Transversals in quasirandom latin squares
Abstract.
A transversal in an latin square is a collection of entries not repeating any row, column, or symbol. Kwan showed that almost every latin square has transversals as . Using a loose variant of the circle method we sharpen this to . Our method works for all latin squares satisfying a certain quasirandomness condition, which includes both random latin squares with high probability as well as multiplication tables of quasirandom groups.
1. Introduction
A transversal in an latin square is a set of entries such that no two of them come from the same row or column or contain the same symbol.
Although there are examples of latin squares with no transversals (e.g., the multiplication table of for even), it is widely believed that these are rare. For example, a famous conjecture of Ryser claims that an latin square contains a transversal whenever is odd. In the same direction, Kwan [kwan] proved that a uniformly random latin square has a transversal with probability . Moreover, he showed that, with probability , the number of transversals is .
In this paper we improve Kwan’s result by finding the precise asymptotic of the number of transversals in a uniformly random latin square.
Theorem 1.1.
Let be a uniformly random latin square. Then has transversals with probability as .
More generally, we find a (deterministic) quasirandomness condition for latin squares which is sufficient to guarantee this same asymptotic number of transversals.
Theorem 1.2.
There is a constant such that the following holds. Let be an latin square which is -quasirandom with parameter . Then has transversals.
The precise definition of “-quasirandom” is in terms of the spectral gap of some operator associated to : see Definition 7.1\wrtusdrfdef:quasirandom. Despite the language, it is not actually obvious that a uniformly random is quasirandom with high probability as , and hence that Theorem 1.2\wrtusdrfthm:main-quasirandom implies Theorem 1.1\wrtusdrfthm:main-random. Indeed, it is incredibly delicate to prove any statistical properties of a uniform random latin square, for a number of reasons: the exact asymptotic count of latin squares is not known; the latin square property is too rigid to make local changes; and no efficient way of sampling uniform random latin squares is known.
However, using a recent result of Kwan, Sah, Sawhney, and Simkin [KSSS] we are indeed able to establish that a random latin square is -quasirandom with parameter , with high probability, and we can thus prove Theorem 1.1\wrtusdrfthm:main-random as a consequence of Theorem 1.2\wrtusdrfthm:main-quasirandom.
Theorem 1.3.
Let be a uniformly random latin square. Then is -quasirandom with parameter , with probability as .
Somewhat opposite to random latin squares are latin squares which are multiplication tables of finite groups. In [EMM] we proved that as long as the underlying group satisfies a necessary condition to have at least one transversal, we have the count as in Theorem 1.2\wrtusdrfthm:main-quasirandom with an extra factor equal to the size of the group’s abelianization. For some groups, this result is implied by Theorem 1.2\wrtusdrfthm:main-quasirandom and the following (easy) result.
Theorem 1.4.
Let be a group and let be the multiplication table of . Then is -quasirandom with parameter , where is the minimal dimension of a nontrivial linear representation of .
This shows that the -quasirandomness condition when restricted to group multiplication tables coincides with the usual notion of quasirandomness for groups due to Gowers [gowers]. Thus together Theorems 1.2 and 1.4\wrtusdrfthm:main-quasirandom,thm:quasirandom-groups-are-quasirandom recover the main result of [EMM] for sufficiently quasirandom groups.
There appears to be no single universal definition of a “quasirandom latin square”, in the same way that there is no single definition of a “quasirandom set of integers”. Instead there are various possible qualitatively inequivalent definitions, some more natural than others, and the correct choice depends on the problem at hand. For this reason we prefer to talk about a quasirandomness condition than about a “definition of quasirandomness”, and we do not claim that the condition in Definition 7.1\wrtusdrfdef:quasirandom is necessarily the correct one for other problems. In particular it is not directly related to the notion introduced in [MR4012871, MR4456029], since that depends on some additional structure (namely, an ordering on the set of symbols) to which our condition is oblivious. See Section 7\wrtusdrfsec:dense-minor for further remarks.
