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

The Distribution of Sandpile Groups of Random Graphs with their Pairings

Eliot Hodges Department of Mathematics, Harvard University, Cambridge, Massachusetts 02138 [email protected]
Abstract.

We determine the distribution of the sandpile group (also known as the Jacobian) of the Erdős-Rényi random graph G(n,q)G(n,q) along with its canonical duality pairing as nn tends to infinity, fully resolving a conjecture from 2015 due to Clancy, Leake, and Payne and generalizing the result by Wood on the groups. In particular, we show that a finite abelian pp-group GG equipped with a perfect symmetric pairing δ\delta appears as the Sylow pp-part of the sandpile group and its pairing with frequency inversely proportional to |G||Aut(G,δ)||G||\operatorname{\mathrm{Aut}}(G,\delta)|, where Aut(G,δ)\operatorname{\mathrm{Aut}}(G,\delta) is the set of automorphisms of GG preserving the pairing δ\delta. While this distribution is related to the Cohen-Lenstra distribution, the two distributions are not the same on account of the additional algebraic data of the pairing. The proof utilizes the moment method: we first compute a complete set of moments for our random variable (the average number of epimorphisms from our random object to a fixed object in the category of interest) and then show the moments determine the distribution. To obtain the moments, we prove a universality result for the moments of cokernels of random symmetric integral matrices whose dual groups are equipped with symmetric pairings that is strong enough to handle both the dependence in the diagonal entries and the additional data of the pairing. We then apply results due to Sawin and Wood to show that these moments determine a unique distribution.

1. Introduction

To any graph Γ\Gamma we may associate a natural abelian group SΓS_{\Gamma}, which is commonly referred to as the sandpile group, the Jacobian, the critical group, or the Picard group of the graph. The sandpile group is an important object of interest in several subjects, including but not limited to statistical mechanics, combinatorics, and arithmetic geometry (hence its several names; see [48] for an overview of these various connections). If Γ\Gamma is connected, then SΓS_{\Gamma} is finite, and |SΓ||S_{\Gamma}| is the number of spanning trees of Γ\Gamma. The sandpile group of a graph comes naturally equipped with an additional important piece of algebraic data: a canonical duality pairing (i.e., a perfect symmetric pairing) SΓ×SΓ/S_{\Gamma}\times S_{\Gamma}\to\mathbb{Q}/\mathbb{Z}, which was first introduced by Bosch and Lorenzini in [13]. In [42], Lorenzini conjectured the frequency that SΓS_{\Gamma} is cyclic and posed several other questions about statistics of the distribution of sandpile groups of random graphs. Initially, some suspected that the distribution of sandpile groups of random graphs obeyed a Cohen-Lenstra [17] heuristic—a natural first guess for the distribution of a random finite abelian group. However, Clancy, Leake, and Payne collected empirical data in [16] suggesting that this was not the case. In particular, Clancy, Leake, and Payne noticed that a finite abelian group with duality pairing (G,,G)(G,\langle\,,\rangle_{G}) seems to appear as the sandpile group and pairing of a random graph with frequency proportional to

1|G||Aut(G,,G)|,\frac{1}{|G||\operatorname{\mathrm{Aut}}(G,\langle\,,\rangle_{G})|},

where Aut(G,,G)\operatorname{\mathrm{Aut}}(G,\langle\,,\rangle_{G}) denotes the automorphisms of GG preserving the pairing. In other words, they observed that the natural duality pairing on SΓS_{\Gamma} seemed to be affecting its distribution. In [15], Clancy, Kaplan, Leake, Payne, and Wood proved an analogous result where the sandpile group is replaced by the cokernel of a Haar-random symmetric matrix with entries in p\mathbb{Z}_{p} (such groups also have duality pairings). Later, in [54], Wood proved the distribution of sandpile groups of Erdős-Rényi random graphs, almost fully resolving the conjecture of Clancy, Leake, and Payne. However, Wood’s approach was unable to determine the distribution of the pairings, and the conjecture regarding the distribution of the pairings has since remained open (this conjecture is restated as Open Problem 3.13 in [56]). In this paper, we extend Wood’s result to show that the sandpile group of an Erdős-Rényi random graph with its pairing is distributed as Clancy, Leake, and Payne conjectured.

Let ΓG(n,q)\Gamma\in G(n,q) be an Erdős-Rényi random graph on nn vertices with edges occurring independently with probability qq, where 0<q<10<q<1. Let SΓS_{\Gamma} denote the sandpile group of Γ\Gamma, and for a prime pp let SΓ,pS_{\Gamma,p} denote the Sylow pp-subgroup of SΓS_{\Gamma}. Also, let ,SΓ\langle\,,\rangle_{S_{\Gamma}} denote the duality pairing on the torsion subgroup of SΓS_{\Gamma} and ,SΓ,p\langle\,,\rangle_{S_{\Gamma},p} its restriction to SΓ,pS_{\Gamma,p}.111As we will see in Section 3, it is actually the torsion subgroup of the sandpile group that naturally carries a duality pairing. When Γ\Gamma is connected, the sandpile group is finite and the torsion subgroup is all of SΓS_{\Gamma}. However, since the graphs in question are Erdős-Rényi random, we must be more careful, since Erdős-Rényi random graphs are not necessarily connected (though they are asymptotically almost surely connected). It follows from Wood’s work in [54] that, for a given finite abelian group GG, we have limn(SΓG)=0\lim_{n\to\infty}\operatorname{\mathbb{P}}(S_{\Gamma}\simeq G)=0. Recalling that a finite abelian group can be decomposed into a direct sum of its Sylow pp-subgroups, we instead ask: for GG a finite abelian pp-group with duality pairing ,G\langle\,,\rangle_{G}, what is the probability as nn\to\infty that (SΓ,p,,SΓ,p)(G,,G)(S_{\Gamma,p},\langle\,,\rangle_{S_{\Gamma},p})\simeq(G,\langle\,,\rangle_{G})?

Theorem 1.1.

Let pp be a prime and GG a finite abelian pp-group with duality pairing ,G\langle\,,\rangle_{G}. For a random graph ΓG(n,q)\Gamma\in G(n,q), let SS denote its sandpile group and ,S\langle\,,\rangle_{S} the duality pairing on its torsion subgroup. Then

(1) limn((Sp,,S,p)(G,,G))=k=1(1p12k)|G||Aut(G,,G)|.\lim_{n\to\infty}\operatorname{\mathbb{P}}((S_{p},\langle\,,\rangle_{S,p})\simeq(G,\langle\,,\rangle_{G}))=\frac{\prod_{k=1}^{\infty}(1-p^{1-2k})}{|G||\operatorname{\mathrm{Aut}}(G,\langle\,,\rangle_{G})|}.

Theorem 1.1 has several interpretations: we may regard it both as a result in combinatorics on random graphs and as a result in arithmetic statistics (there are analogies between sandpile groups and Jacobians of curves over finite fields [13, 39]). In addition to proving Theorem 1.1, we show a stronger result—Theorem 1.3—along the way. Theorem 1.3 is a universality theorem on cokernels of random matrices with duality pairings which implies the analogous result due to Wood (Theorem 1.3 of [54]) by an application of the Orbit-Stabilizer Theorem.

We remark that the numerator on the right-hand side of (1) does not depend on GG and that this product is essentially a normalization constant. Moreover, k=1(1p12k)\prod_{k=1}^{\infty}(1-p^{1-2k}) increases with pp, and for p>1000p>1000 we have k=1(1p12k)>0.999\prod_{k=1}^{\infty}(1-p^{1-2k})>0.999. Note that the right-hand side of (1) is independent of the edge probability qq. Further, in Corollary 10.5, we are able to show a much stronger result than Theorem 1.1: for any finite set of primes, we give the asymptotic probabilities of particular Sylow subgroups with their pairings at these primes.

The proof of Theorem 1.1 uses a result due to Sawin and Wood on the moment problem for random objects in diamond categories, where a moment of a random object is defined to be the average number of epimorphisms to a fixed object in the category. In their paper [50], Sawin and Wood show that as long as the moments do not grow too fast, a unique measure with these moments exists, and a sequence of measures whose moments approach these particular moments in the limit converge weakly to the aforementioned measure.222In our case, the moments grow too quickly to use classical probabilistic methods to show the moments determine a unique distribution. Because ,SΓ\langle\,,\rangle_{S_{\Gamma}} is a duality pairing, there is a corresponding duality pairing on SΓ^=Hom(SΓ,/)\widehat{S_{\Gamma}}=\operatorname{\mathrm{Hom}}(S_{\Gamma},\mathbb{Q}/\mathbb{Z}), the Pontryagin dual group of SΓS_{\Gamma}, which we denote by ,SΓ^\langle\,,\rangle_{\widehat{S_{\Gamma}}}. To prove Theorem 1.1, we first compute a complete set of moments for (SΓ,,SΓ^)(S_{\Gamma},\langle\,,\rangle_{\widehat{S_{\Gamma}}}).

It is not immediately clear which category we should be working in or why we are considering pairings on SΓ^\widehat{S_{\Gamma}} rather than SΓS_{\Gamma}. We define a category 𝒞\mathcal{C} whose objects are finite abelian groups whose dual groups are equipped with symmetric pairings (i.e., the objects of 𝒞\mathcal{C} are finite abelian groups GG such that G^=Hom(G,/)\widehat{G}=\operatorname{\mathrm{Hom}}(G,\mathbb{Q}/\mathbb{Z}) has a symmetric pairing ,G^:G^G^/\langle\,,\rangle_{\widehat{G}}:\widehat{G}\otimes\widehat{G}\to\mathbb{Q}/\mathbb{Z}). Note that if (A,,A^)(A,\langle\,,\rangle_{\widehat{A}}) and (B,,B^)(B,\langle\,,\rangle_{\widehat{B}}) are objects in 𝒞\mathcal{C}, and if F:ABF:A\to B is a group homomorphism, then the transpose of FF, the map Ft:B^A^F^{t}:\widehat{B}\to\widehat{A}, pushes the pairing on A^\widehat{A} forward to a pairing on B^\widehat{B} by precomposition with FtFtF^{t}\otimes F^{t}. A morphism between (A,,A^)(A,\langle\,,\rangle_{\widehat{A}}) and (B,,B^)(B,\langle\,,\rangle_{\widehat{B}}) in 𝒞\mathcal{C} is a group homomorphism between AA and BB that pushes the pairing ,A^\langle\,,\rangle_{\widehat{A}} on A^\widehat{A} forward to the pairing ,B^\langle\,,\rangle_{\widehat{B}} on B^\widehat{B}. The fact that FtF^{t} pushes the pairing on A^\widehat{A} forward to a pairing on B^\widehat{B} is the primary reason for considering pairings on the dual groups. Suppressing the symbols for the pairings on A^\widehat{A} and B^\widehat{B} from our notation, let Sur(A,B)\operatorname{\mathrm{Sur}}^{*}(A,B) denote the set of surjective homomorphisms from AA to BB taking ,A^\langle\,,\rangle_{\widehat{A}} to ,B^\langle\,,\rangle_{\widehat{B}}.

The following gives the moments of random sandpile groups with their pairings.

Theorem 1.2.

Let GG be a finite abelian group, and let ,G^\langle\,,\rangle_{\widehat{G}} be a symmetric pairing on its dual, G^=Hom(G,/)\widehat{G}=\operatorname{\mathrm{Hom}}(G,\mathbb{Q}/\mathbb{Z}). Then for a random graph ΓG(n,q)\Gamma\in G(n,q) with (SΓ,,SΓ^)(S_{\Gamma},\langle\,,\rangle_{\widehat{S_{\Gamma}}}) its sandpile group and symmetric pairing on SΓ^\widehat{S_{\Gamma}}, we have

limn𝔼(#Sur(SΓ,G))=1|G|.\lim_{n\to\infty}\mathbb{E}(\#\operatorname{\mathrm{Sur}}^{*}(S_{\Gamma},G))=\frac{1}{|G|}.

We call 𝔼(#Sur(SΓ,G))\mathbb{E}(\#\operatorname{\mathrm{Sur}}^{*}(S_{\Gamma},G)) the (G,,G^)(G,\langle\,,\rangle_{\widehat{G}})-moment of (SΓ,,SΓ^)(S_{\Gamma},\langle\,,\rangle_{\widehat{S_{\Gamma}}}). The error term for Theorem 1.2 decreases exponentially in nn (see Theorem 8.2).

1.1. Sandpile groups and their pairings

We refer the reader to “What is… a sandpile?” [37] for an overview of sandpile groups. The article [48] also gives an outline of the ways in which sandpile groups are connected to various disparate subjects. While there are many equivalent definitions of the sandpile group of a graph Γ\Gamma, perhaps the easiest is the following. Suppose Γ\Gamma is a simple graph with nn vertices labeled by the first nn integers. The Laplacian of Γ\Gamma, denoted LΓL_{\Gamma}, is the n×nn\times n matrix whose (i,j)(i,j) entry is given by

(LΓ)ij={1ifijand{i,j}isanedgeofΓ;0ifijand{i,j}isnotanedgeofΓ;deg(i)ifi=j.(L_{\Gamma})_{ij}=\begin{cases}1&\mathrm{if\ }i\neq j\ \mathrm{and\ }\{i,j\}\ \mathrm{is\ an\ edge\ of\ }\Gamma;\\ 0&\mathrm{if\ }i\neq j\ \mathrm{and\ }\{i,j\}\ \mathrm{is\ not\ an\ edge\ of\ }\Gamma;\\ -\deg(i)&\mathrm{if\ }i=j.\end{cases}

The reduced Laplacian of Γ\Gamma, denoted ΔΓ\Delta_{\Gamma}, is the (n1)×(n1)(n-1)\times(n-1) matrix given by deleting the final row and column from LΓL_{\Gamma}. We define the sandpile group of Γ\Gamma, denoted SΓS_{\Gamma}, to be the cokernel of the reduced Laplacian ΔΓ\Delta_{\Gamma} of Γ\Gamma. We remark here that the sandpile group does not depend on the row or column removed from LΓL_{\Gamma}—the cokernels of the matrices given by deleting any row or column of LΓL_{\Gamma} are all isomorphic (see [10, 18]). From this definition, it follows immediately that if Γ\Gamma is connected, then |SΓ|=|det(ΔΓ)||S_{\Gamma}|=|\det(\Delta_{\Gamma})|; Kirchhoff’s matrix tree theorem implies that |SΓ||S_{\Gamma}| is the number of spanning trees of Γ\Gamma. Alternatively, if ZnZ\subset\mathbb{Z}^{n} denotes the subspace of vectors whose coordinates sum to 0, then it is equivalent to define SΓS_{\Gamma} as SΓ=Z/LΓS_{\Gamma}=Z/L_{\Gamma}\mathbb{Z}.

Sandpile groups were first studied in 1988 by statistical physicists interested in the dynamics of a sandpile [4]; thus originating the name “sandpile.” The sandpile model is an example of a chip-firing game (see [12]): given a graph Γ\Gamma, we imagine that on each vertex is some number of chips; if a vertex has at least as many chips as its degree, the chips topple, distributing a chip to each neighboring vertex. Supposing that chips fall randomly onto the vertices of Γ\Gamma, toppling whenever they can, the sandpile group of Γ\Gamma is given by the recurrent states of the resulting Markov chain. For more in depth treatments, we refer the reader to [22, 25, 26, 11]. There is also a nice connection between sandpile groups and the Tutte polynomial of a graph; see [39, 25, 26]. An overview of these connections can be found in [31].

The sandpile group of a graph has also been referred to in the literature as the Jacobian (also the Picard group or critical group) of the graph. This moniker originates from a beautiful analogy between Riemann surfaces and graphs, in which the sandpile group corresponds to the Jacobian of a Riemann surface [10, 1]. This analogy is not purely philosophical and is spelled out in detail in [40, 13], where the group of components of the Néron model of the Jacobian of a curve over a local field is realized as the sandpile group of a graph. Moreover, the cardinality of the sandpile group occurs in the “analytic class number formula” for graphs [32], and there also exist analogs of the Riemann-Roch and Riemann-Hurwitz formulas for Jacobians of graphs [5, 6].

Sandpile groups of connected graphs are canonically equipped with a perfect, symmetric bilinear pairing (see [41, 13]), which is given by any generalized inverse of the Laplacian (see [51]). If the graph is not connected, then only the torsion subgroup of the sandpile group carries such a duality pairing. We will show in Section 3 that the duality pairing on SΓ^\widehat{S_{\Gamma}} induced by the pairing on SΓS_{\Gamma} can be described directly using the Laplacian of the graph. In the Riemann surface analogy, the pairing on SΓS_{\Gamma} has been considered as an analog of the Weil pairing [51] and also agrees with a pairing defined by Grothendieck [29, 30] giving an obstruction to extending the Poincaré bundle to the product of an abelian variety and its dual [13].

1.2. The Cohen-Lenstra Heuristics

For a family of random finite abelian groups, a natural first guess for how the groups should be distributed is the Cohen-Lenstra heuristics [17], which conjecture the distribution of ideal class groups of quadratic number fields. Ideal class groups are finite abelian groups measuring the failure of unique factorization in rings of integers of number fields (e.g., [5]\mathbb{Z}[\sqrt{-5}]). The underlying philosophy behind the Cohen-Lenstra heuristics is that objects “in nature” often occur with frequency that is inversely proportional to their number of automorphisms. Hence, in the case of finite abelian groups, given a random finite abelian group XX and a fixed finite abelian group GG, a priori we expect (XG)\operatorname{\mathbb{P}}(X\simeq G) to be inversely proportional to |Aut(G)||\operatorname{\mathrm{Aut}}(G)|.

However, Wood showed in [54] that sandpile groups of random graphs do not satisfy a Cohen-Lenstra heuristic. Instead, Wood proved that they obey a “Cohen-Lenstra-like” distribution, in which the canonical duality pairing on the group influences the way in which they are distributed. Notably, Wood’s result is what we obtain by summing the conjectural distribution of random sandpile groups with their pairings due to Clancy, Leake, and Payne [16] over all possible pairings on a given group. Similarly, Tate-Shafarevich groups of elliptic curves are a class of conjecturally finite and abelian groups that, if finite, also come equipped with a nondegenerate, bilinear pairing. Unlike sandpile groups, the pairing on Tate-Shafarevich groups is alternating rather than symmetric. Delaunay developed heuristics [21, 8] similar to those of Clancy, Leake, and Payne in which the automorphisms of GG preserving the pairing take the place of Aut(G)\operatorname{\mathrm{Aut}}(G).

While the distribution of sandpile groups of random graphs with their pairings is intriguing in its own right, our methods may give insight into how to approach other similar problems in both random group theory and arithmetic statistics. Recently, Lipnowski, Sawin, and Tsimerman showed in [38] that, in some cases, class groups of quadratic number fields come equipped with a bilinear structure closely related to the Weil pairing on abelian varieties and the aforementioned Cassels-Tate pairing on Tate-Shafarevich groups. It may be possible that the methods developed in this paper to deal with the pairing on the sandpile group can be used to determine the distribution of other families of random groups with pairings.

1.3. Connections to random matrices

The Laplacian is a matrix with integral entries, so if Γ\Gamma is an Erdős-Rényi random graph, we inevitably find ourselves working in the realm of random matrix theory, a field in which significant effort has been devoted to finding (asymptotically) universal statistics of random matrices. In other words, it is often the case that some statistics are asymptotically equal for a broad class of random matrices, which often contain matrices from especially nice (e.g., symmetric) distributions for which we can actually explicitly compute these statistics. For example, a considerable amount of effort has been dedicated to finding such universality results for the eigenvalue distributions of random matrices: see [43, 27, 2, 3, 49, 52, 53].

In our scenario, we would like to find the distribution of the Sylow pp-subgroups of the cokernels of random integral matrices with their natural duality pairings on the torsion subgroups of these cokernels. In Section 3, we will recall a theorem of Bosch and Lorenzini stating that the torsion subgroup of the cokernel of any integral symmetric matrix is naturally equipped with a duality pairing. For a random symmetric matrix with pp-adic integral entries, drawn with respect to Haar measure on p\mathbb{Z}_{p}, Clancy, Kaplan, Leake, Payne, and Wood showed in [15] that the cokernel of this matrix with its duality pairing are distributed as in Theorem 1.1. These Haar-random matrices should be viewed as the matrices that are “nicely distributed”—because Haar measure on p\mathbb{Z}_{p} is well-understood, we can actually compute the distribution of the cokernels with their pairings. A similar computation is done in [24] for the cokernels (sans pairings) of random (not necessarily symmetric) pp-adic integral matrices and in [8] for Haar-random alternating matrices.

Because we can compute the aforementioned distribution in the nice Haar case, a natural next line of inquiry is to see whether these asymptotic cokernel and pairing distributions are universal—whether other large families of random symmetric matrices have the same cokernel and pairing distributions. In particular, Theorem 1.3 is a strong universality result for distributions of cokernels of random symmetric integral matrices with their pairings. The proof of Theorem 1.3 is essentially the same as Theorem 1.1.

Theorem 1.3.

Let 0<α<10<\alpha<1 be a real number, and let pp be a prime. Let MM be a symmetric random matrix in Mn()M_{n}(\mathbb{Z}) with independent entries mijm_{ij} for 1ijn1\leq i\leq j\leq n. Suppose also that for any t/pt\in\mathbb{Z}/p\mathbb{Z} the probability (mijtmodp)1α\operatorname{\mathbb{P}}(m_{ij}\equiv t\ \mathrm{mod\ }p)\leq 1-\alpha. Let col(M)\operatorname{\mathrm{col}}(M) denote the column space of MM in n\mathbb{Z}^{n}, let cok(M)p\mathrm{cok}(M)_{p} denote the Sylow pp-subgroup of cok(M)=n/col(M)\mathrm{cok}(M)=\mathbb{Z}^{n}/\operatorname{\mathrm{col}}(M), and let ,tcok(M)\langle\,,\rangle_{\mathrm{tcok}(M)} denote the canonical duality pairing on the torsion subgroup of cok(M)\mathrm{cok}(M) (see Section 3). For any finite abelian pp-group GG equipped with a duality pairing ,G\langle\,,\rangle_{G}, we have

limn((cok(M)p,,tcok(M),p)(G,,G))=k1(1p12k)|G||Aut(G,,G)|,\lim_{n\to\infty}\operatorname{\mathbb{P}}((\mathrm{cok}(M)_{p},\langle\,,\rangle_{\mathrm{tcok}(M),p})\simeq(G,\langle\,,\rangle_{G}))=\frac{\prod_{k\geq 1}(1-p^{1-2k})}{|G||\operatorname{\mathrm{Aut}}(G,\langle\,,\rangle_{G})|},

where ,tcok(M),p\langle\,,\rangle_{\mathrm{tcok}(M),p} denotes the restriction of ,tcok(M)\langle\,,\rangle_{\mathrm{tcok}(M)} to cok(M)p\mathrm{cok}(M)_{p}.

As in the case with the sandpile group, we are able to show a stronger result than Theorem 1.3—for any finite set of primes, we give the asymptotic probabilities of particular Sylow subgroups with their pairings at these primes.

Recently, there has been significant effort devoted to finding universal distributions of various classes of random groups. In 2019, Wood showed that random integral matrices with independent entries are distributed according to a Cohen-Lenstra heuristic [55], and in [57], Yan extended this result to random matrices with independent entries taking values in Dedekind domains with finite quotients. In 2020, Mézáros gave the distribution of the sandpile groups of random regular graphs in [44]. Koplewitz studied the pp-rank distribution of sandpile groups of random bipartite graphs in [33], and Bhargava, DePascale, and Koenig considered the directed bipartite case in [7]. Nguyen and Wood wrote several papers [46, 47] in which they study several other related problems on the distribution of cokernels of certain classes of random matrices. In particular, they found in [47] the asymptotic probability that the sandpile group of an Erdős-Rényi random graph is cyclic, answering a long-open conjecture due to Lorenzini [42]. Lee found the distribution of cokernels of random Hermitian matrices with entries in the ring of integers of a quadratic extension of p\mathbb{Q}_{p} [36] and proved results of similar flavor in [35, 34]. In 2022, Nguyen and Van Peski found in [45] the distribution of the cokernel of the product of random integral square matrices, and in 2023 Cheong and Yu proved a similar result in which they found the distribution of the cokernel of a polynomial evaluated at a random pp-adic square matrix [14]. While much of the work in this area has been devoted to studying universal distributions of abelian groups, there has also been interest in the nonabelian case. For example, Gorokhovsky proved analogous universality results for random quotients of the free group on nn generators in [28], and Lee studied the distributions of similar random nonabelian groups in [34].

1.4. Our methods

The overall structure of our argument is based off of the seminal methods developed by Wood in [54]. For a more detailed outline of the overall procedure, we refer the reader to the introduction of [54] and instead focus on the new techniques developed to deal with the pairing. The most important of these techniques is our usage of duality. As mentioned previously, we will primarily be working with finitely generated abelian groups with symmetric pairings attached to their dual groups. We will convert results about pairings on duals into statements about pairings on the groups themselves in Section 10. In Section 3, we attach a natural symmetric pairing to the Pontryagin dual of the cokernel of a symmetric integral matrix and show that, when the matrix is LΓL_{\Gamma}, the pairing agrees with the duality pairing on SΓ^\widehat{S_{\Gamma}}. To show Theorem 1.2, we compute the moments in a far more general setting where a random matrix as in Theorem 1.3 replaces the Laplacian of a random graph (see Theorem 8.1). Then, using results due to Sawin and Wood in [50], we prove (in Section 9) that these moments in fact determine a unique distribution. Thus, we may use the nice case of cokernels of uniform random symmetric matrices over /a\mathbb{Z}/a\mathbb{Z} with their pairings, along with our universality results, to prove the distribution in the more general case (the distribution in the nice case was computed in [15]).

While the method of proof is inspired by [54], the extra data of the pairing complicates some of the proofs significantly. Indeed, it is not immediately clear that we should consider pairings on the duals of our groups or what the moments of the groups with pairings should be. Thus, to compute the averages Sur(SΓ,G)\operatorname{\mathrm{Sur}}^{*}(S_{\Gamma},G), we first translate the event that the pairing on SΓ^\widehat{S_{\Gamma}} pushes forward onto the pairing on G^\widehat{G} into a piece of algebraic data which is easier to work with. In particular, if F:SΓGF:S_{\Gamma}\to G is a surjection, the event that FtF^{t} takes the pairing on SΓ^\widehat{S_{\Gamma}} to the pairing on G^\widehat{G} occurs if and only if FLΓFtFL_{\Gamma}F^{t} is equal to some fixed symmetric homomorphism from G^G\widehat{G}\to G. Then since we are trying to compute the average number of surjections from SΓGS_{\Gamma}\to G satisfying this property, and since the sandpile group is Z/LΓZZ/L_{\Gamma}Z, it suffices to determine the probability that a surjection F:ZGF:Z\to G descends to SΓS_{\Gamma} and pushes the pairing on SΓ^\widehat{S_{\Gamma}} forward to the pairing on G^\widehat{G}. Equivalently, we determine the probability that FLΓ=0FL_{\Gamma}=0 and that FLΓFtFL_{\Gamma}F^{t} is equal to some fixed symmetric element of Hom(G^,G)\operatorname{\mathrm{Hom}}(\widehat{G},G).

We regard this as a system of linear equations in the entries of LΓL_{\Gamma}, and we note that there are on the order of nn equations in (n2){{n}\choose{2}} variables. While in [54] it was enough to check whether FLΓ=0FL_{\Gamma}=0, the pairing data makes this system particularly complex not only because the equations are more complicated, but because the equations given by FLΓFtFL_{\Gamma}F^{t} are not even well-defined unless FLΓ=0FL_{\Gamma}=0. To work around this inherent dependence between our equations, we choose a lift of F:ZGF:Z\to G to a larger group HH and modify our equations slightly so that we may work exclusively in this new group. Denote the lift by \mathcal{F}. Although this construction seems artificial, it turns out that all of our arguments depend only on FF itself rather than the specific choice of lift.

Our system of equations is parameterized by Hom(Z,G^)×Sym2H^\operatorname{\mathrm{Hom}}(Z,\widehat{G})\times\operatorname{\mathrm{Sym}}^{2}\widehat{H}. Some pairs (C,D)Hom(Z,G^)×Sym2H^(C,D)\in\operatorname{\mathrm{Hom}}(Z,\widehat{G})\times\operatorname{\mathrm{Sym}}^{2}\widehat{H}, give equations in which all of the coefficients are 0; we refer to these as special pairs. The set of special pairs (C,D)(C,D) depends implicitly on FF and the chosen lift \mathcal{F}. Ultimately, we are interested in the structural properties of \mathcal{F} and CC that influence the number of nonzero coefficients (see Section 5). We adapt two concepts from Wood’s work in [54], depth and robustness, to describe this structure. Finally, we count the number of special (C,D)(C,D) for each FF. The special (C,D)(C,D) give the main term in Theorem 8.2 (the limit in Theorem 1.2), while the remaining cases contribute to the error term, which we bound by adapting techniques from [54].

Lastly, because the diagonal entries of LΓL_{\Gamma} depend on the off-diagonal ones, we run the program described above for a matrix with independent diagonal entries. Then we enlarge FF by taking its direct sum with another map that detects whether the column sums of a matrix are 0. This allows us to condition on what we require of the diagonal. However, in order to apply our methods to the enlarged group we must also extend the pairing on G^\widehat{G} to the bigger group and sum over all possible extensions.

1.5. Outline of the paper

In Section 3, for a symmetric integral matrix MM, we define a symmetric pairing on cok^(M)\mathrm{c}\widehat{\mathrm{ok}}(M) and show that this induces a duality pairing on the torsion subgroup of cok(M)\mathrm{cok}(M). The induced duality pairing is equal to the duality pairing on the torsion part of cok(M)\mathrm{cok}(M) introduced by Bosch and Lorenzini in [13]. Theorem 1.2 (along with an analog of this theorem for cokernels of random symmetric matrices with their pairings) is proved in Sections 4, 5, 6, 7 and 8. In Section 9, we state several results from [50] and use these to prove that the moments of Theorem 1.2 in fact determine the distribution in question. Finally, in Section 10, we compare to the case of uniform random symmetric matrices over /a\mathbb{Z}/a\mathbb{Z}; Theorem 1.1 follows from Corollary 10.5.

2. Background

2.1. Finite abelian groups

Let pp be a prime. A finite abelian pp-group is isomorphic to

i=1r/pλi\bigoplus_{i=1}^{r}\mathbb{Z}/p^{\lambda_{i}}\mathbb{Z}

for some positive integers λ1λ2λr\lambda_{1}\geq\lambda_{2}\geq\cdots\geq\lambda_{r}. If λ\lambda is the partition given by the λi\lambda_{i}’s, we call λ\lambda the type of our finite abelian pp-group. When pp is understood, let GλG_{\lambda} denote the pp-group of type λ\lambda.

We may view any two abelian groups GG and HH as \mathbb{Z}-modules, and we may consider their tensor product GHG\otimes_{\mathbb{Z}}H. When G=i/aiG=\bigoplus_{i}\mathbb{Z}/a_{i}\mathbb{Z} is finite and H=/aH=\mathbb{Z}/a\mathbb{Z},

GH=(i/ai)/a=i/(a,ai),G\otimes_{\mathbb{Z}}H=\left(\bigoplus_{i}\mathbb{Z}/a_{i}\mathbb{Z}\right)\otimes_{\mathbb{Z}}\mathbb{Z}/a\mathbb{Z}=\bigoplus_{i}\mathbb{Z}/(a,a_{i})\mathbb{Z},

where (a,ai)(a,a_{i}) denotes the greatest common divisor of aa and aia_{i}. The fact that tensor products distribute over direct sums in conjunction with the classification of finite abelian groups allows us to compute GHG\otimes H for arbitrary finite abelian groups GG and HH. In particular, if GpG_{p} denotes the Sylow pp-subgroup of GG, then GGpGpGp.G\otimes G\simeq\bigoplus_{p}G_{p}\otimes G_{p}. Moreover, if GpG_{p} is of type λ\lambda and is generated by gig_{i} with relations pλigi=0p^{\lambda_{i}}g_{i}=0, then GpGpG_{p}\otimes G_{p} is generated as an abelian group by gigjg_{i}\otimes g_{j} with relations pmin(λi,λj)gigj=0p^{\min(\lambda_{i},\lambda_{j})}g_{i}\otimes g_{j}=0. Thus,

GpGpi(/pλi)2i1.G_{p}\otimes G_{p}\simeq\bigoplus_{i}\left(\mathbb{Z}/p^{\lambda_{i}}\mathbb{Z}\right)^{2i-1}.

Now, GGG\otimes G comes equipped with a natural involution taking ghhgg\otimes h\mapsto h\otimes g. Denote the subgroup of elements fixed under this involution by Sym2G\operatorname{\mathrm{Sym}}_{2}G. For GpG_{p} as above, we have that Sym2GpSym2Gp\operatorname{\mathrm{Sym}}_{2}G\simeq\bigoplus_{p}\operatorname{\mathrm{Sym}}_{2}G_{p} and that Sym2Gp\operatorname{\mathrm{Sym}}_{2}G_{p} is the abelian group generated by the elements gigig_{i}\otimes g_{i} and gigj+gjgig_{i}\otimes g_{j}+g_{j}\otimes g_{i} for i<ji<j with relations pλjgigj=pλjgjgi=0p^{\lambda_{j}}g_{i}\otimes g_{j}=p^{\lambda_{j}}g_{j}\otimes g_{i}=0 for each i<ji<j. Thus, |Sym2Gp|=piiλi|\operatorname{\mathrm{Sym}}_{2}G_{p}|=p^{\sum_{i}i\lambda_{i}}.

