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

Lower Bounds for Symmetric Circuits for the Determinantthanks: Research funded by EPSRC grant EP/S03238X/1. A preliminary version of this paper appeared as [12]

Anuj Dawar and Gregory Wilsenach
Department of Computer Science and Technology
University of Cambridge.
[email protected], [email protected]
Abstract

Dawar and Wilsenach (ICALP 2020) introduce the model of symmetric arithmetic circuits and show an exponential separation between the sizes of symmetric circuits for computing the determinant and the permanent. The symmetry restriction is that the circuits which take a matrix input are unchanged by a permutation applied simultaneously to the rows and columns of the matrix. Under such restrictions we have polynomial-size circuits for computing the determinant but no subexponential size circuits for the permanent. Here, we consider a more stringent symmetry requirement, namely that the circuits are unchanged by arbitrary even permutations applied separately to rows and columns, and prove an exponential lower bound even for circuits computing the determinant. The result requires substantial new machinery. We develop a general framework for proving lower bounds for symmetric circuits with restricted symmetries, based on a new support theorem and new two-player restricted bijection games. These are applied to the determinant problem with a novel construction of matrices that are bi-adjacency matrices of graphs based on the CFI construction. Our general framework opens the way to exploring a variety of symmetry restrictions and studying trade-offs between symmetry and other resources used by arithmetic circuits.

1 Introduction

The central open question in the field of arithmetic circuit complexity is the separation of the complexity classes VP\mathrm{VP} and VNP\mathrm{VNP}. Sometimes known as Valiant’s conjecture, this is also described as the algebraic analogue of the P\mathrm{P} vs. NP\mathrm{NP} question. The conjecture is equivalent to the statement that the permanent of a matrix cannot be expressed by a family of polynomial-size arithmetic circuits. Lower bounds on the size of circuits computing the permanent have been established by imposing certain restrictions on the circuit model. For instance, it is known that there is no subexponential family of monotone circuits for the permanent [20] and an exponential lower bound for the permanent is also known for depth-3 arithmetic circuits [14]. In both these cases, the lower bound obtained for the permanent also applies to the determinant, which is known to be in VP\mathrm{VP}111Note that the determinant is not itself monotone in the usual sense, but there is a suitably adapted notion of monotonicity called syntactic monotonicity in [21] with respect to which the determinant does have circuits, but not subexponential size ones.. In that sense, the lower bounds tell us more about the weakness of the model resulting from the restriction than the difficulty of computing the permanent.

In this paper we focus on another restriction on arithmetic circuits introduced relatively recently: that of symmetry [11]. This has been shown to give an exponential separation in the size of circuits computing the permanent and the determinant. We first introduce this restriction informally (a more formal definition is given in Section 3). Given a field 𝔽\mathbb{F} and a set of variables XX, let CC be a circuit computing a polynomial pp in 𝔽[X]\mathbb{F}[X]. For a group GG acting on the set XX, we say that CC is GG-symmetric if the action of any element gGg\in G on the inputs of CC can be extended to an automorphism of CC. Of course, this makes sense only when the polynomial pp itself is invariant under the action of GG.

For example, both the permanent and the determinant are polynomials in a matrix of variables X={xij1i,jn}X=\{x_{ij}\mid 1\leq i,j\leq n\}. Let GG be the group Symn\text{{Sym}}_{n} acting on XX by the action whereby πG\pi\in G takes xijx_{ij} to xπ(i)π(j)x_{\pi(i)\pi(j)}. We call this the square symmetric action. It corresponds to arbitrary permutations applied simultaneously to the rows and columns of the matrix. We show in [11] that there are polynomial-size GG-symmetric circuits computing the determinant, but any family of GG-symmetric circuits for computing the permanent has exponential size. Both results are established for any field of characteristic zero.

The choice of the group action GG in these results is natural, but certainly not the only possibility. The lower bound immediately applies to any larger group of symmetries of the polynomial as well. Consider, for instance the action of the group Symn×Symn\text{{Sym}}_{n}\times\text{{Sym}}_{n} on XX whereby (π,σ)(\pi,\sigma) takes xijx_{ij} to xπ(i)σ(j)x_{\pi(i)\sigma(j)}. We call this the matrix symmetric action. It corresponds to independent permutations applied to the rows and columns. The permanent is invariant under this action and the exponential lower bound for square-symmetric circuits for the permanent applies a fortiori to matrix-symmetric circuits as well. As it happens, in this case we can strengthen the lower bound, if not in terms of size at least in terms of the range of application. That is, the exponential lower bound has been shown not only for circuits computing the permanent in fields of characteristic zero, but over all fields of characteristic other than two. This exponential lower bound matches known upper bounds as the smallest circuits for computing permanents, those based on Ryser’s formula, are in fact Symn×Symn\text{{Sym}}_{n}\times\text{{Sym}}_{n}-symmetric.

In the case of the determinant, it was left open whether the polynomial upper bound obtained for square symmetric circuits could be improved by requiring larger groups of symmetries on the circuits. The most efficient algorithms for computing the determinant (based on Gaussian elimination) are not square symmetric and the polynomial upper bound is obtained by an application of Le Verrier’s method [11]. It is a natural question to ask how much more stringent a symmetry requirement we can impose and still find efficient algorithms. The determinant is not matrix-symmetric like the permanent is. Let us write 𝐃n\mathbf{D}_{n} for the group of permutations of X={xij1i,jn}X=\{x_{ij}\mid 1\leq i,j\leq n\} which fix the determinant. This can be seen to be DnTD_{n}\ltimes T where DnD_{n} is the subgroup of Symn×Symn\text{{Sym}}_{n}\times\text{{Sym}}_{n} of index 22 consisting of pairs (σ,π)(\sigma,\pi) of permutations with sgn(σ)=sgn(π)\text{sgn}(\sigma)=\text{sgn}(\pi) and T2T\cong\mathbb{Z}_{2} represents the transposition of rows and columns. We prove in the present paper that any family of 𝐃n\mathbf{D}_{n}-symmetric circuits computing the determinant must have exponential size. Indeed, our lower bound is proved even for the subgroup of DnD_{n} given by Altn×Altn\text{{Alt}}_{n}\times\text{{Alt}}_{n}.

Proving this lower bound requires substantially different methods than those of other results, and developing these methods is a central thrust of this paper. The exponential lower bound for square-symmetric circuits for the permanent is established in [11] by proving a lower bound on the orbit size of Boolean circuits computing the permanent of a {0,1}\{0,1\}-matrix. Here, the orbit size of a circuit CC is the maximal size of an orbit of a gate of CC under the action of the automorphism group of CC. This lower bound is proved using a connection between the orbit size of circuits computing a graph parameter and the counting width of the parameter as established in [1]. To be precise, it is shown that if a graph parameter has linear counting width, i.e. it distinguishes graphs on nn vertices which are not distinguished by Ω(n)\Omega(n)-dimensional Weisfeiler-Leman equivalences, then it cannot be computed by symmetric circuits of subexponential orbit size. The Weisfeiler-Leman equivalences are well-studied approximations of the graph isomorphism relation, graded by dimension (see [15, Section 3.5] for an introduction). The equivalences have many equivalent characterizations arising in combinatorics, algebra, logic and linear optimization. The term counting width comes from the connection with counting logic (see [8]). The main technical ingredient in the lower bound proof in [11] is then a proof of a linear lower bound on the counting width of the number of perfect matchings in a bipartite graph.

We were able to rely on lower bounds on the counting width of graph parameters because a Symn\text{{Sym}}_{n}-invariant parameter of a {0,1}\{0,1\}-matrix can be seen as a graph parameter. That is, a graph parameter that does not distinguish between isomorphic graphs is necessarily Symn\text{{Sym}}_{n}-invariant on the adjacency matrices of graphs. Similarly, the Symn×Symn\text{{Sym}}_{n}\times\text{{Sym}}_{n} action on a matrix can be understood as the natural invariance condition of the biadjacency matrix of a bipartite graph. On the other hand, there seems to be no natural graph structure giving rise to an Altn×Altn\text{{Alt}}_{n}\times\text{{Alt}}_{n}-invariance requirement on the matrices. For this reason, we develop here both a general framework for presenting and studying symmetric circuits and new methods for proving lower bounds under some of these symmetry assumptions.

The generality of this framework allows us to consider a variety of different symmetry conditions, providing both a broad vocabulary for working with these circuits and game-based characterizations of the expressive power of these various symmetric models. This opens up the possibility of studying symmetry as a resource. Our results suggest a spectrum of symmetry restrictions and it would be interesting to establish exactly where on this spectrum the boundary of efficient algorithms for the determinant lies. Similar questions can be asked about other polynomials which admit efficient evaluation. For the permanent, the natural question is how much can we relax the symmetry conditions and still prove lower bounds. This is all to say that, quite apart from the main result, we regard the framework developed here as a contribution in its own right, which lays out a landscape to explore symmetry as a resource in this area.

Table 1 summarizes what is currently known about the power of various symmetric circuit models computing the permanent and determinant. The first column is for unrestricted circuits, i.e. those symmetric under the trivial group action. The upper bound for the determinant is by an adapted Gaussian elimination algorithm [5] and the upper bound for the permanent, which in fact holds for every group in the table, is by Ryser’s formula [24]. The lower bound in the first column is the trivial one. The second and fourth column state results established in [11]. There is no result for the determinant in the fourth column as the determinant is not invariant under the Symn×Symn\text{{Sym}}_{n}\times\text{{Sym}}_{n} action. The results in the column for Altn×Altn\text{{Alt}}_{n}\times\text{{Alt}}_{n} are new to this paper. In the last two columns 𝐃n\mathbf{D}_{n} and 𝐏n\mathbf{P}_{n} represent the full invariance groups (as subgroups of Symn×n\text{{Sym}}_{n\times n}) of the determinant and permanent respectively. The lower bounds stated in those columns follow from the ones obtained for their subgroups Altn×Altn\text{{Alt}}_{n}\times\text{{Alt}}_{n} and Symn×Symn\text{{Sym}}_{n}\times\text{{Sym}}_{n} respectively.

A more detailed discussion of some of the main innovations follows.

GG {id}\{\text{id}\} Symn\text{{Sym}}_{n} Altn×Altn\text{{Alt}}_{n}\times\text{{Alt}}_{n} Symn×Symn\text{{Sym}}_{n}\times\text{{Sym}}_{n} 𝐃n\mathbf{D}_{n} 𝐏n\mathbf{P}_{n}
Det O(n3)O(n^{3})
O(n4)O(n^{4})
(char 0)
2Ω(n)2^{\Omega(n)}
(char 0)
N/A
2Ω(n)2^{\Omega(n)}
(char 0)
N/A
Perm
O(n22n)O(n^{2}2^{n})
Ω(n2)\Omega(n^{2})
2Ω(n)2^{\Omega(n)}
(char 0)
2Ω(n)2^{\Omega(n)}
(char 2\neq 2)
2Ω(n)2^{\Omega(n)}
(char 2\neq 2)
2Ω(n)2^{\Omega(n)}
(char 2\neq 2)
2Ω(n)2^{\Omega(n)}
(char 2\neq 2)
Table 1: Table of upper and lower bounds for GG-symmetric circuits for the determinant and permanent, for various group actions GG. Here {id}\{\text{id}\} denotes the trivial group, and thus the column gives upper and lower bound for general circuits.
  1. 1.

    The lower bounds on counting width of graph parameters are proved using the method of bijection games applied to graphs based on a construction due to Cai, Fürer and Immerman [6] (we refer to this class of graph constructions as the CFI construction). The games are parameterized by a natural number kk and are used to establish the indistinguishability of graphs by means of kk-dimensional Weisfeiler-Leman equivalences. We adapt the bijection games and show that they can be directly used to obtain lower bounds for symmetric circuits without reference to graphs or Weisfeiler-Leman equivalences. The parameter kk is related to a parameter of the circuits we call support size. This is based on an idea outlined in [7] which we elaborate further.

  2. 2.

    The support theorem established in [1] and stengthened in [11] is a key tool, which we extend further. It shows that small symmetric circuits have small support size. This was proved for symmetry under the action of the symmetric group originally and we extend the range of group actions for which it can be shown to hold.

  3. 3.

    The original kk-pebble bijection games of Hella [17] are two-player games played on a pair of graphs (or more generally relational structures) between two players called Spoiler and Duplicator. In each move of the game, Duplicator is required to provide a bijection between the two graphs. A winning strategy for Duplicator shows that the two graphs are not distinguishable in the kk-dimensional Weisfeiler-Leman equivalences. We refine the games by restricting Duplicator to play bijections from a restricted set of bijections. This set is obtained by composing an initial bijection with permutations from some group GG. In this game, even an isormophism does not guarantee a Duplicator winning strategy. But, we are able to relate Duplicator winning strategies to indistinguishability by GG-symmetric circuits taking the adjacency matrices as input. At the same time, we generalize the game so it can be played on a much more general notion of structured input rather than a relational structure.

  4. 4.

    A fourth key ingredient is the construction of matrices with distinct determinants on which we can show Duplicator winning strategies in the Altn×Altn\text{{Alt}}_{n}\times\text{{Alt}}_{n}-restricted bijection game. The two matrices are obtained as biadjacency matrices of a single bipartite graph given by the CFI construction. The two matrices differ only through the interchange of two columns. The challenge here is to show that they have non-zero determinant.

  5. 5.

    An important element of the construction of the matrices in (4) are bipartite 33-regular graphs with large tree-width and an odd number of perfect matchings. We show the existence of an infinite family of such graphs in Section 5.2. This may be of independent interest. A strengthening of the conditions on the number of perfect matchings could be used to extend our lower bound to fields of positive characteristic other than two. We expand on this in Section 5.4.

Our work may be compared to that of Landsberg and Ressayre [22] who establish an exponential lower bound on the complexity of the permanent (specifically over the complex field \mathbb{C}) under an assumption of symmetry. Their lower bound is for equivariant determinantal representations of the permanent, that is those that preserve all the symmetries of the permanent function. This approach doesn’t yield any lower bounds for symmetric circuits in the sense we consider here. On the other hand, lower bounds for equivariant determinantal complexity can be derived from the results in [11], albeit not ones as strong as in [22]. For a more detailed comparison of the two approaches see the full version of [11].

This paper is organized as follows. We begin with introducing the notation we need in Section 2. Symmetric circuits working on arbitrary structured inputs are defined in Section 3. The support theorem we need is proved in Section 4.2. The bijection games are defined in Section 4.1 where we also prove that Duplicator winning strategies imply indistinguishability by small circuits. The main construction establishing the lower bound for the determinant is presented in Section 5. The corresponding lower bound for the permanent is given in Section 6. The proof for the last result is only sketched as we only need to note that the proof from [11] can be easily adapted to the new games we define to tighten the symmetry result from Symn×Symn\text{{Sym}}_{n}\times\text{{Sym}}_{n} to Altn×Altn\text{{Alt}}_{n}\times\text{{Alt}}_{n}.

2 Background

In this section we discuss relevant background and introduce notation.

We write \mathbb{N} for the positive integers and 0\mathbb{N}_{0} for the non-negative integers. For m0m\in\mathbb{N}_{0}, [m][m] denotes the set {1,,m}\{1,\ldots,m\}. For a set XX we write 𝒫(X)\mathcal{P}(X) to denote the powerset of XX. We write id\operatorname{id} to denote the identity function on some specified set. For f:XYf:X\rightarrow Y and SXS\subseteq X we write f(S)f(S) to denote the image of SS. Let Bij(A,B)\mathrm{Bij}(A,B) denote the set of bijections from AA to BB.

2.1 Groups

Let SymA\text{{Sym}}_{A} and AltA\text{{Alt}}_{A} denote the symmetric group and alternating group on the set AA. We write Symn\text{{Sym}}_{n} and Altn\text{{Alt}}_{n} to abbreviate Sym[n]\text{{Sym}}_{[n]} and Alt[n]\text{{Alt}}_{[n]}, respectively. Let {id}\{\operatorname{id}\} denote the trivial group. For groups GG and HH we write HGH\leq G to denote that HH is a subgroup of GG. The sign of a permutation σSymA\sigma\in\text{{Sym}}_{A} is defined so that if σ\sigma is even sgn(σ)=1\text{sgn}(\sigma)=1 and otherwise sgn(σ)=1\text{sgn}(\sigma)=-1.

Let GG be a group acting on a set XX. We denote this as a left action, i.e. σx\sigma x for σG\sigma\in G, xXx\in X. The action extends in a natural way to powers of XX. So, for (x,y)X×X(x,y)\in X\times X, σ(x,y)=(σx,σy)\sigma(x,y)=(\sigma x,\sigma y). It also extends to the powerset of XX and functions on XX as follows. The action of GG on 𝒫(X)\mathcal{P}(X) is defined for σG\sigma\in G and S𝒫(X)S\in\mathcal{P}(X) by σS={σxxS}\sigma S=\{\sigma x\mid x\in S\}. For YY any set, the action of GG on YXY^{X} is defined for σG\sigma\in G and fYXf\in Y^{X} by (σf)(x)=f(σx)(\sigma f)(x)=f(\sigma x) for all xXx\in X. We refer to all of these as the natural action of GG on the relevant set.

A set XX with the action of group GG on it is called a GG-set. We do not distinguish notationally between a GG-set XX and the underlying set of elements. Thus, we can say that if XX is a GG-set, the collection of functions YXY^{X} is also a GG-set with the natural action.

