, $`\mathrm{\Delta }_2=\omega _1\omega _3`$ and $`\mathrm{\Delta }_3=\omega _2\omega _3`$. $`\omega _i`$ are determined by the pole equation. This equation indicates that the strong coupling of the two kinds of the QWs exciton states and cavity field results in three new eigenstates. Their eigenvalues are $`\omega _1`$, $`\omega _2`$ and $`\omega _3`$ respectively. These states are just hybrid exciton-polaritons states. Their energy splitting are $`\mathrm{\Delta }_1`$, $`\mathrm{\Delta }_2`$ and $`\mathrm{\Delta }_3`$ respectively 3. Stationary spectrum For the case of the ergodic and stationary process, the emission spectrum of the system is defined as following : $$S(\omega )=_0^{\mathrm{}}e^{i\omega t}<a^+(t)a(0)>dt+c.c.$$ (18) If the cavity field, Wannier exciton and Frenkel exciton are initially in the number states $`|n_c>`$, $`|n_W>`$ and $`|n_F>`$ respectively, then $$<a^+(t)a(0)>=\overline{n}_c[i\frac{E(\omega _1)}{\mathrm{\Delta }_1^{}\mathrm{\Delta }_2^{}}e^{i\omega _1t}+i\frac{E(\omega _2)}{\mathrm{\Delta }_2^{}\mathrm{\Delta }_3^{}}e^{i\omega _2t}i\frac{E(\omega _3)}{\mathrm{\Delta }_1^{}\mathrm{\Delta }_3^{}}e^{i\omega _3t}]$$ (19) where $`\overline{n}_c`$ is mean photon number of the cavity field and $$E(\omega )=(i\gamma _2+\omega \omega _W)(i\gamma _3+\omega \omega _F)$$ (20) As the general exciton-polaritons , If we assume that the damping is moderate, the process is almost ergodic and stationary. It’s deserved to point out that all of the parameters excepting for $`\omega `$ are fixed by the properties of the organic and inorganic QWs as well as microcavity material. We always may choose some moderate parameters so that the stationary condition could be satisfied. So the spectrum of the system is : $$S(\omega )=\frac{A}{(\omega \omega _1^{})]^2+\mathrm{\Gamma }^2_1}+\frac{B}{(\omega \omega _2^{})]^2+\mathrm{\Gamma }^2_2}+\frac{C}{(\omega \omega _3^{})]^2+\mathrm{\Gamma }^3_2}$$ (21) with $`A(\omega )`$ $`=`$ $`2\overline{n}_c{\displaystyle \frac{Re[E(\omega _1)\mathrm{\Delta }_1\mathrm{\Delta }_2(\omega \omega _1)]}{|\mathrm{\Delta }_1|^2|\mathrm{\Delta }_2|^2}}`$ (23) $`B(\omega )`$ $`=`$ $`2\overline{n}_c{\displaystyle \frac{Re[E(\omega _2)\mathrm{\Delta }_3\mathrm{\Delta }_2(\omega \omega _2)]}{|\mathrm{\Delta }_3|^2|\mathrm{\Delta }_2|^2}}`$ (24) $`C(\omega )`$ $`=`$ $`2\overline{n}_c{\displaystyle \frac{Re[E(\omega _3)\mathrm{\Delta }_3\mathrm{\Delta }_1(\omega \omega _3)]}{|\mathrm{\Delta }_3|^2|\mathrm{\Delta }_1|^2}}`$ (25) We find that when the system reaches stability, the hybrid exciton-polaritons spectrum is superposition three Lorentzian lines which are expected. The exciton-polariton splitting may be measured at the peak points of the emission spectrum which are determined by the condition $`\frac{dS(\omega )}{d\omega }=0`$. We apply the general eq.(11) to give a numerical sketch map. We firstly adopt to the assumption of the reference namely $`\omega _F=\mathrm{\Omega }`$, $`\omega _W=\omega _F(1+\delta )`$, $`\delta =10^2`$. Now we give a set of the possible values for the above parameters. $`\overline{n}_c`$ only determines the amplitude of the spectrum, so we set $`\overline{n}_c=1`$. We assume that $`\omega _F=\mathrm{\Omega }=1562meV`$, $`\mathrm{\Gamma }_{23}^2=16meV`$, $`\mathrm{\Gamma }_{13}^2=8meV`$, $`\gamma _1=0.1meV`$, $`\gamma _2=0.18meV`$, $`\gamma _3=0.12meV`$. By using these numbers and eq.(11), we give the sketch map of the spectrum for the system. (Fig.1). This sketch map shows there are sudden change near the three peaks. If we choose moderate parameters, $`A(\omega )`$, $`B(\omega )`$ and $`C(\omega )`$ are slowing varing functions of $`\omega `$ near the peaks and can be considered as constant. So the sudden varing points will disappear, the precise stationary spectrum is given. 4. Conclusion In conclusion, the hybrid exciton-polariton states in a bad microcavity containing the organic and inorganic quantum wells is given. Although we only discuss a single mode model, this approach also could be apply to general case. This paper shows that the hybrid exciton-polaritons decay at three difference rate. The analytical and numerical results of the emission spectrum for the exciton-polaritons are also given.
warning/0001/hep-ph0001287.html
ar5iv
text
# A QCD Analysis of Polarised Parton Densities ## I Introduction It is now more than a decade since the first polarised DIS experiments discovered the strong breaking of a $`SU(6)`$ quark model based sum-rule , and precipitated the “proton spin crisis”. Since then many polarised deep-inelastic scattering (DIS) experiments have reported measurements of the virtual photon asymmetry $$A_1(x,Q^2)=\frac{g_1(x,Q^2)}{F_1(x,Q^2)}$$ (1) on different targets (the structure functions $`g_1`$ and $`F_1`$ are defined later). A polarised proton collider at RHIC will soon begin to constrain the unknown polarised parton distributions even more strongly. Current and future interest in this topic stems partly from the history of the “spin crisis”. However, polarised parton densities are also interesting because of the role they might play in future polarised hadron collider searches for completions of the standard model. Essentially, a large variety of physics beyond the standard model plays with chirality. Some of this freedom can easily be curtailed by polarised scattering experiments, if the polarised parton densities are known with precision. We expect that by the end of the polarised-RHIC program this goal should be reached. The longitudinally polarised structure function $`g_1`$ is defined by $$g_1(x,Q^2)=\stackrel{~}{𝒞}_q\frac{1}{2}\underset{f}{}e_f^2\left[\stackrel{~}{q}_f+\stackrel{~}{\overline{q}}_f\right]+\frac{1}{2}\left(\underset{f}{}e_f^2\right)\stackrel{~}{𝒞}_g\stackrel{~}{g},$$ (2) which is a Mellin convolution of the quark ($`\stackrel{~}{q}_f`$), anti-quark ($`\stackrel{~}{\overline{q}}_f`$) and gluon ($`\stackrel{~}{g}`$) longitudinally polarised distributions with the coefficient functions $`\stackrel{~}{𝒞}_{q,g}`$. The index $`f`$ denotes flavour, and $`e_f`$ is the charge carried by the quark. The unpolarised structure function is given by a similar formula in terms of the corresponding unpolarised densities and coefficient functions. These coefficient functions and the splitting functions (which determine the evolution of the densities) are computable order by order in perturbative QCD. The former are crucial for the Bjorken sum rule , connecting the proton and neutron structure functions, $`g_1^p`$ and $`g_1^n`$, to the neutron $`\beta `$-decay constant $`g_A`$, and are known to NNLO . This makes it possible to use this sum rule for precision measurement of the QCD scale . The polarised splitting functions are known only at NLO . In this paper we analyse the currently available inclusive DIS data in QCD and extract polarised parton distributions from them. In this respect our work is similar to that of . However, our analysis differs in several ways. For one, some of the data we use is more recent than the older fits. More importantly, we relax some of the assumptions which needed to be made in analysing the older data. We allow for flavour asymmetry in the polarised sea quark densities, and let the first moment of the gluon density vary freely in the fit. Furthermore we make a detailed investigation of the uncertainties in these polarised gluon and sea quark densities. The uncertainty in gluon densities may seem puzzling in view of the fact that the $`Q^2`$-dependence of the structure function $`g_1`$ involves the gluon strongly. In fact, at LO we already have— $$\frac{g_1}{\mathrm{log}Q^2}=\frac{\alpha _S}{2\pi }\left[\stackrel{~}{P}_{qq}g_1+\stackrel{~}{P}_{qg}\stackrel{~}{g}\right],$$ (3) where $`\stackrel{~}{P}_{qq}`$ and $`\stackrel{~}{P}_{qg}`$ are polarised splitting functions. Since $`\alpha _S`$ is now very strongly constrained by measurements at LEP and through unpolarised DIS, one might expect that data on $`g_1`$ constrains the polarised gluon densities. Unfortunately, errors on $`g_1`$ are large in the low-$`x`$ region, where the contribution of the gluons dominate, primarily because the asymmetry $`A_1`$ is small at low-$`x`$. For the same reason the flavour singlet sea quark density is also rather loosely constrained by data. We investigate the statistics necessary to improve these constraints through DIS measurements of $`A_1`$ at polarised HERA. The plan of this paper is the following. In the next section we discuss the various technicalities that distinguish different global analyses. This section also serves to set up the notation. This is followed by a section that discusses our choice of data used in the fit. The next section contains our results for the LO and NLO fits, and a detailed consideration of the parameter errors. A section on some applications of our parametrisations follows this. The final section contains a summary of our main results. ## II Constraints on parton densities ### A Parton densities and structure functions With $`N_f`$ flavours of quarks, we need to fix $`2N_f+1`$ parton densities. These are for the $`2N_f`$ flavours of quarks and anti-quarks and the gluon. For protons or neutrons, the quark and anti-quark densities for the strange and heavier flavours are equal. We work with the two (flavour non-singlet) polarised valence quark densities, $`\stackrel{~}{V}_u`$ and $`\stackrel{~}{V}_d`$, corresponding to the up and down flavours. The other non-singlet densities we use are those corresponding to the diagonal generators of $`SU(5)`$ flavour— $`\stackrel{~}{q}_32(\stackrel{~}{\overline{u}}\stackrel{~}{\overline{d}}),`$ $`\stackrel{~}{q}_82(\stackrel{~}{\overline{u}}+\stackrel{~}{\overline{d}}2\stackrel{~}{\overline{s}}),`$ (4) $`\stackrel{~}{q}_{15}2(\stackrel{~}{\overline{u}}+\stackrel{~}{\overline{d}}+\stackrel{~}{\overline{s}}3\stackrel{~}{\overline{c}}),`$ $`\stackrel{~}{q}_{24}2(\stackrel{~}{\overline{u}}+\stackrel{~}{\overline{d}}+\stackrel{~}{\overline{s}}+\stackrel{~}{\overline{c}}4\stackrel{~}{\overline{b}}).`$ (5) The initial conditions for evolution are that below and at each flavour threshold, the density for that flavour of quarks is zero. Thus, below the charm threshold we have $`\stackrel{~}{q}_{24}=\stackrel{~}{q}_{15}=\stackrel{~}{q}_0`$ and below the bottom threshold we set $`\stackrel{~}{q}_{24}=\stackrel{~}{q}_0`$. For the singlet quark density, we use $$\stackrel{~}{q}_02\underset{f}{}\stackrel{~}{\overline{q}}_f=\stackrel{~}{\mathrm{\Sigma }}\underset{f}{}\stackrel{~}{V}_f,$$ (6) in preference to the usual $`\stackrel{~}{\mathrm{\Sigma }}`$ (which is the sum over quark and anti-quark densities of all flavours). The evolution equations couple $`\stackrel{~}{q}_0`$ to the gluon density $`\stackrel{~}{g}`$. We also define similar unpolarised quark and gluon densities<sup>*</sup><sup>*</sup>*Our convention is that polarised quantities are distinguished from the corresponding unpolarised ones by a tilde.. Finally, the structure functions $`g_1^p`$ and $`g_1^n`$, for the proton and neutron, are given by eq. (2). The unpolarised structure functions $`F_1^{p,n}`$ are given by the analogous expression where the polarised parton densities are replaced by the unpolarised densities. An isospin flip, interchanging $`\stackrel{~}{V}_u`$ and $`\stackrel{~}{V}_d`$ and switching the sign of $`\stackrel{~}{q}_3`$, relates $`g_1^p`$ and $`g_1^n`$. After correcting for nuclear effects, the normalised structure function for deuterium is $`g_1^d=(g_1^p+g_1^n)/2`$, and a similar expression for $`F_1^d`$. We shall have occasion to use the first moments of various polarised parton distributions. We introduce the notation— $`\mathrm{\Gamma }_u(Q^2)={\displaystyle _0^1}𝑑x\stackrel{~}{V}_u(x,Q^2),`$ $`\mathrm{\Gamma }_d(Q^2)={\displaystyle _0^1}𝑑x\stackrel{~}{V}_d(x,Q^2),`$ (7) $`\mathrm{\Gamma }_i(Q^2)={\displaystyle _0^1}𝑑x\stackrel{~}{q}_i(x,Q^2),`$ $`\mathrm{\Gamma }_g(Q^2)={\displaystyle _0^1}𝑑x\stackrel{~}{g}(x,Q^2).`$ (8) We will also use the notation $`\mathrm{\Gamma }_V^0=\mathrm{\Gamma }_u+\mathrm{\Gamma }_d`$ and $`\mathrm{\Gamma }_V^3=\mathrm{\Gamma }_u\mathrm{\Gamma }_d`$. The notation $$\mathrm{\Gamma }_1^{n,p}(Q^2)=_0^1𝑑xg_1^{n,p}(x,Q^2)$$ (10) is fairly standard. We shall use it in the text. We also use the notation $`\mathrm{\Gamma }_{\overline{u}}`$, etc., to denote the first moments of the flavoured sea densities. ### B The fitting strategy We use experimental data on the asymmetries $`A_1^p`$, $`A_1^n`$ and $`A_1^d`$, measured on proton, neutron ($`{}_{}{}^{3}\mathrm{He}`$) and deuterium targets to constrain the polarised parton densities. We assume full knowledge of the unpolarised parton densities as given by some global fit, so that the structure function $`F_1`$ can be reconstructed using appropriate NLO coefficient functions. Then the data on $`A_1`$ can be converted to $`g_1`$. We prefer this method to taking the $`g_1`$ values presented by experiments, since different experimental groups may make different assumptions about the unpolarised structure functions. Such effects would lead to additional normalisation uncertainties in any global fit. We have chosen to use the CTEQ4 set of parton densities in our work. We do not expect this choice to affect our conclusions strongly since the unpolarised parton densities now have smaller errors than data on the polarisation asymmetry $`A_1`$. However, with this choice we are constrained to follow some of the assumptions made by the CTEQ group— 1. We work in the $`\overline{\mathrm{MS}}`$ scheme, since the CTEQ group does that. Other possibilities would have been to work in the AB or JET schemes, but then we would have had to transform the CTEQ distributions. We prefer to avoid this procedure, since the best fit parton densities in one scheme do not necessarily transform into the best fit densities in another scheme. 2. We retain the CTEQ choice for the charm quark mass being 1.6 GeV and the bottom quark mass to be 5.0 GeV. At each mass threshold, we increase the number of flavours by one, and treat the newly activated flavour as massless immediately above the threshold. Parton distributions and $`\alpha _S`$ are continuous across these thresholds . 3. We are constrained to use the $`\mathrm{\Lambda }_{QCD}`$ values used in . 4. We take $`Q_0^2=2.56`$ GeV<sup>2</sup> in order to avoid having to evolve the unpolarised parton densities downwards. In future we plan to study the results of relaxing one or more of these restrictions. We follow the parametrisation of CTEQ4 and write— $$\stackrel{~}{f}(x,Q_0^2)=a_0x^{a_1}(1x)^{a_2}(1+a_3x^{a_4}),$$ (11) for all densities apart from $`q_3`$, which is parametrised as $$\stackrel{~}{q}_3(x,Q_0^2)=a_0x^{a_1}(1x)^{a_2}(1+a_3\sqrt{x}+a_4x).$$ (12) We have made the choice that the large-$`x`$ behaviour of any polarised density is the same as that of the unpolarised density; in other words, the parameter $`a_2`$ is the same for corresponding polarised and unpolarised densities (this assumption is sometimes given the name “helicity retention property” ). For simplicity we have also equated the polarised and unpolarised values of $`a_4`$ when this parameter is a power. Finally, at $`Q_0^2`$ we have extended some of the CTEQ assumptions for unpolarised parton densities to polarised. These include equating the values of $`a_1`$ for $`\stackrel{~}{V}_u`$, $`\stackrel{~}{V}_d`$ and $`\stackrel{~}{q}_3`$, taking $`a_4=1`$ for $`\stackrel{~}{q}_0`$, equating the values of $`a_2`$ for $`\stackrel{~}{q}_0`$ and $`\stackrel{~}{q}_3`$. We also retain the choice $`\eta 2\stackrel{~}{s}/(\stackrel{~}{\overline{u}}+\stackrel{~}{\overline{d}})=1/2`$ in some of our fits, but let it vary in others. Although these assumptions seem overly restrictive, the quality of the data does not allow us to fit many of these parameters. We discuss some of these points later in this paper. The main difference between our parametrisation and previous ones is that we explicitly include a non-zero $`\stackrel{~}{q}_3(x,Q_0^2)`$ and break $`SU(2)`$ flavour symmetry in the polarised sea. This part of the sea density is actually quite well constrained, and plays a crucial role in our fits. ### C Sum rules In a three flavour world, we can write down the following sum rule for the first moments of the nucleon structure functions in NLO QCD— $$\mathrm{\Gamma }_1^{p,n}=\left(\pm \frac{1}{6}g_3+\frac{1}{18}g_8+\frac{1}{9}g_0\right)\left\{1\frac{\alpha _S}{\pi }\right\}.$$ (13) The anomalous dimensions on the right hand side of this equation have been calculated in the $`\overline{\mathrm{MS}}`$ scheme . Two facts used are that in the $`\overline{\mathrm{MS}}`$ scheme the first moment of the NLO quark coefficient $`\stackrel{~}{𝒞}_q^{(1)}=2`$, and the first moment of the gluon coefficient function is $`\stackrel{~}{𝒞}_g^{(1)}=0`$. The upper and lower signs belong to protons and neutrons, respectively. The quantities $`g_3`$, $`g_8`$ and $`g_0`$ are baryonic axial couplings. They are defined as matrix elements of axial vector currents between baryon states. Due to the axial anomaly, the singlet axial-vector current is not conserved. As a result, $`g_0`$ picks up a $`Q^2`$ dependence . Hence, $`g_0`$, the moments, and $`\alpha _S`$ have to be evaluated at the same $`Q^2`$ in eq. (13). It is not easy to extract $`g_0`$ from low-energy hadron data, although there have been some attempts to do this using elastic $`\nu p`$ scattering . This gives $`g_0=0.14\pm 0.27`$. Lattice computations and QCD sum rules also give similar numbers, but have systematic uncertainties which have to be removed in future. The quantity $`g_3=1.2670\pm 0.0035`$ is obtained from the neutron beta-decay constant. The coupling $`g_8`$ is extracted from the decay of strange to non-strange baryons. $`SU(3)`$ flavour symmetry is used crucially in this extraction . The PDG result is $`g_8=0.579\pm 0.025`$ . Using the coefficient functions in the $`\overline{\mathrm{MS}}`$ scheme, the moments of the structure functions can be expressed as $$\mathrm{\Gamma }_1^{p,n}(Q^2)=\left\{1\frac{\alpha _S}{\pi }\right\}\left[\frac{5}{18}\mathrm{\Gamma }_V^0(Q^2)\pm \frac{1}{6}\left(\mathrm{\Gamma }_V^3(Q^2)+\mathrm{\Gamma }_3(Q^2)\right)+\frac{1}{18}\mathrm{\Gamma }_8(Q^2)+\frac{2}{9}\mathrm{\Gamma }_0(Q^2)\right],$$ (14) where the upper (lower) sign is for the proton (neutron). Apart from eq. (2), we have used the definitions of the non-singlet densities in eq. (5) and the singlet parton density in eq. (6). There are two sum rules which can be obtained by equating the right-hand sides of eqs. (13) and (14). Alternatively, we could use some linear combinations. The only one that removes the coupling $`g_0`$ is the difference $`\mathrm{\Gamma }_1^p\mathrm{\Gamma }_1^n`$, and gives the Bjorken sum rule . At NLO this is— $$\mathrm{\Gamma }_V^3+\mathrm{\Gamma }_3=g_3.$$ (15) Note that the first moments of non-singlet densities are independent of $`Q^2`$ order by order to all orders. As a result, this equation is also valid in this form to all orders. We impose this form of the Bjorken sum rule on our fits. The sum of the two gives the Ellis-Jaffe sum rule when the additional assumption $`g_8=g_0`$ is made. This cannot be correct in QCD because $`g_0`$ is $`Q^2`$ dependent and $`g_8`$ is not. Moreover, in the absence of a real measurement of $`g_0(Q^2)`$, an Ellis-Jaffe type sum rule cannot constrain the parton densities as the Bjorken sum rule does. Hence, we use such a sum rule to extract $`g_0`$ rather than to impose it as a constraint on parton densities. ### D Positivity Polarisation asymmetries are the ratios of the difference and sum of physically measurable cross sections. Since cross sections are non-negative, asymmetries are bounded by unity in absolute value. In the parton model or in LO QCD, these cross sections are directly related to parton densities. Hence positivity of cross sections imply $$\left|\frac{\stackrel{~}{f}(x,Q^2)}{f(x,Q^2)}\right|1$$ (16) for the ratio of each polarised and unpolarised density to leading order in QCD. In our LO fits, we impose these restrictions. At NLO and beyond, this simple relation between parton densities and cross sections no longer holds. Parton densities are renormalisation scheme dependent objects; although universal, they are not physical. Hence they need not satisfy positivity , instead one must impose positivity on the actual cross sections. This is numerically difficult and requires knowledge of a variety of cross sections evaluated to NLO. Since this knowledge is lacking, and for numerical simplicity, we have instead imposed eq. (16) on all our parton density fits. ### E Choice of numerical techniques Our numerical goal is to evolve parton density functions with absolute errors of at most $`10^3`$. If this design goal were reached, then numerical errors would lie at least an order of magnitude below all other errors. We integrate the evolution equations using a 4-th order Runge-Kutta algorithm. The Mellin convolutions required in the evaluation of the derivative are computed using a Gauss-Legendre integral. The parton densities are evaluated on a grid and interpolated using a cubic spline method. All the numerical algorithms may be found in . The knot points of the cubic spline are selected to give an accuracy of $`10^5`$ in the evaluation of the parton densities. The Mellin convolutions are also accurate to this order. We require the Runge-Kutta to give us integration errors bounded by $`10^4`$. This gives us the error limits we require. We can test these estimates by checking that all sum rules are satisfied to within $`10^3`$. On a 180 MHz R10000 processor, the program takes about 0.15 CPU seconds to evolve the parton densities by $`\mathrm{\Delta }Q^2=1`$ GeV<sup>2</sup>. ## III Selection of data Experiments do not measure the asymmetry $`A_1`$ directly; they measure the asymmetry between the cross sections for lepton and longitudinally polarised hadrons being parallel and anti-parallel— $$A_L=\frac{d\sigma ()d\sigma ()}{d\sigma ()+d\sigma ()},$$ (17) or a similar asymmetry, $`A_T`$, with transversely polarised hadrons. These asymmetries are related to the two that we require by $$A_L=D(A_1+\eta A_2)\mathrm{and}A_T=d(A_2\xi A_1),$$ (18) where $`D`$ and $`d`$ are depolarisation factors for the virtual photon and $`\xi `$ and $`\eta `$ are essentially kinematic constants. In terms of the ratio of the Compton scattering cross sections for transversely and longitudinally polarised virtual photons, $$R=\frac{\sigma _L}{\sigma _T}=\frac{F_22xF_1}{2xF_1},$$ (19) we can write $$D=\frac{y(2y)}{y^2+2(1y)(1+R)},\mathrm{and}\eta =2\gamma \frac{1y}{2y}.$$ (20) Here $`\gamma =2Mx/Q1`$. Using the degree of transverse polarisation of the virtual photon, $$ϵ=\frac{1y}{1y+y^2/2},$$ (21) we can write $$d=D\sqrt{\frac{2ϵ}{1+ϵ}},\mathrm{and}\xi =\eta \frac{1+ϵ}{2ϵ}.$$ (22) Since $`\gamma `$ is very small in the DIS region, the relations $`A_L=DA_1`$ and $`A_T=dA_2`$ are actually satisfied to high accuracy. We then use the further relations $$A_1=(g_1\gamma ^2g_2)/F_1\mathrm{and}A_2=\gamma (g_1+g_2)/F_1,$$ (23) to obtain eq. (1) when $`\gamma 1`$. It is clear from the second equation that $`g_2`$ is difficult to measure. The main theoretical uncertainty in measurements of $`A_1`$ is in the values of $`R`$ used. In fact, many experiments use $`R`$ in two ways. First, it enters the expression for $`D`$ and $`d`$, and hence is used to compute QED corrections for initial state radiationWe thank Abhay Deshpande for drawing our attention to this point. and to construct $`A_1`$ and $`A_2`$ from $`A_L`$ and $`A_T`$. Next, it is used along with measurements of $`F_2`$ to compute $`F_1`$ and thus relate $`g_1`$ to $`A_1`$. We bypass this second use of $`R`$ by utilising experimental data on $`A_1`$ instead of $`g_1`$. We are forced, however, to accept the first use of $`R`$, In any case, differences between experiments in their estimates of $`D`$ should be factored into the overall normalisation errors. The SMC collaboration has data from muon scattering off both proton and deuterium targets. Data was taken in separate runs in 1993 and 1996. The most recent publication for $`A_1`$ is ; this supersedes previously published data. The E-143 experiment at SLAC has data from electron scattering off proton, deuterium and $`{}_{}{}^{3}\mathrm{He}`$ targets. Their most recent publication is , which supersedes all previous published data on $`A_1(x,Q^2)`$ by this collaboration. The E-154 experiment at SLAC has data from electron scattering off $`{}_{}{}^{3}\mathrm{He}`$ targets . The HERMES collaboration in DESY has data from positron scattering off protons and $`{}_{}{}^{3}\mathrm{He}`$ . We have also used data on DIS from $`{}_{}{}^{3}\mathrm{He}`$ taken by the SLAC E-142 collaboration . We have chosen not to utilise data taken by the older EMC collaboration and the E-140 experiments at SLAC. Deuterium is a spin-1 nucleus with the $`p`$ and $`n`$ primarily in a relative $`s`$-wave state. The $`d`$-wave probability is estimated to be $`\omega _D=0.05\pm 0.01`$ . This is used in the relation between the structure function of deuterium and those of $`p`$ and $`n`$$`g_1^d=(13\omega _D/2)(g_1^p+g_1^n)/2`$. In $`{}_{}{}^{3}\mathrm{He}`$, the two protons are essentially paired into a spin singlet, and the asymmetry is largely due to the unpaired neutron. Corrections due to other components of the nuclear wave-function are small . More details are available in . From the chosen experiments, we have retained only the data on $`A_1(x,Q^2)`$ for $`Q^22.56`$ GeV<sup>2</sup>. While this does remove some of the low-$`x`$ data, the error bars in the removed data are pretty large. We have checked by backward evolution that the data which is removed would not have constrained the fits any further. The total number of data points used in our analysis is 224. In most fitting procedures the statistical errors on measurements are combined in some way with the systematic error estimates. Both sets of errors are usually reported in the literature in each bin of data. Whereas this procedure is acceptable for statistical errors, it oversimplifies the nature of systematic errors. These latter are correlated from bin to bin, and one must use the full covariance matrix of errors in the analysis. In the absence of published information on the covariance matrix, one may make the simplifying assumption that the bin-to-bin correlation vanishes, and add the statistical and systematic errors in quadrature. This overestimates the errors on data and hence the errors on the parameters determined by fitting. We have made a different extremal assumption of neglecting the systematic errors altogether. This procedure almost certainly leads us to under-estimate the parameter errors— a point to be borne in mind when we discuss large errors and uncertainties in the fits. In summary, our choice of error analysis is deliberately conservative. Since $`g_2`$ contains a possible twist-3 contribution, which cannot be written in terms of parton distributions, we cannot utilise data on $`g_2`$ for our fits. However, the twist-2 part is completely determined by $`g_1`$. In a later section, we report an attempt to limit the extent of the twist-3 term using our fitted polarised parton densities. For this we have utilised data on proton target from SMC and the E-143 collaboration at SLAC , on deuterium target from SMC, E-143 and SLAC E-155 , and on neutron target from the E-143 and SLAC experiment E-154 . In all cases, we have used the most recent data set and analysis from each collaboration. The quality of data on $`g_2`$ is poorer than that for $`g_1`$. This is because $`A_2`$ is small, and extraction of $`g_2`$ from $`A_2`$ requires the subtraction of $`g_1`$, which itself has significant measurement errors. The errors are dominated by statistical uncertainties. There remains data from semi-inclusive DIS taken by the SMC and HERMES experiments. Analysis of these require fragmentation functions and their $`Q^2`$ evolution. Since such analyses are still to reach the stage that parton densities have, utilising parametrisations for fragmentation functions would introduce larger errors into our fits. For this reason, we have chosen not to use such data in this work. ## IV Results We have made four full analyses— two LO and two NLO, each with and without $`SU(2)`$ flavour symmetry (denoted S and $`\overline{\mathrm{S}}`$ respectively) for the sea quarks, and with fixed $`\eta =0.5`$. In addition, we have made a set of fits with $`SU(2)`$ symmetric sea ($`\stackrel{~}{q}_3=0`$) but $`\eta `$ allowed to vary freely. At LO this had no effect on the fit— $`\eta =0.5`$ gave the best fit and the remaining parameters were identical to LO S. At NLO $`\eta `$ moved to 0.6, but the parameters remained close to the set NLO S. The goodness of fit improved only marginally when $`\eta `$ was allowed to float. Given the uncertainties in the remaining parameters we retain the choice $`\eta =0.5`$ in the main work reported below. The goodness of fit, and the constraints imposed by each set of data are summarised in Table I. The values of $`\chi ^2`$ favour the NLO sets slightly. The HERMES proton and the E-154 neutron data (see Figures 1, 2) express the strongest preferences for the NLO fits; almost the entire change in $`\chi ^2`$ in going from LO to NLO comes from these two data sets. It is probably no coincidence that these two data sets also have the smallest error bars. The parameters and their error estimates are shown in Tables IIV. It is worth noting that the NLO $`\overline{\mathrm{S}}`$ densities lie well within the limits allowed by eq. (16), as do all the densities for NLO S except the gluon. The parameters $`a_0`$ and $`a_3`$ for $`\stackrel{~}{g}`$ in NLO S are at the limit of positivity. The quality of the NLO $`\overline{\mathrm{S}}`$ fit is shown in Figures 13. We recommend that the NLO parametrisations be used with the CTEQ4M set of unpolarised parton densities and the LO with the CTEQ4L set , and with appropriate values of $`\mathrm{\Lambda }_{QCD}`$. For the $`\overline{\mathrm{S}}`$ densities, the normalisation of $`\stackrel{~}{q}_3`$ inherits its error from the valence parameters and the coupling $`g_3`$. It is the best constrained among the parameters describing the sea. Similarly, for the two S densities the normalisation of $`\stackrel{~}{V}_d`$ is fixed by the Bjorken sum rule and its error is inherited from the remaining valence parameters. In Fig 4 we have shown the variation of $$\mathrm{\Delta }\chi ^2=\chi ^2\chi _{\mathrm{min}}^2$$ (24) when one of the parameters in $`\stackrel{~}{g}`$ is varied for fixed values of all the other parameters in the set. The minimum of these curves fixes the best-fit value of the parameter, and the points where $`\mathrm{\Delta }\chi ^2=1`$ give the 68.3% confidence limits on this parameter. Uncertainties in the gluon density are shown in greater detail in Figure 5. This shows contour lines of $`\mathrm{\Delta }\chi ^2=2.3`$, which encloses the area with 68.3% probability of giving a good description of the data. Also shown in the figure are lines of constant $`\mathrm{\Gamma }_g(Q_0^2)`$. Although $`1\sigma `$ contours give negative values of $`\mathrm{\Gamma }_g`$, we note that $`3\sigma `$ contours include positive values as well. Note that our error bars are deliberately conservative, due to our neglect of systematic errors. Since systematic errors are as large as the statistical errors, the naive procedure of summing them in quadrature would have led us to believe that at NLO positive $`\mathrm{\Gamma }_g`$ is allowed at $`1.5\sigma `$. The huge uncertainties in the gluon distribution due to these parameter variations are illustrated in Figure 6. At $`x=0.01`$ the polarised gluon density at NLO can lie anywhere in the range from $`10`$ to $`50`$. This uncertainty at low-$`x`$ prevents us from investigating this theoretically interesting region. The situation is worse at LO where $`\stackrel{~}{g}`$ is at the limit of integrability, since $`a_11`$. Thus $`\mathrm{\Gamma }_g`$ is essentially undetermined. This large uncertainty in $`\stackrel{~}{g}`$ comes because the only constraint on gluon densities at present are the data on $`Q^2`$ variations of $`g_1`$. Furthermore, the data at $`x<0.1`$ are most effective in constraining $`\stackrel{~}{g}`$, and in this range, the data have large errors. We have quantified this in Figure 7, where we show selected data on $`g_1`$ in different bins of $`x`$ as a function of $`Q^2`$ (the normalisation has been made arbitrary for ease of viewing). As the bands of variation due to $`\stackrel{~}{g}`$ (in the NLO $`\overline{\mathrm{S}}`$ set) show, the data does not constrain the gluon density well. A few qualitative statements are in order. It is clear that in a scaling theory, DIS data with nucleon or nuclear targets can at most fix two linear combinations of quark densities. However, DIS structure functions are $`Q^2`$-dependent. Hence, in principle, data of arbitrarily high accuracy fixes these two linear combinations at each $`Q^2`$i.e., an infinite number of functions. In QCD there is a more economical description of this $`Q^2`$ dependence involving the set of parton densities given in Section 2. Of course, in the real world data are never infinitely precise, so the question is how accurately does $`Q^2`$ evolution fix these densities. Some part of the answer is clear from Figure 7— improved data at low $`x`$ will constrain $`\stackrel{~}{g}`$ much better and $`g_1^d`$, being iso-singlet, would present the best constraint. We investigated this question quantitatively by generating fake data at $`x<0.1`$ from the NLO S set. The values of $`A_1^p`$ so generated were smeared randomly over a 10% band to simulate noise in the data, and error bars of 20–30% were assigned to each such data point. This faked set is meant to mimic data that could possibly come from a future polarised HERA experiment. We redid our fits with this faked data set replacing all data on $`A_1^p`$ for $`x<0.1`$. This data brings down the error bars in the parameter $`a_0`$ appearing in $`\stackrel{~}{g}`$ by a factor of 4, and improves the errors in $`a_1`$ by a factor of 2–3. The error estimates obtained with the fake data set are shown in Figure 8. The region of parameter space allowed by this fake data is shown as the grey patch in Figure 5. Taking data at $`x<0.01`$ or over a larger range of $`Q^2`$, both of which would be feasible at polarised HERA, would constrain $`\stackrel{~}{g}`$ even better . The ability of RHIC to fix the gluon densities is, of course, well appreciated. However, if DIS experiments can fix the gluon densities better, then polarised RHIC can be used as a discovery machine. ## V Applications ### A Flavour asymmetry Unlike previous fits of parton densities which had built in the constraint $`\stackrel{~}{\overline{d}}\stackrel{~}{\overline{u}}`$ , we have allowed for sea quark densities that violate flavour $`SU(2)`$ symmetry. The fits show that the data tolerate, and even prefer, strong flavour symmetry violations. This is easily seen in the first moments of various densities (Table VI). Since $`\mathrm{\Gamma }_0`$ and $`\mathrm{\Gamma }_8`$ are small, it is clear that the pattern $`|\mathrm{\Gamma }_s||\mathrm{\Gamma }_{\overline{u}}|`$ and $`\mathrm{\Gamma }_{\overline{d}}\mathrm{\Gamma }_{\overline{u}}`$ must follow whenever $`\mathrm{\Gamma }_3`$ is large. It has been suggested that $`SU(2)`$ flavour asymmetry in the sea be observed through two combinations of cross sections for production of $`W^\pm `$ in longitudinally polarised pp scattering— $$A_{\sigma ,\delta }=\frac{\sigma (W^+,)\pm \sigma (W^{},)\sigma (W^+,)\sigma (W^{},)}{\sigma (W^+,)\pm \sigma (W^{},)+\sigma (W^+,)\pm \sigma (W^{},)},$$ (25) where $`A_\sigma `$ ($`A_\delta `$) is defined with the upper (lower) signs. At LO these asymmetries can be written as the ratio of certain combinations of polarised and unpolarised parton densities. At LO the asymmetry $`A_{DY}`$ for Drell-Yan pairs can also be written in terms of the parton densities. At zero rapidity, $`A_{DY}`$ is a function of $`\sqrt{\tau }=M/\sqrt{S}`$, where $`M`$ is the mass of the pairs and $`\sqrt{S}`$ is the center of mass energy of the colliding protons. For $`W^\pm `$ production at zero rapidity, the parton densities have to be evaluated at $`M=M_W`$, and hence $`\sqrt{S}=M_W/\sqrt{\tau }`$. In Figure 9 we have shown these asymmetries at zero rapidity as a function of $`\sqrt{S}`$, or equivalently of $`\sqrt{\tau }`$. It is clear that at $`\sqrt{S}`$ appropriate to RHIC, the iso-triplet spin asymmetry, $`A_\delta `$, is best suited to distinguish the LO S densities from LO $`\overline{\mathrm{S}}`$. Experimentally studying the dependence of $`A_\delta `$ on $`\sqrt{S}`$ over even a limited range below $`\sqrt{S}=500`$ GeV would be very useful. $`A_{DY}`$ is the least suitable measurement for making this distinction. We have already pointed out that the LO fit is unable to decide on the sign of $`\mathrm{\Gamma }_g`$, and that the NLO fits yield a negative $`\mathrm{\Gamma }_g`$, although positive values are not ruled out. Although previous fits have seen overlapping ranges of allowed $`\mathrm{\Gamma }_g`$, the theoretical bias has been to take large and positive values of $`\mathrm{\Gamma }_g`$. This sign can be easily fixed by various experiments at RHIC or in charm production measurements at HERA or the COMPASS experiment in CERN . We would like to caution that parton densities are renormalisation scheme dependent (and hence unphysical). They are universally applicable to all experiments, as long as each experiment is treated in the same scheme . Our determination of these densities are in the $`\overline{\mathrm{MS}}`$ scheme, and statements about their moments are therefore also restricted to this scheme. When interpreting the moments of unphysical parton densities, their scheme dependence must be held in mind. ### B Structure functions and couplings It is possible to construct physical quantities out of the unphysical first moments of the parton densities. For the first moments of the structure functions as given in eq. (14), we obtain the values— $`\mathrm{\Gamma }_1^p=0.136\pm 0.008,`$ $`\mathrm{\Gamma }_1^n=0.053\pm 0.008(\mathrm{NLO}\overline{\mathrm{S}}),`$ (26) $`\mathrm{\Gamma }_1^p=0.157\pm 0.006,`$ $`\mathrm{\Gamma }_1^n=0.032\pm 0.006(\mathrm{NLO}\mathrm{S})`$ (27) at $`Q_0^2=2.56`$ GeV<sup>2</sup>. These values are within $`1.5\sigma `$ of each other and they compare well with values deduced from experiments . We can also use eq. (13) to extract the value of $`g_0`$. Using as input the above values of $`\mathrm{\Gamma }_1^{p,n}`$ derived from our fits and the PDG value for $`g_8`$, we find $$g_0(Q_0^2)=\{\begin{array}{cc}0.13\pm 0.20\hfill & \text{(NLO }\overline{\mathrm{S}}\text{),}\hfill \\ 0.35\pm 0.15\hfill & \text{(NLO S).}\hfill \end{array}$$ (28) Since $`g_0`$ is a physical quantity, it is only to be expected that our determination of $`g_0`$ should agree with other analyses, such as , even if they use some other scheme to arrive at the same result. We will, of course, disagree with them on any scheme dependent quantity, such as $`\mathrm{\Delta }\mathrm{\Sigma }=\mathrm{\Gamma }_u+\mathrm{\Gamma }_v+\mathrm{\Gamma }_0`$. Our maximally flavour symmetry violating fits give physically reasonable results. ### C The structure function $`g_2`$ Wandzura and Wilczek have derived a sum rule relating the twist-2 part of $`g_2`$ to $`g_1`$ $$g_2^{WW}(x,Q^2)=g_1(x,Q^2)+_x^1\frac{g_1(y,Q^2)}{y}dy.$$ (29) An additional twist-2 contribution to $`g_2`$, suppressed by the ratio of the quark to the nucleon mass , is ignored here. Predictions for the twist-3 contribution have been made using bag models , QCD sum rules as well as from non-perturbative lattice QCD computations . Since some computations predict large twist-3 contributions to moments of $`g_2`$, it becomes interesting to check whether the data on $`g_2`$ allows such contributions. Figure 10 shows our “prediction” for the twist-2 part of $`g_2`$ and compares it to measurements of this structure function. Clearly the data is compatible with the NLO twist-2 prediction (and also with the parton model result, $`g_2=0`$). Between the prediction and the data, there is little room for a twist-3 contribution. Statistics have to be improved vastly in order to study higher-twist effects. In fact, COMPASS hopes to make this measurement . Since the statistical errors are smallest for $`g_2^d`$, it seems that this is the best candidate in which to look for twist-3 effects. However the data quality needs improvement even here. There is considerable scaling violation in the twist-2 part of $`g_2`$, but the large errors prevent any analysis of the $`Q^2`$-dependence. ### D The valence densities Recently the HERMES collaboration has used semi-inclusive polarised DIS data to extract valence and sea quark densities . We have not used these in our fits because this analysis is performed with parton model formalæ. Nevertheless, it is interesting to compare our fits with these numbers. We display this comparison in Figure 11. The rough agreement is heartening, but the small systematic differences between the fit results and the HERMES extraction of the valence densities shows the need for a more accurate QCD analysis of the experimental data, taking into account properly the $`Q^2`$ dependence through NLO evolution. ### E Axion-matter coupling We present an example of the application of polarised proton scattering to physics beyond the standard model. The Peccei-Quinn solution to the strong CP problem postulates a global symmetry whose spontaneous breakdown generates a (nearly) massless pseudo-Goldstone boson called the axion . There is a variant of the original model which is still viable . The axion, $`a`$, whose decay constant is $`f_a`$, couples to fermions, $`\psi _f`$, of mass $`m_f`$ by the term $$_{int}=ig_f\overline{\psi }_f\gamma _5\psi _fa.$$ (30) The coupling $`g_f=C_f(m_f/f_a)`$. The effective Peccei-Quinn charge, $`C_f=X_f/N`$, appears in the coupling instead of the actual charge $`X_f`$. Here, $`N`$ is given by $`X_f`$. There have been several studies of the effective Peccei-Quinn charge of the proton. For three flavours, the LO expression can be written as $$C_{p,n}=\underset{f}{}(C_f\mu _f)\left[\mathrm{\Gamma }_f+\mathrm{\Gamma }_{\overline{f}}\right].$$ (31) where $`\mu _f=M/m_f`$ with $`1/M=1/m_f`$ . $`C_f`$ for quarks and leptons is highly model dependent. In the so-called KSVZ , and other hadronic axion models, $`C_u=C_d=C_s=0`$. Using quark mass ratios $`m_u/m_d=0.568\pm 0.042`$ and $`m_u/m_s=0.0290\pm 0.0043`$, obtained by chiral perturbation theory , and our LO $`\overline{\mathrm{S}}`$ fits, we find that $$C_p=0.402(2)\mathrm{and}C_n=0.058(4).$$ (32) The statistical errors in this coupling are dominated by the errors in the fits to polarised parton densities. The uncertainty in the quark masses give smaller contributions to these errors. However, the real source of uncertainty comes from higher loop corrections. An estimate of this theoretical uncertainty can be obtained by inserting the NLO results for various moments into eq. (31). This changes $`C_p`$ by 25% and $`C_n`$ by 100%. The chiral couplings of neutralinos and charginos in generic supersymmetric extensions of the standard model also give rise to effective couplings with matter which depend on the moments of the parton densities. Such couplings are often needed in astrophysical contexts. Unless these couplings are examined to 2-loop order, they should not be evaluated with the NLO moments. ## VI Conclusions We have made global QCD analyses of data on the asymmetry $`A_1`$ from polarised DIS without making overly restrictive assumptions about the flavour content of the sea. We have extracted polarised parton densities which can be used with the CTEQ4 set of unpolarised densities. Our NLO analyses (in the $`\overline{\mathrm{MS}}`$ scheme) yield the parameter sets given in Tables II and III, and the LO analyses give the sets displayed in Tables IV and V. Strong $`SU(2)`$ flavour symmetry violation in the sea is supported by the sets of densities in which the first moment of the singlet sea quark density is much smaller than that of the triplet sea quark density. We also found in the NLO fits that the first moment of the gluon density is preferentially negative, although positive values are within 2–3 $`\sigma `$ of the best fits. The LO fits, to the contrary, prefer a positive sign for $`\mathrm{\Gamma }_g(Q_0^2)`$, although negative values are allowed within 2–3 $`\sigma `$ of the central values. All physical quantities obtained from the first moments of our fitted densities have completely sensible values, as they must have. We have shown that the polarised HERA option may yield highly accurate measurements of the polarised gluon densities if measurements of $`A_1`$ in the range $`x0.125`$ can be performed with errors of about 25%. Measurements of the asymmetry in the iso-triplet part of $`W^\pm `$ production at RHIC are likely to be able to pin down the flavour content of the sea. Measurements at future facilities for spin physics thus nicely complement each other. We have checked that our parametrisations are roughly consistent with semi-inclusive DIS data, although a full QCD analysis of this data remains to be performed. We have also shown that these parametrisations, when used to determine the twist-2 part of $`g_2`$ leave very little room for a twist-3 part to this structure function. Finally, we have determined the coupling of hadronic axions to matter— an input into several astrophysical constraints on the invisible axion. In a future publication we plan to make a more detailed study of several issues, including the proper inclusion of systematic experimental errors into the analysis and several other technical issues concerning NLO QCD global fits. We would like to thank Willy van Neerven for several discussions and clarifications. We also thank the organisers of the 6th Workshop on High Energy Physics Phenomenology (WHEPP-6), Chennai, January 2000, where a portion of this work was completed.
warning/0001/cond-mat0001248.html
ar5iv
text
# Magnetopolaron effects in light reflection and absorption of a three-level system in a QW ## I Introduction When a polaron state is formed in a magnetic field the role of the electron-phonon interaction grows sharply under the resonant condition $$\omega _{LO}=j\mathrm{\Omega },j=1,2,3,\mathrm{}..,$$ (1) $`\omega _{LO}`$ is the LO phonon frequency, $`\mathrm{\Omega }`$ is the electron (hole) cyclotron frequency. The intersections of the electron-phonon system energy levels as functions of magnetic field stem (see Fig.1). A transition to the magnetopolaron states results in energy levels splitting in crossing points. For the first time magnetopolaron energy level splitting was discovered in light absorption in bulk $`InSb`$ Polaron state formation takes place in both three (3D) and two dimensional (2D) semiconducting systems. These states are very important in frequency dependence formation of magnetooptical effects, such as reflection, interband absorption, cyclotron resonance and Raman scattering (see, for instance, the reviews Refs. ). The systems distinguish by the electron (hole) spectrum: in 3D systems these are one-dimensional Landau bands , whereas in 2D structures these are discrete energy levels. This difference results in different splitting of the energy levels of the electron-phonon system: In 3D systems it is the value of order $`\alpha ^{2/3}\mathrm{}\omega _{LO}`$ whereas in 2D structures it is the value of order $`\alpha ^{1/2}\mathrm{}\omega _{LO}`$ where $`\alpha `$ is Fröhlich’s non-dimensional electron-LO phonon coupling constant. A perfect single QW of a finite depth is considered below as an example of a 2D system. The inhomogeneous broadening effects of the energy levels are ignored. We shall calculate the non-dimensional reflection coefficient $``$ and the non-dimensional light absorption coefficient $`𝒜`$ which are determined as the relation of the reflected or absorbed light flux to the incident one. There are both the EHP discrete energy levels and the continuum spectrum. The optical effects are due to the resonance of the energies of incident light $`\mathrm{}\omega _l`$ and the EHP or magnetopolaron discrete energy levels. Far from the $`H`$ and $`\omega _l`$ values, corresponding to the magnetopolaron resonance, we will consider the system as a two-level one (the first level is the crystal ground state energy and the second is the EHP energy level). In the vicinity of the magnetopolaron resonance we consider the system as a three-level one, including the ground state energy and two polaron levels. Light absorption and reflection in quasi-2D systems had been continually considered earlier. In calculations of absorption usually the low order perturbation theory on the light-electron system interaction (see, for instance, Refs. ) had been used. The result for the light absorption coefficient contains the factor $`\mathrm{\Delta }_{\gamma _\rho }(\mathrm{}\omega _lE_\rho )`$, where $`E_\rho `$ is the electron excitation energy measured from the ground state energy, $$\mathrm{\Delta }_\gamma (E)=\frac{1}{\pi }\frac{\mathrm{}\gamma /2}{E^2+(\mathrm{}\gamma /2)^2}$$ (2) is the function transiting into the Dirac $`\delta `$ \- function when $`\gamma 0`$, $`\gamma _\rho `$ is the non-radiative inverse lifetime on the energy level $`\rho `$. This result implicates some self-contradiction: The value $`\mathrm{\Delta }_\gamma (E)\mathrm{}`$ at $`E=0`$ and $`\gamma 0`$, whereas the non-dimensional absorption coefficient $`𝒜`$ cannot exceed the value of one. This self-contradiction is lifted if one goes out the perturbation theory limits on the light-electron coupling constant, i. e. the contributions of all the orders on this coupling constant are summarized. This summarizing means taking into account the sequence of all the absorption and reradiation processes of the light quantum $`\mathrm{}\omega _l`$. There appears a new concept of the radiation lifetime $`\gamma _{r\rho }^1`$ of an electronic excitation which has been introduced firstly for the excitons in a QW at $`H=0`$ At first, the new approach had been used to describe light reflection from QW at frequencies $`\omega _l`$ close to an exciton energy, then an appropriate theory for light absorption has been created. It has been shown that the previous results obtained with the help of the perturbation theory are applicable under condition $$\gamma _{r\rho }<<\gamma _\rho .$$ (3) The new theory has to be used to describe the features of light reflection and absorption by ideal QWs in magnetic fields because the non-radiative values $`\gamma _\rho `$ for the discrete energy levels of EHPs are small and the condition of Eq. (3) may be unfeasible. We will calculate the coefficients $``$ and $`𝒜`$ for the case of the incident light perpendicular to the QW plane, taking into account the magnetopolaron effects. The paper is organized as follows. The magnetopolaron classification is described in Seq. II, the expressions for the electric fields left and right from a QW at the normal irradiation by the incident light in the case of the multilevel system are obtained in Seq. III. The formulae for light reflection and absorption in a QW are calculated in Seq. IY. In Seq. Y and Seq. YI, the inverse radiative lifetime for an EHP in QW in magnetic field and the inverse radiative lifetimes of two magnetopolaron states are calculated, respectively. The inverse non-radiative lifetimes of the magnetopolaron states are calculated in Seq. YII. The numerical calculation results for light absorption and reflection are given in Seq. YIII. ## II Magnetopolaron classification Fig. 1 shows (the full lines) the non-dimensional electron-phonon system energy levels $`E/\mathrm{}\omega _{LO}`$ in a QW, pertaining to the size-quantization quantum number $`l`$, as functions of $`\mathrm{\Omega }/\omega _{LO}=j`$, $$\mathrm{\Omega }=|e|H/m_{e(h)}c$$ (4) is the cyclotron frequency, $`e`$ is the electron charge, $`m_{e(h)}`$ is the electron (hole) effective mass. $`E`$ is measured from the energy $`\epsilon _l^{e(h)}`$ corresponding to the $`l`$ size-quantization energy level. It is supposed that all the phonons relevant to the magnetopolaron creation have the same non-dispersal frequency $`\omega _{LO}`$. The polaron states correspond to the crossing points. The ”double” polarons are denoted with the black circles and appropriate to crossings of two terms only. Let us consider some crossing point corresponding to the number $`j`$ (see Eq. (1)). $`n`$ is the Landau quantum number of the energy level coming through this crossing point, the phonon number $`N=0`$ (see Fig. 1). Then the condition of the double polaron existence is $$2j>nj.$$ (5) It is easy to note that the only double polaron $`A`$ corresponds to $`j=1`$. Two double polarons $`D`$ and $`E`$ correspond to $`j=2`$, three double polarons $`F`$, $`K`$ and $`L`$ correspond to $`j=3`$ and so on. The polarons on the left of the value $`\mathrm{\Omega }/\omega _{LO}=1/3`$ are not noted in Fig.1. The triple polarons (corresponding to three terms crossings) are above the double polarons, the quateron polarons are above the triple polarons and so on. There are $`j`$ polarons of every sort at the given $`j`$. For the first time the triple polarons in bulk semiconductors and in QWs have been considered in Ref. and in Refs. , respectively. Note that the Landau level’s equidistancy is needed for three and more levels crossings. The triple polaron theory is relevant to the case when the energy corrections due to the band non-parabolicity or excitonic effects are smaller than the triple polaron splitting. But in the case of the double polarons, the equidistancy violation is not an obstacle because crossing of two terms exists at any rate. All the above mentioned polarons correspond to the integer values of $`j`$. At Fig. 1 there are some other crossings of the terms with the size-quantization quantum number $`l`$ (solid lines) marked with empty circles. They are relevant to the fractional values of $`j`$. Because these crossings are characterized by the values $`\mathrm{\Delta }N2`$, the real direct transitions with emitting of one LO phonon are impossible. We call such polarons weak ones. To calculate splittings, one has to take into account the virtual transitions between the crossing terms or two-phonon interaction. Splittings of the weak polarons $`\mathrm{\Delta }E_w`$ are much smaller than for the polarons with integer $`j`$. The contributions of the intermediate transitions into the value $`\mathrm{\Delta }E_w`$ are of higher order on $`\alpha `$ than $`\alpha ^{1/2}`$. The crossing depiction becomes much more complicated when two or more values of the size-quantization quantum numbers $`l`$ are taken into account.Besides the ordinary polarons corresponding to the energy level $`l^{}`$ (for example the polaron $`A^{}`$) the ”combined” polarons appear where electron-phonon interaction bounds two electron energy levels with different Landau quantum numbers $`n`$ and different $`l`$ (or with different $`l`$ and the same $`n`$). Two terms corresponding to the quantum number $`l^{}`$ (dotted lines) and the combined polaron $`P`$ and $`Q`$ positions (tilted crosses) are shown in Fig. 1. The combined weak polaron is marked with $`R`$. The combined polarons have an interesting feature: The corresponding resonant magnetic field values depend on the distance $`\mathrm{\Delta }\epsilon =\epsilon _l^{}\epsilon _l`$ between the size-quantization energy levels $`l`$ and $`l^{}`$ and consequently on the QW’s depth and width. With the help of Fig. 1 one can obtain easily $$\left(\frac{\mathrm{\Omega }}{\omega _{LO}}\right)_P=1\frac{\mathrm{\Delta }\epsilon }{\mathrm{}\omega _{LO}},\left(\frac{\mathrm{\Omega }}{\omega _{LO}}\right)_Q=1+\frac{\mathrm{\Delta }\epsilon }{\mathrm{}\omega _{LO}}.$$ (6) One more combined polaron sort exists under condition $$\mathrm{\Delta }\epsilon =\mathrm{}\omega _{LO},$$ (7) when, for instance, the terms $`l^{},n=0,N=0`$ and $`l,n=0,N=1`$ coincide at any magnetic field values. The definite separation between the energy levels $`l`$ and $`l^{}`$ is necessary to satisfy the condition (7). This could be done by choosing the QW’s depth and width. The situation depicted in Fig. 1 is applicable when the energy separations between the neighbor size-quantization levels $`l`$, $`l1`$, $`l+1`$ are much more than the values $`\mathrm{\Delta }E`$ of polaron splittings. So long as the separations diminish with growing of the QW’s width, some restriction is implicated on the value $`d`$ from above (see the numerical estimates in Ref. ). ## III Electric fields out of QW Let us suppose that an electromagnetic (EM) wave (or some superposition of EM waves, corresponding to a light impulse) falls perpendicular on a single QW from the left. The electric field intensity is <sup>*</sup><sup>*</sup>*One can remove the complex conjugate term and introduce (instead of $`_{0\alpha }(\omega )`$) $$E_{0\alpha }(\omega )=_{0\alpha }(\omega )+_{0\alpha }^{}(\omega )=2\pi E_0[e_{l\alpha }D_0(\omega )+e_{l\alpha }^{}D_0^{}(\omega )].$$ $$E_{0\alpha }(z,t)=\frac{1}{2\pi }_{\mathrm{}}^{\mathrm{}}𝑑\omega e^{i\omega (tzn/c)}_{0\alpha }(\omega )+c.c.,$$ (8) where $`n`$ is the barrier refraction index. Let us introduce the designations $$_{0\alpha }(\omega )=2\pi E_0e_{l\alpha }𝒟_0(\omega ),$$ (9) where $`E_0`$ is the scalar amplitude, $`𝐞_l`$ is the incident light polarization vector ; $`𝒟_0(\omega )`$ is a frequency function; $$𝒟_0(\omega )=\delta (\omega \omega _l)$$ (10) for monochromatic light of frequency $`\omega _l`$. It is supposed that the incident EM wave has a circular polarization $$𝐞_l=\frac{1}{\sqrt{2}}(𝐞_x\pm i𝐞_y),$$ (11) where $`𝐞_x,𝐞_y`$ are the unit vectors along the $`x`$ and $`y`$ axis, respectively. The incident EM wave generates the excited electronic states with the energies characterized by the index $`\rho `$. For instance, in a magnetic field the index $`\rho `$ for the EHP includes $`l_e,l_h,n_e=n_h`$. For the infinitely deep QW $`l_e=l_h`$. The state $`\rho `$ is characterized by the energy $`\mathrm{}\omega _\rho `$, which is measured from the ground state energy, by the inverse radiative lifetime $`\gamma _{r\rho }`$ and by the inverse non-radiative lifetime $`\gamma _\rho `$. Let us consider the QW with $`d<<c/\omega _ln`$. Then the electric field intensities $`𝐄_{left(right)}(z,t)`$ on the left and right of the QW are determined as $$𝐄_{left(right)}(z,t)=𝐄_0(z,t)+\mathrm{\Delta }𝐄_{left(right)}(z,t),$$ (12) $`\mathrm{\Delta }E_{\alpha left(right)}(z,t)`$ $`=`$ $`E_0e_{l\alpha }{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑\omega e^{i\omega (t\pm zn/c)}𝒟(\omega )+`$ (13) $`+`$ $`c.c.,`$ (14) where the upper (lower) sign is relevant to the index $`left`$ ($`right`$). The frequency partitition is determined as It is supposed in Eq. (15) that both circular polarizations corresponds to the excitation of one of two types of EHPs the energies of which are equal each other (see Eq. (88) below). $$𝒟(\omega )=\frac{4\pi \chi (\omega )𝒟_0(\omega )}{1+4\pi \chi (\omega )},$$ (15) $`\chi (\omega )`$ $`=`$ $`{\displaystyle \frac{i}{4\pi }}{\displaystyle \underset{\rho }{}}{\displaystyle \frac{\gamma _{r\rho }}{2}}[(\omega \omega _\rho +i\gamma _\rho /2)^1+`$ (16) $`+`$ $`(\omega +\omega _\rho +i\gamma _\rho /2)^1].`$ (17) Eqs. (12)–(15) determine the electric fields on the both sides of the QW and, consequently, the values of the energy fluxes in transmitted and reflected light for the multilevel system. The second term in the square brackets in the RHS of Eq. (15) is non-resonant, but it is necessary to satisfy the relation $$\chi ^{}(\omega )=\chi (\omega ).$$ (18) In a real QW of a finite depth in magnetic field, there are always discrete energy levels (one at least) together with the continuous spectrum. Thus, the index $`\rho `$ in the RHS of Eq. (15) takes both discrete and continuous values. Sometimes in a magnetic field, the large level number has to be taken into account (see where the results of an excitation with an asymmetrical light impulse have been investigated). Discussing a two-level system we mean a situation when all the energy levels can be neglected, but one which is resonant with $`\mathrm{}\omega _l`$. The absorption and reflection coefficients for a two-level system with one excitonic energy level at $`H=0`$ have been calculated in Refs. . The electric field intensities are expressed as $`E_{\alpha left}`$ $`=`$ $`E_0e_{l\alpha }\{e^{i\omega _l(tzn/c)}`$ (19) $``$ $`{\displaystyle \frac{i\gamma _r}{2(\omega _l\omega +i\mathrm{\Gamma }/2)}}e^{i\omega _l(t+zn/c)}\}+c.c.,`$ (20) $`E_{\alpha right}`$ $`=`$ $`E_0e_{l\alpha }e^{i\omega _l(tzn/c)}\times `$ (21) $`\times `$ $`\left\{1{\displaystyle \frac{i\gamma _r}{2(\omega _l\omega +i\mathrm{\Gamma }/2)}}\right\}+c.c.,`$ (22) $$\mathrm{\Gamma }=\gamma +\gamma _r.$$ (23) One can obtain Eq. (17) and Eq. (18) with the help of Eqs. (10), (12)–(15) if one term in the sum on $`\rho `$ in the RHS of Eq. (15) is preserved using the designations $$\omega _\rho =\omega ,\gamma _{r\rho }=\gamma _r,\gamma _\rho =\gamma ,$$ (24) and neglecting the second non-resonant term in the square brackets in the RHS of Eq. (15). The three-level system consists of the ground state of an electronic system and two energy levels $`\mathrm{}\omega _{1(2)}`$ with the inverse radiative $`\gamma _{r1(2)}`$ and non-radiative $`\gamma _{1(2)}`$ lifetimes. The calculation of the electric field intensities for the three-level system is a little more complicated because one has to solve the square equation (see Eq. (23) below). Applying Eqs. (10), (12)–(15) and omitting the non-resonant term in the RHS of Eq. (15) one obtains $`E`$ $`{}_{\alpha left}{}^{}(z,t)=E_0e_{l\alpha }\{e^{i\omega _l(tzn/c)}ie^{i\omega _l(t+zn/c)}\times `$ (25) $`\times `$ $`[{\displaystyle \frac{\overline{\gamma }_{r1}}{2(\omega _l\mathrm{\Omega }_1+iG_1/2)}}+{\displaystyle \frac{\overline{\gamma }_{r2}}{2(\omega _l\mathrm{\Omega }_2+iG_2/2)}}]\}+`$ (26) $`+`$ $`c.c.,`$ (27) $`E`$ $`{}_{\alpha right}{}^{}(z,t)=E_0e_{l\alpha }e^{i\omega _l(tzn/c)}\times `$ (28) $`\times `$ $`\left\{1i\left[{\displaystyle \frac{\overline{\gamma }_{r1}}{2(\omega _l\mathrm{\Omega }_1+iG_1/2)}}+{\displaystyle \frac{\overline{\gamma }_{r2}}{2(\omega _l\mathrm{\Omega }_2+iG_2/2)}}\right]\right\}`$ (29) $`+`$ $`c.c..`$ (30) The complex quantities $`O_1=\mathrm{\Omega }_1iG_1/2`$ and $`O_2=\mathrm{\Omega }_2iG_2/2`$ ($`\mathrm{\Omega }_{1(2)}`$ and $`G_{1(2)}`$ are real on their definition) are the precise solutions of the equation $`(O`$ $``$ $`\omega _1+i\gamma _1/2)(O\omega _2+i\gamma _2/2)+`$ (31) $`+`$ $`i{\displaystyle \frac{\gamma _{r1}}{2}}(O\omega _2+i\gamma _2/2)+`$ (32) $`+`$ $`i{\displaystyle \frac{\gamma _{r2}}{2}}(O\omega _1+i\gamma _1/2)=0,`$ (33) which are equal $$(\mathrm{\Omega }iG/2)_{1,2}=\frac{1}{2}\left[\stackrel{~}{\omega }_1+\stackrel{~}{\omega }_2\pm \sqrt{(\stackrel{~}{\omega }_1\stackrel{~}{\omega }_2)^2\gamma _{r1}\gamma _{r2}}\right],$$ (34) where $$\stackrel{~}{\omega }_{1(2)}=\omega _{1(2)}i\mathrm{\Gamma }_{1(2)}/2,\mathrm{\Gamma }_{1(2)}=\gamma _{1(2)}+\gamma _{r1(2)}.$$ (35) Note that the value in the square root in Eq. (24) is, generally speaking, a complex one. In the RHS of Eq. (24) the upper (lower) sign corresponds to the index 1(2). The following designations are used $$\overline{\gamma }_{r1}=\gamma _{r1}+\mathrm{\Delta }\gamma ,\overline{\gamma }_{r2}=\gamma _{r2}\mathrm{\Delta }\gamma ,$$ (36) $`\mathrm{\Delta }\gamma `$ $`=`$ $`{\displaystyle \frac{\gamma _{r1}[\mathrm{\Omega }_2\omega _2i(G_2\gamma _2)/2]}{\mathrm{\Omega }_1\mathrm{\Omega }_2+i(G_2G_1)/2}}+`$ (37) $`+`$ $`{\displaystyle \frac{\gamma _{r2}[\mathrm{\Omega }_1\omega _1i(G_1\gamma _1)/2]}{\mathrm{\Omega }_1\mathrm{\Omega }_2+i(G_2G_1)/2}}.`$ (38) Comparing Eq. (21) and Eq. (22) with Eq. (17) and Eq. (18) for the two-level system, one can see that the addition of the contributions from the energy levels 1 and 2 is accompanied by the substitutions: $`\mathrm{\Omega }_{1(2)}`$ instead of $`\omega _{1(2)}`$, $`G_{1(2)}`$ instead of $`\mathrm{\Gamma }_{1(2)}`$ and $`\overline{\gamma }_{r1(2)}`$ instead of $`\gamma _{r1(2)}`$. To calculate the electric field intensities for the four-level system one has to solve a third order equation and so on. It is impossible to solve this task precisely for an arbitrary number $`\rho `$ of the energy levels. However, in the case $$\gamma _{r\rho }<<\gamma _\rho ,$$ (39) when the perturbation theory on light-electron system interaction is applicable (what results in the possibility of neglecting the term $`4\pi \chi (\omega )`$ in the denominator in the RHS of Eq. (14)), we obtain the result for the arbitrary number of the energy levels $`E`$ $`{}_{\alpha left}{}^{many}(z,t)=E_0e_{l\alpha }\{e^{i\omega _l(tzn/c)}`$ (40) $``$ $`ie^{i\omega _l(t+zn/c)}{\displaystyle \underset{\rho }{}}{\displaystyle \frac{\gamma _{r\rho }}{2(\omega _l\omega _\rho +i\gamma _\rho /2)}}\}+c.c.,`$ (41) $`E`$ $`{}_{\alpha right}{}^{many}(z,t)=E_0e_{l\alpha }e^{i\omega _l(tzn/c)}\times `$ (42) $`\times `$ $`\left[1i{\displaystyle \underset{\rho }{}}{\displaystyle \frac{\gamma _{r\rho }}{2(\omega _l\omega _\rho +i\gamma _\rho /2)}}\right]+c.c.,`$ (43) ## IV Light reflection and absorption coefficients. Having the electric field intensities out of the QW we can calculate the reflection and absorption coefficients of light. The Umov-Poynting vector $`𝐒_{left}`$ on the left of the QW is $$𝐒_{left}=𝐒_0+\mathrm{\Delta }𝐒_{left},$$ (44) where $$𝐒_0=\frac{c}{2\pi }E_0^2𝐞_𝐳$$ (45) is the incident light flux, $`\mathrm{\Delta }𝐒_{left}`$ is the reflected light flux directed along $`𝐞_z`$. The light reflection coefficient is determined as $$=\frac{|\mathrm{\Delta }𝐒_{left}|}{|𝐒_0|},$$ (46) the non-dimensional light absorption coefficient is determined as $$𝒜=\frac{|𝐒_{left}𝐒_{right}|}{|𝐒_0|},$$ (47) and the light transition coefficient is $$𝒯=1𝒜=\frac{|𝐒_{right}|}{|𝐒_0|}.$$ (48) With the help of Eq. (17) and Eq. (18) we obtain the results for the two-level system $$=\frac{(\gamma _r/2)^2}{(\omega _l\omega )^2+(\mathrm{\Gamma }/2)^2},$$ (49) $$𝒜=\frac{\gamma \gamma _r/2}{(\omega _l\omega )^2+(\mathrm{\Gamma }/2)^2},$$ (50) obtained earlier in Refs. . Eq. (37) contains the very important result: Light absorption in a QW is absent completely when $`\gamma =0`$ Let us consider two limit cases: $`\gamma _r<<\gamma `$ and $`\gamma _r>>\gamma `$. In the case $`\gamma _r<<\gamma `$ the perturbation theory on light-electron interaction is applicable in the lowest order. In this approximation, the values of the light absorption and reflection coefficients are of the second and fourth order on interaction, respectively. From Eq. (36) and Eq. (37) we obtain $$\frac{\pi \mathrm{}}{2}\frac{\gamma _r^2}{\gamma }\mathrm{\Delta }_\gamma [\mathrm{}(\omega _l\omega )],$$ (51) $$𝒜\pi \mathrm{}\gamma _r\mathrm{\Delta }_\gamma [\mathrm{}(\omega _l\omega )],$$ (52) where $`\mathrm{\Delta }_\gamma (E)`$ is determined in Eq. (2). At $`\gamma _r<<\gamma `$ $$𝒜<<1,<<𝒜.$$ (53) In the opposite case $`\gamma _r>>\gamma `$ in Eq. (36) and Eq. (37) $`\mathrm{\Gamma }`$ has to be substituted approximately by $`\gamma _r`$. Then we obtain that in the resonance $`(\omega _l=\omega )=1`$, which means the total light reflection, $`𝒜<<1`$. Thus, light absorption is small for both cases $`\gamma _r<<\gamma `$ and $`\gamma _r>>\gamma `$. The maximum value $`𝒜(\omega _l=\omega )=1/2`$ is reached in the case $`\gamma _r=\gamma `$. With the help of Eq. (33), Eq. (34) and Eq. (21), Eq. (22) we obtain for the three-level system $``$ $`=`$ $`{\displaystyle \frac{1}{4Z}}\{[\gamma _{r1}(\omega _l\omega _2)+\gamma _{r2}(\omega _l\omega _1)]^2+`$ (54) $`+`$ $`(\gamma _{r1}\gamma _2+\gamma _{r2}\gamma _1)^2/4\},`$ (55) $`𝒜`$ $`=`$ $`{\displaystyle \frac{1}{2Z}}\{\gamma _{r1}\gamma _1[(\omega _l\omega _2)^2+(\gamma _2/2)^2]+`$ (56) $`+`$ $`\gamma _{r2}\gamma _2[(\omega _l\omega _1)^2+(\gamma _1/2)^2]\},`$ (57) where $$Z=[(\omega _l\mathrm{\Omega }_1)^2+(G_1/2)^2][(\omega _l\mathrm{\Omega }_2)^2+(G_2/2)^2].$$ (58) It follows from Eq. (42) that light absorption by the three-level system is equal 0, if $$\gamma _1=\gamma _2=0.$$ (59) Applying the fact that the quantities $`(\mathrm{\Omega }iG/2)_{1,2}`$ are the roots of Eq. (23), we transform the denominator in Eq. (43) to the form $`Z`$ $`=`$ $`[(\omega _l\omega _1)(\omega _l\omega _2)(\gamma _{r1}\gamma _2+\gamma _{r2}\gamma _1+\gamma _1\gamma _2)/4]^2+`$ (60) $`+`$ $`[(\omega _l\omega _1)\mathrm{\Gamma }_2+(\omega _l\omega _2)\mathrm{\Gamma }_1]^2/4.`$ (61) Let us obtain the simplified expressions for $``$ and $`𝒜`$ for the different limits. It is convenient to use Eq. (41) and Eq. (42) with the substitution Eq. (43) or, sometimes, with Eq. (45). The first limit case corresponds to the inequalities $$\gamma _{r1(2)}<<\gamma _{1(2)},$$ (62) when the perturbation theory on light-electron interaction is applicable what corresponds to neglecting by the term $`4\pi \chi (\omega )`$ in the RHS of Eq. (14). Under condition Eq. (46) in the RHS of Eq. (43) we suppose $`\mathrm{\Omega }_{1(2})\omega _{1(2)},G_{1(2)}=\gamma _{1(2)}`$ and obtain $``$ $``$ $`{\displaystyle \frac{(\gamma _{r1}/2)^2}{(\omega _l\omega _1)^2+(\gamma _1/2)^2}}+{\displaystyle \frac{(\gamma _{r2}/2)^2}{(\omega _l\omega _2)^2+(\gamma _2/2)^2}}+`$ (63) $`+`$ $`{\displaystyle \frac{\gamma _{r1}\gamma _{r2}}{2}}\times `$ (64) $`\times `$ $`{\displaystyle \frac{(\omega _l\omega _1)(\omega _l\omega _2)+\gamma _1\gamma _2/4}{[(\omega _l\omega _1)^2+(\gamma _1/2)^2][(\omega _l\omega _2)^2+(\gamma _2/2)^2]}},`$ (65) $$𝒜\frac{\gamma _{r1}\gamma _1/2}{(\omega _l\omega _1)^2+(\gamma _1/2)^2}+\frac{\gamma _{r2}\gamma _2/2}{(\omega _l\omega _2)^2+(\gamma _2/2)^2}.$$ (66) According to Eq. (48) the absorption coefficient $`𝒜`$ is the sum of contributions Eq. (39) from the levels 1 and 2, because under condition Eq. (46) absorption is linear on the constants $`\gamma _{r1}`$ and $`\gamma _{r2}`$. According to Eq. (47) the reflection coefficient $``$ is square-law on $`\gamma _{r1}`$ and $`\gamma _{r2}`$. Therefore it contains an interferential contribution besides the separate levels contributions. The functions $`𝒜(\omega _l)`$ and $`(\omega _l)`$ are represented on Fig. 2 under condition Eq. (46) for the case $`\gamma _{r1}=\gamma _{r2}`$, $`\gamma _1=\gamma _2`$. Generalizing Eq. (48) for the case of the arbitrary number of the energy levels and under condition Eq. (28) one obtains $$𝒜^{many}\frac{1}{2}\underset{\rho }{}\frac{\gamma _{r\rho }\gamma _\rho }{(\omega _l\omega _\rho )^2+(\gamma _\rho /2)^2}.$$ (67) The next limit case (opposite to the preceding one) is determined by the conditions $$\gamma _{r1(2)}>>\gamma _{1(2)}.$$ (68) We suppose $`\gamma _1=\gamma _2=0`$ in the RHS of Eq. (41) and Eq. (42). Then $`𝒜=0`$ and one obtains $``$ $`(\gamma _1=\gamma _2=0)=[(\gamma _{r1}+\gamma _{r2})/2]^2(\omega _l\omega _0)^2\times `$ (69) $`\times `$ $`[(\omega _l\omega _1)^2(\omega _l\omega _2)^2+`$ (70) $`+`$ $`[(\gamma _{r1}+\gamma _{r2})/2]^2(\omega _l\omega _0)^2]^1,`$ (71) where $$\omega _0=\frac{\omega _1\gamma _{r2}+\omega _2\gamma _{r1}}{\gamma _{r1}+\gamma _{r2}}.$$ (72) The peculiar properties of light reflection from the three-level system follow from Eq. (51) in the case of the dominant inverse radiative lifetimes. At any quantities $`\gamma _{r1}`$ and $`\gamma _{r2}`$ in the point $`\omega _l=\omega _0`$ light reflection equals 0; in the points $`\omega _l=\omega _1`$ and $`\omega _l=\omega _2`$ the reflection coefficient equals 1. Let us suppose $`\gamma _{r1}=\gamma _{r2}=\gamma _r`$. Then $`=0`$ at $`\omega _l=(\omega _1+\omega _2)/2`$. At $`\gamma _r>>(\omega _1\omega _2)`$ in the point $`\omega _l=(\omega _1+\omega _2)/2`$ there is a narrow minimum descending to $`=0`$. The half-width of this minimum on the half-depth equals to $`(\omega _1\omega _2)^2/2^{3/2}\gamma _r`$ . The functions $`(\omega _l)`$ are represented in Fig. 3b under conditions $$\gamma _1=\gamma _2=\gamma ,\gamma _{r1}=\gamma _{r2}=\gamma _r,\gamma <<\gamma _r.$$ (73) The functions $`𝒜(\omega _l)`$ are represented on Fig. 3a under the same condition Eq. (53). Note a peculiar turn of $`𝒜(\omega _l)`$ curves at a fixed quantity $`\gamma `$ and at growing $`\gamma _r`$. Transiting from the curve 1 to the curve 6 the quantity $`\mathrm{}\gamma _r`$ takes a successive row of values: 0.002, 0.005, 0.008, 0.04, 0.125, 0.5. There are two maxima on the first curve at $`\mathrm{}\gamma _r=0.002`$. On the consecutive curves there is one maximum which reaches its largest value $`𝒜_{max}=0.5`$ on the curve 5 ($`\mathrm{}\gamma _r=0.125`$), and afterwards its height $`𝒜_{max}`$ diminishes. The quantity $`𝒜_0`$ corresponding to the central point $`\omega _l=(\omega _1+\omega _2)/2`$ and coinciding with the height $`𝒜_{max}`$ on the curves 2 \- 6 is described by the precise formula $$𝒜_0=\frac{4\gamma _r\gamma [(\omega _1\omega _2)^2+\gamma ^2]}{[(\omega _1\omega _2)^2+2\gamma _r\gamma +\gamma ^2]^2},$$ (74) which follows from Eq. (42) at $`\omega _l=(\omega _1+\omega _2)/2`$. The quantity $`𝒜_0`$ maximizes at $$\gamma _{r0}=\frac{(\omega _1\omega _2)^2+\gamma ^2}{2\gamma }.$$ (75) Substituting the values $`\mathrm{}(\omega _1\omega _2)=0.005`$ and $`\mathrm{}\gamma =0.0001`$ in Eq. (55), what corresponds Fig. 3a, one obtains $`\mathrm{}\gamma _{r0}=0.125`$ and $`𝒜_{max}=𝒜_0=0.5`$ (the curve 5 on Fig.3a). The next limit realizes when the reverse lifetimes $`\gamma _{1(2)}`$ and $`\gamma _{r1(2)}`$ are small in comparison with the interlevel separation $`\omega _1\omega _2`$. Let us suppose $$\gamma _{1(2)}<<\omega _1\omega _2,\gamma _{r1(2)}<<\omega _1\omega _2.$$ (76) The interrelation between the non-radiative $`\gamma _{1(2)}`$ and radiative $`\gamma _{r1(2)}`$ lifetimes can be an arbitrary one. Then, if the frequency $`\omega _l`$ is close to the resonance with one of the levels , let us say $`\omega _l\omega _1`$, one obtains from Eq. (41) and Eq. (42) $``$ $``$ $`{\displaystyle \frac{\gamma _{r1}^2/4}{(\omega _l\omega _1)^2+(\mathrm{\Gamma }_1/2)^2}},`$ (77) $`𝒜`$ $``$ $`{\displaystyle \frac{\gamma _{r1}\gamma _1/2}{(\omega _l\omega _1)^2+(\mathrm{\Gamma }_1/2)^2}},`$ (78) which coincides with the results of Eq. (36) and Eq. (37) for the two-level system. Lastly we consider the case of the merging curves. Supposing $$\gamma _1=\gamma _2=\gamma ,\omega _1=\omega _2=\omega ,\gamma _{r1}=\gamma _{r2}=\gamma _r,$$ (79) we obtain from Eq. (41) and Eq. (42) $$\frac{\gamma _r^2}{(\omega _l\omega )^2+(\gamma _r+\gamma /2)^2},$$ (80) $`𝒜{\displaystyle \frac{\gamma _r\gamma }{(\omega _l\omega )^2+(\gamma _r+\gamma /2)^2}}.`$ (81) These results are similar to Eq. (36) and Eq. (37) for the two-level system, but the quantity $`\gamma _r`$ is reduplicated. That means that in the case of a double degenerated excited level the formulae for the two-level system with the reduplicated value $`\gamma _r`$ are applicable. ## V EHP radiative lifetime for QW in magnetic field. One could calculate the EHP inverse radiative lifetime in magnetic field indirectly: calculating the absorption coefficient $`𝒜`$ with the help of the perturbation theory and then, comparing the obtained result with Eq. (39), to calculate $`\gamma _r`$. The second way is the direct calculation of $`\chi (\omega )`$ (see Eq. (15)). But we shall demonstrate the third way: The direct calculation of $`\gamma _r`$ with the help of the Fermi Golden Rule $`\gamma _{r\eta }={\displaystyle \frac{2\pi }{\mathrm{}}}{\displaystyle \underset{s}{}}|s|U|\eta |^2\delta (\mathrm{}\omega _sE_\eta ).`$ (82) We ascertained that all three calculation methods led to the same results for $`\gamma _r`$. In Eq. (61) $`|\eta ,|s`$ are the secondary quantized wave functions. The initial state wave function $`|\eta `$ describes the EHP with the index set $`\eta `$; the final state $`|s`$ describes the photon with the index set $`s`$. The interaction $`U`$ is written as $$U=\frac{1}{c}𝑑\mathrm{𝐫𝐀}(𝐫)𝐣(𝐫),$$ (83) $$𝐀(𝐫)=\left(\frac{2\pi \mathrm{}}{V_0}\right)^{1/2}\frac{c}{n}\underset{s}{}\omega _s^{1/2}(c_s𝐞_se^{i\kappa 𝐫}+c_s^+𝐞_s^{}e^{i\kappa r}).$$ (84) The index set $`s`$ involves the wave vector $`\kappa `$ and the polarization index $`i`$ (accepting two values), $`c_s^+(c_s)`$ is the photon creation (annihilation) operator , $`𝐞_s`$ and $`\mathrm{}\omega _s`$ are the polarization vector and energy, respectively; $`V_0`$ is the normalization volume. The charge current density operator is given by $$𝐣(𝐫)=\frac{e}{m_0}\underset{\xi }{}[𝐩_{cv}F_\xi ^{}(𝐫)a_\xi ^++𝐩_{cv}^{}F_\xi (𝐫)a_\xi ],$$ (85) where $`m_0`$ is the bare electron mass, $`a_\xi ^+(a_\xi )`$ is the EHP creation (annihilation) operator, $`\xi `$ is the index set, $$F_\xi (𝐫)=\mathrm{\Psi }_\xi (𝐫_e=𝐫_h=𝐫),$$ (86) $`\mathrm{\Psi }_\xi (𝐫_e,𝐫_h)`$ is the smooth part of the EHP wave function (in the effective mass approximation), depending on the radius vectors $`𝐫_e,𝐫_h`$ of the electron and hole, respectively, $`𝐩_{cv}`$ is the interband matrix element of the momentum operator. Let us stress that the index set $`\xi `$ includes the indexes $`c`$ and $`v`$ of the conductivity and valence bands, which can be degenerate ones. Let us determine the wave functions $`\mathrm{\Psi }_\xi (𝐫_e,𝐫_h)`$. That is well known that the electron (hole) wave functions, corresponding to the gauge $`𝐀=𝐀(0,xH,0)`$ of the vector-potential, are $$\psi _{c(v),n,k_y,l}^{e(h)}(𝐫)=\varphi _n(x\pm a_H^2k_y)\frac{1}{\sqrt{L_y}}e^{ik_yy}\phi _{c(v)l}(z).$$ (87) The upper sign ”+” corresponds to the electron, $`a_H=\sqrt{c\mathrm{}/|e|H}`$, $`L_y`$ is the normalization length, $`\varphi _n(x)={\displaystyle \frac{1}{\sqrt{a_H}}}f_n(x/a_H),`$ (88) $`f_n(t)={\displaystyle \frac{1}{\sqrt{\pi ^{1/2}2^nn!}}}e^{t^2/2}H_n(t),`$ (89) $`H_n(t)`$ is the Hermitian polynomial. The real functions $`\phi _{c(v)l}(z)`$, corresponding to the size-quantization quantum number $`l`$ for the QW of the finite depth, can be found, for instance, in Ref. . At first we introduce an EHP wave function set consisting of a product of the wave functions Eq. (66) for electrons and holes and characterized by the index set $$\zeta c,v,n_e,n_h,k_{ey},k_{hy},l_e,l_h.$$ (90) $$\mathrm{\Psi }_\zeta (𝐫_e,𝐫_h)=\psi _{c,n_e,k_{ey},l_e}^e(𝐫_e)\psi _{v,n_h,k_{hy},l_h}^h(𝐫_h).$$ (91) However one cannot calculate the EHP radiative lifetime with the help of Eq. (69). One has to introduce the wave functions for a state of the EHP created by light from the ground state of the system in a QW. Therefore we go to the wave functions $`\mathrm{\Psi }_\xi (𝐫_e,𝐫_h)`$ $`=`$ $`\sqrt{{\displaystyle \frac{2\pi a_H^2}{L_xL_y}}}{\displaystyle \underset{k_{ey},k_{hy}}{}}\delta _{K_y,k_{ey}+k_{hy}}e^{ia_H^2K_xk_y}\times `$ (92) $`\times `$ $`\mathrm{\Psi }_\zeta (𝐫_e,𝐫_h),`$ (93) where $$k_y=\frac{k_{ey}m_hk_{hy}m_e}{M},M=m_e+m_h,$$ $`m_e(m_h)`$ is the electron (hole) effective mass. The functions of type of Eq. (70) have been introduced in Ref. , see also Refs. . The index set is $$\xi c,v,n_e,n_h,𝐊_{},l_e,l_h.$$ (94) It has been shown in Ref. that the functions Eq. (70) are the eigenfunctions of the EHP momentum operator $`\widehat{𝐏}_{}`$ in the $`xy`$ plane, corresponding to the eigenvalues $`\mathrm{}𝐊_{}`$. The operator $`\widehat{𝐏}_{}`$ is determined in Ref. . With the help of the coordinates $$𝐫_{}=𝐫_e_{}𝐫_h,𝐑_{}=\frac{m_h𝐫_h+m_e𝐫_e}{M}$$ of the relative and whole movement of the electron and hole in the $`xy`$ plane and summarizing on $`k_{ey},k_{hy}`$ in the RHS of Eq. (70), one obtains $`\mathrm{\Psi }`$ $`{}_{\xi }{}^{}(𝐫_e,𝐫_h)={\displaystyle \frac{1}{\sqrt{2\pi a_H^2L_xL_y}}}\times `$ (95) $`\times `$ $`\mathrm{exp}[i(K_xX+K_yY)i{\displaystyle \frac{yX}{a_H^2}}]\times `$ (96) $`\times `$ $`\mathrm{exp}[i{\displaystyle \frac{m_em_h}{2M}}(𝐊_{}𝐫_{}+a_H^2K_xK_y)+{\displaystyle \frac{xy}{a_H^2}}]\times `$ (97) $`\times `$ $`𝒦_{n_e,n_h}\left({\displaystyle \frac{𝐫_{}𝐫_0}{a_H}}\right)\phi _{cl_e}(z_e)\phi _{vl_h}(z_h),`$ (98) where $$𝐫_0=a_H^2\frac{𝐇\times 𝐊_{}}{H}.$$ The function $`𝒦_{n,m}(𝐩)`$ is determined by $`𝒦_{n,m}(𝐩)`$ $`=`$ $`\left[{\displaystyle \frac{min(n!,m!)}{max(n!,m!)}}\right]^{1/2}i^{|nm|}e^{p^2/4}\times `$ (99) $`\times `$ $`\left({\displaystyle \frac{p}{\sqrt{2}}}\right)^{|nm|}\mathrm{exp}[i(\varphi \pi /2)(nm)]\times `$ (100) $`\times `$ $`L_{min(n,m)}^{|nm|}(p^2/2),`$ (101) where $$p=\sqrt{p_x^2+p_y^2},\varphi =\mathrm{arctan}(p_y/p_x),$$ $`L_m^n(x)`$ is the Laguerre polynomial. The energy eigenvalues of the states Eq. (72) are $`E_\xi =\mathrm{}\omega _\xi `$ $`=`$ $`E_g+\epsilon _{le}^e+\epsilon _{lh}^h+`$ (102) $`+`$ $`\mathrm{}\mathrm{\Omega }_e(n_e+1/2)+\mathrm{}\mathrm{\Omega }_h(n_h+1/2),`$ (103) where $`E_g`$ is the energy gap. With the help of Eq. (72) one obtains $`F_\xi (𝐫)`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2\pi a_H^2L_xL_y}}}\mathrm{exp}\left(i{\displaystyle \frac{m_em_h}{2M}}a_H^2K_xK_y\right)\times `$ (104) $`\times `$ $`𝒦_{n_e,n_h}\left({\displaystyle \frac{𝐫_0}{a_H}}\right)e^{i𝐊_{}𝐫_{}}\phi _{cl_e}(z)\phi _{vl_h}(z).`$ (105) Let us calculate $`\gamma _{r\xi }`$ for the index set $`\xi `$. With the help of Eqs. (61) - (64) one obtains $`\gamma _{r\xi }`$ $`=`$ $`{\displaystyle \frac{(2\pi e)^2}{V_0(nm_0)^2}}{\displaystyle \underset{s}{}}{\displaystyle \frac{|𝐞_s𝐩_{cv}|^2}{\omega _s}}|{\displaystyle }d𝐫e^{i\kappa 𝐫}F_\xi (𝐫)|^2\times `$ (106) $`\times `$ $`\delta (\mathrm{}\omega _sE_\xi ).`$ (107) The substitution of Eq. (75) in Eq. (76) results in $`\gamma _{r\xi }`$ $`=`$ $`\left({\displaystyle \frac{e}{a_Hnm_0}}\right)^2{\displaystyle \frac{1}{\mathrm{}\omega _\xi }}B_{n_e,n_h}(K_{}){\displaystyle \frac{2\pi }{L_z}}{\displaystyle \underset{i,\kappa }{}}\delta _{\kappa _{},𝐊_{}}\times `$ (108) $`\times `$ $`|𝐞_s𝐩_{cv}|^2|R_{l_e,l_h}(\kappa _z)|^2\delta (\omega _s\omega _\xi ),`$ (109) where $`B_{n,m}(K_{})`$ $`=`$ $`|𝒦_{n,m}({\displaystyle \frac{𝐫_0}{a_H}})|^2={\displaystyle \frac{min(n!,m!)}{max(n!,m!)}}\times `$ (110) $`\times `$ $`\mathrm{exp}({\displaystyle \frac{a_H^2K_{}^2}{2}})\left({\displaystyle \frac{a_H^2K_{}^2}{2}}\right)^{|nm|}\times `$ (111) $`\times `$ $`\left[L_{min(n,m)}^{|nm|}\left({\displaystyle \frac{a_H^2K_{}^2}{2}}\right)\right]^2,`$ (112) $$R_{l_e,l_h}(k)=_{\mathrm{}}^{\mathrm{}}𝑑ze^{ikz}\phi _{cl_e}(z)\phi _{vl_h}(z).$$ (113) Summing up in the RHS of Eq. (77) on $`\kappa _{}`$ and taking into account that $`\omega _s=c\kappa /n,\kappa =\sqrt{\kappa _{}^2+\kappa _z^2}`$, one obtains $`\gamma _{r\xi }`$ $`=`$ $`{\displaystyle \frac{2\pi e^2}{L_zm_0^2\mathrm{}\omega _\xi a_H^2cn}}\times `$ (114) $`\times `$ $`B_{n_e,n_h}(K_{})|R_{l_e,l_h}(\sqrt{(\omega _\xi n/c)^2K_{}^2})|^2\times `$ (115) $`\times `$ $`{\displaystyle \underset{i,\kappa _z}{}}|𝐞_s𝐩_{cv}|^2\delta \left(\sqrt{K_{}^2+\kappa _z^2}\omega _\xi n/c\right).`$ (116) Summing up on $`\kappa _z`$ at $`K_{}<\omega _\xi n/c`$, one obtains $`\gamma _{r\xi }`$ $`=`$ $`{\displaystyle \frac{e^2\mathrm{\Omega }_0}{\mathrm{}^2c^2m_0}}B_{n_e,n_h}(K_{})\times `$ (117) $`\times `$ $`|R_{l_e,l_h}(\sqrt{(\omega _\xi n/c)^2K_{}^2})|^2{\displaystyle \frac{1}{\sqrt{(\omega _\xi n/c)^2K_{}^2}}}\times `$ (118) $`\times `$ $`{\displaystyle \underset{i}{}}[|𝐞_i(\kappa ^+)𝐩_{cv}|^2+|𝐞_i(\kappa ^{})𝐩_{cv}|^2];`$ (119) $`\gamma _{r\xi }`$ $`=`$ $`0,K_{}>\omega _\xi n/c,`$ (120) where $`\mathrm{\Omega }_0=|e|H/m_0c`$, $`\kappa ^+(\kappa ^{})`$ is the photon wave vector with the component $`𝐊_{}`$ in the $`xy`$ plane and with the z-component equal to $`\pm \sqrt{(\omega _\xi n/c)^2K_{}^2}`$. Thus the EHP with the energy $`\mathrm{}\omega _\xi `$ and the wave vector $`𝐊_{}`$ in the $`xy`$ plane emits the photons with the energies $`\mathrm{}\omega _\xi `$ and wave vectors $`\kappa ^+`$ and $`\kappa ^{}`$ if $`K_{}<\omega _\xi n/c`$. Let us direct the $`x`$ axis along $`𝐊_{}`$. Summarizing on the photon polarizations at $`K_{}<\omega _\xi n/c`$ one obtains $`\gamma _{r\xi }`$ $`=`$ $`{\displaystyle \frac{2e^2\mathrm{\Omega }_0}{\mathrm{}^2cnm_0\omega _\xi }}B_{n_e,n_h}(K_{})\times `$ (121) $`\times `$ $`|R_{l_e,l_h}(\sqrt{(\omega _\xi n/c)^2K_{}^2})|^2\times `$ (122) $`\times `$ $`[{\displaystyle \frac{\sqrt{(\omega _\xi n/c)^2K_{}^2}}{\omega _\xi n/c}}|p_{cvx}|^2+`$ (123) $`+`$ $`{\displaystyle \frac{\omega _\xi n/c}{\sqrt{(\omega _\xi n/c)^2K_{}^2}}}|p_{cvy}|^2+`$ (124) $`+`$ $`{\displaystyle \frac{K_{}^2|p_{cvz}|^2}{(\omega _\xi n/c)\sqrt{(\omega _\xi n/c)^2K_{}^2}}}].`$ (125) Eq. (82) can be compared with the result from Ref. regarding to the exciton energy level in a QW at $`H=0`$. It turns out that the dependencies from the components $`p_{cvx},p_{cvy}`$ and $`p_{cvz}`$ coincide and the result $`\gamma _{r\xi }=0`$ at $`K_{}>\omega _\xi n/c`$ coincides also. The RHS of Eq. (82) contains the factors $`\mathrm{\Omega }_0`$ and $`B_{n_e,n_h}(K_{})`$ due to magnetic field. The factor $`\left|R_{l_e,l_h}(\sqrt{(\omega _\xi n/c)^2K_{}^2})\right|^2`$ is absent in Ref. because the QWs with $`d<<c/n\omega _l`$ have been considered. We did not apply such a restriction obtaining Eq. (82). For the narrow QWs with $`d<<\lambda `$ one has $`R_{l_e,l_h}\left(\sqrt{(\omega _\xi n/c)^2K_{}^2}\right)`$ (126) $`\pi _{l_e,l_h}={\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑z\phi _{cle}(z)\phi _{vlh}(z),`$ (127) and for infinitely deep QWs $$\pi _{l_e,l_h}=\delta _{l_e,l_h}.$$ (128) For QWs with $`dc/n\omega _l`$ the quantity $`\left|R_{l_e,l_h}(\sqrt{(\omega _\xi n/c)^2K_{}^2})\right|^2`$ depends on $`d`$ and diminishes with growth $`d`$, therefore the quantity $`\gamma _{r\xi }`$ diminishes also. In the limit $`d>>\lambda `$ the quantity $`\gamma _{r\xi }0`$ as in a bulk crystal. We are interested in obtaining the quantities Eq. (82) at $`K_{}=0`$ because we calculate the absorption and reflection coefficients for the light normal incidence. Because $`B_{n_e,n_h}(0)=\delta _{n_e,n_h}`$, let us calculate $`\gamma _{r\xi _0}`$ for the index set $$\xi _0c,v,n_e=n_h,𝐊_{}=0,l_e,l_h.$$ (129) With the help of Eq. (82) one obtains $$\gamma _{r\xi _0}=\frac{2e^2\mathrm{\Omega }_0}{\mathrm{}cn}\left|R_{l_e,l_h}(\omega _{\xi _0}n/c)\right|^2\frac{|p_{cvx}|^2+|p_{cvy}|^2}{m_0\mathrm{}\omega _{\xi _0}},$$ (130) where $$\mathrm{}\omega _{\xi _0}=E_g+\epsilon _{le}^c+\epsilon _{lh}^h+\mathrm{}\mathrm{\Omega }_\mu (n_e+1/2),$$ (131) $$\mathrm{\Omega }_\mu =\frac{|e|H}{\mu c},\mu =\frac{m_em_h}{M}.$$ It follows from Eq. (86) that the EHP inverse radiative lifetime is proportional to magnetic field $`H`$ if the energy $`\mathrm{}\omega _{\xi _0}`$ weakly depends on $`H`$ (approximately $`E_{\xi _0}\stackrel{~}{E}_g,\stackrel{~}{E}_g=E_g+\epsilon _{le}^e+\epsilon _{lh}^h)`$. Let us calculate $`\gamma _{r\xi _0}`$ for the band model of $`GaAs`$. The conductivity band is twice degenerated (on the spin) and the index $`c`$ takes two values : $`c=1`$ or $`c=2`$. The valence band (for heavy holes) is twice degenerate also: $`v=1`$ or $`v=2`$. Two EHP sorts are possible: First, with the indexes $`c=1,v=1`$, second, with indexes $`c=2,v=2`$. These EHP sorts I and II differ by values $`𝐩_{cv}`$, which are $$𝐩_{cv}^I=\frac{1}{\sqrt{2}}p_{cv}(𝐞_xi𝐞_y),𝐩_{cv}^{II}=\frac{1}{\sqrt{2}}p_{cv}(𝐞_x+i𝐞_y).$$ (132) For the proposed model the wave functions $`\phi _{cle}(z)`$ and $`\phi _{vlh}(z)`$ do not depend on indexes $`c`$ and $`v`$. When the circular polarizations are used, every polarization (left or right relatively of the axis $`z`$) is toughly linked with the EHP sort (I ore II) because the EHP – light interaction is proportional to $`𝐞_i𝐩_{cv}`$. For the model Eq. (88) one obtains from Eq. (86) the result for any sort The corresponding formula from Ref. coincides with Eq. (89) at $`n=1`$. $$\gamma _{r\xi _0}=\frac{2e^2\mathrm{\Omega }_0}{\mathrm{}cn}\left|R_{l_e,l_h}(\omega _\xi n/c)\right|^2\frac{p_{cv}^2}{m_0\mathrm{}\omega _{\xi _0}}.$$ (133) Applying the approximation (83) for the narrow QWs with $`d<<\lambda `$, we obtain $$\gamma _{r\xi _0}\frac{2e^2\mathrm{\Omega }_0}{\mathrm{}cn}\pi _{l_e,l_h}^2.\frac{p_{cv}^2}{m_0\mathrm{}\omega _{\xi _0}}.$$ (134) Applying the parameters for GaAs from Ref. and the approximation of Eq. (84) and supposing $`\mathrm{}\omega _{\xi _0}E_g,`$ we obtain the numerical estimation $$\mathrm{}\gamma _r5.3510^5\frac{H}{H_{res}}eV,$$ (135) where $`H_{res}`$ appropriates to the magnetopolaron $`A`$ (Fig. 1), i. e. equals to $`H_{res}=m_ec\omega _{LO}/|e|`$. ## VI Inverse radiative lifetimes of magnetopolaron. Let us consider the polaron $`A`$ (Fig. 1) in a combination with the hole characterized by the size-quantization quantum number $`l_h`$ and the Landau quantum number $`n_h=1`$. Close to the resonance at $`\omega _{LO}=\mathrm{\Omega }_e`$ the EHP energy splits into two magnetopolaron branches with the inverse radiative lifetimes $`\gamma _{ra}`$ and $`\gamma _{rb}`$, according to designations of Ref. . The index $`a(b)`$ corresponds to the upper (lower) magnetopolaron term. To calculate $`\gamma _{ra}`$ and $`\gamma _{rb}`$ we will use Eq. (76), determining the function $`F_\xi (𝐫)`$ from the RHS of Eq. (76). At first we consider the EHP wave functions under conditions of the magnetophonon resonance <sup>§</sup><sup>§</sup>§Here and below the indexes $`c`$ and $`v`$ are omitted. $$\mathrm{\Psi }_\mathrm{\Pi }(𝐫_e,𝐫_h)|0=\psi _{1,k_{hy},l_h}^h(𝐫_h)\mathrm{\Theta }_{p,k_{ey},l_e}(𝐫_e)|0.$$ (136) The hole wave function $`\psi _{1,k_{hy},l}^h(𝐫)`$ is defined in Eq. (66). The polaron wave function $`\mathrm{\Theta }_{p,k_y,l}(𝐫_e)|0`$ is calculated in Ref. . The index set in Eq. (92) is $$\mathrm{\Pi }c,v,p;k_{ey},k_{hy},l_e,l_h.$$ The index $`p`$ takes two values : $`a`$ or $`b`$. The function $`|0`$ appropriates to the phonon vacuum . The operator $`\mathrm{\Theta }_{p,k_y,l}(𝐫_e)`$, according to Ref. , can be written as $$\mathrm{\Theta }_{p,k_y,l}(𝐫)=\mathrm{\Theta }_{p,k_y,l}^0(𝐫)+\mathrm{\Theta }_{p,k_y,l}^1(𝐫),$$ (137) where $$\mathrm{\Theta }_{p,k_y,l}^0(𝐫)=Q_{0p}^{1/2}\psi _{1,k_y,l}^e(𝐫),$$ (138) $$\mathrm{\Theta }_{p,k_y,l}^1(𝐫)=\frac{Q_{1p}^{1/2}}{|A|}\underset{\nu }{}e^{ia_Hq_x(k_yq_y/2)}U^{}(\nu )\psi _{0,k_yq_y}^e(𝐫)b_\nu ^+.$$ (139) In the RHS of Eq. (94) and Eq. (95) $`\psi _{n,k_y,l}^e(𝐫)`$ are the electron wave functions determined in Eq. (66); $`\nu `$ is the phonon index set, including the transverse component $`𝐪_{}`$ ; $`b_\nu ^+`$ is the phonon creation operator; $`U(\nu )`$ is the quantity proportional to electron-phonon interaction and determined in Ref. ; $`A^2=_\nu |U(\nu )|^2`$. The polaron wave function $`\mathrm{\Theta }_{p,k_y,l}(𝐫)|0`$ and the corresponding energy eigenvalue (measured from the energy level $`\epsilon _{le}^e`$) $`E_p=\mathrm{}\mathrm{\Omega }_e+{\displaystyle \frac{1}{2}}\mathrm{}\omega _{LO}\pm \sqrt{(\lambda /2)^2+A^2},`$ (140) $`\lambda =\mathrm{}(\mathrm{\Omega }_e\omega _{LO})`$ (141) have been calculated in Ref. , where it was assumed that all the LO phonons, composing the polaron, had the same non-dispersal frequency $`\omega _{LO}`$. The coefficients in Eq. (94) and Eq. (95) are $`Q_{0p}={\displaystyle \frac{1}{2}}\left(1\pm {\displaystyle \frac{\lambda }{\sqrt{\lambda ^2+4A^2}}}\right),`$ (142) $`Q_{1p}={\displaystyle \frac{1}{2}}\left(1{\displaystyle \frac{\lambda }{\sqrt{\lambda ^2+4A^2}}}\right),`$ (143) where the upper (lower) sign refers to the term $`p=a(p=b)`$. It follows from Eq. (93) that the polaron wave function is represented by the linear combination of two functions of the electron-phonon system, one of which corresponds to the electron with the Landau quantum number $`n=1`$ and to the phonon vacuum, and another corresponds to the electron with $`n=0`$ and the LO phonon. The coefficients $`Q_{0p}`$ and $`Q_{1p}`$ are the probabilities of finding the system in these states designated with indexes 0 and 1, respectively. In the resonance $`\mathrm{\Omega }_e=\omega _{LO}`$ the coefficients $`Q_{0p}^{res}=Q_{1p}^{res}=1/2`$ and the energy $`E_p^{res}={\displaystyle \frac{3}{2}}\mathrm{}\omega _{LO}\pm \sqrt{A^2},`$ (144) $`\mathrm{\Delta }E=E_a^{res}E_b^{res}=2\sqrt{A^2}.`$ (145) To calculate the inverse radiative lifetimes $`\gamma _{ra}`$ and $`\gamma _{rb}`$ it is necessary to introduce the eigenfunctions of the EHP momentum operator $`\widehat{𝐏}`$ . Analogically to Eq. (70) we introduce the functions $`\mathrm{\Psi }_\eta (𝐫_e,𝐫_h)|0=\sqrt{{\displaystyle \frac{2\pi a_H^2}{L_xL_y}}}\times `$ (146) $`\times {\displaystyle \underset{k_{ey},k_{hy}}{}}\delta _{K_y,k_{ey}+k_{hy}}e^{ia_H^2K_xk_y}\mathrm{\Psi }_\mathrm{\Pi }(𝐫_e,𝐫_h)|0,`$ (147) characterized by the index set $$\eta c,v,p;𝐊_{},l_e,l_h.$$ (148) Determining $`\gamma _{r\eta }`$ according to Eq. (76), we substitute $$F_\eta (𝐫)=0|\mathrm{\Psi }_\eta (𝐫_e=𝐫_h=𝐫)|0,$$ (149) instead of $`F_\xi (𝐫)`$ and the energy $$E_\eta =\mathrm{}\omega _\eta =E_g+\epsilon _{le}^e+\epsilon _{lh}^h+\frac{3}{2}\mathrm{}\mathrm{\Omega }_h+E_p.$$ (150) instead of the energy $`E_\xi `$. So long as $`0|b_\nu ^+|0=0`$ only the first term from the RHS of Eq. (93) contributes into the RHS of Eq. (101). And for $`F_\eta (𝐫)`$ we obtain the expression distinguishing from Eq. (75) at $`n_e=n_h=1`$ only by the additional factor $`Q_{0p}^{1/2}`$. If the $`x`$ axis is directed along $`𝐊_{}`$, the result for the inverse radiative lifetimes $`\gamma _{r\eta }`$ is analogical to Eq. (82) $`\gamma _{r\eta }`$ $`=`$ $`{\displaystyle \frac{2Q_{0p}e^2\mathrm{\Omega }_0}{\mathrm{}^2cnm_0\omega _\eta }}B_{11}(K_{})\times `$ (151) $`\times `$ $`|R_{l_e,l_h}\left(\sqrt{(\omega _\eta n/c)^2K_{}^2}\right)|^2\times `$ (152) $`\times `$ $`[{\displaystyle \frac{\sqrt{(\omega _\eta n/c)^2K_{}^2}}{\omega _\eta n/c}}|p_{cvx}|^2+`$ (153) $`+`$ $`{\displaystyle \frac{\omega _\eta n/c}{\sqrt{(\omega _\eta n/c)^2K_{}^2}}}|p_{cvy}|^2+`$ (154) $`+`$ $`{\displaystyle \frac{K_{}^2}{(\omega _\eta n/c)\sqrt{(\omega _\eta n/c)^2K_{}^2}}}|p_{cvz}|^2];K_{}<\omega _\eta n/c;`$ (155) $`\gamma _{r\eta }`$ $`=`$ $`0;K_{}>\omega _\eta n/c.`$ (156) For the case $`𝐊_{}=0`$ we introduce the index set $$\eta _0c,v,p;𝐊_{}=0,l_e,l_h.$$ (157) Then for the quantities $`\gamma _{r\eta _0}`$ we obtain the formulae distinguishing from Eq. (86), Eq. (89) and Eq. (90) only by the substitution the indexes $`\eta `$ instead of $`\xi _0`$ (as long as $`\omega _{\eta _0}=\omega _\eta `$) and by the factor $`Q_{0p}`$. Instead of Eq. (90) we obtain $$\gamma _{r\eta _0}\frac{2Q_{0p}e^2}{\mathrm{}c}\frac{\mathrm{\Omega }_0p_{cv}^2}{nm_0\mathrm{}\omega _\eta }\pi _{l_e,l_h}^2.$$ (158) The substitution $`\omega _\eta `$ instead of $`\omega _{\xi _0}`$ is not essential as long as approximately $`\omega _{\xi _0}\omega _\eta \stackrel{~}{E}_g/\mathrm{}`$. The tough dependence of the inverse radiative lifetime from the index $`p`$ and from the value $`\lambda `$ (characterizing the deviation from the resonance) is determined by the factor $`Q_{0p}`$. In the resonance $`Q_{0p}=1/2`$, and the quantities $`\gamma _{ra}`$ and $`\gamma _{rb}`$ are equal and contain the factor 1/2 in comparison with $`\gamma _r`$ for the appropriate single level (without phonons) coming through the crossing point of energy terms (see Fig.1). ## VII Non-radiative lifetimes of magnetopolaron We will not calculate the EHP inverse non-radiative lifetimes in a QW in a magnetic field far away from the magnetopolaron resonance as long as it is not clear what processes determine them. One-phonon transitions are forbidden by the energy conservation law. Perhaps the main contribution is determined by the two-phonon processes with acoustic phonons participation. But in the magnetophonon resonance vicinity some contribution into $`\gamma _p(p=a,b)`$, appears due to the finite lifetimes of LO phonons composing the magnetopolaron. We will calculate this contribution here and determine the low limit of $`\gamma _p(p=a,b)`$ in the resonance vicinity. The phonon $`\nu `$ in a QW is characterized by the inverse non-radiative lifetime $`\gamma _\nu `$ due to phonon-phonon interaction, for instance, by the LO phonon decay into two acoustic phonons. The quantity $`\gamma _\nu `$ may be written as $$\gamma _\nu =\frac{2\pi }{\mathrm{}}\underset{f}{}|f|_{phph}|i|^2\delta (E_iE_f),$$ (159) where $`_{phph}`$ is the phonon-phonon interaction, $`|i=b_\nu ^+|0`$ is the initial state of the phonon $`\nu `$ with the energy $`E_i=\mathrm{}\omega _{LO},|f=a_\tau ^+|0`$ is the phonon system final state with the set of indexes $`\tau `$ and the energy $`E_f=E_\tau ,a_\tau ^+`$ is the phonon operator corresponding, for instance, to the creation of two acoustic phonons. Taking into account that the component $`𝐪_{}`$ in a QW is preserved in any transitions, we can write the matrix element in the RHS of Eq. (106) as $$0|a_\tau _{phph}b_\nu ^+|0=\delta _{𝐪_{},𝐪_{}^{}}V(𝐪_{},j,\vartheta ),$$ (160) supposing that the initial state index set $`\nu `$ is $$\nu 𝐪_{},j,$$ (161) and the final state index set $`\tau `$ is $$\tau 𝐪_{}^{},\vartheta ,$$ (162) where $`𝐪_{}^{}`$ is the resultant transverse component of the wave vector in the final state, for example, the sum of the corresponding wave vectors of two phonons; $`j(\vartheta )`$ is the initial (final) state index set. Substituting Eq. (107) into Eq. (106) and summarizing on $`𝐪_{}`$, we obtain $$\gamma _\nu =\frac{2\pi }{\mathrm{}}\underset{\vartheta }{}|V(𝐪_{},j,\vartheta )|^2\delta (\mathrm{}\omega _{LO}E_{𝐪_{},\vartheta }).$$ (163) Applying Eq. (106) we can now calculate the magnetopolaron inverse non-radiative lifetime $`\gamma _{p,k_y,l}`$ due to phonon-phonon interaction. The initial state wave function is written as $`\mathrm{\Theta }_{p,k_y,l}(𝐫)|0`$, where $`|0`$ is a phonon vacuum , the operator $`\mathrm{\Theta }_{p,k_y,l}(𝐫)`$ is determined in Eqs. (93) — (95). The final state wave function is the product of the electron and phonon wave functions $$|f=\psi _{n^{},k_y^{},l^{}}^e(𝐫)a_\tau ^+|0,$$ (164) where $`\psi _{n,k_y,l}^e(𝐫)`$ is determined in Eq. (66). Only the second term from the RHS of Eq. (93), describing the one LO phonon state, contributes into the matrix element $`f|_{phph}|i`$. The first term does not contribute because $`0|a_\tau |0=0`$. One obtains $`f|_{phph}|i`$ $`=`$ $`Q_{1p}^{1/2}{\displaystyle \underset{\nu }{}}e^{ia_H^2q_x(k_yq_y/2)}{\displaystyle \frac{U^{}(\nu )}{|A|}}\times `$ (165) $`\times `$ $`{\displaystyle }d𝐫\psi _{n^{},k_y^{},l^{}}^e(𝐫)\psi _{0,k_yq_y,l}^e(𝐫)\times `$ (166) $`\times `$ $`0|a_\tau _{phph}b_\nu ^+|0.`$ (167) The initial (final) state is characterized by the indexes $`p,k_y,l`$ ($`n^{},k_y^{},l^{},𝐪_{}^{},\vartheta `$). The index set $`\nu `$ in Eq. (112) consists of $`𝐪_{}`$ and $`j`$. Taking into account the orthonormalization of the electron wave functions $`\psi _{n,k_y,l}^e`$ and Eq. (107) and summarizing in Eq. (112) on $`𝐪_{}`$, one obtains $`f|_{phph}|i`$ $`=`$ $`{\displaystyle \frac{Q_{1p}^{1/2}}{|A|}}\delta _{n^{},0}\delta _{l^{},l}\delta _{k_y^{},k_yq_y^{}}e^{ia_H^2q_x^{}(k_x^{}+q_y^{}/2)}\times `$ (168) $`\times `$ $`{\displaystyle \underset{j}{}}U^{}(𝐪_{}^{},j)V(𝐪_{}^{},j,\vartheta ).`$ (169) Substituting Eq. (113) into Eq. (106) and summarizing on indexes $`n^{},k_y^{},l^{}`$ of the final state we obtain $`\gamma _p`$ $`=`$ $`{\displaystyle \frac{2\pi }{\mathrm{}}}{\displaystyle \frac{Q_{1p}}{A^2}}{\displaystyle \underset{𝐪_{},\vartheta }{}}{\displaystyle \underset{j_1,j_2}{}}U^{}(𝐪_{},j_1)U(𝐪_{},j_2)V(𝐪_{},j_1,\vartheta )\times `$ (170) $`\times `$ $`V^{}(𝐪_{},j_2,\vartheta )\delta (E_p\mathrm{}\mathrm{\Omega }_e/2E_{𝐪_{},\vartheta }).`$ (171) Eq. (114) is applicable for the model used in Ref. , where it was assumed that the confined and interface phonons, having the same non-dispersal frequency $`\omega _{LO}`$, take part in the polaron creation. The index set $`j`$ contains the phonon sort indexes. It has been shown earlier that interaction with the confined phonons in quite wide QWs may be approximately replaced by the Fröhlih interaction with the bulk LO phonons and that interaction with the interface phonons does not contribute essentially into polaron energy splitting. In the case of interaction with the bulk phonons the index $`j`$ coincide with $`q_z`$, and Eq. (114) is simplified. One obtains instead of Eq. (107) $$0|a_\tau _{phph}b_\nu ^+|0=\delta _{𝐪,𝐪^{}}W(𝐪,\phi ).$$ (172) It is supposed that the phonon final state is characterized by the 3D wave vector $`𝐪`$ and by other indexes $`\phi `$. With the help of Eq. (115) Eq. (114) is transformed into $`\gamma _p`$ $`=`$ $`{\displaystyle \frac{2\pi }{\mathrm{}}}{\displaystyle \frac{Q_{1p}}{A^2}}{\displaystyle \underset{𝐪,\phi }{}}|U(𝐪)|^2|W(𝐪,\phi )|^2\times `$ (173) $`\times `$ $`\delta (E_p\mathrm{}\mathrm{\Omega }_e/2E_{𝐪,\phi }).`$ (174) Eq. (110) for the inverse LO phonon lifetime (if Eq. (115) is satisfied) can be written as $$\gamma _𝐪=\frac{2\pi }{\mathrm{}}\underset{\phi }{}|W(𝐪,\phi )|^2\delta (\mathrm{}\omega _{LO}E_{𝐪,\phi }),$$ (175) where $`E_{𝐪,\phi }`$ is the final state energy. Applying Eq. (96) for the polaron term energy $`E_p`$ we obtain instead of Eq. (116) $`\gamma _p`$ $`=`$ $`{\displaystyle \frac{2\pi }{\mathrm{}}}{\displaystyle \frac{Q_{1p}}{A^2}}{\displaystyle \underset{𝐪,\phi }{}}|U(𝐪)|^2|W(𝐪,\phi )|^2\times `$ (176) $`\times `$ $`\delta \left(\lambda /2\pm \sqrt{(\lambda /2)^2+A^2}+\mathrm{}\omega _{LO}E_{𝐪,\phi }\right),`$ (177) where the upper (lower) sign corresponds to the term $`a(b)`$. Comparing Eq. (117) and Eq. (118) one finds that in the polaron resonance vicinity, i. e. under condition $$\left|\lambda /2\pm \sqrt{(\lambda /2)^2+A^2}\right|<<\mathrm{}\omega _{LO},$$ (178) the quantity $`\gamma _p`$ is expressed through $`\gamma _𝐪`$, i. e. $$\gamma _p=Q_{1p}\overline{\gamma }_{LO},$$ (179) $$\overline{\gamma }_{LO}=\frac{\underset{𝐪}{}\gamma _𝐪|U(𝐪)|^2}{_𝐪|U(𝐪)|^2}.$$ (180) Due to the factor $`Q_{1p}`$ in Eq. (120) the quantity $`\gamma _p`$ toughly depends on the parameter $`\lambda =\mathrm{}(\mathrm{\Omega }_e\omega _{LO})`$. In the resonance $$\gamma _p=\frac{1}{2}\overline{\gamma }_{LO}$$ (181) for the both terms $`p=a`$ and $`p=b`$. One can suppose that the result from Eq. (114) is comparable with Eq. (120). Let us consider the results of Eq. (118) far away from the resonance $`\mathrm{\Omega }_e=\omega _{LO}`$. It is seen on Fig. 1 that the term $`a`$ at $`\lambda >0`$ and the term $`b`$ at $`\lambda <0`$ transits under condition $`|\lambda |>>|A|`$ into the electron level with indexes $`n=1,l`$. According to Eq. (97) the factor $`Q_{1p}`$ approximately equals to $`Q_{1p}A^2/\lambda ^2`$. With the help of Eq. (118) one obtains approximately $$\gamma _p=\frac{2\pi }{\mathrm{}}\frac{A^2}{\lambda ^2}\underset{𝐪,\phi }{}\frac{|U(𝐪)|^2}{A^2}|W(𝐪,\phi )|^2\delta (\mathrm{}\mathrm{\Omega }_eE_{𝐪,\phi }).$$ (182) This value is not expressed through $`\overline{\gamma }_{LO}`$, but one can see that in comparison with $`\overline{\gamma }_{LO}`$ it contains the small factor $`A^2/\lambda ^2`$. Note that the argument of the $`\delta `$-function in Eq. (123) appropriates to the transition $`n=1n=0`$ with emitting, for instance, of two acoustic phonons. The quantity Eq. (123) is of the second order on phonon-phonon interaction. It is one of the contributions into the value of the EHP inverse non-radiative lifetime far away from the magnetophonon resonance. Two other branches of the terms, $`a`$ at $`\lambda <0`$ and $`b`$ at $`\lambda >0,|\lambda |>>|A|`$ , appropriate to the state with the electron on the level $`n=0,l`$ plus one LO phonon. $`Q_{1p}1`$ on these branches, thus we obtain approximately $$\gamma _p\overline{\gamma }_{LO},$$ (183) as it must be for a state including LO phonon. Thus the formula Eq. (118) gives the right limit transitions at $`|\lambda |>>|A|`$. One can conclude that the quantity $`\gamma _p`$ increases sharply around the electronic terms intersections reaching the half of the LO phonon inverse lifetime. In the resonance $`\mathrm{\Omega }_e=\omega _{LO}`$ the value of the inverse non-radiative lifetime for each of two terms is no smaller than $`\overline{\gamma }_{LO}/2`$. ## VIII Numerical calculation results The reflection $`(\omega _l)`$ and absorption $`𝒜(\omega _l)`$ coefficients for the three-level system and for the different interrelations between $`\gamma _{r1(2)}`$, $`\gamma _{1(2)}`$ and $`\omega _1\omega _2`$ are represented in Figs. 2 - 4. These results have to be used in the case of any two excited levels in a QW when they are situated quite close each other. When $`\gamma _{r1(2)}<<\omega _1\omega _2,\gamma _{1(2)}<<\omega _1\omega _2`$, the results for the two-level system are applicable. Fig. 2 show the dependencies at $`\gamma _{r1}=\gamma _{r2}=\gamma _r`$, $`\gamma _1=\gamma _2=\gamma `$, $`\gamma _r<<\gamma `$. The curves 1 are relevant to the case $`\gamma <\omega _1\omega _2`$, the curves 2 are relevant to the case $`\gamma =\omega _1\omega _2`$ and the curves 3 are relevant to the case $`\gamma >\omega _1\omega _2`$. It is seen that in case 1 the transit stems to the results for two-level systems. The maxima are situated in the vicinity of the points $`\omega _l=\omega _1`$ and $`\omega _l=\omega _2`$. In Fig. 3 the functions $`𝒜(\omega _l)`$ and $`(\omega _l)`$ for the three level systems at $`\gamma _{r1}=\gamma _{r2}=\gamma _r`$, $`\gamma _1=\gamma _2=\gamma `$, $`\gamma _r>>\gamma `$ are represented. The curves 1 are relevant to the case $`\gamma _r<\omega _1\omega _2`$, the curves 2 are relevant to the case $`\gamma _r=\omega _1\omega _2`$ and the curves 3-6 in Fig. 3a and the curves 3 in Fig. 3b are relevant to the case $`\gamma _r>\omega _1\omega _2`$. It is seen that in the case 1 the transit stems to the results for two two-level systems. In other cases the peculiar results can be seen. In particular the quantity $`(\omega _0)`$ approaches 0 on all the curves of Fig. 3b in the point $`\omega _l=\omega _0`$ . The reflection and absorption coefficients are much smaller than one and $`(\omega _l)<<𝒜(\omega _l)`$ as in the case of the two-level systems at $`\gamma _{r1(2)}<<\gamma _{1(2)}`$. On the contrary at $`\gamma _{r1(2)}>>\gamma _{1(2)}`$ $`(\omega _l)>>𝒜(\omega _l)`$ and $``$ reaches one. In Fig. 4 the functions $`𝒜(\omega _l)`$ and $`(\omega _l)`$ for the three-level systems at $`\gamma _{r1}=\gamma _{r2}=\gamma _r`$, $`\gamma _1=\gamma _2=\gamma `$, $`\gamma _r=\gamma `$ are represented. The curves 1 are relevant to the case $`\gamma _r<\omega _1\omega _2`$, the curves 2 are relevant to the case $`\gamma _r=\omega _1\omega _2`$ and the curves 3 are relevant to the case $`\gamma _r>\omega _1\omega _2`$. It is seen that the absorption coefficient reaches the maximum values but does not exceed 0.5. In Fig. 5 the energy terms $`a`$ and $`b`$ for the system consisting of the polaron $`A`$ (Fig. 1) and the hole with indexes $`n=1,l`$ are represented. $`\mathrm{\Delta }E`$ is term’s splitting at $`\lambda =0`$. $`_p`$ is the system energy measured from $$E_0=E_g+\epsilon _l^e+\epsilon _l^h+(3/2)(1+m_e/m_h)\mathrm{}\omega _{LO}.$$ According to Eq. (102) and Eq. (96), $$\frac{_p}{\mathrm{\Delta }E}=\frac{\lambda }{\mathrm{\Delta }E}\left(1+\frac{3}{2}\frac{m_e}{m_h}\right)\pm \frac{1}{2}\sqrt{1+\left(\frac{\lambda }{\mathrm{\Delta }E}\right)^2},$$ (184) where the upper (lower) sign corresponds to the term $`a`$($`b`$). The value $`m_e/m_h=0.2`$ is applied, which is relevant to GaAs. Calculating $`𝒜(\omega _l)`$ and $`(\omega _l)`$ for magnetopolarons in QWs one has to substitute the quantities $`_a`$ and $`_b`$ (see Eq. (125)) into all the formulae of the section IY as energies $`\mathrm{}\omega _1`$ and $`\mathrm{}\omega _2`$ when $`\mathrm{}\omega _l`$ is measured from the level $`E_0`$. In Fig. 6 the dependencies $`\gamma _{rp}`$ and $`\gamma _p/\gamma _{rp}`$ from the magnetic field value are represented. Fig. 9 (10) appropriates to the term $`a`$ ($`b`$). The constant $`\gamma _0`$ is the EHP inverse radiative lifetime at $`\lambda =0`$ without the polaron effect. According to the section Y, $$\gamma _0=\frac{2e^2\omega _{LO}m_e}{\mathrm{}cnm_0}\frac{p_{cv}^2}{m_0\stackrel{~}{E}_g}\pi _{ll}^2.$$ (185) With the help of Eq. (105) we obtain $$\frac{\gamma _{rp}}{\gamma _0}=\frac{1}{2}\left(1+\frac{\lambda /\mathrm{\Delta }E}{\mathrm{\Delta }E/\mathrm{}\omega _{LO}}\right)\left[1\pm \frac{\lambda /\mathrm{\Delta }E}{\sqrt{1+(\lambda /\mathrm{\Delta }E)^2}}\right],$$ (186) which was applied in Fig. 6. The upper (lower) sign is relevant to the term $`p=a`$ ($`b`$). The factor in the parenthesis in the RHS of Eq. (127) describes the magnetic field dependence of $`\mathrm{\Omega }_0`$ entering in $`\gamma _{rp}`$ (see Eq. (105)). This factor depends on $`\mathrm{\Delta }E/\mathrm{}\omega _{LO}`$, which is chosen equal to 0.18. According to Ref. in GaAs (the polaron $`A`$) at $`d=300A`$ energy splitting is $`\mathrm{\Delta }E6.6510^3eV`$. As long as in GaAs at $`\mathrm{}\omega _{LO}=0.0367eV`$ one obtains $`\mathrm{\Delta }E/\mathrm{}\omega _{LO}0.181.`$ According to Eq. (105) and Eq. (120) $$\frac{\gamma _p}{\gamma _{rp}}=\frac{\overline{\gamma }_{LO}/\gamma _0}{1+\frac{\lambda /\mathrm{\Delta }E}{\mathrm{\Delta }E/\mathrm{}\omega _{LO}}}\frac{1\frac{\lambda /\mathrm{\Delta }E}{\sqrt{1+(\lambda /\mathrm{\Delta }E)^2}}}{1\pm \frac{\lambda /\mathrm{\Delta }E}{\sqrt{1+(\lambda /\mathrm{\Delta }E)^2}}},$$ (187) where the upper (lower) sign appropriates to the term $`p=a`$ ($`b`$). Eq. (127) and Eq. (128) were applied in Figs. 9, 10. On Figs. 9, 10 the function $`\gamma _p/\gamma _{rp}`$ for the three values of $`\overline{\gamma }_{LO}/\gamma _0`$ is represented: 10 (curve 1), 1 (curve 2), 0.1 (curve 3). According to our estimate with applying the parameter values from Ref. $$\mathrm{}\gamma _05.3510^5eV.$$ (188) Eqs. (125) - (128) and Fig. 5 - 6 allow us in principle to determine $`𝒜(\omega _l)`$ and $`(\omega _l)`$ with the help of the results of the section IY at a fixed value of magnetic field. One can also determine $`𝒜(H)`$ and $`(H)`$ at fixed $`\omega _l`$ values. The quantity $`\overline{\gamma }_{LO}`$ remains practically unknown. One can try to estimate $`\overline{\gamma }_{LO}`$ on the line width of one-phonon scattering in the bulk GaAs. Therefore on Figs. 6 - 8 the results for the different interrelations $`\overline{\gamma }_{LO}/\gamma _0`$ are represented. One can conclude from Fig. 5 that the magnetopolaron effect area spreads approximately from $`\lambda /\mathrm{\Delta }E=2`$ to $`\lambda /\mathrm{\Delta }E=2`$. At $`\lambda /\mathrm{\Delta }E>2`$ the term $`a`$ (and at $`\lambda /\mathrm{\Delta }E<2`$ the term $`b`$) transits into the EHP energy level with quantum numbers of the electron and hole $`n_e=n_h=1,l_e=l_h=l.`$ The term $`b`$ at $`\lambda /\mathrm{\Delta }E>2`$ and the term $`a`$ at $`\lambda /\mathrm{\Delta }E<2`$ transit into the energy level of the system consisting of the EHP with quantum numbers $`n_e=0,n_h=1,l_e=l_h=l`$ and one LO phonon. This state interacts weakly with exciting light out of the magnetophonon resonance. In the case of the polaron $`A`$ the value $`\mathrm{\Delta }E`$ of energy splitting is very large. Therefore the conditions $`\gamma _{rp}<<\omega _1\omega _2,\gamma _p<<\omega _1\omega _2`$ are obviously satisfied and the approximation of two two-level systems is applicable for calculations of $`𝒜`$ and $``$, i. e. Eq. (36) and Eq. (37) of the section IY. An application of the results of the same section, relevant to the three-level systems, is necessary in the case of smaller values $`\mathrm{\Delta }E`$, which would observed, for instance, in the case of weak polarons (see the section II). In the magnetopolaron area, two maxima have to be observed in the functions $`𝒜(\omega _l)`$ and $`(\omega _l)`$ at the definite value $`H`$ of the magnetic field. The maximum points $`\omega _{la}`$ and $`\omega _{lb}`$ one can determine , drawing the vertical sections on Fig. 5. Two maxima have to be observed also in functions $`𝒜(H)`$ and $`(H)`$ at $`\omega _l=const`$. The appropriate magnetic field values $`H_a`$ and $`H_b`$ can be determined drawing the horizontal sections on Fig. 5. The quantities $`\gamma _{rp}`$ and $`\gamma _p`$ are strongly dependent on magnetic field in the magnetophonon resonance vicinity (see Fig. 6). The results for $`𝒜(\omega _l)`$ and $`(\omega _l)`$ are different for the different values of $`\overline{\gamma }_{LO}/\gamma _0`$. At those values of $`\lambda /\mathrm{\Delta }E`$ (i.e. the magnetic fields $`H`$), where $`\gamma _p/\gamma _{rp}>>1`$, light absorption is much smaller than one , but exceeds light reflection. In the case of curves 1 on Figs. 9,10, relevant to the parameter $`\overline{\gamma }_{LO}/\gamma _0=10`$, this area for the both terms $`a`$ and $`b`$ takes almost the whole interval of magnetic field values where the polaron effect is essential. Indeed, for the curves 1 in Fig. 6 $`\gamma _a/\gamma _{ra}=1`$ at $`\lambda _{a1}/\mathrm{\Delta }E=1.24`$ , and $`\gamma _b/\gamma _{rb}=1`$ at $`\lambda _{b1}/\mathrm{\Delta }E=1.82`$ . Around the points $`\lambda _{a1}`$ and $`\lambda _{b1}`$ light reflection by terms $`a`$ and $`b`$ reaches the largest values ($`𝒜`$=1/2 and $``$=1/4 in maximum). In the regions $`\lambda >>\lambda _{a1}`$, and $`\lambda <<\lambda _{b1}`$ the polaron effect is inessential. The second maxima on the curves $`A(\omega _l)`$ and $`R(\omega _l)`$ disappear. These results are alike the results for the EHP with $`n_e=n_h=1,l_e=l_h=l`$ far from the polaron resonance. In this area $`\gamma _p/\gamma _{rp}<<1`$, and light reflection exceeds light absorption. In the opposite case $`\overline{\gamma }_{LO}/\gamma _0=0.1`$ (curves 3) the condition $`\gamma _p/\gamma _{rp}<<1`$ is satisfied in almost the whole region of the magnetophonon resonance. This region includes the resonant point $`\lambda =0`$. Light absorption is much smaller than light reflection which reaches the values $`=1`$ in both maxima on the curve $`(\omega _l)`$. At $`\lambda _{a3}/\mathrm{\Delta }E=1.24`$ and $`\lambda _{b3}/\mathrm{\Delta }E=1.68`$ the inverse non-radiative and radiative lifetimes become equal , which corresponds to maximum absorption. On the left of the point $`\lambda _{a3}`$ the term $`a`$ does not interact with light (note the small values $`\gamma _{ra}`$). The same is relevant to the term $`b`$ on the right of $`\lambda _{b3}`$. The curves 2 on Figs. 6 are relevant to an intermediate case $`\overline{\gamma }_{LO}/\gamma _0=1.`$ The maximum absorption point coincides with the resonance point $`\lambda =0`$. For the term $`a`$ at $`\lambda >0`$ the condition $`\gamma _a/\gamma _{ra}<1`$ is satisfied, at $`\lambda <0`$ the opposite condition $`\gamma _a/\gamma _{ra}>1`$ is satisfied. There is the opposite picture for the term $`b`$. At $`\gamma _p/\gamma _{rp}<1`$ light reflection dominates, at $`\gamma _p/\gamma _{rp}>1`$ light absorption dominates . Figs. 7, 8 $`𝒜(H)`$ and $`(H)`$ in the polaron splitting vicinity. The curves 1, 2, 3 correspond to three different magnitudes of the frequency $`\omega _l`$. The Figs. 7a and 8a correspond to the ratio $`\overline{\gamma }_{LO}/\gamma _0=10`$. The Figs. 7b and 8b correspond to the ratio $`\overline{\gamma }_{LO}/\gamma _0=0.4`$. To summarize, a classification of magnetopolarons in a QW in strong magnetic field has been done. The usual polarons (including double-, triple- and so on), combined polarons (for which the resonant magnetic field depends on the QW width and depth) as well as the weak polarons (for which a polaron energy splitting $`\mathrm{\Delta }E`$ is comparatively small) have been defined. The formulae for the non-dimensional coefficients of light absorption and reflection by QW under normal incidence of light have been obtained for the three-level electronic systems. Our calculations are based on taking into account the radiative lifetimes of the electronic systems on two excited level, which means taking into account all of the absorption and reradiation processes and a transcendence of the perturbation theory on the light-electron coupling constant. The EHP radiative lifetime in a QW in a strong magnetic field under an arbitrary value of the QW inplane EHP wave vector $`𝐊_{}`$ has been calculated far from the magnetophonon resonance. The radiative lifetimes of a system, consisting of a magnetopolaron and a hole, have been calculated. The expressions for the contributions into non-radiative lifetimes of the polaron states due to the finite LO phonon lifetimes are obtained. It turned out that the radiative lifetimes of polarons as well as non-radiative ones are strongly dependent on magnetic field $`H`$ in the magnetopolaron resonance vicinity. The functions $`𝒜(H)`$ and $`(H)`$ have been obtained by taking into account the magnetopolaron effect. ## IX Acknowledgements S.T.P thanks the Zacatecas University and the National Council of Science and Technology (CONACYT) of Mexico for the financial support and hospitality. D.A.C.S. thanks CONACYT (27736-E) for the financial support. Authors are grateful to V. M. Bañuelos-Alvarez for the computer cooperation and to A. D’Amor for a critical reading of the manuscript. This work has been partially supported by the Russian Foundation for Basic Research and by the Program ”Solid State Nanostructures Physics”.