The exterior power 2G\wedge^{2}G is the quotient of GGG\otimes G by the subgroup generated by elements of the form ggg\otimes g. For GpG_{p} as above, we have that 2Gp2Gp\wedge^{2}G\simeq\bigoplus_{p}\wedge^{2}G_{p} and that 2Gp\wedge^{2}G_{p} is generated as an abelian group by gigjg_{i}\wedge g_{j} for i<ji<j with relations pλjgigj=0p^{\lambda_{j}}g_{i}\wedge g_{j}=0 for each i<ji<j. Thus,

2Gpi(/pλi)i1.\wedge^{2}G_{p}\simeq\bigoplus_{i}\left(\mathbb{Z}/p^{\lambda_{i}}\mathbb{Z}\right)^{i-1}.

Likewise, the symmetric power Sym2G\operatorname{\mathrm{Sym}}^{2}G is the quotient of GGG\otimes G by the subgroup generated by elements of the form ghhgg\otimes h-h\otimes g. For GpG_{p} as above, we have that Sym2GpSym2Gp\operatorname{\mathrm{Sym}}^{2}G\simeq\bigoplus_{p}\operatorname{\mathrm{Sym}}^{2}G_{p} and that Sym2Gp\operatorname{\mathrm{Sym}}^{2}G_{p} is generated as an abelian group by gigjg_{i}g_{j} for iji\leq j with relations pλjgigj=0p^{\lambda_{j}}g_{i}g_{j}=0 for iji\leq j. Therefore,

Sym2Gpi(/pλi)i.\operatorname{\mathrm{Sym}}^{2}G_{p}\simeq\bigoplus_{i}\left(\mathbb{Z}/p^{\lambda_{i}}\mathbb{Z}\right)^{i}.

The exponent of a finite abelian group is the smallest positive integer ee such that eG=0eG=0. If R=/aR=\mathbb{Z}/a\mathbb{Z}, then a finite abelian group GG with exponent dividing aa is also an RR-module. If HH is another such group, then HomR(G,H)=Hom(G,H)\operatorname{\mathrm{Hom}}_{R}(G,H)=\operatorname{\mathrm{Hom}}_{\mathbb{Z}}(G,H) (i.e., the RR-module homomorphisms GHG\to H are the same as the group homomorphisms GHG\to H).

For any finite abelian group GG, its Pontryagin dual is the group Hom(G,/)\operatorname{\mathrm{Hom}}_{\mathbb{Z}}(G,\mathbb{Q}/\mathbb{Z}), which we denote by G^\widehat{G}.

2.2. Pairings

Let GG be an abelian group and consider a map ϕ:G×G/\phi:G\times G\to\mathbb{Q}/\mathbb{Z}. We say that ϕ\phi is symmetric if ϕ(g,h)=ϕ(h,g)\phi(g,h)=\phi(h,g) and bilinear if ϕ(g1+g2,h)=ϕ(g1,h)+ϕ(g2,h)\phi(g_{1}+g_{2},h)=\phi(g_{1},h)+\phi(g_{2},h) and ϕ(g,h1+h2)=ϕ(g,h1)+ϕ(g,h2)\phi(g,h_{1}+h_{2})=\phi(g,h_{1})+\phi(g,h_{2}). If ϕ\phi is bilinear, then it is called a pairing on GG. Equivalently, a pairing on GG can be thought of as a homomorphism GG/G\otimes G\to\mathbb{Q}/\mathbb{Z}. Given a pairing ϕ\phi and generators g1,,grg_{1},\ldots,g_{r} of GG, note that ϕ\phi is entirely determined by the values ϕ(gi,gj)\phi(g_{i},g_{j}) for all pairs (i,j)(i,j).

A symmetric pairing ϕ\phi on GG induces a map from GHom(G,/)G\to\operatorname{\mathrm{Hom}}(G,\mathbb{Q}/\mathbb{Z}) taking gϕ(g,)g\mapsto\phi(g,\cdotp) (if the pairing were not symmetric we would have two maps—one for each factor). Such a pairing is said to be nondegenerate if the induced map GHom(G,/)G\to\operatorname{\mathrm{Hom}}(G,\mathbb{Q}/\mathbb{Z}) is injective. If the induced map is not injective, the pairing is said to be degenerate, and the kernel of the map is called the kernel of the pairing. The cokernel of the pairing ϕ\phi, denoted coker(ϕ)\mathrm{coker}(\phi), is defined to be G/ker(ϕ)G/\ker(\phi). Note that there is a natural nondegenerate symmetric pairing on coker(ϕ)\mathrm{coker}(\phi) induced by ϕ\phi.

If the map GG^G\to\widehat{G} induced by ϕ\phi is an isomorphism, then the pairing is said to be perfect. Recall that a perfect symmetric pairing on a group is called a duality pairing. If GG is finite, then a nondegenerate pairing on GG is automatically perfect (this fact follows from counting). Given two groups GG and HH equipped with pairings ,G\langle\,,\rangle_{G} and ,H\langle\,,\rangle_{H}, there is a notion of isomorphism between (G,,G)(G,\langle\,,\rangle_{G}) and (H,,H)(H,\langle\,,\rangle_{H}): an isomorphism between (G,,G)(G,\langle\,,\rangle_{G}) and (H,,H)(H,\langle\,,\rangle_{H}) is an isomorphism of groups f:GHf:G\to H such that f1(h),f1(h)G=h,hH\langle f^{-1}(h),f^{-1}(h^{\prime})\rangle_{G}=\langle h,h^{\prime}\rangle_{H} for all pairs (h,h)H×H(h,h^{\prime})\in H\times H. If (G,,G)(G,\langle\,,\rangle_{G}) is a group equipped with a pairing, an automorphism of GG is said to respect the pairing if it is an isomorphism of groups and pairings between (G,,G)(G,\langle\,,\rangle_{G}) and itself. Denote the set of automorphisms of GG that respect the pairing ,G\langle\,,\rangle_{G} by Aut(G,,G)\operatorname{\mathrm{Aut}}(G,\langle\,,\rangle_{G}).

2.3. Duality

Let R=/bR=\mathbb{Z}/b\mathbb{Z}. When the ring RR is understood, for an RR-module GG, we may consider its dual, G=HomR(G,R)G^{*}=\operatorname{\mathrm{Hom}}_{R}(G,R). If GG is a finite abelian group, there is a canonical isomorphism between GG^{*} and G^\widehat{G}. For finite RR-modules GG and HH, we have (GH)GH(G\oplus H)^{*}\simeq G^{*}\oplus H^{*}. While GG and GG^{*} are not naturally isomorphic, when GG is finite, there is a canonical isomorphism between GG and (G)(G^{*})^{*} given by evaluation.

For a finite abelian group GG, the following are equivalent data:

  1. (1)

    a perfect symmetric pairing on GG;

  2. (2)

    an isomorphism from GG to G^\widehat{G};

  3. (3)

    a perfect symmetric pairing on G^\widehat{G}.

Given a duality pairing ,G\langle\,,\rangle_{G} on GG, let ϕ\phi denote the induced isomorphism from GG^G\to\widehat{G}. Then the duality pairing on G^\widehat{G} is given by x,y=ϕ1(x),ϕ1(y)G\langle x,y\rangle=\langle\phi^{-1}(x),\phi^{-1}(y)\rangle_{G}.

Let GG and HH be two finite abelian groups. Let f:GHf:G\to H be a group homomorphism. The transpose of ff is the map ft:H^G^f^{t}:\widehat{H}\to\widehat{G} given by precomposing Hom(H,/)\ell\in\operatorname{\mathrm{Hom}}(H,\mathbb{Q}/\mathbb{Z}) with ff (a similar definition can be made if GG and HH are RR-modules for R=/bR=\mathbb{Z}/b\mathbb{Z} and ff an RR-module homomorphism). If G^\widehat{G} is equipped with a pairing ,G^\langle\,,\rangle_{\widehat{G}}, then ftf^{t} pushes ,G^\langle\,,\rangle_{\widehat{G}} forward to a pairing ,H^\langle\,,\rangle_{\widehat{H}} on H^\widehat{H} given by

a,bH^:=ft(a),ft(b)G^.\langle a,b\rangle_{\widehat{H}}:=\langle f^{t}(a),f^{t}(b)\rangle_{\widehat{G}}.

In other words, ftf^{t} induces a map from Hom(G^G^,/)Hom(H^H^,/)\operatorname{\mathrm{Hom}}(\widehat{G}\otimes\widehat{G},\mathbb{Q}/\mathbb{Z})\to\operatorname{\mathrm{Hom}}(\widehat{H}\otimes\widehat{H},\mathbb{Q}/\mathbb{Z}); let ft(,G^)f^{t}(\langle\,,\rangle_{\widehat{G}}) denote the image of ,G^\langle\,,\rangle_{\widehat{G}} under this map (denoted ,H^\langle\,,\rangle_{\widehat{H}} above). Note that we do not require ,G^\langle\,,\rangle_{\widehat{G}} to be perfect.

Let (G,,G^)(G,\langle\,,\rangle_{\widehat{G}}) and (H,,H^)(H,\langle\,,\rangle_{\widehat{H}}) be groups whose duals are equipped with pairings. Define

Hom((G,,G^),(H,,H^))\operatorname{\mathrm{Hom}}((G,\langle\,,\rangle_{\widehat{G}}),(H,\langle\,,\rangle_{\widehat{H}}))

to be the set of group homomorphisms f:GHf:G\to H such that ft(,G^)=,H^f^{t}(\langle\,,\rangle_{\widehat{G}})=\langle\,,\rangle_{\widehat{H}}. In the category 𝒞\mathcal{C} defined in Section 1, Hom((G,,G^),(H,,H^))\operatorname{\mathrm{Hom}}((G,\langle\,,\rangle_{\widehat{G}}),(H,\langle\,,\rangle_{\widehat{H}})) is the set of morphisms between (G,,G^)(G,\langle\,,\rangle_{\widehat{G}}) and (H,,H^)(H,\langle\,,\rangle_{\widehat{H}}). When the pairings on both groups are understood, we denote Hom((G,,G^),(H,,H^))\operatorname{\mathrm{Hom}}((G,\langle\,,\rangle_{\widehat{G}}),(H,\langle\,,\rangle_{\widehat{H}})) by Hom(G,H)\operatorname{\mathrm{Hom}}^{*}(G,H) for the sake of brevity. Note that this generalizes the notion of isomorphism between groups with duality pairings defined in the previous subsection. Moreover, given finite abelian groups GG and HH with duality pairings ,G\langle\,,\rangle_{G} and ,H\langle\,,\rangle_{H} and induced duality pairings ,G^\langle\,,\rangle_{\widehat{G}} and ,H^\langle\,,\rangle_{\widehat{H}} on the dual groups, note that (G,,G)(H,,H)(G,\langle\,,\rangle_{G})\simeq(H,\langle\,,\rangle_{H}) if and only if (G,,G^)(H,,H^)(G,\langle\,,\rangle_{\widehat{G}})\simeq(H,\langle\,,\rangle_{\widehat{H}}).

2.4. Cokernels of matrices

Let Mn(R)M_{n}(R) denote the space of n×nn\times n matrices with entries in a commutative ring RR. For MMn(R)M\in M_{n}(R), let col(M)\operatorname{\mathrm{col}}(M) denote the column space of MM (this is the image of the map RnRnR^{n}\to R^{n} given by MM). Define the cokernel of MM to be

cok(M)=Rn/col(M).\mathrm{cok}(M)=R^{n}/\operatorname{\mathrm{col}}(M).

For MMn()M\in M_{n}(\mathbb{Z}), regard MM as a linear map M:nnM:\mathbb{Z}^{n}\to\mathbb{Z}^{n}. Then cok(M)\mathrm{cok}(M) is a finitely generated abelian group; let tcok(M)\mathrm{tcok}(M) denote its torsion subgroup. For any subgroup YnY\subset\mathbb{Z}^{n}, let YY^{\perp} denote its orthogonal with respect to the standard scalar product on n\mathbb{Z}^{n}.

If MM is symmetric, col(M)ker(M)\operatorname{\mathrm{col}}(M)\subset\ker(M)^{\perp}. Because rank(col(M))=rank(ker(M))\mathrm{rank}(\operatorname{\mathrm{col}}(M))=\mathrm{rank}(\ker(M)^{\perp}) and because ker(M)\ker(M)^{\perp} is a saturated submodule of n\mathbb{Z}^{n} (i.e., if rxker(M)rx\in\ker(M)^{\perp} for rr\in\mathbb{Z} and xnx\in\mathbb{Z}^{n}, then xker(M)x\in\ker(M)^{\perp}), it follows that

tcok(M)=ker(M)/col(M).\mathrm{tcok}(M)=\ker(M)^{\perp}/\operatorname{\mathrm{col}}(M).

Since tcok(M)\mathrm{tcok}(M) is pure torsion and cok(M)\mathrm{cok}(M) is finitely generated, tcok(M)\mathrm{tcok}(M) is finite.

2.5. Sandpile groups

Recall the definition of the graph Laplacian. Let [n]={1,,n}[n]=\{1,\ldots,n\}, and let Γ\Gamma be a graph on nn vertices labeled by [n][n]. The Laplacian of Γ\Gamma, denoted LΓL_{\Gamma}, is the n×nn\times n matrix whose (i,j)(i,j) entry is given by

(LΓ)ij={1ifijand{i,j}isanedgeofΓ;0ifijand{i,j}isnotanedgeofΓ;deg(i)ifi=j.(L_{\Gamma})_{ij}=\begin{cases}1&\mathrm{if\ }i\neq j\ \mathrm{and\ }\{i,j\}\ \mathrm{is\ an\ edge\ of\ }\Gamma;\\ 0&\mathrm{if\ }i\neq j\ \mathrm{and\ }\{i,j\}\ \mathrm{is\ not\ an\ edge\ of\ }\Gamma;\\ -\deg(i)&\mathrm{if\ }i=j.\end{cases}

Note that LΓL_{\Gamma} is a symmetric matrix in Mn()M_{n}(\mathbb{Z}). If Γ\Gamma is connected, then LΓL_{\Gamma} has rank n1n-1 and kernel spanned by the all-ones vector (see [51]). If ZnZ\subset\mathbb{Z}^{n} is the subgroup of elements whose coordinates sum to 0, then we see that col(LΓ)Z\operatorname{\mathrm{col}}(L_{\Gamma})\subset Z. Define the sandpile group of Γ\Gamma, denoted SΓS_{\Gamma}, to be Z/col(LΓ)Z/\operatorname{\mathrm{col}}(L_{\Gamma}). We see immediately that SΓS_{\Gamma} must be finitely generated. Moreover, it is finite if and only if Γ\Gamma is connected.

2.6. Random graphs

Let ΓG(n,q)\Gamma\in G(n,q) denote that Γ\Gamma is an Erdős-Rényi random graph on nn labeled vertices with edges occurring independently with probability qq.

2.7. Notation

We use |||\cdot| or #\# to denote the cardinality of sets (we use #\#’s to avoid confusion with the notation || for “divides”; while absolute value signs take up less space). Throughout, \simeq denotes “is isomorphic to,” and \operatorname{\mathbb{P}} and 𝔼\mathbb{E} denote probability and expected value, respectively. Also, pp will always denote a prime. For any integers ii and jj, let δij\delta_{ij} be the Kronecker delta—if i=ji=j, then δij=1\delta_{ij}=1 and if iji\neq j, then δij=0\delta_{ij}=0.

3. Pairings associated to symmetric matrices

It is a well-known fact (see, e.g., [20]) that the cokernel of an invertible symmetric integral matrix MM carries a canonical duality pairing. If x,ycok(M)x,y\in\mathrm{cok}(M), and X,YnX,Y\in\mathbb{Z}^{n} are lifts of x,yx,y respectively, then the pairing is given by

(x,y)XtM1Ymod.(x,y)\mapsto X^{t}M^{-1}Y\mod\mathbb{Z}.

In [13], Bosch and Lorenzini showed that the torsion subgroup of cok(M)\mathrm{cok}(M) is canonically equipped with a duality pairing. When the matrix is invertible, this duality pairing coincides with the one defined above.

We define this canonical duality pairing on tcok(M)\mathrm{tcok}(M) in the following way (we refer the reader to Section 1 of [13] for a more general and detailed setup). Let τ,τtcok(M)\tau,\tau^{\prime}\in\mathrm{tcok}(M), and let TT and TT^{\prime} be lifts of τ\tau and τ\tau^{\prime} to ker(M)\ker(M)^{\perp}, respectively. Since tcok(M)\mathrm{tcok}(M) is pure torsion, there exist k,kk,k^{\prime}\in\mathbb{N} such that kT,kTcol(M)kT,k^{\prime}T^{\prime}\in\operatorname{\mathrm{col}}(M), i.e., there exist S,SnS,S^{\prime}\in\mathbb{Z}^{n} such that MS=kTMS=kT and MS=kTMS^{\prime}=k^{\prime}T^{\prime}. Define a map ,tcok(M):tcok(M)×tcok(M)/\langle\,,\rangle_{\mathrm{tcok}(M)}:\mathrm{tcok}(M)\times\mathrm{tcok}(M)\to\mathbb{Q}/\mathbb{Z} by

(τ,τ)StMSkkmod.(\tau,\tau^{\prime})\mapsto\frac{S^{t}MS^{\prime}}{kk^{\prime}}\mod\mathbb{Z}.

If MM is nonsingular, then cok(M)=tcok(M)\mathrm{cok}(M)=\mathrm{tcok}(M), and this map is the classical pairing on cok(M)\mathrm{cok}(M) discussed above. The map ,tcok(M)\langle\,,\rangle_{\mathrm{tcok}(M)} is a duality pairing on tcok(M)\mathrm{tcok}(M):

Lemma 3.1 (Lemma 1.1 and Theorem 1.3 of [13]).

The map ,tcok(M)\langle\,,\rangle_{\mathrm{tcok}(M)} is well-defined, bilinear, symmetric, and perfect.

When MM is singular, cok(M)\mathrm{cok}(M) has some free part to which ,tcok(M)\langle\,,\rangle_{\mathrm{tcok}(M)} does not naturally extend. The duality pairing on tcok(M)\mathrm{tcok}(M) implies the existence of a corresponding duality pairing on tcok^(M)\mathrm{tc}\widehat{\mathrm{ok}}(M). Now, tcok^(M)\mathrm{tc}\widehat{\mathrm{ok}}(M) is a quotient of cok^(M)\mathrm{c}\widehat{\mathrm{ok}}(M), so the pairing on tcok^(M)\mathrm{tc}\widehat{\mathrm{ok}}(M) does induce a natural pairing on cok^(M)\mathrm{c}\widehat{\mathrm{ok}}(M). However, it can be quite difficult to compute this induced pairing on cok^(M)\mathrm{c}\widehat{\mathrm{ok}}(M) using the definition given above. So, we will define a symmetric pairing on cok^(M)\mathrm{c}\widehat{\mathrm{ok}}(M) given by MM; show that this pairing induces a perfect symmetric pairing on tcok^(M)\mathrm{tc}\widehat{\mathrm{ok}}(M); and prove that the induced duality pairing on tcok(M)\mathrm{tcok}(M) coincides with the pairing of Bosch and Lorenzini.

Recall MM is a linear map nn\mathbb{Z}^{n}\to\mathbb{Z}^{n}. Any ϕcok^(M)\phi\in\mathrm{c}\widehat{\mathrm{ok}}(M) can be extended to ϕ:n/\phi:\mathbb{Z}^{n}\to\mathbb{Q}/\mathbb{Z} via the quotient ncok(M)\mathbb{Z}^{n}\to\mathrm{cok}(M). Note that cok^(M)=Hom(cok(M),/)\mathrm{c}\widehat{\mathrm{ok}}(M)=\operatorname{\mathrm{Hom}}(\mathrm{cok}(M),\mathbb{Q}/\mathbb{Z}) is the subset of maps ϕHom(n,/)\phi\in\operatorname{\mathrm{Hom}}(\mathbb{Z}^{n},\mathbb{Q}/\mathbb{Z}) such that ϕM=0\phi M=0 (i.e., for any lift Φ\Phi of ϕ\phi to Hom(n,)\operatorname{\mathrm{Hom}}(\mathbb{Z}^{n},\mathbb{Q}), we have (ΦM)tn(\Phi M)^{t}\in\mathbb{Z}^{n}). For τcok(M)\tau\in\mathrm{cok}(M), let TT be a lift of τ\tau to n\mathbb{Z}^{n}. Then for ϕcok^(M)\phi\in\mathrm{c}\widehat{\mathrm{ok}}(M), we have ϕ(τ)=ϕ(T)\phi(\tau)=\phi(T). We define a pairing on cok^(M)\mathrm{c}\widehat{\mathrm{ok}}(M) given by the matrix MM in the following way. For x,ycok^(M)x,y\in\mathrm{c}\widehat{\mathrm{ok}}(M), let X,YHom(n,)X,Y\in\operatorname{\mathrm{Hom}}(\mathbb{Z}^{n},\mathbb{Q}) be lifts of x,yx,y, respectively. We define a map ,M:cok^(M)×cok^(M)/\langle\,,\rangle_{M}:\mathrm{c}\widehat{\mathrm{ok}}(M)\times\mathrm{c}\widehat{\mathrm{ok}}(M)\to\mathbb{Q}/\mathbb{Z} by

(x,y)XMYtmod.(x,y)\mapsto XMY^{t}\mod\mathbb{Z}.
Lemma 3.2.

The map ,M\langle\,,\rangle_{M} on cok^(M)\mathrm{c}\widehat{\mathrm{ok}}(M) is well-defined, bilinear, and symmetric.

Proof.

To show that ,M\langle\,,\rangle_{M} is well-defined, we simply need to check that our map does not depend on the chosen lifts of xx and yy. Recall that both MXtMX^{t} and MYtMY^{t} are vectors in n\mathbb{Z}^{n}. Hence, for A,BHom(n,)A,B\in\operatorname{\mathrm{Hom}}(\mathbb{Z}^{n},\mathbb{Z}), we have

(X+A)M(Y+B)t=XMYt+XMBt+AMYt+AMBt=XMYtmod.(X+A)M(Y+B)^{t}=XMY^{t}+XMB^{t}+AMY^{t}+AMB^{t}=XMY^{t}\mod\mathbb{Z}.

Hence, the map is well-defined.

Bilinearity and symmetry are both easily verified, where the symmetry of ,M\langle\,,\rangle_{M} follows from the symmetry of MM itself. ∎

Since the Pontryagin duality functor Hom(,/)\operatorname{\mathrm{Hom}}(-,\mathbb{Q}/\mathbb{Z}) is exact, the inclusion ι:tcok(M)cok(M)\iota:\mathrm{tcok}(M)\hookrightarrow\mathrm{cok}(M) becomes a quotient ιt:cok^(M)tcok^(M)\iota^{t}:\mathrm{c}\widehat{\mathrm{ok}}(M)\twoheadrightarrow\mathrm{tc}\widehat{\mathrm{ok}}(M) after taking the transpose. The kernel of this quotient is the annihilator of tcok(M)\mathrm{tcok}(M), denoted Ann(tcok(M))\operatorname{\mathrm{Ann}}(\mathrm{tcok}(M)). It turns out that ,M\langle\,,\rangle_{M} descends to a well-defined duality pairing on tcok^(M)\mathrm{tc}\widehat{\mathrm{ok}}(M). Recall the definition of coker,M\mathrm{coker}\langle\,,\rangle_{M} from Section 2.2.

Lemma 3.3.

Let MM be a symmetric integral matrix and ,M\langle\,,\rangle_{M} the symmetric pairing on cok^(M)\mathrm{c}\widehat{\mathrm{ok}}(M). Then ker(,M)=Ann(tcok(M))\ker(\langle\,,\rangle_{M})=\operatorname{\mathrm{Ann}}(\mathrm{tcok}(M)), implying coker(,M)=tcok^(M)\mathrm{coker}(\langle\,,\rangle_{M})=\mathrm{tc}\widehat{\mathrm{ok}}(M).

Proof.

Suppose ψker(,M)\psi\in\ker(\langle\,,\rangle_{M}) so that for all xcok^(M)x\in\mathrm{c}\widehat{\mathrm{ok}}(M) we have ψ,xM=0\langle\psi,x\rangle_{M}=0. Let τtcok(M)\tau\in\mathrm{tcok}(M) and choose a lift TT of τ\tau to ker(M)n\ker(M)^{\perp}\subset\mathbb{Z}^{n}. Since τtcok(M)\tau\in\mathrm{tcok}(M), there exist kk\in\mathbb{N} and SnS\in\mathbb{Z}^{n} such that kT=MSkT=MS. Consider the image of (S/k)tHom(n,)(S/k)^{t}\in\operatorname{\mathrm{Hom}}(\mathbb{Z}^{n},\mathbb{Q}) in Hom(n,/)\operatorname{\mathrm{Hom}}(\mathbb{Z}^{n},\mathbb{Q}/\mathbb{Z}), which we denote by [(S/k)t][(S/k)^{t}]. Since (S/k)tM=Tt(S/k)^{t}M=T^{t}, we may view [(S/k)t][(S/k)^{t}] as an element of cok^(M)\mathrm{c}\widehat{\mathrm{ok}}(M). Thus, if Ψ\Psi is any lift of ψ\psi to \mathbb{Q}, then

ψ(τ)=ψ(T)=ψ(MSk)=ΨMSk=ψ,[(S/k)t]M=0mod,\psi(\tau)=\psi(T)=\psi\left(\frac{MS}{k}\right)=\frac{\Psi MS}{k}=\langle\psi,[(S/k)^{t}]\rangle_{M}=0\mod\mathbb{Z},

forcing ψAnn(tcok(M)).\psi\in\operatorname{\mathrm{Ann}}(\mathrm{tcok}(M)).

For the reverse inclusion, assume ϕAnn(tcok(M))\phi\in\operatorname{\mathrm{Ann}}(\mathrm{tcok}(M)) so that for all τtcok(M)\tau\in\mathrm{tcok}(M), we have ϕ(τ)=0\phi(\tau)=0. We would like to show that for all xcok^(M)x\in\mathrm{c}\widehat{\mathrm{ok}}(M)

ϕ,xM=ΦMXt=0mod,\langle\phi,x\rangle_{M}=\Phi MX^{t}=0\mod\mathbb{Z},

where Φ\Phi and XX are lifts of ϕ\phi and xx to \mathbb{Q}, respectively. Suppose xcok^(M)x\in\mathrm{c}\widehat{\mathrm{ok}}(M) so that MXtnMX^{t}\in\mathbb{Z}^{n}. Writing X=Y/kX=Y/k for some kk\in\mathbb{N} and YtnY^{t}\in\mathbb{Z}^{n}, we see that kMXt=MYtcol(M)kMX^{t}=MY^{t}\in\operatorname{\mathrm{col}}(M). Hence, the image of MXtMX^{t} in cok(M)\mathrm{cok}(M) lies in tcok(M)\mathrm{tcok}(M). Since ϕAnn(tcok(M))\phi\in\operatorname{\mathrm{Ann}}(\mathrm{tcok}(M)),

ΦMXt=ϕ([MXt])=0mod,\Phi MX^{t}=\phi([MX^{t}])=0\mod\mathbb{Z},

where [MXt][MX^{t}] denotes the image of MXtMX^{t} in cok(M)\mathrm{cok}(M). Therefore, ϕker(,M)\phi\in\ker(\langle\,,\rangle_{M}). This forces ker(,M)=Ann(tcok(M))\ker(\langle\,,\rangle_{M})=\operatorname{\mathrm{Ann}}(\mathrm{tcok}(M)), as desired. ∎

By abuse of notation, we also use ,M\langle\,,\rangle_{M} to denote the duality pairing on tcok^(M)=coker,M\mathrm{tc}\widehat{\mathrm{ok}}(M)=\mathrm{coker}\langle\,,\rangle_{M}.

Corollary 3.4.

The pairing ,M\langle\,,\rangle_{M} is a duality pairing on tcok^(M)\mathrm{tc}\widehat{\mathrm{ok}}(M).

Since tcok^(M)\mathrm{tc}\widehat{\mathrm{ok}}(M) is a finite abelian group, the duality pairing ,M\langle\,,\rangle_{M} on tcok^(M)\mathrm{tc}\widehat{\mathrm{ok}}(M) induces a duality pairing on Hom(tcok^(M),/)=tcok(M)\operatorname{\mathrm{Hom}}(\mathrm{tc}\widehat{\mathrm{ok}}(M),\mathbb{Q}/\mathbb{Z})=\mathrm{tcok}(M). The following shows that the duality pairing on tcok(M)\mathrm{tcok}(M) induced by ,M\langle\,,\rangle_{M} is Bosch and Lorenzini’s pairing ,tcok(M)\langle\,,\rangle_{\mathrm{tcok}(M)} on tcok(M)\mathrm{tcok}(M).

Lemma 3.5.

The duality pairing on tcok(M)\mathrm{tcok}(M) induced by ,M\langle\,,\rangle_{M} is ,tcok(M)\langle\,,\rangle_{\mathrm{tcok}(M)}.

Proof.

Recall that ,M\langle\,,\rangle_{M} gives an isomorphism, ρ:tcok^(M)tcok(M)\rho:\mathrm{tc}\widehat{\mathrm{ok}}(M)\to\mathrm{tcok}(M) taking ϕ¯tcok^(M)\overline{\phi}\in\mathrm{tc}\widehat{\mathrm{ok}}(M) to ϕ¯,M\langle\overline{\phi},-\rangle_{M}. Let ϕ\phi be any lift of ϕ¯\overline{\phi} to cok^(M)\mathrm{c}\widehat{\mathrm{ok}}(M) (recall that there is a natural quotient cok^(M)tcok^(M)\mathrm{c}\widehat{\mathrm{ok}}(M)\to\mathrm{tc}\widehat{\mathrm{ok}}(M)). If Φ\Phi is any lift of ϕ\phi to Hom(n,)\operatorname{\mathrm{Hom}}(\mathbb{Z}^{n},\mathbb{Q}), then ρ(ϕ¯)=ϕ¯,M=[MΦt]\rho(\overline{\phi})=\langle\overline{\phi},-\rangle_{M}=[M\Phi^{t}], where the brackets denote the image of MΦtnM\Phi^{t}\in\mathbb{Z}^{n} in tcok(M)\mathrm{tcok}(M).

For τ,τtcok(M)\tau,\tau^{\prime}\in\mathrm{tcok}(M), let T,TT,T^{\prime} be lifts of τ,τ\tau,\tau^{\prime} to ker(M)\ker(M)^{\perp}. There exist k,kk,k^{\prime}\in\mathbb{N} and S,SnS,S^{\prime}\in\mathbb{Z}^{n} such that kT=MSkT=MS and kT=MSk^{\prime}T^{\prime}=MS^{\prime}. Thus, the preimages of τ\tau and τ\tau^{\prime} under ρ\rho are the images of St/kS^{t}/k and (S)t/k(S^{\prime})^{t}/k^{\prime} in tcok^(M)\mathrm{tc}\widehat{\mathrm{ok}}(M). The pairing on tcok(M)\mathrm{tcok}(M) induced by ,M\langle\,,\rangle_{M} takes

(τ,τ)ρ1(τ),ρ1(τ)M=StMSkkmod.(\tau,\tau^{\prime})\mapsto\langle\rho^{-1}(\tau),\rho^{-1}(\tau^{\prime})\rangle_{M}=\frac{S^{t}MS^{\prime}}{kk^{\prime}}\mod\mathbb{Z}.

However, recall that this is exactly τ,τtcok(M)\langle\tau,\tau^{\prime}\rangle_{\mathrm{tcok}(M)}. ∎

Later, given a finitely generated abelian group GG, we will be interested in taking the tensor product of GG with /b\mathbb{Z}/b\mathbb{Z} for some positive integer bb. If G^\widehat{G} is equipped with a symmetric pairing ,G^\langle\,,\rangle_{\widehat{G}}, we would like the dual of G/bG\otimes\mathbb{Z}/b\mathbb{Z} to come naturally equipped with a symmetric pairing given by ,G^\langle\,,\rangle_{\widehat{G}}. Let =/b\mathfrak{R}=\mathbb{Z}/b\mathbb{Z}. There is a natural quotient map GGG\to G\otimes\mathfrak{R} given by gg1g\mapsto g\otimes 1, and therefore there is a natural inclusion Hom(G,/)G^\operatorname{\mathrm{Hom}}(G\otimes\mathfrak{R},\mathbb{Q}/\mathbb{Z})\to\widehat{G}. Thus, we obtain a pairing on the dual of GG\otimes\mathfrak{R} by restricting the pairing on G^\widehat{G} to Hom(G,/)G^\operatorname{\mathrm{Hom}}(G\otimes\mathfrak{R},\mathbb{Q}/\mathbb{Z})\subset\widehat{G}. We abuse notation and also denote this pairing on Hom(G,/)\operatorname{\mathrm{Hom}}(G\otimes\mathfrak{R},\mathbb{Q}/\mathbb{Z}) by ,G^\langle\,,\rangle_{\widehat{G}}.

