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

The decomposition formula for Verlinde Sums

Yiannis Loizides  and  Eckhard Meinrenken
Abstract.

We prove a decomposition formula for Verlinde sums (rational trigonometric sums), as a discrete counterpart to the Boysal-Vergne decomposition formula for Bernoulli series. Motivated by applications to fixed point formulas in Hamiltonian geometry, we develop differential form valued version of Bernoulli series and Verlinde sums, and extend the decomposition formula to this wider context.

1. Introduction

The prototypical example of a Verlinde sum is the function, for a given natural number nn,

(1) Vn(λ,)=q=1,q1qλ(1q1)n,(λ,)×.V_{n}(\lambda,\ell)=\sum_{q^{\ell}=1,\ q\neq 1}\frac{q^{\lambda}}{(1-q^{-1})^{n}},\ \ \ (\lambda,\ell)\in{\mathbb{Z}}\times{\mathbb{N}}.

It may be seen as a discrete analogue of the Bernoulli series

(2) Bn(λ)=j0e2πijλ(2πij)n,λ.B_{n}(\lambda)=\sum_{j\in{\mathbb{Z}}_{\neq 0}}\frac{e^{2\pi{\mathrm{i}}j\lambda}}{(2\pi{\mathrm{i}}j)^{n}},\ \ \ \ \lambda\in{\mathbb{R}}.

The Bernoulli series is invariant under integer translation, and its restriction to the open unit interval (0,1)(0,1) is given by a polynomial Bern\operatorname{Ber}_{n} of degree nn in λ\lambda. For n1n\geq 1, these are 1n!-\frac{1}{n!} times the Bernoulli polynomials. Similarly, the Verlinde sum for fixed \ell is \ell{\mathbb{Z}}-periodic in λ\lambda, and its restriction to the set of all (λ,)(\lambda,\ell) such that λ/(0,1)\lambda/\ell\in(0,1) is a polynomial Vern\operatorname{Ver}_{n} of degree nn in the two variables λ,\lambda,\ell. Its part of homogeneity nn in (λ,)(\lambda,\ell) is given by nBern(λ/)\ell^{n}\operatorname{Ber}_{n}(\lambda/\ell). These polynomials may be described through generating functions:

(3) n=0(1ez)nVern(λ,)=1eλzez1ez1\sum_{n=0}^{\infty}(1-e^{-z})^{n}\operatorname{Ver}_{n}(\lambda,\ell)=1-\ell e^{\lambda z}\frac{e^{z}-1}{e^{\ell z}-1}

and

(4) n=0znBern(λ)=1eλzzez1.\sum_{n=0}^{\infty}z^{n}\operatorname{Ber}_{n}(\lambda)=1-e^{\lambda z}\frac{z}{e^{z}-1}.

The Bernoulli series make an appearance in the volume formula for moduli spaces of flat connections of flat SU(2)\operatorname{SU}(2)-bundles over surfaces, due to Thaddeus [20] and Witten [26]. Similarly, the Verlinde sums appear in the Verlinde formulas [25] for the quantization of these moduli spaces.

The versions for higher rank Lie groups lead to higher-dimensional multiple Bernoulli series and Verlinde sums. The multiple Bernoulli series exhibit a piecewise polynomial behavior, while the multiple Verlinde sums are piecewise quasi-polynomial. Aside from the context of 2-dimensional gauge theory, they also appear in the localization formulas for volumes and quantizations of Hamiltonian loop group spaces or, equivalently, quasi-Hamiltonian GG-spaces [1, 2, 11].

The combinatorics of the (multiple) Verlinde sums, also known as rational trigonometric sums, was developed by Szenes [17]. The main result in [17] is a residue formula for the Verlinde sums, similar to his earlier result [16] for Bernoulli series. The Szenes formula allows for efficient computations, and in particular leads to a proof of the (quasi-)polynomial behavior – for example, (3) and (4) may be obtained from his theorem.

The main purpose of this article is to prove a decomposition formula for multiple Verlinde sums, similar to the Boysal-Vergne decomposition formula [5] for Bernoulli series. For the Verlinde sums VnV_{n}, the formula is

Vn(λ,)=Vern(λ,)+μ=1Pn(λμ)+(1)nμ=0Pn(μλn),V_{n}(\lambda,\ell)={\textnormal{Ver}}_{n}(\lambda,\ell)+\ell\sum_{\mu=1}^{\infty}P_{n}(\lambda-\ell\mu)+(-1)^{n}\ell\sum_{\mu=-\infty}^{0}P_{n}(\ell\mu-\lambda-n),

where Pn(μ)P_{n}(\mu) is the partition function, i.e., the number of ways of writing μ\mu as a sum of nn non-negative integers. It expresses VnV_{n} as a central polynomial contribution, plus correction terms supported on affine half lines; the support properties are such that for any given (λ,)(\lambda,\ell) the sum is finite.

We will need the decomposition formula for Verlinde sums for our combinatorial proof of the ‘quantization commutes with reduction’ theorem for Hamiltonian loop group spaces, in the spirit of Szenes-Vergne’s argument [18] for ordinary Hamiltonian spaces. In turn, this approach is motivated by Paradan’s norm-square formulas for Hamiltonian spaces [12, 13]. With these applications in mind, we are led to consider more general Verlinde sums with ‘equivariant parameters’; under the Chern-Weil homomorphism this produces the required formulas for expressions arising in the fixed point formula.


Acknowledgments: We thank Michele Vergne for discussions and helpful comments. Most of the results described in this paper are natural extensions of work by Arzu Boysal and Michele Vergne on Bernoulli series, and a number of steps in this direction are already explained in Vergne’s lecture notes [23]; as well as the slides [21] from her lecture at the 2010 AMS meeting in San Francisco.

2. Verlinde sums

2.1. Lattices.

Let Λ\Lambda be a lattice, with dual lattice Λ=Hom(Λ,)\Lambda^{\ast}=\operatorname{Hom}(\Lambda,{\mathbb{Z}}). We denote by 𝔱=Λ{\mathfrak{t}}=\Lambda\otimes_{{\mathbb{Z}}}{\mathbb{R}} the real vector space spanned by Λ\Lambda and by 𝔱=Λ{\mathfrak{t}}_{{\mathbb{Q}}}=\Lambda\otimes_{{\mathbb{Z}}}{\mathbb{Q}} its rational points. A subspace 𝔥𝔱{\mathfrak{h}}\subseteq{\mathfrak{t}} is rational if it is spanned by 𝔥𝔱{\mathfrak{h}}\cap{\mathfrak{t}}_{{\mathbb{Q}}}; in this case, ann(𝔥)𝔱\operatorname{ann}({\mathfrak{h}})\subseteq{\mathfrak{t}}^{\ast} is rational and (𝔥Λ)Λ/Λann(𝔥)({\mathfrak{h}}\cap\Lambda)^{\ast}\cong\Lambda^{\ast}/\Lambda^{\ast}\cap\operatorname{ann}({\mathfrak{h}}).

Let 𝖳=𝔱/Λ{\mathsf{T}}={\mathfrak{t}}/\Lambda the torus; the quotient map is the exponential map exp:𝔱𝖳\exp\colon{\mathfrak{t}}\to{\mathsf{T}} for this torus. Thus Λ=ker(exp)\Lambda=\operatorname{ker}(\exp) becomes the integral lattice of 𝖳{\mathsf{T}}, and Λ\Lambda^{\ast} is identified with the weight lattice Hom(T,U(1))\mathrm{Hom}(T,\mathrm{U}(1)). For λΛ\lambda\in\Lambda^{\ast}, we denote by 𝖳U(1),ttλ{\mathsf{T}}\to\mathrm{U}(1),\ \ t\mapsto t^{\lambda} the corresponding homomorphism; thus

(5) tλ=e2πiλ,Xt^{\lambda}=e^{2\pi{\mathrm{i}}\langle\lambda,X\rangle}

for t=exp(X)t=\exp(X). Suppose Ξ𝔱\Xi\subseteq{\mathfrak{t}} is a lattice containing Λ\Lambda. For \ell\in{\mathbb{N}} we consider the finite subgroup

𝖳=1Ξ/Λ𝖳.{\mathsf{T}}_{\ell}=\ell^{-1}\Xi/\Lambda\subseteq{\mathsf{T}}.

For example, if BB is an inner product on 𝔱{\mathfrak{t}} which is integral in the sense that it restricts to a {\mathbb{Z}}-valued bilinear form on Λ\Lambda, then the inverse image of Λ\Lambda^{*} under the isomorphism B:𝔱𝔱B^{\flat}\colon{\mathfrak{t}}\to{\mathfrak{t}}^{\ast} can play the role of such a lattice Ξ\Xi.

Example 2.1.

The following setting plays a role for the Verlinde formula: Let GG be a compact, simple, simply connected Lie group, 𝖳G{\mathsf{T}}\subseteq G a maximal torus, and Λ𝔱\Lambda\subseteq{\mathfrak{t}} the integral lattice, so that 𝖳=𝔱/Λ{\mathsf{T}}={\mathfrak{t}}/\Lambda. The basic inner product BB on 𝔤{\mathfrak{g}} is the unique invariant inner product such that the shortest vectors in Λ{0}\Lambda-\{0\} have length 2\sqrt{2}. This is an integral inner product, and we take Ξ=Λ\Xi=\Lambda^{*} under the resulting identification 𝔱𝔱{\mathfrak{t}}\cong{\mathfrak{t}}^{\ast}. Given a level kk\in{\mathbb{N}}, the level kk fusion ring (or Verlinde algebra) Rk(G)R_{k}(G) can be abstractly defined as quotient ring R(G)/k(G)R(G)/\mathcal{I}_{k}(G), where R(G)R(G) is regarded as the ring of characters of GG-representations, and k(G)\mathcal{I}_{k}(G) is the ideal of characters vanishing on the regular elements of

𝖳k+h𝖳.{\mathsf{T}}_{k+{\textnormal{h}^{\vee}}}\subseteq{\mathsf{T}}.

Here h{\textnormal{h}^{\vee}} is the dual Coxeter number of GG, and an element of GG is called regular if its centralizer is a maximal torus.

Recall that a list AA of elements of a set SS is a collection of elements of SS ‘with multiplicities’. Equivalently, it amounts to a map S{0}S\to{\mathbb{N}}\cup\{0\}. For finite lists, it is often convenient to use set-theoretic notation A={s1,,sn}A=\{s_{1},\ldots,s_{n}\}, where elements appear several times, according to their multiplicity. The notation sAs\in A means that ss appears with multiplicity at least 11, and A{s}A-\{s\} is the new list in which the multiplicity of ss has been reduced by 11. The Verlinde sums in the following section will involve a choice of list 𝜶={α1,,αn}{\bm{\alpha}}=\{\alpha_{1},\ldots,\alpha_{n}\} of weights αiΛ\alpha_{i}\in\Lambda^{*}. Note that finite lists of weights in Λ\Lambda^{*} are in 1-1 correspondence with isomorphism classes of finite-dimensional unitary 𝖳{\mathsf{T}}-representations.

Definition 2.2.

For a list 𝜶={α1,,αn}{\bm{\alpha}}=\{\alpha_{1},\ldots,\alpha_{n}\} of weights, we define:

  • \cdot

    The zonotope

    𝜶={k=1ntkαk𝔱| 0tk1}.\Box{\bm{\alpha}}=\Big{\{}\sum_{k=1}^{n}t_{k}\alpha_{k}\in{\mathfrak{t}}^{\ast}|\ 0\leq t_{k}\leq 1\Big{\}}.
  • \cdot

    The collection 𝒮=𝒮(𝜶){\mathcal{S}}={\mathcal{S}}({\bm{\alpha}}) of affine subspaces of 𝔱{\mathfrak{t}}^{\ast}, consisting of subspaces spanned by sublists of 𝜶{\bm{\alpha}}, together with Ξ\Xi^{\ast}-translates of such subspaces. The elements Δ𝒮\Delta\in{\mathcal{S}} are referred to as admissible subspaces [5].

  • \cdot

    For Δ𝒮(𝜶)\Delta\in{\mathcal{S}}({\bm{\alpha}}), we denote by 𝔱Δ𝔱{\mathfrak{t}}_{\Delta}\subseteq{\mathfrak{t}} the rational subspace of vectors orthogonal to Δ\Delta, and by 𝖳Δ=exp(𝔱Δ){\mathsf{T}}_{\Delta}=\exp({\mathfrak{t}}_{\Delta}) the corresponding torus. If Δ\Delta is a Ξ\Xi^{\ast}-translate of the span of some α𝜶\alpha\in{\bm{\alpha}}, these coincide with 𝔱α=ker(α),𝖳α=exp(𝔱α){\mathfrak{t}}_{\alpha}=\operatorname{ker}(\alpha),\ {\mathsf{T}}_{\alpha}=\exp({\mathfrak{t}}_{\alpha}).

  • \cdot

    Given Δ𝒮(𝜶)\Delta\in{\mathcal{S}}({\bm{\alpha}}), an element μΔ\mu\in\Delta is called regular in Δ\Delta if it is not contained in any Δ𝒮(𝜶)\Delta^{\prime}\in{\mathcal{S}}({\bm{\alpha}}) with dimΔ<dimΔ\dim\Delta^{\prime}<\dim\Delta. The connected components of regular elements in Δ\Delta are called the (open) chambers in Δ\Delta. If Δ0:=span𝜶\Delta_{0}:={\textnormal{span}}_{{\mathbb{R}}}\,{\bm{\alpha}} is all of 𝔱{\mathfrak{t}}^{\ast}, the chambers of Δ0𝒮(𝜶)\Delta_{0}\in{\mathcal{S}}({\bm{\alpha}}) will simply be referred to as the chambers.

Example 2.3.

In Example 2.1, consider the list 𝜶={\bm{\alpha}}=\mathfrak{R}_{-} of negative roots of GG. The collection 𝒮{\mathcal{S}} of admissible subspaces consists of subspaces spanned by subsets of roots, together with their ΞΛ\Xi^{\ast}\cong\Lambda-translates. Letting ρ\rho be the half-sum of positive roots, the zonotope is given by

=hull(Wρ)ρ,\Box\mathfrak{R}_{-}=\mathrm{hull}(W\cdot\rho)-\rho,

the ρ-\rho shift of the convex hull of the Weyl group orbit of ρ\rho. This follows from the formula for the character of the ρ\rho-representation,

χρ(t)=tρα(1+tα)\chi_{\rho}(t)=t^{\rho}\prod_{\alpha\in\mathfrak{R}_{-}}(1+t^{\alpha})

by comparing the set of weights appearing on both sides.

2.2. Definition of the Verlinde sum.

Suppose ΛΞ\Lambda\subseteq\Xi are full-rank lattices in a finite-dimensional vector space 𝔱{\mathfrak{t}} of dimension rr, and recall 𝖳=(1Ξ)/Λ{\mathsf{T}}_{\ell}=(\frac{1}{\ell}\Xi)/\Lambda. Consider a list of weights 𝜶={α1,,αn}{\bm{\alpha}}=\{\alpha_{1},\ldots,\alpha_{n}\}, and a list 𝒖={u1,,un}{\bm{u}}=\{u_{1},\ldots,u_{n}\} of complex numbers which are roots of unity (that is, some positive power of uku_{k} is 11). We will call the pairs α¯k=(αk,uk)\underline{\alpha}_{k}=(\alpha_{k},u_{k}) augmented weights, and denote the corresponding list of augmented pairs by 𝜶¯=(𝜶,𝒖){\underline{{\bm{\alpha}}}}=({\bm{\alpha}},{\bm{u}}).

Definition 2.4.

The Verlinde sum associated to the list of augmented weights 𝜶¯{\underline{{\bm{\alpha}}}} is the function V𝜶¯:Λ×V_{{\underline{{\bm{\alpha}}}}}\colon\Lambda^{\ast}\times{\mathbb{N}}\rightarrow{\mathbb{C}} defined by

(6) V𝜶¯(λ,)=t𝖳tλ(α,u)𝜶¯ 1utα.V_{{\underline{{\bm{\alpha}}}}}(\lambda,\ell)=\sideset{}{{}^{\prime}}{\sum}_{t\in{\mathsf{T}}_{\ell}}\frac{t^{\lambda}}{\prod_{(\alpha,u)\in{\underline{{\bm{\alpha}}}}}\,1-u\,t^{-\alpha}}.

Here the prime next to the summation symbol means that the summation only extends over elements t𝖳t\in{\mathsf{T}}_{\ell} for which the denominator does not vanish. If all uku_{k} are equal to 11, we also use the notation V𝜶V_{{\bm{\alpha}}}.

The functions V𝜶¯V_{{\underline{{\bm{\alpha}}}}} are called rational trigonometric sums in [17]; the term Verlinde sum was used in [21]. Since tλt^{\lambda} for t𝖳t\in{\mathsf{T}}_{\ell} depends only on the equivalence class of λ\lambda modulo Ξ\ell\Xi^{\ast}, the Verlinde sum has the periodicity property

V𝜶¯(λ+μ,)=V𝜶¯(λ,),μΞ.V_{{\underline{{\bm{\alpha}}}}}(\lambda+\ell\mu,\ell)=V_{{\underline{{\bm{\alpha}}}}}(\lambda,\ell),\ \ \ \mu\in\Xi^{\ast}.

Two special cases are worth pointing out:

  1. (i)

    If the list 𝜶¯{\underline{{\bm{\alpha}}}} includes the trivial augmented weight α¯=(0,1)\underline{\alpha}=(0,1), then the summation is over an empty set, hence V𝜶¯=0V_{{\underline{{\bm{\alpha}}}}}=0.

  2. (ii)

    If 𝜶¯={\underline{{\bm{\alpha}}}}=\emptyset, then the denominator in Equation (6) is equal to 11, and the summation is over all of 𝖳{\mathsf{T}}_{\ell}. We hence obtain, by finite Fourier transform,

    (7) V(λ,)=(#𝖳)δΞ(λ)V_{\emptyset}(\lambda,\ell)=(\#{\mathsf{T}}_{\ell})\ \delta_{\ell\Xi^{\ast}}(\lambda)

    where δΞ(λ)\delta_{\ell\Xi^{\ast}}(\lambda) equals 11 if λΞ\lambda\in\ell\Xi^{\ast}, 0 otherwise.

The Verlinde sum (6) is the discrete counterpart to the multiple Bernoulli series

(8) B𝜶(λ)=ξΞe2πiλ,ξα𝜶2πiα,ξ,λ𝔱B_{{\bm{\alpha}}}(\lambda)=\sideset{}{{}^{\prime}}{\sum}_{\xi\in\Xi}\frac{e^{2\pi{\mathrm{i}}\langle\lambda,\xi\rangle}}{\prod_{\alpha\in{\bm{\alpha}}}2\pi{\mathrm{i}}\langle\alpha,\xi\rangle},\ \ \ \ \lambda\in{\mathfrak{t}}^{\ast}

where the infinite sum is defined as a generalized function of λ\lambda. These series satisfy B𝜶(λ+μ)=B𝜶(λ)B_{{\bm{\alpha}}}(\lambda+\mu)=B_{{\bm{\alpha}}}(\lambda) for μΞ\mu\in\Xi^{\ast}, and turn out to be piecewise polynomial. The multiple Bernoulli series have been studied in [4, 5, 16, 21].

Example 2.5.

(Cf.  [21]) Let 𝔱={\mathfrak{t}}={\mathbb{R}}, with Λ=Ξ=\Lambda=\Xi={\mathbb{Z}}. Let 𝜶={1,,1}{\bm{\alpha}}=\{1,\ldots,1\} (where 11 denotes the weight 1=Λ1\in{\mathbb{Z}}=\Lambda^{\ast}) be the list consisting of the element 11 repeated nn times. For n>0n>0, the corresponding Bernoulli series Bn:=B𝜶B_{n}:=B_{{\bm{\alpha}}} and Verlinde sum Vn:=V𝜶V_{n}:=V_{{\bm{\alpha}}} are given by

Bn(λ)=j0e2πijλ(2πij)n,Vn(λ,)=j=1l1e2πijλ/(1e2πij/)n.B_{n}(\lambda)=\sum_{j\in{\mathbb{Z}}_{\neq 0}}\frac{e^{2\pi{\mathrm{i}}j\lambda}}{(2\pi{\mathrm{i}}j)^{n}},\ \ \ \ \ \ V_{n}(\lambda,\ell)=\sum_{j=1}^{l-1}\frac{e^{2\pi{\mathrm{i}}j{\lambda}/{\ell}}}{(1-e^{-2\pi{\mathrm{i}}j/\ell})^{n}}.

For BnB_{n}, this is a (generalized) function of λ\lambda\in{\mathbb{R}}; for VnV_{n}, we take λ\lambda to be an integer. If n=0n=0 (corresponding to 𝜶={\bm{\alpha}}=\emptyset) we have

B0(λ)=δ(λ),V0(λ,)=δ(λ).B_{0}(\lambda)=\delta_{\mathbb{Z}}(\lambda),\ \ \ V_{0}(\lambda,\ell)=\ell\delta_{\ell{\mathbb{Z}}}(\lambda).

In the first formula, δ\delta_{\mathbb{Z}} denotes the generalized function on 𝔱{\mathfrak{t}} whose integration against a test function ff gives jf(j)\sum_{j\in{\mathbb{Z}}}f(j); in the second formula, δ\delta_{\ell{\mathbb{Z}}} denotes the characteristic function of \ell{\mathbb{Z}}\subseteq{\mathbb{Z}}. The series BnB_{n} is periodic with period 11 in λ\lambda, while VnV_{n} is periodic with period \ell. The Bernoulli series satisfy a differential equation λBn(λ)=Bn1(λ)δn,1\frac{\partial}{\partial\lambda}B_{n}(\lambda)=B_{n-1}(\lambda)-\delta_{n,1} and a normalization 01Bn(λ)𝑑λ=0\int_{0}^{1}B_{n}(\lambda)d\lambda=0 for n1n\geq 1. These two properties can can be used to compute BnB_{n} recursively, see [21]. It turns out that Bn(λ)B_{n}(\lambda) are given by polynomials Bern(λ)\mathrm{Ber}_{n}(\lambda) on the interval 0<λ<10<\lambda<1. We have that Ber0=0\operatorname{Ber}_{0}=0, while for n1n\geq 1, Bern\operatorname{Ber}_{n} is 1n!\frac{-1}{n!} times the Bernoulli polynomials. For example,

Ber1(λ)\displaystyle\mathrm{Ber}_{1}(\lambda) =\displaystyle= 12λ,\displaystyle\frac{1}{2}-\lambda,
Ber2(λ)\displaystyle\mathrm{Ber}_{2}(\lambda) =\displaystyle= 12λ2+12λ112.\displaystyle-\frac{1}{2}\lambda^{2}+\frac{1}{2}\lambda-\frac{1}{12}.

Similarly, the Verlinde sums satisfy the difference equation

(9) Vn(λ,)Vn(λ1,)=Vn1(λ,)δn,1,n1V_{n}(\lambda,\ell)-V_{n}(\lambda-1,\ell)=V_{n-1}(\lambda,\ell)-\delta_{n,1},\ \ \ n\geq 1

and the normalization

(10) Vn(0,)++Vn(1,)=0.V_{n}(0,\ell)+\ldots+V_{n}(\ell-1,\ell)=0.

By Theorem 2.7 below, there are polynomials Vern\mathrm{Ver}_{n} in λ,\lambda,\ell such that Vern(λ,)=Vn(λ,){\textnormal{Ver}}_{n}(\lambda,\ell)=V_{n}(\lambda,\ell) for n<λ<-n<\lambda<\ell. These can be computed recursively, using (9) and (10). In low degrees,

Ver0(λ,)\displaystyle{\textnormal{Ver}}_{0}(\lambda,\ell) =0,\displaystyle=0,
Ver1(λ,)\displaystyle{\textnormal{Ver}}_{1}(\lambda,\ell) =λ+12,\displaystyle=-\lambda+\frac{\ell-1}{2},
Ver2(λ,)\displaystyle{\textnormal{Ver}}_{2}(\lambda,\ell) =12λ2+(21)λ212+2512.\displaystyle=-\frac{1}{2}\lambda^{2}+(\frac{\ell}{2}-1)\lambda-\frac{\ell^{2}}{12}+\frac{\ell}{2}-\frac{5}{12}.

The polynomials Bern(λ)\operatorname{Ber}_{n}(\lambda) are a ‘classical limit’ of the polynomials Vern(λ,)\operatorname{Ver}_{n}(\lambda,\ell), in the sense that

(11) limnVern(λ,)=Bern(λ).\lim_{\ell\to\infty}\ell^{-n}\operatorname{Ver}_{n}(\ell\lambda,\ell)=\operatorname{Ber}_{n}(\lambda).

See Section 5.4 for an explanation of this fact.

2.3. Basic properties of Verlinde sums.

Throughout this section, ΛΞ𝔱\Lambda\subseteq\Xi\subseteq{\mathfrak{t}} are given lattices of maximal rank, and 𝜶¯=(𝜶,𝒖){\underline{{\bm{\alpha}}}}=({\bm{\alpha}},{\bm{u}}) is a list of augmented weights. We assume that α0\alpha\neq 0 for all α𝜶\alpha\in{\bm{\alpha}}. We will describe some general properties of the resulting Verlinde sum.

