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

Transversals in quasirandom latin squares

Sean Eberhard Sean Eberhard, Mathematical Sciences Research Centre, Queen’s University Belfast, Belfast BT7 1NN, UK [email protected] Freddie Manners Freddie Manners, UCSD Department of Mathematics, 9500 Gilman Drive #0112, La Jolla CA 92093, USA [email protected]  and  Rudi Mrazović Rudi Mrazović, University of Zagreb, Faculty of Science, Department of Mathematics, Zagreb, Croatia [email protected]
Abstract.

A transversal in an n×nn\times n latin square is a collection of nn entries not repeating any row, column, or symbol. Kwan showed that almost every n×nn\times n latin square has ((1+o(1))n/e2)n\bigl{(}(1+o(1))n/e^{2}\bigr{)}^{n} transversals as nn\to\infty. Using a loose variant of the circle method we sharpen this to (e1/2+o(1))n!2/nn(e^{-1/2}+o(1))n!^{2}/n^{n}. 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.

SE has received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement No. 803711) and from the Royal Society. RM is supported in part by the Croatian Science Foundation under the project UIP-2017-05-4129 (MUNHANAP). FM is supported by a Sloan Fellowship.

1. Introduction

A transversal in an n×nn\times n latin square is a set of nn 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 𝐙/n𝐙\mathbf{Z}/n\mathbf{Z} for nn even), it is widely believed that these are rare. For example, a famous conjecture of Ryser claims that an n×nn\times n latin square contains a transversal whenever nn is odd. In the same direction, Kwan [kwan] proved that a uniformly random n×nn\times n latin square has a transversal with probability 1o(1)1-o(1). Moreover, he showed that, with probability 1o(1)1-o(1), the number of transversals is ((1o(1))n/e2)n\bigl{(}(1-o(1))n/e^{2}\bigr{)}^{n}.

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 𝖫\mathsf{L} be a uniformly random n×nn\times n latin square. Then 𝖫\mathsf{L} has (e1/2+o(1))n!2/nn\bigl{(}e^{-1/2}+o(1)\bigr{)}n!^{2}/n^{n} transversals with probability 1o(1)1-o(1) as nn\to\infty.

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 ρ>0\rho>0 such that the following holds. Let 𝖫\mathsf{L} be an n×nn\times n latin square which is 𝒜\mathcal{A}-quasirandom with parameter ρ\rho. Then 𝖫\mathsf{L} has (e1/2+o(1))n!2/nn\bigl{(}e^{-1/2}+o(1)\bigr{)}n!^{2}/n^{n} transversals.

The precise definition of “𝒜\mathcal{A}-quasirandom” is in terms of the spectral gap of some operator associated to 𝖫\mathsf{L}: see Definition 7.1\wrtusdrfdef:quasirandom. Despite the language, it is not actually obvious that a uniformly random n×nn\times n is quasirandom with high probability as nn\to\infty, 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 n×nn\times n 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 𝒜\mathcal{A}-quasirandom with parameter o(1)o(1), 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 𝖫\mathsf{L} be a uniformly random n×nn\times n latin square. Then 𝖫\mathsf{L} is 𝒜\mathcal{A}-quasirandom with parameter o(1)o(1), with probability 1o(1)1-o(1) as nn\to\infty.

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 GG be a group and let 𝖫G\mathsf{L}_{G} be the multiplication table of GG. Then 𝖫G\mathsf{L}_{G} is 𝒜\mathcal{A}-quasirandom with parameter 1/D1/D, where DD is the minimal dimension of a nontrivial linear representation of GG.

This shows that the 𝒜\mathcal{A}-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 XX, YY, ZZ be nn-element sets of rows, columns and symbols. We identify an n×nn\times n latin square 𝖫\mathsf{L} with a subset of X×Y×ZX\times Y\times Z satisfying the latin square property, i.e., every pair from X×YX\times Y, Y×ZY\times Z and Z×XZ\times X is in exactly one triple from 𝖫\mathsf{L}. We let L2(X),L2(Y),L2(Z)L^{2}(X),L^{2}(Y),L^{2}(Z) denote the spaces of complex-valued functions on XX, YY, ZZ (equipped with the standard hermitian inner product). The latin square tensor Λ=Λ𝖫\Lambda=\Lambda_{\mathsf{L}} is defined by

Λ(f,g,h):=𝐄(x,y,z)𝖫f(x)g(y)h(z)\Lambda(f,g,h):=\mathbf{E}_{(x,y,z)\in\mathsf{L}}f(x)g(y)h(z)

for fL2(X),gL2(Y),hL2(Z)f\in L^{2}(X),g\in L^{2}(Y),h\in L^{2}(Z).

We stress that the latin square tensor Λ𝖫\Lambda_{\mathsf{L}} depends on 𝖫\mathsf{L}, but we will always just write Λ\Lambda for brevity. We use the same notation for powers of 𝖫\mathsf{L}, in the following sense. If 𝖫1\mathsf{L}_{1} and 𝖫2\mathsf{L}_{2} are latin squares then 𝖫1×𝖫2\mathsf{L}_{1}\times\mathsf{L}_{2} is also a latin square, where 𝖫1×𝖫2\mathsf{L}_{1}\times\mathsf{L}_{2} is simply the cartesian product of 𝖫1\mathsf{L}_{1} and 𝖫2\mathsf{L}_{2} as subsets of X1×Y1×Z1X_{1}\times Y_{1}\times Z_{1} and X2×Y2×Z2X_{2}\times Y_{2}\times Z_{2}. Accordingly the powers 𝖫m\mathsf{L}^{m} are latin squares of order nmn^{m} for all m0m\geqslant 0, and if fL2(Xm),gL2(Ym),hL2(Zm)f\in L^{2}(X^{m}),g\in L^{2}(Y^{m}),h\in L^{2}(Z^{m}) then we write

Λ(f,g,h):=𝐄(x,y,z)𝖫mf(x)g(y)h(z).\Lambda(f,g,h):=\mathbf{E}_{(x,y,z)\in\mathsf{L}^{m}}f(x)g(y)h(z).

Of particular interest is the latin square 𝖫n\mathsf{L}^{n}. We write SS (or sometimes SXS_{X} to emphasize the domain, and similarly SYS_{Y} and SZS_{Z}) for the subset SXnS\subseteq X^{n} of all bijections [n]X[n]\to X. Then one can check that the number of transversals in 𝖫\mathsf{L} is

Λ(1S,1S,1S)n2nn!.\Lambda(1_{S},1_{S},1_{S})\frac{n^{2n}}{n!}.

Our goal is therefore to show that

Λ(1S,1S,1S)=(e1/2+o(1))(n!nn)3,\Lambda(1_{S},1_{S},1_{S})=\bigl{(}e^{-1/2}+o(1)\bigr{)}{\left(\frac{n!}{n^{n}}\right)}^{3}, (2.1)

provided that 𝖫\mathsf{L} satisfies an appropriate quasirandomness condition.

We approach (2.1) principally by studying how 1S1_{S} deviates from its density n!/nnn!/n^{n}. We do this as follows. For any set A[m]A\subseteq[m], we may identify L2(XA)L^{2}(X^{A}) with the subspace of L2(Xm)L^{2}(X^{m}) consisting of functions Xm𝐂X^{m}\to\mathbf{C} that factor as XmXA𝐂X^{m}\to X^{A}\to\mathbf{C}; i.e., functions f(x1,,xm)f(x_{1},\dots,x_{m}) that only depend on (xi:iA)(x_{i}\colon i\in A). These spaces are nested: if ABA\subseteq B then L2(XA)L2(XB)L^{2}(X^{A})\subseteq L^{2}(X^{B}). We write QAQ_{A} for the orthogonal projection L2(Xm)L2(XA)L^{2}(X^{m})\to L^{2}(X^{A}) and PAP_{A} for the orthogonal projection

PA:L2(Xm)L2(XA)BAL2(XB).P_{A}:L^{2}(X^{m})\to L^{2}(X^{A})\cap\bigcap_{B\subsetneq A}L^{2}(X^{B})^{\perp}. (2.2)

Here L2(XB)L^{2}(X^{B})^{\perp} is the space of functions f(x1,,xm)f(x_{1},\dots,x_{m}) orthogonal to functions depending only on (xi:iB)(x_{i}:i\in B), i.e., such that 𝐄xi:iBf(x1,,xm)=0\mathbf{E}_{x_{i}:i\notin B}f(x_{1},\dots,x_{m})=0 for any choice of (xi:iB)(x_{i}:i\in B). Therefore the space on the right-hand side of (2.2) is the space of functions depending only on (xi:iA)(x_{i}:i\in A) and such that 𝐄xif(x1,,xm)=0\mathbf{E}_{x_{i}}f(x_{1},\dots,x_{m})=0 for any iAi\in A.

The operators PAP_{A}, QAQ_{A} are related via inclusion–exclusion rules:

QA\displaystyle Q_{A} =BAPB,\displaystyle=\sum_{B\subseteq A}P_{B},
PA\displaystyle P_{A} =BA(1)|AB|QB.\displaystyle=\sum_{B\subseteq A}(-1)^{|A\setminus B|}Q_{B}.

Hence we have a kind of “Fourier expansion”

f=A[m]PAf,f=\sum_{A\subseteq[m]}P_{A}f,

for any function fL2(Xm)f\in L^{2}(X^{m}) (which is only truly a Fourier expansion if n=2n=2 and XmX^{m} is identified with 𝐅2m\mathbf{F}_{2}^{m}). Applying this to f=1SL2(Xn)f=1_{S}\in L^{2}(X^{n}),

1S=A[n]PA1S.1_{S}=\sum_{A\subseteq[n]}P_{A}1_{S}.

By the discussion above, PA1SP_{A}1_{S} can be thought of as the component of 1S1_{S} which depends exactly on (xi:iA)(x_{i}:i\in A) (and is orthogonal to all functions depending only on variables in a strict subset of AA). For example, P1SP_{\emptyset}1_{S} is equal to the density n!/nnn!/n^{n}.

The relevance of the PAP_{A} projections is that any latin square tensor Λ𝖫\Lambda_{\mathsf{L}} is diagonal with respect to this decomposition: that is,

Λ(1S,1S,1S)=A[n]Λ(PA1S,PA1S,PA1S).\Lambda(1_{S},1_{S},1_{S})=\sum_{A\subseteq[n]}\Lambda(P_{A}1_{S},P_{A}1_{S},P_{A}1_{S}). (2.3)

This is a consequence of the following lemma.

Lemma 2.1.

Let fL2(Xn),gL2(Yn),hL2(Zn)f\in L^{2}(X^{n}),g\in L^{2}(Y^{n}),h\in L^{2}(Z^{n}) and A,B,C[n]A,B,C\subseteq[n]. Then

Λ(PAf,PBg,PCh)=0\Lambda(P_{A}f,P_{B}g,P_{C}h)=0

unless A=B=CA=B=C.

Proof.

Assume it is not the case that A=B=CA=B=C. By symmetry we may assume ABA\not\subseteq B, say iABi\in A\setminus B. We may also assume PAf=fP_{A}f=f, PBg=gP_{B}g=g, PCh=hP_{C}h=h, by replacing f,g,hf,g,h with their images under PA,PB,PCP_{A},P_{B},P_{C}, respectively. Now consider

Λ(f,g,h)=𝐄(x,y,z)𝖫nf(x)g(y)h(z).\Lambda(f,g,h)=\mathbf{E}_{(x,y,z)\in\mathsf{L}^{n}}f(x)g(y)h(z).

In particular consider the average over the variables (xi,yi,zi)𝖫(x_{i},y_{i},z_{i})\in\mathsf{L}. Since iBi\notin B, there is no dependence on yiy_{i}, so it is equivalent by the latin square property to average over all (xi,zi)X×Z(x_{i},z_{i})\in X\times Z. Since 𝐄xif(x1,,xm)=0\mathbf{E}_{x_{i}}f(x_{1},\dots,x_{m})=0, it follows that Λ(f,g,h)=0\Lambda(f,g,h)=0. ∎

We now divide up the sum (2.3) according to the size mm of AA.

2.1. Major arcs

The terms in this decomposition where AA is very sparse (of size up to cn1/2cn^{1/2}) form the major arcs.

Theorem 2.2.

There is a constant c>0c>0 such that for logn<mcn1/2\log n<m\leqslant cn^{1/2},

A[n]|A|mΛ(PA1S,PA1S,PA1S)=(n!nn)3e1/2(1+O(m2/n)).\sum_{\begin{subarray}{c}A\subseteq[n]\\ |A|\leqslant m\end{subarray}}\Lambda(P_{A}1_{S},\>P_{A}1_{S},\>P_{A}1_{S})={\left(\frac{n!}{n^{n}}\right)}^{3}e^{-1/2}{\left(1+O(m^{2}/n)\right)}.

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 AA of size up to cncn for some small absolute constant cc.

Theorem 2.3.

There is a constant c>0c>0 such that for 1mcn1\leqslant m\leqslant cn,

A[n]|A|=mΛ(|PA1S|,|PA1S|,|PA1S|)(n!nn)3O(1)m+n/m(m/n)m/8.\sum_{\begin{subarray}{c}A\subseteq[n]\\ |A|=m\end{subarray}}\Lambda(|P_{A}1_{S}|,|P_{A}1_{S}|,|P_{A}1_{S}|)\leqslant{\left(\frac{n!}{n^{n}}\right)}^{3}O(1)^{m+n/m}(m/n)^{m/8}.

Note |Λ(PA1S,PA1S,PA1S)|Λ(|PA1S|,|PA1S|,|PA1S|)|\Lambda(P_{A}1_{S},P_{A}1_{S},P_{A}1_{S})|\leqslant\Lambda(|P_{A}1_{S}|,|P_{A}1_{S}|,|P_{A}1_{S}|) by the triangle inequality. The point is we have an exponential-in-mm gain over the main term provided

mlog(n/m)>C(m+n/m),m\log(n/m)>C^{\prime}(m+n/m),

for some large enough C>0C^{\prime}>0. This would be satisfied as long as

C(n/logn)1/2<m<ϵn,C(n/\log n)^{1/2}<m<\epsilon n, (2.4)

for some large enough C>0C>0 and small enough ϵ>0\epsilon>0.

We prove Theorem 2.3\wrtusdrfthm:sparse-minor by exhibiting a majorant for |PA1S||P_{A}1_{S}| and then using generating function methods.

2.3. Dense minor arcs

Finally we have the dense range, where mcnm\geqslant cn. Here we use quasirandomness. To be precise we define a certain Markov chain on X×YX\times Y, with adjacency operator 𝒜\mathcal{A}, and we consider 𝖫\mathsf{L} to be 𝒜\mathcal{A}-quasirandom with parameter ρ\rho if 𝒜\mathcal{A} has a spectral gap at least 1ρ1-\rho, i.e., if the spectral radius of 𝒜𝒰\mathcal{A}-\mathcal{U} is at most ρ\rho, where 𝒰\mathcal{U} is the projection to constants (the uniform distribution). See Definition 7.1\wrtusdrfdef:quasirandom.

Theorem 2.4.

