Hostname: page-component-788cddb947-jbkpb Total loading time: 0 Render date: 2024-10-15T10:16:58.252Z Has data issue: false hasContentIssue false

An energy decomposition theorem for matrices and related questions

Published online by Cambridge University Press:  15 May 2023

Ali Mohammadi
Affiliation:
School of Mathematics and Statistics, University of Sydney, Camperdown, NSW 2006, Australia e-mail: ali.mohammadi.np@gmail.com
Thang Pham*
Affiliation:
University of Science, Vietnam National University, Hanoi 100000, Vietnam
Yiting Wang
Affiliation:
Institute of Science and Technology Austria, Klosterneuburg 3400, Austria e-mail: yiting.wang@ist.ac.at
Rights & Permissions [Opens in a new window]

Abstract

Given $A\subseteq GL_2(\mathbb {F}_q)$, we prove that there exist disjoint subsets $B, C\subseteq A$ such that $A = B \sqcup C$ and their additive and multiplicative energies satisfying

$$\begin{align*}\max\{\,E_{+}(B),\, E_{\times}(C)\,\}\ll \frac{|A|^3}{M(|A|)}, \end{align*}$$

where

$$ \begin{align*} M(|A|) = \min\Bigg\{\,\frac{q^{4/3}}{|A|^{1/3}(\log|A|)^{2/3}},\, \frac{|A|^{4/5}}{q^{13/5}(\log|A|)^{27/10}}\,\Bigg\}. \end{align*} $$
We also study some related questions on moderate expanders over matrix rings, namely, for $A, B, C\subseteq GL_2(\mathbb {F}_q)$, we have
$$\begin{align*}|AB+C|, ~|(A+B)C|\gg q^4,\end{align*}$$
whenever $|A||B||C|\gg q^{10 + 1/2}$. These improve earlier results due to Karabulut, Koh, Pham, Shen, and Vinh ([2019], Expanding phenomena over matrix rings, $Forum Math.$, 31, 951–970).

Type
Article
Copyright
© The Author(s), 2023. Published by Cambridge University Press on behalf of The Canadian Mathematical Society

1 Introduction

Let ${\mathbb F}_q$ denote a finite field of order q and characteristic p, and let $M_2({\mathbb F}_q)$ be the set of two-by-two matrices with entries in ${\mathbb F}_q$ . We write $X\ll Y$ to mean $X\leq CY$ for some absolute constant $C>0$ and use $X\sim Y$ if $Y\ll X\ll Y$ .

Given subsets $A, B\subseteq M_2({\mathbb F}_q)$ , we define the sum set $A+B$ to be the set $\{a+b: (a, b)\in A\times B\}$ and similarly define the product set $AB$ . In this paper, we study various questions closely related to the sum-product problem over $M_2({\mathbb F}_q)$ , which is to determine nontrivial lower bounds on the quantity $\max \{\,|A+A|,\, |AA|\,\}$ , under natural conditions on sets $A\subseteq M_2({\mathbb F}_q)$ .

A result in this direction was proved by Karabulut et al. in [Reference Karabulut, Koh, Pham, Shen and Vinh4, Theorem 1.12], showing that if $A\subseteq M_2({\mathbb F}_q)$ satisfies $|A|\gg q^3$ then

(1.1) $$ \begin{align} \max\{\,|A+A|,\, |AA|\,\}\gg \min\left\{\,\frac{|A|^2}{q^{7/2}},\, q^{2}|A|^{1/2}\,\right\}. \end{align} $$

A closely related quantity is the additive energy $E_+(A, B)$ defined as the number of quadruples $(a, a^{\prime }, b, b^{\prime })\in A^2\times B^2$ such that $a + b = a^{\prime } + b^{\prime }$ . The multiplicative energy $E_{\times }(A, B)$ is defined in a similar manner. We also use, for example, $E_+(A) = E_+(A, A)$ . For $\lambda \in M_2({\mathbb F}_q)$ , we define the representation function $r_{A B}(\lambda ) = |\{\,(a, b)\in A\times B: a b = \lambda \,\}|$ . Note that $r_{A B}$ is supported on the set $AB$ and so we have the identities

(1.2) $$ \begin{align} \sum_{\lambda\in A B}r_{A B}(\lambda) = |A||B|\quad \text{and}\quad \sum_{\lambda\in A B}r_{A B}(\lambda)^2= E_{\times}(A, B). \end{align} $$

A standard application of the Cauchy–Schwarz inequality gives

(1.3) $$ \begin{align} |A+B|\ge \frac{|A|^2|B|^2}{E_+(A, B)},\; |AB|\ge \frac{|A|^2|B|^2}{E_{\times}(A, B)}. \end{align} $$

Thus, if either $E_+(A, B)$ or $E_{\times }(A, B)$ is small, then $\max (|A+B|, |AB|)$ is big. This motivates the study of energy estimates.

