This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

institutetext: Department of Mathematics, King’s College London,
The Strand, London WC2R 2LS, U.K.

From multi-gravitons to Black holes: The role of complex saddles

Alejandro Cabo-Bizet alejandro.cabo_bizet@kcl.ac.uk
Abstract

By applying the Atiyah-Bott-Berline-Vergne equivariant integration formula upon double dimensional integrals, we find a way to compute the matrix integral representations of 4d4d 𝒩=1\mathcal{N}=1 superconformal indices. The final formula allows us to easily extract analytic results in the large-rank expansion of certain theories. As an example, we compute the leading one-loop corrections to the effective action of the known complex saddles in those theories. For a superconformal index of SU(N)SU(N) 𝒩=4\mathcal{N}=4 SYM, we use the equivariant integration formula and the Picard-Lefschetz method to show that at large enough values of NN, only two, among the known complex saddles, dominate the counting of large operators i.e. of operators with charges of order N2N^{2}. Contributions from other known complex saddles are present, but we show they are exponentially suppressed in that range of charges; the smaller the charges the less suppressed they are, and eventually, to count small operators i.e. operators with charges smaller than N23N^{\frac{2}{3}}, like multi-gravitons, they can not be neglected.

1 Introduction and summary

Following the general principles of string theory/gauge theory conjecture Maldacena:1997re2 , and inspired by previous seminal results in the context of string theory Strominger:1996sh ; a fresh perspective has emerged in the last few years regarding the microscopic structure of supersymmetric AdSd+1AdS_{d+1} black holes in d=3d=3, starting with Benini:2015eyy , and d=4d=4, starting with Cabo-Bizet:2018ehj ; Choi:2018hmj ; Benini:2018ywd .

Regardless of which of the current approaches one chooses as tool to study the asymptotic growth of BPS operators at large rank, via the use of matrix integrals Cabo-Bizet:2019eaf ; Cabo-Bizet:2020nkr Benini:2018ywd ; Benini:2018mlo , there seems to be convincing evidence that complex eigenvalue configurations play an important role in understanding the problem. Currently, there are two approaches that work at large-NN and generic values of BPS charges. They are consistent with each other, and in important ways, also with the predictions coming from the gravitational side of the duality Cabo-Bizet:2018ehj , they both include two complex eigenvalue configurations, whose effective actions are entropy functionals dual to the Bekenstein-Hawking entropy of the corresponding AdS5AdS_{5} black holes Gutowski:2004yv ; Cvetic:2004hs ; Cvetic:2004ny ; Kunduri:2005zg ; Kunduri:2006ek ; Silva:2006xv ; Hosseini:2017mds . These two approaches are a)a) the Bethe Ansatz formula of Benini:2018ywd ; Benini:2018mlo ; Closset:2017bse , and b)b) the saddle-point analysis of Cabo-Bizet:2019eaf ; Cabo-Bizet:2020nkr 111A very interesting third perspective has been put forward recently Goldstein:2020yvj . This paper completes the approach of Cabo-Bizet:2019eaf ; Cabo-Bizet:2020nkr by promoting the saddle point approximation to an exact formula at finite values of NN. The problem that remains is that for the theories we have analysed so far, it is hard to evaluate the formula at finite values of NN. However, at large NN the formula gives a natural expansion that matches exactly the current expectations from the string theory side of the duality. 222An incomplete subset of relevant references in both gravity and field theory sides is Cassani:2019mms ; Kantor:2019lfo ; Suh:2018qyv ; Bobev:2019zmz Benini:2020gjh ; Lezcano:2019pae ; Lanir:2019abx ArabiArdehali:2019orz Murthy:2020rbd ; Agarwal:2020zwm Choi:2018vbz ; Honda:2019cio ; ArabiArdehali:2019tdm ; Kim:2019yrz ; Cabo-Bizet:2019osg ; Amariti:2019mgp Crichigno:2020ouj  Goldstein:2019gpz  Copetti:2020dil ; Agarwal:2020pol ; Agarwal:2019crm Larsen:2019oll ; Nian:2020qsk ; David:2020ems ; Melo:2020amq ; David:2020jhp ; Larsen:2020lhg . A pedagogical introduction to the topic can be found in Zaffaroni:2019dhb .

There were two issues that Cabo-Bizet:2019eaf ; Cabo-Bizet:2020nkr left open:

  1. 1.

    A precise analysis of a deformation of the original contour passing through all the complex saddles there studied (This would have justified the assumption of considering all such configurations as competing ones in determining the original integral).

  2. 2.

    An exact formula that could be used to evaluate 1/N1/N perturbative and non-perturbative corrections.

This paper solves these two issues. The natural completion of the large-NN saddle point approach initiated in Cabo-Bizet:2019eaf ; Cabo-Bizet:2020nkr is the Atiyah-Bott-Beligne-Vergne (ABBV) equivariant integration formula Duistermaat:1982vw ; 1983InMat..72..153D ; Witten:1982im ; berline1983 ; Atiyah:1984px 333We have followed as much as possible the presentation given by Atiyah and Bott in Atiyah:1984px . But there are many other useful references. An absolutely incomplete selection that we have found necessary to follow at some points, being Duistermaat:1982vw ; 1983InMat..72..153D ; Witten:1982im ; berline1983 and audin2012torus ; Cordes:1994fc ; Niemi:1994ej ; Blau:1995rs ; Pestun:2016qko ; JeffreyLectures ; Alekseev:2000fe ; Cremonesi:2014dva . The formula does not require us to assume large NN, but to evaluate effective actions and one-loop determinants we will use the large-NN expansion.

It is well known that the superconformal index can be written as a matrix integral Romelsberger:2005eg ; Kinney:2005ej . We will see that the ABBV formula can be used to compute a generic family of matrix integrals that includes the one associated to the 4d superconformal index as a particular case. For simplicity we will focus on unitary matrix integrals but the idea can be explored in other cases as well. The answer is similar to the one of a)a) in that it can be expressed as a sum over configurations that solve a Bethe Ansatz equation. 444That they compute the same quantity does not imply that they are the same formula.

To prove the formula, we will make use of the ambiguity in the choice of complex extension of the integrand of the original matrix integral, away from the real contour of integration. In appendix B, we will find that these extensions can be understood as regularizations of the divergent super-determinants in the original 4d physical problem. The elliptic extension of Cabo-Bizet:2019eaf ; Cabo-Bizet:2020nkr is one of those. The (piecewise-) meromorphic extension corresponds to another.

The second goal of this paper is to understand the competition of the many large-NN complex configurations that are known to compete, at the moment: Is there a finite/small number of saddles dominating the large NN expansion of the microcanonical index? 555For SU(N)SU(N) 𝒩=4\mathcal{N}=4 SYM the number of known fixed points is finite and bounded by a polynomial function of NN Hong:2018viz ; ArabiArdehali:2019orz Cabo-Bizet:2019eaf . The observable we will focus on, will be the logarithm of the number of operators – counted with (1)F(-1)^{F} – at a given value of a one-dimensional section in the space of conserved BPS charges in SU(N)SU(N) 𝒩=4\mathcal{N}=4 SYM. We will call it the microcanonical index. The generating function of those integer numbers is the superconformal index Romelsberger:2005eg ; Kinney:2005ej 666Early relevant work concerning the estimation of growth in the number of 1/161/16-th BPS operators in 𝒩=4\mathcal{N}=4 SYM, can be found in Janik:2007pm ; Janik2008 ; Berkooz:2006wc ; Chang:2013fba .

We will observe that for the counting of operators with charges of order N2N^{2} i.e. large operators, the competition among known saddles is dominated by only two configurations, the (1,0)(1,0) and (1,1)(1,1). For large enough values of NN the contributions from these two configurations become the same and the absolute value of the microcanonical index matches the exponential of the Bekenstein-Hawking entropy of the dual Gutowski-Reall black hole. Contributions from other (m,n)(m,n) saddles are present but they are shown to be exponentially suppressed for such counting problem. As one tries to count operators of lesser charges, the contributions from the remaining (m,n)(m,n) configurations become less suppressed. In principle the formula could be used even to count small operators of charges 𝒪(1)\mathcal{O}(1) 777As shown in Murthy:2020rbd ; Chang:2013fba ; Kinney:2005ej these small operators are in one-to-one correspond to multi-gravitons. , at leading order in large-NN approximation, but doing so with precision, would require the resummation of contributions from all saddle configurations.

Before contributions from other (m,n)(m,n) saddles start being relevant, something interesting happens for operators whose charges are not large enough, but still finite in units of N2N^{2}. For those values of charges, the ABBV formula predicts that the logarithm of the absolute value of the microcanonical index must oscillate around the exponential of the Bekenstein-Hawking entropy and that these oscillations are determined mainly by the superposition of contributions coming from (1,0)(1,0) and (1,1)(1,1) saddles. For smaller values of charges of order 0 in units of N2N^{2}, more precisely of order N2/3N^{2/3} or smaller, the two-saddles approximation becomes invalid and other (m,n)(m,n) configurations need to be considered.

The previous understanding is consistent with recent numerical results that show that for relatively large (but still finite) values of NN (NN of order  1010Murthy:2020rbd ; Agarwal:2020zwm indeed oscillations are present. Moreover, the analytic interpolation of the numerical oscillations given in reference Agarwal:2020zwm , is consistent with the theoretical understanding summarised in the previous paragraph. Specifically, 888Up to a factor of 22 that lies within the margin of systematic error that one makes after discarding subleading contirbutions to the effective action at large NN. such interpolation corresponds to the profile of the previously mentioned two-saddles approximation to the microcanonical index.

To summarise, we will focus on answering the following question: is there a finite number of configurations PP dominating the large-NN counting problem? Our answer will be:

  • For the counting of large enough operators, only two of them dominate the large-NN expansion. Every other fixed point/saddle point/root known to us at the moment is either non-contributing, or exponentially suppressed in that expansion.

  • The presence of oscillations at relatively large (but still finite) values of NN around the Bekenstein-Hawking entropy is a consequence of the sum over saddles/fixed points/roots.

  • At leading order in large-NN expansion the oscillations become negligible and the counting of large operators matches the exponential of the Bekenstein-Hawking entropy.

Our goal with this paper is not to make precise numerical predictions at subleading order in the 1N\frac{1}{N} expansion, but to understand the competition and the relevant hierarchies of charges where the nature of the competition among saddles changes. We will mainly focus on analyzing the leading order in the 1/N1/N expansion, but the ABBV formula does not require assuming large NN. To have analytic control at finite NN999Or large NN and small enough scale of charges. though, a complete classification of the set of fixed points is needed. That classification must be completed in a case-by-case basis as the form of the conditions that determine the positions of the fixed points PP’s, depend on the representation content of the theory one wishes to analyze. Unfortunately, this is an open problem even in the simplest toy example that we study i.e. a simple ensemble of 1/161/16-th BPS operators in SU(N)SU(N) 𝒩=4\mathcal{N}=4 SYM at large NN. Thus, given our current technical reach, we focus on analytically understanding the large-NN competition among known configurations, and use Picard-Lefschetz theory to identify which of them contribute, and finally, among those, which one dominates the large-NN counting. 101010The claim that large-NN counting of large operators is dominated by two complex saddles, should not be extrapolated to a statement about the large-NN asymptotics of the grand canonical index index \mathcal{I}. These are two different quantities. It is possible that the canonical index could take relatively small values at specific regions of the chemical potential τ\tau, meanwhile the underlying Hilbert space could still have an exponentially large number of states at large enough BPS charges, counted with the grading of (1)F(-1)^{F}.

In the remaining part of this introductory section we summarize the main two results of the paper. In subsection 1.1 we introduce the superconformal index and its relation to matrix integrals and elliptic hypergeometric integrals. In subsection 1.3 we explain how the ABBV formula can be used to compute relevant families of matrix integrals. In subsection 1.5 we comment on the relation to the saddle point analysis of Cabo-Bizet:2019eaf ; Cabo-Bizet:2020nkr and the Bethe Ansatz formula of Benini:2018ywd ; Benini:2018mlo ; Closset:2017bse . In subsection 1.6 we explain how the leading large-NN behaviour of a specific microcanonical index is controlled by the competition among complex saddles, and how such competition depends on how large the operators to be counted are. In subsection 1.7 we give interesting open questions that are left as future work. Towards the end we summarize the content of the main sections of the paper. The regularization of the 4d super-determinants is presented in appendix B.

1.1 Superconformal indices, matrix integrals and Elliptic Hypergeometric integrals: a lightening review

The Hamiltonian definition of the 𝒩=1\mathcal{N}=1 superconformal index on S3S_{3} Romelsberger:2005eg ; Kinney:2005ej is

(p,q)= Trphys(1)Feβ{𝒬,𝒬¯}pJ1+Q2qJ2+Q2tQg.\mathcal{I}(p,q)\,=\,\,{\text{ Tr}}_{\mathcal{H}_{\text{phys}}}\,(-1)^{F}\text{e}^{-\beta\{\mathcal{Q},\bar{\mathcal{Q}}\}}p^{J_{1}+\frac{Q}{2}}\,q^{J_{2}+\frac{Q}{2}}\,t^{Q_{g}}\,. (1)

The 𝒬\mathcal{Q} and 𝒬¯\overline{\mathcal{Q}} are complex conjugated supercharges. FF is the Fermionic number. J1J_{1} and J2J_{2} are the left and right angular momenta on S3S_{3}QQ is the U(1)U(1) R-charge. The J1,2+Q2J_{1,2}\,+\,\frac{Q}{2} are the two combinations out of J1,2J_{1,2} and QQ that commute with 𝒬\mathcal{Q} and 𝒬¯\overline{\mathcal{Q}}. The QgQ_{g} denotes a generic global charge. The p=𝐞(σ)p\,=\,{\bf e}(\sigma)q=𝐞(τ)q\,=\,{\bf e}(\tau) and t=𝐞(Δ)t\,=\,{\bf e}(\Delta) are the rapidities associated to the charges J1,2+Q2J_{1,2}\,+\,\frac{Q}{2} and QgQ_{g}, respectively.

The series coefficients of (1) in an expansion around p=q=0p=q=0 (and for t=1t=1) count the difference between bosonic and fermionic gauge invariant states in the cohomology of 𝒬\mathcal{Q}, with fixed values of energy EE. In an abuse of terminology, we will call the energy EE charge, as that is justified by the BPS relation that all relevant BPS states obey

E= 2J+Qwith2JJ1+J2.E\,=\,2J\,+\,Q\,\quad\text{with}\quad 2J\,\equiv\,J_{1}\,+\,J_{2}\,. (2)

For a theory carrying a representation RR of a gauge group GG, which we assume to be a product of unitary groups

=[dM]exp(n=1sl(pn,qn)nχR(G)(Mn)).\mathcal{I}\,=\,\int[dM]\,\exp\Bigl{(}\sum_{n=1}^{\infty}\,\frac{\mathcal{I}_{\text{sl}}(p^{n},q^{n})}{n}\,\chi_{R(G)}(M^{n})\Bigr{)}\,. (3)

The sl\mathcal{I}_{sl} denotes the same index (1) but computed after restricting the trace tr to run over gauge invariant operators that include only single fundamental fields 111111For example, supposing the theory is Lagrangian these are the fundamental fields in the free limit of such Lagrangian. , and any number of derivative insertions, instead of over the entire Hilbert space of the theory. [dM][dM] is the invariant measure of integration over GG which is the product of Haar measures of each of the simple factors in MMχR(M)\chi_{R(M)} is the character of the representation RR. If {𝐞(u)}\{{\bf e}{(u)}\} denotes the set of eigenvalues of MM then χR(U)=ρ𝐞(ρ(u))\chi_{R}(U)\,=\,\sum_{\rho}{\bf e}(\rho(u)) where {ρ}\{\rho\} is the set of weights of the representation RR. From (3) one can recover the index of a generic theory from the single-letter indices of chiral and vector multiplets

slc(t,p,q)=tpq/t(1p)(1q),slv(p,q)= 1(1pq)(1p)(1q).\begin{split}\mathcal{I}^{\text{c}}_{\text{sl}}(t,p,q)\,=\,\frac{t-pq/t}{(1-p)(1-q)}\,,\quad\mathcal{I}^{\text{v}}_{\text{sl}}(p,q)\,=\,1\,-\,\frac{(1-pq)}{(1-p)(1-q)}\,.\end{split} (4)

For example, consider a U(N)U(N) 𝒩=1\mathcal{N}=1 SCFT composed by one vector multiplet and three adjoint chiral multiplets. The vector and chiral multiplets are assumed to be charged under the U(1)RU(1)_{R} superconformal RR-charge. The chiral multiplets are assumed to carry the fundamental representation of the SU(3)SU(3) flavour symmetry group. In this case the superconformal index is a unitary matrix integral

=[dU]exp(n=1sl(t1n,t2n,t3n,pn,qn)nTrUnTrUn),sl= 1(1t1)(1t2)(1t3)(1p)(1q),\begin{split}\mathcal{I}&\,=\,\int[dU]\,\exp\Bigl{(}\sum_{n=1}^{\infty}\,\frac{\mathcal{I}_{sl}(t^{n}_{1},t^{n}_{2},t^{n}_{3},p^{n},q^{n})}{n}\,\text{Tr}U^{n}\text{Tr}U^{\dagger n}\Bigr{)}\,,\\ \mathcal{I}_{sl}&\,=\,1\,-\,\frac{(1\,-\,t_{1})(1\,-\,t_{2})(1\,-\,t_{3})}{(1-p)(1-q)}\,,\end{split} (5)

where the variables ttpp and qq are constrained by t1t2t3pq=1\frac{t_{1}t_{2}t_{3}}{pq}=1 and initially one must assume 0<|t1,2,3|,|p|,|q|<10<|t_{1,2,3}|,\,|p|,|q|\,<1 in order to guaranty this representation of \mathcal{I} to be well defined. The \mathcal{I} can be analytically continued in tt’s, pp and qq to larger regions. Interestingly, it turns out that \mathcal{I} has singularities when p=qp=q becomes a root of unit. These singularities are important in the large-NN counting of BPS operators. Their neighborhood encodes the essential information needed to extract the asymptotic form of such numbers.

After integrating over angular variables, the integral (5) can be recast in terms of integrals of products of elliptic functions Dolan:2008qi Spiridonov:2010em

=κ|ζ|=1i=1rk(G)dζiζi(i<j=1Nθell(ζiζj,p)θell(ζjζi,q))i,j=1ijNa=13Γell(ζiζjta;p,q),κ(p;p)rk(G)(q;q)rk(G)N!a=13Γell(ta;p,q),\begin{split}\mathcal{I}&\,=\,\kappa\,\oint_{|\zeta|=1}\prod_{i=1}^{\text{rk}(G)}{\frac{d\zeta_{i}}{\zeta_{i}}}\,\left(\prod_{i\,<\,j=1}^{N}\,{\theta_{\text{ell}}(\frac{\zeta_{i}}{\zeta_{j}},p)}\,{\theta_{\text{ell}}(\frac{\zeta_{j}}{\zeta_{i}},q)}\right)\,\prod_{i,j=1\atop i\neq j}^{N}\,\,{\prod_{a=1}^{3}\Gamma_{\text{ell}}(\frac{\zeta_{i}}{\zeta_{j}}\,t_{a};\,p,\,q)}\,,\,\\ \kappa&\equiv\frac{(p;p)^{\text{rk}(G)}(q;q)^{\text{rk}(G)}}{N!}\,\prod_{a=1}^{3}\Gamma_{\text{ell}}(t_{a};\,p,\,q)\,,\end{split} (6)
121212The definition of elliptic Gamma function Γell\Gamma_{\text{ell}} is given in equation (161).

where for G=U(N)G=U(N)rk(G)=N\text{rk}(G)\,=\,N, and for G=SU(N)G=SU(N)rk(G)=N1\text{rk}(G)=N-1. Moreover, for G=SU(N)G=~{}SU(N) the ξN1=i=1N1ξi\xi^{-1}_{N}=\prod_{i=1}^{N-1}\xi_{i}.  131313For generic theories similar expressions exist. From this expression and after algebraic manipulations it follows that SU(N)=(p;p)1(q;q)1U(N).\mathcal{I}_{SU(N)}\,=\,(p;p)^{-1}(q;q)^{-1}\,\mathcal{I}_{U(N)}\,. (7) In particular, this implies that the answer at leading order in large NN expansion, is the same for both gauge groups. It is useful to note that the vector multiplet contribution can also be expressed in terms of elliptic Gamma functions by using the following identities

θell(ζ,p)θell(1/ζ,q)=Γell(ζpq;p,q)Γell(1/ζpq;p,q)=1Γell(ζ;p,q)Γell(1/ζ;p,q).\theta_{\text{ell}}(\zeta,p)\theta_{\text{ell}}(1/\zeta,q)\,=\,\Gamma_{\text{ell}}(\zeta\,pq;\,p,\,q)\Gamma_{\text{ell}}(1/\zeta\,pq;\,p,\,q)\,=\,\frac{1}{\Gamma_{\text{ell}}(\zeta;\,p,\,q)\Gamma_{\text{ell}}(1/\zeta;\,p,\,q)}\,. (8)

Sometimes it is convenient to express (6) in terms of the potentials u¯\underline{u}σ\sigmaτ\tau and Δ\Delta’s

ζρ=𝐞(ρ(u)),p=𝐞(σ),q=𝐞(τ),ta=𝐞(Δ(a)+τ+σ2),\zeta_{\rho}\,=\,{\bf e}(\rho(u))\,,\,p\,=\,{\bf e}(\sigma)\,,\,q\,=\,{\bf e}(\tau)\,,\,t_{a}\,=\,{\bf e}(\Delta_{(a)}+\frac{\tau+\sigma}{2})\,, (9)

where u¯=(u1,u2,)\underline{u}=(u^{1},u^{2},\ldots) is an array that collects the NN Cartan angles. For the moment, we can consider the vector ρ\rho to be a weight of the adjoint of either U(N)U(N) or SU(N)SU(N). For the case of SU(N)SU(N) we can consider ρ\rho to be a weight of U(N)U(N) upon the imposition of the constraint uN=i=1Nuiu^{N}\,=\,-\sum^{N}_{i=1}u^{i}.

The Δ\Delta’s encode the dependence on RR-charges, and chemical potentials dual to global symmetries. They can be used to recover refined versions of the superconformal index. 141414The point at which the flavour fugacities vanish Δg=0\Delta_{g}=0 corresponds to Δ1=Δ2=Δ3=τ+σ6+\Delta_{1}=\Delta_{2}=\Delta_{3}=-\frac{\tau\,+\,\sigma}{6}+\mathbb{Z}. As pointed out in Cabo-Bizet:2019eaf for this value of the parameters there is a pole in A×BA\times B and thus in principle one would need to use the ABBV formula plus contributions coming from the non-vanishing residues. This correction seems to be essential at finite NN. However, at large-NN we argue that in principle one can compute the integral (6) before imposing the constraint (10) and then recover the constrained result after using meromorphic flow in the Δ\Delta’s. This is the approach we will follow in this paper. In forthcoming work, we will comeback to analyze the details of the contributions that come from residues. In these variables the constraint below (5) takes the form

Δ(1)+Δ(2)+Δ(3)+τ+σ2.\Delta_{(1)}\,+\,\Delta_{(2)}\,+\,\Delta_{(3)}\,+\,\frac{\tau+\sigma}{2}\,\in\,\mathbb{Z}\,. (10)

As mentioned before, for a generic theory with gauge group GG of rank rk(G)\text{rk}(G), the index can be written as elliptic hypergeometric integrals as well Dolan:2008qi Spiridonov:2010em . The generic answer can be recast in the form

=01𝑑u¯eS(u)κG01𝑑u¯αρα 0Γe(zα(u)+τ+σ2;τ,σ),κG(p;p)rk(G)(q;q)rk(G)|𝒲|ααvectorsρα=0Γe(zα(0)+τ+σ2;p,q),\begin{split}\mathcal{I}&\,=\,\int_{0}^{1}d\underline{u}\,e^{-\,S(u)}\,\equiv\,\kappa_{G}\int^{1}_{0}d\underline{u}\prod_{\alpha}\prod_{\rho_{\alpha}\,\neq\,0}\Gamma_{\text{e}}(z_{\alpha}(u)+\frac{\tau+\sigma}{2};\tau,\sigma)\,,\\ \kappa_{G}&\equiv\frac{(p;p)^{\text{rk}(G)}(q;q)^{\text{rk}(G)}}{|\mathcal{W}|}\,\prod_{\alpha\atop\alpha\neq\text{vectors}}\prod_{\rho_{\alpha}=0}\Gamma_{\text{e}}(z_{\alpha}(0)+\frac{\tau+\sigma}{2};\,p,\,q)\,,\end{split} (11)
151515The definition of elliptic Gamma functions Γe\Gamma_{e} is given in equation (161).

where the integration measure is defined as du¯i=1rk(G)duid{\underline{u}}\equiv\prod_{i=1}^{\text{rk}(G)}du_{i}. The |𝒲||\mathcal{W}| is the dimension of the corresponding Weyl group. The S(u)S(u) is a zero-dimensional analog of QFT quantum effective action, and thus we will call it effective action from now on. The label α\alpha runs over all the multiplets in the theory. We define zαρα(u)+Δαz_{\alpha}\,\equiv\,\rho_{\alpha}(u)\,+\,\Delta_{\alpha}\, where ρα\rho_{\alpha} is the vector of charges, i.e a weight, that the multiplet α\alpha carries (under the Cartan elements of GG). The complex number Δα\Delta_{\alpha} encodes both the dependence on the R-charges of α\alpha and the dependence on the set of potentials associated to global charges QgQ_{g}, as defined in (1). This parameter can be used to recover more refined versions of the superconformal index that are obtained by turning on global symmetry holonomies.

We find convenient to analyze the generic integrals like (11), without imposing constraints like (10). Namely for each multiplet α\alpha we associate a generic Δα\Delta_{\alpha} (even for the vector multiplets) and only after the evaluation of the observable one wishes to study, the Δα\Delta_{\alpha} can be fixed to specific values. For example, in the case of (6) we can work by default with Δα={Δ(1),Δ(2),Δ(3),Δ(4)}\Delta_{\alpha}\,=\,\{\Delta_{(1)},\Delta_{(2)},\Delta_{(3)},\Delta_{(4)}\} and in the very end, once a convergent result is obtained one can fix Δ(3)\Delta_{(3)} from condition (10).

1.2 Recovering the index from double periodic extensions: a first incomplete but useful trial

Let A×BA\times B be a real torus with AA and BB cycles. Let eSλe^{-S_{\lambda}} be a family of smooth complex functions in A×BA\times B labelled by a real parameter λ\lambda\,. Assume without loss of generality that λ\lambda flows from zero to one.

Refer to caption
Figure 1: A first trial to the problem: One starts with integrals of a family of exponentials of moment maps which are smooth on the torus A×BA\times B. These integrals are parameterized by a real number λ\lambda. As λ 1\lambda\,\to\,1^{-} the integrand develops a discontinuity at an AA-period and it is now well-defined and smooth in the cylinder (not in the torus). If the limit integrand is holomorphic and the height of cylinder is one, the integral over the cylinder collapses to an integral over an AA-period. Every AA-period gives the same answer.

Suppose that as λ 1\lambda\,\to\,1^{-} the eSλe^{-S_{\lambda}} develops a discontinuity at an AA-cycle of the torus. We can rephrase the previous statement by saying that in the limit λ 1\lambda\,\to\,1^{-} the domain on which SλS_{\lambda} was smooth, which used to be a torus A×BA\times B before the limit, becomes a cylinder as sketched in Figure 1. If the limit of eSλe^{-S_{\lambda}} is holomorphic, its integral over the cylinder collapses to a middle-dimensional integral over any path homologous to an AA-period. Under such conditions, as it will be clear from the explanation we will give below, the limit of the integral over the original torus equals the integral of the limit function over any of the A-cycles. 161616This last statement follows from the Lebesgue’s dominated convergence theorem (DCT) Baxter1996 . We assume integration to be defined in the Lebesgue’s sense. In summary, this theorem states that the limit of a Lebesgue integral coincides with the Lebesgue integral of the limit if the limit of the integrand is pointwise convergent, and the integral over the same domain of the absolute value of the limit is convergent. In particular, this implies that if the limit of the integrand is bound and the domain of integration is compact, the limit can always be commuted with the Lebesgue integral.

Now, think of the AA-cycle in the previous paragraph as the Cartan torus of a gauge group. Let the torus A×BA\times B be a double copy of AA. Suppose there exists a family of eSλe^{-S_{\lambda}} on A×BA\times B that flows to a holomorphic function living in the cylinder obtained by slicing A×BA\times B across an AA-cycle. If the limit of eSλe^{-S_{\lambda}} matches the integrand of the superconformal index (or the integrand of any other integrals that one wishes to compute) at one of these AA-cycles, any of them, we conclude that in such limit the integral over A×BA\times B converges to the superconformal index.

A trivial example of how an integral over a slicing of A×BA\times B collapses to a middle dimensional one over AA (assuming Im(Δ)>0\text{Im}(\Delta)>0) is

01𝑑u101𝑑u2Γe(u1+u2τ+τ+Δ;τ,τ)=01𝑑u1Γe(u1+τ+Δ;τ,τ).\int^{1}_{0}du_{1}\int^{1}_{0}du_{2}\,\Gamma_{\text{e}}(u_{1}+u_{2}\tau+\tau+\Delta;\tau,\tau)\,=\,\int^{1}_{0}du_{1}\,\Gamma_{\text{e}}(u_{1}+\tau+\Delta;\tau,\tau)\,. (12)
171717This does not mean that there is no dependence on u2u_{2} in the integrand of the left-hand side. On the contrary, such integrand does depend on u2u_{2} but still the equality holds in virtue of the Cauchy theorem.

From inspection of the position of poles of Γe\Gamma_{\text{e}} that are given in appendix A, one reaches the conclusion that there are no poles in the domain of integration of the left-hand side of (12). That implies that the integral over u2u_{2} is spurious. The next obvious example is the n=2n=2 one in the following family

01𝑑u¯101𝑑u¯2i,j=1ijnΓe(u1ij+u2ijτ+τ+Δ;τ,τ)=01𝑑u¯1i,j=1ijnΓe(u1ij+τ+Δ;τ,τ),\begin{split}\int^{1}_{0}d\underline{u}_{1}\int^{1}_{0}\,d\underline{u}_{2}\,\prod_{i,j=1\atop i\neq j}^{n}\Gamma_{\text{e}}(u_{1ij}+u_{2ij}\tau+\tau+\Delta;\tau,\tau)\\ \,=\,\int^{1}_{0}d\underline{u}_{1}\,\prod_{i,j=1\atop i\neq j}^{n}\Gamma_{\text{e}}(u_{1ij}+\tau+\Delta;\tau,\tau)\,,\end{split} (13)

where u1,2ij=u1,2iu1,2ju_{1,2ij}=u_{1,2}^{i}\,-\,u_{1,2}^{j}\,. This can be generalized further to integrals of the following kind (assuming Im(Δα)>0\text{Im}(\Delta_{\alpha})>0)

01𝑑u¯101𝑑u¯2αρα 0Γe(zα(u1+τu2)+τ;τ,τ)=01𝑑u¯1αρα 0Γe(zα(u1)+τ;τ,τ),\begin{split}\int^{1}_{0}d\underline{u}_{1}\int^{1}_{0}d\underline{u}_{2}\prod_{\alpha}&\prod_{\rho_{\alpha}\,\neq\,0}\Gamma_{\text{e}}(z_{\alpha}(u_{1}+\tau u_{2})+\tau;\tau,\tau)\\ &\,=\,\int^{1}_{0}d\underline{u}_{1}\prod_{\alpha}\prod_{\rho_{\alpha}\,\neq\,0}\Gamma_{\text{e}}(z_{\alpha}(u_{1})+\tau;\tau,\tau)\,,\end{split} (14)

with ρα\rho_{\alpha} being weights of a quiver gauge theory with matter in the fundamental, adjoint and/or bi-fundamental of say U(N)U(N) gauge nodes. The analysis for SU(N)SU(N) nodes is more subtle, but one can always obtain it from U(N)U(N) by using the identity (7).

1.3 The index from Atiyah-Bott-Berline-Vergne localization formula

What is the limitation of the previous approach?

As we will explain in due time, integrals over A×BA\times B of the smooth function eSλe^{-S_{\lambda}} can be computed by using the Atiyah-Bott-Berline-Vergne formula. The expressions take the following form

ωnn!eSλ=PeSλ(P)E(νP).\int\,\frac{\omega^{n}}{n!}\,e^{-S_{\lambda}}\,=\,\sum_{P}\,\frac{e^{-S_{\mathcal{\lambda}}(P)}}{{E}(\nu_{P})}\,. (15)

Here n=rk(G)n=\text{rk}(G) and the volume form 1n!ωn=01𝑑u¯101𝑑u¯2\int\frac{1}{n!}\omega^{n}\,=\,\int^{1}_{0}d\underline{u}_{1}\int^{1}_{0}d\underline{u}_{2}\,, is defined in terms of the two-form ω=i=1ndu1idu2i\omega\,=\,\sum_{i=1}^{n}du^{i}_{1}\wedge du^{i}_{2}. The remaining objects will be properly defined in due time. Here we just want to draw attention to a specific point.

The main claim of the previous subsection is summarized by the following equality

=limλ 1ωnn!eSλ(Plimλ 1eSλ(P)E(νP)).\mathcal{I}\,=\,\underset{\lambda\,\to\,1^{-}}{\lim}\,\int\,\frac{\omega^{n}}{n!}\,e^{-S_{\lambda}}\quad\Bigl{(}{\,\neq\,}\sum_{P}\,\underset{\lambda\,\to\,1^{-}}{\lim}\,\frac{e^{-S_{\mathcal{\lambda}}(P)}}{{E}(\nu_{P})}\Bigr{)}\,. (16)

However, as stated by the inequality in parenthesis, the limit λ 1\lambda\,\to\,1^{-} of the previous subsection does not commute with the sum over PP’s. The reason being that

limλ 1E(νP)= 0.\underset{\lambda\,\to\,1^{-}}{\lim}{E}(\nu_{P})\,=\,0\,. (17)

This can be understood before doing computations. The quantity E(νP)E(\nu_{P}), which will be properly defined below, is proportional to the square root of the determinant of the Hessian of SλS_{\lambda}, and inversely proportional to the determinant of the matrix components of the two-form ω\omega. As in that limit SλS_{\lambda} becomes holomorphic, the Hessian develops anti-holomorphic zero modes. Thus, as detω\det\omega is constant, equality (17) follows through.

The integration measure

The previous obstacle can be solved by using an appropriate integration measure, i.e by replacing the two form ω\omega by a λ\lambda-dependent one that we will denote as ωλ\omega_{\lambda}. The ωλ\omega_{\lambda} and SλS_{\lambda} will be defined in such a way that the corresponding integral will be independent of λ\lambda. Thus, for such definitions the limit λ 1\lambda\,\to\,1^{-} is well defined. Finally, from our choice of SλS_{\lambda} it follows that for specific values of chemical potentials the integral along the real contour of integration is equal to the double dimensional integral for any 0λ<10\leq\lambda<1.

In particular, this implies that the ABBV equivariant integration formula can be used to compute 4d unitary matrix integrals of the form (3) or equivalently (11).

Let us go by steps. First we introduce the following measure of integration in A×BA\times B

A×Bωλnn!01𝑑u¯101𝑑u¯2i=1n𝒪i,λ.\int_{A\times B}\frac{\omega^{n}_{\lambda}}{n!}\,\equiv\,\int_{0}^{1}d\underline{u}_{1}\int^{1}_{0}d\underline{u}_{2}\prod_{i=1}^{n}\mathcal{O}_{i,\lambda}\,. (18)

In this equation the closed two-form ωλ\omega_{\lambda} is defined as

ωλ=i=1ndu1idu2i𝒪i,λ.\omega_{\lambda}\,=\,\sum_{i=1}^{n}du^{i}_{1}\wedge du^{i}_{2}\,\mathcal{O}_{i,\lambda}\,. (19)

For 0λ<10\leq\lambda<1

𝒪i,λ=𝒪λ(u2i) 1{u2i}λu2i,\mathcal{O}_{i,\lambda}\,=\,\mathcal{O}_{\lambda}(u_{2i})\,\equiv\,1\,-\,\frac{\partial\{u_{2i}\}_{\lambda}}{\partial u_{2i}}\,, (20)

is a smooth and strictly positive function of a single real variable u2iu_{2i} that is equal to 11 at λ= 0\lambda\,=\,0 i.e.

𝒪λ(u2)> 0,𝒪λ(u2+1)=𝒪λ(u2),𝒪i,λ=0= 1.\begin{split}\mathcal{O}_{\lambda}(u_{2})\,>\,0\,,\qquad\mathcal{O}_{\lambda}(u_{2}+1)\,=\,\mathcal{O}_{\lambda}(u_{2})\,,\qquad\mathcal{O}_{i,\lambda=0}\,=\,1\,.\end{split} (21)

These properties imply that ωλ\omega_{\lambda} is simply a smooth and globally well-defined passive diffeomorphism transformation of the two-form

ω=ωλ=0=i=1ndu1idu2i.\omega\,=\,\omega_{\lambda=0}\,=\,\sum_{i=1}^{n}du^{i}_{1}\,\wedge\,du^{i}_{2}\,. (22)

The diffeomorphism is induced by the transformation u2iu2i{u2i}λu^{i}_{2}\,\mapsto\,u^{i}_{2}\,-\,\{u^{i}_{2}\}_{\lambda}. The definition and properties of the function {}λ\{\cdot\}_{\lambda} will be given in the main body of the paper, in equation (75). Here we just note that from that definition it follows that

𝒪λ(u2)λ 1mδ(u2+m),\mathcal{O}_{\lambda}(u_{2})\,\underset{\lambda\,\to\,1^{-}}{\longrightarrow}\,\sum_{m\in\mathbb{Z}}\,\delta(u_{2}\,+\,m)\,, (23)

as illustrated by the plots in figure 2. In fact, the specific details of the function {}λ\{\cdot\}_{\lambda} are irrelevant as long as we use a {}λ\{\cdot\}_{\lambda} that is periodic, smooth and that preserves the conditions (21), (22) and (23). This is consistent with the results presented in appendix B. There, it was shown that the choice of {}λ\{\cdot\}_{\lambda} corresponds to a choice of regularisation of the divergent one-loop determinants of the original 4d physical problem. Consequently different such choices should not affect the outcome of physical observables. We have chosen to work with the explicit form (75) for concreteness.

