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

On the Complexity of Identifying Strongly Regular Graphs

Michael Levet 111 This work was partially supported by J. A. Grochow’s startup funds and NSF award CISE-2047756. I wish to thank C.J. Colbourn for helping me to find a copy of [Col79]. I also wish to thank J.N. Cooper and J.A. Grochow for helpful discussions. Department of Computer Science- University of Colorado Boulder
Abstract

In this paper, we show that Graph Isomorphism (GI) is not AC0\textsf{AC}^{0}-reducible to several problems, including the Latin Square Isotopy problem, isomorphism testing of several families of Steiner designs, and isomorphism testing of conference graphs. As a corollary, we obtain that GI is not AC0\textsf{AC}^{0}-reducible to isomorphism testing of Latin square graphs and strongly regular graphs arising from special cases of Steiner 22-designs. We accomplish this by showing that the generator-enumeration technique for each of these problems can be implemented in β2FOLL\beta_{2}\textsf{FOLL}, which cannot compute Parity (Chattopadhyay, Torán, & Wagner, ACM Trans. Comp. Theory, 2013).

Keywords: Latin squares, Quasigroups, Graph Isomorphism, Steiner designs, Conference graphs.

1 Introduction

In one of the original papers on Graph Isomorphism (GI), by Corneil & Gotlieb, the authors observed that in practice, strongly regular graphs serve as difficult instances [CG70]. A common approach for producing such instances is to construct strongly regular graphs from combinatorial objects, such as Latin squares, nets (partial geometries), and combinatorial designs [CG70, SD76]. While strongly-regular graphs have been perceived as hard instances, there is little evidence to suggest that they are GI-complete.

There is a beautiful classification of strongly-regular graphs (of valency ρ<n/2\rho<n/2, which is essentially without loss of generality by taking complements) due to Neumaier [Neu79]. In using this classification in isomorphism testing, Spielman [Spi96] and Babai–Wilmes [BW13] organize it as follows:

  1. (a)

    Latin square graphs,

  2. (b)

    Line graphs of Steiner 22-designs satisfying ρ<f(n)\rho<f(n), for a certain function f(n)n3/4f(n)\sim n^{3/4},

  3. (c)

    Strongly-regular graphs of degree ρ=(n1)/2\rho=(n-1)/2 (a.k.a. conference graphs),

  4. (d)

    Graphs satisfying a certain eigenvalue inequality referred to as Neumaier’s claw bound.

In this paper, we show (in particular) that (a) Latin square graphs, (b) line graphs arising from special cases of Steiner 22-designs, and (c) conference graphs are not GI-hard under AC0\textsf{AC}^{0}-computable many-one reductions. As with [CTW13], we show this by improving the parallel complexity of isomorphism testing in these classes of graphs to β2FOLL\beta_{2}\textsf{FOLL}, which does not compute Parity (in the case of conference graphs, we achieve a stronger bound of β2AC0\beta_{2}\textsf{AC}^{0}). As Parity is AC0\textsf{AC}^{0}-reducible to GI [Tor04], this rules out AC0\textsf{AC}^{0}-reductions (and more strongly, for any i,c0i,c\geq 0, we rule out βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c})-reductions).

Prior to our work, there were few pieces of complexity-theoretic evidence to suggest that strongly-regular graphs are not GI-complete. Babai showed that there is no functorial reduction from GI to the isomorphism testing of strongly-regular graphs [Bab14]. We note that almost all known reductions between isomorphism problems are functorial (c.f. [Bab14]). An example where this is not the case is the reduction from 11-Block Conjugacy of shifts of finite type to kk-Block Conjugacy [SF21, Theorem 18]. In the case of Quasigroup Isomorphism, Chattopadhyay, Torán, and Wagner improved the upper bound to β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL}. In particular, they showed that for any i,c0i,c\geq 0, GI is not βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c})-reducible to Quasigroup Isomorphism [CTW13].

Let us briefly discuss why problems such as Quasigroup Isomorphism, Latin Square Isotopy, and isomorphism testing of Steiner 22-designs are important from a complexity-theoretic perspective. Several families of strongly regular graphs (e.g., Latin square graphs, and the block-intersection graphs of Steiner 22-designs) arise naturally from the isomorphism problems from the underlying combinatorial objects. Precisely, many of these problems, including Quasigroup Isomorphism, Latin Square Isotopy, and isomorphism testing of Steiner 22-designs, are polynomial-time (and in fact, AC0\textsf{AC}^{0}) reducible to GI. Polynomial-time solutions for these problems have been elusive. In particular, each of these problems are candidate NP-intermediate problems; that is, problems that belong to NP, but are neither in P nor NP-complete [Lad75].

There is considerable evidence suggesting that GI is not NP-complete [Sch88, BH92, IPZ01, Bab16, KST92, AK06, Mat79]. As Quasigroup Isomorphism, Latin Square Isotopy, and isomorphism testing of Steiner 22-designs reduce to GI, this evidence also suggests that none of these problems is NP-complete. Furthermore, Chattopadhyay, Torán, & Wagner [CTW13] showed that Quasigroup Isomorphism is not NP-complete under AC0\textsf{AC}^{0}-reductions, and our result shows the same for Latin Square Isotopy. In light of Babai’s nΘ(log2(n))n^{\Theta(\log^{2}(n))}-time algorithm [Bab16], these problems serve as key barries to placing GI into P.

Main Results. In this paper, we show that several families of isomorphism problems characterized by the generator enumeration technique are not GI-complete under βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c})-reductions, for any choice of i,c0i,c\geq 0. In particular, this rules out AC0\textsf{AC}^{0}-reductions from GI to these problems. We also show that these results extend to the corresponding families of strongly regular graphs, providing complexity-theoretic evidence that isomorphism testing of strongly regular graphs may not be GI-complete.

The first problem we consider is the Latin Square Isotopy problem, which takes as input two quasigroups L1,L2L_{1},L_{2} given by their multiplication tables and asks if there exist bijections α,β,γ:L1L2\alpha,\beta,\gamma:L_{1}\to L_{2} such that α(x)β(y)=γ(xy)\alpha(x)\beta(y)=\gamma(xy) for all x,yL1x,y\in L_{1}.

Theorem 1.1.

Latin Square Isotopyβ2Lβ2FOLL\textsc{Latin Square Isotopy}\in\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL}.

Remark 1.2.

This improves the previous bound of β2NC2\beta_{2}\textsf{NC}^{2} due to Wolf [Wol94].

To prove Theorem 1.1, we leverage cube generating sequences, in a similar manner as Chattopadhyay, Torán, & Wagner [CTW13, Theorem 3.4]. After non-deterministically guessing cube generating sequences, we can check in LFOLL\textsf{L}\cap\textsf{FOLL} whether the induced map extends to an isotopy of the Latin squares.

Now for any i,c0i,c\geq 0, we have that βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c}) cannot compute Parity [CTW13]. Thus, we obtain the following corollary.

Corollary 1.3.

For any i,c0i,c\geq 0, GI is not βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c})-reducible to Latin Square Isotopy.

Latin squares yield a corresponding family of strongly regular graphs, known as Latin square graphs, where two Latin square graphs G1G_{1} and G2G_{2} are isomorphic if and only if their corresponding Latin squares L(G1)L(G_{1}) and L(G2)L(G_{2}) are main class isotopic [Mil78, Lemma 3] (that is, the Latin squares can be put into compatible normal forms that correspond to isomorphic quasigroups). Miller previously showed that it is possible to recover the corresponding Latin square from a Latin square graph in polynomial-time [Mil78], and Wolf strengthened the analysis to show that this reduction is in fact NC1\textsf{NC}^{1}-computable [Wol94]. A closer analysis yields that this reduction is actually AC0\textsf{AC}^{0}-computable (see Lemma 3.9). Furthermore, we can place a Latin square into its main isotopy class in AC0\textsf{AC}^{0} (see Remark 1.6). Thus, we obtain the following corollary.

Corollary 1.4.

Deciding whether two Latin square graphs are isomorphic is in β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL}. Consequently, for any i,c0i,c\geq 0, GI is not βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c})-reducible to isomorphism testing of Latin square graphs.

Remark 1.5.

It is also possible to construct a Latin square graph from a Latin square using an AC0\textsf{AC}^{0}-circuit. Thus, Latin Square Isotopy and isomorphism-testing of Latin square graphs are equivalent under AC0\textsf{AC}^{0}-reductions.

Remark 1.6.

We sketch an alternative proof of Theorem 1.1 and Corollary 1.4. In Miller’s [Mil78] first proof that Latin Square Isotopy is solvable in time nlog(n)+O(1)n^{\log(n)+O(1)}, Miller takes one Latin square LL^{\prime} and places it in a normal form. Miller shows that this step is polynomial-time computable. A closer analysis shows that it is in fact AC0\textsf{AC}^{0}-computable: we may label the first row and first column as 1,2,,n1,2,\ldots,n. We then in AC0\textsf{AC}^{0} fill in the remaining cells of the normal form. For the second Latin square LL, Miller places LL in n2n^{2} possible normal forms by guessing the element to label as 11. We may consider all such guesses in parallel, and so this step is also AC0\textsf{AC}^{0}-computable. Miller then checks whether the normal form of LL^{\prime} and the normal form of LL are isomorphic as quasigroups. This step is β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL}-computable [CTW13]. Thus, we have an AC0\textsf{AC}^{0}-computable disjunctive truth-table reduction (see Section 2.2) from Latin Square Isotopy to Quasigroup Isomorphism, which places Latin Square Isotopy into β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL}. We note that Wolf’s [Wol94] proof that Latin Square Isotopyβ2NC2\textsc{Latin Square Isotopy}\in\beta_{2}\textsf{NC}^{2} followed a similar strategy as Miller’s [Mil78] proof that Quasigroup Isomorphism can be solved in time nlog(n)+O(1)n^{\log(n)+O(1)}.

As we can recover a Latin square from a Latin square graph in AC0\textsf{AC}^{0} (see Lemma 3.9), we also have an AC0\textsf{AC}^{0}-computable disjunctive truth table reduction from isomorphism testing of Latin square graphs to Quasigroup Isomorphism. So isomorphism testing of Latin square graphs belongs to β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL}.

Instead, the proof we give of Theorem 1.1 (see Section 3) exhibits how the technique of cube generating sequences [CTW13] can be fruitfully applied directly to isotopy (rather than isomorphism) of algebraic structures, which we do by combining cube generating sequences with Miller’s [Mil78] second proof that Latin Square Isotopy can be solved in time nlog(n)+O(1)n^{\log(n)+O(1)}.

We now turn our attention to isomorphism testing of Steiner designs.

Theorem 1.7.

For any i,c0i,c\geq 0, GI is not βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c})-reducible to isomorphism testing of Steiner (t,t+1)(t,t+1)-designs.

We prove Theorem 1.7 in two steps. First, we establish the case of t=2t=2, which is the case of Steiner triple systems. It is well-known that Steiner triple systems correspond to quasigroups, and that this reduction is polynomial-time computable [Mil78]. A careful analysis shows that this reduction is in fact AC0\textsf{AC}^{0}-computable. Furthermore, we observe that the reduction in [BW13, Theorem 33] from isomorphism testing of Steiner tt-designs to isomorphism testing of Steiner 22-designs is β2AC0\beta_{2}\textsf{AC}^{0}-computable. As a corollary, isomorphism testing of Steiner (t,t+1)(t,t+1)-designs is β2AC0\beta_{2}\textsf{AC}^{0}-reducible to isomorphism testing of Steiner triple systems, and hence Quasigroup Isomorphism.

Steiner designs yield a family of graphs known as block intersection graphs. In the case of Steiner 22-designs, these graphs are strongly regular. When the block size is bounded, we observe that we may recover a Steiner 22-design from its block-incidence graph in AC0\textsf{AC}^{0}. If the block size is not bounded, this reduction is TC0\textsf{TC}^{0}-computable. In the case of Steiner triple systems, this yields the following corollary.

