Lower Bounds for Symmetric Circuits for the Determinant††thanks: Research funded by EPSRC grant EP/S03238X/1. A preliminary version of this paper appeared as [12]
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 and . Sometimes known as Valiant’s conjecture, this is also described as the algebraic analogue of the vs. 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 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 and a set of variables , let be a circuit computing a polynomial in . For a group acting on the set , we say that is -symmetric if the action of any element on the inputs of can be extended to an automorphism of . Of course, this makes sense only when the polynomial itself is invariant under the action of .
For example, both the permanent and the determinant are polynomials in a matrix of variables . Let be the group acting on by the action whereby takes to . 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 -symmetric circuits computing the determinant, but any family of -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 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 on whereby takes to . 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 -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 for the group of permutations of which fix the determinant. This can be seen to be where is the subgroup of of index consisting of pairs of permutations with and represents the transposition of rows and columns. We prove in the present paper that any family of -symmetric circuits computing the determinant must have exponential size. Indeed, our lower bound is proved even for the subgroup of given by .
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 -matrix. Here, the orbit size of a circuit is the maximal size of an orbit of a gate of under the action of the automorphism group of . 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 vertices which are not distinguished by -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 -invariant parameter of a -matrix can be seen as a graph parameter. That is, a graph parameter that does not distinguish between isomorphic graphs is necessarily -invariant on the adjacency matrices of graphs. Similarly, the 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 -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 action. The results in the column for are new to this paper. In the last two columns and represent the full invariance groups (as subgroups of ) of the determinant and permanent respectively. The lower bounds stated in those columns follow from the ones obtained for their subgroups and respectively.
A more detailed discussion of some of the main innovations follows.
Det |
|
|
N/A |
|
N/A | |||||||||||||
Perm |
|
|
|
|
|
|
-
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 and are used to establish the indistinguishability of graphs by means of -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 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.
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.
The original -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 -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 . 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 -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.
A fourth key ingredient is the construction of matrices with distinct determinants on which we can show Duplicator winning strategies in the -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.
An important element of the construction of the matrices in (4) are bipartite -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 ) 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 to .
2 Background
In this section we discuss relevant background and introduce notation.
We write for the positive integers and for the non-negative integers. For , denotes the set . For a set we write to denote the powerset of . We write to denote the identity function on some specified set. For and we write to denote the image of . Let denote the set of bijections from to .
2.1 Groups
Let and denote the symmetric group and alternating group on the set . We write and to abbreviate and , respectively. Let denote the trivial group. For groups and we write to denote that is a subgroup of . The sign of a permutation is defined so that if is even and otherwise .
Let be a group acting on a set . We denote this as a left action, i.e. for , . The action extends in a natural way to powers of . So, for , . It also extends to the powerset of and functions on as follows. The action of on is defined for and by . For any set, the action of on is defined for and by for all . We refer to all of these as the natural action of on the relevant set.
A set with the action of group on it is called a -set. We do not distinguish notationally between a -set and the underlying set of elements. Thus, we can say that if is a -set, the collection of functions is also a -set with the natural action.
Let and for each let be a group acting on . The action of the direct product on is defined for and by . If instead then the action of on is defined for and such that if then . Again, we refer to either of these as the natural action of on .
Let be a -set. Let . Let denote the pointwise stabilizer of . Let denote the setwise stabilizer of . If is a singleton we omit set braces and write . Note that is the pointwise stabilizer of in the -set . For let denote the orbit of . In all cases, we omit the subscript where it is clear from context.
2.2 Matrices
Let and be finite non-empty sets. We identify matrices with rows indexed by and columns by and entries from some set with functions of the form . So for , , . We also denote by , or just when the index sets are clear from context.
Let be a commutative ring and be a matrix with . The permanent of over is . Suppose . The determinant of over is . The trace of over is . We omit reference to the ring when it is obvious from context. When is a field, we write to denote the rank of the matrix .
We always use to denote a field and to denote the characteristic of . For any prime power we write for the finite field of order . We are often interested in polynomials and circuits defined over a set of variables with a natural matrix structure, i.e. . We identify with this matrix. We also identify any function of the form with the matrix with entries in defined by replacing each with .
2.3 Graphs
Given a graph , the adjacency matrix of is the -matrix with if, and only if, . If is bipartite, with bipartition , then the biadjacency matrix of is the -matrix with if, and only if, . For a set we write to denote the subgraph induced by .
A -factor of a graph is a spanning -regular subgraph. A perfect matching is a -factor. It is well known that for a bipartite graph , over any field of characteristic zero counts the number of perfect matchings in [16] and for prime , for a field of characteristic counts the number of perfect matchings in modulo . For bipartite , is a block anti-diagonal matrix with two blocks corresponding to and , and so .
Let be a graph. For let . We say is an -expander for if for every of size at most we have . For an introduction to expander graphs see [19]. A set of vertices in a graph is a balanced separator if no connected component of contains more than half the vertices of . It is easy to see that for any there is a constant such that if is an -expander then has no balanced separator of size less than .