Refer to caption
Figure 2: The profile of Oλ(u)O_{\lambda}(u) that follows from the specific choice of function {}λ\{\cdot\}_{\lambda} that we have chosen to work with. The plot shows three different values of λ\lambdaOλ(u)O_{\lambda}(u) is positive for any value of 0λ<10\leq\lambda<1Oλ(u)1O_{\lambda}(u)\to 1 for small enough values of λ\lambdaOλ(u)O_{\lambda}(u) tends to the periodic Dirac-delta function centered at integer values of uu for λ 1\lambda\,\to\,1^{-}. In fact, the specific details of the function {}λ\{\cdot\}_{\lambda}, the one that defines the dependence on λ\lambda in both the action SλS_{\lambda}, and the two-form ωλ\omega_{\lambda} (through OλO_{\lambda}), are irrelevant as long as {}λ\{\cdot\}_{\lambda} is periodic, smooth and it preserves the conditions (21), (22) and (23).

From property (23) and (18) it follows that for any smooth action 𝒮\mathcal{S} on A×BA\times B

A×Bωλnn!e𝒮λ 1u2i=0i=1ndu1ie𝒮.\int_{A\times B}\,\frac{\omega^{n}_{\lambda}}{n!}\,e^{-\mathcal{S}}\,\,\underset{\lambda\,\to\,1^{-}}{\longrightarrow}\,\int_{u^{i}_{2}=0}\,\prod_{i=1}^{n}du^{i}_{1}\,e^{-\mathcal{S}}\,. (24)

If when restricted to the real contour of integration u2i=0u^{i}_{2}=0 the action 𝒮\mathcal{S} equals the effective action SS defined in equation (11) i.e.

𝒮(u)=S(u)atu2i= 0,\mathcal{S}(u)\,=\,{S}(u)\qquad\text{at}\qquad u^{i}_{2}\,=\,0\,, (25)

then the desired integral e.g. the superconformal index, is recovered via the following limit

A×Bωλnn!e𝒮λ 1.\int_{A\times B}\,\frac{\omega^{n}_{\lambda}}{n!}\,e^{-\mathcal{S}}\,\,\underset{\lambda\,\to\,1^{-}}{\longrightarrow}\,\mathcal{I}\,. (26)
The story is slightly more involved though

To guaranty smoothness of 𝒮\mathcal{S} at intermediate stages we will need to introduce a cut-off Λ\Lambda BrownLevin that eventually must be fixed to Λ= 1\Lambda\,=\,1^{-}. The limit Λ=1\Lambda=1^{-} introduces discontinuities at u2i= 0u^{i}_{2}\,=\,0. Thus, the correct version of (24) in that case is a product of semi-sums of two single dimensional contour integrals i.e.

A×Bωλnn!e𝒮λ 1i=1n(C+i+Ci)2e𝒮(u),\int_{A\times B}\,\frac{\omega^{n}_{\lambda}}{n!}\,e^{-\mathcal{S}}\,\,\underset{\lambda\,\to\,1^{-}}{\longrightarrow}\,\prod_{i=1}^{n}\frac{\Bigl{(}\int_{C^{i}_{+}}\,+\,\int_{C^{i}_{-}}\Bigr{)}}{2}\,{e^{-\,\mathcal{S}(u)}}\,, (27)

where

C+i=0, at u2i= 0+1𝑑u1i,Ci=0, at u2i= 11𝑑u1i.\int_{C^{i}_{+}}\,=\,\int^{1}_{0\,,\text{ at }u^{i}_{2}\,=\,0^{+}}du^{i}_{1}\,,\qquad\int_{C^{i}_{-}}\,=\,\int^{1}_{0\,,\text{ at }u^{i}_{2}\,=\,1^{-}}du^{i}_{1}\,. (28)

(27) comes from the fact the the periodic delta function (23) picks up 1/21/2 times the value of the corresponding integrand at u2i=0+u^{i}_{2}=0^{+} plus 1/21/2 times the value of the corresponding integrand at u2i=1u^{i}_{2}=1^{-}.

At this point the global property (e.g. away from the AA-period u2i= 0u^{i}_{2}\,=\,0) of our choice of function 𝒮\mathcal{S} plays an essential role. Thus, we choose the action mentioned in subsection 1.3𝒮=Sλ\mathcal{S}\,=\,S_{\lambda}\, where SλS_{\lambda} will be defined in equation (82). In the limit λ,Λ 1\lambda,\,\Lambda\,\to\,1^{-}

𝒮S(u),\mathcal{S}\,\to\,S(u)\,, (29)

where eS(u)e^{-S(u)} is the meromorphic extension of the integrand of (11). In that case due to the periodicity property  eS(u)=eS(u+ 2π)e^{-S(u)}\,=\,e^{-S(u\,+\,2\pi)} and Cauchy theorem, it follows that

Ci=C+i+Res^i[].\int_{C^{i}_{-}}\,\blacksquare\,=\,\int_{C^{i}_{+}}\,\blacksquare\,+\,\widehat{\text{Res}}_{i}[\blacksquare]\,. (30)

Both linear operators, C+i\int_{C^{i}_{+}} and Re^i\widehat{\text{Re}}_{i}\,, annihilate the dependence on a single complex variable uiu1i+τu2iu^{i}\,\equiv\,u^{i}_{1}\,+\,\tau\,u^{i}_{2} of the function they act upon and give a new function of the remaining ujiu^{j\,\neq\,i} complex variables. The integral fixes u2i= 0u^{i}_{2}\,=\,0 and integrates u1iu^{i}_{1} from 0 to 11. The operator Res^i\widehat{\text{Res}}_{i} when acting on a function f(ui)f(u^{i}) of the variable uiu^{i}’s, gives back a new function of the remaining variables  ujiu^{j\,\neq\,i} which equates to the sum over residues

ξ0Res[f(u(ξ))ξi,ξi=ξ0],ξi=𝐞(ui).\sum_{\xi_{0}}\text{Res}\Bigl{[}\frac{f(u(\xi))}{\xi^{i}},\,\xi^{i}\,=\,\xi_{0}\Bigr{]},\qquad\xi^{i}\,\,=\,\,{\bf e}{(u^{i})}\,. (31)

ξ0\xi_{0} runs over all the simple poles of the function f(ui(ξi)){f(u^{i}(\xi^{i}))} located in the annular domain |q|<|ξi|<1|q|<|\xi^{i}|<1. The corresponding residues can be computed systematically, but we will not do that in here. Plugging (30) into (27) and from the fact that =C+ieS(u)\mathcal{I}\,=\,\int_{C^{i}_{+}}\,e^{-S(u)} we obtain

A×Bωλnn!eSλλ 1+Residues.\int_{A\times B}\,\frac{\omega^{n}_{\lambda}}{n!}\,e^{-{S}_{\lambda}}\,\,\underset{\lambda\,\to\,1^{-}}{\longrightarrow}\,\mathcal{I}\,+\,\text{Residues}. (32)

By Residues we mean contributions that involve the application of at least one of the operators Res^i\widehat{\text{Res}}_{i}i=1,,rk(G)i=1,\ldots,\text{rk}(G). These contributions vanish if eS(u)e^{-S(u)} is a holomorphic function in the sliced A×BA\times B. We will assume there is always a chamber of Δ\Delta’s, where there is no contribution from simple poles. 181818This very same assumption was used in reaching the Bethe Ansatz formula of Benini:2018mlo . The improved integration formula includes contributions that involve the action of at least one of the residue operators defined in equation (30). We will analyze that point in forthcoming work. From now on we focus on the simplest variant of (32)

A×Bωλnn!eSλ(u)λ 1.\int_{A\times B}\,\frac{\omega^{n}_{\lambda}}{n!}\,e^{-\,S_{\lambda}(u)}\,\,\underset{\lambda\,\to\,1^{-}}{\longrightarrow}\,\mathcal{I}\,. (33)

A complete study of chambers of Δ\Delta’s where residues do contribute is left for future work.

1.4 The final formula

Let us wrap up and move forward. The relation (33) can be used to compute unitary matrix integrals like (11) or (3) with the help of the Atiyah-Bott-Berline-Vergne integration formula. The first step is to define the following sequences of integrals

λA×Bωλnn!eSλ.\mathcal{I}_{\lambda}\equiv\int_{A\times B}\,\frac{\omega^{n}_{\lambda}}{n!}\,e^{-S_{\lambda}}\,. (34)

As said earlier, the action SλS_{\lambda} will be defined in equation (82). The deformation parameter λ\lambda ranges from 0 to 11, including 0 but not 11! As we will explain in details in section 2 the sequence of integrals (34) is computed exactly by the ABBV equivariant integration formula i.e.

λ=PeSλ(P)E(νP)Pλ(P),\mathcal{I}_{\lambda}\,=\,\sum_{P}\,\frac{e^{-S_{\mathcal{\lambda}}(P)}}{{E}(\nu_{P})}\,\,\equiv\,\,\sum_{P}\,\mathcal{I}_{\lambda}(P)\,, (35)

where the PP’s are configurations of complex eigenvalues that are fixed points of the moment maps defined by the real and imaginary parts of the action SλS_{\lambda}. For 0λ< 10\,\leq\,\lambda\,<\,1 they are also saddle points (or stationary phase points) of SλS_{\lambda}.

The factor in the denominator is an equivariant Euler class that equates to

E(νP)=1(2πi)rk(G)detBi(P)u1j.{E}(\nu_{P})\,=\,\frac{1}{(2\pi{\rm i})^{\text{rk}(G)}}\det{\frac{\,\partial B^{i}(P)}{\partial u_{1}^{j}}}\,. (36)

For the examples we will analyze, half of the fixed-point (saddle-point) conditions can be written in terms of the objects Bi(u)B^{i}(u) (See definition (84)). They take the form of Bethe Ansatz equations Bi(P)=1B^{i}(P)=1. This object does not seem to be the same as the Bethe operator of Benini:2018mlo , but it is closely related. 191919A more detailed comparison will be left as future work.

We will assume G=SU(N)G=\otimes SU(N) and choose SλS_{\lambda} and ωλ\omega_{\lambda} in such a way that

dλ(P)dλ=(λ+PλP)λ(P)= 0.\frac{\text{d}\mathcal{I}_{\lambda}(P)}{\text{d}\lambda}\,=\,\Bigl{(}\frac{\partial}{\partial\lambda}\,+\,\frac{\partial P}{\partial\lambda}\cdot\nabla_{P}\Bigr{)}\,\mathcal{I}_{\lambda}(P)\,=\,0\,. (37)

Here /λ\partial/\partial\lambda means derivative with respect to λ\lambda at fixed values of PP.

For fixed points PP that are not continuously connected to other, the proof will go in two steps. First, we will show that for any such PP

λSλ(P)= 0,λE(νP)= 0.\partial_{\lambda}S_{\lambda}(P)\,=\,0\,,\qquad\partial_{\lambda}E(\nu_{P})\,=\,0\,. (38)

and thus λ(P)λ= 0\frac{\partial\mathcal{I}_{\lambda}(P)}{\partial\lambda}\,=\,0. The first equation follows from the fact that λSλ(u)\partial_{\lambda}S_{\lambda}(u) is proportional to a linear combination of fixed point conditions and thus it vanishes at any PP (not just for those with non-vanishing Hessian). The second condition follows trivially from the fact λBi(u)= 0\partial_{\lambda}B^{i}(u)\,=\,0. Second, we will show that the position of that type of fixed points PP can not depend on λ\lambda in virtue of ABBV theorem. That will complete the proof of (37) for such cases.

For a fixed point whose Hessian of the action vanishes, and thus it is continuously connected to others, the variation in λ\lambda generates motion in the corresponding connected manifold of fixed points. As the ABBV formula includes integration over such manifolds, it follows that even in that case there is no dependence on λ\lambda.

The final conclusion is that in appropriate chambers of chemical potentials  dλdλ= 0\dfrac{d\mathcal{I}_{\lambda}}{d\lambda}\,=\,0 and thus in virtue of (33)

=λ for 0λ< 1,\mathcal{I}\,=\,\mathcal{I}_{\lambda}\quad\text{ for }\quad 0\,\leq\,\lambda\,<\,1\,, (39)

where the superconformal index \mathcal{I}, was defined in (11). At last, plugging (35) into (39) we obtain

=PeSλ(P)E(νP).\mathcal{I}\,=\,\sum_{P}\,\frac{e^{-S_{\mathcal{\lambda}}(P)}}{E(\nu_{P})}\,. (40)

As just mentioned, the PP’s can be part of connected manifolds and in that case one needs to integrate over such components with the measure induced by the inclusion map of the corresponding submanifold in A×BA\times B upon the Liouville measure of A×BA\times B. We will elaborate on this later on.

Comment about complex fixed points

We focus on quiver theories with ν\nu SU(N)SU(N) gauge nodes and matter in the adjoint and bi-fundamental representations. As SλS_{\lambda} is smooth and double periodic, there will always exist complex fixed points PP carrying finite Abelian group structure. That follows from the analysis of Cabo-Bizet:2019eaf ; Cabo-Bizet:2020nkr . For instance for NN prime, only N\mathbb{Z}_{N} complex fixed points of such kind exist. They take the form

uiβ=ui=(iNN+12N)T,i= 1,,N1.u^{i\beta}\,=\,u^{i}\,=\,\Bigl{(}\frac{i}{N}\,-\,\frac{N+1}{2N}\Bigr{)}\,T\,,\qquad i\,=\,1,\ldots,N-1\,. (41)

T=mτ+nT\,=\,m\tau\,+\,n and mmnn are co-prime integers s.t. 0m<N0\leq m<N and 0n<N0\leq n<Nβ=1,,ν\beta=1,\ldots,\nu is the label of the SU(N)SU(N) gauge nodes. For large-NN the (m,n)=(1,0)(m,n)=(1,0) and (m,n)=(1,1)(m,n)=(1,1) fixed points are the ones whose on-shell action corresponds to the dual AdS5 BPS black holes. For fixed NN the number of solutions are finite. Note that these fixed points lie outside the AA-period u¯2=0\underline{u}_{2}=0 i.e they are not on the middle dimensional contour of integration that defines the superconformal index.

1.5 Relation to other approaches and questions

Elliptic extension

Our initial motivation was to strengthen the basis of a large-NN saddle-point approach of Cabo-Bizet:2019eaf ; Cabo-Bizet:2020nkr . We have found that the method put forward in those references is completed by the use of the ABBV formula at λ= 0\lambda\,=\,0. The saddle-point or stationary phase approximation turns out to be exact in this case due to equivariant integration. In this way we have shown that complex saddles do contribute to the original integral. On the way, we have fixed minor ambiguities regarding the computation of contributions coming from vector multiplets, and τ\tau-independent constant phases in the effective action.

Bethe Ansatz formula

Possibly, formula (35) and the Bethe Ansatz formula of Benini:2018mlo ; Closset:2017bse are the same. 202020As previously said, the fact fact they compute the same quantity does not mean that they are the same formula. There are three observations we should make though:

  • The Bethe operator BiB^{i} does not seem a priori to be the same one of Benini:2018mlo ; Closset:2017bse . Finite NN checks are perhaps a good way to start a more concrete comparison. We have computed the leading contribution of E(νP)E(\nu_{P}) at large-NN for any (m,n)(m,n) fixed point, and for the only case at disposal for comparison, the (1,0)(1,0), the result matches an analogous semi-analytic result obtained via the Bethe Ansatz formula GonzalezLezcano:2020yeb . That is a large-NN check. Finite NN checks from both perspectives will be addressed elsewhere.

  • The Bethe condition Bi=1B^{i}=1 is only half of the saddle point conditions. If all saddle point solutions carry Abelian group structure then the other half is automatically satisfied. However, should other solutions of the Bethe root conditions of Benini:2018mlo ; Closset:2017bse not carrying Abelian group structure exist, we do not know if they will also be fixed points of our moment maps. This is an important issue that needs to be understood.

  • A complete classification of saddle points is missing. This is the main obstacle to make checks at finite and generic values of NN212121 It would be interesting to explore if already-known integrability techniques could be of help to complete such classification Nekrasov:2009uh ; Nekrasov:2009ui .

1.6 The large-NN expansion of the microcanonical index

In the second part of the paper, the ABBV formula will be combined with Picard-Lefschetz method to study how complex eigenvalue configurations compete at leading order in the large-NN expansion of the microcanonical index. So far, that is the only case where numerical results have been reported in the literature Murthy:2020rbd ; Agarwal:2020zwm . The analysis will focus on the toy example of the case of 𝒩=4\mathcal{N}=4 SYM with σ=τ\sigma=\tau and no flavour potentials. More general cases στ\sigma\neq\tau can be analyzed by using the results in appendix B.

Before entering in details let us explain in simple terms the conclusions of our study. As we will explain below, the perturbative parameter controlling the competition among saddles will be the exponential of minus the Bekenstein-Hawking entropy of the Gutowski-Reall black holes. To ease presentation let us denote this parameter as geAH4g\,\equiv\,e^{-\frac{A_{H}}{4}}, where AHA_{H} is a function of the single independent charge carried by this BPS black hole, and which from now on we identify with the very same charge of the dual operators in the field theory side. For charges of order N2N^{2}, it turns out that g<<1g<<1 for large enough values of NN, and thus only the two saddle contributions dominate the complete sum over PP. The contributions from other complex saddles to the microcanonical index are exponentially suppressed and then can be safely truncated when counting large operators. For small enough charges, the gg starts to approach one and indeed for charges of order N23N^{\frac{2}{3}}logg\log g becomes of order one. When the charges become of order one in the large-NN expansion ((.e for multi-gravitons) the g1g\sim 1 and then every known complex saddle contributes.

Next, we introduce some notation and explain our conclusions in more detail. Further details are postponed until section 5.3.

We will find that at large NN the microcanonical index

d(𝔮)=trE(1)Fd(\mathfrak{q})\,=\,\text{tr}_{E}(-1)^{F} (42)

defined as the trace over 116\frac{1}{16}-th BPS operators with charge E=N2𝔮E=N^{2}\,\mathfrak{q} for 𝔮\mathfrak{q} finite, is defined by the competition of only two fixed points (1,0)(1,0) and (1,1)(1,1) out of the many (m,n)(m,n)’s. We should mention that the associated grand canonical index is a function of a single chemical potential τ\tau, namely

(τ)=E= 0d(𝔮)qE.\mathcal{I}(\tau)\,=\,\,\sum_{E\,=\,0}^{\infty}\,d(\mathfrak{q})\,q^{E}. (43)

This quantity can be obtained from the more general ensemble =(Δ1,Δ2,Δ3,σ,τ)\mathcal{I}\,=\,\mathcal{I}(\Delta_{1},\Delta_{2},\Delta_{3},\sigma,\tau) (See (6)) after imposing the following linear constraint among chemical potentials Δ1+Δ2+Δ3+σ+τ2=p\Delta_{1}+\Delta_{2}+\Delta_{3}+\frac{\sigma+\tau}{2}\,=\,p, pp\,\in\,\mathbb{Z}. For instance pp can be fixed to an integer that we denote as n0=±1n_{0}=\pm 1 e.g. p=n0=±1p=n_{0}=\pm 1. The ensemble to be compared versus the numerical results obtained in references Murthy:2020rbd ; Agarwal:2020zwm corresponds to fixing Δ1=Δ2=Δ3\Delta_{1}\,=\,\Delta_{2}\,=\,\Delta_{3}σ=τ\sigma\,=\,\tau and finally pp to a specific integer value. We will fix p=n0=1p\,=\,n_{0}=-1 and proceed.

The formula (35) predicts that the microcanonical index is a linear combination of contributions from each fixed point PP i.e.

d(𝔮)=PdP(𝔮).d(\mathfrak{q})\,=\,\sum_{P}d_{P}(\mathfrak{q})\,. (44)

The single contributions dPd_{P} are extracted out of the grand canonical index by a contour integral in τ\tau-plane. We denote such integral as CηC_{\eta}. For each PP, this integral can be decomposed in an integer combination of 66 Lefschetz-thimbles in τ\tau-plane. To understand whether a fixed point PP contributes or not to d(𝔮)d(\mathfrak{q}), we need to compute the intersection numbers of CηC_{\eta} with its 66 thimbels.  In that way one reaches a large-NN expansion of competing exponential terms the form

d(𝔮)=0m,n<Ngcd(m,n)= 1η(m,n)eH(m,n)(𝔮)+ei(I(m,n)(𝔮)+)+,\begin{split}d(\mathfrak{q})&\,=\,\sum_{0\,\leq\,m,\,n\,<\,N\atop\text{gcd}(m,n)\,=\,1}\,\eta_{(m,n)}\,e^{H_{(m,n)}(\mathfrak{q})\,+\,\dots}e^{{\rm i}\,(I_{(m,n)}(\mathfrak{q})\,+\,\ldots)}\,+\,\ldots\,,\end{split} (45)

where the magnitude of the (m,n)(m,n)-th exponential is controlled by the Morse function

H(m,n)(𝔮)=1mH(1,0)(𝔮)=14mAH 0,H_{(m,n)}(\mathfrak{q})\,=\,\frac{1}{m}\,H_{(1,0)}(\mathfrak{q})\,=\,\frac{1}{4m}\,A_{H}\,\geq\,0\,, (46)

and I(m,n)I_{(m,n)} are two real functions. H(m,n)H_{(m,n)} is the would-be entropy of the supposed-to-exist gravitational configuration dual to (m,n)(m,n) fixed points. 222222As computed in section 4.3 of Cabo-Bizet:2019eaf . The \ldots denote sub-leading contributions in the large-NN expansion at finite 𝔮\mathfrak{q}. It denotes also potential contributions from other large-NN fixed points which are unknown to us at the moment.

The integer intersection numbers η(m,n)\eta_{(m,n)} take three possible values

η(m,n)=±1or 0.\eta_{(m,n)}\,=\,\pm 1\quad\text{or }\quad 0\,. (47)

For instance for (m,n)=(1,0)(m,n)=(1,0) or (1,1),(1,1), η(m,n)= 1\eta_{(m,n)}\,=\,1232323The (0,1)(0,1) does not contribute at leading order in large NN, so effectively we can take η(0,1)=0\eta_{(0,1)}=0. The sufficient condition for (m,n)(0,1)(m,n)\neq(0,1) and χ1(m+n)0\chi_{1}(m+n)\neq 0 to contribute, i.e for ηm,n0\eta_{m,n}\neq 0, is that CηC_{\eta} must enclose its corresponding Cady-like point τ=nm\tau=-\frac{n}{m} Cabo-Bizet:2019eaf . For the chosen CηC_{\eta} that condition is equivalent to

2<nm< 1.-2<-\,\frac{n}{m}\,<\,1\,. (48)

In that case the intersection numbers will be non vanishing i.e. |η(m,n)|=1|\eta_{(m,n)}|=1. The (m,n)(m,n) fixed points for which (48) does not hold, have vanishing intersection number η(m,n)=0\eta_{(m,n)}=0, and thus they do not contribute to the counting. That implies that replica 242424The grand canonical index we study has periodicity ττ+3\tau\to\tau+3. This implies that the effective action of saddle (m,n)(m,n) is mapped to the effective action of (m,n+3m)(m,n+3m) by such transformation. We say then that they are replica of each other. fixed points of (1,0)(1,0) e.g. like (1,3)(1,3), do not contribute to the microcanonical index expansion defined by the contour CηC_{\eta}, as expected.

Then, at large enough values of NN with fixed and finite 𝔮=EN2=3N2\mathfrak{q}\,=\,\frac{E}{N^{2}}\,=\,\frac{\ell}{3N^{2}}, and up to exponentially suppressed contributions, the prediction coming from the ABBV formula to the microcanonical index (45) is

d(𝔮)=eH(1,0)(𝔮)+× 2cos(I(1,0)(𝔮)+)+.d(\mathfrak{q})\,=\,e^{\,H_{(1,0)}(\mathfrak{q})\,+\,\ldots}\,\times\,2\,\cos{\bigl{(}I_{(1,0)}(\mathfrak{q})\,+\,\ldots\bigr{)}}\,+\,\ldots\qquad\,. (49)

Here the \ldots denote sub-leading corrections in large-NN expansion. Note that in virtue of (46), the (m,n)(m,n) with m>1m>1 are exponentially suppressed. We will call (49) the two-saddles approximation to the microcanonical index.

The Morse function H(1,0)H_{(1,0)} coincides with the Bekenstein-Hawking entropy of the dual AdS5 black hole Gutowski:2004ez . The cosine modulation

cos(I(1,0)(𝔮)+)\cos{\bigl{(}I_{(1,0)}(\mathfrak{q})\,+\,\ldots\bigr{)}} (50)

explains the oscillations of d(3N2)d(\frac{\ell}{3N^{2}}) as a function of \ell at relatively large but still finite values of NN e.g. N10N\sim 10. These oscillations have been recently observed in the U(N)U(N) index 252525At leading order in large NN approximation the U(N)U(N) and SU(N)SU(N) indices are equivalent. Differences come in at finite values of NN. The two saddle approximation can be used to approximate the microcanonical index of both, U(N)U(N) and SU(N)SU(N), finite values of NN e.g. N10N\sim 10. The approximation should be even better for larger values of the charge \ell\,, or equivalently for finite values of 𝔮=3N2\mathfrak{q}=\frac{\ell}{3N^{2}}. This is consistent with the outcome of the fit made in Agarwal:2020zwm to their numerical results Agarwal:2020zwm ; Murthy:2020rbd . The reason is that the small parameter that controls the order of the corrections coming from other fixed points is eH(1,0)(𝔮)e^{-H_{(1,0)}(\mathfrak{q})}; and this parameter increases when 𝔮\mathfrak{q} decreases, at fixed NN. In particular it approaches eπN2𝔮3/21e^{-\pi N^{2}\mathfrak{q}^{3/2}}\,\sim 1 when 𝔮0\mathfrak{q}\to 0. Roughly speaking, for values of charges \ell (resp. 𝔮\mathfrak{q}) smaller enough than N2/3N^{2/3} (resp. 1N43\frac{1}{N^{\frac{4}{3}}}) we need to start considering contributions from other fixed points in order to increase precision. Implementing this procedure to refine the counting lies beyond the scope of this paper. Here we always assume that charges are large enough in such a way that the two-fixed point approximation to the oscillations is a good approximation (e.g. 𝔮\mathfrak{q} finite at large NN). via explicit evaluation of the Fourier coefficients up to rank N=10N=10 Murthy:2020rbd ; Agarwal:2020zwm .

Here we claim that oscillations around the Bekenstein-Hawking curve, like the example (50) are an interference effect produced by the superposition of contributions of different complex fixed points of the matrix integral that include the (m,n)(m,n) fixed points. However, for large enough value of NN, but not too large e.g. N10N\sim 10, the main contribution comes from the couple of fixed points (1,0)(1,0) and (1,1)(1,1). This is, we claim that the oscillations are a consequence of

  • the grand canonical index (τ)\mathcal{I}(\tau) having an exact decomposition as a sum (40) over fixed points PP262626Note that (τ)\mathcal{I}(\tau) depends only on a single parameter, τ\tau.

In the region of charges where the two-fixed point approximation to the oscillations can be trusted, we note that at large enough values of NN the oscillatory pattern becomes negligible, and the curve d(𝔮)d(\mathfrak{q}) approaches the profile of the Bekenstein-Hawking entropy of the Gutowski-Reall solution Gutowski:2004ez ; Gutowski:2004yv as shown in the plot in figure 3.

Refer to caption
Figure 3: In the figure in the top we have plotted values of the two-saddles approximation to the graded number of states d(𝔮)d(\mathfrak{q}) for N=100N=100, and the Bekenstein-Hawking entropy of the dual black holes. In this plot q=𝔮q\,=\,\mathfrak{q}. In this scale, both plots are superposed and both are the same. In this scale of charges and entropy, the two saddles oscillations are negligible. For smaller values of NN oscillation become noticeable as shown “experimentally” in Agarwal:2020zwm Murthy:2020rbd ) and as analytically predicted by the two-saddles approximation. Something interesting happens at the point q=0q=0. This point encodes information about small operators, i.e roughly speaking, operators with charges of order N23N^{\frac{2}{3}} or less. If we zoom in close enough into the point q=0q=0, oscillations start to enhance, first caused by the interference among contributions coming from (1,0)(1,0) and (1,1)(1,1), and at some point contributions from generic (m,n)(m,n) saddles start to become important and can not be neglected. In principle, that flow evolution can be exactly probed (at least the one driven by the competition among (m,n)(m,n) fixed points) with the help of the ABBV formula, but implementing that with numerical precision lies beyond the scope of the present paper and it is left for future work.

1.7 Final comments

Before entering in technical details, let us share few questions/problems that we plan to study in the future:

  • 1.

    To check the ABBV formula for SU(N)SU(N) 𝒩=4\mathcal{N}=4 SYM against numerical data for the lower rank N=2, 3,N=2,\,3,\ldots.

  • 2.

    As mentioned in the caption of figure 3, in principle, for SU(N)SU(N) 𝒩=4\mathcal{N}=4 SYM, the flow evolution from large to small charges EE, always at large NN, can be exactly probed (at least the approximation to the exact flow driven by the competition among (m,n)(m,n) fixed points) with the help of the ABBV formula. It would be extremely interesting to implement this with numerical precision and to compare against numerical data.

  • 3.

    To understand how this approach relates to the more orthodox large-NN methods Brezin:1977sv ; Marino:2015yie 272727A fresh revision of many related topics have been recently given in Anninos:2020ccj . Is the continuum limit of the fixed points PP related to large-NN cuts or not? Can one understand the large-NN expansion of our approach from within the frame of loop-equations, WKB expansions and/or resurgence? Answers to these questions could uncover interesting observations. 282828Interesting approaches to the analysis of unitary matrix models of the kind addressed in this paper, have been recently put forward in GonzalezLezcano:2020yeb ; Copetti:2020dil . They suggest points of contact between classical large-NN saddle point approaches and the one presented in this paper.

  • 4.

    Do these techniques give a hint on how to construct an explicit and tractable representation of the Hilbert space of 116\frac{1}{16}-th BPS operators (in 𝒩=4\mathcal{N}=4 SYM language) in a large-NN expansion? The order zero problem would be to try to understand the form of the large operators (in terms of fundamental letters) that dominate the counting.

  • 5.

    The exact solvability of this subsector of SU(N)SU(N) 𝒩=4\mathcal{N}=4 SYM, 292929In relation with point 4.4., similar exact formulas, to the one here presented, can be analogously found for other BPS observables. is indicative of the existence of an analogous subsector in the string theory side of the duality. It would be very interesting if such subsector could be explicitly identified.

  • 6.

    It would be very interesting to explore how this sector of large BPS operators embeds and interact with non-BPS operators in the Hilbert space of weakly coupled 𝒩=4\mathcal{N}=4 SYM. This will allow us to probe how the complete physical system behaves when perturbed away from zero temperature and supersymmetry.

This paper is organized as follows. In section 2 we introduce the ABBV equivariant integration formula. In section 3 contact is made with the superconformal index. In section 4 we re-compute the effective actions of (m,n)(m,n) fixed point and in the way fix previous ambiguities  Cabo-Bizet:2019eaf ; Cabo-Bizet:2020nkr associated to vector multiplet contributions. In the second part of that section, in subsection 4.1, we compute the equivariant Euler class of generic (m,n)(m,n) fixed points at leading order in large-NN expansion and compare against a known result in the simplest ensemble of SU(N)SU(N) 𝒩=4\mathcal{N}=4 SYM for the of (1,0)(1,0) fixed point. In section 5 the ABBV formula and Picard-Lefschetz method are combined to analyze the competition among fixed points. The focus is on the simplest ensemble of SU(N)SU(N) 𝒩=4\mathcal{N}=4 SYM that captures the growth of states of interest. In appendix B we revisit the issue of regularization of the supersymmetric partition function Cabo-Bizet:2018ehj and show that there exists a scheme on which the latter matches the superconformal index. Although we mainly work in the case σ=τ\sigma\,=\,\tau that condition can be relaxed. In appendix B.2 we show how the results obtained in this paper, can be easily generalized to infinitely many other cases στ\sigma\,\neq\,\tau. In appendix D the deformed action SλS_{\lambda} that will be used to apply the equivariant localization formula, is shown to correspond to a particular choice of regularization scheme. Other supporting material is relegated to the remaining appendices.

2 Equivariant integration on a real torus

This section introduces the Atiyah-Bott-Berline-Vergne (ABBV) equivariant integration formula. This formula can be used, for instance, to compute integrals over symplectic manifolds. The relevant integrands can contain exponentials of a linear combination of the so-called moment maps faf^{a}. These are real analytic functions whose gradients generate isometries and preserve the symplectic structure of the manifold in question. The original integral localises to lower dimensional ones over sub-manifolds that are left invariant by the flows generated by the faf^{a}’s. As pointed out by Atiyah and Bott in Atiyah:1984px , one convenient feature of the formula is that the linear combination of faf^{a}’s can be complex. That will be advantageous to our purposes.

2.1 The moment maps and the equivariant Euler class

Let AA be the Cartan torus of the gauge group GG. Let =A×B\mathcal{M}\,=\,A\times B denote a real torus of rank 2n 2rk(G)2n\,\equiv\,2\text{rk}(G). The original Cartan torus is a middle dimensional cycle AA in A×BA\times BA×BA\times B can be thought as a product of nn two-tori with labels i=1,,ni=1,\ldots,n i.e a double copy of a period AA with real angles

u1iu1i+ 1,u2iu2i+ 1.u^{i}_{1}\,\sim\,{u}^{i}_{1}\,+\,1\,,\qquad{u}^{i}_{2}\,\sim\,{u}^{i}_{2}\,+\,{1}\,. (51)

We find convenient to define the following symplectic 2-form on A×BA\times B

ωλi=1ndu1idu2i𝒪i,λ(u2i)=i=1nduidvi𝒪i,λ(vi),\omega_{\lambda}\,\equiv\,\sum_{i=1}^{n}du_{1}^{i}\,\wedge\,du_{2}^{i}\,\mathcal{O}_{i,\lambda}(u_{2}^{i})\,=\,\sum_{i=1}^{n}du^{i}\,\wedge\,dv^{i}\,\mathcal{O}_{i,\lambda}(v^{i})\,, (52)

with complex twisted differentials

dudu1+τdu2,dvdu2.du\,\equiv\,du_{1}\,+\,\tau du_{2}\,,\qquad dv\,\equiv\,du_{2}\,. (53)

This 2-form allows to define integrals via the corresponding Liouville measure 1n!ωλn\frac{1}{n!}\,\omega_{\lambda}^{n}\,,

A×Bωλnn!e𝒮01𝑑u¯101𝑑u¯2i=1n𝒪i,λe𝒮.\int_{A\times B}\,\frac{\omega_{\lambda}^{n}}{n!}\,e^{-\,\mathcal{S}}\,\equiv\,\int_{0}^{1}\,d\underline{u}_{1}\int^{1}_{0}\,d\underline{u}_{2}\,\prod_{i=1}^{n}\,\mathcal{O}_{i,\lambda}\,e^{-\mathcal{S}}\,. (54)

For any value of λ\lambda01𝑑u2i𝒪iλ= 1\int^{1}_{0}du^{i}_{2}\,\mathcal{O}_{i\,\lambda}\,=\,1.

Let 𝒮\mathcal{S} be an arbitrary linear combination of dd functions i.e. 𝒮a=1ditafa\mathcal{S}\equiv\sum_{a=1}^{d}{\rm i}t_{a}f^{a}\,. The fa:f^{a}:\mathcal{M}\mapsto\mathbb{R} with a= 1,,da\,=\,1,\ldots,d\,, are smooth and real functions on \mathcal{M}. The functions faf^{a} are called moment maps of a dd-torus action on \mathcal{M}, if there exists a set of globally defined dd vector fields VaV^{a} on \mathcal{M} that solve the following set of equations

ı(Va)ωλ=dfa.{\imath}(V^{a})\,\omega_{\lambda}\,=\,df^{a}\,. (55)

In that case, a particularization of the Atiyah-Bott-Beligne-Vergne (ABBV) integration formula Atiyah:1984px states that for tat^{a}\in\mathbb{C}

ωλnn!eiatafa=PPıP(ωλnn!)eiatafa(P)E(νP).\begin{split}\int_{\mathcal{M}}\frac{\omega_{\lambda}^{n}}{n!}\,e^{-\,{\rm i}\sum_{a}t_{a}\,f^{a}}\,=\,\sum_{P}\,\int_{P}\,\imath_{P}^{*}\Bigl{(}\frac{\omega_{\lambda}^{n}}{n!}\Bigr{)}\,\frac{e^{-\,{\rm i}\sum_{a}t_{a}f^{a}(P)}}{E(\nu_{P})}\,.\end{split} (56)

Usually, the parameters tat^{a} are assumed to be pure imaginary numbers, and thus the action in the exponent is a real function, a Morse function. That it is possible to use complex tat^{a}’s, was noticed by Atiyah and Bott in their seminal work Atiyah:1984px . Here we build upon their suggestion (See footnote 32 below).