Let X=iIXiX=\prod_{i\in I}X_{i} and for each iIi\in I let GiG_{i} be a group acting on XiX_{i}. The action of the direct product G:=iIGiG:=\prod_{i\in I}G_{i} on XX is defined for x=(xi)iIXx=(x_{i})_{i\in I}\in X and σ=(σi)iIG\sigma=(\sigma_{i})_{i\in I}\in G by σx=(σixi)iI\sigma x=(\sigma_{i}x_{i})_{i\in I}. If instead X=iIXiX=\biguplus_{i\in I}X_{i} then the action of GG on XX is defined for xXx\in X and σ=(σi)iIG\sigma=(\sigma_{i})_{i\in I}\in G such that if xXix\in X_{i} then σx=σix\sigma x=\sigma_{i}x. Again, we refer to either of these as the natural action of GG on XX.

Let XX be a GG-set. Let SXS\subseteq X. Let StabG(S):={σGxSσx=x}\text{{Stab}}_{G}(S):=\{\sigma\in G\mid\forall x\in S\,\,\sigma x=x\} denote the pointwise stabilizer of SS. Let SetStabG(S):={σGσ(S)=S}\text{{SetStab}}_{G}(S):=\{\sigma\in G\mid\sigma(S)=S\} denote the setwise stabilizer of SS. If S={x}S=\{x\} is a singleton we omit set braces and write StabG(x)\text{{Stab}}_{G}(x). Note that SetStabG(S)\text{{SetStab}}_{G}(S) is the pointwise stabilizer of {S}\{S\} in the GG-set 𝒫(X)\mathcal{P}(X). For xXx\in X let OrbG(x):={σ(x)σG}\text{{Orb}}_{G}(x):=\{\sigma(x)\mid\sigma\in G\} denote the orbit of xx. In all cases, we omit the subscript GG where it is clear from context.

2.2 Matrices

Let AA and BB be finite non-empty sets. We identify matrices with rows indexed by AA and columns by BB and entries from some set XX with functions of the form M:A×BXM:A\times B\rightarrow X. So for aAa\in A, bBb\in B, Mab=M(a,b)M_{ab}=M(a,b). We also denote MM by (Mab)aA,bB(M_{ab})_{a\in A,b\in B}, or just (Mab)(M_{ab}) when the index sets are clear from context.

Let RR be a commutative ring and M:A×BRM:A\times B\rightarrow R be a matrix with |A|=|B||A|=|B|. The permanent of MM over RR is permR(M)=σBij(A,B)aAMaσ(a)\text{perm}_{R}(M)=\sum_{\sigma\in\mathrm{Bij}(A,B)}\prod_{a\in A}M_{a\sigma(a)}. Suppose A=BA=B. The determinant of MM over RR is detR(M)=σSymAsgn(σ)aAMaσ(a)\det_{R}(M)=\sum_{\sigma\in\text{{Sym}}_{A}}\text{sgn}(\sigma)\prod_{a\in A}M_{a\sigma(a)}. The trace of MM over RR is TrR(M)=aAMaa\operatorname{Tr}_{R}(M)=\sum_{a\in A}M_{aa}. We omit reference to the ring when it is obvious from context. When RR is a field, we write rk(M)\text{{rk}}(M) to denote the rank of the matrix MM.

We always use 𝔽\mathbb{F} to denote a field and char(𝔽)\text{char}(\mathbb{F}) to denote the characteristic of 𝔽\mathbb{F}. For any prime power qq we write 𝔽q\mathbb{F}_{q} for the finite field of order qq. We are often interested in polynomials and circuits defined over a set of variables XX with a natural matrix structure, i.e. X={xab:aA,bB}X=\{x_{ab}:a\in A,b\in B\}. We identify XX with this matrix. We also identify any function of the form f:XYf:X\rightarrow Y with the A×BA\times B matrix with entries in YY defined by replacing each xabx_{ab} with f(xab)f(x_{ab}).

2.3 Graphs

Given a graph Γ=(V,E)\Gamma=(V,E), the adjacency matrix AΓA_{\Gamma} of Γ\Gamma is the V×VV\times V {0,1}\{0,1\}-matrix with AΓ(u,v)=1A_{\Gamma}(u,v)=1 if, and only if, {u,v}E\{u,v\}\in E. If Γ\Gamma is bipartite, with bipartition V=UWV=U\cup W, then the biadjacency matrix BΓB_{\Gamma} of Γ\Gamma is the U×WU\times W {0,1}\{0,1\}-matrix with BΓ(u,v)=1B_{\Gamma}(u,v)=1 if, and only if, {u,v}E\{u,v\}\in E. For a set UVU\subseteq V we write Γ[U]\Gamma[U] to denote the subgraph induced by UU.

A kk-factor of a graph Γ\Gamma is a spanning kk-regular subgraph. A perfect matching is a 11-factor. It is well known that for a bipartite graph Γ\Gamma, perm(BΓ)\text{perm}(B_{\Gamma}) over any field of characteristic zero counts the number of perfect matchings in Γ\Gamma [16] and for prime pp, perm𝔽(BΓ)\text{perm}_{\mathbb{F}}(B_{\Gamma}) for a field 𝔽\mathbb{F} of characteristic pp counts the number of perfect matchings in Γ\Gamma modulo pp. For bipartite Γ\Gamma, AΓA_{\Gamma} is a block anti-diagonal matrix with two blocks corresponding to BΓB_{\Gamma} and BΓTB_{\Gamma}^{T}, and so perm(AΓ)=perm(BΓ)2\text{perm}(A_{\Gamma})=\text{perm}(B_{\Gamma})^{2}.

Let Γ=(V,E)\Gamma=(V,E) be a graph. For SVS\subseteq V let N+(S):={vVSsS,(v,s)E}N^{+}(S):=\{v\in V\setminus S\mid\exists s\in S,\,\,(v,s)\in E\}. We say Γ\Gamma is an α\alpha-expander for α[0,1]\alpha\in[0,1] if for every SVS\subseteq V of size at most |V|/2|V|/2 we have |N+(S)|α|S||N^{+}(S)|\geq\alpha|S|. For an introduction to expander graphs see [19]. A set SS of vertices in a graph Γ\Gamma is a balanced separator if no connected component of ΓS\Gamma\setminus S contains more than half the vertices of Γ\Gamma. It is easy to see that for any α\alpha there is a constant τ\tau such that if Γ\Gamma is an α\alpha-expander then Γ\Gamma has no balanced separator of size less than τ|V|\tau|V|.

The tree width of a graph is a well-known graph parameter which can be characterized in terms of the cops and robbers game [25]. This game is played on a graph Γ=(V,E)\Gamma=(V,E), with one player controlling the single robber and another a team of kk cops. The cops and robbers both sit on nodes in the graph. At the start of a round the cops are sitting on nodes XVX\subset V and the round begins with the cop player announcing he intends to move cops from XXX^{\prime}\subseteq X to occupy the nodes YY. The robber player must respond with a path from the robbers current location that does not intersect XXX\setminus X^{\prime}. At the end of the round the cops and robber are moved to their new locations, and if any cop is on the same node as the robber the cops win. The characterization is as follows.

Theorem 2.1 ([25]).

There is a winning strategy for the cop player with kk cops if, and only if, the tree-width of Γ\Gamma is at most k1k-1.

2.4 Circuits

We give a general definition of a circuit that incorporates both Boolean and arithmetic circuits.

Definition 2.2 (Circuit).

A circuit over the basis 𝔹\mathbb{B} with variables XX and values KK is a directed acyclic graph with a labelling where each vertex of in-degree 0 is labelled by an element of XKX\cup K and each vertex of in-degree greater than 0 is labelled by an element of 𝔹\mathbb{B}.

Let CC be circuit that is the labelling of a directed graph (D,W)(D,W), where WD×DW\subset D\times D with values KK. We call the elements of DD gates, and the elements of WW wires. We call the gates with in-degree 0 input gates and gates with out-degree 0 output gates. We call those input gates labelled by elements of KK constant gates. We call those gates that are not input gates internal gates. For g,hDg,h\in D we say that hh is a child of gg if (h,g)W(h,g)\in W. We write child(g)\text{child}(g) to denote the set of children of gg. We write CgC_{g} to denote the sub-circuit of CC rooted at gg, i.e the sub-circuit induced by those gates with a directed path to gg. Unless otherwise stated we always assume a circuit has exactly one output gate.

If KK is a field 𝔽\mathbb{F}, and 𝔹\mathbb{B} is the set {+,×}\{+,\times\}, we have an arithmetic circuit over 𝔽\mathbb{F}. If K={0,1}K=\{0,1\}, and 𝔹\mathbb{B} is a collection of Boolean functions, we have a Boolean circuit over the basis 𝔹\mathbb{B}. We define two bases here. The first is the standard basis 𝔹std\mathbb{B}_{\text{std}} containing the functions \land, \lor, and ¬\neg. The second is the threshold basis 𝔹t\mathbb{B}_{t} which is the union of 𝔹std\mathbb{B}_{\text{std}} and {tk:k}\{t_{\geq k}:k\in\mathbb{N}\}, where for each kk\in\mathbb{N}, tkt_{\geq k} is defined for a string x{0,1}\vec{x}\in\{0,1\}^{*} so that tk(x)=1t_{\geq k}(\vec{x})=1 if, and only if, the number of 11s in x\vec{x} at least kk. We call a circuit defined over this basis a threshold circuit. Another useful Boolean function is t=kt_{=k}, which is defined by t=k(x)=tk(x)¬tk+1(x)t_{=k}(x)=t_{\leq k}(x)\land\neg t_{\leq k+1}(x). We do not explicitly include it in the basis as it is easily defined in 𝔹t\mathbb{B}_{t}.

In general, we require that a basis contain only functions that are invariant under all permutations of their inputs. That is, if f𝔹f\in\mathbb{B} is such that f:KnKf:K^{n}\rightarrow K for some nn\in\mathbb{N} then for all σSymn\sigma\in\text{{Sym}}_{n} and MKnM\in K^{n} we have f(Mσ)=f(M)f(M\circ\sigma)=f(M). We call such a function ff fully symmetric. The arithmetic functions ++ and ×\times and all of the Boolean functions in 𝔹t\mathbb{B}_{t} and 𝔹std\mathbb{B}_{\text{std}} are fully symmetric. Let CC be a circuit defined over such a basis with variables XX and values KK. We evaluate CC for an assignment MKXM\in K^{X} by evaluating each gate labelled by some xXx\in X to M(x)M(x) and each gate labelled by some kKk\in K to kk, and then recursively evaluating each gate gg according to its corresponding basis element. That is gg gets the value obtained by applying the function labelling gg to the set of values of child(g)\text{child}(g). We write C[M](g)C[M](g) to denote the value of the gate gg and C[M]C[M] to denote the value of the output gate. We say that CC computes the function MC[M]M\mapsto C[M].

It is conventional to consider an arithmetic circuit CC over 𝔽\mathbb{F} with variables XX to be computing a polynomial in 𝔽[X]\mathbb{F}[X], rather than a function 𝔽X𝔽\mathbb{F}^{X}\rightarrow\mathbb{F}. This polynomial is defined via a similar recursive evaluation, except that now each gate labelled by a variable evaluates to the corresponding formal variable, and we treat addition and multiplication as ring operations in 𝔽[X]\mathbb{F}[X]. Each gate then evaluates to some polynomial in 𝔽[X]\mathbb{F}[X]. The polynomial computed by CC is the value of the output gate.

For more details on arithmetic circuits see [26] and for Boolean circuits see [27].

3 Symmetric Circuits

Complexity theory is often concerned with computation models that take as input binary strings. In practice these strings are almost always taken to encode some structured input (e.g. graphs, matrices, numbers, etc.). In order to study the symmetries that arise from these structures we forgo this encoding. More precisely, we consider computation models such as circuits whose inputs are themselves functions of type KXK^{X} where we think of XX as a set of variables and KK a domain of values that the variables can take. The set XX may have some further structure to reflect the intended structured input, but no more. In particular, we do not assume that XX is linearly ordered. For example, if X=V2X=V^{2} and K={0,1}K=\{0,1\} then the elements of {0,1}X\{0,1\}^{X} may be naturally interpreted as directed graphs over the vertex set VV. Or, with the same XX, if we let K=𝔽K=\mathbb{F}, we can think of the elements of KXK^{X} as matrices over 𝔽\mathbb{F} with rows and columns indexed by VV.

The symmetries of interest for a given class of structures correspond to group actions on XX, which lift to actions on KXK^{X}. In this section we introduce the definitions of GG-symmetric functions, i.e. functions which are invariant under the action of GG on its input, and GG-symmetric circuits, which are circuits computing GG-symmetric functions but where the structure of the circuit itself and not just the output is invariant under the action of GG. The definitions are variations and generalizations of those from [11]. We illustrate them with examples.

3.1 Group Actions and Symmetric Functions

Definition 3.1.

For a group GG acting on the domain of a function FF we say FF is GG-symmetric if for every σG\sigma\in G, σF=F\sigma F=F. We omit mention of the group when it is obvious from context.

We are interested in functions of type KXKK^{X}\rightarrow K with some group GG acting on XX, which then induces an action on KXK^{X}. We think of elements of KXK^{X} as “generalized structures”, and define notions of homomorphism and isomorphism below. We first consider some examples. Note that whenever HH is a subgroup of GG, then any GG-symmetric function is also HH-symmetric. In particular every function is {id}\{\operatorname{id}\}-symmetric.

Example 3.2.
  1. 1.

    For a vertex set VV, a function F:{0,1}V2{0,1}F:\{0,1\}^{V^{2}}\rightarrow\{0,1\} defines a property of (directed) graphs if it is SymV\text{{Sym}}_{V}-symmetric, with the action of SymV\text{{Sym}}_{V} being defined simultaneously on both coordinates. We get properties of simple undirected graphs by considering SymV\text{{Sym}}_{V}-symmetric functions F:{0,1}X{0,1}F:\{0,1\}^{X}\rightarrow\{0,1\} where X=(V2)X={V\choose 2} is the collection of 22-element subsets of VV. Examples of graph properties include connectedness, Hamiltonicity and the existence of a perfect matching. A graph parameter is a function F:{0,1}V2F:\{0,1\}^{V^{2}}\rightarrow\mathbb{R} which is SymV\text{{Sym}}_{V}-symmetric. Examples include the number of connected components, the number of Hamiltonian cycles and the number of perfect matchings.

  2. 2.

    The elementary symmetric polynomial of degree kk in the set of variables XX is the polynomial:

    ek(X)=S(Xk)xSx.e_{k}(X)=\sum_{S\in{X\choose k}}\prod_{x\in S}x.

    For any field 𝔽\mathbb{F}, ek(X)e_{k}(X) defines a function ek𝔽:𝔽X𝔽e_{k}^{\mathbb{F}}:\mathbb{F}^{X}\rightarrow\mathbb{F} which is SymX\text{{Sym}}_{X}-symmetric.

  3. 3.

    If X={xij1i,jn}X=\{x_{ij}\mid 1\leq i,j\leq n\} is a matrix of variables, then the trace tr(X)\text{tr}(X), determinant det(X)\det(X) and permanent perm(X)\text{perm}(X) are polynomials that define, for any field 𝔽\mathbb{F}, functions of type 𝔽X\mathbb{F}^{X} which are Symn\text{{Sym}}_{n}-symmetric where the group action is defined simultaneously on both coordinates.

  4. 4.

    The permanent is invariant under separate row/column permutations. In other words, for (π,σ)Symn×Symn(\pi,\sigma)\in\text{{Sym}}_{n}\times\text{{Sym}}_{n}, and matrix M𝔽n×nM\in\mathbb{F}^{n\times n}, we have perm((Mxy))=perm((Mπ(x)σ(y)))\text{perm}((M_{xy}))=\text{perm}((M_{\pi(x)\sigma(y)})). So the permanent for n×nn\times n-matrices is a Symn×Symn\text{{Sym}}_{n}\times\text{{Sym}}_{n}-symmetric function.

  5. 5.

    The trace and determinant are not Symn×Symn\text{{Sym}}_{n}\times\text{{Sym}}_{n}-symmetric under the above action, but the symmetry group of the determinant is richer than just Symn\text{{Sym}}_{n}. Define DnSymn×SymnD_{n}\leq\text{{Sym}}_{n}\times\text{{Sym}}_{n} to be the group {(σ,π)sgn(σ)=sgn(π)}\{(\sigma,\pi)\mid\text{sgn}(\sigma)=\text{sgn}(\pi)\}. This is a subgroup of Symn×Symn\text{{Sym}}_{n}\times\text{{Sym}}_{n} of index 22. The determinant is DnD_{n}-symmetric and so, in particular Altn×Altn\text{{Alt}}_{n}\times\text{{Alt}}_{n}-symmetric, since Altn×AltnDn\text{{Alt}}_{n}\times\text{{Alt}}_{n}\leq D_{n}.

Let GG and HH be groups. A homomorphism from the GG-set XX to HH-set YY is a pair of functions (f,ϕ)(f,\phi) where f:XYf:X\rightarrow Y is a function and ϕ:GH\phi:G\rightarrow H is a group homomorphism such that for each xXx\in X and πG\pi\in G, f(πx)=ϕ(π)f(x)f(\pi x)=\phi(\pi)f(x). We say (f,ϕ)(f,\phi) is an isomorphism if both ff and ϕ\phi are bijective. We abuse terminology and refer to a bijection f:XYf:X\rightarrow Y as an isomorphism if there exists ϕ:GH\phi:G\rightarrow H such that (f,ϕ)(f,\phi) is an isomorphism.

Let M:XKM:X\rightarrow K with XX a GG-set and let N:YKN:Y\rightarrow K with YY an HH-set. A homomorphism from (M,G)(M,G) to (N,H)(N,H) is a homomorphism (f,ϕ):(X,G)(Y,H)(f,\phi):(X,G)\rightarrow(Y,H) such that Nf=MN\circ f=M. It is an isomorphism from (M,G)(M,G) to (N,H)(N,H) if both ff and ϕ\phi are bijective. We omit mention of the groups and refer just to a homomorphisms or isomorphisms from MM to NN when the groups are clear from context (see Section 4.1.