warning/0001/hep-ph0001188.html
ar5iv
text
# 1 Introduction ## 1 Introduction Together with low-energy $`\pi `$$`\pi `$ scattering , the semi-leptonic $`K_\mathrm{}3`$ and $`K_{e4}`$ decays have been among the first applications of standard chiral perturbation theory (S$`\chi `$PT) to be studied at the one-loop level. At this order, the mesonic form factors that describe these decays do not contain the poorly known low energy constants $`L_8`$ and $`L_6`$, and consequently they may be expected to be less sensitive to the size of the chiral condensate $`\overline{q}q`$ than, e.g., the $`\pi \pi `$ s-wave . In particular, the $`K_\mathrm{}3`$ form factors at $`𝒪(\mathrm{p}^4)`$ merely involve, besides the masses and decay constants of the pseudoscalar states, the low energy constant $`L_9`$, whose value can be obtained from the experimentally measured charge radius of the pion . This fortunate circumstance has been used in the past to extract the CKM matrix element $`V_{us}`$ from the $`K_{e3}`$ decay rates , and presently this extraction is still considered by the Particle Data Group (PDG) compilation to remain the most accurate and least model dependent. Yet this determination of $`V_{us}`$ relies on a model-dependent estimate of $`𝒪(m_{\mathrm{quark}}^2)`$ contributions to the form factor $`f_+(0)`$ (the notation will be given below), which, in S$`\chi `$PT, arise as contributions of order $`𝒪(\mathrm{p}^6)`$. With the new careful and accurate determinations of $`V_{ud}`$ from superallowed nuclear $`\beta `$-decays , the size of these corrections is required to be comparable to the (parameter-free) genuine $`𝒪(\mathrm{p}^4)`$ contribution in order to preserve the unitarity of the CKM matrix. It is then legitimate to ask whether similarly important $`𝒪(m_{\mathrm{quark}}^2)`$ contributions would not affect the parameter-free S$`\chi `$PT prediction for the slope $`\lambda _0`$ of the $`K_{\mu 3}`$ scalar form factor at order $`𝒪(\mathrm{p}^4)`$. In principle, these questions can be answered by performing the full two-loop S$`\chi `$PT calculation of the $`K_\mathrm{}3`$ form factors, provided the several new $`𝒪(\mathrm{p}^6)`$ counterterms that will contribute can be measured independently or estimated in a reliable way <sup>1</sup><sup>1</sup>1The structure of the order $`𝒪(\mathrm{p}^6)`$ effective lagrangian of S$`\chi `$PT has been discussed in Refs. .. The present status of this enterprise is limited <sup>2</sup><sup>2</sup>2Unfortunately, the $`𝒪(\mathrm{p}^6)`$ calculation of Ref. only considers a very specific combination of the slopes of the $`f_+^{K\pi }`$ and of the mesonic electromagnetic form factors. to the evaluation of the so-called chiral double-logs , which, although only part of the full two-loop corrections, do not seem to point towards huge effects coming from the chiral loops themselves (see in particular the numerical results in Table 1 of Ref. and the comments preceding it). The purpose of the present work is to address some of these issues from the point of view of generalized chiral perturbation theory (G$`\chi `$PT) . The expressions for the form factors of semileptonic decay of kaons to $`𝒪(\mathrm{p}^4)`$ in G$`\chi `$PT have never been published (a discussion of the $`K_\mathrm{}4`$ form factors at order $`𝒪(\mathrm{p}^2)`$ in G$`\chi `$PT can be found in Ref. ), although they have existed for some time and have been partially reported on various occasions. Keeping in mind the well-known and important exception of the phase of the $`K_{e4}^+`$ decay amplitude, the $`K_\mathrm{}3`$ and $`K_{e4}`$ form factors are indeed found to be to a large extent independent of the size of the condensate $`\overline{q}q`$. The main consequence of the modified chiral counting ($`m_q𝒪(\mathrm{p}))`$ is that in G$`\chi `$PT the form factors receive a contribution quadratic in quark masses already at order $`𝒪(\mathrm{p}^4)`$, in addition to the standard $`𝒪(\mathrm{p}^4)`$ expressions. Furthermore, these new $`𝒪(m_{\mathrm{quark}}^2)`$ contributions, which within S$`\chi `$PT would count as $`𝒪(\mathrm{p}^6)`$, are all related. They all stem from a single term of the $`_{(2,2)}`$ component of the effective Lagrangian $`_{\mathrm{eff}}`$ and they are described by a single divergence-free low-energy constant $`A_3`$ (see Ref. and Appendix A for notation). (This statement is exact in the case of $`K_\mathrm{}3`$ form factors $`f_+`$ and $`f_{}`$, whereas in the case of $`K_{e4}`$ it remains true for the dominant $`𝒪(m_s^2)`$ terms.) It thus appears that the $`𝒪(\mathrm{p}^4)`$ G$`\chi `$PT offers a predictive description of the terms quadratic in quark masses which is non-trivial compared to the one-loop S$`\chi `$PT (no quadratic terms) and yet much simpler than the standard $`𝒪(\mathrm{p}^6)`$ order, in which other unknown constants should contribute in addition to the $`𝒪(\mathrm{p}^4)`$ G$`\chi `$PT terms driven by the constant $`A_3`$. This framework, which is as systematic as the standard expansion, suggests a new simultaneous analysis of $`V_{us}`$, $`K_\mathrm{}3`$ decay rates, $`F_K/F_\pi `$, and of the scalar slope $`\lambda _0`$ which may be of interest in connection with the constraint of unitarity of the CKM matrix and with the forthcoming new $`K_\mathrm{}3`$ data. A closely related application, namely the determination of the $`𝒪(\mathrm{p}^4)`$ constant $`L_3`$ from the $`K_{e4}^0`$ decay rate, will be presented elsewhere , together with a full discussion of the $`K_\mathrm{}4`$ form factors at order $`𝒪(\mathrm{p}^4)`$ in G$`\chi `$PT. The present stage of our analysis involves one additional limitation, to the extent that no electromagnetic corrections are included, unless explicitly stated. For this reason we postpone a detailed analysis of the isospin asymmetry in the $`K_{e3}^0`$ and $`K_{e3}^+`$ decay rates that is due to the mass difference $`m_dm_u`$. We just check that this asymmetry is consistent with the G$`\chi `$PT treatment of $`\pi ^0\eta `$ mixing within errors. The paper is organized as follows: Section 2 provides the necessary expressions of the $`K_\mathrm{}3`$ form factors and of the $`\pi ^0\eta `$ mixing angles in G$`\chi `$PT. Implications for the determinations of $`V_{us}`$ and of the ratio $`F_K/F_\pi `$ of pseudoscalar decay constants are discussed in Section 3. The slopes of the $`K_\mathrm{}3`$ form factors are considered in Section 4. Concluding remarks are presented in Section 5. Details on the $`𝒪(\mathrm{p}^4)`$ structure of the G$`\chi `$PT effective Lagrangian and on its renormalization are presented in Appendix A. Useful expressions for the pseudoscalar masses and decay constants have been gathered in Appendix B. ## 2 $`K_\mathrm{}3`$ form factors in G$`\chi `$PT to $`𝒪(\mathrm{p}^4)`$ We consider the two semileptonic decay channels $`K^+(p)`$ $``$ $`\pi ^0(p^{})\mathrm{}^+(p_{\mathrm{}})\nu _{\mathrm{}}(p_\nu )[K_\mathrm{}3^+]`$ $`K^0(p)`$ $``$ $`\pi ^{}(p^{})\mathrm{}^+(p_{\mathrm{}})\nu _{\mathrm{}}(p_\nu )[K_\mathrm{}3^0].`$ (2.1) The symbol $`\mathrm{}`$ stands for $`\mu `$ or $`e`$. As stated before, we do not consider electromagnetic corrections. The processes (2.1) are then described by four form factors, $`f_\pm ^{K^+\pi ^0}(t)`$ and $`f_\pm ^{K^0\pi ^{}}(t)`$, which depend on $`t=(p^{}p)^2=(p_l+p_\nu )^2`$, the square of the four momentum transfer to the leptons, and which are defined in terms of the hadronic matrix elements of the charged strangeness changing QCD vector current as follows: $`<\pi ^0(p^{})|(\overline{s}\gamma _\mu u)(0)|K^+(p)>`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\left[(p+p^{})_\mu f_+^{K^+\pi ^0}+(pp^{})_\mu f_{}^{K^+\pi ^0}\right]`$ $`<\pi ^{}(p^{})|(\overline{s}\gamma _\mu u)(0)|K^0(p)>`$ $`=`$ $`(p+p^{})_\mu f_+^{K^0\pi ^{}}+(pp^{})_\mu f_{}^{K^0\pi ^{}}.`$ (2.2) ### 2.1 G$`\chi `$PT expressions of the $`K_\mathrm{}3`$ form factors The one-loop G$`\chi `$PT expressions (using the notation of reference ) for the form factors are summarized below. We start with the two form factors $`f_+^{K^0\pi ^{}}(t)`$ and $`f_+^{K^+\pi ^0}(t)`$, which are in practice sufficient for the description of the electron decay modes $`K_{e3}^+`$ and $`K_{e3}^0`$, and keeping isospin breaking contributions due to the quark mass difference $`m_dm_u`$. For the $`K_\mathrm{}3^0`$ channel, which is somewhat simpler, since $`\pi ^0\eta `$ mixing only enters the loop contributions, we obtain $`f_+^{K^0\pi ^{}}(t)`$ $`=`$ $`1+H_{\pi ^+K^0}(t)+{\displaystyle \frac{1}{2}}H_{\pi ^0K^+}(t)+{\displaystyle \frac{3}{2}}H_{\eta K^+}(t)`$ (2.3) $`+`$ $`\sqrt{3}\stackrel{o}{\epsilon }\left[H_{\pi K}(t)H_{\eta K}(t)\right]`$ $`+`$ $`\left[{\displaystyle \frac{1}{8}}\widehat{m}^2\xi ^2+{\displaystyle \frac{1}{2}}\widehat{m}^2A_3\right](r1)^2\left(1+{\displaystyle \frac{1}{R}}\right),`$ with (the definitions of the loop functions $`M_{PQ}^\mathrm{r}(t)`$, $`K_{PQ}(t)`$ and $`L_{PQ}(t)`$ that we use below can be found in Ref. ) $$H_{PQ}(t)=\frac{1}{F_\pi ^2}\left[tM_{PQ}^\mathrm{r}(t)L_{PQ}(t)+\frac{2}{3}L_9^\mathrm{r}t\right].$$ (2.4) Here $`\stackrel{o}{\epsilon }`$ denotes the leading order $`\pi ^0\eta `$ mixing angle, $$\stackrel{o}{\epsilon }=\frac{\sqrt{3}}{4R}[1\frac{\mathrm{\Delta }_{GMO}}{M_\eta ^2M_\pi ^2}+\frac{2}{3}(r_2r)\frac{r1}{r+1}\frac{M_\pi ^2}{M_\eta ^2M_\pi ^2}],$$ (2.5) where $`r`$ $``$ $`m_s/\widehat{m}`$ $`r_2`$ $``$ $`2{\displaystyle \frac{M_K^2}{M_\pi ^2}}1`$ (2.6) $`\mathrm{\Delta }_{GMO}`$ $``$ $`3M_\eta ^24M_K^2+M_\pi ^2.`$ We work only at first order in the quark mass difference $`(m_dm_u)`$, i.e. we consider only terms that are at most of order $`𝒪(1/R)`$ <sup>3</sup><sup>3</sup>3We have however kept the pion mass difference, which is mainly an electromagnetic effect , in the loop contributions., where $$R=\frac{m_s\widehat{m}}{m_dm_u}.$$ (2.7) The lowest order S$`\chi `$PT value for the $`\pi ^0\eta `$ mixing angle, $`\underset{\mathrm{st}}{\overset{o}{\epsilon }}=\frac{\sqrt{3}}{4R}`$, is recovered by dropping, in Eq. (2.5), the last two terms, which are counted as order $`𝒪(\mathrm{p}^4)`$ in S$`\chi `$PT. The $`\xi ^2`$ and $`A_3`$ terms in Eq. (2.3) are $`𝒪(\mathrm{p}^4)`$ contributions in G$`\chi `$PT (see the detailed formulas for the effective Lagrangian in Appendix A), but are absent at this order in S$`\chi `$PT. For the $`K_\mathrm{}3^+`$ decay mode, we find $`f_+^{K^+\pi ^0}(t)`$ $`=`$ $`f_+^{K^0\pi ^{}}(t)\times \{1+{\displaystyle \frac{\sqrt{3}}{2}}(\epsilon _1+\epsilon _2)`$ $`{\displaystyle \frac{1}{\sqrt{3}}}\widehat{m}^2\xi ^2(r1)^2\stackrel{o}{\epsilon }\widehat{m}^2C_2^P(r1)^2[{\displaystyle \frac{1}{R}}+{\displaystyle \frac{2\stackrel{o}{\epsilon }}{\sqrt{3}}}]\},`$ where the $`\pi ^0\eta `$ mixing angles at order $`𝒪(\mathrm{p}^4)`$, $`\epsilon _1`$, $`\epsilon _2`$, are defined in section 2.2 below. Although the constant $`C_2^P`$ corresponds to a counterterm of $`_{(2,2)}`$ (see Eq. (A) in Appendix A) that violates the Zweig rule, it is not expected to be suppressed, since this violation occurs in the $`0^{}`$ channel. At zero momentum transfer, Eq. (2.3) gives $`f_+^{K^0\pi ^{}}(0)`$ $`=`$ $`1{\displaystyle \frac{1}{256\pi ^2F_\pi ^2}}\{2(M_{\pi ^+}^2+M_{K^0}^2)h_0\left({\displaystyle \frac{M_{\pi ^+}^2}{M_{K^0}^2}}\right)`$ $`+(M_{\pi ^0}^2+M_{K^+}^2)h_0\left({\displaystyle \frac{M_{\pi ^0}^2}{M_{K^+}^2}}\right)+3(M_{K^+}^2+M_\eta ^2)h_0\left({\displaystyle \frac{M_{K^+}^2}{M_\eta ^2}}\right)`$ $`+2\sqrt{3}\stackrel{o}{\epsilon }[(M_\pi ^2+M_K^2)h_0\left({\displaystyle \frac{M_\pi ^2}{M_K^2}}\right)(M_K^2+M_\eta ^2)h_0\left({\displaystyle \frac{M_K^2}{M_\eta ^2}}\right)]\}`$ $`+`$ $`\left[{\displaystyle \frac{1}{8}}\widehat{m}^2\xi ^2+{\displaystyle \frac{1}{2}}\widehat{m}^2A_3\right](r1)^2\left(1+{\displaystyle \frac{1}{R}}\right),`$ with $$h_0(x)=\mathrm{\hspace{0.17em}1}+\frac{2x}{1x^2}\mathrm{ln}x.$$ (2.10) To recover the expression at order $`𝒪(\mathrm{p}^4)`$ in S$`\chi `$PT for $`f_+^{K^0\pi ^{}}(0)`$, one replaces $`\stackrel{o}{\epsilon }`$ by the leading order S$`\chi `$PT value $`\underset{\mathrm{st}}{\overset{o}{\epsilon }}`$, and one simply drops the last line in Eq. (2.1). Notice that this last contribution is the only correction of order $`𝒪(\mathrm{p}^4)`$ that does not vanish in the large-$`N_c`$ limit of QCD. The fact that the $`𝒪(\mathrm{p}^4)`$ corrections to $`f_+^{K^0\pi ^{}}(0)`$ in S$`\chi `$PT vanish altogether in the large-$`N_c`$ limit might provide a natural explanation why contributions of the $`𝒪(\mathrm{p}^6)`$ counterterms could be comparatively sizeable. The expressions of the form factors $`f_{}^{K^0\pi ^{}}(t)`$ and $`f_{}^{K^+\pi ^0}(t)`$ are rather cumbersome if $`𝒪(m_dm_u)`$ effects are included. Since we do not need the latter in this case, we give only the common expression of these two form factors in the isospin limit, which is $`f_{}^{K\pi }(t)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\widehat{m}\xi (r1){\displaystyle \frac{1}{4}}\widehat{m}^2\xi ^2(r1)(r+3)\widehat{m}^2\xi \stackrel{~}{\xi }(r1)(r+2)`$ $`+`$ $`{\displaystyle \frac{1}{2}}\widehat{m}^2(A_1+B_1)(r^21)+{\displaystyle \frac{1}{2}}\widehat{m}^2(A_22B_2)(r1)`$ $`+`$ $`\widehat{m}^2D^S(r1)(r+2)+{\displaystyle \frac{1}{4}}[5\mu _\pi 2\mu _K3\mu _\eta ]`$ $``$ $`{\displaystyle \frac{1}{4F_0^2}}\left[5t2M_\pi ^22M_K^2+16\widehat{m}^2(A_0+2Z_0^S)(r+1)\right]K_{\pi K}(t)`$ $``$ $`{\displaystyle \frac{1}{4F_0^2}}\left[3t2M_\pi ^22M_K^2+4\widehat{m}^2(A_0+2Z_0^P)(r1)(r+3)\right]K_{K\eta }(t)`$ $``$ $`{\displaystyle \frac{3}{2F_0^2}}(M_K^2M_\pi ^2)\left\{\left[M_{K\pi }^\mathrm{r}(t)+{\displaystyle \frac{2}{3}}L_9^\mathrm{r}\right]+\left[M_{K\eta }^\mathrm{r}(t)+{\displaystyle \frac{2}{3}}L_9^\mathrm{r}\right]\right\}.`$ For the ease of comparison, we may rewrite this G$`\chi `$PT expression for the form factor $`f_{}(t)`$ in terms of the corresponding S$`\chi `$PT expression: $`f_{}^{K\pi ;std}(t)`$ $`=`$ $`{\displaystyle \frac{F_K}{F_\pi }}1{\displaystyle \frac{2L_9^r\left(M_K^2M_\pi ^2\right)}{F_\pi ^2}}`$ (2.12) $`+{\displaystyle \frac{\left(2M_K^2+2M_\pi ^23t\right)}{4F_\pi ^2}}K_{K\eta }(t)`$ $`+{\displaystyle \frac{\left(2M_K^2+2M_\pi ^25t\right)}{4F_\pi ^2}}K_{\pi K}(t)`$ $`{\displaystyle \frac{3\left(M_K^2M_\pi ^2\right)}{2F_\pi ^2}}\left[M_{}^{r}{}_{K\eta }{}^{}(t)+M_{}^{r}{}_{K\pi }{}^{}(t)\right],`$ $`f_{}^{K\pi }(t)`$ $`=`$ $`f_{}^{K\pi ;std}(t)`$ (2.13) $`{\displaystyle \frac{1}{4}}\left({\displaystyle \frac{F_K}{F_\pi }}1\right)^2\left({\displaystyle \frac{F_K^2}{F_{\pi }^{}{}_{}{}^{2}}}+2{\displaystyle \frac{F_K}{F_\pi }}1\right)`$ $`{\displaystyle \frac{1}{2}}m^2A_3(r1)(r+3)`$ $`{\displaystyle \frac{\mathrm{\Delta }_{GMO}\left(r+3\right)}{4F_\pi ^2\left(r1\right)}}K_{K\eta }(t){\displaystyle \frac{M_\pi ^2\left(1+r\right)\widehat{ϵ}(r)}{F_\pi ^2}}K_{\pi K}(t),`$ where $`\widehat{ϵ}(r)`$ $``$ $`{\displaystyle \frac{4\widehat{m}^2(A_0+2Z_0^S)}{M_\pi ^2}}=2{\displaystyle \frac{r_2r}{r^21}}(1+2\zeta )`$ $`\zeta `$ $`=`$ $`Z_0^S/A_0,`$ (2.14) and the $`𝒪(\mathrm{p}^4)`$ expression $`\widehat{m}\xi `$ $`=`$ $`{\displaystyle \frac{1}{r1}}\left({\displaystyle \frac{F_K^2}{F_\pi ^2}}1\right)`$ $`(r+1)\widehat{m}^2(A_1+B_1)\widehat{m}^2(A_22B_2)(r+3)\widehat{m}^2A_3`$ $`2(r+2)\widehat{m}^2D^S`$ $`+2\widehat{m}^2\xi ^2+2(r+2)\widehat{m}^2\xi \stackrel{~}{\xi }`$ $`{\displaystyle \frac{1}{2(r1)}}(5\mu _\pi 2\mu _K3\mu _\eta )`$ has been used. It is remarkable that, in Eq. (2.13), all $`_{(2,2)}`$ constants – except for $`A_3`$ – have been absorbed into the renormalization of the decay constants. Notice also that, in contrast to $`f_+(t)`$, $`f_{}(t)`$ starts only at next-to-leading order in the chiral expansion, and that furthermore this contribution is not suppressed at $`t=0`$ for $`N_c\mathrm{}`$. ### 2.2 Mixing angles The expressions of the $`K_\mathrm{}3`$ form factors given in the preceding subsection involve the $`\pi ^0\eta `$ mixing angles $`\stackrel{o}{\epsilon },\epsilon _1,\epsilon _2`$. In defining them we ignore isospin breaking through electromagnetic effects, so that the only source of isospin violation is the quark mass difference $`m_dm_u`$. If $`m_dm_u`$, the isosinglet and isovector axial currents $`A_\mu ^8`$ and $`A_\mu ^3`$ have nonvanishing off-diagonal matrix elements between the vacuum and one-meson states. The $`two`$ $`\pi ^0\eta `$ mixing angles $`\epsilon _1`$ and $`\epsilon _2`$ are introduced such as to define combinations of the axial currents having vanishing off-diagonal matrix elements: $`\mathrm{\Omega }|\mathrm{cos}\epsilon _1A_\mu ^8(0)\mathrm{sin}\epsilon _1A_\mu ^3(0)|\pi ^0(p)`$ $`=`$ $`0`$ $`\mathrm{\Omega }|\mathrm{cos}\epsilon _2A_\mu ^3(0)+\mathrm{sin}\epsilon _2A_\mu ^8(0)|\eta (p)`$ $`=`$ $`0.`$ Both $`\epsilon _1`$ and $`\epsilon _2`$ are of order $`𝒪(m_dm_u)`$. Note that we were not forced to define mixing angles in terms of matrix elements of the axial currents: we could have chosen to use the pseudoscalar densities, or any other operators with the appropriate quantum numbers; however, the corresponding expressions for the mixing angles would in general differ from those given below. Keeping only contributions which are at most linear in the quark mass difference $`m_dm_u`$, the off-diagonal matrix elements of the flavor neutral axial currents read $`\mathrm{\Omega }|A_\mu ^8(0)|\pi ^0(p)`$ $`=`$ $`ip_\mu F_\pi \epsilon _1`$ $`\mathrm{\Omega }|A_\mu ^3(0)|\eta (p)`$ $`=`$ $`ip_\mu F_\eta \epsilon _2.`$ The decay constants $`F_\pi `$ and $`F_\eta `$ define the diagonal matrix elements of the same currents, $`\mathrm{\Omega }|A_\mu ^3(0)|\pi ^0(p)`$ $`=`$ $`ip_\mu F_\pi `$ $`\mathrm{\Omega }|A_\mu ^8(0)|\eta (p)`$ $`=`$ $`ip_\mu F_\eta .`$ Up to corrections of order $`𝒪((m_um_d)^2)`$ that we neglect, $`F_\pi `$ can be identified with the charged pion decay constant, $`F_\pi `$ = 92.4 MeV. At order $`𝒪(\mathrm{p}^2)`$, the two mixing angles coincide and are equal to $`\stackrel{o}{\epsilon }`$. At order $`𝒪(\mathrm{p}^3)`$, they both receive $`𝒪(m_{\mathrm{quark}})`$ corrections, and are no longer equal. Explicitly, $`{\displaystyle \frac{\epsilon _1+\epsilon _2}{2}}`$ $`=`$ $`\stackrel{o}{\epsilon }+{\displaystyle \frac{1}{\sqrt{3}R(M_\eta ^2M_\pi ^2)}}\{\widehat{m}^3X_\rho (r)`$ (2.19) $`({\displaystyle \frac{F_K^2}{F_\pi ^2}}1)[{\displaystyle \frac{\mathrm{\Delta }_{GMO}}{2}}+{\displaystyle \frac{1}{3}}(r_2r)\left({\displaystyle \frac{r+2}{r+1}}\right)M_\pi ^2]\},`$ and $`\epsilon _1\epsilon _2`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{3}R(M_\eta ^2M_\pi ^2)}}\left({\displaystyle \frac{F_K^2}{F_\pi ^2}}1\right)\left[\mathrm{\Delta }_{GMO}{\displaystyle \frac{2}{3}}(r_2r)\left({\displaystyle \frac{r2}{r+1}}\right)M_\pi ^2\right],`$ (2.20) where we have defined $$X_\rho (r)=(r1)^2[(3\rho _1+4\rho _3)(r1)\rho _2(r+1)]$$ (2.21) in terms of the $`_{(0,3)}`$ quantities $`\rho _{1,2,3}`$. ## 3 Decay rates, $`V_{us}`$ and $`F_K/F_\pi `$ revisited ### 3.1 The $`K^0\pi ^{}e^+\nu `$ rate For the $`K_{e3}`$ decay we only need to consider the form factor $`f_+(t)`$, and for $`K^0\pi ^{}e^+\nu `$ Eq. (2.3) is all we need. Ignoring the tiny term in Eq. (2.3) proportional to $`\stackrel{o}{\epsilon }`$, we may write the decay rate in units of $`10^{15}\text{MeV}`$ as $`\mathrm{\Gamma }[K^0\pi ^{}e^+\nu ]`$ $`=`$ $`V_{us}^2[105.056`$ $`+\mathrm{\hspace{0.17em}203.031}(r1)^2\left({\displaystyle \frac{\widehat{m}^2A_3}{2}}+{\displaystyle \frac{\widehat{m}^2\xi ^2}{8}}\right)`$ $`+\mathrm{\hspace{0.17em}98.267}(r1)^4({\displaystyle \frac{\widehat{m}^2A_3}{2}}+{\displaystyle \frac{\widehat{m}^2\xi ^2}{8}})^2].`$ Strictly speaking, the last contribution on the right-hand side of this expression, although coming from the square of the $`𝒪(\mathrm{p}^4)`$ expression of $`f_+^{K^0\pi ^{}}(t)`$, represents an order $`𝒪(\mathrm{p}^6)`$ effect in the chiral expansion of the decay rate. However, as can be checked e.g. on Figs. 1 and 2 below, it does not affect our analysis in the range of values considered for the Cabibbo angle. We express $`\widehat{m}\xi `$ in terms of $`V_{us}`$ as follows: first, we use Eq. (2.1), but only at lowest order, $$\widehat{m}\xi =\frac{1}{r1}\left(\frac{F_K^2}{F_\pi ^2}1\right),$$ (3.2) and then use the formula for the ratio of the branching rates $$\frac{\mathrm{\Gamma }[K\mu \nu ]}{\mathrm{\Gamma }[\pi \mu \nu ]}=\frac{\left|V_{us}\right|^2}{\left|V_{ud}\right|^2}\frac{F_K^2}{F_\pi ^2}\frac{M_{K^+}}{M_{\pi ^+}}\frac{\left\{1(M_\mu /M_{K^+})^2\right\}^2}{\left\{1(M_\mu /M_{\pi ^+})^2\right\}^2}(1+\delta _K\delta _\pi ),$$ (3.3) where the radiative corrections $`\delta _K=0.0020\pm 0.0002,\delta _\pi =0.0017\pm 0.0002`$ nearly cancel . Next, we use the experimental values for the branching rates to fix the combination of constants $$\frac{F_K}{F_\pi }\frac{V_{us}}{V_{ud}}=0.2758\pm 0.0005,$$ (3.4) together with the unitarity of the Cabibbo-Kobayashi-Maskawa (CKM) matrix $$V_{ud}^2+V_{us}^2=1.$$ (3.5) We ignore $`|V_{ub}|`$, which has been obtained from a model-dependent analysis of data from B semileptonic decays to be (3.25 $`\pm `$ 0.6) $`\times 10^3`$; however, even if the central value of $`V_{ub}`$ turned out to be three times larger, this would not affect our results. After expanding Eq. (LABEL:GammaK0) in $`V_{us}`$ around $`V_{us}=0.220`$, this implies $`\mathrm{\Gamma }[K^0\pi ^{}e^+\nu ]`$ $`=`$ $`5.390+31.10(V_{us}0.220)`$ (3.6) $`+(r1)^2\widehat{m}^2A_3[5.059+37.58(V_{us}0.220)]`$ $`+(r1)^4\widehat{m}^4A_3^2[1.189+10.81(V_{us}0.220)],`$ to be compared with the experimental rate of $`4.937\pm 0.052`$. Consequently, from the knowledge of $`V_{us}`$ one may extract a value for the $`_{(2,2)}`$ quantity $`\widehat{m}^2(r1)^2A_3=(m_s\widehat{m})^2A_3`$. For example, in Figure 1 we show the band in the ($`V_{us},A_3`$) plane indicated by experiment, together with lines of constant $`V_{us}`$ corresponding to the Particle Data Group (PDG) value $`V_{us}=0.2196\pm 0.0023`$. Accepting this constraint on $`V_{us}`$ would imply $`\widehat{m}^2(r1)^2A_3=0.089\pm 0.024`$. A naive dimensional analysis (NDA) would give $$|(r1)^2\widehat{m}^2A_3||_{NDA}\frac{m_s^2(\mathrm{\Lambda }_H)}{\mathrm{\Lambda }_H^2}0.04,$$ (3.7) where we have taken $`m_s(\mathrm{\Lambda }_H)`$ 200 MeV, and $`\mathrm{\Lambda }_H`$ 1 GeV is a typical hadronic mass scale. ### 3.2 Determination of $`V_{ud}`$ from nuclear beta decay In its 1998 update, PDG recommends for $`V_{us}`$ only the value determined from $`K_{e3}`$ decay, arguing that the value obtained from hyperon decays (see for an early attempt along these lines) suffers from theoretical uncertainties due to first-order SU(3) symmetry-breaking effects in the axial-vector couplings. However, within the context of the present G$`\chi `$PT analysis, we may not, without independent knowledge of $`A_3`$, use the values for $`V_{us}`$ thereby extracted from $`K_{e3}`$ decay. Moreover, $`F_K`$ is determined from the $`K_{\mu 2}`$ decay together with the knowledge of $`V_{us}`$, and we note that $`F_K/F_\pi `$ appears implicitly in our theoretical expressions for the $`K_\mathrm{}3`$ form factors. The origin of this is, of course, our re-expression Eq. (3.2) of the $`_{(0,3)}`$ constant $`\xi `$ in terms of $`F_K/F_\pi `$. ¿From the unitarity condition for the CKM matrix, it follows that $`|V_{us}|`$ may be fixed from the knowledge of the up-down quark-mixing matrix element of the CKM matrix, $`|V_{ud}|`$, alone. The value of $`V_{ud}`$ can be determined from several independent sources: nuclear superallowed Fermi beta decays, free neutron decay, and pion beta decay. Currently, superallowed Fermi $`0^+0^+`$ nuclear beta decays , together with the muon lifetime, provide the most accurate value, $$V_{ud}=0.9740\pm 0.0005.$$ (3.8) The precision is limited not by experimental error but by the estimated uncertainty in theoretical corrections . In Figure 2 we show the band in the $`V_{ud},A_3`$ plane indicated by experiment, together with lines of constant $`V_{ud}`$ corresponding to these values. The determinations of $`V_{ud}`$ from free neutron decay data are approximately a factor of four poorer in precision (see the discussion in and references therein), $$V_{ud}=0.9759\pm 0.0021,$$ (3.9) due to the difficulty in separating the vector from the axial piece, but planned experiments aiming at an accurate measurement of the electron emission asymmetry could change this situation qualitatively, since the error in this case is primarily of experimental origin. Finally, the theoretical corrections in the nuclear Fermi transitions are absent in the case of the pion beta decay. The present status of this type of experiments results in the value $$V_{ud}=0.9670\pm 0.0161.$$ (3.10) Here also, the situation might improve in the future . Combining the above results, one obtains $$V_{ud}=0.9741\pm 0.0005.$$ (3.11) Accepting this value for $`V_{ud}`$ would imply $`\widehat{m}^2(r1)^2A_3=0.124\pm 0.022`$, somewhat larger than the NDA estimate Eq. (3.7), but still acceptable. Finally, we note that recent (model-dependent) analyses of hyperon semileptonic decays give $$V_{ud}=0.9750\pm 0.0004,$$ (3.12) and $$V_{ud}=0.9743\pm 0.0009.$$ (3.13) Taking the value Eq. (3.11) for $`V_{ud}`$ (i.e., excluding results from hyperon decays), the unitarity relation Eq. (3.5) gives $$V_{us}=0.2261\pm 0.0023.$$ (3.14) Incorporating the above values for the CKM matrix elements into Eq. (3.4) then implies $$\frac{F_K}{F_\pi }=1.189\pm 0.012[V_{ud}\mathrm{from}\mathrm{nuclear}\mathrm{beta}\mathrm{decay}],$$ (3.15) which may be compared with the corresponding result using the PDG values for $`V_{us}`$, $$\frac{F_K}{F_\pi }=1.226\pm 0.014[V_{us}\mathrm{from}\mathrm{PDG}].$$ (3.16) Using Eq. (3.4), which relates $`F_K/F_\pi `$ to $`V_{us}`$ and $`V_{ud}`$, Eq. (3.5) which relates $`V_{us}`$ and $`V_{ud}`$, and Eq. (3.6) which relates $`V_{us}`$ and $`\widehat{m}^2(r1)^2A_3`$, we may directly relate $`F_K/F_\pi `$ and $`\widehat{m}^2(r1)^2A_3`$, see Figure 3. ### 3.3 $`m_um_d`$ effects on the $`K^+\pi ^0e^+\nu `$ rate We may treat the decay of the $`K^+`$ similarly; the result is most conveniently expressed in terms of the ratio $`{\displaystyle \frac{\mathrm{\Gamma }[K^+\pi ^0e^+\nu ]}{\mathrm{\Gamma }[K^0\pi ^{}e^+\nu ]}}`$ $`=`$ $`(R^{+0})^2[0.496150.00240(r1)^2({\displaystyle \frac{\widehat{m}^2A_3}{2}}+{\displaystyle \frac{\widehat{m}^2\xi ^2}{8}})`$ $`+\mathrm{\hspace{0.17em}0.00234}(r1)^4({\displaystyle \frac{\widehat{m}^2A_3}{2}}+{\displaystyle \frac{\widehat{m}^2\xi ^2}{8}})^2],`$ where $`R^{+0}`$ is the ratio $$R^{+0}\frac{f_+^{K^+\pi ^0}(0)}{f_+^{K^0\pi ^{}}(0)}.$$ (3.18) Note that the last two terms in the brackets in Eq. (3.3) arise from the small differences in phase space which are due to mass differences between $`K^0`$ and $`K^+`$, and between $`\pi ^+`$ and $`\pi ^0`$. Proceeding as we did for the $`K^0`$ rate above, i.e., using Eqs. (3.2) and (3.4), we obtain $`{\displaystyle \frac{\mathrm{\Gamma }[K^+\pi ^0e^+\nu ]}{\mathrm{\Gamma }[K^0\pi ^{}e^+\nu ]}}`$ $`=`$ $`(R^{+0})^2\{0.4961+0.0040(V_{us}0.22)`$ (3.19) $`(r1)^2\widehat{m}^2A_3[0.0011+0.0041(V_{us}0.22)]`$ $`+0.0005(r1)^4\widehat{m}^4A_3^2\}.`$ The correction terms in the curly brackets are completely negligible for $`A_3`$ and $`V_{us}`$ in the range determined above, since they are at most of the order of 0.0001, compared to the leading term 0.4961. The measured rates may be deduced from the data given by the Particle Data Group : $`\mathrm{\Gamma }[K^+\pi ^0e^+\nu ]_{exp}`$ $`=`$ $`2.561\pm 0.032`$ $`\mathrm{\Gamma }[K^0\pi ^{}e^+\nu ]_{exp}`$ $`=`$ $`4.937\pm 0.052,`$ where, as above, we express all rates in units of $`10^{15}`$ MeV; consequently, $$\frac{\mathrm{\Gamma }[K^+\pi ^0e^+\nu ]_{exp}}{\mathrm{\Gamma }[K^0\pi ^{}e^+\nu ]_{exp}}=0.519\pm 0.016.$$ (3.21) Consequently, the experimental value for $`R^{+0}`$ is 1.023 $`\pm `$ 0.016. Up to and including terms of order $`𝒪(\mathrm{p}^4)`$, the theoretical expression for $`R^{+0}`$ is given by Eq. (2.1); the terms proportional to $`\widehat{m}^2\xi ^2(r1)^2`$ and to $`\widehat{m}^2C_2^P`$ are of order $`𝒪(\mathrm{p}^4)`$. We first estimate the $`𝒪(\mathrm{p}^3)`$ correction terms in Eq. (2.19) by NDA. For the first one, we obtain $$\left|\frac{\widehat{m}^3X_\rho (r)}{M_\pi ^2}\right|_{NDA}\left|\frac{8r^3\widehat{m}^3\rho }{M_\pi ^2}\right|_{NDA}\frac{8m_s^3(\mathrm{\Lambda }_H)}{M_\pi ^2\mathrm{\Lambda }_H}3.2,$$ (3.22) where $`\rho `$ characterizes the size of the $`_{(0,3)}`$ low-energy constants $`\rho _\alpha ,\alpha =1,2,3`$. This implies (using $`R=43.5`$, from reference ) $$\frac{\widehat{m}^3X_\rho (r)}{R(M_\eta ^2M_\pi ^2)}\frac{0.22}{R}0.005.$$ (3.23) Similarly, the last term in Eq. (2.19) is $`{\displaystyle \frac{1}{\sqrt{3}R(M_\eta ^2M_\pi ^2)}}\left({\displaystyle \frac{F_K^2}{F_\pi ^2}}1\right)\left[{\displaystyle \frac{\mathrm{\Delta }_{GMO}}{2}}+{\displaystyle \frac{1}{3}}(r_2r)\left({\displaystyle \frac{r+2}{r+1}}\right)M_\pi ^2\right]`$ $`{\displaystyle \frac{0.033}{R}}+{\displaystyle \frac{0.006(r_2r)}{R}}\left({\displaystyle \frac{r+2}{r+1}}\right)`$ (3.24) $`0.00075+0.00014(r_2r)\left({\displaystyle \frac{r+2}{r+1}}\right),`$ which is quite negligible. Finally, we turn to the $`𝒪(\mathrm{p}^4)`$ corrections to $`R^{+0}`$. From $$\widehat{m}^2\xi ^2(r1)^2\left(\frac{F_K^2}{F_\pi ^2}1\right)^20.24,$$ (3.25) we obtain $$\frac{\widehat{m}^2\xi ^2(r1)^2\stackrel{o}{\epsilon }}{\sqrt{3}}0.14\stackrel{o}{\epsilon }.$$ (3.26) The NDA estimate for the $`C_2^P`$ term gives $$\widehat{m}^2\left|C_2^P\right|(r1)^2\left[\frac{1}{R}+\frac{2\stackrel{o}{\epsilon }}{\sqrt{3}}\right]0.04\left[\frac{1}{R}+\frac{2\stackrel{o}{\epsilon }}{\sqrt{3}}\right],$$ (3.27) where we note that no assumption has been made here of any suppression of $`C_2^P`$ due to Zweig rule violation. We have verified that $`𝒪(\mathrm{p}^4)`$ corrections to $`\epsilon _1+\epsilon _2`$ are negligibly small. Using the numerical form of Eq. (2.5), $$\stackrel{o}{\epsilon }=\frac{1}{R}[0.533+0.502\frac{r1}{r+1}(1\frac{r}{r_2})],$$ (3.28) the upshot is that $`R^{+0}`$ may be written (recall Eq. (2.1)) $`R^{+0}`$ $`=`$ $`1+\sqrt{3}\stackrel{o}{\epsilon }\mathrm{\hspace{0.17em}0.14}\stackrel{o}{\epsilon }\pm {\displaystyle \frac{0.22}{R}},`$ (3.29) where the term $`0.14\stackrel{o}{\epsilon }`$ comes from Eq. (3.26); the indicated uncertainty, coming from other higher-order corrections, is dominated by the estimation of the contributions from the $`_{(0,3)}`$ terms (see Eq. (3.23) above). Numerically, this implies $$R^{+0}=\mathrm{\hspace{0.17em}1}+\frac{1}{R}\left[0.85+0.80\frac{r1}{r+1}\left(1\frac{r}{r_2}\right)\right]\pm \frac{0.22}{R},$$ (3.30) and taking, for example, the commonly accepted value $`R`$ = 43.5 $`\pm `$ 2.2, $$R^{+0}=1.020\pm 0.005+0.018\frac{r1}{r+1}\left(1\frac{r}{r_2}\right),$$ (3.31) which is consistent, for any permissible value of $`r`$, with the experimental value 1.023 $`\pm `$ 0.016 which we deduced above. In view of this experimental uncertainty, we will leave for a later time the careful investigation of the $`𝒪(\mathrm{p}^4)`$ effects mentioned above, together with the analysis of electromagnetic corrections (the radiative corrections to $`R^{+0}`$ in the standard case have been investigated in Ref. ). Note that some (but not all) of the existing experimental data has been published with radiative corrections, but often without mention of how these corrections have been implemented . More precise knowledge of $`R^{+0}`$ would be useful in constraining the relationship displayed in Eq. (3.29) between the two quark-mass ratios $`R`$ and $`r`$, thereby testing the relevance of G$`\chi `$PT. ## 4 Form factor slopes Analyses of $`K_\mathrm{}3`$ data frequently assume a linear dependence $$f_{+,0}^{K\pi }(t)=f_+^{K\pi }(0)\left[1+\lambda _{+,0}\frac{t}{M_{\pi ^+}^2}\right],$$ (4.1) where, as usual, the scalar form factor is defined as $$f_0^{K\pi }(t)=f_+^{K\pi }(t)+\frac{t}{M_K^2M_\pi ^2}f_{}^{K\pi }(t).$$ (4.2) The parameter $`\lambda _+^{K\pi }`$ is identical to that of S$`\chi `$PT, $`{\displaystyle \frac{\lambda _+^{K\pi }}{M_{\pi ^+}^2}}`$ $`=`$ $`{\displaystyle \frac{1}{6}}\mathrm{r}^2_\mathrm{V}^\pi `$ (4.3) $`+{\displaystyle \frac{1}{64\pi ^2F_0^2}}\left[1{\displaystyle \frac{1}{2}}h_1\left({\displaystyle \frac{M_\pi ^2}{M_K^2}}\right){\displaystyle \frac{1}{2}}h_1\left({\displaystyle \frac{M_K^2}{M_\eta ^2}}\right)+{\displaystyle \frac{5}{12}}\mathrm{ln}\left({\displaystyle \frac{M_\pi ^2}{M_K^2}}\right)+{\displaystyle \frac{1}{4}}\mathrm{ln}\left({\displaystyle \frac{M_K^2}{M_\eta ^2}}\right)\right],`$ with $$h_1(x)+\frac{1}{2}\frac{(x^33x^23x+1)}{(x1)^3}\mathrm{ln}x+\frac{1}{2}\left(\frac{x+1}{x1}\right)^2\frac{1}{3},$$ (4.4) and $$F_\pi ^2\mathrm{r}^2_\mathrm{V}^\pi =\mathrm{\hspace{0.17em}12}L_9^r\frac{1}{32\pi ^2}\left\{2\mathrm{ln}\frac{M_\pi ^2}{\mu ^2}+\mathrm{ln}\frac{M_K^2}{\mu ^2}+3\right\}.$$ (4.5) On the other hand, the G$`\chi `$PT expression for $`\lambda _0^{K\pi }`$, $`{\displaystyle \frac{\lambda _0^{K\pi }}{M_{\pi ^+}^2}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left[\left({\displaystyle \frac{F_K^2}{F_\pi ^2}}\mathrm{\hspace{0.17em}1}\right){\displaystyle \frac{1}{2}}\left({\displaystyle \frac{F_K^2}{F_\pi ^2}}\mathrm{\hspace{0.17em}1}\right)^2\widehat{m}^2(r1)^2A_3{\displaystyle \frac{r+3}{r1}}\right]{\displaystyle \frac{1}{M_K^2M_\pi ^2}}`$ $``$ $`{\displaystyle \frac{1}{384\pi ^2F_0^2}}\left[5{\displaystyle \frac{4M_\pi ^2(r+1)\widehat{ϵ}(r)}{M_K^2+M_\pi ^2}}\right]h_2\left({\displaystyle \frac{M_\pi ^2}{M_K^2}}\right)`$ $``$ $`{\displaystyle \frac{1}{384\pi ^2F_0^2}}[3+`$ $`\mathrm{\hspace{0.17em}2}\left({\displaystyle \frac{M_\eta ^2M_K^2}{M_\eta ^2+M_K^2}}\right)\left({\displaystyle \frac{M_K^2+M_\pi ^2}{M_K^2M_\pi ^2}}\right)(1{\displaystyle \frac{\mathrm{\Delta }_{GMO}(r+3)}{2(M_K^2+M_\pi ^2)(r1)}})\left]h_2\right({\displaystyle \frac{M_K^2}{M_\eta ^2}})`$ $`+`$ $`{\displaystyle \frac{1}{64\pi ^2F_0^2}},`$ where $$h_2(x)=\frac{3}{2}\left(\frac{1+x}{1x}\right)^2+\frac{3x(1+x)}{(1x)^3}\mathrm{ln}x,$$ (4.7) differs from the S$`\chi `$PT result. This expression for $`\lambda _0`$ explicitly displays the dependence on $`\widehat{m}^2(r1)^2A_3`$ and on $`r`$, while its dependence on $`V_{us}`$ is implicit in the $`F_K/F_\pi `$ terms. However, as we saw above in Figure 3, $`F_K/F_\pi `$ and $`\widehat{m}^2(r1)^2A_3`$ are correlated once we know the rate for $`K^0\pi ^{}e^+\nu `$. Therefore, for any choice of the quark mass ratio $`r`$, Eq. (4) gives a range of values for $`\lambda _0`$ corresponding to a given region in the ($`F_K/F_\pi ,\widehat{m}^2(r1)^2A_3`$) plane. For example, Eq. (3.15) defines one region in Figure 3, while Eq. (3.16) defines another. We display in Figure 4 the corresponding regions in the ($`\lambda _0,r`$) plane. This analysis was done assuming that the Zweig-violating parameter $`Z_0^S`$=0. The whole dependence on $`Z_0^S`$ is contained (through $`\zeta =Z_0^S/A_0`$) in the parameter $`\widehat{ϵ}(r)`$. As pointed out in , the vacuum stability requirement $`B_00`$ implies a upper bound on $`\zeta `$, which yields $$\widehat{ϵ}(r)\widehat{ϵ}(r)|_{\zeta =0}<(1\widehat{ϵ}(r)|_{\zeta =0})\frac{2}{r+2}$$ (4.8) We find that taking $`\zeta =0`$ or its maximal allowed value makes a difference of less than 0.0004 in $`\lambda _0`$, for any choice of $`r`$. In S$`\chi `$PT, including one-loop corrections, the result is $$\lambda _0=0.017\pm 0.004,$$ (4.9) where the error is an estimate of the uncertainties due to higher-order contributions. The experimental situation remains unclear, in view of the inconsistency between some more recent data and the result ($`\lambda _0=0.019\pm 0.004`$) of the high-statistics experiment of Donaldson et al . ## 5 Summary and discussion In the preceding pages, we have studied the $`K_\mathrm{}3`$ form factors at $`𝒪(\mathrm{p}^4)`$ precision within the generalized framework of chiral perturbation theory. In this connection, three (related) issues have been discussed: the extraction of the CKM matrix element $`V_{us}`$ from the experimental value of the $`K_\mathrm{}3`$ branching ratios, the determination of the ratio $`F_K/F_\pi `$ of pseudoscalar decay constants, and the prediction for the slope $`\lambda _0`$ of the scalar form factor. In the Leutwyler-Roos analysis , the expression of $`f_+^{K^0\pi ^{}}(0)`$ is written as $$f_+^{K^0\pi ^{}}(0)=1+f_1+f_2+\mathrm{},$$ (5.1) where the contributions $`f_1`$ and $`f_2`$ are of order $`m_{\mathrm{quark}}`$ and $`m_{\mathrm{quark}}^2`$ respectively. The one-loop S$`\chi `$PT contribution $`f_1`$ arises from Goldstone boson loops only, a counterterm contribution at this order of the standard counting being forbidden by the Ademollo-Gatto theorem. This leads to a parameter free prediction $`f_1=0.023`$. Furthermore, $`f_2`$, which would arise at chiral order $`𝒪(\mathrm{p}^6)`$ in S$`\chi `$PT, has been estimated in to be $``$0.016 $`\pm `$ 0.008 by using a model for pion and kaon wave functions to compute matrix elements in the infinite momentum frame. This leads to a value $`V_{us}=0.2196\pm 0.0023`$. Assuming unitarity of the CKM matrix in a three generation standard model, and using the existing estimates of $`|V_{ub}|`$, the determination $`V_{ud}=0.9741\pm 0.0005`$ from superallowed nuclear $`\beta `$-decays leads instead to $`V_{us}=0.2261\pm 0.0022`$. Therefore, unitarity of the CKM matrix can only be restored at the expense of having the S$`\chi `$PT two-loop correction $`f_2`$ at least as large as the one-loop contribution $`f_1`$. This is far from signaling a failure of the chiral expansion in the present case, since $`f_1`$ might be anomalously small, being, for instance, suppressed, even compared to $`f_2`$, in the large-$`N_c`$ limit (another consequence of the Ademollo-Gatto theorem). In G$`\chi `$PT, the corresponding expansion of $`f_+^{K^0\pi ^{}}(0)`$ reads $$f_+^{K^0\pi ^{}}(0)=1+\stackrel{~}{f}_1+\stackrel{~}{f}_2+\mathrm{},$$ (5.2) where $`\stackrel{~}{f}_1`$ collects all $`𝒪(\mathrm{p}^4)`$ contributions in the generalized chiral counting, and contains $`𝒪(m_{\mathrm{quark}}^2)`$ terms. The difference between $`\stackrel{~}{f}_1`$ and $`f_1`$ $$\stackrel{~}{f}_1f_1=\left[\frac{1}{8}\widehat{m}^2\xi ^2+\frac{1}{2}\widehat{m}^2A_3\right](r1)^2\left(1+\frac{1}{R}\right),$$ (5.3) involves a single low-energy constant from $`_{(2,2)}`$, $`A_3`$, which would appear only at order $`𝒪(\mathrm{p}^6)`$ in S$`\chi `$PT. The contribution of the $`_{(2,1)}`$ low-energy constant $`\xi `$ is determined by the value of the ratio $`F_K/F_\pi `$. The value of $`V_{ud}`$ from the nuclear $`\beta `$-decay data can be accommodated by $$(r1)^2\widehat{m}^2A_3=0.124\pm 0.022,\frac{F_K}{F_\pi }=1.189\pm 0.012.$$ (5.4) The corresponding difference is $`\stackrel{~}{f}_1f_1=0.04\pm 0.01`$, and it represents the value which would be required in the Leutwyler-Roos analysis for the two-loop contribution $`f_2`$ (instead of the estimate $`f_2=0.016\pm 0.008`$ of ) in order to maintain the unitarity of the CKM matrix. Of course, a confirmation, with a comparable accuracy, from other sources (neutron decay, pion $`\beta `$-decay) of the value of $`V_{ud}`$ obtained from nuclear $`\beta `$-decays can only be welcome. This determination of $`A_3`$ and the decrease in the value of the ratio of decay constants, as compared to the number $`F_K/F_\pi =1.22\pm 0.01`$ , is compatible with the present experimental information concerning the difference in the $`K_{e3}^0`$ and $`K_{e3}^+`$ decay rates, and induces only a mild modification in the prediction for the slope of the scalar form factor $`\lambda _0`$, which, as a function of the quark mass ratio $`m_s/\widehat{m}`$, varies between 0.0018 and 0.0022, well within the range set by the high-statistics $`K_L^0`$ experiment of Donaldson et al. . The higher values of $`\lambda _0`$ obtained by some of the more recent experiments are therefore difficult to understand at the theoretical level, and cannot be ascribed, within G$`\chi `$PT, to the manifestation of a smaller value of the bilinear light quark condensate. We thus conclude that the nuclear $`\beta `$-decay determination of $`V_{ud}=0.9741\pm 0.0005`$ need not be in contradiction with the present values of the $`K_{e3}`$ decay rates and with chiral perturbation theory. One should then ask how the corresponding increase of $`|V_{us}|`$ by about 2.5 standard deviations would manifest itself in various observables. We have already mentioned that the present understanding of hyperon semi-leptonic decays is compatible with the suggested update of $`V_{us}`$. The effect on the $`K_{e4}`$ decay rates should be analyzed separately . Some effect on the extraction of the $`𝒪(\mathrm{p}^4)`$ low-energy constants $`L_1,L_2`$ and $`L_3`$ is to be expected a priori, but a precise statement requires a closer analysis. Finally, it is worth mentioning the possible effect on hadronic spectral functions which are extracted from the decays $`\tau \mathrm{hadrons}+\nu _\tau `$ and used for a determination of fundamental QCD parameters . While the non-strange ($`\overline{u}d`$) spectral functions should be only barely affected by an increase of $`0.25\%`$, the recently published strange ($`\overline{u}s`$) spectral functions should be reduced by $`4.6\%`$. Consequently, we would expect no notable influence on the determination of $`\alpha _S(M_\tau ^2)`$ , whereas the central value of the running strange quark mass $`\mathrm{m}_\mathrm{s}`$ determined recently could increase by $`1520\%`$. The issue certainly deserves a more detailed study. Acknowledgments Useful correspondence with W.S. Woolcock, W.C. Haxton and J. Comfort is kindly acknowledged. We also thank N. Vinh Mau for clarifying discussions. Appendices ## A Expansion and renormalization of the effective Lagrangian In this Appendix, we display the structure of the effective action of G$`\chi `$PT up to order $`𝒪(\mathrm{p}^4)`$. For a general discussion of the G$`\chi `$PT expansion framework, we refer the reader to the existing literature . At leading order, the generalized expansion is described by $`\stackrel{~}{}^{(2)}`$, which was first given in Ref. : $`\stackrel{~}{}^{(2)}`$ $`=`$ $`{\displaystyle \frac{1}{4}}F_0^2\{D_\mu U^+D^\mu U+2B_0U^+\chi +\chi ^+U`$ $`+A_0(U^+\chi )^2+(\chi ^+U)^2+Z_0^SU^+\chi +\chi ^+U^2`$ $`+Z_0^PU^+\chi \chi ^+U^2+H_0\chi ^+\chi \}.`$ The notation is as in Refs. , except for the consistent removal of the factor $`2B_0`$ from $`\chi `$, the parameter that collects the scalar and pseudoscalar sources, $$\chi =s+ip=+\mathrm{},=\mathrm{diag}(m_u,m_d,m_s).$$ (A.6) In G$`\chi `$PT, the next-to-leading-order corrections are of order $`𝒪(\mathrm{p}^3)`$, and still occur at tree level only. They are embodied in $`\stackrel{~}{}^{(3)}=_{(2,1)}+_{(0,3)}`$, which reads $`\stackrel{~}{}^{(3)}`$ $`=`$ $`{\displaystyle \frac{1}{4}}F_0^2\{\xi D_\mu U^+D^\mu U(\chi ^+U+U^+\chi )+\stackrel{~}{\xi }D_\mu U^+D^\mu U\chi ^+U+U^+\chi `$ $`+\rho _1(\chi ^+U)^3+(U^+\chi )^3+\rho _2(\chi ^+U+U^+\chi )\chi ^+\chi `$ $`+\rho _3\chi ^+UU^+\chi (\chi ^+U)^2(U^+\chi )^2`$ $`+\rho _4(\chi ^+U)^2+(U^+\chi )^2\chi ^+U+U^+\chi `$ $`+\rho _5\chi ^+\chi \chi ^+U+U^+\chi `$ $`+\rho _6\chi ^+UU^+\chi ^2\chi ^+U+U^+\chi +\rho _7\chi ^+U+U^+\chi ^3\}.`$ The tree-level contributions at order $`𝒪(\mathrm{p}^4)`$ are contained in <sup>4</sup><sup>4</sup>4Contributions from the odd intrinsic parity sector are also present in the effective Lagrangian; at order $`𝒪(\mathrm{p}^4)`$ they are given by the Wess-Zumino term and are the same for S$`\chi `$PT and for G$`\chi `$PT, so that we do not display them here. $$\stackrel{~}{}^{(4)}=_{(4,0)}+_{(2,2)}+_{(0,4)}+B_0^2_{(0,2)}^{}+B_0_{(2,1)}^{}+B_0_{(0,3)}^{}.$$ (A.8) The part without explicit chiral symmetry breaking, $`_{(4,0)}`$, is described by the same low-energy constants $`L_1`$, $`L_2`$, $`L_3`$, $`L_9`$ and $`L_{10}`$ as in S$`\chi `$PT : $`_{(4,0)}`$ $`=`$ $`L_1D_\mu U^+D^\mu UD_\nu U^+D^\nu U+L_2D_\mu U^+D^\nu UD_\mu U^+D^\nu U`$ $`+L_3D_\mu U^+D^\mu UD_\nu U^+D^\nu UiL_9F_{\mu \nu }^RD^\mu UD^\nu U^++F_{\mu \nu }^LD^\mu U^+D^\nu U`$ $`+L_{10}U^+F_{\mu \nu }^RUF^{L\mu \nu }+H_1F_{\mu \nu }^RF^{R\mu \nu }+F_{\mu \nu }^LF^{L\mu \nu }.`$ The new term $`_{(2,2)}`$ would count as $`𝒪(\mathrm{p}^6)`$ in S$`\chi `$PT, and is given by: $`_{(2,2)}`$ $`=`$ $`{\displaystyle \frac{1}{4}}F_0^2\{A_1D_\mu U^+D^\mu U(\chi ^+\chi +U^+\chi \chi ^+U)`$ $`+A_2D_\mu U^+U\chi ^+D^\mu UU^+\chi `$ $`+A_3D_\mu U^+U(\chi ^+D^\mu \chi D^\mu \chi ^+\chi )+D_\mu UU^+(\chi D^\mu \chi ^+D^\mu \chi \chi ^+)`$ $`+A_4D_\mu U^+D^\mu U\chi ^+\chi `$ $`+B_1D_\mu U^+D^\mu U(\chi ^+U\chi ^+U+U^+\chi U^+\chi )`$ $`+B_2D_\mu U^+\chi D^\mu U^+\chi +\chi ^+D_\mu U\chi ^+D^\mu U`$ $`+B_3U^+D_\mu \chi U^+D^\mu \chi +D_\mu \chi ^+UD^\mu \chi ^+U`$ $`+B_4D_\mu U^+D^\mu U\chi ^+U\chi ^+U+U^+\chi U^+\chi `$ $`+C_1^SD_\mu U\chi ^++\chi D^\mu U^+D_\mu U\chi ^++\chi D^\mu U^+`$ $`+C_2^SD_\mu \chi ^+U+U^+D^\mu \chi D_\mu U^+\chi +\chi ^+D^\mu U`$ $`+C_3^SD_\mu \chi ^+U+U^+D^\mu \chi D_\mu \chi ^+U+U^+D^\mu \chi `$ $`+C_1^PD_\mu U\chi ^+\chi D^\mu U^+D_\mu U\chi ^+\chi D^\mu U^+`$ $`+C_2^PD_\mu \chi ^+UU^+D^\mu \chi D_\mu U^+\chi \chi ^+D^\mu U`$ $`+C_3^PD_\mu \chi ^+UU^+D^\mu \chi D_\mu \chi ^+UU^+D^\mu \chi `$ $`+D^SD_\mu U^+D^\mu U(\chi ^+U+U^+\chi )\chi ^+U+U^+\chi `$ $`+D^PD_\mu U^+D^\mu U(\chi ^+UU^+\chi )\chi ^+UU^+\chi `$ $`+H_2D_\mu \chi ^+D^\mu \chi \}.`$ Notice that the number of counterterms (17) involved in $`_{(2,2)}`$ agrees with Refs. . However, in both cases, different bases have been used. Finally, the tree-level contributions which behave as $`𝒪(m_{\mathrm{quark}}^4)`$ in the chiral limit are contained in $`_{(0,4)}`$, $`_{(0,4)}`$ $`=`$ $`{\displaystyle \frac{1}{4}}F_0^2\{E_1(\chi ^+U)^4+(U^+\chi )^4`$ $`+E_2\chi ^+\chi (\chi ^+U\chi ^+U+U^+\chi U^+\chi )`$ $`+E_3\chi ^+\chi U^+\chi \chi ^+U`$ $`+F_1^S\chi ^+U\chi ^+U+U^+\chi U^+\chi ^2`$ $`+F_2^S(\chi ^+U)^3+(U^+\chi )^3\chi ^+U+U^+\chi `$ $`+F_3^S\chi ^+\chi (\chi ^+U+U^+\chi )\chi ^+U+U^+\chi `$ $`+F_4^S(\chi ^+U)^2+(U^+\chi )^2\chi ^+\chi `$ $`+F_1^P\chi ^+U\chi ^+UU^+\chi U^+\chi ^2`$ $`+F_2^P(\chi ^+U)^3(U^+\chi )^3\chi ^+UU^+\chi `$ $`+F_3^P\chi ^+\chi (\chi ^+UU^+\chi )\chi ^+UU^+\chi `$ $`+F_5^{SS}(\chi ^+U)^2+(U^+\chi )^2\chi ^+U+U^+\chi ^2`$ $`+F_6^{SS}\chi ^+\chi \chi ^+U+U^+\chi ^2`$ $`+F_5^{SP}(\chi ^+U)^2+(U^+\chi )^2\chi ^+UU^+\chi ^2`$ $`+F_6^{SP}\chi ^+\chi \chi ^+UU^+\chi ^2`$ $`+F_7^{SP}(\chi ^+U)^2(U^+\chi )^2\chi ^+UU^+\chi \chi ^+U+U^+\chi `$ $`+H_3\chi ^+\chi \chi ^+\chi +H_4\chi ^+\chi ^2\}.`$ Contributions from $`_{(0,4)}`$ would only appear at order $`𝒪(\mathrm{p}^8)`$ in S$`\chi `$PT, which, to the best of our knowledge, have not been discussed in the literature. The $`𝒪(\mathrm{p}^4)`$ loop corrections to the processes studied here involve only graphs with one or two vertices from $`\stackrel{~}{}^{(2)}`$: $$\stackrel{~}{𝒵}_{1\mathrm{loop}}^{(4)}=\stackrel{~}{𝒵}_{\mathrm{tadpole}}^{(4)}+\stackrel{~}{𝒵}_{\mathrm{unitarity}}^{(4)}+\mathrm{}$$ (A.12) The divergent parts of these one-loop graphs have been subtracted at a scale $`\mu `$ in the same dimensional renormalization scheme as described in . Accordingly, the low energy constants of $`_{(4,0)}`$, $`_{(2,2)}`$, and $`_{(0,4)}`$ stand for the renormalized quantities, with an explicit logarithmic scale dependence ($`X(\mu )`$ denotes generically any of these renormalized low-energy constants) $$X(\mu )=X(\mu ^{})+\frac{\mathrm{\Gamma }_X}{(4\pi )^2}\mathrm{ln}(\mu ^{}/\mu ).$$ (A.13) At order $`𝒪(\mathrm{p}^4)`$, the low-energy constants of $`\stackrel{~}{}^{(2)}`$ and $`\stackrel{~}{}^{(3)}`$ also need to be renormalized. The corresponding counterterms, however, are of order $`𝒪(B_0^2)`$ and $`𝒪(B_0)`$, respectively, and they are gathered in the three last terms of Eq. (A.8): in G$`\chi `$PT, renormalization proceeds order by order in the expansion in powers of $`B_0`$ . Alternatively, one may think of Eqs. (A) and (A) as standing for the combinations $`\stackrel{~}{}^{(2)}+B_0^2_{(0,2)}^{}`$ and $`\stackrel{~}{}^{(3)}+B_0_{(2,1)}^{}+B_0_{(0,3)}^{}`$, respectively, with the corresponding low-energy constants representing the renormalized quantities. The full list of $`\beta `$-function coefficients $`\mathrm{\Gamma }_X`$ are tabulated below. ## B Masses and Decay Constants For the reader’s convenience and for later reference, we provide, in this Appendix, the expressions of the pseudoscalar decay constants and masses at order $`𝒪(\mathrm{p}^4)`$ <sup>5</sup><sup>5</sup>5We also take this opportunity to correct a misprint in an earlier published expression of the decay constant $`F_K`$.. Introducing the notation ($`P=\pi ,K,\eta `$) $$\mu _P=\frac{M_P^2}{32\pi ^2F_\pi ^2}\mathrm{ln}\frac{M_P^2}{\mu ^2},$$ (B.1) where $`\mu `$ is the renormalisation scale, we obtain the following formulae for the decay constants (in the limit $`m_u=m_d`$), $$F_\pi ^2=F_0^2\left[1+2\widehat{m}\delta _{F,\pi }^{(2,1)}+2\widehat{m}^2\delta _{F,\pi }^{(2,2)}4\mu _\pi 2\mu _K\right],$$ (B.2) with $`\delta _{F,\pi }^{(2,1)}`$ $`=`$ $`\xi +(2+r)\stackrel{~}{\xi }`$ (B.3) $`\delta _{F,\pi }^{(2,2)}`$ $`=`$ $`A_1+{\displaystyle \frac{1}{2}}A_2+2A_3+B_1B_2`$ $`+{\displaystyle \frac{1}{2}}(2+r^2)(A_4+2B_4)+2(2+r)D^S`$ $$F_K^2=F_0^2\left[1+2\widehat{m}\delta _{F,K}^{(2,1)}+2\widehat{m}^2\delta _{F,K}^{(2,2)}\frac{3}{2}\mu _\pi 3\mu _K\frac{3}{2}\mu _\eta \right],$$ (B.4) with $`\delta _{F,K}^{(2,1)}`$ $`=`$ $`{\displaystyle \frac{1}{2}}(1+r)\xi +(2+r)\stackrel{~}{\xi }`$ (B.5) $`\delta _{F,K}^{(2,2)}`$ $`=`$ $`{\displaystyle \frac{1}{2}}(1+r^2)(A_1+A_3+B_1)+{\displaystyle \frac{r}{2}}(A_2+2A_32B_2)`$ $`+{\displaystyle \frac{1}{2}}(2+r^2)(A_4+2B_4)+(1+r)(2+r)D^S`$ $$F_\eta ^2=F_0^2\left[1+2\widehat{m}\delta _{F,\eta }^{(2,1)}+2\widehat{m}^2\delta _{F,\eta }^{(2,2)}6\mu _K\right],$$ (B.6) with $`\delta _{F,\eta }^{(2,1)}`$ $`=`$ $`{\displaystyle \frac{1}{3}}(1+2r)\xi +(2+r)\stackrel{~}{\xi }`$ (B.7) $`\delta _{F,\eta }^{(2,2)}`$ $`=`$ $`{\displaystyle \frac{1}{3}}(1+2r^2)(A_1+{\displaystyle \frac{1}{2}}A_2+2A_3+B_1B_2)`$ $`+{\displaystyle \frac{1}{2}}(2+r^2)(A_4+2B_4)+{\displaystyle \frac{2}{3}}(1+2r)(2+r)D^S`$ $`{\displaystyle \frac{4}{3}}(r1)^2(C_1^P+C_2^P)`$ For the masses, we obtain: $`F_\pi ^2M_\pi ^2`$ $`=`$ $`F_0^2\{2\widehat{m}B_0+4\widehat{m}^2A_0+4\widehat{m}^2(2+r)Z_0^S`$ $`+2\widehat{m}^3\delta _{M,\pi }^{(0,3)}+2\widehat{m}^4\delta _{M,\pi }^{(0,4)}+4\widehat{m}^2A_3M_\pi ^2`$ $`\mu _\pi [3M_\pi ^2+5M_\pi ^2\widehat{ϵ}(r)]`$ $`2\mu _K[M_\pi ^2+(1+r)M_\pi ^2\widehat{ϵ}(r)]`$ $`{\displaystyle \frac{1}{3}}\mu _\eta [M_\pi ^2+(1+2r)M_\pi ^2\widehat{ϵ}(r)2{\displaystyle \frac{\mathrm{\Delta }_{GMO}}{r1}}]\}`$ $`F_K^2M_K^2`$ $`=`$ $`F_0^2\{(1+r)\widehat{m}B_0+(1+r)^2\widehat{m}^2A_0+2(1+r)(2+r)\widehat{m}^2Z_0^S`$ $`+(1+r)\widehat{m}^3\delta _{M,K}^{(0,3)}+(1+r)\widehat{m}^4\delta _{M,K}^{(0,4)}+(1+r)^2\widehat{m}^2A_3M_K^2`$ $`{\displaystyle \frac{3}{2}}\mu _\pi [M_K^2+M_\pi ^2\widehat{ϵ}(r)]`$ $`3\mu _K[M_K^2+{\displaystyle \frac{1}{2}}(1+r)^2M_\pi ^2\widehat{ϵ}(r)]`$ $`{\displaystyle \frac{1}{6}}\mu _\eta [5M_K^2+(r+1)(1+2r)M_\pi ^2\widehat{ϵ}(r)+{\displaystyle \frac{r+1}{r1}}\mathrm{\Delta }_{GMO}]\}`$ $`F_\eta ^2M_\eta ^2`$ $`=`$ $`F_0^2\{{\displaystyle \frac{2}{3}}(1+2r)\widehat{m}B_0+{\displaystyle \frac{4}{3}}(1+2r^2)\widehat{m}^2A_0+{\displaystyle \frac{4}{3}}(1+2r)(2+r)\widehat{m}^2Z_0^S+{\displaystyle \frac{8}{3}}(1r)^2\widehat{m}^2Z_0^P`$ $`+2\widehat{m}^3\delta _{M,\eta }^{(0,3)}+2\widehat{m}^4\delta _{M,\eta }^{(0,4)}+{\displaystyle \frac{4}{3}}(1+2r^2)\widehat{m}^2A_3M_\eta ^2{\displaystyle \frac{8}{3}}(1r)^2\widehat{m}^2C_2^PM_\eta ^2`$ $`\mu _\pi \left[M_\pi ^2+(2r+1)M_\pi ^2\widehat{ϵ}(r)2{\displaystyle \frac{\mathrm{\Delta }_{GMO}}{r1}}\right]`$ $`{\displaystyle \frac{2}{3}}\mu _K\left[8M_K^23M_\pi ^2+(r+1)(2r+1)M_\pi ^2\widehat{ϵ}(r)+2{\displaystyle \frac{2r1}{r1}}\mathrm{\Delta }_{GMO}\right]`$ $`{\displaystyle \frac{1}{9}}\mu _\eta [16M_K^27M_\pi ^2+3(2r+1)^2M_\pi ^2\widehat{ϵ}(r)+4{\displaystyle \frac{4r1}{r1}}\mathrm{\Delta }_{GMO}]\},`$ with $`\delta _{M,\pi }^{(0,3)}`$ $`=`$ $`{\displaystyle \frac{9}{2}}\rho _1+{\displaystyle \frac{1}{2}}\rho _2+(10+4r+r^2)\rho _4+{\displaystyle \frac{1}{2}}(2+r^2)\rho _5+6(2+r)^2\rho _7`$ (B.11) $`\delta _{M,K}^{(0,3)}`$ $`=`$ $`{\displaystyle \frac{3}{2}}(1+r+r^2)\rho _1+{\displaystyle \frac{1}{2}}(1r+r^2)\rho _2+3(2+2r+r^2)\rho _4+{\displaystyle \frac{1}{2}}(2+r^2)\rho _5+6(2+r)^2\rho _7`$ (B.12) $`\delta _{M,\eta }^{(0,3)}`$ $`=`$ $`{\displaystyle \frac{3}{2}}(1+2r^3)\rho _1+{\displaystyle \frac{1}{6}}(1+2r^3)\rho _2+{\displaystyle \frac{8}{3}}(1+r)(1r)^2\rho _3+`$ (B.13) $`{\displaystyle \frac{1}{3}}(10+8r+17r^2+10r^3)\rho _4+{\displaystyle \frac{1}{6}}(1+2r)(2+r^2)\rho _5+`$ $`{\displaystyle \frac{8}{3}}(1r)^2(2+r)\rho _6+2(1+2r)(2+r)^2\rho _7,`$ while the $`𝒪(\mathrm{p}^4)`$ contributions are $`\delta _{M,\pi }^{(0,4)}`$ $`=`$ $`8E_1+2E_2+4(1+r)(2+r^2)F_1^S+(20+9r+r^3)F_2^S`$ $`+(4+r+r^3)F_3^S+2(2+r^2)F_4^S+4(2+r)(6+2r+r^2)F_5^{SS}`$ $`+2(2+r)(2+r^2)F_6^{SS},`$ $`\delta _{M,K}^{(0,4)}`$ $`=`$ $`2(1+r)(1+r^2)E_1+(1+r^3)E_2+{\displaystyle \frac{1}{2}}(1+r)(1r)^2E_3`$ $`+4(1r)(2+r^2)F_1^S+(8+9r+9r^2+4r^3)F_2^S`$ $`+(4r+r^2+2r^3)F_3^S+(1+r)(2+r^2)F_4^S`$ $`+4(2+r)(4+3r+2r^2)F_5^{SS}+2(2+r)(2+r^2)F_6^{SS},`$ $`\delta _{M,\eta }^{(0,4)}`$ $`=`$ $`{\displaystyle \frac{8}{3}}(1+2r^4)E_1+{\displaystyle \frac{2}{3}}(1+2r^4)E_2`$ $`+{\displaystyle \frac{8}{3}}(1+2r^2)(2+r^2)F_1^S+{\displaystyle \frac{1}{3}}(20+13r+37r^3+20r^4)F_2^S`$ $`+{\displaystyle \frac{1}{3}}(4+5r+5r^3+4r^4)F_3^S+{\displaystyle \frac{2}{3}}(1+2r^2)(2+r^2)F_4^S`$ $`+{\displaystyle \frac{16}{3}}(1+r)^2(1r)^2F_1^P+4(1r)^2(1+r+r^2)F_2^P`$ $`+{\displaystyle \frac{4}{3}}(1r)^2(1+r+r^2)F_3^P+4(2+r)(2+2r+3r^2+2r^3)F_5^{SS}`$ $`+{\displaystyle \frac{2}{3}}(2+r)(2+r^2)(1+2r)F_6^{SS}+{\displaystyle \frac{8}{3}}(1r)^2(2+r^2)F_5^{SP}`$ $`+{\displaystyle \frac{4}{3}}(1r)^2(2+r^2)F_6^{SP}+{\displaystyle \frac{16}{3}}(1r)^2(1+r)(2+r)F_7^{SP}.`$
warning/0001/quant-ph0001092.html
ar5iv
text
# 1 Introduction ## 1 Introduction Models described by one-dimensional Schrödinger equations with quasi-periodic potentials display interesting spectra: this kind of potential is intermediate between periodic ones, leading to energy bands and extended states, and truly random potentials, which cause localisation . A super-lattice e.g. may be made of two species of doped semiconductors producing a one-dimensional chain of quantum wells. A qualitative model to describe the corresponding wave-functions is given by the tight-binding model, $`\widehat{H}\psi (n)=\psi (n+1)+\psi (n1)+\lambda V(n)\psi (n),\psi l^2(𝐙)`$ , where $`V(n)`$ represents the effect of quantum well $`n`$ , and $`\lambda `$ is a positive parameter playing the role of a coupling constant . An interesting description results when this super-lattice is constructed by means of a deterministic rule. The simplest rule obtains when the two species alternate in a periodic way. But in general the rule will be aperiodic. One widely studied example is the Fibonacci sequence, which is quasi-periodic : given two ‘letters’ $`a`$ and $`b`$, one substitutes $`a\xi (a)=ab`$ and $`b\xi (b)=a`$. Iterating this rule on $`a`$ one thus generates the sequence $`abaababaabaababa\mathrm{}`$ , in which the frequency of $`a`$’s is given by the golden mean $`(5^{1/2}1)/2`$. Other examples of such substitution rules are the Thue-Morse sequence (non quasi-periodic), which is obtained through the substitution $`a\xi (a)=ab,b\xi (b)=ba`$ giving $`abbabaabbaababba\mathrm{}`$ , and the period-doubling sequence (non quasi-periodic), $`a\xi (a)=ab,b\xi (b)=aa`$ . Accordingly, a piece-wise constant potential $`\{V_n;n𝐙\}`$ based on such such a rule, e.g. $`V_1=a,V_2=b,V_3=a,\mathrm{}`$ , is called a ‘substitution potential’. The potential of the Fibonacci sequence has one-dimensional quasi-crystalline properties. In recent years problems of quantum computing (QC) and information processing have received increasing attention. To solve certain classes of problems in a potentially very powerful way, one tries to utilize in QC the quantum-mechanical superposition principle and the (non-classical) entanglement. In the models of QC based on quantum Turing machines (QTM) , the computation is characterized by sequences of unitary transformations (i.e. by the corresponding Hamiltonians $`\widehat{H}`$ acting during finite time interval steps). Benioff has studied the tight-binding model in a generalized QTM , where QC is associated with different potentials at different steps as an ‘environmental effect’. These kinds of influences may introduce deterministic disorder which would degrade performance by causing reflections at various steps and decay of the transmitted component . Here we investigate an iterative map on qubits which can be interpreted as a QTM architecture : local transformations of the Turing head controlled by a sequence of rotation angles $`\{\alpha _m\},m=1,2,\mathrm{}`$ (parameter inputs) alternate with a quantum-controlled NOT-operation (QCNOT) with a second spin on the Turing tape. Those angles at steps $`n=2m1`$ are reminiscent of the potentials $`V_m`$ introduced before. In the present paper we will investigate the Fibonacci rule and the Thue-Morse case mainly with respect to the local dynamics of the Turing head, which will be shown to reflect the degree of ‘randomness’ of the substitution sequences. The various types of aperiodic structures have been characterized up to now by the nature of their Fourier spectra only . ## 2 Quantum Turing machine driven by substitution sequences The quantum network to be considered here is composed of two pseudo-spins $`|p^{(\mu )};p=1,1;\mu =S,1`$ (Turing-head $`S`$, Turing-tape spin $`1`$, see figure 1) so that its network state $`|\psi `$ lives in the four-dimensional Hilbert space spanned by the product wave-functions $`|j^{(S)}k^{(1)}=|jk`$. Correspondingly, any (unitary) network operator can be expanded as a sum of product operators. The latter may be based on the $`SU(2)`$-generators, the Pauli matrices $`\widehat{\sigma }_j^{(\mu )},j=1,2,3`$, together with the unit operator $`\widehat{1}^{(\mu )}`$. The initial state $`|\psi _0`$ will be taken to be a product of the Turing-head and tape wave-functions. For the discretized dynamical description of the QTM we identify the unitary operators $`\widehat{U}_n,n=1,2,3,\mathrm{}`$ (step number) with the local unitary transformation on the Turing head $`S`$, $`\widehat{U}_{\alpha _m}^{(S)}`$, and the QCNOT on ($`S,1`$), $`\widehat{U}^{(S,1)}`$, respectively, as follows: $`\widehat{U}_{2m1}=\mathrm{exp}\left(i\widehat{\sigma }_1^{(S)}\alpha _m/2\right)`$ (1) $`\widehat{U}_{2m}=\widehat{U}^{(S,1)}=\widehat{P}_{1,1}^{(S)}\widehat{\sigma }_1^{(1)}+\widehat{P}_{1,1}^{(S)}\widehat{1}^{(1)}=\left(\widehat{U}^{(S,1)}\right)^{},`$ (2) where $`P_{j,j}^{(S)}=|j^{(S)}^{(S)}j|`$ is a (local) projection operator, and the Turing head is externally driven by substitution sequences $`\{\alpha _m\}`$ specified by $`\alpha _1,\alpha _2`$ . Here we restrict ourselves to the quasi-periodic Fibonacci - (qf) and Thue-Morse sequence (tm), respectively: $`\alpha _1^{\text{qf}}=\alpha _1,\alpha _2^{\text{qf}}=\alpha _2,\alpha _3^{\text{qf}}=\alpha _1,\alpha _4^{\text{qf}}=\alpha _1,\alpha _5^{\text{qf}}=\alpha _2,\mathrm{}`$ (3) $`\alpha _1^{\text{tm}}=\alpha _1,\alpha _2^{\text{tm}}=\alpha _2,\alpha _3^{\text{tm}}=\alpha _2,\alpha _4^{\text{tm}}=\alpha _1,\alpha _5^{\text{tm}}=\alpha _2,\mathrm{}.`$ (4) First, we consider the reduced state-space dynamics of the head $`S`$ and tape-spin $`1`$, respectively, $`\sigma _j^{(S)}(n)=`$ $`\text{Tr}\left(\widehat{\rho }_n^{(S)}\widehat{\sigma }_j^{(S)}\right)`$ $`=\psi _n|\widehat{\sigma }_j^{(S)}\widehat{1}^{(1)}|\psi _n,`$ $`\sigma _k^{(1)}(n)=`$ $`\text{Tr}\left(\widehat{\rho }_n^{(1)}\widehat{\sigma }_k^{(1)}\right)`$ $`=\psi _n|\widehat{1}^{(S)}\widehat{\sigma }_k^{(1)}|\psi _n,`$ (5) where $`|\psi _n`$ is the total network state at step $`n`$, and $`\sigma _j^{(\mu )}(n)`$ are the respective Bloch-vectors. Due to the entanglement between the head and tape, both will, in general, appear to be in a ‘mixed-state’, which means that the length of the Bloch-vectors in (5) is less than $`1`$. However, for specific initial states $`|\psi _0`$ the state of head and tape will remain pure: As $`|\pm ^{(1)}=\frac{1}{\sqrt{2}}\left(|1^{(1)}\pm |1^{(1)}\right)`$ are the eigenstates of $`\widehat{\sigma }_1^{(1)}`$ with $`\widehat{\sigma }_1^{(1)}|\pm ^{(1)}=\pm |\pm ^{(1)}`$, the QCNOT-operation $`\widehat{U}^{(S,1)}`$ of equation (2) cannot create any entanglement, irrespective of the head state $`|\phi ^{(S)}`$, i.e. $`\widehat{U}^{(S,1)}|\phi ^{(S)}|+^{(1)}`$ $`=`$ $`|\phi ^{(S)}|+^{(1)}`$ $`\widehat{U}^{(S,1)}|\phi ^{(S)}|^{(1)}`$ $`=`$ $`\widehat{\sigma }_3^{(S)}|\phi ^{(S)}|^{(1)}.`$ (6) As a consequence, the state $`|\psi _n`$ remains a product state for any step $`n`$ and initial product state $`|\psi _0^\pm =|\phi _0^{(S)}|\pm ^{(1)}`$ with $`|\phi _0^{(S)}=\mathrm{exp}\left(i\widehat{\sigma }_1^{(S)}\phi _0/2\right)|1^{(S)}`$ , so that the Turing head performs a pure-state trajectory (‘primitive’, see ) on the Bloch-circle $`\left(\sigma _1^{(S)}(n)=0\right)`$ $`|\psi _n^\pm =|\phi _n^\pm ^{(S)}|\pm ^{(1)},\left(\sigma _2^{(S)}(n|\pm )\right)^2+\left(\sigma _3^{(S)}(n|\pm )\right)^2=1.`$ Here $`\sigma _j^{(S)}(n|\pm )`$ denotes the Bloch-vector of the Turing head $`S`$ conditioned by the initial state $`|\psi _0^\pm `$. From the Fibonacci sequence (3) and the property (2) it is found for $`|\phi _n^+^{(S)}|+^{(1)},n=2m`$, and $`\phi _0^\pm =\alpha _0=0`$ that $$\sigma _2^{(S)}(2m|+)=\mathrm{sin}𝒞_{2m}(+),\sigma _3^{(S)}(2m|+)=\mathrm{cos}𝒞_{2m}(+),$$ (7) and $`\sigma _k^{(S)}(2m1|+)=\sigma _k^{(S)}(2m|+)`$ , where the cumulative rotation angle is $`𝒞_{2m}(+)={\displaystyle \underset{j=1}{\overset{m}{}}}\alpha _j^{\text{qf}}=\alpha _1m+(\alpha _2\alpha _1)m^{},`$ with $`m^{}(m)`$ being the total number of angles $`\alpha _2`$ up to step $`2m`$. For the cumulative rotation angle $`𝒞_n()`$ up to step $`n`$ we utilize the following recursion relations $`𝒞_{2m}()=𝒞_{2m1}(),𝒞_{2m1}()=\alpha _m^{\text{qf}}+𝒞_{2m2}().`$ Then it is easy to verify that for $`|\phi _n^{}^{(S)}|^{(1)}`$ and $`\phi _0^\pm =\alpha _0=0`$ $$\left|𝒞_n()\right|\mathrm{\hspace{0.33em}2}\mathrm{max}(|\alpha _1|,|\alpha _2|)=:M,$$ (8) yielding $`\sigma _2^{(S)}(n|)=\mathrm{sin}𝒞_n(),\sigma _3^{(S)}(n|)=\mathrm{cos}𝒞_n()`$. From any initial state, $`|\psi _0=a^{(+)}|\phi _0^+^{(S)}|+^{(1)}+a^{()}|\phi _0^{}^{(S)}|^{(1)}`$, we then obtain at step $`n`$ $`|\psi _n=a^{(+)}|\phi _n^+^{(S)}|+^{(1)}+a^{()}|\phi _n^{}^{(S)}|^{(1)}`$ and, observing the orthogonality of the $`|\pm ^{(1)}`$, $$\sigma _k^{(S)}(n)=|a^{(+)}|^2\sigma _k^{(S)}(n|+)+|a^{()}|^2\sigma _k^{(S)}(n|).$$ (9) This trajectory of the Turing-head $`S`$ represents a non-orthogonal pure-state decomposition. By using (7), (8), (9) $`\left(\text{with}a^{(+)}=a^{()}=1/\sqrt{2}\right)`$ we finally get for $`|\psi _0=|1^{(S)}|1^{(1)}`$ $`(\sigma _2^{(S)}(n),\sigma _3^{(S)}(n))=\mathrm{cos}_n(\mathrm{sin}𝒜_n,\mathrm{cos}𝒜_n),`$ (10) where $`(𝒞_n(+)M)/2𝒜_n=(𝒞_{2m}(+)+𝒞_{2m}())/2,_n=(𝒞_{2m}(+)𝒞_{2m}())/2(𝒞_{2m}(+)+M)/2`$ , $`n=2m`$ or $`2m1`$. Thus the expression (10) indicates that for the local dynamics of the Turing head in the ‘non-classical’ regime the cumulative control loss due to any small perturbation $`\delta `$ of the given $`\alpha _1^{\text{qf}},\alpha _1^{\text{qf}}`$ grows at most linearly with $`n`$, so that all periodic orbits on the plane $`\{0,\sigma _2^{(S)},\sigma _3^{(S)}\}`$ are stable (see figure 1$`a`$ \- $`c`$), as in the case of the ‘regular’ control $`\alpha _m=\alpha _1`$ . This may be contrasted with the chaotic Fibonacci-rule (cf), $`\alpha _{m+1}^{\text{cf}}=\alpha _m^{\text{cf}}+\alpha _{m1}^{\text{cf}}`$ (Lyapunov exponent: $`\mathrm{ln}(1+\sqrt{5})/2>0`$) , which can be interpreted as a temporal random (chaotic) analogue to one-dimensional chaotic potentials : each step $`\alpha _m`$ is controlled by the cumulative information of the two previous steps. For a small perturbation of the initial phase angle $`\alpha _0`$ the cumulative angles $`𝒜_m,_m`$ , respectively, grow exponentially with $`m`$, and so do the deviation terms $`\mathrm{\Delta }𝒞_{2m}^{\text{cf}}(\pm )=𝒞_{2m}^{\text{cf}^{}}(\pm )𝒞_{2m}^{\text{cf}_{\text{po}}}(\pm )`$ from the periodic orbits (po). Thus the total cumulative control loss induced by the perturbation can show chaotic quantum behaviour on the Turing head . For the Thue-Morse control (4) we easily find that for $`|\psi _0=|1^{(S)}|1^{(1)}`$ and $`n=8m`$ $`𝒞_n(+)=\mathrm{\hspace{0.33em}2}(\alpha _1+\alpha _2)m,𝒞_n()=\mathrm{\hspace{0.33em}0},`$ respectively, which is very similar to the result of the ‘regular’ machine with $`\alpha _m^{\text{reg}}=(\alpha _1+\alpha _2)/2`$, implying $`𝒞_{8m}^{\text{reg}}(+)=2(\alpha _1+\alpha _2)`$ and $`𝒞_{8m}^{\text{reg}}()=0`$. In all cases considered we thus find characteristic local invariants with respect to the Turing head (figure 1$`d`$). ## 3 Parameter sensitivity The distance between density operators, $`\widehat{\rho }`$ and $`\widehat{\rho }^{}`$, defined by the so-called Bures metric $`D_{\rho \rho ^{}}^2:=\text{Tr}\left\{(\widehat{\rho }\widehat{\rho }^{})^2\right\}.`$ lies, independent of the dimension of the Liouville space, between 0 and 2. For pure states we can rewrite $`D^2=\mathrm{\hspace{0.33em}2}(1|\psi |\psi ^{}|^2)=\mathrm{\hspace{0.33em}2}(1O^{})`$ $`O^{}:=|\psi _0(\delta )|\widehat{U}^{}(\delta )\widehat{U}(0)|\psi _0(0)|^2,`$ where the perturbed unitary evolution, $`\widehat{U}(\delta )`$, connects the initial state, $`|\psi _0(\delta )`$, and $`|\psi ^{}`$. This metric can be applied likewise to the total-network-state space or any subspace. In any case it is a convenient additional means to characterize various QTMs: for the regular case, $`\alpha _m^{\text{reg}}=\alpha _1`$ (Lyapunov exponent $`=0`$), and any initial perturbation $`\delta `$ for $`\widehat{\rho }^{}`$ the distance remains almost constant ; for the chaotic Fibonacci rule (cf), on the other hand, $`(\alpha _m(\widehat{\rho })=\alpha _m^{\text{cf}},\alpha _m(\widehat{\rho }^{})=\alpha _m(\widehat{\rho })+\delta _m^{\text{cf}}`$ , where $`\delta _m^{\text{cf}}`$ is the cumulative perturbation of the angle $`\alpha _m^{\text{cf}}`$ at step $`n=2m1)`$ we obtain an initial exponential sensitivity . In the case of the present substitution sequences we observe for a small perturbation of the given $`\alpha _1,\alpha _2`$ no initial exponential sensitivity in the evolution of $`D^2`$, which confirms that any periodic orbit is stable (figure 1$`a`$). Finally we display the evolution of $`D^2`$ for the total network state $`|\psi _n`$, which also shows no exponential sensitivity (figure 1$`b`$ , cf. figure 3$`c`$ in ). The respective distances for tape-spin $`1`$ are similar to those shown. The corresponding behaviour under the Thue-Morse control is qualitatively the same. This parameter-sensitivity has been proposed as a measure to distinguish quantum chaos from regular quantum dynamics. From the results of the present analysis and those in satisfying this criterion we conclude that, indeed, only classical chaotic input makes the quantum dynamics in QTM architectures chaotic, too (figure 1). ## 4 Summary In conclusion, we have studied the quantum dynamics of a small QTM driven by substitution sequences based on a decoherence-free Hamiltonian. As quantum features we utilized the superposition principle and the physics of entanglement. Quantum dynamics manifests itself in the superposition and entanglement of a pair of ‘classical’ (i.e. disentangled) state-sequences. The generalized QTM under this kind of control connects two fields of much recent interest, quantum computation and motion in one-dimensional structures with ‘deterministically aperiodic’ potential distributions. No chaotic quantum dynamics results in this case, as shown by the lack of exponential parameter-sensitivity. Local invariants leading to one-dimensional point manifolds (patterns) exist for $`\alpha _1=\alpha _2`$ only. For $`\alpha _1\alpha _2`$ a continuous destruction of these patterns sets in (figure 1$`b`$ ; hardly visible yet in figure 1$`a`$). This reminds us of the disappearing of KAM tori in the classical phase space resulting from a small perturbation (see e.g. ). Patterns in reduced Bloch-planes $`\{0,\sigma _2^{(\mu )},\sigma _3^{(\mu )}\}`$ (a quantum version of a Poincaré-cut) should thus be similarly useful to characterize quantum dynamics in a broad class of quantum networks. Due to the entanglement, we can see regular, chaotic, and intermediate quantum dynamics, respectively. Furthermore, the parameter sensitivity gives a sensitive criterion for testing induced quantum chaos in a pure quantum regime. This might be contrasted with the usual quantum chaology, which is concerned essentially only with semiclassical spectrum analysis of classically chaotic systems (e.g. level spacing, spectral rigidity) . It is expected that a QTM architecture with a larger number of pseudo-spins on the Turing tape would still exhibit the same type of dynamical behaviour under the corresponding driving conditions. ## 5 Acknowledgements We thank J. Gemmer, A. Otte, P. Pangritz, and F. Tonner for fruitful discussions. One of us (I. K.) is grateful to Prof. P. L. Knight, Dr. M. B. Plenio, and their co-workers at Imperial College for their hospitality and to the European Science Foundation (Quantum information theory and quantum computation) for financial support during his visit. Figure 1: Input-output-scheme of our quantum Turing machine (QTM). Figure 1: Turing-head patterns $`\{0,\sigma _2(n),\sigma _3(n)\}`$ for initial state $`|\psi _0=|1^{(S)}`$$`|1^{(1)}`$ under the control of substitution sequences: ($`a`$) quasi-periodic Fibonacci (qf) with $`\alpha _1=\frac{2}{5}\pi ,\alpha _2=\alpha _1+0.0005\pi `$ , ($`b`$) as in $`a`$) but for $`\alpha _2=\alpha _1+0.03\pi `$ , ($`c`$) as in $`a`$) but for $`\alpha _2=\alpha _1+0.05\pi `$ ; ($`d`$) Thue-Morse (tm) control with $`\alpha _1=\frac{2}{5}\pi ,\alpha _2=\alpha _1+0.1001\pi `$ . For each simulation the total step number is $`n=10000`$. Figure 1: Evolution of the (squared) distance $`D_{\rho \rho ^{}}^2`$ between the perturbed, $`\widehat{\rho }^{}`$ , and the reference QTM state, $`\widehat{\rho }`$ , under the quasi-periodic Fibonacci (qf) control: ($`a`$) for Turing head; ($`b`$) for total network state $`|\psi _n`$ . For $`\widehat{\rho }`$ we take $`|\psi _0=|1^{(S)}|1^{(1)}`$ and $`\alpha _1=\frac{2}{5}\pi ,\alpha _2=\alpha _1+0.03\pi `$ . Line A: $`|\psi _0^{}=\mathrm{exp}\left(i\widehat{\sigma }_1^{(S)}\delta /2\right)|\psi _0`$ for $`\widehat{\rho }^{}`$ , $`\delta =0.001`$. Line B: $`|\psi _0^{}=|\psi _0`$ , but $`\alpha _1^{}=\alpha _1+0.001\pi ,\alpha _2^{}=\alpha _2+0.001\pi `$ for $`\left(\widehat{\rho }^{}\right)`$ .
warning/0001/cond-mat0001335.html
ar5iv
text
# The Physics of the Glass Transition ## 1 Introduction Glasses are roughly speaking liquids that do not crystallize also at very low temperature (to be precise: glasses also do not quasi-crystallize). These liquids can avoid crystalization mainly for two reasons: * The liquid does not crystallize because it is cooled very fast: the crystallisation time may become very large at low temperature (e.g. hard spheres at high pressure). The system should be cooled very fast at temperatures near the melting point; however if crystalization is avoided, and the temperature is low enough, (e.g. near the glass transition) the system may be cooled very slowly without producing crystalization. * The liquid does not crystallize even at equilibrium. An example is a binary mixture of hard spheres with different radius: 50% type A (radius $`r_A`$, 50% type B (radius $`r_A`$, where $`R`$ denotes $`r_B/r_A`$. If $`.77<R<.89`$ (the bounds may be not precise), the amorphous packing is more dense than a periodic packing, distorted by defects. Which of the two mechanism is present is irrelevant for understanding the liquid glass transition. Other examples of glassy systems are mixtures: in this case we have $`M`$ kind of particles and the Hamiltonian of $`N`$ particles is given by $$H=\underset{a,b=1,M}{}\underset{i=1,N(a)}{}\underset{k=1,N(b)}{}V_{a,b}\left(x_a(i)x_b(k)\right),$$ (1) where $`a=1,M`$; $`N(a)=Nc(a)`$; $`_ac(a)=1`$. In this case the values of the $`N`$ concentrations $`c`$ and the $`M(M+1)/2`$ functions $`V_{a,b}(x)`$ describe the model. Well studied case are: * Soft spheres (non-realistic, but simple), e.g. $`M=2`$, $`c(1)=.5`$, $`c(2)=.5`$, $`V_{a,b}(x)=R_{a,b}x^{12}`$. * Lennard-Jones spheres (more realistic), e.g. $`M=2`$, $`c(1)=.8`$, $`c(2)=.2`$, $`V_{a,b}(x)=R_{a,b}x^{12}A_{a,b}x^6`$. There are some choices of the parameters which are well studied and in these case it is known that the system does not crystallize, one of these has been introduced by Kob and Anderson in the L-J case and correspond to a particular choice of the parameters $`R`$ and $`A`$. There are many other material which are glass forming, e.g. short polymers, asymmetric molecules (e.g. OTP) … The behaviour of the viscosity in glass forming liquid is very interesting. We recall that the viscosity can be defined microscopically as follows. We consider a system in a box of large volume $`V`$ and we call $`T_{\mu ,\nu }(t)`$ the total stress tensor at time $`t`$. We define the correlation function of the stress tensor at different times: $$T_{\mu ,\nu }(t)T_{\rho ,\sigma }(0)=VS_{\mu ,\nu ,\rho ,\sigma }(t)$$ (2) One finds, neglecting indices, that $`\eta 𝑑tS(t)\tau ^\alpha `$, where $`\tau `$ is the characteristic time of the system (in a first approximation we can suppose that $`\alpha =1`$). In fragile glasses the behaviour of the viscosity as function of the temperature can be phenomenological described by the two following regimes: * In an high temperature region the mode coupling theory is valid: it predicts $`\eta (TT_c)^\gamma `$, where $`\gamma `$ is not an universal quantity and it a of the orders of a few units. * In the low temperature region by the Vogel Fulcher law is satisfied: it predicts that $`\eta \mathrm{exp}(A(TT_K)^1)`$. Nearly tautologically fragile glasses can be defined as those glasses that have $`T_K0`$; strong glasses have $`T_K0`$. The glass temperature ($`T_g`$) is the temperature at which the viscosity becomes so large that it cannot be any more measured. This happens after an increase of about 18 order of magnitude (which correspond to a microscopic time changing from $`10^{14}`$ to $`10^4`$ seconds): the relaxation time becomes larger than the experimental tine. A characteristic of glasses is the dependance of the specific heat on the cooling rate. There is a (slightly rounded) discontinuity in the specific heat that it is shifted at lower temperatures when we increase the cooling time. For systems that do crystallize if cooled very slowly, one can plot $`\mathrm{\Delta }SS(liquid)S(crystal)`$ versus $`T`$ in the thermalized region. One gets a very smooth curve whose extrapolation becomes negative at a finite temperature. A negative $`\mathrm{\Delta }S`$ does not make sense, so there is a wide spread belief that a phase transition is present before (and quite likely near) the point where the entropy becomes negative. Such a thermodynamic transition (suggested by Kaufmann) would be characterized by a jump of the specific heat. Quite often this temperature is very similar to the temperature found by fitting the viscosity by the Volker Fulcher law and the two temperature are believed to coincide. Generally speaking it is expected that a thermodynamic phase transition induces a divergent correlation time. However the exponential dependance of the correlation time on the temperature is not so common (in conventional critical slowing down we should have a power like behaviour); moreover an other strange property of the glass transition is the apparent absence of a detectable correlation length or susceptibility (linear or non-linear) which diverges when we approach the transition point. In section 2, I will firstly present the many valley picture which has been the inspiration point of many works; in the subsequent section I will summarize some of the results that have been obtained using the replica theory on the organization of these valley in phase space, on the temperature dependence of the free energy in each valley, in brief of the free energy landscape. A strategy for applying these ideas to glasses will be presented in section 4 and in the next section I will show some results that have been obtained in the case of a Lennard-Jones binary mixture. ## 2 The many valleys picture A quite old statement is that the glass is frozen liquid, which I prefer to reformulate it as that a low temperature a liquid is an unfrozen glass. The general idea is the following: the phase space of the systems can be approximately decomposed into valleys, labelled by $`a`$, such that the barrier among these valleys is large in the liquid (in the low temperature region where the viscosity is large) the barriers become infinite at $`T_K`$ . For simplicity I will assume that valleys do maintain their identity when changing the temperature. (in reality they split into smaller valleys when we decrease the temperature ). Under this assumption there is a one to one correspondence among valleys and inherent structures, i.e. minima of the Hamiltonian $`H`$. The properties of the liquid are therefore connected to the properties of the system near the minima of the Hamiltonian. The partition function can be approximately written as $$Z=\underset{a}{}\mathrm{exp}\left(\beta Nf_a(\beta )\right),$$ (3) where $`f_a(\beta )`$ is the free energy density of the $`a^{th}`$ valley. As we shall see later in the infinite volume limit the previous sum is dominated by those valleys having a given free energy density $`f=f^{}(\beta )`$. The number of relevant valleys at $`f=f^{}(\beta )`$ (i.e. $`𝒩^{}(\beta )`$) is supposed to exponentially increase with the size ($`N`$) of the system: $$cN^{}\mathrm{exp}(N\mathrm{\Sigma }^{}(\beta )).$$ (4) It can be argued that the configurational entropy or complexity $`\mathrm{\Sigma }^{}(\beta )`$ is approximately near to $`\mathrm{\Delta }S`$: indeed it has suggest long time ago that the configurational entropy $`\mathrm{\Sigma }^{}`$, goes to zero at the phase transition point . Recently using the techniques stemming from replica theory we have added new ingredients to this old scenario: * We have construct a microscopic realization of the previous ideas in the mean field models . In other words there are soluble infinite range models in which there is an exponentially large number of valleys and their properties can be computed analytically. * If we assume that minimal corrections are present in finite dimensions to mean field predictions, the mean field results may be extended to three dimensional glasses. * The replica method gives tools to do the appropriate computations, both in mean the mean field approach and in short range models. As outcome of these advances a many valleys picture with detailed predictions on the landscape has been constructed and the properties of glasses can be computed in the framework of replicated liquid theory . ## 3 The free energy landscape Let us suppose that we can define a temperature free energy functional ($`F[\rho ]`$) as function of the density $`\rho (x)`$. It is natural to assume that the different valleys are in a first approximation associated to different local minima of this functional, the free energy evaluated at the local minima being the free energy of the corresponding valley. I will summarize which are in mean field the properties of the local minima the free energy that have been computed explicitly in some long range microscopic models. Although these results have been obtained in some particular models, one can show that they are valid for a quite large class of models (at least in the mean field approximation ). * For $`T>T_c`$ there is one relevant minimum with $`\rho (x)=const`$. There are many solutions of the equation $$\frac{\delta F}{\delta \rho }=0.$$ (5) which are not minima (there are some negative eigenvalues of the Hessian $`(x,y)=\frac{^2H}{xy}`$. * For $`T=T_c`$ the negative eigenvalues of the Hessian become positive or zero. There is an exponentially large number of minima, connected by flat regions. * For $`T_c>T>T_K`$ the number of relevant minima becomes proportional to $`\mathrm{exp}(N\mathrm{\Sigma }^{}(\beta ))`$. where $`\mathrm{\Sigma }^{}(\beta )`$ is the configurational entropy or complexity. The minima are separated by barriers that diverge with the number of particles $`N`$ (in mean field theory), however the barriers are are finite in real life (i.e. beyond mean field theory). * For $`T_K>T`$ the number of relevant minima is finite. The minima are separated by barriers that diverge with $`N`$. For $`T_c>T>T_K`$ there is a dual description: the system may be described as a liquid and also as a solid (with an exponentially large number of different solid structures). This duality tell us that all the information needed to study the glass can be extracted from the liquid phase. ## 4 A more quantitative approach As we have seen we can write $$Z(\beta )=\underset{a}{}\mathrm{exp}(\beta Nf_a(\beta ))=𝑑𝒩(f,\beta )\mathrm{exp}(\beta Nf),$$ (6) where $`f_a(\beta )`$ is the free energy density of the valley labeled by $`a`$ at the temperature $`\beta ^1`$, and $`𝒩(f,\beta )`$ is the number of valleys with free energy density less than $`f`$. We can assume that $`𝒩(f,\beta )=\mathrm{exp}(N\mathrm{\Sigma }(f,\beta ))`$, where the configurational entropy, or complexity, $`\mathrm{\Sigma }(f,\beta )`$ is positive in the region $`f.f_0(\beta )`$ and vanishes at $`f=f_0(\beta )`$. The quantity $`f_0(\beta )`$ is the minimum value of the free energy: $`𝒩(f,\beta )`$ is zero for $`f<f_0(\beta )`$ . If the equation $$\frac{\mathrm{\Sigma }}{f}=\beta $$ (7) has a solution at $`f=f^{}(\beta )`$ (obviously this may happens only for $`f^{}(\beta )>f_0(\beta )`$), we stay in the liquid phase the free density is given by $$F=f^{}\beta ^1\mathrm{\Sigma }(f^{},\beta )$$ (8) and $`\mathrm{\Sigma }(f^{},\beta )=\mathrm{\Sigma }^{}(\beta )`$. Otherwise we stay in the glass phase and $$F=f_0(\beta ).$$ (9) In order to compute the properties in the glass phase we need to know $`\mathrm{\Sigma }(f,\beta )`$: a simple strategy is the following. We introduce the modified partition function $$Z(\gamma ;\beta )\mathrm{exp}(N\gamma \mathrm{\Phi }(\gamma ;\beta ))=\underset{a}{}\mathrm{exp}(\gamma Nf_a(\beta )).$$ (10) Using standard thermodynamical arguments it can be easily proven that in the limit $`N\mathrm{}`$ one has: $`\gamma \mathrm{\Phi }(\gamma ;\beta )=\gamma f\mathrm{\Sigma }(\beta ,f),f={\displaystyle \frac{(\gamma \mathrm{\Phi }(\gamma ;\beta ))}{\gamma }}.`$ The complexity is obtained from $`\mathrm{\Phi }(\gamma ;\beta )`$ in the same way as the entropy is obtained from the usual free energy : $$\mathrm{\Sigma }(\beta ,f)=\frac{\mathrm{\Phi }(\gamma ;\beta )}{\gamma }.$$ (11) A few observations are in order: * In the new formalism $`\gamma `$, the free energy and the complexity play respectively the same role of $`\beta `$, the internal energy and the entropy in the usual formalism. * In the new formalism $`\beta `$ only indicates the value of the temperature which is used to compute the free energy and $`\gamma `$ controls which part of the free energy landscape is sampled. * When $`\beta \mathrm{}`$ we sample the energy landscape: $$Z(\gamma ;\mathrm{})=\underset{a}{}\mathrm{exp}(\gamma Ne_a)=\nu (e)𝑑e\mathrm{exp}(\gamma Ne)$$ (12) where $`e_a`$ are the minima of the Hamiltonian and $`\nu (e)`$ the density of the minima of the Hamiltonian. * The equilibrium complexity is given by $`\mathrm{\Sigma }^{}(\beta )=\mathrm{\Sigma }(\beta ;\beta )`$. On the other hand $`\mathrm{\Sigma }(\gamma ;\mathrm{})`$ give us information on the minima of the Hamiltonian. It is convenient to write $`Z(\gamma ;\beta )={\displaystyle 𝑑C\mathrm{exp}\left(\gamma H(C)N\gamma f(\beta ,C)+N\gamma f(\gamma ,C)\right)}=`$ $`{\displaystyle 𝑑C\mathrm{exp}\left(\gamma H(C)N\gamma \widehat{f}(\beta ,C)+N\gamma \widehat{f}(\gamma ,C)\right)},`$ (13) where $`\widehat{f}(\beta ,C)=f(\beta ,C)f(\mathrm{},C)`$, $`f(\beta ,C)`$ is is constant in each valley and it is equal to the free energy density of the valley to which the configuration $`C`$ belongs. and $`\widehat{f}(\beta ,C)`$ is the temperature dependent part of free energy density of the valley to which the configuration $`C`$ belongs. Our strategy is to use duality and liquid theory to evaluate the properties of $`\widehat{f}(\beta ,C)`$. A simple approximation is the so called quenched approximation which in this case consists in assuming that $$\mathrm{exp}(A\widehat{f}(\beta ,C))\mathrm{exp}(A\widehat{f}_\gamma (\beta )),$$ (14) where $`\widehat{f}_\gamma (\beta )`$ is the expectation value of $`\widehat{f}(\beta ,C)`$ taken with the probability distribution proportional to $`\mathrm{exp}(\gamma H(C))`$. The quenched approximation would be exact if all the valleys at a given temperature (also with different free energy) would have the same entropy. In other words the fluctuations of the energy are more relevant than the fluctuations of the entropy. It is now a matter of simple algebra to show that $`\mathrm{\Phi }(\gamma ;\beta )=F_{liquid}(\gamma )+\widehat{f}_\gamma (\beta )\widehat{f}_\gamma (\gamma ),`$ $`\mathrm{\Sigma }(\gamma ;\beta )=S_{liquid}(\gamma )S_\gamma (\gamma )+\widehat{f}_\gamma ^{}(\gamma )\widehat{f}_\gamma ^{}(\beta ),`$ (15) where $`\widehat{f}_\gamma ^{}(\beta )=\widehat{f}_\gamma (\beta )/\gamma `$. Everything is now computable using liquid theory. Different approximations can be done for $`\widehat{f}_\gamma (\beta )`$ and slightly different results may be obtained for the transition temperature and for the specific heat in the glass region. The replica formalism allow us to study directly the function $`\mathrm{\Phi }(\gamma ;\beta )`$ and most of the computation are done (for sake of compactness and simplicity) using this powerful tool. ## 5 Some results Analytic and numerical computation have been done in many case. I will present here only the results for the Kob Andersen Lennard Jones binary mixture described in the introduction at density $`\rho =1.2`$. In order to compute the free energy (or the entropy of the valleys) a simple and useful approximation is to assume that the potential near each minimum of the Hamiltonian is approximately an harmonic one and therefore the free energy of a valley can be computed using an harmonic approximation. In this way one finds that the previous formulae in the liquid phase become $$\mathrm{\Sigma }^{}(\beta ))=S_{liquid}(\beta )S_{harmonic}(\beta )+\mathrm{},$$ (16) where $$\beta S_{harmonic}=\frac{1}{2}\mathrm{ln}det\left(\frac{^2\beta H}{x(i)x(k)}\right).$$ (17) Similar formulae holds below the transition temperature. The relevant quantities can be or computed analytically (using liquid theory) or extracted by numerical simulations in the liquid phase. Both the analytic and the numerical results of for the configurational are shown in fig. 4 (from ). The correctness of the numerical estimate have been later confirmed by the more accurate computations of . These microscopic analytic computations support the proposal that there is a thermodynamic liquid glass transition characterized by vanishing of the complexity and that at low temperature below $`T_K`$ the system stays only in a few valleys. In other words in the glass phase replica symmetry is broken and we aspect that there are characteristic violations of the fluctuation dissipation theorem in off-equilibrium dynamics in the aging regime . Indeed in the aging regime the fluctuation dissipation theorem cannot be applied and generalized dissipation relations are valid. The form of these generalized dissipation relations is fixed by the theory and the resulting predictions are confirmed by numerical simulations. Direct measurements of these fluctuation dissipation relations in real (not numerical ) experiments both for glasses are possible with the present technology. Some experiments are under the way. The outcome will be a crucial test for this theoretical approach and I hope that some results will be available in the next future.