The PP’s are connected sub-manifolds of \mathcal{M} that solve the fixed-point conditions

ı(Va)ωλ=dfa= 0,a= 1,,d.\imath(V^{a})\,\omega_{\lambda}\,=\,df^{a}\,=\,0\,,\qquad a\,=\,1,\ldots,d\,. (57)

The vector fields VaV^{a} generate isometries of ωλ\omega_{\lambda}. That can be shown by using Cartan’s formula upon the Lie derivative along VaV^{a} of ωλ\omega_{\lambda}, and the defining equation (55) i.e.

Vaωλ=i(Va)dωλ+d(i(Va)ωλ)=d(i(Va)ωλ)=d2fa= 0.\begin{split}\mathcal{L}_{V^{a}}\,\omega_{\lambda}&\,=\,i(V^{a})d\omega_{\lambda}\,+\,d(i(V^{a})\omega_{\lambda})\\ &\,=\,d(i(V^{a})\,\omega_{\lambda})\,=\,d^{2}f^{a}\,=\,0\,.\end{split} (58)

The measure ıP(ωλnn!)\imath_{P}^{*}(\frac{\omega_{\lambda}^{n}}{n!}) is the pullback induced by the inclusion map ıP:P\imath_{P}:P\hookrightarrow\mathcal{M} over the Liouville measure. Let us assume the PP’s to be flat submanifolds of \mathcal{M}. The E(νP)E(\nu_{P}) is the equivariant Euler class of the normal bundle νP\nu_{P} to PP in \mathcal{M}. Let the real co-dimension of PP be 2k2k then Niemi:1994ej ; Blau:1995rs

E(νP)=1(2π)kdet(μν(p0))det((ıPωλ)μν),p0P,E(\nu_{P})\,=\,\frac{1}{(2\pi)^{k}}\,\frac{\sqrt{\det\Bigl{(}\mathcal{H}_{\mu\nu}(p_{0})\Bigr{)}}}{\sqrt{\det\Bigl{(}(\imath_{P}^{*}\omega_{\lambda})_{\mu\nu}\Bigr{)}}}\,,\qquad p_{0}\in P\,, (59)

where the greek indices μ\mu and ν= 1,,2k\nu\,=\,1,\ldots,2k label the coordinates normal to PP on \mathcal{M} with a=1,,ka=1,\ldots,k. We note that for a complex choice of equivariant parameters tat_{a}’s the Hessian needs not to be real. 303030We obtained this expression by following the analysis in section 3 of Niemi:1994ej but implementing the new feature that we are working with complex equivariant parameters and thus complex actions.

(ıPω)μν(\imath_{P}^{*}\omega)_{\mu\nu} stands for the components of the two-form ıPω\imath_{P}^{*}\omega. The Hessian matrix \mathcal{H} is defined as

=(2𝒮uaub2𝒮uavb2𝒮vaub2𝒮vavb),a,b=1,,k,\mathcal{H}\,=\,\left(\begin{array}[]{r@{}c|c@{}l}&\mbox{$\frac{\partial^{2}{\mathcal{S}}}{\partial u^{a}\partial{u}^{b}}$}&\mbox{$\frac{\partial^{2}{\mathcal{S}}}{\partial u^{a}\partial{v}^{b}}$}\\ \hline\cr&\mbox{$\frac{\partial^{2}{\mathcal{S}}}{\partial v^{a}\partial{u}^{b}}$}&\mbox{$\frac{\partial^{2}{\mathcal{S}}}{\partial v^{a}\partial{v}^{b}}$}\end{array}\right)\,,\qquad a,b=1\,,\ldots,k\,, (60)

where the twisted derivatives are

u¯u¯1,v¯u¯2τu¯1.\partial_{\underline{u}}\,\equiv\,\partial_{\underline{u}_{1}}\,,\qquad\partial_{\underline{v}}\,\equiv\,\partial_{\underline{u}_{2}}-\tau\,\partial_{\underline{u}_{1}}\,. (61)

This twisted derivatives need not to be interpreted as coming from a complex change of variables in A×BA\times B. The coordinate will always be assumed to be u1iu^{i}_{1} and u2iu^{i}_{2}.

It is not rare to find continuous fixed sub-manifolds and not just isolated fixed points PP. For example, for U(N)U(N) 𝒩=4\mathcal{N}=4 SYM the fixed manifolds are not points but two dimensional real surfaces. To avoid this extra technical involvement only SU(N)SU(N) gauge groups will be analyzed. 313131The U(N)U(N) case requires a minor redefinition of the closed 22-from  ωλ\omega_{\lambda} in (52) that we will not try to obtain in this paper.

2.2 The fixed point conditions

Let us focus on the case of a two-torus action i.e fix d=2d=2. Fix the moment maps faf^{a} to be the real and imaginary parts of the action 𝒮\mathcal{S} and the equivariant parameters accordingly i.e.

f1=Re𝒮,f2=Im𝒮,it1=t2=1.f^{1}\,=\text{Re}\,\mathcal{S}\,,\qquad f^{2}\,=\,\text{Im}\,\mathcal{S}\,,\qquad{\rm i}t_{1}=t_{2}=1\,. (62)

As announced before our choice of equivariant parameters is complex. 323232As pointed out in Atiyah:1984px in page 15 in the paragraph below the comment numbered as 2). For the case of a 2-torus action, the component uau_{a}\in\mathbb{C} in the ll-array mentioned in that paragraph (in this case l=2l=2), corresponds to the ita{\rm i}t_{a} in here.

From this choice of moment maps and equivariant parameters tat_{a} in (62), it is immediate to see that the fixed point conditions (57) are the saddle-point condition of the corresponding action 𝒮\mathcal{S} i.e. u¯1𝒮=u¯2𝒮=0\partial_{\underline{u}_{1}}\mathcal{S}=\partial_{\underline{u}_{2}}\mathcal{S}=0. Or equivalently

u¯𝒮= 0,v¯𝒮= 0,\partial_{\underline{u}}\,\mathcal{S}\,=\,0\,,\qquad\partial_{\underline{v}}\,\mathcal{S}\,=\,0\,, (63)

when written in terms of the twisted derivatives u¯\partial_{\underline{u}} and v¯\partial_{\underline{v}}. This form of the equations will simplify the analysis later on.

2.3 Examples of fixed points: quiver gauge theories

Let 𝒮\mathcal{S} be a smooth and double periodic extension to the double torus A×BA\times B, of the effective action S(u)S(u) in (11).

Any such 𝒮\mathcal{S}, with the spectrum of gauge charges ρ\rho of the adjoint and/or bi-fundamental of SU(N)SU(N) nodes, contains a simple set of stationary configurations that carries a finite Abelian group structure of rank NN. That was shown in Cabo-Bizet:2019eaf ; Cabo-Bizet:2020nkr . See section 66 of that reference for a proof. 333333For finite NN there are a finite number of these fixed points. For instance for solutions with N\mathbb{Z}_{N} structure, which are those labelled by a couple of integers (m,n)(m,n) with gcd(m,n)=1(m,n)=1, and rank N=2N=2 there are only three solutions (1,0)(1,0)(1,1)(1,1) and (0,1)(0,1). For simplicity in this paper we will always assume NN to be prime. In that case, among solutions carrying finite Abelian group representations, only those carrying N\mathbb{Z}_{N} representation exist. It is possible that there are other kind of solutions that do not carry these previously mentioned finite Abelian group structure.

Let us assume the action 𝒮\mathcal{S} to be of the form

𝒮(u)=αρα 0𝒮α(ρ(u))logκG,\mathcal{S}(u)\,=\,\sum_{\alpha}\sum_{\rho_{\alpha}\,\neq\,0}\mathcal{S}_{\alpha}(\rho(u))\,-\,\log\kappa_{G}\,, (64)

where for z=z1+τz2z\,=\,z_{1}\,+\,\tau z_{2} and the single-particle action 𝒮α(ρ(u))\mathcal{S}_{\alpha}(\rho(u)) is a smooth and double periodic complex function of the real components z1z_{1} and z2z_{2} of zα(u)=ρα(u)+Δαz_{\alpha}(u)=\rho_{\alpha}(u)+\Delta_{\alpha}, and ρα\rho_{\alpha} are weights with respect to the total gauge group G=βSUβ(N)G={\oplus}_{\beta}\,SU_{\beta}(N). The index β=1,,ν{\beta}=1,\ldots,\nu labels the gauges nodes. ρα\rho_{\alpha} is the array of ν(N1)\nu(N-1) charges under the action Cartan generators of GG. We will always assume multiplets in the the adjoint, or the bi-fundamental representations. The variable uu stands for the array of ν(N1)\nu(N-1) complexified Cartan angles i.e the complexified Cartan angles of all gauge nodes are collected in a single array of dimension ν×(N1)\nu\times(N-1) and denoted by the letter uu in this equation.

For any smooth and double periodic single particle action 𝒮α\mathcal{S}_{\alpha} the solutions to (63) include configurations

uiβ=ui=(iNN+12N)T,i= 1,,N1.u^{i\beta}\,=\,u^{i}\,=\,\Bigl{(}\frac{i}{N}\,-\,\frac{N+1}{2N}\Bigr{)}\,T\,,\qquad i\,=\,1,\ldots,N-1\,. (65)

To simplify presentation we can insert an extra auxiliary NN-th gauge variable in some intermediate equations

uN=(1N+12N)T.u^{N}\,=\,\Bigl{(}1\,-\,\frac{N+1}{2N}\Bigr{)}\,T\,. (66)

In the previous equations

T=mτ+n,0mN1,0nN1,gcd(m,n)=1.T\,=\,m\tau+n\,,\qquad~{}0\leq m\leq N-1\,,\qquad~{}0\leq n\leq N-1\,,\qquad gcd(m,n)=1\,. (67)

As the ansatz (65) is the same for every node β\beta, effectively, we can consider all weights ρα\rho_{\alpha} to be weights of the Adjoint of the diagonal SU(N)SU(N) subgroup of GG. Thus, for any such quiver gauge theory, the effective action of ansatz (65) can be written in the convenient form

ναρAdj(SU(N))ρ 0𝒮α(ρ(u))=ναi,j= 1ijN𝒮α(ijNT),\nu\,\sum_{\alpha}\sum_{\rho\,\in\,\text{Adj}(SU(N))\atop\rho\,\neq\,0}\mathcal{S}_{\alpha}\Bigl{(}\rho(u)\Bigr{)}\,\,=\,\,\nu\,\sum_{\alpha}\sum_{i,\,j\,=\,1\atop i\,\neq\,j}^{N}\mathcal{S}_{\alpha}\Bigl{(}\frac{i-j}{N}\,T\Bigr{)}\,, (68)

where all the effective weights are in the adjoint of the diagonal SU(N)GSU(N)\subset G and the index α\alpha runs over every multiplet in the quiver theory.

In the large-NN limit the set of fixed points (65) becomes a continuum i.e. iNx[0,1)\frac{i}{N}\to x\in[0,1) and iN01𝑑x\sum_{i}\to N\int^{1}_{0}dx. The effective action in that limit is computed by the Fourier averages along the period T=mτ+nT=m\tau+n

𝒮eff(m,n)=α𝒮effα(m,n)νN2α0101𝑑x𝑑y𝒮α((xy)T)=νN2α01𝑑x𝒮α(xT).\begin{split}\mathcal{S}_{\text{eff}}(m,n)&\,=\,\,\sum_{\alpha}\mathcal{S}_{\text{eff}\,\alpha}(m,n)\,\\ &\,\equiv\,\nu N^{2}\,\sum_{\alpha}\int_{0}^{1}\int_{0}^{1}dxdy\,\mathcal{S}_{\alpha}\bigl{(}(x-y)T\bigr{)}\\ &\,=\,\nu N^{2}\,\sum_{\alpha}\int_{0}^{1}dx\,\mathcal{S}_{\alpha}(xT)\,.\end{split} (69)

The index α\alpha runs over every multiplet in the theory. Note that we have not specified the form of the action 𝒮α\mathcal{S}_{\alpha} yet.

3 The superconformal index

As mentioned in section 1.3 the first step to obtain a formula for the superconformal index \mathcal{I} is to construct the action SλS_{\lambda} that together with the measure ωλ\omega_{\lambda} defines the λ\lambda-independent quantity λ\mathcal{I}_{\lambda} introduced in equation (34). The real and imaginary parts of the action SλS_{\lambda} must be smooth moment maps in A×BA\times B and thus, we can use the ABBV equivariant integration formula to compute λ\mathcal{I}_{\lambda} at any value of λ\lambda in between zero and one, including 0 and excluding 11. If by construction λ\mathcal{I}_{\lambda} reduces to \mathcal{I} in the limit λ 1\lambda\,\to\,1^{-}, and if the former happens to be independent of λ\lambda, then the ABBV formula for λ\mathcal{I}_{\lambda} is guarantied to compute \mathcal{I}343434 In specific regions of chemical potentials for which there are no poles emerging in the limit λ1\lambda\to 1^{-}(and Λ=1\Lambda=1^{-}) of the integrand of λ\mathcal{I}_{\lambda} inside the fundamental domain A×BA\times B. In case there are poles, the formula needs to be ammended as explained in section 1.3.

To simplify the presentation, at first we will focus on a single multiplet α\alpha with gauge charge vector ρα\rho_{\alpha}. Later on a theory will be assembled.

For 0λ< 10\,\leq\,\lambda\,<\,1 we define

eSλα=Γλ(zα)RΛ1(zα)Q0Λ(zα)(P0Λ(zα))z2α({u2}λ).e^{-S_{\lambda\alpha}}\,=\,\Gamma_{\lambda}(z_{\alpha})\,\equiv\,\frac{{R}_{\Lambda}^{-1}(z_{\alpha})}{Q_{0\Lambda}(z_{\alpha})}\,\Bigl{(}P_{0\Lambda}(z_{\alpha})\Bigr{)}^{\,z_{2\alpha}(\{u_{2}\}_{\lambda})}\,. (70)

Here zα=zα(u)=ρα(u)+Δαz_{\alpha}\,=\,z_{\alpha}(u)\,=\,\rho_{\alpha}(u)\,+\,\Delta_{\alpha}\,, and Δ=Δα1+τΔα2\Delta\,=\,\Delta_{\alpha 1}\,+\,\tau\Delta_{\alpha 2} with real Δα1\Delta_{\alpha 1} and Δα2\Delta_{\alpha 2}. From now on we assume 1<Δ2α< 1-1\,<\,\Delta_{2\alpha}\,<\,1\,.

The double periodic functions P0ΛP_{0\Lambda} and Q0ΛQ_{0\Lambda} are defined via the logarithms

logP0Λ(z)=logP0Λ(z1,z2)i2πm,nm0Λ|m|+|n|𝐞(nz2mz1)m(mτ+n),\log P_{0\Lambda}(z)\,=\,\log P_{0\Lambda}(z_{1},z_{2})\,\equiv-\frac{{\rm i}}{2\pi}\!{\underset{m,n\in\mathbb{Z}\atop m\neq 0}{\sum}}\;\Lambda^{|m|\,+\,|n|}\,\frac{{\bf e}(nz_{2}-mz_{1})}{m(m\tau+n)}\,, (71)

and

logQ0Λ(z)=logQ0Λ(z1,z2)14π2m,nm0Λ|m|+|n|𝐞(nz2mz1)m(mτ+n)2.\log Q_{0\Lambda}(z)\,=\,\log Q_{0\Lambda}(z_{1},z_{2})\,\equiv\,-\frac{1}{4\pi^{2}}\!{\underset{m,n\in\mathbb{Z}\atop m\neq 0}{\sum}}\;\Lambda^{|m|\,+\,|n|}\,\frac{{\bf e}(nz_{2}-mz_{1})}{m(m\tau+n)^{2}}\,. (72)

Here 0Λ< 10\,\leq\,\Lambda\,<\,1. The role of Λ\Lambda is explained in appendix F.

The pre-factor RΛ1R_{\Lambda}^{-1} has two contributions. One is a power of q=𝐞(τ)q={\bf e}(\tau) and the other a pure τ\tau-independent phase

RΛ1(z)=RΛ1(z1,z2)=qEΛ𝐞(ΦΛ2).R_{\Lambda}^{-1}(z)\,=\,R_{\Lambda}^{-1}(z_{1},z_{2})\,=\,q^{-E_{\Lambda}}\,\,{\bf e}\bigl{(}-\frac{\Phi_{\Lambda}}{2}\bigr{)}\,. (73)

Explicit definitions of EΛE_{\Lambda} and ΦΛ\Phi_{\Lambda} are given in appendix D equation (238). But for concrete purposes we will only be interested in their form at Λ= 1\Lambda\,=\,1^{-}.

From (70) we obtain the following expression for the action of a generic chiral multiplet

Sλα=logQ0Λ(zα(u))z2α({u2}λ)logP0Λ(zα(u))+ 2πiτEΛ(z2α(u))+πiΦΛ(zα(u)).\begin{split}S_{\lambda\alpha}&\,=\,\,\log Q_{0\Lambda}(z_{\alpha}(u))\,-\,z_{2\alpha}(\{u_{2}\}_{\lambda})\,\log P_{0\Lambda}(z_{\alpha}(u))\\ &\qquad\qquad+\,2\pi{\rm i}\tau E_{\Lambda}(z_{2\alpha}(u))\,+\,\pi{\rm i}\Phi_{\Lambda}(z_{\alpha}(u))\,.\end{split} (74)

We define the bracket function of a real variable xx as

{x}λB1λ({x})+12.\{x\}_{\lambda}\,\equiv\,B_{1\lambda}(\{x\})+\,\frac{1}{2}\,. (75)

It is a smooth version of the usual {x}xx\{x\}\,\equiv\,x\,-\,\lfloor x\rfloor{x}\{x\} is almost equal to limλ 1{x}λ\lim_{\lambda\,\to\,1^{-}}\{x\}_{\lambda}\,. The difference is located at integer values of xx. For such xx’s

limλ 1{x}λ={x+}+{x}2.\lim_{\lambda\,\to\,1^{-}}\{x\}_{\lambda}\,\,=\,\frac{\{x^{+}\}\,+\,\{x^{-}\}}{2}\,. (76)

We define the bracket of the vector x¯\underline{x} as the vector of the brackets of the components {x¯i}\{\underline{x}^{i}\}. The function B1λB_{1\lambda} is defined in (174).

The phase function ΦΛ(z(u))\Phi_{\Lambda}(z(u)) can be any smooth function that in the limit λ 1\lambda\,\to\,1^{-} equates to a specific discrete periodization of the phase Φ\Phi in identity (292). As mentioned before, one such an example is given in the second line of equation (238). However, for concrete purposes there is no unique such choice i.e one can choose any real ΦΛ(z(u))\Phi_{\Lambda}(z(u)) s.t. in the limit Λ1\Lambda\to 1^{-} the exponential 𝐞(ΦΛ2){\bf e}\Bigl{(}\frac{\Phi_{\Lambda}}{2}\Bigr{)} goes to

𝐞(12B1(ρ(u1)+Δ1K)(B2(ρ({u2})+Δ2+1)))𝐞(12B1(ρ(u1)+Δ1K)(B2(Δ2+1))).\begin{split}\frac{{\bf e}\Bigl{(}\frac{1}{2}\,B_{1}(\rho(u_{1})+\Delta_{1}-K)\Bigl{(}B_{2}(\lfloor\rho(\{u_{2}\})+\Delta_{2}\rfloor+1)\Bigr{)}\Bigr{)}}{\,{\bf e}\Bigl{(}\frac{1}{2}\,B_{1}(\rho(u_{1})+\Delta_{1}-K)\Bigl{(}B_{2}(\lfloor\Delta_{2}\rfloor+1)\Bigr{)}\Bigr{)}}\,.\end{split}\, (77)

For large-NN evaluation, we will take the weights ρ\rho to be the ones of the adjoint representation of U(N)U(N) which are labelled by two integers i,j=1,Ni,j=1,\ldots N and are s.t. ρ(u)uij=uiuj\rho(u)\,\equiv\,u_{ij}\,=\,u^{i}-u^{j}. Although we are interested in SU(N)SU(N), in order to obtain closed expressions for the contribution coming from the phase function ΦΛ\Phi_{\Lambda}, it will be technically convenient to effectively work with U(N)U(N) instead of SU(N)SU(N). In virtue of identity (7), the final contribution is bound to be the same at leading order in large NN expansion. Thus, to compute the large-NN contribution to the effective action of an (m,n)(m,n) fixed point out of equation (77) we will consider

ρ(u1)=u1iu1j,ρ({u2})=u2{i}{j}{u2i}{u2j}.\rho(u_{1})\,=\,u^{i}_{1}\,-\,u^{j}_{1}\,,\qquad\rho(\{u_{2}\})\,=\,u_{2\{i\}\{j\}}\,\equiv\,\{u^{i}_{2}\}-\{u^{j}_{2}\}\,. (78)

The object KK\in\mathbb{Z} is a choice of branch. It is an arbitrary piece-wise constant and integer function, which guaranty the exponent to be double periodic. When computing large-NN objects by using the continuum limit, different definitions of KK could seem to lead to different contributions to the phase of the effective actions of (m,n)(m,n) fixed points. However, coming from the discrete finite NN case, one can check that the difference between any two such choices is always an integer multiple of 2πi2\pi{\rm i}, as it is bound to be the case.  This implies that in order to compute values of the on-shell actions, we can even choose a branch KK that breaks double periodicity, and that could even depend on specific saddles. This is because the value of the imaginary part of the effective action is only determined up to an integer multiple of 2πi2\pi{\rm i}, thus we can always compute the double periodic value of the exponential by relaxing double periodicity of the exponent up to 2πi2\pi{\rm i} identifications. In due time 353535This is a trick and it will only be used in the very end, to evaluate effective actions of specific fixed points. we will use this feature for technical convenience.

Before moving on, let us summarize three relevant properties of eSλ(u)e^{-S_{\lambda}(u)} that follow from definition (70)

  • a.

    At finite 0λ<10\leq\lambda<1 and Λ\Lambda, it is smooth and double periodic, but non-meromorphic in u¯\underline{u}.

  • b.

    In the limits λ=1\lambda=1^{-} and Λ=1\Lambda=1^{-}, it becomes locally meromorphic i.e.

    Γλ(zα(u))=Γe(z1α+τz2α({u2})+τ;τ,τ).\Gamma_{\lambda}(z_{\alpha}(u))\,=\,\Gamma_{\text{e}}\bigl{(}z_{1\alpha}\,+\,\tau z_{2\alpha}(\{u_{2}\})\,+\,\tau;\,\tau,\,\tau\bigr{)}\,. (79)
  • c.

    For any 0λ<10\leq\lambda<1 and Λ= 1\Lambda\,=\,1^{-}, it matches the elliptic gamma function in the limit to the AA-cycle u2i= 0+u^{i}_{2}\,=\,0^{+} i.e at Λ=1\Lambda=1^{-}

    Γλ(z1α(u)+τz2α(0+))=Γe(z1α(u)+τΔ2α+τ;τ,τ).\Gamma_{\lambda}(z_{1\alpha}(u)\,+\,\tau z_{2\alpha}(0^{+}))\,=\,\Gamma_{\text{e}}\bigl{(}z_{1\alpha}(u)\,+\,\tau\Delta_{2\alpha}\,+\,\tau;\,\tau,\,\tau\bigr{)}\,. (80)

Property a. b. and c. were the initial assumptions needed in order to reach (33).

3.1 Half of the fixed point conditions: A Bethe Ansatz form

In this subsection we show that for the choice of action SλS_{\lambda} to be defined below, half of the fixed point conditions can be written as a Bethe Ansatz condition.

Now that we have defined SλαS_{\lambda\alpha} for a multiplet α\alpha, we can further refine the definition of λ\mathcal{I}_{\lambda} initially given in equation (34) for the case of a generic quiver theory

λ=A×Bωλnn!eSλ(u)κG01𝑑u¯101𝑑u¯2i=1n𝒪i,λαρα 0Γλ(z1α(u)+τz2α(u)).\begin{split}\mathcal{I}_{\lambda}&\,=\,\int_{A\times B}\,\frac{\omega^{n}_{\lambda}}{n!}\,e^{-\,S_{\lambda}(u)}\,\\ &\equiv\,\kappa_{G}\int_{0}^{1}d\underline{u}_{1}\int_{0}^{1}d\underline{u}_{2}\,\prod_{i=1}^{n}\,\mathcal{O}_{i,\lambda}\,\prod_{\alpha}\prod_{\rho_{\alpha}\,\neq\,0}\Gamma_{\lambda}(z_{1\alpha}(u)\,+\,\tau\,z_{2\alpha}(u))\,.\end{split} (81)

The action of the complete theory is

Sλ(u)αραSλα(u),S_{\lambda}(u)\,\equiv\,\sum_{\alpha}\sum_{\rho_{\alpha}}\,S_{\lambda\alpha}(u)\,, (82)

where z1α=ρα(u1)+Δ1αz_{1\alpha}\,=\,\rho_{\alpha}(u_{1})\,+\,\Delta_{1\alpha} and z2α=ρα(u2)+Δ2αz_{2\alpha}\,=\,\rho_{\alpha}(u_{2})\,+\,\Delta_{2\alpha}. The function Γλ(zα)=eSλα\Gamma_{\lambda}(z_{\alpha})=e^{-S_{\lambda\alpha}} was defined in (70). It is a combination of P0ΛP_{0\Lambda}, Q0ΛQ_{0\Lambda} and Polylogs.

The main goal of this subsection is to complement the explanations given in subsection 1.4. To do so let us highlight the explicit form of half of the fixed point conditions. Once Λ=1\Lambda=1^{-}, such half takes the form

vi𝒮λ=αραρi𝒪i,λlogP0Λ=1(zα(u))+c.t.=𝒪i,λlogBi+c.t.= 0.\begin{split}\partial_{v^{i}}\,\mathcal{S_{\lambda}}&\,=\,\sum_{\alpha}\sum_{\rho_{\alpha}}\rho^{i}\,\mathcal{O}_{i,\lambda}\log P_{0\Lambda=1^{-}}(z_{\alpha}(u))\,+\,\text{c.t.}\,\\ &\,=\,\mathcal{O}_{i,\lambda}\,\log B^{i}\,+\,\text{c.t.}\,=\,0\,.\end{split} (83)

Recall that viu2iτu1i\partial_{v^{i}}\,\equiv\,\partial_{u^{i}_{2}}\,-\,\tau\partial_{u^{i}_{1}}\,. In reaching (83) we have used the definition of SλαS_{\lambda\alpha} (70), the equations in appendix F.3, and also (239). The c.t. stands for contact terms contributions. The BiB^{i} stands for

BiαραP0Λ=1ραi(zα(u)).\begin{split}B^{i}&\,\equiv\,\prod_{\alpha}\,\prod_{\rho_{\alpha}}\,P^{\rho_{\alpha}^{i}}_{0\Lambda=1^{-}}(z_{\alpha}(u))\,.\end{split} (84)

The vector multiplets do not contribute to the definition of BiB^{i} and can be ignored. That can be seen even before taking the limit Λ=1\Lambda=1^{-}. It follows from property logP0Λ(z)=logP0Λ(z)\log P_{0\Lambda}(z)\,=\,\log P_{0\Lambda}(-z), and from the fact that vector multiplets carry the weights of the adjoint representation which is a real representation. For the same reason, the same happens for the Bethe Ansatz equations of Closset:2017bse ; Benini:2018mlo .

From the positivity property (21) i.e. 𝒪i,λ>0\mathcal{O}_{i,\lambda}>0 and the condition viSλ=0\partial_{v^{i}}S_{\lambda}=0, it follows that the fixed points PP must necessarily solve a Bethe Ansatz like condition

logBi(P)= 0Bi(P)= 1.\log B^{i}(P)\,=\,0\,\implies\,B^{i}(P)\,=\,1\,. (85)

Moreover, they must not intersect the positions of the contact terms c.t.. 363636In case the latter do not vanish identically after summing over the corresponding matter content, which is the case for the examples we have studied. Note that (85) is only half of the fixed point conditions.

The explicit form of these c.t. contributions can be obtained from identity (273). From the sufficient condition

i,j,αuiujΔαc.t.= 0,uN=i=1N1ui.\forall_{i\,,\,j\,,\,\alpha}\,u^{i}\,\,-u^{j}\,\neq\,\Delta_{\alpha}\,\implies\,\text{c.t.}\,=\,0\,,\,\qquad u^{N}\,=\,-\sum^{N-1}_{i=1}\,u^{i}\,. (86)

where i,j= 1,Ni\,,\,j\,=\,1\,,\,\ldots\,N\,, it follows that c.t.= 0\,=\,0 is solved by demanding the constraint to the left of the implication symbol upon any solution of (85). For the theories we analyze, and the solutions of (85) that we are aware of (See subsection 2.3), that condition can be always satisfied. 373737For instance by restricting to certain domains of the Δα\Delta_{\alpha}’s and/or by assuming NN to be prime. From now on we assume such choices and thus ignore the contact terms.

Before moving on, let us complete the proof of the first equation (38), λSλ(P)= 0\partial_{\lambda}S_{\lambda}(P)\,=\,0. Recall that λ\partial_{\lambda} in that equation means at fixed PP. This equation follows trivially from (85) as the only contribution of Sλ(P)S_{\lambda}(P) that depends explicitly on λ\lambda is a linear combination of the logarithm of (84) which by definition vanishes at any fixed point PP.

3.2 The final form of the equivariant Euler class

In this subsection we refine the definition of the equivariant Euler class associated to the action SλS_{\lambda} and the symplectic form ωλ\omega_{\lambda}.

For simplicity we focus on the generic case in which PP is a point. That means that k=nk=n. From the definition of BiB^{i} given in (84) and identity (273), it follows that for Λ=1\Lambda=1^{-} and after avoiding contact terms

2Sλvivj= 0.\frac{\partial^{2}S_{\lambda}}{\partial{v^{i}}\partial{v^{j}}}\,=\,0\,. (87)

Recall that u¯u¯1,v¯u¯2τu¯1\partial_{\underline{u}}\,\equiv\,\partial_{\underline{u}_{1}}\,,\,\partial_{\underline{v}}\,\equiv\,\partial_{\underline{u}_{2}}-\tau\,\partial_{\underline{u}_{1}}\,. Thus from the definition of the Hessian μν\mathcal{H}_{\mu\nu} of the action SλS_{\lambda} (60) and (87) it follows that only the off-diagonal blocks contribute to the corresponding determinant and thus

det(μν(p0))=±indet(2Bi(u)uj)×i=1n𝒪i,λ.\sqrt{\det\Bigl{(}\mathcal{H}_{\mu\nu}(p_{0})\Bigr{)}}\,=\,\pm\,{\rm i}^{n}\,\det\Bigl{(}\frac{\partial^{2}B^{i}(u)}{\partial u^{j}}\Bigr{)}\,\times\,\prod_{i=1}^{n}\mathcal{O}_{i,\lambda}\,. (88)
det((ıPωλ)μν)=det((ωλ)μν)=±i=1n𝒪i,λ.\sqrt{\det\Bigl{(}(\imath_{P}^{*}\omega_{\lambda})_{\mu\nu}\Bigr{)}}\,=\,\sqrt{\det\Bigl{(}(\omega_{\lambda})_{\mu\nu}\Bigr{)}}\,=\,\pm\,\prod_{i=1}^{n}\mathcal{O}_{i,\lambda}\,. (89)

From these equations and the definition of E(νP)E(\nu_{P}) in (59), it follows that the equivariant Euler class equates to

E(νP)=1(2πi)ndetBi(P)u1j.{E}(\nu_{P})\,=\,\frac{1}{(2\pi{\rm i})^{n}}\det{\frac{\,\partial B^{i}(P)}{\partial u_{1}^{j}}}\,. (90)

In this definition we have fixed an overall constant ambiguity by comparing against other results in the literature. Note that this expression does not depend on λ\lambda, thus it implies the second equation in (38).

3.3 The final integration formula: independence of λ\lambda

In this subsection we show that the integral λ\mathcal{I}_{\lambda} is independent of λ\lambda.

Again focus on Λ=1\Lambda=1^{-}. Assume that pλp_{\lambda} is a fixed point whose position depends on λ\lambda, i.e.

pλ=(u¯λ,v¯λ),pλ+δλ=(uλ+δλu(1),vλ+δλv(1)).p_{\lambda}\,=\,(\underline{u}_{\lambda},\underline{v}_{\lambda})\,,\qquad p_{\lambda+\delta\lambda}\,=\,(u_{\lambda}+\delta\lambda\,{u}^{(1)},v_{\lambda}+\delta\lambda\,{v}^{(1)})\,. (91)

It then follows that

Bi(pλ)=Bi(pλ+δλ)= 1.B^{i}(p_{\lambda})\,=\,B^{i}(p_{\lambda+\delta\lambda})\,=\,1\,. (92)

The linear variation of this equation gives

j=1nu(1)jBi(pλ)uj+v(1)jBi(pλ)vj=j=1nu(1)jBi(pλ)uj= 0,\sum_{j=1}^{n}\,u^{(1)j}\,\frac{\partial B^{i}(p_{\lambda})}{\partial{u^{j}}}+\,v^{(1)j}\,\frac{\partial B^{i}(p_{\lambda})}{\partial{v^{j}}}\,=\,\sum_{j=1}^{n}\,u^{(1)j}\,\frac{\partial B^{i}(p_{\lambda})}{\partial{u^{j}}}\,=\,0\,, (93)

The first equality follows from (273) and from the fact that only solutions of (92) which do not collide with contact terms are fixed points. The second equality is telling us that the matrix Bi(Pλ)u1j\frac{\partial B^{i}(P_{\lambda})}{\partial u_{1}^{j}} has a zero eigenvalue and thus

detBi(pλ)u1j= 0=E(νpλ).\det{\frac{\,\partial B^{i}(p_{\lambda})}{\partial u_{1}^{j}}}\,=\,0\,=\,{E}(\nu_{p_{\lambda}})\,. (94)

In virtue of ABBV theorem, this can not be true for a fixed point pλp_{\lambda} that is not smoothly connected with others. Thus we conclude that for such class, pλp_{\lambda} can not depend on λ\lambda i.e.

dpλdλ= 0.\frac{dp_{\lambda}}{d\lambda}\,=\,0\,. (95)

When pλp_{\lambda} is part of a connected set of fixed points PP, Atiyah-Bott localization theorem tells us that the matrix Bi(pλ)u1j\frac{\,\partial B^{i}(p_{\lambda})}{\partial u_{1}^{j}} can indeed have zero eigenvalues and the corresponding eigenvectors should be tangent vectors to the connected manifold PP at pλp_{\lambda}. The second equality in (93) is telling us that the variation induced by λ\lambda is constrained to move along the connected manifold of fixed points PP. But as the Atiyah-Bott prescription includes a covariant integration over such manifold, it then follows that even when pλp_{\lambda} belongs to a continuum set of fixed points PP, the associated contribution to the equivariant integral is independent of λ\lambda. Of course it would be great to check this in a concrete fixed point, but we do not have an example of this type of solution yet.

In summary, for Λ= 1\Lambda\,=\,1^{-} the integral (81) does not depend on λ\lambda, and in specific chambers in the space of potentials Δs\Delta^{\prime}s, it is by definition equal to the limit limλ 1λ\underset{\lambda\,\to\,1^{-}}{\lim}\mathcal{I}_{\lambda} (at Λ=1\Lambda=1^{-}). If we restrict attention to those specific regions in the domain of potentials Δ\Delta’s then

=λ=PeSλ(P)E(νP).\mathcal{I}\,\,=\,\mathcal{I}_{\lambda}\,=\,\sum_{P}\,\frac{e^{-S_{\lambda}(P)}}{\,E(\nu_{P})}\,. (96)

E(νP)E{(\nu_{P})} was defined in (90). We should point out that generic fixed points PP need to obey the saddle conditions of SλS_{\lambda}. The Bethe Ansatz condition (85) is only half of the latter. (m,n)(m,n) fixed points satisfy both conditions trivially, but there could be other kind of solutions not carrying Abelian group structure.

4 The leading contributions in large-NN expansion

In this subsection we revisit the evaluation of the large-NN effective action

Seff(p)=Sλ(p),S_{\text{eff}}(p)\,=\,S_{\lambda}(p)\,, (97)

when pp is an (m,n)(m,n) fixed point. In Cabo-Bizet:2019eaf ; Cabo-Bizet:2020nkr the answer for (97) was fixed up to a τ\tau independent phase. Here we constraint such ambiguity.