For every ϵ>0\epsilon>0 there is ρ>0\rho>0 such that if 𝖫\mathsf{L} is 𝒜\mathcal{A}-quasirandom with parameter ρ\rho, then

A[n]|A|ϵn|Λ(PA1S,PA1S,PA1S)|(n!nn)310n.\sum_{\begin{subarray}{c}A\subseteq[n]\\ |A|\geqslant\epsilon n\end{subarray}}|\Lambda(P_{A}1_{S},P_{A}1_{S},P_{A}1_{S})|\leqslant{\left(\frac{n!}{n^{n}}\right)}^{3}10^{-n}.

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 𝖫\mathsf{L} is the multiplication table of a group GG we compute the entire spectrum of 𝒜\mathcal{A} and find ρ=1/D\rho=1/D where DD is the minimal dimension of a nontrivial representation of GG, 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 tr𝒜6=1+o(1)\operatorname{tr}\mathcal{A}^{6}=1+o(1) with high probability, and this implies that ρ=o(1)\rho=o(1).

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.

Let CC and ϵ\epsilon be as in (2.4) and M:=C(n/logn)1/2M:=C(n/\log n)^{1/2}. Theorems 2.22.3 and 2.4 give us that for some c>0c>0

Λ(1S,1S,1S)\displaystyle\Lambda(1_{S},1_{S},1_{S}) =(e1/2+O(M2/n))(n!nn)3\displaystyle=\bigl{(}e^{-1/2}+O(M^{2}/n)\bigr{)}\,{\left(\frac{n!}{n^{n}}\right)}^{3}
+M<mϵnO(ecm)(n!nn)3\displaystyle\qquad+\sum_{M<m\leqslant\epsilon n}O(e^{-cm})\,{\left(\frac{n!}{n^{n}}\right)}^{3}
+10n(n!nn)3,\displaystyle\qquad+10^{-n}{\left(\frac{n!}{n^{n}}\right)}^{3},

as long as 𝖫\mathsf{L} is 𝒜\mathcal{A}-quasirandom with parameter ρ\rho for small enough ρ\rho (depending on ϵ\epsilon). The choice of MM implies (2.1) and hence 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 QAQ_{A}, PAP_{A} 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. (1)

    If AA is a set, ΠA\Pi_{A} is the set of all partitions of AA. If A=[m]A=[m] we will conserve brackets by writing simply Πm\Pi_{m}.

  2. (2)

    ΠA(k)\Pi_{A}^{(k)} is the set of partitions all of whose cells have size at most kk.

  3. (3)

    If ABA\subseteq B then any partition of AA is identified with a partition of BB by adding singletons {{b}:bBA}\bigl{\{}\{b\}\colon b\in B\setminus A\bigr{\}}. With this convention, ΠAΠB\Pi_{A}\subseteq\Pi_{B}.

  4. (4)

    The support suppπ\operatorname{supp}\pi of a partition πΠA\pi\in\Pi_{A} is the union of the nonsingleton cells of π\pi. It is the smallest set BAB\subseteq A such that πΠB\pi\in\Pi_{B}.

  5. (5)

    ΠA\Pi^{\prime}_{A} is the set of πΠA\pi\in\Pi_{A} with suppπ=A\operatorname{supp}\pi=A.

  6. (6)

    If π,πΠA\pi,\pi^{\prime}\in\Pi_{A}, ππ\pi\leqslant\pi^{\prime} means that π\pi is a refinement of π\pi^{\prime} (i.e., every cell of π\pi^{\prime} is a union of cells of π\pi). Synonymously, π\pi^{\prime} is a coarsening of π\pi.

  7. (7)

    The meet ππ\pi\wedge\pi^{\prime} is the coarsest partition refining both π\pi and π\pi^{\prime}; the join ππ\pi\vee\pi^{\prime} is the finest partition coarsening both π\pi and π\pi^{\prime}.

  8. (8)

    The partition {{a}:aA}\{\{a\}:a\in A\} is the discrete partition; the partition {A}\{A\} is the indiscrete partition.

  9. (9)

    The rank of a partition πΠA\pi\in\Pi_{A} is rank(π)=|A||π|\operatorname{rank}(\pi)=|A|-|\pi|; equivalently it is the greatest rr such that there are partitions π0<π1<<πr=π\pi_{0}<\pi_{1}<\cdots<\pi_{r}=\pi. (Note that rank(π)\operatorname{rank}(\pi) is meaningful without specifying AA, unlike |π||\pi|; i.e., it is invariant under adding or removing singletons.)

  10. (10)

    The Möbius function μ\mu at πΠA\pi\in\Pi_{A} is given by μ(π)=(1)rank(π)pπ(|p|1)!\mu(\pi)=(-1)^{\operatorname{rank}(\pi)}\prod_{p\in\pi}(|p|-1)!.

  11. (11)

    A function f:AXf:A\to X is π\pi-measurable if ff is constant on the cells of π\pi. A subset SAS\subseteq A is called π\pi-measurable if 1S1_{S} is π\pi-measurable.

  12. (12)

    If πΠA\pi\in\Pi_{A}, cπL2(XA)c_{\pi}\in L^{2}(X^{A}) is the indicator of π\pi-measurability (i.e., cπ(f)c_{\pi}(f) is 11 if ff is π\pi-measurable, and 0 otherwise).

The exponential formula for partitions states

m01m!πΠmpπx|p|=exp(k11k!xk).\sum_{m\geqslant 0}\frac{1}{m!}\sum_{\pi\in\Pi_{m}}\prod_{p\in\pi}x_{|p|}=\exp{\left(\sum_{k\geqslant 1}\frac{1}{k!}x_{k}\right)}. (3.1)

Here x1,x2,x_{1},x_{2},\dots are formal variables. We will apply (3.1) several times in Section 6\wrtusdrfsec:true-sparse-minor.

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 Λ(cπ1,cπ2,cπ3)\Lambda(c_{\pi_{1}},c_{\pi_{2}},c_{\pi_{3}}) for A[n]A\subseteq[n] and various choices π1,π2,π3ΠA\pi_{1},\pi_{2},\pi_{3}\in\Pi_{A}. This motivates the following definitions.

  1. (1)

    A partition triple on a set AA is a triple 𝔓=(π1,π2,π3)ΠA3\mathfrak{P}=(\pi_{1},\pi_{2},\pi_{3})\in\Pi_{A}^{3}.

  2. (2)

    We call 𝔓\mathfrak{P} a partition system if suppπ1=suppπ2=suppπ3\operatorname{supp}\pi_{1}=\operatorname{supp}\pi_{2}=\operatorname{supp}\pi_{3}.

  3. (3)

    The support of 𝔓\mathfrak{P} is supp𝔓=suppπ1suppπ2suppπ3\operatorname{supp}\mathfrak{P}=\operatorname{supp}\pi_{1}\cup\operatorname{supp}\pi_{2}\cup\operatorname{supp}\pi_{3}.

Definition 3.1 (Combinatorial rank).

Let 𝔓=(π1,π2,π3)ΠA3\mathfrak{P}=(\pi_{1},\pi_{2},\pi_{3})\in\Pi_{A}^{3} be a partition triple. We write S𝔓S\subseteq\mathfrak{P} to mean that Sπ1π2π3S\subseteq\pi_{1}\sqcup\pi_{2}\sqcup\pi_{3}, i.e., SS is a collection of cells labelled 1, 2, or 3. A subset S𝔓S\subseteq\mathfrak{P} is closed (with respect to 𝔓\mathfrak{P}) if whenever piπip_{i}\in\pi_{i} for i=1,2,3i=1,2,3 and p1p2p3p_{1}\cap p_{2}\cap p_{3}\neq\emptyset, if two of p1,p2,p3p_{1},p_{2},p_{3} are in SS then so is the third. The closure S\langle S\rangle of SS is the intersection of all closed sets containing SS. The combinatorial rank of 𝔓=(π1,π2,π3)\mathfrak{P}=(\pi_{1},\pi_{2},\pi_{3}) is defined as

crank(𝔓)=2|A|min{|S|:S𝔓,S=𝔓}.\operatorname{crank}(\mathfrak{P})=2|A|-\min\left\{|S|:S\subseteq\mathfrak{P},\langle S\rangle=\mathfrak{P}\right\}.

The motivation for combinatorial rank is the following bound.

Lemma 3.2.

For a set AA, partitions π1,π2,π3ΠA\pi_{1},\pi_{2},\pi_{3}\in\Pi_{A}, and latin square 𝖫X×Y×Z\mathsf{L}\subseteq X\times Y\times Z,

0Λ(cπ1,cπ2,cπ3)ncrank(π1,π2,π3).0\leqslant\Lambda(c_{\pi_{1}},c_{\pi_{2}},c_{\pi_{3}})\leqslant n^{-\operatorname{crank}(\pi_{1},\pi_{2},\pi_{3})}.

The idea of the proof is the same as for the related result [EMM, Lemma 4.6].

Proof.

The Λ\Lambda value is, by definition, n2|A|n^{-2|A|} times the number of triples of functions

f1:AX,f2:AY,f3:AZf_{1}\colon A\to X,\qquad f_{2}\colon A\to Y,\qquad f_{3}\colon A\to Z

such that fif_{i} is πi\pi_{i}-measurable for i=1,2,3i=1,2,3 and such that (f1(a),f2(a),f3(a))𝖫(f_{1}(a),f_{2}(a),f_{3}(a))\in\mathsf{L} for all aAa\in A. Note we can think of fif_{i} as a function on the cells of πi\pi_{i}, since it is πi\pi_{i}-measurable.

We claim that, given S𝔓S\subseteq\mathfrak{P} with S=𝔓\langle S\rangle=\mathfrak{P}, the triple (f1,f2,f3)(f_{1},f_{2},f_{3}) is determined by the values of fif_{i} on cells in SS. Hence the number of such triples is at most n|S|n^{|S|}, giving the result.

Indeed, suppose f1,f2,f3f^{\prime}_{1},f^{\prime}_{2},f^{\prime}_{3} is another triple of measurable functions with the same restriction to SS. Let W𝔓W\subseteq\mathfrak{P} be the set of all cells piπip_{i}\in\pi_{i} such that fi|pi=fi|pif_{i}|_{p_{i}}=f^{\prime}_{i}|_{p_{i}}. By hypothesis WSW\supseteq S. If piπip_{i}\in\pi_{i} for i=1,2,3i=1,2,3, ap1p2p3a\in p_{1}\cap p_{2}\cap p_{3}, and two of p1,p2,p3p_{1},p_{2},p_{3} are in WW, then so is the third, as the triples (f1(a),f2(a),f3(a)),(f1(a),f2(a),f3(a))𝖫(f_{1}(a),f_{2}(a),f_{3}(a)),(f^{\prime}_{1}(a),f^{\prime}_{2}(a),f^{\prime}_{3}(a))\in\mathsf{L} agree at two coordinates and so are equal by the latin square property. Hence WW is a closed set, so WS=𝔓W\supseteq\langle S\rangle=\mathfrak{P} and fi=fif_{i}=f^{\prime}_{i}, as required. ∎

This reduces the problem of bounding Λ(cπ1,cπ2,cπ3)\Lambda(c_{\pi_{1}},c_{\pi_{2}},c_{\pi_{3}}) from above to the problem of bounding crank(π1,π2,π3)\operatorname{crank}(\pi_{1},\pi_{2},\pi_{3}) from below. In [EMM] we did this using two slightly weaker notions of rank, called triple rank and lower rank, defined respectively as

trank(𝔓)\displaystyle\operatorname{trank}(\mathfrak{P}) =maxσS3(rank(πσ(1))+rank(πσ(2)πσ(3)))\displaystyle=\max_{\sigma\in S_{3}}\bigl{(}\operatorname{rank}(\pi_{\sigma(1)})+\operatorname{rank}(\pi_{\sigma(2)}\vee\pi_{\sigma(3)})\bigr{)}
lrank(𝔓)\displaystyle\operatorname{lrank}(\mathfrak{P}) =(rank(π1)+rank(π2)+rank(π3)+rank(π1π2π3))/2.\displaystyle=\bigl{(}\operatorname{rank}(\pi_{1})+\operatorname{rank}(\pi_{2})+\operatorname{rank}(\pi_{3})+\operatorname{rank}(\pi_{1}\vee\pi_{2}\vee\pi_{3})\bigr{)}/2.
Lemma 3.3.

crank(𝔓)trank(𝔓)lrank(𝔓)\operatorname{crank}(\mathfrak{P})\geqslant\operatorname{trank}(\mathfrak{P})\geqslant\operatorname{lrank}(\mathfrak{P}).

Proof.

For the first inequality, let S𝔓S\subseteq\mathfrak{P} contain all of π1\pi_{1} and one cell of π2\pi_{2} from each cell of π2π3\pi_{2}\vee\pi_{3}. Then |S|=|π1|+|π2π3||S|=|\pi_{1}|+|\pi_{2}\vee\pi_{3}| and S=𝔓\langle S\rangle=\mathfrak{P}, so crank(𝔓)rank(π1)+rank(π2π3)\operatorname{crank}(\mathfrak{P})\geqslant\operatorname{rank}(\pi_{1})+\operatorname{rank}(\pi_{2}\vee\pi_{3}), 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 𝔓\mathfrak{P} to be

cx(𝔓)=trank(𝔓)|supp𝔓|.\operatorname{cx}(\mathfrak{P})=\operatorname{trank}(\mathfrak{P})-|\operatorname{supp}\mathfrak{P}|.

The complexity of a partition system is nonnegative, and it is zero if and only if 𝔓=(π,π,π)\mathfrak{P}=(\pi,\pi,\pi) for some matching π\pi, i.e., a partition of A=suppπA=\operatorname{supp}\pi into |A|/2|A|/2 pairs.

3.3. Combinatorial rank of matching systems

In this subsection we compute crank(π1,π2,π3)\operatorname{crank}(\pi_{1},\pi_{2},\pi_{3}) for all (π1,π2,π3)ΠA(2)(\pi_{1},\pi_{2},\pi_{3})\in\Pi^{(2)}_{A}, i.e., partition triples such that all cells of π1,π2,π3\pi_{1},\pi_{2},\pi_{3} 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 π1,π2,π3ΠA(2)\pi_{1},\pi_{2},\pi_{3}\in\Pi_{A}^{(2)}. Suppose there are precisely kk cells pπ1π2π3p\in\pi_{1}\vee\pi_{2}\vee\pi_{3} such that πi|p\pi_{i}|_{p} has full support (i.e., is a matching) for each i[3]i\in[3]. Then

crank(π1,π2,π3)=rank(π1)+rank(π2)+rank(π3)k.\operatorname{crank}(\pi_{1},\pi_{2},\pi_{3})=\operatorname{rank}(\pi_{1})+\operatorname{rank}(\pi_{2})+\operatorname{rank}(\pi_{3})-k.
Proof.

Since all terms are additive across cells of π1π2π3\pi_{1}\vee\pi_{2}\vee\pi_{3}, we may assume π1π2π3\pi_{1}\vee\pi_{2}\vee\pi_{3} is indiscrete. In particular k{0,1}k\in\{0,1\}, and k=0k=0 if and only if one of π1,π2,π3\pi_{1},\pi_{2},\pi_{3} has a singleton.

Case k=1k=1: In this case π1,π2,π3\pi_{1},\pi_{2},\pi_{3} are matchings, so

rank(π1)=rank(π2)=rank(π3)=|A|/2,\operatorname{rank}(\pi_{1})=\operatorname{rank}(\pi_{2})=\operatorname{rank}(\pi_{3})=|A|/2,

and we must show

crank(π1,π2,π3)=3|A|/21.\operatorname{crank}(\pi_{1},\pi_{2},\pi_{3})=3|A|/2-1.

