Average Ehrhart coefficients & symplectic Hecke eigenbases

Joshua Maglione

How to Navigate:

Spacebar: Forward

Shift + Spacebar: Backward

Escape: Jump around

Arrow keys: Move around

Polytopes

A polytope is the convex hull of finitely many points in \(\mathbf{R}^n\).

A lattice polytope is the convex hull of finitely many points in \(\mathbf{Z}^n\subset \mathbf{R}^n\).

For us: "polytope" means "lattice polytope"

Discrete Volumes

For a polytope \(P\) and \(k\in \mathbf{Z}_{\geqslant 0}\) the \(k\)th dilate is \(kP = \{k\cdot p ~|~ p \in P\}\)

The discrete volume of a polytope \(P\) is \(\# (P\cap \mathbf{Z}^n)\).

\lim_{k\to \infty} \dfrac{\#(kP\cap \mathbf{Z}^n)}{k^n} = \mathrm{vol}(P)

Theorem (Ehrhart (1962)). The function 

 

is given by a polynomial (in \(k\)).

E_P : \mathbf{Z}_{\geqslant 0} \to \mathbf{Z}, \qquad k\mapsto \#(kP\cap\mathbf{Z}^n)

Corollary. The degree is \(\leqslant n\). The coefficient of \(t^n\) is \(\mathrm{vol}(P)\).

Ehrhart

polynomial

Ehrhart polynomials

\(E(t) = 10t^2 + 5t + 1\)

\(E_P^{\Lambda}(t)\) : Ehrhart polynomial

\(t \mapsto \#(tP\cap \Lambda)\)

lattice: \(\Lambda = \mathbf{Z}^2\)

\(E(1)=16\)

\(E(3)=106\)

\(E(2)=51\)

Other coefficients with meaning?

Let \(E_P^{\Lambda}(t) = c_0 + c_1t + \cdots + c_nt^n\). For \(\ell\in \{0,\dots, n\}=[n]_0\), define

\mathscr{E}_{n, \ell}^{\Lambda}(P) = c_{\ell}.

Thus, \(\mathscr{E}_{n,\ell}^{\Lambda}\) is the function on polytopes in \(\mathbf{R}^n\) extracting the \(\ell\)th coefficient of its Ehrhart polynomial.

\begin{aligned} \mathscr{E}_{n,n} &= \mathrm{vol} \\ \mathscr{E}_{n,n-1} &= \frac{1}{2}\mathrm{vol}_{\partial} \\ &~~\vdots \\ \mathscr{E}_{n,0} &= 1 \end{aligned}
\Lambda = \mathbf{Z}^n

\(\mathbf{Z}^n\subseteq \Lambda\subset \mathbf{Q}^n\)

\begin{aligned} \mathscr{E}_{n,n}^{\Lambda} &= \mathrm{vol}_{\Lambda} \\ \mathscr{E}_{n,n-1}^{\Lambda} &= \frac{1}{2}\mathrm{vol}_{\Lambda,\partial} \\ &~~\vdots \\ \mathscr{E}_{n,0}^{\Lambda} &= 1 \end{aligned}

More

complicated

Example: \((a,b,c)\)-simplex

\mathscr{E}_{3,0}(P) = 1 \qquad \mathscr{E}_{3,2}(P) = \frac{ab+bc+ca+1}{4} \qquad \mathscr{E}_{3,3}(P) = \frac{abc}{6}
(0,0,0)
(0,0,0)
(0,0,c)
(0,b,0)
(a,0,0)
P =

\((a,b,c)\) co-prime

Here, \(S\) is a Dedekind sum. With

\mathscr{E}_{3,1}(P) = \frac{1}{12}\left(\frac{ab}{c} + \frac{bc}{a} + \frac{ca}{b} + \frac{1}{abc} + 3(a+b+c) + 9\right) - S
s(p, q) = \sum_{i=1}^q \left(\left(\frac{i}{q}\right)\right) \left(\left(\frac{pi}{q}\right)\right)
((x)) = \begin{cases} 0 & x\in \mathbf{Z}, \\ x - \lfloor x\rfloor -1/2 & x\not\in \mathbf{Z}, \end{cases}
S = s(ab, c) + s(bc, a) + s(ca, b)

Brion & Vergne (1997) proved that \(\mathscr{E}_{n,\ell}\) are given by certain differential Todd operators acting on \(\mathrm{vol}\) for simple polytopes.

