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

Time-Inhomogeneous Random Walks on Finite Groups and Cokernels of Random Integer Block Matrices

Elia Gorokhovsky Harvard University, Cambridge, USA
Abstract.

We study time-inhomogeneous random walks on finite groups in the case where each random walk step need not be supported on a generating set of the group. When the supports of the random walk steps satisfy a natural condition involving normal subgroups of quotients of the group, we show that the random walk converges to the uniform distribution on the group and give bounds for the convergence rate using spectral properties of the random walk steps. As an application, we use the moment method of Wood to prove a universality theorem for cokernels of random integer matrices allowing some dependence between entries.

1. Introduction

The work in this paper is motivated by a question in the theory of integer random matrices but is of independent interest to the study of random walks on groups.

Random walks on finite groups are well-studied in the reversible, time-homogeneous, ergodic regime, where the random walk on a group GG consists of a product X1X2XnX_{1}X_{2}\dots X_{n} for i.i.d. XiX_{i} drawn from a distribution supported on a generating set of GG. Such random walks are known to converge to the uniform distribution π\pi on GG exponentially quickly. Namely, if we denote by νn\nu_{n} the distribution of X1X2XnX_{1}X_{2}\dots X_{n}, then

νnπL2σn,||\nu_{n}-\pi||_{L^{2}}\leq\sigma^{n},

where σ\sigma is the second-largest singular value of the Markov operator of the random walk and ||||L2||\cdot||_{L^{2}} denotes the L2L^{2} norm of signed measures as functions GG\to\mathbb{R}. See [Sal04] for an excellent review of these kinds of walks.

The above inequality comes from looking at norms of convolution operators on the space of signed measures on GG. If XX and YY are random elements of GG distributed according to μ\mu and ν\nu respectively, then XYXY is distributed according to the convolution

(μν)(g)=hGμ(h)ν(h1g).(\mu*\nu)(g)=\sum_{h\in G}\mu(h)\nu(h^{-1}g).

In particular, if the XiX_{i} are distributed according to μ\mu, then νn\nu_{n} is the nn-fold convolution μn=μμn times\mu^{*n}=\underbrace{\mu*\dots*\mu}_{n\text{ times}}.

Some of the assumptions can be relaxed; for instance, Saloff-Coste and Zúñiga [SZ07] studied convergence of time-inhomogeneous Markov chains, including random walks on finite groups, in the case where each step of the random walk is irreducible (in particular, supported on a generating set of GG). In that case, if we denote by σi\sigma_{i} the second-largest singular value of the iith step,

νnπL2i=1nσi.||\nu_{n}-\pi||_{L^{2}}\leq\prod_{i=1}^{n}\sigma_{i}.

The condition that each step of the random walk is supported on a generating set is crucial because if the subgroup generated by the supports of the steps is a proper subgroup of GG, the random walk will surely stay in that subgroup. Nevertheless, one may expect that if the supports of all steps taken together generate GG, the random walk might still equilibrate to the uniform distribution on GG.

A consequence of our first main result is the following theorem, which relaxes this “generating” assumption by extending part of [SZ07, Theorem 3.5] to some time-inhomogeneous random walks where the probability measures driving each step need not be irreducible:

Theorem 1.1.

Let GG be a finite group, and let μ1,μ2,,μn\mu_{1},\mu_{2},\dots,\mu_{n} be probability measures on GG. For each subgroup HH of GG, let IH={iH=suppμi}I_{H}=\{i\mid H=\langle\operatorname{supp}\mu_{i}\rangle\}. Let 𝒮\mathcal{S} be a finite set of normal subgroups of GG such that G=H𝒮HG=\left\langle\bigcup_{H\in\mathcal{S}}H\right\rangle. Write νn=μ1μn\nu_{n}=\mu_{1}*\dots*\mu_{n}. Also, for each ii, let σi\sigma_{i} be the second-largest singular value of μi*\mu_{i} as an operator on L2(suppμi)L^{2}(\langle\operatorname{supp}\mu_{i}\rangle). Let π\pi be the uniform distribution on GG.

If IHI_{H} is nonempty for each H𝒮H\in\mathcal{S}, we have

νnπL2H𝒮(iIHσi).||\nu_{n}-\pi||_{L^{2}}\leq\sum_{H\in\mathcal{S}}\left(\prod_{i\in I_{H}}\sigma_{i}\right).

We prove a more general version of this result in Theorem 3.1.

In particular, if a time-inhomogeneous random walk on a finite group has steps supported on enough subgroups, then it converges to the uniform distribution on the group with an exponential rate controlled by subgroups that appear infrequently or mix very slowly. Adding more probability measures to the convolution νn\nu_{n} may not improve the convergence rate, but it never makes the bound worse because convolution with a probability measure is non-expansive in the L2L^{2} norm. A nice consequence of this is that 𝒮\mathcal{S} need not be an exhaustive list of every normal subgroup for which IHI_{H} is nonempty.

The main difference between this result and [SZ07, Theorem 3.5] is that [SZ07] relaxes the time-homogeneity assumption for random walks but not the assumption that each step is supported on a generating set for the group. The new condition that the supports of the steps jointly generate GG is substantially weaker than the assumption that the support of each step generates GG.

The conditions of Theorem 1.1 can be weakened so that not all the subgroups HiH_{i} need to be normal (see Theorem 3.1), but see Example 3.3 for why some hypothesis on the subgroups is necessary.

Our main interest in developing this theorem is an application to limiting distributions of cokernels of random matrices. Wood [Woo19, Theorem 2.9] and Nguyen and Wood [NW22, Theorem 1.1] showed that cokernels of integer-valued random matrices approach a universal limiting distribution in the following sense. Let (Mn)n=1(M_{n})_{n=1}^{\infty} be a sequence with each MnM_{n} a random n×(n+u)n\times(n+u) integer matrix with independent entries (u0u\geq 0). Wood showed that, under very weak conditions on the distributions of the entries of the MnM_{n}, the distribution of the isomorphism class of the random group coker(Mn)n/Mn(n+u)\operatorname{coker}(M_{n})\coloneqq\mathbb{Z}^{n}/M_{n}(\mathbb{Z}^{n+u}) converges weakly (i.e., at any finite collection of primes) as nn\to\infty to the distribution λu\lambda_{u} on isomorphism classes of profinite abelian groups defined as follows: if AλuA\sim\lambda_{u} and BB is a finite abelian pp-group, then

[ApB]=1|B|u|Aut(B)|k=u+1(1pk)\mathbb{P}[A_{p}\cong B]=\frac{1}{|B|^{u}|\operatorname{Aut}(B)|}\prod_{k=u+1}^{\infty}(1-p^{-k})

independently for all primes pp, where ApA_{p} is the pp-part of AA. If further u1u\geq 1, then λu\lambda_{u} is supported on isomorphism classes of finite abelian groups, and for finite abelian BB we have

λu(B)=1|B|u|Aut(B)|k=u+1ζ(k)1,\lambda_{u}(B)=\frac{1}{|B|^{u}|\operatorname{Aut}(B)|}\prod_{k=u+1}^{\infty}\zeta(k)^{-1},

where ζ\zeta denotes the Riemann zeta function. Nguyen and Wood weakened the conditions on the entries and showed strong (pointwise) convergence to λu\lambda_{u}. The phenomenon that the limiting distribution of n/Mn(n+u)\mathbb{Z}^{n}/M_{n}(\mathbb{Z}^{n+u}) is rather insensitive to the distributions of the entries of MnM_{n} is an example of universality. The distributions λ0\lambda_{0} and λ1\lambda_{1} are known as the Cohen-Lenstra distributions, and are conjectured to describe the distributions of class groups of imaginary and real quadratic number fields, respectively.

In her 2022 ICM talk, Wood [Woo23, Open Problem 3.10] asks if the universality class of λu\lambda_{u} can be extended to cokernels of matrices with some dependent entries. There are a few specific results in this direction. Most recently, Nguyen and Wood [NW22, Theorem 1.1] show that the distribution λ1\lambda_{1} is universal for Laplacians of Erdős-Rényi random directed graphs. Mészáros [Més20] shows that λ0\lambda_{0} is universal for Laplacians of random regular directed graphs. Friedman and Washington [FW89] show that the cokernels of the random matrices IMI-M, where MM is drawn at random from the multiplicative Haar measure on GL2g(p)\operatorname{GL}_{2g}(\mathbb{Z}_{p}), approach the pp-part of λ0\lambda_{0} as gg\to\infty. However, when there is too much dependence in the entries of the random matrices, one gets different (but often related) limiting distributions, for example in the case of symmetric matrices ([Woo14]), Laplacians of random regular undirected graphs ([Més20]), products of independent random integral matrices ([NV24]), and quadratic polynomials in Haar-random matrices ([CK24]). It is natural to ask just how much (and what kind of) dependence is allowed between the entries of sequences of random matrices before their cokernels leave the universality class of λu\lambda_{u}.

The main application of Theorem 3.1 in this paper is a Theorem 1.2 below, which extends the result of [Woo19] to matrices with some dependence in their rows and columns. We introduce a regularity condition on matrices, (w,h,ε)(w,h,\varepsilon)-balanced, in Definitions 4.2 and 4.7. Generally, it means that the matrix can be written as a block matrix where the blocks have height at most hh, width at most ww, are all independent, and each satisfy some regularity condition depending on ε\varepsilon. The key detail is that the blocks of the matrix may have dependent entries, as long as there is no dependence between blocks. (The (w,h,ε)(w,h,\varepsilon)-balanced condition is invariant under permutation of rows and columns, so one can also think of a (w,h,ε)(w,h,\varepsilon)-balanced matrix as a block matrix which is at most hh blocks tall, at most ww blocks wide, and such that the entries of each block are independent of each other.) With this condition, we have:

Theorem 1.2.

Let u0u\geq 0 be an integer. Let (wn)n,(hn)n(w_{n})_{n},(h_{n})_{n} be sequences of real numbers such that wn=o(logn)w_{n}=o(\log n), hn=O(n1α)h_{n}=O(n^{1-\alpha}), and εnnβ\varepsilon_{n}\geq n^{-\beta} for some 0<α10<\alpha\leq 1 and 0<β<α/20<\beta<\alpha/2.

For each integer n0n\geq 0, let MnM_{n} be an (wn,hn,εn)(w_{n},h_{n},\varepsilon_{n})-balanced n×(n+u)n\times(n+u) random matrix with entries in \mathbb{Z}. Then the distribution of coker(Mn)\operatorname{coker}(M_{n}) converges weakly to λu\lambda_{u} as nn\to\infty. In other words, if YλuY\sim\lambda_{u}, then for every positive integer aa and every abelian group HH with exponent dividing aa we have

limn[coker(Mn)/aH]=[Y/aH].\lim_{n\to\infty}\mathbb{P}[\operatorname{coker}(M_{n})\otimes\mathbb{Z}/a\mathbb{Z}\cong H]=\mathbb{P}[Y\otimes\mathbb{Z}/a\mathbb{Z}\cong H].

The key idea of the proof of Theorem 1.2 uses the moment method developed in [Woo14] and [Woo19]. Understanding the cokernel of a random integer matrix reduces to finding the probability that each random column maps to zero under an arbitrary surjective group homomorphism f:nGf\colon\mathbb{Z}^{n}\to G for an arbitrary abelian group GG. To handle dependent columns, we treat several columns at a time and look at the induced surjection (n)mGm(\mathbb{Z}^{n})^{m}\to G^{m}. We view the image of a random element of (n)m(\mathbb{Z}^{n})^{m} as a random walk in GmG^{m} and apply Theorem 3.1 to approximate the distribution of this image. Since the surjection ff is arbitrary, we have very little control over the distribution of the steps of this walk. In particular, they are almost never supported on all of GmG^{m}, which is why we need Theorem 3.1 to handle random walk steps supported on proper subgroups. The (w,h,ε)(w,h,\varepsilon)-balanced condition allows us to bound the singular values of the associated convolution operators and get quantitative bounds on the error in terms of ww, hh, and ε\varepsilon.

There is a considerable body of literature pertaining to random matrices with complex entries, with analogous universality results about distributions of eigenvalues. If {Mn}\{M_{n}\} is a sequence of n×nn\times n random complex matrices whose entries are independent, with appropriately normalized mean and variance, the empirical distribution of the eigenvalues of MnM_{n} converges to the circular law, which is the uniform distribution on the unit disc in \mathbb{C} [TVK10]. The universality of the circular law for spectra of a wide class of random complex block matrices was proved by Nguyen and O’Rourke in [NO15].

2. Notation and Terminology

For a finite set SS and p>0p>0, we use L2(S)L^{2}(S) to denote the space of signed measures (equivalently, real-valued functions) on SS, equipped with the norm fL2(S)2=sS|f(s)|2||f||_{L^{2}(S)}^{2}=\sum_{s\in S}|f(s)|^{2}. When the set SS is implicit, we write fL2||f||_{L^{2}} for fL2(S)||f||_{L^{2}(S)}. Any set map f:STf\colon S\to T defines a pushforward map f:L2(S)L2(T)f_{*}\colon L^{2}(S)\to L^{2}(T) by fμ(t)=μ(f1(t))f_{*}\mu(t)=\mu(f^{-1}(t)). We say a signed measure ν\nu is uniform on TST\subset S if ν(t)=ν(t)\nu(t)=\nu(t^{\prime}) for t,tTt,t^{\prime}\in T. For a point fnf\in\mathbb{R}^{n} with the Euclidean metric and a linear subspace WnW\subset\mathbb{R}^{n}, we write dL2(f,W)d_{L^{2}}(f,W) for the distance between ff and its orthogonal projection onto WW. Note that this is equal to infgW|fg|\inf_{g\in W}|f-g|. If GG is a finite group, any signed measure μ\mu defines a linear convolution operator (or, if μ\mu is a probability measure, Markov operator) μ*\mu on L2(G)L^{2}(G) given by ννμ\nu\mapsto\nu*\mu. When we discuss the second-largest singular value of an operator, we are counting with multiplicity; for example, if the singular values of MM are 1,1,01,1,0, then its second-largest singular value is 11.

For two finite or profinite groups G,GG,G^{\prime}, we write Hom(G,G)\operatorname{Hom}(G,G^{\prime}) for the set of (continuous) group homomorphisms from GG to GG^{\prime} and Sur(G,G)\operatorname{Sur}(G,G^{\prime}) for the set of (continuous) surjective group homomorphisms from GG to GG^{\prime}. For a subset SGS\subseteq G, we denote by S\langle S\rangle the (closed) subgroup of GG generated by SS. We refer to the identity element of a group as ee.

A probability measure or distribution is a measure with total mass 1 (not signed). The uniform distribution on GG is usually denoted π\pi and is the measure on GG with π(g)=1/|G|\pi(g)=1/|G| for gGg\in G. We use []\mathbb{P}[\cdot] for probability and 𝔼[]\mathbb{E}[\cdot] for expectation. We denote by suppμ\operatorname{supp}\mu the support of a measure μ\mu. If a random variable XX has law μ\mu, we write XμX\sim\mu.

For a positive integer nn, we write [n][n] for the set {1,,n}\{1,\dots,n\}.

3. Random Walks

This section is devoted to proving the following stronger version of Theorem 1.1 from the introduction:

Theorem 3.1.

Let GG be a finite group and suppose we have a sequence of surjective homomorphisms

G=G0\xtwoheadrightarrowQ1G1\xtwoheadrightarrowQ2G2\xtwoheadrightarrowQ3\xtwoheadrightarrowQk1Gk1\xtwoheadrightarrowQkGk={e}.G=G_{0}\xtwoheadrightarrow{Q_{1}}G_{1}\xtwoheadrightarrow{Q_{2}}G_{2}\xtwoheadrightarrow{Q_{3}}\dots\xtwoheadrightarrow{Q_{k-1}}G_{k-1}\xtwoheadrightarrow{Q_{k}}G_{k}=\{e\}.