Let 𝒢\mathcal{G} be the multigraph whose vertex set is AA and with edges given by the cells of π1,π2,π3\pi_{1},\pi_{2},\pi_{3} (which are all 2-cells). Clearly 𝒢\mathcal{G} is 33-regular, with |A||A| vertices and 3|A|/23|A|/2 edges. Since π1π2π3\pi_{1}\vee\pi_{2}\vee\pi_{3} is indiscrete, 𝒢\mathcal{G} 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 11. Hence, equivalently, we want to delete as few edges as possible to get a forest.

Since 𝒢\mathcal{G} has 3|A|/23|A|/2 edges and any forest has at most |A|1|A|-1 edges, we must delete at least |A|/2+1|A|/2+1 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 |A|/2+1|A|/2+1, so crank(π1,π2,π3)=2|A|(|A|/2+1)=3|A|/21\operatorname{crank}(\pi_{1},\pi_{2},\pi_{3})=2|A|-(|A|/2+1)=3|A|/2-1, as claimed.

Case k=0k=0: In this case at least one of π1,π2,π3\pi_{1},\pi_{2},\pi_{3} has a singleton, and we must show that

crank(π1,π2,π3)=rank(π1)+rank(π2)+rank(π3).\operatorname{crank}(\pi_{1},\pi_{2},\pi_{3})=\operatorname{rank}(\pi_{1})+\operatorname{rank}(\pi_{2})+\operatorname{rank}(\pi_{3}).

We define a graph 𝒢\mathcal{G} as in the previous case but additionally for every singleton {v}π1π2π3\{v\}\in\pi_{1}\sqcup\pi_{2}\sqcup\pi_{3} we add an edge {v,}\{v,*\}, where * is a special additional vertex at which infection does not spread. Since π1π2π3\pi_{1}\vee\pi_{2}\vee\pi_{3} is indiscrete, 𝒢\mathcal{G}\setminus* is connected. Since there is at least one singleton, 𝒢\mathcal{G} is connected. Again we want to delete as few edges as possible to get a forest. The number of vertices in 𝒢\mathcal{G} is |A|+1|A|+1 and the number of edges is |π1|+|π2|+|π3||\pi_{1}|+|\pi_{2}|+|\pi_{3}|, so the number of edges we must delete is precisely |π1|+|π2|+|π3||A|.|\pi_{1}|+|\pi_{2}|+|\pi_{3}|-|A|. Hence

crank(π1,π2,π3)=3|A||π1||π2||π3|=rank(π1)+rank(π2)+rank(π3),\operatorname{crank}(\pi_{1},\pi_{2},\pi_{3})=3|A|-|\pi_{1}|-|\pi_{2}|-|\pi_{3}|=\operatorname{rank}(\pi_{1})+\operatorname{rank}(\pi_{2})+\operatorname{rank}(\pi_{3}),

as claimed. ∎

4. The “Fourier” expansion of 1S1_{S}

Recall from Section 2\wrtusdrfsec:outline that QAQ_{A} denotes the orthogonal projection L2(Xm)L2(XA)L^{2}(X^{m})\to L^{2}(X^{A}) and PAP_{A} denotes the orthogonal projection

PA:L2(Xm)L2(XA)BAL2(XB),P_{A}:L^{2}(X^{m})\to L^{2}(X^{A})\cap\bigcap_{B\subsetneq A}L^{2}(X^{B})^{\perp},

and these operators are related via inclusion–exclusion rules:

QA\displaystyle Q_{A} =BAPB,\displaystyle=\sum_{B\subseteq A}P_{B},
PA\displaystyle P_{A} =BA(1)|AB|QB.\displaystyle=\sum_{B\subseteq A}(-1)^{|A\setminus B|}Q_{B}. (4.1)

In this section we study the terms in the expansion

1S=A[n]PA1S.1_{S}=\sum_{A\subseteq[n]}P_{A}1_{S}.

To express some of the results it is convenient to use the linear map U:𝐂[z]𝐂U:\mathbf{C}[z]\to\mathbf{C} defined by

U(zk)={nk/(n)k:kn,0:k>n.U(z^{k})=\begin{cases}n^{k}/(n)_{k}&:k\leqslant n,\\ 0&:k>n.\end{cases}

Here (n)k=n(n1)(nk+1)(n)_{k}=n(n-1)\cdots(n-k+1).

4.1. Formulas for PA1SP_{A}1_{S}

Let SAXAS_{A}\subseteq X^{A} denote the set of injections AXA\to X. Thus if |A|=m|A|=m, |SA|=(n)m|S_{A}|=(n)_{m}.

Lemma 4.1.

If A[n]A\subseteq[n] and |A|=m|A|=m,

QA1S=n!nnnm(n)m1SA.Q_{A}1_{S}=\frac{n!}{n^{n}}\frac{n^{m}}{(n)_{m}}1_{S_{A}}.
Proof.

A function f:AXf\colon A\to X can be extended to a bijection [n]X[n]\to X in (nm)!(n-m)! ways if ff is injective and 0 ways otherwise, and by definition QA1S(f)Q_{A}1_{S}(f) is the number of such extensions normalized by 1/nnm1/n^{n-m}. ∎

Lemma 4.2 ([EMM, Lemma 4.3]).
1SA=πΠAμ(π)cπ.1_{S_{A}}=\sum_{\pi\in\Pi_{A}}\mu(\pi)c_{\pi}.
Lemma 4.3.

If A[n]A\subseteq[n] and |A|=m|A|=m then

PA1S=n!nnnm(n)mπΠAμ(π)PAcπ.P_{A}1_{S}=\frac{n!}{n^{n}}\frac{n^{m}}{(n)_{m}}\sum_{\pi\in\Pi^{\prime}_{A}}\mu(\pi)P_{A}c_{\pi}.
Proof.

Combining the previous two lemmas,

PA1S=PAQA1S=n!nnnm(n)mπΠAμ(π)PAcπ.P_{A}1_{S}=P_{A}Q_{A}1_{S}=\frac{n!}{n^{n}}\frac{n^{m}}{(n)_{m}}\sum_{\pi\in\Pi_{A}}\mu(\pi)P_{A}c_{\pi}.

Since cπL2(Xsuppπ)c_{\pi}\in L^{2}(X^{\operatorname{supp}\pi}), only the terms with suppπ=A\operatorname{supp}\pi=A survive. ∎

We can use UU to give another formula for PA1SP_{A}1_{S}. If xXAx\in X^{A} (i.e., x:AXx:A\to X), the kernel kerxΠA\ker x\in\Pi_{A} of xx is the level set partition

kerx={x1(t):tX,x1(t)}.\ker x=\bigl{\{}x^{-1}(t):t\in X,x^{-1}(t)\neq\emptyset\bigr{\}}.

Note that

cπ(x)=1xis π-measurableπkerx.c_{\pi}(x)=1\iff x~{}\text{is $\pi$-measurable}\iff\pi\leqslant\ker x.
Lemma 4.4.

Let A[n]A\subseteq[n], |A|=m|A|=m. For xXnx\in X^{n}, let π=ker(x|A)ΠA\pi=\ker(x|_{A})\in\Pi_{A}. Then

PA1S(x)=(1)rank(π)n!nnUpπ(|p|z1).P_{A}1_{S}(x)=(-1)^{\operatorname{rank}(\pi)}\frac{n!}{n^{n}}U\prod_{p\in\pi}(|p|z-1).
Proof.

From (4.1) and Lemma 4.1\wrtusdrflem:Q_AS_B we have

PA1S=n!nnBA(1)|AB|n|B|(n)|B|1SB.P_{A}1_{S}=\frac{n!}{n^{n}}\sum_{B\subseteq A}(-1)^{|A\setminus B|}\frac{n^{|B|}}{(n)_{|B|}}1_{S_{B}}.

Now, the sets BB such that x|Bx|_{B} is injective are precisely those which intersect each cell of π\pi in at most one point. Hence

PA1S(x)\displaystyle P_{A}1_{S}(x) =n!nnUBA(1)|AB|z|B|1SB(x)\displaystyle=\frac{n!}{n^{n}}U\sum_{B\subseteq A}(-1)^{|A\setminus B|}z^{|B|}1_{S_{B}}(x)
=n!nn(1)|A||π|Upπ(|p|z1).\displaystyle=\frac{n!}{n^{n}}(-1)^{|A|-|\pi|}U\prod_{p\in\pi}(|p|z-1).\qed

4.2. Sparseval

The word sparseval is our playful term for the computation of PAf22\|P_{A}f\|_{2}^{2} for any A[n]A\subseteq[n]. This is possible by inclusion–exclusion and orthogonality: since

QAf22=BAPBf22,\|Q_{A}f\|_{2}^{2}=\sum_{B\subseteq A}\|P_{B}f\|_{2}^{2},

it follows that

PAf22=BA(1)|AB|QBf22.\|P_{A}f\|_{2}^{2}=\sum_{B\subseteq A}(-1)^{|A\setminus B|}\|Q_{B}f\|_{2}^{2}. (4.2)
Lemma 4.5.

If A[n]A\subseteq[n] and |A|=m|A|=m,

PA1S22=(n!nn)2U((z1)m).\|P_{A}1_{S}\|_{2}^{2}={\left(\frac{n!}{n^{n}}\right)}^{2}U\bigl{(}(z-1)^{m}\bigr{)}.
Proof.

Note that 1SB22=(n)|B|/n|B|\|1_{S_{B}}\|_{2}^{2}=(n)_{|B|}/n^{|B|} for every BAB\subseteq A. Hence, from (4.2) and Lemma 4.1\wrtusdrflem:Q_AS_B,

PA1S22(n!nn)2=BA(1)|AB|n|B|(n)|B|=U((z1)m).\|P_{A}1_{S}\|_{2}^{2}{\left(\frac{n!}{n^{n}}\right)}^{-2}=\sum_{B\subseteq A}(-1)^{|A\setminus B|}\frac{n^{|B|}}{(n)_{|B|}}=U\bigl{(}(z-1)^{m}\bigr{)}.

Proposition 4.6.

Assume 0mn0\leqslant m\leqslant n and let t=m/nt=m/n. Then

0U((z1)m)(nm)1es(t)n,0\leqslant U\bigl{(}(z-1)^{m}\bigr{)}\ll\binom{n}{m}^{-1}e^{s(t)n},

where

s(t)=t1/2tlogt1/2(1t)log(1+t1/2).s(t)=t^{1/2}-t\log t^{1/2}-(1-t)\log(1+t^{1/2}).

In particular

U((z1)m)eO(m)(m/n)m/2.U\bigl{(}(z-1)^{m}\bigr{)}\leqslant e^{O(m)}(m/n)^{m/2}.
Sketch.

The inequality U((z1)m)0U\bigl{(}(z-1)^{m}\bigr{)}\geqslant 0 follows from the previous lemma. For the main claim, by expanding we have

U((z1)m)=1n!k=0m(mk)(1)mknk(nk)!,U\bigl{(}(z-1)^{m}\bigr{)}=\frac{1}{n!}\sum_{k=0}^{m}\binom{m}{k}(-1)^{m-k}n^{k}(n-k)!,

and this can be identified as (nm)1\binom{n}{m}^{-1} times the coefficient of XmX^{m} in enX/(1+X)nm+1e^{nX}/(1+X)^{n-m+1}. 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 tt near 1, but we omit the details because we will not use the claim for t>1/2t>1/2. 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 PA1S(x)P_{A}1_{S}(x) is (1)rank(kerx|A)(-1)^{\operatorname{rank}(\ker x|_{A})}.

Proof.

Let π=ker(x|A)\pi=\ker(x|_{A}). By Lemma 4.4\wrtusdrfprop:PA1S-sparseval-like-expression it suffices to prove that

Upπ(|p|z1)>0.U\prod_{p\in\pi}(|p|z-1)>0.

There are nonnegative integers rω0r_{\omega}\geqslant 0 such that

pπ(|p|z1)=pπ(|p|(z1)+(|p|1))=ωπrω(z1)|ω|.\prod_{p\in\pi}(|p|z-1)=\prod_{p\in\pi}\bigl{(}|p|(z-1)+(|p|-1)\bigr{)}=\sum_{\omega\subseteq\pi}r_{\omega}(z-1)^{|\omega|}.

Hence the claim follows from U((z1)m)0U\bigl{(}(z-1)^{m}\bigr{)}\geqslant 0. ∎

5. Major arcs

The goal in this section is to prove Theorem 2.2. Define

𝔖m=2km(1)k2kk!,\displaystyle\mathfrak{S}_{m}=\sum_{2k\leqslant m}\frac{(-1)^{k}}{2^{k}k!},
Mm=A[n]|A|mΛ(PA1S,PA1S,PA1S).\displaystyle M_{m}=\sum_{\begin{subarray}{c}A\subseteq[n]\\ |A|\leqslant m\end{subarray}}\Lambda(P_{A}1_{S},P_{A}1_{S},P_{A}1_{S}).

Our aim is to prove that, for mcn1/2m\leqslant cn^{1/2},

Mm=(𝔖m+O(m2/n))(n!nn)3.M_{m}=\bigl{(}\mathfrak{S}_{m}+O(m^{2}/n)\bigr{)}{\left(\frac{n!}{n^{n}}\right)}^{3}. (5.1)

In particular this implies Theorem 2.2\wrtusdrfthm:major-arcs.

5.1. The quantities γ\gamma and γ0\gamma_{0}

From Lemma 4.3\wrtusdrflem:P_A-on-1_S it is clear that to estimate Λ(PA1S,PA1S,PA1S)\Lambda(P_{A}1_{S},P_{A}1_{S},P_{A}1_{S}) it suffices to estimate Λ(PAcπ1,PAcπ2,PAcπ3)\Lambda(P_{A}c_{\pi_{1}},P_{A}c_{\pi_{2}},P_{A}c_{\pi_{3}}) for every partition system 𝔓=(π1,π2,π3)\mathfrak{P}=(\pi_{1},\pi_{2},\pi_{3}) with support AA and aggregate the results with the appropriate weighting. For continuity with [EMM, Section 4], we define the normalized quantities

γ0(𝔓)=ntrank(𝔓)Λ(cπ1,cπ2,cπ3)\gamma_{0}(\mathfrak{P})=n^{\operatorname{trank}(\mathfrak{P})}\Lambda(c_{\pi_{1}},c_{\pi_{2}},c_{\pi_{3}})

and

γ(𝔓)=ntrank(𝔓)Λ(PAcπ1,PAcπ2,PAcπ3)\gamma(\mathfrak{P})=n^{\operatorname{trank}(\mathfrak{P})}\Lambda(P_{A}c_{\pi_{1}},P_{A}c_{\pi_{2}},P_{A}c_{\pi_{3}})

for any partition triple 𝔓=(π1,π2,π3)\mathfrak{P}=(\pi_{1},\pi_{2},\pi_{3}). Note that

0γ0(𝔓)10\leqslant\gamma_{0}(\mathfrak{P})\leqslant 1

by Lemmas 3.2 and 3.3\wrtusdrflem:crank-bound,lem:crank¿=trank. Since cπL2(Xsuppπ)c_{\pi}\in L^{2}(X^{\operatorname{supp}\pi}), γ(𝔓)=0\gamma(\mathfrak{P})=0 unless 𝔓\mathfrak{P} is a partition system.

Lemma 5.1.