Balog and Wooley [Reference Balog and Wooley2] initiated the investigation into a type of energy variant of the sum-product problem by proving that given a finite set $A\subset \mathbb {R}$ , one may write $A= B \sqcup C$ such that $\max \{E_+(B), E_{\times }(C)\}\ll |A|^{3-\delta }(\log |A|)^{1-\delta }$ for $\delta =2/33$ . In the prime field setting, they also provided similar results, namely:

  1. (1) If $|A|\le p^{\frac {101}{161}}(\log p)^{\frac {71}{161}}$ , then

    $$\begin{align*}\max\{E_+(B), E_{\times}(C)\}\ll |A|^{3-\delta}(\log |A|)^{1-\delta/2}, ~~\delta=4/101.\end{align*}$$
  2. (2) If $|A|> p^{\frac {101}{161}}(\log p)^{\frac {71}{161}}$ , then

    $$\begin{align*}\max\{E_+(B), E_{\times}(C)\}\ll |A|^{3}(|A|/p)^{1/15}(\log |A|)^{14/15}.\end{align*}$$

These results have been improved by Rudnev, Shkredov, and Stevens in [Reference Rudnev, Shkredov and Stevens10]. In particular, they increased $\delta $ from $2/33$ to $1/4$ over the reals, and from $4/101$ to $1/5$ over prime fields. We note that this type of result has many applications in different areas, for instance, bounding exponential sums [Reference Mohammadi and Stevens5, Reference Roche-Newton, Shparlinski and Winterhof8, Reference Shkredov12Reference Swaenepoel and Winterhof15] or studying structures in Heisenberg groups [Reference Anh, Ham, Koh, Pham and Vinh1, Reference Hegyvári and Hennecart3].

The main goals of this paper are to study energy variants of the sum-product problem, and to obtain new exponents on two moderate expanding functions in the matrix ring $M_2({\mathbb F}_q)$ . While the results in [Reference Balog and Wooley2, Reference Rudnev, Shkredov and Stevens10] mainly relies on a number of earlier results on the sum-product problem or Rudnev’s point–plane incidence bound [Reference Rudnev9], our proofs rely on graph theoretic methods. It follows from our results in the next section that there exists a different phenomenon between problems over finite fields and over the matrix ring $M_2(\mathbb {F}_q)$ .

2 Main results

Our first theorem is on an energy decomposition of a set of matrices in $M_2(\mathbb {F}_q)$ .

Theorem 2.1 Given $A\subseteq GL_2({\mathbb F}_q)$ , there exist disjoint subsets $B, C\subseteq A$ such that $A = B \sqcup C$ and

$$\begin{align*}\max\{E_{+}(B), E_{\times}(C)\}\ll \frac{|A|^3}{M(|A|)}, \end{align*}$$

where

(2.1) $$ \begin{align} M(|A|) = \min\Bigg\{\,\frac{q^{4/3}}{|A|^{1/3}(\log|A|)^{2/3}},\, \frac{|A|^{4/5}}{q^{13/5}(\log|A|)^{27/10}}\,\Bigg\}. \end{align} $$

It follows from this theorem that for any set A of matrices in $M_2(\mathbb {F}_q)$ , we always can find a subset with either small additive energy or small multiplicative energy. By the Cauchy–Schwarz inequality, we have the following direct consequence on a sum-product estimate, namely, for $A\subseteq GL_2(\mathbb {F}_q)$ , we have

(2.2) $$ \begin{align} \max \left\lbrace |A+A|, |AA| \right\rbrace \gg |A|\cdot M(|A|). \end{align} $$

By a direct computation, one can check that this is better than the estimate (1.1) in the range $|A|\ll q^{3+5/8}/(\log |A|)^{1/2}$ .

In the next theorem, we show that the lower bound of (2.2) can be improved by a direct energy estimate.

Theorem 2.2 Let $A, B\subseteq M_2({\mathbb F}_q)$ and $C\subseteq GL_2({\mathbb F}_q)$ . Then

$$\begin{align*}E_+(A, B) \ll \frac{|A|^2|BC|^2}{q^4} + q^{13/2}\frac{|A||BC|}{|C|}. \end{align*}$$

Corollary 2.3 For $A\subseteq M_2({\mathbb F}_q)$ , with $|A|\gg q^{3}$ , we have

(2.3) $$ \begin{align} \max\{\,|A+A|,\, |AA|\,\}\gg \min\left\{\,\frac{|A|^2}{q^{13/4}},\, q^{4/3}|A|^{2/3}\,\right\}. \end{align} $$

In addition, if $|AA|\ll |A|$ and $|A|\gg q^{3+1/2}$ , then

(2.4) $$ \begin{align} |A+A|\gg q^4. \end{align} $$

If $|AA|\ll |A|$ and $|A|\gg q^{3+2/5}$ , then

(2.5) $$ \begin{align} |A+A+A|\gg q^4. \end{align} $$

We point out that the arguments of the proof of Corollary 2.3 could be used iteratively to give stronger results for expansion of k-fold sum sets $A + \cdots + A$ of sets $A\subseteq M_2(F_q)$ with $|AA|\ll |A|$ , as k gets larger.