2. Outline
Our approach is analytical rather than combinatorial. Let , , be -element sets of rows, columns and symbols. We identify an latin square with a subset of satisfying the latin square property, i.e., every pair from , and is in exactly one triple from . We let denote the spaces of complex-valued functions on , , (equipped with the standard hermitian inner product). The latin square tensor is defined by
for .
We stress that the latin square tensor depends on , but we will always just write for brevity. We use the same notation for powers of , in the following sense. If and are latin squares then is also a latin square, where is simply the cartesian product of and as subsets of and . Accordingly the powers are latin squares of order for all , and if then we write
Of particular interest is the latin square . We write (or sometimes to emphasize the domain, and similarly and ) for the subset of all bijections . Then one can check that the number of transversals in is
Our goal is therefore to show that
(2.1) |
provided that satisfies an appropriate quasirandomness condition.
We approach (2.1) principally by studying how deviates from its density . We do this as follows. For any set , we may identify with the subspace of consisting of functions that factor as ; i.e., functions that only depend on . These spaces are nested: if then . We write for the orthogonal projection and for the orthogonal projection
(2.2) |
Here is the space of functions orthogonal to functions depending only on , i.e., such that for any choice of . Therefore the space on the right-hand side of (2.2) is the space of functions depending only on and such that for any .
The operators , are related via inclusion–exclusion rules:
Hence we have a kind of “Fourier expansion”
for any function (which is only truly a Fourier expansion if and is identified with ). Applying this to ,
By the discussion above, can be thought of as the component of which depends exactly on (and is orthogonal to all functions depending only on variables in a strict subset of ). For example, is equal to the density .
The relevance of the projections is that any latin square tensor is diagonal with respect to this decomposition: that is,
(2.3) |
This is a consequence of the following lemma.
Lemma 2.1.
Let and . Then
unless .
Proof.
Assume it is not the case that . By symmetry we may assume , say . We may also assume , , , by replacing with their images under , respectively. Now consider
In particular consider the average over the variables . Since , there is no dependence on , so it is equivalent by the latin square property to average over all . Since , it follows that . ∎
We now divide up the sum (2.3) according to the size of .
2.1. Major arcs
The terms in this decomposition where is very sparse (of size up to ) form the major arcs.
Theorem 2.2.
There is a constant such that for ,
The proof is a mostly mechanical adaptation of [EMM, Section 4], which did not use group theory in an essential way.
2.2. Sparse minor arcs
The next range, the sparse minor arcs, concerns of size up to for some small absolute constant .
Theorem 2.3.
There is a constant such that for ,
Note by the triangle inequality. The point is we have an exponential-in- gain over the main term provided
for some large enough . This would be satisfied as long as
(2.4) |
for some large enough and small enough .
We prove Theorem 2.3\wrtusdrfthm:sparse-minor by exhibiting a majorant for and then using generating function methods.
2.3. Dense minor arcs
Finally we have the dense range, where . Here we use quasirandomness. To be precise we define a certain Markov chain on , with adjacency operator , and we consider to be -quasirandom with parameter if has a spectral gap at least , i.e., if the spectral radius of is at most , where is the projection to constants (the uniform distribution). See Definition 7.1\wrtusdrfdef:quasirandom.
Theorem 2.4.
For every there is such that if is -quasirandom with parameter , then
2.4. Quasirandomness
It remains (for Theorems 1.1 and 1.4\wrtusdrfthm:main-random,thm:quasirandom-groups-are-quasirandom) to demonstrate that the latin squares in scope are quasirandom in this sense. If is the multiplication table of a group we compute the entire spectrum of and find where is the minimal dimension of a nontrivial representation of , which shows that our notion of quasirandomness is equivalent to the usual one due to Gowers [gowers] in the case of groups. For genuinely random latin squares we use recent work of Kwan, Sah, Sawhney, and Simkin [KSSS] to show that with high probability, and this implies that .
