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

New upper bounds for the Erdős-Gyárfás problem on generalized Ramsey numbers

Alex Cameron 111Department of Mathematics, Vanderbilt University, Nashville, TN 37212, USA.
Email: [email protected]
   Emily Heath 222Department of Mathematics, University of Illinois at Urbana-Champaign, Urbana, IL 61801, USA. Email: [email protected]
Abstract

A (p,q)(p,q)-coloring of a graph GG is an edge-coloring of GG which assigns at least qq colors to each pp-clique. The problem of determining the minimum number of colors, f(n,p,q)f(n,p,q), needed to give a (p,q)(p,q)-coloring of the complete graph KnK_{n} is a natural generalization of the well-known problem of identifying the diagonal Ramsey numbers rk(p)r_{k}(p). The best-known general upper bound on f(n,p,q)f(n,p,q) was given by Erdős and Gyárfás in 1997 using a probabilistic argument. Since then, improved bounds in the cases where p=qp=q have been obtained only for p{4,5}p\in\{4,5\}, each of which was proved by giving a deterministic construction which combined a (p,p1)(p,p-1)-coloring using few colors with an algebraic coloring.

In this paper, we provide a framework for proving new upper bounds on f(n,p,p)f(n,p,p) in the style of these earlier constructions. We characterize all colorings of pp-cliques with p1p-1 colors which can appear in our modified version of the (p,p1)(p,p-1)-coloring of Conlon, Fox, Lee, and Sudakov. This allows us to greatly reduce the amount of case-checking required in identifying (p,p)(p,p)-colorings, which would otherwise make this problem intractable for large values of pp. In addition, we generalize our algebraic coloring from the p=5p=5 setting and use this to give improved upper bounds on f(n,6,6)f(n,6,6) and f(n,8,8)f(n,8,8).

1 Introduction

Let pp and qq be positive integers such that 1q(p2)1\leq q\leq\binom{p}{2}. We say that an edge-coloring of a graph GG is a (p,q)(p,q)-coloring if any pp-clique of GG contains edges of at least qq distinct colors. Let f(n,p,q)f(n,p,q) denote the minimum number of colors needed to give a (p,q)(p,q)-coloring of the complete graph on nn vertices, KnK_{n}.

This function f(n,p,q)f(n,p,q) is known as the Erdős-Gyárfás function after the authors of the first paper [5] to systematically study (p,q)(p,q)-colorings. The majority of their work focused on understanding the asymptotic behavior of this function as nn\rightarrow\infty for fixed values of pp and qq. One of their primary results was a general upper bound of

f(n,p,q)=O(np2(p2)q+1)f(n,p,q)=O\left(n^{\frac{p-2}{\binom{p}{2}-q+1}}\right)

obtained using the Lovász Local Lemma, while one of the main problems they left open was the determination of qq, given a fixed value of pp, for which f(n,p,q)=Ω(nϵ)f(n,p,q)=\Omega(n^{\epsilon}) for some constant ϵ\epsilon, but f(n,p,q1)=no(1)f(n,p,q-1)=n^{o(1)}. Towards this end, they found that

n1p21f(n,p,p)cn2p1,n^{\frac{1}{p-2}}-1\leq f(n,p,p)\leq cn^{\frac{2}{p-1}},

where the lower bound is given by a simple induction argument and the upper bound is a special case of their general upper bound. However, they did not determine whether f(n,p,p1)=no(1)f(n,p,p-1)=n^{o(1)}.

In 2015, Conlon, Fox, Lee, and Sudakov [3], building on work done on small cases by Mubayi and Eichhorn [4, 6], showed that f(n,p,p1)=no(1)f(n,p,p-1)=n^{o(1)} by constructing an explicit (p,p1)(p,p-1)-coloring using very few colors. In [2], we slightly modified their coloring, which we call the CFLS coloring, and paired it with an “algebraic” construction to show that f(n,5,5)n1/3+o(1)f(n,5,5)\leq n^{1/3+o(1)}. This improves on the general upper bound found by Erdős and Gyárfás and comes close to matching their lower bound in terms of order of growth. Our construction built on the ideas of Mubayi in [7], where he gave an explicit construction showing that f(n,4,4)n1/2+o(1)f(n,4,4)\leq n^{1/2+o(1)}.

In this paper, we push these ideas further. In Section 2, we prove the following result.

Theorem 1.1.

For any p3p\geq 3, there is a (p,p1)(p,p-1)-coloring of KnK_{n} using no(1)n^{o(1)} colors such that the only pp-cliques that contain exactly p1p-1 distinct edge-colors are isomorphic (as edge-colored graphs) to one of the edge-colored pp-cliques given in the definition below.

Definition.

Given an edge-coloring f:E(Kn)Cf:E(K_{n})\rightarrow C, we say that a subset SV(Kn)S\subseteq V(K_{n}) has a leftover structure under ff if either |S|=1|S|=1 or there exists a bipartition (which we will call the initial bipartition) of SS into nonempty sets AA and BB for which

  • AA and BB each have a leftover structure under ff;

  • f(A)f(B)=f(A)\cap f(B)=\emptyset; and

  • there is a fixed color αC\alpha\in C such that f(a,b)=αf(a,b)=\alpha for all aAa\in A and all bBb\in B and αf(A)\alpha\not\in f(A) and αf(B)\alpha\not\in f(B).

pkp-kkk
Figure 1: An example of a pp-clique with a leftover structure.

Alternatively, a more constructive definition is to say that a pp-clique SS is leftover if either p=1p=1 or if it can be formed from a leftover (p1)(p-1)-clique by taking one of its vertices xx, making a copy xx^{\prime}, coloring xxxx^{\prime} with a new color, and coloring xyx^{\prime}y with the same color as xyxy for each ySy\in S for which yxy\neq x. Note that it is easy to see by induction that these pp-cliques always contain exactly p1p-1 colors.

One of the general difficulties in producing explicit (p,q)(p,q)-colorings is dealing with the large number of possible non-isomorphic ways to color the edges of a pp-clique with fewer than qq colors in order to demonstrate that a construction avoids them. By identifying the “bad” structures that are leftover after using only no(1)n^{o(1)} colors, we are able to greatly reduce the amount of case-checking required in identifying (p,p)(p,p)-colorings, which would otherwise make this problem intractable for large pp.

More precisely, one of the nice properties of these leftover structures is that any subset of vertices of a leftover clique induces a clique that is itself leftover. Therefore, any edge-coloring of KnK_{n} that eliminates leftover pp-cliques also eliminates all leftover PP-cliques for any PpP\geq p. Moreover, by Theorem 1.1, if this coloring uses nϵ+o(1)n^{\epsilon+o(1)} colors, then f(n,P,P)nϵ+o(1)f(n,P,P)\leq n^{\epsilon+o(1)}, as the product of this coloring with the one guaranteed in Theorem 1.1 will avoid any PP-clique with fewer than PP colors for each PpP\geq p.

As a specific example, in [2] we gave a (5,5)(5,5)-coloring of KnK_{n} that used n1/3+o(1)n^{1/3+o(1)} colors. Since this coloring avoids leftover 5-cliques, then it also avoids leftover PP-cliques for all P5P\geq 5. Therefore, if we take the product of this coloring with the appropriate one developed in Section 2 that eliminates all 66-cliques with 5 or fewer colors other than leftover 66-cliques, then we have a (6,6)(6,6)-coloring that uses only n1/3+o(1)n^{1/3+o(1)} colors, improving the best known upper bound given above, O(n2/5)O(n^{2/5}).

In Section 3, we generalize the algebraic portion of our coloring in [2], the “Modified Dot Product” coloring, to a version that eliminates leftover 66-cliques with O(n1/3)O(n^{1/3}) colors (making the above example redundant) and eliminates leftover 88-cliques with O(n1/4)O(n^{1/4}) colors. By taking the product of these colorings with the appropriate ones developed in Section 2, this gives us the following theorem.

Theorem 1.2.

We have the following upper bounds:

f(n,6,6)=n1/3+o(1);f(n,8,8)=n1/4+o(1).f(n,6,6)=n^{1/3+o(1)};\quad f(n,8,8)=n^{1/4+o(1)}.