warning/0001/math0001160.html
ar5iv
text
# 1 Introduction ## 1 Introduction There are 3 generalized Kac-Moody algebras or superalgebras which represent the physical states of a string moving on a certain variety, namely the monster algebra, the fake monster algebra \[B\] and the the fake monster superalgebra \[S\]. The no-ghost theorem from string theory can be used to construct actions of finite groups on these algebras. For example the monster group acts on the monster algebra. Applying elements of the finite groups to the Weyl denominator identity of these algebras gives twisted denominator identities. They can be calculated explicitly because the simple roots of the algebras are known. For the monster algebra and the fake monster algebra this has been done in \[B\]. There Borcherds uses twisted denominator identities of the monster algebra to prove the moonshine conjectures. In this paper we calculate twisted denominator identities of the fake monster superalgebra and use them to construct new examples of supersymmetric generalized Kac-Moody superalgebras. They have similar properties as the fake monster superalgebra. For example they have no real roots and the Weyl vector is zero. Their denominator identities give new infinite product expansions. In a forthcoming paper we show that they define automorphic forms of singular weight. We describe the sections of this paper. In the second section we recall some facts about the fake monster superalgebra and describe the action of an extension of the Weyl group of $`E_8`$ on this algebra. In the third section we derive the general expression of the twisted denominator identity corresponding to elements in this group of odd order. In the last section we calculate the twisted denominator identities explicitly for certain elements of order 3 and 7. We construct two supersymmetric generalized Kac-Moody superalgebras of rank 6 and 4 and describe their simple roots and the multiplicities. ## 2 The fake monster superalgebra In this section we recall some results about the fake monster superalgebra from \[S\] and construct an action of $`2^8.2.Aut(E_8)`$ on it. The main difference to the bosonic case described in \[B\] is that we have to work with a double cover of the automorphism group of the light cone lattice rather than with the automorphism group of that lattice. The fake monster superalgebra $`G`$ can be constructed as the space of physical states of a chiral N$`=`$1 superstring moving on a 10 dimensional torus. $`G`$ is a generalized Kac-Moody superalgebra. The root lattice of $`G`$ is the 10 dimensional even unimodular Lorentzian lattice $`II_{9,1}=E_8II_{1,1}`$. A nonzero element $`\alpha II_{9,1}`$ is a root of $`G`$ if and only if $`\alpha ^20`$. In particular $`G`$ has no real roots. This reflects the fact that the superstring has no tachyons. The multiplicity of a root $`\alpha `$ is given by $`mult_0(\alpha )=mult_1(\alpha )=c(\alpha ^2/2)`$ where $`c(n)`$ is the coefficient of $`q^n`$ in $`8\eta (q^2)^8/\eta (q)^{16}=8+128q+1152q^2+7680q^3+42112q^4+\mathrm{}`$ and $`\eta (q)`$ is the Dedekind eta function. There are 2 cones of negative norm vectors in $`II_{9,1}`$. We define one of them as the positive cone and denote the closure of the positive cone by $`II_{9,1}^+`$. Then the positive roots of $`G`$ are the nonzero vectors in $`II_{9,1}^+`$ and the simple roots of $`G`$ are the positive roots of zero norm. The simple roots have multiplicity 8 as even and odd roots. The Cartan subalgebra of $`G`$ is isomorphic to the vector space generated by $`II_{9,1}`$. Since $`G`$ has no real roots the Weyl group is trivial. The Weyl vector of $`G`$ is zero. We describe some results about the 8 dimensional real spin group. For more details confer \[J\]. Let $`𝕆`$ be the real 8 dimensional algebra of octonions, i.e. the unique real alternative division algebra of dimension 8. $`𝕆`$ has a quadratic form $`N`$ permitting composition. For $`b𝕆`$ define the left multiplication $`L_b`$, the right multiplication $`R_b`$ and the operator $`U_b=L_bR_b=R_bL_b`$. Let $`𝕆_0`$ be the orthogonal complement of $`1`$ and denote $`C(𝕆,N)`$ the Clifford algebra generated by $`𝕆`$ with relations $`a^2=N(a)1`$. There is an isomorphism $`\epsilon `$ from the even subalgebra $`C^e(𝕆,N)`$ of $`C(𝕆,N)`$ to the Clifford algebra $`C(𝕆_0,N)`$ mapping $`1a`$, $`a𝕆`$, to $`a`$, where we have identified the unit in $`C(𝕆_0,N)`$ with the unit in $`𝕆`$. $`C(𝕆_0,N)`$ acts naturally on $`𝕆`$ by left and right multiplication. An element $`u`$ of the spin group $`\mathrm{\Gamma }_0^e(𝕆,N)`$ can be written as $`u=1b_1\mathrm{}1b_n`$ with $`N(b_i)=1`$ and $`\epsilon (u)=b_1\mathrm{}b_n`$. The actions $`\rho _L(u)=L_{b_1}\mathrm{}L_{b_n},\rho _R(u)=R_{b_1}\mathrm{}R_{b_n}`$ and $`\rho _V(u)=U_{b_1}\mathrm{}U_{b_n}`$ give three irreducible and inequivalent 8 dimensional representations of the spin group called conjugate spinor, spinor and vector representation. They are related by triality, i.e. $`\rho _V(u)(ab)=(\rho _L(u)a)(\rho _R(u)b)`$ holds for all $`a,b𝕆`$. The image of $`\mathrm{\Gamma }_0^e(𝕆,N)`$ under each of these representations is $`SO(8)`$. The kernel of $`\rho _V:\mathrm{\Gamma }_0^e(𝕆,N)SO(8)`$ is $`\{1,1\}`$ so that $`\mathrm{\Gamma }_0^e(𝕆,N)`$ is a double cover of $`SO(8)`$. The representations $`\rho _L,\rho _R`$ and $`\rho _V`$ of the spin group induce representations of the Lie algebra $`so(8)`$ with weights $`\frac{1}{2}(\pm 1,\mathrm{},\pm 1)`$ with an odd number of $``$ signs, $`\frac{1}{2}(\pm 1,\mathrm{},\pm 1)`$ with an even number of $``$ signs and the permutations of $`(\pm 1,0,\mathrm{},0)`$. We embed $`E_8`$ into $`𝕆`$. Let $`Aut(E_8)`$ be the group of automorphisms of $`E_8`$ leaving the bilinear form invariant. Then $`Aut(E_8)SO(8)`$ and the inverse image of $`Aut(E_8)`$ under $`\rho _V`$ is a double cover of $`Aut(E_8)`$. $`Aut(E_8)`$ can also be constructed using the ring of integral octonions (cf. \[C\]). This description implies that $`E_8𝕆`$ is also invariant under the actions $`\rho _L`$ and $`\rho _R`$. Now we construct an action of an extension of the double cover $`2.Aut(E_8)`$ on the fake monster superalgebra. The extension $`2^8.Aut(E_8)`$ of $`Aut(E_8)`$ by $`Hom(E_8,_2)`$ acts naturally on the vertex algebra $`V_{E_8}`$ of the lattice $`E_8`$. The same holds for the extension of $`2.Aut(E_8)`$ by $`2^8`$ where $`2.Aut(E_8)`$ acts by $`\rho _V`$. The vector space $`E_8[t^1]t^{\frac{1}{2}}`$ is an abelian subalgebra of the Heisenberg algebra $`E_8[t,t^1]t^{\frac{1}{2}}`$. The exterior algebra $`V_{NS}`$ of $`E_8[t^1]t^{\frac{1}{2}}`$ is a vertex superalgebra carrying a representation of the Virasoro algebra. $`V_{NS}`$ decomposes into eigenspaces of $`L_0`$ with eigenvalues in $`\frac{1}{2}`$. $`2.Aut(E_8)`$ acts on $`V_{NS}`$ in the vector representation. We define the vertex superalgebra $`V_0=V_{NS}V_{E_8}`$. This algebra carries a representation of the Virasoro algebra of central charge $`4+8=12`$. We write $`V_{0,n}`$ for the subspace of $`L_0`$-degree $`n\frac{1}{2}`$. The vector space $`E_8[t^1]t^1`$ is an abelian subalgebra of the Heisenberg algebra $`E_8[t,t^1]`$. We define $`V_R`$ as the tensor product of the exterior algebra of $`E_8[t^1]t^1`$ with the sum $`SC`$ of two $`8`$ dimensional spaces. The double cover of $`Aut(E_8)`$ acts in the vector representation on the first tensor factor and in the spinor resp. conjugate spinor representation on the second factor. $`V_R`$ can be given the structure of a $`V_{NS}`$-module. We decompose $`V_R=V_R^+V_R^{}`$ where $`V_R^+`$ is the subspace generated by vectors $`d_{n_1}\mathrm{}d_{n_k}v`$ where $`v`$ is in $`C`$ if $`k`$ is even and in $`S`$ if $`k`$ is odd and analogous for $`V_R^{}`$. The projection of $`V_R`$ on $`V_R^+`$ is called GSO projection. We define $`V_1=V_R^+V_{E_8}`$ and adopt the same notations as for $`V_0`$. $`V_1`$ carries a representation the Virasoro algebra of central charge $`12`$. Now $`2^8.2.Aut(E_8)`$ acts on the fake monster superalgebra in the following way. We decompose the Cartan subalgebra $`II_{9,1}`$ by writing $`II_{9,1}=E_8II_{1,1}`$. $`2.Aut(E_8)`$ acts in the vector representation on the vector space generated by $`E_8`$. Since $`G`$ is graded by $`II_{9,1}=E_8II_{1,1}`$ it is also graded by $`II_{1,1}`$. We denote the corresponding spaces $`G_a`$ with $`aII_{1,1}`$ and the even resp. odd subspace $`G_{0,a}`$ resp. $`G_{1,a}`$. All these spaces are $`E_8`$-graded. By the no-ghost theorem (cf. \[GSW\],\[P\] or \[B\]) the even subspace $`G_{0,a}`$ is isomorphic to $`V_{0,(1a^2)/2}`$ and $`G_{1,a}`$ is isomorphic to $`V_{1,(1a^2)/2}`$ as $`E_8`$-graded $`2^8.2.Aut(E_8)`$-module. This gives us a natural action of $`2^8.2.Aut(E_8)`$ on the fake monster superalgebra. ## 3 The twisted denominator identities In this section we calculate twisted denominator identities of the fake monster superalgebra corresponding to elements in $`2.Aut(E_8)`$ of odd order. The fake monster superalgebra $`G`$, like any generalized Kac-Moody superalgebra, can be written as direct sum $`EHF`$ where $`H`$ is the Cartan subalgebra and $`E`$ and $`F`$ are the subalgebras corresponding to the positive and negative roots. We have the standard sequence $$\mathrm{}\mathrm{\Lambda }^2(E)\mathrm{\Lambda }^1(E)\mathrm{\Lambda }^0(E)0$$ with homology groups $`H_i(E)`$. Note that $`E=E_0E_1`$ is a superspace so that the exterior algebra $`\mathrm{\Lambda }(E)`$ is defined as the tensor algebra of $`E`$ divided by the two sided ideal generated by $`uv+(1)^{|u||v|}vu`$. The Euler-Poincaré principle implies $$\mathrm{\Lambda }^{}(E)=H(E)$$ where $`\mathrm{\Lambda }^{}(E)=_{n0}(1)^n\mathrm{\Lambda }^n(E)`$ is the alternating sum of exterior powers of $`E`$ and $`H(E)=_{n0}(1)^nH_n(E)`$ is the alternating sum of homology groups of $`E`$. Both sides of this identity are graded by the root lattice $`II_{9,1}`$ of $`G`$ and the homogeneous subspaces are finite dimensional. The homology groups $`H_n(E)`$ can be calculated in the same way as for Kac-Moody algebras (cf. \[B\]). The result is that $`H_n(E)`$ is the subspace of $`\mathrm{\Lambda }^n(E)`$ spanned by the homogeneous vectors of $`\mathrm{\Lambda }^n(E)`$ of degree $`\alpha II_{9,1}`$ with $`\alpha ^2=0`$. We can work out the homology groups of $`G`$ explicitly because we know the simple roots. Denote the subspace of $`H(E)`$ with degree $`\alpha II_{9,1}`$ by $`H(E)_\alpha `$ and analogous for $`E`$. Let $`\lambda `$ be a primitive norm zero vector in $`II_{9,1}^+`$. Then $`_{n>0}H(E)_{n\lambda }=\mathrm{\Lambda }^{}(_{n>0}E_{n\lambda })`$. The denominator identity of the fake monster superalgebra now follows easily by calculating the character on both sides of $`\mathrm{\Lambda }^{}(E)=H(E)`$. More generally if $`g`$ is an automorphism of $`G`$ then $`g`$ commutes with the derivations $`d_i:\mathrm{\Lambda }^i(E)\mathrm{\Lambda }^{i1}(E)`$ and we get a sequence $$\mathrm{}g(\mathrm{\Lambda }^2(E))g(\mathrm{\Lambda }^1(E))g(\mathrm{\Lambda }^0(E))0.$$ $`g`$ induces an isomorphism between the homology groups of the two complexes above. Hence we can apply $`g`$ to both sides of the equation $`\mathrm{\Lambda }^{}(E)=H(E)`$ and take the trace. This gives a twisted denominator identity. It depends only on the conjugacy class of $`g`$ in the automorphism group of $`G`$. Let $`u`$ be an element of $`2.Aut(E_8)`$ of odd order $`N`$. The lattice $`E_8`$ has a unique central extension $`\widehat{E}_8`$ by $`\{1,1\}`$ such that the commutator of any inverse images of $`r,v`$ in $`E_8`$ is $`(1)^{(r,v)}`$. Since $`\rho _V(u)`$ has odd order it has a lift to $`Aut(\widehat{E}_8)=2^8.Aut(E_8)`$ such that $`\rho _V(u)^n`$ fixes all elements of $`\widehat{E}_8`$ which are in the inverse image of the vectors of $`E_8`$ fixed by $`\rho _V(u)^n`$ (cf. Lemma 12.1 in \[B\]). We define $`g`$ as the automorphism of $`G`$ induced by this lift. Let $`E_8^u`$ be the sublattice of $`E_8`$ fixed by $`\rho _V(u)`$. The natural projection $`\pi :E_8E_8^u`$ maps $`E_8`$ onto the dual lattice $`E_8^u`$ because $`E_8`$ is unimodular. We define the Lorentzian lattice $`L=E_8^uII_{1,1}`$ with dual lattice $`L^{}=E_8^uII_{1,1}`$ and denote the closures of the canonical positive cones by $`L^+`$ and $`L^+`$. The fake monster superalgebra has a natural $`L^{}`$-grading. For $`\alpha =(r^{},a)L^{}`$ we define $$\stackrel{~}{E}_{0,\alpha }=_{\pi (r)=r^{}}E_{0,(r,a)}$$ and analogous for $`\stackrel{~}{E}_{1,\alpha }`$. Let $`\epsilon _i`$ and $`\sigma _i`$ denote the eigenvalues of $`\rho _V(u)`$ and $`\rho _L(u)`$. Then we have ###### Theorem 3.1 The twisted denominator identity corresponding to g is given by $$\underset{\alpha L^+}{}\frac{(1e^\alpha )^{mult_0(\alpha )}}{(1+e^\alpha )^{mult_1(\alpha )}}=1+a(\lambda )e^\lambda $$ where $`\text{mult}_0(\alpha )`$ $`=`$ $`{\displaystyle \underset{ds|((\alpha ,L),N)}{}}{\displaystyle \frac{\mu (s)}{ds}}tr(g^d|\stackrel{~}{E}_{0,\alpha /ds})`$ $`\text{mult}_1(\alpha )`$ $`=`$ $`{\displaystyle \underset{ds|((\alpha ,L),N)}{}}{\displaystyle \frac{\mu (s)}{ds}}tr(g^d|\stackrel{~}{E}_{1,\alpha /ds})`$ and $`a(\lambda )`$ is the coefficient of $`q^m`$ in $$\underset{n1}{}\underset{\mathrm{\hspace{0.25em}1}i8}{}\frac{(1\epsilon _iq^n)}{(1+\sigma _iq^n)}$$ if $`\lambda `$ is $`m`$ times a primitive norm zero vector in $`L^+`$ and zero else. Proof: We consider both sides of $`\mathrm{\Lambda }^{}(E)=H(E)`$ as $`L^{}`$-graded $`g`$-modules. $`\mathrm{\Lambda }^{}(E)`$ is isomorphic to $`\mathrm{\Lambda }^{}(E_0)S^{}(E_1)`$ if we forget the superstructure on $`E_1`$. First we calculate the trace of $`g`$ on $`S^{}(E_1)`$. For that we recall some formulas. Let $`V`$ be a finite dimensional vector space. Then we have $$\underset{n0}{}(1)^n(dimS^n(V))q^n=(1+q)^{dimV}=exp\left\{\underset{n>0}{}(1)^n(dimV)q^n/n\right\}.$$ The second equality can be proven by taking logarithms and using the expansion $`log(1+q)=_{n>0}(1)^nq^n/n`$. This can be generalized to $$\underset{n0}{}(1)^ntr(g|S^n(V))q^n=exp\left\{\underset{n>0}{}(1)^ntr(g^n|V)q^n/n\right\}.$$ Another formula we need is $`S^{}(V_1V_2)=S^{}(V_1)S^{}(V_2)`$. Using these two formulas we find that the trace of $`g`$ on $`S^{}(E_1)`$ is given by $$exp\left\{\underset{\beta L^+}{}\underset{n>0}{}(1)^ntr(g^n|\stackrel{~}{E}_{1,\beta })e^{n\beta }/n\right\}.$$ We want to express the trace in the form $$\underset{\beta L^+}{}(1+e^\beta )^{mult_1(\beta )}=\text{exp}\left\{\underset{\beta L^+}{}\underset{n>0}{}(1)^nmult_1(\beta )e^{n\beta }/n\right\}.$$ Taking logarithms and comparing coefficients at $`e^\alpha `$ this implies $$\underset{\genfrac{}{}{0pt}{}{\beta L^+}{n\beta =\alpha }}{}(1)^ntr(g^n|\stackrel{~}{E}_{1,\beta })/n=\underset{\genfrac{}{}{0pt}{}{\beta L^+}{n\beta =\alpha }}{}(1)^nmult_1(\beta )/n.$$ The vector $`\alpha L^{}`$ is a positive multiple $`m`$ of a primitive vector in $`L^{}`$. This vector generates a lattice that we can identify with $``$. Then the right hand side of the last equation is the convolution product of the arithmetic functions $`h(n)=(1)^n/n`$ and $`mult_1(n)`$. Hence $$mult_1(\alpha )=\underset{ds|m}{}h^1(s)(1)^dtr(g^d|\stackrel{~}{E}_{1,\alpha /ds})/d.$$ Let $`\mu `$ be the Möbiusfunction and $`f(n)`$ the arithmetic function which is zero if $`n`$ contains an odd square and $`(1)^k/p_1\mathrm{}p_k`$, where the $`p_i`$ are the different primes dividing $`n`$, else. Then convolution inverse of $`h(n)`$ is given by $$h^1(n)=\{\begin{array}{cc}(\mu h)(n)& n\text{odd}\hfill \\ f(n)& n\text{even}\hfill \end{array}$$ $`g`$ has the same order as $`u`$ because $`N`$ is odd. The trace $`tr(g^d|\stackrel{~}{E}_{1,\alpha })`$ depends only on $`\alpha `$ and $`(d,N)`$. This implies that the sum in the expression for $`mult_1(\alpha )`$ extends only over $`ds|(m,N)`$. It is easy to see that $`m`$ is equal to the highest common factor $`(\alpha ,L)`$ of the numbers $`(\alpha ,\beta ),\beta L`$. Hence we get the following formula $$mult_1(\alpha )=\underset{ds|((\alpha ,L),N)}{}\mu (s)tr(g^d|\stackrel{~}{E}_{1,\alpha /ds})/ds.$$ Note that this formula is wrong for even $`N`$. This is another reason why we restrict to elements in $`2.Aut(E_8)`$ of odd order. If we express the contributions of $`\mathrm{\Lambda }^{}(E_0)`$ to the trace in the form $$\underset{\alpha L^+}{}(1e^\alpha )^{mult_0(\alpha )}$$ then we find $$mult_0(\alpha )=\underset{ds|((\alpha ,L),N)}{}h^1(s)tr(g^d|\stackrel{~}{E}_{0,\alpha /ds})/d$$ with $`h(n)=1/n`$. This function is strongly multiplicative so that its convolution inverse is given by $`\mu h`$. An argument as above then gives the expression for $`mult_0(\alpha )`$. Next we calculate the trace of $`g`$ on $`H(E)`$. We have $$H(E)=_{\alpha L^+}\stackrel{~}{H}(E)_\alpha $$ with $$\stackrel{~}{H}(E)_\alpha =_{\pi (r)=r^{}}H(E)_{(r,a)}$$ for $`\alpha =(r^{},a)`$ in $`L^+`$. Clearly $`tr(g|\stackrel{~}{H}(E)_\alpha )`$ is zero if $`\alpha `$ is not in $`L`$ and $`tr(g|\stackrel{~}{H}(E)_\alpha )=tr(g|H(E)_\alpha )`$ for $`\alpha L`$. We can also restrict to $`\alpha ^2=0`$. Let $`\alpha L^+II_{9,1}^+`$ be m times a primitive norm zero vector $`\lambda `$ in $`II_{9,1}^+`$. Then $`\lambda `$ is also a primitive vector in $`L`$. $`H(E)_\alpha `$ is determined by $$_{n>0}H(E)_{n\lambda }=\mathrm{\Lambda }^{}(_{n>0}E_{0,n\lambda })S^{}(_{n>0}E_{1,n\lambda }).$$ g acts by $`\rho _V(u)`$ on $`E_{0,n\lambda }E_8`$ and by $`\rho _L(u)`$ on $`E_{1,n\lambda }C`$ (cf. section 2). Going over to complexifications this implies that $`tr(g|H(E)_\alpha )`$ is given by the coefficient of $`q^m`$ in $$1+\underset{n1}{}tr(g|H(E)_{n\lambda })q^n=\underset{n1}{}\underset{\mathrm{\hspace{0.25em}1}i8}{}\frac{(1\epsilon _iq^n)}{(1+\sigma _iq^n)}.$$ This finishes the proof of the theorem. For $`u=1`$ the theorem gives the denominator identity of the fake monster superalgebra. If the multiplicities are nonnegative integers then the twisted denominator identity is the untwisted denominator identity of a generalized Kac-Moody superalgebra with the following properties. The root lattice is $`L^{}`$ or a sublattice thereof. The algebra has no real roots so that the Weyl group is trivial. The multiplicities of a root $`\alpha `$ are given by $`mult_0(\alpha )`$ and $`mult_1(\alpha )`$. There are no real simple roots and the imaginary simple roots are the norm zero vectors in $`L^+`$. The Weyl vector is zero. We describe the multiplicities of the simple roots. Let $`\alpha =n\lambda `$ be a simple root where $`\lambda `$ is a primitive vector in $`L^+`$. Suppose that $`\rho _V(u)`$ and $`\rho _L(u)`$ have cycle shapes $`a_1^{b_1}\mathrm{}a_k^{b_k}`$ and $`c_1^{d_1}\mathrm{}c_l^{d_l}`$. Then the multiplicities of $`\alpha `$ as even and odd root are given by $`_{a_k|n}b_k`$ and $`_{c_k|n}d_k`$. The formulas for the multiplicities simplify if $`u`$ has in addition prime order. In this case we only need to calculate the trace of $`g`$ on $`\stackrel{~}{E}_{0,\alpha }`$ and $`\stackrel{~}{E}_{1,\alpha }`$ with $`\alpha L^{}`$ and the dimensions of these spaces. We find ###### Proposition 3.2 For $`\alpha L^{}`$ the trace $`tr(g|\stackrel{~}{E}_{0,\alpha })`$ is given by the coefficient of $`q^{(1\alpha ^2)/2}`$ in $$\frac{1}{2}\left(\underset{n1}{}\underset{\mathrm{\hspace{0.25em}1}i8}{}\frac{(1+\epsilon _iq^{n1/2})}{(1\epsilon _iq^n)}\underset{n1}{}\underset{\mathrm{\hspace{0.25em}1}i8}{}\frac{(1\epsilon _iq^{n1/2})}{(1\epsilon _iq^n)}\right)$$ if $`\alpha L`$ and zero else. Let $`tr(\rho _L(u))=tr(\rho _R(u))`$. Then $`tr(g|\stackrel{~}{E}_{1,\alpha })`$ is the coefficient of $`q^{(1\alpha ^2)/2}`$ in $$tr(\rho _L(u))q^{1/2}\underset{n1}{}\underset{\mathrm{\hspace{0.25em}1}i8}{}\frac{(1+\epsilon _iq^n)}{(1\epsilon _iq^n)}$$ if $`\alpha L`$ and zero else. Proof: Clearly the trace of $`g`$ is zero on $`\stackrel{~}{E}_{0,\alpha }`$ if $`\alpha `$ is not in $`L`$. For $`\alpha L`$ we have $`tr(g|\stackrel{~}{E}_{0,\alpha })=tr(g|E_{0,\alpha })`$. Write $`\alpha =(r,a)`$. Then $`E_{0,\alpha }`$ is isomorphic as $`g`$-module to the subspace of $`V_{0,(1a^2)/2}`$ of degree $`r`$ (cf. section 2). This space is generated by products of fermionic and bosonic oscillators and $`e^r`$. The sum of the $`L_0`$-contribution of the oscillators and the $`L_0`$-contribution of $`e^r`$ is $`\frac{1}{2}\frac{1}{2}a^2`$. The vector $`e^r`$ has $`L_0`$-eigenvalue $`\frac{1}{2}r^2`$ so that the $`L_0`$-contribution of the oscillators is $`\frac{1}{2}\frac{1}{2}a^2\frac{1}{2}r^2=\frac{1}{2}\frac{1}{2}\alpha ^2`$. Now we go over to complexifications and choose a basis of $`E_8`$ in which $`\rho _V(u)`$ is diagonal. Then the trace of $`g`$ on $`E_{0,\alpha }`$ is given by the coefficient of $`q^{(1\alpha ^2)/2}`$ in $$\underset{n1}{}\underset{\mathrm{\hspace{0.25em}1}i8}{}\frac{(1+\epsilon _iq^{n1/2})}{(1\epsilon _iq^n)}.$$ Since we only need the half integral exponents of $`q`$ in this expression we can subtract the integral exponents. This proves the first statement. The argument for the second statement is similar. Note that there are two types of ground states on which $`u`$ may act differently. We avoid this problem by assuming that $`u`$ has in both representations the same trace. This proves the proposition. ###### Proposition 3.3 Let $`\alpha =(r^{},a)L^{}`$ and let $`r^{}E_8^u`$ such that $`r^{}+r^{}E_8`$. Then the dimension of $`\stackrel{~}{E}_{0,\alpha }`$ and $`\stackrel{~}{E}_{1,\alpha }`$ is given by the coefficient of $`q^{(1\alpha ^2)/2}`$ in $$8q^{1/2}\frac{\eta (q^2)^8}{\eta (q)^{16}}\theta _{r^{}+E_8^u}(q)$$ where $`\theta _{r^{}+E_8^u}(q)`$ is the theta function of the translated lattice $`r^{}+E_8^u`$. Proof: Clearly $`\stackrel{~}{E}_{0,\alpha }`$ and $`\stackrel{~}{E}_{1,\alpha }`$ have the same dimension. The inverse image of $`r^{}`$ in $`E_8`$ under $`\pi `$ is $`r^{}+(r^{}+E_8^u)`$. Hence $`\stackrel{~}{E}_{1,\alpha }=_{\pi (r)=r^{}}E_{1,(r,a)}`$ is isomorphic to the direct sum of the $`V_{1,(1a^2)/2}(r^{}+s)`$ where $`s`$ is in the translated lattice $`r^{}+E_8^u`$. As above the sum of the $`L_0`$-contribution of the bosonic and fermionic oscillators and $`\frac{1}{2}s^2`$ is $`\frac{1}{2}\frac{1}{2}\alpha ^2`$. The proposition now follows from simple counting. We will use the shorter notation $`\theta _r^{}(q)`$ in the following. ## 4 Two supersymmetric algebras In this section we calculate explicitly twisted denominator identities corresponding to elements in $`2.Aut(E_8)`$ of order 3 and 7. These identities are the untwisted denominator identities of 2 supersymmetric generalized Kac-Moody superalgebras. We choose an orthonormal basis $`\{e_0,e_1,\mathrm{},e_7\}`$ of $`^8`$ and embed the lattice $`E_8`$ as the set of points $`m_ie_i`$ where all $`m_i`$ are in $``$ or all $`m_i`$ are in $`+\frac{1}{2}`$ and $`m_i`$ is even. In these coordinates the automorphism group of $`E_8`$ is generated by the permutations of the coordinates, even sign changes and an involution generated by the Hadamard matrix. We also identify $`^8`$ with the alternative algebra $`𝕆`$ of octonions by defining $`e_0`$ as the identity and $`e_ie_j=a_{ijk}e_k\delta _{ij}1`$ for $`1i,j7`$ where $`a_{ijk}`$ is the totally antisymmetric tensor with $`a_{ijk}=1`$ for $`ijk=123,154,264,374,176,257,365`$. The element $`u=\frac{1}{4}1(e_2e_3)1(e_1e_2)1(e_6e_7)1(e_5e_6)`$ in $`2.Aut(E_8)`$ has order $`3`$. It is easy to check that the transformations corresponding to $`\rho _V(u),\rho _L(u)`$ and $`\rho _R(u)`$ are all equal and $$\begin{array}{cccc}\rho _V(u)1=1\hfill & \rho _V(u)e_1=e_3\hfill & \rho _V(u)e_2=e_1\hfill & \rho _V(u)e_3=e_2\hfill \\ \rho _V(u)e_4=e_4\hfill & \rho _V(u)e_5=e_7\hfill & \rho _V(u)e_6=e_5\hfill & \rho _V(u)e_7=e_6.\hfill \end{array}$$ Hence $`\rho _V(u),\rho _L(u)`$ and $`\rho _R(u)`$ all have cycle shape $`1^23^2`$. The following proposition collects some results on $`E_8^u`$. ###### Proposition 4.1 The sublattice $`E_8^u`$ of $`E_8`$ fixed by $`\rho _V(u)`$ is the 4 dimensional lattice with elements $`(m_1,m_2,m_3,m_4)`$ where all $`m_i`$ are in $``$ or all $`m_i`$ are in $`+\frac{1}{2}`$ and $`m_i`$ is even. The norm is $`(m_1,m_2,m_3,m_4)^2=m_1^2+m_2^2+3m_3^2+3m_4^2`$. $`E_8^u`$ has determinant $`3^2`$ and the quotient $`E_8^u/E_8^u`$ is $`_3\times _3`$. The level of $`E_8^u`$ is $`3`$ so that $`3E_8^u`$ is a sublattice of $`E_8^u`$. The orthogonal complement of $`E_8^u`$ in $`E_8`$ is $`A_2A_2`$. We recall some results about modular forms. The congruence subgroup of $`SL_2()`$ of level $`3`$ is defined as $`\mathrm{\Gamma }(3)=\{\left(\genfrac{}{}{0pt}{}{a}{c}\genfrac{}{}{0pt}{}{b}{d}\right)SL_2()|\left(\genfrac{}{}{0pt}{}{a}{c}\genfrac{}{}{0pt}{}{b}{d}\right)=\left(\genfrac{}{}{0pt}{}{1}{0}\genfrac{}{}{0pt}{}{0}{1}\right)\text{ mod }3\}`$. The vector space of modular forms for $`\mathrm{\Gamma }(3)`$ with even positive weight $`n`$ has dimension $`n+1`$. A modular form in this vector space is zero if and only if the coefficients of $`q^0,q^{1/3},\mathrm{},q^{n/3}`$ in its Fourier expansion are zero. For $`rA_2^{}A_2^{}`$ we define $`\delta (r)=1`$ if $`rA_2A_2`$ and $`0`$ else. We have ###### Proposition 4.2 Let $`rA_2^{}A_2^{}`$. Then $$\theta _{r+A_2A_2}(q)=\frac{1}{4}\frac{\eta (q)^{12}}{\eta (q^2)^6}\left\{\frac{\eta (q^6)^2}{\eta (q^3)^4}\delta (r)+\underset{j=0}{\overset{2}{}}\epsilon ^{3jr^2/2}\frac{\eta \left((\epsilon ^jq^{1/3})^2\right)^2}{\eta \left(\epsilon ^jq^{1/3}\right)^4}\right\}$$ where $`\epsilon =e^{2\pi i/3}`$. Proof: $`A_2A_2`$ has dimension $`4`$ and level $`3`$ so that $`\theta _{r+A_2A_2}`$ is a modular form for $`\mathrm{\Gamma }(3)`$ of weight $`2`$. The same holds for the right hand side of the formula which can be shown by calculating the transformations under a set of generators. The proposition follows from comparing the coefficients at $`q^0,q^{1/3}`$ and $`q^{2/3}`$. The next identity is a twisted version of Jacobi’s identity. ###### Proposition 4.3 Let $`|q|<1`$. Then $$\frac{1}{2q^{1/2}}\left\{\underset{n1}{}(1+q^{3n3/2})^2(1+q^{n1/2})^2\underset{n1}{}(1q^{3n3/2})^2(1q^{n1/2})^2\right\}$$ $$=2\underset{n1}{}(1+q^{3n})^2(1+q^n)^2.$$ Proof: This is an identity between modular forms and therefore can be proven by comparing sufficiently many coefficients in their Fourier expansions. The proposition implies that the generating functions for $`tr(g|\stackrel{~}{E}_{0,\alpha })`$ and $`tr(g|\stackrel{~}{E}_{1,\alpha })`$ are equal. Define $`c(n)`$ by $$\underset{n0}{}c(n)q^n=2\underset{n1}{}\frac{(1+q^{3n})^2(1+q^n)^2}{(1q^{3n})^2(1q^n)^2}=2\frac{\eta (q^6)^2\eta (q^2)^2}{\eta (q^3)^4\eta (q)^4}$$ $$=2+8q+24q^2+72q^3+184q^4+432q^5+984q^6+2112q^7+\mathrm{}$$ Let $`g`$ be the automorphism induced by $`u`$. Then we have ###### Theorem 4.4 The twisted denominator identity corresponding to $`g`$ is $$\underset{\alpha L^+}{}\frac{(1e^\alpha )^{c(\alpha ^2/2)}}{(1+e^\alpha )^{c(\alpha ^2/2)}}\underset{\alpha L^+3L^{}}{}\frac{(1e^\alpha )^{c(\alpha ^2/6)}}{(1+e^\alpha )^{c(\alpha ^2/6)}}=1+a(\lambda )e^\lambda $$ where $`a(\lambda )`$ is the coefficient of $`q^n`$ in $$\underset{n1}{}\frac{(1q^{3n})^2(1q^n)^2}{(1+q^{3n})^2(1+q^n)^2}=14q+4q^24q^3+20q^424q^5+4q^6\mathrm{}$$ if $`\lambda `$ is $`n`$ times a primitive norm zero vector in $`L^+`$ and zero else. Proof: Recall that $`3L^{}L`$ because $`L`$ has level 3. Furthermore $`3`$ divides $`(\alpha ,L)`$ if and only if $`\alpha 3L^{}`$. We consider now 4 cases. $`\alpha L`$. Then $`\alpha 3L^{}`$ and by Proposition 3.2 the multiplicities $`mult_0(\alpha )`$ and $`mult_1(\alpha )`$ are zero. $`\alpha L`$ and $`\alpha 3L^{}`$. Then $`mult_0(\alpha )=tr(g|\stackrel{~}{E}_{0,\alpha })`$ and $`mult_1(\alpha )=tr(g|\stackrel{~}{E}_{1,\alpha })`$. Using Propositions 3.2 and 4.3 we find $`mult_0(\alpha )=mult_1(\alpha )=c(\alpha ^2/2)`$. $`\alpha 3L^{}`$ and $`\alpha 3L`$. Then $`mult_0(\alpha )=tr(g|\stackrel{~}{E}_{0,\alpha })+tr(1|\stackrel{~}{E}_{0,\alpha /3})/3`$. The first term gives $`c(\alpha ^2/2)`$. Write $`\alpha /3=(r^{},a)`$ where $`r^{}`$ is in $`E_8^u`$ but not in $`E_8^u`$ and choose $`r^{}`$ as in Proposition 3.3. Then the dimension of $`\stackrel{~}{E}_{0,\alpha /3}`$ is the coefficient of $`q^{\alpha ^2/18}`$ in $`8\eta (q^2)^8\theta _r^{}(q)/\eta (q)^{16}`$ or equivalently the coefficient of $`q^{\alpha ^2/6}`$ in $`8\eta (q^6)^8\theta _r^{}(q^3)/\eta (q^3)^{16}`$. Note that $`\alpha ^26`$. We have $`\alpha ^2/9=r^2=r^2`$ mod $`2`$ and $`\alpha ^2/6=3r^2/2`$ mod $`3`$ so that by Proposition 4.2 $$\theta _r^{}(q^3)=\frac{1}{4}\frac{\eta (q^3)^{12}}{\eta (q^6)^6}\underset{j=0}{\overset{2}{}}\epsilon ^{j\alpha ^2/6}\frac{\eta \left((\epsilon ^jq)^2\right)^2}{\eta \left(\epsilon ^jq\right)^4}.$$ The coefficient of $`q^{\alpha ^2/6}`$ in $`8\eta (q^6)^8\theta _r^{}(q^3)/\eta (q^3)^{16}`$ is equal to the coefficient of $`q^{\alpha ^2/6}`$ in $$6\frac{\eta (q^6)^2\eta (q^2)^2}{\eta (q^3)^4\eta (q)^4}.$$ This shows that $`mult_0(\alpha )=c(\alpha ^2/2)+c(\alpha ^2/6)`$. The result for $`mult_1(\alpha )`$ is clear. $`\alpha 3L`$. Then $`mult_0(\alpha )=tr(g|\stackrel{~}{E}_{0,\alpha })tr(g|\stackrel{~}{E}_{0,\alpha /3})/3+tr(1|\stackrel{~}{E}_{0,\alpha /3})/3`$. Here we have an additional term in $`\theta _r^{}`$ which cancels exactly with the term from $`tr(g|\stackrel{~}{E}_{0,\alpha /3})`$ so that we get the same result for $`mult_0(\alpha )`$ as in the case before. The result for $`mult_1(r)`$ is again clear. This proves the theorem. Since the multiplicities are all nonnegative integers there is a generalized Kac-Moody superalgebra whose denominator identity is the identity given in the theorem. ###### Corollary 4.5 There is a generalized Kac-Moody superalgebra with root lattice $`L`$ and root multiplicities given by $$mult_0(\alpha )=mult_1(\alpha )=c(\alpha ^2/2)\alpha L,\alpha 3L^{}$$ and $$mult_0(\alpha )=mult_1(\alpha )=c(\alpha ^2/2)+c(\alpha ^2/6)\alpha 3L^{}.$$ The simple roots of are the norm zero vectors in $`L^+`$. Their multiplicities as even and odd roots are equal. Let $`\lambda `$ be a simple root. Then $`mult_0(\lambda )=4`$ if $`\lambda `$ is 3n times a primitive vector in $`L^+`$ and $`mult_0(\lambda )=2`$ else. As the fake monster superalgebra this algebra is supersymmetric and has no real roots. The Weyl group is trivial and the Weyl vector is zero. The supersymmetry is a consequence of the twisted Jacobi identity in Proposition 4.3. In a forthcoming paper we show that the denominator function of this algebra defines an automorphic form for a discrete subgroup of $`O_{6,2}()`$ of weight $`2`$. Now we consider the next example. The analysis is similar to the above one. The element $`u=\frac{1}{8}1(e_6e_7)1(e_5e_6)1(e_4e_5)1(e_3e_4)1(e_2e_3)1(e_1e_2)`$ in $`2.Aut(E_8)`$ has order $`7`$. The transformations corresponding to $`\rho _V(u),\rho _L(u)`$ and $`\rho _R(u)`$ are all equal and $$\begin{array}{cccc}\rho _V(u)1=1\hfill & \rho _V(u)e_1=e_7\hfill & \rho _V(u)e_2=e_1\hfill & \rho _V(u)e_3=e_2\hfill \\ \rho _V(u)e_4=e_3\hfill & \rho _V(u)e_5=e_4\hfill & \rho _V(u)e_6=e_5\hfill & \rho _V(u)e_7=e_6.\hfill \end{array}$$ Hence $`\rho _V(u),\rho _L(u)`$ and $`\rho _R(u)`$ have cycle shape $`1^17^1`$. We have ###### Proposition 4.6 The sublattice $`E_8^u`$ of $`E_8`$ fixed by $`\rho _V(u)`$ is the 2 dimensional lattice with elements $`(m_1,m_2)`$, where either $`m_1`$ and $`m_2`$ are in $``$ and $`m_1+m_2`$ is even or $`m_1`$ and $`m_2`$ are in $`+\frac{1}{2}`$ and $`m_1+m_2`$ is odd, and norm $`(m_1,m_2)^2=m_1^2+7m_2^2`$. The quotient $`E_8^u/E_8^u`$ is $`_7`$ and $`E_8^u`$ has level $`7`$. The orthogonal complement of $`E_8^u`$ in $`E_8`$ is isomorphic to $`A_6`$. The theta function $`\theta _{r+A_6}`$ of a coset $`r+A_6`$ of $`A_6`$ in its dual depends only on $`r^2`$ mod $`2`$. ###### Proposition 4.7 Let $`rA_6^{}`$. Then $$\theta _{r+A_6}(q)=\frac{1}{8}\frac{\eta (q)^{14}}{\eta (q^2)^7}\left\{\frac{\eta (q^{14})}{\eta (q^7)^2}\delta (r)+\underset{j=0}{\overset{6}{}}\epsilon ^{7jr^2/2}\frac{\eta \left((\epsilon ^jq^{1/7})^2\right)^2}{\eta \left(\epsilon ^jq^{1/7}\right)^4}\right\}$$ where $`\epsilon =e^{2\pi i/7}`$. The following supersymmetry relation holds. ###### Proposition 4.8 Let $`|q|<1`$. Then $$\frac{1}{2q^{1/2}}\left\{\underset{n1}{}(1+q^{7n7/2})(1+q^{n1/2})\underset{n1}{}(1q^{7n7/2})(1q^{n1/2})\right\}$$ $$=\underset{n1}{}(1+q^{7n})(1+q^n)$$ Here we define the numbers $`c(n)`$ by $$\underset{n0}{}c(n)q^n=\underset{n1}{}\frac{(1+q^{7n})(1+q^n)}{(1q^{7n})(1q^n)}=\frac{\eta (q^{14})\eta (q^2)}{\eta (q^7)^2\eta (q)^2}$$ $$=1+2q+4q^2+8q^3+14q^4+24q^5+40q^6+66q^7+\mathrm{}$$ Write $`g`$ for the automorphism induced by $`u`$. Then ###### Theorem 4.9 The twisted denominator identity corresponding to $`u`$ is $$\underset{\alpha L^+}{}\frac{(1e^\alpha )^{c(\alpha ^2/2)}}{(1+e^\alpha )^{c(\alpha ^2/2)}}\underset{\alpha L^+7L^{}}{}\frac{(1e^\alpha )^{c(\alpha ^2/14)}}{(1+e^\alpha )^{c(\alpha ^2/14)}}=1+a(\lambda )e^\lambda $$ where $`a(\lambda )`$ is the coefficient of $`q^n`$ in $$\underset{n1}{}\frac{(1q^{7n})(1q^n)}{(1+q^{7n})(1+q^n)}=12q+2q^42q^7+4q^82q^94q^{11}+6q^{16}\mathrm{}$$ if $`\lambda `$ is $`n`$ times a primitive norm zero vector in $`L^+`$ and zero else. Again the multiplicities are all nonnegative integers and we have ###### Corollary 4.10 There is a generalized Kac-Moody superalgebra with root lattice $`L`$ and root multiplicities given by $$mult_0(\alpha )=mult_1(\alpha )=c(\alpha ^2/2)\alpha L,\alpha 7L^{}$$ and $$mult_0(\alpha )=mult_1(\alpha )=c(\alpha ^2/2)+c(\alpha ^2/14)\alpha 7L^{}.$$ The simple roots of are the norm zero vectors in $`L^+`$. Let $`\lambda `$ be a simple root. Then $`mult_0(\lambda )=mult_1(\lambda )=2`$ if $`\lambda `$ is 7n times a primitive vector in $`L^+`$ and $`mult_0(\lambda )=mult_1(\lambda )=1`$ else. The denominator identity of this algebra is given by the identity in the above theorem. It defines an automorphic form for a subgroup of $`O_{4,2}()`$ of weight $`1`$. ## Acknowledgments I thank R. E. Borcherds for stimulating discussions.
warning/0001/nucl-ex0001003.html
ar5iv
text
# Untitled Document Magnetic trapping of neutrons P. R. Huffman<sup>∗†</sup>, C. R. Brome, J. S. Butterworth<sup>∗‡</sup>, K. J. Coakley<sup>§</sup>, M. S. Dewey, S. N. Dzhosyuk, R. Golub, G. L. Greene, K. Habicht, S. K. Lamoreaux, C. E. H. Mattoni, D. N. McKinsey, F. E. Wietfeldt<sup>∗†</sup>, & J. M. Doyle <sup>3</sup><sup>3</sup>footnotetext: Present address: Institut Laue-Langevin, B.P. 156-6 rue Jules Horowitz, 38042, Grenoble (Cedex 9), France. $``$ Harvard University, 17 Oxford Street, Cambridge, Massachusetts 02138, USA. $``$ National Institute of Standards and Technology, 100 Bureau Drive, MS 8461, Gaithersburg, Maryland 20899, USA. $`\mathrm{\S }`$ National Institute of Standards and Technology, 325 Broadway, MS 898.02, Boulder, Colorado 80303, USA. $``$ Hahn-Meitner-Institut, Glienicker Straße 100, D-14109 Berlin, Germany. $`\mathrm{}`$ University of California, Los Alamos National Laboratory, P.O. Box 1663, Los Alamos, New Mexico 87545, USA. Accurate measurement of the lifetime of the neutron (which is unstable to beta decay) is important for understanding the weak nuclear force and the creation of matter during the Big Bang. Previous measurements of the neutron lifetime have mainly been limited by certain systematic errors; however, these could in principle be avoided by performing measurements on neutrons stored in a magnetic trap. Neutral and charged particle traps are widely used tool for studying both composite and elementary particles, because they allow long interaction times and isolation from perturbing environments. Here we report the magnetic trapping of neutrons. The trapping region is filled with superfluid <sup>4</sup>He, which is used to load neutrons into the trap and as a scintillator to detect their decay. Neutrons have a lifetime in the trap of 750$`{}_{\mathrm{𝟐𝟎𝟎}}{}^{}{}_{}{}^{+\mathrm{𝟑𝟑𝟎}}`$ seconds, mainly limited by their beta decay rather than trap losses. Our experiment verifies theoretical predictions regarding the loading process and magnetic trapping of neutrons. Further refinement of this method should lead to improved precision in the neutron lifetime measurement. A more precisely known value of the neutron lifetime will benefit both weak interaction and Big Bang Nucleosynthesis (BBN) theory. Combined with the measured neutron beta-decay asymmetry, the neutron lifetime yields the $`V_{ud}`$ element of the Cabibbo-Kobayashi-Maskawa (CKM) quark mixing matrix. Non-unitarity of the CKM matrix would indicate physics beyond the Standard Model. In addition, the theoretical uncertainty in the predicted <sup>4</sup>He abundance in the BBN model is dominated by the uncertainty in the neutron lifetime. Many experiments, using different methods, have measured the neutron beta-decay lifetime. The most precise measurements use ultracold neutrons (UCN, neutrons with energies $`E/k_b1`$ mK, where $`k_b`$ is Boltzmann’s constant) confined in a ‘bottle’ made of materials with high UCN reflectivity. The largest uncertainties are due to UCN interactions with the material walls and fluctuations in the initial filling densities. Extrapolation methods are necessary to extract the neutron beta-decay lifetime. In another experiment, neutrons with specific trajectories and in a band of velocities (typically between 10–14 m s<sup>-1</sup>) were kept in a magnetic storage ring similar to those used in high-energy particle physics. Since storage rings only confine neutrons in two dimensions, coupling between the large longitudinal momentum and small radial momentum by way of betatron oscillations allows neutrons to escape. In contrast, neutrons in magnetic traps are not susceptible to wall losses or betatron oscillations. Static magnetic traps are formed by creating a magnetic field minimum in free space. The confining potential depth ($`D`$) of such a trap is determined by the magnetic moment of the trapped species ($`\mu `$) and the difference ($`\mathrm{\Delta }B`$) between the magnitude of the field at the edge of the trap and at the minimum, $`D=\mu \mathrm{\Delta }B`$. A particle in a low-field-seeking state (one with its magnetic moment anti-parallel with the local magnetic field vector) is pushed towards the trap minimum. Low-field-seeking particles with total energy less than $`D`$ are energetically forbidden from leaving the trapping region. For atoms and molecules with a magnetic moment of one Bohr magneton it is possible to produce trap depths of $`1`$ K. The trap depth for a neutron in the same trap would be only $`1`$ mK, because of its much smaller magnetic moment. Despite this difficulty, magnetic trapping of the neutron was proposed as early as 1961 by Vladimirskiĭ. His proposed technique was later used to confine neutrons using a combination of gravity and magnets. A separate effort to trap neutrons using a similar loading method to our work (but different detection scheme) was unsuccessful because of the high temperature of the helium during the loading phase. Crucial to the utility of traps are the techniques used to load them. In order to catch a particle in a static conservative trap, its energy must be lowered while it is in the potential well. Atoms and molecules have been cooled and loaded into magnetic traps by scattering with either cryogenic surfaces, cold gases or photons from a laser beam (laser cooling). Neutrons, however, cannot be loaded by such methods because they cannot be excited optically and interact too weakly with atoms to be effectively cooled by a gas. Direct thermalization with a cold solid or liquid is generally precluded by the high probability for neutron absorption in the vast majority of materials. Our trapping of neutrons relies on a loading technique that employs the “superthermal process”. A neutron with kinetic energy near 11 K (where the free neutron and superfluid helium dispersion curves cross) that passes through the helium-filled trapping region can lose nearly all of its energy through the creation of a single phonon. Neutrons that scatter to energies less than the trap depth ($`1`$ mK) and in the appropriate spin state are trapped. Upscattering of UCN in liquid helium by single phonon absorption is highly suppressed due to the low density of 11 K phonons in the superfluid (proportional to $`e^{11\mathrm{K}/T}`$). The rate of higher-order (multiple-phonon) upscattering processes is proportional to $`T^7`$(ref. 19). The helium is maintained below a temperature of 250 mK to minimize upscattering of the trapped neutrons (total calculated upscattering rate per neutron $`<10^6`$s<sup>-1</sup>). Snother possible loss mechanism for trapped neutrons is nuclear absorption due to <sup>3</sup>He contamination of the superfluid <sup>4</sup>He bath. Unlike <sup>4</sup>He which does not absorb thermal neutrons at all, <sup>3</sup>He readily absorbs neutrons with a thermal cross section, $`\sigma =5330\times 10^{24}`$ cm<sup>2</sup>. To minimize these losses, our superfluid-helium bath was isotopically purified (<sup>3</sup>He/<sup>4</sup>He $`<5\times 10^{13}`$; ref. 20), giving a calculated absorption loss rate per neutron of $`5\times 10^6`$ s<sup>-1</sup>. Thus, the trapped neutrons travel unimpeded through the helium and reside in the trap for a time limited primarily by the beta-decay lifetime of the neutron. In our experiment, the superfluid helium is contained in a tube that is axially centered within a superconducting magnetic trap (see Fig. 1). The incident cold neutron beam is collimated by a neutron-absorbing ring, passes through the trapping region, and is absorbed by the beam stop. As the beam traverses the trapping region, about 1% of the 11 K neutrons scatter in the helium. Some of these neutrons are trapped and the remainder are absorbed by shielding materials that surround the helium. When a trapped neutron decays (into an electron, proton and anti-neutrino), the resulting high-energy electron (up to 782 keV) travels through the helium leaving a trail of ionized helium atoms. These ions quickly combine with neutral helium atoms to form excimer diatomic molecules, about half in singlet states and half in triplet states. About 15 singlet molecules are created per kiloelectronvolt of electron energy and the trail is 0–20 mm long. The singlet molecules decay within 10 ns, emitting a pulse of light in the extreme ultraviolet (EUV), $`\lambda 7090`$ nm. This pulse of scintillation light is the signal of a neutron decaying in the trap. Radiative decay of the triplet states has a much longer lifetime, which we have measured to be $`13\pm 2`$ s (ref. 22), and does not contribute to our signal. The EUV scintillation light is wavelength-shifted to the visible and transported to the outside of the cryostat where it is detected. As shown in Fig. 1, the superfluid-helium-filled trapping region is surrounded by an acrylic tube. The inside surface of this tube is coated with a thin layer of polystyrene doped with the organic fluor tetraphenyl butadiene (TPB; ref. 23). The TPB converts the EUV scintillation light into blue light, a portion of which is internally reflected down the length of the tube. The tube is optically coupled to a solid light guide which transports the light to the end of the 250 mK region. The light then passes through a window at 4 K and into a second light guide which exits the dewar. The light is split into two guides that are each coupled to a photomultiplier tube (PMT). Background from uncorrelated photons is suppressed by requiring coincident detection of at least two photoelectrons in each PMT. This is required because of significant time-dependent luminescence (initial counting rate without coincidence detection of $`10^4`$ s<sup>-1</sup>), induced by neutron absorption in the shielding materials. In a series of calibration runs (using a <sup>113</sup>Sn, 360-keV beta line source placed in the center of the trapping region) we have determined that a total of $`14.5\pm 1.5`$ photoelectrons per megaelectronvolt of beta energy are detected in a pulse 20 ns wide. We estimate that a coincidence signal is recorded for $`31\pm 4`$% of neutron decays in the trapping region. In each experimental run, the cold neutron beam is allowed to pass through the trapping region for 1,350 s, after which the beam is blocked and pulses of light are counted for 3,600 s. While the beam is on, neutrons interact with the helium and the number of UCN in the trap increases. After the beam is turned off, the PMTs are turned on and a decreasing rate of scintillation events is observed. A background signal, consisting of both constant and time-varying components, obscures the trapped neutron signal. These backgrounds are subtracted by taking data in two different kinds of run, trapping (signal + background) and non-trapping (background only), and subtracting the two. In a trapping run the magnetic field is on during the initial loading phase, so that UCN are confined in the trap, while in a non-trapping run the magnetic field is initially off, so that no neutrons can be trapped. In a non-trapping run, the field is turned on during detection so that any residual luminescence (which depends on the magnetic field) will be properly subtracted. Equal numbers of each kind of run were taken, pooled and subtracted to give the background subtracted data. Two sets of background subtracted data were collected: set 1 with a trap depth of 0.76 mK (figure 2a) and set 2 with a lower trap depth of 0.50 mK (figure 2c). (The lower trap depth was used due to problems with the magnet.) Most of the run-to-run variation in background rate is eliminated by excluding the first two pairs of runs in which the background rate is changing quickly due to activation of materials with lifetimes $`>12`$ h. The remaining 23 pairs in set 1 (from about five days of running) and 120 pairs in set 2 (from about three weeks of running) are pooled and modeled as: $$W_1=\alpha _1e^{t/\tau }+C_1,W_2=\alpha _2e^{t/\tau }+C_2.$$ (1) The subscripts 1 and 2 refer to sets 1 and 2, $`\alpha _i=ϵN_i/\tau `$, $`N_i`$ is the initial number of trapped neutrons and $`i=1`$ or 2, $`ϵ`$ is the detection efficiency, and $`\tau `$ is the lifetime of neutrons in the trap. The constant $`C_i`$ is present to account for the possible remaining effect of the changing background rate due to long-lived activation. However, in all of our fits the value of $`C_i`$ is consistent with zero. The fit is performed simultaneously on the two data sets, minimizing the total $`\chi ^2`$ while varying five parameters: $`\alpha _1`$, $`\alpha _2`$, $`C_1`$, $`C_2`$ and $`\tau `$. The only parameter connecting the two data sets is $`\tau `$. The best fit values indicate $`N_1=560\pm 160`$ and $`N_2=240\pm 65`$. Calculations using the known beam flux, trap geometry and the theory of the superthermal process predict $`N_1=480\pm 100`$ and $`N_2=255\pm 50`$, in good agreement with the measured values. The best fit value for the lifetime, $`\tau =750_{200}^{+330}`$ s is consistent with the presently accepted value of the neutron beta-decay lifetime of $`886.7\pm 1.9`$ s (ref. 5). All of the errors quoted correspond to a 68% confidence interval. To verify that our signal is due to trapped neutrons, we doped the isotopically pure <sup>4</sup>He with <sup>3</sup>He at a concentration of $`2\times 10^7`$ <sup>3</sup>He/<sup>4</sup>He. This amount of <sup>3</sup>He absorbs the trapped neutrons in less than 1 s without affecting anything else in the experiment (less than 1% of the 11 K neutrons are absorbed by <sup>3</sup>He). The <sup>3</sup>He background subtracted data (see Fig. 2b and d) is consistent with zero, and differs from the signal data by more than four standard deviations. This confirms that our signal is due to trapped neutrons. This work demonstrates the loading, trapping, and detection techniques necessary for performing a neutron lifetime measurement using magnetically trapped UCN. Another important result is the direct confirmation of the theoretical prediction of the UCN density in the trap. Our value for the number of trapped neutrons at a 0.76 mK trap depth corresponds to a density of 2 UCN/cm<sup>3</sup>, compared to the density of 1 UCN/cm<sup>3</sup> obtained in the UCN material bottle experiments with a comparable UCN cut-off energy and higher flux reactor. Our measured density is consistant with previous measurements of the UCN production rate based on observation of upscattered UCN at higher temperatures and confirms that the low number of UCN detected in ref. 26 was not due to an intrinsically low UCN production rate. Magnetic trapping should allow significant improvement in the measurement of the neutron lifetime $`\tau _n`$. There are several loss mechanisms specific to this method, including thermal upscattering and <sup>3</sup>He absorption. Both of these can be suppressed to a high level as discussed above. Another possible loss mechanism is neutron depolarization. When the neutron passes through a region where the field changes direction quickly compared to the neutron’s spin precession frequency, the neutron can flip spin and be ejected from the trap. In our trap geometry, the trap has no zero-field regions and this loss rate can be minimized by the application of a bias field. Using our method with a higher-flux source of cold neutrons, a $`10^5`$ measurement of $`\tau _n`$ should be possible. In the near term, improvements such as increased trap size and depth (resluting in an improved ratio of signal-to-background events) should allow an improved accuracy measurement of $`\tau _n`$ at the $`5\times 10^4`$ level. Other potential applications of superthermal UCN sources include searches for time-reversal violation and new methods of neutron scattering especially suited for the study of large (biological-scale) molecules. Received 5 August; accepted 16 November 1999. Acknowledgments. We thank J. M. Rowe, D. M. Gilliam, G. L. Jones, J. S. Nico, N. Clarkson, G. P. Lamaze, C. Chin, C. Davis, D. Barkin, A. Black, V. Dinu, J. Higbie, H. Park, R. Ramakrishnan, I. Siddiqi, and G. Brandenburg for their help with this project. We thank P. McClintock, D. Meredith and P. Hendry for supplying the isotopically pure helium. We acknowledge the support of the NIST, US DOC, in providing the neutron facilities used in this work. This work is supported in part by the US NSF. The NIST authors acknowledge the support of the US DOE. Correspondence and requests for materials should be addressed to P.R.H. (e-mail: paul.huffman@nist.gov).
warning/0001/hep-ph0001151.html
ar5iv
text
# Azimuthal Angle Distribution in 𝐵→𝐾^∗(→𝐾𝜋)ℓ⁺ℓ⁻ at low invariant 𝑚_{ℓ⁺⁢ℓ⁻} Region ## I Introduction Rare $`B`$ decays are suitable for testing the standard model (SM) and the models beyond the SM. Exclusive decay $`BK^{}\gamma `$ and the corresponding inclusive decay $`BX_s\gamma `$ place strong constraints on the parameters of models beyond the SM, for example, the left-right symmetric model (LRSM), SUSY, the multi-Higgs doublet model, etc. . However, if the decay rate is not changed drastically from the prediction of the SM, it would be very difficult to probe new physics effects from the $`BK^{}\gamma `$ decay. In this regard, new methods have been proposed, which consist of observables sensitive to chiral structure, such as mixing-induced CP asymmetry in $`B_{d,s}M^0\gamma `$ decay and $`\mathrm{\Lambda }`$ polarization in the $`\mathrm{\Lambda }_b\mathrm{\Lambda }\gamma `$ decay . And these methods have been also applied to search for the new physics, as shown in : The $`BK^{}\gamma `$ decay occurs through the effective interaction of two magnetic moment operators, $`m_b(C_{7L}\overline{s}_L\sigma _{\mu \nu }b_RF^{\mu \nu }+C_{7R}\overline{s}_R\sigma _{\mu \nu }b_LF^{\mu \nu }).`$ (1) In the SM, the first term is dominant and the second term is suppressed by $`𝒪(m_s/m_b)`$. In the LRSM, the contribution of both operators can be equally important . The new contributions for $`C_{7R}`$ and $`C_{7L}`$ in this model are enhanced as $`m_t/m_b`$. Because the probability for $`B`$ meson decaying to left-handed (or right-handed) circular polarized $`K^{}`$ is proportional to $`|C_{7L}|^2`$ (or $`|C_{7R}|^2`$), the polarization measurement of $`K^{}`$ and $`\gamma `$ is useful for extracting the ratio of $`\frac{|C_{7L}|}{|C_{7R}|}`$. However, since the polarizations of high energy real photon ($`\gamma `$) cannot be measured easily, we have to develop more elaborated method for extracting the afore-mentioned ratio. Therefore, we propose another new method, which is very efficient when we cannot find the new physics effects from the total decay rate of $`BK^{}\gamma `$. Let us imagine the decay configuration when $`K^{}`$ from the decay $`BK^{}\gamma ^{}`$ is emitted to the direction of $`+z`$ and $`\gamma ^{}`$ is emitted to the opposite direction in the rest frame of $`B`$ meson. Here $`\gamma ^{}`$ is off-shell photon and it further decays into $`\mathrm{}^+\mathrm{}^{}`$, and $`K^{}`$ subsequently decays into $`K\pi `$. If we ignore small mixture of the longitudinal component, the angular momentum of $`K^{}`$ is either $`J_z=+1`$ or $`J_z=1`$, and the corresponding production amplitude is proportional to $`C_{7R}`$ or $`C_{7L}`$, respectively. Suppose the final $`K`$ meson is emitted to the direction of $`(\theta _K,\varphi )`$ in the rest frame of $`K^{}`$, where $`\theta _K`$ is a polar angle and $`\varphi `$ is an azimuthal angle between the decay plane of ($`K\pi `$) and the decay plane of ($`\mathrm{}^+\mathrm{}^{}`$). The decay amplitude for the whole process is proportional to $$AC_{7L}\mathrm{exp}(i\varphi )+BC_{7R}\mathrm{exp}(+i\varphi )+C.$$ Here, $`A`$, $`B`$ and $`C`$ are the real functions of the other angles and $`C`$ corresponds to the amplitude for the $`B`$ meson decaying into the longitudinally polarized $`K^{}`$, which is possible only for the off-shell photon. By squaring the amplitude, we can show that in the azimuthal distribution the coefficient of $`\mathrm{cos}(2\varphi )`$ (and that of $`\mathrm{sin}(2\varphi )`$) is $`Re(C_{7R}C_{7L}^{})`$ (and $`Im(C_{7R}C_{7L}^{})`$). Therefore, from the angular dependence we may extract the ratio $`\frac{C_{7L}}{C_{7R}}`$. Note that for on-shell photon the dependence on the azimuthal angle $`\varphi `$ does not appear for the $`BK^{}(K\pi )+\gamma `$ decay because of rotational symmetry of the decay configuration with respect to $`z`$-axis. This naturally leads us to investigate the angular distribution of the $`BK^{}(K\pi )+\gamma ^{}(\mathrm{}^+\mathrm{}^{})`$. However, once we consider off-shell photon, some complications arise and the argument discussed above has to be modified: The other diagrams like box and $`Z`$ penguin diagrams now contribute to the same final state through the process, $`BK^{}(K\pi )+\mathrm{}^++\mathrm{}^{}`$. They do not contribute to $`BK^{}\gamma `$. Though in the low invariant mass region of dileptons the decay through the magnetic moment interactions may be dominant, we still have to take into account the effect of the box and $`Z`$ penguin diagrams, which have the form of the local four-fermi interactions in the effective Hamiltonian. The paper is organized as follows: In section 2, we derive the angular distribution formulae in terms of the helicity amplitudes. In section 3, our numerical analyses for azimuthal angle are shown. Concluding remarks are also in section 3. ## II Angular Distribution of $`BK^{}(K\pi )+\mathrm{}^++\mathrm{}^{}`$ The short distance contribution to decay $`BK^{}(K\pi )\mathrm{}^+\mathrm{}^{}`$ is governed by the quark level decay $`bs\mathrm{}^+\mathrm{}^{}`$ as $`(bs\mathrm{}^+\mathrm{}^{})={\displaystyle \frac{G_F\alpha }{\sqrt{2}\pi }}V_{ts}^{}V_{tb}`$ $`[`$ $`(C_9^{eff}C_{10})\overline{s}_L\gamma _\mu b_L\overline{\mathrm{}}_L\gamma ^\mu \mathrm{}_L`$ (2) $`+`$ $`(C_9^{eff}+C_{10})\overline{s}_L\gamma _\mu b_L\overline{\mathrm{}}_R\gamma ^\mu \mathrm{}_R`$ (3) $``$ $`2C_{7L}\overline{s_L}i\sigma _{\mu \nu }{\displaystyle \frac{L^\nu }{L^2}}m_bb_R\overline{\mathrm{}}\gamma ^\mu \mathrm{}`$ (4) $``$ $`2C_{7R}\overline{s_R}i\sigma _{\mu \nu }{\displaystyle \frac{L^\nu }{L^2}}m_bb_L\overline{\mathrm{}}\gamma ^\mu \mathrm{}],`$ (5) where we assume that new physics effect does not change the Wilson coefficients $`C_9`$ and $`C_{10}`$ and only can change the coefficients of non-local four-fermi interactions which are denoted by $`C_7`$. The latter also contributes to $`bs\gamma `$. In the SM, $`C_{7L}=C_{7eff},C_{7R}=\frac{ms}{mb}C_{7eff}`$, where $`C_{7eff}`$ is given in Ref. . Although there are overwhelming resonance contributions from $`J/\psi `$ and $`\psi ^{}`$, etc., the short distance contribution still dominates the low invariant mass region of the lepton pair . The effective Hamiltonian for the corresponding $`bs\mathrm{}^+\mathrm{}^{}`$ is $`_{eff}(bs\mathrm{}^+\mathrm{}^{})`$ $`=`$ $`_{eff}(bs\gamma ){\displaystyle \frac{G_F}{\sqrt{2}}}V_{tb}V_{ts}^{}{\displaystyle \frac{\alpha }{\pi }}{\displaystyle \underset{i=9}{\overset{10}{}}}C_iO_i,`$ (6) $`\mathrm{where}_{eff}(bs\gamma )`$ $`=`$ $`{\displaystyle \frac{G_F}{\sqrt{2}}}V_{tb}V_{ts}^{}[C_{7L}O_{7L}+C_{7R}O_{7R}].`$ (7) The operators $`O_i`$ relevant for us are $`O_9`$ $`=`$ $`(\overline{s}b)_L(\overline{\mathrm{}}\mathrm{})_V,`$ (8) $`O_{10}`$ $`=`$ $`(\overline{s}b)_L(\overline{\mathrm{}}\mathrm{})_A,`$ (9) $`O_{7L}`$ $`=`$ $`{\displaystyle \frac{em_b}{4\pi ^2}}(\overline{s}_L\sigma _{\mu \nu }b_R)F^{\mu \nu },`$ (10) $`O_{7R}`$ $`=`$ $`{\displaystyle \frac{em_b}{4\pi ^2}}(\overline{s}_R\sigma _{\mu \nu }b_L)F^{\mu \nu },`$ (11) where in addition to the SM operators $`O_9`$, $`O_{10}`$ and $`O_{7L}`$, we include also a new operator $`O_{7R}`$. The new physics effects can contribute to any of the operators. For example, the LRSM based upon the electroweak gauge group $`SU(2)_L\times SU(2)_R\times U(1)`$ can lead to interesting new physics effects in the operators $`O_{7L}`$ and $`O_{7R}`$. Due to the extended gauge structure there are both new neutral and charged gauge bosons, $`Z_R`$ and $`W_R`$, as well as a right-handed gauge coupling, $`g__R`$. After the symmetry breaking, the charged $`W_R`$ mixes with $`W_L`$ of the SM to form the mass eigenstates $`W_{1,2}`$ with eigenvalues $`M_{1,2}`$. And this mixing is described by two parameters; a real mixing angle $`\zeta `$ and a phase $`\alpha `$, $`\left(\begin{array}{c}W_1^+\\ W_2^+\end{array}\right)=\left(\begin{array}{cc}\mathrm{cos}\zeta & e^{i\alpha }\mathrm{sin}\zeta \\ \mathrm{sin}\zeta & e^{i\alpha }\mathrm{cos}\zeta \end{array}\right)\left(\begin{array}{c}W_L^+\\ W_R^+\end{array}\right).`$ (18) In this model the charged current interactions of the right-handed quarks are governed by a right-handed CKM matrix $`V_R`$, which, in principle, need not be related to its left-handed counterpart $`V_L`$. If we neglect the charged physical scalar contributions, the magnetic moment operator coefficients in the LRSM are given by $`C_{7L}(m_b)`$ $`=`$ $`C_{7L}^{SM}(m_b)+A^{tb}\left[\eta ^{16/23}\stackrel{~}{F}(x_t)+{\displaystyle \frac{8}{3}}(\eta ^{14/23}\eta ^{16/23})\stackrel{~}{G}(x_t)\right]+A^{cb}{\displaystyle \underset{i}{}}h_i^{}\eta ^{p_i^{}},`$ (19) $`C_{7R}(m_b)`$ $`=`$ $`(A^{ts})^{}\left[\eta ^{16/23}\stackrel{~}{F}(x_t)+{\displaystyle \frac{8}{3}}(\eta ^{14/23}\eta ^{16/23})\stackrel{~}{G}(x_t)\right]+(A^{cs})^{}{\displaystyle \underset{i}{}}h_i^{}\eta ^{p_i^{}},`$ (20) where $`C_{7L}^{SM}(m_b)=\eta ^{16/23}F(x_t)+{\displaystyle \frac{8}{3}}(\eta ^{14/23}\eta ^{16/23})G(x_t)+{\displaystyle \underset{i}{}}h_i\eta ^{p_i},`$ (21) $`A^{tq}=\zeta e^{i\alpha }{\displaystyle \frac{m_t}{m_b}}{\displaystyle \frac{V_R^{tq}}{V_L^{tq}}},A^{cq}=\zeta e^{i\alpha }{\displaystyle \frac{m_c}{m_b}}{\displaystyle \frac{V_R^{cq}}{V_L^{cq}}},`$ (22) with $`\eta =\alpha _s(M_{W_1})/\alpha _s(m_b)`$ and $`x_t=(m_t/m_b)^2`$. The various functions of $`x_t`$ and the coefficients $`h_i^{()}`$ and powers $`p_i^{()}`$ can be found in Ref. . In this paper, we will not constrain ourselves to the LRSM, but discuss the general effects of new physics. Working for the exclusive decay $`BK^{}\mathrm{}^+\mathrm{}^{}`$, we need form factors for the $`BK^{}`$ transition. These form factors can be written as $`K^{}(p^{})|\overline{s}\gamma _\mu b|B(p)`$ $`=`$ $`ig\epsilon _{\mu \nu \lambda \sigma }ϵ^\nu (p+p^{})^\lambda (pp^{})^\sigma ,`$ (23) $`K^{}(p^{})|\overline{s}\gamma _\mu \gamma _5b|B(p)`$ $`=`$ $`fϵ_\mu ^{}+a_+(ϵ^{}p)(p+p^{})_\mu +a_{}(ϵ^{}p)(pp^{})_\mu ,`$ (24) $`K^{}(p^{})|\overline{s}\sigma _{\mu \nu }b|B(p)`$ $`=`$ $`g_+\epsilon _{\mu \nu \lambda \sigma }ϵ^\lambda (p+p^{})^\sigma +g_{}\epsilon _{\mu \nu \lambda \sigma }ϵ^\lambda (pp^{})^\sigma `$ (26) $`+h\epsilon _{\mu \nu \lambda \sigma }(p+p^{})^\lambda (pp^{})^\sigma (ϵ^{}p),`$ $`K^{}(p^{})|\overline{s}\sigma _{\mu \nu }\gamma _5b|B(p)`$ $`=`$ $`ig_+\left(ϵ_\nu ^{}(p+p^{})_\mu ϵ_\mu ^{}(p+p^{})_\nu \right)ig_{}\left(ϵ_\nu ^{}(pp^{})_\mu ϵ_\mu ^{}(pp^{})_\nu \right)`$ (28) $`2ih\left(p_\mu p_\nu ^{}p_\nu p_\mu ^{}\right)(ϵ^{}p),`$ where we have used $`\sigma ^{\mu \nu }=\frac{i}{2}\epsilon ^{\mu \nu \lambda \sigma }\sigma _{\lambda \sigma }\gamma _5`$. We also use the following definitions, $`\gamma _5=i\gamma ^0\gamma ^1\gamma ^2\gamma ^3`$ and $`\epsilon _{0123}=1`$. The $`K^{}`$ meson subsequently decays to $`K`$ and $`\pi `$, with effective Hamiltonian $$_{eff}=g_{K^{}K\pi }(p_Kp_\pi )ϵ_K^{}.$$ (29) In the following analysis, we neglect the masses of leptons, kaon and pion. The final 4-body decay amplitude can be written as the sum of two amplitudes, $$𝒜=A_R+A_L,$$ where $`A_R=`$ $`{\displaystyle \frac{G_F}{\sqrt{2}}}V_{tb}V_{ts}^{}g_{K^{}K\pi }{\displaystyle \frac{\alpha m_b}{\pi L^2}}(\overline{\mathrm{}}_R\gamma ^\mu l_R)\left(a_Rg_{\mu \nu }b_RP_\mu L_\nu +ic_Rϵ_{\mu \nu \alpha \beta }P^\alpha L^\beta \right)`$ (31) $`{\displaystyle \frac{g^{\nu \alpha }P^\nu P^\alpha /m_K^{}^2}{P^2m_K^{}^2+im_K^{}\mathrm{\Gamma }_K^{}}}(p_Kp_\pi )_\alpha ,`$ $`A_L=`$ $`{\displaystyle \frac{G_F}{\sqrt{2}}}V_{tb}V_{ts}^{}g_{K^{}K\pi }{\displaystyle \frac{\alpha m_b}{\pi L^2}}(\overline{\mathrm{}}_L\gamma ^\mu l_L)\left(a_Lg_{\mu \nu }b_LP_\mu L_\nu +ic_Lϵ_{\mu \nu \alpha \beta }P^\alpha L^\beta \right)`$ (33) $`{\displaystyle \frac{g^{\nu \alpha }P^\nu P^\alpha /m_K^{}^2}{P^2m_K^{}^2+im_K^{}\mathrm{\Gamma }_K^{}}}(p_Kp_\pi )_\alpha ,`$ with $`P=p_K+p_\pi `$, $`L=p_++p_{}`$. The $`a_R,b_R,c_R`$ and $`a_L,b_L,c_L`$ can be expressed as $`a_L`$ $`=`$ $`C_7\left[2(PL)g_++L^2(g_++g_{})\right]{\displaystyle \frac{(C_9C_{10})f}{2m_b}}L^2,`$ (34) $`b_L`$ $`=`$ $`2C_7(g_+L^2h)+{\displaystyle \frac{(C_9C_{10})a_+}{m_b}}L^2,`$ (35) $`c_L`$ $`=`$ $`2C_{7+}g_++{\displaystyle \frac{(C_9C_{10})g}{m_b}}L^2,`$ (36) $`a_R`$ $`=`$ $`C_7\left[2(PL)g_++L^2(g_++g_{})\right]{\displaystyle \frac{(C_9+C_{10})f}{2m_b}}L^2,`$ (37) $`b_R`$ $`=`$ $`2C_7(g_+L^2h)+{\displaystyle \frac{(C_9+C_{10})a_+}{m_b}}L^2,`$ (38) $`c_R`$ $`=`$ $`2C_{7+}g_++{\displaystyle \frac{(C_9+C_{10})g}{m_b}}L^2,`$ (39) where $`C_7=C_{7R}C_{7L}`$ and $`C_{7+}=C_{7R}+C_{7L}`$. The decay rate is computed and the result is $`{\displaystyle \frac{d^5\mathrm{\Gamma }}{dp^2dl^2d\mathrm{cos}\theta _Kd\mathrm{cos}\theta _+d\varphi }}={\displaystyle \frac{2\sqrt{\lambda }}{128\times 256\pi ^6m_B^3}}(|A_R|^2+|A_L|^2),`$ (40) with $`p=\sqrt{P^2}`$, $`l=\sqrt{L^2}`$, and $`\lambda =(m_B^2p^2l^2)^2/4p^2l^2`$. We introduce the various angles as: $`\theta _K`$ is the polar angle of the $`K`$ momentum in the rest system of the $`K^{}`$ meson with respect to the helicity axis, i.e. the outgoing direction of $`K^{}`$. Similarly $`\theta _+`$ is the polar angle of the positron in the $`\gamma ^{}`$ rest system with respect to the helicity axis of the $`\gamma ^{}`$. And $`\varphi `$ is the azimuthal angle between the planes of the two decays $`K^{}K\pi `$ and $`\gamma ^{}\mathrm{}^+\mathrm{}^{}`$. And then, $`|A_R|^2`$ $`=`$ $`\left|{\displaystyle \frac{G_F}{\sqrt{2}}}V_{tb}V_{ts}^{}g_{K^{}K\pi }{\displaystyle \frac{\alpha m_b}{\pi L^2}}\right|^2{\displaystyle \frac{1}{(P^2m_K^{}^2)^2+(m_K^{}\mathrm{\Gamma }_K^{})^2}}`$ (42) $`[|a_R|^2\{(QL)^2(QN)^2{\displaystyle \frac{(L^2N^2)Q^2}{2}}\}`$ $`+`$ $`2Re(a_Rb_R^{})\left\{(QL)^2(PL)+(QN)(PN)(QL)\right\}`$ (43) $`+`$ $`|b_R|^2\left\{(PL)^2(QL)^2(PN)^2(QL)^2{\displaystyle \frac{L^2N^2}{2}}P^2(QL)^2\right\}`$ (44) $`+`$ $`|c_R|^2\left\{(\stackrel{~}{NPLQ})(\stackrel{~}{NPLQ}){\displaystyle \frac{L^2N^2}{2}}(\stackrel{~}{PLQ})(\stackrel{~}{PLQ})\right\}`$ (45) $`+`$ $`2Im(b_Rc_R^{})(PN)(QL)(\stackrel{~}{NPLQ})`$ (46) $`+`$ $`2Im(c_Ra_R^{})(QN)(\stackrel{~}{NPLQ})`$ (47) $``$ $`2Im(b_Ra_R^{})(QL)(\stackrel{~}{NPLQ})`$ (48) $``$ $`2Re(c_Ra_R^{})(\stackrel{~}{LQN})(\stackrel{~}{QPL})`$ (49) $`+`$ $`2Re(b_Rc_R^{})(QL)(\stackrel{~}{LPN})(\stackrel{~}{QPL})],`$ (50) and $`|A_L|^2`$ $`=`$ $`\left|{\displaystyle \frac{G_F}{\sqrt{2}}}V_{tb}V_{ts}^{}g_{K^{}K\pi }{\displaystyle \frac{\alpha m_b}{\pi L^2}}\right|^2{\displaystyle \frac{1}{(P^2m_K^{}^2)^2+(m_K^{}\mathrm{\Gamma }_K^{})^2}}`$ (52) $`[|a_L|^2\{(QL)^2(QN)^2{\displaystyle \frac{(L^2N^2)Q^2}{2}}\}`$ $`+`$ $`2Re(a_Lb_L^{})\left\{(QL)^2(PL)+(QN)(PN)(QL)\right\}`$ (53) $`+`$ $`|b_L|^2\left\{(PL)^2(QL)^2(PN)^2(QL)^2{\displaystyle \frac{L^2N^2}{2}}P^2(QL)^2\right\}`$ (54) $`+`$ $`|c_L|^2\left\{(\stackrel{~}{NPLQ})(\stackrel{~}{NPLQ}){\displaystyle \frac{L^2N^2}{2}}(\stackrel{~}{PLQ})(\stackrel{~}{PLQ})\right\}`$ (55) $`+`$ $`2Im(b_Lc_L^{})(PN)(QL)(\stackrel{~}{NPLQ})`$ (56) $`+`$ $`2Im(c_La_L^{})(QN)(\stackrel{~}{NPLQ})`$ (57) $`+`$ $`2Im(b_La_L^{})(QL)(\stackrel{~}{NPLQ})`$ (58) $`+`$ $`2Re(c_La_L^{})(\stackrel{~}{LQN})(\stackrel{~}{QPL})`$ (59) $``$ $`2Re(b_Lc_L^{})(QL)(\stackrel{~}{LPN})(\stackrel{~}{QPL})],`$ (60) where $`(\stackrel{~}{ABC})_\mu =\epsilon _{\mu \alpha \beta \gamma }A^\alpha B^\beta C^\gamma `$, $`(\stackrel{~}{ABCD})=\epsilon _{\alpha \beta \gamma \delta }A^\alpha B^\beta C^\gamma D^\delta `$, $`Q=p_Kp_\pi `$, and $`N=p_+p_{}`$. We use $`Tr(\gamma ^\alpha \gamma ^\beta \gamma ^\gamma \gamma ^\delta \gamma _5)=+4i\epsilon ^{\alpha \beta \gamma \delta }`$. Comparing $`|A_L|^2`$ with $`|A_R|^2`$, we see that the signs of the corresponding last three terms are opposite to each other. We can simplify the expression by introducing the helicity amplitudes. The helicity amplitudes are defined as, $`H_{(\pm 1,0)}^L`$ $`=`$ $`ϵ_{V}^{(\pm ,0)}{}_{}{}^{\nu }ϵ_{\gamma }^{(\pm ,0)}{}_{}{}^{\mu }\left(a_Lg_{\mu \nu }b_LP_\mu L_\nu +ic_L\epsilon _{\mu \nu \alpha \beta }P^\alpha L^\beta \right),`$ (61) $`H_{(\pm 1,0)}^R`$ $`=`$ $`ϵ_{V}^{(\pm ,0)}{}_{}{}^{\nu }ϵ_{\gamma }^{(\pm ,0)}{}_{}{}^{\mu }\left(a_Rg_{\mu \nu }b_RP_\mu L_\nu +ic_R\epsilon _{\mu \nu \alpha \beta }P^\alpha L^\beta \right).`$ (62) We define the following polarization vectors: $$\begin{array}{cccccccc}ϵ_V^+\hfill & =& (\hfill & 0,& 1,& \hfill i,& 0& )/\sqrt{2},\hfill \\ ϵ_V^{}\hfill & =& (\hfill & 0,& 1,& \hfill i,& 0& )/\sqrt{2},\hfill \\ ϵ_V^0\hfill & =& (\hfill & \frac{\sqrt{\lambda }}{m_B},& 0,& \hfill 0,& \sqrt{\frac{\lambda }{m_B^2}+p^2}& )/p,\hfill \\ ϵ_\gamma ^+\hfill & =& (\hfill & 0,& 1,& \hfill i,& 0& )/\sqrt{2},\hfill \\ ϵ_\gamma ^{}\hfill & =& (\hfill & 0,& 1,& \hfill +i,& 0& )/\sqrt{2},\hfill \\ ϵ_\gamma ^0\hfill & =& (\hfill & \frac{\sqrt{\lambda }}{m_B},& 0,& \hfill 0,& \sqrt{\frac{\lambda }{m_B^2}+l^2}& )/l.\hfill \end{array}$$ (63) Substituting them into Eq. (62), we obtain the following helicity amplitudes, $`H_{+1}^L`$ $`=`$ $`\left(a_L+c_L\sqrt{\lambda }\right),`$ (64) $`H_1^L`$ $`=`$ $`\left(a_Lc_L\sqrt{\lambda }\right),`$ (65) $`H_0^L`$ $`=`$ $`a_L{\displaystyle \frac{PL}{pl}}+{\displaystyle \frac{b_L\lambda }{pl}},`$ (66) $`H_{+1}^R`$ $`=`$ $`\left(a_R+c_R\sqrt{\lambda }\right),`$ (67) $`H_1^R`$ $`=`$ $`\left(a_Rc_R\sqrt{\lambda }\right),`$ (68) $`H_0^R`$ $`=`$ $`a_R{\displaystyle \frac{PL}{pl}}+{\displaystyle \frac{b_R\lambda }{pl}}.`$ (69) Applying the Eqs. (34-36), we have $`H_{+1}^L`$ $`=`$ $`2g_+\left(C_7(PL)C_{7+}\sqrt{\lambda }\right)C_7l^2(g_++g_{}){\displaystyle \frac{(C_9C_{10})l^2}{2m_b}}(f2g\sqrt{\lambda }),`$ (70) $`H_1^L`$ $`=`$ $`2g_+\left(C_7(PL)+C_{7+}\sqrt{\lambda }\right)C_7l^2(g_++g_{}){\displaystyle \frac{(C_9C_{10})l^2}{2m_b}}(f+2g\sqrt{\lambda }),`$ (71) $`H_0^L`$ $`=`$ $`{\displaystyle \frac{lC_7}{p}}\left[2p^2g_++(PL)(g_++g_{})+2\lambda h\right]+{\displaystyle \frac{(C_9C_{10})l}{2m_bp}}\left[f(PL)+2a_+\lambda \right],`$ (72) where $`PL=\sqrt{\lambda +p^2l^2}=(m_B^2p^2l^2)/2`$. The formulae for $`H_{+1}^R`$, $`H_1^R`$, $`H_0^R`$ are the same as above except that $`C_{10}C_{10}`$. Using the variables $`\theta _K`$, $`\theta _+`$, $`\varphi `$, $`p`$ and $`l`$, we find: $`QL`$ $`=`$ $`\sqrt{\lambda }\mathrm{cos}\theta _K,`$ (73) $`PN`$ $`=`$ $`\sqrt{\lambda }\mathrm{cos}\theta _+,`$ (74) $`PL`$ $`=`$ $`\sqrt{\lambda +p^2l^2},`$ (75) $`QN`$ $`=`$ $`\sqrt{\lambda +p^2l^2}\mathrm{cos}\theta _K\mathrm{cos}\theta _+pl\mathrm{sin}\theta _K\mathrm{sin}\theta _+\mathrm{cos}\varphi ,`$ (76) $`(\stackrel{~}{NPLQ})`$ $`=`$ $`pl\sqrt{\lambda }\mathrm{sin}\theta _K\mathrm{sin}\theta _+\mathrm{sin}\varphi .`$ (77) Using these equations, we can get the results for Eqs. (50,60), whose sum makes the decay angular distribution of $`BK^{}(K\pi )\mathrm{}^+\mathrm{}^{}`$, $`{\displaystyle \frac{d^5\mathrm{\Gamma }}{dp^2dl^2d\mathrm{cos}\theta _Kd\mathrm{cos}\theta _+d\varphi }}`$ $`=`$ $`{\displaystyle \frac{\alpha ^2G_F^2g_{K^{}K\pi }^2\sqrt{\lambda }p^2m_b^2|V_{tb}V_{ts}^{}|^2}{64\times 8(2\pi )^8m_B^3l^2[(p^2m_K^{}^2)^2+m_K^{}^2\mathrm{\Gamma }_K^{}^2]}}`$ $`\times `$ $`\{4\mathrm{cos}^2\theta _K\mathrm{sin}^2\theta _+(|H_0^R|^2+|H_0^L|^2)`$ (86) $`+\mathrm{sin}^2\theta _K(1+\mathrm{cos}^2\theta _+)(|H_{+1}^L|^2+|H_1^L|^2+|H_{+1}^R|^2+|H_1^R|^2)`$ $`2\mathrm{sin}^2\theta _K\mathrm{sin}^2\theta _+[\mathrm{cos}2\varphi Re(H_{+1}^RH_1^R+H_{+1}^LH_1^L)`$ $`\mathrm{sin}2\varphi Im(H_{+1}^RH_1^R+H_{+1}^LH_1^L)]`$ $`\mathrm{sin}2\theta _K\mathrm{sin}2\theta _+[\mathrm{cos}\varphi Re(H_{+1}^RH_0^R+H_1^RH_0^R+H_{+1}^LH_0^L+H_1^LH_0^L)`$ $`\mathrm{sin}\varphi Im(H_{+1}^RH_0^RH_1^RH_0^R+H_{+1}^LH_0^LH_1^LH_0^L)]`$ $`2\mathrm{sin}^2\theta _K\mathrm{cos}\theta _+(|H_{+1}^R|^2|H_1^R|^2|H_{+1}^L|^2+|H_1^L|^2)`$ $`+2\mathrm{sin}\theta _+\mathrm{sin}2\theta _K[\mathrm{cos}\varphi Re(H_{+1}^RH_0^RH_1^RH_0^RH_{+1}^LH_0^L+H_1^LH_0^L)`$ $`\mathrm{sin}\varphi Im(H_{+1}^RH_0^R+H_1^RH_0^RH_{+1}^LH_0^LH_1^LH_0^L)]\}.`$ If we integrate out the angles $`\theta _K`$ and $`\theta _+`$, we get the $`\varphi `$ distribution $`{\displaystyle \frac{d\mathrm{\Gamma }}{d\varphi }}`$ $`=`$ $`{\displaystyle }{\displaystyle \frac{\alpha ^2G_F^2g_{K^{}K\pi }^2\sqrt{\lambda }p^2m_b^2|V_{tb}V_{ts}^{}|^2}{9\times 16(2\pi )^8m_B^3l^2[(p^2m_K^{}^2)^2+m_K^{}^2\mathrm{\Gamma }_K^{}^2]}}\{|H_0^R|^2+|H_{+1}^R|^2+|H_1^R|^2`$ (89) $`|H_0^L|^2+|H_{+1}^L|^2+|H_1^L|^2\mathrm{cos}2\varphi Re(H_{+1}^RH_1^R+H_{+1}^LH_1^L)`$ $`+\mathrm{sin}2\varphi Im(H_{+1}^RH_1^R+H_{+1}^LH_1^L)\}dp^2dl^2.`$ Even if the new physics gives the same total decay rate for $`bs\gamma `$ compared to the SM, i.e., we cannot see new physics from the $`bs\gamma `$ decay, we can still tell new physics effects from the angular distribution of $`BK\pi \mathrm{}^+\mathrm{}^{}`$. If we integrate out the angles $`\theta _+`$ and $`\varphi `$, we get the $`\theta _K`$ distribution $`{\displaystyle \frac{d\mathrm{\Gamma }}{d\mathrm{cos}\theta _K}}`$ $`=`$ $`{\displaystyle }{\displaystyle \frac{(2\pi )\alpha ^2G_F^2g_{K^{}K\pi }^2\sqrt{\lambda }p^2m_b^2|V_{tb}V_{ts}^{}|^2}{3\times 64(2\pi )^8m_B^3l^2[(p^2m_K^{}^2)^2+m_K^{}^2\mathrm{\Gamma }_K^{}^2]}}\{2\mathrm{cos}^2\theta _K(|H_0^R|^2`$ (91) $`+|H_0^L|^2)+\mathrm{sin}^2\theta _K(|H_{+1}^R|^2+|H_1^R|^2+|H_{+1}^L|^2+|H_1^L|^2)\}dp^2dl^2.`$ Taking the narrow resonance limit of $`K^{}`$ meson, i.e., using the equations $`\mathrm{\Gamma }_K^{}={\displaystyle \frac{g_{K^{}K\pi }^2m_K^{}}{48\pi }},`$ (92) $`\underset{\mathrm{\Gamma }_K^{}0}{lim}{\displaystyle \frac{\mathrm{\Gamma }_K^{}m_K^{}}{(p^2m_K^{}^2)^2+m_K^{}^2\mathrm{\Gamma }_K^{}^2}}=\pi \delta (p^2m_K^{}^2),`$ (93) we can perform the integration over $`p^2`$ and obtain the double differential branching ratio with respect to dilepton mass squared $`l^2`$ and azimuthal angle $`\varphi `$, $`{\displaystyle \frac{dBr}{dl^2d\varphi }}`$ $`=`$ $`\tau _B{\displaystyle \frac{\alpha ^2G_{F}^{}{}_{}{}^{2}}{384\pi ^5}}\sqrt{\lambda }{\displaystyle \frac{m_{b}^{}{}_{}{}^{2}}{m_{B}^{}{}_{}{}^{3}l^2}}|V_{ts}V_{tb}|^2{\displaystyle \frac{1}{2\pi }}\{|H_0^R|^2+|H_{+1}^R|^2+|H_1^R|^2`$ (96) $`+|H_0^L|^2+|H_{+1}^L|^2+|H_1^L|^2\mathrm{cos}2\varphi Re(H_{+1}^RH_1^R+H_{+1}^LH_1^L)`$ $`+\mathrm{sin}2\varphi Im(H_{+1}^RH_1^R+H_{+1}^LH_1^L)\},`$ and $`{\displaystyle \frac{dBr}{dl^2}}`$ $`=`$ $`\tau _B{\displaystyle \frac{\alpha ^2G_{F}^{}{}_{}{}^{2}}{384\pi ^5}}\sqrt{\lambda }{\displaystyle \frac{m_{b}^{}{}_{}{}^{2}}{m_{B}^{}{}_{}{}^{3}l^2}}|V_{ts}V_{tb}|^2\{|H_0^R|^2+|H_{+1}^R|^2+|H_1^R|^2`$ (98) $`+|H_0^L|^2+|H_{+1}^L|^2+|H_1^L|^2\},`$ where $`\tau _B`$ is the life time of $`B`$ meson, and we replace all $`p`$ by $`m_K^{}`$ due to the $`\delta `$ function. We further define the distribution $`r(\varphi ,\widehat{s})`$ as $`r(\varphi ,\widehat{s})`$ (99) $``$ $`\left[{\displaystyle \frac{dBr}{dl^2d\varphi }}\right]/\left[{\displaystyle \frac{dBr}{dl^2}}\right]`$ (100) $`=`$ $`{\displaystyle \frac{1}{2\pi }}\left\{1{\displaystyle \frac{\mathrm{cos}2\varphi Re(H_{+1}^RH_1^R+H_{+1}^LH_1^L)\mathrm{sin}2\varphi Im(H_{+1}^RH_1^R+H_{+1}^LH_1^L)}{|H_0^R|^2+|H_{+1}^R|^2+|H_1^R|^2+|H_0^L|^2+|H_{+1}^L|^2+|H_1^L|^2}}\right\},`$ (101) where $`\widehat{s}=l^2/m_B^2`$. The distribution $`r(\varphi ,\widehat{s})`$ is the probability for finding $`K`$ meson per unit radian region in the direction of azimuthal angle $`\varphi `$. Therefore $`r(\varphi ,\widehat{s})`$ oscillates around its average value given by $`\frac{1}{2\pi }0.16`$. ## III Numerical Analyses and Conclusions In the numerical calculations, we use the form factors calculated in Ref. . They are listed in Table I for zero momentum transfer. The revolution formula for these form factors is $$f_i(l^2)=\frac{f_i(0)}{1\sigma _1l^2+\sigma _2l^4},$$ (102) where $`l^2=(p_\mathrm{}^++p_{\mathrm{}^{}})^2`$. The corresponding values $`\sigma _1`$ and $`\sigma _2`$ for each form factors are also listed in Table 1. The analytic Wilson coefficients $`C_7^{eff}(\mu )`$, $`C_9^{eff}(\mu )`$, and $`C_{10}(\mu )`$ in the SM are given in Ref. . Under the leading logarithmic approximation, we get the numerical results at $`\mu =m_b`$: $$C_7^{eff}=0.311,C_{10}=4.546,$$ (103) and to the next-to-leading order, $`C_9^{eff}`$ $`=`$ $`4.153+0.381g(m_c/m_b,\widehat{s})+0.033g(1,\widehat{s})+0.032g(0,\widehat{s}),`$ (104) where $`\widehat{s}=l^2/m_b^2`$. The function $`g(z,\widehat{s})`$ can be found in Ref. . Here for numerical evaluation, we use $`m_{top}=175`$ GeV, $`m_b=4.8`$ GeV, $`m_c=1.4`$ GeV, $`\mathrm{\Lambda }_{QCD}=214`$ MeV. We include the $`J/\psi `$ contribution as done in , $$C_9^{eff}C_9^{eff}=C_9^{eff}C^0\kappa \frac{3\pi \mathrm{\Gamma }[\psi \mathrm{}^+\mathrm{}^{}]m_\psi }{\alpha ^2(l^2m_\psi ^2+im_\psi \mathrm{\Gamma }_\psi )}.$$ (105) where $`\kappa =2.3`$ and $`C^{(0)}=0.381.`$ The decay width for inclusive $`bs\gamma `$ decay in terms of operators $`O_{7L}`$ and $`O_{7R}`$ is given by $$\mathrm{\Gamma }(bs\gamma )=\frac{G_F^2m_b^5}{32\pi ^4}\alpha _{em}|V_{ts}^{}V_{tb}|^2\left(|C_{7L}(m_b)|^2+|C_{7R}(m_b)|^2\right).$$ (106) It is convenient to normalize this radiative partial width to the semileptonic rate $$\mathrm{\Gamma }(bce\overline{\nu })=\frac{G_F^2m_b^5}{192\pi ^3}|V_{cb}|^2f(m_c/m_b)\left[1\frac{2}{3\pi }\alpha _s(m_b)g(m_c/m_b)\right],$$ (107) where $`f(x)=18x^224x^4\mathrm{ln}x+8x^6x^8`$ represents a phase space factor, and the function $`g(x)`$ encodes next-to-leading order QCD correction effects . In terms of the ratio $`R`$, $$R\frac{\mathrm{\Gamma }(bs\gamma )}{\mathrm{\Gamma }(bce\overline{\nu })}=\frac{6}{\pi }\frac{|V_{ts}^{}V_{tb}|^2}{|V_{cb}|^2}\frac{\alpha _{em}}{f(m_c/m_b)}\frac{|C_{7L}(m_b)|^2+|C_{7R}(m_b)|^2}{1\frac{2}{3\pi }\alpha _s(m_b)g(m_c/m_b)},$$ (108) the $`bs\gamma `$ branching fraction is obtained by $$(bs\gamma )(BX_cl\nu )_{\mathrm{exp}.}\times R(0.105)\times R.$$ (109) For $`(bs\gamma )`$, we use the present experimental value of the branching fraction for $`BX_s\gamma `$ decay, $$(BX_s\gamma )=(3.15\pm 0.35\pm 0.32\pm 0.26)\times 10^4.$$ (110) Constrained by this experiment, we derive from Eq. (108) $$|C_{7L}(m_b)|^2+|C_{7R}(m_b)|^2=0.081\pm 0.014.$$ (111) In general, we can parameterize $`C_{7L}`$ and $`C_{7R}`$ as follows by introducing parameters $`(x,u,v)`$, $`C_{7L}`$ $`=`$ $`\sqrt{|C_{7L}(m_b)|^2+|C_{7R}(m_b)|^2}\mathrm{cos}x\mathrm{exp}i(u+v),`$ (112) $`C_{7R}`$ $`=`$ $`\sqrt{|C_{7L}(m_b)|^2+|C_{7R}(m_b)|^2}\mathrm{sin}x\mathrm{exp}i(uv),`$ (113) where $`u`$ is a common phase of $`C_{7L}`$ and $`C_{7R}`$, and $`v`$ denotes the relative phase between $`C_{7L}`$ and $`C_{7R}`$. In Figs. 1–6, we show the distribution $`r(\varphi ,\widehat{s})`$ for different sets of $`(x,u,v)`$. The minimum of the invariant mass is set to be $`0.7`$ GeV in the figures. We can understand the qualitative features in the region of small invariant mass by comparing with an approximate formula for the azimuthal angle distribution. By using Eq. (113), we can show that in the small invariant mass limit, $`r(\varphi ,\widehat{s})`$ defined in Eq. (101) is written as, $`r(\varphi ,\widehat{s})`$ $``$ $`{\displaystyle \frac{1}{2\pi }}\left\{1+\mathrm{cos}2\varphi {\displaystyle \frac{Re(C_{7R}C_{7L}^{})}{|C_{7R}|^2+|C_{7L}|^2}}\mathrm{sin}2\varphi {\displaystyle \frac{Im(C_{7R}C_{7L}^{})}{|C_{7R}|^2+|C_{7L}|^2}}\right\}`$ (114) $`=`$ $`{\displaystyle \frac{1}{2\pi }}\left\{1{\displaystyle \frac{1}{2}}\mathrm{sin}2x\mathrm{cos}2(\varphi v)\right\}.`$ (115) The equation follows from the fact that the helicity amplitudes are dominated by the two coefficients $`C_{7R}`$ and $`C_{7L}`$ in the region of low invariant mass, $`H_{+1}^{L,R}`$ $``$ $`4g_+C_{7R}\sqrt{\lambda },`$ (116) $`H_1^{L,R}`$ $``$ $`4g_+C_{7L}\sqrt{\lambda },`$ (117) $`H_0^{L,R}`$ $``$ $`0.`$ (118) The SM case ($`C_{7R}0`$) corresponds to $`(x,u,v)=(0,0,0)`$, and $`r(\varphi ,\widehat{s})`$ is shown in Fig. 1. In the SM there are only small phase shifts from the $`g(z,\widehat{s})`$ in Eq. (104) , which are practically negligible because of $`Im(H_{+1}^RH_1^R+H_{+1}^LH_1^L)0`$. The last term of (101) vanishes for any $`\widehat{s}`$. It is shown in Fig. 1 that there is only $`(\mathrm{cos}2\varphi )`$ behavior for larger $`\widehat{s}`$. We can also note that as $`\widehat{s}`$ is getting smaller, the $`\varphi `$ dependence even vanishes. This is consistent with formula (115), since $`x=0`$ in the SM. Another extreme case, $`C_{7L}=0`$, is shown in Fig. 2. There still remains $`\varphi `$ dependence even in low invariant mass region. We checked that $`\varphi `$ dependence vanishes by going further to smaller invariant mass $`l1`$ GeV, which is not shown in the figure. This shows that there is large contribution from $`C_9`$ and $`C_{10}`$ even for rather low invariant mass $`l1`$ GeV. For larger $`\widehat{s}`$, near 0.4, there is some disorder appearing in Fig. 2. It represents the interference effect of the short distance contribution with the long distance contribution from $`J/\psi `$ resonance. If $`\frac{|C_{7R}|}{|C_{7L}|}=O(1)`$, the approximate formula (115) works qualitatively well \[see Figs. 3–6\]. There we change the relative phase of $`C_{7R}`$ and $`C_{7L}`$ by setting $`\frac{|C_{7R}|}{|C_{7L}|}=1`$. In Fig. 3, $`u=v=0`$, then there is no imaginary part. We can read from Fig. 3 the $`(\mathrm{cos}2\varphi )`$ behavior for $`\frac{C_{7R}}{C_{7L}}=1`$ in the region of small $`\widehat{s}`$. For larger $`\widehat{s}`$, there is interference from the $`C_9`$ and $`C_{10}`$ contributions, and the resulting figure is not so simple. From Fig. 4, we can see the $`(+\mathrm{cos}2\varphi )`$ behavior for $`\frac{C_{7R}}{C_{7L}}=1`$ in the region of small $`\widehat{s}`$. This is consistent with the approximate formula (115). It is the inverse case of Fig. 3. Finally we introduce CP violating phase $`v`$ between $`C_{7L}`$ and $`C_{7R}`$, which leads to the phase shifts. In Figs. 5 and 6, we choose $`v=\pm \pi /8`$. According to Eq. (115), it amounts to $`\pm \pi /4`$ in the phase shift, which can be seen in Figs. 5 and 6. We do not show figures for non-zero values of $`u`$, which is the relative phase between $`C_{7i}`$’s and $`C_9`$ ($`C_{10}`$). The non-zero value of this angle $`u`$ will not change the $`r(\widehat{s},\varphi )`$ behavior at low $`\widehat{s}`$ \[see Eq. (115)\], but will change it at higher $`\widehat{s}`$. This area is affected by the interference of $`C_{7i}`$’s and $`C_9`$ ($`C_{10}`$). Using eq.(98), we do the integration with $`l`$ from 0.4 GeV to 1.2 GeV, we get the branching ratio of $`BK\pi \mathrm{}^+\mathrm{}^{}`$ at this region: $`1\times 10^7`$. From the figures we know that in the above region, it is effective to distinguish the new physics contribution. The number of B mesons we need is around $`10^{10}`$, which can not be produced in the current B factories, but possible in the future LHC-B etc. More concretely, dividing the region of $`\varphi `$ into 10 bins, we expect $`10^2`$ events in each bin in the standard model. If the distribution follows from the formulae Eq.(115) with $`\mathrm{sin}2x=1`$, the numbers of the event of each bin are no more flat and it oscillates between 50 and 150. If this is the case, we can surely distinguish the distribution from the flat one of the standard model. To summarize, we studied the angular distribution of $`BK^{}(K\pi )\mathrm{}^+\mathrm{}^{}`$. We showed that the azimuthal angle $`(\varphi )`$ distribution is very useful for probing possible new physics effects and for confirming the SM through this flavor-changing neutral current process. Here $`\varphi `$ is the angle between the decay plane of $`(K\pi )`$ and the decay plane of $`(\mathrm{}^+\mathrm{}^{})`$. In particular, if the two operators $`O_{7L}`$ and $`O_{7R}`$, which contribute to $`BK^{}\gamma `$, are equally important, then the $`\varphi `$ dependence is significant. In the SM case, there is only a weak $`(\mathrm{cos}2\varphi )`$ dependence for the region of small $`\widehat{s}`$, but the term proportional to $`(\mathrm{cos}2\varphi )`$ becomes dominant for the region of larger $`\widehat{s}`$. When new physics is introduced without changing the decay rate of the $`bs\gamma `$, we can nontheless have quite different angular distribution for $`BK\pi \mathrm{}^+\mathrm{}^{}`$. We also showed that the phase shift results in the appearance of $`(\mathrm{sin}2\varphi )`$ term, the latter-thus being a clear signature of the presence of CP violating phase. Even if we cannot probe new physics from $`BK^{}\gamma `$, it is possible to see the new physics effects through the azimuthal angle distribution of $`BK\pi \mathrm{}^+\mathrm{}^{}`$. We also note that $`C_9`$ and $`C_{10}`$ are about ten times larger than $`C_7`$’s. Therefore, even in the region of small dilepton mass, their effect cannot be neglected. In our analysis, their effect has been fully incorporated. ## Acknowledgement We thank G. Cvetic for careful reading of the manuscript and his valuable comments. The work of C.S.K. was supported in part by grant No. 1999-2-111-002-5 from the Interdisciplinary Research Program of the KOSEF, in part by the BSRI Program, Ministry of Education, Project No. 99-015-DI0032, and in part by Sughak program of Korean Reasearch Foundation, Project No. 1997-011-D00015. The work of Y.G.K. is supported by KOSEF Postdoctoral Program. The work of T.M. is supported by Grant-in-Aid for Scientific Research on Priority Areas (Physics of CP violation) and the work of C.L. is supported by JSPS.
warning/0001/math0001151.html
ar5iv
text
# Deformations of algebras over operads and Deligne’s conjecture ## 1 Generalities on operads ### 1.1 Polynomial functors, operads, algebras The material of this subsection is very well-known (see for ex. \[GiKa\], \[GeJ\], \[GeKa\], \[Ma\]). We present it here for completeness and in order to fix the notation. Let $`k`$ be a field of characteristic zero. All vector spaces below will be $`k`$-vector spaces unless we say otherwise. We fix a category $`𝒞`$ which is assumed to be $`k`$-linear abelian symmetric monoidal and closed under infinite sums and products. We will also assume that it has internal $`Hom^{}s`$. Our main examples will be the category of $`k`$-vector spaces, the category $`Vect_𝐙`$ of $`𝐙`$-graded vector spaces (with Koszul rule of signs), and the category of complexes of $`k`$-vector spaces. Suppose we have a collection of representations $`F=(F_n)_{n0}`$ of the symmetric groups $`S_n,n=0,1,\mathrm{}`$ in $`𝒞`$ (i.e. we have a sequence of objects $`F_n`$ together with an action of the group $`S_n`$ on $`F_n`$ for each $`n`$). ###### Definition 1 A polynomial functor $`F:𝒞𝒞`$ is defined on objects by the formula $$F(V)=_{n0}(F_nV^n)_{S_n}$$ where for a group $`H`$ and an $`H`$-module $`W`$ we denote by $`W_H`$ the space of coinvariants. Functor $`F`$ is defined on morphisms in an obvious way. Notice that having a sequence $`F_n`$ as above we can define $`F_I`$ for any finite set $`I`$ using isomorphisms of $`I`$ with the standard set $`\{1,\mathrm{},|I|\}`$, where $`|I|`$ is the cardinality of $`I`$. Thus $`F_{\{1,\mathrm{},n\}}=F_n`$. Technically speaking, we consider a functor $`\mathrm{\Phi }`$ from the groupoid of finite sets (morphisms are bijections) to the symmetric monoidal category $`𝒞`$. Then we set $`F_I=\mathrm{\Phi }(I)`$. Polynomial functors on $`𝒞`$ form a category $`𝒫`$ if we define morphisms between two such functors $`F`$ and $`G`$ as a vector space of $`S_n`$-intertwiners $$Hom(F,G)=\underset{n=0}{\overset{\mathrm{}}{}}Hom_{S_n}(F_n,G_n)$$ There is a composition operation $``$ on polynomial functors such that $`(FG)(V)`$ is naturally isomorphic to $`F(G(V))`$ for any $`V𝒞`$. We also have a polynomial functor $`\mathrm{𝟏}`$ such that $`\mathrm{𝟏}_1=1_𝒞`$ and $`\mathrm{𝟏}_n=0`$ for all $`n1`$. Here $`1_𝒞`$ is the unit object in the monoidal category $`𝒞`$. It is easy to see that in this way we get a monoidal structure on $`𝒫`$. ###### Definition 2 An operad in $`𝒞`$ is a monoid in the monoidal category $`𝒫`$. In other words it is a polynomial functor $`R𝒫`$ together with morphisms $`m:RRR`$ and $`u:\mathrm{𝟏}R`$ satisfying the associativity and the unit axioms. To shorten the notation we will denote the operad $`(R,m,u)`$ simply by $`R`$. An operad $`R`$ gives rise to a triple in the category $`𝒞`$. There is the notion of an algebra over a triple in a category. Hence we can use it in order to give a definition of an algebra in $`𝒞`$ over the operad $`R`$. It is given by an object $`V𝒞`$ and a morphism $`R(V)V`$ satisfying natural properties of compatibility with the structure of a triple. Equivalently, $`V`$ is an $`R`$-algebra iff there is a morphism of operads $`Rnd(V)`$, where $`nd(V)`$ is the endomorphism operad of $`V`$ defined by $`(nd(V))_n=\underset{¯}{Hom}(V^n,V)`$, and $`\underset{¯}{Hom}`$ denotes the internal $`Hom`$ in $`𝒞`$. The category of $`R`$-algebras will be denoted by $`Ralg`$. There are two adjoint functors $`Forget_R:Ralg𝒞`$ and $`Free_R:𝒞Ralg`$ such that $`Forget_RFree_R=R`$. ###### Definition 3 For $`XOb(𝒞)`$ we call $`Free_R(X)`$ the free $`R`$-algebra generated by $`X`$. We remind to the reader that there are operads $`As`$, $`Lie`$, $`Comm`$ such that the algebras over them in the category of vector spaces are associative, Lie and commutative algebras correspondingly. ### 1.2 Colored operads There is a generalization of the notion of operad. It is useful in order to describe in operadic terms pairs (associative algebra A, A-module), homomorphisms of algebras over operads, etc. Let $`I`$ be set. We consider the category $`𝒞^I`$ consisting of families $`(V_i)_{iI}`$ of objects of $`𝒞`$. A polynomial functor $`F:𝒞^I𝒞^I`$ is defined by the following formula: $$(F((V_i)_{iI}))_j=_{a:I𝐙_\mathrm{𝟎}}F_{a,j}_{_iS_{a(i)}}_{iI}(V_i^{a(i)})$$ where $`a:I𝐙_+`$ is a map with the finite support, and $`F_{a,j}`$ is a representation in $`𝒞`$ of the group $`_{iI}S_{a(i)}`$. Polynomial functors in $`𝒞^I`$ form a monoidal category with the tensor product given by the composition of functors. ###### Definition 4 A colored operad is a monoid in this category. Similarly to the case of usual operads it defines a triple in the category $`𝒞^I`$. Therefore we have the notion of an algebra over a colored operad. There exists a colored operad $`𝒪𝒫`$ such that the category of $`𝒪𝒫`$-algebras is equivalent to the category of operads. Namely, let us consider the forgetful functor $`Operads𝒫`$. It has a left adjoint functor. Thus we have a triple in $`𝒫`$. As we have noticed before, the category $`𝒫`$ can be described as a category of sequences $`(P_n)_{n0}`$ of $`S_n`$-modules. Then using the representation theory of symmetric groups, we conclude that the category $`𝒫`$ is equivalent to the category $`𝒞^{I_0}`$, where $`I_0`$ is the set of all Young diagrams (partitions). Hence a polynomial functor $`F:𝒫𝒫`$ can be described as a collection $`F_{(m_i),n}`$ of the representations of the groups $`S_{n,(m_k)}:=S_n\times _{k0}(S_{m_k}S_k^{m_k})`$, where $``$ denotes the semidirect product of groups. Having these data we can express any polynomial functor $`F`$ on $`𝒫`$ by the formula: $$(F((U_k)_{k0}))_n=\underset{(m_k)}{}F_{(m_k),n}_{S_{1,(m_k)}}\underset{k0}{}(U_k^{m_k})$$ In particular, one has a functor $`𝒪𝒫:𝒫𝒫`$, which is the composition of the forgetful functor $`Operads𝒫`$ with its adjoint. It gives rise to an $`I_0`$-colored operad $`𝒪𝒫=(𝒪𝒫_{(m_i),n})`$. We will describe it explicitly in the subsection devoted to trees. ### 1.3 Non-linear operads We remark that operads and algebras over operads can be defined for any symmetric monoidal category $`𝒞`$, not necessarily $`k`$-linear. In particular, we are going to use operads in the categories of sets, topological spaces, etc. Namely, an operad in $`𝒞`$ is a collection $`(F_n)_{n0}`$ of objects in $`𝒞`$, each equipped with an $`S_n`$-action, as well as composition maps: $$F_nF_{k_1}\mathrm{}F_{k_n}F_{k_1+\mathrm{}+k_n}$$ for any $`n0,k_1,\mathrm{},k_n0`$. Another datum is the unit, which is a morphism $`\mathrm{𝟏}_𝒞F_1`$. All the data are required to satisfy certain axioms (see \[Ma\]). Analogously one describes colored operads and algebras over operads. Notice that in this framework one cannot speak about polynomial functors and free algebras. This approach has some advantages and drawbacks (like description of analytic functions in terms of Taylor series vs their description in terms of Taylor coefficients). ### 1.4 Pseudo-tensor categories The notion of colored operad has been rediscovered many times. In \[BD\] the notion of pseudo-tensor category was introduced as a generalization of the notion of symmetric monoidal (=tensor) category. The terminology stresses the similarity of operads with tensor categories. Similar notion was introduced in \[B\] under the name multi-linear category. This notion is essentially equivalent to the notion of colored operad. We recall it below, using the name suggested in \[BD\]. ###### Definition 5 A pseudo-tensor category is given by the following data: 1. A class $`𝒜`$ called the class of objects, and a symmetric monoidal category $`𝒱`$ called the category of operations. 2. For every finite set $`I`$, a family $`(X_i)_{iI}`$ of objects, and an object $`Y`$, an object $`P_I((X_i),Y)𝒱`$ called the space of operations from $`(X_i)_{iI}`$ to $`Y`$. 3. For any map of finite sets $`\pi :JI`$ , two families of objects $`(Y_i)_{iI},(X_j)_{jJ}`$ and an object $`Z`$, a morphism in $`𝒱`$ $$P_I((Y_i),Z)(_iP_{\pi ^1(i)}((X_{j_i}),Y_i))P_J((X_j),Z)$$ called composition of operations. Here we denote by $``$ the tensor product in $``$. 4. For an $`1`$-element set $``$ and an object $`X`$, a unit morphism $`\mathrm{𝟏}_𝒱P_{}((X),X)`$. These data are required to satisfy natural conditions. In particular, compositions of operations are associative with respect to morphisms of finite sets, and the unit morphisms satisfy the properties analogous to those of the identity morphisms (see \[BD\] or \[So\] for details). If $`𝒜`$ is a set, then a pseudo-tensor category is exactly the same as an $`𝒜`$-colored operad in the tensor category $`𝒱`$. If we take $`𝒱`$ to be the category of sets, and take $`I`$ above to be $`1`$-element sets only, we obtain a category with the class of objects equal to $`𝒜`$. Colored operad with one color gives rise to an ordinary operad. A symmetric monoidal category $`𝒜`$ produces the colored operad with $`P_I((X_i),Y)=Hom_𝒜(_iX_i,Y)`$. The notion of pseudo-tensor category admits a generalization to the case when no action of symmetric group is assumed. This means that we consider sequences of objects instead of families (see \[So\]). The new notion generalizes monoidal categories. In terms of the next subsection this would mean that one uses planar trees instead of all trees. One can make one step further generalizing braided categories. This leads to colored braided operads (or pseudo-braided categories). In this case trees in $`𝐑^3`$ should be used. The deformation theory of such structures can be developed along the lines of present paper. It will be explained in detail elsewhere. ### 1.5 Trees ###### Definition 6 A tree $`T`$ is defined by the following data: 1) a finite set $`V(T)`$ whose elements are called vertices; 2) a distinguished element $`root_TV(T)`$ called root vertex; 3) subsets $`V_i(T)`$ and $`V_t(T)`$ of $`V(T)\{root_T\}`$ called the set of internal vertices and the set of tails respectively. Their elements are called internal and tail vertices respectively; 4) a map $`N=N_T:V(T)V(T)`$. These data are required to satisfy the following properties: a) $`V(T)=\{root_T\}V_i(T)V_t(T)`$; b) $`N_T(root_T)=root_T`$, and $`N_T^k(v)=root_T`$ for all $`vV(T)`$ and $`k1`$; c) $`N_T(V(T))V_t(T)=\mathrm{}`$ ; d) there exists a unique vertex $`vV(T),vroot_T`$ such that $`N_T(v)=root_T`$. We denote by $`|v|`$ the $`valency`$ of a vertex $`v`$, which we understand as the cardinality of the set $`N_T^1(v)`$. We call the pairs $`(v,N(v))`$ $`edges`$ in the case if $`vroot_T`$. If both elements of the pair belong to $`V_i(T)`$ we call the corresponding edge $`internal`$. The only edge $`e_r`$ defined by the condition d) above is called the $`root`$ $`edge`$. All edges of the type $`(v,N(v)),vV_t(T)`$ are called $`tail`$ $`edges`$. We use the notation $`E_i(T)`$ and $`E_t(T)`$ for the sets of internal and tail edges respectively. We have a decomposition of the set of all edges $`E(T)=E_i(T)(E_t(T)\{e_r\})`$. There is a unique tree $`T_e`$ such that $`|V_t(T_e)|=1`$ and $`|V_i(T_e)|=0`$. It has the only tail edge which is also the root edge. A $`numbered`$ tree with $`n`$ tails is by definition a tree $`T`$ together with a bijection of sets $`\{1,\mathrm{},n\}V_t(T)`$. We can picture trees as follows Let $`R`$ be an operad. Any tree $`T`$ gives a natural way to compose elements of $`R`$, $`comp_T:_{iV_i(T)}R_{N^1(v)}R_{V_t(T)}`$. Let us return to the colored operad $`𝒪𝒫`$ and give its description using the language of trees. Namely, $`𝒪𝒫_{(m_i),n}`$ is a $`k`$-vector space generated by the isomorphism classes of trees $`T`$ such that: a) $`T`$ has $`n`$ tails numbered from $`1`$ to $`n`$; b) $`T`$ has $`_im_i`$ internal vertices all numbered in such a way that first $`m_0`$ vertices have valency $`0`$, and they are numbered from $`1`$ to $`m_0`$, next $`m_1`$ internal vertices have valency $`1`$, and they are numbered from $`1`$ to $`m_1`$, and so on; c) for every internal vertex $`vV_i(T)`$ the set of incoming edges $`N_T^1(v)`$ is also numbered. An action of the group $`S_{(m_k),n}`$ is defined naturally: the factor $`S_n`$ permutes numbered tails, the factor $`S_{m_k}`$ permutes numbered internal vertices and the factor $`S_k^{m_k}`$ permutes their incoming edges numbered from $`1`$ to $`k`$. The composition is given by the procedure of inserting of a tree into an internal vertex of another one. The new numeration is clear. We leave these details as well as the proof of the following proposition to the reader. ###### Proposition 1 The category of $`𝒪𝒫`$-algebras is equivalent to the category of $`k`$-linear operads. Let $`F`$ be a polynomial functor on $`𝒞`$ (see Section 1.1). Let us consider a category $`𝒞_F`$ objects of which are pairs $`(V,\varphi :F(V)V)`$ where $`V`$ is an object of $`𝒞`$ and $`\varphi `$ is a morphism in $`𝒞`$. Morphisms of pairs are defined in the natural way. The following lemma is easy to prove. ###### Lemma 1 The category $`𝒞_F`$ is equivalent to the category of $`Free_{𝒪𝒫}(F)`$-algebras. We call $`P=Free_{𝒪𝒫}(F)`$ the $`free`$ $`operad`$ generated by $`F`$. Components $`P_n`$ of the functor $`P`$ can be defined explicitly as follows. Let $`Tree(n)`$ denotes the groupoid of numbered trees with $`n`$ tails, $`|Tree(n)|`$ denotes the set of classes of isomorphisms of these trees . We denote the class of isomorphism of $`T`$ by $`[T]`$. Then we have $$P_n=Free_{𝒪𝒫}(F)_n=_{[T]|Tree(n)|}(_{vV_i(T)}F_{N^1(v)})_{AutT}$$ ## 2 Deformations and differentials in free operads Let $`F`$ be a polynomial functor, $`P=Free_{𝒪𝒫}(F)`$ be the corresponding free operad. Let $`g_P`$ be the Lie algebra (in the symmetric monoidal category $`𝒞`$) of derivations of the operad $`P`$. Then, as an object of $`𝒞`$: $$g_P=\underset{n0}{}\underset{¯}{Hom}(F_n,P_n)^{S_n}$$ where $`W^H`$ denotes the space of $`H`$-invariants of an $`H`$-module $`W`$ and $`\underset{¯}{Hom}`$ denotes the internal $`Hom`$ in $`𝒞`$. This observation follows from the fact that $`Hom_𝒫(F,Forget_{𝒪𝒫}(G))=Hom_{𝒪𝒫alg}(Free_{𝒪𝒫}(F),G)`$. ¿From now on we suppose that $`𝒞`$ is the category $`Vect_𝐙`$ of $`𝐙`$-graded vector spaces. Then $`g_P`$ is a graded Lie algebra with the graded components $`g_P^n`$. ###### Definition 7 A structure of a differential-graded operad on $`P`$ which is free as a graded operad is given by an element $`d_Pg_P^1`$ such that $`[d_P,d_P]=0`$. The definition means that $`P`$ can be considered as an operad in the symmetric monoidal category of complexes, and it is free as an operad in the category $`Vect_𝐙`$. Sometimes we will denote the corresponding operad in the category of complexes by $`\widehat{P}`$. One of our purposes will be to use $`\widehat{P}`$ for constructing resolutions of dg-operads, and subsequently the deformation theory of algebras over them. ###### Definition 8 A dg-algebra over $`(P,d_P)`$ (or simply over $`P`$) is an algebra over $`\widehat{P}`$ in the category of complexes. Notice that the deformation theory of the pair $`(P,d_P)`$ is the same as the deformation theory of $`d_P`$ (since $`P`$ is free and therefore rigid). ###### Definition 9 The formal pointed dg-manifold associated with the differential graded Lie algbera $`(g_P,[d_P,])`$ controls the deformation theory of $`(P,d_P)`$. Now we are going to describe the deformation theory of dg-algebras over $`P`$ ($`\widehat{P}`$-algebras). In what follows we will often speak about points of $`𝐙`$-graded manifolds. It will always mean $`\mathrm{\Lambda }`$-points, where $`\mathrm{\Lambda }`$ is an auxiliary $`𝐙`$-graded commutative associative algebra (in general without the unit). In the case of formal manifolds we take $`\mathrm{\Lambda }`$ to be nilpotent. Let us describe the formal pointed dg-manifold controlling deformations of a $`P`$-algebra $`V`$. We have the following graded vector space $$=(P,V)=(\underset{¯}{Hom}(V,V))[1]\underset{¯}{Hom}(F(V),V)$$ We denote by $`^n,n𝐙`$ the graded components of $``$. The structures of a complex on a graded vector space $`V`$ and an action of $`P`$ on $`V`$ define a point $`(d_V,\rho )^0=Hom_{Vect_𝐙}(k,)`$. We consider here $`d_V`$ and $`\rho `$ as morphisms of graded vector spaces. The equation $`d_V^2=0`$ and the condition of compatibility of $`d_V`$ and $`\rho `$ can be written in the form $`d_{}(d_V,\rho )=0`$, where $`d_{}(d_V,\rho )=(d_V^2,\xi (d_V,\rho ))^1`$, for some $`\xi (d_V,\rho )\underset{¯}{Hom}(F(V),V)`$. It is easy to see that the assignment $`(d_V,\rho )d_{}(d_V,\rho )`$ defines an odd vector field $`d_{}`$ on the “infinite-dimensional graded manifold” $``$. A zero of this vector field corresponds to a structure of a complex on $`V`$ together with a compatible structure of a dg-algebra over $`P`$. This gives a bijection between the set of zeros and the set of such structures. It is easy to check that $`[d_{},d_{}]=0`$. Therefore a formal neighborhood of a fixed point $`(d_V,\rho )`$ of $`d_{}`$ becomes a formal pointed dg-manifold. ###### Definition 10 The deformation theory of a dg-algebra $`V`$ is controlled by this formal pointed dg-manifold. ###### Remark 1 Operads are algebras over the colored operad $`𝒪𝒫`$. One can show that the deformation theories for an $`𝒪𝒫`$-algebra $`\widehat{P}`$ described in the last two definitions are in fact equivalent. Let $`P^{(0)}`$ be an operad in $`Vect_k`$. In order to define the deformation theory of $`P^{(0)}`$-algebras, one needs to choose a resolution $`\varphi :PP^{(0)}`$, where $`P=Free_{𝒪𝒫}(F)`$ as a $`𝐙`$-graded operad, and $`\varphi `$ is a quasi-isomorphism. One wants to be sure that the deformation theory does not depend on the choice of the resolution. This can be achieved by assuming that (see for ex. \[M2\]) : a) $`F`$ admits a filtration (as a polynomial functor) $`F=_{j1}F^{(j)},F^{(j)}F^{(j+1)}`$ such that $`d_P(F^{(0)})=0`$ and $`d_P(F^{(j)})Free_{𝒪𝒫}(F^{(j1)}),j1`$; b) $`\varphi :PP^{(0)}`$ is a an epimorphism. All resolutions used in the paper will satisfy these properties. ### 2.1 Example: $`A_{\mathrm{}}`$-operad and $`A_{\mathrm{}}`$-algebras Let $`VVect_𝐙`$ and $`m_n:V^nV[n2],n2`$ be a sequence of morphisms. It gives rise to an action on $`V`$ of the free operad $`P=Free_{𝒪𝒫}(F)`$ where $$F(V)=_{n2}V^n[n2].$$ Then $`F_n=k[S_n]m_nk[1]^{(n2)}`$. This notation means that we consider $`F_n`$ as a space (with the grading shifted by $`n2`$) of the regular representation of the group algebra of the symmetric group $`S_n`$. This space is generated by an element which we denote by $`m_n`$. The differential $`d_Pg_P`$ (equivalently, a structure of a dg-operad on $`P`$) is defined by the standard formulas: $$d_P(m_2)=0,$$ $$d_P(m_n)(v_1\mathrm{}v_n)=\underset{k+l=n}{}\pm m_k(v_1\mathrm{}v_im_l(v_{i+1}\mathrm{}v_{i+l})\mathrm{}v_n),n>2.$$ We do not specify signs in these well-known formulas. In Section 5 we propose a general framework allowing to fix signs in the formulas like the one above. ###### Definition 11 The dg-operad $`𝒜_{\mathrm{}}=(P,d_P)`$ is called the $`A_{\mathrm{}}`$-operad. Algebras over this dg-operad are called $`A_{\mathrm{}}`$-algebras. Deformations of an $`A_{\mathrm{}}`$-algebra $`A`$ are controlled by the truncated Hochschild complex $$C_+^{}(A,A)=\underset{n1}{}Hom_{Vect_𝐙}(A^n,A)[n]$$ More precisely, let $`A`$ be a graded vector space. We define a graded vector space of all Hochschild cochains of $`A`$ as $$C^{}(A,A)=\underset{n0}{}Hom_{Vect_𝐙}(A^n,A)[n]$$ Then $`C^{}(A,A)[1]`$ can be equipped with the structure of a graded Lie algebra with the Gerstenhaber bracket (the latter appears naturally if we interpret Hochschild cochains as coderivations of the free coalgebra $`_{n0}(A[1])^n`$). Let us consider an element $`m=(m_1,m_2\mathrm{})C_+^{}(A,A)[1]`$ of degree $`+1`$ such that $`[m,m]=0`$. Such an element defines a differenitial $`d=m_1`$ on $`A`$, and the sequence $`(m_2,m_3,\mathrm{})`$ gives rise to a structure of an $`A_{\mathrm{}}`$-algebra on $`(A,m_1)`$. Then we can make $`C^{}(A,A)`$ into a complex (Hochschild complex) with the differential $`d_m=[m,]`$. It is easy to see that in this way we get a differential graded Lie algebra (DGLA for short) $`(C^{}(A,A)[1],d_m)`$. The truncated Hochschild complex $`C_+^{}(A,A)[1]`$ is a DGLA subalgebra. According to the general theory (see \[Ko1\]) both DGLAs define formal pointed dg-manifolds, and therefore give rise to the deformation functors. This is a straightforward generalization of the well-known deformation theory of associative algebras. In a sense the full Hochschild complex controls deformations of the $`A_{\mathrm{}}`$-category with one object, such that its endomorphism space is equal to $`A`$. The deformation theory of $`A_{\mathrm{}}`$-categories is not in the scope of present paper. It will be explained elsewhere. Nevertheless we will refer to the formal dg-manifold associated with $`C^{}(A,A)[1]`$ as to the moduli space of $`A_{\mathrm{}}`$-categories. Similarly, the formal dg-manifold associated with $`C_+^{}(A,A)[1]`$ will be called the moduli space of $`A_{\mathrm{}}`$-algebras. (All the terminology assumes that we deform a given $`A_{\mathrm{}}`$-algebra $`A`$). The moduli space of $`A_{\mathrm{}}`$-algebras is the same as $`(𝒜_{\mathrm{}},A)`$ in the previous notation. Similarly we will denote the moduli space of $`A_{\mathrm{}}`$-categories by $`_{cat}(𝒜_{\mathrm{}},A)`$. The natural inclusion of DGLAs $`C_+^{}(A,A)[1]C^{}(A,A)[1]`$ induces a dg-map $`(𝒜_{\mathrm{}},A)_{cat}(𝒜_{\mathrm{}},A)`$ (dg-map is a morphism of dg-manifolds). Let us remark that the operad $`𝒜_{\mathrm{}}`$ gives rise to a free resolution of the operad $`As`$. Algebras over the latter are associative algebras without the unit. ###### Remark 2 It is interesting to describe deformation theories of free resolutions of the classical operads $`As`$, $`Lie`$, $`Comm`$. It seems that for an arbitrary free resolution $`P`$ of either of these operads the following is true: $`H^i(g_P)=0`$ for $`i0`$, $`H^0(g_P)=k`$. This one-dimensional vector space gives rise to the rescaling of operations, like $`m_n\lambda ^nm_n`$ in the case of $`A_{\mathrm{}}`$-algebras. ### 2.2 Homotopical actions of Lie algebras Let $`g`$ be a Lie algebra acting on a formal dg-manifold $`(Y,d_Y)`$. This means that we have a homomorphism of Lie algebras $`gDer(Y),\gamma \widehat{\gamma }`$ where $`Der(Y)`$ is the Lie algebra of vector fields on $`Y`$ preserving $`𝐙`$-grading an $`d_Y`$. We can make $`Z=Y\times g[1]`$ into a formal dg-manifold introducing an odd vector field by the following formula $$d_Z(y,\gamma )=(d_Y(y)+\widehat{\gamma },[\gamma ,\gamma ]/2)$$ Then $`[d_Z,d_Z]=0`$. We can make $`g[1]`$ into a formal dg-manifold using the odd vector field $`d_{g[1]}`$ arising from the Lie bracket. Then the natural projection $`(Z,d_Z)(g[1],d_{g[1]})`$ becomes a dg-bundle (cf. \[Ko2\]). This contsruction of a dg-bundle out of a group or Lie algebra acting on a dg-manifold motivates the following definition. ###### Definition 12 Let $`g`$ be a Lie algebra. Homotopical $`g`$-action on a formal dg-manifold $`(Y,d_Y)`$ is a dg-bundle $`\pi :(Z,d_Z)(g[1],d_{g[1]})`$ together with an isomorphism of dg-manifolds $`(\pi ^1(0),d_Z)(Y,d_Y)`$. ###### Remark 3 It was pointed out in \[Ko2\] that in this case $`g`$ acts on the cohomology of all complexes naturally associated with $`(Y,d_Y)`$ ( like the tangent space at a zero point of $`d_Y`$, the space of formal functions on $`Y`$, etc.). We remark also that if $`g`$ is a DGLA then the same definition can be given. It can be also generalized to the case when the base of a equivariant dg-bundle is an arbitrary dg-manifold with a marked $`d`$-stable point. Suppose that $`F`$ is a polynomial functor in the category of $`𝐙`$-graded vector spaces, $`P=Free(F),VVect_𝐙`$. We apply the general scheme outlined above to the case $`Y=(P,V),g=g_P`$. Obviously $`g`$ acts on the dg-manifold $`\underset{¯}{Hom}(F(V),V)`$, equipped with the trivial odd vector field. Let us consider the graded vector space $$𝒩=\underset{¯}{Hom}(V,V)[1]\underset{¯}{Hom}(F(V),V)g_P[1]$$ Let $`d_V\underset{¯}{Hom}(V,V)[1]`$ makes $`V`$ into a complex, $`\gamma =d_Pg_P[1]`$ satisfies the equation $`[d_P,d_P]=0`$ and $`\rho \underset{¯}{Hom}(F(V),V)`$ makes $`V`$ into a dg-algebra over $`(P,d_P)`$. We consider the formal neighborhood of the point $`(d_V,\rho ,d_P)`$ in $`𝒩`$, and define an odd vector field by the formula $$d_𝒩(d_V,\rho ,d_P)=(d_V^2,\xi (d_V,\rho )+\widehat{d}_P,[d_P,d_P]/2)$$ The notation here is compatible with the one for $``$. One can check that $`[d_𝒩,d_𝒩]=0`$. Thus the formal neighborhood becomes a formal dg-manifold. It controls deformations of pairs (an operad, an algebra over this operad). The natural projection $`\pi :𝒩g_P[1]`$ is a morphism of formal dg-manifolds . Here on $`g_P[1]`$ we use the odd vector field $`d_{g_P[1]}`$ defined by the Lie bracket. Then the formal scheme of zeros of $`d_{g_P[1]}`$ corresponds to the structures of a dg-operad on $`P`$. The fiber over a fixed point $`xg_P[1]`$ is a dg-manifold with the differential induced from $`𝒩`$. Then the formal neighborhood of a fixed point in $`\pi ^1(x)`$ controls deformations of $`\widehat{P}`$-algebras. We conclude that the Lie algebra of derivations of an operad acts homotopically on the moduli space of algebras over this operad. ## 3 Hochschild complex and operads This section serves as a sort of a “second introduction”, outlining objectives and the strategy of the rest of the paper. One of our aims will be to construct a dg-operad of the type $`\widehat{P}`$ (i.e. free as a graded operad) acting naturally on the Hochschild complex of an arbitrary $`A_{\mathrm{}}`$\- algebra. In Section 5 we are going to construct an operad $`M`$ which acts naturally on the full Hochschild complex $`C^{}(A,A)`$ of an $`A_{\mathrm{}}`$-algebra $`A`$ as well as on $`C_+^{}(A,A)`$. There is a natural free resolution $`P`$ of the operad $`M`$, so that $`C:=C^{}(A,A)`$ becomes a $`\widehat{P}`$-algebra. Then we can say that there is a dg-map of the moduli space of $`A_{\mathrm{}}`$-categories to the moduli space $`(P,C)`$ of structures of $`\widehat{P}`$-algebras on the graded vector space $`C`$. ¿From the point of view of deformation theory it is not very natural to make constructions of the type algebraic structure $``$ another algebraic structure (like our construction $`A_{\mathrm{}}algebrasMalgebras`$). It is more natural to extend them to morphisms between the formal pointed dg-manifolds controlling the deformation theories of structures. We will construct an explicit dg-map $`(𝒜_{\mathrm{}},A)(P,C)`$ as well as a dg-map $`_{cat}(𝒜_{\mathrm{}},A)(P,C)`$, such that the one is obtained from another by the restriction from the moduli space of algebras to the moduli space of categories. The operad $`𝒜_{\mathrm{}}`$ is augmented, i.e. equipped with a morphism of dg-operads $`\eta :𝒜_{\mathrm{}}Free(0)`$. Here $`Free(0)`$ is the trivial operad : $`Free(0)_1=k,Free(0)_{(n1)}=0`$. Since $`A`$ (as any graded vector space) is an algebra over $`Free(0)`$, it becomes also an algebra over $`𝒜_{\mathrm{}}`$. Any structure of an $`𝒜_{\mathrm{}}`$-algebra on $`A`$ can be considered as a deformation of this trivial structure. Notice also that in the previous notation the augmentation morphism defines a point in the dg-manifold $`(𝒜_{\mathrm{}},C)`$, where $`C=C^{}(A,A)`$. Therefore it is sufficient to work in the formal neighborhood of this point. Notice that we can consider also the moduli space of structures of a complex on the graded vector space $`C^{}(A,A)`$ where $`A`$ is an arbitrary graded vector space. It gives rise to a formal dg-manifold. There are natural morphisms to it from the formal dg-manifold of the moduli space of $`A_{\mathrm{}}`$-categories and from the formal dg-manifold of the moduli space of structures of $`\widehat{P}`$-algebras on $`C^{}(A,A)`$. Theorem $`1`$ below combines all three morphisms discussed above into a commutative diagram. Let us make it more precise. First we formulate a simple general lemma, which will be applied in the case $`V=C[2]`$. Let $`V`$ be an arbitrary graded vector space, $`d_V`$ an odd vector field on $`V`$ (considered as a graded manifold) such that $`[d_V,d_V]=0`$. Thus we get a dg-manifold. The graded vector space $`H=H(V)=\underset{¯}{Hom}(V,V)[1]`$ is a dg-manifold with $`d_H(\gamma )=\gamma ^2`$. To every point $`vV`$ we assign a point in $`H`$ by taking the first Taylor coefficient $`d_V^{(1)}(v)`$ of $`d_V`$ at $`v`$. In this way we obtain a map $`\nu :VH`$. ###### Lemma 2 The map $`\nu `$ is a morphism of dg-manifolds. $`Proof`$. Let us write in local coordinates $`x=(x_1,\mathrm{},x_n)`$ the vector field $`d_V=_i\varphi _i_i`$ where $`_i`$ denotes the partial derivative with respect to $`x_i`$, and $`\varphi _i`$ are functions on $`V`$. Then the map $`\nu `$ assigns to a point $`x`$ the matrix $`M=(M_{ij}(x))`$ with $`M_{ij}=_j\varphi _i`$. Then direct computation shows that the condition $`[d_V,d_V]=0`$ implies that the vector field $`\dot{x}=d_V(x)`$ is mapped to the vector field $`\dot{M}=d_H(M)=M^2`$. $`\mathrm{}`$ Let $`A`$ be a graded vector space endowed with the trivial $`A_{\mathrm{}}`$-structure, and $`C=C^{}(A,A)=_{n0}Hom_{Vect_𝐙}(A^n,A)`$ be the graded space of Hochschild cochains. Since $`C[1]`$ carries a structure of a graded Lie algebra (with the Gerstenhaber bracket), it gives rise to the structure of a dg-manifold on $`C[2]`$, which is the same as $`_{cat}(𝒜_{\mathrm{}},A)`$. We will denote it by $`(X,d_X)`$ (or simply by $`X`$ for short). Even for the trivial $`A_{\mathrm{}}`$-algebra structure on $`A`$, we get a non-trivial $`P`$-algebra structure on $`C`$. The corresponding moduli space $`(P,C)`$ will be denoted by $`(Y,d_Y)`$ (or $`Y`$ for short). There is a natural morphism of dg-manifolds $`p:Y\underset{¯}{Hom}(C,C)[1]=H`$ (projection of $`Y=(P,C)`$ to the first summand). The following theorem will be proved later in the paper. ###### Theorem 1 There exists a $`GL(A)`$-equivariant morphism of dg-manifolds $`f:XY`$ such that $`pf=\nu `$. Moreover we will present an explicit construction of the morphism. Suppose that $`A`$ is an $`A_{\mathrm{}}`$-algebra. Geometrically the structure of an $`A_{\mathrm{}}`$-algebra on the graded vector space $`A`$ gives rise to a point $`\gamma X=C[2],C=C^{}(A,A)`$ such that $`d_X(\gamma )=0`$. Indeed, the definition can be written as $`[\gamma ,\gamma ]=0`$. Thus we get a differential in $`C`$ (commutator with $`\gamma `$) making it into a complex. The structure of a complex on the graded vector space $`C`$ gives rise to a zero of the field $`d_H`$ in the dg-manifold $`H=\underset{¯}{Hom}(C,C)[1]`$. Theorem $`1`$ implies that $`f(\gamma )`$ is a zero of the vector field $`d_{(P,C)}`$. Therefore the Hochschild complex $`(C,[\gamma ,])`$ carries a structure of a dg-algebra over $`P`$. Our next aim is to uncover the geometric origin of the operad $`P`$. It will be related to the configuration space of discs inside of the unit discs in the plane. More precisely, the operad $`Chains(E_2)`$ of chains on the little discs operad (see \[Ko3\] and Section 7 below) is quasi-isomorphic to $`\widehat{P}`$. Here we use either usual singular chains or piecewise algebraic chains (see Appendix). In fact we are going to construct explicitly a morphism $`\widehat{P}Chains(E_2)`$ which gives the homotopy equivalence (to be more precise we will do that for the operad $`Chains(FM_2)`$ which is quasi-isomorphic to $`E_2`$). Then using the fact that both dg-operads are free as graded operads, we invert this quasi-isomorphism. This gives a structure of an $`Chains(E_2)`$-algebra on the Hochschild complex of an $`A_{\mathrm{}}`$-algebra. This result is known as Deligne’s conjecture. Let us recall that there is a notion of $`d`$-algebra, $`d𝐙_+`$ (see for ex. \[Ko3\]). Namely, a graded vector space $`V`$ is called a $`d`$-algebra if it is an algebra over the operad $`Chains(E_d)`$ (chains of the topological operad of little $`d`$-dimensional discs). Thus the moduli space $`(P,C)`$ can be thought as a moduli space of structures of a $`2`$-algebra on a graded vector space $`C`$. Then our Theorem $`1`$ says that there is a $`GL(A)`$-equivariant morphism of the moduli space of $`A_{\mathrm{}}`$-categories with one object to the moduli space of $`2`$-algebras. ###### Remark 4 In the unpublished paper \[GJ\] the name $`d`$-algebras was reserved for algebras over the operad $`H_{}(Chains(E_d))`$. It was proved by Tamarkin in \[T\] and by the first author in \[Ko3\] that there exists a (non-canonical) quasi-isomorphism between the operad $`Chains(E_d)`$ and its homology operad $`H_{}(Chains(E_d))`$ (in other words the operad $`Chains(E_d)`$ is formal). Let $`g_P=\underset{¯}{Der}P`$ means as before the DGLA of derivations of $`\widehat{P}`$. Then $`g_P`$ acts on the moduli space of $`\widehat{P}`$-algebras. On the other hand, from the point of view of deformation theory, we can replace the operad of little discs $`E_2`$ by the operad of configurations of points in the plane (properly compactified). This is the operad $`FM_2`$ mentioned above. There is a natural action of the Grothendieck-Tiechmüller groups on the rational homotopy type of the latter. It gives rise to a morphism of $`L_{\mathrm{}}`$-algebras $`Lie(GT)[1](g_P,[d_P,])`$ where $`GT`$ is the Grothendieck-Teichmüller group. Therefore one has a homotopical action of the Lie algebra $`Lie(GT)`$ on the moduli space of $`\widehat{P}`$-algebras. ## 4 Free resolution of a dg-operad In this subsection we recall well-known constructions of free resolutions of operads (see e.g. \[GJ\]). Let $`k`$ be a field as before, $`R`$ be a dg-operad over $`k`$. The aim of this subsection is to construct canonically a dg-operad $`P_R`$ over $`k`$, which is free as a graded operad, and a quasi-isomorphism $`P_RR`$. Then $`P_R`$ will be a free resolution of $`R`$. In this subsection we will assume that $`R`$ is non-trivial, which means that the unit operation from $`R_1`$ is not equal to zero. ### 4.1 Topological construction We will mainly follow \[BV\]. Let $`O=(O_n)_{n0}`$ be a topological operad (i.e. all $`O_n`$ are $`S_n`$-topological spaces and all operadic morphisms are continuous). We describe (following \[BV\]) a construction of a topological operad $`B(O)=(B(O)_n)_{n0}`$ together with a morphism of topological operads $`B(O)O`$ which is homotopy equivalence. To simplify the exposition we assume that $`S_n`$ acts freely on $`O_n`$ for all $`n`$. Each space $`B(O)_n`$ will be the quotient of $$\overline{B(O)}_n=\underset{[T],TTree(n)}{}([0,+\mathrm{}]^{E_i(T)}\times \underset{vV_i(T)}{}O_{N^1(v)})/AutT$$ under the relations described in the following way. Let us consider the elements of $`\overline{B(O)}_n`$ as numbered trees with elements of $`O`$ attached to the internal vertices, and the $`length`$ $`l(e)[0,+\mathrm{}]`$ attached to every edge $`e`$. We require that all external edges (i.e. root edge and the tail edges) have lengths $`+\mathrm{}`$. We impose two type of relations. 