2.5. Proof of Theorem 1.2\wrtusdrfthm:main-quasirandom
Putting Theorems 2.2, 2.3 and 2.4\wrtusdrfthm:major-arcs,thm:sparse-minor,thm:dense-minor together it is straightforward to deduce Theorem 1.2\wrtusdrfthm:main-quasirandom.
Proof of Theorem 1.2\wrtusdrfthm:main-quasirandom.
2.6. Layout of the paper
To prove Theorems 2.2, 2.3 and 2.4\wrtusdrfthm:major-arcs,thm:sparse-minor,thm:dense-minor we need some background material on partition systems (Section 3\wrtusdrfsec:partitions) and on the primitive “Fourier analysis” of coordinate projections , discussed above (Section 4\wrtusdrfsec:fourier). This builds on similar material from [EMM].
Then, Sections 5, 6 and 7\wrtusdrfsec:major-arcs,sec:true-sparse-minor,sec:dense-minor give the proofs of the three key theorems above. Finally, Section 8\wrtusdrfsec:quasirandomness proves the quasirandomness properties from Section 2.4\wrtusdrfsub:outline-quasi.
3. Partitions and partition systems
3.1. Partitions
Most of our language relating to the partition lattice is standard.
-
(1)
If is a set, is the set of all partitions of . If we will conserve brackets by writing simply .
-
(2)
is the set of partitions all of whose cells have size at most .
-
(3)
If then any partition of is identified with a partition of by adding singletons . With this convention, .
-
(4)
The support of a partition is the union of the nonsingleton cells of . It is the smallest set such that .
-
(5)
is the set of with .
-
(6)
If , means that is a refinement of (i.e., every cell of is a union of cells of ). Synonymously, is a coarsening of .
-
(7)
The meet is the coarsest partition refining both and ; the join is the finest partition coarsening both and .
-
(8)
The partition is the discrete partition; the partition is the indiscrete partition.
-
(9)
The rank of a partition is ; equivalently it is the greatest such that there are partitions . (Note that is meaningful without specifying , unlike ; i.e., it is invariant under adding or removing singletons.)
-
(10)
The Möbius function at is given by .
-
(11)
A function is -measurable if is constant on the cells of . A subset is called -measurable if is -measurable.
-
(12)
If , is the indicator of -measurability (i.e., is if is -measurable, and otherwise).
3.2. Partition systems
In Sections 5 and 6\wrtusdrfsec:major-arcs,sec:true-sparse-minor it will be essential to have good bounds on the quantity for and various choices . This motivates the following definitions.
-
(1)
A partition triple on a set is a triple .
-
(2)
We call a partition system if .
-
(3)
The support of is .
Definition 3.1 (Combinatorial rank).
Let be a partition triple. We write to mean that , i.e., is a collection of cells labelled 1, 2, or 3. A subset is closed (with respect to ) if whenever for and , if two of are in then so is the third. The closure of is the intersection of all closed sets containing . The combinatorial rank of is defined as
The motivation for combinatorial rank is the following bound.
Lemma 3.2.
For a set , partitions , and latin square ,
The idea of the proof is the same as for the related result [EMM, Lemma 4.6].
Proof.
The value is, by definition, times the number of triples of functions
such that is -measurable for and such that for all . Note we can think of as a function on the cells of , since it is -measurable.
We claim that, given with , the triple is determined by the values of on cells in . Hence the number of such triples is at most , giving the result.
Indeed, suppose is another triple of measurable functions with the same restriction to . Let be the set of all cells such that . By hypothesis . If for , , and two of are in , then so is the third, as the triples agree at two coordinates and so are equal by the latin square property. Hence is a closed set, so and , as required. ∎
This reduces the problem of bounding from above to the problem of bounding from below. In [EMM] we did this using two slightly weaker notions of rank, called triple rank and lower rank, defined respectively as
Lemma 3.3.
.
Proof.
For the first inequality, let contain all of and one cell of from each cell of . Then and , so , and equally for other permutations of 1, 2, 3. The second inequality was proved in [EMM, Lemma 4.8], and in any case will not be used in this paper. ∎
For continuity with [EMM], we define the complexity of a partition system to be
The complexity of a partition system is nonnegative, and it is zero if and only if for some matching , i.e., a partition of into pairs.