This improves the best-known upper bound f(n,8,8)=O(n2/7)f(n,8,8)=O(n^{2/7}) as well.

2 Modified CFLS coloring

In this section, we define an edge-coloring ψp\psi_{p} of the complete graph with vertex set {0,1}α\{0,1\}^{\alpha} for some positive integer α\alpha. This construction is the product of two colorings, ψp=cp×Δp\psi_{p}=c_{p}\times\Delta_{p}, where cpc_{p} is the (p+3,p+2)(p+3,p+2)-coloring defined in [3]. In many places, this section tracks parts of the proof given in [3], and we have attempted to keep the notation consistent with that paper to make cross-referencing easier.

We will prove the following lemma about the coloring cpc_{p}.

Lemma 2.1.

Let pp be a fixed positive integer. Any subset S{0,1}αS\subseteq\{0,1\}^{\alpha} with |S|p+3|S|\leq p+3 vertices that contains exactly |S|1|S|-1 distinct colors under the edge-coloring cpc_{p} either has a leftover structure under cpc_{p} or contains a striped K4K_{4} under cpc_{p}.

A striped K4K_{4}, as described by the following definition, was first defined in [7].

Definition.

Let f:E(G)Cf:E(G)\rightarrow C be an edge-coloring of a graph GG. We call any 4-clique of GG, {a,b,c,d}V(G)\{a,b,c,d\}\subseteq V(G), for which f(ab)=f(cd)f(ab)=f(cd), f(ac)=f(bd)f(ac)=f(bd), f(ad)=f(bc)f(ad)=f(bc), f(ab)f(ac)f(ab)\neq f(ac), f(ab)f(ad)f(ab)\neq f(ad), and f(ac)f(ad)f(ac)\neq f(ad) a striped K4K_{4}.

We will also prove the following result about the coloring ψp\psi_{p}.

Lemma 2.2.

There is no striped K4K_{4} under the edge-coloring ψp\psi_{p}.

These two lemmas are enough to conclude that ψp\psi_{p} is a (p+3,p+2)(p+3,p+2)-coloring for which any clique SS with |S|p+3|S|\leq p+3 that contains exactly |S|1|S|-1 colors must have a leftover structure.

2.1 The construction

For some positive integer pp, let

1r1r2rp1\leq r_{1}\leq r_{2}\leq\cdots\leq r_{p}

be fixed positive integers such that rd|rd+1r_{d}|r_{d+1} for each d=1,,p1d=1,\ldots,p-1. These rir_{i} will be called the parameters of our edge-coloring.

For any αrp\alpha\geq r_{p}, let n=2αn=2^{\alpha}, and associate each vertex of the complete graph KnK_{n} with its own unique binary string of length α\alpha. For each d=1,,pd=1,\ldots,p, let α=adrd+bd\alpha=a_{d}r_{d}+b_{d} for positive integers ad,bda_{d},b_{d} such that 1bdrd1\leq b_{d}\leq r_{d}. For each string x{0,1}αx\in\{0,1\}^{\alpha}, we let

x=(x1(d),x2(d),,xad+1(d))x=\left(x_{1}^{(d)},x_{2}^{(d)},\ldots,x_{a_{d}+1}^{(d)}\right)

where xi(d)x_{i}^{(d)} denotes a binary string in {0,1}rd\{0,1\}^{r_{d}} for each i=1,,adi=1,\ldots,a_{d} and xad+1(d)x_{a_{d}+1}^{(d)} denotes a binary string from {0,1}bd\{0,1\}^{b_{d}}. We will call these substrings rdr_{d}-blocks of xx, including the final one which may or may not actually have length equal to rdr_{d}.

In the following definitions, we let r0=1r_{0}=1 and rp+1=αr_{p+1}=\alpha. First, we define a function ηd\eta_{d} for any d=0,,pd=0,\ldots,p on domain {0,1}β×{0,1}β\{0,1\}^{\beta}\times\{0,1\}^{\beta} where β\beta is any positive integer as

ηd(x,y)={(i,{xi(d),yi(d)})xy0x=y\eta_{d}(x,y)=\left\{\begin{array}[]{ll}\left(i,\{x_{i}^{(d)},y_{i}^{(d)}\}\right)&\quad x\neq y\\ 0&\quad x=y\end{array}\right.

where ii denotes the minimum index for which xi(d)yi(d)x_{i}^{(d)}\neq y_{i}^{(d)}.

For x,y{0,1}αx,y\in\{0,1\}^{\alpha} and 0dp0\leq d\leq p, let

ξd(x,y)=(ηd(x1(d+1),y1(d+1)),,ηd(xad+1+1(d+1),yad+1+1(d+1))).\xi_{d}(x,y)=\left(\eta_{d}\left(x_{1}^{(d+1)},y_{1}^{(d+1)}\right),\ldots,\eta_{d}\left(x_{a_{d+1}+1}^{(d+1)},y_{a_{d+1}+1}^{(d+1)}\right)\right).

And let

cp(x,y)=(ξp(x,y),,ξ0(x,y)).c_{p}(x,y)=\left(\xi_{p}(x,y),\ldots,\xi_{0}(x,y)\right).

Next, we assume that the binary strings of {0,1}β\{0,1\}^{\beta} are lexicographically ordered for every positive integer β\beta. For 1iap+11\leq i\leq a_{p}+1 and binary strings x<yx<y, define

δp,i(x,y)={+1if xi(p)yi(p)1if xi(p)>yi(p).\delta_{p,i}(x,y)=\left\{\begin{array}[]{ll}+1&\quad\text{if }x_{i}^{(p)}\leq y_{i}^{(p)}\\ -1&\quad\text{if }x_{i}^{(p)}>y_{i}^{(p)}.\end{array}\right.

Let

Δp(x,y)=(δp,1(x,y),,δp,ap+1(x,y)).\Delta_{p}(x,y)=\left(\delta_{p,1}(x,y),\ldots,\delta_{p,a_{p}+1}(x,y)\right).

Finally, let

ψp(x,y)=(cp(x,y),Δp(x,y)).\psi_{p}(x,y)=\left(c_{p}(x,y),\Delta_{p}(x,y)\right).

2.2 Number of colors

For any positive integer nn, let β\beta be the positive integer for which

2(β1)p+1<n2βp+1.2^{(\beta-1)^{p+1}}<n\leq 2^{\beta^{p+1}}.

For each d=1,,p+1d=1,\ldots,p+1, let rd=βdr_{d}=\beta^{d} in the construction of ψp\psi_{p}. Specifically, this means we are constructing the coloring on the complete graph with vertex set {0,1}α\{0,1\}^{\alpha} where α=βp+1\alpha=\beta^{p+1}. We can apply this coloring to KnK_{n} by arbitrarily associating each vertex of KnK_{n} with a unique binary string from {0,1}α\{0,1\}^{\alpha} and taking the induced coloring.

As shown in [3], for these choices of parameters rdr_{d}, the coloring cpc_{p} uses at most 24(p+1)βplog2β2^{4(p+1)\beta^{p}\log_{2}\beta} colors. On the other hand, Δp\Delta_{p} uses

2ap+12β2^{a_{p}+1}\leq 2^{\beta}

colors. So all together, ψp\psi_{p} uses at most 24(p+1)βplog2β+β2^{4(p+1)\beta^{p}\log_{2}\beta+\beta} colors, where

(log2n)1/(p+1)β<(log2n)1/(p+1)+1.(\log_{2}n)^{1/(p+1)}\leq\beta<(\log_{2}n)^{1/(p+1)}+1.

Thus, for any fixed pp, ψp\psi_{p} uses a total of no(1)n^{o(1)} colors.

2.3 Refinement of functions

Before we prove Lemma 2.1, it will be helpful to give the following definition and results about refinement of functions. The definition and Lemma 2.3 are paraphrased from [3].

Definition.

Let f:ABf:A\rightarrow B and g:ACg:A\rightarrow C. We say that ff refines gg if f(a1)=f(a2)f(a_{1})=f(a_{2}) implies that g(a1)=g(a2)g(a_{1})=g(a_{2}) for all a1,a2Aa_{1},a_{2}\in A.

