Hostname: page-component-76fb5796d-zzh7m Total loading time: 0 Render date: 2024-04-28T06:25:52.189Z Has data issue: false hasContentIssue false

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

Published online by Cambridge University Press:  24 November 2022

Alex Cameron
Affiliation:
Department of Mathematics, Vanderbilt University, Nashville, TN 37212, USA
Emily Heath*
Affiliation:
Department of Mathematics, Iowa State University, Ames, IA 50011, USA
*
*Corresponding author. Email: eheath@iastate.edu
Rights & Permissions [Opens in a new window]

Abstract

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

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

Type
Paper
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2022. Published by Cambridge University Press

1. Introduction

Let $p$ and $q$ be positive integers such that $1 \leq q \leq \binom{p}{2}$ . We say that an edge-colouring of a graph $G$ is a $(p,q)$ -colouring if any $p$ -clique of $G$ contains edges of at least $q$ distinct colours. In 1975, Erdős and Shelah [Reference Erdős5] posed the question of determining $f(n,p,q)$ , the minimum number of colours needed to give a $(p,q)$ -colouring of the complete graph on $n$ vertices, $K_n$ .

This function $f(n,p,q)$ is known as the Erdős-Gyárfás function after the authors of the first paper [Reference Erdős and Gyárfás6] to systematically study $(p,q)$ -colourings. The majority of their work focused on understanding the asymptotic behaviour of this function as $n \rightarrow \infty$ for fixed values of $p$ and $q$ . One of their primary results was a general upper bound of

\begin{align*} f(n,p,q) = O\left (n^{\frac{p-2}{\binom{p}{2}-q+1}}\right ) \end{align*}

obtained using the Lovász Local Lemma, while one of the main problems they left open was the determination of $q$ , given a fixed value of $p$ , for which $f(n,p,q) = \Omega (n^{\varepsilon })$ for some constant $\varepsilon$ , but $f(n,p,q-1)=n^{o(1)}$ . Towards this end, they found that

\begin{align*} n^{\frac{1}{p-2}}-1 \leq f(n,p,p) \leq cn^{\frac{2}{p-1}}, \end{align*}

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 $f(n,p,p-1)=n^{o(1)}$ .

In 2015, Conlon, Fox, Lee, and Sudakov [Reference Conlon, Fox, Lee and Sudakov3], building on work done on small cases by Mubayi and Eichhorn [Reference Eichhorn and Mubayi4, Reference Mubayi7], showed that $f(n,p,p-1)=n^{o(1)}$ by constructing an explicit $(p,p-1)$ -colouring using very few colours. In [Reference Cameron and Heath2], we slightly modified their colouring, which we call the CFLS colouring, and paired it with an ‘algebraic’ construction to show $f(n,5,5) \leq n^{1/3+o(1)}$ . 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 [Reference Mubayi8], where he gave an explicit construction showing $f(n,4,4) \leq n^{1/2+o(1)}$ .

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

Theorem 1.1. For any $p \geq 3$ , there is a $(p,p-1)$ -colouring of $K_n$ using $n^{o(1)}$ colours such that the only $p$ -cliques that contain exactly $p-1$ distinct edge colours are isomorphic (as edge-coloured graphs) to one of the edge-coloured $p$ -cliques given in the definition below.

Definition 1.1. Given an edge-colouring $f\;:\;E(K_n) \rightarrow C$ , we say that a subset $S \subseteq V(K_n)$ has a leftover structure under $f$ if either $|S|=1$ or there exists a bipartition (which we will call the initial bipartition) of $S$ into nonempty sets $A$ and $B$ for which

  • $A$ and $B$ each have a leftover structure under $f$ ;

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

  • there is a fixed colour $\alpha \in C$ such that $f(a,b)=\alpha$ for all $a \in A$ and all $b \in B$ and $\alpha \not \in f(A)$ and $\alpha \not \in f(B)$ .

See Figure 1 for an example of a leftover structure. Alternatively, a more constructive definition is to say that a $p$ -clique $S$ is leftover if either $p=1$ or if it can be formed from a leftover $(p-1)$ -clique by taking one of its vertices $x$ , making a copy $x'$ , colouring $xx'$ with a new colour, and colouring $x'y$ with the same colour as $xy$ for each $y \in S$ for which $y \neq x$ . Note that it is easy to see by induction that these $p$ -cliques always contain exactly $p-1$ colours.

Figure 1. An example of a $p$ -clique with a leftover structure.

One of the general difficulties in producing explicit $(p,q)$ -colourings is dealing with the large number of possible non-isomorphic ways to colour the edges of a $p$ -clique with fewer than $q$ colours in order to demonstrate that a construction avoids them. By identifying the ‘bad’ structures that are leftover after using only $n^{o(1)}$ colours, we are able to greatly reduce the amount of case-checking required in identifying $(p,p)$ -colourings, which would otherwise make this problem intractable for large $p$ .

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-colouring of $K_n$ that eliminates leftover $p$ -cliques also eliminates all leftover $P$ -cliques for any $P \geq p$ . Moreover, by Theorem 1.1, if this colouring uses $n^{\varepsilon +o(1)}$ colours, then $f(n,P,P) \leq n^{\varepsilon +o(1)}$ , as the product of this colouring with the one guaranteed in Theorem 1.1 will avoid any $P$ -clique with fewer than $P$ colours for each $P \geq p$ .

As a specific example, in [Reference Cameron and Heath2] we gave a $(5,5)$ -colouring of $K_n$ that used $n^{1/3+o(1)}$ colours. Since this colouring avoids leftover 5-cliques, then it also avoids leftover $P$ -cliques for all $P \geq 5$ . Therefore, if we take the product of this colouring with the appropriate one developed in Section 2 that eliminates all $6$ -cliques with 5 or fewer colours other than leftover $6$ -cliques, then we have a $(6,6)$ -colouring that uses only $n^{1/3+o(1)}$ colours, improving the best-known upper bound given above, $O(n^{2/5})$ .

In Section 3, we generalize the algebraic portion of our colouring in [Reference Cameron and Heath2], the ‘Modified Dot Product’ colouring, to a version that eliminates leftover $6$ -cliques with $O(n^{1/3})$ colours (making the above example redundant) and eliminates leftover $8$ -cliques with $O(n^{1/4})$ colours. By taking the product of these colourings with the appropriate ones developed in Section 2, this gives us the following theorem.