For 0jk0\leq j\leq k, define Q~j:GGj\tilde{Q}_{j}\colon G\twoheadrightarrow G_{j} by Q~j=QjQj1Q1\tilde{Q}_{j}=Q_{j}\circ Q_{j-1}\circ\dots\circ Q_{1} (so Q~0=idG\tilde{Q}_{0}=\operatorname{id}_{G}), and for 1jk1\leq j\leq k define HjGj1H_{j}\trianglelefteq G_{j-1} by Hj=kerQjH_{j}=\ker Q_{j}.

Let μ1,,μn\mu_{1},\dots,\mu_{n} be probability measures on GG. Let νn=μ1μn\nu_{n}=\mu_{1}*\dots*\mu_{n}. For each j=1,,kj=1,\dots,k, let Ij={isupp(Q~j1)μi=Hj}I_{j}=\{i\mid\langle\operatorname{supp}(\tilde{Q}_{j-1})_{*}\mu_{i}\rangle=H_{j}\}. Let π\pi be the uniform distribution on GG.

For iIji\in I_{j}, let σi\sigma_{i} be the second largest singular value of the (Q~j1)μi(\tilde{Q}_{j-1})_{*}\mu_{i}-random walk on HjH_{j}. If each IjI_{j} is nonempty, we have

νnπL22j=1k|Gj1|1|G|(iIjσi2)=j=1ki=jk|Hi|1|G|(iIjσi2).||\nu_{n}-\pi||_{L^{2}}^{2}\leq\sum_{j=1}^{k}\frac{|G_{j-1}|-1}{|G|}\left(\prod_{i\in I_{j}}\sigma_{i}^{2}\right)=\sum_{j=1}^{k}\frac{\prod_{i=j}^{k}|H_{i}|-1}{|G|}\left(\prod_{i\in I_{j}}\sigma_{i}^{2}\right).

In the case where k=1k=1 and H1=GH_{1}=G, we recover the first part of [SZ07, Theorem 3.5]. We postpone the proof that Theorem 3.1 implies Theorem 1.1 until the end of this section.

The following example gives a case that is covered by Theorem 3.1 but not by Theorem 1.1:

Example 3.2.

Consider the dihedral group G=D2n=r,srn=s2=(rs)2=eG=D_{2n}=\langle r,s\mid r^{n}=s^{2}=(rs)^{2}=e\rangle with n>2n>2. The subgroup H1=rH_{1}=\langle r\rangle is normal, but the subgroup H~2=s\tilde{H}_{2}=\langle s\rangle is not normal. We have G1=D2n/r=/2G_{1}=D_{2n}/\langle r\rangle=\mathbb{Z}/2\mathbb{Z} generated by the image of ss, so the image of s\langle s\rangle in the quotient is normal. Let Q:G/2Q\colon G\twoheadrightarrow\mathbb{Z}/2\mathbb{Z} be the quotient map. Let μ\mu be a measure on r\langle r\rangle with second-largest singular value σ\sigma. Consider the following random walk on D8D_{8}:

  • For ii odd, μi=μ\mu_{i}=\mu.

  • For ii even, μi(s)=p\mu_{i}(s)=p and μi(e)=1p\mu_{i}(e)=1-p.

Say ii is even. In matrix form the operator (Qμi)*(Q_{*}\mu_{i}) on L2(/2)2L^{2}(\mathbb{Z}/2\mathbb{Z})\cong\mathbb{R}^{2} is given by (1ppp1p)\begin{pmatrix}1-p&p\\ p&1-p\end{pmatrix}. It is symmetric, so the singular values are just the absolute values of the eigenvalues. The vector (11)\begin{pmatrix}1\\ -1\end{pmatrix} is a (12p)(1-2p)-eigenvector. Thus, the singular values of μi*\mu_{i} are 1 and |12p||1-2p|, so σi=|12p|\sigma_{i}=|1-2p|. Theorem 3.1 says that

μ1μ2kπL22σk+|12p|k.||\mu_{1}*\dots*\mu_{2k}-\pi||_{L^{2}}^{2}\leq\sigma^{k}+|1-2p|^{k}.

In particular, if p=1/2p=1/2, the random walk mixes on D2nD_{2n} half as fast as the μ\mu-random walk mixes on r/n\langle r\rangle\cong\mathbb{Z}/n\mathbb{Z}.

Although Theorem 3.1 makes weaker assumptions on the subgroups than Theorem 1.1, it is not possible to fully remove the normality assumption, as the following example shows:

Example 3.3.

Consider the alternating group A5A_{5}. Recall that A5A_{5} is generated by the 3-cycles (1 2 3),(1 2 4),(1 2 5)(1\;2\;3),(1\;2\;4),(1\;2\;5). Consider the following three-step time-inhomogeneous “random walk” on A5A_{5}: X1X_{1} is uniformly distributed on (1 2 3)\langle(1\;2\;3)\rangle, X2X_{2} is uniformly distributed on (1 2 4)\langle(1\;2\;4)\rangle, and X3X_{3} is uniformly distributed on (1 2 5)\langle(1\;2\;5)\rangle. The step distributions μ1,μ2,μ3\mu_{1},\mu_{2},\mu_{3} on the respective cyclic groups all have second-largest singular value zero. However, the product X1X2X3X_{1}X_{2}X_{3} is not uniformly distributed on A5A_{5}. Indeed, when X1X2X3X_{1}X_{2}X_{3} acts on the tuple (1,2,3,4,5)(1,2,3,4,5), 33 can never end up in the fourth or fifth position, whereas if X1X2X3X_{1}X_{2}X_{3} were uniform on A5A_{5}, 33 would end up in the fourth and fifth position with probability 1/51/5 each.

Consider the space =L2(G)\mathcal{M}=L^{2}(G) of \mathbb{R}-valued functions on a finite group GG. Since GG is finite, G\mathcal{M}\cong\mathbb{R}^{G} (with the Euclidean norm). Let 0={νν(G)=0}\mathcal{M}_{0}=\{\nu\in\mathcal{M}\mid\nu(G)=0\}. Let 𝒫\mathcal{P}\subseteq\mathcal{M} be the set of signed measures ν\nu on GG with ν(G)=1\nu(G)=1. Note that 𝒫\mathcal{P} and \mathcal{M} are parallel affine hyperplanes in G\mathbb{R}^{G}. Probability measures on GG are points in the simplex formed by the part of 𝒫\mathcal{P} in the positive orthant. The orthogonal complement to 0\mathcal{M}_{0} is the line spanned by the uniform probability measure π\pi on GG, and span{π}\operatorname{span}\{\pi\} intersects 𝒫\mathcal{P} at π\pi and nowhere else.

Any measure μi\mu_{i} on GG acts by on \mathcal{M} by convolution on the right. If μi\mu_{i} is a probability measure, the convolution operator Mi:ννμiM_{i}\colon\nu\mapsto\nu*\mu_{i} also sends 𝒫\mathcal{P} into itself. The following lemma tells us that, in this case, MiM_{i} contracts the distance between points of 𝒫\mathcal{P} and π\pi.

Lemma 3.4.

Let GG be any finite group and μ\mu a probability measure on GG. Let MM be the convolution operator ννμ\nu\mapsto\nu*\mu and let σ\sigma be the second-largest singular value of MM on L2(G)L^{2}(G).

  1. (1)

    If ν\nu is a signed measure on GG, then

    MνL2νL2.||M\nu||_{L^{2}}\leq||\nu||_{L^{2}}.
  2. (2)

    If ν,ν\nu,\nu^{\prime} are signed measures on GG with ν(G)=ν(G)\nu(G)=\nu^{\prime}(G), then

    MνMνL2σννL2.||M\nu-M\nu^{\prime}||_{L^{2}}\leq\sigma||\nu-\nu^{\prime}||_{L^{2}}.
Proof.

Part (1) is a case of Young’s convolution inequality for unimodular groups . To prove part (2), we will show that σ\sigma is the L2L^{2} operator norm of MM on the subspace of L2(G)L^{2}(G) consisting of signed measures with total mass 0.

Let MM^{*} be the adjoint operator to MM. Observe that MM^{*} is also a convolution operator, given by ννμˇ\nu\mapsto\nu*\check{\mu}, where μˇ(g)=μ(g1)\check{\mu}(g)=\mu(g^{-1}) for gGg\in G. Thus, MMM^{*}M is the convolution operator given by ννμμˇ\nu\mapsto\nu*\mu*\check{\mu}. In particular, MM and MMM^{*}M are each given by convolution with a probability measure, so they have a shared 1-eigenvector: the uniform measure π\pi on GG. The largest singular value of a real matrix coincides with its L2L^{2} operator norm. By part (1), the operator norm of MM is at most 1, so the largest eigenvalue of MMM^{*}M is exactly 1. Let L2(G)0=span{π}L^{2}(G)_{0}=\operatorname{span}\{\pi\}^{\perp}. Since π\pi is an eigenvector of MM, the operator MM restricts to an operator on L2(G)0L^{2}(G)_{0}, and since (M|L2(G)0)(M|L2(G)0)=(MM)|L2(G)0(M|_{L^{2}(G)_{0}})^{*}(M|_{L^{2}(G)_{0}})=(M^{*}M)|_{L^{2}(G)_{0}}, the singular values of M|L2(G)0M|_{L^{2}(G)_{0}} are the singular values of MM with a copy of 1 (the largest singular value of MM) excluded. Thus, the operator norm and largest singular value of M|L2(G)0M|_{L^{2}(G)_{0}} is the second-largest singular value of MM, which is σ\sigma. If ν(G)=ν(G)\nu(G)=\nu^{\prime}(G), then ννL2(H)0\nu-\nu^{\prime}\in L^{2}(H)_{0}, so

M(νν)L2σννL2.||M(\nu-\nu^{\prime})||_{L^{2}}\leq\sigma||\nu-\nu^{\prime}||_{L^{2}}.

Remark 3.5.

When the support of μ\mu does not generate GG, the conclusion of Lemma 3.4(2) still holds. However, we have σ=1\sigma=1, so Lemma 3.4(2) gives no useful information. For this reason, in Theorem 3.1, we write Ij={isupp(Q~j1)μi=Hj}I_{j}=\{i\mid\langle\operatorname{supp}(\tilde{Q}_{j-1})_{*}\mu_{i}\rangle=H_{j}\} rather than Ij={isupp(Q~j1)μiHj}I_{j}=\{i\mid\operatorname{supp}(\tilde{Q}_{j-1})_{*}\mu_{i}\subseteq H_{j}\}.

The second part of Lemma 3.4 implies that if the support of μi\mu_{i} generates GG and ν\nu is a probability measure, then MiνπL2σiνπL2||M_{i}\nu-\pi||_{L^{2}}\leq\sigma_{i}||\nu-\pi||_{L^{2}}, so applying MiM_{i} to a probability measure nn times contracts the distance to π\pi by a factor of σin\sigma_{i}^{n}. More generally, if the support of each μ\mu generates GG, then applying any combination of MiM_{i} in any order contracts the distance to π\pi by the appropriate product of σi\sigma_{i} factors. However, when the support of μi\mu_{i} is contained in a proper subgroup of GG, the second-largest singular value of MiM_{i} as an operator on \mathcal{M} is always 1. In this case, Lemma 3.4(2) applied to GG gives no useful information.

The key idea in the proof of Theorem 3.1 is that even though we cannot say outright that applying the operator MiM_{i} moves a probability measure closer to uniform, we will show in Lemma 3.8 that MiM_{i} moves a probability measure closer to some (explicit) subspace of \mathcal{M} containing π\pi. This subspace depends on the subgroup generated by the support of μi\mu_{i}. If we choose enough subspaces that their intersection is just {π}\{\pi\}, then we will be able to show that successive application of different MiM_{i}’s moves a probability measure closer to that intersection, that is, to the uniform probability measure. The condition that our chosen subspaces intersect in {π}\{\pi\} is exactly the condition G=HSHG=\langle\bigcup_{H\in S}H\rangle in Theorem 1.1, or the condition that Gk={e}G_{k}=\{e\} in Theorem 3.1.

For each subgroup HGH\leq G, let H\mathcal{M}_{H}\subseteq\mathcal{M} be the space of functions on \mathcal{M} which are uniform on each left coset of HH (i.e., for νH\nu\in\mathcal{M}_{H} and g1,g2Gg_{1},g_{2}\in G with g11g2Hg_{1}^{-1}g_{2}\in H, ν(g1)=ν(g2)\nu(g_{1})=\nu(g_{2})).

Lemma 3.6.

Let GG be a finite group and HGH\leq G be a subgroup. Let ν\nu\in\mathcal{M}. Let ν~H\tilde{\nu}\in\mathcal{M}_{H} be the signed measure on GG given by ν~(gh)=ν(gH)|H|\tilde{\nu}(gh)=\frac{\nu(gH)}{|H|} for hHh\in H. Then ν~\tilde{\nu} is the orthogonal projection of ν\nu onto the subspace H\mathcal{M}_{H} of \mathcal{M}. In particular, dL2(ν,H)=νν~L2d_{L^{2}}(\nu,\mathcal{M}_{H})=||\nu-\tilde{\nu}||_{L^{2}}.

Proof.

We have decompositions

H=gHG/Hspan{πgH}gHG/HL2(gH)=,\mathcal{M}_{H}=\bigoplus_{gH\in G/H}\operatorname{span}\{\pi_{gH}\}\subset\bigoplus_{gH\in G/H}L^{2}(gH)=\mathcal{M},

where πgHL2(gH)\pi_{gH}\in L^{2}(gH) is given by π(gh)=1/|H|\pi(gh)=1/\sqrt{|H|} for all hHh\in H. The projection operator H\mathcal{M}\to\mathcal{M}_{H} decomposes as a direct sum of projection operators, one for each coset of HH. In L2(gH)L^{2}(gH), projection onto span{πgH}\operatorname{span}\{\pi_{gH}\} is given by inner product with πgH\pi_{gH}, and we have ν|gH,πgHπgH=ν~|gH\left\langle\nu|_{gH},\pi_{gH}\right\rangle\pi_{gH}=\tilde{\nu}|_{gH}, which means the projection of ν\nu onto H\mathcal{M}_{H} is ν~\tilde{\nu}. ∎

Lemma 3.7.

Let GG be a finite group and HH a normal subgroup of GG. Let μL2(G)\mu\in L^{2}(G). Then μ:L2(G)L2(G)*\mu\colon L^{2}(G)\to L^{2}(G) preserves H\mathcal{M}_{H}, i.e., HμH\mathcal{M}_{H}*\mu\subseteq\mathcal{M}_{H}.

Proof.

Suppose νH\nu\in\mathcal{M}_{H}, so ν\nu is uniform on each left coset of HH. Since HH is normal, its left and right cosets coincide, so ν\nu is uniform on each right coset of HH. Say XνX\sim\nu and YμY\sim\mu are independent, so νμ\nu*\mu is the distribution of XYXY. We have [X=hg]=[X=hg]\mathbb{P}[X=hg]=\mathbb{P}[X=h^{\prime}g] for all h,hHh,h^{\prime}\in H and gGg\in G. For yGy\in G, we therefore have

[XY=hg|Y=y]=[X=hgy1]=[X=hgy1]=[XY=hg|Y=y]\mathbb{P}[XY=hg|Y=y]=\mathbb{P}[X=hgy^{-1}]=\mathbb{P}[X=h^{\prime}gy^{-1}]=\mathbb{P}[XY=h^{\prime}g|Y=y]

for all h,hHh,h^{\prime}\in H and gGg\in G. Summing over yy shows (νμ)(hg)=(νμ)(hg)(\nu*\mu)(hg)=(\nu*\mu)(h^{\prime}g) for all h,hHh,h^{\prime}\in H and gGg\in G. So, νμ\nu*\mu is uniform on right cosets (hence also left cosets) of HH and νμH\nu*\mu\in\mathcal{M}_{H}. ∎

Lemma 3.8.