Lemma 2.3 (Lemma 4.1(vi) from [3]).

Let f,gf,g be functions on domain AA. If ff refines gg, then for all AAA^{\prime}\subseteq A, we have |f(A)||g(A)||f(A^{\prime})|\geq|g(A^{\prime})|.

Lemma 2.4.

Let f,gf,g be functions on domain AA. If ff refines gg and SAS\subseteq A is a finite subset for which |f(S)|=|g(S)||f(S)|=|g(S)|, then

f(s1)=f(s2)g(s1)=g(s2)f(s_{1})=f(s_{2})\iff g(s_{1})=g(s_{2})

for all s1,s2Ss_{1},s_{2}\in S.

Proof.

The forward direction follows from the definition of ff refining gg. Conversely, if we have g(s1)=g(s2)g(s_{1})=g(s_{2}) but f(s1)f(s2)f(s_{1})\neq f(s_{2}) for some s1,s2Ss_{1},s_{2}\in S, then |f(S)||g(S)|+1|f(S)|\geq|g(S)|+1, a contradiction. ∎

In particular, Lemma 2.4 implies that if some edge-coloring of a clique SS is refined by another edge-coloring, but SS contains the same number of colors under each, then the edge-colorings must be isomorphic.

2.4 Proof of Lemma 2.1

Let S{0,1}αS\subseteq\{0,1\}^{\alpha} be a set of |S|p+3|S|\leq p+3 vertices which contains exactly |S|1|S|-1 distinct edge colors under cpc_{p}. We will prove that SS either has a leftover structure or contains a striped K4K_{4} by induction on α\alpha, similar to the proof of Theorem 2.2 from [3].

For the base case, consider αrp\alpha\leq r_{p}. Then for any x,ySx,y\in S, the first component of cp(x,y)c_{p}(x,y) is

ξp(x,y)=(ηp(x,y))=((1,{x,y})).\xi_{p}(x,y)=\left(\eta_{p}(x,y)\right)=\left((1,\{x,y\})\right).

Therefore, all of the edges of SS receive distinct colors. So it must be that |S|1=(|S|2)|S|-1=\binom{|S|}{2}, which happens only when |S|=1,2|S|=1,2. In either case, SS trivially has a leftover structure.

Now assume that α>rp\alpha>r_{p} and that the statement is true for shorter binary strings. For each d=1,,pd=1,\ldots,p, let αd\alpha_{d} be the largest integer strictly less than α\alpha that is divisible by rdr_{d}. For any xSx\in S, let x=(xd,xd′′)x=(x^{\prime}_{d},x^{\prime\prime}_{d}) for xd{0,1}αdx^{\prime}_{d}\in\{0,1\}^{\alpha_{d}} and xd′′{0,1}ααdx^{\prime\prime}_{d}\in\{0,1\}^{\alpha-\alpha_{d}}.

Let SdS_{d} denote the set of αd\alpha_{d}-prefixes of SS,

Sd={xd{0,1}αd|xS,x=(xd,xd′′)}.S_{d}=\left\{x_{d}^{\prime}\in\{0,1\}^{\alpha_{d}}|\exists x\in S,x=(x_{d}^{\prime},x_{d}^{\prime\prime})\right\}.

For each xdSdx_{d}^{\prime}\in S_{d}, let

Txd={xS|x=(xd,xd′′)}.T_{x_{d}^{\prime}}=\{x\in S|x=(x_{d}^{\prime},x_{d}^{\prime\prime})\}.

Let ΛI(d)\Lambda_{I}^{(d)} be the set of colors contained in SS found on edges that go between vertices from two distinct TT-sets,

ΛI(d)={cp(x,y)|x,yS;xdyd}.\Lambda_{I}^{(d)}=\{c_{p}(x,y)|x,y\in S;x_{d}^{\prime}\neq y_{d}^{\prime}\}.

Similarly, let ΛE(d)\Lambda_{E}^{(d)} denote the set of colors contained in SS found on edges between vertices from the same TT-set,

ΛE(d)={cp(x,y)|x,yS;xy;xd=yd}.\Lambda_{E}^{(d)}=\{c_{p}(x,y)|x,y\in S;x\neq y;x_{d}^{\prime}=y_{d}^{\prime}\}.

Note that these sets of colors, ΛI(d)\Lambda_{I}^{(d)} and ΛE(d)\Lambda_{E}^{(d)}, partition all of the colors contained in SS. Therefore,

|S|1=|ΛI(d)|+|ΛE(d)|.|S|-1=|\Lambda_{I}^{(d)}|+|\Lambda_{E}^{(d)}|.

Next, define

CI(d)\displaystyle C_{I}^{(d)} ={(cp(xd,yd),ηd1(xd′′,yd′′))|x,yS;xdyd}\displaystyle=\left\{(c_{p}(x_{d}^{\prime},y_{d}^{\prime}),\eta_{d-1}(x_{d}^{\prime\prime},y_{d}^{\prime\prime}))|x,y\in S;x_{d}^{\prime}\neq y_{d}^{\prime}\right\}
CE(d)\displaystyle C_{E}^{(d)} ={{xd′′,yd′′}|x,yS;xy;xd=yd}.\displaystyle=\left\{\{x_{d}^{\prime\prime},y_{d}^{\prime\prime}\}|x,y\in S;x\neq y;x_{d}^{\prime}=y_{d}^{\prime}\right\}.

It is shown in [3] that |ΛI(d)||CI(d)||\Lambda_{I}^{(d)}|\geq|C_{I}^{(d)}| and that |ΛE(d)||CE(d)||\Lambda_{E}^{(d)}|\geq|C_{E}^{(d)}|. The second inequality is easier to see since any distinct x,ySx,y\in S for which xd=ydx_{d}^{\prime}=y_{d}^{\prime} give ξd=(0,,0,(i,{xd′′,yd′′}))\xi_{d}=\left(0,\ldots,0,(i,\{x_{d}^{\prime\prime},y_{d}^{\prime\prime}\})\right) as the appropriate component of cp(x,y)c_{p}(x,y). Although the first inequality seems intuitively true, its proof is a bit more subtle. The following Fact 2.5 (proved in [3]) together with Lemma 2.3 give us the desired inequality.

Fact 2.5 (Lemma 4.3 from [3]).

For x,y{0,1}αx,y\in\{0,1\}^{\alpha}, let

γd(x,y)=(cp(xd,yd),ηd1(xd′′,yd′′)).\gamma_{d}(x,y)=(c_{p}(x_{d}^{\prime},y_{d}^{\prime}),\eta_{d-1}(x_{d}^{\prime\prime},y_{d}^{\prime\prime})).

Then cpc_{p} refines γd\gamma_{d} as functions on domain {0,1}α×{0,1}α\{0,1\}^{\alpha}\times\{0,1\}^{\alpha}.

We will also use the following Fact 2.6 which is proven in [3], although not stated as a claim or lemma that can be easily cited. (See the final sentence of the second-to-final paragraph on page 11.)

Fact 2.6 (proved in [3]).

There exists an integer 1dp1\leq d\leq p for which

|CI(d)|+|CE(d)||S|1.|C_{I}^{(d)}|+|C_{E}^{(d)}|\geq|S|-1.

Therefore,

|S|1=|ΛI(d)|+|ΛE(d)||CI(d)|+|CE(d)||S|1,|S|-1=|\Lambda_{I}^{(d)}|+|\Lambda_{E}^{(d)}|\geq|C_{I}^{(d)}|+|C_{E}^{(d)}|\geq|S|-1,

which implies that

|S|1=|ΛI(d)|+|ΛE(d)|=|CI(d)|+|CE(d)|.|S|-1=|\Lambda_{I}^{(d)}|+|\Lambda_{E}^{(d)}|=|C_{I}^{(d)}|+|C_{E}^{(d)}|.