Let 𝔓\mathfrak{P} be a partition system with support supp𝔓=A\operatorname{supp}\mathfrak{P}=A of size mm, and suppose mm^{\prime} points of AA are contained in cells π1π2π3\pi_{1}\vee\pi_{2}\vee\pi_{3} of size at least 33. Then

|γ(𝔓)|2m.|\gamma(\mathfrak{P})|\leqslant 2^{m^{\prime}}.
Sketch.

The idea is that

Λ(PAcπ1,PAcπ2,PAcπ3)\displaystyle\Lambda(P_{A}c_{\pi_{1}},P_{A}c_{\pi_{2}},P_{A}c_{\pi_{3}}) =Λ(cπ1,cπ2,PAcπ3)\displaystyle=\Lambda(c_{\pi_{1}},c_{\pi_{2}},P_{A}c_{\pi_{3}})
=BA(1)|AB|Λ(cπ1,cπ2,QBcπ3)\displaystyle=\sum_{B\subseteq A}(-1)^{|A\setminus B|}\Lambda(c_{\pi_{1}},c_{\pi_{2}},Q_{B}c_{\pi_{3}})
=BA(1)|AB|Λ(QBcπ1,QBcπ2,QBcπ3),\displaystyle=\sum_{B\subseteq A}(-1)^{|A\setminus B|}\Lambda(Q_{B}c_{\pi_{1}},Q_{B}c_{\pi_{2}},Q_{B}c_{\pi_{3}}),

and

QBcπ=nrank(π)+rank(π|B)cπ|B,Q_{B}c_{\pi}=n^{-\operatorname{rank}(\pi)+\operatorname{rank}(\pi|_{B})}c_{\pi|_{B}},

where

π|B={pB:pπ,pB}.\pi|_{B}=\{p\cap B:p\in\pi,p\cap B\neq\emptyset\}.

Let 𝔓|B=(π1|B,π2|B,π3|B)\mathfrak{P}|_{B}=(\pi_{1}|_{B},\pi_{2}|_{B},\pi_{3}|_{B}). Then, normalizing,

γ(𝔓)=BA(1)|AB|γ0(𝔓|B)nt(𝔓,B),\gamma(\mathfrak{P})=\sum_{B\subseteq A}(-1)^{|A\setminus B|}\gamma_{0}(\mathfrak{P}|_{B})n^{-t(\mathfrak{P},B)},

where

t(𝔓,B)=trank(𝔓|B)trank(𝔓)+i=13(rank(πi)rank(πi|B)).t(\mathfrak{P},B)=\operatorname{trank}(\mathfrak{P}|_{B})-\operatorname{trank}(\mathfrak{P})+\sum_{i=1}^{3}\bigl{(}\operatorname{rank}(\pi_{i})-\operatorname{rank}(\pi_{i}|_{B})\bigr{)}.

In [EMM, Section 4] we showed t(𝔓,B)0t(\mathfrak{P},B)\geqslant 0. Since γ0(𝔓|B)[0,1]\gamma_{0}(\mathfrak{P}|_{B})\in[0,1] for all BB this shows |γ(𝔓)|2m|\gamma(\mathfrak{P})|\leqslant 2^{m}. The stronger bound with mm^{\prime} in place of mm follows by separating off the doubleton cells of π1π2π3\pi_{1}\vee\pi_{2}\vee\pi_{3}. See [EMM, Section 4] for details. ∎

5.2. The Mm(z)M_{m}(z) series

For a partition triple 𝔓=(π1,π2,π3)\mathfrak{P}=(\pi_{1},\pi_{2},\pi_{3}) we use the shorthand

μ(𝔓)=μ(π1)μ(π2)μ(π3).\mu(\mathfrak{P})=\mu(\pi_{1})\,\mu(\pi_{2})\,\mu(\pi_{3}).

From Lemma 4.3 we have

Mm=(n!nn)3|supp𝔓|m(n|supp𝔓|(n)|supp𝔓|)3μ(𝔓)γ(𝔓)ntrank(𝔓),M_{m}={\left(\frac{n!}{n^{n}}\right)}^{3}\sum_{|\operatorname{supp}\mathfrak{P}|\leqslant m}{\left(\frac{n^{|\operatorname{supp}\mathfrak{P}|}}{(n)_{|\operatorname{supp}\mathfrak{P}|}}\right)}^{3}\mu(\mathfrak{P})\gamma(\mathfrak{P})n^{-\operatorname{trank}(\mathfrak{P})},

where the sum is over all partition systems on [n][n]. For z𝐂z\in\mathbf{C} define

Mm(z)=(n!nn)3|supp𝔓|m(n|supp𝔓|(n)|supp𝔓|)3μ(𝔓)γ(𝔓)n|supp𝔓|zcx(𝔓).M_{m}(z)={\left(\frac{n!}{n^{n}}\right)}^{3}\sum_{|\operatorname{supp}\mathfrak{P}|\leqslant m}{\left(\frac{n^{|\operatorname{supp}\mathfrak{P}|}}{(n)_{|\operatorname{supp}\mathfrak{P}|}}\right)}^{3}\mu(\mathfrak{P})\gamma(\mathfrak{P})n^{-|\operatorname{supp}\mathfrak{P}|}z^{\operatorname{cx}(\mathfrak{P})}.

As we have mentioned, cx(𝔓)0\operatorname{cx}(\mathfrak{P})\geqslant 0 for any partition system 𝔓\mathfrak{P}, so Mm(z)M_{m}(z) is a polynomial such that Mm=Mm(1/n)M_{m}=M_{m}(1/n). By bounding Mm(z)M_{m}(z) and using some complex analysis we will show Mm(1/n)Mm(0)M_{m}(1/n)\approx M_{m}(0), and then we will directly estimate Mm(0)M_{m}(0).

Proposition 5.2.

There is a constant c>0c>0 such that, for |z|1/2c/m|z|^{1/2}\leqslant c/m, we have

|Mm(z)|(nm(n)m)2(n!nn)3.|M_{m}(z)|\ll{\left(\frac{n^{m}}{(n)_{m}}\right)}^{2}{\left(\frac{n!}{n^{n}}\right)}^{3}.
Proof.

By the definition of Mm(z)M_{m}(z), the triangle inequality, and Lemma 5.1\wrtusdrflem:trivial-gamma-bound, the quantity |Mm(z)|/(n!nn)3|M_{m}(z)|/{\left(\frac{n!}{n^{n}}\right)}^{3} is bounded by

|A|mn|A|(n|A|(n)|A|)3supp𝔓=A2m(𝔓)|μ(𝔓)||z|cx(𝔓),\sum_{|A|\leqslant m}n^{-|A|}{\left(\frac{n^{|A|}}{(n)_{|A|}}\right)}^{3}\sum_{\operatorname{supp}\mathfrak{P}=A}2^{m^{\prime}(\mathfrak{P})}|\mu(\mathfrak{P})|\ |z|^{\operatorname{cx}(\mathfrak{P})},

where m(𝔓)m^{\prime}(\mathfrak{P}) is the number of points of supp𝔓\operatorname{supp}\mathfrak{P} contained in cells of π1π2π3\pi_{1}\vee\pi_{2}\vee\pi_{3} of size at least 3. This exact sum was analyzed in [EMM, Section 4.4], and we showed that it is O(nm/(n)m)2O(n^{m}/(n)_{m})^{2} provided |z|1/2<c/m|z|^{1/2}<c/m. The proposition follows. ∎

Corollary 5.3.

There is a constant c>0c>0 such that, for m<cn1/2m<cn^{1/2},

|MmMm(0)|(m2/n)(n!nn)3.|M_{m}-M_{m}(0)|\ll(m^{2}/n){\left(\frac{n!}{n^{n}}\right)}^{3}.
Proof.

By the residue theorem,

Mm(u)Mm(0)=12πi|z|=RMm(z)u(zu)z𝑑zM_{m}(u)-M_{m}(0)=\frac{1}{2\pi i}\oint_{|z|=R}\frac{M_{m}(z)u}{(z-u)z}\,dz

as long as |u|<R|u|<R. Hence

|Mm(u)Mm(0)|max|z|=R|Mm(z)||u|/R1|u|/R.\left|M_{m}(u)-M_{m}(0)\right|\leqslant\max_{|z|=R}|M_{m}(z)|\frac{|u|/R}{1-|u|/R}.

Take u=1/nu=1/n and R=c2/m2R=c^{2}/m^{2}, where cc is as in the previous proposition. Then as long as 1/n<c2/m21/n<c^{2}/m^{2}, i.e., m<cn1/2m<cn^{1/2}, we get

|MmMm(0)|(1+n1/2)2m(nm(n)m)2(n!nn)3m2/n1m2/(c2n).|M_{m}-M_{m}(0)|\ll(1+n^{-1/2})^{2m}{\left(\frac{n^{m}}{(n)_{m}}\right)}^{2}{\left(\frac{n!}{n^{n}}\right)}^{3}\frac{m^{2}/n}{1-m^{2}/(c^{2}n)}.

Hence as long as say m<(c/2)n1/2m<(c/2)n^{1/2} we get the claimed bound. ∎

5.3. The constant term Mm(0)M_{m}(0)

By definition,

Mm(0)=(n!nn)3|supp𝔓|mcx(𝔓)=0(n|supp𝔓|(n)|supp𝔓|)3μ(𝔓)γ(𝔓)n|supp𝔓|.M_{m}(0)={\left(\frac{n!}{n^{n}}\right)}^{3}\sum_{\begin{subarray}{c}|\operatorname{supp}\mathfrak{P}|\leqslant m\\ \operatorname{cx}(\mathfrak{P})=0\end{subarray}}{\left(\frac{n^{|\operatorname{supp}\mathfrak{P}|}}{(n)_{|\operatorname{supp}\mathfrak{P}|}}\right)}^{3}\mu(\mathfrak{P})\gamma(\mathfrak{P})n^{-|\operatorname{supp}\mathfrak{P}|}.

As remarked, cx(𝔓)=0\operatorname{cx}(\mathfrak{P})=0 if and only if 𝔓=(π,π,π)\mathfrak{P}=(\pi,\pi,\pi) for some matching π\pi. In this case, if say |supp𝔓|=2k|\operatorname{supp}\mathfrak{P}|=2k,

n|supp𝔓|(n)|supp𝔓|=n2k(n)2k=1+O(k2/n),\displaystyle\frac{n^{|\operatorname{supp}\mathfrak{P}|}}{(n)_{|\operatorname{supp}\mathfrak{P}|}}=\frac{n^{2k}}{(n)_{2k}}=1+O(k^{2}/n),
μ(𝔓)=μ(π)3=(1)k,\displaystyle\mu(\mathfrak{P})=\mu(\pi)^{3}=(-1)^{k},
γ(𝔓)=(11/n)k.\displaystyle\gamma(\mathfrak{P})=(1-1/n)^{k}.

The last identity holds by a direct calculation analogous to [EMM, Lemma 4.10]. The number of matchings π\pi in [n][n] of support size 2k2k is

(n)2k2kk!=n2k2kk!(1+O(k2/n)).\frac{(n)_{2k}}{2^{k}k!}=\frac{n^{2k}}{2^{k}k!}\bigl{(}1+O(k^{2}/n)\bigr{)}.

Thus

Mm(0)\displaystyle M_{m}(0) =(n!nn)3k=0m/2n2k2kk!(1)kn2k(1+O(k2/n))\displaystyle={\left(\frac{n!}{n^{n}}\right)}^{3}\sum_{k=0}^{\left\lfloor{m/2}\right\rfloor}\frac{n^{2k}}{2^{k}k!}(-1)^{k}n^{-2k}\bigl{(}1+O(k^{2}/n)\bigr{)}
=(n!nn)3(𝔖m+O(1/n)).\displaystyle={\left(\frac{n!}{n^{n}}\right)}^{3}\bigl{(}\mathfrak{S}_{m}+O(1/n)\bigr{)}.

By combining with Corollary 5.3 we have

Mm=(n!nn)3(𝔖m+O(m2/n))M_{m}={\left(\frac{n!}{n^{n}}\right)}^{3}\bigl{(}\mathfrak{S}_{m}+O(m^{2}/n)\bigr{)}

provided m<cn1/2m<cn^{1/2}. This finishes the proof of (5.1).

6. Sparse minor arcs

To prove Theorem 2.3\wrtusdrfthm:sparse-minor we need a bound on Λ(|PA1S|,|PA1S|,|PA1S|)\Lambda(|P_{A}1_{S}|,|P_{A}1_{S}|,|P_{A}1_{S}|) for larger |A||A|. Note that in any latin square 𝖫(X,Y,Z)\mathsf{L}^{\prime}\subseteq(X^{\prime},Y^{\prime},Z^{\prime}),

|Λ(f,g,h)|\displaystyle|\Lambda(f,g,h)| =|𝐄(x,y,z)𝖫f(x)g(y)h(z)|\displaystyle=\lvert\mathbf{E}_{(x,y,z)\in\mathsf{L}^{\prime}}f(x)g(y)h(z)\rvert
𝐄xX|f(x)||𝐄y,z:(x,y,z)𝖫g(y)h(z)|f1g2h2\displaystyle\leqslant\mathbf{E}_{x\in X^{\prime}}|f(x)|\left\lvert\mathbf{E}_{y,z\colon(x,y,z)\in\mathsf{L}^{\prime}}g(y)h(z)\right\rvert\leqslant\|f\|_{1}\|g\|_{2}\|h\|_{2} (6.1)

using the latin square property and Cauchy–Schwarz, and similarly permuting f,g,hf,g,h. One approach to Theorem 2.3\wrtusdrfthm:sparse-minor might be to find upper bounds on |PA1S(x)||P_{A}1_{S}(x)|, pointwise or in L1L^{1}, 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 |PA1S||P_{A}1_{S}| of the form

|PA1S|πΠAtπcπ|P_{A}1_{S}|\leqslant\sum_{\pi\in\Pi_{A}}t_{\pi}c_{\pi} (6.2)

for some coefficients tπ0t_{\pi}\geqslant 0. Then

Λ(|PA1S|,|PA1S|,|PA1S|)π1,π2,π3ΠAtπ1tπ2tπ3Λ(cπ1,cπ2,cπ3)\Lambda(|P_{A}1_{S}|,|P_{A}1_{S}|,|P_{A}1_{S}|)\leqslant\sum_{\pi_{1},\pi_{2},\pi_{3}\in\Pi_{A}}t_{\pi_{1}}t_{\pi_{2}}t_{\pi_{3}}\Lambda(c_{\pi_{1}},c_{\pi_{2}},c_{\pi_{3}})

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 πiΠA(2)\pi_{i}\in\Pi_{A}^{(2)}, given Lemma 3.4\wrtusdrfclaim:cool-crank-value.

Again this approach does not succeed by itself. Our final argument works by decomposing |PA1S||P_{A}1_{S}| into two pieces and combining the two techniques discussed above.

6.1. A majorant for |PA1S||P_{A}1_{S}|

Throughout this section let C>0C>0 be some large enough constant, A[n]A\subseteq[n] and |A|=mn/C|A|=m\leqslant n/C. Additionally, we let

δ:=(Cm/n)1/2.\delta:=(Cm/n)^{1/2}.