  1. (a)

    Since the function V𝜶¯(,)V_{{\underline{{\bm{\alpha}}}}}(\cdot,\ell) on Λ\Lambda^{\ast} is Ξ\ell\Xi^{\ast}-periodic, it descends to a function on the finite group Λ/Ξ\Lambda^{\ast}/\ell\Xi^{\ast} (the dual group to 𝖳=l1Ξ/Λ{\mathsf{T}}_{\ell}=l^{-1}\Xi/\Lambda). By definition, this function is the finite Fourier transform of the function 𝖳{\mathsf{T}}_{\ell}\rightarrow{\mathbb{C}}, given by t(α,u)𝜶¯(1utα)1t\mapsto\prod_{(\alpha,u)\in{\underline{{\bm{\alpha}}}}}\,(1-u\,t^{-\alpha})^{-1} if all utα1u\,t^{-\alpha}\neq 1, and t0t\mapsto 0 otherwise. By inverse Fourier transform, the value of this function at t=et=e is the sum of V𝜶¯(,)V_{{\underline{{\bm{\alpha}}}}}(\cdot,\ell) over Λ/Ξ\Lambda^{\ast}/\ell\Xi^{\ast}. That is,

    [λ]Λ/ΞV𝜶¯(λ,)={u𝒖(1u)1 if 1𝒖,0 if 1𝒖,\sum_{[\lambda]\in\Lambda^{\ast}/\ell\Xi^{\ast}}V_{{\underline{{\bm{\alpha}}}}}(\lambda,\ell)=\begin{cases}\prod_{u\in{\bm{u}}}(1-u)^{-1}&\mbox{ if }1\not\in{\bm{u}},\\ 0&\mbox{ if }1\in{\bm{u}},\\ \end{cases}

    where the sum picks one representative λ\lambda from each equivalence class. This generalizes (10).

  2. (b)

    The Bernoulli series (2) is real-valued, since complex conjugation amounts to replacing ξ\xi with ξ-\xi in the sum. Similarly, the Verlinde sum satisfies

    V𝜶¯(λ,)=V𝜶¯(λ,),V_{{\underline{{\bm{\alpha}}}}}(\lambda,\ell)^{*}=V_{{\underline{{\bm{\alpha}}}}^{*}}(\lambda,\ell),

    where 𝜶¯{\underline{{\bm{\alpha}}}}^{*} is the list of all (α,u)(\alpha,u^{*}) with (α,u)𝜶¯(\alpha,u)\in{\underline{{\bm{\alpha}}}}.

  3. (c)

    One knows [5] that the Bernoulli series B𝜶B_{{\bm{\alpha}}} is supported on the union of (top-dimensional) admissible subspaces Δ𝒮(𝜶)\Delta\in{\mathcal{S}}({\bm{\alpha}}). That is, supp(B𝜶)Δ0+Ξ\mathrm{supp}(B_{{\bm{\alpha}}})\subseteq\Delta_{0}+\Xi^{\ast} where Δ0=span𝜶\Delta_{0}={\textnormal{span}}_{{\mathbb{R}}}{\bm{\alpha}}. We will see that similarly, supp(V𝜶¯)\mathrm{supp}(V_{{\underline{{\bm{\alpha}}}}}) is contained in the set of all λ,\lambda,\ell such that λ/Δ0+Ξ\lambda/\ell\in\Delta_{0}+\Xi^{\ast}. If Δ0\Delta_{0} is a proper subspace of 𝔱{\mathfrak{t}}^{\ast}, then the Verlinde sum for 𝜶¯=(𝜶,u){\underline{{\bm{\alpha}}}}=({\bm{\alpha}},u) is related to lower-dimensional Verlinde sums, as follows. Let 𝔱=𝔱/𝔱Δ0{\mathfrak{t}}^{\prime}={\mathfrak{t}}/{\mathfrak{t}}_{\Delta_{0}}, and denote by Λ,Ξ\Lambda^{\prime},\Xi^{\prime} be the images of Λ,Ξ\Lambda,\Xi under the quotient map, so that (Λ)=ΛΔ0(\Lambda^{\prime})^{*}=\Lambda^{*}\cap\Delta_{0}, and similarly for (Ξ)(\Xi^{\prime})^{*}. Denote by 𝜶{\bm{\alpha}}^{\prime} the list of weights 𝜶{\bm{\alpha}}, but regarded as weights for 𝖳=𝔱/Λ{\mathsf{T}}^{\prime}={\mathfrak{t}}^{\prime}/\Lambda^{\prime}, and put 𝜶¯=(𝜶,𝒖){\underline{{\bm{\alpha}}}}^{\prime}=({\bm{\alpha}}^{\prime},{\bm{u}}).

    Proposition 2.6.

    The support of the Verlinde sum satisfies

    supp(V𝜶¯){(λ,)Λ×|1λΔ0+Ξ}.\mathrm{supp}(V_{{\underline{{\bm{\alpha}}}}})\subseteq\left\{(\lambda,\ell)\in\Lambda^{\ast}\times{\mathbb{N}}\,\Big{|}\ \frac{1}{\ell}\lambda\in\Delta_{0}+\Xi^{\ast}\right\}.

    That is, V𝛂¯(λ,)V_{{\underline{{\bm{\alpha}}}}}(\lambda,\ell) is zero unless there exists λ(Λ)\lambda^{\prime}\in(\Lambda^{\prime})^{*} with λλΞ\lambda-\lambda^{\prime}\in\ell\Xi^{\ast}. In the latter case,

    V𝜶¯(λ,)=#(T𝖳Δ0)V𝜶¯(λ,).V_{{\underline{{\bm{\alpha}}}}}(\lambda,\ell)=\#(T_{\ell}\cap{\mathsf{T}}_{\Delta_{0}})\ V_{{\underline{{\bm{\alpha}}}}^{\prime}}(\lambda^{\prime},\ell).
    Proof.

    If t1𝖳t_{1}\in{\mathsf{T}}_{\ell} and t2𝖳𝖳Δ0t_{2}\in{\mathsf{T}}_{\ell}\cap{\mathsf{T}}_{\Delta_{0}}, then (t1t2)α=t1αt2α=t1α(t_{1}t_{2})^{-\alpha}=t_{1}^{-\alpha}\,t_{2}^{-\alpha}=t_{1}^{-\alpha} for all α𝜶\alpha\in{\bm{\alpha}}. We can therefore carry out the Verlinde sum in stages, by first summing over all elements in a fixed equivalence class of 𝖳{\mathsf{T}}_{\ell} modulo 𝖳𝖳Δ0{\mathsf{T}}_{\ell}\cap{\mathsf{T}}_{\Delta_{0}}, followed by a sum over 𝖳/(T𝖳Δ0){\mathsf{T}}_{\ell}/(T_{\ell}\cap{\mathsf{T}}_{\Delta_{0}}). For fixed t1𝖳t_{1}\in{\mathsf{T}}_{\ell} such that ut1α1ut_{1}^{-\alpha}\neq 1 for all (α,u)𝜶¯(\alpha,u)\in{\underline{{\bm{\alpha}}}}, we have that

    t2𝖳𝖳Δ0(t1t2)λ(α,u)𝜶¯ 1u(t1t2)α=t1λ(α,u)𝜶¯ 1ut1αt2𝖳𝖳Δ0t2λ.\sum_{t_{2}\in{\mathsf{T}}_{\ell}\cap{\mathsf{T}}_{\Delta_{0}}}\frac{(t_{1}t_{2})^{\lambda}}{\prod_{(\alpha,u)\in{\underline{{\bm{\alpha}}}}}\,1-u\,(t_{1}t_{2})^{-\alpha}}=\frac{t_{1}^{\lambda}}{\prod_{(\alpha,u)\in{\underline{{\bm{\alpha}}}}}\,1-u\,t_{1}^{-\alpha}}\sum_{t_{2}\in{\mathsf{T}}_{\ell}\cap{\mathsf{T}}_{\Delta_{0}}}t_{2}^{\lambda}.

    The sum on the right hand side vanishes unless t2λ=1t_{2}^{\lambda}=1 for all t2𝖳𝖳Δ0t_{2}\in{\mathsf{T}}_{\ell}\cap{\mathsf{T}}_{\Delta_{0}}, in which case it is #𝖳𝖳Δ0\#{\mathsf{T}}_{\ell}\cap{\mathsf{T}}_{\Delta_{0}}. But this condition is equivalent to λ,ξ\langle\lambda,\xi\rangle\in\ell{\mathbb{Z}} for all ξΞ𝔱Δ0\xi\in\Xi\cap{\mathfrak{t}}_{\Delta_{0}}, that is, λΛ(Δ0+Ξ)=(Λ)+Ξ\lambda\in\Lambda^{\ast}\cap({\Delta_{0}}+\ell\Xi^{\ast})=(\Lambda^{\prime})^{\ast}+\ell\Xi^{\ast}. Assuming that this is the case, pick λ(Λ)\lambda^{\prime}\in(\Lambda^{\prime})^{\ast} such that λλΞ\lambda-\lambda^{\prime}\in\ell\Xi^{\ast}. Then t1λ=(t1)λt_{1}^{\lambda}=(t_{1}^{\prime})^{\lambda^{\prime}}, where t1𝖳/(T𝖳Δ0)t_{1}^{\prime}\in{\mathsf{T}}_{\ell}/(T_{\ell}\cap{\mathsf{T}}_{\Delta_{0}}) is the image of t1t_{1}. Similarly, t1αk=(t1)αkt_{1}^{-\alpha_{k}}=(t_{1}^{\prime})^{-\alpha^{\prime}_{k}}. Carrying out the sum over 𝖳/(T𝖳Δ0){\mathsf{T}}_{\ell}/(T_{\ell}\cap{\mathsf{T}}_{\Delta_{0}}), we obtain #(T𝖳Δ0)V𝜶¯(λ,)\#(T_{\ell}\cap{\mathsf{T}}_{\Delta_{0}})\ V_{{\underline{{\bm{\alpha}}}}^{\prime}}(\lambda^{\prime},\ell) as desired. ∎

  4. (d)

    We may always reduce to the case that the lattice vectors from the list 𝜶{\bm{\alpha}} are primitive, i.e, not positive multiples of shorter lattice vectors. Indeed, suppose (α,u)𝜶¯(\alpha,u)\in{\underline{{\bm{\alpha}}}} with 1mαΛ\frac{1}{m}\alpha\in\Lambda^{*} for some integer m>1m>1. Let 𝜶¯~\tilde{{\underline{{\bm{\alpha}}}}} be obtained from 𝜶¯{\underline{{\bm{\alpha}}}} by replacing (α,u)(\alpha,u) with (αm,ζ1),(αm,ζm)(\frac{\alpha}{m},\zeta_{1}),\ldots(\frac{\alpha}{m},\zeta_{m}), where ζ1,,ζm\zeta_{1},\ldots,\zeta_{m} are the distinct solutions of ζm=u\zeta^{m}=u. Then V𝜶¯=V𝜶¯~V_{{\underline{{\bm{\alpha}}}}}=V_{\tilde{{\underline{{\bm{\alpha}}}}}}, since

    (1utα)=k=1m(1ζktαm).(1-u\,t^{-\alpha})=\prod_{k=1}^{m}(1-\zeta_{k}\ t^{\frac{-\alpha}{m}}).
  5. (e)

    A function f:Γf\colon\Gamma\rightarrow{\mathbb{C}} on a lattice Γ\Gamma is quasi-polynomial if there is a sublattice ΓΓ\Gamma^{\prime}\subseteq\Gamma of finite index such that ff is given by a polynomial on each coset of Γ\Gamma^{\prime}. More generally, given a subset SΓS\subseteq\Gamma, a function f:Sf\colon S\to{\mathbb{C}} is quasi-polynomial on SS if it is the restriction of a quasi-polynomial on Γ\Gamma. (Note that this condition is vacuous if SS is a finite subset.) If SS contains ‘sufficiently many points’, for example if it contains the intersection of Γ\Gamma with an open cone in the underlying vector space Γ\Gamma\otimes_{\mathbb{Z}}{\mathbb{R}}, then the quasi-polynomial on SS extends uniquely to a quasi-polynomial on all of Γ\Gamma.

    Theorem 2.7 (Szenes[14]).

    Suppose Δ\Delta is a top-dimensional affine subspace in 𝒮(𝛂){\mathcal{S}}({\bm{\alpha}}), and 𝔠Δ{\mathfrak{c}}\subseteq\Delta is a chamber (cf.  Definition 2.2). Then the restriction of the Verlinde sum V𝛂¯V_{{\underline{{\bm{\alpha}}}}} to the set

    (12) {(λ,)Λ×|λ𝔠𝜶}\{(\lambda,\ell)\in\Lambda^{\ast}\times{\mathbb{N}}|\ \lambda\in\ell{\mathfrak{c}}-\Box{\bm{\alpha}}\}

    is quasi-polynomial.

    Szenes’ paper does not state the result in this particular form (which we took from Vergne’s lectures [21]), but it is a simple consequence of the residue formula proved in [14]. See Section 5 below for details. Notice that the theorem gives a quasi-polynomial behaviour not only on the lattice points of the cone >0(𝔠×{1}){\mathbb{R}}_{>0}({\mathfrak{c}}\times\{1\}), but even on a slightly larger region

    {(x,t)𝔱×>0|xt𝔠𝜶}.\{(x,t)\in{\mathfrak{t}}^{\ast}\times{\mathbb{R}}_{>0}|\ x\in t{\mathfrak{c}}-\Box{\bm{\alpha}}\}.

    If Δ0=span𝜶=𝔱\Delta_{0}={\textnormal{span}}_{\mathbb{R}}{\bm{\alpha}}={\mathfrak{t}}^{\ast}, then these regions give an open cover of 𝔱×>0{\mathfrak{t}}^{\ast}\times{\mathbb{R}}_{>0}, as 𝔠{\mathfrak{c}} varies over all open chambers. The appearance of the zonotope may be understood as follows: Note that if if 𝜶¯~\tilde{{\underline{{\bm{\alpha}}}}} is obtained from 𝜶¯{\underline{{\bm{\alpha}}}} by replacing some (β,v)(\beta,v) with (β,v1)(-\beta,v^{-1}), then V𝜶¯(λ,)=v1V𝜶¯~(λ+β,)V_{{\underline{{\bm{\alpha}}}}}(\lambda,\ell)=-v^{-1}\ V_{\tilde{{\underline{{\bm{\alpha}}}}}}(\lambda+\beta,\ \ell). Hence, the quasi-polynomial behavior of V𝜶¯~V_{\tilde{\underline{{\bm{\alpha}}}}} on the cone over 𝔠×{1}\mathfrak{c}\times\{1\} implies the quasi-polynomial behavior of V𝜶¯V_{{\underline{{\bm{\alpha}}}}} on the β\beta-shifted cone. Applying this observation to all the weights, one obtains a quasi-polynomial behavior on the shifted cone.

2.4. Difference equation.

Recall that the Verlinde sums VnV_{n} in Example 2.5 satisfy a difference equation (9), expressing Vn(λ,)Vn(λ1,)V_{n}(\lambda,\ell)-V_{n}(\lambda-1,\ell) in terms of Vn1(λ,)V_{n-1}(\lambda,\ell). We are interested in a version of this equation for the general Verlinde sums associated to a list 𝜶¯{\underline{{\bm{\alpha}}}} of augmented weights. For β¯=(β,v)𝜶¯\underline{\beta}=(\beta,v)\in{\underline{{\bm{\alpha}}}}, denote by 𝜶¯\β¯{\underline{{\bm{\alpha}}}}\backslash\underline{\beta} the list obtained by removing β¯\underline{\beta}. Define a finite difference operator on functions fMap(Λ,)f\in{\textnormal{Map}}(\Lambda^{*},{\mathbb{C}}) (see Appendix A.4):

(β¯f)(λ)=f(λ)vf(λβ).(\nabla_{\underline{\beta}}f)(\lambda)=f(\lambda)-v\,f(\lambda-\beta).

We will see that β¯V𝜶¯\nabla_{\underline{\beta}}V_{{\underline{{\bm{\alpha}}}}} is equal to V𝜶¯\β¯V_{{\underline{{\bm{\alpha}}}}\backslash\underline{\beta}} modulo correction terms involving a lower-dimensional Verlinde sum. Let pp\in{\mathbb{N}} be the smallest natural number such that there exists t0𝖳pt_{0}\in{\mathsf{T}}_{p} with

(13) vt0β=1.v\,t_{0}^{-\beta}=1.

For any t𝖳t\in{\mathsf{T}}, denote by 𝖾t:Λ\mathsf{e}_{t}\colon\Lambda^{*}\to{\mathbb{C}} the map λtλ\lambda\mapsto t^{\lambda}. Recall the notation 𝔱β=kerβ,𝖳β=exp(𝔱β){\mathfrak{t}}_{\beta}=\operatorname{ker}\beta,\ {\mathsf{T}}_{\beta}=\exp({\mathfrak{t}}_{\beta}) from Definition 2.2, and put Λβ=Λ𝔱β\Lambda_{\beta}=\Lambda\cap{\mathfrak{t}}_{\beta} and Ξβ=Ξ𝔱β\Xi_{\beta}=\Xi\cap{\mathfrak{t}}_{\beta}.

Proposition 2.8 (Difference equation).

Given β¯𝛂¯\underline{\beta}\in{\underline{{\bm{\alpha}}}}, define pp\in{\mathbb{N}} and t0𝖳pt_{0}\in{\mathsf{T}}_{p} by (13). Then

(14) β¯V𝜶¯=V𝜶¯β¯𝖾t0δpπV𝜶¯.\nabla_{\underline{\beta}}V_{{\underline{{\bm{\alpha}}}}}=V_{{\underline{{\bm{\alpha}}}}{\setminus}\underline{\beta}}-\mathsf{e}_{t_{0}}\ \delta_{p{\mathbb{N}}}\ \pi^{*}V_{{\underline{{\bm{\alpha}}}}^{\prime}}.

Here π:ΛΛβ\pi\colon\Lambda^{*}\to\Lambda_{\beta}^{*} is the projection, and 𝛂¯{\underline{{\bm{\alpha}}}}^{\prime} is the list of all augmented weights (π(α),t0αu)(\pi(\alpha),\,t_{0}^{-\alpha}\,u) with (α,u)𝛂¯\β¯(\alpha,u)\in{\underline{{\bm{\alpha}}}}\backslash\underline{\beta}.

Proof.

Under finite Fourier transform, the difference operator β¯\nabla_{\underline{\beta}} amounts to multiplication by 1vtβ1-vt^{-\beta}. (See Section A.5.) Thus, the summands in the formula for β¯V𝜶¯(λ,)\nabla_{\underline{\beta}}V_{{\underline{{\bm{\alpha}}}}}(\lambda,\ell) are the same as those for V𝜶¯β¯(λ,)V_{{\underline{{\bm{\alpha}}}}{\setminus}\underline{\beta}}(\lambda,\ell), but one is summing over t𝖳t\in{\mathsf{T}}_{\ell} such that utα1ut^{-\alpha}\neq 1 for all (α,u)α(\alpha,u)\in\bf\alpha, whereas in the sum defining V𝜶¯β¯(λ,)V_{{\underline{{\bm{\alpha}}}}{\setminus}\underline{\beta}}(\lambda,\ell) this is only required for (α,u)𝜶¯\β¯(\alpha,u)\in{\underline{{\bm{\alpha}}}}\backslash\underline{\beta}. Thus,

(15) V𝜶¯β¯(λ,)β¯V𝜶¯(λ,)=t𝖳′′tλ(α,u)𝜶¯\β¯(1utα)V_{{\underline{{\bm{\alpha}}}}{\setminus}\underline{\beta}}(\lambda,\ell)-\nabla_{\underline{\beta}}V_{{\underline{{\bm{\alpha}}}}}(\lambda,\ell)={\sum}^{\prime\prime}_{t\in{\mathsf{T}}_{\ell}}\frac{t^{\lambda}}{\prod_{(\alpha,u)\in{\underline{{\bm{\alpha}}}}\backslash\underline{\beta}}\,(1-u\,t^{-\alpha})}

where the sum ′′\sum^{\prime\prime} is over all t𝖳t\in{\mathsf{T}}_{\ell} such that utα1ut^{-\alpha}\neq 1 for α¯β¯\underline{\alpha}\neq\underline{\beta}, but vtβ=1vt^{-\beta}=1. After fixing t0t_{0} satisfying (13), one can find t𝖳t\in{\mathsf{T}}_{\ell} with vtβ=1v\,t^{-\beta}=1 if and only if \ell is a multiple of pp, and any such tt differs from t0t_{0} by an element of 𝖳𝖳β{\mathsf{T}}_{\ell}\cap{\mathsf{T}}_{\beta}. Hence, if \ell is not a multiple of pp, then (15) is an empty sum. If \ell is a multiple of pp, write t=t0tt=t_{0}t^{\prime} as above. As tt ranges over all elements of 𝖳{\mathsf{T}}_{\ell} such that vtβ=1vt^{-\beta}=1, the elements tt^{\prime} range over 𝖳𝖳β{\mathsf{T}}_{\ell}\cap{\mathsf{T}}_{\beta}. Furthermore,

tλ=t0λ(t)π(λ),utα=u(t)αt^{\lambda}=t_{0}^{\lambda}\,(t^{\prime})^{\pi(\lambda)},\ \ \ u\,t^{-\alpha}=u^{\prime}\,(t^{\prime})^{-\alpha^{\prime}}

for (α,u)𝜶¯\β¯(\alpha,u)\in{\underline{{\bm{\alpha}}}}\backslash\underline{\beta}. Hence the sum on the right hand side becomes t0λV𝜶¯(π(λ),)t_{0}^{\lambda}\ V_{{\underline{{\bm{\alpha}}}}^{\prime}}(\pi(\lambda),\ell). ∎

Remark 2.9.
  1. (a)

    In Proposition 2.8, the Verlinde sum V𝜶¯V_{{\underline{{\bm{\alpha}}}}^{\prime}} depends on the choice of t0t_{0}, but its product with 𝖾t0\mathsf{e}_{t_{0}} does not.

  2. (b)

    If there exists (α,u)(β,v)(\alpha,u)\neq(\beta,v) in the list 𝜶¯{\underline{{\bm{\alpha}}}} such that utα=1ut^{-\alpha}=1 whenever vtβ=1vt^{-\beta}=1, then the summation in (15) is over an empty set, hence the sum is zero. In particular , this is the case if the list 𝜶¯{\underline{{\bm{\alpha}}}} contains a second copy of β¯=(β,v)\underline{\beta}=(\beta,v).

Remark 2.10.

The counterpart to (14) for multiple Bernoulli series expresses the directional derivative βB𝜶\partial_{\beta}B_{{\bm{\alpha}}} in terms of B𝜶\βB_{{\bm{\alpha}}\backslash\beta}, and possibly a correction term involving a lower-dimensional Bernoulli series. See [5, Proposition 3.1].

3. Partition functions.

Let 𝜶¯=(𝜶,𝒖){\underline{{\bm{\alpha}}}}=({\bm{\alpha}},{\bm{u}}) be a list of augmented weights. We will assume that all α𝜶\alpha\in{\bm{\alpha}} are nonzero. Replacing summation over 𝖳{\mathsf{T}}_{\ell} in the definition of the Verlinde sum by an integration over 𝖳{\mathsf{T}}, we obtain at a formal expression

Ttλ(α,u)𝜶¯(1utα)𝑑t.\int_{T}\frac{t^{\lambda}}{\prod_{(\alpha,u)\in{\underline{{\bm{\alpha}}}}}\,(1-u\,t^{-\alpha})}\ dt.

As it stands, the integral is not well-defined due to the zeroes of the denominator. To ‘regularize’ the integral, choose a polarizing vector τ𝔱\tau\in{\mathfrak{t}} such that α,τ0\langle\alpha,\tau\rangle\neq 0 for all α𝜶\alpha\in{\bm{\alpha}}. Then

(16) limϵ0+(α,u)𝜶¯(1utαeϵα,τ))1\lim_{\epsilon\to 0^{+}}\prod_{(\alpha,u)\in{\underline{{\bm{\alpha}}}}}\,\big{(}1-u\,t^{-\alpha}\ e^{-\epsilon\langle\alpha,\tau\rangle})\big{)}^{-1}

is a well-defined generalized function of t𝖳t\in{\mathsf{T}}. Taking its Fourier transform, we arrive at the following definition.

Definition 3.1.

The generalized partition function P(𝜶¯,τ):ΛP_{({\underline{{\bm{\alpha}}}},\tau)}\colon\Lambda^{\ast}\rightarrow{\mathbb{C}} is defined by

P(𝜶¯,τ)(λ)=limϵ0+Ttλ(α,u)𝜶¯(1utαeϵα,τ)𝑑t.P_{({\underline{{\bm{\alpha}}}},\tau)}(\lambda)=\lim_{\epsilon\to 0^{+}}\int_{T}\frac{t^{\lambda}}{\prod_{(\alpha,u)\in{\underline{{\bm{\alpha}}}}}\,(1-u\,t^{-\alpha}\ e^{-\epsilon\langle\alpha,\tau\rangle})}\ dt.