We remark that the estimate (2.3) improves (1.1) in the range $|A|\ll q^{3+5/8}$ and is stronger than (2.2) in the range of $|A|\gg q^{13/4}$ . We also note that our assumption to get the estimate (2.4) is reasonable. For instance, let G be a subgroup of $\mathbb {F}_q^*$ , and let A be the set of matrices with determinants in G, then we have $|A|\sim q^3\cdot |G|$ and $|AA|=|A|$ .

It has been proved in [Reference Karabulut, Koh, Pham, Shen and Vinh4, Theorems 1.8 and 1.9] that for $A, B, C\subseteq M_2(\mathbb {F}_q)$ , if $|A||B||C|\ge q^{11}$ , then we have

$$\begin{align*}|AB+C|, ~|(A+B)C|\gg q^4.\end{align*}$$

In the following theorem, we provide improvements of these results.

Theorem 2.4 Let $A, B, C \subseteq M_2({\mathbb F}_q)$ , we have

$$\begin{align*}|AB+C|\gg \min \left\lbrace\, q^4,\, \frac{|A||B||C|}{q^{13/2}} \,\right\rbrace. \end{align*}$$

If $C\subseteq GL_2(\mathbb {F}_q)$ , the same conclusion holds for $(A+B)C$ , i.e.,

$$\begin{align*}|(A+B)C|\gg \min \left\lbrace\, q^4,\, \frac{|A||B||C|}{q^{13/2}} \,\right\rbrace. \end{align*}$$

In particular:

  1. (1) If $|A||B||C|\gg q^{10 + 1/2}$ , then $|AB + C|\gg q^4.$

  2. (2) If $|A||B||C|\gg q^{10 + 1/2}$ and $C\subseteq GL_2(\mathbb {F}_q)$ , then $|(A+B)C|\gg q^4.$

The condition $C\subseteq GL_2(\mathbb {F}_q)$ is necessary, since, for instance, one can take C being the set of matrices with zero determinant and $A=B=M_2(\mathbb {F}_q)$ , then $|(A+B)C|\sim q^3$ and $|A||B||C|\sim q^{11}$ .

We expect that the exponent $q^{10 + 1/2}$ , in the final conclusions of the above theorem, could be further improved to $q^{10}$ , which, as we shall demonstrate, is sharp. For $AB+C$ , let A and B be the set of lower triangular matrices in $M_2({\mathbb F}_q)$ and for arbitrary $0<\delta <1$ , let $X\subseteq {\mathbb F}_q$ be any set with $|X|= q^{1-\delta }$ , and let

$$\begin{align*}C = \left\{\,\begin{pmatrix} c_1 & c_2 \\ c_3 & c_4 \end{pmatrix}:c_1, c_3, c_4\in {\mathbb F}_q, c_2\in X\,\right\}. \end{align*}$$

Then $|A||B||C| = q^{10-\delta }$ and $|AB+C| = |C| = q^{4-\delta }$ .

For $(A+B)C$ , the construction is as follows: For arbitrary k, let $q=p^k$ , and let V be the set of elements corresponding to a $(k-1)$ -dimensional vector space over ${\mathbb F}_p$ in ${\mathbb F}_q$ . Thus, we have $|V| = p^{k-1} = q^{1-1/k}$ . Now, let

$$\begin{align*}A = B = \left\{\,\begin{pmatrix} x_1 & x_2 \\ x_3 & x_4 \end{pmatrix}:x_1, x_2\in V, x_3, x_4\in {\mathbb F}_q\,\right\}, \end{align*}$$

and

$$\begin{align*}C = \left\{\,\begin{pmatrix} c_1 & c_2 \\ c_3 & c_4 \end{pmatrix}:c_1, c_3\in {\mathbb F}_q, c_2, c_4\in {\mathbb F}_p\,\right\}. \end{align*}$$

Note that $A+B = A=B$ and so

$$\begin{align*}(A+B)C = AC = \left\{\,\begin{pmatrix} y_1 & y_2 \\ y_3 & y_4 \end{pmatrix}:y_1, y_3, y_4\in {\mathbb F}_q, y_2\in V\,\right\}, \end{align*}$$

where we have used that $V\cdot {\mathbb F}_p + V\cdot {\mathbb F}_p = V+V = V.$

Thus, $|A||B||C| = (q^2\cdot q^{2-2/k})^2 \cdot (q^2\cdot q^{2/k}) = q^{10-2/k}$ while $|(A+B)C| = q^{4-1/k}$ .