For a graph Γ\Gamma, recall from Section 2 the definition of its graph Laplacian LΓL_{\Gamma} and associated sandpile group SΓS_{\Gamma}. If Γ\Gamma is connected, in the language of the above, we see that SΓS_{\Gamma} is simply the torsion subgroup of the cokernel of the Laplacian, since ker(LΓ)=Z\ker(L_{\Gamma})^{\perp}=Z (this is a well-known fact [9]; recall that ZZ is the subspace of vectors in n\mathbb{Z}^{n} whose coordinates sum to 1). Hence, using Bosch and Lorenzini’s construction, SΓS_{\Gamma} comes equipped with a canonical duality pairing.

However, since we will be dealing with Erdős-Rényi random graphs later on, we would like for there to exist a pairing on SΓ^\widehat{S_{\Gamma}} for any graph Γ\Gamma. We claim that for any Γ\Gamma, not necessarily connected, SΓ^\widehat{S_{\Gamma}} is equipped with a natural symmetric pairing given by LΓL_{\Gamma}, which we will denote using ,SΓ^\langle\,,\rangle_{\widehat{S_{\Gamma}}}. To see why, recall that we have the following chain of inclusions:

tcok(LΓ)=ker(LΓ)/col(LΓ)SΓ=Z/col(LΓ)cok(LΓ)=n/col(LΓ).\mathrm{tcok}(L_{\Gamma})=\ker(L_{\Gamma})^{\perp}/\operatorname{\mathrm{col}}(L_{\Gamma})\subset S_{\Gamma}=Z/\operatorname{\mathrm{col}}(L_{\Gamma})\subset\mathrm{cok}(L_{\Gamma})=\mathbb{Z}^{n}/\operatorname{\mathrm{col}}(L_{\Gamma}).

As a result, the canonical duality pairing on tcok^(LΓ)\mathrm{tc}\widehat{\mathrm{ok}}(L_{\Gamma}) pushes forward to a symmetric pairing on SΓ^\widehat{S_{\Gamma}}, which itself pushes forward to the symmetric pairing ,LΓ\langle\,,\rangle_{L_{\Gamma}} on cok^(LΓ)\mathrm{c}\widehat{\mathrm{ok}}(L_{\Gamma}). Of these three pairings, we will be most interested in the symmetric pairing on SΓ^\widehat{S_{\Gamma}}.

The following lemma will be useful later.

Lemma 3.6.

Let Γ\Gamma be a graph on nn vertices and GG any finite abelian group whose dual is equipped with a symmetric pairing ,G^\langle\,,\rangle_{\widehat{G}}. Suppose F:nGF:\mathbb{Z}^{n}\to G is a surjection whose restriction to ZZ remains onto. Also assume that FLΓ=0FL_{\Gamma}=0 so that FF descends to a map F:SΓ=Z/col(LΓ)GF:S_{\Gamma}=Z/\operatorname{\mathrm{col}}(L_{\Gamma})\to G with transpose Ft:G^SΓ^F^{t}:\widehat{G}\to\widehat{S_{\Gamma}}. Then Ft(,SΓ^)=,G^F^{t}(\langle\,,\rangle_{\widehat{S_{\Gamma}}})=\langle\,,\rangle_{\widehat{G}} if and only if Ft(,LΓ)=,G^F^{t}(\langle\,,\rangle_{L_{\Gamma}})=\langle\,,\rangle_{\widehat{G}}.

Proof.

We start with the following commutative diagram and its dual:

Z{Z}n{\mathbb{Z}^{n}}SΓ{S_{\Gamma}}cok(LΓ){\mathrm{cok}(L_{\Gamma})}G{G}F\scriptstyle{F}F\scriptstyle{F}

       Z^{\widehat{Z}}(/)n{(\mathbb{Q}/\mathbb{Z})^{n}}SΓ^{\widehat{S_{\Gamma}}}cok^(LΓ){\mathrm{c}\widehat{\mathrm{ok}}(L_{\Gamma})}G^{\widehat{G}}Ft\scriptstyle{F^{t}}Ft\scriptstyle{F^{t}}

Recall that FtF^{t} pushes the pairing ,SΓ^\langle\,,\rangle_{\widehat{S_{\Gamma}}} on SΓ^\widehat{S_{\Gamma}} forward to some pairing on G^\widehat{G}. Likewise, the transpose of the inclusion ι:SΓcok(LΓ)\iota:S_{\Gamma}\hookrightarrow\mathrm{cok}(L_{\Gamma}) pushes the pairing ,SΓ^\langle\,,\rangle_{\widehat{S_{\Gamma}}} forward to the pairing ,LΓ\langle\,,\rangle_{L_{\Gamma}} on cok^(LΓ)\mathrm{c}\widehat{\mathrm{ok}}(L_{\Gamma}). The commutativity of the diagram implies the result. ∎

4. Obtaining the moments I: Finding the equations

Let GG be a finite abelian group whose dual G^=Hom(G,/)\widehat{G}=\operatorname{\mathrm{Hom}}(G,\mathbb{Q}/\mathbb{Z}) is equipped with a symmetric pairing ,G^\langle\,,\rangle_{\widehat{G}}. Suppose ΓG(n,q)\Gamma\in G(n,q), and let SS denote its sandpile group SΓS_{\Gamma} and LL the Laplacian LΓL_{\Gamma}. Because SS is defined as a quotient of ZZ, any surjection SGS\to G lifts to a surjection ZGZ\to G. Therefore,

𝔼(#Sur(S,G))=FSur(Z,G)(col(L)ker(F)andFt(,S^)=,G^).\mathbb{E}(\#\operatorname{\mathrm{Sur}}^{*}(S,G))=\sum_{F\in\operatorname{\mathrm{Sur}}(Z,G)}\operatorname{\mathbb{P}}(\operatorname{\mathrm{col}}(L)\subset\ker(F)\mathrm{\ and\ }F^{t}(\langle\,,\rangle_{\widehat{S}})=\langle\,,\rangle_{\widehat{G}}).

We will estimate the probabilities on the right-hand side of the above; to do so, we utilize the following more general setup.

Let bb denote the exponent of GG, and let a=b2a=b^{2}. Let RR be the ring /a\mathbb{Z}/a\mathbb{Z}. This notation will be used through Section 8. Note that Sur(S,G)=Sur(S/b,G)\operatorname{\mathrm{Sur}}(S,G)=\operatorname{\mathrm{Sur}}(S\otimes\mathbb{Z}/b\mathbb{Z},G), i.e., whether col(L)ker(F)\operatorname{\mathrm{col}}(L)\subset\ker(F) only depends on the entries of LL modulo bb. We will see later in this section that whether Ft(,S^)=,G^F^{t}(\langle\,,\rangle_{\widehat{S}})=\langle\,,\rangle_{\widehat{G}} also only depends on the entries of LL modulo bb. However, for reasons that will become apparent later, we will mostly work over RR (modulo aa) in the following. In this and the following four sections, all of our linear algebra will be done over RR. However, since RR is not a domain, we are forced to work more abstractly rather than just with matrices.

Let GG be the finite abelian group fixed above with symmetric pairing ,G^\langle\,,\rangle_{\widehat{G}} on its dual G^\widehat{G}. Let GpG_{p} denote the Sylow pp-subgroup of GG, and write G=pGpG=\bigoplus_{p}G_{p}. For a prime pp, suppose GpG_{p} is a pp-group of type λ\lambda, where λ\lambda is the partition given by λ1λ2λr\lambda_{1}\geq\lambda_{2}\geq\cdots\geq\lambda_{r}. Because a pairing on G^\widehat{G} may be viewed as an element of Hom(G^G^,/)\operatorname{\mathrm{Hom}}(\widehat{G}\otimes\widehat{G},\mathbb{Q}/\mathbb{Z}), any pairing on G^\widehat{G} is entirely determined by its restriction to (G^)p(G^)p(\widehat{G})_{p}\otimes(\widehat{G})_{p} for each pp.

Let g1,,grg_{1},\ldots,g_{r} generate GpG_{p} as an abelian group with pλigi=0p^{\lambda_{i}}g_{i}=0 and no other relations, and let gi^\widehat{g_{i}} be the elements of G^p\widehat{G}_{p} such that gi^(gj)=δij/pλi\widehat{g_{i}}(g_{j})=\delta_{ij}/p^{\lambda_{i}}, where recall δij\delta_{ij} is the indicator of i=ji=j. Also, suppose that gi^,gj^G^=aij/pλj\langle\widehat{g_{i}},\widehat{g_{j}}\rangle_{\widehat{G}}=a_{ij}/p^{\lambda_{j}} for 1ijr1\leq i\leq j\leq r, where aija_{ij} is an integer 0aij<pλj0\leq a_{ij}<p^{\lambda_{j}}. Note that the restriction of the pairing to (G^)p(G^)p(\widehat{G})_{p}\otimes(\widehat{G})_{p} is entirely determined by these values aija_{ij}. Let ,G^p\langle\,,\rangle_{\widehat{G}_{p}} denote the restriction of the pairing ,G^\langle\,,\rangle_{\widehat{G}} to (G^)p(G^)p(\widehat{G})_{p}\otimes(\widehat{G})_{p}.

Recall R=/aR=\mathbb{Z}/a\mathbb{Z}, and let V=RnV=R^{n}. Suppose MM is a symmetric matrix in Mn(R)M_{n}(R) with entries mijRm_{ij}\in R for 1ijn1\leq i\leq j\leq n. We distinguish a basis of VV, say v1,,vnv_{1},\ldots,v_{n}; let v1^,,vn^\widehat{v_{1}},\ldots,\widehat{v_{n}} denote the corresponding dual basis of V^\widehat{V}, where vi^(vj)=δij/a\widehat{v_{i}}(v_{j})=\delta_{ij}/a. With respect to the bases v1^,,vn^\widehat{v_{1}},\ldots,\widehat{v_{n}} and v1,,vnv_{1},\ldots,v_{n}, we may view MM as a linear map V^V\widehat{V}\to V where M(vi^)=jmijvjM(\widehat{v_{i}})=\sum_{j}m_{ij}v_{j}. Recall from Section 3 that cok^(M)=Hom(cok(M),/)\mathrm{c}\widehat{\mathrm{ok}}(M)=\operatorname{\mathrm{Hom}}(\mathrm{cok}(M),\mathbb{Q}/\mathbb{Z}) comes equipped with a symmetric pairing given by MM, which we’ll denote by ,M\langle\,,\rangle_{M}. Let F:VGpF:V\to G_{p} be a surjective homomorphism and denote the transpose of FF by Ft:G^pV^F^{t}:\widehat{G}_{p}\to\widehat{V}. Since cok(M)\mathrm{cok}(M) is a quotient of RnR^{n}, if col(M)ker(F)\operatorname{\mathrm{col}}(M)\subset\ker(F), then FF factors through the quotient of VV by col(M)\operatorname{\mathrm{col}}(M), i.e., FF descends to a map F:cok(M)GpF:\mathrm{cok}(M)\to G_{p}. Let F(vj)=i=1rfijgiF(v_{j})=\sum_{i=1}^{r}f_{ij}g_{i} for fij/pλif_{ij}\in\mathbb{Z}/p^{\lambda_{i}}\mathbb{Z}.

Lemma 4.1.

With notation as above, col(M)ker(F)\operatorname{\mathrm{col}}(M)\subset\ker(F) if and only if

(2) k=1nfikmk=0modpλi\sum_{k=1}^{n}f_{ik}m_{k\ell}=0\mod p^{\lambda_{i}}

for 1ir1\leq i\leq r and 1n1\leq\ell\leq n. If col(M)ker(F)\operatorname{\mathrm{col}}(M)\subset\ker(F), then FF descends to a map cok(M)Gp\mathrm{cok}(M)\to G_{p}, and we have Ft(,M)=,G^pF^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}_{p}} if and only if

(3) k,=1nfikmkfj=pλiaijmodpλi+λj\sum_{k,\ell=1}^{n}f_{ik}m_{k\ell}f_{j\ell}=p^{\lambda_{i}}a_{ij}\mod p^{\lambda_{i}+\lambda_{j}}

for all 1ijr1\leq i\leq j\leq r (we will use (2)\eqref{cok eqns} to show that the left-hand side of (3) is well-defined modulo pλi+λjp^{\lambda_{i}+\lambda_{j}}).

Proof.

We first note that col(M)ker(F)\operatorname{\mathrm{col}}(M)\subset\ker(F) if and only if the composition FMHom(V^,Gp)FM\in\operatorname{\mathrm{Hom}}(\widehat{V},G_{p}) is 0. Moreover, we see that FM=0FM=0 if and only if

k=1nfikmk=0modpλi\sum_{k=1}^{n}f_{ik}m_{k\ell}=0\mod p^{\lambda_{i}}

for all 1ir1\leq i\leq r and 1n1\leq\ell\leq n.

Now, assume FM=0FM=0 so that F:VGpF:V\to G_{p} descends to a map F:cok(M)GpF:\mathrm{cok}(M)\to G_{p}. Recall from Section 3 that FtF^{t} takes the pairing on cok^(M)\mathrm{c}\widehat{\mathrm{ok}}(M) to a pairing on G^p\widehat{G}_{p} and that a symmetric pairing on G^p\widehat{G}_{p} is entirely determined by where it sends the pairs (gi^,gj^)(\widehat{g_{i}},\widehat{g_{j}}) for 1ijr1\leq i\leq j\leq r. Thus, we first compute Ft(gi^)=gi^FV^F^{t}(\widehat{g_{i}})=\widehat{g_{i}}\circ F\in\widehat{V} for 1ir1\leq i\leq r. We see that

gi^(F(vj))=gi^(kfkjgk)=fijpλimod.\widehat{g_{i}}(F(v_{j}))=\widehat{g_{i}}\left(\sum_{k}f_{kj}g_{k}\right)=\frac{f_{ij}}{p^{\lambda_{i}}}\mod\mathbb{Z}.

Hence, with respect to the aforementioned bases, we may view Ft(gi^)F^{t}(\widehat{g_{i}}) as the row vector

(fi1/pλifin/pλi)mod.\begin{pmatrix}f_{i1}/p^{\lambda_{i}}&\cdots&f_{in}/p^{\lambda_{i}}\end{pmatrix}\mod\mathbb{Z}.

It follows that

Ft(gi^),Ft(gj^)M=Ft(gi^)MFt(gj^)t=k,=1nfikmkfjpλi+λjmod\langle F^{t}(\widehat{g_{i}}),F^{t}(\widehat{g_{j}})\rangle_{M}=F^{t}(\widehat{g_{i}})MF^{t}(\widehat{g_{j}})^{t}=\sum_{k,\ell=1}^{n}\frac{f_{ik}m_{k\ell}f_{j\ell}}{p^{\lambda_{i}+\lambda_{j}}}\mod\mathbb{Z}

(recall we can do this computation by lifting Ft(gi^)F^{t}(\widehat{g_{i}}) to \mathbb{Q}). Therefore, given that FM=0FM=0, we see that Ft(,M)=,G^pF^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}_{p}} if and only if

k,=1nfikmkfjpλi+λj=aijpλjmod\sum_{k,\ell=1}^{n}\frac{f_{ik}m_{k\ell}f_{j\ell}}{p^{\lambda_{i}+\lambda_{j}}}=\frac{a_{ij}}{p^{\lambda_{j}}}\mod\mathbb{Z}

for all 1ijr1\leq i\leq j\leq r. ∎

Note that Sur(n,G)=Sur(nR,G)=Sur(V,G)\operatorname{\mathrm{Sur}}(\mathbb{Z}^{n},G)=\operatorname{\mathrm{Sur}}(\mathbb{Z}^{n}\otimes R,G)=\operatorname{\mathrm{Sur}}(V,G). Hence, given FSur(V,G)F\in\operatorname{\mathrm{Sur}}(V,G), Lemma 4.1 and the fact that p2λ1|ap^{2\lambda_{1}}|a imply whether col(M)ker(F)\operatorname{\mathrm{col}}(M)\subset\ker(F) and Ft(,M)=,G^F^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}} depends only on the entries of MM modulo aa. As before, let VV denote the RR-module RnR^{n} with distinguished basis v1,,vnv_{1},\ldots,v_{n}. Consider V=Hom(V,R)V^{*}=\operatorname{\mathrm{Hom}}(V,R) with dual basis v1,,vnv_{1}^{*},\ldots,v_{n}^{*}, where vi(vj)=δijv_{i}^{*}(v_{j})=\delta_{ij}. Note that there is an isomorphism between VV^{*} and V^\widehat{V} given by sending viv_{i}^{*} to vi^\widehat{v_{i}}.

Recall GpG_{p} is a pp-group of type λ\lambda. Define Hp=(/p2λ1)rH_{p}=(\mathbb{Z}/p^{2\lambda_{1}}\mathbb{Z})^{r}, and, noting that λ1\lambda_{1} and rr depend implicitly on pp, set H=pHpH=\bigoplus_{p}H_{p}. For any p|#Gp|\#G, let h1,,hrh_{1},\ldots,h_{r} denote the standard basis of Hp=RprH_{p}=R_{p}^{r}, and define a surjective map HGH\to G by summing the maps HpGpH_{p}\to G_{p} taking higih_{i}\mapsto g_{i} for each pp. We will use this map throughout the argument—unless stated otherwise, when we refer to a surjection HGH\to G, we refer to this map.

Let FHom(V,G)F\in\operatorname{\mathrm{Hom}}(V,G), and let Hom(V,H)\mathcal{F}\in\operatorname{\mathrm{Hom}}(V,H) be a lift of FF to HH (with respect to the aforementioned surjection HGH\to G).

Remark 4.2.

For the reader disturbed by this arbitrary choice of lift, we remark here that while there are many choices for \mathcal{F}, our arguments in the following section do not depend on this choice—the proofs only depend on the values of \mathcal{F} in HH after projecting to the quotient GG, i.e., the proofs only depend on FF itself. The lift \mathcal{F} solves the issue of whether the equations (3) from Lemma 4.1 are well-defined. By considering a lift of \mathcal{F}, we can instead work with equations that are always well-defined.

For example, consider the 1×11\times 1 matrix M=[ep]M1(/p2)M=[ep]\in M_{1}(\mathbb{Z}/p^{2}\mathbb{Z}) for some e0e\neq 0 mod pp. The cokernel of MM can be written as

cok^(M)={ϕHom(/p2,/)|ϕM=0}={ϕHom(/p2,/)|pϕ=0}\mathrm{c}\widehat{\mathrm{ok}}(M)=\{\phi\in\operatorname{\mathrm{Hom}}(\mathbb{Z}/p^{2}\mathbb{Z},\mathbb{Q}/\mathbb{Z})\;|\;\phi M=0\}=\{\phi\in\operatorname{\mathrm{Hom}}(\mathbb{Z}/p^{2}\mathbb{Z},\mathbb{Q}/\mathbb{Z})\;|\;p\phi=0\}

and is isomorphic to /p\mathbb{Z}/p\mathbb{Z}. The pairing on cok^(M)\mathrm{c}\widehat{\mathrm{ok}}(M) is given by

x/p,y/pM=xepyp2=xeyp/,\langle x/p,y/p\rangle_{M}=\frac{xepy}{p^{2}}=\frac{xey}{p}\in\mathbb{Q}/\mathbb{Z},

and we see that ,M\langle\,,\rangle_{M} depends on the entries of MM modulo p2p^{2} rather than the entries of MM modulo pp.

Consider the Sylow pp-part of im()\operatorname{\mathrm{im}}(\mathcal{F}), i.e., :VHp\mathcal{F}:V\to H_{p}. Let F(vj)=i=1rfijgiGpF(v_{j})=\sum_{i=1}^{r}f_{ij}g_{i}\in G_{p}, where fij/pλif_{ij}\in\mathbb{Z}/p^{\lambda_{i}}\mathbb{Z}, and let (vj)=i=1rFijhiHp\mathcal{F}(v_{j})=\sum_{i=1}^{r}F_{ij}h_{i}\in H_{p}, where FijR=/p2λ1F_{ij}\in R=\mathbb{Z}/p^{2\lambda_{1}}\mathbb{Z}. Since \mathcal{F} is a lift of FF, note that FijfijF_{ij}\equiv f_{ij} modulo pλip^{\lambda_{i}}.

Define Λp:HpHp\Lambda_{p}:H_{p}\to H_{p} to be the map sending hip2λ1λihih_{i}\mapsto p^{2\lambda_{1}-\lambda_{i}}h_{i} for 1ir1\leq i\leq r. Let Λ:HH\Lambda:H\to H be the direct sum of the Λp\Lambda_{p}’s. Note that the image of Λ\Lambda is isomorphic to GG, giving us a copy of GG living in HH—i.e., the map imΛG\operatorname{\mathrm{im}}\Lambda\to G given by p2λ1λihigip^{2\lambda_{1}-\lambda_{i}}h_{i}\mapsto g_{i} is an isomorphism. Moreover, the surjection HGH\to G defined earlier is equal to the composition of Λ\Lambda with this isomorphism imΛG\operatorname{\mathrm{im}}\Lambda\simeq G. Define Ωp:HpHp\Omega_{p}:H_{p}\to H_{p} to be the map taking hipλ1λihih_{i}\mapsto p^{\lambda_{1}-\lambda_{i}}h_{i} for 1ir1\leq i\leq r. Let Ω:HH\Omega:H\to H be the sum of the Ωp\Omega_{p}’s.

We define an element AA of Sym2HHom(H,H)\operatorname{\mathrm{Sym}}_{2}H\subset\operatorname{\mathrm{Hom}}(H^{*},H) that is equivalent to the data of the pairing ,G^\langle\,,\rangle_{\widehat{G}}. For each p|#Gp|\#G, recall that gi^,gj^G^=aij/pλj\langle\widehat{g_{i}},\widehat{g_{j}}\rangle_{\widehat{G}}=a_{ij}/p^{\lambda_{j}} for all 1ijr1\leq i\leq j\leq r. Let ApA_{p} denote the following symmetric element of Hom(Hp,Hp)\operatorname{\mathrm{Hom}}(H_{p}^{*},H_{p}), which we view as HpHpH_{p}\otimes H_{p}. Let

Ap=1i<jrp2λ1λjaij(hihj+hjhi)+i=1rp2λ1λiaiihihiHpHp,A_{p}=\sum_{1\leq i<j\leq r}p^{2\lambda_{1}-\lambda_{j}}a_{ij}(h_{i}\otimes h_{j}+h_{j}\otimes h_{i})+\sum_{i=1}^{r}p^{2\lambda_{1}-\lambda_{i}}a_{ii}h_{i}\otimes h_{i}\in H_{p}\otimes H_{p},

and let A=pApA=\bigoplus_{p}A_{p}. We may think of AA as a way of expressing the pairing ,G^\langle\,,\rangle_{\widehat{G}} algebraically.

Now, note that for each vVv_{\ell}^{*}\in V^{*}, where 1n1\leq\ell\leq n, we have

ΛM(v)=p|#Gi=1r(p2λ1λik=1nFikmk)hi.\Lambda\mathcal{F}M(v_{\ell}^{*})=\sum_{p|\#G}\sum_{i=1}^{r}\left(p^{2\lambda_{1}-\lambda_{i}}\sum_{k=1}^{n}F_{ik}m_{k\ell}\right)h_{i}.

Note here that in the above we are abusing notation slightly—both the hih_{i}’s and rr implicitly depend on pp. We note that

(4) p2λ1λik=1nFikmk=0modp2λ1p^{2\lambda_{1}-\lambda_{i}}\sum_{k=1}^{n}F_{ik}m_{k\ell}=0\mod p^{2\lambda_{1}}

for 1ir1\leq i\leq r and 1n1\leq\ell\leq n if and only if (2) holds. If we view GG as the subgroup ΛH\Lambda H of HH, then note that Λ\Lambda\mathcal{F} is simply the map F:VGF:V\to G defined earlier. Likewise, we see that

ΩMtΩt(hi)=p|#G1jrk,=1np2λ1λiλjFikFjmkhj.\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}(h_{i}^{*})=\sum_{p|\#G}\sum_{1\leq j\leq r}\sum_{k,\ell=1}^{n}p^{2\lambda_{1}-\lambda_{i}-\lambda_{j}}F_{ik}F_{j\ell}m_{k\ell}h_{j}.

Note that ΩMtΩt\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t} is a map HHH^{*}\to H and is symmetric when viewed as an element of HHH\otimes H, since (ΩMtΩt)t=ΩMtΩt(\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t})^{t}=\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}. Moreover,

(5) k,=1np2λ1λiλjFikFjmk=p2λ1λjaijmodp2λ1\sum_{k,\ell=1}^{n}p^{2\lambda_{1}-\lambda_{i}-\lambda_{j}}F_{ik}F_{j\ell}m_{k\ell}=p^{2\lambda_{1}-\lambda_{j}}a_{ij}\mod p^{2\lambda_{1}}

for 1ijr1\leq i\leq j\leq r if and only if (3) holds. Hence, ΛM=0\Lambda\mathcal{F}M=0 and ΩMtΩt=A\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}=A if and only if both (2) and (3) hold for all p|#Gp|\#G. It follows that

(6) (col(M)ker(F)andFt(,M)=,G^)=(ΛM=0andΩMtΩt=A).\operatorname{\mathbb{P}}(\operatorname{\mathrm{col}}(M)\subset\ker(F)\ \mathrm{and\ }F^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}})=\operatorname{\mathbb{P}}(\Lambda\mathcal{F}M=0\mathrm{\ and\ }\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}=A).

While working with the lift \mathcal{F} seems artificial and notationally clumsy, we do so because it removes a dependence between our equations—we saw previously that the equations (3) for Ft(,M)=,G^F^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}} are not a priori well-defined. By considering a lift of FF, the equations (5) can be studied without first having to condition on (4). The maps Λ\Lambda and Ω\Omega are introduced so that the lifted equations (4) and (5) hold if and only if our original equations (2) and (3) hold. Thus, we may consider all of the equations simultaneously; doing so will be helpful when we apply the discrete Fourier transform in the next section.

5. Obtaining the moments II: The structural properties of the equations

In the previous section, we showed that for any random symmetric matrix MMn(R)M\in M_{n}(R), we have

𝔼(#Sur(cok(M),G))=FSur(V,G)(ΛM=0andΩMtΩt=A),\mathbb{E}(\#\operatorname{\mathrm{Sur}}^{*}(\mathrm{cok}(M),G))=\sum_{F\in\operatorname{\mathrm{Sur}}(V,G)}\operatorname{\mathbb{P}}(\Lambda\mathcal{F}M=0\mathrm{\ and\ }\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}=A),

where \mathcal{F} is any lift of FF to HH. The goal of this section and the two that follow is to estimate the probabilities in the right-hand side of the equation above. Let ζ\zeta be a primitive aath root of unity, and let \mathcal{E} denote the event

ΛM=0andΩMtΩt=A.\Lambda\mathcal{F}M=0\mathrm{\ and\ }\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}=A.

Recall that we may view MM as an element of Hom(V,V)\operatorname{\mathrm{Hom}}(V^{*},V) and note that ΛMHom(V,G)\Lambda\mathcal{F}M\in\operatorname{\mathrm{Hom}}(V^{*},G) (where we view GΛHG\simeq\Lambda H as a subgroup of HH) and that ΩMtΩtHom(H,H)\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}\in\operatorname{\mathrm{Hom}}(H^{*},H), which, recall, is naturally isomorphic to HHH\otimes H. Since (ΩMtΩt)t=ΩMtΩt(\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t})^{t}=\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}, it follows that ΩMtΩtSym2H\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}\in\operatorname{\mathrm{Sym}}_{2}H. Thus, the Fourier expansion implies

𝟙=1|G|n|Hom(Sym2H,R)|CHom(Hom(V,G),R)DHom(Sym2H,R)ζC(ΛM)+D(ΩMtΩtA).\mathbbm{1}_{\mathcal{E}}=\frac{1}{|G|^{n}\cdot|\operatorname{\mathrm{Hom}}(\operatorname{\mathrm{Sym}}_{2}H,R)|}\sum_{\begin{subarray}{c}C\in\operatorname{\mathrm{Hom}}(\operatorname{\mathrm{Hom}}(V^{*},G),R)\\ D\in\operatorname{\mathrm{Hom}}(\operatorname{\mathrm{Sym}}_{2}H,R)\end{subarray}}\zeta^{C(\Lambda\mathcal{F}M)+D(\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}-A)}.

Hence,

()=𝔼(𝟙)=1|G|n|Hom(Sym2H,R)|CHom(Hom(V,G),R)DHom(Sym2H,R)𝔼(ζC(ΛM)+D(ΩMtΩtA)).\operatorname{\mathbb{P}}(\mathcal{E})=\mathbb{E}(\mathbbm{1}_{\mathcal{E}})=\frac{1}{|G|^{n}\cdot|\operatorname{\mathrm{Hom}}(\operatorname{\mathrm{Sym}}_{2}H,R)|}\sum_{\begin{subarray}{c}C\in\operatorname{\mathrm{Hom}}(\operatorname{\mathrm{Hom}}(V^{*},G),R)\\ D\in\operatorname{\mathrm{Hom}}(\operatorname{\mathrm{Sym}}_{2}H,R)\end{subarray}}\mathbb{E}(\zeta^{C(\Lambda\mathcal{F}M)+D(\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}-A)}).

The pairs (C,D)(C,D) give the equations that a matrix must satisfy in order for FF to factor through the cokernel of the matrix and for FF to push the pairing ,M\langle\,,\rangle_{M} on cok^(M)\mathrm{c}\widehat{\mathrm{ok}}(M) forward to the fixed pairing on G^\widehat{G}. Viewing C(ΛM)+D(ΩMtΩtA)C(\Lambda\mathcal{F}M)+D(\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}-A) as a function of MM, we see that it is in fact a linear function of the entries mijm_{ij} of MM (for iji\leq j). In the following, we will compute the coefficient of each mijm_{ij}. When the entries mijm_{ij} are independent, the expected value function is multiplicative, and we can factor the expected value in the right-hand side of the above (see (7) below).

For the rest of this section, unless specified otherwise, when we refer to GG we are referencing the copy of GG living in HH as ΛH\Lambda H. Because we are viewing GG as a subgroup of HH, we may view elements of HH^{*} as elements of GG^{*}, since a linear map HRH\to R restricts to a linear map G=ΛHRG=\Lambda H\to R. Further, any map GRG\to R yields a map HRH\to R via precomposition with Λ\Lambda. However, note that an element of GG^{*} cannot a priori be applied to an arbitrary element of HH.

Because VRnV^{*}\simeq R^{n} (noncanonically), there is a natural isomorphism Hom(V,R)GHom(V,G)\operatorname{\mathrm{Hom}}(V^{*},R)\otimes G\to\operatorname{\mathrm{Hom}}(V^{*},G). Therefore, there is a natural isomorphism from Hom(Hom(V,G),R)Hom(VG,R)\operatorname{\mathrm{Hom}}(\operatorname{\mathrm{Hom}}(V^{*},G),R)\to\operatorname{\mathrm{Hom}}(V\otimes G,R), which is itself canonically isomorphic to Hom(V,Hom(G,R))Hom(V,G)\operatorname{\mathrm{Hom}}(V,\operatorname{\mathrm{Hom}}(G,R))\simeq\operatorname{\mathrm{Hom}}(V,G^{*}). Hence, we regard CC as an element of Hom(V,G)\operatorname{\mathrm{Hom}}(V,G^{*}). Let e:G×GRe:G^{*}\times G\to R denote the evaluation map. Similarly, there is a natural isomorphism between Hom(Sym2H,R)\operatorname{\mathrm{Hom}}(\operatorname{\mathrm{Sym}}_{2}H,R) and Sym2H\operatorname{\mathrm{Sym}}^{2}H^{*}. Let [ab][a^{*}\otimes b^{*}] denote the class of abHHa^{*}\otimes b^{*}\in H^{*}\otimes H^{*} in Sym2H\operatorname{\mathrm{Sym}}^{2}H^{*}, and recall that Sym2H\operatorname{\mathrm{Sym}}_{2}H is generated by the elements of the form ab+baa\otimes b+b\otimes a for aba\neq b along with the elements aaa\otimes a. The aforementioned isomorphism between Hom(Sym2H,R)\operatorname{\mathrm{Hom}}(\operatorname{\mathrm{Sym}}_{2}H,R) and Sym2H\operatorname{\mathrm{Sym}}^{2}H^{*} is witnessed by letting [ab][a^{*}\otimes b^{*}] correspond to the map taking

xy+yxa(x)b(y)+a(y)b(x)Randxxa(x)b(x)R.x\otimes y+y\otimes x\mapsto a^{*}(x)b^{*}(y)+a^{*}(y)b^{*}(x)\in R\quad\textrm{and}\quad x\otimes x\mapsto a^{*}(x)b^{*}(x)\in R.

Use this isomorphism to regard DD as an element of Sym2H\operatorname{\mathrm{Sym}}^{2}H^{*}, and note that the evaluation does not depend on the chosen representative of DD in HHH^{*}\otimes H^{*}. Let 𝔢:Sym2H×Sym2HR\mathfrak{e}:\operatorname{\mathrm{Sym}}^{2}H^{*}\times\operatorname{\mathrm{Sym}}_{2}H\to R denote the evaluation map described above. For each element DD of Sym2H\operatorname{\mathrm{Sym}}^{2}H^{*} and choice of basis for HH, there exists some unique upper-triangular representative of DD in HHH^{*}\otimes H^{*} with respect to this basis. We will use this fact to simplify computations.