\mathrm{Todd}(z) = \dfrac{z}{1 - \mathrm{exp}(-z)} = 1 + \frac{1}{2}z + \sum_{k=1}^{\infty}(-1)^{k-1}\dfrac{B_{2k}}{(2k)!}z^{2k}

Unimodular invariant valuations

An \(\mathbf{R}\)-valued function \(\varphi\) on polytopes is a valuation if

\varphi(P) + \varphi(Q) = \varphi(P\cup Q) + \varphi(P\cap Q)

whenever \(P, Q, P\cap Q, P\cup Q\) are polytopes and \(\varphi(\varnothing)=0\).

A valuation \(\varphi\) is unimodular invariant if for all \(\alpha\in \mathrm{Aut}(\Lambda)\!\leqslant\! \mathrm{Aff}_{n}(\mathbf{R})\)

\varphi = \varphi\alpha.

Valuations capture

"geometric" measurements

Examples:

  • Volume
  • Surface area
  • Mean width
  • Euler characteristic

is a basis for the space of unimodular invariant valuations.

\left\{\mathscr{E}_{n,0},\; \mathscr{E}_{n,1},\; \dots,\; \mathscr{E}_{n,n}\right\}

Theorem (Betke & Kneser (1985)). The set

The Ehrhart coefficients encode all geometric measurements.

Betke & Kneser is a discrete analogue of Hadwiger's Theorem (1957).

The analogy, unfortunately, does not carry over to the properties of the valuations \(\mathscr{E}_{n,\ell}\).

The \(n+1\) intrinsic volumes are a basis for the space of certain valuations on compact convex bodies in \(\mathbf{R}^n\).

\sum_{m=0}^{\infty} E_P(m) t^m = \dfrac{h^*_0(P) + h^*_1(P)t + \cdots + h^*_n(P)t^n}{(1 - t)^{n+1}}

We can form the Ehrhart series:

Ehrhart series & \(h^*\)-polynomial

The \(h_i^*\) are unimodular invariant but not valuations.

Stanley (1980) proved non-negativity:

\begin{aligned} h_0^*(P) &= 1, & h_k^*(P) &\in \mathbf{Z}_{\geqslant 0} \end{aligned}

and, in (1993), monotonicity:

P\subseteq Q \implies \forall i,\ h_i^*(P)\leqslant h_i^*(Q)

Is there anything special about Ehrhart coefficients?

\sum_{i=0}^n{\color{blue}\mathscr{E}_{n,i}(P)}t^i = \sum_{i=0}^n {\color{blue}h_i^*(P)} \binom{t + n - i}{n} = \sum_{i=0}^n {\color{blue}f_i^*(P)} \binom{t - 1}{i}

Breuer (2012) defined the \(f^*\)-polynomial writing Ehrhart polynomial in a different polynomial basis:

Jochemko & Sanyal (2018) proved that the \(f^*_i\) form the essentially unique basis for \(\mathrm{UIV}_n\) which are non-negative and monotone.

unimodular invariant

valuations

Unique* Hecke eigenfunctions 

Theorem (Alfes, M, & Voll (2025)). The valuations

\{\mathscr{E}_{2n,0}, \mathscr{E}_{2n, 1}, \dots, \mathscr{E}_{2n, 2n}\}

form a symplectic Hecke eigenbasis for the space of \(\mathrm{UIV}_{2n}\).

Theorem (Gunnells & Rodriguez Villegas (2007)). The valuations

\{\mathscr{E}_{n,0}, \mathscr{E}_{n, 1}, \dots, \mathscr{E}_{n, n}\}

form a Hecke eigenbasis for the space of \(\mathrm{UIV}_n\).

Up to independent scaling, this is the unique basis of such valuations.

Lattice perspective

\(\mathbf{Z}^n\)

\(\frac{1}{2}\mathbf{Z}^n\)

\(\frac{1}{3}\mathbf{Z}^n\)

\#(kP \cap \mathbf{Z}^n) = \#(P \cap \tfrac{1}{k}\mathbf{Z}^n)

Consider how \(\mathscr{E}_{n,\ell}^{\Lambda}\) changes as we refine \(\mathbf{Z}^n\subseteq \Lambda\subset \mathbf{Q}^n\).

Represent a lattice \(\Lambda\) by a matrix, whose rows generate \(\Lambda\).