Corollary 1.8.

Let G1,G2G_{1},G_{2} be block-incidence graphs arising from Steiner triple systems. We can decide whether G1G2G_{1}\cong G_{2} in β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL}. Consequently, for any i,c0i,c\geq 0, GI is not βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c})-reducible to isomorphism testing of block-incidence graphs arising from Steiner triple systems.

We finally consider the case of conference graphs.

Theorem 1.9.

Isomorphism testing between a conference graph GG and an arbitrary graph HH belongs to β2AC0\beta_{2}\textsf{AC}^{0}. Consequently, for any i,c0i,c\geq 0, there is no βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c})-reduction from GI to isomorphism testing of conference graphs.

To prove Theorem 1.9, we use a result of Babai [Bab80], who showed that conference graphs admit so called distinguishing sets (see Section 5) of size O(logn)O(\log n). Distinguishing sets are quite powerful: after indvidiualizing the vertices in a distinguishing set, only two rounds of the count-free Color Refinement algorithm are needed to assign each vertex in the graph a unique label. The β2AC0\beta_{2}\textsf{AC}^{0} algorithm thus works as follows: we first guess such a set SS using O(log2n)O(\log^{2}n) non-deterministic bits, individualize the vertices in SS, and then run two rounds of the count-free Color Refinement algorithm.

Remark 1.10.

Theorem 1.9 provides evidence that isomorphism testing of conference graphs might in fact be easier than Group Isomorphism (and hence, isomorphism testing of Latin square graphs and Steiner designs). The best known complexity-theoretic upper bound for Group Isomorphism is β2Lβ2FOLLβ2SC2\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL}\cap\beta_{2}\textsf{SC}^{2} (the β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL} is due to Chattopadhyay, Torán, & Wagner [CTW13]; Tang [Tan13] established the β2SC2\beta_{2}\textsf{SC}^{2} bound and provided an alternative proof of the β2L\beta_{2}\textsf{L} bound). Even after fixing a cube generating sequence, it is not clear that an AC circuit of depth o(loglogn)o(\log\log n) and polynomial size could uniquely identify each group element.

These differences are quite opaque when examined from the algorithmic perspective. Both conference graphs and groups admit nΘ(logn)n^{\Theta(\log n)}-time isomorphism tests that have seen little improvement in the worst case in over 40 years. For conference graphs, this bound is achieved using Babai’s [Bab80] result showing the existence of size O(logn)O(\log n) distinguishing sets. In the setting of groups, Tarjan (c.f., [Mil78]) leveraged the fact that every group admits a generating set of size logpn\leq\lceil\log_{p}n\rceil (where pp is the smallest prime dividing nn) to obtain an nlogp(n)+O(1)n^{\log_{p}(n)+O(1)} isomorphism test via the generator enumeration strategy. Rosenbaum [Ros13] (see [LGR16, Sec. 2.2]) improved this to n(1/4)logp(n)+O(1)n^{(1/4)\log_{p}(n)+O(1)}. And even the impressive body of work on practical algorithms for this problem, led by Eick, Holt, Leedham-Green and O’Brien (e. g., [BEO02, ELGO02, BE99, CH03]) still results in an nΘ(logn)n^{\Theta(\log n)}-time222Where we stress n=|G|n=|G|. In the setting of practical algorithms, the groups are often given by generating sequences, and so polynomial time in the size of the succinct input is polylog|G|\operatorname{poly}\log|G|. algorithm in the general case (see [Wil19, Page 2]).

Further Related Work. There has been significant work on isomorphism testing of strongly regular graphs. In 1980, Babai used a simple combinatorial approach to test strongly regular graphs on nn vertices and degree ρ<n/2\rho<n/2 in time exp(O((n/ρ)log2n))\text{exp}(O((n/\rho)\log^{2}n)) [Bab80, Bab81]. We note that the complement of a strongly-regular graph is strongly regular, hence the assumptions on the degree. As ρn1\rho\geq\sqrt{n-1}, this gives an algorithm in moderately exponential time exp(O(nlog2n))\text{exp}(O(\sqrt{n}\log^{2}n)) for all ρ\rho. Spielman [Spi96] subsequently improved this exp(O((n/ρ)log2n))\text{exp}(O((n/\rho)\log^{2}n)) bound to exp(O(n1/3log2n))\exp(O(n^{1/3}\log^{2}n)). This improved upon what was at the time, the best known bound of exp(O(nlogn))\exp(O(\sqrt{n\log n})) for isomorphism testing of general graphs [BL83, BKL83, ZKT85] based on Luks’ group theoretic method [Luk82].

We note that Babai’s [Bab80] work already handled the case of conference graphs in time nO(logn)n^{O(\log n)}. Furthermore, the works of Babai and Spielman sufficed to handle isomorphism testing of strongly regular graphs that satisfy Neumaier’s claw bound. Namely, Spielman [Spi96] handled the case of ρo(n2/3)\rho\in o(n^{2/3}) in time exp(O(n1/4log2n))\text{exp}(O(n^{1/4}\log^{2}n)), and Babai [Bab80] handled the case of ρΩ(n2/3)\rho\in\Omega(n^{2/3}) in time exp(O(n1/3log2n))\text{exp}(O(n^{1/3}\log^{2}n)).

In the case of Latin square graphs, Miller [Mil78] resolved this in time nO(logn)n^{O(\log n)}. Wolf improved the complexity theoretic bound to β2NC2\beta_{2}\textsf{NC}^{2} (uniform NC2\textsf{NC}^{2}-circuits that accept O(log2n)O(\log^{2}n) existentially quantified non-deterministic bits).

The isomorphism problem for Steiner 22-designs also dates back to Miller [Mil78], who considered the special cases of Steiner triple systems (Steiner (2,3,n)(2,3,n)-designs), projective planes (Steiner (2,q+1,q2+q+1)(2,q+1,q^{2}+q+1)-designs), and affine planes (Steiner (2,q,q2)(2,q,q^{2})-designs). In the case of Steiner triple systems, Miller leveraged a standard reduction to Quasigroup Isomorphism to obtain an nO(logn)n^{O(\log n)} upper bound. M. Colbourn [Col79] subsequently extended this result to obtain an vO(logv)v^{O(\log v)} canonization procedure for Steiner (t,t+1)(t,t+1)-designs.

For the cases of projective and affine planes, Miller [Mil78] showed that isomorphism testing can be done in time nO(loglogn)n^{O(\log\log n)}. In particular, Miller showed that the corresponding structures may be recovered from the block-intersection graphs in polynomial-time, providing the block-intersection graphs are of sufficiently high degree [Mil78]. Thus, isomorphism testing of block-intersection graphs arising from Steiner triple systems can be done in time nO(logn)n^{O(\log n)}, and the block-intersection graphs arising from projective and affine planes can be identified in time nO(loglogn)n^{O(\log\log n)}.

In 1983, Babai & Luks showed that Steiner 22-designs with blocks of bounded size admit an nO(logn)n^{O(\log n)} canonization procedure (and hence, isomorphism test). In paritcular, isomorphism testing of the corresponding block-intersection graphs is solvable in time nO(logn)n^{O(\log n)} [BL83]. Here, Babai & Luks utilized Luks’ group theoretic method [Luk82]. Huber later extended this result to obtain an nO(logn)n^{O(\log n)}-runtime isomorphism test for arbitrary tt-designs, where both tt and the block size are bounded, as well as the corresponding block-intersection graphs [Hub11].

Spielman solved the general case of strongly-regular graphs of degree f(n)n3/4f(n)\sim n^{3/4} arising from Steiner 22-designs in time exp(O(n1/4log2n))\text{exp}(O(n^{1/4}\log^{2}n)) [Spi96]. Independently, Babai & Wilmes [BW13] and Chen, Sun, & Teng [CST13] exhibited an nO(logn)n^{O(\log n)}-runtime isomorphism test for both Steiner 22-designs and the corresponding block-intersection graphs. Babai & Wilmes extended their result to Steiner tt-designs. Here, each of these papers utilized the individualize and refine technique [Spi96, BW13, CST13].

The isomorphism problem for combinatorial designs is another candidate NP-intermediate problem. In the general case, testing whether two combinatorial designs are isomorphic is GI-complete under polynomial-time Turing reductions [CC81]. Cyclic Steiner 22-designs of prime order are known to admit a polynomial-time isomorphism test [CM80]. To the best of our knowledge, the complexity of deciding whether two cyclic Steiner 22-designs are isomorphic remains open for arbitrary orders.

2 Preliminaries

2.1 Algebra and Combinatorics

Graph Theory. A strongly regular graph with parameters (n,k,λ,μ)(n,k,\lambda,\mu) is a simple, undirected kk-regular, nn-vertex graph G(V,E)G(V,E) where any two adjacent vertices share λ\lambda neighbors, and any two non-adjacent vertices share μ\mu neighbors. The complement of a strongly regular graph is also strongly regular, with parameters (n,nk1,n22k+μ,n2k+λ)(n,n-k-1,n-2-2k+\mu,n-2k+\lambda).

Quasigroups. A quasigroup consists of a set GG and a binary operation :G×GG\star:G\times G\to G satisfying the following. For every a,bGa,b\in G, there exist unique x,yx,y such that ax=ba\star x=b and ya=by\star a=b. When the multiplication operation is understood, we simply write axax for axa\star x.

Unless otherwise stated, all quasigroups are assumed to be finite and represented using their Cayley (multiplication) tables. For a quasigroup of order nn, thet Cayley table has n2n^{2} entries, each represented by a string of size log2(n)\lceil\log_{2}(n)\rceil.

A Latin square of order nn is an n×nn\times n matrix LL where each cell Lij[n]L_{ij}\in[n], and each element of [n][n] appears exactly once in a given row or a given column. Latin squares are precisely the Cayley tables corresponding to quasigroups. We will abuse notation by referring to a quasigroup and its multiplication table interchangeably. An isotopy of Latin squares L1L_{1} and L2L_{2} is an ordered triple (α,β,γ)(\alpha,\beta,\gamma), where α,β,γ:L1L2\alpha,\beta,\gamma:L_{1}\to L_{2} are bijections satisfying the following: whenever ab=cab=c in L1L_{1}, we have that α(a)β(b)=γ(c)\alpha(a)\beta(b)=\gamma(c) in L2L_{2}. Alternatively, we may view α\alpha as a permutation of the rows of L1L_{1}, β\beta as a permutation of the rows of L2L_{2}, and γ\gamma as a permutation of the values in the table. Here, L1L_{1} and L2L_{2} are isotopic precisely if (i) the (i,j)(i,j) entry of L1L_{1} is the (α(i),β(j))(\alpha(i),\beta(j)) entry of L2L_{2}, and (ii) xx is the (i,j)(i,j) entry of L1L_{1} if and only if γ(x)\gamma(x) is the (α(i),β(j))(\alpha(i),\beta(j)) entry of L2L_{2}.

As quasigroups are non-associative, the parenthesization of a given expression may impact the resulting value. We restrict attention to balanced parenthesizations, which ensure that words of the form g0g1gkg_{0}g_{1}\cdots g_{k} are evaluated using a balanced binary tree with k+1k+1 leves and therefore depth O(loglogn)O(\log\log n).

For a sequence S:=(s0,s1,,sk)S:=(s_{0},s_{1},\ldots,s_{k}) from a quasigroup, define:

Cube(S)={s0s1e1skek:e1,,ek{0,1}}.\text{Cube}(S)=\{s_{0}s_{1}^{e_{1}}\cdots s_{k}^{e_{k}}:e_{1},\ldots,e_{k}\in\{0,1\}\}.

We say that SS is a cube generating sequence if each element gg in the quasigroup can be written as g=s0s1e1skekg=s_{0}s_{1}^{e_{1}}\cdots s_{k}^{e_{k}}, for e1,,ek{0,1}e_{1},\ldots,e_{k}\in\{0,1\}. Every quasigroup is known to admit a cube generating sequence of size O(logn)O(\log n) [CTW13].