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 , with one player controlling the single robber and another a team of cops. The cops and robbers both sit on nodes in the graph. At the start of a round the cops are sitting on nodes and the round begins with the cop player announcing he intends to move cops from to occupy the nodes . The robber player must respond with a path from the robbers current location that does not intersect . 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 cops if, and only if, the tree-width of is at most .
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 with variables and values is a directed acyclic graph with a labelling where each vertex of in-degree is labelled by an element of and each vertex of in-degree greater than is labelled by an element of .
Let be circuit that is the labelling of a directed graph , where with values . We call the elements of gates, and the elements of wires. We call the gates with in-degree input gates and gates with out-degree output gates. We call those input gates labelled by elements of constant gates. We call those gates that are not input gates internal gates. For we say that is a child of if . We write to denote the set of children of . We write to denote the sub-circuit of rooted at , i.e the sub-circuit induced by those gates with a directed path to . Unless otherwise stated we always assume a circuit has exactly one output gate.
If is a field , and is the set , we have an arithmetic circuit over . If , and is a collection of Boolean functions, we have a Boolean circuit over the basis . We define two bases here. The first is the standard basis containing the functions , , and . The second is the threshold basis which is the union of and , where for each , is defined for a string so that if, and only if, the number of s in at least . We call a circuit defined over this basis a threshold circuit. Another useful Boolean function is , which is defined by . We do not explicitly include it in the basis as it is easily defined in .
In general, we require that a basis contain only functions that are invariant under all permutations of their inputs. That is, if is such that for some then for all and we have . We call such a function fully symmetric. The arithmetic functions and and all of the Boolean functions in and are fully symmetric. Let be a circuit defined over such a basis with variables and values . We evaluate for an assignment by evaluating each gate labelled by some to and each gate labelled by some to , and then recursively evaluating each gate according to its corresponding basis element. That is gets the value obtained by applying the function labelling to the set of values of . We write to denote the value of the gate and to denote the value of the output gate. We say that computes the function .
It is conventional to consider an arithmetic circuit over with variables to be computing a polynomial in , rather than a function . 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 . Each gate then evaluates to some polynomial in . The polynomial computed by is the value of the output gate.
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 where we think of as a set of variables and a domain of values that the variables can take. The set may have some further structure to reflect the intended structured input, but no more. In particular, we do not assume that is linearly ordered. For example, if and then the elements of may be naturally interpreted as directed graphs over the vertex set . Or, with the same , if we let , we can think of the elements of as matrices over with rows and columns indexed by .
The symmetries of interest for a given class of structures correspond to group actions on , which lift to actions on . In this section we introduce the definitions of -symmetric functions, i.e. functions which are invariant under the action of on its input, and -symmetric circuits, which are circuits computing -symmetric functions but where the structure of the circuit itself and not just the output is invariant under the action of . 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 acting on the domain of a function we say is -symmetric if for every , . We omit mention of the group when it is obvious from context.
We are interested in functions of type with some group acting on , which then induces an action on . We think of elements of as “generalized structures”, and define notions of homomorphism and isomorphism below. We first consider some examples. Note that whenever is a subgroup of , then any -symmetric function is also -symmetric. In particular every function is -symmetric.
Example 3.2.
-
1.
For a vertex set , a function defines a property of (directed) graphs if it is -symmetric, with the action of being defined simultaneously on both coordinates. We get properties of simple undirected graphs by considering -symmetric functions where is the collection of -element subsets of . Examples of graph properties include connectedness, Hamiltonicity and the existence of a perfect matching. A graph parameter is a function which is -symmetric. Examples include the number of connected components, the number of Hamiltonian cycles and the number of perfect matchings.
-
2.
The elementary symmetric polynomial of degree in the set of variables is the polynomial:
For any field , defines a function which is -symmetric.
-
3.
If is a matrix of variables, then the trace , determinant and permanent are polynomials that define, for any field , functions of type which are -symmetric where the group action is defined simultaneously on both coordinates.
-
4.
The permanent is invariant under separate row/column permutations. In other words, for , and matrix , we have . So the permanent for -matrices is a -symmetric function.
-
5.
The trace and determinant are not -symmetric under the above action, but the symmetry group of the determinant is richer than just . Define to be the group . This is a subgroup of of index . The determinant is -symmetric and so, in particular -symmetric, since .
Let and be groups. A homomorphism from the -set to -set is a pair of functions where is a function and is a group homomorphism such that for each and , . We say is an isomorphism if both and are bijective. We abuse terminology and refer to a bijection as an isomorphism if there exists such that is an isomorphism.
Let with a -set and let with an -set. A homomorphism from to is a homomorphism such that . It is an isomorphism from to if both and are bijective. We omit mention of the groups and refer just to a homomorphisms or isomorphisms from to when the groups are clear from context (see Section 4.1.
Let be a group acting on and . We are usually interested in such functions up to isomorphism. Let and be a group acting on such that and are isomorphic. Then we abuse notation and sometimes write to denote for some isomorphism .
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 , where is a -set, and are symmetric in the sense that the computation itself, not just the value of the output, remains unchanged under the action of . We first need to formalize what it means for a permutation in 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 be a circuit over the basis with variables and values . For , we say that a bijection is an automorphism extending if for every gate in we have that
-
•
if is a constant gate then ,
-
•
if is a non-constant input gate then ,
-
•
if is a wire, then so is
-
•
if is labelled by , then so is .