The random matrices we will consider have entries that are independent with respect to a specific basis of VV. Therefore, occasionally we are forced to do some computations using this distinguished basis, which we denote by v1,,vnv_{1},\ldots,v_{n}. At times, we will work with an alternate basis more conveniently aligned with GG or HH (through \mathcal{F}). Because we are working over R=/aR=\mathbb{Z}/a\mathbb{Z}, which is not necessarily an integral domain, we prioritize making our arguments basis-independent as often as possible.

We begin by computing the coefficient on each mijm_{ij} and rewriting it in more convenient notation. For h1,h2Hh_{1},h_{2}\in H, let h1h2{h_{1}\odot h_{2}} be shorthand for the element h1h2+h2h1Sym2Hh_{1}\otimes h_{2}+h_{2}\otimes h_{1}\in\operatorname{\mathrm{Sym}}_{2}H. We see that

C(ΛM)\displaystyle C(\Lambda\mathcal{F}M) +D(ΩMtΩt)\displaystyle+D(\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t})
=i=1nj=i+1n(e(C(vj),Λ(vi))+e(C(vi),Λ(vj))+𝔢(D,Ω(vi)Ω(vj)))mij\displaystyle=\sum_{i=1}^{n}\sum_{j=i+1}^{n}\Big{(}e(C(v_{j}),\Lambda\mathcal{F}(v_{i}))+e(C(v_{i}),\Lambda\mathcal{F}(v_{j}))+\mathfrak{e}(D,{\Omega\mathcal{F}(v_{i})\odot\Omega\mathcal{F}(v_{j})})\Big{)}m_{ij}
+i=1n(e(C(vi),Λ(vi))+𝔢(D,Ω(vi)Ω(vi)))mii.\displaystyle\qquad\qquad\qquad+\sum_{i=1}^{n}\Big{(}e(C(v_{i}),\Lambda\mathcal{F}(v_{i}))+\mathfrak{e}(D,\Omega\mathcal{F}(v_{i})\otimes\Omega\mathcal{F}(v_{i}))\Big{)}m_{ii}.

For i<ji<j, define

Eij(C,D,)=e(C(vj),Λ(vi))+e(C(vi),Λ(vj))+𝔢(D,Ω(vi)Ω(vj)),E_{ij}(C,D,\mathcal{F})=e(C(v_{j}),\Lambda\mathcal{F}(v_{i}))+e(C(v_{i}),\Lambda\mathcal{F}(v_{j}))+\mathfrak{e}(D,{\Omega\mathcal{F}(v_{i})\odot\Omega\mathcal{F}(v_{j})}),

and we also define

Eii(C,D,)=e(C(vi),Λ(vi))+𝔢(D,Ω(vi)Ω(vi)).E_{ii}(C,D,\mathcal{F})=e(C(v_{i}),\Lambda\mathcal{F}(v_{i}))+\mathfrak{e}(D,\Omega\mathcal{F}(v_{i})\otimes\Omega\mathcal{F}(v_{i})).

Hence, when the entries mijm_{ij} (for iji\leq j) are independent,

(7) (ΛM=0andΩMtΩt=A)=1|G|n|Sym2H|(C,D)𝔼(ζD(A))ij𝔼(ζEij(C,D,)mij).\operatorname{\mathbb{P}}(\Lambda\mathcal{F}M=0\mathrm{\ and\ }\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}=A)=\frac{1}{|G|^{n}\cdot|\operatorname{\mathrm{Sym}}^{2}H|}\sum_{(C,D)}\mathbb{E}(\zeta^{-D(A)})\prod_{i\leq j}\mathbb{E}(\zeta^{E_{ij}(C,D,\mathcal{F})m_{ij}}).

For x0x\neq 0, we know from [54] that |𝔼(ζumij)||\mathbb{E}(\zeta^{um_{ij}})| will be quite small (Lemma 4.2 of [54]; restated as Lemma 6.2). To get the bounds required to prove Theorem 1.2, we would like most of the coefficients Eij(C,D,)E_{ij}(C,D,\mathcal{F}) to be nonzero. Thus, this section is devoted to pinpointing the characteristics of \mathcal{F} and (C,D)(C,D) that control the number of nonzero Eij(C,D,)E_{ij}(C,D,\mathcal{F}). Note there are (n+12)\binom{n+1}{2} total coefficients, and, to get the bounds we want, we would like quadratically many Eij(C,D,)E_{ij}(C,D,\mathcal{F})’s to be nonzero. For certain “good” FF, we will show that there are three possible outcomes: on the order of n2n^{2} coefficients are nonzero; on the order of nn coefficients are nonzero; all of the coefficients are zero. To show this, we identify the desired structural property of (C,D)(C,D) that affects the number of nonzero coefficients. We also show that there are relatively few (C,D)(C,D) such that only linearly many of the coefficients are nonzero; we count explicitly the (C,D)(C,D) for which all of the coefficients are zero. In the following, we will assume FF is “good” and define the aforementioned structural properties of (C,D)(C,D). We will also explain the structural property which makes FF “good.” Later, in Section 7, we will consider the remaining FF.

Now, we will write the coefficients Eij(C,D,)E_{ij}(C,D,\mathcal{F}) more equivariantly via a pairing. For elements ΛHom(V,G)\Lambda\mathcal{F}\in\operatorname{\mathrm{Hom}}(V,G) and CHom(V,G)C\in\operatorname{\mathrm{Hom}}(V,G^{*}), define a map ϕΛ,CHom(V,GG)\phi_{\Lambda\mathcal{F},C}\in\operatorname{\mathrm{Hom}}(V,G\oplus G^{*}) given by taking the direct sum of Λ\Lambda\mathcal{F} and CC. Likewise, define a map ϕC,ΛHom(V,GG)\phi_{C,\Lambda\mathcal{F}}\in\operatorname{\mathrm{Hom}}(V,G^{*}\oplus G) by switching the order of summation. There is a map

t:(GG)×(GG)Rt:(G\oplus G^{*})\times(G^{*}\oplus G)\to R

given by

((h,ϕ),(ϕ,h))ϕ(h)+ϕ(h).((h,\phi),(\phi^{\prime},h^{\prime}))\mapsto\phi^{\prime}(h)+\phi(h^{\prime}).

Note that for all u,vVu,v\in V,

t(ϕC,Λ(u),ϕΛ,C(v))=e(C(u),Λ(v))+e(C(v),Λ(u)).t(\phi_{C,\Lambda\mathcal{F}}(u),\phi_{\Lambda\mathcal{F},C}(v))=e(C(u),\Lambda\mathcal{F}(v))+e(C(v),\Lambda\mathcal{F}(u)).

Similarly, for Hom(V,H)\mathcal{F}\in\operatorname{\mathrm{Hom}}(V,H) and CHom(V,G)C\in\operatorname{\mathrm{Hom}}(V,G^{*}), we have a map ϕ,CHom(V,HG)\phi_{\mathcal{F},C}\in\operatorname{\mathrm{Hom}}(V,H\oplus G^{*}) given by taking the direct sum of \mathcal{F} and CC. Immediately, we see that for any submodule UU of VV,

ker(ϕ,C|U)ker(|U).\ker(\phi_{\mathcal{F},C}|_{U})\subset\ker(\mathcal{F}|_{U}).

Given any subset σ[n]\sigma\subset[n], let VσV_{\sigma} denote the submodule of VV generated by the viv_{i} such that iσi\not\in\sigma. In other words, VσV_{\sigma} is the submodule given by not using the coordinates in σ\sigma. The following definition gives the key structural property of the pair (C,D)(C,D) (for some fixed \mathcal{F}) that determines if enough of the coefficients Eij(C,D,)E_{ij}(C,D,\mathcal{F}) are nonzero.

Definition 5.1.

Let 0<γ<10<\gamma<1 be a real number (which we will specify later). For a fixed Hom(V,H)\mathcal{F}\in\operatorname{\mathrm{Hom}}(V,H), we say that the pair (C,D)(C,D) is robust for \mathcal{F} if every σ[n]\sigma\subset[n] such that |σ|<γn|\sigma|<\gamma n has the property that

ker(ϕ,C|Vσ)ker(|Vσ).\ker(\phi_{\mathcal{F},C}|_{V_{\sigma}})\subsetneq\ker(\mathcal{F}|_{V_{\sigma}}).

Otherwise, the pair (C,D)(C,D) is said to be weak for \mathcal{F}.

We give an upper bound for the number of weak pairs (C,D)(C,D).

Lemma 5.2 (Estimate for number of weak (C,D)(C,D)).

Given GG, there is a constant CGC_{G} such that for all nn the following holds. Let HH be defined as in Section 4, and let Hom(V,H)\mathcal{F}\in\operatorname{\mathrm{Hom}}(V,H). The number of pairs (C,D)Hom(V,G)×Sym2H(C,D)\in\operatorname{\mathrm{Hom}}(V,G^{*})\times\operatorname{\mathrm{Sym}}^{2}H^{*} that are weak for \mathcal{F} is at most

CG(nγn1)|G|γn.C_{G}{{n}\choose{\lceil\gamma n\rceil-1}}|G|^{\gamma n}.
Proof.

Suppose (C,D)(C,D) is weak for \mathcal{F}. Then there exists some σ[n]\sigma\subset[n] with |σ|<γn|\sigma|<\gamma n such that

ker(ϕ,C|Vσ)=ker(|Vσ).\ker(\phi_{\mathcal{F},C}|_{V_{\sigma}})=\ker(\mathcal{F}|_{V_{\sigma}}).

Because σμ\sigma\subset\mu implies VμVσV_{\mu}\subset V_{\sigma}, we see that the above equality still holds for σ\sigma enlarged. Thus, without loss of generality, we may assume that |σ|=γn1|\sigma|=\lceil\gamma n\rceil-1. Now, since ker(ϕ,C|Vσ)=ker(|Vσ)\ker(\phi_{\mathcal{F},C}|_{V_{\sigma}})=\ker(\mathcal{F}|_{V_{\sigma}}), it follows that CsCs is determined by s\mathcal{F}s for sVσs\in V_{\sigma}: if s=s\mathcal{F}s=\mathcal{F}s^{\prime} for s,sVσs,s^{\prime}\in V_{\sigma} and CsCsCs\neq C{s^{\prime}}, then ssker()s-s^{\prime}\in\ker(\mathcal{F}) while ssker(C)s-s^{\prime}\not\in\ker(C). In other words, for each value in im(|Vσ)\operatorname{\mathrm{im}}(\mathcal{F}|_{V_{\sigma}}) there is exactly one corresponding value in im(C|Vσ)\operatorname{\mathrm{im}}(C|_{V_{\sigma}}). Moreover, we have a well-defined homomorphism ψ:im(|Vσ)G\psi:\operatorname{\mathrm{im}}(\mathcal{F}|_{V_{\sigma}})\to G^{*} taking ψ(s)=Cs\psi(\mathcal{F}s)=Cs for all sVσs\in V_{\sigma}.

Recall that CC is determined by the values CviCv_{i} for iσi\in\sigma in addition to im(C|Vσ)\operatorname{\mathrm{im}}(C|_{V_{\sigma}}). Therefore, given \mathcal{F}, to specify a CHom(V,G)C\in\operatorname{\mathrm{Hom}}(V,G^{*}) such that (C,D)(C,D) is weak for \mathcal{F} for any DSym2HD\in\operatorname{\mathrm{Sym}}^{2}H^{*}, it suffices to choose a subset σ[n]\sigma\subset[n] with |σ|=γn1|\sigma|=\lceil\gamma n\rceil-1, values CviCv_{i} for iσi\in\sigma, and a homomorphism ψ:im(|Vσ)G\psi:\operatorname{\mathrm{im}}(\mathcal{F}|_{V_{\sigma}})\to G^{*}. Thus, there are (nγn1){{n}\choose{\lceil\gamma n\rceil-1}} choices for σ\sigma and |G||G| choices for each CviCv_{i} and |Hom(im(|Vσ),G)||\operatorname{\mathrm{Hom}}(\operatorname{\mathrm{im}}(\mathcal{F}|_{V_{\sigma}}),G^{*})| choices for ψ\psi, and then CC is determined. There are |Sym2H||\operatorname{\mathrm{Sym}}^{2}H| possibilities for DD. Note that because HH is determined by GG and because im(|Vσ)H\operatorname{\mathrm{im}}(\mathcal{F}|_{V_{\sigma}})\subset H, we can find some constant CGC_{G} such that |Sym2H||Hom(im(|Vσ),G)|CG|\operatorname{\mathrm{Sym}}^{2}H|\cdot|\operatorname{\mathrm{Hom}}(\operatorname{\mathrm{im}}(\mathcal{F}|_{V_{\sigma}}),G^{*})|\leq C_{G}. ∎

The following gives a sufficient condition for (C,D)(C,D) to be weak in terms of 𝔢\mathfrak{e} and tt.

Lemma 5.3.

Let Hom(V,H)\mathcal{F}\in\operatorname{\mathrm{Hom}}(V,H), let CHom(V,G)C\in\operatorname{\mathrm{Hom}}(V,G^{*}), and let DSym2HD\in\operatorname{\mathrm{Sym}}^{2}H^{*}. Suppose UU is a submodule of VV such that U=H\mathcal{F}U=H. Then if UU^{\prime} is a submodule of VV such that for all uUu\in U and uUu^{\prime}\in U^{\prime}

t(ϕC,Λ(u),ϕΛ,C(u))+𝔢(D,Ω(u)Ω(u))=0,t(\phi_{C,\Lambda\mathcal{F}}(u),\phi_{\Lambda\mathcal{F},C}(u^{\prime}))+\mathfrak{e}(D,\Omega\mathcal{F}(u)\odot\Omega\mathcal{F}(u^{\prime}))=0,

then ker(ϕ,C|U)=ker(|U)\ker(\phi_{\mathcal{F},C}|_{U^{\prime}})=\ker(\mathcal{F}|_{U^{\prime}}).

Proof.

Suppose for a contradiction that there exists some wUw\in U^{\prime} with (w)=0H\mathcal{F}(w)=0\in H but C(w)=ψ0GC(w)=\psi\neq 0\in G^{*}. Since ψ0\psi\neq 0, there exists some element gGHg\in G\subset H such that ψ(g)0\psi(g)\neq 0. Because U=H\mathcal{F}U=H, we have ΛU=ΛH=G\Lambda\mathcal{F}U=\Lambda H=G, and it follows that there must be some xUx\in U such that Λ(x)=g\Lambda\mathcal{F}(x)=g. Let C(x)=ψC(x)=\psi^{\prime}. Then

t(ϕC,Λ(x),ϕΛ,C(w))+𝔢(D,Ω(x)Ω(w))\displaystyle t(\phi_{C,\Lambda\mathcal{F}}(x),\phi_{\Lambda\mathcal{F},C}(w))+\mathfrak{e}(D,\Omega\mathcal{F}(x)\odot\Omega\mathcal{F}(w)) =ψ(g)+ψ(0)+𝔢(D,Ω(x)Ω(w))\displaystyle=\psi(g)+\psi^{\prime}(0)+\mathfrak{e}(D,{\Omega\mathcal{F}(x)\odot\Omega\mathcal{F}(w)})
=ψ(g)+𝔢(D,0)=ψ(g)0.\displaystyle=\psi(g)+\mathfrak{e}(D,0)=\psi(g)\neq 0.

By hypothesis,

t(ϕC,Λ(u),ϕΛ,C(u))+𝔢(D,Ω(u)Ω(u))=0,t(\phi_{C,\Lambda\mathcal{F}}(u),\phi_{\Lambda\mathcal{F},C}(u^{\prime}))+\mathfrak{e}(D,\Omega\mathcal{F}(u)\odot\Omega\mathcal{F}(u^{\prime}))=0,

for any uUu\in U and uUu^{\prime}\in U^{\prime}, so we have a contradiction. ∎

Corollary 5.4.

Let Hom(V,H)\mathcal{F}\in\operatorname{\mathrm{Hom}}(V,H), let CHom(V,G)C\in\operatorname{\mathrm{Hom}}(V,G^{*}), and let DSym2HD\in\operatorname{\mathrm{Sym}}^{2}H^{*}. Suppose UU is some submodule of VV such that U=H\mathcal{F}U=H. Let

σ={i[n]|t(ϕC,Λ(u),ϕΛ,C(vi))+𝔢(D,Ω(u)Ω(vi))0for some uU}.\sigma=\{i\in[n]\;|\;t(\phi_{C,\Lambda\mathcal{F}}(u),\phi_{\Lambda\mathcal{F},C}(v_{i}))+\mathfrak{e}(D,\Omega\mathcal{F}(u)\odot\Omega\mathcal{F}(v_{i}))\neq 0\ \textrm{for some }u\in U\}.

Then ker(ϕ,C|Vσ)=ker(|Vσ)\ker(\phi_{\mathcal{F},C}|_{V_{\sigma}})=\ker(\mathcal{F}|_{V_{\sigma}}). In particular, if |σ|<γn|\sigma|<\gamma n, then (C,D)(C,D) is weak for \mathcal{F}.

Proof.

By assumption, t(ϕC,Λ(u),ϕΛ,C(v))+𝔢(D,Ω(u)Ω(v))=0t(\phi_{C,\Lambda\mathcal{F}}(u),\phi_{\Lambda\mathcal{F},C}(v))+\mathfrak{e}(D,\Omega\mathcal{F}(u)\odot\Omega\mathcal{F}(v))=0 for all uUu\in U and vVσv\in V_{\sigma}. Applying Lemma 5.3 with U=VσU^{\prime}=V_{\sigma} tells us that ker(ϕ,C|Vσ)=ker(|Vσ)\ker(\phi_{\mathcal{F},C}|_{V_{\sigma}})=\ker(\mathcal{F}|_{V_{\sigma}}). ∎

Now, we define the important structural property of FF, which is a generalization of the notion of a linear code to RR-modules.

Definition 5.5.

For FHom(V,G)F\in\operatorname{\mathrm{Hom}}(V,G), we call FF a code of distance ww if for every σ[n]\sigma\subset[n] such that |σ|<w|\sigma|<w, we have FVσ=GFV_{\sigma}=G. Said another way, FF is not only a surjection, but would remain so if we threw out any fewer than ww standard basis vectors from VV.333We note that this definition is dual to the usual notion of a code. If RR is a field, then FF is a code if and only if the transpose Ft:VGF^{t}:V^{*}\to G^{*} is injective and its image im(Ft)V\operatorname{\mathrm{im}}(F^{t})\subset V^{*} is a code of distance ww in the usual sense.

The lift \mathcal{F} of a code FF to HH is still a code; importantly, note that the notion of being a code is a property we define for FHom(V,G)F\in\operatorname{\mathrm{Hom}}(V,G) rather than for the lift of FF. However, the methods we use to bound the multiplicands in (7) require \mathcal{F} to be a code as well.

Lemma 5.6.

Let FHom(V,G)F\in\operatorname{\mathrm{Hom}}(V,G) be a code of distance ww, and let Hom(V,H)\mathcal{F}\in\operatorname{\mathrm{Hom}}(V,H) be a lift of FF to HH. Then \mathcal{F} is a code of the same distance.

Proof.

Write both VV and HH as direct sums of their Sylow pp-subgroups, and note that any map VHV\to H is entirely determined by its restriction to VpV_{p} for each prime pp. Also note that (Vp)Hp\mathcal{F}(V_{p})\subset H_{p}. Suppose GpG_{p} is a pp-group of type λ\lambda. Since V=RnV=R^{n}, where R=/aR=\mathbb{Z}/a\mathbb{Z}, we see that Vp=Rpn=(/p2λ1)nV_{p}=R_{p}^{n}=(\mathbb{Z}/p^{2\lambda_{1}}\mathbb{Z})^{n}. Moreover, |Vσ,p=|(Vp)σ\mathcal{F}|_{V_{\sigma,p}}=\mathcal{F}|_{(V_{p})_{\sigma}}. Thus, it suffices to prove the lemma for GG a finite abelian pp-group.

Suppose GG is of type λ\lambda, where λ\lambda is the partition λ1λr\lambda_{1}\geq\cdots\geq\lambda_{r}. Then H=(/p2λ1)rH=(\mathbb{Z}/p^{2\lambda_{1}}\mathbb{Z})^{r}. Recall from Section 4 that we have a surjection Λ:HG\Lambda:H\twoheadrightarrow G and that Λ=F\Lambda\circ\mathcal{F}=F. Let σ[n]\sigma\subset[n] be a set of cardinality less than ww. We have the following sequence of maps

Vσ/pVσH/pHG/pG,V_{\sigma}/pV_{\sigma}\to H/pH\to G/pG,

where the first map is given by \mathcal{F} and the second by Λ\Lambda. The composite map Vσ/pVσG/pGV_{\sigma}/pV_{\sigma}\to G/pG is given by FF and is surjective since FF is a code of distance ww. Because H/pHH/pH and G/pGG/pG are 𝔽p\mathbb{F}_{p}-vector spaces of rank rr, it follows that the map Vσ/pVσH/pHV_{\sigma}/pV_{\sigma}\to H/pH must be surjective. By Nakayama’s lemma, it follows that :VσH\mathcal{F}:V_{\sigma}\to H must be surjective as well. Therefore, Vσ=H\mathcal{F}V_{\sigma}=H for any such σ\sigma. ∎

The following lemma will be useful often. In short, it allows us to choose a suitable basis of HH that works particularly nicely with the viv_{i}’s in VV in tandem with \mathcal{F} and CC. We will use what it says about codes along with the property of robustness to obtain a good lower bound on the number of Eij(C,D,)E_{ij}(C,D,\mathcal{F}) that are nonzero.

Lemma 5.7 (Lemma 3.4, [54]).

Let HH be a finite RR-module with Sylow pp-subgroup of type λ\lambda, where λ\lambda is a partition with rr parts. Suppose FHom(V,H)F\in\operatorname{\mathrm{Hom}}(V,H) is a code of distance δn\delta n, and let CHom(V,H)C\in\operatorname{\mathrm{Hom}}(V,H^{*}). Then we can find A1,,ArHA_{1},\ldots,A_{r}\in H and B1,,BrHB_{1},\ldots,B_{r}\in H^{*} such that for every 1ir1\leq i\leq r

#{j[n]|Fvj=AiandCvj=Bi}δn/|H|2,\#\{j\in[n]\;|\;Fv_{j}=A_{i}\ \mathrm{and\ }Cv_{j}=B_{i}\}\geq\delta n/|H|^{2},

and after the projection to the Sylow pp-subgroup HpH_{p} of HH, the elements A1,,ArA_{1},\ldots,A_{r} generate HpH_{p}.

Lemma 5.8 (Quadratically many nonzero coefficients for robust (C,D)(C,D)).

Let GG be a finite abelian group, and let PP be the set of primes dividing |G||G|. If Hom(V,H)\mathcal{F}\in\operatorname{\mathrm{Hom}}(V,H) is a code of distance δn\delta n, and if (C,D)Hom(V,G)×Sym2H(C,D)\in\operatorname{\mathrm{Hom}}(V,G^{*})\times\operatorname{\mathrm{Sym}}^{2}H^{*} is robust for \mathcal{F}, then there are at least γδn2/(2|H|2|P|)\gamma\delta n^{2}/(2|H|^{2}|P|) pairs (i,j)(i,j) with iji\leq j such that Eij(C,D,)0E_{ij}(C,D,\mathcal{F})\neq 0.

Proof.

For each pPp\in P, let HpH_{p} denote the Sylow pp-subgroup of HH, as usual. By taking the transpose of the surjection Λ:HG\Lambda:H\twoheadrightarrow G, there is an inclusion of GG^{*} into HH^{*}. Postcomposing with Λt\Lambda^{t} gives us an element ΛtCHom(V,H)\Lambda^{t}C\in\operatorname{\mathrm{Hom}}(V,H^{*}). With this in mind, apply Lemma 5.7 to HH, pp, \mathcal{F}, and ΛtC\Lambda^{t}C, and consider the resulting Ai(p)A_{i}(p) and Bi(p)B_{i}(p). Let

(8) τi(p)={j[n]|vj=Ai(p)andΛtCvj=Bi(p)},\tau_{i}(p)=\{j\in[n]\;|\;\mathcal{F}v_{j}=A_{i}(p)\ \mathrm{and\ }\Lambda^{t}Cv_{j}=B_{i}(p)\},

and let τ(p)=iτi(p)\tau(p)=\bigcup_{i}\tau_{i}(p). Let VpV_{p} be the submodule generated by the vjv_{j} for jτ(p)j\in\tau(p) so that Vp\mathcal{F}V_{p} in the projection to HpH_{p} is the entirety of HpH_{p}.

Next, let WW be the submodule of VV generated by the VpV_{p} for all pPp\in P; note that W=H\mathcal{F}W=H. Now, we apply Corollary 5.4 to WW. Since (C,D)(C,D) is robust for \mathcal{F}, we have

S:=#{i[n]|t(ϕC,Λ(w),ϕΛ,C(vi))+𝔢(D,Ω(w)Ω(vi))0 for some wW}γn.S:=\#\{i\in[n]\;|\;t(\phi_{C,\Lambda\mathcal{F}}(w),\phi_{\Lambda\mathcal{F},C}(v_{i}))+\mathfrak{e}(D,{\Omega\mathcal{F}(w)\odot\Omega\mathcal{F}(v_{i})})\neq 0\textrm{ for some }w\in W\}\geq\gamma n.

If viv_{i} pairs nontrivially with wWw\in W, then it must pair nontrivially with an element of one of the submodules generating WW, implying that

pP\displaystyle\sum_{p\in P} #{i[n]|t(ϕC,Λ(x),ϕΛ,C(vi))+𝔢(D,Ω(x)Ω(vi))0 for some xVp}S.\displaystyle\#\{i\in[n]\;|\;t(\phi_{C,\Lambda\mathcal{F}}(x),\phi_{\Lambda\mathcal{F},C}(v_{i}))+\mathfrak{e}(D,{\Omega\mathcal{F}(x)\odot\Omega\mathcal{F}(v_{i})})\neq 0\textrm{ for some }x\in V_{p}\}\geq S.

Hence, for some pPp\in P,

#{i[n]|t(ϕC,Λ(x),ϕΛ,C(vi))+𝔢(D,Ω(x)Ω(vi))0 for some xVp}γn/|P|.\#\{i\in[n]\;|\;t(\phi_{C,\Lambda\mathcal{F}}(x),\phi_{\Lambda\mathcal{F},C}(v_{i}))+\mathfrak{e}(D,{\Omega\mathcal{F}(x)\odot\Omega\mathcal{F}(v_{i})})\neq 0\textrm{ for some }x\in V_{p}\}\geq\gamma n/|P|.

For that particular pp, it follows that

#{i[n]|t(ϕC,Λ(vj),ϕΛ,C(vi))+𝔢(D,Ω(vj)Ω(vi))0forsomejτ(p)}γn/|P|.\#\{i\in[n]\;|\;t(\phi_{C,\Lambda\mathcal{F}}(v_{j}),\phi_{\Lambda\mathcal{F},C}(v_{i}))+\mathfrak{e}(D,{\Omega\mathcal{F}(v_{j})\odot\Omega\mathcal{F}(v_{i})})\neq 0\ \mathrm{for\ some\ }j\in\tau(p)\}\geq\gamma n/|P|.

By (8), for any jτ(p)j\in\tau(p), there are at least δn/|H|2\delta n/|H|^{2} indices kτ(p)k\in\tau(p) such that vk=vj\mathcal{F}v_{k}=\mathcal{F}v_{j} and ΛtCvk=ΛtCvj\Lambda^{t}Cv_{k}=\Lambda^{t}Cv_{j}. Moreover, note that e((u),ΛtC(v))=C(v)(Λ(u))=e(Λ(u),C(v))e(\mathcal{F}(u),\Lambda^{t}C(v))=C(v)(\Lambda\mathcal{F}(u))=e(\Lambda\mathcal{F}(u),C(v)) for any u,vVu,v\in V. Thus, for all indices kτ(p)k\in\tau(p) such that vk=vj\mathcal{F}v_{k}=\mathcal{F}v_{j} and ΛtCvk=ΛtCvj\Lambda^{t}Cv_{k}=\Lambda^{t}Cv_{j}, we have

t(ϕC,Λ(vj),ϕΛ,C(vi))\displaystyle t(\phi_{C,\Lambda\mathcal{F}}(v_{j}),\phi_{\Lambda\mathcal{F},C}(v_{i})) +𝔢(D,ΩvjΩvi)\displaystyle+\mathfrak{e}(D,{\Omega\mathcal{F}v_{j}\odot\Omega\mathcal{F}v_{i}})
=t(ϕC,Λ(vk),ϕΛ,C(vi))+𝔢(D,Ω(vk)Ω(vi)).\displaystyle=t(\phi_{C,\Lambda\mathcal{F}}(v_{k}),\phi_{\Lambda\mathcal{F},C}(v_{i}))+\mathfrak{e}(D,{\Omega\mathcal{F}(v_{k})\odot\Omega\mathcal{F}(v_{i})}).

For all i<ji<j, notice that

Eij(C,D,)=t(ϕC,Λ(vj),ϕΛ,C(vi))+𝔢(D,Ω(vj)Ω(vi))E_{ij}(C,D,\mathcal{F})=t(\phi_{C,\Lambda\mathcal{F}}(v_{j}),\phi_{\Lambda\mathcal{F},C}(v_{i}))+\mathfrak{e}(D,{\Omega\mathcal{F}(v_{j})\odot\Omega\mathcal{F}(v_{i})})

and also that

2Eii(C,D,)=t(ϕC,Λ(vi),ϕΛ,C(vi))+𝔢(D,Ω(vi)Ω(vi)).2E_{ii}(C,D,\mathcal{F})=t(\phi_{C,\Lambda\mathcal{F}}(v_{i}),\phi_{\Lambda\mathcal{F},C}(v_{i}))+\mathfrak{e}(D,{\Omega\mathcal{F}(v_{i})\odot\Omega\mathcal{F}(v_{i})}).

Hence, there must be at least γδn2/(2|H|2|P|)\gamma\delta n^{2}/(2|H|^{2}|P|) pairs (i,j)(i,j) such that Eij(C,D,)0E_{ij}(C,D,\mathcal{F})\neq 0. ∎

We next study the number of nonzero coefficients for weak (C,D)(C,D). To do so, we will consider the pairs (C,D)(C,D) such that all of the coefficients are 0 (e.g., (0,0)(0,0) is one such pair, but there are many others). We will first need to adopt a more equivariant view of the coefficients Eij(C,D,)E_{ij}(C,D,\mathcal{F}). The evaluation map e:GGRe:G^{*}\otimes G\to R gives rise to a natural map

Hom(V,G)Hom(V,G)(VG)(VG)VV,\operatorname{\mathrm{Hom}}(V,G)\otimes\operatorname{\mathrm{Hom}}(V,G^{*})\simeq(V^{*}\otimes G)\otimes(V^{*}\otimes G^{*})\to V^{*}\otimes V^{*},

and we may further compose with the quotient VVSym2VV^{*}\otimes V^{*}\to\operatorname{\mathrm{Sym}}^{2}V^{*} to get a map

Φ:Hom(V,G)Hom(V,G)VVSym2V.\Phi:\operatorname{\mathrm{Hom}}(V,G)\otimes\operatorname{\mathrm{Hom}}(V,G^{*})\to V^{*}\otimes V^{*}\to\operatorname{\mathrm{Sym}}^{2}V^{*}.

For Hom(V,H)\mathcal{F}\in\operatorname{\mathrm{Hom}}(V,H), composing with Λ\Lambda gives us a map ΛHom(V,G)\Lambda\mathcal{F}\in\operatorname{\mathrm{Hom}}(V,G).

Likewise, the evaluation 𝔢:Sym2HSym2HR\mathfrak{e}:\operatorname{\mathrm{Sym}}_{2}H\otimes\operatorname{\mathrm{Sym}}^{2}H^{*}\to R induces a natural map

Ψ:Hom(Sym2V,Sym2H)Sym2H(Sym2V)Sym2HSym2H(Sym2V).\Psi:\operatorname{\mathrm{Hom}}(\operatorname{\mathrm{Sym}}_{2}V,\operatorname{\mathrm{Sym}}_{2}H)\otimes\operatorname{\mathrm{Sym}}^{2}H^{*}\simeq(\operatorname{\mathrm{Sym}}_{2}V)^{*}\otimes\operatorname{\mathrm{Sym}}_{2}H\otimes\operatorname{\mathrm{Sym}}^{2}H^{*}\to(\operatorname{\mathrm{Sym}}_{2}V)^{*}.

