96
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Sandpile groups of Cayley graphs of 𝔽r2

, , &
Received 06 Aug 2023, Accepted 19 Apr 2024, Published online: 15 May 2024

Abstract

The sandpile group of a connected graph G, defined to be the torsion part of the cokernel of the graph Laplacian, is a subtle graph invariant with combinatorial, algebraic, and geometric descriptions. Extending and improving previous works on the sandpile group of hypercubes, we study the sandpile groups of the Cayley graphs of F2r, focusing on their poorly understood Sylow-2 component. We find the number of Sylow-2 cyclic factors for “generic” Cayley graphs and deduce a bound for the non-generic ones. Moreover, we provide a sharp upper bound for their largest Sylow-2 cyclic factors. In the case of hypercubes, we give exact formulae for the largest n–1 Sylow-2 cyclic factors. Some key ingredients of our work include the natural ring structure on these sandpile groups from representation theory, and calculation of the 2-adic valuations of binomial sums via the combinatorics of carries.

2020 MATHEMATICS SUBJECT CLASSIFICATION:

1 Introduction and notation

Let G=(V,E) be a connected graph on n (ordered) vertices with no loops. The Laplacian L(G) of graph G is the n×n matrix with entries L(G)u,v={deg(u)u=vm(u,v)uv,where m(u, v) is the number of edges between u and v. The Laplacian is an important matrix associated with the graph that has been studied from different perspectives, ranging from more theoretical aspects to real life algorithms [4, 16]. Among these perspectives, as L(G) is an integer matrix, we can study its properties as a linear map of Z-modules L(G):ZnZn.

From the definition of L(G) and the assumption on G, kerZL(G) is spanned by the vector (1,1,,1)T. Therefore Im L(G) is some codimension 1 sublattice of Zn. It follows that the cokernel can be written as coker L(G)=Zn/Im L(G)ZK(G),where K(G) is a finite abelian group, known as the sandpile group (also critical group or Jacobian in the literature) of G. Kirchhoff’s Matrix–Tree Theorem shows that the size of K(G) is the number of spanning trees of G. K(G) is a subtle isomorphism invariant of a graph [Citation7, Citation12], and is our main object of study. We are interested in understanding K(G) for Cayley graphs of the group F2r, with an emphasis on their Sylow-2 subgroups.

A motivation for studying this family of graphs is their connection with representation theory. Roughly speaking (see Section 1.1 for the details in our setting and see [Citation8] for the more general case), for a Cayley graph G of a finite abelian group and a prime p that does not divide the order of the group, one can use the complex character table of the group to diagonalize L(G) so that its eigenvalues are expressed as character sums, then an algebraic argument shows that one can read off the p-adic valuations of the invariant factors of K(G) from these eigenvalues. Further generalizations have been studied for the McKay–Cartan matrices of representations of finite groups [Citation3, Citation9], and representations of finite-dimensional Hopf algebras [Citation10]. However, in our setting, the strategy fails when p = 2, which echoes the difficulties in modular representation theory.Footnote1

Nevertheless, partial progress has been made along this direction, particularly for the hypercube graphs Qn, which are the Cayley graphs of F2n with respect to the standard basis {e1,,en}F2n. In 2003, Bai [Citation2] determined the Sylow-p subgroups of K(Qn) for p2 with an elementary argument, and derived formulae for the number of Sylow-2 cyclic factors and the number of Z/2Z-factors in K(Qn). In 2015, Chandler et al. [Citation6] determined the cokernel of the adjacency matrix (known as the Smith group) of Qn in terms of n. Anzis and Prasad [Citation1] made the next step in 2016 by bounding the largest Sylow-2 cyclic factor of K(Qn).

We also note the other work by Chandler et al. [Citation5], where they determined the critical groups of Paley graphs, a special family of Cayley graphs of Fq for prime powers q=pk1(mod4). They were able to approach the Sylow-p subgroup of the critical group via Jacobi sums, and by applying Stickelberger’s theorem, they further reduced the problem to the combinatorics of “carries” when adding two p-ary numbers. Our approach of Theorem 1.4 thus has a similar flavor of their work.

We now summarize our contributions in this paper. It is worth emphasizing that many of our arguments are based on the natural ring structure of K(G) coming from representation theory [Citation3] and the corresponding polynomial algebra. We begin by defining a generic set of generators M for a Cayley graph of F2r as set of generators whose sum is nonzero (see Definition 2.1). For example, hypercubes have generic generating sets. We then have the following result that generalizes Bai’s result [Citation2, Theorem 1.1].

Theorem 1.1.

Suppose that M is generic. Then the number of Sylow-2 cyclic factors of K(G(F2r,M)) is 2r11.

On the other hand, we provide a qualitative result for non-generic cases that complements Theorem 1.1:

Theorem 1.2.

Suppose that M is not generic. Then the number of Sylow-2 cyclic factors of K(G(F2r,M)) is strictly larger than 2r11.

We then investigate the size of the largest Sylow-2 cyclic factor in K(G(F2r,M)). Suppose that K(G) has the invariant factor decomposition K(G)=i=1dZ/ci(G)Z, where cd(G)|cd1(G)|cd2(G)||c1(G), thus c1(G) is the largest cyclic factor. Define v2(x), the 2-adic valuation of x, as the largest n such that 2n|x. We adopt the methods of Anzis and Prasad to extend and improve their upper bound to arbitrary Cayley graphs G of F2r.

Theorem 1.3.

The largest cyclic factor Z/c1(G)Z of K(G(F2r,M)) satisfies v2(c1(G))log2n+r1,where n is the number of generators in M.

In the case of Qn, we go further to explicitly determine the top cyclic factor.

Theorem 1.4.

For n2, the largest cyclic factor Z/c1(Qn)Z of K(Qn) satisfies v2(c1(Qn))=max{maxx<n{v2(x)+x},v2(n)+n1}.

We then continue to determine the 2nd through (n1)th cyclic factors, and conjecture a formula for the nth factor (see Theorem 4.2 and Conjecture 4.14). Finally, we conclude with some remaining conjectures about the structure of K(G) in Section 5.

Remark 1.5.

Theorem 1.2 was posed as a conjecture in an earlier version of this paper. Iga et al. proved partial results of the conjecture [Citation11] (see the discussion at the end of Section 2 for more background) before the conjecture was solved in full generality in the current version.

1.1 Background and previous results

We first define what a Cayley graph is in our context. Given a multiset of nonzero generators M=(||v1vn||)of F2r, we form the Cayley graph G=G(F2r,M) with vertex set V=F2r and multiedge set {(w,w+vi)} for wV and viM. G is connected because M is a generating set, and there are no loops since all vi0. As addition is performed in F2r, we also have w+vi+vi=w+2vi=w. Therefore, we can think of this graph as undirected. If we index the matrix representation of L(G) by the binary tuples u,vF2r then we can say that L(G)u,v={nu=v(# of generators, vi, such that u+vi=v)uv since G is an n-regular graph.

Example 1.6.

We give three examples of Cayley graphs in the case r = 3.

  1. The cube Q3 with M=(100010001) is shown in .

  2. Define Q3,vert to be the Cayley graph with associated generating set

    M=(100001000011).

    Then the graph is a cube with two sets of vertical edges shown in .

  3. We can recover the complete graph K8 by setting

    M=(100101101011010010111).

Fig. 1 Examples of Cayley graphs in dimension r = 3.

