The Distribution of Sandpile Groups of Random Graphs with their Pairings
Abstract.
We determine the distribution of the sandpile group (also known as the Jacobian) of the Erdős-Rényi random graph along with its canonical duality pairing as 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 -group equipped with a perfect symmetric pairing appears as the Sylow -part of the sandpile group and its pairing with frequency inversely proportional to , where is the set of automorphisms of preserving the pairing . 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 we may associate a natural abelian group , 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 is connected, then is finite, and is the number of spanning trees of . 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) , which was first introduced by Bosch and Lorenzini in [13]. In [42], Lorenzini conjectured the frequency that 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 seems to appear as the sandpile group and pairing of a random graph with frequency proportional to
where denotes the automorphisms of preserving the pairing. In other words, they observed that the natural duality pairing on 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 (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 be an Erdős-Rényi random graph on vertices with edges occurring independently with probability , where . Let denote the sandpile group of , and for a prime let denote the Sylow -subgroup of . Also, let denote the duality pairing on the torsion subgroup of and its restriction to .111As we will see in Section 3, it is actually the torsion subgroup of the sandpile group that naturally carries a duality pairing. When is connected, the sandpile group is finite and the torsion subgroup is all of . 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 , we have . Recalling that a finite abelian group can be decomposed into a direct sum of its Sylow -subgroups, we instead ask: for a finite abelian -group with duality pairing , what is the probability as that ?
Theorem 1.1.
Let be a prime and a finite abelian -group with duality pairing . For a random graph , let denote its sandpile group and the duality pairing on its torsion subgroup. Then
(1) |
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 and that this product is essentially a normalization constant. Moreover, increases with , and for we have . Note that the right-hand side of (1) is independent of the edge probability . 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 is a duality pairing, there is a corresponding duality pairing on , the Pontryagin dual group of , which we denote by . To prove Theorem 1.1, we first compute a complete set of moments for .
It is not immediately clear which category we should be working in or why we are considering pairings on rather than . We define a category whose objects are finite abelian groups whose dual groups are equipped with symmetric pairings (i.e., the objects of are finite abelian groups such that has a symmetric pairing ). Note that if and are objects in , and if is a group homomorphism, then the transpose of , the map , pushes the pairing on forward to a pairing on by precomposition with . A morphism between and in is a group homomorphism between and that pushes the pairing on forward to the pairing on . The fact that pushes the pairing on forward to a pairing on is the primary reason for considering pairings on the dual groups. Suppressing the symbols for the pairings on and from our notation, let denote the set of surjective homomorphisms from to taking to .
The following gives the moments of random sandpile groups with their pairings.
Theorem 1.2.
Let be a finite abelian group, and let be a symmetric pairing on its dual, . Then for a random graph with its sandpile group and symmetric pairing on , we have
We call the -moment of . The error term for Theorem 1.2 decreases exponentially in (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 , perhaps the easiest is the following. Suppose is a simple graph with vertices labeled by the first integers. The Laplacian of , denoted , is the matrix whose entry is given by
The reduced Laplacian of , denoted , is the matrix given by deleting the final row and column from . We define the sandpile group of , denoted , to be the cokernel of the reduced Laplacian of . We remark here that the sandpile group does not depend on the row or column removed from —the cokernels of the matrices given by deleting any row or column of are all isomorphic (see [10, 18]). From this definition, it follows immediately that if is connected, then ; Kirchhoff’s matrix tree theorem implies that is the number of spanning trees of . Alternatively, if denotes the subspace of vectors whose coordinates sum to 0, then it is equivalent to define as .
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 , 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 , toppling whenever they can, the sandpile group of 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 induced by the pairing on can be described directly using the Laplacian of the graph. In the Riemann surface analogy, the pairing on 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., ). 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 and a fixed finite abelian group , a priori we expect to be inversely proportional to .
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 preserving the pairing take the place of .
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 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 -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 -adic integral entries, drawn with respect to Haar measure on , 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 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) -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 be a real number, and let be a prime. Let be a symmetric random matrix in with independent entries for . Suppose also that for any the probability . Let denote the column space of in , let denote the Sylow -subgroup of , and let denote the canonical duality pairing on the torsion subgroup of (see Section 3). For any finite abelian -group equipped with a duality pairing , we have
where denotes the restriction of to .
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 -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 [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 -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 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 , the pairing agrees with the duality pairing on . 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 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 , we first translate the event that the pairing on pushes forward onto the pairing on into a piece of algebraic data which is easier to work with. In particular, if is a surjection, the event that takes the pairing on to the pairing on occurs if and only if is equal to some fixed symmetric homomorphism from . Then since we are trying to compute the average number of surjections from satisfying this property, and since the sandpile group is , it suffices to determine the probability that a surjection descends to and pushes the pairing on forward to the pairing on . Equivalently, we determine the probability that and that is equal to some fixed symmetric element of .
We regard this as a system of linear equations in the entries of , and we note that there are on the order of equations in variables. While in [54] it was enough to check whether , the pairing data makes this system particularly complex not only because the equations are more complicated, but because the equations given by are not even well-defined unless . To work around this inherent dependence between our equations, we choose a lift of to a larger group and modify our equations slightly so that we may work exclusively in this new group. Denote the lift by . Although this construction seems artificial, it turns out that all of our arguments depend only on itself rather than the specific choice of lift.
Our system of equations is parameterized by . Some pairs , give equations in which all of the coefficients are 0; we refer to these as special pairs. The set of special pairs depends implicitly on and the chosen lift . Ultimately, we are interested in the structural properties of and 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 for each . The special 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 depend on the off-diagonal ones, we run the program described above for a matrix with independent diagonal entries. Then we enlarge 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 to the bigger group and sum over all possible extensions.
1.5. Outline of the paper
In Section 3, for a symmetric integral matrix , we define a symmetric pairing on and show that this induces a duality pairing on the torsion subgroup of . The induced duality pairing is equal to the duality pairing on the torsion part of 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 ; Theorem 1.1 follows from Corollary 10.5.
2. Background
2.1. Finite abelian groups
Let be a prime. A finite abelian -group is isomorphic to
for some positive integers . If is the partition given by the ’s, we call the type of our finite abelian -group. When is understood, let denote the -group of type .