Recall that (Sym2V)(\operatorname{\mathrm{Sym}}_{2}V)^{*} is naturally isomorphic to Sym2V\operatorname{\mathrm{Sym}}^{2}V^{*}. Given Hom(V,G)\mathcal{F}\in\operatorname{\mathrm{Hom}}(V,G), we get an element of Hom(Sym2V,Sym2H)\operatorname{\mathrm{Hom}}(\operatorname{\mathrm{Sym}}_{2}V,\operatorname{\mathrm{Sym}}_{2}H) by considering ΩΩ\Omega\mathcal{F}\otimes\Omega\mathcal{F}. Taking the sum of the aforementioned maps Φ\Phi and Ψ\Psi (not the direct sum), given Hom(V,H)\mathcal{F}\in\operatorname{\mathrm{Hom}}(V,H), we can construct a map

m:Hom(V,G)Sym2HSym2Vm_{\mathcal{F}}:\operatorname{\mathrm{Hom}}(V,G^{*})\oplus\operatorname{\mathrm{Sym}}^{2}H^{*}\to\operatorname{\mathrm{Sym}}^{2}V^{*}

given by

m(C,D)=Φ(ΛC)+Ψ((ΩΩ)D).m_{\mathcal{F}}(C,D)=\Phi(\Lambda\mathcal{F}\otimes C)+\Psi((\Omega\mathcal{F}\otimes\Omega\mathcal{F})\otimes D).

We see that mm_{\mathcal{F}} can be written explicitly as

(9) m(C,D)=i=1nj=i+1n(e(C(vj),Λ(vi))+e(C(vi),Λ(vj))+𝔢(D,Ω(vi)Ω(vj)))vivj+i=1n(e(C(vi),Λ(vi))+𝔢(D,Ω(vi)Ω(vi)))(vi)2.\begin{split}m_{\mathcal{F}}(C,D)=&\sum_{i=1}^{n}\sum_{j=i+1}^{n}\Big{(}e(C(v_{j}),\Lambda\mathcal{F}(v_{i}))+e(C(v_{i}),\Lambda\mathcal{F}(v_{j}))+\mathfrak{e}(D,{\Omega\mathcal{F}(v_{i})\odot\Omega\mathcal{F}(v_{j})})\Big{)}v_{i}^{*}v_{j}^{*}\\ &\qquad\qquad+\sum_{i=1}^{n}\Big{(}e(C(v_{i}),\Lambda\mathcal{F}(v_{i}))+\mathfrak{e}(D,\Omega\mathcal{F}(v_{i})\otimes\Omega\mathcal{F}(v_{i}))\Big{)}(v_{i}^{*})^{2}.\end{split}

We remark that the definition of mm_{\mathcal{F}} is canonical and basis independent. In particular, the above could be rewritten by replacing the viv_{i}’s with the elements of any other basis of VV. Note that the elements in the kernel of mm_{\mathcal{F}} are precisely the (C,D)(C,D) for which all of the coefficients Eij(C,D,)E_{ij}(C,D,\mathcal{F}) are zero. We call the (C,D)ker(m)(C,D)\in\ker(m_{\mathcal{F}}) special for \mathcal{F}.

Recall that the data of the pairing on G^\widehat{G} is equivalent to an element AA of Sym2H\operatorname{\mathrm{Sym}}_{2}H whose precise definition is given in Section 4.

Lemma 5.9.

Given Hom(V,H)\mathcal{F}\in\operatorname{\mathrm{Hom}}(V,H) with V=H\mathcal{F}V=H, there are |Sym2H|/|G|{|\operatorname{\mathrm{Sym}}^{2}H|}/{|G|} special (C,D)(C,D) for \mathcal{F}. Moreover, if (C,D)(C,D) is special for Hom(V,H)\mathcal{F}\in\operatorname{\mathrm{Hom}}(V,H), then D(A)=0D(-A)=0.

Proof.

Because everything in sight can be written as a direct sum of Sylow pp-subgroups, we can assume without loss of generality that GG is a pp-group of type λ\lambda (and, accordingly, that R=/p2λ1R=\mathbb{Z}/p^{2\lambda_{1}}\mathbb{Z}). Suppose λ\lambda has rr parts, and recall that H=RrH=R^{r}. Because \mathcal{F} is a surjection, there exist w1,,wrVw_{1},\ldots,w_{r}\in V that can be completed to a free RR-module basis w1,,wnw_{1},\ldots,w_{n} of VV such that (wi)=hi\mathcal{F}(w_{i})=h_{i} and (wi)=0\mathcal{F}(w_{i})=0 for i>ri>r (recall hih_{i} is the iith standard basis vector in HH). With respect to this basis, mm_{\mathcal{F}} takes (C,D)(C,D) to

i=1nj=i+1n(e(C(wj),Λ(wi))+e(C(wi),Λ(wj))+𝔢(D,Ω(wi)Ω(wj)))wiwj\displaystyle\sum_{i=1}^{n}\sum_{j=i+1}^{n}\Big{(}e(C(w_{j}),\Lambda\mathcal{F}(w_{i}))+e(C(w_{i}),\Lambda\mathcal{F}(w_{j}))+\mathfrak{e}(D,{\Omega\mathcal{F}(w_{i})\odot\Omega\mathcal{F}(w_{j})})\Big{)}w_{i}^{*}w_{j}^{*}
+i=1n(e(C(wi),Λ(wi))+𝔢(D,Ω(wi)Ω(wi)))(wi)2.\displaystyle\qquad\qquad+\sum_{i=1}^{n}\Big{(}e(C(w_{i}),\Lambda\mathcal{F}(w_{i}))+\mathfrak{e}(D,\Omega\mathcal{F}(w_{i})\otimes\Omega\mathcal{F}(w_{i}))\Big{)}(w_{i}^{*})^{2}.

We compute the coefficients of each wiwjw_{i}^{*}w_{j}^{*} explicitly. For 1i<jr1\leq i<j\leq r, the coefficient on wiwjw_{i}^{*}w_{j}^{*} is

e(C(wj),p2λ1λihi)+e(C(wi),p2λ1λjhj)+p2λ1λiλj𝔢(D,hihj).e(C(w_{j}),p^{2\lambda_{1}-\lambda_{i}}h_{i})+e(C(w_{i}),p^{2\lambda_{1}-\lambda_{j}}h_{j})+p^{2\lambda_{1}-\lambda_{i}-\lambda_{j}}\mathfrak{e}(D,{h_{i}\odot h_{j}}).

Recall that DD has a unique upper-triangular representative D0HHD_{0}\in H^{*}\otimes H^{*} with respect to the basis h1,,hrh_{1}^{*},\ldots,h_{r}^{*} and that we may do all our computations with respect to this representative element D0D_{0}. Since D0D_{0} is upper-triangular with respect to the basis h1,,hrh_{1}^{*},\ldots,h_{r}^{*}, the coefficient of wiwjw_{i}^{*}w_{j}^{*} for 1i<jr1\leq i<j\leq r is

Cji+Cij+p2λ1λiλjDij,C_{ji}+C_{ij}+p^{2\lambda_{1}-\lambda_{i}-\lambda_{j}}D_{ij},

where CijC_{ij} denotes C(wi)(p2λ1λjhj)C(w_{i})(p^{2\lambda_{1}-\lambda_{j}}h_{j}) and DijD_{ij} denotes D0(hi)(hj)D_{0}(h_{i})(h_{j}) (recall that CC takes elements of VV to G=Hom(ΛH,R)G^{*}=\operatorname{\mathrm{Hom}}(\Lambda H,R)). Note that pλjCij=0Rp^{\lambda_{j}}C_{ij}=0\in R, i.e., Cij0C_{ij}\equiv 0 mod p2λ1λjp^{2\lambda_{1}-\lambda_{j}}. For i=ji=j, the coefficient on wiwjw_{i}^{*}w_{j}^{*} is Cii+p2λ12λiDii.C_{ii}+p^{2\lambda_{1}-2\lambda_{i}}D_{ii}. If 1ir<j1\leq i\leq r<j, the coefficient on wiwjw_{i}^{*}w_{j}^{*} is Cji.C_{ji}.

For (C,D)(C,D) to be special for \mathcal{F}, all of these coefficients must be 0 in RR. Setting all of these coefficients to 0 gives us the following system of equations

(10) {Cji+Cij+p2λ1λiλjDij=0for 1i<jr;Cii+p2λ12λiDii=0for 1ir;Cji=0for 1ir<jn.\begin{cases}C_{ji}+C_{ij}+p^{2\lambda_{1}-\lambda_{i}-\lambda_{j}}D_{ij}=0&\textrm{for }1\leq i<j\leq r;\\ C_{ii}+p^{2\lambda_{1}-2\lambda_{i}}D_{ii}=0&\textrm{for }1\leq i\leq r;\\ C_{ji}=0&\textrm{for }1\leq i\leq r<j\leq n.\end{cases}

modulo p2λ1p^{2\lambda_{1}}, which the entries of CC and DD satisfy if and only if (C,D)(C,D) is special. We count the pairs (C,D)(C,D) satisfying this system of equations. Recall that G=ΛHG=\Lambda H and that (C,D)Hom(V,G)×Sym2H(C,D)\in\operatorname{\mathrm{Hom}}(V,G^{*})\times\operatorname{\mathrm{Sym}}^{2}H^{*}. Hence, the possible CHom(V,G)C\in\operatorname{\mathrm{Hom}}(V,G^{*}) are given by choosing CijC_{ij} to be any multiple of p2λ1λjp^{2\lambda_{1}-\lambda_{j}} in RR for each 1in1\leq i\leq n and 1jr1\leq j\leq r (recall that Cij0C_{ij}\equiv 0 mod p2λ1λjp^{2\lambda_{1}-\lambda_{j}}). Because DD is determined by its unique upper-triangular representative D0D_{0}, the possible DD are given by choosing the entries DijRD_{ij}\in R for 1ijr1\leq i\leq j\leq r. The DijD_{ij}’s determine D0D_{0}, which determines DD. Since DSym2HD\in\operatorname{\mathrm{Sym}}^{2}H^{*}, the DijD_{ij}’s can be any elements of RR.

Equation 10 determines CjiC_{ji} for 1ir<jn1\leq i\leq r<j\leq n. In order for there to be solutions DijD_{ij}, we need Cji+Cij0C_{ji}+C_{ij}\equiv 0 mod p2λ1λiλjp^{2\lambda_{1}-\lambda_{i}-\lambda_{j}}. This holds because Cij0C_{ij}\equiv 0 modulo p2λ1λjp^{2\lambda_{1}-\lambda_{j}}. Thus, for 1i,jr1\leq i,j\leq r, choose the CjiC_{ji}’s freely; there are pλip^{\lambda_{i}} choices for each CjiC_{ji}. Equation 10 then determines the DijD_{ij} modulo pλi+λjp^{\lambda_{i}+\lambda_{j}} for all 1ijr1\leq i\leq j\leq r. It follows that there are p2λ1λiλjp^{2\lambda_{1}-\lambda_{i}-\lambda_{j}} choices for each DijRD_{ij}\in R such that (10) is satisfied. Therefore, there are

1irprλi1ijrp2λ1λiλj=|G|rp2λ1(r+12)p1ijr(λi+λj)=|G|r|Sym2H||G|r+1=|Sym2H||G|\prod_{1\leq i\leq r}p^{r\lambda_{i}}\cdot\prod_{1\leq i\leq j\leq r}p^{2\lambda_{1}-\lambda_{i}-\lambda_{j}}=\frac{|G|^{r}p^{2\lambda_{1}{{r+1}\choose{2}}}}{p^{\sum_{1\leq i\leq j\leq r}(\lambda_{i}+\lambda_{j})}}=\frac{|G|^{r}|\operatorname{\mathrm{Sym}}^{2}H|}{|G|^{r+1}}=\frac{|\operatorname{\mathrm{Sym}}^{2}H|}{|G|}

special (C,D)(C,D), as desired.

To prove the second part of the lemma, we again reduce to the case to where GG is a pp-group of type λ\lambda, where λ\lambda is the partition λ1λr\lambda_{1}\geq\cdots\geq\lambda_{r}, and R=/p2λ1R=\mathbb{Z}/p^{2\lambda_{1}}\mathbb{Z}. Recall ASym2HA\in\operatorname{\mathrm{Sym}}_{2}H is the element

1i<jrp2λ1λjaijhihj+i=1rp2λ1λiaiihihi.\sum_{1\leq i<j\leq r}p^{2\lambda_{1}-\lambda_{j}}a_{ij}h_{i}\odot h_{j}+\sum_{i=1}^{r}p^{2\lambda_{1}-\lambda_{i}}a_{ii}h_{i}\otimes h_{i}.

Then we may compute D(A)=𝔢(D,A)D(-A)=\mathfrak{e}(D,-A) with respect to the unique upper-triangular representative D0D_{0} of DD from above. By hypothesis, (C,D)(C,D) is special for \mathcal{F}. Hence, for 1i<jr1\leq i<j\leq r, we have

Cji+Cij+p2λ1λiλjDij=0modp2λ1,C_{ji}+C_{ij}+p^{2\lambda_{1}-\lambda_{i}-\lambda_{j}}D_{ij}=0\mod p^{2\lambda_{1}},

and for 1i=jr1\leq i=j\leq r, we have Cii+p2λ12λiDii=0modp2λ1C_{ii}+p^{2\lambda_{1}-2\lambda_{i}}D_{ii}=0\mod p^{2\lambda_{1}}. Since Cij0C_{ij}\equiv 0 mod p2λ1λjp^{2\lambda_{1}-\lambda_{j}}, we have that Dij0D_{ij}\equiv 0 mod pλjp^{\lambda_{j}} for all iji\leq j. Thus,

𝔢(D,A)\displaystyle\mathfrak{e}(D,A) =1i<jrp2λ1λjaij(D0(hi)(hj)+D0(hj)(hi))+i=1rp2λ1λiaiiD0(hi)(hi)\displaystyle=\sum_{1\leq i<j\leq r}p^{2\lambda_{1}-\lambda_{j}}a_{ij}(D_{0}(h_{i})(h_{j})+D_{0}(h_{j})(h_{i}))+\sum_{i=1}^{r}p^{2\lambda_{1}-\lambda_{i}}a_{ii}D_{0}(h_{i})(h_{i})
=1i<jrp2λ1λjaijDij+i=1rp2λ1λiaiiDii\displaystyle=\sum_{1\leq i<j\leq r}p^{2\lambda_{1}-\lambda_{j}}a_{ij}D_{ij}+\sum_{i=1}^{r}p^{2\lambda_{1}-\lambda_{i}}a_{ii}D_{ii}
=0R,\displaystyle=0\in R,

where the second equality follows from the fact that D0D_{0} is upper-triangular. ∎

Now, we will use Lemma 5.9 to prove that as long as (C,D)(C,D) is not special for \mathcal{F} there are linearly many nonzero coefficients Eij(C,D,)E_{ij}(C,D,\mathcal{F}).

Lemma 5.10 (Linearly many nonzero coefficients for non-special (C,D)(C,D)).

Let Hom(V,H)\mathcal{F}\in\operatorname{\mathrm{Hom}}(V,H) be a code of distance δn\delta n, and suppose (C,D)Hom(V,G)×Sym2H(C,D)\in\operatorname{\mathrm{Hom}}(V,G^{*})\times\operatorname{\mathrm{Sym}}^{2}H^{*} is not special for \mathcal{F} so that (C,D)ker(m)(C,D)\not\in\ker(m_{\mathcal{F}}). Then there are at least δn/2\delta n/2 pairs (i,j)[n]2(i,j)\in[n]^{2} with iji\leq j such that Eij(C,D,)0E_{ij}(C,D,\mathcal{F})\neq 0.

Proof.

Suppose for a contradiction that there are fewer than δn/2\delta n/2 pairs. Let π\pi be the set of pairs (i,j)[n]2(i,j)\in[n]^{2} with iji\leq j and Eij(C,D,)0E_{ij}(C,D,\mathcal{F})\neq 0. By assumption, |π|<δn/2|\pi|<\delta n/2. Let σ\sigma be the set of all indices a[n]a\in[n] occurring in a pair (i,j)π(i,j)\in\pi. Immediately, we see that |σ|<δn|\sigma|<\delta n, and if (i,j)(i,j) is a pair with ii or jj not in σ\sigma, then Eij(C,D,)E_{ij}(C,D,\mathcal{F}) must be 0.

We will use this setup to give a lower bound on the size of im(m)\operatorname{\mathrm{im}}(m_{\mathcal{F}}). Recall the definition of mm_{\mathcal{F}} from (9): (C,D)(C,D) is sent to the element of Sym2V\operatorname{\mathrm{Sym}}^{2}V^{*} given by

i=1nj=i+1n(e(C(vj),Λ(vi))+e(C(vi),Λ(vj))+𝔢(D,Ω(vi)Ω(vj)))vivj\displaystyle\sum_{i=1}^{n}\sum_{j=i+1}^{n}\Big{(}e(C(v_{j}),\Lambda\mathcal{F}(v_{i}))+e(C(v_{i}),\Lambda\mathcal{F}(v_{j}))+\mathfrak{e}(D,{\Omega\mathcal{F}(v_{i})\odot\Omega\mathcal{F}(v_{j})})\Big{)}v_{i}^{*}v_{j}^{*}
+i=1n(e(C(vi),Λ(vi))+𝔢(D,Ω(vi)Ω(vi)))(vi)2.\displaystyle\qquad\qquad+\sum_{i=1}^{n}\Big{(}e(C(v_{i}),\Lambda\mathcal{F}(v_{i}))+\mathfrak{e}(D,\Omega\mathcal{F}(v_{i})\otimes\Omega\mathcal{F}(v_{i}))\Big{)}(v_{i}^{*})^{2}.

Compose mm_{\mathcal{F}} with the quotient map taking vivj0v_{i}^{*}v_{j}^{*}\mapsto 0 if both i,jσi,j\in\sigma, and call the resulting map m:Hom(V,G)×Sym2HZm_{\mathcal{F}}^{\prime}:\operatorname{\mathrm{Hom}}(V,G^{*})\times\operatorname{\mathrm{Sym}}^{2}H^{*}\to Z, where ZZ denotes our new quotient space. Under mm_{\mathcal{F}}^{\prime}, we see that (C,D)0(C,D)\mapsto 0 (by hypothesis (C,D)ker(m)(C,D)\not\in\ker(m_{\mathcal{F}}) since it is not special). We will arrive at a contradiction by proving that #im(m)=#(Hom(V,G)×Sym2H)/#ker(m)\#\operatorname{\mathrm{im}}(m_{\mathcal{F}})=\#(\operatorname{\mathrm{Hom}}(V,G^{*})\times\operatorname{\mathrm{Sym}}^{2}H^{*})/\#\ker(m_{\mathcal{F}}) divides #im(m)\#\operatorname{\mathrm{im}}(m_{\mathcal{F}}^{\prime}). Then, since #im(m)|#im(m)\#\operatorname{\mathrm{im}}(m_{\mathcal{F}}^{\prime})|\#\operatorname{\mathrm{im}}(m_{\mathcal{F}}), this will force ker(m)=ker(m)\ker(m_{\mathcal{F}})=\ker(m_{\mathcal{F}}^{\prime}), which is a contradiction. By Lemma 5.9,

|im(m)|=|Hom(V,G)×Sym2H||ker(m)|=|G|n+1.|\operatorname{\mathrm{im}}(m_{\mathcal{F}})|=\frac{|\operatorname{\mathrm{Hom}}(V,G^{*})\times\operatorname{\mathrm{Sym}}^{2}H^{*}|}{|\ker(m_{\mathcal{F}})|}=|G|^{n+1}.

As in the proof of Lemma 5.9, we can prove |G|n+1|G|^{n+1} divides |im(m)||\operatorname{\mathrm{im}}(m_{\mathcal{F}}^{\prime})| by reducing to the case where GG is a pp-group of type λ\lambda, where λ\lambda is the partition λ1λr\lambda_{1}\geq\cdots\geq\lambda_{r}. Accordingly, assume that R=/p2λ1R=\mathbb{Z}/p^{2\lambda_{1}}\mathbb{Z}.

Because \mathcal{F} is a code of distance δn\delta n, and because |σ|<δn|\sigma|<\delta n, we have that Vσ=H\mathcal{F}V_{\sigma}=H. Hence |Vσ\mathcal{F}|_{V_{\sigma}} is a code of some positive distance. Apply Lemma 5.7 to |Vσ\mathcal{F}|_{V_{\sigma}} to get τ[n]σ\tau\subset[n]\setminus\sigma such that |τ|=r|\tau|=r and vi\mathcal{F}v_{i} generate HH for iτi\in\tau. Let WW be the RR-submodule of VV generated by the viv_{i} for iτi\in\tau. Let h1,hrh_{1}\ldots,h_{r} be generators for HH with relations p2λ1hi=0p^{2\lambda_{1}}h_{i}=0, and let hih_{i}^{*} denote the corresponding dual basis for HH^{*} so that hi(hj)=δijh_{i}^{*}(h_{j})=\delta_{ij}. For 1jr1\leq j\leq r, suppose wjWw_{j}\in W is such that wj=hj\mathcal{F}w_{j}=h_{j}, and let UWU\subset W be the submodule of WW generated by the wjw_{j}. Now, we have maps

U/pUW/pWH/pH,U/pU\to W/pW\to H/pH,

where the second map is given by \mathcal{F}. Each of the above spaces is an 𝔽p\mathbb{F}_{p}-vector space of rank at most rr, exactly rr, and exactly rr, respectively. Because the composite map U/pUH/pHU/pU\to H/pH is a surjection, our considerations on the rank of the spaces tell us that the map U/pUW/pWU/pU\to W/pW is a surjection. Thus, Nakayama’s lemma forces U=WU=W. The elements w1,,wrw_{1},\ldots,w_{r} generate the free RR-module UU of rank rr, so they must form a basis for WW as a free RR-module. Thus, we have found a basis for WW such that (wi)=hi\mathcal{F}(w_{i})=h_{i}.

Consider the following basis for VV, given by replacing the viv_{i} for iτi\in\tau with the wkw_{k}’s. Let τ={τ1,,τr}\tau=\{\tau_{1},\ldots,\tau_{r}\}. For iτi\in\tau, let zi=wτiz_{i}=w_{\tau_{i}}, and for iτi\not\in\tau set zi=viz_{i}=v_{i}. Because the wjw_{j}’s are a basis for WW, it follows that z1,,znz_{1},\ldots,z_{n} forms a basis for VV; denote the corresponding dual basis by z1,,znz_{1}^{*},\ldots,z_{n}^{*}, where zi(zj)=δijz_{i}^{*}(z_{j})=\delta_{ij}.

View Hom(V,H)\mathcal{F}\in\operatorname{\mathrm{Hom}}(V,H) as an element of VHV^{*}\otimes H and write

=1in1jrfijzihj\mathcal{F}=\sum_{\begin{subarray}{c}1\leq i\leq n\\ 1\leq j\leq r\end{subarray}}f_{ij}z_{i}^{*}\otimes h_{j}

so that (z)=1krfkhk\mathcal{F}(z_{\ell})=\sum_{1\leq k\leq r}f_{\ell k}h_{k}. Because mm_{\mathcal{F}} was defined equivariantly, we see that m(C,D)m_{\mathcal{F}}(C,D) is given by

i=1nj>in(e(C(zj),Λ(zi))+e(C(zi),Λ(zj))+𝔢(D,Ω(zj)Ω(zi)))zizj\displaystyle\sum_{i=1}^{n}\sum_{j>i}^{n}\Big{(}e(C(z_{j}),\Lambda\mathcal{F}(z_{i}))+e(C(z_{i}),\Lambda\mathcal{F}(z_{j}))+\mathfrak{e}(D,{\Omega\mathcal{F}(z_{j})\odot\Omega\mathcal{F}(z_{i})})\Big{)}z_{i}^{*}z_{j}^{*}
+i=1n(e(C(zi),Λ(zi))+𝔢(D,Ω(zi)Ω(zi)))(zi)2\displaystyle\quad\quad\quad+\sum_{i=1}^{n}\Big{(}e(C(z_{i}),\Lambda\mathcal{F}(z_{i}))+\mathfrak{e}(D,\Omega\mathcal{F}(z_{i})\otimes\Omega\mathcal{F}(z_{i}))\Big{)}(z_{i}^{*})^{2}

which we may rewrite as

i=1nj>in(e(C(zj),1krΛ(fikhk))+e(C(zi),1krΛ(fjkhk))+𝔢(D,Ω(zj)Ω(zj)))zizj\displaystyle\sum_{i=1}^{n}\sum_{j>i}^{n}\Big{(}e(C(z_{j}),\textstyle\sum_{1\leq k\leq r}\Lambda(f_{ik}h_{k}))+e(C(z_{i}),\textstyle\sum_{1\leq k\leq r}\Lambda(f_{jk}h_{k}))+\mathfrak{e}(D,{\Omega\mathcal{F}(z_{j})\odot\Omega\mathcal{F}(z_{j})})\Big{)}z_{i}^{*}z_{j}^{*}
+i=1n(e(C(zi),1krΛ(fikhk))+𝔢(D,Ω(zi)Ω(zi)))(zi)2.\displaystyle\quad\quad\quad+\sum_{i=1}^{n}\Big{(}e(C(z_{i}),\textstyle\sum_{1\leq k\leq r}\Lambda(f_{ik}h_{k}))+\mathfrak{e}(D,\Omega\mathcal{F}(z_{i})\otimes\Omega\mathcal{F}(z_{i}))\Big{)}(z_{i}^{*})^{2}.

Hence,

m(zhm,0)\displaystyle m_{\mathcal{F}}(z_{\ell}^{*}\otimes h_{m}^{*},0) =i=1nj=1,jin(e(z(zj)hm,1krp2λ1λkfikhk))zizj\displaystyle=\sum_{i=1}^{n}\sum_{j=1,j\neq i}^{n}\Big{(}e(z_{\ell}^{*}(z_{j})\otimes h_{m}^{*},\textstyle\sum_{1\leq k\leq r}p^{2\lambda_{1}-\lambda_{k}}f_{ik}h_{k})\Big{)}z_{i}^{*}z_{j}^{*}
+i=1n(e(z(zi)hm,1krp2λ1λkfikhk))(zi)2.\displaystyle\quad\quad\quad+\displaystyle\sum_{i=1}^{n}\Big{(}e(z_{\ell}^{*}(z_{i})\otimes h_{m}^{*},\textstyle\sum_{1\leq k\leq r}p^{2\lambda_{1}-\lambda_{k}}f_{ik}h_{k})\Big{)}(z_{i}^{*})^{2}.
=i=1nj=1,jinp2λ1λmfimz(zj)(zizj)+i=1np2λ1λmfimz(zi)(zi)2\displaystyle=\sum_{i=1}^{n}\sum_{j=1,j\neq i}^{n}p^{2\lambda_{1}-\lambda_{m}}f_{im}z_{\ell}^{*}(z_{j})(z_{i}^{*}z_{j}^{*})+\sum_{i=1}^{n}p^{2\lambda_{1}-\lambda_{m}}f_{im}z_{\ell}^{*}(z_{i})(z_{i}^{*})^{2}
=i=1np2λ1λmfimziz.\displaystyle=\sum_{i=1}^{n}p^{2\lambda_{1}-\lambda_{m}}f_{im}z_{i}^{*}z_{\ell}^{*}.

Now, further quotient ZZ by the submodule generated by the vivjv_{i}^{*}v_{j}^{*} such that neither ii nor jj are in τ\tau. Call the resulting space ZZ^{\prime} and denote the composite Hom(V,G)×Sym2HZ\operatorname{\mathrm{Hom}}(V,G^{*})\times\operatorname{\mathrm{Sym}}^{2}H^{*}\to Z^{\prime} by m′′m_{\mathcal{F}}^{\prime\prime}. Because (zt)=1krftkhk\mathcal{F}(z_{t})=\sum_{1\leq k\leq r}f_{tk}h_{k} and (zτi)=hi\mathcal{F}(z_{\tau_{i}})=h_{i}, it follows that fτij=δijf_{\tau_{i}j}=\delta_{ij}. Therefore, if τ\ell\not\in\tau,

(11) m′′(zhm,0)=iτp2λ1λmfimziz=i=1rp2λ1λmfτimzτiz=p2λ1λmwmz.m_{\mathcal{F}}^{\prime\prime}(z_{\ell}^{*}\otimes h_{m}^{*},0)=\sum_{i\in\tau}p^{2\lambda_{1}-\lambda_{m}}f_{im}z_{i}^{*}z_{\ell}^{*}=\sum_{i=1}^{r}p^{2\lambda_{1}-\lambda_{m}}f_{\tau_{i}m}z_{\tau_{i}}^{*}z_{\ell}^{*}=p^{2\lambda_{1}-\lambda_{m}}w_{m}^{*}z_{\ell}^{*}.

The element p2λ1λmwmzp^{2\lambda_{1}-\lambda_{m}}w_{m}^{*}z_{\ell}^{*} has order pλmp^{\lambda_{m}}, implying that im(m′′)\operatorname{\mathrm{im}}(m_{\mathcal{F}}^{\prime\prime}) has a subgroup of order |G|nr|G|^{n-r}, as we can take any [n]τ\ell\in[n]\setminus\tau and 1mr1\leq m\leq r.

Continuing, further quotient ZZ^{\prime} by the submodule generated by the zizjz_{i}^{*}z_{j}^{*} such that one of ii or jj is not in τ\tau. Denote the resulting space by Z′′Z^{\prime\prime} and map by m′′′m_{\mathcal{F}}^{\prime\prime\prime}. Note that the only basis elements zizjz_{i}^{*}z_{j}^{*} of Sym2V\operatorname{\mathrm{Sym}}^{2}V^{*} whose images are nonzero in Z′′Z^{\prime\prime} are those with both i,jτi,j\in\tau. Also, note that the subgroup of size |G|nr|G|^{n-r} we identified previously is sent to 0 under m′′′m_{\mathcal{F}}^{\prime\prime\prime}. Suppose 1yxr1\leq y\leq x\leq r and let Dxy:HHD_{xy}:H\to H^{*} be the map hxhyh_{x}^{*}\otimes h_{y}^{*} taking hxhyh_{x}\mapsto h_{y}^{*} and all other hj0h_{j}\mapsto 0. Consider the image of DxyD_{xy} in Sym2H\operatorname{\mathrm{Sym}}^{2}H^{*}, which we will denote using [Dxy]=[hxhy][D_{xy}]=[h_{x}^{*}\otimes h_{y}^{*}].

Suppose =τsτ\ell=\tau_{s}\in\tau. We see that

(12) m′′′(zhm,[Dxy])=iτp2λ1λmfimziz+i,jτi<j𝔢([hxhy],(kpλ1λkfjkhk)(kpλ1λkfikhk))zizj+iτ𝔢([hxhy],(kpλ1λkfikhk)(kpλ1λkfikhk))(zi)2=i=1rp2λ1λmfτimwiz+i,jτi<jp2λ1λxλy(fjxfiy+fjyfix)zizj+iτp2λ12λifixfiy(zi)2=p2λ1λmwmz+p2λ1λxλywxwy.\begin{split}m_{\mathcal{F}}^{\prime\prime\prime}&(z_{\ell}^{*}\otimes h_{m}^{*},[D_{xy}])\\ &=\sum_{i\in\tau}p^{2\lambda_{1}-\lambda_{m}}f_{im}z_{i}^{*}z_{\ell}^{*}+\sum_{\begin{subarray}{c}i,j\in\tau\\ i<j\end{subarray}}\mathfrak{e}\left([h_{x}^{*}\otimes h_{y}^{*}],{\left(\textstyle\sum_{k}p^{\lambda_{1}-\lambda_{k}}f_{jk}h_{k}\right)\odot\left(\textstyle\sum_{k}p^{\lambda_{1}-\lambda_{k}}f_{ik}h_{k}\right)}\right)z_{i}^{*}z_{j}^{*}\\ &\quad\quad+\sum_{i\in\tau}\mathfrak{e}\left([h_{x}^{*}\otimes h_{y}^{*}],{\left(\textstyle\sum_{k}p^{\lambda_{1}-\lambda_{k}}f_{ik}h_{k}\right)\otimes\left(\textstyle\sum_{k}p^{\lambda_{1}-\lambda_{k}}f_{ik}h_{k}\right)}\right)(z_{i}^{*})^{2}\\ &=\sum_{i=1}^{r}p^{2\lambda_{1}-\lambda_{m}}f_{\tau_{i}m}w_{i}^{*}z_{\ell}^{*}+\sum_{\begin{subarray}{c}i,j\in\tau\\ i<j\end{subarray}}p^{2\lambda_{1}-\lambda_{x}-\lambda_{y}}(f_{jx}f_{iy}+f_{jy}f_{ix})z_{i}^{*}z_{j}^{*}+\sum_{i\in\tau}p^{2\lambda_{1}-2\lambda_{i}}f_{ix}f_{iy}(z_{i}^{*})^{2}\\ &=p^{2\lambda_{1}-\lambda_{m}}w_{m}^{*}z_{\ell}^{*}+p^{2\lambda_{1}-\lambda_{x}-\lambda_{y}}w_{x}^{*}w_{y}^{*}.\end{split}

For all mm such that msm\geq s, set x=mx=m and y=sy=s. Then

m′′′(zhm,[Dms])=(p2λ1λm+p2λ1λmλs)wmws=p2λ1λmλs(pλs+1)wmws.m_{\mathcal{F}}^{\prime\prime\prime}(z_{\ell}^{*}\otimes h_{m}^{*},[D_{ms}])=(p^{2\lambda_{1}-\lambda_{m}}+p^{2\lambda_{1}-\lambda_{m}-\lambda_{s}})w_{m}^{*}w_{s}^{*}=p^{2\lambda_{1}-\lambda_{m}-\lambda_{s}}(p^{\lambda_{s}}+1)w_{m}^{*}w_{s}^{*}.