For a given Latin square LL of order nn, we associate a Latin square graph G(L)G(L) that has n2n^{2} vertices; one for each triple (a,b,c)(a,b,c) that satisfies ab=cab=c. Two vertices (a,b,c)(a,b,c) and (x,y,z)(x,y,z) are adjacent in G(L)G(L) precisely if a=xa=x, b=yb=y, or c=zc=z. Miller showed that two Latin square graphs G1G_{1} and G2G_{2} are isomorphic if and only if the corresponding Latin squares, L1L_{1} and L2L_{2}, are main class isotopic; that is, if L1L_{1} and L2L_{2} can be placed into compatible normal forms that correspond to isomorphic quasigroups [Mil78]. A Latin square graph on n2n^{2} vertices is a strongly regular graph with parameters (n2,3(n1),n,6)(n^{2},3(n-1),n,6). Conversely, a strongly regular graph with these same parameters (n2,3(n1),n,6)(n^{2},3(n-1),n,6) is called a pseudo-Latin square graph. Bruck showed that for n>23n>23, a pseudo-Latin square graph is a Latin square graph [Bru63].

Albert showed that a quasigroup QQ is isotopic to a group GG if and only if QQ is isomorphic to GG. In general, isotopic quasigroups need not be isomorphic [Alb43].

Nets. A net or partial geometry of order n1n\geq 1 and degree k1k\geq 1, denoted 𝒩(n,k)\mathcal{N}(n,k), consists of a set PP of n2n^{2} points and a set LL of knkn lines satisfying the following.

  1. (a)

    Each line in LL has nn points.

  2. (b)

    Each point in PP lies on exactly kk distinct lines.

  3. (c)

    The knkn lines of LL fall into kk parallel classes. No two lines in the same parallel class intersect. If 1,2L\ell_{1},\ell_{2}\in L are in different parallel classes, then 1\ell_{1} and 2\ell_{2} share exactly one common point.

Necessarily, kn+1k\leq n+1. If k=n+1k=n+1, then the net is called an affine plane of order nn. A projective plane is the extension of an affine plane 𝒩\mathcal{N} by adding n+1n+1 new points p1,,pn+1p_{1},\ldots,p_{n+1} (one per parallel class). The point pip_{i} is then added to each line of the iith parallel class. We finally add a new line containing {p1,,pn+1}\{p_{1},\ldots,p_{n+1}\}. We may recover an affine plane from a projective plane by removing a line and the associated points from the projective plane. We assume that nets are given by the point-block incidence matrix.

Given a net 𝒩(n,k)\mathcal{N}(n,k) where knk\leq n, we may construct a net graph of order nn and degree kk G(V,E)G(V,E) on n2n^{2} vertices corresponding to the points in 𝒩\mathcal{N}. Two vertices are adjacent in GG if their corresponding points in 𝒩\mathcal{N} determine a line. We note that a net graph uniquely determines the net 𝒩(n,k)\mathcal{N}(n,k) [Bru63]. Miller showed that, provided n>(k1)2n>(k-1)^{2}, this equivalence holds under polynomial-time reductions [Mil78]. A net graph is strongly-regular with parameters (n2,k(n1),n2+(k1)(k2),k(k1))(n^{2},k(n-1),n-2+(k-1)(k-2),k(k-1)). Conversely, a strongly regular graph with parameters (n2,k(n1),n2+(k1)(k2),k(k1))(n^{2},k(n-1),n-2+(k-1)(k-2),k(k-1)) is referred to as a pseudo-net graph. Bruck showed that for nn sufficiently large, a pseudo-net graph of order nn and degree kk is a net graph [Bru63].


Designs. Let tkvt\leq k\leq v and λ\lambda be positive integers. A (t,k,λ,v)(t,k,\lambda,v) design is an incidence structure 𝒟=(X,,I)\mathcal{D}=(X,\mathcal{B},I), where XX denotes our set of vv points. Now \mathcal{B} is a subset of (Xk)\binom{X}{k}, where each element of \mathcal{B} is referred to as a block, satisfying the property that each tt-element subset of XX belongs to exactly λ\lambda blocks. If t<k<vt<k<v, we say that the design is non-trivial. If λ=1\lambda=1, the design is referred to as a Steiner design. We denote Steiner designs as (t,k,v)(t,k,v)-designs when we want to specify vv the number of points, or Steiner (t,k)(t,k)-designs when referring to a family of designs. We note that Steiner (2,3)(2,3)-designs are known as Steiner triple systems. Projective planes are Steiner (2,q+1,q2+q+1)(2,q+1,q^{2}+q+1)-designs, and affine planes are Steiner (2,q,q2)(2,q,q^{2})-designs. We assume that designs are given by the point-block incidence matrix.

Let 𝒟=(X,,I)\mathcal{D}=(X,\mathcal{B},I) be a design, and let AXA\subseteq X. The derived design at AA, denoted 𝒟(A)\mathcal{D}(A), has the set of points XAX\setminus A and blocks {BA:B,AB}\{B\setminus A:B\in\mathcal{B},A\subsetneq B\}. If 𝒟\mathcal{D} is a Steiner (t,k,v)(t,k,v)-design, then 𝒟(A)\mathcal{D}(A) is a Steiner (t|A|,k|A|,v|A|)(t-|A|,k-|A|,v-|A|)-design.

For a design 𝒟=(X,,I)\mathcal{D}=(X,\mathcal{B},I), we may define a block intersection graph (also known as a line graph) G(V,E)G(V,E), where V(G)=V(G)=\mathcal{B} and two blocks B1,B2B_{1},B_{2} are adjacent in GG if and only if B1B2B_{1}\cap B_{2}\neq\emptyset. In the case of a Steiner 22-design, the block-intersection graph is strongly regular. For Steiner triple systems, the block-intersection graphs are strongly regular with parameters (n(n1)/6,3(n3)/2,(n+3)/2,9)(n(n-1)/6,3(n-3)/2,(n+3)/2,9). Conversely, strongly regular graphs with the parameters (n(n1)/6,3(n3)/2,(n+3)/2,9)(n(n-1)/6,3(n-3)/2,(n+3)/2,9) are referred to as pseudo-STS graphs. Bose showed that pseudo-STS graphs graphs with strictly more than 6767 vertices are in fact STS graphs [Bos63].

2.2 Computational Complexity

We assume familiarity with Turing machines and the complexity classes P,NP,L\textsf{P},\textsf{NP},\textsf{L}, and NL- we refer the reader to standard references [Zoo, AB09].

Circuit Complexity. For a standard reference, see [Vol99]. We consider Boolean circuits using the AND, OR, NOT, and Majority, where Majority(x1,,xn)=1\textsf{Majority}(x_{1},\ldots,x_{n})=1 if and only if n/2\geq n/2 of the inputs are 11. Otherwise, Majority(x1,,xn)=0\textsf{Majority}(x_{1},\ldots,x_{n})=0. In this paper, we will consider logspace uniform circuit families (Cn)n(C_{n})_{n\in\mathbb{N}}, in which a deterministic logspace Turing machine can compute the map 1nCn1^{n}\mapsto\langle C_{n}\rangle (here, Cn\langle C_{n}\rangle denotes an encoding of the circuit CnC_{n}).

Definition 2.1.

Fix k0k\geq 0. We say that a language LL belongs to (logspace uniform) NCk\textsf{NC}^{k} if there exist a (logspace uniform) family of circuits (Cn)n(C_{n})_{n\in\mathbb{N}} over the AND,OR,NOT\textsf{AND},\textsf{OR},\textsf{NOT} gates such that the following hold:

  • The AND and OR gates take exactly 22 inputs. That is, they have fan-in 22.

  • CnC_{n} has depth O(logkn)O(\log^{k}n) and uses (has size) nO(1)n^{O(1)} gates. Here, the implicit constants in the circuit depth and size depend only on LL.

  • xLx\in L if and only if C|x|(x)=1C_{|x|}(x)=1.

The complexity class ACk\textsf{AC}^{k} is defined analogously as NCk\textsf{NC}^{k}, except that the AND,OR\textsf{AND},\textsf{OR} gates are permitted to have unbounded fan-in. That is, a single AND gate can compute an arbitrary conjunction, and a single OR gate can compute an arbitrary disjunction. The complexity class TCk\textsf{TC}^{k} is defined analogously as ACk\textsf{AC}^{k}, except that our circuits are now permitted Majority gates of unbounded fan-in. For every kk, the following containments are well-known:

NCkACkTCkNCk+1.\textsf{NC}^{k}\subseteq\textsf{AC}^{k}\subseteq\textsf{TC}^{k}\subseteq\textsf{NC}^{k+1}.

In the case of k=0k=0, we have that:

NC0AC0TC0NC1LNLAC1.\textsf{NC}^{0}\subsetneq\textsf{AC}^{0}\subsetneq\textsf{TC}^{0}\subseteq\textsf{NC}^{1}\subseteq\textsf{L}\subseteq\textsf{NL}\subseteq\textsf{AC}^{1}.

We note that functions that are NC0\textsf{NC}^{0}-computable can only depend on a bounded number of input bits. Thus, NC0\textsf{NC}^{0} is unable to compute the AND function. It is a classical result that AC0\textsf{AC}^{0} is unable to compute Parity [Smo87]. The containment TC0NC1\textsf{TC}^{0}\subseteq\textsf{NC}^{1} (and hence, TCkNCk+1\textsf{TC}^{k}\subseteq\textsf{NC}^{k+1}) follows from the fact that NC1\textsf{NC}^{1} can simulate the Majority gate. The class NC is:

NC:=kNCk=kACk=kTCk.\textsf{NC}:=\bigcup_{k\in\mathbb{N}}\textsf{NC}^{k}=\bigcup_{k\in\mathbb{N}}\textsf{AC}^{k}=\bigcup_{k\in\mathbb{N}}\textsf{TC}^{k}.

It is known that NCP\textsf{NC}\subseteq\textsf{P}, and it is believed that this containment is strict.

Let d:d:\mathbb{N}\to\mathbb{N} be a function. The complexity class FO(d(n))\textsf{FO}(d(n)) is the set of languages decidable by uniform AC circuit families of depth O(d(n))O(d(n)) and polynomial size. The complexity class FOLL=FO(loglogn)\textsf{FOLL}=\textsf{FO}(\log\log n). It is known that AC0FOLLAC1\textsf{AC}^{0}\subsetneq\textsf{FOLL}\subsetneq\textsf{AC}^{1}, and it is open as to whether FOLL is contained in NL [BKLM01].

For a complexity class 𝒞\mathcal{C}, we define βi𝒞\beta_{i}\mathcal{C} to be the set of languages LL such that there exists an L𝒞L^{\prime}\in\mathcal{C} such that xLx\in L if and only if there exists yy of length at most O(logi|x|)O(\log^{i}|x|) such that (x,y)L(x,y)\in L^{\prime}. For any i,c0i,c\geq 0, βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c}) cannot compute Parity [CTW13].

Reductions. We will consider several notions of reducibility. Let L1,L2L_{1},L_{2} be languages. We say that L1L_{1} is many-one reducible to L2L_{2}, denoted L1mL2L_{1}\leq_{m}L_{2}, if there exists a computable function φ:{0,1}{0,1}\varphi:\{0,1\}^{*}\to\{0,1\}^{*} such that xL1φ(x)L2x\in L_{1}\iff\varphi(x)\in L_{2}. We will often ask that, for a given complexity 𝒞\mathcal{C}, φ\varphi is 𝒞\mathcal{C}-computable.