1) We can delete every vertex $`v`$ of valency 1 if it contains the unit of the operad, replacing it and the attached two edges of lengths $`l_i,i=1,2`$ by the edge with the length $`l_1+l_2`$ . We use here the usual assumption: $`l+\mathrm{}=\mathrm{}+l=\mathrm{}`$. 2) We can contract every internal edge $`e=(v_1,v_2),v_2=N(v_1)`$ of the length $`0`$ and compose in $`O`$ the operations attached to $`v_i,i=1,2`$. We depict the trees and relations below. Let us describe how $`B(O)`$ can act naturally on a topological space. Let $`X`$ be a topological space, and $`g^t:XX,t[0,+\mathrm{})`$ a $`1`$-parametric semigroup of continuous maps acting on $`X`$. We assume that the map $`[0,+\mathrm{})\times XX,(t,x)g^tx`$ extends continuosly to $`[0,+\mathrm{}]\times X`$. We denote by $`Y`$ the image of $`g^{\mathrm{}}:=lim_t\mathrm{}g^t`$. Clearly the subspace $`Y`$ is a homotopy retract of $`X`$. Suppose that a topological operad $`O`$ acts on $`X`$, i.e. that we are given continuous maps $`O_n\times X^nX,n0`$, satisfying the usual properties. We can construct an action of $`B(O)`$ on $`Y`$ as follows. Let $`\gamma O_n,t,t_i𝐑_+\{+\mathrm{}\},x_iX,1in`$. Then we assign to these data the point $`x=g^t\gamma (g^{t_1}x_1,\mathrm{},g^{t_n}x_n)`$ of $`X`$. We define the composition of such operations in the natural way. We can interpret the parameters $`t,t_i`$ above as lengths of edges of trees . Putting $`t=+\mathrm{}`$ we obtain an action of $`B(O)`$ on the homotopy retract $`Y`$. ### 4.2 Free resolutions of linear operads Let $`R`$ be a dg-operad over a field $`k`$. To describe its free resolution $`P_R`$ we need a special class of trees described below. For every $`n0`$ we introduce a groupoid $`𝒯(n)`$ of $`marked`$ $`trees`$ with $`n`$ tails. An object of $`𝒯(n)`$ is a numbered tree $`TTree(n)`$ and a map to a $`3`$-element set $`l=l_T:E(T)\{0,finite,+\mathrm{}\}`$ such that $`l_T(\{e_r\}E_t(T))=\{+\mathrm{}\}`$. Notice that in the case of topological operads the component $`\overline{B(O)_n}`$ is stratified naturally with the strata labeled by equivalence classes $`|𝒯(n)|`$ of marked trees. The label of an edge $`e`$ of the corresponding marked tree is $`0`$ if $`l(e)=0`$, is $`finite`$ if $`l(e)(0,+\mathrm{})`$ and is $`+\mathrm{}`$ if $`l(e)=+\mathrm{}`$. According to this description, we call them zero, finite, infinite edges respectively. We denote these sets of edges by $`E_{zero},E_{finite}`$ and $`E_{infinite}`$ correspondingly. We will give three different but equivalent descriptions of the operad $`P=P_R`$ as a graded operad. Then we define a differential on $`P`$. Description 1. Let $$\overline{P}_n=\underset{[T],T𝒯(n)}{}(_{vV(T)}R_{N^1(v)}[J_T])_{AutT}$$ where $`AutT`$ is the group of automorphisms of the tree $`T`$, $`J_T=l_T^1(finite)`$, and for any graded vector space $`W`$ and a finite set $`J`$ we use the notation $`W[J]=Wk[1]^J`$ (shift of the grading by $`J`$). Notice that the dimension of the corresponding stratum of $`𝒯(n)`$ is the cardinality of the set $`J_T`$, i.e. the number of finite edges. Then $`(\overline{P}_n)_{n0}`$ evidently form a graded operad $`\overline{P}`$. It is a $`k`$-linear analog of the operad $`\overline{BO}`$. The operad $`\overline{P}`$ contains a subspace $`I_P`$ generated by the following relations 1) if the length of an edge $`(w,v)`$ is $`0`$ we contract it and make the composition in $`R`$ of the operations attached to $`w`$ and $`v`$ (cf. description for $`B(O)`$); 2) if $`v`$ is a vertex of $`T\overline{P}`$ such that $`|v|=1`$ and $`1_RR_1`$ is attached to $`v`$, then $`T`$ belongs to $`I_P`$ if the following holds: lengths of both edges attached to $`v`$ are non-zero edges, and at least one of them is finite. If both edges are infinite, we remove the vertex and two attached edges, replacing them by an infinite edge. One can check easily that $`I_P`$ is a graded ideal in $`\overline{P}`$. We denote by $`P`$ the quotient operad $`\overline{P}/I_P`$. Description 2. We define $`P_n`$ in the same way, but making the summation over all trees without edges of zero length. We also drop the relation 1) from the list of imposed relations (there are no edges with $`l=0`$ ). Description 3. We define an operad $`R^{}`$ such as follows: $`R_n^{}=R_n`$ for $`n1`$, $`R_1^{}`$ is a complement to the subspace $`k1_R`$ in $`R_1`$. Then we define $`P_n`$ as in Description 2, but using $`R^{}`$ instead of $`R`$ and dropping both relations 1) and 2). It is clear that this description defines a free graded operad. Equivalently, it can be described as a free graded operad $`P`$ such that $$P=Free(Cofree^{}(R^{}[1]))[1]$$ Here $`Cofree(L)`$ means a dg-cooperad generated by $`L`$ which is cofree as a graded co-operad, and $``$ denotes the procedure of taking a (non-canonical) complement to the subspace generated by the unit (or counit in the case of a co-operad) as described above in the case of $`R`$. In this description the generators of $`P`$ correspond to such trees $`T`$ in $`𝒯=(𝒯(n))_{n0}`$ that every $`T`$ has at least one internal vertex, all internal edges are finite and there are no zero-edges in $`T`$. ###### Proposition 2 All three descriptions give rise to isomorphic graded free operads over $`k`$. $`Proof`$. Straitforward. $`\mathrm{}`$ We can make $`\overline{P}`$ into a dg-operad introducing a differential $`d_{\overline{P}}`$. We use the Description 1 for this purpose. The differential $`d_{\overline{P}}`$ is naturally decomposed into the sum of two differentials: $`d_{\overline{P}}=\stackrel{~}{d}_R+d_𝒯`$ where a) the differential $`\stackrel{~}{d}_R`$ arises from the differential $`d_R`$ in $`R`$; b) the differential $`d_𝒯`$ arising from the stratification of $`𝒯(n)`$: it either contracts a finite edge or makes it into an infinite edge . To be more precise, let us consider the following object $`\mathrm{\Delta }`$ in $`𝒞=Vect_𝐙`$: $`\mathrm{\Delta }^1=1_𝒞,\mathrm{\Delta }^0=1_𝒞1_𝒞`$ where $`1_𝒞`$ is the unit object in the monoidal category $`𝒞`$. Then $`\mathrm{\Delta }`$ can be made into a chain complex of the CW complex $`[0,+\mathrm{}]=\{0\}(0,+\mathrm{})\{+\mathrm{}\}`$. We see that as a graded space $`\overline{P}_n`$ is given by the formula $$\overline{P}_n=_{[T],TTree(n)}(_{vV_i(T)}R_{N^1(v)}\mathrm{\Delta }^{E_i(T)})_{AutT}$$ Since we have here a tensor product of complexes, we get the corresponding differential $`d_{\overline{P}}`$ in $`\overline{P}`$. ###### Proposition 3 The ideal $`I_P`$ is preserved by $`d_{\overline{P}}`$. $`Proof`$. Straitforward computation. $`\mathrm{}`$ Therefore $`P=P_R`$ is a dg-operad which is free as a graded operad. There is a natural morphism of dg-operads $`\varphi :PR`$. In terms of the Description 2 it can be defined as follows: $`\varphi `$ sends to zero all generators of $`P`$ corresponding to trees with at least one finite edge. Let $`TP`$ be a tree with all infinite edges. Then $`T`$ gives rise to a natural rule of composing in $`R`$ elements of $`R_{N^1(v)}`$ assigned to the vertices of $`T`$. We define $`\varphi (T)R`$ as the result of this composition. It is easy to check that $`\varphi `$ is a well-defined morphism of dg-operad. ###### Proposition 4 The morphism $`\varphi `$ is a quasi-isomorphism of dg-operads. $`Proof`$. Follows from the spectral sequence arising from the natural stratification of $`𝒯`$. To say it differently, let us consider the tautological embedding $`\psi `$ of $`R`$ into $`P`$. Then $`\psi `$ is a right inverse to $`\varphi `$. It gives a splitting of $`P`$ into the sum $`P=\psi (R)P^{(0)}`$. Here $`P^{(0)}`$ is spanned by the operations corresponding to trees with finite edges only. Such a tree can be contracted to a point which means that $`P^{(0)}`$ is contractible as a complex. Hence $`\varphi `$ defines a quasi-isomorphism of complexes and dg-operads. $`\mathrm{}`$ ### 4.3 Example Let us discuss an example when $`R`$ is the operad of associative algebras without the unit. We denote it by $`As`$. Then for any $`n1`$ we have: $`As_n`$ is isomorphic to the regular representation of the symmetric group $`S_n`$. In this case the complex $`P_n`$ from the previous subsection can be identified with the chain complex of the CW-complex $`K_n,n2`$ described below. The cells of $`K_n`$ are parametrized by $`planar`$ $`trees`$ with an additional structure on edges. By a planar tree here we understand a numbered tree $`T`$ such that for any $`vV_i(T)`$ the cardinality of $`N^1(v)`$ is at least $`2`$ and this set is completely ordered. The additional structure is a map $`E_i(T)\{finite,infinite\}`$. We call an edge finite or infinite according to its image under this map. Dimension of the cell is equal to the number of finite edges of the corresponding planar tree. We can either contract a finite edge or make it infinite. This defines an incidence relation on the set of cells. We can picture planar trees as follows Here the dashed lines show the complete orders on set of incoming edges. We will not show them on other figures in the text. Instead, we will tacitly assume that for a given vertex the incoming edges are completely ordered from the left to the right. In this way we obtain cubical subdivisons of the Stasheff polyhedra. We depict the case $`n=4`$ below ## 5 Minimal operad In this subsection we describe a dg-operad $`M=(M_n)_{n1}`$ which acts naturally on the Hochschild complex of an $`A_{\mathrm{}}`$-algebra. We call it $`minimal`$ operad. Let us describe this operad informally. We treat elements $`\gamma C^{}(A,A)`$ as polylinear operations on $`A`$. For given operations $`\gamma _1,\mathrm{},\gamma _nC^{}(A,A)`$, and an $`A_{\mathrm{}}`$-structure $`mC^{}(A,A)`$ we can make compostions in all possible ways, reading the arguments from left to right. For example we can make an expression like this: $`\gamma (a_1\mathrm{}a_5)=\gamma _2(a_1a_2m_2(a_3\gamma _1(a_4))\gamma _3(a_5))`$, etc. Such compostions can be depicted by planar trees. The operad $`M`$ is spanned by operations corresponding to certain trees, which we call admissible (see below). It seems that the operad $`M`$ is close to what is described in \[MS\] as “natural transformations $`(C^{}(A,A))^nC^{}(A,A)`$” (this terminology is confusing because the assignement $`AC^{}(A,A)`$ is not a functor). In all previous works dealing with Deligne’s conjecture the authors used operads which act on $`C^{}(A,A)`$ and generated by the operations called braces. It seems that the braces generate the operad $`M`$, but it is not clear what is the complete list of relations. The advantage of our operad $`M`$ is that it is described directly, not as a quotient of a free operad. The defining properties of a dg-operad (i.e. associativity of the composition, Leibniz rule for the differential) will become clear later. We will give two definitions of the operad $`M`$. The first one is suitable for the pure algebraic descriptions of the operadic composition and the differential. But the signs in the formulas are not very transparent. Second description takes care about “parity of edges”, so the correct signs come out automatically. In some formulas related to the first description we will write $`\pm `$ having in mind that the correct sign follows from the second description. ### 5.1 Basis of $`M`$ ###### Definition 13 For a finite set $`I`$ we define an $`I`$-labeled planar tree as a triple $`(T,lab,ord)`$ where $`T`$ is a tree in the sense of Section 1.2, $`lab:IV_i(T)`$ is an embedding, and $`ord`$ is a complete order on the sets $`N^1(v),vV_i(T)`$. We call labeled a vertex from the image of the map $`lab`$. All other internal vertices are called non-labeled. ###### Definition 14 We will call an $`I`$-labeled tree admissible if it has no tail vertices, and for every non-labeled internal vertex $`v`$ we have $`|v|2`$. We denote the set of isomorphism classes of $`I`$-labeled planar trees by $`Tree^{(p)}(I)`$, where the upper script $`p`$ stays for “planar”. Notice that the automorphism group of an $`I`$-labeled planar tree is trivial. For $`I=\{1,\mathrm{},n\}`$ we will use the notation $`Tree^{(p)}(n)`$. We are going to use admissible trees unless we say otherwise. If it will not lead to a confusion, we will simply call them trees. Slightly abusing the notation we will denote an $`I`$-labeled tree by $`T`$, skipping lab and ord. Notice that terminology here is different from the one in Section 1.5. In particular, we label here internal vertices, not tails. Since we do not consider here numbered trees (in the terminology of Section 1.5), this change of terminology should not lead to a confusion. We can depict trees from $`Tree^{(p)}=(Tree^{(p)}(n))_{n1}`$ as follows Labeled vertices are depicted as circles with numbers inscribed, non-labeled vertices are depicted as black vertices. We define $`M_I`$ to be a $`k`$-vector space spanned by all elements of $`Tree^{(p)}(I)`$. For $`I=\{1,\mathrm{},n\}`$ we will use the notation $`M_n`$. The symmetric group $`S_n`$ acts on $`M_n`$ permuting labeled vertices. Abusing the notation further, we will denote the element of $`M_n`$ corresponding to a tree $`T`$ simply by $`T`$. Thus we have: $`M_0=0`$ and $`M_1`$ is a 1-dimensional vector space generated by $`T_e`$, the first tree on the figure above. In fact $`T_e`$ corresponds to the unit $`1_MM_1`$ of the operad $`M`$. The operadic composition in $`M`$ and the differential will be described below. The degree of the basis element corresponding to a tree $`T`$ is equal to $$deg(T)=\underset{vV_{lab}(T)}{}(|v|)+\underset{vV_{nonl}(T)}{}(2|v|)$$ where $`V_{lab}(T)`$ and $`V_{nonl}(T)`$ denote the sets of labeled and non-labeled vertices respectively , and $`|v|`$ is the cardinality of the set $`N^1(v)`$. ### 5.2 Composition in $`M`$ We need to define an element of $`M`$ which corresponds to a tree $`T_2`$ glued to a tree $`T_1`$ at a labeled vertex $`vV_{lab}(T_1)`$. The trees $`T_i,i=1,2`$ correspond to some elements of $`M`$. The resulting element will be by definition their operadic composition. It is given by the sum $$T_1_vT_2=\underset{\beta }{}\pm (T_1_vT_2)_\beta $$ where the trees $`(T_1_vT_2)_\beta `$ are defined below. First, with the tree $`T_2`$ we associate a set $`A(T_2)=_{vV_i(T_2)}\{0,\mathrm{},|v|\}`$. We call it the set of $`angles`$ of $`T_2`$. Obviously there is a natural map $`\kappa :A(T_2)V_i(T_2)`$. The path in $`𝐑^2`$ which goes from the left to the right and surrounds $`T_2`$ defines a complete order on $`A(T_2)`$. On the following figure angles are marked by asteriscs. The datum $`\beta `$ above is a (non-strictly) monotonic map $`\beta :N^1(v)A(T_2)`$. We will think of this map as about the way to insert a vertex $`wN^1(v)`$ and the edge $`(w,v)`$ inside of an angle formed by two edges incoming to $`\beta (v)`$. Schematically it is shown on the figure below. Let $`T_\beta =(T_1_vT_2)_\beta `$. Then we define the set of vertices $`V(T_\beta )`$ as $`(V(T_1)\{v\})(V(T_2)\{root\})`$. The map $`N=N_{T_\beta }`$ is defined such as follows: for all $`wV(T_1)\{v\}`$ such that $`N_{T_1}(w)v`$ we put $`N(w)=N_{T_1}(w)`$ in the self-explained notation. Similarly if $`wV(T_2),N_{T_2}(w)=root`$ we put $`N(w)=N_{T_1}(v)`$. If $`wV(T_2),N_{T_2}(w)root_{T_2}`$ then we put $`N(w)=N_{T_2}(w)`$. Let us suppose that $`wN_{T_1}^1(v)`$. Then we define $`N(w)`$ as $`\kappa (\beta (w))`$. The root vertex of $`T_\beta `$ is the same as for $`T_1`$. The labeling and complete orders on the sets $`N^1(x)`$ are defined in the natural way. Informally speaking, $`T_\beta `$ is obtained by removing from $`T_1`$ the vertex $`v`$ together with all incoming edges and vertices, and gluing $`T_2`$ to $`v`$. Then we use the map $`\beta `$ in order to “insert” removed vertices. With such a composition we obtain the structure of a graded operad on $`M`$. We depict an example of the composition in $`M`$ below ### 5.3 Differential in $`M`$ For a generator $`T`$ we define $`d_M(T)=_{vV_i(T)}d_v(T)`$ where each $`d_v(T)`$ will be of the form $$d_v(T)=\underset{i,j}{}\pm d_{v,i,j}(T).$$ We need to explain the set of indices of summation and each summand. Let us recall that for every vertex $`vV_i(T),|v|=k`$ there is a bijection of sets $`\{1,\mathrm{},k\}N^1(v)`$ which defines a complete order on the set $`N^1(v)`$. We will identify an element of $`N^1(v)`$ with the corresponding number. The indices $`i,j`$ in the sum above will be half-integers: $`1/2ijk+1/2`$. For a pair $`i,j`$, we define $`d_{v,i,j}(T)`$ to be a tree $`T^{}`$ such that: a) $`V(T^{})=(V(T)\{v\})\{v^{up},v^{down}\}`$ where $`\{v^{up},v^{down}\}`$ are new vertices; b) $`root_T^{}=root_T`$; We put $$N_T^{}(v^{down})=v^{up},$$ $$N_T^{}(l)=v^{down},$$ if $`i<l<j`$; $$N_T^{}(l)=v^{up},$$ if $`l<i`$ or $`l>j`$. For all other vertices $`u`$ we put $`N_T^{}(u)=N_T(u)`$. Complete orders on the sets $`N_T^{}^1(u)`$ are defined in the natural way. The tree $`T^{}`$ has one new edge $`(v^{down},v^{up})`$ which we denote in pictures as “new”. We can depict these definitions as follows The range of summation is defined differently in the case of labeled and non-labeled vertices $`v`$. The idea is to keep admissible graphs only. 1) If $`v`$ is non-labeled then we take sum over $`1/2ijk+1/2`$ such that $`2ji`$ and $`1k(ji)`$, in $`d_{v,i,j}(T)`$ both vertices $`v^{up}`$ and $`v^{down}`$ are non-labeled. 2) If $`v`$ is labeled, then $`d_v(T)=d_v^{(1)}(T)+d_v^{(2)}(T)`$ where the first summand $`d_v^{(1)}(T)`$ is the sum of $`d_{v,i,j}(T)`$ with $`1k(ji)`$ and with the vertex $`v^{down}`$ appearing with the old label of $`v`$, and $`v^{up}`$ being non-labeled. The second summand $`d_v^{(2)}(T)`$ is the sum of $`d_{v,i,j}(T)`$ with $`2ji`$, the vertex $`v^{down}`$ is non-labeled, and $`v^{up}`$ has the same label as $`v`$ in $`T`$. Labeling of all vertices different from $`v^{up}`$ and $`v^{down}`$ remains the same. ### 5.4 Action of $`M`$ on the Hochschild complex Let $`(A,m)`$ be an $`A_{\mathrm{}}`$-algebra. Our next step is to define an action of a dg-operad $`M`$ on $`C^{}(A,A)`$ so the latter becomes a dg-algebra over $`M`$. Let $`TM_n,\gamma _iC^{}(A,A),1in`$. We need to define an element $`T(\gamma _1,\mathrm{},\gamma _n)C^{}(A,A)`$. It can be expressed as a sequence of morphisms of graded vector spaces $`T(\gamma _1,\mathrm{},\gamma _n)_N:A^kA`$, $`k=0,1,\mathrm{}`$ Let $`T_{(k)}`$ be a unique (non-admissible) planar tree with only one internal vertex $`v`$, which is labeled, and $`k`$ tails, numbered from the left to the right. Then we have the composition $$T_{(k)}_vT=\underset{\beta }{}\pm (T_{(k)}_vT)_\beta $$ defined in the same way as for admissible trees. Here as before $`\beta :\{1,\mathrm{},k\}A(T_{(k)}_vT)`$ is a monotonic map. Every tree $`(T_{(k)}_vT)_\beta `$ has $`k`$ tails. For every $`vV_i((T_{(k)}_vT)_\beta )`$ we define a polylinear map $`\gamma _v:A^{|v|}A`$ in the following way: 1) if $`v`$ is labeled by $`j`$, $`1jn`$, we define $`\gamma _v`$ as the component of $`\gamma _j`$ which belongs to $`Hom_{Vect_𝐙}(A^{|v|},A)`$; 2) if $`v`$ is non-labeled we define $`\gamma _v=m_{|v|}`$. The tree $`(T_{(k)}_vT)_\beta `$ defines the way to compose operations $`\gamma _v`$ into an operation $`\gamma _\beta :A^kA`$. We define $`T(\gamma _1,\mathrm{},\gamma _n)`$ to be equal to the sum $`_\beta \pm \gamma _\beta `$. We claim that in this way we get on $`C^{}(A,A)`$ a structure of a dg-algebra over $`M`$. This can be checked by a straightforward computation. It is more or less clear from the definitions that $`C^{}(A,A)`$ is an algebra over the graded operad $`M`$. Hence the question is about the compatibility with differentials. The latter follows from a more general result (the Theorem $`1`$), and will be proved in Section 6. ### 5.5 Signs in the minimal operad Now we would like to discuss the second description of the operad $`M`$. It is based on the following result from the theory of Strebel’s differentials (see \[St\]). ###### Theorem 2 Let $`I=\{z_1,\mathrm{},z_n\}`$ be a finite non-empty subset of the complex line $`𝐂`$. Then there exists a unique quadratic differential $`\alpha =f(z)(dz)^2`$ with $`f(z)=_i\frac{a_i}{zz_i}`$, $`a_i𝐂,i=1,\mathrm{},n`$ , $`_ia_i=1`$, satisfying the following property: There exists a unique tree $`TTree^{(p)}(I)`$ and an embedding $`j:T\mathrm{𝐂𝐏}^1`$, $`\{root\}\{+i\mathrm{}\}`$ such that the pair $`(\mathrm{𝐂𝐏}^1j(T),\alpha )`$ is equivalent (as a complex curve with quadratic differential) to the lower half-plane $`Im(z)<0`$ equipped with the quadratic differential $`i(dz)^2`$. We will call such $`\alpha =f(z)(dz)^2`$ the Strebel differential associated with the finite set $`I𝐂`$. To every Strebel differential $`\alpha `$ we can assign canonically the $`\{1,\mathrm{},n\}`$-labeled planar tree $`T`$. Conversely, having a tree $`TTree^{(p)}(n)`$ we can look for sequences of pairwise distinct points $`(z_1,\mathrm{},z_n)`$ in $`𝐂`$ which can appear as the sets of possible poles of the Strebel differential from the Theorem. We remark that the set of all possible $`(z_1,\mathrm{},z_n)𝐂^n\{diag\}`$ form an open cell $`Str_T`$. This follows from the fact that this space is a vector bundle of rank $`2`$ over a cell. The latter cell is defined by the lengths of internal edges (lengths are taken with respect to the metric $`|\alpha |`$ which is well-defined on $`𝐂I`$). Then the embedding of $`T`$ is fixed up to parallel traslations. The latter span the fiber of a vector bundle. The total space of this bundle is the cell $`Str_T𝐂^n\{diag\}`$. It is easy to see that $`codim(Str_T)=deg(T)=(|E_i(T)|+22|V_{lab}(T)|)`$. One can show that there is a finite CW-complex $`\mathrm{\Sigma }_{(n)}𝐂^n\{diag\}`$ with the cells $`\mathrm{\Sigma }_T`$ labeled by $`TTree^{(p)}(n)`$, and such that $`\mathrm{\Sigma }_T`$ intersects $`Str_T`$ transversally and at exactly one point. Now we can give the second definition of the operad $`M`$: For $`n2`$ the complex $`M_n`$ is defined as the chain complex of $`\mathrm{\Sigma }_{(n)}`$. Using our agreement about degrees of homological complexes, one can check again that every complex $`M_n`$ is a finite-dimensional complex concentrated in degrees $`\{(n1),\mathrm{},0\}`$. We use the second definition of in order to derive the following combinatorial description of the operad $`M`$. One associates to a tree $`TTree^{(p)}(n)`$ a one-dimensional vector space (in fact an abelian group, so everything can be done over integers). It is defined by the formula $`U_T=((\stackrel{~}{H}_{}(𝐑))^{})^{E_i(T)}((\stackrel{~}{H}_{}(𝐑^2)^{})((\stackrel{~}{H}_{}(𝐑^2))^{V_{lab}(T)}`$. In the formula $`\stackrel{~}{H}_{}(X)`$ denotes the reduced homology of the one-point compactification of $`X`$ (Borel-Moore homology). The graded vector space $`M_n`$ coincides with the sum of $`U_T`$ over all trees belonging to $`Tree^{(p)}(n)`$. Next step is to interpret various shifts in complexes as tensor products with the reduced homology of vector spaces. For example the shift by $`1`$ is the tensor product with the reduced homology of $`𝐑`$. It is more convenient to use a different notation for the same spaces of reduced homology. Let $`L_1:=\stackrel{~}{H}_{}(𝐑)`$ and $`L_2:=\stackrel{~}{H}_{}(𝐑^2)`$. These are $`1`$-dimensional graded vector spaces of pure degrees $`1`$ and $`2`$ correspondingly. For a finite set $`I`$ we define a vector space $`M_I:=_{TTree^{(p)}(I)}(L_1^{})^{E_i(T)}L_2^{}L_2^I=_{TTree^{(p)}(I)}U_T.`$ Geometrically the new notation is related to the picture with the Strebel differentials. We think about $`𝐑^2`$ as about direct sum of two lines $`𝐑^2=𝐑_{hor}𝐑_{vert}`$ (horizontal and vertical lines). The vertical line is the $`y`$-axis in $`𝐂=𝐑^2`$. The vertical line corresponds to $`L_1`$ in the notation above, and the horizontal line corresponds to $`L_1^{}L_2`$ (we think of it as about $`𝐑^2/𝐑_{vert}`$). There is a natural structure of a complex on $`M_I`$. We define a linear map $`d_I:M_IM_IL_1`$ as a sum $$d_I=\underset{T_1T}{}d_{T_1,T}.$$ Here $`T_1T`$ means that $`T_1`$ is obtained from $`T`$ by adding an internal edge. The summand $`d_{T_1,T}`$ is a linear map from $`U_T`$ to $`U_{T_1}L_1=(U_TL_1^{})L_1`$. It is given by $`id_{U_T}\epsilon _{L_1}`$ where $`\epsilon _{L_1}:\mathrm{𝟏}L_1^{}L_1`$ is the canonical morphism in the symmetric monoidal category $`Vect_𝐙`$. ###### Lemma 3 The linear map $`d_I`$ defines a differential in $`M_I`$. Proof. We need to prove that $`(d_Iid_{L_1})d_I=0`$. This can be checked directly using the fact that $`L_1`$ has degree $`1`$ , so the commutativity constraint acts on $`L_1L_1`$ as $`\sigma `$ where $`\sigma `$ is the permutation map. $`\mathrm{}`$ The compostion maps for the operad $`M`$ can be naturally described in the new notation. We remark that in order to glue a tree $`T_2`$ to a tree $`T_1`$ at a vertex $`v`$ first we need to define a sequence of trees $`\{(T_1_vT_2)_\beta \}`$ (they were also denoted by $`T_\beta `$ in the previous description of the gluing). For every $`(T_1_vT_2)_\beta `$ the set of internal edges is the disjoint union of the corresponding sets for $`T_1`$ and $`T_2`$. The set of labeled vertices for $`(T_1_vT_2)_\beta `$ is $`(V_{lab}(T_1)V_{lab}(T_2))\{v\}`$. Then we need to define a morphism of vector spaces $`U_{T_1}U_{T_2}U_{(T_1_vT_2)_\beta }`$. Equivalently we need to define a morphism of the graded vector space $`(L_1^{})^{E_i(T_1)}L_2^{}L_2^{I_1}(L_1^{})^{E_i(T_2)}L_2^{}L_2^{I_2}`$ to the graded vector space $`(L_1^{})^{(E_i(T_1)E_i(T_2))}L_2^{}L_2^{((I_1\{v\})I_2)}.`$ (The notation is self-explained). We define the morphism to be the identity on the tensor factors marked by same edges or same elements of the sets $`I_k,k=1,2`$. For example $`(L_1^{})^{E_i(T_1)}(L_1^{})^{E_i(T_2)}`$ is identically mapped to $`(L_1^{})^{(E_i(T_1)E_i(T_2))}`$. Hence it is enough to define a morphism $`L_2^{}L_2^{}L_2L_2^{}`$, where first and third tensor factors in the LHS correspond to the tree $`T_1`$, middle tensor factor corresponds to the tree $`T_2`$ and the space $`L_2^{}`$ in the RHS corresponds to the tree $`(T_1_vT_2)_\beta `$. Then one uses the canonical evaluation map $`ev:L_2^{}L_2\mathrm{𝟏}`$ in the symmetric monoidal category $`Vect_𝐙`$. ### 5.6 Signs in the Hochschild complex In this subsection we are going to reformulate the definition of the Hochschild complex using the reduced homology spaces $`L_1`$ and $`L_2`$. Then one gets the sign agreement with the computations in the previous subsection. For a graded vector space $`A`$ we define $$C=C^{}(A,A)=_I\underset{¯}{Hom}(A^I,A)(L_2^{}L_1)^I$$ where the sum is taken over all non-empty completly ordered finite sets, and $`\underset{¯}{Hom}`$ is the internal $`Hom`$ in the tensor category $`Vect_𝐙`$. Let us demonstrate how the structures in $`C`$ can be reformulated by means of this language. The Gerstenhaber bracket is a map $`CCC(L_2^{}L_1)`$. Now the multiplication $`m`$ can be described as a point of a dg-manifold: $`mCL_2`$. The bracket $`[m,m]`$ defines a point of the dg-manifold $`CL_2^{}L_1L_2L_2=(CL_2)L_1`$. If $`[m,m]=0`$ then we have a differential $`d_m=[m,]:CCL_1`$. The corresponding structure of DGLA on $`C`$ can be reformulated such as follows. Consider the class $``$ of $`simple`$ $`forests`$. A simple forest $`F`$ is a finite collection of planar trees $`F=\{T_\alpha \}_{\alpha \mathrm{\Omega }}`$ with no internal edges. For every $`F`$ we define the graded vector space $$W_F:=_{\alpha \mathrm{\Omega }}Hom(A^{V_t(T_\alpha )},A)(L_2^{}L_1)^{V_t(T_\alpha )}L_2.$$ We have the natural groupoid structure on $``$. The correspondence $`FW_F`$ defines a functor from this groupoid to the category $`Vect_𝐙`$. The free cocommutative coalgebra cogenerated by $`CL_2`$ can be described as a colimit of this functor: $`Coalg(CL_2)=colim_FW_F=_{F/iso}(W_F)_{AutF}`$. The coalgebra structure on $`Coalg(CL_2)`$ can be described in these terms. Namely $`\mathrm{\Delta }pr=_{F_1F}(prpr)i(F_1,F_2)`$. Let us explain the notation. Here $`\mathrm{\Delta }`$ is the coproduct on the free cocommutative coalgebra cogenerated by $`CL_2`$, $`pr`$ is the projection to the coinvariants. The sum is taken over all subforests $`F_1`$ (unions of some connected components of $`F`$), and we fix the splitting $`i(F_1,F_2):W_FW_{F_1}W_{F_2}`$, where $`F_2`$ is the complementary forest. We leave to the reader straightforward reformulations of other structures on the Hochschild complex and checking the signs. ###### Remark 5 For an $`A_{\mathrm{}}`$-algebra $`A`$ one can define the opposite algebra $`A^{op}`$. In the geometric language of this section it corresponds to the antipodal involution on $`𝐑_{hor}`$. ## 6 Morphism of dg-manifolds ### 6.1 Generators of the free operad The main purpose of this section is to prove the Theorem 1. It will be done in subsection 6.2. This subsection is devoted to some technical preparatory material. Applying the general theory of Section 4 to the case of the operad $`M`$ we obtain its free resolution $`P`$. It is a dg-operad which is free as a graded operad. Its graded components $`P_n`$ can be described explicitly in terms of the operad $`M`$. We know that $`P_0=0`$ and $`P_1=k1_P`$. One can easily describe the space $`G_n=(Free^{}(M^{}[1])[1])_n`$ of generators of $`P_n,n2`$. Namely, $`G_n`$ is a direct sum of $`1`$-dimensional graded vector spaces $`W_𝐓`$. The sum is taken over the set of equivalence classes of collections $`𝐓=(T,\{T_v\}_{vV_i(T)})`$ where: a) $`TTree(n),n2`$ such that for every $`vV_i(T)`$ we have $`|v|2`$; b) $`T_vTree^{(p)}(N_T^1(v))`$. In order to be consistent with the notation of Section 7, we should also label all internal edges of $`𝐓`$ by an additional label finite. We omit the labeling since it will not be used in this section. Clearly the automorphism group of any such $`𝐓`$ is trivial. The symmetric group $`S_n`$ acts freely on $`G_n`$ permuting the tails of $`T`$. Any $`𝐓`$ gives rise to the $`1`$-dimensional vector space $`W_𝐓=L_1^{E_i(T)}U_{T_v}`$, where the space $`U_T^{}`$ was defined in 5.5. In what follows we will identify $`𝐓`$ with the corresponding (up to a sign) generator of $`G_n`$. Notice that if $`E_i(T)=\mathrm{}`$ then $`T`$ has the only internal vertex $`v`$ and $`W_𝐓=U_{T_v}`$. Moreover, the set $`N^1(v)`$ is naturally identified with the set $`\{1,\mathrm{},n\}`$, so that we can write $`T_vTree^{(p)}(n)`$. There is a natural morphism of $`S_n`$-modules $`pr_n:G_nM_n`$, such that $`pr_n(W_𝐓)=0`$ if $`E_i(T)=\mathrm{}`$, otherwise $`pr_n(W_𝐓)=id_{T_v}`$ where $`v`$ is the only internal vertex of $`T`$. A generator $`𝐓`$ is pictured by a tree $`T`$ with $`n`$ numbered tails, and with generators of $`M`$ inscribed into all internal vertices of $`T`$. If $`T_vM`$ is inscribed into a vertex $`vV_i(T)`$ then the cardinality of $`V_{lab}(T_v)`$ is equal to the cardinality of $`N^1(v)`$. Now we can return to the Theorem 1. We have constructed the minimal operad $`M`$. The Hochschild complex $`C^{}(A,A)`$ of an $`A_{\mathrm{}}`$-algebra $`A`$ is an algebra over $`M`$ (if we forget the differentials). Then the free resolution $`P`$ of $`M`$ acts on $`C^{}(A,A)`$, so the latter becomes an algebra over the graded operad $`P`$ (again we forget the differentials). On the other hand, let $`A`$ be a graded vector space and $`mC=C^{}(A,A)`$, but not necessarily $`[m,m]=0`$. Then the constructions of the Section 5 give rise to a sequence of elements $`\rho (m)=(\rho (m)_n)_{n1}`$ of $`\underset{¯}{Hom}(M_n_{S_n}C^n,C)`$ (no conditions on $`m`$ were imposed by the construction). The following lemma is easy to prove ###### Lemma 4 The sequence $`\rho (m)`$ defines a structure of a graded $`M`$-operad on $`C`$. Compositions $`\gamma (m)_n=\rho (m)_npr_n`$ for $`n2`$ give rise to a sequence of $`S_n`$-equivariant maps $`G_n_{S_n}CC`$. Since the graded operad $`P`$ is freely generated by $`G=(G_n)_{n2}`$, these compositions define a point in $`(P,C)`$. We denote the sequence $`(\gamma (m)_n)_{n2}`$ by $`\gamma (m)`$. Thus we can write $`\gamma (m)=\rho (m)pr`$ where $`pr=(pr_n)_{n2}`$. We can define an element $`d_m\underset{¯}{Hom}(C,C)[1]`$ in the natural way: $`d_m=[m,]`$. In general $`d_m^20`$. In any case it can be naturally extended to a map $`f:_{cat}(𝒜_{\mathrm{}},A)(P,C)`$ such that $`f(m)=(d_m,\gamma (m))`$. We claim that this map is a morphism of dg-manifolds and it satisfies the condition $`pf=\nu `$ of the Theorem 1. It will be proved in the next subsection. ### 6.2 Proof of the theorem In the course of the proof we will not pay much attention to the signs. The reason for that was explained in the previous section. Namely, our second description of the operad $`M`$ (with reduced homology) gives automatically the agreement of signs. Proof of the theorem will occupy the rest of this subsection. We start with some general considerations. Let us recall that points of the moduli space $`Y=(P,C)`$ parametrize pairs $`(d_C,\gamma )`$ where $`d_C:CCL_1`$ is a morphism in $`Vect_𝐙`$, and $`\gamma `$ is an action of the space of generators $`G=(G_n)_{n2}`$ on $`C`$. Having the action $`\gamma (m)=\rho (m)pr`$ of $`P`$ on $`C`$ we would like to compute the odd vector field $`d_Y`$ at the point $`(d_m,\gamma (m))`$. Clearly $`d_Y(d_C,\gamma )=(d_C^2,\overline{\gamma })`$ where $`\overline{\gamma }=(\overline{\gamma }_2,\overline{\gamma }_3,\mathrm{})`$ and $`\overline{\gamma }_n\underset{¯}{Hom}(G_n_{S_n}C^n,CL_1)`$. We need to compute $`\overline{\gamma }`$. One can extend $`d_C`$ to $`\underset{¯}{Hom}(C^n,C)`$ using the Leibniz rule. We denote this extension by $`d_C^{(n)}`$. To write down $`\overline{\gamma }`$ we need to know the images of the generators of $`P`$ under all $`\gamma _n`$. We know that the maps $`pr_n`$ send to zero all of them except of the trees with one non-labeled vertex and $`n`$ numbered tail vertices. The latter is depicted below Then the direct computation together the previous lemma show that the following result holds ###### Lemma 5 The component $`\overline{\gamma }_n`$ of $`d_Y(f(m))`$ is equal to $$(d_m^{(n)}\rho (m)_n+\rho (m)_nd_{M_n})pr_n$$ where $`d_{M_n}`$ is the differential $`d_M`$ being restricted to $`M_n`$. Proof. We sketch the proof. First we observe that all components of $`\overline{\gamma }_n`$, which corresponds to the generators of $`G_n`$ with more than two internal edges, vanish. Using the previous lemma one can show that the generators with one internal edge give no input as well. After that one can make computations with elements of the operad $`M`$ only. Then the direct computation proves the lemma. $`\mathrm{}`$ We would like to prove that the map $`f:X=_{cat}(𝒜_{\mathrm{}},A)Y=(P,C)`$ is a dg-map, i.e. it transforms the odd vector field $`d_X`$ into the odd vector field $`d_Y`$. Vector field on $`X`$ is given by $`\dot{m}=d_X(m)=\frac{1}{2}[m,m]`$. The image of the map $`f`$ belongs to the vector subspace $`\underset{¯}{Hom}(C,C)[1]\underset{¯}{Hom}(M^{}(C),C)`$ of $`Y`$. Therefore the image $`\dot{f}(m)=f_{}(d_X(m))`$ of the tangent vector $`d_X(m)`$ belongs to the same vector space. Thus, we have to show that the second component $`\dot{\gamma }`$ of $`\dot{f}(m)`$ is equal to $`\overline{\gamma }`$. Its first component is equal to $`\frac{1}{2}[\dot{m},]`$ which is the same as $`\frac{1}{2}[[m,m],]=d_m^2`$. Notice that $`\overline{\gamma }\dot{\gamma }`$ can be considered as an action of an action of $`M^{}`$ on $`C`$. We decompose it into the sum of terms corresponding to planar trees from $`Tree^{(p)}`$. For such a tree $`T`$ the component of $`\overline{\gamma }\dot{\gamma }`$ is a sum of four terms described below. A. These terms correspond to the differential $`d_M`$ of the operad $`M`$. We can schematically write them as $`(d_MT)((c_i),m)`$. We are inserting $`c_iC`$ in labeled vertices and $`m`$ in non-labeled vertices. B. These terms can be schematically written as $`_iT([m,c_i],(c_j)_{ji},m)`$. We are inserting $`m`$ in non-labeled vertices, elements $`c_j`$ and $`[m,c_i]`$ in the corresponding labeled vertices. C. These terms can be schematically written as $`[m,T((c_i),m)]`$ in the notation above. These terms appear when we apply the differential (= commutator with $`m`$) to the tree with $`c_i`$ inserted in labeled vertices and $`m`$ inserted in non-labeled vertices. D. These terms correspond to $`\dot{\gamma }`$. Each of them consists of the replacement of $`m`$ in one non-labeled vertex $`v`$ by $`\dot{m}`$. The latter can be in turn replaced by $`\frac{1}{2}[m,m]`$. We can schematically write the resulting sum as $`_{vV_{nonl}(T)}T((c_i),\frac{1}{2}[m,m])`$. We are inserting the element $`m`$ in all non-labeled vertices of $`T`$ except $`v`$, the element $`\frac{1}{2}[m,m]`$ in the non-labeled vertex $`v`$ and $`c_i`$ to all labeled vertices. Notice that although a tree $`T`$ is always admissible, planar trees appearing in the decomposition of $`\dot{\gamma }`$ are not necessarily admissible. This means that valencies of some non-labeled vertices can be either $`0`$ or $`1`$. In order to depict all four cases we use the following notation: composition of an operation $`\alpha `$ sitting in a vertex $`v`$ with $`m`$ produces a new tree with a new non-labeled vertex $`w`$, as well as a new edge $`(w,v)`$ where $`N(w)=v`$. Similarly, a composition of $`m`$ with $`\alpha `$ produces a new tree with a non-labeled vertex $`w`$, as well as a new edge $`(v,w)`$ such that $`N(v)=w`$. The commutator $`[\alpha ,m]`$ corresponds to the difference of the above-mentioned trees. This agreement will be used also in the case when $`\alpha =m`$ thus giving the way to depict trees with $`\frac{1}{2}[m,m]`$ inserted. In the pictures below we show neighborhoods of vertices where the original tree changes. Then we split these terms in the following way: $`A=A_1+A_2+A_3`$, where: a) the terms $`A_1`$ correspond to the two adjoint non-labeled vertices of valency $`2`$ with $`m`$ inserted in each; b) the terms $`A_2`$ correspond to the labeled vertex adjoint to a non-labeled one of valency $`2`$ with $`m`$ inserted in the latter; c) the terms $`A_3`$ correspond to the non-labeled vertex of valency $`2`$ with $`m`$ inserted in it adjoint to a labeled vertex. These cases can be depicted as follows Similarly we split the terms $`B`$ such as follows: $`B=B_{+,1}+B_{+,2}+B_{,0}+B_{,1}+B_{,2}`$, where individual summands are depicted below. We split the terms $`C`$ such as follows: $`C=C_{root}^{}+C_{root}^{\prime \prime }+C_{}+C_{}`$ where individual summands are depicted below. We split the terms $`D`$ such as follows: $`D=D_0+D_{1,A}+D_{1,B}+D_2`$ where individual summands are depicted below. We see that $`A_1+D_2=0`$, $`A_2+B_{+,2}=0`$, $`A_3+B_{,2}=0`$, $`B_{,0}+C_{}=0`$, $`D_0+C_{}=0`$. Furthermore we can split each of the remaining terms into summands $`B_{+,1}=B_{+,1}^{}+B_{+,1}^{\prime \prime }+B_{+,1}^{\prime \prime \prime },B_{,1}=B_{,1}^{}+B_{,1}^{\prime \prime }`$, $`D_{1,A}=D_{1,A}^{}+D_{1,A}^{\prime \prime }+D_{1,A}^{\prime \prime \prime }`$, $`D_{1,B}=D_{1,B}^{}+D_{1,B}^{\prime \prime }`$. This splitting is depicted below. We combine these summands into six groups corresponding to the six types of edges depicted below We see that $`B_{+,1}^{}+C_{root}^{}=0,C_{root}^{\prime \prime }+D_{1,A}^{}=0,B_{+,1}^{\prime \prime }+B_{,1}^{}=0`$, $`B_{,1}^{\prime \prime }+D_{1,A}^{\prime \prime }=0,B_{+,1}^{\prime \prime \prime }+D_{1,B}^{}=0,D_{1,A}^{\prime \prime \prime }+D_{1,B}^{\prime \prime }=0`$. Then we conclude that $`B_{+,1}+C_{root}^{}+C_{root}^{\prime \prime }+D_{1,A}+B_{,1}+D_{1,B}=0`$. Thus $`A+B+C+D=0`$. This means that the map $`f:XY`$ is a dg-map. Obviously it is $`GL(A)`$-equivariant and satisfies the condition $`pf=\nu `$. This concludes the proof of the Theorem 1. ### 6.3 Remark about a generalization Let $`A`$ be an $`A_{\mathrm{}}`$-algebra, and $`C=C^{}(A,A)`$ be its Hochschild complex. Admissible planar trees with $`n`$ labeled vertices give rise to operations $`C^nC,n1`$. Analogously, planar trees with $`n`$ labeled vertices and $`m`$ tails, $`n0,m1`$ give rise to operations $`C^nA^mA`$. Let us restrict ourselves to such trees that $`|v|2`$ for all non-labeled internal vertices $`v`$. Thus we obtain a colored operad $`M^{(2)}`$ with two colors (Alg, Hoch) acting on the set of pairs $`(A,C^{}(A,A))`$, where $`A`$ is an $`A_{\mathrm{}}`$-algebra. Clearly $`M^{(2)}`$ contains as suboperads both $`𝒜_{\mathrm{}}`$ and $`M`$. Presumably a result analogous to the Theorem $`1`$ holds for the colored operad $`M^{(2)}`$. ## 7 Deligne’s conjecture ### 7.1 Preliminaries We are going to prove the following result. ###### Theorem 3 Let $`P`$ be the free resolution of the minimal operad $`M`$ constructed in Section 6. Then there is a homomorphism of dg-operads $`PChains(FM_2)`$ which induces an isomorphism on cohomology (i.e. it is a quasi-isomorphism of dg-operads). Here $`Chains(FM_2)`$ is the chain operad for the Fulton-Macpherson operad of configurations of points in $`𝐑^2`$ (see Section 7.2 below). We will construct a homomorphism of the dg-operads which induces a quasi-isomorphism of the chain complexes. Such a homomorphism is not defined canonically. Different choices are naturally parametrized by a contractible topological space. Using the fact that the operad $`Chains(FM_2)`$ is free as an operad (not as dg-operad), one can invert the quasi-isomorphism mentioned in the theorem. Since the Hochschild complex is a $`P`$-algebra, we obtain the following Corollary known as Deligne’s conjecture (see for example \[Ko3\], \[V\], \[MS\]). ###### Corollary 1 a) The Hochschild complex $`C^{}(A,A)`$ of an $`A_{\mathrm{}}`$-algebra $`A`$ can be equipped with a structure of an algebra over the operad $`Chains(FM_2)`$. b) The corresponding structure of a $`H^{}(Chains(FM_2))`$-algebra on the Hochschild cohomology $`H^{}(A,A)`$ coincides with the standard structure of a Gerstenhaber algebra on the Hochschild cohomology of an $`A_{\mathrm{}}`$-algebra. Same results remain true with $`Chains(FM_2)`$ being replaced by the operad $`Chains(E_2)`$ of chains on the little disc operad $`E_2`$ (see Section 7.2 for the definition). ###### Remark 6 It is easy to see that the operad $`P`$ can be defined over $`𝐙`$. It follows that such an operad acts on the Hochschild complex of an $`A_{\mathrm{}}`$-algebra defined over a field of arbitrary characteristic. The quasi-isomorphism $`PChains(FM_2)`$ can be also defined over the ring of integers (see also \[MS\]). The proof of the Theorem and the Corollary will occupy the rest of the section. We are going to use the following strategy. Let $`𝐓`$ be a “meta-tree” corresponding to a generator of $`P`$ (see Section 6 and Definition 17 below). 1) To every $`𝐓`$ we are going to associate a contractible closed subspace $`X_𝐓`$ in the Fulton-Macpherson compactification of the configuration space of points in $`𝐑^2`$ modulo shifts and dilations. 2) The collection of subspaces $`(X)_𝐓`$ will satisfy the following properties: a) The correspondence $`𝐓X_𝐓`$ is $`S_n`$-equivariant; b) if a tree $`𝐓^{}`$ appears as a summand in the formula for $`d_P(𝐓)`$ then $`X_𝐓^{}X_𝐓`$; c) if a composition of trees $`𝐓_1𝐓_2P`$ appears as a summand in the formula for $`d_P(𝐓)`$ then the operadic composition $`X_{𝐓_1}X_{𝐓_2}`$ belongs to the stratum $`X_𝐓`$. The latter composition of the strata has meaning because the Fulton-Macpherson compactifications form a topological operad. 3) For any generator $`𝐓`$ of the operad $`P`$ we will choose inductively chains $`\gamma _𝐓Chains(FM_2)`$ such that $`Supp(\gamma _𝐓)X_𝐓`$, where $`Supp`$ means the support of a chain. In this way we obtain a homomorphism of dg-operads $`PChains(FM_2)`$. 4) This homomorphism is a quasi-isomorphism. This will follow from the fact that $`P`$ is quasi-isomorphc to the minimal operad $`M`$, and on the level of chain complexes (not operads) every $`M_n`$ is quasi-isomorphic to $`Chains(FM_2(n))`$. ###### Remark 7 The theorem and its proof seem to admit a generalization to the case of higher dimensions. ### 7.2 Little discs operad and Fulton-Macpherson operad We recall here the definitions of both operads following \[Ko3\]. We fix the dimension $`d1`$. Let us denote by $`G_d`$ the $`(d+1)`$-dimensional Lie group acting on $`𝐑^d`$ by affine transformations $`u\lambda u+v`$, where $`\lambda >0`$ is a real number and $`v𝐑^d`$ is a vector. This group acts simply transitively on the space of closed discs in $`𝐑^d`$ (in the usual Euclidean metric). The disc with center $`v`$ and with radius $`\lambda `$ is obtained from the standard disc $$D_0:=\{(x_1,\mathrm{},x_d)𝐑^d|x_1^2+\mathrm{}+x_d^21\}$$ by a transformation from $`G_d`$ with parameters $`(\lambda ,v)`$. ###### Definition 15 The little discs operad $`E_d=\{E_d(n)\}_{n0}`$ is a topological operad defined such as follows: 1) $`E_d(0)=\mathrm{}`$, 2) $`E_d(1)=\mathrm{point}=\{\mathrm{id}_{E_d}\}`$, 3) for $`n2`$ the space $`E_d(n)`$ is the space of configurations of $`n`$ disjoint discs $`(D_i)_{1in}`$ inside the standard disc $`D_0`$. The composition $`E_d(k)\times E_d(n_1)\times \mathrm{}\times E_d(n_k)E_d(n_1+\mathrm{}+n_k)`$ is obtained by applying elements from $`G_d`$ associated with discs $`(D_i)_{1ik}`$ in the configuration in $`E_d(k)`$ to configurations in all $`E_d(n_i),i=1,\mathrm{},k`$ and putting the resulting configurations together. The action of the symmetric group $`S_n`$ on $`E_d(n)`$ is given by renumeration of indices of discs $`(D_i)_{1in}`$. The space $`E_d(n)`$ is homotopy equivalent to the configuration space of $`n`$ pairwise distinct points in $`𝐑^d`$. There is an obvious continuous map $`E_d(n)\mathrm{𝖢𝗈𝗇𝖿}_n(Int(D_0))`$ which associates to a collection of disjoint discs the collection of their centers. This map induces a homotopy equivalence because its fibers are contractible. The little discs operad and homotopy equivalent little cubes operad were introduced in topology by J. P. May in order to describe homotopy types of iterated loop spaces. The Fulton-Macpherson operad defined below is homotopy equivalent to the little discs operad. For $`n2`$ we denote by $`\stackrel{~}{E}_d(n)`$ the quotient space of the configuration space of $`n`$ points in $`𝐑^d`$ $$\mathrm{𝖢𝗈𝗇𝖿}_n(𝐑^d):=\{(x_1,\mathrm{},x_n)(𝐑^d)^n|x_ix_j\mathrm{for}\mathrm{any}ij\}$$ by the action of the group $`G_d`$. The space $`\stackrel{~}{E}_d(n)`$ is a smooth manifold of dimension $`d(n1)1`$. For $`n=2`$, the space $`\stackrel{~}{E}_d(n)`$ coincides with the $`(d1)`$-dimensional sphere $`S^{d1}`$. There is an obvious free action of $`S_n`$ on $`\stackrel{~}{E}_d(n)`$. We define the spaces $`\stackrel{~}{E}_d(0)`$ and $`\stackrel{~}{E}_d(1)`$ to be empty. The collection of spaces $`\stackrel{~}{E}_d(n)`$ does not form an operad because there is no identity element, and compositions are not defined. Now we are ready to define the operad $`FM_d=\{FM_d(n)\}_{n0}`$ The components of the operad $`FM_d`$ are 1) $`FM_d(0):=\mathrm{}`$, 2) $`FM_d(1)=`$point, 3) $`FM_d(2)=\stackrel{~}{E}_d(2)=S^{d1}`$, 4) for $`n3`$ the space $`FM_d(n)`$ is a manifold with corners, its interior is $`\stackrel{~}{E}_d(n)`$, and all boundary strata are certain products of copies of $`\stackrel{~}{E}_d(n^{})`$ for $`n^{}<n`$. The spaces $`FM_d(n),n2`$ can be defined explicitly. ###### Definition 16 For $`n2`$, the manifold with corners $`FM_d(n)`$ is the closure of the image of $`\stackrel{~}{E}_d(n)`$ in the compact manifold $`\left(S^{d1}\right)^{n(n1)/2}\times [0,+\mathrm{}]^{n(n1)(n2)}`$ under the map $`G_d(x_1,\mathrm{},x_n)(\left(\frac{x_jx_i}{|x_jx_i|}\right)_{1i<jn},\frac{|x_ix_j|}{|x_ix_k|})`$ where $`i,j,k`$ are pairwise distinct indices. One can define the natural structure of operad on the collection of spaces $`FM_d(n)`$. We skip here the obvious definition. It is easy to check that in this way we obtain a topological operad (in fact an operad in the category of real compact piecewise algebraic sets defined in Appendix). We call it the Fulton-Macpherson operad and denote by $`FM_d`$. Set-theoretically, the operad $`FM_d`$ is the same as the free operad generated by the collection of sets $`(\stackrel{~}{E}_d(n))_{n0}`$ endowed with the $`S_n`$-actions discribed above. Using piecewise algebraic chains from Appendix we define dg-operads $`Chains(E_d)`$ and $`Chains(FM_d)`$ (they are operads in the symmetric monoidal category of complexes of abelian groups). Since we are working over the ground field $`k`$ of characteristic zero, the dg-operads will be complexes of $`k`$-vector spaces. Notice that when we write $`Chains(Z)`$ we mean the $`cohomological`$ complex with graded components $`Chains_i(Z)[i],i0`$ concentrated in $`negative`$ degrees. The complex $`P_n,n2`$ is concentrated in degrees $`[2n3,\mathrm{},0]`$. The same is true for the complex $`Chains(FM_2(n))`$ (we use piecewise algebraic chains) because $`dim_𝐑FM_2(n)=2n3`$. For every $`n2`$ the space $`P_n`$ has a canonical (up to signs) basis, called the standard basis, with elements labeled by certain combinatorial objects, which we will call meta-trees. ###### Definition 17 A meta-tree $`𝐓`$ with $`n`$ tails is given by the following data: a) an abstract tree $`TTree(n)`$ together with a marking $`E_i(T)\{finite,infinite\}`$; b) for every internal vertex $`v`$ of $`T`$ we have: $`|v|2`$; c) to every $`vV_i(T)`$ we assign an admissible labeled tree $`T_v`$ together with a bijection $`V_{lab}(T_v)N_{T_v}^1(v)`$. We denote by $`MT(n)`$ the set of isomorphism classes of meta-trees with $`n`$ tails. ###### Conjecture 1 There is a piecewise algebraic (see Appendix) cell decomposition of the spaces $`FM_2(n),n2`$, with the cells $`\sigma _𝐓`$ labeled by $`MT(n)`$ such that: 1) the correspondence $`𝐓\sigma _𝐓`$ is $`S_n`$-equivariant; 2) the operadic composition of any two cells is again a cell; 3) there is a morphism of dg-operads $`Chains_{(\sigma _𝐓)}(FM_2)P`$ such that every cell $`\sigma _𝐓`$ is mapped (up to a sign) to the corresponding element of the standard basis of $`P`$. Here $`Chains_{(\sigma _𝐓)}(FM_2)`$ denotes the chain subcomplex of $`Chains(FM_2)`$ formed by $`k`$-linear combinations of cells $`\sigma _𝐓`$; 4) for any $`n2`$ the cell decomposition of $`FM_2(n)`$ formed by cells $`(\sigma _𝐓),𝐓MT(n)`$ is regular, i.e. the closure of every cell is homeomorphic to a closed ball. This conjecture implies Deligne’s conjecture. Let us now introduce a partial order on the set $`MT(n),n2`$. ###### Definition 18 Let $`𝐓,𝐓^{}`$ be meta-trees with $`n`$ tails. We say that $`𝐓<𝐓^{}`$ if there exists a sequence of meta-trees $`(𝐓_0,\mathrm{},𝐓_m)`$ such that $`𝐓_0=𝐓,𝐓_m=𝐓^{}`$ and for any $`i`$ such that $`0im1`$ we have: $`𝐓_i`$ appears as a summand in the decomposition of $`d_P𝐓_{i+1}`$ with respect to the standard basis of $`P_n`$. It follows from the condition $`4)`$ of the Conjecture above, that the nerve of the partially ordered set $`(MT(n),<)`$ is homeomorphic to $`FM_2(n)`$. The following conjecture also follows from the Conjecture 1. ###### Conjecture 2 The nerve of $`(MT(n),<)`$ is a PL-manifold with the boundary. The above-mentioned homemorphism with $`FM_2(n)`$ is a homeomorphism of PL-manifolds with boundaries. We checked this Conjecture for small $`n`$. ### 7.3 Partial orders induced by trees We recall here the structure of the space $`G_n`$ of generators of the components $`P_n,n1`$ of the free operad $`P`$ (see Section 6). There is a basis of $`G_n`$ elements of which are (up to signs) parametrized by meta-trees $`𝐓=(T,\{T_v\}_{vV_i(T)})`$ such that all internal edges of $`T`$ are finite. The degree of the generator $`𝐓=(T,\{T_v\}_{vV_i(T)})`$ is equal to $$deg(𝐓)=\underset{vV_i(T)}{}deg(T_v)|E_i(T)|$$ Let $`𝐓`$ be a generator of $`P_n`$. We are going to introduce on the set $`\{1,2,\mathrm{},n\}`$ two partial orders $`<_{h,𝐓}`$ and $`<_{v,𝐓}`$ (called horizontal and vertical). Although these orders will depend on $`𝐓`$, we will skip $`𝐓`$ from the notation if it does not lead to a confusion. Let $`i,j\{1,\mathrm{},n\},ij`$. We have two tail vertices of $`T`$ labeled by $`i`$ and $`j`$ respectively. Then there exists a unique internal vertex $`v`$ of $`T`$ satisfying the following properties: a) $`N_T^k(i)=N_T^l(j)=v`$ where $`k,l`$ are positive integers; b) the vertex $`v`$ is minimal among those satisfying a) (which means that $`k,l`$ are both minimal in a). Then there exist unique labeled vertices $`x,yT_v`$ such that $`N_T^{k1}(i)=x,N_T^{l1}(j)=y`$. Since $`T_v`$ is a planar tree, we can compare $`x`$ and $`y`$ with respect to exactly one of the following partial orders: “$`x`$ is to the left of $`y`$ in $`T_v`$” or “$`x`$ is above $`y`$ in $`T_v`$”. We will call them the “horizonal” and “vertical” order respectively. Let us describe the orders in $`T_v`$ more precisely. We say that $`x`$ is above $`y`$ (or $`y`$ is below $`x`$) in $`T_v`$ if there exists a positive integer $`a`$ such that $`N_{T_v}^a(y)=x`$. We say that $`x`$ is to the left of $`y`$ (or $`y`$ is to the right of $`x`$) if there exist positive integers $`a,b`$ such that $`N_{T_v}^a(x)=N_{T_v}^b(y)=w`$, but $`N_{T_v}^{a1}(x)N_{T_v}^{b1}(y)`$ and $`N_{T_v}^{a1}(x)`$ preceeds $`N_{T_v}^{b1}(y)`$ in $`N_{T_v}^1(w)`$ with respect to the order on the latter set given by the planar structure on $`T_v`$. Thus, we have defined two partial orders on vertices of $`(T_v)_{vV_i(T)}`$. They induce partial orders $`<_{h,𝐓}`$ (horizontal) and $`<_{v,𝐓}`$ (vertical) on the set $`\{1,\mathrm{},n\}`$. We have identified the latter with the set of tails of $`T`$. Namely, we say that $`i<_{h,𝐓}j`$ if (in the above notation) the vertex $`x`$ is to the left of $`y`$. We say that $`i<_{v,𝐓}j`$ if $`y`$ is above $`x`$ (equivalently we say that $`x`$ is below $`y`$). The following lemma is easy to prove. ###### Lemma 6 Both horizontal and vertical orders are indeed partial orders (i.e. they satisfy all the axioms of orders). ###### Definition 19 Let $`S`$ be a set, $`<_1`$ and $`<_2`$ be partial orders on $`S`$. We will call them $`complementary`$ if any two elements of $`S`$ can be compared with respect to exactly one of them. This means that for any two elements $`i,jS,ij`$ exactly one of the following properties holds: $`i<_1j`$ , $`j<_1i`$, $`i<_2j`$, $`j<_2i`$. Suppose that we have two partial orders as in the Definition. We define on $`S`$ two new pre-orders $`<_{1+2}`$ and $`<_{12}`$ such as follows: (i) $`x<_{1+2}y`$ if $`x<_1y`$ or $`x<_2y`$; (ii) $`x<_{12}y`$ if $`x<_1y`$ or $`y<_2x`$. Notice that we can reconstruct $`<_1`$ and $`<_2`$ from these new orders. For example $`x<_1y`$ is equivalent to the conjunction: $`(x<_{1+2}y)`$ and $`(x<_{12}y)`$. ###### Proposition 5 Formulas (i) and (ii) define complete orders on the set $`S`$. Proof. Straightforward.$`\mathrm{}`$ Using this result one can easily prove the following one. ###### Proposition 6 Let $`<_1`$ and $`<_2`$ be a pair of two complementary orders given on a finite set $`S`$. Then there exists a unique element $`s_0S`$ such that for any $`iS,is_0`$ we have: either $`s_0<_1i`$ or $`s_0<_2i`$. $`Proof`$. The element $`s_0`$ is minimal with respect to the complete order $`<_{1+2}`$.$`\mathrm{}`$ Summarizing, we can say that on the set of tails of a generator of $`P_n`$ we have defined two complementary partial orders (or, equivalently, two complete orders). These orders will be used below when we will construct the closed sets $`X_𝐓`$. ### 7.4 Closed sets $`X_𝐓`$ Let us recall that for any $`n2,1i,jn,ij`$ we have a natural projection $`p_{i,j}:FM_2(n)FM_2(2)`$ (forgetting all points in $`(x_1,\mathrm{},x_n)`$ except $`x_i`$ and $`x_j`$). As a topological space $`FM_2(2)`$ is identified to the unit circle $`S^1𝐑^2`$ via the map $`G_2(x_1,x_2)\frac{x_2x_1}{|x_2x_1|}`$. We denote by $`S_{+,v}^1FM_2`$ the closed upper-half circle, and by $`S_{+,h}^1S^1`$ the one-element subset consisting of the point $`\{(1,0)\}`$. Then $`S_{+,v}^1`$ corresponds to the configurations $`(x_1,x_2)𝖢onf_2(𝐑^2)`$ such that if we put $`x_2x_1=re^{i\alpha },0\alpha <2\pi `$ then $`\alpha [0,\pi ]`$. Similarly the subset $`S_{+,h}^1`$ corresponds to the configurations $`(x_1,x_2)`$ such that both $`x_j,j=1,2`$ belong to the same horizontal line, and $`x_1`$ is positioned to the left of $`x_2`$. Suppose that we are given two complementary orders $`<_h`$ and $`<_v`$ on the set $`\{1,\mathrm{},n\},n2`$. Then we define the following subset of $`FM_2(n)`$: $`X_{<_h,<_v}=\{xFM_2(n)|i<_hjp_{i,j}(x)S_{+,h}^1,i<_vjp_{i,j}(x)S_{+,v}^1\}.`$ For a generator $`𝐓P_n`$ we define $`X_𝐓`$ as $`X_{<_{h,𝐓},<_{v,𝐓}}`$. First of all we would like to prove that $`X_𝐓`$ is contractible. This is a special case of a more general statement. ###### Proposition 7 For any pair of complementary orders $`<_h`$ and $`<_v`$ given on the set $`\{1,\mathrm{},n\},n2`$ the subspace $`X_{<_h,<_v}`$ is non-empty and contractible. Proof. It can be done by induction. For $`n=2`$ the result is clear, since subsets $`S_{+,v}^1`$ and $`S_{+,h}^1`$ are contractible. Suppose that the Proposition is true for $`n1`$ points. Let us take the element $`i_0`$ which is minimal in $`I_n=\{1,\mathrm{},n\}`$ with respect to $`<_{h+v}`$. We have already proved that it exists. Let us consider the induced complementary partial orders $`<_h^{},<_v^{}`$ on the set $`I_n\{i_0\}`$. They define the subset $`X_{<_h^{},<_v^{}}FM_2(n1)`$ which is non-empty and contractible by induction. Then the result is a corollary of the following observation: the fibers of the natural projection $`\pi :X_{<_h,<_v}X_{<_h^{},<_v^{}}`$ are contractible. Basically it follows from the fact that the point $`x_{i_0}`$ is either left or below of all the points $`x_i,1in,ii_0`$. Let us prove that the fibers of $`\pi `$ are non-empty. Indeed, for any $`x^{}X_{<_h^{},<_v^{}}`$ the operadic composition $`(1,0)_jx^{}`$ belongs to the fiber $`\pi ^1(x^{})`$. Here $`(1,0)S^1`$ is considered as a point in $`FM_2(\{i_0,j\})`$, and $`j`$ is an auxiliary index. We leave to the reader the proof of contractibility of the fibers of $`\pi `$. $`\mathrm{}`$ ###### Remark 8 a) One can prove that all homotopies can be taken in the category of piecewise algebraic sets. b) It follows from the construction that the map $`𝐓X_𝐓`$ is $`S_n`$-equivariant. Now we would like to explain the Property $`2`$ of $`X_𝐓`$ stated in Section 7.1. First of all, the $`S_n`$-equivariance is obvious. Let $`𝐓G`$ be a generator of $`P`$. We recall that the differential in $`P`$ can be written schematically (up to signs) as $$d_P(𝐓)=\underset{vV_i(T),l}{}𝐓_{v,l}+\underset{\alpha E_{infinite}(T)}{}𝐓_\alpha ^{}𝐓_\alpha ^{\prime \prime }+\underset{\alpha E_{finite}(T),j}{}𝐓_{\alpha ,j}.$$ Here meta-trees $`𝐓_{v,l}G`$ arise from the application of the differential of $`M`$ to the tree $`T_v`$ inscribed into the vertex $`vV_i(T)`$. This differential was described in Section $`5`$. Index $`l`$ runs over all possible insertions of a new edge. The second summand corresponds to the tree, obtained from $`T`$ by making a finite edge $`\alpha `$ into an infinite edge. The result is a composition of two generators of $`P`$ which we denote by $`𝐓_\alpha ^{}`$ and $`𝐓_\alpha ^{\prime \prime }`$. The last sum corresponds to the operation “contract a finite edge $`\alpha `$” in $`T`$. Then the planar trees from $`M`$ inscribed into the vertices which are endpoints of $`\alpha `$ must be composed and inscribed into the new vertex. The result is a sum of generators $`𝐓_{\alpha ,j}`$ of $`P`$. ###### Proposition 8 In the above notation we have: $`X_{𝐓_{v,l}},X_{𝐓_\alpha ^{}}X_{𝐓_\alpha ^{\prime \prime }},X_{𝐓_{\alpha ,j}}`$ belong to $`X_𝐓`$. Proof. Straightforward check which uses the fact that $`S_{+,h}^1S_{+,v}^1`$. $`\mathrm{}`$ ### 7.5 Morphism $`PChains(FM_2)`$ We would like to consctruct the chains $`\gamma _𝐓Chains(FM_2)`$ where $`𝐓`$ runs through the set of generators of $`P`$. Let us explain the idea. We will construct $`\gamma _𝐓`$ by induction in the degree of $`𝐓`$. a) Let us assume that $`deg(𝐓)=0`$. Then $`T`$ has the only internal vertex $`v`$ and the corresponding planar tree $`T_v`$ is a binary tree. The corresponding element of the operad $`P`$ is defined up to a sign. We make the following choice: it is just the composition of several copies of the operation $`m_2M_2.`$ The corresponding chain $`\gamma _𝐓`$ should have the dimension zero. Therefore it must be a linear combination of points with integer coefficients. We choose $`\gamma _𝐓`$ to be just one point in $`X_𝐓`$ with the multiplicity equal to $`+1`$. b) Let us assume that $`deg(𝐓)=1`$. Then the formula for $`d_P(𝐓)`$ contains two summands $`𝐓^{},𝐓^{\prime \prime }`$ corresponding to the two binary planar trees. It is not difficult to check that they appear with the opposite coefficients $`\pm 1`$. According to a) these summands define $`0`$-dimensional chains $`\gamma _𝐓^{},\gamma _{𝐓^{\prime \prime }}`$. We define $`\gamma _𝐓`$ to be the only (up to a boundary) $`1`$-chain having $`\gamma _𝐓^{}\gamma _{𝐓^{\prime \prime }}`$ as the boundary. This $`1`$-chain exists because of the condition imposed on multiplicities of $`0`$-chains. c) Suppose that we have a generator $`𝐓`$, $`deg(𝐓)=k,k2`$. Then we can find a chain $`\gamma _𝐓`$ of degree $`k`$ such that $$(\gamma _𝐓)=\underset{v,l}{}\gamma _{𝐓_{v,l}}+\underset{\alpha }{}\gamma _{𝐓_\alpha ^{}𝐓_\alpha ^{\prime \prime }}+\underset{\alpha ,j}{}\gamma _{𝐓_{\alpha ,j}}$$ where $``$ is the boundary operator in the chain complex for $`FM_2`$, and rest of the notation is self-explained. Indeed, the RHS of this formula is known by the induction assumption. We also know that it is a closed chain because $`d_P^2=0`$. Then one can always find a chain $`\gamma _𝐓`$ with the given boundary. Indeed, $`X_𝐓`$ is contractible, and we consider chains of negative degrees (zero degree case was considered in a)). The space parametrizing all different choices of $`\gamma _𝐓`$ is contractible, so our choice is unique in a given homology class. The map $`𝐓\gamma _𝐓`$ extends to a homomorphism $`\mathrm{\Phi }:PChains(FM_2)`$ of graded operads in such a way that $`\mathrm{\Phi }(𝐓)=\gamma _𝐓`$. We have checked that $`\mathrm{\Phi }`$ is compatible with the differentials. Therefore it is a homomorphism of dg-operads. ###### Theorem 4 The morphism $`\mathrm{\Phi }`$ is a quasi-isomorphism of complexes. $`Proof.`$ We will give only a sketch of the proof. The idea is to consider a subcomplex $`L`$ of the complex $`P`$ spanned by the trees without internal edges. This subcomplex is isomorphic to the complex of the minimal dg-operad $`M`$. Indeed, $`P`$ can be decomposed into a direct sum of $`L`$ and a contractible complex (see Proposition $`4`$ in Section $`4`$). Therefore it is enough to prove that the restriction $`\mathrm{\Phi }|_L:LChains(FM_2)`$ is a quasi-isomorphism of complexes. In order to do that we need to describe chains from the subcomplex $`\mathrm{\Phi }(L)`$. Our constructions were based on certain non-canonical choices of chains $`\gamma _𝐓,𝐓G`$. Now, for $`𝐓L`$ we will make specific choices of $`\gamma _𝐓`$. Let $`𝐓L`$ corresponds to a labeled tree $`T_vM`$. We will denote it simply by $`T_v`$, where $`T_v`$ is a labeled planar tree inscribed in the only internal vertex $`v`$ of $`𝐓`$. We associate with $`T_v`$ an abstract tree $`\widehat{T}_v`$ such that $`V_t(\widehat{T}_v)=V_i(T_v)`$, and the valency of each internal vertex of $`\widehat{T}_v`$ is at least $`2`$. Internal vertices of $`\widehat{T}_v`$ correspond to subsets $`T_{x_0}`$ described below. Let us choose an internal vertex $`x_0V_i(T_v)`$ and consider the set of such $`xV_i(T_v)`$ that $`N_{T_v}^j(x)=x_0`$ for some $`j0`$. Then internal vertices of $`\widehat{T}_v`$ correspond to such sets $`T_{x_0}`$ for which the cardinality $`|T_{x_0}|2`$. Moreover, there is a path in $`\widehat{T}_v`$ from a tail vertex $`vV_t(\widehat{T}_v)`$ to an internal vertex $`T_{x_0}`$ iff $`vT_{x_0}`$. It is well-known that any abstract tree $`\widehat{T}`$ with valencies of internal vertices at least $`2`$ gives rise to a stratum $`J(\widehat{T})FM_2(|V_t(\widehat{T})|)`$ (see \[FM\]). It is the operadic composition of $`FM_2(N_{\widehat{T}_v}^1(u)),uV_i(\widehat{T}_v)`$. We are going to construct a subspace $`X_{\widehat{T}}J(\widehat{T}_v)`$ The space $`X_{\widehat{T}}`$ will be constructed as the operadic composition of certain subspaces $$X_uFM_2(|N_{\widehat{T}_v}^1(u)|),uV_i(\widehat{T}_v).$$ Let $`u=T_{x_0}`$ be an internal vertex of $`\widehat{T}`$. The set of edges having $`u`$ as an endpoint is in one-to-one correspondence with the set $`\{x_0\}N_{T_v}^1(x_0).`$ The subspace $`X_u`$ consists of configurations of points $`G_2(p_{x_0},(p_y)_{yN_T^1(x_0)})`$ such that $`p_{x_0}=(0,1)𝐑^2`$, all points $`p_y`$ belong to the horizontal line $`\{(x,0)|x𝐑\}𝐑^2`$, and their order on this line is the same as their order in $`N_{T_v}^1(x_0)`$. One can check that: a) $`X_{\widehat{T}_v}`$ is an open cell. b) The natural projection (forgetting map) $`FM_2(|V_i(T_v)|)FM_2(|V_{lab}(T_v)|)`$ maps $`X_{\widehat{T}_v}`$ onto an open cell $`X_{T_v}X_𝐓`$. c) the closure $`\overline{X_{T_v}}`$ is a manifold with corners (more precisely, a real piecewise algebraic manifold). d) the boundary of $`\overline{X_{T_v}}`$ is the union of cells of the same type . Using a)-d) one checks that the construction above gives rise to a homomorphism of complexes $`\chi :LChains(FM_2)`$, and moreover, it coincides with the homomorphism $`\mathrm{\Phi }|_L`$. To be more precise, one observes that the construction of $`\mathrm{\Phi }`$ was not canonical. We made certain choices when constructed a chain with the prescribed boundary. The point is that one can choose inductively the chains in such a way that the restriction $`\mathrm{\Phi }|_L`$ of the homomorphism $`\mathrm{\Phi }:PChains(FM_2)`$ coincides with $`\chi `$. We claim that $`\chi `$ induces a homotopy equivalence. This follows from the ###### Lemma 7 Let $`MT_1(n)`$ denotes the subset of $`MT(n)`$ consisting of meta-trees with only one internal vertex $`v`$. Then the natural embedding of the CW-complex $`X_n=_{𝐓MT_1(n)}X_{T_v}`$ into $`FM_2(n)`$ is a homotopy equivalence for all $`n2`$. Proof. It is not difficult to show that $`X_n`$ is isomorphic to the CW-complex $`\mathrm{\Sigma }_{(n)}`$ constructed in Section 5.5 via Strebel differentials. It follows from the fact that both are regular complexes and posets of their cells are isomorphic. Moreover, both complexes are $`K(\pi ,1)`$ spaces, classifying spaces for the pure braid group of $`n`$-strings. Moreover, the map $`\mathrm{\Phi }`$ induces an equivalence of the fundamental groupoids. Hence it is a homotopy equivalence. $`\mathrm{}`$ This Lemma conludes the proof of the Theorem $`2`$.$`\mathrm{}`$ ### 7.6 Proof of Deligne’s conjecture The results of the previous subsection establish the Theorem 2. In this subsection we will prove the Corollary (Deligne’s conjecture). First of all, we remark that both $`P`$ and $`Chains(FM_2)`$ are dg-operads which are free as graded operads. Therefore the quasi-isomorphism morphism of graded operads $`\mathrm{\Phi }:PChains(FM_2)`$ admits a homotopy inverse $`\mathrm{\Psi }:Chains(FM_2)P`$ (see for ex. \[M2\]). We have already proved that the Hochschild complex of an $`A_{\mathrm{}}`$-algebra is an algebra over the operad $`P`$. Indeed, $`P`$ is a free resolution of $`M`$, and the latter dg-operad acts on the Hochschild complex. Using the morphism $`\mathrm{\Psi }`$ we make the Hochschild complex into an algebra over the dg-operad $`Chains(FM_2)`$. This concludes the proof of Deligne’s conjecture. ## 8 Appendix: Singular chains and differential forms In the Appendix we are going to describe a theory of singular chains which is suitable for work with manifolds with corners. This formalism is helpful in the proof of formality of the operad of chains on the little disc operad (see \[Ko3\]). We used it in the proof of Deligne’s conjecture. The idea of the theory of singular chains developed below is the following: one can construct chains which produce complexes quasi-isomorphic to the standard complexes of singular chains, and built out of some kind of “piecewise algebraic spaces”. ### 8.1 Spaces We recall that real semialgebraic sets in $`𝐑^n`$ are subsets defined by a finite number of polynomial equations and inequalities. Constructible sets are obtained from semialgebraic ones by boolean operations. We define the category $`𝒫`$ of compact piecewise algebraic spaces (compact PA-spaces for short) in the following way. Objects of $`𝒫`$ are pairs $`(X,n)`$, $`n=1,2,\mathrm{}`$ such that $`X𝐑^n`$ is a compact constructible set (it is the same as a compact real semialgebraic set). For two objects $`(X,n)`$ and $`(Y,m)`$ the space of morphisms $`Hom((X,n),(Y,m))`$ is formed by continuous maps $`f:XY`$ such that $`graph(f)𝐑^n\times 𝐑^m`$ is constructible. In the future we will skip the index $`n`$ in the notation $`(X,n)`$ if it will not lead to a confusion. Obviously we have a functor from $`𝒫`$ to the category of compact Hausdorff topological spaces. An isomorphism in $`𝒫`$ is a morphism which is a homeomorphism of topological spaces. Let $`X𝒫`$. Then one can define a sheaf $`𝒪_X`$ of piecewise algebraic functions on $`X`$. To do this we note first that one can speak about constructible subsets of $`X`$. By definition they are constructible sets in the bigger space $`𝐑^n`$. Then for any open $`UX`$ we define $`𝒪_X`$ to be an $`𝐑`$-algebra of continuous functions $`f`$ on $`U`$ such that for any compact constructible $`VU`$ we have: $`graph(f|_V)`$ is constructible. It is easy to see that we get a sheaf $`𝒪_X`$ of algebras. ###### Lemma 8 Morphisms from $`X`$ to $`Y`$ in the category $`𝒫`$ are in one-to-one correspondence with homomorphisms of algebras $`𝒪(Y)𝒪(X)`$. Proof. Exercise. $`\mathrm{}`$ Similarly, we can consider non-compact case. ###### Definition 20 Piecewise algebraic space (PA-space for short) is a locally compact Hausdorff topological space $`X`$, equipped with the sheaf $`𝒪_X`$ of $`𝐑`$-algebras which is locally isomorphic to $`𝒪_X^{}`$ for some compact $`X^{}𝒫`$. Clearly PA-spaces form a category. Compact PA-spaces are exactly objects of $`𝒫`$. We define a $`d`$-dimensional PA-manifold with boundary as a PA-space $`X`$ which is modeled locally by the closed half-space $`𝐑_+^d`$. ### 8.2 Singular chains We would like to define an appropriate version of singular chains for PA-spaces. We will give two equivalent descriptions. First description. Let $`X`$ be a PA-space. We define the group of $`n`$-chains $`Chains_n(X):=C_n(X,𝐙)`$ as an abelian group generated by equivalence classes of triples $`(M,or,f)`$ such that: i) $`M`$ is a compact PA-manifold of dimension $`n`$; ii) $`f:MX`$ is a morphism in $`𝒫`$; iii) $`or`$ is an orientation of $`M`$. We need to define an equivalence relation. It is the same as to say when a finite linear combination $`_in_i(M_i,or_i,f_i)`$ is equal to zero in $`C_n(X,𝐙)`$. Notice that $`Y=_if_i(M_i)`$ carries the structure of a compact PA-space, of the dimension $`n`$. Then there exists a constructible subset $`Y_0Y`$ such that $`dim(YY_0)n1`$, and for any point $`yY_0`$, any $`i`$ and $`x_{i,\alpha }f_i^1(y)`$ there exists a neighborhood $`U_{i,\alpha }`$ of $`x_{i,\alpha }`$ such that $`Y_0f_i(U_{i,\alpha })`$ is a PA-manifold and the morphism $`f_i|_{U_{i,\alpha }}`$ is a homeomorphism of $`U_{i,\alpha }`$ onto its image. We choose an orientation $`or_y`$ of $`Y_0`$ near the point $`y`$. Then the above- mentioned linear combination is declared to be zero iff for every point $`yY_0`$ we have $`_{i,\alpha }n_isgn(i,\alpha ,y)=0`$ where $`sgn(i,\alpha ,y)`$ is defined as $`f_i|_{U_{i,\alpha }}(or_i)=sgn(i,\alpha ,y)or_y`$, and sum is taken over all points $`x_{i,\alpha }`$ such that $`f_i(x_{i,\alpha })=y`$. Notice that $`sgn(i,\alpha ,y)`$ always takes values $`\pm 1`$. Second description. We will define $`C_n(X)=C_n(X,𝐙)`$ as a quotient by certain equivalence relation of the set of quadruples $`(Y,Y_0,or,mult)`$ such that: a) $`YX`$ is a compact PA-subspace, $`dimYn`$, which contains an open dense constructible subset $`Y_0`$ without boundary which is a PA-manifold of dimension $`n`$; b) $`or`$ is an orientation of $`Y_0`$; c) $`mult:Y_0\{1,2,\mathrm{},\}`$ is a locally constant map (multiplicity). The equivalence relation is defined such as follows. We say that $`(Y,Y_0,or,mult)`$ is equivalent to $`(Y^{},Y_0^{},or^{},mult^{})`$ iff (i) $`Y=Y^{}`$ and there exists an open contsructible $`Y_0^{\prime \prime }Y_0^{}Y_0^{\prime \prime }`$ such that $`Y`$ is equal to the closure of $`Y_0^{\prime \prime }`$; (ii) restrictions of orientations and multiplicity functions of $`Y_0`$ and $`Y_0^{}`$ to $`Y_0^{\prime \prime }`$ coincide. Using the pairing between chains and differential forms (see 8.3), one can show that the first and the second descriptions give canonically isomorphic sets of chains. One can also check the following properties of singular chains: 1) there is a naturally defined differential $`:C_n(X,𝐙)C_{n1}(X,𝐙),^2=0`$ (in the Description 1 it is well-defined by the formula $`(M,or,f)=(M,(or),f|_M)`$). 2) the correspondence $`XC_{}(X,𝐙)`$ is a functor from the category of PA-spaces to the category of abelian groups; 3) $`C_i(X,𝐙)`$ vanishes for $`i<0`$ and $`i>dimX`$; 4) if $`X`$ is a compact oriented PA-manifold, then there exists a canonically defined chain $`[X]C_n(X,𝐙)`$. In the Description 1 it is defined by the formula $`[X]=(X,or,1)`$; 5) for any finite collection of PA-spaces $`(X_i)_{iI}`$ there is a natural homomorphism of complexes of abelian groups $`_{iI}C_{}(X_i,𝐙)C_{}(_{iI}X_i,𝐙)`$; 6) there is a naturally defined soft sheaf of complexes $`\underset{¯}{C}_{}^{closed}`$ on $`X`$ such that for a compact $`X`$ the abelian group $`\mathrm{\Gamma }(X,\underset{¯}{C}_{}^{closed})`$ coincides with $`C_{}(X,𝐙)`$. 7) the homology of $`C_{}(X,𝐙)`$ is naturally isomorphic to the usual singular homology $`H_{}(X,𝐙)`$. For non-compact PA-spaces we define locally finite chains as global sections of the sheaf $`\underset{¯}{C}_{}^{closed}`$. Then the homology of $`C_{}^{closed}(X,𝐙)`$ is isomorphic to the usual singular homology with locally compact support $`H_{}^{closed}(X,𝐙)`$. ###### Remark 9 a) Mayer-Vietoris sequence trivializes for PA chains: if $`X=Y_1Y_2`$ is a union of locally closed constructible subsets, then the corresponding short sequence of abelian groups $`0C_n(Y_1Y_2,𝐙)C_n(Y_1,𝐙)C_n(Y_2,𝐙)C_n(X,𝐙)0`$ is exact. b) The Property 5 is very convenient in order to formulate Deligne’s conjecture. It allows to avoid Eilenberg-Zilber theorem. c) Property 4 seems to be useful for the higher-dimensional generalization of Deligne’s conjecture (see \[Ko3\]). Namely, if $`A`$ is a $`d`$-algebra (i.e. an algebra over the dg-operad $`Chains(FM_d)`$), then $`A[d1]`$ carries a natural structure of an $`L_{\mathrm{}}`$-algebra. More precisely, for any $`n2`$ the fundamental cycle $`[FM_d(n)]`$ gives the $`n`$-th higher bracket on $`A[d1]`$. ### 8.3 Differential forms: first approximation Let $`X`$ be a PA-space. We would like to define an appropriate notion of differential form on $`X`$. In this subsection we construct a first approximation to the future algebra of differential forms. ###### Definition 21 Sheaf $`\mathrm{\Omega }_{X,min}^k`$ is locally defined as a vector subsubspace in $`Hom(C_k(X,𝐙),𝐑)`$ generated by functionals $`l=(f_0,f_1,\mathrm{},f_k),f_i𝒪_X`$ such that $$l(M,or,\varphi )=_{M_0}\varphi ^{}(f_0df_1df_2\mathrm{}df_k).$$ Here $`M_0=M`$ is a dense open constructible subset. We can assume that $`M`$ is a subspace of some $`𝐑^N`$, and all functions $`f_i`$ are smooth on a smooth dense open submanifold $`M_0M`$. ###### Proposition 9 Functional $`l`$ from the definition above is well-defined (i.e. the integral absolutely converges). Proof. We can use a sequence of functions $`(f_0,f_1,\mathrm{},f_k)`$ in order to map $`M`$ to $`𝐑^{k+1}`$. Then we obtain a singular chain $`\gamma C_k(𝐑^{k+1})`$. The support of $`\gamma `$ is a compact constructible subset of dimension $`k`$. Moreover, the volume of $`\gamma `$ with respect to the induced metric is finite. It follows that $`_\gamma \omega `$ absolutely converges for an arbitrary smooth differential $`k`$-form $`\omega `$. In particular converges the integral in question.$`\mathrm{}`$ It is easy to check that the differential $`d:\mathrm{\Omega }_{X,min}^k\mathrm{\Omega }_{X,min}^{k+1}`$ is well-defined (as the adjoint to the boundary operator on the chains). Moreover the following Stokes formula holds. ###### Lemma 9 In the previous notation we have: $$(1,f_0,\mathrm{},f_k)(M,or,\varphi )=(f_0,\mathrm{},f_k)(M,or|_M,\varphi |_M)$$ Hence we obtain a complex of sheaves $`\mathrm{\Omega }_{X,min}^k`$ for every $`k0`$. Our (first approximation to) $`k`$-forms are sections of these sheaves. One can take pull-backs of forms. Moreover, we can introduce wedge product of forms. Indeed the space of $`k`$-forms $`\mathrm{\Omega }_{min}^k(X)`$ is naturally embedded into the direct limit of spaces of smooth $`k`$-forms $`\mathrm{\Omega }_C^{\mathrm{}}^k(X_\alpha )`$. Here $`X_\alpha X`$ are open dense locally constructible subspaces, $`X_\alpha =_iX_{i,\alpha }`$, where $`X_{i,\alpha }`$ is a PA-manifold. Moreover, it is assumed that we have fixed an embedding $`X_{i,\alpha }𝐑^{N_{i,\alpha }}`$ identifying $`X_{i,\alpha }`$ with a $`C^{\mathrm{}}`$-submanifolds in $`𝐑^{N_{i,\alpha }}`$. Using the embeddings we define $`\mathrm{\Omega }_C^{\mathrm{}}^k(X_\alpha )`$. Therefore the wedge product given by the formula $`(f_0,\mathrm{},f_k)(g_0,\mathrm{},g_n)=(f_0g_0,f_1,\mathrm{},f_k,g_1,\mathrm{},g_n)`$ is well-defined. The above-defined forms are sections of soft sheaves because one can use piecewise algebraic functions to produce a partition of unity, so the standard proofs work. One can show that Mayer-Vietoris sequence degenerates as in the case with chains above: if $`X=Y_1Y_2`$ is a union of locally-closed piece-wise algebraic subspaces then one has a short exact sequence $$0\mathrm{\Omega }_{min}^k(X)\mathrm{\Omega }_{min}^k(Y_1)\mathrm{\Omega }_{min}^k(Y_2)\mathrm{\Omega }_{min}^k(Y_1Y_2)0$$ In the next subsection we will extend algebras $`\mathrm{\Omega }_{min}^{}(X)`$ adding push-forwards of such forms. The following example illustrates one of the reasons for that. ###### Example 1 Let $`Y=[0,1]\times [0,1]`$, $`X=[0,1]`$ and $`f:YX`$ be the map $`(x,y)t=xy`$. Then, outside of $`t=1`$, we have a bundle in the category $`𝒫`$ (i.e. base, fibers and total space are objects, projection is a morphism in $`𝒫`$). Take the form $`\omega =xdy`$. Then $`f_{}\omega =t(logt)`$ is a continuous function, but does not belong to $`𝒪(X)`$. ### 8.4 Full algebra of forms We are going to define a dg-algebra $`\mathrm{\Omega }_{PA}^{}(X)`$ which contains $`\mathrm{\Omega }_{min}^{}(X)`$, satisfies Poincare lemma, etc. Elements of $`\mathrm{\Omega }_{PA}^{}(X)`$ will be called PA-forms, although their coefficients are not PA-functions (see the previous Example). It is interesting that only $`0`$-forms will be forms with continuous coefficients. Higher order forms can have coefficients in $`L_1^{loc}(X)`$. But they are still closed under the wedge product. We start with some preliminaries. Let $`f:YX`$ be a proper PA-map of PA-spaces. For simplicity we will give the definition below in the case when $`X`$ is compact. ###### Definition 22 Continuous family of PA $`k`$-cycles (PA-family of cycles for short) is defined as an element of the abelian group $`C_k(YX)`$ described below. The group $`C_k(YX)`$ consists of maps $`\mathrm{\Phi }:XC_k(X)`$ such that: a) For any $`xX`$ the set $`Supp\mathrm{\Phi }(x)`$ belongs to $`f^1(x)`$. b) The set $`Z=_{xX}Supp\mathrm{\Phi }(x)`$ is constructible with the compact closure $`\overline{Z}Y`$. c) There exists a dense constructible $`Z_0\overline{Z}`$ such that for any $`xX`$ the intersection $`Z_0f^1(x)`$ is a PA-manifold without boundary, dense in $`\overline{Z}f^1(x)`$. Moreover, the chain $`\mathrm{\Phi }(x)`$ is obtained from some orientation of $`Z_0f^1(x)`$ and locally constant multiplicity map $`Z_0f^1(x)𝐙`$ as in the second description of PA-chains. d) For any $`xX`$ and $`zf^1(x)Z_0`$, and any sequence $`(x_i),x_ix`$ the multiplicity at the point $`z`$ is the “natural limit” of the multiplicities of $`f^1(x_i)Z_0`$. In the example below $`X`$ not compact. It easy to make necessary modifications of our definition, so it will be valid in the non-compact case. In that case we get a sheaf on $`X`$. ###### Example 2 Let $`Y=𝐑^2,X=𝐑`$ and $`f:YX`$ is the projection $`(x_1,x_2)x_1`$. Let us define $`\mathrm{\Phi }(x_1)`$ to be equal $`\{x_1\}+\{x_1\}`$ for $`x_1>0`$ and equal to $`2\{0\}`$ otherwise. Then this map gives an element of $`C_0(YX)`$. We will briefly describe some properties of PA-families of cycles below. 1) PA-families $`C_{}(YX)=\{C_k(YX)\}_{k0}`$ form a complex of sheaves (abelian groups in the case when $`X`$ is compact). 2) There are natural operations $`C_k(YX)C_l(X)C_{k+l}(Y)`$. Informally we will denote them by $`(\gamma ,\alpha )\gamma \times \alpha `$. 3) For any Cartesian square $$\begin{array}{cc}Y^{}\hfill & \hfill Y\\ \hfill & \hfill \\ X^{}\hfill & \hfill X\end{array}$$ one has the natural morphisms $$C_k(YX)C_k(Y^{}X^{})$$ and $$C_k(YX)C_l(X^{}X)C_{k+l}(Y^{}X).$$ Now we are ready to define piecewise PA-forms. ###### Definition 23 PA-forms of degree $`k`$ are sections of soft sheaves which are locally given by functionals $`l:C_k(X)𝐑`$ such that $$l(\alpha )=_{\gamma \times \alpha }\omega $$ where $`\gamma C_l(YX),l0`$ for some proper $`YX`$, and $`\omega \mathrm{\Omega }^{k+l}(Y)_{min}`$. We will denote the space of PA-forms on $`X`$ by $`\mathrm{\Omega }_{PA}^{}(X)`$. Similarly to the case of $`\mathrm{\Omega }_{min}^{}(X)`$ one can see that at the “generic point” the space $`\mathrm{\Omega }_{PA}^{}(X)`$ can be naturally embedded into the space of differential forms. In fact we obtain a soft sheaf of dg-algebras. It is closed under pushforwards $`f_{}`$ where $`f:YX`$ is a locally trivial bundle in the category $`𝒫`$ with fibers which are compact oriented PA-manifolds. The latter fact can be generalized further. One can consider a family $`f:YX`$ where $`Y`$ and $`X`$ are oriented compact PA-manifolds, and all fibers of $`f`$ have the same dimension $`dim(Y)dim(X)`$. Then one get a continuous family of cycles (“fundamental cycle” of $`f^1(x),xX`$) over $`X`$. This gives a pushforward of PA-forms for certain maps (“flat morphisms”) which are not necessarily fibrations. Notice that the Poincare lemma holds for piecewise forms. One can imitate the usual proof based on the integration over rays $`(x,t)xt`$, where $`xU𝐑^n`$, $`U`$ is a convex domain containing $`0𝐑^n`$, and $`t[0,1]`$. REFERENCES \[BD\] A. Beilinson, V. Drinfeld, Chiral algebras, preprint (1995). \[B\] R. Borcherds, Vertex algebras, q-alg/9706008. \[BV\] J. Boardman, R. Vogt, Homotopy invariant algebraic structures on topological spaces, Lect. Notes Math. 347 (1973). \[De\] P. Deligne, letter to Drinfeld (1990) \[Dr\] V. Drinfeld, On quasi-triangular quasi-Hopf algebras and a group closely related with $`Gal(\overline{𝐐}/𝐐)`$, Leningrad Math. J vol. 2 (1991), 829-860. \[FM\] W. Fulton, R. MacPherson, A compactification of configuration spaces, Annals Math., 139(1994), 183-225. \[GeJ\] E. Getzler, J. Jones, Operads, homotopy algebra, and iterated integrals for double loop spaces. Preprint 1994, hep-th/9403055. \[GeKa\] E. Getzler, M. Kapranov, Cyclic operads and cyclic homology. In: Geometry, Topology and Physics for Raul Bott. International Press, 1995, p. 167-201. \[GiKa\] V. Ginzburg, M. Kapranov, Koszul duality for operads, Duke Math. J. 76 (1994), 203-272. \[KM\] M. Kontsevich, Yu. Manin, Gromov-Witten classes, quantum cohomology, and enumerative geometry, Comm. Math. Phys., 164:3 (1994), 525-562. \[Ko1\] M. Kontsevich, Deformation quantization of Poisson manifolds,I. Preprint q-alg/9709040. \[Ko2\] M. Kontsevich, Rozansky-Witten invariants via formal geometry. Preprint q-alg/9704009. \[Ko3\] M.Kontsevich, Operads and motives in deformation quantization. Letters in Math.Phys., 48:1 (1999). Preprint math.QA/9904055 \[M\] M. Markl, Cotangent cohomology of a category and deformations. J. Pure Appl. Alg., 113:2 (1996), 195-218. \[M2\] M. Markl, Homotopy Algebras are Homotopy Algebras. Preprint math.AT/9907138 \[Ma\] J. May, Infinite loop space theory, Bull. AMS, 83:4 (1977), 456-494. \[MS\] J. McClure, J. Smith, A solution of Deligne’s conjecture, math.QA/9910126. \[So\] Y. Soibelman, Meromorphic tensor categories, q-alg/9709030. To appear in Contemporary Math. \[St\] K. Strebel, Quadratic differentials. Berlin: Springer-Verlag, 1984. \[T\] D. Tamarkin, Formality of chain operad of small squares, math.QA/9809164. \[V\] A. Voronov, Homotopy Gerstenhaber algebras, math.QA/9908040 Authors addresses: I.H.E.S., 35 route de Chartres, Bures-sur-Yvette 91440, France (M.K. and Y.S.), and Department of Mathematics, Kansas State University, Manhattan, KS 66506, USA (Y.S.). emails: maxim@ihes.fr, soibel@ihes.fr
warning/0001/hep-ph0001231.html
ar5iv
text
# Resonant Amplification of Gauge Fields in Expanding Universe ## I Introduction In these last years the concept of parametric resonance (henceforth PR) has played a fundamental role in cosmology for the development of the theory of preheating , the explosive decay of the inflaton after inflation. There have been investigations of several types of fields coupled to the inflaton: scalars , fermions , gravitational waves , scalar gravitational perturbations , gravitinos . Systems which includes nearly conformal invariant fields can be efficiently amplified in an expanding universe . This occurs because the Robertson-Walker metric is conformally related to the flat metric, and therefore the equations of motion can be mapped into similar ones in Minkowski space-time. In these systems the resonance can be efficient without resorting to large couplings (it is sufficient to have $`q𝒪(1)`$, where $`q`$ is the amplitude of the driving term). The resonance is not weakened by the expansion of the universe and it persists until nonlinear effects will shut off it . It is interesting to investigate the effect of PR on fields which are conformally coupled to gravity. For these fields particle production due to the expansion of the universe is absent . Therefore PR can be an alternative and efficient way to amplify fluctuations of these fields. Examples of conformally invariant fields are massless vector fields and massless fermions . In the case of fermion the Pauli blocking is present, but this does not prevent the occurring of novel interesting effects . For a massless vector field an exponential Bose enhancement occurs. In this Letter we investigate the evolution of gauge field coupled to a charged scalar field and to an axion. ## II Scalar Electrodynamics The Lagrangian density for scalar electrodynamics is given by $$=\frac{1}{16\pi }F_{\mu \nu }F^{\mu \nu }(D_\mu \mathrm{\Phi })^{}(D^\mu \mathrm{\Phi })V(\mathrm{\Phi }^{}\mathrm{\Phi })$$ (1) where $`D_\mu =_\mu ieA_\mu `$ is the gauge and metric covariant derivative, $`A_\mu `$ is the gauge potential and $`F_{\mu \nu }`$ is the antysimmetric field tensor, defined as $`F_{\mu \nu }=_\mu A_\nu _\nu A_\mu `$. From this Lagrangian the equations of motion are: $`\mathrm{}\mathrm{\Phi }+2ieA_\mu ^\mu \mathrm{\Phi }+ie\mathrm{\Phi }_\mu A^\mu +{\displaystyle \frac{V}{\mathrm{\Phi }^{}}}+e^2A_\mu A^\mu \mathrm{\Phi }=0`$ (2) $`_\nu F^{\nu \mu }=4\pi j^\mu +8\pi e^2A^\mu |\mathrm{\Phi }|^2`$ (3) where $`j_\mu =ie(\mathrm{\Phi }_\mu \mathrm{\Phi }^{}\mathrm{\Phi }^{}_\mu \mathrm{\Phi })`$. We shall consider inhomogeneous linearized fluctuations around homogeneous quantities $`f(t,𝐱)=\overline{f}(t)+\delta f(t,𝐱)`$ (where $`f`$ is any field) in the Robertson-Walker metric $`ds^2=dt^2+a^2(t)d𝐱^2`$. We shall study the evolution of the gauge field in the Coulomb gauge $`𝐀=0`$. This gauge singles simply out the dynamical degrees of freedom for the problem (1) and its close relation to the gauge invariant formalism has been pointed out in . By writing the homogeneous condensate as $`\overline{\mathrm{\Phi }}(t)=e^{i\mathrm{\Theta }(t)}\rho (t)/\sqrt{2}`$ one has at the homogeneous level: $$\ddot{\rho }+3H\dot{\rho }+\frac{V}{\rho }+e^2\frac{|\overline{𝐀}|^2}{a^2}\rho =0$$ (4) plus an equation of motion for $`\overline{𝐀}(t)`$ and the constraint $`\overline{A}_0=\dot{\mathrm{\Theta }}/e`$. We shall consider fluctuations $`\delta A_\mu (t,𝐱)`$ around a background $`\overline{A}(t)`$ that is chosen to be zero: $`\delta \ddot{𝐀}_T+H\delta \dot{𝐀}_T{\displaystyle \frac{^2}{a^2}}\delta 𝐀_T+4\pi e^2\rho ^2\delta 𝐀_T=4\pi \delta 𝐣_T`$ (5) $`{\displaystyle \frac{^2}{a^2}}\delta A_0=4\pi \delta j_04\pi e^2\rho ^2\delta A_0,`$ (6) where $`T`$ denotes the transverse part of a vector and $`\delta 𝐣_T`$ is a source term which can be different from zero only if one consider quantum or statistical correlation of the currents because of the symmetry of the space-time background. The constraint equation (6) in Fourier space is $$\left[\frac{k^2}{a^2}+4\pi e^2\rho ^2\right]\delta A_{0k}=4\pi \delta j_{0k}$$ (7) with $`\delta j_{0k}=e(\rho \dot{\delta \varphi _{2k}}\dot{\rho }\delta \varphi _{2k})`$ where $`\delta \varphi =(\delta \varphi _1+i\delta \varphi _2)/\sqrt{2}`$. We consider now the homogeneous part of the differential equation (5) in Fourier space: $$\delta 𝐀_{Tk}^{\prime \prime }+\omega _{Tk}^2\delta 𝐀_{Tk}=0$$ (8) where $`\omega _{Tk}^2=k^2+4\pi e^2a^2\rho ^2`$ and the prime denotes the derivative with respect the conformal time $`\eta `$ ($`d\eta =dt/a`$). The general solution $`y`$ of the inhomogeneous differential equation (5) can be obtained through the Green function method : $$𝐲=𝐲_h+4\pi e^2^\eta \left[\frac{y_1(\eta ^{})y_2(\eta )y_1(\eta )y_2(\eta ^{})}{W}\right]a^2\delta \stackrel{~}{𝐣}_{Tk}d\eta $$ (9) where $`y_h`$ is the homogeneous solution, $`y_1`$, $`y_2`$ are the two linear independent homogeneous solutions and $`W`$ is their wronskian. Equation (8) describes a harmonic oscillator with time dependent frequency: during the oscillation of the complex scalar field Eq. (8) can be reduced to a Mathieu-like equation . The solutions to this type of equation show an exponential instability $`e^{\mu _k\eta }`$, for some interval of frequencies, called resonance bands. There is a known correspondence at the level of equations of motion between Eq. (8) and the equation of a massless real scalar field $`\chi `$ (rescaled by its conformal weigth) interacting with a real scalar field $`\varphi `$ by a term $`g^2\varphi ^2\chi ^2`$: $$(a\chi )_k^{\prime \prime }+\left[k^2+g^2a^2\varphi ^2+(\xi \frac{1}{6})a^2R\right](a\chi )_k=0,$$ (10) where $`R=6a^{\prime \prime }/a^3`$ is the Ricci curvature. In the case of conformal coupling ($`\xi =1/6`$) this correspondence is exact, while for a quartic potential for $`\varphi `$ (or $`\mathrm{\Phi }`$) this correspondence is nearly exact since $`R0`$ . The behaviour of parametric resonance depends strongly on the time behaviour of the homogeneous scalar field, which on turn depends on the form of the potential $`V(\mathrm{\Phi })`$. In the following we consider power law potentials as: $$V=\lambda _n(\mathrm{\Phi }^{}\mathrm{\Phi })^{2n}$$ (11) and we briefly illustrate the different behaviours depending on the parameters $`n`$ and $`\lambda _n`$, restricting ourselves to the case in which the scalar field is the dominant component of energy density in the universe . For a quadratic potential ($`n=1`$) the driving term $`a^2\rho ^2`$ in Eq. (8) decays as $`\eta ^2`$. PR is efficient for $`4\pi e^2\rho ^2\lambda _1`$: the resonance is stochastic and broad, and the largest Floquet exponent $`\mu _k`$ occurs for small $`k`$ fluctuations . In the conformally invariant quartic case ($`n=2`$) the oscillations of $`\stackrel{~}{\rho }a\rho `$ are given by an elliptic cosine and Eq. (8) reads as an equation in Minkowski space-time (technically speaking it is a Lame equation). The resonance structure and the relative Floquet exponents $`\mu _k`$ depend non-monotonically by the ratio $`e^2/\lambda _2`$ : long wavelengths $`k^2\lambda _2\stackrel{~}{\rho }_0^2`$ are resonant for $`1/2\pi <e^2/\lambda _2<3/2\pi `$, as well for other values , where $`\stackrel{~}{\rho }_0`$ is the initial amplitude for $`\stackrel{~}{\rho }`$. For $`n3`$ the driving term $`a^2\rho ^2`$ oscillates and grows in time as $`\eta ^{\frac{2(n2)}{2n1}}`$. In this case the resonance for $`A_{Tk}`$ (as well for $`a\chi `$) is very efficient . The case of a simmetry breaking potential $$V=m^2\mathrm{\Phi }^{}\mathrm{\Phi }+\lambda (\mathrm{\Phi }^{}\mathrm{\Phi })^2$$ (12) with $`m^2<0`$, is definitively the closest to the electroweak phase transition: $`\mathrm{\Phi }`$ would play the role of the Higgs field, and $`\rho `$ rolls down from the zero value (false vacuum) towards $`\rho _{min}=\pm \sqrt{m^2/\lambda }`$ (true vacuum). Let us note that the most important feature in the case of symmetry breaking in this toy model of scalar electrodynamics is the Higgs mechanism: the gauge field gets an effective mass proportional to the value of the field in the broken phase $`\sqrt{m^2/\lambda }`$. Besides the unpleasant feature of considering a massive photon, this fact could completely inhibit the resonance in an expanding universe or narrow the resonance shifting it to scales proportional to $`\sqrt{m^2/\lambda }`$. However, in the realistic electroweak theory the photon does not acquire a mass: the symmetry breaking mechanism gives masses to the $`W`$ and $`Z`$ bosons, leaving the photon massless. This is an important condition for the efficiency of parametric resonance in expanding universe. We note that in this case the driving term in Eq. (8) is given by the $`W`$ condensate . Let us discuss for completeness also the behaviour of the scalar field fluctuations for the potential (12) without any constraints on the sign of $`m^2`$. The Fourier components of the real part of the rescaled field fluctuation $`\stackrel{~}{\delta \varphi _1}a\delta \varphi _1`$ satisfy the following equation: $$\stackrel{~}{\delta \varphi }_{1k}^{\prime \prime }+\left[\omega _{1k}^2\frac{a^{\prime \prime }}{a}\right]\stackrel{~}{\delta \varphi }_{1k}=0$$ (13) where $`\omega _{1k}^2=k^2+a^2m^2+3\lambda a^2\rho ^2`$. The Fourier components of the imaginary part of the rescaled field fluctuation $`\stackrel{~}{\delta \varphi _2}a\delta \varphi _2`$ satisfy $`\stackrel{~}{\delta \varphi }_{2k}^{\prime \prime }+\left[\omega _{2k}^2{\displaystyle \frac{a^{\prime \prime }}{a}}\right]\stackrel{~}{\delta \varphi }_{2k}=`$ (14) $`ea^2(2\delta A_{0k}\rho ^{}+\rho \delta A_{0k}^{}+3aH\rho \delta A_{0k})`$ (15) where $`\omega _{2k}^2=k^2+m^2a^2+\lambda a^2\rho ^2`$. By using the constraint equation (7) and introducing $`q_k=\stackrel{~}{\delta \varphi }_{2k}/\omega _T`$ one obtains $`q_k^{\prime \prime }+[\omega _{k\mathrm{\hspace{0.17em}2}}^2+\omega _T^2k^2{\displaystyle \frac{a^{\prime \prime }}{a}}+{\displaystyle \frac{8\pi e^2a^2}{\omega _T^2}}(\rho ^{}+aH\rho )^2`$ (16) $`+{\displaystyle \frac{\omega _T^{\prime \prime }}{\omega _T}}2{\displaystyle \frac{\omega _T^2}{\omega _T^2}}]q_k=0.`$ (17) We note that the gauge coupling affects the resonance structure of the scalar field: while $`\omega _{1k}^2`$ is the driving term for $`\stackrel{~}{\delta \varphi }_{1k}`$, $`\omega _{2k}^2`$ does not determine the resonance bands for $`\stackrel{~}{\delta \varphi }_{2k}`$, as would do if the charged scalar field were uncoupled to a gauge field . ## III Axion Electrodynamics We consider the effect of parametric resonance for another type of coupling described by: $$=\frac{1}{16\pi }F_{\mu \nu }F^{\mu \nu }\frac{1}{2}_\mu \varphi ^\mu \varphi V(\varphi )\frac{g}{4}\varphi F_{\mu \nu }\stackrel{~}{F}^{\mu \nu }$$ (18) The neutral scalar field $`\varphi `$ can represent an axion in (in such a case the vacuum angle $`\theta `$ is defined to be $`\theta =\varphi /f`$, where $`f`$ is the Peccei-Quinn simmetry scale) or a general pseudo-Goldstone boson in . By considering again the scalar field as $`\varphi (t,𝐱)=\varphi (t)+\delta \varphi (t,𝐱)`$ one has at the homogeneous level: $$\ddot{\varphi }+3H\dot{\varphi }+\frac{V}{\varphi }=0$$ (19) and for the transverse degrees of freedom $$\delta \ddot{𝐀}_T+H\delta \dot{𝐀}_T\frac{^2}{a^2}\delta 𝐀_T+4\pi g\dot{\varphi }\frac{}{a}\times \delta 𝐀_T=0.$$ (20) In this case it is more convenient to consider the circular polarized Fourier component perpendicular to $`𝐤`$, $`𝐀_{\pm k}`$, whose equation of motion is $$\delta 𝐀_{\pm k}^{\prime \prime }+(k^2\pm 4\pi g\varphi ^{}k)\delta 𝐀_{\pm k}=0$$ (21) In the regime of coherent oscillation of the scalar field one has again a Mathieu-like equation where the driving term is proportional to the time derivative of the scalar field and goes as $`\eta ^1`$ for a power law potential analogous to Eq. (11) (always under the assumption that the scalar field dominates the energy density of the universe ). The efficiency of the resonance depends on the value of the adimensional quantity $`g\varphi `$. The analysis of Eq. (21), describing the axion decay into photons, was probably one of the first application of parametric resonance in cosmology , and it has been recently reanalyzed in . By considering $`\varphi `$ as the QCD axion, the smallness of $`gf10^4`$ prevents the efficiency of the resonance in the case of a potential with temperature-dependent mass due to QCD effects . In general the resonance is on $`k4\pi g\varphi \omega `$, where $`\omega `$ is the oscillation frequency of $`\varphi `$. However, for $`V=\lambda \varphi ^4/4`$ and $`4\pi gf=1`$ we observe a linear growth of $`A_\pm `$ for $`k/\omega 4\pi gf`$ by numerically integrating Eq. (19) and (21). Such growth is slower than the quasi-exponential amplification of the field fluctuations $`\delta \varphi `$ on smaller scales . Besides the PR effect, one of the two circular polarization exhibit also a negative effective mass (depending on the sign of $`\dot{\varphi }`$). This feature could be interesting before the stage during which the field coherently oscillates, when $`\dot{\varphi }`$ has a fixed sign. ## IV Plasma effects We now comment on effects due to the presence of other charged particles <sup>*</sup><sup>*</sup>*We are referring to charged fields which are uncoupled from the scalar field which oscillates coherently. These fields produce currents $`J_\mu `$ which are incoherent and are coupled in a $`J_\mu A^\mu `$ to the gauge field.. These effects can be important since the universe is considered a good conductor after reheating until decoupling. For wavelengths larger than the collision length one could use the Ohm relation for the additional current $`𝐣_{\mathrm{add}}=\sigma 𝐄`$. This would add to Eqs. (8) and (20) a damping term: $$\delta 𝐀_k^{\prime \prime }+4\pi (a\sigma )\delta 𝐀_k^{}+\omega _k^2(\eta )\delta 𝐀_k=0,$$ (22) where we have denoted collectively the transverse component as $`\delta 𝐀`$ and with $`\omega _k^2`$ their time dependent frequency. The time behaviour of conductivity is considered in the literature to be $`\sigma T1/a`$ and so the term multiplying the first derivative in Eq. (22) is constant: in this way the resonance on long wavelengths could be washed out if $`a\sigma `$ is larger than the Floquet exponent. For wavelengths smaller than the collision length charge separation should be taken into account and a term qualitatively similar to the classical plasma frequency would enter as: $$\omega _k^2(\eta )\omega _k^2(\eta )+4\pi e^2\frac{n(\eta )}{m}$$ (23) where $`n(\eta )`$ and $`m`$ are respectively the number density and the mass of the charged particles. This term would play the role of an effective mass, but which would decay as $`a(t)^3`$ (which is more rapid than the decay of the driving term considered in this paper), leaving open the possibility of resonance, in particular for large coupling costants. Let us observe that these qualitative estimates for conductivity and plasma frequency are obtained by thermodynamic considerations. Even in the case of a preexisting thermal equilibrium the resonance would amplify the gauge fluctuations. Infact a thermal mass $`T^2`$ for the gauge fiels will not be able to prevent the resonance and it has the effect of shifting the resonance on smaller scales ($`k^2`$ must be replaced with $`k^2+g_{\mathrm{eff}}T^2a^2`$). There is an epoch where the the resonance could proceed just as worked out in the vacuum case: in preheating after inflation, when it is assumed that the ”creation” of charged particles is contemporaneous to the amplification of gauge fluctuations. This amplification will last until rescattering and backreaction set in to shut off the resonance , and terms such as conductivity and plasma frequency will appear in this out-of-equilibrium process by a self-consistent study of the problem (although their time-dependence will be different from the corresponding thermal equilibrium quantities). In this case the constraints on the coupling costants $`e^2`$ and $`g`$ come only from requirement to protect the potential of the scalar field from radiative corrections: this argument would give the usual $`e^210^6`$ for the scalar electrodynamic case. If one requires that the charged inflaton gives the right amount of e-folds and CMB fluctuations ($`m10^6M_{pl}`$ for a massive inflaton, $`\lambda 10^{13}`$ for a self-interacting inflaton), then a broad resonance regime would be allowed for the gauge field. For the axion electrodynamic case, the same argument leads to $`gf10^4`$, opening a possible new efficient channel for the decay of the scalar field $`\varphi `$ into gauge fluctuations. ## V Primordial Magnetic Field by Parametric Resonance? Even if the conformal property of gauge fields in cosmological backgrounds is useful for PR, it is the main reason why gauge fluctuations are not amplifield by the expansion of the universe , in contrast to what happens for minimally coupled scalar fields. Therefor the conformal invariance of gauge fields is the main issue for the explanation of primordial magnetic fields within the inflationary paradigm . PR could be interesting for the generation of primordial magnetic fields for two reasons. First, it is intringuing to note how the exponential growth due to parametric resonance is of the same type as the dynamo effect , one of the astrophysical processes postulated to explain the observational evidence for galactic magnetic fields. As a second point, there are several examples in which significant amplification occurs on the maximum causal scale allowed by the problem (the coherence scale of the field). Since the scale of the magnetic seeds is always a crucial issue for a model which aims to explain their origin, this feature of PR is very interesting. As a straightforward result of section II preheating could provide an extra growth in the amplitude of gauge fluctuations which are produced during inflation : for parameters in which long wavelength gauge fluctuations are amplified, this occurs directly on observable scales because of the coherence of the inflaton on the particle horizon scale. We illustrate how the mechanism could work in a simple scalar electrodynamic toy model where the charged scalar field is a massless self-interacting inflaton with $`2\pi e^2=\lambda `$ During inflation, assuming $`\mathrm{\Phi }`$ in slow rollover, a solution for Eq. (8) is: $$\delta 𝐀_{Tk}=(H\eta )^{1/2}\left(\frac{\pi }{4H}\right)^{1/2}H_\nu ^{(1)}(k\eta )$$ (24) where $`\eta =1/(Ha)`$ is the conformal time used to model a de Sitter era ($`\mathrm{}<\eta <\eta _0<0`$, where $`\eta _0`$ is some arbitrary time) and the index of the Hankel function is so defined $$\nu ^2=\frac{1}{4}4\pi e^2\frac{\rho ^2}{H^2}\frac{1}{4}\frac{3e^2}{2\lambda }\frac{M_{\mathrm{pl}}^2}{\rho ^2}$$ (25) where the last equality holds during slow-rollover and $`M_{\mathrm{pl}}^2`$ is the inverse of the Newton constant $`G`$. The solution (24) matches the adiabatic vacuum $`e^{ik\eta }/\sqrt{2k}`$ for $`\eta \mathrm{}`$. For $`\nu ^2>0`$ Eq. (24) leads to a slight shift of the initial vacuum for the long wavelength fluctuations ($`k\eta 0`$) of the gauge field: $$\delta 𝐀_{Tk}i\frac{\mathrm{\Gamma }(\nu )}{\sqrt{2\pi k}}\left(\frac{k\eta }{2}\right)^{\frac{1}{2}\nu }.$$ (26) By using the definition $`B_i=ϵ_{imn}_mA_n/a`$ one can relate the fluctuation of the magnetic field to the gauge field fluctuations $$|𝐁_k|^2=\frac{k^2|\delta 𝐀_{Tk}|^2}{a^4}$$ (27) Parametric resonance after inflation amplifies the gauge fluctuations in Eq. (26) and $`|𝐁_k|^2`$ in Eq. (27) exponentially up to a maximum factor $`10^{12}`$ . However, this last amplification is not sufficient to explain the value required by the observation. Indeed, the simple toy model of a self-interacting inflaton is very useful in order to follow gauge fluctuations at a linear order from inflation through preheating, but it misses a super-adiabatic amplification during inflation, which was investigated by Turner and Widrow . This amplification can occur due to an effective negative mass for the gauge field during inflation, which can be obtained by breaking the conformal invariance through a explicit coupling to the curvature, for instance . In string cosmology an effective negative mass for the gauge field is generated by the dilaton coupling . Instead, as we see from Eq. (26), long wavelength gauge fluctuations are slightly modified in amplitude and spectrum with respect to the initial adiabatic spectrum by the coupling with inflaton during slow-rollover. Indeed the positive mass generated by the coupling of the gauge field to the charged scalar field lead to a slight suppression. When slow-rollover is ended and $`\nu ^2<0`$ the gauge modes are suppressed. This phase is just a short transient, after which the resonant amplification starts immediately (roughly when $`\rho `$ first cross zero). The main point is the decay of $`B^2`$ as $`1/a^4`$ once a gauge fluctuation has left the Hubble radius during inflation, leading to a suppression which the following preheating phase cannot fill up . Now we quantify our previous statements and we estimate the order of magnitude of the amplitude for the magnetic field at decoupling. We shall consider the effect of preheating as a k-independent amplification factor $`\mathrm{\Omega }^2`$ which multiplies the spectrum at the end of inflation (26). We also neglect the finite time duration of the preheating phase and we identify the end of inflation with the beginning of the usual radiation dominated Friedmann era. At the beginning of the radiation dominated phase (denoted in the following by $`\eta ^{}`$) the spectrum of the magnetic field is: $$k^3|𝐁_k(\eta ^{})|^2=\mathrm{\Omega }^2\frac{k^5}{a^4(\eta ^{})}|\delta 𝐀_{Tk}(\eta ^{})|^2$$ (28) where we consider $`\delta 𝐀_{Tk}(\eta ^{})1/\sqrt{2k}`$, which has been obtained from Eq. (26) by setting $`\nu =1/2`$. By using the the relation $`a^4B^2=const.`$ for the magnetic field evolution in the radiation dominated era, the spectrum at the decoupling time $`\eta _{\mathrm{DEC}}`$ is: $$k^3|𝐁_k(\eta _{\mathrm{DEC}})|^2=k^3|𝐁_k(\eta ^{})|^2\frac{T_{\mathrm{DEC}}^4}{T_{\mathrm{REH}}^4}$$ (29) since the temperature $`T`$ scales as the inverse of the scale factor $`a`$. By using Eqs. (28) and (29) we estimate at time of decoupling a magnetic field of size $`|_{\mathrm{DEC}}(k)`$ $`=`$ $`(k^3|𝐁_k(\eta _{\mathrm{DEC}})|^2)^{1/2}`$ (30) $``$ $`{\displaystyle \frac{\mathrm{\Omega }}{\sqrt{2}}}(k\mathrm{kpc})^2{\displaystyle \frac{T_{\mathrm{DEC}}^2}{\mathrm{kpc}^2T_{\mathrm{NOW}}^2}}`$ (31) $``$ $`10^{42}\mathrm{Gauss}`$ (32) where it has been assumed $`a(\eta _{\mathrm{NOW}})=1`$, a comoving wavenumber $`k=0.1\mathrm{kpc}^1=0.64\times \mathrm{\hspace{0.17em}10}^{36}\mathrm{GeV}`$, $`\mathrm{\Omega }=10^6`$, $`T_{\mathrm{DEC}}/T_{\mathrm{NOW}}10^3`$, $`1\mathrm{Gauss}=3\times 10^{20}\mathrm{GeV}^2`$ . The estimate (32) is lower than the magnetic seed field required by observation, which is $`10^{34}`$G for a flat, low density universe . We note that $``$ is proportional to $`\mathrm{\Omega }`$, but not to $`T_{\mathrm{REH}}`$ or to the Hubble constant during inflation, as observed in . This is due to the the approximation $`\nu =1/2`$ which restablishes the conformal invariant spectrum in the estimate (32). ## VI Conclusions In this paper we have studied the possibility of an enhancement of gauge fields due to parametric resonance. For both the couplings to a scalar charged field and to an axion, the resonance can counteract the redshift due to the expansion of the universe and exponential growing modes appear. The resonance is generically on small scales in the axion case, while it can occur for scales up to the coherence scale of the scalar field in the scalar electrodynamic case. Therefore the latter scalar field coupling is more interesting for the problem of cosmological magnetic fields, in which the large scale observed seems a crucial issue. In the simplest toy model with a massless charged inflaton, a linear calculation based on preheating does not provide directly an amplitude in agreement with observations because the magnetic field fluctuations are redshifted during inflation. However, this latter effect may be circumvented in many ways. In more realistic inflationary models motivated by particle physics the gauge field can be sustained during inflation and then amplified during preheating. A sustain during inflation could be also realized by breaking conformal invariance with explicit coupling between the gauge field and the Riemann curvature in the lagrangian . Conformal invariance is also effectively broken by the inhomogeneities genarated during inflation (a second order effect in perturbations). Moreover, PR would also provide an additional growth in the generation mechanism based on the statistical correlation of the currents studied in . A final amplitude and spectrum of gauge field fluctuations generated by PR is however beyond the scope of the present paper. Infact, its calculation would include the analysis of nonlinear effects and a simple estimate based on matching techniques is inadequate. In models with interacting scalar fields, nonlinear effects as rescattering play an important role in the generation of the final power spectra. Indeed, such power spectra, as $`|\delta 𝐀_k|^2k^\alpha `$ with $`\alpha >1`$, would be very interesting for the problem of large scale magnetic fields, because steeper than (26) in the infrared region. Acknowledgments It is a pleasure to thank R. Brandenberger for discussions and many valuable comments on earlier versions of this draft, and S. Khlebnikov and G. Venturi for discussions and suggestions. We thank R. Balbinot, B. Bassett, K. Dimopoulos, K. Enqvist, G. Field, E. Lazzaro, A. Riotto, A. Starobinsky, K. Subramanian, O. Tornkvist, G. Veneziano for discussions and comments. F. F. was supported in part by the DOE under Grant no. DE-FG02-91ER40681 (Task B).