\mathbf{Z}^2 \leftrightsquigarrow \begin{pmatrix} 1 & 0 \\ 0 & 1 \end{pmatrix}

Lattices \(\Lambda\supseteq\mathbf{Z}^2\)

\Lambda = 2^{-1}\left(\begin{smallmatrix} \phantom{0} & \phantom{0} \\ \phantom{0} & \phantom{0} \end{smallmatrix}\right)
\Lambda = 4^{-1}\left(\begin{smallmatrix} \phantom{0} & \phantom{0} \\ \phantom{0} & \phantom{0} \end{smallmatrix}\right)
\Lambda \mapsto \mathcal{E}_{2, 1}^{\Lambda}(\phantom{poly})
\begin{aligned} \mathrm{conv}&\{(0,0), (0,3), (2,3),\\ &\;(3,0), (4,2)\} \end{aligned}
\begin{aligned} \mathrm{conv}&\{(0,0), (0,2), (1,4),\\ &\;(3,2), (4,0)\} \end{aligned}
\begin{aligned} \mathrm{conv}&\{(0,0), (0,2), (1,4),\\ &\;(3,2), (4,0)\} \end{aligned}
\begin{aligned} \mathrm{conv}&\{(0,0), (0,1), (1,0)\} \end{aligned}

Normalized sums of coefficients

Fix a polytope \(P\), with \(\mathscr{E}_{n,\ell}(P)\neq 0\). Define

where \(\Lambda\supseteq\mathbf{Z}^n\) runs over all lattices in \(\mathbf{Q}^n\) with \(\#(\Lambda/\mathbf{Z}^n)=m\).

Just as a lattice generated from a matrix from \(\mathrm{GL}_n(\mathbf{Q})\cap\mathrm{Mat}_n(\mathbf{Z})\),

a symplectic lattice is generated from one in \(\mathrm{GSp}_{2n}(\mathbf{Q})\cap\mathrm{Mat}_{2n}(\mathbf{Z})\).

\mathcal{C}_{P,\ell}^{\mathsf{A}} : \mathbf{N} \longrightarrow \mathbf{Q}
m \longmapsto \dfrac{1}{\mathscr{E}_{n,\ell}(P)} \sum_{\Lambda} \mathscr{E}_{n,\ell}^{\Lambda}(P),

Define \(\mathcal{C}_{P,\ell}^{\mathsf{C}} : \mathbf{N}\to \mathbf{Q}\) similarly but restricted to symplectic lattices.

We computed \(\mathcal{C}_{P,1}^{\mathsf{A}}(2)\) and \(\mathcal{C}_{P,1}^{\mathsf{A}}(4)\).

An arithmetic function

Write \(\mathsf{X}\) for either \(\mathsf{A}\) or \(\mathsf{C}\).

Theorem (Alfes, M, & Voll (2025)). 

(Independence) For polytopes \(P\) and \(Q\), with \(\dim P=\dim Q\),

 

(Multiplicativity) For \(a,b\in \mathbf{N}\) with \(\mathrm{gcd}(a,b)=1\), 

\mathcal{C}_{P,\ell}^{\mathsf{X}}(ab) = \mathcal{C}_{P,\ell}^{\mathsf{X}}(a)\mathcal{C}_{P,\ell}^{\mathsf{X}}(b).
\mathcal{C}_{P,\ell}^{\mathsf{X}}=\mathcal{C}_{Q,\ell}^{\mathsf{X}}.

Growth rate of coefficients

For functions \(f,g : \mathbb{R}\to \mathbb{R}\), we write

f(N) \sim g(N) \quad \iff \quad \lim_{N\to \infty} \dfrac{f(N)}{g(N)} = 1.

Proposition (Alfes, M, & Voll (2025)).

Let \(\ell\in \{0,1,\dots, n\}\). Assuming \(\mathscr{E}_{n,\ell}(P)\neq 0\), 

\sum_{m=1}^{N} \mathcal{C}_{P,\ell}^{\mathsf{A}}(m) \,\sim\, \begin{cases} \frac{\zeta(n-\ell)}{n} \zeta(2)\zeta(3) \cdots \zeta(n-1) N^n & \text{if }\ell\leqslant n-2, \\[0.5em] \frac{1}{n}\zeta(2)\zeta(3) \cdots \zeta(n-2) N^n \log N & \text{if }\ell=n-1, \\[0.5em] \frac{1}{n+1} \zeta(2)\zeta(3) \cdots \zeta(n) N^{n+1} & \text{if }\ell=n. \end{cases}

Here, \(\zeta\) is the Riemann zeta function.

Averaging over all lattices

Corollary. 