If all u=1u=1 for all u𝒖u\in{\bm{u}}, we will also use the notation P(𝜶,τ)P_{({\bm{\alpha}},\tau)}.

Generalized partition functions were studied, e.g., in [8, 9, 19]. They are a discrete counterpart to the generalized Heaviside function

H(𝜶,τ)(λ)=limϵ0+𝔱e2πiλ,ξα𝜶α,2πiξ+ϵτ𝑑ξ,H_{({\bm{\alpha}},\tau)}(\lambda)=\lim_{\epsilon\to 0^{+}}\int_{{\mathfrak{t}}}\frac{e^{2\pi{\mathrm{i}}\langle\lambda,\xi\rangle}}{\prod_{\alpha\in{\bm{\alpha}}}\langle\alpha,2\pi{\mathrm{i}}\xi+\epsilon\tau\rangle}\ d\xi,

here dξd\xi is the measure on 𝔱{\mathfrak{t}} corresponding to the normalized Haar measure on 𝖳{\mathsf{T}}, and the integral is defined as a generalized function of λ𝔱\lambda\in{\mathfrak{t}}^{\ast}.

Example 3.2.

We continue Example 2.5, thus 𝔱=,Λ=Ξ={\mathfrak{t}}={\mathbb{R}},\ \Lambda=\Xi={\mathbb{Z}}, with 𝜶{\bm{\alpha}} the list consisting of the weight 11, repeated nn times. Let τ=1\tau=1, and write Hn(λ)=H(𝜶,τ)(λ)H_{n}(\lambda)=H_{({\bm{\alpha}},\tau)}(\lambda) and Pn(λ)=P(𝜶,τ)(λ)P_{n}(\lambda)=P_{({\bm{\alpha}},\tau)}(\lambda). One finds that

Hn(λ)={λn1(n1)!λ0,0λ<0.H_{n}(\lambda)=\begin{cases}\frac{\lambda^{n-1}}{(n-1)!}&\lambda\geq 0,\\ 0&\lambda<0.\end{cases}

Similarly, Pn(λ)P_{n}(\lambda) is the number of ways of writing λ\lambda\in\mathbb{Z} as a sum λ=k1++kn\lambda=k_{1}+\ldots+k_{n} where all ki0k_{i}\geq 0. Thus,

P1(λ)=1,P2(λ)=λ+1,P3(λ)=12λ2+32λ+1,P_{1}(\lambda)=1,\ \ P_{2}(\lambda)=\lambda+1,\ \ P_{3}(\lambda)=\frac{1}{2}\lambda^{2}+\frac{3}{2}\lambda+1,\ldots

for λ0\lambda\geq 0, while Pn(λ)=0P_{n}(\lambda)=0 for λ<0\lambda<0.

Here are some of the basic properties of the generalized partition functions.

  1. (a)

    Write 𝜶¯={α¯1,,α¯n}{\underline{{\bm{\alpha}}}}=\{\underline{\alpha}_{1},\ldots,\underline{\alpha}_{n}\} with α¯k=(αk,uk)\underline{\alpha}_{k}=(\alpha_{k},u_{k}). If the list 𝜶{\bm{\alpha}} is polarized in the sense that αk,τ>0\langle\alpha_{k},\tau\rangle>0 for all kk, then the integrand (16) can be expanded into a geometric series, and the limit ϵ0\epsilon\to 0 gives

    k=1n(jk=0ukjktjkαk),\prod_{k=1}^{n}\big{(}\sum_{j_{k}=0}^{\infty}u_{k}^{j_{k}}\ t^{-j_{k}\alpha_{k}}\big{)},

    where the sum is defined as a generalized function on 𝖳{\mathsf{T}}. The integration against tλt^{\lambda} picks out the coefficient of tλt^{-\lambda} in this expansion, i.e., P(𝜶¯,τ)(λ)=u1j1unjnP_{({\underline{{\bm{\alpha}}}},\tau)}(\lambda)=\sum u_{1}^{j_{1}}\cdots u_{n}^{j_{n}} where the sum is over all solutions of j1α1++jnαn=λj_{1}\alpha_{1}+\ldots+j_{n}\alpha_{n}=\lambda, with jk0j_{k}\in\mathbb{Z}_{\geq 0}. If all uk=1u_{k}=1, then it is simply the number of such solutions; thus P(𝜶¯,τ)P_{({\underline{{\bm{\alpha}}}},\tau)} becomes a vector partition function. If 𝜶{\bm{\alpha}} is the set +\mathfrak{R}_{+} of positive roots of a simple Lie algebra, and τ\tau is in the positive Weyl chamber, then P(𝜶,τ)P_{({\bm{\alpha}},\tau)} is the Kostant partition function.

    In a similar fashion, H(𝜶,τ)(λ)H_{({\bm{\alpha}},\tau)}(\lambda) for a polarized list 𝜶{\bm{\alpha}} is a constant (not depending on λ\lambda) times the volume of the n1n-1-dimensional polytope in n\mathbb{R}^{n} defined by j1α1++jnαn=λj_{1}\alpha_{1}+\ldots+j_{n}\alpha_{n}=\lambda, with jk0j_{k}\in\mathbb{R}_{\geq 0}. See [8] for a detailed discussion.

  2. (b)

    Suppose 𝜶¯~\tilde{\underline{{\bm{\alpha}}}} is obtained from 𝜶¯{\underline{{\bm{\alpha}}}} by replacing some element (β,v)𝜶¯(\beta,v)\in{\underline{{\bm{\alpha}}}} with (β,v1)(-\beta,v^{-1}). Then

    P(𝜶¯,τ)(λ)=v1P(𝜶¯~,τ)(λ+β).P_{({\underline{{\bm{\alpha}}}},\tau)}(\lambda)=-v^{-1}P_{(\tilde{\underline{{\bm{\alpha}}}},\tau)}(\lambda+\beta).

    This follows from (1z)1=z1(1z1)1(1-z)^{-1}=-z^{-1}(1-z^{-1})^{-1}.

  3. (c)

    Decompose the list 𝜶¯{\underline{{\bm{\alpha}}}} into two sublists 𝜶¯=𝜶¯+𝜶¯{\underline{{\bm{\alpha}}}}={\underline{{\bm{\alpha}}}}_{+}\cup{\underline{{\bm{\alpha}}}}_{-}, where α𝜶¯+\alpha\in{\underline{{\bm{\alpha}}}}_{+} if α,τ>0\langle\alpha,\tau\rangle>0 and α𝜶¯\alpha\in{\underline{{\bm{\alpha}}}}_{-} if α,τ<0\langle\alpha,\tau\rangle<0. Let 𝜶¯pol{\underline{{\bm{\alpha}}}}^{pol} be the polarized list, consisting of 𝜶¯+{\underline{{\bm{\alpha}}}}_{+} together with all (α,u1)(-\alpha,u^{-1}) such that (α,u)𝜶¯(\alpha,u)\in{\underline{{\bm{\alpha}}}}_{-}. Define

    σ=α𝜶α.\sigma=-\sum_{\alpha\in{\bm{\alpha}}_{-}}\alpha.

    Then σ,τ>0\langle\sigma,\tau\rangle>0, and item (b) above shows

    P(𝜶¯,τ)(λ)=(1)#𝜶u𝒖uP(𝜶¯pol,τ)(λσ).P_{({\underline{{\bm{\alpha}}}},\tau)}(\lambda)=\frac{(-1)^{\#{\bm{\alpha}}_{-}}}{\prod_{u\in{\bm{u}}_{-}}u}\ P_{({\underline{{\bm{\alpha}}}}^{pol},\tau)}(\lambda-\sigma).

    In particular, using item (a) above, we see that the generalized partition function is supported in the closed cone spanned by the polarized weights, shifted by σ\sigma:

    supp(P(𝜶¯,τ))σ+0𝜶pol.\mathrm{supp}(P_{({\underline{{\bm{\alpha}}}},\tau)})\subseteq\sigma+{\mathbb{Z}}_{\geq 0}\ {\bm{\alpha}}^{pol}.
  4. (d)

    In the case 𝜶{\bm{\alpha}} spans 𝔱\mathfrak{t}^{\ast}, the functions H(𝜶,τ)H_{({\bm{\alpha}},\tau)} are piecewise polynomial L1L^{1}-functions on 𝔱{\mathfrak{t}}^{\ast}, where the polynomials are supported on cones. Similarly, the partition functions P(𝜶¯,τ)P_{({\underline{{\bm{\alpha}}}},\tau)} exhibit a piecewise quasi-polynomial behavior. In the case of vector partition functions, this goes back to Dahmen-Micchelli [7]; see Szenes-Vergne [19, Theorem 3.6] for the more general case.

  5. (e)

    If 𝜶¯{\underline{{\bm{\alpha}}}} is a disjoint union of two sublists 𝜶¯,𝜶¯′′{\underline{{\bm{\alpha}}}}^{\prime},{\underline{{\bm{\alpha}}}}^{\prime\prime}, then

    P(𝜶¯,τ)=P(𝜶¯,τ)P(𝜶¯′′,τ).P_{({\underline{{\bm{\alpha}}}},\tau)}=P_{({\underline{{\bm{\alpha}}}}^{\prime},\tau)}\star P_{({\underline{{\bm{\alpha}}}}^{\prime\prime},\tau)}.

    Here the convolution (of functions on Λ\Lambda^{\ast}) is well-defined, due to the support properties.

  6. (f)

    For β¯=(β,v)𝜶¯\underline{\beta}=(\beta,v)\in{\underline{{\bm{\alpha}}}}, since the finite-difference operator β¯\nabla_{\underline{\beta}} has the effect of multiplying the inverse Fourier transform by (1vtβ)(1-vt^{-\beta}), we see that

    β¯P(𝜶¯,τ)=P(𝜶¯\β¯,τ).\nabla_{\underline{\beta}}P_{({\underline{{\bm{\alpha}}}},\tau)}=P_{({\underline{{\bm{\alpha}}}}\backslash\underline{\beta},\tau)}.
  7. (g)

    For h𝖳h\in{\mathsf{T}}, multiplication by the function 𝖾h:Λ,λhλ\mathsf{e}_{h}\colon\Lambda^{*}\to{\mathbb{C}},\ \lambda\mapsto h^{\lambda} has the effect of shifting the scalars of the augmented weights: That is,

    𝖾hP(𝜶¯,τ)=P(𝜶¯~,τ)\mathsf{e}_{h}\ P_{({\underline{{\bm{\alpha}}}},\tau)}=P_{(\tilde{\underline{{\bm{\alpha}}}},\tau)}

    where 𝜶¯~\underline{\tilde{{\bm{\alpha}}}} consists of all (α,uhα)(\alpha,\,u\,h^{\alpha}) with (α,u)𝜶¯(\alpha,u)\in\underline{{\bm{\alpha}}}.

  8. (h)

    Suppose 𝔥𝔱\mathfrak{h}\subseteq{\mathfrak{t}} is a rational hyperplane, and let π:Λ(Λ𝔥)=Λ/(Λann(𝔥))\pi\colon\Lambda^{\ast}\to(\Lambda\cap\mathfrak{h})^{\ast}=\Lambda^{*}/(\Lambda^{*}\cap\mathrm{ann}(\mathfrak{h})) be the quotient map. Suppose the restrictions α|𝔥=π(α)\alpha|_{\mathfrak{h}}=\pi(\alpha) for α𝜶\alpha\in{\bm{\alpha}} are all non-zero. If τ𝔥\tau\in{\mathfrak{h}} is polarizing for 𝜶{\bm{\alpha}}, then τ\tau is also polarizing for the list 𝜶|𝔥{\bm{\alpha}}|_{\mathfrak{h}} consisting of the restrictions, and

    P(𝜶¯|𝔥,τ)=πP(𝜶¯,τ)P_{({\underline{{\bm{\alpha}}}}|_{\mathfrak{h}},\tau)}=\pi_{*}P_{({\underline{{\bm{\alpha}}}},\tau)}

    where 𝜶¯|𝔥=(𝜶|𝔥,𝒖){\underline{{\bm{\alpha}}}}|_{\mathfrak{h}}=({\bm{\alpha}}|_{\mathfrak{h}},{\bm{u}}). Here the push-forward (cf. Appendix A.2) is well-defined, since π\pi is proper on the support of P(𝜶¯,τ)P_{({\underline{{\bm{\alpha}}}},\tau)} by item (c) above.

4. Decomposition formula.

Consider a list of weights 𝜶{\bm{\alpha}}, defining a Bernoulli series B𝜶B_{{\bm{\alpha}}}. For any admissible affine subspace Δ𝒮(𝜶)\Delta\in{\mathcal{S}}({\bm{\alpha}}), let 𝜶Δ{\bm{\alpha}}_{\Delta} be the sublist consisting of all α𝜶ann(𝔱Δ)\alpha\in{\bm{\alpha}}\cap\mathrm{ann}({\mathfrak{t}}_{\Delta}); that is, α\alpha is parallel to Δ\Delta. Its complement is denoted 𝜶Δ𝖼=𝜶\𝜶Δ{\bm{\alpha}}^{\mathsf{c}}_{\Delta}={\bm{\alpha}}\backslash{\bm{\alpha}}_{\Delta}.

Fix an inner product on 𝔱{\mathfrak{t}}, used to identify 𝔱𝔱{\mathfrak{t}}\simeq{\mathfrak{t}}^{\ast}, and let γ𝔱\gamma\in{\mathfrak{t}}. Given Δ𝒮(𝜶)\Delta\in{\mathcal{S}}({\bm{\alpha}}), let

γΔΔ\gamma_{\Delta}\in\Delta

denote the orthogonal projection of γ\gamma onto Δ\Delta, and put

τΔ=γΔγ𝔱Δ.\tau_{\Delta}=\gamma_{\Delta}-\gamma\in\mathfrak{t}_{\Delta}.

For a generic choice of γ\gamma, all the projections γΔ\gamma_{\Delta} for Δ𝒮(𝜶)\Delta\in{\mathcal{S}}({\bm{\alpha}}) are in open chambers for Δ\Delta, while the τΔ\tau_{\Delta} are polarizing vectors for the complement 𝜶Δ𝖼{\bm{\alpha}}^{\mathsf{c}}_{\Delta}.

Suppose μΔ\mu\in\Delta is regular in Δ\Delta, i.e., contained in one of the chambers 𝔠{\mathfrak{c}} of Δ\Delta. Boysal-Vergne defined

Ber(𝜶Δ;μ):𝔱\operatorname{Ber}_{({\bm{\alpha}}_{\Delta};\mu)}\colon{\mathfrak{t}}^{*}\to{\mathbb{C}}

to be the generalized function on 𝔱{\mathfrak{t}}^{*} with support on the affine subspace Δ\Delta, which is given by a polynomial on Δ\Delta times a δ\delta-distribution in directions transverse to Δ\Delta, and which coincides with B𝜶ΔB_{{\bm{\alpha}}_{\Delta}} on the chamber 𝔠{\mathfrak{c}}. We will call this the polynomial germ of B𝛂B_{{\bm{\alpha}}} with respect to 𝔠{\mathfrak{c}}. In [5, Theorem 9.3], A. Boysal and M. Vergne proved the following decomposition formula for Bernoulli series:

(17) B𝜶=Δ𝒮Ber(𝜶Δ;γΔ)H(𝜶Δ𝖼,τΔ),B_{{\bm{\alpha}}}=\sum_{\Delta\in{\mathcal{S}}}\operatorname{Ber}_{({\bm{\alpha}}_{\Delta};\gamma_{\Delta})}\star H_{({\bm{\alpha}}_{\Delta}^{\mathsf{c}},\tau_{\Delta})},

(equality of generalized functions on 𝔱{\mathfrak{t}}^{\ast}). The formula expresses the periodic generalized function B𝜶B_{{\bm{\alpha}}} as a locally finite sum of generalized functions. The term indexed by Δ\Delta is supported on the affine cone Δ+0{𝜶Δ𝖼}\Delta+\mathbb{R}_{\geq 0}\{{\bm{\alpha}}_{\Delta}^{\mathsf{c}}\}; a given point of 𝔱{\mathfrak{t}}^{*} is contained in only finitely many of these affine cones. In particular, none of the cones with Δ𝔱\Delta\neq{\mathfrak{t}}^{*} contains the center of expansion. If Δ0=span(𝜶)=𝔱\Delta_{0}={\textnormal{span}}({\bm{\alpha}})={\mathfrak{t}}^{\ast}, then Δ0\Delta_{0} gives a leading polynomial contribution supported on 𝔱{\mathfrak{t}}^{\ast}, while the other terms give successive corrections away from the center of expansion.

In this section, we will give an analogous result for the Verlinde sums. Let 𝜶¯{\underline{{\bm{\alpha}}}} be a list of augmented weights. For Δ𝒮(𝜶)\Delta\in{\mathcal{S}}({\bm{\alpha}}), denote by 𝜶¯Δ=(𝜶Δ,𝒖Δ){\underline{{\bm{\alpha}}}}_{\Delta}=({\bm{\alpha}}_{\Delta},{\bm{u}}_{\Delta}) the sublist consisting of all (α,u)𝜶¯(\alpha,u)\in{\underline{{\bm{\alpha}}}} such that αann(𝔱Δ)\alpha\in\mathrm{ann}({\mathfrak{t}}_{\Delta}). According to Theorem 2.7, if 𝔠{\mathfrak{c}} is a chamber of Δ\Delta, the Verlinde sum V𝜶¯ΔV_{{\underline{{\bm{\alpha}}}}_{\Delta}} is quasi-polynomial on the set of (λ,)(\lambda,\ell) such that λ𝔠\lambda\in\ell{\mathfrak{c}}. Given μ𝔠\mu\in{\mathfrak{c}}, we define

(18) Ver(𝜶¯Δ;μ):Λ×{\textnormal{Ver}}_{({\underline{{\bm{\alpha}}}}_{\Delta};\mu)}\colon\Lambda^{\ast}\times{\mathbb{N}}\to{\mathbb{C}}

to be the unique function supported on the set of (λ,)(\lambda,\ell) with λΔ\lambda\in\ell\Delta, such that this function is quasi-polynomial on this set and agrees with V𝜶¯ΔV_{{\underline{{\bm{\alpha}}}}_{\Delta}} on the set of all (λ,)(\lambda,\ell) with λ𝔠\lambda\in\ell\mathfrak{c}. We refer to Ver(𝜶¯Δ;μ){\textnormal{Ver}}_{({\underline{{\bm{\alpha}}}}_{\Delta};\mu)} as the quasi-polynomial germ of V𝜶¯ΔV_{{\underline{{\bm{\alpha}}}}_{\Delta}} at μ\mu.

Remark 4.1.

By Theorem 2.7, the Verlinde sum V𝜶¯ΔV_{{\underline{{\bm{\alpha}}}}_{\Delta}} agrees with Ver(𝜶¯Δ;μ){\textnormal{Ver}}_{({\underline{{\bm{\alpha}}}}_{\Delta};\mu)} on the slightly larger set where λ𝔠𝜶Δ\lambda\in\ell\mathfrak{c}-\Box{\bm{\alpha}}_{\Delta}.

Pick a generic γ𝔱\gamma\in{\mathfrak{t}}, and define γΔ,τΔ\gamma_{\Delta},\ \tau_{\Delta} for Δ𝒮(𝜶)\Delta\in{\mathcal{S}}({\bm{\alpha}}) as above. The polarizing vectors τΔ\tau_{\Delta} define partition functions P(𝜶¯Δ𝖼,τΔ)P_{({\underline{{\bm{\alpha}}}}_{\Delta}^{\mathsf{c}},\tau_{\Delta})}, with support in a shifted cone σΔ+0(𝜶Δ𝖼)pol\sigma_{\Delta}+{\mathbb{R}}_{\geq 0}({\bm{\alpha}}_{\Delta}^{\mathsf{c}})^{pol} where (see Section 3(c))

(19) σΔ=α(𝜶Δ𝖼)α,\sigma_{\Delta}=-\sum_{\alpha\in({\bm{\alpha}}^{\mathsf{c}}_{\Delta})_{-}}\alpha,

a sum over weights α\alpha in 𝜶Δ𝖼{\bm{\alpha}}_{\Delta}^{\mathsf{c}} such that α,τΔ<0\langle\alpha,\tau_{\Delta}\rangle<0. Then σΔ,τΔ0\langle\sigma_{\Delta},\tau_{\Delta}\rangle\geq 0; in fact μ,τΔ0\langle\mu,\tau_{\Delta}\rangle\geq 0 for all elements in this shifted cone.

Theorem 4.2 (Decomposition formula for Verlinde sums).

Let 𝛂¯{\underline{{\bm{\alpha}}}} be a list of augmented weights, and let γ𝔱\gamma\in{\mathfrak{t}}^{\ast} be generic. Then

(20) V𝜶¯=Δ𝒮Ver(𝜶¯Δ;γΔ)P(𝜶¯Δ𝖼,τΔ)V_{{\underline{{\bm{\alpha}}}}}=\sum_{\Delta\in{\mathcal{S}}}{\textnormal{Ver}}_{({\underline{{\bm{\alpha}}}}_{\Delta};\gamma_{\Delta})}\star P_{({\underline{{\bm{\alpha}}}}_{\Delta}^{\mathsf{c}},\tau_{\Delta})}

(a locally finite sum). Here the term indexed by Δ\Delta is is a (well-defined) convolution of functions on Λ\Lambda^{\ast}, for fixed \ell\in{\mathbb{N}}.

Proof.

We first show that the right hand side of the decomposition formula is locally finite, and hence well defined. The first factor in each term

(21) V𝜶¯Δ=Ver(𝜶¯Δ;γΔ)P(𝜶¯Δ𝖼,τΔ)V_{{\underline{{\bm{\alpha}}}}}^{\Delta}={\textnormal{Ver}}_{({\underline{{\bm{\alpha}}}}_{\Delta};\gamma_{\Delta})}\star P_{({\underline{{\bm{\alpha}}}}_{\Delta}^{\mathsf{c}},\tau_{\Delta})}

is supported in the set of (λ,)(\lambda,\ell) such that λΔ\lambda\in\ell\Delta, while according to Section 3 (c), the second factor is supported in σΔ+0(𝜶Δ𝖼)pol\sigma_{\Delta}+{\mathbb{R}}_{\geq 0}({\bm{\alpha}}_{\Delta}^{\mathsf{c}})^{pol}, using the polarization of 𝜶Δ𝖼{\bm{\alpha}}_{\Delta}^{\mathsf{c}} defined by τΔ\tau_{\Delta}. This shows that the convolution has support in the the set of all (λ,)Λ×(\lambda,\ell)\in\Lambda^{\ast}\times{\mathbb{N}} such that λ\lambda is contained in the affine cone

(22) σΔ+Δ+0(𝜶Δ𝖼)pol.\sigma_{\Delta}+\ell\Delta+{\mathbb{R}}_{\geq 0}({\bm{\alpha}}_{\Delta}^{\mathsf{c}})^{pol}.

For fixed λΛ\lambda\in\Lambda^{*}, there are only finitely many Δ\Delta such that λ\lambda is contained in (22). Hence the sum in (20) is locally finite.

The proof of the equality V𝜶¯=ΔV𝜶¯ΔV_{{\underline{{\bm{\alpha}}}}}=\sum_{\Delta}V_{{\underline{{\bm{\alpha}}}}}^{\Delta} involves an induction on the number nn of elements in 𝜶{\bm{\alpha}}. Consider first the base case n=0n=0, so that the admissible subspaces are all 0-dimensional: 𝒮()=Ξ{\mathcal{S}}(\emptyset)=\Xi^{\ast}. For Δ={μ}\Delta=\{\mu\} with μΞ\mu\in\Xi^{\ast}, we have that 𝖳Δ=𝖳{\mathsf{T}}_{\Delta}={\mathsf{T}}, γΔ=μ\gamma_{\Delta}=\mu, and consequently VΔ(λ,)=#𝖳δμ(λ)V_{\emptyset}^{\Delta}(\lambda,\ell)=\#{\mathsf{T}}_{\ell}\ \delta_{\ell\mu}(\lambda). Summation over all Δ={μ}\Delta=\{\mu\} gives the expression (7) for VV_{\emptyset}.

Suppose now that 𝜶{\bm{\alpha}} has n>0n>0 elements, and that the decomposition formula has been proved for lists with at most n1n-1 elements. The argument for V𝜶V_{{\bm{\alpha}}} splits into two stages. In the first step, we will show that the difference V𝜶¯ΔV𝜶¯ΔV_{{\underline{{\bm{\alpha}}}}}-\sum_{\Delta}V_{{\underline{{\bm{\alpha}}}}}^{\Delta} is annihilated by all difference operators β¯\nabla_{\underline{\beta}}, for β¯𝜶¯\underline{\beta}\in{\underline{{\bm{\alpha}}}}. By definition of the difference operator, this implies that for any given \ell, this difference is quasi-invariant under the action of the sublattice 𝜶Λ{\mathbb{Z}}{\bm{\alpha}}\subseteq\Lambda^{\ast}, i.e. invariant up to some character 𝜶U(1){\mathbb{Z}}{\bm{\alpha}}\to\mathrm{U}(1). In the second step, we show that V𝜶V_{{\bm{\alpha}}} coincides with ΔV𝜶¯Δ\sum_{\Delta}V_{{\underline{{\bm{\alpha}}}}}^{\Delta} on the lattice points of some fundamental domain for the action of 𝜶{\mathbb{Z}}{\bm{\alpha}}. Together, these two steps show that they agree everywhere.