Fig. 1 Examples of Cayley graphs in dimension r = 3.

We are interested in K(G) for such graphs. Kirchhoff’s Matrix–Tree Theorem tells us that if we label the eigenvalues of L(G) as 0=λ1<λ2λ3λ2r, then |K(G)|=12ri=22r|λi|, which is also the number of spanning trees of G [Citation15]. On the other hand, by the structure theorem for finite abelian groups, K(G)pe1(Z/peZ)m(pe), where m(pe) is the multiplicity of Z/peZ in K(G). Thus, we can try to determine the group prime by prime in terms of its Sylow subgroups. We note that, while eigenvalues are useful in the study of K(G) for Cayley graphs G, the spectrum of L(G) provides rather little information about K(G) in general; some of these connections can be found in [Citation14].

We detail some basic properties about the sandpile group of an arbitrary Cayley graph of F2r. As mentioned in the introduction, much of this generalizes to finite abelian groups with slightly more complicated algebraic formalism, cf. [Citation8]. Fixing r, when considering matrices over R, there is an eigenbasis for all of L(G(F2r,M)) simultaneously.

Definition 1.7.

For uF2r, define fu=vF2r(1)u·vevwhere ev is the standard basis vector (0,,0,1,0,,0)TR2r with the only 1 in the v-coordinate.

Lemma 1.8.

eu’s can be expressed in terms of fv’s as follows: eu=12rvF2r(1)u·vfv.

Hence, {fu} is a basis for R2r. In fact, it is an orthogonal eigenbasis for any generating set M, with the eigenvalue corresponding to fu being λu,M=ni=1n(1)u·vi.

Proof.

For a proof of the second statement and a more thorough exposition of the setting of this problem, see [Citation15, Chapter 2]. We only prove the first statement here.

We have v(1)u·vfv=v(1)u·v[w(1)v·wew]=w[v(1)(u+w)·v]ew. For w = u, all (1)(u+w)·v’s are 1, so the coefficient of ew is 2r; for wu, say they differ over the i-th coordinate, then we can pair up the elements of F2r by flipping their i-th coordinate, the sum of (1)(u+w)·v associated with a pair is 0, so the coefficient of ew is 0. □

One can use this information to determine the Sylow-p structure for odd primes p. Lemma 1.8 implies we can diagonalize L(G) to a matrix D=diag(λu,M,uF2r) using matrices whose entries are from Z[12], which provides a matrix equivalence that relates D to the Smith Normal Form of L(G) over Z[12]. This implies the following proposition.

Proposition 1.9.

For any odd prime p2, SylpK(G)=Sylp(uF2r{0}Z/λu,MZ).

Thus, we have a nice description of the Sylow-p subgroups for p2 in terms of the eigenvalues, but the Sylow-2 subgroups exhibit a more irregular structure.

Example 1.10.

These are the sandpile groups for our 3 examples in the case r = 3:

  1. The eigenvalues are (0,2,2,2,4,4,4,6) and K(Q3)=Z/2Z(Z/8Z)2Z/3Z

  2. The eigenvalues are (0,2,2,4,4,6,6,8) and K(Q3,vert)=(Z/4Z)2Z/16Z(Z/3Z)2,

  3. The eigenvalues are (0,8,8,8,8,8,8) and K(K8)=(Z/8Z)6.

In order to deal with the case p = 2, we adopt the approach of Benkart, Klivans, and Reiner [Citation3] that gives a natural ring structure on K(G):

Proposition 1.11.

For G=G(F2r,M), we have an isomorphism of Z-modules cokerL(G)K(G)ZZ[x1,x2,,xr]/(x121,,xr21,ni=1nj=1rxj(vi)j).

The isomorphism is given by evj=1rxj(v)j for all vF2r. We denote the ideal (x121,,xr21,ni=1nj=1rxj(vi)j) by I(G).

Proof.

First, there is an isomorphism of Z-modules Z2rZ[x1,,xr]/(x121,,xr21)under the map evj=1rxj(v)j for all vF2r. To prove the proposition, it suffices to examine the image of imL(G) under this isomorphism, and show that it is the ideal generated by ni=1nj=1rxj(vi)j. Based on the definition of L(G), the Laplacian maps eu to n·eui=1neu+vi. Under our isomorphism, this implies that L(G) maps monomial j=1rxj(u)j to nj=1rxj(u)ji=1nj=1rxj(u)j+(vi)j=j=1rxj(u)j(ni=1nj=1rxj(vi)j).

As a result, the image of imL(G) in Z[x1,,xr]/(xi21) is the ideal generated by ni=1nj=1rxj(vi)j. □

Remark 1.12.

As [Citation3] does not explicitly use the language of Cayley graphs, we explain how to deduce Proposition 1.11 from its results, and use this opportunity to explain the representation theoretic background of it.

In the paper, the authors considered the notion of a critical group attached to a faithful representation γ of a finite group Γ, which is the torsion part of the cokernel of the extended McKay–Cartan matrix associated with γ: the rows and columns of the matrix are indexed by the irreducible (complex) representations of Γ, with dimγ on the diagonal entries, and each off-diagonal (α,β)-entry equals minus the multiplicity of β in γα. This group is isomorphic to (the additive group of) the quotient of the virtual representation ring of Γ [Citation3, Proposition 5.22], where its elements correspond to the complex representations of Γ, with addition and multiplication coming from taking direct sum and tensor product of representations, respectively.

In our case, the variable xi corresponds to the irreducible representation F2rC given by (b1,,br)(1)bi, so all irreducible representations of F2r are in bijection with the squarefree monomials in xi’s. The map evj=1rxj(v)j then gives a bijection between elements of F2r and the irreducible representations of the very group. Under this identification, we can consider the representation i=1nevii=1nj=1rxj(vi)j of F2r, which is faithful as vi’s generate F2r by assumption. It is routine to check that the Laplacian of G(F2r,M) is exactly the extended McKay–Cartan matrix associated with this representation, and Proposition 1.11 follows from [Citation3, Proposition 5.22].

Example 1.13.

In the case G = Qn, this polynomial ring is Z[x1,,x2,,xn]/(x121,x221,,xn21,x1+x2++xnn).

The ring and the ideal I(Qn)=(x121,x221,,xn21,x1+x2++xnn) are symmetric under the action of Sn, which is an important fact that we use later.

The two descriptions of the sandpile group K(G) in Proposition 1.11 each has its own advantages: the matrix-theoretic definition allows us to apply linear algebra techniques, and the representation-theoretic (ring-theoretic) definition has a special multiplicative structure. We will switch between the two descriptions throughout this paper.

Remark 1.14.

It is important to note the sandpile group K(G(F2r,M)) is invariant under left multiplication by elements of GLr(F2). That is, given TGLr(F2), we have K(G(F2r,M))K(G(F2r,T·M)). This is because M and T·M form Cayley graphs that are isomorphic via the bijective map uT·u,uF2r of vertices. As a result, we are free to replace a given M by any representative from the GLr(F2)-orbit of M.

Convention: It is often convenient to label the elements of F2r as w0,,w2r1, where the index of the vector (d1,,dr)T is the binary number d1dr¯. For the generator set M, we denote the multiplicity of wi by μi, so n=i=02r1μi. In Section 4, we also label the elements of F2r as wS for all subsets S[r], where wS has entry 1 at indices sS and 0 elsewhere. For example, when r = 5, w{1,3,4}=(1,0,1,1,0)TF2r.