Theorem 1.2. We have the following upper bounds:

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

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

2. Modified CFLS colouring

In this section, we define an edge-colouring $\psi _p$ of the complete graph with vertex set $\{0,1\}^{\alpha }$ for some positive integer $\alpha$ . This construction is the product of two colourings, $\psi _p = c_p \times \Delta _p$ , where $c_p$ is the $(p+3,p+2)$ -colouring defined in [Reference Conlon, Fox, Lee and Sudakov3]. In many places, this section tracks parts of the proof given in [Reference Conlon, Fox, Lee and Sudakov3], 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 colouring $c_p$ .

Lemma 2.1. Let $p$ be a fixed positive integer. Any subset $S \subseteq \{0,1\}^{\alpha }$ with $|S| \leq p+3$ vertices that contains exactly $|S|-1$ distinct colours under the edge-colouring $c_p$ either has a leftover structure under $c_p$ or contains a striped $K_4$ under $c_p$ .

A striped $K_4$ , as described by the following definition, was first defined in [Reference Mubayi8].

Definition 2.1. Let $f\;:\;E(G) \rightarrow C$ be an edge-colouring of a graph $G$ . We call any 4-clique of $G$ , $\{a,b,c,d\} \subseteq V(G)$ , for which $f(ab)=f(cd)$ , $f(ac)=f(bd)$ , $f(ad)=f(bc)$ , $f(ab) \neq f(ac)$ , $f(ab) \neq f(ad)$ , and $f(ac) \neq f(ad)$ a striped $K_4$ .

We will also prove the following result about the colouring $\psi _p$ .

Lemma 2.2. There is no striped $K_4$ under the edge-colouring $\psi _p$ .

These two lemmas are enough to conclude that $\psi _p$ is a $(p+3,p+2)$ -colouring for which any clique $S$ with $|S| \leq p+3$ that contains exactly $|S|-1$ colours must have a leftover structure.

2.1 The construction

For some positive integer $p$ , let

\begin{align*} 1 \leq r_1 \leq r_2 \leq \cdots \leq r_p \end{align*}

be fixed positive integers such that $r_d | r_{d+1}$ for each $d=1,\ldots,p-1$ . These $r_i$ will be called the parameters of our edge-colouring.

For any $\alpha \geq r_p$ , let $n=2^{\alpha }$ , and associate each vertex of the complete graph $K_n$ with its own unique binary string of length $\alpha$ . For each $d=1,\ldots,p$ , let $\alpha = a_dr_d+b_d$ for positive integers $a_d,b_d$ such that $1 \leq b_d \leq r_d$ . For each string $x \in \{0,1\}^{\alpha }$ , we let

\begin{align*} x=\left (x_1^{(d)},x_2^{(d)},\ldots,x_{a_d+1}^{(d)}\right ) \end{align*}

where $x_i^{(d)}$ denotes a binary string in $\{0,1\}^{r_d}$ for each $i=1,\ldots,a_d$ and $x_{a_d+1}^{(d)}$ denotes a binary string from $\{0,1\}^{b_d}$ . We will call these substrings $r_d$ -blocks of $x$ , including the final one which may or may not actually have length equal to $r_d$ .

In the following definitions, we let $r_0=1$ and $r_{p+1}=\alpha$ . First, we define a function $\eta _d$ for any $d=0,\ldots,p$ on domain $\{0,1\}^{\beta } \times \{0,1\}^{\beta }$ where $\beta$ is any positive integer as

\begin{align*} \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 . \end{align*}

where $i$ denotes the minimum index for which $x_i^{(d)} \neq y_i^{(d)}$ .

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

\begin{align*} \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 ). \end{align*}

And let

\begin{align*} c_p(x,y) = \left (\xi _p(x,y),\ldots,\xi _0(x,y)\right ). \end{align*}

Next, we assume that the binary strings of $\{0,1\}^{\beta }$ are lexicographically ordered for every positive integer $\beta$ . For $1 \leq i \leq a_p+1$ and binary strings $x\lt y$ , define

\begin{align*} \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)} \gt y_i^{(p)}. \end{array} \right . \end{align*}

Let

\begin{align*} \Delta _p(x,y) = \left (\delta _{p,1}(x,y),\ldots,\delta _{p,a_p+1}(x,y)\right ). \end{align*}

Finally, let

\begin{align*} \psi _p(x,y) = \left (c_p(x,y),\Delta _p(x,y)\right ). \end{align*}

2.2 Number of colours

For any positive integer $n$ , let $\beta$ be the positive integer for which

\begin{align*} 2^{(\beta -1)^{p+1}}\lt n \leq 2^{\beta ^{p+1}}. \end{align*}

For each $d=1,\ldots,p+1$ , let $r_d=\beta ^d$ in the construction of $\psi _p$ . Specifically, this means we are constructing the colouring on the complete graph with vertex set $\{0,1\}^{\alpha }$ where $\alpha =\beta ^{p+1}$ . We can apply this colouring to $K_n$ by arbitrarily associating each vertex of $K_n$ with a unique binary string from $\{0,1\}^{\alpha }$ and taking the induced colouring.

As shown in [Reference Conlon, Fox, Lee and Sudakov3], for these choices of parameters $r_d$ , the colouring $c_p$ uses at most $2^{4(p+1)\beta ^p\log _2 \beta }$ colours. On the other hand, $\Delta _p$ uses

\begin{align*} 2^{a_p+1} \leq 2^{\beta } \end{align*}

colours. So all together, $\psi _p$ uses at most $2^{4(p+1)\beta ^p\log _2 \beta +\beta }$ colours, where

\begin{align*} (\log _2n)^{1/(p+1)} \leq \beta \lt (\log _2n)^{1/(p+1)} +1. \end{align*}

Thus, for any fixed $p$ , $\psi _p$ uses a total of $n^{o(1)}$ colours.

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 [Reference Conlon, Fox, Lee and Sudakov3].

Definition 2.2. Let $f\;:\;A \rightarrow B$ and $g\;:\;A \rightarrow C$ . We say that $f$ refines $g$ if $f(a_1)=f(a_2)$ implies that $g(a_1)=g(a_2)$ for all $a_1,a_2 \in A$ .