We may view any two abelian groups and as -modules, and we may consider their tensor product . When is finite and ,
where denotes the greatest common divisor of and . The fact that tensor products distribute over direct sums in conjunction with the classification of finite abelian groups allows us to compute for arbitrary finite abelian groups and . In particular, if denotes the Sylow -subgroup of , then Moreover, if is of type and is generated by with relations , then is generated as an abelian group by with relations . Thus,
Now, comes equipped with a natural involution taking . Denote the subgroup of elements fixed under this involution by . For as above, we have that and that is the abelian group generated by the elements and for with relations for each . Thus, .
The exterior power is the quotient of by the subgroup generated by elements of the form . For as above, we have that and that is generated as an abelian group by for with relations for each . Thus,
Likewise, the symmetric power is the quotient of by the subgroup generated by elements of the form . For as above, we have that and that is generated as an abelian group by for with relations for . Therefore,
The exponent of a finite abelian group is the smallest positive integer such that . If , then a finite abelian group with exponent dividing is also an -module. If is another such group, then (i.e., the -module homomorphisms are the same as the group homomorphisms ).
For any finite abelian group , its Pontryagin dual is the group , which we denote by .
2.2. Pairings
Let be an abelian group and consider a map . We say that is symmetric if and bilinear if and . If is bilinear, then it is called a pairing on . Equivalently, a pairing on can be thought of as a homomorphism . Given a pairing and generators of , note that is entirely determined by the values for all pairs .
A symmetric pairing on induces a map from taking (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 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 , denoted , is defined to be . Note that there is a natural nondegenerate symmetric pairing on induced by .
If the map induced by 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 is finite, then a nondegenerate pairing on is automatically perfect (this fact follows from counting). Given two groups and equipped with pairings and , there is a notion of isomorphism between and : an isomorphism between and is an isomorphism of groups such that for all pairs . If is a group equipped with a pairing, an automorphism of is said to respect the pairing if it is an isomorphism of groups and pairings between and itself. Denote the set of automorphisms of that respect the pairing by .
2.3. Duality
Let . When the ring is understood, for an -module , we may consider its dual, . If is a finite abelian group, there is a canonical isomorphism between and . For finite -modules and , we have . While and are not naturally isomorphic, when is finite, there is a canonical isomorphism between and given by evaluation.
For a finite abelian group , the following are equivalent data:
-
(1)
a perfect symmetric pairing on ;
-
(2)
an isomorphism from to ;
-
(3)
a perfect symmetric pairing on .
Given a duality pairing on , let denote the induced isomorphism from . Then the duality pairing on is given by .
Let and be two finite abelian groups. Let be a group homomorphism. The transpose of is the map given by precomposing with (a similar definition can be made if and are -modules for and an -module homomorphism). If is equipped with a pairing , then pushes forward to a pairing on given by
In other words, induces a map from ; let denote the image of under this map (denoted above). Note that we do not require to be perfect.
Let and be groups whose duals are equipped with pairings. Define
to be the set of group homomorphisms such that . In the category defined in Section 1, is the set of morphisms between and . When the pairings on both groups are understood, we denote by 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 and with duality pairings and and induced duality pairings and on the dual groups, note that if and only if .
2.4. Cokernels of matrices
Let denote the space of matrices with entries in a commutative ring . For , let denote the column space of (this is the image of the map given by ). Define the cokernel of to be
For , regard as a linear map . Then is a finitely generated abelian group; let denote its torsion subgroup. For any subgroup , let denote its orthogonal with respect to the standard scalar product on .
If is symmetric, . Because and because is a saturated submodule of (i.e., if for and , then ), it follows that
Since is pure torsion and is finitely generated, is finite.
2.5. Sandpile groups
Recall the definition of the graph Laplacian. Let , and let be a graph on vertices labeled by . The Laplacian of , denoted , is the matrix whose entry is given by
Note that is a symmetric matrix in . If is connected, then has rank and kernel spanned by the all-ones vector (see [51]). If is the subgroup of elements whose coordinates sum to 0, then we see that . Define the sandpile group of , denoted , to be . We see immediately that must be finitely generated. Moreover, it is finite if and only if is connected.
2.6. Random graphs
Let denote that is an Erdős-Rényi random graph on labeled vertices with edges occurring independently with probability .
2.7. Notation
We use 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, denotes “is isomorphic to,” and and denote probability and expected value, respectively. Also, will always denote a prime. For any integers and , let be the Kronecker delta—if , then and if , then .
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 carries a canonical duality pairing. If , and are lifts of respectively, then the pairing is given by
In [13], Bosch and Lorenzini showed that the torsion subgroup of 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 in the following way (we refer the reader to Section 1 of [13] for a more general and detailed setup). Let , and let and be lifts of and to , respectively. Since is pure torsion, there exist such that , i.e., there exist such that and . Define a map by
If is nonsingular, then , and this map is the classical pairing on discussed above. The map is a duality pairing on :
Lemma 3.1 (Lemma 1.1 and Theorem 1.3 of [13]).
The map is well-defined, bilinear, symmetric, and perfect.
When is singular, has some free part to which does not naturally extend. The duality pairing on implies the existence of a corresponding duality pairing on . Now, is a quotient of , so the pairing on does induce a natural pairing on . However, it can be quite difficult to compute this induced pairing on using the definition given above. So, we will define a symmetric pairing on given by ; show that this pairing induces a perfect symmetric pairing on ; and prove that the induced duality pairing on coincides with the pairing of Bosch and Lorenzini.
Recall is a linear map . Any can be extended to via the quotient . Note that is the subset of maps such that (i.e., for any lift of to , we have ). For , let be a lift of to . Then for , we have . We define a pairing on given by the matrix in the following way. For , let be lifts of , respectively. We define a map by
Lemma 3.2.
The map on is well-defined, bilinear, and symmetric.
Proof.
To show that is well-defined, we simply need to check that our map does not depend on the chosen lifts of and . Recall that both and are vectors in . Hence, for , we have
Hence, the map is well-defined.
Bilinearity and symmetry are both easily verified, where the symmetry of follows from the symmetry of itself. ∎
Since the Pontryagin duality functor is exact, the inclusion becomes a quotient after taking the transpose. The kernel of this quotient is the annihilator of , denoted . It turns out that descends to a well-defined duality pairing on . Recall the definition of from Section 2.2.