3.3. Combinatorial rank of matching systems
In this subsection we compute for all , i.e., partition triples such that all cells of have size at most 2. Where it applies, this is a significant improvement on what Lemma 3.3\wrtusdrflem:crank¿=trank gives us.
Lemma 3.4.
Let . Suppose there are precisely cells such that has full support (i.e., is a matching) for each . Then
Proof.
Since all terms are additive across cells of , we may assume is indiscrete. In particular , and if and only if one of has a singleton.
Case : In this case are matchings, so
and we must show
Let be the multigraph whose vertex set is and with edges given by the cells of (which are all 2-cells). Clearly is -regular, with vertices and edges. Since is indiscrete, is connected.
According to Definition 3.1\wrtusdrfdef:crank, we want to infect as few edges as possible in such a way that, if two infected edges incident at a vertex always spread infection to the third edge, then infection spreads to all edges. Note that for this to happen, it is necessary and sufficient to infect at least one edge in each cycle, since the edges that are uninfected at the end of the process form a subgraph with no vertex of degree . Hence, equivalently, we want to delete as few edges as possible to get a forest.
Since has edges and any forest has at most edges, we must delete at least edges. Conversely, given any connected 3-regular multigraph, we can delete edges until we have a (simple) tree. Hence the minimal number of generators is precisely , so , as claimed.
Case : In this case at least one of has a singleton, and we must show that
We define a graph as in the previous case but additionally for every singleton we add an edge , where is a special additional vertex at which infection does not spread. Since is indiscrete, is connected. Since there is at least one singleton, is connected. Again we want to delete as few edges as possible to get a forest. The number of vertices in is and the number of edges is , so the number of edges we must delete is precisely Hence
as claimed. ∎
4. The “Fourier” expansion of
Recall from Section 2\wrtusdrfsec:outline that denotes the orthogonal projection and denotes the orthogonal projection
and these operators are related via inclusion–exclusion rules:
(4.1) |
In this section we study the terms in the expansion
To express some of the results it is convenient to use the linear map defined by
Here .
4.1. Formulas for
Let denote the set of injections . Thus if , .
Lemma 4.1.
If and ,
Proof.
A function can be extended to a bijection in ways if is injective and ways otherwise, and by definition is the number of such extensions normalized by . ∎
Lemma 4.2 ([EMM, Lemma 4.3]).
Lemma 4.3.
If and then
Proof.
Combining the previous two lemmas,
Since , only the terms with survive. ∎
We can use to give another formula for . If (i.e., ), the kernel of is the level set partition
Note that
Lemma 4.4.
Let , . For , let . Then
4.2. Sparseval
The word sparseval is our playful term for the computation of for any . This is possible by inclusion–exclusion and orthogonality: since
it follows that
(4.2) |
Lemma 4.5.
If and ,
Proposition 4.6.
Assume and let . Then
where
In particular
Sketch.
The inequality follows from the previous lemma. For the main claim, by expanding we have
and this can be identified as times the coefficient of in . The stated bound follows by taking a contour integral (chosen in the spirit of the saddle-point method) to extract the coefficient. For details see [EMM-cyclic, bound for the sum in (5.4)]. Extra care is needed for near 1, but we omit the details because we will not use the claim for . The second bound follows by Stirling’s formula. ∎
The following corollary will not be used but is included for interest.
Corollary 4.7.
The sign of is .
Proof.
Let . By Lemma 4.4\wrtusdrfprop:PA1S-sparseval-like-expression it suffices to prove that
There are nonnegative integers such that
Hence the claim follows from . ∎
5. Major arcs
The goal in this section is to prove Theorem 2.2. Define
Our aim is to prove that, for ,
(5.1) |
In particular this implies Theorem 2.2\wrtusdrfthm:major-arcs.