Step 1. Claim: For all β¯𝜶¯\underline{\beta}\in{\underline{{\bm{\alpha}}}}, we have that β¯V𝜶¯=Δ𝒮(𝜶)β¯V𝜶¯Δ\nabla_{\underline{\beta}}V_{{\underline{{\bm{\alpha}}}}}=\sum_{\Delta\in{\mathcal{S}}({\bm{\alpha}})}\nabla_{\underline{\beta}}V_{{\underline{{\bm{\alpha}}}}}^{\Delta}.

Recall the difference equation (14): β¯V𝜶¯=V𝜶¯β¯𝖾t0δpπV𝜶¯\nabla_{\underline{\beta}}V_{{\underline{{\bm{\alpha}}}}}=V_{{\underline{{\bm{\alpha}}}}{\setminus}\underline{\beta}}-\mathsf{e}_{t_{0}}\ \delta_{p{\mathbb{N}}}\ \pi^{*}V_{{\underline{{\bm{\alpha}}}}^{\prime}}. To compare with the sum over all β¯V𝜶¯Δ\nabla_{\underline{\beta}}V_{{\underline{{\bm{\alpha}}}}}^{\Delta} with Δ𝒮(𝜶)\Delta\in{\mathcal{S}}({\bm{\alpha}}), we consider two cases:

Case (i): β¯𝜶¯Δ𝖼\underline{\beta}\in{\underline{{\bm{\alpha}}}}_{\Delta}^{\mathsf{c}}. Apply β¯\nabla_{\underline{\beta}} to (21), using the property β¯(fg)=fβ¯(g)\nabla_{\underline{\beta}}(f\star g)=f\star\nabla_{\underline{\beta}}(g) of the convolution. Since β¯P(𝜶¯Δ𝖼,τΔ)=P(𝜶¯Δ𝖼β¯;τΔ)\nabla_{\underline{\beta}}P_{({\underline{{\bm{\alpha}}}}_{\Delta}^{\mathsf{c}},\tau_{\Delta})}=P_{({\underline{{\bm{\alpha}}}}_{\Delta}^{\mathsf{c}}{\setminus}\underline{\beta};\,\tau_{\Delta})} by Section 3(f), and since 𝜶¯Δ=(𝜶¯β¯)Δ{\underline{{\bm{\alpha}}}}_{\Delta}=({\underline{{\bm{\alpha}}}}{\setminus}\underline{\beta})_{\Delta}, we obtain

(23) β¯V𝜶¯Δ=V𝜶¯\β¯Δ.\nabla_{\underline{\beta}}V_{{\underline{{\bm{\alpha}}}}}^{\Delta}=V_{{\underline{{\bm{\alpha}}}}\backslash\underline{\beta}}^{\Delta}.

Case (ii): β¯𝜶¯Δ\underline{\beta}\in{\underline{{\bm{\alpha}}}}_{\Delta}. Apply β¯\nabla_{\underline{\beta}} to (21), using the property β¯(fg)=β¯(f)g\nabla_{\underline{\beta}}(f\star g)=\nabla_{\underline{\beta}}(f)\star g of the convolution. Lemma 4.3 below gives

(24) β¯Ver(𝜶¯Δ;γΔ)=Ver(𝜶¯Δ\β¯;γΔ)𝖾t0δpπVer(𝜶¯π(Δ);π(γΔ)).\nabla_{\underline{\beta}}{\textnormal{Ver}}_{({\underline{{\bm{\alpha}}}}_{\Delta};\gamma_{\Delta})}={\textnormal{Ver}}_{({\underline{{\bm{\alpha}}}}_{\Delta}\backslash\underline{\beta};\gamma_{\Delta})}-\mathsf{e}_{t_{0}}\ \delta_{p{\mathbb{N}}}\ \pi^{*}\,{\textnormal{Ver}}_{({\underline{{\bm{\alpha}}}}^{\prime}_{\pi(\Delta)};\pi(\gamma_{\Delta}))}.

Taking a convolution with P(𝜶¯Δ𝖼,τΔ)P_{({\underline{{\bm{\alpha}}}}_{\Delta}^{\mathsf{c}},\tau_{\Delta})}, the first term produces V𝜶¯\β¯ΔV_{{\underline{{\bm{\alpha}}}}\backslash\underline{\beta}}^{\Delta}. For the second term, we use the property 𝖾t(fg)=(𝖾tf)(𝖾tg)\mathsf{e}_{t}(f\star g)=(\mathsf{e}_{t}\ f)\star(\mathsf{e}_{t}\ g) and (πf)g=π(fπg)(\pi^{*}f)\star g=\pi^{*}(f\star\pi_{*}g) of the convolution. Thus

(25) β¯V𝜶¯Δ=V𝜶¯\β¯Δδp𝖾t0π(Ver(𝜶¯π(Δ);π(γΔ))π(𝖾t0P(𝜶¯Δ𝖼,τΔ))).\nabla_{\underline{\beta}}V_{{\underline{{\bm{\alpha}}}}}^{\Delta}=V_{{\underline{{\bm{\alpha}}}}\backslash\underline{\beta}}^{\Delta}-\delta_{p{\mathbb{N}}}\ \mathsf{e}_{t_{0}}\ \pi^{*}\big{(}{\textnormal{Ver}}_{({\underline{{\bm{\alpha}}}}^{\prime}_{\pi(\Delta)};\pi(\gamma_{\Delta}))}\ast\pi_{*}\big{(}\mathsf{e}_{-t_{0}}P_{({\underline{{\bm{\alpha}}}}_{\Delta}^{\mathsf{c}},\tau_{\Delta})}\big{)}\big{)}.

Multiplying the partition function P(𝜶¯Δ𝖼,τΔ)P_{({\underline{{\bm{\alpha}}}}_{\Delta}^{\mathsf{c}},\tau_{\Delta})} by 𝖾t01\mathsf{e}_{t_{0}^{-1}} amounts to replacing every (α,u)𝜶¯Δ𝖼(\alpha,u)\in{\underline{{\bm{\alpha}}}}_{\Delta}^{\mathsf{c}} with (α,t0αu)(\alpha,\,t_{0}^{-\alpha}u), while push-forward π\pi_{\ast} amounts to a restriction to 𝔱β{\mathfrak{t}}_{\beta}. Hence

π(𝖾t01P(𝜶¯Δ𝖼,τΔ))=P((𝜶¯π(Δ))𝖼;π(τΔ)).\pi_{\ast}\big{(}\mathsf{e}_{t_{0}^{-1}}\ P_{({\underline{{\bm{\alpha}}}}_{\Delta}^{\mathsf{c}},\tau_{\Delta})}\big{)}=P_{(({\underline{{\bm{\alpha}}}}^{\prime}_{\pi(\Delta)})^{\mathsf{c}};\pi(\tau_{\Delta}))}.

The inner product on 𝔱{\mathfrak{t}} restricts to an inner product on 𝔱β𝔱β{\mathfrak{t}}_{\beta}\simeq{\mathfrak{t}}_{\beta}^{\ast} satisfying π(γΔ)=π(γ)π(Δ)\pi(\gamma_{\Delta})=\pi(\gamma)_{\pi(\Delta)}, which implies that

τπ(Δ):=π(τΔ)=π(γ)π(Δ)π(γ)\tau_{\pi(\Delta)}:=\pi(\tau_{\Delta})=\pi(\gamma)_{\pi(\Delta)}-\pi(\gamma)

is the polarizing vector defined by π(γ)\pi(\gamma). We hence arrive at

(26) β¯V𝜶¯Δ=V𝜶¯\β¯Δ𝖾t0δpπV𝜶¯π(Δ),\nabla_{\underline{\beta}}V_{{\underline{{\bm{\alpha}}}}}^{\Delta}=V_{{\underline{{\bm{\alpha}}}}\backslash\underline{\beta}}^{\Delta}-\mathsf{e}_{t_{0}}\ \delta_{p{\mathbb{N}}}\ \pi^{\ast}V_{{\underline{{\bm{\alpha}}}}^{\prime}}^{\pi(\Delta)},

where the generic element used in defining V𝜶¯π(Δ)V_{{\underline{{\bm{\alpha}}}}^{\prime}}^{\pi(\Delta)} is π(γ)\pi(\gamma).

Having worked out β¯V𝜶¯Δ\nabla_{\underline{\beta}}V_{{\underline{{\bm{\alpha}}}}}^{\Delta} for the two cases, let us now take the sum over all admissible affine subspaces Δ𝒮(𝜶)\Delta\in{\mathcal{S}}({\bm{\alpha}}).

Observe that if Δ𝒮(𝜶\β)\Delta\not\in{\mathcal{S}}({\bm{\alpha}}\backslash\beta), then the term V𝜶¯\β¯ΔV_{{\underline{{\bm{\alpha}}}}\backslash\underline{\beta}}^{\Delta} is zero. Indeed, Δ𝒮(𝜶\β)\Delta\not\in{\mathcal{S}}({\bm{\alpha}}\backslash\beta) means that ann(𝔱Δ)\mathrm{ann}({\mathfrak{t}}_{\Delta}) is spanned by a subset of 𝜶{\bm{\alpha}}, but not by a subset of 𝜶β{\bm{\alpha}}{\setminus}\beta. As a consequence, the restriction of the Verlinde sum V𝜶Δ\βV_{{\bm{\alpha}}_{\Delta}\backslash\beta} to Δ\Delta is supported on a union of affine subspaces of codimension at least one in Δ\Delta, and consequently Ver(απ(Δ);γΔ)=0{\textnormal{Ver}}\big{(}\alpha^{\prime}_{\pi(\Delta)};\gamma_{\Delta}\big{)}=0, thus V𝜶¯\β¯Δ=0V_{{\underline{{\bm{\alpha}}}}\backslash\underline{\beta}}^{\Delta}=0.

On the other hand, the projection π:𝔱𝔱β\pi\colon{\mathfrak{t}}^{\ast}\to{\mathfrak{t}}_{\beta}^{\ast} gives a bijection from the set of Δ𝒮(𝜶)\Delta\in{\mathcal{S}}({\bm{\alpha}}) for which β𝜶Δ\beta\in{\bm{\alpha}}_{\Delta} (case (ii) above), with the set 𝒮(𝜶){\mathcal{S}}({\bm{\alpha}}^{\prime}) of admissible affine subspaces for the list 𝜶{\bm{\alpha}}^{\prime}. From equations (23), (26) we therefore obtain

(27) β¯Δ𝒮(𝜶)V𝜶¯Δ=Δ𝒮(𝜶\β)V𝜶¯\β¯Δ𝖾t0δpπΔ𝒮(𝜶)V𝜶¯Δ.\nabla_{\underline{\beta}}\ \sum_{\Delta\in{\mathcal{S}}({\bm{\alpha}})}V_{{\underline{{\bm{\alpha}}}}}^{\Delta}=\sum_{\Delta\in{\mathcal{S}}({\bm{\alpha}}\backslash\beta)}V_{{\underline{{\bm{\alpha}}}}\backslash\underline{\beta}}^{\Delta}-\mathsf{e}_{t_{0}}\ \delta_{p{\mathbb{N}}}\ \pi^{\ast}\sum_{\Delta^{\prime}\in{\mathcal{S}}({\bm{\alpha}}^{\prime})}V_{{\underline{{\bm{\alpha}}}}^{\prime}}^{\Delta^{\prime}}.

Applying the induction hypothesis to the right hand side, this is becomes

(28) β¯Δ𝒮(𝜶)V𝜶¯Δ=V𝜶¯\β¯𝖾t0δpπV𝜶¯.\nabla_{\underline{\beta}}\ \sum_{\Delta\in{\mathcal{S}}({\bm{\alpha}})}V_{{\underline{{\bm{\alpha}}}}}^{\Delta}=V_{{\underline{{\bm{\alpha}}}}\backslash\underline{\beta}}-\mathsf{e}_{t_{0}}\ \delta_{p{\mathbb{N}}}\ \pi^{\ast}V_{{\underline{{\bm{\alpha}}}}^{\prime}}.

Finally, by the difference equation (14) this equals β¯V𝜶¯\nabla_{\underline{\beta}}V_{{\underline{{\bm{\alpha}}}}}.

Step 2. Claim: V𝜶¯V_{{\underline{{\bm{\alpha}}}}} agrees with the sum ΔV𝜶¯Δ\sum_{\Delta}V_{{\underline{{\bm{\alpha}}}}}^{\Delta} on a set of representatives for the translation action of the lattice 𝜶Λ{\mathbb{Z}}{\bm{\alpha}}\subseteq\Lambda^{\ast} in Λ×\Lambda^{\ast}\times{\mathbb{N}}.

Recall that by Proposition (2.6), the Verlinde sum V𝜶¯(,)V_{{\underline{{\bm{\alpha}}}}}(\cdot,\ell) for fixed \ell\in{\mathbb{N}} is supported on the union of all Δ1\ell\Delta_{1}, for Δ1𝒮(𝜶)\Delta_{1}\in{\mathcal{S}}({\bm{\alpha}}) top-dimensional (i.e., a Ξ\Xi^{\ast}-translate of Δ0=span𝜶\Delta_{0}={\textnormal{span}}\,{\bm{\alpha}}). The same is true for each of the summands V𝜶¯ΔV_{{\underline{{\bm{\alpha}}}}}^{\Delta}. Since each Δ1\ell\Delta_{1} is invariant under the 𝜶{\mathbb{Z}}{\bm{\alpha}}-action, it suffices to show that the functions V𝜶(,)V_{{\bm{\alpha}}}(\cdot,\ell) and ΔV𝜶Δ(,)\sum_{\Delta}V^{\Delta}_{{\bm{\alpha}}}(\cdot,\ell) agree on a set of representatives for the 𝜶Λ{\mathbb{Z}}{\bm{\alpha}}\subseteq\Lambda^{\ast} action on Δ1\ell\Delta_{1}, for any given \ell and Δ1\Delta_{1}. The set

(29) Λ(γΔ1𝜶)\Lambda^{*}\cap\big{(}\ell\,\gamma_{\Delta_{1}}-\Box{\bm{\alpha}}\big{)}

contains such a set of representatives. By Remark 4.1, and the definition of the quasi-polynomial germ, V𝜶¯(,)V_{{\underline{{\bm{\alpha}}}}}(\cdot,\ell) and Ver(𝜶¯Δ1;γΔ1)(,){\textnormal{Ver}}_{({\underline{{\bm{\alpha}}}}_{\Delta_{1}};\gamma_{\Delta_{1}})}(\cdot,\ell) agree on (29). Since 𝜶Δ1𝖼={\bm{\alpha}}_{\Delta_{1}}^{\mathsf{c}}=\emptyset, the corresponding partition function P(𝜶¯Δ1𝖼,τΔ1)P_{({\underline{{\bm{\alpha}}}}_{\Delta_{1}}^{\mathsf{c}},\tau_{\Delta_{1}})} is simply δ0:Λ\delta_{0}\colon\Lambda^{*}\to{\mathbb{C}}. That is,

V𝜶¯(,)=V𝜶¯Δ1(,)V_{{\underline{{\bm{\alpha}}}}}(\cdot,\ell)=V_{{\underline{{\bm{\alpha}}}}}^{\Delta_{1}}(\cdot,\ell)

on (29). We claim that the supports of all other terms V𝜶¯Δ(,)V_{{\underline{{\bm{\alpha}}}}}^{\Delta}(\cdot,\ell) with ΔΔ1\Delta\neq\Delta_{1} do not meet (29). This is immediate when Δ\Delta is not contained in Δ1\Delta_{1}, thus suppose ΔΔ1\Delta\subseteq\Delta_{1}. Note that

(30) 𝗇:=γΔγΔ1=τΔτΔ1\mathsf{n}:=\gamma_{\Delta}-\gamma_{\Delta_{1}}=\tau_{\Delta}-\tau_{\Delta_{1}}

is a normal vector to Δ\Delta as an affine subspace of Δ1\Delta_{1}. Since τΔ1\tau_{\Delta_{1}} is normal to Δ1\Delta_{1}, it is orthogonal to all α𝜶\alpha\in{\bm{\alpha}}. In the defining the polarization of 𝜶Δ𝖼{\bm{\alpha}}_{\Delta}^{\mathsf{c}}, and hence in the description (22) of supp(V𝜶¯Δ(,))\mathrm{supp}(V_{{\underline{{\bm{\alpha}}}}}^{\Delta}(\cdot,\ell)), we may therefore replace τΔ\tau_{\Delta} with 𝗇\mathsf{n}. It follows that the elements λΛ\lambda\in\Lambda^{\ast} in the support of V𝜶¯Δ(,)V_{{\underline{{\bm{\alpha}}}}}^{\Delta}(\cdot,\ell) satisfy

λγΔ1,𝗇>λγΔ,𝗇σΔ,𝗇.\langle\lambda-\ell\gamma_{\Delta_{1}},\mathsf{n}\rangle>\langle\lambda-\ell\gamma_{\Delta},\mathsf{n}\rangle\geq\langle\sigma_{\Delta},\mathsf{n}\rangle.

On the other hand, if λγΔ1α𝜶[0,1]α\lambda\in\ell\gamma_{\Delta_{1}}-\sum_{\alpha\in{\bm{\alpha}}}[0,1]\alpha, then

λγΔ1,𝗇α(𝜶Δ𝖼)α,𝗇=σΔ,𝗇.\langle\lambda-\ell\gamma_{\Delta_{1}},\mathsf{n}\rangle\leq-\sum_{\alpha\in({\bm{\alpha}}^{\mathsf{c}}_{\Delta})_{-}}\langle\alpha,\mathsf{n}\rangle=\langle\sigma_{\Delta},\mathsf{n}\rangle.

Hence, these elements are never in the support of V𝜶¯Δ(,)V_{{\underline{{\bm{\alpha}}}}}^{\Delta}(\cdot,\ell) when Δ\Delta is properly contained in Δ1\Delta_{1}.

This concludes the argument for step 2, and completes the proof of the theorem. ∎

In the proof we used the following difference equation for the quasi-polynomial germs, which is obtained as a consequence of the difference equation for Verlinde sums in Proposition 2.8. Using the notation from that proposition we have:

Lemma 4.3.

Suppose β¯=(β,v)𝛂¯Δ\underline{\beta}=(\beta,v)\in{\underline{{\bm{\alpha}}}}_{\Delta}, and let pp\in{\mathbb{N}} be the smallest number such that there exists t0𝖳pt_{0}\in{\mathsf{T}}_{p} with vt0β=1v\,t_{0}^{-\beta}=1. Let 𝛂¯π(Δ){\underline{{\bm{\alpha}}}}^{\prime}_{\pi(\Delta)} be the list of augmented weights (π(α),ut0α)(\pi(\alpha),\ u\,t_{0}^{-\alpha}) for (α,u)𝛂Δ(\alpha,u)\in{\bm{\alpha}}_{\Delta} with α¯β¯\underline{\alpha}\neq\underline{\beta}. Then

Ver(𝜶¯Δβ¯;μ)=β¯Ver(𝜶¯Δ;μ)+𝖾t0δpπVer(𝜶¯π(Δ);π(μ)).{\textnormal{Ver}}_{({\underline{{\bm{\alpha}}}}_{\Delta}{\setminus}\underline{\beta};\mu)}=\nabla_{\underline{\beta}}{\textnormal{Ver}}_{({\underline{{\bm{\alpha}}}}_{\Delta};\mu)}+\mathsf{e}_{t_{0}}\ \delta_{p{\mathbb{N}}}\ \pi^{*}{\textnormal{Ver}}_{({\underline{{\bm{\alpha}}}}^{\prime}_{\pi(\Delta)};\pi(\mu))}.

Here π:ΛΛβ\pi\colon\Lambda^{*}\to\Lambda_{\beta}^{*} is the projection.

Proof.

On the set of all (λ,)(\lambda,\ell) such that λ𝔠\lambda\in\ell\mathfrak{c}, this formula coincides with the difference equation for Verlinde sums (Proposition 2.8, with 𝜶¯{\underline{{\bm{\alpha}}}} replaced by 𝜶¯Δ{\underline{{\bm{\alpha}}}}_{\Delta}). Since all terms in the deletion formula are quasi-polynomial on this set, the result follows. ∎

Example 4.4.

We illustrate the decomposition formulas for the Bernoulli series Bn(λ)B_{n}(\lambda) and the Verlinde sums Vn(λ,)V_{n}(\lambda,\ell), using the notation from Examples 2.5, 3.2. Take γ𝔠=(0,1)\gamma\in\mathfrak{c}=(0,1). The contribution from Δ0=𝔱\Delta_{0}={\mathfrak{t}}^{\ast} give the leading term

H𝜶Δ0(λ)=Bern(λ),V𝜶Δ0(λ,)=Vern(λ,).H^{\Delta_{0}}_{{\bm{\alpha}}}(\lambda)={\textnormal{Ber}}_{n}(\lambda),\ \ \ \ V^{\Delta_{0}}_{{\bm{\alpha}}}(\lambda,\ell)={\textnormal{Ver}}_{n}(\lambda,\ell).

The other contributions come from Δ={μ}\Delta=\{\mu\} with μΞ=\mu\in\Xi^{\ast}={\mathbb{Z}}. We have 𝜶Δ={\bm{\alpha}}_{\Delta}=\emptyset, hence

Ber(𝜶Δ,γΔ)(λ)=δμ(λ),Ver(𝜶Δ,γΔ)(λ,)=δμ(λ).{\textnormal{Ber}}_{({\bm{\alpha}}_{\Delta},\gamma_{\Delta})}(\lambda)=\delta_{\mu}(\lambda),\ \ {\textnormal{Ver}}_{({\bm{\alpha}}_{\Delta},\gamma_{\Delta})}(\lambda,\ell)=\ell\delta_{\ell\mu}(\lambda).

Since τΔ=μγ\tau_{\Delta}=\mu-\gamma, the weight α=+1\alpha=+1 is polarized if and only if μ>0\mu>0. With the partition functions from Example 3.2, this gives

H(𝜶Δ𝖼,τΔ)(λ)=Hn(λ),P(𝜶Δ𝖼,τΔ)(λ)=Pn(λ)H_{({\bm{\alpha}}^{\mathsf{c}}_{\Delta},\tau_{\Delta})}(\lambda)=H_{n}(\lambda),\ \ \ P_{({\bm{\alpha}}^{\mathsf{c}}_{\Delta},\tau_{\Delta})}(\lambda)=P_{n}(\lambda)

for Δ={μ}\Delta=\{\mu\} with μ>0\mu>0 while

H(𝜶Δ𝖼,τΔ)(λ)=(1)nHn(λ),P(𝜶Δ𝖼,τΔ)(λ)=(1)nPn(λn)H_{({\bm{\alpha}}^{\mathsf{c}}_{\Delta},\tau_{\Delta})}(\lambda)=(-1)^{n}H_{n}(-\lambda),\ \ \ P_{({\bm{\alpha}}^{\mathsf{c}}_{\Delta},\tau_{\Delta})}(\lambda)=(-1)^{n}P_{n}(-\lambda-n)

for Δ={μ}\Delta=\{\mu\} with μ0\mu\leq 0. The resulting decomposition formulas hence read as

Bn(λ)=Bern(λ)+n=1Hn(λμ)+(1)nμ=0Hn(μλ)B_{n}(\lambda)=\operatorname{Ber}_{n}(\lambda)+\sum_{n=1}^{\infty}H_{n}(\lambda-\mu)+(-1)^{n}\sum_{\mu=-\infty}^{0}H_{n}(\mu-\lambda)

and

Vn(λ,)=Vern(λ,)+μ=1Pn(λμ)+(1)nμ=0Pn(μλn),V_{n}(\lambda,\ell)={\textnormal{Ver}}_{n}(\lambda,\ell)+\ell\sum_{\mu=1}^{\infty}P_{n}(\lambda-\ell\mu)+(-1)^{n}\ell\sum_{\mu=-\infty}^{0}P_{n}(\ell\mu-\lambda-n),

respectively.

See also Example 6 further below, where we describe the decomposition formula for Example 2.3 in the case G=SU(3)G=SU(3).

5. Szenes’ Residue Theorem.

In this Section, we will review Szenes’ residue theorem for the Verlinde sums [14] (see also [15]). The formula is a powerful tool for explicit computations of such sums. We will explain how it implies the quasi-polynomial behaviour, in the form stated in Theorem 2.7.

5.1. The constant term functional.

The formula in [14] is expressed in terms of iterated ‘constant-term’ functionals. We describe these briefly here; for further background see [6, 16, 14, 19]. For a meromorphic function ff of a single variable zz, defined on a neighborhood of 00\in{\mathbb{C}}, the constant term CTz=0(f)\operatorname*{CT}_{z=0}(f) is defined to be the constant term in the Laurent expansion of f(z)f(z) about the point 0. In terms of residues,

(31) CTz=0(f)=Resz=0(f(z)dzz).\operatorname*{CT}_{z=0}(f)=\mathrm{Res}_{z=0}\Big{(}f(z)\frac{dz}{z}\Big{)}.