Let GG be a group acting on XX and F:KXLF:K^{X}\rightarrow L. We are usually interested in such functions up to isomorphism. Let M:YKM:Y\rightarrow K and HH be a group acting on YY such that (X,G)(X,G) and (Y,H)(Y,H) are isomorphic. Then we abuse notation and sometimes write F(M)F(M) to denote F(Mf)F(M\circ f) for some isomorphism f:XYf:X\rightarrow Y.

3.2 Symmetric Circuits

We next define the notion of a symmetric circuit as it appears in [11]. These circuits take as input functions of the form M:XKM:X\rightarrow K, where XX is a GG-set, and are symmetric in the sense that the computation itself, not just the value of the output, remains unchanged under the action of GG. We first need to formalize what it means for a permutation in GG to act on the gates of a circuit, and for this we define the notion of a circuit automorphism extending a permutation.

Definition 3.3 (Circuit Automorphism).

Let C=(D,W)C=(D,W) be a circuit over the basis 𝔹\mathbb{B} with variables XX and values KK. For σSymX\sigma\in\text{{Sym}}_{X}, we say that a bijection π:DD\pi:D\rightarrow D is an automorphism extending σ\sigma if for every gate gg in DD we have that

  • if gg is a constant gate then π(g)=g\pi(g)=g,

  • if gg is a non-constant input gate then π(g)=σ(g)\pi(g)=\sigma(g),

  • if (h,g)W(h,g)\in W is a wire, then so is (πh,πg)(\pi h,\pi g)

  • if gg is labelled by b𝔹b\in\mathbb{B}, then so is πg\pi g.

We say that a circuit CC with variables XX is rigid if for every permutation σSym(X)\sigma\in\text{{Sym}}(X) there is at most one automorphism of CC extending σ\sigma. The argument used to prove [10, Lemma 5.5] suffices to show that any symmetric circuit may be transformed into an equivalent rigid one in time polynomial in the size of the circuit. As such, when proving lower bounds we often assume the circuit is rigid without a loss of generality.

We are now ready to define the key notion of a symmetric circuit.

Definition 3.4 (Symmetric Circuit).

Let GG be a group acting on a set XX and CC be a circuit with variables XX. We say CC is a GG-symmetric circuit if for every σG\sigma\in G the action of σ\sigma on XX extends to an automorphism of CC.

The following can be shown via a straightforward induction.

Proposition 3.5.

Let CC be a GG-symmetric circuit. Then CC defines a GG-symmetric function.

4 Games and Supports

Hella’s bijection game [17] is a two-player game played on relational structures, such as graphs. It was defined to demonstrate indistinguishability of structures in logics with counting and extensions thereof. The indistinguishability relations it defines on graphs are closely tied to Weisfeiler-Leman equivalences [6]. The games are played on a pair of structures AA and BB by two players called Spoiler and Duplicator. Spoiler aims to show that the two structures are different while Duplicator pretends they are the same. We associate with each structure a sequence of kk pebbles which are placed in the course of the game on elements of the two structures. The game proceeds in a sequence of rounds. The number of rounds can be greater than kk and so pebbles can be moved from one element to another during the course of the game. At each round, Spoiler chooses a pebble pip_{i} (i[k]i\in[k]) from AA and the matching pebble qiq_{i} from BB. Duplicator provides a bijection hh between the two structures. Spoiler then chooses an element aa of AA on which to place pip_{i} and qiq_{i} is placed on h(a)h(a). In each round, Duplicator must ensure that the partial map between the two structures defined by the placement of the pebbles is a partial isomorphism, otherwise Spoiler wins. For more details on this game see [15].

The connection between bijection games and symmetric circuits is first made in [1] which showed a connection between the expressive power of counting logics and symmetric Boolean circuits in the threshold basis 𝔹t\mathbb{B}_{t}. This leads to the suggestion (made in [7]) of using bijection games as a tool directly to prove circuit lower bounds. The main contribution of this section is the generalization of these bijection games in two different directions, as well as results establishing how this tool can be used to prove more general circuit lower bounds.

We noted earlier that we may think of functions on GG-sets as some sort of generalised structure. The development of the theory of bijection games (and supports) requires us to further restrict our attention to the case where GG is a subgroup of some symmetric group. The domain of the symmetric group corresponds in some sense to the universe of the structure in question.

We also develop in Section 4.2 a theory of supports. The support of a gate gg in a GG-symmetric circuit for GSymAG\leq\text{{Sym}}_{A} is a subset of AA which determines both the evaluation of gg and its orbit. We establish in Section 4.4 the support theorem, which connects the minimum size of these supports with the minimum size of the orbits of a gate (and so the size of the circuit). We show in Section 4.3 that if Duplicator has a winning strategy in the game with 2k2k pebbles on a pair of indexed functions then these functions cannot be distinguished by any pair of GG-symmetric circuits with supports of size at most kk. In Section 4.4 we combine this result with the support theorem to show that to prove exponential lower bounds it suffices to establish a linear lower bound on the number of pebbles needed by Spoiler.

4.1 Bijection Games on Indexed function

We now introduce our generalization of Hella’s bijection game. The generalization is in two directions. The first is that we allow the game to be played on arbitrary indexed functions, rather than just graphs or relational structures. The second is that Duplicator is restricted on which bijections it is allowed to play. Specifically, the bijection must be obtained from a fixed initial bijection by composition with a permutation from a fixed group. We recover the usual requirement in the bijection game by just taking this group to be the full symmetric group.

We begin by introducing the key notions of indexed set and function.

Definition 4.1.

Let AA be a set, GSymAG\leq\text{{Sym}}_{A}, and XX be a GG-set. We call the triple (X,A,G)(X,A,G) an indexed set. We call a triple (M,A,G)(M,A,G), where M:XKM:X\rightarrow K is a function on XX, an indexed function.

All of the structures discussed so far may be identified with indexed functions. For example, we can identify a (directed) graph (V,E2)(V,E^{2}) with the indexed function (E:V2{0,1},V,SymV)(E:V^{2}\rightarrow\{0,1\},V,\text{{Sym}}_{V}). Similarly, a bipartitioned graph (V,U,E)(V,U,E) can be identified with the indexed function (E:U×V{0,1},UV,SymU×SymV)(E:U\times V\rightarrow\{0,1\},U\uplus V,\text{{Sym}}_{U}\times\text{{Sym}}_{V}).

We now want to generalize the notion of isomorphism between structures to an equivalence notion on indexed functions, which is suitably restricted by the permutation group GG. Suppose AA and BB are index sets and f:ABf:A\rightarrow B is a bijection. For a group GSymAG\leq\text{{Sym}}_{A}, let HSymBH\leq\text{{Sym}}_{B} be the group of permutations {fπf1πG}\{f\pi f^{-1}\mid\pi\in G\}. Let =(M:XK,A,G)\mathcal{M}=(M:X\rightarrow K,A,G) and 𝒩=(M:YK,B,H)\mathcal{N}=(M:Y\rightarrow K,B,H) be a pair of indexed functions and f^:XY\hat{f}:X\rightarrow Y be a bijection which lifts the bijection ff in the sense that f^(πx)=(fπf1)f^(x)\hat{f}(\pi x)=(f\pi f^{-1})\hat{f}(x) for all πG\pi\in G. We say that the indexed functions \mathcal{M} and 𝒩\mathcal{N} are (f,f^)(f,\hat{f})-isomorophic, if there is a permutation πH\pi\in H such that for all xXx\in X, M(x)=N(πf^(x))M(x)=N(\pi\hat{f}(x)).

In particular, for a pair of directed graphs (V1,E1)(V_{1},E_{1}) and (V2,E2)(V_{2},E_{2}), if ff is any bijection between V1V_{1} and V2V_{2}, and f^\hat{f} is its natural lift to a bijection between V12V_{1}^{2} and V22V_{2}^{2} (i.e. f^(u,v)=(f(u),f(v))\hat{f}(u,v)=(f(u),f(v)), then the pair of indexed functions (E1:V12{0,1},V1,SymV1)(E_{1}:V_{1}^{2}\rightarrow\{0,1\},V_{1},\text{{Sym}}_{V_{1}}) and (E2:V22{0,1},V2,SymV2)(E_{2}:V_{2}^{2}\rightarrow\{0,1\},V_{2},\text{{Sym}}_{V_{2}}) are (f,f^)(f,\hat{f})-isomorphic if, and only if, (V1,E1)(V_{1},E_{1}) and (V2,E2)(V_{2},E_{2}) are isomorphic in the usual sense. Similar statements hold for bipartitioned graphs and other relational structures. As a further example, we can identify n×nn\times n matrices over a field 𝔽\mathbb{F} with structured functions of the form (M:[n]×[n]𝔽,[n][n],G)(M:[n]\times[n]\rightarrow\mathbb{F},[n]\uplus[n],G), for some group GG. Unlike the case of structures considered above, there is not a canonical choice of GG, and it depends on what invariants of the matrix we wish to take into account. For example, if G=SymnG=\text{{Sym}}_{n} then isomorphism corresponds to equivalence under simultaneous row and column permutations and if G=Symn×SymnG=\text{{Sym}}_{n}\times\text{{Sym}}_{n} then isomorphism corresponds to equivalence under separate row and column permutations.

Our notion of isomorphism is parameterized by a fixed bijection between the index sets AA and BB and its lift to the indexed sets XX and YY. To simplify notation, we often simply identify the sets AA and BB, and the sets XX and YY, so that the base bijection can be taken to be the identity. This is akin to fixing the universe of any nn vertex graph to be the set [n][n], so that isomorphisms can be identified with permutations of [n][n].

In order to define the bijection game, we need a notion of partial isomorphism on the set of pebbled positions in an indexed function. In order to introduce this, we first define the notion of a lift.

Definition 4.2.

Let (X,A,G)(X,A,G) be an indexed set. For SAS\subseteq A we define the lift of SS to be the set XS={xXStab(S)Stab(x)}X_{S}=\{x\in X\mid\text{{Stab}}(S)\leq\text{{Stab}}(x)\}.

Thus, XSX_{S} is those elements of XX that are fixed by any permutation that fixes SS pointwise. The following is now an easy observation.

Lemma 4.3.

Let (X,A,G)(X,A,G) be an indexed set. If XSX_{S} is the lift of SAS\subseteq A then SetStab(S)SetStab(XS)\text{{SetStab}}(S)\leq\text{{SetStab}}(X_{S}).

Proof.

Let σSetStab(S)\sigma\in\text{{SetStab}}(S), πStab(S)\pi\in\text{{Stab}}(S), and sSs\in S. Then σ1πσ(s)=s\sigma^{-1}\pi\sigma(s)=s, so σ1πσStab(S)\sigma^{-1}\pi\sigma\in\text{{Stab}}(S). Hence, by definition of XSX_{S}, we have σ1πσ(x)=x\sigma^{-1}\pi\sigma(x)=x for any xXSx\in X_{S}. Hence πσ(x)=σ(x)\pi\sigma(x)=\sigma(x) and therefore πStab(σ(x))\pi\in\text{{Stab}}(\sigma(x)). Since πStab(S)\pi\in\text{{Stab}}(S) was arbitrary, σ(x)XS\sigma(x)\in X_{S} and therefore σSetStab(XS)\sigma\in\text{{SetStab}}(X_{S}). ∎

We define the notion of partial isomorphism specifically for two indexed functions over the same index set. It can easily be generalized to the case with distinct index sets with a fixed bijection between them.

Definition 4.4.

Let (M:XK,A,G)(M:X\rightarrow K,A,G) and (N:XK,A,G)(N:X\rightarrow K,A,G) be two indexed functions, and let SAS\subseteq A and πG\pi\in G. We say that π\pi induces a partial isomorphism on SS if, for each xXSx\in X_{S}, we have M(x)=N(πx)M(x)=N(\pi x).

It is easily seen that this gives the usual notion of partial isomorphism when the indexed functions are graphs or other relational structures. With this, we are ready to define the bijection game on indexed functions. The game is played by two players Spoiler and Duplicator on a pair of indexed functions, using a set of pebbles. During the course of the game, the pebbles are placed on elements of the indexing sets. Where it does not cause confusion, we do not distinguish notationally between the pebbles and the elements on which they are placed.

Definition 4.5.

Let 𝒩=(M:XK,A,G)\mathcal{N}=(M:X\rightarrow K,A,G) and 𝒩=(N:XK,A,G)\mathcal{N}=(N:X\rightarrow K,A,G) be indexed functions. The kk-pebble GG-bijection game on (M,N)(M,N) is defined as follows. The game is played between two players called the Spoiler and Duplicator using two sequences of pebbles a1,,aka_{1},\ldots,a_{k} and b1,,bkb_{1},\ldots,b_{k}. In each round of the game:

  1. 1.

    Spoiler picks a pair of pebbles aia_{i} and bib_{i},

  2. 2.

    Duplicator picks a permutation πG\pi\in G such that for each jij\neq i, π(aj)=bj\pi(a_{j})=b_{j}, and

  3. 3.

    Spoiler chooses some aAa\in A and places aia_{i} on aa and bib_{i} on π(a)\pi(a).

Spoiler has won the game at the end of the round if the permutation π\pi does not induce a partial isomorphism on the set {a1,,ak}\{a_{1},\ldots,a_{k}\}. We say that Duplicator has a winning strategy for the kk-pebble GG-bijection game if it has a strategy to play the game forever without Spoiler winning.

In the sequel, we abbreviate “kk-pebble GG-bijection game” to (G,k)(G,k)-bijection game.

We recover the ordinary bijection game on graphs by taking A=[n]A=[n], X=[n]2X=[n]^{2} and G=SymnG=\text{{Sym}}_{n}. It is clear in this case that if \mathcal{M} and 𝒩\mathcal{N} are isomorphic graphs, then Duplicator has a winning strategy by choosing a fixed isomorphism π\pi between them and playing it at every move. However, if GSymAG\leq\text{{Sym}}_{A} is a more restricted group that does not contain an isomorphism between \mathcal{M} and 𝒩\mathcal{N}, a Duplicator winning strategy is not guaranteed even when \mathcal{M} and 𝒩\mathcal{N} are isomorphic as graphs. This is just the case in the application in Section 5.

4.2 Supports

Lower bounds for symmetric circuits rely on the notion of a support. We define the notion formally below but it is worthwhile developing an intuition. If we have a GG-symmetric circuit CC then the function computed by CC is invariant under permutations in GG on the input. This is not the case for individual gates gg in CC other than the output gate: applying a permutation πG\pi\in G to the inputs of the circuit might yield a different function at gg. But, by the symmetry condition, there is then another gate πg\pi g which computes this other function. If the circuit is small, the orbit of gg is small and thus the stabilizer group of gg is large. That is, for many π\pi we have g=πgg=\pi g. What the support theorem tells us is that in this case, there is a small subset SS (which we call a support of gg) of the permutation domain of GG such that the function computed at gg only depends on SS. This is in the sense that permutations that only move elements of the permutation domain outside of SS do not change the function computed at gg. This support theorem can be proved as long as the group GG is, in a sense, large enough. We now define the notion of a support formally and prove the relationship between support size and orbit size in Section 4.3.

Definition 4.6.

Let XX be a GG-set and HGH\leq G a subgroup of GG. We say that SXS\subseteq X is a support of HH if StabG(S)H\text{{Stab}}_{G}(S)\leq H.

We extend this notion to indexed sets, where the support is now a subset of the indices. A key example is the action of Symn\text{{Sym}}_{n} on the set X=[n]×[n]X=[n]\times[n] or on the collection of matrices M:X{0,1}M:X\rightarrow\{0,1\}.

Definition 4.7.

Let (X,A,G)(X,A,G) be an indexed set. We say that a set SAS\subseteq A is a support of xXx\in X if it is a support of Stab(x)\text{{Stab}}(x).

Note that SS is a support of xx just in case xx is in the lift of SS, in the sense of Definition 4.2.

Let (X,A,G)(X,A,G) be an indexed set. We write max-orbit(X)\text{{max-orbit}}(X) to denote the maximum size of the orbit of xx over all xXx\in X. For xXx\in X we write min-supp(x)\text{{min-supp}}(x) to denote the minimum size of a support of xx and max-supp(X)\text{{max-supp}}(X) to denote the maximum of min-supp(x)\text{{min-supp}}(x) over all xXx\in X.

We now specialise our discussion to circuits. Let CC be a rigid GG-symmetric circuit with gate set DD, where GSymnG\leq\text{{Sym}}_{n} for some natural number nn. Then (D,[n],G)(D,[n],G) is an indexed set. In this way we can speak of the supports and orbits of a gate. We abuse notation slightly and write max-orbit(C)\text{{max-orbit}}(C) and max-supp(C)\text{{max-supp}}(C) to denote max-orbit(D)\text{{max-orbit}}(D) and max-supp(D)\text{{max-supp}}(D), respectively.

4.3 Playing Games on Circuits

We are now ready to prove the first major theorem of this section. This links the number of pebbles in a bijection game with the size of the supports. Structures in which Duplicator has a 2k2k-pebble winning strategy cannot be distinguished by symmetric circuits with supports limited to size kk. The statement of the theorem and argument used generalize that in [7].

Theorem 4.8.

Let (X,A,G)(X,A,G) be an indexed set and let CC be a GG-symmetric circuit with values from KK and variables XX, such that max-supp(C)k\text{{max-supp}}(C)\leq k. If Duplicator has a winning strategy for the (G,2k)(G,2k)-bijection game played on (M,A,G)(M,A,G) and (N,A,G)(N,A,G) for functions M:XKM:X\rightarrow K and N:XKN:X\rightarrow K then C[M]=C[N]C[M]=C[N].

Proof.