We say that a circuit with variables is rigid if for every permutation there is at most one automorphism of extending . 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 be a group acting on a set and be a circuit with variables . We say is a -symmetric circuit if for every the action of on extends to an automorphism of .
The following can be shown via a straightforward induction.
Proposition 3.5.
Let be a -symmetric circuit. Then defines a -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 and 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 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 and so pebbles can be moved from one element to another during the course of the game. At each round, Spoiler chooses a pebble () from and the matching pebble from . Duplicator provides a bijection between the two structures. Spoiler then chooses an element of on which to place and is placed on . 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 . 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 -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 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 in a -symmetric circuit for is a subset of which determines both the evaluation of 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 pebbles on a pair of indexed functions then these functions cannot be distinguished by any pair of -symmetric circuits with supports of size at most . 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 be a set, , and be a -set. We call the triple an indexed set. We call a triple , where is a function on , an indexed function.
All of the structures discussed so far may be identified with indexed functions. For example, we can identify a (directed) graph with the indexed function . Similarly, a bipartitioned graph can be identified with the indexed function .
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 . Suppose and are index sets and is a bijection. For a group , let be the group of permutations . Let and be a pair of indexed functions and be a bijection which lifts the bijection in the sense that for all . We say that the indexed functions and are -isomorophic, if there is a permutation such that for all , .
In particular, for a pair of directed graphs and , if is any bijection between and , and is its natural lift to a bijection between and (i.e. , then the pair of indexed functions and are -isomorphic if, and only if, and are isomorphic in the usual sense. Similar statements hold for bipartitioned graphs and other relational structures. As a further example, we can identify matrices over a field with structured functions of the form , for some group . Unlike the case of structures considered above, there is not a canonical choice of , and it depends on what invariants of the matrix we wish to take into account. For example, if then isomorphism corresponds to equivalence under simultaneous row and column permutations and if 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 and and its lift to the indexed sets and . To simplify notation, we often simply identify the sets and , and the sets and , so that the base bijection can be taken to be the identity. This is akin to fixing the universe of any vertex graph to be the set , so that isomorphisms can be identified with permutations of .
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 be an indexed set. For we define the lift of to be the set .
Thus, is those elements of that are fixed by any permutation that fixes pointwise. The following is now an easy observation.
Lemma 4.3.
Let be an indexed set. If is the lift of then .
Proof.
Let , , and . Then , so . Hence, by definition of , we have for any . Hence and therefore . Since was arbitrary, and therefore . ∎
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 and be two indexed functions, and let and . We say that induces a partial isomorphism on if, for each , we have .
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 and be indexed functions. The -pebble -bijection game on is defined as follows. The game is played between two players called the Spoiler and Duplicator using two sequences of pebbles and . In each round of the game:
-
1.
Spoiler picks a pair of pebbles and ,
-
2.
Duplicator picks a permutation such that for each , , and
-
3.
Spoiler chooses some and places on and on .
Spoiler has won the game at the end of the round if the permutation does not induce a partial isomorphism on the set . We say that Duplicator has a winning strategy for the -pebble -bijection game if it has a strategy to play the game forever without Spoiler winning.
In the sequel, we abbreviate “-pebble -bijection game” to -bijection game.
We recover the ordinary bijection game on graphs by taking , and . It is clear in this case that if and are isomorphic graphs, then Duplicator has a winning strategy by choosing a fixed isomorphism between them and playing it at every move. However, if is a more restricted group that does not contain an isomorphism between and , a Duplicator winning strategy is not guaranteed even when and 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 -symmetric circuit then the function computed by is invariant under permutations in on the input. This is not the case for individual gates in other than the output gate: applying a permutation to the inputs of the circuit might yield a different function at . But, by the symmetry condition, there is then another gate which computes this other function. If the circuit is small, the orbit of is small and thus the stabilizer group of is large. That is, for many we have . What the support theorem tells us is that in this case, there is a small subset (which we call a support of ) of the permutation domain of such that the function computed at only depends on . This is in the sense that permutations that only move elements of the permutation domain outside of do not change the function computed at . This support theorem can be proved as long as the group 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 be a -set and a subgroup of . We say that is a support of if .
We extend this notion to indexed sets, where the support is now a subset of the indices. A key example is the action of on the set or on the collection of matrices .
Definition 4.7.
Let be an indexed set. We say that a set is a support of if it is a support of .
Note that is a support of just in case is in the lift of , in the sense of Definition 4.2.
Let be an indexed set. We write to denote the maximum size of the orbit of over all . For we write to denote the minimum size of a support of and to denote the maximum of over all .
We now specialise our discussion to circuits. Let be a rigid -symmetric circuit with gate set , where for some natural number . Then is an indexed set. In this way we can speak of the supports and orbits of a gate. We abuse notation slightly and write and to denote and , 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 -pebble winning strategy cannot be distinguished by symmetric circuits with supports limited to size . The statement of the theorem and argument used generalize that in [7].