Let Λ𝔱\Lambda\subseteq{\mathfrak{t}} be a maximal rank lattice, and consider a Λ\Lambda-periodic affine hyperplane arrangement 𝒜{\mathcal{A}} in 𝔱{\mathfrak{t}}. For p𝔱p\in{\mathfrak{t}} we denote by 𝒜p𝒜{\mathcal{A}}_{p}\subseteq{\mathcal{A}} the collection of affine hyperplanes passing through pp. The point pp is called a vertex of the arrangement if the intersection of affine hyperplanes in 𝒜p{\mathcal{A}}_{p} is {p}\{p\}. Let vx(𝒜){\textnormal{vx}}({\mathcal{A}}) denote the set of all vertices. A subset of 𝒜p{\mathcal{A}}_{p} is called linearly independent if the conormal directions to the hyperplanes in this subset are linearly independent.

Let M𝒜M_{\mathcal{A}} be the ring of meromorphic functions on 𝔱{\mathfrak{t}}_{\mathbb{C}} whose singularities are contained in H𝒜H\bigcup_{H\in{\mathcal{A}}}H. Suppose pvx(𝒜)p\in{\textnormal{vx}}({\mathcal{A}}), and consider an ordered rr-tuple 𝑯=(H1,,Hr){\bm{H}}=(H_{1},\ldots,H_{r}) of r=dim𝔱r=\dim{\mathfrak{t}} linearly independent affine hyperplanes Hk𝒜pH_{k}\in{\mathcal{A}}_{p}. The iterated constant term functional

iCT𝑯:M𝒜\operatorname*{iCT}_{{\bm{H}}}\colon M_{\mathcal{A}}\to{\mathbb{C}}

is defined as follows: Choose affine-linear functions βk\beta_{k} with Hk=βk1(0)H_{k}=\beta_{k}^{-1}(0); the collection 𝜷=(β1,,βr){\bm{\beta}}=(\beta_{1},\ldots,\beta_{r}) defines an isomorphism 𝜷:𝔱r{\bm{\beta}}\colon{\mathfrak{t}}_{\mathbb{C}}\to{\mathbb{C}}^{r}, taking pp to 0. Thus f𝜷1f\circ{\bm{\beta}}^{-1} is a function of z1,,zrz_{1},\ldots,z_{r}. One defines

(32) iCT𝑯(f)=CTz1=0CTzr=0(f𝜷1).\operatorname*{iCT}_{{\bm{H}}}(f)=\operatorname*{CT}_{z_{1}=0}\ \cdots\operatorname*{CT}_{z_{r}=0}\,(f\circ{\bm{\beta}}^{-1}).

That is, one first computes CTzr=0(f𝜷1)\operatorname*{CT}_{z_{r}=0}(f\circ{\bm{\beta}}^{-1}) as a meromorphic function of z1,,zr1z_{1},\ldots,z_{r-1}, one next computes CTzr1=0\operatorname*{CT}_{z_{r-1}=0} of the result, and so forth. This definition is independent of the choice of βk\beta_{k} defining HkH_{k}, due to the fact that the constant term (31) is invariant under coordinate changes of the form zczz\leadsto cz. On the other hand, it does in general depend on the ordering of the hyperplanes. Note that the definition of iCT𝑯\operatorname*{iCT}_{{\bm{H}}} extends to the space M𝒜,pM_{{\mathcal{A}},p} of germs at pp of meromorphic functions with poles in 𝒜{\mathcal{A}}.

5.2. nbc bases.

Szenes’ Theorem 5.3 expresses the Verlinde sum V𝜶V_{{\bm{\alpha}}} as a finite sum over iterated residues, for suitable rr-tuples 𝑯{\bm{H}} of affine hyperplanes. Given pvx(𝒜)p\in{\textnormal{vx}}({\mathcal{A}}), choose an ordering on the set 𝒜p{\mathcal{A}}_{p}. This determines a unique collection BpB_{p} of rr-tuples 𝑯{\bm{H}} of linearly independent hyperplanes in 𝒜p{\mathcal{A}}_{p}, with the following property: For every H𝒜pH\in{\mathcal{A}}_{p}, the set

{H𝑯:H<H}H\{H^{\prime}\in{\bm{H}}\colon H<H^{\prime}\}\cup H

is linearly independent. The collection BpB_{p} is known in the theory of hyperplane arrangements as an nbc basis of 𝒜p{\mathcal{A}}_{p}, where nbc stands for no broken circuit basis (cf.  [14, 16]).

Example 5.1.

If dim(𝔱)=2\dim({\mathfrak{t}})=2, the nbc basis BpB_{p} for a given ordering of 𝒜p{\mathcal{A}}_{p} is the collection of all 2-tuples 𝑯=(H0,H){\bm{H}}=(H_{0},H), where H0H_{0} is the unique smallest element of 𝒜p{\mathcal{A}}_{p}, and HH is any of the other elements.

Remark 5.2.
  1. (a)

    In [6], it is shown that BpB_{p} corresponds naturally to a vector space basis for the subspace of simple fractions, in the ring of rational functions generated by the inverses of the affine-linear functions defining the hyperplanes in 𝒜p{\mathcal{A}}_{p}.

  2. (b)

    In Szenes’ formula one can also use more general orthogonal bases, which need not arise from a linear ordering of 𝒜p{\mathcal{A}}_{p}, see [14]. For simplicity we have only described nbc bases, which are a special case.

  3. (c)

    For interesting and non-trivial examples, one might consult [4] where the case of hyperplane arrangements arising from root systems of the classical Lie algebras are discussed in detail, as well as many computations of the associated multiple Bernoulli series.

5.3. Statement of Szenes’ theorem.

For the remainder of this section, we assume that Ξ𝔱\Xi\subseteq{\mathfrak{t}} is a lattice containing Λ\Lambda, and that 𝜶¯=(𝜶,𝒖){\underline{{\bm{\alpha}}}}=({\bm{\alpha}},{\bm{u}}) is a list of augmented weights, where the weights α𝜶\alpha\in{\bm{\alpha}} are all non-zero. Let 𝒜=𝒜(𝜶¯){\mathcal{A}}={\mathcal{A}}({\underline{{\bm{\alpha}}}}) be the resulting affine hyperplane arrangement in 𝔱{\mathfrak{t}}, consisting of the affine hyperplanes α1(s)\alpha^{-1}(s) such that ss\in{\mathbb{Q}} with (α,e2πis)=(α,u)𝜶¯(\alpha,e^{2\pi{\mathrm{i}}s})=(\alpha,u)\in{\underline{{\bm{\alpha}}}}. Using the procedure from Subsection 2.3 (d), we will assume that the α𝜶\alpha\in{\bm{\alpha}} are primitive vectors in the weight lattice. Note that this procedure does not change the arrangement 𝒜(𝜶¯){\mathcal{A}}({\underline{{\bm{\alpha}}}}).

We will also assume span𝜶=𝔱{\textnormal{span}}_{{\mathbb{R}}}{\bm{\alpha}}={\mathfrak{t}}^{\ast}, so that the set vx(𝒜){\textnormal{vx}}({\mathcal{A}}) of vertices is non-empty. For every vertex pvx(𝒜)p\in{\textnormal{vx}}({\mathcal{A}}), choose an ordering of the set 𝒜p{\mathcal{A}}_{p}, and let BpB_{p} be the resulting nbc basis, as in Subsection 5.2. Recall also the set 𝒮=𝒮(𝜶){\mathcal{S}}={\mathcal{S}}({\bm{\alpha}}) of admissible affine subspaces Δ𝔱\Delta\subseteq{\mathfrak{t}}^{\ast}, and that a chamber is a component of 𝔱Δ𝔱Δ{\mathfrak{t}}^{\ast}-\bigcup_{\Delta\neq{\mathfrak{t}}^{\ast}}\Delta (cf. Definition 2.2).

For any chamber 𝔠𝔱{\mathfrak{c}}\subseteq{\mathfrak{t}}^{\ast}, and any 𝑯Bp{\bm{H}}\in B_{p} and (λ,)Λ×(\lambda,\ell)\in\Lambda^{\ast}\times{\mathbb{N}}, define a meromorphic function

XT𝔠,𝑯(λ,)(X)X\mapsto T_{{\mathfrak{c}},{\bm{H}}}(\lambda,\ell)(X)

of X𝔱X\in{\mathfrak{t}}_{\mathbb{C}}, as follows. Choose affine-linear functions β:𝔱\beta\colon{\mathfrak{t}}\to{\mathbb{R}} defining the hyperplanes H𝑯H\in{\bm{H}}, in such a way that their linear parts β0=ββ(0)\beta^{0}=\beta-\beta(0) lie in ΞΛ\Xi^{\ast}\subseteq\Lambda^{\ast}. Let 𝜷{\bm{\beta}} be the list of these functions; the list of their linear parts forms a basis 𝜷0{\bm{\beta}}^{0} of 𝔱{\mathfrak{t}}^{\ast}. Let volΞ(𝜷0){\textnormal{vol}}_{\Xi^{\ast}}(\Box{\bm{\beta}}^{0}) be the volume of the zonotope with respect to Lebesgue measure on 𝔱{\mathfrak{t}}^{\ast} (normalized such that 𝔱/Ξ{\mathfrak{t}}^{\ast}/\Xi^{\ast} has volume 11). Equivalently,

volΞ(𝜷0)=#(Ξ(ν+𝜷0)){\textnormal{vol}}_{\Xi^{\ast}}(\Box{\bm{\beta}}^{0})=\#\big{(}\Xi^{\ast}\cap(\nu+\Box{\bm{\beta}}^{0})\big{)}

for generic ν𝔱\nu\in{\mathfrak{t}}^{*}. (The intersection on the right hand side does not change as ν\nu varies in a chamber of 𝒮(𝜷0){\mathcal{S}}({\bm{\beta}}^{0}); in particular, we may also write it as Ξ(𝔠+𝜷0)\Xi^{\ast}\cap({\mathfrak{c}}+\Box{\bm{\beta}}^{0}).) With these preparations, put

(33) T𝔠,𝑯(λ,)(X)=1volΞ(𝜷0)μe2πiλμ,Xβ𝜷2πiβ(X)1e2πiβ0,X;T_{{\mathfrak{c}},{\bm{H}}}(\lambda,\ell)(X)=\frac{1}{{\textnormal{vol}}_{\Xi^{\ast}}(\Box{\bm{\beta}}^{0})}\sum_{\mu}e^{2\pi{\mathrm{i}}\langle\lambda-\ell\mu,X\rangle}\prod_{\beta\in{\bm{\beta}}}\frac{2\pi{\mathrm{i}}\ell\beta(X)}{1-e^{2\pi{\mathrm{i}}\ell\langle\beta^{0},X\rangle}};

here the sum is over all μΞ(𝔠𝜷0)\mu\in\Xi^{\ast}\cap({\mathfrak{c}}-\Box{\bm{\beta}}^{0}). As shown by Szenes [14], this function does not depend on the choice of affine-linear functions β\beta defining HH.

Observe that the function (33) is holomorphic for XX near pp. Indeed, the singularities of X(1e2πiβ0,X)1X\mapsto(1-e^{2\pi{\mathrm{i}}\ell\langle\beta^{0},X\rangle})^{-1} are contained in the subset where β0,X\ell\langle\beta^{0},X\rangle\in{\mathbb{Z}}; but for X=pX=p, any such singularity is compensated by the zero of Xβ(X)X\mapsto\beta(X) in the enumerator. Let f𝜶¯M𝒜f_{{\underline{{\bm{\alpha}}}}}\in M_{\mathcal{A}} be the meromorphic function defined by

(34) f𝜶¯(X)=(α,u)𝜶¯(1ue2πiα,X)1.f_{{\underline{{\bm{\alpha}}}}}(X)=\prod_{(\alpha,u)\in{\underline{{\bm{\alpha}}}}}\Big{(}1-u\,e^{-2\pi{\mathrm{i}}\langle\alpha,X\rangle}\Big{)}^{-1}.

We are now in position to state Szenes’ theorem for Verlinde sums:

Theorem 5.3 (Szenes [14, Theorem 4.2]).

Let 𝛂¯{\underline{{\bm{\alpha}}}} be a list of augmented weights, where all α𝛂\alpha\in{\bm{\alpha}} are primitive lattice vectors in Λ\Lambda^{\ast}, such that span𝛂=𝔱{\textnormal{span}}_{{\mathbb{R}}}{\bm{\alpha}}={\mathfrak{t}}^{\ast}. Define 𝒜=𝒜(𝛂¯){\mathcal{A}}={\mathcal{A}}({\underline{{\bm{\alpha}}}}) as above. Then for every chamber 𝔠\mathfrak{c} of 𝒮(𝛂){\mathcal{S}}({\bm{\alpha}}), and all (λ,)Λ×(\lambda,\ell)\in\Lambda^{*}\times{\mathbb{N}} with λ𝔠𝛂\lambda\in\ell{\mathfrak{c}}-\Box{\bm{\alpha}}, the Verlinde sum is given by the formula

(35) V𝜶¯(λ,)=[p]vx(𝒜)/Λ𝑯BpiCT𝑯(T𝔠,𝑯(λ,)f𝜶¯).V_{{\underline{{\bm{\alpha}}}}}(\lambda,\ell)=\sum_{[p]\in{\textnormal{vx}}({\mathcal{A}})/\Lambda}\,\,\,\,\sum_{{\bm{H}}\in B_{p}}\operatorname*{iCT}_{{\bm{H}}}\Big{(}T_{{\mathfrak{c}},{\bm{H}}}(\lambda,\ell)f_{{\underline{{\bm{\alpha}}}}}\Big{)}.

Here the summation over [p][p] uses one representative pvx(𝒜)p\in{\textnormal{vx}}({\mathcal{A}}) from each equivalence class modulo Λ\Lambda.

Example 5.4.

In Example 2.5, let 𝔱=𝔱={\mathfrak{t}}={\mathfrak{t}}^{\ast}={\mathbb{R}} with pairing given by multiplication, and let Λ=Ξ=\Lambda=\Xi={\mathbb{Z}}. Let 𝜶={1,,1}{\bm{\alpha}}=\{1,\ldots,1\} be the weight 1Λ=1\in\Lambda^{\ast}={\mathbb{Z}}, repeated nn times, defining the Verlinde sum V𝜶=VnV_{{\bm{\alpha}}}=V_{n}. Thus f𝜶(X)=(1e2πiX)nf_{{\bm{\alpha}}}(X)=\big{(}1-e^{-2\pi{\mathrm{i}}X}\big{)}^{-n}, and 𝜶=(0,n)\Box{\bm{\alpha}}=(0,n). The affine hyperplane arrangement 𝒜{\mathcal{A}} is the set {\mathbb{Z}} of integers, and has a single vertex (up to translation) p=0p=0. The basis BpB_{p} consists of just one element 𝑯{\bm{H}}, which is given by the single hyperplane H={0}H=\{0\}. For the chamber 𝔠=(0,1)\mathfrak{c}=(0,1), we obtain

T𝔠,𝑯(λ,)(X)=e2πiλX2πiX1e2πiX.T_{{\mathfrak{c}},{\bm{H}}}(\lambda,\ell)(X)=e^{2\pi{\mathrm{i}}\lambda X}\frac{2\pi{\mathrm{i}}\ell X}{1-e^{2\pi{\mathrm{i}}\ell X}}.

Writing z=2πiXz=2\pi{\mathrm{i}}X, Szenes’ formula hence shows that Vn(λ,)=Vern(λ,)V_{n}(\lambda,\ell)=\operatorname{Ver}_{n}(\lambda,\ell) for all n1n\geq 1 and all λ,\lambda,\ell such that λ(n,)\lambda\in(-n,\ell), with

(36) Vern(λ,)=CTz=0(eλzz1ez1(1ez)n).\operatorname{Ver}_{n}(\lambda,\ell)=\operatorname*{CT}_{z=0}\Big{(}e^{\lambda z}\frac{\ell z}{1-e^{\ell z}}\ \ \frac{1}{(1-e^{-z})^{n}}\Big{)}.

See [23, Theorem 4] for a direct proof of this formula, which may be reformulated as the generating function (3) stated in the introduction. In a similar way, Szenes’ formula for Bernoulli series reduces to the fact that Bn(λ)B_{n}(\lambda) agrees with Bern(λ)\operatorname{Ber}_{n}(\lambda) for 0<λ<10<\lambda<1, where Bern(λ)\operatorname{Ber}_{n}(\lambda) is given by the generating series (4).

A more elaborate example will be given in Section 6 below.

5.4. Relation between Bernoulli series and Verlinde sums

We briefly describe a Khovanskii-Pukhlikov-type formula, also due to Szenes ([17, Proposition 4.9] with somewhat different notation), relating Verlinde sums to Bernoulli series. See also [23, Theorem 5] for discussion of the 1-dimensional case.

We first need the analogue of Bernoulli series centred at vertices p0p\neq 0. Let pΛ𝔱p\in\Lambda\otimes{\mathbb{Q}}\subseteq\mathfrak{t} and 𝜶=(α1,,αn){\bm{\alpha}}=(\alpha_{1},...,\alpha_{n}) a list of weights. Let 𝜶^=(α^1,,α^n)\widehat{{\bm{\alpha}}}=(\widehat{\alpha}_{1},...,\widehat{\alpha}_{n}) be the list of affine linear functions vanishing at pp, with linear parts given by 𝜶{\bm{\alpha}}. Define the Bernoulli-like sum

B𝜶^(λ,)=ξ1Ξe2πiλ,ξα^𝜶^2πiα^(ξ),B_{\widehat{{\bm{\alpha}}}}(\lambda,\ell)=\sideset{}{{}^{\prime}}{\sum}_{\xi\in\tfrac{1}{\ell}\Xi}\frac{e^{2\pi{\mathrm{i}}\langle\lambda,\xi\rangle}}{\prod_{\widehat{\alpha}\in\widehat{{\bm{\alpha}}}}2\pi{\mathrm{i}}\widehat{\alpha}(\xi)},

where the prime next to the summation means to omit terms such that the denominator vanishes. If =1\ell=1 and pΞp\in\Xi, then this is related to the Bernoulli series B𝜶B_{{\bm{\alpha}}} by a multiplicative factor. Szenes [16, 17] gave a version of Theorem 5.3 for the Bernoulli series B𝜶^B_{\widehat{{\bm{\alpha}}}}; the result has the same form, but with f𝜶¯(X)f_{{\underline{{\bm{\alpha}}}}}(X) replaced by α𝜶^(2πiα^(X))1\prod_{\alpha\in\widehat{{\bm{\alpha}}}}(2\pi{\mathrm{i}}\widehat{\alpha}(X))^{-1}.

For α𝔱\alpha\in\mathfrak{t}^{\ast}, let α()=α(λ):Pol(𝔱)Pol(𝔱)\alpha(\partial)=\alpha(\partial_{\lambda})\colon{\textnormal{Pol}}(\mathfrak{t}^{\ast})\rightarrow{\textnormal{Pol}}(\mathfrak{t}^{\ast}) denote the directional derivative in the direction α\alpha, a first-order linear differential operator acting on the space of polynomial functions in 𝔱\mathfrak{t}^{\ast}. The assignment αα()\alpha\mapsto\alpha(\partial) extends multiplicatively to a map from the completion of the symmetric algebra Sym(𝔱){\textnormal{Sym}}(\mathfrak{t}^{\ast}) to infinite-order differential operators, acting on Pol(𝔱){\textnormal{Pol}}(\mathfrak{t}^{\ast}). In other words, to each formal power series 𝖯(X)\mathsf{P}(X) in X𝔱X\in\mathfrak{t}, we assign an infinite-order differential operator 𝖯()\mathsf{P}(\partial). More generally, if 𝖯(X)\mathsf{P}(X) is the Taylor expansion of a function at p𝔱p\in\mathfrak{t}, we can define an infinite-order differential operator 𝖯()\mathsf{P}(\partial), acting on the Pol(𝔱){\textnormal{Pol}}(\mathfrak{t}^{\ast})-module eλ,pPol(𝔱)e^{\langle\lambda,p\rangle}\cdot{\textnormal{Pol}}(\mathfrak{t}^{\ast}). The equation

α(λp)eλ,pg=eλ,pα(λ)g,α𝔱,\alpha(\partial_{\lambda}-p)e^{\langle\lambda,p\rangle}g=e^{\langle\lambda,p\rangle}\alpha(\partial_{\lambda})g,\qquad\alpha\in\mathfrak{t}^{\ast},

implies that only finitely many terms of the infinite series 𝖯()\mathsf{P}(\partial) act non-trivially on any fixed element of this module.

Let 𝜶¯{\underline{{\bm{\alpha}}}} be a list of augmented weights. Choose representatives pΛp\in\Lambda\otimes{\mathbb{Q}} for vx(𝒜)/Λ{\textnormal{vx}}(\mathcal{A})/\Lambda. Let 𝜶¯p{\underline{{\bm{\alpha}}}}_{p} be the sublist of 𝜶¯{\underline{{\bm{\alpha}}}} consisting of the α¯=(α,u){\underline{\alpha}}=(\alpha,u) such that ue2πiα,p=1ue^{-2\pi{\mathrm{i}}\langle\alpha,p\rangle}=1. For each α¯=(α,u){\underline{\alpha}}=(\alpha,u) in 𝜶¯p{\underline{{\bm{\alpha}}}}_{p}, let α^\widehat{\alpha} be the unique affine linear function vanishing at pp and with linear part α\alpha. Denote the corresponding list of affine linear functions by 𝜶^p\widehat{{\bm{\alpha}}}_{p}.

Let Td𝜶¯,p(X){\textnormal{Td}}_{{\underline{{\bm{\alpha}}}},p}(X) denote the Taylor series at pp of

α¯𝜶¯p2πiα^(X)1e2πiα^(X).\prod_{{\underline{\alpha}}\in{\underline{{\bm{\alpha}}}}_{p}}\frac{2\pi{\mathrm{i}}\widehat{\alpha}(X)}{1-e^{-2\pi{\mathrm{i}}\widehat{\alpha}(X)}}.

The function

f𝜶¯p𝖼(X)=(α,u)𝜶¯p𝖼(1ue2πiα,X)1f_{{\underline{{\bm{\alpha}}}}_{p}^{\mathsf{c}}}(X)=\prod_{(\alpha,u)\in{\underline{{\bm{\alpha}}}}_{p}^{\mathsf{c}}}\big{(}1-ue^{-2\pi{\mathrm{i}}\langle\alpha,X\rangle}\big{)}^{-1}

is holomorphic on a neighborhood of pp, and we will use the same symbol f𝜶¯p𝖼(X)f_{{\underline{{\bm{\alpha}}}}_{p}^{\mathsf{c}}}(X) to denote its Taylor series at this point. By differentiating inside the constant term of the residue formula for Bernoulli series and comparing with the residue formula for Verlinde sums, one has

(37) Ver(𝜶¯;γ)(λ,)=pvx(𝒜)/ΛTd𝜶¯,p((2πi)1)f𝜶¯p𝖼((2πi)1)Ber(𝜶^p;γ)(λ,).{\textnormal{Ver}}_{({\underline{{\bm{\alpha}}}};\gamma)}(\lambda,\ell)=\sum_{p\in{\textnormal{vx}}(\mathcal{A})/\Lambda}{\textnormal{Td}}_{{\underline{{\bm{\alpha}}}},p}\big{(}(2\pi{\mathrm{i}})^{-1}\partial\big{)}f_{{\underline{{\bm{\alpha}}}}_{p}^{\mathsf{c}}}\big{(}(2\pi{\mathrm{i}})^{-1}\partial\big{)}{\textnormal{Ber}}_{(\widehat{{\bm{\alpha}}}_{p};\gamma)}(\lambda,\ell).

In case there is only a single vertex at p=0p=0, equation (37) simplifies:

(38) nVer(𝜶;γ)(λ,)=(Td𝜶((2πi)1)Ber(𝜶;γ))(λ/),\ell^{-n}{\textnormal{Ver}}_{({\bm{\alpha}};\gamma)}(\lambda,\ell)=\Big{(}{\textnormal{Td}}_{{\bm{\alpha}}}\big{(}(2\pi{\mathrm{i}}\ell)^{-1}\partial\big{)}{\textnormal{Ber}}_{({\bm{\alpha}};\gamma)}\Big{)}(\lambda/\ell),

where nn is number of elements in the list 𝜶{\bm{\alpha}}. Equation (11) follows from (38). As a simple example of (38), consider Ber2(λ){\textnormal{Ber}}_{2}(\lambda), Ver2(λ,){\textnormal{Ver}}_{2}(\lambda,\ell) introduced in Example 2.5. One has

(1+12λ+11222λ2)2Ber2(λ)=12λ2+12λ112λ+125122.\Big{(}1+\frac{1}{2\ell}\frac{\partial}{\partial\lambda}+\frac{1}{12\ell^{2}}\frac{\partial^{2}}{\partial\lambda^{2}}\Big{)}^{2}{\textnormal{Ber}}_{2}(\lambda)=-\frac{1}{2}\lambda^{2}+\frac{1}{2}\lambda-\frac{1}{12}-\frac{\lambda}{\ell}+\frac{1}{2\ell}-\frac{5}{12\ell^{2}}.