Also, we remark here that in the setting of finite fields, our approach and that of Karabulut et al. in [Reference Karabulut, Koh, Pham, Shen and Vinh4] imply the same result. Namely, for $A,B,C \subseteq \mathbb {F}_q$ , we have $|(A+B)C|, |AB+C| \gg q$ whenever $|A||B||C| \gg q^2$ . However, this is not true in the matrix ring. Let us briefly sketch the proof. For $\lambda \in AB +C$ , write

$$\begin{align*}t(\lambda) = |\{\,(a, b, c)\in A\times B \times C: a b + c = \lambda\,\}|. \end{align*}$$

By the Cauchy–Schwarz inequality, we have

$$\begin{align*}(|A||B||C|)^2 = \left(\sum_{\lambda\in AB+C} t(\lambda)\right)^2 \leq |AB +C|\sum_{\lambda\in AB+C} t(\lambda)^2. \end{align*}$$

Thus, the main task is to bound $\sum _{\lambda }t(\lambda )^2$ , i.e., the number of tuples $(a, b, c, a', b', c')\in (A\times B\times C)^2$ such that $ab+c=a'b'+c'$ . In [Reference Karabulut, Koh, Pham, Shen and Vinh4], instead of bounding $\sum _{\lambda }t(\lambda )^2$ , they bounded the number of quadruples $(a, b, c, \lambda )\in A\times B\times C\times (AB+C)$ such that $ab+c=\lambda $ . These two approaches imply the same lower bounds for $(A+B)C$ and $AB+C$ when $A, B, C\subset \mathbb {F}_q$ , but in the matrix rings, bounding $\sum _{\lambda }t(\lambda )^2$ is more effective. In other words, there exists a different phenomenon between problems over finite fields and over the matrix ring $M_2(\mathbb {F}_q)$ .

We now state a corollary of the above theorem with $C=AA$ which might be of independent interest.

Corollary 2.5 Let $A\subset M_2({\mathbb F}_q)$ with $|A|\gg q^{3+ 7/16}$ , then

$$\begin{align*}\max\{|AA(A+A)|, |AA+A+A|\}\gg q^4. \end{align*}$$

Let $A, B, C, D\subseteq M_2({\mathbb F}_q)$ , our last theorem is devoted for the solvability of the equation

(2.6) $$ \begin{align} x+y=zt,\quad (x, y, z, t)\in A\times B\times C\times D. \end{align} $$

Let $\mathcal {J}(A,B,C,D)$ denote the number of solutions to this equation.

One can check that by using Lemma 4.1 and Theorem 4.2 from [Reference Karabulut, Koh, Pham, Shen and Vinh4], one has

(2.7) $$ \begin{align} \left\vert \mathcal{J}(A,B,C,D)-\frac{|A||B||C||D|}{q^4} \right\vert\ll q^{7/2}(|A||B||C||D|)^{1/2}. \end{align} $$

Thus, when $|A||B||C||D|\gg q^{15}$ , then $\mathcal {J}(A,B,C,D)\sim \frac {|A||B||C||D|}{q^4}$ . We refer the interested reader to [Reference Sárközy11] for a result on this problem over finite fields. In our last theorem, we are interested in bounding $\mathcal {J}(A,B,C,D)$ from above when $|A||B||C||D|$ is smaller.

Theorem 2.6 Let $A, B, C, D\subseteq M_2({\mathbb F}_q)$ , and let $\mathcal {J}(A,B,C,D)$ denote the number of solutions to equation (2.6). Then, we have

$$ \begin{align*} \mathcal{J}(A,B,C,D) \ll \frac{|A||B|^{1/2}|C||D|}{q^2} + q^{13/4} (|A||B||C||D|)^{1/2}. \end{align*} $$

Assume $|A|=|B|=|C|=|D|$ , the upper bound of this theorem is stronger than that of (2.7) when $|A|\ll q^{11/3}$ .

2.1 Structure

The rest of this paper is structured as follows: In Section 3, we prove a preliminary lemma, which is one of the key ingredients in the proof of our energy decomposition theorem. Section 4 is devoted to proving Theorem 2.1. The proofs of Theorem 2.2 and Corollary 2.3 will be presented in Section 5. Section 6 contains proofs of Theorem 2.4, Corollary 2.5, and Theorem 2.6.

3 A preliminary lemma

Given sets $A, B, C, D, E, F\subseteq M_2({\mathbb F}_q)$ , let $\mathcal {I}(A, B, C, D, E, F)$ be the number of solutions

$$\begin{align*}(a, e, c, b, f, d)\in A\times B\times C\times D\times E\times F: \quad ab+ ef = c + d. \end{align*}$$

The main purpose of this section is to prove an estimate for $\mathcal {I}(A, B, C, D, E, F)$ , which is one of the key ingredients in the proof of Theorem 2.1.

