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

\stackMath

Van Lint–MacWilliams’ conjecture and maximum cliques in Cayley graphs over finite fields

Shamil Asgarli Department of Mathematics
University of British Columbia
1984 Mathematics Road
Canada V6T 1Z2
[email protected]
 and  Chi Hoi Yip Department of Mathematics
University of British Columbia
1984 Mathematics Road
Canada V6T 1Z2
[email protected]
Abstract.

A well-known conjecture due to van Lint and MacWilliams states that if AA is a subset of 𝔽q2\mathbb{F}_{q^{2}} such that 0,1A0,1\in A, |A|=q|A|=q, and aba-b is a square for each a,bAa,b\in A, then AA must be the subfield 𝔽q\mathbb{F}_{q}. This conjecture is often phrased in terms of the maximum cliques in Paley graphs. It was first proved by Blokhuis and later extended by Sziklai to generalized Paley graphs. In this paper, we give a new proof of the conjecture and its variants, and show how this Erdős-Ko-Rado property of Paley graphs extends to a larger family of Cayley graphs, which we call Peisert-type graphs. These results resolve the conjectures by Mullin and Yip.

Key words and phrases:
Cayley graph, Paley graph, Peisert graph, maximum clique, Erdős-Ko-Rado theorem
2020 Mathematics Subject Classification:
Primary 05C25, 11T30; Secondary 05C69, 11T24, 51E15

1. Introduction

Throughout the paper, pp denotes an odd prime, qq denotes a positive power of pp, and 𝔽q\mathbb{F}_{q} denotes the finite field with qq elements. For a finite field 𝔽q\mathbb{F}_{q}, we denote 𝔽q=𝔽q{0}\mathbb{F}_{q}^{*}=\mathbb{F}_{q}\setminus\{0\}.

In 1978, van Lint and MacWilliams [vLM78] studied the vectors of minimum weight in certain generalized quadratic residue codes, and they proposed the following conjecture: if AA is a subset of 𝔽q2\mathbb{F}_{q^{2}} such that 0,1A0,1\in A, |A|=q|A|=q, and aba-b is a square for each a,bAa,b\in A, then AA is the subfield 𝔽q\mathbb{F}_{q}. Although this conjecture is purely a statement about the structure of the subfield, it is often phrased in terms of the maximum cliques in Paley graphs of square order, and sometimes in terms of the analog of Erdős-Ko-Rado theorem for Paley graphs [GM]*Chapter 5. The conjecture was first completely proved by Blokhuis [Blo84]; see Section 2 for a brief history of the conjecture and related results by other authors.

In this paper, we give a new proof of this conjecture (for pp sufficiently large with respect to logpq\log_{p}q), explore its variant for other Cayley graphs defined on finite fields, including Peisert graphs, and confirm two related conjectures by Mullin and the second author. In addition, we exhibit several new results on maximal cliques in both generalized Paley and Peisert graphs.

We recall a few relevant terms from graph theory. Given an abelian group GG and a set SG{0}S\subset G\setminus\{0\} with S=SS=-S, the Cayley graph Cay(G,S)\operatorname{Cay}(G,S) is the graph whose vertices are elements of GG, such that two vertices gg and hh are adjacent if and only if ghSg-h\in S. Such a set SS is called a connection set, and the condition S=SS=-S guarantees that the graph is undirected. In this paper, we introduce a new family of Cayley graphs.

Definition 1.1 (Peisert-type graphs).

Let qq be an odd prime power. Let S𝔽q2S\subset\mathbb{F}_{q^{2}}^{*} be a union of at most q+12\frac{q+1}{2} cosets of 𝔽q\mathbb{F}_{q}^{*} in 𝔽q2\mathbb{F}_{q^{2}}^{*} such that 𝔽qS\mathbb{F}_{q}^{*}\subset S, that is,

S=c1𝔽qc2𝔽qcm𝔽q,S=c_{1}\mathbb{F}_{q}^{*}\cup c_{2}\mathbb{F}_{q}^{*}\cup\cdots\cup c_{m}\mathbb{F}_{q}^{*}, (1)

where c1=1c_{1}=1 and mq+12m\leq\frac{q+1}{2}. Then the Cayley graph X=Cay(𝔽q2+,S)X=\operatorname{Cay}(\mathbb{F}_{q^{2}}^{+},S) is said to be a Peisert-type graph.

Such graphs have been implicitly studied within a larger family of strongly regular Cayley graphs in [BWX99]. In Lemma 2.13, we will see that Paley graphs, Peisert graphs, and more general versions of such graphs, are all special instances of Peisert-type graphs. The precise definitions of these graphs will be given in Section 2.

In order to state our first main result, we review a few additional definitions. A clique in a graph XX is a subset of vertices in XX such that every two distinct vertices in the clique are adjacent. A maximum clique is a clique with the maximum size, while a maximal clique is a clique that is not contained in a strictly larger clique. The clique number of XX, denoted by ω(X)\omega(X), is the size of a maximum clique in a graph XX. We prove the following theorem on the structure of maximum cliques in Peisert-type graphs.

Theorem 1.2.

Let XX be a Peisert-type graph of order q2q^{2}, where qq is a power of an odd prime pp. Then ω(X)=q\omega(X)=q, and any maximum clique in XX containing 0 is an 𝔽p\mathbb{F}_{p}-subspace of 𝔽q2\mathbb{F}_{q^{2}}.

Our second main result strengthens the conclusion of Theorem 1.2 under additional assumptions.

Theorem 1.3.

Let n2n\geq 2 be an integer and ε>0\varepsilon>0 a real number. Let X=Cay(𝔽q2+,S)X=\operatorname{Cay}(\mathbb{F}_{q^{2}}^{+},S) be a Peisert-type graph, where q=pnq=p^{n} and p>4.1n2/ϵ2p>4.1n^{2}/\epsilon^{2}. Suppose that there is a nontrivial multiplicative character χ\chi of 𝔽q2\mathbb{F}_{q^{2}}, such that the set {χ(x):xS}\{\chi(x):x\in S\} is ε\varepsilon-lower bounded. Then in the Cayley graph XX, the only maximum clique containing 0,10,1 is the subfield 𝔽q\mathbb{F}_{q}.

Roughly speaking, a set AA is ε\varepsilon-lower bounded if the sum over any subset of AA is not too small. A precise definition of ε\varepsilon-lower bounded appears in Definition 2.16 in Section 2. The sharpness of Theorem 1.3 is discussed in Example 2.21. In Section 2.4, as a straightforward application of Theorem 1.3, we will prove the van Lint–MacWilliams’ conjecture for all sufficiently large pp.

We mention here one new application of Theorem 1.3. Mullin [NM]*Chapter 8 conjectured that every maximum clique in the Peisert graph of square order containing 0,10,1 must be the subfield 𝔽q\mathbb{F}_{q}. Note that this is an exact analogue of the van Lint–MacWilliams’ conjecture for Peisert graphs. We prove Mullin’s conjecture when pp is sufficiently large.

Theorem 1.4.

Let q=pnq=p^{n} be a prime power such that q3(mod4)q\equiv 3\pmod{4}. Assume that p>8.2n2p>8.2n^{2}. Then the only maximum clique containing 0,10,1 in the Peisert graph of order q2q^{2} is given by the subfield 𝔽q\mathbb{F}_{q}.

Another objective of the paper is to prove new results on the structure of maximal cliques in Paley and Peisert graphs. While these results are not direct applications of our main Theorem 1.3, their proofs share similar ideas involving character sum estimates and the subspace structure. In particular, we confirm the conjecture of the second author on the maximal cliques in Peisert graphs of order q4q^{4}.

Theorem 1.5 ([Yip3]*Conjecture 4.4).

If qq is a power of a prime p3(mod4)p\equiv 3\pmod{4} and q>3q>3, then 𝔽q\mathbb{F}_{q} is a maximal clique in the Peisert graph of order q4q^{4}.

The paper is organized as follows. In Section 2, we provide additional background, and further motivate the new definitions appearing in our paper. We also review the history of the van Lint–MacWilliams’ conjecture, and put our main theorems in context. In Section 3, we explain the two main techniques used in our proof. Specifically, we borrow results regarding the number of directions determined by a point set, and estimates on character sums over finite fields. In Section 4, we give proofs of Theorem 1.2 and Theorem 1.3; afterwards, we extend Theorem 1.4 to generalized Peisert graphs. In Section 5, we prove Theorem 1.5 and several other results on maximal cliques in certain Cayley graphs.

2. Background and overview of the paper

This section is organized into four subsections. In Section 2.1, we give additional context for the conjecture due to van Lint and MacWilliams. In Section 2.2, we recall the definitions of Peisert and generalized Peisert graphs, mention Mullin’s conjecture, and describe what is currently known about the clique number of Peisert graphs. In Section 2.3, we give further motivation for defining Peisert-type graphs. In Section 2.4, we explain our main results in more detail, and deduce Theorem 1.4 as a quick application.

2.1. Van Lint–MacWilliams’ conjecture and its variants

Definition 2.1.

Let qq be a prime power satisfying q1(mod4)q\equiv 1\pmod{4}. The Paley graph on 𝔽q\mathbb{F}_{q}, denoted by PqP_{q}, is the Cayley graph Cay(𝔽q+,(𝔽q)2)\operatorname{Cay}(\mathbb{F}_{q}^{+},(\mathbb{F}_{q}^{*})^{2}).

In other words, PqP_{q} is the graph whose vertices are the elements of 𝔽q\mathbb{F}_{q}, such that two vertices are adjacent if and only if the difference between two vertices is a square in 𝔽q\mathbb{F}_{q}^{\ast}.