Lemma 3.3.
Let be a symmetric integral matrix and the symmetric pairing on . Then , implying .
Proof.
Suppose so that for all we have . Let and choose a lift of to . Since , there exist and such that . Consider the image of in , which we denote by . Since , we may view as an element of . Thus, if is any lift of to , then
forcing
For the reverse inclusion, assume so that for all , we have . We would like to show that for all
where and are lifts of and to , respectively. Suppose so that . Writing for some and , we see that . Hence, the image of in lies in . Since ,
where denotes the image of in . Therefore, . This forces , as desired. ∎
By abuse of notation, we also use to denote the duality pairing on .
Corollary 3.4.
The pairing is a duality pairing on .
Since is a finite abelian group, the duality pairing on induces a duality pairing on . The following shows that the duality pairing on induced by is Bosch and Lorenzini’s pairing on .
Lemma 3.5.
The duality pairing on induced by is .
Proof.
Recall that gives an isomorphism, taking to . Let be any lift of to (recall that there is a natural quotient ). If is any lift of to , then , where the brackets denote the image of in .
For , let be lifts of to . There exist and such that and . Thus, the preimages of and under are the images of and in . The pairing on induced by takes
However, recall that this is exactly . ∎
Later, given a finitely generated abelian group , we will be interested in taking the tensor product of with for some positive integer . If is equipped with a symmetric pairing , we would like the dual of to come naturally equipped with a symmetric pairing given by . Let . There is a natural quotient map given by , and therefore there is a natural inclusion . Thus, we obtain a pairing on the dual of by restricting the pairing on to . We abuse notation and also denote this pairing on by .
For a graph , recall from Section 2 the definition of its graph Laplacian and associated sandpile group . If is connected, in the language of the above, we see that is simply the torsion subgroup of the cokernel of the Laplacian, since (this is a well-known fact [9]; recall that is the subspace of vectors in whose coordinates sum to 1). Hence, using Bosch and Lorenzini’s construction, 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 for any graph . We claim that for any , not necessarily connected, is equipped with a natural symmetric pairing given by , which we will denote using . To see why, recall that we have the following chain of inclusions:
As a result, the canonical duality pairing on pushes forward to a symmetric pairing on , which itself pushes forward to the symmetric pairing on . Of these three pairings, we will be most interested in the symmetric pairing on .
The following lemma will be useful later.
Lemma 3.6.
Let be a graph on vertices and any finite abelian group whose dual is equipped with a symmetric pairing . Suppose is a surjection whose restriction to remains onto. Also assume that so that descends to a map with transpose . Then if and only if .
Proof.
We start with the following commutative diagram and its dual:
Recall that pushes the pairing on forward to some pairing on . Likewise, the transpose of the inclusion pushes the pairing forward to the pairing on . The commutativity of the diagram implies the result. ∎
4. Obtaining the moments I: Finding the equations
Let be a finite abelian group whose dual is equipped with a symmetric pairing . Suppose , and let denote its sandpile group and the Laplacian . Because is defined as a quotient of , any surjection lifts to a surjection . Therefore,
We will estimate the probabilities on the right-hand side of the above; to do so, we utilize the following more general setup.
Let denote the exponent of , and let . Let be the ring . This notation will be used through Section 8. Note that , i.e., whether only depends on the entries of modulo . We will see later in this section that whether also only depends on the entries of modulo . However, for reasons that will become apparent later, we will mostly work over (modulo ) in the following. In this and the following four sections, all of our linear algebra will be done over . However, since is not a domain, we are forced to work more abstractly rather than just with matrices.
Let be the finite abelian group fixed above with symmetric pairing on its dual . Let denote the Sylow -subgroup of , and write . For a prime , suppose is a -group of type , where is the partition given by . Because a pairing on may be viewed as an element of , any pairing on is entirely determined by its restriction to for each .
Let generate as an abelian group with and no other relations, and let be the elements of such that , where recall is the indicator of . Also, suppose that for , where is an integer . Note that the restriction of the pairing to is entirely determined by these values . Let denote the restriction of the pairing to .
Recall , and let . Suppose is a symmetric matrix in with entries for . We distinguish a basis of , say ; let denote the corresponding dual basis of , where . With respect to the bases and , we may view as a linear map where . Recall from Section 3 that comes equipped with a symmetric pairing given by , which we’ll denote by . Let be a surjective homomorphism and denote the transpose of by . Since is a quotient of , if , then factors through the quotient of by , i.e., descends to a map . Let for .
Lemma 4.1.
With notation as above, if and only if
(2) |
for and . If , then descends to a map , and we have if and only if
(3) |
for all (we will use to show that the left-hand side of (3) is well-defined modulo ).
Proof.
We first note that if and only if the composition is 0. Moreover, we see that if and only if
for all and .
Now, assume so that descends to a map . Recall from Section 3 that takes the pairing on to a pairing on and that a symmetric pairing on is entirely determined by where it sends the pairs for . Thus, we first compute for . We see that
Hence, with respect to the aforementioned bases, we may view as the row vector
It follows that
(recall we can do this computation by lifting to ). Therefore, given that , we see that if and only if
for all . ∎
Note that . Hence, given , Lemma 4.1 and the fact that imply whether and depends only on the entries of modulo . As before, let denote the -module with distinguished basis . Consider with dual basis , where . Note that there is an isomorphism between and given by sending to .
Recall is a -group of type . Define , and, noting that and depend implicitly on , set . For any , let denote the standard basis of , and define a surjective map by summing the maps taking for each . We will use this map throughout the argument—unless stated otherwise, when we refer to a surjection , we refer to this map.
Let , and let be a lift of to (with respect to the aforementioned surjection ).
Remark 4.2.
For the reader disturbed by this arbitrary choice of lift, we remark here that while there are many choices for , our arguments in the following section do not depend on this choice—the proofs only depend on the values of in after projecting to the quotient , i.e., the proofs only depend on itself. The lift solves the issue of whether the equations (3) from Lemma 4.1 are well-defined. By considering a lift of , we can instead work with equations that are always well-defined.
For example, consider the matrix for some mod . The cokernel of can be written as
and is isomorphic to . The pairing on is given by
and we see that depends on the entries of modulo rather than the entries of modulo .
Consider the Sylow -part of , i.e., . Let , where , and let , where . Since is a lift of , note that modulo .