Theorem 4.8.
Let be an indexed set and let be a -symmetric circuit with values from and variables , such that . If Duplicator has a winning strategy for the -bijection game played on and for functions and then .
Proof.
We prove this result by contraposition, i.e. we assume and show that Spoiler has a winning strategy for the -game on .
The strategy we define for Spoiler can be understood informally as follows. We start at the output gate and note that for any bijection , . The goal of Spoiler now over the next (at most) rounds is to pebble a support of some child of such that for Duplicator’s most recent chosen bijection we have that . Since the bijection chosen by Duplicator can be interpreted as a permutation respecting the pebble configuration, and the fact that an entire support of is pebbled and so any such permutation fixes , it follows that for any choice of . We now iterate this process, aiming next to pebble a support of an appropriate child of over the next (at most) rounds. We terminate when we reach an input gate, which witnesses that Spoiler has won the game.
So, fix for each gate of a support of size at most . We write to denote this support.
Claim 4.9.
Suppose the pebbled position is and . Let be a gate such that and for some that maps to we have that . There exists a strategy for Spoiler such that after at most rounds where the pebbled positions are now and there is such that and for all that maps to we have that .
Proof.
We describe an invariant that Spoiler is able to maintain inductively. Suppose that after rounds and have been pebbled. Let be the pointwise stabiliser of in . Then . Spoiler ensures there is a gate with and a such that
for all that maps to . This suffices to prove the claim since and therefore for some , we will have and in this case, is a singleton.
We first consider the case . We have for any that maps to that . Since the basis element labelling is a fully symmetric function, there must be some such that is different from . Moreover, since , partitions into orbits. It follows that for some
Now, take any that maps to . Let . Then , and so . Moreover, for any , , and so
Thus, taking suffices.
Inductively, suppose Spoiler has maintained the invariant after rounds. At the start of round Spoiler picks up a pair of pebbles where is not placed on or an element of . Suppose Duplicator plays a bijection . Then, by the induction hypothesis there is an and an such that
and . Let enumerate . Then is partitioned into sets , where . Thus there is a value and for which
Spoiler then places on and on . Let be any element of . The required invariant is satisfied by construction. ∎
It follows from Claim 4.9 that in at most rounds Spoiler can force the pebbles to a position such that for some input gate labelled by some variable we have: that for any that Duplicator may play we have that . Thus , and so Spoiler wins.
∎
4.4 Bounds on Supports
The main result of this subsection establishes that, for suitable choice of groups , if a family of -symmetric circuits has subexponential size orbits than it has sublinear size supports. We prove this specifically for the case when is an alternating group. The argument extends easily to cases where 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 is a gate with orbit size (relative to ) bounded by . Then by the orbit-stabilizer theorem, this is equivalent to . One way for to have small index in this way is for it to always contain a large copy of the alternating group. That is contains the alternating group restricted to elements, or equivalently has a support of size at most . Theorem 4.10 asserts that indeed this is the only way.
Theorem 4.10 ([13], Theorem 5.2B).
Let be a set such that , and let be an integer with . Suppose that has index then there exists with such that .
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 be a family of rigid -symmetric circuits. If then .
Proof.
Let be the least value such that . By the assumption that , we have that is . Indeed, otherwise there is a constant with , such that for infinitely many . And since for all it follows that . Since is the least value such that , it follows that for infinitely many , contradicting the assumption that .
From it follows that for all large enough , . Then, for any gate of , by the orbit-stabilizer theorem, we have and so by Theorem 4.10, has a support of size less than . ∎
The following corollary establishes the analogue of Theorem 4.11 needed to prove our lower bounds for -symmetric circuits.
Corollary 4.12.
Let be such that for each , is defined over the matrix of variables and is a rigid -symmetric circuit. If then .
Proof.
Fix and let and be the restriction of the action of to the two coordinates. We can think of them as the actions of on the rows and columns respectively. For any gate in , let and be supports of under the action of and respectively. Then it is easily seen that is a support of .
Suppose for all gates in we have . Since the orbit of under the action of includes its orbits under the action of its subgroups and , we have that each of and has size and therefore by Theorem 4.11 and can be chosen of size 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 -symmetric circuits taking as input -matrices, but note that it holds more generally.
Theorem 4.13.
Fix a set , and for each , let denote the group . Fix a -set and let be a family of functions , where is -symmetric. Suppose there are infinitely many for which there are pairs such that and Duplicator has a strategy to win the -bijection game played on and for . Then there is no family of -symmetric circuits that computes and has size .
Proof.
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 be a field of characteristic . There is no family of -symmetric circuits over of size computing the determinant over .
To prove Theorem 5.1 we need to construct for each , a pair of -matrices and with , and such that Duplicator has a winning strategy in the -bijection game played on and .
We construct the matrix as the biadjacency matrix of a bipartite graph . The matrix is obtained from by interchanging a pair of columns of . Hence, is also a biadjacency matrix of and by construction. Thus, as long as the two determinants are different.