Van Lint–MacWilliams’ conjecture is often phrased as follows: in the Paley graph with order q2q^{2}, the only maximum clique containing 0,10,1 is the subfield 𝔽q\mathbb{F}_{q}. Van Lint and MacWilliams [vLM78] proved their conjecture in the special case q=pq=p. A shorter proof for the case q=pq=p was obtained by Lovász and Schrijver [LS83] using Rédei’s result on the number of directions; see Theorem 3.2 for a related discussion. The conjecture was first proved completely by Blokhuis [Blo84].

Theorem 2.2 ([Blo84]).

If qq is an odd prime power, then in the Paley graph of order q2q^{2}, the only maximum clique containing 0,10,1 is the subfield 𝔽q\mathbb{F}_{q}.

Later, Bruen and Fisher [BF91] gave a different proof using the so-called “Jamison method”. Sziklai [Szi99] generalized Blokhuis’s proof and extended the conjecture to certain generalized Paley graphs; see Theorem 2.4. See Remark 2.19 for discussion on the proofs given by Blokhuis and Sziklai, and how they differ from the proof in the present paper.

One can define generalized Paley graphs in an analogous way. They were first introduced by Cohen [SC] in 1988, and reintroduced by Lim and Praeger [LP] in 2009.

Definition 2.3.

Let d>1d>1 be a positive integer. If q1(mod2d)q\equiv 1\pmod{2d}, the dd-Paley graph on 𝔽q\mathbb{F}_{q}, denoted by GP(q,d)GP(q,d), is the Cayley graph Cay(𝔽q+,(𝔽q)d)\operatorname{Cay}(\mathbb{F}_{q}^{+},(\mathbb{F}_{q}^{*})^{d}), where (𝔽q)d(\mathbb{F}_{q}^{*})^{d} is the set of dd-th powers in 𝔽q\mathbb{F}_{q}^{*}.

In the literature, 33-Paley graphs and 44-Paley graphs are also known as cubic Paley graphs and quadruple Paley graphs, respectively.

If d(q+1)d\mid(q+1), the following theorem of Sziklai characterizes all maximum cliques in GP(q2,d)GP(q^{2},d).

Theorem 2.4 ([Szi99]*Theorem 1.2).

If qq is an odd prime power and dd is a divisor of (q+1)(q+1) such that d>1d>1, then in the generalized Paley graph GP(q2,d)GP(q^{2},d), the only maximum clique containing 0,10,1 is the subfield 𝔽q\mathbb{F}_{q}; in particular, each maximum clique in GP(q2,d)GP(q^{2},d) is an affine transformation of the subfield 𝔽q\mathbb{F}_{q}.

Remark 2.5.

The assumption that d(q+1)d\mid(q+1) in Theorem 2.4 is crucial. In the case d(q+1)d\nmid(q+1), it is easy to verify that 𝔽q\mathbb{F}_{q} does not form a clique in GP(q2,d)GP(q^{2},d) (see for example [Yip4]*Lemma 1.5); moreover, the second author [Yip4] recently proved that ω(GP(q2,d))q1\omega(GP(q^{2},d))\leq q-1. Therefore, it is unlikely that there is a nice characterization of maximum cliques in GP(q2,d)GP(q^{2},d), where d(q+1)d\nmid(q+1).

Sziklai [Szi99] also proved the following embeddability result on maximum cliques.

Theorem 2.6 ([Szi99]*Theorem 1.3).

Let qq be an odd prime power and dd a divisor of (q+1)(q+1) such that d3d\geq 3. If CC is a clique in the generalized Paley graph GP(q2,d)GP(q^{2},d), such that 0,1C0,1\in C, and |C|>q(11/d)q|C|>q-(1-1/d)\sqrt{q}, then C𝔽qC\subset\mathbb{F}_{q}.

Many authors have attempted to give a new proof of van Lint–MacWilliams’ conjecture. Godsil and Meagher [GM] suggested an approach using algebraic graph theory, in particular the module method, to prove Theorem 2.2. The module method is a powerful tool to prove Erdős-Ko-Rado type theorems; for more details, see the discussion in [GM]*Section 5.9. Recently, Goryainov, Lin, and the authors [AGLY22] have proved that all Peisert-type graphs satisfy the EKR-module property, which is the first step in applying the module method. In addition, the second author [Yip4] gave a Fourier analytic characterization of maximum cliques, which suggests an alternative approach to prove Theorem 2.2.

2.2. Peisert graphs and generalized Peisert graphs

It is well-known that Paley graphs are self-complementary and symmetric. In [WP2], Peisert discovered a new infinite family of self-complementary symmetric graphs, which he called PP^{*}-graphs. This new family of graphs also goes under the name of Peisert graphs in the literature.

Definition 2.7.

The Peisert graph of order q=prq=p^{r}, where pp is a prime such that p3(mod4)p\equiv 3\pmod{4} and rr is even, denoted by PqP^{*}_{q}, is the Cayley graph Cay(𝔽q+,Mq)\operatorname{Cay}(\mathbb{F}_{q}^{+},M_{q}) with

Mq\colonequals{gj:j0,1(mod4)},M_{q}\colonequals\{g^{j}:j\equiv 0,1\pmod{4}\},

where gg is a primitive root of the field 𝔽q\mathbb{F}_{q}.

While the construction of PqP^{*}_{q} depends on the primitive root gg, the isomorphism type of PqP^{*}_{q} is independent of the choice of gg [WP2]. Note that MqM_{q} is not closed under multiplication since gg=g2Mqg\cdot g=g^{2}\notin M_{q}. The condition p3(mod4)p\equiv 3\pmod{4} is needed to ensure that PqP_{q}^{*} is symmetric.

Kisielewicz and Peisert [KP] showed that Paley graphs and Peisert graphs are similar in many aspects. However, we know very little about the structure of cliques of Peisert graphs other than Theorem 2.8.

Theorem 2.8 ([KP]*Theorem 5.1).

Let q=psq=p^{s}, where p3(mod4)p\equiv 3\pmod{4} and s=2ks=2k. If kk is odd, then ω(Pq)=q\omega(P^{*}_{q})=\sqrt{q} and the subfield 𝔽q\mathbb{F}_{\sqrt{q}} forms a clique in PqP^{*}_{q}; if kk is even, then ω(Pq)q1/4\omega(P^{*}_{q})\geq q^{1/4} and the subfield 𝔽q1/4\mathbb{F}_{q^{1/4}} forms a clique in PqP^{*}_{q}.

Motivated by Theorem 2.2 and some numerical evidence, Mullin proposed the following conjecture [NM]*Chapter 8. Note that it is an analog of Theorem 2.2 on the characterization of maximum cliques in a Paley graph with square order.

Conjecture 2.9 (Mullin).

Let q3(mod4)q\equiv 3\pmod{4} be a prime power. Then the only maximum clique containing 0,10,1 in the Peisert graph of order q2q^{2} is given by the subfield 𝔽q\mathbb{F}_{q}.

She remarked that if qq is a prime, then Conjecture 2.9 can be proved using a similar technique as Theorem 2.4; see [NM]*Corollary 3.4. She also remarked that the proof of Theorem 2.4 fails to extend to Peisert graphs straightforwardly, since the connection set of a Peisert graph is not closed under multiplication, while for Paley graphs and generalized Paley graphs it is closed; see also Remark 2.19. We show that Conjecture 2.9 holds for sufficiently large qq; this is Theorem 1.4 and its proof is given in Section 4.

In [Yip4], the second author studied the connection between maximal cliques and maximum cliques in Peisert graphs of order q4q^{4}, and conjectured the following based on numerical evidence.

Conjecture 2.10 ([Yip3]*Conjecture 4.4).

If qq is a power of a prime p3(mod4)p\equiv 3\pmod{4} and q>3q>3, then 𝔽q\mathbb{F}_{q} is a maximal clique in the Peisert graph of order q4q^{4}.

Note that the Peisert graph of order q4q^{4} is the edge-disjoint union of two copies of the quadruple Paley graph GP(q4,4)GP(q^{4},4). Therefore, Conjecture 2.10 strengthens [Yip3]*Theorem 1.6, which states that 𝔽q\mathbb{F}_{q} is a maximal clique in GP(q4,4)GP(q^{4},4). We show that the conjecture above holds; see Theorem 1.5.

Motivated by the similarity shared among Paley graphs, generalized Paley graphs, and Peisert graphs, Mullin introduced generalized Peisert graphs; see [NM]*Section 5.3. In order to ensure that GP(q,d)GP^{*}(q,d) is an undirected graph, we also need a further hypothesis q1(mod2d)q\equiv 1\pmod{2d}; in Mullin’s thesis, this last congruence condition is q1(modd)q\equiv 1\pmod{d}, which appears to be a typo.

Definition 2.11.

Let dd be a positive even integer, and qq a prime power such that q1(mod2d)q\equiv 1\pmod{2d}. The dd-th power Peisert graph of order qq, denoted by GP(q,d)GP^{*}(q,d), is the Cayley graph Cay(𝔽q+,Mq,d)\operatorname{Cay}(\mathbb{F}_{q}^{+},M_{q,d}), where

Mq,d={gdk+j:0jd21,k},M_{q,d}=\bigg{\{}g^{dk+j}:0\leq j\leq\frac{d}{2}-1,k\in\mathbb{Z}\bigg{\}},

and gg is a primitive root of 𝔽q\mathbb{F}_{q}.

Remark 2.12.