Let

c~p(x,y)={(cp(xd,yd),ηd1(xd′′,yd′′))if xdyd{xd′′,yd′′}otherwise.\tilde{c}_{p}(x,y)=\left\{\begin{array}[]{ll}(c_{p}(x_{d}^{\prime},y_{d}^{\prime}),\eta_{d-1}(x_{d}^{\prime\prime},y_{d}^{\prime\prime}))&\quad\text{if }x_{d}^{\prime}\neq y_{d}^{\prime}\\ \{x_{d}^{\prime\prime},y_{d}^{\prime\prime}\}&\quad\text{otherwise}.\end{array}\right.

Then by Fact 2.5 we know that c~p\tilde{c}_{p} refines cpc_{p}. And since |ΛI(d)|+|ΛE(d)|=|CI(d)|+|CE(d)||\Lambda_{I}^{(d)}|+|\Lambda_{E}^{(d)}|=|C_{I}^{(d)}|+|C_{E}^{(d)}|, then by Lemma 2.4 we know that the structure of SS under c~p\tilde{c}_{p} must be the same as the structure of SS under cpc_{p}. Therefore, we need only show that SS either has a leftover structure or contains a striped K4K_{4} under c~p\tilde{c}_{p} to complete the proof. We consider two cases: either there exists some ωCE(d)\omega\in C_{E}^{(d)} that appears more than once in SS under c~p\tilde{c}_{p} or each ωCE(d)\omega\in C_{E}^{(d)} appears exactly once in SS under c~p\tilde{c}_{p}.

Case 1: Let ωCE(d)\omega\in C_{E}^{(d)} appear on at least two edges in SS. This implies that ω={xd′′,yd′′}\omega=\{x_{d}^{\prime\prime},y_{d}^{\prime\prime}\} and so there must exist a,b,c,eSa,b,c,e\in S such that a=(xd,xd′′)a=(x_{d}^{\prime},x_{d}^{\prime\prime}), b=(xd,yd′′)b=(x_{d}^{\prime},y_{d}^{\prime\prime}), c=(yd,xd′′)c=(y_{d}^{\prime},x_{d}^{\prime\prime}), and e=(yd,yd′′)e=(y_{d}^{\prime},y_{d}^{\prime\prime}) for some xdydx_{d}^{\prime}\neq y_{d}^{\prime}. Therefore,

c~p(a,b)=c~p(c,e)\displaystyle\tilde{c}_{p}(a,b)=\tilde{c}_{p}(c,e) ={xd′′,yd′′}\displaystyle=\{x_{d}^{\prime\prime},y_{d}^{\prime\prime}\}
c~p(a,c)=c~p(b,e)\displaystyle\tilde{c}_{p}(a,c)=\tilde{c}_{p}(b,e) =(cp(xd,yd),0)\displaystyle=(c_{p}(x_{d}^{\prime},y_{d}^{\prime}),0)
c~p(a,e)=c~p(b,c)\displaystyle\tilde{c}_{p}(a,e)=\tilde{c}_{p}(b,c) =(cp(xd,yd),ηd1(xd′′,yd′′)),\displaystyle=(c_{p}(x_{d}^{\prime},y_{d}^{\prime}),\eta_{d-1}(x_{d}^{\prime\prime},y_{d}^{\prime\prime})),

and all three colors are distinct. Hence, SS contains a striped K4K_{4} under c~p\tilde{c}_{p}.

Case 2: If each ωCE(d)\omega\in C_{E}^{(d)} appears exactly once in SS under c~p\tilde{c}_{p}, then we know that

|CE(d)|=xdSd(|Txd|2)|C_{E}^{(d)}|=\sum_{x_{d}^{\prime}\in S_{d}}\binom{|T_{x_{d}^{\prime}}|}{2}

since each edge within a given TT-set receives a unique color. Moreover, if we let

CB(d)={cp(xd,yd)|xd,ydSd},C_{B}^{(d)}=\{c_{p}(x_{d}^{\prime},y_{d}^{\prime})|x_{d}^{\prime},y_{d}^{\prime}\in S_{d}\},

then we know that

|CI(d)||CB(d)||Sd|1.|C_{I}^{(d)}|\geq|C_{B}^{(d)}|\geq|S_{d}|-1.

Therefore,

|Sd|1+xdSd(|Txd|2)\displaystyle|S_{d}|-1+\sum_{x_{d}^{\prime}\in S_{d}}\binom{|T_{x_{d}^{\prime}}|}{2} |S|1\displaystyle\leq|S|-1
xdSd(|Txd|2)\displaystyle\sum_{x_{d}^{\prime}\in S_{d}}\binom{|T_{x_{d}^{\prime}}|}{2} |S||Sd|\displaystyle\leq|S|-|S_{d}|
xdSd(|Txd|2)\displaystyle\sum_{x_{d}^{\prime}\in S_{d}}\binom{|T_{x_{d}^{\prime}}|}{2} xdSd(|Txd|1)\displaystyle\leq\sum_{x_{d}^{\prime}\in S_{d}}(|T_{x_{d}^{\prime}}|-1)
xdSd(|Txd|1)(|Txd|2)\displaystyle\sum_{x_{d}^{\prime}\in S_{d}}(|T_{x_{d}^{\prime}}|-1)(|T_{x_{d}^{\prime}}|-2) 0.\displaystyle\leq 0.

Hence, |Txd|=1,2|T_{x_{d}^{\prime}}|=1,2 for each xdSdx_{d}^{\prime}\in S_{d}. This implies that |CE(d)|=xdSd(|Txd|1)|C_{E}^{(d)}|=\sum_{x_{d}^{\prime}\in S_{d}}(|T_{x_{d}^{\prime}}|-1) and |CI(d)|=|CB(d)|=|Sd|1|C_{I}^{(d)}|=|C_{B}^{(d)}|=|S_{d}|-1. So by induction, SdS_{d} either has a leftover structure or contains a striped K4K_{4} under cpc_{p}. Furthermore, the coloring defined by

cp(x,y)={cp(xd,yd)if xdyd{xd′′,yd′′}otherwisec^{\prime}_{p}(x,y)=\left\{\begin{array}[]{ll}c_{p}(x_{d}^{\prime},y_{d}^{\prime})&\quad\text{if }x_{d}^{\prime}\neq y_{d}^{\prime}\\ \{x_{d}^{\prime\prime},y_{d}^{\prime\prime}\}&\quad\text{otherwise}\end{array}\right.

is refined by c~p\tilde{c}_{p}, and SS contains exactly |S|1|S|-1 colors under both cpc^{\prime}_{p} and c~p\tilde{c}_{p}. So by Lemma 2.4, the edge-coloring of SS under c~p\tilde{c}_{p} is isomorphic to the one under cpc^{\prime}_{p}, and hence it is sufficient to show that SS has either a leftover structure or contains a striped K4K_{4} under cpc^{\prime}_{p}.

If SdS_{d} has a leftover structure under cpc_{p}, then we see that SS also has a leftover structure under cpc^{\prime}_{p} since we can form SS under cpc^{\prime}_{p} from SdS_{d} under cpc_{p} by a sequence of splits as described in the definition of a leftover structure. That is, for each xdSdx_{d}^{\prime}\in S_{d} for which |Txd|=2|T_{x_{d}^{\prime}}|=2, we replace xdx_{d}^{\prime} with two vertices with a new edge color between them, and the same edge colors that xdx_{d}^{\prime} already had to the rest of the vertices.

On the other hand, if SdS_{d} contains a striped K4K_{4} under cpc_{p}, then SS must contain a striped K4K_{4} under cpc_{p}^{\prime} with colors entirely from CB(d)C_{B}^{(d)}. This concludes the proof.

2.5 Proof of Lemma 2.2

Let a,b,c,d{0,1}αa,b,c,d\in\{0,1\}^{\alpha} be four distinct vertices, and assume towards a contradiction that they form a striped K4K_{4} under ψp\psi_{p}. Specifically, assume that ψp(a,b)=ψp(c,d)\psi_{p}(a,b)=\psi_{p}(c,d), ψp(a,c)=ψp(b,d)\psi_{p}(a,c)=\psi_{p}(b,d), and ψp(a,d)=ψp(b,c)\psi_{p}(a,d)=\psi_{p}(b,c).