Let GG be a finite group and HH a normal subgroup of GG. Let μ1,,μn\mu_{1},\dots,\mu_{n} be probability measures on GG and νn=μ1μn\nu_{n}=\mu_{1}*\dots*\mu_{n}. Let IH={1insuppμi=H}I_{H}=\{1\leq i\leq n\mid\langle\operatorname{supp}\mu_{i}\rangle=H\}. For each iIHi\in I_{H}, let σi\sigma_{i} be the second-largest singular value of μi*\mu_{i} on HH. Let HL2(G)\mathcal{M}_{H}\subseteq L^{2}(G) be the set of signed measures on GG that are uniform on left cosets of HH. Then

dL2(νn,H)2|G|1|G|iIHσi2d_{L^{2}}(\nu_{n},\mathcal{M}_{H})^{2}\leq\frac{|G|-1}{|G|}\prod_{i\in I_{H}}\sigma_{i}^{2}
Proof.

We say ν0=δe\nu_{0}=\delta_{e} is the Dirac measure on the identity of GG, so that νm+1=νmμm+1\nu_{m+1}=\nu_{m}*\mu_{m+1} for all 0m<n0\leq m<n.

We will show the following statement by induction on nn:

dL2(νn,H)2|G|1|G|iIHinσi2.d_{L^{2}}(\nu_{n},\mathcal{M}_{H})^{2}\leq\frac{|G|-1}{|G|}\prod_{\begin{subarray}{c}i\in I_{H}\\ i\leq n\end{subarray}}\sigma_{i}^{2}.

First, we have dL2(δe,H)2δeπL22=11|G|d_{L^{2}}(\delta_{e},\mathcal{M}_{H})^{2}\leq||\delta_{e}-\pi||_{L^{2}}^{2}=1-\frac{1}{|G|}. This proves P(0)P(0). .

Now suppose P(n)P(n) holds. Let ν~n\tilde{\nu}_{n} be the orthogonal projection of νn\nu_{n} onto H\mathcal{M}_{H}, described explicitly in Lemma 3.6. Then note that νnν~nL2=dL2(νn,H)||\nu_{n}-\tilde{\nu}_{n}||_{L^{2}}=d_{L^{2}}(\nu_{n},\mathcal{M}_{H}).

There are two cases, depending on whether n+1IHn+1\in I_{H}:

Suppose suppμn+1H\langle\operatorname{supp}\mu_{n+1}\rangle\neq H, so n+1IHn+1\notin I_{H}. By Lemma 3.4(1), νn+1ν~nμn+1L2νnν~nL2||\nu_{n+1}-\tilde{\nu}_{n}*\mu_{n+1}||_{L^{2}}\leq||\nu_{n}-\tilde{\nu}_{n}||_{L^{2}}.

By Lemma 3.7, we have ν~nμn+1H\tilde{\nu}_{n}*\mu_{n+1}\in\mathcal{M}_{H}, and

dL2(νn+1,H)νn+1ν~nμn+1L2νnν~nL2=dL2(νn,H).d_{L^{2}}(\nu_{n+1},\mathcal{M}_{H})\leq||\nu_{n+1}-\tilde{\nu}_{n}*\mu_{n+1}||_{L^{2}}\leq||\nu_{n}-\tilde{\nu}_{n}||_{L^{2}}=d_{L^{2}}(\nu_{n},\mathcal{M}_{H}).

Now suppose suppμn+1=H\langle\operatorname{supp}\mu_{n+1}\rangle=H, so n+1IHn+1\in I_{H}.

For gGg\in G, define g:L2(G)L2(G)g_{*}\colon L^{2}(G)\to L^{2}(G) by gν(h)=ν(g1h)g_{*}\nu(h)=\nu(g^{-1}h). Note that gg_{*} is an automorphism of normed spaces, and for signed measures ν\nu and μ\mu, we have g(νμ)=gνμg_{*}(\nu*\mu)=g_{*}\nu*\mu. For each left coset gHgH of HH we have

||(νnμn+1)|gH(ν~nμn+1)|gH||L2(gH)\displaystyle||(\nu_{n}*\mu_{n+1})|_{gH}-(\tilde{\nu}_{n}*\mu_{n+1})|_{gH}||_{L^{2}(gH)} =||(g1(νnμn+1))|H(g1(ν~nμn+1))|H||L2(H)\displaystyle=||(g^{-1}_{*}(\nu_{n}*\mu_{n+1}))|_{H}-(g^{-1}_{*}(\tilde{\nu}_{n}*\mu_{n+1}))|_{H}||_{L^{2}(H)}
=||(g1νnμn+1)|H(g1ν~nμn+1)|H||L2(H)\displaystyle=||(g^{-1}_{*}\nu_{n}*\mu_{n+1})|_{H}-(g^{-1}_{*}\tilde{\nu}_{n}*\mu_{n+1})|_{H}||_{L^{2}(H)}

Since μn+1\mu_{n+1} is supported on a subset of HH, for any νL2(G)\nu\in L^{2}(G) we have (νμn+1)|H=ν|Hμn+1|H(\nu*\mu_{n+1})|_{H}=\nu|_{H}*\mu_{n+1}|_{H}. Thus,

||(νnμn+1)|gH(ν~nμn+1)|gH||L2(gH)=||g1νn|Hμn+1|Hg1ν~n|Hμn+1|Hj||L2(H).||(\nu_{n}*\mu_{n+1})|_{gH}-(\tilde{\nu}_{n}*\mu_{n+1})|_{gH}||_{L^{2}(gH)}=||g^{-1}_{*}\nu_{n}|_{H}*\mu_{n+1}|_{H}-g^{-1}_{*}\tilde{\nu}_{n}|_{H}*\mu_{n+1}|_{H_{j}}||_{L^{2}(H)}.

By Lemma 3.6, ν~n|gH\tilde{\nu}_{n}|_{gH} is uniform on gHgH with total mass νn(gH)\nu_{n}(gH), so g1ν~n|Hg^{-1}_{*}\tilde{\nu}_{n}|_{H} is uniform on HH with total mass νn(gH)\nu_{n}(gH). Thus, g1ν~n|Hμn+1|H=g1ν~n|Hg^{-1}_{*}\tilde{\nu}_{n}|_{H}*\mu_{n+1}|_{H}=g^{-1}_{*}\tilde{\nu}_{n}|_{H}.

Applying Lemma 3.4(2) on HH, we get

||g1νn|Hμn+1|Hg1ν~n|H||L2(H)\displaystyle||g^{-1}_{*}\nu_{n}|_{H}*\mu_{n+1}|_{H}-g^{-1}_{*}\tilde{\nu}_{n}|_{H}||_{L^{2}(H)} σn+1||g1νn|Hg1ν~n|H||L2(H)\displaystyle\leq\sigma_{n+1}||g^{-1}_{*}\nu_{n}|_{H}-g^{-1}_{*}\tilde{\nu}_{n}|_{H}||_{L^{2}(H)}
=σn+1||(νnν~n)|gHj||L2(gH).\displaystyle=\sigma_{n+1}||(\nu_{n}-\tilde{\nu}_{n})|_{gH_{j}}||_{L^{2}(gH)}.

Adding up over cosets of HH gives

νn+1ν~nL2(G)σn+1νnν~nL2(G).||\nu_{n+1}-\tilde{\nu}_{n}||_{L^{2}(G)}\leq\sigma_{n+1}||\nu_{n}-\tilde{\nu}_{n}||_{L^{2}(G)}.

Hence,

dL2(νn+1,H)νn+1ν~nL2(G)σn+1νnν~nL2(G)=σn+1dL2(νn,H).d_{L^{2}}(\nu_{n+1},\mathcal{M}_{H})\leq||\nu_{n+1}-\tilde{\nu}_{n}||_{L^{2}(G)}\leq\sigma_{n+1}||\nu_{n}-\tilde{\nu}_{n}||_{L^{2}(G)}=\sigma_{n+1}d_{L^{2}}(\nu_{n},\mathcal{M}_{H}).

By induction, we get

dL2(νn,H)2|G|1|G|iIHinσi2.d_{L^{2}}(\nu_{n},\mathcal{M}_{H})^{2}\leq\frac{|G|-1}{|G|}\prod_{\begin{subarray}{c}i\in I_{H}\\ i\leq n\end{subarray}}\sigma_{i}^{2}.

Note that if HH is a subgroup of GG and P:GG/HP\colon G\twoheadrightarrow G/H is the projection onto the left coset space, then one can identify L2(G/H)L^{2}(G/H) with H\mathcal{M}_{H} as follows:

Lemma 3.9.

Let GG be a finite group and HH a subgroup. Let HL2(G)\mathcal{M}_{H}\subset L^{2}(G) be the space of measures uniform on left cosets of HH. Let P:GG/HP\colon G\twoheadrightarrow G/H send each element to the corresponding left coset of HH. Then the map ϕ:L2(G)L2(G/H)\phi\colon L^{2}(G)\to L^{2}(G/H) sending μ\mu to |H|1/2Pν|H|^{-1/2}P_{*}\nu restricts to an isometry of normed spaces ϕ|H:HL2(G/H)\phi|_{\mathcal{M}_{H}}\colon\mathcal{M}_{H}\cong L^{2}(G/H). Moreover, (ϕ|H)1ϕ(\phi|_{\mathcal{M}_{H}})^{-1}\circ\phi is the orthogonal projection map L2(G)HL^{2}(G)\twoheadrightarrow\mathcal{M}_{H}.

Proof.

The map ϕ|H\phi|_{\mathcal{M}_{H}} is norm-preserving because if ν\nu is uniform on left cosets of HH, then ν(gH)=|H|ν(g)\nu(gH)=|H|\nu(g) for gGg\in G. Indeed, we have

|H|1/2PνL2(G/H)=1|H|gHG/H|ν(gH)|2=1|H|2gG|ν(gH)|2=1|H|2gG|H|2|ν(g)|2=νL2(G).|||H|^{-1/2}P_{*}\nu||_{L^{2}(G/H)}=\frac{1}{|H|}\sum_{gH\in G/H}|\nu(gH)|^{2}=\frac{1}{|H|^{2}}\sum_{g\in G}|\nu(gH)|^{2}=\frac{1}{|H|^{2}}\sum_{g\in G}|H|^{2}|\nu(g)|^{2}=||\nu||_{L^{2}(G)}.

The inverse map (ϕ|H)1(\phi|_{\mathcal{M}_{H}})^{-1} is given by (ϕ|H)1(μ)(g)=μ(gH)/|H|(\phi|_{\mathcal{M}_{H}})^{-1}(\mu)(g)=\mu(gH)/|H|. The fact that (ϕ|H)1ϕ(\phi|_{\mathcal{M}_{H}})^{-1}\circ\phi is the orthogonal projection map L2(G)HL^{2}(G)\to\mathcal{M}_{H} follows from Lemma 3.6. ∎

Identifying L2(G/H)L^{2}(G/H) with H\mathcal{M}_{H} will allow us to prove Theorem 3.1 by induction. Using Lemma 3.8, we can say that a random walk approaches the subspace H\mathcal{M}_{H}. Then, we can consider its projection onto H\mathcal{M}_{H} as a random walk on G/HG/H. This allows us to ignore all random walk steps supported in HH. The key ingredient that allows us to combine Lemma 3.8 with the inductive hypothesis is the following lemma:

Lemma 3.10.

Let GG be a finite group and HGH\leq G. Let π\pi be the uniform distribution on GG and let μ\mu be any signed measure on GG. Let P:GG/HP\colon G\twoheadrightarrow G/H be the set map sending each element of GG to the corresponding left coset of HH. Let HL2(G)\mathcal{M}_{H}\subseteq L^{2}(G) be the set of signed measures uniform on left cosets of HH. Then

μπL2(G)2=1|H|PμPπL2(G/H)2+dL2(μ,H)2||\mu-\pi||_{L^{2}(G)}^{2}=\frac{1}{|H|}||P_{*}\mu-P_{*}\pi||_{L^{2}(G/H)}^{2}+d_{L^{2}}(\mu,\mathcal{M}_{H})^{2}
Proof.

Let μ~\tilde{\mu} be the orthogonal projection of μ\mu onto H\mathcal{M}_{H}. By Lemma 3.6, we have Pμ=Pμ~P_{*}\mu=P_{*}\tilde{\mu}. Then

μπL2(G)2\displaystyle||\mu-\pi||_{L^{2}(G)}^{2} =μ~πL2(G)2+μμ~L2(G)2\displaystyle=||\tilde{\mu}-\pi||_{L^{2}(G)}^{2}+||\mu-\tilde{\mu}||_{L^{2}(G)}^{2}
=1|H|Pμ~PπL2(G/H)2+dL2(μ,H)2\displaystyle=\frac{1}{|H|}||P_{*}\tilde{\mu}-P_{*}\pi||_{L^{2}(G/H)}^{2}+d_{L^{2}}(\mu,\mathcal{M}_{H})^{2}
=1|H|PμPπL2(G/H)2+dL2(μ,H)2\displaystyle=\frac{1}{|H|}||P_{*}\mu-P_{*}\pi||_{L^{2}(G/H)}^{2}+d_{L^{2}}(\mu,\mathcal{M}_{H})^{2}

The final lemma before the proof of Theorem 3.1 is a pair of facts about pushforwards to quotients:

Lemma 3.11.

Let GG be a finite group and HH a normal subgroup. Let P:GG/HP\colon G\twoheadrightarrow G/H be the projection. Then:

  1. (1)

    If μ,νL2(G)\mu,\nu\in L^{2}(G), we have P(μν)=PμPνP_{*}(\mu*\nu)=P_{*}\mu*P_{*}\nu.

  2. (2)

    Suppose μL2(G)\mu\in L^{2}(G) and the second-largest singular value of μ*\mu on suppμ\langle\operatorname{supp}\mu\rangle is σ\sigma. Then the second-largest singular value of (Pμ)*(P_{*}\mu) on P(suppμ)P(\langle\operatorname{supp}\mu\rangle) is at most σ\sigma.

Proof.
  1. (1)

    We have

    P(μν)(gH)=hH(μν)(gh)=hHkGμ(k)ν(k1gh)=kHG/HhHμ(kh)ν(h1k1gH)P_{*}(\mu*\nu)(gH)=\sum_{h\in H}(\mu*\nu)(gh)=\sum_{h\in H}\sum_{k\in G}\mu(k)\nu(k^{-1}gh)=\sum_{kH\in G/H}\sum_{h\in H}\mu(kh)\nu(h^{-1}k^{-1}gH)

    Since left cosets of HH are right cosets too, h1k1gH=k1gHh^{-1}k^{-1}gH=k^{-1}gH, so

    P(μν)(gH)=kHG/HhHμ(kh)ν(k1gH)=kHG/Hμ(kH)ν(k1gH)=PμPν.P_{*}(\mu*\nu)(gH)=\sum_{kH\in G/H}\sum_{h\in H}\mu(kh)\nu(k^{-1}gH)=\sum_{kH\in G/H}\mu(kH)\nu(k^{-1}gH)=P_{*}\mu*P_{*}\nu.
  2. (2)

    By restricting to the projection suppμP(suppμ)\langle\operatorname{supp}\mu\rangle\twoheadrightarrow P(\langle\operatorname{supp}\mu\rangle) we may as well assume suppμ=G\langle\operatorname{supp}\mu\rangle=G.