2 The number of 2-cyclic factors

In this section we wish to compute the number of 2-cyclic factors appearing in the sandpile group for G=G(F2r,M). Given a sandpile group K(G)pe1(Z/peZ)m(pe),we can tensor with Z/2Z to get K(G)(Z/2Z)e1(Z/2eZ)m(2e)Z/2Z(Z/2Z)em(2e),

where we used that (Z/peZ)(Z/2Z)=0 for p2 and (Z/2eZ)(Z/2Z)=Z/2Z for all e1. We define d(M):=e1m(2e),which is the number of even invariant factors.

As stated in the introduction, we establish a dichotomy of Cayley graphs in terms of their values of d(M). First, we define the genericity condition on the generating set M that only depends on the parity of μi’s:

Definition 2.1.

We say that M is generic if i=1nvn0.

Example 2.2.

In the examples for r = 3 in Example 1.6, Q3 and Q3,vert are generic, but K8 is not generic.

We prove a few useful observations.

Lemma 2.3.

d(M) equals 2r1 minus the rank of L(G(F2r,M)) over F2.

Proof.

Tensoring the exact sequence Z2rLZ2rK(G)Z0 with Z/2Z yields (Z/2Z)2rL(Z/2Z)2r(Z/2Z)d(M)+10, hence (Z/2Z)2rIm L(Z/2Z)d(M)+1. Taking the dimension of both sides proves the lemma. □

Proposition 2.4.

d(M) only depends on the parity of the multiplicities μi’s.

Proof.

The rank of L(G(F2r,M)) over F2 only depends on the parity of its entries, which only depends on the parity of μi’s. The statement follows from Lemma 2.3. □

Proposition 2.5.

Let (μ1,,μ2r1) be the multiplicities of the generators associated to M, and assume not all μi have the same parity. Let M be the generating set with multiplicities (μ1+1,,μ2r1+1). Then d(M)=d(M).

In the following proof, we use the term support in two different senses, which should be clear from the context. For a vector wiF2r, its support is the set Si={j:(wi)j=1}; for a polynomial, its support consists of all monomials whose coefficients are nonzero.

Proof.

Apply the change of variables ui:=xi1 to the presentation in Proposition 1.11, and tensor the ring with Z/2Z. We thus obtain the ring R:=(Z/2Z)[u1,,ur]/(u12,,ur2,p(u1,u2,,ur):=ni=12r1μij=1r(uj+1)(wi)j),which is isomorphic to (K(G)Z)(Z/2Z).

For M, we similarly obtain the ring R:=(Z/2Z)[u1,,ur]/(u12,,ur2,p(u1,,ur)),where p(u1,,ur)=n+(2r1)i=12r1(μi+1)j=1r(uj+1)(wi)j.

We want to show that p(u1,,ur)I(G) and p(u1,,ur)I(G), as they imply I(G)=I(G), hence R=R and d(M)=dimZ/2ZR1=dimZ/2ZR1=d(M). We claim that p(u1,ur)p(u1,,ur)=1+i=12r1j=1r(uj+1)(wi)jI(G) and I(G), which would give the desired containment. Set uK=kKuk. Then expanding yields j=1r(uj+1)(wi)j=KSikKuk=KSiuK.

From this we get 1+i=12r1j=1r(uj+1)(wi)j=1+i=12r1KSiuK=1+K{1,,r}#(i:KSi)·uK=(1+(2r1))+K{1,,r},K2r|K|·uK=u1ur.

Thus it remains to show that u1urI(G); since our only assumption on G is that μi’s do not have the same parity, which G satisfies by construction, the argument below also shows u1urI(G).

From the assumption on μi’s, p(u1,,ur) contains some monomial that is not u1ur: if μ2r1 is odd, then we pick, among all wj’s whose multiplicity is even, any wi whose Si is inclusion-maximal. There are 2r|Si|1 elements in F2r whose supports properly contain Si, and their multiplicities are odd. Since j=1r(uj+1)(wi)j contains the monomial uSi if and only if SiSi, uSi is a monomial in the support of p(u1,,ur). The case when μ2r1 is even follows from an analogous argument.

Next we pick, among all wj’s such that uSj is in the support of p(u1,,ur), any wi whose Si is inclusion-minimal. In particular, for any other wj such that uSj is in the support of p(u1,,ur), SjSi is non-empty, and uSj·u1uruSi is not squarefree. Now I(G)p(u1,,ur)·u1uruSi=u1ur+[p(u1,,ur)uSi]·u1uruSi,using the fact that every term in [p(u1,,ur)uSi]·u1uruSi is the multiple of some uj2 hence in I(G), we have u1urI(G) as wanted. □

Example 2.6.

Clearly Proposition 2.5 follows from Theorem 1.1 in the generic case, since adding multiplicity one to each vector does not change the sum. Consider then the case M=(100101010011)N=(110000111010011001111000001111011)

In these cases, a simple Sage computation tells us that K(G(F23,M))=(Z/4Z)4Z/16ZK(G(F23,N))=(Z/4Z)4Z/32Z(Z/3Z)6.

Even though these two cases do not have the same number of invariant factors, they have the same number of Sylow-2 invariant factors.

Remark 2.7.

In linear algebra terms, Proposition 2.5 states that if one changes the parity of every entry of L(G(F2r,M)), the new matrix is of the same rank over F2. However, Proposition 2.5 is not quite a straightforward linear algebra result as it not true for integer matrices in general. Indeed, even within the scope of this paper, the Laplacian of K4G(F22,(e1,e2,e1+e2)) shows that the parity assumption on μi’s is necessary, and the ring theoretic proof is justified.

2.1 Generic cases

We now prove Theorem 1.1, which we restate here:

Theorem 2.8.

Suppose that M is generic. Then the number of Sylow-2 cyclic factors of K(G(F2r,M)) is 2r11.

Proof.

We want to find the dimension of (K(G)Z)(Z/2Z)(Z/2Z)[x1,,xr]/(x121,,xr21,ni=1nj=1rxj(vi)j)as a vector space over Z/2Z.

Again making the change of variables ui:=xi1 yields (K(G)Z)(Z/2Z)R:=(Z/2Z)[u1,,ur]/(u12,,ur2,p(u1,,ur):=ni=1nj=1r(uj+1)(vi)j).

Since i=1nvi0, without loss of generality, i=1n(vi)r=1. As a result, the coefficient of ur in p(u1,,ur) is 1. p(u1,,ur) has trivial constant term, so we can rewrite p(u1,,ur)=ur·fg+(non‐squarefree terms), where f, g are polynomials in u1,,ur1 and f has constant term 1. Since f has nonzero constant term and ui20,i, we in fact have f21, so f is invertible and urfg in R. We can now construct a ring homomorphism T:R(Z/2Z)[u1,,ur1]/(u12,,ur12)by mapping utut for t < r and urfg. T is surjective for every monomial in (Z/2Z)[u1,,ur1] is the image of the same monomial in (Z/2Z)[u1,,ur].