Since pλs+1p^{\lambda_{s}}+1 is invertible in RR, this tells us that im(m′′′)\operatorname{\mathrm{im}}(m_{\mathcal{F}}^{\prime\prime\prime}) has a subgroup of size

1smrpλm+λs=p1smrλm+λs=|G|r+1.\prod_{1\leq s\leq m\leq r}p^{\lambda_{m}+\lambda_{s}}=p^{\sum_{1\leq s\leq m\leq r}\lambda_{m}+\lambda_{s}}=|G|^{r+1}.

Therefore, we conclude that #Gn+1=#Gr+1#Gnr|#im(m′′)|#im(m)\#G^{n+1}=\#G^{r+1}\cdot\#G^{n-r}|\#\operatorname{\mathrm{im}}(m_{\mathcal{F}}^{\prime\prime})|\#\operatorname{\mathrm{im}}(m_{\mathcal{F}}^{\prime}). By our prior explanation, this completes the proof of the lemma. ∎

6. Obtaining the moments III: A good bound for codes

We combine the results proved in the previous section to bound

(FM=0andFt(,M)=,G^)=(ΛM=0andΩMtΩt=A)\operatorname{\mathbb{P}}(FM=0\mathrm{\ and\ }F^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}})=\operatorname{\mathbb{P}}(\Lambda\mathcal{F}M=0\mathrm{\ and\ }\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}=A)

when FF is a code and MM is a random matrix. If MM is a random symmetric matrix with integer entries mijm_{ij} and bb is an integer, we say MM is α\alpha-balanced mod bb if for any prime divisor pp of bb and t/pt\in\mathbb{Z}/p\mathbb{Z}, the probability (mijtmodp)1α\operatorname{\mathbb{P}}(m_{ij}\equiv t\ \mathrm{mod\ }p)\leq 1-\alpha.

Lemma 6.1.

Let 0<α<10<\alpha<1, let δ>0\delta>0, and let GG be a finite abelian group. Let aa, HH, and RR be defined as in Section 4. Then there exists some c,K>0c,K>0 with the following significance.

Let MM be a random symmetric n×nn\times n matrix whose entries mij/am_{ij}\in\mathbb{Z}/a\mathbb{Z} for 1ijn1\leq i\leq j\leq n are independent and α\alpha-balanced mod aa. Suppose FHom(V,G)F\in\operatorname{\mathrm{Hom}}(V,G) is a code of distance δn\delta n, and let Hom(V,H)\mathcal{F}\in\operatorname{\mathrm{Hom}}(V,H) be a lift of \mathcal{F} to HH. For all nn, we have that

|(ΛM=0andΩMtΩt=A)|G|(n+1)|Kexp(cn)|G|n.|\operatorname{\mathbb{P}}(\Lambda\mathcal{F}M=0\mathrm{\ and\ }\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}=A)-|G|^{-(n+1)}|\leq\frac{K\exp(-cn)}{|G|^{n}}.

To prove the above, we will need the following result due to Wood.

Lemma 6.2 (Lemma 4.2, [54]).

Suppose ξ\xi is a primitive bbth root of unity, and let zz be a random variable taking values in /b\mathbb{Z}/b\mathbb{Z} such that zz takes each value with probability at most 1α1-\alpha. Then |𝔼(ξz)|exp(α/b2)|\mathbb{E}(\xi^{z})|\leq\exp(-\alpha/b^{2}).

Proof of Lemma 6.1.

Recall that

(ΛM=0andΩMtΩt=A)=1|G|n|Sym2H|CHom(V,G)DSym2H𝔼(ζC(ΛM)+D(ΩMtΩtA)),\operatorname{\mathbb{P}}(\Lambda\mathcal{F}M=0\mathrm{\ and\ }\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}=A)=\frac{1}{|G|^{n}|\operatorname{\mathrm{Sym}}^{2}H|}\sum_{\begin{subarray}{c}C\in\operatorname{\mathrm{Hom}}(V,G^{*})\\ D\in\operatorname{\mathrm{Sym}}^{2}H^{*}\end{subarray}}\mathbb{E}(\zeta^{C(\Lambda\mathcal{F}M)+D(\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}-A)}),

where ζ\zeta is a primitive aath root of unity. Further, by Lemma 5.6, we know that \mathcal{F} is a code of distance δn\delta n, so we may apply our results from the previous section.

To prove the theorem, we break the sum into three pieces (we will choose 0<γ<δ0<\gamma<\delta later):

  1. (1)

    (C,D)(C,D) is special for \mathcal{F};

  2. (2)

    (C,D)(C,D) is not special and weak for \mathcal{F};

  3. (3)

    (C,D)(C,D) is robust for \mathcal{F}.

For a fixed \mathcal{F}, there are |Sym2H|/|G||\operatorname{\mathrm{Sym}}^{2}H|/|G| special pairs (C,D)(C,D) by Lemma 5.9. We may write

𝔼(ζC(ΛM)+D(ΩMtΩtA))=𝔼(ζD(A))𝔼(ζC(ΛM)+D(ΩMtΩt)).\mathbb{E}(\zeta^{C(\Lambda\mathcal{F}M)+D(\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}-A)})=\mathbb{E}(\zeta^{D(-A)})\mathbb{E}(\zeta^{C(\Lambda\mathcal{F}M)+D(\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t})}).

If (C,D)(C,D) is special for \mathcal{F}, recall from Lemma 5.9 that

C(ΛM)+D(ΩMΩtt)=D(A)=0R.C(\Lambda\mathcal{F}M)+D(\Omega\mathcal{F}M\Omega^{t}\mathcal{F}^{t})=D(-A)=0\in R.

Therefore, for (C,D)(C,D) special for \mathcal{F},

𝔼(ζC(ΛM)+D(ΩMtΩtA))=1\mathbb{E}(\zeta^{C(\Lambda\mathcal{F}M)+D(\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}-A)})=1

for any MM.

For the second case, we will rely on the fact that there are not too many weak pairs (C,D)(C,D) in conjunction with our bound from Lemma 5.10. Recall from Lemma 5.2 that for a given \mathcal{F} there are at most

CG(nγn1)|G|γnC_{G}{{n}\choose{\lceil\gamma n\rceil-1}}|G|^{\gamma n}

weak (C,D)Hom(V,G)×Sym2H(C,D)\in\operatorname{\mathrm{Hom}}(V,G^{*})\times\operatorname{\mathrm{Sym}}^{2}H^{*}. Now, factor the expected value as in (7) to get

𝔼(\displaystyle\mathbb{E}( ζC(ΛMB)+D(ΩMtΩtA))=𝔼(ζD(A))1i<jn𝔼(ζEij(C,D,)mij)1in𝔼(ζEii(C,D,)mii).\displaystyle\zeta^{C(\Lambda\mathcal{F}M-B)+D(\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}-A)})=\mathbb{E}(\zeta^{D(-A)})\prod_{1\leq i<j\leq n}\mathbb{E}(\zeta^{E_{ij}(C,D,\mathcal{F})m_{ij}})\prod_{1\leq i\leq n}\mathbb{E}(\zeta^{E_{ii}(C,D,\mathcal{F})m_{ii}}).

Recall that we are assuming (C,D)(C,D) is not special for \mathcal{F}. Lemma 5.10 tells us that at least δn/2\delta n/2 of the Eij(C,D,)E_{ij}(C,D,\mathcal{F}) are nonzero. Hence, if (C,D)(C,D) is not special for \mathcal{F}, we conclude using Lemma 6.2 that

|𝔼(ζC(ΛMB)+D(ΩMtΩtA))|exp(αδn/(2a2)).|\mathbb{E}(\zeta^{C(\Lambda\mathcal{F}M-B)+D(\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}-A)})|\leq\exp(-\alpha\delta n/(2a^{2})).

For the third case, we recall that given \mathcal{F} and robust (C,D)(C,D) for \mathcal{F}, Lemma 5.8 implies that we have at least γδn2/(2|H|2)\gamma\delta n^{2}/(2|H|^{2}) pairs (i,j)(i,j) such that Eij(C,D,)0E_{ij}(C,D,\mathcal{F})\neq 0. Thus, if (C,D)(C,D) is robust for \mathcal{F}, then

|𝔼(ζC(ΛMB)+D(ΩMtΩtA))|exp(αγδn2/(2|H|a2)).|\mathbb{E}(\zeta^{C(\Lambda\mathcal{F}M-B)+D(\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}-A)})|\leq\exp(-\alpha\gamma\delta n^{2}/(2|H|a^{2})).

Putting all three cases together, we get

|\displaystyle\bigg{|} (ΛM=0andΩMtΩt=A)1|G|n|Sym2H|(C,D)special𝔼(ζC(ΛM)+D(ΩMtΩtA))|\displaystyle\operatorname{\mathbb{P}}(\Lambda\mathcal{F}M=0\mathrm{\ and\ }\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}=A)-\frac{1}{|G|^{n}|\operatorname{\mathrm{Sym}}^{2}H|}\sum_{(C,D)\;\mathrm{special}}\mathbb{E}(\zeta^{C(\Lambda\mathcal{F}M)+D(\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}-A)})\bigg{|}
|G|n|Sym2H|(C,D)notspecial|𝔼(ζC(ΛM)+D(ΩMtΩtA))|\displaystyle\leq\frac{|G|^{-n}}{|\operatorname{\mathrm{Sym}}^{2}H|}\sum_{(C,D)\;\mathrm{not\ special}}|\mathbb{E}(\zeta^{C(\Lambda\mathcal{F}M)+D(\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}-A)})|
|G|n|Sym2H|(CG(nγn1)|G|γnexp(αδn/(2a2))+|G|n|Sym2H|exp(αγδn2/(2|H|2a2))).\displaystyle\leq\frac{|G|^{-n}}{|\operatorname{\mathrm{Sym}}^{2}H|}\bigg{(}C_{G}{{n}\choose{\lceil\gamma n\rceil-1}}|G|^{\gamma n}\exp(-\alpha\delta n/(2a^{2}))+|G|^{n}|\operatorname{\mathrm{Sym}}^{2}H|\exp(-\alpha\gamma\delta n^{2}/(2|H|^{2}a^{2}))\bigg{)}.

Let c>0c>0 be such that c<αδ/(2a2)c<\alpha\delta/(2a^{2}). Given δ,α,G,c\delta,\alpha,G,c, we can choose γ\gamma sufficiently small so that

|(ΛM=0\displaystyle\bigg{|}\operatorname{\mathbb{P}}(\Lambda\mathcal{F}M=0 andΩMtΩt=A)|G|n|Sym2H|(C,D)special𝔼(ζC(ΛM)+D(ΩMtΩtA))|\displaystyle\mathrm{\ and\ }\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}=A)-\frac{|G|^{-n}}{|\operatorname{\mathrm{Sym}}^{2}H|}\sum_{(C,D)\;\mathrm{special}}\mathbb{E}(\zeta^{C(\Lambda\mathcal{F}M)+D(\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}-A)})\bigg{|}
1|G|n(CGexp(cn)+exp(log(|G|)nαγδn2/(2|H|2a2))).\displaystyle\leq\frac{1}{|G|^{n}}\left(C_{G}\exp(-cn)+\exp(\log(|G|)n-\alpha\gamma\delta n^{2}/(2|H|^{2}a^{2}))\right).

Now, for fixed α,G,δ,γ,c\alpha,G,\delta,\gamma,c, the inequality

log(|G|)nαγδn2/(2|H|2a2)cn\log(|G|)n-\alpha\gamma\delta n^{2}/(2|H|^{2}a^{2})\leq-cn

holds for nn sufficiently large. Hence, we have

|(ΛM=0andΩMtΩt=A)|G|(n+1)|(CG+1)exp(cn)|G|n\left|\operatorname{\mathbb{P}}(\Lambda\mathcal{F}M=0\mathrm{\ and\ }\Omega\mathcal{F}M\mathcal{F}^{t}\Omega^{t}=A)-|G|^{-(n+1)}\right|\leq\frac{(C_{G}+1)\exp(-cn)}{|G|^{n}}

for nn large enough. For the finitely many nn such that the above is false, just increase the constant KK in the statement of the result. ∎

7. Obtaining the moments IV: The structural properties of the surjections

In this section, we treat the case where FHom(V,G)F\in\operatorname{\mathrm{Hom}}(V,G) is not a code. To do so, we use a division of the FF based on the subgroups of GG from [54]. Let DD be an integer with prime factorization ipiei\prod_{i}p_{i}^{e_{i}}, and let (D)=iei\ell(D)=\sum_{i}e_{i}. The following is an amalgamation of useful ideas and results developed by Wood in [54] that we will apply later on in the next section.

Definition 7.1.

The depth of a homomorphism FHom(V,G)F\in\operatorname{\mathrm{Hom}}(V,G) is the maximal positive DD such that there exists σ[n]\sigma\subset[n] with |σ|<(D)δn|\sigma|<\ell(D)\delta n such that D=[G:FVσ]D=[G:FV_{\sigma}]. If no such DD exists, the depth of FF is 1. (When applying this definition, we always assume δ<(|G|)1\delta<\ell(|G|)^{-1}.) If the depth of FF is 1, then note that FF is a code of distance δn\delta n: if FF has depth 1, then for every σ[n]\sigma\subset[n] with |σ|<δn|\sigma|<\delta n we have FVσ=GFV_{\sigma}=G (otherwise, ([G:FVσ])1\ell([G:FV_{\sigma}])\geq 1).

Note that if the depth of FF is DD, then D|#GD|\#G. The following bounds the number of FF of depth DD.

Lemma 7.2 (Estimate for FF of depth DD; Lemma 5.2 in [54]).

There is a constant KK depending on GG such that if D>1D>1, then the number of FHom(V,G)F\in\operatorname{\mathrm{Hom}}(V,G) of depth DD is at most

K(n(D)δn1)|G|nDn+(D)δn.K{{n}\choose{\lceil\ell(D)\delta n\rceil-1}}|G|^{n}D^{-n+\ell(D)\delta n}.

When we are working with the Laplacian of a random graph, we also need a bound on the number of FF of depth DD, which is given by the following lemma from [54]. If GG is a finite abelian group with exponent bb, then so is G/bG\oplus\mathbb{Z}/b\mathbb{Z}. Thus, we may apply the definition of depth to a homomorphism FHom(V,G/b)F\in\operatorname{\mathrm{Hom}}(V,G\oplus\mathbb{Z}/b\mathbb{Z}).

Lemma 7.3 (Lemma 5.3 of [54]).

Let π2:G/b/b\pi_{2}:G\oplus\mathbb{Z}/b\mathbb{Z}\to\mathbb{Z}/b\mathbb{Z} denote projection onto the second factor. Given a finite abelian pp-group GG of exponent bb, there is a constant KK, which depends on GG, such that if D>1D>1, then the number of FHom(V,G/b)F\in\operatorname{\mathrm{Hom}}(V,G\oplus\mathbb{Z}/b\mathbb{Z}) of depth DD such that π2(Fvi)=1\pi_{2}(Fv_{i})=1 for all i[n]i\in[n] is at most

K(n(D)δn1)|G|n|D|n+(D)δn.K{{n}\choose{\lceil\ell(D)\delta n\rceil}-1}|G|^{n}|D|^{-n+\ell(D)\delta n}.

For each FF of depth DD, we use the following lemma to bound (FM=0andFt(,M)=,G^)\operatorname{\mathbb{P}}(FM=0\mathrm{\ and\ }F^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}}). In particular, because

(FM=0andFt(,M)=,G^)(FM=0)\operatorname{\mathbb{P}}(FM=0\mathrm{\ and\ }F^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}})\leq\operatorname{\mathbb{P}}(FM=0)

and because the FF that are not codes only contribute to our error term, we can simply apply the bounds for (FM=0)\operatorname{\mathbb{P}}(FM=0) due to Wood from [54].

Lemma 7.4.

Let α,δ,G,a\alpha,\delta,G,a be as in Lemma 6.1. Let FHom(V,G)F\in\operatorname{\mathrm{Hom}}(V,G) have depth DD, and suppose [G:FV]<D[G:FV]<D (e.g., this holds if FV=GFV=G). Then there is a real number KK with the following significance. For all MM as in Lemma 6.1 and all nn,

(FM=0)Keα(1(D)δ)n(|G|/D)(1(D)δ)n.\operatorname{\mathbb{P}}(FM=0)\leq Ke^{-\alpha(1-\ell(D)\delta)n}(|G|/D)^{-(1-\ell(D)\delta)n}.

8. Obtaining the moments V: Oh yeah, it’s all coming together

We aggregate the work from the previous four sections to produce a universality result on (and the actual values of) the moments of the cokernels of random matrices with their pairings as well as the moments of sandpile groups of random graphs with their pairings (see the following two theorems, respectively). It will be helpful to recall the definition of an α\alpha-balanced random matrix from Section 6.

Theorem 8.1.

Suppose 0<α<10<\alpha<1, and let GG be a finite abelian group such that G^\widehat{G} is equipped with a symmetric pairing. For cc sufficiently small, there exists some K>0K>0, depending on α,G,c\alpha,G,c, with the following significance. Let MM be a random symmetric n×nn\times n integral matrix, α\alpha-balanced mod |G||G|, whose entries mijm_{ij}\in\mathbb{Z} for iji\leq j are independent. Then

|𝔼(#Sur(cok(M),G))|G|1|Kecn.|\mathbb{E}(\#\operatorname{\mathrm{Sur}}^{*}(\mathrm{cok}(M),G))-|G|^{-1}|\leq Ke^{-cn}.
Proof.

We skip several details in this proof, since the strategy is almost identical to (and, in many cases, simpler than) that of Theorem 8.2. Let bb be the exponent of GG, and let a=b2a=b^{2}. Reduce the matrix MM modulo aa so that our notation agrees with that of previous sections. Recall that

𝔼(#Sur(cok(M),G))=FSur(V,G)(FM=0andFt(,M)=,G^).\mathbb{E}(\#\operatorname{\mathrm{Sur}}^{*}(\mathrm{cok}(M),G))=\sum_{F\in\operatorname{\mathrm{Sur}}(V,G)}\operatorname{\mathbb{P}}(FM=0\mathrm{\ and\ }F^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}}).

By Lemma 7.4 and Lemma 7.2, we have that

FSur(V,G)F~notcodeofdistanceδn(FM=0andFt(,M)=,G^)Kecn.\sum_{\begin{subarray}{c}F\in\operatorname{\mathrm{Sur}}(V,G)\\ \tilde{F}\mathrm{\ not\ code\ of\ distance\ }\delta n\end{subarray}}\operatorname{\mathbb{P}}(FM=0\mathrm{\ and\ }F^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}})\leq Ke^{-cn}.

Lemma 7.2 also implies that

FSur(V,G)F~notcodeofdistanceδn|G|n1Kecn.\sum_{\begin{subarray}{c}F\in\operatorname{\mathrm{Sur}}(V,G)\\ \tilde{F}\mathrm{\ not\ code\ of\ distance\ }\delta n\end{subarray}}|G|^{-n-1}\leq Ke^{-cn}.

We can also show that

FHom(V,G)Sur(V,G)|G|n1K2n.\sum_{F\in\operatorname{\mathrm{Hom}}(V,G)\setminus\operatorname{\mathrm{Sur}}(V,G)}|G|^{-n-1}\leq K2^{-n}.

Finally, applying Lemma 6.1, we see that

FSur(V,G)F~codeofdistanceδn|(FM=0andFt(,M)=,G^)|G|n1|Kecn;\sum_{\begin{subarray}{c}F\in\operatorname{\mathrm{Sur}}(V,G)\\ \tilde{F}\mathrm{\ code\ of\ distance\ }\delta n\end{subarray}}\left|\operatorname{\mathbb{P}}(FM=0\mathrm{\ and\ }F^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}})-|G|^{-n-1}\right|\leq Ke^{-cn};

putting this all together yields the desired result. ∎

Theorem 8.2.

Let 0<q<10<q<1, and let GG be a finite abelian group equipped with a symmetric pairing on its dual G^=Hom(G,/)\widehat{G}=\operatorname{\mathrm{Hom}}(G,\mathbb{Q}/\mathbb{Z}). Then there exist positive constants c,Kc,K (depending on GG) with the following significance. Suppose ΓG(n,q)\Gamma\in G(n,q) is a random graph and SS its sandpile group. Let LL be the Laplacian of Γ\Gamma, and let ,S^\langle\,,\rangle_{\widehat{S}} denote the canonical symmetric pairing on S^\widehat{S}. Then for all nn we have that

|𝔼(#Sur(S,G))|G|1|Kecn.|\mathbb{E}(\#\operatorname{\mathrm{Sur}}^{*}(S,G))-|G|^{-1}|\leq Ke^{-cn}.
Proof.

Suppose bb is the exponent of GG, and let =/b\mathfrak{R}=\mathbb{Z}/b\mathbb{Z}. Let a=b2a=b^{2}, and let R=/aR=\mathbb{Z}/a\mathbb{Z}. Let SR=SRS_{R}=S\otimes R and S=SS_{\mathfrak{R}}=S\otimes\mathfrak{R}, and recall that #Sur(S,G)=#Sur(S,G)=#Sur(SR,G)\#\operatorname{\mathrm{Sur}}(S,G)=\#\operatorname{\mathrm{Sur}}(S\otimes\mathfrak{R},G)=\#\operatorname{\mathrm{Sur}}(S\otimes R,G). Likewise, we let LRL_{R} and LL_{\mathfrak{R}} be the reductions of the Laplacian of Γ\Gamma modulo aa and bb, respectively; we may think of LRL_{R} and LL_{\mathfrak{R}} as n×nn\times n matrices with coefficients in RR and \mathfrak{R}, respectively. Recall from Section 3 that cok^(L)\mathrm{c}\widehat{\mathrm{ok}}(L) carries a symmetric pairing given by LL, which we denote by ,L\langle\,,\rangle_{L}. Likewise cok^(LR)\mathrm{c}\widehat{\mathrm{ok}}(L_{R}) and cok^(L)\mathrm{c}\widehat{\mathrm{ok}}(L_{\mathfrak{R}}) are both equipped with symmetric pairings given by restricting the pairing on cok^(L)\mathrm{c}\widehat{\mathrm{ok}}(L) to cok^(LR)\mathrm{c}\widehat{\mathrm{ok}}(L_{R}) and cok^(L)\mathrm{c}\widehat{\mathrm{ok}}(L_{\mathfrak{R}}), respectively.

Let MM be a random n×nn\times n symmetric matrix with coefficients in RR. Suppose the mijm_{ij} modulo bb are distributed as (L)ij(L_{\mathfrak{R}})_{ij} for i<ji<j, and suppose the miim_{ii}’s are distributed uniformly in \mathfrak{R}, with all mijm_{ij} (i<ji<j) and miim_{ii} independent. Set W=nW=\mathfrak{R}^{n}. Recall that we may view LL_{\mathfrak{R}} as a linear map WWW^{*}\to W, and let w1,,wnw_{1},\ldots,w_{n} be our distinguished basis of WW (i.e., the basis of WW such that the strict upper-triangular entries of LL_{\mathfrak{R}} with respect to the bases wiw_{i} and wiw_{i}^{*} of WW and WW^{*} are independent). Let F0Hom(Wn,)F_{0}\in\operatorname{\mathrm{Hom}}(W^{n},\mathfrak{R}) be the map that sends each basis vector wi1w_{i}\mapsto 1 (we see that F0MF_{0}M is the row vector whose entries are the column sums of MM). Note that MM and LL_{\mathfrak{R}} do not have the same distribution, since the column sums of MM can be anything and the column sums of LL_{\mathfrak{R}} are zero. To fix this, we’ll condition on F0M=0F_{0}M=0 so that the conditioned distribution of MM is the same as the distribution of LL_{\mathfrak{R}}. Let bijb_{ij}\in\mathfrak{R} for each 1i<jn1\leq i<j\leq n. We see that

(F0M=0|mij=bij for all 1i<jn)=bn,\operatorname{\mathbb{P}}(F_{0}M=0\;|\;m_{ij}=b_{ij}\textrm{ for all }1\leq i<j\leq n)=b^{-n},

so (F0M=0)=bn\operatorname{\mathbb{P}}(F_{0}M=0)=b^{-n}. Hence, given F0M=0F_{0}M=0, we see that any choice of off-diagonal entries is equally likely in LL_{\mathfrak{R}} as in MM.

For FHom(W,G)F\in\operatorname{\mathrm{Hom}}(W,G), we have that

(FL=0andFt(,L)=,G^)=(FM=0andFt(,M)=,G^|F0M=0),\operatorname{\mathbb{P}}(FL_{\mathfrak{R}}=0\mathrm{\ and\ }F^{t}(\langle\,,\rangle_{L_{\mathfrak{R}}})=\langle\,,\rangle_{\widehat{G}})=\operatorname{\mathbb{P}}(FM=0\mathrm{\ and\ }F^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}}\;|\;F_{0}M=0),

where ,L\langle\,,\rangle_{L_{\mathfrak{R}}} and ,M\langle\,,\rangle_{M} denote the symmetric pairings on cok^(L)\mathrm{c}\widehat{\mathrm{ok}}(L_{\mathfrak{R}}) and cok^(M)\mathrm{c}\widehat{\mathrm{ok}}(M), respectively. Also,

(FM=0\displaystyle\operatorname{\mathbb{P}}(FM=0 andFt(,M)=,G^|F0M=0)(F0M=0)\displaystyle\mathrm{\ and\ }F^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}}\;|\;F_{0}M=0)\operatorname{\mathbb{P}}(F_{0}M=0)
=(FM=0andFt(,M)=,G^andF0M=0).\displaystyle=\operatorname{\mathbb{P}}(FM=0\mathrm{\ and\ }F^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}}\ \mathrm{and\ }F_{0}M=0).

Let F~Hom(W,G)\tilde{F}\in\operatorname{\mathrm{Hom}}(W,G\oplus\mathfrak{R}) be the sum of FF and F0F_{0}.

Let UWU\subset W denote the vectors whose coordinates sum to 0. In other words, let U={wW|F0w=0}U=\{w\in W\;|\;F_{0}w=0\}. Denote by SurU(W,G)\operatorname{\mathrm{Sur}}_{U}(W,G) the maps from WGW\to G that are surjections when restricted to UU. We would like to estimate

𝔼(#Sur(S,G))=𝔼(Sur(U/col(L)),G)\displaystyle\mathbb{E}(\#\operatorname{\mathrm{Sur}}^{*}(S_{\mathfrak{R}},G))=\mathbb{E}(\operatorname{\mathrm{Sur}}^{*}(U/\operatorname{\mathrm{col}}(L_{\mathfrak{R}})),G) =FSur(U,G)(FL=0andFt(,S^)=,G^).\displaystyle=\sum_{F\in\operatorname{\mathrm{Sur}}(U,G)}\operatorname{\mathbb{P}}(FL_{\mathfrak{R}}=0\ \mathrm{and\ }F^{t}(\langle\,,\rangle_{\widehat{S_{\mathfrak{R}}}})=\langle\,,\rangle_{\widehat{G}}).

Now, we have

𝔼(#Sur(S,G))\displaystyle\mathbb{E}(\#\operatorname{\mathrm{Sur}}^{*}(S_{\mathfrak{R}},G)) =FSur(U,G)(FL=0andFt(,S^)=,G^)\displaystyle=\sum_{F\in\operatorname{\mathrm{Sur}}(U,G)}\operatorname{\mathbb{P}}(FL_{\mathfrak{R}}=0\ \mathrm{and\ }F^{t}(\langle\,,\rangle_{\widehat{S_{\mathfrak{R}}}})=\langle\,,\rangle_{\widehat{G}})
=1|G|FSurU(W,G)(FL=0andFt(,S^)=,G^).\displaystyle=\frac{1}{|G|}\sum_{F\in\operatorname{\mathrm{Sur}}_{U}(W,G)}\operatorname{\mathbb{P}}(FL_{\mathfrak{R}}=0\mathrm{\ and\ }F^{t}(\langle\,,\rangle_{\widehat{S_{\mathfrak{R}}}})=\langle\,,\rangle_{\widehat{G}}).

Given FSurU(W,G)F\in\operatorname{\mathrm{Sur}}_{U}(W,G) such that FL=0FL_{\mathfrak{R}}=0, recall from Lemma 3.6 that Ft(,S^)=,G^F^{t}(\langle\,,\rangle_{\widehat{S_{\mathfrak{R}}}})=\langle\,,\rangle_{\widehat{G}} if and only if Ft(,L)=,G^F^{t}(\langle\,,\rangle_{L_{\mathfrak{R}}})=\langle\,,\rangle_{\widehat{G}}. In other words, for any such FF, we have that

(FL=0andFt(,S^)=,G^)=(FL=0andFt(,L)=,G^),\operatorname{\mathbb{P}}(FL_{\mathfrak{R}}=0\mathrm{\ and\ }F^{t}(\langle\,,\rangle_{\widehat{S_{\mathfrak{R}}}})=\langle\,,\rangle_{\widehat{G}})=\operatorname{\mathbb{P}}(FL_{\mathfrak{R}}=0\mathrm{\ and\ }F^{t}(\langle\,,\rangle_{L_{\mathfrak{R}}})=\langle\,,\rangle_{\widehat{G}}),

and we can check whether the symmetric pairing on S^\widehat{S_{\mathfrak{R}}} pushes forward to ,G^\langle\,,\rangle_{\widehat{G}} by checking whether the symmetric pairing on cok^(L)\mathrm{c}\widehat{\mathrm{ok}}(L_{\mathfrak{R}}) pushes forward to ,G^\langle\,,\rangle_{\widehat{G}}.

Now, let V=RnV=R^{n}, and let ZVZ\subset V denote the submodule of vectors whose coordinates sum to 0 modulo bb, i.e., Z={vV|F0v=0modb}Z=\{v\in V\;|\;F_{0}v=0\mathrm{\ mod\ }b\}, where we abuse notation and define F0:VF_{0}:V\to\mathfrak{R} to be the map taking each vi1v_{i}\mapsto 1 modulo bb. As before, set SurZ(V,G)\operatorname{\mathrm{Sur}}_{Z}(V,G) to be the set of maps VGV\to G that are surjections when restricted to ZZ. Note that because Sur(W,G)=Sur(V,G)\operatorname{\mathrm{Sur}}(W,G)=\operatorname{\mathrm{Sur}}(V,G), it follows that Sur(U,G)=Sur(Z,G)\operatorname{\mathrm{Sur}}(U,G)=\operatorname{\mathrm{Sur}}(Z,G). Therefore,

𝔼(#Sur(S,G))\displaystyle\mathbb{E}(\#\operatorname{\mathrm{Sur}}^{*}(S_{\mathfrak{R}},G)) =1|G|FSurU(W,G)(FL=0andFt(,L)=,G^)\displaystyle=\frac{1}{|G|}\sum_{F\in\operatorname{\mathrm{Sur}}_{U}(W,G)}\operatorname{\mathbb{P}}(FL_{\mathfrak{R}}=0\mathrm{\ and\ }F^{t}(\langle\,,\rangle_{L_{\mathfrak{R}}})=\langle\,,\rangle_{\widehat{G}})
=|G|1bnFSurU(W,G)(F~M=0andFt(,M)=,G^)\displaystyle=|G|^{-1}b^{n}\sum_{F\in\operatorname{\mathrm{Sur}}_{U}(W,G)}\operatorname{\mathbb{P}}(\tilde{F}M=0\mathrm{\ and\ }F^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}})
=|G|1bnFSurZ(V,G)(F~M=0andFt(,M)=,G^).\displaystyle=|G|^{-1}b^{n}\sum_{F\in\operatorname{\mathrm{Sur}}_{Z}(V,G)}\operatorname{\mathbb{P}}(\tilde{F}M=0\mathrm{\ and\ }F^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}}).

Note that if F:VGF:V\to G is a surjection when restricted to ZZ, then F~\tilde{F} is a surjection from VV to GG\oplus\mathfrak{R}. Write the above as

𝔼(#Sur(S,G))=|G|1bnFSurZ(V,G)extensions,G^^of,G^toG^^(F~M=0andF~t(,M)=,G^^),\mathbb{E}(\#\operatorname{\mathrm{Sur}}^{*}(S_{\mathfrak{R}},G))=|G|^{-1}b^{n}\sum_{F\in\operatorname{\mathrm{Sur}}_{Z}(V,G)}\sum_{\begin{subarray}{c}\mathrm{extensions}\ \langle\,,\rangle_{\widehat{G}\oplus\widehat{\mathfrak{R}}}\mathrm{\ of}\\ \langle\,,\rangle_{\widehat{G}}\mathrm{\ to\ }\widehat{G}\oplus\widehat{\mathfrak{R}}\end{subarray}}\operatorname{\mathbb{P}}(\tilde{F}M=0\mathrm{\ and\ }\tilde{F}^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}\oplus\widehat{\mathfrak{R}}}),

where by extensions ,G^^\langle\,,\rangle_{\widehat{G}\oplus\widehat{\mathfrak{R}}} of ,G^\langle\,,\rangle_{\widehat{G}} we really mean symmetric extensions of ,G^\langle\,,\rangle_{\widehat{G}} (we truncate for the sake of brevity). For the rest of the argument, when we refer to extensions of ,G^\langle\,,\rangle_{\widehat{G}} we mean symmetric extensions of this pairing. If g1,,grg_{1},\ldots,g_{r} generate GG, and if tt generates \mathfrak{R}, then we see that symmetric extensions of ,G^\langle\,,\rangle_{\widehat{G}} to G^^\widehat{G}\oplus\widehat{\mathfrak{R}} are entirely determined by where the pairs (gi^,t^)(\widehat{g_{i}},\widehat{t}) and (t^,t^)(\widehat{t},\widehat{t}) are sent for all 1ir1\leq i\leq r. Hence, in total, there are b|G|b|G| possible extensions of ,G^\langle\,,\rangle_{\widehat{G}} to G^^\widehat{G}\oplus\widehat{\mathfrak{R}}.