Define to be the map sending for . Let be the direct sum of the ’s. Note that the image of is isomorphic to , giving us a copy of living in —i.e., the map given by is an isomorphism. Moreover, the surjection defined earlier is equal to the composition of with this isomorphism . Define to be the map taking for . Let be the sum of the ’s.
We define an element of that is equivalent to the data of the pairing . For each , recall that for all . Let denote the following symmetric element of , which we view as . Let
and let . We may think of as a way of expressing the pairing algebraically.
Now, note that for each , where , we have
Note here that in the above we are abusing notation slightly—both the ’s and implicitly depend on . We note that
(4) |
for and if and only if (2) holds. If we view as the subgroup of , then note that is simply the map defined earlier. Likewise, we see that
Note that is a map and is symmetric when viewed as an element of , since . Moreover,
(5) |
for if and only if (3) holds. Hence, and if and only if both (2) and (3) hold for all . It follows that
(6) |
While working with the lift seems artificial and notationally clumsy, we do so because it removes a dependence between our equations—we saw previously that the equations (3) for are not a priori well-defined. By considering a lift of , the equations (5) can be studied without first having to condition on (4). The maps and 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 , we have
where is any lift of to . 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 be a primitive th root of unity, and let denote the event
Recall that we may view as an element of and note that (where we view as a subgroup of ) and that , which, recall, is naturally isomorphic to . Since , it follows that . Thus, the Fourier expansion implies
Hence,
The pairs give the equations that a matrix must satisfy in order for to factor through the cokernel of the matrix and for to push the pairing on forward to the fixed pairing on . Viewing as a function of , we see that it is in fact a linear function of the entries of (for ). In the following, we will compute the coefficient of each . When the entries 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 we are referencing the copy of living in as . Because we are viewing as a subgroup of , we may view elements of as elements of , since a linear map restricts to a linear map . Further, any map yields a map via precomposition with . However, note that an element of cannot a priori be applied to an arbitrary element of .
Because (noncanonically), there is a natural isomorphism . Therefore, there is a natural isomorphism from , which is itself canonically isomorphic to . Hence, we regard as an element of . Let denote the evaluation map. Similarly, there is a natural isomorphism between and . Let denote the class of in , and recall that is generated by the elements of the form for along with the elements . The aforementioned isomorphism between and is witnessed by letting correspond to the map taking
Use this isomorphism to regard as an element of , and note that the evaluation does not depend on the chosen representative of in . Let denote the evaluation map described above. For each element of and choice of basis for , there exists some unique upper-triangular representative of in 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 . Therefore, occasionally we are forced to do some computations using this distinguished basis, which we denote by . At times, we will work with an alternate basis more conveniently aligned with or (through ). Because we are working over , 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 and rewriting it in more convenient notation. For , let be shorthand for the element . We see that
For , define
and we also define
Hence, when the entries (for ) are independent,
(7) |
For , we know from [54] that 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 to be nonzero. Thus, this section is devoted to pinpointing the characteristics of and that control the number of nonzero . Note there are total coefficients, and, to get the bounds we want, we would like quadratically many ’s to be nonzero. For certain “good” , we will show that there are three possible outcomes: on the order of coefficients are nonzero; on the order of coefficients are nonzero; all of the coefficients are zero. To show this, we identify the desired structural property of that affects the number of nonzero coefficients. We also show that there are relatively few such that only linearly many of the coefficients are nonzero; we count explicitly the for which all of the coefficients are zero. In the following, we will assume is “good” and define the aforementioned structural properties of . We will also explain the structural property which makes “good.” Later, in Section 7, we will consider the remaining .
Now, we will write the coefficients more equivariantly via a pairing. For elements and , define a map given by taking the direct sum of and . Likewise, define a map by switching the order of summation. There is a map
given by
Note that for all ,
Similarly, for and , we have a map given by taking the direct sum of and . Immediately, we see that for any submodule of ,
Given any subset , let denote the submodule of generated by the such that . In other words, is the submodule given by not using the coordinates in . The following definition gives the key structural property of the pair (for some fixed ) that determines if enough of the coefficients are nonzero.
Definition 5.1.
Let be a real number (which we will specify later). For a fixed , we say that the pair is robust for if every such that has the property that
Otherwise, the pair is said to be weak for .
We give an upper bound for the number of weak pairs .
Lemma 5.2 (Estimate for number of weak ).
Given , there is a constant such that for all the following holds. Let be defined as in Section 4, and let . The number of pairs that are weak for is at most
Proof.
Suppose is weak for . Then there exists some with such that
Because implies , we see that the above equality still holds for enlarged. Thus, without loss of generality, we may assume that . Now, since , it follows that is determined by for : if for and , then while . In other words, for each value in there is exactly one corresponding value in . Moreover, we have a well-defined homomorphism taking for all .
Recall that is determined by the values for in addition to . Therefore, given , to specify a such that is weak for for any , it suffices to choose a subset with , values for , and a homomorphism . Thus, there are choices for and choices for each and choices for , and then is determined. There are possibilities for . Note that because is determined by and because , we can find some constant such that . ∎
The following gives a sufficient condition for to be weak in terms of and .
Lemma 5.3.
Let , let , and let . Suppose is a submodule of such that . Then if is a submodule of such that for all and
then .
Proof.
Suppose for a contradiction that there exists some with but . Since , there exists some element such that . Because , we have , and it follows that there must be some such that . Let . Then
By hypothesis,
for any and , so we have a contradiction. ∎
Corollary 5.4.
Let , let , and let . Suppose is some submodule of such that . Let
Then . In particular, if , then is weak for .
Proof.
By assumption, for all and . Applying Lemma 5.3 with tells us that . ∎
Now, we define the important structural property of , which is a generalization of the notion of a linear code to -modules.
Definition 5.5.
For , we call a code of distance if for every such that , we have . Said another way, is not only a surjection, but would remain so if we threw out any fewer than standard basis vectors from .333We note that this definition is dual to the usual notion of a code. If is a field, then is a code if and only if the transpose is injective and its image is a code of distance in the usual sense.
The lift of a code to is still a code; importantly, note that the notion of being a code is a property we define for rather than for the lift of . However, the methods we use to bound the multiplicands in (7) require to be a code as well.
Lemma 5.6.