    Recall (from the proof of Lemma 3.4(2)) that the second-largest singular value of μ*\mu is the operator norm of μ*\mu acting on the subspace L2(G)0L^{2}(G)_{0} of measures with total mass 0. Suppose σ\sigma^{\prime} is the second-largest singular value of (Pμ)*(P_{*}\mu). Then

    σ=supνL2(G/H)ν(G/H)=0νPμL2(G/H)νL2(G/H)\sigma^{\prime}=\sup_{\begin{subarray}{c}\nu\in L^{2}(G/H)\\ \nu(G/H)=0\end{subarray}}\frac{||\nu*P_{*}\mu||_{L^{2}(G/H)}}{||\nu||_{L^{2}(G/H)}}

    Let HL2(G)\mathcal{M}_{H}\subseteq L^{2}(G) be the space of signed measures uniform on cosets of HH. By Lemma 3.9 and part (1) of this lemma, we have

    supνL2(G/H)ν(G/H)=0νPμL2(G/H)νL2(G/H)=supν~Hν~(G)=0Pν~PμL2(G/H)ν~L2(G/H)=supν~Hν~(G)=0P(ν~μ)L2(G/H)Pν~L2(G/H)\sup_{\begin{subarray}{c}\nu\in L^{2}(G/H)\\ \nu(G/H)=0\end{subarray}}\frac{||\nu*P_{*}\mu||_{L^{2}(G/H)}}{||\nu||_{L^{2}(G/H)}}=\sup_{\begin{subarray}{c}\tilde{\nu}\in\mathcal{M}_{H}\\ \tilde{\nu}(G)=0\end{subarray}}\frac{||P_{*}\tilde{\nu}*P_{*}\mu||_{L^{2}(G/H)}}{||\tilde{\nu}||_{L^{2}(G/H)}}=\sup_{\begin{subarray}{c}\tilde{\nu}\in\mathcal{M}_{H}\\ \tilde{\nu}(G)=0\end{subarray}}\frac{||P_{*}(\tilde{\nu}*\mu)||_{L^{2}(G/H)}}{||P_{*}\tilde{\nu}||_{L^{2}(G/H)}}

    Then by Lemma 3.9 again, we have

    σ=supν~Hν~(G)=0P(ν~μ)L2(G/H)Pν~L2(G/H)=supν~Hν~(G)=0ν~μL2(G)ν~L2(G)supν~L2(G)ν~(G)=0ν~μL2(G)ν~L2(G)=σ\sigma^{\prime}=\sup_{\begin{subarray}{c}\tilde{\nu}\in\mathcal{M}_{H}\\ \tilde{\nu}(G)=0\end{subarray}}\frac{||P_{*}(\tilde{\nu}*\mu)||_{L^{2}(G/H)}}{||P_{*}\tilde{\nu}||_{L^{2}(G/H)}}=\sup_{\begin{subarray}{c}\tilde{\nu}\in\mathcal{M}_{H}\\ \tilde{\nu}(G)=0\end{subarray}}\frac{||\tilde{\nu}*\mu||_{L^{2}(G)}}{||\tilde{\nu}||_{L^{2}(G)}}\leq\sup_{\begin{subarray}{c}\tilde{\nu}\in L^{2}(G)\\ \tilde{\nu}(G)=0\end{subarray}}\frac{||\tilde{\nu}*\mu||_{L^{2}(G)}}{||\tilde{\nu}||_{L^{2}(G)}}=\sigma

Proof of Theorem 3.1.

We will prove the following statement by induction on rr:

(P(r)P(r)) (Q~r)νn(Q~r)πL2(Gr)2j=r+1k|Gj1|1|Gr|(iIjσi2).\displaystyle||(\tilde{Q}_{r})_{*}\nu_{n}-(\tilde{Q}_{r})_{*}\pi||_{L^{2}(G_{r})}^{2}\leq\sum_{j=r+1}^{k}\frac{|G_{j-1}|-1}{|G_{r}|}\left(\prod_{i\in I_{j}}\sigma_{i}^{2}\right).

When r=kr=k, the right hand side of P(r)P(r) is 0. Since both (Q~r)νn(\tilde{Q}_{r})_{*}\nu_{n} and (Q~r)π(\tilde{Q}_{r})_{*}\pi are the unique probability measure on Gr={e}G_{r}=\{e\}, the left hand side is also 0, so P(k)P(k) holds.

Now suppose P(r+1)P(r+1) holds. We will show P(r)P(r) holds.

Since (Q~r)π(\tilde{Q}_{r})_{*}\pi is the uniform distribution on GrG_{r}, Lemma 3.10 applied to GrG_{r} and Hr+1H_{r+1} says

(Q~r)νn(Q~r)πL2(Gr)2\displaystyle||(\tilde{Q}_{r})_{*}\nu_{n}-(\tilde{Q}_{r})_{*}\pi||_{L^{2}(G_{r})}^{2} =1|Hr+1|(Q~r+1)νn(Q~r+1)πL2(Gr+1)2+dL2((Q~r)νn,Hr+1)2\displaystyle=\frac{1}{|H_{r+1}|}||(\tilde{Q}_{r+1})_{*}\nu_{n}-(\tilde{Q}_{r+1})_{*}\pi||_{L^{2}(G_{r+1})}^{2}+d_{L^{2}}((\tilde{Q}_{r})_{*}\nu_{n},\mathcal{M}_{H_{r+1}})^{2}

where Hr+1\mathcal{M}_{H_{r+1}} is the subspace of L2(Gr)L^{2}(G_{r}) consisting of measures uniform on cosets of Hr+1H_{r+1}. By the inductive hypothesis,

1|Hr+1|(Q~r+1)νn(Q~r+1)πL2(Gr+1)2\displaystyle\frac{1}{|H_{r+1}|}||(\tilde{Q}_{r+1})_{*}\nu_{n}-(\tilde{Q}_{r+1})_{*}\pi||_{L^{2}(G_{r+1})}^{2} j=r+2k|Gj1|1|Gr+1||Hr+1|(iIjσi2)\displaystyle\leq\sum_{j=r+2}^{k}\frac{|G_{j-1}|-1}{|G_{r+1}||H_{r+1}|}\left(\prod_{i\in I_{j}}\sigma_{i}^{2}\right)
=j=r+2k|Gj1|1|Gr|(iIjσi2).\displaystyle=\sum_{j=r+2}^{k}\frac{|G_{j-1}|-1}{|G_{r}|}\left(\prod_{i\in I_{j}}\sigma_{i}^{2}\right).

By Lemma 3.11(1), we have (Q~r)νn=(Q~r)μ1(Q~r)μn(\tilde{Q}_{r})_{*}\nu_{n}=(\tilde{Q}_{r})_{*}\mu_{1}*\dots*(\tilde{Q}_{r})_{*}\mu_{n}, so by Lemma 3.8 applied to GrG_{r}, Hr+1H_{r+1}, and the measures (Q~r)μi(\tilde{Q}_{r})_{*}\mu_{i}, we get

dL2((Q~r)νn,Hr+1)2|Gr|1|Gr|iIr+1σi2.d_{L^{2}}((\tilde{Q}_{r})_{*}\nu_{n},\mathcal{M}_{H_{r+1}})^{2}\leq\frac{|G_{r}|-1}{|G_{r}|}\prod_{i\in I_{r+1}}\sigma_{i}^{2}.

Hence,

(Q~r)νn(Q~r)πL2(Gr)2\displaystyle||(\tilde{Q}_{r})_{*}\nu_{n}-(\tilde{Q}_{r})_{*}\pi||_{L^{2}(G_{r})}^{2} j=r+2k|Gj1|1|Gr|(iIjσi2)+|Gr|1|Gr|iIr+1σi2\displaystyle\leq\sum_{j=r+2}^{k}\frac{|G_{j-1}|-1}{|G_{r}|}\left(\prod_{i\in I_{j}}\sigma_{i}^{2}\right)+\frac{|G_{r}|-1}{|G_{r}|}\prod_{i\in I_{r+1}}\sigma_{i}^{2}
=j=r+1k|Gj1|1|Gr|(iIjσi2),\displaystyle=\sum_{j=r+1}^{k}\frac{|G_{j-1}|-1}{|G_{r}|}\left(\prod_{i\in I_{j}}\sigma_{i}^{2}\right),

completing the induction. When r=0r=0, we get

νnπL22j=1k|Gj1|1|G|(iIjσi2).||\nu_{n}-\pi||_{L^{2}}^{2}\leq\sum_{j=1}^{k}\frac{|G_{j-1}|-1}{|G|}\left(\prod_{i\in I_{j}}\sigma_{i}^{2}\right).

Now we show how Theorem 3.1 implies Theorem 1.1 by giving a corollary slightly stronger than Theorem 1.1.

Corollary 3.12.

Let GG be a finite group, and let μ1,μ2,,μn\mu_{1},\mu_{2},\dots,\mu_{n} be probability measures on GG. For each subgroup HH of GG, let IH={iH=suppμi}I_{H}=\{i\mid H=\langle\operatorname{supp}\mu_{i}\rangle\}. Let H1,,HkH_{1},\dots,H_{k} be a finite set of subgroups of GG such that G=j=1kHjG=\left\langle\bigcup_{j=1}^{k}H_{j}\right\rangle and the image of HjH_{j} in G/H1Hj1G/H_{1}\cdots H_{j-1} is a normal subgroup for all 1jk1\leq j\leq k. Write νn=μ1μn\nu_{n}=\mu_{1}*\dots*\mu_{n}. Also, for each ii, let σi\sigma_{i} be the second-largest singular value of μi*\mu_{i} as an operator on L2(suppμi)L^{2}(\langle\operatorname{supp}\mu_{i}\rangle). Let π\pi be the uniform distribution on GG.

If IHjI_{H_{j}} is nonempty for each 1jk1\leq j\leq k, we have

νnπL2j=1k(iIHjσi).||\nu_{n}-\pi||_{L^{2}}\leq\sum_{j=1}^{k}\left(\prod_{i\in I_{H_{j}}}\sigma_{i}\right).
Proof.

We have j=1kHjH1Hk\bigcup_{j=1}^{k}H_{j}\subseteq H_{1}\cdots H_{k}, so H1Hk=GH_{1}\cdots H_{k}=G. Let Q~j:GG/H1Hj\tilde{Q}_{j}\colon G\to G/H_{1}\cdots H_{j} be the projection. For iIHji\in I_{H_{j}}, let σi\sigma_{i}^{\prime} be the second-largest singular value of (Q~j1)μi*(\tilde{Q}_{j-1})_{*}\mu_{i} on Q~j1(suppμi)=Q~j1(Hj)\tilde{Q}_{j-1}(\langle\operatorname{supp}\mu_{i}\rangle)=\tilde{Q}_{j-1}(H_{j}). By Lemma 3.11(2), we have σiσi\sigma_{i}^{\prime}\leq\sigma_{i}. Then applying Theorem 3.1 to the sequence

GG/H1G/H1H2G/H1Hk={e}G\longrightarrow G/H_{1}\longrightarrow G/H_{1}H_{2}\longrightarrow\dots\longrightarrow G/H_{1}\cdots H_{k}=\{e\}

we get that

νnπL22j=1k|G/H1Hj1|1|G|(iIHj(σi)2)j=1k(iIHjσi2).||\nu_{n}-\pi||_{L^{2}}^{2}\leq\sum_{j=1}^{k}\frac{|G/H_{1}\cdots H_{j-1}|-1}{|G|}\left(\prod_{i\in I_{H_{j}}}(\sigma_{i}^{\prime})^{2}\right)\leq\sum_{j=1}^{k}\left(\prod_{i\in I_{H_{j}}}\sigma_{i}^{2}\right).

Then the corollary follows by subadditivity of square root. ∎

We obtain Theorem 1.1 from Corollary 3.12 because the image of a normal subgroup under a surjection is normal.

4. Universality for Random Groups

The goal of this section is to prove Theorem 1.2.

To prove Theorem 1.2, we will use the moment method of Wood (see [Woo14, Woo19]) as follows. Let X1,X2,X_{1},X_{2},\dots be a sequence of random finitely generated abelian groups and YY be a random finitely generated abelian group. Let a>0a>0 be an integer and AA the set of isomorphism classes of abelian groups with exponent dividing aa. If for every GAG\in A we have

limn𝔼[#Sur(Xn,G)]=𝔼[#Sur(Y,G)]|2G|\lim_{n\to\infty}\mathbb{E}[\#\operatorname{Sur}(X_{n},G)]=\mathbb{E}[\#\operatorname{Sur}(Y,G)]\leq|\wedge^{2}G|

then for every HAH\in A we have

limn[Xn/aH]=[Y/aH]\lim_{n\to\infty}\mathbb{P}[X_{n}\otimes\mathbb{Z}/a\mathbb{Z}\cong H]=\mathbb{P}[Y\otimes\mathbb{Z}/a\mathbb{Z}\cong H]

[Woo19, Theorem 3.1]. The quantity 𝔼[#Sur(Xn,G)]\mathbb{E}[\#\operatorname{Sur}(X_{n},G)] is called the GG-moment of XnX_{n}.

Remark 4.1.

We can put a topology on the set of (isomorphism classes of) finitely generated abelian groups given by a basis of open sets of the form

Ua,H={X finitely generated abelianX/aH}U_{a,H}=\{X\text{ finitely generated abelian}\mid X\otimes\mathbb{Z}/a\mathbb{Z}\cong H\}

indexed by positive integers aa and abelian groups HH of exponent dividing aa. The assertion that ()(**) holds for all choices of aa and HH is equivalent to the assertion that the distribution of XnX_{n} converges weakly to the distribution of YY in this topology. In particular, if ()(*) holds for all abelian groups GG, then the distribution of XnX_{n} converges weakly to the distribution of YY.

If YλuY\sim\lambda_{u}, then [Woo19, Lemma 3.2] gives

𝔼[#Sur(Y,G)]=|G|u.\mathbb{E}[\#\operatorname{Sur}(Y,G)]=|G|^{-u}.

Following this strategy, we obtain Theorem 1.2 as a corollary of Theorem 4.18, which states that if XnX_{n} are the cokernels of n×(n+u)n\times(n+u) random matrices satisfying appropriate conditions, then limn𝔼[#Sur(Xn,G)]=|G|u\lim_{n}\mathbb{E}[\#\operatorname{Sur}(X_{n},G)]=|G|^{-u}.

When XnX_{n} is the cokernel of a random n×mn\times m matrix MM, the problem of counting surjections from XnX_{n} into GG can be attacked with combinatorics. Say Xn=n/ΛX_{n}=\mathbb{Z}^{n}/\Lambda, where Λ\Lambda is a random subgroup of n\mathbb{Z}^{n} (e.g., the column space of a random integer matrix). Then surjections XnGX_{n}\to G correspond one-to-one with surjections nG\mathbb{Z}^{n}\to G which vanish on Λ\Lambda. It follows from linearity of expectation that

𝔼[#Sur(n/Λ,G)]=fSur(n,G)[f(Λ)=0].\mathbb{E}[\#\operatorname{Sur}(\mathbb{Z}^{n}/\Lambda,G)]=\sum_{f\in\operatorname{Sur}(\mathbb{Z}^{n},G)}\mathbb{P}[f(\Lambda)=0].

In the case of cokernels of random matrices, Λ\Lambda is the subgroup generated by the columns of the random matrix, viewed as random elements of n\mathbb{Z}^{n}. But we can also view MM as a random element of (n)m(\mathbb{Z}^{n})^{m}. Given a map f:nGf\colon\mathbb{Z}^{n}\to G, we get by abuse of notation a map f:(n)mGmf\colon(\mathbb{Z}^{n})^{m}\to G^{m} applying ff to each component. Then we have that f(Λ)=0f(\Lambda)=0 if and only if f(M)=0f(M)=0. Thus, we want to bound the probabilities f(M)=0f(M)=0. Past work on random matrices with independent entries (e.g., [NW22]) has observed that if ZZ is a random tuple in n\mathbb{Z}^{n} with independent, sufficiently regular components, then for most fSur(n,G)f\in\operatorname{Sur}(\mathbb{Z}^{n},G), the element f(Z)Gf(Z)\in G is close to uniformly distributed. Applying this to each column independently allows us to compute [f(M)=0]\mathbb{P}[f(M)=0]. In this work, we apply the same principle to consider several columns of a random matrix at a time.

4.1. Balanced elements

The following definition captures the idea that a random element in a group is not too concentrated in a particular coset.

Definition 4.2.

Let GG be a group. A GG-valued random variable XX is ε\varepsilon-balanced if for any proper subgroup H<GH<G and element gGg\in G, we have [XgH]1ε\mathbb{P}[X\in gH]\leq 1-\varepsilon.

This definition agrees with the definition in [Woo19] when GG is a finite cyclic group. Here is an example of an ε\varepsilon-balanced random variable that does not take values in a cyclic group.

Example 4.3.

Let GG be a finitely generated group with finite generating set SS containing the identity, and let XX be a random variable supported on SS with mingS[X=g]=ε\min_{g\in S}\mathbb{P}[X=g]=\varepsilon. Then XX is ε\varepsilon-balanced.

Indeed, suppose HH is a subgroup of GG and gGg\in G such that [XgH]>1ε\mathbb{P}[X\in gH]>1-\varepsilon. Then [XgH]=1\mathbb{P}[X\in gH]=1, so SgHS\subset gH. Since SS contains the identity element of GG, we must have gH=HgH=H, and since SgH=HS\subset gH=H, we must have H=GH=G.