Note that GP(q,2)GP^{*}(q,2) is precisely the Paley graph PqP_{q}, and GP(q,4)GP^{*}(q,4) is the Peisert graph PqP^{*}_{q} if qq is an even power of a prime p3(mod4)p\equiv 3\pmod{4}. Moreoever, the generalized Peisert graph GP(q,d)GP^{*}(q,d) is the edge-disjoint union of d2\frac{d}{2} copies of the generalized Paley graph GP(q,d)GP(q,d), which again suggests that it is much harder to characterize the maximum cliques in GP(q,d)GP^{*}(q,d) than doing so for GP(q,d)GP(q,d). Inspired by Conjecture 2.9, one can propose a similar conjecture on maximum cliques in a generalized Peisert graph; we prove such a result in Theorem 2.20.

2.3. Peisert-type graphs

One motivation of this paper is to provide a new and short proof for Theorem 2.4, and explore the structure of maximum cliques of those Cayley graphs which are similar to Paley graphs and generalized Paley graphs. In particular, we consider Cayley graphs defined on 𝔽q2+\mathbb{F}_{q^{2}}^{+} with connection set being a union of cosets of 𝔽q\mathbb{F}_{q}^{*} in 𝔽q2\mathbb{F}_{q^{2}}^{*} as in equation (1).

In our Definition 1.1 of a Peisert-type graph X=Cay(𝔽q2+,S)X=\operatorname{Cay}(\mathbb{F}_{q^{2}}^{+},S) we assume that 𝔽qS\mathbb{F}_{q}\subset S; in particular, these graphs enjoy the property that ω(X)q\omega(X)\geq q. As we will see in the proof of Theorem 1.2, the condition on the number of cosets mq+12m\leq\frac{q+1}{2} is needed. Another reason for assuming mq+12m\leq\frac{q+1}{2} is that we want |S|q212|S|\leq\frac{q^{2}-1}{2}; see Example 2.18.

We remark that our definition of Peisert-type graphs shares some similarities with the definition of Peisert graphs. Recall that a Peisert graph is simply an edge-disjoint union of two copies of the quadruple Paley graph with the same order. Note that Cay(𝔽q2+,𝔽q)=GP(q2,q+1)\operatorname{Cay}(\mathbb{F}_{q^{2}}^{+},\mathbb{F}_{q}^{*})=GP(q^{2},q+1) is a (q+1)(q+1)-Paley graph. Therefore, a Peisert-type graph is simply an edge-disjoint union of copies of (q+1)(q+1)-Paley graph. Notice also that we do not require the connection set SS to be closed under multiplication in Definition 1.1.

The following lemma allows us to unify the notions of generalized Paley graphs GP(q2,d)GP(q^{2},d) and generalized Peisert graphs GP(q2,d)GP^{*}(q^{2},d) under the assumption that d(q+1)d\mid(q+1). Note that when dd is even, GP(q,d)GP(q,d) already embeds inside GP(q,d)GP^{\ast}(q,d); however, for odd dd, there is no such obvious embedding as GP(q,d)GP^{\ast}(q,d) is defined only for even values of dd. The lemma shows that the Peisert-type graph is general enough to encompass both cases.

Lemma 2.13.

The following families of Cayley graphs are Peisert-type graphs:

  • Paley graphs of square order;

  • Peisert graph with order q2q^{2}, where q3(mod4)q\equiv 3\pmod{4};

  • Generalized Paley graphs GP(q2,d)GP(q^{2},d), where d(q+1)d\mid(q+1) and d>1d>1;

  • Generalized Peisert graphs GP(q2,d)GP^{*}(q^{2},d), where d(q+1)d\mid(q+1) and dd is even.

The proof of the preceding lemma will be presented in Section 4. Both (generalized) Paley graphs and (generalized) Peisert graphs have many nice properties, and they have many applications in coding theory and design theory; see for example [JAlex15, JL, KR, WQWX]. We also expect that Peisert-type graphs will have a new set of applications.

2.4. Main results and their implications

Our goal is to characterize maximum cliques in a Peisert-type graph X=Cay(𝔽q2+,S)X=\operatorname{Cay}(\mathbb{F}_{q^{2}}^{+},S). Our first main result, Theorem 1.2, asserts that the clique number of XX is qq, and every maximum clique in XX has vector space structure; see [Yip3] for a related discussion on the vector space structure of maximal cliques in Cayley graphs. Later, we will show that such subspace must be in fact the subfield 𝔽q\mathbb{F}_{q} under additional assumptions.

When the order of the graph is p2p^{2}, Theorem 1.2 immediately implies the following corollary.

Corollary 2.14.

Let XX be a Peisert-type graph of order p2p^{2}. Then the only maximum clique containing 0,10,1 is the subfield 𝔽p\mathbb{F}_{p}.

It is natural to ask whether the conclusion in Theorem 1.2 can be strengthened to the assertion that any maximum clique is an 𝔽q\mathbb{F}_{q}-affine space. In Example 2.21 below, we will see a counterexample for this stronger conclusion.

In the same spirit as Theorem 2.6, we prove an embeddability result for cliques in Peisert-type graphs.

Theorem 2.15.

Let X=Cay(𝔽q2+,S)X=\operatorname{Cay}(\mathbb{F}_{q^{2}}^{+},S) be a Peisert-type graph, where q=pnq=p^{n}, and m=|S|/(q1)<q+12m=|S|/(q-1)<\frac{q+1}{2}. If CC is a clique in XX, such that 0C0\in C and |C|>q(1m/(q+1))q|C|>q-\big{(}1-m/(q+1)\big{)}\sqrt{q}, then CC is contained in a 𝔽p\mathbb{F}_{p}-subspace of 𝔽q2\mathbb{F}_{q^{2}} with dimension nn.

We want to find minimal assumptions on Peisert-type graphs X=Cay(𝔽q2+,S)X=\operatorname{Cay}(\mathbb{F}_{q^{2}}^{+},S) for which the only maximum clique containing 0,10,1 is the subfield 𝔽q\mathbb{F}_{q}. Such a clique is already an 𝔽p\mathbb{F}_{p}-subspace by Theorem 1.2, and to show that it is exactly the subfield 𝔽q\mathbb{F}_{q}, we will use character sum estimates in Section 3.2. The following definition is helpful for our discussion.

Definition 2.16.

Let ε>0\varepsilon>0. A set MM\subset\mathbb{C} is said to be ε\varepsilon-lower bounded if for every integer kk\in\mathbb{N}, and for every choice of x1,x2,,xkMx_{1},x_{2},\ldots,x_{k}\in M, we have

|j=1kxj|εk.\displaystyle\bigg{|}\sum_{j=1}^{k}x_{j}\bigg{|}\geq\varepsilon k.
Example 2.17.

If a,ba,b are non-negative integers, then

|a+bi|=a2+b2(a+b)2/2=a+b2.\displaystyle|a+bi|=\sqrt{a^{2}+b^{2}}\geq\sqrt{(a+b)^{2}/2}=\frac{a+b}{\sqrt{2}}.

Therefore, the set {1,i}\{1,i\} is 12\frac{1}{\sqrt{2}}-lower bounded.

Our Theorem 1.3 has the hypothesis that {χ(x):xS}\{\chi(x):x\in S\} is ε\varepsilon-lower bounded for some ε>0\varepsilon>0. The following example explains the necessity of the condition m(q+1)/2m\leq(q+1)/2, or equivalently, |S|(q21)/2|S|\leq(q^{2}-1)/2 in Definition 1.1.

Example 2.18.

Suppose that S𝔽q2S\subset\mathbb{F}_{q^{2}}^{\ast} with |S|>(q21)/2|S|>(q^{2}-1)/2. If χ\chi is an odd multiplicative character of 𝔽q2\mathbb{F}_{q^{2}}, meaning that χ(1)=1\chi(-1)=-1, then we claim that SS is not ε\varepsilon-lower bounded for any ε>0\varepsilon>0. Indeed, by the pigeonhole principle, {b,b}S\{b,-b\}\in S for some b𝔽q2b\in\mathbb{F}_{q^{2}}^{\ast}, and χ(b)+χ(b)=0\chi(-b)+\chi(b)=0, showing that the condition in Definition 2.16 is not satisfied for k=2k=2.

In Lemma 2.13, we saw that certain generalized Paley graphs and Peisert graphs are Peisert-type graphs. Next we present two quick applications of Theorem 1.3.

It is straightforward to apply Theorem 1.3 to generalized Paley graph GP(q2,d)GP(q^{2},d), where d(q+1)d\mid(q+1). Let χ\chi to be a multiplicative character with order dd; then the set {χ(xd):x𝔽q2}={1}\{\chi(x^{d}):x\in\mathbb{F}_{q^{2}}^{*}\}=\{1\} is trivially 11-lower bounded. This immediately implies a slightly weaker version of Theorem 2.4, that is, we obtain a simple proof that Theorem 2.4 holds if pp is sufficiently large.

Remark 2.19.

The original proofs of Theorem 2.2 by Blokhuis and Theorem 2.4 by Sziklai relied on the algebraic properties of the connection sets being squares and dd-th powers, which are in fact subgroups. Although the implication of our main result to Paley graphs is slightly weaker than their results, it seems difficult to extend their method to a general Peisert-type graph.

Similarly, we can apply Theorem 1.3 to Peisert graph with order q2q^{2}, where q3(mod4)q\equiv 3\pmod{4}, and deduce that Conjecture 2.9 holds as long as pp is sufficiently large.

Proof of Theorem 1.4.

Let gg be a primitive root of 𝔽q2\mathbb{F}_{q^{2}} and let χ\chi be a multiplicative character such that χ(g)=i\chi(g)=i. Then the set {χ(g4k):k}{χ(g4k+1):k}={1,i}\{\chi(g^{4k}):k\in\mathbb{N}\}\cup\{\chi(g^{4k+1}):k\in\mathbb{N}\}=\{1,i\} is 12\frac{1}{\sqrt{2}}-lower bounded, as shown in Example 2.17. The required claim follows from Theorem 1.3. ∎

In Section 4.3, we will apply Theorem 1.3 to study maximum cliques in a generalized Peisert graph GP(q2,d)GP^{*}(q^{2},d), where d(q+1)d\mid(q+1). Note that Theorem 2.20 is a generalization of Theorem 1.4. It also strengthens Sziklai’s result (Theorem 2.4), because GP(q2,d)GP(q^{2},d) is a subgraph of GP(q2,d)GP^{*}(q^{2},d); see Remark 2.12 for more details.