Without loss of generality, we may assume the following: that aa is the minimum element of the four under the lexicographic ordering of {0,1}α\{0,1\}^{\alpha}; that for some ij,ki\leq j,k,

ηp(a,b)=ηp(c,d)\displaystyle\eta_{p}(a,b)=\eta_{p}(c,d) =(i,{x,y})\displaystyle=(i,\{x,y\})
ηp(a,c)=ηp(b,d)\displaystyle\eta_{p}(a,c)=\eta_{p}(b,d) =(j,{z,w})\displaystyle=(j,\{z,w\})
ηp(a,d)=ηp(b,c)\displaystyle\eta_{p}(a,d)=\eta_{p}(b,c) =(k,{s,t});\displaystyle=(k,\{s,t\});

and that ai(p)=ci(p)=xa_{i}^{(p)}=c_{i}^{(p)}=x while bi(p)=di(p)=yb_{i}^{(p)}=d_{i}^{(p)}=y. It follows from the ordering that x<yx<y and that a<c<b,da<c<b,d. Furthermore, we have i<ji<j since aa and cc do not differ in the ithi^{th} block. Similarly, we see that (k,{s,t})=(i,{x,y})(k,\{s,t\})=(i,\{x,y\}). Without loss of generality, we may let aj(p)=bj(p)=za_{j}^{(p)}=b_{j}^{(p)}=z and cj(p)=dj(p)=wc_{j}^{(p)}=d_{j}^{(p)}=w. Therefore, z<wz<w and a<c<b<da<c<b<d.

Now, it follows that δj(a,d)=+1\delta_{j}(a,d)=+1 and that δj(c,b)=1\delta_{j}(c,b)=-1, a contradiction since we assume that ψp(a,d)=ψp(c,b)\psi_{p}(a,d)=\psi_{p}(c,b).

3 Modified Dot Product coloring

Fix an odd prime power qq and a positive integer dd. In this section, we prove Theorem 1.2 by giving an edge-coloring φd\varphi_{d} for the complete graph on n=(q1)dn=(q-1)^{d} vertices that uses (3d+1)q1(3d+1)q-1 colors and contains no leftover 6-cliques when d=3d=3 and no leftover 8-cliques when d=4d=4.

In what follows, we make use of several standard concepts and results from linear algebra without providing explicit definitions or proofs. We highly recommend Linear Algebra Methods in Combinatorics by László Babai and Péter Frankl [1] for a detailed treatment of these ideas. In particular, Chapter 2 covers all of the necessary background for our argument.

3.1 The construction

Let 𝔽q\mathbb{F}_{q}^{*} denote the nonzero elements of the finite field with qq elements, and let (𝔽q)d(\mathbb{F}_{q}^{*})^{d} denote the set of ordered dd-tuples of elements from 𝔽q\mathbb{F}_{q}^{*}. In other words, (𝔽q)d(\mathbb{F}_{q}^{*})^{d} is the set of dd-dimensional vectors over the field 𝔽q\mathbb{F}_{q} without zero components. In what follows, we will assume that the set 𝔽q\mathbb{F}_{q}^{*} is endowed with a linear order which can be arbitrarily chosen. We then order the set (𝔽q)d(\mathbb{F}_{q}^{*})^{d} with lexicographic ordering based on the order applied to 𝔽q\mathbb{F}_{q}^{*}.

Define a set of colors CdC_{d} as the disjoint union

Cd=DOTZEROUPDOWN,C_{d}=\text{DOT}\sqcup\text{ZERO}\sqcup\text{UP}\sqcup\text{DOWN},

where DOT=𝔽q\text{DOT}=\mathbb{F}_{q}^{*}, and ZERO, UP, and DOWN are each copies of the set {1,,d}×𝔽q\{1,\ldots,d\}\times\mathbb{F}_{q}. Let

φd:((𝔽q)d2)Cd\varphi_{d}:\binom{(\mathbb{F}_{q}^{*})^{d}}{2}\rightarrow C_{d}

be a coloring function of pairs of distinct vectors, x<yx<y, defined by

φd(x,y)={(i,xi+yi)ZEROif xy=0(i,xi+yi)UPif xy0 and xy=xx(i,xi+yi)DOWNif xy{0,xx} and xy=yyxyotherwise\varphi_{d}\left(x,y\right)=\left\{\begin{array}[]{ll}(i,x_{i}+y_{i})_{\text{ZERO}}&\quad\text{if }x\cdot y=0\\ (i,x_{i}+y_{i})_{\text{UP}}&\quad\text{if }x\cdot y\neq 0\text{ and }x\cdot y=x\cdot x\\ (i,x_{i}+y_{i})_{\text{DOWN}}&\quad\text{if }x\cdot y\not\in\{0,x\cdot x\}\text{ and }x\cdot y=y\cdot y\\ x\cdot y&\quad\text{otherwise}\\ \end{array}\right.

where ii is the first coordinate for which x=(x1,,xd)x=(x_{1},\ldots,x_{d}) differs from y=(y1,,yd)y=(y_{1},\ldots,y_{d}) and xyx\cdot y denotes the standard inner product (dot product).

3.2 Number of colors

Let nn be a positive integer. Let qq be the smallest odd prime power for which n(q1)dn\leq(q-1)^{d}. Then we can color the edges of KnK_{n} by arbitrarily associating each vertex with a unique vector from (𝔽q)d(\mathbb{F}_{q}^{*})^{d} and taking the coloring induced by φd\varphi_{d}. By Bertrand’s Postulate, q2(n1/d+1)q\leq 2(n^{1/d}+1). Therefore, the number of colors used by φd\varphi_{d} on KnK_{n} is at most

(3d+1)q1(6d+2)n1/d+(6d+1).(3d+1)q-1\leq(6d+2)n^{1/d}+(6d+1).

3.3 Definitions and lemmas

Definition.

Given a subset of vectors S𝔽dS\subseteq\mathbb{F}^{d}, let rk(S)\text{rk}(S) denote the rank of the subset, the dimension of the linear subspace spanned by the vectors of SS. Let af(S)\text{af}(S) denote the affine dimension of SS, the dimension of the affine subspace (also known as the affine hull) spanned by SS.

Definition.

A color αCd\alpha\in C_{d} has the dot property if αDOTZERO\alpha\in\text{DOT}\cup\text{ZERO}. Note that if α\alpha has the dot property, then φd(a,b)=φd(e,f)=α\varphi_{d}\left(a,b\right)=\varphi_{d}\left(e,f\right)=\alpha implies that ab=efa\cdot b=e\cdot f for any a,b,e,f(𝔽q)da,b,e,f\in(\mathbb{F}_{q}^{*})^{d}.

Lemma 3.1.

Let {s1,,st}(𝔽q)d\{s_{1},\ldots,s_{t}\}\subseteq(\mathbb{F}_{q}^{*})^{d} be a set of tt linearly independent vectors and let a,b(𝔽q)da,b\in(\mathbb{F}_{q}^{*})^{d} such that

φd(a,b)=φd(a,si)\displaystyle\varphi_{d}\left(a,b\right)=\varphi_{d}\left(a,s_{i}\right) =α\displaystyle=\alpha
φd(b,si)\displaystyle\varphi_{d}\left(b,s_{i}\right) =β\displaystyle=\beta

for some α,βCd\alpha,\beta\in C_{d} and for each 1it1\leq i\leq t. Then s1,,st,bs_{1},\ldots,s_{t},b are linearly independent.

Proof.

Assume towards a contradiction that b=j=1tλjsjb=\sum_{j=1}^{t}\lambda_{j}s_{j} for some scalars λ1,,λt𝔽q\lambda_{1},\ldots,\lambda_{t}\in\mathbb{F}_{q}. We will first show that j=1tλj=1\sum_{j=1}^{t}\lambda_{j}=1.

If α\alpha\in DOT, then b=j=1tλjsjb=\sum_{j=1}^{t}\lambda_{j}s_{j} implies that

α=ab=j=1tλj(asj)=j=1tλjα.\alpha=a\cdot b=\sum_{j=1}^{t}\lambda_{j}(a\cdot s_{j})=\sum_{j=1}^{t}\lambda_{j}\alpha.