As a vector space over Z/2Z, the dimension of (Z/2Z)[u1,,ur1]/(u12,,ur12) is 2r1 as the standard monomials (with respect to any monomial order) are precisely the squarefree monomials in u1,,ur1. On the other hand, we claim that dimZ/2ZR2r1 by showing that the standard monomials must be squarefree and ur-free with respect to the lexicographic order, with ur being the largest variable: squarefree-ness is routine; for any monomial of the form ur·u where u is a squarefree monomial in u1,,ur1, consider (uf)p=(ur·u)f2ufg+uf(non‐squarefree terms)I(G), every term other than ur·u is either not squarefree or does not contain ur, so ur·u is the leading term of [(uf)p subtracting the non-squarefree terms], which is still in I(G). Thus, the map T must be an isomorphism by comparing dimensions. Therefore, (ZK(G))Z/2Z(Z/2Z)2r1 as vector spaces, and K(G) has (2r11) Sylow-2 cyclic factors. □

2.2 Non-generic cases

In the non-generic cases, the last generator of I(G) no longer has a degree 1 term, so we cannot construct the isomorphism in the proof above. Nevertheless, we in turn prove that Theorem 1.1 actually characterizes generic sets of generators. The following theorem restates Theorem 1.2 with a stronger bound when n is not too big compared to r.

Theorem 2.9.

Suppose that M is not generic. Then the number of Sylow-2 cyclic factors of K(G(F2r,M)) is at least 2r11+2rn/2; when r<n/2, the inequality simply states that the number is at least 2r1.

For the rest of this section, we consider the purely linear algebraic point of view of the problem as in Lemma 2.3, although it would also be interesting to find a Gröbner theoretic proof as in the generic case. In particular, we view every matrix of interest as a matrix over F2 and we need to prove that the rank of the Laplacian is at most 2r12rn/2.

The simple but essential observation is that every Cayley graph of F2r can be considered as the quotient of some hypercube.

Definition 2.10.

Let G=G(F2r,M) be a Cayley graph and UF2r be a subspace, such that UM=. The quotient Cayley graph of G by U, denoted by G/U=G(F2r/U,M), is the graph with vertex set F2r/U and multiedge set {(w+U,vi+w+U)} for coset w + U and viM.

In other words, for any coset X and Y, fix element xX, then every edge between two vertices xX,yY in G descends to an edge between X, Y in G/U.

Lemma 2.11.

[Citation11, Proposition 37] Let M be the matrix whose columns are the generators of the Cayley graph. Then G(F2r,M)Qn/kerM as graphs.

By definition, M is non-generic if and only if the all one vector 1 is contained in kerM, and we can construct any non-generic Cayley graph of F2r from Qn/1 by taking successive codimensional one quotients. This allows us to prove Theorem 2.9 by induction.

Proposition 2.12.

Let r2 and Mr=(e1ere1++er). Then the rank of L(G(F2r,Mr)) is equal to 2r12(r1)/2.

Proof.

We prove by induction on r. For r = 2 and respectively r = 3, the corresponding Cayley graph is K4 and respectively K4,4, their Laplacians’ ranks are 1 and 2, respectively.

To ease the notation, we order the rows and columns of the Laplacian in the order of w0,,w2r1F2r. Denote by L the Laplacian of Qr plus the (2r×2r) identity matrix I, and denote by L¯ and I¯ the matrices obtained from L and I by reversing the ordering of the rows. Note that since L and I are also symmetric with respect to the main skew diagonal, L¯ and I¯ can be obtained from L and I by reversing the ordering of their columns as well. Now the Laplacian of G(F2r+2,Mr+2) has the block structure (LIII¯ILI¯III¯LII¯IIL).

We then perform the following row and column operations in blocks. Here CC+C¯ means we add the (2ri+1)-th column of the block C to the i-th column of C, the meanings of RR+R¯,RR¯,CC¯ are similar. (LIII¯ILI¯III¯LII¯IIL)C3C3+C2,C4C4+C1R3R2+R3,R4R4+R1(LI0L+I¯ILL+I¯00L+I¯00L+I¯000)C1C1+C3¯,C2C2+C3,C4C4+C3¯R2R2+R1¯(LI0L+I¯00L+I¯00L+I¯00L+I¯000)C2C2+C1¯R1R1+R3¯,R4R4+R3¯(L00L+I¯00L+I¯00L+I¯00L+I¯000).

The matrix L+I¯ is the Laplacian of G(F2r,Mr), hence the second and third block of columns each has rank 2r12(r1)/2 by induction hypothesis. For the first and forth block of columns, we have the 2r+1×2r+1 nonzero submatrix (LL+I¯L+I¯0), in which we can perform further row and column operations: (LL+I¯L+I¯0)C2C2+C1R2R2+R1(LI¯I¯L)C1C1¯R1R1¯(LIIL). This is the Laplacian of Qr+1, and is of rank 2r by Theorem 1.1.

Summing the ranks of all the blocks here, we can see that L(G(F2r+2,Mr+2)) is of rank 2(2r12(r1)/2)+2r=2r+12(r+1)/2. □

Remark 2.13.

The number 2r12(r1)/2 also occurs as the number of Z/2Z summands in the elementary divisor decomposition of K(Qr+1) [Citation2, Theorem 1.3]. We are not sure if that is a coincident, or a special case of some deeper connection between the elementary divisors of hypercubes and the number of 2-cyclic factors of other Cayley graphs.

Proof of Theorem 2.9.

As discussed after the statement of Lemma 2.11, we consider a non-generic Cayley graph as the quotient graph Qn/U for some subspace UF2n that contains 1, and prove the theorem by induction on dimU. When dimU=1, we may assume M = Mr by a left multiplication of suitable element of GLr(F2), and the base case of the induction is Proposition 2.12 as r(r+1)/2=(r1)/2.

Suppose dimU>1. Pick a codimension one subspace U of U that contains 1, and pick an arbitrary uU. Since x,yF2n descend to the same vertex in Qn/U if and only if x+U,y+UF2n/U descend to the same vertex in (Qn/U)/u+U, we can see that Qn/U(Qn/U)/u+U. By induction hypothesis, the rank of the Laplacian of Qn/U is equal to 2rm for some m2r+1n/2.

Choose an arbitrary (2r1(m1)/2)×(2r1(m1)/2)-submatrix M of L(Qn/U), whose rows and columns are indexed by R,CF2n/U, respectively. Pick any (2r1(m1)/2)-subset R1F2n/U whose image is R under the projection map F2n/UF2n/U, and let R2=R1+(u+U) be its complement in the preimage of R, order both subsets according to the ordering of their images in R; define and order C1 and C2 for C similarly.

Consider the (2r2(m1)/2)×(2r2(m1)/2)-submatrix M of L(Qn/U), whose rows and columns are indexed R1R2 and C1C2, respectively. The (ordered) subsets R1 and R2 (respectively, C1 and C2) differ coordinate-wise by the translation of u+U, which implies that M has a block structure (ABBA). Moreover, M itself is equal to A + B by the definition of quotient. Since 2r2(m1)/2>2rm, M is singular. Pick an arbitrary nonzero element (v1v2) from the kernel of M, i.e., (1) Av1+Bv2=Av2+Bv1=0.(1)

If v1v2, then we have M(v1+v2)=(Av1+Bv2)+(Av2+Bv1)=0, so M is singular as v1+v20 is in its kernel. If v1=v20, then 1 implies Av1+Bv1=0, which shows that v1 is in the kernel of M and again shows that M is singular.