Although we will always have this specific value of δ\delta in mind, most of the results in this section only rely on δ1\delta\leqslant 1. The next proposition gives a useful bound for |PA1S||P_{A}1_{S}|. For δ:=(Cm/n)1/2\delta:=(Cm/n)^{1/2}, r1r\geqslant 1, and π\pi a partition define

σr(δ)\displaystyle\sigma_{r}^{(\delta)} ={δ:r=1,r1:r>1,\displaystyle=\begin{cases}\delta&:r=1,\\ r-1&:r>1,\end{cases}
σπ(δ)\displaystyle\sigma_{\pi}^{(\delta)} =pπσ|p|(δ).\displaystyle=\prod_{p\in\pi}\sigma_{|p|}^{(\delta)}.
Proposition 6.1.

We have

|PA1S(x)|n!nneδmσkerx(δ)(xXA).|P_{A}1_{S}(x)|\leqslant\frac{n!}{n^{n}}e^{\delta m}\sigma_{\ker x}^{(\delta)}\qquad(x\in X^{A}).
Proof.

From Lemma 4.4\wrtusdrfprop:PA1S-sparseval-like-expression,

PA1S(x)=(1)rank(π)n!nnUϕ,P_{A}1_{S}(x)=(-1)^{\operatorname{rank}(\pi)}\frac{n!}{n^{n}}U\phi,

where π=kerx\pi=\ker x and

ϕ\displaystyle\phi =pπ(|p|z1)=ωπrω(z1)|ω|,\displaystyle=\prod_{p\in\pi}(|p|z-1)=\sum_{\omega\subseteq\pi}r_{\omega}(z-1)^{|\omega|},
rω\displaystyle r_{\omega} =pπω(|p|1)pω|p|.\displaystyle=\prod_{p\in\pi\setminus\omega}(|p|-1)\prod_{p\in\omega}|p|.

From Proposition 4.6\wrtusdrfprop:umbral-sparseval and crude estimates (Stirling’s formula), for 0dm0\leqslant d\leqslant m,

U((z1)d)(Cd/n)d/2(Cm/n)d/2=δdU\bigl{(}(z-1)^{d}\bigr{)}\leqslant(Cd/n)^{d/2}\leqslant(Cm/n)^{d/2}=\delta^{d}

provided CC is large enough. Then

Uϕ\displaystyle U\phi ωπrωδ|ω|\displaystyle\leqslant\sum_{\omega\subseteq\pi}r_{\omega}\delta^{|\omega|}
=pπ(|p|1+|p|δ)\displaystyle=\prod_{p\in\pi}(|p|-1+|p|\delta)
=σπ(δ)pπ:|p|>1(1+|p||p|1δ)\displaystyle=\sigma_{\pi}^{(\delta)}\prod_{p\in\pi:|p|>1}{\left(1+\frac{|p|}{|p|-1}\delta\right)}
σπ(δ)(1+2δ)m/2\displaystyle\leqslant\sigma_{\pi}^{(\delta)}{\left(1+2\delta\right)}^{m/2}
σπ(δ)eδm\displaystyle\leqslant\sigma_{\pi}^{(\delta)}e^{\delta m}

as required. ∎

In light of the proposition, to find majorants for |PA1S||P_{A}1_{S}| of the form (6.2) it suffices to find analogous bounds for σπ(δ)\sigma_{\pi}^{(\delta)}. Recall that ΠA(k)\Pi_{A}^{(k)} is the set of all πΠA\pi\in\Pi_{A} having no part of size greater than kk. Let rk(π)r_{k}(\pi) be the number of kk-cells in π\pi and let r3+(π)=k3rk(π)r_{3+}(\pi)=\sum_{k\geqslant 3}r_{k}(\pi).

Lemma 6.2.

Let π\pi be a partition.

  1. (1)
    σπ(δ){σπ(δ):ππ,πΠ(3)}.\sigma_{\pi}^{(\delta)}\leqslant\sum\Bigl{\{}\sigma_{\pi^{\prime}}^{(\delta)}:\pi^{\prime}\leqslant\pi,\pi^{\prime}\in\Pi^{(3)}\Bigr{\}}.
  2. (2)
    σπ(δ){σπ(δ):ππ,πΠ(4),r3+(π)=r3+(π)}.\sigma_{\pi}^{(\delta)}\leqslant\sum\Bigl{\{}\sigma_{\pi^{\prime}}^{(\delta)}:\pi^{\prime}\leqslant\pi,\pi^{\prime}\in\Pi^{(4)},r_{3+}(\pi^{\prime})=r_{3+}(\pi)\Bigr{\}}.
  3. (3)
    σπ(δ)δr3+(π){σπ(δ):ππ,πΠ(2)}.\sigma_{\pi}^{(\delta)}\leqslant\delta^{-r_{3+}(\pi)}\sum\Bigl{\{}\sigma_{\pi^{\prime}}^{(\delta)}:\pi^{\prime}\leqslant\pi,\pi^{\prime}\in\Pi^{(2)}\Bigr{\}}.
Proof.

Consider the first inequality. Both sides are multiplicative across cells of π\pi, so we may assume π\pi is a single cell, say of size rr. The inequality is trivial for r3r\leqslant 3 (since σπ(δ)\sigma_{\pi}^{(\delta)} is one of the summands on the right-hand side), so we may assume r4r\geqslant 4. Then it suffices to check

r12a+3b=rr!2!aa!3!bb!2b.r-1\leqslant\sum_{\begin{subarray}{c}2a+3b=r\end{subarray}}\frac{r!}{2!^{a}a!3!^{b}b!}2^{b}.

This is a calculation for r10r\leqslant 10 (say) and an uninteresting exercise for r>10r>10.

Now consider the second inequality. This time, the right-hand side is not itself multiplicative over cells of π\pi, but if we replace the condition r3+(π)=r3+(π)r_{3+}(\pi^{\prime})=r_{3+}(\pi) by the stronger one

pπ,|p|3:there is exactly one pπ with pp and |p|3\forall p\in\pi,\ |p|\geqslant 3:\ \text{there is exactly one }p^{\prime}\in\pi^{\prime}\text{ with }p^{\prime}\subseteq p\text{ and }|p^{\prime}|\geqslant 3

then it becomes so, and it suffices to prove the corresponding stronger inequality. Now we may again assume that π\pi is an rr-cell, and we may assume r5r\geqslant 5. Then we must check

r12a+3b+4c=rb+c=1r!2!aa!3!bb!4!cc!2b3c.r-1\leqslant\sum_{\begin{subarray}{c}2a+3b+4c=r\\ b+c=1\end{subarray}}\frac{r!}{2!^{a}a!3!^{b}b!4!^{c}c!}2^{b}3^{c}.

Again we omit further details.

Now consider the third inequality. Again it suffices to consider the case of an rr-cell, and we may assume r3r\geqslant 3. Then the assertion is

r1δ1a+2b=rr!a!2!bb!δa.r-1\leqslant\delta^{-1}\sum_{a+2b=r}\frac{r!}{a!2!^{b}b!}\delta^{a}.

Since δ1\delta\leqslant 1, it suffices to check

r1a+2b=ra1r!a!2!bb!,r-1\leqslant\sum_{\begin{subarray}{c}a+2b=r\\ a\leqslant 1\end{subarray}}\frac{r!}{a!2!^{b}b!},

which is again essentially a calculation. ∎

Lemma 6.3.

Let xXAx\in X^{A}. Then

|PA1SA(x)|n!nneδmπΠA(3)σπ(δ)cπ(x).|P_{A}1_{S_{A}}(x)|\leqslant\frac{n!}{n^{n}}e^{\delta m}\sum_{\pi\in\Pi_{A}^{(3)}}\sigma_{\pi}^{(\delta)}c_{\pi}(x).
Proof.

By Proposition 6.1\wrtusdrfprop:PA1S-physical-bound,

|PA1SA(x)|n!nneδmσkerx(δ).|P_{A}1_{S_{A}}(x)|\leqslant\frac{n!}{n^{n}}e^{\delta m}\sigma_{\ker x}^{(\delta)}.

By Lemma 6.2(1),

σkerx(δ){σπ(δ):πkerx,πΠA(3)}=πΠA(3)σπ(δ)cπ(x).\sigma_{\ker x}^{(\delta)}\leqslant\sum\Bigl{\{}\sigma_{\pi}^{(\delta)}:\pi\leqslant\ker x,\pi\in\Pi_{A}^{(3)}\Bigr{\}}=\sum_{\pi\in\Pi_{A}^{(3)}}\sigma_{\pi}^{(\delta)}c_{\pi}(x).\qed

6.2. A splitting of |PA1S||P_{A}1_{S}|

We can use the bound on |PA1S||P_{A}1_{S}| given in the previous section to bound the L1L^{1} norm of PA1SP_{A}1_{S}, but the bound would not be strong enough for what we need. To go further, we break up R:=|PA1S|R:=|P_{A}1_{S}| into two parts, a part whose L1L^{1} norm we can control better, and a part we can analyze separately. Fix ϵ0\epsilon\geqslant 0 and let

Π={πΠA:r3+(π)<ϵm}.\Pi^{\sharp}=\{\pi\in\Pi_{A}:r_{3+}(\pi)<\epsilon m\}.

Let Π=ΠAΠ\Pi^{\flat}=\Pi_{A}\setminus\Pi^{\sharp}. Define

R(x)\displaystyle R^{\sharp}(x) =1Π(kerx)R(x),\displaystyle=1_{\Pi^{\sharp}}(\ker x)R(x),
R(x)\displaystyle R^{\flat}(x) =1Π(kerx)R(x).\displaystyle=1_{\Pi^{\flat}}(\ker x)R(x).

Clearly R=R+RR=R^{\sharp}+R^{\flat}.

Lemma 6.4.

We have

R\displaystyle R^{\flat} n!nneδmπΠΠ(4)σπ(δ)cπ,\displaystyle\leqslant\frac{n!}{n^{n}}e^{\delta m}\sum_{\pi\in\Pi^{\flat}\cap\Pi^{(4)}}\sigma_{\pi}^{(\delta)}c_{\pi},
R\displaystyle R^{\sharp} n!nneδmπΠΠ(4)σπ(δ)cπ,\displaystyle\leqslant\frac{n!}{n^{n}}e^{\delta m}\sum_{\pi\in\Pi^{\sharp}\cap\Pi^{(4)}}\sigma_{\pi}^{(\delta)}c_{\pi},
R\displaystyle R^{\sharp} n!nneδmδϵmπΠ(2)σπ(δ)cπ.\displaystyle\leqslant\frac{n!}{n^{n}}e^{\delta m}\delta^{-\epsilon m}\sum_{\pi\in\Pi^{(2)}}\sigma_{\pi}^{(\delta)}c_{\pi}.
Proof.

By Proposition 6.1\wrtusdrfprop:PA1S-physical-bound,

R(x)n!nneδmσkerx(δ).R(x)\leqslant\frac{n!}{n^{n}}e^{\delta m}\sigma_{\ker x}^{(\delta)}.

Suppose kerxΠ\ker x\in\Pi^{\flat}. Then by Lemma 6.2(2),

σkerx{σπ:πkerx,πΠ(4),r3+(π)ϵm}.\sigma_{\ker x}\leqslant\sum\{\sigma_{\pi}:\pi\leqslant\ker x,\pi\in\Pi^{(4)},r_{3+}(\pi)\geqslant\epsilon m\}.

This proves the bound on RR^{\flat}. The first bound on RR^{\sharp} is proved identically. The second is proved in the same way using instead Lemma 6.2(3). ∎

Corollary 6.5.

We have

R1n!nneO(m)(m/n)(1+ϵ)m/2.\|R^{\flat}\|_{1}\ll\frac{n!}{n^{n}}e^{O(m)}(m/n)^{(1+\epsilon)m/2}.
Proof.

Using the previous lemma, δ1\delta\leqslant 1 and cπ1=nrank(π)\|c_{\pi}\|_{1}=n^{-\operatorname{rank}(\pi)},

R1n!nnemπΠΠ(4)σπ(δ)nrank(π).\|R^{\flat}\|_{1}\leqslant\frac{n!}{n^{n}}e^{m}\sum_{\pi\in\Pi^{\flat}\cap\Pi^{(4)}}\sigma_{\pi}^{(\delta)}n^{-\operatorname{rank}(\pi)}.

Let

αr(x,w)=πΠr(4)σπ(δ)xrank(π)wr3+(π).\alpha_{r}(x,w)=\sum_{\pi\in\Pi_{r}^{(4)}}\sigma_{\pi}^{(\delta)}x^{\operatorname{rank}(\pi)}w^{r_{3+}(\pi)}.

Then, for real w1w\geqslant 1,

πΠΠ(4)σπ(δ)nrank(π)wϵmαm(1/n,w).\sum_{\pi\in\Pi^{\flat}\cap\Pi^{(4)}}\sigma_{\pi}^{(\delta)}n^{-\operatorname{rank}(\pi)}\leqslant w^{-\epsilon m}\alpha_{m}(1/n,w).

Using the exponential formula (3.1) with xk=σk(δ)xk1ykx_{k}=\sigma_{k}^{(\delta)}x^{k-1}y^{k} for k=1,2k=1,2, xk=σk(δ)wxk1ykx_{k}=\sigma_{k}^{(\delta)}wx^{k-1}y^{k} for k=3,4k=3,4, and xk=0x_{k}=0 for k5k\geqslant 5, we obtain

r01r!αr(x,w)yr=exp(δy+xy2/2+wx2y3/3+wx3y4/8).\sum_{r\geqslant 0}\frac{1}{r!}\alpha_{r}(x,w)y^{r}=\exp(\delta y+xy^{2}/2+wx^{2}y^{3}/3+wx^{3}y^{4}/8).

Hence for real y>0y>0,

wϵmαm(x,w)m!wϵmymexp(δy+xy2/2+wx2y3/3+wx3y4/8).w^{-\epsilon m}\alpha_{m}(x,w)\leqslant\frac{m!}{w^{\epsilon m}y^{m}}\exp(\delta y+xy^{2}/2+wx^{2}y^{3}/3+wx^{3}y^{4}/8).

Putting x=1/nx=1/n, y=(mn)1/2y=(mn)^{1/2}, and w=(n/m)1/2w=(n/m)^{1/2}, we get

wϵmαm(1/n,w)m!(n/m)ϵm/2(mn)m/2eO(m).w^{-\epsilon m}\alpha_{m}(1/n,w)\leqslant\frac{m!}{(n/m)^{\epsilon m/2}(mn)^{m/2}}e^{O(m)}.

This proves what we want. ∎

Corollary 6.6.

We have

Λ(R,R,R)Λ(R,R,R)+(n!nn)3O(1)m(m/n)(1+ϵ/2)m.\Lambda(R,R,R)\leqslant\Lambda(R^{\sharp},R^{\sharp},R^{\sharp})+{\left(\frac{n!}{n^{n}}\right)}^{3}O(1)^{m}(m/n)^{(1+\epsilon/2)m}.
Proof.

We have R2R2\|R^{\sharp}\|_{2}\leqslant\|R\|_{2} since 0RR0\leqslant R^{\sharp}\leqslant R pointwise. Hence, from (6.1),

Λ(R,R,R)\displaystyle\Lambda(R,R,R) =Λ(R,R,R)+Λ(R,R,R)+Λ(R,R,R)+Λ(R,R,R)\displaystyle=\Lambda(R^{\flat},R,R)+\Lambda(R^{\sharp},R^{\flat},R)+\Lambda(R^{\sharp},R^{\sharp},R^{\flat})+\Lambda(R^{\sharp},R^{\sharp},R^{\sharp})
Λ(R,R,R)+3R22R1.\displaystyle\leqslant\Lambda(R^{\sharp},R^{\sharp},R^{\sharp})+3\|R\|_{2}^{2}\|R^{\flat}\|_{1}.