Let be a code of distance , and let be a lift of to . Then is a code of the same distance.
Proof.
Write both and as direct sums of their Sylow -subgroups, and note that any map is entirely determined by its restriction to for each prime . Also note that . Suppose is a -group of type . Since , where , we see that . Moreover, . Thus, it suffices to prove the lemma for a finite abelian -group.
Suppose is of type , where is the partition . Then . Recall from Section 4 that we have a surjection and that . Let be a set of cardinality less than . We have the following sequence of maps
where the first map is given by and the second by . The composite map is given by and is surjective since is a code of distance . Because and are -vector spaces of rank , it follows that the map must be surjective. By Nakayama’s lemma, it follows that must be surjective as well. Therefore, for any such . ∎
The following lemma will be useful often. In short, it allows us to choose a suitable basis of that works particularly nicely with the ’s in in tandem with and . We will use what it says about codes along with the property of robustness to obtain a good lower bound on the number of that are nonzero.
Lemma 5.7 (Lemma 3.4, [54]).
Let be a finite -module with Sylow -subgroup of type , where is a partition with parts. Suppose is a code of distance , and let . Then we can find and such that for every
and after the projection to the Sylow -subgroup of , the elements generate .
Lemma 5.8 (Quadratically many nonzero coefficients for robust ).
Let be a finite abelian group, and let be the set of primes dividing . If is a code of distance , and if is robust for , then there are at least pairs with such that .
Proof.
For each , let denote the Sylow -subgroup of , as usual. By taking the transpose of the surjection , there is an inclusion of into . Postcomposing with gives us an element . With this in mind, apply Lemma 5.7 to , , , and , and consider the resulting and . Let
(8) |
and let . Let be the submodule generated by the for so that in the projection to is the entirety of .
Next, let be the submodule of generated by the for all ; note that . Now, we apply Corollary 5.4 to . Since is robust for , we have
If pairs nontrivially with , then it must pair nontrivially with an element of one of the submodules generating , implying that
Hence, for some ,
For that particular , it follows that
By (8), for any , there are at least indices such that and . Moreover, note that for any . Thus, for all indices such that and , we have
For all , notice that
and also that
Hence, there must be at least pairs such that . ∎
We next study the number of nonzero coefficients for weak . To do so, we will consider the pairs such that all of the coefficients are 0 (e.g., is one such pair, but there are many others). We will first need to adopt a more equivariant view of the coefficients . The evaluation map gives rise to a natural map
and we may further compose with the quotient to get a map
For , composing with gives us a map .
Likewise, the evaluation induces a natural map
Recall that is naturally isomorphic to . Given , we get an element of by considering . Taking the sum of the aforementioned maps and (not the direct sum), given , we can construct a map
given by
We see that can be written explicitly as
(9) |
We remark that the definition of is canonical and basis independent. In particular, the above could be rewritten by replacing the ’s with the elements of any other basis of . Note that the elements in the kernel of are precisely the for which all of the coefficients are zero. We call the special for .
Recall that the data of the pairing on is equivalent to an element of whose precise definition is given in Section 4.
Lemma 5.9.
Given with , there are special for . Moreover, if is special for , then .
Proof.
Because everything in sight can be written as a direct sum of Sylow -subgroups, we can assume without loss of generality that is a -group of type (and, accordingly, that ). Suppose has parts, and recall that . Because is a surjection, there exist that can be completed to a free -module basis of such that and for (recall is the th standard basis vector in ). With respect to this basis, takes to
We compute the coefficients of each explicitly. For , the coefficient on is
Recall that has a unique upper-triangular representative with respect to the basis and that we may do all our computations with respect to this representative element . Since is upper-triangular with respect to the basis , the coefficient of for is
where denotes and denotes (recall that takes elements of to ). Note that , i.e., mod . For , the coefficient on is If , the coefficient on is
For to be special for , all of these coefficients must be 0 in . Setting all of these coefficients to 0 gives us the following system of equations
(10) |
modulo , which the entries of and satisfy if and only if is special. We count the pairs satisfying this system of equations. Recall that and that . Hence, the possible are given by choosing to be any multiple of in for each and (recall that mod ). Because is determined by its unique upper-triangular representative , the possible are given by choosing the entries for . The ’s determine , which determines . Since , the ’s can be any elements of .
Equation 10 determines for . In order for there to be solutions , we need mod . This holds because modulo . Thus, for , choose the ’s freely; there are choices for each . Equation 10 then determines the modulo for all . It follows that there are choices for each such that (10) is satisfied. Therefore, there are
special , as desired.
To prove the second part of the lemma, we again reduce to the case to where is a -group of type , where is the partition , and . Recall is the element
Then we may compute with respect to the unique upper-triangular representative of from above. By hypothesis, is special for . Hence, for , we have
and for , we have . Since mod , we have that mod for all . Thus,
where the second equality follows from the fact that is upper-triangular. ∎
Now, we will use Lemma 5.9 to prove that as long as is not special for there are linearly many nonzero coefficients .
Lemma 5.10 (Linearly many nonzero coefficients for non-special ).
Let be a code of distance , and suppose is not special for so that . Then there are at least pairs with such that .
Proof.
Suppose for a contradiction that there are fewer than pairs. Let be the set of pairs with and . By assumption, . Let be the set of all indices occurring in a pair . Immediately, we see that , and if is a pair with or not in , then must be 0.
We will use this setup to give a lower bound on the size of . Recall the definition of from (9): is sent to the element of given by
Compose with the quotient map taking if both , and call the resulting map , where denotes our new quotient space. Under , we see that (by hypothesis since it is not special). We will arrive at a contradiction by proving that divides . Then, since , this will force , which is a contradiction. By Lemma 5.9,
As in the proof of Lemma 5.9, we can prove divides by reducing to the case where is a -group of type , where is the partition . Accordingly, assume that .
Because is a code of distance , and because , we have that . Hence is a code of some positive distance. Apply Lemma 5.7 to to get such that and generate for . Let be the -submodule of generated by the for . Let be generators for with relations , and let denote the corresponding dual basis for so that . For , suppose is such that , and let be the submodule of generated by the . Now, we have maps
where the second map is given by . Each of the above spaces is an -vector space of rank at most , exactly , and exactly , respectively. Because the composite map is a surjection, our considerations on the rank of the spaces tell us that the map is a surjection. Thus, Nakayama’s lemma forces . The elements generate the free -module of rank , so they must form a basis for as a free -module. Thus, we have found a basis for such that .