Lemma 2.3 (Lemma 4.1(vi) from [Reference Conlon, Fox, Lee and Sudakov3]). Let $f,g$ be functions on domain $A$ . If $f$ refines $g$ , then for all $A' \subseteq A$ , we have $|f(A')| \geq |g(A')|$ .

Lemma 2.4. Let $f,g$ be functions on domain $A$ . If $f$ refines $g$ and $S \subseteq A$ is a finite subset for which $|f(S)|=|g(S)|$ , then

\begin{align*} f(s_1)=f(s_2) \iff g(s_1)=g(s_2) \end{align*}

for all $s_1,s_2 \in S$ .

Proof. The forward direction follows from the definition of $f$ refining $g$ . Conversely, if we have $g(s_1)=g(s_2)$ but $f(s_1) \neq f(s_2)$ for some $s_1,s_2 \in S$ , then $|f(S)| \geq |g(S)|+1$ , a contradiction.

In particular, Lemma 2.4 implies that if some edge-colouring of a clique $S$ is refined by another edge-colouring, but $S$ contains the same number of colours under each, then the edge-colourings must be isomorphic.

2.4 Proof of Lemma 2.1

Let $S \subseteq \{0,1\}^{\alpha }$ be a set of $|S| \leq p+3$ vertices which contains exactly $|S|-1$ distinct edge colours under $c_p$ . We will prove that $S$ either has a leftover structure or contains a striped $K_4$ by induction on $\alpha$ , similar to the proof of Theorem 2.2 from [Reference Conlon, Fox, Lee and Sudakov3].

For the base case, consider $\alpha \leq r_p$ . Then for any $x,y \in S$ , the first component of $c_p(x,y)$ is

\begin{align*} \xi _p(x,y)=\left (\eta _p(x,y)\right )=\left ((1,\{x,y\})\right ). \end{align*}

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

Now assume that $\alpha \gt r_p$ and that the statement is true for shorter binary strings. For each $d=1,\ldots,p$ , let $\alpha _d$ be the largest integer strictly less than $\alpha$ that is divisible by $r_d$ . For any $x \in S$ , let $x=(x'_{\!\!d},x''_{\!\!\!d})$ for $x'_{\!\!d} \in \{0,1\}^{\alpha _d}$ and $x''_{\!\!\!d} \in \{0,1\}^{\alpha -\alpha _d}$ .

Let $S_d$ denote the set of $\alpha _d$ -prefixes of $S$ ,

\begin{align*} S_d=\left \{x'_{\!\!d} \in \{0,1\}^{\alpha _d} | \exists x \in S, x=(x'_{\!\!d},x''_{\!\!\!d}) \right \}. \end{align*}

For each $x'_{\!\!d} \in S_d$ , let

\begin{align*} T_{x'_{\!\!d}} = \{x \in S | x=(x'_{\!\!d},x''_{\!\!\!d})\}. \end{align*}

Let $\Lambda _I^{(d)}$ be the set of colours contained in $S$ found on edges that go between vertices from two distinct $T$ -sets,

\begin{align*} \Lambda _I^{(d)}=\{c_p(x,y) | x,y \in S;\; x'_{\!\!d} \neq y'_{\!\!d}\}. \end{align*}

Similarly, let $\Lambda _E^{(d)}$ denote the set of colours contained in $S$ found on edges between vertices from the same $T$ -set,

\begin{align*} \Lambda _E^{(d)} = \{c_p(x,y)| x,y \in S;\; x \neq y;\; x'_{\!\!d} = y'_{\!\!d}\}. \end{align*}

Note that these sets of colours, $\Lambda _I^{(d)}$ and $\Lambda _E^{(d)}$ , partition all of the colours contained in $S$ . Therefore,

\begin{align*} |S|-1=|\Lambda _I^{(d)}|+|\Lambda _E^{(d)}|. \end{align*}

Next, define

\begin{align*} C_I^{(d)} &= \left \{(c_p(x'_{\!\!d},y'_{\!\!d}),\eta _{d-1}(x''_{\!\!\!d},y''_{\!\!\!d}))|x,y \in S;\; x'_{\!\!d} \neq y'_{\!\!d}\right \}\\ C_E^{(d)} &= \left \{\{x''_{\!\!\!d},y''_{\!\!\!d}\}|x,y \in S;\; x \neq y;\; x'_{\!\!d} = y'_{\!\!d}\right \}.\\ \end{align*}

It is shown in [Reference Conlon, Fox, Lee and Sudakov3] that $|\Lambda _I^{(d)}| \geq |C_I^{(d)}|$ and that $|\Lambda _E^{(d)}| \geq |C_E^{(d)}|$ . The second inequality is easier to see since any distinct $x,y \in S$ for which $x'_{\!\!d} = y'_{\!\!d}$ give $\xi _d=\left (0,\ldots,0,(i,\{x''_{\!\!\!d},y''_{\!\!\!d}\})\right )$ as the appropriate component of $c_p(x,y)$ . Although the first inequality seems intuitively true, its proof is a bit more subtle. The following fact (proved in [Reference Conlon, Fox, Lee and Sudakov3]) together with Lemma 2.3 gives us the desired inequality.

Fact 2.1 (Lemma 4.3 from [Reference Conlon, Fox, Lee and Sudakov3]). For $x,y \in \{0,1\}^{\alpha }$ , let

\begin{align*} \gamma _d(x,y)=(c_p(x'_{\!\!d},y'_{\!\!d}),\eta _{d-1}(x''_{\!\!\!d},y''_{\!\!\!d})). \end{align*}

Then $c_p$ refines $\gamma _d$ as functions on domain $\{0,1\}^{\alpha } \times \{0,1\}^{\alpha }$ .