In Section 5.1 we describe the construction of the graph which gives rise to the biadjacency matrix . This graph is obtained by a CFI construction from a base graph satisfying a number of conditions. We show the existence of graphs satisfying these conditions in Section 5.2. Then, in Section 5.3 we argue that Duplicator has a winning strategy in the -bijection game played on and . 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 -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 -dimensional Weisfeiler-Leman algorithm is the same as saying that Duplicator has a winning strategy in the -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 be a -regular bipartite graph with bipartition . We obtain the graph by replacing each vertex , with neigbours with the ten-vertex gadget depicted in Figure 1. This gadget is described as follows.
-
•
There is a set denoted of four inner vertices: a vertex for each set of even size.
-
•
There is a set denoted of six outer vertices: two for each .
-
•
There is an edge between and if and an edge between and if .
Corresponding to each edge there is a pair of edges that we denote and in : connects the vertex with and connects the vertex with
Note that the graph is bipartite. Indeed, if we let and we let then it is easily seen that all edges in are between and . Since is a -regular bipartite graph, it follows that and therefore . Writing for and for , we obtain a biadjacency matrix representing the graph by fixing a pair of bijections and . The action of the group divides the collection of all such matrices into two orbits. Letting and be representatives of the two orbits, we have . We next aim to show that , provided that has an odd number of perfect matchings.
Suppose we are given bijections and which together determine a biadjacency matrix representation of . We can also identify each perfect matching in with a bijection such that is a neighbour of for all . We write for the collection of all perfect matchings of . Then, the determinant of is given by:
From now on, we take and to be fixed and write and talk of the sign of a matching as short hand for .
To show that is non-zero, we analyze the structure of the set . Note that since and are fixed, this imposes a linear order on the sets and . It also induces a linear order on the sets and , for instance by saying that , for if the least element of is less than the least element of . We make use of this order without further elaboration.
Perfect Matchings
In any perfect matching of , all four vertices in for any must be matched to vertices in as they have no neighbours outside this set. Thus, exactly two of the vertices of are matched to vertices in other gadgets. These two could be two vertices in a pair, e.g. in Figure 1 or they could be from different pairs, e.g. . It is easily checked, by inspection of the gadget, that in either case, removing the two vertices of 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 the resulting graph is a two-regular bipartite graph on eight vertices. The graph resulting from removing and from the gadget is depicted in Figure 2 and all other cases of removing two vertices of from different pairs result in a graph isomorphic to this one. We now classify perfect matchings in according to which edges between gadgets are included in the matching.
Say that a matching is uniform if for each edge , at most one of the two edges and is included in . Thus, is non-uniform if for some both and are included in . Our first aim is to show that the non-uniform matchings contribute a net zero to the determinant of . Write to denote the set of uniform matchings of .
Lemma 5.2.
Proof.
For a non-uniform matching , let be an edge for which both and are included in and and a vertex such that is incident on . Then, the four vertices in are matched with the remaining four vertices in and there are exactly two ways this can be done. To see this, consider the gadget in Figure 1 and suppose that so and are matched outside the gadget. Then the only two possible matchings of the remaining eight vertices are and . Note that the second one is obtained from the first by composing with an odd permutation: namely the -cycle . We can then define an involution on the set of non-uniform matchings as follows: map to the unique non-uniform matching which differs from only in the set corresponding to the vertex which is minimal (in the order on ) among all vertices of incident on some edge for which are in . This is easily seen to be an involution by our previous observation. Moroever, it takes any matching to one of opposite sign.
We conclude that all non-uniform matchings come in pairs of opposite sign and therefore cancel out in the expression for the determinant of . ∎
Uniform Perfect Matchings.
Our next aim is to count uniform perfect matchings in and classify them by sign. Suppose then that is a perfect matching that includes for each at most one of the two edges and . Let be the set of those such that exactly one of and is in . Furthermore, let be the function given by if, and only if, is in .
Since for each , exactly two of the vertices of are matched to vertices in other gadgets we can see that includes exactly two edges incident on every vertex . In other words, is a -factor of and therefore has exactly edges.
For a -factor of and a function , write for the collection of all matchings of with and .
Lemma 5.3.
There are exactly perfect matchings in , for any -factor of and any function .
Proof.
Let be a vertex of with neighbours and consider the gadget in Figure 1. Exactly two of the edges incident on are included in and suppose that it is the edges and . Further, suppose . This means that the matching must pair the vertices and 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 and 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 must be matched with . This leaves a six-cycle which admits two matchings. Entirely analogously, if we can consider the gadget with the vertices and 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 .
Since the choice of the matching at each gadget is independent, and there are gadgets, one for each vertex in , we see that there are distinct matchings for the fixed choice of -factor and function . ∎
The proof of Lemma 5.3 shows that the two matchings and in 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 -cycle . We can immediately conclude that all matchings in have the same sign.
Lemma 5.4.
For any -factor of , any function and any matchings , .
We next show that we can go further and show that the sign of any matching in does not depend on the choice of .
Lemma 5.5.
For any -factor of , any functions and any matchings and , .
Proof.