Consider the following basis for , given by replacing the for with the ’s. Let . For , let , and for set . Because the ’s are a basis for , it follows that forms a basis for ; denote the corresponding dual basis by , where .
View as an element of and write
so that . Because was defined equivariantly, we see that is given by
which we may rewrite as
Hence,
Now, further quotient by the submodule generated by the such that neither nor are in . Call the resulting space and denote the composite by . Because and , it follows that . Therefore, if ,
(11) |
The element has order , implying that has a subgroup of order , as we can take any and .
Continuing, further quotient by the submodule generated by the such that one of or is not in . Denote the resulting space by and map by . Note that the only basis elements of whose images are nonzero in are those with both . Also, note that the subgroup of size we identified previously is sent to 0 under . Suppose and let be the map taking and all other . Consider the image of in , which we will denote using .
Suppose . We see that
(12) |
For all such that , set and . Then
Since is invertible in , this tells us that has a subgroup of size
Therefore, we conclude that . 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
when is a code and is a random matrix. If is a random symmetric matrix with integer entries and is an integer, we say is -balanced mod if for any prime divisor of and , the probability .
Lemma 6.1.
Let , let , and let be a finite abelian group. Let , , and be defined as in Section 4. Then there exists some with the following significance.
Let be a random symmetric matrix whose entries for are independent and -balanced mod . Suppose is a code of distance , and let be a lift of to . For all , we have that
To prove the above, we will need the following result due to Wood.
Lemma 6.2 (Lemma 4.2, [54]).
Suppose is a primitive th root of unity, and let be a random variable taking values in such that takes each value with probability at most . Then .
Proof of Lemma 6.1.
Recall that
where is a primitive th root of unity. Further, by Lemma 5.6, we know that is a code of distance , so we may apply our results from the previous section.
To prove the theorem, we break the sum into three pieces (we will choose later):
-
(1)
is special for ;
-
(2)
is not special and weak for ;
-
(3)
is robust for .
For a fixed , there are special pairs by Lemma 5.9. We may write
If is special for , recall from Lemma 5.9 that
Therefore, for special for ,
for any .
For the second case, we will rely on the fact that there are not too many weak pairs in conjunction with our bound from Lemma 5.10. Recall from Lemma 5.2 that for a given there are at most
weak . Now, factor the expected value as in (7) to get
Recall that we are assuming is not special for . Lemma 5.10 tells us that at least of the are nonzero. Hence, if is not special for , we conclude using Lemma 6.2 that
For the third case, we recall that given and robust for , Lemma 5.8 implies that we have at least pairs such that . Thus, if is robust for , then
Putting all three cases together, we get
Let be such that . Given , we can choose sufficiently small so that
Now, for fixed , the inequality
holds for sufficiently large. Hence, we have
for large enough. For the finitely many such that the above is false, just increase the constant 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 is not a code. To do so, we use a division of the based on the subgroups of from [54]. Let be an integer with prime factorization , and let . 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 is the maximal positive such that there exists with such that . If no such exists, the depth of is 1. (When applying this definition, we always assume .) If the depth of is 1, then note that is a code of distance : if has depth 1, then for every with we have (otherwise, ).
Note that if the depth of is , then . The following bounds the number of of depth .
Lemma 7.2 (Estimate for of depth ; Lemma 5.2 in [54]).
There is a constant depending on such that if , then the number of of depth is at most
When we are working with the Laplacian of a random graph, we also need a bound on the number of of depth , which is given by the following lemma from [54]. If is a finite abelian group with exponent , then so is . Thus, we may apply the definition of depth to a homomorphism .
Lemma 7.3 (Lemma 5.3 of [54]).
Let denote projection onto the second factor. Given a finite abelian -group of exponent , there is a constant , which depends on , such that if , then the number of of depth such that for all is at most
For each of depth , we use the following lemma to bound . In particular, because
and because the that are not codes only contribute to our error term, we can simply apply the bounds for due to Wood from [54].
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 -balanced random matrix from Section 6.
Theorem 8.1.
Suppose , and let be a finite abelian group such that is equipped with a symmetric pairing. For sufficiently small, there exists some , depending on , with the following significance. Let be a random symmetric integral matrix, -balanced mod , whose entries for are independent. Then
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 be the exponent of , and let . Reduce the matrix modulo so that our notation agrees with that of previous sections. Recall that
By Lemma 7.4 and Lemma 7.2, we have that
Lemma 7.2 also implies that
We can also show that
Finally, applying Lemma 6.1, we see that
putting this all together yields the desired result. ∎
Theorem 8.2.
Let , and let be a finite abelian group equipped with a symmetric pairing on its dual . Then there exist positive constants (depending on ) with the following significance. Suppose is a random graph and its sandpile group. Let be the Laplacian of , and let denote the canonical symmetric pairing on . Then for all we have that
Proof.
Suppose is the exponent of , and let . Let , and let . Let and , and recall that . Likewise, we let and be the reductions of the Laplacian of modulo and , respectively; we may think of and as matrices with coefficients in and , respectively. Recall from Section 3 that carries a symmetric pairing given by , which we denote by . Likewise and are both equipped with symmetric pairings given by restricting the pairing on to and , respectively.
Let be a random symmetric matrix with coefficients in . Suppose the modulo are distributed as for , and suppose the ’s are distributed uniformly in , with all () and independent. Set . Recall that we may view as a linear map , and let be our distinguished basis of (i.e., the basis of such that the strict upper-triangular entries of with respect to the bases and of and are independent). Let be the map that sends each basis vector (we see that is the row vector whose entries are the column sums of ). Note that and do not have the same distribution, since the column sums of can be anything and the column sums of are zero. To fix this, we’ll condition on so that the conditioned distribution of is the same as the distribution of . Let for each . We see that
so . Hence, given , we see that any choice of off-diagonal entries is equally likely in as in .
For , we have that
where and denote the symmetric pairings on and , respectively. Also,
Let be the sum of and .
Let denote the vectors whose coordinates sum to 0. In other words, let . Denote by the maps from that are surjections when restricted to . We would like to estimate
Now, we have
Given such that , recall from Lemma 3.6 that if and only if . In other words, for any such , we have that
and we can check whether the symmetric pairing on pushes forward to by checking whether the symmetric pairing on pushes forward to .