We now turn towards recalling the notion of a Turing reduction. Intuitively, we say that the language L1L_{1} is Turing-reducible to L2L_{2} if, given an algorithm to solve L2L_{2}, we can design an algorithm to solve L1L_{1}. This is made precise with oracle Turing machines, which we recall now. An oracle Turing machine is a Turing machine with a separate oracle tape and oracle query state. When the Turing machine enters the query state, it transitions to one of two specified states based on whether the string on the oracle tape belongs to the oracle. When the oracle is unspecified, we write MM^{\square}. When the oracle 𝒪\mathcal{O} is specified, we write M𝒪M^{\mathcal{O}}. We now make precise the notion of a Turing reduction.

A Turing reduction from L1L_{1} to L2L_{2} is an oracle Turing machine MM^{\square} such that ML2M^{L_{2}} decides L1L_{1}. If there exists a Turing reduction from L1L_{1} to L2L_{2}, we write L1TL2L_{1}\leq_{T}L_{2}. We will be particularly interested in Turing reductions from L1L_{1} to L2L_{2}, where ML2M^{L_{2}} runs in polynomial time.

We finally recall the notion of a truth-table reduction. Again let L1,L2L_{1},L_{2} be languages. We say that L1L_{1} is truth-table reducible to L2L_{2}, denoted L1ttL2L_{1}\leq_{tt}L_{2}, if there exists a computable function gg mapping an input XX to inputs Y1,,YkY_{1},\ldots,Y_{k}, and a computable function ff (called the truth-table predicate), which depends on XX and maps {0,1}k{0,1}\{0,1\}^{k}\to\{0,1\} such that XL1X\in L_{1} if and only if f(χL2(Y1),,χL2(Yk))=1f(\chi_{L_{2}}(Y_{1}),\ldots,\chi_{L_{2}}(Y_{k}))=1 (where χL2\chi_{L_{2}} is the characteristic function for L2L_{2}). The truth-table reduction is disjunctive if ff is the OR function. We say that the truth-table reduction is 𝒞\mathcal{C}-computable if both f,gf,g are 𝒞\mathcal{C}-computable.

It is known that:

L1mL2L1ttL2L1TL2,L_{1}\leq_{m}L_{2}\implies L_{1}\leq_{tt}L_{2}\implies L_{1}\leq_{T}L_{2},

and that this chain of implications holds even in the presence of resource constraints (e.g., if we were to insist that all steps be P-computable). Some care has to be taken to formulate a precise notion of a Turing reduction for circuit classes. Fortunately, we will not need the notion of a Turing reduction for circuit classes, and so we will not discuss this further.


Graph Isomorphism. We refer to GI as the decision problem Graph Isomorphism, which takes as input two graphs and asks whether there is an isomorphism between them. The complexity class GI is the set of decision problems that are polynomial-time Turing reducible to GI.

2.3 Color Refinement

Let GG be a graph. The Color Refinement (or 11-dimensional Weisfeiler–Leman) algorithm works by iteratively coloring the vertices of GG in an isomorphism invariant manner. The algorithm begins by assigning each vertex an initial color based on their degree; that is, χ0(v)=deg(v)\chi_{0}(v)=\text{deg}(v). Now for r0r\geq 0, we refine the coloring in the following manner:

χr+1(u)=(χr(u),{{χr(v):vN(u)}}),\chi_{r+1}(u)=(\chi_{r}(u),\{\!\{\chi_{r}(v):v\in N(u)\}\!\}),

where {{}}\{\!\{\cdot\}\!\} denotes a multiset. The count-free variant of Color Refinement works analogously, except the refinement step considers the set rather than multiset of colors at each round:

χr+1(u)=(χr(u),{χr(v):vN(u)}).\chi_{r+1}(u)=(\chi_{r}(u),\{\chi_{r}(v):v\in N(u)\}).

The algorithm terminates when the partition on the vertices induced by the coloring is not refined. We can use Color Refinement can as a non-isomorphism test by running the algorithm on G˙HG\dot{\cup}H and comparing the multiset of colors. Note that if at round rr, if the multiset of colors for GG differs from HH, then we can conclude that G≇HG\not\cong H. Each iteration of the standard counting Color Refinement can be implemented using a logspace uniform TC0\textsf{TC}^{0}-circuit, and each iteration of the count-free Color Refinement can be implemented using a logspace uniform AC0\textsf{AC}^{0}-circuit (c.f., the work of Grohe & Verbitsky [GV06] who observed this for kk-dimensional Weisfeiler–Leman where k2k\geq 2. The analogous parallel implementation holds for Color Refinement).

The individualize-and-refine paradigm works as follows. Let SS be a sequence of vertices in GG. We first assign each vertex in SS a unique color (so no two vertices in SS have the same color, and no vertex in V(G)SV(G)-S has the same color as any vertex in SS), and then run Weisfeiler–Leman. The refinement step is usually performed using Color Refinement (11-WL), but may be performed using higher-dimensional Weisfeiler–Leman. For the purposes of this paper, we will use the count-free variant of Color Refinement in the refinement step.

It is known that even higher-dimensional Weisfeiler–Leman fails to place GI into P [CFI92, NS18]. We refer to Sandra Kiefer’s dissertation [Kie19] for a thorough and current survey of Weisfeiler–Leman.

3 Latin Square Isotopy

In this section, we show that the Latin Square Isotopy problem is in β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL}. The key technique is to guess cube generating sequences AA and BB for the Latin square L1L_{1}, and cube generating sequences A,BA^{\prime},B^{\prime} for the Latin square L2L_{2}. We then (attempt to) construct appropriate bijections α,β:L1L2\alpha,\beta:L_{1}\to L_{2}, where α\alpha extends the map from AAA\to A^{\prime} and β\beta extends the map from BBB\to B^{\prime}. We can construct such bijections in FOLLL\textsf{FOLL}\cap\textsf{L} using the techniques from Chattopadhyay, Torán, and Wagner [CTW13].

The key step remains in checking whether the map sending the product of each pair (x,y)L1×L1(x,y)\in L_{1}\times L_{1}, xyα(x)β(y)xy\mapsto\alpha(x)\beta(y) a bijection. Wolf approaches this in the following manner. First, construct sets C={aibi:aiA,biB}C=\{a_{i}b_{i}:a_{i}\in A,b_{i}\in B\} and C={aibi:aiA,biB}C^{\prime}=\{a_{i}^{\prime}b_{i}^{\prime}:a_{i}^{\prime}\in A^{\prime},b_{i}^{\prime}\in B^{\prime}\}. Then check whether the map aibiaibia_{i}b_{i}\mapsto a_{i}^{\prime}b_{i}^{\prime} extends to a bijection γ:L1L2\gamma:L_{1}\to L_{2}. Finally, check whether α(x)β(y)=γ(xy)\alpha(x)\beta(y)=\gamma(xy) for all x,yL1x,y\in L_{1}. We note that Wolf is able to do this in NC2\textsf{NC}^{2} [Wol94]. If CC and CC^{\prime} are cube generating sequences, then we would be able to apply the technique of Chattophadyay, Torán, and Wagner [CTW13] to compute the bijection γ\gamma in FOLLL\textsf{FOLL}\cap\textsf{L}. However, CC and CC^{\prime} need not be cube generating sequences. As an example in the case of groups, suppose that B=A1B=A^{-1}. Then BB is a cube generating sequence. Furthermore, aibi=1a_{i}b_{i}=1 for each ii, and so CC has only the identity element. In general, we do not expect CC and CC^{\prime} to be cube generating sequences, even if they do generate L1L_{1} and L2L_{2} respectively.

Instead of extending Wolf’s technique, we extend Miller’s technique in his second proof that
Latin Square Isotopy is decidable in time nlog(n)+O(1)n^{\log(n)+O(1)}. Miller constructs the relation

R={(xy,α(x)β(y)):x,yL1},R=\{(xy,\alpha(x)\beta(y)):x,y\in L_{1}\},

and checks whether RR is a bijection. If RR is a bijection; then by construction, the triple (α,β,R)(\alpha,\beta,R) is an isotopy [Mil78, Theorem 2, Proof 2]. Using the fact that AA and BB are cube generating sequences, we can compute xx and yy in LFOLL\textsf{L}\cap\textsf{FOLL}. In the process of computing α\alpha and β\beta, we obtain a data structure which allows us to compute α(x)\alpha(x) and β(y)\beta(y) in LFOLL\textsf{L}\cap\textsf{FOLL}.

Theorem 3.1.

Latin Square Isotopy is in β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL}.

We begin with the following lemmas.

Lemma 3.2.

Let A,BA,B be finite sets of the same size, and let RA×BR\subseteq A\times B be a relation. We can decide whether RR is a well-defined surjection (and as |A|=|B||A|=|B|, consequently whether RR is a well-defined bijection) in AC0\textsf{AC}^{0}.

Proof.

For each bBb\in B, define the relation:

Y(b)=aA1(a,b)R.Y(b)=\bigvee_{a\in A}1_{(a,b)\in R}.

Observe that Y(b)Y(b) is AC0\textsf{AC}^{0}-computable. Now RR is surjective if and only if Y(b)=1Y(b)=1 for all bBb\in B, which is defined by the following condition:

φ:=bBY(b).\varphi:=\bigwedge_{b\in B}Y(b).

Observe that φ\varphi is AC0\textsf{AC}^{0} computable. ∎

Lemma 3.3.

Let L1L_{1} and L2L_{2} be Latin squares of order nn, and let k=4log(n)k=4\lceil\log(n)\rceil. Suppose that (g0,g1,,gk)(g_{0},g_{1},\ldots,g_{k}) and (h0,,hk)(h_{0},\ldots,h_{k}) are cube generating sequences for L1L_{1} and L2L_{2} respectively, with balanced parenthesization PP. Deciding whether the map gihig_{i}\mapsto h_{i} for all i{0,,k}i\in\{0,\ldots,k\} extends to a bijection is in LFOLL\textsf{L}\cap\textsf{FOLL}.

Proof.

For each (g,h)L1×L2(g,h)\in L_{1}\times L_{2}, define X(g,h)=1X(g,h)=1 if and only if there exists (ϵ1,,ϵk){0,1}k(\epsilon_{1},\ldots,\epsilon_{k})\in\{0,1\}^{k} such that:

g:=g0g1ϵ1gkϵk,\displaystyle g:=g_{0}g_{1}^{\epsilon_{1}}\cdots g_{k}^{\epsilon_{k}},
h:=h0h1ϵ1hkϵk.\displaystyle h:=h_{0}h_{1}^{\epsilon_{1}}\cdots h_{k}^{\epsilon_{k}}.

Chattophadyay, Torán, and Wagner showed that the cube words for gg and hh can be computed in LFOLL\textsf{L}\cap\textsf{FOLL} [CTW13], so X(g,h)X(g,h) is computable in LFOLL\textsf{L}\cap\textsf{FOLL}. We note that the FOLL bound follows from the fact that PP is a balanced parenthesization. As the two quasigroups are the same size, the map on the quasigroups induced by (g0,g1,,gk)(h0,,hk)(g_{0},g_{1},\ldots,g_{k})\mapsto(h_{0},\ldots,h_{k}) is a well-defined bijection iff the induced map is injective iff the induced map is surjective. So it now suffices to check whether the induced map is surjective. By Lemma 3.2, we may check whether the induced map is surjective in AC0\textsf{AC}^{0}. ∎

Proof of Theorem 3.1.

Let k=4log2(n)k=4\lceil\log_{2}(n)\rceil. We use 4k24k^{2} non-deterministic bits to guess cube generating sequences A,BLA,B\subseteq L and A,BLA^{\prime},B^{\prime}\subseteq L^{\prime}, where A={a0,a1,,ak}A=\{a_{0},a_{1},\ldots,a_{k}\}, B={b0,b1,,bk}B=\{b_{0},b_{1},\ldots,b_{k}\}, A={a0,a1,,ak}A^{\prime}=\{a_{0}^{\prime},a_{1}^{\prime},\ldots,a_{k}^{\prime}\}, and B={b0,b1,,bk}B^{\prime}=\{b_{0}^{\prime},b_{1}^{\prime},\ldots,b_{k}^{\prime}\}. We may then in LFOLL\textsf{L}\cap\textsf{FOLL} check the following.