We prove this result by contraposition, i.e. we assume C[M]C[N]C[M]\neq C[N] and show that Spoiler has a winning strategy for the (G,2k)(G,2k)-game on (M,N)(M,N).

The strategy we define for Spoiler can be understood informally as follows. We start at the output gate gg and note that for any bijection σG\sigma\in G, C[M](g)C[σN](g)C[M](g)\neq C[\sigma N](g). The goal of Spoiler now over the next (at most) kk rounds is to pebble a support of some child hh of gg such that for Duplicator’s most recent chosen bijection σ\sigma we have that C[M](h)C[σN](h)C[M](h)\neq C[\sigma N](h). Since the bijection chosen by Duplicator can be interpreted as a permutation σG\sigma\in G respecting the pebble configuration, and the fact that an entire support of hh is pebbled and so any such permutation fixes hh, it follows that C[M](h)C[σN](h)C[M](h)\neq C[\sigma N](h) for any choice of σ\sigma. We now iterate this process, aiming next to pebble a support of an appropriate child of hh over the next (at most) kk rounds. We terminate when we reach an input gate, which witnesses that Spoiler has won the game.

So, fix for each gate gg of CC a support of size at most kk. We write sp(g)\text{sp}(g) to denote this support.

Claim 4.9.

Suppose the pebbled position is a\vec{a} and b\vec{b}. Let gg be a gate such that sp(g)a\text{sp}(g)\subseteq\vec{a} and for some σ\sigma that maps a\vec{a} to b\vec{b} we have that C[M](g)C[σN](g)C[M](g)\neq C[\sigma N](g). There exists a strategy for Spoiler such that after at most kk rounds where the pebbled positions are now a\vec{a}^{\prime} and b\vec{b}^{\prime} there is hchild(g)h\in\text{child}(g) such that sp(h)a\text{sp}(h)\subseteq\vec{a}^{\prime} and for all σG\sigma^{\prime}\in G that maps a\vec{a}^{\prime} to b\vec{b}^{\prime} we have that C[M](h)C[σN](h)C[M](h)\neq C[\sigma^{\prime}N](h).

Proof.

We describe an invariant that Spoiler is able to maintain inductively. Suppose that after i0i\geq 0 rounds c1,,cic_{1},\ldots,c_{i} and d1,,did_{1},\ldots,d_{i} have been pebbled. Let SiS_{i} be the pointwise stabiliser of c1,,cic_{1},\ldots,c_{i} in Stab(a)\text{{Stab}}(\vec{a}). Then S0=Stab(a)S_{0}=\text{{Stab}}(\vec{a}). Spoiler ensures there is a gate hichild(g)h_{i}\in\text{child}(g) with c1,,cisp(hi)c_{1},\ldots,c_{i}\in\text{sp}(h_{i}) and a αK\alpha\in K such that

|{hOrbSi(hi)C[M](h)=α}||{hOrbSi(hi)C[σN](h)=α}|\displaystyle|\{h\in\text{{Orb}}_{S_{i}}(h_{i})\mid C[M](h)=\alpha\}|\neq|\{h\in\text{{Orb}}_{S_{i}}(h_{i})\mid C[\sigma N](h)=\alpha\}|

for all σG\sigma\in G that maps a,c1,,ci\vec{a},c_{1},\ldots,c_{i} to b,d1,,di\vec{b},d_{1},\ldots,d_{i}. This suffices to prove the claim since |sp(hi)|k|\text{sp}(h_{i})|\leq k and therefore for some ii, we will have i|sp(hi)|i\geq|\text{sp}(h_{i})| and in this case, OrbSi(hi)\text{{Orb}}_{S_{i}}(h_{i}) is a singleton.

We first consider the case i=0i=0. We have for any σG\sigma\in G that maps a\vec{a} to b\vec{b} that C[M](g)C[σN](g)C[M](g)\neq C[\sigma N](g). Since the basis element labelling gg is a fully symmetric function, there must be some αK\alpha\in K such that |{hchild(g)C[M]h=α}||\{h\in\text{child}(g)\mid C[M]h=\alpha\}| is different from |{hchild(g)C[σN]h=α}||\{h\in\text{child}(g)\mid C[\sigma N]h=\alpha\}|. Moreover, since S0Stab(g)SetStab(child(g))S_{0}\leq\text{{Stab}}(g)\leq\text{{SetStab}}(\text{child}(g)), S0S_{0} partitions child(g)\text{child}(g) into orbits. It follows that for some hchild(g)h\in\text{child}(g)

|{hOrbS0(h)C[M](h)=α}||{hOrbS0(h)C[σN](h)=α}|.\displaystyle|\{h^{\prime}\in\text{{Orb}}_{S_{0}}(h)\mid C[M](h^{\prime})=\alpha\}|\neq|\{h^{\prime}\in\text{{Orb}}_{S_{0}}(h)\mid C[\sigma N](h^{\prime})=\alpha\}|.

Now, take any σG\sigma^{\prime}\in G that maps a\vec{a} to b\vec{b}. Let ρ:=σ1σ\rho:=\sigma^{-1}\sigma^{\prime}. Then ρS0\rho\in S_{0}, and so OrbS0(h)=OrbS0(ρh)\text{{Orb}}_{S_{0}}(h)=\text{{Orb}}_{S_{0}}(\rho h). Moreover, for any hOrbS0(h)h^{\prime}\in\text{{Orb}}_{S_{0}}(h), C[σN](h)=C[σρN](ρh)=C[σN](ρh)C[\sigma N](h^{\prime})=C[\sigma\rho N](\rho h^{\prime})=C[\sigma^{\prime}N](\rho h^{\prime}), and so

|{hOrbS0(h)C[M](h)=α}||{hOrbS0(h)C[σN](h)=α}|.\displaystyle|\{h^{\prime}\in\text{{Orb}}_{S_{0}}(h)\mid C[M](h^{\prime})=\alpha\}|\neq|\{h^{\prime}\in\text{{Orb}}_{S_{0}}(h)\mid C[\sigma^{\prime}N](h^{\prime})=\alpha\}|.

Thus, taking h0:=hh_{0}:=h suffices.

Inductively, suppose Spoiler has maintained the invariant after i>0i>0 rounds. At the start of round i+1i+1 Spoiler picks up a pair of pebbles (aj,bj)(a_{j},b_{j}) where aja_{j} is not placed on c1,,cic_{1},\ldots,c_{i} or an element of sp(g)\text{sp}(g). Suppose Duplicator plays a bijection σi+1\sigma_{i+1}. Then, by the induction hypothesis there is an hih_{i} and an αi\alpha_{i} such that

|{hOrbSi(hi)C[M](h)=αi}||{hOrbSi(hi)C[σiN](h)=αi}|.\displaystyle|\{h^{\prime}\in\text{{Orb}}_{S_{i}}(h_{i})\mid C[M](h^{\prime})=\alpha_{i}\}|\neq|\{h^{\prime}\in\text{{Orb}}_{S_{i}}(h_{i})\mid C[\sigma_{i}N](h^{\prime})=\alpha_{i}\}|.

and c1,,cisp(hi)c_{1},\ldots,c_{i}\in\text{sp}(h_{i}). Let c1,,ci,si+1,,suc_{1},\ldots,c_{i},s_{i+1},\ldots,s_{u} enumerate sp(hi)\text{sp}(h_{i}). Then OrbSi(hi)\text{{Orb}}_{S_{i}}(h_{i}) is partitioned into sets (Oc)cAa{c1,,ci}(O_{c})_{c\in A\setminus\vec{a}\cup\{c_{1},\ldots,c_{i}\}}, where Oc={πhiπ(si+1)=c}O_{c}=\{\pi h_{i}\mid\pi(s_{i+1})=c\}. Thus there is a value ci+1c_{i+1} and αi+1\alpha_{i+1} for which

|{hOci+1C[M](h)=αi+1}||{hOci+1C[σN](h)=αi+1}|.\displaystyle|\{h^{\prime}\in O_{c_{i+1}}\mid C[M](h^{\prime})=\alpha_{i+1}\}|\neq|\{h^{\prime}\in O_{c_{i+1}}\mid C[\sigma N](h^{\prime})=\alpha_{i+1}\}|.

Spoiler then places aja_{j} on ci+1c_{i+1} and bjb_{j} on σi+1(ci+1)\sigma_{i+1}(c_{i+1}). Let hi+1h_{i+1} be any element of Oc+1O_{c+1}. The required invariant is satisfied by construction. ∎

It follows from Claim 4.9 that in at most k0pt(C)k\cdot 0pt(C) rounds Spoiler can force the pebbles to a position such that for some input gate gg labelled by some variable xXx\in X we have: that for any σG\sigma\in G that Duplicator may play we have that C[M](g)C[σN](g)C[M](g)\neq C[\sigma N](g). Thus M(x)σN(x)M(x)\neq\sigma N(x), and so Spoiler wins.

4.4 Bounds on Supports

The main result of this subsection establishes that, for suitable choice of groups GG, if a family of GG-symmetric circuits has subexponential size orbits than it has sublinear size supports. We prove this specifically for the case when GG is an alternating group. The argument extends easily to cases where GG contains a large alternating group, and we derive one such instance in Corollary 4.12. The proof relies on a standard fact about permutation groups, attributed to Jordan and Liebeck, that translates a bound on orbit size to a bound on support size. To understand this, suppose gg is a gate with orbit size (relative to Altn\text{{Alt}}_{n}) bounded by (nk){n\choose k}. Then by the orbit-stabilizer theorem, this is equivalent to [Altn:Stab(g)](nk)[\text{{Alt}}_{n}:\text{{Stab}}(g)]\leq{n\choose k}. One way for Stab(g)\text{{Stab}}(g) to have small index in this way is for it to always contain a large copy of the alternating group. That is Stab(g)\text{{Stab}}(g) contains the alternating group restricted to nkn-k elements, or equivalently gg has a support of size at most kk. Theorem 4.10 asserts that indeed this is the only way.

Theorem 4.10 ([13], Theorem 5.2B).

Let YY be a set such that n:=|Y|>8n:=|Y|>8, and let kk be an integer with 1kn41\leq k\leq\frac{n}{4}. Suppose that GAltYG\leq\text{{Alt}}_{Y} has index [AltY:G]<(nk)[\text{{Alt}}_{Y}:G]<{n\choose k} then there exists XYX\subset Y with |X|<k|X|<k such that StabAltY(X)G\text{{Stab}}_{\text{{Alt}}_{Y}}(X)\leq G.

We derive from Theorem 4.10 the following asymptotic relationship between orbit and support size. An analogous version of this with respect to the symmetric group is stated in the full version of [11] and the proof is largely the same. We include a proof for completeness.

Theorem 4.11.

Let (Cn)(C_{n}) be a family of rigid Altn\text{{Alt}}_{n}-symmetric circuits. If max-orbit(Cn)=2o(n)\text{{max-orbit}}(C_{n})=2^{o(n)} then max-supp(Cn)=o(n)\text{{max-supp}}(C_{n})=o(n).

Proof.

Let kk be the least value such that max-orbitCn(nk)\text{{max-orbit}}{C_{n}}\leq{n\choose k}. By the assumption that max-orbit(Cn)=2o(n)\text{{max-orbit}}(C_{n})=2^{o(n)}, we have that kk is o(n)o(n). Indeed, otherwise there is a constant cc with 0<c<120<c<\frac{1}{2}, such that k1cnk-1\geq cn for infinitely many nn. And since (nl)(nl)l{n\choose l}\geq(\frac{n}{l})^{l} for all ll it follows that (nk1)(ncn)cn>2cn{n\choose{k-1}}\geq(\frac{n}{cn})^{cn}>2^{cn}. Since kk is the least value such that max-orbit(Cn)(nk)\text{{max-orbit}}(C_{n})\leq{n\choose k}, it follows that max-orbit(Cn)2cn\text{{max-orbit}}(C_{n})\geq 2^{cn} for infinitely many nn, contradicting the assumption that max-orbit(Cn)=2o(n)\text{{max-orbit}}(C_{n})=2^{o(n)}.

From k=o(n)k=o(n) it follows that for all large enough nn, kn4k\leq\frac{n}{4}. Then, for any gate gg of CnC_{n}, by the orbit-stabilizer theorem, we have [Altn:Stab(g)](nk)[\text{{Alt}}_{n}:\text{{Stab}}(g)]\leq{n\choose k} and so by Theorem 4.10, gg has a support of size less than kk. ∎

The following corollary establishes the analogue of Theorem 4.11 needed to prove our lower bounds for Altn×Altn\text{{Alt}}_{n}\times\text{{Alt}}_{n}-symmetric circuits.

Corollary 4.12.

Let (Cn)n(C_{n})_{n\in\mathbb{N}} be such that for each nn, CnC_{n} is defined over the matrix of variables Xn={xi,ji,j[n]}X_{n}=\{x_{i,j}\mid i,j\in[n]\} and CnC_{n} is a rigid Altn×Altn\text{{Alt}}_{n}\times\text{{Alt}}_{n}-symmetric circuit. If max-orbit(Cn)=2o(n)\text{{max-orbit}}(C_{n})=2^{o(n)} then max-supp(Cn)=o(n)\text{{max-supp}}(C_{n})=o(n).

Proof.

Fix G=Altn×AltnG=\text{{Alt}}_{n}\times\text{{Alt}}_{n} and let Gr=Altn×{id}GG_{r}=\text{{Alt}}_{n}\times\{\operatorname{id}\}\leq G and Gc={id}×AltnGG_{c}=\{\operatorname{id}\}\times\text{{Alt}}_{n}\leq G be the restriction of the action of GG to the two coordinates. We can think of them as the actions of GG on the rows and columns respectively. For any gate gg in CnC_{n}, let SrS_{r} and ScS_{c} be supports of gg under the action of GrG_{r} and GcG_{c} respectively. Then it is easily seen that SrScS_{r}\cup S_{c} is a support of gg.

Suppose for all gates gg in CnC_{n} we have max-orbit(Cn)=2o(n)\text{{max-orbit}}(C_{n})=2^{o(n)}. Since the orbit of gg under the action of GG includes its orbits under the action of its subgroups GrG_{r} and GcG_{c}, we have that each of OrbGr(g)\text{{Orb}}_{G_{r}}(g) and OrbGc(g)\text{{Orb}}_{G_{c}}(g) has size 2o(n)2^{o(n)} and therefore by Theorem 4.11 SrS_{r} and ScS_{c} can be chosen of size o(n)o(n) and the result follows. ∎

We now combine these results along with Theorem 4.8 to establish the crucial connection between bijection games and exponential lower bounds for symmetric circuits. We prove the result for the particular case of interest to us in this paper, namely for Altn×Altn\text{{Alt}}_{n}\times\text{{Alt}}_{n}-symmetric circuits taking as input n×nn\times n-matrices, but note that it holds more generally.

Theorem 4.13.

Fix a set KK, and for each nn\in\mathbb{N}, let GnG_{n} denote the group Altn×Altn\text{{Alt}}_{n}\times\text{{Alt}}_{n}. Fix a GnG_{n}-set XnX_{n} and let P=(Pn)nP=(P_{n})_{n\in\mathbb{N}} be a family of functions Pn:KXnKP_{n}:K^{X_{n}}\rightarrow K, where PnP_{n} is GnG_{n}-symmetric. Suppose there are infinitely many nn for which there are pairs Mn,Nn:XnKM_{n},N_{n}:X_{n}\rightarrow K such that Pn(Mn)Pn(Nn)P_{n}(M_{n})\neq P_{n}(N_{n}) and Duplicator has a strategy to win the (Gn,k)(G_{n},k)-bijection game played on (Mn,[n][n],Altn×Altn)(M_{n},[n]\uplus[n],\text{{Alt}}_{n}\times\text{{Alt}}_{n}) and (Nn,[n][n],Altn×Altn)(N_{n},[n]\uplus[n],\text{{Alt}}_{n}\times\text{{Alt}}_{n}) for k=Ω(n)k=\Omega(n). Then there is no family of GnG_{n}-symmetric circuits that computes PP and has size 2o(n)2^{o(n)}.

Proof.

From Theorem 4.8 any GnG_{n}-symmetric circuit CnC_{n} that has max-supp(Cn)k/2\text{{max-supp}}(C_{n})\leq k/2 must have that C[Mn]=C[Nn]C[M_{n}]=C[N_{n}], and so cannot compute PnP_{n}. It follows then that any family of GnG_{n}-symmetric circuits that computes PnP_{n} must have max-supp(Cn)=Ω(n)\text{{max-supp}}(C_{n})=\Omega(n) and so from Corollary 4.12 cannot have orbits of size 2o(n)2^{o(n)}, and hence cannot have size 2o(n)2^{o(n)}. ∎

Notice that both the definition of the bijection games and Theorem 4.8 place almost no restriction on the group actions considered. The link between the number of pebbles in the game and the size of the support is robust. However, the application in Theorem 4.13 is for a severely limited group action. This is because the link between support size and orbit size proved in Theorem 4.11 requires the presence of a large alternating group.

5 Lower Bound for the Determinant

We now deploy the machinery we have developed to prove the main lower bound result of this paper.

Theorem 5.1 (Main Theorem).

Let 𝔽\mathbb{F} be a field of characteristic 0. There is no family of Altn×Altn\text{{Alt}}_{n}\times\text{{Alt}}_{n}-symmetric circuits (Cn)n(C_{n})_{n\in\mathbb{N}} over 𝔽\mathbb{F} of size 2o(n)2^{o(n)} computing the determinant over 𝔽\mathbb{F}.

To prove Theorem 5.1 we need to construct for each kk, a pair of n×nn\times n {0,1}\{0,1\}-matrices MkM_{k} and NkN_{k} with n=O(k)n=O(k), det(Mk)det(Nk)\det(M_{k})\neq\det(N_{k}) and such that Duplicator has a winning strategy in the (Altn×Altn,k)(\text{{Alt}}_{n}\times\text{{Alt}}_{n},k)-bijection game played on MkM_{k} and NkN_{k}.