Now, let , and let denote the submodule of vectors whose coordinates sum to 0 modulo , i.e., , where we abuse notation and define to be the map taking each modulo . As before, set to be the set of maps that are surjections when restricted to . Note that because , it follows that . Therefore,
Note that if is a surjection when restricted to , then is a surjection from to . Write the above as
where by extensions of we really mean symmetric extensions of (we truncate for the sake of brevity). For the rest of the argument, when we refer to extensions of we mean symmetric extensions of this pairing. If generate , and if generates , then we see that symmetric extensions of to are entirely determined by where the pairs and are sent for all . Hence, in total, there are possible extensions of to .
We start by considering the part of the sum where is not a code of distance for some that we will choose later. We allow to change in each line, as long as it is a constant depending only on . Let . Then
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 , we can choose so small that
We also have that
where the last inequality holds because, for any , we may choose sufficiently small so that
Similarly, we also have that
where indicates that is a proper subgroup of .
Recall that the number of symmetric extensions of to is . Putting this all together, we have
Recalling that
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 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 is the category of finite abelian groups whose dual groups are equipped with symmetric pairings. For two objects , recall that a morphism between these objects in is a group homomorphism such that . To apply the moment methods of Sawin and Wood, we would like to be a diamond category.
Before we give the definition of a diamond category, we first collect some notation from [50]. Suppose is a small category. Given an object , a quotient of is the data of a object and an epimorphism . Two quotients and of are said to be isomorphic if there exists an isomorphism such that . We may define a partial order on the set of isomorphism classes quotients of an object , where if there is some compatible with the epimorphisms from . Given a quotient of , let be the partially ordered set of (isomorphism classes of) quotients of that are greater than or equal to . 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 and in the lattice have a least upper bound, or join, and a greatest lower bound, or meet. Denote the join of and (respectively, meet) by (respectively, ). We say a lattice is modular if implies for all in the lattice.
For a set of objects of , stable under isomorphism, is said to be downward-closed if every quotient of an object is also in . We say that is join-closed if for every , given two quotients and of such that exists (in the poset of isomorphism classes of quotients of ), we have implies . Given a set of isomorphism classes of objects in , the set generated by is the smallest downward-closed and join-closed subset of containing and every minimal object of . A level of is a set of isomorphism classes of objects of generated by a finite set. An epimorphism is called simple if the interval contains only and .
Definition 9.1.
A diamond category is a small category satisfying the following three properties.
-
(1)
For each , the set of isomorphism classes of quotients of form a finite modular lattice.
-
(2)
Each object in has finitely many automorphisms.
-
(3)
For each level of and each object in the level , there are only finitely many elements of with a simple epimorphism to .
To see that is a diamond category, we will apply the following lemma from [50].
Lemma 9.2 (Lemma 6.22 of [50]).
Let be a diamond category, and let be a functor from to the category of finite sets. Then the category of pairs of an object together with an element , with morphisms given by morphisms in such that , is a diamond category.
That is a diamond category follows from applying Lemma 9.2 to the category of finite abelian groups and the functor taking each group to the set of symmetric pairings on its dual group. Let denote the category of finite abelian groups. The functor takes a group homomorphism to the map of sets
given by
We note that the category is precisely the enriched category of pairs 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 be a diamond category, and let be a level in . For each , let be a real number such that is well-behaved at .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 and are two sequences of measures on such that for each , we have
then and exist and are equal for any .
Note that if we set in the above, then for . Before showing that the -moments computed in Section 8 are well-behaved, we must first better understand the levels of via the following lemma.
Lemma 9.4.
Let be the level of generated by the set of objects whose underlying group is for and whose pairing is any symmetric pairing on . Then is the following set of isomorphism classes of objects: finite abelian groups whose exponent divides with all possible symmetric pairings on their dual groups.
Proof.
We begin by noting that, given two groups , both and are quotients of (we have , for example). Moreover, is the join of and in the lattice of quotients of .
Now, given any two objects , let be a pairing on whose restriction to is and whose restriction to is . For any such pairing , we claim that is the join of and in the lattice of quotients of . This follows from our remark in the previous paragraph and the fact that a quotient of is a quotient (of groups) such that the transpose pushes the pairing forward to the pairing .
Let be a level of , and suppose for any symmetric pairings and on and , respectively. Given any symmetric pairing on , we note that is simply the join of and , where denotes pairing on given by the restriction of to (similarly for ). Therefore, .
Since any quotient of for has exponent dividing , and since the join of with any other such group (in ) has exponent dividing , we see that only contains objects whose underlying group has exponent dividing . Now, any group whose exponent divides can be written as
where each . In other words, is the join (as groups) of the . By the above, the join-closedness of forces to contain equipped with any possible symmetric pairing on its dual, since is generated by with any symmetric pairing on its dual, where is any divisor of . ∎
Ultimately, we would like to apply Lemma 9.3 to the category at the level for an arbitrary positive integer . Let be the level in generated by , i.e., the level of finite abelian groups of exponent dividing . Note that is equal to the level of generated by for . To prove that the titular sequence of -moments computed in Section 8 are well-behaved, we reduce the question of well-behavedness at to one of well-behavedness at in using the following lemma.
Lemma 9.5 (Proof of Lemma 6.23 of [50]).
Let be a diamond category and a functor from to the category of finite sets. Let be a level of the category of pairs, , and let be the level of generated by the underlying objects of a finite set of generators of . Let be a set of moments. Then is well-behaved at if is well-behaved at .
When applied to , Lemma 9.5 implies the following. For a sequence of moments indexed by elements of , if the sequence of moments indexed by given by
is well-behaved at for , then the sequence is well-behaved at .
Recall that for an object , Theorem 8.2 tells us that the -moment of the sandpile group of a random graph converges to as , where is the number of vertices in the graph. Hence, we have reduced the problem to showing that the sequence in whose -entry is given by
is well-behaved.
For any prime , let denote the category of finite -modules, i.e., the category of finite abelian -groups. For any set of primes , let be the product of categories . If is the set of all primes, note that .
Lemma 9.6 (Lemma 6.16 of [50]).