In this paper, we consider n×mn\times m integer matrices as elements of the abelian group (n)m(\mathbb{Z}^{n})^{m}. For each subset SS of [n]×[m][n]\times[m], we have a quotient map πS\pi_{S} from (n)m(\mathbb{Z}^{n})^{m} onto S\mathbb{Z}^{S} given by taking the entries of a matrix indexed by pairs in SS. We say that a subset of the entries of a random matrix MM with indices SS is jointly ε\varepsilon-balanced if πS(M)\pi_{S}(M) is ε\varepsilon-balanced in S\mathbb{Z}^{S}.

The new definition of ε\varepsilon-balanced has some desirable properties that help construct new examples of ε\varepsilon-balanced random variables.

Lemma 4.4.
  1. (1)

    If π:GQ\pi\colon G\twoheadrightarrow Q is a surjective homomorphism of groups and XX is ε\varepsilon-balanced in GG, then π(X)\pi(X) is ε\varepsilon-balanced in QQ.

  2. (2)

    Let G,GG,G^{\prime} be groups, XX be ε\varepsilon-balanced in GG, and YY be ε\varepsilon-balanced in GG^{\prime}. If XX and YY are independent, then (X,Y)(X,Y) is ε\varepsilon-balanced in G×GG\times G^{\prime}.

Proof.
  1. (1)

    Let qKQqK\subsetneq Q be a coset of a proper subgroup of QQ. Let q~π1(q)\tilde{q}\in\pi^{-1}(q), so π1(qK)=q~π1(K)\pi^{-1}(qK)=\tilde{q}\pi^{-1}(K) is a coset of a proper subgroup of GG. Since XX is ε\varepsilon-balanced,

    [π(X)qK][Xq~π1(K)]1ε,\mathbb{P}[\pi(X)\in qK]\leq\mathbb{P}[X\in\tilde{q}\pi^{-1}(K)]\leq 1-\varepsilon,

    as desired.

  2. (2)

    Let kHkH be a coset of a proper subgroup of G×GG\times G^{\prime}. Note that

    [(X,Y)kH]=[(X,e)(e,Y1)kH]=[(X,e)(e,Y1)kH(G×{e})].\mathbb{P}[(X,Y)\in kH]=\mathbb{P}[(X,e)\in(e,Y^{-1})kH]=\mathbb{P}[(X,e)\in(e,Y^{-1})kH\cap(G\times\{e\})].

    Recall that the intersection of two cosets in a group is either empty or a coset of their intersection. In particular, (e,Y1)kH(G×{e})(e,Y^{-1})kH\cap(G\times\{e\}) is either empty or a coset of a subgroup of G×{e}G\times\{e\}.

    There are two cases, depending on whether (e,Y1)kH(G×{e})(e,Y^{-1})kH\cap(G\times\{e\}) is always a proper subset of G×{e}G\times\{e\}:

    1. i.

      If (e,y1)kH(G×{e})G×{e}(e,y^{-1})kH\cap(G\times\{e\})\subsetneq G\times\{e\} for all yGy\in G^{\prime}:

      Condition on Y=yY=y for some fixed yGy\in G^{\prime}. Since XX and YY are independent, and XX is ε\varepsilon-balanced,

      [(X,e)(e,Y1)kH(G×{e})Y=y]=[(X,e)(e,y1)kH(G×{e})]1ε.\mathbb{P}[(X,e)\in(e,Y^{-1})kH\cap(G\times\{e\})\mid Y=y]=\mathbb{P}[(X,e)\in(e,y^{-1})kH\cap(G\times\{e\})]\leq 1-\varepsilon.
    2. ii.

      If G×{e}(e,y1)kHG\times\{e\}\subseteq(e,y^{-1})kH for some yYy\in Y, then (e,e)(e,y1)kH(e,e)\in(e,y^{-1})kH, so in particular (e,y1)kH(e,y^{-1})kH is a subgroup of G×GG\times G^{\prime} and we must have (e,y1)kH=H(e,y^{-1})kH=H. We claim that H=G×HH=G\times H^{\prime} for some proper subgroup HH^{\prime} of GG^{\prime}.

      Indeed, let π:G×GG\pi\colon G\times G^{\prime}\to G^{\prime} be the projection and let H=π(H)H^{\prime}=\pi(H). On one hand, clearly HG×π(H)H\subseteq G\times\pi(H). On the other, if (g,h)G×H(g,h^{\prime})\in G\times H^{\prime}, then h=π(g,h)h^{\prime}=\pi(g^{\prime},h) for some (g,h)H(g^{\prime},h)\in H. Then (g,h)=(g(g)1,e)(g,h)(g,h^{\prime})=(g(g^{\prime})^{-1},e)(g^{\prime},h). Since (g(g)1,e)G×{e}H(g(g^{\prime})^{-1},e)\in G\times\{e\}\subseteq H, we have (g,h)H(g,h^{\prime})\in H. Hence H=G×HH=G\times H^{\prime}. Note that HGH^{\prime}\lneq G^{\prime}, or else H=G×GH=G\times G^{\prime} is not a proper subgroup.

      Then

      [(X,Y)kH]=[YH]1ε.\mathbb{P}[(X,Y)\in kH]=\mathbb{P}[Y\in H^{\prime}]\leq 1-\varepsilon.

    Hence, in both cases we have [(X,Y)kH]1ε\mathbb{P}[(X,Y)\in kH]\leq 1-\varepsilon and since this holds for every proper coset kHkH, we have that (X,Y)(X,Y) is balanced.

Note that Lemma 4.4 gives us a nice way to build up ε\varepsilon-balanced matrices. If the entries of a random matrix can be partitioned into independent subsets and each of these subsets of the entries is jointly ε\varepsilon-balanced, then the whole matrix is ε\varepsilon-balanced. For example, any matrix with independent, ε\varepsilon-balanced entries (as in [Woo19]) is ε\varepsilon-balanced as a matrix.

When a random variable is ε\varepsilon-balanced, we can get an upper bound on the associated singular value.

Lemma 4.5.

Suppose GG is a finite group and XX is ε\varepsilon-balanced in GG with distribution μ\mu. Let σ\sigma be the second largest singular value of the operator μ*\mu on L2(G)L^{2}(G). Then

σexp(ε2|G|3).\sigma\leq\exp\left(-\frac{\varepsilon}{2|G|^{3}}\right).
Proof.

Note that σ\sigma is the square root of the second largest eigenvalue of the operator νμμˇ:L2(G)L2(G)*\nu\coloneqq*\mu*\check{\mu}\colon L^{2}(G)\to L^{2}(G), where μˇ*\check{\mu} is the adjoint to the operator μ*\mu, given by μˇ(g)=μ(g1)\check{\mu}(g)=\mu(g^{-1}). The operator ν*\nu is the transition operator for a random walk on GG, where each step is a difference of two independent copies of XX.

In particular, note that ν=νˇ\nu=\check{\nu}. For any generating set Σ\Sigma of GG, [Sal04, Theorem 6.2] applied to ΣΣ1\Sigma\cup\Sigma^{-1} shows that the second-largest eigenvalue σ2\sigma^{2} of ν*\nu is bounded above by

σ21mD2,\sigma^{2}\leq 1-\frac{m}{D^{2}},

where m=minxΣν(x)m=\min_{x\in\Sigma}\nu(x) and DD is the diameter of the Cayley graph of (G,Σ)(G,\Sigma). In particular, D|G|D\leq|G|.

The goal is to choose an appropriate Σ\Sigma to bound mm from below. Note that if X1X_{1} and X2X_{2} are ε\varepsilon-balanced and independent, then so is X1X21X_{1}X_{2}^{-1} (via conditioning on X2X_{2}). In particular, ν\nu is ε\varepsilon-balanced.

We proceed iteratively. Having chosen x1,,xn1x_{1},\dots,x_{n-1} (including the empty set n=1n=1), if x1,,xn1=G\langle x_{1},\dots,x_{n-1}\rangle=G then we are done. Otherwise, since ν\nu is ε\varepsilon-balanced, ν(x1,,xn1)1ε\nu(\langle x_{1},\dots,x_{n-1}\rangle)\leq 1-\varepsilon. Choose

xn=argmaxxGx1,,xn1ν(x).x_{n}=\operatorname{argmax}_{x\in G\setminus\langle x_{1},\dots,x_{n-1}\rangle}\nu(x).

Since ν(x1,,xn1)1ε\nu(\langle x_{1},\dots,x_{n-1}\rangle)\leq 1-\varepsilon, we have ν(Gx1,,xn1)ε\nu(G\setminus\langle x_{1},\dots,x_{n-1}\rangle)\geq\varepsilon, so ν(xn)ε|Gx1,,xn1|ε|G|\nu(x_{n})\geq\frac{\varepsilon}{|G\setminus\langle x_{1},\dots,x_{n-1}\rangle|}\geq\frac{\varepsilon}{|G|}.

Hence we have mε|G|m\geq\frac{\varepsilon}{|G|}, so

σ1ε|G|31ε2|G|3exp(ε2|G|3),\sigma\leq\sqrt{1-\frac{\varepsilon}{|G|^{3}}}\leq 1-\frac{\varepsilon}{2|G|^{3}}\leq\exp\left(-\frac{\varepsilon}{2|G|^{3}}\right),

as desired. ∎

Now we will use the ε\varepsilon-balanced condition to give a related balancedness condition for matrices that contains information about how balanced and independent the entries are.

Definition 4.6.

Let SS be a finite set. A partition of SS is a collection 𝒫={P1,,Pk}2S\mathcal{P}=\{P_{1},\dots,P_{k}\}\subseteq 2^{S}, such that S=P1P2PkS=P_{1}\sqcup P_{2}\sqcup\dots\sqcup P_{k} and each PiP_{i} is nonempty. We say |𝒫|=maxi#Pi|\mathcal{P}|=\max_{i}\#P_{i} and #𝒫=k\#\mathcal{P}=k. If σ2S\sigma\subseteq 2^{S}, write σ\cup\sigma for SσS\bigcup_{S\in\sigma}S.

Note that #𝒫|𝒫|#S\#\mathcal{P}\cdot|\mathcal{P}|\geq\#S.

The next definition specifies the kinds of restrictions we will give for the matrices in our universality class. The idea is that we can split up the columns of the matrix and then the rows, so that the resulting sections of the matrix are ε\varepsilon-balanced.

If MM is an n×mn\times m matrix, S={s1<<sk}[n]S=\{s_{1}<\dots<s_{k}\}\subset[n], and T={t1<<t}[m]T=\{t_{1}<\dots<t_{\ell}\}\subset[m], then MS,TM_{S,T} is the k×k\times\ell matrix (Msi,tj)1ik,1j(M_{s_{i},t_{j}})_{1\leq i\leq k,1\leq j\leq\ell}.

Definition 4.7.

An n×mn\times m random matrix MM with entries in a ring RR is (w,h,ε)(w,h,\varepsilon)-balanced if there is a partition 𝒬={Q1,,Qr}\mathcal{Q}=\{Q_{1},\dots,Q_{r}\} of [m][m] and a partition 𝒫={P1,,P}\mathcal{P}=\{P_{1},\dots,P_{\ell}\} of [n][n] with |𝒬|w|\mathcal{Q}|\leq w, |𝒫|h|\mathcal{P}|\leq h, and such that each random matrix MPi,QjM_{P_{i},Q_{j}} is ε\varepsilon-balanced in the additive abelian group (R#Pi)#Qj(R^{\#P_{i}})^{\#Q_{j}} and the random matrices MPi,QjM_{P_{i},Q_{j}} are independent.

If |𝒫|=|𝒬|=1|\mathcal{P}|=|\mathcal{Q}|=1 then we recover the definition of ε\varepsilon-balanced from [Woo19] and other related work.

Now we are ready to state the main theorem of this section:

Theorem 4.8.

Let u0u\geq 0 be an integer. Let (wn)n,(hn)n(w_{n})_{n},(h_{n})_{n} be sequences of real numbers such that wn=o(logn)w_{n}=o(\log n), hn=O(n1α)h_{n}=O(n^{1-\alpha}), and εnnβ\varepsilon_{n}\geq n^{-\beta} for some 0<α10<\alpha\leq 1 and 0<β<α/20<\beta<\alpha/2.

For each integer n0n\geq 0, let MnM_{n} be an (wn,hn,εn)(w_{n},h_{n},\varepsilon_{n})-balanced n×(n+u)n\times(n+u) random matrix with entries in \mathbb{Z}. Let YλuY\sim\lambda_{u}. Then for all positive integers aa and abelian groups HH of exponent dividing aa we have

limn[coker(Mn)/aH]=[Y/aH]=λu(Ua,H).\lim_{n\to\infty}\mathbb{P}[\operatorname{coker}(M_{n})\otimes\mathbb{Z}/a\mathbb{Z}\cong H]=\mathbb{P}[Y\otimes\mathbb{Z}/a\mathbb{Z}\cong H]=\lambda_{u}(U_{a,H}).

Together with Remark 4.1, this gives Theorem 1.2.

As discussed at the beginning of this section, we will prove this by computing the limiting moments of coker(Mn)\operatorname{coker}(M_{n}), which involves estimating [f(Mn)=0]\mathbb{P}[f(M_{n})=0] for maps nG\mathbb{Z}^{n}\to G.

Remark 4.9.

The same proof will work as written when the entries of MnM_{n} come from any ring RR with at most one quotient to /a\mathbb{Z}/a\mathbb{Z} for any positive integer aa. Some examples of interest are the pp-adic integers p\mathbb{Z}_{p} or a product ipi\prod_{i}\mathbb{Z}_{p_{i}} for some collection of distinct primes pip_{i}. We will find that when RR has exactly one quotient to /a\mathbb{Z}/a\mathbb{Z}, then for any finite abelian group GG of exponent dividing aa, the limiting GG-moment of coker(Mn)\operatorname{coker}(M_{n}) is |G|u|G|^{-u}. Then we get the conclusion of Theorem 4.8 for those aa for which RR has a quotient to /a\mathbb{Z}/a\mathbb{Z}.

4.2. Bounds for most maps

It turns out that (w,h,ε)(w,h,\varepsilon)-balanced is a strong enough condition that we can get bounds on [f(M)=0]\mathbb{P}[f(M)=0] for the vast majority of maps ff.

Definition 4.10.

If VV is an abelian group with generating set SS and TST\subseteq S, we write VTV_{\setminus T} for the subgroup ST\langle S\setminus T\rangle of VV. When V=(/a)nV=(\mathbb{Z}/a\mathbb{Z})^{n} or n\mathbb{Z}^{n} we implicitly take SS to be the “standard basis”.

Let 𝒫={P1,,P}\mathcal{P}=\{P_{1},\dots,P_{\ell}\} be a partition of SS and GG be a finite abelian group. A function f:VGf\colon V\to G is a 𝒫\mathcal{P}-code of distance ww if for any σ𝒫\sigma\subset\mathcal{P} with |σ|<w|\cup\sigma|<w, we have f(Vσ)=Gf(V_{\setminus\cup\sigma})=G.

To approximate [f(M)=0]\mathbb{P}[f(M)=0] for codes ff, we will split the matrices MM into independent sets of columns. Each such set of rr random columns gets mapped to something close to uniform in GrG^{r}. The following lemma is analogous to [Woo19, Lemma 2.1].

Lemma 4.11.

Let n,r1n,r\geq 1 be integers. Let GG be a finite abelian group and let aa be a multiple of the exponent of GG. Let NN be the number of subgroups of GG. Let ε>0\varepsilon>0 be a real number. Let V=(/a)nV=(\mathbb{Z}/a\mathbb{Z})^{n}. Let 𝒫={Pi}\mathcal{P}=\{P_{i}\} be a partition of [n][n] and let =|𝒫|\ell=|\mathcal{P}|. Let fHom(V,G)f\in\operatorname{Hom}(V,G) be a 𝒫\mathcal{P}-code of distance w<nw<n.