We construct the matrix MkM_{k} as the biadjacency matrix of a bipartite graph Γ\Gamma. The matrix NkN_{k} is obtained from MkM_{k} by interchanging a pair of columns of MkM_{k}. Hence, NkN_{k} is also a biadjacency matrix of Γ\Gamma and det(Mk)=det(Nk)\det(M_{k})=-\det(N_{k}) by construction. Thus, as long as det(Mk)0\det(M_{k})\neq 0 the two determinants are different.

In Section 5.1 we describe the construction of the graph which gives rise to the biadjacency matrix MkM_{k}. This graph is obtained by a CFI construction from a base graph Γ\Gamma satisfying a number of conditions. We show the existence of graphs Γ\Gamma satisfying these conditions in Section 5.2. Then, in Section 5.3 we argue that Duplicator has a winning strategy in the (Altn×Altn,k)(\text{{Alt}}_{n}\times\text{{Alt}}_{n},k)-bijection game played on MkM_{k} and NkN_{k}. We bring it all together in Section 5.4 for a proof of Theorem 5.1.

5.1 Constructing the Graph

In proving the lower bound for symmetric circuits computing the permanent in [11], we adapted a construction due to Cai, Fürer and Immerman [6] of pairs of graphs not distinguished by the kk-dimensional Weisfeiler-Leman algorithm. We showed that we could obtain such a pair of bipartite graphs with different numbers of perfect matchings. Note that saying a pair of graphs are not distinguished by the kk-dimensional Weisfeiler-Leman algorithm is the same as saying that Duplicator has a winning strategy in the (k+1)(k+1)-pebble bijection game, using the full symmetric group. In the present construction, we consider a game played on a pair of isomorphic bipartite graphs but where the set of permissible bijections does not include any isomorphisms between them. Equivalently, we play the game on two distinct biadjacency matrices for the same graph. The graphs we consider are, indeed, exactly the graphs used in [6] except that we have to ensure they are bipartite.

CFI graphs and Determinants

Let Γ=(UV,E)\Gamma=(U\cup V,E) be a 33-regular bipartite graph with bipartition U,VU,V. We obtain the graph Γ^\hat{\Gamma} by replacing each vertex vv, with neigbours x,y,zx,y,z with the ten-vertex gadget depicted in Figure 1. This gadget is described as follows.

  • There is a set denoted IvI_{v} of four inner vertices: a vertex vSv_{S} for each set S{x,y,z}S\subseteq\{x,y,z\} of even size.

  • There is a set denoted OvO_{v} of six outer vertices: two u0,u1u_{0},u_{1} for each u{x,y,z}u\in\{x,y,z\}.

  • There is an edge between vSv_{S} and u1u_{1} if uSu\in S and an edge between vSv_{S} and u0u_{0} if uSu\not\in S.

Corresponding to each edge e={u,v}Ee=\{u,v\}\in E there is a pair of edges that we denote e0e_{0} and e1e_{1} in Γ^\hat{\Gamma}: e0e_{0} connects the vertex u0Ovu_{0}\in O_{v} with v0Ouv_{0}\in O_{u} and e1e_{1} connects the vertex u1Ovu_{1}\in O_{v} with v1Ouv_{1}\in O_{u}

Refer to caption
x0x_{0}
x1x_{1}
y1y_{1}
y0y_{0}
z1z_{1}
z0z_{0}
vv_{\emptyset}
v{y,z}v_{\{y,z\}}
Refer to caption
v{x,z}v_{\{x,z\}}
v{x,y}v_{\{x,y\}}
Refer to caption
Figure 1: A gadget in Γ^\hat{\Gamma} corresponding to vertex vv with neighbours x,y,zx,y,z

Note that the graph Γ^\hat{\Gamma} is bipartite. Indeed, if we let X:=vUIvvVOvX:=\bigcup_{v\in U}I_{v}\cup\bigcup_{v\in V}O_{v} and we let Y:=vVIvvUOvY:=\bigcup_{v\in V}I_{v}\cup\bigcup_{v\in U}O_{v} then it is easily seen that all edges in Γ^\hat{\Gamma} are between XX and YY. Since Γ\Gamma is a 33-regular bipartite graph, it follows that |U|=|V||U|=|V| and therefore |X|=|Y||X|=|Y|. Writing mm for |U||U| and nn for |X|=10m|X|=10m, we obtain a biadjacency n×nn\times n matrix representing the graph Γ^\hat{\Gamma} by fixing a pair of bijections η:X[n]\eta:X\rightarrow[n] and η:Y[n]\eta^{\prime}:Y\rightarrow[n]. The action of the group DnD_{n} divides the collection of all such matrices into two orbits. Letting MM and NN be representatives of the two orbits, we have det(M)=det(N)\det(M)=-\det(N). We next aim to show that det(M)0\det(M)\neq 0, provided that Γ\Gamma has an odd number of perfect matchings.

Suppose we are given bijections η:X[n]\eta:X\rightarrow[n] and η:Y[n]\eta^{\prime}:Y\rightarrow[n] which together determine a biadjacency matrix MM representation of Γ^\hat{\Gamma}. We can also identify each perfect matching in Γ^\hat{\Gamma} with a bijection μ:XY\mu:X\rightarrow Y such that μ(x)\mu(x) is a neighbour of xx for all xXx\in X. We write match(Γ^)\text{match}(\hat{\Gamma}) for the collection of all perfect matchings of Γ^\hat{\Gamma}. Then, the determinant of MM is given by:

det(M)\displaystyle\det(M) =\displaystyle= |{μμmatch(Γ^) with sgn(ημη1)=1}|\displaystyle|\{\mu\mid\mu\in\text{match}(\hat{\Gamma})\text{ with }\text{sgn}(\eta^{\prime}\mu\eta^{-1})=1\}|
|{μμmatch(Γ^) with sgn(ημη1)=1}|.\displaystyle-|\{\mu\mid\mu\in\text{match}(\hat{\Gamma})\text{ with }\text{sgn}(\eta^{\prime}\mu\eta^{-1})=-1\}|.

From now on, we take η\eta and η\eta^{\prime} to be fixed and write sgn(μ)\text{sgn}(\mu) and talk of the sign of a matching μ\mu as short hand for sgn(ημη1)\text{sgn}(\eta^{\prime}\mu\eta^{-1}).

To show that det(M)\det(M) is non-zero, we analyze the structure of the set match(Γ^)\text{match}(\hat{\Gamma}). Note that since η\eta and η\eta^{\prime} are fixed, this imposes a linear order on the sets XX and YY. It also induces a linear order on the sets UU and VV, for instance by saying that u<vu<v, for u,vUu,v\in U if the least element of η(IuOu)\eta(I_{u}\cup O_{u}) is less than the least element of η(IvOv)\eta(I_{v}\cup O_{v}). We make use of this order without further elaboration.

Perfect Matchings

In any perfect matching μ\mu of Γ^\hat{\Gamma}, all four vertices in IvI_{v} for any vUVv\in U\cup V must be matched to vertices in OvO_{v} as they have no neighbours outside this set. Thus, exactly two of the vertices of OvO_{v} are matched to vertices in other gadgets. These two could be two vertices in a pair, e.g. {x0,x1}\{x_{0},x_{1}\} in Figure 1 or they could be from different pairs, e.g. {x0,y0}\{x_{0},y_{0}\}. It is easily checked, by inspection of the gadget, that in either case, removing the two vertices of OvO_{v} from the gadget results in an 8-vertex graph that admits a perfect matching. Indeed, if we remove two vertices in a pair, such as {x0,x1}\{x_{0},x_{1}\} the resulting graph is a two-regular bipartite graph on eight vertices. The graph resulting from removing x0x_{0} and y0y_{0} from the gadget is depicted in Figure 2 and all other cases of removing two vertices of OvO_{v} from different pairs result in a graph isomorphic to this one. We now classify perfect matchings in match(Γ^)\text{match}(\hat{\Gamma}) according to which edges between gadgets are included in the matching.

Say that a matching μmatch(Γ^)\mu\in\text{match}(\hat{\Gamma}) is uniform if for each edge eEe\in E, at most one of the two edges e0e_{0} and e1e_{1} is included in μ\mu. Thus, μ\mu is non-uniform if for some eEe\in E both e0e_{0} and e1e_{1} are included in μ\mu. Our first aim is to show that the non-uniform matchings contribute a net zero to the determinant of MM. Write uni-match(Γ^)\text{uni-match}(\hat{\Gamma}) to denote the set of uniform matchings of Γ^\hat{\Gamma}.

Lemma 5.2.
det(M)\displaystyle\det(M) =\displaystyle= |{μμuni-match(Γ^) with sgn(μ)=1}|\displaystyle|\{\mu\mid\mu\in\text{uni-match}(\hat{\Gamma})\text{ with }\text{sgn}(\mu)=1\}|
|{μμuni-match(Γ^) with sgn(μ)=1}|.\displaystyle-|\{\mu\mid\mu\in\text{uni-match}(\hat{\Gamma})\text{ with }\text{sgn}(\mu)=-1\}|.
Proof.

For a non-uniform matching μ\mu, let eEe\in E be an edge for which both e0e_{0} and e1e_{1} are included in μ\mu and and vUVv\in U\cup V a vertex such that ee is incident on vv. Then, the four vertices in IvI_{v} are matched with the remaining four vertices in OvO_{v} and there are exactly two ways this can be done. To see this, consider the gadget in Figure 1 and suppose that e={v,x}e=\{v,x\} so x0x_{0} and x1x_{1} are matched outside the gadget. Then the only two possible matchings of the remaining eight vertices are vy0;v{x,y}z0;v{y,z}y1;v{x,z}z1v_{\emptyset}-y_{0};v_{\{x,y\}}-z_{0};v_{\{y,z\}}-y_{1};v_{\{x,z\}}-z_{1} and vz0;v{x,y}y1;v{y,z}z1;v{x,z}y0v_{\emptyset}-z_{0};v_{\{x,y\}}-y_{1};v_{\{y,z\}}-z_{1};v_{\{x,z\}}-y_{0}. Note that the second one is obtained from the first by composing with an odd permutation: namely the 44-cycle (y0z0y1z1)(y_{0}z_{0}y_{1}z_{1}). We can then define an involution on the set of non-uniform matchings as follows: map μ\mu to the unique non-uniform matching μ\mu^{\prime} which differs from μ\mu only in the set OvO_{v} corresponding to the vertex vUv\in U which is minimal (in the order on UU) among all vertices of UU incident on some edge ee for which e0,e1e_{0},e_{1} are in μ\mu. This is easily seen to be an involution by our previous observation. Moroever, it takes any matching μ\mu to one of opposite sign.

We conclude that all non-uniform matchings μ\mu come in pairs of opposite sign and therefore cancel out in the expression for the determinant of MM. ∎

Uniform Perfect Matchings.

Our next aim is to count uniform perfect matchings in Γ^\hat{\Gamma} and classify them by sign. Suppose then that μ\mu is a perfect matching that includes for each eEe\in E at most one of the two edges e0e_{0} and e1e_{1}. Let FμEF_{\mu}\subseteq E be the set of those eEe\in E such that exactly one of e0e_{0} and e1e_{1} is in μ\mu. Furthermore, let fμ:Fμ{0,1}f_{\mu}:F_{\mu}\rightarrow\{0,1\} be the function given by fμ(e)=if_{\mu}(e)=i if, and only if, eie_{i} is in μ\mu.

Since for each vUVv\in U\cup V, exactly two of the vertices of OvO_{v} are matched to vertices in other gadgets we can see that FμF_{\mu} includes exactly two edges incident on every vertex vv. In other words, FμF_{\mu} is a 22-factor of Γ\Gamma and therefore has exactly 2m2m edges.

For a 22-factor FF of Γ\Gamma and a function f:F{0,1}f:F\rightarrow\{0,1\}, write μ(F,f)\mu(F,f) for the collection of all matchings μ\mu of Γ^\hat{\Gamma} with Fμ=FF_{\mu}=F and fμ=ff_{\mu}=f.

Lemma 5.3.

There are exactly 22m2^{2m} perfect matchings in μ(F,f)\mu(F,f), for any 22-factor FF of Γ\Gamma and any function f:F{0,1}f:F\rightarrow\{0,1\}.

Proof.

Let vv be a vertex of Γ\Gamma with neighbours x,y,zx,y,z and consider the gadget in Figure 1. Exactly two of the edges incident on vv are included in FF and suppose that it is the edges {v,x}\{v,x\} and {v,y}\{v,y\}. Further, suppose f({v,x})=f({v,y})=0f(\{v,x\})=f(\{v,y\})=0. This means that the matching must pair the vertices x0x_{0} and y0y_{0} with vertices outside the gadget and we verify that the gadget with these two vertices removed admits two distinct perfect matchings. The gadget with vertices x0x_{0} and y0y_{0} removed is pictured in Figure 2, where for clarity, we have removed the set brackets in the subscripts of the vertex labels. It is clear that vv_{\emptyset} must be matched with z0z_{0}. This leaves a six-cycle v{x,y}x1v{x,z}z1v{y,z}y1v_{\{x,y\}}-x_{1}-v_{\{x,z\}}-z_{1}-v_{\{y,z\}}-y_{1} which admits two matchings. Entirely analogously, if f({v,x})=f({v,y})=1f(\{v,x\})=f(\{v,y\})=1 we can consider the gadget with the vertices x1x_{1} and y1y_{1} removed and it is easily checked that the resulting graph is isomorphic to the one depicted in Figure 2, as is also the case when f({v,x})f({v,y})f(\{v,x\})\neq f(\{v,y\}).

Refer to caption
vv_{\emptyset}
z0z_{0}
vxyv_{xy}
vxzv_{xz}
vyzv_{yz}
z1z_{1}
x1x_{1}
y1y_{1}
Figure 2: The gadget from Figure 1 with x0x_{0} and y0y_{0} removed

Since the choice of the matching at each gadget is independent, and there are 2m2m gadgets, one for each vertex in UVU\cup V, we see that there are 22m2^{2m} distinct matchings μ\mu for the fixed choice of 22-factor FF and function f:F{0,1}f:F\rightarrow\{0,1\}. ∎

The proof of Lemma 5.3 shows that the two matchings μ\mu and μ\mu^{\prime} in μ(F,f)\mu(F,f) obtained by varying the choice at just one gadget are related to each other by an even permutation. For instance, the two matchings in Figure 2 are related by the 33-cycle x1y1z1x_{1}y_{1}z_{1}. We can immediately conclude that all 22m2^{2m} matchings in μ(F,f)\mu(F,f) have the same sign.

Lemma 5.4.

For any 22-factor FF of Γ\Gamma, any function f:F{0,1}f:F\rightarrow\{0,1\} and any matchings μ1,μ2μ(F,f)\mu_{1},\mu_{2}\in\mu(F,f), sgn(μ1)=sgn(μ2)\text{sgn}(\mu_{1})=\text{sgn}(\mu_{2}).

We next show that we can go further and show that the sign of any matching in μ(F,f)\mu(F,f) does not depend on the choice of ff.

Lemma 5.5.

For any 22-factor FF of Γ\Gamma, any functions f,g:F{0,1}f,g:F\rightarrow\{0,1\} and any matchings μ1μ(F,f)\mu_{1}\in\mu(F,f) and μ1μ(F,g)\mu_{1}\in\mu(F,g), sgn(μ1)=sgn(μ2)\text{sgn}(\mu_{1})=\text{sgn}(\mu_{2}).

Proof.

It suffices to show the lemma for functions ff and gg which differ at exactly one edge eFe\in F as the result then follows by transitivity. Moreover, by Lemma 5.4, it suffices to choose any μ1μ(F,f)\mu_{1}\in\mu(F,f) and μ2μ(F,g)\mu_{2}\in\mu(F,g) and show that they have the same sign

So, assume f(e)=0f(e)=0 and g(e)=1g(e)=1 and ff and gg agree on all other edges. Let e={v,x}e=\{v,x\} where vv has neighbours x,y,zx,y,z and xx has neighbours v,u,wv,u,w. Without loss of generality, assume {v,y}\{v,y\} and {x,w}\{x,w\} are in FF with f({v,y})=f({x,w})=g({v,y})=g({x,w})=0f(\{v,y\})=f(\{x,w\})=g(\{v,y\})=g(\{x,w\})=0. By assumption, μ1\mu_{1} includes the edge x0v0x_{0}-v_{0} and we can assume further that it includes the edges vz0v_{\emptyset}-z_{0}, v{x,y}x1v_{\{x,y\}}-x_{1} in the gadget corresponding to vv and the edges xu0x_{\emptyset}-u_{0}, v{v,w}v1v_{\{v,w\}}-v_{1} in the gadget corresponding to xx. In other words, it contains alternating edges along the 1010-cycle depicted in Figure 3.

Refer to caption
z0z_{0}
vv_{\emptyset}
vxyv_{xy}
x1x_{1}
x0x_{0}
v0v_{0}
v1v_{1}
xx_{\emptyset}
xvwx_{vw}
u0u_{0}
Figure 3: An alternating cycle between two matchings in μ(F,f)\mu(F,f) and μ(F,g)\mu(F,g) where ff and gg differ on the edge {v,x}\{v,x\}.

Now, choose μ2\mu_{2} to be the symmetric difference between μ1\mu_{1} and the cycle in Figure 3. This is also a perfect matching and it is in μ(F,g)\mu(F,g) as the only difference with μ1\mu_{1} as far as edges between gadgets is concerned is that μ2\mu_{2} contains x1v1x_{1}-v_{1} rather than x0v0x_{0}-v_{0}. Moreover, it is easily seen that μ2\mu_{2} is obtained from μ1\mu_{1} by composition with an even permutation: namely the 55-cycle: (x0z0x1x{v,w}x)(x_{0}z_{0}x_{1}x_{\{v,w\}}x_{\emptyset}). ∎

With this, we are now ready to establish the main result of this section.