  • We first check that the map aiaia_{i}\mapsto a_{i}^{\prime} extends to a bijection of A\langle A\rangle and A\langle A^{\prime}\rangle. In particular, we may check in LFOLL\textsf{L}\cap\textsf{FOLL} whether L1=Cube(A)L_{1}=\text{Cube}(A) and L2=Cube(A)L_{2}=\text{Cube}(A^{\prime}) (see [CTW13, Theorem 5]). The procedure in Lemma 3.3 that decides whether the map aiaia_{i}\mapsto a_{i}^{\prime} for all i{0,,k}i\in\{0,\ldots,k\} extends to a bijection L1L2L_{1}\to L_{2}, also explicitly computes a bijection φA:L1L2\varphi_{A}:L_{1}\to L_{2} if one exists.

  • We proceed analogously for the map bibib_{i}\mapsto b_{i}^{\prime} for all i{0,,k}i\in\{0,\ldots,k\}. In the case that L1=Cube(B)L_{1}=\text{Cube}(B) and L2=Cube(B)L_{2}=\text{Cube}(B^{\prime}), let φB:L1L2\varphi_{B}:L_{1}\to L_{2} be the bijection computed by Lemma 3.3.

Suppose now that the bijections φA,φB:L1L2\varphi_{A},\varphi_{B}:L_{1}\to L_{2} have been constructed. We now attempt to construct a relation φC:L1L2\varphi_{C}:L_{1}\to L_{2} in such a way that φC\varphi_{C} is a bijection if and only if (φA,φB,φC)(\varphi_{A},\varphi_{B},\varphi_{C}) is an isotopism. For each pair (,m)L1×L2(\ell,m)\in L_{1}\times L_{2}, define X(,m)=1X(\ell,m)=1 if and only if we define φC\varphi_{C} to map m\ell\mapsto m. For each (ϵ1,,ϵk),(ν1,,νk){0,1}k(\epsilon_{1},\ldots,\epsilon_{k}),(\nu_{1},\ldots,\nu_{k})\in\{0,1\}^{k}, we do the following:

  1. (a)

    Compute:

    g:=a0a1ϵ1akϵk,\displaystyle g:=a_{0}a_{1}^{\epsilon_{1}}\cdots a_{k}^{\epsilon_{k}},
    h:=b0b1ν1bkνk.\displaystyle h:=b_{0}b_{1}^{\nu_{1}}\cdots b_{k}^{\nu_{k}}.

    We note that computing gg and hh can be done in LFOLL\textsf{L}\cap\textsf{FOLL} (see Chattopadhyay, Toràn, and Wagner [CTW13]).

  2. (b)

    Compute :=gh\ell:=gh, φA(g)\varphi_{A}(g), φB(h)\varphi_{B}(h), and m:=φA(g)φB(h)m:=\varphi_{A}(g)\varphi_{B}(h). We set φC()=m\varphi_{C}(\ell)=m, which we indicate by defining X(,m)=1X(\ell,m)=1. The computations at this stage are computable using an AC0\textsf{AC}^{0} circuit.

  3. (c)

    It remains to check whether φC\varphi_{C} is a bijection. By Lemma 3.2, we may test whether φC\varphi_{C} is a bijection in AC0\textsf{AC}^{0}.

Now φC\varphi_{C} was constructed so that (φA,φB,φC)(\varphi_{A},\varphi_{B},\varphi_{C}) satisfies the homotopy condition, so, as they are also bijective, they are an isotopy. Thus, checking whether L1L_{1} and L2L_{2} are isotopic is in β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL}. ∎

Now for any i,c0i,c\geq 0, we have that βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c}) cannot compute Parity [CTW13]. As Latin Square Isotopy belongs to β2FOLL\beta_{2}\textsf{FOLL}, we obtain the following corollary.

Corollary 3.4.

For any i,c0i,c\geq 0, GI is not βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c})-reducible to Latin Square Isotopy.

Miller [Mil78] showed that isomorphism testing of Latin square graphs is polynomial-time reducible to the Latin square isotopy problem. Wolf [Wol94] improved this bound, showing that isomorphism testing of Latin square graphs is NC1\textsf{NC}^{1}-reducible to testing for Latin square isotopy. We recall Wolf’s result below.

Lemma 3.5 ([Wol94, Lemma 4.11]).

Let GG be a Latin square graph derived from a Latin square of size nn. We can retrieve a Latin square from GG with a polynomial-sized NC circuit with O(logn)O(\log n) depth.

Remark 3.6.

The statement of Lemma 4.11 in Wolf actually claims a depth of O(log2n)O(\log^{2}n). However, in the proof of Lemma 4.11, Wolf shows that only O(logn)O(\log n) depth is needed [Wol94].

Now by Remark 1.6, we can place the Latin squares into their main isotopy class in AC0\textsf{AC}^{0}. In light of this, Wolf’s result, Theorem 3.1, and Bruck’s result that for n>23n>23, a pseudo-Latin square graph is a Latin square graph [Bru63], we obtain the following corollary.

Corollary 3.7.

Isomorphism of pseudo-Latin square graphs can be decided in β2L\beta_{2}\textsf{L}.

Remark 3.8.

To the best of the author’s knowledge, the largest pseudo-Latin square graph that is not a Latin square graph is the Hall–Janko graph, where n=10n=10 (that is, the Hall–Janko graph has n2=100n^{2}=100 vertices) [Suz69, Nak06]. The author would be grateful for references to larger examples.

We next show that the reduction from [Wol94, Lemma 4.11], which effectively parallelizes the reduction found in [Mil78], can be implemented in AC0\textsf{AC}^{0}. Furthermore, we can place the recovered Latin square into its main isotopy class in AC0\textsf{AC}^{0} (see Remark 1.6). It follows that Latin Square Graph Isomorphism is also in β2FOLL\beta_{2}\textsf{FOLL}. As β2FOLL\beta_{2}\textsf{FOLL} cannot compute Parity, we obtain that for any i,c0i,c\geq 0, GI is not βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c})-reducible to Latin Square Graph Isomorphism.

Lemma 3.9.

Let GG be a Latin square graph obtained from a Latin square of order nn. We can recover a Latin square from GG using an AC0\textsf{AC}^{0} circuit.

Proof.

We carefully analyze the reduction from [Mil78, Wol94]. By construction, two vertices uu and vv in GG are adjacent precisely if uu and vv correspond to elements in the Latin square that are either in the same row, the same column, or have the same value. As a result, each row, each column, and the nodes corresponding to a fixed given value all induce cliques of size nn. The algorithm effectively identifies these cliques and their relations to the other cliques in order to recover a Latin square.

Denote L=(ij)1i,jnL=(\ell_{ij})_{1\leq i,j\leq n} to be the n×nn\times n matrix, which our algorithm will populate with values for the Latin square. We proceed as follows.

  1. 1.

    We begin by selecting two adjacent vertices x1x_{1} and x2x_{2}. Without loss of generality, we may assume x1x_{1} and x2x_{2} belong to the same row of the Latin square. We next find the nn vertices v1,,vnv_{1},\ldots,v_{n} adjacent to both x1x_{1} and x2x_{2}. Precisely, for a vertex vv, we have the indicator

    X(v)=E(x1,v)E(x2,v),X(v)=E(x_{1},v)\land E(x_{2},v),

    where E(u,v)=1E(u,v)=1 precisely if uu and vv are adjacent. Exactly nn indicators will evaluate to 11. All but two of these nodes form a clique of size nn with x1x_{1} and x2x_{2}. As v1,,vnv_{1},\ldots,v_{n} are adjacent to x1x_{1} and x2x_{2}, it suffices to check which n2n-2 vertices of v1,,vnv_{1},\ldots,v_{n} form a clique of size (n2)(n-2). For a given set S([n]n2)S\in\binom{[n]}{n-2}, it suffices to check:

    i,jSijE(vi,vj),\bigwedge_{\begin{subarray}{c}i,j\in S\\ i\neq j\end{subarray}}E(v_{i},v_{j}),

    which is AC0\textsf{AC}^{0}-computable. There are (nn2)Θ(n2)\binom{n}{n-2}\in\Theta(n^{2}) such sets to check. Thus, identifying the clique is AC0\textsf{AC}^{0}-computable.

    In parallel, we label the vertices of the clique as x3,,xnx_{3},\ldots,x_{n}. One node not adjacent to any of x3,,xnx_{3},\ldots,x_{n} is labeled y2y_{2}.

  2. 2.

    We associate 1j\ell_{1j} with xjx_{j}. Precisely, we set (in parallel) 1j=j\ell_{1j}=j for each j[n]j\in[n]. This step is AC0\textsf{AC}^{0}-computable.

  3. 3.

    We next find the nn-clique associated with x1x_{1} and y2y_{2}, and label the additional vertices as y3,,yny_{3},\ldots,y_{n}. By similar argument as in Step 1, this step is AC0\textsf{AC}^{0}-computable. Here, we view x1,y2,y3,,ynx_{1},y_{2},y_{3},\ldots,y_{n} as the first column of the Latin square.

  4. 4.

    For each 3in3\leq i\leq n, there exists a 3jn3\leq j\leq n such that xix_{i} and yjy_{j} are adjacent. In particular, as xix_{i} and yjy_{j} are neither in the same row nor the same column, it must be the case that xix_{i} and yjy_{j} correspond to elements in the Latin square with the same value. It follows that our choice of jj is in fact unique. We reorder y3,,yny_{3},\ldots,y_{n} so that xix_{i} is adjacent to yiy_{i}. This step is AC0\textsf{AC}^{0}-computable.

  5. 5.

    For 2jn2\leq j\leq n, we associate j1\ell_{j1} with yjy_{j}. Precisely, we set (in parallel), j1=j\ell_{j1}=j for each 2jn2\leq j\leq n. This step is AC0\textsf{AC}^{0}-computable.

  6. 6.

    For each of the remaining (n1)2(n-1)^{2} nodes zz, we do the following in parallel:

    1. (a)

      If zz is adjacent to x1x_{1}, then the edge {x1,z}\{x_{1},z\} is a value edge (as zz is not in the same row or column as x1x_{1}). So there exist unique i,j>1i,j>1 such that zz is adjacent to xjx_{j} and yiy_{i}. In this case, we set ij=1\ell_{ij}=1. This case is AC0\textsf{AC}^{0}-computable.

    2. (b)

      Suppose that zz is not adjacent to x1x_{1}. As each row, each column, and each value induce an nn-clique, there exist unique i,j,k[n]i,j,k\in[n] such that zz is adjacent to xj,yi,xkx_{j},y_{i},x_{k}, and yky_{k}. Unless i=j=ki=j=k, we set ij=k\ell_{ij}=k. If i=j=ki=j=k, we do nothing at this step and defer to step 7. This case is AC0\textsf{AC}^{0}-computable.

  7. 7.

    We note that step 6b does not account for diagonal entries where ii=i\ell_{ii}=i. To this end, we do the following. For each i2i\geq 2, we (in parallel) set ii\ell_{ii} to be the value that does not appear in row ii. This step is AC0\textsf{AC}^{0}-computable.

As we have a finite number of steps, each of which are AC0\textsf{AC}^{0}-computable, it follows that we may recover a Latin square from GG using an AC0\textsf{AC}^{0} circuit. ∎

Proposition 3.10.

Isomorphism testing of pseudo-Latin square graphs is in β2FOLL\beta_{2}\textsf{FOLL}. In particular, for any i,c0i,c\geq 0, GI is not βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c})-reducible to isomorphism testing of pseudo-Latin square graphs.

Proof.