Since M is arbitrary, the rank of L(Qn/U) is at most 2r1(m1)/21=2r1(m+1)/2. If r+1>n/2, then 2r+1n/22 is an even integer, and (m+1)/2(2r+1n/2+1)/2=2rn/2. Otherwise, we at least have m1, and (m+1)/21>2rn/2 for r<n/2. In both cases, the inequality in the theorem statement is verified. □

We conclude this section with some implications in the other setting that was mentioned in Remark 1.5. A signed graph is an ordinary graph with some edges labeled as negative. The Laplacian of a signed (simple) graph can be obtained from the ordinary Laplacian by flipping the signs of the off-diagonal entries corresponding to the negative edges; the critical group of the signed graph is the cokernel of the signed Laplacian [Citation13]. In [Citation11], Iga et al. studied a class of signed graphs known as Adinkras, which are graphical gadgets that encode special supersymmetry and Clifford algebras (much like Cayley graphs encode finite groups); it is known that the underlying graph of any Adinkra must be a Cayley graph of some F2r. In that work, the authors were able to determine the Sylow-p subgroup of the critical group of an Adinkra for every odd prime p, and the Sylow-2 subgroup when the rank of the signed Laplacian over Z/2Z is exactly 2r1. Since the rank over Z/2Z is independent of the signs of the entries, the results of this section imply that their results on Sylow-2 subgroups apply precisely when the underlying Cayley graph is generic; see [Citation17, Section 4] for further discussion on the role of genericity in the theory of Adinkras.

3 Bounding the largest cyclic factor

We are interested in the largest powers of 2 dividing the cyclic factors ci (i.e., their 2-adic valuations), since all of the p2 information is determined by Proposition 1.9. Anzis and Prasad [Citation1] proved that for G = Qn, the largest cyclic factor c1(Qn) must divide 2nlcm(1,,n). As an immediate corollary, the largest 2-cyclic factor is bounded by 2log2n+n. In this section, we adopt and improve their approach, and prove a tighter bound for all Cayley graphs of F2r.

Theorem 3.1.

Let G=G(F2r,M) be a Cayley graph, and let (λ1=0),λ2,,λ2r be the eigenvalues of the L(G). Then c1(G)|2r2lcm(λi:i2).

For G = Qn, the eigenvalues of L(G) are of the form 2k for 0kn with multiplicity (nk). In this case our bound is 2n2lcm(2,,2n)=2n1lcm(1,,n), which improves the bound in [Citation1] by a factor of 2. This also improves the general bound in [Citation14, Corollary 3], which states that lcm(λi) divides c1(G), and c1(G) in turn divides the product of all distinct nonzero eigenvalues λ{λi:λi0}λ for any integer matrix with integer eigenvalues.

First, using the ring-theoretic description in Proposition 1.11, we have the following lemma.

Lemma 3.2.

Suppose a nonzero p(x1,,xr)¯Z[x1,,xr]/I(G)ZK(G) has finite additive order and let w be the vector in Z2r corresponding to p(x1,,xr) under the isomorphism Z2rZ[x1,,xr]/(x121,,xr21).

Then the additive order of p(x1,,xr)¯ is |w|:=min{CZ>0:CwIm L(G)}.

Proof.

This follows by the definition of cokernel and considering the cokernel as a Z-module. □

From now on, we denote ord(p(x1,,xr)) as the additive order of p(x1,,xr)¯ in Z[x1,,xr]/I(G).

Lemma 3.3.

c1(G)=lcm(ord(xk1):1kr).

Proof.

Consider the map π:Z[x1,,xr]/I(G)ZK(G)Z that sends all xi1. This is a well-defined map because (xi=1) is a common zero for all polynomials in I(G). Since π is surjective, (Z[x1,,xr]/I(G))/ker(π)Z, which forces ker(π) to be the torsion part K(G), and an element has finite additive order in the ring Z[x1,,xr]/I(G) iff it lies in the kernel of π. Furthermore, any polynomial with finite additive order is a linear combination of xI1, where xI denotes a squarefree monomial: we can write any polynomial in the ring as a sum IcIxI of squarefree monomials since xi2=1, and if the polynomial is in the kernel of π, then IcI=0 thus we can rewrite it as IcI(xI1).

Now let C=lcm(ord(xk1):1kr), and let xI be any squarefree monomial. We wish to show that C(xI1)0, i.e., C(xI1)I(G). Indeed, suppose xj|xI. Then we have C(xI1)=C(xj1)·xIxj+C(xIxj1), which we can reduce inductively to a linear combination of xI1 with degxI=1. This shows that the largest cyclic factor is equal to the lcm of the orders taken over all xi1, which is the desired result. □

Remark 3.4.

This lemma can actually be slightly generalized. Namely, let v1,,vn be any generating set of F2r. Then the largest possible additive order equals the lcm of ord(jxj(vi)j1) over all 1in. Anzis and Prasad’s original argument shows that for the hypercube, we can take any xk1 for 1kn. The argument relies on showing that xi1’s have the same additive order, which follows from symmetry of the generators under permutation. However, this is no longer the case for general Cayley graphs.

Denote the lowest common denominator (lcd) of a set of fractions AQ as the smallest positive integer lcd(A) such that lcd(A)·aZ for all aA; lcd obviously exists when A is finite. The next lemma gives an explicit way for calculating ord(xj1), Theorem 3.1 follows from it.

Lemma 3.5.

The order of xj1 in K(G) is equal to the lowest common denominator of the following rational numbers: 12r2uj=1u·v=11λu,for vF2r.

Proof.

Following from Lemma 3.2, the order of xj1 is the smallest positive integer C such that there is a solution vZ2r to L(G)v=Cw for w=xj1(1,0,,1,,0)=ewiew0. We first solve the equation L(G)v=w by expressing w as a linear combination of the eigenbasis of L(G). By Lemma 1.8, we change to the eigenbasis and get w=xj112r(vF2r((1)v·ej(1)v·0)fv)=12r1uj=1fu.

Since fu is an eigenbasis, the equation L(G)v=w has the following particular solution v=v0: v0=12r1uj=11λufu=12r1vF2ruj=1(1)u·vλu·ev.

Since the graph G is connected, all solutions to L(G)v=w are of the form v0+k1T where 1=(1,,1) lies in the kernel. Therefore, C is the smallest integer such that Cv0+k1 is an integer vector for some kR. Equivalently speaking, C is the lowest common denominator of pairwise-differences of v0’s entries given by (v0)v(v0)0=12r1uj=1(1)u·v1λu=12r2uj=1u·v=11λu,for all vF2r.

Proof of Theorem 3.1.

This follows immediately from Lemmas 3.3 and 3.5. □

Corollary 3.6.

The largest 2-cyclic factor Z/2eZ of K(G) for the Cayley Graph G=G(F2r,M) has bound e=v2(c1(G))log2(n)+r1.

Proof.

From Theorem 3.1 we know that c1(G) must divide 2r2lcm(λi:i2). Therefore, v2(c1(G)) must be capped by the 2-adic valuation of the right hand side, which is e=v2(c1(G))maxi2(v2(λi))+r2.

However, note that 2λi2n for i2, so all v2(λi) is upper-bounded by log2(2n)=log2(n)+1. □

What is especially nice about this improvement is that it is asymptotically tight: We can see in Corollary 4.3 that this upper bound on the 2-adic valuation is achieved for all hypercubes of dimension 2k or 2k+1. This is an immediate consequence of the main result of the next section, which completely determines the top cyclic factor of the hypercube. We now discuss some preliminary results that lead to this sharper Qn result.

Remark 3.7.