Making the substitution λλ/\lambda\leadsto\lambda/\ell, and then multiplying by 2\ell^{2}, one arrives at the formula for Ver2(λ,){\textnormal{Ver}}_{2}(\lambda,\ell).

5.5. Proof of Theorem 2.7


We now explain how to deduce the quasi-polynomial behavior (Theorem 2.7) from Szenes’ Theorem 5.3. Using Proposition 2.6, and since #(𝖳𝖳Δ0)\#({\mathsf{T}}_{\ell}\cap{\mathsf{T}}_{\Delta_{0}}) is a polynomial in \ell, we may assume that span𝜶=𝔱{\textnormal{span}}_{{\mathbb{R}}}{\bm{\alpha}}={\mathfrak{t}}^{*}. Furthermore, by 2.3 (d), we may assume that the linear parts α\alpha are primitive lattice vectors in Λ\Lambda^{\ast}. We will show that each of the terms

(39) iCT𝑯(T𝔠,𝑯(λ,)f𝜶¯)\operatorname*{iCT}_{{\bm{H}}}\big{(}T_{{\mathfrak{c}},{\bm{H}}}(\lambda,\ell)f_{{\underline{{\bm{\alpha}}}}}\big{)}

with T𝔠,𝑯(λ,)(X)T_{{\mathfrak{c}},{\bm{H}}}(\lambda,\ell)(X) given by (33) and f𝜶¯(X)f_{{\underline{{\bm{\alpha}}}}}(X) given by (34), is quasi-polynomial in λ,\lambda,\ell. Write 𝜷=(β1,,βr){\bm{\beta}}=(\beta_{1},\ldots,\beta_{r}). We have to compute the iterated constant term with respect to the coordinates zk=2πiβk(X)z_{k}=2\pi{\mathrm{i}}\beta_{k}(X). We have that

e2πiβk0,X=vk1ezk,e^{2\pi{\mathrm{i}}\langle\beta_{k}^{0},X\rangle}=v_{k}^{-1}e^{z_{k}},

with the phase factor vk=e2πiβk(0)v_{k}=e^{2\pi{\mathrm{i}}\beta_{k}(0)}. The term

β𝜷2πiβ(X)1e2πiβ0,X=k=1rzk1vkezk\prod_{\beta\in{\bm{\beta}}}\frac{2\pi{\mathrm{i}}\ell\beta(X)}{1-e^{2\pi{\mathrm{i}}\ell\langle\beta^{0},X\rangle}}=\prod_{k=1}^{r}\frac{\ell z_{k}}{1-v_{k}^{-\ell}e^{\ell z_{k}}}

in (39) is holomorphic in z1,,zrz_{1},\ldots,z_{r}, and the coefficients of its power series expansion in the zkz_{k} are quasi-polynomial functions of \ell. (Here we are using that the map vk\ell\mapsto v_{k}^{\ell} is periodic in \ell, since βk(0)\beta_{k}(0) are rational.) Next, letting λk\lambda_{k} be the components of λ\lambda with respect to the basis βk0\beta^{0}_{k}, and similarly for μ\mu, we see that

e2πiλμ,X=(e2πiμ,p)tpλk=1re(λkμk)zk,e^{2\pi{\mathrm{i}}\langle\lambda-\ell\mu,X\rangle}=(e^{-2\pi{\mathrm{i}}\langle\mu,p\rangle})^{\ell}\ t_{p}^{\lambda}\prod_{k=1}^{r}e^{(\lambda_{k}-\ell\mu_{k})z_{k}},

with tp=exp(p)𝖳t_{p}=\exp(p)\in{\mathsf{T}}. The phase factor (e2πiμ,p)tpλ(e^{-2\pi{\mathrm{i}}\langle\mu,p\rangle})^{\ell}\ t_{p}^{\lambda}, is quasi-polynomial in λ,\lambda,\ell, while k=1re(λkμk)zk\prod_{k=1}^{r}e^{(\lambda_{k}-\ell\mu_{k})z_{k}} is a holomorphic function of z1,,zrz_{1},\ldots,z_{r} whose power series coefficients are polynomial in λ,\lambda,\ell. In summary, T𝔠,𝑯(λ,)𝜷1T_{\mathfrak{c},{\bm{H}}}(\lambda,\ell)\circ{\bm{\beta}}^{-1}, is a meromorphic function of z1,,zrz_{1},\ldots,z_{r} whose Laurent series coefficients are quasipolynomial in λ,\lambda,\ell. Finally, f𝜶¯𝜷1f_{{\underline{{\bm{\alpha}}}}}\circ{\bm{\beta}}^{-1} is a meromorphic function of the z1,,zrz_{1},\ldots,z_{r} that does not depend on λ,\lambda,\ell. Hence, (39) is a quasi-polynomial function of λ,\lambda,\ell.

Remark 5.5.

If 𝜶¯=(𝜶,𝒖){\underline{{\bm{\alpha}}}}=({\bm{\alpha}},{\bm{u}}) is such that all u=1u=1, and 𝜶{\bm{\alpha}} is unimodular (i.e., any vector space basis consisting of elements of 𝜶{\bm{\alpha}} is also a basis of Λ\Lambda^{*}), then there is only the zero vertex p=0p=0, up to Λ\Lambda-translation. Thus all vv’s are equal to 11 as well, and the argument above shows that V𝜶V_{{\bm{\alpha}}} is not only quasi-polynomial on the region (12), but is in fact as a polynomial on that region.

6. Example: Verlinde sum associated to G=SU(3)G=\operatorname{SU}(3)

In this Section, we work out the terms of the decomposition formula for the case of G=SU(3)G=\mathrm{SU}(3), continuing Example 2.3. For a wealth of examples of this type in the case of Bernoulli series, see [4]; for calculations of generalized Heaviside functions and partition functions associated to root systems, see [3].

We begin with a general compact, simple, simply connected Lie group GG. We fix the standard notation: 𝖳{\mathsf{T}} is a maximal torus, WW the corresponding Weyl group, Λ𝔱\Lambda\subseteq{\mathfrak{t}} is the integral (i.e., root) lattice, Ξ=Λ\Xi=\Lambda^{*} is the weight lattice (using the basic inner product to identify 𝔱{\mathfrak{t}} and 𝔱{\mathfrak{t}}^{\ast}). We denote by 𝔱+{\mathfrak{t}}_{+} the closed positive Weyl chamber, corresponding to the set +\mathfrak{R}_{+} of positive roots, and let 𝔱=𝔱+{\mathfrak{t}}_{-}=-{\mathfrak{t}}_{+} and =+\mathfrak{R}_{-}=-\mathfrak{R}_{+}. We are interested in the Verlinde sum V𝜶V_{{\bm{\alpha}}} for 𝜶={\bm{\alpha}}=\mathfrak{R}_{-}.

Let ZTZ\subseteq T be the center of GG. The function ttαt\mapsto t^{\alpha} defined by roots α\alpha descends to T/ZT/Z; hence the Verlinde sum can be written as a sum

V𝜶(λ,)=(ζZζλ)[t]T/Ztλ𝜶1tα.V_{{\bm{\alpha}}}(\lambda,\ell)=\Big{(}\sum_{\zeta\in Z}\zeta^{\lambda}\Big{)}\sum_{[t]\in T_{\ell}/Z}\frac{t^{\lambda}}{\prod_{{\bm{\alpha}}}1-t^{-\alpha}}.

using a representative tt for each equivalence class [t][t]. The sum ζZζλ\sum_{\zeta\in Z}\zeta^{\lambda} vanishes unless λ\lambda is in the root lattice, in which case it is equal to #Z\#Z. In particular, we see that V𝜶(,)V_{{\bm{\alpha}}}(\cdot,\ell) is supported on the root lattice. if GG is simply laced, then the root lattice is equal to Λ\Lambda, and the prefactor is #ZδΛ(λ)\#Z\ \delta_{\Lambda}(\lambda).

We now specialize to the case G=SU(3)G=\operatorname{SU}(3). Here Λ\Lambda has a basis α1,α2\alpha_{1},\alpha_{2} given by the simple roots, while Λ\Lambda^{*} has the dual basis ϖ1,ϖ2\varpi_{1},\varpi_{2} given by the fundamental weights. It will be convenient to denote by α3=α1+α2\alpha_{3}=\alpha_{1}+\alpha_{2} the highest root, so that 𝜶={α1,α2,α3}{\bm{\alpha}}=\{-\alpha_{1},-\alpha_{2},-\alpha_{3}\}.

The collection 𝒮{\mathcal{S}} of admissible subspaces consists of subspaces spanned by subsets of roots, together with their translates under the integral (i.e., root) lattice Ξ=Λ\Xi^{*}=\Lambda. Take the open chamber 𝔠\mathfrak{c} to be the the interior of the triangle spanned by 0,α2,α30,\alpha_{2},\alpha_{3}. An element

γ𝔠\gamma\in\mathfrak{c}

satisfies the genericity assumption from Section 4 if and only if it lies in the interior of an alcove from the Stiefel diagram; for convenience, we will will take it to be in the interior of the fundamental Weyl alcove (the triangle spanned by 0,ϖ1,ϖ20,\varpi_{1},\varpi_{2}).

The decomposition formula is a sum V𝜶=ΔV𝜶ΔV_{{\bm{\alpha}}}=\sum_{\Delta}V_{{\bm{\alpha}}}^{\Delta} over admissible subspaces, where each V𝜶ΔV_{{\bm{\alpha}}}^{\Delta} is a convolution of a polynomial Ver(αΔ,γΔ){\textnormal{Ver}}_{(\alpha_{\Delta},\gamma_{\Delta})} and a partition function. We will discuss these terms separately, according to the dimension of Δ\Delta.

6.1. Contribution from Δ=𝔱\Delta={\mathfrak{t}}^{*}

The leading term in the decomposition formula for V𝜶V_{{\bm{\alpha}}} is the polynomial expression Ver(𝜶;γ){\textnormal{Ver}}_{({\bm{\alpha}};\gamma)} associated to the admissible subspace Δ=𝔱\Delta=\mathfrak{t}^{\ast}. To compute Ver(𝜶;γ){\textnormal{Ver}}_{({\bm{\alpha}};\gamma)}, we will use the Szenes formula.

The affine hyperplane arrangement 𝒜{\mathcal{A}} is given by the set of affine root hyperplanes (the Stiefel diagram). Up to Λ\Lambda-translation, it has three vertices given by the vertices p=0,ϖ1,ϖ2p=0,\varpi_{1},\varpi_{2} of the Weyl alcove.

Let p{0,ϖ1,ϖ2}p\in\{0,\varpi_{1},\varpi_{2}\}. Using the ordering of root hyperplanes Hα1<Hα2<Hα3H_{\alpha_{1}}<H_{\alpha_{2}}<H_{\alpha_{3}}, we obtain an nbc basis Bp={𝑯p,1,𝑯p,2}B_{p}=\{{\bm{H}}_{p,1},{\bm{H}}_{p,2}\} where 𝑯p,1,𝑯p,2{\bm{H}}_{p,1},{\bm{H}}_{p,2} are defined by the linear functionals

𝜷p,1={α1c1,α2c2},𝜷p,2={α1c1,α3c3},{\bm{\beta}}_{p,1}=\{\alpha_{1}-c_{1},\alpha_{2}-c_{2}\},\hskip 28.45274pt{\bm{\beta}}_{p,2}=\{\alpha_{1}-c_{1},\alpha_{3}-c_{3}\},

respectively, with ck=αk,p{0,1}c_{k}=\langle\alpha_{k},p\rangle\in\{0,1\}. Explicitly (c1,c2,c3)=(0,0,0)(c_{1},c_{2},c_{3})=(0,0,0) for p=0p=0, (1,0,1)(1,0,1) for p=ϖ1p=\varpi_{1}, and (0,1,1)(0,1,1) for p=ϖ2p=\varpi_{2}. Note that Ξ=Λ\Xi^{\ast}=\Lambda, and

(𝔠𝜷p,10)Ξ={0},(𝔠𝜷p,20)Ξ={α1};(\mathfrak{c}-\Box{\bm{\beta}}_{p,1}^{0})\cap\Xi^{\ast}=\{0\},\ \ (\mathfrak{c}-\Box{\bm{\beta}}_{p,2}^{0})\cap\Xi^{\ast}=\{-\alpha_{1}\};

in particular volΞ(𝜷p,k0)=1{\textnormal{vol}}_{\Xi^{\ast}}(\Box{\bm{\beta}}^{0}_{p,k})=1 for k=1,2k=1,2. The Szenes formula for V𝜶V_{{\bm{\alpha}}} reads as

V𝜶=piCT𝑯p,1(Tp,1f𝜶)+iCT𝑯p,2(Tp,2f𝜶)V_{{\bm{\alpha}}}=\sum_{p}\operatorname*{iCT}_{{\bm{H}}_{p,1}}(T_{p,1}f_{{\bm{\alpha}}})+\operatorname*{iCT}_{{\bm{H}}_{p,2}}(T_{p,2}f_{{\bm{\alpha}}})

where

Tp,1(X)=e2πiλ,X(2πi)2(α1,Xc1)(α2,Xc2)(1e2πiα1,X)(1e2πiα2,X);T_{p,1}(X)=\frac{e^{2\pi{\mathrm{i}}\langle\lambda,X\rangle}(2\pi{\mathrm{i}}\ell)^{2}(\langle\alpha_{1},X\rangle-c_{1})(\langle\alpha_{2},X\rangle-c_{2})}{(1-e^{2\pi{\mathrm{i}}\ell\langle\alpha_{1},X\rangle})(1-e^{2\pi{\mathrm{i}}\ell\langle\alpha_{2},X\rangle});}
Tp,2(X)=e2πiα1+λ,X(2πi)2(α1,Xc1)(α3,Xc3)(1e2πiα1,X)(1e2πiα3,X);T_{p,2}(X)=\frac{e^{2\pi{\mathrm{i}}\langle\ell\alpha_{1}+\lambda,X\rangle}(2\pi{\mathrm{i}}\ell)^{2}(\langle\alpha_{1},X\rangle-c_{1})(\langle\alpha_{3},X\rangle-c_{3})}{(1-e^{2\pi{\mathrm{i}}\ell\langle\alpha_{1},X\rangle})(1-e^{2\pi{\mathrm{i}}\ell\langle\alpha_{3},X\rangle});}

To compute the iterated residue, introduce variables zk=2πi(αk,Xck)z_{k}=2\pi{\mathrm{i}}(\langle\alpha_{k},X\rangle-c_{k}), let tp=exp(p)Zt_{p}=\exp(p)\in Z, and write λ=μ1ϖ1+μ2ϖ2\lambda=\mu_{1}\varpi_{1}+\mu_{2}\varpi_{2}. Using ϖ1=23α1+13α2\varpi_{1}=\tfrac{2}{3}\alpha_{1}+\tfrac{1}{3}\alpha_{2}, ϖ2=13α1+23α2\varpi_{2}=\tfrac{1}{3}\alpha_{1}+\tfrac{2}{3}\alpha_{2} the first term reads as

iCT𝑯p,1(Tp,1f𝜶)=CTz1=0CTz2=0(tpλe(23μ1+13μ2)z1+(13μ1+23μ2)z22z1z2(1ez1)(1ez2)(1ez1+z2)(1ez1)(1ez2)).\operatorname*{iCT}_{{\bm{H}}_{p,1}}(T_{p,1}f_{{\bm{\alpha}}})=\operatorname*{CT}_{z_{1}=0}\operatorname*{CT}_{z_{2}=0}\left(\frac{t_{p}^{\lambda}e^{(\frac{2}{3}\mu_{1}+\frac{1}{3}\mu_{2})z_{1}+(\frac{1}{3}\mu_{1}+\frac{2}{3}\mu_{2})z_{2}}\,\ell^{2}\,z_{1}z_{2}}{(1-e^{z_{1}})(1-e^{z_{2}})(1-e^{z_{1}+z_{2}})(1-e^{\ell z_{1}})(1-e^{\ell z_{2}})}\right).

For the second term, one uses the variable z3=z1+z2z_{3}=z_{1}+z_{2} in place of z2z_{2}, and finds

iCT𝑯p,2(Tp,2f𝜶)=CTz1=0CTz3=0(tpλe(13μ113μ2)z1+(13μ1+23μ2)z3ez12z1z3(1ez1)(1ez3z1)(1ez3)(1ez1)(1ez3)).\operatorname*{iCT}_{{\bm{H}}_{p,2}}(T_{p,2}f_{{\bm{\alpha}}})=\operatorname*{CT}_{z_{1}=0}\operatorname*{CT}_{z_{3}=0}\left(\frac{t_{p}^{\lambda}e^{(\frac{1}{3}\mu_{1}-\frac{1}{3}\mu_{2})z_{1}+(\frac{1}{3}\mu_{1}+\frac{2}{3}\mu_{2})z_{3}}e^{\ell z_{1}}\,\ell^{2}\,z_{1}z_{3}}{(1-e^{z_{1}})(1-e^{z_{3}-z_{1}})(1-e^{z_{3}})(1-e^{\ell z_{1}})(1-e^{\ell z_{3}})}\right).

Note that the terms for p=0,ϖ1,ϖ2p=0,\varpi_{1},\varpi_{2} are the same except for the factor tpλt_{p}^{\lambda}. These elements tpt_{p} are just the elements of the center ZZ of SU(3)\operatorname{SU}(3), hence ptpλ=3δΛ(λ)\sum_{p}t_{p}^{\lambda}=3\delta_{\Lambda}(\lambda). Carrying out the calculations of the constant terms with the help of Wolfram Alpha, we arrive at the following expression:

Ver(𝜶;γ)(λ,)\displaystyle{\textnormal{Ver}}_{({\bm{\alpha}};\gamma)}(\lambda,\ell) =12(1μ1)(2(μ1+2μ23)+(μ21)(μ1+μ22))δΛ(λ).\displaystyle={\frac{1}{2}}(1-\mu_{1})\Big{(}\ell^{2}-(\mu_{1}+2\mu_{2}-3)\ell+(\mu_{2}-1)(\mu_{1}+\mu_{2}-2)\Big{)}\ \delta_{\Lambda}(\lambda).

This may also be written,

(40) Ver(𝜶;γ)(λ,)=12(μ11)(μ21)(μ1+μ22)δΛ(λ){\textnormal{Ver}}_{({\bm{\alpha}};\gamma)}(\lambda,\ell)=-{\frac{1}{2}}(\mu_{1}-1)(\mu_{2}-\ell-1)(\mu_{1}+\mu_{2}-\ell-2)\delta_{\Lambda}(\lambda)

One may verify (as we did) that a different choice of nbc basis gives the same result.

By the general theory, V𝜶(λ,)V_{{\bm{\alpha}}}(\lambda,\ell) coincides with Ver(𝜶;γ)(λ,){\textnormal{Ver}}_{({\bm{\alpha}};\gamma)}(\lambda,\ell) for all (λ,)Λ×(\lambda,\ell)\in\Lambda^{*}\times{\mathbb{N}} with λ\lambda in the region 𝔠𝜶=𝔠++\ell{\mathfrak{c}}-\Box{\bm{\alpha}}=\ell{\mathfrak{c}}+\Box{\mathcal{R}}_{+}. A simple calculation shows that this region is the interior of the hexagon spanned by the orbit of 0 under the shifted Weyl group action

(41) wλ=w(λτ)+τ,w\bullet\lambda=w(\lambda-\tau)+\tau,

where τ=ϖ2+ρ\tau=\ell\varpi_{2}+\rho is the barycenter of the hexagonal region. One can check that the polynomial is anti-invariant under this action:

Ver(𝜶;γ)(wλ,)=(1)length(w)Ver(𝜶;γ)(λ,).{\textnormal{Ver}}_{({\bm{\alpha}};\gamma)}(w\bullet\lambda,\ell)=(-1)^{\operatorname{length}(w)}{\textnormal{Ver}}_{({\bm{\alpha}};\gamma)}(\lambda,\ell).
Remark 6.1.

Making the replacement μiμi\mu_{i}\leadsto\ell\mu_{i} in (40), dividing by 3\ell^{3} and taking the limit as \ell\rightarrow\infty, the polynomial in the expression above becomes

12μ1(μ21)(μ1+μ21).-\tfrac{1}{2}\mu_{1}(\mu_{2}-1)(\mu_{1}+\mu_{2}-1).

This is 33 times the polynomial (for the same chamber 𝔠\mathfrak{c}) for the Bernoulli series Bα(λ)B_{\bf\alpha}(\lambda), as computed by Baldoni-Boysal-Vergne [4, Equation (2.5.2)]. The factor of 33 comes from the δΛ\delta_{\Lambda} factor, and since #(Λ/Λ)=#Z=3\#(\Lambda^{*}/\Lambda)=\#Z=3.

6.2. Contributions with dimΔ=1\dim\Delta=1.

The 11-dimensional admissible subspaces are of the form Δ=ξ+α\Delta=\xi+{\mathbb{R}}\alpha where α𝜶\alpha\in{\bm{\alpha}} and ξΞ=Λ\xi\in\Xi^{\ast}=\Lambda. The corresponding contribution V𝜶ΔV_{{\bm{\alpha}}}^{\Delta} is a convolution of Ver(𝜶Δ;γΔ){\textnormal{Ver}}_{({\bm{\alpha}}_{\Delta};\gamma_{\Delta})}, where 𝜶Δ={α}{\bm{\alpha}}_{\Delta}=\{\alpha\}, with the partition function P(𝜶Δ𝖼;τΔ)P_{({\bm{\alpha}}_{\Delta}^{\mathsf{c}};\tau_{\Delta})}, where 𝜶Δ𝖼{\bm{\alpha}}_{\Delta}^{\mathsf{c}} are the two remaining roots in 𝜶{\bm{\alpha}}. For the calculation below, we will find it convenient to work with the basis α1,α2\alpha_{1},\alpha_{2} of 𝔱{\mathfrak{t}}, so we will write λ=λ1α1+λ2α2\lambda=\lambda_{1}\alpha_{1}+\lambda_{2}\alpha_{2} and only later switch to the basis ϖ1,ϖ2\varpi_{1},\varpi_{2}.

We compute the contribution from Δ=α1\Delta={\mathbb{R}}\alpha_{1}; thus 𝜶Δ={α1}{\bm{\alpha}}_{\Delta}=\{-\alpha_{1}\}. The chambers of Δ\Delta are the intervals (k,k+1)α1(k,k+1)\alpha_{1} for kk\in{\mathbb{Z}}; here the chamber of γΔ\gamma_{\Delta} is the interval for k=0k=0, while the polarizing vector τΔ\tau_{\Delta} is a positive multiple of ϖ2-\varpi_{2}.

By Proposition 2.6, the restriction of V𝜶ΔV_{{\bm{\alpha}}_{\Delta}} to Δ\Delta is #(𝖳𝖳Δ)\#({\mathsf{T}}_{\ell}\cap{\mathsf{T}}_{\Delta}) times a lower-dimensional Verlinde sum. In our case, #(𝖳𝖳Δ)=3\#({\mathsf{T}}_{\ell}\cap{\mathsf{T}}_{\Delta})=3\ell, while the lower-dimensional Verlinde sum is

λζ=1ζλ11ζ.\lambda\mapsto\sideset{}{{}^{\prime}}{\sum}_{\zeta^{\ell}=1}\frac{\zeta^{\lambda_{1}}}{1-\zeta}.

This is nearly the same as V1V_{1} defined in Section 2.2, and one finds similarly that it is supported on ΔΛΔΛ\Delta\cap\Lambda\subseteq\Delta\cap\Lambda^{*}, and equals the polynomial λλ112(+1)\lambda\mapsto\lambda_{1}-\tfrac{1}{2}(\ell+1) for λ1(0,+1)\lambda_{1}\in(0,\ell+1)\cap{\mathbb{Z}}. It follows that

Ver(𝜶Δ;γΔ)(λ,)=3(λ112(+1))δ(λ1)δ0(λ2).{\textnormal{Ver}}_{({\bm{\alpha}}_{\Delta};\gamma_{\Delta})}(\lambda,\ell)=3\ell\Big{(}\lambda_{1}-\frac{1}{2}(\ell+1)\Big{)}\delta_{{\mathbb{Z}}}(\lambda_{1})\delta_{0}(\lambda_{2}).