We will also use the following fact which is proven in [Reference Conlon, Fox, Lee and Sudakov3], 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 2.2 (proved in [Reference Conlon, Fox, Lee and Sudakov3]). There exists an integer $1 \leq d \leq p$ for which

\begin{align*} |C_I^{(d)}|+|C_E^{(d)}| \geq |S|-1. \end{align*}

Therefore,

\begin{align*} |S|-1=|\Lambda _I^{(d)}|+|\Lambda _E^{(d)}| \geq |C_I^{(d)}|+|C_E^{(d)}| \geq |S|-1, \end{align*}

which implies that

\begin{align*} |S|-1=|\Lambda _I^{(d)}|+|\Lambda _E^{(d)}| = |C_I^{(d)}|+|C_E^{(d)}|. \end{align*}

Let

\begin{align*} \tilde{c}_p(x,y) = \left \{ \begin{array}{ll} (c_p(x'_{\!\!d},y'_{\!\!d}),\eta _{d-1}(x''_{\!\!\!d},y''_{\!\!\!d})) & \quad \text{if }\; x'_{\!\!d} \neq y'_{\!\!d}\\ \{x''_{\!\!\!d},y''_{\!\!\!d}\} & \quad \text{otherwise}. \end{array} \right . \end{align*}

Then by Fact 2.1, we know that $\tilde{c}_p$ refines $c_p$ . And since $|\Lambda _I^{(d)}|+|\Lambda _E^{(d)}| = |C_I^{(d)}|+|C_E^{(d)}|$ , then by Lemma 2.4, we know that the structure of $S$ under $\tilde{c}_p$ must be the same as the structure of $S$ under $c_p$ . Therefore, we need only show that $S$ either has a leftover structure or contains a striped $K_4$ under $\tilde{c}_p$ to complete the proof. We consider two cases: either there exists some $\omega \in C_E^{(d)}$ that appears more than once in $S$ under $\tilde{c}_p$ or each $\omega \in C_E^{(d)}$ appears exactly once in $S$ under $\tilde{c}_p$ .

Case 1: Let $\omega \in C_E^{(d)}$ appear on at least two edges in $S$ . This implies that $\omega =\{x''_{\!\!\!d},y''_{\!\!\!d}\}$ and so there must exist $a,b,c,e \in S$ such that $a=(x'_{\!\!d},x''_{\!\!\!d})$ , $b=(x'_{\!\!d},y''_{\!\!\!d})$ , $c=(y'_{\!\!d},x''_{\!\!\!d})$ , and $e=(y'_{\!\!d},y''_{\!\!\!d})$ for some $x'_{\!\!d} \neq y'_{\!\!d}$ . Therefore,

\begin{align*} \tilde{c}_p(a,b)=\tilde{c}_p(c,e)&=\{x''_{\!\!\!d},y''_{\!\!\!d}\}\\ \tilde{c}_p(a,c)=\tilde{c}_p(b,e)&=(c_p(x'_{\!\!d},y'_{\!\!d}),0)\\ \tilde{c}_p(a,e)=\tilde{c}_p(b,c)&=(c_p(x'_{\!\!d},y'_{\!\!d}),\eta _{d-1}(x''_{\!\!\!d},y''_{\!\!\!d})), \end{align*}

and all three colours are distinct. Hence, $S$ contains a striped $K_4$ under $\tilde{c}_p$ .

Case 2: If each $\omega \in C_E^{(d)}$ appears exactly once in $S$ under $\tilde{c}_p$ , then we know that

\begin{align*} |C_E^{(d)}| = \sum _{x'_{\!\!d} \in S_d}\binom{|T_{x'_{\!\!d}}|}{2} \end{align*}

since each edge within a given $T$ -set receives a unique colour. Moreover, if we let

\begin{align*} C_B^{(d)} = \{c_p(x'_{\!\!d},y'_{\!\!d}) | x'_{\!\!d},y'_{\!\!d} \in S_d\}, \end{align*}

then we know that

\begin{align*} |C_I^{(d)}| \geq |C_B^{(d)}| \geq |S_d|-1. \end{align*}

Therefore,

\begin{align*} |S_d|-1+\sum _{x'_{\!\!d} \in S_d}\binom{|T_{x'_{\!\!d}}|}{2} &\leq |S|-1\\[4pt] \sum _{x'_{\!\!d} \in S_d}\binom{|T_{x'_{\!\!d}}|}{2} &\leq |S|-|S_d|\\[4pt] \sum _{x'_{\!\!d} \in S_d}\binom{|T_{x'_{\!\!d}}|}{2} &\leq \sum _{x'_{\!\!d} \in S_d} (|T_{x'_{\!\!d}}|-1)\\[4pt] \sum _{x'_{\!\!d} \in S_d}(|T_{x'_{\!\!d}}|-1)(|T_{x'_{\!\!d}}|-2) &\leq 0. \end{align*}

Hence, we have $|T_{x'_{\!\!d}}|=1,2$ for each $x'_{\!\!d} \in S_d$ . This implies that $|C_E^{(d)}|=\sum _{x'_{\!\!d} \in S_d} (|T_{x'_{\!\!d}}|-1)$ and $|C_I^{(d)}|=|C_B^{(d)}|=|S_d|-1$ . So by induction, $S_d$ either has a leftover structure or contains a striped $K_4$ under $c_p$ . Furthermore, the colouring defined by

\begin{align*} c'_{\!\!p}(x,y) = \left \{ \begin{array}{ll} c_p(x'_{\!\!d},y'_{\!\!d}) & \quad \text{if }\; x'_{\!\!d} \neq y'_{\!\!d}\\ \{x''_{\!\!\!d},y''_{\!\!\!d}\} & \quad \text{otherwise} \end{array} \right . \end{align*}

is refined by $\tilde{c}_p$ , and $S$ contains exactly $|S|-1$ colours under both $c'_{\!\!p}$ and $\tilde{c}_p$ . So by Lemma 2.4, the edge-colouring of $S$ under $\tilde{c}_p$ is isomorphic to the one under $c'_{\!\!p}$ , and hence it is sufficient to show that $S$ has either a leftover structure or contains a striped $K_4$ under $c'_{\!\!p}$ .

If $S_d$ has a leftover structure under $c_p$ , then we see that $S$ also has a leftover structure under $c'_{\!\!p}$ since we can form $S$ under $c'_{\!\!p}$ from $S_d$ under $c_p$ by a sequence of splits as described in the definition of a leftover structure. That is, for each $x'_{\!\!d} \in S_d$ for which $|T_{x'_{\!\!d}}| = 2$ , we replace $x'_{\!\!d}$ with two vertices with a new edge colour between them, and the same edge colours that $x'_{\!\!d}$ already had to the rest of the vertices.