Let and be diamond categories. Then the product category , whose objects are ordered pairs of an object from and an object from and whose morphisms are ordered pairs of morphisms, is a diamond category.
Lemma 9.7 (Lemma 6.17 of [50]).
Let and be diamond categories. Let and be tuples of nonnegative real numbers indexed by the respective isomorphism classes of and . Define by the formula . If and are both well-behaved for and , then is well-behaved for the product category .
Let be a finite set of primes. For each , consider a sequence of moments in , and form a sequence of moments in by taking all products of the moments in the component categories . By Lemma 9.7, if each of the component sequences of moments are well-behaved, so is this sequence of moments in . Therefore, for the sequence of moments , it suffices to check that the sequence is well-behaved for each , since
To do this, we simply apply the following lemma from [50] to the category of finite -modules.
Lemma 9.8 (Lemma 6.9 of [50]).
Let be a discrete valuation ring with finite residue field. For each isomorphism class of finite -modules, let be a real number. Then is well-behaved if there exists some and such that for all finite -modules .
Applying Lemma 9.8 to , the category of finite -modules, it follows that the moments in given by for each are well-behaved. Hence, the corresponding moments in are also well-behaved. To conclude the section, we repackage all of our above work in the following explicit theorem.
Theorem 9.9.
Let and be sequences of random variables taking values in the category of finitely generated abelian groups whose duals are equipped with symmetric pairings. Let be a positive integer and the set of isomorphism classes of finite abelian groups of exponent dividing with symmetric pairings on their duals. Assume that for every ,
and that the analogous statement holds for . Then for every the limits and exist, and
Proof.
Recall that is the level of generated by for with every possible symmetric pairing on its dual. We have a sequence of -moments given by for , since for each we have
By Lemma 9.5 and our previous discussion, to show that the moments are well-behaved at , it suffices to check well-behavedness at the level of (recall that is the level of all finite abelian groups of exponent dividing ).
Let be the finite set of primes dividing , and consider . The level in generated by is exactly the same as , so we may check well-behavedness at this level in . By Lemma 9.7 and our work in the above, we see that the sequence of moments given by for is well-behaved at , implying the sequence of moments is well-behaved at . 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 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 be a finite abelian -group with a duality pairing on , and let be a random symmetric matrix whose entries are distributed with respect to additive Haar measure on . Then the asymptotic probability (as ) that the cokernel of with its duality pairing is isomorphic to is
By extending the measure in the above result by 0, we can extend the above to give probabilities for when the pairing on is any symmetric pairing. Notably, we do not require the pairing to be perfect.
Corollary 10.2 (of Theorem 9.9).
Let be a finite abelian group of exponent dividing , and suppose has a symmetric pairing on its dual , which we will denote using . Let , let be a random graph, and let be its sandpile group. Denote the canonical symmetric pairing on by . Also let be a uniform random symmetric matrix with entries in . Then
In particular, if is perfect, then
Otherwise, .
Proof.
That
follows from applying Theorem 9.9. In particular, we use the fact that both and have the same asymptotic -moment for any object (this is a consequence of Theorem 8.2).
The second part of the lemma follows from noting that everything factors over , allowing us to reduce to the case where is a -group. ∎
To prove Theorem 1.1, we will turn Corollary 10.2 into a statement about the Sylow -subgroups of the sandpile group. In particular, we convert information about for any into information about the Sylow -subgroups of for , and we translate statements about pairings on the dual of to statements about pairings on the Sylow subgroups of .
Remark 10.3.
First, we will need to know that is finite with asymptotic probability 1 as . While this result is well-known, we offer an alternate proof below. Recall that is finite if and only if is connected; is connected with probability 1 as for fixed (this is a corollary of results proved by Erdős and Rényi in their 1960 paper [23]). Hence, as , we know that 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 in Lemma 10.4 by , where is a random matrix satisfying the hypotheses of Theorem 8.1, we get an analogous result for . The result for 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 be the sandpile group of a random graph on vertices as in Theorem 8.2. As , we have that .
Proof.
Let be any finite set of primes, and let . By Corollary 9.2 of [54], we have
for any finite abelian -group .555The proof of Corollary 9.2 of [54] works for our purposes, but note that it contains a slight mistake—throughout, “Sylow -subgroup of ” should be replaced by . Proposition 7 of [15] tells us that the sum over all finite abelian -groups of the right-hand side of the above is 1. By Fatou’s lemma, it follows that
as desired. ∎
Corollary 10.5.
Let be a finite abelian group equipped with a duality pairing . Suppose is a random graph with Laplacian and sandpile group . Denote the canonical duality pairing on by . Suppose is the finite set of prime divisors of . Let denote the sum of the Sylow -subgroups of for and the restriction of to . Then
Proof.
Write the exponent of as , and let . Let be the pairing on induced by . Recall that , so is also perfect. Also recall that the dual of comes equipped with a symmetric pairing, denoted , given by pushing the pairing on forward by the transpose of the natural quotient . We have the following commutative diagram:
where the vertical maps are isomorphisms induced by the duality pairing .
Assume is finite. We will show that
To prove this, first note that finite implies that if and only if . Thus, it is sufficient to show that the data of the pairing on the dual of is equivalent to the data of the pairing on . Suppose . This forces the composition of maps in the first row of the above diagram to be an isomorphism between and . The isomorphism gives us a duality pairing on , which in turn gives us a duality pairing on . We will prove that this induced pairing on is just .
Recall from Lemma 3.5 that the pairing on is induced by the duality pairing on . Moreover, recall that we may view the pairing on as the restriction of to . Hence, to compute on , we may look at the corresponding elements in and do the computation using . Since the elements of all have order dividing , their images in are -torsion. In other words, the image of in lies in . It follows that is exactly the pairing on induced by , proving that .
By Remark 10.3, we know that is asymptotically almost surely finite. Therefore, as , we have that
which, in conjunction with Corollary 10.2, implies the result. ∎
Remark 10.6.
Because is finite with probability 1 as , in the limit, the pairing is simply the canonical duality pairing on . Hence, Corollary 10.5 can actually be viewed as a statement about the distribution of the Sylow -part of the sandpile group and its canonical duality pairing (more accurately, the restriction of this pairing to the Sylow -subgroup).
Moreover, all of the results in Section 10 hold if 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 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 . 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 -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 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.