As for the partition function P(𝜶Δ𝖼;τΔ)P_{({\bm{\alpha}}_{\Delta}^{\mathsf{c}};\tau_{\Delta})}, note that the list 𝜶Δ𝖼={α2,α3}{\bm{\alpha}}^{\mathsf{c}}_{\Delta}=\{-\alpha_{2},-\alpha_{3}\} is polarized already. Hence, the partition function has support in the intersection of Λ\Lambda with the cone generated by α2,α3-\alpha_{2},-\alpha_{3} (without shifts). In terms of the coordinates λ1,λ2\lambda_{1},\lambda_{2} it is given by

P(𝜶Δ𝖼;τΔ)(λ)={1 if λ2λ10,λi,0 otherwise.P_{({\bm{\alpha}}_{\Delta}^{\mathsf{c}};\tau_{\Delta})}(\lambda)=\begin{cases}1&\mbox{ if }\lambda_{2}\leq\lambda_{1}\leq 0,\ \lambda_{i}\in{\mathbb{Z}},\\ 0&\text{ otherwise.}\end{cases}

Taking the convolution, one obtains

(42) V𝜶α1(λ,)=32(1λ2)(2λ1λ21)H(λ2)δΛ(λ),V_{{\bm{\alpha}}}^{\mathbb{R}\alpha_{1}}(\lambda,\ell)=\tfrac{3}{2}\ell(1-\lambda_{2})\big{(}2\lambda_{1}-\lambda_{2}-\ell-1\big{)}\,H(-\lambda_{2})\,\delta_{\Lambda}(\lambda),

where HH is the Heaviside function supported on [0,)[0,\infty). Finally, switching to the coordinates μ1,μ2\mu_{1},\mu_{2} defined by λ=μ1ϖ1+μ2ϖ2\lambda=\mu_{1}\varpi_{1}+\mu_{2}\varpi_{2} we arrive at the following formula for the contribution from Δ=α1\Delta={\mathbb{R}}\alpha_{1}:

(43) V𝜶α1(λ,)=(3μ12μ2)(μ11)2H(μ1+2μ23)δΛ(λ).V_{{\bm{\alpha}}}^{\mathbb{R}\alpha_{1}}(\lambda,\ell)=\frac{\ell(3-\mu_{1}-2\mu_{2})(\mu_{1}-\ell-1)}{2}H\big{(}-\frac{\mu_{1}+2\mu_{2}}{3}\big{)}\,\delta_{\Lambda}(\lambda).

The contribution V𝜶ΔV_{{\bm{\alpha}}}^{\Delta} of general affine subspaces Δ\Delta, is obtained from V𝜶α1V_{{\bm{\alpha}}}^{\mathbb{R}\alpha_{1}} as follows: Let :𝔱𝔱\mathcal{L}\colon{\mathfrak{t}}\to{\mathfrak{t}} be the unique orientation preserving affine-linear transformation, taking α1\mathbb{R}\alpha_{1} to Δ\Delta, taking ϖ2-\varpi_{2} to a positive multiple of τΔ\tau_{\Delta}, and taking (0,1)α1(0,1)\alpha_{1} to the chamber of Δ\Delta containing γΔ\gamma_{\Delta}. For each \ell, let (λ)=(λ/)\mathcal{L}_{\ell}(\lambda)=\ell\mathcal{L}(\lambda/\ell) be the corresponding transformation at level \ell. Let σΔ\sigma_{\Delta} be the shift vector (cf.  (19)), given as minus the sum of roots in 𝜶Δ𝖼{\bm{\alpha}}_{\Delta}^{\mathsf{c}} that are not polarized, and let ±1\pm 1 be the determined by the number of sign changes. (It is easy to see that one gets 1-1 if Δ\Delta is parallel to α3\alpha_{3}, equal to +1+1 otherwise.).

V𝜶Δ(λ,)=±V𝜶α1(1(λσΔ),).V_{{\bm{\alpha}}}^{\Delta}(\lambda,\ell)=\pm V_{{\bm{\alpha}}}^{\mathbb{R}\alpha_{1}}(\mathcal{L}_{\ell}^{-1}(\lambda-\sigma_{\Delta}),\ell).

6.3. Contributions with dimΔ=0\dim\Delta=0

The 0-dimensional admissible subspaces are given by Δ={ξ}\Delta=\{\xi\} for ξΞ=Λ\xi\in\Xi^{\ast}=\Lambda. Consider the contribution of a given ξ\xi, Recall that γΔ=ξ\gamma_{\Delta}=\xi and τΔ=ξγ\tau_{\Delta}=\xi-\gamma. The term V{ξ}(,)V^{\{\xi\}}(\cdot,\ell) is the convolution of δξ\delta_{\ell\xi} with a partition function P𝜶Δ,τΔP_{{\bm{\alpha}}_{\Delta},\tau_{\Delta}}. Convolution with δξ\delta_{\ell\xi} amounts to shifting the argument λ\lambda by ξ\ell\xi, but the partition function depends on the polarization.

Since we took γ\gamma to be in the interior of the Weyl alcove, adding γ-\gamma to ξ\xi shifts it into the interior of one of the Weyl chambers.

Let us now first consider the case {ξ}=0\{\xi\}=0. Then τΔ=γ\tau_{\Delta}=-\gamma, so that 𝜶={\bm{\alpha}}=\mathfrak{R}_{-} is already polarized. Let Λ,λ𝔓(λ)\Lambda^{*}\to\mathbb{Z},\ \lambda\mapsto\mathfrak{P}(\lambda) be the Kostant partition function, given as the number of ways of writing μ\mu as a linear combination of elements of +\mathfrak{R}_{+} with non-negative coefficients. Writing λ=μ1ϖ1+μ2ϖ2\lambda=\mu_{1}\varpi_{1}+\mu_{2}\varpi_{2}, one has that

𝔓(μ)={13(μ2μ1)δΛ(λ)μ2μ112μ213(μ1μ2)δΛ(λ)μ1μ212μ10 otherwise\mathfrak{P}(\mu)=\begin{cases}\frac{1}{3}(\mu_{2}-\mu_{1})\ \delta_{\Lambda}(\lambda)&\mu_{2}\geq\mu_{1}\geq-{\frac{1}{2}}\mu_{2}\\ \frac{1}{3}(\mu_{1}-\mu_{2})\ \delta_{\Lambda}(\lambda)&\mu_{1}\geq\mu_{2}\geq-{\frac{1}{2}}\mu_{1}\\ 0&\mbox{ otherwise}\end{cases}

Then

V{0}=P𝜶{0},τ{0}(λ)=𝔓(λ).V^{\{0\}}=P_{{\bm{\alpha}}_{\{0\}},\tau_{\{0\}}}(\lambda)=\mathfrak{P}(-\lambda).

More generally, for ξΛ𝔱\xi\in\Lambda\cap\mathfrak{t}_{-}, we obtain a shifted version,

V{ξ}=𝔓(ξλ).V^{\{\xi\}}=\mathfrak{P}(\ell\xi-\lambda).

For more general ξΛ\xi\in\Lambda^{*}, we have that τΔw𝔱\tau_{\Delta}\in w\mathfrak{t}_{-} for a unique wWw\in W, also characterized as the shortest Weyl group element with w1ξ𝔱w^{-1}\xi\in\mathfrak{t}_{-}. Equivalently, ww is the unique element such that α,w1τΔ>0\langle\alpha,w^{-1}\tau_{\Delta}\rangle>0 for all negative roots. We hence see that

𝜶pol=w,𝜶+=w,𝜶=w+.{\bm{\alpha}}^{\operatorname{pol}}=w\mathfrak{R}_{-},\ \ {\bm{\alpha}}_{+}=\mathfrak{R}_{-}\cap w\mathfrak{R}_{-},\ \ {\bm{\alpha}}_{-}=\mathfrak{R}_{-}\cap w\mathfrak{R}_{+}.

As is well-known, the set w+\mathfrak{R}_{-}\cap w\mathfrak{R}_{+} has cardinality l(w)l(w), and

σ:=α𝜶α=ρwρ.\sigma:=-\sum_{\alpha\in{\bm{\alpha}}_{-}}\alpha=\rho-w\rho.

We hence obtain

P𝜶,τΔ(λ)=(1)length(w)P𝜶pol,τΔ(λρ+wρ).P_{{\bm{\alpha}},\tau_{\Delta}}(\lambda)=(-1)^{\operatorname{length}(w)}P_{{\bm{\alpha}}^{\operatorname{pol}},\tau_{\Delta}}(\lambda-\rho+w\rho).

Observe that 𝔓(μ)\mathfrak{P}(-\mu) is the number of ways of writing μ\mu as a linear combination of elements of \mathfrak{R}_{-} with non-negative coefficients, and is thus also the number of ways of writing wμw\mu as a linear combination of elements of w=𝜶polw\mathfrak{R}_{-}={\bm{\alpha}}^{\operatorname{pol}} with non-negative coefficients. That is, 𝔓(μ)=P𝜶pol,τΔ(wμ)\mathfrak{P}(-\mu)=P_{{\bm{\alpha}}^{\operatorname{pol}},\tau_{\Delta}}(w\mu). This shows

P𝜶,τΔ(λ)=(1)l(w)𝔓(w1(λρ)ρ).P_{{\bm{\alpha}},\tau_{\Delta}}(\lambda)=(-1)^{l(w)}\mathfrak{P}(-w^{-1}(\lambda-\rho)-\rho).

Finally, convolution with δξ\delta_{\ell\xi} amounts to replacing λ\lambda with λξ\lambda-\ell\xi. We conclude that the contribution from Δ={ξ}\Delta=\{\xi\}, with ξw𝔱\xi\in w\mathfrak{t}_{-}, is

V𝜶Δ(λ,)=(1)l(w)𝔓(w1(ξλ+ρ)ρ).V^{\Delta}_{{\bm{\alpha}}}(\lambda,\ell)=(-1)^{l(w)}\mathfrak{P}(w^{-1}(\ell\xi-\lambda+\rho)-\rho).

7. Equivariant Bernoulli series and Verlinde sums.

In this section we define Verlinde sums and partition functions with additional equivariant parameters. The decomposition formula for these sums is relevant for applications to localization formulas in differential geometry, with the parameters as the recipients for ‘curvatures’. Note that the idea of using curvatures variables in the various partition functions is also used in Vergne’s articles [22, 24].

7.1. Equivariant Bernoulli series.

It will be convenient to consider the lists of weights 𝜶={α1,,αn}{\bm{\alpha}}=\{\alpha_{1},\ldots,\alpha_{n}\} in Λ\Lambda^{*} as the weights of a unitary 𝖳{\mathsf{T}}-representation

A:𝖳U(n),tA(t),A\colon{\mathsf{T}}\to\operatorname{U}(n),\ t\mapsto A(t),

where A(t)ei=tαieiA(t)e_{i}=t^{\alpha_{i}}\,e_{i}. We will use the same letter to denote the infinitesimal representation A:𝔱𝔲(n)A\colon{\mathfrak{t}}\to\mathfrak{u}(n). With this notation, the Bernoulli series can be written as a sum BA(λ)=ξΞe2πiλ,ξdet(A(ξ))1B_{A}(\lambda)={\sum}^{\prime}_{\xi\in\Xi}e^{2\pi{\mathrm{i}}\langle\lambda,\xi\rangle}\det(A(\xi))^{-1}. Letting 𝖧U(n){\mathsf{H}}\subseteq\operatorname{U}(n) be the Lie group of transformations commuting with all A(t)A(t), and 𝔥{\mathfrak{h}} its Lie algebra, we define the ‘𝖧{\mathsf{H}}-equivariant’ multiple Bernoulli sum as the following (generalized) function of λ\lambda, for X𝔥X\in{\mathfrak{h}} sufficiently small:

(44) BA(λ;X)=ξΞe2πiλ,ξdet(A(ξ)+X).B_{A}(\lambda;\,X)=\sideset{}{{}^{\prime}}{\sum}_{\xi\in\Xi}\frac{e^{2\pi{\mathrm{i}}\langle\lambda,\xi\rangle}}{\det(A(\xi)+X)}.

Here {\sum}^{\prime} signifies, as before, a sum over all ξΞ\xi\in\Xi such that det(A(ξ))0\det(A(\xi))\neq 0. A vector τ𝔱\tau\in{\mathfrak{t}} is polarizing for the list 𝜶{\bm{\alpha}} if and only if detA(τ)0\det A(\tau)\neq 0, and in this case we may define

(45) HA,τ(λ;X)=limϵ0+ξ𝔱e2πiλ,ξdet(A(ξiϵτ)+X)dξ.H_{A,\tau}(\lambda;\,X)=\lim_{\epsilon\to 0^{+}}\int_{\xi\in\mathfrak{t}}\frac{e^{2\pi{\mathrm{i}}\langle\lambda,\xi\rangle}}{\det(A(\xi-{\mathrm{i}}\epsilon\tau)+X)}\ {\mbox{d}}\xi.

Both XBA(λ;X)X\mapsto B_{A}(\lambda;\,X) and XHA(λ;X)X\mapsto H_{A}(\lambda;\,X) depend analytically on XX, for X𝔥X\in{\mathfrak{h}} in a sufficiently small neighborhood of zero, in the sense that the integral against test functions is analytic. But for our purposes, it will be quite enough to treat XX as a formal parameter, and thus to consider the Taylor expansions.

Consider the decomposition of the symmetric algebra as sum over multi-indices J=(j1,,jn)J=(j_{1},\ldots,j_{n}) with jk0j_{k}\geq 0,

S(n)=JSJ(n),S({\mathbb{C}}^{n})=\bigoplus_{J}S^{J}({\mathbb{C}}^{n}),

where SJ(n)=Sj1()Sjn()S^{J}({\mathbb{C}}^{n})=S^{j_{1}}({\mathbb{C}})\otimes\cdots\otimes S^{j_{n}}({\mathbb{C}}). Since A(t)A(t) are diagonal matrices, we obtain representations AJA^{J} of 𝖳{\mathsf{T}} on SJ(n)S^{J}({\mathbb{C}}^{n}).

Lemma 7.1.

We have the expansion, for X𝔲(1)nX\in\mathfrak{u}(1)^{n},

det(A(ξ)+X)1=JfJ(X)det(AJ(ξ))1.\det(A(\xi)+X)^{-1}=\sum_{J}f_{J}(X)\,\det(A^{J}(\xi))^{-1}.

Here the sum is over multi-indices J=(j1,,jn)J=(j_{1},\ldots,j_{n}) with jk>0j_{k}>0, and the Taylor expansion of fJ(X)f_{J}(X) starts with terms of order |J|n\geq|J|-n.

Proof.

Since both A(t)A(t) and XX are diagonal it is enough to consider the case n=1n=1. But for n=1n=1, the claim just follows from (t+z)1=j=0zj(1)jtj1(t+z)^{-1}=\sum_{j=0}^{\infty}z^{j}(-1)^{j}t^{-j-1}. ∎

Using the lemma, we have the expansion

(46) BA(λ;X)=JfJ(X)BAJ(λ)B_{A}(\lambda;\,X)=\sum_{J}f_{J}(X)\,B_{A^{J}}(\lambda)

for the Bernoulli series, and similarly

HA,τ(λ;X)=JfJ(X)HAJ,τ(λ)H_{A,\tau}(\lambda;X)=\sum_{J}f_{J}(X)H_{A^{J},\tau}(\lambda)

(with the same fJf_{J}). In particular, the Taylor coefficients of BA(λ;X)B_{A}(\lambda;X) are linear combinations of Bernoulli series for AJA^{J}. Since only JJ with jk>0j_{k}>0 appear, the list 𝜶J{\bm{\alpha}}^{J} of weights appearing in the representation AJA^{J} is obtained from the original list 𝜶{\bm{\alpha}} of weights appearing in AA by increasing some of the multiplicities. Hence, 𝒮(𝜶J)=𝒮(𝜶){\mathcal{S}}({\bm{\alpha}}^{J})={\mathcal{S}}({\bm{\alpha}}). Consequently, the the Taylor coefficients of BA(λ;X)B_{A}(\lambda;X) are supported on the union of Ξ\Xi^{\ast}-translates of Δ0\Delta_{0}, and are polynomial on each chamber of Δ0\Delta_{0} and its translates. A similar discussion applies to HA,τH_{A,\tau}.

As in Section 4, we fix an integral inner product on 𝔱{\mathfrak{t}}, thus identifying 𝔱𝔱{\mathfrak{t}}^{*}\cong{\mathfrak{t}}, and we pick γ𝔱\gamma\in{\mathfrak{t}}, with the property that for all Δ𝒮\Delta\in{\mathcal{S}}, the orthogonal projection γΔ\gamma_{\Delta} onto Δ\Delta lies in a chamber for 𝜶Δ{\bm{\alpha}}_{\Delta}, and τΔ=γΔγ\tau_{\Delta}=\gamma_{\Delta}-\gamma is polarizing for 𝜶Δ𝖼{\bm{\alpha}}_{\Delta}^{\mathsf{c}}. Consider the orthogonal decomposition n=Δn(n)Δ{\mathbb{C}}^{n}={\mathbb{C}}^{n}_{\Delta}\oplus({\mathbb{C}}^{n})_{\Delta}^{\perp}, where the first summand is the sum of the coordinate lines ej\mathbb{C}e_{j} with αj𝜶Δ\alpha_{j}\in{\bm{\alpha}}_{\Delta}, and similarly the second summand is a sum of coordinate lines with αj𝜶Δ𝖼\alpha_{j}\in{\bm{\alpha}}_{\Delta}^{\mathsf{c}}. The two summands are subrepresentations AΔA_{\Delta}, AΔ𝖼A_{\Delta}^{\mathsf{c}} of 𝖳{\mathsf{T}}; the summand Δn{\mathbb{C}}^{n}_{\Delta} is the subspace on which the subtorus 𝖳Δ𝖳{\mathsf{T}}_{\Delta}\subseteq{\mathsf{T}} acts trivially.

The 𝖧{\mathsf{H}}-action preserves this decomposition. We hence obtain an 𝖧{\mathsf{H}}-equivariant Bernoulli series BAΔ(λ;X)B_{A_{\Delta}}(\lambda;\,X), whose Taylor coefficients are polynomial on each chamber of Δ\Delta. We define Ber(AΔ;γΔ)(λ;X)\operatorname{Ber}_{(A_{\Delta};\gamma_{\Delta})}(\lambda;\,X) with equivariant variables X𝔥X\in{\mathfrak{h}}, by requiring that its Taylor coefficients are generalized functions supported on Δ\Delta, given by polynomials on Δ\Delta, and agreeing with the Taylor coefficients of BAΔ(λ;X)B_{A_{\Delta}}(\lambda;\,X) on the chamber containing γΔ\gamma_{\Delta}.

Proposition 7.2 (Equivariant Boysal-Vergne decomposition formula).

For generic choices of γ\gamma as above, the equivariant Bernoulli series decomposes as

(47) BA(X)=Δ𝒮Ber(AΔ;γΔ)(X)H(AΔ𝖼,τΔ)(X),B_{A}(X)=\sum_{\Delta\in{\mathcal{S}}}\operatorname{Ber}_{(A_{\Delta};\gamma_{\Delta})}(X)\star H_{(A_{\Delta}^{\mathsf{c}},\tau_{\Delta})}(X),

(as an equality of formal power series in XX, with coefficients that are functions of λ𝔱\lambda\in{\mathfrak{t}}^{*}).

Proof.

By 𝖧{\mathsf{H}}-invariance, it suffices to prove the equality of both sides over 𝔲(1)n\mathfrak{u}(1)^{n}. For X𝔲(1)nX\in\mathfrak{u}(1)^{n}, we have the expansions

Ber(AΔ;γΔ)(λ;X)=JfJΔ(X)Ber(AΔJ;γΔ)(λ),\operatorname{Ber}_{(A_{\Delta};\gamma_{\Delta})}(\lambda;\,X)=\sum_{J^{\prime}}f_{J^{\prime}}^{\Delta}(X)\operatorname{Ber}_{(A_{\Delta}^{J^{\prime}};\gamma_{\Delta})}(\lambda),

where the fJΔ(X)f_{J^{\prime}}^{\Delta}(X) are defined similar to the fJ(X)f_{J}(X), but using only the basis elements belonging to Δn\mathbb{C}^{n}_{\Delta}. Similarly,

H(AΔ𝖼,τΔ)(λ;X)=J′′fJ′′Δ,𝖼(X)H(AΔ,𝖼J′′,τΔ)(λ)H_{(A_{\Delta}^{\mathsf{c}},\tau_{\Delta})}(\lambda;X)=\sum_{J^{\prime\prime}}f_{J^{\prime\prime}}^{\Delta,\mathsf{c}}(X)H_{(A^{J^{\prime\prime}}_{\Delta,\mathsf{c}},\tau_{\Delta})}(\lambda)

where fJ′′Δ,𝖼(X)f_{J^{\prime\prime}}^{\Delta,\mathsf{c}}(X) are defined in terms of the representation on (Δn)(\mathbb{C}^{n}_{\Delta})^{\perp}. Taking the convolution of these expressions, and using,

fJ(X)=fJΔ(X)fJ′′Δ,𝖼(X)f_{J}(X)=f_{J^{\prime}}^{\Delta}(X)f_{J^{\prime\prime}}^{\Delta,\mathsf{c}}(X)

for J=JJ′′J=J^{\prime}\cup J^{\prime\prime}, we obtain

Ber(AΔ;γΔ)(X)H(AΔ𝖼,τΔ)(X)=JfJ(X)Ber(AΔJ;γΔ)H(AΔ,𝖼J′′,τΔ).\operatorname{Ber}_{(A_{\Delta};\gamma_{\Delta})}(X)\star H_{(A_{\Delta}^{\mathsf{c}},\tau_{\Delta})}(X)=\sum_{J}f_{J}(X)\operatorname{Ber}_{(A_{\Delta}^{J^{\prime}};\gamma_{\Delta})}\star H_{(A^{J^{\prime\prime}}_{\Delta,\mathsf{c}},\tau_{\Delta})}.

Summing over all Δ\Delta, and using the non-equivariant Boysal-Vergne decomposition formula we recover (46). ∎

7.2. Equivariant Verlinde sums.

should one For the Verlinde sums, we have the 𝖳{\mathsf{T}}-representation AA as in the previous section, as well as possibly an additional diagonal matrix UU(n)U\in\operatorname{U}(n) such that some power of UU is the identity. We denote by A¯=(A,U)\underline{A}=(A,U) the ‘augmented representation’; it determines a list 𝜶¯{\underline{{\bm{\alpha}}}} of augmented weights (α,u)(\alpha,u) for the 1-dimensional weight spaces.

Take 𝖧U(n){\mathsf{H}}\subseteq\operatorname{U}(n) to be the subgroup of transformations commuting with UU and with all A(t)A(t), and 𝔥𝔲(n){\mathfrak{h}}\subseteq\mathfrak{u}(n) its Lie algebra. As before, 𝖧{\mathsf{H}} contains U(1)n\operatorname{U}(1)^{n}. For X𝔥X\in{\mathfrak{h}} sufficiently small, put

(48) VA¯(λ,;X)=t𝖳tλdet(1UA(t1)exp(X))V_{\underline{A}}(\lambda,\ell;\,X)=\sideset{}{{}^{\prime}}{\sum}_{t\in{\mathsf{T}}_{\ell}}\frac{t^{\lambda}}{\det(1-U\,A(t^{-1})\exp(-X))}

where the sum is over all t𝖳t\in{\mathsf{T}}_{\ell} such that UA(t1)U\,A(t^{-1}) has no eigenvalue equal to 11. This is a is a well-defined 𝖧{\mathsf{H}}-invariant analytic function of XX, for XX sufficiently small, which reduces to the Verlinde sum V𝜶¯(λ,)V_{{\underline{{\bm{\alpha}}}}}(\lambda,\ell) when X=0X=0. We also have an equivariant version of the generalized partition function, for a given polarizing vector τ𝔱\tau\in{\mathfrak{t}} such that α,τ0\langle\alpha,\tau\rangle\neq 0 for all weights α\alpha:

(49) P(A¯,τ)(λ;X)=limϵ0+𝖳tλdet(1UA(t1exp(iϵτ))exp(X))dt.P_{(\underline{A},\tau)}(\lambda;\,X)=\lim_{\epsilon\to 0^{+}}\int_{{\mathsf{T}}}\frac{t^{\lambda}}{\det(1-U\,A(t^{-1}\exp({\mathrm{i}}\epsilon\tau))\exp(-X))}\ {\mbox{d}}t.

As in the case of the Bernoulli series, we will consider XX as a ‘formal variable’, hence we will work with the Taylor expansions with respect to XX.

For multi-indices J=(j1,,jn)J=(j_{1},\ldots,j_{n}) with jk>0j_{k}>0, we obtain 𝖳{\mathsf{T}}-representations AJA^{J} on SJ(n)S^{J}({\mathbb{C}}^{n}), with commuting endomorphisms UJU^{J}.

Lemma 7.3.

For X𝔲(1)nX\in\mathfrak{u}(1)^{n}, we have that

(det(1UA(t1)exp(X)))1=JgJ(X)det(1UJAJ(t1))1,\big{(}\det(1-UA(t^{-1})\exp(-X))\big{)}^{-1}=\sum_{J}g_{J}(X)\det(1-U^{J}A^{J}(t^{-1}))^{-1},

where the sum is over multi-indices J=(j1,,jn)J=(j_{1},\ldots,j_{n}) with jk>0j_{k}>0, and where the Taylor expansion of gJ(X)g_{J}(X) starts with terms of order |J|n|J|-n.

Proof.

Similar to the proof of Lemma 7.1, it suffices to consider the case n=1n=1, where it follows from

(1a1ez)1=j=0(1)j(1ez)j(1a1)j1.(1-a^{-1}\ e^{-z})^{-1}=\sum_{j=0}^{\infty}(-1)^{j}(1-e^{-z})^{j}(1-a^{-1})^{-j-1}.\qed

It follows that

(50) VA¯(λ,;X)=JgJ(X)VA¯J(λ,),V_{\underline{A}}(\lambda,\ell;\,X)=\sum_{J}g_{J}(X)\ V_{\underline{A}^{J}}(\lambda,\ell),

and similarly for the partition function PA¯(λ;X)P_{\underline{A}}(\lambda;\,X), with the same coefficients gJ(X)g_{J}(X). By a discussion parallel to that for Bernoulli series, the Taylor coefficients of VA¯(λ,;X)V_{\underline{A}}(\lambda,\ell;\,X) are quasi-polynomial on the same regions (e.g., (12)) as for X=0X=0.