We start by considering the part of the sum where F~\tilde{F} is not a code of distance δn\delta n for some δ>0\delta>0 that we will choose later. We allow KK to change in each line, as long as it is a constant depending only on q,G,δq,G,\delta. Let α=max(q,1q)\alpha=\max(q,1-q). Then

bn|G|\displaystyle\frac{b^{n}}{|G|} FSurZ(V,G)F~notcodeofdistanceδnextensions,G^^of,G^toG^^(F~M=0andF~t(,M)=,G^^)\displaystyle\sum_{\begin{subarray}{c}F\in\operatorname{\mathrm{Sur}}_{Z}(V,G)\\ \tilde{F}\mathrm{\ not\ code\ of\ distance\ }\delta n\end{subarray}}\sum_{\begin{subarray}{c}\mathrm{extensions}\ \langle\,,\rangle_{\widehat{G}\oplus\widehat{\mathfrak{R}}}\mathrm{\ of}\\ \langle\,,\rangle_{\widehat{G}}\mathrm{\ to\ }\widehat{G}\oplus\widehat{\mathfrak{R}}\end{subarray}}\operatorname{\mathbb{P}}(\tilde{F}M=0\mathrm{\ and\ }\tilde{F}^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}\oplus\widehat{\mathfrak{R}}})
bn|G|D>1D|#GFSurZ(V,G)F~depthDextensions,G^^of,G^toG^^(F~M=0andF~t(,M)=,G^^)\displaystyle\leq\frac{b^{n}}{|G|}\sum_{\begin{subarray}{c}D>1\\ D|\#G\end{subarray}}\sum_{\begin{subarray}{c}F\in\operatorname{\mathrm{Sur}}_{Z}(V,G)\\ \tilde{F}\ \mathrm{depth\ }D\end{subarray}}\sum_{\begin{subarray}{c}\mathrm{extensions}\ \langle\,,\rangle_{\widehat{G}\oplus\widehat{\mathfrak{R}}}\mathrm{\ of}\\ \langle\,,\rangle_{\widehat{G}}\mathrm{\ to\ }\widehat{G}\oplus\widehat{\mathfrak{R}}\end{subarray}}\operatorname{\mathbb{P}}(\tilde{F}M=0\mathrm{\ and\ }\tilde{F}^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}\oplus\widehat{\mathfrak{R}}})
bn|G|D>1D|#GFSurZ(V,G)F~depthDb|G|(F~M=0)\displaystyle\leq\frac{b^{n}}{|G|}\sum_{\begin{subarray}{c}D>1\\ D|\#G\end{subarray}}\sum_{\begin{subarray}{c}F\in\operatorname{\mathrm{Sur}}_{Z}(V,G)\\ \tilde{F}\ \mathrm{depth\ }D\end{subarray}}b|G|\operatorname{\mathbb{P}}(\tilde{F}M=0)
bn+1D>1D|#G#{F~Hom(V,G)depthD|π2(vi)=1foralli}\displaystyle\leq b^{n+1}\sum_{\begin{subarray}{c}D>1\\ D|\#G\end{subarray}}\#\{\tilde{F}\in\operatorname{\mathrm{Hom}}(V,G\oplus\mathfrak{R})\ \mathrm{depth}\ D\;|\;\pi_{2}(v_{i})=1\ \mathrm{for\ all\ }i\}
×Keα(1(D)δ)n(b|G|/D)(1(D)δ)n\displaystyle{}\quad\quad\quad\quad\quad\quad\times Ke^{-\alpha(1-\ell(D)\delta)n}(b|G|/D)^{-(1-\ell(D)\delta)n}
bn+1D>1D|#GK(n(D)δn1)|G|nDn+(D)δn×eα(1(D)δ)n(b|G|/D)(1(D)δ)n\displaystyle{}\leq b^{n+1}\sum_{\begin{subarray}{c}D>1\\ D|\#G\end{subarray}}K{{n}\choose{\lceil\ell(D)\delta n\rceil-1}}|G|^{n}D^{-n+\ell(D)\delta n}\times e^{-\alpha(1-\ell(D)\delta)n}(b|G|/D)^{-(1-\ell(D)\delta)n}
K(n(|G|)δn1)eα(1(|G|)δ)n(b|G|)δ(|G|)n\displaystyle{}\leq K{{n}\choose{\lceil\ell(|G|)\delta n\rceil-1}}e^{-\alpha(1-\ell(|G|)\delta)n}(b|G|)^{\delta\ell(|G|)n}
Kecn,\displaystyle{}\leq Ke^{-cn},

where the third and fourth inequalities in the above follow from Lemmas 7.3 and 7.4. The last inequality in the above follows from the fact that for any 0<c<α0<c<\alpha, we can choose δ\delta so small that

(n(|G|)δn1)eα(1(|G|)δ)n(b|G|)δ(|G|)necn.{{n}\choose{\lceil\ell(|G|)\delta n\rceil-1}}e^{-\alpha(1-\ell(|G|)\delta)n}(b|G|)^{\delta\ell(|G|)n}\leq e^{-cn}.

We also have that

FSurZ(V,G)F~notcodeofdistanceδn\displaystyle\sum_{\begin{subarray}{c}F\in\operatorname{\mathrm{Sur}}_{Z}(V,G)\\ \tilde{F}\mathrm{\ not\ code\ of\ distance\ }\delta n\end{subarray}} |G|n1D>1D|#GFSurZ(V,G)F~depthD|G|n1\displaystyle|G|^{-n-1}\leq\sum_{\begin{subarray}{c}D>1\\ D|\#G\end{subarray}}\sum_{\begin{subarray}{c}F\in\operatorname{\mathrm{Sur}}_{Z}(V,G)\\ \tilde{F}\ \mathrm{depth\ }D\end{subarray}}|G|^{-n-1}
D>1D|#GK(n(D)δn1)|G|n|D|n+(D)δn|G|n1\displaystyle\leq\sum_{\begin{subarray}{c}D>1\\ D|\#G\end{subarray}}K{{n}\choose{\lceil\ell(D)\delta n\rceil-1}}|G|^{n}|D|^{-n+\ell(D)\delta n}|G|^{-n-1}
K(n(|G|)δn1)2n+(|G|)δn\displaystyle\leq K{{n}\choose{\lceil\ell(|G|)\delta n\rceil-1}}2^{-n+\ell(|G|)\delta n}
Kecn,\displaystyle\leq Ke^{-cn},

where the last inequality holds because, for any 0<c<log(2)0<c<\log(2), we may choose δ\delta sufficiently small so that

(n(|G|)δn1)2n+(|G|)δnecn.{{n}\choose{\lceil\ell(|G|)\delta n\rceil-1}}2^{-n+\ell(|G|)\delta n}\leq e^{-cn}.

Similarly, we also have that

FHom(V,G)SurZ(V,G)|G|n1\displaystyle\sum_{F\in\operatorname{\mathrm{Hom}}(V,G)\setminus\operatorname{\mathrm{Sur}}_{Z}(V,G)}|G|^{-n-1} <GFSur(Z,)|G|n<G||n1|G|nK2n,\displaystyle\leq\sum_{\mathcal{H}<G}\sum_{F\in\operatorname{\mathrm{Sur}}(Z,\mathcal{H})}|G|^{-n}\leq\sum_{\mathcal{H}<G}|\mathcal{H}|^{n-1}|G|^{-n}\leq K2^{-n},

where <G\mathcal{H}<G indicates that \mathcal{H} is a proper subgroup of GG.

Then, using Lemma 6.1 and (6), we see that

FSurZ(V,G)F~codeofdistanceδnextensions,G^^of,G^toG^^|(F~M=0andF~t(,M)=,G^^)(b|G|)n1|\displaystyle\sum_{\begin{subarray}{c}F\in\operatorname{\mathrm{Sur}}_{Z}(V,G)\\ \tilde{F}\mathrm{\ code\ of\ distance\ }\delta n\end{subarray}}\sum_{\begin{subarray}{c}\mathrm{extensions}\ \langle\,,\rangle_{\widehat{G}\oplus\widehat{\mathfrak{R}}}\mathrm{\ of}\\ \langle\,,\rangle_{\widehat{G}}\mathrm{\ to\ }\widehat{G}\oplus\widehat{\mathfrak{R}}\end{subarray}}\left|\operatorname{\mathbb{P}}(\tilde{F}M=0\mathrm{\ and\ }\tilde{F}^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}\oplus\widehat{\mathfrak{R}}})-(b|G|)^{-n-1}\right|
FSurZ(V,G)F~codeofdistanceδnextensions,G^^of,G^toG^^Kecn(b|G|)n\displaystyle\qquad\leq\sum_{\begin{subarray}{c}F\in\operatorname{\mathrm{Sur}}_{Z}(V,G)\\ \tilde{F}\mathrm{\ code\ of\ distance\ }\delta n\end{subarray}}\sum_{\begin{subarray}{c}\mathrm{extensions}\ \langle\,,\rangle_{\widehat{G}\oplus\widehat{\mathfrak{R}}}\mathrm{\ of}\\ \langle\,,\rangle_{\widehat{G}}\mathrm{\ to\ }\widehat{G}\oplus\widehat{\mathfrak{R}}\end{subarray}}Ke^{-cn}(b|G|)^{-n}
Kecnbn.\displaystyle\qquad\leq Ke^{-cn}b^{-n}.

Recall that the number of symmetric extensions of ,G^\langle\,,\rangle_{\widehat{G}} to G^^\widehat{G}\oplus\widehat{\mathfrak{R}} is b|G|b|G|. Putting this all together, we have

|\displaystyle\Bigg{|} bn|G|(FSurZ(V,G)extensions,G^^of,G^toG^^(F~M=0andF~t(,M)=,G^^))|G|1|\displaystyle\frac{b^{n}}{|G|}\Bigg{(}\sum_{F\in\operatorname{\mathrm{Sur}}_{Z}(V,G)}\sum_{\begin{subarray}{c}\mathrm{extensions}\ \langle\,,\rangle_{\widehat{G}\oplus\widehat{\mathfrak{R}}}\mathrm{\ of}\\ \langle\,,\rangle_{\widehat{G}}\mathrm{\ to\ }\widehat{G}\oplus\widehat{\mathfrak{R}}\end{subarray}}\operatorname{\mathbb{P}}(\tilde{F}M=0\mathrm{\ and\ }\tilde{F}^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}\oplus\widehat{\mathfrak{R}}})\Bigg{)}-|G|^{-1}\Bigg{|}
|bn|G|FSurZ(V,G)F~notcodeofdistanceδnextensions,G^^of,G^toG^^(F~M=0andF~t(,M)=,G^^)|\displaystyle\leq\left|\frac{b^{n}}{|G|}\sum_{\begin{subarray}{c}F\in\operatorname{\mathrm{Sur}}_{Z}(V,G)\\ \tilde{F}\mathrm{\ not\ code\ of\ distance\ }\delta n\end{subarray}}\sum_{\begin{subarray}{c}\mathrm{extensions}\ \langle\,,\rangle_{\widehat{G}\oplus\widehat{\mathfrak{R}}}\mathrm{\ of}\\ \langle\,,\rangle_{\widehat{G}}\mathrm{\ to\ }\widehat{G}\oplus\widehat{\mathfrak{R}}\end{subarray}}\operatorname{\mathbb{P}}(\tilde{F}M=0\mathrm{\ and\ }\tilde{F}^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}\oplus\widehat{\mathfrak{R}}})\right|
+bn|G|FSurZ(V,G)F~codeofdistanceδnextensions,G^^of,G^toG^^|(F~M=0andF~t(,M)=,G^^)(b|G|)n1|\displaystyle\qquad+\frac{b^{n}}{|G|}\sum_{\begin{subarray}{c}F\in\operatorname{\mathrm{Sur}}_{Z}(V,G)\\ \tilde{F}\mathrm{\ code\ of\ distance\ }\delta n\end{subarray}}\sum_{\begin{subarray}{c}\mathrm{extensions}\ \langle\,,\rangle_{\widehat{G}\oplus\widehat{\mathfrak{R}}}\mathrm{\ of}\\ \langle\,,\rangle_{\widehat{G}}\mathrm{\ to\ }\widehat{G}\oplus\widehat{\mathfrak{R}}\end{subarray}}\left|\operatorname{\mathbb{P}}(\tilde{F}M=0\mathrm{\ and\ }\tilde{F}^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}\oplus\widehat{\mathfrak{R}}})-(b|G|)^{-n-1}\right|
+||G|1+bn|G|FSurZ(V,G)F~codeofdistanceδnextensions,G^^of,G^toG^^(b|G|)n1|\displaystyle\qquad\qquad+\left|-|G|^{-1}+\frac{b^{n}}{|G|}\sum_{\begin{subarray}{c}F\in\operatorname{\mathrm{Sur}}_{Z}(V,G)\\ \tilde{F}\mathrm{\ code\ of\ distance\ }\delta n\end{subarray}}\sum_{\begin{subarray}{c}\mathrm{extensions}\ \langle\,,\rangle_{\widehat{G}\oplus\widehat{\mathfrak{R}}}\mathrm{\ of}\\ \langle\,,\rangle_{\widehat{G}}\mathrm{\ to\ }\widehat{G}\oplus\widehat{\mathfrak{R}}\end{subarray}}(b|G|)^{-n-1}\right|
Kecn+FSurZ(V,G)F~notcodeofdistanceδn|G|n1+FHom(V,G)SurZ(V,G)|G|n1\displaystyle\leq Ke^{-cn}+\sum_{\begin{subarray}{c}F\in\operatorname{\mathrm{Sur}}_{Z}(V,G)\\ \tilde{F}\mathrm{\ not\ code\ of\ distance\ }\delta n\end{subarray}}|G|^{-n-1}+\sum_{F\in\operatorname{\mathrm{Hom}}(V,G)\setminus\operatorname{\mathrm{Sur}}_{Z}(V,G)}|G|^{-n-1}
Kecn.\displaystyle\leq Ke^{-cn}.

Recalling that

𝔼(#Sur(S,G))=bn|G|FSurZ(V,G)extensions,G^^of,G^toG^^(F~M=0andF~t(,M)=,G^^),\mathbb{E}(\#\operatorname{\mathrm{Sur}}^{*}(S_{\mathfrak{R}},G))=\frac{b^{n}}{|G|}\sum_{F\in\operatorname{\mathrm{Sur}}_{Z}(V,G)}\sum_{\begin{subarray}{c}\mathrm{extensions}\ \langle\,,\rangle_{\widehat{G}\oplus\widehat{\mathfrak{R}}}\\ \mathrm{of\ }\langle\,,\rangle_{\widehat{G}}\mathrm{\ to\ \widehat{G}\oplus\widehat{\mathfrak{R}}}\end{subarray}}\operatorname{\mathbb{P}}(\tilde{F}M=0\mathrm{\ and\ }\tilde{F}^{t}(\langle\,,\rangle_{M})=\langle\,,\rangle_{\widehat{G}\oplus\widehat{\mathfrak{R}}}),

we may conclude the theorem. ∎

9. The moments determine the distribution

In the following, we show that the moments we obtained in the previous section indeed determine the distributions of our random variables taking values in the category 𝒞\mathcal{C} of finite abelian groups whose duals are equipped with symmetric pairings. To do so, we combine several results from [50] due to Sawin and Wood.

Recall from Section 1 that 𝒞\mathcal{C} is the category of finite abelian groups whose dual groups are equipped with symmetric pairings. For two objects (A,,A^),(B,,B^)𝒞(A,\langle\,,\rangle_{\widehat{A}}),(B,\langle\,,\rangle_{\widehat{B}})\in\mathcal{C}, recall that a morphism between these objects in 𝒞\mathcal{C} is a group homomorphism F:ABF:A\to B such that Ft(,A^)=,B^F^{t}(\langle\,,\rangle_{\widehat{A}})=\langle\,,\rangle_{\widehat{B}}. To apply the moment methods of Sawin and Wood, we would like 𝒞\mathcal{C} to be a diamond category.

Before we give the definition of a diamond category, we first collect some notation from [50]. Suppose 𝒟\mathcal{D} is a small category. Given an object G𝒟G\in\mathcal{D}, a quotient of GG is the data of a object F𝒟F\in\mathcal{D} and an epimorphism π:GF\pi:G\to F. Two quotients (F,π)(F,\pi) and (F,π)(F^{\prime},\pi^{\prime}) of GG are said to be isomorphic if there exists an isomorphism ρ:FF\rho:F\to F^{\prime} such that π=ρπ\pi^{\prime}=\rho\circ\pi. We may define a partial order on the set of isomorphism classes quotients of an object G𝒟G\in\mathcal{D}, where FHF\leq H if there is some h:HFh:H\to F compatible with the epimorphisms from GG. Given a quotient FF of G𝒟G\in\mathcal{D}, let [F,G][F,G] be the partially ordered set of (isomorphism classes of) quotients of GG that are greater than or equal to FF. Call an object minimal if its only quotient is itself.

In our category of interest, we would like the set of quotients of an object with respect to this partial order to form a lattice. A lattice is a poset such that any two elements xx and yy in the lattice have a least upper bound, or join, and a greatest lower bound, or meet. Denote the join of xx and yy (respectively, meet) by xyx\vee y (respectively, xyx\wedge y). We say a lattice is modular if aba\leq b implies a(xb)=(ax)ba\vee(x\wedge b)=(a\vee x)\wedge b for all xx in the lattice.

For a set XX of objects of 𝒟\mathcal{D}, stable under isomorphism, XX is said to be downward-closed if every quotient of an object GXG\in X is also in XX. We say that XX is join-closed if for every GXG\in X, given two quotients FF and HH of GG such that FHF\vee H exists (in the poset of isomorphism classes of quotients of GG), we have F,HXF,H\in X implies FHXF\vee H\in X. Given a set XX of isomorphism classes of objects in 𝒟\mathcal{D}, the set generated by XX is the smallest downward-closed and join-closed subset of 𝒟\mathcal{D} containing XX and every minimal object of 𝒟\mathcal{D}. A level DD of 𝒟\mathcal{D} is a set of isomorphism classes of objects of 𝒟\mathcal{D} generated by a finite set. An epimorphism GFG\to F is called simple if the interval [F,G][F,G] contains only FF and GG.

Definition 9.1.

A diamond category is a small category 𝒟\mathcal{D} satisfying the following three properties.

  1. (1)

    For each G𝒟G\in\mathcal{D}, the set of isomorphism classes of quotients of GG form a finite modular lattice.

  2. (2)

    Each object in 𝒟\mathcal{D} has finitely many automorphisms.

  3. (3)

    For each level DD of 𝒟\mathcal{D} and each object in the level DD, there are only finitely many elements of DD with a simple epimorphism to GG.

To see that 𝒞\mathcal{C} is a diamond category, we will apply the following lemma from [50].

Lemma 9.2 (Lemma 6.22 of [50]).

Let 𝒟\mathcal{D} be a diamond category, and let 𝒢\mathcal{G} be a functor from 𝒟\mathcal{D} to the category of finite sets. Then the category (𝒟,𝒢)(\mathcal{D},\mathcal{G}) of pairs of an object G𝒟G\in\mathcal{D} together with an element s𝒢(G)s\in\mathcal{G}(G), with morphisms (G,sG)(H,sH)(G,s_{G})\to(H,s_{H}) given by morphisms f:GHf:G\to H in 𝒟\mathcal{D} such that 𝒢(f)(sG)=sH\mathcal{G}(f)(s_{G})=s_{H}, is a diamond category.

That 𝒞\mathcal{C} is a diamond category follows from applying Lemma 9.2 to the category of finite abelian groups and the functor Φ\Phi taking each group to the set of symmetric pairings on its dual group. Let FinAb\mathrm{FinAb} denote the category of finite abelian groups. The functor Φ:FinAbFinSet\Phi:\mathrm{FinAb}\to\mathrm{FinSet} takes a group homomorphism F:ABF:A\to B to the map of sets

Ft:{symmetricpairingsonA^}{symmetricpairingsonB^}F^{t}:\{\mathrm{symmetric\ pairings\ on\ }\widehat{A}\}\to\{\mathrm{symmetric\ pairings\ on\ }\widehat{B}\}

given by

Ft(,A^)=Ft(),Ft()A^:B^B^/.F^{t}(\langle\,,\rangle_{\widehat{A}})=\langle F^{t}(-),F^{t}(-)\rangle_{\widehat{A}}:\widehat{B}\otimes\widehat{B}\to\mathbb{Q}/\mathbb{Z}.

We note that the category 𝒞\mathcal{C} is precisely the enriched category of pairs (FinAb,Φ)(\mathrm{FinAb},\Phi) as in Lemma 9.2.

The following is the key result from [50] that will allow us to prove the moments determine the distribution.

Lemma 9.3 (Theorem 1.6 of [50]).

Let 𝒟\mathcal{D} be a diamond category, and let DD be a level in 𝒟\mathcal{D}. For each GDG\in D, let MGM_{G} be a real number such that (MG)G(M_{G})_{G} is well-behaved at DD.444The notion of well-behavedness at a level is not something we will precisely define. Roughly, well-behavedness captures whether or not the moments “grow too quickly.” In other words, if the moments in a diamond category are well-behaved at a level, then they will determine a unique measure on that level. The version of Lemma 9.3 stated in [50] is actually much stronger than the one stated above. However, for the sake of brevity, we use a weaker version of the theorem that is sufficient for our purposes. Then if μDt\mu_{D}^{t} and νDt\nu_{D}^{t} are two sequences of measures on DD such that for each GDG\in D, we have

limtFD|Epi(F,G)|dμDt=MG=limtFD|Epi(F,G)|dνDt,\lim_{t\to\infty}\int_{F\in D}|\mathrm{Epi}(F,G)|\mathrm{d}\mu_{D}^{t}=M_{G}=\lim_{t\to\infty}\int_{F\in D}|\mathrm{Epi}(F,G)|\mathrm{d}\nu_{D}^{t},

then limtμDt({F})\lim_{t\to\infty}\mu_{D}^{t}(\{F\}) and limtνDt({F})\lim_{t\to\infty}\nu_{D}^{t}(\{F\}) exist and are equal for any FDF\in D.

Note that if we set 𝒟=𝒞\mathcal{D}=\mathcal{C} in the above, then Epi((A,,A^),(B,,B^))=Sur(A,B)\mathrm{Epi}((A,\langle\,,\rangle_{\widehat{A}}),(B,\langle\,,\rangle_{\widehat{B}}))=\operatorname{\mathrm{Sur}}^{*}(A,B) for (A,,A^),(B,,B^)𝒞(A,\langle\,,\rangle_{\widehat{A}}),(B,\langle\,,\rangle_{\widehat{B}})\in\mathcal{C}. Before showing that the (G,,G^)(G,\langle\,,\rangle_{\widehat{G}})-moments computed in Section 8 are well-behaved, we must first better understand the levels of 𝒞\mathcal{C} via the following lemma.

Lemma 9.4.

Let CbC_{b} be the level of 𝒞\mathcal{C} generated by the set of objects whose underlying group is /d\mathbb{Z}/d\mathbb{Z} for d|bd|b and whose pairing is any symmetric pairing on Hom(/d,/)\operatorname{\mathrm{Hom}}(\mathbb{Z}/d\mathbb{Z},\mathbb{Q}/\mathbb{Z}). Then CbC_{b} is the following set of isomorphism classes of objects: finite abelian groups whose exponent divides bb with all possible symmetric pairings on their dual groups.

Proof.

We begin by noting that, given two groups A,BFinAbA,B\in\mathrm{FinAb}, both AA and BB are quotients of A×BA\times B (we have B(A×B)/(A×{0})B\simeq(A\times B)/(A\times\{0\}), for example). Moreover, A×BA\times B is the join of AA and BB in the lattice of quotients of A×BA\times B.

Now, given any two objects (A,,A^),(B,,B^)𝒞(A,\langle\,,\rangle_{\widehat{A}}),(B,\langle\,,\rangle_{\widehat{B}})\in\mathcal{C}, let ,\langle\,,\rangle be a pairing on A^×B^Hom(A×B,/)\widehat{A}\times\widehat{B}\simeq\operatorname{\mathrm{Hom}}(A\times B,\mathbb{Q}/\mathbb{Z}) whose restriction to A^×{0}\widehat{A}\times\{0\} is ,A^\langle\,,\rangle_{\widehat{A}} and whose restriction to {0}×B^\{0\}\times\widehat{B} is ,B^\langle\,,\rangle_{\widehat{B}}. For any such pairing ,\langle\,,\rangle, we claim that (A×B,,)(A\times B,\langle\,,\rangle) is the join of (A,,A^)(A,\langle\,,\rangle_{\widehat{A}}) and (B,,B^)(B,\langle\,,\rangle_{\widehat{B}}) in the lattice of quotients of (A×B,,)(A\times B,\langle\,,\rangle). This follows from our remark in the previous paragraph and the fact that a quotient (G,,G^)(G,\langle\,,\rangle_{\widehat{G}}) of (A×B,,)(A\times B,\langle\,,\rangle) is a quotient (of groups) π:A×BG\pi:A\times B\to G such that the transpose πt:G^A^×B^\pi^{t}:\widehat{G}\to\widehat{A}\times\widehat{B} pushes the pairing ,\langle\,,\rangle forward to the pairing ,G^\langle\,,\rangle_{\widehat{G}}.

Let CC be a level of 𝒞\mathcal{C}, and suppose (A,,A^),(B,,B^)C(A,\langle\,,\rangle_{\widehat{A}}),(B,\langle\,,\rangle_{\widehat{B}})\in C for any symmetric pairings ,A^\langle\,,\rangle_{\widehat{A}} and ,B^\langle\,,\rangle_{\widehat{B}} on A^\widehat{A} and B^\widehat{B}, respectively. Given any symmetric pairing ϕ\phi on A^×B^\widehat{A}\times\widehat{B}, we note that (A×B,ϕ)(A\times B,\phi) is simply the join of (A,ϕ|A^)(A,\phi|_{\widehat{A}}) and (B,ϕ|B^)(B,\phi|_{\widehat{B}}), where ϕ|A^\phi|_{\widehat{A}} denotes pairing on A^\widehat{A} given by the restriction of ϕ\phi to A^×{0}\widehat{A}\times\{0\} (similarly for ϕ|B^\phi|_{\widehat{B}}). Therefore, (A×B,ϕ)C(A\times B,\phi)\in C.

Since any quotient of /d\mathbb{Z}/d\mathbb{Z} for d|bd|b has exponent dividing bb, and since the join of /d\mathbb{Z}/d\mathbb{Z} with any other such group (in FinAb\mathrm{FinAb}) has exponent dividing bb, we see that CbC_{b} only contains objects whose underlying group has exponent dividing bb. Now, any group GG whose exponent divides bb can be written as

G=i=1r/di,G=\bigoplus_{i=1}^{r}\mathbb{Z}/d_{i}\mathbb{Z},

where each di|bd_{i}|b. In other words, GG is the join (as groups) of the /di\mathbb{Z}/d_{i}\mathbb{Z}. By the above, the join-closedness of CbC_{b} forces CbC_{b} to contain GG equipped with any possible symmetric pairing on its dual, since CbC_{b} is generated by /d\mathbb{Z}/d\mathbb{Z} with any symmetric pairing on its dual, where dd is any divisor of bb. ∎

Ultimately, we would like to apply Lemma 9.3 to the category 𝒞\mathcal{C} at the level CbC_{b} for an arbitrary positive integer bb. Let DbD_{b} be the level in FinAb\mathrm{FinAb} generated by /b\mathbb{Z}/b\mathbb{Z}, i.e., the level of finite abelian groups of exponent dividing bb. Note that DbD_{b} is equal to the level of FinAb\mathrm{FinAb} generated by /d\mathbb{Z}/d\mathbb{Z} for d|bd|b. To prove that the titular sequence of (G,,G^)(G,\langle\,,\rangle_{\widehat{G}})-moments computed in Section 8 are well-behaved, we reduce the question of well-behavedness at CbC_{b} to one of well-behavedness at DbD_{b} in FinAb\mathrm{FinAb} using the following lemma.

Lemma 9.5 (Proof of Lemma 6.23 of [50]).

Let 𝒟\mathcal{D} be a diamond category and 𝒢\mathcal{G} a functor from 𝒟\mathcal{D} to the category of finite sets. Let CC be a level of the category of pairs, (𝒟,𝒢)(\mathcal{D},\mathcal{G}), and let DD be the level of 𝒟\mathcal{D} generated by the underlying objects of a finite set of generators of CC. Let (MH)HC(M_{H})_{H}\in\mathbb{R}^{C} be a set of moments. Then (MH)H(M_{H})_{H} is well-behaved at CC if (s𝒢(G)M(G,s))GD(\sum_{s\in\mathcal{G}(G)}M_{(G,s)})_{G}\in\mathbb{R}^{D} is well-behaved at DD.

When applied to 𝒞=(FinAb,Φ)\mathcal{C}=(\mathrm{FinAb},\Phi), Lemma 9.5 implies the following. For a sequence of moments (M(G,,G^))(G,,G^)(M_{(G,\langle\,,\rangle_{\widehat{G}})})_{(G,\langle\,,\rangle_{\widehat{G}})} indexed by elements of CbC_{b}, if the sequence of moments indexed by GDbG\in D_{b} given by

,G^asymmetricpairingonG^M(G,,G^)\sum_{\begin{subarray}{c}\langle\,,\rangle_{\widehat{G}}\mathrm{\ a\ symmetric}\\ \mathrm{pairing\ on\ }\widehat{G}\end{subarray}}M_{(G,\langle\,,\rangle_{\widehat{G}})}

is well-behaved at DbD_{b} for FinAb\mathrm{FinAb}, then the sequence (M(G,,G^))(G,,G^)(M_{(G,\langle\,,\rangle_{\widehat{G}})})_{(G,\langle\,,\rangle_{\widehat{G}})} is well-behaved at CbC_{b}.

Recall that for an object (G,,G^)Cb(G,\langle\,,\rangle_{\widehat{G}})\in C_{b}, Theorem 8.2 tells us that the (G,,G^)(G,\langle\,,\rangle_{\widehat{G}})-moment of the sandpile group of a random graph converges to |G|1|G|^{-1} as nn\to\infty, where nn is the number of vertices in the graph. Hence, we have reduced the problem to showing that the sequence in Db\mathbb{R}^{D_{b}} whose GG-entry is given by

#{symmetricpairingsonG^}|G|=|Sym2G||G|=|2G|\frac{\#\{\mathrm{symmetric\ pairings\ on\ }\widehat{G}\}}{|G|}=\frac{|\operatorname{\mathrm{Sym}}_{2}G|}{|G|}=|\wedge^{2}G|

is well-behaved.

For any prime pp, let p\mathcal{M}_{p} denote the category of finite p\mathbb{Z}_{p}-modules, i.e., the category of finite abelian pp-groups. For any set of primes PP, let FinAbP\mathrm{FinAb}_{P} be the product of categories pPp\prod_{p\in P}\mathcal{M}_{p}. If PP is the set of all primes, note that FinAb=FinAbP\mathrm{FinAb}=\mathrm{FinAb}_{P}.

Lemma 9.6 (Lemma 6.16 of [50]).

Let 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} be diamond categories. Then the product category 𝒟1×𝒟2\mathcal{D}_{1}\times\mathcal{D}_{2}, whose objects are ordered pairs of an object from 𝒟1\mathcal{D}_{1} and an object from 𝒟2\mathcal{D}_{2} and whose morphisms are ordered pairs of morphisms, is a diamond category.

Lemma 9.7 (Lemma 6.17 of [50]).

Let 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} be diamond categories. Let (MG1)G𝒟1/(M_{G}^{1})_{G\in\mathcal{D}_{1}/\simeq} and (MG2)G𝒟2/(M_{G}^{2})_{G\in\mathcal{D}_{2}/\simeq} be tuples of nonnegative real numbers indexed by the respective isomorphism classes of 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2}. Define (M(G1,G2))(G1,G2)𝒟1×𝒟2/(M_{(G_{1},G_{2})})_{(G_{1},G_{2})\in\mathcal{D}_{1}\times\mathcal{D}_{2}/\simeq} by the formula M(G1,G2)=MG11MG22M_{(G_{1},G_{2})}=M_{G_{1}}^{1}M_{G_{2}}^{2}. If (MG1)G𝒟1/(M_{G}^{1})_{G\in\mathcal{D}_{1}/\simeq} and (MG2)G𝒟2/(M_{G}^{2})_{G\in\mathcal{D}_{2}/\simeq} are both well-behaved for 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2}, then (M(G1,G2))(G1,G2)𝒟1×𝒟2/(M_{(G_{1},G_{2})})_{(G_{1},G_{2})\in\mathcal{D}_{1}\times\mathcal{D}_{2}/\simeq} is well-behaved for the product category 𝒟1×𝒟2\mathcal{D}_{1}\times\mathcal{D}_{2}.