From sparseval (Lemma 4.5\wrtusdrflem:L2-to-sparseval and Proposition 4.6\wrtusdrfprop:umbral-sparseval),

R22(n!nn)2eO(m)(m/n)m/2.\|R\|_{2}^{2}\ll{\left(\frac{n!}{n^{n}}\right)}^{2}e^{O(m)}(m/n)^{m/2}.

Combining with Corollary 6.5\wrtusdrfcor:rflat-l1-bound gives the bound. ∎

6.3. The contribution from RR^{\sharp}

Finally we must bound Λ(R,R,R)\Lambda(R^{\sharp},R^{\sharp},R^{\sharp}). From Lemma 6.4,

Rn!nneδmδϵmQn!nneO(m)(m/n)ϵm/2Q,R^{\sharp}\leqslant\frac{n!}{n^{n}}e^{\delta m}\delta^{-\epsilon m}Q\leqslant\frac{n!}{n^{n}}e^{O(m)}(m/n)^{-\epsilon m/2}Q, (6.3)

where

Q=πΠ(2)σπ(δ)cπ.Q=\sum_{\pi\in\Pi^{(2)}}\sigma_{\pi}^{(\delta)}c_{\pi}.

Hence it suffices to bound Λ(Q,Q,Q)\Lambda(Q,Q,Q). The key ingredient for this is the knowledge of the exact value of combinatorial rank for π1,π2,π3Π(2)\pi_{1},\pi_{2},\pi_{3}\in\Pi^{(2)} (Lemma 3.4\wrtusdrfclaim:cool-crank-value).

Lemma 6.7.
Λ(Q,Q,Q)(m/n)3m/2eO(m+n/m).\Lambda(Q,Q,Q)\leqslant(m/n)^{3m/2}e^{O(m+n/m)}.
Proof.

Let AΠA(2)\mathcal{M}_{A}\subseteq\Pi_{A}^{(2)} be the set of matchings (partitions all of whose cells have size 2). For π1,π2,π3ΠA(2)\pi_{1},\pi_{2},\pi_{3}\in\Pi^{(2)}_{A}, let k(π1,π2,π2)k(\pi_{1},\pi_{2},\pi_{2}) be the number of cells pπ1π2π3p\in\pi_{1}\vee\pi_{2}\vee\pi_{3} such that πi|pp\pi_{i}|_{p}\in\mathcal{M}_{p} for each i[3]i\in[3]. Then, from Lemma 3.2\wrtusdrflem:crank-bound and Lemma 3.4\wrtusdrfclaim:cool-crank-value,

Λ(cπ1,cπ2,cπ3)nk(π1,π2,π3)rank(π1)rank(π2)rank(π3).\Lambda(c_{\pi_{1}},c_{\pi_{2}},c_{\pi_{3}})\leqslant n^{k(\pi_{1},\pi_{2},\pi_{3})-\operatorname{rank}(\pi_{1})-\operatorname{rank}(\pi_{2})-\operatorname{rank}(\pi_{3})}.

Hence

Λ(Q,Q,Q)\displaystyle\Lambda(Q,Q,Q) =π1,π2,π3ΠA(2)σπ1(δ)σπ2(δ)σπ3(δ)Λ(cπ1,cπ2,cπ3)\displaystyle=\sum_{\pi_{1},\pi_{2},\pi_{3}\in\Pi^{(2)}_{A}}\sigma_{\pi_{1}}^{(\delta)}\sigma_{\pi_{2}}^{(\delta)}\sigma_{\pi_{3}}^{(\delta)}\Lambda(c_{\pi_{1}},c_{\pi_{2}},c_{\pi_{3}})
π1,π2,π3ΠA(2)σπ1(δ)σπ2(δ)σπ3(δ)nk(π1,π2,π3)rank(π1)rank(π2)rank(π3)\displaystyle\leqslant\sum_{\pi_{1},\pi_{2},\pi_{3}\in\Pi^{(2)}_{A}}\sigma_{\pi_{1}}^{(\delta)}\sigma_{\pi_{2}}^{(\delta)}\sigma_{\pi_{3}}^{(\delta)}n^{k(\pi_{1},\pi_{2},\pi_{3})-\operatorname{rank}(\pi_{1})-\operatorname{rank}(\pi_{2})-\operatorname{rank}(\pi_{3})}
=πΠApππ1,π2,π3Πp(2)π1π2π3={p}nk(π1,π2,π3)i[3]σπi(δ)nrank(πi).\displaystyle=\sum_{\pi\in\Pi_{A}}\prod_{p\in\pi}\sum_{\begin{subarray}{c}\pi_{1},\pi_{2},\pi_{3}\in\Pi^{(2)}_{p}\\ \pi_{1}\vee\pi_{2}\vee\pi_{3}=\{p\}\end{subarray}}n^{k(\pi_{1},\pi_{2},\pi_{3})}\prod_{i\in[3]}\sigma_{\pi_{i}}^{(\delta)}n^{-{\operatorname{rank}(\pi_{i})}}.

In the last sum above, since π1π2π3={p}\pi_{1}\vee\pi_{2}\vee\pi_{3}=\{p\}, k(π1,π2,π3)k(\pi_{1},\pi_{2},\pi_{3}) is 0 or 11 according to whether π1,π2,π3p\pi_{1},\pi_{2},\pi_{3}\in\mathcal{M}_{p}. Splitting the sum according to these cases,

Λ(Q,Q,Q)\displaystyle\Lambda(Q,Q,Q) πΠApπ(π1,π2,π3Πp(2)π1π2π3={p}i[3]σπi(δ)nrank(πi)\displaystyle\leqslant\sum_{\pi\in\Pi_{A}}\prod_{p\in\pi}\bigg{(}\sum_{\begin{subarray}{c}\pi_{1},\pi_{2},\pi_{3}\in\Pi^{(2)}_{p}\\ \pi_{1}\vee\pi_{2}\vee\pi_{3}=\{p\}\end{subarray}}\prod_{i\in[3]}\sigma_{\pi_{i}}^{(\delta)}n^{-{\operatorname{rank}(\pi_{i})}}
+π1,π2,π3pπ1π2π3={p}ni[3]σπi(δ)nrank(πi)).\displaystyle\qquad+\sum_{\begin{subarray}{c}\pi_{1},\pi_{2},\pi_{3}\in\mathcal{M}_{p}\\ \pi_{1}\vee\pi_{2}\vee\pi_{3}=\{p\}\end{subarray}}n\prod_{i\in[3]}\sigma_{\pi_{i}}^{(\delta)}n^{-{\operatorname{rank}(\pi_{i})}}\bigg{)}.

In the second sum we will ignore the constraint π1π2π3={p}\pi_{1}\vee\pi_{2}\vee\pi_{3}=\{p\}; in the first sum we will use only rank(π1)+rank(π2)+rank(π3)rank(π1π2π3)=|p|1\operatorname{rank}(\pi_{1})+\operatorname{rank}(\pi_{2})+\operatorname{rank}(\pi_{3})\geqslant\operatorname{rank}(\pi_{1}\vee\pi_{2}\vee\pi_{3})=|p|-1.

Fix parameters wr1w_{r}\geqslant 1 for all r1r\geqslant 1. Define

αr(x)\displaystyle\alpha_{r}(x) =πΠr(2)σπ(δ)xrank(π),\displaystyle=\sum_{\pi\in\Pi_{r}^{(2)}}\sigma_{\pi}^{(\delta)}x^{\operatorname{rank}(\pi)},
αr(x)\displaystyle\alpha_{r}^{\prime}(x) =πrσπ(δ)xrank(π)=|r|xr/2,\displaystyle=\sum_{\pi\in\mathcal{M}_{r}}\sigma_{\pi}^{(\delta)}x^{\operatorname{rank}(\pi)}=|\mathcal{M}_{r}|x^{r/2},
βr(x)\displaystyle\beta_{r}(x) =πΠrpπ(w|p|(|p|1)α|p|(w|p|x)3+x1α|p|(x)3).\displaystyle=\sum_{\pi\in\Pi_{r}}\prod_{p\in\pi}{\left(w_{|p|}^{-(|p|-1)}\alpha_{|p|}(w_{|p|}x)^{3}+x^{-1}\alpha^{\prime}_{|p|}(x)^{3}\right)}.

Then, by the discussion above,

Λ(Q,Q,Q)βm(1/n).\Lambda(Q,Q,Q)\leqslant\beta_{m}(1/n).

Three applications of the exponential formula (3.1) give

r0yrr!αr(x)\displaystyle\sum_{r\geqslant 0}\frac{y^{r}}{r!}\alpha_{r}(x) =exp(δy+xy2/2),\displaystyle=\exp(\delta y+xy^{2}/2), (6.4)
r0yrr!αr(x)\displaystyle\sum_{r\geqslant 0}\frac{y^{r}}{r!}\alpha_{r}^{\prime}(x) =exp(xy2/2),\displaystyle=\exp(xy^{2}/2), (6.5)
r0yrr!βr(x)\displaystyle\sum_{r\geqslant 0}\frac{y^{r}}{r!}\beta_{r}(x) =exp(r1yrwrr+1αr(wrx)3r!+r2evenyrx1αr(x)3r!).\displaystyle=\exp{\left(\sum_{r\geqslant 1}\frac{y^{r}w_{r}^{-r+1}\alpha_{r}(w_{r}x)^{3}}{r!}+\sum_{r\geqslant 2~{}\text{even}}\frac{y^{r}x^{-1}\alpha_{r}^{\prime}(x)^{3}}{r!}\right)}. (6.6)

From (6.4), for real y>0y>0,

αr(x)r!yrexp(δy+xy2/2).\alpha_{r}(x)\leqslant\frac{r!}{y^{r}}\exp(\delta y+xy^{2}/2).

Replacing xx with wrxw_{r}x, putting wr=δ2/(xr)w_{r}=\delta^{2}/(xr) (we will ensure later that wr1w_{r}\geqslant 1 for 1rm1\leqslant r\leqslant m) and y=r/δy=r/\delta gives

wrr+1αr(wrx)3eO(r)rrδr+2xr1.w_{r}^{-r+1}\alpha_{r}(w_{r}x)^{3}\leqslant e^{O(r)}r^{r}\delta^{r+2}x^{r-1}.

From (6.5) with y=(r/x)1/2y=(r/x)^{1/2} we have

αr(x)r!yrexp(xy2/2)r1/2(rx/e)r/2\alpha^{\prime}_{r}(x)\leqslant\frac{r!}{y^{r}}\exp(xy^{2}/2)\asymp r^{1/2}(rx/e)^{r/2}

(alternatively, this follows directly from αr(x)=|r|xr/2\alpha^{\prime}_{r}(x)=|\mathcal{M}_{r}|x^{r/2}). Hence, from (6.6) for x,y>0x,y>0,

βm(x)m!ymexpb(x,y),\beta_{m}(x)\leqslant\frac{m!}{y^{m}}\exp b(x,y), (6.7)

where bb is the truncated sum

b(x,y)\displaystyle b(x,y) =r=1myrwrr+1αr(wrx)3r!+r=2myrx1αr(x)3r!\displaystyle=\sum_{r=1}^{m}\frac{y^{r}w_{r}^{-r+1}\alpha_{r}(w_{r}x)^{3}}{r!}+\sum_{r=2}^{m}\frac{y^{r}x^{-1}\alpha_{r}^{\prime}(x)^{3}}{r!}
r=1meO(r)δr+2xr1yr+r=2mrO(1)(e1/2r1/2x3/2y)rx1.\displaystyle\ll\sum_{r=1}^{m}e^{O(r)}\delta^{r+2}x^{r-1}y^{r}+\sum_{r=2}^{m}r^{O(1)}(e^{-1/2}r^{1/2}x^{3/2}y)^{r}x^{-1}.

Inserting x=1/nx=1/n and δ=(Cm/n)1/2\delta=(Cm/n)^{1/2},

b(1/n,y)r=1mO(m1/2y/n3/2)rm+r=2mrO(1)(e1/2r1/2y/n3/2)rn.b(1/n,y)\ll\sum_{r=1}^{m}O(m^{1/2}y/n^{3/2})^{r}m+\sum_{r=2}^{m}r^{O(1)}(e^{-1/2}r^{1/2}y/n^{3/2})^{r}n.

Note that wr=Cm/rw_{r}=Cm/r, and this is indeed at least 1 for rmr\leqslant m since we may assume C1C\geqslant 1. Finally, let y=cn3/2/m1/2y=cn^{3/2}/m^{1/2} for a sufficiently small constant c>0c>0. Then

b(1/n,y)m+n/m.b(1/n,y)\ll m+n/m.

Hence, from (6.7),

Λ(Q,Q,Q)βm(1/n)m!ymexpb(1/n,y)(m/n)3m/2eO(m+n/m),\Lambda(Q,Q,Q)\leqslant\beta_{m}(1/n)\leqslant\frac{m!}{y^{m}}\exp b(1/n,y)\ll(m/n)^{3m/2}e^{O(m+n/m)},

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

Λ(|PA1S|,|PA1S|,|PA1S||)(n!nn)3(m/n)9m/8eO(m+n/m).\Lambda(|P_{A}1_{S}|,|P_{A}1_{S}|,|P_{A}1_{S}||)\leqslant{\left(\frac{n!}{n^{n}}\right)}^{3}(m/n)^{9m/8}e^{O(m+n/m)}.
Proof.

From Corollary 6.6,

Λ(R,R,R)Λ(R,R,R)+(n!nn)3eO(m)(m/n)(1+ϵ/2)m.\Lambda(R,R,R)\leqslant\Lambda(R^{\sharp},R^{\sharp},R^{\sharp})+{\left(\frac{n!}{n^{n}}\right)}^{3}e^{O(m)}(m/n)^{(1+\epsilon/2)m}.

By (6.3) and the previous lemma the main term is

Λ(R,R,R)\displaystyle\Lambda(R^{\sharp},R^{\sharp},R^{\sharp}) (n!nn)3eO(m)(m/n)3ϵm/2Λ(Q,Q,Q)\displaystyle\leqslant{\left(\frac{n!}{n^{n}}\right)}^{3}e^{O(m)}(m/n)^{-3\epsilon m/2}\Lambda(Q,Q,Q)
(n!nn)3(m/n)(1ϵ)3m/2eO(m+n/m).\displaystyle\leqslant{\left(\frac{n!}{n^{n}}\right)}^{3}(m/n)^{(1-\epsilon)3m/2}e^{O(m+n/m)}.

Set ϵ=1/4\epsilon=1/4. ∎

7. Dense minor arcs

xxyy^{\prime}yyxx^{\prime}zz
Figure 1. A transition (x,y)(x,y)(x,y)\mapsto(x^{\prime},y^{\prime}) in the Markov chain

Define a Markov chain on X×YX\times Y as follows. If the current state is (x,y)(x,y), pick uniformly at random zZz\in Z. The next state is (x,y)(x^{\prime},y^{\prime}), where xx^{\prime} and yy^{\prime} are the unique solutions to

(x,y,z),(x,y,z)𝖫(x,y^{\prime},z),(x^{\prime},y,z)\in\mathsf{L}

(see Figure 1\wrtusdrffig:markov-chain). Let 𝒜\mathcal{A} be the transition operator for this Markov chain:

𝒜(f)(x,y)=1n(x,y,z),(x,y,z)𝖫f(x,y).\mathcal{A}(f)(x,y)=\frac{1}{n}\sum_{(x,y^{\prime},z),(x^{\prime},y,z)\in\mathsf{L}}f(x^{\prime},y^{\prime}).