Once again, we fix an element γ𝔱\gamma\in{\mathfrak{t}} that is generic with respect to 𝒮{\mathcal{S}}. For Δ𝒮\Delta\in{\mathcal{S}} we obtain an equivariant Verlinde sum VA¯Δ(X)V_{\underline{A}_{\Delta}}(X), and a 𝖧{\mathsf{H}}-equivariant function

Ver(A¯Δ;γΔ)(λ,;X){\textnormal{Ver}}_{(\underline{A}_{\Delta};\gamma_{\Delta})}(\lambda,\ell;\,X)

whose Taylor coefficients are quasi-polynomial on {(λ,):λΔ}\{(\lambda,\ell)\colon\lambda\in\Delta\} and agrees with VA¯Δ(λ,;X)V_{\underline{A}_{\Delta}}(\lambda,\ell;X) whenever 1λ\frac{1}{\ell}\lambda lies in the same chamber as γΔ\gamma_{\Delta}. The decomposition formula extends to the setting with parameters:

Proposition 7.4 (Equivariant decomposition formula for Verlinde sums).
(51) VA¯(X)=Δ𝒮Ver(A¯Δ,γΔ)(X)P(A¯Δ𝖼,τΔ)(X)V_{\underline{A}}(X)=\sum_{\Delta\in{\mathcal{S}}}{\textnormal{Ver}}_{(\underline{A}_{\Delta},\gamma_{\Delta})}(X)\star P_{(\underline{A}_{\Delta}^{\mathsf{c}},\tau_{\Delta})}(X)

(convolution of functions on Λ\Lambda^{\ast}, for fixed \ell).

Outline of proof.

The proof is parallel to that of Proposition 7.2, hence we will be brief. By 𝖧{\mathsf{H}}-invariance, it suffices to prove the identity for X𝔲(1)nX\in\mathfrak{u}(1)^{n}. The terms on the right hand side have expansions

Ver(A¯Δ;γΔ)(X)=JgJΔ(X)Ver(A¯ΔJ;γΔ),{\textnormal{Ver}}_{(\underline{A}_{\Delta};\gamma_{\Delta})}(X)=\sum_{J^{\prime}}g_{J^{\prime}}^{\Delta}(X)\ {\textnormal{Ver}}_{(\underline{A}^{J^{\prime}}_{\Delta};\gamma_{\Delta})},

and

P(AΔ𝖼,τΔ)(X)=J′′gJ′′Δ,𝖼(X)P(AΔ,𝖼J′′,τΔ).P_{(A_{\Delta}^{\mathsf{c}},\tau_{\Delta})}(X)=\sum_{J^{\prime\prime}}g_{J^{\prime\prime}}^{\Delta,\mathsf{c}}(X)P_{(A^{J^{\prime\prime}}_{\Delta,\mathsf{c}},\tau_{\Delta})}.

Taking their convolution product, and using

gJΔ(X)gJ′′Δ,𝖼(X)=gJ(X)g_{J^{\prime}}^{\Delta}(X)g_{J^{\prime\prime}}^{\Delta,\mathsf{c}}(X)=g_{J}(X)

one may carry out the summation over Δ\Delta by using the non-equivariant decomposition formula for Verlinde sums. The resulting expression is the expansion for VA¯(X)V_{\underline{A}}(X). ∎

7.3. Differential form valued Bernoulli series and Verlinde sums.

Suppose now that EME\to M is a 𝖳{\mathsf{T}}-equivariant Hermitian vector bundle over a connected base, where the action on the base MM is trivial. The TT-action will be denoted AE:𝖳Aut(E)A_{E}\colon{\mathsf{T}}\to\operatorname{Aut}(E); it may be regarded as a family of unitary 𝖳{\mathsf{T}}-representations, smoothly labeled by the points of MM. Letting REΩ2(M,End(E))R_{E}\in\Omega^{2}(M,\operatorname{End}(E)) be the curvature of a 𝖳{\mathsf{T}}-invariant connection on EE, and Eul(E,){\textnormal{Eul}}(E,\cdot) the corresponding 𝖳{\mathsf{T}}-equivariant Euler form, we have for all ξ𝔱\xi\in{\mathfrak{t}}

Eul(E,2πiξ)=det(AE(ξ)+i2πRE)Ω(M).{\textnormal{Eul}}(E,-2\pi{\mathrm{i}}\xi)=\det\big{(}A_{E}(\xi)+\tfrac{{\mathrm{i}}}{2\pi}R_{E}\big{)}\in\Omega(M).

The expression

BE(λ)=ξΞe2πiλ,ξEul(E,2πiξ)Ω(M)B_{E}(\lambda)=\sideset{}{{}^{\prime}}{\sum}_{\xi\in\Xi^{\ast}}\frac{e^{2\pi{\mathrm{i}}\langle\lambda,\xi\rangle}}{{\textnormal{Eul}}(E,-2\pi{\mathrm{i}}\xi)}\in\Omega(M)

appears in the fixed point formula for Duistermaat-Heckman measures of Hamiltonian loop group spaces (see [10]). It may be regarded as the characteristic form corresponding to an equivariant Bernoulli series. To see this more clearly, note that by rigidity of actions of compact Lie groups, the fibers of EE are equivariantly isomorphic to a fixed 𝖳{\mathsf{T}}-representation tA(t)t\mapsto A(t) on some n{\mathbb{C}}^{n}, which we may take to be a representation by diagonal matrices. The associated frame bundle Fr(E)\operatorname{Fr}(E), with fibers the 𝖳{\mathsf{T}}-equivariant unitary isomorphisms from fibers of EE to n{\mathbb{C}}^{n}, has structure group 𝖧{\mathsf{H}} the unitary automorphisms commuting with 𝖳{\mathsf{T}}. The Chern-Weil map takes 𝖧{\mathsf{H}}-invariant formal power series on 𝔥{\mathfrak{h}} to differential forms on the base. Applying the Chern-Weil map to BA(λ,X)B_{A}(\lambda,X) we obtain the closed differential form BE(λ)B_{E}(\lambda). As an Ω(M)\Omega(M)-valued function of λ\lambda, it is polynomial on chambers of 𝒮{\mathcal{S}}. Given Δ𝒮\Delta\in{\mathcal{S}}, and a generic choice of γ𝔱\gamma\in{\mathfrak{t}} as in previous sections, we define Ω(M)\Omega(M)-valued generalized functions of λ\lambda

(52) Ber(EΔ;γΔ)(λ),H(EΔ,τΔ)(λ)\operatorname{Ber}_{(E_{\Delta};\gamma_{\Delta})}(\lambda),\ \ H_{(E_{\Delta}^{\perp},\tau_{\Delta})}(\lambda)

where EΔEE_{\Delta}\subseteq E is the subbundle fixed by TΔT_{\Delta}, and EΔE_{\Delta}^{\perp} its orthogonal complement in EE. Both are associated bundles to Fr(E)\operatorname{Fr}(E), and the differential forms (52) are obtained by applying the Chern-Weil map to the corresponding 𝖧{\mathsf{H}}-equivariant invariant functions. We obtain the form-valued decomposition formula,

BE=Δ𝒮Ber(EΔ;γΔ)H(EΔ,τΔ).B_{E}=\sum_{\Delta\in{\mathcal{S}}}\operatorname{Ber}_{(E_{\Delta};\gamma_{\Delta})}\star H_{(E_{\Delta}^{\perp},\tau_{\Delta})}.

For the Verlinde sums, we allow for a slightly more general setting where EE comes with a unitary automorphism UEU_{E}, fixing the base and commuting with AEA_{E}, and such that some power of UEU_{E} is the identity. We write A¯E=(AE,UE)\overline{A}_{E}=(A_{E},U_{E}). By rigidity of actions of compact groups, the fibers of EE are isomorphic to n\mathbb{C}^{n} with a given 𝖳{\mathsf{T}}-action by diagonal matrices and an additional unitary automorphism UU, also by diagonal matrices. Take Fr(E)\operatorname{Fr}(E) to be the corresponding frame bundle, consisting of unitary maps from fibers of EE to n\mathbb{C}^{n} intertwining AE,UEA_{E},U_{E} with A,UA,U, respectively; it is a principal 𝖧{\mathsf{H}}-bundle, where 𝖧U(n){\mathsf{H}}\subseteq\operatorname{U}(n) is the group of unitary automorphism commuting with both AA and UU.

Choose a principal connection on Fr(E)\operatorname{Fr}(E), and let REΩ2(M,End(E))R_{E}\in\Omega^{2}(M,\operatorname{End}(E)) be the curvature form of the resulting linear connection on EE. Then

𝒟(E,t)=det(1UEAE(t)1ei2πRE)Ω(M),{\mathcal{D}_{\mathbb{C}}}(E,t)=\det(1-U_{E}\,A_{E}(t)^{-1}e^{-\frac{{\mathrm{i}}}{2\pi}R_{E}})\in\Omega(M),

for t𝖳t\in{\mathsf{T}} is the characteristic form corresponding to det(1UA(t)1exp(X))\det(1-U\,A(t)^{-1}\exp(X)). The expression

VE(λ,):=t𝖳tλ𝒟(E,t)Ω(M)V_{E}(\lambda,\ell):=\sideset{}{{}^{\prime}}{\sum}_{t\in{\mathsf{T}}_{\ell}}\frac{t^{\lambda}}{{\mathcal{D}_{\mathbb{C}}}(E,t)}\in\Omega(M)

arises from the localization formula for the quantization of Hamiltonian loop group spaces [1, 11]; it is also obtained by applying the Chern-Weil map to the equivariant Verlinde sum VA(λ,;X)V_{A}(\lambda,\ell;X). The decomposition formula for these differential form-valued Verlinde sums reads as

(53) VE=Δ𝒮Ver(EΔ;γΔ)P(EΔ,τΔ),V_{E}=\sum_{\Delta\in{\mathcal{S}}}{\textnormal{Ver}}_{(E_{\Delta};\gamma_{\Delta})}\star P_{(E_{\Delta}^{\perp},\tau_{\Delta})},

for generic choice of γ𝔱\gamma\in{\mathfrak{t}}. Here Ver(EΔ;γΔ){\textnormal{Ver}}_{(E_{\Delta};\gamma_{\Delta})} and P(EΔ,τΔ)P_{(E_{\Delta}^{\perp},\tau_{\Delta})} are obtained by applying the Chern-Weil map to the corresponding 𝖧{\mathsf{H}}-equivariant functions.


Appendix A Functions on lattices

Let Λ\Lambda be a lattice, with dual lattice Λ=Hom(Λ,)\Lambda^{\ast}=\operatorname{Hom}(\Lambda,{\mathbb{Z}}). We denote by 𝔱=Λ{\mathfrak{t}}=\Lambda\otimes_{{\mathbb{Z}}}{\mathbb{R}} the vector space spanned by Λ\Lambda and by 𝖳=𝔱/Λ{\mathsf{T}}={\mathfrak{t}}/\Lambda the torus.

A.1. Functions on Λ\Lambda^{\ast}.

Consider the vector space Map(Λ,){\textnormal{Map}}(\Lambda^{\ast},{\mathbb{C}}) of functions f:Λf\colon\Lambda^{\ast}\to{\mathbb{C}}, and its subspace Map0(Λ,){\textnormal{Map}}_{0}(\Lambda^{\ast},{\mathbb{C}}) of functions of compact (i.e., finite) support. Every subset SΛS\subseteq\Lambda^{\ast} defines a delta-function

δSMap(Λ,),δS(λ)={1:λS0:λS\delta_{S}\in{\textnormal{Map}}(\Lambda^{\ast},{\mathbb{C}}),\ \ \delta_{S}(\lambda)=\begin{cases}1&\colon\lambda\in S\\ 0&\colon\lambda\not\in S\end{cases}

Every t𝖳t\in{\mathsf{T}} defines an evaluation function

𝖾tMap(Λ,),𝖾t(λ)=tλ.\mathsf{e}_{t}\in{\textnormal{Map}}(\Lambda^{\ast},{\mathbb{C}}),\ \ \mathsf{e}_{t}(\lambda)=t^{\lambda}.

A.2. Push-forward, pull-back.

Given a lattice homomorphism ϕ:(Λ)Λ\phi\colon(\Lambda^{\prime})^{\ast}\to\Lambda^{\ast} (dual to some morphism χ:ΛΛ\chi\colon\Lambda\to\Lambda^{\prime}) we define the usual pull-back map ϕ:Map(Λ,)Map((Λ),)\phi^{*}\colon{\textnormal{Map}}(\Lambda^{\ast},{\mathbb{C}})\to{\textnormal{Map}}((\Lambda^{\prime})^{\ast},{\mathbb{C}}), as well as the push-forward

ϕ:Map0((Λ),)Map0(Λ,)\phi_{*}\colon{\textnormal{Map}}_{0}((\Lambda^{\prime})^{\ast},{\mathbb{C}})\to{\textnormal{Map}}_{0}(\Lambda^{\ast},{\mathbb{C}})

by duality to the pull-back under χ\chi. Explicitly, (ϕf)(λ)=λϕ1(λ)f(λ)(\phi_{*}f)(\lambda)=\sum_{\lambda^{\prime}\in\phi^{-1}(\lambda)}f(\lambda^{\prime}).

A.3. Convolution.

As a special case, letting Add:Λ×ΛΛ\mathrm{Add}\colon\Lambda^{\ast}\times\Lambda^{\ast}\to\Lambda^{\ast} be the addition (with dual map ΛΛ×Λ\Lambda\to\Lambda\times\Lambda the diagonal inclusion), we define the convolution

f1f2=Add(f1f2).f_{1}\star f_{2}=\mathrm{Add}_{*}(f_{1}\otimes f_{2}).

Thus, (f1f2)(λ)=λ1+λ2=λf1(λ1)f2(λ2)(f_{1}\star f_{2})(\lambda)=\sum_{\lambda_{1}+\lambda_{2}=\lambda}f_{1}(\lambda_{1})f_{2}(\lambda_{2}). More generally, f1f2f_{1}*f_{2} is defined even for functions of non-compact support, provided Add\mathrm{Add} restricts to a proper map on the support of f1f2f_{1}*f_{2}. Some basic properties of the convolution are

(54) (𝖾tf)(𝖾tg)=𝖾t(fg),(\mathsf{e}_{t}\ f)\star(\mathsf{e}_{t}\ g)=\mathsf{e}_{t}\ (f\star g),
(55) (ϕfg)=ϕ(fϕg).(\phi^{*}f\star g)=\phi^{*}(f\star\phi_{*}g).

Convolution with δμ\delta_{\mu} for μΛ\mu\in\Lambda^{\ast} acts as a translation: (δμf)(λ)=f(λμ)(\delta_{\mu}\star f)(\lambda)=f(\lambda-\mu).

A.4. Finite difference operators.

For any pair β¯=(β,v)\underline{\beta}=(\beta,v), where βΛ\beta\in\Lambda^{*} and vv\in{\mathbb{C}} (usually taken to be a root of unity), define the following finite difference operator on the space of functions f:Λf\colon\Lambda^{\ast}\to{\mathbb{C}}:

(56) (β¯f)(λ)=f(λ)vf(λβ).(\nabla_{\underline{\beta}}f)(\lambda)=f(\lambda)-v\,f(\lambda-\beta).

The finite difference operators for any two such functions β1,β2\beta_{1},\beta_{2} commute. Under convolution of functions

(57) β(fg)=(βf)g=f(βg)\nabla_{\beta}(f\star g)=(\nabla_{\beta}f)\star g=f\star(\nabla_{\beta}g)

A.5. Finite Fourier transform.

Suppose Ξ𝔱\Xi\subseteq{\mathfrak{t}} is a lattice containing Λ\Lambda. Then Ξ/Λ\Xi/\Lambda is a finite group, with dual group Λ/Ξ\Lambda^{*}/\Xi^{*}. If fMap(Λ,)f\in{\textnormal{Map}}(\Lambda^{*},{\mathbb{C}}) is Ξ\Xi^{*}-periodic, it has a finite Fourier expansion

f(λ)=tΛ/Ξf(t)tλ;f(t)=1#(Ξ/Λ)Λ/Ξf(λ)tλ.f(\lambda)=\sum_{t\in\Lambda/\Xi}f^{\vee}(t)\ t^{\lambda};\ \ \ f^{\vee}(t)=\frac{1}{\#(\Xi/\Lambda)}\sum_{\Lambda^{*}/\Xi^{*}}\ f(\lambda)\,t^{-\lambda}.

We have that (β¯f)(t)=(1vtβ)f(t)(\nabla_{\underline{\beta}}f)^{\vee}(t)=(1-v\,t^{-\beta})\,f^{\vee}(t).

A.6. Poisson summation formula.

Given a rational subspace 𝔥𝔱{\mathfrak{h}}\subseteq{\mathfrak{t}}, the (possibly disconnected) group exp(Ξ+𝔥)𝖳\exp(\Xi+{\mathfrak{h}})\subseteq{\mathsf{T}} has a normalized Haar measure. By pull-back, this defines a measure on Ξ+𝔥𝔱\Xi+{\mathfrak{h}}\subseteq{\mathfrak{t}}, which in turn pushes forward to a delta-measure

δΞ+𝔥𝒟(𝔱)\delta_{\Xi+{\mathfrak{h}}}\in{\mathcal{D}}^{\prime}({\mathfrak{t}})

supported on Ξ+𝔥\Xi+{\mathfrak{h}}. We may write this distribution as a sum UδU\sum_{U}\delta_{U} of delta-measures supported on affine subspaces ξ+𝔥\xi+{\mathfrak{h}}, where ξ\xi ranges over representatives of Ξ/𝔥\Xi/{\mathfrak{h}}. We have the following version of the Poisson summation formula:

Proposition A.1.

Let ΞΛ\Xi\supseteq\Lambda be full rank lattices in 𝔱{\mathfrak{t}}, and let 𝔥𝔱{\mathfrak{h}}\subseteq{\mathfrak{t}} be a rational subspace. Let dXdX be Lebesgue measure on 𝔱{\mathfrak{t}}, normalized such that the induced measure on 𝖳=𝔱/Λ{\mathsf{T}}={\mathfrak{t}}/\Lambda is normalized Haar measure. Then we have an equality of distributions on 𝔱{\mathfrak{t}},

(νΞann(𝔥)e2πiν,X)dX=δΞ+𝔥.\Big{(}\sum_{\nu\in\Xi^{\ast}\cap\operatorname{ann}({\mathfrak{h}})}e^{2\pi{\mathrm{i}}\langle\nu,X\rangle}\Big{)}\ dX=\delta_{\Xi+{\mathfrak{h}}}.
Proof.

Both sides are Λ\Lambda-invariant, hence are pullbacks of distributions on 𝖳=𝔱/Λ{\mathsf{T}}={\mathfrak{t}}/\Lambda. The desired identity is equivalent to the equality of distributions on 𝖳{\mathsf{T}},

νΞann(𝔥)tνdt=δexp(Ξ+𝔥)\sum_{\nu\in\Xi^{\ast}\cap\operatorname{ann}({\mathfrak{h}})}t^{\nu}\,dt=\delta_{\exp(\Xi+{\mathfrak{h}})}

where dtdt is the normalized Haar measure on 𝖳{\mathsf{T}} and δexp(Ξ+𝔥)\delta_{\exp(\Xi+{\mathfrak{h}})} is the push-forward of normalized Haar measure on exp(Ξ+𝔥)\exp(\Xi+{\mathfrak{h}}). To compare the Fourier coefficients of these two distributions, integrate against the function ttλt\mapsto t^{-\lambda} for λΛ\lambda\in\Lambda^{\ast}. If λΞann(𝔥)\lambda\in\Xi^{\ast}\cap\operatorname{ann}({\mathfrak{h}}), the Fourier coefficient on the left hand side is 11, and likewise for the right hand side because tλt^{-\lambda} restricts to the constant function 11 on exp(Ξ+𝔥)\exp(\Xi+\mathfrak{h}). If λΞann(𝔥)\lambda\not\in\Xi^{\ast}\cap\operatorname{ann}({\mathfrak{h}}), both sides integrate to zero against tλt^{-\lambda}. ∎

References

  • [1] A. Alekseev, E. Meinrenken, and C. Woodward, The Verlinde formulas as fixed point formulas, J. Symplectic Geom. 1 (2001), no. 1, 1–46.
  • [2] A. Alekseev, E. Meinrenken, and C. Woodward, Duistermaat-Heckman measures and moduli spaces of flat bundles over surfaces, Geom. and Funct. Anal. 12 (2002), 1–31.
  • [3] M. Baldoni, M. Beck, C. Cochet, and M. Vergne, Volume computation for polytopes and partition functions for classical root systems, Discrete Comput. Geom. 35 (2006), no. 4, 551–595.
  • [4] V. Baldoni, A. Boysal, and M. Vergne, Multiple Bernoulli series and volumes of moduli spaces of flat bundles over surfaces, J. Symbolic Comput. 68 (2015), no. part 2, 27–60.
  • [5] A. Boysal and M. Vergne, Multiple Bernoulli series, an Euler-Maclaurin formula, and wall crossings, Ann. Inst. Fourier (Grenoble) 62 (2012), no. 2, 821–858.
  • [6] M. Brion and M. Vergne, Arrangement of hyperplanes. I. Rational functions and Jeffrey-Kirwan residue, Ann. Sci. École Norm. Sup. (4) 32 (1999), no. 5, 715–741.
  • [7] W. Dahmen and C. Micchelli, The number of solutions to linear Diophantine equations and multivariate splines, Trans. Amer. Math. Soc. 308 (1988), no. 2, 509–532.
  • [8] V. Guillemin, E. Lerman, and S. Sternberg, On the Kostant multiplicity formula, J. Geom. Phys. 5 (1988), no. 4, 721–750.
  • [9] V. Guillemin and E. Prato, Heckman, Kostant, and Steinberg formulas for symplectic manifolds, Adv. in Math. 82 (1990), no. 2, 160–179.
  • [10] Y. Loizides, Norm-square localization for Hamiltonian LGLG-spaces, J.  Geom.  and Phys.  114 (2017), 420–449.
  • [11] E. Meinrenken, Twisted KK-homology and group-valued moment maps, International Mathematics Research Notices 2012 (20) (2012), 4563–4618.
  • [12] P.-E. Paradan, Localization of the Riemann-Roch character, J. Funct. Anal. 187 (2001), no. 2, 442–509.
  • [13] by same author, Wall-crossing formulas in Hamiltonian geometry, Geometric aspects of analysis and mechanics, Progr. Math., vol. 292, Birkhäuser/Springer, New York, 2011, pp. 295–343.
  • [14] A. Szenes, Verification of Verlinde’s formulas for SU(2){\rm SU}(2), Internat. Math. Res. Notices (1991), no. 7, 93–98.
  • [15] by same author, The combinatorics of the Verlinde formulas, Vector bundles in algebraic geometry (Durham, 1993), London Math. Soc. Lecture Note Ser., no. 208, Cambridge Univ. Press, 1995, pp. 241–253.
  • [16] by same author, Iterated residues and multiple Bernoulli polynomials, Internat. Math. Res. Notices (1998), no. 18, 937–956.
  • [17] by same author, Residue theorem for rational trigonometric sums and Verlinde’s formula, Duke Math. J. 118 (2003), no. 2, 189–227.
  • [18] A. Szenes and M. Vergne, [Q,R]=0[Q,R]=0 and Kostant partition functions, Preprint, 2010. arXiv:1006.4149.
  • [19] by same author, Residue formulae for vector partitions and Euler-MacLaurin sums, Adv. in Appl. Math. 30 (2003), no. 1-2, 295–342, Formal power series and algebraic combinatorics (Scottsdale, AZ, 2001).
  • [20] M. Thaddeus, Conformal field theory and the cohomology of the moduli space of stable bundles, J. Differential Geom. 35 (1992), no. 1, 131–149.
  • [21] M. Vergne, Multiple Bernoulli series and wall crossing, Slides for AMS meeting in San Francisco, 2010.
  • [22] by same author, Poisson summation formula and box splines, Preprint (2013). arXiv:1302.6599.
  • [23] by same author, Residue formulae for Verlinde sums, and for number of integral points in convex rational polytopes, European women in mathematics (Malta, 2001), World Sci. Publ., River Edge, NJ, 2003, pp. 225–285.
  • [24] by same author, Formal equivariant A^\widehat{A} class, splines and multiplicities of the index of transversally elliptic operators, Izv. Ross. Akad. Nauk Ser. Mat. 80 (2016), no. 5, 157–192.
  • [25] E. Verlinde, Fusion rules and modular transformations in 22d conformal field theory, Nuclear Phys. B 300 (1988), 360–376.
  • [26] E. Witten, On quantum gauge theories in two dimensions, Comm. Math. Phys. 141 (1991), 153–209.