It suffices to show the lemma for functions and which differ at exactly one edge as the result then follows by transitivity. Moreover, by Lemma 5.4, it suffices to choose any and and show that they have the same sign
So, assume and and and agree on all other edges. Let where has neighbours and has neighbours . Without loss of generality, assume and are in with . By assumption, includes the edge and we can assume further that it includes the edges , in the gadget corresponding to and the edges , in the gadget corresponding to . In other words, it contains alternating edges along the -cycle depicted in Figure 3.
Now, choose to be the symmetric difference between and the cycle in Figure 3. This is also a perfect matching and it is in as the only difference with as far as edges between gadgets is concerned is that contains rather than . Moreover, it is easily seen that is obtained from by composition with an even permutation: namely the -cycle: . ∎
With this, we are now ready to establish the main result of this section.
Lemma 5.6.
If has an odd number of perfect matchings, then .
Proof.
By Lemma 5.2, we have
For any -factor of write for and note that by Lemma 5.3 for all . Moreover, by Lemma 5.5 all matchings in have the same sign. Thus, we define to be for any . Hence, we have that
where the sum is over all -factors of .
Since is -regular, the number of -factors of is exactly the number of perfect matchings. Indeed, the complement of any -factor is a perfect matching and this gives a bijection between the collection of perfect matchings and -factors. Thus, since the number of perfect matchings of is odd, so is the number of -factors and we conclude that the sum cannot be zero, proving the result. ∎
5.2 Graphs with Odd Number of Perfect Matchings
We have seen that if has an odd number of perfect matchings, then the matrices and have different determinant. In order to play the bijection game on and we also need to be well connected. We now show that we can find suitable graphs that satisfy both of these conditions simultaneously.
For a positive integer , say that is -well-connected if any balanced separator of has size greater than . For our construction, we need -regular bipartite graphs on vertices which are -well-connected for 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 there is a bipartite graph satisfying the following conditions:
-
1.
;
-
2.
is -regular;
-
3.
is -well-connected for ; and
-
4.
has an odd number of perfect matchings.
We prove the existence of these graphs by a randomized construction. To be precise, a random -regular bipartite graph on 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 -regular bipartite graphs.
Fix and to be two disjoint sets of vertices, and we are interested in the uniform distribution on -regular bipartite graphs on the vertices and . 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 is disjoint if there is no with . Now consider a random graph obtained by the following process:
-
1.
choose, uniformly at random, three bijections ;
-
2.
if for some with , and are not disjoint discard this choice of bijections; otherwise
-
3.
let be the bipartite graph with parts and edges .
The random graph model obtained in this way is known to be contiguous to the uniform distribution on -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 -regular bipartite graph is an expander with probability tending to . 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 .
Lemma 5.8.
There is a constant such that with probability tending to , is an -expander.
An immediate consequence of this is that with high probability, is -well-connected for some constant . We now describe how we can obtain from a graph which also has an odd number of perfect matchings.
Let denote the biadjacency matrix of with rows indexed by and columns by . Then the permanent of over a field of characteristic is exactly the number of perfect matchings in modulo . In particular, when , since the permanent is the same as the determinant, we have that the number of perfect matchings in is odd if, and only if, , where the determinant is over . We do not expect that 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 is at least . And, we then show that we can transform any graph with to a graph so that and is still well-connected if 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 of a graph as being a matrix over 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 such that for all sufficiently large , .
Proof.
For any matrix , if , then the dimension of the null space of is , so there are non-zero vectors such that . We show that for the random graph , the expected number of vectors such that is at most linear in 222This 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 as goes to infinity.. Let denote the random variable that is the number of such vectors.
For an element , write for the row of indexed by . For a vector , let be the set and note that the condition is equivalent to the statement .
For any set and a bijection , we write to denote the image of under . Say that is a zero-sum set if is non-empty with . If is a zero-sum set, it must be the case that any vertex in has exactly two neighbours in . If , this implies that . In particular is even, and .
We now estimate the expected number of zero-sum sets of size in the random graph . Fix a set of size and choose three permutations of independently at random. Note that a random choice of means that is a uniformly random subset of of size . Thus, the probability that is:
For to be a zero-sum set, we further require that . The probability that a randomly chosen gives exactly this set is . Summing over the choices of the set , we get that the expected number of sets of size satisfying the condition , taken over all choices of is . 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 from above as goes to infinity. Hence we see that the probability that all three permutations are disjoint is at least and therefore the expected number of zero-sum sets in of size is at most .
Hence the total expectation of the number of zero-sum sets is
Since for all and , we get . By Markov’s inequality, it follows that the probability that exceeds is less than . Since the dimension of the null space of is , the theorem follows. ∎
To complete the construction, we show that if is a -regular bipartite graph with biadjacency matrix and , then under mild assumptions satisfied by almost all such graphs, we can edit to get so that and is at least -well-connected if is -well-connected.
Assume then, that is a -regular graph on two sets and of vertices each and is its biadjacency matrix. As before, we write to denote the row of indexed by . We drop the superscript where it is clear from context. We always treat these rows as vectors in .
We say that a pair of edges and of with and are switchable if they are disjoint and neither of nor is an edge. For a switchable pair we denote by the graph obtained from by exchanging the two edges and . That is, is the bipartite graph on the vertices , with edge set
Note that is also a -regular bipartite graph. We write for the biadjacency matrix of .