Let MM be an n×rn\times r random matrix in VrV^{r} such that the matrices MPi,[r]M_{P_{i},[r]} are independent and ε\varepsilon-balanced as random elements of ((/a)#Pi)r((\mathbb{Z}/a\mathbb{Z})^{\#P_{i}})^{r}.

Let g1,,grGg_{1},\dots,g_{r}\in G. Then

|[f(M)=(g1,,gr)]|G|r|Nexp(εw2N|G|3r)|\mathbb{P}[f(M)=(g_{1},\dots,g_{r})]-|G|^{-r}|\leq N\exp\left(-\frac{\varepsilon w}{2\ell N|G|^{3r}}\right)
Proof.

Let e1,,ene_{1},\dots,e_{n} be the standard generating set for VV. For i=1,,#𝒫i=1,\dots,\#\mathcal{P}, let Vi=ejjPi#PiV_{i}=\langle e_{j}\mid j\in P_{i}\rangle\cong\mathbb{Z}^{\#P_{i}}.

The idea is to treat f(M)f(M) as a random walk in GrG^{r}. We have

f(M)=i=1#𝒫f(MPi,[r]),f(M)=\sum_{i=1}^{\#\mathcal{P}}f(M_{P_{i},[r]}),

where MPi,[r]M_{P_{i},[r]} is interpreted as an ε\varepsilon-balanced random element of Vir((/a)#Pi)rV_{i}^{r}\cong((\mathbb{Z}/a\mathbb{Z})^{\#P_{i}})^{r}, a subgroup of ((/a)n)r((\mathbb{Z}/a\mathbb{Z})^{n})^{r}.

Let S={HGrH=f(Vir) for at least w/N values of i}S=\{H\leq G^{r}\mid H=f(V_{i}^{r})\text{ for at least }w/\ell N\text{ values of }i\}. Note that f(Vir)=f(Vi)rf(V_{i}^{r})=f(V_{i})^{r}, so as ii ranges over 1,,#𝒫1,\dots,\#\mathcal{P} there are at most NN possible values for f(Vir)f(V_{i}^{r}), each an rrth power of a subgroup of GG. Let I={if(Vir)S}I=\{i\mid f(V_{i}^{r})\notin S\}. Then #Iw/\#I\leq w/\ell, and so |iIPi|w|\bigcup_{i\in I}P_{i}|\leq w. Since ff is a 𝒫\mathcal{P}-code of distance ww, it remains surjective if we discard all of these indices, which means the images of the VirV_{i}^{r}s with f(Vir)Sf(V_{i}^{r})\in S generate GrG^{r}. In other words, we have HSH=Gr\langle\bigcup_{H\in S}H\rangle=G^{r}. The subgroups in SS will be the ones we use in the random walk, applying Theorem 1.1.

By the definition of SS, for each HH in SS we have #IHw/N\#I_{H}\geq w/\ell N. By Lemma 4.4, the steps f(MPi,[r])f(M_{P_{i},[r]}) are ε\varepsilon-balanced, which means that by Lemma 4.5 the second largest singular value σi\sigma_{i} of the iith step f(MPi,[r])f(M_{P_{i},[r]}) is bounded above: σiexp(ε2|G|3r)\sigma_{i}\leq\exp\left(-\frac{\varepsilon}{2|G|^{3r}}\right) (using the fact that each f(MPi,[r])f(M_{P_{i},[r]}) is supported on a subgroup of GrG^{r}).

Hence by Theorem 1.1 we have

|[f(M)=(g1,,gr)]|G|r|HSexp(εw2N|G|3r)Nexp(εw2N|G|3r),|\mathbb{P}[f(M)=(g_{1},\dots,g_{r})]-|G|^{-r}|\leq\sum_{H\in S}\exp\left(-\frac{\varepsilon w}{2\ell N|G|^{3r}}\right)\leq N\exp\left(-\frac{\varepsilon w}{2\ell N|G|^{3r}}\right),

as desired. ∎

To combine these estimates we will use a result in the flavor of [Woo19, Lemma 2.3]:

Lemma 4.12.

Let x1,,xm1x_{1},\dots,x_{m}\geq-1 be real numbers such that i=1mmax{0,xi}log2\sum_{i=1}^{m}\max\{0,x_{i}\}\leq\log 2. Then

|i=1m(1+xi)1|2i=1m|xi|\left|\prod_{i=1}^{m}(1+x_{i})-1\right|\leq 2\sum_{i=1}^{m}|x_{i}|

and

i=1mmin{0,xi}i=1m(1+xi)12i=1mmax{0,xi}.\sum_{i=1}^{m}\min\{0,x_{i}\}\leq\prod_{i=1}^{m}(1+x_{i})-1\leq 2\sum_{i=1}^{m}\max\{0,x_{i}\}.
Proof.

The first statement follows from the second statement because max{0,xi}|xi|\max\{0,x_{i}\}\leq|x_{i}| and min{0,xi}|xi|\min\{0,x_{i}\}\geq-|x_{i}|. So, we will show the second statement.

First, assume xi0x_{i}\leq 0 for all ii. In that case,

i=1m(1+xi)1+i=1mxi.\prod_{i=1}^{m}(1+x_{i})\geq 1+\sum_{i=1}^{m}x_{i}.

Next, assume xi0x_{i}\geq 0 for all ii. Using the fact that 1+xiexi1+x_{i}\leq e^{x_{i}}, we get

i=1m(1+xi)ei=1mxi.\prod_{i=1}^{m}(1+x_{i})\leq e^{\sum_{i=1}^{m}x_{i}}.

We have ex1=2xe^{x}-1=2x at x=0x=0 and ddx(ex1)ddx(2x)\frac{d}{dx}(e^{x}-1)\leq\frac{d}{dx}(2x) for xlog2x\leq\log 2, so ex12xe^{x}-1\leq 2x for 0xlog20\leq x\leq\log 2. Hence, if i=1mxilog2\sum_{i=1}^{m}x_{i}\leq\log 2, then exp(i=1mxi)12i=1mxi\exp\left(\sum_{i=1}^{m}x_{i}\right)-1\leq 2\sum_{i=1}^{m}x_{i}.

Now consider the general case. By replacing each negative xix_{i} with zero, we can only increase the product i=1m(1+xi)\prod_{i=1}^{m}(1+x_{i}). On the other hand, by replacing each positive xix_{i} with zero, we can only decrease it. Hence, for general xix_{i}, we get

i=1mmin{0,xi}i=1m(1+min{0,xi})1i=1m(1+xi)1i=1m(1+max{0,xi})12i=1mmax{0,xi}.\sum_{i=1}^{m}\min\{0,x_{i}\}\leq\prod_{i=1}^{m}(1+\min\{0,x_{i}\})-1\leq\prod_{i=1}^{m}(1+x_{i})-1\leq\prod_{i=1}^{m}(1+\max\{0,x_{i}\})-1\leq 2\sum_{i=1}^{m}\max\{0,x_{i}\}.

Applying this lemma with xix_{i} being the error in Lemma 4.11 multiplied by |G|r|G|^{r} yields an estimate on the probability that the whole matrix maps to zero:

Lemma 4.13.

Let u0u\geq 0 be an integer. Let GG be a finite abelian group and let aa be a multiple of the exponent of GG. Let (wn)n,(hn)n,(δn)n,(εn)n(w_{n})_{n},(h_{n})_{n},(\delta_{n})_{n},(\varepsilon_{n})_{n} be sequences of real numbers such that wn=o(logn)w_{n}=o(\log n), hn=O(n1α)h_{n}=O(n^{1-\alpha}), and εnδnnα+β\varepsilon_{n}\delta_{n}\geq n^{-\alpha+\beta} for some 0<βα10<\beta\leq\alpha\leq 1.

For a natural number nn, let V=(/a)nV=(\mathbb{Z}/a\mathbb{Z})^{n}. Let MM be an (wn,hn,εn)(w_{n},h_{n},\varepsilon_{n})-balanced n×(n+u)n\times(n+u) random matrix with entries in /a\mathbb{Z}/a\mathbb{Z}. Let 𝒫\mathcal{P} be the row partition associated to MM and let fHom(V,G)f\in\operatorname{Hom}(V,G) be a 𝒫\mathcal{P}-code of distance nδnn\delta_{n}.

Then there are constants K,c,γ>0K,c,\gamma>0 depending only on GG, α\alpha, β\beta, and the sequences (wn)n,(hn)n(w_{n})_{n},(h_{n})_{n} such that for all g1,,gn+uGg_{1},\dots,g_{n+u}\in G,

|[f(M)=(g1,,gn+u)]|G|nu|Kexp(cnγ)|G|n+u|\mathbb{P}[f(M)=(g_{1},\dots,g_{n+u})]-|G|^{-n-u}|\leq\frac{K\exp(-cn^{\gamma})}{|G|^{n+u}}
Proof.

Let 𝒫\mathcal{P} and 𝒬\mathcal{Q} be the row and column partitions for MM as in the definition of (wn,hn,εn)(w_{n},h_{n},\varepsilon_{n})-balanced. Let Mi=M[n],QiM_{i}=M_{[n],Q_{i}} for each ii. Let gQi=(gjjQi)g_{Q_{i}}=(g_{j}\mid j\in Q_{i}). By independence,

[f(M)=(g1,,gn+u)]=i[f(Mi)=gQi].\mathbb{P}[f(M)=(g_{1},\dots,g_{n+u})]=\prod_{i}\mathbb{P}[f(M_{i})=g_{Q_{i}}].

For each ii, let xi=|G|#Qi[f(Mi)=gQi]1x_{i}=|G|^{\#Q_{i}}\mathbb{P}[f(M_{i})=g_{Q_{i}}]-1. By Lemma 4.11, we have

|xi|\displaystyle|x_{i}| N|G|#Qiexp(nεnδn2Nhn|G|3#Qi)\displaystyle\leq N|G|^{\#Q_{i}}\exp\left(-\frac{n\varepsilon_{n}\delta_{n}}{2Nh_{n}|G|^{3\#Q_{i}}}\right)
N|G|wnexp(nεnδn2Nhn|G|3wn).\displaystyle\leq N|G|^{w_{n}}\exp\left(-\frac{n\varepsilon_{n}\delta_{n}}{2Nh_{n}|G|^{3w_{n}}}\right).

Hence we have

log|xi|logN+wnlog|G|nεnδn2Nhn|G|3wn.\log|x_{i}|\leq\log N+w_{n}\log|G|-\frac{n\varepsilon_{n}\delta_{n}}{2Nh_{n}|G|^{3w_{n}}}.

Since hn=O(n1α)h_{n}=O(n^{1-\alpha}) and εnδnnα+β\varepsilon_{n}\delta_{n}\geq n^{-\alpha+\beta}, there is a constant CC depending only on the proportionality constant in hnh_{n} such that for large enough nn we have εnδnhnCnβ1\frac{\varepsilon_{n}\delta_{n}}{h_{n}}\geq Cn^{\beta-1} so that nεnδn2Nhn|G|3wnCnβ2N|G|3wn\frac{n\varepsilon_{n}\delta_{n}}{2Nh_{n}|G|^{3w_{n}}}\geq\frac{Cn^{\beta}}{2N|G|^{3w_{n}}}

Since wn=o(logn)w_{n}=o(\log n), for large enough nn we have wnβlogn6log|G|w_{n}\leq\frac{\beta\log n}{6\log|G|} so that |G|3wn=e3wnlog|G|nβ/2|G|^{3w_{n}}=e^{3w_{n}\log|G|}\leq n^{\beta/2} and, for large enough nn, nεnδn2Nhn|G|3wnCnβ/22N\frac{n\varepsilon_{n}\delta_{n}}{2Nh_{n}|G|^{3w_{n}}}\geq\frac{Cn^{\beta/2}}{2N}.

Finally, since logN+wnlog|G|=o(logn)\log N+w_{n}\log|G|=o(\log n), we also have that for nn large enough, log|xi|C4Nnβ/2\log|x_{i}|\leq-\frac{C}{4N}n^{\beta/2} and |xi|exp(C4Nnβ/2)|x_{i}|\leq\exp\left(-\frac{C}{4N}n^{\beta/2}\right). In particular, for nn large enough,

i=1m|xi|mexp(C4Nnβ/2)nexp(C4Nnβ/2)log2.\sum_{i=1}^{m}|x_{i}|\leq m\exp\left(-\frac{C}{4N}n^{\beta/2}\right)\leq n\exp\left(-\frac{C}{4N}n^{\beta/2}\right)\leq\log 2.

By Lemma 4.12, we therefore have that for such nn,

||G|n+u[f(M)=(g1,,gn+u)]1|\displaystyle||G|^{n+u}\mathbb{P}[f(M)=(g_{1},\dots,g_{n+u})]-1| =|i=1m|G|#Qi[f(Mi)=gQi]|1|\displaystyle=\left|\prod_{i=1}^{m}|G|^{\#Q_{i}}\mathbb{P}[f(M_{i})=g_{Q_{i}}]|-1\right|
=|i=1m(1+xi)1|\displaystyle=\left|\prod_{i=1}^{m}(1+x_{i})-1\right|
2i=1m|xi|\displaystyle\leq 2\sum_{i=1}^{m}|x_{i}|
2nexp(C4Nnβ/2)\displaystyle\leq 2n\exp\left(-\frac{C}{4N}n^{\beta/2}\right)
=2nexp(C8Nnβ/2)exp(C8Nnβ/2).\displaystyle=2n\exp\left(-\frac{C}{8N}n^{\beta/2}\right)\cdot\exp\left(-\frac{C}{8N}n^{\beta/2}\right).

Since limn2nexp(C8Nnβ/2)=0\lim_{n\to\infty}2n\exp\left(-\frac{C}{8N}n^{\beta/2}\right)=0, the expression 2nexp(C8Nnβ/2)2n\exp\left(-\frac{C}{8N}n^{\beta/2}\right) is uniformly bounded above by some constant for all n0n\geq 0. Then the appropriate constant KK can be chosen so that

||G|n+u[f(M)=(g1,,gn+u)]1|Kexp(C8Nnβ/2),||G|^{n+u}\mathbb{P}[f(M)=(g_{1},\dots,g_{n+u})]-1|\leq K\exp\left(-\frac{C}{8N}n^{\beta/2}\right),

for all nn, as desired. ∎

4.3. Bounds for the rest of the maps

This gives results for the case when ff is a code, but we still need to account for non-codes. To do this, we will show that non-codes make up a negligible proportion of all maps VGV\to G and thus contribute only a small error term to the sum 𝔼[#Sur(coker(M),G)]\mathbb{E}[\#\operatorname{Sur}(\operatorname{coker}(M),G)]. However it turns out that splitting maps into codes and non-codes is not enough to get this bound. Instead, as in [Woo19], [NW22], and similar work, we will categorize non-codes by how far they are from being codes.

If DD is an integer with prime factorization ipiei\prod_{i}p_{i}^{e_{i}}, we write (D)=iei\ell(D)=\sum_{i}e_{i}.

Definition 4.14.

If V=(/a)nV=(\mathbb{Z}/a\mathbb{Z})^{n} and 𝒫\mathcal{P} is a partition of the “standard basis” of VV, the (𝒫,δ)(\mathcal{P},\delta)-depth of fHom(V,G)f\in\operatorname{Hom}(V,G) is the maximal positive DD such that there is a σ𝒫\sigma\subset\mathcal{P} with |σ|<(D)δn|\cup\sigma|<\ell(D)\delta n such that D=[G:f(Vσ)]D=[G:f(V_{\setminus\cup\sigma})], or 1 if there is no such DD.

We can count the number of ff that have given (𝒫,δ)(\mathcal{P},\delta)-depth:

Lemma 4.15.

If D>1D>1, then the number of fHom(V,G)f\in\operatorname{Hom}(V,G) of (𝒫,δ)(\mathcal{P},\delta)-depth DD is at most

K(n(D)δn1)2(D)δn|G|nDn+(D)δn,K\binom{n}{\lceil\ell(D)\delta n\rceil-1}2^{\ell(D)\delta n}|G|^{n}D^{-n+\ell(D)\delta n},

where KK is the number of subgroups of GG of index DD.

Proof.