Theorem 2.20.

Let n2n\geq 2 be an integer, and d4d\geq 4 an even integer. Let X=GP(q2,d)X=GP^{*}(q^{2},d) be a generalized Peisert graph, where q=pnq=p^{n}, p>4.1n2d4/π2(d1)2p>4.1n^{2}d^{4}/\pi^{2}(d-1)^{2}, and d(q+1)d\mid(q+1). Then in XX, the only maximum clique containing 0,10,1 is the subfield 𝔽q\mathbb{F}_{q}.

The following example illustrates the necessity of the assumption that pp is sufficiently large stated in Theorem 1.3.

Example 2.21.

We consider the maximum cliques in a generalized Peisert graph GP(q2,q+1)GP^{*}(q^{2},q+1), which was shown to be a conference graph by Mullin [NM]*Lemma 5.3.6 and also a Peisert-type graph in Lemma 2.13. Let gg be a primitive root of 𝔽q2\mathbb{F}_{q^{2}}^{*}; then gq+1g^{q+1} is a primitive root of 𝔽q𝔽q2\mathbb{F}_{q}^{*}\subset\mathbb{F}_{q^{2}}^{*}. Therefore, GP(q2,q+1)=Cay(𝔽q2+,S)GP^{*}(q^{2},q+1)=\operatorname{Cay}(\mathbb{F}_{q^{2}}^{+},S) is a Peisert-type graph, where

S=g0𝔽qg1𝔽qg(q1)/2𝔽q.S=g^{0}\mathbb{F}_{q}^{*}\cup g^{1}\mathbb{F}_{q}^{*}\cup\cdots\cup g^{(q-1)/2}\mathbb{F}_{q}^{*}.

Let CC be a maximum clique containing 0; then C{0}SC\setminus\{0\}\subset S. Let jj be the smallest non-negative integer such that Cgj𝔽qC\cap g^{j}\mathbb{F}_{q}^{*}\neq\emptyset, and say agjCag^{j}\in C, where a𝔽qa\in\mathbb{F}_{q}^{*}. Then (agj)1C{0}S(ag^{j})^{-1}C\setminus\{0\}\subset S; in particular, (agj)1C(ag^{j})^{-1}C is a maximum clique containing 0,10,1.

Suppose 𝔽q\mathbb{F}_{q} was the only maximum clique containing 0,10,1; then the above argument would show that all maximum cliques containing 0 are given by gj𝔽qg^{j}\mathbb{F}_{q}, where 0j(q1)/20\leq j\leq(q-1)/2. This would imply that the number of maximum cliques containing 0 is exactly q+12\frac{q+1}{2}, which violates the computational results listed in [NM]*Section 5.4. For example, GP(34,10)GP^{*}(3^{4},10) has 99 maximum cliques containing 0 which is more than q+12=9+12=5\frac{q+1}{2}=\frac{9+1}{2}=5; see [AGLY22]*Section 5.2 for a detailed study of the maximum cliques in GP(34,10)GP^{*}(3^{4},10). Similarly, GP(54,26)GP^{*}(5^{4},26) has 1919 maximum cliques containing 0, which is more than q+12=25+12=13\frac{q+1}{2}=\frac{25+1}{2}=13. This shows that Theorem 1.3 can only hold for sufficiently large pp.

3. Main tools

In this section, we introduce two classical tools, which are both crucial for our proof.

3.1. Number of directions determined by a point set

Let AG(2,q)AG(2,q) denote the affine Galois plane over the finite field 𝔽q\mathbb{F}_{q}. Let UAG(2,q)U\subset AG(2,q). We use Cartesian coordinates in AG(2,q)AG(2,q) so that U={(xi,yi):1i|U|}U=\{(x_{i},y_{i}):1\leq i\leq|U|\}. The set of directions determined by UAG(2,q)U\subset AG(2,q) is

𝒟U={yjyixjxi:1i<j|U|}𝔽q{},\mathcal{D}_{U}=\left\{\frac{y_{j}-y_{i}}{x_{j}-x_{i}}\colon 1\leq i<j\leq|U|\right\}\subset\mathbb{F}_{q}\cup\{\infty\},

where \infty is the vertical direction. The theory of directions has been developed by Rédei [LR73], Szőnyi [Sz], and many other authors. The main methods in the theory revolve around studying certain lacunary polynomials and symmetric polynomials. In modern language, the theory was built via the application of the polynomial method over finite fields. We remark that it has been used several times to study cliques of Paley graphs and generalized Paley graphs. One highlight is that it can be used to deduce the best-known upper bounds on the clique number of Paley graphs and generalized Paley graphs; see recent papers [DSW, Yip2].

The standard way to identify 𝔽q2\mathbb{F}_{q^{2}} and AG(2,q)AG(2,q) is via an embedding with respect to a basis of 𝔽q2\mathbb{F}_{q^{2}} over 𝔽q\mathbb{F}_{q} seen as a 22-dimensional vector space over 𝔽q\mathbb{F}_{q}. We say that π:𝔽q2AG(2,q)\pi:\mathbb{F}_{q^{2}}\to AG(2,q) is an embedding if

π(au+bv)=(a,b),a,b𝔽q,\pi(au+bv)=(a,b),\forall a,b\in\mathbb{F}_{q},

where {u,v}\{u,v\} forms a basis of 𝔽q2\mathbb{F}_{q^{2}} over 𝔽q\mathbb{F}_{q}. In the definition below, the term FF-subspace means a vector subspace defined over a field FF.

Definition 3.1 ([BBBSS]).

Let UU be a subset of AG(2,q)AG(2,q) containing the origin, and let KK be a subfield of 𝔽q\mathbb{F}_{q}. We say UU is KK-linear if there is an embedding π0\pi_{0} of 𝔽q2\mathbb{F}_{q^{2}} into the AG(2,q)AG(2,q), such that π01(U)\pi_{0}^{-1}(U) forms a KK-subspace of 𝔽q2\mathbb{F}_{q^{2}}.

The following theorem, due to Ball [Ball03], characterizes the number of directions determined by the graph of a function. It was built on previous works [LR73, BBS, BBBSS]. We present below the simplified version of Ball’s original formulation.

Theorem 3.2 ([Ball03]).

Let f:𝔽q𝔽qf:\mathbb{F}_{q}\to\mathbb{F}_{q} be any function such that f(0)=0f(0)=0, where qq is an odd prime power. Let NN be the number of directions determined by the graph of ff. Then either Nq+32N\geq\frac{q+3}{2}, or there is a subfield KK of 𝔽q\mathbb{F}_{q} such that the graph of ff is KK-linear. Moreover, in the latter case, if KK is the largest subfield over which the graph is KK-linear and K𝔽qK\neq\mathbb{F}_{q}, then Nq/|K|+1N\geq q/|K|+1.

If UAG(2,q)U\subset AG(2,q) and |U|=q|U|=q, we can try to apply Theorem 3.2 to study 𝒟U\mathcal{D}_{U}. This is because an affine transformation preserves the number of directions; if UU does not determine all the directions, then UU is a graph of a function after an affine transformation. The following stability result, due to Szőnyi [Sz1] and Sziklai [Szi99], enables us to extend Theorem 3.2 to any UAG(2,q)U\subset AG(2,q) with size qO(q)q-O(\sqrt{q}).

Theorem 3.3 ([Sz1]*Theorem 4, [Szi99]*Theorem 3.1).

Let UAG(2,q)U\subset AG(2,q) with |U|=qk|U|=q-k, where 0kαq0\leq k\leq\alpha\sqrt{q}, where α[1/2,1]\alpha\in[1/2,1]. If UU determines less than (q+1)(1α)(q+1)(1-\alpha) directions, then it can be extended to a set UU^{\prime} with |U|=q|U^{\prime}|=q, which determines the same set of directions as UU.

3.2. Character sums over subspaces

Recall that a multiplicative character of 𝔽q\mathbb{F}_{q} is a group homomorphism from 𝔽q\mathbb{F}_{q}^{*} to the multiplicative group of complex numbers with modulus 1. It is customary to define χ(0)=0\chi(0)=0. We will use χ0\chi_{0} to denote the trivial multiplicative character of 𝔽q\mathbb{F}_{q}. For a multiplicative character χ\chi, its order dd is the smallest positive integer such that χd=χ0\chi^{d}=\chi_{0}.

Many problems in number theory and combinatorics rely on non-trivial character sum estimates over certain subsets of interest. We refer to [Chang] for a nice survey paper by Chang, consisting of known results on character sums over a subset of a given finite field, such as the extension of the classical Pólya-Vinogradov inequalities and Burgess’ bounds to a general finite field. For our purpose, we need nontrivial character sum estimates over a subspace of a finite field.

Weil’s estimate (see [LN]*Theorem 5.41) is a classical tool of bounding complete character sums with a polynomial argument over a finite field. The following theorem, due to Katz [Katz], essentially generalizes Weil’s estimate to character sums over a subspace. Recall an element θ𝔽qn\theta\in\mathbb{F}_{q^{n}} is said to have degree kk over 𝔽q\mathbb{F}_{q} if 𝔽qk\mathbb{F}_{q^{k}} is the smallest extension of 𝔽q\mathbb{F}_{q} that contains θ\theta.

Theorem 3.4 ([Katz]).