On the other hand, if $S_d$ contains a striped $K_4$ under $c_p$ , then $S$ must contain a striped $K_4$ under $c'_{\!\!p}$ with colours entirely from $C_B^{(d)}$ . This concludes the proof.

2.5 Proof of Lemma 2.2

Let $a,b,c,d \in \{0,1\}^{\alpha }$ be four distinct vertices, and assume towards a contradiction that they form a striped $K_4$ under $\psi _p$ . Specifically, assume that $\psi _p(a,b)=\psi _p(c,d)$ , $\psi _p(a,c)=\psi _p(b,d)$ , and $\psi _p(a,d)=\psi _p(b,c)$ .

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

\begin{align*} \eta _p(a,b) = \eta _p(c,d) &= (i,\{x,y\})\\ \eta _p(a,c) = \eta _p(b,d) &= (j,\{z,w\})\\ \eta _p(a,d) = \eta _p(b,c) &= (k,\{s,t\}); \end{align*}

and that $a_i^{(p)}=c_i^{(p)}=x$ while $b_i^{(p)}=d_i^{(p)}=y$ . It follows from the ordering that $x \lt y$ and that $a\lt c\lt b,d$ . Furthermore, we have $i\lt j$ since $a$ and $c$ do not differ in the $i^{th}$ block. Similarly, we see that $(k,\{s,t\})=(i,\{x,y\})$ . Without loss of generality, we may assume $a_j^{(p)}=b_j^{(p)}=z$ and $c_j^{(p)}=d_j^{(p)}=w$ . Therefore, $z\lt w$ and $a\lt c\lt b\lt d$ .

Now, it follows that $\delta _j(a,d) = +1$ and that $\delta _j(c,b)=-1$ , contradicting our assumption that $\psi _p(a,d)=\psi _p(c,b)$ .

3. Modified Dot Product colouring

Fix an odd prime power $q$ and a positive integer $d$ . In this section, we prove Theorem 1.2 by giving an edge-colouring $\varphi _d$ for the complete graph on $n=(q-1)^d$ vertices that uses $(3d+1)q-1$ colours and contains no leftover 6-cliques when $d=3$ and no leftover 8-cliques when $d=4$ .

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 [Reference Babai and Frankl1] 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 ${\mathbb{F}}_q^*$ denote the nonzero elements of the finite field with $q$ elements, and let $({\mathbb{F}}_q^*)^d$ denote the set of ordered $d$ -tuples of elements from ${\mathbb{F}}_q^*$ . In other words, $({\mathbb{F}}_q^*)^d$ is the set of $d$ -dimensional vectors over the field ${\mathbb{F}}_q$ without zero components. In what follows, we will assume that the set ${\mathbb{F}}_q^*$ is endowed with a linear order which can be arbitrarily chosen. We then order the set $({\mathbb{F}}_q^*)^d$ with lexicographic ordering based on the order applied to ${\mathbb{F}}_q^*$ .

Define a set of colours $C_d$ as the disjoint union

\begin{align*} C_d = \text{DOT} \sqcup \text{ZERO} \sqcup \text{UP} \sqcup \text{DOWN}, \end{align*}

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

\begin{align*} \varphi _d\;:\;\binom{({\mathbb{F}}_q^*)^d}{2} \rightarrow C_d \end{align*}

be a colouring function of pairs of distinct vectors, $x\lt y$ , defined by

\begin{align*} \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 . \end{align*}

where $i$ is the first coordinate for which $x=(x_1,\ldots,x_d)$ differs from $y=(y_1,\ldots,y_d)$ and $x \cdot y$ denotes the standard inner product (dot product).

3.2 Number of colours

Let $n$ be a positive integer. Let $q$ be the smallest odd prime power for which $n \leq (q-1)^d$ . Then we can colour the edges of $K_n$ by arbitrarily associating each vertex with a unique vector from $({\mathbb{F}}_q^*)^d$ and taking the colouring induced by $\varphi _d$ . By Bertrand’s Postulate, $q \leq 2(n^{1/d}+1)$ . Therefore, the number of colours used by $\varphi _d$ on $K_n$ is at most

\begin{align*} (3d+1)q-1 \leq (6d+2)n^{1/d}+(6d+1). \end{align*}

3.3 Proof of Theorem 1.2

Definition 3.1. Given a subset of vectors $S \subseteq{\mathbb{F}}^d$ , let $\text{rk}(S)$ denote the rank of the subset, the dimension of the linear subspace spanned by the vectors of $S$ . Let ${\textrm{af}}(S)$ denote the affine dimension of $S$ , the dimension of the affine subspace (also known as the affine hull) spanned by $S$ .

Definition 3.2. A colour $\alpha \in C_d$ has the dot property if $\alpha \in \text{DOT} \cup \text{ZERO}$ . Note that if $\alpha$ has the dot property, then $\varphi _d\left (a,b\right ) = \varphi _d\left (e,f\right )=\alpha$ implies that $a \cdot b = e \cdot f$ for any $a,b,e,f\in ({\mathbb{F}}_q^*)^d$ .

Lemma 3.1. Let $\{s_1,\ldots,s_t\} \subseteq ({\mathbb{F}}_q^*)^d$ be a set of linearly independent vectors and let $a,b \in ({\mathbb{F}}_q^*)^d$ such that

\begin{align*} \varphi _d\left (a,b\right )=\varphi _d\left (a,s_i\right )&=\alpha \\ \varphi _d\left (b,s_i\right )&=\beta \end{align*}

for some $\alpha,\beta \in C_d$ and for each $1\leq i \leq t$ . Then $s_1,\ldots, s_{t},b$ are linearly independent.

Proof. Assume towards a contradiction that $b=\sum _{j=1}^{t} \lambda _j s_j$ for some scalars $\lambda _1,\ldots,\lambda _{t} \in{\mathbb{F}}_q$ . We will first show that $\sum _{j=1}^{t}\lambda _j=1$ .

If $\alpha \in$ DOT, then $b=\sum _{j=1}^{t}\lambda _j s_j$ implies that

\begin{align*} \alpha =a \cdot b = \sum _{j=1}^{t}\lambda _j(a\cdot s_j)=\sum _{j=1}^{t}\lambda _j\alpha . \end{align*}

Therefore, $\sum _{j=1}^{t}\lambda _j=1$ since $\alpha \notin$ ZERO.