We may handle the cases when n23n\leq 23 in AC0\textsf{AC}^{0}. So suppose n23n\geq 23, and let GG and HH be pseudo-Latin square graphs. As n23n\geq 23, we have by Bruck that GG and HH are Latin square graphs [Bru63]. By Lemma 3.9, we may in AC0\textsf{AC}^{0} recover canonical Latin squares LGL_{G} and LHL_{H} corresponding to GG and HH. Now by [Mil78, Lemma 3], GHG\cong H if and only if LGL_{G} and LHL_{H} are main class isotopic. By Remark 1.6, we may place LGL_{G} (respectively, LHL_{H}) into a normal form corresponding to its main isotopy class in AC0\textsf{AC}^{0}. By Theorem 3.1, we can test whether LGL_{G} and LHL_{H} are isotopic in β2FOLL\beta_{2}\textsf{FOLL}. Chattopadhyay, Torán, and Wagner showed that β2FOLL\beta_{2}\textsf{FOLL} cannot compute Parity [CTW13]. The result now follows. ∎

4 Isomorphism Testing of Steiner Designs

In this section, we show that for any i,c0i,c\geq 0, GI is not βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c})-reducible to isomorphism testing of several families of Steiner designs.

4.1 Nets

In this section, we show that for any i,c0i,c\geq 0, GI is not βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c})-reducible to isomorphism testing of nets or the corresponding strongly regular graphs. We note that projective and affine planes are special cases of nets.

Theorem 4.1.

Deciding whether two kk-nets are isomorphic is in β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL}.

Proof.

For k=0,1,2k=0,1,2, the pair (k,n)(k,n) determines the net uniquely, and so deciding isomorphism is trivial in these cases. So assume that n+1k3n+1\geq k\geq 3. Let 𝒩1(n,k),𝒩2(n,k)\mathcal{N}_{1}(n,k),\mathcal{N}_{2}(n,k) be nets. We now start by guessing three non-parallel lines a,b,cL(𝒩1)\ell_{a},\ell_{b},\ell_{c}\in L(\mathcal{N}_{1}) and three non-parallel lines a,b,cL(𝒩2)\ell_{a}^{\prime},\ell_{b}^{\prime},\ell_{c}^{\prime}\in L(\mathcal{N}_{2}). By definition, no two lines in the same parallel class share any points in common, and two lines in different classes share exactly one point in common. As there are knkn lines in 𝒩i\mathcal{N}_{i}, we only require O(log(kn))O(\log(kn)) bits to identify a given line. As kn+1k\leq n+1, in fact only O(log(n))O(\log(n)) bits are required.

We may identify whether two lines are disjoint in AC0\textsf{AC}^{0}. In particular, we may check in AC0\textsf{AC}^{0} whether a,b\ell_{a},\ell_{b}, and c\ell_{c} (respectively, a,b\ell_{a}^{\prime},\ell_{b}^{\prime}, and c\ell_{c}^{\prime}) belong to different parallel classes. Suppose now that a,b\ell_{a},\ell_{b}, and c\ell_{c} (respectively, a,b\ell_{a}^{\prime},\ell_{b}^{\prime}, and c\ell_{c}^{\prime}) belong to different parallel classes. As there are (kn2)\binom{kn}{2} such pairs to check in a given 𝒩i\mathcal{N}_{i}, we may identify in AC0\textsf{AC}^{0} the lines belonging to the three parallel classes a,b,ca,b,c in each 𝒩i\mathcal{N}_{i}.

Now our three parallel classes determine a net X1(n,3)X_{1}(n,3) in 𝒩1\mathcal{N}_{1} and a net X2(n,3)X_{2}(n,3) in 𝒩2\mathcal{N}_{2}. We note that a net of degree 33 identifies a quasigroup (Latin square) up to isomorphism. As Quasigroup Isomorphism is in β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL} [CTW13, Theorem 3.4], we may now in β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL} decide whether X1X2X_{1}\cong X_{2}. We note that [CTW13, Theorem 3.4] actually yields an isomorphism φ:X1X2\varphi:X_{1}\to X_{2}. We may now in LFOLL\textsf{L}\cap\textsf{FOLL} check whether φ\varphi extends to an isomorphism of 𝒩1\mathcal{N}_{1} and 𝒩2\mathcal{N}_{2}. The result follows. ∎

Corollary 4.2.

For any i,c0i,c\geq 0, GI is not βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c})-reducible to isomorphism testing of two kk-nets.

Miller previously showed that isomorphism testing of nets and the corresponding net graphs are equivalent under polynomial-time reductions when n>(k1)2n>(k-1)^{2} [Mil78, Theorem 8]. A closer analysis shows that this equivalence holds under TC0\textsf{TC}^{0}-reductions in general; and that when kk is bounded, the equivalence holds under AC0\textsf{AC}^{0}-reductions.

Lemma 4.3.

Suppose that G(V,E)G(V,E) is a kk-net graph of order nn and n>(k1)2n>(k-1)^{2}. We can reconstruct the net 𝒩(n,k)\mathcal{N}(n,k) associated with GG in TC0\textsf{TC}^{0}. If kk is bounded, we reconstruct the net 𝒩(n,k)\mathcal{N}(n,k) associated with GG in AC0\textsf{AC}^{0}.

Proof.

Suppose that x1,x2V(G)x_{1},x_{2}\in V(G) are adjacent. As there are knkn lines, it suffices to show that in AC0\textsf{AC}^{0}, we can identify the remaining vertices of V(G)V(G) that are on the same line as x1,x2x_{1},x_{2}. We first note that we can, in AC0\textsf{AC}^{0}, identify the vertices adjacent to both x1x_{1} and x2x_{2}.

Let HH be the subgraph induced by these vertices together with x1x_{1} and x2x_{2}. We note that HH contains the maximum clique (of nn vertices) containing both x1x_{1} and x2x_{2}. The vertices of this clique have degree n2n-2. As two adjacent vertices have (n2)+(k1)(k2)(n-2)+(k-1)(k-2) neighbors in common, there are (k1)(k2)(k-1)(k-2) vertices of HH that do not belong to this clique. Recall that a net has kk parallel classes, and any two lines in a given parallel class have empty intersection. Furthermore, two lines from different parallel classes share exactly one point in common. Thus, each nonclique vertex of HH is adjacent to exactly (k1)(k-1) elements of the clique. Thus, each nonclique vertex of HH has degree at most:

(k1)+(k1)(k2)1=(k1)21.(k-1)+(k-1)(k-2)-1=(k-1)^{2}-1.

In general, using the difference in degrees, we may identify the clique and nonclique vertices in TC0\textsf{TC}^{0}. If kk is bounded, then we may identify the clique and nonclique vertices in AC0\textsf{AC}^{0}. The result follows. ∎

We combine Lemma 4.3 with Bruck’s result [Bru63, Theorem 7] that for fixed kk, pseudo-net graphs with sufficiently many vertices are net graphs to obtain the following.

Corollary 4.4.

For fixed kk, isomorphism testing of pseudo-net graphs of order nn and degree kk is in β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL}. In particular, for any i,c0i,c\geq 0, GI is not βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c})-reducible to isomorphism testing of pseudo-net graphs of order nn and degree kk.

Remark 4.5.

Let p(x):=(1/2)x4+x3+x2+(3/2)xp(x):=(1/2)x^{4}+x^{3}+x^{2}+(3/2)x. Bruck [Bru63, Theorem 7] showed that if n>p(k1)n>p(k-1) and k>1k>1, then a pseudo-net graph of order nn and degree kk is a net graph.

4.2 Steiner Triple Systems

Using the standard connection between Steiner triple systems and quasigroups in tandem with [CTW13], we observe the following.

Observation 4.6.

Deciding whether two Steiner triple systems are isomorphic is in β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL}.

Proof.

Let 𝒮\mathcal{S} be a Steiner triple system of order nn. We define a quasigroup QQ on the set [n][n], with the multiplication operation satisfying the following: (i) for each x[n]x\in[n], define x2=xx^{2}=x, and (ii) for each block {a,b,c}\{a,b,c\}, define ab=cab=c (note that as the blocks are unordered, all such products ba=c,ac=ca=b,bc=cb=aba=c,ac=ca=b,bc=cb=a are required). The Steiner triple system determines QQ up to isomorphism. In particular, this construction is AC0\textsf{AC}^{0}-computable. As Quasigroup Isomorphism belongs to β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL} [CTW13], it follows that deciding whether two Steiner triple systems are isomorphic is in β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL}. ∎

Proposition 4.7.

Let GG be a block-incidence graph on nn vertices derived from a Steiner 22-design, in which each block has kk points and n2>(k1)2\sqrt{n}-2>(k-1)^{2}. We can reconstruct the Steiner 22-design in TC0\textsf{TC}^{0}. Furthermore, if kk is bounded, then we may reconstruct the Steiner 22-design in AC0\textsf{AC}^{0}.

Proof.

We closely analyze the proof of [Spi96, Proposition 10]. We note that nn is the number of blocks in the Steiner design. Let vv be the number of points in the Steiner design, and let R=(v1)/(k1)R=(v-1)/(k-1) be the number of blocks containing a given point. Let B1,B2B_{1},B_{2} be two blocks that intersect uniquely at the point pp. There are R2+(k1)2R-2+(k-1)^{2} blocks that intersect both B1B_{1} and B2B_{2}. We note that R2R-2 of these blocks also go through pp, and the remaining (k1)2(k-1)^{2} blocks go through points other than pp.

We note that pp corresponds to the edge {B1,B2}E(G)\{B_{1},B_{2}\}\in E(G). The RR blocks that intersect pp (including B1,B2B_{1},B_{2}) form a clique. Furthermore, these R2R-2 blocks intersecting B1,B2B_{1},B_{2} at pp do not intersect with the remaining (k1)2(k-1)^{2} blocks that intersect with B1,B2B_{1},B_{2} in points other than pp. Let NN be the set of mutual neighbors for B1,B2B_{1},B_{2}. Now if R2>(k1)2R-2>(k-1)^{2}, the RR-clique is distinguished from the remaining (k1)2(k-1)^{2} blocks that don’t contain pp by their degrees in the induced subgraph G[N]G[N]. We may distinguish these vertices in TC0\textsf{TC}^{0} in general. If kk is bounded, then we may distinguish these vertices in AC0\textsf{AC}^{0}.

The fact that R>nR>\sqrt{n} was established in [Spi96, Proposition 10]. ∎

We combine Proposition 4.7 with Bose’s result that pseudo-STS graphs with strictly more than 6767 vertices are STS graphs [Bos63] to obtain the following.

Corollary 4.8.

Deciding whether two pseudo-STS graphs are isomorphic is in β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL}. In particular, for any i,c0i,c\geq 0, GI is not βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c})-reducible to isomorphism testing of pseudo-STS graphs.

Remark 4.9.

To the best of the author’s knowledge, it remains open whether the bound of 6767 due to Bose [Bos63] can be decreased.

4.3 Reduction to Steiner 22-Designs

Babai & Wilmes [BW13] previously exhibited a reduction from isomorphism testing of Steiner tt-designs to the case of Steiner 22-designs. A careful analysis shows that their reduction is β2AC0\beta_{2}\textsf{AC}^{0}-computable. As a corollary, we obtain that GI is not AC0\textsf{AC}^{0}-reducible to isomorphism testing of Steiner (t,t+1)(t,t+1)-designs.

Observation 4.10 (c.f. [BW13]).

If isomorphism testing of Steiner 22-designs belongs to β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL}, then testing isomorphism of Steiner tt-designs also belongs to β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL}.

Proof.