Proposition 3.1 We have

$$\begin{align*}\bigg|\mathcal{I}(A, B, C, D, E, F) - \frac{|A||B||C||D||E||F|}{q^4}\bigg| \ll q^{13/2}\sqrt{|A||B||C||D||E||F|}\,. \end{align*}$$

To prove Proposition 3.1, we define the sum-product digraph $G=(V, E)$ with the vertex set $V=M_2(\mathbb {F}_q)\times M_2(\mathbb {F}_q)\times M_2(\mathbb {F}_q)$ , and there is a directed edge going from $(a, e, c)$ to $(b, f, d)$ if and only if $ab+ ef = c+d$ . The setting of this digraph is a generalization of that in [Reference Karabulut, Koh, Pham, Shen and Vinh4, Section 4.1]

Let G be a digraph on n vertices. Suppose that G is regular of degree d, i.e., the in-degree and out-degree of each vertex are equal to d. Let $m_G$ be the adjacency matrix of G, where $(m_G)_{ij}=1$ if and only if there is a directed edge from i to j. Let $\mu _1=d, \mu _2, \ldots , \mu _n$ be the eigenvalues of $m_G$ . Notice that these eigenvalues can be complex numbers, and for all $2\le i\le n$ , we have $|\mu _i|\le d$ . Define $\mu (G):=\max _{|\mu _i|\ne d}|\mu _i|$ . This value is referred to as the second largest eigenvalue of $m_G$ .

A digraph G is called an $(n, d, \mu )$ -digraph if G is a d-regular digraph of n vertices, and the second largest eigenvalue of $m_G$ is at most $\mu $ .

We recall the following lemma from [Reference Vu16] on the distribution of edges between two vertex sets on an $(n, d, \mu )$ -digraph.

Lemma 3.2 Let $G = (V, E)$ be an $(n, d, \mu )$ -digraph. For any two sets $B, C \subseteq V$ , the number of directed edges from B to C, denoted by $e(B, C)$ satisfies

$$\begin{align*}\left| e(B, C) - \frac{d}{n}|B | |C| \right| \leq \mu \sqrt{|B| |C|}\,.\end{align*}$$

With Lemma 3.2 in hand, to prove Proposition 3.1, it is enough to study properties of the sum-product digraph G.

Definition 3.1 Let $a, b \in M_2(\mathbb {F}_q)$ . We say they are equivalent, if whenever the ith row of a is not all-zero, neither is the ith row of b and vice versa, for $1\leq i \leq 2$ .

Proposition 3.3 The sum product graph G is a $(q^{12},\, q^8,\, c\cdot q^{13/2})$ -digraph, for some positive constant c.

Proof The number of vertices is $|M_2(\mathbb {F}_q)|^3 = q^{12}$ . Moreover, for each vertex $(a, e, c)$ , with each choice of $(b, f)$ , d is determined uniquely from $d = ab + ef - c$ . Thus, there are $|M_2(\mathbb {F}_q)|^2 = q^8$ directed edges going out of each vertex. The number of incoming directed edges can be argued in the same way. To conclude, the digraph G is $q^8$ -regular. Let $m_G$ denote the adjacency matrix of G. It remains to bound the magnitude of the second largest eigenvalue of the adjacency matrix of G, i.e., $\mu (m_G)$ .

In the next step, we are going to show that $m_G$ is a normal matrix, i.e., $m_G^Tm_G=m_Gm_G^T$ , where $m_G^T$ is the conjugate transpose of $m_G$ . For a normal matrix m, we know that if $\lambda $ is an eigenvalue of m, then $|\lambda |^2$ is an eigenvalue of $mm^T$ and $m^Tm$ . Thus, for a normal matrix m, it is enough to give an upper bound for the second largest eigenvalue of $mm^T$ or $m^Tm$ .

There is a simple way to check whenever G is normal. For any two vertices u and v, let $\mathcal {N}^+(u,v)$ be the set of vertices w such that $\overrightarrow {uw}, \overrightarrow {vw}$ are directed edges, and $\mathcal {N}^-(u, v)$ be the set of vertices $w'$ such that $\overrightarrow {w'u}, \overrightarrow {w'v}$ are directed edges. It is not hard to check that $m_G$ is normal if and only if $|\mathcal {N}^+(u,v)| = |\mathcal {N}^-(u,v)|$ for any two vertices u and v.

Given two vertices $(a, e, c)$ and $(a^{\prime }, e^{\prime }, c^{\prime })$ , where $(a, e, c) \neq (a^{\prime }, e^{\prime }, c^{\prime })$ , the number of $(x,y,z)$ that lies in the common outgoing neighborhood of both vertices is characterized by

For each pair $(x, y)$ satisfying this equation, z is determined uniquely. Thus, the problem is reduced to computing the number of such pairs $(x, y)$ .