5.1. The quantities and
From Lemma 4.3\wrtusdrflem:P_A-on-1_S it is clear that to estimate it suffices to estimate for every partition system with support and aggregate the results with the appropriate weighting. For continuity with [EMM, Section 4], we define the normalized quantities
and
for any partition triple . Note that
by Lemmas 3.2 and 3.3\wrtusdrflem:crank-bound,lem:crank¿=trank. Since , unless is a partition system.
Lemma 5.1.
Let be a partition system with support of size , and suppose points of are contained in cells of size at least . Then
Sketch.
The idea is that
and
where
Let . Then, normalizing,
where
In [EMM, Section 4] we showed . Since for all this shows . The stronger bound with in place of follows by separating off the doubleton cells of . See [EMM, Section 4] for details. ∎
5.2. The series
For a partition triple we use the shorthand
From Lemma 4.3 we have
where the sum is over all partition systems on . For define
As we have mentioned, for any partition system , so is a polynomial such that . By bounding and using some complex analysis we will show , and then we will directly estimate .
Proposition 5.2.
There is a constant such that, for , we have
Proof.
By the definition of , the triangle inequality, and Lemma 5.1\wrtusdrflem:trivial-gamma-bound, the quantity is bounded by
where is the number of points of contained in cells of of size at least 3. This exact sum was analyzed in [EMM, Section 4.4], and we showed that it is provided . The proposition follows. ∎
Corollary 5.3.
There is a constant such that, for ,
Proof.
By the residue theorem,
as long as . Hence
Take and , where is as in the previous proposition. Then as long as , i.e., , we get
Hence as long as say we get the claimed bound. ∎
5.3. The constant term
By definition,
As remarked, if and only if for some matching . In this case, if say ,
The last identity holds by a direct calculation analogous to [EMM, Lemma 4.10]. The number of matchings in of support size is
Thus
By combining with Corollary 5.3 we have
provided . This finishes the proof of (5.1).
6. Sparse minor arcs
To prove Theorem 2.3\wrtusdrfthm:sparse-minor we need a bound on for larger . Note that in any latin square ,
(6.1) |
using the latin square property and Cauchy–Schwarz, and similarly permuting . One approach to Theorem 2.3\wrtusdrfthm:sparse-minor might be to find upper bounds on , pointwise or in , and simply apply (6.1). However, by itself this approach is too crude, even assuming optimal upper bounds.
Another idea is to seek a majorant for of the form
(6.2) |
for some coefficients . Then
and Lemma 3.2\wrtusdrflem:crank-bound, together with generating function techniques, gives a way to control the right-hand side. This bound is particularly effective if , given Lemma 3.4\wrtusdrfclaim:cool-crank-value.
Again this approach does not succeed by itself. Our final argument works by decomposing into two pieces and combining the two techniques discussed above.
6.1. A majorant for
Throughout this section let be some large enough constant, and . Additionally, we let
Although we will always have this specific value of in mind, most of the results in this section only rely on . The next proposition gives a useful bound for . For , , and a partition define
Proposition 6.1.
We have
Proof.
From Lemma 4.4\wrtusdrfprop:PA1S-sparseval-like-expression,
where and
From Proposition 4.6\wrtusdrfprop:umbral-sparseval and crude estimates (Stirling’s formula), for ,
provided is large enough. Then
as required. ∎
In light of the proposition, to find majorants for of the form (6.2) it suffices to find analogous bounds for . Recall that is the set of all having no part of size greater than . Let be the number of -cells in and let .
Lemma 6.2.
Let be a partition.
-
(1)
-
(2)
-
(3)
Proof.
Consider the first inequality. Both sides are multiplicative across cells of , so we may assume is a single cell, say of size . The inequality is trivial for (since is one of the summands on the right-hand side), so we may assume . Then it suffices to check
This is a calculation for (say) and an uninteresting exercise for .
Now consider the second inequality. This time, the right-hand side is not itself multiplicative over cells of , but if we replace the condition by the stronger one
then it becomes so, and it suffices to prove the corresponding stronger inequality. Now we may again assume that is an -cell, and we may assume . Then we must check
Again we omit further details.