Therefore, j=1tλj=1\sum_{j=1}^{t}\lambda_{j}=1 since α\alpha\notin ZERO.

If α\alpha\notin DOT, then

ai+bi=ai+s1,i==ai+st,ia_{i}+b_{i}=a_{i}+s_{1,i}=\cdots=a_{i}+s_{t,i}

where ii is the first index of difference between aa and bb. Thus, sj,i=bis_{j,i}=b_{i} for all 1jt1\leq j\leq t. But then b=j=1tλjsjb=\sum_{j=1}^{t}\lambda_{j}s_{j} implies that

bi=j=1tλjsj,i=j=1tλjbi.b_{i}=\sum_{j=1}^{t}\lambda_{j}s_{j,i}=\sum_{j=1}^{t}\lambda_{j}b_{i}.

Hence, j=1tλj=1\sum_{j=1}^{t}\lambda_{j}=1 since bi0b_{i}\neq 0. Therefore, for any αCd\alpha\in C_{d} we have j=1tλj=1\sum_{j=1}^{t}\lambda_{j}=1.

Now, if β\beta has the dot property, then let β\beta^{\prime} denote bsjb\cdot s_{j} for all j=1,,tj=1,\ldots,t. We have

bb=j=1tλj(bsj)=j=1tλjβ=β.b\cdot b=\sum_{j=1}^{t}\lambda_{j}(b\cdot s_{j})=\sum_{j=1}^{t}\lambda_{j}\beta^{\prime}=\beta^{\prime}.

But this implies that βUPDOWN\beta\in\text{UP}\cup\text{DOWN}, contradicting that β\beta has the dot property.

So we must assume that β\beta does not have the dot property. It follows that

bk+s1,k==bk+st,kb_{k}+s_{1,k}=\cdots=b_{k}+s_{t,k}

where kk is the first index of difference between bb and s1s_{1}. Therefore, s1,k==st,ks_{1,k}=\cdots=s_{t,k}, and so

bk=j=1tλjsj,k=j=1tλjs1,k=s1,k,b_{k}=\sum_{j=1}^{t}\lambda_{j}s_{j,k}=\sum_{j=1}^{t}\lambda_{j}s_{1,k}=s_{1,k},

contradicting our choice of kk.

Since we reach a contradiction for all colors β\beta, it must be the case that s1,,st,bs_{1},\ldots,s_{t},b are linearly independent vectors, as desired. ∎

We now define a particular instance of leftover structure that will be useful in our arguments.

Definition.

We call the set of vectors S={s1,,st}(𝔽q)dS=\{s_{1},\ldots,s_{t}\}\subseteq(\mathbb{F}_{q}^{*})^{d} a tt-falling star under the coloring φd\varphi_{d} if φd(si,sj)=αi\varphi_{d}\left(s_{i},s_{j}\right)=\alpha_{i} for all 1j<it1\leq j<i\leq t. For any set of vectors T(𝔽q)dT\subseteq(\mathbb{F}_{q}^{*})^{d} under φd\varphi_{d}, let FS(T)\text{FS}(T) denote the maximum tt such that TT contains a tt-falling star.

sts_{t}st1s_{t-1}st2s_{t-2}s2s_{2}s1s_{1}αt\alpha_{t}αt1\alpha_{t-1}αt2\alpha_{t-2}α2\alpha_{2}\iddots
Figure 2: A tt-falling star.

The following result about these falling stars is an easy consequence of Lemma 3.1 which can be shown by induction on the number of vectors.

Corollary 3.2.

Let S={s1,,st}(𝔽q)dS=\{s_{1},\ldots,s_{t}\}\subseteq(\mathbb{F}_{q}^{*})^{d} be a tt-falling star under φd\varphi_{d}. Then the vectors s1,,st1s_{1},\ldots,s_{t-1} are linearly independent. Consequently, for any subset T(𝔽q)dT\subseteq(\mathbb{F}_{q}^{*})^{d},

rk(T)FS(T)1.\text{rk}(T)\geq\text{FS}(T)-1.

Moreover, if TT is contained within a monochromatic neighborhood of some other vector, then

rk(T)FS(T).\text{rk}(T)\geq\text{FS}(T).
Definition.

Let A,B𝔽qdA,B\subseteq\mathbb{F}_{q}^{d} be disjoint sets of vectors. We say that AA confines BB if for each aAa\in A, ax=aya\cdot x=a\cdot y for every x,yBx,y\in B.

Lemma 3.3.

Let A,B𝔽qdA,B\subseteq\mathbb{F}_{q}^{d} be disjoint sets of vectors such that AA confines BB. Then

af(B)drk(A).\text{af}(B)\leq d-\text{rk}(A).
Proof.

Let t=rk(A)t=\text{rk}(A), and let a1,,ata_{1},\ldots,a_{t} be linearly independent vectors from AA. Since AA confines BB, then for each aia_{i}, there exists an αi𝔽q\alpha_{i}\in\mathbb{F}_{q} such that aib=αia_{i}\cdot b=\alpha_{i} for all bBb\in B. Therefore, BB is a subset of the solution space for the matrix equation,

(a1at)(x)=(α1αt).\left(\begin{array}[]{c}-a_{1}-\\ \vdots\\ -a_{t}-\end{array}\right)\left(\begin{array}[]{c}\mid\\ x\\ \mid\end{array}\right)=\left(\begin{array}[]{c}\alpha_{1}\\ \vdots\\ \alpha_{t}\end{array}\right).

Since a1,,ata_{1},\ldots,a_{t} are linearly independent, the matrix of these tt vectors has full rank, and hence, the solution set is an affine space of dimension dtd-t, as desired. ∎

Lemma 3.4.

Let A,B(𝔽q)dA,B\subseteq(\mathbb{F}_{q}^{*})^{d} be disjoint sets of vectors and αCd\alpha\in C_{d} such that φd(a,b)=α\varphi_{d}(a,b)=\alpha for all aAa\in A and bBb\in B. Then either AA confines BB or BB confines AA (or both).

Proof.

If α\alpha has the dot property, then it is trivial that AA and BB confine one another. So assume that αUPDOWN\alpha\in\text{UP}\cup\text{DOWN}. It follows that the first position of difference ii is the same between any aAa\in A and any bBb\in B. Moreover, every vector of AA has the same ithi^{th} component, every vector of BB has the same ithi^{th} component, and every vector of ABA\cup B has the same jthj^{th} component for each 1j<i1\leq j<i if i>1i>1. Since the vectors are ordered lexicographically based on an underlying linear order of 𝔽q\mathbb{F}_{q}^{*}, it follows that either a<ba<b for all aAa\in A and bBb\in B, or b<ab<a for all aAa\in A and bBb\in B.

Without loss of generality, assume that a<ba<b for all aAa\in A and bBb\in B. If αUP\alpha\in\text{UP}, then for any particular aAa\in A, ab=aaa\cdot b=a\cdot a for every bBb\in B. Therefore, AA confines BB. Similarly, if αDOWN\alpha\in\text{DOWN}, then for any particular bBb\in B, ba=bbb\cdot a=b\cdot b for every aAa\in A, so BB confines AA. ∎

Lemma 3.5.

Let t2t\geq 2 be an integer. An affine subspace of 𝔽qd\mathbb{F}_{q}^{d} of dimension t2t-2 will contain no tt-falling stars of (𝔽q)d(\mathbb{F}_{q}^{*})^{d} under φd\varphi_{d}. Therefore,

af(S)FS(S)1\text{af}(S)\geq\text{FS}(S)-1

for any subset of vectors S(𝔽q)dS\subseteq(\mathbb{F}_{q}^{*})^{d}.

Proof.

We will proceed by induction on tt. The base case t=2t=2 is trivial since an affine subspace of dimension 0 is just one vector while a 22-falling star contains two distinct vectors.