If $\alpha \notin$ DOT, then

\begin{align*} a_i+b_i=a_i+s_{1,i}=\cdots =a_i+s_{t,i} \end{align*}

where $i$ is the first index of difference between $a$ and $b$ . Thus, $s_{j,i}=b_i$ for all $1\leq j\leq t$ . But then $b=\sum _{j=1}^{t}\lambda _j s_j$ implies that

\begin{align*} b_i=\sum _{j=1}^{t}\lambda _j s_{j,i}=\sum _{j=1}^{t}\lambda _j b_i. \end{align*}

Hence, $\sum _{j=1}^{t}\lambda _j=1$ since $b_i \neq 0$ . Therefore, for any $\alpha \in C_d$ we have $\sum _{j=1}^{t} \lambda _j=1$ .

Now, if $\beta$ has the dot property, then let $\beta '$ denote $b \cdot s_j$ for all $j=1,\ldots,t$ . We have

\begin{align*} b \cdot b =\sum _{j=1}^{t}\lambda _j(b\cdot s_j)=\sum _{j=1}^{t}\lambda _j\beta '=\beta '. \end{align*}

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

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

\begin{align*} b_k+s_{1,k}=\cdots =b_k+s_{t,k} \end{align*}

where $k$ is the first index of difference between $b$ and $s_1$ . Therefore, $s_{1,k}=\cdots =s_{t,k}$ , and so

\begin{align*} b_k = \sum _{j=1}^t \lambda _j s_{j,k} = \sum _{j=1}^t \lambda _j s_{1,k} = s_{1,k}, \end{align*}

contradicting our choice of $k$ .

Since we reach a contradiction for all colours $\beta$ , it must be the case that $s_1,\ldots, s_{t},b$ are linearly independent vectors, as desired.

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

Definition 3.3. We call the set of vectors $S=\{s_1,\ldots,s_t\} \subseteq ({\mathbb{F}}_q^*)^d$ a $t$ -falling star under the colouring $\varphi _d$ if $\varphi _d\left (s_i,s_j\right )=\alpha _i$ for all $1\leq j \lt i\leq t$ , as shown in Figure 2. For any set of vectors $T \subseteq ({\mathbb{F}}_q^*)^d$ under $\varphi _d$ , let $FS(T)$ denote the maximum $t$ such that $T$ contains a $t$ -falling star.

Figure 2. A $t$ -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 $S=\{s_1,\ldots,s_t\} \subseteq ({\mathbb{F}}_q^*)^d$ be a $t$ -falling star under $\varphi _d$ . Then the vectors $s_1, \ldots, s_{t-1}$ are linearly independent. Consequently, for any subset $T \subseteq ({\mathbb{F}}_q^*)^d$ ,

\begin{align*}{\text{rk}}(T) \geq FS(T)-1. \end{align*}

Moreover, if $T$ is contained within a monochromatic neighbourhood of some other vector, then

\begin{align*}{\text{rk}}(T) \geq FS(T). \end{align*}

Definition 3.4. Let $A,B \subseteq{\mathbb{F}}_q^d$ be disjoint sets of vectors. We say that $A$ confines $B$ if for each $a \in A$ , $a \cdot x = a \cdot y$ for every $x,y \in B$ .

Lemma 3.3. Let $A,B \subseteq{\mathbb{F}}_q^d$ be disjoint sets of vectors such that $A$ confines $B$ . Then

\begin{align*}{\text{af}}(B) \leq d-{\text{rk}}(A). \end{align*}

Proof. Let $t={\textrm{rk}}(A)$ , and let $a_1,\ldots, a_t$ be linearly independent vectors from $A$ . Since $A$ confines $B$ , then for each $a_i$ , there exists an $\alpha _i \in{\mathbb{F}}_q$ such that $a_i \cdot b = \alpha _i$ for all $b \in B$ . Therefore, $B$ is a subset of the solution space for the matrix equation,

\begin{align*} \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 ). \end{align*}

Since $a_1,\ldots, a_t$ are linearly independent, the matrix of these $t$ vectors has full rank, and hence, the solution set is an affine space of dimension $d-t$ , as desired.

Lemma 3.4. Let $A,B \subseteq ({\mathbb{F}}_q^*)^d$ be disjoint sets of vectors and $\alpha \in C_d$ such that $\varphi _d(a,b)=\alpha$ for all $a \in A$ and $b \in B$ . Then either $A$ confines $B$ or $B$ confines $A$ (or both).

Proof. If $\alpha$ has the dot property, then it is trivial that $A$ and $B$ confine one another. So assume that $\alpha \in \text{UP} \cup \text{DOWN}$ . It follows that the first position of difference $i$ is the same between any $a \in A$ and any $b \in B$ . Moreover, every vector of $A$ has the same $i^{th}$ component, every vector of $B$ has the same $i^{th}$ component, and every vector of $A \cup B$ has the same $j^{th}$ component for each $1 \leq j \lt i$ if $i\gt 1$ . Since the vectors are ordered lexicographically based on an underlying linear order of ${\mathbb{F}}_q^*$ , it follows that either $a\lt b$ for all $a \in A$ and $b \in B$ , or $b\lt a$ for all $a \in A$ and $b \in B$ .

Without loss of generality, assume that $a\lt b$ for all $a \in A$ and $b \in B$ . If $\alpha \in \text{UP}$ , then for any particular $a \in A$ , $a \cdot b = a \cdot a$ for every $b \in B$ . Therefore, $A$ confines $B$ . Similarly, if $\alpha \in \text{DOWN}$ , then for any particular $b \in B$ , $b \cdot a = b \cdot b$ for every $a \in A$ , so $B$ confines $A$ .

Lemma 3.5. Let $t \geq 2$ be an integer. An affine subspace of ${\mathbb{F}}_q^d$ of dimension $t-2$ will contain no $t$ -falling stars of $({\mathbb{F}}_q^*)^d$ under $\varphi _d$ . Therefore,

\begin{align*}{\text{af}}(S) \geq FS(S)-1 \end{align*}

for any subset of vectors $S \subseteq ({\mathbb{F}}_q^*)^d$ .