warning/0001/astro-ph0001309.html
ar5iv
text
# Weak lensing observations of the “dark” cluster MG2016+112 ## 1 Introduction The QSO lens MG2016$`+`$112 is a long-standing enigma. Spanning a few arcseconds in size, the system contains two QSO images (designated as components A and B, see Fig. 4) and an extended structure (designated C), all at redshift $`z=3.273`$ (Lawrence et al. Lawrence84 (1984), Schneider et al. Schneider85 (1985), Schneider86 (1986)). The extended structure appears as an arc in the infrared (Langston et al. Langston (1991); Lawrence et al. Lawrence93 (1993); Benitez et al. Benitez (2000), hereafter B99), but consists of four subcomponents in the radio, elongated along the arc (Garrett et al. Garrett (1996)). Roughly in the middle of the lens structure is a giant elliptical galaxy at redshift $`z=1.01`$ (designated D, Schneider et al. Schneider85 (1985)), and another galaxy of unknown redshift lies just outside the giant arc. Various lensing models have been applied to this system involving a single elliptical galaxy as lens, a giant elliptical galaxy plus a cluster core lens, a giant elliptical plus companion galaxy lens, and a double lens model involving two galaxies at different redshift each lensing parts of the system (Narasimha & Chitre Narasimha (1989); Langston et al. Langston (1991); Nair & Garrett Nair (1997); B99), but no model has successfully reproduced all of the observed structures, positions, and brightness ratios of the system. An observation with the ASCA satellite revealed a strong X-ray source, designated AXJ2019$`+`$112, towards MG2016$`+`$112 with an emission line at 3.5 keV which could correspond to FeXXV at $`z=0.92`$ or FeXXVI at $`z=1.00`$, the redshift of the lensing galaxy (Hattori et al. Hattori (1997), hereafter H97). The X-ray spectra results in an 8.6 keV gas temperature for a $`z=1.0`$ source, implying a massive cluster ($`M2\times 10^{14}h^1M_{}`$ at $`r=250h^1`$ kpc) at the redshift of the lens. Infrared and optical observations of the field (Lawrence et al. Lawrence93 (1993), Langston et al. Langston (1991), Schneider et al. Schneider85 (1985)) revealed no excess of galaxies around the giant elliptical lens, the supposed brightest cluster galaxy (BCG), as would normally be found in a cluster. This led to the suggestion of a “dark cluster” associated with the lens, in which a cluster-sized dark matter and hot gas overdensity exists with few optically bright galaxies ($`M/L_V>2000M_{}/L_{}`$). These observations were even more puzzling as the observed iron line in the X-ray spectra indicates a near-solar metallicity for the cluster, which implies a long history of star formation in the region (H97). A handful of galaxies at $`z=1.01`$ have been found spectroscopically within a few arcminutes of the lens (Soucail et al. Soucail (2000)), but these are not concentrated strongly around the lens. Recently, B99 have found a slight overdensity of red galaxies using deep $`V`$, $`I`$, and $`K_s`$ imaging in a region 200 kpc ($``$47″) in radius around the QSO lens. Assuming that the galaxies they identified were the most luminous galaxies in a Schechter cluster luminosity function, they calculated that the total cluster luminosity was roughly ten times that of the BCG. Thus the proposed cluster has a mass-to-light ratio of only a few hundred solar, typical of what is seen in other massive high-redshift clusters (Luppino & Kaiser Luppino (1997), Clowe et al. Clowe98 (1998)). The observed galaxies are red in both $`VI`$ and $`IK_s`$, suggesting a old stellar population and consistent with what is observed in other high-redshift clusters (Trentham & Mobasher Trentham98 (1998), van Dokkum et al. Dokkum (1999)). These observations, however, are based on only nine galaxies selected by two colors and confirmed by a third. In this paper we present deep optical and infrared images in $`V`$, $`R`$, $`I`$, and $`K`$ of the MG2016$`+`$112 field with the goals of testing the B99 selected galaxies with additional colors, detecting any additional, fainter cluster galaxies, and measuring the weak lensing signal in the field to confirm the presence and measure the mass of any cluster. In §2 we discuss the images taken and the image reduction process. We discuss the colors and magnitudes we measure for the proposed cluster galaxies of B99 as well as others in the field in §3. Section 4 contains a weak lensing analysis of the field. We summarize and discuss our major conclusions in §5. Throughout this paper, unless stated otherwise, we assume $`\mathrm{\Omega }_0=1`$, $`\mathrm{\Lambda }=0`$, and $`H_0=`$100 km/s/Mpc. ## 2 Observations and Data Reduction Deep $`I`$, $`R`$, and $`V`$-band imaging was performed on a field centered on MG2016$`+`$112 using LRIS (Oke et al. Oke (1995)) in direct imaging mode at the Keck II telescope on the nights of 1998 July 22-23. The resulting images cover a 6′$`\times `$7$`\stackrel{}{.}`$5 area in $`I`$ and $`V`$ and 7$`\stackrel{}{.}`$5$`\times `$7$`\stackrel{}{.}`$5 area in $`R`$ with 0$`\stackrel{}{.}`$215 pixels. The total integration times are 1500s in $`I`$ and $`V`$ and 5400s in $`R`$, with seeing of 0$`\stackrel{}{.}`$50, 0$`\stackrel{}{.}`$56, and 0$`\stackrel{}{.}`$54 FWHM in $`I`$, $`R`$, and $`V`$ respectively. The individual images in each filter were de-biased using the overscan strip and flattened with dome flats. The $`I`$-band images were then re-flattened with a medianed night-sky flat from a (relatively) empty piece of sky taken during the same nights to remove fringing. The images were then remapped with a bi-cubic polynomial to correct for focal plane curvature. The polynomial for the mapping was determined by requiring that the positions of the stars in each remapped frame were consistent to within 0.1 pixels ($`0\stackrel{}{.}02`$, the estimated rms error in the centroiding algorithm) rms with each other and within the errors of the positions of the stars in a Digital Sky Survey image and a catalog of stellar positions from the United States Naval Observatory. The mappings were then checked using a second group of stars which were not used to generate the polynomial. The re-mapping was performed using linear interpolation and a triangle method to distribute the original pixel values onto the new map. The method preserved surface brightness and has been tested to ensure that in the case of a fractional pixel shift with no change in the shape or size of the pixel that the second moments of the objects in the image are not shifted in a systematic manner (although there is some noise added). The re-mapped images then had their sky subtracted by fitting a low order polynomial to the minima in the image. A final image was produced by averaging the sky-subtracted images with a sigma-clipping algorithm to remove cosmic rays. Photometry was determined by comparing the magnitudes of the bright but unsaturated stars with those in images of the field from the UH88<sup>′′</sup> telescope and Isaac Newton telescope which were calibrated from Landolt standards (Landolt Landolt (1992)). We assume a Galactic extinction $`A(B)=0.81\pm 0.12`$ (Schlegel et al. Schlegel (1998)) and derive $`A(V)0.61`$, $`A(R)0.46`$, $`A(I)0.29`$, and $`A(K)0.07`$, using the color conversions of Cardelli et al. (Cardelli (1989)). The 5$`\sigma `$ limiting magnitudes within a square 1$`\stackrel{}{.}`$0 aperture are $`I`$=25.1, $`R`$=27.1, and $`V`$=27.1. We also obtained deep $`K`$-band images using IRCAM3 at UKIRT on 1998 July 19-20. Each image is 1$`\stackrel{}{.}`$5 on a side with 0$`\stackrel{}{.}`$286 pixels. A 2$`\times `$2 mosaic of pointings resulted in a 3′$`\times `$3′ field with 9720s exposure time in each quadrant and an additional 3240s exposure time in the central 1$`\stackrel{}{.}`$5$`\times `$1$`\stackrel{}{.}`$5. Individual exposures were 120s long and dithered by 8$`\stackrel{}{.}`$0. The median seeing was about 0$`\stackrel{}{.}`$8. Flat-field images were constructed using median-filtered dithered object frames from the whole night and sky images were constructed using sets of twenty-six 120s exposures. The individual frames were then flat-fielded, sky-subtracted and combined using a clipping algorithm that rejected bad pixels. Instrumental magnitudes were computed from observations of UKIRT faint standards (Casali & Hawarden Casali (1992)), giving a photometric accuracy in the zero-point of better than 2%. The 5$`\sigma `$ limiting magnitude within a 1$`\stackrel{}{.}`$0 square aperture in the reduced image is $`K`$=21.3. ## 3 Galaxy Counts and Color-Selection The IMCAT (http://www.ifa.hawaii.edu/$``$kaiser/imcat) hfindpeaks detection algorithm was used to search for objects in the above $`VRIK`$ images. The objects were each inspected by eye in all images and spurious detections (due to stellar diffraction peaks, for example) were eliminated from the catalog of objects. Objects were identified as stars or galaxies on the basis of this inspection as well as using the measured half-light radii and central surface brightnesses compared to the total (large-aperture) magnitudes of the objects. Those objects which were classified as galaxies had their magnitudes remeasured using the IRAF apphot package, following a local sky subtraction. Uncertainties in these numbers are substantial at the faint end and come from uncertainties in the local sky due to scattered starlight, faint galaxy clustering, and Poisson noise. For example, objects at $`R>24`$ typically have uncertainties of at least 0.5 magnitudes in each color, measured by using apertures of various sizes and radii around the objects to determine the local sky value. Fig. 1 shows the galaxy number counts as a function of magnitude for this field in each filter. There is no obvious excess of galaxies above the background, as was seen in the rich cluster MS1054$``$03 at $`z=0.83`$ (Trentham & Mobasher Trentham98 (1998)), but the background uncertainty on these small scales is large. Thus, the background counts might be anomalously low, and a cluster at $`z=1`$ may still exist despite the lack of an excess. The most straightforward way to look for such a cluster is to (i) look for dense groups of galaxies very close to the lensing galaxy D, and (ii) inspect color-color and color-magnitude diagrams. These are presented in Fig. 2 and Fig. 3. We identified one group of seven faint objects all between 19 and 27 kpc from galaxy D in the $`R`$-band image. The brightest of these is $`R24`$ and the rest around $`R26`$ and all, except the brightest, were too faint to be detected in the other filters. The brightest has $`RI`$ and $`IK`$ colors of 1.57 and 2.81, which are somewhat bluer than those of galaxy D (1.81 and 3.83). From the colors it is unlikely that these would be the brightest galaxies in a higher redshift cluster. From the observed number density of $`24<R<26`$ galaxies in this field, the chances of projecting seven unrelated galaxies in such a small area is only 0.14%. However, with the exception of the brightest object, which is spatially extended and therefore, probably, a galaxy, we cannot determine if the objects are stars or galaxies. Two other objects also around $`R26`$ are located on the eastern side of galaxy D at roughly the same distance. Even assuming that these objects are galaxies and associated with galaxy D, the proposed 8 keV cluster with galaxy D as the brightest cluster galaxy would still have a low surface density of cluster galaxies near the BCG as compared to high-redshift clusters of similar mass (both MS1137$`+`$66 and RXJ1716$`+`$67, $``$6 keV clusters at $`z0.8`$, have roughly twice as many galaxies within 30 kpc of the BCG). We note, however, that as MG2016$`+`$112 was originally discovered as a medium-separation QSO lens, that this deficiency of bright cluster galaxies near the BCG may be a selection effect as a large number of massive galaxies would perturb the lensing caustics sufficiently to preclude a double image of the QSO from lensing by the BCG. From Fig. 2 and Fig. 3, one can find a number of red galaxies having approximately the correct $`IK`$ to be old stellar populations at $`z=1`$ (Kodama et al. Kodama (1998)) within 225$`h^1`$ kpc of galaxy D. Of these, eight are the red objects described by B99, including galaxy D, and three are not found in the B99 sample. One of the B99 objects, the eighth in their Table 1, we identified as a star. Our revised count of the total number of candidate members from color selection is thus eleven. There are only fifteen galaxies in the image outside the selected region with colors and magnitudes similar to the selected galaxies. This gives an overdensity around galaxy D of $`1160_{260}^{+440}`$% compared to the rest of the image, and the probability this is due to Poissonian fluctuations in the density is 3$`\times 10^{12}`$. If cluster members, as argued by B99, these are presumably responsible for generating the metals seen at X-ray wavelengths by H97. Spectroscopic redshifts would be needed to confirm this interpretation. We note, however, that these galaxies are up to a magnitude redder in $`RI`$ than would be predicted for passively-evolving ellipticals. This is probably due in part to photometric uncertainties, but could also be due to small amounts of internal extinction. Note that at $`z=1`$ the $`RI`$ color of a pure old stellar population is anomalously red since at this redshift the 4000Å break falls exactly between these two passbands, which in turn means that this color is particularly susceptible to being made significantly different by internal extinction. In addition, galaxies bluer than the ones discussed by B99 could in principle also be members that are undergoing current star formation. Again, spectroscopic information will be useful to investigate these possibilities further. ## 4 Weak Lensing Analysis ### 4.1 Techniques The first step in the weak lensing analysis of the field is to detect and measure the shapes of the background galaxies. This was done using the $`R`$-band image, which is of similar depth and resolution to other Keck fields which have successfully reconstructed the surface mass density of other high-redshift clusters (Clowe et al. Clowe98 (1998), Clowe00 (2000)). Most of this analysis was done using the IMCAT software package. Prior to using a detection algorithm on the image, we first subtracted small-scale background fluctuations in the image, caused primarily by stellar halos and scattered star light. These fluctuations were determined by smoothing an image of the local minima of the sky values with a 5″ Gaussian, and dividing by an image smoothed on the same scale which contained the mask of the pixels detected as minima. In addition to removing any sky variations, this routine typically removed extended wings of bright stars in the image as well as any large, low surface brightness objects which might have been present. Performing this step was necessary as the detection algorithm can get confused by rapid changes in the sky background. To detect the faint galaxies we used a hierarchical peak-finding algorithm which provided a position, luminosity, and size estimate for each object (Kaiser et al. Kaiser95 (1995), hereafter KSB). Local sky levels around each object were calculated after excluding all pixels within three detection radii of an object in the catalog. The luminosity was measured using a circular aperture equal to three times the smoothing radius at which the object achieved maximum significance. $`I`$ and $`V`$ magnitudes were measured using the same sky-subtraction technique and the same aperture radius. Noise peaks and merged objects were removed from the catalog by requiring a minimum signal-to-noise at maximum detection significance of 10, rejecting abnormally large and small objects, rejecting objects with extremely high ellipticities, and by visual inspection. The ellipticities of each object were measured using optical polarizations $`e_\alpha =\{I_{11}I_{22},2I_{12}\}/(I_{11}+I_{22})`$ formed from the quadrupole moments $`I_{ij}=d^2\mathrm{\Theta }W(\mathrm{\Theta })\mathrm{\Theta }_i\mathrm{\Theta }_jf(\mathrm{\Theta })`$ where $`f`$ is the flux density and $`W(\mathrm{\Theta })`$ is a Gaussian weighting function of a scale equal to the smoothing radius at which the object was detected with maximum significance (KSB). The next step in the lensing analysis was to remove the PSF anisotropies and smearing. Due to the large number of stars in the field (the final sample was 675 bright, unsaturated stars with no neighbors within 10″), we were able to fit the stellar ellipticities across the field with an arbitrarily high-order two-dimensional polynomial. We used third, fifth, and seventh order polynomials for the fits, and subtracted the fit values from all detected objects via the method in KSB. We then applied the boost factor discussed below to the stellar ellipticities, and used the boosted ellipticities to measure any false signal which will be induced in the data by the residual PSF anisotropies. As can be seen in Fig. 5, there is a small positive signal in the center of the fields; however the magnitude of this signal is small compared to the noise caused by the intrinsic ellipticity distribution of the background galaxies, and thus should not significantly affect any observed lensing signal. The amplitude of the false residual signal decreases somewhat with the increasing order of the polynomial fit to the ellipticities, as can be seen in Fig. 5, and thus we used the bi-septic polynomial fit to correct the ellipticities of the galaxies. This correction was done by subtracting the value of the fit at the position of each galaxy from the measured ellipticity, after scaling the fit value by $`P_{sm}/P_{sm}^{}`$. As defined in KSB (corrections in Hoekstra et al. Hoekstra (1998)), $`P_{sm}`$ is a quasi-tensor which describes how the ellipticity of an object changes under an applied anisotropy and the asterisk denotes the typical value for a star in the image. To correct for the reduction of the ellipticities from circular smearing by the PSF, we calculated a boost factor for each galaxy $`P_\gamma =P_{sh}P_{sm}(P_{sh}^{}/P_{sm}^{})`$ (Luppino & Kaiser Luppino (1997)), where $`P_{sh}`$ is a quasi-tensor describing how the ellipticity of an object changes with an applied shear (KSB), and $`P_{sm}`$ and the asterisk are defined as above. Due to the large anisotropy in the stars at the edges of the Keck field, we computed $`P_{sh}^{}`$ and $`P_{sm}^{}`$ only in the central region with a diameter of one thousand pixels, in which there was no significant anisotropy. As the individual values of $`P_\gamma `$ are quite noisy for faint galaxies, we fit $`P_\gamma `$ with a third order polynomial as a function of size and ellipticity. Because of noise and a tendency of the sky-subtraction algorithm to measure slightly too large a sky value (owing to the presence of many faint, unresolved objects), very faint galaxies tend to have artificially low, sometimes even negative, $`P_\gamma `$ values. We corrected for this by applying a cut to the fitted $`P_\gamma `$ values such that any object with a detected size smaller than that of the stars in the field was assigned a $`P_\gamma `$ equal to that of galaxies with detected sizes similar to that of stars, 0.2 in this case. We choose to apply this correction instead of removing the objects from the catalog as the faintest galaxies should be at the highest mean redshift, and therefore show the greatest amount of gravitational shearing. This artificial cutoff in $`P_\gamma `$ is equivalent to applying a weight function in all subsequent operations such that the galaxies with measured sizes smaller than stellar are given a lower weight. We then created an estimate of the shear for each galaxy $`\widehat{\gamma }_\alpha =e_\alpha /P_\gamma `$. Both the shear $`\gamma `$ and convergence $`\kappa `$ are second derivatives of the surface potential, $`\gamma _\alpha ={}_{}{}^{1}/_{2}^{}\{\varphi _{,11}\varphi _{,22},2\varphi _{,12}\}`$ and $`\kappa =\mathrm{\Sigma }/\mathrm{\Sigma }_{crit}={}_{}{}^{1}/_{2}^{}^2\varphi `$ where $`\mathrm{\Sigma }_{crit}^1=4\pi Gc^2D_lD_{ls}D_s^1`$. Thus, one can create a two dimensional map of the convergence by Fourier transforming shear estimates, multiplying by the appropriate conversion factors, and retransforming (Kaiser & Squires Kaiser93 (1993)). The surface density can then be extracted from the convergence, providing one knows both the redshift of the lens and the redshift distribution of the background galaxies. The intrinsic ellipticities of the galaxies cannot be removed from the shear estimate, and are the dominant source of noise in the analysis. ### 4.2 Results Previous work on lensing by $`z0.8`$ clusters has shown that to maximize the signal-to-noise of the lensing one must use faint blue galaxies (Luppino & Kaiser Luppino (1997); Clowe et al. Clowe98 (1998), Clowe00 (2000)) as these are presumably at higher redshifts than the brighter and redder galaxies, and thus have been subjected to a stronger shear. Also, because of the large number of stars in this field, one must also apply a selection in color to exclude the majority of the stellar population, which still has a significant contribution at the faintest observable magnitudes (as seen in Fig. 6). As a result we selected as our background sample all galaxies with $`24.20<R<26.70`$ and $`RI<0.35`$ ($`23.64<R<26.14`$ and $`RI<0.55`$ after correcting for Galactic extinction). This resulted in a catalog of 910 galaxies ($`17`$ per sq. arcminute), which is less than half the counts expected from the color and magnitude cuts because of incompleteness at the faint end and loss of area due to extended stellar halos. The rms ellipticity of this galaxy sample, after the corrections for psf effects described above, is $`0.4`$, which combined with the number counts gives an rms shear in a square arcminute area of $`0.1`$. The weak lensing mass reconstruction from the background galaxy catalog is shown in Fig. 7. This was created using the KS93 algorithm (Kaiser & Squires Kaiser93 (1993)), which uses the Fourier transform method described earlier to convert shear to convergence. This algorithm can only determine $`\kappa `$ to an unknown additive constant and suffers from biasing at the edges of the frame, but the intrinsic ellipticity distribution of the galaxies is translated into white noise across the field. As can be seen in Fig. 7, we do detect an overdensity of mass in the field, although its center is 60″north and 24″east of the position of the QSO lens. The rest of the peaks seen in the reconstruction are at the level of the expected noise due to the low number density of background galaxies. One way to check that the peak is real and not caused by a small number of high ellipticity galaxies is to randomly divide the background galaxy catalog in two and see if the peak is present in both of the sub-catalogs. We did this ten times, and in all twenty sub-catalogs we detect the mass peak albeit with changes in the shape, position, and amplitude as expected by the increased noise in the reconstruction due to the lower background galaxy number density. The other peaks in the field tended to vary in amplitude and often disappear in the sub-catalog reconstructions, which suggests that they are caused primarily by a handful of galaxies and not a real lensing signal. We also performed a lensing analysis using an image made from both the $`R`$ and $`V`$ images added together. This allowed us to detect $`20\%`$ more galaxies using the same signal-to-noise, $`R`$ magnitude, and $`RI`$ color cuts as given above. The lensing reconstruction for this is also shown in Fig. 7. The extra galaxies and change in the noise properties of the original galaxies results in only a small change in the amplitude of the main peak but a much larger change in the smaller, presumably noise, peaks. We also attempted the same using $`R`$ and $`I`$, and $`V`$, $`R`$, and $`I`$ images added together, but the higher sky noise in the $`I`$ frames resulted in a severe decrease in the number of faint galaxies detected. For the rest of the results we will quote the numbers taken from just the $`R`$ image, but all of the results from the $`R+V`$ image agree within errors to those of the $`R`$ image. To better quantify the strength and significance of the observed overdensity, as well as to detect any low-significance signal at the position of the QSO lens, we computed the azimuthally-averaged tangential shear of the background galaxies as a function of radius from a given position. We can then fit these radial shear profiles with various models and measure the $`\delta \chi ^2`$ between the best fit model and a zero shear model to determine the significance of the peaks. The radial shear profiles centered on the peak seen in the mass reconstruction and on the QSO lens are given in Fig. 8, as well as the best fit isothermal sphere models. For these models we assumed a lens redshift of 1 and background galaxy redshift of 1.5. The $`z_{bg}`$ = 1.5 assumption was derived by calculating $$\overline{\mathrm{\Sigma }}_{crit}(z_bg)=\frac{n(z)𝑑z}{\frac{n(z)}{\mathrm{\Sigma }_{crit}(z)}𝑑z}$$ (1) where $`n(z)`$ is from the HDF south photometric redshift survey of Fontana et al. (Fontana (1999)) using the same magnitude and color cuts as used in the data, after correction for Galactic extinction. Altering these assumptions will change the velocity dispersion, and thus the mass, of the lens model, but does not significantly change the $`\chi ^2`$ or the significance of the peak. The best fit isothermal sphere model for the observed peak, centered on the peak in the KS93 reconstruction, has $`v=`$ 970 km/s, $`\chi ^2/(n1)=`$ 1.02, and $`\delta \chi ^2`$(0 km/s $``$ 970 km/s) = 8.95, and thus is at nearly 3$`\sigma `$ significance. The highest significance in the field using this statistic occurs 5$`\stackrel{}{.}`$6 east and 14$`\stackrel{}{.}`$2 north of the KS93 peak (18$`\stackrel{}{.}`$3 west and 71$`\stackrel{}{.}`$6 north of the QSO lens), with $`\sigma =`$ 1040 km/s, $`\chi ^2/(n1)=`$ 0.62, and $`\delta \chi ^2`$(0 km/s - 1040 km/s) = 9.23. The difference between the maxima in the two techniques is a result of the KS93 reconstruction using all of the galaxies in the frame while the tangential shear uses only those between 17$`\stackrel{}{.}`$2 and 150$`\stackrel{}{.}`$5 from the central position, and that the contributions from the galaxies are weighted in different manners. The best fit isothermal sphere model when the shear profile is centered on the QSO lens has $`\sigma =`$ 560 km/s, $`\chi ^2/(n1)=`$ 1.13, and $`\delta \chi ^2`$(0 km/s $``$ 560 km/s) = 1.05, and therefore has only $`1\sigma `$ significance. The best fit velocity dispersion, as well as the one-$`\sigma `$ deviations, as a function of background galaxy redshift is given in Fig. 9. To test the significances of the peaks given by the $`\delta \chi ^2`$ of the isothermal sphere fits we performed Monte Carlo simulations in which we kept the positions and moduli of the ellipticities of the background galaxies fixed, but applied a random spin to their orientations. We then measured the best isothermal sphere fits to the tangential shear profiles using the central position from the KS93 reconstruction. The resulting best fit isothermal sphere models were a Gaussian distribution around $`v^2=`$ 0 km<sup>2</sup>/s<sup>2</sup>. Only 128 of the 100,000 realizations resulted in a best fit model with $`\sigma `$ 970 km/s, which corresponds to a significance level of 3$`\sigma `$, in agreement with the $`\delta \chi ^2`$ significance level. To determine the significance of not detecting a peak at the position of the QSO lens we performed a similar process to the above. We again kept the positions and moduli of the ellipticities of the background galaxies fixed and randomized their orientations, but then we also applied a shear to the background galaxies. The applied shear was calculated for each galaxy’s position by using a 1000 km/s isothermal sphere centered on the QSO lens with $`z_{lens}=1`$ and $`z_{bg}=1.5`$. Only 5031 of the 100,000 realizations resulted in a best fit model with $`\sigma `$ 560 km/s, which corresponds to $`2.0\sigma `$ significance. ## 5 Conclusions We have detected in $`VRIK`$ images ten galaxies with colors consistent with the lensing galaxy of the MG2016$`+`$112 QSO lens within $`225h^1`$ kpc of the lensing galaxy. This represents an increase in number density of $`1160_{260}^{+440}`$% (1$`\sigma `$ error on the background density) compared to galaxies of similar colors and magnitudes in the rest of the image. We also find nine faint objects, seven of them grouped in a small area, within $`30h^1`$ kpc of the lens galaxy D. Due to their faintness, we are unable to measure colors or determine if they are galaxies or stars except for the brightest of the nine which is a galaxy and somewhat bluer than galaxy D. Even assuming all of these galaxies are associated with the lens galaxy D, they would total a small fraction of the galaxy counts associated with the clusters MS1137$`+`$66 and RXJ1716$`+`$67, both at $`z0.8`$ and of similar X-ray properties as those measured for AXJ2019$`+`$112 (Donahue et al. Donahue (2000); Henry et al. Henry (1998); H97), using the same radial, magnitude, and color cuts adjusted to the cluster BCGs and redshifts. We have also made a mass reconstruction of the field from faint blue galaxies. We find a peak of comparable mass to that given by the X-ray observations, but the center of the peak is over an arcminute north-west of the lens galaxy D (272$`h^1`$ kpc at $`z=1`$). This peak has a best fit singular isothermal sphere model of $`\sigma =`$ 970 km/s and is significant at $`3\sigma `$. We can rule out a 1000 km/s singular isothermal sphere, the mass expected from the X-ray observations, centered on the lens galaxy D at 2.0$`\sigma `$ significance. We do not see any evidence for a group of similarly colored galaxies in the vicinity of the observed mass peak, but this region is highly obscured by bright stars. We plot in Fig. 10 the velocity dispersion of a singular isothermal sphere needed to provide the observed lensing signal as a function of redshift of the lens. Finally, we note that the 1$`\sigma `$ upper mass limit centered on galaxy D from weak lensing is near the 1$`\sigma `$ lower mass limit from the X-ray data. Thus the best measure of the mass of the cluster from the two data sets would give a velocity dispersion of $`\sigma `$800 km/s. ###### Acknowledgements. We wish to thank Genevieve Soucail, Jean-Paul Kneib, Narciso Benitez, Tom Broadhurst, and Piero Rosati for providing information and data on this field prior to publication. We also wish to thank Tadayuki Kodama and Peter Schneider for their advice and help. Some of the data presented herein were obtained at the W.M. Keck Observatory, which is operated as a scientific partnership among the California Institute of Technology, the University of California and the National Aeronautics and Space Administration. The Observatory was made possible by the generous financial support of the W.M. Keck Foundation. The United Kingdom Infrared Telescope is operated by the Joint Astronomy Centre on behalf of the U.K. Partical Physics and Astronomy Research Council. We acknowledge assistance from the Isaac Newton Group Service Programme in providing photometric zero points for the optical data, from the Isaac Newton Telescope on La Palma. DC acknowledges the “Sonderforschungsbereich 375-95 für Astro–Teilchenphysik” der Deutschen Forschungsgemeinschaft for finacial support. NT acknowldedges the PPARC for financial support. 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.
warning/0001/hep-ph0001313.html
ar5iv
text
# CP Violation Beyond the Standard Model in Hadronic B Decays ## Abstract In this talk I discuss three different methods, using $`B_dJ/\psi K_S`$, $`J/\psi K_S\pi ^0`$, $`B_dK^{}\pi ^+,\pi ^+\pi ^{}`$ and $`B_uK^{}\pi ^0,\overline{K}^0\pi ^{},\pi ^{}\pi ^0`$, to extract hadronic model independent information about new physics. 1. Introduction The Gold-plated mode for the study of CP violation in the Standard Model (SM) is $`BJ/\psi K_S`$. This decay mode can be used to measure $`\mathrm{sin}2\beta `$ in the SM without hadronic uncertainties. CDF and ALEPH data obtain $`\mathrm{sin}2\beta =0.91\pm 0.35`$ which is consistent with $`\mathrm{sin}2\beta =0.75_{0.065}^{+0.055}`$ from fitting other experimental data . The best fit value for $`\gamma `$ from the same fitting gives $`\gamma _{best}=59.9^{}`$. Rare hadronic B decays also provide important information . Using factorization calculations for rare hadronic B decays measured by CLEO , $`\gamma `$ is determined to be $`114_{21}^{+25}`$ degrees. It seems that there is a conflict between $`\gamma `$ obtained from rare B decay data and from other fits. At present it is not possible to conclude whether this possible inconsistence is due to experimental and theoretical uncertainties or due to new physics. It is important to find ways to test the SM and to study new physics without large hadronic uncertainties. In this talk I discuss three methods, using $`B_d`$ $`J/\psi K_S`$, $`J/\psi K_S\pi ^0`$ , $`B_dK^{}\pi ^+,\pi ^+\pi ^{}`$ and $`B_uK^{}\pi ^0,\overline{K}^0\pi ^{},\pi ^{}\pi ^0`$ , to extract hadronic model independent information about the SM and models beyond. In the SM the Hamiltonian for hadronic B decays is given by, $`H_{eff}=(4G_F/\sqrt{2})[V_{fb}V_{fq}^{}(c_1O_1^{f(q)}(L)+c_2O_2^{f(q)}(L))_{i=(312),j=(u,c,t)}V_{jb}^{}V_{jq}c_i^jO_i]`$. $`O_{11,12}`$ are the gluonic and electromagnetic dipole operators and $`O_{110}`$ are defined in Ref. . I use the values of $`c_i`$ obtained for the SM in Ref. with $`(c_1,c_2,`$ $`c_3^t,c_4^t`$, $`c_5^t,c_6^t`$, $`c_8^t,c_9^t,c_{10^t},c_{11}^t,c_{12}^t)`$ =(-0.313,1.150,0.017,-0.037, 0.010, -0.046, -0.001$`\alpha _{em}`$, 0.049$`\alpha _{em}`$, -1.321$`\alpha _{em}`$, 0.267 $`\alpha _{em}`$, -0.3, -0.6). $`c_i^{c,u}`$ are due to c and u quarks in the loop and can be found in Ref. . When going beyond the SM, there are new contributions. I take three models for illustrations. Model i): R-parity violation model In R-parity violating supersymmetric (SUSY) models, there are new CP violating phases. Here I consider the effects due to, $`L=(\lambda _{ijk}^{\prime \prime }/2)U_R^{ci}D_R^{cj}D_R^{ck}`$, R-parity violating interaction. Exchange of $`\stackrel{~}{d}_i`$ squark can generate the following effective Hamiltonian at tree level, $`H_{eff}=(4G_F/\sqrt{2})`$ $`V_{fb}^{}V_{fq}`$ $`c^{f(q)}[O_1^{f(q)}(R)O_2^{f(q)}(R)]`$. Here $`O_1^{f(q)}(R)=\overline{f}\gamma _\mu Rf\overline{q}\gamma ^\mu Rb`$ and $`O_2^{f(q)}(R)`$ $`=\overline{f}_\alpha \gamma _\mu Rf_\beta \overline{q}_\beta \gamma ^\mu Rb_\alpha `$. The operators $`O_{1,2}(R)`$ have the opposite chirality as those of the tree operators $`O_{1,2}(L)`$ in the SM. The coefficients $`c^{f(q)}`$ with QCD corrections are given by $`(c_1c_2)(\sqrt{2}/(4G_FV_{fb}V_{fq}^{}))(\lambda _{fqi}^{\prime \prime }\lambda _{f3i}^{\prime \prime }/2m_{\stackrel{~}{d}^i}^2)`$. Here $`m_{\stackrel{~}{d}^i}`$ is the squark mass. $`B_u\pi ^{}\overline{K}^0`$ and $`B_uK^{}\overline{K}^0`$ data constrain $`|c^{c(q)}|`$ to be less than $`O(1)`$ . The new contributions can be larger than the SM ones. I will take the values to be 10% of the corresponding values for the SM with arbitrary phases $`\delta ^{f(q)}`$ for later discussions. Model ii): SUSY with large gluonic dipole interaction In SUSY models with R-parity conservation, potential large contributions to B decays may come from gluonic dipole interaction by exchanging gluino at loop level with left- and right-handed squark mixing. In the mass insertion approximation , $`c_{11}^{new}=(\sqrt{2}\pi \alpha _s/(G_FV_{tb}V_{ts}m_{\stackrel{~}{g}}))(m_{\stackrel{~}{q}}/m_b)\delta _{qb}^{LR}G(m_{\stackrel{~}{g}}^2/m_{\stackrel{~}{q}}^2)`$, where $`m_{\stackrel{~}{g}}`$ is the gluino mass, and $`\delta _{qb}^{LR}`$ is the squark mixing parameter. $`G(x)`$ is from loop integral. $`c_{11}^{new}`$ is constrained by experimental data from $`bs\gamma `$ which, however, still allows $`c_{11}^{new}`$ to be as large as three times of the SM contribution in magnitude with an arbitrary CP violating phase $`\delta _{dipole}`$ . I will take $`c_{11}^{new}`$ to be 3 times of the SM value with an arbitrary $`\delta _{dipole}`$. Model iii): Anomalous gauge boson couplings Anomalous gauge boson couplings can modify $`c_i`$ (mainly on $`c_{710}^t`$) with the same CP violating source as that for the SM . The largest contribution may come from the $`WWZ`$ anomalous coupling $`\mathrm{\Delta }g_1^Z`$. To the leading order in QCD corrections, the new contributions to $`c_{710}^{AC}`$ due to $`\mathrm{\Delta }g_1^Z`$ are given by , $`(c_{7AC}^t,c_{8AC}^t,c_{9AC}^t,c_{10AC}^t)`$ = $`\alpha _{em}\mathrm{\Delta }g_1^Z`$(1.397, 0.391,-5.651,1.143). LEP data constrain $`\mathrm{\Delta }g_1^Z`$ to be within $`0.113<\mathrm{\Delta }g_1^Z<0.126`$ at the 95% c.l.. $`c_{iAC}`$ can be very different from those in the SM. 2. Test new physics using $`\mathrm{sin}2\beta `$ from $`BJ/\psi K_S,J/\psi K_S\pi ^0`$ The usual $`CP`$ violation measure for $`B`$ decays to CP eigenstates is $`\mathrm{Im}\xi =\mathrm{Im}\{(q/p)(A^{}\overline{A}/|A|^2)\}`$, where $`q/p=e^{2i\varphi _B}`$ is from $`B^0`$$`\overline{B}^0`$ mixing, while $`A`$, $`\overline{A}`$ are $`B`$, $`\overline{B}`$ decay amplitudes. For $`BJ/\psi K_S`$, the final state is $`P`$-wave hence $`CP`$ odd. Setting the weak phase in the decay amplitude to be $`2\varphi _0=Arg(A/\overline{A})`$, one has , $`\mathrm{Im}\xi (BJ/\psi K_S)=\mathrm{sin}(2\varphi _B+2\varphi _0)\mathrm{sin}2\beta _{J/\psi K_S}`$. For $`BJ/\psi K^{}J/\psi K_S\pi ^0`$, the final state has both $`P`$-wave ($`CP`$ odd) and $`S`$\- and $`D`$-wave ($`CP`$ even) components. If $`S`$\- and $`D`$-wave have a common weak phase $`\stackrel{~}{\varphi }_1`$ and P-wave has a weak phase $`\varphi _1`$ , $`\mathrm{Im}\xi (BJ/\psi K_S\pi ^0)`$ $`=`$ $`\mathrm{Im}\{e^{2i\varphi _B}[e^{2i\varphi _1}|P|^2e^{2i\stackrel{~}{\varphi }_1}(1|P|^2)]\}`$ (1) $``$ $`(12|P|^2)\mathrm{sin}2\beta _{J/\psi K_S\pi ^0},`$ where $`|P|^2`$ is the fraction of P-wave component. In the SM one has $`\varphi _B=\beta `$ and $`2\varphi _0=2\varphi _1(2\stackrel{~}{\varphi }_1)=Arg[\{V_{cb}V_{cs}^{}\{c_1+a(a^{})c_2\}\}/\{V_{cb}^{}V_{cs}\{c_1+a(a^{})c_2\}\}]=0`$, in the Wolfenstein phase convention. Here $`a`$ and $`a^{}`$ are parameters which indicate the relative contribution from $`O_2^{c(s)}(L)`$ compared with $`O_1^{c(s)}(L)`$ for the P-wave and (S-, D-) wave. In the factorization approximation $`1/a=1/a^{}=N_c`$ (the number of colors). Therefore $`\mathrm{sin}2\beta _{J/\psi K_S}=\mathrm{sin}2\beta _{J/\psi K_S\pi ^0}=\mathrm{sin}2\beta `$. $`|P|^2`$ has been measured with a small value $`0.16\pm 0.08\pm 0.04`$ by CLEO which implies that the measurement of $`\mathrm{sin}2\beta `$ using $`BJ/\psi K_S\pi ^0`$ is practical although there is a dilution factor of 30%. When one goes beyond the SM, $`\mathrm{sin}2\beta _{J/\psi K_S}=\mathrm{sin}2\beta _{J/\psi K_S\pi ^0}`$ is not necessarily true. The difference $`\mathrm{\Delta }\mathrm{sin}2\beta \mathrm{sin}2\beta _{J/\psi K_S}\mathrm{sin}2\beta _{J/\psi K_S\pi ^0}`$ is a true measure of CP violation beyond the SM without hadronic uncertainties. Let us now analyze the possible values for $`\mathrm{\Delta }\mathrm{sin}2\beta `$ for the three models described in the previous section. Because $`BJ/\psi K_S,J/\psi K^{}`$ are tree dominated processes, Models ii) and iii) would not change the SM predictions significantly. $`\mathrm{\Delta }\mathrm{sin}2\beta `$ is not sensitive to new physics in Models ii) and iii). However, for Model i), the contributions can be large. The weak phases are given by $`2\varphi _0=2\varphi _1(2\stackrel{~}{\varphi }_1)=Arg[\{V_{cb}V_{cs}^{}\{c_1+ac_2+()c^{c(s)}\{1a(a^{})\}\}\}/\{V_{cb}^{}V_{cs}\{c_1+ac_2+()c^{c(s)}\{1a(a^{})\}\}\}]`$. Taking the new contributions to be 10% of the SM ones, one obtains $`\varphi _0=\varphi _1\stackrel{~}{\varphi }_10.1\mathrm{sin}\delta ^{c(s)}`$. From this, $`\mathrm{\Delta }\mathrm{sin}2\beta 4((1|P|^2)/(12|P|^2))\mathrm{cos}2\varphi _B(0.1\mathrm{sin}\delta ^{c(s)})0.5\mathrm{cos}2\varphi _B\mathrm{sin}\delta ^{c(s)}`$. $`\varphi _B`$ may be different from the SM one due to new contributions. Using the central value $`\mathrm{sin}2\beta _{J/\psi K_S}=0.91`$ measured from CDF and ALEPH, $`\mathrm{\Delta }\mathrm{sin}2\beta 0.2\mathrm{sin}\delta ^{c(s)}`$ which can be as large as 0.2. Such a large difference can be measured at B factories. Information about new CP violating phase $`\delta ^{c(s)}`$ can be obtained. 3. Test new physics using rate differences between $`B_d\pi ^+K^{},\pi ^+\pi ^{}`$ Theoretical calculations for rare hadronic B decays have large uncertainties. Information extracted using model calculations is unreliable. I now show that hadronic model independent information about CP violation can be obtained using SU(3) analysis for rare hadronic B decays. The operators $`O_{1,2}`$, $`O_{36,11,12}`$, and $`O_{710}`$ transform under SU(3) symmetry as $`\overline{3}_a+\overline{3}_b+6+\overline{15}`$, $`\overline{3}`$, and $`\overline{3}_a+\overline{3}_b+6+\overline{15}`$, respectively. These properties enable one to write the decay amplitudes for $`BPP`$ in only a few SU(3) invariant amplitudes. There are relations between different decays. When small annihilation contributions are neglected, one has $`A(B_d\pi ^+\pi ^{})=V_{ub}V_{ud}^{}T+V_{tb}V_{td}^{}P,`$ $`A(B_d\pi ^+K^{})=V_{ub}V_{us}^{}T+V_{tb}V_{ts}^{}P.`$ Due to different KM matrix elements involved in these decays, although the amplitudes have similarities, the branching ratios are not simply related. However, when considering the rate difference, $`\mathrm{\Delta }(PP)=\mathrm{\Gamma }(B_dPP)\mathrm{\Gamma }(\overline{B}_d\overline{P}\overline{P})`$, the situation is dramatically different because the simple relation $`Im(V_{ub}V_{ud}^{}V_{tb}^{}V_{td})=Im(V_{ub}V_{us}^{}V_{tb}^{}V_{ts})`$. In the SU(3) limit, $`\mathrm{\Delta }(\pi ^+\pi ^{})=\mathrm{\Delta }(\pi ^+K^{}).`$ (2) This non-trivial equality dose not depend on detailed models for the hadronic physics and provides test for the SM . Including SU(3) breaking effect from factorization calculation, one has, $`\mathrm{\Delta }(\pi ^+\pi ^{})\frac{f_\pi ^2}{f_K^2}\mathrm{\Delta }(\pi ^+K^{})`$. Although there is correction, the relative sign is not changed. When going beyond the SM, there are new CP violating phases leading to violation of the equality above. Among the three models described in the first section, Models i) and ii) can alter the equality significantly, while Model iii) can not because the CP violating source is the same as that in the SM. To illustrate how the situation is changed in Models i) and ii), I calculate the normalized asymmetry $`A_{norm}(PP)=`$$`\mathrm{\Delta }(PP)/\mathrm{\Gamma }(\pi ^+K^{})`$ using factorization approximation following Ref. . The new effects may come in such a way that only $`B_d\pi ^+K^{}`$ is changed but not $`B_d\pi ^+\pi ^{}`$. This scenario leads to maximal violation of the equality discussed here. I will take this iscenario for illustration. The results are shown in Figure 1. The solid curve is the SM prediction for $`A_{norm}(\pi ^+K^{})`$ as a function of $`\gamma `$. For $`\gamma _{best}`$ $`A_{norm}(\pi ^+K^0)10\%`$. It is clear from Figure 1 that within the allowed range of the parameters, new physics effects can dramatically violate the equality discussed above. Experimental study of the rate differences $`\mathrm{\Delta }(\pi ^+\pi ^{},\pi ^+K^{})`$ can provide model independent information about CP violation beyond the SM in the near future. 4. Test new physics using SU(3) relation for $`B_u\pi ^{}\overline{K}^0,\pi ^0K^{},\pi ^0\pi ^{}`$ I now discuss another method which provides important information about new physics using $`B_u\pi ^{}\overline{K}^0,\pi ^0K^{},\pi ^0\pi ^{}`$. Using SU(3) relation and factorization estimate for the breaking effects, one obtains $`A(B_u\pi ^{}\overline{K}^0)+\sqrt{2}A(B_u\pi ^0K^{})`$ $`=`$ $`ϵA(B_u\pi ^{}\overline{K}^0)e^{i\mathrm{\Delta }\varphi }(e^{i\gamma }\delta _{EW}),`$ $`\delta _{EW}={\displaystyle \frac{3}{2}}{\displaystyle \frac{|V_{cb}||V_{cs}|}{|V_{ub}||V_{us}|}}{\displaystyle \frac{c_9+c_{10}}{c_1+c_2}},`$ $`ϵ=\sqrt{2}{\displaystyle \frac{|V_{us}|}{|V_{ud}|}}{\displaystyle \frac{f_K}{f_\pi }}{\displaystyle \frac{|A(\pi ^+\pi ^0)|}{|A(\pi ^+K^0)|}},`$ where $`\mathrm{\Delta }\varphi `$ is the difference of the final state rescattering phases for $`I=3/2,1/2`$ amplitudes. For $`f_K/f_\pi =1.22`$ and $`Br(B^\pm \pi ^\pm \pi ^0)=(0.54_{0.20}^{+0.21}\pm 0.15)\times 10^5`$ , one obtains $`ϵ=0.21\pm 0.06`$. Neglecting small tree contribution to $`B_u\pi ^{}\overline{K}^0`$, one obtains $`\mathrm{cos}\gamma =\delta _{EW}{\displaystyle \frac{(r_+^2+r_{}^2)/21ϵ^2(1\delta _{EW}^2)}{2ϵ(\mathrm{cos}\mathrm{\Delta }\varphi +ϵ\delta _{EW})}},r_+^2r_{}^2=4ϵ\mathrm{sin}\mathrm{\Delta }\varphi \mathrm{sin}\gamma ,`$ where $`r_\pm ^2=4Br(\pi ^0K^\pm )/[Br(\pi ^+K^0)+Br(\pi ^{}\overline{K}^0)]=1.33\pm 0.45`$ . The relation between $`\gamma `$ and $`\delta _{EW}`$ is complicated. However it is interesting to note that even in the most general case, bound on $`\mathrm{cos}\gamma `$ can be obtained . For $`\mathrm{\Delta }=(r_+^2+r_{}^2)/21ϵ^2(1\delta _{EW}^2)()>0`$, we have $`\mathrm{cos}\gamma ()\delta _{EW}{\displaystyle \frac{\mathrm{\Delta }}{2ϵ(1+ϵ\delta _{EW})}},\text{or}\mathrm{cos}\gamma ()\delta _{EW}{\displaystyle \frac{\mathrm{\Delta }}{2ϵ(1+ϵ\delta _{EW})}}.`$ (3) The bounds on $`\mathrm{cos}\gamma `$ as a function of $`\delta _{EW}`$ are shown in Fig. 2 by the solid curves for three representative cases: a) Central values for $`ϵ`$ and $`r_\pm ^2`$; b) Central values for $`ϵ`$ and $`1\sigma `$ upper bound $`r_\pm ^2=1.78`$; and c) Central value for $`ϵ`$ and $`1\sigma `$ lower bound $`r_\pm ^2=0.88`$. The bounds with $`|\mathrm{cos}\gamma |1`$ for a), b) and c) are indicated by the curves (a1, a2), (b) and (c1, c2), respectively. For cases a) and c) there are two allowed regions, the regions below (a1, c1) and the regions above (a2, c2). For case b) the allowed range is below (b). When the decay amplitudes for $`B^\pm K\pi `$, $`B^\pm \pi ^\pm \pi ^0`$ and the rate asymmetries for these decays are determined to a good accuracy, $`\gamma `$ can be determined, one obtains, $`(1\mathrm{cos}^2\gamma )[1({\displaystyle \frac{\mathrm{\Delta }}{2ϵ(\delta _{EW}\mathrm{cos}\gamma )}}ϵ\delta _{EW})^2]{\displaystyle \frac{(r_+^2r_{}^2)^2}{16ϵ^2}}=0.`$ (4) To have some idea about the details, I analyze the solutions of $`\mathrm{cos}\gamma `$ as a function of $`\delta _{EW}`$ for the three cases discussed earlier with a given value for the asymmetry $`A_{asy}=(r_+^2r_{}^2)/(r_+^2+r_{}^2)=15\%`$. In the SM for $`r_v=|V_{ub}|/|V_{cb}|=0.08`$ and $`|V_{us}|=0.2196`$, $`\delta _{EW}=0.81`$. The central values for $`r_\pm `$ and $`ϵ`$ prefers $`\mathrm{cos}\gamma <0`$ which is different from the result obtained in Ref. by fitting other data. The parameter $`\delta _{EW}`$ is sensitive to new physics in the tree and electroweak sectors. Model i) has large corrections to the tree level contributions. However, the contribution is proportional to the sum of the coefficients of operators $`O_{1,2}^{u(s)}(R)`$ which is zero in Model i). The above method does not provide information about new physics due to Model i). This method would not provide information about new physics due to Model ii) neither because the gluonic dipole interaction transforms as $`\overline{3}`$ which does not affect $`\delta _{EW}`$. Model iii) can have large effect on $`\delta _{EW}`$. In this model $`\delta _{EW}=0.81(1+4.33\mathrm{\Delta }g_1^Z)`$ which can vary in the range $`0.401.25`$. Constraint on $`\mathrm{cos}\gamma `$ can be very different from the SM prediction. For case a), in the SM $`\mathrm{cos}\gamma <0.18`$ which is inconsistent with $`\mathrm{cos}\gamma _{best}0.5`$. In Model iii) $`\mathrm{cos}\gamma `$ can be consistent with $`\mathrm{cos}\gamma _{best}`$. For case b), $`\mathrm{cos}\gamma `$ is less than zero in both the SM and Model iii). If this is indeed the case, other types of new physics is needed. For case c) $`\mathrm{cos}\gamma `$ can be close to $`\mathrm{cos}\gamma _{best}`$ for both the SM and Model iii). Improved measurements of $`ϵ`$ and $`r_\pm `$ can provide important information about $`\gamma `$ and new physics. 5. Conclusion From discussions in previous sections, it is clear that using $`B_dJ/\psi K_S`$, $`J/\psi K_S\pi ^0`$, $`B_u\pi ^{}K^+,\pi ^+\pi ^{}`$ and $`B^{}\pi ^0K^{},\pi ^{}\overline{K}^0,\pi ^0\pi ^{}`$ important information free from uncertainties in hadronic physics about the Standard Model and models beyond can be obtained. These analyses should be carried out at B factories. Acknowledgements This work was partially supported by National Science Council of R.O.C. under grant number NSC 89-2112-M-002-016. I thank Deshpande, Hou, Hsueh and Shi for collaborations on materials presented in this talk.
warning/0001/hep-ph0001143.html
ar5iv
text
# 1 Introduction ## 1 Introduction Perhaps the most persuasive indication of structure “Beyond the Standard Model” comes from the observation that the strong, and electroweak gauge couplings unify in the simplest supersymmetric extension of the Standard Model. Computing the radiative corrections to the gauge couplings coming from the virtual states of the MSSM one finds the $`SU(3),SU(2)`$ and $`U(1)`$ couplings become very nearly equal at the scale $`(13).10^{16}`$ GeV provided one adopts the $`SU(5)`$ normalisation of the $`U(1)`$ factor. A measure of the accuracy of this prediction may be obtained by using the assumed unification to predict the value of the strong coupling given the weak and electromagnetic couplings. One finds $`\alpha _3(M_z)=0.126\pm 0.003`$ to be compared with the experimental value $`\alpha _3(M_z)=0.119\pm 0.002`$. Only slightly less impressive is the fact that the SUSY radiative corrections, assuming the same unification scale, lead to a prediction of the bottom mass in the range $`3.12<m_b(M_z)<3.76GeV`$ to be compared to the range $`2.72<m_b(M_z)<3.16GeV`$ obtained from the pole mass $`M_b=4.8\pm 0.2GeV`$ . Again this prediction assumes the $`SU(5)`$ prediction for the equality of the bottom to the tau mass at the unification scale. It is largely on the basis of these successes and the fact that supersymmetry is needed to solve the mass hierarchy problem that there has been so much interest in supersymmetric models and in compactification schemes that preserve a low energy supersymmetry. Weakly coupled heterotic string compactifications accommodate these features in a very natural way. The string and gauge unification scales are predicted to be very large, of the order of the Planck scale. The $`SU(5)`$ normalisation of the $`U(1)`$ factor emerges for a wide variety of compactifications even without a stage of Grand Unification below the Planck scale. The relation of the bottom to tau masses is more model dependent but, because there is an underlying $`E_6`$ symmetry above the compactification scale, it is easy to find specific models which do predict equality even without a stage of Grand Unification. This picture has been challenged by the realisation that it is possible to build (open) string theories in which the string scale is much smaller -, possibly no more than the TeV scale which might be small enough to avoid the gauge hierarchy problem. Associated with this is the possibility that there are new large dimensions, possibly as large as a millimetre. At first sight such theories seem to be inconsistent with the MSSM unification of couplings which requires a very large unification scale. However there are several possible ways that gauge unification may be maintained. One is that some, or all, of the Standard Model states may propagate in new large dimensions, leading to power law running of the gauge couplings and the possibility that unification is achieved at a much lower scale. A second possibility is that the normal logarithmic running applies but the ultraviolet cut-off scale is not the string scale but is associated with a very heavy Kaluza Klein or winding mode. This is a particularly interesting possibility because the cut-off scale is related to the infra-red properties of the transverse (closed string) channel. The mass, $`m_{KK},`$ of the lightest Kaluza Klein states in this channel is inversely proportional to the new large compactification radius and is of $`O(10^4eV)`$ for a compactification radius of $`O(1mm).`$ A state of this mass corresponds to a cutoff scale in the open string channel of $`O(M_s^2/m_{KK})`$ and if one sets the string scale, $`M_s,`$ to 1 TeV one obtains $`10^{17}GeV`$ for the ultra violet cut-off. This is tantalisingly close to the gauge unification scale in the MSSM. In this letter we will consider both these possibilities for the case of Type I/I string compactifications. We will concentrate on the question whether it is possible to maintain the remarkable success of the MSSM predictions in the sense that the running of the gauge couplings is governed by the MSSM beta functions. Our conclusion is somewhat pessimistic suggesting that MSSM unification with a low string scale does not follow as nicely as it did in the heterotic string case. Of course it may be that nature is perverse and that the success of unification in the MSSM could be just an accident. As an illustration of this we also consider a promising Type I string compactification in which gauge unification can be made to work. ## 2 Type I/I models and unification with low string scale In Type I (Type I) strings at weak coupling the Standard Model fields are described by open strings confined to a $`D_p`$ brane with Kaluza Klein modes with respect to $`(p3)`$ compact dimensions in the brane and winding modes with respect to $`(9p)`$ dimensions orthogonal to the brane. Closed strings mediate the gravitational interaction and are free to propagate in the full ten dimensional space-time. Upon compactification to four dimensional space-time, the following relations emerge between the volume<sup>3</sup><sup>3</sup>3in Type I/I string length units. $`v_p`$ of the compact dimensions of the SM brane, the volume $`v_o`$ of the dimensions orthogonal to it and the string scale $`M_I`$ $$M_Ig^2M_P\left(\frac{v_p}{v_o}\right)^{\frac{1}{2}}$$ (1) so the string scale may be reduced and may be very low (“TeV scale”) if the transverse space is large . It should however be noted that lowering the fundamental scale does not, by itself, solve the hierarchy problem for now we must explain why $`v_ov_p`$. In addition to the relation of eq.(1) the string coupling is related to the four dimensional gauge coupling $$\lambda _Ig^2v_p$$ (2) where $`v_p`$ should be of order unity to keep $`\lambda _I<1`$. Particularly interesting is the case in which the MSSM is embedded in a $`D3`$ brane, for in this situation one may have extra (transverse) dimensions as large as $`mm`$ (not ruled out by experiment) while the Standard Model field has only winding modes corresponding to the large transverse dimensions. In the case the MSSM is embedded in a $`D_p`$ brane with $`p>3`$ the $`p3`$ additional dimensions cannot be much larger than $`1TeV^1`$ due to the presence of Standard Model Kaluza Klein modes. However in asymmetric compactifications some or all of the $`(9p)`$ orthogonal dimensions may be much larger. Unlike Grand Unification, string theories predict the gauge unification scale. In the heterotic string the Standard Model fields are closed string states and the ultraviolet cutoff scale is determined by the geometry of the string world sheet. At one loop this is the torus and one may readily verify that the cutoff is at the string scale . In the case the Standard Model fields are described by open strings however the world sheet geometry does not regulate the divergences the latter only being eliminated after summing over the possible geometries. After eliminating the divergences the UV behaviour in the open string channel is determined by the IR behaviour in the transverse open string channel. This follows due to the closed string/open string duality and leads to alternative, but equivalent, descriptions of the behaviour of the radiative corrections to gauge couplings. The alternative open and closed string descriptions are summarised in Table 1. Looking at the first two entries in Table 1 one may see that in the open string channel the running of the couplings is due to the familiar radiative corrections to the gauge couplings while in the closed string channel it is due to the modification of the vacuum expectation value of the bulk field, $`\varphi ,`$ which is the coefficient of the gauge kinetic term. The latter running is due to the $`\varphi `$ propagator in the bulk, emitted by distant branes, evaluated at the transverse position of the SM brane. This provides a classical interpretation of the quantum effects on the gauge couplings (UV behaviour) on the SM brane. Logarithmic evolution in the direct channel is due to the radiative corrections involving the propagation of virtual Standard Model states in four dimensions. In the transverse channel this is due to $`\varphi `$ propagation in two transverse dimensions. If the open string states propagate in more than four (“in-brane”) dimensions the couplings will run like a power rather than a logarithm. From the four dimensional point of view an (“in-brane”) extra dimension shows up as a Kaluza Klein (KK) tower of excitations which each contribute additional logarithmic running to the coupling which sum to a (linear) power once sufficient KK states can be excited. In the transverse closed string channel this corresponds to $`\varphi `$ propagation in one transverse dimension (the Green functions fall as $`r^{2d_{}}`$ for $`d_{}`$ transverse dimensions). Similarly if there are two KK towers of excitations in the direct channel the couplings evolve quadratically. In the closed string channel this corresponds to the case $`d_{}=0.`$ In the weakly coupled heterotic string with orbifold or Calabi Yau supersymmetric compactifications from 10 to 4 dimensions it might seem there could be running proportional to $`r^6.`$ However, the massive KK modes giving rise to the extra dimensions fill out $`N=2`$ and $`N=4`$ supermultiplets. The latter do not contribute to radiative corrections and so the power law running is limited to $`r^2.`$ In the transverse channel this result follows simply from the observation that $`d_{}=0`$ is the limiting case. The prediction of gauge unification also requires that the tree level coupling constant ratios be predicted. The gauge couplings are related to the dilaton couplings to the various gauge kinetic terms. In the heterotic string the reason for unification is the presence of a single dilaton v.e.v. which is the string coupling at the unification scale. For level-1 Kac-Moody theories one has the usual $`SU(5)`$ relations. More generally such a relation may come from an underlying Grand Unified symmetry above the compactification scale. In Type I/I theories it has been suggested that in the closed string channel it is an underlying geometric symmetry of the brane sector that is responsible for relating the couplings. Finally string theories predict the gauge unification scale. The prediction for the tree level ratios apply at the scale at which the radiative corrections are cut-off. As discussed above this is determined by the lightest relevant state in the transverse channel. This turns out either to be the string scale or the lightest KK scale depending on which sector of the theory one is considering . ### 2.1 Power-law unification The first discussion of the possibility of a low string and unification scale following from power law running employed an effective field theory approach. In this the individual logarithmic corrections of Kaluza Klein modes (with respect to $`\delta =(p3)`$ ($`p3`$) “in-brane” dimensions) below the string scale are summed . This leads to the following (one-loop) RGE evolution $`\alpha _i^1(\mu )`$ $`=`$ $`\alpha ^1(\mathrm{\Lambda })+{\displaystyle \frac{\overline{b}_i}{2\pi }}{\displaystyle \underset{\stackrel{}{m}}{\overset{|\stackrel{}{m}|\mathrm{\Lambda }/\mu _0}{}}}\mathrm{ln}{\displaystyle \frac{\mathrm{\Lambda }}{\mu _0|\stackrel{}{m}|}}+{\displaystyle \frac{b_i}{2\pi }}\mathrm{ln}{\displaystyle \frac{\mathrm{\Lambda }}{\mu }}`$ (3) $``$ $`\alpha ^1(\mathrm{\Lambda })+{\displaystyle \frac{\overline{b}_i}{2\pi }}\left\{{\displaystyle \frac{\pi ^{\delta /2}}{\delta \mathrm{\Gamma }(1+\delta /2)}}\left[\left({\displaystyle \frac{\mathrm{\Lambda }}{\mu _0}}\right)^\delta 1\right]\mathrm{ln}{\displaystyle \frac{\mathrm{\Lambda }}{\mu _0}}\right\}+{\displaystyle \frac{b_i}{2\pi }}\mathrm{ln}{\displaystyle \frac{\mathrm{\Lambda }}{\mu }}`$ (4) which shows explicitly the origin of the “power-law” behaviour in the high degeneracy of Kaluza-Klein modes. Here $`\overline{b}_i`$ is the $`N=2`$ beta function of KK sector and $`b_i`$ is the one loop contribution, $`\delta `$ is the effective number of (“in-brane”) extra dimensions, $`\mathrm{\Lambda }`$ is the cut-off scale and $`\mu _0`$ is the mass of the lowest KK excitation. Although the number of extra dimensions may be as large as six, as discussed above, we expect at most quadratic running, $`\delta =1,2`$. The constraint (2) restricts the in-brane volume to values close to unity in string length units. In general the power-law running takes place over a very small range of energy and thus the radii ($`1/\mu _0`$) are close to the string scale and (2) may be satisfied by symmetric unifications with the remaining $`6\delta `$ dimensions also of order the string length scale. In the effective field theory approach unification occurs at the cutoff scale $`\mathrm{\Lambda }`$ and this scale must be identified. The usual assumption is to identify it with the string scale, $`M_I.`$ However, following the discussion above, in Type I theories it may be one should identify $`\mathrm{\Lambda }`$ with the larger winding mode mass. The correct identification depends on the details of the string theory - in some cases the winding mode excitations fill out complete $`N=4`$ multiplets and in this case the lower string scale is the appropriate choice . In either case the power law running can lead to unification at a scale which may be smaller than the ordinary MSSM unification scale. While power law running can reduce the unification scale there are two difficulties in obtaining the usual MSSM predictions. The first is the sensitivity power law running introduces to the thresholds at the unification scale . This is readily seen from eqs.(3), (4) where, for a low (TeV) unification scale $`\mu _0(\alpha _i^1/\mu _0)\delta (\alpha _i^1(\mu )\alpha ^1(\mathrm{\Lambda })).`$ This is to be compared to the case of normal logarithmic running in the MSSM where $`\mu _0(\alpha _i^1/\mu _0)(\alpha _i^1(\mu )\alpha ^1(\mathrm{\Lambda }))/\mathrm{ln}(\mathrm{\Lambda }/\mu ).`$ Thus the sensitivity to the unification scale is roughly enhanced by a factor $`\delta \mathrm{ln}(\mathrm{\Lambda }/\mu )`$ which is $`32.`$ A more careful analysis gives $`\delta \mu _o/\mu _o=1/224`$ . This means that in order to make a precision prediction for the gauge coupling one must know the detailed spectrum of the states at the unification scale. Although this does not rule out the possibility of low scale unification through power law running, it does make the success of the detailed prediction of the gauge couplings in the MSSM seem surprising. The second difficulty one encounters in trying to duplicate the MSSM results via power law running comes from the fact that in eq.(3) the rate of running is controlled by the coefficients $`\overline{b}_i`$ which are determined by the $`N=2`$ sector of the theory. The difficulty is that these are not expected to be proportional to the $`N=1`$ beta functions and thus we may expect substantial deviations from the MSSM predictions. To see this note that if one has towers of KK states for the entire MSSM spectrum, $`\overline{b}_i=2T_i(G)+2_\psi T_i(\psi )`$ which is not proportional to $`b_i^{MSSM}=3T_i(G)+_\psi T_i(\psi )`$ of the ($`N=1`$) MSSM sector. Instead we have $`\overline{b}_i=2b_i^{MSSM}+4T_i(G)`$. To avoid this we should look for a model for which $$\overline{b}_i=Kb_i^{MSSM}+C$$ (5) with $`K`$ and $`C`$ some gauge group independent constants. In this case eq.(3) gives the same value for $`\alpha _3(M_z)`$ as in the MSSM at one loop level (in such a case we would have an explanation for the “accidental” success of the MSSM gauge unification predictions). Attempts to obtain a spectrum satisfying this condition have been made by Kakushadze by including further KK states which, together with the KK excitations of the MSSM, have one loop coefficients which satisfy eq.(5). A simple example with $`K=1`$ was given by a $`Z_2`$ orbifold model which breaks an underlying $`N=2`$ supersymmetry to $`N=1`$. The $`N=2`$ spectrum is that of the MSSM KK excitations together with those of additional superfields $`F_\pm `$, $`SU(3)\times SU(2)`$ singlets with $`U(1)_Y`$ charge $`\pm 2`$ respectively. However the addition of the $`F_\pm ,`$ while leading to a the $`N=2`$ beta function which satisfies eq.(5), does not reproduce the $`N=1`$ MSSM spectrum because it contains additional light $`N=1`$ fields with the quantum numbers of $`F_\pm `$. As a result one finds $$\alpha _3^1(M_z)=\alpha _3^{o1}(M_z)+\frac{1}{2\pi }\frac{6}{7}\mathrm{ln}\frac{M_s}{M_F}$$ (6) Here $`\alpha _3^0(M_z)`$ is the MSSM (one loop) value while the last term is the (one-loop) contribution of $`F_\pm `$ states to the running of the gauge couplings. This term depends on the (unknown) ratio $`M_s/M_F`$ and thus eq.(6) makes no prediction for $`\alpha _3(M_z)`$ following from the unification condition. A detailed mechanism or further input is needed to fix the value $`M_s/M_F`$. Thus, although the model may allow the presence of a low (TeV) string scale, it does not predict $`\alpha _3(M_z)`$. It may be that one may construct models of this type which do not modify the $`N=1`$ spectrum but one may see from this example the difficulty in accommodating the MSSM predictions in the framework of power-law running. ### 2.2 Logarithmic unification We turn now to the possibility that the gauge couplings unify with logarithmic running. A Type I/I full string calculation for the threshold corrections to the gauge couplings was performed for the case of an $`N=2`$ orientifold based on the $`T^4/Z_2`$ orbifold and other $`N=2`$ orientifolds obtained by toroidal compactification of six-dimensional vacua . These results were also generalised to four dimensional compactifications with $`N=1`$ supersymmetry. In the latter case the cut-off for $`N=2`$ sector is not the string scale itself but the first winding mode above the string scale associated with the (“off-the-brane”) two-torus while for the light states $`N=1`$ the cut-off is set by the string scale itself. Two possibilities exist for the RGE behaviour at the high scale in this case. One of these is the linear “power-law” regime, corresponding to an “asymmetric” two-torus compactification, while the other has only logarithmic dependence on $`R`$ (and ultimately on $`M_P`$) in the RGE. We will now concentrate on the second case as the first one was discussed above. The interest for phenomenology in these models is two-fold. Firstly, the large transverse volume and low string scale brings the possibility (e.g. for the case the MSSM is confined to a D3 brane) of new gravitons and their Kaluza Klein towers in the $`mm^1`$ range , a situation not ruled out by the experiment. Secondly, the mild logarithmic dependence of the couplings on the high scale raises the possibility of gauge coupling unification at a large scale above $`M_I`$ given by the string cut-off (first winding mode). This may be just the original unification scale of the MSSM thus elegantly providing an alternative (Type I) interpretation of the successful MSSM predictions. Turning to a more detailed discussion of the latter possibility note that the corrections to gauge couplings come only from the $`N=1`$ (massless) sector and from the $`N=2`$ massive winding (Type I) (or Kaluza Klein, Type I) sector. Let us consider the $`N=2`$ sector first. The massive $`N=2`$ threshold corrections to $`\alpha ^1`$ at the string scale $`M_I`$ have the form (for both $`𝒜`$ (annulus) and $``$ (Möbius strip) geometries relevant at one loop) $$\mathrm{\Delta }_i^{TypeI}=\frac{1}{4\pi }\underset{m}{}\overline{b}_{i,m}\left\{\mathrm{ln}\left(\sqrt{G_m}ImU_mM_I^2|\eta (U_m)|^4\right)\right\}$$ (7) Here $`G_m`$ is the metric on the torus<sup>4</sup><sup>4</sup>4 We consider in the following only compactification on a single torus, $`T^1`$. $`T^m`$, and $`\sqrt{G_1}=\stackrel{~}{R}_1\stackrel{~}{R}_2`$ (for a rectangular torus). The behaviour of the thresholds<sup>5</sup><sup>5</sup>5For Type I case we should take $`R=1/\stackrel{~}{R}`$ when $`ImU=\stackrel{~}{R}_1/\stackrel{~}{R}_2`$ is of order one is indeed logarithmic, $`\mathrm{\Delta }_a\mathrm{ln}(\stackrel{~}{R}_1\stackrel{~}{R}_2)`$. We note that unlike the weakly coupled heterotic case where quadratic dependence on the common radius of the two torus (associated with the $`N=2`$ sector) exists, in the present case this does not happen. The condition of global tadpole cancellation ensures that after adding the contributions (each regulated by an infrared cut-off) of massless closed string states emitted into the transverse channel, the two $`R^2`$ dependent results obtained for $`𝒜`$ and $`𝒢`$ cancel. Therefore there is no quadratic term present in (7). In this calculation the ultraviolet (would-be quadratic) behaviour in the open channel is regulated by the infra-red regime of the closed string channel (c.f. Table 1). From eq.(7), after adding the usual $`N=1`$ one loop term proportional to $`b_i`$ we have $$\alpha _i^1(M_z)=\alpha _I^1+\frac{b_i}{2\pi }\mathrm{ln}\frac{M_I}{M_z}+\frac{\overline{b}_i}{2\pi }\mathrm{ln}\frac{\stackrel{~}{\mu }}{M_I}$$ (8) where we have taken $`\stackrel{~}{R}_1=\stackrel{~}{R}_2`$ and $`\stackrel{~}{\mu }`$ =$`1/(\stackrel{~}{R}_1|\eta (U_m)|^2).`$ Here $`\overline{b}_i`$ is the $`N=2`$ one-loop beta function coefficient. From this equation we can immediately identify a problem in obtaining the usual MSSM results with a low string scale. As in the case of power law running, the dominant evolution is governed by the $`N=2`$ sector. We therefore have a similar difficulty to that discussed above in obtaining the MSSM results because in general $`\overline{b}_i`$ is not proportional to $`b_i^{MSSM}.`$ In ref. an orbifold projection was used to break $`N=2`$ supersymmetry to $`N=1`$. In this case the running up to the unification scale (i.e. first winding mode scale $`\stackrel{~}{\mu }`$) is indeed given by the $`N=1`$ sector apparently solving this problem. However the construction works only for modding by freely acting groups and this implies the $`N=1`$ matter fields giving rise to $`\overline{b}_i`$ originate from the adjoint representation of the underlying gauge group before modding. Although in the orbifolded model there are $`N=1`$ matter fields transforming as the fundamental representations $`F`$ under one of the group factors, there are also fields transforming as $`\overline{F}`$ . This does not allow for an identification of these fields with the MSSM matter fields. As a result we still have the problem of explaining why the $`\overline{b}_i`$ should satisfy eq.(5). If we assume however that this problem can be solved, eq.(8) together with eq.(5) reproduces the MSSM unification at one loop level with $`(\stackrel{~}{\mu }/M_I)^KM_I10^{16}`$ GeV. For the case $`K=1`$ the unification occurs at $`\stackrel{~}{\mu }10^{16}`$ GeV. Somewhat surprisingly logarithmic unification also has a fine tuning problem. Naively our previous discussion suggests logarithmic unification will be no more sensitive to threshold effects than the MSSM. However this is misleading because here the string cut-off scale $`\stackrel{~}{\mu }10^{16}GeV`$ corresponds to a cross channel exchange state with mass $`M_I^2/\stackrel{~}{\mu }`$. For a 1 TeV string scale this is $`10^{10}`$ $`GeV`$ and corresponds to the lightest mass of Kaluza Klein modes in the closed string channel. However the state controlling the UV behaviour of the open string channel has vacuum quantum numbers. Such a state has no symmetry (local gauge symmetry etc) to protect it from receiving contributions to its mass at scales below the supersymmetry breaking scale. We therefore expect that such a state will acquire mass from supersymmetry breaking effects. The magnitude of such a mass depends on the supersymmetry breaking mechanism but has a lower bound governed by the supersymmetry breaking communicated by gravitational effects. If supersymmetry is broken in a hidden sector (e.g. on another brane) and communicated to the visible sector by gravitational effects then we know such effects must be of $`TeV`$ scale otherwise some of the SUSY partners of the SM should have been seen. However the supersymmetry breaking in the bulk must be at least of $`TeV`$ scale as well because, in this case, SUSY breaking is communicated to the visible sector via the bulk. As a result logarithmic running is not a viable mechanism for gauge unification because now the closed string state mass $`(M_I^2/\stackrel{~}{\mu })1TeV`$ so with $`M_I1TeV`$ we have $`\stackrel{~}{\mu }1TeV`$ instead of the $`10^{16}`$ GeV needed for unification. If, to avoid this problem, one requires the closed string exchange state with mass $`M_I^2/\stackrel{~}{\mu }`$ be $``$ $`TeV`$ scale, while maintaining high scale unification ($`\stackrel{~}{\mu }10^{16}GeV)`$ one finds the string scale is bounded by $`M_I10^{10}`$ GeV. The only possibility to maintain logarithmic unification with a low string scale is that supersymmetry breaking in the visible sector is much larger than in the bulk. This happens in gauge mediated scenarios where the supersymmetry breaking is communicated to the visible sector by gauge interactions much stronger than gravity. Such a scheme requires that supersymmetry breaking occurs on the SM brane itself. However, even in such a scheme there is a problem in keeping the string scale as low as $`1TeV`$ as gravitational effects still generate a supersymmetry breaking mass of order the gravitino mass for fields in the bulk. The gravitino mass is of order $`M_{Susy}^2/M_P`$ where<sup>6</sup><sup>6</sup>6We denote by $`M_P`$ the Planck scale. $`M_{Susy}`$ is the supersymmetry breaking scale on our brane. Demanding that the gravitino mass is less than $`10^{10}`$ $`GeV`$ to avoid the fine tuning problem implies $`M_{Susy}10^4GeV.`$ Provided one remains in the perturbative domain this is probably too small for a viable gauge mediated scheme because SUSY breaking in the visible sector is given by $`\gamma M_{Susy}^2/M`$ where $`\gamma `$ is the effective coupling of the visible sector to the SUSY breaking sector and $`M`$ is the messenger mass $`(MM_{Susy})`$. In viable models it occurs at one loop order for gauginos suggesting $`\gamma `$ is too small to generate acceptable masses for the MSSM sparticle spectrum. Even if it proves to be possible one sees that the resulting model is heavily constrained requiring gauge mediated supersymmetry breaking with the gauge mediator mass, $`M`$, in the $`TeV`$ range too. Gauge mediation with a light messenger sector $`(<10^5GeV)`$ is however unnatural . To summarise, in both the case of power law and logarithmic running there are difficulties in accommodating the success of the MSSM unification predictions in models with a low string scale. In the case of power law unification there is also an enhanced sensitivity to the details of the mass spectrum at the unification scale making the success of the detailed prediction of the gauge couplings in the MSSM quite surprising. In the case of logarithmic running there is a fine tuning problem which can be alleviated by raising the string scale significantly above the $`TeV`$ scale. However in both cases it is difficult to generate the MSSM beta function via the N=2 sector. To avoid this difficulty in these models it is thus necessary to give up the MSSM structure. From this point of view, in this case the MSSM remarkable success in predicting gauge unification appears to be accidental. To illustrate this last possibility we consider unification in a promising Type I which has a multiplet structure close to that of the MSSM. ### 2.3 A Type I model with intermediate unification The model we wish to consider is a specific example of a larger class of $`D=4`$ Type I vacua obtained from a standard $`D=4,`$ $`N=1`$ compact Type IIB orientifold with D<sub>p</sub> branes and anti-D<sub>p</sub> branes located at different points of the underlying orbifold . We consider here the class of models of which have gravity mediated supersymmetry breaking with the full Standard Model embedded in a 9-brane sector and with an additional set of 5-branes trapped at the fixed points of the $`Z_3`$ orientifold. This model has no $`N=2`$ sector and the UV cutoff scale for the $`N=1`$ sector is the string scale . As we shall discuss this model can give gauge unification via the radiative corrections of the $`N=1`$ sector with an intermediate string scale and without introducing significant threshold sensitivity. The price to achieve this is to depart from the MSSM light matter content and to start with a non-MSSM relation between the gauge couplings. In particular the normalisation of the hypercharge of these states is $`\alpha _Y/\alpha _s=3/11`$ at the string scale (instead of the usual “SU(5) factor” $`3/5`$). The massless spectrum contains the spectrum of the MSSM. In addition there are five further pairs of Higgs doublets and three vector-like right handed colour triplets ($`d^c+\overline{d^c}`$). The remaining states of the model may have mass of the order of string scale. In determining the unification prediction of this model we follow and assume that these additional states have a common bare mass<sup>7</sup><sup>7</sup>7 For the purpose of two loop RGE only the bare mass of the fields is needed., $`\stackrel{~}{m}`$, which must not be greatly different from the electroweak scale. Including the threshold effects at $`\stackrel{~}{m}`$, the two loop RGE equations are given by $$\alpha _i^1(M_z)=\delta _i+\alpha _s^1+\frac{b_i}{2\pi }\mathrm{ln}\left[\frac{M_s}{M_z}\right]+\frac{\delta b_i}{2\pi }\mathrm{ln}\left[\frac{M_s}{\stackrel{~}{m}}\right]+\frac{1}{4\pi }\underset{j=1}{\overset{3}{}}Y_{ij}\mathrm{ln}\left[\frac{\alpha _g}{\alpha _j(\stackrel{~}{m})}\right]+\frac{1}{4\pi }\underset{j=1}{\overset{3}{}}\frac{b_{ij}}{b_j}\mathrm{ln}\left[\frac{\alpha _s}{\alpha _j(M_z)}\right]$$ (9) where $$Y_{ij}=\frac{\delta b_j}{b_jb_j^{}}\left[2b_jT_j(G)\delta _{ij}b_{ij}\right]$$ (10) with $`T_j(G)=\{0,2,3\}_j`$, $`b_{ij}`$ is the two loop beta function as in the MSSM but with $`3/11`$ hypercharge normalisation, $`b_j=\{11\xi ,1,3\}_j`$, ($`\xi =3/11`$), $`b_j^{}=b_j+\delta b_j`$, with $`\delta b_j=\{7\xi ,5,3\}`$ to account for the additional five Higgs pairs and three pairs of right handed colour triplets. For the terms $`Y_{i3}`$ equations (9) should be understood as limits of $`b_3^{}0`$ which indeed give the appropriate (finite) results when multiplied by the term $`\mathrm{log}(\alpha _g/\alpha _3(\stackrel{~}{m}))`$. Finally one should include the low energy supersymmetric thresholds $`\delta _i`$ and regularisation scheme conversion factors ($`\overline{MS}\overline{DR}`$). Usually $`\delta _i`$’s are included as an overall effect on $`\alpha _3(M_z)`$ through an effective (low-energy) supersymmetric threshold . In this case the different hypercharge normalisation of $`\delta _1`$ prevents us from using this approach, although their overall effect on $`\alpha _3(M_z)`$ is not expected to change significantly from the “normal” hypercharge case (when $`\alpha _3(M_z)`$ is reduced by an amount of $`24\%`$). The predictions for the string coupling and the string scale $`M_s`$ can be read from the plots in Figure 1 to see that the model indeed allows for a logarithmic unification at about $`10^{10}M_z`$ GeV for a strong coupling consistent with the experimental value $`\alpha _3(M_z)=0.119\pm 0.002`$. This requires an intermediate threshold at about<sup>8</sup><sup>8</sup>8The results are subject to additional thresholds at the string scale, which may change this picture significantly. $`\stackrel{~}{m}(100300)M_z`$. One can also compute the bottom to tau mass ratio, taking account of the different normalisation of $`\alpha _Y/\alpha _s`$. For the model with the above spectrum set-up, we find the following one loop level result (including only gauge effects) $$\frac{R(M_s)}{R(M_z)}=\left(1+\frac{\alpha _sb_1^{}}{2\pi }\mathrm{ln}\frac{M_s}{\mu _o}\right)^{\frac{1}{b_1^{}}\frac{2N_1\delta b_1}{b_1}}\left(\frac{M_s}{\mu _o}\right)^{\frac{\alpha _s}{2\pi }\frac{2N_3\delta b_3}{b_3}}\left(\frac{\alpha _s}{\alpha _1(M_z)}\right)^{\frac{2N_1}{b_1}}\left(\frac{\alpha _s}{\alpha _3(M_z)}\right)^{\frac{2N_3}{b_3}}$$ (11) where $`N_1=10/9\xi `$, $`N_3=8/3`$ and $`R(Q)`$ denotes the ratio bottom to tau mass at the scale Q. For $`\alpha _s0.1`$ (needed to generate $`\alpha _3(M_z)`$) we find that $`R(M_z)5R(M_s)`$. If we assume bottom-tau unification at the string scale, this implies $`m_b(M_z)8.731GeV`$. To correct this, model dependent Yukawa effects not considered here must be very significant, but then the MSSM result for the bottom-tau mass ratio should be regarded as yet another accident. This example shows that it is possible in a realistic model to obtain good results for gauge unification in a manner incompatible with the original MSSM, demonstrating that the latter’s success may indeed be an illusion. However it also illustrates the difficulties a non-MSSM scheme must surmount in order to approach the accuracy of the MSSM prediction. Although the unification is logarithmic the sensitivity to thresholds is considerably enhanced due to the light states additional to the MSSM. For example the uncertainty in the additional Higgs doublets mass translates to the uncertainty $`\delta \alpha _2^1=(5/2\pi )\delta M_H/M_H.`$ If one is to predict $`\alpha _3(M_z)`$ to the present experimental accuracy then one must know these masses to better than $`\delta M_H/M_H=0.1.`$ In the light of this the precision of the MSSM prediction for the strong coupling seems even more remarkable and makes it harder for us to accept it is just an accident. ## 3 Conclusions We have considered the possibility that Type I string theories with a low (TeV) string scale may accommodate the successful MSSM predictions for the unification of the gauge couplings. Our conclusions are pessimistic. Models with power-law running have a significant sensitivity to the unification scale thresholds. Models with logarithmic running up to the string cut-off (winding) scale far above the string scale reduce the sensitivity to this scale but have a fine tuning problem associated with the closed string spectrum which requires either a very special Supersymmetry breaking sector or an intermediate string scale. In both cases it is difficult to generate the MSSM beta function via the N=2 sector. Of course it may be that the success of the MSSM predictions is illusory and the existence of a realistic Type I theory which makes a non-MSSM prediction for the gauge couplings consistent with the measured values demonstrates this possibility. However, due to the threshold dependence on the non-MSSM states, it is not possible in this model to make a prediction for the gauge couplings to the same accuracy as the MSSM result without a detailed knowledge of the masses at the unification scale. All this emphasizes the fact that if the MSSM result is regarded as an accident it is a surprisingly accurate accident! In the light of all this it seems to us that the original heterotic string (possibly strongly coupled), being able naturally to accommodate the MSSM results, still offers the best explanation of the unification of gauge couplings. ## 4 Acknowledgments We would like to thank I. Antoniadis and L. Ibáñez for useful discussions. D.G. acknowledges the financial support from the part of the University of Oxford (Leverhulme Trust research grant). The research is supported in part by the EEC under TMR contract ERBFMRX-CT96-0090.
warning/0001/astro-ph0001077.html
ar5iv
text
# * ## 1 The Mimimum Mass Required to Form a Black Hole For the study of black holes, the relevance of neutron stars is mostly that below a certain maximum mass, degeneracy pressure due to nucleons is sufficient to prevent an object from becoming a black hole. Unfortunately, the equation of state of matter at densities above nuclear-matter density is rather uncertain and theoretical estimates of the minimum mass required to form a black hole range from just over $`1.4M_{}`$ to about $`2.5M_{}`$ . Constraints on the equation of state can be obtained from a variety of observed properties of neutron stars. Among the more direct measurements that have been used or proposed are: (i) the maximum mass inferred from dynamical measurements in binaries; (ii) the minimum spin period among milli-second pulsars; (iii) identification of kHz quasi-periodic oscillations with the orbital frequency in the last stable orbit; (iv) identification of quasi-periodic oscillations with the Lens-Thirring precession period; (v) influence of gravitational light bending on X-ray light curves in radio pulsars; (vi) model-atmosphere fitting of X-ray lightcurves during X-ray bursts (resulting from to thermonuclear run-aways on the neutron-star surface); (vii) gravitational redshift from $`\gamma `$-ray spallation lines in accreting systems; (viii) gravitational redshift and surface gravity from model-atmosphere analysis of spectra of isolated neutron stars. Less direct measurements include: (ix) matching observed neutron-star temperatures and inferred ages to cooling curves; (x) neutrino light curves in supernova explosions; (xi) pulsar glitches; (xii) moment of inertia from accretion torques in combination with knowledge of the magnetic field strength from cyclotron lines; (xiii) comparing X-ray fluxes between states in which a neutron star is accreting and in which matter is stopped at the magnetosphere and “propelled” away. For references and somewhat more detail, see . The strongest constraints on the equation of state are still set by dynamical mass measurements, so I will restrict myself to those below. ## 2 Dynamical measurements Most mass determinations have come from radio timing studies of pulsars; see for an excellent review. The most accurate ones are for pulsars that are in eccentric, short-period orbits with other neutron stars, in which several non-Keplerian effects on the orbit can be observed: the advance of periastron, the combined effect of variations in the second-order Doppler shift and gravitational redshift, the shape and amplitude of the Shapiro delay curve shown by the pulse arrival times as the pulsar passes behind its companion, and the decay of the orbit due to the emission of gravitational waves. The most famous of the double neutron-star binaries is the Hulse-Taylor pulsar, PSR B1913+16, for which recent measurements give $`M_{\mathrm{PSR}}=1.4411\pm 0.0007M_{}`$ and $`M_{\mathrm{comp}}=1.3874\pm 0.0007M_{}`$ . Almost as accurate masses have been inferred for PSR B1534+12, for which the pulsar and its companion are found to have very similar mass: for both, $`M=1.339\pm 0.003M_{}`$ . Neutron-star masses can also be determined for some binaries containing an accreting X-ray pulsar, from the amplitudes of the X-ray pulse delay and optical radial-velocity curves in combination with constraints on the inclination (the latter usually from the duration of the X-ray eclipse, if present). This method has been applied to about half a dozen systems , but the masses are generally not very precise. So far, for all but one of the neutron stars, the masses are consistent with being in a surprisingly narrow range, which can be approximated with a Gaussian distribution with a standard deviation of only $`0.04M_{}`$ . The mean of the distribution is $`1.35M_{}`$, close to the “canonical” value of $`1.4M_{}`$. The one exception is the X-ray pulsar Vela X-1, which is in a 9-day orbit with the B0.5 Ib supergiant HD 77581. For this system, a rather higher mass of around $`1.8M_{}`$ has consistently<sup>1</sup><sup>1</sup>1One study based on IUE spectra of HD 77581 appeared to find a lower neutron-star mass . However, this was found to be due to a bug in the cross-correlation software used been found ever since the first detailed study in the late seventies . A problem with this system is that the measured radial velocities show strong deviations from a pure Keplerian radial-velocity curve, which are correlated within one night, but not from one night to another. A possible cause could be that the varying tidal force exerted by the neutron star in its eccentric orbit excites high-order pulsation modes in the optical star which interfere constructively for short time intervals. We were granted time at ESO to improve the mass determination of this possibly very massive neutron star from 200 new spectra, taken in as many nights. These cover more than 20 orbits, and make it possible to average out the velocity excursions and to constrain possible systematic effects with orbital phase. In combination with measurements from our old photographic plates, earlier CCD spectroscopy, and high resolution IUE spectra, we derived a 95% confidence constraint on the mass of the neutron star of $`M_{\mathrm{NS}}=1.87_{0.17}^{+0.23}`$ . Our constraints are illustrated graphically in Fig. 1. One sees that even at 99% confidence, $`M_{\mathrm{NS}}>1.6M_{}`$. It should be noted, however, that from the data it appears that while the excursions in radial velocity are mostly random, there is also a component that is systematic, locked to orbital phase. Since we do not understand these effects, it may be that our mass estimate is biased. In our trials with excluding the worst-affected phase ranges, however, we consistently found that the fitted mass became even higher . ## 3 Trying for Bias The narrow range in masses inferred for most neutron stars might be seen as evidence for a relatively low maximum neutron-star mass. It could also be, however, that it reflects the formation mechanism. Indeed, from models, it appears that supernova explosions result in neutron stars with masses preferentially in two narrow peaks, one around $`1.3M_{}`$ and one around $`1.65M_{}`$ . It is tempting to associate the latter with Vela X-1, but for honesty it should be noted that from the same calculations it is expected that only lighter neutron stars can be formed by stars which lose their envelope during their evolution, as would likely have happened for the progenitor of Vela X-1. If the narrow range indeed reflects the formation process, it seems worthwile to focus especially on systems in which the neutron star is expected to have accreted a lot of matter since it was formed. Such systems are the low-mass X-ray binaries and their descendents, the pulsars with low-mass white dwarf companions. In the latter systems, the white dwarfs typically have masses of $`0.3M_{}`$. However, in order for mass transfer to have happened, these stars need to have evolved to at least the end of the main sequence life. For this to happen in a Hubble time, their masses need to have been at least $`0.8M_{}`$ initially. Since the mass transfer in these systems is thought to be stable, the neutron star should thus have accreted more than $`0.6M_{}`$, i.e., have become substantially more massive than it was initially. From radio timing measurements, it is generally more difficult to measure masses for these systems, because the orbits are circular and relatively wide. However, for one system, PSR B1855+09, the orbit is very close to edge on, and it has been possible to derive constraints from the Shapiro delay . These constraints are indicated in Fig. 2. Another way of determining the masses in these systems uses optical spectroscopy of the white-dwarf companion. From a model-atmosphere fit to the spectrum, one can determine the effective temperature and surface gravity. From the latter, the white-dwarf mass follows, assuming a theoretical mass-radius relation. If one can also measure the radial-velocity orbit, and determine the mass ratio (in combination with the pulse-delay orbit), then one has a constraint on the neutron-star mass. So far, the only pulsar for which this has been possible is PSR J1012+5307 , whose companion is particularly bright ($`V20`$). The present constraints are shown in Fig. 3. One sees that for both systems, the mass determinations are at present not precise enough to provide meaningful additional constraints on the maximum mass a neutron star can have. The situation unfortunately is no better for determinations in low-mass X-ray binaries. A problem for those is that the X-ray sources generally do not pulse, and hence all information has to be gleaned from observations of the companions. So far, a meaningful determination has been possible only for Cyg X-2 . A relatively high mass is found, but again the uncertainties are rather large. Despite the above, prospects for progress appear brightest for further studies of systems in which the neutron star should have accreted a substantial amount of matter. Some other white-dwarf companions are bright enough, and more may be found in the on-going pulsar searches.
warning/0001/hep-th0001094.html
ar5iv
text
# USACH-FM-00-02 Bosonized supersymmetric quantum mechanics and supersymmetry of parabosons (parafermions)Talk given at the International Workshop “Supersymmetry and Quantum Symmetries”, JINR, Dubna, Russia, July 26-31,1999 ## Acknowledgements The work was supported in part by the grant 1980619 from FONDECYT (Chile) and by DICYT (USACH).
warning/0001/quant-ph0001042.html
ar5iv
text
# IS SUPERSYMMETRIC QUANTUM MECHANICS COMPATIBLE WITH DUALITY ? ## Abstract Supersymmetry applied to quantum mechanics has given new insights in various topics of theoretical physics like analytically solvable potentials, WKB approximation or KdV solitons. Duality plays a central role in many supersymmetric theories such as Yang-Mills theories or strings models. We investigate the possible existence of some duality within supersymmetric quantum mechanics. M. CAPDEQUI PEYRAN RE Laboratoire de Physique Math matique et Th orique CNRS - UMR 5825 Universit de Montpellier II, 34095 Montpellier Cedex.5 (France) PM/99-43. To be published in Modern Physics Letters A Introduction Supersymmetry is an exciting idea to relate bosons and fermions. Many particle physicists think that supersymmetry could be found in Nature, not too far from the electroweak scale, typically the TeV scale. From a theoretical point of view, supersymmetry naturally explains, for example, the large ratio between the electroweak scale and the other higher physical scales as the Planck or the GUT scales. Moreover, supersymmetry favours the unification for the three fundamental coupling constants responsible for the electromagnetic, the weak and the strong interactions. With the present data, such an unification is not possible within the Standard Model but becomes probable in its supersymmetric version like the MSSM (Minimal Supersymmetric Standard Model). Supersymmetry can also enrich quantum mechanics and gives the so called supersymmetric quantum mechanics (SQM) which is a nice framework to test new concepts in Physics. SQM is famous due to the pioneering papers of Witten on supersymmetry breaking. SQM also gives new understandings of the analytical solvability of certain potentials and of the (super)BKW approximation which turns out to be exact for a large class of potentials, or allows connections between isospectral hamiltonians and the multi-soliton solution of the Korteweg-De Vries non linear equation. Much more about SQM can be found in and references therein. Duality is a very fascinating concept whose original aim was to connect two different regimes of the same dynamics. The electric-magnetic duality is the most popular example where a weakly coupled version of a theory is exchanged for a strongly coupled formulation. The concept of duality has been extended and means an equivalence between two formulations of a model or between two different models. Such a case is the duality between the sine - Gordon model and the massive Thirring model , connecting a two-dimensional solitonic object and a two-dimensional fermionic field. That means that there is no guarantee at all that the connection is simple. Supersymmetry goes well with duality. As a matter of fact, some problems in duality invariant gauge theories disappear when supersymmetry comes in. This is exactely what happens in \[N=4\] Supersymmetric Yang-Mills theory . Supersymmetry is also an essential ingredient to study duality in string theories . In this paper we will elaborate on the following simple idea: is it possible to find a strong/weak coupling duality in the framework of supersymmetric quantum mechanics ? However clear the question may be, the definition remains vague. Thus we will discuss a bit more this point later, in the first section of this paper. In the second section we will develop duality in SQM with a precise definition. After the conclusion, we will work out in an appendix two examples to illustrate section II. Section I The idea we wish to explore is the mixing of duality and supersymmetric quantum mechanics. What does this mean? To be more precise, we will try to find duality in SQM and not to link by duality SQM to some other theory, and more exactly to stay in a model of SQM, and not to map a model of SQM with a strong coupling in another model of SQM with a weak coupling. Any model of SQM can be written in a superspace formalism like a field theory. Here we will use the hamiltonian formulation where the physics of the model lies in an hamiltonian $`H`$ which is a non negative operator. One has $`H=\{Q,Q^{}\}`$ where $`Q`$ and $`Q^{}`$, the “supercharges”, can be seen as operators transforming "bosons" into "fermions" and vice versa. Also any SQM hamiltonian can be cast in a matrix form as follows : $$H=\left|\begin{array}{cc}H_{}& 0\\ 0& H_+\end{array}\right|$$ These two hamiltonians $`H_{}`$ and $`H_+`$ are connected by supersymmetry which in turn gives a relation between the potential $`V_{}`$ and $`V_+`$ via the so called superpotential $`W`$ . The question is to know whether one of them can be strongly coupled and the other one weakly coupled. Let us denote by $`x`$ the (one dimensional) space variable and by g a real positive coupling constant. One has : $$V_{}(x,g)=W^2(x,g)W^{}(x,g)$$ and $$V_+(x,g)=W^2(x,g)+W^{}(x,g),$$ where $`W^{}=dW/dx.`$ Naively, if $`V_{}`$ is zero, the coupling could be said "weak" ; then $`V_+`$ will be different from zero and the coupling could be said "strong". The solution is $`W=1/(a+x)`$, where a is an arbitrary constant, and $`V_+`$ is obviously $`2/(a+x)^2`$. As a consequence of this rough example, it seems possible to see a path through the solution but perhaps not really interesting. Indeed, in this example, there are some drawbacks : \- $`g`$ does not appear at all ; \- the potential $`V_+`$ is a singular potential on the line ; \- the Hamiltonian $`H_{}`$ is just a second derivative, without any dynamical content. One can be a bit less naive in modifying our guess : is it possible to find a couple of supersymmetric potentials such as : $$V_{}(x,g)=gf(x,g)\text{and}V_+(x,g)=f(x,g)/g\mathrm{?}$$ (I.1) In such a way it seems that one forces, if $`g0`$, a weak coupling for $`V_{}`$ and a strong coupling for $`V_+`$. In solving the system (I.1), one successively gets the superpotential : $$W(x,g)=\frac{1}{s(a+x)},$$ and the potentials: $$V_{}(x,g)=\frac{(1s)}{s^2(a+x)^2},$$ and $$V_+(x,g)=\frac{(1+s)}{s^2(a+x)^2},$$ where $`s=\left(1g^2\right)/\left(1+g^2\right)`$ and a is an arbitrary constant. Although the potentials are still singular on a line, this result is quite interesting. Indeed, one obtains the following identity : $$V_{}(x,1/g)=V_+(x,g)$$ since s is changed in $`s`$ when one changes $`g`$ in $`1/g`$. This is a rather good feature. Now both potentials depend only on s and not on $`g`$, and this is a rather bad feature, because when g runs from zero to infinity, $`s`$ goes between $`1`$ and $`+1`$ and the strength of the couplings is somewhat washed off. This is a general consequence of our proposal to search for duality within the same SQM model. Actually, assuming there is only one coupling constant $`g(1)`$ in the model, the duality we look for can be written as follows : $$H_{}(x,1/g)=H_+(x,g)$$ (I.2) That means that a SQM model determined by the two hamiltonians $`(H_{}(x,1/g),H_+(x,1/g))`$, connected by susy thanks to the superpotential $`W(x,1/g)`$ in a regime of strong coupling (if $`g0`$), is exchanged in the SQM model determined by $`(H_+(x,g),H_{}(x,g))`$, which works in a weak coupling regime. The formula I.2 is just equivalent to $`V_{}(x,1/g)=V_+(x,g)`$, or : $$W^2(x,1/g)W^{}(x,1/g)=W^2(x,g)+W^{}(x,g).$$ One can re-write this formula as : $$W^2(x,1/g)W^2(x,g)=W^{}(x,1/g)+W^{}(x,g).$$ (I.3) When one changes $`g`$ in $`1/g`$ in this formula, the left hand side gets the opposite value while the right hand side is invariant. Hence both sides vanish and the unique superpotential solution of I.2 is such that: $$W(x,1/g)=W(x,g).$$ (I.4) So any function of $`x`$ and $`g`$ verifying I.4 is an answer to our question. If one looks for such functions, the simplest form is $`W(x,g)=a(x)b(g)`$, with the condition $$b(1/g)=b(g).$$ (I.5) The general solution of I.5 is $`b(g)=B(Log(g))`$ where $`B`$ is any odd function, but one can find rather simple solutions without logarithm. For example, $`b(g)=g1/g`$ is a solution but not suitable for a perturbative point a view (as $`Log(g)`$) since it has no expansion neither in $`g=0`$ nor in $`g=`$ infinity. Better solutions are $`b(g)=(1g)/(1+g)`$, which is $`tanh((Log(g))/2)`$, or $`b(g)=arctan(g)\pi /4`$. Anyway, the result (I.5) induces two remarks : \- the true coupling constant is $`b(g)`$ and not $`g`$. \- if the true coupling constant $`b(g)`$ belongs to the intervalle $`[c,d]`$, then $`b(1/g)`$ belongs to the intervalle $`[d,c]`$. The consequence is evident: it is impossible to talk about strong/weak duality in such a simple case. Of course, $`W(x,g)=a(x)b(g)`$ is far from being the most general solution of (I.4) but one cannot find other solutions which escape the previous consequence, because for a fixed $`x`$ the relation I.4 between $`W(x,g)`$ and $`W(x,1/g)`$ falls in a formula similar to I.5. Thus it is impossible to find a strong/weak duality in a same model of SQM, but one can search for a class of models where the couple $`(H_{},H_+)`$ is changed in the couple $`(H_+,H_{})`$, depending on the sign of the (true) coupling constant value. In these cases, the duality would mean a correspondence between two different regimes of the same SQM model. Section II In order to naturally introduce the class of models that we look for, we will continue to follow for a while the ingenuous path presented in the previous section. Further, we note that, for any quantum mechanical hamiltonian, there are several supersymmetric extensions. For example, starting from the harmonic oscillator hamiltonian $`H_0`$ ( in convenient units : $`2m=\mathrm{}=1`$ ) : $$H_0=d^2/dx^2+\omega ^2x^2,$$ one has $$H_{}=d^2/dx^2+\omega ^2x^2\omega ,$$ and $$H_+=d^2/dx^2+\omega ^2x^2+\omega ,$$ when the superpotential is $`W=\omega x`$ . But it is possible to obtain other superpotentials and hamiltonians $`H_{}`$ and $`H_+`$ (see for instance the second example of the Appendix ). So we will assume that we know a "free" hamiltonian $`H_0=d^2/dx^2+k(x)`$, where $`k(x)`$ is a fonction of $`x`$ and of some other parameters, and we add extra terms to simulate the weak and strong couplings as we did in I.1. We write : $$H_{}(x,g)=d^2/dx^2+k(x)+gf(x,g)$$ and $$H_+(x,g)=d^2/dx^2+k(x)+f(x,g)/g.$$ (II.1) In accordance with the idea of $`H_0`$ as a free hamiltonian, we suppose also that $`k(x)`$ is independent of $`g`$ ($`g1`$). One could probably imagine other forms that II.1 but the main virtue of our type is its solvability, without altering the physical meaning. One could also slightly modify the proposal II.1 by some sign or by multiplying $`f`$ by $`g`$ ; these changes are really minor, thus we will not consider them any longer . In solving II.1 in terms of the superpotential, one obtains the following Riccati equation : $$W^{}(x,g)=s\left(W^2(x,g)k(x)\right),$$ (II.2) where again we get the parameter $`s=\left(1g^2\right)/\left(1+g^2\right)`$. We use the usual trick to linearize the Riccati equation II.2 ; introducing a function $`y`$ of $`x`$ such as $`W=y^{}/(sy)`$, one readily gets the following Sturm - Liouville equation : $$y^{\prime \prime }s^2k(x)y=0.$$ (II.3) Once a solution $`y`$ is found for this equation II.3, we have the potentials solution of the problem (II.1) : $$V_{}(x,g)=+sk(x)+(1s)(y^{}/(sy))^2,$$ and $$V_+(x,g)=sk(x)+(1+s)(y^{}/(sy))^2.$$ (II.4) The supersymmetry is not broken if one can choose the sign of $`s`$ such that only one of the following sets of equations is true : $$H_{}(x,g)y^{(+1/s)}=0\text{and}H_+(x,g)y^{(+1/s)}0,$$ or $$H_+(x,g)y^{(1/s)}=0\text{and}H_{}(x,g)y^{(1/s)}0.$$ Furthermore to get discrete spectra for $`H_{}`$ and $`H_+`$, one needs that $`y^{(1/s)}`$ or $`y^{(1/s)}`$ be normalizable $`\left(L^2(x,R)\right)`$, depending on the signe of $`s`$. Then the corresponding hamiltonians $`H_{}`$ and $`H_+`$ will be isospectral (up to the lowest bound state). Let us be more precise about the solution II.4. At first sight, it seems that when we perform the following change : $`g1/g`$, then $`ss`$ and $`V_{}(x,g)V_{}(x,1/g)`$, which would be equal to $`V_+(x,g)`$. Actually, this is not true because it is necessary to take into account the eventuality that the boundary conditions of the equation II.3 depend on $`g`$. In such a way, $`y`$ would be $`g`$ dependent and $`V_{}(x,1/g)`$ different from $`V_+(x,g)`$. So, in general, $`V_{}(x,1/g)V_+(x,g)`$ and it is not suitable for duality ; fortunately there are simple situations where we get it. Assuming that $`k(x)`$ is an even function of $`x`$, then the second order differential equation II.3 (without first order term) has two independent solutions of given parity, one is even and the other is odd in the variable $`x`$, and both depend on $`s^2`$. The odd one is zero at $`x=0`$, so leading possibly to singular potentials. Thus it is highly preferable to use the even solution, and then the ratio $`y^{}/y`$ will be well defined and insensitive to a possible $`g`$ dependence coming from the boundary conditions. This is a case where we get the wished duality : \- first regime : $$0<s<1(0<g<1)$$ $$H_{}(x,s)=\frac{d^2}{dx^2}+V_{}(x,s)$$ $$V_{}(x,s)=+sk(x)+(1s)(y^{}/(sy))^2,$$ $$H_+(x,s)=\frac{d^2}{dx^2}+V_+(x,s)$$ $$V_+(x,s)=sk(x)+(1+s)(y^{}/(sy))^2,$$ and (for example) $`H_{}(x,s)y^{(+1/s)}=0`$ with $`H_+(x,s)y^{(+1/s)}0`$ and $`y^{(+1/s)}`$ is normalizable. In this regime, the ground state energy of $`H_{}`$ is zero and the other energy eigenvalues of both $`H_{}`$ and $`H_+`$ are paired. \- second regime : $$1<s<0(1<g<\mathrm{})$$ $$H_{}(x,s)=\frac{d^2}{dx^2}+V_{}(x,s)=H_+(x,s)$$ $$V_{}(x,s)=V_+(x,s)$$ $$H_+(x,s)=\frac{d^2}{dx^2}+V_+(x,s)=H_{}(x,s)$$ $$V_+(x,s)=V_{}(x,s),$$ and (for example) $`H_+(x,s)y^{(1/s)}=0`$ with $`H_{}(x,s)y^{(1/s)}0`$ and $`y^{(1/s)}`$ is normalizable. In this regime, the ground state energy of $`H_+`$ is zero, and the other energy eigenvalues of both $`H_{}`$ and $`H_+`$ are paired. In other words, this duality exchanges the role of $`H_+`$ and $`H_{}`$, when one goes from the positive s regime to the negative s regime. Roughly speaking, one could say that this duality exchanges the "boson" spectrum and the "fermion" spectrum (see figure 1). In the appendix, we work out two illustrative examples, one based on $`k(x)=`$ constant and the other one on the harmonic oscillator. Finally, thanks to a theorem of Poincaré one can prove that the solutions of the Sturm - Liouville equation II.3 are analytic in $`s^2`$, whatever $`k(x)`$ is, if the boundary conditions are independent of $`s`$. In these situations, the ratio $`y^{}/y`$ will be $`s^2`$ dependent and $`V_{}(x,s)`$ will be equal to $`V_+(x,s)`$. That extends the cases of duality. Conclusion In this study we have shown that it is impossible to find a strong/weak duality within a model of Supersymmetric Quantum Mechanics and we have developed a strategy to find a class of models of SQM where the duality means a correspondence between two regimes of the coupling constant. Acknowledgments I would like to thank Guy Auberson (LPMT), Avinash Khare (Bhubaneswar Institute of Physics, India) and Michel Talon (LPTHE, Paris VI) for their enlightening remarks and encouragements. I am also grateful to members of the LPMT for many discussions, especially Christophe Le Mouel, Gerard Mennessier and Gilbert Moultaka. References (1) E. Witten, Nucl.Phys. B 185 (1981) 513 ; Nucl.Phys. B 202 (1982) 253. (2) A. Lahiri, P.K. Roy and B. Bagchi, Int.J.Mod.Phys. A Vol.5 No.8 (1990) 1383. (3) F.Cooper, A. Khare and U. Sukhatme, Phys.Rep. 251 (1995) 267. (4) S. Coleman, Phys.Rev.D 11(1975) 2088. (5) C. Wafa and E. Witten, Nucl.Phys. B 431 (1994) 3. (6) C. Hull and P. Townsend, Nucl.Phys. B 438 (1995) 109. (7) V. De Alfaro and T. Regge, Potential Scattering, North Holland Publishing Company (1965), page 9. Appendix Example 1 : $`k(\omega )=\omega ^2(\omega >0)`$ The master equation II.3 reads : $$y^{\prime \prime }s^2\omega ^2y=0,$$ and an even solution is $`\mathrm{cosh}(s\omega x)`$. One gets the potentials : $$V_{}(x,s)=\omega ^2\left(+s+(1s)(\mathrm{tanh}(s\omega x))^2\right)$$ and $$V_+(x,s)=\omega ^2\left(s+(1+s)(\mathrm{tanh}(s\omega x))^2\right).$$ (A.1) It is easy to check that $`H_{}(x,s)`$ $`\left(\mathrm{cosh}(s\omega x)\right)^{\left(\frac{1}{s}\right)}=0`$ and that $`(\mathrm{cosh}(s\omega x))^{\left(\frac{1}{s}\right)}`$ is normalizable for negative $`s`$. The duality is guaranteed since $`V_{}(x,s)=V_+(x,s)`$. The surprise in this example is that it is quite easy to find the full (discrete) spectra of the hamiltonians because these potentials are in the S.I.P class (Shape Invariant Potential) . Indeed, potentials in A.1 can be derived from the superpotential $`W=a\mathrm{tanh}(bx)`$ when $`a=\omega `$ and $`b=s\omega `$. We refer to the reference for details. One finds that the energy levels are given by $$E_n(s,\omega )=ns\omega ^2(2+ns).$$ Now, since $`s`$ is negative and $`n`$ is a non negative integer, and because $`E_n`$ must be an increasing function of $`n`$, one gets that $`n`$ belongs to the interval $`[0,E(1/s)]`$. \[$`E(x)`$ $``$ greatest integer less than or equal to $`x`$\]. For example, if $`s=1/4,H_{}`$ has 5 eigenvalues $`(0,\omega ^2\frac{7}{16},\omega ^2\frac{12}{16},\omega ^2\frac{15}{16},\omega ^2)`$ and $`H_+`$ has 4 eigenvalues $`(\omega ^2\frac{7}{16},\omega ^2\frac{12}{16},\omega ^2\frac{15}{16},\omega ^2)`$. See figure A1. If one chooses $`k(x)=\omega ^2`$, one gets trigonometric functions instead of hyperbolic functions. With this choice, $`x`$ must be in an interval such as $`]\pi /2,+\pi /2[`$. Example 2 : $`k(x)=\omega ^2x^2(\omega >0)`$ The master equation II.3 reads: $$y^{\prime \prime }s^2\omega ^2x^2y=0,$$ and an even solution is given in terms of a Bessel function : $$\sqrt{\left|x\right|}\text{Bessel}\text{I}[\frac{1}{4},(x^2\sqrt{\omega ^2s^2)}/2]$$ which has the following expansion (up to an overall numerical factor) $$1+\underset{p=1}{\overset{\mathrm{}}{}}\left(\frac{s^{2p}\omega ^{2p}x^{4p}}{c_p}\right)$$ where $`c_p=((3.4)(7.8)\mathrm{}((4p1).(4p)))`$. Handling with care, the potentials read : $$V_{}(x,s)=+s\omega x^2+(1s\omega )x^2\left(\frac{\text{Bessel}\text{I}[+\frac{3}{4},(s\omega x^2)/2]}{\text{Bessel}\text{I}[\frac{1}{4},(s\omega x^2)/2]}\right)^2$$ and $$V_+(x,s)=s\omega x^2+(1+s\omega )x^2\left(\frac{\text{Bessel}\text{I}[+\frac{3}{4},(s\omega x^2)/2]}{\text{Bessel}\text{I}[\frac{1}{4},(s\omega x^2)/2]}\right)^2$$ (A.2) Depending on $`s`$, one obtains the same behaviour as in the example 1 and the duality is got. The potentials are (probably) not S.I.P . In figure A2 are drawn these potentials for $`s=1/3`$ and $`\omega =1`$. For these values, $`V_{}`$ shows two degenerate classically stable minima and one classically unstable maximum. Tunneling effects in quantum mechanics make that the energy of the ground state, which is of course $`0`$ for supersymmetric reasons, is at the level of the classical unstable extremum. Figure Captions Figure 1 : typical discrete spectrum of a model in the negative s regime on the left side and in the positive s regime on the right side showing the "Bosonic" and "Fermionic" states . Figure A1 : plots of $`V_{}`$(thick) and $`V_+`$ (thin) and the energy eigenvalues (example A1 of the Appendix ; $`\omega =2,s=1/4`$). Figure A2 : plots of $`V_{}`$ (thick) and $`V_+`$(thin) and the ground state energy (example A2 of the Appendix ; $`\omega =1,s=1/3`$). B F B F Figure 1 Figure A1 Figure A2
warning/0001/hep-th0001122.html
ar5iv
text