In the large-NN limit, we can introduce the NN-th auxiliary variable uNu^{N} that at (m,n)(m,n) saddles is defined as (66) and work with the weights of U(N)U(N) instead of with the ones of SU(N)SU(N). After substituting zz by zα(u)=uij+Δαz_{\alpha}(u)=u_{ij}\,+\,\Delta_{\alpha} with the ansatz

uij=ijN(mτ+n),u_{ij}=\frac{i-j}{N}\,(m\tau\,+\,n)\,, (98)

the discrete variables become continuum and the sums become integrals i.e.

iNx(0,1],i=1NN01𝑑x.\frac{i}{N}\rightarrow x\in(0,1]\,,\qquad\sum_{i=1}^{N}\rightarrow N\int_{0}^{1}dx\,. (99)

In the large NN for m0m\neq 0 one obtains

0101𝑑x𝑑ylogQ0Λ((xy)(mτ+n)+Δα)=πiB3Λ({mΔα1nΔα2})3m(mτ+n)2,0101𝑑x𝑑ylogP0Λ((xy)(mτ+n)+Δα)=πiB2Λ({mΔα1nΔα2})m(mτ+n),0101𝑑x𝑑y 2πiτEΛ((xy)(mτ+n)+Δα)=πiτ6(2Δα23Δα2)+.\begin{split}\int_{0}^{1}\int_{0}^{1}dxdy\,\log Q_{0\Lambda}\bigl{(}(x-y)(m\tau+n)+\Delta_{\alpha}\bigr{)}&\,=\,\frac{\pi{\rm i}\,B_{3\Lambda}(\{m\Delta_{\alpha 1}\,-\,n\Delta_{\alpha 2}\})}{3m(m\tau\,+\,n)^{2}}\,,\\ \int_{0}^{1}\int_{0}^{1}dxdy\,\log P_{0\Lambda}\bigl{(}(x-y)(m\tau+n)+\Delta_{\alpha}\bigr{)}&\,=\,-\frac{\pi{\rm i}\,B_{2\Lambda}(\{m\Delta_{\alpha 1}\,-\,n\Delta_{\alpha 2}\})}{m(m\tau\,+\,n)}\,,\\ \int_{0}^{1}\int_{0}^{1}dxdy\,2\pi{\rm i}\tau\,E_{\Lambda}\bigl{(}(x-y)(m\tau+n)+\Delta_{\alpha}\bigr{)}&\,=\,\frac{\pi{\rm i}\tau}{6}\,\bigl{(}2\Delta_{\alpha 2}^{3}\,-\,\Delta_{\alpha 2}\bigr{)}\,+\,\ldots\,.\end{split} (100)

The Polylogs in (100) become periodic Bernoulli polynomials at Λ= 1\Lambda\,=\,1^{-}. The \ldots stands for corrections that go to zero at Λ=1\Lambda=1^{-}.

From the limit (77) of ΦΛ\Phi_{\Lambda} and after making a specific choice of branch KK that we will detail below, we obtain a definite answer for the phase

limΛ1i,j= 0N1πiΦΛ(ijN(mτ+n)+Δα)=πiN2Φ(Δa).\lim_{\Lambda\to 1^{-}}\sum_{i,j\,=\,0}^{N-1}\,\pi{\rm i}\Phi_{\Lambda}\Bigl{(}\frac{i-j}{N}\,(m\tau+n)+\Delta_{\alpha}\Bigr{)}\,=\,\pi{\rm i}N^{2}\Phi(\Delta_{a})\,. (101)

Let us go step by step. For convenience we work in continuum variables xx and yy instead of discrete ones ii and jj. To ease notation, let us drop the index α\alpha for a moment. We recall that Δ=Δ1+τΔ2\Delta\,=\,\Delta_{1}\,+\,\tau\,\Delta_{2}\,. For 0<Δ2< 10\,<\,\Delta_{2}\,<\,1 we can choose (for m>0m>0)

K((xy)(mτ+n)+Δ)= 2mm(Δ1 1)n(Δ2 1)+m(m+ 1)K\Bigl{(}(x\,-\,y)\,(m\tau\,+\,n)\,+\,\Delta\Bigr{)}\,=\,2m\lfloor m(\Delta_{1}\,-\,1)\,-\,n(\Delta_{2}\,-\,1)\rfloor\,+\,m(m\,+\,1)\, (102)

for 1Δ2<(mxmy)< 1112mΔ21\,-\,\Delta_{2}\,<\,(mx\,-\,my)\,<\,{1-\sqrt{1\,-\,\frac{1}{2m}}\,\Delta_{2}} and zero otherwise. Below it will become clear why we have made this choice of KK. A similar KK can be defined for the case 1<Δ2<0-1<\Delta_{2}<0.

For 1Δ2<1-1\leq\Delta_{2}<1 we can take the definition of ΦΛ\Phi_{\Lambda} in the exponent in the second line of (77) and obtain

πiN2Φ(Δ)=πiN2Φ(Δ1,Δ2)πiN20101dxdy(B1(nxny+Δ1)K)×(B2({mx}{my}+Δ2+1)B2(Δ2+1))=πiN2(n3mΔ23+1mΔ22B1({mΔ1nΔ2})).\begin{split}\pi{\rm i}N^{2}\Phi(\Delta)&\,=\,\pi{\rm i}N^{2}\Phi(\Delta_{1},\Delta_{2})\,\\ &\equiv\,\pi{\rm i}N^{2}\,\int_{0}^{1}\int_{0}^{1}dxdy\,\bigl{(}B_{1}(nx\,-\,ny\,+\,\Delta_{1})\,-\,K\bigr{)}\times\\ &\qquad\qquad\hfill\Bigl{(}B_{2}(\lfloor\{mx\}-\{my\}\,+\,\Delta_{2}\rfloor+1)\,-\,B_{2}(\lfloor\Delta_{2}\rfloor+1)\Bigr{)}\\ &\,=\,\pi{\rm i}N^{2}\,\Bigl{(}\frac{n}{3m}\,\Delta^{3}_{2}\,+\,\frac{1}{m}\,\Delta^{2}_{2}\,B_{1}(\{m\Delta_{1}\,-\,n\Delta_{2}\})\Bigr{)}\,.\end{split} (103)

The choice of KK mentioned before was chosen in order to reach the result in the third line. That will be convenient later on.

Next, we show an useful identity. Introducing the NN-th auxiliary variable uNu1N+τu2N=i=1N1uiu^{N}\,\equiv\,u^{N}_{1}\,+\,\tau u^{N}_{2}\,=\,-\sum_{i=1}^{N-1}u^{i}, and defining its bracket as {u2N}λi= 1N1{u2i}λ\{u^{N}_{2}\}_{\lambda}\,\equiv\,-\sum_{i\,=\,1}^{N-1}\,\{u^{i}_{2}\}_{\lambda}\, it follows that

ρAdj(SU(N))ρ({u2}λ)logP0Λ(za(u))=i,j=1N({u2i}λ{u2j}λ)logP0Λ(za(u))=i,j=1N({iNm}λ{jNm}λ)logP0Λ((ij)NT+Δα)=i,j=1N{iNm}λ(logP0Λ((ij)NT+Δα)ij)=i=1N{iNm}λj=1N(logP0Λ((ij)NT+Δα)ij)=i=1N{iNm}λ×0= 0.\begin{split}\sum_{\rho\in\text{Adj}(SU(N))}\rho(\{u_{2}\}_{\lambda})\log P_{0\Lambda}(z_{a}(u))&\,=\,\sum_{i,j=1}^{N}\,(\{u^{i}_{2}\}_{\lambda}-\{u^{j}_{2}\}_{\lambda})\log P_{0\Lambda}(z_{a}(u))\\ &\,=\,\sum_{i,j=1}^{N}\,(\{\frac{i}{N}m\}_{\lambda}-\{\frac{j}{N}m\}_{\lambda})\log P_{0\Lambda}(\frac{(i-j)}{N}T+\Delta_{\alpha})\\ &\,=\,\sum_{i,j=1}^{N}\,\{\frac{i}{N}m\}_{\lambda}\,\Bigl{(}\log P_{0\Lambda}(\frac{(i-j)}{N}T+\Delta_{\alpha})\,-\,i\leftrightarrow j\Bigr{)}\,\\ &=\sum_{i=1}^{N}\,\{\frac{i}{N}m\}_{\lambda}\,\sum_{j=1}^{N}\Bigl{(}\log P_{0\Lambda}(\frac{(i-j)}{N}T+\Delta_{\alpha})\,-\,i\leftrightarrow j\Bigr{)}\,\\ &\,=\,\sum_{i=1}^{N}\,\{\frac{i}{N}m\}_{\lambda}\,\times 0\,=\,0\,.\end{split} (104)

In the last step we have used that

j=1N(logP0Λ((ij)NT+Δα)ij)\sum_{j=1}^{N}\Bigl{(}\log P_{0\Lambda}(\frac{(i-j)}{N}T+\Delta_{\alpha})\,-\,i\leftrightarrow j\Bigr{)}\, (105)

is a combination of the following vanishing quantities

j=1Nsin2πp(ij)N.\sum_{j=1}^{N}\sin{2\pi p\frac{(i-j)}{N}}\,. (106)

The integer numbers pp’s depend on mm and nn. One way to prove this analytically, is just to split the sinsin in the difference of exponentials, and then use the geometric sum i=1NXi=X(1XN)1X\sum_{i=1}^{N}X^{i}\,=\,\frac{X(1-X^{N})}{1-X} for X=𝐞(pN)1X={\bf e}(\frac{p}{N})\neq 1, then as XN=1X^{N}=1 the sum cancels. When X=1X=1 every term in the sum (106) vanishes. Note that identity (104) holds for any finite NNλ\lambda and Λ\Lambda.

For every chiral multiplet, after using (69) we can obtain a closed expression for the deformed effective action of a chiral multiplet in terms of Polylogs:

Sλα(p)N2Seffα(m,n)N2=πiB3Λ({mΔα1nΔα2})3mT2+Δα2πiB2Λ({mΔα1nΔα2})mT+πiτ6(2Δα23Δα2)+πiΦΛ.\begin{split}\frac{S_{\lambda\alpha}(p)}{N^{2}}\,\equiv\,\frac{{S}_{\text{eff}\alpha}(m,n)}{N^{2}}&\,=\,\frac{\pi{\rm i}\,B_{3\Lambda}(\{m\Delta_{\alpha 1}\,-\,n\Delta_{\alpha 2}\})}{3mT^{2}}\\ &\,+\,\Delta_{\alpha 2}\,\frac{\pi{\rm i}\,B_{2\Lambda}(\{m\Delta_{\alpha 1}\,-\,n\Delta_{\alpha 2}\})}{mT}\,\\ &\,+\,\frac{\pi{\rm i}\tau}{6}\,\,\bigl{(}2\Delta_{\alpha 2}^{3}\,-\,\Delta_{\alpha 2}\bigr{)}\,+\,\pi{\rm i}\Phi_{\Lambda}\,.\end{split} (107)

Essentially, this effective action reduces to the one computed in Cabo-Bizet:2020nkr at Λ=1\Lambda=1^{-}. This time the phase Φ\Phi is fixed up to the previously mentioned ambiguity in the choice of central term KK. See (103).

After taking the limits, we can use property Cabo-Bizet:2020nkr

B3(x+y)=B3(x)+ 3B2(x)y+ 3B1(x)y2+y3,x,y,B_{3}(x+y)\,=\,B_{3}(x)\,+\,3B_{2}(x)y\,+\,3B_{1}(x)y^{2}\,+\,y^{3}\,,\qquad x,\,y\,\in\,\mathbb{C}\,, (108)

to obtain the following form of the effective action Cabo-Bizet:2020nkr

Seffλα(m,n)=πiN23mT2[Δα]Tm([Δα]Tm+12)([Δα]Tm+1)πiτ6N2Δα2+πiN2(C+Φ)+.\begin{split}{S}_{\text{eff}\lambda\alpha}(m,n)&\,=\,\frac{\pi{\rm i}N^{2}}{3mT^{2}}\,[\Delta_{\alpha}]^{m}_{T}\,\big{(}[\Delta_{\alpha}]^{m}_{T}+\frac{1}{2}\big{)}\big{(}[\Delta_{\alpha}]^{m}_{T}+1\big{)}\\ &\,-\,\frac{\pi{\rm i}\tau}{6}\,N^{2}\,\Delta_{\alpha 2}\,+\,\pi{\rm i}N^{2}(C+\Phi)\,+\,\ldots.\end{split}

For mΔα1nΔα2m\Delta_{\alpha 1}\,-\,n\Delta_{\alpha 2}\notin\mathbb{Z}, the symbol bracket is the same as the one defined in Cabo-Bizet:2020nkr i.e.

[Δα]mTTΔα2+{mΔα1nΔα2} 1,[\Delta_{\alpha}]^{T}_{m}\,\equiv\,T\Delta_{\alpha 2}\,+\,\{m\Delta_{\alpha 1}\,-\,n\Delta_{\alpha 2}\}\,-\,1\,, (109)

for T=mτ+nT=m\tau+n\,383838To simplify the presentation, we have changed the definition of Δ\Delta relative to the one we used in Cabo-Bizet:2020nkr . The translation is Δthere=Δhere+τ\Delta_{there}=\Delta_{here}+\tau. However, in virtue of property (76), for γmΔα1nΔα2\gamma\equiv m\Delta_{\alpha 1}\,-\,n\Delta_{\alpha 2}\,\in\,\mathbb{Z} that definition needs to be slightly modified i.e for γ\gamma\in\mathbb{Z}

{γ}{γ+}+{γ}2.\{\gamma\}\,\equiv\,\frac{\{\gamma^{+}\}+\{\gamma^{-}\}}{2}\,. (110)

This follows from the fact that the corresponding discontinuities arise from limits of smooth and double-periodic Fourier expansions where uniform convergence is lost, and thus Dirichlet theorem applies (See around F below equation (269) for more information on this theorem). The constant

Cn3mΔα231mΔα22B1({γ})=Φ,C\,\equiv\,-\,\frac{n}{3m}\,\Delta^{3}_{\alpha_{2}}\,\,-\,\frac{1}{m}\,\Delta^{2}_{\alpha 2}\,B_{1}(\{\gamma\})\,=\,-\Phi\,, (111)

where the Φ\Phi was defined in (103). Thus we reach

Seffλα(m,n)=πiN23mT2[Δα]Tm([Δα]Tm+12)([Δα]Tm+1)πiτ6N2Δα2{S}_{\text{eff}\lambda\alpha}(m,n)\,=\,\frac{\pi{\rm i}N^{2}}{3mT^{2}}\,[\Delta_{\alpha}]^{m}_{T}\,\big{(}[\Delta_{\alpha}]^{m}_{T}+\frac{1}{2}\big{)}\big{(}[\Delta_{\alpha}]^{m}_{T}+1\big{)}\,-\,\frac{\pi{\rm i}\tau}{6}\,N^{2}\,\Delta_{\alpha 2}\, (112)

but with the modification (110).

Contribution from a vector multiplet

The result for a vector multiplet can be recovered from the limit Δ2α1\Delta_{2\alpha}\to 1^{-} and Δ1α0±\Delta_{1\alpha}\to 0^{\pm} of (112), by using the modified definition of (109). In this case the effective action is

Seffλv(m,n)=πiN26mT+πiτN23+πiN2n3m.{S}_{\text{eff}\lambda\text{v}}(m,n)\,=\,\frac{\pi{\rm i}N^{2}}{6mT}\,+\,\frac{\pi{\rm i}\tau N^{2}}{3}\,+\,\frac{\pi{\rm i}N^{2}n}{3m}\,. (113)

Thus, having the previous detail in mind, we can write, for a complete gauge-anomaly free theory

Seff(m,n)Sefflambda(m,n)=απiN23mT2[Δα]Tm([Δα]Tm+12)([Δα]Tm+1)πiτ6N2Δα2=απiN23mT2[Δα]Tm([Δα]Tm+12)([Δα]Tm+1).\begin{split}{S}_{\text{eff}}(m,n)\,\equiv\,{S}_{\text{eff}lambda}(m,n)&\,=\,\sum_{\alpha}\,\frac{\pi{\rm i}N^{2}}{3mT^{2}}\,[\Delta_{\alpha}]^{m}_{T}\,\big{(}[\Delta_{\alpha}]^{m}_{T}+\frac{1}{2}\big{)}\big{(}[\Delta_{\alpha}]^{m}_{T}+1\big{)}\,-\,\frac{\pi{\rm i}\tau}{6}\,N^{2}\,\Delta_{\alpha 2}\,\\ &\,=\,\sum_{\alpha}\,\frac{\pi{\rm i}N^{2}}{3mT^{2}}\,[\Delta_{\alpha}]^{m}_{T}\,\big{(}[\Delta_{\alpha}]^{m}_{T}+\frac{1}{2}\big{)}\big{(}[\Delta_{\alpha}]^{m}_{T}+1\big{)}\,.\end{split} (114)

In going to the second line we have used that

αΔ2α=TrR= 0+𝒪(1N),\sum_{\alpha}\Delta_{2\alpha}\,=\,\text{Tr}{R}\,=\,0\,+\,\mathcal{O}\Bigl{(}\frac{1}{N}\Bigr{)}, (115)

where TrR{R} is the sum of superconformal R-charges of the fundamental fermionic fields of the corresponding theory. As noticed in Cabo-Bizet:2020nkr , this quantity vanishes at large NN due to ABJ anomaly cancellation. Please refer to subsection 4.1 of Cabo-Bizet:2020nkr around equations (4.12) and (4.21) for a detailed explanation.

The simplest ensemble of 𝒩=4\mathcal{N}=4 SYM

In this case we have one vector and three chiral multiplets with Δα13(τn0)\Delta_{\alpha}\to\frac{1}{3}(-\tau-n_{0}) where we can choose n0=1n_{0}=1 or n0=1n_{0}=-1393939 As it follows from setting up N=4N=4 SYM on the conformal boundary conditions dictated by the dual BPS black hole solution Cabo-Bizet:2018ehj . We then proceed to complete the effective action computation of Cabo-Bizet:2019eaf

Seff(m,n;τ)=1mπi27N2(2T+χ1(n0m+n))3T2+N2πiφ(m,n).S_{\text{eff}}(m,n;\tau)\,=\,\frac{1}{m}\,\frac{\pi{\rm i}}{27}\,N^{2}\,\frac{\bigl{(}2T+\chi_{1}(-n_{0}m\,+\,n)\bigr{)}^{3}}{T^{2}}+N^{2}\pi{\rm i}\,\varphi(m,n)\,. (116)

For our choice of branch KK

φ(m,n)=χ1(n0m+n)2m.\varphi(m,n)\,=\,-\frac{\chi_{1}(-n_{0}m\,+\,n)}{2m}\,. (117)

This answer is unique up to an addition of τ\tau-independent real numbers that after multiplication by N2N^{2} become an even integer. From now on we will fix n0=1n_{0}=-1.

4.1 The equivariant Euler class at leading order in the large-NN expansion

In this subsection we study the equivariant Euler class E(νP)E(\nu_{P}). The goal is to compute its leading behaviour in large-NN expansion.

For the Sλ(u)S_{\lambda}(u) of a theory with a single SU(N)SU(N) node and fields in the adjoint, let us say for 𝒩=4\mathcal{N}=4 SYM (if PP is isolated),

E(νP)=1(2πi)N1deti,jujBi(P)=1(2πi)N1deti,ju1jBi(P)=1(2πi)N1deti,ju1jlogBi(P).\begin{split}E(\nu_{P})&\,=\,\frac{1}{(2\pi{\rm i})^{N-1}}\,\underset{i,j}{\det}\,\frac{\partial}{\partial u^{j}}\,B^{i}(P)\,\\ &\,=\,\frac{1}{(2\pi{\rm i})^{N-1}}\,\underset{i,j}{\det}\,\frac{\partial}{\partial u_{1}^{j}}\,B^{i}(P)\,\\ &\,=\,\frac{1}{(2\pi{\rm i})^{N-1}}\,\underset{i,j}{\det}\,\frac{\partial}{\partial u_{1}^{j}}\,\log B^{i}(P)\,.\end{split} (118)

In going to the second line we have used (61). In going to the third line we have used Bi(P)=1B^{i}(P)=1 .

Basic algebraic manipulations show that the matrix elements can be arranged as follows

u1jlogBi(u)=α(k= 1kiN1α(ujk)δij+k= 1kikjN1α(uNk))+.\partial_{u^{j}_{1}}\log B^{i}(u)\,=\,\sum_{\alpha}\,\Bigl{(}\sum_{k\,=\,1\atop k\neq i}^{N-1}\,\mathcal{B}_{\alpha}(u_{jk})\,\delta_{ij}\,+\,\sum_{k\,=\,1\,\atop\,k\,\neq\,i\,\wedge\,k\,\neq\,j}^{N-1}\mathcal{B}_{\alpha}(u_{Nk})\Bigr{)}\,+\,\ldots\,. (119)

For x=x1+τx2x=x_{1}+\tau x_{2}, with x1x_{1} and x2x_{2} being real, one obtains

α(x)x1(logP0Λ(x+Δα)logP0Λ(x+Δα))= 2im,nm0Λ|m|+|n|sin(2π(nΔα2mΔα1))(mτ+n)cos(2π(nx2mx1)).\begin{split}\mathcal{B}_{\alpha}(x)&\,\equiv\,\frac{\partial}{\partial x_{1}}\Bigl{(}\log P_{0\Lambda}(x+\Delta_{\alpha})\,-\,\log P_{0\Lambda}(-x+\Delta_{\alpha})\Bigr{)}\,\\ &\,=\,\!-\,2{\rm i}{\underset{m,n\in\mathbb{Z}\atop m\neq 0}{\sum}}\;\Lambda^{|m|\,+\,|n|}\,\frac{\sin\Bigl{(}2\pi(n\Delta_{\alpha 2}-m\Delta_{\alpha 1})\Bigr{)}}{(m\tau+n)}\,\cos\Bigl{(}2\pi(nx_{2}-mx_{1})\Bigr{)}\,.\end{split} (120)

The \ldots in (119) stand for terms with single elements, namely without sum over N1N-1. These terms are suppressed in the large-NN expansion. We write them for completeness

=α(uij)|ϵij|+ 2(α(uNi)+α(uNj))|ϵij|+ 4α(uNi)δij,\ldots\,=\,\,-\,\mathcal{B}_{\alpha}(u_{ij})\,|\epsilon_{ij}|\,+\,2\,\bigl{(}\mathcal{B}_{\alpha}(u_{Ni})+\mathcal{B}_{\alpha}(u_{Nj})\bigr{)}|\epsilon_{ij}|\,+\,4\,\mathcal{B}_{\alpha}(u_{Ni})\,\delta_{ij}\,, (121)

where 1i,jN11\leq i,\,j\leq N-1. At finite NN these terms are relevant. For instance at N=2N=2 the only non-vanishing contribution comes from the last term in (121).

At large NN the (N1)×(N1)(N-1)\times(N-1) matrix (119) takes the form

DN1×(2112112),D^{N-1}\times\begin{pmatrix}2&1&\dots&1\\ \vdots&&2&\vdots\\ 1&\ldots&1&2\end{pmatrix}\,, (122)

with 22’s as diagonal elements and 11’s otherwise. The matrix that multiplies DN1D^{N-1} has determinant N1N-1 and so the final form of E(νP)E(\nu_{P}) has the following form at leading order in large-NN expansion

(N1)DN1NDN1.(N-1)\,D^{N-1}\,\sim\,N\,D^{N-1}\,. (123)

Let us show these statements and determine DD. The first contribution in (119) is

k= 1kjN1α(ujk).\sum_{k\,=\,1\atop k\neq j}^{N-1}\,\mathcal{B}_{\alpha}(u_{jk})\,.

At large NNjNw[0,1)\frac{j}{N}\to w\in[0,1) and kNx[0,1)\frac{k}{N}\to x\in[0,1) and the sums become integrals. In that limit (4.1) becomes

N01𝑑xα((wx)T)=N01𝑑xα((xw)T)=N01𝑑xα(xT).\begin{split}N\,\int^{1}_{0}dx\mathcal{B}_{\alpha}((w-x)T)\,=\,N\,\int^{1}_{0}dx\mathcal{B}_{\alpha}((x-w)T)\,=\,N\,\int^{1}_{0}dx\mathcal{B}_{\alpha}(xT)\,.\end{split} (124)

After fixing Λ= 1\Lambda\,=\,1^{-}, the last term in (124) becomes

N01𝑑xα(xT)=4πiNT(TΔα2+[Δα]mT+12).N\,\int^{1}_{0}dx\mathcal{B}_{\alpha}(xT)\,=\,\frac{4\pi{\rm i}N}{T}\,\Bigl{(}-T\,\Delta_{\alpha 2}\,+\,[\Delta_{\alpha}]^{T}_{m}\,+\,\frac{1}{2}\Bigr{)}\,. (125)

There is still a second contribution to analyze from (119). Evaluating it at fixed points, the sum becomes an integral at leading order in large-NN expansion

k= 1kikjN1α(uNk)N01𝑑xα(xT).\sum_{k\,=\,1\,\atop\,k\neq i\,\wedge\,k\neq j}^{N-1}\,\mathcal{B}_{\alpha}(u_{Nk})\longrightarrow N\,\int^{1}_{0}dx\mathcal{B}_{\alpha}(xT)\,. (126)

This integral equals the contribution from the first term in (119). Thus, we conclude that at leading order in large-NN expansion, (119) takes the form (122), with

D=4πiNTα(TΔα2+[Δα]mT+12).D\,=\,\frac{4\pi{\rm i}N}{T}\,\sum_{\alpha}\Bigl{(}-T\,\Delta_{\alpha 2}\,+\,[\Delta_{\alpha}]^{T}_{m}\,+\,\frac{1}{2}\Bigr{)}\,. (127)

The final result for (m,n)(m,n) solutions is

deti,jujBi(P)=NN(2Tα(TΔα2+[Δα]mT+12))N1+sub-leading.\underset{i,j}{\det}\,{\partial_{u_{j}}B^{i}}(P)\,=\,N^{N}\,\Bigl{(}\frac{2}{T}\,\sum_{\alpha}\Bigl{(}-T\,\Delta_{\alpha 2}\,+\,[\Delta_{\alpha}]^{T}_{m}\,+\,\frac{1}{2}\Bigr{)}\Bigr{)}^{N-1}\,+\,\text{sub-leading}\,. (128)

In the limit Λ1\Lambda\to 1^{-} the effective action SλS_{\lambda} (112) develops discontinuities at walls defined by the condition mΔ1nΔ2m\Delta_{1}\,-\,n\Delta_{2}\in\mathbb{Z}\,. Dirichlet theorem fixes the value of Seffα(m,n)S_{\text{eff}\alpha}(m,n) and its derivatives Seffα(m,n)\partial^{\ell}\,S_{\text{eff}\alpha}(m,n), at such discontinuities to

(Seffα(m,n))L+(Seffα(m,n))R2,\frac{(\partial^{\ell}\,S_{\text{eff}\alpha}(m,n))^{L}\,+\,(\partial^{\ell}\,S_{\text{eff}\alpha}(m,n))^{R}}{2}\,, (129)

where LL and RR denote the limit values from the left and right sides of the corresponding discontinuity of Seffα(m,n)\partial^{\ell}\,S_{\text{eff}\alpha}(m,n)\,.

To compare with other literature, let us analyze the previously introduced case of 𝒩=4\mathcal{N}=4 SYM. Focus on the saddle/fixed point (1,0)(1,0) i.e on T=τT=\tau, then

α(TΔα2+[Δα]1τ+12)=12.\sum_{\alpha}\,\Bigl{(}-T\,\Delta_{\alpha 2}\,+\,[\Delta_{\alpha}]^{\tau}_{1}\,+\,\frac{1}{2}\Bigr{)}\,=\,-\,\frac{1}{2}\,. (130)

We recall that the function bracket was defined in (109). For a generic gauge anomaly-free theory αΔα2=TrR\sum_{\alpha}\Delta_{\alpha 2}\,=\,\text{Tr}R where R{R} denotes the superconformal R-charge of the fermionic fields in a given multiplet and Tr denotes sum over all multiplets. This quantity vanishes at large NN in virtue of ABJ anomaly cancellation Cabo-Bizet:2020nkr . For 𝒩=4\mathcal{N}=4 SYM this quantity vanishes at any value of NN.

Finally, we use (130) in the logarithm of (128) and obtain at leading order

logE(ν(1,0))=NlogN(N1)log(τ)+sub-leading.\log E(\nu_{(1,0)})\,=\,N\,\log N\,-\,(N-1)\log(\tau)\,+\,\text{sub-leading}\,. (131)

This answer matches a recent result obtained for the fixed point (1,0)(1,0) with a combination of a small τ\tau expansion and numerical extrapolation, via the use of the Bethe Ansatz formula GonzalezLezcano:2020yeb .

5 Towards an analytic approach to large-NN counting

In this section we combine the ABBV formula with Picard-Lefschetz method to study how complex eigenvalue configurations compete at leading order in the large-NN expansion of the microcanonical index. We will use as toy example the case of 𝒩=4\mathcal{N}=4 SYM with σ=τ\sigma=\tau and no flavour potentials. The generalization to the cases στ\sigma\neq\tau is explained in appendix B.2. The outline of this section was given in the introduction section 1. Here we will focus on the details.

5.1 The contour of integration

We focus on the simplest ensemble of SU(N)SU(N) 𝒩=4\mathcal{N}=4 SYM that captures the graded counting of 116\frac{1}{16}-th BPS operators. That is the family of integrals (6) with t(1)=t(2)=t(3)=tqt_{(1)}=t_{(2)}=t_{(3)}=tq and p=qp=q and the constraint (10t3=q1t^{3}=q^{-1}\,. This is a function of a single chemical potential.

For finite NN, the integral \mathcal{I} (6) can be computed order by order in a Taylor expansion around p=q=0p=q=0. The coefficients ana_{n} in such expansion can be computed as a contour integral over the Cartan torus Murthy:2020rbd ; Agarwal:2020zwm or via decomposition in characters of SU(N)SU(N) Murthy:2020rbd ; Dolan:2007rq . As the R-charges are quantized in multiples of 13\frac{1}{3}, the variable with the correct monodromy properties is q1/3q^{1/3}. The Fourier coefficients can then be extracted from the q13q^{\frac{1}{3}}-series, via the Laurent integral

a()=C𝑑q13(q13)q+13,j.a(\ell)\,=\,\oint_{C}\,{dq^{\frac{1}{3}}}\,\frac{\mathcal{I}(q^{\frac{1}{3}})\,}{\,q^{\frac{\ell+1}{3}}\,}\,,\qquad j\in\mathbb{Z}\,. (132)

The contour CC is a closed contour surrounded by |q|=1|q|=1. In terms of τ\tau the contour CC is a segment that for convenience we take to run in between the vertical lines τ1=2\tau_{1}=-2 and τ1=1\tau_{1}=1. As (τ)=(τ+3)\mathcal{I}(\tau)\,=\,\mathcal{I}(\tau+3) we can add to CC two vertical contours running in opposite sense from (resp. to) τ2=η\tau_{2}=-\eta to (resp. from) the left (resp. right) extremum of CC. The positive number η\eta is assumed to be as large as wished, see Figure 4. On the contrary to CCCηC_{\eta} has a natural decomposition in terms of Lefschetz thimbles. We will explain this later on. As explained in appendix F.1 the vertical pieces of CηC_{\eta} must be slightly deformed (while preserving their mutual cancellation) in such a way they cross the real axis across two irrational values of τ\tau at distance 33 of each other.

Plugging the ABBV formula (35) into (132), changing integration variables from q13τq^{\frac{1}{3}}\mapsto\tau, using the fact that the integral of (τ)\mathcal{I}(\tau) over CC is equal to the integral over CηC_{\eta}, and finally commuting the sum over PP’s with the integral over CηC_{\eta} 404040 If we consider only (m,n)(m,n) saddles, as we are doing, then the number of terms in the sum over fixed points, is less than N2N^{2}. Thus for large but finite NN, if the integrals of the summands is finite, then the integral can be commuted with the sum. one obtains the microcanonical index as a sum over fixed points PP

d(𝔮)=PdP,\begin{split}d(\mathfrak{q})\,=\,\sum_{P}\,d_{P}\,,\end{split} (133)

where d(𝔮)a(+1)d(\mathfrak{q})\equiv a(\ell+1) and 𝔮=EN2+13N2>0\mathfrak{q}\,=\,\frac{{E}}{N^{2}}\equiv\frac{\ell+1}{3N^{2}}>0 and the contributions of single fixed point PPdPd_{P}, are defined as

dPCη𝑑τeSλ(P) 2πiτEE(νP)=Cη𝑑τeSλ(P) 2πiτE+.{d}_{P}\,\equiv\,\int_{C_{\eta}}d\tau\,\frac{e^{-S_{\lambda}(P)\,-\,2\pi{\rm i}\tau E}}{E(\nu_{P})}\,=\,\int_{C_{\eta}}d\tau\,e^{-S_{\lambda}(P)\,-\,2\pi{\rm i}\tau{E}\,+\,\ldots}\,. (134)

The \ldots denote sub-leading contributions in the large-NN expansion. The goal of this section is to study the leading in NN asymptotics of the full sum by comparing the individual contributions of the dPd_{P}.

Refer to caption
Figure 4: The contour CηC_{\eta} in q13q^{\frac{1}{3}} and τ\tau-planes. The original contour in qq-plane is a circle inside the unit disk |q|<1|q|<1. The grey dashed circle (segment) represents |q|=1|q|=1 (Im(τ)=0\text{Im}(\tau)=0). The integration along the two vertical contours in τ\tau-plane cancel each other due to the periodicity properties of the index. Due to reasons explained in F.1 they must cross the real axis across irrational values. Thus, the integral along CηC_{\eta} equals the integral along the original contour CC.

5.2 Picard-Lefschetz method

In this section we use Picard-Lefschetz method to analyze the integrals at large NN\,. Our approach follows the presentation of Witten:2010cx . The large-NN limit of integrals dPd_{P} is determined by the critical points of

P(τ)S(P) 2πiτE.\mathcal{E}_{P}(\tau)\,\equiv\,-\,S(P)\,-\,2\pi{\rm i}\tau{E}\,. (135)

This function is analytic in the region

𝒳P{τ:ττP},\mathcal{X}_{P}\,\equiv\,\{\tau\,:\,\tau\in\mathbb{C}-\tau_{P}\}\,, (136)

and the point τP\tau_{P} is an essential singularity. The behaviour of P\mathcal{E}_{P} in a limit to τP\tau_{P} depends on how the limit is taken. For the example that we are considering, τP\tau_{P} is always a rational number.

We need to study real one-dimensional curves in the space 𝒳P\mathcal{X}_{P}, say τ=τ(s)𝒳P\tau\,=\,\tau(s)\in\mathcal{X}_{P}\,, with worldline parameter s(0,)s\in(0,\infty). They must end or start at a non-degenerate critical τc\tau_{c} of P\mathcal{E}_{P} i.e. they must satisfy the asymptotic initial condition

τc+τ(0+)=τcorτc=τ()=τc.\tau_{c}^{+}\,\equiv\,\tau(0^{+})\,=\,\tau_{c}\quad\text{or}\quad\tau^{-}_{c}\,=\,\tau(\infty)\,=\,\tau_{c}\,. (137)

Moreover, they should preserve the imaginary part of P\mathcal{E}_{P} along the trajectory, and thus are determined by the differential equation

sIP(τ(s))= 0,IP(τ)ImP(τ),\partial_{s}\,I_{P}(\tau(s))\,=\,0\,,\qquad I_{P}(\tau)\,\equiv\,\text{Im}\mathcal{E}_{P}(\tau)\,, (138)

subject to one of the initial conditions in (137). If the Morse function

HP(τ)=ReP(τ)H_{P}(\tau)\,=\,\text{Re}\mathcal{E}_{P}(\tau) (139)

is strictly decreasing along the flow i.e.

sHP(τ(s))<0,\partial_{s}H_{P}(\tau(s))<0\,, (140)

then such flow 𝒥\mathcal{J} is called a Lefschetz thimble and it satisfies the first of the conditions in equation (137). A more refined way of imposing condition (140) upon solutions to (138) is instead, to demand the flow to be a solution of the so-called downward flow equations Witten:2010cx with the initial condition (137). In the end, every thimble 𝒥\mathcal{J} obtained directly from downward flow equations can be obtained from restricting the space of solutions of (138) to the ones satisfying (140), and viceversa.

Instead of solving the flow equations, it is convenient to solve the algebraic relations in between imaginary and real part of the complex curves τ(s)\tau(s) that arise from demanding the imaginary part of the entropy function, IP(τ(s))I_{P}(\tau(s)), to coincide with the imaginary part of IP(τc))I_{P}(\tau_{c})) of the fixed point τc\tau_{c} of interest. The problem then reduces to that of solving algebraic equations. Out of those solutions one then needs to understand which ones are ascent or descent paths. Below we explain how that is done in this particular example.

The next step is to decompose the contour CηC_{\eta} in Figure 4 into an integral combination of Lefschetz thimbles 𝒥P\mathcal{J}_{P}

Cη=𝒥PnCη,𝒥P𝒥P,n𝒥P.C_{\eta}\,=\,\sum_{\mathcal{J}_{P}}\,n_{\langle C_{\eta},\mathcal{J}_{P}\rangle}\,\mathcal{J}_{P}\,,\qquad n_{\mathcal{J}_{P}}\in\mathbb{Z}\,. (141)

That this decomposition exists can be understood in two ways. One homological, and the other uses meromorphy of P\mathcal{E}_{P}. In the example we study, the latter turns out to be convenient to obtain the decomposition in thimbles (141). Let us briefly comment on why the former turns out to be more involved in this case. This requires to recall the concept of relative homology. Again, we follow the presentation of Witten:2010cx .

The thimbles 𝒥P\mathcal{J}_{P} are a basis of the relative homology group H1(𝒳P,𝒳PT,)H_{1}(\mathcal{X}_{P},\mathcal{X}_{P_{-T}},\mathbb{Z}). This is the group of one-real dimensional paths that originate and end at the given connected piece of “boundary” of 𝒳P\mathcal{X}_{P} (the space defined in (136)). One-cycles that can be contracted to a point without crossing any obstruction are identified as trivial. We will see an instance of this in the next subsection. By one-cycles one means the following. Select a very large positive real number TT, and excise from 𝒳P\mathcal{X}_{P} all the regions at which the Morse function HP<TH_{P}<-T. Then, identify each connected subpart of such region as a “point” and define the new punctured domain as 𝒳PT\mathcal{X}_{P_{-T}}. The one-cycles are the one-curves in 𝒳P\mathcal{X}_{P} that start and end at the “same” puncture of 𝒳PT\mathcal{X}_{P_{-T}}. Those cycles can be expanded as an integral combination of Lefschetz-thimbles 𝒥P\mathcal{J}_{P}. Usually the integers are computed as intersection numbers between the corresponding one-cycle and the dual ascent paths 𝒦P\mathcal{K}_{P}414141Ascent paths are one-dimensional curves that solve the flow equation (138). They are also asymptotically close to a critical point, but this time they obey the opposite condition to (140) i.e they flow towards the critical point. Equivalently, they solve the so-called upward flow equations Witten:2010cx . However, that method needs to be refined in the presence of Stokes’ lines i.e when there exists a flow running among two different critical points. In that case a thimble can be an ascent path as well, and one needs to slightly deform P\mathcal{E}_{P} to solve this issue (See Witten:2010cx for a detailed discussion on this, and many other related issues).