Assume now that . Then has a zero-sum set, i.e. a set such that . Moreover, by the -regularity of , we have . We can now state the lemma we aim to prove.
Lemma 5.10.
If has a zero-sum set with , then there are switchable edges so that .
Proof.
Fix a zero-sum set in with . Let denote the set of elements of which are neighbours in of vertices in . The assumption on the size of implies that and so .
For , write for the vector in which is at positions and and everywhere else. Now consider the following set of vectors in :
Observe that where is the subspace of consisting of vectors with even Hamming weight. To see this, first observe that for all pairs . When and , this is true by definition of . For , choose any and note that . Similarly, if , we can pick a and again . Since the collection of vectors clearly spans , we are done.
Let be the set of rows of . Since each has Hamming weight , so . Since , we conclude that . Let us fix a such that .
Since and is a zero-sum set, has exactly two neighbours in and one in . On the other hand, has three neighbours, all in . Thus, we can choose a which is a neighbour of but not and an which is a neighbour of but not of . Let and and observe that this pair is switchable by construction.
To prove the lemma, it then suffices to prove that . We do this by establishing the following two facts.
-
1.
; and
-
2.
for any , if , then .
Together these establish that the null space of is strictly smaller than that of and hence the claim.
Note that for all , we have , while and .
Thus, to prove the first fact, just note that since is in and is not, .
To prove the second fact, consider any set such that . We consider the following cases.
-
•
If neither nor is in , then , so since the former sum is , so is the latter.
-
•
If both and are in , then , and the same argument applies.
-
•
If exactly one of and is in , then . Hence, if , we must have . However, and were chosen so that , so this is impossible.
∎
Proof of Theorem 5.7.
By Lemma 5.8, for large enough values of , the random -regular graph is -well-connected for some constant with probability tending to . Thus, with high probability, the first three conditions are satisfied. If the biadjacency matrix of the resulting graph has rank , we are done.
If , then with probability at least , we have by Lemma 5.9. Moreover, since the expected number of zero-sum sets of size exactly is at most by the argument in the proof of Lemma 5.9, and this value tends to as grows, with high probability, has no zero-sum sets of this size. Hence, with positive probability, satisfies the pre-conditions of Lemma 5.10.
Note that if satisfies the conditions of Lemma 5.10, then any zero-sum set in the graph is also a zero-sum set in . Hence, if contains no zero-sum sets of size exactly , the same is true of . We can thus repeatedly apply the construction of Lemma 5.10 to obtain a graph such that . It then follows that has an odd number of perfect matchings. It remains to argue that is still well-connected. Note that since , we get from by at most applications of Lemma 5.10. At each step edges incident at most vertices are modified. Thus, if is a balanced separator in , then adding these four vertices to it gives us a balanced separator in . Thus, if is -well-connected, then is at least -well-connected and the result is proved.
∎
5.3 Playing the Game
Suppose is a bipartite -regular graph on two sets and of vertices each that is -well-connected. Let be the CFI-graph constructed from as described in Section 5.1, and and be two biadjacency matrices for where is obtained from by interchanging exactly one pair of columns. We aim to prove that Duplicator has a winning strategy in the -bijection game played on and .
Recall that is a bipartite graph on two sets and of vertices each with and . To say that Duplicator has a winning strategy in the -bijection game played on and is the same as saying that Duplicator has a winning strategy in the -bijection game played on the pair of graphs and where the latter is obtained from by swapping two elements . It does not matter which two elements we choose as any swap can be obtained from by composing with a permutation in . So, for some , fix two vertices which form a single pair in the gadget corresponding to and let . We write for the graph on which is exactly the same as except the neighbours of in are exactly the neighbours of in and the neighbours of in are exactly the neighbours of in .
Lemma 5.11.
Duplicator has a winning strategy in the -bijection game played on the graphs and .
To prove this lemma, we first introduce some notation and some observations about . Consider permutations of the vertices in the gadget in Figure 1 which fix each of the sets , and 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 , that exchanges with and with . The action of this automorphism on is to swap with and with . Each such automorphism consists of two swaps in and two in and so is a permutation in .
We can compose such automorphisms of the individual gadgets to get certain automorphisms of . Let be a simple cycle in the graph . That is, there are edges from to for each with and an edge from to . We define the permutation of as the composition
This is easily seen to be an automorphism of . Since it is the compostion of permutations each of which is in , .
Now, say that a permutation of is coherent if for each , and . We are only interested in coherent permutations. Let vertices and be neighbours in and let be the pair of vertices in which connect to vertices in . We say that a coherent permutation is good bar if composing it with the swap yields an automorphism of . 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 pebbled vertices of along with a permutation of which is obtained from the initial permutation by means of composing it with an element of . If is the vertex covered by pebble , let be the vertex of such that . In other words enumerate the vertices of whose gadgets in contain the pebbled vertices. Note that by the connectedness assumption, has a component which contains more than half of the vertices of , and is -connected. We call the large component at this game position.
We show that Duplicator can play to maintain the following invariant:
- (*)
-
there is an edge of such that and are both in the large component and is a coherent permutation that is good bar .