Let \(P\) be an \(n\)-dimensional polytope in \(\mathbf{Z}^n\) and \(\ell\in \{0,1,\dots, n-2\}\). The average \(\ell\)th Ehrhart coefficient of \(P\), running over all \(\Lambda\supseteq\mathbf{Z}^n\) in \(\mathbf{Q}^n\) with finite co-index, approaches

\dfrac{\zeta(n-\ell)}{\zeta(n)} \mathscr{E}_{n,\ell}(P).

Example. For \(n=3\), the average linear coefficient approaches

\dfrac{\pi^2}{6\zeta(3)}\mathscr{E}_{3,1}(P) \approx 1.37\cdot\mathscr{E}_{3,1}(P)

Example. For \(n=4\), the average quadratic coefficient approaches

\dfrac{15}{\pi^2}\mathscr{E}_{4,2}(P) \approx 1.52\cdot \mathscr{E}_{4,2}(P)

Growth rate (symplectic edition)

Theorem (Alfes, M, & Voll (2025)).

For each \(\ell\in \{0,1,\dots, 2n\}\) there exists \(c_{n,\ell}\in\mathbb{R}\) such that 

\sum_{m=1}^{N}\mathcal{C}_{P,\ell}^{\mathsf{C}}(m) \,\sim\, c_{n,\ell}\sqrt{N^{n+1}}\sqrt[n]{N^{\max(0,\ell-n)+1}}(\log N)^{\delta_{n,\ell}}

For \(i,j\in\mathbb{N}_0\), write \(\delta_{i,j}\) for the Kronecker delta symbol.

For \(n=2\) (i.e. in \(4\) dimensions) we completely know \(c_{2,\ell}\):

\begin{aligned} c_{2,0} &= \frac{7\zeta(3)}{8}, & c_{2,1} &= \frac{\zeta(2)^2\zeta(3)}{12\zeta(5)}, & c_{2,2} &= \frac{5}{8}, & c_{2,3} &= \frac{2\zeta(2)^2\zeta(3)}{5\zeta(5)}, & c_{2,4} &= \frac{7\zeta(3)}{12}. \end{aligned}

Averaging over symplectic lattices

When \(n=2\) and \(\ell=1\), the average linear coefficient is 

Corollary.

Let \(P\) be \(2n\)-dimensional in \(\mathbf{Z}^{2n}\). The average \(\ell\)th Ehrhart coefficient, running over all symplectic lattices, converges only when \(\ell\in\{0,1,\dots, n-1\}\). Its limit is 

\dfrac{c_{n,\ell}}{c_{n,0}} \mathscr{E}_{2n,\ell}(P).
\dfrac{\pi^4}{63\zeta(5)}\mathscr{E}_{4,1}(P) \approx 1.49 \cdot \mathscr{E}_{4,1}(P).

Modular form analogue

Previous theorems stem from the understanding extracted from the Hecke operators.

Points in \(\mathrm{SL}_2(\mathbf{Z})\setminus \mathbf{H}\)

\(\mathrm{GL}_n(\mathbf{Z})\setminus\{\text{Polytopes}\}\)

Modular form weight \(k\)

\(\mathscr{E}_{n,k}\)

\leftrightsquigarrow
\begin{aligned} G^+ &= \mathrm{GSp}_{2n}(\mathbf{Q}_p)\cap \mathrm{Mat}_{2n}(\mathbf{Z}_p), & \Gamma &= \mathrm{GSp}_{2n}(\mathbf{Z}_p) \end{aligned}

The (symplectic) Hecke ring \(\mathcal{H}\) is the set of functions \(G^+ \to \mathbf{Z}\) that:

  • are continuous (thus, locally constant),
  • have compact support,
  • are constant on \(\Gamma\)-double cosets.

\(\mathbf{Z}\)-basis for \(\mathcal{H}\) : characteristic functions on \(\Gamma\)-double cosets.

\mathcal{H} = \mathbf{Z}[T_0, T_1, \dots, T_n].
\begin{aligned} D_0 &= \begin{pmatrix} I_n & \\ & pI_n \end{pmatrix}, & D_k &= \begin{pmatrix} I_{n-k} & & & \\ & pI_k & & \\ & & p^2I_{n-k} & \\ & & & pI_k \end{pmatrix}. \end{aligned}

We define matrices \(D_0,\dots, D_n\in G^+\): for \(k\in \{1,\dots, n\}\),

Let \(T_k\in\mathcal{H}\) be the characteristic function on \(\Gamma D_k\Gamma\). Thus,

(symplectic)

elemtary

divisors

For \(f\in\mathrm{UIV}_{2n}\), define \(Tf\in\mathrm{UIV}_{2n}\) via

where

\(\Gamma\setminus \Gamma g \Gamma\)  :  set of lattices with prescribed elementary divisors,

Extend linearly to define an action of \(\mathcal{H}\) on \(\mathrm{UIV}_{2n}\).