Lemma 5.6.

If Γ\Gamma has an odd number of perfect matchings, then det(M)0\det(M)\neq 0.

Proof.

By Lemma 5.2, we have

det(M)\displaystyle\det(M) =\displaystyle= |{μμuni-match(Γ^) with sgn(μ)=1}|\displaystyle|\{\mu\mid\mu\in\text{uni-match}(\hat{\Gamma})\text{ with }\text{sgn}(\mu)=1\}|
|{μμuni-match(Γ^) with sgn(μ)=1}|.\displaystyle-|\{\mu\mid\mu\in\text{uni-match}(\hat{\Gamma})\text{ with }\text{sgn}(\mu)=-1\}|.

For any 22-factor FF of Γ\Gamma write μ(F)\mu(F) for f:F{0,1}μ(F,f)\bigcup_{f:F\rightarrow\{0,1\}}\mu(F,f) and note that by Lemma 5.3 |μ(F)|=24m|\mu(F)|=2^{4m} for all FF. Moreover, by Lemma 5.5 all matchings in μ(F)\mu(F) have the same sign. Thus, we define sgn(F)\text{sgn}(F) to be sgn(μ)\text{sgn}(\mu) for any μμ(F)\mu\in\mu(F). Hence, we have that

det(M)=24mFsgn(F),\det(M)=2^{4m}\sum_{F}\text{sgn}(F),

where the sum is over all 22-factors of Γ\Gamma.

Since Γ\Gamma is 33-regular, the number of 22-factors of Γ\Gamma is exactly the number of perfect matchings. Indeed, the complement of any 22-factor is a perfect matching and this gives a bijection between the collection of perfect matchings and 22-factors. Thus, since the number of perfect matchings of Γ\Gamma is odd, so is the number of 22-factors and we conclude that the sum Fsgn(F)\sum_{F}\text{sgn}(F) cannot be zero, proving the result. ∎

5.2 Graphs with Odd Number of Perfect Matchings

We have seen that if Γ\Gamma has an odd number of perfect matchings, then the matrices MM and NN have different determinant. In order to play the bijection game on MM and NN we also need Γ\Gamma to be well connected. We now show that we can find suitable graphs that satisfy both of these conditions simultaneously.

For a positive integer kk, say that Γ\Gamma is kk-well-connected if any balanced separator of Γ\Gamma has size greater than kk. For our construction, we need 33-regular bipartite graphs on 2n2n vertices which are kk-well-connected for k=Ω(n)k=\Omega(n) and which have an odd number of perfect matchings. The main purpose of this section is to prove the existence of such a family of graphs.

Theorem 5.7.

For all positive integers nn there is a bipartite graph Γn=(U,V,E)\Gamma_{n}=(U,V,E) satisfying the following conditions:

  1. 1.

    |U|=|V|=n|U|=|V|=n;

  2. 2.

    Γn\Gamma_{n} is 33-regular;

  3. 3.

    Γn\Gamma_{n} is kk-well-connected for k=Ω(n)k=\Omega(n); and

  4. 4.

    Γn\Gamma_{n} has an odd number of perfect matchings.

We prove the existence of these graphs by a randomized construction. To be precise, a random 33-regular bipartite graph on 2n2n vertices satisfies the third condition with high probability. We show that it can be modified to satisfy the fourth condition while keeping the connectivity high. To do this, we need some facts about the distribution of random 33-regular bipartite graphs.

Fix UU and VV to be two disjoint sets of nn vertices, and we are interested in the uniform distribution on 33-regular bipartite graphs on the vertices UU and VV. This distribution is not easy to sample from but it is known to be well-approximated by a number of other random models, including the union of disjoint random matchings, which we now describe. We say that a pair of bijections π,σ:UV\pi,\sigma:U\rightarrow V is disjoint if there is no uUu\in U with π(u)=σ(u)\pi(u)=\sigma(u). Now consider a random graph 𝒢\mathcal{G} obtained by the following process:

  1. 1.

    choose, uniformly at random, three bijections π1,π2,π3:UV\pi_{1},\pi_{2},\pi_{3}:U\rightarrow V;

  2. 2.

    if for some j{1,2,3}j\in\{1,2,3\} with iji\neq j, πi\pi_{i} and πj\pi_{j} are not disjoint discard this choice of bijections; otherwise

  3. 3.

    let 𝒢\mathcal{G} be the bipartite graph with parts UU and VV edges {{u,πi(u)}i{1,2,3}}\{\{u,\pi_{i}(u)\}\mid i\in\{1,2,3\}\}.

The random graph model obtained in this way is known to be contiguous to the uniform distribution on 33-regular bipartite graphs [23]. This means that any property that holds asymptotically almost surely in one also holds so in the other. The property we are interested in is that of being an expander. It is known [4] that a random 33-regular bipartite graph is an expander with probability tending to 11. This result is, in fact, proved in the configuration model of Bollobas [3] but this is also known to be contiguous to the uniform distribution. We can therefore conclude that the same is true for the random graph 𝒢\mathcal{G}.

Lemma 5.8.

There is a constant α>0\alpha>0 such that with probability tending to 11, 𝒢\mathcal{G} is an α\alpha-expander.

An immediate consequence of this is that with high probability, 𝒢\mathcal{G} is ϵn\epsilon n-well-connected for some constant ϵ\epsilon. We now describe how we can obtain from 𝒢\mathcal{G} a graph which also has an odd number of perfect matchings.

Let BΓB_{\Gamma} denote the biadjacency matrix of Γ=(U,V,E)\Gamma=(U,V,E) with rows indexed by UU and columns by VV. Then the permanent of BΓB_{\Gamma} over a field of characteristic pp is exactly the number of perfect matchings in Γ\Gamma modulo pp. In particular, when p=2p=2, since the permanent is the same as the determinant, we have that the number of perfect matchings in Γ\Gamma is odd if, and only if, det(BΓ)0\det(B_{\Gamma})\neq 0, where the determinant is over 𝔽2\mathbb{F}_{2}. We do not expect that det(B𝒢)0\det(B_{\mathcal{G}})\neq 0 with high probability. To prove Theorem 5.7 it would suffice to show that this is the case with positive probability and this does seem likely to be true. However, we adopt an indirect approach. We show that with probability bounded away from zero rk(B𝒢)\text{{rk}}(B_{\mathcal{G}}) is at least no(n)n-o(n). And, we then show that we can transform any graph Γ\Gamma with rk(BΓ)<n\text{{rk}}(B_{\Gamma})<n to a graph Γ\Gamma^{\prime} so that rk(BΓ)>rk(BΓ)+1\text{{rk}}(B_{\Gamma^{\prime}})>\text{{rk}}(B_{\Gamma})+1 and Γ\Gamma^{\prime} is still well-connected if Γ\Gamma is. Together these give us the graphs we want. We next introduce some notation and terminology that is helpful in establishing these two facts.

In what follows, we treat the biadjacency matrix BΓB_{\Gamma} of a graph Γ\Gamma as being a matrix over 𝔽2\mathbb{F}_{2} and so all arithmetic operations on elements of the matrix should be taken as being over this field.

Lemma 5.9.

There is a constant ϵ\epsilon such that for all sufficiently large nn, Pr[rk(B𝒢)nϵlogn]1/2\Pr[\text{{rk}}(B_{\mathcal{G}})\geq n-\epsilon\log n]\geq 1/2.

Proof.

For any matrix AA, if rk(A)=nc\text{{rk}}(A)=n-c, then the dimension of the null space of AA is cc, so there are 2c12^{c}-1 non-zero vectors 𝐱\mathbf{x} such that AT𝐱=0A^{T}\mathbf{x}=0. We show that for the random graph 𝒢\mathcal{G}, the expected number of vectors 𝐱𝔽2\mathbf{x}\in\mathbb{F}_{2} such that B𝒢T𝐱=0B_{\mathcal{G}}^{T}\mathbf{x}=0 is at most linear in nn222This is a very loose upper bound that suffices for our purposes. By more careful analysis, we can easily show that this expectation is bounded by a constant. Numerical simulations suggests that it in fact tends to 11 as nn goes to infinity.. Let 𝐗\mathbf{X} denote the random variable that is the number of such vectors.

For an element uUu\in U, write rur_{u} for the row of B𝒢B_{\mathcal{G}} indexed by uu. For a vector 𝐱𝔽2U\mathbf{x}\in\mathbb{F}_{2}^{U}, let R𝐱R_{\mathbf{x}} be the set {uU𝐱u=1}\{u\in U\mid\mathbf{x}_{u}=1\} and note that the condition B𝒢T𝐱=0B_{\mathcal{G}}^{T}\mathbf{x}=0 is equivalent to the statement uR𝐱ru=0\sum_{u\in R_{\mathbf{x}}}r_{u}=0.

For any set SUS\subseteq U and a bijection π:UV\pi:U\rightarrow V, we write π(S)\pi(S) to denote the image of SS under π\pi. Say that SS is a zero-sum set if SS is non-empty with uSru=0\sum_{u\in S}r_{u}=0. If SS is a zero-sum set, it must be the case that any vertex vv in π1(S)π2(S)π3(S)\pi_{1}(S)\cup\pi_{2}(S)\cup\pi_{3}(S) has exactly two neighbours in SS. If |S|=m|S|=m, this implies that |π1(S)π2(S)|=|π2(S)π3(S)|=|π1(S)π3(S)|=m/2|\pi_{1}(S)\cap\pi_{2}(S)|=|\pi_{2}(S)\cap\pi_{3}(S)|=|\pi_{1}(S)\cap\pi_{3}(S)|=m/2. In particular mm is even, and m2n/3m\leq 2n/3.

We now estimate the expected number of zero-sum sets of size m=2lm=2l in the random graph 𝒢\mathcal{G}. Fix a set SUS\subseteq U of size mm and choose three permutations π1,π2,π3\pi_{1},\pi_{2},\pi_{3} of [n][n] independently at random. Note that a random choice of π1\pi_{1} means that π1(S)\pi_{1}(S) is a uniformly random subset of VV of size mm. Thus, the probability that |π1(S)π2(S)|=m/2=l|\pi_{1}(S)\cap\pi_{2}(S)|=m/2=l is:

p(n,l):=(2ll)(n2ll)(n2l).p(n,l):=\frac{{{2l}\choose l}{{n-2l}\choose l}}{{n\choose{2l}}}.

For SS to be a zero-sum set, we further require that π3(S)=π1(S)π2(S)\pi_{3}(S)=\pi_{1}(S)\triangle\pi_{2}(S). The probability that a randomly chosen π3\pi_{3} gives exactly this set is 1/(n2l)1/{n\choose{2l}}. Summing over the (n2l){n\choose{2l}} choices of the set SS, we get that the expected number of sets of size m=2lm=2l satisfying the condition |π1(S)π2(S)|=|π2(S)π3(S)|=|π1(S)π3(S)|=m/2|\pi_{1}(S)\cap\pi_{2}(S)|=|\pi_{2}(S)\cap\pi_{3}(S)|=|\pi_{1}(S)\cap\pi_{3}(S)|=m/2, taken over all choices of π1,π2,π3\pi_{1},\pi_{2},\pi_{3} is p(n,l)p(n,l). Now, the probability that two random permutations are not disjoint is the same as the probability that a random permutation contains a fixed-point, which is well known to tend to 1/e1/e from above as nn goes to infinity. Hence we see that the probability that all three permutations are disjoint is at least 1/e31/e^{3} and therefore the expected number of zero-sum sets in 𝒢\mathcal{G} of size 2l2l is at most e3p(n,l)e^{3}p(n,l).

Hence the total expectation of the number of zero-sum sets is

E[𝐗]=1ln/3e3p(n,l).\text{E}[\mathbf{X}]=\sum_{1\leq l\leq n/3}e^{3}p(n,l).

Since p(n,l)<1p(n,l)<1 for all nn and ll, we get E[𝐗]<e3n/3\text{E}[\mathbf{X}]<e^{3}n/3. By Markov’s inequality, it follows that the probability that 𝐗\mathbf{X} exceeds 2e3n/32e^{3}n/3 is less than 1/21/2. Since the dimension of the null space of BΓB_{\Gamma} is log(𝐗+1)\log(\mathbf{X}+1), the theorem follows. ∎

To complete the construction, we show that if Γ\Gamma is a 33-regular bipartite graph with n×nn\times n biadjacency matrix BΓB_{\Gamma} and rk(B)<n\text{{rk}}(B)<n, then under mild assumptions satisfied by almost all such graphs, we can edit Γ\Gamma to get Γ\Gamma^{\prime} so that rk(BΓ)>rk(BΓ)\text{{rk}}(B_{\Gamma^{\prime}})>\text{{rk}}(B_{\Gamma}) and Γ\Gamma^{\prime} is at least (k4)(k-4)-well-connected if Γ\Gamma is kk-well-connected.

Assume then, that Γ\Gamma is a 33-regular graph on two sets UU and VV of nn vertices each and BB is its biadjacency matrix. As before, we write ruBr_{u}^{B} to denote the row of BB indexed by uUu\in U. We drop the superscript BB where it is clear from context. We always treat these rows as vectors in 𝔽2V\mathbb{F}_{2}^{V}.

We say that a pair of edges e1={u1,v1}e_{1}=\{u_{1},v_{1}\} and e2={u2,v2}e_{2}=\{u_{2},v_{2}\} of Γ\Gamma with u1,u2Uu_{1},u_{2}\in U and v1,v2Vv_{1},v_{2}\in V are switchable if they are disjoint and neither of {u2,v1}\{u_{2},v_{1}\} nor {u1,v2}\{u_{1},v_{2}\} is an edge. For a switchable pair e1,e2e_{1},e_{2} we denote by Γ~e1,e2\tilde{\Gamma}_{e_{1},e_{2}} the graph obtained from Γ\Gamma by exchanging the two edges e1e_{1} and e2e_{2}. That is, Γ~e1,e2\tilde{\Gamma}_{e_{1},e_{2}} is the bipartite graph on the vertices UU, VV with edge set

E(Γ){e1,e2}{{u1,v2},{u2,v1}}.E(\Gamma)\setminus\{e_{1},e_{2}\}\cup\{\{u_{1},v_{2}\},\{u_{2},v_{1}\}\}.

Note that Γ~e1,e2\tilde{\Gamma}_{e_{1},e_{2}} is also a 33-regular bipartite graph. We write B~e1,e2\tilde{B}_{e_{1},e_{2}} for the biadjacency matrix of Γ~e1,e2\tilde{\Gamma}_{e_{1},e_{2}}.

Assume now that rk(B)<n\text{{rk}}(B)<n. Then BB has a zero-sum set, i.e. a set SUS\subseteq U such that uSru=0\sum_{u\in S}r_{u}=0. Moreover, by the 33-regularity of Γ\Gamma, we have 2|S|2n/32\leq|S|\leq 2n/3. We can now state the lemma we aim to prove.

Lemma 5.10.

If BB has a zero-sum set SS with |S|<2n/3|S|<2n/3, then there are switchable edges e1,e2E(Γ)e_{1},e_{2}\in E(\Gamma) so that rk(B~e1,e2)>rk(B)\text{{rk}}(\tilde{B}_{e_{1},e_{2}})>\text{{rk}}(B).

Proof.

Fix a zero-sum set SS in BB with |S|<2n/3|S|<2n/3. Let N(S)VN(S)\subseteq V denote the set of elements of VV which are neighbours in Γ\Gamma of vertices in SS. The assumption on the size of SS implies that |N(S)|<n|N(S)|<n and so VN(S)V\setminus N(S)\neq\emptyset.

For i,jVi,j\in V, write tijt_{ij} for the vector in 𝔽2V\mathbb{F}_{2}^{V} which is 11 at positions ii and jj and 0 everywhere else. Now consider the following set of vectors in 𝔽2V\mathbb{F}_{2}^{V}:

T={tijiN(S) and jVN(S)}.T=\{t_{ij}\mid i\in N(S)\text{ and }j\in V\setminus N(S)\}.

Observe that span(T)=E\mathrm{span}(T)=E where EE is the subspace of 𝔽2V\mathbb{F}_{2}^{V} consisting of vectors with even Hamming weight. To see this, first observe that tijspan(T)t_{ij}\in\mathrm{span}(T) for all pairs i,jVi,j\in V. When iN(S)i\in N(S) and jN(S)j\not\in N(S), this is true by definition of TT. For i,jN(S)i,j\in N(S), choose any kN(S)k\not\in N(S) and note that tik+tjk=tijt_{ik}+t_{jk}=t_{ij}. Similarly, if i,jN(S)i,j\not\in N(S), we can pick a kN(S)k\in N(S) and again tki+tkj=tijt_{ki}+t_{kj}=t_{ij}. Since the collection of vectors {tiji,jV}\{t_{ij}\mid i,j\in V\} clearly spans EE, we are done.

Let R={ruuU}R=\{r_{u}\mid u\in U\} be the set of rows of BB. Since each rur_{u} has Hamming weight 33, ruEr_{u}\not\in E so span(R)E\mathrm{span}(R)\not\subseteq E. Since dim(E)=n1rk(B)=dim(span(R))\dim(E)=n-1\geq\text{{rk}}(B)=\dim(\mathrm{span}(R)), we conclude that Tspan(R)T\not\subseteq\mathrm{span}(R). Let us fix a tijTt_{ij}\in T such that tijspan(R)t_{ij}\not\in\mathrm{span}(R).