For convenience, let $\bar {a} = a - a^{\prime }$ , $\bar {c} = c - c^{\prime }$ , and $\bar {e} = e - e^{\prime }$ . Also, let $t = \begin {pmatrix} \bar {a} & \bar {e}\end {pmatrix}_{2\times 4}$ . Then, the above relation is equivalent to

(3.1) $$ \begin{align} \begin{pmatrix} \bar{a} & \bar{e}\end{pmatrix} \begin{pmatrix} x \\ y\end{pmatrix} = t \begin{pmatrix} x \\ y\end{pmatrix}_{4\times 2}= \bar{c}\,. \end{align} $$

We now have the following cases:

  • (Case 1: $\operatorname {\mathrm {rank}}(t) = 0$ ) Note that in this case, we need $a = a^{\prime }$ , $c = c^{\prime }$ , and $e = e^{\prime }$ , which is a contradiction to our assumption that $(a, e, c) \neq (a^{\prime }, e^{\prime }, c^{\prime })$ . Thus, we simply exclude this case.

  • (Case 2: $\operatorname {\mathrm {rank}}(t) = 1$ ) As t is not an all-zero matrix, there is at least one nonzero row. Without loss of generality, assume it is the first row. Then,

    ${t = \begin {pmatrix} a_1 & a_2 & e_1 & e_2 \\ \alpha a_1 & \alpha a_2 & \alpha e_1 & \alpha e_2\end {pmatrix}}$ , where $(a_1, a_2, e_1, e_2) \neq \textbf {0}$ and $\alpha \in \mathbb {F}_q$ .

    • (Case 2.1: $\operatorname {\mathrm {rank}}(\bar {c}) = 2$ ) In this case, there is no solution, as $\operatorname {\mathrm {rank}}\left (t \begin {pmatrix} x \\ y\end {pmatrix}\right ) \leq \operatorname {\mathrm {rank}}(t) = 1$ but $\operatorname {\mathrm {rank}}(\bar {c}) = 2$ .

    • (Case 2.2: $\operatorname {\mathrm {rank}}(\bar {c}) = 1$ ) Let $x = \begin {pmatrix} x_1 & x_2 \\ x_3 & x_4 \end {pmatrix}$ , $y = \begin {pmatrix} y_1 & y_2 \\ y_3 & y_4 \end {pmatrix}$ . We discuss two sub-cases:

      (a) $\bar {c} = \begin {pmatrix} c_1 & c_2 \\ \alpha c_1 & \alpha c_2 \end {pmatrix}$ with the same factor $\alpha $ , where $(c_1, c_2) \neq (0, 0)$ .

      In this case, we have the following set of equations:

      $$ \begin{align*} \begin{cases} a_1 x_1 + a_2 x_3 + e_1 y_1 + e_2 y_3 = c_1, \\ a_1 x_2 + a_2 x_4 + e_1 y_2 + e_2 y_4 = c_2. \end{cases} \, \end{align*} $$
      Since we assume $(a_1, a_2, e_1, e_2) \neq \textbf {0}$ , without loss of generality, let $a_1 \neq 0$ . Then,
      $$ \begin{align*} \begin{cases} x_1 = (a_1)^{-1}(c_1 -a_2 x_3 -e_1 y_1 - e_2 y_3), \\ x_2 = (a_1)^{-1}(c_2 - a_2 x_4 - e_1 y_2 - e_2 y_4), \end{cases} \end{align*} $$
      which means that for each $(x_3, y_1, y_3)$ there is a unique $x_1$ and for each $(x_4, y_2, y_4)$ there is a unique $x_2$ . Thus, there are $q^6$ different $(x,y,z)$ solutions.

      (b) In all other sub-cases, there is no solution. If $\bar {c} = \begin {pmatrix} c_1 & c_2 \\ \beta c_1 & \beta c_2 \end {pmatrix}$ , where $\beta \neq \alpha $ and $(c_1, c_2) \neq (0, 0)$ , then we get the following two equations:

      $$ \begin{align*} \begin{cases} a_1 x_1 + a_2 x_3 + e_1 y_1 + e_2 y_3 = c_1, \\ \alpha a_1 x_1 + \alpha a_2 x_3 + \alpha e_1 y_1 + \alpha e_2 y_3 = \beta c_1, \end{cases}\, \end{align*} $$
      which obviously do not have any solution.

      Otherwise, $\bar {c} = \begin {pmatrix} \beta c_1 & \beta c_2 \\ c_1 & c_2 \end {pmatrix}$ , where $(c_1, c_2) \neq (0, 0)$ . Note that if $\alpha \neq 0$ , then $\beta \neq \alpha ^{-1}$ , because this case is covered in Case 2.2(a) implicitly. We get the following equations.

      $$ \begin{align*} \begin{cases} a_1 x_1 + a_2 x_3 + e_1 y_1 + e_2 y_3 = \beta c_1, \\ \alpha a_1 x_1 + \alpha a_2 x_3 + \alpha e_1 y_1 + \alpha e_2 y_3 = c_1, \end{cases}\, \end{align*} $$
      which obviously do not have any solution. Notice that $\alpha = 0$ or $\beta = 0$ corresponds to t and $\bar {c}$ not being equivalent.
    • (Case 2.3: $\operatorname {\mathrm {rank}}(\bar {c}) = 0$ ) This case is similar to the Case 2.2(a), except $c_1 = c_2 = 0$ . We have the following two equations:

      $$ \begin{align*} \begin{cases} a_1 x_1 + a_2 x_3 + e_1 y_1 + e_2 y_3 = 0, \\ a_1 x_2 + a_2 x_4 + e_1 y_2 + e_2 y_4 =0. \end{cases}\, \end{align*} $$
      Following the same analysis, we conclude there are $q^6$ solutions.