Let PP be a finite set of primes. For each pPp\in P, consider a sequence of moments in p\mathcal{M}_{p}, and form a sequence of moments in FinAbP\mathrm{FinAb}_{P} by taking all products of the moments in the component categories p\mathcal{M}_{p}. By Lemma 9.7, if each of the component sequences of moments are well-behaved, so is this sequence of moments in FinAbP\mathrm{FinAb}_{P}. Therefore, for the sequence of moments (|2G|)GFinAbP/(|\wedge^{2}G|)_{G}\in\mathbb{R}^{\mathrm{FinAb}_{P}/\simeq}, it suffices to check that the sequence (|2G|)Gp/(|\wedge^{2}G|)_{G}\in\mathbb{R}^{\mathcal{M}_{p}/\simeq} is well-behaved for each pp, since

2G=p2(Gp).\wedge^{2}G=\bigoplus_{p}\wedge^{2}(G_{p}).

To do this, we simply apply the following lemma from [50] to the category of finite p\mathbb{Z}_{p}-modules.

Lemma 9.8 (Lemma 6.9 of [50]).

Let RR be a discrete valuation ring with finite residue field. For each isomorphism class NN of finite RR-modules, let MNM_{N} be a real number. Then MNM_{N} is well-behaved if there exists some ϵ>0\epsilon>0 and cc such that MNc|N|1ϵ|2N|M_{N}\leq c|N|^{1-\epsilon}|\wedge^{2}N| for all finite RR-modules NN.

Applying Lemma 9.8 to p\mathcal{M}_{p}, the category of finite p\mathbb{Z}_{p}-modules, it follows that the moments in p\mathcal{M}_{p} given by |2G||\wedge^{2}G| for each GpG\in\mathcal{M}_{p} are well-behaved. Hence, the corresponding moments in FinAbP\mathrm{FinAb}_{P} are also well-behaved. To conclude the section, we repackage all of our above work in the following explicit theorem.

Theorem 9.9.

Let (Xn,,Xn^)(X_{n},\langle\,,\rangle_{\widehat{X_{n}}}) and (Yn,,Yn^)(Y_{n},\langle\,,\rangle_{\widehat{Y_{n}}}) be sequences of random variables taking values in the category of finitely generated abelian groups whose duals are equipped with symmetric pairings. Let bb be a positive integer and CbC_{b} the set of isomorphism classes of finite abelian groups of exponent dividing bb with symmetric pairings on their duals. Assume that for every (G,,G^)Cb(G,\langle\,,\rangle_{\widehat{G}})\in C_{b},

limn𝔼(#Sur(Xn,G))=|G|1\lim_{n\to\infty}\mathbb{E}(\#\operatorname{\mathrm{Sur}}^{*}(X_{n},G))=|G|^{-1}

and that the analogous statement holds for (Yn,,Yn^)(Y_{n},\langle\,,\rangle_{\widehat{Y_{n}}}). Then for every (G,,G^)Cb(G,\langle\,,\rangle_{\widehat{G}})\in C_{b} the limits limn((Xn/b,,Xn^)(G,,G^))\lim_{n\to\infty}\operatorname{\mathbb{P}}((X_{n}\otimes\mathbb{Z}/b\mathbb{Z},\langle\,,\rangle_{\widehat{X_{n}}})\simeq(G,\langle\,,\rangle_{\widehat{G}})) and limn((Yn/b,,Yn^)(G,,G^))\lim_{n\to\infty}\operatorname{\mathbb{P}}((Y_{n}\otimes\mathbb{Z}/b\mathbb{Z},\langle\,,\rangle_{\widehat{Y_{n}}})\simeq(G,\langle\,,\rangle_{\widehat{G}})) exist, and

limn((Xn/b,,Xn^)(G,,G^))=limn((Yn/b,,Yn^)(G,,G^)).\lim_{n\to\infty}\operatorname{\mathbb{P}}((X_{n}\otimes\mathbb{Z}/b\mathbb{Z},\langle\,,\rangle_{\widehat{X_{n}}})\simeq(G,\langle\,,\rangle_{\widehat{G}}))=\lim_{n\to\infty}\operatorname{\mathbb{P}}((Y_{n}\otimes\mathbb{Z}/b\mathbb{Z},\langle\,,\rangle_{\widehat{Y_{n}}})\simeq(G,\langle\,,\rangle_{\widehat{G}})).
Proof.

Recall that CbC_{b} is the level of 𝒞\mathcal{C} generated by /d\mathbb{Z}/d\mathbb{Z} for d|bd|b with every possible symmetric pairing on its dual. We have a sequence of (G,,G^)(G,\langle\,,\rangle_{\widehat{G}})-moments given by |G|1|G|^{-1} for (G,,G^)Cb(G,\langle\,,\rangle_{\widehat{G}})\in C_{b}, since for each (G,,G^)Cb(G,\langle\,,\rangle_{\widehat{G}})\in C_{b} we have

limn𝔼(#Sur(Xn/b,G))=|G|1.\lim_{n\to\infty}\mathbb{E}(\#\operatorname{\mathrm{Sur}}^{*}(X_{n}\otimes\mathbb{Z}/b\mathbb{Z},G))=|G|^{-1}.

By Lemma 9.5 and our previous discussion, to show that the moments are well-behaved at CbC_{b}, it suffices to check well-behavedness at the level DbD_{b} of FinAb\mathrm{FinAb} (recall that DbD_{b} is the level of all finite abelian groups of exponent dividing bb).

Let PP be the finite set of primes dividing bb, and consider FinAbP\mathrm{FinAb}_{P}. The level in FinAbP\mathrm{FinAb}_{P} generated by /b\mathbb{Z}/b\mathbb{Z} is exactly the same as DbD_{b}, so we may check well-behavedness at this level in FinAbP\mathrm{FinAb}_{P}. By Lemma 9.7 and our work in the above, we see that the sequence of moments given by |2G||\wedge^{2}G| for GDbG\in D_{b} is well-behaved at DbD_{b}, implying the sequence of moments (|G|1)(G,,G^)Cb(|G|^{-1})_{(G,\langle\,,\rangle_{\widehat{G}})\in C_{b}} is well-behaved at CbC_{b}. Thus, we may apply Lemma 9.3 and conclude the theorem. ∎

10. Comparison to uniform random matrices and their pairings

In the previous section, we showed that the moments we obtained for the sandpile groups of random graphs with their pairings determine a unique measure on certain levels of the category of finite abelian groups whose duals carry symmetric pairings. In order to determine the values of this measure at specific groups with pairings, we compare to the uniform case. In particular, Theorem 8.1 implies that cokernels of uniform random matrices over /a\mathbb{Z}/a\mathbb{Z} with their pairings have the same moments as sandpile groups with their pairings. (Although the values of the moments of cokernels of uniform random matrices and their pairings are given by Theorem 8.1, we note that it is possible to compute these explicitly by adapting methods from Section 4 and the proof of Theorem 11 of [15].) In [15], Clancy, Kaplan, Leake, Payne, and Wood computed several of the aforementioned probabilities in the uniform case explicitly.

We will need the following result of [15], which gives the (asymptotic) distribution of the cokernel of a uniform random symmetric matrix with its natural duality pairing.

Lemma 10.1 (Theorem 2 of [15]).

Let GG be a finite abelian pp-group with a duality pairing ,G^\langle\,,\rangle_{\widehat{G}} on G^\widehat{G}, and let MM be a random n×nn\times n symmetric matrix whose entries are distributed with respect to additive Haar measure on p\mathbb{Z}_{p}. Then the asymptotic probability (as nn\to\infty) that the cokernel of MM with its duality pairing is isomorphic to (G,,G^)(G,\langle\,,\rangle_{\widehat{G}}) is

limn((cok(M),,M)(G,,G^))=k1(1p12k)|G||Aut(G,,G)|.\lim_{n\to\infty}\operatorname{\mathbb{P}}((\mathrm{cok}(M),\langle\,,\rangle_{M})\simeq(G,\langle\,,\rangle_{\widehat{G}}))=\frac{\prod_{k\geq 1}(1-p^{1-2k})}{|G||\operatorname{\mathrm{Aut}}(G,\langle\,,\rangle_{G})|}.

By extending the measure in the above result by 0, we can extend the above to give probabilities for when the pairing on G^\widehat{G} is any symmetric pairing. Notably, we do not require the pairing to be perfect.

Corollary 10.2 (of Theorem 9.9).

Let GG be a finite abelian group of exponent dividing bb, and suppose GG has a symmetric pairing on its dual G^=Hom(G,/)\widehat{G}=\operatorname{\mathrm{Hom}}(G,\mathbb{Q}/\mathbb{Z}), which we will denote using ,G^\langle\,,\rangle_{\widehat{G}}. Let 0<q<10<q<1, let ΓG(n,q)\Gamma\in G(n,q) be a random graph, and let SS be its sandpile group. Denote the canonical symmetric pairing on S^\widehat{S} by ,S^\langle\,,\rangle_{\widehat{S}}. Also let HnH_{n} be a uniform random n×nn\times n symmetric matrix with entries in /b\mathbb{Z}/b\mathbb{Z}. Then

limn((S/b,,S^)(G,,G^))\displaystyle\lim_{n\to\infty}\operatorname{\mathbb{P}}((S\otimes\mathbb{Z}/b\mathbb{Z},\langle\,,\rangle_{\widehat{S}})\simeq(G,\langle\,,\rangle_{\widehat{G}})) =limn((cok(Hn),,Hn)(G,,G^)).\displaystyle=\lim_{n\to\infty}\operatorname{\mathbb{P}}((\mathrm{cok}(H_{n}),\langle\,,\rangle_{H_{n}})\simeq(G,\langle\,,\rangle_{\widehat{G}})).

In particular, if ,G^\langle\,,\rangle_{\widehat{G}} is perfect, then

limn((S/b,,S^)(G,,G^))=p|bk1(1p12k)|G||Aut(G,,G^)|.\lim_{n\to\infty}\operatorname{\mathbb{P}}((S\otimes\mathbb{Z}/b\mathbb{Z},\langle\,,\rangle_{\widehat{S}})\simeq(G,\langle\,,\rangle_{\widehat{G}}))=\frac{\prod_{p|b}\prod_{k\geq 1}(1-p^{1-2k})}{|G||\operatorname{\mathrm{Aut}}(G,\langle\,,\rangle_{\widehat{G}})|}.

Otherwise, limn((S/b,,S^)(G,,G^))=0\lim_{n\to\infty}\operatorname{\mathbb{P}}((S\otimes\mathbb{Z}/b\mathbb{Z},\langle\,,\rangle_{\widehat{S}})\simeq(G,\langle\,,\rangle_{\widehat{G}}))=0.

Proof.

That

limn((S/b,,S^)(G,,G^))\displaystyle\lim_{n\to\infty}\operatorname{\mathbb{P}}((S\otimes\mathbb{Z}/b\mathbb{Z},\langle\,,\rangle_{\widehat{S}})\simeq(G,\langle\,,\rangle_{\widehat{G}})) =limn((cok(Hn),,Hn)(G,,G^)).\displaystyle=\lim_{n\to\infty}\operatorname{\mathbb{P}}((\mathrm{cok}(H_{n}),\langle\,,\rangle_{H_{n}})\simeq(G,\langle\,,\rangle_{\widehat{G}})).

follows from applying Theorem 9.9. In particular, we use the fact that both (S/b,,S^)(S\otimes\mathbb{Z}/b\mathbb{Z},\langle\,,\rangle_{\widehat{S}}) and (cok(Hn),,Hn)(\mathrm{cok}(H_{n}),\langle\,,\rangle_{H_{n}}) have the same asymptotic (,,^)(\mathcal{H},\langle\,,\rangle_{\widehat{\mathcal{H}}})-moment for any object (,,^)Cb(\mathcal{H},\langle\,,\rangle_{\widehat{\mathcal{H}}})\in C_{b} (this is a consequence of Theorem 8.2).

The second part of the lemma follows from noting that everything factors over p|bp|b, allowing us to reduce to the case where GG is a pp-group. ∎

To prove Theorem 1.1, we will turn Corollary 10.2 into a statement about the Sylow pp-subgroups of the sandpile group. In particular, we convert information about S/bS\otimes\mathbb{Z}/b\mathbb{Z} for any bb into information about the Sylow pp-subgroups of SS for p|bp|b, and we translate statements about pairings on the dual of S/bS\otimes\mathbb{Z}/b\mathbb{Z} to statements about pairings on the Sylow subgroups of SS.

Remark 10.3.

First, we will need to know that SS is finite with asymptotic probability 1 as nn\to\infty. While this result is well-known, we offer an alternate proof below. Recall that SS is finite if and only if Γ\Gamma is connected; Γ\Gamma is connected with probability 1 as nn\to\infty for fixed qq (this is a corollary of results proved by Erdős and Rényi in their 1960 paper [23]). Hence, as nn\to\infty, we know that SS is finite with probability 1. Using Fatou’s lemma and results from [54] and [15], we give an alternate proof of this fact. If we replace SS in Lemma 10.4 by cok(M)\mathrm{cok}(M), where MM is a random matrix satisfying the hypotheses of Theorem 8.1, we get an analogous result for cok(M)\mathrm{cok}(M). The result for cok(M)\mathrm{cok}(M) is newer but also known—it is implied by Theorem 6.5 of [19] and also by Theorem 5.1 of [46].

Lemma 10.4.

Let SS be the sandpile group of a random graph on nn vertices as in Theorem 8.2. As nn\to\infty, we have that (|S|<)1\operatorname{\mathbb{P}}(|S|<\infty)\to 1.

Proof.

Let PP be any finite set of primes, and let P=pPp\mathbb{Z}_{P}=\prod_{p\in P}\mathbb{Z}_{p}. By Corollary 9.2 of [54], we have

limn(SPG)=#{symmetric,perfectϕ:GG/}|G||Aut(G)|pPk1(1p12k)\lim_{n\to\infty}\operatorname{\mathbb{P}}(S\otimes\mathbb{Z}_{P}\simeq G)=\frac{\#\{\mathrm{symmetric,\ perfect\ }\phi:G\otimes G\to\mathbb{Q}/\mathbb{Z}\}}{|G||\operatorname{\mathrm{Aut}}(G)|}\prod_{p\in P}\prod_{k\geq 1}(1-p^{1-2k})

for any finite abelian PP-group GG.555The proof of Corollary 9.2 of [54] works for our purposes, but note that it contains a slight mistake—throughout, “Sylow PP-subgroup of SS” should be replaced by SPS\otimes\mathbb{Z}_{P}. Proposition 7 of [15] tells us that the sum over all finite abelian PP-groups of the right-hand side of the above is 1. By Fatou’s lemma, it follows that

limn(|S|<)=1,\lim_{n\to\infty}\operatorname{\mathbb{P}}(|S|<\infty)=1,

as desired. ∎

Corollary 10.5.

Let GG be a finite abelian group equipped with a duality pairing ,G\langle\,,\rangle_{G}. Suppose ΓG(n,q)\Gamma\in G(n,q) is a random graph with Laplacian LL and sandpile group SS. Denote the canonical duality pairing on tcok(L)\mathrm{tcok}(L) by ,tcok(L)\langle\,,\rangle_{\mathrm{tcok}(L)}. Suppose PP is the finite set of prime divisors of |G||G|. Let SPS_{P} denote the sum of the Sylow pp-subgroups of SS for pPp\in P and ,tcok(L),P\langle\,,\rangle_{\mathrm{tcok}(L),P} the restriction of ,tcok(L)\langle\,,\rangle_{\mathrm{tcok}(L)} to SPS_{P}. Then

limn((SP,,tcok(L),P)(G,,G))=pPk1(1p12k)|G||Aut(G,,G)|.\lim_{n\to\infty}\operatorname{\mathbb{P}}((S_{P},\langle\,,\rangle_{\mathrm{tcok}(L),P})\simeq(G,\langle\,,\rangle_{G}))=\frac{\prod_{p\in P}\prod_{k\geq 1}(1-p^{1-2k})}{|G||\operatorname{\mathrm{Aut}}(G,\langle\,,\rangle_{G})|}.
Proof.

Write the exponent of GG as pPpap\prod_{p\in P}p^{a_{p}}, and let e=pPpap+1e=\prod_{p\in P}{p^{a_{p}+1}}. Let ,G^\langle\,,\rangle_{\widehat{G}} be the pairing on G^\widehat{G} induced by ,G\langle\,,\rangle_{G}. Recall that SPtcok(L)S_{P}\subset\mathrm{tcok}(L), so ,tcok(L),P\langle\,,\rangle_{\mathrm{tcok}(L),P} is also perfect. Also recall that the dual of S/eS\otimes\mathbb{Z}/e\mathbb{Z} comes equipped with a symmetric pairing, denoted ,S^\langle\,,\rangle_{\widehat{S}}, given by pushing the pairing on S^\widehat{S} forward by the transpose of the natural quotient SS/eS\twoheadrightarrow S\otimes\mathbb{Z}/e\mathbb{Z}. We have the following commutative diagram:

SP{S_{P}}tcok(L){\mathrm{tcok}(L)}S{S}S/e{S\otimes\mathbb{Z}/e\mathbb{Z}}SP^{\widehat{S_{P}}}tcok^(L){\mathrm{tc}\widehat{\mathrm{ok}}(L)}S^{\widehat{S}}Hom(S/e,/),{\operatorname{\mathrm{Hom}}(S\otimes\mathbb{Z}/e\mathbb{Z},\mathbb{Q}/\mathbb{Z}),}

where the vertical maps are isomorphisms induced by the duality pairing ,tcok(L)\langle\,,\rangle_{\mathrm{tcok}(L)}.

Assume SS is finite. We will show that

(S/e,,S^)(G,,G^) if and only if (SP,,tcok(L),P)(G,,G).(S\otimes\mathbb{Z}/e\mathbb{Z},\langle\,,\rangle_{\widehat{S}})\simeq(G,\langle\,,\rangle_{\widehat{G}})\textrm{ if and only if }(S_{P},\langle\,,\rangle_{\mathrm{tcok}(L),P})\simeq(G,\langle\,,\rangle_{G}).

To prove this, first note that SS finite implies that SPGS_{P}\simeq G if and only if S/eGS\otimes\mathbb{Z}/e\mathbb{Z}\simeq G. Thus, it is sufficient to show that the data of the pairing ,S^\langle\,,\rangle_{\widehat{S}} on the dual of S/eS\otimes\mathbb{Z}/e\mathbb{Z} is equivalent to the data of the pairing ,tcok(L),P\langle\,,\rangle_{\mathrm{tcok}(L),P} on SPS_{P}. Suppose S/eGSPS\otimes\mathbb{Z}/e\mathbb{Z}\simeq G\simeq S_{P}. This forces the composition of maps in the first row of the above diagram to be an isomorphism between SPS_{P} and S/eS\otimes\mathbb{Z}/e\mathbb{Z}. The isomorphism SPS/eS_{P}\simeq S\otimes\mathbb{Z}/e\mathbb{Z} gives us a duality pairing on S/eS\otimes\mathbb{Z}/e\mathbb{Z}, which in turn gives us a duality pairing on Hom(S/e,/)\operatorname{\mathrm{Hom}}(S\otimes\mathbb{Z}/e\mathbb{Z},\mathbb{Q}/\mathbb{Z}). We will prove that this induced pairing on Hom(S/e,/)\operatorname{\mathrm{Hom}}(S\otimes\mathbb{Z}/e\mathbb{Z},\mathbb{Q}/\mathbb{Z}) is just ,S^\langle\,,\rangle_{\widehat{S}}.

Recall from Lemma 3.5 that the pairing ,S^\langle\,,\rangle_{\widehat{S}} on S^\widehat{S} is induced by the duality pairing ,tcok(L)\langle\,,\rangle_{\mathrm{tcok}(L)} on tcok(L)\mathrm{tcok}(L). Moreover, recall that we may view the pairing ,S^\langle\,,\rangle_{\widehat{S}} on Hom(S/e,/)\operatorname{\mathrm{Hom}}(S\otimes\mathbb{Z}/e\mathbb{Z},\mathbb{Q}/\mathbb{Z}) as the restriction of ,S^\langle\,,\rangle_{\widehat{S}} to Hom(S/e,/)\operatorname{\mathrm{Hom}}(S\otimes\mathbb{Z}/e\mathbb{Z},\mathbb{Q}/\mathbb{Z}). Hence, to compute ,S^\langle\,,\rangle_{\widehat{S}} on Hom(S/e,/)\operatorname{\mathrm{Hom}}(S\otimes\mathbb{Z}/e\mathbb{Z},\mathbb{Q}/\mathbb{Z}), we may look at the corresponding elements in tcok(L)\mathrm{tcok}(L) and do the computation using ,tcok(L)\langle\,,\rangle_{\mathrm{tcok}(L)}. Since the elements of Hom(S/e,/Z)\operatorname{\mathrm{Hom}}(S\otimes\mathbb{Z}/e\mathbb{Z},\mathbb{Q}/Z) all have order dividing ee, their images in tcok(L)\mathrm{tcok}(L) are PP-torsion. In other words, the image of Hom(S/e,/Z)\operatorname{\mathrm{Hom}}(S\otimes\mathbb{Z}/e\mathbb{Z},\mathbb{Q}/Z) in tcok(L)\mathrm{tcok}(L) lies in SPS_{P}. It follows that ,S^\langle\,,\rangle_{\widehat{S}} is exactly the pairing on Hom(S/e,/)SP^\operatorname{\mathrm{Hom}}(S\otimes\mathbb{Z}/e\mathbb{Z},\mathbb{Q}/\mathbb{Z})\simeq\widehat{S_{P}} induced by ,tcok(LΓ),P\langle\,,\rangle_{\mathrm{tcok}(L_{\Gamma}),P}, proving that (S/e,,S^)(G,,G^) if and only if (SP,,tcok(L),P)(G,,G)(S\otimes\mathbb{Z}/e\mathbb{Z},\langle\,,\rangle_{\widehat{S}})\simeq(G,\langle\,,\rangle_{\widehat{G}})\textrm{ if and only if }(S_{P},\langle\,,\rangle_{\mathrm{tcok}(L),P})\simeq(G,\langle\,,\rangle_{G}).

By Remark 10.3, we know that SS is asymptotically almost surely finite. Therefore, as nn\to\infty, we have that

((S/e,,S^)(G,,G^))=((SP,,tcok(L),P)(G,,G)),\operatorname{\mathbb{P}}((S\otimes\mathbb{Z}/e\mathbb{Z},\langle\,,\rangle_{\widehat{S}})\simeq(G,\langle\,,\rangle_{\widehat{G}}))=\operatorname{\mathbb{P}}((S_{P},\langle\,,\rangle_{\mathrm{tcok}(L),P})\simeq(G,\langle\,,\rangle_{G})),

which, in conjunction with Corollary 10.2, implies the result. ∎

Remark 10.6.

Because SΓS_{\Gamma} is finite with probability 1 as nn\to\infty, in the limit, the pairing ,tcok(L)\langle\,,\rangle_{\mathrm{tcok}(L)} is simply the canonical duality pairing on SΓS_{\Gamma}. Hence, Corollary 10.5 can actually be viewed as a statement about the distribution of the Sylow PP-part of the sandpile group and its canonical duality pairing (more accurately, the restriction of this pairing to the Sylow PP-subgroup).

Moreover, all of the results in Section 10 hold if SS is replaced with the cokernel of a random symmetric integral matrix satisfying the hypotheses of Theorem 8.1, where we use Theorem 8.1 rather than Theorem 8.2.

Acknowledgements

This research was conducted at Harvard University and the Duluth Summer Mathematics Research Program for Undergraduates at the University of Minnesota Duluth with support from Jane Street Capital, the National Security Agency, the National Science Foundation (grants 2140043 and 2052036), Harvard University, and the Harvard College Research Program. I am very grateful to Joe Gallian and Colin Defant for their support and for giving me the opportunity to return to Duluth. I would also like to extend special thanks to Nathan Kaplan, Gilyoung Cheong, and Dino Lorenzini for their insightful suggestions that greatly improved the quality of the paper. Finally, I am most deeply grateful to my research advisor, Melanie Matchett Wood, who, in addition to suggesting this problem, provided invaluable support and guidance throughout the research process.

References

  • [1] Bacher, R., La Harpe, P. d., and Nagnibeda, T. The lattice of integral flows and the lattice of integral cuts on a finite graph. Bulletin de la Société Mathématique de France 125, 2 (1997), 167–198.
  • [2] Bai, Z. Circular law. The Annals of Probability 25, 1 (1997), 494–529.
  • [3] Bai, Z., and Silverstein, J. W. Spectral Analysis of Large Dimensional Random Matrices. Springer, New York, 2010.
  • [4] Bak, P., Tang, C., and Wiesenfeld, K. Self-organized criticality: An explanation of the 1/f1/f noise. Phys. Rev. Lett. 59 (Jul 1987), 381–384.
  • [5] Baker, M., and Norine, S. Riemann–Roch and Abel–Jacobi theory on a finite graph. Advances in Mathematics 215, 2 (2007), 766–788.
  • [6] Baker, M., and Norine, S. Harmonic morphisms and hyperelliptic graphs. International Mathematics Research Notices 2009 (08 2009), 2914–2955.
  • [7] Bhargava, A., DePascale, J., and Koenig, J. The rank of the sandpile group of random directed bipartite graphs. Annals of Combinatorics (Apr 2023).
  • [8] Bhargava, M., Kane, D., Lenstra, H., Rains, E., and Poonen, B. Modeling the distribution of ranks, Selmer groups, and Shafarevich–Tate groups of elliptic curves. Cambridge Journal of Math 3, 3 (08 2015), 257–321.
  • [9] Biggs, N. Algebraic Graph Theory, 2 ed. Cambridge Mathematical Library. Cambridge University Press, 1974.
  • [10] Biggs, N. Algebraic potential theory on graphs. Bulletin of the London Mathematical Society 29, 6 (1997), 641–682.
  • [11] Biggs, N. Chip-firing and the critical group of a graph. Journal of Algebraic Combinatorics 9 (1999), 25–45.
  • [12] Björner, A., Lovász, L., and Shor, P. W. Chip-firing games on graphs. European Journal of Combinatorics 12, 4 (1991), 283–291.
  • [13] Bosch, S., and Lorenzini, D. Grothendieck’s pairing on component groups of Jacobians. Inventiones mathematicae 148 (05 2002), 353–396.
  • [14] Cheong, G., and Yu, M. The distribution of the cokernel of a polynomial evaluated at a random integral matrix, preprint.
  • [15] Clancy, J., Kaplan, N., Leake, T., Payne, S., and Matchett Wood, M. On a Cohen–Lenstra heuristic for Jacobians of random graphs. Journal of Algebraic Combinatorics 42 (2015), 701–723.
  • [16] Clancy, J., Leake, T., and Payne, S. A note on Jacobians, Tutte polynomials, and two-variable zeta functions of graphs. Experimental Mathematics 24, 1 (2015), 1–7.
  • [17] Cohen, H., and Lenstra, H. W. Heuristics on class groups of number fields. In Number Theory Noordwijkerhout 1983 (Berlin, Heidelberg, 1984), H. Jager, Ed., Springer Berlin Heidelberg, pp. 33–62.
  • [18] Corry, S., and Perkinson, D. Divisors and sandpiles. American Mathematical Society, Providence, RI, 2018. An introduction to chip-firing.
  • [19] Costello, K. P., Tao, T., and Vu, V. Random symmetric matrices are almost surely nonsingular. Duke Math. J. 135, 2 (2006), 395–413.
  • [20] Dauns, J. Module types. The Rocky Mountain Journal of Mathematics 27, 2 (1997), 503–557.
  • [21] Delaunay, C. Heuristics on Tate-Shafarevitch groups of elliptic curves defined over \mathbb{Q}. Experimental Mathematics 10 (2001), 191–196.
  • [22] Dhar, D. Self-organized critical state of sandpile automaton models. Phys. Rev. Lett. 64 (Apr 1990), 1613–1616.
  • [23] Erdős, P., and Rényi, A. On the evolution of random graphs. Princeton University Press, Princeton, 2006, pp. 38–82.
  • [24] Friedman, E., and Washington, L. C. On the distribution of divisor class groups of curves over a finite field. In Théorie des nombres (Quebec, PQ, 1987). de Gruyter, Berlin, 1989, pp. 227–239.
  • [25] Gabrielov, A. Abelian avalanches and Tutte polynomials. Physica A: Statistical Mechanics and its Applications 195, 1 (1993), 253–274.
  • [26] Gabrielov, A. Avalanches, Sandpiles and Tutte Decomposition. Birkhäuser Boston, Boston, MA, 1993, pp. 19–26.
  • [27] Girko, V. L. The strong circular law. Twenty years later. Part ii. Random Operators and Stochastic Equations 12, 3 (2004), 255–312.
  • [28] Gorokhovsky, E. Convergence of time-inhomogeneous random walks on finite groups with applications to universality for random groups. Master’s thesis, California Institute of Technology, 2023.
  • [29] Grothendieck, A. Complements sur les biextensions. proprietes generales des biextensions des schemas en groupes. In Groupes de Monodromie en Géométrie Algébrique (Berlin, Heidelberg, 1972), Springer Berlin Heidelberg, pp. 218–312.
  • [30] Grothendieck, A., and Raynaud, M. Modeles de Neron et monodromie. In Groupes de Monodromie en Géométrie Algébrique (Berlin, Heidelberg, 1972), Springer Berlin Heidelberg, pp. 313–523.
  • [31] Holroyd, A. E., Levine, L., Mészáros, K., Peres, Y., Propp, J., and Wilson, D. B. Chip-Firing and Rotor-Routing on Directed Graphs. Birkhäuser Basel, Basel, 2008, pp. 331–364.
  • [32] Horton, M., Stark, H., and Terras, A. What are zeta functions of graphs and what are they good for? Contemporary Mathematics 415 (01 2006), 173–189.
  • [33] Koplewitz, S. Sandpile groups of random bipartite graphs. Ann. Comb. 27, 1 (2023), 1–18.
  • [34] Lee, J. Joint distribution of the cokernels of random p-adic matrices. Forum Mathematicum 35, 4 (2023), 1005–1020.
  • [35] Lee, J. Mixed moments and the joint distribution of random groups, preprint.
  • [36] Lee, J. Universality of the cokernels of random pp-adic Hermitian matrices, preprint.
  • [37] Levine, L., and Propp, J. What is… a sandpile ? Notices of the American Mathematical Society 57, 8 (2010), 976–979.
  • [38] Lipnowski, M., Sawin, W., and Tsimerman, J. Cohen-Lenstra heuristics and bilinear pairings in the presence of roots of unity, preprint.
  • [39] Lopez, C. M. Chip firing and the Tutte polynomial. Annals of Combinatorics 1 (1997), 253–259.
  • [40] Lorenzini, D. Arithmetical graphs. Mathematische Annalen 285, 3 (1989), 481–502.
  • [41] Lorenzini, D. Arithmetical properties of Laplacians of graphs. Linear and Multilinear Algebra 47, 4 (2000), 281–306.
  • [42] Lorenzini, D. Smith normal form and Laplacians. Journal of Combinatorial Theory, Series B 98, 6 (2008), 1271–1300.
  • [43] Mehta, M. L. Random matrices and the Statistical Theory of Energy Levels. Academic Press, New York-London, 1967.
  • [44] Mészáros, A. The distribution of sandpile groups of random regular graphs. Transactions of the American Mathematical Society 379, 9 (2020), 6529–6594.
  • [45] Nguyen, H. H., and Peski, R. V. Universality for cokernels of random matrix products, preprint.
  • [46] Nguyen, H. H., and Wood, M. M. Random integral matrices: universality of surjectivity and the cokernel. Inventiones mathematicae 228 (2022), 1–76.
  • [47] Nguyen, H. H., and Wood, M. M. Local and global universality of random matrix cokernels, preprint.
  • [48] Norine, S., and Whalen, P. Jacobians of nearly complete and threshold graphs. European Journal of Combinatorics 32, 8 (2011), 1368–1376.
  • [49] Pan, G., and Zhou, W. Circular law, extreme singular values and potential theory. Journal of Multivariate Analysis 101, 3 (2010), 645–656.
  • [50] Sawin, W., and Wood, M. M. The moment problem for random objects in a category, preprint.
  • [51] Shokrieh, F. The monodromy pairing and discrete logarithm on the Jacobian of finite graphs. Journal of Mathematical Cryptology 4, 1 (2010), 43–56.
  • [52] Tao, T., and Vu, V. On random ±1\pm 1 matrices: Singularity and determinant. Random Structures & Algorithms 28, 1 (2006), 1–23.
  • [53] Tao, T., and Vu, V. On the singularity probability of random Bernoulli matrices. Journal of the American Mathematical Society 20, 3 (2007), 603–628.
  • [54] Wood, M. The distribution of sandpile groups of random graphs. Journal of the American Mathematical Society 30 (02 2014), 915–958.
  • [55] Wood, M. Random integral matrices and the Cohen Lenstra heuristics. American Journal of Mathematics 141, 2 (04 2019), 383–398.
  • [56] Wood, M. Probability theory for random groups arising in number theory. Proceedings of the International Congress of Mathematicians (2022), to appear.
  • [57] Yan, E. Universality for cokernels of Dedekind domain valued random matrices, preprint.