In particular, this guarantees that is a partial isomorphism on the pebbled positions. Indeed, is a partial isomorphism on the graph excluding , 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 satisfies (*) as no vertices are pebbled.
At each subsequent move, Spoiler places a pebble on a vertex . Duplicator chooses any edge in the large component (this could be if they are still in the large component). We then distinguish two cases.
-
1.
If then Duplicator’s response is to compose with . This is a valid move as none of the four vertices is pebbled and the pemutation is in since all four of are in . Moreover, the fact that is good bar implies that composing it with yields an automorphism of . Composing this automorphism with gives us a permutation that is good bar . Thus, the invariant (*) is maintained.
-
2.
If Duplicator must compose with a permutation that fixes , so in fact fixes both and . By the fact that the large component before the pebble is placed on is -connected, we know that there is a path in from to that does not use the edge . Combining this with we obtain a simple cycle . Then, is an automorphism of which, in particular, swaps and . Duplicator’s move is to compose with the permutation . Call the resulting permutation . Observe first that this is a valid move, as and we are composing it with two swaps of elements of . It remains to argue that is good bar . By assumption composing with yields an automorphism of and is also an automorphism of . Thus, is an automorphism of and is just the composition of this with .
∎
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 there exists a -regular balanced bipartite graph with vertices that is -well-connected for and has an odd number of perfect matchings. Let be the CFI-graph constructed from as described in Section 5.1, and and be two biadjacency matrices for where is obtained from by interchanging exactly one pair of columns. From Lemma 5.11 we have that Duplicator has a winning strategy for the -bijection game on and . From Lemma 5.6 and the fact that has an odd number of perfect matchings, it follows that and so, since , we have . 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 , when is an odd prime, provided that the matrices we construct have non-zero determinant modulo . As noted in the proof of Lemma 5.6, , where the sum is over all -factors (or equivalently over all perfect matchings) of the bipartite graph . Of course, is just the determinant of the bi-adjacency matrix of . Hence, for odd prime , provided that . Now, the construction in Section 5.2 is aimed at ensuring that . It seems plausible that we could just as well ensure that for some odd prime . This would immediately yield the analogue of Theorem 5.1 in characteristic . 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 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 , a pair of bipartite graphs and on which Duplicator has a winning strategy in the -pebble bijection game and which have different numbers of perfect matchings. The graphs and are on two sets and of vertices each and the difference between the number of perfect matchings in and is a power of . The -pebble bijection game for which a Duplicator winning strategy is shown is essentially the game. This shows that the biadjacency matrices of and cannot be distinguished by -symmetric circuits of subexponential size and also that the adjacency matrices of and cannot be distinguished by -symmetric circuits of subexponential size. Since and have different numbers of perfect matchings, their biadjacency matrices and have distinct permanents. Since the number of perfect matchings differ by a power of , they have distinct permanents modulo for any odd prime . 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 and 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 and even in the restricted game . 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 from given in Section 5.1 is essentially the original construction given by Cai et al. [6]. Their construction gives from a pair of non-isomorphic graphs which are not distinguishable by -dimensional Weisfeiler-Leman equivalence. We use only one graph 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 by “twisting” one of the edges. That is, for some edge of we replace the two edges and by the edges and . 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 is obtained from by first, for each edge of , contracting the two edges and in and secondly, for each vertex of , adding a new vertex to which is adjacent to all four vertices in . Overall, this is equivalent to replacing each vertex of with incident edges with the gadget in Figure 4 where the dashed lines indicate edges whose endpoints are in other gadgets.
The resulting graph is a -regular bipartite graph and is obtained from it by taking one vertex of and in the corresponding gadget, for one edge incident on , interchanging the connections of and . The fact that and 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 -bijection game played on and . The winning strategy in the ordinary -pebble bijection game, is described in [9] and is based on the fact that the graph has tree-width greater than and so there is a winning strategy for robber in the -cops-and-robbers game played on . This is lifted to a winning strategy for Duplicator in the -bijection game which sees Duplicator maintain a bijection that is an isomorphism except at one gadget corresponding to a vertex of . The position is given as a robber winning position in a -cops-and-robbers game played on the graph . At each move, Duplicator takes a path from to in describing a robber move and changes to a bijection by composing it with automorphisms for the gadgets corresponding to the vertices along the path. This results in being an isomorphism except at the gadget corresponding to . It can be easily checked that is obtained from by composing with a permutation in provided that the path from to is of even length. It is easy to ensure that the graph of tree-width greater than used is bipartite and robber has a winning strategy in the -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 be a field of characteristic other than . There is no family of -symmetric circuits over of size computing the permanent over .
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 -complete graph problems including -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 matrix 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 is more subtle. If is a symmetric matrix, then we can see at as the adjacency matrix of a graph on vertices and the determinant is a graph invariant, that is to say it only depends on the isomorphism class of . 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 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 , so simultaneous permutations of the rows and columns of and the upper bounds obtained still apply. But, we can also think of as a biadjacency matrix of a bipartite graph , and now the determinant of is not an invariant of . 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 that fixes the determinant of . Indeed, we do this for the smaller group . 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 , 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
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.