Since iN(S)i\in N(S) and SS is a zero-sum set, ii has exactly two neighbours in SS and one in USU\setminus S. On the other hand, jj has three neighbours, all in USU\setminus S. Thus, we can choose a kSk\in S which is a neighbour of ii but not jj and an lUSl\in U\setminus S which is a neighbour of jj but not of ii. Let e1={i,k}e_{1}=\{i,k\} and e2={j,l}e_{2}=\{j,l\} and observe that this pair is switchable by construction.

To prove the lemma, it then suffices to prove that rk(B~e1,e2)>rk(B)\text{{rk}}(\tilde{B}_{e_{1},e_{2}})>\text{{rk}}(B). We do this by establishing the following two facts.

  1. 1.

    uSruB~e1,e20\sum_{u\in S}r_{u}^{\tilde{B}_{e_{1},e_{2}}}\neq 0; and

  2. 2.

    for any SUS^{\prime}\subseteq U, if uSruB~e1,e2=0\sum_{u\in S^{\prime}}r_{u}^{\tilde{B}_{e_{1},e_{2}}}=0, then uSruB=0\sum_{u\in S^{\prime}}r_{u}^{B}=0.

Together these establish that the null space of B~e1,e2\tilde{B}_{e_{1},e_{2}} is strictly smaller than that of BB and hence the claim.

Note that for all uU{k,l}u\in U\setminus\{k,l\}, we have ruB~e1,e2=ruBr_{u}^{\tilde{B}_{e_{1},e_{2}}}=r_{u}^{B}, while rkB~e1,e2=rkB+tijr_{k}^{\tilde{B}_{e_{1},e_{2}}}=r_{k}^{B}+t_{ij} and rlB~e1,e2=rlB+tijr_{l}^{\tilde{B}_{e_{1},e_{2}}}=r_{l}^{B}+t_{ij}.

Thus, to prove the first fact, just note that since kk is in SS and ll is not, uSruB~e1,e2=uSruB+tij=tij\sum_{u\in S}r_{u}^{\tilde{B}_{e_{1},e_{2}}}=\sum_{u\in S}r_{u}^{B}+t_{ij}=t_{ij}.

To prove the second fact, consider any set SUS^{\prime}\subseteq U such that uSruB~e1,e2=0\sum_{u\in S^{\prime}}r_{u}^{\tilde{B}_{e_{1},e_{2}}}=0. We consider the following cases.

  • If neither kk nor ll is in SS^{\prime}, then uSruB~e1,e2=uSruB\sum_{u\in S^{\prime}}r_{u}^{\tilde{B}_{e_{1},e_{2}}}=\sum_{u\in S^{\prime}}r_{u}^{B}, so since the former sum is 0, so is the latter.

  • If both kk and ll are in SS^{\prime}, then uSruB~e1,e2=uSruB+2tij=uSruB\sum_{u\in S^{\prime}}r_{u}^{\tilde{B}_{e_{1},e_{2}}}=\sum_{u\in S^{\prime}}r_{u}^{B}+2t_{ij}=\sum_{u\in S^{\prime}}r_{u}^{B}, and the same argument applies.

  • If exactly one of kk and ll is in SS^{\prime}, then uSruB~e1,e2=uSruB+tij\sum_{u\in S^{\prime}}r_{u}^{\tilde{B}_{e_{1},e_{2}}}=\sum_{u\in S^{\prime}}r_{u}^{B}+t_{ij}. Hence, if uSruB~e1,e2=0\sum_{u\in S^{\prime}}r_{u}^{\tilde{B}_{e_{1},e_{2}}}=0, we must have uSruB=tij\sum_{u\in S^{\prime}}r_{u}^{B}=t_{ij}. However, ii and jj were chosen so that tijspan(R)t_{ij}\not\in\mathrm{span}(R), so this is impossible.

Proof of Theorem 5.7.

By Lemma 5.8, for large enough values of nn, the random 33-regular graph 𝒢\mathcal{G} is τn\tau n-well-connected for some constant τ>0\tau>0 with probability tending to 11. Thus, with high probability, the first three conditions are satisfied. If the biadjacency matrix BΓB_{\Gamma} of the resulting graph Γ\Gamma has rank nn, we are done.

If rk(BΓ)<n\text{{rk}}(B_{\Gamma})<n, then with probability at least 1/21/2, we have rk(BΓ)nϵlogn\text{{rk}}(B_{\Gamma})\geq n-\epsilon\log n by Lemma 5.9. Moreover, since the expected number of zero-sum sets of size exactly 2n/32n/3 is at most e3p(n,n/3)e^{3}p(n,n/3) by the argument in the proof of Lemma 5.9, and this value tends to 0 as nn grows, with high probability, BΓB_{\Gamma} has no zero-sum sets of this size. Hence, with positive probability, 𝒢\mathcal{G} satisfies the pre-conditions of Lemma 5.10.

Note that if Γ\Gamma satisfies the conditions of Lemma 5.10, then any zero-sum set in the graph Γ~e1,e2\tilde{\Gamma}_{e_{1},e_{2}} is also a zero-sum set in Γ\Gamma. Hence, if Γ\Gamma contains no zero-sum sets of size exactly 2n/32n/3, the same is true of Γ~e1,e2\tilde{\Gamma}_{e_{1},e_{2}}. We can thus repeatedly apply the construction of Lemma 5.10 to obtain a graph Γ\Gamma^{\prime} such that rk(BΓ)=n\text{{rk}}(B_{\Gamma})=n. It then follows that Γ\Gamma^{\prime} has an odd number of perfect matchings. It remains to argue that Γ\Gamma^{\prime} is still well-connected. Note that since rk(BΓ)nϵlogn\text{{rk}}(B_{\Gamma})\geq n-\epsilon\log n , we get Γ\Gamma^{\prime} from Γ\Gamma by at most ϵlogn\epsilon\log n applications of Lemma 5.10. At each step edges incident at most 44 vertices are modified. Thus, if SS is a balanced separator in Γ\Gamma^{\prime}, then adding these four vertices to it gives us a balanced separator in Γ\Gamma. Thus, if Γ\Gamma is kk-well-connected, then Γ\Gamma^{\prime} is at least (k4)(k-4)-well-connected and the result is proved.

5.3 Playing the Game

Suppose Γ\Gamma is a bipartite 33-regular graph on two sets UU and VV of mm vertices each that is (k+3)(k+3)-well-connected. Let Γ^\hat{\Gamma} be the CFI-graph constructed from Γ\Gamma as described in Section 5.1, and MM and NN be two biadjacency matrices for Γ^\hat{\Gamma} where NN is obtained from MM by interchanging exactly one pair of columns. We aim to prove that Duplicator has a winning strategy in the (Altn×Altn,k)(\text{{Alt}}_{n}\times\text{{Alt}}_{n},k)-bijection game played on MM and NN.

Recall that Γ^\hat{\Gamma} is a bipartite graph on two sets XX and YY of n=10mn=10m vertices each with X=uUIuvVOvX=\bigcup_{u\in U}I_{u}\cup\bigcup_{v\in V}O_{v} and Y=vVIuuUOuY=\bigcup_{v\in V}I_{u}\cup\bigcup_{u\in U}O_{u}. To say that Duplicator has a winning strategy in the (Altn×Altn,k)(\text{{Alt}}_{n}\times\text{{Alt}}_{n},k)-bijection game played on MM and NN is the same as saying that Duplicator has a winning strategy in the (Altn×Altn,k)(\text{{Alt}}_{n}\times\text{{Alt}}_{n},k)-bijection game played on the pair of graphs Γ^\hat{\Gamma} and Γ^\hat{\Gamma}^{\prime} where the latter is obtained from Γ^\hat{\Gamma} by swapping two elements y,yYy,y^{\prime}\in Y. It does not matter which two elements y,yy,y^{\prime} we choose as any swap can be obtained from (yy)(yy^{\prime}) by composing with a permutation in AltY\text{{Alt}}_{Y}. So, for some uUu\in U, fix two vertices x0,x1Oux_{0},x_{1}\in O_{u} which form a single pair in the gadget corresponding to uu and let α=(x0x1)\alpha=(x_{0}x_{1}). We write αΓ^\alpha\hat{\Gamma} for the graph on XYX\cup Y which is exactly the same as Γ^\hat{\Gamma} except the neighbours of x0x_{0} in αΓ^\alpha\hat{\Gamma} are exactly the neighbours of x1x_{1} in Γ^\hat{\Gamma} and the neighbours of x1x_{1} in αΓ^\alpha\hat{\Gamma} are exactly the neighbours of x0x_{0} in Γ^\hat{\Gamma}.

Lemma 5.11.

Duplicator has a winning strategy in the (AltX×AltY,k)(\text{{Alt}}_{X}\times\text{{Alt}}_{Y},k)-bijection game played on the graphs Γ^\hat{\Gamma} and αΓ^\alpha\hat{\Gamma}.

To prove this lemma, we first introduce some notation and some observations about Γ^\hat{\Gamma}. Consider permutations of the vertices IvOvI_{v}\cup O_{v} in the gadget in Figure 1 which fix each of the sets {x0,x1}\{x_{0},x_{1}\}, {y0,y1}\{y_{0},y_{1}\} and {z0,z1}\{z_{0},z_{1}\} setwise. It is easily checked (and this is the key property of the gadget) that for any two of these three sets there is an autormorphism of the gadget that swaps the two vertices inside the two sets while leaving the two vertices in the third set fixed. For instance, there is an automorphism which we denote βxyv\beta^{v}_{xy}, that exchanges x0x_{0} with x1x_{1} and y0y_{0} with y1y_{1}. The action of this automorphism on IvI_{v} is to swap vv_{\emptyset} with v{x,y}v_{\{x,y\}} and v{x,z}v_{\{x,z\}} with v{y,z}v_{\{y,z\}}. Each such automorphism consists of two swaps in IvI_{v} and two in OvO_{v} and so is a permutation in AltX×AltY\text{{Alt}}_{X}\times\text{{Alt}}_{Y}.

We can compose such automorphisms of the individual gadgets to get certain automorphisms of Γ^\hat{\Gamma}. Let C=v1vlC=v_{1}\cdots v_{l} be a simple cycle in the graph Γ\Gamma. That is, there are edges from v1v_{1} to vi+1v_{i+1} for each ii with 1i<l1\leq i<l and an edge from vlv_{l} to v1v_{1}. We define the permutation βC\beta_{C} of XYX\cup Y as the composition

βvlv2v1βv1v3v2βvi1vi+1viβvl1v1vl.\beta^{v_{1}}_{v_{l}v_{2}}\beta^{v_{2}}_{v_{1}v_{3}}\cdots\beta^{v_{i}}_{v_{i-1}v_{i+1}}\cdots\beta^{v_{l}}_{v_{l-1}v_{1}}.

This is easily seen to be an automorphism of Γ^\hat{\Gamma}. Since it is the compostion of permutations each of which is in AltX×AltY\text{{Alt}}_{X}\times\text{{Alt}}_{Y}, βCAltX×AltY\beta_{C}\in\text{{Alt}}_{X}\times\text{{Alt}}_{Y}.

Now, say that a permutation β\beta of XYX\cup Y is coherent if for each vUVv\in U\cup V, β(Iv)=Iv\beta(I_{v})=I_{v} and β(Ov)=Ov\beta(O_{v})=O_{v}. We are only interested in coherent permutations. Let vertices uUu\in U and vVv\in V be neighbours in Γ\Gamma and let v0,v1v_{0},v_{1} be the pair of vertices in OuO_{u} which connect to vertices in OvO_{v}. We say that a coherent permutation β\beta is good bar uvuv if composing it with the swap (v0v1)(v_{0}v_{1}) yields an automorphism of Γ^\hat{\Gamma}. We are now ready to describe the Duplicator winning strategy.

Proof of Lemma 5.11.

We describe Duplicator’s winning strategy. The position at any point of the game with Spoiler to move, consists of up to kk pebbled vertices of Γ^\hat{\Gamma} along with a permutation β\beta of XYX\cup Y which is obtained from the initial permutation α\alpha by means of composing it with an element of AltX×AltY\text{{Alt}}_{X}\times\text{{Alt}}_{Y}. If xx is the vertex covered by pebble ii, let piUVp_{i}\in U\cup V be the vertex of Γ\Gamma such that xIpiOpix\in I_{p_{i}}\cup O_{p_{i}}. In other words p1,,pkp_{1},\ldots,p_{k} enumerate the vertices of Γ\Gamma whose gadgets in Γ^\hat{\Gamma} contain the pebbled vertices. Note that by the connectedness assumption, Γ{p1,,pk}\Gamma\setminus\{p_{1},\ldots,p_{k}\} has a component Δ\Delta which contains more than half of the vertices of Γ\Gamma, and Δ\Delta is 22-connected. We call Δ\Delta the large component at this game position.

We show that Duplicator can play to maintain the following invariant:

(*)

there is an edge {u,v}\{u,v\} of Γ\Gamma such that uUu\in U and vVv\in V are both in the large component and β\beta is a coherent permutation that is good bar uvuv.

In particular, this guarantees that β\beta is a partial isomorphism on the pebbled positions. Indeed, β\beta is a partial isomorphism on the graph Γ^\hat{\Gamma} excluding IuOuIvOvI_{u}\cup O_{u}\cup I_{v}\cup O_{v}, and none of these vertices is pebbled. Thus, if Duplicator can maintian the invariant (*) it is a winning strategy.

It is clear that the initial bijection α\alpha satisfies (*) as no vertices are pebbled.

At each subsequent move, Spoiler places a pebble on a vertex xx. Duplicator chooses any edge {u,v}\{u^{\prime},v^{\prime}\} in the large component (this could be {u,v}\{u,v\} if they are still in the large component). We then distinguish two cases.

  1. 1.

    If x{v0,v1}x\not\in\{v_{0},v_{1}\} then Duplicator’s response is to compose β\beta with (v0v1)(v0v1)(v_{0}v_{1})(v_{0}^{\prime}v_{1}^{\prime}). This is a valid move as none of the four vertices is pebbled and the pemutation (v0v1)(v0v1)(v_{0}v_{1})(v_{0}^{\prime}v_{1}^{\prime}) is in AltX×AltY\text{{Alt}}_{X}\times\text{{Alt}}_{Y} since all four of v0,v1,v0,v1v_{0},v_{1},v_{0}^{\prime},v_{1}^{\prime} are in YY. Moreover, the fact that β\beta is good bar uvuv implies that composing it with (v0v1)(v_{0}v_{1}) yields an automorphism of Γ^\hat{\Gamma}. Composing this automorphism with (v0v1)(v_{0}^{\prime}v_{1}^{\prime}) gives us a permutation that is good bar uvu^{\prime}v^{\prime}. Thus, the invariant (*) is maintained.

  2. 2.

    If x{v0,v1}x\in\{v_{0},v_{1}\} Duplicator must compose β\beta with a permutation that fixes xx, so in fact fixes both v0v_{0} and v1v_{1}. By the fact that the large component Δ\Delta before the pebble is placed on xx is 33-connected, we know that there is a path in Δ\Delta from uu to vv that does not use the edge e={u,v}e=\{u,v\}. Combining this with ee we obtain a simple cycle CC. Then, βC\beta_{C} is an automorphism of Γ^\hat{\Gamma} which, in particular, swaps v0v_{0} and v1v_{1}. Duplicator’s move is to compose β\beta with the permutation (v0,v1)βC(v0,v1)(v_{0},v_{1})\beta_{C}(v_{0}^{\prime},v_{1}^{\prime}). Call the resulting permutation β\beta^{\prime}. Observe first that this is a valid move, as βCAltX×AltY\beta_{C}\in\text{{Alt}}_{X}\times\text{{Alt}}_{Y} and we are composing it with two swaps of elements of YY. It remains to argue that β\beta^{\prime} is good bar uvu^{\prime}v^{\prime}. By assumption composing β\beta with (v0v1)(v_{0}v_{1}) yields an automorphism of Γ^\hat{\Gamma} and βC\beta_{C} is also an automorphism of Γ^\hat{\Gamma}. Thus, β(v0,v1)βC\beta(v_{0},v_{1})\beta_{C} is an automorphism of Γ^\hat{\Gamma} and β\beta^{\prime} is just the composition of this with (v0,v1)(v_{0}^{\prime},v_{1}^{\prime}).

5.4 Bringing it Together

We pull things together to prove Theorem 5.1

Proof of Theorem 5.1.

We have from Theorem 5.7 that for each nn\in\mathbb{N} there exists a 33-regular balanced bipartite graph Γn\Gamma_{n} with 2n2n vertices that is k(n)k(n)-well-connected for k(n)=Ω(n)k(n)=\Omega(n) and has an odd number of perfect matchings. Let Γ^n\hat{\Gamma}_{n} be the CFI-graph constructed from Γn\Gamma_{n} as described in Section 5.1, and MnM_{n} and NnN_{n} be two biadjacency matrices for Γ^n\hat{\Gamma}_{n} where NnN_{n} is obtained from MnM_{n} by interchanging exactly one pair of columns. From Lemma 5.11 we have that Duplicator has a winning strategy for the (Altn×Altn,k(n)3)(\text{{Alt}}_{n}\times\text{{Alt}}_{n},k(n)-3)-bijection game on MnM_{n} and NnN_{n}. From Lemma 5.6 and the fact that Γn\Gamma_{n} has an odd number of perfect matchings, it follows that det(Mn)0\det(M_{n})\neq 0 and so, since det(Mn)=det(Nn)\det(M_{n})=-\det(N_{n}), we have det(Mn)det(Nn)\det(M_{n})\neq\det(N_{n}). The result now follows from Theorem 4.13. ∎

Theorem 5.1 is stated and proved specifically for fields of characteristic zero. We could prove the result for fields of characteristic pp, when pp is an odd prime, provided that the matrices MnM_{n} we construct have non-zero determinant modulo pp. As noted in the proof of Lemma 5.6, det(Mn)=24mFsgn(F)\det(M_{n})=2^{4m}\sum_{F}\text{sgn}(F), where the sum is over all 22-factors (or equivalently over all perfect matchings) of the bipartite graph Γn\Gamma_{n}. Of course, Fsgn(F)\sum_{F}\text{sgn}(F) is just the determinant of the bi-adjacency matrix BΓnB_{\Gamma_{n}} of Γn\Gamma_{n}. Hence, for odd prime pp, detMn0(modp)\det{M_{n}}\not\equiv 0\pmod{p} provided that det(BΓn)(modp)\det(B_{\Gamma_{n}})\not\equiv\pmod{p}. Now, the construction in Section 5.2 is aimed at ensuring that det(BΓn)0(mod2)\det(B_{\Gamma_{n}})\not\equiv 0\pmod{2}. It seems plausible that we could just as well ensure that det(BΓn)0(modp)\det(B_{\Gamma_{n}})\not\equiv 0\pmod{p} for some odd prime pp. This would immediately yield the analogue of Theorem 5.1 in characteristic pp. We leave this extension to future work.