Now consider the third inequality. Again it suffices to consider the case of an -cell, and we may assume . Then the assertion is
Since , it suffices to check
which is again essentially a calculation. ∎
Lemma 6.3.
Let . Then
Proof.
6.2. A splitting of
We can use the bound on given in the previous section to bound the norm of , but the bound would not be strong enough for what we need. To go further, we break up into two parts, a part whose norm we can control better, and a part we can analyze separately. Fix and let
Let . Define
Clearly .
Lemma 6.4.
We have
Proof.
By Proposition 6.1\wrtusdrfprop:PA1S-physical-bound,
Suppose . Then by Lemma 6.2(2),
This proves the bound on . The first bound on is proved identically. The second is proved in the same way using instead Lemma 6.2(3). ∎
Corollary 6.5.
We have
Proof.
Using the previous lemma, and ,
Let
Then, for real ,
Using the exponential formula (3.1) with for , for , and for , we obtain
Hence for real ,
Putting , , and , we get
This proves what we want. ∎
Corollary 6.6.
We have
Proof.
We have since pointwise. Hence, from (6.1),
From sparseval (Lemma 4.5\wrtusdrflem:L2-to-sparseval and Proposition 4.6\wrtusdrfprop:umbral-sparseval),
Combining with Corollary 6.5\wrtusdrfcor:rflat-l1-bound gives the bound. ∎
6.3. The contribution from
Finally we must bound . From Lemma 6.4,
(6.3) |
where
Hence it suffices to bound . The key ingredient for this is the knowledge of the exact value of combinatorial rank for (Lemma 3.4\wrtusdrfclaim:cool-crank-value).
Lemma 6.7.
Proof.
Let be the set of matchings (partitions all of whose cells have size 2). For , let be the number of cells such that for each . Then, from Lemma 3.2\wrtusdrflem:crank-bound and Lemma 3.4\wrtusdrfclaim:cool-crank-value,
Hence
In the last sum above, since , is or according to whether . Splitting the sum according to these cases,
In the second sum we will ignore the constraint ; in the first sum we will use only .
Fix parameters for all . Define
Then, by the discussion above,
Three applications of the exponential formula (3.1) give
(6.4) | ||||
(6.5) | ||||
(6.6) |
From (6.4), for real ,
Replacing with , putting (we will ensure later that for ) and gives
From (6.5) with we have
(alternatively, this follows directly from ). Hence, from (6.6) for ,
(6.7) |
where is the truncated sum
Inserting and ,
Note that , and this is indeed at least 1 for since we may assume . Finally, let for a sufficiently small constant . Then
Hence, from (6.7),
as claimed. ∎
Putting the last few results together, we have the following theorem, which clearly implies Theorem 2.3\wrtusdrfthm:sparse-minor.
Theorem 6.8.
We have
7. Dense minor arcs
Define a Markov chain on as follows. If the current state is , pick uniformly at random . The next state is , where and are the unique solutions to
(see Figure 1\wrtusdrffig:markov-chain). Let be the transition operator for this Markov chain:
The Markov chain is reversible with uniform stationary distribution, so is self-adjoint and has the constant function on as a -eigenvector. Let be the projection to constants:
Definition 7.1.
We say is -quasirandom with parameter if has spectral radius at most .
In particular, if and only if the Markov chain is connected, and in general measures the rate of mixing.
Remark 7.2.
For a finite set , let denote the subspace . Then equivalently, is -quasirandom with parameter if the restriction has spectral radius at most .
All our applications of quasirandomness go through the following lemma.
Lemma 7.3.
Assume is -quasirandom with parameter and111Note that the case of (7.1) does not obviously imply the general case: the operator-type norm for trilinear forms does not behave well under taking tensor powers. let . Then
(7.1) |
for all , , .
Remark 7.4.
Identifying with in the usual way, is identified with the subspace ; see (2.2).
Proof of Lemma 7.3\wrtusdrflem:quasi-use.