Let θ\theta be an element of degree nn over 𝔽q\mathbb{F}_{q} and χ\chi a non-trivial multiplicative character of 𝔽qn.\mathbb{F}_{q^{n}}. Then

|a𝔽qχ(θ+a)|(n1)q.\left|\sum_{a\in\mathbb{F}_{q}}\chi(\theta+a)\right|\leq(n-1)\sqrt{q}.

Recently, Reis [Reis] used Theorem 3.4 to prove the following character sum estimate over an affine space.

Theorem 3.5 ([Reis]).

Let 𝒜=u+𝒱𝔽qn\mathcal{A}=u+\mathcal{V}\subseteq\mathbb{F}_{q^{n}} be an 𝔽q\mathbb{F}_{q}-affine space of dimension t1t\geq 1, where n>1.n>1. Suppose that there exists a nonzero element y𝒱y\in\mathcal{V} such that the set 𝒜y:={ay1a𝒜}\mathcal{A}_{y}:=\left\{ay^{-1}\mid a\in\mathcal{A}\right\} contains an element of degree nn over 𝔽q.\mathbb{F}_{q}. If χ\chi is a non-trivial multiplicative character of 𝔽qn\mathbb{F}_{q^{n}}, then

|a𝒜χ(a)|<nqt1/2.\left|\sum_{a\in\mathcal{A}}\chi(a)\right|<n\cdot q^{t-1/2}. (2)

The hypothesis on the existence of a degree nn element cannot be completely dropped. For equation (2) to hold, we must have some assumptions on the structure of the affine space 𝒜\mathcal{A}. Typically, we need to assume that 𝒜\mathcal{A} does not have a special algebraic structure. As a counter-example, let 𝒜=𝒱=𝔽qt\mathcal{A}=\mathcal{V}=\mathbb{F}_{q^{t}} be a proper subfield. Then we can find a non-trivial multiplicative character χ\chi on 𝔽qn\mathbb{F}_{q^{n}} such that χ|A\chi|_{A} is the trivial character; in this case, a𝒜χ(a)=qt1>nqt1/2\sum_{a\in\mathcal{A}}\chi(a)=q^{t}-1>nq^{t-1/2} holds for sufficiently large qq, violating the conclusion in Theorem 3.5. This construction is an example of the so-called subfield obstruction.

Chang [Chang]*Section 5 also gave a somewhat weaker nontrivial estimate of character sums over subspaces of finite fields. Here we use Theorem 3.5 since it is stronger to some extent, and it does not involve any implicit constants, which can be challenging to determine.

The following corollary gives a non-trivial upper bound on the character sums over a subspace when 𝒜\mathcal{A} is a subspace but not a subfield. Corollary 3.6 is a consequence of Theorem 3.5 and [Reis]*Proposition 2.2; here we include the proof for the sake of completeness.

Corollary 3.6.

Let nn be an integer such that n2n\geq 2, and qq an odd prime power. Let 𝒱𝔽q2n\mathcal{V}\subseteq\mathbb{F}_{q^{{2n}}} be an 𝔽q\mathbb{F}_{q}-space of dimension nn, with 1𝒱1\in\mathcal{V}, and 𝒱𝔽qn\mathcal{V}\neq\mathbb{F}_{q^{n}}. Then for any non-trivial multiplicative character χ\chi of 𝔽q2n\mathbb{F}_{q^{2n}},

|x𝒱χ(x)|<2nq|𝒱|.\left|\sum_{x\in\mathcal{V}}\chi(x)\right|<\frac{2n}{\sqrt{q}}\cdot|\mathcal{V}|. (3)
Proof.

It suffices to show that if 𝒱\mathcal{V} has size qnq^{n} and 𝒱𝔽qn\mathcal{V}\neq\mathbb{F}_{q^{n}}, then 𝒱\mathcal{V} has an element with degree 2n2n. To prove this assertion, note that the number of elements in the union of 𝔽qd\mathbb{F}_{q^{d}} as d2nd\mid 2n with d<nd<n is less than qn1q^{n-1}. Suppose that 𝒱\mathcal{V} has size qnq^{n}, and 𝒱𝔽qn\mathcal{V}\neq\mathbb{F}_{q^{n}}. If 𝒱\mathcal{V} did not have any element of degree 2n2n, then all the elements of 𝒱𝔽qn\mathcal{V}\setminus\mathbb{F}_{q^{n}} would be contained in d2n,d<n𝔽qd\bigcup_{\begin{subarray}{c}d\mid 2n,d<n\end{subarray}}\mathbb{F}_{q^{d}}. Now, 𝒱𝔽qn\mathcal{V}\cap\mathbb{F}_{q^{n}} has dimension at most n1n-1, which means that at least qnqn1q^{n}-q^{n-1} many elements of 𝒱\mathcal{V} must belong to d2n,d<n𝔽qd\bigcup_{\begin{subarray}{c}d\mid 2n,d<n\end{subarray}}\mathbb{F}_{q^{d}}. This is a contradiction because for q3q\geq 3, we have

|d2n,d<n𝔽qd|i=1n1qi=qn1q11qn121<qnqn1.\bigg{|}\bigcup_{\begin{subarray}{c}d\mid 2n,d<n\end{subarray}}\mathbb{F}_{q^{d}}\bigg{|}\leq\sum_{i=1}^{n-1}q^{i}=\frac{q^{n}-1}{q-1}-1\leq\frac{q^{n}-1}{2}-1<q^{n}-q^{n-1}.\qed

Note that if nn is an odd prime, then any element in 𝔽qn𝔽q\mathbb{F}_{q^{n}}\setminus\mathbb{F}_{q} has degree nn. Thus, we obtain the following corollary of Theorem 3.5, which will be used in Section 5.

Corollary 3.7.

Let 𝒱𝔽qn\mathcal{V}\subseteq\mathbb{F}_{q^{n}} be an 𝔽q\mathbb{F}_{q}-subspace of dimension 22 , where qq is an odd prime power, nn is an odd prime with q>n2q>n^{2}, and 1𝒱1\in\mathcal{V}. If χ\chi is a non-trivial multiplicative character of 𝔽qn\mathbb{F}_{q^{n}}, then

|v𝒱χ(v)|<|𝒱|1.\left|\sum_{v\in\mathcal{V}}\chi(v)\right|<|\mathcal{V}|-1. (4)
Proof.

In view of Corollary 3.6, it suffices to show that,

nq|𝒱|<|𝒱|1\displaystyle\frac{n}{\sqrt{q}}|\mathcal{V}|<|\mathcal{V}|-1

for qn2+1q\geq n^{2}+1. Using |𝒱|=q2|\mathcal{V}|=q^{2}, the above inequality is equivalent to:

n2q<(11q2)2\displaystyle\frac{n^{2}}{q}<\left(1-\frac{1}{q^{2}}\right)^{2} (5)

The inequality (5) follows directly from combining the two inequalities below:

n2qq1q=11qand 11q<(11q2)2\displaystyle\frac{n^{2}}{q}\leq\frac{q-1}{q}=1-\frac{1}{q}\ \ \ \ \text{and}\ \ \ 1-\frac{1}{q}<\left(1-\frac{1}{q^{2}}\right)^{2}

for each q2q\geq 2. ∎

4. Maximum cliques in a Peisert-type graph

We start the section by proving Lemma 2.13. In Section 4.2, we prove our main theorems and mention a few corollaries. Finally, in Section 4.3, we classify maximum cliques in generalized Peisert graphs as an application of our main theorem.

4.1. Proof of Lemma 2.13

As a motivation for Definition 1.1, Lemma 2.13 states that (generalized) Paley and Peisert graphs are special instances of Peisert-type graphs.

Proof of Lemma 2.13.

The first two families of graphs are special cases of the last two. Thus, it suffices to prove the claim for generalized Paley graphs and generalized Peisert graphs.

A generalized Paley graph GP(q2,d)GP(q^{2},d), where d(q+1)d\mid(q+1) and d>1d>1, is of the form Cay(𝔽q2+,S)\operatorname{Cay}(\mathbb{F}_{q^{2}}^{+},S) where SS consists of all dd-th powers in 𝔽q2\mathbb{F}_{q^{2}}^{\ast}. Since d(q+1)d\mid(q+1), we have 𝔽qS\mathbb{F}_{q}^{\ast}\subset S. Note that SS is the subgroup obtained by taking the image of the group homomorphism 𝔽q2𝔽q2\mathbb{F}_{q^{2}}^{\ast}\to\mathbb{F}_{q^{2}}^{\ast} sending xxdx\mapsto x^{d}. It follows that |S|=q21d|S|=\frac{q^{2}-1}{d}. Finally, 𝔽q\mathbb{F}_{q}^{\ast} is a subgroup of SS with index

|S||𝔽q|=(q21)/dq1=q+1d\frac{|S|}{|\mathbb{F}_{q}^{\ast}|}=\frac{(q^{2}-1)/d}{q-1}=\frac{q+1}{d}

which allows us to write,

S=c1𝔽qcm𝔽qS=c_{1}\mathbb{F}_{q}^{\ast}\cup\cdots\cup c_{m}\mathbb{F}_{q}^{\ast}

with c1=1c_{1}=1 and m=q+1dq+12m=\frac{q+1}{d}\leq\frac{q+1}{2}. Thus, GP(q2,d)GP(q^{2},d) is a Peisert-type graph.

For the generalized Peisert graph GP(q2,d)GP^{*}(q^{2},d), where d(q+1)d\mid(q+1) and dd is even, the connection set SS can be expressed as a disjoint union:

S=i=0d/21gi(𝔽q2)d.S=\bigsqcup_{i=0}^{d/2-1}g^{i}(\mathbb{F}_{q^{2}}^{\ast})^{d}.