Let 𝒟1=(X1,1,I1)\mathcal{D}_{1}=(X_{1},\mathcal{B}_{1},I_{1}) and 𝒟2=(X2,2,I2)\mathcal{D}_{2}=(X_{2},\mathcal{B}_{2},I_{2}) be Steiner (t,k,v)(t,k,v)-designs. We begin by guessing subsequences a1,,at2X1a_{1},\ldots,a_{t-2}\in X_{1} and b1,,bt2X2b_{1},\ldots,b_{t-2}\in X_{2}. While we think of a1,,at2a_{1},\ldots,a_{t-2} and b1,,bt2b_{1},\ldots,b_{t-2} as determining subsets A={a1,,at2}X1A=\{a_{1},\ldots,a_{t-2}\}\subseteq X_{1} and B={b1,,bt2}X2B=\{b_{1},\ldots,b_{t-2}\}\subseteq X_{2}, we stress that we guess both the elements and an ordering. As tO(logn)t\in O(\log n) (see the proof of [BW13, Theorem 30], in which Babai & Wilmes cite [RCW75]), guessing AA and BB requires O(log2n)O(\log^{2}n) bits. We may write down the derived designs 𝒟1(A)\mathcal{D}_{1}(A) and 𝒟2(B)\mathcal{D}_{2}(B) in AC0\textsf{AC}^{0} (see Section 2 for the definition of a derived design).

Suppose first that 𝒟1𝒟2\mathcal{D}_{1}\cong\mathcal{D}_{2}. Let φ:X1X2\varphi:X_{1}\to X_{2} be an isomorphism. Then for any AX1A\subseteq X_{1} of size t2t-2, the derived designs 𝒟1(A)\mathcal{D}_{1}(A) and 𝒟2(φ(A))\mathcal{D}_{2}(\varphi(A)) are isomorphic.

Conversely, let ψ:X1AX2B\psi:X_{1}\setminus A\to X_{2}\setminus B be an isomorphism of the derived designs 𝒟1(A)𝒟2(B)\mathcal{D}_{1}(A)\cong\mathcal{D}_{2}(B). Define:

ψ^(x)={ψ(x):xA,bi:x=aiA.\widehat{\psi}(x)=\begin{cases}\psi(x)&:x\not\in A,\\ b_{i}&:x=a_{i}\in A.\end{cases}

We may easily check in AC0\textsf{AC}^{0} whether ψ^\widehat{\psi} is an isomorphism of 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2}. In particular, if isomorphism testing of Steiner 22-designs belongs to β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL}, we may use the added non-determinism to guess an isomorphism of 𝒟1(A)𝒟1(B)\mathcal{D}_{1}(A)\cong\mathcal{D}_{1}(B) that lifts to an isomorphism of 𝒟1𝒟2\mathcal{D}_{1}\cong\mathcal{D}_{2}. In total, we are still using at most O(log2n)O(\log^{2}n) non-deterministic bits. ∎

We obtain the following corollaries.

Corollary 4.11.

The problem of deciding whether two Steiner (t,t+1)(t,t+1)-designs are isomorphic is β2AC0\beta_{2}\textsf{AC}^{0}-reducible to the problem of finding an isomorphism of two Steiner triple systems.

Now deciding isomorphism testing of Steiner triple systems is AC0\textsf{AC}^{0}-reducible to Quasigroup Isomorphism (see Theorem 4.6). Furthermore, Chattopadhyay, Torán, and Wagner solved the search version of Quasigroup Isomorphism; that is, their β2Lβ2FOLL\beta_{2}\textsf{L}\cap\beta_{2}\textsf{FOLL} procedure for Quasigroup Isomorphism returns an isomorphism of the two quasigroups whenever an isomorphism exists. So in particular GI is not AC0\textsf{AC}^{0}-reducible to the search version of Quasigroup Isomorphism [CTW13]. We may use the isomorphism for the quasigroups to construct an isomorphism of the Steiner triple systems, which may in turn be used to construct an isomorphism of the two Steiner (t,t+1)(t,t+1)-designs. We summarize this observation with the following corollaries.

Corollary 4.12.

The problem of deciding whether two Steiner (t,t+1)(t,t+1)-designs are isomorphic is β2AC0\beta_{2}\textsf{AC}^{0}-reducible to the problem of finding an isomorphism of two quasigroups.

Corollary 4.13.

For any i,c0i,c\geq 0, GI is not βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c})-reducible to the problem of deciding whether two Steiner (t,t+1)(t,t+1)-designs are isomorphic.

5 Conference Graphs

In this section, we consider the complexity of identifying conference graphs. A conference graph is a strongly regular graph with parameters (n,(n1)/2,(n5)/4,(n1)/4)(n,(n-1)/2,(n-5)/4,(n-1)/4). For a graph GG, a distinguishing set SS is a subset of V(G)V(G) such that for every u,vV(G)u,v\in V(G), we have that uS,vSu\in S,v\in S, or N(u)SN(v)SN(u)\cap S\neq N(v)\cap S. We have the following observation.

Observation 5.1.

Let GG be a graph, and let SS be a distinguishing set. After individualizing each vertex in SS, we have that 22 rounds of the count-free Color Refinement algorithm will assign each vertex in GG a unique color.

Babai [Bab80, Lemma 3.2] (take k=(n5)/4k=(n-5)/4) showed that conference graphs admit distinguishing sets of size O(logn)O(\log n) (in fact, almost all such subsets of this size are distinguishing). As a consequence of this and Observation 5.1, we obtain the following.

Theorem 5.2.

Let GG be a conference graph, and let HH be arbitrary. We can decide isomorphism between GG and HH in β2AC0\beta_{2}\textsf{AC}^{0}.

Proof.

We use m:=O(log2n)m:=O(\log^{2}n) non-deterministic bits to guess a sequence S=(s1,,sm)S=(s_{1},\ldots,s_{m}) for GG– while it would suffice for {s1,,sm}\{s_{1},\ldots,s_{m}\} to be a distinguishing set for GG, we only need that after individualizing the elements of SS, two rounds of count-free Color Refinement will assign each vertex in GG a unique color. There may be non-distinguishing sets that acccomplish this. Let S=(s1,,sm)S^{\prime}=(s_{1}^{\prime},\ldots,s_{m}^{\prime}) be the vertices of HH that were guessed. We now individualize SS and SS^{\prime} so that sisis_{i}\mapsto s_{i}^{\prime} get the same color, and si,sjs_{i},s_{j} get different colors whenever iji\neq j. The individualization step is AC0\textsf{AC}^{0}-computable. We now run two rounds of count-free Color Refinement, which is AC0\textsf{AC}^{0}-computable (c.f., [GV06] and the discussion in Section 2). Lastly, we check that each vertex in G,HG,H has a unique color, and whether the map induced by the colors is an isomorphism. This last step is AC0\textsf{AC}^{0}-computable. The result now follows. ∎

In light of the previous work of Chattopadhyay, Torán, & Wagner [CTW13], we obtain the following corollary.

Corollary 5.3.

For any i,c0i,c\geq 0, there is no βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c})-computable reduction from GI to isomorphism testing of conference graphs.

6 Conclusion

We showed that for any i,c0i,c\geq 0, GI is not βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c})-reducible to several isomorphism problems characterized by the generator enumeration technique, including Latin Square Isotopy, isomorphism testing of nets (which includes affine and projective planes), isomorphism testing of Steiner (t,t+1)(t,t+1)-designs, and isomorphism testing of conference graphs. As a corollary, we obtained that GI is not βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c})-reducible to isomorphism testing of Latin square graphs, kk-net graphs (for fixed kk), and the block-intersection graphs arising from Steiner triple systems. Our work leaves several directions for further research.

In Proposition 4.7, we showed that a Steiner 22-design can be recovered from its block-incidence graph in AC0\textsf{AC}^{0} when the block size is bounded. Otherwise, the procedure is TC0\textsf{TC}^{0}-computable, as we need to distinguish vertices by their degrees. As a step towards ruling out βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c}) reductions from GI, it would be of interest to show that we can recover a Steiner 22-design from its block-incidence graph in a complexity class that cannot compute Parity. To this end, we ask the following.

Question 6.1.

Can Steiner 22-designs of unbounded block size be recovered from their block-incidence graphs in AC0\textsf{AC}^{0}?

It would also be of interest to find additional families of Steiner 22-designs where the isomorphism problem belongs to βiFO((loglogn)c)\beta_{i}\textsf{FO}((\log\log n)^{c}). As a starting point, we ask the following.

Question 6.2.

For Steiner 22-designs with bounded block size, can we decide isomorphism in β2FOLL\beta_{2}\textsf{FOLL}?

While a Steiner 22-design 𝒟\mathcal{D} admits generating set SS of O(logv)O(\log v) points [BL83], we have that AutS(𝒟)\operatorname{Aut}_{S}(\mathcal{D}) may in general be non-trivial. This is the key barrier for the techniques here.

Babai & Wilmes [BW13] and Chen, Sun, & Teng [CST13] independently showed that Steiner 22-designs admit a set of O(logv)O(\log v) points where, once individualized, the Color Refinement algorithm assigns a unique color to each point. A priori, it seems plausible that only polyloglogv\operatorname{poly}\log\log v iterations are required. However, Color Refinement takes into account the multiset of colors, and so each round can be implemented with a TC0\textsf{TC}^{0}-circuit. In particular, Babai & Wilmes rely crucially on counting [BW13, Target 2]. So we ask the following.

Question 6.3.

Does there exist an absolute constant cc, such that Steiner 22-designs admit a set SS of O(logcv)O(\log^{c}v) points where, after individuaizing SS, the coloring from polyloglogn\operatorname{poly}\log\log n rounds of count-free Color Refinement assigns each point a unique color?

In the remark following the proof of [Bab80, Lemma 3.2], Babai outines a deterministic proof that leverages the greedy set cover algorithm (see [Lov75]) to obtain a distinguishing set of the prescribed size. Now suppose that GG and HH are isomorphic, and the greedy set cover algorithm returns distinguishing sequences SS and SS^{\prime} of the same size for G,HG,H respectively. A priori, SS and SS^{\prime} need not be canonical in the sense that there need not be an isomorphism φ:V(G)V(H)\varphi:V(G)\to V(H) mapping φ(S)=S\varphi(S)=S^{\prime}. Thus, we ask the following.

Question 6.4.

Is it possible to deterministically construct a canonical distinguishing set of the size prescribed by [Bab80, Lemma 3.2] for a graph in polynomial time?

An answer of yes to Question 6.4 in tandem with the work in Section 5 would immediately yield a polynomial-time isomorphism test for conference graphs. Babai’s work [Bab80] implies an nO(logn)n^{O(\log n)}-time algorithm, and to the best of my knowledge, no further improvements have been made to the runtime.

Alternatively, we ask the following.

Question 6.5.

Let GG be a conference graph. Does there exists a set of vertices SS of size O(1)O(1) such that, after individualizing SS and running Color Refinement, each vertex of GG would receive a unique color?

An answer of yes to Question 6.5 would also yield a polynomial-time isomorphism test for conference graphs, using the individualize-and-refine paradigm. Even finding a such a set SS of size o(logn)o(\log n) would be a major advance, as it would yield an no(logn)n^{o(\log n)}-time isomorphism test.