The present case “sits at” a Stokes line i.e. there is a thimble 𝒥P\mathcal{J}_{P} that it is also ascent path 𝒦P\mathcal{K}_{P}. That happens for every PP. The reason is that for every PP there are two critical points of P\mathcal{E}_{P} with the same imaginary part IPI_{P}. Using meromorphy this will turn out not to be an issue to compute the intersection numbers.

Using the contour decomposition (141) upon the integrals dPd_{P} defined in (134) one obtains

dP=𝒥PnCη,𝒥P𝒥P𝑑τeP(τ)+.d_{P}\,=\,\sum_{\mathcal{J}_{P}}n_{\langle C_{\eta},\mathcal{J}_{P}\rangle}\,\int_{\mathcal{J}_{P}}\,d\tau\,e^{\mathcal{E}_{P}(\tau)\,+\,\ldots\,}\,. (142)

We will see that for many PP’s the intersection numbers nn_{\langle\ldots\rangle} vanish. But many others could contribute. The \ldots denote sub-leading contributions in the large-NN expansion.

5.3 The competition among (m,n)(m,n) fixed points

Let NN be a large integer, that we assume to be prime. 424242This implies the (m,n)(m,n) configurations here-analyzed to be the only fixed points carrying finite Abelian group structure of rank NN Cabo-Bizet:2019eaf ; Cabo-Bizet:2020nkr . The fixed points PP are the (m,n)(m,n)’s configurations (65). The on-shell action of the (m,n)(m,n) fixed point with 0m,n<N0\leq m,n<N integer co-primes and (m,n)(0,0)(m,n)\neq(0,0), was computed in Cabo-Bizet:2019eaf

S(m,n)Seff(m,n;τ)=1mπi27N2(2T+χ1(m+n))3T2+N2πiφ(m,n).S(m,n)\equiv S_{\text{eff}}(m,n;\tau)\,=\,\frac{1}{m}\,\frac{\pi{\rm i}}{27}\,N^{2}\,\frac{\bigl{(}2T+\chi_{1}(m+n)\bigr{)}^{3}}{T^{2}}+N^{2}\pi{\rm i}\,\varphi(m,n)\,. (143)

In this formula T=mτ+nT=m\tau+n and χ1()={1,0,1}\chi_{1}(\ell)=\{-1,0,1\} if ={1,0,1}\ell=\{-1,0,1\} mod 33 respectively. From now on we focus on the cases χ1 0\chi_{1}\,\neq\,0 i.e. m+n 0m\,+\,n\,\neq\,0 mod 33. Incorporating the cases χ1= 0\chi_{1}\,=\,0 will not affect the conclusions that will be presented below. It will be useful to note the following conjugation relation

Seff(1,0;τR)=Seff(1,1;τ)S_{\text{eff}}(1,0;\tau_{R})\,=\,S_{\text{eff}}(1,1;\tau)^{*}\, (144)

where τR= 1τ1+iτ2\tau_{R}\,=\,-\,1\,-\,\tau_{1}\,+\,{\rm i}\,\tau_{2}\,. For a given (m,n)(m,n) we must extremise the functional (135)

(m,n)(τ)=Seff(m,n;τ)2πiτE.\mathcal{E}_{(m,n)}(\tau)\,=\,-S_{\text{eff}}(m,n;\tau)-2\pi{\rm i}\,\tau{E}\,. (145)

The extremization of this functional was analyzed in some detail in Cabo-Bizet:2019eaf . Here we will complete the analysis by following the Picard-Lefschetz method.

For any (m,n)(m,n) there are three critical points in 𝒳(m,n)\mathcal{X}_{(m,n)}. Let us denote them as

τ+,τ,τ0.\tau^{*}_{+}\,,\qquad\tau^{*}_{-}\,,\qquad\tau^{*}_{0}\,. (146)

We have not written down their explicit dependence on (m,n)(m,n), but they do depend on mm and nn Cabo-Bizet:2019eaf . For the critical points τ±\tau^{*}_{\pm} the Morse function H(m,n)H_{(m,n)} has the same absolute value and opposite sign: say plus for τ+\tau^{*}_{+} and minus for τ\tau^{*}_{-}. For the critical point τ0\tau^{*}_{0}

H(m,n)(τ0)= 0.H_{(m,n)}(\tau^{*}_{0})\,=\,0\,. (147)

The critical points τ±\tau^{*}_{\pm} have equal phase function I(m,n)I_{(m,n)}. Indeed for all (m,n)(m,n) there will always be a descent/ascent flow from τ+\tau^{*}_{+} to τ\tau^{*}_{-}\,.

Let us start by studying the thimbles associated to the fixed point (1,0)(1,0). For the critical point τ+\tau^{*}_{+} there are two thimbles, one flows to τ2=\tau_{2}=-\infty and the other flows to τ= 0\tau\,=\,0. The point τ= 0\tau\,=\,0 is the essential singularity of (1,0)\mathcal{E}_{(1,0)}, namely the difference between 𝒳(1,0)\mathcal{X}_{(1,0)} and \mathbb{C}, as defined previously in equation (136). For a limit that approaches τ=0\tau=0 from the first and third quadrants, H(1,0)H_{(1,0)} goes to -\infty. For a limit that approaches τ=0\tau=0 from the second and fourth quadrants, H(1,0)H_{(1,0)} goes to ++\infty. This means that Lefschetz thimbles of (1,0)(1,0) that flow to τ=0\tau=0, must approach it from the first and third quadrants, meanwhile ascent paths emanating from τ=0\tau=0 must do it through the second and/or fourth quadrants. Indeed this is what happens.

For (1,1)(1,1) the pattern is analogous, but this time the essential singularity is located at τ=1\tau\,=\,-1 and now the first (second) and third (fourth) quadrants take the roles of the second (first) and the fourth (third) had in the previous case, respectively. For a generic (m,n)(m,n) fixed point, the story is also the same. Now the essential singularity is located at τ=nm\tau=-\frac{n}{m} and the two opposite quadrants of that singularity through which the thimbles can enter, are determined by the sign of χ1(m+n)\chi_{1}(m+n). If χ1(m+n)=+1\chi_{1}(m+n)=+1 then thimbles should approach it through the first and third quadrants, otherwise through the second and fourth. The other region, apart from the essential singularity τ=nm\tau=-\frac{n}{m}, to which the thimbles can flow towards is the asymptotic region τ=i\tau\,=\,-{\rm i}\infty. Namely in this region H(m,n)H_{(m,n)} goes to -\infty. This is the same region where our integration cycle starts and ends at.

Refer to caption
Refer to caption
Figure 5: The plots of the analytic solutions (of the relevant algebraic equations) obtained for the Lefschetz thimbles and ascent paths associated to the (1,1)(1,1) (left) and (1,0)(1,0) (right) fixed points. The solution for generic (m,n)(m,n) has the same form as one of the two plots above. Each critical point of the entropy functional (m,n)\mathcal{E}_{(m,n)} in (145), has two thimbles and two ascent paths associated. In the figure we have only indicated the direction of the flows corresponding to the thimbles that appear (have non-zero intersection number) in the decomposition of the original integration contour CηC_{\eta} which we have denoted as IIIIIIIIIIIIIVIVVV. The different HH’s in the plots denote the values of the Morse function at the closest critical point and also at τ2±\tau_{2}\approx\pm\infty where H=±TH=\pm T, respectively, with TT a very large positive number.

To know whether the intersection numbers nn_{\langle\ldots\rangle} that determine the contribution of fixed points (m,n)(m,n) to the integral along CηC_{\eta} vanish or not, we need to know whether the cycle CηC_{\eta} is trivial in the relative homology H1(𝒳(m,n),𝒳(m,n)T;)H_{1}(\mathcal{X}_{(m,n)},\mathcal{X}_{(m,n)_{-T}};\mathbb{Z})\,. That means to check whether the CηC_{\eta} encloses the essential singularity τ=nm\tau=-\frac{n}{m}. If the cycle CηC_{\eta} does not enclose τ=nm\tau=-\frac{n}{m} the integral (m,n)\mathcal{I}_{(m,n)} vanishes. It vanishes because the integrand is meromorphic and the contour can be contracted to a point at τ=i\tau=-{\rm i}\infty (where the integrand vanishes). If on the contrary the cycle CηC_{\eta} does enclose τ=nm\tau=-\frac{n}{m} then the fixed point (m,n)(m,n) does contribute and the precise intersection numbers can be computed easily.

Given our choice of contour CηC_{\eta}, only fixed points for which

2<nm< 1,-2<-\,\frac{n}{m}\,<\,1\,, (148)

contribute to the entropy counting formula (133). Take an (m,n)(m,n) that obeys this condition. The symmetry transformation ττ+3\tau\mapsto\tau+3 maps (m,n)(m,n+3m)(m,n)\mapsto(m,n+3m). Naively, these two saddles should contribute with the same exponential weight, namely m,n=m,n+3m\mathcal{E}_{m,n}\,=\,\mathcal{E}_{m,n+3m}, however (m,n+3m)(m,n+3m) does not satisfy (148) and thus its intersection numbers vanish. Thus, these replica saddles do not contribute to the large-NN microcanonical index defined by the contour integral CηC_{\eta}.

In figure 4 we have plotted the analytic solutions of all the three thimbles and three ascent paths, together with the corresponding critical points, for fixed points (1,0)(1,0) and (1,1)(1,1). In both cases the contour CηC_{\eta} can be explicitly checked to have intersection number equal to either 111-1 or 0 associated to thimbles 𝒥(1,0)\mathcal{J}_{(1,0)} and 𝒥(1,1)\mathcal{J}_{(1,1)} i.e.

Cη=𝒥(1,0)I𝒥(1,0)II+𝒥(1,0)III𝒥(1,0)IV+𝒥(1,0)V,=𝒥(1,1)I+𝒥(1,1)II𝒥(1,1)III+𝒥(1,1)IV𝒥(1,1)V.\begin{split}C_{\eta}&\,\,=\,\,-\,\mathcal{J}_{(1,0){I}}\,-\,\mathcal{J}_{(1,0){II}}\,+\,\mathcal{J}_{(1,0){III}}\,-\,\mathcal{J}_{(1,0){IV}}\,+\,\mathcal{J}_{(1,0){V}}\,,\\ &\,\,=\,\,\quad\,\mathcal{J}_{(1,1){I}}\,+\,\mathcal{J}_{(1,1){II}}\,-\,\mathcal{J}_{(1,1){III}}\,+\,\mathcal{J}_{(1,1){IV}}\,-\,\mathcal{J}_{(1,1){V}}\,.\end{split} (149)

This decomposition can be obtained by simply comparing the contour CηC_{\eta} with the thimbles in figure 5.

Refer to caption
Figure 6: For large enough NN we can focus on combinations of subregions of the dominating thimbles that include τ+\tau^{*}_{+}. This is the plot of two such combinations, one of them associated to (1,1)(1,1) (to the left), the other to (1,0)(1,0) (to the right). These two combinations transform into each other under the 2\mathbb{Z}_{2} operation ττR\tau\mapsto\tau^{R}. The arrows denote the sense in which the original contour CηC_{\eta} flows. For convenience we define 𝒥~(1,0)\widetilde{\mathcal{J}}_{(1,0)} and 𝒥~(1,1)\widetilde{\mathcal{J}}_{(1,1)} to flow in the same sense CηC_{\eta} does.

Integrals along these contours can be computed exactly with all the data we have. However, in this paper we are interested in the leading approximation at large NN so it will be enough to identify the contribution of the leading critical point. That is always the τ+\tau^{*}_{+}, thus, for instance, to analyze the contribution of the fixed points (1,0)(1,0) and (1,1)(1,1) we can focus on the following two portions of thimble

Cη𝒥~(1,0)+,Cη𝒥~(1,1)+.\begin{split}C_{\eta}\,\sim\widetilde{\mathcal{J}}_{(1,0)}\,+\,\ldots\,,\\ C_{\eta}\,\sim\widetilde{\mathcal{J}}_{(1,1)}\,+\,\ldots\,.\end{split} (150)

A computation shows that 𝒥~(1,0)\widetilde{\mathcal{J}}_{(1,0)} and 𝒥~(1,1)\widetilde{\mathcal{J}}_{(1,1)} transform into each other under the 2\mathbb{Z}_{2} operation

(τ1,τ2(τ1))(τ1R,τ2(τ1R)).(\tau_{1},\tau_{2}(\tau_{1}))\mapsto(\tau^{R}_{1},\tau_{2}(\tau^{R}_{1}))\,. (151)

Moreover, from property (144), definition (145) and the fact E{E}\in\mathbb{R}, it follows that

𝒥~(1,1)𝑑τ1e(1,1)(τ(τ1))=𝒥~(1,0)𝑑τ1e((1,0)(τ(τ1))).\int_{\widetilde{\mathcal{J}}_{(1,1)}}d\tau_{1}\,e^{\mathcal{E}_{(1,1)}(\tau(\tau_{1}))}\,=\,\int_{\widetilde{\mathcal{J}}_{(1,0)}}d\tau_{1}\,e^{\bigl{(}\mathcal{E}_{(1,0)}(\tau(\tau_{1}))\bigr{)}^{*}}\,. (152)

This relation implies that contributions to the large-NN microcanonical index that come from (1,0)(1,0) and (1,1)(1,1), are complex conjugated to each other. We expect that this property will be preserved by sub-leading corrections in the large-NN expansion as well. That is because by direct evaluation one can check that the Fourier coefficients ana_{n} are always real integers.

As explained before, generic (m,n)(m,n) fixed points have analogous thimble structure. Thus, we must compare the contribution coming from thimbles associated to generic (m,n)(m,n) fixed points that obey the condition (148). For those (m,n)(m,n), the intersection number nCη,n_{\langle C_{\eta},\ldots\rangle}\, with the leading thimble will always have absolute value one. Thus, the contribution of those fixed points is determined by the value of the Morse function at τ+\tau^{*}_{+} which is

H(m)(𝔮)=H(m,n)(τ+)N2mπ(2T+ 1)3(1 6T)243T2,T54𝔮+23D2+ 82 22/3(27𝔮+ 4)D,D(27𝔮+4)2+33𝔮(27𝔮+4)33.\begin{split}H_{(m)}(\mathfrak{q})\,=\,H_{(m,n)}(\tau^{*}_{+})&\,\equiv\,\frac{N^{2}}{m}\,\frac{\pi\sqrt{(2T^{*}+\,1)^{3}\,(1\,-\,6T^{*})}}{24\sqrt{3}T^{*2}}\,,\\ T^{*}&\,\equiv\,-\frac{54\,\mathfrak{q}\,+\,\sqrt[3]{2}\,D^{2}\,+\,8}{2\,2^{2/3}\,(27\,\mathfrak{q}\,+\,4)\,D}\,,\\ D&\,\equiv\,\sqrt[3]{(27\mathfrak{q}+4)^{2}+3\sqrt{3}\sqrt{\mathfrak{q}(27\mathfrak{q}+4)^{3}}}\,.\end{split} (153)

The Morse function is inversely proportional to mm and it does not depend on nn. At a given level mm there are only a finite number of nn’s with |χ(m+n)|=1|\chi(m+n)|=1 in the intersection domain (148). For m=1m=1 there are two, n=0n=0 and n=1n=1. These are the two dominating contributions at large NN, and for charges of order N2N^{2}. As explained above, both have the same leading exponential contribution in absolute value, but their phases are complex conjugated to each other and are given by the imaginary part

I(1,0)=I(1,1)=πN2(2T+1)(10T1)72T2+πN22.I_{(1,0)}\,=\,-\,I_{(1,1)}\,=\,-\,\frac{\pi\,N^{2}\,(2T^{*}+1)(10T^{*}-1)}{72T^{*2}}\,+\,\frac{\pi N^{2}}{2}. (154)
434343For large 𝔮\mathfrak{q} after large-NN H(1,0)N2123π𝔮2/3π𝔮33+O(1𝔮3),I(1,0)N212π𝔮2/3+π𝔮33+π6+O(1𝔮3).\frac{H_{(1,0)}}{N^{2}}\sim\frac{1}{2}\sqrt{3}\pi\mathfrak{q}^{2/3}-\frac{\pi\sqrt[3]{\mathfrak{q}}}{\sqrt{3}}+O\left(\sqrt[3]{\frac{1}{\mathfrak{q}}}\right)\,,\qquad\frac{I_{(1,0)}}{N^{2}}\sim\frac{1}{2}\pi\mathfrak{q}^{2/3}+\frac{\pi\sqrt[3]{\mathfrak{q}}}{3}+\frac{\pi}{6}+O\left(\sqrt[3]{\frac{1}{\mathfrak{q}}}\right)\,. (155) We should note that in terms of the charge EE of CFT operators, H(1,0)N23E23H_{(1,0)}\sim N^{\frac{2}{3}}E^{\frac{2}{3}}.

If we discard sub-leading corrections coming from E(νP)E(\nu_{P})444444These were analyzed in subsection 4.1. and exponentially suppressed contributions coming from other fixed points, we obtain

d(𝔮)=eH(1)(𝔮)+×2cos(I(1,0)(𝔮)+)+.d(\mathfrak{q})\,=\,e^{H_{(1)}(\mathfrak{q})\,+\,\ldots}\,\times 2\cos{\bigl{(}I_{(1,0)}(\mathfrak{q})\,+\,\ldots\bigr{)}}\,+\,\ldots\,. (156)

where we should stress that other fixed point configurations (m,n)(m,n) also interfere to produce extra oscillations on top of (156). However, at finite but relatively large values of NN and at finite 𝔮\mathfrak{q}, these are exponentially suppressed with respect to the contributions coming from the pair (1,0)(1,0) and (1,1)(1,1), as one can note immediately from equation (153). Moreover, at very first leading order in large-NN expansion all oscillations attenuate and the microcanonical index matches the exponential of the Bekenstein-Hawking entropy of the dual black holes (See figure 3)

d(𝔮)eH(1)(𝔮)=eAH4.d(\mathfrak{q})\,\sim\,e^{H_{(1)}(\mathfrak{q})}\,=\,e^{\frac{A_{H}}{4}}\,. (157)

Here AHA_{H} is the area of the horizon of the Gutowski-Reall black hole. 454545We have absorbed the 5d5d Newton constant G5DG_{5D}, which is to be identified as proportional to 1N2\frac{1}{N^{2}}, in the definition of AHA_{H}.

For small enough values of 𝔮\mathfrak{q} after the large-NN limit is taken, the physical picture changes. In that case the relevant expansion is

H(1,0)N2π𝔮3/2218π𝔮5/2+O(𝔮7/2),I(1,0)N2π2+π𝔮3π𝔮22+ 5π𝔮3+O(𝔮7/2).\frac{H_{(1,0)}}{N^{2}}\sim\pi\mathfrak{q}^{3/2}\,-\,\frac{21}{8}\pi\mathfrak{q}^{5/2}+O\left(\mathfrak{q}^{7/2}\right)\,,\qquad\frac{I_{(1,0)}}{N^{2}}\,\sim\,\frac{\pi}{2}\,+\,\pi\mathfrak{q}\,-\,\frac{3\pi\mathfrak{q}^{2}}{2}\,+\,5\pi\mathfrak{q}^{3}\,+\,O\left(\mathfrak{q}^{7/2}\right)\,. (158)

In terms of the charge EE of the CFT operators, H(1,0)N1E32H_{(1,0)}\sim N^{-1}E^{\frac{3}{2}}. Roughly speaking, this implies that to count operators with charges of the order EN23E\sim N^{\frac{2}{3}} or smaller, the competition of other (m,n)(m,n) saddles can not be neglected. Moreover, other critical points of the entropy functional (m,n)\mathcal{E}_{(m,n)} in (145), mainly the real one (out of the triad of them), should also be analyzed in the regime of small charges (at large NN). 464646Interestingly, note that for small enough values of 𝔮\mathfrak{q}, at large NN, the interference among (1,0)(1,0) and (1,1)(1,1) is destructive i.e they tend to cancel each other.

Let us summarize in words what we have learnt. The parameter that controls the magnitude of the exponential corrections coming from fixed points is exp[AH/4]\exp[-A_{H}/4] i.e. the inverse of the exponential of the entropy of the Gutowski-Reall black hole. Generically, this parameter is small for large enough NN, meaning by this, values of NN such as the discrete sums over eigenvalues are well approximated by the integrals. However, for the two-fixed point approximation to work, at relatively large but still finite values of NN, operator charges need to be large enough in units of the specific values NN, in such a way that the perturbation parameter exp[AH/4]\exp[-A_{H}/4] remains small. That means that for small enough operators, the two-fixed point approximation breaks down and one needs to start considering many more configurations to approximate the correct micro-canonical index. This means that when one approaches the transition from the black hole phase to the multigraviton phase, the “two- fixed-point” approximation breaks down and competition from all (m,n)(m,n) saddles must be considered. This observation deserves deeper analysis and exploration. The ABBV and the Bethe ansatz formulae are tools that could be useful to make this analytic picture precise at the numerical level (for instance at large NN but for small operators).

Acknowledgements

We thank Cyril Closset and Heeyeon Kim for a useful discussion regarding regularization procedures. We are grateful to Davide Cassani and Dario Martelli for useful comments and discussions. We are especially grateful to Sameer Murthy for his careful reading of the manuscript, as well as for many useful comments and discussions. This work is supported by the ERC Consolidator Grant N. 681908, “Quantum black holes: A microscopic window into the microstructure of gravity”.

Appendix A Definitions and identities

The quasi-elliptic function θ0(z)=θ0(z;τ)\theta_{0}(z)=\theta_{0}(z;\tau) has the following product representation

θell(ζ;q)=θ0(z;τ)=(1ζ)j=1(1qjζ)(1qjζ1).\theta_{\text{ell}}(\zeta;q)\,=\,\theta_{0}(z;\tau)\,=\,(1-\zeta)\prod_{j=1}^{\infty}(1-q^{j}\zeta)\,(1-q^{j}\zeta^{-1})\,. (159)

The zeroes of θ0\theta_{0} are located at z=jτ+kz=j\tau+k where jj and kk run over the integers. Representation (159) is convergent for τ2>0\tau_{2}>0. It has also the following quasi-periodicity properties

θ0(z)=θ0(z+1)=𝐞(z)θ0(z+τ)θ0(τz)=θ0(z).\theta_{0}(z)\,=\,\theta_{0}(z+1)\,=\,-{\bf e}(z)\,\theta_{0}(z+\tau)\,\quad\,\theta_{0}(\tau-z)\,=\,\theta_{0}(z)\,. (160)

The elliptic Gamma functions are defined out of the following product representation

Γell(ζ;p,q)=Γe(z;σ,τ)=j,k=01ζ1pj+1qj+11ζpjqk.\Gamma_{\text{ell}}(\zeta;p,q)\,=\,\Gamma_{\text{e}}(z;\sigma,\tau)\,=\,\prod_{j,\,k=0}^{\infty}\frac{1\,-\,\zeta^{-1}p^{j+1}q^{j+1}}{1\,-\,\zeta\,p^{j}\,q^{k}}\,. (161)

The poles and zeroes of Γe\Gamma_{\text{e}} are located at z=jσkτ+z=-j\sigma-k\tau+\ell and z=(j+1)σ(k+1)τ+z=(j+1)\sigma-(k+1)\tau+\ell respectively where the indices jjkk run over the non-negative integers and \ell runs over the integers Felder2000 . It has also the following quasi-periodicity properties

Γe(z+σ;σ,τ)=θ0(z;τ)Γe(z;σ,τ),Γe(z+τ;σ,τ)=θ0(z;σ)Γe(z;σ,τ).\begin{split}\Gamma_{\text{e}}(z+\sigma;\sigma,\tau)&\,=\,\theta_{0}(z;\tau)\,\Gamma_{\text{e}}(z;\sigma,\tau)\,,\\ \Gamma_{\text{e}}(z+\tau;\sigma,\tau)&\,=\,\theta_{0}(z;\sigma)\,\Gamma_{\text{e}}(z;\sigma,\tau)\,.\end{split} (162)

The first three Bernoulli polynomials are

B1(z)\displaystyle B_{1}(z) \displaystyle\,\equiv\, z12,\displaystyle z-\frac{1}{2}\,, (163)
B2(z)\displaystyle B_{2}(z) \displaystyle\,\equiv\, z2z+16,\displaystyle z^{2}-z+\frac{1}{6}\,, (164)
B3(z)\displaystyle B_{3}(z) \displaystyle\,\equiv\, z33z22+z2.\displaystyle z^{3}-\frac{3\,z^{2}}{2}+\frac{z}{2}\,. (165)

Their Fourier series decomposition, for k>1k>1 and 0z<10\leq z<1 is

Bk(z)=k!(2πi)kj0𝐞(jz)jk.B_{k}(z)\,=\,-\frac{k!}{(2\pi{\rm i})^{k}}\,\sum_{j\neq 0}\frac{{\bf e}(jz)}{j^{k}}\,. (166)

The k=1k=1 is peculiar in the sense that (166) is almost fine, but fails to hold at z=0z=0. The equality (166) doesn’t hold for k=1k=1 because at z=0z=0 the Fourier expansion vanishes meanwhile the Bernoulli polynomial B1({0})=12B_{1}(\{0\})=-\frac{1}{2}. Thus, (166) fails to hold at z=0z=0. This will have zero impact in generic computations, and thus, generically, we will only highlight the difference when it will matter .

Clearly this also equals Bk({x})B_{k}(\{x\}) where {x}xx\{x\}\,\equiv\,x-\lfloor x\rfloor is the fractional part of xx. In particular we have for xx\in\mathbb{R}

B0({x})\displaystyle B_{0}(\{x\}) \displaystyle\,\equiv\, 1jδ(x+j),\displaystyle 1-\sum_{j}\delta(x+j)\,, (167)
B1({x})\displaystyle B_{1}(\{x\}) \displaystyle\,\equiv\, 12πij0𝐞(jx)jforx,\displaystyle-\frac{1}{2\pi{\rm i}}\sum_{j\neq 0}\frac{{\bf e}(jx)}{j}\,\quad\text{for}\,\quad x\notin\mathbb{Z}\,, (168)
B2({x})\displaystyle B_{2}(\{x\}) \displaystyle\,\equiv\, 12π2j0𝐞(jx)j2,\displaystyle\frac{1}{2\pi^{2}}\,\sum_{j\neq 0}\frac{{\bf e}(jx)}{j^{2}}\,, (169)
B3({x})\displaystyle B_{3}(\{x\}) \displaystyle\,\equiv\, 3i4π3j0𝐞(jx)j3.\displaystyle-\frac{3\,{\rm i}}{4\pi^{3}}\,\sum_{j\neq 0}\frac{{\bf e}(jx)}{j^{3}}\,. (170)

Equality (168) does not hold at xx\in\mathbb{Z}. These are the positions of the discontinuities of both he left and the right-hand objects. The value of the function in the left-hand side at xx\in\mathbb{Z} coincides with the limit from the right. The value of the function in the right-hand side, matches the semisum of the left and right limits.

The generic polynomial BkB_{k} for k= 1, 2,k\,=\,1\,,\,2\,,\,\ldots obeys the following three properties (the first and second also apply to p=0p=0)

01𝑑xBk({x})\displaystyle\int_{0}^{1}dx\,B_{k}(\{x\}) =\displaystyle\,=\, 0,\displaystyle 0\,, (171)
Bk({x})\displaystyle B_{k}(\{x\}) =\displaystyle\,=\, (k+1)ddxBk+1({x}),\displaystyle(k+1)\,\frac{d}{dx}B_{k+1}(\{x\})\,, (172)
Bk(z+1)Bk(z)\displaystyle B_{k}(z+1)\,-\,B_{k}(z) =\displaystyle\,=\, kzk1.\displaystyle kz^{k-1}\,. (173)

At some points we will need to use the following smooth versions of the periodic Bernoulli polynomials

B1Λ({x})\displaystyle B_{1\Lambda}(\{x\}) \displaystyle\,\equiv\, 12πij0Λ|j|𝐞(jx)j=12πi(Li1(Λ𝐞(x))Li1(Λ𝐞(x))),\displaystyle-\frac{1}{2\pi{\rm i}}\sum_{j\neq 0}\frac{\Lambda^{|j|}{\bf e}(jx)}{j}\,=\,-\,\frac{1}{2\pi{\rm i}}\Bigl{(}\text{Li}_{1}(\Lambda{\bf e}(x))\,-\,\text{Li}_{1}(\Lambda{\bf e}(-x))\Bigr{)}\,, (174)
B2Λ({x})\displaystyle B_{2\Lambda}(\{x\}) \displaystyle\,\equiv\, 12π2j0Λ|j|𝐞(jx)j2=12π2(Li2(Λ𝐞(x))+Li2(Λ𝐞(x))),\displaystyle\frac{1}{2\pi^{2}}\,\sum_{j\neq 0}\frac{\Lambda^{|j|}{\bf e}(jx)}{j^{2}}\,\,=\,\,\,\,\,\,\,\,\frac{1}{2\pi^{2}}\Bigl{(}\text{Li}_{2}(\Lambda{\bf e}(x))\,+\,\text{Li}_{2}(\Lambda{\bf e}(-x))\Bigr{)}\,, (175)
B3Λ({x})\displaystyle B_{3\Lambda}(\{x\}) \displaystyle\,\equiv\, 3i4π3j0Λ|j|𝐞(jx)j3=3i4π3(Li3(Λ𝐞(x))Li3(Λ𝐞(x))).\displaystyle-\frac{3\,{\rm i}}{4\pi^{3}}\,\sum_{j\neq 0}\frac{\Lambda^{|j|}{\bf e}(jx)}{j^{3}}\,=\,-\,\frac{3\,{\rm i}}{4\pi^{3}}\,\Bigl{(}\text{Li}_{3}(\Lambda{\bf e}(x))\,-\,\text{Li}_{3}(\Lambda{\bf e}(-x))\Bigr{)}\,. (176)

For xx\in\mathbb{R} we will assume 0Λ<10\leq\Lambda<1. In cases where the argument xx happens to be complex we will assume 0<Λe2π|Imx|<10<\Lambda e^{2\pi|\text{Im}x|}<1 .

Recalling the functions logP\log PlogQ\log Q

The function logP\log P is defined by the following double expansion

logP(z)=logP(z1,z2;τ)=logP0(z1,z2;τ)+πiτB2({z2})+πiΨP(z)=i2πm,nm0𝐞(nz2mz1)m(mτ+n)+πiτB2({z2})+πiΨP(z),\begin{split}\log P(z)&\,=\,\log P(z_{1},z_{2};\tau)\,=\,\log P_{0}(z_{1},z_{2};\tau)\,+\,{\pi{\rm i}\tau}B_{2}\bigl{(}\{z_{2}\}\bigr{)}+\pi{\rm i}\Psi_{P}(z)\,\\ &\,=\,-\frac{{\rm i}}{2\pi}\!{\underset{m,n\in\mathbb{Z}\atop m\neq 0}{\sum}}\;\frac{{\bf e}(nz_{2}-mz_{1})}{m(m\tau+n)}\,+\,{\pi{\rm i}\tau}B_{2}\bigl{(}\{z_{2}\}\bigr{)}+\pi{\rm i}\Psi_{P}(z)\,,\end{split} (177)

where ΨP(z)\Psi_{P}(z) is an arbitrary τ\tau independent double periodic real function. The real part of logP\log P is the Kronecker-Eisenstein series. Without loss of generality, in the main body of the manuscript we fix ΨP= 0\Psi_{P}\,=\,0. This is equivalent to a trivial redefinition of the regularization ambiguity Ψ0\Psi_{0} in (205).

The function logQ\log Q is defined by the following double expansion

logQ(z)=logQ(z1,z2;τ)=logQ0(z1,z2;τ)+2πiτ3B3({z2})+πiΨQ(z)14π2m,nm0𝐞(nz2mz1)m(mτ+n)2+2πiτ3B3({z2})+πiΨQ(z),\begin{split}\log Q(z)&\,=\,\log Q(z_{1},z_{2};\tau)\,=\,\log Q_{0}(z_{1},z_{2};\tau)\,+\,\frac{2\pi{\rm i}\tau}{3}B_{3}\bigl{(}\{z_{2}\}\bigr{)}\,+\,\pi{\rm i}\,\Psi_{Q}(z)\,\\ &\,\equiv\,-\frac{1}{4\pi^{2}}\!{\underset{m,n\in\mathbb{Z}\atop m\neq 0}{\sum}}\;\frac{{\bf e}(nz_{2}-mz_{1})}{m(m\tau+n)^{2}}\,+\,\frac{2\pi{\rm i}\tau}{3}B_{3}\bigl{(}\{z_{2}\}\bigr{)}\,+\,\pi{\rm i}\,\Psi_{Q}(z)\,,\end{split} (178)

where ΨQ(z)\Psi_{Q}(z) is an arbitrary τ\tau independent double periodic real function. The real part of logQ\log Q is the Bloch-Wigner elliptic dilogarithm. Without loss of generality, in the main body of the manuscript we fix ΨQ= 0\Psi_{Q}\,=\,0. This is equivalent to a trivial redefinition of the regularization ambiguity Ψ0\Psi_{0} in (205).

Appendix B The supersymmetric partition function and the Index

In this appendix we revisit the problem of regularization of the divergent super-determinants of the underlying four-dimensional field-theory problem from a different perspective than usual. 474747The gravitational counterpart of this perspective will not be studied here Assel:2014tba ; Genolini:2016ecx ; Papadimitriou:2017kzw ; An:2017ihs . This gives a first principles origin of the sequence of actions SλS_{\lambda} that was used to find a fixed point expansion of the superconformal index.

Before entering in technical discussions, let us take a couple of paragraphs to comment about regularization schemes in the context of supersymmetric partition functions. The superconformal index and the supersymmetric partition function that the gravitational side of the AdS/CFTAdS/CFT duality sets as the appropriate observable to study Cabo-Bizet:2018ehj , can differ by a pre-factor. We will show that there exist regularization schemes that, distinctly to the schemes used in Cabo-Bizet:2018ehj ; Closset:2013sxa ; Assel:2014paa ; Assel:2015nca , give the superconformal index out of the divergent 4d super-determinants without the pre-factor. In that sense the schemes below will play an analogous role to the one played by the background-subtraction method used in the gravitational side of the duality, when evaluating the onshell action Chen:2005zj of the dual BPS black holes Cabo-Bizet:2018ehj ; Cassani:2019mms .

The difference in between this regularization and the ones of Cabo-Bizet:2018ehj ; Closset:2013sxa ; Assel:2014paa ; Assel:2015nca  484848The method used in the first three references also differs from the one used in the last one, by a finite quantity. is a finite contribution. It would be interesting to better understand whether the counterterms accounting for such finite differences can go beyond the hypothesis of the no-go arguments of Assel:2014tba , or not; and in case the answer is yes, then how. In particular the former scheme cancels the pre-factor, meanwhile the latter two do not. As these schemes are used after supersymmetric paired contributions are cancelled, then in principle the three of them could naively be thought to be consistent with supersymmetry, and thus from this perspective, it is not clear to us why one of them is supersymmetric and the others are not. This is a puzzle to us. Perhaps an analysis along the lines of Kuzenko:2019vvi ; Bzowski:2020tue could help to resolve it.

After this small disgresion, let us summarize. The first outcome of this appendix is to show that there exists a regularization, after supersymmetric cancellations have been already taken into consideration, that cancels the relative pre-factor between superconformal index and supersymmetric partition function. 494949Note that this does not mean that the pre-factor lacks physical meaning. Indeed, the pre-factor is known to be related to ’t Hooft anomaly coefficients of the corresponding theory, as it will be recalled in appendix C.1. To claim that this scheme is consistent with supersymmetry a more detailed analysis along the lines of Kuzenko:2019vvi ; Bzowski:2020tue needs to be pursued. 505050It would be also interesting to compare with the scheme presented in Closset:2019ucb . That lies beyond the scope of this paper. The most important outcome of this section is that the freedom in choice of that regularization scheme, includes the freedom in choice of extension of the integrand (of the superconformal index) to the complex plane. This gives a first principles derivation of the extensions SλS_{\lambda} that we used to find the fixed point expansion of the superconformal index.

For a while we will focus on the contribution of a single chiral multiplet. In the following appendices we will reinstate the weights ρ\rho\, and the multiplet index α\alpha and sum over them.

B.1 Two examples of regularization

We start by regularizing the one-loop determinants of the supersymmetric partition functions on S1×S3S_{1}\times S_{3}. Firstly, we illustrate two cases, that we call the elliptic and meromorphic schemes. The elliptic scheme corresponds to the double periodic extension of Cabo-Bizet:2019eaf ; Cabo-Bizet:2020nkr . Both regularization schemes match the superconformal index at the contour of integration u¯2=0\underline{u}_{2}=0.