It follows that GP(q2,d)GP^{*}(q^{2},d) is the edge-disjoint union of d2\frac{d}{2} copies of GP(q2,d)GP(q^{2},d). Since each gi(𝔽q2)dg^{i}(\mathbb{F}_{q^{2}}^{\ast})^{d} can be expressed as a union of q+1d\frac{q+1}{d} cosets of 𝔽q\mathbb{F}_{q}^{\ast}, we can express SS above as,

S=c1𝔽qcm𝔽qS=c_{1}\mathbb{F}_{q}^{\ast}\cup\cdots\cup c_{m}\mathbb{F}_{q}^{\ast}

with c1=1c_{1}=1 and m=q+1dd2=q+12m=\frac{q+1}{d}\cdot\frac{d}{2}=\frac{q+1}{2}. Thus, GP(q2,d)GP^{*}(q^{2},d) is a Peisert-type graph. ∎

4.2. Proof of main results and consequences

With the two tools described in the previous section, we are ready to prove Theorem 1.2.

Proof of Theorem 1.2.

Let v𝔽q2Sv\in\mathbb{F}_{q^{2}}^{*}\setminus S. Since 𝔽qS\mathbb{F}_{q}^{*}\subset S, {1,v}\{1,v\} forms a basis of 𝔽q2\mathbb{F}_{q^{2}} over 𝔽q\mathbb{F}_{q}. We consider the following standard embedding π\pi of 𝔽q2\mathbb{F}_{q^{2}} into AG(2,q)AG(2,q), where

π(a+bv)=(a,b),a,b𝔽q.\pi(a+bv)=(a,b),\quad\forall a,b\in\mathbb{F}_{q}.

It is clear that the subfield 𝔽q\mathbb{F}_{q} forms a clique in the Cayley graph XX, so ω(X)q\omega(X)\geq q.

Let CC be a maximum clique in XX containing 0. We can express UU, the image of CC under the embedding π\pi as

U=π(C)={(aj,bj):1j|C|}AG(2,q).U=\pi(C)=\{(a_{j},b_{j}):1\leq j\leq|C|\}\subset AG(2,q).

If there are 1j<k|C|1\leq j<k\leq|C| such that aj=aka_{j}=a_{k}, then bjbkb_{j}\neq b_{k}, and (bjbk)vSv𝔽q(b_{j}-b_{k})v\in S\cap v\mathbb{F}_{q}^{*} since CC is a clique. Since SS is a union of cosets of 𝔽q\mathbb{F}_{q}^{*}, this implies that vSv\in S, contradicting our assumption. This also rules out the possibility that |C|q+1|C|\geq q+1 by the pigeonhole principle. Therefore, |C|=ω(X)=q|C|=\omega(X)=q.

Let S=j=1mcj𝔽qS=\bigcup_{j=1}^{m}c_{j}\mathbb{F}_{q}^{*}. Note that C={aj+bjv:1jq}C=\{a_{j}+b_{j}v:1\leq j\leq q\}. As we saw above, for any 1j<kq1\leq j<k\leq q, we have ajaka_{j}\neq a_{k}. Since CC is a clique, we have (aj+bjv)(ak+bkv)=(ajak)+(bjbk)vS(a_{j}+b_{j}v)-(a_{k}+b_{k}v)=(a_{j}-a_{k})+(b_{j}-b_{k})v\in S; it follows that

1+bjbkajakvS(1+𝔽qv),\displaystyle 1+\frac{b_{j}-b_{k}}{a_{j}-a_{k}}v\in S\cap(1+\mathbb{F}_{q}v),

since SS is a union of cosets of 𝔽q\mathbb{F}_{q}^{*}, ajak𝔽qa_{j}-a_{k}\in\mathbb{F}_{q}^{*}, and bjbk𝔽qb_{j}-b_{k}\in\mathbb{F}_{q}.

Notice that

S(1+𝔽qv)=j=1m(cj𝔽q(1+𝔽qv)).S\cap(1+\mathbb{F}_{q}v)=\bigcup_{j=1}^{m}\big{(}c_{j}\mathbb{F}_{q}^{*}\cap(1+\mathbb{F}_{q}v)\big{)}.

We claim that |cj𝔽q(1+𝔽qv)|1|c_{j}\mathbb{F}_{q}^{*}\cap(1+\mathbb{F}_{q}v)|\leq 1 for each 1jm1\leq j\leq m. Indeed, if x,ycj𝔽q(1+𝔽qv)x,y\in c_{j}\mathbb{F}_{q}^{*}\cap(1+\mathbb{F}_{q}v), then xy=tvx-y=tv for some t𝔽qt\in\mathbb{F}_{q}; combining tvcj𝔽qtv\in c_{j}\mathbb{F}_{q} with vSv\notin S, we get t=0t=0, and so x=yx=y.

As we saw above, UU does not determine the vertical direction. Since |U|=q|U|=q, we must have that

U={(x,f(x)):x𝔽q}\displaystyle U=\{(x,f(x)):x\in\mathbb{F}_{q}\}

is the graph of some function f:𝔽q𝔽qf:\mathbb{F}_{q}\to\mathbb{F}_{q} such that f(0)=0f(0)=0. Recall that the direction set 𝒟U\mathcal{D}_{U} consists of all ratios bjbkajak\frac{b_{j}-b_{k}}{a_{j}-a_{k}} for 1j<kq1\leq j<k\leq q. We have shown that,

1+𝒟UvS(1+𝔽qv)=j=1m(cj𝔽q(1+𝔽qv))1+\mathcal{D}_{U}\cdot v\subset S\cap(1+\mathbb{F}_{q}v)=\bigcup_{j=1}^{m}\big{(}c_{j}\mathbb{F}_{q}^{*}\cap(1+\mathbb{F}_{q}v)\big{)}

has size at most mq+12m\leq\frac{q+1}{2}. Equivalently, |𝒟U|q+12|\mathcal{D}_{U}|\leq\frac{q+1}{2}. By Theorem 3.2, it follows that UU is KK-linear for some subfield K𝔽qK\subset\mathbb{F}_{q}. In particular, UU is 𝔽p\mathbb{F}_{p}-linear, and so CC is an 𝔽p\mathbb{F}_{p}-subspace of 𝔽q2\mathbb{F}_{q^{2}}. ∎

Inspired by the proof of Theorem 1.2, we provide a sufficient condition for a maximum clique in a Peisert-type graph on 𝔽q2\mathbb{F}_{q^{2}} to be the subfield 𝔽q\mathbb{F}_{q}.

Corollary 4.1.

Let X=Cay(𝔽q2+,S)X=\operatorname{Cay}(\mathbb{F}_{q^{2}}^{+},S) be a Peisert-type graph, where q=pnq=p^{n}. Suppose that m=|S|/(q1)pnkm=|S|/(q-1)\leq p^{n-k}, where kk is the largest proper divisor of nn. Then the only maximum clique in XX containing 0,10,1 is the subfield 𝔽q\mathbb{F}_{q}.

Proof.

Let CC be a maximum clique in XX such that 0C0\in C. Following the proof of Theorem 1.2, if we embed CC into AG(2,q)AG(2,q), then its image UU determines at most mm directions and UU is KK-linear, where KK is a subfield of 𝔽q\mathbb{F}_{q}. If K𝔽qK\neq\mathbb{F}_{q}, then Theorem 3.2 implies that mq/|K|+1pnk+1m\geq q/|K|+1\geq p^{n-k}+1, contradicting the assumption. Therefore, UU is 𝔽q\mathbb{F}_{q}-linear, that is, CC is an 𝔽q\mathbb{F}_{q}-subspace. In particular, if 1C1\in C, then CC must be the subfield 𝔽q\mathbb{F}_{q}. ∎

Remark 4.2.

There are counterexamples of Peisert-type graphs with m=pn1pk1m=\frac{p^{n}-1}{p^{k}-1} where 𝔽q\mathbb{F}_{q} is not the only maximum clique containing 0 and 11. Explicit constructions can be found in [AGLY22]*Section 5.1. Note that pn1pk1\frac{p^{n}-1}{p^{k}-1} is only slightly bigger than pnkp^{n-k}, which means that the previous corollary is nearly sharp.

With the help of Theorem 3.3, namely the stability result on the direction set, the proof of Theorem 1.2 can be easily modified to prove Theorem 2.15.

Proof of Theorem 2.15.

Let CC be a clique in XX, such that 0C0\in C and |C|>q(1m/(q+1))q|C|>q-\big{(}1-m/(q+1)\big{)}\sqrt{q}. Similar to the proof of Theorem 1.2, we can embed CC into AG(2,q)AG(2,q) using the same map π\pi to obtain UAG(2,q)U\subset AG(2,q). By the exact same reasoning, UU determines at most m<q+12m<\frac{q+1}{2} directions, and UU does not determine the vertical direction.

We can now apply Theorem 3.3 with |U|=qk|U|=q-k where k<(1mq+1)qk<\left(1-\frac{m}{q+1}\right)\sqrt{q} and α=1mq+1ε[1/2,1]\alpha=1-\frac{m}{q+1}-\varepsilon\in[1/2,1] for a sufficiently small ε>0\varepsilon>0. This allows us to extend UU to a larger set UAG(2,q)U^{\prime}\subset AG(2,q), such that |U|=q|U^{\prime}|=q, and UU^{\prime} determines the same set of directions as UU; in particular, 1+𝒟Uv=1+𝒟Uv1+\mathcal{D}_{U}v=1+\mathcal{D}_{U^{\prime}}v. Let (a,b)(a,b)U(a,b)\neq(a^{\prime},b^{\prime})\in U^{\prime}; then aaa\neq a^{\prime}, and

1+bbaav1+𝒟UvS.1+\frac{b^{\prime}-b}{a^{\prime}-a}v\in 1+\mathcal{D}_{U^{\prime}}v\subset S.