Since we only care about the Sylow-2 factor in these maximal orders, it actually suffices to find the minimal C such that for any vF2r, C2r2uj=1u·v=11λuZ(2)where Z(2) is the localization of the integers away from the prime ideal (2). This way, we do not actually care about odd denominators.

Although Lemma 3.5 gives us a closed form to calculate the order of xj1, the alternating sums in Lemma 3.5 are too complicated to help us prove useful properties of these numbers or perform calculations. Our solution is to transform these sums into a set of simpler numbers that have the same lowest common denominator. In particular, we use the following claim.

Claim 3.8. Given two (finite) sets of rational numbers A and B, if spanZ(A)=spanZ(B), then A and B have the same lowest common denominator.

The next change-of-basis lemma simplifies the complicated sums in Lemma 3.5, thus helps us to prove a simpler closed form for c1(G) by the end of this section.

Lemma 3.9.

In the algebra Z[aS:S[r]], we have the following equality: spanZ{|ST| oddaS|T[r]}=spanZ{2|T|1TSaS|T[r]}.

Proof.

Denote the sums on the left hand side as xT=|ST| oddaS and the sums on the right hand side as yT=2|T|1TSaS. We will first show that every xT can be expressed as an integer linear combination of the yW’s.

Claim 3.10. xT=WT(1)|W|1yW.

We can verify this by computation. (2) WT(1)|W|1yW=WT((2)|W|1·WSaS)=S[r](aS·WST(2)|W|1).(2)

Now we investigate the rightmost sum in 2. If |ST|=k, then WST(2)|W|1=1pk(2)p1(kp)=12·((2+1)k1)={0,ifkiseven,1,ifkisodd.

Plug it back into 2 and we get exactly the equation in the claim.

Now if we look at the change-of-basis matrix between xT’s and yT’s, it is an upper-triangular matrix with all ±1’s on the main diagonal (if we order variables by inclusion). Therefore the matrix is invertible over Z. □

With the help of Lemma 3.9, we can rewrite the equation for ord(xj1) in Lemma 3.5 into the following corollary.

Corollary 3.11.

The order of xj1 in K(G) is equal to the lowest common denominator of the following rational numbers: 12r|S|sSus=11λu,for jS[r] and |S|2 and 12r2uj=11λu.

Proof.

Recall that we define wSF2r to be the 0-1 vector with entry 1 at indices sS and 0 elsewhere. In Lemma 3.9, we assign aS=1/λwS if jS and aS = 0 if jS. Then both sides in Lemma 3.9 have the same lowest common denominators. The left hand side is the order of xj1 by Lemma 3.5. The right hand side is the lowest common denominator of the following numbers: 12r|T|1uj=1uT=11λu,for Tj and 12r|T|1uT=11λu,for Tj.

Let T=T{j}. Then we can rewrite the numbers above as: 12r|T|uT=11λu,for jT and |T|2 and 22r|T|uT=11λu,for jT.

The left hand side covers the right hand side except when |T|=1, or T={j}. The left hand side corresponds to LHS in the corollary, and the case T={j} corresponds to RHS in the corollary. □

Finally, we can combine the orders for all xj1 and prove the following closed form of the largest cyclic factor c1(G) of K(G).

Corollary 3.12.

For Cayley Graph G=G(F2r,M), c1(G) is equal to the lowest common denominator of the following rational numbers: 12r|S|sSus=11λu,for S[r] and |S|2 and 12r2uj=11λu,for 1jr.

In fact, this is equivalent to saying that if we draw a r-dimensional hypercube, and put 1/λu on each vertex u0, then c1(G) is equal to the lowest common denominator of the arithmetic means of numbers on a face for all faces with codimension 2 that do not pass through the origin.

4 The top cyclic factors for hypercube graphs

In this section, we specialize to the case G = Qn. We will use the techniques developed in the past section to prove the following theorems.

Theorem 4.1.

For n2, let c1(Qn) be the size of the largest cyclic factor in K(Qn). Then, v2(c1(Qn))=max{maxx<n{v2(x)+x},v2(n)+n1}.

Theorem 4.2.

For n3, the 2nd to the (n1)th largest 2-cyclic factor in K(Qn) all have the same size v2(c2(Qn)). Moreover, v2(c2(Qn))=v2(c3(Qn))==v2(cn1(Qn))=maxx<n{v2(x)+x}.

Specifically, when n=2k or 2k+1, Theorems 4.1 and 4.2 implies the following corollary, which shows that the bound c1(G)r+log2n1 in Corollary 3.6 is asymptotically sharp over all Cayley graphs of F2r.

Corollary 4.3.

  1. For n=2k1, the top (n1) Sylow-2 cyclic factors have exponent 2k1.

  2. For n=2k, the top Sylow-2 factor is 2k+k1. The 2nd through (n1)st Sylow-2 factors are 2k1.

  3. For n=2k+1, the top (n1) Sylow-2 cyclic factors are 2k+k.

Remark 4.4.

We note that one only needs to inspect at most log2n many candidates for x when finding maxx<n{v2(x)+x}. Rewriting maxx<n{v2(x)+x}=maxklog2n{maxv2(x)k,x<n{k+x}}, for every k, we may assume x is equal to n–1 setting the last k digits in binary representation zeros, as it is the largest possible x < n to have v2(x)k.

In particular, writing n–1 in binary, one only needs to compare the values of v2(x)+x for x’s obtained from n–1 (in binary representation) by successively changing the last nonzero digit to zero. For example, if n = 54 and n1=53=(110101)2, we only need to inspect x=(110101)2,(110100)2,(110000)2,(100000)2=53,52,48,32.

In order to prove Theorem 4.1, we first translate it into a question about binomial coefficients.

Lemma 4.5.

c1(Qn) is the lowest common denominator of the following rational numbers: Fn(a):=12nai=0na(nai)2(a+i),for 2an and Fn(1):=12n2i=0n1(n1i)2(1+i).

Proof.

This follows from Corollary 3.12 and the fact that when G = Qn, λu=2w(u) (recall w(u) is the number of 1s in the vector uF2r). In particular, there are (nai) ways to choose i 1’s in a size na vector. □

Notice that there are binomial coefficients in the expressions in the lemma above. Therefore, we will make extensive use of a classical theorem of Kummer’s that computes the 2-adic valuation of binomial coefficients in terms of their binary expansions.

Theorem 4.6.

(Kummer’s Theorem) For any non-negative integers ab, v2((ab))=number of carries when adding ab to b in binary.

Example 4.7.

Therefore, v2((500317))=6, since there are 6 carries.

The next lemma shows that the 2-adic valuations of each sum in Lemma 4.5 is equal to the 2-adic valuation of a single term within the sum.

Lemma 4.8.

For any p1, q0, assume u is the unique element in the interval [p,p+q] that maximizes v2(u), then v2(i=0q(qi)p+i)=v2((qup)u).

Proof.

First we claim that for any c[p,p+q], (3) v2((qcp)c)v2((qup)u).(3)

Denote q<2k+1, then q<2k+1. Now we compare v2((qcp)) and v2((qup)), which by Kummer’s theorem is the number of carries in the sums (cp)+(p+qc) and (up)+(p+qu), respectively. Notice that the binary form of cp and up are the same in the last v2(c) bits, therefore both cases have the same number of carries in the last v2(c) bits. In the remaining kv2(c) bits, the case up has at most kv2(c) carries and the case cp has at least 0 carries. This worst scenario exactly results in equality.