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

So assume that $t \geq 3$ and that the statement is true for $t-1$ . Let $s_1,\ldots,s_t$ be $t$ distinct vectors that form a $t$ -falling star. That is, let $\alpha _1,\ldots,\alpha _{t-1} \in C_d$ and let $\varphi _d\left (s_i,s_j\right ) = \alpha _i$ when $1 \leq i \lt j \leq t$ . Assume towards a contradiction that these vectors are contained inside an affine subspace of dimension $t-2$ . Then there exist scalars $\lambda _1,\ldots,\lambda _{t-1} \in{\mathbb{F}}_q$ such that $s_t=\sum _{j=1}^{t-1} \lambda _j s_j$ and $\sum _{j=1}^{t-1} \lambda _j=1$ .

First, note that if $\lambda _1=0$ , then the vectors $s_2,\ldots,s_t$ form a $(t-1)$ -falling star and are contained in an affine subspace of dimension $t-3$ , a contradiction of the inductive hypothesis. So we must assume in what follows that $\lambda _1 \neq 0$ .

Now, we consider two cases: either $\alpha _1 \in \text{DOT}$ or $\alpha _1 \not \in \text{DOT}$ . If $\alpha _1 \in \text{DOT}$ , then

\begin{align*} \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). \end{align*}

Therefore, $\lambda _1(s_1 \cdot s_1 - \alpha _1)=0$ . Since $\lambda _1 \neq 0$ , it follows that

\begin{align*} s_1 \cdot s_1 = \alpha _1 = s_1 \cdot s_2, \end{align*}

which that implies $\alpha _1 \not \in \text{DOT}$ , a contradiction.

So assume that $\alpha _1 \not \in \text{DOT}$ , and let $i$ denote the index of the first component where $s_1$ differs from the other vectors. In this case,

\begin{align*} s_{1,i}+s_{2,i}=\cdots =s_{1,i}+s_{t,i}, \end{align*}

and hence $s_{2,i}=\cdots =s_{t,i}$ . Therefore,

\begin{align*} 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). \end{align*}

So $\lambda _1(s_{1,i}-s_{t,i})=0$ . Since $\lambda _1 \neq 0$ , we have $s_{1,i}=s_{t,i}$ , a contradiction of our choice of $i$ .

Lemma 3.6. Let $S \subseteq ({\mathbb{F}}_q^*)^d$ be a set of $p \geq 1$ vectors with a leftover structure under the colouring $\varphi _d$ . Then

\begin{align*} FS(S) \geq \left \lceil \log _2{p} \right \rceil + 1. \end{align*}

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

\begin{align*} FS(S) \geq 1+ \max \left (FS(A),FS(B)\right ). \end{align*}

Since $|A|,|B|\lt p$ , then by induction $FS(T) \geq \left \lceil \log _2(|T|) \right \rceil + 1$ for $T=A,B$ . Thus, we have

\begin{align*} FS(S) \geq \left \lceil \log _2\left (\max (|A|,|B|)\right ) \right \rceil + 2, \end{align*}

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

\begin{align*} 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. \end{align*}

Lemma 3.7. Let $p\geq 2$ and $T\geq 0$ be integers. Let $S \subseteq ({\mathbb{F}}_q^*)^d$ be a subset of $p$ vectors with a leftover structure under $\varphi _d$ . If $T \geq 1$ , let $a_1,\ldots,a_T \in ({\mathbb{F}}_q^*)^d$ and $\alpha _1,\ldots,\alpha _T \in C_d$ such that $\varphi _d\left (a_i,a_j\right )=\alpha _i$ for all $1\leq i\lt j\leq T$ and $\varphi _d\left (a_i,s\right )=\alpha _i$ for all $1\leq i \leq T$ and all $s \in S$ .

Then there exists a sequence of positive integers, $x_1,\ldots,x_t$ such that $\sum _{i=1}^t x_i=p-1$ and for each $i=1,\ldots,t$ , the following three conditions hold:

  1. (1) $1 \leq x_i \leq \left \lfloor \frac{p-s_i}{2} \right \rfloor$ ;

  2. (2) $\left \lceil \log _2(x_i) \right \rceil + \left \lceil \log _2(p-s_i-x_i) \right \rceil \leq d-1$ ;

  3. (3) $\left \lceil \log _2(p-s_i-x_i) \right \rceil \leq d-i-T$ ,

where $s_i=0$ if $i=1$ and $s_i=\sum _{j=1}^{i-1}x_j$ otherwise.

Proof. We will prove this by induction on $p$ . For the base case, let $p=2$ . Let $x_1=1$ be the entire sequence. Then the first two conditions hold trivially since the sum of the sequence is 1, and since

\begin{align*} \left \lceil \log _2(1) \right \rceil + \left \lceil \log _2(1) \right \rceil =0 \leq d-1 \end{align*}

for any $d \geq 1$ . For the third condition, since $\left \lceil \log _2(1) \right \rceil = 0$ , it suffices to show that $T+1 \leq d$ . This follows from Corollary 3.2, since $S \cup \{a_1,\ldots,a_T\}$ forms a $(T+2)$ -falling star, and hence $d \geq{\textrm{rk}}\left (S \cup \{a_1,\ldots,a_T\}\right ) \geq T+1$ .

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

\begin{align*} \left \lceil \log _2(|A|) \right \rceil + \left \lceil \log _2(|B|) \right \rceil \leq d-1. \end{align*}

Therefore, setting $x_1=\min \{|A|,|B|\}$ guarantees that $1 \leq x_1 \leq \left \lfloor \frac{p}{2} \right \rfloor$ and that

\begin{align*} \left \lceil \log _2(x_1) \right \rceil + \left \lceil \log _2(p-x_1) \right \rceil \leq d-1. \end{align*}

This gives us a positive integer $x_1$ which satisfies the first two conditions. Moreover, by Corollary 3.2 and Lemma 3.6,

\begin{align*} d \geq{\textrm{rk}}\left (S \cup \{a_1,\ldots,a_T\}\right ) &\geq FS(S\cup \{a_1,\ldots, a_T\})-1\\ & \geq (T+1+\max\!\left (FS(A),FS(B)\right ))-1\\ &\geq T + \left \lceil \log _2(p-x_1) \right \rceil +1. \end{align*}

Thus, $x_1$ also satisfies the third condition.