It follows that (aa)+(bb)vS(a^{\prime}-a)+(b^{\prime}-b)v\in S; in other words, the preimage C=π1(U)C^{\prime}=\pi^{-1}(U^{\prime}) forms a maximum clique in XX. The required claim follows from Theorem 1.2. ∎

Now we are ready to present the proof of Theorem 1.3.

Proof of Theorem 1.3.

We can always assume ε1\varepsilon\leq 1 in the statement of the theorem since characters have absolute value at most 11. Suppose that there is a nontrivial multiplicative character χ\chi of 𝔽q2\mathbb{F}_{q^{2}} such that the set {χ(x):xS}\{\chi(x):x\in S\} is ε\varepsilon-lower bounded. By Theorem 1.2, if 0C0\in C is a clique in the Peisert-type graph XX, with |C|=q|C|=q, then CC is an 𝔽p\mathbb{F}_{p}-subspace of 𝔽q2\mathbb{F}_{q^{2}}.

Suppose 0,1C0,1\in C, and C𝔽qC\neq\mathbb{F}_{q}. Then CC is an 𝔽p\mathbb{F}_{p}-subspace of 𝔽q2=𝔽p2n\mathbb{F}_{q^{2}}=\mathbb{F}_{p^{2n}} and C{0}SC\setminus\{0\}\subset S. By Corollary 3.6,

|xCχ(x)|<2np|C|.\left|\sum_{x\in C}\chi(x)\right|<\frac{2n}{\sqrt{p}}\cdot|C|.

On the other hand, the set {χ(x):xS}\{\chi(x):x\in S\} is ε\varepsilon-lower bounded, which guarantees that

|xCχ(x)|ε(|C|1).\left|\sum_{x\in C}\chi(x)\right|\geq\varepsilon(|C|-1).

Comparing these two inequalities, we obtain

p<4n2ε2(qq1)2.p<\frac{4n^{2}}{\varepsilon^{2}}\left(\frac{q}{q-1}\right)^{2}.

Now, p>4n216p>4n^{2}\geq 16 and thus qp2172=289q\geq p^{2}\geq 17^{2}=289. Thus,

p<(4n2ε2)(289288)24.02783n2ε2\displaystyle p<\left(\frac{4n^{2}}{\varepsilon^{2}}\right)\left(\frac{289}{288}\right)^{2}\approx\frac{4.02783n^{2}}{\varepsilon^{2}}

which contradicts our hypothesis. Therefore, if 0,1C0,1\in C, then C=𝔽qC=\mathbb{F}_{q}. ∎

Remark 4.3.

Under additional conditions on the size of the connection set, the range for the values of pp in Theorem 1.3 can be improved. We already saw this phenomenon in Corollary 4.1. More generally, suppose that |S|q1q1δ\frac{|S|}{q-1}\leq q^{1-\delta} where δ=1/s\delta=1/s and ss is a proper divisor of nn. Then using Theorem 3.2 and applying Theorem 3.5 over the base field 𝔽qδ=𝔽pn/s\mathbb{F}_{q^{\delta}}=\mathbb{F}_{p^{n/s}}, we see that Theorem 1.3 will hold under the weaker hypothesis pn/s>9s2ε2p^{n/s}>\frac{9s^{2}}{\varepsilon^{2}}.

The following corollary follows immediately from Theorem 1.3 and Theorem 2.15.

Corollary 4.4.

Under the same assumptions as in Theorem 1.3, if |S|<q212|S|<\frac{q^{2}-1}{2}, then for any clique CC in XX with |C|>q(1mq+1)q|C|>q-\big{(}1-\frac{m}{q+1}\big{)}\sqrt{q} and 0,1C0,1\in C, we have C𝔽qC\subset\mathbb{F}_{q}.

4.3. Application to generalized Peisert graphs

We start with the following lemma, concerning the connection set of a generalized Peisert graph.

Lemma 4.5.

Let d4d\geq 4 be an even integer, and θ=exp(2πi/d)\theta=\exp(2\pi i/d). Then the set M={θj:0jd/21}M=\{\theta^{j}:0\leq j\leq d/2-1\} is (πdπd2)\left(\frac{\pi}{d}-\frac{\pi}{d^{2}}\right)-lower bounded.

Proof.

Let aja_{j} be a non-negative integer for each 0jd/210\leq j\leq d/2-1, and b=j=1d/21ajb=\sum_{j=1}^{d/2-1}a_{j}. We have

(j=0d/21ajθj)=j=1d/21ajsin(2πjd)j=1d/21ajsin(2πd)=bsin2πd.\displaystyle\Im\bigg{(}\sum_{j=0}^{d/2-1}a_{j}\theta^{j}\bigg{)}=\sum_{j=1}^{d/2-1}a_{j}\sin\left(\frac{2\pi j}{d}\right)\geq\sum_{j=1}^{d/2-1}a_{j}\sin\left(\frac{2\pi}{d}\right)=b\sin\frac{2\pi}{d}.

If a0bcos2πda_{0}\leq b\cos\frac{2\pi}{d}, then

|j=0d/21ajθj|(j=0d/21ajθj)bsin2πd=b(1+cos2πd1+cos2πd)sin(2πd)(b+a0)sin2πd1+cos2πd.\displaystyle\bigg{|}\sum_{j=0}^{d/2-1}a_{j}\theta^{j}\bigg{|}\geq\Im\bigg{(}\sum_{j=0}^{d/2-1}a_{j}\theta^{j}\bigg{)}\geq b\sin\frac{2\pi}{d}=b\left(\frac{1+\cos\frac{2\pi}{d}}{1+\cos\frac{2\pi}{d}}\right)\sin\left(\frac{2\pi}{d}\right)\geq(b+a_{0})\frac{\sin\frac{2\pi}{d}}{1+\cos\frac{2\pi}{d}}.

If a0>bcos2πda_{0}>b\cos\frac{2\pi}{d}, then

(j=0d/21ajθj)=a0+j=1d/21ajcos(2πjd)a0j=1d/21ajcos(2πd)=a0bcos2πd>0;\displaystyle\Re\bigg{(}\sum_{j=0}^{d/2-1}a_{j}\theta^{j}\bigg{)}=a_{0}+\sum_{j=1}^{d/2-1}a_{j}\cos\left(\frac{2\pi j}{d}\right)\geq a_{0}-\sum_{j=1}^{d/2-1}a_{j}\cos\left(\frac{2\pi}{d}\right)=a_{0}-b\cos\frac{2\pi}{d}>0;

therefore,

|j=0d/21ajθj|2(a0bcos2πd)2+(bsin2πd)2=(a0+b)22a0b(1+cos2πd)(a0+b)21cos2πd2.\bigg{|}\sum_{j=0}^{d/2-1}a_{j}\theta^{j}\bigg{|}^{2}\geq\bigg{(}a_{0}-b\cos\frac{2\pi}{d}\bigg{)}^{2}+\bigg{(}b\sin\frac{2\pi}{d}\bigg{)}^{2}=(a_{0}+b)^{2}-2a_{0}b\bigg{(}1+\cos\frac{2\pi}{d}\bigg{)}\geq(a_{0}+b)^{2}\frac{1-\cos\frac{2\pi}{d}}{2}.

Note that we always have

sin2πd1+cos2πd1cos2πd2.\frac{\sin\frac{2\pi}{d}}{1+\cos\frac{2\pi}{d}}\geq\sqrt{\frac{1-\cos\frac{2\pi}{d}}{2}}.

Therefore, MM is 1cos2πd2\sqrt{\frac{1-\cos\frac{2\pi}{d}}{2}}-lower bounded. As dd\to\infty, MM is πd+O(d2)\frac{\pi}{d}+O(d^{-2})-lower bounded. In particular, one can check that 1cos2πd2πdπd2\sqrt{\frac{1-\cos\frac{2\pi}{d}}{2}}\geq\frac{\pi}{d}-\frac{\pi}{d^{2}} using Taylor series expansion, and thus, MM is (πdπd2)\left(\frac{\pi}{d}-\frac{\pi}{d^{2}}\right)-lower bounded. ∎

Now we are ready to prove Theorem 2.20.

Proof of Theorem 2.20.

Let gg be a primitive root of 𝔽q2\mathbb{F}_{q^{2}}^{*}, and θ=exp(2πi/d)\theta=\exp(2\pi i/d). Let χ\chi be the multiplicative character in 𝔽q2\mathbb{F}_{q^{2}} such that χ(g)=θ\chi(g)=\theta. Recall that the connection set SS of the generalized Peisert graph X=GP(q2,d)X=GP^{*}(q^{2},d) is given by

S={gdk+j:0jd21,k}.S=\bigg{\{}g^{dk+j}:0\leq j\leq\frac{d}{2}-1,k\in\mathbb{Z}\bigg{\}}.

Therefore, {χ(x):xS}={θj:0jd/21}\{\chi(x):x\in S\}=\{\theta^{j}:0\leq j\leq d/2-1\} is (πdπd2)\left(\frac{\pi}{d}-\frac{\pi}{d^{2}}\right)-lower bounded by Lemma 4.5. We have shown that XX is a Peisert-type graph in Lemma 2.13; therefore, we can apply Theorem 1.3 to get the desired characterization of maximum cliques in XX provided that

p>4.1n2(πdπd2)2=4.1n2d4π2(d1)2.p>\frac{4.1n^{2}}{\big{(}\frac{\pi}{d}-\frac{\pi}{d^{2}}\big{)}^{2}}=\frac{4.1n^{2}d^{4}}{\pi^{2}(d-1)^{2}}.\qed

5. Maximal cliques in generalized Paley graphs and Peisert graphs

Recall that our strategy to classify maximum cliques is the following: first show that every maximum clique has a subspace structure, and then apply character sum estimates over a subspace to show a maximum clique must be a subfield. In this section, we employ a similar idea to study maximal cliques. Assuming that a given clique is not maximal, if we can extend the clique and construct a maximal clique with a subspace structure, then it is possible that we can apply character sum estimates to derive a contradiction.