General strategy to regularize

The contribution of a 4d4d 𝒩=1\mathcal{N}=1 chiral multiplet to the supersymmetric partition function on S1×S3S_{1}\times S_{3} has been computed via supersymmetric localization. The formal expression is Cabo-Bizet:2018ehj  Assel:2014paa ; Closset:2013sxa

Z=eS(z)k=m,n= 0z+k(m+12)τ(n+12)σz+k+(m+12)τ+(n+12)σZ\,=\,e^{-S(z)}\,\equiv\,\prod_{k\,=\,-\infty}^{\infty}\,\prod_{m,\,n\,=\,0}^{\infty}\frac{z\,+\,k\,-\,(m\,+\,\frac{1}{2})\tau\,-\,(n\,+\,\frac{1}{2})\sigma}{z\,+\,k\,+\,(m+\frac{1}{2})\tau\,+\,(n\,+\,\frac{1}{2})\sigma}\, (179)

where SS is the divergent one-loop effective action. The complex parameters σ=σ1+iσ2\sigma=\sigma_{1}\,+\,{\rm i}\sigma_{2} and τ=τ1+iτ2\tau=\tau_{1}\,+\,{\rm i}\,\tau_{2} are in a beginning assumed to have positive imaginary part σ2>0\sigma_{2}>0 and τ2>0\tau_{2}>0\,. These parameters can be thought of as a complexification of the angular velocities of rotation around the two independent rotational axis of the three-sphere. The three integer numbers kkmm and nn label the modes associated to the non-contractible Euclidean time cycle and two contractible cycles of the three-sphere. For the moment zz is just a complex variable that in due time will be identified with a linear combination of the gauge variables uu and flavour and RR-symmetry chemical potentials. Momentarily, we will assume that σ=τ\sigma=\tau. Later on in appendix B.2 we will show that this analysis can be extended to infinitely many other cases.

For σ=τ\sigma=\tau the potential in (179) can be written as the following divergent series

S=(k,m)2(m+ 1)log(z+k+τ(m+1)).S\,=\,\sum_{(k,\,m)\,\in\,\mathbb{Z}^{2}}\,(m\,+\,1)\,\log\Bigl{(}z\,+\,k\,+\,\tau\,(m+1)\Bigr{)}\,. (180)

The series SS is formally related to

S^(k,m)2log(z+k+τ(m+1)),\widehat{S}\,\equiv\,\sum_{(k,\,m)\,\in\,\mathbb{Z}^{2}}\,\,\log\,\Bigl{(}z\,+\,k\,+\,\tau\,(m+1)\Bigl{)}\,, (181)

by the following linear partial differential equation

zS=τS^.\partial_{z}\,S\,=\,\nabla_{\tau}\,\widehat{S}\,. (182)

The derivative

ττeτ,eττzz,\nabla_{\tau}\,\equiv\,\partial_{\tau}\,-\,e_{\tau}\,,\qquad\qquad e_{\tau}\,\equiv\,\partial_{\tau}z\,\partial_{z}\,, (183)

is invariant under holomorphic re-parameterizations of zz as a function of τ\tau. The choice of “frame” eτe_{\tau} is part of the ambiguity in regularization. Different such choices define different regularizations. Let us choose a linear and holomorphic (in τ\tau) re-parameterization of the complex variable zz

z=χ+τξ,z\,=\,\chi\,+\,\tau\,\xi\,, (184)

with χ\chi and ξ\xi being two auxiliary complex variables that are assumed to be independent of τ\tau. Then, the differential equation (182) takes the form

zS=(τξz)S^.\partial_{z}\,S\,=\,(\partial_{\tau}\,-\,\xi\,\partial_{z})\,\widehat{S}\,. (185)

To ease the work, we restrict (185) to the domain

ξ,χ.\xi\,\in\,\mathbb{R}\,,\qquad\chi\,\in\,\mathbb{C}\,. (186)

We recall that one can always write a generic complex variable 𝔛\mathfrak{X} such as z,χz,\,\chi\, or ξ\xi, as a linear combination of 11 and τ\tau. The corresponding components are non-holomorphic functions of 𝔛\mathfrak{X} and τ\tau

𝔛1=𝔛¯τ𝔛τ¯ττ¯𝔛2=𝔛¯𝔛ττ¯.\begin{split}\mathfrak{X}_{1}&\,=\,\frac{\overline{\mathfrak{X}}\,\tau\,-\,\mathfrak{X}\,\overline{\tau}}{\tau\,-\,\overline{\tau}}\,\qquad\mathfrak{X}_{2}\,=\,\frac{\overline{\mathfrak{X}}\,-\,\mathfrak{X}}{\tau\,-\,\overline{\tau}}\,.\end{split} (187)

We note that χ\chi and ξ\xi are not the real components z1z_{1} and z2z_{2}. In fact z1=χ1z_{1}\,=\,\chi_{1} and z2=χ2+ξz_{2}\,=\,\chi_{2}\,+\,\xi, where χ1\chi_{1} and χ2\chi_{2} are the components 11 and 22 of χ\chi.

So far zz in (184) is just a mere complex variable. We need to define its relation to the physical variables i.e to the gauge variables uu and chemical potentials Δ\Delta’s . The relation will be

u+Δz=χ+τξ.u\,+\,\Delta\,\,\equiv\,z\,=\,\chi\,+\,\tau\,\xi. (188)

In particular we can always identify the components 11 and 22 of u+Δu\,+\,\Delta with the respective ones of z=χ+τξz\,=\,\chi\,+\,\tau\xi. For short, from now on we will use the notation z(u)u+Δz(u)\,\equiv\,u\,+\,\Delta.

The independent variable ξ\xi\in\mathbb{R} labels different regularization schemes, i.e defines different regularization conditions, each of them, associated to a differential equation (185) in the way that will be explained below.

To regularize we will follow three steps :

  1. 1.

    Regularize the real part of SS by demanding either double periodicity (in subsection B), or meromorphy (in subsection B) in uu. In both cases the regular answer turns out to be a real-harmonic function of τ\tau

    (τ12+τ22)ReS= 0.(\partial^{2}_{\tau_{1}}\,+\,\partial^{2}_{\tau_{2}})\,\text{Re}S\,=\,0\,. (189)
  2. 2.

    The regularized imaginary part of SS is constrained by requiring meromorphy in τ\tau.

  3. 3.

    There is a remaining τ\tau-meromorphic ambiguity that we denote as RR. We bound it to be, in both cases, of the form

    logR+iτfor τ+i,logR 0forτ+i 0.\begin{split}\log R\,\sim\,+{\rm i}\tau\sim-\infty\,\quad\text{for }\quad\tau\to+\,{\rm i}\,\infty\,,\qquad\log R\,\sim\,0\,\quad\text{for}\quad\tau\to+\,{\rm i}\,0\,.\end{split} (190)
515151As explained before, this condition can be relaxed. In principle, the only constraint on RR is that its real part must be an arbitrary real function of u2u_{2} and τ\tau. This ambiguity has to be fixed by imposition of a physical condition. In our case such condition is matching to the superconformal index. As will be shown in the next section, the condition (190) is only at half-way of the constraint that we must impose. To illustrate connections to previous regularization schemes we find illustrative to leave this ambiguity free for the moment.

We call RR the remainder function. The function logR\log R determines the large-τ\tau asymptotic behaviour of the regularized version of the action S(z)S(z).

There are two technical issues that we need to highlight before moving on

  • The infinities to regularize, come from a zero Fourier mode (m,n)=0(m,n)=0 in the double Fourier expansions below. The Poisson summation dual of the divergent expressions will make that clear. The first step will be to excise such mode. After that, finite results are obtained, but there is still a finite ambiguity that needs to be constrained by the properties of the physical observable that one wishes to compute.

  • Some of the regular series below do not converge fast enough. This means they converge to functions which are non smooth. To render such limit functions smooth at intermediate steps, further “regularization” is required. 525252In order to apply ABBV integration formula it is safer to use smooth functions. To solve that issue we insert an intermediate real cut-off 0Λ< 10\,\leq\,\Lambda\,<\,1. We will implicitly assume Λ= 1\Lambda\,=\,1^{-} at every step. See appendix F.

An elliptic scheme

We start by working out two particular schemes. The first one we call the elliptic scheme. It corresponds to the following choice of frame

ξ=Δ2,Δ2,\xi\,=\,\Delta_{2}\,,\qquad\Delta_{2}\,\in\,\mathbb{R}\,, (191)

where Δ2\Delta_{2} is the second real component of the complex chemical potential Δ=Δ1+Δ2τ\Delta\,=\,\Delta_{1}\,+\,\Delta_{2}\tau.

The divergent series SS and S^\widehat{S} are analytic in the independent complex variables zz and τ\tau. To avoid clumsiness in our presentation it is convenient to introduce the following two real variables xx and yy to denote the real and imaginary parts of zz i.e z=x+iyz=\,x\,+\,{\rm i}y\,. Thus, from (182) it follows that 535353For a holomorphic function h(z)h(z) of zz it follows that Rezh(z)=RezReh.\text{Re}\,\partial_{z}h(z)=\partial_{\text{Re}\,z}\text{Re}\,h\,.

xReS=(ξx+τ1)ReS^.\partial_{x}\,\text{Re}S\,=\,\,(-\,\xi\,\partial_{x}+\partial_{\tau_{1}})\,\text{Re}\widehat{S}\,. (192)

The series S^\widehat{S} is an average over the lattice τ+\mathbb{Z}\tau+\mathbb{Z}. Thus, at formal level we can recast it as a double Fourier expansion with the use of Poisson summation formula

(m,n)2f(x+k,y+m)=(m,n)2f^(m,n)𝐞(xm+yn).\sum_{(m,\,n)\,\in\,\mathbb{Z}^{2}}\,f(x+k,\,y+m)\,=\,\,\sum_{(m,\,n)\,\in\,\mathbb{Z}^{2}}\widehat{f}(m,n)\,{\bf e}(xm+yn)\,.

We define the Fourier transform of f(x,y)f(x,y) as f^(m,n)01𝑑x𝑑y𝐞(xmyn)f(x,y).\widehat{f}(m,n)\,\equiv\,\int_{0}^{1}dxdy\,{\bf e}(-xm-yn)f(x,y)\,. A computation shows that

ReS^=(k,m)2τ22π𝐞(nz2mz1)|mτ+n|2.\text{Re}\widehat{S}\,=\,-\sum_{(k,\,m)\,\in\,\mathbb{Z}^{2}}\,\frac{\tau_{2}}{2\pi}\,\frac{{\bf e}\bigl{(}nz_{2}\,-\,mz_{1}\bigr{)}}{|m\tau\,+\,n|^{2}}\,. (193)

This series is divergent due to the presence of the zero mode (m,n)=0(m,n)=0. Indeed, this is the origin of the divergences in the original series SS. From now on we excise the zero mode (m,n)=(0,0)(m,n)=(0,0) and proceed with the new convergent expression of S^\widehat{S} which can be presented in the form

ReS^log|P(z1,z2;τ)|,\text{Re}\widehat{S}\,\equiv\,\log|P(z_{1},z_{2};\tau)|\,, (194)

where

log|P(z1,z2;τ)|τ22πm,n(m,n)(0,0)𝐞(nz2mz1)|mτ+n|2,\log|P(z_{1},z_{2};\tau)|\equiv-\frac{\tau_{2}}{2\pi}\!{\underset{m,n\in\mathbb{Z}\atop(m,n)\neq(0,0)}{\sum}}\;\frac{{\bf e}(nz_{2}-mz_{1})}{|m\tau+n|^{2}}\,, (195)

is a well known real harmonic and real analytic modular invariant function of τ\tau.

If we plug (194) in the right-hand side of (192) we can integrate to obtain a regular expression for ReSS. We obtain two contributions. One contribution comes from the antiderivative with respect to xx of xReS^\partial_{x}\text{Re}\widehat{S}, which equals ξ-\xi times S^\widehat{S}, up to an arbitrary function of z2z_{2} and τ\tau,

ξ𝑑xxReS^=ξlog|P(z1,z2;τ)|+,-\xi\int\,dx\,\partial_{x}\,\text{Re}\widehat{S}\,=\,\,\,-\,\xi\,\log|P(z_{1},z_{2};\tau)|\,+\,\ldots\,, (196)

and a second comes from the antiderivative with respect to xx of the derivative of (194) with respect to τ1\tau_{1}, which equals

𝑑xτ1ReS^=log|Q(z1,z2;τ)|+,\int\,dx\,\partial_{\tau_{1}}\text{Re}\widehat{S}\,=\,\,\,\,\log|Q(z_{1},z_{2};\tau)|\,+\,\ldots\,, (197)

where

log|Q(z1,z2;τ)|iτ22π2m,n(m,n)(0,0)(mτ1+n)𝐞(nz2mz1)|mτ+n|4.\begin{split}\log|Q(z_{1},z_{2};\tau)|&\equiv\frac{{\rm i}\,\tau_{2}}{2\pi^{2}}\,{\underset{m,n\in\mathbb{Z}\atop(m,n)\neq(0,0)}{\sum}}\frac{\left(m\,\tau_{1}\,+\,n\right)\,{\bf e}\bigl{(}nz_{2}-mz_{1}\bigr{)}}{|m\tau+n|^{4}}\,.\end{split} (198)

This double periodic function is closely related to the Bloch-Wigner elliptic dilogarithm ZagierOnBloch , as was pointed out in Cabo-Bizet:2019eaf , see section 3.13.1 in that reference. This function is a real harmonic function of τ\tau but it is not modular covariant.

Collecting partial results we obtain the regularized expression for the real part of the action after evaluating ξ=Δ2\xi=\Delta_{2}

ReSell=log|Q(z1,z2;τ)|Δ2log|P(z1,z2;τ)|+log|R(z2;τ)|,\text{Re}S_{\text{ell}}\,=\,\log|Q(z_{1},z_{2};\tau)|\,-\,\Delta_{2}\,\log|P(z_{1},z_{2};\tau)|\,+\,\log|R(z_{2};\tau)|\,, (199)

where log|R|\log|R| depends on z2z_{2} and τ\tau. If we assume log|R|\log|R| to be the real part of a function that is meromorphic in τ\tau, then we must assume it to be a real harmonic function of τ\tau\,.

Holomorphy in τ\tau constraints the imaginary part of the action ImS\text{Im}S\,. As log|P|\log|P| and log|Q|\log|Q| are the real parts of (177) and (178) we obtain

Sell(z(u))=logQ(z(u))Δ2logP(z(u))+logR(z(u);τ).S_{\text{ell}}(z(u))\,=\,\log Q(z(u))\,-\,\Delta_{2}\,\log P(z(u))\,+\,\log R(z(u);\tau)\,. (200)

The asymptotic conditions (190) constraint logR\log R to be of the form

logR= 2πiτE(z2)+πiΦ(z1,z2),\log R\,=\,2\pi{\rm i}\tau E(z_{2})\,\,+\,\pi{\rm i}\Phi(z_{1},z_{2})\,, (201)

where Φ\Phi and EE are real functions. Note that the term proportional to EE dictates the asymptotic behaviour of SellS_{\text{ell}} at τ+i\tau\,\to\,+\,{\rm i}\,\infty. Note that when the function EE is a constant, we can interpret it as an insertion of a background charge (or energy). That is because 2πiτ2\pi{\rm i}\tau is the chemical potential dual to the BPS charge EE.

A meromorphic scheme

In this subsection we present a second scheme. It corresponds to the following choice

ξ=z2=u2+Δ2,z2,\xi\,=\,z_{2}\,=\,u_{2}\,+\,\Delta_{2}\,,\qquad z_{2}\,\in\,\mathbb{R}\,, (202)

where u2u_{2} and Δ2\Delta_{2} are one of the components of uu and Δ\Delta in their linear decomposition in terms of 11 and τ\tauu=u1+τu2u\,=\,u_{1}\,+\,\tau u_{2} and Δ=Δ1+τΔ2\Delta\,=\,\Delta_{1}\,+\,\tau\Delta_{2}\,.

The first two steps are identical to the ones of the previous section. Compute the anti-derivative with respect to xx of the term coming from ξx\xi\,\partial_{x} and obtain again (196). Then, compute the antiderivative with respect to xx of the term coming from τ1\partial_{\tau_{1}} which is again (197). Collecting partial results, and finally substituting ξ=z2=u2+Δ2\xi\,=\,z_{2}=u_{2}+\Delta_{2}, we obtain

ReS=ReSellu2log|P(z(u))|=log|Q(z(u))|(u2+Δ2)log|P(z(u))|+log|R(z2(u))|,\begin{split}\text{Re}S&\,=\,\text{Re}S_{\text{ell}}\,-\,u_{2}\log|P(z(u))|\\ &\,=\,\log|Q(z(u))|\,-\,(u_{2}+\Delta_{2})\,\log|P(z(u))|\,+\,\log|R(z_{2}(u))|\,,\end{split} (203)

where log|R|\log|R| is (before imposing boundary conditions (190)) an arbitrary function of z2z_{2} and harmonic in τ\tau. Finally, by demanding meromorphy in τ\tau we obtain

S=Sellu2logP(z(u))=logQ(z(u))z2(u)logP(z(u))+logR(z2(u)),\begin{split}S&\,=\,S_{\text{ell}}\,-\,u_{2}\log P(z(u))\\ &\,=\,\log Q(z(u))\,-\,z_{2}(u)\,\log P(z(u))\,+\,\log R(z_{2}(u))\,,\end{split} (204)

where again the asymptotic conditions (190) constraint logR\log R to be of the form

logR= 2πiτE(z2)+πiΦ(z1,z2),\log R\,=\,2\pi{\rm i}\tau E(z_{2})\,+\,\pi{\rm i}\Phi(z_{1},z_{2}), (205)

where Φ\Phi and EE are real functions. Note that this term dictates the asymptotic behaviour of SS at τ+i\tau\to+{\rm i}\infty\,. In the next section we will show that there exists a choice of RR such that for uu\in\mathbb{C}

eS(z(u))=Γe(z(u)+τ;τ,τ).\,e^{-S(z(u))}\,=\,\Gamma_{\text{e}}(z(u)+\tau;\tau,\tau)\,. (206)

In these two subsections, we have assumed σ=τ\sigma=\tau but that condition can be relaxed and the same regularization process applies (See appendix B.2).

B.2 Generalization to the case στ\sigma\,\neq\,\tau

In this appendix we show that the same regularization used for the cases σ=τ\sigma\,=\,\tau in section B can be adapted to more general cases with στ\sigma\neq\tau. This implies that everything we have said in this paper relative to the case σ=τ\sigma=\tau can be straightforwardly generalized to the more general case studied in this subsection.

Let σ=aω\sigma\,=\,a\omega and τ=bω\tau\,=\,b\omega with aa and bb positive integers. For simplicity we will assume aa and bb to be primes, but the same result applies to the generic case. Let us define z~=z+τ+σ2\widetilde{z}\,=\,z\,+\,\frac{\tau\,+\,\sigma}{2} and analyze the logarithm of (179)

km,n= 0log(z~+k+(mb+na)abω)log(z~+k(m+1b+n+1a)abω).\sum_{k\in\mathbb{Z}}\sum_{m,n\,=\,0}^{\infty}\log\Bigl{(}\widetilde{z}\,+\,k\,+\,\Bigl{(}\frac{m}{b}+\frac{n}{a}\Bigr{)}ab\omega\Bigr{)}\,-\,\log\Bigl{(}\widetilde{z}\,+\,k-\Bigl{(}\frac{m+1}{b}+\frac{n+1}{a}\Bigr{)}ab\omega\Bigr{)}\,. (207)

Here we have conveniently written maω+nbω=(mb+na)abωma\omega+nb\omega\,=\,\Bigl{(}\frac{m}{b}+\frac{n}{a}\Bigr{)}ab\omega. For positive integers mm and nn it is always possible to rewrite

mb=rmb+m~,na=sna+n~,\frac{m}{b}\,=\,\frac{r_{m}}{b}\,+\,\widetilde{m}\,,\qquad\frac{n}{a}\,=\,\frac{s_{n}}{a}\,+\,\widetilde{n}\,, (208)

with integers m~\widetilde{m}n~\widetilde{n}rmr_{m} and sns_{n} ranging over

m~ 0,n~ 0,rm= 0,,b1,sn= 1,,a1.\widetilde{m}\,\geq\,0\,,\quad\widetilde{n}\,\geq\,0\,,\quad r_{m}\,=\,0,\,\ldots\,,b-1\,,\quad s_{n}\,=\,1,\ldots,a-1\,. (209)

It is straightforward to check that (207) takes the form

r=0a1s=0b1km~,n~= 0log(z~r,s+k+(m~+n~)abω)log(z~r,s+k(m~+n~+ 2)abω),\sum_{r=0}^{a-1}\sum_{s=0}^{b-1}\sum_{k\in\mathbb{Z}}\sum_{\widetilde{m},\,\widetilde{n}\,=\,0}^{\infty}\log\Bigl{(}\widetilde{z}_{r,s}\,+\,k\,+\,\bigl{(}\widetilde{m}\,+\,\widetilde{n}\bigr{)}ab\omega\Bigr{)}\,-\,\log\Bigl{(}\widetilde{z}_{r,s}\,+\,k-\bigl{(}\widetilde{m}\,+\,\widetilde{n}\,+\,2\bigr{)}ab\omega\Bigr{)}\,, (210)

where

z~r,s=z~+raω+sbω.\widetilde{z}_{r,s}\,=\,\widetilde{z}\,+\,ra\omega\,+\,sb\omega\,. (211)

In reaching this expansion we have used the following redefinition in the argument of the second logarithm

m~after=b 1m~before,n~after=a 1n~before.\widetilde{m}_{\text{after}}\,=\,b\,-\,1\,-\,\widetilde{m}_{\text{before}}\,,\qquad\widetilde{n}_{\text{after}}\,=\,a\,-\,1\,-\,\widetilde{n}_{\text{before}}\,. (212)

As the summand of (210) only depends on the combination m~+n~\widetilde{m}\,+\,\widetilde{n} we can reorganize the sum as follows, by collecting degeneracies

r=0a1s=0b1(k,m)2(m+ 1)log(zr,sa,b+k+(m+1)abω).\sum_{r=0}^{a-1}\,\sum_{s=0}^{b-1}\,\sum_{(k,m)\in\mathbb{Z}^{2}}\,(m\,+\,1)\,\log\Bigl{(}z^{a,b}_{\,r,s}\,+\,k\,+\,(m+1)\,ab\omega\Bigr{)}\,. (213)

In this expression

zrsabz~r,sabω=z+(r+12)aω+(s+12)bωabω.z^{ab}_{rs}\,\equiv\,\widetilde{z}_{r,s}\,-\,ab\omega\,=\,z\,+\,\Bigl{(}r+\frac{1}{2}\Bigr{)}a\omega\,+\,\Bigl{(}s+\frac{1}{2}\Bigr{)}b\omega\,-\,ab\omega\,. (214)

Note that (213) is a combination of terms like the one we regularized before in subsection B, see equation (180). Thus, assuming (213) as starting point, one can regularize as explained in the previous subsections, but for generic values of τ\tau and σ\sigma, with a common complex divisor ω\omega.

Appendix C The real functions EE and Φ\Phi

Here we fix RR in such a way that the finite result coincides with the elliptic Gamma function Γe(z(u)+τ;τ,τ)\Gamma_{\text{e}}(z(u)+\tau;\tau,\tau) in the entire plane (u1,u2)(u_{1},u_{2}) with u=u1+τu2u\,=\,u_{1}\,+\,\tau u_{2}. The details are given in appendix F.2.

We choose the same scheme as before but modifying z(u)z(u) in (188) by substituting uu by a linear combination of NN complexified Cartan angles u¯\underline{u}ρ(u)\rho(u)ρ\rho is a weight. This gives again (204) with uu substituted by ρ(u)\rho(u). Next we must fix RR by demanding that for all uiu^{i}\in\mathbb{C}

Γe(z(u)+τ;τ,τ)=R1P0(z1,z2)z2Q0(z1,z2).\begin{split}\Gamma_{\text{e}}(z(u)+\tau;\tau,\tau)\,=\,{R}^{-1}\,\,\frac{\,\,P_{0}(z_{1},z_{2})^{z_{2}}}{Q_{0}(z_{1},z_{2})}\,.\\ \end{split} (215)

This equation is the exponential version of (204), after assuming (206), with the substitution uρ(u)u\,\mapsto\,\rho(u). This is the completion of the equation (2.28) of Cabo-Bizet:2020nkr Zudilin . There, this identity was defined up to a τ\tau-independent imaginary constant. The derivation of the remainder RR that fixes this pure phase ambiguity, will be postponed until appendix F.2. Here we will just quote the result.

We define the functions P0P_{0} and Q0Q_{0} via their logarithms

logP0=logP0(z1,z2)i2πm,nm0𝐞(nz2mz1)m(mτ+n),\log P_{0}\,=\,\log P_{0}(z_{1},z_{2})\,\equiv-\frac{{\rm i}}{2\pi}\!{\underset{m,n\in\mathbb{Z}\atop m\neq 0}{\sum}}\;\frac{{\bf e}(nz_{2}-mz_{1})}{m(m\tau+n)}\,, (216)

and

logQ0=logQ0(z1,z2)14π2m,nm0𝐞(nz2mz1)m(mτ+n)2.\log Q_{0}\,=\,\log Q_{0}(z_{1},z_{2})\,\equiv\,-\frac{1}{4\pi^{2}}\!{\underset{m,n\in\mathbb{Z}\atop m\neq 0}{\sum}}\;\frac{{\bf e}(nz_{2}-mz_{1})}{m(m\tau+n)^{2}}\,. (217)

They are related, via a redefinition of RR, to the functions logP\log P and logQ\log Q (See appendix A). In the left-hand side of (215) we define

z(u)=ρ(u)+Δ.z(u)\,=\,\rho(u)\,+\,\Delta\,. (218)

In the right-hand side, the variables z1z_{1} and z2z_{2} are related to the 22rk(G)(G) variables u¯=u¯1+τu¯2\underline{u}\,=\,\underline{u}_{1}\,+\,\tau\underline{u}_{2} and v¯\underline{v} as follows

z1(u,v)=z(u)ρ(v)τΔ2τ,z2(v)=ρ(v)+Δ2.\begin{split}\,z_{1}(u,v)&\,=\,z(u)\,-\,\rho(v)\,\tau\,-\,\Delta_{2}\,\tau\,,\\ \,z_{2}(v)&\,=\,\rho(v)\,+\,\Delta_{2}\,.\end{split} (219)

In the right-hand side of (215) the auxiliary variable vv is fixed in terms of uu i.e.

v¯=u¯2.\underline{v}\,=\,\underline{u}_{2}\,. (220)

Note that z1z_{1} is a real function of the real variables u¯1\underline{u}_{1} and not of u¯2\underline{u}_{2}\,. The R1R^{-1} then takes the following form

R1=R1(u,v)qE𝐞(Φ2),\begin{split}R^{-1}\,&\,=\,R^{-1}(u,v)\,\,\equiv\,q^{-E}\,\,{\bf e}\bigl{(}-\frac{\Phi}{2}\bigr{)}\,,\end{split} (221)

where EE is related to the following cubic polynomial in two variables

𝒜(u,z2(v))=B3(z2(v))3(ρ(u2)+Δ2)B2(z2(v))2\mathcal{A}(u,z_{2}(v))\,=\,\frac{B_{3}(z_{2}(v))}{3}\,-\,\bigl{(}\rho(u_{2})+\Delta_{2}\bigr{)}\,\frac{B_{2}(z_{2}(v))}{2} (222)

by the following relation

E𝒜(u,z2(v))+𝒜(u,{z2(v)}).E\,\equiv\,-\mathcal{A}(u,z_{2}(v))\,+\,\mathcal{A}(u,\{z_{2}(v)\})\,. (223)

The polynomial 𝒜\mathcal{A} is related to the contribution of the given multiplet to the cubic polynomial of anomaly coefficients. If we define

𝒜(u,z2(u))=16(z2(u)3z2(u)2),\mathcal{A}(u,z_{2}(u))\,\,=\,\,-\,\frac{1}{6}\,\Bigl{(}z_{2}(u)^{3}\,-\,\frac{z_{2}(u)}{2}\Bigr{)}\,, (224)

then the cubic polynomial of anomaly coefficients for flavour and R-charges can be defined as

𝒜anom=αρα𝒜(0,z2α(0)),\mathcal{A}_{anom}\,=\,\sum_{\alpha}\sum_{\rho_{\alpha}}\mathcal{A}(0,z_{2\alpha}(0))\,, (225)

where zα=ρα(u)+Δαz_{\alpha}\,=\,\rho_{\alpha}(u)+\Delta_{\alpha}. More details on this can be found in section 2.4 of Cabo-Bizet:2020nkr . Finally, the pure phase eπiΦe^{\pi{\rm i}\Phi} takes the form

𝐞(Φ2)𝐞(12B1({z1(u)})B2(z2(v)+1))𝐞(12B1({z1(u)})B2(z2(0+)+1)),𝐞(12B1(z1(u))B2(z2(v)+1))𝐞(12B1(z1(u))B2(z2(0+)+1)).\begin{split}{\bf e}\bigl{(}-\frac{\Phi}{2}\bigr{)}&\,\equiv\,\frac{{\bf e}\Bigl{(}-\,\frac{1}{2}B_{1}(\{z_{1}(u)\})\,B_{2}(\lfloor z_{2}(v)\rfloor+1)\Bigr{)}}{{\bf e}\Bigl{(}-\,\frac{1}{2}B_{1}(\{z_{1}(u)\})\,B_{2}(\lfloor z_{2}(0^{+})\rfloor+1)\Bigr{)}}\,,\\ &\equiv\frac{{\bf e}\Bigl{(}-\,\frac{1}{2}B_{1}(z_{1}(u))\,B_{2}(\lfloor z_{2}(v)\rfloor+1)\Bigr{)}}{{\bf e}\Bigl{(}-\,\frac{1}{2}B_{1}(z_{1}(u))\,B_{2}(\lfloor z_{2}(0^{+})\rfloor+1)\Bigr{)}}\,.\end{split} (226)

In going to the second line we have used the 2πi2\pi{\rm i}\mathbb{Z}\,-shift ambiguity of the exponent. In the main body of the paper we reinstated this ambiguity in the form of a piecewise integer function that we denoted with the letter KK (See equation (77)).

C.1 The universal contribution at large-NN of RR: The susy Casimir energy

In a gauge anomaly-free theory, the u2u_{2}-dependence cancels out after summing over multiplets. In those theories, the non-vanishing contribution of EE to the supersymmetric partition function comes from the following two terms in (223). From the first, one obtains

𝒜anom.-\,\mathcal{A}_{anom}\,. (227)

The other contribution comes from the sum over matter multiplets of the term

𝒜(u,{z2(u2)})=B3({ρ(u2)+Δ2})3(ρ(u2)+Δ2)B2({ρ(u2)+Δ2})2.\mathcal{A}(u,\{z_{2}(u_{2})\})\,=\,\frac{B_{3}(\{\rho(u_{2})+\Delta_{2}\})}{3}\,-\,\bigl{(}\rho(u_{2})+\Delta_{2}\bigr{)}\,\frac{B_{2}(\{\rho(u_{2})+\Delta_{2}\})}{2}\,. (228)

The contribution (227), is the cubic polynomial of flavour and R-symmetry anomaly coefficients. This is a universal contribution to the large-NN effective action of complex saddles, that is linear in τ\tau Cabo-Bizet:2020nkr . As pointed out in Cabo-Bizet:2020nkr , this contribution matches the susy Casimir energy of Assel:2015nca .

For the complex fixed-points PP that we know contribute,

𝒜(u,{z2(u2)})|u=PN0.\mathcal{A}(u,\{z_{2}(u_{2})\})\Big{|}_{u\,=\,P}\quad\underset{N\,\to\,\infty}{\longrightarrow}\quad 0\,. (229)

Let us explain how this happens. For (m,n)(m,n) fixed points with m>0m>0, the ρ(u2)\rho(u_{2}) is always of the form

ijK,i,j=1,,K,\frac{i-j}{K},\qquad i,j=1,\ldots,K\in\mathbb{Z}\,, (230)

where KK is a divisor of NN such that KK\to\infty as NN\to\infty. In that limit the sum over ii and jj can be reduced to an integral. (229) follows from the identities

01𝑑xBk(x)= 0,01𝑑xxBk(x)= 0.\int^{1}_{0}dxB_{k}(x)\,=\,0\,,\qquad\int^{1}_{0}dx\,x\,B_{k}(x)\,=\,0\,. (231)

Thus, for the meromorphic scheme

E(P)N𝒜anom.E(P)\quad\underset{N\,\to\,\infty}{\longrightarrow}\quad-\mathcal{A}_{anom}\,. (232)

This contribution is the same for every PP. For the (1,0)(1,0) this contribution is not present.

Appendix D Another choice of regularization: Γλ\Gamma_{\lambda}

As we have mentioned before. The regularization ambiguity that we have in the form of a choice of frame eτe_{\tau} and remainder RR, includes other schemes that are not the elliptic and meromorphic ones. 545454 Note that part of the degeneracy corresponds to reparametrizations of S1S_{1} upon S1S_{1} that preserve one point. The point being a representation of the real contour of integration. In this subsection we define a family of schemes that corresponds to the deformed action SλS_{\lambda} of section 3.

The choice of frame in this case is

ξ=ρ({u2}λ)+Δ2,ρ({u2}λ)+Δ2.\xi\,=\,\rho(\{u_{2}\}_{\lambda})\,+\,\Delta_{2}\,,\qquad\rho(\{u_{2}\}_{\lambda})\,+\,\Delta_{2}\in\,\mathbb{R}\,. (233)

By {u2¯}λ\{\underline{u_{2}}\}_{\lambda} we mean the vector with components {u2i}λ\{u^{i}_{2}\}_{\lambda}\,; and for xx\in\mathbb{R} and 0λ<10\leq\lambda<1 was defined in (75).

Here we comment on three relevant properties of this function:

  • a.

    The {x}λ\{x\}_{\lambda} is a smooth function for 0λ<10\leq\lambda<1, and it is periodic {x+1}λ={x}λ\{x+1\}_{\lambda}=\{x\}_{\lambda}.

  • b.

    When x0+x\to 0^{+} the {xi}λ0\{x^{i}\}_{\lambda}\to 0\,.

  • c.

    When λ0+\lambda\to 0^{+} the {x}λ0\{x\}_{\lambda}\to 0\,.

Repeating the same regularization steps used in appendix B.1, and introducing the cut-off Λ\Lambda in the final double series expansions we obtain

eSλ=Γλ(z(u))RΛ1P0Λ(z1,z2)z2({u2}λ)Q0Λ(z1,z2).e^{-S_{\lambda}}\,=\,{\Gamma_{\lambda}}(z(u))\,\equiv\,{R}_{\Lambda}^{-1}\,\,\frac{\,\,P_{{0}\Lambda}(z_{1},z_{2})^{z_{2}(\{u_{2}\}_{\lambda})}}{Q_{0\Lambda}(z_{1},z_{2})}\,. (234)

As it will be explained in great detail in appendix F, to ensure continuity and smoothness of their derivatives, we will need to introduce a cut-off Λ\Lambda in the double series representation of the logarithms of P0P_{0}Q0Q_{0}, and R1R^{-1} i.e for 0λ<10\leq\lambda<1\,, we define

logP0Λ=i2πm,nm0Λ|m|+|n|𝐞(nz2mz1)m(mτ+n).\begin{split}\log P_{0\Lambda}&\,=\,-\frac{{\rm i}}{2\pi}\!{\underset{m,n\in\mathbb{Z}\atop m\neq 0}{\sum}}\;\Lambda^{|m|\,+\,|n|}\,\frac{{\bf e}(nz_{2}-mz_{1})}{m(m\tau+n)}\,.\end{split} (235)
logQ0Λ(z;τ)14π2m,nm0Λ|m|+|n|𝐞(nz2mz1)m(mτ+n)2.\begin{split}\log Q_{0\Lambda}(z;\tau)&\,\equiv\,-\frac{1}{4\pi^{2}}\!{\underset{m,n\in\mathbb{Z}\atop m\neq 0}{\sum}}\;\Lambda^{|m|\,+\,|n|}\,\frac{{\bf e}(nz_{2}-mz_{1})}{m(m\tau+n)^{2}}\,.\end{split} (236)

Moreover, we choose

RΛ1=qEΛ𝐞(ΦΛ2),\begin{split}R^{-1}_{\Lambda}&=q^{-E_{\Lambda}}\,\,{\bf e}{\left(-\frac{\Phi_{\Lambda}}{2}\right)}\,,\end{split} (237)

where

qEΛqB3(z2({u2}Λ)Λ+1)B3(z2(0+)+1)3××𝐞(τ({u2}Λ+Δ2)Δ2{u2}Λ+Δ2𝑑ξξ(mΛ|m|𝐞(mξ))).𝐞(ΦΛ2)𝐞(B1Λ(z1(u)))Δ2{u2}Λ+Δ2dξξ(mΛ|m|𝐞(mξ))).\begin{split}q^{-E_{\Lambda}}&\,\equiv\,q^{\frac{B_{3}(\lfloor z_{2}(\{u_{2}\}_{\Lambda})\rfloor_{\Lambda}+1)\,-\,B_{3}(\lfloor z_{2}(0^{+})\rfloor+1)}{3}}\,\times\,\\ &\qquad\qquad\hfill\times\,{\bf e}\left(-\,\tau\,(\{u_{2}\}_{\Lambda}+\Delta_{2})\,\int^{\{u_{2}\}_{\Lambda}\,+\,\Delta_{2}}_{\Delta_{2}}d\xi\,\xi\,\,\Bigl{(}\sum_{m\in\mathbb{Z}}\,{\Lambda}^{|m|}\,{\bf e}(m\,\xi)\Bigr{)}\right)\,.\\ {\bf e}{\left(-\frac{\Phi_{\Lambda}}{2}\right)}&\,\equiv\,{\bf e}\left(-\,B_{1\Lambda}(z_{1}(u)))\,\,\int^{\{u_{2}\}_{\Lambda}\,+\,\Delta_{2}}_{\Delta_{2}}d\xi\,\xi\,\,\Bigl{(}\sum_{m\in\mathbb{Z}}\,{\Lambda}^{|m|}\,{\bf e}(m\,\xi)\Bigr{)}\right)\,.\end{split} (238)

Here by Λ\lfloor\cdot\rfloor_{\Lambda} denotes a generic smooth version of the floor function. We note that for 0Λ<10\leq\Lambda<1\,, the right-hand side of equation (237) is not holomorphic in uu when v=u2v=u_{2}, but that is not a problem. Basic algebraic manipulations upon (238) imply that

limΛ 1vi(2QΛ+ΦΛ)= 0+c.t..\underset{\Lambda\,\to\,1}{\lim}\,\partial_{v^{i}}(2\,Q_{\Lambda}\,+\,\Phi_{\Lambda})\,=\,0\,+\,\text{c.t.}\,. (239)

Here vi=u2iτu1i\partial_{v^{i}}\,=\,\partial_{u^{i}_{2}}\,-\,\tau\,\partial_{u^{i}_{1}} . This relation was used in reaching (83) in the main body of the paper.