So assume that t3t\geq 3 and that the statement is true for t1t-1. Let s1,,sts_{1},\ldots,s_{t} be tt distinct vectors that form a tt-falling star. That is, let α1,,αt1Cd\alpha_{1},\ldots,\alpha_{t-1}\in C_{d} and let φd(si,sj)=αi\varphi_{d}\left(s_{i},s_{j}\right)=\alpha_{i} when 1i<jt1\leq i<j\leq t. Assume towards a contradiction that these vectors are contained inside an affine subspace of dimension t2t-2. Then there exist scalars λ1,,λt1𝔽q\lambda_{1},\ldots,\lambda_{t-1}\in\mathbb{F}_{q} such that st=j=1t1λjsjs_{t}=\sum_{j=1}^{t-1}\lambda_{j}s_{j} and j=1t1λj=1\sum_{j=1}^{t-1}\lambda_{j}=1.

First, note that if λ1=0\lambda_{1}=0, then the vectors s2,,sts_{2},\ldots,s_{t} form a (t1)(t-1)-falling star and are contained in an affine subspace of dimension t3t-3, a contradiction of the inductive hypothesis. So we must assume in what follows that λ10\lambda_{1}\neq 0.

Now, we consider two cases: either α1DOT\alpha_{1}\in\text{DOT} or α1DOT\alpha_{1}\not\in\text{DOT}. If α1DOT\alpha_{1}\in\text{DOT}, then

α1=s1st=s1j=1t1λjsj=λ1(s1s1)+α1j=2t1λj=λ1(s1s1)+α1(1λ1).\alpha_{1}=s_{1}\cdot s_{t}=s_{1}\cdot\sum_{j=1}^{t-1}\lambda_{j}s_{j}=\lambda_{1}(s_{1}\cdot s_{1})+\alpha_{1}\sum_{j=2}^{t-1}\lambda_{j}=\lambda_{1}(s_{1}\cdot s_{1})+\alpha_{1}(1-\lambda_{1}).

Therefore, λ1(s1s1α1)=0\lambda_{1}(s_{1}\cdot s_{1}-\alpha_{1})=0. Since λ10\lambda_{1}\neq 0, it follows that

s1s1=α1=s1s2,s_{1}\cdot s_{1}=\alpha_{1}=s_{1}\cdot s_{2},

which that implies α1DOT\alpha_{1}\not\in\text{DOT}, a contradiction.

So assume that α1DOT\alpha_{1}\not\in\text{DOT}, and let ii denote the index of the first component where s1s_{1} differs from the other vectors. In this case,

s1,i+s2,i==s1,i+st,i,s_{1,i}+s_{2,i}=\cdots=s_{1,i}+s_{t,i},

and hence s2,i==st,is_{2,i}=\cdots=s_{t,i}. Therefore,

st,i=j=1t1λjsj,i=λ1s1,i+st,ij=2t1λj=λ1s1,i+st,i(1λ1).s_{t,i}=\sum_{j=1}^{t-1}\lambda_{j}s_{j,i}=\lambda_{1}s_{1,i}+s_{t,i}\sum_{j=2}^{t-1}\lambda_{j}=\lambda_{1}s_{1,i}+s_{t,i}(1-\lambda_{1}).

So λ1(s1,ist,i)=0\lambda_{1}(s_{1,i}-s_{t,i})=0. Since λ10\lambda_{1}\neq 0, we have s1,i=st,is_{1,i}=s_{t,i}, a contradiction of our choice of ii. ∎

Lemma 3.6.

Let S(𝔽q)dS\subseteq(\mathbb{F}_{q}^{*})^{d} be a set of p1p\geq 1 vectors with a leftover structure under the coloring φd\varphi_{d}. Then

FS(S)log2p+1.\text{FS}(S)\geq\left\lceil\log_{2}{p}\right\rceil+1.
Proof.

We will prove this by induction on pp. The base case when p=1p=1 is trivial, so assume that SS has p2p\geq 2 vectors. Then SS has an initial bipartition, S=ABS=A\cup B, and we note that

FS(S)1+max(FS(A),FS(B)).\text{FS}(S)\geq 1+\max\left(\text{FS}(A),\text{FS}(B)\right).

Since |A|,|B|<p|A|,|B|<p, then by induction FS(T)log2(|T|)+1\text{FS}(T)\geq\left\lceil\log_{2}(|T|)\right\rceil+1 for T=A,BT=A,B. Thus, we have

FS(S)log2(max(|A|,|B|))+2,\text{FS}(S)\geq\left\lceil\log_{2}\left(\max(|A|,|B|)\right)\right\rceil+2,

and since max(|A|,|B|)p2\max(|A|,|B|)\geq\left\lceil\frac{p}{2}\right\rceil, then

FS(S)log2(p2)+2=log2p+1.FS(S)\geq\left\lceil\log_{2}\left(\left\lceil\frac{p}{2}\right\rceil\right)\right\rceil+2=\left\lceil\log_{2}{p}\right\rceil+1.

Lemma 3.7.

Let p2p\geq 2 and T0T\geq 0 be integers. Let S(𝔽q)dS\subseteq(\mathbb{F}_{q}^{*})^{d} be a subset of pp vectors with a leftover structure under φd\varphi_{d}. If T1T\geq 1, let a1,,aT(𝔽q)da_{1},\ldots,a_{T}\in(\mathbb{F}_{q}^{*})^{d} and α1,,αTCd\alpha_{1},\ldots,\alpha_{T}\in C_{d} such that φd(ai,aj)=αi\varphi_{d}\left(a_{i},a_{j}\right)=\alpha_{i} for all 1i<jT1\leq i<j\leq T and φd(ai,s)=αi\varphi_{d}\left(a_{i},s\right)=\alpha_{i} for all 1iT1\leq i\leq T and all sSs\in S.

Then there exists a sequence of positive integers, x1,,xtx_{1},\ldots,x_{t} such that i=1txi=p1\sum_{i=1}^{t}x_{i}=p-1 and for each i=1,,ti=1,\ldots,t, the following three conditions hold:

  1. 1.

    1xipsi21\leq x_{i}\leq\left\lfloor\frac{p-s_{i}}{2}\right\rfloor;

  2. 2.

    log2(xi)+log2(psixi)d1\left\lceil\log_{2}(x_{i})\right\rceil+\left\lceil\log_{2}(p-s_{i}-x_{i})\right\rceil\leq d-1;

  3. 3.

    log2(psixi)diT\left\lceil\log_{2}(p-s_{i}-x_{i})\right\rceil\leq d-i-T,

where si=0s_{i}=0 if i=1i=1 and si=j=1i1xjs_{i}=\sum_{j=1}^{i-1}x_{j} otherwise.

Proof.

We will prove this by induction on pp. For the base case, let p=2p=2. Let x1=1x_{1}=1 be the entire sequence. Then the first two conditions hold trivially since the sum of the sequence is 1, and since

log2(1)+log2(1)=0d1\left\lceil\log_{2}(1)\right\rceil+\left\lceil\log_{2}(1)\right\rceil=0\leq d-1

for any d1d\geq 1. For the third condition, since log2(1)=0\left\lceil\log_{2}(1)\right\rceil=0, it suffices to show that T+1dT+1\leq d. This follows from Corollary 3.2, since S{a1,,aT}S\cup\{a_{1},\ldots,a_{T}\} forms a (T+2)(T+2)-falling star, and hence drk(S{a1,,aT})T+1d\geq\text{rk}\left(S\cup\{a_{1},\ldots,a_{T}\}\right)\geq T+1.

So assume that SS is a set of pp vertices for p3p\geq 3 and that the statement is true for smaller sets. Let the initial bipartition of SS be S=ABS=A\cup B. By Lemma 3.4, we may assume without loss of generality that AA confines BB. Therefore, af(B)drk(A)\text{af}(B)\leq d-\text{rk}(A) by Lemma 3.3. By Corollary 3.2, we know that rk(A)FS(A)\text{rk}(A)\geq\text{FS}(A) since AA is in a monochromatic neighborhood of any vector from BB. And by Lemma 3.5, we know that af(B)FS(B)1\text{af}(B)\geq\text{FS}(B)-1. Thus, FS(A)+FS(B)1d\text{FS}(A)+\text{FS}(B)-1\leq d. So by Lemma 3.6, we can conclude that

log2(|A|)+log2(|B|)d1.\left\lceil\log_{2}(|A|)\right\rceil+\left\lceil\log_{2}(|B|)\right\rceil\leq d-1.