# 1 Introduction ## 1 Introduction AdS/CFT correspondence may be realized in a sufficiently simple form as d5 gauged supergravity/boundary gauge theory correspondence. The reason is very simple: different versions of five-dimensional gauged SG (for example, $`N=8`$ gauged SG which contains 42 scalars and non-trivial scalar potential) could be obtained as compactification (reduction) of ten-dimensional IIB SG. Then, in practice it is enough to consider 5d gauged SG classical solutions (say, AdS-like backgrounds) in AdS/CFT set-up instead of the investigation of much more involved, non-linear equations of IIB SG. Moreover, such solutions describe RG flows in boundary gauge theory (for a very recent discussion of such flows see and refs. therein). To simplify the situation in extended SG one can consider the symmetric (special) RG flows where scalars lie in one-dimensional submanifold of total space. Then, such theory is effectively described as d5 dilatonic gravity with non-trivial dilatonic potential. Nevertheless, it is still extremely difficult to make the explicit identification of deformed SG solution with the dual (non-conformal exactly) gauge theory. As a rule , only indirect arguments may be suggested in such identification<sup>4</sup><sup>4</sup>4Such dual theory in massless case is, of course, classically conformally invariant and it has well-defined conformal anomaly. However, among the interacting theories only $`𝒩=4`$ SYM is known to be exactly conformally invariant. Its conformal anomaly is not renormalized. For other, d4 QFTs there is breaking of conformal invariance due to radiative corrections which give contribution also to conformal anomaly. Hence, one can call such theories as non-conformal ones or not exactly conformally invariant. The conformal anomalies for such theories are explicitly unknown. Only for few simple theories (like scalar QED or gauge theory without fermions) the calculation of radiative corrections to conformal anomaly has been done up to two or three loops. It is a challenge to find exact conformal anomaly. Presumbly, only SG description may help to resolve this problem.. From another side, the fundamental holographic principle in AdS/CFT form enriches the classical gravity itself (and here also classical gauged SG). Indeed, instead of the standard subtraction of reference background in making the gravitational action finite and the quasilocal stress tensor well-defined one introduces more elegant, local surface counterterm prescription . Within it one adds the coordinate invariant functional of the intrinsic boundary geometry to gravitational action. Clearly, that does not modify the equations of motion. Moreover, this procedure has nice interpretation in terms of dual QFT as standard regularization. The specific choice of surface counterterm cancels the divergences of bulk gravitational action. As a by-product, it also defines the conformal anomaly of boundary QFT. Local surface counterterm prescription has been successfully applied to construction of finite action and quasilocal stress tensor on asymptotically AdS space in Einstein gravity and in higher derivative gravity . Moreover, the generalization to asymptotically flat spaces is possible as it was first mentioned in ref.. Surface counterterm has been found for domain-wall black holes in gauged SG in diverse dimensions . However, actually only the case of asymptotically constant dilaton has been investigated there. In the present paper we discuss the construction of finite action, consistent gravitational stress tensor and dilaton-dependent Weyl anomaly for boundary QFT (from bulk side) in three- and five-dimensional gauged supergravity with single scalar (dilaton) on asymptotically AdS background. Note that dilaton is not constant and the potential is chosen to be arbitrary. The implications of results for the study of RG flows in boundary QFT are presented, in particular, the candidate c-function is suggested. The next section is devoted to the evaluation of Weyl anomaly from gauged supergravity with arbitrary dilatonic potential via AdS/CFT correspondence. We present explicit result for d3 and d5 gauged SGs. Such SG side conformal anomaly should correspond to dual QFT with broken conformal invariance in two and in four dimensions, respectively. The explicit form of d4 conformal anomaly takes few pages, so its lengthy dilaton-dependent coefficients are listed in Appendix. The comparison with similar AdS/CFT calculation of conformal anomaly in the same theory but with constant dilatonic potential is given. The candidates for c-function in two and four dimensions are proposed. Section three is devoted to presentation of acceptable proposal for candidate c-fnction given in terms of dilatonic potential. It is shown that for numberof potentials such c-function is monotonic and positively defined. It has fixed point in asymptotically AdS region. The comparison with other c-functions is given. In section four we construct surface counterterms for d3 and d5 gauged SGs. As a result, the gravitational action in asymptotically AdS space is finite. On the same time, the gravitational stress tensor around such space is well defined. It is interesting that conformal anomaly defined in second section directly follows from the gravitational stress tensor with account of surface terms. Section five is devoted to the application of finite gravitational action found in previous section in the calculation of thermodynamical quantities in dilatonic AdS black hole. The dilatonic AdS black hole is constructed approximately, using the perturbations around constant dilaton AdS black hole. The entropy, mass and free energy of such black hole are found using the local surface counterterm prescription to regularize these quantities. The comparison is done with the case when standard prescription: regularization with reference background is used. The explicit regularization dependence of the result is mentioned. Finally, in the Discussion the summary of results is presented and some open problems are mentioned. ## 2 Weyl anomaly for gauged supergravity with general dilaton potential In the present section the derivation of dilaton-dependent Weyl anomaly from gauged SG will be given. As we note in section 4 this derivation can be made also from the definition of finite action in asymptotically AdS space. We start from the bulk action of $`d+1`$-dimensional dilatonic gravity with the potential $`\mathrm{\Phi }`$ $$S=\frac{1}{16\pi G}_{M_{d+1}}d^{d+1}x\sqrt{\widehat{G}}\left\{\widehat{R}+X(\varphi )(\widehat{}\varphi )^2+Y(\varphi )\widehat{\mathrm{\Delta }}\varphi +\mathrm{\Phi }(\varphi )+4\lambda ^2\right\}.$$ (1) Here $`M_{d+1}`$ is $`d+1`$ dimensional manifold whose boundary is $`d`$ dimensional manifold $`M_d`$ and we choose $`\mathrm{\Phi }(0)=0`$. Such action corresponds to (bosonic sector) of gauged SG with single scalar (special RG flow). In other words, one considers RG flow in extended SG when scalars lie in one-dimensional submanifold of complete scalars space. Note also that classical vacuum stability restricts the form of dilaton potential . As well-known, we also need to add the surface terms to the bulk action in order to have well-defined variational principle. At the moment, for the purpose of calculation of Weyl anomaly (via AdS/CFT correspondence) the surface terms are irrelevant. The equations of motion given by variation of (1) with respect to $`\varphi `$ and $`G^{\mu \nu }`$ are $`0`$ $`=`$ $`\sqrt{\widehat{G}}\mathrm{\Phi }^{}(\varphi )\sqrt{\widehat{G}}V^{}(\varphi )\widehat{G}^{\mu \nu }_\mu \varphi _\nu \varphi `$ (2) $`+2_\mu \left(\sqrt{\widehat{G}}\widehat{G}^{\mu \nu }V(\varphi )_\nu \varphi \right)`$ $`0`$ $`=`$ $`{\displaystyle \frac{1}{d1}}\widehat{G}_{\mu \nu }\left(\mathrm{\Phi }(\varphi )+{\displaystyle \frac{d(d1)}{l^2}}\right)+\widehat{R}_{\mu \nu }+V(\varphi )_\mu \varphi _\nu \varphi .`$ (3) Here $$V(\varphi )X(\varphi )Y^{}(\varphi ).$$ (4) We choose the metric $`\widehat{G}_{\mu \nu }`$ on $`M_{d+1}`$ and the metric $`\widehat{g}_{\mu \nu }`$ on $`M_d`$ in the following form $$ds^2\widehat{G}_{\mu \nu }dx^\mu dx^\nu =\frac{l^2}{4}\rho ^2d\rho d\rho +\underset{i=1}{\overset{d}{}}\widehat{g}_{ij}dx^idx^j,\widehat{g}_{ij}=\rho ^1g_{ij}.$$ (5) Here $`l`$ is related with $`\lambda ^2`$ by $`4\lambda ^2=d(d1)/l^2`$. If $`g_{ij}=\eta _{ij}`$, the boundary of AdS lies at $`\rho =0`$. We follow to method of calculation of conformal anomaly as it was done in refs. where dilatonic gravity with constant dilaton potential has been considered. Part of results of this section concerning Weyl anomaly with no dilaton derivatives has been presented already in letter . The action (1) diverges in general since it contains the infinite volume integration on $`M_{d+1}`$. The action is regularized by introducing the infrared cutoff $`ϵ`$ and replacing $$d^{d+1}xd^dx_ϵ𝑑\rho ,_{M_d}d^dx\left(\mathrm{}\right)d^dx\left(\mathrm{}\right)|_{\rho =ϵ}.$$ (6) We also expand $`g_{ij}`$ and $`\varphi `$ with respect to $`\rho `$: $$g_{ij}=g_{(0)ij}+\rho g_{(1)ij}+\rho ^2g_{(2)ij}+\mathrm{},\varphi =\varphi _{(0)}+\rho \varphi _{(1)}+\rho ^2\varphi _{(2)}+\mathrm{}.$$ (7) Then the action is also expanded as a power series on $`\rho `$. The subtraction of the terms proportional to the inverse power of $`ϵ`$ does not break the invariance under the scale transformation $`\delta g_{\mu \nu }=2\delta \sigma g_{\mu \nu }`$ and $`\delta ϵ=2\delta \sigma ϵ`$ . When $`d`$ is even, however, the term proportional to $`\mathrm{ln}ϵ`$ appears. This term is not invariant under the scale transformation and the subtraction of the $`\mathrm{ln}ϵ`$ term breaks the invariance. The variation of the $`\mathrm{ln}ϵ`$ term under the scale transformation is finite when $`ϵ0`$ and should be canceled by the variation of the finite term (which does not depend on $`ϵ`$) in the action since the original action (1) is invariant under the scale transformation. Therefore the $`\mathrm{ln}ϵ`$ term $`S_{\mathrm{ln}}`$ gives the Weyl anomaly $`T`$ of the action renormalized by the subtraction of the terms which diverge when $`ϵ0`$ (d=4) $$S_{\mathrm{ln}}=\frac{1}{2}d^4x\sqrt{g}T.$$ (8) The conformal anomaly can be also obtained from the surface counterterms, which is discussed in Section 4. First we consider the case of $`d=2`$, i.e. three-dimensional gauged SG. The anomaly term $`S_{\mathrm{ln}}`$ proportional to $`\mathrm{ln}ϵ`$ in the action is $`S_{\mathrm{ln}}={\displaystyle \frac{1}{16\pi G}}{\displaystyle \frac{l}{2}}{\displaystyle }d^2x\sqrt{g_{(0)}}\{R_{(0)}+X(\varphi _{(0)})(\varphi _{(0)})^2+Y(\varphi _{(0)})\mathrm{\Delta }\varphi _{(0)}`$ $`+\varphi _{(1)}\mathrm{\Phi }^{}(\varphi _{(0)})+{\displaystyle \frac{1}{2}}g_{(0)}^{ij}g_{(1)ij}\mathrm{\Phi }(\varphi _{(0)})\}.`$ (9) The terms proportional to $`\rho ^0`$ with $`\mu ,\nu =i,j`$ in (3) lead to $`g_{(1)ij}`$ in terms of $`g_{(0)ij}`$ and $`\varphi _{(1)}`$. $`g_{(1)ij}`$ $`=`$ $`[R_{(0)ij}V(\varphi _{(0)})_i\varphi _{(0)}_j\varphi _{(0)}g_{(0)ij}\mathrm{\Phi }^{}(\varphi _{(0)})\varphi _{(1)}`$ (10) $`+{\displaystyle \frac{g_{(0)ij}}{l^2}}\left\{2\mathrm{\Phi }^{}(\varphi _{(0)})\varphi _{(1)}+R_{(0)}+V(\varphi _{(0)})g_{(0)}^{kl}_k\varphi _{(0)}_l\varphi _{(0)}\right\}`$ $`\times (\mathrm{\Phi }(\varphi _{(0)})+{\displaystyle \frac{2}{l^2}})^1]\times \mathrm{\Phi }(\varphi _{(0)})^1`$ In the equation (2), the terms proportional to $`\rho ^1`$ lead to $`\varphi _{(1)}`$ as following. $`\varphi _{(1)}`$ $`=`$ $`[V^{}(\varphi _{(0)})g_{(0)}^{ij}_i\varphi _{(0)}_j\varphi _{(0)}+2{\displaystyle \frac{V(\varphi _{(0)})}{\sqrt{g_{(0)}}}}_i\left(\sqrt{g_{(0)}}g_{(0)}^{ij}_j\varphi _{(0)}\right)`$ (11) $`+{\displaystyle \frac{1}{2}}\mathrm{\Phi }^{}(\varphi _{(0)})(\mathrm{\Phi }(\varphi _{(0)})+{\displaystyle \frac{2}{l^2}})^1\{R_{(0)}+V(\varphi _{(0)})g_{(0)}^{ij}_i\varphi _{(0)}_j\varphi _{(0)}\}]`$ $`\times \left(\mathrm{\Phi }^{\prime \prime }(\varphi _{(0)})\mathrm{\Phi }^{}(\varphi _{(0)})^2\left(\mathrm{\Phi }(\varphi _{(0)})+{\displaystyle \frac{2}{l^2}}\right)^1\right)^1`$ Then anomaly term takes the following form using (10), (11) $`T`$ $`=`$ $`{\displaystyle \frac{1}{8\pi G}}{\displaystyle \frac{l}{2}}\{R_{(0)}+X(\varphi _{(0)})(\varphi _{(0)})^2+Y(\varphi _{(0)})\mathrm{\Delta }\varphi _{(0)}`$ (12) $`+{\displaystyle \frac{1}{2}}\left\{{\displaystyle \frac{2\mathrm{\Phi }^{}(\varphi _{(0)})}{l^2}}\left(\mathrm{\Phi }^{\prime \prime }(\varphi _{(0)})\left(\mathrm{\Phi }(\varphi _{(0)})+{\displaystyle \frac{2}{l^2}}\right)\mathrm{\Phi }^{}(\varphi _{(0)})^2\right)^1\mathrm{\Phi }(\varphi _{(0)})\right\}`$ $`\times \left(R_{(0)}+V(\varphi _{(0)})g_{(0)}^{ij}_i\varphi _{(0)}_j\varphi _{(0)}\right)\left(\mathrm{\Phi }(\varphi _{(0)})+{\displaystyle \frac{2}{l^2}}\right)^1`$ $`+{\displaystyle \frac{2\mathrm{\Phi }^{}(\varphi _{(0)})}{l^2}}\left(\mathrm{\Phi }^{\prime \prime }(\varphi _{(0)})\left(\mathrm{\Phi }(\varphi _{(0)})+{\displaystyle \frac{2}{l^2}}\right)\mathrm{\Phi }^{}(\varphi _{(0)})^2\right)^1`$ $`\times (V^{}(\varphi _{(0)})g_{(0)}^{ij}_i\varphi _{(0)}_j\varphi _{(0)}+2{\displaystyle \frac{V(\varphi _{(0)})}{\sqrt{g_{(0)}}}}_i\left(\sqrt{g_{(0)}}g_{(0)}^{ij}_j\varphi _{(0)}\right))\}.`$ For $`\mathrm{\Phi }(\varphi )=0`$ case, the central charge of two-dimensional conformal field theory is defined by the coefficient of $`R`$. Then it might be natural to introduce the candidate c-function $`c`$ for the case when the conformal symmetry is broken by the deformation in the following way : $`c`$ $`=`$ $`{\displaystyle \frac{3}{2G}}[l+{\displaystyle \frac{l}{2}}\{{\displaystyle \frac{2\mathrm{\Phi }^{}(\varphi _{(0)})}{l^2}}(\mathrm{\Phi }^{\prime \prime }(\varphi _{(0)})(\mathrm{\Phi }(\varphi _{(0)})`$ (13) $`+{\displaystyle \frac{2}{l^2}})\mathrm{\Phi }^{}(\varphi _{(0)})^2)^1\mathrm{\Phi }(\varphi _{(0)})\}\times (\mathrm{\Phi }(\varphi _{(0)})+{\displaystyle \frac{2}{l^2}})^1].`$ Comparing this with radiatively-corrected c-function of boundary QFT ($`\mathrm{AdS}_3/\mathrm{CFT}_2`$) may help in correct bulk description of such theory. Clearly, that in the regions (or for potentials) where such candidate c-function is singular or not monotonic it cannot be the acceptable c-function. Presumbly, the appearence of such regions indicates to the breaking of SG description. Four-dimensional case is more interesting but also much more involved. The anomaly terms which proportional to $`\mathrm{ln}ϵ`$ are $`S_{\mathrm{ln}}`$ $`=`$ $`{\displaystyle \frac{1}{16\pi G}}{\displaystyle }d^4x\sqrt{g_{(0)}}[{\displaystyle \frac{1}{2l}}g_{(0)}^{ij}g_{(0)}^{kl}(g_{(1)ij}g_{(1)kl}g_{(1)ik}g_{(1)jl})`$ $`+{\displaystyle \frac{l}{2}}\left(R_{(0)}^{ij}{\displaystyle \frac{1}{2}}g_{(0)}^{ij}R_{(0)}\right)g_{(1)ij}`$ $`{\displaystyle \frac{2}{l}}V(\varphi _{(0)})\varphi _{(1)}^2+{\displaystyle \frac{l}{2}}V^{}(\varphi _{(0)})\varphi _{(1)}g_{(0)}^{ij}_i\varphi _{(0)}_j\varphi _{(0)}`$ $`+lV(\varphi _{(0)})\varphi _{(1)}{\displaystyle \frac{1}{\sqrt{g_{(0)}}}}_i\left(\sqrt{g_{(0)}}g_{(0)}^{ij}_j\varphi _{(0)}\right)`$ $`+{\displaystyle \frac{l}{2}}V(\varphi _{(0)})\left(g_{(0)}^{ik}g_{(0)}^{jl}g_{(1)kl}{\displaystyle \frac{1}{2}}g_{(0)}^{kl}g_{(1)kl}g_{(0)}^{ij}\right)_i\varphi _{(0)}_j\varphi _{(0)}`$ $`{\displaystyle \frac{l}{2}}\left({\displaystyle \frac{1}{2}}g_{(0)}^{ij}g_{(2)ij}{\displaystyle \frac{1}{4}}g_{(0)}^{ij}g_{(0)}^{kl}g_{(1)ik}g_{(1)jl}+{\displaystyle \frac{1}{8}}(g_{(0)}^{ij}g_{(1)ij})^2\right)\mathrm{\Phi }(\varphi _{(0)})`$ $`{\displaystyle \frac{l}{2}}(\mathrm{\Phi }^{}(\varphi _{(0)})\varphi _{(2)}+{\displaystyle \frac{1}{2}}\mathrm{\Phi }^{\prime \prime }(\varphi _{(0)})\varphi _{(1)}^2+{\displaystyle \frac{1}{2}}g_{(0)}^{kl}g_{(1)kl}\mathrm{\Phi }^{}(\varphi _{(0)})\varphi _{(1)})].`$ The terms proportional to $`\rho ^0`$ with $`\mu ,\nu =i,j`$ in the equation of the motion (3) lead to $`g_{(1)ij}`$ in terms of $`g_{(0)ij}`$ and $`\varphi _{(1)}`$. $`g_{(1)ij}`$ $`=`$ $`[R_{(0)ij}V(\varphi _{(0)})_i\varphi _{(0)}_j\varphi _{(0)}{\displaystyle \frac{1}{3}}g_{(0)ij}\mathrm{\Phi }^{}(\varphi _{(0)})\varphi _{(1)}`$ (15) $`+{\displaystyle \frac{g_{(0)ij}}{l^2}}\left\{{\displaystyle \frac{4}{3}}\mathrm{\Phi }^{}(\varphi _{(0)})\varphi _{(1)}+R_{(0)}+V(\varphi _{(0)})g_{(0)}^{kl}_k\varphi _{(0)}_l\varphi _{(0)}\right\}`$ $`\times ({\displaystyle \frac{1}{3}}\mathrm{\Phi }(\varphi _{(0)})+{\displaystyle \frac{6}{l^2}})^1]\times ({\displaystyle \frac{1}{3}}\mathrm{\Phi }(\varphi _{(0)})+{\displaystyle \frac{2}{l^2}})^1.`$ In the equation (2), the terms proportional to $`\rho ^2`$ lead to $`\varphi _{(1)}`$ as follows: $`\varphi _{(1)}`$ $`=`$ $`[V^{}(\varphi _{(0)})g_{(0)}^{ij}_i\varphi _{(0)}_j\varphi _{(0)}+2{\displaystyle \frac{V(\varphi _{(0)})}{\sqrt{g_{(0)}}}}_i\left(\sqrt{g_{(0)}}g_{(0)}^{ij}_j\varphi _{(0)}\right)`$ (16) $`+{\displaystyle \frac{1}{2}}\mathrm{\Phi }^{}(\varphi _{(0)})({\displaystyle \frac{1}{3}}\mathrm{\Phi }(\varphi _{(0)})+{\displaystyle \frac{6}{l^2}})^1\{R_{(0)}+V(\varphi _{(0)})g_{(0)}^{ij}_i\varphi _{(0)}_j\varphi _{(0)}\}]`$ $`\times \left({\displaystyle \frac{8V(\varphi _{(0)})}{l^2}}+\mathrm{\Phi }^{\prime \prime }(\varphi _{(0)}){\displaystyle \frac{2}{3}}\mathrm{\Phi }^{}(\varphi _{(0)})^2\left({\displaystyle \frac{1}{3}}\mathrm{\Phi }(\varphi _{(0)})+{\displaystyle \frac{6}{l^2}}\right)^1\right)^1.`$ In the equation (3), the terms proportional to $`\rho ^1`$ with $`\mu ,\nu =i,j`$ lead to $`g_{(2)ij}`$. $`g_{(2)ij}`$ $`=`$ $`[{\displaystyle \frac{1}{3}}\{g_{(1)ij}\mathrm{\Phi }^{}(\varphi _{(0)})\varphi _{(1)}+g_{(0)ij}(\mathrm{\Phi }^{}(\varphi _{(0)})\varphi _{(2)}+{\displaystyle \frac{1}{2}}\mathrm{\Phi }^{\prime \prime }(\varphi _{(0)})\varphi _{(1)}^2)\}`$ (17) $`{\displaystyle \frac{2}{l^2}}g_{(0)}^{kl}g_{(1)ki}g_{(1)lj}+{\displaystyle \frac{1}{l^2}}g_{(0)}^{km}g_{(0)}^{nl}g_{(1)mn}g_{(1)kl}g_{(0)ij}`$ $`{\displaystyle \frac{2}{l^2}}g_{(0)ij}({\displaystyle \frac{1}{3}}\mathrm{\Phi }(\varphi _{(0)})+{\displaystyle \frac{8}{l^2}})^1\times \{{\displaystyle \frac{2}{l^2}}g_{(0)}^{mn}g_{(0)}^{kl}g_{(1)km}g_{(1)ln}`$ $`{\displaystyle \frac{4}{3}}\left(\mathrm{\Phi }^{}(\varphi _{(0)})\varphi _{(2)}+{\displaystyle \frac{1}{2}}\mathrm{\Phi }^{\prime \prime }(\varphi _{(0)})\varphi _{(1)}^2\right){\displaystyle \frac{1}{3}}g_{(0)}^{ij}g_{(1)ij}\mathrm{\Phi }^{}(\varphi _{(0)})\varphi _{(1)}`$ $`+V^{}(\varphi _{(0)})\varphi _{(1)}g_{(0)}^{ij}_i\varphi _{(0)}_j\varphi _{(0)}+{\displaystyle \frac{2V(\varphi _{(0)})\varphi _{(1)}}{\sqrt{g_{(0)}}}}_i\left(\sqrt{g_{(0)}}g_{(0)}^{ij}_j\varphi _{(0)}\right)\}`$ $`+V^{}(\varphi _{(0)})\varphi _{(1)}_i\varphi _{(0)}_j\varphi _{(0)}+2V(\varphi _{(0)})\varphi _{(1)}_i_j\varphi _{(0)}]`$ $`\times \left({\displaystyle \frac{1}{3}}\mathrm{\Phi }(\varphi _{(0)})\right)^1.`$ And the terms proportional to $`\rho ^1`$ in the equation (2), lead to $`\varphi _{(2)}`$ as follows: $`\varphi _{(2)}`$ $`=`$ $`[V^{\prime \prime }(\varphi _{(0)})\varphi _{(1)}g_{(0)}^{ij}_i\varphi _{(0)}_j\varphi _{(0)}`$ (18) $`+V^{}(\varphi _{(0)})\left(g_{(0)}^{ik}g_{(0)}^{jl}{\displaystyle \frac{1}{2}}g_{(0)}^{ij}g_{(0)}^{kl}\right)g_{(1)kl}_i\varphi _{(0)}_j\varphi _{(0)}`$ $`+{\displaystyle \frac{2V^{}(\varphi _{(0)})\varphi _{(1)}}{\sqrt{g_{(0)}}}}_i\left(\sqrt{g_{(0)}}g_{(0)}^{ij}_j\varphi _{(0)}\right)`$ $`{\displaystyle \frac{4}{l^2}}V_{(0)}^{}\varphi _{(1)}^2{\displaystyle \frac{1}{2}}\mathrm{\Phi }^{\prime \prime \prime }(\varphi _{(0)})\varphi _{(1)}^2{\displaystyle \frac{1}{2}}g_{(0)}^{kl}g_{(1)kl}\mathrm{\Phi }^{\prime \prime }(\varphi _{(0)})\varphi _{(1)}`$ $`\left({\displaystyle \frac{1}{4}}g_{(0)}^{ij}g_{(0)}^{kl}g_{(1)ik}g_{(1)jl}+{\displaystyle \frac{1}{8}}(g_{(0)}^{ij}g_{(1)ij})^2\right)\mathrm{\Phi }^{}(\varphi _{(0)})`$ $`{\displaystyle \frac{1}{2}}\mathrm{\Phi }^{}(\varphi _{(0)})({\displaystyle \frac{1}{3}}\mathrm{\Phi }(\varphi _{(0)})+{\displaystyle \frac{8}{l^2}})^1\times \{{\displaystyle \frac{2}{l^2}}g_{(0)}^{mn}g_{(0)}^{kl}g_{(1)km}g_{(1)ln}`$ $`{\displaystyle \frac{2}{3}}\mathrm{\Phi }^{\prime \prime }(\varphi _{(0)})\varphi _{(1)}^2{\displaystyle \frac{1}{3}}g_{(0)}^{ij}g_{(1)ij}\mathrm{\Phi }^{}(\varphi _{(0)})\varphi _{(1)}`$ $`+V^{}(\varphi _{(0)})\varphi _{(1)}g_{(0)}^{ij}_i\varphi _{(0)}_j\varphi _{(0)}+{\displaystyle \frac{2V(\varphi _{(0)})\varphi _{(1)}}{\sqrt{g_{(0)}}}}_i\left(\sqrt{g_{(0)}}g_{(0)}^{ij}_j\varphi _{(0)}\right)\}]`$ $`\times \left(\mathrm{\Phi }^{\prime \prime }(\varphi _{(0)}){\displaystyle \frac{2}{3}}\mathrm{\Phi }^{}(\varphi _{(0)})^2\left({\displaystyle \frac{1}{3}}\mathrm{\Phi }(\varphi _{(0)})+{\displaystyle \frac{8}{l^2}}\right)^1\right)^1`$ Then we can get the anomaly (2) in terms of $`g_{(0)ij}`$ and $`\varphi _{(0)}`$, which are boundary values of metric and dilaton respectively by using (15), (16), (17), (18). In the following, we choose $`l=1`$, denote $`\mathrm{\Phi }(\varphi _{(0)})`$ by $`\mathrm{\Phi }`$ and abbreviate the index $`(0)`$ for the simplicity. Then substituting (16) into (15), we obtain $`g_{(1)ij}`$ $`=`$ $`\stackrel{~}{c}_1R_{ij}+\stackrel{~}{c}_2g_{ij}R+\stackrel{~}{c}_3g_{ij}g^{kl}_k\varphi _l\varphi `$ (19) $`+\stackrel{~}{c}_4g_{ij}{\displaystyle \frac{_k}{\sqrt{g}}}\left(\sqrt{g}g^{kl}_l\varphi \right)+\stackrel{~}{c}_5_i\varphi _j\varphi .`$ The explicit form of $`\stackrel{~}{c}_1`$, $`\stackrel{~}{c}_2`$, $`\mathrm{}`$ $`\stackrel{~}{c}_5`$ is given in Appendix A. Further, substituting (16) and (19) into (18), one gets $`\varphi _{(2)}`$ $`=`$ $`d_1R^2+d_2R_{ij}R^{ij}+d_3R^{ij}_i\varphi _j\varphi `$ (20) $`+d_4Rg^{ij}_i\varphi _j\varphi +d_5R{\displaystyle \frac{1}{\sqrt{g}}}_i(\sqrt{g}g^{ij}_j\varphi )`$ $`+d_6(g^{ij}_i\varphi _j\varphi )^2+d_7\left({\displaystyle \frac{1}{\sqrt{g}}}_i(\sqrt{g}g^{ij}_j\varphi )\right)^2`$ $`+d_8g^{kl}_k\varphi _l\varphi {\displaystyle \frac{1}{\sqrt{g}}}_i(\sqrt{g}g^{ij}_j\varphi ).`$ Here, the explicit form of $`d_1`$, $`\mathrm{}`$ $`d_8`$ is given in Appendix A. Substituting (16), (19) and (20) into (17), one gets $`g^{ij}g_{(2)ij}`$ $`=`$ $`f_1R^2+f_2R_{ij}R^{ij}+f_3R^{ij}_i\varphi _j\varphi `$ (21) $`+f_4Rg^{ij}_i\varphi _j\varphi +f_5R{\displaystyle \frac{1}{\sqrt{g}}}_i(\sqrt{g}g^{ij}_j\varphi )`$ $`+f_6(g^{ij}_i\varphi _j\varphi )^2+f_7\left({\displaystyle \frac{1}{\sqrt{g}}}_i(\sqrt{g}g^{ij}_j\varphi )\right)^2`$ $`+f_8g^{kl}_k\varphi _l\varphi {\displaystyle \frac{1}{\sqrt{g}}}_i(\sqrt{g}g^{ij}_j\varphi ).`$ Again, the explicit form of very complicated functions $`f_1`$, $`\mathrm{}`$ $`f_8`$ is given in Appendix A. Finally substituting (16), (19), (20) and (21) into the expression for the anomaly (2), we obtain, $`T`$ $`=`$ $`{\displaystyle \frac{1}{8\pi G}}[h_1R^2+h_2R_{ij}R^{ij}+h_3R^{ij}_i\varphi _j\varphi `$ (22) $`+h_4Rg^{ij}_i\varphi _j\varphi +h_5R{\displaystyle \frac{1}{\sqrt{g}}}_i(\sqrt{g}g^{ij}_j\varphi )`$ $`+h_6(g^{ij}_i\varphi _j\varphi )^2+h_7\left({\displaystyle \frac{1}{\sqrt{g}}}_i(\sqrt{g}g^{ij}_j\varphi )\right)^2`$ $`+h_8g^{kl}_k\varphi _l\varphi {\displaystyle \frac{1}{\sqrt{g}}}_i(\sqrt{g}g^{ij}_j\varphi )].`$ Here $`h_1`$ $`=`$ $`[3\{(2410\mathrm{\Phi })\mathrm{\Phi }^6`$ $`+\left(62208+22464\mathrm{\Phi }+2196\mathrm{\Phi }^2+72\mathrm{\Phi }^3+\mathrm{\Phi }^4\right)\mathrm{\Phi }^{\prime \prime }(\mathrm{\Phi }^{\prime \prime }+8V)^2`$ $`+2\mathrm{\Phi }^4\left\{\left(108+162\mathrm{\Phi }+7\mathrm{\Phi }^2\right)\mathrm{\Phi }^{\prime \prime }+72\left(8+14\mathrm{\Phi }+\mathrm{\Phi }^2\right)V\right\}`$ $`2\mathrm{\Phi }^2\{(6912+2736\mathrm{\Phi }+192\mathrm{\Phi }^2+\mathrm{\Phi }^3)\mathrm{\Phi }^{\prime \prime 2}`$ $`+4\left(11232+6156\mathrm{\Phi }+552\mathrm{\Phi }^2+13\mathrm{\Phi }^3\right)\mathrm{\Phi }^{\prime \prime }V`$ $`+32(2592+468\mathrm{\Phi }+96\mathrm{\Phi }^2+5\mathrm{\Phi }^3)V^2\}`$ $`3(24+\mathrm{\Phi })(6+\mathrm{\Phi })^2\mathrm{\Phi }^3(\mathrm{\Phi }^{\prime \prime \prime }+8V^{})\}]/`$ $`[16(6+\mathrm{\Phi })^2\{2\mathrm{\Phi }^2+(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }\}\{2\mathrm{\Phi }^2`$ $`+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V)\}^2]`$ $`h_2`$ $`=`$ $`{\displaystyle \frac{3\left\{(125\mathrm{\Phi })\mathrm{\Phi }^2+(288+72\mathrm{\Phi }+\mathrm{\Phi }^2)\mathrm{\Phi }^{\prime \prime }\right\}}{8(6+\mathrm{\Phi })^2\left\{2\mathrm{\Phi }^2+(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }\right\}}}.`$ (23) We also give the explicit forms of $`h_3`$, $`\mathrm{}`$ $`h_8`$ in Appendix A. Thus, we found the complete Weyl anomaly from bulk side. This expression which should describe dual d4 QFT of QCD type, with broken SUSY looks really complicated. The interesting remark is that Weyl anomaly is not integrable in general. In other words, it is impossible to construct the anomaly induced action. This is not strange, as it is usual situation for conformal anomaly when radiative corrections are taken into account. In case of the dilaton gravity in corresponding to $`\mathrm{\Phi }=0`$ (or more generally in case that the axion is included as in ), we have the following expression: $`T`$ $`=`$ $`{\displaystyle \frac{l^3}{8\pi G}}{\displaystyle }d^4x\sqrt{g_{(0)}}[{\displaystyle \frac{1}{8}}R_{(0)ij}R_{(0)}^{ij}{\displaystyle \frac{1}{24}}R_{(0)}^2`$ (24) $`{\displaystyle \frac{1}{2}}R_{(0)}^{ij}_i\phi _{(0)}_j\phi _{(0)}+{\displaystyle \frac{1}{6}}R_{(0)}g_{(0)}^{ij}_i\phi _{(0)}_j\phi _{(0)}`$ $`+{\displaystyle \frac{1}{4}}\left\{{\displaystyle \frac{1}{\sqrt{g_{(0)}}}}_i\left(\sqrt{g_{(0)}}g_{(0)}^{ij}_j\phi _{(0)}\right)\right\}^2+{\displaystyle \frac{1}{3}}\left(g_{(0)}^{ij}_i\phi _{(0)}_j\phi _{(0)}\right)^2].`$ Here $`\phi `$ can be regarded as dilaton. In the limit of $`\mathrm{\Phi }0`$, we obtain $`h_1`$ $``$ $`{\displaystyle \frac{362208\mathrm{\Phi }^{\prime \prime }(8V)^2}{166^22418\mathrm{\Phi }^{\prime \prime }(8V)^2}}={\displaystyle \frac{1}{24}}`$ $`h_2`$ $``$ $`{\displaystyle \frac{3288\mathrm{\Phi }^{\prime \prime }}{86^224\mathrm{\Phi }^{\prime \prime }}}={\displaystyle \frac{1}{8}}`$ $`h_3`$ $``$ $`{\displaystyle \frac{3288(\mathrm{\Phi }^{\prime \prime }V\mathrm{\Phi }^{}V^{})}{46^224\mathrm{\Phi }^{\prime \prime }}}={\displaystyle \frac{1}{4}}{\displaystyle \frac{(\mathrm{\Phi }^{\prime \prime }V\mathrm{\Phi }^{}V^{})}{\mathrm{\Phi }^{\prime \prime }}}`$ $`h_4`$ $``$ $`{\displaystyle \frac{362208\mathrm{\Phi }^{\prime \prime }V(8V)^2+6\mathrm{\Phi }^{}384(5184)V^2V^{}}{86^224\mathrm{\Phi }^{\prime \prime }(188V)^2}}={\displaystyle \frac{1}{12}}{\displaystyle \frac{(\mathrm{\Phi }^{\prime \prime }V\mathrm{\Phi }^{}V^{})}{\mathrm{\Phi }^{\prime \prime }}}`$ $`h_5`$ $``$ $`0`$ $`h_6`$ $``$ $`\{\mathrm{\Phi }^{\prime \prime }64V(373248V^3139968V_{}^{}{}_{}{}^{2})`$ $`+26\mathrm{\Phi }^{}V^{}(2)(432)(4608V^3+864V_{}^{}{}_{}{}^{2}1728VV^{\prime \prime })\}`$ $`/166^224\mathrm{\Phi }^{\prime \prime }(188V)^2`$ $`=`$ $`{\displaystyle \frac{\left\{\mathrm{\Phi }^{\prime \prime }V\left(V^3\frac{3}{8}V_{}^{}{}_{}{}^{2}\right)+2\mathrm{\Phi }^{}V^{}\left(V^3+\frac{3}{16}V_{}^{}{}_{}{}^{2}\frac{3}{8}VV^{\prime \prime }\right)\right\}}{12\mathrm{\Phi }^{\prime \prime }V^2}}`$ $`h_7`$ $``$ $`{\displaystyle \frac{V818^2\mathrm{\Phi }^{\prime \prime }V212V}{24\mathrm{\Phi }^{\prime \prime }(188V)^2}}={\displaystyle \frac{V}{8}}`$ $`h_8`$ $``$ $`{\displaystyle \frac{3218^2\mathrm{\Phi }^{\prime \prime }V212V^{}}{424\mathrm{\Phi }^{\prime \prime }(188V)^2}}={\displaystyle \frac{V^{}}{8V}}.`$ (25) Especially if we choose $$V=2,$$ (26) we obtain, $`h_1{\displaystyle \frac{1}{24}},h_2{\displaystyle \frac{1}{8}},h_3{\displaystyle \frac{1}{2}},h_4{\displaystyle \frac{1}{6}}`$ $`h_50,h_6{\displaystyle \frac{1}{3}},h_7{\displaystyle \frac{1}{4}},h_80`$ (27) and we find that the standard result (conformal anomaly of $`𝒩=4`$ super YM theory covariantly coupled with $`𝒩=4`$ conformal supergravity ) in (24) is reproduced . We should also note that the expression (22) cannot be rewritten as a sum of the Gauss-Bonnet invariant $`G`$ and the square of the Weyl tensor $`F`$, which are given as $`G`$ $`=`$ $`R^24R_{ij}R^{ij}+R_{ijkl}R^{ijkl}`$ $`F`$ $`=`$ $`{\displaystyle \frac{1}{3}}R^22R_{ij}R^{ij}+R_{ijkl}R^{ijkl},`$ (28) This is the signal that the conformal symmetry is broken already in classical theory. When $`\varphi `$ is constant, only two terms corresponding to $`h_1`$ and $`h_2`$ survive in (22) : $`T`$ $`=`$ $`{\displaystyle \frac{1}{8\pi G}}\left[h_1R^2+h_2R_{ij}R^{ij}\right]`$ (29) $`=`$ $`{\displaystyle \frac{1}{8\pi G}}\left[\left(h_1+{\displaystyle \frac{1}{3}}h_2\right)R^2+{\displaystyle \frac{1}{2}}h_2\left(FG\right)\right].`$ As $`h_1`$ depends on $`V`$, we may compare the result with the conformal anomaly from, say, scalar or spinor QED, or QCD in the phase where there are no background scalars and (or) spinors.. The structure of the conformal anomaly in such a theory has the following form $$T=\widehat{a}G+\widehat{b}F+\widehat{c}R^2.$$ (30) where $$\widehat{a}=\text{constant}+a_1e^2,\widehat{b}=\text{constant}+a_2e^2,\widehat{c}=a_3e^2.$$ (31) Here $`e^2`$ is the electric charge (or $`g^2`$ in case of QCD). Imagine that one can identify $`e`$ with the exponential of the constant dilaton (using holographic RG ). $`a_1`$, $`a_2`$ and $`a_3`$ are some numbers. Comparing (29) and (30), we obtain $$\widehat{a}=\widehat{b}=\frac{h_2}{16\pi G},\widehat{c}=\frac{1}{8\pi G}\left(h_1+\frac{1}{3}h_2\right).$$ (32) When $`\mathrm{\Phi }`$ is small, one gets $`h_1`$ $`=`$ $`{\displaystyle \frac{1}{24}}[1{\displaystyle \frac{1}{8}}\mathrm{\Phi }+{\displaystyle \frac{1}{8}}{\displaystyle \frac{\left(\mathrm{\Phi }^{}\right)^2}{\mathrm{\Phi }^{\prime \prime }}}`$ $`+{\displaystyle \frac{25}{2592}}\mathrm{\Phi }^2{\displaystyle \frac{17}{216}}{\displaystyle \frac{\left(\mathrm{\Phi }^{}\right)^2\mathrm{\Phi }}{\mathrm{\Phi }^{\prime \prime }}}+{\displaystyle \frac{1}{576}}{\displaystyle \frac{\left(\mathrm{\Phi }^{}\right)^2}{V}}+{\displaystyle \frac{1}{96}}{\displaystyle \frac{\left(\mathrm{\Phi }^{}\right)^4}{\left(\mathrm{\Phi }^{\prime \prime }\right)^2}}+𝒪\left(\mathrm{\Phi }^3\right)]`$ $`h_2`$ $`=`$ $`{\displaystyle \frac{1}{8}}[1{\displaystyle \frac{1}{8}}\mathrm{\Phi }+{\displaystyle \frac{1}{8}}{\displaystyle \frac{\left(\mathrm{\Phi }^{}\right)^2}{\mathrm{\Phi }^{\prime \prime }}}`$ (33) $`+{\displaystyle \frac{5}{576}}\mathrm{\Phi }^2{\displaystyle \frac{3}{64}}{\displaystyle \frac{\left(\mathrm{\Phi }^{}\right)^2\mathrm{\Phi }}{\mathrm{\Phi }^{\prime \prime }}}+{\displaystyle \frac{1}{96}}{\displaystyle \frac{\left(\mathrm{\Phi }^{}\right)^4}{\left(\mathrm{\Phi }^{\prime \prime }\right)^2}}+𝒪\left(\mathrm{\Phi }^3\right)].`$ If one assumes $$\mathrm{\Phi }(\varphi )=a\mathrm{e}^{b\varphi },(|a|1),$$ (34) then $`h_2`$ $`=`$ $`{\displaystyle \frac{1}{8}}\left[1{\displaystyle \frac{a^2}{36}}\mathrm{e}^{2b\varphi }+𝒪\left(a^3\right)\right]`$ $`h_1+{\displaystyle \frac{1}{3}}h_2`$ $`=`$ $`{\displaystyle \frac{a^2}{24}}\left({\displaystyle \frac{5}{162}}+{\displaystyle \frac{b^2}{576V}}\right)\mathrm{e}^{2b\varphi }+𝒪\left(a^3\right).`$ (35) Comparing (2) with (31) and (32) and assuming $$e^2=\mathrm{e}^{2b\varphi },$$ (36) we find $`a_1=a_2`$ $`=`$ $`{\displaystyle \frac{1}{16\pi G}}{\displaystyle \frac{1}{8}}{\displaystyle \frac{a^2}{36}},`$ $`a_3`$ $`=`$ $`{\displaystyle \frac{1}{8\pi G}}{\displaystyle \frac{a^2}{24}}\left({\displaystyle \frac{5}{162}}+{\displaystyle \frac{b^2}{576V}}\right).`$ (37) Here $`V`$ should be arbitrary but constant. We should note $`\mathrm{\Phi }(0)0`$. One can absorb the difference into the redefinition of $`l`$ since we need not to assume $`\mathrm{\Phi }(0)=0`$ in deriving the form of $`h_1`$ and $`h_2`$ in (2). Hence, this simple example suggests the way of comparison between SG side and QFT descriptions of non-conformal boundary theory. In order that the region near the boundary at $`\rho =0`$ is asymptotically AdS, we need to require $`\mathrm{\Phi }0`$ and $`\mathrm{\Phi }^{}0`$ when $`\rho 0`$. One can also confirm that $`h_1\frac{1}{24}`$ and $`h_2\frac{1}{8}`$ in the limit of $`\mathrm{\Phi }0`$ and $`\mathrm{\Phi }^{}0`$ even if $`\mathrm{\Phi }^{\prime \prime }0`$ and $`\mathrm{\Phi }^{\prime \prime \prime }0`$. In the AdS/CFT correspondence, $`h_1`$ and $`h_2`$ are related with the central charge $`c`$ of the conformal field theory (or its analog for non-conformal theory). Since we have two functions $`h_1`$ and $`h_2`$, there are two ways to define the candidate c-function when the conformal field theory is deformed: $$c_1=\frac{24\pi h_1}{G},c_2=\frac{8\pi h_2}{G}.$$ (38) If we put $`V(\varphi )=4\lambda ^2+\mathrm{\Phi }(\varphi )`$, then $`l=\left(\frac{12}{V(0)}\right)^{\frac{1}{2}}`$. One should note that it is chosen $`l=1`$ in (38). We can restore $`l`$ by changing $`hl^3h`$ and $`kl^3k`$ and $`\mathrm{\Phi }^{}l\mathrm{\Phi }^{}`$, $`\mathrm{\Phi }^{\prime \prime }l^2\mathrm{\Phi }^{\prime \prime }`$ and $`\mathrm{\Phi }^{\prime \prime \prime }l^3\mathrm{\Phi }^{\prime \prime \prime }`$ in (22). Then in the limit of $`\mathrm{\Phi }0`$, one gets $$c_1,c_2\frac{\pi }{G}\left(\frac{12}{V(0)}\right)^{\frac{3}{2}},$$ (39) which agrees with the proposal of the previous work in the limit. The c-function $`c_1`$ or $`c_2`$ in (38) is, of course, more general definition. It is interesting to study the behaviour of candidate c-function for explicit values of dilatonic potential at different limits. It also could be interesting to see what is the analogue of our dilaton-dependent c-function in non-commutative YM theory (without dilaton, see ). ## 3 Properties of c-function The definitions of the c-functions in (13) and (38), are, however, not always good ones since our results are too wide. That is, we have obtained the conformal anomaly for arbitrary dilatonic background which may not be the solution of original $`d=5`$ gauged supergravity. As only solutions of d5 gauged supergravity describe RG flows of dual QFT it is not strange that above candidate c-functions are not acceptable. They quickly become non-monotonic and even singular in explicit examples. They presumbly measure the deviations from SG description and should not be taken seriously. As pointed in , it might be necessary to impose the condition $`\mathrm{\Phi }^{}=0`$ on the conformal boundary. Such condition follows from the equations of motion of d5 gauged SG. Anyway as $`\mathrm{\Phi }^{}=0`$ on the boundary in the solution which has the asymptotic AdS region, we can add any function which proportional to the power of $`\mathrm{\Phi }^{}=0`$ to the previous expressions of the c-functions in (13) and (38). As a trial, if we put $`\mathrm{\Phi }^{}=0`$, we obtain $`c`$ $`=`$ $`{\displaystyle \frac{3}{2G}}\left[{\displaystyle \frac{l}{2}}+{\displaystyle \frac{1}{l}}{\displaystyle \frac{1}{\mathrm{\Phi }(\varphi _{(0)})+\frac{2}{l^2}}}\right]`$ (40) instead of (13) and $`c_1`$ $`=`$ $`{\displaystyle \frac{2\pi }{3G}}{\displaystyle \frac{62208+22464\mathrm{\Phi }+2196\mathrm{\Phi }^2+72\mathrm{\Phi }^3+\mathrm{\Phi }^4}{(6+\mathrm{\Phi })^2(24+\mathrm{\Phi })(18+\mathrm{\Phi })}}`$ $`c_2`$ $`=`$ $`{\displaystyle \frac{3\pi }{G}}{\displaystyle \frac{288+72\mathrm{\Phi }+\mathrm{\Phi }^2}{(6+\mathrm{\Phi })^2(24+\mathrm{\Phi })}}`$ (41) instead of (38). We should note that there disappear the higher derivative terms like $`\mathrm{\Phi }^{\prime \prime }`$ or $`\mathrm{\Phi }^{\prime \prime \prime }`$. That will be our final proposal for acceptable c-function in terms of dilatonic potential. The given c-functions in (3) also have the property (39) and reproduce the known result for the central charge on the boundary. Since $`\frac{d\mathrm{\Phi }}{dz}0`$ in the asymptotically AdS region even if the region is UV or IR, the given c-functions in (40) and (3) have fixed points in the asymptotic AdS region $`\frac{dc}{dU}=\frac{dc}{d\mathrm{\Phi }}\frac{d\mathrm{\Phi }}{d\varphi }\frac{d\varphi }{dU}0`$, where $`U=\rho ^{\frac{1}{2}}`$ is the radius coordinate in AdS or the energy scale of the boundary field theory. We can now check the monotonity in the c-functions. For this purpose, we consider some examples. In and , the following dilaton potentials appeared: $`4\lambda ^2+\mathrm{\Phi }_{\mathrm{FGPW}}(\varphi )`$ $`=`$ $`4\left(\mathrm{exp}\left[\left({\displaystyle \frac{4\varphi }{\sqrt{6}}}\right)\right]+2\mathrm{exp}\left[\left({\displaystyle \frac{2\varphi }{\sqrt{6}}}\right)\right]\right)`$ (42) $`4\lambda ^2+\mathrm{\Phi }_{\mathrm{GPPZ}}(\varphi )`$ $`=`$ $`{\displaystyle \frac{3}{2}}\left(3+\left(\mathrm{cosh}\left[\left({\displaystyle \frac{2\varphi }{\sqrt{3}}}\right)\right]\right)^2+4\mathrm{cosh}\left[\left({\displaystyle \frac{2\varphi }{\sqrt{3}}}\right)\right]\right).`$ (43) In both cases $`V`$ is a constant and $`V=2`$. In the classical solutions for the both cases, $`\varphi `$ is the monotonically decreasing function of the energy scale $`U=\rho ^{\frac{1}{2}}`$ and $`\varphi =0`$ at the UV limit corresponding to the boundary. Then in order to know the energy scale dependences of $`c_1`$ and $`c_2`$, we only need to investigate the $`\varphi `$ dependences of $`c_1`$ and $`c_2`$ in (3). As the potentials and also $`\mathrm{\Phi }`$ have a minimum $`\mathrm{\Phi }=0`$ at $`\varphi =0`$, which corresponds to the UV boundary in the solutions in and , and $`\mathrm{\Phi }`$ is monotonicaly increasing function of the absolute value $`|\varphi |`$, we only need to check the monotonities of $`c_1`$ and $`c_2`$ with respect to $`\mathrm{\Phi }`$ when $`\mathrm{\Phi }0`$. From (3), we find $`{\displaystyle \frac{d\left(\mathrm{ln}c_1\right)}{d\mathrm{\Phi }}}`$ $`={\displaystyle \frac{18\left(622080+383616\mathrm{\Phi }+64296\mathrm{\Phi }^2+4548\mathrm{\Phi }^3+130\mathrm{\Phi }^4+\mathrm{\Phi }^5\right)}{(6+\mathrm{\Phi })(18+\mathrm{\Phi })(24+\mathrm{\Phi })(62208+22464\mathrm{\Phi }+2196\mathrm{\Phi }^2+72\mathrm{\Phi }^3+\mathrm{\Phi }^4)}}`$ $`<0`$ $`{\displaystyle \frac{d\left(\mathrm{ln}c_2\right)}{d\mathrm{\Phi }}}={\displaystyle \frac{5184+2304\mathrm{\Phi }+138\mathrm{\Phi }^2+\mathrm{\Phi }^3}{(6+\mathrm{\Phi })(24+\mathrm{\Phi })(288+72\mathrm{\Phi }+\mathrm{\Phi }^2)}}<0.`$ (44) Therefore the c-functions $`c_1`$ and $`c_2`$ are monotonically decreasing functions of $`\mathrm{\Phi }`$ or increasings function of the energy scale $`U`$ as the c-function in . We should also note that the c-functions $`c_1`$ and $`c_2`$ are positive definite for non-negative $`\mathrm{\Phi }`$. For $`c`$ in (40) for $`d=2`$ case, it is very straightforward to check the monotonity and the positivity. In , another c-function has been proposed in terms of the metric as follows: $$c_{\mathrm{GPPZ}}=\left(\frac{dA}{dz}\right)^3,$$ (45) where the metric is given by $$ds^2=dz^2+\mathrm{e}^{2A}dx_\mu dx^\mu .$$ (46) The c-function (45) is positive and has a fixed point in the asymptotically AdS region again and the c-function is also monotonically increasing function of the energy scale. The c-functions (40) and (3) proposed in this paper are given in terms of the dilaton potential, not in terms of metric, but it might be interesting that the c-functions in (40) and (3) have the similar properties (positivity, monotonity and fixed point in the asymptotically AdS region). These properties could be understood from the equations of motion. When the metric has the form (46), the equations of motion are: $`\varphi ^{\prime \prime }+dA^{}\varphi ^{}={\displaystyle \frac{\mathrm{\Phi }}{\varphi }},`$ (47) $`dA^{\prime \prime }+d(A^{})^2+{\displaystyle \frac{1}{2}}(\varphi ^{})^2={\displaystyle \frac{4\lambda ^2+\mathrm{\Phi }}{d1}},`$ (48) $`A^{\prime \prime }+d(A^{})^2={\displaystyle \frac{4\lambda ^2+\mathrm{\Phi }}{d1}}.`$ (49) Here $`{}_{}{}^{}\frac{d}{dz}`$. From (47) and (48), we obtain $$0=2(d1)A^{\prime \prime }+\varphi _{}^{}{}_{}{}^{2}$$ (50) Then if $`A^{\prime \prime }=0`$, $`\varphi ^{}=0`$, which tells that if $`\frac{dc_{\mathrm{GPPZ}}}{dz}=0`$, then $`\frac{dc_1}{dz}=\frac{dc_2}{dz}=0`$. Then $`c_{\mathrm{GPPZ}}`$ has a fixed point, $`c_1`$ and $`c_2`$ have a fixed point. From (47) and (48), we also obtain $$0=d(d1)A_{}^{}{}_{}{}^{2}+4\lambda ^2+\mathrm{\Phi }\frac{1}{2}\varphi _{}^{}{}_{}{}^{2}.$$ (51) Then at the fixed point where $`\varphi ^{}=0`$, we obtain $$0=d(d1)A_{}^{}{}_{}{}^{2}+4\lambda ^2+\mathrm{\Phi }.$$ (52) Therefore if $`c_{\mathrm{GPPZ}}`$ and $`A^{}`$ is the monotonic function of $`z`$, $`V`$ and also $`c_1`$ and $`c_2`$ are also monotonic function at least at the fixed point. We have to note that above considerations do not give the proof of equivalency of our proposal c-functions with other proposals. However, it is remarkable (at least, for a number of potentials) that they enjoy the similar properties: positivity, monotonity and existance of fixed points. We can also consider other examples of c-function for different choices of dilatonic potential. In , several examples of the potentials in gauged supergravity are given. They appeared as a result of sphere reduction in M-theory or string theory, down to three or five dimensions. Their properties are described in detail in refs.. The potentials have the following form: $$4\lambda ^2+\mathrm{\Phi }(\varphi )=\frac{d(d1)}{\frac{1}{a_1^2}\frac{1}{a_1a_2}}\left(\frac{1}{a_1^2}\mathrm{e}^{a_1\varphi }\frac{1}{a_1a_2}\mathrm{e}^{a_2\varphi }\right).$$ (53) Here $`a_1`$ and $`a_2`$ are constant parameters depending on the model. We also normalize the potential so that $`4\lambda ^2+\mathrm{\Phi }(\varphi )d(d1)`$ when $`\varphi 0`$. For simplicity, we choose $`G=l=1`$ in this section. For $`𝒩=1`$ model in $`D=d+1=3`$ dimensions $$a_1=2\sqrt{2},a_2=\sqrt{2},$$ (54) for $`D=3`$, $`𝒩=2`$, one gets $$a_1=\sqrt{6},a_2=2\sqrt{\frac{2}{3}},$$ (55) and for $`D=3`$, $`𝒩=3`$ model, we have $$a_1=\frac{4}{\sqrt{3}},a_2=\sqrt{3}.$$ (56) On the other hand, for $`D=d+1=5`$, $`𝒩=1`$ model, $`a_1`$ and $`a_2`$ are $$a_1=2\sqrt{\frac{5}{3}},a_2=\frac{4}{\sqrt{15}}.$$ (57) The proposed c-functions have not acceptable behaviour for above potentials. (There seems to be no problem for 2d case.) The problem seems to be that the solutions in above models have not asymptotic AdS region in UV but in IR. On the same time the conformal anomaly in (22) is evaluated as UV effect. If we assume that $`\mathrm{\Phi }`$ in the expression of c-functions $`c_1`$ and $`c_2`$ vanishes at IR AdS region, $`\mathrm{\Phi }`$ becomes negative. When $`\mathrm{\Phi }`$ is negative, the properties of the c-functions $`c_1`$ and $`c_2`$ become bad, they are not monotonic nor positive, and furthermore they have a singularity in the region given by the solutions in . Thus, for such type of potential other proposal for c-function which isnot related with conformal nomaly should be made. Hence, we discussed the typical behaviour of candidate c-functions. However, it is not clear which role should play dilaton in above expressions as holographic RG coupling constant in dual QFT. It could be induced mass, quantum fields or coupling constants (most probably, gauge coupling), but the explicit rule with what it should be identified is absent. The big number of usual RG parameters in dual QFT suggests also that there should be considered gauged SG with few scalars. ## 4 Surface Counterterms and Finite Action As well-known, we need to add the surface terms to the bulk action in order to have the well-defined variational principle. Under the variation of the metric $`\widehat{G}^{\mu \nu }`$ and the scalar field $`\varphi `$, the variation of the action (1) is given by $`\delta S=\delta S_{M_{d+1}}+\delta S_{M_d}`$ (58) $`\delta S_{M_{d+1}}={\displaystyle \frac{1}{16\pi G}}{\displaystyle _{M_{d+1}}}d^{d+1}x\sqrt{\widehat{G}}[\delta \widehat{G}^{\zeta \xi }\{{\displaystyle \frac{1}{2}}G_{\zeta \xi }\{\widehat{R}`$ $`+(X(\varphi )Y^{}(\varphi ))(\widehat{}\varphi )^2+\mathrm{\Phi }(\varphi )+4\lambda ^2\}+\widehat{R}_{\zeta \xi }+(X(\varphi )Y^{}(\varphi ))_\zeta \varphi _\xi \varphi \}`$ $`+\delta \varphi \{(X^{}(\varphi )Y^{\prime \prime }(\varphi ))(\widehat{}\varphi )^2+\mathrm{\Phi }^{}(\varphi )`$ $`{\displaystyle \frac{1}{\sqrt{\widehat{G}}}}_\mu \left(\sqrt{\widehat{G}}\widehat{G}^{\mu \nu }(X(\varphi )Y^{}(\varphi ))_\nu \varphi \right)\}].`$ $`\delta S_{M_d}={\displaystyle \frac{1}{16\pi G}}{\displaystyle _{M_d}}d^dx\sqrt{\widehat{g}}n_\mu \left[^\mu \left(\widehat{G}_{\xi \nu }\delta \widehat{G}^{\xi \nu }\right)D_\nu \left(\delta \widehat{G}^{\mu \nu }\right)+Y(\varphi )^\mu \left(\delta \varphi \right)\right].`$ Here $`\widehat{g}_{\mu \nu }`$ is the metric induced from $`\widehat{G}_{\mu \nu }`$ and $`n_\mu `$ is the unit vector normal to $`M_d`$. The surface term $`\delta S_{M_d}`$ of the variation contains $`n^\mu _\mu \left(\delta \widehat{G}^{\xi \nu }\right)`$ and $`n^\mu _\mu \left(\delta \varphi \right)`$, which makes the variational principle ill-defined. In order that the variational principle is well-defined on the boundary, the variation of the action should be written as $$\delta S_{M_d}=\underset{\rho 0}{lim}_{M_d}d^dx\sqrt{\widehat{g}}\left[\delta \widehat{G}^{\xi \nu }\left\{\mathrm{}\right\}+\delta \varphi \left\{\mathrm{}\right\}\right]$$ (59) after using the partial integration. If we put $`\left\{\mathrm{}\right\}=0`$ for $`\left\{\mathrm{}\right\}`$ in (59), one could obtain the boundary condition corresponding to Neumann boundary condition. We can, of course, select Dirichlet boundary condition by choosing $`\delta \widehat{G}^{\xi \nu }=\delta \varphi =0`$, which is natural for AdS/CFT correspondence. The Neumann type condition becomes, however, necessary later when we consider the black hole mass etc. by using surface terms. If the variation of the action on the boundary contains $`n^\mu _\mu \left(\delta \widehat{G}^{\xi \nu }\right)`$ or $`n^\mu _\mu \left(\delta \varphi \right)`$, however, we cannot partially integrate it on the boundary in order to rewrite the variation in the form of (59) since $`n_\mu `$ expresses the direction perpendicular to the boundary. Therefore the “minimum” of the action is ambiguous. Such a problem was well studied in for the Einstein gravity and the boundary term was added to the action. It cancels the term containing $`n^\mu _\mu \left(\delta \widehat{G}^{\xi \nu }\right)`$. We need to cancel also the term containing $`n^\mu _\mu \left(\delta \varphi \right)`$. Then one finds the boundary term $$S_b^{(1)}=\frac{1}{8\pi G}_{M_d}d^dx\sqrt{\widehat{g}}\left[D_\mu n^\mu +Y(\varphi )n_\mu ^\mu \varphi \right].$$ (60) We also need to add surface counterterm $`S_b^{(2)}`$ which cancels the divergence coming from the infinite volume of the bulk space, say AdS. In order to investigate the divergence, we choose the metric in the form (5). In the parametrization (5), $`n^\mu `$ and the curvature $`R`$ are given by $`n^\mu `$ $`=`$ $`({\displaystyle \frac{2\rho }{l}},0,\mathrm{},0)`$ $`R`$ $`=`$ $`\stackrel{~}{R}+{\displaystyle \frac{3\rho ^2}{l^2}}\widehat{g}^{ij}\widehat{g}^{kl}\widehat{g}_{ik}^{}\widehat{g}_{jl}^{}{\displaystyle \frac{4\rho ^2}{l^2}}\widehat{g}^{ij}\widehat{g}_{ij}^{\prime \prime }{\displaystyle \frac{\rho ^2}{l^2}}\widehat{g}^{ij}\widehat{g}^{kl}\widehat{g}_{ij}^{}\widehat{g}_{kl}^{}.`$ (61) Here $`\stackrel{~}{R}`$ is the scalar curvature defined by $`g_{ij}`$ in (5). Expanding $`g_{ij}`$ and $`\varphi `$ with respect to $`\rho `$ as in (7), we find the following expression for $`S+S_b^{(1)}`$: $`S+S_b^{(1)}`$ $`=`$ $`{\displaystyle \frac{1}{16\pi G}}\underset{\rho 0}{lim}{\displaystyle }d^dxl\rho ^{\frac{d}{2}}\sqrt{g_{(0)}}[{\displaystyle \frac{22d}{l^2}}{\displaystyle \frac{1}{d}}\mathrm{\Phi }(\varphi _0)`$ (62) $`+\rho \{{\displaystyle \frac{1}{d2}}R_{(0)}{\displaystyle \frac{1}{l^2}}g_{(0)}^{ij}g_{(1)ij}`$ $`{\displaystyle \frac{1}{d2}}(X(\varphi _{(0)})(_{(0)}\varphi _{(0)})^2+Y(\varphi _{(0)})\mathrm{\Delta }\varphi _{(0)}`$ $`+\mathrm{\Phi }^{}(\varphi _{(0)})\varphi _{(1)})\}+𝒪\left(\rho ^2\right)].`$ Then for $`d=2`$ $$S_b^{(2)}=\frac{1}{16\pi G}d^dx\sqrt{\widehat{g}}\left[\frac{2}{l}+\frac{l}{2}\mathrm{\Phi }(\varphi )\right]$$ (63) and for $`d=3,4`$, $`S_b^{(2)}={\displaystyle \frac{1}{16\pi G}}{\displaystyle }d^dx[\sqrt{\widehat{g}}\{{\displaystyle \frac{2d2}{l}}+{\displaystyle \frac{l}{d2}}R{\displaystyle \frac{2l}{d(d2)}}\mathrm{\Phi }(\varphi )`$ $`+{\displaystyle \frac{l}{d2}}(X(\varphi )\left(\widehat{}\varphi \right)^2+Y(\varphi )\widehat{\mathrm{\Delta }}\varphi )\}{\displaystyle \frac{l^2}{d(d2)}}n^\mu _\mu \left(\sqrt{\widehat{g}}\mathrm{\Phi }(\varphi )\right)].`$ (64) Note that the last term in above expression does not look typical from the AdS/CFT point of view. The reason is that it does not depend from only the boundary values of the fields. Its presence may indicate to breaking of AdS/CFT conjecture in the situations when SUGRA scalars significally deviate from constants or are not asymptotic constants <sup>5</sup><sup>5</sup>5We thank the referee for adressing this issue.. Here $`\widehat{\mathrm{\Delta }}`$ and $`\widehat{}`$ are defined by using $`d`$-dimensional metric and we used $`\sqrt{\widehat{g}}\mathrm{\Phi }(\varphi )=\rho ^{\frac{d}{2}}\sqrt{g_{(0)}}\{\mathrm{\Phi }(\varphi _{(0)})`$ $`+\rho ({\displaystyle \frac{1}{2}}g_{(0)}^{ij}g_{(1)ij}\mathrm{\Phi }(\varphi _{(0)})+\mathrm{\Phi }^{}(\varphi _{(0)})\varphi _{(1)})+𝒪\left(\rho ^2\right)\}`$ $`n^\mu _\mu \left(\sqrt{\widehat{g}}\mathrm{\Phi }(\varphi )\right)={\displaystyle \frac{2}{l}}\rho ^{\frac{d}{2}}\sqrt{g_{(0)}}\{{\displaystyle \frac{d}{2}}\mathrm{\Phi }(\varphi _{(0)}`$ $`+\rho (1{\displaystyle \frac{d}{2}})({\displaystyle \frac{1}{2}}g_{(0)}^{ij}g_{(1)ij}\mathrm{\Phi }(\varphi _{(0)})+\mathrm{\Phi }^{}(\varphi _{(0)})\varphi _{(1)})+𝒪\left(\rho ^2\right)\}.`$ (65) Note that $`S_b^{(2)}`$ in (63) or (4) is only given in terms of the boundary quantities except the last term in (4). The last term is necessary to cancel the divergence of the bulk action and it is, of course, the total derivative in the bulk theory: $$d^dxn^\mu _\mu \left(\sqrt{\widehat{g}}\mathrm{\Phi }(\varphi )\right)=d^{d+1}x\sqrt{\widehat{G}}\mathrm{}\mathrm{\Phi }(\varphi ).$$ (66) Thus we got the boundary counterterm action for gauged SG. Using these local surface counterterms as part of complete action one can show explicitly that bosonic sector of gauged SG in dimensions under discussion gives finite action in asymptotically AdS space. The corresponding example will be given in next section. Recently the surface counterterms for the action with the dilaton (scalar) potential are discussed in . Their counterterms seem to correspond to the terms cancelling the leading divergence when $`\rho 0`$ in (62). However, they seem to have only considered the case where the dilaton becomes asymptotically constant $`\varphi \varphi _0`$. If we choose $`\varphi _0=0`$, the total dilaton potential including the cosmological term $`V_{\mathrm{dilaton}}(\varphi )4\lambda ^2+\mathrm{\Phi }(\varphi )`$ approaches to $`V_{\mathrm{dilaton}}(\varphi )4\lambda ^2=d(d1)/l^2`$. Then if we only consider the leading $`\rho `$ behavior and the asymptotically constant dilaton, the counterterm action in (63) and/or (4) has the following form $`S_b^{(2)}`$ $`=`$ $`{\displaystyle \frac{1}{16\pi G}}{\displaystyle d^dx\sqrt{\widehat{g}}\left(\frac{2d2}{l}\right)},`$ (67) which coincides with the result in when the spacetime is aymptotically AdS. Let us turn now to the discussion of deep connection between surface counterterms and holographic conformal anomaly. It is enough to mention only $`d=4`$. In order to control the logarithmically divergent terms in the bulk action $`S`$, we choose $`d4=ϵ<0`$. Then $$S+S_b=\frac{1}{ϵ}S_{\mathrm{ln}}+\text{finite terms}.$$ (68) Here $`S_{\mathrm{ln}}`$ is given in (2). We also find $$g_{(0)}^{ij}\frac{\delta }{\delta g_{(0)}^{ij}}S_{\mathrm{ln}}=\frac{ϵ}{2}_{\mathrm{ln}}+𝒪\left(ϵ^2\right).$$ (69) Here $`_{\mathrm{ln}}`$ is the Lagrangian density corresponding to $`S_{\mathrm{ln}}`$ : $`S_{\mathrm{ln}}=d^{d+1}_{\mathrm{ln}}`$. Then combining (68) and (69), we obtain the trace anomaly : $$T=\underset{ϵ0}{lim}\frac{2\widehat{g}_{(0)}^{ij}}{\sqrt{\widehat{g}_{(0)}}}\frac{\delta (S+S_b)}{\delta \widehat{g}_{(0)}^{ij}}=\frac{1}{2}_{\mathrm{ln}},$$ (70) which is identical with the result found in (8). We should note that the last term in (4) does not lead to any ambiguity in the calculation of conformal anomaly since $`g_{(0)}`$ does not depend on $`\rho `$. If we use the equations of motion (15), (16), (17) and (18), we finally obtain the expression (22) or (111). Hence, we found the finite gravitational action (for asymptotically AdS spaces) in 5 dimensions by adding the local surface counterterm. This action correctly reproduces holographic trace anomaly for dual (gauge) theory. In principle, one can also generalize all results for higher dimensions, say, d6, etc. With the growth of dimension, the technical problems become more and more complicated as the number of structures in boundary term is increasing. ## 5 Dilatonic AdS Black Hole and its Mass Let us consider the black hole or “throat” type solution for the equations of the motion (2) and (3) when $`d=4`$. The surface term (4) may be used for calculation of the finite black hole mass and/or other thermodynamical quantities. For simplicity, we choose $$X(\varphi )=\alpha (\text{constant}),Y(\varphi )=0$$ (71) and we assume the spacetime metric in the following form: $$ds^2=\mathrm{e}^{2\rho }dt^2+\mathrm{e}^{2\sigma }dr^2+r^2\underset{i=1}{\overset{d1}{}}\left(dx^i\right)^2$$ (72) and $`\rho `$, $`\sigma `$ and $`\varphi `$ depend only on $`r`$. The equations (2) and (3) can be rewritten in the following form: $`0`$ $`=`$ $`\mathrm{e}^{\rho +\sigma }\mathrm{\Phi }^{}(\varphi )2\alpha \left(\mathrm{e}^{\rho \sigma }\varphi ^{}\right)^{}`$ (73) $`0`$ $`=`$ $`{\displaystyle \frac{1}{3}}\mathrm{e}^{2\rho }\left(\mathrm{\Phi }(\varphi )+{\displaystyle \frac{12}{l^2}}\right)+\left(\rho ^{\prime \prime }+\left(\rho ^{}\right)^2\rho ^{}\sigma ^{}+{\displaystyle \frac{3\rho ^{}}{r}}\right)\mathrm{e}^{2\rho 2\sigma }`$ (74) $`0`$ $`=`$ $`{\displaystyle \frac{1}{3}}\mathrm{e}^{2\sigma }\left(\mathrm{\Phi }(\varphi )+{\displaystyle \frac{12}{l^2}}\right)\rho ^{\prime \prime }\left(\rho ^{}\right)^2+\rho ^{}\sigma ^{}+{\displaystyle \frac{3\sigma ^{}}{r}}+\alpha \left(\varphi ^{}\right)^2`$ (75) $`0`$ $`=`$ $`{\displaystyle \frac{1}{3}}\mathrm{e}^{2\sigma }\left(\mathrm{\Phi }(\varphi )+{\displaystyle \frac{12}{l^2}}\right)r^2+k+\left\{r\left(\sigma ^{}\rho ^{}\right)2\right\}\mathrm{e}^{2\sigma }.`$ (76) Here $`{}_{}{}^{}\frac{d}{dr}`$. If one defines new variables $`U`$ and $`V`$ by $$U=\mathrm{e}^{\rho +\sigma },V=r^2\mathrm{e}^{\rho \sigma },$$ (77) we obtain the following equations from (73-76): $`0`$ $`=`$ $`r^3U\mathrm{\Phi }^{}(\varphi )2\alpha \left(rV\varphi ^{}\right)^{}`$ (78) $`0`$ $`=`$ $`{\displaystyle \frac{1}{3}}\mathrm{e}^{2\sigma }\left(\mathrm{\Phi }(\varphi )+{\displaystyle \frac{12}{l^2}}\right)r^3U+krV^{}`$ (79) $`0`$ $`=`$ $`{\displaystyle \frac{3U^{}}{rU}}+\alpha \left(\varphi \right)^{}.`$ (80) We should note that only three equations in (73-76) are independent. There is practical problem in the construction of AdS BH with non-trivial dilaton, especially for arbitrary dilatonic potential. That is why we use below the approximate technique which was developed in ref. for constant dilatonic potential. When $`\mathrm{\Phi }(0)=\mathrm{\Phi }^{}(0)=\varphi =0`$, a solution corresponding to the throat limit of D3-brane is given by $$U=1,V=V_0\frac{r^4}{l^2}\mu .$$ (81) In the following, we use large $`r`$ expansion and consider the perturbation around (81). It is assumed $$\mathrm{\Phi }(\varphi )=\stackrel{~}{\mu }\varphi ^2+𝒪\left(\varphi ^3\right).$$ (82) Then one can neglect the higher order terms in (82). We obtain from (78) $$0\stackrel{~}{\mu }r^3\varphi +\alpha \left(\frac{r^5}{l^2}\varphi ^{}\right)^{}.$$ (83) The solution of eq.(83) is given by $$\varphi =cr^\beta ,(c\text{ is a constant}),\beta =2\pm \sqrt{4\frac{\stackrel{~}{\mu }l^2}{\alpha }}.$$ (84) Consider $`r`$ is large or $`c`$ is small, and write $`U`$ and $`V`$ in the following form: $$U=1+c^2u,V=V_0+c^2v.$$ (85) Then from (79) and(80), one gets $$u=u_0+\frac{\alpha \beta }{6}r^{2\beta },v=v_0\frac{\stackrel{~}{\mu }(\beta 6)}{6(\beta 4)(\beta 2)}r^{2\beta +4}.$$ (86) Here $`u_0`$ and $`v_0`$ are constants of the integration. Here we choose $$v_0=u_0=0.$$ (87) The horizon which is defined by $$V=0$$ (88) lies at $$r=r_hl^{\frac{1}{2}}\mu ^{\frac{1}{4}}+c^2\frac{\stackrel{~}{\mu }(\beta 6)l^{\frac{5}{2}\beta }\mu ^{\frac{1}{4}\frac{\beta }{2}}}{24(\beta 4)(\beta 2)}.$$ (89) And the Hawking temperature is $`T`$ $`=`$ $`{\displaystyle \frac{1}{4\pi }}\left[{\displaystyle \frac{1}{r^2}}{\displaystyle \frac{dV}{dr}}\right]_{r=r_h}`$ (90) $`=`$ $`{\displaystyle \frac{1}{4\pi }}\left\{4l^{\frac{3}{2}}\mu ^{\frac{1}{4}}+c^2{\displaystyle \frac{\stackrel{~}{\mu }(\beta 6)(2\beta 3)}{6(\beta 4)(\beta 2)}}l^{\frac{1}{2}\beta }\mu ^{\frac{1}{4}\frac{\beta }{2}}\right\}.`$ We now evaluate the free energy of the black hole within the standard prescription . The free energy $`F`$ can be obtained by substituting the classical solution into the action $`S`$: $$F=TS.$$ (91) Here $`T`$ is the Hawking temperature. Using the equations of motion in (2) ($`X=\alpha `$, $`Y=0`$, $`4\lambda ^2=\frac{12}{l^2}`$), we obtain $$0=\frac{5}{3}\left(\mathrm{\Phi }(\varphi )+\frac{12}{l^2}\right)+\widehat{R}+\alpha \left(\varphi \right)^2.$$ (92) Substituting (92) into the action (1) after Wick-rotating it to the Euclid signature $`S`$ $`=`$ $`{\displaystyle \frac{1}{16\pi G}}{\displaystyle \frac{2}{3}}{\displaystyle _{M_5}}d^5\sqrt{G}\left(\mathrm{\Phi }(\varphi )+{\displaystyle \frac{12}{l^2}}\right)`$ (93) $`=`$ $`{\displaystyle \frac{1}{16\pi G}}{\displaystyle \frac{2}{3}}{\displaystyle \frac{V_{(3)}}{T}}{\displaystyle _{r_h}^{\mathrm{}}}𝑑rr^3U\left(\mathrm{\Phi }(\varphi )+{\displaystyle \frac{12}{l^2}}\right).`$ Here $`V_{(3)}`$ is the volume of the 3d space ($`d^5x\mathrm{}=\beta V_{(3)}𝑑rr^3\mathrm{}`$) and $`\beta `$ is the period of time, which can be regarded as the inverse of the temperature $`T`$ ($`\frac{1}{T}`$). The expression (93) contains the divergence. We regularize the divergence by replacing $$^{\mathrm{}}𝑑r^{r_{\mathrm{max}}}𝑑r$$ (94) and subtract the contribution from a zero temperature solution, where we choose $`\mu =c=0`$, and the solution corresponds to the vacuum or pure AdS: $$S_0=\frac{1}{16\pi G}\frac{2}{3}\frac{12}{l^2}\frac{V_{(3)}}{T}\sqrt{\frac{G_{tt}\left(r=r_{\mathrm{max}},\mu =c=0\right)}{G_{tt}\left(r=r_{\mathrm{max}}\right)}}_{r_h}^{\mathrm{}}𝑑rr^3.$$ (95) The factor $`\sqrt{\frac{G_{tt}\left(r=r_{\mathrm{max}},\mu =c=0\right)}{G_{tt}\left(r=r_{\mathrm{max}}\right)}}`$ is chosen so that the proper length of the circles which correspond to the period $`\frac{1}{T}`$ in the Euclid time at $`r_{\mathrm{max}}`$ coincides with each other in the two solutions. Then we find the following expression for the free energy, $`F`$ $`=`$ $`\underset{r_{\mathrm{max}}\mathrm{}}{lim}T\left(SS_0\right)`$ (96) $`=`$ $`{\displaystyle \frac{V_{(3)}}{2\pi Gl^2T^2}}\left[{\displaystyle \frac{l^2\mu }{8}}+c^2\mu ^{1\frac{\beta }{2}}\stackrel{~}{\mu }\left\{{\displaystyle \frac{(\beta 1)}{12\beta (\beta 4)(\beta 2)}}\right\}+\mathrm{}\right].`$ Here we assume $`\beta >2`$ or the expression $`SS_0`$ still contains the divergences and we cannot get finite results. However, the inequality $`\beta >2`$ is not always satisfied in the gauged supergravity models. In that case the expression in (96) would not be valid. One can express the free energy $`F`$ in (96) in terms of the temperature $`T`$ instead of $`\mu `$: $$F=\frac{V_{(3)}}{16\pi G}\left[\pi T^4l^6+c^2l^{84\beta }T^{42\beta }\stackrel{~}{\mu }\left(\frac{2\beta ^315\beta ^2+22\beta 4}{6\beta (\beta 4)(\beta 2)}\right)+\mathrm{}\right].$$ (97) Then the entropy $`𝒮`$ and the energy (mass) $`E`$ is given by $`𝒮`$ $`=`$ $`{\displaystyle \frac{dF}{dT}}={\displaystyle \frac{V_{(3)}}{16\pi G}}[4\pi T^3l^6`$ $`+c^2l^{84\beta }T^{32\beta }\stackrel{~}{\mu }\left({\displaystyle \frac{2\beta ^315\beta ^2+22\beta 4}{3\beta (\beta 4)}}\right)+\mathrm{}]`$ $`E`$ $`=`$ $`F+TS={\displaystyle \frac{V_{(3)}}{16\pi G}}[3\pi T^4l^6`$ (98) $`+c^2l^{84\beta }\left(\pi T^4\right)^{1\frac{\beta }{2}}\stackrel{~}{\mu }\left({\displaystyle \frac{(2\beta 3)(2\beta ^315\beta ^2+22\beta 4)}{6\beta (\beta 4)(\beta 2)}}\right)+\mathrm{}].`$ We now evaluate the mass using the surface term of the action in (4), i.e. within local surface counterterm method. The surface energy momentum tensor $`T_{ij}`$ is now defined by ($`d=4`$)<sup>6</sup><sup>6</sup>6 $`S`$ does not contribute due to the equation of motion in the bulk. The variation of $`S+S_b^{(1)}`$ gives a contribution proportional to the extrinsic curvature $`\theta _{ij}`$ at the boundary: $$\delta \left(S+S_b^{(2)}\right)=\frac{\sqrt{\widehat{g}}}{16\pi G}\left(\theta _{ij}\theta \widehat{g}_{ij}\right)\delta \widehat{g}^{ij}$$ The contribution is finite even in the limit of $`r\mathrm{}`$. Then the finite part does not depend on the parameters characterizing the black hole. Therefore after subtracting the contribution from the reference metric, which could be that of AdS, the contribution from the variation of $`S+S_b^{(1)}`$ vanishes. $`\delta S_b^{(2)}`$ $`=`$ $`\sqrt{\widehat{g}}\delta \widehat{g}^{ij}T_{ij}`$ (99) $`=`$ $`{\displaystyle \frac{1}{16\pi G}}[\sqrt{\widehat{g}}\delta \widehat{g}^{ij}\{{\displaystyle \frac{1}{2}}\widehat{g}_{ij}({\displaystyle \frac{6}{l}}+{\displaystyle \frac{l}{2}}\widehat{R}+{\displaystyle \frac{l}{4}}\mathrm{\Phi }(\varphi ))\}`$ $`+{\displaystyle \frac{l^2}{4}}n^\mu _\mu \left\{\sqrt{\widehat{g}}\delta \widehat{g}^{ij}\widehat{g}_{ij}\mathrm{\Phi }(\varphi )\right\}].`$ Note that the energy-momentum tensor is still not well-defined due to the term containing $`n^\mu _\mu `$. If we assume $`\delta \widehat{g}^{ij}𝒪\left(\rho ^{a_1}\right)`$ for large $`\rho `$ when we choose the coordinate system (5), then $$n^\mu _\mu (\delta \widehat{g}^{ij})\frac{2}{l}\delta \widehat{g}^{ij}(a_1+_\rho )().$$ (100) Or if $`\delta \widehat{g}^{ij}𝒪\left(r^{a_2}\right)`$ for large $`r`$ when we choose the coordinate system (72), then $$n^\mu _\mu (\delta \widehat{g}^{ij})\delta \widehat{g}^{ij}\mathrm{e}^\sigma (\frac{a_2}{r}+_r)().$$ (101) As we consider the black hole-like object in this section, one chooses the coordinate system (72) and assumes Eq.(101). Then mass $`E`$ of the black hole like object is given by $$E=d^{d1}x\sqrt{\stackrel{~}{\sigma }}N\delta T_{tt}\left(u^t\right)^2.$$ (102) Here we assume the metric of the reference spacetime (e.g. AdS) has the form of $`ds^2=f(r)dr^2N^2(r)dt^2+_{i,j=1}^{d1}\stackrel{~}{\sigma }_{ij}dx^idx^j`$ and $`\delta T_{tt}`$ is the difference of the $`(t,t)`$ component of the energy-momentum tensor in the spacetime with black hole like object from that in the reference spacetime, which we choose to be AdS, and $`u^t`$ is the $`t`$ component of the unit time-like vector normal to the hypersurface given by $`t=`$constant. By using the solution in (85) and (86), the $`(t,t)`$ component of the energy-momentum tensor in (99) has the following form: $`T_{tt}`$ $`=`$ $`{\displaystyle \frac{3}{16\pi G}}{\displaystyle \frac{r^2}{l^3}}[1{\displaystyle \frac{l^3\mu }{r^4}}+l^2\stackrel{~}{\mu }c^2({\displaystyle \frac{1}{12}}{\displaystyle \frac{1}{6\beta (\beta 6)}}`$ (103) $`{\displaystyle \frac{\beta 6}{6(\beta 4)(\beta 2)}}{\displaystyle \frac{(3\beta )(1+a_2)}{12}})r^{2\beta }+\mathrm{}].`$ If we assume the mass is finite, $`\beta `$ should satisfy the inequality $`\beta >2`$, as in the case of the free energy in (96) since $`\sqrt{\sigma }N\left(u^t\right)^2=lr^2`$ for the reference AdS space. Then the $`\beta `$-dependent term in (103) does not contribute to the mass and one gets $$E=\frac{3\mu V_{(3)}}{16\pi G}.$$ (104) Using (90) $$E=\frac{3l^6V_{(3)}\pi T^4}{16\pi G}\left\{1c^2\stackrel{~}{\mu }l^{24\beta }\left(\pi T^4\right)^{\frac{\beta }{2}}\frac{(\beta 6)(2\beta 3)}{(\beta 4)(\beta 2)}\right\},$$ (105) which does not agree with the result in (5). This might express the ambiguity in the choice of the regularization to make the finite action. A possible origin of it might be following. We assumed $`\varphi `$ can be expanded in the (integer) power series of $`\rho `$ in (7) when deriving the surface terms in (4). However, this assumption seems to conflict with the classical solution in (84), where the fractional power seems to appear since $`r^2\frac{1}{\rho }`$. In any case, in QFT there is no problem in regularization dependence of the results. In many cases (see example in ref.) the explicit choice of free parameters of regularization leads to coincidence of the answers which look different in different regularizations. As usually happens in QFT the renormalization is more universal as the same answers for beta-functions may be obtained while using different regularizations. That suggests that holographic renormalization group should be developed and the predictions of above calculations should be tested in it. As in the case of the c-function, we might be drop the terms containing $`\mathrm{\Phi }^{}`$ in the expression of $`S_b^{(2)}`$ in (4). Then we obtain $`S_b^{(2)}={\displaystyle \frac{1}{16\pi G}}{\displaystyle }d^dx[\sqrt{\widehat{g}}\{{\displaystyle \frac{2d2}{l}}+{\displaystyle \frac{l}{d2}}R+{\displaystyle \frac{2l}{d(d2)}}\mathrm{\Phi }(\varphi )`$ $`+{\displaystyle \frac{l}{d2}}(X(\varphi )\left(\widehat{}\varphi \right)^2+Y(\varphi )\widehat{\mathrm{\Delta }}\varphi )\}{\displaystyle \frac{l^2\mathrm{\Phi }(\varphi )}{d(d2)}}n^\mu _\mu \left(\sqrt{\widehat{g}}\right)].`$ (106) If we use the expression (5), however, the result of the mass $`E`$ in (105) does not change. ## 6 Discussion In summary, we constructed surface counterterm for gauged supergravity with single scalar and arbitrary scalar potential in three and five dimensions. As a result, the finite gravitational action and consistent stress tensor in asymptotically AdS space is found. Using this action, the regularized expressions for free energy, entropy and mass are derived for d5 dilatonic AdS black hole. From another side, finite action may be used to get the holographic conformal anomaly of boundary QFT with broken conformal invariance. Such conformal anomaly is calculated from d5 and d3 gauged SG with arbitrary dilatonic potential with the use of AdS/CFT correspondence. Due to dilaton dependence it takes extremely complicated form. Within holographic RG where identification of dilaton with some coupling constant is made, we suggested the candidate c-function for d2 and d4 boundary QFT from holographic conformal anomaly. It is shown that such proposal gives monotonic and positive c-function for few examples of dilatonic potential. We expect that our results may be very useful in explicit identification of supergravity description (special RG flow) with the particular boundary gauge theory (or its phase) which is very non-trivial task in AdS/CFT correspondence. We show that on the example of constant dilaton and special form of dilatonic potential where qualitative agreement of holographic conformal anomaly and QFT conformal anomaly (with the account of radiative corrections) from QED-like theory with single coupling constant may be achieved. Our work may be extended in various directions. First of all, we can consider large number of scalars, say 42 as in $`N=8`$ d5 SG, and construct the corresponding Weyl anomaly from the bulk side. However, this is technically very complicated problem as even in case of single scalar the complete answer for d4 anomaly takes few pages. The calculation of surface counterterm in d5 gauged SG with many scalars is slightly easier task. However, again the application of surface counterterm for the derivation of regularized thermodynamical quantities in multi-scalar AdS black holes (when they will be constructed) is complicated. Second, the generalization of surface counterterm for higher dimensions (say, $`d=7,9`$) is possible. Third, in general the extension of AdS/CFT set-up to non-conformal boundary theories is challenging problem. In this respect, better investigation of candidate c-functions from bulk and from boundary is required as well as their comparison in all detail. The related question is bulk calculation of Casimir effect in the presence of dilaton and comparison of it with QFT result, including radiative corrections. ## Acknoweledgements The work by SDO has been supported in part by CONACyT (CP, ref.990356) and in part by RFBR grant N99-02-16617. ## Appendix A Coefficients of conformal anomaly In this appendix, we give the explicit values of the coefficients appeared in the calculation of $`d=4`$ conformal anomaly. Substituting (16) into (15), we obtain $`g_{(1)ij}`$ $`=`$ $`\stackrel{~}{c}_1R_{ij}+\stackrel{~}{c}_2g_{ij}R+\stackrel{~}{c}_3g_{ij}g^{kl}_k\varphi _l\varphi `$ (107) $`+\stackrel{~}{c}_4g_{ij}{\displaystyle \frac{_k}{\sqrt{g}}}\left(\sqrt{g}g^{kl}_l\varphi \right)+\stackrel{~}{c}_5_i\varphi _j\varphi `$ $`\stackrel{~}{c}_1`$ $`=`$ $`{\displaystyle \frac{3}{6+\mathrm{\Phi }}}`$ $`\stackrel{~}{c}_2`$ $`=`$ $`{\displaystyle \frac{3\left\{\mathrm{\Phi }^26(\mathrm{\Phi }^{\prime \prime }+8V)\right\}}{2(6+\mathrm{\Phi })\left\{2\mathrm{\Phi }^2+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V)\right\}}}`$ $`\stackrel{~}{c}_3`$ $`=`$ $`{\displaystyle \frac{3\mathrm{\Phi }^2V+18V(\mathrm{\Phi }^{\prime \prime }+8V)2(6+\mathrm{\Phi })\mathrm{\Phi }^{}V^{}}{2(6+\mathrm{\Phi })(2\mathrm{\Phi }^2+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V))}}`$ $`\stackrel{~}{c}_4`$ $`=`$ $`{\displaystyle \frac{2\mathrm{\Phi }^{}V}{2\mathrm{\Phi }^2+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V)}}`$ $`\stackrel{~}{c}_5`$ $`=`$ $`{\displaystyle \frac{V}{2+\frac{\mathrm{\Phi }}{3}}}.`$ (108) Further, substituting (16) and (107) into (18), we obtain $`\varphi _{(2)}`$ $`=`$ $`d_1R^2+d_2R_{ij}R^{ij}+d_3R^{ij}_i\varphi _j\varphi `$ (109) $`+d_4Rg^{ij}_i\varphi _j\varphi +d_5R{\displaystyle \frac{1}{\sqrt{g}}}_i(\sqrt{g}g^{ij}_j\varphi )`$ $`+d_6(g^{ij}_i\varphi _j\varphi )^2+d_7\left({\displaystyle \frac{1}{\sqrt{g}}}_i(\sqrt{g}g^{ij}_j\varphi )\right)^2`$ $`+d_8g^{kl}_k\varphi _l\varphi {\displaystyle \frac{1}{\sqrt{g}}}_i(\sqrt{g}g^{ij}_j\varphi )`$ $`d_1`$ $`=`$ $`[9\mathrm{\Phi }^{}\{2(12+\mathrm{\Phi })\mathrm{\Phi }^4(864+36\mathrm{\Phi }+24\mathrm{\Phi }^2+\mathrm{\Phi }^3)\mathrm{\Phi }^{\prime \prime 2}`$ $`+192(12+\mathrm{\Phi })^2\mathrm{\Phi }^{\prime \prime }V+64\left(2592+612\mathrm{\Phi }+48\mathrm{\Phi }^2+\mathrm{\Phi }^3\right)V^2`$ $`2\mathrm{\Phi }^2\left(\left(216+30\mathrm{\Phi }+\mathrm{\Phi }^2\right)\mathrm{\Phi }^{\prime \prime }+144(10+\mathrm{\Phi })V\right)`$ $`+(6+\mathrm{\Phi })^2(24+\mathrm{\Phi })\mathrm{\Phi }^{}(\mathrm{\Phi }^{\prime \prime \prime }+8V^{})\}]/`$ $`[8(6+\mathrm{\Phi })^2\{2\mathrm{\Phi }^2+(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }\}`$ $`\times \{2\mathrm{\Phi }^2+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V)\}^2]`$ $`d_2`$ $`=`$ $`{\displaystyle \frac{9(12+\mathrm{\Phi })\mathrm{\Phi }^{}}{4(6+\mathrm{\Phi })^2\left\{2\mathrm{\Phi }^2+(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }\right\}}}`$ $`d_3`$ $`=`$ $`{\displaystyle \frac{3(3(12+\mathrm{\Phi })\mathrm{\Phi }^{}V2(144+30\mathrm{\Phi }+\mathrm{\Phi }^2)V^{})}{2(6+\mathrm{\Phi })^2(2\mathrm{\Phi }^2+(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime })}}`$ $`d_4`$ $`=`$ $`(3(6(12+\mathrm{\Phi })\mathrm{\Phi }^5V+6(108+24\mathrm{\Phi }+\mathrm{\Phi }^2)\mathrm{\Phi }^4V^{}`$ $`+4(2592+684\mathrm{\Phi }+48\mathrm{\Phi }^2+\mathrm{\Phi }^3)(\mathrm{\Phi }^{\prime \prime }+8V)((9+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }`$ $`+4(12+\mathrm{\Phi })V)V^{}(6+\mathrm{\Phi })\mathrm{\Phi }^2(3(144+30\mathrm{\Phi }+\mathrm{\Phi }^2)\mathrm{\Phi }^{\prime \prime \prime }V`$ $`+(1980\mathrm{\Phi }^{\prime \prime }+216\mathrm{\Phi }\mathrm{\Phi }^{\prime \prime }+5\mathrm{\Phi }^2\mathrm{\Phi }^{\prime \prime }+27360V+4176\mathrm{\Phi }V`$ $`+128\mathrm{\Phi }^2V)V^{})+2\mathrm{\Phi }^3(3(216+30\mathrm{\Phi }+\mathrm{\Phi }^2)\mathrm{\Phi }^{\prime \prime }V`$ $`2(2160V^2216\mathrm{\Phi }V^2+864V^{\prime \prime }+324\mathrm{\Phi }V^{\prime \prime }`$ $`+36\mathrm{\Phi }^2V^{\prime \prime }+\mathrm{\Phi }^3V^{\prime \prime }))+\mathrm{\Phi }^{}(3(864+36\mathrm{\Phi }+24\mathrm{\Phi }^2+\mathrm{\Phi }^3)\mathrm{\Phi }^{\prime \prime 2}V`$ $`+2\mathrm{\Phi }^{\prime \prime }(41472V^26912\mathrm{\Phi }V^2288\mathrm{\Phi }^2V^2+15552V^{\prime \prime }`$ $`+6696\mathrm{\Phi }V^{\prime \prime }+972\mathrm{\Phi }^2V^{\prime \prime }+54\mathrm{\Phi }^3V^{\prime \prime }+\mathrm{\Phi }^4V^{\prime \prime })`$ $`2(248832V^3+58752\mathrm{\Phi }V^3+4608\mathrm{\Phi }^2V^3`$ $`+96\mathrm{\Phi }^3V^3+15552\mathrm{\Phi }^{\prime \prime \prime }V^{}+6696\mathrm{\Phi }\mathrm{\Phi }^{\prime \prime \prime }V^{}+972\mathrm{\Phi }^2\mathrm{\Phi }^{\prime \prime \prime }V^{}`$ $`+54\mathrm{\Phi }^3\mathrm{\Phi }^{\prime \prime \prime }V^{}+\mathrm{\Phi }^4\mathrm{\Phi }^{\prime \prime \prime }V^{}+124416V^2+53568\mathrm{\Phi }V^2`$ $`+7776\mathrm{\Phi }^2V^2+432\mathrm{\Phi }^3V^2+8\mathrm{\Phi }^4V^2124416VV^{\prime \prime }`$ $`53568\mathrm{\Phi }VV^{\prime \prime }7776\mathrm{\Phi }^2VV^{\prime \prime }432\mathrm{\Phi }^3VV^{\prime \prime }`$ $`8\mathrm{\Phi }^4VV^{\prime \prime }))))/`$ $`(4(6+\mathrm{\Phi })^2(2\mathrm{\Phi }^2+(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime })(2\mathrm{\Phi }^2`$ $`+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V))^2)`$ $`d_5`$ $`=`$ $`(3(2\mathrm{\Phi }^4V+2(432+42\mathrm{\Phi }+\mathrm{\Phi }^2)\mathrm{\Phi }^{\prime \prime }V(\mathrm{\Phi }^{\prime \prime }+8V)`$ $`+\mathrm{\Phi }^2V((6+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }8(162+7\mathrm{\Phi })V)4(24+\mathrm{\Phi })\mathrm{\Phi }^3V^{}`$ $`2(432+42\mathrm{\Phi }+\mathrm{\Phi }^2)\mathrm{\Phi }^{}(\mathrm{\Phi }^{\prime \prime \prime }V\mathrm{\Phi }^{\prime \prime }V^{})))/`$ $`\left(2\left(2\mathrm{\Phi }^2(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }\right)\left(2\mathrm{\Phi }^2+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V)\right)^2\right)`$ $`d_6`$ $`=`$ $`(54(12+\mathrm{\Phi })\mathrm{\Phi }^5V^2+12(828+168\mathrm{\Phi }+5\mathrm{\Phi }^2)\mathrm{\Phi }^4VV^{}`$ $`+4\left(2592+684\mathrm{\Phi }+48\mathrm{\Phi }^2+\mathrm{\Phi }^3\right)`$ $`V^{}(54\mathrm{\Phi }^{\prime \prime 2}V+4608V^3+192\mathrm{\Phi }V^3+108\mathrm{\Phi }^{\prime \prime \prime }V^{}+24\mathrm{\Phi }\mathrm{\Phi }^{\prime \prime \prime }V^{}`$ $`+\mathrm{\Phi }^2\mathrm{\Phi }^{\prime \prime \prime }V^{}+864V^2+192\mathrm{\Phi }V^2+8\mathrm{\Phi }^2V^21728VV^{\prime \prime }`$ $`384\mathrm{\Phi }VV^{\prime \prime }16\mathrm{\Phi }^2VV^{\prime \prime }+2\mathrm{\Phi }^{\prime \prime }(504V^2+12\mathrm{\Phi }V^2108V^{\prime \prime }`$ $`24\mathrm{\Phi }V^{\prime \prime }\mathrm{\Phi }^2V^{\prime \prime }))+(6+\mathrm{\Phi })\mathrm{\Phi }^2(9(144+30\mathrm{\Phi }+\mathrm{\Phi }^2)\mathrm{\Phi }^{\prime \prime \prime }V^2`$ $`2V^{}(14796\mathrm{\Phi }^{\prime \prime }V+1368\mathrm{\Phi }\mathrm{\Phi }^{\prime \prime }V+33\mathrm{\Phi }^2\mathrm{\Phi }^{\prime \prime }V`$ $`+88992V^2+4680\mathrm{\Phi }V^2+36\mathrm{\Phi }^2V^220736V^{\prime \prime }`$ $`5472\mathrm{\Phi }V^{\prime \prime }384\mathrm{\Phi }^2V^{\prime \prime }8\mathrm{\Phi }^3V^{\prime \prime }))`$ $`+2\mathrm{\Phi }^3(27(216+30\mathrm{\Phi }+\mathrm{\Phi }^2)\mathrm{\Phi }^{\prime \prime }V^2+4(12312V^3`$ $`+1836\mathrm{\Phi }V^3+72\mathrm{\Phi }^2V^3+2376V^2+864\mathrm{\Phi }V^2+90\mathrm{\Phi }^2V^2`$ $`+2\mathrm{\Phi }^3V^2+2592VV^{\prime \prime }+972\mathrm{\Phi }VV^{\prime \prime }+108\mathrm{\Phi }^2VV^{\prime \prime }`$ $`+3\mathrm{\Phi }^3VV^{\prime \prime }))\mathrm{\Phi }^{}(27(2304+516\mathrm{\Phi }+40\mathrm{\Phi }^2+\mathrm{\Phi }^3)\mathrm{\Phi }^{\prime \prime 2}V^2`$ $`+4\mathrm{\Phi }^{\prime \prime }(217728V^3+44064\mathrm{\Phi }V^3+3024\mathrm{\Phi }^2V^3+72\mathrm{\Phi }^3V^3`$ $`+81648V^2+34992\mathrm{\Phi }V^2+5040\mathrm{\Phi }^2V^2+276\mathrm{\Phi }^3V^2`$ $`+5\mathrm{\Phi }^4V^2+46656VV^{\prime \prime }+20088\mathrm{\Phi }VV^{\prime \prime }`$ $`+2916\mathrm{\Phi }^2VV^{\prime \prime }+162\mathrm{\Phi }^3VV^{\prime \prime }+3\mathrm{\Phi }^4VV^{\prime \prime })`$ $`+4V(746496V^3+129600\mathrm{\Phi }V^3+6912\mathrm{\Phi }^2V^3`$ $`+144\mathrm{\Phi }^3V^346656\mathrm{\Phi }^{\prime \prime \prime }V^{}20088\mathrm{\Phi }\mathrm{\Phi }^{\prime \prime \prime }V^{}2916\mathrm{\Phi }^2\mathrm{\Phi }^{\prime \prime \prime }V^{}`$ $`162\mathrm{\Phi }^3\mathrm{\Phi }^{\prime \prime \prime }V^{}3\mathrm{\Phi }^4\mathrm{\Phi }^{\prime \prime \prime }V^{}404352V^2177984\mathrm{\Phi }V^2`$ $`26784\mathrm{\Phi }^2V^21584\mathrm{\Phi }^3V^232\mathrm{\Phi }^4V^2+373248VV^{\prime \prime }`$ $`+160704\mathrm{\Phi }VV^{\prime \prime }+23328\mathrm{\Phi }^2VV^{\prime \prime }+1296\mathrm{\Phi }^3VV^{\prime \prime }`$ $`+24\mathrm{\Phi }^4VV^{\prime \prime })))/(8(6+\mathrm{\Phi })^2(2\mathrm{\Phi }^2+(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime })(2\mathrm{\Phi }^2`$ $`+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V))^2)`$ $`d_7`$ $`=`$ $`(2V(36\mathrm{\Phi }^3V3(18+\mathrm{\Phi })\mathrm{\Phi }^{}V`$ $`((26+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }8(18+\mathrm{\Phi })V)+4\left(432+42\mathrm{\Phi }+\mathrm{\Phi }^2\right)\mathrm{\Phi }^2V^{}`$ $`+(18+\mathrm{\Phi })^2(24+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime \prime }V2(\mathrm{\Phi }^{\prime \prime }+4V)V^{})))/`$ $`\left(\left(2\mathrm{\Phi }^2(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }\right)\left(2\mathrm{\Phi }^2+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V)\right)^2\right)`$ $`d_8`$ $`=`$ $`(6\mathrm{\Phi }^4V^24(156+5\mathrm{\Phi })\mathrm{\Phi }^3VV^{}2(18+\mathrm{\Phi })\mathrm{\Phi }^{}V`$ $`(3(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime \prime }V+(276\mathrm{\Phi }^{\prime \prime }11\mathrm{\Phi }\mathrm{\Phi }^{\prime \prime }+480V+32\mathrm{\Phi }V)V^{})`$ $`+2(432+42\mathrm{\Phi }+\mathrm{\Phi }^2)(3\mathrm{\Phi }^{\prime \prime 2}V^2`$ $`+2(18+\mathrm{\Phi })V(\mathrm{\Phi }^{\prime \prime \prime }V^{}+8VV^{\prime \prime })`$ $`+2\mathrm{\Phi }^{\prime \prime }(12V^3+18V^2+\mathrm{\Phi }V^2+18VV^{\prime \prime }+\mathrm{\Phi }VV^{\prime \prime }))`$ $`+\mathrm{\Phi }^2(3(6+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }V^28(486V^3+21\mathrm{\Phi }V^3+432V^2`$ $`+42\mathrm{\Phi }V^2+\mathrm{\Phi }^2V^2+432VV^{\prime \prime }+42\mathrm{\Phi }VV^{\prime \prime }+\mathrm{\Phi }^2VV^{\prime \prime })))/`$ $`\left(2\left(2\mathrm{\Phi }^2(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }\right)\left(2\mathrm{\Phi }^2+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V)\right)^2\right).`$ Substituting (16), (107) and (109) into (17), one gets $`g^{ij}g_{(2)ij}`$ $`=`$ $`f_1R^2+f_2R_{ij}R^{ij}+f_3R^{ij}_i\varphi _j\varphi `$ (110) $`+f_4Rg^{ij}_i\varphi _j\varphi +f_5R{\displaystyle \frac{1}{\sqrt{g}}}_i(\sqrt{g}g^{ij}_j\varphi )`$ $`+f_6(g^{ij}_i\varphi _j\varphi )^2+f_7\left({\displaystyle \frac{1}{\sqrt{g}}}_i(\sqrt{g}g^{ij}_j\varphi )\right)^2`$ $`+f_8g^{kl}_k\varphi _l\varphi {\displaystyle \frac{1}{\sqrt{g}}}_i(\sqrt{g}g^{ij}_j\varphi )`$ $`f_1`$ $`=`$ $`[9\{2\mathrm{\Phi }^672(12+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }(\mathrm{\Phi }^{\prime \prime }+8V)^2`$ $`2\mathrm{\Phi }^4\left((24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }+8(18+\mathrm{\Phi })V\right)`$ $`+\mathrm{\Phi }^2((324+12\mathrm{\Phi }\mathrm{\Phi }^2)\mathrm{\Phi }^{\prime \prime 2}`$ $`+8(540+48\mathrm{\Phi }+\mathrm{\Phi }^2)\mathrm{\Phi }^{\prime \prime }V+64(180+24\mathrm{\Phi }+\mathrm{\Phi }^2)V^2)`$ $`+(6+\mathrm{\Phi })^2\mathrm{\Phi }^3(\mathrm{\Phi }^{\prime \prime \prime }+8V^{})\}]/`$ $`[2(6+\mathrm{\Phi })^2\{2\mathrm{\Phi }^2+(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }\}`$ $`\times \{2\mathrm{\Phi }^2+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V)\}^2]`$ $`f_2`$ $`=`$ $`{\displaystyle \frac{9(\mathrm{\Phi }^26\mathrm{\Phi }^{\prime \prime })}{(6+\mathrm{\Phi })^2\left\{2\mathrm{\Phi }^2+(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }\right\}}}`$ $`f_3`$ $`=`$ $`{\displaystyle \frac{6(3\mathrm{\Phi }^2V+18\mathrm{\Phi }^{\prime \prime }V+2(6+\mathrm{\Phi })\mathrm{\Phi }^{}V^{})}{(6+\mathrm{\Phi })^2(2\mathrm{\Phi }^2+(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime })}}`$ $`f_4`$ $`=`$ $`(3(12\mathrm{\Phi }^6V+432(12+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }V(\mathrm{\Phi }^{\prime \prime }+8V)^2`$ $`+8(6+\mathrm{\Phi })\mathrm{\Phi }^5V^{}+(6+\mathrm{\Phi })\mathrm{\Phi }^{}((1044+168\mathrm{\Phi }+7\mathrm{\Phi }^2)\mathrm{\Phi }^{\prime \prime 2}`$ $`+8\left(1476+192\mathrm{\Phi }+7\mathrm{\Phi }^2\right)\mathrm{\Phi }^{\prime \prime }V`$ $`+256(216+30\mathrm{\Phi }+\mathrm{\Phi }^2)V^2)V^{}`$ $`2(6+\mathrm{\Phi })\mathrm{\Phi }^3(3(6+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime \prime }V`$ $`+(66\mathrm{\Phi }^{\prime \prime }+3\mathrm{\Phi }\mathrm{\Phi }^{\prime \prime }+912V+88\mathrm{\Phi }V)V^{})`$ $`+4\mathrm{\Phi }^4(3(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }V2(216V^212\mathrm{\Phi }V^2+36V^{\prime \prime }`$ $`+12\mathrm{\Phi }V^{\prime \prime }+\mathrm{\Phi }^2V^{\prime \prime }))+2\mathrm{\Phi }^2(3(32412\mathrm{\Phi }+\mathrm{\Phi }^2)\mathrm{\Phi }^{\prime \prime 2}V`$ $`+2(18+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }\left(360V^212\mathrm{\Phi }V^2+36V^{\prime \prime }+12\mathrm{\Phi }V^{\prime \prime }+\mathrm{\Phi }^2V^{\prime \prime }\right)`$ $`2(17280V^3+2304\mathrm{\Phi }V^3`$ $`+96\mathrm{\Phi }^2V^3+648\mathrm{\Phi }^{\prime \prime \prime }V^{}+252\mathrm{\Phi }\mathrm{\Phi }^{\prime \prime \prime }V^{}+30\mathrm{\Phi }^2\mathrm{\Phi }^{\prime \prime \prime }V^{}`$ $`+\mathrm{\Phi }^3\mathrm{\Phi }^{\prime \prime \prime }V^{}+5184V^2+2016\mathrm{\Phi }V^2+240\mathrm{\Phi }^2V^2+8\mathrm{\Phi }^3V^2`$ $`5184VV^{\prime \prime }2016\mathrm{\Phi }VV^{\prime \prime }240\mathrm{\Phi }^2VV^{\prime \prime }8\mathrm{\Phi }^3VV^{\prime \prime }))))/`$ $`\left(2(6+\mathrm{\Phi })^2\left(2\mathrm{\Phi }^2+(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }\right)\left(2\mathrm{\Phi }^2+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V)\right)^2\right)`$ $`f_5`$ $`=`$ $`(3\mathrm{\Phi }^{}(\mathrm{\Phi }^{\prime \prime }V(3(10+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }+8(42+\mathrm{\Phi })V)`$ $`+\mathrm{\Phi }^2\left(6\mathrm{\Phi }^{\prime \prime }V+32V^2\right)+8\mathrm{\Phi }^3V^{}`$ $`+4(18+\mathrm{\Phi })\mathrm{\Phi }^{}(\mathrm{\Phi }^{\prime \prime \prime }V\mathrm{\Phi }^{\prime \prime }V^{})))/`$ $`\left(\left(2\mathrm{\Phi }^2(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }\right)\left(2\mathrm{\Phi }^2+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V)\right)^2\right)`$ $`f_6`$ $`=`$ $`(54\mathrm{\Phi }^6V^2+72(6+\mathrm{\Phi })\mathrm{\Phi }^5VV^{}`$ $`+2\mathrm{\Phi }^{\prime \prime }(54(252+30\mathrm{\Phi }+\mathrm{\Phi }^2)\mathrm{\Phi }^{\prime \prime 2}V^2`$ $`+24V(36288V^3+4320\mathrm{\Phi }V^3+144\mathrm{\Phi }^2V^3`$ $`+11664V^2+5184\mathrm{\Phi }V^2+792\mathrm{\Phi }^2V^2+48\mathrm{\Phi }^3V^2+\mathrm{\Phi }^4V^2)`$ $`+\mathrm{\Phi }^{\prime \prime }(217728V^3+25920\mathrm{\Phi }V^3+864\mathrm{\Phi }^2V^3`$ $`+11664V^2+5184\mathrm{\Phi }V^2+792\mathrm{\Phi }^2V^2+48\mathrm{\Phi }^3V^2+\mathrm{\Phi }^4V^2))`$ $`+(6+\mathrm{\Phi })\mathrm{\Phi }^3(9(6+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime \prime }V^22V^{}(666\mathrm{\Phi }^{\prime \prime }V+39\mathrm{\Phi }\mathrm{\Phi }^{\prime \prime }V`$ $`+4392V^2+156\mathrm{\Phi }V^2864V^{\prime \prime }192\mathrm{\Phi }V^{\prime \prime }8\mathrm{\Phi }^2V^{\prime \prime }))`$ $`+(6+\mathrm{\Phi })\mathrm{\Phi }^{}V^{}(3(1548+120\mathrm{\Phi }+\mathrm{\Phi }^2)\mathrm{\Phi }^{\prime \prime 2}V`$ $`+8\mathrm{\Phi }^{\prime \prime }(11124V^2+1152\mathrm{\Phi }V^2`$ $`+27\mathrm{\Phi }^2V^21944V^{\prime \prime }540\mathrm{\Phi }V^{\prime \prime }`$ $`42\mathrm{\Phi }^2V^{\prime \prime }\mathrm{\Phi }^3V^{\prime \prime })+4(18+\mathrm{\Phi })`$ $`(4608V^3+192\mathrm{\Phi }V^3+108\mathrm{\Phi }^{\prime \prime \prime }V^{}`$ $`+24\mathrm{\Phi }\mathrm{\Phi }^{\prime \prime \prime }V^{}+\mathrm{\Phi }^2\mathrm{\Phi }^{\prime \prime \prime }V^{}+864V^2`$ $`+192\mathrm{\Phi }V^2+8\mathrm{\Phi }^2V^21728VV^{\prime \prime }384\mathrm{\Phi }VV^{\prime \prime }16\mathrm{\Phi }^2VV^{\prime \prime }))`$ $`+6\mathrm{\Phi }^4(9(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }V^2+4(324V^3+18\mathrm{\Phi }V^3`$ $`+36V^2+12\mathrm{\Phi }V^2+\mathrm{\Phi }^2V^2+36VV^{\prime \prime }+12\mathrm{\Phi }VV^{\prime \prime }+\mathrm{\Phi }^2VV^{\prime \prime }))`$ $`\mathrm{\Phi }^2(27(396+36\mathrm{\Phi }+\mathrm{\Phi }^2)\mathrm{\Phi }^{\prime \prime 2}V^2+4\mathrm{\Phi }^{\prime \prime }(29160V^3`$ $`+2592\mathrm{\Phi }V^3+54\mathrm{\Phi }^2V^3+4104V^2+1620\mathrm{\Phi }V^2+198\mathrm{\Phi }^2V^2`$ $`+7\mathrm{\Phi }^3V^2+1944VV^{\prime \prime }+756\mathrm{\Phi }VV^{\prime \prime }+90\mathrm{\Phi }^2VV^{\prime \prime }+3\mathrm{\Phi }^3VV^{\prime \prime })`$ $`+4V(67392V^3+6912\mathrm{\Phi }V^3+144\mathrm{\Phi }^2V^3`$ $`1944\mathrm{\Phi }^{\prime \prime \prime }V^{}756\mathrm{\Phi }\mathrm{\Phi }^{\prime \prime \prime }V^{}`$ $`90\mathrm{\Phi }^2\mathrm{\Phi }^{\prime \prime \prime }V^{}3\mathrm{\Phi }^3\mathrm{\Phi }^{\prime \prime \prime }V^{}5184V^22016\mathrm{\Phi }V^2240\mathrm{\Phi }^2V^2`$ $`8\mathrm{\Phi }^3V^2+15552VV^{\prime \prime }+6048\mathrm{\Phi }VV^{\prime \prime }`$ $`+720\mathrm{\Phi }^2VV^{\prime \prime }+24\mathrm{\Phi }^3VV^{\prime \prime })))/`$ $`(2(6+\mathrm{\Phi })^2(2\mathrm{\Phi }^2+(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime })(2\mathrm{\Phi }^2`$ $`+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V))^2)`$ $`f_7`$ $`=`$ $`(4V(4\mathrm{\Phi }^4V2(78+5\mathrm{\Phi })\mathrm{\Phi }^2\mathrm{\Phi }^{\prime \prime }V`$ $`+(18+\mathrm{\Phi })^2\mathrm{\Phi }^{\prime \prime }V(\mathrm{\Phi }^{\prime \prime }+24V)`$ $`+8(18+\mathrm{\Phi })\mathrm{\Phi }^3V^{}+2(18+\mathrm{\Phi })^2\mathrm{\Phi }^{}(\mathrm{\Phi }^{\prime \prime \prime }V2(\mathrm{\Phi }^{\prime \prime }+4V)V^{})))/`$ $`\left(\left(2\mathrm{\Phi }^2(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }\right)\left(2\mathrm{\Phi }^2+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V)\right)^2\right)`$ $`f_8`$ $`=`$ $`(56\mathrm{\Phi }^4VV^{}4(18+\mathrm{\Phi })^2\mathrm{\Phi }^{\prime \prime }V(\mathrm{\Phi }^{\prime \prime }+24V)V^{}`$ $`4\mathrm{\Phi }^2V(3(18+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime \prime }V`$ $`(246\mathrm{\Phi }^{\prime \prime }+15\mathrm{\Phi }\mathrm{\Phi }^{\prime \prime }+288V+16\mathrm{\Phi }V)V^{})`$ $`+2\mathrm{\Phi }^3(9\mathrm{\Phi }^{\prime \prime }V^28(6V^3+18V^2+\mathrm{\Phi }V^2`$ $`+18VV^{\prime \prime }+\mathrm{\Phi }VV^{\prime \prime }))+\mathrm{\Phi }^{}`$ $`(9(10+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime 2}V^2`$ $`+8(18+\mathrm{\Phi })^2V(\mathrm{\Phi }^{\prime \prime \prime }V^{}+8VV^{\prime \prime })8\mathrm{\Phi }^{\prime \prime }(126V^3+3\mathrm{\Phi }V^3`$ $`324V^236\mathrm{\Phi }V^2\mathrm{\Phi }^2V^2`$ $`324VV^{\prime \prime }36\mathrm{\Phi }VV^{\prime \prime }\mathrm{\Phi }^2VV^{\prime \prime })))/`$ $`\left(\left(2\mathrm{\Phi }^2(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }\right)\left(2\mathrm{\Phi }^2+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V)\right)^2\right).`$ Finally substituting (16), (107), (109) and (110) into the expression for the anomaly (2), we obtain, $`T`$ $`=`$ $`{\displaystyle \frac{1}{8\pi G}}[h_1R^2+h_2R_{ij}R^{ij}+h_3R^{ij}_i\varphi _j\varphi `$ (111) $`+h_4Rg^{ij}_i\varphi _j\varphi +h_5R{\displaystyle \frac{1}{\sqrt{g}}}_i(\sqrt{g}g^{ij}_j\varphi )`$ $`+h_6(g^{ij}_i\varphi _j\varphi )^2+h_7\left({\displaystyle \frac{1}{\sqrt{g}}}_i(\sqrt{g}g^{ij}_j\varphi )\right)^2`$ $`+h_8g^{kl}_k\varphi _l\varphi {\displaystyle \frac{1}{\sqrt{g}}}_i(\sqrt{g}g^{ij}_j\varphi )]`$ $`h_1`$ $`=`$ $`[3\{(2410\mathrm{\Phi })\mathrm{\Phi }^6`$ $`+\left(62208+22464\mathrm{\Phi }+2196\mathrm{\Phi }^2+72\mathrm{\Phi }^3+\mathrm{\Phi }^4\right)\mathrm{\Phi }^{\prime \prime }(\mathrm{\Phi }^{\prime \prime }+8V)^2`$ $`+2\mathrm{\Phi }^4\left\{\left(108+162\mathrm{\Phi }+7\mathrm{\Phi }^2\right)\mathrm{\Phi }^{\prime \prime }+72\left(8+14\mathrm{\Phi }+\mathrm{\Phi }^2\right)V\right\}`$ $`2\mathrm{\Phi }^2\{(6912+2736\mathrm{\Phi }+192\mathrm{\Phi }^2+\mathrm{\Phi }^3)\mathrm{\Phi }^{\prime \prime 2}`$ $`+4\left(11232+6156\mathrm{\Phi }+552\mathrm{\Phi }^2+13\mathrm{\Phi }^3\right)\mathrm{\Phi }^{\prime \prime }V`$ $`+32(2592+468\mathrm{\Phi }+96\mathrm{\Phi }^2+5\mathrm{\Phi }^3)V^2\}`$ $`3(24+\mathrm{\Phi })(6+\mathrm{\Phi })^2\mathrm{\Phi }^3(\mathrm{\Phi }^{\prime \prime \prime }+8V^{})\}]/`$ $`[16(6+\mathrm{\Phi })^2\{2\mathrm{\Phi }^2+(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }\}\{2\mathrm{\Phi }^2`$ $`+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V)\}^2]`$ $`h_2`$ $`=`$ $`{\displaystyle \frac{3\left\{(125\mathrm{\Phi })\mathrm{\Phi }^2+(288+72\mathrm{\Phi }+\mathrm{\Phi }^2)\mathrm{\Phi }^{\prime \prime }\right\}}{8(6+\mathrm{\Phi })^2\left\{2\mathrm{\Phi }^2+(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }\right\}}}`$ $`h_3`$ $`=`$ $`(3((125\mathrm{\Phi })\mathrm{\Phi }^2V+(288+72\mathrm{\Phi }+\mathrm{\Phi }^2)\mathrm{\Phi }^{\prime \prime }V`$ $`+2(14418\mathrm{\Phi }+\mathrm{\Phi }^2)\mathrm{\Phi }^{}V^{}))/`$ $`\left(4(6+\mathrm{\Phi })^2\left(2\mathrm{\Phi }^2+(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }\right)\right)`$ $`h_4`$ $`=`$ $`(6(12+5\mathrm{\Phi })\mathrm{\Phi }^6V`$ $`+3\left(62208+22464\mathrm{\Phi }+2196\mathrm{\Phi }^2+72\mathrm{\Phi }^3+\mathrm{\Phi }^4\right)\mathrm{\Phi }^{\prime \prime }V(\mathrm{\Phi }^{\prime \prime }+8V)^2`$ $`+2\left(68448\mathrm{\Phi }+11\mathrm{\Phi }^2\right)\mathrm{\Phi }^5V^{}`$ $`+(6+\mathrm{\Phi })\mathrm{\Phi }^{}((311042772\mathrm{\Phi }+120\mathrm{\Phi }^2+13\mathrm{\Phi }^3)\mathrm{\Phi }^{\prime \prime 2}`$ $`+8\left(622087092\mathrm{\Phi }132\mathrm{\Phi }^2+7\mathrm{\Phi }^3\right)\mathrm{\Phi }^{\prime \prime }V`$ $`+384(5184504\mathrm{\Phi }+6\mathrm{\Phi }^2+\mathrm{\Phi }^3)V^2)V^{}`$ $`(6+\mathrm{\Phi })\mathrm{\Phi }^3(9(14418\mathrm{\Phi }+\mathrm{\Phi }^2)\mathrm{\Phi }^{\prime \prime \prime }V`$ $`+(3492\mathrm{\Phi }^{\prime \prime }+252\mathrm{\Phi }\mathrm{\Phi }^{\prime \prime }+19\mathrm{\Phi }^2\mathrm{\Phi }^{\prime \prime }`$ $`71712V4944\mathrm{\Phi }V+208\mathrm{\Phi }^2V)V^{})`$ $`+6\mathrm{\Phi }^4((108+162\mathrm{\Phi }+7\mathrm{\Phi }^2)\mathrm{\Phi }^{\prime \prime }V+2(288V^2`$ $`+504\mathrm{\Phi }V^2+36\mathrm{\Phi }^2V^2+864V^{\prime \prime }+252\mathrm{\Phi }V^{\prime \prime }+12\mathrm{\Phi }^2V^{\prime \prime }\mathrm{\Phi }^3V^{\prime \prime }))`$ $`6\mathrm{\Phi }^2((6912+2736\mathrm{\Phi }+192\mathrm{\Phi }^2+\mathrm{\Phi }^3)\mathrm{\Phi }^{\prime \prime 2}V`$ $`82944V^3+14976\mathrm{\Phi }V^3`$ $`+3072\mathrm{\Phi }^2V^3+160\mathrm{\Phi }^3V^315552\mathrm{\Phi }^{\prime \prime \prime }V^{}`$ $`5400\mathrm{\Phi }\mathrm{\Phi }^{\prime \prime \prime }V^{}468\mathrm{\Phi }^2\mathrm{\Phi }^{\prime \prime \prime }V^{}`$ $`+6\mathrm{\Phi }^3\mathrm{\Phi }^{\prime \prime \prime }V^{}+\mathrm{\Phi }^4\mathrm{\Phi }^{\prime \prime \prime }V^{}124416V^2`$ $`43200\mathrm{\Phi }V^23744\mathrm{\Phi }^2V^2`$ $`+48\mathrm{\Phi }^3V^2+8\mathrm{\Phi }^4V^2+124416VV^{\prime \prime }`$ $`+43200\mathrm{\Phi }VV^{\prime \prime }+3744\mathrm{\Phi }^2VV^{\prime \prime }`$ $`48\mathrm{\Phi }^3VV^{\prime \prime }8\mathrm{\Phi }^4VV^{\prime \prime }`$ $`+\mathrm{\Phi }^{\prime \prime }(44928V^2+24624\mathrm{\Phi }V^2+2208\mathrm{\Phi }^2V^2`$ $`+52\mathrm{\Phi }^3V^2+15552V^{\prime \prime }+5400\mathrm{\Phi }V^{\prime \prime }`$ $`+468\mathrm{\Phi }^2V^{\prime \prime }6\mathrm{\Phi }^3V^{\prime \prime }\mathrm{\Phi }^4V^{\prime \prime })))/`$ $`(8(6+\mathrm{\Phi })^2(2\mathrm{\Phi }^2+(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime })(2\mathrm{\Phi }^2`$ $`+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V))^2)`$ $`h_5`$ $`=`$ $`(\mathrm{\Phi }^{}(10\mathrm{\Phi }^4V+\mathrm{\Phi }^2V((426+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }8(270+\mathrm{\Phi })V)`$ $`+\mathrm{\Phi }\mathrm{\Phi }^{\prime \prime }V(7(6+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }+8(174+5\mathrm{\Phi })V)+12(24+\mathrm{\Phi })\mathrm{\Phi }^3V^{}`$ $`+6(4326\mathrm{\Phi }+\mathrm{\Phi }^2)\mathrm{\Phi }^{}(\mathrm{\Phi }^{\prime \prime \prime }V\mathrm{\Phi }^{\prime \prime }V^{})))/`$ $`\left(4\left(2\mathrm{\Phi }^2(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }\right)\left(2\mathrm{\Phi }^2+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V)\right)^2\right)`$ $`h_6`$ $`=`$ $`(18(12+5\mathrm{\Phi })\mathrm{\Phi }^6V^2+4(2772+384\mathrm{\Phi }13\mathrm{\Phi }^2)\mathrm{\Phi }^5VV^{}`$ $`\mathrm{\Phi }^{\prime \prime }(3(124416+44928\mathrm{\Phi }`$ $`+4212\mathrm{\Phi }^2+144\mathrm{\Phi }^3+\mathrm{\Phi }^4)\mathrm{\Phi }^{\prime \prime 2}V^2`$ $`+48\mathrm{\Phi }^{\prime \prime }(124416V^3+44928\mathrm{\Phi }V^3+4212\mathrm{\Phi }^2V^3+144\mathrm{\Phi }^3V^3+\mathrm{\Phi }^4V^3`$ $`23328V^210368\mathrm{\Phi }V^21584\mathrm{\Phi }^2V^296\mathrm{\Phi }^3V^22\mathrm{\Phi }^4V^2)`$ $`+64V(373248V^3+134784\mathrm{\Phi }V^3`$ $`+12636\mathrm{\Phi }^2V^3+432\mathrm{\Phi }^3V^3+3\mathrm{\Phi }^4V^3139968V^2`$ $`50544\mathrm{\Phi }V^24320\mathrm{\Phi }^2V^2+216\mathrm{\Phi }^3V^2+36\mathrm{\Phi }^4V^2+\mathrm{\Phi }^5V^2))`$ $`(6+\mathrm{\Phi })\mathrm{\Phi }^3(9(14418\mathrm{\Phi }+\mathrm{\Phi }^2)\mathrm{\Phi }^{\prime \prime \prime }V^2`$ $`2V^{}(17244\mathrm{\Phi }^{\prime \prime }V540\mathrm{\Phi }\mathrm{\Phi }^{\prime \prime }V+29\mathrm{\Phi }^2\mathrm{\Phi }^{\prime \prime }V99360V^2`$ $`+1992\mathrm{\Phi }V^2+212\mathrm{\Phi }^2V^2+20736V^{\prime \prime }+3744\mathrm{\Phi }V^{\prime \prime }8\mathrm{\Phi }^3V^{\prime \prime }))`$ $`+2(6+\mathrm{\Phi })\mathrm{\Phi }^{}V^{}((62208+3708\mathrm{\Phi }24\mathrm{\Phi }^2+\mathrm{\Phi }^3)\mathrm{\Phi }^{\prime \prime 2}V`$ $`4\mathrm{\Phi }^{\prime \prime }(248832V^211736\mathrm{\Phi }V^2+840\mathrm{\Phi }^2V^2+34\mathrm{\Phi }^3V^2`$ $`+46656V^{\prime \prime }+11016\mathrm{\Phi }V^{\prime \prime }+468\mathrm{\Phi }^2V^{\prime \prime }18\mathrm{\Phi }^3V^{\prime \prime }\mathrm{\Phi }^4V^{\prime \prime })`$ $`2(4326\mathrm{\Phi }+\mathrm{\Phi }^2)(4608V^3+192\mathrm{\Phi }V^3`$ $`+108\mathrm{\Phi }^{\prime \prime \prime }V^{}+24\mathrm{\Phi }\mathrm{\Phi }^{\prime \prime \prime }V^{}+\mathrm{\Phi }^2\mathrm{\Phi }^{\prime \prime \prime }V^{}+864V^2`$ $`+192\mathrm{\Phi }V^2+8\mathrm{\Phi }^2V^21728VV^{\prime \prime }`$ $`384\mathrm{\Phi }VV^{\prime \prime }16\mathrm{\Phi }^2VV^{\prime \prime }))`$ $`2\mathrm{\Phi }^4(3(180+438\mathrm{\Phi }+17\mathrm{\Phi }^2)\mathrm{\Phi }^{\prime \prime }V^2+4(4752V^3`$ $`+1116\mathrm{\Phi }V^3+66\mathrm{\Phi }^2V^33240V^21008\mathrm{\Phi }V^266\mathrm{\Phi }^2V^2`$ $`+2\mathrm{\Phi }^3V^22592VV^{\prime \prime }756\mathrm{\Phi }VV^{\prime \prime }36\mathrm{\Phi }^2VV^{\prime \prime }+3\mathrm{\Phi }^3VV^{\prime \prime }))`$ $`+4\mathrm{\Phi }^2(6(2484+1197\mathrm{\Phi }+84\mathrm{\Phi }^2+2\mathrm{\Phi }^3)\mathrm{\Phi }^{\prime \prime 2}V^2`$ $`+\mathrm{\Phi }^{\prime \prime }(88128V^3+67608\mathrm{\Phi }V^3+5040\mathrm{\Phi }^2V^3+90\mathrm{\Phi }^3V^3125712V^2`$ $`46656\mathrm{\Phi }V^24896\mathrm{\Phi }^2V^272\mathrm{\Phi }^3V^2+5\mathrm{\Phi }^4V^246656VV^{\prime \prime }`$ $`16200\mathrm{\Phi }VV^{\prime \prime }1404\mathrm{\Phi }^2VV^{\prime \prime }+18\mathrm{\Phi }^3VV^{\prime \prime }+3\mathrm{\Phi }^4VV^{\prime \prime })`$ $`+3V(82944V^3+30528\mathrm{\Phi }V^3`$ $`+3840\mathrm{\Phi }^2V^3+80\mathrm{\Phi }^3V^3+15552\mathrm{\Phi }^{\prime \prime \prime }V^{}`$ $`+5400\mathrm{\Phi }\mathrm{\Phi }^{\prime \prime \prime }V^{}+468\mathrm{\Phi }^2\mathrm{\Phi }^{\prime \prime \prime }V^{}6\mathrm{\Phi }^3\mathrm{\Phi }^{\prime \prime \prime }V^{}\mathrm{\Phi }^4\mathrm{\Phi }^{\prime \prime \prime }V^{}+72576V^2`$ $`+28224\mathrm{\Phi }V^2+3360\mathrm{\Phi }^2V^2+112\mathrm{\Phi }^3V^2124416VV^{\prime \prime }`$ $`43200\mathrm{\Phi }VV^{\prime \prime }3744\mathrm{\Phi }^2VV^{\prime \prime }+48\mathrm{\Phi }^3VV^{\prime \prime }+8\mathrm{\Phi }^4VV^{\prime \prime })))/`$ $`(16(6+\mathrm{\Phi })^2(2\mathrm{\Phi }^2+(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime })(2\mathrm{\Phi }^2`$ $`+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V))^2)`$ $`h_7`$ $`=`$ $`(V(84\mathrm{\Phi }^4V8(18+\mathrm{\Phi })^2\mathrm{\Phi }^{\prime \prime }V(3\mathrm{\Phi }^{\prime \prime }+2(12+\mathrm{\Phi })V)`$ $`+\mathrm{\Phi }^2V\left(3\left(87640\mathrm{\Phi }+\mathrm{\Phi }^2\right)\mathrm{\Phi }^{\prime \prime }+8(18+\mathrm{\Phi })^2V\right)`$ $`4\left(4326\mathrm{\Phi }+\mathrm{\Phi }^2\right)\mathrm{\Phi }^3V^{}`$ $`(24+\mathrm{\Phi })(18+\mathrm{\Phi })^2\mathrm{\Phi }^{}(\mathrm{\Phi }^{\prime \prime \prime }V2(\mathrm{\Phi }^{\prime \prime }+4V)V^{})))/`$ $`\left(\left(2\mathrm{\Phi }^2(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }\right)\left(2\mathrm{\Phi }^2+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V)\right)^2\right)`$ $`h_8`$ $`=`$ $`(10\mathrm{\Phi }^5V^2+4(204+5\mathrm{\Phi })\mathrm{\Phi }^4VV^{}`$ $`+32(18+\mathrm{\Phi })^2\mathrm{\Phi }^{\prime \prime }V(3\mathrm{\Phi }^{\prime \prime }+2(12+\mathrm{\Phi })V)V^{}`$ $`+2\mathrm{\Phi }^2V(3(4326\mathrm{\Phi }+\mathrm{\Phi }^2)\mathrm{\Phi }^{\prime \prime \prime }V`$ $`+(7416\mathrm{\Phi }^{\prime \prime }+270\mathrm{\Phi }\mathrm{\Phi }^{\prime \prime }11\mathrm{\Phi }^2\mathrm{\Phi }^{\prime \prime }`$ $`+1728V480\mathrm{\Phi }V32\mathrm{\Phi }^2V)V^{})`$ $`+\mathrm{\Phi }^3((426+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }V^28(270V^3+\mathrm{\Phi }V^3`$ $`+432V^2+6\mathrm{\Phi }V^2\mathrm{\Phi }^2V^2+432VV^{\prime \prime }+6\mathrm{\Phi }VV^{\prime \prime }\mathrm{\Phi }^2VV^{\prime \prime }))`$ $`+\mathrm{\Phi }^{}(6\mathrm{\Phi }(7\mathrm{\Phi }^{\prime \prime 2}V^2232\mathrm{\Phi }^{\prime \prime }V^3+360\mathrm{\Phi }^{\prime \prime \prime }VV^{}`$ $`360\mathrm{\Phi }^{\prime \prime }V^2360\mathrm{\Phi }^{\prime \prime }VV^{\prime \prime }2880V^2V^{\prime \prime })`$ $`+4\mathrm{\Phi }^3\left(\mathrm{\Phi }^{\prime \prime \prime }VV^{}\mathrm{\Phi }^{\prime \prime }V^2\mathrm{\Phi }^{\prime \prime }VV^{\prime \prime }8V^2V^{\prime \prime }\right)`$ $`+31104\left(\mathrm{\Phi }^{\prime \prime \prime }VV^{}+\mathrm{\Phi }^{\prime \prime }V^2+\mathrm{\Phi }^{\prime \prime }VV^{\prime \prime }+8V^2V^{\prime \prime }\right)`$ $`\mathrm{\Phi }^2(7\mathrm{\Phi }^{\prime \prime 2}V^240\mathrm{\Phi }^{\prime \prime }V^348\mathrm{\Phi }^{\prime \prime \prime }VV^{}+48\mathrm{\Phi }^{\prime \prime }V^2`$ $`+48\mathrm{\Phi }^{\prime \prime }VV^{\prime \prime }+384V^2V^{\prime \prime })))/`$ $`\left(4\left(2\mathrm{\Phi }^2(24+\mathrm{\Phi })\mathrm{\Phi }^{\prime \prime }\right)\left(2\mathrm{\Phi }^2+(18+\mathrm{\Phi })(\mathrm{\Phi }^{\prime \prime }+8V)\right)^2\right).`$ The c functions proposed in this paper for $`d=4`$ case is given by $`h_1`$ and $`h_2`$ by putting $`\mathrm{\Phi }^{}`$ to vanish: $`c_1`$ $`=`$ $`{\displaystyle \frac{2\pi }{3G}}{\displaystyle \frac{62208+22464\mathrm{\Phi }+2196\mathrm{\Phi }^2+72\mathrm{\Phi }^3+\mathrm{\Phi }^4}{(6+\mathrm{\Phi })^2(24+\mathrm{\Phi })(18+\mathrm{\Phi })}}`$ $`c_2`$ $`=`$ $`{\displaystyle \frac{3\pi }{G}}{\displaystyle \frac{288+72\mathrm{\Phi }+\mathrm{\Phi }^2}{(6+\mathrm{\Phi })^2(24+\mathrm{\Phi })}}`$ (112) Note also that using of above condition on the zero value of dilatonic potential derivative on conformal boundary significally simplifies the conformal anomaly as many terms vanish. ## Appendix B Comparison with other counterterm schemes In this Appendix, we compare the counter terms and the trace anomaly obtained here with those in ref. which appeared after this work has been submitted to hepth. For simplicity, we consider the case that the spacetime dimension is 4 and the boundary is flat and the metric $`g_{ij}`$ in (5) on the boundary is given by $$g_{ij}=F(\rho )\eta _{ij}.$$ (113) We also assume the dilaton $`\varphi `$ only depends on $`\rho `$: $`\varphi =\varphi (\rho )`$. This is exactly the case of ref.. Then the conformal anomaly (22) vanishes on such background. Let us demonstrate that this is consistent with results of ref.. In the metric (113), the equation of motion (2) given by the variation over the dilaton $`\varphi `$ and the Einstein equations in (3) have the following forms: $`0`$ $`=`$ $`{\displaystyle \frac{l}{2\rho ^3}}F^2\mathrm{\Phi }^{}(\varphi ){\displaystyle \frac{2}{l}}_\rho \left({\displaystyle \frac{F^2}{\rho }}_\rho \varphi \right)`$ (114) $`0`$ $`=`$ $`{\displaystyle \frac{l^2}{12\rho ^2}}\left(\mathrm{\Phi }(\varphi )+{\displaystyle \frac{12}{l^2}}\right){\displaystyle \frac{1}{\rho ^2}}{\displaystyle \frac{2}{F}}_\rho ^2F+{\displaystyle \frac{1}{F^2}}\left(_\rho F\right)^2{\displaystyle \frac{1}{2}}\left(_\rho \varphi \right)^2`$ (115) $`0`$ $`=`$ $`{\displaystyle \frac{F}{3\rho }}\left(\mathrm{\Phi }(\varphi )+{\displaystyle \frac{12}{l^2}}\right){\displaystyle \frac{2\rho }{l^2}}_\rho ^2F{\displaystyle \frac{2\rho }{l^2F}}\left(_\rho F\right)^2+{\displaystyle \frac{6}{l^2}}_\rho F{\displaystyle \frac{4F}{l^2\rho }}.`$ (116) Eq.(115) corresponds to $`\mu =\nu =\rho `$ component in (3) and (116) to $`\mu =\nu =i`$. Other components equations in (3) vanish identically. Combining (115) and (116), we obtain $`0`$ $`=`$ $`{\displaystyle \frac{l^2}{4\rho ^2}}\left(\mathrm{\Phi }(\varphi )+{\displaystyle \frac{12}{l^2}}\right)+{\displaystyle \frac{3}{\rho ^2}}+{\displaystyle \frac{3}{F^2}}\left(_\rho F\right)^2{\displaystyle \frac{6}{\rho F}}_\rho F{\displaystyle \frac{1}{2}}\left(_\rho \varphi \right)^2`$ (117) $`0`$ $`=`$ $`{\displaystyle \frac{6}{F}}_\rho ^2F+{\displaystyle \frac{6}{F^2}}\left(_\rho F\right)^2{\displaystyle \frac{6}{\rho F}}_\rho F2\left(_\rho \varphi \right)^2.`$ (118) If we define a new variable $`A`$, which corresponds to the exponent in the warp factor by $$F=\rho \mathrm{e}^{2A},$$ (119) Eq.(118) can be rewritten as $$0=\frac{6}{\rho }_\rho \left(\rho _\rho A\right)\left(_\rho \varphi \right)^2.$$ (120) Now we further define a new variable $`B`$ by $$B\rho _\rho A.$$ (121) If $`\frac{\varphi }{\rho }0`$, we can regard $`B`$ as a function of $`\varphi `$ instead of $`\rho `$ and one obtains $$_\rho B=\frac{B}{\varphi }\frac{\varphi }{\rho }.$$ (122) By substituting (121) and (122) into (120), we find (by assuming $`\frac{\varphi }{\rho }0`$) $$\frac{B}{\varphi }=\frac{1}{6\rho }_\rho \varphi .$$ (123) Using (117) and (123) (and also (119) and (121)), we find that the dilaton equations motion (114) is automatically satisfied. In , another counterterms scheme is proposed $$S_{\mathrm{BGM}}^{(2)}=\frac{1}{16\pi G}d^4x\sqrt{\widehat{g}}\left\{\frac{6u(\varphi )}{l}+\frac{l}{2u(\varphi )}R\right\},$$ (124) instead of (4). Here $`u`$ is obtained in terms of this paper as follows: $$u(\varphi )^2=1+\frac{l^2}{12}\mathrm{\Phi }(\varphi ).$$ (125) Then based on the counter terms in (124), the following expression of the trace anomaly is given in :<sup>7</sup><sup>7</sup>7 The radial coordinate $`r`$ in is related to $`\rho `$ by $`dr=\frac{ld\rho }{4\rho }`$. Therefore $`_r=\frac{2\rho }{l}_\rho `$, especially $`_rA=\frac{2\rho }{l}_\rho A=\frac{2}{l}B`$. $$T=\frac{3}{2\pi Gl}(2Bu).$$ (126) The above trace anomaly was evaluated for fixed but finite $`\rho `$. If the boundary is asymptotically AdS, $`F`$ in (113) goes to a constant $`FF_0`$ ($`F_0`$: a constant). Then from (119) and (121, we find the behaviors of $`A`$ and $`B`$ as $$A\frac{1}{2}\mathrm{ln}\frac{F_0}{\rho },B\frac{1}{2}.$$ (127) Then (123) tells that the dilaton $`\varphi `$ becomes a constant. Then (117) tells that $$u=\sqrt{1+\frac{l^2}{12}\mathrm{\Phi }(\varphi )}1.$$ (128) Eqs.(127) and (128) tell that the trace anomaly (126) vanishes on the boundary. Thus, we demonstrated that trace anomaly of vanishes in the UV limit what is expected also from AdS/CFT correspondence. We should note that the trace anomaly (22) is evaluated on the boundary, i.e., in the UV limit. We evaluated the anomaly by expandind the action in the power series of $`ϵ`$ in (6) and subtracting the divergent terms in the limit of $`ϵ0`$. If we evaluate the anomaly for finite $`\rho `$ as in , the terms with positive power of $`ϵ`$ in the expansion do not vanish and we would obtain non-vanishing trace anomaly in general. Thus, the trace anomaly obtained in this paper does not not have any contradiction with that in . ## Appendix C Remarks on boundary values From the leading order term in the equations of motion $$0=\sqrt{\widehat{G}}\frac{\mathrm{\Phi }(\varphi _1,\mathrm{},\varphi _N)}{\varphi _\beta }_\mu \left(\sqrt{\widehat{G}}\widehat{G}^{\mu \nu }_\nu \varphi _\beta \right),$$ (129) which are given by variation of the action (130) $`S={\displaystyle \frac{1}{16\pi G}}{\displaystyle _{M_{d+1}}}d^{d+1}x\sqrt{\widehat{G}}\left\{\widehat{R}{\displaystyle \underset{\alpha =1}{\overset{N}{}}}{\displaystyle \frac{1}{2}}(\widehat{}\varphi _\alpha )^2+\mathrm{\Phi }(\varphi _1,\mathrm{},\varphi _N)+4\lambda ^2\right\}.`$ (130) with respect to $`\varphi _\alpha `$, we obtain $$\frac{\mathrm{\Phi }(\varphi _{(0)})}{\varphi _{(0)\alpha }}=0.$$ (131) The equation (131) gives one of the necessary conditions that the spacetime is asymptotically AdS. The equation (131) also looks like a constraint that the boundary value $`\varphi _{(0)}`$ must take a special value satisfying (131) for the general fluctuations but it is not always correct. The condition $`\varphi =\varphi _{(0)}`$ at the boundary is, of course, the boundary condition, which is not a part of the equations of motion. Due to the boundary condition, not all degrees of freedom of $`\varphi `$ are dynamical. Here the boundary value $`\varphi _{(0)}`$ is, of course, not dynamical. This tells that we should not impose the equations given only by the variation over $`\varphi _{(0)}`$. The equation (131) is, in fact, only given by the variation of $`\varphi _{(0)}`$. In order to understand the situation, we choose the metric in the following form $$ds^2\widehat{G}_{\mu \nu }dx^\mu dx^\nu =\frac{l^2}{4}\rho ^2d\rho d\rho +\underset{i=1}{\overset{d}{}}\widehat{g}_{ij}dx^idx^j,\widehat{g}_{ij}=\rho ^1g_{ij},$$ (132) (If $`g_{ij}=\eta _{ij}`$, the boundary of AdS lies at $`\rho =0`$.) and we use the regularization for the action (130) by introducing the infrared cutoff $`ϵ`$ and replacing $$d^{d+1}xd^dx_ϵ𝑑\rho ,_{M_d}d^dx\left(\mathrm{}\right)d^dx\left(\mathrm{}\right)|_{\rho =ϵ}.$$ (133) Then the action (130) has the following form: $$S=\frac{l}{16\pi G}\frac{1}{d}ϵ^{\frac{d}{2}}_{M_d}d^dx\sqrt{\widehat{g}_{(0)}}\left\{\mathrm{\Phi }(\varphi _{1(0)},\mathrm{},\varphi _{N(0)})\frac{8}{l^2}\right\}+𝒪\left(ϵ^{\frac{d}{2}+1}\right).$$ (134) Then it is clear that Eq.(131) can be derived only from the variation over $`\varphi _{(0)}`$ but not other components $`\varphi _{(i)}`$ ($`i=1,2,3,\mathrm{}`$). Furthermore, if we add the surface counterterm $`S_b^{(1)}`$ $$S_b^{(1)}=\frac{1}{16\pi G}\frac{d}{2}ϵ^{\frac{d}{2}}_{M_d}d^dx\sqrt{\widehat{g}_{(0)}}\mathrm{\Phi }(\varphi _{1(0)},\mathrm{},\varphi _{N(0)})$$ (135) to the action (130), the first $`\varphi _{(0)}`$ dependent term in (134) is cancelled and we find that Eq.(131) cannot be derived from the variational principle. The surface counterterm in (135) is a part of the surface counterterms, which are necessary to obtain the well-defined AdS/CFT correspondence. Since the volume of AdS is infinte, the action (130) contains divergences, a part of which appears in (134). Then in order that we obtain the well-defined AdS/CFT set-up, we need the surface counterterms to cancell the divergence.
warning/0001/cond-mat0001362.html
ar5iv
text
# The jamming transition of Granular Media ## I Introduction Despite their importance for industrial applications, non-thermal disordered systems as granular media have just recently begun to be systematically studied in the physics community . In particular, concepts from Statistical Mechanics seems to be successful to describe these systems as suggested in his pioneering work by Edwards . Actually, granular media are composed by a large number of single grains and, as much as systems of standard Statistical Mechanics, each of their macroscopic states corresponds to a huge number of microstates. Furthermore, they show very general reproducible macroscopic behaviours whose general properties are not material specific and which are statistically characterised by very few control parameters, such as their density, load or amplitude of external drive . Granular media are “non-thermal” systems since thermal energy plays no role with respect, for instance, to gravitational energies involved in grain displacements, however, thermal motion can be replaced by agitation induced by shaking or other driving and this allows the system to explore its space of configurations. A “tapping” dynamics has several control parameters. A very important one, related to the amplitude of vibrations of the external driving, has appeared to be the adimensional ratio $`\mathrm{\Gamma }_{ex}=a/g`$ , where $`a`$ is the shake peak acceleration and $`g`$ the gravity acceleration constant. It has been suggested that $`\mathrm{\Gamma }_{ex}`$ plays a role very similar to “temperature” in thermal systems. For simplicity, below we suppose the characteristic frequency, $`\omega `$, of external driving is fixed, and actually in typical experiments as those in Ref.s we are interested in, it plays a minor role with respect to $`\mathrm{\Gamma }_{ex}`$ (see ). Material parameters such as restitution, friction coefficients, presence of moisture, etc., are also to be usually considered, but they generally do not affect the overall behaviours. Experimentally it is found that at low $`\mathrm{\Gamma }_{ex}`$ relaxation processes as density compaction under tapping are logarithmically slow. An essential ingredient in the dynamics of dense granular media is, instead, the presence of “frustrating” mechanisms on the motion of grains due to steric hindrance and hard core repulsion. We have introduced Frustrated Lattice Gases to describe the dynamics of granular materials in the regime of high densities or not too strong shaking . They interestingly allow to interpret, in a single framework, several different properties of granular media ranging from logarithmic compaction, density fluctuations, “irreversible-reversible” cycles, aging, breakdown of fluctuation dissipation relation, segregation, to avalanches effects, the Reynolds transition and several others . In particular they predict the existence of a definite “jamming” transition in granular media, in strict correspondence with the glass transition of glass formers and spin glasses . Furthermore, they are new important applications of the statistical mechanics to powders. ## II Frustrated Lattice Gas models for granular media The very simple schematic models for gently shaken granular media we introduced are based on a drastic reduction of the degrees of freedom of the systems to those we suppose are essential. The models consist of a system of elongated grains which move on a lattice. Grains are subject to gravity and, eventually, to shaking, which is simulated with a driven diffusion like Monte Carlo dynamics. The crucial ingredient of the models is the presence of geometric constraints on the motion of the grains, thus, they have been called Frustrated Lattice Gases . Particles in our models occupy the sites of, say, a square (or cubic) lattice (see Fig.1). They have also an internal degree of freedom $`S_i=\pm 1`$ corresponding to their possible “orientations” along the two lattice axes. Two nearest neighbour sites can be both occupied only if the particles do not overlap. Each “experimental” tap can be divided in two parts: one where the grains average kinetic energy is finite and a second where it goes to zero. Thus, in our models grains undergo a schematic driven diffusive Monte Carlo (MC) dynamics: in absence of vibrations they are subject only to gravity and they can move just downwards always fulfilling the non overlap condition; the presence of vibration is introduced by allowing the particles to diffuse with a probability $`p_{up}`$ to move upwards and a probability $`p_{down}=1p_{up}`$ to move downwards. The quantity $`\mathrm{ln}(x_0)/2`$, with $`x_0=p_{up}/p_{down}`$, as we will see, plays the role of an effective temperature and can be related to the experimental tap vibrations intensity, $`\mathrm{\Gamma }_{ex}`$. ### A Hamiltonian description The general model introduced above can be described in terms of a standard lattice gas of Statistical Mechanics . The system Hamiltonian must have an hard core repulsion term ($`J\mathrm{}`$): $$H_{HC}=J\underset{ij}{}f_{ij}(S_i,S_j)n_in_j.$$ (1) Here $`n_i=0,1`$ are occupancy variables describing the positions of grains, $`S_i=\pm 1`$ are “spin” variables associated to the orientations of the particles, $`J`$ represents the infinite repulsion felt by the particles when they have the wrong orientations. The hard core repulsion function $`f_{ij}(S_i,S_j)`$ is 0 or 1 depending whether the configuration $`S_i,S_j`$ is right (allowed) or wrong (not allowed), see Fig.1. The choice of $`f_{ij}(S_i,S_j)`$ depends on the particular model. In particular, here we consider two models: the Tetris and the Ising Frustrated Lattice Gas (IFLG). The Tetris model is made of elongated particles (see Fig. 1), which may point in two (orthogonal) directions coinciding with the two lattice bond orientations. In this case $`f_{ij}(S_i,S_j)`$ is given by $`f_{ij}^{Tetris}(S_i,S_j)=1/2(S_iS_jϵ_{ij}(S_i+S_j)+1)`$ here $`ϵ_{ij}=+1`$ for bonds along one direction of the lattice and $`ϵ_{ij}=1`$ for bonds on the other. In order to have a non trivial behaviour the dynamics of the Tetris has a crucial purely kinetic constraint: particles can flip their “spin” only if three of their own neighbours are empty. A real granular system may contain more disorder due to the presence of a wider grain shape distribution or to the absence of a regular underlying lattice. Typically, each grain moves in the disordered environment generated by its neighbours. In order to schematically consider these effects within the above context, an other kind of “frustrated lattice gas” was introduced, made of grains moving in a lattice with quenched geometric disorder. Such a model, the Ising Frustrated Lattice Gas (IFLG), has the following hard core repulsion function, $`f_{ij}(S_i,S_j)`$ : $`f_{ij}^{IFLG}(S_i,S_j)=1/2(ϵ_{ij}S_iS_j1)`$ where $`ϵ_{ij}=\pm 1`$ are quenched random interactions associated to the edges of the lattice, describing the fact that particles must satisfy the geometric constraint of the environment considered as “practically” quenched (see Fig.1). The IFLG shows a non trivial dynamics without the necessity to introduce kinetic constraints. The phase diagram of the Tetris Hamiltonian corresponds to the usual antiferromagnetic Ising model with dilution. The Hamiltonian of the IFLG exhibits richer behaviours. In the limit where all sites are occupied ($`n_i=1`$ $`i`$), it becomes equal to the usual $`\pm J`$ Ising Edwards-Anderson Spin Glass . In the limit $`J\mathrm{}`$ (which we consider below), a version of Site Frustrated Percolation is recovered . The other important contribution to the full Hamiltonian of a granular pack we consider below must be gravitational energy: $`=H_{HC}+H_G`$, where $`H_G=g_in_iy_i`$, and $`g`$ is the gravity constant and $`y_i`$ is the height of particle $`i`$ (grains mass and lattice spacing are set to unity). The temperature, $`T`$, of the present Hamiltonian system (with $`J=\mathrm{}`$) is related to the ratio $`x_0=p_{up}/p_{down}`$ via the following relation: $`e^{2g/T}=x_0`$. It is useful to define the adimensional quantity $`\mathrm{\Gamma }1/\mathrm{ln}(x_0^{1/2})=T/g`$, which we assume plays the same role as the amplitude of the vibrations in real granular matter, that is $`\mathrm{\Gamma }`$ is a smooth function of $`\mathrm{\Gamma }_{ex}`$. It is easy to show that if the system reaches equilibrium, the temperature $`T`$ is: $$T^1=\frac{\mathrm{ln}\mathrm{\Omega }}{E}$$ (2) where $`\mathrm{\Omega }`$ is the number of configurations corresponding to the gravitational energy E. Note that for samples with constant particle density, $`E`$ is proportional to the volume and thus $`1/T`$ coincides with Edwards’ compactivity apart from a proportionality constant. ## III The dynamics of compaction To describe experimental observations about grain density relaxation under a sequence of taps a logarithmic law was proposed in Ref. : $$\rho (t_n)=\rho _{\mathrm{}}\mathrm{\Delta }\rho _{\mathrm{}}/[1+B\mathrm{ln}(t_n/\tau _1+1)]$$ (3) This law has proved to be satisfied very well by relaxation data in the IFLG model , which can be excellently rescaled with experimental data. In Fig.2 MC compaction data for four different amplitudes, $`\mathrm{\Gamma }`$, as well as the experimental data for three different amplitudes, $`\mathrm{\Gamma }_{ex}`$, are collapsed on a single curve using eq.(3). The agreement is very satisfactory (details about these data are in Ref. ). Interestingly, the results from Tetris are similar , but the asymptotic density $`\rho _{\mathrm{}}`$ is numerically indistinguishable from 1, thus almost independent on $`x_0`$, a fact in contrast with both IFLG and experimental results . ## IV Glassy behaviours of granular media We have seen that compaction shows extremely long relaxation times. Actually we show that granular systems are typically in off equilibrium configurations and that they may undergo a “jamming” transition. In particular, in the same framework of the above models, we show how experimental results about “memory” effects (the so called “irreversible-reversible” cycles) outline the existence of a dynamical glassy transition, which can be defined in a similar way to the glassy transition in real glass formers. It has been shown in mean field approximation and numerically for finite dimensional systems , that the IFLG, in absence of gravity, exhibits a spin glass (SG) transition at high density (or low temperature) similar to the one found in the p-spin model. J.J. Arenzon has extended the mean field solution of the IFLG in presence of gravity , showing that at low $`\mathrm{\Gamma }`$ the system is frozen in a SG-like phase, but at higher $`\mathrm{\Gamma }`$ it separates in a frozen SG phase at the bottom and a fluid phase on top. Above a critical $`\mathrm{\Gamma }`$ the system is entirely fluid . ### A “Irreversible-reversible” cycles and the dynamical glass transition Tapping experiments typically show “irreversible-reversible” cycles (see Fig.3): during a sequence of taps, if the system is successively shaken at increasing vibration amplitudes, its bulk density typically grows, as in compaction, and then, after a characteristic $`\mathrm{\Gamma }_{ex}`$ starts decreasing. However, if the amplitude of shaking is decreased back to zero, the density generally doesn’t follow the same path since it keeps growing. In this sense these observations outlines the existence of “memory effects”. In analogy to real experiments, cycles of taps were performed in the IFLG in which the vibration amplitude $`\mathrm{\Gamma }`$ was varied in sequences of increments $`\gamma =\mathrm{\Delta }\mathrm{\Gamma }/\tau `$, with constant taps duration $`\tau `$. After each tap the static bulk density of the system $`\rho (\mathrm{\Gamma }_n)`$ ($`n`$ is the $`n`$-th tap number) was then measured. The data are qualitatively very similar to those reported in real experiments on dry granular packs . Furthermore, they allow to define a “jamming” transition point $`\mathrm{\Gamma }_g(\gamma )`$ in analogy to the glass transition in glass formers , as explained in Fig.4. ### B Two-times correlation functions Granular media in the above dynamical situations show “aging” too . To see this, as much as in glassy systems, it is useful to consider two time correlation functions as ($`tt^{}`$): $`C(t,t^{})=[\rho (t)\rho (t^{})\rho (t)\rho (t^{})]/[\rho (t^{})^2\rho (t^{})^2]`$ where $`\rho (t)`$ is the bulk density of the system at time $`t`$. The fit, at low $`\mathrm{\Gamma }`$, of the two time correlation function, $`C(t,t^{})`$, (on five decades in MC time), reveals the following interesting approximate scaling form: $$C(t,t^{})=(1c_{\mathrm{}})\frac{\mathrm{ln}[(t^{}+t_s)/\tau ]}{\mathrm{ln}[(t+t_s)/\tau ]}+c_{\mathrm{}}$$ (4) where $`\tau `$ (which now is not the “tap duration” as before), $`t_s`$ and $`c_{\mathrm{}}`$ are fit parameters. Very interesting is the fact that the above behaviour is found in both the discussed models (Tetris and IFLG) . The data for the two models, for several values of $`\mathrm{\Gamma }`$, rescaled on a single universal master function, are plotted in Fig. 5. It is interesting the existence of such a scaling behaviours in off equilibrium dynamics of apparently different systems . ### C Response functions and the “fluctuation-dissipation” relation In Ref. it was argued that in granular media it is possible to formulate a link between response and fluctuations, i.e., the analogous of a “fluctuation dissipation theorem” (FDT) (see ), where the amplitude of external vibrations plays the role of the usual “temperature”. In typical situations the FDT coincides neither with its usual version at equilibrium nor with the extensions valid in off equilibrium thermal systems in the so called “small entropy production” limit . The origin of universalities in off equilibrium dynamics in the above heterogeneous classes of materials is still unanswered. Our systems are “shaken” at a given amplitude $`x_0`$ and the average grains height, $`h_0(t)=H(t)`$ (with $`H(t)=_iy_i(t)/N`$ ), recorded as long as the “mean square displacement” $`B(t,t^{})[H(t)H(t^{})]^2`$. To measure the response function, the average height, $`h_1(t,t_w)`$, was also recorded in an identical copy of the system (a “replica”) perturbed by a small increase of the shaking amplitude, after a time $`t_w`$. The difference in heights between the perturbed and unperturbed systems $`\mathrm{\Delta }h(t,t_w)=h_1(t,t_w)h_0(t)`$ is by definition the integrated response. FDT or its generalisations concern the relation between $`\mathrm{\Delta }h(t,t_w)`$ and the displacement, $`B(t,t_w)`$. In analogy to thermal systems , in Ref. was proposed that: $$\mathrm{\Delta }h(t,t_w)\frac{X}{2}\mathrm{\Delta }(\mathrm{\Gamma }^1)B(t,t_w)$$ (5) Eq.(5) states that, after transients, in a granular system the measure of the height variation after a change in shaking amplitude, should be proportional to the “displacement” recorded during an unperturbed run. In the simplest situations, as equilibrium thermal systems, the proportionality factor, $`X`$, is a constant, but more generally, $`X`$ is function of $`t_w`$ and $`t`$ themselves. Interestingly, in glassy systems the quantity $`g\mathrm{\Gamma }/X`$ has the meaning of an “effective temperature” . In the present model, eq. (5) seems to be approximately valid, also if in typical off-equilibrium situations $`X`$ slowly depends on $`\mathrm{\Gamma }`$ and $`t_w`$. This is shown, for high $`x_0`$, in the top panels of Fig.6: in the long time regime $`X`$ is equal to 1, showing that, in the high “temperature” and low density region, the usual equilibrium version of FDT is obeyed. In the low $`x_0`$ region, the above picture changes, as shown in the bottom panels Fig.6: after an early transient the response, $`\mathrm{\Delta }h`$, is negative and, thus, $`X`$ is negative, which would asymptotically correspond to a negative “effective temperature”. ## V Conclusions In conclusion, the present paper has dealt with the understanding of the “jamming” transition in granular media via the introduction of models from standard statistical mechanics to describe these “non-thermal” systems (the IFLG and Tetris, ). They exhibit a logarithmic compaction when subject to gentle shaking in presence of gravity, a compaction extremely close to what is experimentally observed in granular packs . The presence of such slow dynamics is linked to the existence of a “jamming” transition in granular media. In facts, in the high density region, at $`\rho _m`$, the models undergo a structural arrest where grains self-diffusivity becomes zero . Self-diffusion suppression at $`\rho _m`$ signals that, above such a density, it is impossible to obtain a macroscopic rearrangement of grains without increasing the system volume, a fact interestingly similar to the phenomenon of the Reynolds dilatancy transition . A very important fact is that results from the present models are in excellent agreement with known experimental ones, and their many new outcomes must be still experimentally investigated. We have also discussed the intriguing connections of granular media with others materials, as glassy systems or spin glasses, where geometrical disorder and frustration play a crucial role. ## VI Acknowledgements We thank INFM-CINECA for CPU time on Cray-T3D/E. Work partially supported by the TMR Network ERBFMRXCT980183, MURST-PRIN 97 and INFM-PRA 99.