References

  • [AB09] Sanjeev Arora and Boaz Barak. Computational Complexity: A Modern Approach. 01 2009. doi:10.1017/CBO9780511804090.
  • [AK06] V. Arvind and Piyush P. Kurur. Graph isomorphism is in SPP. Information and Computation, 204(5):835–852, 2006. doi:10.1016/j.ic.2006.02.002.
  • [Alb43] A. A. Albert. Quasigroups. I. Transactions of the American Mathematical Society, 54(3):507–519, 1943. doi:10.2307/1990259.
  • [Bab80] László Babai. On the complexity of canonical labeling of strongly regular graphs. SIAM Journal on Computing, 9(1):212–216, 1980. doi:10.1137/0209018.
  • [Bab81] Laszlo Babai. On the order of uniprimitive permutation groups. Annals of Mathematics, 113(3):553–568, 1981. doi:10.2307/2006997.
  • [Bab14] László Babai. On the automorphism groups of strongly regular graphs I. In Proceedings of the 5th Conference on Innovations in Theoretical Computer Science, ITCS ’14, page 359–368, New York, NY, USA, 2014. Association for Computing Machinery. doi:10.1145/2554797.2554830.
  • [Bab16] László Babai. Graph isomorphism in quasipolynomial time [extended abstract]. In STOC’16—Proceedings of the 48th Annual ACM SIGACT Symposium on Theory of Computing, pages 684–697. ACM, New York, 2016. Preprint of full version at arXiv:1512.03547v2 [cs.DS]. doi:10.1145/2897518.2897542.
  • [BE99] Hans Ulrich Besche and Bettina Eick. Construction of finite groups. J. Symb. Comput., 27(4):387–404, 1999. doi:10.1006/jsco.1998.0258.
  • [BEO02] Hans Ulrich Besche, Bettina Eick, and E.A. O’Brien. A millennium project: Constructing small groups. Intern. J. Alg. and Comput, 12:623–644, 2002. doi:10.1142/S0218196702001115.
  • [BH92] Harry Buhrman and Steven Homer. Superpolynomial circuits, almost sparse oracles and the exponential hierarchy. In R. K. Shyamasundar, editor, Foundations of Software Technology and Theoretical Computer Science, 12th Conference, New Delhi, India, December 18-20, 1992, Proceedings, volume 652 of Lecture Notes in Computer Science, pages 116–127. Springer, 1992. doi:10.1007/3-540-56287-7\_99.
  • [BKL83] L. Babai, W. M. Kantor, and E. M. Luks. Computational complexity and the classification of finite simple groups. In 24th Annual Symposium on Foundations of Computer Science (sfcs 1983), pages 162–171, 1983. doi:10.1109/SFCS.1983.10.
  • [BKLM01] David A. Mix Barrington, Peter Kadau, Klaus-Jörn Lange, and Pierre McKenzie. On the complexity of some problems on groups input as multiplication tables. J. Comput. Syst. Sci., 63(2):186–200, 2001. doi:10.1006/jcss.2001.1764.
  • [BL83] László Babai and Eugene M. Luks. Canonical labeling of graphs. In Proceedings of the Fifteenth Annual ACM Symposium on Theory of Computing, STOC ’83, page 171–183, New York, NY, USA, 1983. Association for Computing Machinery. doi:10.1145/800061.808746.
  • [Bos63] R. C. Bose. Strongly regular graphs, partial geometries and partially balanced designs. Pacific Journal of Mathematics, 13(2):389 – 419, 1963. doi:pjm/1103035734.
  • [Bru63] R. H. Bruck. Finite nets. II. Uniqueness and imbedding. Pacific Journal of Mathematics, 13(2):421 – 457, 1963. doi:10.2140/pjm.1963.13.421.
  • [BW13] László Babai and John Wilmes. Quasipolynomial-time canonical form for Steiner designs. In STOC 2013, pages 261–270, New York, NY, USA, 2013. Association for Computing Machinery. doi:10.1145/2488608.2488642.
  • [CC81] Marlene J. Colbourn and Charles J. Colbourn. Concerning the complexity of deciding isomorphism of block designs. Discrete Applied Mathematics, 3(3):155–162, July 1981. doi:10.1016/0166-218X(81)90012-3.
  • [CFI92] Jin-Yi Cai, Martin Fürer, and Neil Immerman. An optimal lower bound on the number of variables for graph identification. Combinatorica, 12(4):389–410, 1992. Originally appeared in SFCS ’89. doi:10.1007/BF01305232.
  • [CG70] D. G. Corneil and C. C. Gotlieb. An efficient algorithm for graph isomorphism. J. ACM, 17(1):51–64, January 1970. doi:10.1145/321556.321562.
  • [CH03] John J. Cannon and Derek F. Holt. Automorphism group computation and isomorphism testing in finite groups. J. Symb. Comput., 35:241–267, March 2003. doi:10.1016/S0747-7171(02)00133-5.
  • [CM80] Marlene J. Colbourn and Rudolf A. Mathon. On cyclic Steiner 2-designs. In C.C. Lindner and A. Rosa, editors, Topics on Steiner Systems, volume 7 of Annals of Discrete Mathematics, pages 215–253. Elsevier, 1980. doi:10.1016/S0167-5060(08)70182-1.
  • [Col79] M J Colbourn. An analysis technique for Steiner triple systems. Proc. Tenth Southeastern Conf. Combin., Graph Theory, Computing, page 289–303, 1979.
  • [CST13] Xi Chen, Xiaorui Sun, and Shang-Hua Teng. Multi-stage design for quasipolynomial-time isomorphism testing of steiner 2-systems. In Proceedings of the Forty-Fifth Annual ACM Symposium on Theory of Computing, STOC ’13, page 271–280, New York, NY, USA, 2013. Association for Computing Machinery. doi:10.1145/2488608.2488643.
  • [CTW13] Arkadev Chattopadhyay, Jacobo Torán, and Fabian Wagner. Graph isomorphism is not AC0\rm AC^{0}-reducible to group isomorphism. ACM Trans. Comput. Theory, 5(4):Art. 13, 13, 2013. Preliminary version appeared in FSTTCS ’10; ECCC Tech. Report TR10-117. doi:10.1145/2540088.
  • [ELGO02] Bettina Eick, C. R. Leedham-Green, and E. A. O’Brien. Constructing automorphism groups of pp-groups. Comm. Algebra, 30(5):2271–2295, 2002. doi:10.1081/AGB-120003468.
  • [GV06] Martin Grohe and Oleg Verbitsky. Testing graph isomorphism in parallel by playing a game. In Michele Bugliesi, Bart Preneel, Vladimiro Sassone, and Ingo Wegener, editors, Automata, Languages and Programming, 33rd International Colloquium, ICALP 2006, Venice, Italy, July 10-14, 2006, Proceedings, Part I, volume 4051 of Lecture Notes in Computer Science, pages 3–14. Springer, 2006. doi:10.1007/11786986_2.
  • [Hub11] Michael Huber. Computational complexity of reconstruction and isomorphism testing for designs and line graphs. Journal of Combinatorial Theory, Series A, 118(2):341–349, 2011. doi:10.1016/j.jcta.2010.06.006.
  • [IPZ01] Russell Impagliazzo, Ramamohan Paturi, and Francis Zane. Which problems have strongly exponential complexity? Journal of Computer and System Sciences, 63(4):512–530, 2001. doi:10.1006/jcss.2001.1774.
  • [Kie19] Sandra Kiefer. Power and Limits of the Weisfeiler–Leman Algorithm. PhD thesis, RWTH Aachen University, 2019. URL: https://publications.rwth-aachen.de/record/785831/files/785831.pdf.
  • [KST92] Johannes Köbler, Uwe Schöning, and Jacobo Torán. Graph isomorphism is low for PP. Comput. Complex., 2:301–330, 1992. doi:10.1007/BF01200427.
  • [Lad75] Richard E. Ladner. On the structure of polynomial time reducibility. J. ACM, 22(1):155–171, January 1975. doi:10.1145/321864.321877.
  • [LGR16] François Le Gall and David J. Rosenbaum. On the group and color isomorphism problems. arXiv:1609.08253 [cs.CC], 2016.
  • [Lov75] L. Lovász. On the ratio of optimal integral and fractional covers. Discrete Mathematics, 13(4):383–390, 1975. doi:10.1016/0012-365X(75)90058-8.
  • [Luk82] Eugene M. Luks. Isomorphism of graphs of bounded valence can be tested in polynomial time. Journal of Computer and System Sciences, 25(1):42–65, 1982. doi:10.1016/0022-0000(82)90009-5.
  • [Mat79] Rudolf Mathon. A note on the graph isomorphism counting problem. Information Processing Letters, 8(3):131–136, 1979. doi:10.1016/0020-0190(79)90004-8.
  • [Mil78] Gary L. Miller. On the nlognn^{\log n} isomorphism technique (a preliminary report). In Proceedings of the Tenth Annual ACM Symposium on Theory of Computing, STOC ’78, pages 51–58, New York, NY, USA, 1978. Association for Computing Machinery. doi:10.1145/800133.804331.
  • [Nak06] Hiroyuki Nakasora. Mutually orthogonal latin squares and self-complementary designs. Mathematical Journal of Okayama University, 01 2006.
  • [Neu79] A. Neumaier. Strongly regular graphs with smallest eigenvalue m-m. Archiv der Mathematik, 33:392–400, 1979. doi:10.1007/BF01222774.
  • [NS18] Daniel Neuen and Pascal Schweitzer. An exponential lower bound for individualization-refinement algorithms for graph isomorphism. In Ilias Diakonikolas, David Kempe, and Monika Henzinger, editors, Proceedings of the 50th Annual ACM SIGACT Symposium on Theory of Computing, STOC 2018, Los Angeles, CA, USA, June 25-29, 2018, pages 138–150. ACM, 2018. doi:10.1145/3188745.3188900.
  • [RCW75] Dijen K. Ray-Chaudhuri and Richard M. Wilson. On tt-designs. Osaka Journal of Mathematics, 12(3):737–744, 1975. doi:10.18910/7296.
  • [Ros13] David J. Rosenbaum. Bidirectional collision detection and faster deterministic isomorphism testing. arXiv:1304.3935 [cs.DS], 2013.
  • [Sch88] Uwe Schöning. Graph isomorphism is in the low hierarchy. Journal of Computer and System Sciences, 37(3):312 – 323, 1988. doi:10.1016/0022-0000(88)90010-4.
  • [SD76] Douglas C. Schmidt and Larry E. Druffel. A fast backtracking algorithm to test directed graphs for isomorphism using distance matrices. J. ACM, 23(3):433–445, July 1976. doi:10.1145/321958.321963.
  • [SF21] Tyler Schrock and Rafael Frongillo. Computational complexity of k-block conjugacy. Theoretical Computer Science, 856:21–40, 2021. doi:10.1016/j.tcs.2020.12.009.
  • [Smo87] Roman Smolensky. Algebraic methods in the theory of lower bounds for boolean circuit complexity. In Alfred V. Aho, editor, Proceedings of the 19th Annual ACM Symposium on Theory of Computing, 1987, New York, New York, USA, pages 77–82. ACM, 1987. doi:10.1145/28395.28404.
  • [Spi96] Daniel A. Spielman. Faster Isomorphism Testing of Strongly Regular Graphs. In Proceedings of the Twenty-Eighth Annual ACM Symposium on Theory of Computing, STOC ’96, page 576–584, New York, NY, USA, 1996. Association for Computing Machinery. doi:10.1145/237814.238006.
  • [Suz69] Michio Suzuki. A simple group of order 448,345,497,600448,345,497,600, 1969.
  • [Tan13] Bangsheng Tang. Towards Understanding Satisfiability, Group Isomorphism and Their Connections. PhD thesis, Tsinghua University, 2013. URL: http://papakonstantinou.org/periklis/pdfs/bangsheng_thesis.pdf.
  • [Tor04] Jacobo Torán. On the hardness of graph isomorphism. SIAM J. Comput., 33(5):1093–1108, 2004. doi:10.1137/S009753970241096X.
  • [Vol99] Heribert Vollmer. Introduction to Circuit Complexity - A Uniform Approach. Texts in Theoretical Computer Science. An EATCS Series. Springer, 1999. doi:10.1007/978-3-662-03927-4.
  • [Wil19] James B. Wilson. The threshold for subgroup profiles to agree is logarithmic. Theory of Computing, 15(19):1–25, 2019. doi:10.4086/toc.2019.v015a019.
  • [Wol94] Marty J. Wolf. Nondeterministic circuits, space complexity and quasigroups. Theoretical Computer Science, 125(2):295–313, 1994. doi:10.1016/0304-3975(92)00014-I.
  • [ZKT85] V. N. Zemlyachenko, N. M. Korneenko, and R. I. Tyshkevich. Graph isomorphism problem. Journal of Soviet Mathematics, 29:1426–1481, 1985. doi:10.1007/BF02104746.
  • [Zoo] Complexity zoo. URL: https://complexityzoo.net.