By Cauchy–Schwarz,
Note , and that . Since has spectral radius at most , the tensor power has spectral radius (and hence operator norm) at most , so the last expression above is bounded by . ∎
Remark 7.5.
As stated in the introduction, while Definition 7.1\wrtusdrfdef:quasirandom has some nice properties (e.g., the spectral radius of can be computed efficiently), it is chosen for mainly practical rather than philosophical reasons, and there are similar but qualitatively inequivalent conditions that would work equally well.
One notable criticism of this definition is that latin squares associated to Steiner triple systems (i.e., where and contains the diagonal and is invariant under the -action on triples) always fail to be -quasirandom with parameter (since the diagonal of is a closed set for the Markov chain). On the other hand, a random Steiner triple system is far from having algebraic structure and presumably satisfies (7.1) for with high probability as .
One point of view is that (7.1) itself is the more natural quasirandomness condition (but harder to verify), and Definition 7.1\wrtusdrfdef:quasirandom is a convenient sufficient condition.
Proof of Theorem 2.4\wrtusdrfthm:dense-minor.
Let and . By Lemmas 7.3 and 7.2\wrtusdrflem:quasi-use,rem:l20,
Hence, for ,
Taking and so that , the result follows. ∎
8. Quasirandomness
In this section we will verify that two natural classes of latin squares are -quasirandom with parameter :
-
•
multiplication tables of quasirandom groups;
-
•
uniformly random latin squares, with high probability as .
In the case of a group we can compute the whole spectrum of using representation theory. In the case of a random latin square we will use the bound
which holds because the spectrum of is real and is even. By interpreting as counting certain kinds of configuration in (and using a recent result of [KSSS]) we will show that with high probability, which implies that . (Using the same method one can show that with high probability, so is the smallest even integer that we can use for this argument.)
8.1. Quasirandom groups
The following proposition shows that our quasirandomness condition generalizes the definition of a quasirandom group (see [gowers]), implying Theorem 1.4.
Proposition 8.1.
Suppose is the multiplication table of a group . Then the spectrum of consists of copies of and copies of for every -dimensional irreducible representation of , and zeros. In particular where is the minimal dimension of a nontrivial representation of .
Proof.
Here and , so and is the operator defined by
By representation theory, has an orthogonal basis consisting of the functions of the form , where is an irreducible unitary representation of and is an orthonormal basis of .
It follows that has an orthogonal basis consisting of functions of the form
where and are two irreducible unitary representations of and , .
To find we recall the Schur orthogonality relation for matrix coefficients, which states that for irreducible , as above, and ,
and thereby compute
In the case we get an eigenfunction with eigenvalue . When and we get a -eigenfunction. Finally when and , the functions
are eigenfunctions of with eigenvalues respectively.
Altogether we have copies of and copies of , and the rest , as claimed.
∎
8.2. Random latin squares
We will use a recent result of Kwan, Sah, Sawhney, and Simkin [KSSS] on configuration counts in random latin squares. A triple system is a 3-uniform 3-partite hypergraph with vertex classes . The number of vertices is and the number of triples (hyperedges) is . We say is latin if every pair of vertices is in at most one triple. (A latin square of order is then a latin triple system with 3 classes of vertices and triples.)
Let be a fixed triple system. A copy of in a triple system is a triple of injective maps
which maps triples to triples. Let denote the number of copies of in .
Let denote the random triple system in which each possible triple is present independently with probability . Note that (when is fixed and is large). We say is -stable if and
for any latin triple system with at most triples.
Theorem 8.2 ([KSSS, Theorem 7.2]).
Fix an -stable latin triple system with vertices and triples. Let be a uniformly random latin square. Then
with high probability as .
In order to use this theorem effectively we need a computable form of stability. Let be a latin triple system. A subset of the vertices is called closed if whenever two vertices of a triple of is in , so is the third. The closure of a subset if the smallest closed set containing it. If let denote the vertices incident with at least one member of , and let and . We say generates if
Let
For example, if is the latin triple system shown in Figure 2\wrtusdrffig:trAdj6, one generating set consists of both triples containing , one triple containing , and one triple containing , and there is no smaller generating set, so .