warning/0001/cond-mat0001069.html
ar5iv
text
# Untitled Document 1. Introduction In everyday life we face several complex problems, classified as combinatorial optimization problems, the solutions of which are of great practical importance. Research in this area tries to find different efficient techniques for finding the extremum (maximum or minimum) values of a function of many different independent variables \[1-3\]. The travelling salesman problem (TSP) is a simple example of a combinatorial optmization problem and perhaps the most famous one. Given a certain set of cities and the intercity distance metric, a travelling salesman must find the shortest tour in which he visits all the cities and comes back to his starting point. It is a non-deterministic polynomial complete (NP- complete) problem. NP problems are those for which a potential solution can be checked efficiently for correctness, but finding such a solution appears to take time which scales exponentially with the size $`N`$ in the worst case. The completeness property of NP- complete problems means that if it is possible to find a deterministic algorithm that solves one NP- complete problem in polynomial time, then the other NP- complete problems could also be solved in polynomial time. In the TSP, the most naive algorithm for finding the optimal tour would have to consider all the $`(N1)!/2`$ possible tours for $`N`$ number of cities and check for the shortest of them. Working this way, the fastest computer available today would require more time than the current age of the universe to solve a case with about $`30`$ cities. The typical-case behaviour is difficult to characterize for the the TSP though it is believed to require exponential time to solve in the worst case. For this reason the TSP serves as a prototype problem for the study of the combinatorial optimization problems in general. In the normal TSP, we have $`N`$ number of cities distributed in some continuum space and we determine the average optimal travel distance per city $`\overline{l}_E`$ in the Euclidean metric (with $`\mathrm{\Delta }r_E=\sqrt{\mathrm{\Delta }x^2+\mathrm{\Delta }y^2}`$), or $`\overline{l}_M`$ in the Manhattan metric (with $`\mathrm{\Delta }r_M=|\mathrm{\Delta }x|+|\mathrm{\Delta }y|`$). Since the average distance per city (for fixed area) scales with the number of cities $`N`$ as $`1/\sqrt{N}`$, we find that the normalized travel distance per city $`\mathrm{\Omega }_E=\overline{l}_E\sqrt{N}`$ or $`\mathrm{\Omega }_M=\overline{l}_M\sqrt{N}`$ become the optimized constants and their values depend on the method used to optimize the travel distance. In section 2, we discuss some algorithms used to determine the optimal tour and find the values of the constants $`\mathrm{\Omega }_E`$ and $`\mathrm{\Omega }_M`$ for the optimized travel. In section 3, we present an analytic method to estimate the upper and lower bounds of $`\mathrm{\Omega }_E`$ and $`\mathrm{\Omega }_M`$. In the lattice version of the TSP, the cities are represented by randomly occupied lattice sites of a two- dimensional square lattice; the fractional number of occupied sites being $`p`$ (lattice occupation concentration). In this case the average optimal travel distance in the Euclidean metric $`\overline{l}_E`$, and in the Manhattan metric $`\overline{l}_M`$, vary with the lattice concentration $`p`$. Then the normalised travel distance per city are defined as $`\mathrm{\Omega }_E=\overline{l}_E\sqrt{p}`$ and $`\mathrm{\Omega }_M=\overline{l}_M\sqrt{p}`$. In section 4, we study the variation of $`\mathrm{\Omega }_E`$ and $`\mathrm{\Omega }_M`$, and the ratio $`\mathrm{\Omega }_M/\mathrm{\Omega }_E`$ with $`p`$. Finally, we draw conclusions in section 5. 2. Some heuristic algorithms The most naive method to obtain an approximate solution of travelling salesman problem is the “greedy” heuristic algorithm . Suppose we have a random arrangement of $`N`$ cities in a square (country) of fixed area (taken to be unity). Let us think of any tour to start-with and then make a local exchange of a pair of cities in the tour. We compute the new tour length and if it is lower than the previous one, then the greedy algorithm accepts the new tour as the starting point for further such modifications. The “Lin- Kernighan” algorithm considers local exchange between three or more cities. The essential drawback of such local search algorithms is the obvious one of getting stuck at a local minimum, where any local rearrangement in the tour does not improve the optimized tour length. The “simulated annealing method” is an ingenious method in analogy with the thermodynamic way of avoiding such local minima in free energy (glass formations) and achieving the global minimum of a many-body system by slow cooling or annealing. The rapid quenching of the system leads to the trapping of the system in a local minimum (or glass) state. The system cannot get out of it, since the Boltzmann probability to get out of the minimum drops to zero, as the temperature becomes zero due to quenching. This is similar to the greedy or other local search algorithms. In the annealing, the system is slowly cooled so that as the system falls in a local trap, the finite Boltzmann probability ($`\mathrm{exp}(\mathrm{E}^{}\mathrm{E})/\mathrm{kT}`$, for trap energy $`E`$ and barrier height $`E^{}`$) allows the system to get out of the trap, maintaining a general flow to lower energy states as temperature decreases. Eventually the system anneals to the ground state at the lowest temperature. In the TSP case, one takes the total tour length $`L(=Nl)`$ as the energy $`E`$ and one introduces a fictitious temperature $`T`$. Initially one takes $`T`$ very high such that the average total tour length $`\overline{L}`$ is much higher than the global minimum. The tours are then modified locally and the modified tours are accepted with probability $`\mathrm{exp}(\mathrm{\Delta }\mathrm{L}/\mathrm{kT})`$ where $`\mathrm{\Delta }L`$ is the change in the tour length. In greedy algorithm the probability is unity for negative $`\mathrm{\Delta }L`$ and it is zero for positive $`\mathrm{\Delta }L`$ cases. Here, probability is non-vanishing even for $`\mathrm{\Delta }L`$ positive as long as the temperature is nonzero! Simulated annealing and numerous heuristic generalizations of the local search algorithm optimize very effectively on small scales involving a small number of variables, but fail for the larger scales that require the modification of many variables simultaneously. To deal with the large scales, “genetic algorithms” use a “crossing” procedure which takes two good configurations $``$ “parents”, from a population and finds sub-paths that are common to the parents. It generates a “child” by reconnecting those sub-paths, either randomly or by using large parts of its parents. A population of configurations is evolved from one generation to the next using these crossings followed by a selection of the best children. However, this approach supposedly does not work well in practice since it is extremely difficult to produce two parents and cross them to make a child as good as them. This is a major drawback of the genetic algorithms and is responsible for their limited use. So far, careful analysis of the numerical results obtained indicates that $`\mathrm{\Omega }_E0.72`$ for TSP on continuum. 3. Some analytical results for the bounds for $`\mathrm{\Omega }`$ Although the TSP problem is a multivariable optimization problem (real number of variables $`N!`$ in an $`N`$ city problem), we now look for an approximate analytical solution (upper bound) by expressing the travel distance as a function of a single variable and optimizing the distance with respect to that variable . As is obvious, the problem is trivial in one dimensional case where any directed tour will solve it. In two dimensions, one can again reduce it (approximately) to an one dimensional problem, where the square (country) is divided into strips of width $`W`$ and within each strip, the salesman visits the cities in a directed way. The total travel distance is then optimized with respect to $`W`$. Let the strip width be $`W`$ and the probability density of cities be $`p(=N`$ for unit area). We have a city at $`(0,y_1)`$ \[See Fig. 1\]. The probability that the next city is between distances $`x`$ and $`x+\mathrm{\Delta }x`$, is $`pW\mathrm{\Delta }x`$. The probability that there is no city in the distance $`x=n\mathrm{\Delta }x`$, is $`(1pW\mathrm{\Delta }x)^ne^{(pWx)}`$. The probability that there is a city between $`y`$ and $`y+\mathrm{\Delta }y`$, is $`\mathrm{\Delta }y/W`$. Hence the probability that there is no other city within distance $`y`$ is $`(1y/W)`$. The average distance between any two consecutive cities is therefore $$\overline{l}_E=2_{x=0}^{\mathrm{}}_{y=0}^W\sqrt{x^2+y^2}pW𝑑xe^{(pWx)}\frac{dy}{W}(1\frac{y}{W}).$$ $`(1)`$ The factor $`2`$ arises to take care of the fact that $`y`$ can be both positive and negative. We make the substitutions: $`u=pWx`$ and $`v=y/W`$, so that $$\overline{l}_E=2_{u=0}^{\mathrm{}}_{v=0}^1\frac{1}{pW}\sqrt{u^2+p^2W^4v^2}e^u(1v)𝑑u𝑑v.$$ We introduce two dimensionless quantities $`\mathrm{\Omega }_E=\sqrt{p}\overline{l}_E`$ and $`\stackrel{~}{W}=\sqrt{p}W`$, so that $$\mathrm{\Omega }_E=\frac{2}{\stackrel{~}{W}}_{u=0}^{\mathrm{}}_{v=0}^1\sqrt{u^2+\stackrel{~}{W}^4v^2}e^u(1v)𝑑u𝑑v.$$ $`(2)`$ Using the method of Monte Carlo integration to evaluate the above integral, we get the minimum $`\mathrm{\Omega }_E0.92`$ at normalized strip width $`\stackrel{~}{W}1.73`$ \[See Fig. 2\]. In the Manhattan metric the average distance between any two consecutive cities is $$\overline{l}_M=2_{x=0}^{\mathrm{}}_{y=0}^W(x+y)pW𝑑xe^{(pWx)}\frac{dy}{W}(1\frac{y}{W}).$$ $`(3)`$ As before we introduce $`u=pWx`$ and $`v=y/W`$, so that $$\overline{l}_M=2_{u=0}^{\mathrm{}}_{v=0}^1\frac{1}{pW}(u+pW^2v)e^u(1v)𝑑u𝑑v,$$ and then introduce the dimensionless quantities $`\mathrm{\Omega }_M=\sqrt{p}\overline{l}_M`$ and $`\stackrel{~}{W}=\sqrt{p}W`$, so that $$\mathrm{\Omega }_M=\frac{2}{\stackrel{~}{W}}_{u=0}^{\mathrm{}}_{v=0}^1(u+\stackrel{~}{W}^2v)e^u(1v)𝑑u𝑑v.$$ $`(4)`$ Using the method of Monte Carlo integration, we get the minimum $`\mathrm{\Omega }_M1.15`$ at the normalized strip width $`\stackrel{~}{W}1.73`$ \[See Fig. 3\]. Note that the relation $$\mathrm{\Omega }_M\frac{4}{\pi }\mathrm{\Omega }_E$$ can be explained as follows. Let $$x=l_E\mathrm{sin}\theta \mathrm{and}y=l_E\mathrm{cos}\theta .$$ Then, $$l_M=x+y=l_E(\mathrm{cos}\theta +\mathrm{sin}\theta ).$$ Since $`x=y`$, $$\overline{l}_M=2\overline{l}_E\mathrm{cos}\theta .$$ We have now $$\mathrm{cos}\theta =\frac{2}{\pi }_0^{\pi /2}\mathrm{cos}\theta d\theta =\frac{2}{\pi }[\mathrm{sin}\theta ]_0^{\pi /2}=\frac{2}{\pi }.$$ Hence $$\overline{l}_M=\frac{4}{\pi }\overline{l}_E,\mathrm{or}\mathrm{\Omega }_\mathrm{M}=\frac{4}{\pi }\mathrm{\Omega }_\mathrm{E}.$$ $`(5)`$ Let us now estimate the lower bound of the minimum travel distance per city. Let the distance between any two cities be denoted by $`l`$. Then the probability that there is a city between $`l`$ and $`l+dl(p1)2\pi ldl2p\pi ldl`$. Now, the probability that there is no other city in the distance $`l(1\pi l^2)^{p2}e^{(p2)\pi l^2}e^{p\pi l^2}`$. Therefore, $`P(l)dl=(2p\pi l)e^{p\pi l^2}dl`$. Note that $`P(l)𝑑l=1`$. Hence the average distance is $$\overline{l}_E=_0^{\mathrm{}}lP(l)𝑑l=2p\pi _0^{\mathrm{}}l^2e^{\pi pl^2}𝑑l=\frac{1}{2}\frac{1}{\sqrt{p}}.$$ $`(6)`$ Therefore, the lower bound for $`\mathrm{\Omega }_E=1/2`$. The lower bound for $`\mathrm{\Omega }_M`$ can then easily be estimated to be $`2/\pi `$ in a similar manner. 4. The TSP on randomly diluted lattices The lattice version of the TSP was first studied by Chakrabarti . In the lattice version of the TSP, the cities are represented by randomly occupied lattice sites of a two- dimensional square lattice ($`L\times L`$), the fractional number of sites occupied being $`p`$ (lattice concentration) \[11-13\]. In this case, the average optimal travel distance in the Euclidean metric $`\overline{l}_E`$, and in the Manhattan metric $`\overline{l}_M`$, vary with the lattice concentration $`p`$. We intend to study in this case the variation of the normalised travel distance per city, $`\mathrm{\Omega }_E=\overline{l}_E\sqrt{p}`$ and $`\mathrm{\Omega }_M=\overline{l}_M\sqrt{p}`$, with the lattice occupation (city) concentration $`p`$. We generate the randomly diluted lattice configuration following the standard Monte Carlo procedure for 64($`=N`$) randomly positioned (on the lattice) cities. We vary the lattice size from $`(8\times 8)`$ to $`(48\times 48)`$ so that the lattice concentration $`p`$ varies from $`1.000`$ to $`0.028`$. For each such lattice configuration, the exact optimum tour \[See Fig. 4\] is obtained with the help of the GNU tsp\_ solve . We then calculate $`l_E`$ and $`l_M`$. At each lattice concentration $`p`$, we take different lattice configurations and then obtain the averages, $`\overline{l}_E`$ and $`\overline{l}_M`$. We then determine $`\mathrm{\Omega }_E=\overline{l}_E\sqrt{p}`$ and $`\mathrm{\Omega }_M=\overline{l}_M\sqrt{p}`$ and study the variation of $`\mathrm{\Omega }_E`$ and $`\mathrm{\Omega }_M`$, and of the ratio $`\mathrm{\Omega }_M/\mathrm{\Omega }_E`$ with $`p`$. We find that $`\mathrm{\Omega }_E`$ has monotonic variation from $`1`$ (for $`p=1`$) to a constant $`0.79`$ (for $`p0`$) and $`\mathrm{\Omega }_M`$ has monotonic variation from $`1`$ (for $`p=1`$) to the constant $`1.01`$ (for $`p0`$) respectively. We believe, with bigger $`N`$ the value of $`\mathrm{\Omega }_E`$ eventually reduces to about $`0.72`$ as in continuum TSP. Results for higher values of $`N`$ ($`100`$) indeed suggest the same. The ratio $`\mathrm{\Omega }_M/\mathrm{\Omega }_E`$ changes from $`1`$ to $`1.26(4/\pi )`$, as $`p`$ varies from $`1`$ to $`0`$ \[See Fig. 5\]. We note that the TSP on randomly diluted lattice is certainly a trivial problem when $`p=1`$ (lattice limit) as it reduces to the one-dimensional TSP (the connections in the optimal tour are between the nearest neighbours along the lattice; Hamiltonian walks). However, it is certainly hard at the $`p0`$ (continuum) limit. It is clear that the problem crosses from triviality (for $`p=1`$) to the NP- hard problem (for $`p0`$) at a certain value of $`p`$. It seems the transition occurs at $`p<1`$. This requires further investigation. 5. Conclusions If one places $`N`$ cities randomly on a continuum in an unit area, the best numerical results and their analysis (scaling, etc.) suggest that the best optimized travel distance per city becomes $`l_E0.72/\sqrt{N}`$ for the Euclidean metric and $`l_M0.92/\sqrt{N}`$ for the Manhattan metric. The analytic bounds we discussed in section 3, gives $`\mathrm{\Omega }_E(=l_E\sqrt{N})<0.92`$ and $`\mathrm{\Omega }_M(=l_M\sqrt{N})<1.17`$. When the cities are randomly placed on a lattice with concentration $`p`$, as discussed in section 4, we find (with $`N=p`$ for unit area of the country) that $`\mathrm{\Omega }_E(p)`$ and $`\mathrm{\Omega }_M(p)`$ are monotonically varying with $`p`$. The problem is trivial for $`p=1`$ where $`\mathrm{\Omega }_E(p)=\mathrm{\Omega }_M(p)=1`$ and it certainly reduces to the continuum TSP discussed before for $`p0`$ ($`\mathrm{\Omega }_E0.72`$ and $`\mathrm{\Omega }_M0.92`$; although we observed higher values, viz., $`\mathrm{\Omega }_E0.79`$ and $`\mathrm{\Omega }_M1.01`$, since $`N`$ is not suffiently large). The variations of $`\mathrm{\Omega }`$ with $`p`$ are found to be monotonic without any irregular behaviour at any intermediate point like the percolation point, etc. The crossover from the triviality to the NP- hard problem seems to occur at $`p<1`$. However, this requires further investigation. Acknowledgement : We are grateful to R. B. Stinchcombe, A. Percus and O. C. Martin for very useful comments and suggestions. References e-mail addresses : <sup>(1)</sup>anirban@cmp.saha.ernet.in <sup>(2)</sup>bikas@cmp.saha.ernet.in M. R. Garey and D. S. Johnson, Computers and Intractability: A Guide to the Theory of NP- Completeness (1979). S. Kirkpatrick, C. D. Gelatt, Jr., and M. P. Vecchi, Science, 220, 671 (1983). M. Mezard, G. Parisi and M. A. Virasoro, Spin Glass Theory and Beyond (1987). Y. Usami and M. Kitaoka, Int. J. of Mod. Phys. B, 11, 1519 (1997). S. Lin and B. W. Kernighan, Oper. Res., 21, 498 (1973). W. H. Press, S. A. Teukolsky, W. T. Vetterling, B. P. Flannery, Numerical Recipes in C, Second Edition, 444 (1992). D. E. Goldberg, Genetic Algorithms in Search, Optimization and Learning (1989). A. Percus and O. C. Martin, Phys. Rev. Lett., 76, 1188 (1996). J. Beardwood, J. H. Halton and J. M. Hammersley, Proc. Camb. Phil. Soc., 55, 299, (1959); R. S. Armour and J. A. Wheeler, Am. J. Phys., 51 (5), 405 (1983). B. K. Chakrabarti, J. Phys. A: Math. Gen., 19, 1273 (1986). D. Dhar, M. Barma, B. K. Chakrabarti and A. Tarapder, J. Phys. A: Math. Gen., 20, 5289 (1987). M. Ghosh, S. S. Manna and B. K. Chakrabarti, J. Phys. A: Math. Gen., 21, 1483 (1988). P. Sen and B. K. Chakrabarti, J. Phys. (Paris), 50, 255, 1581 (1989). C. Hurtwitz, GNU tsp\_ solve, available at : http://www.cs.sunysb.edu/\̃\ algorith/implement/tsp/implement.shtml A. Chakraborti and B. K. Chakrabarti (to be published). Figure captions Fig. 1 : Calculating the average distance between two nearest neighbours along a strip of width $`W`$. Fig. 2 : Plot of $`l_E\sqrt{p}`$ against $`W\sqrt{p}`$ from eqn. (2). Fig. 3 : Plot of $`l_M\sqrt{p}`$ against $`W\sqrt{p}`$ from eqn. (4). Fig. 4 : A typical optimized tour for TSP on dilute lattice in the Euclidean metric for $`N=64`$ cities. Fig. 5 : Plot of $`\mathrm{\Omega }_E`$, $`\mathrm{\Omega }_M`$ and $`\mathrm{\Omega }_M/\mathrm{\Omega }_E`$ against $`p`$ for TSP on dilute lattice, obtained using the optimization programs (exact) for $`N=64`$ cities (fixed). The error bars are due to configuration to configuration variations.
warning/0001/math0001048.html
ar5iv
text
# 𝐴_∞-structures on an elliptic curve ## 1. $`A_{\mathrm{}}`$-structures and their homotopies In the following definitions we use the sign convention of which is different from the one in the original definition of . ### 1.1. $`A_{\mathrm{}}`$-algebras A ($``$-graded) $`A_{\mathrm{}}`$-algebra is a $``$-graded vector space $`A`$ equipped with linear maps $`m_k:A^kA`$ for $`k1`$ of degree $`2k`$ satisfying for every $`n1`$ the following $`A_{\mathrm{}}`$-constraint $`\mathrm{Ax}_n`$: $$\underset{k+l=n+1}{}\underset{j=1}{\overset{k}{}}(1)^{l(\stackrel{~}{a}_1+\mathrm{}+\stackrel{~}{a}_{j1})+j(l+1)}m_k(a_1,\mathrm{},a_{j1},m_l(a_j,\mathrm{},a_{j+l1}),a_{j+l},\mathrm{},a_n)=0$$ where $`\stackrel{~}{a}_i=\mathrm{deg}(a_i)mod(2)`$. For example, $`\mathrm{Ax}_1`$ says that $`m_1^2=0`$, $`\mathrm{Ax}_2`$ gives the Leibnitz identity for $`m_1`$ and $`m_2`$, etc. One can consider $`m_n`$ as components of a coderivation $`d`$ of the coalgebra $`T(sA)`$ where $`s`$ denote the suspension. The elements of $`T(sA)`$ are denoted traditionally as follows: $$[a_1|a_2|\mathrm{}|a_k]=(sa_1)(sa_2)\mathrm{}(sa_k).$$ The coderivation $`d`$ has a component $`d_k:(sA)^ksA`$ given by $`sm_k(s^1)^k`$, so that $$d([a_1|\mathrm{}|a_n])=\underset{k+l=n+1}{}\underset{j=1}{\overset{k}{}}(1)^{\stackrel{~}{a}_1+\mathrm{}+\stackrel{~}{a}_{j1}+j1+\mu (a_j,\mathrm{},a_{j+l1})}[a_1|\mathrm{}|a_{j1}|m_l(a_j,\mathrm{},a_{j+l1})|a_{j+l}|\mathrm{}|a_n].$$ where for every collection of elements $`(a_1,\mathrm{},a_k)`$ in $`A`$ we denote $$\mu (a_1,\mathrm{},a_k)=(k1)\stackrel{~}{a}_1+(k2)\stackrel{~}{a}_2+\mathrm{}+\stackrel{~}{a}_{k1}+\frac{k(k1)}{2}.$$ The $`A_{\mathrm{}}`$-constraints are equivalent to the condition $`d^2=0`$. For a pair of $`A_{\mathrm{}}`$-algebras $`(A,m^A)`$ and $`(B,m^B)`$ there is a natural notion of a $`A_{\mathrm{}}`$-morphism from $`A`$ to $`B`$. Namely, such a morphism consists of the data $`(f_n,n1)`$ where $`f_n:A^nB`$ is a linear map of degree $`1n`$ such that $`{\displaystyle \underset{1k_1<k_2<\mathrm{}<k_i=n}{}}(1)^{ϵ_L}m_i^B(f_{k_1}(a_1,\mathrm{},a_{k_1}),f_{k_2k_1}(a_{k_1+1},\mathrm{},a_{k_2}),\mathrm{},f_{nk_{i1}}(a_{k_{i1}+1},\mathrm{},a_n))`$ $`={\displaystyle \underset{k+l=n+1}{}}{\displaystyle \underset{j=1}{\overset{k}{}}}(1)^{ϵ_R}f_k(a_1,\mathrm{},a_{j1},m_l^A(a_j,\mathrm{},a_{j+l1}),a_{j+l},\mathrm{},a_n)`$ where the signs $`ϵ_L`$ and $`ϵ_R`$ are defined as follows: $`ϵ_L=\mu (a_1,\mathrm{},a_{k_1})+\mu (a_{k_1+1},\mathrm{},a_{k_2})+\mathrm{}+\mu (a_{k_{i1}+1},\mathrm{},a_n)+`$ $`\mu (f_{k_1}(a_1,\mathrm{},a_{k_1}),\mathrm{},f_{nk_{i1}}(a_{k_{i1}+1},\mathrm{},a_n)),`$ $$ϵ_R=\stackrel{~}{a}_1+\mathrm{}+\stackrel{~}{a}_{j1}+j1+\mu (a_j,\mathrm{},a_{j+l1})+\mu (a_1,\mathrm{},a_{j1},m_l^A(a_j,\mathrm{},a_{j+l1}),a_{j+l},\mathrm{},a_n).$$ Again one can consider $`(f_n)`$ as components of a coalgebra homomorphism $`F:T(sA)T(sB)`$, so that the above equation is equivalent to $$Fd^A=d^BF,$$ where $`d_A`$ (resp. $`d^B`$) is the coderivation on $`A`$ (resp. $`B`$) defined by $`m^A`$ (resp. $`m^B`$). In particular, there is a natural composition of $`A_{\mathrm{}}`$-morphisms defined as follows: $`(fg)_n(a_1,\mathrm{},a_n)=`$ $`{\displaystyle \underset{1k_1<k_2<\mathrm{}<k_i=n}{}}(1)^ϵf_i(g_{k_1}(a_1,\mathrm{},a_{k_1}),g_{k_2k_1}(a_{k_1+1},\mathrm{},a_{k_2}),\mathrm{},g_{nk_{i1}}(a_{k_{i1}+1},\mathrm{},a_n))`$ where $`ϵ=\mu (a_1,\mathrm{},a_{k_1})+\mu (a_{k_1+1},\mathrm{},a_{k_2})+\mathrm{}+\mu (a_{k_{i1}+1},\mathrm{},a_n)+`$ $`\mu (g_{k_1}(a_1,\mathrm{},a_{k_1}),\mathrm{},g_{nk_{i1}}(a_{k_{i1}+1},\mathrm{},a_n)).`$ In the case when $`B`$ and $`A`$ have the same underlying spaces and $`f_1=\mathrm{id}`$ we will call the data $`f=(f_n,n2)`$ a homotopy between two $`A_{\mathrm{}}`$-structures $`m=m^A`$ and $`m^{}=m^B`$ on the same space. Note that for homotopic $`m`$ and $`m^{}`$ we necessarily have $`m_1=m_1^{}`$. If $`f`$ is a homotopy between $`m`$ and $`m^{}`$, $`g`$ is a homotopy between $`m^{}`$ and $`m^{\prime \prime }`$ then $`gf`$ is a homotopy between $`m`$ and $`m^{\prime \prime }`$. ###### Lemma 1.1. Let $`m=(m_n)`$ be an $`A_{\mathrm{}}`$-structure on $`A`$, $`(f_n:A^nA,n2)`$ be an arbitrary family of maps, $`\mathrm{deg}f_n=1n`$. Then there exists a unique $`A_{\mathrm{}}`$-structure $`m^{}`$ on $`A`$ such that $`f=(f_n)`$ (where $`f_1=\mathrm{id}`$) is a homotopy between $`m`$ and $`m^{}`$. Proof. This follows immediately from the fact that the coalgebra homomorphism $`T(sA)T(sA)`$ defined by $`(f_n)`$ is an isomophism. ∎ We denote the $`A_{\mathrm{}}`$-structure $`m^{}`$ constructed in the above lemma by $`m+\delta f`$ (note that it depends non-linearly on $`f`$). ### 1.2. $`A_{\mathrm{}}`$-categories The definition of an $`A_{\mathrm{}}`$-category is similar to that of an $`A_{\mathrm{}}`$-algebra (see , ). Namely, an $`A_{\mathrm{}}`$-category $`𝒞`$ consists of a class of objects $`\mathrm{Ob}𝒞`$, for every pair of objects $`X`$ and $`X^{}`$ a graded space of morphisms $`\mathrm{Hom}(X,X^{})`$, and a collection of linear maps (compositions) $$m_k:\mathrm{Hom}(X_0,X_1)\mathrm{Hom}(X_1,X_2)\mathrm{}\mathrm{Hom}^{}(X_{k1},X_k)\mathrm{Hom}(X_0,X_k)$$ of degree $`2k`$ for all $`k1`$. The associativity constraint is that these compositions define a structure of $`A_{\mathrm{}}`$-algebra on $`_{ij}\mathrm{Hom}(X_i,X_j)`$ for every collection $`X_0,\mathrm{},X_n\mathrm{Ob}𝒞`$. An $`A_{\mathrm{}}`$-functor (see , ) $`\varphi :𝒞𝒞^{}`$ between $`A_{\mathrm{}}`$-categories consists of a map $`\varphi :\mathrm{Ob}𝒞\mathrm{Ob}𝒞^{}`$ and of a collection of linear maps $$f_k:\mathrm{Hom}_𝒞(X_0,X_1)\mathrm{Hom}_𝒞(X_1,X_2)\mathrm{}\mathrm{Hom}_𝒞(X_{k1},X_k)\mathrm{Hom}_𝒞^{}(\varphi (X_0),\varphi (X_k))$$ of degree $`1k`$ for $`k1`$, which define $`A_{\mathrm{}}`$-morphisms $`_{ij}\mathrm{Hom}_𝒞(X_i,X_j)_{ij}\mathrm{Hom}_𝒞^{}(\varphi (X_i),\varphi (X_j))`$. Now assume that we are given two structures of $`A_{\mathrm{}}`$-category with the same class of objects $`𝒞`$ and with the same morphism spaces. Let $`m=(m_n)`$ and $`m^{}=m_n^{}`$ be the collections of the corresponding composition maps. A homotopy between $`m`$ and $`m^{}`$ is an $`A_{\mathrm{}}`$-functor $`\varphi :(𝒞,m)(𝒞,m^{})`$ such that the corresponding map on objects is identity and such that $`f_1`$ is the identity map on morphisms. The analogue of lemma 1.1 is valid in this situation. In the case when $`m_1=0`$ for an $`A_{\mathrm{}}`$-category $`𝒞`$ the products $`m_2`$ define a structure of the usual category on $`𝒞`$ (without units). If we have two such $`A_{\mathrm{}}`$-categories $`𝒞`$ and $`𝒞^{}`$ and a functor $`\varphi _0:𝒞𝒞^{}`$ between them considered as usual categories then we say that $`\varphi _0`$ is strictly compatible with $`A_{\mathrm{}}`$-structures if it extends to an $`A_{\mathrm{}}`$-functor with $`f_k=0`$ for $`k>1`$. Let $`X`$ be an object of an $`A_{\mathrm{}}`$-category which has $`m_1=0`$ equipped with a decomposition $`X=X_1X_2`$ into a direct sum. By definition (here we deal with the usual category structure without units) this means that we have functorial in $`Y`$ isomorphisms $$\mathrm{Hom}(X,Y)\mathrm{Hom}(X_1,Y)\mathrm{Hom}(X_2,Y)$$ and $$\mathrm{Hom}(Y,X)\mathrm{Hom}(Y,X_1)\mathrm{Hom}(Y,X_2).$$ We say that the decomposition $`X=X_1X_2`$ is strictly compatibile with an $`A_{\mathrm{}}`$-structure if every composition $`m_n`$ involving the spaces $`\mathrm{Hom}(X,Y)`$ or $`\mathrm{Hom}(Y,X)`$ is a direct sum of the corresponding compositions with the spaces $`\mathrm{Hom}(X_i,Y)`$ and $`\mathrm{Hom}(Y,X_i)`$. ### 1.3. Cyclic $`A_{\mathrm{}}`$-structures We will consider a special class of $`A_{\mathrm{}}`$-algebras, namely, those equipped with a cyclic symmetry. ###### Definition 1.2. Let $`A`$ be an $`A_{\mathrm{}}`$-algebra equipped with a bilinear form $`b:AAk`$. We will call $`A`$ cyclic if for every $`n1`$ the following identity is satisfied: (1.1) $$b(m_n(a_1,\mathrm{},a_n),a_{n+1})=(1)^{n(\stackrel{~}{a}_1+1)}b(a_1,m_n(a_2,\mathrm{},a_{n+1})).$$ Remark. Assume in addition that $`b`$ satisfies the following symmetry: $$b(a_1,a_2)=(1)^{\stackrel{~}{a_1}\stackrel{~}{a_2}}b(a_2,a_1).$$ Then (1.1) can be rewritten as follows: $$b(m_n(a_1,\mathrm{},a_n),a_{n+1})=(1)^{n+\stackrel{~}{a_1}(\stackrel{~}{a_2}+\mathrm{}+\stackrel{~}{a_{n+1}})}b(m_n(a_2,\mathrm{},a_{n+1}),a_1).$$ Using $`b`$ we can define a linear functional $`\xi `$ on $`T(sA)`$ by setting $`\xi =b(s^1)^2`$ on $`(sA)^2`$ while $`\xi =0`$ on $`(sA)^n`$ for $`n2`$. Thus, we have $$\xi ([a_1|a_2])=(1)^{\stackrel{~}{a}_1+1}b(a_1,a_2),$$ Then the equation (1.1) is equivalent to the condition $`\xi d=0`$ where $`d`$ is the coderivation defined by $`(m_n)`$. The collection of maps $`f=(f_n:A^nA,n1)`$, $`\mathrm{deg}f_n=1n`$, $`f_1=\mathrm{id}`$ is called a cyclic homotopy if (1.2) $$\underset{k+l=n}{}(1)^{(l+1)(\stackrel{~}{a}_1+\mathrm{}+\stackrel{~}{a}_k)+nk}b(f_k(a_1,\mathrm{},a_k),f_l(a_{k+1},\mathrm{},a_n))=0$$ for $`n3`$. This is equivalent to the condition $`\xi F=\xi `$ where $`F:T(sA)T(sA)`$ is the coalgebra homomorphism defined by $`(f_n)`$. Let $`m`$ be a cyclic $`A_{\mathrm{}}`$-structure, $`f`$ be a cyclic homotopy. Then the $`A_{\mathrm{}}`$-structure $`m+\delta f`$ is cyclic. If $`f`$ and $`g`$ are cyclic homotopies then $`fg`$ is also cyclic. Remark. Assume that $`f_k=0`$ unless $`k=1`$ or $`k=n`$ for some $`n2`$. Then $`f`$ is a cyclic homotopy if and only if $$b(f_n(a_1,\mathrm{},a_n),a_{n+1})=(1)^{(n+1)\stackrel{~}{a}_1+n}b(a_1,f_n(a_2,\mathrm{},a_{n+1}))$$ and $$b(f_n(a_1,\mathrm{},a_n),f_n(a_{n+1},\mathrm{},a_{2n}))=0.$$ The definition of cyclic $`A_{\mathrm{}}`$-categories follows the same pattern. We assume that there is a bilinear form $$b:\mathrm{Hom}(X,Y)\mathrm{Hom}(Y,X)k$$ for every pair of objects $`(X,Y)`$. Then an $`A_{\mathrm{}}`$-category is called cyclic if the identity (1.1) is satisfied whenever $`a_1\mathrm{Hom}(X_1,X_2),\mathrm{},a_n\mathrm{Hom}(X_n,X_{n+1})`$, $`a_{n+1}\mathrm{Hom}(X_{n+1},X_1)`$. Similarly we define cyclic homotopy between two cyclic $`A_{\mathrm{}}`$-structures with the same objects and morphism spaces (and the same bilinear form $`b`$). ### 1.4. Transversal $`A_{\mathrm{}}`$-structures Assume that we are given a class of objects and a notion of transversality for pairs of objects. We will call an $`n`$-tuple of objects $`(X_1,\mathrm{},X_n)`$ transversal if for every $`1i<jn`$ the pair $`(X_i,X_j)`$ is transversal. Then the structure of transversal (cyclic) $`A_{\mathrm{}}`$-category on this class of objects consists of the following data. For every pair of transversal objects $`(X,Y)`$ a graded space of morphisms $`\mathrm{Hom}(X,Y)`$ is given. For every transversal collection $`(X_0,\mathrm{},X_n)`$, $`n1`$ we have linear maps $$m_n:\mathrm{Hom}(X_0,X_1)\mathrm{Hom}(X_1,X_2)\mathrm{}\mathrm{Hom}(X_{n1},X_n)\mathrm{Hom}(X_0,X_n)$$ of degree $`2n`$ such that the axioms $`\mathrm{Ax}_n`$ and the identity (1.1) are satisfied whenever the objects involved in it form a transversal collection. Similarly we define a notion of homotopy between transversal $`A_{\mathrm{}}`$-structures and the cyclic analogues of these notions. The motivating example is that of Fukaya category (see ) where objects are Lagrangain submanifolds in a symplectic manifold with some additional structure. Then we have the standard notion of transversality for pairs of Lagrangians. However, notice that the notion of transversality for $`n`$-tuples we use is weaker than the standard one: we just require every pair of them to intersect transversally but, for example, we allow three Lagrangians to intersect in one point. ### 1.5. Two $`A_{\mathrm{}}`$-structures on $`\mathrm{Vect}(E)`$ The first $`A_{\mathrm{}}`$-structure (or rather a class of equivalent structures) can be defined on the category $`\mathrm{Vect}(M)`$ where $`M`$ is a variety over $`k`$ or a complex manifold as follows. Let us choose some functorial acyclic resolution $`VR^{}(V)`$ for every vector bundle $`V`$ on $`M`$ such that for every pair of bundles there are functorial morphisms $$R^{}(V_1)R^{}(V_2)R^{}(V_1V_2)$$ inducing the identity map on $`V_1V_2`$ and satisfying the natural associativity condition (e.g. one can take Cech complexes of acyclic covering or in the case $`k=`$ Dolbeault complexes). Then we can define a dg-category whose objects are vector bundles with $`\mathrm{Hom}(V_1,V_2)=R^{}(V_1^{}V_2)`$. By homotopic invariance of the notion of $`A_{\mathrm{}}`$-algebra there exists an equivalent $`A_{\mathrm{}}`$-category structure on $`\mathrm{Vect}(M)`$ with morphisms $`\mathrm{Hom}^{}(V_1,V_2)=_i\mathrm{Ext}^i(V_1,V_2)`$ which has $`m_1=0`$ (see ,,,,). We will use a particular representative of this class of $`A_{\mathrm{}}`$-structures in the case when $`M`$ is a compact complex manifold equipped with a hermitian metric. This $`A_{\mathrm{}}`$-structure appears naturally on the category $`\mathrm{Vect}^h(M)`$ of holomorphic vector bundles equipped with a hermitian metric. Starting with the dg-category given by Dolbeault complexes one can use metrics to write an explicit formula for higher compositions involving some Hodge theory operators (see ). We will denote this $`A_{\mathrm{}}`$-structure by $`m^H=(m_n^H)`$. Note that since $`m_2^H`$ is the standard composition (while $`m_1=0`$) the choices of hermitian metrics on bundles are not really important. More precisely, by the standard argument in the homotopy theory the objects $`(V,h)`$ and $`(V,h^{})`$, where $`h`$ and $`h^{}`$ are different metrics on the same bundle $`V`$, are equivalent objects of this $`A_{\mathrm{}}`$-category. By the definition (that appeared in ), this means that there exists an $`A_{\mathrm{}}`$-functor from the category with two isomorphic objects $`O_1O_2`$ and no other non-trivial morphisms to our $`A_{\mathrm{}}`$-category, that sends $`O_1`$ to $`(V,h)`$ and $`O_2`$ to $`(V,h^{})`$. Assume in addition that $`\omega _M`$ is trivialized. Then the Serre duality gives a non-degenerate pairing $$\mathrm{Hom}^{}(V_1,V_2)\mathrm{Hom}^{}(V_2,V_1).$$ The main feature of our particular choice of an $`A_{\mathrm{}}`$-structure is the cyclic symmetry (1.1) of $`m_n^H`$ with respect to the Serre duality (see ). Note also that the higher products $`m_n^H`$ are compatible with Massey products when the latter are well-defined. To define the second $`A_{\mathrm{}}`$-structure on $`\mathrm{Vect}E`$ (or rather, transversal $`A_{\mathrm{}}`$-structure) let us recall the definition of the Fukaya $`A_{\mathrm{}}`$-category of the torus $`T=^2/^2`$ with the (complexified) symplectic form $`2\pi i\tau dxdy`$ where $`\tau `$ is an element of the upper half-plane. We give here a very concrete version of the general definition which can be found in ,,. The objects of this category are pairs $`(\overline{L},A)`$ where $`\overline{L}=p(L)`$ is the image of a non-vertical line $`L`$ with rational slope under the natural projection $`p:^2T`$ (a geodesic circle), $`A:VV`$ is an operator on a finite dimensional complex vector space $`V`$ with real eigenvalues <sup>3</sup><sup>3</sup>3The slight difference with is that we don’t attach to $`L`$ an integer and don’t consider vertical lines.. We call a pair of objects $`(\overline{L_1},A_1)`$ and $`(\overline{L_2},A_2)`$ transversal if $`\overline{L_1}`$ and $`\overline{L_2}`$ are different. For such a pair the morphism space is $$\mathrm{Hom}((\overline{L_1},A_1),(\overline{L_2},A_2))=\mathrm{Hom}(V_1,V_2)\mathrm{Hom}(\overline{L_1},\overline{L_2})$$ where $$\mathrm{Hom}(\overline{L_1},\overline{L_2})=_{P\overline{L_1}\overline{L_2}}[P]$$ ($`[P]`$ is a basis vector attached to a point $`P`$). Note that there is a natural pairing (1.3) $$\mathrm{Hom}((\overline{L_1},A_1),(\overline{L_2},A_2))\mathrm{Hom}((\overline{L_2},A_2),(\overline{L_1},A_1))$$ induced by the natural duality between $`\mathrm{Hom}(V_1,V_2)`$ and $`\mathrm{Hom}(V_2,V_1)`$ and by the self-duality of $`\mathrm{Hom}(\overline{L_1},\overline{L_2})=\mathrm{Hom}(\overline{L_2},\overline{L_1})`$ (such that the basis $`([P])`$ is autodual). Let $`\lambda _i`$ be the slope of the line $`L_i`$ ($`i=1,2`$). Then $`\mathrm{Hom}((\overline{L_1},A_1),(\overline{L_2},A_2))0`$ only if $`\lambda _1\lambda _2`$. This space has grading $`0`$ if $`\lambda _1<\lambda _2`$ and grading $`1`$ if $`\lambda _1>\lambda _2`$. By definition the differential $`m_1`$ is zero. The compositions $`m_k`$ for $`k2`$ are defined as follows. Let $`(\overline{L_i},A_i)`$, $`i=0,\mathrm{},k`$ be objects of the Fukaya categories such that the corresponding circles are pairwise different. Below it will be convenient to identify the set of indices $`[0,k]`$ with $`/(k+1)`$. For every $`i/(k+1)`$ let $`d_i[0,1]`$ be the grading of $`\mathrm{Hom}((\overline{L_i},A_i),(L_{i+1},A_{i+1}))`$. The composition $$m_k^F:\mathrm{Hom}((\overline{L_0},A_0),(\overline{L_1},A_1))\mathrm{}\mathrm{Hom}((\overline{L_{k1}},A_{k1}),(\overline{L_k},A_k))\mathrm{Hom}((\overline{L_0},A_0),(\overline{L_k},A_k))$$ is non-zero only if $`_{i=0}^kd_i=k1`$. Let $`P_{i,i+1}`$ be some intersection points of $`\overline{L_i}`$ and $`\overline{L_{i+1}}`$ ($`i=0,\mathrm{},k1`$). For every $`i=0,\mathrm{},k1`$ let $`M_{i,i+1}`$ be an element of $`\mathrm{Hom}(V_i,V_{i+1})`$. Then $`m_k^F(M_{0,1}[P_{0,1}],M_{1,2}[P_{1,2}],\mathrm{},M_{k1,k}[P_{k1,k}])=`$ $`{\displaystyle \underset{P_{0,k},\mathrm{\Delta }}{}}\pm \mathrm{exp}(2\pi i\tau {\displaystyle _\mathrm{\Delta }}𝑑xdy)\mathrm{exp}(2\pi i(x(p_k)x(p_{k1}))A_k)M_{k1,k}\mathrm{exp}(2\pi i(x(p_{k1})x(p_{k2}))A_{k1})`$ $`\mathrm{}M_{1,2}\mathrm{exp}(2\pi i(x(p_1)x(p_0))A_1)M_{0,1}\mathrm{exp}(2\pi i(x(p_0)x(p_k))A_0)[P_{0,k}]`$ where the sum is taken over points of intersections $`P_{0,k}`$ of $`\overline{L_0}`$ with $`\overline{L_k}`$ and over all $`(k+1)`$-gons $`\mathrm{\Delta }`$ (considered up to translation by $`^2`$) with vertices $`p_iP_{i,i+1}mod^2`$, $`i/(k+1)`$, such that the edge $`[p_{i1},p_i]`$ belongs to $`p^1(\overline{L_i})`$ (the restriction on degrees $`d_i`$ implies that $`\mathrm{\Delta }`$ is convex). We also require that the path formed by the edges $`[p_0,p_1],[p_1,p_2],\mathrm{},[p_k,p_0]`$ goes in the clockwise direction. The sign in the RHS is given by the following rule. If $`k`$ is even then all signs are “plus”. If $`k`$ odd then the sign is equal to the sign of $`x(p_0)x(p_k)`$ (recall that we do not allow vertical lines). It is not difficult to check that $`m^F=(m_k^F)`$ is a (transversal) cyclic $`A_{\mathrm{}}`$-category with respect to the pairing (1.3). This $`A_{\mathrm{}}`$-structure is strictly compatible with decomposition of an operator $`A`$ into a direct sum of operators. The main theorem of identifies the corresponding usual category given by $`m_2^F`$ with a full subcategory of $`\mathrm{Vect}(E)`$ where $`E=/+\tau `$ (which contains all indecomposable bundles). In order to get all vector bundles one has to modify the Fukaya category by adding formally direct sums. We extend the $`A_{\mathrm{}}`$-structure to this larger category using strict compatibility with direct sums. Thus, we get a transversal $`A_{\mathrm{}}`$-structure (which we still denote $`m^F`$) on $`\mathrm{Vect}(E)`$. The construction of identifies the pairing (1.3) with the Serre duality (for some trivialization of $`\omega _E`$), so the obtained $`A_{\mathrm{}}`$-structure is cyclic with respect to it. We will remind some details of the correspondence between vector bundles on $`E`$ and objects of the Fukaya category later. Let us only mention here that the slope of a line corresponding to an indecomposable bundle $`V`$ is equal to the slope of $`V`$ (the ratio of the degree and the rank). Stable bundles correspond to objects $`(\overline{L},A)`$ where $`A`$ is a real number (considered as an operator on a one-dimensional space). ## 2. Transversal $`A_{\mathrm{}}`$-structures on the category of line bundles over an elliptic curve ### 2.1. Transversality and admissibility Let $`E`$ be an elliptic curve over a field $`k`$. Let $``$ be the full subcategory in $`\mathrm{Vect}(E)`$ consisting of line bundles. One can consider extensions of the (strictly associative) composition $`m_2`$ on $``$ to $`A_{\mathrm{}}`$-structures. The following definition gives some natural restrictions one can impose on such an extension. Let us fix a trivialization of $`\omega _E`$. Then the Serre duality gives a non-degenerate pairing $$\mathrm{Hom}^{}(V_1,V_2)\mathrm{Hom}^{}(V_2,V_1)k.$$ ###### Definition 2.1. Let us call a cyclic (with respect to the Serre duality) $`A_{\mathrm{}}`$-structure $`m`$ on the category $``$ admissible if $`m_1=0`$, $`m_2`$ is the standard composition, and the functor of tensor multiplication by a line bundle is strictly compatible with $`m`$. Note that if $`m_1=0`$ then for any $`A_{\mathrm{}}`$-structure $`m^{}`$ which is homotopic to $`m`$ one has $`m_2^{}=m_2`$. So it makes sense to try to classify admissible $`A_{\mathrm{}}`$-structures on $``$ up to cyclic homotopy, strictly compatible with tensor multiplication by any line bundle. We refer to such homotopies as admissible ones. We also define an admissible transversal $`A_{\mathrm{}}`$-structure on $``$ by similar restrictions provided that we have some notion of transversality for pairs of line bundles. We assume that such a notion is given and that it has the following properties: (i) $`(L,M)`$ is transversal if and only if $`(M,L)`$ is transversal; (ii) $`(L,M)`$ is transversal if and only if $`(L^1,M^1)`$ is transversal; (iii) $`(L_1,L_2)`$ is transversal if and only if $`(L_1M,L_2M)`$ is transversal; (iv) for every finite collection of line bundle $`(L_1,\mathrm{},L_n)`$ and every integer $`d`$ there exists an infinite number of isomorphism classes of line bundles $`L`$ of degree $`d`$ such that $`L`$ and $`L^2`$ are transversal to all $`L_i`$; (v) if $`(L,M)`$ is transversal then $`L\simeq ̸M`$. For example, assume that $`E(k)`$ is infinite (this is necessary for the property (iv) to hold). Then one can call $`(L,M)`$ transversal if $`L\simeq ̸M`$. Another example arises from the correspondence between line bundles and objects of Fukaya category defined in . In this example the complex parameter describing an isomorphism class of a line bundle splits into two real parameters: one describes the position of the corresponding geodesic circle and another specifies the connection on it. Then the pair $`(L,M)`$ is transversal if the first real parameter takes different values at $`L`$ and $`M`$ (see section 3.2 for details). The data of an admissible transversal $`A_{\mathrm{}}`$-structure are encoded in the sequence of maps $$m_n:H^{i_1}(L_1)H^{i_2}(L_2)\mathrm{}H^{i_n}(L_n)H^{i_1+i_2+\mathrm{}+i_n+2n}(L_1L_2\mathrm{}L_n)$$ for line bundles $`(L_i)`$ such that the collection $`(𝒪,L_1,L_1L_2,\mathrm{},L_1\mathrm{}L_n)`$ is transversal. ### 2.2. Construction of the homotopy ###### Theorem 2.2. Let $`m`$ and $`m^{}`$ be admissible transversal $`A_{\mathrm{}}`$-structures on the category of line bundles on $`E`$. Assume that for every triple of line bundles $`(L_1,M,L_2)`$ where $`\mathrm{deg}(L_1)=\mathrm{deg}(L_2)=1`$, $`\mathrm{deg}(M)=1`$, such that $`(𝒪,L_1,L_1M,L_1ML_2)`$ is transversal, the maps (2.1) $$H^0(L_1)H^1(M)H^0(L_2)H^0(L_1L_2M)$$ given by $`m_3`$ and $`m_3^{}`$ coincide. Then there exist a unique admissible homotopy between $`m`$ and $`m^{}`$. The following lemma is the main ingredient of the proof. ###### Lemma 2.3. Let $`L`$ be a line bundle of degree $`3`$ on $`E`$, $`S\mathrm{Pic}(E)`$ be a subset such that for every $`d`$ and every isomorphism classes $`[L_1],[L_2]\mathrm{Pic}(E)`$ there exists an infinite number of $`[M]S`$ such that $`\mathrm{deg}(M)=d`$, $`2[M]S`$, $`[L_1]+[M]S`$ and $`[L_2][M]S`$. Then the following sequence is exact: (2.2) $$_{L_1L_2L_3=L,[L_3]S,[L_2L_3]S}H^0(L_1)H^0(L_2)H^0(L_3)\stackrel{\alpha }{}_{L_1L_2=L,[L_2]S}H^0(L_1)H^0(L_2)\stackrel{\beta }{}H^0(L)0$$ where $`L_i`$ denote line bundles of positive degrees, the map $`\alpha `$ sends $`s_1s_2s_3`$ to $`s_1s_2s_3s_1s_2s_3`$, $`\beta `$ sends $`s_1s_2`$ to $`s_1s_2`$. Proof. Clearly, $`\beta `$ is surjective. Thus, it suffices to prove the following statement. Assume that for every pair of line bundles of positive degree $`(L_1,L_2)`$ such that $`[L_2]S`$ and $`L_1L_2L`$ we are given a linear map $$b_{L_1,L_2}:H^0(L_1)H^0(L_2)k$$ such that for every triple $`(L_1,L_2,L_3)`$ such that $`\mathrm{deg}(L_i)>0`$, $`i=1,2,3`$, $`L_1L_2L_3L`$, $`[L_3]S`$, $`[L_2L_3]S`$, one has $$b_{L_1,L_2L_3}(s_1,s_2s_3)=b_{L_1L_2,L_3}(s_1s_2,s_3)$$ where $`s_iH^0(L_i)`$, $`i=1,2,3`$. Then there exists a functional $`\varphi `$ on $`H^0(L)`$ such that for every $`(L_1,L_2)`$ (with $`[L_2]S`$) (2.3) $$b_{L_1,L_2}(s_1,s_2)=\varphi (s_1s_2).$$ We will consider separately several cases. (i) $`\mathrm{deg}(L)=3`$. Let $`p_1,p_2E`$ be a pair of distinct points such that $`[𝒪(p_i)]S`$, $`i=1,2`$, and $`[𝒪(p_1+p_2)]S`$. Let $`s_{p_i}H^0(𝒪(p_i))`$, $`i=1,2`$, be non-zero sections. Then we have the following exact sequence: $$0H^0(L(p_1p_2))\stackrel{\alpha ^{}}{}H^0(L(p_1))H^0(L(p_2))\stackrel{\beta ^{}}{}H^0(L)0$$ where $`\alpha ^{}(s)=(ss_{p_2},ss_{p_1})`$, $`\beta ^{}(t_1,t_2)=t_1s_{p_1}+t_2s_{p_2}`$. Let us define a functional $`\stackrel{~}{\varphi }`$ on $`H^0(L(p_1))H^0(L(p_2))`$ by the formula $$\stackrel{~}{\varphi }(t_1,t_2)=b_{L(p_1),𝒪(p_1)}(t_1,s_{p_1})+b_{L(p_2),𝒪(p_2)}(t_2,s_{p_2}).$$ Note that $`\stackrel{~}{\varphi }`$ vanishes on the image of $`\alpha ^{}`$. Indeed, we have $$b(ss_{p_2},s_{p_1})=b(s,s_{p_2}s_{p_1})=b(s,s_{p_1}s_{p_2})=b(ss_{p_1},s_{p_2}).$$ Therefore, there exists a functional $`\varphi `$ on $`H^0(L)`$ such that $`\stackrel{~}{\varphi }=\varphi \beta ^{}`$. We are going to show that this functional is the one we are looking for. Let $`L_1`$ and $`L_2`$ be line bundles of degrees $`1`$ and $`2`$ respectively such that $`L_1L_2L`$, $`[L_2]S`$. Assume in addition that $`L_2\simeq ̸𝒪(p_1+p_2)`$. Then we claim that $$b_{L_1,L_2}(s,t)=\varphi (st)$$ for any $`sH^0(L_1)`$, $`tH^0(L_2)`$. Indeed, the space $`H^0(L_2)`$ is a direct sum of subspaces $`H^0(L_2(p_1))s_{p_1}`$ and $`H^0(L_2(p_2))s_{p_2}`$. Thus, it suffices to prove that $`b_{L_1,L_2}(s,t)=\varphi (st)`$ for $`t`$ in any of these subspaces. For example, let $`t=t^{}s_{p_1}`$, where $`t^{}H^0(L_2(p_1))`$. Then we have, $$b(s,t)=b(s,t^{}s_{p_1})=b(st^{},s_{p_1})=\varphi (st^{}s_{p_1})$$ as required. Now we claim that if $`M_1`$ and $`M_2`$ are arbitrary line bundles of degrees $`2`$ and $`1`$ respectively such that $`M_1M_2L`$ and $`[M_2]S`$, then $`b_{M_1,M_2}(s,t)=\varphi (st)`$ for $`sH^0(M_1)`$, $`tH^0(M_2)`$. Indeed, let us choose points $`q_1,q_2E`$ such that $`M_1\simeq ̸𝒪(q_1+q_2)`$, $`M_2(q_i)\simeq ̸𝒪(p_1+p_2)`$ and $`[M_2(q_i)]S`$ for $`i=1,2`$. Then $`H^0(M_1)`$ is a direct sum of $`H^0(M_1(q_1))H^0(𝒪(q_1))`$ and $`H^0(M_1(q_2))H^0(𝒪(q_2))`$, so we can assume that $`s`$ is in one of these subspaces. For example, assume that $`s=s^{}s_{q_1}`$ where $`s_{q_1}H^0(𝒪(q_1))`$, $`s^{}H^0(M_1(q_1))`$. Then we have $$b(s,t)=b(s^{}s_{q_1},t)=b(s^{},s_{q_1}t).$$ Applying the previous part of the proof to $`L_1=M_1(q_1)`$, $`L_2=M_2(q_1)`$ we obtain that $$b(s^{},s_{q_1}t)=\varphi (s^{}s_{q_1}t)=\varphi (st)$$ as required. Finally, a similar argument shows that for arbitrary line bundles $`L_1`$ and $`L_2`$ of degrees $`1`$ and $`2`$ one has $`b_{L_1,L_2}(s,t)=\varphi (st)`$ where $`sH^0(L_1)`$, $`tH^0(L_2)`$. (ii) $`\mathrm{deg}(L)=4`$. Let us choose a pair of line bundles $`L_1`$ and $`L_2`$ both of degree $`2`$ such that $`L_1\simeq ̸L_2`$, $`[L_2]S`$, and $`L_1L_2L`$. Then the product map $$H^0(L_1)H^0(L_2)H^0(L)$$ is an isomorphism, so we can define $`\varphi `$ by setting $$\varphi (s_1s_2)=b_{L_1,L_2}(s_1,s_2)$$ where $`s_1H^0(L_1)`$, $`s_2H^0(L_2)`$. Let $`L_1^{}`$, $`L_2^{}`$ be line bundles of degree $`2`$ such that $`L_1^{}L_2^{}L`$ and $`[L_2^{}]S`$. Let $`p,qE`$ be a pair of points such that $`𝒪(p+q)L_1`$, $`[L_2^{}(q)]S`$. Then we claim that the equality (2.4) $$b_{L_1^{},L_2^{}}(s,t)=\varphi (st)$$ holds whenever $`sH^0(L_1^{}(p))`$, $`tH^0(L_2^{}(q))`$. Indeed, assume $`s=s^{}s_p`$, $`t=t^{}s_q`$ where $`s_pH^0(𝒪(p))`$, $`s_qH^0(𝒪(q))`$, $`s^{}H^0(L_1^{}(p))`$, $`t^{}H^0(L_2^{}(q))`$. Then $$b(s,t)=b(s^{}s_p,t^{}s_q)=b(s^{}s_ps_q,t^{})=b(s_ps_q,s^{}t^{})=\varphi (s_ps_qs^{}t^{})=\varphi (st).$$ Now let $`p_1,p_2E`$ be a pair of distinct points such that $`L_1^{}𝒪(p_1+p_2)`$ and for $`q_1,q_2E`$ defined by $`𝒪(p_1+q_1)𝒪(p_2+q_2)L_1`$ one has $`[L_2^{}(q_1)]S`$, $`[L_2^{}(q_2)]S`$. Note that we have $`L_1^{}(q_1+q_2)L_1^2\simeq ̸L`$ since $`L_2\simeq ̸L_1`$. Hence, $`𝒪(q_1+q_2)\simeq ̸L_2^{}`$ and $`H^0(L_2^{})`$ has a basis $`(t_1,t_2)`$ such that $`t_1`$ vanishes at $`q_1`$ and $`t_2`$ vanishes at $`q_2`$. Therefore, if $`s`$ is a section of $`L_1^{}`$ vanishing at $`p_1`$ and $`p_2`$ then by the previous part of the proof we have $$b(s,t_i)=\varphi (st_i)$$ for $`i=1,2`$. Repeating this argument for another pair of points $`(p_1,p_2)`$ as above we get a similar statement for a section of $`L_1^{}`$ linearly independent from $`s`$. Thus, we conclude that (2.4) holds for all $`sH^0(L_1^{})`$, $`tH^0(L_2^{})`$. Now let $`M_1`$ be a line bundle of degree $`1`$, $`M_2`$ be a line bundle of degree $`3`$ such that $`M_1M_2L`$, $`[M_2]S`$. Let $`s_1H^0(M_1)`$, $`s_2H^0(M_2)`$, $`p`$ be a point in the divisor of $`s_2`$. Then assuming that $`[M_2(p)]S`$ we can write $`s_2=s_ps_2^{}`$ where $`s_pH^0(𝒪(p))`$, $`s_2^{}H^0(M_2(p))`$ and $$b(s_1,s_2)=b(s_1,s_ps_2^{})=b(s_1s_p,s_2^{})=\varphi (s_1s_ps_2^{})=\varphi (s_1s_2).$$ Since $`H^0(M_2)`$ is spanned by $`H^0(M_2(p))`$ and $`H^0(M_2(p^{}))`$ for two distinct points $`p,p^{}E`$, this proves that $`b_{M_1,M_2}(s_1,s_2)=\varphi (s_1s_2)`$ for all $`s_1`$ and $`s_2`$. The case when $`\mathrm{deg}(M_1)=3`$, $`\mathrm{deg}(M_2)=1`$ is completely analogous. (iii) $`\mathrm{deg}(L)=d5`$. Let us fix a line bundle $`L_2`$ of degree $`2`$ such that $`[L_2]S`$ and $`[L_2^2]S`$. Then there is an exact sequence $$0L_2^1H^0(L_2)𝒪L_20,$$ which induces for every line bundle $`M`$ of degree $`3`$ an exact sequence $$0H^0(ML_2^1)H^0(M)H^0(L_2)H^0(ML_2)0.$$ Let $`L_1=LL_1^1`$. Consider the following diagram with exact rows (2.5) $$\begin{array}{ccccccc}0& H^0(𝒪(p))H^0(L_1L_2^1(p))& & H^0(𝒪(p))H^0(L_1(p))H^0(L_2)& \stackrel{\alpha _2}{}& H^0(𝒪(p))H^0(L(p))& 0\\ & \text{}& & \text{}& & \text{}& \\ 0& H^0(L_1L_2^1)& & H^0(L_1)H^0(L_2)& \stackrel{\beta }{}& H^0(L)& 0\end{array}$$ where the direct sums in first row are taken over all $`pE`$ such that $`[L(p)]S`$. Notice that $`\gamma `$ is surjective. Indeed, if $`d6`$ this is clear, while for $`d=5`$ we have to check that for the unique point $`p`$ such that $`𝒪(p)L_1L_2^1`$ one has $`[L(p)]S`$. But this follows from our assumptions of $`L_2`$, since $`L(p)L_2^2`$ for such $`p`$. He have $`b_{L_1,L_2}\alpha _1=_pb_{𝒪(p),L(p)}\alpha _2`$. From this by an easy diagram chasing (using the surjectivity of $`\gamma `$) we obtain that $`b_{L_1,L_2}`$ vanishes on the kernel of $`\beta `$, hence, there exists a functional $`\varphi `$ on $`H^0(L)`$ such that $`b_{L_1,L_2}=\varphi \beta `$. It follows that for any $`pE`$ such that $`[L(p)]S`$ one has $$b_{𝒪(p),L(p)}(s,t)=\varphi (st)$$ for $`sH^0(𝒪(p))`$, $`tH^0(L(p))`$. Indeed, we can assume that $`t=t_1t_2`$ with $`t_1H^0(L_1(p))`$, $`t_2L_2`$, in which case $$b(s,t)=b(s,t_1t_2)=b(st_1,t_2)=\varphi (st).$$ Now we can deduce (2.3) in the general case using the same argument as in the end of case (ii). ∎ Remark. It is easy to see from the proof that our assumptions on the set $`S\mathrm{Pic}(E)`$ can be weakened. Let us denote by $`S_d`$ the subset of elements of $`S`$ of degree $`d`$. Then it suffices to require that: (1) for any $`[L]\mathrm{Pic}(E)`$ and any $`d`$ one has $`|S_d(S+[L])|>4`$, $`|S_d([L]S)|>5`$; (2) there exists $`[L]S_2`$ such that $`2[L]S`$. Proof of theorem 2.2. Let us prove the existence first. Clearly we can replace $`m`$ by $`m+\delta f`$ where $`f=(f_n,n2)`$ is an admissible homotopy. Therefore, we can argue by induction: for every $`n3`$, assuming that $`m_k=m_k^{}`$ for $`k<n`$ we will construct an admissible homotopy $`f^n`$ such that $`f_k^n=0`$ for $`k<n1`$ and $`(m+\delta f^n)_n=m_n^{}`$ (this implies that $`(m+\delta f^n)_k=m_k^{}`$ for all $`kn`$). Using the cyclic symmetry we can reduce various types of non-zero transversal $`n`$-tuple products to the following two types: (i) $$m_n:H^0(L_1)H^1(M_1)\mathrm{}H^1(M_i)H^0(L_2)H^1(M_{i+1})\mathrm{}H^1(M_{n2})H^0(L_1L_2M_1\mathrm{}M_{n2})$$ where $`1in2`$, (ii) $$m_n:H^0(L_1)H^0(L_2)H^1(M_1)\mathrm{}H^1(M_{n2})H^0(L_1L_2M_1\mathrm{}M_{n2}).$$ Let us call $`w=\mathrm{deg}(L_1)+\mathrm{deg}(L_2)`$ the weight of the corresponding $`n`$-tuple product type. Note that we have $`w2`$. The first observation is that any $`n`$-tuple product of type (i) of weight $`>2`$ can be expressed via $`k`$-tuple products with $`k<n`$ and via $`n`$-tuple products of smaller weight. Indeed, if $`\mathrm{deg}(L_1)+\mathrm{deg}(L_2)>2`$ then either $`\mathrm{deg}(L_1)>1`$ or $`\mathrm{deg}(L_2)>1`$. Assume for example that $`\mathrm{deg}(L_1)>1`$. Then $`H^0(L_1)`$ is spanned by various products $`s_ps`$ where $`s_p𝒪(p)`$, $`sL_1(p)`$, a point $`pE`$ is such that the collection $$(𝒪,𝒪(p),L_1,L_1M_1,\mathrm{},L_1M_1\mathrm{}M_i,L_1L_2M_1\mathrm{}M_i,L_1L_2M_1\mathrm{}M_{i+1},\mathrm{},L_1L_2M_1\mathrm{}M_{n2})$$ is transversal. Now for any collection of elements $`e_jH^1(M_j)`$, $`j=1,\mathrm{},n2`$, $`tH^0(L_2)`$ and any $`1j<n2`$ we have $`m_n(s_ps,e_1,\mathrm{},e_i,t,e_{i+1},\mathrm{},e_{n2})=m_n(s_p,se_1,\mathrm{},e_i,t,e_{i+1},\mathrm{},e_{n2})\pm `$ $`m_n(s_p,s,e_1,\mathrm{},e_it,e_{i+1},\mathrm{},e_{n2})\pm m_n(s_p,s,e_1,\mathrm{},e_i,te_{i+1},\mathrm{},e_{n2})\pm `$ $`s_pm_n(s,e_1,\mathrm{},e_i,t,e_{i+1},\mathrm{})+\mathrm{}`$ where the unwritten terms contain only $`m_k`$ with $`k<n`$, while the weights of three $`n`$-tuple products in the RHS are smaller than $`w`$. If $`j=n2`$ then there is an additional term $`m_n(s_p,s,e_1,\mathrm{},e_{n2})t`$ which doesn’t affect our argument. Similarly one considers the case when $`\mathrm{deg}(L_2)>1`$. On the other hand, the only non-zero transversal products of type (i) and weight $`2`$ are those of type (2.1). As we’ll see below this will allow us to restrict our attention to products of type (ii). To construct the homotopy $`f^n`$ we again apply induction. Namely, assuming that $`m_n=m_n^{}`$ for all products (of types (i) and (ii)) of weight $`<w`$ (and $`m_k=m_k^{}`$ for $`k<n`$) we will construct a homotopy $`f^{n,w}`$ such that $`(m+\delta f^{n,w})_n=m_n^{}`$ for all products of weight $`w`$ and such that the only non-zero component $`f^{n,w}`$ (other than $`f_1^{n,w}=\mathrm{id}`$) reduces by cyclic symmetry to the following type: $$f_{n1}^{n,w}:H^0(L)H^1(M_1)\mathrm{}H^1(M_{n2})H^0(LM_1\mathrm{}M_{n2})$$ where $`\mathrm{deg}(L)=w`$. Note that $`f^{n,w}`$ is automatically cyclic. Indeed, any non-zero value of $`f^{n,w}`$ is an element of $`H^i(M)`$ where the degree of $`M`$ is either $`w`$ or $`d`$, such that $`0<d<w`$. On the other hand, by definition of Serre duality $`b(H^i(M),H^j(M^{}))=0`$ unless $`\mathrm{deg}(M)+\mathrm{deg}(M^{})=0`$. It follows that one has $$b(f_{n1}^{n,w}(a_1,\mathrm{},a_{n1}),f_{n1}^{n,w}(a_n,\mathrm{},a_{2n2}))=0,$$ so $`f^{n,w}`$ is cyclic. By the above observation it will be sufficient to check the relation $`(m+\delta f^{n,w})_n=m_n^{}`$ only for products of type (ii) (and weight $`w`$). Assume first that $`w=2`$. Then we necessarily have $`n=3`$. Let us fix line bundles $`L`$ and $`M`$, $`\mathrm{deg}(L)=2`$, $`\mathrm{deg}(M)=1`$, such that the triple $`(𝒪,L,LM)`$ is transversal. We want to construct a map $$f_2^{3,2}:H^0(L)H^1(M)H^0(LM)$$ such that for every pair of line bundles $`L_1`$, $`L_2`$ of degree $`1`$, where $`L_1L_2L`$ and the quadruple $`(𝒪,L_1,L,LM)`$ is transversal, the map $$m_3^{}m_3:H^0(L_1)H^0(L_2)H^1(M)H^0(LM)$$ is a composition of the product map $`H^0(L_1)H^0(L_2)H^0(L)`$ with $`f_2^{3,2}`$. Let us fix line bundles $`M^{}`$ and $`L^{}`$ such that $`\mathrm{deg}(M^{})=2`$, $`\mathrm{deg}(L^{})=1`$, $`M^{}L^{}M`$ and the quadruple $`(𝒪,L,LM^{},LM)`$ is transversal. Let $`eH^1(M)`$ be a non-zero element. Then $`e=e^{}s^{}`$ for some $`e^{}H^1(M^{})`$, $`s^{}H^0(L^{})`$. Now for every line bundles $`L_1`$ and $`L_2`$ such that $`L_1L_2L`$, where the quintuple $`(𝒪,L_1,L,LM^{},LM)`$ is transversal, and every $`s_1H^0(L_1)`$ and $`s_2H^0(L_2)`$ we have $$m_3(s_1,s_2,e)=m_3(s_1,s_2,e^{}s^{})=m_3(s_1,s_2e^{},s^{})m_3(s_1s_2,e^{},s^{})$$ and the similar equality holds for $`m_3^{}`$. Note that we have $$m_3(s_1,s_2e^{},s^{})=m_3^{}(s_1,s_2e^{},s^{})$$ by the assumption of the theorem. Therefore, (2.6) $$(m_3^{}m_3)(s_1,s_2,e)=(m_3^{}m_3)(s_1s_2,e^{},s^{}).$$ Let us define the linear map $$f_{e^{},s^{}}:H^0(L)H^1(M)H^0(LM)$$ by the formula $`f_{e^{},s^{}}(s,e)=(m_3^{}m_3)(s,e^{},s^{})`$. We claim that $`f_{e^{},s^{}}`$ doesn’t depend on a choice of $`(M^{},L^{})`$ and $`e^{},s^{}`$ such that $`e^{}s^{}=e`$. Indeed, $`H^0(L)`$ is generated by sections of the form $`s=s_1s_2`$ where $`s_1`$ and $`s_2`$ are as above and the equality (2.6) shows that for such sections $`f_{e^{},s^{}}(s,e)`$ doesn’t depend on $`(e^{},s^{})`$. Thus, we can set $`f_2^{3,2}=f_{e^{},s^{}}`$. Now the same equality shows that for any line bundles $`L_1`$, $`L_2`$ such that $`\mathrm{deg}(L_i)=1`$, $`L_1L_2L`$ and the quadruple $`(𝒪,L_1,L,LM)`$ is transversal one has $$(m_3^{}m_3)(s_1,s_2,e)=f_2^{3,2}(s_1s_2,e).$$ Now assume that $`w3`$. Let us fix line bundles $`M_1,\mathrm{},M_{n2}`$ and elements $`e_iH^1(M_i)`$ for $`i=1,\mathrm{},n2`$. Let us also fix a line bundle $`L`$ of degree $`w`$, such that the collection $$(𝒪,L,LM_1,\mathrm{},LM_1\mathrm{}M_{n2})$$ is transversal. Then for every pair of line bundles $`L_1`$ and $`L_2`$ of positive degree such that $`L_1L_2L`$ and the collection $`(𝒪,L_2,L_2M_1,\mathrm{},L_2M_1\mathrm{}M_{n2})`$ is transversal, consider the map $$b_{L_1,L_2}:H^0(L_1)H^0(L_2)H^0(L_1L_2M_1\mathrm{}M_{n2}):(s_1,s_2)(m_n^{}m_n)(s_1,s_2,e_1,\mathrm{},e_{n2}).$$ We claim that these maps satisfy the condition $$b(s_1s_2,s_3)=b(s_1,s_2s_3)$$ for any sections $`s_iL_i`$, $`i=1,2,3`$, where $`L_1L_2L_3L`$, $`\mathrm{deg}(L_i)>0`$, the collection $$(𝒪,L_2,L_2L_3,L_2L_3M_1,\mathrm{},L_2L_3M_1\mathrm{}M_{n2})$$ is transversal. Indeed, the constraint $`\mathrm{Ax}_n`$ implies that $$m_n(s_1s_2,s_3,e_1,\mathrm{},e_{n2})m_n(s_1,s_2s_3,e_1,\mathrm{},e_{n2})$$ is a linear combination of terms either involving only $`m_k`$ with $`k<n`$ or involving products $`m_n`$ of weight $`<w`$. The same is true for $`m^{}`$, so our claim follows from the induction assumptions on $`m`$ and $`m^{}`$. Therefore, we can apply Lemma 2.3 to the line bundle $`L`$ and the set of isomorphism classes $$S=\{[M]:(𝒪,M,MM_1,\mathrm{},MM_1\mathrm{}M_{n2})\text{is transversal}\}.$$ We conclude that there exists a linear map $$f_{e_1,\mathrm{},e_{n2}}:H^0(L)H^0(LM_1\mathrm{}M_{n2})$$ satisfying $$m_n^{}(s_1,s_2,e_1,\mathrm{},e_{n2})m_n(s_1,s_2,e_1,\mathrm{},e_{n2})=(1)^nf_{e_1,\mathrm{},e_{n2}}(s_1s_2,e_1,\mathrm{},e_{n2}).$$ One can see from this defining property that the map $$f_{n1}^{n,w}:H^0(L)H^1(M_1)\mathrm{}H^1(M_{n2})H^0(LM_1\mathrm{}M_{n2}):se_1\mathrm{}e_{n2}f_{e_1,\mathrm{},e_{n2}}(s)$$ is linear and gives the required homotopy. The proof of uniqueness is also achieved by induction. It suffices to check that an admissible transversal homotopy $`f=(f_n)`$ from $`m`$ to $`m`$ such that $`f_k=0`$ for $`2k<n`$ has also $`f_n=0`$. By cyclic symmetry it suffices to consider the maps $$f_n:H^0(L)H^1(M_1)\mathrm{}H^1(M_{n1})H^0(LM_1\mathrm{}M_{n1})$$ where $`(𝒪,L,LM_1,\mathrm{},LM_1\mathrm{}M_{n1})`$ is transversal. Now we use the induction in degree of $`L`$. If $`\mathrm{deg}(L)=1`$ then such a map is automatically zero. If $`\mathrm{deg}(L)>1`$ then it suffices to consider elements of $`H^0(L)`$ of the form $`ss_p`$ where $`s_pH^0(𝒪(p))`$, $`sH^0(L(p))`$ (where $`𝒪(p)`$ is transversal to all the relevant bundles). Then we can use the identity for $`f_n`$ and the induction assumption to prove the desired vanishing. ∎ ### 2.3. An identity between triple products Assume that we are given a transversal admissible $`A_{\mathrm{}}`$-structure on the category of line bundles on $`E`$. Let $`(L_1,M,L_2)`$ be a triple of line bundles such that $`\mathrm{deg}(L_1)=\mathrm{deg}(L_2)=n>0`$, $`\mathrm{deg}(M)=n`$, and the collection $`(𝒪,L_1,L_1M,L_1ML_2)`$ is transversal. Then the triple products $$m_3:H^0(L_1)H^1(M)H^0(L_2)H^0(L_1ML_2)$$ are invariant under any homotopy. However, in theorem 2.2 only such triple products with $`n=1`$ appear. The reason is that one can express all triple products as above in terms of those with $`n=1`$. This is done by induction in $`n`$ using the identity below. Assume that $`L_i=L_i^{}L_i^{\prime \prime }`$ for $`i=1,2`$, where $`\mathrm{deg}(L_i^{})=n^{}`$, $`\mathrm{deg}(L_i^{\prime \prime })=n^{\prime \prime }`$ for some positive integers $`n^{}`$, $`n^{\prime \prime }`$ such that $`n=n^{}+n^{\prime \prime }`$. Assume also that the collection $`(𝒪,L_1^{},L_1,L_1M,L_1ML_2^{},L_1ML_2)`$ is transversal. ###### Proposition 2.4. One has the following identity $$m_3(s_1^{}s_1^{\prime \prime },e,s_2^{}s_2^{\prime \prime })=m_3(s_1^{},s_1^{\prime \prime }e,s_2^{})s_2^{\prime \prime }+s_1^{}m_3(s_1^{\prime \prime },es_2^{},s_2^{\prime \prime })$$ where $`s_i^{}H^0(L_i^{})`$, $`s_i^{\prime \prime }H^0(L_i^{\prime \prime })`$, $`eH^1(M)`$. Proof. Applying the $`A_{\mathrm{}}`$-constraint $`\mathrm{Ax}_3`$ we get $$m_3(s_1^{}s_1^{\prime \prime },e,s_2)=m_3(s_1^{},s_1^{\prime \prime }e,s_2^{}s_2^{\prime \prime })+s_1^{}m_3(s_1^{\prime \prime },e,s_2^{}s_2^{\prime \prime }).$$ Applying $`\mathrm{Ax}_3`$ again we obtain the following expressions for the terms in the RHS: $$m_3(s_1^{},s_1^{\prime \prime }e,s_2^{}s_2^{\prime \prime })=m_3(s_1^{},s_1^{\prime \prime }e,s_2^{})s_2^{\prime \prime }+s_1^{}m_3(s_1^{\prime \prime }e,s_2^{},s_2^{\prime \prime }),$$ $$m_3(s_1^{\prime \prime },e,s_2^{}s_2^{\prime \prime })=m_3(s_1^{\prime \prime },es_2^{},s_2^{\prime \prime })m_3(s_1^{\prime \prime }e,s_2^{},s_2^{\prime \prime }).$$ Substituting these expressions in the above equality we get the result. ∎ ## 3. Application to homological mirror symmetry ### 3.1. Adding unipotent bundles By a unipotent bundle we mean a vector bundle which has a filtration by subbundles such that the associated graded bundle is trivial. Let $`𝒰=𝒰(E)`$ be the full subcategory in $`\mathrm{Vect}(E)`$ consisting of bundles of the form $`LU`$, where $`L`$ is a line bundle, $`U`$ is a unipotent bundle. Note that a decomposition of $`LU`$ into a tensor product of a line bundle and a unipotent bundle is unique up to an isomorphism. Assume that we are given a notion of transversality for pairs of line bundles. We can extend it to the category $`𝒰`$ by calling a pair $`(LU,L^{}U^{})`$ transversal if and only if $`(L,L^{})`$ is transversal. Then we define an admissible transversal $`A_{\mathrm{}}`$-structure on $`𝒰`$ as a transversal $`A_{\mathrm{}}`$-structure on $`𝒰`$ which is cyclic with respect to Serre duality and is strictly compatible with tensor multiplication by a line bundle, has $`m_1=0`$ and $`m_2`$ equal to the standard product. One defines a notion of admissible homotopy between admissible $`A_{\mathrm{}}`$-structures on $`𝒰`$ similarly to the case of the category $``$. The proof of the following theorem is very similar to that of theorem 2.2 so we omit it. ###### Theorem 3.1. Let $`m`$ and $`m^{}`$ be admissible transversal $`A_{\mathrm{}}`$-structures on the category $`𝒰`$. Assume that for every triple of line bundles $`(L_1,M,L_2)`$ such that $`\mathrm{deg}(L_1)=\mathrm{deg}(L_2)=1`$, $`\mathrm{deg}(M)=1`$ and such that $`(𝒪,L_1,L_1M,L_1ML_2)`$ is transversal, and for every quadruple of unipotent bundles $`U_0`$, $`U_1`$, $`U_2`$ and $`U_3`$ the maps (3.1) $$\mathrm{Hom}(U_0,L_1U_1)\mathrm{Ext}^1(L_1U_1,L_1MU_2)\mathrm{Hom}(L_1MU_2,L_1ML_2U_3)\mathrm{Hom}(U_0,L_1ML_2U_3)$$ given by $`m_3`$ and $`m_3^{}`$ coincide. Then there exist a unique admissible homotopy between $`m`$ and $`m^{}`$. ### 3.2. Connection with the Fukaya category Let $`\tau `$ be an element in the upper half-plane, $`E=/+\tau `$, be the corresponding elliptic curve. Then as shown in the (usual) category $`𝒰`$ is equivalent to the subcategory in the Fukaya category (with compositions $`m_2^F`$) consisting of objects $`(\overline{L},\lambda \mathrm{Id}+N)`$ where $`\overline{L}`$ has an integer slope, $`\lambda `$, $`N`$ is a nilpotent operator. Let $`L(0)`$ be the line bundle on $`E`$ such that the theta-function $$\theta (z)=\theta (z,\tau )=\underset{n}{}\mathrm{exp}(\pi i\tau n^2+2\pi inz)$$ is the pull-back of a section of $`L`$. So $`L(0)𝒪_E(z_0)`$ where $`z_0=\frac{\tau +1}{2}mod(+\tau )`$. For every $`u`$ let us denote $`L(u)=t_u^{}L(0)`$, where $`t_u:EE`$ is the translation by $`u`$. Then every line bundle of degree $`n`$ is isomorphic to a line bundle of the form $`L(0)^{(n1)}L(u)`$. For a nilpotent operator $`N:VV`$ we denote by $`𝒱_N`$ the unipotent bundle on $`E`$, such that the sections of $`𝒱_N`$ correspond to $`V`$-valued functions on $``$ satisfying the quasi-periodicity equations $`f(z+1)=f(z)`$, $`f(z+\tau )=\mathrm{exp}(2\pi iN)f(z)`$. Then every unipotent bundle is isomorphic to a bundle of the form $`𝒱_N`$. The correspondence between bundles in $`𝒰`$ and objects of the Fukaya category constructed in associates to the bundle $`𝒱=L(0)^{(n1)}L(u)𝒱_N`$ the object $`O=(\overline{L},u_1\mathrm{Id}+N)`$, where $`u=u_1+\tau u_2`$, $`u_i`$, $`\overline{L}=\{(u_2+x,(n1)u_2+nx),x/\}`$. This correspondence extends to a functor from $`𝒰`$ to the Fukaya category (with $`m_2^F`$ as a composition) as follows. Let $`𝒱^{}=L(0)^{(n^{}1)}L(u^{})𝒱_N^{}`$ be another bundle in $`𝒰`$, where $`n^{}`$, $`u^{}=u_1^{}+\tau u_2^{}`$, $`N^{}:V^{}V^{}`$ is a nilpotent operator. Let $`O^{}=(\overline{L^{}},u_1^{}\mathrm{Id}+N^{})`$ be the corresponding object in the Fukaya category. Note that $`O`$ and $`O^{}`$ are transversal if and only if either $`n^{}n`$, or $`n^{}=n`$ and $`u_2^{}u_2`$. In the latter case $`\mathrm{Hom}(𝒱,𝒱^{})=\mathrm{Hom}(O,O^{})=0`$ so we can assume that $`nn^{}`$. Assume first that $`n<n^{}`$. Then $`\mathrm{Hom}(O,O^{})=\mathrm{Hom}(V,V^{})\mathrm{Hom}(\overline{L},\overline{L^{}})`$ has degree zero. We can enumerate the points of intersection $`\overline{L}\overline{L^{}}`$ by residues $`k/(n^{}n)`$. Namely, this intersection consists of the points $$P_k=(\frac{k+u_2^{}u_2}{n^{}n},\frac{nk+nu_2^{}n^{}u_2}{n^{}n})$$ where $`k/(n^{}n)`$. On the other hand, we have $$\mathrm{Hom}(𝒱,𝒱^{})=H^0(E,L(0)^{n^{}n1}L(u^{}u)𝒱_{N^{}N^{}})$$ where we consider $`N^{}`$ and $`N^{}`$ as operators on $`V^{}V^{}`$ (acting trivially on one component). Note that if $`M`$ is a line bundle on $`E`$ of the form $`L(0)^{(m1)}L(u)`$ where $`m0`$ and $`N:VV`$ is a nilpotent operator then there is a natural isomorphism between Dolbeault complexes of bundles $`MV`$ and $`M_𝒪𝒱_N`$. Indeed, using the trivialization of the pull-backs of $`M`$ and $`𝒱_N`$ to $``$ we can define the map from the Dolbeault complex of $`MV`$ to that of $`M_𝒪𝒱_N`$ by sending $`\eta (z)`$ to $`\eta (zN/m)`$ where $$(fv)(z\frac{N}{m})=\mathrm{exp}(_z\frac{N}{m})(f)v$$ In particular, we can identify $`\mathrm{Hom}(𝒱,𝒱^{})`$ with the space $`\mathrm{Hom}(V,V^{})H^0(E,L(0)^{n^{}n1}L(u^{}u))`$. The space of global sections of the line bundle $`L(0)^{n^{}n1}L(u^{}u)`$ has a natural basis of theta functions $$\theta _k(z)=\underset{m(n^{}n)+k}{}\mathrm{exp}(\frac{1}{n^{}n}(\pi i\tau m^2+2\pi im((n^{}n)z+u^{}u)))$$ where $`k/(n^{}n)`$. Now we can identify $`\mathrm{Hom}(𝒱,𝒱^{})`$ with $`\mathrm{Hom}(O,O^{})`$ by sending $`T[P_k]`$ (where $`T\mathrm{Hom}(V,V^{})`$) to $$\mathrm{exp}(\frac{1}{n^{}n}(\pi i\tau (u_2^{}u_2)^2\mathrm{Id}+2\pi i(u_2^{}u_2)(N^{}N^{}(u_1^{}u_1)\mathrm{Id})))T\theta _k.$$ To construct similar identification in the case $`n>n^{}`$ we use Serre duality and its natural analogue on the Fukaya category to reduce to the case considered above. As shown in this identification is compatible with compositions $`m_2`$. Using it we can consider $`m^F`$ as a transversal $`A_{\mathrm{}}`$-structure on $`𝒰`$. Furthermore, it is easy to see that $`m^F`$ is admissible. The main point is that the functor of tensoring with a line bundle on $`𝒰`$ corresponds to an automorphism of the Fukaya category given by some symplectic automorphism of the torus. As we will see in section 3.4 the assumptions of the theorem 3.1 are satisfied for the transversal $`A_{\mathrm{}}`$-structures $`m^F`$ and $`m^H`$ on $`𝒰`$. Hence, they are homotopic. The equivalence of $`𝒰`$ with a subcategory of the Fukaya category (with $`m_2^F`$ as a composition) is extended to all bundles in using the construction of vector bundles on $`E`$ as push-forwards of objects in $`𝒰`$ under isogenies. Below we consider the corresponding extension of equivalence between $`A_{\mathrm{}}`$-structures. ### 3.3. Equivalence Let $`\tau `$ be an element in the upper half-plane, $`E=/+\tau `$, be the corresponding elliptic curve. We are going to prove that the two transversal $`A_{\mathrm{}}`$-structures $`m^F`$ and $`m^H`$ on $`\mathrm{Vect}(E)`$ considered in section 1.5 are equivalent. More precisely, the definition of $`m^H`$ requires us to work with bundles equipped with hermitian metrics. Since the different choices of metrics on a vector bundle really give equivalent objects of the $`A_{\mathrm{}}`$-category $`(\mathrm{Vect}^h(E),m^H)`$ we can restrict to some preferred class of hermitian metrics (which we’ll define below). Note that both $`A_{\mathrm{}}`$-structures are cyclic and are strictly compatible with tensor multiplication by a hermitian line bundle and with decompositions of bundles into orthogonal direct sums. For every positive integer $`r`$ we consider the elliptic curve $`E_r=/+r\tau `$. Then we have a natural isogeny $`\pi ^r:E_rE`$ of degree $`r`$ and for every $`r|s`$ an isogeny $`\pi _r^s:E_sE_r`$ such that $`\pi ^s=\pi ^r\pi _r^s`$. We can consider two transversal $`A_{\mathrm{}}`$-structures $`m^F`$ and $`m^H`$ on any of these elliptic curves. An important observation is that both $`m^F`$ and $`m^H`$ are strictly compatible with the functors of pull-back and push-forward with respect to isogenies $`\pi _r^s`$ and $`\pi ^r`$ (these functors extend naturally to vector bundles with metric). For the structure $`m^H`$ this is clear while for $`m^F`$ this follows from the construction of equivalence in . The idea of the proof is to use the decomposition of every bundle on elliptic curve into a direct sum $`V=V_iU_i`$ where $`(V_i)`$ are pairwise non-isomorphic stable bundles, $`(U_i)`$ are unipotent bundles. Then we want to use the fact that every stable bundle of rank $`r`$ on $`E`$ is the push-forward of a line bundle on $`E_r`$. Since our $`A_{\mathrm{}}`$-structures are strictly compatible with isogenies we can derive the desired homotopy from theorem 3.1. More precisely, we need a slight modification of this theorem for the category of bundles with metrics: the assumption should be that the triple products (3.1) given by two $`A_{\mathrm{}}`$-structures coincide for all choices of metrics on the bundles in question. To be able to apply this theorem in our case we will compute explicitly products $`m_3^F`$ and $`m_3^H`$ of the type (3.1) in section 3.4 and will see that they are equal (at this point it will be important to use a particular trivialization of $`\omega _E`$ on which the construction of equivalence in depends). Then the uniqueness of the homotopies constructed in theorem 3.1 will imply that these homotopies are compatible with isogenies, hence, descend to a homotopy on the category $`\mathrm{Vect}(E)`$. Let us call a vector bundle on $`E`$ almost stable if it has form $`VU`$ where $`V`$ is a stable bundle, $`U`$ is a unipotent bundle (thus, every bundle on $`E`$ is a direct sum of almost stable bundles). Let $`r`$ be the rank of $`V`$. Then $`V=\pi _{}^r(L)`$ for some line bundle $`L`$ on $`E_r`$. Hence $`VU=\pi _{}^r(L(\pi ^r)^{}U)`$. We call a hermitian metric on $`V`$ preferred if it comes from a metric on $`L(\pi ^r)^{}U`$. Let $`(V_1,\mathrm{},V_{n+1})`$ be a collection of almost stable bundles on $`E`$ equipped with preferred metrics. From the strict compatibility of our $`A_{\mathrm{}}`$-structures with isogenies we have the following commutative diagram (3.2) $$\begin{array}{ccccccc}\mathrm{Hom}^{}(V_1,V_2)\mathrm{}\mathrm{Hom}^{}(V_n,V_{n+1})& \text{}& \mathrm{Hom}^{}(V_1,V_{n+1})& & & & \\ \text{}& & \text{}& & & & \\ \mathrm{Hom}^{}((\pi ^r)^{}V_1,(\pi ^r)^{}V_2)\mathrm{}\mathrm{Hom}((\pi ^r)^{}V_n,(\pi ^r)^{}V_{n+1})& \text{}& \mathrm{Hom}^{}((\pi ^r)^{}V_1,(\pi ^r)^{}V_{n+1})& & & & \end{array}$$ where $`m=m^F`$ or $`m=m^H`$. Now if $`r`$ is divisible by ranks of all bundles $`V_i`$ then $`(\pi ^r)^{}(V_i)`$ is an orthogonal direct sum of bundles of the form $`LU`$ where $`L`$ is a line bundle, $`U`$ is a unipotent bundle. We will check in section 3.4 that the conditions of theorem 3.1 are satisfied for $`m^F`$ and $`m^H`$. Therefore, we get a unique admissible homotopy $`f^r`$ between these structures on $`𝒰(E^r)`$ for every $`r`$. We can extend this homotopy to orthogonal direct sums of bundles in $`𝒰(E^r)`$ in an obvious way. Note that for every isogeny of elliptic curves $`\pi :E^{}E^{\prime \prime }`$ we have a canonical splitting of the natural embedding $`𝒪_{E^{\prime \prime }}\pi _{}𝒪_E^{}`$, hence for every pair of bundles $`(V_1,V_2)`$ on $`E^{\prime \prime }`$ we get a canonical splitting $$T(\pi ):\mathrm{Hom}(\pi ^{}V_1,\pi ^{}V_2)\mathrm{Hom}(V_1,V_2)$$ of the natural embedding $`\pi ^{}:\mathrm{Hom}(V_1,V_2)\mathrm{Hom}(\pi ^{}V_1,\pi ^{}V_2)`$. Now we claim that the homotopies $`f^r`$ and $`f^s`$ where $`r|s`$ are compatible in the following way: for any bundles $`W_1=L_1U_1,\mathrm{},W_{n+1}=L_{n+1}U_{n+1}`$ in $`𝒰(E^r)`$ one has the commutative diagram (3.3) $$\begin{array}{ccccccc}\mathrm{Hom}^{}(W_1,W_2)\mathrm{}\mathrm{Hom}^{}(W_n,W_{n+1})& \text{}& \mathrm{Hom}^{}(W_1,W_{n+1})& & & & \\ \text{}& & \text{}& & & & \\ \mathrm{Hom}^{}((\pi _r^s)^{}(W_1),(\pi _r^s)^{}(W_2))\mathrm{}\mathrm{Hom}((\pi _r^s)^{}(W_n),(\pi _r^s)^{}(W_{n+1}))& \text{}& \mathrm{Hom}^{}((\pi _r^s)^{}(W_1),(\pi _r^s)^{}(W_{n+1}))& & & & \end{array}$$ Indeed, the compatibility of $`m^F`$ and $`m^H`$ with the isogeny $`\pi _r^s`$ implies that $`T(\pi _r^s)f^s(\pi _r^s)^{}`$ is an admissible homotopy between $`m^F`$ and $`m^H`$ on $`𝒰(E^r)`$, hence it coincides with $`f^r`$. Now we define the homotopy $`f`$ between $`m^F`$ and $`m^H`$ on the category of almost stable vector bundles on $`E`$ with preferred metrics using commutativity of diagrams of the type (3.2). Namely, choosing $`r`$ which is divisible by all ranks of bundles $`V_i`$ we define the map $$f_n:\mathrm{Hom}(V_1,V_2)\mathrm{}\mathrm{Hom}(V_n,V_{n+1})\mathrm{Hom}(V_1,V_{n+1})$$ by the formula $`f_n=T(\pi ^r)f_n^r(\pi ^r)^{}`$. The compatibility (3.3) ensures that this definition doesn’t depend on a choice $`r`$. Now to check that $`f`$ is indeed a homotopy from $`m^F`$ to $`m^H`$ we choose $`r`$ divisible by ranks of all the bundles involved and use the commutativity of (3.2). Since every bundle $`V`$ on $`E`$ is a direct sum of almost stable bundles we have a class of preferred metrics on $`V`$ coming from preferred metrics on almost stable bundles (so that the direct sum becomes orthogonal). We can extend the homotopy $`f`$ to all bundles with preferred metrics in a natural way. ### 3.4. Massey products It remains to compute explicitly the products $`m_3^F`$ and $`m_3^H`$ of the type (3.1). Let us trivialize $`\omega _E`$ in such a way that the Serre duality induces the pairing $$b:\mathrm{Hom}(V_1,V_2)\mathrm{Ext}^1(V_2,V_1)$$ given by the formula $$b(f,gd\overline{z})=_E𝑑z\mathrm{Tr}(fgd\overline{z})$$ where $`f\mathrm{Hom}(V_1,V_2)`$, $`gd\overline{z}\mathrm{\Omega }^{0,1}(\mathrm{Hom}(V_2,V_1))`$. First let us compute $`m_3^H`$. We start with the case when all $`U_i`$ are trivial of rank $`1`$. Then we have to compute the product $$m_3^H:H^0(L_1)H^1(M)H^0(L_2)H^0(L_1ML_2)$$ where $`L_1M\simeq ̸𝒪`$, $`L_2M\simeq ̸𝒪`$. Using a translation on $`E`$ we can assume without loss of generality that $`M=L(0)^1`$. Let $`L_1=L(t)`$, $`L_2=L(u)`$ where $`t,u`$. Let $`z_1`$ and $`z_2`$ be the real components of the complex variable $`z`$ defined by the equality $`z=z_1+\tau z_2`$. The transversality condition means that $`t_2,u_2`$. We will compute the above product under the weaker assumption $`t,u+\tau `$. It is easy to check that the $`(0,1)`$-form with values in $`L(0)^1`$ $$\alpha (z)=\frac{i}{\sqrt{2\mathrm{Im}(\tau )}}\overline{\theta (z)}\mathrm{exp}(2\pi \mathrm{Im}(\tau )(z_2^2))d\overline{z}$$ is a representative of the class in $`H^1(L(0)^1)`$ dual to the class in $`H^0(L(0))`$ given by $`\theta (z)`$. Now for every $`u`$, such that $`u+\tau `$ there exists a unique section $`h(z,u)`$ of $`L(0)^1L(u)`$ such that $$\theta (z+u)\alpha (z)=\overline{}h(z,u)$$ where $`\overline{}=\overline{}_z`$. Indeed, this follows from the fact that all the cohomologies of $`L(0)^1L(u)`$ vanish. One can write an explicit formula for $`h(z,u)`$(see ): $$h(z,u)=\frac{1}{2\pi i}\underset{m,n}{}(1)^{mn}\frac{\mathrm{exp}(\frac{\pi }{2\mathrm{Im}(\tau )}(|\gamma |^2+2\overline{\gamma }u+u^2)+2\pi i(mz_1+(nu)z_2))}{\gamma +u}$$ where $`\gamma =m\tau n`$. Now we have $$m_3^H(\theta (z+t),\alpha ,\theta (z+u))=h(z,t)\theta (z+u)h(z,u)\theta (z+t).$$ As a function of $`z`$ up to a constant factor this should be equal to $`\theta (z+u+v)`$, so we have (3.4) $$h(z,t)\theta (z+u)h(z,u)\theta (z+t)=H(t,u)\theta (z+t+u)$$ for some meromorphic function $`H`$. We have $`H(t,u)=H(u,t)`$. Also it is easy to see that the function $`H(t,u)`$ satisfies the following quasi-periodicity equations: $$H(t+1,u)=H(t,u),$$ $$H(t+\tau ,u)=\mathrm{exp}(2\pi iu)H(t,u).$$ The only poles of $`H(t,u)`$ are poles of order $`1`$ along the divisors $`t=\gamma `$ and $`u=\gamma `$ where $`\gamma +\tau `$. It follows that $`H(t,u)`$ is equal up to a constant to the function $$F(t,u)=\frac{\theta ^{}(\frac{\tau +1}{2})\theta (tu+\frac{\tau +1}{2})}{2\pi i\theta (t+\frac{\tau +1}{2})\theta (u+\frac{\tau +1}{2})}.$$ Furthermore, comparing the residues at $`t=0`$ we conclude that $`H(t,u)=F(t,u)`$. Now let us compute the product $$m_3^H:H^0(L_1U_0^{}U_1)H^1(MU_1^{}U_2)H^0(L_2U_2^{}U_3)H^0(L_1ML_2U_0^{}U_3)$$ where $`U_i`$ are unipotent bundles. As before we can take $`M=L(0)^1`$, $`L_1=L(t)`$, $`L_2=L(u)`$. Let $`U_i=𝒱_{N_i}`$ where $`N_i:V_iV_i`$ are nilpotent operators. Then $`U_i^{}U_{i+1}𝒱_{N_{i+1}N_i^{}}`$ where $`N_{i+1}N_i^{}`$ is an operator on $`V_i^{}V_{i+1}`$. As in section 3.2 we use the isomorphisms between the Dolbeault complexes of bundles $`LV`$ and $`L𝒱_N`$, where $`L`$ is one of line bundles of degree $`1`$ above, $`N:VV`$ is the corresponding nilpotent operator, sending $`\eta (z)`$ to $`\eta (zN)`$. Similarly, we have an isomorphism between the Dolbeault complexes of $`L(0)^1V_1^{}V_2`$ and $`L(0)^1𝒱_{N_2N_1^{}}`$ given by $`\eta (z)\eta (z+N_2N_1^{})`$. Let $`v_{i,i+1}V_i^{}V_{i+1}`$ be some elements. Then we have $`(\alpha (z+N_2N_1^{})v_{1,2})(\theta (z+tN_1+N_0^{}))v_{0,1})=`$ $`\mathrm{Tr}_{V_1}(\overline{}h(z+N_2N_1^{},tN_2+N_1^{}N_1+N_0^{}))v_{0,1}v_{1,2})=`$ $`\mathrm{Tr}_{V_1}(\overline{}h(z+N_2N_1^{},tN_2+N_0^{})v_{0,1}v_{1,2}),`$ since we can replace $`N_1^{}`$ by $`N_1`$ under the sign of $`\mathrm{Tr}_{V_1}`$. Similarly, we get $$(\theta (z+uN_3+N_2^{}))v_{2,3})(\alpha (z+N_2N_1^{})v_{1,2})=\mathrm{Tr}_{V_2}(\overline{}h(z+N_2N_1^{},uN_3+N_1^{})v_{1,2}v_{2,3}).$$ Hence, $`m_3^H(\theta (z+tN_1+N_0^{}))v_{0,1},\alpha (z+N_2N_1^{})v_{1,2},\theta (z+uN_3+N_2^{})v_{2,3})=`$ $`\mathrm{Tr}_{V_1V_2}((\theta (z+uN_3+N_2^{})h(z+N_2N_1^{},tN_2+N_0^{})`$ $`h(z+N_2N_1^{},uN_3+N_1^{})\theta (z+tN_1+N_0^{}))v_{0,1}v_{1,2}v_{2,3}).`$ Making a substitution $`zz+N_2N_1^{}`$, $`ttN_2+N_0^{}`$, $`uuN_3+N_1^{}`$ in the identity (3.4) and using the equality $`H=F`$ we can rewrite the above formula as follows: (3.5) $$\begin{array}{c}m_3^H(\theta (z+tN_1+N_0^{})v_{0,1},\alpha (z+N_2N_1^{})v_{1,2},\theta (z+uN_3+N_2^{})v_{2,3})=\\ \mathrm{Tr}_{V_1V_2}(F(tN_2+N_0^{}),uN_3+N_1^{}))\theta (z+t+uN_3+N_0^{})v_{0,1}v_{1,2}v_{2,3}).\end{array}$$ Now let us compute the corresponding product $`m_3^F`$. The objects of the Fukaya category corresponding to our four bundles $`U_0=𝒱_{N_0}`$, $`L_1U_1=L(t)𝒱_{N_1}`$, $`L_1MU_2=L(0)^1L(t)𝒱_{N_2}`$ and $`L_1ML_2U_3=L(t+u)𝒱_{N_3}`$ are $`((x,0),N_0)`$, $`((x+t_2,x),t_1+N_1)`$, $`((x,t_2),t_1+N_2)`$ and $`((x+t_2+u_2,x),t_1u_1+N_3)`$, where $`t=t_1+\tau t_2`$, $`u=u_1+\tau u_2`$, $`t_2,u_2`$. Note that any two of these circles either don’t intersect or intersect at a unique point. So we can identify morphisms between these objects with spaces $`\mathrm{Hom}(V_0,V_1)`$, $`\mathrm{Hom}(V_1,V_2)`$, etc. Now we have $`m_3^F(v_{0,1}[P_{0,1}],v_{1,2}[P_{1,2}],v_{2,3}[P_{2,3}])=\mathrm{Tr}_{V_1V_2}{\displaystyle \underset{(m,n)^2,(mt_2)(n+u_2)>0}{}}\mathrm{sign}(mt_2)`$ $`\mathrm{exp}(2\pi i\tau (mt_2)(n+u_2)+2\pi i(mt_2)(t_1+N_2N_0^{})+2\pi i(n+u_2)(u_1N_3+N_1^{}))v_{0,1}v_{1,2}v_{2,3})[P_{0,3}]`$ $`=\mathrm{Tr}_{V_1V_2}({\displaystyle \mathrm{sign}(mt_2)\mathrm{exp}(2\pi i\tau mn+2\pi im(uN_3+N_1^{})+2\pi in(t+N_2N_0^{}))Cv_{0,1}v_{1,2}v_{2,3}})`$ where $`C=\mathrm{exp}(2\pi i\tau t_2u_22\pi it_2(u_1N_3+N_1^{})+2\pi iu_2(t_1+N_2N_0^{}))`$. At this point we need the following identity (which essentially coincides with the formula (2.3.4) of ): $$\underset{(m,n)^2,(mt_2)(n+u_2)>0}{}\mathrm{sign}(mt_2)\mathrm{exp}(2\pi i\tau mn+2\pi i(munt))=F(t,u)$$ for arbitrary $`t=t_1+\tau t_2`$, $`u=u_1+\tau u_2`$ such that $`t_2,u_2`$. This identity which is due to Kronecker can be proven as follows: first, one has to check that the left hand side extends to a meromorphic function of $`u`$ and $`t`$ with poles at the lattice points, then one has to compare its quasi-periodicity properties and residues at poles with those of $`F`$. Hence, we get (3.6) $$m_3^F(v_{0,1}[P_{0,1}],v_{1,2}[P_{1,2}],v_{2,3}[P_{2,3}])=\mathrm{Tr}_{V_1V_2}(F(tN_2+N_0^{},uN_3+N_1^{})Cv_{0,1}v_{1,2}v_{2,3})[P_{0,3}].$$ Now it easy to see that the exponential factors involved in the identification of morphisms in $`𝒰`$ with morphisms in the Fukaya category (see section 3.2) kill the factor $`C`$ and we get $`m_3^H=m_3^F`$ on the products of the type (3.1).