  • (Case 3: $\operatorname {\mathrm {rank}}(t) = 2$ ) In this case, we always have solutions, for any $\bar {c}$ .

    1. (Case 3.1: $\operatorname {\mathrm {rank}}(\bar {a}) = 2$ or $\operatorname {\mathrm {rank}}(\bar {e}) = 2$ ) In this case, let us look back on equation (3.1). If $\operatorname {\mathrm {rank}}(\bar {a}) = 2$ , then we can rewrite (3.1) as $\bar {a}x = \bar {c} - \bar {e} y$ . Observe that, for any $y\in M_2(\mathbb {F}_q)$ , there is a unique x. Thus, the number of solutions is $q^4$ . The case where $\operatorname {\mathrm {rank}}(\bar {e}) = 2$ is similar.

    2. (Case 3.2: $\operatorname {\mathrm {rank}}(\bar {a}) \leq 1$ and $\operatorname {\mathrm {rank}}(\bar {e}) \leq 1$ ) In this case, it is not hard to observe that t must be one of the following four types:

      1. (i) $\begin {pmatrix} a_1 & a_2 & e_1 & e_2 \\ \alpha a_1 & \alpha a_2 &\beta e_1 &\beta e_2\end {pmatrix}$ , where $(a_1, a_2), (e_1, e_2)\neq (0,0)$ , $\alpha \neq \beta $ , $(\alpha , \beta ) \neq (0,0)$ .

      2. (ii) $\begin {pmatrix} \alpha a_1 & \alpha a_2 &\beta e_1 &\beta e_2 \\ a_1 & a_2 & e_1 & e_2 \end {pmatrix}$ , where $(a_1, a_2), (e_1, e_2) \neq (0,0)$ , $\alpha \neq \beta $ , $(\alpha , \beta ) \neq (0,0)$ .

      3. (iii) $\begin {pmatrix} a_1 & a_2 &0 &0 \\ 0& 0 & e_1 & e_2 \end {pmatrix}$ , where $(a_1, a_2),(e_1, e_2) \neq (0,0)$ .

      4. (iv) $\begin {pmatrix} 0 & 0 & e_1 & e_2 \\ a_1& a_2 & 0 & 0 \end {pmatrix}$ , where $(a_1, a_2), (e_1, e_2)\neq (0,0)$ .

      Since $(i)$ and $(ii)$ are symmetric and so is $(iii)$ and $(iv)$ , we only argue for $(i)$ and $(iii)$ . For $(iii)$ , reusing notations from Case 2.2(a), we have

      $$ \begin{align*} \begin{cases} a_1 x_1 + a_2 x_3 = c_1, \\ a_1 x_2 + a_2 x_4 = c_2, \\ e_1 y_1 + e_2 y_3 = c_3, \\ e_1 y_2 + e_2 y_4 = c_4. \end{cases} \end{align*} $$

      As $(a_1, a_2) \neq (0,0)$ and $(e_1, e_2) \neq (0,0)$ , without loss of generality, we assume $a_1 \neq 0$ and $e_1 \neq 0$ . Then, it means for each $(x_3, x_4, y_3, y_4)$ there is a unique $(x_1, x_2, y_1, y_2)$ . Thus, the system has $q^4$ solutions.

      For $(i)$ , we have

      Again, assume $a_1 \neq 0$ and $e_1 \neq 0$ . Now, take , we get $(\alpha - \beta )(e_1 y_1 + e_2 y_3) = \alpha c_1 - c_3 $ . As $\alpha \neq \beta $ , this means $e_1 y_1 + e_2 y_3 = (\alpha - \beta )^{-1}(\alpha c_1 - c_3)$ . Thus, for each $y_3$ , there is a unique $y_1$ . Similarly, compute , and we get $ a_1 x_1 + a_2 x_3 = (\beta - \alpha )^{-1}(\beta c_1 - c_3)$ , which means that for each $x_3$ , we get a unique $x_1$ . We can do the same for and and conclude that there are $q^4$ solutions.