For each ff of (𝒫,δ)(\mathcal{P},\delta)-depth DD, there is a σ𝒫\sigma\subset\mathcal{P} as described in Definition 4.14. There must be some set S[n]S\subset[n] with #S=(D)δn1\#S=\lceil\ell(D)\delta n\rceil-1 and σS\cup\sigma\subseteq S. There are (n(D)δn1)\binom{n}{\lceil\ell(D)\delta n\rceil-1} choices of SS, and for each choice of SS, there are certainly at most 2#S=2(D)δn12(D)δn2^{\#S}=2^{\lceil\ell(D)\delta n\rceil-1}\leq 2^{\ell(D)\delta n} choices of σ\cup\sigma. Since 𝒫\mathcal{P} is a partition, σ\cup\sigma uniquely determines σ\sigma, so there are at most 2(D)δn2^{\ell(D)\delta n} choices of σ\sigma for each choice of SS.

Now we count how many ff of (𝒫,δ)(\mathcal{P},\delta)-depth DD have each choice of σ\sigma, so fix σ\sigma. There are KK subgroups of GG of index DD, so there are KK options for f(Vσ)f(V_{\setminus\cup\sigma}).

Fix a subgroup HH of GG with index DD. We now count the number of ff with f(Vσ)Hf(V_{\setminus\cup\sigma})\subseteq H. There are at most |H|n|σ||H|^{n-|\cup\sigma|} maps from VσV_{\setminus\cup\sigma} to HH, and for each such map, there are at most |G||σ||G|^{|\cup\sigma|} homomorphisms from VV to GG which restrict appropriately. Hence, there are at most

|H|n|σ||G||σ|=|G|n|σ|Dn+|σ||G||σ|=|G|nDn+|σ||G|nDn+(D)δn|H|^{n-|\cup\sigma|}|G|^{|\cup\sigma|}=|G|^{n-|\cup\sigma|}D^{-n+|\cup\sigma|}|G|^{|\cup\sigma|}=|G|^{n}D^{-n+|\cup\sigma|}\leq|G|^{n}D^{-n+\ell(D)\delta n}

maps ff with f(Vσ)=Hf(V_{\setminus\cup\sigma})=H. Combined with the counts of choices of σ\sigma and subgroups of GG of index DD, we get the lemma. ∎

For non-codes, we do not get precise estimates on [f(M)=0]\mathbb{P}[f(M)=0], but we can get upper bounds.

Lemma 4.16.

Let r1r\geq 1 be an integer. Let GG be a finite abelian group and let aa be a multiple of the exponent of GG. Let NN be the number of subgroups of GG. Let ε>0\varepsilon>0 and δ>0\delta>0 be real numbers. Let V=(/a)nV=(\mathbb{Z}/a\mathbb{Z})^{n}. Let 𝒫={P1,,Pm}\mathcal{P}=\{P_{1},\dots,P_{m}\} be a partition of [n][n] and let =|𝒫|\ell=|\mathcal{P}|. Let fHom(V,G)f\in\operatorname{Hom}(V,G) have (𝒫,δ)(\mathcal{P},\delta)-depth D>1D>1 with [G:f(V)]<D[G:f(V)]<D.

Let MM be an n×rn\times r random matrix in VrV^{r} such that the matrices MPi,[r]M_{P_{i},[r]} are independent and ε\varepsilon-balanced as random elements of ((/a)#Pi)r((\mathbb{Z}/a\mathbb{Z})^{\#P_{i}})^{r}.

Then

[f(M)=0](1ε)(Dr|G|r+Nexp(εδn2N(D1|G|)3r))\mathbb{P}[f(M)=0]\leq(1-\varepsilon)\left(D^{r}|G|^{-r}+N\exp\left(-\frac{\varepsilon\delta n}{2N\ell(D^{-1}|G|)^{3r}}\right)\right)
Proof.

Since ff has (𝒫,δ)(\mathcal{P},\delta)-depth DD, there is a σ𝒫\sigma\subset\mathcal{P} with |σ|<(D)δn|\cup\sigma|<\ell(D)\delta n such that D=[G:f(Vσ)]D=[G:f(V_{\setminus\cup\sigma})]. Let f(Vσ)=:Hf(V_{\setminus\cup\sigma})=\colon H. Since [G:f(V)]<D[G:f(V)]<D, we cannot have that σ\sigma is empty.

Write f(M)=jσf(MPj,[r])+jσf(MPj,[r])f(M)=\sum_{j\notin\sigma}f(M_{P_{j},[r]})+\sum_{j\in\sigma}f(M_{P_{j},[r]}). So,

[f(M)=0]=[f(M)H][jσf(MPj,[r])=jσf(MPj,[r])|f(M)H].\mathbb{P}[f(M)=0]=\mathbb{P}[f(M)\in H]\mathbb{P}\left[\sum_{j\notin\sigma}f(M_{P_{j},[r]})=-\sum_{j\in\sigma}f(M_{P_{j},[r]})\ \middle|\ f(M)\in H\right].

We bound the two probabilities on the right side separately. Note that since jσf(MPj,[r])H\sum_{j\in\sigma}f(M_{P_{j},[r]})\in H, we have f(M)Hf(M)\in H exactly when jσf(MPj,[r])H\sum_{j\notin\sigma}f(M_{P_{j},[r]})\in H. Since [G:f(V)]<[G:H][G:f(V)]<[G:H], there must be some iσi\in\sigma such that f(MPi,[r])f(M_{P_{i},[r]}) reduces to a nonzero element of G/HG/H. Conditioning on all other MPk,[r]M_{P_{k},[r]} for kik\neq i, by the ε\varepsilon-balanced assumption we have that

[f(M)H]=[f(MPi,[r])jσ{k}f(MPj,[r])(modH)]1ε.\mathbb{P}\left[f(M)\in H\right]=\mathbb{P}\left[f(M_{P_{i},[r]})\equiv-\sum_{j\in\sigma\setminus\{k\}}f(M_{P_{j},[r]})\pmod{H}\right]\leq 1-\varepsilon.

For the second probability, let 𝒫\mathcal{P}^{\prime} be the partition of [n]σ[n]\setminus\cup\sigma induced by 𝒫\mathcal{P}. Notice that f|Vσf|_{V_{\setminus\cup\sigma}} is a 𝒫\mathcal{P}^{\prime}-code of distance δn\delta n. Indeed, suppose there is some τ𝒫\tau\subset\mathcal{P}^{\prime} with |τ|<δn|\tau|<\delta n inducing some τ𝒫\tau^{\prime}\subset\mathcal{P} with f(V(στ))Hf(V_{\setminus\cup(\sigma\cup\tau)})\neq H. Then the image of f|V(στ)f|_{V_{\setminus\cup(\sigma\cup\tau)}} would have index strictly greater than DD, contradicting maximality of DD.

Now we can apply Lemma 4.11 to the submatrix M[n]σ,[r]M_{[n]\setminus\cup\sigma,[r]} and the code ff mapping it into HrH^{r}. If NN^{\prime} is the number of subgroups of HH and =|𝒫|\ell^{\prime}=|\mathcal{P}^{\prime}|, then conditioning on MPj,[r]M_{P_{j},[r]} for jσj\in\sigma gives

[jσf(MPj,[r])=jσf(MPj,[r])|f(M)H]\displaystyle\mathbb{P}\left[\sum_{j\notin\sigma}f(M_{P_{j},[r]})=-\sum_{j\in\sigma}f(M_{P_{j},[r]})\ \middle|\ f(M)\in H\right] |H|r+Nexp(εδn2N|H|3r)\displaystyle\leq|H|^{-r}+N^{\prime}\exp\left(-\frac{\varepsilon\delta n}{2N^{\prime}\ell^{\prime}|H|^{3r}}\right)
Dr|G|r+Nexp(εδn2N(D1|G|)3r),\displaystyle\leq D^{r}|G|^{-r}+N\exp\left(-\frac{\varepsilon\delta n}{2N\ell(D^{-1}|G|)^{3r}}\right),

and the lemma follows. ∎

Finally, we use Lemma 4.12 again to get a bound for the full n×(n+u)n\times(n+u) matrix:

Lemma 4.17.

Let u0u\geq 0 be an integer. Let GG be a finite abelian group and let aa be a multiple of the exponent of GG. Let (wn)n,(hn)n,(δn)n,(εn)n(w_{n})_{n},(h_{n})_{n},(\delta_{n})_{n},(\varepsilon_{n})_{n} be sequences of real numbers such that wn=o(logn)w_{n}=o(\log n), hn=O(n1α)h_{n}=O(n^{1-\alpha}), and εnδnnα+β\varepsilon_{n}\delta_{n}\geq n^{-\alpha+\beta} for some 0<βα10<\beta\leq\alpha\leq 1.

For a natural number nn, let V=(/a)nV=(\mathbb{Z}/a\mathbb{Z})^{n}. Let MM be an (wn,hn,εn)(w_{n},h_{n},\varepsilon_{n})-balanced n×n+un\times n+u random matrix with entries in /a\mathbb{Z}/a\mathbb{Z}. Let 𝒫\mathcal{P} be the row partition associated to MM and let fHom(V,G)f\in\operatorname{Hom}(V,G) have (𝒫,δn)(\mathcal{P},\delta_{n})-depth D>1D>1, with [G:f(V)]<D[G:f(V)]<D.

Then there is a constant K>0K>0 depending only on uu, GG, α\alpha, β\beta, and the sequences hnh_{n}, wnw_{n} such that for all nn,

[f(M)=0]Kexp(εnnlogn)Dn|G|n.\mathbb{P}[f(M)=0]\leq K\exp\left(-\varepsilon_{n}\frac{n}{\log n}\right)D^{n}|G|^{-n}.
Proof.

Let 𝒬\mathcal{Q} be the column partition for MM as in the definition of (wn,hn,εn)(w_{n},h_{n},\varepsilon_{n})-balanced. Let Mi=M[n],QiM_{i}=M_{[n],Q_{i}} for each ii. By independence,

[f(M)=0]=i[f(Mi)=0].\mathbb{P}[f(M)=0]=\prod_{i}\mathbb{P}[f(M_{i})=0].

For each ii, let xi=|G|#QiD#Qi1εn[f(Mi)=0]1x_{i}=\frac{|G|^{\#Q_{i}}D^{-\#Q_{i}}}{1-\varepsilon_{n}}\mathbb{P}[f(M_{i})=0]-1. By Lemma 4.16, we have

max{0,xi}\displaystyle\max\{0,x_{i}\} N|G|#QiD#Qiexp(nεnδn2Nhn(D1|G|)3#Qi)\displaystyle\leq N|G|^{\#Q_{i}}D^{-\#Q_{i}}\exp\left(-\frac{n\varepsilon_{n}\delta_{n}}{2Nh_{n}(D^{-1}|G|)^{3\#Q_{i}}}\right)
N|G|wnDwnexp(nεnδn2Nhn(D1|G|)3wn).\displaystyle\leq N|G|^{w_{n}}D^{-w_{n}}\exp\left(-\frac{n\varepsilon_{n}\delta_{n}}{2Nh_{n}(D^{-1}|G|)^{3w_{n}}}\right).

By the same argument as in the proof of Lemma 4.13, there is some constant CC depending only on α\alpha and the sequence hnh_{n} such that for large enough nn (where “large enough” depends on uu, GG, α\alpha, β\beta, and the sequences hnh_{n} and wnw_{n}), we have

i=1mmax{0,xi}nexp(C4Nnβ/2)log2.\sum_{i=1}^{m}\max\{0,x_{i}\}\leq n\exp\left(\frac{C}{4N}n^{\beta/2}\right)\leq\log 2.

By Lemma 4.12, we therefore have that for such nn,

(D1|G|)n+u(1εn)#𝒬[f(M)=0]1\displaystyle\frac{(D^{-1}|G|)^{n+u}}{(1-\varepsilon_{n})^{\#\mathcal{Q}}}\mathbb{P}[f(M)=0]-1 =i=1m(D1|G|)#Qi1εn[f(Mi)=0]1\displaystyle=\prod_{i=1}^{m}\frac{(D^{-1}|G|)^{\#Q_{i}}}{1-\varepsilon_{n}}\mathbb{P}[f(M_{i})=0]-1
=i=1m(1+xi)1\displaystyle=\prod_{i=1}^{m}(1+x_{i})-1
2i=1mmax{0,xi}\displaystyle\leq 2\sum_{i=1}^{m}\max\{0,x_{i}\}
2nexp(C4Nnβ/2)\displaystyle\leq 2n\exp\left(-\frac{C}{4N}n^{\beta/2}\right)
=2nexp(C8Nnβ/2)exp(C8Nnβ/2).\displaystyle=2n\exp\left(-\frac{C}{8N}n^{\beta/2}\right)\cdot\exp\left(-\frac{C}{8N}n^{\beta/2}\right).

Since limn2nexp(C8Nnβ/2)=0\lim_{n\to\infty}2n\exp\left(-\frac{C}{8N}n^{\beta/2}\right)=0, the expression 2nexp(C8Nnβ/2)2n\exp\left(-\frac{C}{8N}n^{\beta/2}\right) is uniformly bounded above by some constant for all n0n\geq 0. Then the appropriate constant KK^{\prime} can be chosen so that

(D1|G|)n+u(1εn)#𝒬[f(M)=0]1Kexp(C8Nnβ/2),\frac{(D^{-1}|G|)^{n+u}}{(1-\varepsilon_{n})^{\#\mathcal{Q}}}\mathbb{P}[f(M)=0]-1\leq K^{\prime}\exp\left(-\frac{C}{8N}n^{\beta/2}\right),

for all nn. Hence we have

[f(M)=0]\displaystyle\mathbb{P}[f(M)=0] Dn+u|G|nu(1εn)#𝒬(1+Kexp(C8Nnβ/2))\displaystyle\leq D^{n+u}|G|^{-n-u}(1-\varepsilon_{n})^{\#\mathcal{Q}}\left(1+K^{\prime}\exp\left(-\frac{C}{8N}n^{\beta/2}\right)\right)
Dn+u|G|nuexp(εn#𝒬)(1+Kexp(C8Nnβ/2))\displaystyle\leq D^{n+u}|G|^{-n-u}\exp(-\varepsilon_{n}\#\mathcal{Q})\left(1+K^{\prime}\exp\left(-\frac{C}{8N}n^{\beta/2}\right)\right)
(K+1)Dn+u|G|nuexp(εn#𝒬).\displaystyle\leq(K^{\prime}+1)D^{n+u}|G|^{-n-u}\exp(-\varepsilon_{n}\#\mathcal{Q}).

The lemma follows from the fact that for large enough nn, we have wnlognw_{n}\leq\log n, so #𝒬nwnnlogn\#\mathcal{Q}\geq\frac{n}{w_{n}}\geq\frac{n}{\log n}. ∎

4.4. Computing the moments

Finally, we can combine all these results to compute the limiting moments for cokernels of (wn,hn,εn)(w_{n},h_{n},\varepsilon_{n})-balanced random matrices. The most delicate part of this proof is the part where we handle the non-codes. This will involve a careful choice of the sequence δn\delta_{n}.

Theorem 4.18.

Let u0u\geq 0 be an integer. Let GG be a finite abelian group and let aa be a multiple of the exponent of GG (including zero). Let (wn)n,(hn)n(w_{n})_{n},(h_{n})_{n} be sequences of real numbers such that wn=o(logn)w_{n}=o(\log n), hn=O(n1α)h_{n}=O(n^{1-\alpha}), and εnnβ\varepsilon_{n}\geq n^{-\beta} for some 0<α10<\alpha\leq 1 and 0<β<α/20<\beta<\alpha/2.

Then there are c,K,γ>0c,K,\gamma>0 such that the following holds for every sufficiently large natural number nn. Let MM be an (wn,hn,εn)(w_{n},h_{n},\varepsilon_{n})-balanced n×(n+u)n\times(n+u) random matrix with entries in /a\mathbb{Z}/a\mathbb{Z}. Then

|𝔼[#Sur(coker(M),G)]|G|u|Kecnγ.|\mathbb{E}[\#\operatorname{Sur}(\operatorname{coker}(M),G)]-|G|^{-u}|\leq Ke^{-cn^{\gamma}}.
Proof.