Lemma 8.3.
Let be a latin triple system with vertices and triples. Then is -stable provided and
Remark 8.4.
A much simpler model problem is the following: given a fixed graph and a random graph , does contain copies of (i.e., close to the expected number) with high probability? The answer might be no if contains a subgraph with much greater density than in some sense: indeed, if then with high probability contains zero copies of , and hence of . However, this is essentially all that can go wrong. The condition for -stability in the lemma captures a similar intuition.
Remark 8.5.
Given a triple system , one can construct a partition triple in the sense of Section 3.1 (i.e., the ground set has size ) where two triples lie in the same cell of if and only if , and similarly for and , and and .
The construction can be reversed (up to the issue of repeated edges). In other words, triple systems and partition triples are more-or-less the same objects. Under this analogy, the notion of closure here coincides with that in Definition 3.1, and .
Although using both languages is strictly speaking redundant, it is useful to keep the two notions separate, partly for minor technical reasons, but mainly because using partition systems follows our previous work in [EMM-cyclic, EMM] while using triple systems follows [KSSS].
Proof of Lemma 8.3\wrtusdrflem:stability-lemma.
(Cf. [KSSS, p. 15]) Let be a latin triple system with at most triples. For a copy of in , say one of its triples is forced if it appears in . The difference
(8.1) |
arises from copies of with at least one forced triple. Let be a nonempty subsystem and consider copies of whose forced triples are precisely the images of those in . Let be a generating subsystem of size . Because satisfies the latin property, any copy of in is determined by the image of . Therefore the number of copies of in is at most . There are vertices of outside , each with possible images in , and the image of each of the triples outside has probability (independently) of being present in . Hence the contribution to (8.1) from is bounded by
This is provided the stated condition is satisfied. ∎
Now we can show that random latin squares are -quasirandom with parameter with high probability (Theorem 1.3\wrtusdrfthm:random-implies-quasirandom). This follows from the following proposition and the bound .
Proposition 8.6.
For a uniformly random latin square ,
with high probability as .
Proof (computer-assisted).
For , let denote the iterates of under the Markov chain defining . Then
where is the number of configurations in of the form shown in Figure 2\wrtusdrffig:trAdj6 with and . We do not assume the other vertices are distinct.
Let be the latin triple system depicted in Figure 2\wrtusdrffig:trAdj6 and let (where is bounded) be all the degenerations obtainable by identifying some (like-colored) vertices and identifying triangles as necessary to preserve the latin property.
Formally, we consider all triples of partitions where , , satisfying the following closure property: if and are two triples of and two of the pairs , , are in the same cell of , , respectively then so is the third. Number such triples of partitions , where corresponds to three copies of the discrete partition. Then denotes the quotient hypergraph of with respect to partition .
Let . Then . Let and . Then . Now the proposition follows from Theorem 8.2\wrtusdrfthm:KSSS, Lemma 8.3\wrtusdrflem:stability-lemma, and the following two claims:
-
(1)
for each ,
-
(2)
for each .
Indeed, provided (1) and (2) hold, Lemma 8.3\wrtusdrflem:stability-lemma shows that is -stable for each , so Theorem 8.2\wrtusdrfthm:KSSS implies that with high probability for each , so with high probability.
Both claims can be verified by exhaustive search. We find by starting with and iteratively identifying pairs of vertices, using breadth-first search. Thus we verify (1). Now for each we check all subsystems and compute by checking all , and thus we verify (2).
It turns out , and there are distinct isomorphism classes among the degenerations . The quantity in (2) turns out to be at most 4 in all cases except , for which it is . There are just eight degenerations (up to isomorphism) for which . Of these, four are just with a single pair of vertices identified (so and ). The other four cases are shown in Figure 3\wrtusdrffig:degenerations. These cases are therefore the dominant contributors to the error term. ∎