For generalized Paley graphs, this approach is made possible via the following lemma, established by the second author [Yip3].

Lemma 5.1 ([Yip3]*Corollary 3.1).

Let GP(q,d)GP(q,d) be a generalized Paley graph. If KK is a proper subfield such that KK forms a clique in GP(q,d)GP(q,d), then there is a KK-subspace 𝒱\mathcal{V} of 𝔽q\mathbb{F}_{q}, such that K𝒱K\subset\mathcal{V}, and 𝒱\mathcal{V} forms a maximal clique in GP(q,d)GP(q,d).

Our aim is to provide partial progress towards the following main conjecture in [Yip3]:

Conjecture 5.2 ([Yip3]*Conjecture 1.4).

Let dd be a positive integer greater than 11. Let q1(mod2d)q\equiv 1\pmod{2d} be a power of a prime pp, and let rr be the largest integer such that 𝔽pr\mathbb{F}_{p^{r}} is a subfield of 𝔽q\mathbb{F}_{q} and dq1pr1d\mid\frac{q-1}{p^{r}-1}. Then the subfield 𝔽pr\mathbb{F}_{p^{r}} forms a maximal clique in GP(q,d)GP(q,d).

Conjecture 5.2 has been shown to be true for cubic Paley graphs with cubic order and quadruple Paley graphs with quartic order in [Yip3]. We establish the following special case of Conjecture 5.2.

Theorem 5.3.

Let dd be a positive integer greater than 11. Let nn be an odd prime. If q>n2q>n^{2} and dqn1q1d\mid\frac{q^{n}-1}{q-1}, then the subfield 𝔽q\mathbb{F}_{q} forms a maximal clique in GP(qn,d)GP(q^{n},d).

Proof.

Since qq is an odd prime power, dqn1q1d\mid\frac{q^{n}-1}{q-1} implies that qn1(mod2d)q^{n}\equiv 1\pmod{2d}, and thus GP(qn,d)GP(q^{n},d) is a well-defined dd-Paley graph. It is easy to verify that the condition dqn1q1d\mid\frac{q^{n}-1}{q-1} guarantees that 𝔽q\mathbb{F}_{q} forms a clique in GP(qn,d)GP(q^{n},d); see [BDR]*Theorem 1.

Suppose that 𝔽q\mathbb{F}_{q} is not a maximal clique in GP(qn,d)GP(q^{n},d). Then Lemma 5.1 implies that there is a clique 𝒱\mathcal{V} in GP(qn,d)GP(q^{n},d), such that 𝒱\mathcal{V} is a 2-dimensional 𝔽q\mathbb{F}_{q}-subspace, and 𝔽q𝒱\mathbb{F}_{q}\subset\mathcal{V}. Therefore, if we let χ\chi to be a multiplicative character of 𝔽qn\mathbb{F}_{q^{n}} with order dd, then

|v𝒱χ(v)|=|𝒱|1,\left|\sum_{v\in\mathcal{V}}\chi(v)\right|=|\mathcal{V}|-1, (6)

which violates inequality (4) in Corollary 3.7, provided q>n2q>n^{2}. This shows that 𝔽q\mathbb{F}_{q} is a maximal clique in GP(qn,d)GP(q^{n},d). ∎

To prove Theorem 1.5, we use a similar strategy. We need the following result from [Yip3], which explores the connection between maximal cliques and maximum cliques in a Peisert graph of order q4q^{4}.

Theorem 5.4 ([Yip3]*Theorem 1.7).

Let qq be a power of a prime p3(mod4)p\equiv 3\pmod{4}. If 𝔽q\mathbb{F}_{q} is not a maximal clique in the Peisert graph of order q4q^{4}, then ω(Pq4)=q2\omega(P^{*}_{q^{4}})=q^{2}; moreover, there exists h𝔽q4𝔽qh\in\mathbb{F}_{q^{4}}\setminus\mathbb{F}_{q}, such that 𝔽qh𝔽q\mathbb{F}_{q}\oplus h\mathbb{F}_{q} forms a maximum clique.

We now have the tools to prove Theorem 1.5.

Proof of Theorem 1.5.

Let gg be a primitive root of 𝔽q4\mathbb{F}_{q^{4}}. Let χ\chi be the multiplicative character of 𝔽q4\mathbb{F}_{q^{4}} such that χ(g)=i\chi(g)=i. It follows that for any kk\in\mathbb{N}, χ(g4k)=1\chi(g^{4k})=1, and χ(g4k+1)=i\chi(g^{4k+1})=i.

Suppose that 𝔽q\mathbb{F}_{q} is not a maximal clique in the Peisert graph Pq4P^{*}_{q^{4}}. By Theorem 5.4, there exists h𝔽q4𝔽qh\in\mathbb{F}_{q^{4}}\setminus\mathbb{F}_{q}, such that 𝒱=𝔽qh𝔽q\mathcal{V}=\mathbb{F}_{q}\oplus h\mathbb{F}_{q} forms a maximum clique. In particular, for each x𝒱{0}x\in\mathcal{V}\setminus\{0\}, we have χ(x)=i\chi(x)=i or χ(x)=1\chi(x)=1. Consequently,

x𝒱χ(x)=a+bi,\sum_{x\in\mathcal{V}}\chi(x)=a+bi,

where a,ba,b are non-negative integers such that a+b=|𝒱|1a+b=|\mathcal{V}|-1. Therefore,

|x𝒱χ(x)|=a2+b2(a+b)2/2=|𝒱|12=q212.\bigg{|}\sum_{x\in\mathcal{V}}\chi(x)\bigg{|}=\sqrt{a^{2}+b^{2}}\geq\sqrt{(a+b)^{2}/2}=\frac{|\mathcal{V}|-1}{\sqrt{2}}=\frac{q^{2}-1}{\sqrt{2}}. (7)

On the other hand, 𝒱\mathcal{V} is an 𝔽q\mathbb{F}_{q}-subspace with dimension 22. Moreover, note that gq2+1g^{q^{2}+1} is a primitive root of the subfield 𝔽q2\mathbb{F}_{q^{2}}; in particular, since q2+12(mod4)q^{2}+1\equiv 2\pmod{4}, the subfield 𝔽q2\mathbb{F}_{q^{2}} does not form a clique in the Peisert graph of order q4q^{4}. Therefore, 𝒱𝔽q2\mathcal{V}\neq\mathbb{F}_{q^{2}}. By Corollary 3.6,

|x𝒱χ(x)|<4q3/2.\bigg{|}\sum_{x\in\mathcal{V}}\chi(x)\bigg{|}<4q^{3/2}. (8)

Combining inequalities (7) and (8), we must have q<33q<33. This implies that when q33q\geq 33, 𝔽q\mathbb{F}_{q} is a maximal clique in the Peisert graph of order q4q^{4}.

Finally, in [Yip3], using SageMath, we have verified that for each q{7,9,11,19,23,27,31}q\in\{7,9,11,19,23,27,31\}, 𝔽q\mathbb{F}_{q} is a maximal clique in the Peisert graph of order q4q^{4}. ∎

Remark 5.5.

The assumption q>3q>3 is necessary in Theorem 1.5. Indeed, one can check that when q=3q=3, the subfield 𝔽3\mathbb{F}_{3} is not maximal in the Peisert graph of order 8181. Indeed, there is a clique which is a two-dimensional 𝔽3\mathbb{F}_{3}-space containing 𝔽3\mathbb{F}_{3}.

We conclude the section with a similar statement regarding maximal cliques in a generalized Peisert graph.

Theorem 5.6.

Let d4d\geq 4 be a positive even integer. Let nn be an odd prime, and qq be an even power of an odd prime pp. If q>1.03n2d4/π2(d1)2q>1.03n^{2}d^{4}/\pi^{2}(d-1)^{2} and dqn1q1d\mid\frac{q^{n}-1}{q-1}, then the subfield 𝔽q\mathbb{F}_{q} forms a maximal clique in GP(qn,d)GP^{*}(q^{n},d).

The proof is similar to the proof of Theorem 2.20 and Theorem 5.3, except that we need to establish the following variant of Theorem 5.4 for GP(qn,d)GP^{*}(q^{n},d) (which is straightforward by following the proof of Theorem 5.4 step by step), and use Lemma 4.5 to get a lower bound on the character sum estimate.

Proposition 5.7.

Let d4d\geq 4 be a positive even integer. Let nn be an odd prime, and qq be an even power of an odd prime pp. If dqn1q1d\mid\frac{q^{n}-1}{q-1}, and if the subfield 𝔽q\mathbb{F}_{q} is not a maximal clique in GP(qn,d)GP^{*}(q^{n},d), then there exists h𝔽qn𝔽qh\in\mathbb{F}_{q^{n}}\setminus\mathbb{F}_{q}, such that 𝔽qh𝔽q\mathbb{F}_{q}\oplus h\mathbb{F}_{q} forms a clique in XX.

Acknowledgements

The authors thank Greg Martin, Daniel Panario, Zinovy Reichstein, József Solymosi, and Joshua Zahl for helpful discussions. The authors are also grateful to the referees for valuable comments, corrections, and suggestions. The research of the first author was supported by a postdoctoral research fellowship from the University of British Columbia. The research of the second author was supported in part by a Four Year Doctoral Fellowship from the University of British Columbia.

References