warning/0001/astro-ph0001067.html
ar5iv
text
# An asymmetric relativistic model for classical double radio sources ## 1 Introduction The classical double FRII radio sources are among the most luminous extragalactic radio sources (Fanaroff & Riley 1974). They are characterised by two steep spectrum radio lobes, symmetrically disposed with respect to the host galaxy or quasar, and, according to the standard picture, these are powered by beams or jets originating in an active galactic nucleus. The identification of flat-spectrum central radio cores in most of these sources has enabled a number of analyses of the kinematics of their lobes and hot-spots to be undertaken (Longair & Riley 1979, Zieba & Chyzy 1991, Best et al. 1995). Structural asymmetries and misalignments of the hot-spots and radio lobes have been used to investigate their intrinsic properties and to test unification schemes for radio quasars and radio galaxies (Scheuer 1987; Barthel 1987, 1989). In the simplest picture, it was conjectured that the radio lobes and hot-spots moved out symmetrically from an active nucleus at a significantly relativistic speed and the observed structural asymmetries could then be attributed to the differences in light travel times from the lobes to the observer (Ryle & Longair 1967). Many studies have been made of the probability distribution of the velocities of the radio source components from the observed distributions of the ratio of core–hot-spot distances (Longair & Riley 1979; Katgert-Meikelijn et al. 1980; Banhatti 1980; Best et al. 1995). The mean velocity of advance of the hot-spots was found to be $`0.2c`$, with a considerable spread about the mean velocity, some values greater than $`0.4c`$ being found. There are however significant problems with this model. McCarthy et al. (1991) studied the spatial distribution of thermal emission-line gas about powerful FRII radio sources and found that the asymmetry in the distribution of the ionised gas was correlated with the structural asymmetry of the radio lobes, strongly suggesting that environmental asymmetries play a significant rôle. This correlation between optical and radio asymmetries can be naturally attributed to clumpy environmental effects (Pedelty et al. 1989a, b). A further possibility is that the jets of powerful FRII sources might well be intrinsically asymmetric. The analysis of Wardle & Aaron (1997) of the jet-counterjet flux ratios of 13 3CR quasars observed with high resolution by the VLA by Bridle et al. (1994) showed, however, that, where jets and counterjets are observed in quasars, the jet-counterjet brightness ratios can be attributed almost entirely to relativistic beaming and, at most, modest intrinsic asymmetries in the kiloparsec-scale jets are allowed. Further problems arise from observations of one-sided radio jets, which have now been observed in many FRII radio sources. Following Wardle & Aaron (1997), it is natural to attribute this one-sidedness to the effects of relativistic beaming. Consequently, the lobe on the same side as the jet should be approaching the observer and should be longer than the lobe on the counterjet side. Significant discrepances from this rule have been found for a number of FRII radio galaxies and quasars (Saikia 1981, 1984; Black, et al. 1992; Fernini, et al. 1993, 1997; Bridle et al. 1994; Scheuer 1995; Leahy, et al. 1997; Hardcastle et al. 1997, 1998). A further discrepancy is that the expansion speeds of the FRII radio sources found from the simple model are systematically greater, by at least a factor of two, than those found from spectral ageing arguments. This problem is discussed in Section 2.2. Best et al. (1995) realised that it is likely that both relativistic, environmental and intrinsic asymmetries play a rôle in determining the observed structures of FRII radio sources. To test the likely contributions of relativistic and environmental/intrinsic effects upon the observed structures of FRII sources, an *asymmetric relativistic model* has been developed in which account is taken of the contribution of both relativistic and intrinsic/environmental asymmetry effects. By the term ‘relativistic’ we mean that light-travel time differences play a significant rôle in determining the observed structural asymmetries. The key parameter in the present analysis is the jet-side of the radio sources, which is estimated for $`80\%`$ of those in the 3CRR complete sample (Section 4). In Section 2, the problems of the symmetric relativistic model are reviewed and, in Section 3, the asymmetric relativistic model is analysed quantitatively. The sample of FRII radio sources from the 3CRR catalogue is described in Section 4. In Section 5, correlations of source asymmetry with radio luminosity and linear size are discussed. A statistical analysis of the properties of FRII sources in the context of the asymmetric relativistic model is discussed in Section 6. The mean expansion speed of the lobes and the critical angle for orientation-based unification schemes for radio galaxies and quasars are the subjects of Section 7. Except where otherwise stated, Hubble’s constant has been taken to be $`H_0=50`$ km $`\mathrm{s}^1`$ $`\mathrm{Mpc}^1`$ and the deceleration parameter $`q_0=0.5`$. ## 2 Symmetric model In the symmetric relativistic model, the two hot-spots recede at a constant speed $`v_0`$ in opposite directions from an active galactic nucleus and any observed asymmetry is attributed to differences in light travel times to the observer. According to this model, the hot-spot approaching the observer at projected distance, $`r_\mathrm{a}`$, from the nucleus is older, and observed further from the nucleus, than the receding hot-spot which has projected distance $`r_\mathrm{r}`$. Banhatti (1980) introduced the *fractional separation difference*, $`x=(r_\mathrm{a}r_\mathrm{r})/(r_\mathrm{a}+r_\mathrm{r})`$, to describe the structural asymmetry of the radio hot-spots. According to the model, $`x`$ is proportional to radial component of the velocity of hot-spots, $$x\frac{r_\mathrm{a}r_\mathrm{r}}{r_\mathrm{a}+r_\mathrm{r}}=\beta \mathrm{cos}\theta ,$$ (1) where $`\beta =v_0/c`$, $`c`$ is the speed of light and $`\theta `$ is the angle between the radio axis and the line of sight to the observer. Assuming that the orientations of radio axes with respect to the line of sight are random, the simple integral equation relating the observed probability distribution of the radial velocities $`x`$, $`g(x)`$, to the probability distribution of true space velocities of the hot-spots $`G(\beta )`$ is $$g(x)=_x^1\frac{G(\beta )}{\beta }d\beta .$$ (2) (Banhatti 1980). The solution of this equation is $$G(\beta )=\beta g^{}(\beta ).$$ (3) The determination of $`G(\beta )`$ was the subject of the papers by Banhatti (1980) and Best et al. (1995). ### 2.1 Mean and dispersion of hot-spot velocities It is useful to introduce moments of the probability distribution $`G(\beta )`$. Multiplying (3) by $`\beta ^n\mathrm{d}\beta `$ and integrating from zero to one, the $`k^{\mathrm{th}}`$ moment of the observed fractional separation difference, $`\nu _k={\displaystyle \frac{\nu _{0k}}{k+1}},`$ (4) where the $`k^{\mathrm{th}}`$ moment of the distribution function of $`\beta `$ is $$\nu _{0k}=_0^1\beta ^kG(\beta )d\beta .$$ (5) The first and second moments of $`G(\beta )`$ and $`g(x)`$ are equal to the mean velocity and the mean square velocity, that is, $`\nu _{01}=\overline{\beta }`$, $`\nu _1=\overline{x}`$, $`\nu _{02}=\overline{\beta ^2}`$ and $`\nu _2=\overline{x^2}`$. Hence the mean space velocity and mean square velocity are, $$\overline{\beta }=2\overline{x}\mathrm{and}\overline{\beta ^2}=3\overline{x^2}$$ (6) The variance of the velocity about its mean value is $$\sigma _\beta ^2=\overline{\beta ^2}\overline{\beta }^2=3\overline{x^2}4\overline{x}^2.$$ (7) ### 2.2 Problems with the symmetric model #### 2.2.1 Expansion speed discrepancies The mean intrinsic speeds of the source components according to the symmetric relativistic model can be compared with those derived from synchrotron spectral ageing arguments. The latter speeds were estimated for 33 3CRR FRII radio sources in the power range $`10^{25}P_{178}10^{29}\mathrm{WHz}^1\mathrm{sr}^1`$ by Alexander & Leahy (1987) and Liu et al. (1992), assuming that the source axes lay in the plane of sky, thus providing estimates of the tangential velocities ($`y=v_\mathrm{t}/c`$) of the lobes. The true mean speed $`\overline{\beta }`$ is related to $`\overline{y}`$ by $$\overline{\beta }=\frac{4}{\pi }\overline{y}.$$ (8) In the same way, the standard deviations of the projected and true velocity distributions are related by $$\sigma _\beta =\sqrt{\frac{3}{2}\overline{y^2}\left(\frac{4}{\pi }\overline{y}\right)^2},$$ (9) where $`y=v_\mathrm{t}/c`$. The true mean speed of the source components derived from spectral ageing arguments from the above papers is $`\overline{\beta }=0.13\pm 0.08`$, compared to a mean value of $`\overline{\beta }=0.27\pm 0.16`$ from the observed distribution of fractional separation differences for the 132 3CRR FRII sources. Thus, the mean speed and its standard deviation of the source components are about two times greater than the same quantities derived from spectral ageing arguments. To check this result, the source sample described in Section 4 was divided into four equal logarithmic bins in radio power and mean velocity estimates were made using both techniques (Fig. 1). In this analysis, the cosmologies adopted by the above authors have been adopted, $`H_0=50`$ km $`\mathrm{s}^1`$ $`\mathrm{Mpc}^1`$ and deceleration parameter $`q_0=0`$. There were roughly equal numbers of sources in each of the bins. For all four bins, the mean speeds estimated from the relativistic symmetric model are systematically greater, by factors of between 1.5 and 5, than the mean speeds obtained from the spectral ageing analyses. The differences in the estimates becomes smaller for the most powerful sources. Whilst recognising that there are a number of reasons why the spectral ageing estimates might be in error, these discrepancies suggest that the velocities derived from the symmetric model are overestimated. The obvious interpretation of these data is that not all the source asymmetry can be attributed to the relativistic bulk motion of the source components. On the other hand, the discrepancy is less than a factor of two for the most luminous sources and so, although the asymmetries may not be wholly attributed to relativistic effects, they cannot be neglected. #### 2.2.2 Jet-side discrepancies Laing (1988) and Garrington et al. (1988) discovered the important correlation between the degree of depolarisation of a radio lobe and the presence of a one-sided radio jet emanating from the active nucleus. The sense of the correlation is such that the jet-side lobe is significantly less depolarized than the counterjet-side lobe and can be attributed to the fact that the radiation from the approaching lobe passes through a smaller path length of the depolarising halo surrounding the radio source than does the receding lobe. The Laing-Garrington effect thus enables the approaching and receding lobes to be identifed and suggests that the one-sidedness of the jet is the result of relativistic beaming. The lobe approaching the observer should be longer than the receding lobe and should be the side on which the one-sided radio jet is found. Saikia (1981) used the jet-side to test the symmetric model, initially for a small sample of radio galaxies, and later for a larger sample of 36 quasars (Saikia, 1984). For about half the quasars in his sample, the jet was found to lie on the shorter side. The same effect was also present in the larger sample of quasars and radio galaxies selected by Scheuer (1995). The latest observations of jets from high-resolution VLA observations indicate that, in sources with double-sided jets, there is no tendency for the brighter to lie in the longer radio lobe. According to the studies listed in Section 1, only 18 out of 33 radio galaxies, or 55%, show the expected correlation and, in the quasar sample of Bridle et al. (1994), only 6 out of 13 sources follow the expectation of the relativistic symmetric model. ## 3 The asymmetric relativistic model The *asymmetric relativistic model* is designed to take account of the contributions of both relativistic and intrinsic/environmental asymmetries to the structures of the radio sources. The jets advance through a clumpy asymmetric environment at an angle $`\theta `$ to the line of sight (Fig. 2). Asymmetries associated with both the environment and the intrinsic properties of the jets can be described by assuming that the mean velocity of the lobe in the jet (approaching) direction $`v_\mathrm{j}`$ and that in the counterjet (receding) direction $`v_{\mathrm{cj}}`$ are different. Throughout this analysis, it is assumed that the jet-side, or the brighter of a two-sided jet, lies in the lobe approaching the observer and we will consistently use the term ‘jet-side’ to have this meaning, even if the jets are absent, but the orientiation of the source can be found from other arguments. As in the symmetric case, a *fractional separation difference* $`x`$ can be defined, but now the relationship between $`x`$, $`v_\mathrm{j}`$ and $`v_{\mathrm{cj}}`$ is slightly more complex than (1): $$x\frac{r_\mathrm{j}r_{\mathrm{cj}}}{r_\mathrm{j}+r_{\mathrm{cj}}}=\frac{v_\mathrm{j}v_{\mathrm{cj}}}{v_\mathrm{j}+v_{\mathrm{cj}}}+\frac{2}{c}\frac{v_\mathrm{j}v_{\mathrm{cj}}}{v_\mathrm{j}+v_{\mathrm{cj}}}\mathrm{cos}\theta ,$$ (10) where $`r_\mathrm{j}`$ and $`r_{\mathrm{cj}}`$ are the projected distances between the active nucleus and the furthest ends of the lobes on the jet and counterjet sides respectively, and $`c`$ is the speed of light. In the symmetric relativistic model, $`v_\mathrm{j}=v_{\mathrm{cj}}=v_0`$ and the value of $`x`$ is always positive, $`x[0;1]`$. In the asymmetric relativistic model, however, negative values of $`x`$ can be found, $`x[1;1]`$. Specifically, if the first term on the right-hand side of (10) is negative and of greater magnitude than the second, negative values of the fractional separation difference are found. The speeds $`v_\mathrm{j}`$ and $`v_{\mathrm{cj}}`$ can be written, $`v_\mathrm{j}=v_0+v_{\mathrm{jd}}`$ and $`v_{\mathrm{cj}}=v_0+v_{\mathrm{cjd}}`$, where $`v_{\mathrm{jd}}`$ and $`v_{\mathrm{cjd}}`$ are the changes to $`v_0`$ due to intrinsic/environmental asymmetries on the jet and counterjet sides respectively. Introducing new variables $`\delta _\mathrm{j}=1+v_{\mathrm{jd}}/v_0`$ and $`\delta _{\mathrm{cj}}=1+v_{\mathrm{cjd}}/v_0`$, (10) can be written $$x=\frac{\delta _\mathrm{j}\delta _{\mathrm{cj}}}{\delta _\mathrm{j}+\delta _{\mathrm{cj}}}+\frac{2}{c}\frac{\delta _\mathrm{j}\delta _{\mathrm{cj}}}{\delta _\mathrm{j}+\delta _{\mathrm{cj}}}v_0\mathrm{cos}\theta .$$ (11) In the symmetric case, $`\delta _\mathrm{j}=\delta _{\mathrm{cj}}=1`$, we recover (1) and $`v_0`$ is the average speed of advance of the lobes. In the asymmetric model, the same interpretation is correct, provided the mean value of the joint distribution of $`\delta _\mathrm{j}`$ and $`\delta _{\mathrm{cj}}`$ is unity. In other words, $`v_0`$ is the average advance speed of the lobes, provided averages are taken over large samples of sources normalised so that $`\delta =1`$. In physical terms, we would expect $`v_0`$ to depend on the mean jet-counterjet power and the mean environmental density, whereas $`\delta _\mathrm{j}`$ and $`\delta _{\mathrm{cj}}`$ are measures of the intrinsic asymmetry of the jets and inhomogeneities in the environments of the radio sources. Let us examine (11) for positive and negative values of $`x`$. * If $`\delta _\mathrm{j}\delta _{\mathrm{cj}}0`$, $`x0`$ for all $`\theta `$. * If $`\delta _\mathrm{j}\delta _{\mathrm{cj}}0`$, two cases are possible: $`x0`$ for $`\theta <\theta _\mathrm{n}`$ and $`x0`$ for $`\theta >\theta _\mathrm{n}`$ where $$\theta _n=\mathrm{arccos}\left(\frac{c}{v_0}\frac{|\delta _\mathrm{j}\delta _{\mathrm{cj}}|}{2\delta _\mathrm{j}\delta _{\mathrm{cj}}}\right).$$ (12) The viewing angles for FRII sources with the jet on the long lobe side ($`x0`$, hereafter, +FRII sources) span all angles from $`0`$ to $`\pi /2`$, whereas the viewing angles for FRII sources with the jet on the short lobe side ($`x0`$, hereafter $``$FRII sources) are restricted to the range $`\theta _\mathrm{n}\theta \pi /2`$ (see Fig. 3). If relativistic effects are negligible ($`v_0c`$), the value of $`\theta _\mathrm{n}0`$; if relativistic effects are the predominant cause of the observed asymmetries, $`\theta _\mathrm{n}\pi /2`$. Thus, the radio axes of *relativistic* $``$FRII sources lie preferentially close to the plane of sky, while the axes of *non-relativistic* $``$FRII sources can be seen at almost all angles. Notice that, if the sources are non-relativistic, $`v_0c`$, and are significantly asymmetric intrinsically, the argument of arccos in (12) can become greater than 1 and then negative values of $`x`$ can be found for all angles in the range $`0<\theta <\pi /2`$. For sources in which the relativistic effects are dominant, these should be greater for the +FRII sources than for $``$FRII sources; the intrinsic/environmental asymmetry is more significant for $``$FRII sources and should be dominant at low velocities for both +FRII and $``$FRII sources. Large positive values of $`x`$ are found in FRII sources which expand with the highest speeds $`v_0`$ at angles close to the line of sight $`\theta 0`$, while large negative values of $`x`$ are associated with intrinsically asymmetric sources lying close to the plane of the sky. Let us consider how the ratio of the numbers of $``$FRII to +FRII sources depends upon the value of $`v_0`$. As $`v_00`$, $`x`$ is an indicator of the intrinsic/environmental asymmetry of FRII sources, $`x(v_{\mathrm{jd}}v_{\mathrm{cjd}})/(v_{\mathrm{jd}}+v_{\mathrm{cjd}})`$. If the jet axes and the intrinsic/environmental asymmetries are distributed isotropically, the velocities $`v_{\mathrm{jd}}`$ and $`v_{\mathrm{cjd}}`$ on the jet and counterjet sides should be randomly distributed with the same probability distribution. Then, the ratio of the number of $``$FRII to +FRII sources should be equal, that is, $`N(\text{FRII})/N(+\text{FRII})1`$. If the relativistic effect is dominant, $`v_{\mathrm{jd}}`$, $`v_{\mathrm{cjd}}v_0`$ and $`x\mathrm{cos}\theta 0`$. In this case, the ratio $`N(\text{FRII})/N(+\text{FRII})`$ tends to $`0`$. We can define an *asymmetry parameter* $`\epsilon `$, where $$\epsilon =12\frac{N(\text{FRII})}{N(+\text{FRII)}}.$$ (13) Relativistic effects are more important than intrinsic/environmental asymmetries if $`0<\epsilon 1`$, while, if $`1\epsilon <0`$, the latter effects are dominant. If $`\epsilon 0`$, the contribution of each effect is of comparable importance. Let us consider how the observed distribution of $`x_\mathrm{l}`$ depends upon the properties of the asymmetric model. The mean speeds of the jet and counterjet lobes are the sums of the mean intrinsic speed of the lobes, $`v_0`$, and the random dispersion about the mean, $`v_\mathrm{j}=v_0+v_{\mathrm{jd}}`$ and $`v_{\mathrm{cj}}=v_0+v_{\mathrm{cjd}}`$. If we assume that the intrinsic asymmetry is distributed isotropically in the sky, the distribution function of the space velocities of the lobes on the jet- and counterjet-sides should be the same, $`F(v_\mathrm{j})=F(v_{\mathrm{cj}})=F(v_\mathrm{l})`$, implying that the distribution functions $`G(v_{\mathrm{jd}})=G(v_{\mathrm{cjd}})=G(v_\mathrm{d})`$. For illustrative purposes, let us assume that (i) the distribution function of the intrinsic velocities $`v_0`$ is Gaussian with the mean $`\overline{v}_0`$ and standard deviation $`\sigma _{v_0}`$ and (ii) the distribution function of random disturbed velocities is also a Gaussian with constant standard deviation $`\sigma _{v_\mathrm{d}}`$ on the jet- and counterjet sides. We assume that these normal distribution functions are independent. Then, the distribution function of their sum, the lobe speeds $`v_\mathrm{j}`$ and $`v_{\mathrm{cj}}`$ are also normal with a mean speed $`\overline{v}_0`$ and dispersion $`\sigma _{v_\mathrm{l}}^2=\sigma _{v_0}^2+\sigma _{v_\mathrm{d}}^2`$, $$F(v_\mathrm{l})=F(v_0+v_\mathrm{d})=\frac{1}{\sigma _{v_\mathrm{l}}\sqrt{2\pi }}\mathrm{exp}^{\frac{(v_\mathrm{l}\overline{v}_0)^2}{2\sigma _{v_\mathrm{l}}^2}}.$$ (14) For illustrative purposes, let us adopt a mean lobe speed $`\overline{v}_\mathrm{l}=0.15c`$ and $`\sigma _{v_\mathrm{l}}=0.05c`$. Then, in Fig. 4 and Table 1, the predicted distributions of $`x_\mathrm{l}`$, and the mean values of $`\epsilon `$ and $`x_\mathrm{l}`$, are shown for different combinations of the intrinsic dispersions in the lobe speeds $`\sigma _{v_0}`$ and the intrinsic/environmental asymmetry $`\sigma _{v_\mathrm{d}}`$. The first and fourth lines of Table 1 correspond to the limiting cases of $`\sigma _{v_\mathrm{d}}0`$ and 0.05 respectively. It can be seen that the distribution function of $`x_\mathrm{l}`$ becomes more asymmetric for large values of $`\sigma _{v_\mathrm{d}}`$, leading to a decrease in the value of $`\overline{x}_\mathrm{l}`$. Therefore the simple relation (6) between the mean speed and the mean of the observed fractional separation difference underestimates $`\overline{v}_\mathrm{l}`$ and should be written $$\overline{v}_\mathrm{l}/c=\overline{\beta }2\overline{x}_\mathrm{l}.$$ (15) The asymmetry parameter is a better indicator of the intrinsic/environmental asymmetry than $`\overline{x}_\mathrm{l}`$ (see Table 1). Notice that the value $`\epsilon 0`$ is attained when $`\sigma _{v_\mathrm{d}}\sigma _{v_0}`$, that is, the intrinsic/environmental and relativistic effects are of comparable importance. It should be noted that the asymmetry parameter $`\epsilon `$ is a rather crude quantitative measure of the role of intrinsic/environmental asymmetries because of the small number statistics generally involved in determining $`N(\mathrm{FRII})`$ and $`N(+\mathrm{FRII})`$, resulting in quite large error bars in the estimates of $`\epsilon `$. ## 4 Sample of FRII radio sources ### 4.1 Statistics The analysis was based upon an analysis of the sample of 132 FRII 3CR radio sources selected from the catalogue of Laing, Riley and Longair (1993) which lay in the area of sky $`\delta 10^{}`$ and $`|b|10^{}`$. The jet-side could be determined for 103 of the 3CRR sources using the following criteria in order of preference: * Direct VLA/VLBI images of kiloparsec and/or parsec-scale jets (82%). As discussed by Hardcastle et al. (1998), where high resolution, high sensitivity maps of sources are available, well-defined radio jets are remarkably common. We use the term ‘definite’ jets (D) when a jet can be defined according to the definition of Bridle and Perley (1984), namely, that the radio feature is at least four times longer than it is wide. In a number of cases, evidence for radio jets has been found at a reasonable level of confidence within the extended radio source structure and we refer to these as ‘probable’ jets: these correspond to the ‘possible’ jets of Hardcastle et al. (1997) and the ‘candidate’ jets of Fernini et al. (1997). Furthermore, Pearson & Readhead (1988) found that, in cases where both a VLBI (parsec-scale) jet and kiloparsec jet are seen, the VLBI jet is connected with, or is on the same side as, the kiloparsec jet. The same result was found in the sample of sources studied by Wardle & Aaron (1997). Parsec-scale jets could therefore be used to define the jet-side. * Depolarization asymmetry – the Laing-Garrington effect (13%). It is assumed that the less depolarized lobe is approaching the observer. * Spectral index asymmetry – the Liu-Pooley effect (5%). Liu & Pooley (1991) found a strong correlation between the depolarization and spectral index asymmetries of the lobes for a small sample of FRII radio sources, in the sense that the lobe which is less depolarized has the flatter spectrum. The lobe with the flatter spectrum is assumed to be approaching the observer. The remaining 29 sources could not be classified for a number of reasons. In a number of cases, excellent radio maps were obtained by Fernini et al. (1993, 1997), but no evidence for jets could be discerned. In other cases, radio maps of sufficiently high quality have not been published. Scheuer (1995) used the statistics of the structural asymmetries of the radio lobes, rather than hot-spots, to define the fractional separation differences, arguing that the hot-spots are ephemeral structures which can change their location within the volume of the lobe; the largest lobe length is therefore likely to be a more stable measure of the mean growth rate of the lobe. Means and standard deviations of the fractional separation differences for the lobes and hot-spots in the present sample are $`\overline{x}_\mathrm{l}=0.07\pm 0.18`$ and $`\overline{x}_\mathrm{h}=0.06\pm 0.22`$, respectively. The greater standard deviation of $`x_\mathrm{h}`$ as compared with that of $`x_\mathrm{l}`$ shows that the locations of hot-spots are on the average more asymmetric than the lobe lengths, consistent with Scheuer’s argument. For this reason, the fractional separation difference of lobes has been used in the rest of the analysis. It should be noted that the larger dispersion in $`x_\mathrm{h}`$ as compared with $`x_\mathrm{l}`$ means that the high mean velocities of expansion found by Longair & Riley (1979), Banhatti (1980) and Best et al. (1995) are overestimates. Details of the sources included in this analysis and the criteria used, as well as the adopted values of the fractional separation differences $`x`$ for the lobes and hot-spots are given in the Appendix. Of the 103 FRII sources, 71 were radio galaxies and 32 were quasars. Among the 71 radio galaxies, the jet-side was determined (i) by the presence of jets (34 cases) and probable jets (21 cases), (ii) from depolarization asymmetries (12 cases) and (iii) from spectral index asymmetries (4 cases). Among the 32 quasars, 28 cases of jets and two probable jets were found; the jet-side was found from depolarization asymmetries and from spectral index asymmetries in one case each. Two jets among the 36 radio galaxies were detected on the parsec-scale from VLBI images. Among the 103 sources, there are 67 +FRII sources (65%) and 36 $``$FRII sources (35%). Of the 71 radio galaxies, 43 are +FRII and 28 are $``$FRII sources; among the 32 quasars, 24 are +FRII and 8 $``$FRII sources. Fig. 5 shows the distributions of the fractional separation differences for the symmetric ($`0<|x_\mathrm{l}|<1`$) and asymmetric pictures ($`1<x_\mathrm{l}<1`$). The latter distribution, containing 36 $``$FRII sources (35%), is asymmetric, suggesting that both intrinsic/environmental and relativistic effects are important. ### 4.2 Selection effects The 3CRR sample is a complete unbiased sample of sources selected at a low radio frequency. It is orientation-independent since the flux densities of the sources in the sample are generally dominated by the diffuse emission of the radio lobes. The presence of jets or probable jets has been found in virtually all 32 quasars in the sample and in 70% of the radio galaxies. Therefore, the uncertainties are associated with the presence or otherwise of jets in the remaining 20% of the complete sample and these are not likely to be large. Let us consider first the reliability of using the ‘probable’ jets in the analysis. The sample was divided into two subsamples, the first containing all sources with probable jets, and the second the remaining 84 sources with definite jets. The statistics of probable jets shows that 16/22 (72%) are +FRII sources while 6 (28%) are $``$FRII sources. For the second sample, 51/81 (62%) and 30/81 (38%) are +FRII and $``$FRII sources respectively. We conclude that no bias is introduced by including the probable jets in the analysis. According to the relativistic beaming model for the radio jets, there is a strong selection effect favouring the detection of jets in FRII sources which are oriented close to the line of sight. Within the context of orientation-based unified schemes, it is therefore natural that the sources without detectable jets should be radio galaxies in which the axis of ejection lies close to the plane of the sky. For these ‘missing’ sources, the fractional separation difference should be predominantly due to intrinsic/environmental asymmetries and the distributions of the space velocities of lobes $`v_\mathrm{j}`$ and $`v_{\mathrm{cj}}`$ should be symmetric about $`x_\mathrm{l}=0`$. Therefore, jet orientation selection effect would not change the symmetry of the observed distributions of $`x_\mathrm{l}`$, but would slightly decrease the asymmetry parameter. In the worst case, the 29 sources for which the jet side has not been determined would be distributed exactly symmetrically and so we would expect 14.5 more $``$FRII and 14.5 more $`+`$FRII sources, reducing the estimate of $`\epsilon `$ from the value $`\epsilon =0.3`$ found in Section 6.1 to $`0.47`$. The sample of 103 sources was derived from a literature search of all available radio maps and it is important to compare the statistics with samples which have been observed with more uniform criteria of sensitivity and angular resolution. High-resolution observations of a sample of 50 3CR FRII radio galaxies with $`z<0.3`$ are presented by Hardcastle et al. (1998), representing the combined samples of Black et al. (1992), Leahy et al. (1997) and Hardcastle et al. (1997). 34 of the 50 FRII radio galaxies belong to the 3CRR complete sample. Of these 34 radio galaxies, only 6 (or 18%) did not possess jets or probable jets, a somewhat greater success rate than for our sample, as expected. For 12 radio galaxies, depolarisation asymmetries have been used to determine the jet-side where no radio jet has been detected. Since no jet was detected, the axes of these sources are likely to lie close to the plane of the sky and intrinsic/environmental asymmetries are likely to be important. These may also be more important than the Laing-Garrington effect in causing the observed depolarisations (Pedelty et al. 1989a and McCarthy et al. 1991). Taking the jet-side to be the less depolarized side may change the true sign of $`x_\mathrm{l}`$ from positive to negative or vice versa with equal probability, as the orientation of the jet axes are assumed to be isotropic. This selection effect is not likely to influence strongly the observed distribution of $`x_\mathrm{l}`$ and the asymmetry parameter because it concerns only 12 sources. We have repeated the complete analysis excluding these 12 radio galaxies and found that their exclusion makes no difference to our conclusions. Scheuer (1995) noted a number of other effects which could bias the analysis of the jet-side of the FRII sources. For example, misalignment of the jets, intrinsic one-sidedness of the radio jets, and misalignment and distortion of the radio lobes, can all complicate the analysis. These could all be modelled (see for example Best et al. 1995), but the effects were found to be quite small in that analysis. ## 5 Fractional separation differences of the radio lobes ### 5.1 Correlations with radio luminosity A plot of radio luminosity against fractional separation difference is presented in Fig. LABEL:fig-6. The diagram is V-shaped with a tendency for the spread in fractional separation difference to increase with increasing luminosity up to $`P_{178}10^{29}\mathrm{WHz}^1`$. Intrinsic/environmental asymmetries are necessary to account for negative values of $`x_\mathrm{l}`$ and there is a trend for this asymmetry to increase with luminosity, or with redshift since, in the present flux density limited sample, radio luminosity and redshift are correlated. A possible explanation for this increase is that high redshift ($`z>0.5`$) FRII radio galaxies are associated with the brightest galaxies in rich clusters, while their low redshift counterparts tend to lie in isolated environments or in small groups (Lilly & Prestage 1987, Hill & Lilly 1991, Best et al. 1998). Gradients in the intergalactic gas density should be present in these rich clusters on the scale of the overall structure of the radio sources, which might lead to asymmetric environments on large scales. Furthermore, Best et al. (1996, 1997) present evidence that the interstellar medium in the vicinity of 3CR radio galaxies at $`z1`$ must contain clumpy cold gas to account for the pronounced alignment effect displayed by these galaxies, which is not observed at small redshifts. Both intrinsic/environmental and relativistic effects contribute to positive values of $`x_\mathrm{l}`$, and there is considerable spread of these values at high luminosities. The luminosity–fractional separation difference correlation for high luminosity quasars $`P_{178}>10^{29}\mathrm{WHz}^1`$ is of special interest since jet-sides have been determined for all of them. There are two important features of their distribution. First, there are only 2 $``$FRII quasars and, second, there is only one +FRII quasars with $`x_\mathrm{l}0.25`$ (Fig. LABEL:fig-6). The lack of $``$FRII quasars suggests that the relativistic effect is dominant for luminous quasars, that the axes of quasars must lie at reasonably small angles to the line of sight and that the ratios $`v_{\mathrm{jd}}/v_0`$ and $`v_{\mathrm{cjd}}/v_0`$ should tend to zero. The lack of quasars with large positive values of $`x`$ suggests that most values of $`x`$ must lie in the restricted range $`0<x_\mathrm{l}x_{\mathrm{max}}`$. In the case of the symmetric relativistic model, $`x_\mathrm{l}=v_0\mathrm{cos}\theta /c`$. Recognising that there must be some residual environmental/intrinsic effect present at these high luminosities, a reasonable upper limit to the value of $`x`$ is $`x_{\mathrm{max}}=0.26`$, which is shown by the vertical dashed line. Then, assuming $`\theta 0`$, we obtain an firm *upper limit* to the expansion speed of the lobes of $`v_0x_\mathrm{l}c=0.26c`$ for high luminosity quasars. At lower luminosities, the $`+`$FRII sources with $`x_\mathrm{l}0.26`$ and the $``$FRII sources may be accounted for by a combination of intrinsic/environmental asymmetries in mildly relativistic sources. This picture is supported by the corresponding values of the asymmetry parameter $`\epsilon `$. For all quasars with $`P_{178}>10^{29}`$ W Hz<sup>-1</sup>, the asymmetry parameter is $`\overline{\epsilon }_\mathrm{Q}(>10^{29})0.7\pm 0.24`$, whereas for quasars with smaller luminosities, $`\overline{\epsilon }_\mathrm{Q}(<10^{29})0.1\pm 0.5`$. Thus, the relativistic effects appear to be dominant at high luminosities, whereas at lower luminosities, both effects are of comparable importance. This last result appears to differ from that found by Scheuer (1995) who found that relativistic effects are small for the most luminous sources. His analysis was based upon the ratios of lobe lengths for three samples of quasars, including a sample of 19 quasars from the LRL sample. For this sample, Scheuer found ratios of the arm lengths which correspond to $`\overline{x}_\mathrm{l}=0.097`$, in excellent agreement with our estimate for all the quasars in our sample of $`\overline{x}_\mathrm{l}(\mathrm{QSO})=0.102`$. The value of the asymmetry parameter for the LRL quasars studied by Scheuer is $`\epsilon =0.29`$, compared with our value of 0.33. Thus, there is no discrepancy between Scheuer’s results and the present analysis for corresponding sets of data. It may be significant that one of the other samples of discussed by Scheuer, the BMSL sample, was restricted to quasars with redshifts $`z>1.5`$, whereas the present sample of 103 sources has only four objects with redshifts greater than this value. We note that Scheuer’s sample was largely restricted to quasars, whereas our sample includes all radio galaxies and quasars in the 3CRR sample and should be free from systematic bias. We also note that the results of multi-epoch global VLBI observations of Compact Symmetric Objects, which are presumed to be young radio-loud sources, show strong evidence for mildly relativistic hotspot speeds, $`v_\mathrm{h}>0.1c`$ (Owsianik & Conway 1998, Owsianik et al. 1998). ### 5.2 Correlation with projected linear size Fig. LABEL:fig-7 shows a plot of projected linear size against fractional separation difference. This diagram has roughly the form of an inverted V-shape, which is not particularly surprising in view of the well-known inverse correlation between project linear size and radio luminosity. The spread of negative values of $`x_\mathrm{l}`$ for small sources indicates that the intrinsic asymmetries are important on small physical scales, and many of these are associated with radio galaxies. The majority of the quasars have positive values of $`x`$. There is some suggestion that the dispersion in $`x`$ becomes smaller at larger physical sizes. The sample has been divided into two equal samples of small and large sources and these sub-samples have asymmetry parameters $`\epsilon _{\mathrm{G}+\mathrm{Q}}(<250\text{kpc})0.3\pm 0.2`$ and $`\epsilon _{\mathrm{G}+\mathrm{Q}}(>250\text{kpc})0.7\pm 0.4`$ respectively. These results indicate that relativistic and environmental/intrinsic effects are important for smaller physical sizes, while intrinsic symmetries are more important for larger sources. These differences in asymmetry parameters can be at least partly explained by projection effects. If the larger sources lie predominantly close to the plane of the sky, they will be observed to be larger than those seen at a smaller angles to the line of sight and their structures would be dominated by intrinsic/environmental asymmetries. The smaller sources are observed closer to the line of sight and so the relativistic effects are of greater importance, although the intrinsic asymmetries still play a significant rôle. The fact that the quasars all lie towards the lower part of the diagram is consistent with an orientation-based unification picture. Thus the results of Section 5.2 and 5.3 are not independent, but are part of a consistent picture. ## 6 Statistical analysis ### 6.1 Asymmetry parameter The asymmetry parameter for the sample of 103 FRII sources is $`\epsilon _{\mathrm{G}+\mathrm{Q}}=0.07\pm 0.22`$, while for radio galaxies and quasars separately the values are $`\epsilon _\mathrm{G}=0.3\pm 0.32`$ and $`\epsilon _\mathrm{Q}=0.33\pm 0.36`$ respectively. There are no quasars of low radio luminosity in the 3CRR sample and so a smaller sample limited to high luminosity sources $`P_{178}10^{27.7}\mathrm{WHz}^1`$ has also been studied, this luminosity corresponding to the lower luminosity limit for quasars in the 3CRR sample. The values of the asymmetry parameters for these samples are $`\epsilon _{\mathrm{G}+\mathrm{Q}}=0\pm 0.24`$, $`\epsilon _\mathrm{G}=0.27\pm 0.36`$ and $`\epsilon _\mathrm{Q}=0.33\pm 0.36`$ for 81, 49 and 32 sources respectively. Thus, the results for the whole and restricted samples are similar. These results suggest that intrinsic/environmental asymmetries and relativistic effects are of comparable importance for the sample overall, but the intrinsic asymmetries are more significant for the radio galaxies and relativistic effects are for the quasars, $`\epsilon _\mathrm{G}<\epsilon _\mathrm{Q}`$, consistent with the asymmetric relativistic model and in qualitative agreement with the expectations of orientation-based unification schemes. Furthermore, the mean value of $`x_\mathrm{l}`$ for the sample of radio galaxies and quasars is $`\overline{x}_{\mathrm{G}+\mathrm{Q}}=0.07\pm 0.018`$, while the values for radio galaxies and quasars separately are $`\overline{x}_\mathrm{G}=0.056\pm 0.022`$ and $`\overline{x}_\mathrm{Q}=0.102\pm 0.028`$, again consistent with the above picture. ### 6.2 One- and two-sided jets For the sources in which jets have been detected, the proportion of two-sided jets is roughly the same for $``$FRII and +FRII radio galaxies – 6/28, or 21%, for $``$FRII and 8/43, or 19%, for +FRII radio galaxies. For the quasars, on the other hand, 4/8 (50%) of the $``$FRII sources and 3/24 (13%) of the +FRII sources have two-sided jets. This difference between the fractions of $``$FRII and +FRII quasars is consistent with the $``$FRII quasars lying closer to the plane of the sky than the +FRII quasars, as well as with the expectations of the asymmetric relativistic model and with orientation-based unification schemes for radio galaxies and quasars. Of the 12 highest luminosity FRII quasars, 11 have one-sided jets and 1 is two sided (92% and 8%). The asymmetry parameter for these sources is $`\epsilon _\mathrm{Q}(P_{178}>10^{29})0.7\pm 0.2`$ which strongly supports the supposition that relativistic effects are dominant for these sources. The only anomaly is that the percentage of $``$FRII quasars with two-sided jets is greater than that of $``$FRII radio galaxies ($`50\%>21\%`$), since we would expect more two-sided jets among the $``$FRII galaxies. The small statistics of $``$FRII quasars may partly be the cause of this discrepancy. Furthermore, the detection of jets is strongly dependent upon the quality of the radio maps and generally the quasars have been studied with higher resolution and sensitivity than the radio galaxies. ### 6.3 Mean linear size The mean linear sizes of $``$FRII and +FRII sources have been calculated for two samples. The first consists of 19 low luminosity sources with $`P_{178}<10^{27.7}\mathrm{WHz}^1`$, the lower radio luminosity limit for the quasars in the sample. Three sources DA240, 3C236 and 3C326 were excluded from the calculation because of their extremely large linear sizes (2070, 5950 and 2660 kpc). For these low luminosity radio galaxies, the mean linear size of $`(450\pm 330)`$ kpc for +FRII sources is larger than that $`(300\pm 140)`$ kpc for $``$FRII sources (Table 2), where the ‘errors’ are the standard deviations of the distributions. The large negative asymmetry parameter, $`\epsilon 0.8`$, for this low luminosity sample as a whole shows that intrinsic asymmetries play the dominant rôle in their structural asymmetries. The inference is that the lobes of these radio galaxies have non-relativistic expansion speeds. Furthermore, these low luminosity radio galaxies have hot-spot advance speeds determined from synchrotron ageing arguments which are significantly smaller than those of the higher luminosity sources, as can be seen from Fig. 1. For the sample of 81 high luminosity radio galaxies and quasars, the mean linear size of the $``$FRII radio galaxies is approximately equal to the mean linear size of +FRII radio galaxies $`(290280\mathrm{kpc})`$, while for quasars the mean size of $``$FRII quasars is twice as large as the mean size of the +FRII quasars (Table 2). This is consistent with a model in which the axes of the radio galaxies lie at larger angles to the line of sight than those of the quasars, as illustrated in Fig. 8. For large angles, the projected size varies slowly and therefore the mean sizes of both $``$FRII and +FRII radio galaxies should be comparable on average. On the other hand, at smaller angles to the line of sight, the projected size varies more rapidly. According to the model, the $``$FRII quasars are observed on average at greater angles ($`\theta _\mathrm{c}>\theta >\theta _\mathrm{n}`$), than the +FRII quasars, ($`\theta <\theta _\mathrm{n}`$), and so should have larger projected sizes. The high luminosity sample has been divided into two equal high- and low-luminosity samples to test these predictions (Table 3). For high-luminosity radio galaxies, the difference between mean sizes of $``$FRII and +FRII sources ($`280>220`$ kpc) is greater than for low luminosities ($`330>300`$ kpc), suggesting that the difference in the mean sizes increases with luminosity. In addition, the asymmetry parameters are different for the two samples, $`0.44`$ for the low luminosity sample and $`0`$ for the high luminosity sample, although the statistical significance is not great. This further supports the inference that relativistic effects are more important for the highest luminosity sources; this relation is consistent with the results of spectral ageing analyses which indicate that expansion speeds of lobes increases with luminosity (Fig. 1). ### 6.4 Spectral classification The asymmetry parameters and mean linear sizes for broad-line radio galaxies and quasars, narrow-line radio galaxies and low-excitation radio galaxies have been evaluated for both the $``$FRII and +FRII sources (Table 4). The same three giant sources were excluded from the sample. For broad-line objects, the mean projected size of $``$FRII is larger than the mean projected size of +FRII sources ($`320>180`$ kpc), while for narrow-line galaxies the mean sizes are similar $`(280320\mathrm{kpc})`$, consistent with orientation-based unified schemes. This result is not particularly surprising since the broad-line radio galaxies are naturally identified with the low redshift ‘quasar’ population. There are 11 low-excitation radio galaxies in the sample, most of which are relatively low-luminosity sources $`(P_{178}<10^{28.5}\mathrm{WHz}^1)`$. The mean sizes of low-excitation and high-excitation radio galaxies are 240 kpc and 380 kpc respectively. The same result was obtained by Hardcastle et al. (1997) for a combined sample of powerful radio galaxies with $`z<0.3`$. The mean projected size of low-excitation radio galaxies is smaller than that of the narrow-line radio galaxies of the same luminosities. The roughly equal numbers of the $``$FRII and +FRII low-excitation radio galaxies (Table 4) indicates that their structural asymmetry is caused almost entirely by intrinsic asymmetries, which implies non-relativistic expansion speeds $`(v_0c)`$ and/or preferential orientation of their axes close to the plane of the sky ($`\theta \pi /2`$). This supports the view that the low-excitation radio galaxies are a separate group of FRII sources with non-relativistic expansion velocities and that they are not unified with high-excitation sources (Laing et al. 1994). ## 7 Numerical simulations The expected distribution of the fractional separation differences for the radio galaxies and quasars have been modelled, with the aim of estimating quantitatively the mean speed of the lobes, its standard deviation, the typical intrinisic/environmental asymmetries of the FRII sources and the critical angle separating the radio galaxies and the quasars according to the orientiation-based unification scenario. ### 7.1 Expansion speeds and intrinsic/environmental asymmetries A lower limit to the lobe advance speeds can be found from the expression (15). Adopting $`\overline{x}_\mathrm{l}=0.07\pm 0.018`$, we find $$\overline{v}_\mathrm{l}/c2\overline{x}_\mathrm{l}=0.14\pm 0.036$$ (16) Adopting the simple Gaussian assumptions made in Section 3, a procedure has been developed for determining the best fitting values of the mean lobe speed standard, $`\overline{v}_\mathrm{l}`$, the standard deviation of the lobe speeds $`\sigma _{v_\mathrm{l}}`$ and the standard deviation of the disturbed velocities $`\sigma _{v_\mathrm{d}}`$. For a given set of parameters, 10,000 random values of these velocities were selected from the joint Gaussian distributions for each source component and the value of $`\mathrm{cos}\theta `$ were selected at random from a uniform distribution between 0 and 1. The best fitting values were found by minimising the parameter $`M=(|z_{\mathrm{obs}}(x_\mathrm{l})z_{\mathrm{pr}}(x_\mathrm{l})|)/N`$; in this expression $`z(x_\mathrm{l})\mathrm{d}x_\mathrm{l}=N(x_\mathrm{l})/N`$, where $`N(x_\mathrm{l})`$ represents the observed (obs) or predicted (pr) numbers of sources in intervals of 0.1 in $`x_\mathrm{l}`$ and $`N`$ is the total number of sources. The two-sigma lower limit for the asymmetry parameter and the mean lobe speed correspond to $`M0.17`$ and the ranges of values of these parameters with $`M<0.17`$ are $`0.47<\epsilon <0.07`$ and $`0.07<\overline{v}_\mathrm{l}/c<0.15`$. The best-fitting values of the parameters are: $`\overline{v}_\mathrm{l}/c=0.11\pm 0.013`$, $`\sigma _{v_0}=(0.032\pm 0.002)c`$, $`\sigma _{v_\mathrm{d}}=(0.023\pm 0.002)c`$ and $`\sigma _{v_\mathrm{l}}=(0.04\pm 0.004)c`$. A comparison of the predictions of the model with these best-fitting values of the parameter with the observations is shown in Fig. 9(a). The mean expansion speed of the lobes is comparable to the results of spectral ageing analyses, $`\overline{v}=(0.13\pm 0.08)c`$, and, as expected, less than the mean speeds estimated for the symmetric model ($`>0.2c`$). ### 7.2 Unified scheme We use (11) to model the fractional separation differences for radio galaxies and quasars separately. For illustrative purposes, the best-fitting values of the distributions found in section 7.1 have been used, that is, normal distributions for $`v_0`$, $`v_{\mathrm{jd}}`$ and $`v_{\mathrm{cjd}}`$ for both radio galaxies and quasars, the only free parameter being the critical angle $`\theta _\mathrm{c}`$. Values of $`\mathrm{cos}\theta `$ were selected at random from the uniform distribution from 1 to $`\mathrm{cos}\theta _\mathrm{c}`$ for quasars and from $`\mathrm{cos}\theta _\mathrm{c}`$ to 0 for radio galaxies. The results of varying $`\theta _\mathrm{c}`$ are presented in Table 5. It can be seen that good agreement between predicted and observed asymmetry parameters, $`\epsilon _\mathrm{G}=0.3`$ and $`\epsilon _\mathrm{Q}=0.33`$, is found for $`\theta _\mathrm{c}45^{}`$; furthermore, the predicted distribution functions of $`x_\mathrm{l}`$ are separately in good agreement with the distributions for galaxies and quasars separately, as can be seen in Fig. 9(b), and with the mean observed values of $`\overline{x}_\mathrm{G}`$ and $`\overline{x}_\mathrm{Q}`$, namely, $`0.056\pm 0.022`$ and $`0.102\pm 0.028`$ respectively (see Table 5). We conclude that the asymmetric relativistic model is in agreement with a unified scheme for $`\theta _\mathrm{c}45^{}`$, the value favoured by Barthel (1989). The success of the modelling also indirectly confirms the validity of the estimates of the mean lobe speed and its standard deviation. ## 8 Conclusions and discussion The principal conclusions of this analysis are as follows: 1. A statistical analysis of the sample of 3CRR sources shows that (a) roughly one third of the FRII sources have a jet on the short lobe side; (b) the predictions of the asymmetric relativistic model can provide a good description of the structural properties of the radio sources in the sample; (c) the contributions of relativistic effects and intrinsic/environmental asymmetries to the structural asymmetry are of comparable importance. Relativistic effects are more important for quasars and high-luminosities, while intrinsic asymmetries are more important for low-luminosities and radio galaxies. 2. Analysis of the correlations of the fractional separation difference with radio luminosity and linear size shows, that (a) there is a trend for the expansion speeds of the lobes and intrinsic/environmental asymmetries to increase with increasing radio luminosity, and (b) intrinsic/environmental asymmetries are more important on small physical scales but they become less significant on larger physical scales. 3. The mean lobe speed of FRII sources and its standard deviation, $`\overline{v}_\mathrm{l}(0.11\pm 0.013)c`$ and $`\sigma _{v_\mathrm{l}}=0.04`$, have been determined from numerical simulations of the observed distribution function of fractional separation differences for both radio galaxies and quasars. These speeds are consistent with ages determined by synchrotron ageing arguments. 4. The asymmetric relativistic model is in agreement with an orientation-based unified scheme in which the critical angle separating radio galaxies from quasars is about $`45^{}`$. The tendency for the intrinsic/environmental asymmetry effects to decrease at low-luminosities and large physical sizes is a direct reflection of the negative correlation between luminosity and linear size. The decrease of intrinsic asymmetries on large physical scales can be naturally explained if the intergalactic gas is more uniform at large distances from the active nuclei. However, an important question is whether environmental effects are responsible for all the observed asymmetries, or whether intrinsic jet asymmetries also need to be taken into account. Some observations suggest that intrinsically the jets may well be symmetric. In the Compact Symmetric Objects (CSO), two hot-spots are located symmetrically with respect to the active galactic nuclei and these are thought to be formed by symmetric jets. Readhead et al. (1996) and Conway & Owsianik (1998) have shown that CSOs are very young objects with ages of only 300 – 3000 years and represent a very early phase in the evolution of classical double FRII radio sources. Saripalli et al. (1997) have detected two parsec scale jets in the giant double radio source DA240 with a flux density ratio of unity. This is the clearest instance of parsec scale jets of equal strengths in FRII sources. McCarthy et al. (1991) quantified the degree of asymmetry in the extended emission-line regions by the parameter $`R_{\text{[OII]}}`$, which is the ratio of the emission-line brightness on either side of the nucleus. They showed that the brighter emission-line region lies on the short lobe side in essentially all cases and considered this as evidence that environmental effects determine the fractional separation differences. In this case, the large emission-line asymmetry parameters should correspond on average to those sources in which the intrinsic/environmental asymmetries dominate, that is, those FRII sources with $`x_\mathrm{l}<0`$ and $`x_\mathrm{l}0.26`$. For the 21 FRII sources which are common to both the sample of McCarthy et al. (1991) and our own, the fractional separation differences are plotted against the emission-line asymmetry parameter in Fig. 10. Excluding the source 3C244.1 which has an extremely large value of $`R_{\text{[OII]}}>20`$, the mean emission-line asymmetry parameters, are found to be $`>4.4`$, $`<2.5`$ and $`>3.9`$, for three parameter regions $`x_\mathrm{l}<0`$, $`0<x_\mathrm{l}0.26`$ and $`x_\mathrm{l}>0.26`$, respectively. It is noteworthy that the mean values of $`R_{\text{[OII]}}`$ in the first and third regions are larger than in the middle one, which illustrates both the important rôle of environmental asymmetries and the validity of the asymmetric relativistic model. The magnitude of the intrinsic/environmental asymmetry is quite small relative to the true mean space velocities of advance of the lobes. Adopting the standard deviation of 0.023$`c`$ for $`\sigma _{v_\mathrm{d}}`$, this amounts to about 20% of the mean advance speed of the lobes. This is not particularly unexpected since the remarkable feature of the FRII sources is the fact that they are as symmetric as they are. This small fractional velocity difference occurs, despite the large differences in the \[OII\] emission on either side of the radio source. The particle densities in the emission line regions should be $`10^2`$ cm<sup>-3</sup>, compared with typical densities of $`10^2`$ cm<sup>-3</sup> for the interstellar and intergalactic environments of FRII radio sources. If the power supplied by the radio jet is assumed to be constant, the rate of advance of the front edge of the radio lobe is given by the balance of the ram pressure of the jet and internal pressure of the hot-spots $`p\rho v^2`$. Keeping $`p`$ constant, it can be seen that the rate of advance of the jet would be much less if it had to penetrate clouds with densities $`10^4`$ times greater than the intervening medium. The answer almost certainly lies in the fact that the emission line gas has a very small filling factor, $`10^6`$, and so there is only a small, but significant, probability of the jet encountering a region of high density gas. Detailed modelling needs to be carried out to investigate the implications of the small value of $`\sigma _{v_\mathrm{d}}/v_0`$ for the properties of the \[OII\] clouds. ### Acknowledgements We are very grateful to Dr. Julia Riley for valuable discussions and for kindly supplying up-to-date information about 3CRR sample, and to Drs. Philip Best and Peter Scheuer for valuable comments. We are also very grateful to Dr. Simon Garrington for careful reading of the text and helpful comments, which have significantly improved this paper. TGA gratefully acknowledges the award of a Royal Society-NATO Post-doctoral Fellowship and a Royal Society ex-quota award. ### APPENDIX: DETAIL AND DESCRIPTION OF THE SAMPLE Information about the jet-side was available from the literature for 103 FRII radio sources from the 3CRR complete sample defined by Laing et al. (1983). All 103 radio sources were of high radio luminosity $`P_{178}2.0\times 10^{26}\mathrm{W}\mathrm{Hz}^1`$. The structural asymmetry of the sources was defined by the *fractional separation difference* of lobes $`x_\mathrm{l}=(r_\mathrm{j}r_{\mathrm{cj}})/(r_\mathrm{j}+r_{\mathrm{cj}})`$, where $`r_\mathrm{j}`$ and $`r_{\mathrm{cj}}`$ are the angular distances from the core to the furthest end of the lobes on the jet and counterjet sides respectively. The fractional separation difference of hot-spots $`v_\mathrm{h}`$ is the ratio of the difference and the sum of the of the core–hot-spots angular distances on the jet and counterjet sides respectively. The value $`x_\mathrm{l}`$ is negative if a jet (or approaching) side is shorter than counterjet side and is positive if jet side is longer. For the most of the FRII sources, values of fractional separation difference of hot-spots given in Table 6 are in agreement with the values published by Best et al. (1995) and McCarthy et al. (1991) within about 5%. Significant discrepancies for 13 sources (3C: 9, 13, 153, 175.1, 184, 205, 217, 226, 236, 324, 352, 368, 432) are generally the result of new deep VLA observations providing improved estimates of $`x_\mathrm{h}`$. The core–lobe angular distances for more than half the FRII sources are taken from the published data of H2, BHL, FBB, FBP and PRM (see Table 6). For the remaining sources, we have measured the distances from the radio images following the procedure of Scheuer (1995): in our analysis, angular distances from the core have been measured to the second or third contour of the radio isophots rather than the lowest. There are uncertainties in the core–lobe angular distances for 3 objects: 3C171, 3C215 and 3C309.1. The last two objects have a strong bent lobe structures. We included these three sources in the final sample measuring the fifth lowest contour. The data are presented in the Table 6 in the following format: IAU name. Common name. The jet or bright jet side. Presence of jet evaluated from VLA/VLBI jet detections (J), depolarization (DM) and spectral index (SI) asymmetries. Definite (D) or probable (P) jets. Fractional separation difference of hot-spots $`x_\mathrm{h}`$. Fractional separation difference of the lobes, including an indication of whether the jet lobe size is larger ($`x_\mathrm{l}0`$) or smaller ($`x_\mathrm{l}0`$) than counterjet lobe size. Identification of the active galaxy with radio galaxy (G) or quasar (Q). Emission class of the active galaxy nuclei: narrow-line (N), broad-line (B) and low-excitation (L) radio galaxies. Redshift of source. Logarithm of the radio luminosity at $`178\mathrm{MHz}`$. Linear sizes in kiloparsec (kpc), calculated from the data of Laing & Riley (personal communication). First and second references indicate the jet-side and radio maps of the FRII sources which were selected to calculate the fractional separation difference of lobes and hot-spots, respectively. Single reference means that the jet-side and radio maps were taken from one source.
warning/0001/hep-th0001115.html
ar5iv
text
# Fourth Order Theories Without Ghosts*footnote **footnote *hep-th/0001115, January 19, 2000 ## Abstract Using the Dirac constraint method we show that the pure fourth-order Pais-Uhlenbeck oscillator model is free of observable negative norm states. Even though such ghosts do appear when the fourth order theory is coupled to a second order one, the limit in which the second order action is switched off is found to be a highly singular one in which these states move off shell. Given this result, construction of a fully unitary, renormalizable, gravitational theory based on a purely fourth order action in 4 dimensions now appears feasible. As a theory conformal gravity is somewhat enigmatic. With it being based on a local invariance (viz. $`g_{\mu \nu }(x)e^{2\alpha (x)}g_{\mu \nu }(x)`$) and with it possessing a gravitational action (viz. $`I_W=\alpha _gd^4x(g)^{1/2}C_{\lambda \mu \nu \kappa }C^{\lambda \mu \nu \kappa }`$ where $`C^{\lambda \mu \nu \kappa }`$ is the conformal Weyl tensor) whose coupling constant $`\alpha _g`$ is dimensionless, it has a structure which is remarkably similar to that which underlies the other fundamental interactions. Moreover, not only is the conformal gravitational theory power counting renormalizable in four spacetime dimensions, as a quantum theory it turns out to even be asymptotically free , while as a classical theory it has the confining $`V(r)=\beta /r+\gamma r/2`$ as its exact classical potential, to thus reproduce some of the most desirable properties of non-Abelian gauge theories. And further, through use of this linear potential the theory has been found capable of accounting for the perplexing galactic rotation curves without the need to introduce any galactic dark matter, with the cosmology associated with the theory being found to be free of any flatness, horizon, universe age or cosmological dark matter problem, with it being a cosmology which naturally (i.e. without fine tuning) explains the recently detected cosmic acceleration, and which, through its underlying scale invariance, naturally solves the cosmological constant problem. Despite all of these attractive features, standing on the debit side is a potential ghost problem which could render the theory non-unitary. Specifically, being a fourth order theory it has a $`D(k^2)=1/k^4`$ propagator, a propagator which by being rewritten as the $`M^20`$ limit of $`D(k^2,M^2)=[1/k^21/(k^2+M^2)]/M^2`$ can immediately be seen as having altogether better ultraviolet behavior than the familiar $`1/k^2`$ propagator of the standard (non-renormalizable) second order gravitational theory, but which achieves this at the price of a possible negative norm ghost state. Prior to actually taking the $`M^20`$ limit, the $`D(k^2,M^2)`$ propagator may be associated with the wave equation $`\mathrm{}(\mathrm{}M^2)\varphi =0`$ ($`\mathrm{}=^\mu _\mu `$), and is thus associated not with a pure fourth order $`\mathrm{}^2\varphi =0`$ theory, but rather with a theory which contains both fourth and second order gravitational components; with it actually being this hybrid theory which is in fact the one which has explicitly been shown to have a ghost. However, since the $`M^20`$ limit of $`D(k^2,M^2)`$ is a singular one, one cannot immediately conclude that the pure fourth order theory itself necessarily also has a ghost, and in fact we shall show (within the simplified fourth order model first considered by Pais and Uhlenbeck ) that in this limit the ghost states then actually decouple from the physical states and do not then lead to a non-unitary S-matrix. With a $`g_{\mu \nu }=\eta _{\mu \nu }+h_{\mu \nu }`$ linearization around flat spacetime reducing the general conformal gravity rank two gravitational tensor $`(g)^{1/2}\delta I_W/\delta g_{\mu \nu }=2\alpha _gW^{\mu \nu }=T^{\mu \nu }/2`$ to $`W^{\mu \nu }=\mathrm{\Pi }^{\mu \rho }\mathrm{\Pi }^{\nu \sigma }K_{\rho \sigma }/2\mathrm{\Pi }^{\mu \nu }\mathrm{\Pi }^{\rho \sigma }K_{\rho \sigma }/6`$ where $`K^{\mu \nu }=h^{\mu \nu }\eta ^{\mu \nu }h_\alpha ^\alpha /4`$ and $`\mathrm{\Pi }^{\mu \nu }=\eta ^{\mu \nu }^\alpha _\alpha ^\mu ^\nu `$, we find that in the conformal gaugeViz. a gauge condition which is left invariant under $`g_{\mu \nu }(x)e^{2\alpha (x)}g_{\mu \nu }(x)`$. $`_\nu g^{\mu \nu }g^{\mu \sigma }g_{\nu \rho }_\sigma g^{\nu \rho }/4=0`$ the free space gravitational wave equation reduces to $`\alpha _g\mathrm{}^2K^{\mu \nu }=0`$, a pure fourth order wave equation in which the tensor indices nicely decouple.With these linearized equations being found to be diagonal in the independent degrees of freedom, we note in passing that conformal gravity would thus appear to have a good initial value problem. With this linearized wave equation having solutions of the form $`[a_{\mu \nu }+b_{\mu \nu }(nx)]e^{ikx}`$ where $`n_\mu =(1,0,0,0)`$, we see that the key aspect of this linearization is the presence of runaway solutions which grow linearly in time. To explore the implications of such temporal runaways great simplification is achieved if we ignore both the spatial dependence and the tensor indices and study the dynamics associated with the embryonic Pais and Uhlenbeck fourth order Lagrangian $$L=\gamma \ddot{q}^2/2\gamma (\omega _1^2+\omega _2^2)\dot{q}^2/2+\gamma \omega _1^2\omega _2^2q^2/2$$ (1) instead. With this Lagrangian having equation of motion $$d^4q/dt^4+(\omega _1^2+\omega _2^2)\ddot{q}+\omega _1^2\omega _2^2q=0,$$ (2) we see that for unequal frequencies (where the solution is $`q(t)=a_1e^{i\omega _1t}+a_2e^{i\omega _2t}`$) the theory is analogous to coupled fourth and second order gravitational theories, while for equal ($`\omega =\omega _1=\omega _2`$) frequencies (where the solution is $`q(t)=c_1e^{i\omega t}+c_2te^{i\omega t}`$) the theory is analogous to a pure fourth order theory all on its own. We shall thus quantize the theory associated with the unequal frequency Eq. (1) and then study what happens when the equal frequency limit is taken. Since the Lagrangian of Eq. (1) contains higher order time derivatives we will have to quantize it using the method of Dirac constraints . To bring Eq. (1) into a form appropriate for canonical quantization we introduce a new variable $`x(t)`$ and a Lagrange multiplier $`\lambda (t)`$ according to $$L=\gamma \dot{x}^2/2\gamma (\omega _1^2+\omega _2^2)x^2/2+\gamma \omega _1^2\omega _2^2q^2/2+\lambda (\dot{q}x).$$ (3) With this new Lagrangian possessing three coordinate variables, we must introduce three canonical momenta, which we find to be given by $`p_x=L/\dot{x}=\gamma \dot{x}`$, $`p_q=L/\dot{q}=\lambda `$ and $`p_\lambda =L/\dot{\lambda }=0`$. Thus we set $`p_q=\gamma \dot{x}`$ and introduce two primary constraint functions $`\varphi _1=p_q\lambda `$, $`\varphi _2=p_\lambda `$, and replace the canonical Hamiltonian $`H_c=p_x\dot{x}+p_q\dot{q}+p_\lambda \dot{\lambda }L`$ by $`H_1=H_c+u_1\varphi _1+u_2\varphi _2`$ where $$H_1=p_x^2/2\gamma +\gamma (\omega _1^2+\omega _2^2)x^2/2\gamma \omega _1^2\omega _2^2q^2/2+\lambda x+u_1(p_q\lambda )+u_2p_\lambda .$$ (4) On introducing the canonical equal time Poisson brackets $`\{x,p_x\}=\{q,p_q\}=\{\lambda ,p_\lambda \}=1`$, we find that both of the quantities $`\{\varphi _1,H_1\}=\gamma \omega _1^2\omega _2^2qu_2+\varphi _1\{\varphi _1,u_1\}+\varphi _2\{\varphi _1,u_2\}`$ and $`\{\varphi _2,H_1\}=x+u_1+\varphi _1\{\varphi _2,u_1\}+\varphi _2\{\varphi _2,u_2\}`$ will vanish weakly (in the sense of Dirac) if we set $`u_1=x`$, $`u_2=\gamma \omega _1^2\omega _2^2q`$, thus enabling us to define a new Hamiltonian $$H_2=p_x^2/2\gamma +\gamma (\omega _1^2+\omega _2^2)x^2/2\gamma \omega _1^2\omega _2^2q^2/2+p_qx+\gamma \omega _1^2\omega _2^2qp_\lambda .$$ (5) For this Hamiltonian $`\{\varphi _2,H_2\}=\{p_\lambda ,H_2\}`$ is found to be zero, but $`\{\varphi _1,H_2\}`$ is found to take the non-zero value $`\gamma \omega _1^2\omega _2^2p_\lambda `$. Hence, finally, if we now set $`p_\lambda =0`$, the resulting algebra associated with the four-dimensional $`q,p_q,x,p_x`$ sector of the theory will now (as befits a fourth order theory) be closed under commutation, with the requisite classical Hamiltonian then being given by $$H=p_x^2/2\gamma +p_qx+\gamma (\omega _1^2+\omega _2^2)x^2/2\gamma \omega _1^2\omega _2^2q^2/2,$$ (6) and with the requisite equal time Poisson bracket relations which define the classical theory being given by $`\{x,p_x\}=1,\{q,p_q\}=1,\{x,H\}=p_x/\gamma ,`$ (7) $`\{q,H\}=x,\{p_x,H\}=p_q\gamma (\omega _1^2+\omega _2^2)x,\{p_q,H\}=\gamma \omega _1^2\omega _2^2q.`$ (8) The canonical equations of motion thus take the form $$\dot{x}=p_x/\gamma ,\dot{q}=x,\dot{p}_x=p_q\gamma (\omega _1^2+\omega _2^2)x,\dot{p}_q=\gamma \omega _1^2\omega _2^2q,$$ (9) to then enable us to both recover Eq. (2) and make the identification $`x=\dot{q}`$ in the solution, with the Legendre transform $`L=p_x\dot{x}+p_q\dot{q}H`$ reducing to Eq. (1) upon imposition of these equations of motion. The Hamiltonian $`H`$ of Eq. (6) is thus the requisite one for the fourth order theory. In constructing the quantum theory associated with Eq. (6) it is important to distinguish between the general Hamiltonian $`H`$ as defined by Eq. (6) and the particular value $`H_{stat}`$ that it takes in the stationary path in which the equations of motion of Eq. (9) are imposed, with the great virtue of the Dirac procedure being that it allows us to define a Hamiltonian $`H`$ which takes a meaning (as the canonical generator used in Eq. (8)) even for non-stationary field configurations (i.e. even for configurations for which $`p_x\dot{x}+p_q\dot{q}H`$ does not reduce to the Lagrangian of Eq. (1)). Thus, for instance, it is the Hamiltonian of Eq. (6) which defines the appropriate phase space for the problem, so that the path integral for the theory is then uniquely given by $`[dq][dp_q][dx][dp_x]exp[i𝑑t(p_x\dot{x}+p_q\dot{q}H)]`$ as integrated over a complete set of classical paths associated with these four independent coordinates and momenta. As regards the stationary value that $`H_{stat}`$ takes when the equations of motion are imposed, viz. $$H_{stat}=\gamma \ddot{q}^2/2\gamma (\omega _1^2+\omega _2^2)\dot{q}^2/2\gamma \omega _1^2\omega _2^2q^2/2\gamma \dot{q}d^3q/dt^3,$$ (10) we note that not only is $`H_{stat}`$ time independent (even in the runaway solution), it is also recognized as being the Ostrogradski generalized higher derivative Hamiltonian (viz. $`H_{ost}=\dot{q}L/\dot{q}+\ddot{q}L/\ddot{q}\dot{q}d(L/\ddot{q})/dtL`$) associated with the Lagrangian of Eq. (1) as well as being its associated time translation generator . With the Hamiltonian of Eq. (6) being defined for both stationary and non-stationary classical paths, a canonical quantization of the theory can readily be obtained by replacing ($`i`$ times) the Poisson brackets of Eq. (8) by canonical equal time commutators. However, without reference to the explicit structure of the Hamiltonian itself, we note first that the identification (as suggested but not required by the equations of motion<sup>§</sup><sup>§</sup>§In general, equal time commutators such as $`[q(t),p_q(t)]=i`$ can be satisfied at all times by an $`[a,a^{}]=1`$ commutator defined via $`q(t)=af(t)/2^{1/2}+H.c.`$, $`p_q(t)=ia^{}/2^{1/2}f(t)+H.c.`$ with the function $`f(t)`$ being arbitrary. The occupation number Fock space can thus be defined independent of the structure of $`H`$, and even independent of any interaction terms that might also be added on to the free particle $`H`$. Thus in general, and as we shall explicitly see below in particular, the dimensionality of the Fock space basis need not be the same as that of the eigenspectrum of $`H`$.) $`q(t)=a_1e^{i\omega _1t}+a_2e^{i\omega _2t}+H.c.,p_q(t)=i\gamma \omega _1\omega _2^2a_1e^{i\omega _1t}+i\gamma \omega _1^2\omega _2a_2e^{i\omega _2t}+H.c.,`$ (11) $`x(t)=i\omega _1a_1e^{i\omega _1t}i\omega _2a_2e^{i\omega _2t}+H.c.,p_x(t)=\gamma \omega _1^2a_1e^{i\omega _1t}\gamma \omega _2^2a_2e^{i\omega _2t}+H.c.`$ (12) then furnishes us with a Fock space representation of the quantum-mechanical commutation relations $`[x,p_x]=[q,p_q]=i`$, $`[x,q]=[x,p_q]=[q,p_x]=[p_x,p_q]=0`$ at all times provided that $$[a_1,a_1^{}]=[2\gamma \omega _1(\omega _1^2\omega _2^2)]^1,[a_2,a_2^{}]=[2\gamma \omega _2(\omega _2^2\omega _1^2)]^1,[a_1,a_2^{}]=0,[a_1,a_2]=0.$$ (13) Then in this convenient Fock representation the quantum-mechanical Hamiltonian is found to take the form $$H=2\gamma (\omega _1^2\omega _2^2)(\omega _1^2a_1^{}a_1\omega _2^2a_2^{}a_2)+(\omega _1+\omega _2)/2$$ (14) with its associated commutators as inferred from Eq. (8) then automatically being satisfied. With the quantity $`\gamma (\omega _1^2\omega _2^2)`$ being taken to be positive for definitiveness, we see that the $`[a_2,a_2^{}]`$ commutator is negative definite, and with $`H`$ being diagonal in the $`a_1,a_2`$ occupation number basis, we see that the state defined by $`a_1|\mathrm{\Omega }=0`$, $`a_2|\mathrm{\Omega }=0`$ is the ground state of $`H`$,Identifying $`a^{}`$ as the annihilator of the Fock vacuum would eliminate negative norm states but would leave $`H`$ without any lower bound on the ground state energy. that the states $`|+1=[2\gamma \omega _1(\omega _1^2\omega _2^2)]^{1/2}a_1^{}|\mathrm{\Omega }`$, $`|1=[2\gamma \omega _2(\omega _1^2\omega _2^2)]^{1/2}a_2^{}|\mathrm{\Omega }`$ are both positive energy eigenstates with respective energies $`\omega _1`$ and $`\omega _2`$ above the ground state, that the state $`|+1`$ has a norm equal to plus one, but that the state $`|1`$ has norm minus one, a ghost state. With the eigenstates of $`H`$ labeling the asymptotic states associated with scattering in the presence of an interaction Lagrangian $`L_I`$ which might be added on to the original Lagrangian $`L`$ of the theory, the action of $`L_I`$ would induce transitions between asymptotic in and out states of opposite norm, with the unequal frequency theory thus explicitly being seen to be non-unitary.While it is widely believed that the quantum theory associated with Eq. (1) is not unitary, it is possible that our demonstration here might be the first explicit constructive proof. However, even though the Hamiltonian of the unequal frequency theory does have ghost eigenstates, since the commutation relations given in Eq. (13) become singular in the equal frequency limit while both the Hamiltonian of Eq. (14) and the normalized $`|+1`$ and $`|1`$ states develop zeroes, we will have to exercise some caution in trying to discover exactly what happens to the ghost states when this limit is taken. And in fact, we will find below that there are actually two possible ways to take this limit, ways which will lead to differing energy eigenspectra even as they lead to the same commutator algebra, with both ways being found to lead to a unitary theory. To explore the limit it is convenient to introduce new Fock space variables according to $`a_1+a_2=ab`$, $`a_1a_2=2b\omega /ϵ`$ where $`\omega =(\omega _1+\omega _2)/2`$ and $`ϵ=(\omega _1\omega _2)/2`$. These variables are found to obey commutation relations $`[a,a^{}]=\lambda `$, $`[a,b^{}]=\mu `$, $`[b,b^{}]=\nu `$, where $$\lambda =ϵ^2/16\gamma \omega _1\omega _2\omega ^3,\mu =(2\omega ^2ϵ^2)/16\gamma \omega _1\omega _2\omega ^3,\nu =ϵ^2/16\gamma \omega _1\omega _2\omega ^3.$$ (15) In terms of these variables $`q(t)`$ of Eq. (12) can be reexpressed as $$q(t)=e^{i\omega t}[(ab)cosϵt2ib\omega ϵ^1sinϵt]+H.c.,$$ (16) and thus has a well defined $`ϵ0`$ limit, viz. $$q(t)=e^{i\omega t}(ab2ib\omega t)+H.c.,$$ (17) a limit in which the Hamiltonian of Eq. (14) takes the form $$H(ϵ=0)=8\gamma \omega ^4(2b^{}b+a^{}b+b^{}a)+\omega ,$$ (18) and in which the following commutators of interest take the form $`[H(ϵ=0),a^{}]=\omega (a^{}+2b^{}),[H(ϵ=0),b^{}]=\omega b^{},`$ (19) $`[a+b,a^{}+b^{}]=2\mu ,[ab,a^{}b^{}]=2\mu ,[a+b,a^{}b^{}]=0.`$ (20) Thus we see that use of the $`a^{},b^{}`$ operators enables us to construct an $`ϵ=0`$ theory which is well defined, with the $`ϵ=0`$ Fock space then being built on the Fock vacuum $`|\mathrm{\Omega }`$ defined by $`a|\mathrm{\Omega }=b|\mathrm{\Omega }=0`$ (we need to use this basis since states built on the unequal frequency Fock vacuum become undefined in the $`ϵ=0`$ limit). Moreover, in the $`ϵ=0`$ limit we see that the $`[a,a^{}]`$ and $`[b,b^{}]`$ commutators both vanish, with, as we shall see below, there actually being two ways rather than one in which the Fock space states can implement this vanishing. However, before constructing these two ways, we note that because the action of the Hamiltonian is to shift $`a^{}`$ by $`2b^{}`$, and because (unlike the unequal frequency case) neither of the operators ($`a^{}\pm b^{}`$) which diagonalize the Fock space basis acts as a ladder operator for the Hamiltonian, we can anticipate that the dimensionality of the eigenspectrum of $`H(ϵ=0)`$ will be lower than that of $`[x,p_x]=[q,p_q]=i`$, $`[x,q]=[x,p_q]=[q,p_x]=[p_x,p_q]=0`$ commutator algebra, i.e. lower than the dimensionality of the two-dimensional harmonic oscillator, even though that precisely was the dimensionality of the unequal frequency eigenspectrum. Because $`b^{}`$ does act as a ladder operator for $`H(ϵ=0)`$, it is possible to construct a one particle energy eigenstate, viz. the state $`b^{}|\mathrm{\Omega }`$, but if we try to normalize it (our first way to realize the $`ϵ=0`$ Fock space) the vanishing of the $`[b,b^{}]`$ commutator would entail that this state would have to have zero norm. On its own a zero norm state (unlike a negative norm state) does not lead to loss of probability, but its very existence entails the existence of negative norm states elsewhere in the theory.<sup>\**</sup><sup>\**</sup>\**The unequal frequency theory states $`|+1\pm |1`$ both have zero norm. Thus, with the state $`a^{}|\mathrm{\Omega }`$ also having zero norm, we immediately construct the states $`|\pm =(a^{}\pm b^{})|\mathrm{\Omega }/(2\mu )^{1/2}`$, states which obey $`+|+=1`$, $`|=1`$, $`+|=0`$. However, unlike the orthogonal positive and negative norm states $`|\pm 1`$ associated with the unequal frequency case, this time we find that neither of the $`|\pm `$ states is an eigenstate of the Hamiltonian ($`H(ϵ=0)|\pm =2\omega (|\pm +2b^{}|\mathrm{\Omega }`$)). Thus even while the $`ϵ=0`$ Fock space possesses negative norm ghost states, this time they can only exist off shell (where they can still regulate Feynman diagrams) but cannot materialize as on shell asymptotic in and out states. With a similar situation being found in the two particle sector<sup>††</sup><sup>††</sup>††The zero norm states $`(b^{})^2|\mathrm{\Omega }`$ and $`(a^{})^2|\mathrm{\Omega }`$ can be combined into positive and negative norm states while the state $`a^{}b^{}|\mathrm{\Omega }/\mu `$ has norm plus one. where only $`(b^{})^2|\mathrm{\Omega }`$ is an energy eigenstate, we see that the complete spectrum of eigenstates of $`H(ϵ=0)`$ is the set of all states $`(b^{})^n|\mathrm{\Omega }`$, a spectrum whose dimensionality is that of a one rather than a two-dimensional harmonic oscillator, even while the complete Fock space has the dimensionality of the two-dimensional oscillator. The equal frequency fourth order theory thus has a propagator which acts not like a $`D(k^2,M^2)=[1/k^21/(k^2+M^2)]/M^2`$ propagator with poles at $`k_0=\pm k,k_0=\pm (k^2+M^2)^{1/2}`$, but rather as the derivative operator $`D(k^2)=d(k^2)/dk^2)`$, i.e. as the derivative of a propagator which only has poles at $`k_0=\pm k`$.<sup>‡‡</sup><sup>‡‡</sup>‡‡A further unusual property of the equal frequency theory is that the one particle matrix element $`\mathrm{\Omega }|q(t)b^{}|\mathrm{\Omega }`$ is given by $`\mu e^{i\omega t}`$ rather than by the coefficient of the $`b`$ field operator in Eq. (17), with there thus being a mismatch between the second quantized states and the first quantized wave functions. Then finally, with the free and interacting Fock spaces being the same (the Fock space can be defined once and for all by the $`[x,p_x]=[q,p_q]=i`$, $`[x,q]=[x,p_q]=[q,p_x]=[p_x,p_q]=0`$ commutator algebra) and with $`(b^{})^n|\mathrm{\Omega }`$ subset of Fock states serving to label the asymptotic in and out energy eigenstates, we see that the equal frequency theory is unitary. While the $`ϵ=0`$ theory is thus seen to be unitary, it is nonetheless a somewhat unusual theory in that all of its energy eigenstates have zero norm. Thus it would be of interest to see if we could construct an alternate limiting theory in which the energy eigenstates have positive norm instead. To this end we note that since we obtained the zero norm states by first going to the $`ϵ=0`$ limit and then constructing the states, an alternate procedure would be to first construct normalized states and then take the limit. With the algebra of the $`a`$ and $`b`$ operators already being defined prior to taking the limit, we can thus construct a basis for the $`ϵ0`$ Fock space via states such as (the choice $`\gamma <0`$ assures the positivity of $`\lambda `$ and $`\nu `$) the one particle $`|1,0=a^{}|\mathrm{\Omega }/\lambda ^{1/2}`$ and $`|0,1=b^{}|\mathrm{\Omega }/\nu ^{1/2}`$, the two particle $`|2,0=(a^{})^2|\mathrm{\Omega }/2^{1/2}\lambda `$, $`|1,1=a^{}b^{}|\mathrm{\Omega }/(\mu ^2+\lambda \nu )^{1/2}`$ and $`|0,2=(b^{})^2|\mathrm{\Omega }/2^{1/2}\nu `$, and so on, a complete basis in which, remarkably, every single state is found to have positive norm, but not an orthogonal one since overlaps such as $`1,0|0,1=\mu /(\lambda \nu )^{1/2}`$ are non-zero, overlaps which actually become singular in the $`ϵ0`$ limit. Since this basis is complete we can use it to define a limiting procedure in which each of the basis states is held fixed while $`ϵ`$ is allowed to go to zero. Then in such a limit we find that $`a^{}|\mathrm{\Omega }`$ and $`b^{}|\mathrm{\Omega }`$ both become null vectors. However, despite this, we cannot conclude that the creation operators annihilate the vacuum identically in this limit since matrix elements such as $`\mathrm{\Omega }|[a,b^{}]|\mathrm{\Omega }=\mu `$ do not vanish in the limit, with product operator actions such as $`a`$ acting on $`b^{}|\mathrm{\Omega }`$ being singular. Instead, the creation operators must be thought of as annihilating the vacuum weakly (i.e. in some but not all matrix elements), but not strongly as an operator identity. Now since neither of the states $`a^{}|\mathrm{\Omega }`$ or $`b^{}|\mathrm{\Omega }`$ survives in the limit, $`H(ϵ=0)`$ now has no one particle eigenstates at all. With the two particle states $`(a^{})^2|\mathrm{\Omega }`$ and $`(b^{})^2|\mathrm{\Omega }`$ also becoming null in the limit, the only two particle state which is found to survive in this limit is the positive norm state $`a^{}b^{}|\mathrm{\Omega }/\mu `$ (since the $`[a,b^{}]`$ commutator does not vanish), and, quite remarkably, in this limit the state is also found to actually become an energy eigenstate ($`H(ϵ=0)a^{}b^{}|\mathrm{\Omega }/\mu =3\omega a^{}b^{}|\mathrm{\Omega }/\mu +2\omega (b^{})^2|\mathrm{\Omega }/\mu 3\omega a^{}b^{}|\mathrm{\Omega }/\mu `$). With this analysis immediately generalizing to the higher multiparticle states as well, we see that the only states which then survive in the limit are states of the form $`(a^{}b^{})^n|\mathrm{\Omega }`$, positive norm states which also become energy eigenstates in the limit, with the observable sector of the theory again being unitary, and with it again having the dimensionality of a one-dimensional harmonic oscillator and not that of a two-dimensional one.<sup>\**</sup><sup>\**</sup>\**In the conformal gravity case such a situation would correspond to there only being one observable graviton rather than two, with the observable graviton being composite rather than fundamental. We thus recognize two limiting procedures, first taking the limit of the algebra and then constructing the Hilbert space, or first constructing the Hilbert space and then taking the limit of the algebra, with this latter limiting procedure being an extremely delicate one in which the only states which survive as observable ones are composite.<sup>\*†</sup><sup>\*†</sup>\*†A model in which a field has a positive frequency part which annihilates the vacuum strongly and a negative frequency part which annihilates the vacuum weakly in a way such that only certain multiparticle states survive would appear to be a possible candidate mechanism for quark confinement, with quarks themselves then only existing off shell. Thus, to conclude, we see that in the equal frequency limit the quantum theory based on the fourth order Lagrangian of Eq. (1) is in fact unitary (in fact, technically, what we have shown is that the lack of unitarity of the unequal frequency theory is simply not a reliable indicator as to the unitarity status of the equal frequency one), and that construction of a fully unitary and renormalizable quantum gravitational theory in four spacetime dimensions would now appear feasible. One of the authors (P.D.M) wishes to thank Dr. E. E. Flanagan for useful discussions. The work of P.D.M. has been supported in part by the Department of Energy under grant No. DE-FG02-92ER4071400.
warning/0001/hep-ph0001191.html
ar5iv
text
# References Recently, Pineda and Ynduráin presented a re-analysis of heavy quarkonia. Their investigation is based on the main assumption that bound systems of heavy quarks may be reasonably described by nonrelativistic kinematics and only the perturbative contribution to the quark–antiquark interaction potential $`V`$ if all nonperturbative effects are taken into account by some appropriate correction to the energy. In order to describe a system of a heavy quark and antiquark, both with constituent mass $`m`$, forming a bound state with total spin $`s=0`$ or $`s=1`$, Pineda and Ynduráin consider the Hamiltonian $`H`$ $`=`$ $`2m{\displaystyle \frac{1}{m}}\mathrm{\Delta }{\displaystyle \frac{1}{4m^3}}\mathrm{\Delta }^2+V_0^\mathrm{P}(r)+{\displaystyle \frac{C_\mathrm{F}\alpha (\mu )}{m^2r}}\mathrm{\Delta }+{\displaystyle \frac{C_\mathrm{F}\alpha ^2(\mu )}{4mr^2}}(C_\mathrm{F}2C_\mathrm{A})`$ (1) $`+`$ $`{\displaystyle \frac{4\pi C_\mathrm{F}\alpha (\mu )}{3m^2}}s(s+1)\delta ^{(3)}(𝐱),r|𝐱|.`$ The perturbative contribution to the static quark–antiquark interaction potential, $`V_0^\mathrm{P}(r),`$ is known up to and including the two-loop level : $`V_0^\mathrm{P}(r)`$ $`=`$ $`C_\mathrm{F}{\displaystyle \frac{\alpha (\mu )}{r}}\{1+{\displaystyle \frac{\alpha (\mu )}{4\pi }}[{\displaystyle \frac{5}{3}}\beta _0{\displaystyle \frac{8}{3}}C_\mathrm{A}+2\beta _0[\mathrm{ln}(\mu r)+\gamma _\mathrm{E}]]`$ $`+`$ $`\left({\displaystyle \frac{\alpha (\mu )}{4\pi }}\right)^2[\beta _0^2(4[\mathrm{ln}(\mu r)+\gamma _\mathrm{E}]^2+{\displaystyle \frac{\pi ^2}{3}})+(2\beta _1+{\displaystyle \frac{20}{3}}\beta _0^2{\displaystyle \frac{32}{3}}\beta _0C_\mathrm{A})[\mathrm{ln}(\mu r)+\gamma _\mathrm{E}]`$ $`+`$ $`\left({\displaystyle \frac{4343}{162}}+4\pi ^2{\displaystyle \frac{\pi ^4}{4}}+{\displaystyle \frac{22}{3}}\zeta (3)\right)C_\mathrm{A}^2\left({\displaystyle \frac{1798}{81}}+{\displaystyle \frac{56}{3}}\zeta (3)\right)n_\mathrm{f}C_\mathrm{A}T_\mathrm{F}`$ $``$ $`({\displaystyle \frac{55}{3}}16\zeta (3))n_\mathrm{f}C_\mathrm{F}T_\mathrm{F}+{\displaystyle \frac{400}{81}}n_\mathrm{f}^2T_\mathrm{F}^2]\}.`$ Here, the following notations have been adopted: $`\alpha (\mu )`$ denotes the strong fine-structure constant in the modified minimal-subtraction ($`\overline{\mathrm{MS}}`$) renormalization scheme. For a non-Abelian gauge theory for $`n_\mathrm{f}`$ Dirac fermions, invariant w. r. t. gauge transformations forming a Lie group SU($`N`$) describing $`N`$ colour degrees of freedom, the quadratic Casimir invariants read, for the fundamental representation, $$C_\mathrm{F}=\frac{N^21}{2N}$$ and, for the adjoint representation, $$C_\mathrm{A}=N,$$ if the generators of the Lie group SU($`N`$) are normalized such that the second-order Dynkin index of the fundamental representation is $$T_\mathrm{F}=\frac{1}{2}.$$ The dependence of the effective (running) fine-structure constant $`\alpha (\mu )`$ on the renormalization scale $`\mu `$ is described in terms of the Gell-Mann–Low $`\beta `$ function according to $$\frac{\mu }{2}\frac{}{\mu }\alpha (\mu )=\frac{\alpha ^2(\mu )}{4\pi }\beta _0\frac{\alpha ^3(\mu )}{(4\pi )^2}\beta _1\frac{\alpha ^4(\mu )}{(4\pi )^3}\beta _2+O(\alpha ^5),$$ involving the well-known expressions for the (gauge-invariant) one-, two-, and three-loop expansion coefficients in the $`\overline{\mathrm{MS}}`$ scheme $`\beta _0`$ $`=`$ $`{\displaystyle \frac{11}{3}}C_\mathrm{A}{\displaystyle \frac{4}{3}}n_\mathrm{f}T_\mathrm{F},`$ $`\beta _1`$ $`=`$ $`{\displaystyle \frac{34}{3}}C_\mathrm{A}^2{\displaystyle \frac{20}{3}}n_\mathrm{f}C_\mathrm{A}T_\mathrm{F}4n_\mathrm{f}C_\mathrm{F}T_\mathrm{F},`$ $`\beta _2`$ $`=`$ $`{\displaystyle \frac{2857}{54}}C_\mathrm{A}^3{\displaystyle \frac{1415}{27}}n_\mathrm{f}C_\mathrm{A}^2T_\mathrm{F}+{\displaystyle \frac{158}{27}}n_\mathrm{f}^2C_\mathrm{A}T_\mathrm{F}^2{\displaystyle \frac{205}{9}}n_\mathrm{f}C_\mathrm{A}C_\mathrm{F}T_\mathrm{F}+{\displaystyle \frac{44}{9}}n_\mathrm{f}^2C_\mathrm{F}T_\mathrm{F}^2+2n_\mathrm{f}C_\mathrm{F}^2T_\mathrm{F}.`$ The resulting dependence of the fine-structure constant $`\alpha (\mu )`$ on the chosen renormalization scale $`\mu `$, expressed in terms of the (standard) scale parameter $`\mathrm{\Lambda }`$, reads up to and including the three-loop level $`\alpha (\mu )`$ $`=`$ $`{\displaystyle \frac{4\pi }{\beta _0\mathrm{ln}(\mu ^2/\mathrm{\Lambda }^2)}}\{1{\displaystyle \frac{\beta _1}{\beta _0^2}}{\displaystyle \frac{\mathrm{ln}(\mathrm{ln}(\mu ^2/\mathrm{\Lambda }^2))}{\mathrm{ln}(\mu ^2/\mathrm{\Lambda }^2)}}`$ $`+`$ $`{\displaystyle \frac{\beta _1^2}{\beta _0^4\mathrm{ln}^2(\mu ^2/\mathrm{\Lambda }^2)}}[(\mathrm{ln}(\mathrm{ln}(\mu ^2/\mathrm{\Lambda }^2)){\displaystyle \frac{1}{2}})^2+{\displaystyle \frac{\beta _2\beta _0}{\beta _1^2}}{\displaystyle \frac{5}{4}}]\}.`$ $`\gamma _\mathrm{E}`$ is known as Euler–Mascheroni constant. In the case of quantum chromodynamics, clearly, $`N=3.`$ Now, in order to stick to an entirely analytical analysis and following the philosophy developed in an earlier treatment of heavy quarkonia, in Ref. the static potential $`V_0^\mathrm{P}(r)`$ is split, according to $$V_0^\mathrm{P}(r)=\stackrel{~}{V}(r)+\widehat{V}(r),$$ (2) into the Coulomb-like contribution $$\stackrel{~}{V}(r)C_\mathrm{F}\frac{\stackrel{~}{\alpha }(\mu )}{r},$$ with the effective fine-structure constant $`\stackrel{~}{\alpha }(\mu )`$ $``$ $`\alpha (\mu )\{1+{\displaystyle \frac{\alpha (\mu )}{4\pi }}({\displaystyle \frac{5}{3}}\beta _0{\displaystyle \frac{8}{3}}C_\mathrm{A}+2\beta _0\gamma _\mathrm{E})`$ $`+`$ $`\left({\displaystyle \frac{\alpha (\mu )}{4\pi }}\right)^2[\beta _0^2(4\gamma _\mathrm{E}^2+{\displaystyle \frac{\pi ^2}{3}})+(2\beta _1+{\displaystyle \frac{20}{3}}\beta _0^2{\displaystyle \frac{32}{3}}\beta _0C_\mathrm{A})\gamma _\mathrm{E}`$ $`+`$ $`\left({\displaystyle \frac{4343}{162}}+4\pi ^2{\displaystyle \frac{\pi ^4}{4}}+{\displaystyle \frac{22}{3}}\zeta (3)\right)C_\mathrm{A}^2\left({\displaystyle \frac{1798}{81}}+{\displaystyle \frac{56}{3}}\zeta (3)\right)n_\mathrm{f}C_\mathrm{A}T_\mathrm{F}`$ $``$ $`({\displaystyle \frac{55}{3}}16\zeta (3))n_\mathrm{f}C_\mathrm{F}T_\mathrm{F}+{\displaystyle \frac{400}{81}}n_\mathrm{f}^2T_\mathrm{F}^2]\},`$ and an obvious remainder involving logarithms of the radial coordinate $`r`$, $`\widehat{V}(r)`$ $`=`$ $`C_\mathrm{F}{\displaystyle \frac{\alpha ^2(\mu )}{4\pi r}}\{2\beta _0\mathrm{ln}(\mu r){\displaystyle \frac{}{}}`$ $`+`$ $`{\displaystyle \frac{\alpha (\mu )}{4\pi }}[4\beta _0^2\mathrm{ln}^2(\mu r)+(2\beta _1+{\displaystyle \frac{20}{3}}\beta _0^2{\displaystyle \frac{32}{3}}\beta _0C_\mathrm{A}+8\beta _0^2\gamma _\mathrm{E})\mathrm{ln}(\mu r)]\}.`$ The eigenvalue problem for the Coulombic Hamiltonian $$\stackrel{~}{H}2m\frac{1}{m}\mathrm{\Delta }+\stackrel{~}{V}(r)$$ is solved exactly, yielding, for instance, for the ground state, the energy eigenvalue $`\stackrel{~}{E}_0=2m+\epsilon _0,`$ with the Coulomb binding energy $$\epsilon _0=\frac{C_\mathrm{F}^2\stackrel{~}{\alpha }^2(\mu )m}{4}.$$ The “non-Coulombic” part $`H\stackrel{~}{H}`$ of the Hamiltonian (1) is treated perturbatively. Counting carefully the powers of $`\alpha `$ yields analytical expressions for the eigenvalues of $`H`$ correct up to and including the order $`\alpha ^4`$. Nonperturbative effects are incorporated by adding the Leutwyler–Voloshin correction , which involves the gluon condensate $`\alpha G^2=0.06\pm 0.02\text{GeV}^4`$. For the ground state, this correction amounts to the energy shift $$\delta E_0=\frac{624}{425}\frac{\pi \alpha G^2m}{(C_\mathrm{F}\stackrel{~}{\alpha }(\mu )m)^4}.$$ By inversion of the expression for the (ground-state) bound-state mass emerging from this procedure, the corresponding quark pole mass $`m`$ is computed. For instance, with $`\mathrm{\Lambda }(n_\mathrm{f}=4)=0.23\begin{array}{c}+0.08\hfill \\ 0.05\hfill \end{array}\text{GeV}`$ and $`\mu =\sqrt{6.632\pm 25\%}\text{GeV},`$ implying $`\alpha (\mu )=0.246`$ and $`\stackrel{~}{\alpha }(\mu )=0.386,`$ the experimental value of the $`\mathrm{{\rm Y}}`$ mass, $`M(\mathrm{{\rm Y}})_{\mathrm{exp}}=9.46037\pm 0.00021\text{GeV},`$ translates into the b quark mass $`m_\mathrm{b}=5.001\begin{array}{c}+0.104\hfill \\ 0.066\hfill \end{array}\text{GeV}`$ . However, a perturbative treatment as implied by the splitting (2) of the static potential $`V_0^\mathrm{P}(r)`$ is by no means mandatory, obligatory, or even desirable. We may also adopt the following point of view. Given the operator $`H`$ defined by Eq. (1) (accurate up to a certain order in $`\alpha `$), compute (numerically, if necessary) its discrete spectrum, i. e., the set of eigenvalues, irrespective of the involved powers of $`\alpha `$. Of course, the terms in Eq. (1) proportional to $$\mathrm{\Delta }^2,+\frac{1}{r}\mathrm{\Delta },$$ and, if the effective coupling strength multiplying this term exceeds some critical value, also the term in Eq. (1) proportional to $$\frac{1}{r^2}$$ render the operator $`H`$ unbounded from below and have therefore to be treated perturbatively anyway. The Hamiltonian $$\widehat{H}2m\frac{1}{m}\mathrm{\Delta }+V_0^\mathrm{P}(r),$$ (3) on the other hand, may certainly be analyzed without adhering to some perturbative approximation. In order to obtain a first idea of the differences brought about by these two approaches, let us start by considering only the Hamiltonian (3). The perturbative calculation is straightforward. Introducing, for notational brevity, the generalized Bohr radius $$a(\mu )\frac{2}{C_\mathrm{F}\stackrel{~}{\alpha }(\mu )m},$$ the expectation values of the non-Coulombic interaction $`\widehat{V}(r)`$ w. r. t. the ground state of $`\stackrel{~}{H}`$ (indicated by the subscript $`\stackrel{~}{0}`$) may be evaluated with the help of the relations (see also Appendix B of Ref. ) $`{\displaystyle \frac{\mathrm{ln}(\mu r)}{r}}_{\stackrel{~}{0}}`$ $`=`$ $`{\displaystyle \frac{1}{a(\mu )}}\left[\mathrm{ln}\left({\displaystyle \frac{\mu a(\mu )}{2}}\right)+1\gamma _\mathrm{E}\right],`$ $`{\displaystyle \frac{\mathrm{ln}^2(\mu r)}{r}}_{\stackrel{~}{0}}`$ $`=`$ $`{\displaystyle \frac{1}{a(\mu )}}\left[\mathrm{ln}^2\left({\displaystyle \frac{\mu a(\mu )}{2}}\right)+2(1\gamma _\mathrm{E})\mathrm{ln}\left({\displaystyle \frac{\mu a(\mu )}{2}}\right)+(1\gamma _\mathrm{E})^2+{\displaystyle \frac{\pi ^2}{6}}1\right].`$ The nonperturbative evaluation of $`\widehat{H}`$ is performed with some numerical procedure<sup>1</sup><sup>1</sup>1 The desired accuracy of the (numerically determined) bound-state energies and wave functions may be adjusted in the routine used for the solution of the Schrödinger equation. For the present analysis, the uncertainty of these energies has been required to be less than $`10^7\text{GeV}=100\text{eV}.`$ This accuracy should be, by far, sufficient for our purposes. developed for the treatment of the nonrelativistic Schrödinger equation . In this way, we find, for the parameter values used in the second of Refs. and focusing our interest to the ground state, for the Coulomb binding energy, $`\epsilon _0=0.33146\text{GeV},`$ for the expectation value of $`\widehat{V}(r)`$, $`\widehat{V}(r)_{\stackrel{~}{0}}=0.13714\text{GeV},`$ and thus, for the perturbatively calculated ground-state energy,<sup>2</sup><sup>2</sup>2 For the perturbative treatment of $`\widehat{H}`$, we truncate the Rayleigh–Schrödinger series for $`\widehat{E}_0^\mathrm{P}`$ at lowest non-trivial order in $`\widehat{V}(r)`$. Inclusion of the next order reduces the observed discrepancies but does not change qualitatively our findings. $`\widehat{E}_0^\mathrm{P}2m+\epsilon _0+\widehat{V}(r)_{\stackrel{~}{0}}=9.5334\text{GeV},`$ while the numerically computed “exact” lowest eigenvalue of the operator (3) is $`\widehat{E}_0^{\mathrm{NP}}\widehat{H}_0=9.5198\text{GeV}.`$ Consequently, for the parameters of Ref. the difference in the lowest bound-state mass predicted by the Hamiltonian (3) within perturbative and nonperturbative approaches is $`\widehat{E}_0^\mathrm{P}\widehat{E}_0^{\mathrm{NP}}=13.6\text{MeV}.`$ Unfortunately, this discrepancy is roughly 65 times larger than the experimental error on the $`\mathrm{{\rm Y}}`$ mass. Phrased the other way round, the perturbative ground-state eigenvalue $`\widehat{E}_0^\mathrm{P}`$ of $`\widehat{H}`$ can be reproduced by the nonperturbative evaluation of $`\widehat{H}`$ for a mass $`m`$ of the bound-state constituents of $`m=5.008\text{GeV}.`$ For the b quark mass of Ref. , the theoretical error different from the one induced by variation of the renormalization scale $`\mu ,`$ attributed to neglected higher-order perturbative as well as nonperturbative corrections, is estimated to be $`\pm 0.006\text{GeV}.`$ Obviously, this error is entirely consumed already by the difference of the masses of the bound-state constituents obtained by perturbative and nonperturbative evaluations of $`\widehat{H}.`$ In view of these findings, let’s try to improve the theoretical value of the b quark pole mass $`m_\mathrm{b}`$ by approaching the part $`\widehat{H}`$ of the Hamiltonian (1) nonperturbatively. The numerical computation of the expectation values of the operator $`H\widehat{H}`$ is considerably facilitated by the following two observations: * The eigenvalue equation for the “toy-model” (Hamiltonian) operator $`\widehat{H}`$ defined in Eq. (3) reads, for some generic (energy) eigenvalue $`\widehat{E}`$ and its corresponding eigenstate $`|\psi `$ of $`\widehat{H}`$, $`\widehat{H}|\psi =\widehat{E}|\psi .`$ Thus the expectation values of all terms in the Hamiltonian (1) involving the Laplacian $`\mathrm{\Delta },`$ taken w. r. t. $`|\psi ,`$ may be evaluated by substituting $`\mathrm{\Delta }`$ according to $`\mathrm{\Delta }|\psi =m\left[2m+V_0^\mathrm{P}(r)\widehat{E}\right]|\psi .`$ * The expectation value of the $`\delta `$ function entering in the “spin–spin term” of the Hamiltonian (1), taken w. r. t. $`|\psi ,`$ is the modulus squared of the corresponding wave function $`\psi (𝐱)`$ at the origin: $`\psi |\delta ^{(3)}(𝐱)|\psi =|\psi (\mathrm{𝟎})|^2.`$ For states with vanishing orbital angular momentum $`\mathrm{}`$ (the so-called “S waves”), $`|\psi (\mathrm{𝟎})|^2`$ may be expressed in terms of the first derivative of the relevant interaction potential $`V(r)`$ w. r. t. the radial coordinate $`r`$ according to (for a derivation, see, e. g., Ref. ) $$|\psi (\mathrm{𝟎})|^2=\frac{m}{4\pi }\psi \left|\frac{\mathrm{d}}{\mathrm{d}r}V(r)\right|\psi .$$ In this way, the lowest eigenvalue of the Hamiltonian $`H`$ may be computed from the expression (where the subscript 0 of the expectation values indicates, as before, the ground state of the Hamiltonian $`\widehat{H}`$) $`E_0H_0`$ $`=`$ $`\widehat{E}_0^{\mathrm{NP}}{\displaystyle \frac{1}{4m^3}}\mathrm{\Delta }^2_0+{\displaystyle \frac{C_\mathrm{F}\alpha (\mu )}{m^2}}{\displaystyle \frac{1}{r}}\mathrm{\Delta }_0+{\displaystyle \frac{C_\mathrm{F}\alpha ^2(\mu )}{4m}}(C_\mathrm{F}2C_\mathrm{A}){\displaystyle \frac{1}{r^2}}_0`$ $`+`$ $`{\displaystyle \frac{4\pi C_\mathrm{F}\alpha (\mu )}{3m^2}}s(s+1)|\psi _0(\mathrm{𝟎})|^2.`$ Adding the nonperturbative shift $`\delta E_0`$ gives our final result for the ground-state energy: $`_0=E_0+\delta E_0.`$ For the parameter values of Ref. and a b quark mass of $`m_\mathrm{b}=5.001\text{GeV},`$ this expression entails the bound-state energy $`_0=9.4953\text{GeV}.`$ Hence, the error of the predicted $`\mathrm{{\rm Y}}`$ mass brought about by the perturbative approximation in the treatment of the Hamiltonian (1), which amounts to $`35\text{MeV},`$ is of the order of the hyperfine splittings in the bottomonium system. For the latter, the second of Refs. quotes, e. g., $`M(\mathrm{{\rm Y}})M(\eta _\mathrm{b})=46.6\begin{array}{c}+14.8\hfill \\ 12.7\hfill \end{array}\text{MeV}.`$ Consequently, a re-evaluation of the b quark mass appears to be in order. Fitting the ground-state bound-state energy $`_0`$, for the numerical values of the parameters $`\mathrm{\Lambda },`$ $`\mu ,`$ and $`\alpha G^2`$ adopted in Ref. , to the experimental mass of the $`\mathrm{{\rm Y}},`$ we obtain for the pole mass of the b quark $$m_\mathrm{b}=4.983\text{GeV}.$$ The errors caused by the uncertainties of $`\mathrm{\Lambda }`$ and $`\alpha G^2`$ and by the variation of $`\mu `$ should be practically the same as in Ref. . Hence, dropping the requirement of analytical accessibility of the Hamiltonian (1) reduces the extracted b quark mass by some $`18\text{MeV}.`$ A very similar result is expected to be found for the determination of the c quark mass. We arrive at the conclusion that in Refs. the theoretical errors on the quark masses have been somewhat underestimated. Note that the above considerations apply, of course, also to the analysis presented in Ref. . Note also that the correct numerical factor in the numerator of the first term in the expression for the three-loop $`\beta `$ function coefficient $`\beta _2`$ differs slightly from the one used in Refs. .
warning/0001/hep-ph0001045.html
ar5iv
text
# Pion dispersion relation at finite temperature in the linear sigma model from chiral Ward identities ## Abstract We develop one-loop effective vertices and propagators in the linear sigma model at finite temperature satisfying the chiral Ward identities. We use these in turn to compute the pion dispersion relation in a pion medium for small momentum and temperatures on the order of the pion mass at next to leading order in the parameter $`m_\pi ^2/4\pi ^2f_\pi ^2`$ and to zeroth order in the parameter $`(m_\pi /m_\sigma )^2`$. We show that this expansion reproduces the result obtained from chiral perturbation theory at leading order. The main effect is a perturbative, temperature-dependent increase of the pion mass. We find no qualitative changes in the pion dispersion curve with respect to the leading order behavior in this kinematical regime. PACS numbers: 11.10.Wx, 11.30.Rd, 11.55.Fv, 25.75.-q The propagation properties of pions in a thermal hadronic environment are a key ingredient for the proper understanding of several physical processes taking place in dense and hot plasmas. The scenarios include dense astrophysical objects such as neutron stars –where the inclusion of mesonic degrees of freedom is essential to determine the equation of state of matter in the inner core– and the evolution of the highly interacting region formed in the aftermath of a high-energy heavy-ion collision. In the latter context, knowledge of the pion dispersion relation would help to properly account for the component of the dilepton spectrum originated in the hadronic phase, below the critical temperatures and densities for the formation of a quark-gluon plasma (QGP) , and to shed light on possible collective surface phenomena which are speculated to occur as a consequence of the mainly attractive interaction between pions and nuclei . In order to account for the hadronic degrees of freedom at temperatures and densities below the QCD phase transition to a QGP, one needs to resort to effective chiral theories whose basic ingredient is the fact that pions are Goldston bosons, originated in the spontaneous breakdown of chiral symmetry. Chiral perturbation theory ($`\chi `$PT) is one of such effective theories that has been successfully employed to show the well known result that at leading perturbative order and at low momentum, the modification of the pion dispersion curve in a pion medium is just a constant, temperature dependent, increase of the pion mass . Any deviation from this leading order behavior in perturbation theory can only be looked for at higher orders. Second order calculations of the pion dispersion curve can be found in Refs. . The simplest realization of chiral symmetry is nevertheless provided by the much studied linear sigma model which possesses the convenient feature of being a renormalizable field theory, both at zero and at finite temperature . This model has lately received a boost of attention in view of recent theoretical results and analyses of data that seem to confirm, though not without controversy , that a broad scalar resonance, with a mass in the vicinity of 600 MeV –that can be identified with the $`\sigma `$-meson– indeed exists. In addition to the above mentioned characteristics, the linear sigma model also reproduces –as we will show– the leading order modification to the pion mass in a thermal pion medium (for temperatures on the order of the pion mass), when use is made of a systematic expansion in the ratio $`(m_\pi /m_\sigma )^2`$ at zeroth order, where $`m_\sigma `$, $`m_\pi `$ are the vacuum sigma and pion masses, respectively. This important point, often missed in the literature , provides the connection between the linear sigma model and $`\chi `$PT at finite temperature. In this work, we use the linear sigma model to construct effective vertices and propagators to one loop. We use these in turn to compute the one loop modification to the pion propagator in a pion medium. We require that the effective vertices and propagators thus constructed satisfy the chiral Ward identities . The net result is a next to leading order calculation in the parameter $`m_\pi ^2/4\pi ^2f_\pi ^2`$, where $`f_\pi `$ is the pion decay constant and to zeroth order in the parameter $`(m_\pi /m_\sigma )^2`$. From the effective pion propagator, we explore the behavior of the pion dispersion curve at small momentum and for temperatures on the order of the pion mass. Though the momentum dependence of the correction term is non-trivial, we find no qualitative difference from the leading order result in this regime. Possible effects due to a high nucleonic density as well as general details of the calculation are left out for a future work. The Lagrangian for the linear sigma model, including only the mesonic degrees of freedom and after the explicit inclusion of the chiral symmetry breaking term, can be expressed as $`={\displaystyle \frac{1}{2}}\left[(\pi )^2+(\sigma )^2m_\pi ^2\pi ^2m_\sigma ^2\sigma ^2\right]\lambda ^2f_\pi \sigma (\sigma ^2+\pi ^2){\displaystyle \frac{\lambda ^2}{4}}(\sigma ^2+\pi ^2)^2,`$ (1) where $`\pi `$ and $`\sigma `$ are the pion and sigma fields, respectively, and the coupling $`\lambda ^2`$ is given by $`\lambda ^2={\displaystyle \frac{m_\sigma ^2m_\pi ^2}{2f_\pi ^2}}.`$ (2) From the Lagrangian in Eq. (1), one obtains the Green’s functions and the Feynman rules to be used in perturbative calculations, in the usual manner. Thus, in particular, the bare pion and sigma propagators $`\mathrm{\Delta }_\pi (P)`$, $`\mathrm{\Delta }_\sigma (Q)`$ and the bare one-sigma two-pion and four-pion vertices $`\mathrm{\Gamma }_{12}^{ij}`$, $`\mathrm{\Gamma }_{04}^{ijkl}`$ are given by (hereafter, capital letters are used to denote four momenta) $`i\mathrm{\Delta }_\pi (P)\delta ^{ij}`$ $`=`$ $`{\displaystyle \frac{i}{P^2m_\pi ^2}}\delta ^{ij}`$ (3) $`i\mathrm{\Delta }_\sigma (Q)`$ $`=`$ $`{\displaystyle \frac{i}{Q^2m_\sigma ^2}}`$ (4) $`i\mathrm{\Gamma }_{12}^{ij}`$ $`=`$ $`2i\lambda ^2f_\pi \delta ^{ij}`$ (5) $`i\mathrm{\Gamma }_{04}^{ijkl}`$ $`=`$ $`2i\lambda ^2(\delta ^{ij}\delta ^{kl}+\delta ^{ik}\delta ^{jl}+\delta ^{il}\delta ^{jk}).`$ (6) These Green’s functions are sufficient to construct the modification to the pion propagator, both at zero and finite temperature, at any given perturbative order. An alternative approach consists on exploiting the relations that chiral symmetry imposes among different n-point Green’s functions. These relations, better known as chiral Ward identities ($`\chi `$WIs), are a direct consequence of the fact that the divergence of the axial current may be used as an interpolating field for the pion. For example, two of the $`\chi `$WIs satisfied –order by order in perturbation theory– by the functions $`\mathrm{\Delta }_\pi (P)`$, $`\mathrm{\Delta }_\sigma (Q)`$, $`\mathrm{\Gamma }_{12}^{ij}`$ and $`\mathrm{\Gamma }_{04}^{ijkl}`$ are $`f_\pi \mathrm{\Gamma }_{04}^{ijkl}(;0,P_1,P_2,P_3)`$ $`=`$ $`\mathrm{\Gamma }_{12}^{kl}(P_1;P_2,P_3)\delta ^{ij}+\mathrm{\Gamma }_{12}^{lj}(P_2;P_3,P_1)\delta ^{ik}+\mathrm{\Gamma }_{12}^{jk}(P_3;P_1,P_2)\delta ^{il}`$ (7) $`f_\pi \mathrm{\Gamma }_{12}^{ij}(Q;0,P)`$ $`=`$ $`\left[\mathrm{\Delta }_\sigma ^1(Q)\mathrm{\Delta }_\pi ^1(P)\right]\delta ^{ij},`$ (8) where momentum conservation at the vertices is implied, that is $`P_1+P_2+P_3=0`$ and $`Q+P=0`$. Therefore, any perturbative modification of one of these functions introduces modifications in other, when the former are related to the latter through $`\chi `$WIs. At one loop and after renormalization, the sigma propagator is modified by finite terms. At zero temperature this modification is purely imaginary and its physical origin is that a sigma particle, with a mass larger than twice the mass of the pion, is unstable and has a (large) non-vanishing width coming from its decay channel into two pions. At finite temperature the modification results on real and imaginary parts. The real part modifies the sigma dispersion curve whereas the imaginary part represents a temperature dependent contribution to the sigma width. After renormalization, the leading one-loop contribution to the sigma self-energy, in an expansion in the parameter $`(m_\pi /m_\sigma )^2`$, can be shown to arise from the diagram depicted in Fig. 1. In the imaginary-time formalism of Thermal Field Theory (TFT), the expression for this diagram is given as $`6\lambda ^4f_\pi ^2I(\omega ,q)`$, where the function $`I`$ is defined by $`I(\omega ,q)=T{\displaystyle \underset{n}{}}{\displaystyle \frac{d^3k}{(2\pi )^3}\frac{1}{\omega _n^2+k^2+m_\pi ^2}\frac{1}{(\omega _n\omega )^2+(𝐤𝐪)^2+m_\pi ^2}}.`$ (9) Here $`\omega =2m\pi T`$ and $`\omega _n=2n\pi T`$ ($`m`$, $`n`$ integers) are discrete boson frequencies and $`q=|𝐪|`$. From Eq. (9) we obtain the time-ordered version $`I^t`$ of the function $`I`$, after analytical continuation to Minkowski space. The imaginary part of $`I^t`$ is given by $`\text{I}\text{m}I^t(q_0,q)`$ $`=`$ $`{\displaystyle \frac{\epsilon (q_o)}{2i}}\left[I(i\omega q_o+iϵ,q)I(i\omega q_oiϵ,q)\right]`$ (10) $`=`$ $`{\displaystyle \frac{1}{16\pi }}\left\{a(Q^2)+{\displaystyle \frac{2T}{q}}\mathrm{ln}\left({\displaystyle \frac{1e^{\omega _+(q_0,q)/T}}{1e^{\omega _{}(q_0,q)/T}}}\right)\right\}\mathrm{\Theta }(Q^24m_\pi ^2),`$ (11) where $`Q^2=q_0^2q^2`$, $`\epsilon `$ and $`\mathrm{\Theta }`$ are the sign and step functions, respectively, and the functions $`a`$ and $`\omega _\pm `$ are given by $`a(Q^2)`$ $`=`$ $`\sqrt{1{\displaystyle \frac{4m_\pi ^2}{Q^2}}}`$ (12) $`\omega _\pm (q_0,q)`$ $`=`$ $`{\displaystyle \frac{|q_0|\pm a(Q^2)q}{2}},`$ (13) whereas the real part of $`I^t`$ at $`Q=0`$ is given by $`\text{R}\text{e}I^t(0)={\displaystyle \frac{1}{8\pi ^2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dk}{E_k}}\left[1+2f(E_k)\right],`$ (14) where $`E_k=\sqrt{k^2+m_\pi ^2}`$ and the function $`f`$ is the Bose-Einstein distribution $`f(E_k)={\displaystyle \frac{1}{e^{E_k/T}1}}.`$ (15) Therefore, the one-loop effective sigma propagator becomes $`i\mathrm{\Delta }_\sigma ^{}(Q)`$ $`=`$ $`{\displaystyle \frac{i}{Q^2m_\sigma ^2+6\lambda ^4f_\pi ^2I^t(Q)}}.`$ (16) The temperature-independent infinities are absorbed into the redefinition of the physical masses and coupling constants by the introduction of suitable counterterms in the usual manner. In order to preserve the $`\chi `$WIs expressed in Eq. (8), the corresponding one-loop effective one-sigma two-pion and four-pion vertices become $`i\mathrm{\Gamma }_{12}^{ij}(Q;P_1,P_2)`$ $`=`$ $`2i\lambda ^2f_\pi \delta ^{ij}\left[13\lambda ^2I^t(Q)\right]`$ (17) $`i\mathrm{\Gamma }_{04}^{ijkl}(;P_1,P_2,P_3,P_4)`$ $`=`$ $`2i\lambda ^2\{[13\lambda ^2I^t(P_1+P_2)]\delta ^{ij}\delta ^{kl}`$ (18) $`+`$ $`\left[13\lambda ^2I^t(P_1+P_3)\right]\delta ^{ik}\delta ^{jl}`$ (19) $`+`$ $`[13\lambda ^2I^t(P_1+P_4)]\delta ^{il}\delta ^{jk}\}.`$ (20) It can be shown that the functions in Eq. (20) arise from considering all of the possible one-loop contributions to the one-sigma two-pion and four-pion vertices, when maintaining only the zeroth order terms in a systematic expansion in the parameter $`(m_\pi /m_\sigma )^2`$. We now use the above effective vertices and propagator to construct the one-loop modification to the pion self-energy. The leading contributions come from the diagrams depicted in Fig. 2. These are, explicitly $`\mathrm{\Pi }_a(P)\delta ^{ij}`$ $`=`$ $`\lambda ^2\delta ^{ij}{\displaystyle _\beta }{\displaystyle \frac{d^4K}{(2\pi )^4}}{\displaystyle \frac{1}{K^2m_\pi ^2}}\left\{59I^t(0)6\lambda ^2I^t(P+K)\right\}`$ (21) $`\mathrm{\Pi }_b(P)\delta ^{ij}`$ $`=`$ $`\lambda ^2\delta ^{ij}\left({\displaystyle \frac{6\lambda ^2f_\pi ^2}{m_\sigma ^2}}\right){\displaystyle _\beta }{\displaystyle \frac{d^4K}{(2\pi )^4}}{\displaystyle \frac{1}{K^2m_\pi ^2}}\left\{{\displaystyle \frac{[13\lambda ^2I^t(0)]^2}{[1\frac{6\lambda ^4f_\pi ^2}{m_\sigma ^2}I^t(0)]}}\right\}`$ (22) $`\mathrm{\Pi }_c(P)\delta ^{ij}`$ $`=`$ $`\lambda ^2\delta ^{ij}\left({\displaystyle \frac{4\lambda ^2f_\pi ^2}{m_\sigma ^2}}\right){\displaystyle _\beta }{\displaystyle \frac{d^4K}{(2\pi )^4}}{\displaystyle \frac{1}{K^2m_\pi ^2}}\left\{{\displaystyle \frac{[13\lambda ^2I^t(P+K)]^2}{[1\frac{6\lambda ^4f_\pi ^2}{m_\sigma ^2}I^t(P+K)\frac{(P+K)^2}{m_\sigma ^2}]}}\right\},`$ (23) where the subindex $`\beta `$ means that the integrals are to be computed at finite temperature. We now expand the denominators in the second and third of Eqs. (23), keeping only the leading order contribution when considering $`m_\pi `$, $`T`$ and $`P`$ as small compared to $`m_\sigma `$. Adding up the above three terms and to zeroth order in $`(m_\pi /m_\sigma )^2`$ where $`\lambda ^2\left(1{\displaystyle \frac{2\lambda ^2f_\pi ^2}{m_\sigma ^2}}\right){\displaystyle \frac{m_\pi ^2}{2f_\pi ^2}},`$ (24) the pion self-energy can be written as $`\mathrm{\Pi }(P)`$ $`=`$ $`\left({\displaystyle \frac{m_\pi ^2}{2f_\pi ^2}}\right)T{\displaystyle \underset{n}{}}{\displaystyle }{\displaystyle \frac{d^3k}{(2\pi )^3}}{\displaystyle \frac{1}{K^2+m_\pi ^2}}\{5+2\left({\displaystyle \frac{P^2+K^2}{m_\pi ^2}}\right)`$ (25) $``$ $`\left({\displaystyle \frac{m_\pi ^2}{2f_\pi ^2}}\right)[9I^t(0)+6I^t(P+K)]\},`$ (26) where we carry out the calculation in the imaginary-time formalism of TFT with $`K=(\omega _n,𝐤)`$ and $`P=(\omega ,𝐩)`$. The pion dispersion relation is thus obtained from the solution to $`P^2+m_\pi ^2+\text{R}\text{e}\mathrm{\Pi }(P)=0,`$ (27) after the analytical continuation $`i\omega p_o+iϵ`$. As it stands, Eq. (26) contains a temperature-dependent infinity coming from the product of the vacuum piece in $`\text{R}\text{e}I^t(0)`$ and the temperature-dependent piece in the indicated integral, as well as vacuum infinities. However, it is well known that finite temperature does not introduce new divergences and that whatever regularization and renormalization is needed at zero temperature, it will also be necessary and sufficient at finite temperature. Let us recall that at one loop, renormalization is carried out by introducing appropriate counterterms into the original Lagrangian and that at this level, no temperature-dependent infinities arise. Going to two loops, one will always encounter temperature-dependent infinities in integrals involving only the bare terms of the original Lagrangian. The remarkable property of renormalizable theories is that these temperature-dependent infinities are exactly canceled by the contribution from the loops computed by using the counterterms that were introduced at one loop. The above was explicitly shown for the case of the self-energy in the $`\varphi ^4`$ theory by Kislinger and Morley and for the sigma vacuum expectation value in the linear sigma model by Mohan . For this reason, here we omit the renormalization procedure and consider only the finite, temperature-dependent terms throughout the rest of the calculation and refer the reader to the cited references for details. Let us first look at the dispersion relation at leading order. After analytical continuation all the terms are real. The integrals involved are $`T{\displaystyle \underset{n}{}}{\displaystyle \frac{d^3k}{(2\pi )^3}\frac{1}{K^2+m_\pi ^2}}`$ $``$ $`{\displaystyle \frac{1}{2\pi ^2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dkk^2}{E_k}}f(E_k)`$ (28) $``$ $`{\displaystyle \frac{m_\pi ^2}{2\pi ^2}}g(T/m_\pi )`$ (29) $`T{\displaystyle \underset{n}{}}{\displaystyle \frac{d^3k}{(2\pi )^3}\frac{K^2}{K^2+m_\pi ^2}}`$ $``$ $`m_\pi ^2\left({\displaystyle \frac{m_\pi ^2}{2\pi ^2}}\right)g(T/m_\pi ),`$ (30) where $`g`$ is a dimensionless function of the ratio $`T/m_\pi `$ and the arrows indicate only the temperature dependence of the expressions. Thus, the dispersion relation results from $`\left[1+2\xi g(T/m_\pi )\right](p_0^2p^2)\left[1+3\xi g(T/m_\pi )\right]m_\pi ^2=0,`$ (31) with $`\xi =m_\pi ^2/4\pi ^2f_\pi ^21`$. For $`Tm_\pi `$, $`g(T/m_\pi )1`$, therefore, at leading order and in the kinematical regime that we are considering, Eq. (31) can be written as $`p_0^2=p^2+m_\pi ^2\left[1+\xi g(T/m_\pi )\right],`$ (32) which coincides with the result obtained from $`\chi `$PT . We now look at the next to leading order terms in Eq. (26). The first of these is purely real and represents a constant, second order shift to the pion mass squared $`9\left({\displaystyle \frac{m_\pi ^2}{2f_\pi ^2}}\right)^2T{\displaystyle \underset{n}{}}{\displaystyle \frac{d^3k}{(2\pi )^3}\frac{I^t(0)}{K^2+m_\pi ^2}}{\displaystyle \frac{9}{2}}\xi ^2g(T/m_\pi )h(T/m_\pi )m_\pi ^2,`$ (33) where $`h`$ is a dimensionless function of the ratio $`T/m_\pi `$ defined by $`h(T/m_\pi ){\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dk}{E_k}}f(E_k).`$ (34) The remaining term in Eq. (26) shows a non-trivial dependence on $`P`$. It involves the function $`S`$ defined by $`S(P)T{\displaystyle \underset{n}{}}{\displaystyle \frac{d^3k}{(2\pi )^3}\frac{I^t(P+K)}{K^2+m_\pi ^2}}.`$ (35) The sum is performed by resorting to the spectral representation of $`I^t`$ and $`(K^2+m_\pi ^2)^1`$. Thus, the real part of the retarded version of $`S`$, after analytical continuation is $`\text{R}\text{e}S^r(p_0,p)`$ $``$ $`{\displaystyle \frac{1}{2}}\left[S(i\omega p_0+iϵ,p)+S(i\omega p_0iϵ,p)\right]`$ (36) $`=`$ $`𝒫{\displaystyle \frac{d^3k}{(2\pi )^3}_{\mathrm{}}^{\mathrm{}}\frac{dk_0}{2\pi }_{\mathrm{}}^{\mathrm{}}\frac{dk_0^{}}{2\pi }\left[1+f(k_0)+f(k_0^{})\right]}`$ (38) $`{\displaystyle \frac{2\pi \epsilon (k_0^{})\delta [k_{0}^{}{}_{}{}^{2}(𝐤𝐩)^2m_\pi ^2]2\text{I}\text{m}I^t(k_0,k)}{p_0k_0k_0^{}}},`$ where $`𝒫`$ represents the principal part of the integral. The integration in Eq. (38) can only be performed numerically. Figure 3 shows a plot of the temperature-dependent terms of the function $`\stackrel{~}{S}(\stackrel{~}{p}_0,p)=(24\pi ^4)\text{R}\text{e}S^r(\stackrel{~}{p}_0,p)`$ for $`T=m_\pi `$ and $`\stackrel{~}{p}_0`$ taken as the solution of Eq. (27) up to the $`P`$-independent terms of Eq. (26). ¿From this figure we see that $`\stackrel{~}{S}`$ is non-vanishing at $`p=0`$ and thus it also contributes to the perturbative increase of the pion mass. We also notice that $`\stackrel{~}{S}`$ is a monotonically increasing function of $`p`$ in the kinematical range considered. However, this increase is not strong enough to cause a qualitatively significant change on the shape of the dispersion curve, given that this function contributes multiplied by the perturbative parameter $`\xi ^2`$. Including all the terms, the dispersion relation up to next to leading order, for $`Tm_\pi `$ and in the small momentum region is obtained as the solution to $`p_0^2=p^2+\left\{1+\xi g(T/m_\pi )+{\displaystyle \frac{\xi ^2}{2}}g(T/m_\pi )\left[9h(T/m_\pi )4g(T/m_\pi )\right]\right\}m_\pi ^2+\xi ^2\stackrel{~}{S}(p_0,p).`$ (39) Figure 4 shows the dispersion relation obtained from Eq. (39) for $`T=m_\pi `$ where we also display the solution without the term $`\xi ^2\stackrel{~}{S}(p_0,p)`$. As mentioned before, inclusion of this last term does not alter the shape of the dispersion relation in this kinematical regime. In conclusion, we have shown that in the linear sigma model at finite temperature, the one-loop modification of the sigma propagator induces a modification in the one-sigma two-pion and four-pion vertices in such a way as to preserve the $`\chi `$WIs. Strictly speaking, chiral Ward identities fully constrain vertices only in the limit where all the external momenta are zero . Nevertheless, we have also checked that these functions arise from considering all of the possible contributions to the vertices of interest, when maintaining only the zeroth order terms in an expansion in the parameter $`(m_\pi /m_\sigma )^2`$ . We have used these objects to compute the next to leading order correction to the pion propagator in a pion medium for small momentum and for $`Tm_\pi `$. We have shown that the linear sigma model yields the same result as $`\chi `$PT at leading order in the parameter $`\xi =m_\pi ^2/4\pi ^2f_\pi ^2`$ when use is made of a systematic expansion in the parameter $`(m_\pi /m_\sigma )^2`$ at zeroth order. This result was to be expected since the kinematical regime we consider is that where the temperature, the pion momentum and the pion mass are treated as small quantities, such as in the case of $`\chi `$PT. The main modification to the pion dispersion curve in the considered kinematical regime is a perturbative increase of the in-medium pion mass. The shape of the curve is not significantly altered. For a sigma meson with a mass on the order of 600 MeV and in the mentioned kinematical regime, corrections in powers of the parameter $`(m_\pi /m_\sigma )^2`$ are small. If a lighter sigma meson exists, corrections of order $`(m_\pi /m_\sigma )^2`$ could be important. It remains to include possible effects introduced by a high nucleonic density. Finally, it is interesting to note that the formalism thus developed could be employed to explore the behavior of the pion dispersion curve in the large momentum region where, in principle, the lowest order $`\chi `$PT Lagrangian cannot be used. These issues will be treated in a following up work. Support for this work has been received in part by CONACyT México under grant numbers I27604-E, 29273-E and 32279-E.
warning/0001/astro-ph0001371.html
ar5iv
text
# A direct view of the AGN powering IRAS12393+3520 ## 1 Introduction The correlation between the ROSAT All Sky Survey (RASS) and the IRAS Point Source Catalog unveiled the existence of many galaxies with X-ray luminosity in the range $`10^{42}`$$`10^{43}`$ erg s<sup>-1</sup>, which had not been previously classified as Active Galactic Nuclei (AGN) (Boller et al. 1992). Subsequent optical observations indicated that a large fraction of them were previously unrecognized Seyfert galaxies (Moran et al. 1994, 1996; Dennefeld et al. in preparation). However, the nature of a few of them is far from being understood in terms of simple nuclear activity and deserves further investigations. IRAS12393+3520 (NGC 4619) is one of those, which we started to study in more details in 1992. It is a nearby ($`\mathrm{z}=0.023`$) barred spiral galaxy of morphological type SB(r)b, with a luminous nucleus and relatively weak spiral arms. Our first optical spectrum (Boller et al. 1993) was obtained at low resolution in March 1992 at the 1.93m telescope of Haute-Provence Observatory (OHP) and is displayed in Fig. 1. It shows a red continuum, typical of IRAS galaxies, with many stellar absorption features and conspicuous emission only in the red part of the spectrum (H<sub>α</sub>, \[Nii\]). H<sub>β</sub> is in absorption, and the \[OIII\] emission barely visible. The existence of a strong NaD absorption line is an indication that significant internal extinction is present. No clear evidence is seen of a broad H<sub>α</sub> component: a hint of it may be seen on the red side of the narrow emission line, but the presence of the atmospheric B band does not allow a proper estimate of the level of the continuum; this prompted further observations. In the mean time, other spectra were published by Moran et al. (1994) and Mas-Hesse et al. (1996, hereinafter M96), showing a broad H<sub>α</sub> component attributed to a Seyfert 1 nucleus. While their two low-dispersion spectra are compatible with each other, their strong, broad H<sub>α</sub> line is not seen on ours. A high-resolution spectrum taken by M96 only 25 days after our first spectrum displays a faint, H<sub>α</sub> broad component, but apparently much fainter than in their low-dispersion spectrum taken about two years later. M96 noted also that the Nii/H<sub>α</sub> ratio in IRAS12393+3520 was reminiscent of a LINER (Filippenko 1993 and references therein), but that, apart of it, the spectrum was remarkably featureless. From the lack of any detectable broad component of the H<sub>β</sub>, they estimated an amount of extinction $`\mathrm{E}(\mathrm{B}\mathrm{V})\stackrel{>}{}0.4`$, corresponding to $`\mathrm{N}_\mathrm{H}\stackrel{>}{}2\times 10^{21}`$ cm<sup>-2</sup> (Lequeux et al. 1981; Prehdel & Schmitt 1995); such a value is not inconsistent with the reddening estimated from the narrow components. The UV spectra obtained by M96 show also broad Ly<sub>α</sub> and Mgii emission, as well as some absorption lines. The $`\mathrm{E}(\mathrm{B}\mathrm{V})`$ inferred from the UV measurements is $``$0.20. The optical to X-ray Spectral Energy Distribution (SED) of IRAS12393+3520 has been studied in detail by M96. They pointed out that a scenario consisting of intense starburst can well explain the infrared, optical and UV emissions and the optical emission lines, but underestimates the soft X-rays. On the other hand, if UV and X-rays are dominated by a non-stellar contribution, the FIR emission is then highly in excess over the one typically observed in Seyfert galaxies. The presence of both sources, with comparable amount of energy output, seemed the most viable solution for this object. Secular optical variability in itself is not discriminating between the two mechanisms if timescales are not determined, as Terlevich et al. (1992) have argued that massive star formation could also account both for variability and broad lines. IRAS12393+3520 is bright in soft X-rays \[$`\mathrm{log}(\mathrm{L}_{0.12.4\mathrm{keV}})=42.80`$\], with a photon spectral index $`\mathrm{\Gamma }1.90`$ (Moran et al. 1996). Studies in the X-rays are in principle of the uppermost importance to address properly the true origin of the output power, because they can pierce throughout the innermost regions of the galaxy, and unveil the high-energy processes that occur in the immediate proximity of the nucleus. An ASCA observational program was started to search for evidence of nuclear activity in a sub-set of the Boller et al. (1992) soft X-ray luminous galaxies sample. In this paper, we report the results on IRAS12393+3520 (the only target for which time was allocated), which provided the first hard X-ray measure for this object. The results are presented together with a re-analysis of a pointed archival ROSAT/PSPC observation and with new optical data obtained in the mean time. We will assume $`\mathrm{H}_0=50`$ km s<sup>-1</sup> Mpc<sup>-1</sup>. Energies are quoted in the source rest frame and errors are at 90% confidence level for one interesting parameter ($`\mathrm{\Delta }\chi ^2=2.71`$). ## 2 X-ray Data Reduction and analysis IRAS12393+3520 was observed by ROSAT/PSPC and ASCA on 1993, 16–17 December, and 1997, 4–5 June, respectively. The data have been retrieved from the HEASARC archive, as cleaned event lists. The ROSAT PSPC is a proportional counter sensitive in the 0.1–2 keV band, with moderate energy resolution ($`\mathrm{\Delta }\mathrm{E}/\mathrm{E}50\%`$ at 1 keV) and a spatial resolution of about 20$`\mathrm{}`$ for on-axis sources. The ASCA scientific payload comprises two Solid Imaging Spectrometers (SIS0 and SIS1, sensitive in the 0.6–9 keV band, Gendreau 1995) and two Gas Imaging Spectrometers (GIS2 and GIS3, sensitive in the 0.7–10 keV band, Ohashi et al. 1996). SIS grades 0,1,2 and 4 and 1-CCD mode were used. The best ROSAT/PSPC centroid position ($`\alpha _{2000.0}`$=$`12^h41^m44^s.8`$ and $`\delta _{2000.0}`$=$`35^o`$03$`\mathrm{}`$46$`\mathrm{}`$) is within 22$`\mathrm{}`$ from the optical position (Santagata et al. 1987) and 40$`\mathrm{}`$ from the ASCA centroid. The latter quantity is comparable with the typical ASCA positional uncertainties (Gotthelf & Ishibashi 1997). Several serendipitous sources are present in the PSPC field of view (see Fig. 2). Two of them could in principle contaminate the ASCA source spots. Their properties are reported in Table 1. We expect that any contamination to the IRAS12393+3520 ASCA flux is lower than 10% and 5% in the 0.5–2 and 2–10 keV bands, respectively, well within the statistical uncertainties of the ASCA measurement. PSPC scientific products have been extracted from a circular region of about 2$`\mathrm{}`$ around the apparent centroid of the source. Background spectra have been extracted in an annulus of radii 3$`\mathrm{}`$20$`\mathrm{}`$ and 5$`\mathrm{}`$20$`\mathrm{}`$, after removing circular areas of 2$`\mathrm{}`$ radius around any serendipitous source included in the annulus. In the SIS0 or SIS1 images no source is detected other than IRAS12393+3520, the 3$`\sigma `$ upper limit at the position of source #1 in PSPC field being $`8\times 10^3`$ s<sup>-1</sup>. The closest serendipitous source in the GIS field of view ($`\alpha _{2000.0}`$=$`12^h41^m44^s.8`$ and $`\delta _{2000.0}`$=$`35^{}`$03$`\mathrm{}`$46$`\mathrm{}`$; distance 5.8$`\mathrm{}`$) is probably associated with the foreground galaxy NGP9 F268-1081523. ASCA source scientific products have been then extracted from circles of radii 4$`\mathrm{}`$, 3$`\mathrm{}`$.25 and 3$`\mathrm{}`$.75, in the SIS0, SIS1 and GIS, respectively. Background scientific products have been extracted from regions of the field of view free from contaminating sources. None of the presented results changes significantly if the background is extracted from blank sky fields. Total exposure times are about 33, 32 and 13 ks for the SIS, GIS and PSPC, respectively. Full band count rates were $`(3.79\pm 0.12)`$, $`(2.77\pm 0.11)`$, $`(1.89\pm 0.09)`$, $`(2.38\pm 0.10)`$ and $`(6.4\pm 0.3)\times 10^2`$s<sup>-1</sup>, for the SIS0, SIS1, GIS2, GIS3 and PSPC, respectively. ### 2.1 The X-ray light curve In Fig. 3 the light curves in the 0.1–2 keV (PSPC) and 0.5–4 keV bands (SIS) are shown. The latter shows a gentle rise and decay before about 50 ks from the start of the observation. A similar trend is observed in the GIS light curves. If a fit with a constant line is performed on the light curve, $`\chi ^2|_{\mathrm{ASCA}}=30/15`$ degrees of freedom (dof). From the flux change rate of the light curve in the 5–35 and 35–50 ks intervals, we estimate a minimum doubling/halving time in the range 30–75 ks. Variability on lower timescales cannot be ruled out, but the limited available statistics prevented us to reach a firm conclusion on its existence. This variability is not associated with any of the serendipitous sources detected in the PSPC and laying at the border of the ASCA extraction regions. We have extracted SIS0 images in the time intervals between 10-40 and 50-80 ks after the beginning of the ASCA observation. The source spot does not show sign of elongation and the centroid best-fit positions agree within 6$`\mathrm{}`$. Although an eye inspection of the PSPC light curve might suggest that a similar variability was observed by ROSAT as well, the evidence is not statistically that clear ($`\chi ^2|_{\mathrm{ROSAT}}=18/21`$ dof). We have searched for spectral dynamics associated with the ASCA flux changes, without success. We will therefore focus in the next section on the time averaged spectrum of IRAS12393+3520. ### 2.2 Spectral analysis All spectra have been rebinned in order to have at least 20 counts per energy channel, in order to ensure the applicability of the $`\chi ^2`$ statistics. Proper matrices for the date of the observations have been retrieved from the HEASARC archive or built with the software available in the Ftools 4.0 version. First, we fitted the PSPC and ASCA spectra in the overlapping 0.8–2 keV band with a simple photoelectric absorbed power–law model, leaving all parameters free (except the $`\mathrm{N}_\mathrm{H}`$, which has been tight to be the same for all instruments), to check if significant spectral variability arose between the epochs of the two observations. The spectral indices turn out to be well consistent within the statistical uncertainties ($`\mathrm{\Gamma }_{\mathrm{ROSAT}}^{0.82\mathrm{k}\mathrm{e}\mathrm{V}}=1.7\pm _{0.8}^{1.0}`$; $`\mathrm{\Gamma }_{\mathrm{ASCA}}^{0.82\mathrm{k}\mathrm{e}\mathrm{V}}=1.5\pm _{0.4}^{0.7}`$), despite of an increase in the flux of $`130\%`$. In the following, we will therefore fit the spectra of all detectors simultaneously, only allowing a relative normalization factor as a free parameter among all the instruments. In Fig. 4, the results of a spectral fit with a simple power-law with photoelectric absorption are shown. The fit is rather good ($`\chi ^2=188/177`$ dof), with $`\mathrm{\Gamma }=1.55\pm _{0.05}^{0.08}`$ and $`\mathrm{N}_\mathrm{H}=(0.8\pm _{0.2}^{0.3})\times 10^{20}`$ cm<sup>-2</sup>. Further spectral complexity is not strongly required by the data. The only feature, whose addition is required at $``$99% level, is an absorption edge ($`\mathrm{\Delta }\chi ^2=9`$ for two more parameters), with threshold energy $`\mathrm{E}_{\mathrm{th}}=0.71\pm _{0.06}^{0.05}`$, formally consistent with the K-shell photoionization energy of Ovii. The edge is detected with comparable significance in both the ROSAT and ASCA spectra separately. Similar features have been commonly observed in the spectra of Seyfert 1 galaxies (Reynolds 1997; George et al. 1998). The addition of this feature makes the spectrum steeper and therefore consistent with that typically observed in Seyfert 1 galaxies (Nandra & Pounds 1994; Nandra et al. 1997b; see Table 2). We have therefore performed a fit with a Seyfert-like model, constituted by a Compton-reflected power-law (model `pexrav` in Xspec, Magdziarz & Zdziarski 1995), an absorption edge and a fluorescent broad (i.e.: $`\sigma =0.43`$ keV, Nandra et al. 1997b) iron line from neutral iron (i.e: centroid energy held fixed to 6.4 keV). The line is actually not required by the fit, but we have included it because it is expected on theoretical grounds (George & Fabian 1991; Matt et al. 1992). The fit is comparably good as in the simple power-law case. The amount of reflection, parameterized through the ratio between the reflected and the transmitted components, R<sup>1</sup><sup>1</sup>1R is equal to 1 when the reflection occurs in a plane-parallel infinite slab. A higher value might imply either a higher solid angle subtended by the disk to the source - as e.g. in a disk “warped” geometry -, a delay in the disk response to flux changes of the primary continuum, or an anisotropy of the nuclear emission, such that the disk sees more flux than emitted along our cone-of-sight, is $`>`$1.3, which is not inconsistent with the upper limit on the EW of the iron line (600 eV), if standard cosmic abundances are assumed (Matt et al. 1992). If we add to this model a 1 keV thermal emission from an optically thin plasma, its 0.1–10 keV luminosity is constrained to be lower than $`3.8\times 10^{41}`$ erg s<sup>-1</sup> at the 90% confidence level. The cold absorbing column density is broadly consistent with the Galactic contribution along the IRAS12393+3520 line of sight ($`\mathrm{N}_{\mathrm{H},\mathrm{Gal}}=1.4\times 10^{20}`$ cm<sup>-2</sup>, Dickey & Lockman 1990). The fluxes in the 0.5–2, 0.5–4.5 and 2–10 keV energy bands are $`5.2\times 10^{13}`$, 1.00 and $`1.12\times 10^{12}`$ erg cm<sup>-2</sup> s<sup>-1</sup>, implying rest frame unabsorbed luminosities of 1.29, 2.4 and $`2.6\times 10^{42}`$ erg s<sup>-1</sup>, respectively. Alternatively, a good fit can be formally obtained also with a double-temperature optically thin plasma (code `mekal` in Xspec), with temperatures $`100`$ eV and $`8`$ keV and abundances consistent with solar. In this scenario, the contribution of a $`\mathrm{\Gamma }=1.9`$ (1.0) power-law to the 0.5-4 keV flux is lower than 7% (6%). ## 3 Additional optical data An intermediate resolution spectrum (1.8 Åper pixel) of the H<sub>α</sub> region of NGC 4619 was obtained in January 1996 at OHP. The calibrated one dimensional spectrum has been extracted in two ways, shown in Fig. 5. The first one (upper spectrum in Fig. 5) integrates over all the length were H<sub>α</sub> is detected (about 40”): the emission lines are clearly detected, with a \[Nii\]/H<sub>α</sub> ratio stronger than in standard Hii regions. This ratio was the basis for the LINER claim in M96. Unfortunately, the \[Sii\] lines, which could provide an additional diagnostic, fall into an atmospheric absorption line. As any clear broad component, revealing an AGN, could be hidden in a much larger star forming region, a second extraction was performed over a smaller extension (about 10”), corresponding to the strong continuum. This is the lower spectrum of Fig.5, where only weak \[Nii\] emission lines are seen, and almost no H<sub>α</sub>. As the latter may however be hidden in the corresponding stellar absorption, it is difficult to use the \[Nii\]/H<sub>α</sub> line ratio alone to claim for a diagnostic of Seyfert or LINER. In both spectra a broad H<sub>α</sub> component may be marginally seen but, if present, is certainly not as strong as seen in the high dispersion spectrum of M96 (and even less so than in their low dispersion one). We finally obtained a new, low dispersion spectrum, in June 1999 at OHP, and the result is shown in Fig. 6. This spectrum, extracted over a similar 10” width as the lower one in Fig. 5, shows a clear, broad, asymmetric wing to H<sub>α</sub>, with a strong component on the blue side (where nothing was suspected in our 1992 spectrum) very similar to the one seen in the low dispersion spectra of M96 and Moran et al. (1994). While a quantitative comparison between the various spectra is difficult, as slit orientations and extraction lengths are different (and not always well documented), we have here reasonable evidence for variations of a broad-line component in a spectrum otherwise dominated by intermediate-type stars. The broad H<sub>α</sub> component is absent or weak in the first available spectra (ours in March 1992, the high-resolution one of April 1992 in M96), is strong in the low dispersion spectra of Moran et al. (1994) and of M96 (July 1994), dimmed again in our data of January 1996 and strong again in our last spectrum in June 1999. The variations seen are between a Seyfert 1.9 spectrum and a spectrum dominated by stellar features. Some indications exist for a LINER identification: we see a strong \[Nii\]/H<sub>α</sub> ratio, as M96 did, and our 1999 spectrum shows also an \[Oii\]/\[Oiii\] ratio much greater than one. We have however no detection of the \[Oi\] line with a strength which could confirm the LINER diagnostic (Heckman, 1980). We note that up to now, only one case is known where broad lines developed in a LINER (NGC 1097; Storchi-Bergmann et al. 1993). The substantial reddening derived in our last spectrum \[$`\mathrm{E}(\mathrm{B}\mathrm{V})0.80.9`$, consistent between the Balmer decrement in the narrow component and the NaD absorption\], is certainly an element of importance in the following discussion. ## 4 Discussion ### 4.1 The hard X-ray properties of IRAS12393+3520 X-rays provide a strong evidence in favor of the existence of an AGN in IRAS12393+3520. We have detected a significant variability of the X-ray light curve. The minimum observed associated doubling/halving time is comprised in the range 30–75 ks. If the variability is intrinsic to the radiation emitted in the nuclear region of an AGN, usual light crossing arguments constrain the mass of the nuclear black hole to be $`<10^9\eta _5^1\mathrm{M}_{}`$, where $`\eta _5\mathrm{r}/(5\mathrm{R}_\mathrm{S})`$ is the typical size of the region emitting the bulk of the primary non-thermal continuum in unit of five Schwarzschild radii. The joint ROSAT/ASCA 0.1–10 keV spectrum has also a shape which closely resembles the one typically observed in radio-quiet AGN. Our data suggest that the solid angle subtended by the accretion disk is higher than due to a plane-parallel infinite slab. The upper limit on the EW of a fluorescent neutral iron line (600 eV) is not inconsistent with this scenario if solar abundances are assumed. However, much better quality of the data is required to confirm this hint. The data require at 99% level of confidence the presence of an absorption edge, whose threshold energy is consistent with the K-shell photoionization energy of Ovii. In at least 50% of Seyfert 1s the nuclear radiation is absorbed by substantially ionized matter (Reynolds 1997; George et al. 1998). In IRAS12393+3520 the properties of the absorbing matter are statistically poorly constrained and are somewhat dependent on the assumed continuum model. We conservatively derive a column density in the range 2.5–$`7.5\times 10^{21}`$ cm<sup>-2</sup> \[or $`\mathrm{E}(\mathrm{B}\mathrm{V})0.5`$–1.5 under standard assumptions\], if the physical conditions required to sustain an oxygen state dominated by He-like are assumed. If the warm absorber contains dust in the Galactic dust-to-gas ratio, such a medium may account for the observed optical reddening and explain the missing H<sub>β</sub> broad line, even in the presence of a strong broad H<sub>α</sub> (which is actually detected only in some of the IRAS12393+3520 optical spectra, cf. Sect. 3). “Dusty” warm absorbers have already been invoked to explain the discrepancy between the amount of X-ray cold absorption and the optical reddening in a few Seyfert galaxies (Reynolds et al. 1997; Komossa & Fink 1998 and references therein), although Siebert et al. (1999) have recently shown that such a ”simple” representation was not able to reproduce all the observed optical and X-rays data in the case of a well studied example, IRAS13349+2438. The observed spectra can be formally accounted by a two-temperature optically thin plasma, in analogy with recent evidence on the X-ray emission of starburst galaxies (Della Ceca et al. 1998; Cappi et al. 1999). In this scenario, however, the observed short-term variability cannot be explained, unless it represents the diluted appearance of a much more variable underlying unresolved source. The lack of spectral variability during the ASCA observation points against the emerging of a further source in the spectrum in correspondence to the flux variation. The contribution of a power-law source in the 0.5-4 keV band is constrained to be lower than $`1.8\times 10^{41}`$ erg s<sup>-1</sup>. To reproduce the observed flux change, it should vary in $`10^4`$ seconds by an amount $`\frac{12}{L_{41}}`$, where $`\mathrm{L}_{41}`$ is the X-ray luminosity in units of $`10^{41}`$ erg s<sup>-1</sup>. This rules out the possible serendipitous detection of local X-ray binaries or superluminal sources; the latter ones exhibit indeed variability in flux by an amount up to two orders of magnitude, but have normally luminosities well within $`10^{39}`$ erg s<sup>-1</sup>. On the other hand, X-ray variability on such timescales is rather common in Seyfert galaxies (Nandra et al. 1997a and references therein). The timescales are too fast to be explained by mechanisms which do not invoke accretion on a supermassive black hole (e.g. the starburst model of Terlevich et al. 1992). ### 4.2 Comparison between X-ray and other wavelengths An energetic argument points against the idea that hard X-rays are dominated by a nuclear starburst. No known starburst galaxy is so X-ray luminous (Ptak et al. 1999). The soft-X vs. 60$`\mu `$m and soft-X vs. UV luminosity ratios are $`1.9\times 10^2`$ and $`1.2\times 10^2`$, more than one order of magnitude higher than typical values observed in star forming galaxies (SFG, Mass-Hesse et al. 1995). The observed X-ray luminosity $`4\times 10^{42}`$ erg s<sup>-1</sup> is, on the other hand, not uncommon among Seyfert galaxies. In Figure 7 the IRAS12393+3520 SED is shown, superimposed to the average “tracks” of the Mass-Hesse et al. (1995) sample for Seyfert 1s, Seyfert 2s, QSO and SFG. Caution must be employed when evaluating the SED in IRAS12393+3520, because it refers to non-simultaneous measurements. Optical spectroscopy suggests that the nuclear activity (as seen by the BLR), and therefore also the energy budget, might be variable with time. Given these caveats, one notes that IRAS12393+3520 follows apparently better a Seyfert 2 SED (or eventually a SFG one), being strongly under-luminous in UV and X-rays in comparison to the IR, if the normalization is done in the far-IR. These pieces of information are summarized in the IR, UV and X-ray color-color plot of Figure 8, where IRAS12393+3520 is seen to lie in the region of Seyfert galaxies, and apparently type 2 rather than 1. The X vs. \[OIII\] luminosity ratio is $`3.7\times 10^2`$, an order of magnitude greater than observed in Seyfert 1s (cf. Fig. 6 in Maiolino et al. 1998) and would also point towards a Seyfert 2 identification. Several pieces of evidence point, however, against an identification of IRAS12393+3520 as a Seyfert 2. First, there is no evidence of the high excitation typical of Sey2’s: the \[OIII\] over H<sub>β</sub> line ratio is difficult to measure accurately, as \[OIII\] is faint and H<sub>β</sub> is hidden in the corresponding stellar absorption, but is not larger than three. In addition, the \[OII\]/\[OIII\] ratio is much larger than one, and it is the combination of these line ratios, with the \[NII\]/H<sub>α</sub> one, which lead to the LINER claim for this object. This is not compatible with a Seyfert 2 identification. Moreover, a high excitation is not seen either when the broad H<sub>α</sub> component is absent, so that the variations cannot be described as a transition between a Seyfert 1 and a Seyfert 2 phase, as seen in some other cases (e.g. Mkn 993, Tran et al. 1992; Mkn 6, Khachikian & Weedman 1971; Mkn 1018, Cohen et al. 1986). Second, the X-ray intrinsic neutral absorption column density is negligible. No model where the nuclear power-law is strongly absorbed yields comparably good fits to the ROSAT/ASCA spectrum as the simple models of Table 2, even if one allows the covering fraction of the absorbing matter to be lower than 1. One might suppose that the X-ray spectrum is totally dominated by scattered radiation from an otherwise invisible nucleus, as observed in several “reflection-dominated” Seyfert 2s (Turner et al. 1997; Matt et al. 1997; Guainazzi et al. 1999). The observed variability, however, introduces severe constraints on this possibility. Assuming a typical variability timescale of $`\mathrm{\Delta }\mathrm{t}4\times 10^4`$ s, light crossing arguments imply a lower limit on the electron numerical density $`\mathrm{n}_\mathrm{e}\stackrel{>}{}(\mathrm{c}\mathrm{\Delta }\mathrm{t}\sigma _{\mathrm{sc}})^11.2\times 10^9`$ cm<sup>-3</sup>. If the absorption occurs in the same medium, an optical depth to scattering of the order of one implies an optical path $`N_H/n_e\stackrel{<}{}4\times 10^{12}`$ cm. Such a low value would imply that the scattering/absorption occurs in the proximity of the galactic nucleus, hardly compatible with the idea that the matter hiding the nuclear region is located at distances $``$1–100 pc, in the shape of a, more or less homogeneous, azimuthally symmetric structure (Antonucci & Miller 1985; Maiolino & Rielke 1995; Greenhill et al. 1996). Finally, the appearance sometimes of a broad H<sub>α</sub> component is a definite sign in favor of a Seyfert 1 classification. One clue to explain the observed SED is to assume that the AGN is a weak Seyfert 1, which is over-luminous in IR in comparison to the objects of the same class of comparable X-ray luminosity. This interpretation can be naturally linked to the puzzling absorption pattern emerging from the optical/UV spectroscopy. Very broad Ly<sub>α</sub> and H<sub>α</sub> lines have been observed in (not simultaneous) observations of the core of IRAS12393+3520 (M96). All these evidences suggest that the geometrical distribution of the circumnuclear neutral absorbing matter is far more complex than the simple equatorially symmetric pattern, which is assumed in the 0-th order Seyfert unification theories (Antonucci & Miller 1985; Antonucci 1993). The Narrow Line Region (NLR) could be absorbed by dusty, optically thick matter distributed on spatial scale of a few hundreds of parsec, which does not (not always?) intercept the line of sight towards the nuclear environment. This dust might simultaneously be responsible for the IR emission through reprocessing of the incoming nuclear radiation during particularly active phases (or when the line of sight between the nucleus and the dust is unobscured). In the same framework, the under-luminosity in $`\mathrm{L}_{[\mathrm{OIII}]}`$ and the relative over-luminosity in IR in comparison to the directly observed X-ray of nuclear origin can be simultaneously explained. A similar scenario had been originally suggested by Maiolino & Rieke (1995), who proposed that “intermediate” Seyfert galaxies are seen through a 100 pc-scale torus coplanar with the plane of the galaxies. This idea has received a further support after the results of a recent HST slew survey, which lead to the discovery that the distribution of matter in the nuclear environment of radio-quiet, nearby AGN is indeed highly patchy, with dusty lanes protruding from several kilo-parsecs to a few hundreds of parsecs towards the center (Malkan et al. 1998). Recently, Maiolino et al. (1999) pointed out that barred galaxies (like IRAS12393+3520) tend to exhibit the highest values of IR to X-ray luminosity ratio (see, however, a different point of view in Regan & Mulchaey (1999). This might suggest that stellar bars are very efficient in driving gas to the circumnuclear region and therefore to provide high amount of warm (AGN-heated) dust in the nuclear environment, which would reprocess the nuclear high-energy continuum. The bulk of the IR radiation could alternatively be produced by an intense star formation episode, occurring on spatial scales of the order or larger than the NLR. The SED in the IR alone is typical of starburst galaxies and does not satisfy the various criteria defined to select AGN in IRAS data (de Grijp et al. 1985; Désert & Dennefeld 1988). Also, the standard IR/radio parameter q, discussed by Condon at al. (1995), has a value of 2.85 for IRAS12393+3520, showing that the AGN, if present, is not dominating the IR emission and/or that the object is radio-weak. Finally, the observed infrared luminosity ($`\mathrm{L}_{\mathrm{IR}}5\times 10^{44}`$ erg s<sup>-1</sup>; M96) is almost two orders of magnitude higher than expected from the AGN 1–10 keV luminosity ($`\mathrm{L}_\mathrm{X}3\times 10^{42}`$ erg s<sup>-1</sup>), if the IRAS12393+3520 AGN SED has a $`\mathrm{L}_{\mathrm{IR}}/\mathrm{L}_\mathrm{X}`$ ratio typical for quasars with $`\mathrm{L}_\mathrm{X}<10^{45}`$ erg s<sup>-1</sup> ($``$4.5; Elvis et al. 1994). While both explanations for the IR emission are plausible, the energetic argument just discussed favors the hypothesis that the IR emission is dominated by star-formation. However, even in this case, an additional (nuclear) source of high-energy radiation is required to explain the hard X-rays. As shown in Fig. 8, IRAS12393+3520 exhibits indeed a much higher X-ray to IR or UV luminosity flux than typically observed in SFG galaxies. It is likely to be provided by a separate component, which is straightforward to identify, in the light of the ASCA/ROSAT results, with an active nucleus. On the other hand, a spatially and physically structured and time varying absorber is also required to yield the observed different absorptions towards the nucleus, of the BLR and of the NLR lines. The dusty environment of a star forming region could provide only the last. IRAS12393+3520 is therefore best described by a central, weak, AGN, with a BLR partly obscured by a structured absorber, and a well absorbed NLR, mixed with a region of intense star formation of perhaps larger extension. ###### Acknowledgements. MG acknowledges the receipt of an ESA Research Fellowship. The authors acknowledge discussions with F.Bocchino. This research has made use of the NASA/IPAC Extragalactic Database, which is operated by the Jet Propulsion Laboratory under contract with NASA, and of data obtained through the High Energy Astrophysics Science Archive Research Center Online Service, provided by the NASA/Goddard Space Flight Center.