Let $S'$ denote the larger of the two parts $A$ and $B$ , and let $a_{T+1}$ denote an arbitrary vector from $S \setminus S'$ . Then $S'$ contains $p-x_1\lt p$ vectors and has a leftover structure under $\varphi _d$ . Moreover, $S'$ and $a_1,\ldots,a_{T},a_{T+1}$ satisfy the monochromatic neighbourhood conditions of the hypothesis. Hence, by induction there exists a sequence of positive integers $x_1',\ldots,x'_{\!\!t^{\prime}}$ such that $\sum _{i=1}^{t'}x'_{\!\!i}=p-x_1-1$ and for each $i=1,\ldots,t'$ , the following three conditions hold:

  1. (1) $1 \leq x'_{\!\!i} \leq \left \lfloor \frac{p-x_1-s'_{\!\!i}}{2} \right \rfloor$ ;

  2. (2) $\left \lceil \log _2(x'_{\!\!i}) \right \rceil + \left \lceil \log _2(p-x_1-s'_{\!\!i}-x'_{\!\!i}) \right \rceil \leq d-1$ ;

  3. (3) $\left \lceil \log _2(p-x_1-s'_{\!\!i}-x'_{\!\!i}) \right \rceil \leq d-i-(T+1)$ ,

where $s'_{\!\!i}=0$ if $i=1$ and $s'_{\!\!i}=\sum _{j=1}^{i-1}x'_{\!\!j}$ otherwise.

Let $x_i=x'_{i-1}$ for $2\leq i \leq t'+1$ and let $t=t'+1$ to get a sequence $x_1,\ldots,x_t$ for which

\begin{align*} \sum _{i=1}^t x_i=x_1+\sum _{i=1}^{t'}x_i'=x_1+p-x_1-1=p-1. \end{align*}

For each $i=2,\ldots,t$ , the first two conditions are satisfied since $x_1+s'_{\!\!i}=s_{i+1}$ , and the third condition is satisfied since $d-i-(T+1)=d-(i+1)-T$ .

Corollary 3.8. Let $S \subseteq ({\mathbb{F}}_q^*)^3$ be a set of 6 vectors. Then $S$ cannot have a leftover structure under the colouring $\varphi _3$ .

Proof. If such a set exists, then by Lemma 3.7 with $T=0$ , a positive integer $x_1$ exists such that $1 \leq x_1 \leq 3$ and

\begin{align*} \left \lceil \log _2(x_1) \right \rceil + \left \lceil \log _2(6-x_1) \right \rceil \leq 2. \end{align*}

It is simple to check that no such integer exists.

Corollary 3.9. Let $S \subseteq ({\mathbb{F}}_q^*)^4$ be a set of 8 vectors. Then $S$ cannot have a leftover structure under the colouring $\varphi _4$ .

Proof. If such a set exists, then by Lemma 3.7 with $T=0$ , we must be able to find a sequence of positive integers $x_1,x_2,\ldots,x_t$ that satisfy the conditions given in the Lemma. In particular, $1 \leq x_1 \leq 4$ and

\begin{align*} \left \lceil \log _2(x_1) \right \rceil + \left \lceil \log _2(8-x_1) \right \rceil \leq 3. \end{align*}

We can check and find that $x_1=1$ is the only possibility. Therefore, $1 \leq x_2 \leq 3$ such that

\begin{align*} \left \lceil \log _2(7-x_2) \right \rceil &\leq 2\\ \left \lceil \log _2(x_2) \right \rceil + \left \lceil \log _2(7-x_2) \right \rceil &\leq 3.\\ \end{align*}

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 $p$ -cliques can appear under $\varphi _d$ for a particular $d$ . For example, this proof implies that the only leftover $5$ -clique that can appear under $\varphi _3$ is a monochromatic $C_4$ contained inside a monochromatic neighbourhood of one vertex (that is, an initial $(1,4)$ -bipartition with a $(2,2)$ -bipartition inside the part with four vertices). In [Reference Cameron and Heath2], we handled this specific leftover structure by splitting each colour class of $\varphi _3$ into four new colours determined by certain relations between vectors. While the current paper can be viewed as our attempt to fully generalize the colouring techniques used in [Reference Cameron and Heath2] and [Reference Mubayi8], it does not generalize the splitting that was crucial for handling the final leftover $5$ -clique. Perhaps such a generalized splitting would be enough to give $f(n,p,p) \leq n^{1/(p-2) +o(1)}$ for $p \geq 6$ or at least improve the best-known upper bounds for values of $p$ other than the two addressed in this paper.

Acknowledgements

This research was partially supported by NSF RTG Grant DMS-1937241 and NSF RTG Grant DMS-1839918.

References

Babai, L. and Frankl, P. (1992) Linear Algebra Methods in Combinatorics: With Applications to Geometry and Computer Science. Department of Computer Science, University of Chicago.Google Scholar
Cameron, A. and Heath, E. (2018) A $(5,5)$ -colouring of ${K}_n$ with few colours. Comb. Probab. Comput. 27(6) 892912.CrossRefGoogle Scholar
Conlon, D., Fox, J., Lee, C. and Sudakov, B. (2015) The Erdős–Gyárfás problem on generalized Ramsey numbers. Proc. Lond. Math. Soc. 110(1) 118.CrossRefGoogle Scholar
Eichhorn, D. and Mubayi, D. (2000) Edge-coloring cliques with many colors on subcliques. Combinatorica 20(3) 441444.CrossRefGoogle Scholar
Erdős, P. (1975) Problems and results on finite and infinite graphs. Recent Advances in Graph Theory (Proc. Second Czechoslovak Symposium. Prague, 1974) 183192.Google Scholar
Erdős, P. and Gyárfás, A. (1997) A variant of the classical Ramsey problem. Combinatorica 17(4) 459467.CrossRefGoogle Scholar
Mubayi, D. (1998) Edge-coloring cliques with three colors on all 4-cliques. Combinatorica 18(2) 293296.CrossRefGoogle Scholar
Mubayi, D. (2004) An explicit construction for a Ramsey problem. Combinatorica 24(2) 313324.CrossRefGoogle Scholar
Figure 0

Figure 1. An example of a $p$-clique with a leftover structure.

Figure 1

Figure 2. A $t$-falling star.