Therefore, setting x1=min{|A|,|B|}x_{1}=\min\{|A|,|B|\} guarantees that 1x1p21\leq x_{1}\leq\left\lfloor\frac{p}{2}\right\rfloor and that

log2(x1)+log2(px1)d1.\left\lceil\log_{2}(x_{1})\right\rceil+\left\lceil\log_{2}(p-x_{1})\right\rceil\leq d-1.

This gives us a positive integer x1x_{1} which satisfies the first two conditions. Moreover, by Corollary 3.2 and Lemma 3.6,

drk(S{a1,,aT})\displaystyle d\geq\text{rk}\left(S\cup\{a_{1},\ldots,a_{T}\}\right) FS(S{a1,,aT})1\displaystyle\geq\text{FS}(S\cup\{a_{1},\ldots,a_{T}\})-1
(T+1+max(FS(A),FS(B)))1\displaystyle\geq(T+1+\max\left(\text{FS}(A),\text{FS}(B)\right))-1
T+log2(px1)+1.\displaystyle\geq T+\left\lceil\log_{2}(p-x_{1})\right\rceil+1.

Thus, x1x_{1} also satisfies the third condition.

Let SS^{\prime} denote the larger of the two parts AA and BB, and let aT+1a_{T+1} denote an arbitrary vector from SSS\setminus S^{\prime}. Then SS^{\prime} contains px1<pp-x_{1}<p vectors and has a leftover structure under φd\varphi_{d}. Moreover, SS^{\prime} and a1,,aT,aT+1a_{1},\ldots,a_{T},a_{T+1} satisfy the monochromatic neighborhood conditions of the hypothesis. Hence, by induction there exists a sequence of positive integers x1,,xtx_{1}^{\prime},\ldots,x_{t^{\prime}}^{\prime} such that i=1txi=px11\sum_{i=1}^{t^{\prime}}x_{i}^{\prime}=p-x_{1}-1 and for each i=1,,ti=1,\ldots,t^{\prime}, the following three conditions hold:

  1. 1.

    1xipx1si21\leq x^{\prime}_{i}\leq\left\lfloor\frac{p-x_{1}-s^{\prime}_{i}}{2}\right\rfloor;

  2. 2.

    log2(xi)+log2(px1sixi)d1\left\lceil\log_{2}(x^{\prime}_{i})\right\rceil+\left\lceil\log_{2}(p-x_{1}-s^{\prime}_{i}-x^{\prime}_{i})\right\rceil\leq d-1;

  3. 3.

    log2(px1sixi)di(T+1)\left\lceil\log_{2}(p-x_{1}-s^{\prime}_{i}-x^{\prime}_{i})\right\rceil\leq d-i-(T+1),

where si=0s^{\prime}_{i}=0 if i=1i=1 and si=j=1i1xjs^{\prime}_{i}=\sum_{j=1}^{i-1}x^{\prime}_{j} otherwise.

Let xi=xi1x_{i}=x^{\prime}_{i-1} for 2it+12\leq i\leq t^{\prime}+1 and let t=t+1t=t^{\prime}+1 to get a sequence x1,,xtx_{1},\ldots,x_{t} for which

i=1txi=x1+i=1txi=x1+px11=p1.\sum_{i=1}^{t}x_{i}=x_{1}+\sum_{i=1}^{t^{\prime}}x_{i}^{\prime}=x_{1}+p-x_{1}-1=p-1.

For each i=2,,ti=2,\ldots,t, the first two conditions are satisfied since x1+si=si+1x_{1}+s^{\prime}_{i}=s_{i+1}, and the third condition is satisfied since di(T+1)=d(i+1)Td-i-(T+1)=d-(i+1)-T. ∎

Corollary 3.8.

Let S(𝔽q)3S\subseteq(\mathbb{F}_{q}^{*})^{3} be a set of 6 vectors. Then SS cannot have a leftover structure under the coloring φ3\varphi_{3}.

Proof.

If such a set exists, then by Lemma 3.7 with T=0T=0, a positive integer x1x_{1} exists such that 1x131\leq x_{1}\leq 3 and

log2(x1)+log2(6x1)2.\left\lceil\log_{2}(x_{1})\right\rceil+\left\lceil\log_{2}(6-x_{1})\right\rceil\leq 2.

It is simple to check that no such integer exists. ∎

Corollary 3.9.

Let S(𝔽q)4S\subseteq(\mathbb{F}_{q}^{*})^{4} be a set of 8 vectors. Then SS cannot have a leftover structure under the coloring φ4\varphi_{4}.

Proof.

If such a set exists, then by Lemma 3.7 with T=0T=0, we must be able to find a sequence of positive integers x1,x2,,xtx_{1},x_{2},\ldots,x_{t} that satisfy the conditions given in the Lemma. In particular, 1x141\leq x_{1}\leq 4 and

log2(x1)+log2(8x1)3.\left\lceil\log_{2}(x_{1})\right\rceil+\left\lceil\log_{2}(8-x_{1})\right\rceil\leq 3.

We can check and find that x1=1x_{1}=1 is the only possibility. Therefore, 1x231\leq x_{2}\leq 3 such that

log2(7x2)\displaystyle\left\lceil\log_{2}(7-x_{2})\right\rceil 2\displaystyle\leq 2
log2(x2)+log2(7x2)\displaystyle\left\lceil\log_{2}(x_{2})\right\rceil+\left\lceil\log_{2}(7-x_{2})\right\rceil 3.\displaystyle\leq 3.

A quick check reveals that no such integer exists. ∎

Theorem 1.2 follows from Theorem 1.1 and Corollaries 3.8 and 3.9.

4 Conclusion

The proof of Lemma 3.7 actually shows which leftover pp-cliques can appear under φd\varphi_{d} for a particular dd. For example, this proof implies that the only leftover 55-clique that can appear under φ3\varphi_{3} is a monochromatic C4C_{4} contained inside a monochromatic neighborhood of one vertex (that is, an initial (1,4)(1,4)-bipartition with a (2,2)(2,2)-bipartition inside the part with four vertices). In [2], we handled this specific leftover structure by splitting each color class of φ3\varphi_{3} into four new colors determined by certain relations between vectors. While the current paper can be viewed as our attempt to fully generalize the coloring techniques used in [2] and [7], it does not generalize the splitting that was crucial for handling the final leftover 55-clique. Perhaps such a generalized splitting would be enough to give f(n,p,p)n1/(p2)+o(1)f(n,p,p)\leq n^{1/(p-2)+o(1)} for p6p\geq 6 or at least improve the best-known upper bounds for values of pp other than the two addressed in this paper.

Remark.

Since the completion of this work, Conlon, Pohoata, and Tyomkyn have emailed us that they have obtained a version of Theorem 1.1 independently.

References

  • [1] László Babai and Péter Frankl. Linear Algebra Methods in Combinatorics: With Applications to Geometry and Computer Science. Department of Computer Science, Univ. of Chicago, 1992.
  • [2] Alex Cameron and Emily Heath. A (5,5)(5,5)-colouring of Kn{K}_{n} with few colours. Combinatorics, Probability & Computing, 27(6):892–912, 2018.
  • [3] David Conlon, Jacob Fox, Choongbum Lee, and Benny Sudakov. The Erdős–Gyárfás problem on generalized Ramsey numbers. Proceedings of the London Mathematical Society, 110(1):1–18, 2015.
  • [4] Dennis Eichhorn and Dhruv Mubayi. Edge-coloring cliques with many colors on subcliques. Combinatorica, 20(3):441–444, 2000.
  • [5] Paul Erdős and András Gyárfás. A variant of the classical Ramsey problem. Combinatorica, 17(4):459–467, 1997.
  • [6] Dhruv Mubayi. Edge-coloring cliques with three colors on all 4-cliques. Combinatorica, 18(2):293–296, 1998.
  • [7] Dhruv Mubayi. An explicit construction for a Ramsey problem. Combinatorica, 24(2):313–324, 2004.