For example, when p = 134, q = 101, we have u = 192 and k = 6. Now we analyze the case when c = 168, v2(c)=3. In the two vertical additions, last v2(c)=3 bits (indicated by the red box) are identical, and therefore the first 3 carries (indicated by the blue box) are identical. The number of carries differ in at most 3, which is the same as kv2(c).

Finally, we show that the equality in 3 will be achieved an odd numbers of times. According to the analysis above, equality occurs when cp does not carry in the highest kv2(c) bits. However, we can negate the top bit of cp. We reconsider the example above where c = 168. In such case, we can switch the top bit (indicated by the orange box) of c = 168 to get c=c+2k=232, such that v2((10134))=v2((10198)).

Therefore, there is a pairing of the equality cases c+2kp and cp when cu. This ends the proof. □

Now we have all the tools to calculate the 2-adic valuation of c1(Qn).

Proof of Theorem 4.1.

n = 1 is trivial. Assume n2, and as usual denote u=2k as the largest power of 2 smaller or equal to n. From Lemma 4.5 we know that v2(c1(Qn))=min{v2(Fn(1)),v2(Fn(2)),,v2(F(n))}. First we show that we can rule out Fn(1) and all Fn(a) for a > u.

  • Claim 1: v2(Fn(1))v2(Fn(2)). From Lemma 4.8, v2(Fn(1))=v2((n1u1))nk+1. Similarly v2(Fn(2))=v2((n2u2))nk+1. Therefore, v2(Fn(1))v2(Fn(2))=v2((n1u1))v2((n2u2))=v2(n1u1)0 since u is even.

  • Claim 2: v2(Fn(a))>v2(Fn(u)) for all a > u. From Lemma 4.8, v2(Fn(u))=n+u1k. In the definition of Fn(a), however, all denominators 2(a+i) in the sum have 2-adic value k. Therefore, v2(Fn(a))n+ak>v2(Fn(u)).

As a result, we only need to consider the minimum of v2(Fn(2)),,v2(Fn(u)), which by Lemma 4.8 is v2(c1(Qn))=min2au{v2(Fn(a))}=n+1+kmin2au{a+v2((naua))}.

We then compute the value of the last term in the equation above:

Claim 4.9. min2au{a+v2((naua))}=min2au{a+kv2(na+1),2+kv2(n)}.

To prove the claim, assume the binary expansion of n is n=2p1+2p2++2pd for 0p1<p2<<pd=k. Denote ni=2p1+2p2++2pi for i=1,2,,d1 and nd=u=2pd. We only need to prove the following two subclaims:

  1. For any i=1,2,,d1, we have

    minni<ani+1{a+v2((naua))}=ni+1+kpi+1=minni<ani+1{a+kv2(na+1)}.

    The first equation comes from the fact that when subtracting ua from na, since na and nu are the same in all but the last pi+1+1 bits, and ua have all 1 except in the last pi+1+1 bits, there are always kpi+1 carries in the first kpi+1 bits. By Kummer’s theorem, v2((naua))kpi+1, and equality is achieved when a=ni+1. The second equation comes from the fact that v2(na+1)pi+1 for a(ni,ni+1] and minimum is reached when a=ni+1.

  2. When n is even, we have

    min2a2p1{a+v2((naua))}=2+kp1=min2a2p1{a+kv2(na+1),2+kv2(n)}.

    The first equation comes from the fact that, similar to case 1, there are always kp1 carries in the first kp1 bits, so v2((naua))kp1, and equality is achieved when a = 2. The second equation comes from the fact that v2(na+1)p1 for a[2,2p1] and the minimum is reached at 2+kv2(n).

By combining the two cases, we have the claim, and using the formula from before: v2(c1(Qn))=max2an{v2(x)+x,v2(n)+n1}

We now work toward the proof of Theorem 4.2. Note that finding the 2nd largest cyclic factor of K(Qn) is the same as finding the maximal additive order in the quotient Z-module (Z[x1,,xn]/I(Qn))/x11. Here · represents the Z-submodule generated by “·” instead of the ring ideal generated by it. The elements of finite additive order are still the elements that are in the kernel of the map π:xi1 after the quotient, and it turns out that the second largest Sylow-2 cyclic factor is the order of x21 in the new quotient Z-module, which we prove in the next lemma.

In the rest of this section, equations of polynomials p(x1,,xn)=q(x1,,xn) means they are equal in the quotient ring Z[x1,,xn]/I(Qn), and the order of a polynomial p(x1,,xn) is the additive order of p(x1,,xn)¯Z[x1,,xn]/I(Qn), unless explicitly specified otherwise.

To find the order of x21 in the quotient Z-module, we want to find C,DZ such that C(x11)+D(x21)=0, with D > 0 as small as possible. The symmetry of the hypercube allows us to deduce the following nice result even though it is not a priori obvious which value should C be.

Lemma 4.10.

For 2kn1, the kth largest cyclic factor of K(Qn) is equal to the smallest positive integer D such that D(xkx1)=0. In the notation above, this means we can choose C=D.

Proof.

To prove this result, we induct on k and adopt the following strategy:

  • We first show that the order of xk1 in (Z[x1,,xn]/I(Qn))/x11,x21,,xk11 is equal to the smallest positive integer D such that D(xkx1)=0.

  • Then we show that xk1 has the maximal additive order in the quotient Z-module above. As a result, the D above is exactly the kth largest cyclic factor of K(Qn).

The base case is k = 2. The order of x21 in the quotient Z-module (Z[x1,,xn]/I(Qn))/x11 is the smallest positive integer D such that C(x11)+D(x21)=0. Note, however, by symmetry that if C(x11)+D(x21)=0 then so C(x31)+D(x11),C(x31)+D(x21)=0. Subtracting shows D(x2x1)=0. Conversely, if D(x2x1)=D(x11)+D(x21)=0 then we can take C=D and so the order of x21 in the quotient Z-module must be the order of x2x1.

Then we show x21 is the element with maximal additive order in the quotient Z-module. Namely, we want to show that D(xI1)0mod (x11) for any squarefree monomial xI. When |I|2, suppose xixj|xI. Then we have D(xI1)=D(xixj)xIxi+D(xIxixj1)=D(xIxixj1) which we can reduce inductively to the base case where |I|1.

For general k, we wish to solve for the minimal Ck such that there exist integer constants r1,,rk1 such that r1(x11)++rk1(xk11)+Ck(xk1)=0. Since kn1, xn is not amongst x1,,xk, and so by symmetry we have that r1(xn1)+r2(x21)+r3(x31)++rk1(xk11)+Ck(xk1)=0 and r1(xn1)+r2(x21)+r3(x31)++rk1(xk11)+Ck(x11)=0. Subtracting yields Ck(xkx1)=0, which implies that Ck must just be the order of xkx1, as desired. The proof that Ck is the kth largest factor follows from the same argument in k = 2 case. □

Remark 4.11.

The proof above heavily depends on the symmetry of Qn. We wonder if similar arguments work to some extent as long as G has some symmetry with respect to generator classes of edges, so that we might derive similar results for certain general Cayley graphs?

Note that the order of xkx1 is just the order of xkx11, since x12=1. By symmetry, all these elements have the same additive order, so this lemma implies that the 2nd through (n1)st largest cyclic factors are all the same. It would thus suffice to compute the 2nd largest cyclic factor.