Finally, properties a.a.b.b. and cc imply three relevant properties of the deformed action SλS_{\lambda}

  • a. implies that the SλS_{\lambda} is smooth in the torus A×BA\times B.

  • b. implies that the value of SλS_{\lambda} at the real cycle of integration u2i=vi=0+u^{i}_{2}=v^{i}=0^{+} is independent of λ\lambda\,.

  • c. implies that in limit λ0+\lambda\to 0^{+} the SλSellS_{\lambda}\to S_{\text{ell}}\,. Where SellS_{\text{ell}} is essentially the elliptic deformation of Cabo-Bizet:2019eaf ; Cabo-Bizet:2020nkr 555555Up to a term linear in τ\tau that does not contribute to the on-shell effective actions of (m,n)(m,n) fixed points and that can be reinstated by a trivial redefinition of the remainder function RR.

Appendix E Cauchy-Riemann conditions

In this section we will prove equations (269) and (280), in that order, with the use of Cauchy-Riemann relations. Let zz\in\mathbb{C}, we define its real and imaginary parts as

Rez=xz1+τ1z2,Im(z)=y=τ2z2.\text{Re}z\,=\,x\equiv z_{1}+\tau_{1}z_{2}\,,\quad\text{Im}(z)\,=\,y\,=\,\tau_{2}\,z_{2}\,. (240)

Let logG=U+iV\log G\,=\,U+{\rm i}V be a holomorphic function of zz. Then, its real and imaginary parts UU and VV, obey the Cauchy-Riemann relations

UxVy= 0,Uy+Vx= 0.\frac{\partial U}{\partial x}\,-\,\frac{\partial V}{\partial y}\,=\,0\,,\quad\frac{\partial U}{\partial y}\,+\,\frac{\partial V}{\partial x}\,=\,0\,. (241)

In terms of the variables z1z_{1} and z2z_{2} these relations read

Uz1+τ1τ2Vz11τ2Vz2= 0,1τ2Uz2τ1τ2Uz2+Vz1= 0.\frac{\partial U}{\partial z_{1}}\,+\,\frac{\tau_{1}}{\tau_{2}}\frac{\partial V}{\partial z_{1}}\,-\,\frac{1}{\tau_{2}}\frac{\partial V}{\partial z_{2}}\,=\,0\,,\quad\frac{1}{\tau_{2}}\frac{\partial U}{\partial z_{2}}\,-\,\frac{\tau_{1}}{\tau_{2}}\frac{\partial U}{\partial z_{2}}\,+\,\frac{\partial V}{\partial z_{1}}\,=\,0\,. (242)

Let us define the real function UU as

U=RelogGRelogQ0(z1,z2;τ)+z2RelogP0(z1,z2;τ).\begin{split}U\,=\,\text{Re}\log G\,&\equiv\,-\,\text{Re}\log Q_{0}(z_{1},z_{2};\tau)\,+\,z_{2}\,\text{Re}\log P_{0}(z_{1},z_{2};\tau)\,.\end{split} (243)

For z2z_{2}\notin\mathbb{Z} the harmonic dual to UU can be checked to be, up to a τ\tau dependent constant

V=(m,n)2m 02πz2Re(mτ+n)cos(2π(mz1nz2))4π2m|mτ+n|2(m,n)2m 0(Re2(mτ+n)Im2(mτ+n))sin(2π(mz1nz2))4π2m|mτ+n|4=ImlogQ0(z1,z2;τ)+z2ImlogP0(z1,z2;τ).\begin{split}V\,=\,&-\sum_{(m,n)\,\in\,\mathbb{Z}^{2}\atop m\,\neq\,0}\frac{2\pi z_{2}\text{Re}(m\tau+n)\,\cos\left(2\pi\left(mz_{1}-nz_{2}\right)\right)}{4\pi^{2}m|m\tau+n|^{2}}\\ &\,-\,\sum_{(m,n)\,\in\,\mathbb{Z}^{2}\atop m\,\neq\,0}\,\frac{\Bigl{(}\text{Re}^{2}\bigl{(}m\tau+n\bigr{)}-\text{Im}^{2}\bigl{(}m\tau+n\bigr{)}\Bigr{)}\sin\left(2\pi\left(mz_{1}-nz_{2}\right)\right)}{4\pi^{2}m|m\tau+n|^{4}}\\ &\,=\,-\,\text{Im}\log Q_{0}(z_{1},z_{2};\tau)\,+\,z_{2}\,\text{Im}\log P_{0}(z_{1},z_{2};\tau)\,.\end{split} (244)

Thus, the function

G(z;τ)P0(z1,z2;τ)z2Q0(z1,z2;τ)G(z;\tau)\equiv\frac{P_{0}(z_{1},z_{2};\tau)^{z_{2}}}{Q_{0}(z_{1},z_{2};\tau)}\, (245)

is meromorphic for z2z_{2}\notin\mathbb{Z}. Moreover, in Cabo-Bizet:2019eaf (See below equation (3.20) in that reference) we have shown that

Γe(z1+τ;τ,τ)=1Q0(z1,0+;τ).\Gamma_{\text{e}}(z_{1}+\tau;\tau,\tau)\,=\,\frac{1}{Q_{0}(z_{1},0^{+};\tau)}\,. (246)

Let us assume that 0<z2<10<z_{2}<1. In this region we can take the limit z20+z_{2}\to 0^{+}. From equations (245) and (246) it follows that in such limit

G(z1+0+τ;τ)=Γe(z1+0+τ+τ;τ,τ).G(z_{1}+0^{+}\tau;\tau)\,=\,\Gamma_{\text{e}}(z_{1}+0^{+}\tau+\tau;\tau,\tau)\,. (247)

From this initial condition and uniqueness of solutions to the Cauchy-Riemann equations it follows that for zz\in\mathbb{C} and 0<z2<10<z_{2}<1

G(z;τ)=Γe(z+τ;τ,τ)=P0(z1,z2;τ)z2Q0(z1,z2;τ).G(z;\tau)\,=\,\Gamma_{\text{e}}(z\,+\,\tau;\tau,\tau)\,=\,\frac{P_{0}(z_{1},z_{2};\tau)^{z_{2}}}{Q_{0}(z_{1},z_{2};\tau)}\,. (248)

Next we note that

Γe(z1+1τ+τ;τ,τ)Γe(z1+0+τ+τ;τ,τ)=θ0(z1+1τ;τ)=P0(z1,1;τ).\frac{\Gamma_{\text{e}}(z_{1}+1^{-}\tau+\tau;\tau,\tau)}{\Gamma_{\text{e}}(z_{1}+0^{+}\tau+\tau;\tau,\tau)}\,=\,\theta_{0}(z_{1}+1^{-}\tau;\tau)\,=\,P_{0}(z_{1},1^{-};\tau)\,. (249)

A computation shows that the real and imaginary parts of P0(z1,z2;τ)P_{0}(z_{1},z_{2};\tau) obey the Cauchy-Riemann relations (242) for z2z_{2}\notin\mathbb{Z}. Then, from the initial condition in the last equality in (249) and uniqueness theorem, it follows that for 0<z2<10<z_{2}<1

θ0(z1+z2τ;τ)=P0(z1,z2;τ).\theta_{0}(z_{1}+z_{2}\tau;\tau)\,=\,P_{0}(z_{1},z_{2};\tau)\,. (250)

From (248) and (250) it follows that for z2z_{2}\notin\mathbb{Z}

Γe(z1+{z2}τ+τ;τ,τ)=P0(z1,z2;τ){z2}Q0(z1,z2;τ),\Gamma_{\text{e}}(z_{1}\,+\,\{z_{2}\}\tau\,+\,\tau;\tau,\tau)\,=\,\frac{P_{0}(z_{1},z_{2};\tau)^{\{z_{2}\}}}{Q_{0}(z_{1},z_{2};\tau)}\,, (251)

and

θ0(z1+{z2}τ;τ)=P0(z1,z2;τ).\theta_{0}(z_{1}+\{z_{2}\}\tau;\tau)\,=\,P_{0}(z_{1},z_{2};\tau)\,. (252)

Using quasi-periodicity properties of Γe\Gamma_{\text{e}} and θ0\theta_{0} one can find other identities of these kind like for instance

Γe(z1+τ({z2}+1);τ,τ)=P0(z;τ){z2}Q0(z;τ).\Gamma_{\text{e}}(z_{1}+\tau(-\{-z_{2}\}+1);\tau,\tau)\,=\,\frac{\,\,P_{0}(z;\tau)^{-\{-z_{2}\}}}{Q_{0}(z;\tau)}\,. (253)

E.1 Cancelling discontinuities and contact terms

The property (173) is particularly useful to generate contact terms that need to be cancelled in order to relate the double periodic and meromorphic functions. From this identity it is easy to note that if one composes the function BpB_{p} with the ceiling function evaluated at real number xx the new composed function is piecewise constant with jumps (from left to right) of length (1)p+1mp(-1)^{p+1}m^{p} at every integer value mm of xx. This new function can be used to generate primitive functions Bp(x)B_{p}(x) of contact terms of the form

m(1)pmpδ(x+m)=1p+1xBp+1(x+ 1),\sum_{m}\,(-1)^{p}\,m^{p}\delta(x\,+\,m)\,=\,\frac{1}{p+1}\,\partial_{x}B_{p+1}(\lfloor x\rfloor\,+\,1)\,, (254)

where the primitive is defined as

Bp+1(x+ 1){Bp+1(x+ 1)forxBp+1(m+ 1)+Bp+1(m)2forx=m.B^{\circ}_{p+1}(\lfloor x\rfloor\,+\,1)\,\equiv\,\left\{\begin{array}[]{lllr}B_{p+1}(\lfloor x\rfloor\,+\,1)&&\text{for}&\qquad x\notin\mathbb{Z}\\ \frac{B_{p+1}(m\,+\,1)\,+\,B_{p+1}(m)}{2}&&\text{for}&\qquad x\,=\,m\in\mathbb{Z}\end{array}\right.\,. (255)

The case x=mx=m\in\mathbb{Z} will be of relevance in due time, but not for the technical manipulations that will be presented in this appendix. Thus, from now on in this appendix we will ignore the supra-index \circ i.e. we will ignore the difference between BpB^{\circ}_{p}\, and BpB_{p}. Only in due time (in the main part of the manuscript) when remarking the difference between the two will turn out to be of relevance to reach partial conclusions, we will reinstate the symbol \circ.

In the main body of the paper relation (254) turns out to be relevant at various instances. One example being the following. Suppose we want to compute the primitive of

B1(u1)mδ(v+m)=(B1(u)vτ)mδ(v+m)=B1(u)mδ(v+m)+τmmδ(v+m),\begin{split}B_{1}(u_{1})\sum_{m}\delta(v\,+\,m)&\,=\,(B_{1}(u)\,-\,v\tau)\sum_{m}\delta(v\,+\,m)\\ &\,=\,B_{1}(u)\sum_{m}\delta(v\,+\,m)\,+\,\tau\sum_{m}\,m\,\delta(v\,+\,m)\,,\end{split} (256)

with respect to the variable vv at fixed uu and evaluate the result at the section u2=vu_{2}=v. Then, from property (254) it follows that the answer is

B1(u)B1(v+1)τ2B2(v+1).B_{1}(u)B_{1}(\lfloor v\rfloor+1)\,-\,\frac{\tau}{2}B_{2}(\lfloor v\rfloor+1)\,. (257)

A second relevant example is the following one. Suppose we want to compute the primitive of

vB1(u1)mδ(v+m)=v(B1(u)vτ)mδ(v+m)=B1(u)mmδ(v+m)τmm2δ(v+m),\begin{split}v\,B_{1}(u_{1})\sum_{m}\delta(v\,+\,m)&\,=\,v\,(B_{1}(u)\,-\,v\tau)\sum_{m}\delta(v\,+\,m)\\ &\,=\,-\,B_{1}(u)\sum_{m}m\delta(v\,+\,m)\,-\,\tau\sum_{m}\,m^{2}\,\delta(v\,+\,m)\,,\end{split} (258)

with respect to the variable vv at fixed uu and evaluate the result on the section u2=vu_{2}=v. From property (254) it follows that the answer is

12B1(u)B2(v+1)τ3B3(v+1).\frac{1}{2}B_{1}(u)B_{2}(\lfloor v\rfloor+1)\,-\,\frac{\tau}{3}B_{3}(\lfloor v\rfloor+1)\,. (259)
A more subtle case

These two previous examples are closely related to another two examples that turn out to be relevant, from a technical viewpoint, to ones for our scope. The first example is computing the primitive of

B1({u1})mδ(v+m)=B1({uvτ})mδ(v+m)|u2=v,B_{1}(\{u_{1}\})\sum_{m}\delta(v\,+\,m)\,=\,B_{1}(\{u-v\tau\})\sum_{m}\delta(v\,+\,m)\Big{|}_{u_{2}=v}\,, (260)

with respect to the variable vv at fixed uu and evaluating the result at the section u2=vu_{2}=v. We first must define what we mean by

{uvτ},\{u-v\tau\}\,, (261)

for u2vu_{2}\neq v when the argument of the bracket is not real. In that case we define for xx\in\mathbb{C}

{x}xP1[x],P1[x]τ¯xτx¯τ¯τ,\{x\}\equiv x-\lfloor P_{1}[x]\rfloor\,,\qquad P_{1}[x]\,\,\equiv\frac{\overline{\tau}x-\tau\overline{x}}{\overline{\tau}-\tau}\,\in\,\mathbb{R}\,, (262)

where P1P_{1} is the projector operator i.e. P1=P12P_{1}=P_{1}^{2}, that projects x=x1+x2τx=x_{1}+x_{2}\tau to x1x_{1}.

From definition (262) it follows that {uvτ}={u}vτ\{u-v\tau\}\,=\,\{u\}\,-\,v\tau, and the primitive of (260) can be written as

12πilogRP(u,v)B1({u})(B1(v+1)+c1)τ2(B2(v+1)+c2).\frac{1}{2\pi{\rm i}}\log R_{P}(u,v)\,\equiv\,B_{1}(\{u\})\bigl{(}B_{1}(\lfloor v\rfloor+1)\,+\,c_{1}\bigr{)}\,-\,\frac{\tau}{2}\bigl{(}B_{2}(\lfloor v\rfloor+1)\,+\,c_{2}\bigr{)}\,. (263)

The c1=c1(u)c_{1}=c_{1}(u) and c2=c2(u)c_{2}=c_{2}(u) are vv-independent quantities that we can fix at convenience. For any such choice we get

(v(12πilogRP(u,v))))|u2=v=B1({u1})mδ(v+m).\left(\partial_{v}\Bigl{(}\frac{1}{2\pi{\rm i}}\log R_{P}(u,v))\Bigr{)}\right)\Big{|}_{u_{2}=v}\,=\,B_{1}(\{u_{1}\})\sum_{m}\delta(v\,+\,m)\,. (264)

Suppose we want to compute the primitive of

vB1({u1})mδ(v+m),\begin{split}v\,B_{1}(\{u_{1}\})\sum_{m}\delta(v\,+\,m)\,,\end{split} (265)

with respect to the variable vv at fixed uu, and evaluate the result at the section u2=vu_{2}=v. Then, the answer is

12πilogRvP(u,v)12B1({u})(B2(v+1)+c1)τ3(B3(v+1)+c2).\frac{1}{2\pi{\rm i}}\log R_{vP}(u,v)\,\equiv\,\frac{1}{2}B_{1}(\{u\})\bigl{(}B_{2}(\lfloor v\rfloor+1)\,+\,c_{1}\bigr{)}\,-\,\frac{\tau}{3}\bigl{(}B_{3}(\lfloor v\rfloor+1)\,+\,c_{2}\bigl{)}\,. (266)

The c1=c1(u)c_{1}=c_{1}(u) and c2=c2(u)c_{2}=c_{2}(u) are vv-independent quantities that we can fix at convenience. Namely

(v(12πilogRvP(u,v))))|u2=v=vB1({u1})mδ(v+m).\left(\partial_{v}\Bigl{(}\frac{1}{2\pi{\rm i}}\log R_{vP}(u,v))\Bigr{)}\right)\Big{|}_{u_{2}=v}\,=\,vB_{1}(\{u_{1}\})\sum_{m}\delta(v\,+\,m)\,. (267)

Appendix F Double periodic and meromorphic functions

In this section we study the relation between the double periodic functions P0P_{0} (essentially the PP that was introduced in the elliptic regularization scheme) and Q0Q_{0} (essentially the QQ that was introduced in the elliptic regularization scheme) and θ0\theta_{0} and Γe\Gamma_{\text{e}}\,565656We shall call these double periodic functions and sometimes (non-holomorphic) elliptic functions ZagierOnBloch .

The precise definition of P0P_{0} and Q0Q_{0} will be given below. The log|P0|\log|P_{0}| is closely related to the real-analytic Kronecker-Eisenstein series (See equation (177)). The function log|Q0|\log|Q_{0}| is closely related to the Bloch-Wigner elliptic dilogarithm (See equation (178)). Understanding the relations with θ0\theta_{0} and Γe\Gamma_{\text{e}}, is useful to understand how to construct a family of functions on the torus A×BA\times B that flows to the meromorphic extension of the integrand of (11) (truncated at a portion of a cylinder). The functions logQ0\log Q_{0} and logP0\log P_{0} can be used as building blocks of such family of double periodic extensions. Interestingly, the relation of θ0\theta_{0} and Γe\Gamma_{\text{e}} to P0P_{0} and Q0Q_{0} allows us to use the latter two to understand how the former two behave in a vicinity of the real axis Imτ=0\text{Im}\tau=0 and further extend them to the lower-half plane. Existence of this extension is important in order to be able to define the contour of integration in τ\tau-plane that could be used to extract the microcanonical index out of the single exponential blocks of the ABBV formula for the grand canonical one, see section 5.

The properties of P0P_{0}

The P0P_{0} is a double-periodic extension of the Jacobi theta function θ0\theta_{0}. Equivalently, it can be defined via the following logarithmic branch

logP0(z)=logP0(z1,z2)i2πm,nm0𝐞(nz2mz1)m(mτ+n).\log P_{0}(z)\,=\,\log P_{0}(z_{1},z_{2})\,\equiv-\frac{{\rm i}}{2\pi}\!{\underset{m,n\in\mathbb{Z}\atop m\neq 0}{\sum}}\;\frac{{\bf e}(nz_{2}-mz_{1})}{m(m\tau+n)}\,. (268)

This function is discontinuous at z2z_{2}\in\mathbb{Z}. The discontinuity is clear from the following identity, which holds for z2z_{2}\notin\mathbb{Z}

P0(z)=θ0(z1+{z2}τ;τ).P_{0}(z)\,=\,\theta_{0}(z_{1}+\{z_{2}\}\tau;\tau)\,. (269)

The curly bracket means {z2}z2z2\{z_{2}\}\equiv z_{2}-\lfloor z_{2}\rfloor.

The identity (269) fails to hold at z2z_{2}\in\mathbb{Z} but it holds in the limits from the left and the right to z2z_{2}\in\mathbb{Z}. However, that is not an issue for the purpose of computing integrals. That the (269) fails to hold at z2z_{2}\in\mathbb{Z} has a simple explanation: The θ0(z1+z2τ;τ)\theta_{0}(z_{1}+z_{2}\tau;\tau) is a continuous function of z2z_{2}. After composition with {z2}\{z_{2}\} the limits from the right at z2z_{2}\in\mathbb{Z} coincide with the value of the function at the point due to continuity of θ0\theta_{0} and continuity from the right of the function {}\{\cdot\}.

On the other hand the function P0P_{0} is not continuous neither from the right nor from the left at z2z_{2}\in\mathbb{Z}. The series (268) diverges at (z1,z2)2(z_{1},z_{2})\in\mathbb{Z}^{2}. Its real part diverges to -\infty at such positions. These are the positions of the zeroes of θ0\theta_{0}. For any other value of zsz^{\prime}s this series converges. However, if z2z_{2}\in\mathbb{Z} and z1z_{1}\notin\mathbb{Z} this series does not converge uniformly. In that case Dirichlet theorem states that the series converges to a function that is discontinuous at z2z_{2}\in\mathbb{Z}, and the value of the function at the discontinuity equals the semi-sum of the limits from both sides of the discontinuity. This is, the function P0P_{0} is not continuous at z2z_{2}\in\mathbb{Z}, neither from the right nor from the left, and its value at the discontinuity differs from that of θ0(z1+{z2};τ)\theta_{0}(z_{1}+\{z_{2}\};\tau).

The discontinuity at z2z_{2}\in\mathbb{Z} from the left and the right of the double series representation (268) of logP0\log P_{0}, originates from the fact that the Fourier coefficients of these series do not decrease fast enough for large values of mm and nn. We solve this with the insertion of a cut-off Λ\Lambda. We will not make it explicit at every step but we will work, unless we explicitly say otherwise, with the Λ\Lambda-smoothened Levin ; BrownLevin double Fourier expansions

logP0Λ=i2πm,nm0Λ1|m|Λ2|n|𝐞(nz2mz1)m(mτ+n).\begin{split}\log P_{0\Lambda}&\,=\,-\frac{{\rm i}}{2\pi}\!{\underset{m,n\in\mathbb{Z}\atop m\neq 0}{\sum}}\;\Lambda_{1}^{|m|}\,\Lambda_{2}^{|n|}\,\frac{{\bf e}(nz_{2}-mz_{1})}{m(m\tau+n)}\,.\end{split} (270)

The real cut-offs 0<Λ1<10<\Lambda_{1}<1 and 0<Λ2<10<\Lambda_{2}<1, must be taken to Λ1=1\Lambda_{1}=1^{-} and Λ2=1\Lambda_{2}=1^{-} at the very end. The smoothened version logP0Λ\log P_{0\Lambda} and its derivatives are absolutely and thus uniformly convergent for every real z1z_{1} and z2z_{2}. Thus, logP0Λ\log P_{0\Lambda} is smooth i.e. even divergences are also resolved by the insertion of Λ\Lambda. There is a price to pay though as the piecewise meromorphy of logP0\log P_{0} gets spoiled in an infinitesimal band of width of order 1Λ1-\Lambda centered at z2z_{2}\in\mathbb{Z}. We give an analytic description of this in terms of the cut-offs in appendix F.3. Next we will do it with distributions

Recovering θ0\theta_{0} from P0P_{0}

The right-hand side of (269) can be triviallly de-periodised by using quasi-periodic properties of θ0\theta_{0} (See appendix F). Here we would like to prove that, starting from the double Fourier expansion of P0P_{0} and upon the requirement of meromorphy. Quasi-periodicity will be an outcome. It is also a convenient introduction for the slightly less trivial case of elliptic Gamma functions that will be presented next.

Let us introduce the following new auxiliary complex variables u=u1+τu2u=u_{1}+\tau u_{2}\,, v=u2v=u_{2} and the derivatives

u=u1,v=u2τu1.\partial_{u}\,=\,\partial_{u_{1}}\,,\qquad\partial_{v}\,=\,\partial_{u_{2}}\,-\,\tau\,\partial_{u_{1}}\,. (271)

A function of uu and vv is a holomorphic function of uu if its derivative with respect to vv vanishes.

After acting on logP0\log P_{0}, with v\partial_{v}, using the Fourier expansion of B1({z})B_{1}(\{z\}) given in (168), and the Poisson summation identity

m𝐞(mz)=mδ(z+m),\sum_{m\in\mathbb{Z}}{\bf e}(mz)\,=\,\sum_{m\in\mathbb{Z}}\delta(z+m)\,, (272)

we obtain the following identity

vlogP0=  2πiB1({u1})(mδ(v+m)).\begin{split}\partial_{v}\,\log P_{0}&\,=\,\,2\pi{\rm i}\,B_{1}(\{u_{1}\})\,\Bigl{(}\sum_{m\in\mathbb{Z}}\,\delta(v\,+\,m)\Bigr{)}\,.\end{split} (273)

where the derivative v\partial_{v} is taken at fixed uu\,\in\,\mathbb{C} and u2=vu_{2}\,=\,v. In this expression u1u_{1}\,\in\,\mathbb{R} stands for the combination uvτu\,-\,v\,\tau with uu\in\mathbb{C} and u2=vu_{2}\,=\,v. Equation (273) is telling us that P0P_{0} is meromorphic in uu for every u2u_{2}\,\notin\,\mathbb{Z}. It is also telling us that the lack of meromorphy corresponds to a discontinuity located at u2u_{2}\,\in\,\mathbb{Z}.

This discontinuity can be cancelled with the help of the function RPR_{P} that we define in (263). Namely, the combination P0RP1P_{0}\,R^{-1}_{P} is meromorphic in uu. Moreover, from the relation (269) and uniqueness of solutions to Cauchy-Riemann relations, it follows that for complex uu

θ0(u;τ)=Rθ01P0(uvτ,v),v=u2,\begin{split}\theta_{0}(u;\tau)\,=\,R^{-1}_{\theta_{0}}\,P_{0}(u-v\tau,v)\,,\qquad v=u_{2}\,,\\ \end{split} (274)

where

Rθ01RP1(u,v)RP1(u,0+)=𝐞(12v(12{u}+(v+1)τ)).R^{-1}_{\theta_{0}}\,\equiv\,\frac{R^{-1}_{P}(u,v)}{R^{-1}_{P}(u,0^{+})}\,=\,{\bf e}\Bigl{(}\frac{1}{2}\lfloor v\rfloor(1-2\{u\}+(\lfloor v\rfloor+1)\tau)\Bigr{)}\,. (275)

Note that {u}\{u\} can be safely substituted by uu when evaluating the exponential. Using the bracket corresponds to a vv-dependent choice of branch for the corresponding logarithm.

Relation (274) can proven as follows: in virtue of (269) the function

P0(uvτ,v)RP1(u,v)RP1(u,0+)\frac{P_{0}(u-v\tau,v)R^{-1}_{P}(u,v)}{R^{-1}_{P}(u,0^{+})} (276)

equals the function θ0(u)\theta_{0}(u) at v=0+v=0^{+} and uu\in\mathbb{R}. Finally, Cauchy-Riemann relations in the complex extension of the variable uu, with the component u2u_{2} identified with vv, imply equality (284) for all uu\in\mathbb{C} with v=u2v=u_{2}. A much simpler way to prove it, by using quasi-periodicity properties of θ0\theta_{0} can be found in appendix F.

Properties of Q0Q_{0}

The P0P_{0} and Q0Q_{0} determine a double-periodic extension of the elliptic Gamma function Γe\Gamma_{e} (as will be recalled below). We can define Q0Q_{0} via the following logarithmic branch

logQ0(z)=logQ0(z1,z2)14π2m,nm0𝐞(nz2mz1)m(mτ+n)2.\log Q_{0}(z)\,=\,\log Q_{0}(z_{1},z_{2})\,\equiv\,-\frac{1}{4\pi^{2}}\!{\underset{m,n\in\mathbb{Z}\atop m\neq 0}{\sum}}\;\frac{{\bf e}(nz_{2}-mz_{1})}{m(m\tau+n)^{2}}\,. (277)

This function is continuous, and it has no divergences for u1u_{1}u2u_{2}\,\in\,\mathbb{R}. Its derivatives have discontinuities at u2u_{2}\,\in\,\mathbb{Z}. Indeed, after acting with v=u2τu1\partial_{v}\,=\,\partial_{u_{2}}\,-\,\tau\partial_{u_{1}} on the right-hand side of (277) one obtains

vlogQ0=logP0.\partial_{v}\log Q_{0}\,=\,\log P_{0}\,. (278)

Then, the previous analysis for logP0\log P_{0} applies. We should stress that P0P_{0} and  Q0Q_{0} are always thought to be functions of u1u_{1} and u2u_{2}. The meaning of v\partial_{v} in (278) is that of considering the complex linear combination of u1\partial_{u_{1}} and u2\partial_{u_{2}} as stated in equation (271).

Although we will not make it explicit at every step, we will always work, unless we indicate explicitly otherwise, with the smooth functions

logQ0Λ(z;τ)14π2m,nm0Λ1|m|Λ2|n|𝐞(nz2mz1)m(mτ+n)2.\begin{split}\log Q_{0\Lambda}(z;\tau)&\,\equiv\,-\frac{1}{4\pi^{2}}\!{\underset{m,n\in\mathbb{Z}\atop m\neq 0}{\sum}}\;\Lambda_{1}^{|m|}\,\Lambda_{2}^{|n|}\,\frac{{\bf e}(nz_{2}-mz_{1})}{m(m\tau+n)^{2}}\,.\end{split} (279)

The real cut-offs Λ1\Lambda_{1} and Λ2\Lambda_{2} are the same as before. The smooth version logQ0Λ\log Q_{0\Lambda} and its derivatives are absolutely and thus uniformly convergent for every real z1z_{1} and z2z_{2}. This is, the function logQ0Λ\log Q_{0\Lambda} is smooth. In the very end the cut-offs must be taken to one.

Towards fixing the remainder function in (204)

In appendix E we have shown the following identity which holds for z2z_{2}\notin\mathbb{Z}

P0(z;τ){z2}Q0(z;τ)=Γe(z1+τ({z2}+1);τ,τ).\frac{\,\,P_{0}(z;\tau)^{\{z_{2}\}}}{Q_{0}(z;\tau)}\,=\,\Gamma_{\text{e}}(z_{1}+\tau(\{z_{2}\}+1);\tau,\tau)\,. (280)

Here  {z2}z2z2\{z_{2}\}\equiv z_{2}-\lfloor z_{2}\rfloor and

P0{z2}exp({z2}logP0).P_{0}^{\{z_{2}\}}\,\equiv\,\exp(\{z_{2}\}\,\log P_{0})\,. (281)

The function Q0Q_{0} is understood to be the exponential of logQ0\log Q_{0}. This identity is saying that the combination in the left-hand side is piecewise meromorphic in zz. Here, we construct a combination of P0P_{0} and Q0Q_{0} that is meromorphic. This is, let us de-periodise relation (280). In appendix F we do so by using quasi-periodic properties of θ0\theta_{0} and Γe\Gamma_{\text{e}}. Here, we would like to do it starting from a deformation of the natural “de-periodization” of the left-hand side of (280), upon demanding meromorphy. Quasi-periodicity will be an outcome.

We start by noticing the following identity which holds for vv\in\mathbb{R}

vlog(P0vQ0)=  2πivB1({u1})(mδ(v+m)).\partial_{v}\log\Bigl{(}\frac{\,\,P_{0}^{\,v\,}}{Q_{0}}\Bigr{)}\,=\,\,2\pi{\rm i}\,v\,B_{1}(\{u_{1}\})\,\Bigl{(}\sum_{m\in\mathbb{Z}}\,\delta(v\,+\,m)\Bigr{)}\,. (282)

We define

P0vexp(vlogP0).P_{0}^{\,v\,}\equiv\,\exp(v\,\log P_{0})\,. (283)

Equation (282) follows from the definitions of P0vP^{v}_{0}Q0Q_{0} and (273). As before, the derivative v\partial_{v} is taken at fixed uu\in\mathbb{C} and  u2=vu_{2}=v. In this expression u1u_{1}\in\mathbb{R} stands for the combination uvτu\,-\,v\tau with  uu\in\mathbb{C} and u2=vu_{2}=v. Equation (273) is telling us that P0vQ0\frac{P_{0}^{v}}{Q_{0}} is holomorphic in uu for every u2u_{2}\notin\mathbb{Z}\,. It is also telling us that the lack of holomorphy corresponds to a discontinuity located at u2u_{2}\in\mathbb{Z}\,.

This discontinuity can be cancelled with the help of the function RvPR_{vP} that we define in (266). Namely, the combination P0vQ0RvP1\frac{P^{v}_{0}}{Q_{0}}\,R^{-1}_{vP} is holomorphic in uu\,575757…. whenever the expression in terms of double Fourier expansions converges. If such expression diverges at some point, then that point is guarantied to be a pole. Moreover, from the relation (269) and uniqueness of solutions to Cauchy-Riemann relations, it follows that for uu\in\mathbb{C}

Γe(u+τ;τ,τ)=RΓe1P0(uvτ,v)vQ0(uvτ,v),v=u2.\begin{split}\Gamma_{\text{e}}(u+\tau;\tau,\tau)\,=\,R_{\Gamma_{\text{e}}}^{-1}\,\,\frac{P_{0}(u-v\tau,v)^{v}}{Q_{0}(u-v\tau,v)}\,,\qquad v=u_{2}\,.\end{split} (284)

The remainder function (205) in this case is fixed to be

RΓe1RvP1(u,v)RvP1(u,0+)=qB3(v+1)3𝐞(12B1({u})B2(v+1))qB3(0++1)3𝐞(12B1({u})B2(0++1))=qB3(v+1)3𝐞(12B1({u})(B2(v+1)16))=qB3(v+1)3𝐞(12B1(u)(B2(v+1)16)),\begin{split}R^{-1}_{\Gamma_{\text{e}}}\,\equiv\,\frac{R^{-1}_{vP}(u,v)}{R^{-1}_{vP}(u,0^{+})}&\,=\,\frac{q^{\frac{B_{3}(\lfloor v\rfloor+1)}{3}}\,{\bf e}\Bigl{(}-\,\frac{1}{2}B_{1}(\{u\})B_{2}(\lfloor v\rfloor+1)\Bigr{)}}{q^{\frac{B_{3}(\lfloor 0^{+}\rfloor+1)}{3}}\,{\bf e}\Bigl{(}-\,\frac{1}{2}B_{1}(\{u\})B_{2}(\lfloor 0^{+}\rfloor+1)\Bigr{)}}\,\\ &\,=\,q^{\frac{B_{3}(\lfloor v\rfloor+1)}{3}}\,{\bf e}\Bigl{(}-\,\frac{1}{2}B_{1}(\{u\})\bigl{(}B_{2}(\lfloor v\rfloor+1)\,-\,\frac{1}{6}\bigr{)}\Bigr{)}\,\\ &\,=\,q^{\frac{B_{3}(\lfloor v\rfloor+1)}{3}}\,{\bf e}\Bigl{(}-\,\frac{1}{2}B_{1}(u)\bigl{(}B_{2}(\lfloor v\rfloor+1)\,-\,\frac{1}{6}\bigr{)}\Bigr{)}\,,\end{split} (285)

where RvP1R^{-1}_{vP} was defined in (266).

Note that if u1u_{1}\notin\mathbb{Z}, then {u}\{u\} in the second line can be safely substituted by uu in the third, when evaluating the exponential. 585858Definition of the {u}\{u\} when uu is complex has been given in appendix E.1 equation (262). Using the bracket corresponds to picking up a uu-dependent choice of branch for the corresponding logarithm.

Relation (284) is proven as follows: In virtue of (280)

RvP1(u,v)RvP1(u,0+)P0(uvτ,v)vQ0(uvτ,v)\,\frac{R^{-1}_{vP}(u,v)}{R^{-1}_{vP}(u,0^{+})}\,\frac{P_{0}(u-v\tau,v)^{v}}{Q_{0}(u-v\tau,v)} (286)

equals the function Γe(u+τ;τ,τ)\Gamma_{\text{e}}(u+\tau;\tau,\tau) at  v=0+v=0^{+} and uu\in\mathbb{R}. Finally, Cauchy-Riemann relations in the complex extension of the variable uu, with the component u2u_{2} identified with vv, implies equality (284) for all uu\in\mathbb{C} with v=u2v=u_{2}.

Comments on RΓeR_{\Gamma_{\text{e}}}

Note that if we change the definition of (285) by multiplying it by a phase 𝐞(K){\bf e}(K) where we assume KK to be an arbitrary piecewise-constant function taking values in the integers, then identity (284) also applies to the new definition of RΓeR_{\Gamma_{\text{e}}}. Obviously, this is because 𝐞(K)= 1{\bf e}(K)\,=\,1. As we have said, by using this ambiguity, which is a choice of branch, one can remove the bracket from {u}\{u\} in the exponent of (285).

F.1 Holomorphic extension of θ0\theta_{0} and Γe\Gamma_{\text{e}} to the lower-half of τ\tau-plane

We pause and comment about a subtle but important result that is implicit in the previously proven identities. Notice that the product representation (161) is absolutely convergent for Im(τ)>0\text{Im}(\tau)>0. This does not mean that the elliptic gamma functions Γe\Gamma_{\text{e}} can not be continued to Im(τ)=0\text{Im}(\tau)=0. In fact Γe\Gamma_{\text{e}} is well-defined at infinitely many points in the real axis Im(τ)=0\text{Im}(\tau)=0. That was shown in subsection 3.5 of Felder2000 . Specifically they showed that (at least) for irrational values of τ\tau the function Γe(z;τ+iϵ)\Gamma_{\text{e}}(z;\tau+{\rm i}\epsilon) is continuous as the infinitesimal real ϵ\epsilon crosses 0 from the negatives to the positives and viceversa. This result is consistent with Identity (284).