The Markov chain is reversible with uniform stationary distribution, so 𝒜\mathcal{A} is self-adjoint and has the constant function on X×YX\times Y as a 11-eigenvector. Let 𝒰\mathcal{U} be the projection to constants:

𝒰(f)(x,y)=1n2x,yf(x,y).\mathcal{U}(f)(x,y)=\frac{1}{n^{2}}\sum_{x^{\prime},y^{\prime}}f(x^{\prime},y^{\prime}).
Definition 7.1.

We say 𝖫\mathsf{L} is 𝒜\mathcal{A}-quasirandom with parameter ρ\rho if 𝒜𝒰\mathcal{A}-\mathcal{U} has spectral radius at most ρ\rho.

In particular, ρ<1\rho<1 if and only if the Markov chain is connected, and in general ρ\rho measures the rate of mixing.

Remark 7.2.

For a finite set TT, let L2(T)0L^{2}(T)_{0} denote the subspace {fL2(T):𝐄f=0}\bigl{\{}f\in L^{2}(T)\colon\mathbf{E}f=0\bigr{\}}. Then equivalently, 𝖫\mathsf{L} is 𝒜\mathcal{A}-quasirandom with parameter ρ\rho if the restriction 𝒜|L2(X×Y)0\mathcal{A}|_{L^{2}(X\times Y)_{0}} has spectral radius at most ρ\rho.

All our applications of quasirandomness go through the following lemma.

Lemma 7.3.

Assume 𝖫\mathsf{L} is 𝒜\mathcal{A}-quasirandom with parameter ρ\rho and111Note that the m=1m=1 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 m1m\geqslant 1. Then

|Λ(f,g,h)|ρm/2f2g2h2|\Lambda(f,g,h)|\leqslant\rho^{m/2}\|f\|_{2}\|g\|_{2}\|h\|_{2} (7.1)

for all fL2(X)0mf\in L^{2}(X)_{0}^{\otimes m}, gL2(Y)0mg\in L^{2}(Y)_{0}^{\otimes m}, hL2(Z)0mh\in L^{2}(Z)_{0}^{\otimes m}.

Remark 7.4.

Identifying L2(X)mL^{2}(X)^{\otimes m} with L2(Xm)L^{2}(X^{m}) in the usual way, L2(X)0mL^{2}(X)_{0}^{\otimes m} is identified with the subspace imP[m]L2(Xm)\operatorname{im}P_{[m]}\subseteq L^{2}(X^{m}); see (2.2).

Proof of Lemma 7.3\wrtusdrflem:quasi-use.

By Cauchy–Schwarz,

|Λ(f,g,h)|\displaystyle|\Lambda(f,g,h)| =|𝐄(x,y,z)𝖫mf(x)g(y)h(z)|\displaystyle=\left|\mathbf{E}_{(x,y,z)\in\mathsf{L}^{m}}f(x)g(y)h(z)\right|
(𝐄z|𝐄x,y:(x,y,z)𝖫mf(x)g(y)|2)1/2h2\displaystyle\leqslant{\left(\mathbf{E}_{z}\left|\mathbf{E}_{x,y:(x,y,z)\in\mathsf{L}^{m}}f(x)g(y)\right|^{2}\right)}^{1/2}\|h\|_{2}
=(𝐄z,x,y,x,y:(x,y,z),(x,y,z)𝖫mf(x)g(y)f¯(x)g¯(y))1/2h2\displaystyle={\left(\mathbf{E}_{z,x,y,x^{\prime},y^{\prime}:(x,y,z),(x^{\prime},y^{\prime},z)\in\mathsf{L}^{m}}f(x)g(y){\bar{f}}(x^{\prime}){\bar{g}}(y^{\prime})\right)}^{1/2}\|h\|_{2}
=𝒜m(fg¯),fg¯1/2h2.\displaystyle=\langle\mathcal{A}^{\otimes m}(f\otimes\bar{g}),f\otimes\bar{g}\rangle^{1/2}\|h\|_{2}.

Note fg¯2=f2g2\|f\otimes\bar{g}\|_{2}=\|f\|_{2}\|g\|_{2}, and that fg¯L2(X×Y)0mf\otimes\overline{g}\in L^{2}(X\times Y)_{0}^{\otimes m}. Since 𝒜|L2(X×Y)0\mathcal{A}|_{L^{2}(X\times Y)_{0}} has spectral radius at most ρ\rho, the tensor power 𝒜m|L2(X×Y)0m\mathcal{A}^{\otimes m}|_{L^{2}(X\times Y)_{0}^{\otimes m}} has spectral radius (and hence operator norm) at most ρm\rho^{m}, so the last expression above is bounded by ρm/2f2g2h2\rho^{m/2}\|f\|_{2}\|g\|_{2}\|h\|_{2}. ∎

Remark 7.5.

As stated in the introduction, while Definition 7.1\wrtusdrfdef:quasirandom has some nice properties (e.g., the spectral radius of 𝒜𝒰\mathcal{A}-\mathcal{U} 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 X=Y=ZX=Y=Z and 𝖫\mathsf{L} contains the diagonal {(x,x,x):xX}\{(x,x,x)\colon x\in X\} and is invariant under the S3S_{3}-action on triples) always fail to be 𝒜\mathcal{A}-quasirandom with parameter ρ<1\rho<1 (since the diagonal {(x,x):xX}\{(x,x):x\in X\} of X×XX\times X 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 ρ=o(1)\rho=o(1) with high probability as nn\to\infty.

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 A[n]A\subseteq[n] and |A|=m|A|=m. By Lemmas 7.3 and 7.2\wrtusdrflem:quasi-use,rem:l20,

|Λ(PA1S,PA1S,PA1S)|ρm/2PA1S23ρm/21S23=ρm/2(n!nn)3.|\Lambda(P_{A}1_{S},P_{A}1_{S},P_{A}1_{S})|\leqslant\rho^{m/2}\|P_{A}1_{S}\|_{2}^{3}\leqslant\rho^{m/2}\|1_{S}\|_{2}^{3}=\rho^{m/2}{\left(\frac{n!}{n^{n}}\right)}^{3}.

Hence, for ρ1\rho\leqslant 1,

|A|m|Λ(PA1S,PA1S,PA1S)|2nρm/2(n!nn)3.\sum_{|A|\geqslant m}|\Lambda(P_{A}1_{S},P_{A}1_{S},P_{A}1_{S})|\leqslant 2^{n}\rho^{m/2}{\left(\frac{n!}{n^{n}}\right)}^{3}.

Taking m=ϵnm=\epsilon n and ρ\rho so that 2ρϵ/21/102\rho^{\epsilon/2}\leqslant 1/10, the result follows. ∎

8. Quasirandomness

In this section we will verify that two natural classes of latin squares are 𝒜\mathcal{A}-quasirandom with parameter o(1)o(1):

  • multiplication tables of quasirandom groups;

  • uniformly random n×nn\times n latin squares, with high probability as nn\to\infty.

In the case of a group we can compute the whole spectrum of 𝒜\mathcal{A} using representation theory. In the case of a random latin square we will use the bound

1+ρ6tr𝒜61+\rho^{6}\leqslant\operatorname{tr}\mathcal{A}^{6}

which holds because the spectrum of 𝒜\mathcal{A} is real and 66 is even. By interpreting n6tr𝒜6n^{6}\operatorname{tr}\mathcal{A}^{6} as counting certain kinds of configuration in 𝖫\mathsf{L} (and using a recent result of [KSSS]) we will show that tr𝒜6=1+o(1)\operatorname{tr}\mathcal{A}^{6}=1+o(1) with high probability, which implies that ρ=o(1)\rho=o(1). (Using the same method one can show that tr𝒜4=3+o(1)\operatorname{tr}\mathcal{A}^{4}=3+o(1) with high probability, so 66 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 𝖫\mathsf{L} is the multiplication table of a group GG. Then the spectrum of 𝒜\mathcal{A} consists of d3(d+1)/2d^{3}(d+1)/2 copies of 1/d1/d and d3(d1)/2d^{3}(d-1)/2 copies of 1/d-1/d for every dd-dimensional irreducible representation of GG, and n2χIrr(G)χ(1)4n^{2}-\sum_{\chi\in\operatorname{Irr}(G)}\chi(1)^{4} zeros. In particular ρ=1/D\rho=1/D where DD is the minimal dimension of a nontrivial representation of GG.

Proof.

Here X=Y=Z=GX=Y=Z=G and 𝖫={(x,y,z)G3:xy=z}\mathsf{L}=\{(x,y,z)\in G^{3}:xy=z\}, so L2(X×Y)=L2(G×G)L^{2}(X\times Y)=L^{2}(G\times G) and 𝒜\mathcal{A} is the operator defined by

𝒜(f)(x,y)=1nzGf(zy1,x1z).\mathcal{A}(f)(x,y)=\frac{1}{n}\sum_{z\in G}f(zy^{-1},x^{-1}z).

By representation theory, L2(G)L^{2}(G) has an orthogonal basis consisting of the functions of the form xρ(x)ei,ejx\mapsto\langle\rho(x)e_{i},e_{j}\rangle, where ρ:GU(V)\rho\colon G\to U(V) is an irreducible unitary representation of GG and e1,,edimVe_{1},\dots,e_{\dim V} is an orthonormal basis of VV.

It follows that L2(G×G)L2(G)L2(G)L^{2}(G\times G)\cong L^{2}(G)\otimes L^{2}(G) has an orthogonal basis consisting of functions of the form

fρ,ρ,i,j,k,(x,y)=ρ(x)ei,eje,ρ(y)ekf_{\rho,\rho^{\prime},i,j,k,\ell}(x,y)=\bigl{\langle}\rho(x)e_{i},e_{j}\bigr{\rangle}\bigl{\langle}e^{\prime}_{\ell},\rho^{\prime}(y)e^{\prime}_{k}\bigr{\rangle}

where ρ:GU(V)\rho\colon G\to U(V) and ρ:GU(V)\rho^{\prime}\colon G\to U(V^{\prime}) are two irreducible unitary representations of GG and 1i,jdimV1\leqslant i,j\leqslant\dim V, 1k,dimV1\leqslant k,\ell\leqslant\dim V^{\prime}.

To find 𝒜(fρ,ρ,i,j,k,)\mathcal{A}(f_{\rho,\rho^{\prime},i,j,k,\ell}) we recall the Schur orthogonality relation for matrix coefficients, which states that for irreducible VV, VV^{\prime} as above, a,bVa,b\in V and a,bVa^{\prime},b^{\prime}\in V^{\prime},

1nzGρ(z)a,bb,ρ(z)a={0:(ρ,V)(ρ,V)1dimVa,ab,b:(ρ,V)=(ρ,V),\frac{1}{n}\sum_{z\in G}\langle\rho(z)a,b\rangle\langle b^{\prime},\rho^{\prime}(z)a^{\prime}\rangle=\begin{cases}0&\colon(\rho,V)\ncong(\rho^{\prime},V^{\prime})\\ \frac{1}{\dim V}\langle a,a^{\prime}\rangle\langle b^{\prime},b\rangle&\colon(\rho,V)=(\rho^{\prime},V^{\prime}),\end{cases}

and thereby compute

𝒜(fρ,ρ,i,j,k,)(x,y)\displaystyle\mathcal{A}(f_{\rho,\rho^{\prime},i,j,k,\ell})(x,y) =1nzGρ(z)ρ(y1)ei,ejρ(x)e,ρ(z)ek\displaystyle=\frac{1}{n}\sum_{z\in G}\bigl{\langle}\rho(z)\rho(y^{-1})e_{i},e_{j}\bigr{\rangle}\bigl{\langle}\rho(x)e^{\prime}_{\ell},\rho(z)e^{\prime}_{k}\bigr{\rangle}
={0:(ρ,V)(ρ,V)1dimVρ(x)e,ejei,ρ(y)ek:(ρ,V)=(ρ,V)\displaystyle=\begin{cases}0&\colon(\rho,V)\ncong(\rho^{\prime},V^{\prime})\\ \frac{1}{\dim V}\bigl{\langle}\rho(x)e_{\ell},e_{j}\bigr{\rangle}\bigl{\langle}e_{i},\rho(y)e_{k}\bigr{\rangle}&\colon(\rho,V)=(\rho^{\prime},V^{\prime})\end{cases}
={0:(ρ,V)(ρ,V)1dimVfρ,ρ,,j,k,i(x,y):(ρ,V)=(ρ,V).\displaystyle=\begin{cases}0&\colon(\rho,V)\ncong(\rho^{\prime},V^{\prime})\\ \frac{1}{\dim V}f_{\rho,\rho,\ell,j,k,i}(x,y)&\colon(\rho,V)=(\rho^{\prime},V^{\prime}).\end{cases}

In the case ρρ\rho\neq\rho^{\prime} we get an eigenfunction with eigenvalue 0. When ρ=ρ\rho=\rho^{\prime} and i=i=\ell we get a (1/dimV)(1/\dim V)-eigenfunction. Finally when ρ=ρ\rho=\rho^{\prime} and ii\neq\ell, the functions

fρ,ρ,i,j,k,±fρ,ρ,,j,k,if_{\rho,\rho,i,j,k,\ell}\pm f_{\rho,\rho,\ell,j,k,i}

are eigenfunctions of 𝒜\mathcal{A} with eigenvalues ±1/dimV\pm 1/\dim V respectively.

Altogether we have d3+d3(d1)/2=d3(d+1)/2d^{3}+d^{3}(d-1)/2=d^{3}(d+1)/2 copies of 1/d1/d and d3(d1)/2d^{3}(d-1)/2 copies of 1/d-1/d, and the rest 0, 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 𝖧X𝖧×Y𝖧×Z𝖧\mathsf{H}\subseteq X_{\mathsf{H}}\times Y_{\mathsf{H}}\times Z_{\mathsf{H}} with vertex classes X𝖧,Y𝖧,Z𝖧X_{\mathsf{H}},Y_{\mathsf{H}},Z_{\mathsf{H}}. The number of vertices is v=|X𝖧|+|Y𝖧|+|Z𝖧|v=|X_{\mathsf{H}}|+|Y_{\mathsf{H}}|+|Z_{\mathsf{H}}| and the number of triples (hyperedges) is e=|𝖧|e=|\mathsf{H}|. We say 𝖧\mathsf{H} is latin if every pair of vertices is in at most one triple. (A latin square of order nn is then a latin triple system with 3 classes of nn vertices and n2n^{2} triples.)

Let 𝖧\mathsf{H} be a fixed triple system. A copy of 𝖧\mathsf{H} in a triple system 𝖫\mathsf{L} is a triple of injective maps

X𝖧X𝖫,Y𝖧Y𝖫,Z𝖧Z𝖫X_{\mathsf{H}}\to X_{\mathsf{L}},\qquad Y_{\mathsf{H}}\to Y_{\mathsf{L}},\qquad Z_{\mathsf{H}}\to Z_{\mathsf{L}}

which maps triples to triples. Let N𝖧(𝖫)N_{\mathsf{H}}(\mathsf{L}) denote the number of copies of 𝖧\mathsf{H} in 𝖫\mathsf{L}.