Proof of Theorem 4.2.

Using Lemma 3.5, we want to find the lowest common denominator of the following numbers 12n2u1+uk=1u·v=11λu.

We will be once again using Lemma 3.9, which tells us we need to find the lowest common denominator of 12n|S|u1+uk=1uS=11λu.

By plugging in the values for λu we obtain the following analogue of Lemma 4.5.

Claim 4.12. c2(Qn) is the lowest common denominator of the following rational numbers: 12na1i=0na1(na1i)2(a+i),for 1an1.

The rational numbers in the lemma above are exactly the same as 12Fn1(1),Fn1(2),,Fn1(n1) in the c1(Qn1) case. The fraction in 12Fn1(1) adds one to the last valuation v2(n1)+(n1)1. Following similar arguments, our factor is just equal to v2(c2(Qn))=maxx<n{v2(x)+x}.

Remark 4.13.

One can show that the nth largest cyclic factor of K(Qn) is the additive order of x1x21 in the quotient Z-module Z[x1,,xn]/(x121,,xn21,xin)/x11,x21,,xn1. This corresponds to the smallest positive integer D such that there exist integer constants r1,,rn where D(x1x21)+r1(x11)++rn(xn1)=0, but we are unable to nail down the exact values for r1,,rn. However, we believe the value of D should be given by the following conjecture, which is supported by data for n11 ().

Table 1 Sylow-2 cyclic factors of K(Qn) for n11 (table from [2]).

Conjecture 4.14. For n3, we have v2(cn(Qn))=max{maxx<n1{v2(x)+x},v2(n1)+n3}.

5 Conclusion and remaining conjectures

In this paper we have analyzed the sandpile groups of Cayley graphs by careful looking at the underlying structure of cyclic factors in the corresponding polynomial ring. While our results partially characterize Syl2(K(G)), determining the complete structure still seems out of reach at this moment.

One possible route to extract finer information could be via finding Gröbner bases for these polynomials. However, we were not able to find strong patterns in the Gröbner bases even for K(Qn), and had difficulty working over Z with these objects. Nevertheless, this appears to be a place for further exploration.

We list remaining conjectures we have gathered based on data.

Conjecture 5.1. The number of Sylow-2 cyclic factors d(M) is odd unless all of the eigenvalues have the same power of 2, in which case d(M)=2n2.

Conjecture 5.2. When the greatest common divisor of all generator multiplicities in M is 1, the sandpile group depends only on the set of eigenvalues generated by M, and independent of the specific choice of M.

Conjecture 5.3. Two Cayley graphs have the same sandpile group if and only if their generator multiplicities are the same up to GL-equivalence.

Conjecture 5.4. The Sylow-2 component of the sandpile group for Q2k1 and Q2k differs as follows: Syl2(K(Q2k)) equals a top cyclic factor as determined in Section 4 and then the remaining factors come from taking Syl2(K(Q2k1)) and doubling the multiplicity of each factor. That is, we have Syl2(K(Q2k))Syl2(K(Q2k1))2×Z/(22k+k1Z).

We note that [Citation6, Remark 5.2] allows one to transfer some information from the (well-understood) Smith group of Qn to Syl2(K(Qn)). Since the amount of information depends on the value of v2(n), their method provides some nontrivial information for the L.H.S. of the above conjecture, but unfortunately provides essentially nothing for the R.H.S. Hence, proving the conjecture (assuming not by computing the sandpile groups directly) can help building a bridge there.

Acknowledgments

A portion of this research was carried out as part of the 2018 REU program at the School of Mathematics at University of Minnesota, Twin Cities. The first three authors would also like to thank Eric Stucky for his edits to this paper and Amal Mattoo for his contributions to this research. The fourth author thanks Victor Reiner for introducing him to the first three authors, and Kevin Iga for the discussion on critical groups. All authors thank the anonymous referee for the thorough reading and comments.

Additional information

Funding

The first three authors are grateful for the support of NSF RTG grant DMS-1745638 and making the program possible, as well as to thank Professor Victor Reiner for providing both guidance and independence in their research efforts. The fourth author was supported by the Trond Mohn Foundation project “Algebraic and Topological Cycles in Complex and Tropical Geometries”.

Notes

1 The failure of the strategy comes from the fact that |G| is not invertible in Zp (or any finite extension thereof) when p||G|, analogous to the situation when |G| is not invertible in characteristic p.

References

  • Anzis, B., Prasad, R. (2016). On the critical groups of cubes. http://www-users.math.umn.edu/∼reiner/REU/AnzisPrasad2016.pdf (accessed: 07.23.2018).
  • Bai, H. (2003). On the critical group of the n-cube. Linear Algebra Appl. 369:251–261. DOI: 10.1016/S0024-3795(02)00727-9.
  • Benkart, G., Klivans, C., Reiner, V. (2018). Chip firing on Dynkin diagrams and McKay quivers. Math. Zeitschrift 290:615–648.
  • Brouwer, A. E., Haemers, W. H. (2012). Spectra of Graphs. Universitext. New York: Springer.
  • Chandler, D. B., Sin, P., Xiang, Q. (2015). The Smith and critical groups of Paley graphs. J. Algebraic Combin. 41(4):1013–1022.
  • Chandler, D. B., Sin, P., Xiang, Q. (2017). The Smith group of the hypercube graph. Designs, Codes Cryptogr. 84(1–2):283–294.
  • Corry, S., Perkinson, D. (2018). Divisors and Sandpiles. Providence, RI: American Mathematical Society.
  • Ducey, J. E., Jalil, D. M. (2014). Integer invariants of abelian Cayley graphs. Linear Algebra Appl. 445:316–325. DOI: 10.1016/j.laa.2013.12.004.
  • Gaetz, C. (2016). Critical groups of group representations. Linear Algebra Appl. 508:91–99. DOI: 10.1016/j.laa.2016.07.001.
  • Grinberg, D., Huang, J., Reiner, V. (2020). Critical groups for Hopf algebra modules. Math. Proc. Cambridge Philos. Soc. 168(3):473–503.
  • Iga, K., Klivans, C., Kostiuk, J., Yuen, C. H. (2023). Eigenvalues and critical groups of Adinkras. Adv. Appl. Math. 143:Paper No. 102450, 45. DOI: 10.1016/j.aam.2022.102450.
  • Klivans, C. J. (2019). The Mathematics of Chip-Firing. Discrete Mathematics and its Applications (Boca Raton). Boca Raton, FL: CRC Press.
  • Reiner, V., Tseng, D. (2014). Critical groups of covering, voltage and signed graphs. Discrete Math. 318:10–40. DOI: 10.1016/j.disc.2013.11.008.
  • Rushanan, J. J. (1995). Eigenvalues and the Smith normal form. Linear Algebra Appl. 216:177–184. DOI: 10.1016/0024-3795(93)00131-I.
  • Stanley, R. P. (2013). Algebraic Combinatorics. Undergraduate Texts in Mathematics. New York: Springer, Walks, trees, tableaux, and more.
  • Vishnoi, N. K. (2012). Lx = b Laplacian solvers and their algorithmic applications. Found. Trends Theor. Comput. Sci. 8(1–2):front matter, 1–141 (2013).
  • Yuen, C. H. (2024). The critical groups of Adinkras up to 2-rank of Cayley graphs. Electron. J. Combin. 31(1):Paper No. 1.38, 9pp. DOI: 10.37236/11758.