Moreover, the right-hand side of (284) can be understood as the definition of Γe\Gamma_{\text{e}} in the lower-half plane Im(τ)<0\text{Im}(\tau)<0 i.e. as its analytic continuation to the lower half-plane. From (284) it follows that rational τ\tau’s are singularities of Γe\Gamma_{\text{e}}. This means that holomorphic flows to the lower-half τ\tau-plane are possible when they go across irrational values of τ\tau. That claim is consistent with the results of Felder2000 .

Everything we have just said for (284) relative to Γe\Gamma_{\text{e}}, applies also to the representation (274) of θ0\theta_{0}.

The existence of these holomorphic flows from the upper to the lower half τ\tau-plane is a highly non-trivial property that proves to be very hard to show existence of, in many cases. Identities (274) and (284) demonstrate that such extensions exist for the quasi-elliptic functions θ0\theta_{0} and Γe\Gamma_{\text{e}}. Their existence will be important to the analysis regarding large-NN counting of operators that will be presented in section 5. It will allow us to define the contour of integration CηC_{\eta} in section 5.

From these identities one can see that the index \mathcal{I} will have essential singularities at rational values of τ\tau. Take for instance τ=0\tau=0. The asymptotic behaviour of the index at such point depends on how one approaches it. This was previously pointed out in Cabo-Bizet:2019osg . More generally, for rational values τ=nm\tau=-\frac{n}{m} an analogous conclusion holds, as pointed out in Cabo-Bizet:2020nkr . We encountered this issue in section 5. There it was further argued that these essential singularities are of paramount importance for operator counting.

Moreover, the relation of θ0\theta_{0} and Γe\Gamma_{\text{e}} to P0P_{0} and Q0Q_{0} allows us to use the latter two to understand how the former two behave in a vicinity of the real axis Im(τ)=0\text{Im}(\tau)=0 and further extend them to the lower-half plane.

Relation between Q0Q_{0} and Γe\Gamma_{\text{e}} from quasi-periodicity

In this appendix we prove (284) with the use of quasi-periodicity properties of Γe\Gamma_{\text{e}}. We start from the logarithm of relation (251)

logG(z1+τ{z2};τ)=logΓe(z1+τz2+τ;τ,τ)=logQ0(z;τ)+{z2}logP0(z;τ)=logQ0(z;τ)+{z2}logθ0(z1+τ{z2};τ).\begin{split}\log G(z_{1}+\tau\{z_{2}\};\tau)&\,=\,\log\Gamma_{\text{e}}(z_{1}+\tau z_{2}+\tau;\tau,\tau)\,=\,-\log Q_{0}(z;\tau)+\{z_{2}\}\log P_{0}(z;\tau)\,\\ &\,=\,-\log Q_{0}(z;\tau)+\{z_{2}\}\log\theta_{0}(z_{1}+\tau\{z_{2}\};\tau)\,.\end{split} (287)

Next we prove a couple of identities that will help in “de-periodising” (287). Assume without loss of generality that z2>0\lfloor z_{2}\rfloor>0 then

logθ0(z1+τ{z2};τ)=logθ0(z;τ)+2πii=1z2(ziτn2)=logθ0(z;τ)+2πiz2(zn2)2πiτz2(z2+1)2.\begin{split}\log\theta_{0}(z_{1}+\tau\{z_{2}\};\tau)&\,=\,\log\theta_{0}(z;\tau)+2\pi{\rm i}\sum_{i=1}^{\lfloor z_{2}\rfloor}(z-i\tau-\frac{n}{2})\\ &\,=\,\log\theta_{0}(z;\tau)+2\pi{\rm i}\lfloor z_{2}\rfloor(z-\frac{n}{2})-2\pi{\rm i}\tau\frac{\lfloor z_{2}\rfloor(\lfloor z_{2}\rfloor+1)}{2}\,.\end{split} (288)

The final answer for the case z2<0\lfloor z_{2}\rfloor<0 is the same. In this equation nn\in\mathbb{Z} is an arbitrary quantity that corresponds to a choice of branch. Note that nn can be a discontinuous function of zz. Note that identity (288) is a version of (275) in terms of logarithms after identifying zz and uu and z2z_{2} with vv. The second identity is

logG(z1+τ{z2};τ)=logG(z;τ)i=0z21logθ0(ziτ;τ)=logG(z;τ)i=0z21(logθ0(z;τ)+2πij=1i(zjτn2))=logG(z;τ)z2logθ0(z;τ)2πiz2(z21)2(zn2)+2πiτ6z2(z221).\begin{split}\log G(z_{1}+\tau\{z_{2}\};\tau)&\,=\,\log G(z;\tau)-\sum_{i=0}^{\lfloor z_{2}\rfloor-1}\log\theta_{0}(z-i\tau;\tau)\\ &\,=\,\log G(z;\tau)-\sum_{i=0}^{\lfloor z_{2}\rfloor-1}\bigl{(}\log\theta_{0}(z;\tau)+2\pi{\rm i}\sum_{j=1}^{i}(z-j\tau-\frac{n}{2})\bigr{)}\\ &\,=\,\log G(z;\tau)-\lfloor z_{2}\rfloor\log\theta_{0}(z;\tau)-2\pi{\rm i}\frac{\lfloor z_{2}\rfloor(\lfloor z_{2}\rfloor-1)}{2}(z-\frac{n}{2})\\ &\qquad\,+\frac{2\pi{\rm i}\tau}{6}\lfloor z_{2}\rfloor(\lfloor z_{2}\rfloor^{2}-1)\,.\end{split} (289)

Finally, from (287), (288) and (289) it follows that

logΓe(z1+τz2+τ;τ,τ)=logQ0(z1,z2;τ)+z2logP0(z1,z2;τ)+logRΓe1,\log\Gamma_{\text{e}}(z_{1}+\tau z_{2}+\tau;\tau,\tau)\,=\,-\log Q_{0}(z_{1},z_{2};\tau)\,+\,z_{2}\log P_{0}(z_{1},z_{2};\tau)\,+\,\log R_{\Gamma_{\text{e}}}^{-1}\,, (290)

where

logRΓe1=2πiτ3B3(z2+1)πi(zn2)(B2(z2+1)16).\log R_{\Gamma_{\text{e}}}^{-1}\,=\,\,\frac{2\pi{\rm i}\tau}{3}B_{3}(\lfloor z_{2}\rfloor+1)\,-\,\pi{\rm i}\,\bigl{(}z\,-\,\frac{n}{2}\bigr{)}\Bigl{(}B_{2}(\lfloor z_{2}\rfloor+1)\,-\,\frac{1}{6}\Bigr{)}. (291)

After the following identifications z=u=u1+τvz\,=\,u\,=\,u_{1}+\tau v, one can check that the exponential of (291) maps to (285). We note that the two equations were computed in different ways. Equation (285) was computed via “differential” methods, meanwhile (291) was computed via quasi-periodicity properties. They both represent the same object, which as a priori expected, can be recast as a remainder function of the regularization procedure introduced in section B.1.

F.2 Placing the discontinuities at u=0u=0 for every ρ\rho

This corresponds to a proper choice of integration constants c1(u)c_{1}(u) and c2(u)c_{2}(u), piece-wise continuous functions of uu, to be more precise, in the solution to equation (267) given in (266). Let us start by assuming 1<Δ2<1-1<\Delta_{2}<1\,. A choice of remainder function RR that solves equation (215) is

R1=R1(u,v)RvP1(z(u),z2(v))RvP1(z(u),z2(0+))=qB3(z2(v)+1)3𝐞(12B1({z(u)})B2(z2(v)+1))qB3(z2(0+)+1)3𝐞(12B1({z(u)})B2(z2(0+)+1)),=qB3(z2(v)+1)3𝐞(12B1(z(u))B2(z2(v)+1))qB3(z2(0+)+1)3𝐞(12B1(z(u))B2(z2(0+)+1)),\begin{split}R^{-1}\,&\,=\,R^{-1}(u,v)\,\equiv\,\,\frac{R^{-1}_{vP}(z(u),z_{2}(v))}{R^{-1}_{vP}(z(u),z_{2}(0^{+}))}\\ \ &\,=\,\frac{q^{\frac{B_{3}(\lfloor z_{2}(v)\rfloor+1)}{3}}\,{\bf e}\Bigl{(}-\,\frac{1}{2}B_{1}(\{z(u)\})B_{2}(\lfloor z_{2}(v)\rfloor+1)\Bigr{)}}{q^{\frac{B_{3}(\lfloor z_{2}(0^{+})\rfloor+1)}{3}}\,{\bf e}\Bigl{(}-\,\frac{1}{2}B_{1}(\{z(u)\})B_{2}(\lfloor z_{2}(0^{+})\rfloor+1)\Bigr{)}}\,,\\ &\,=\,\frac{q^{\frac{B_{3}(\lfloor z_{2}(v)\rfloor+1)}{3}}\,{\bf e}\Bigl{(}-\,\frac{1}{2}B_{1}(z(u))B_{2}(\lfloor z_{2}(v)\rfloor+1)\Bigr{)}}{q^{\frac{B_{3}(\lfloor z_{2}(0^{+})\rfloor+1)}{3}}\,{\bf e}\Bigl{(}-\,\frac{1}{2}B_{1}(z(u))B_{2}(\lfloor z_{2}(0^{+})\rfloor+1)\Bigr{)}}\,,\end{split} (292)

where RvP1R_{vP}^{-1} was defined in (266). In going from the second to the third line we have used the 2πi2\pi{\rm i}\mathbb{Z}\,-shift ambiguity of the exponent. Moreover, as explained in appendix E.1 below (255), when the argument of the function floorfloor hits integer values, we must use the definitions of B2B^{\circ}_{2} and B3B^{\circ}_{3} that were given in (255), instead of the Bernoulli polynomials B2B_{2} and B3B_{3}\,. This is what Dirichlet theorem indicates to do. The bracket function of the complex variable z(u)z(u) is defined by replacing the real component z1z_{1} by {z1}\{z_{1}\} and leaving z2z_{2} unchanged (See definition (262)). The term in the denominator of (292) is independent of vv. It corresponds to a choice of integration constant that guarranties R11R^{-1}\to 1 to vanish as vi0+v^{i}\to 0^{+}.

For 1<Δ2<1-1<\Delta_{2}<1 this choice of R1R^{-1} corresponds to the choice c1=16c_{1}=-\frac{1}{6} and c2= 0c_{2}\,=\,0 in (266).

To prove that the choice (292) solves (215) we proceed as follows. Start by noticing that for vv\in\mathbb{R}

vilog(P0z2(v)Q0)=  2πiρiz2(v)B1({z(u)})(mδ(z2+m)).\partial_{v^{i}}\log\Bigl{(}\frac{\,\,P_{0}^{\,z_{2}(v)\,}}{Q_{0}}\Bigr{)}\,=\,\,2\pi{\rm i}\,\rho^{i}\,z_{2}(v)\,B_{1}(\{z(u)\})\,\Bigl{(}\sum_{m\in\mathbb{Z}}\,\delta(z_{2}\,+\,m)\Bigr{)}\,. (293)

This derivative is taken at fixed uiu^{i}\,. The Dirac-delta term comes from a discontinuity of the function upon which vi\partial_{v^{i}} is acting. Such discontinuity is cancelled by adding a multiplying pre-factor R1R^{-1}, the one defined in (292), to the discontinuous function P0z2(v)/Q0{\,\,P_{0}^{\,z_{2}(v)\,}}/{Q_{0}}. This was explained in appendix E.1.

Finally, in virtue of identity (280), it follows that upon substitution of (292) in (215), the latter holds at least for v=0+v=0^{+} for 0Δ2<10\leq\Delta_{2}<1. This together with uniqueness of solutions to Cauchy-Riemann relations, imply that (215) holds for any uu\in\mathbb{C} and v=u2v=u_{2}. More precisely, in virtue of (280) and for 0Δ2<10\leq\Delta_{2}<1 the function

RvP1(z(u),z2(v))RvP1(z(u),z2(0+))P0(z1(u,v),z2(v))z2(v)Q0(z1(u,v),z2(v))\,\frac{R^{-1}_{vP}(z(u),z_{2}(v))}{R^{-1}_{vP}(z(u),z_{2}(0^{+}))}\,\frac{P_{0}(z_{1}(u,v),z_{2}(v))^{z_{2}(v)}}{Q_{0}(z_{1}(u,v),z_{2}(v))} (294)

equals Γe(ρ(u)+Δ+τ;τ,τ)\Gamma_{\text{e}}(\rho(u)+\Delta+\tau;\tau,\tau) at v=0+v=0^{+} and uu\in\mathbb{R}. At last, Cauchy-Riemann relations in the complex extension of the variable uu, with the component u2u_{2} identified with vv, imply equality (215) to hold upon using the definition of RR given in (292), for all uu\in\mathbb{C} and v=u2v=u_{2}\,.

Refer to caption
Refer to caption
Figure 7: The plots of the real (left) and imaginary (right) parts of the left- and right-hand sides of (215) for gauge group U(1)U(1) and ρ(u)=u\rho(u)=u, as a function of 0<u1<10<u_{1}<1 and 0<u2<10<u_{2}<1, and for τ=1+i/3\tau=1+{\rm i}/3Δ1=0.4325\Delta_{1}=0.4325Δ2=0.54731\Delta_{2}=0.54731. The blue (resp. green) points are obtained out of the left (resp. right)-hand side of (215). We have used the product representation of Γe(z;τ,τ)\Gamma_{\text{e}}(z;\tau,\tau) for the left hand side. Both product and double Fourier representations were truncated at level 2020. More precisely, the product representation has been truncated at the term q20q^{20}.

If Δ2\Delta_{2} is not in between 0 and 11 we can not identify the limit v=0+v=0^{+} of (294) with the left-hand side of (280) (for uu\in\mathbb{R}). In that case the proof of (215) needs to be slightly modified by using an alternative to (280) that we wrote in (253). After identifying the limit v=0+v=0^{+} of (294) with the left-hand side of that equation (for uu\in\mathbb{R}) it follows that (215) holds also when Δ2\Delta_{2} ranges in between 1-1 and 0 i.e. 1<Δ20-1<\Delta_{2}\leq 0\,.

A non Abelian formula for θ0\theta_{0} in terms of P0P_{0}

From completeness we present a non Abelian formula that is analogous to (215).

This one relates the θ0\theta_{0} and P0P_{0} when evaluated at non Abelian arguments z(u)=ρ(u)+Δz(u)=\rho(u)+\Delta, for any value of uiu^{i}’s and 0Δ2<10\leq\Delta_{2}<1. The formula is

θ0(z(u);τ)=Rθ01P0(z1,z2),v=u2,\begin{split}\theta_{0}(z(u);\tau)\,=\,{R}_{\theta_{0}}^{-1}\,P_{0}(z_{1},z_{2})\,,\qquad v=u_{2}\,,\\ \end{split} (295)

where z1z_{1} and z2z_{2} are the ones defined in (219) and

Rθ01=Rθ01(u,v)RP1(z(u),z2(v))RP1(z(u),z(0+))=qB2(z2(v)+ 1)2𝐞(B1({z(u)})B1(z2(v)+1))qB2(z2(0+)+ 1)2𝐞(B1({z(u)})B1(z2(0+)+1))=qB2(z2(v)+ 1)2𝐞(B1(z(u))B1(z2(v)+1))qB2(z2(0+)+ 1)2𝐞(B1(z(u))B1(z2(0+)+1)).\begin{split}{R}_{\theta_{0}}^{-1}&\,=\,{R}^{-1}_{\theta_{0}}(u,v)\,\equiv\,\frac{R^{-1}_{P}(z(u),z_{2}(v))}{R^{-1}_{P}(z(u),z(0^{+}))}\,\\ &\,=\,\frac{q^{\frac{B_{2}(\lfloor z_{2}(v)\rfloor\,+\,1)}{2}}\,{\bf e}{\Bigl{(}-B_{1}(\{z(u)\})B_{1}(\lfloor z_{2}(v)\rfloor+1)\Bigr{)}}}{q^{\frac{B_{2}(\lfloor z_{2}(0^{+})\rfloor\,+\,1)}{2}}\,{\bf e}{\Bigl{(}-B_{1}(\{z(u)\})B_{1}(\lfloor z_{2}(0^{+})\rfloor+1)\Bigr{)}}}\,\\ &\,=\,\frac{q^{\frac{B_{2}(\lfloor z_{2}(v)\rfloor\,+\,1)}{2}}\,{\bf e}{\Bigl{(}-B_{1}(z(u))B_{1}(\lfloor z_{2}(v)\rfloor+1)\Bigr{)}}}{q^{\frac{B_{2}(\lfloor z_{2}(0^{+})\rfloor\,+\,1)}{2}}\,{\bf e}{\Bigl{(}-B_{1}(z(u))B_{1}(\lfloor z_{2}(0^{+})\rfloor+1)\Bigr{)}}}\,.\end{split} (296)

In going from the second to the third line we have used the 2πi2\pi{\rm i}\mathbb{Z}\,-shift ambiguity of the exponent. This formula can be proven by following analogous steps to the ones followed in reaching (215). It can also be checked to follow from (269) and the use of quasi-periodic properties of θ0\theta_{0}\,.

F.3 Analysis in the presence of cut-off

In this subsection we arrive to the relations (273) and (278) with the use of cut-offs 0<Λ1,2<10<\Lambda_{1,2}<1595959We are abusing notation in the sense that now we consider two independent cut-offs Λ1\Lambda_{1} and Λ2\Lambda_{2} instead of one. The single cut-off mentioned in the main body of this paper corresponds to Λ1=Λ2=Λ\Lambda_{1}=\Lambda_{2}=\Lambda.

The second of these relations follows trivially from (270) and (279)

vlogQ0Λ=logP0Λ.\begin{split}\partial_{v}\log Q_{0\Lambda}\,=\,\log P_{0\Lambda}\,.\end{split} (297)

The first one follows from

vlogP0Λ=m,nm0Λ1|m|Λ2|n|𝐞(nu2mu1)m= 2πiB1Λ({u1})nΛ2|n|𝐞(nu2)= 2πiB1Λ({u1})1Λ221+Λ22 2Λ2cos(2πu2).\begin{split}\,\partial_{v}\log P_{0\Lambda}&\,=\,\!{\underset{m,n\in\mathbb{Z}\atop m\neq 0}{\sum}}\;\Lambda_{1}^{|m|}\,\Lambda_{2}^{|n|}\,\frac{{\bf e}(nu_{2}-mu_{1})}{m}\,\\ &\,=\,2\pi{\rm i}\,B_{1\Lambda}(\{u_{1}\})\sum_{n}\Lambda^{|n|}_{2}{\bf e}(nu_{2})\,\\ &\,=\,2\pi{\rm i}\,\,B_{1\Lambda}(\{u_{1}\})\,\frac{1\,-\,\Lambda_{2}^{2}}{1\,+\,\Lambda_{2}^{2}\,-\,2\,\Lambda_{2}\cos(2\pi u_{2})}\,.\\ \end{split} (298)

Note that the right-hand side is a pure phase, and it is τ\tau-independent at fixed u1u_{1} and u2u_{2}. Note also that upon taking the limit Λ21\Lambda_{2}\to 1 (298) vanishes everywhere except for at the points z2z_{2}\in\mathbb{Z} where the result goes to infinity. Moreover, a computation shows that if one integrates the pre-factor in the right-hand side of (298) along the segment ϵ<v<ϵ-\epsilon<v<\epsilon, with ϵ\epsilon being an infinitesimal positive real number, and extracts the leading contribution to such result after expanding it around Λ=1\Lambda=1, the final result turns out to be independent of ϵ\epsilon and it matches the analogous integral of the right-hand side of (273). These facts together with periodicity under the shifts u2u2+1u_{2}\mapsto u_{2}+1, imply that in the limit Λ1\Lambda\to 1 the right-hand side of (298) approaches the Dirac-delta term in the right-hand side of (273).

References

  • (1) J. M. Maldacena, The Large N limit of superconformal field theories and supergravity, Int. J. Theor. Phys. 38 (1999) 1113–1133, [hep-th/9711200].
  • (2) A. Strominger and C. Vafa, Microscopic origin of the Bekenstein-Hawking entropy, Phys. Lett. B379 (1996) 99–104, [hep-th/9601029].
  • (3) F. Benini, K. Hristov, and A. Zaffaroni, Black hole microstates in AdS4 from supersymmetric localization, JHEP 05 (2016) 054, [arXiv:1511.04085].
  • (4) A. Cabo-Bizet, D. Cassani, D. Martelli, and S. Murthy, Microscopic origin of the Bekenstein-Hawking entropy of supersymmetric AdS5 black holes, JHEP 10 (2019) 062, [arXiv:1810.11442].
  • (5) S. Choi, J. Kim, S. Kim, and J. Nahmgoong, Large AdS black holes from QFT, arXiv:1810.12067.
  • (6) F. Benini and P. Milan, Black Holes in 4D 𝒩\mathcal{N}=4 Super-Yang-Mills Field Theory, Phys. Rev. X 10 (2020), no. 2 021037, [arXiv:1812.09613].
  • (7) A. Cabo-Bizet and S. Murthy, Supersymmetric phases of 4d 𝒩\mathcal{N} = 4 SYM at large NN, JHEP 09 (2020) 184, [arXiv:1909.09597].
  • (8) A. Cabo-Bizet, D. Cassani, D. Martelli, and S. Murthy, The large-NN limit of the 4d 𝒩\mathcal{N} = 1 superconformal index, JHEP 11 (2020) 150, [arXiv:2005.10654].
  • (9) F. Benini and P. Milan, A Bethe Ansatz type formula for the superconformal index, Commun. Math. Phys. 376 (2020), no. 2 1413–1440, [arXiv:1811.04107].
  • (10) J. B. Gutowski and H. S. Reall, General supersymmetric AdS(5) black holes, JHEP 04 (2004) 048, [hep-th/0401129].
  • (11) M. Cvetic, H. Lu, and C. Pope, Charged Kerr-de Sitter black holes in five dimensions, Phys. Lett. B 598 (2004) 273–278, [hep-th/0406196].
  • (12) M. Cvetic, H. Lu, and C. Pope, Charged rotating black holes in five dimensional U(1)**3 gauged N=2 supergravity, Phys. Rev. D 70 (2004) 081502, [hep-th/0407058].
  • (13) H. K. Kunduri and J. Lucietti, Notes on non-extremal, charged, rotating black holes in minimal D=5 gauged supergravity, Nucl. Phys. B724 (2005) 343–356, [hep-th/0504158].
  • (14) H. K. Kunduri, J. Lucietti, and H. S. Reall, Supersymmetric multi-charge AdS5 black holes, JHEP 04 (2006) 036, [hep-th/0601156].
  • (15) P. J. Silva, Thermodynamics at the BPS bound for Black Holes in AdS, JHEP 10 (2006) 022, [hep-th/0607056].
  • (16) S. M. Hosseini, K. Hristov, and A. Zaffaroni, An extremization principle for the entropy of rotating BPS black holes in AdS5, JHEP 07 (2017) 106, [arXiv:1705.05383].
  • (17) C. Closset, H. Kim, and B. Willett, 𝒩\mathcal{N} = 1 supersymmetric indices and the four-dimensional A-model, JHEP 08 (2017) 090, [arXiv:1707.05774].
  • (18) K. Goldstein, V. Jejjala, Y. Lei, S. van Leuven, and W. Li, Residues, modularity, and the Cardy limit of the 4d 𝒩=4\mathcal{N}=4 superconformal index, arXiv:2011.06605.
  • (19) D. Cassani and L. Papini, The BPS limit of rotating AdS black hole thermodynamics, JHEP 09 (2019) 079, [arXiv:1906.10148].
  • (20) G. Kántor, C. Papageorgakis, and P. Richmond, AdS7 Black-Hole Entropy and 5D 𝒩=2\mathcal{N}=2 Yang-Mills, arXiv:1907.02923.
  • (21) M. Suh, On-shell action and the Bekenstein-Hawking entropy of supersymmetric black holes in AdS6AdS_{6}, arXiv:1812.10491.
  • (22) N. Bobev and P. M. Crichigno, Universal spinning black holes and theories of class \mathcal{R}, JHEP 12 (2019) 054, [arXiv:1909.05873].
  • (23) A. González Lezcano and L. A. Pando Zayas, Microstate counting via Bethe Ansätze in the 4d 𝒩\mathcal{N} = 1 superconformal index, JHEP 03 (2020) 088, [arXiv:1907.12841].
  • (24) F. Benini, E. Colombo, S. Soltani, A. Zaffaroni, and Z. Zhang, Superconformal indices at large NN and the entropy of AdS5 ×\times SE5 black holes, Class. Quant. Grav. 37 (2020), no. 21 215021, [arXiv:2005.12308].
  • (25) A. Lanir, A. Nedelin, and O. Sela, Black hole entropy function for toric theories via Bethe Ansatz, JHEP 04 (2020) 091, [arXiv:1908.01737].
  • (26) A. Arabi Ardehali, J. Hong, and J. T. Liu, Asymptotic growth of the 4d 𝒩\mathcal{N} = 4 index and partially deconfined phases, JHEP 07 (2020) 073, [arXiv:1912.04169].
  • (27) S. Murthy, The growth of the 116\frac{1}{16}-BPS index in 4d 𝒩=4\mathcal{N}=4 SYM, arXiv:2005.10843.
  • (28) P. Agarwal, S. Choi, J. Kim, S. Kim, and J. Nahmgoong, AdS black holes and finite N indices, arXiv:2005.11240.
  • (29) S. Choi, J. Kim, S. Kim, and J. Nahmgoong, Comments on deconfinement in AdS/CFT, arXiv:1811.08646.
  • (30) M. Honda, Quantum Black Hole Entropy from 4d Supersymmetric Cardy formula, Phys. Rev. D100 (2019), no. 2 026008, [arXiv:1901.08091].
  • (31) A. Arabi Ardehali, Cardy-like asymptotics of the 4d 𝒩=4\mathcal{N}=4 index and AdS5 blackholes, JHEP 06 (2019) 134, [arXiv:1902.06619].
  • (32) J. Kim, S. Kim, and J. Song, A 4d N=1N=1 Cardy Formula, arXiv:1904.03455.
  • (33) A. Cabo-Bizet, D. Cassani, D. Martelli, and S. Murthy, The asymptotic growth of states of the 4d 𝒩=1\mathcal{N}=1 superconformal index, JHEP 08 (2019) 120, [arXiv:1904.05865].
  • (34) A. Amariti, I. Garozzo, and G. Lo Monaco, Entropy function from toric geometry, arXiv:1904.10009.
  • (35) P. M. Crichigno and D. Jain, The 5d Superconformal Index at Large NN and Black Holes, JHEP 09 (2020) 124, [arXiv:2005.00550].
  • (36) K. Goldstein, V. Jejjala, Y. Lei, S. van Leuven, and W. Li, Probing the EVH limit of supersymmetric AdS black holes, JHEP 02 (2020) 154, [arXiv:1910.14293].
  • (37) C. Copetti, A. Grassi, Z. Komargodski, and L. Tizzano, Delayed Deconfinement and the Hawking-Page Transition, arXiv:2008.04950.
  • (38) P. Agarwal, K.-H. Lee, and J. Song, Classification of large N superconformal gauge theories with a dense spectrum, arXiv:2007.16165.
  • (39) P. Agarwal and J. Song, Large N Gauge Theories with a Dense Spectrum and the Weak Gravity Conjecture, arXiv:1912.12881.
  • (40) F. Larsen, J. Nian, and Y. Zeng, AdS5 black hole entropy near the BPS limit, JHEP 06 (2020) 001, [arXiv:1907.02505].
  • (41) J. Nian and L. A. Pando Zayas, Toward an effective CFT2 from 𝒩\mathcal{N} = 4 super Yang-Mills and aspects of Hawking radiation, JHEP 07 (2020) 120, [arXiv:2003.02770].
  • (42) M. David, J. Nian, and L. A. Pando Zayas, Gravitational Cardy Limit and AdS Black Hole Entropy, JHEP 11 (2020) 041, [arXiv:2005.10251].
  • (43) J. a. F. Melo and J. E. Santos, Stringy corrections to the entropy of electrically charged supersymmetric black holes with AdS5×S5\mathrm{AdS}_{5}\times S^{5} asymptotics, arXiv:2007.06582.
  • (44) M. David and J. Nian, Universal Entropy and Hawking Radiation of Near-Extremal AdS4 Black Holes, arXiv:2009.12370.
  • (45) F. Larsen and S. Paranjape, Thermodynamics of Near BPS Black Holes in AdS4 and AdS7, arXiv:2010.04359.
  • (46) A. Zaffaroni, Lectures on AdS Black Holes, Holography and Localization, 2019. arXiv:1902.07176.
  • (47) J. Duistermaat and G. Heckman, On the Variation in the cohomology of the symplectic form of the reduced phase space, Invent. Math. 69 (1982) 259–268.
  • (48) J. J. Duistermaat and G. J. Heckman, Addendum to : On the variation in the cohomology of the symplectic form of the reduced phase space., Inventiones Mathematicae 72 (Feb., 1983) 153–158.
  • (49) E. Witten, Supersymmetry and Morse theory, J. Diff. Geom. 17 (1982), no. 4 661–692.
  • (50) N. Berline and M. Vergne, Zeros d’un champ de vecteurs et classes caracteristiques equivariantes, Duke Math. J. 50 (06, 1983) 539–549.
  • (51) M. Atiyah and R. Bott, The Moment map and equivariant cohomology, Topology 23 (1984) 1–28.
  • (52) M. Audin, Torus Actions on Symplectic Manifolds. Progress in Mathematics. Birkhäuser Basel, 2012.
  • (53) S. Cordes, G. W. Moore, and S. Ramgoolam, Lectures on 2-d Yang-Mills theory, equivariant cohomology and topological field theories, Nucl. Phys. B Proc. Suppl. 41 (1995) 184–244, [hep-th/9411210].
  • (54) A. J. Niemi and K. Palo, Equivariant Morse theory and quantum integrability, hep-th/9406068.
  • (55) M. Blau and G. Thompson, Localization and diagonalization: A review of functional integral techniques for low dimensional gauge theories and topological field theories, J. Math. Phys. 36 (1995) 2192–2236, [hep-th/9501075].
  • (56) V. Pestun, Review of localization in geometry, J. Phys. A 50 (2017), no. 44 443002, [arXiv:1608.02954].
  • (57) L. Jeffrey. http://www.math.toronto.edu/~jeffrey/mat1312/lec10.eqcoh.pdf.
  • (58) A. Alekseev, Notes on equivariant localization, Lect. Notes Phys. 543 (2000) 1–24.
  • (59) S. Cremonesi, An Introduction to Localisation and Supersymmetry in Curved Space, PoS Modave2013 (2013) 002.
  • (60) C. Romelsberger, Counting chiral primaries in N = 1, d=4 superconformal field theories, Nucl. Phys. B747 (2006) 329–353, [hep-th/0510060].
  • (61) J. Kinney, J. M. Maldacena, S. Minwalla, and S. Raju, An Index for 4 dimensional super conformal theories, Commun. Math. Phys. 275 (2007) 209–254, [hep-th/0510251].
  • (62) J. Hong and J. T. Liu, The topologically twisted index of 𝒩\mathcal{N} = 4 super-Yang-Mills on T×2S2{}^{2}\times S^{2} and the elliptic genus, JHEP 07 (2018) 018, [arXiv:1804.04592].
  • (63) R. A. Janik and M. Trzetrzelewski, Supergravitons from one loop perturbative N=4 SYM, Phys. Rev. D77 (2008) 085024, [arXiv:0712.2714].
  • (64) R. A. Janik and M. Trzetrzelewski, Supergravitons from one loop perturbative N=4 super Yang-Mills theory, Physical Review D - Particles, Fields, Gravitation and Cosmology 77 (2008), no. 8 1–22, [arXiv:0712.2714].
  • (65) M. Berkooz, D. Reichmann, and J. Simon, A Fermi Surface Model for Large Supersymmetric AdS(5) Black Holes, JHEP 01 (2007) 048, [hep-th/0604023].
  • (66) C.-M. Chang and X. Yin, 1/16 BPS states in 𝒩=\mathcal{N}= 4 super-Yang-Mills theory, Phys. Rev. D88 (2013), no. 10 106005, [arXiv:1305.6314].
  • (67) F. Dolan and H. Osborn, Applications of the Superconformal Index for Protected Operators and q-Hypergeometric Identities to N=1 Dual Theories, Nucl. Phys. B 818 (2009) 137–178, [arXiv:0801.4947].
  • (68) V. P. Spiridonov, Elliptic beta integrals and solvable models of statistical mechanics, Contemp. Math. 563 (2012) 181–211, [arXiv:1011.3798].
  • (69) L. A. Baxter and R. G. Bartle, The Elements of Integration and Lebesgue Measure., The Statistician 45 (1996), no. 4 528.
  • (70) F. C. S. Brown and A. Levin, Multiple Elliptic Polylogarithms, arXiv e-prints (Oct, 2011) arXiv:1110.6917, [arXiv:1110.6917].
  • (71) A. González Lezcano, J. Hong, J. T. Liu, and L. A. Pando Zayas, Sub-leading Structures in Superconformal Indices: Subdominant Saddles and Logarithmic Contributions, arXiv:2007.12604.
  • (72) N. A. Nekrasov and S. L. Shatashvili, Supersymmetric vacua and Bethe ansatz, Nucl. Phys. B Proc. Suppl. 192-193 (2009) 91–112, [arXiv:0901.4744].
  • (73) N. A. Nekrasov and S. L. Shatashvili, Quantum integrability and supersymmetric vacua, Prog. Theor. Phys. Suppl. 177 (2009) 105–119, [arXiv:0901.4748].
  • (74) J. B. Gutowski and H. S. Reall, Supersymmetric AdS5 black holes, JHEP 02 (2004) 006, [hep-th/0401042].
  • (75) E. Brezin, C. Itzykson, G. Parisi, and J. Zuber, Planar Diagrams, Commun. Math. Phys. 59 (1978) 35.
  • (76) M. Mariño, Instantons and Large N: An Introduction to Non-Perturbative Methods in Quantum Field Theory. Cambridge University Press, 9, 2015.
  • (77) D. Anninos and B. Mühlmann, Notes on matrix models (matrix musings), J. Stat. Mech. 2008 (2020) 083109, [arXiv:2004.01171].
  • (78) F. Dolan, Counting BPS operators in N=4 SYM, Nucl. Phys. B 790 (2008) 432–464, [arXiv:0704.1038].
  • (79) E. Witten, Analytic Continuation Of Chern-Simons Theory, AMS/IP Stud. Adv. Math. 50 (2011) 347–446, [arXiv:1001.2933].
  • (80) G. Felder and A. Varchenko Advances in Mathematics 156 (2000), no. 1 44–76, [9907061].
  • (81) B. Assel, D. Cassani, and D. Martelli, Supersymmetric counterterms from new minimal supergravity, JHEP 11 (2014) 135, [arXiv:1410.6487].
  • (82) P. Benetti Genolini, D. Cassani, D. Martelli, and J. Sparks, Holographic renormalization and supersymmetry, JHEP 02 (2017) 132, [arXiv:1612.06761].
  • (83) I. Papadimitriou, Supercurrent anomalies in 4d SCFTs, JHEP 07 (2017) 038, [arXiv:1703.04299].
  • (84) O. S. An, Anomaly-corrected supersymmetry algebra and supersymmetric holographic renormalization, JHEP 12 (2017) 107, [arXiv:1703.09607].
  • (85) C. Closset and I. Shamir, The 𝒩=1\mathcal{N}=1 Chiral Multiplet on T2×S2T^{2}\times S^{2} and Supersymmetric Localization, JHEP 03 (2014) 040, [arXiv:1311.2430].
  • (86) B. Assel, D. Cassani, and D. Martelli, Localization on Hopf surfaces, JHEP 08 (2014) 123, [arXiv:1405.5144].
  • (87) B. Assel, D. Cassani, L. Di Pietro, Z. Komargodski, J. Lorenzen, and D. Martelli, The Casimir Energy in Curved Space and its Supersymmetric Counterpart, JHEP 07 (2015) 043, [arXiv:1503.05537].
  • (88) W. Chen, H. Lu, and C. N. Pope, Mass of rotating black holes in gauged supergravities, Phys. Rev. D73 (2006) 104036, [hep-th/0510081].
  • (89) S. M. Kuzenko, A. Schwimmer, and S. Theisen, Comments on Anomalies in Supersymmetric Theories, J. Phys. A 53 (2020), no. 6 064003, [arXiv:1909.07084].
  • (90) A. Bzowski, G. Festuccia, and V. Procházka, Consistency of supersymmetric ’t Hooft anomalies, arXiv:2011.09978.
  • (91) C. Closset, L. Di Pietro, and H. Kim, ’t Hooft anomalies and the holomorphy of supersymmetric partition functions, JHEP 08 (2019) 035, [arXiv:1905.05722].
  • (92) D. Zagier, The Bloch-Wigner-Ramakrishnan polylogarithm function, Math. Annalen 286 (1990) 613–624.
  • (93) V. Pa\textcommabelowsol and W. Zudilin, A study of elliptic gamma function and allies, arXiv e-prints (Dec., 2017) arXiv:1801.00210, [arXiv:1801.00210].
  • (94) A. Levin, Elliptic polylogarithms: An analytic theory, Compositio Mathematica 12 (1997) 106: 267.