Let 𝖡n\mathsf{B}_{n} denote the random triple system 𝖡n[n]×[n]×[n]\mathsf{B}_{n}\subseteq[n]\times[n]\times[n] in which each possible triple is present independently with probability 1/n1/n. Note that 𝐄[N𝖧(𝖡n)]=(1o(1))nve\mathbf{E}[N_{\mathsf{H}}(\mathsf{B}_{n})]=(1-o(1))n^{v-e} (when 𝖧\mathsf{H} is fixed and nn is large). We say 𝖧\mathsf{H} is α\alpha-stable if αve\alpha\geqslant v-e and

𝐄[N𝖧(𝖡n)𝖰𝖡n]𝐄[N𝖧(𝖡n)]=o(nα)\mathbf{E}[N_{\mathsf{H}}(\mathsf{B}_{n})\mid\mathsf{Q}\subseteq\mathsf{B}_{n}]-\mathbf{E}[N_{\mathsf{H}}(\mathsf{B}_{n})]=o(n^{\alpha})

for any latin triple system 𝖰[n]×[n]×[n]\mathsf{Q}\subseteq[n]\times[n]\times[n] with at most n(logn)3n(\log n)^{3} triples.

Theorem 8.2 ([KSSS, Theorem 7.2]).

Fix an α\alpha-stable latin triple system 𝖧\mathsf{H} with vv vertices and ee triples. Let 𝖫\mathsf{L} be a uniformly random latin square. Then

N𝖧(𝖫)nve+o(nα)N_{\mathsf{H}}(\mathsf{L})\leqslant n^{v-e}+o(n^{\alpha})

with high probability as nn\to\infty.

In order to use this theorem effectively we need a computable form of stability. Let 𝖧\mathsf{H} be a latin triple system. A subset of the vertices SX𝖧Y𝖧Z𝖧S\subseteq X_{\mathsf{H}}\cup Y_{\mathsf{H}}\cup Z_{\mathsf{H}} is called closed if whenever two vertices of a triple of 𝖧\mathsf{H} is in SS, so is the third. The closure S𝖧\langle S\rangle_{\mathsf{H}} of a subset SS if the smallest closed set containing it. If 𝖥𝖧\mathsf{F}\subseteq\mathsf{H} let X𝖥,Y𝖥,Z𝖥X_{\mathsf{F}},Y_{\mathsf{F}},Z_{\mathsf{F}} denote the vertices incident with at least one member of 𝖥\mathsf{F}, and let v(𝖥)=|X𝖥|+|Y𝖥|+|Z𝖥|v(\mathsf{F})=|X_{\mathsf{F}}|+|Y_{\mathsf{F}}|+|Z_{\mathsf{F}}| and e(𝖥)=|𝖥|e(\mathsf{F})=|\mathsf{F}|. We say 𝖥𝖧\mathsf{F}\subseteq\mathsf{H} generates 𝖧\mathsf{H} if

X𝖥Y𝖥Z𝖥𝖧=X𝖧Y𝖧Z𝖧.\langle X_{\mathsf{F}}\cup Y_{\mathsf{F}}\cup Z_{\mathsf{F}}\rangle_{\mathsf{H}}=X_{\mathsf{H}}\cup Y_{\mathsf{H}}\cup Z_{\mathsf{H}}.

Let

d(𝖧)=min{e(𝖥):𝖥generates𝖧}.d(\mathsf{H})=\min\{e(\mathsf{F}):\mathsf{F}~{}\text{generates}~{}\mathsf{H}\}.

For example, if 𝖧1\mathsf{H}_{1} is the latin triple system shown in Figure 2\wrtusdrffig:trAdj6, one generating set consists of both triples containing z1z_{1}, one triple containing z3z_{3}, and one triple containing z5z_{5}, and there is no smaller generating set, so d(𝖧1)=4d(\mathsf{H}_{1})=4.

Lemma 8.3.

Let 𝖧\mathsf{H} be a latin triple system with vv vertices and ee triples. Then 𝖧\mathsf{H} is α\alpha-stable provided αve\alpha\geqslant v-e and

α>ve+max𝖥𝖧(d(𝖥)v(𝖥)+e(𝖥)).\alpha>v-e+\max_{\emptyset\neq\mathsf{F}\subseteq\mathsf{H}}\bigl{(}d(\mathsf{F})-v(\mathsf{F})+e(\mathsf{F})\bigr{)}.
Remark 8.4.

A much simpler model problem is the following: given a fixed graph HH and a random graph Gn,pG_{n,p}, does GG contain nv(H)pe(H)(1+o(1))n^{v(H)}p^{e(H)}(1+o(1)) copies of HH (i.e., close to the expected number) with high probability? The answer might be no if HH contains a subgraph HH^{\prime} with much greater density than HH in some sense: indeed, if nv(H)pe(H)=o(1)n^{v(H^{\prime})}p^{e(H^{\prime})}=o(1) then with high probability G(n,p)G(n,p) contains zero copies of HH^{\prime}, and hence of HH. However, this is essentially all that can go wrong. The condition for α\alpha-stability in the lemma captures a similar intuition.

Remark 8.5.

Given a triple system 𝖧X𝖧×Y𝖧×Z𝖧\mathsf{H}\subseteq X_{\mathsf{H}}\times Y_{\mathsf{H}}\times Z_{\mathsf{H}}, one can construct a partition triple 𝔓=(π1,π2,π3)Π𝖧3\mathfrak{P}=(\pi_{1},\pi_{2},\pi_{3})\in\Pi_{\mathsf{H}}^{3} in the sense of Section 3.1 (i.e., the ground set has size e(𝖧)e(\mathsf{H})) where two triples (x,y,z),(x,y,z)𝖧(x,y,z),(x^{\prime},y^{\prime},z^{\prime})\in\mathsf{H} lie in the same cell of π1\pi_{1} if and only if x=xx=x^{\prime}, and similarly for π2\pi_{2} and y=yy=y^{\prime}, and π3\pi_{3} and z=zz=z^{\prime}.

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 crank(𝔓)=2e(𝖧)d(𝖧)\operatorname{crank}(\mathfrak{P})=2e(\mathsf{H})-d(\mathsf{H}).

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 𝖰[n]3\mathsf{Q}\subseteq[n]^{3} be a latin triple system with at most n1+o(1)n^{1+o(1)} triples. For a copy of 𝖧\mathsf{H} in 𝖡n\mathsf{B}_{n}, say one of its triples is forced if it appears in 𝖰\mathsf{Q}. The difference

𝐄[N𝖧(𝖡n)𝖰𝖡n]𝐄[N𝖧(𝖡n)]\mathbf{E}[N_{\mathsf{H}}(\mathsf{B}_{n})\mid\mathsf{Q}\subseteq\mathsf{B}_{n}]-\mathbf{E}[N_{\mathsf{H}}(\mathsf{B}_{n})] (8.1)

arises from copies of 𝖧\mathsf{H} with at least one forced triple. Let 𝖥𝖧\mathsf{F}\subseteq\mathsf{H} be a nonempty subsystem and consider copies of 𝖧\mathsf{H} whose forced triples are precisely the images of those in 𝖥\mathsf{F}. Let 𝖥0𝖥\mathsf{F}_{0}\subseteq\mathsf{F} be a generating subsystem of size d(𝖥)d(\mathsf{F}). Because 𝖰\mathsf{Q} satisfies the latin property, any copy of 𝖥\mathsf{F} in 𝖰\mathsf{Q} is determined by the image of 𝖥0\mathsf{F}_{0}. Therefore the number of copies of 𝖥\mathsf{F} in 𝖰\mathsf{Q} is at most |𝖰||𝖥0||\mathsf{Q}|^{|\mathsf{F}_{0}|}. There are vv(𝖥)v-v(\mathsf{F}) vertices of 𝖧\mathsf{H} outside 𝖥\mathsf{F}, each with nn possible images in [n]3[n]^{3}, and the image of each of the ee(𝖥)e-e(\mathsf{F}) triples outside 𝖥\mathsf{F} has probability 1/n1/n (independently) of being present in 𝖡n\mathsf{B}_{n}. Hence the contribution to (8.1) from 𝖥\mathsf{F} is bounded by

|𝖰||𝖥0|nvv(𝖥)(1/n)ee(𝖥)=nve+d(𝖥)v(𝖥)+e(𝖥)+o(1).|\mathsf{Q}|^{|\mathsf{F}_{0}|}n^{v-v(\mathsf{F})}(1/n)^{e-e(\mathsf{F})}=n^{v-e+d(\mathsf{F})-v(\mathsf{F})+e(\mathsf{F})+o(1)}.

This is o(nα)o(n^{\alpha}) provided the stated condition is satisfied. ∎

x0x_{0}y1y_{1}x2x_{2}y3y_{3}x4x_{4}y5y_{5}x6x_{6}y0y_{0}x1x_{1}y2y_{2}x3x_{3}y4y_{4}x5x_{5}y6y_{6}z1z_{1}z2z_{2}z3z_{3}z4z_{4}z5z_{5}z6z_{6}
Figure 2. The chain (x0,y0),,(x6,y6)(x_{0},y_{0}),\dots,(x_{6},y_{6}) and the latin triple system 𝖧1\mathsf{H}_{1} defined by identifying x0x_{0} with x6x_{6} and y0y_{0} with y6y_{6}

Now we can show that random latin squares are 𝒜\mathcal{A}-quasirandom with parameter o(1)o(1) with high probability (Theorem 1.3\wrtusdrfthm:random-implies-quasirandom). This follows from the following proposition and the bound 1+ρ6tr𝒜61+\rho^{6}\leqslant\operatorname{tr}\mathcal{A}^{6}.

Proposition 8.6.

For a uniformly random latin square 𝖫\mathsf{L},

tr𝒜6=1+o(1)\operatorname{tr}\mathcal{A}^{6}=1+o(1)

with high probability as nn\to\infty.

Proof (computer-assisted).

For (x0,y0)X×Y(x_{0},y_{0})\in X\times Y, let (xi,yi)(x_{i},y_{i}) denote the iterates of (x0,y0)(x_{0},y_{0}) under the Markov chain defining 𝒜\mathcal{A}. Then

tr𝒜6=x0,y0𝐏((x6,y6)=(x0,y0))=N/n6,\operatorname{tr}\mathcal{A}^{6}=\sum_{x_{0},y_{0}}\mathbf{P}\bigl{(}(x_{6},y_{6})=(x_{0},y_{0})\bigr{)}=N/n^{6},

where NN is the number of configurations in 𝖫\mathsf{L} of the form shown in Figure 2\wrtusdrffig:trAdj6 with x0=x6x_{0}=x_{6} and y0=y6y_{0}=y_{6}. We do not assume the other vertices are distinct.

Let 𝖧1\mathsf{H}_{1} be the latin triple system depicted in Figure 2\wrtusdrffig:trAdj6 and let 𝖧2,,𝖧k\mathsf{H}_{2},\dots,\mathsf{H}_{k} (where kk 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 (πX,πY,πZ)(\pi_{X},\pi_{Y},\pi_{Z}) where πXΠX𝖧1\pi_{X}\in\Pi_{X_{\mathsf{H}_{1}}}, πYΠY𝖧1\pi_{Y}\in\Pi_{Y_{\mathsf{H}_{1}}}, πZΠZ𝖧1\pi_{Z}\in\Pi_{Z_{\mathsf{H}_{1}}} satisfying the following closure property: if (x,y,z)(x,y,z) and (x,y,z)(x^{\prime},y^{\prime},z^{\prime}) are two triples of 𝖧1\mathsf{H}_{1} and two of the pairs (x,x)(x,x^{\prime}), (y,y)(y,y^{\prime}), (z,z)(z,z^{\prime}) are in the same cell of πX\pi_{X}, πY\pi_{Y}, πZ\pi_{Z} respectively then so is the third. Number such triples of partitions 1,,k1,\dots,k, where 11 corresponds to three copies of the discrete partition. Then 𝖧i\mathsf{H}_{i} denotes the quotient hypergraph of 𝖧1\mathsf{H}_{1} with respect to partition ii.

Let Ni=N𝖧i(𝖫)N_{i}=N_{\mathsf{H}_{i}}(\mathsf{L}). Then N=N1++NkN=N_{1}+\cdots+N_{k}. Let vi=v(𝖧i)v_{i}=v(\mathsf{H}_{i}) and ei=e(𝖧i)e_{i}=e(\mathsf{H}_{i}). Then v1e1=1812=6v_{1}-e_{1}=18-12=6. Now the proposition follows from Theorem 8.2\wrtusdrfthm:KSSS, Lemma 8.3\wrtusdrflem:stability-lemma, and the following two claims:

  1. (1)

    viei5v_{i}-e_{i}\leqslant 5 for each i>1i>1,

  2. (2)

    viei+max𝖥𝖧(d(𝖥)v(𝖥)+e(𝖥))5v_{i}-e_{i}+\max_{\emptyset\neq\mathsf{F}\subseteq\mathsf{H}}\bigl{(}d(\mathsf{F})-v(\mathsf{F})+e(\mathsf{F})\bigr{)}\leqslant 5 for each i1i\geqslant 1.

Indeed, provided (1) and (2) hold, Lemma 8.3\wrtusdrflem:stability-lemma shows that 𝖧i\mathsf{H}_{i} is 66-stable for each i1i\geqslant 1, so Theorem 8.2\wrtusdrfthm:KSSS implies that Ninviei+o(n6)N_{i}\leqslant n^{v_{i}-e_{i}}+o(n^{6}) with high probability for each ii, so N(1+o(1))n6N\leqslant(1+o(1))n^{6} with high probability.

Both claims can be verified by exhaustive search. We find 𝖧2,,𝖧k\mathsf{H}_{2},\dots,\mathsf{H}_{k} by starting with 𝖧1\mathsf{H}_{1} and iteratively identifying pairs of vertices, using breadth-first search. Thus we verify (1). Now for each 𝖧i\mathsf{H}_{i} we check all subsystems 𝖥𝖧i\mathsf{F}\subseteq\mathsf{H}_{i} and compute d(𝖥)d(\mathsf{F}) by checking all 𝖥0𝖥\mathsf{F}_{0}\subseteq\mathsf{F}, and thus we verify (2).

It turns out k=1206k=1206, and there are 154154 distinct isomorphism classes among the degenerations 𝖧i\mathsf{H}_{i}. The quantity in (2) turns out to be at most 4 in all cases except 𝖧1\mathsf{H}_{1}, for which it is 55. There are just eight degenerations 𝖧i\mathsf{H}_{i} (up to isomorphism) for which viei=5v_{i}-e_{i}=5. Of these, four are just 𝖧1\mathsf{H}_{1} with a single pair of vertices identified (so vi=17v_{i}=17 and ei=12e_{i}=12). The other four cases are shown in Figure 3\wrtusdrffig:degenerations. These cases are therefore the dominant contributors to the error term. ∎

(a) v=16v=16, e=11e=11
(b) v=15v=15, e=10e=10
(c) v=11v=11, e=6e=6
(d) v=11v=11, e=6e=6
Figure 3. Degenerations of 𝖧1\mathsf{H}_{1} with viei=5v_{i}-e_{i}=5 and ei<12e_{i}<12. Some triangles are shown flat.

References