Let V=(/a)nV=(\mathbb{Z}/a\mathbb{Z})^{n}. Following the discussion at the beginning of this section, we have

𝔼[#Sur(coker(M),G)]=fSur(V,G)[f(M)=0]\mathbb{E}[\#\operatorname{Sur}(\operatorname{coker}(M),G)]=\sum_{f\in\operatorname{Sur}(V,G)}\mathbb{P}[f(M)=0]

Let 𝒫,𝒬\mathcal{P},\mathcal{Q} be the row and column partitions witnessing the (wn,hn,εn)(w_{n},h_{n},\varepsilon_{n})-balancedness of MM.

Let δn=nα/2\delta_{n}=n^{-\alpha/2}. Note that then εnδnnβα/2\varepsilon_{n}\delta_{n}\geq n^{-\beta-\alpha/2} with βα/2>α-\beta-\alpha/2>-\alpha, so δn\delta_{n} satisfies the conditions for Lemmas 4.13 and 4.17.

For notational convenience, we will allow KK to change in each line as long as it remains a constant depending only on a,u,α,β,(hn)n,(wn)n,Ga,u,\alpha,\beta,(h_{n})_{n},(w_{n})_{n},G.

We have

|𝔼[#Sur(coker(M),G)]1|G|u|\displaystyle\left|\mathbb{E}[\#\operatorname{Sur}(\operatorname{coker}(M),G)]-\frac{1}{|G|^{u}}\right|
=|fSur(V,G)[f(M)=0]1|G|u|\displaystyle\qquad\qquad=\left|\sum_{f\in\operatorname{Sur}(V,G)}\mathbb{P}[f(M)=0]-\frac{1}{|G|^{u}}\right|
=|fSur(V,G)[f(M)=0]fHom(V,G)1|G|n+u|\displaystyle\qquad\qquad=\left|\sum_{f\in\operatorname{Sur}(V,G)}\mathbb{P}[f(M)=0]-\sum_{f\in\operatorname{Hom}(V,G)}\frac{1}{|G|^{n+u}}\right|
(1) fSur(V,G)f code of distance nδn|[f(M¯)=0]1|G|n+u|\displaystyle\qquad\qquad\leq\sum_{\begin{subarray}{c}f\in\operatorname{Sur}(V,G)\\ f\text{ code of distance }n\delta_{n}\end{subarray}}\left|\mathbb{P}[f(\bar{M})=0]-\frac{1}{|G|^{n+u}}\right|
(2) +D>1D|G|fSur(V,G)f of (𝒫,δn)-depth D[f(M¯)=0]\displaystyle\qquad\qquad\qquad+\sum_{\begin{subarray}{c}D>1\\ D\mid|G|\end{subarray}}\sum_{\begin{subarray}{c}f\in\operatorname{Sur}(V,G)\\ f\text{ of }(\mathcal{P},\delta_{n})\text{-depth }D\end{subarray}}\mathbb{P}[f(\bar{M})=0]
(3) +D>1D|G|fSur(V,G)f of (𝒫,δn)-depth D1|G|n+u\displaystyle\qquad\qquad\qquad+\sum_{\begin{subarray}{c}D>1\\ D\mid|G|\end{subarray}}\sum_{\begin{subarray}{c}f\in\operatorname{Sur}(V,G)\\ f\text{ of }(\mathcal{P},\delta_{n})\text{-depth }D\end{subarray}}\frac{1}{|G|^{n+u}}
(4) +fHom(V,G)Sur(V,G)1|G|n+u\displaystyle\qquad\qquad\qquad+\sum_{f\in\operatorname{Hom}(V,G)\setminus\operatorname{Sur}(V,G)}\frac{1}{|G|^{n+u}}

Wood showed in the proof of [Woo19, Theorem 2.9] that (4) is bounded above by Kenlog2Ke^{-n\log 2}. By Lemma 4.13, we can bound (1):

fSur(V,G)f code of distance nδn|[f(M)=0]1|G|n+u|\displaystyle\sum_{\begin{subarray}{c}f\in\operatorname{Sur}(V,G)\\ f\text{ code of distance }n\delta_{n}\end{subarray}}\left|\mathbb{P}[f(M)=0]-\frac{1}{|G|^{n+u}}\right| fSur(V,G)f code of distance nδnKexp(cnγ)|G|n+u\displaystyle\leq\sum_{\begin{subarray}{c}f\in\operatorname{Sur}(V,G)\\ f\text{ code of distance }n\delta_{n}\end{subarray}}\frac{K\exp(-cn^{\gamma})}{|G|^{n+u}}
|G|nKexp(cnγ)|G|n+u\displaystyle\leq|G|^{n}\frac{K\exp(-cn^{\gamma})}{|G|^{n+u}}
=Kexp(cnγ).\displaystyle=K\exp(-cn^{\gamma}).

To bound (2) and (3) we use Lemma 4.15. For each D>1D>1, there are at most

K(n(D)nδn1)2(D)nδn|G|nDn+(D)nδnK\binom{n}{\lceil\ell(D)n\delta_{n}\rceil-1}2^{\ell(D)n\delta_{n}}|G|^{n}D^{-n+\ell(D)n\delta_{n}}

maps of (𝒫,δn)(\mathcal{P},\delta_{n})-depth DD. A standard inequality says that (nk)(nek)k\binom{n}{k}\leq\left(\frac{ne}{k}\right)^{k}, so for (D)nδn2\lceil\ell(D)n\delta_{n}\rceil\geq 2 (which is the case for nn large enough, independent of DD)

(n(D)nδn1)\displaystyle\binom{n}{\lceil\ell(D)n\delta_{n}\rceil-1} (ne(D)nδn1)(D)nδn1\displaystyle\leq\left(\frac{ne}{\lceil\ell(D)n\delta_{n}\rceil-1}\right)^{\lceil\ell(D)n\delta_{n}\rceil-1}
(2ne(D)nδn)(D)nδn\displaystyle\leq\left(\frac{2ne}{\ell(D)n\delta_{n}}\right)^{\ell(D)n\delta_{n}}
=(2e(D)δn)(D)nδn\displaystyle=\left(\frac{2e}{\ell(D)\delta_{n}}\right)^{\ell(D)n\delta_{n}}
=exp((D)nδn(1+log2log(D)logδn)).\displaystyle=\exp\left(\ell(D)n\delta_{n}\left(1+\log 2-\log\ell(D)-\log\delta_{n}\right)\right).

Hence, the number of maps of (𝒫,δn)(\mathcal{P},\delta_{n})-depth DD is at most

K|G|nDnexp((D)nδn(log4eD(D)logδn))\displaystyle K|G|^{n}D^{-n}\exp\left(\ell(D)n\delta_{n}\left(\log\frac{4eD}{\ell(D)}-\log\delta_{n}\right)\right) =K|G|nexp((D)nδn(log4eD(D)logδn)nlogD)\displaystyle=K|G|^{n}\exp\left(\ell(D)n\delta_{n}\left(\log\frac{4eD}{\ell(D)}-\log\delta_{n}\right)-n\log D\right)
K|G|nexp((|G|)nδn(log4e|G|(|G|)logδn)nlog2)\displaystyle\leq K|G|^{n}\exp\left(\ell(|G|)n\delta_{n}\left(\log\frac{4e|G|}{\ell(|G|)}-\log\delta_{n}\right)-n\log 2\right)

Since limδ0δlogδ=0\lim_{\delta\to 0}\delta\log\delta=0 and δn0\delta_{n}\to 0 as nn\to\infty, for large enough nn (depending only on β\beta and |G||G|) we have (|G|)δn(log4e|G|(|G|)logδn)12log2\ell(|G|)\delta_{n}\left(\log\frac{4e|G|}{\ell(|G|)}-\log\delta_{n}\right)\leq\frac{1}{2}\log 2, which means that for large enough nn,

D|G|fSur(V,G)f of (𝒫,δn)-depth D1|G|n+u\displaystyle\sum_{D\mid|G|}\sum_{\begin{subarray}{c}f\in\operatorname{Sur}(V,G)\\ f\text{ of }(\mathcal{P},\delta_{n})\text{-depth }D\end{subarray}}\frac{1}{|G|^{n+u}} D|G|K|G|uexp((|G|)nδn(log4e|G|(|G|)logδn)nlog2)\displaystyle\leq\sum_{D\mid|G|}K|G|^{-u}\exp\left(\ell(|G|)n\delta_{n}\left(\log\frac{4e|G|}{\ell(|G|)}-\log\delta_{n}\right)-n\log 2\right)
D|G|Kexp(log22n)\displaystyle\leq\sum_{D\mid|G|}K\exp\left(-\frac{\log 2}{2}n\right)
Kexp(log22n),\displaystyle\leq K\exp\left(-\frac{\log 2}{2}n\right),

bounding (3) as desired.

Finally, we need to bound (2). From Lemma 4.17, we have that if ff has (𝒫,δn)(\mathcal{P},\delta_{n})-depth DD,

[f(M)=0]Kexp(εnnlogn)Dn|G|n,\mathbb{P}[f(M)=0]\leq K\exp\left(-\varepsilon_{n}\frac{n}{\log n}\right)D^{n}|G|^{-n},

which means

fSur(V,G)f of (𝒫,δn)-depth D[f(M)=0]\displaystyle\sum_{\begin{subarray}{c}f\in\operatorname{Sur}(V,G)\\ f\text{ of }(\mathcal{P},\delta_{n})\text{-depth }D\end{subarray}}\mathbb{P}[f(M)=0] Kexp((D)nδn(log4eD(D)logδn)εnnlogn)\displaystyle\leq K\exp\left(\ell(D)n\delta_{n}\left(\log\frac{4eD}{\ell(D)}-\log\delta_{n}\right)-\varepsilon_{n}\frac{n}{\log n}\right)
Kexp((|G|)n1α/2(log4e|G|(|G|)+α2logn)n1βlogn).\displaystyle\leq K\exp\left(\ell(|G|)n^{1-\alpha/2}\left(\log\frac{4e|G|}{\ell(|G|)}+\frac{\alpha}{2}\log n\right)-\frac{n^{1-\beta}}{\log n}\right).

Since β<α/2\beta<\alpha/2, we have that n1α/2(logn)2=o(n1β)n^{1-\alpha/2}(\log n)^{2}=o(n^{1-\beta}), so

limn(|G|)n1α/2(log4e|G|(|G|)+α2logn)n1β/logn=0.\lim_{n\to\infty}\frac{\ell(|G|)n^{1-\alpha/2}\left(\log\frac{4e|G|}{\ell(|G|)}+\frac{\alpha}{2}\log n\right)}{n^{1-\beta}/\log n}=0.

Hence for large enough nn (depending only on GG, α\alpha, and β\beta), we have (|G|)n1α/2(log4e|G|(|G|)+α2logn)12n1βlogn\ell(|G|)n^{1-\alpha/2}\left(\log\frac{4e|G|}{\ell(|G|)}+\frac{\alpha}{2}\log n\right)\leq\frac{1}{2}\frac{n^{1-\beta}}{\log n} and

fSur(V,G)f of (𝒫,δn)-depth D[f(M)=0]Kexp(n1β2logn).\sum_{\begin{subarray}{c}f\in\operatorname{Sur}(V,G)\\ f\text{ of }(\mathcal{P},\delta_{n})\text{-depth }D\end{subarray}}\mathbb{P}[f(M)=0]\leq K\exp\left(-\frac{n^{1-\beta}}{2\log n}\right).

For the same reason, for nn large enough (depending only on β\beta) we have

fSur(V,G)f of (𝒫,δn)-depth D[f(M)=0]Kexp(12n1β2),\sum_{\begin{subarray}{c}f\in\operatorname{Sur}(V,G)\\ f\text{ of }(\mathcal{P},\delta_{n})\text{-depth }D\end{subarray}}\mathbb{P}[f(M)=0]\leq K\exp\left(-\frac{1}{2}n^{\frac{1-\beta}{2}}\right),

which means

D>1D|G|fSur(V,G)f of (𝒫,δn)-depth D[f(M)=0]Kexp(12n1β2),\sum_{\begin{subarray}{c}D>1\\ D\mid|G|\end{subarray}}\sum_{\begin{subarray}{c}f\in\operatorname{Sur}(V,G)\\ f\text{ of }(\mathcal{P},\delta_{n})\text{-depth }D\end{subarray}}\mathbb{P}[f(M)=0]\leq K\exp\left(-\frac{1}{2}n^{\frac{1-\beta}{2}}\right),

giving us a bound on (2).

Finally, choose cc and γ\gamma appropriately to obtain the desired result. ∎

Acknowledgements

The author was supported by the NSF Graduate Research Fellowship Program, the Caltech Summer Undergraduate Research Fellowship program, and the Samuel P. and Frances Krown SURF Fellowship. The author thanks Melanie Wood and Omer Tamuz for mentorship and Alexander Gorokhovsky, Seth Berman, Sandra O’Neill, and Hoi Nguyen for insightful conversations. The author also thanks Gilyoung Cheong, Yifeng Huang, Hoi Nguyen, Roger Van Peski, Will Sawin, and Melanie Wood for helpful comments on an earlier draft of this manuscript.

References

  • [CK24] Gilyoung Cheong and Nathan Kaplan “Generalizations of results of Friedman and Washington on cokernels of random p-adic matrices” In Journal of Algebra 604.C, 2024 DOI: 10.1016/j.jalgebra.2022.03.035
  • [FW89] Eduardo Friedman and Lawrence C. Washington “On the distribution of divisor class groups of curves over a finite field” In Proceedings of the International Number Theory Conference held at Université Laval, July 5-18, 1987 Berlin, New York: De Gruyter, 1989, pp. 227–239 DOI: 10.1515/9783110852790.227
  • [Més20] András Mészáros “The Distribution of Sandpile Groups of Random Regular Graphs” In Transactions of the American Mathematical Society 373.9, 2020, pp. 6529–6594 DOI: 10.1090/tran/8127
  • [NO15] Hoi H. Nguyen and Sean O’Rourke “On the Concentration of Random Multilinear Forms and the Universality of Random Block Matrices” In Probability Theory and Related Fields 162.1, 2015, pp. 97–154 DOI: 10.1007/s00440-014-0567-7
  • [NV24] Hoi H. Nguyen and Roger Van Peski “Universality for cokernels of random matrix products” In Advances in Mathematics 438, 2024, pp. 109451 DOI: 10.1016/j.aim.2023.109451
  • [NW22] Hoi H. Nguyen and Melanie Matchett Wood “Random Integral Matrices: Universality of Surjectivity and the Cokernel” In Inventiones mathematicae 228.1, 2022, pp. 1–76 DOI: 10.1007/s00222-021-01082-w
  • [Sal04] Laurent Saloff-Coste “Random Walks on Finite Groups” Series Title: Encyclopaedia of Mathematical Sciences In Probability on Discrete Structures 110 Berlin, Heidelberg: Springer Berlin Heidelberg, 2004, pp. 263–346 DOI: 10.1007/978-3-662-09444-0˙5
  • [SZ07] Laurent Saloff-Coste and Jesse Zúñiga “Convergence of some time inhomogeneous Markov chains via spectral techniques” In Stochastic Processes and their Applications 117.8, 2007, pp. 961–979 DOI: 10.1016/j.spa.2006.11.004
  • [TVK10] Terence Tao, Van Vu and Manjunath Krishnapur “Random matrices: Universality of ESDs and the circular law” In The Annals of Probability 38.5 Institute of Mathematical Statistics, 2010, pp. 2023–2065 DOI: 10.1214/10-AOP534
  • [Woo14] Melanie Matchett Wood “The distribution of sandpile groups of random graphs” In Journal of the American Mathematical Society 30.4, 2014
  • [Woo19] Melanie Matchett Wood “Random integral matrices and the Cohen-Lenstra heuristics” In American Journal of Mathematics 141.2, 2019, pp. 383–398
  • [Woo23] Melanie Matchett Wood “Probability Theory for Random Groups Arising in Number Theory” arXiv, 2023 DOI: 10.48550/arXiv.2301.09687