6 Lower Bound for the Permanent

We previously established in [11] lower bounds on symmetric circuits for the permanent showing that there are no subexponential square-symmetric circuits computing the permanent of an n×nn\times n matrix in any field of characteristic zero, along with a similar result for matrix-symmetric circuits in any field of characteristic other than two.

The two bounds are consequences of the same construction: we give, for each kk, a pair of bipartite graphs XkX_{k} and X~k\tilde{X}_{k} on which Duplicator has a winning strategy in the kk-pebble bijection game and which have different numbers of perfect matchings. The graphs XkX_{k} and X~k\tilde{X}_{k} are on two sets AA and BB of n=O(k)n=O(k) vertices each and the difference between the number of perfect matchings in XkX_{k} and X~k\tilde{X}_{k} is a power of 22. The kk-pebble bijection game for which a Duplicator winning strategy is shown is essentially the (SymA×SymB,k)(\text{{Sym}}_{A}\times\text{{Sym}}_{B},k) game. This shows that the biadjacency matrices of XkX_{k} and X~k\tilde{X}_{k} cannot be distinguished by SymA×SymB\text{{Sym}}_{A}\times\text{{Sym}}_{B}-symmetric circuits of subexponential size and also that the adjacency matrices of XkX_{k} and X~k\tilde{X}_{k} cannot be distinguished by SymAB\text{{Sym}}_{A\cup B}-symmetric circuits of subexponential size. Since XkX_{k} and X~k\tilde{X}_{k} have different numbers of perfect matchings, their biadjacency matrices BXkB_{X_{k}} and BX~kB_{\tilde{X}_{k}} have distinct permanents. Since the number of perfect matchings differ by a power of 22, they have distinct permanents modulo pp for any odd prime pp. Moreover, since the permanent of the adjacency matrix of a bipartite graph is the square of the permanent of its biadjacency matrix, we have that the adjacency matrices AXkA_{X_{k}} and AX~kA_{\tilde{X}_{k}} also have distinct permanents. Together, these give us the stated lower bounds for circuits computing the permanent in the second and fourth columns of Table 1. To establish the lower bound in the third column, it suffices to observe that Duplicator has a winning strategy on XkX_{k} and X~k\tilde{X}_{k} even in the restricted game (AltA×AltB,k)(\text{{Alt}}_{A}\times\text{{Alt}}_{B},k). To see this, we give a brief account of the construction, and it is instructive to contrast it with the graphs we use in Section 5.1.

The construction of the graph Γ^\hat{\Gamma} from Γ\Gamma given in Section 5.1 is essentially the original construction given by Cai et al. [6]. Their construction gives from Γ\Gamma a pair of non-isomorphic graphs which are not distinguishable by kk-dimensional Weisfeiler-Leman equivalence. We use only one graph Γ^\hat{\Gamma} from the pair, as our game is played on two different biadjacency matrices of the same graph and our main concern is that this matrix has non-zero determinant. Of course, the different biadjacency matrices have the same permanent, and for a lower bound for the latter, we do need to play the game on a pair of non-isomorphic graphs. So, we look again at the pairs of graphs given by the CFI construction. The other graph in the pair would be obtained from Γ^\hat{\Gamma} by “twisting” one of the edges. That is, for some edge e={u,v}e=\{u,v\} of Γ\Gamma we replace the two edges e0={u0,v0}e_{0}=\{u_{0},v_{0}\} and e1={u1,v1}e_{1}=\{u_{1},v_{1}\} by the edges {u0,v1}\{u_{0},v_{1}\} and {u1,v0}\{u_{1},v_{0}\}. It can be verified that the number of perfect matchings in the two graphs is the same. Thus, the permanents of the biadjacency matrices of the two graphs are the same and they cannot be used directly to establish the lower bounds we want. In the construction we presented in [11] we adapted the CFI construction in two ways. The graph X(Γ)X(\Gamma) is obtained from Γ^\hat{\Gamma} by first, for each edge e={u,v}e=\{u,v\} of Γ\Gamma, contracting the two edges e0={u0,v0}e_{0}=\{u_{0},v_{0}\} and e1={u1,v1}e_{1}=\{u_{1},v_{1}\} in Γ^\hat{\Gamma} and secondly, for each vertex vv of Γ\Gamma, adding a new vertex vbv_{b} to Γ^\hat{\Gamma} which is adjacent to all four vertices in IvI_{v}. Overall, this is equivalent to replacing each vertex vv of Γ\Gamma with incident edges f,g,hf,g,h with the gadget in Figure 4 where the dashed lines indicate edges whose endpoints are in other gadgets.

Refer to caption
f0f_{0}
f1f_{1}
g1g_{1}
g0g_{0}
h1h_{1}
h0h_{0}
vv_{\emptyset}
v{g,h}v_{\{g,h\}}
Refer to caption
v{f,h}v_{\{f,h\}}
Refer to caption
v{f,g}v_{\{f,g\}}
Refer to caption
vbv_{b}
Refer to caption
Figure 4: A gadget in X(Γ)X(\Gamma) corresponding to vertex vv with incident edges f,g,hf,g,h

The resulting graph X(Γ)X(\Gamma) is a 44-regular bipartite graph and X~(Γ)\tilde{X}(\Gamma) is obtained from it by taking one vertex vv of Γ\Gamma and in the corresponding gadget, for one edge ee incident on vv, interchanging the connections of e0e_{0} and e1e_{1}. The fact that X(Γ)X(\Gamma) and X~(Γ)\tilde{X}(\Gamma) have biadjacency matrices with different permanents is proved in [11]. It is interesting to note, however, that these matrices have determinant zero, so do not yield a lower bound on the determinant.

Finally, we briefly describe a winning strategy for Duplicator in the (AltA×AltB,k)(\text{{Alt}}_{A}\times\text{{Alt}}_{B},k)-bijection game played on X(Γ)X(\Gamma) and X~(Γ)\tilde{X}(\Gamma). The winning strategy in the ordinary kk-pebble bijection game, is described in [9] and is based on the fact that the graph Γ\Gamma has tree-width greater than kk and so there is a winning strategy for robber in the kk-cops-and-robbers game played on Γ\Gamma. This is lifted to a winning strategy for Duplicator in the kk-bijection game which sees Duplicator maintain a bijection β\beta that is an isomorphism except at one gadget corresponding to a vertex vv of Γ\Gamma. The position vv is given as a robber winning position in a kk-cops-and-robbers game played on the graph Γ\Gamma. At each move, Duplicator takes a path from vv to vv^{\prime} in Γ\Gamma describing a robber move and changes β\beta to a bijection β\beta^{\prime} by composing it with automorphisms for the gadgets corresponding to the vertices along the path. This results in β\beta^{\prime} being an isomorphism except at the gadget corresponding to vv^{\prime}. It can be easily checked that β\beta^{\prime} is obtained from β\beta by composing with a permutation in AltA×AltB\text{{Alt}}_{A}\times\text{{Alt}}_{B} provided that the path from vv to vv^{\prime} is of even length. It is easy to ensure that the graph Γ\Gamma of tree-width greater than kk used is bipartite and robber has a winning strategy in the kk-cops and robber game in which robber always ends up on one side of the bipartition. This gives us the desired result.

Theorem 6.1.

Let 𝔽\mathbb{F} be a field of characteristic other than 22. There is no family of Altn×Altn\text{{Alt}}_{n}\times\text{{Alt}}_{n}-symmetric circuits (Cn)n(C_{n})_{n\in\mathbb{N}} over 𝔽\mathbb{F} of size 2o(n)2^{o(n)} computing the permanent over 𝔽\mathbb{F}.

7 Concluding Discussion

The study of the complexity of symmetric circuits began in the context of logic. Specifications of decision problems on graphs (or similar structures) formulated in formal logic translate naturally into algorithms that respect the symmetries of the graphs. This yields a restricted model of computation based on symmetric circuits for which we are able to prove concrete lower bounds, in a fashion similar to the restriction to monotone circuits. Methods developed in the realm of logic for proving inexpressibility results can be reinterpreted as circuit lower bound results.

One step in this direction was the connection established in [1] between polynomial-size Boolean threshold circuits on the one hand and fixed-point logic with counting on the other. This shows that the power of symmetric Boolean threshold circuits to decide graph properties is delimited by the counting width of those properties. In particular, this shows that a number of NP\mathrm{NP}-complete graph problems including 33-colourability and Hamiltonicity cannot be decided by polynomial-size symmetric Boolean threshold circuits. This is particularly interesting as the power of such symmetric circuits has been shown to encompass many strong algorithmic methods based on linear and semidefinite programming (see, for instance, [2]). This methodology was extended to graph parameters beyond decision problems, and to arithmetic circuits rather than Boolean circuits in [11]. Together these extensions established that no subexponential size square symmetric (i.e. unchanged by simultaneous row and column permutations) arithmetic circuits could compute the permanent.

The permanent of a {0,1}\{0,1\} matrix MM has a natural interpretation as a graph invariant and so lends itself easily to methods for proving lower bounds on graph parameters. The situation with the determinant of MM is more subtle. If MM is a symmetric n×nn\times n matrix, then we can see at as the adjacency matrix of a graph Γ\Gamma on nn vertices and the determinant is a graph invariant, that is to say it only depends on the isomorphism class of Γ\Gamma. Moreover, it is a graph invariant that can be computed efficiently by symmetric circuits (at least in characteristic zero) as was shown for Boolean circuits in [18] and for arithmetic circuits in [11]. When MM is not a symmetric matrix, we could think of it as the adjacency matrix of a directed graph. In this case, the symmetries that a circuit must preserve are still the permutations of nn, so simultaneous permutations of the rows and columns of MM and the upper bounds obtained still apply. But, we can also think of MM as a biadjacency matrix of a bipartite graph Γ\Gamma, and now the determinant of MM is not an invariant of Γ\Gamma. We have a richer set of symmetries, and methods for proving bounds on the complexity of graph parameters do not directly apply.

What we have sought to do in the present paper is to develop the methods for proving lower bounds on the counting width of graph parameters to a general methodology for proving circuit lower bounds for polynomials or more generally functions invariant under certain permutations of their input variables. To do this, we prove a support theorem for circuits which is for a more general collection of symmetry groups than proved in prior literature; we adapt the Spoiler-Duplicator bijection game to work for more general invariance groups and more general structured inputs than those arising as symmetries of graph matrices; and we show a direct relationship between these games and orbit size of circuits that bypasses connections with width measures on graphs. This methodology is then applied to arithmetic circuits computing the determinant and we are able to prove an exponential lower bound for circuits symmetric under the full permutation group DnSymn×SymnD_{n}\leq\text{{Sym}}_{n}\times\text{{Sym}}_{n} that fixes the determinant of MM. Indeed, we do this for the smaller group Altn×Altn\text{{Alt}}_{n}\times\text{{Alt}}_{n}. The application requires considerable work in constructing the example matrices and applying the bijection games.

We see one main contribution to be establishing the general methodology for proving circuit lower bounds under various notions of symmetry. There are many ways in which this could be pushed further. First, our proof of the support theorem requires the presence of large alternating groups in the symmetry group under consideration. Perhaps more sophisticated notions of support could be developed which would allow us to consider smaller groups. Secondly, while we state our main result for the group Altn×Altn\text{{Alt}}_{n}\times\text{{Alt}}_{n}, the bijection game itself uses rather fewer symmetries. It would be interesting to establish tighter bounds on the symmetry group for which we get exponential lower bounds. Indeed, the results can be seen as giving a trade-off between circuit size and symmetries and this suggests an interesting terrain in which to explore the symmetry requirements of the circuit as a resource.

Acknowledgements

We are grateful to Benedikt Pago, whose comments resulted in a much improved version of the definitions and results in Section 4.1. We are also grateful to Albert Atserias for useful discussions on the construction in Section 5.2.

References

  • [1] M. Anderson and A. Dawar. On symmetric circuits and fixed-point logics. Theory Comput. Syst., 60(3):521–551, 2017.
  • [2] A. Atserias, A. Dawar, and J. Ochremiak. On the power of symmetric linear programs. J.ACM, 2021. to appear. arxiv:1901.07825.
  • [3] Béla Bollobás. The distribution of the maximum degree of a random graph. Discret. Math., 32:201–203, 1980.
  • [4] G. Brito, I. Dumitriu, and K. D. Harris. Spectral gap in random bipartite biregular graphs and applications. Combinatorics, Probability and Computing, 2021. to appear. arXiv:1804.07808.
  • [5] P. Bürgisser, M. Clausen, and M. A. Shokrollahi. Algebraic Complexity Theory. Springer, Berlin, Heidelberg, 1 edition, 1997.
  • [6] J-Y. Cai, M. Fürer, and N. Immerman. An optimal lower bound on the number of variables for graph identification. Combinatorica, 12(4):389–410, 1992.
  • [7] A. Dawar. On symmetric and choiceless computation. In Mohammad Taghi Hajiaghayi and Mohammad Reza Mousavi, editors, Topics in Theoretical Computer Science, pages 23–29, Cham, 2016. Springer International Publishing.
  • [8] A. Dawar. Symmetric computation (invited talk). In 28th EACSL Annual Conference on Computer Science Logic, CSL 2020, 2020.
  • [9] A. Dawar and D. Richerby. The power of counting logics on restricted classes of finite structures. In CSL 2007:Computer Science Logic, volume 4646 of LNCS, pages 84–98. Springer, 2007.
  • [10] A. Dawar and G. Wilsenach. Symmetric circuits for rank logic. In 27th EACSL Annual Conference on Computer Science Logic, CSL 2018, pages 20:1–20:16, 2018. Full version at arXiv:1804.02939.
  • [11] A. Dawar and G. Wilsenach. Symmetric Arithmetic Circuits. In 47th International Colloquium on Automata, Languages, and Programming (ICALP 2020), volume 168 of Leibniz International Proceedings in Informatics (LIPIcs), pages 36:1–36:18, 2020. Full version at arXiv:2002.06451.
  • [12] A. Dawar and G. Wilsenach. Lower bounds for symmetric circuits for the determinant. In 13th Innovations in Theoretical Computer Science Conference, ITCS, volume 215 of LIPIcs, pages 52:1–52:22. Schloss Dagstuhl - Leibniz-Zentrum für Informatik, 2022.
  • [13] J.D. Dixon and B. Mortimer. Permutation Groups. Graduate Texts in Mathematics. Springer New York, 1996.
  • [14] D. Grigoriev and M. Karpinski. An exponential lower bound for depth 3 arithmetic circuits. In Proceedings of the Thirtieth Annual ACM Symposium on the Theory of Computing, pages 577–582, 1998.
  • [15] M. Grohe. Descriptive Complexity, Canonisation, and Definable Graph Structure Theory. Lecture Notes in Logic. Cambridge University Press, 2017.
  • [16] F. Harary. Determinants, permanents and bipartite graphs. Mathematics Magazine, 42:146–148, 1969.
  • [17] L. Hella. Logical hierarchies in PTIME. Information and Computation, 129(1):1 – 19, 1996.
  • [18] B. Holm. Descriptive Complexity of Linear Algebra. PhD thesis, University of Cambridge, 2010.
  • [19] S. Hoory, N. Linial, and A. Wigderson. Expander graphs and their applications. Bulletin of the American Mathematical Society, 43(4):439–561, 2006.
  • [20] M. Jerrum and M. Snir. Some exact complexity results for straight-line computations over semirings. J. ACM, 29:874–897, 1982.
  • [21] N. Kayal and R. Saptharishi. A selection of lower bounds for arithmetic circuits. In M. Agrawal and V. Arvind, editors, Perspectives in Computational Complexity. Birkhäuser Basel, 2014.
  • [22] J.M. Landsberg and N. Ressayre. Permanent v. determinant: An exponential lower bound assuming symmetry. In Proc. ACM Conference on Innovations in Theoretical Computer Science, pages 29–35. ACM, 2016.
  • [23] M. Molloy, H. D. Robalewska, R. W. Robinson, and N. C. Wormald. 1-factorizations of random regular graphs. Random Struct. Algorithms, 10:305–321, 1997.
  • [24] H.J. Ryser. Combinatorial Mathematics, volume 14. Mathematical Association of America, 1 edition, 1963.
  • [25] P.D. Seymour and R. Thomas. Graph searching and a min-max theorem for tree-width. Journal of Combinatorial Theory, Series B, 58(1):22–33, 1993.
  • [26] A. Shpilka and A. Yehudayoff. Arithmetic circuits: A survey of recent results and open questions. Foundations and Trends in Theoretical Computer Science, 5(3-4):207–388, 2010.
  • [27] H. Vollmer. Introduction to Circuit Complexity - A Uniform Approach. Texts in Theoretical Computer Science. An EATCS Series. Springer, 1999.