Let \(T\in\mathcal{H}\) be the characteristic function for \(\Gamma g\Gamma\), so associated with some elementary divisors.

What does this action look like on our generators \(T_k\)?

Tf(P) = \sum_{\Gamma h \,\in\, \Gamma \setminus \Gamma g\Gamma} f(h\cdot P)

\(h\cdot P\)  :  given by some integer matrix in \(\Gamma h\) acting on vertices.

Each matrix corresponds to a \(\Gamma\)-left coset.

Colors correspond to \(\Gamma\)-double cosets.

\Gamma\begin{pmatrix} 1 \\ &1 \\ &&p \\ &&&p \end{pmatrix}\Gamma
T_0
\Gamma\begin{pmatrix} 1 \\ &p \\ &&p^2 \\ &&&p \end{pmatrix}\Gamma
T_1
\Gamma\begin{pmatrix} p \\ &p \\ &&p \\ &&&p \end{pmatrix}\Gamma
T_2

Knowing that the \(\mathscr{E}_{2n,\ell}\) form a symplectic Hecke eigenbasis means we can take advantage of the theory of spherical functions.

Leads to a nice interpretation of the associated zeta function:

\mathcal{Z}_{n,\ell,p}^{\mathsf{X}}(s) = \sum_{\Gamma g \,\in\, \Gamma\setminus \Gamma G^+\Gamma} \dfrac{\mathscr{E}_{2n,\ell}(g\cdot P)}{\mathscr{E}_{2n,\ell}(P)} \left|\Lambda_{g} : \Lambda_0\right|^{-s}.

EhrhartHecke zeta functions

Lattice

from \(\Gamma g\)

To understand growth of \(\mathcal{C}_{P,\ell}^{\mathsf{X}}\), need to understand \(\mathcal{Z}_{n,\ell,p}^{\mathsf{X}}\) for all \(p\):

\mathcal{Z}_{n,\ell}^{\mathsf{X}}(s) := \sum_{m=1}^{\infty} \mathcal{C}_{P,\ell}^{\mathsf{X}}(m)\, m^{-s} = \prod_{p\text{ prime}} \mathcal{Z}_{n,\ell,p}^{\mathsf{X}}(s)

Main method of proof: lattice enumeration from M & Voll (2024).

\mathcal{Z}_{n,\ell}^{\mathsf{A}}(s) = \zeta(s - \ell) \cdot \zeta(s - 1) \zeta(s-2)\cdots \zeta(s - n + 1)

Type A case is simpler. We know the global zeta function:

\mathcal{Z}_{n,\ell,p}^{\mathsf{C}}(s) = \sum_{I,J\subseteq [n]} \Psi_{n,I,J}(p^{-1}) \prod_{i\in I} \dfrac{p^{\binom{i+1}{2} + i(n-i)}t^n}{1-p^{\binom{i+1}{2} + i(n-i)}t^n} \prod_{j\in J} \dfrac{p^{\binom{j+1}{2} + j(n-j) - n + \ell}t^n}{1-p^{\binom{j+1}{2} + j(n-j) - n + \ell}t^n}

Theorem (Alfes, M, & Voll (2025)).

Setting \(t = p^{-s}\),

\Psi_{n,I,J}(Y) = \text{$Y$-multinomial coefficient} \times \prod (1 - Y^k)^{e_k}

Type C case is more subtle. We know the local zeta function:

Summary

The \(\mathscr{E}_{2n,\ell}\) form a symplectic Hecke eigenbasis of \(\mathrm{UIV}_{2n}\).

Theory of spherical functions on \(p\)-adic groups provide tools to study growth of coefficients.

Thank you!

  • Unique* valuations with this property
  • is \(\zeta(n-\ell)/\zeta(n)\) for \(\ell\in\{0,\dots, n-2\}\) over all lattices,

The average \(\ell\)th coefficient of the Ehrhart polynomial...

  • converges for \(\ell\in \{0,\dots, \frac{2n}{2} - 1\}\) over sympletic lattices.