Observe that all cases are disjoint and they together enumerate all possible relations between vertices $(a, e, c)$ and $(a^{\prime }, e^{\prime }, c^{\prime })$ . We computed $\mathcal {N}^+((a, e, c), (a^{\prime }, e^{\prime }, c^{\prime }))$ above and the computation for $\mathcal {N}^-((a, e, c), (a^{\prime }, e^{\prime }, c^{\prime }))$ is the same. Thus, we know $m_G$ is normal. Note that each entry of $m_Gm_G^T$ can be interpreted as counting the number of common outgoing neighbors between two vertices. We can write $m_Gm_G^T$ as

$$ \begin{align*} \begin{aligned} m_Gm_G^T &= q^8 I + 0 E_{21} + q^6 E_{22a} + 0 E_{22b} + q^6 E_{23} + q^4 E_{31} + q^4 E_{32} \\ &=(q^8 - q^4)I + q^4 J - q^4 E_{21} + (q^6 - q^4)E_{22a}\\ &\qquad\qquad -q^4 E_{22b} + (q^6 - q^4) E_{23} + (q^4 - q^4) E_{31} + (q^4 - q^4) E_{32} \\ &= (q^8 - q^4)I + q^4 J - q^4 E_{21} + (q^6 - q^4)E_{22a}- q^4 E_{22b} + (q^6 - q^4) E_{23}\,, \end{aligned} \end{align*} $$

where I is the identity matrix, J is the all one matrix and $E_{ij}$ s are adjacency matrices, specifying which entries are involved. For example, for Case 2.3, all pairs $(a,e,c), (a^{\prime }, e^{\prime }, c^{\prime })$ with $c = c^{\prime }$ and $\operatorname {\mathrm {rank}}(t) = 1$ are involved. Thus, the $E_{23}$ is an adjacency matrix of size $q^{12}\times q^{12}$ (containing all triples $(a,e,c)$ ), with pairs of vertices satisfying this property marked 1 and all others marked 0.

Finally, observe that each subgraph defined by the corresponding adjacency matrix $E_{ij}$ is regular. This is due to the fact that the condition does not depend on specific value of $(a,e,c)$ . Starting from any vertex $(a,e,c)$ , we can get to all possible $\bar {a}, \bar {e},\bar {c}$ by subtracting the correct $(a^{\prime }, e^{\prime }, c^{\prime })$ . Thus, for each case, we get the same number of $(a^{\prime }, e^{\prime }, c^{\prime })$ that satisfies the condition.

Let $\kappa _{ij}$ be the maximum number of 1s in a row in $E_{ij}$ . Obviously, $\kappa _{ij}$ is an upper bound on the largest eigenvalue of $E_{ij}$ . It is not difficult to see that $\kappa _{21} \ll q^9$ , $\kappa _{22a} \ll q^7$ , $\kappa _{22b} \ll q^8$ and $\kappa _{23}\ll q^5$ . For example, in Case 2.1, we have $\operatorname {\mathrm {rank}}(t) = 1$ and $\operatorname {\mathrm {rank}}(\bar {c}) = 2$ . For a fixed $(a, e,c)$ , the former implies that there are $O(q^5)$ possibilities for $a^{\prime }$ and $e^{\prime }$ while the latter implies there are $O(q^4)$ possibilities for $c^{\prime }$ . Altogether, there are $O(q^9)$ possibilities for $(a^{\prime }, e^{\prime }, c^{\prime })$ in Case 2.1. Because the graph induced by $E_{21}$ is regular, we have $\kappa _{21}\ll q^9$ . Other cases can be deduced accordingly.

The rest follows from a routine computation: let $v_2$ be an eigenvector corresponding to $\mu (G)$ . Then, because G is regular and connected (easy to see, there is no isolated vertex), $v_2$ is orthogonal to the all 1 vector, which means $J\cdot v_2 = \mathbf {0}$ . We now have

$$ \begin{align*} \begin{aligned} \mu(m_G)^2 v_2 = m_Gm_G^T \cdot v_2 &= (q^8-q^4)I\cdot v_2 + (- q^4 E_{21} + (q^6 - q^4)E_{22a} - q^4 E_{22b} + (q^6 - q^4) E_{23}) \cdot v_2 \\ &= ((q^8 - q^4)- q^4 \kappa_{21} + (q^6 - q^4) \kappa_{22a} - q^4 \kappa_{22b} + (q^6 - q^4)\kappa_{23})\cdot v_2\\ &\ll q^{13} \cdot v_2 \,. \end{aligned} \end{align*} $$

Thus, $\mu (m_G)\ll q^{13/2}$ .

Proof of Proposition 3.1

It follows directly from Proposition 3.3 and Lemma 3.2 that

$$ \begin{align*} \left\vert \mathcal{I}(A, B, C, D, E, F)- \frac{1}{q^4}|A||B||C||D||E||F| \right\vert \ll q^{13/2} \sqrt{|A||B||C||D||E||F|}\,. \end{align*} $$

This completes the proof.

4 Proof of Theorem 2.1

To prove Theorem 2.1, we will also need several technical results. A proof of the following inequality may be found in [Reference Roche-Newton, Shparlinski and Winterhof8, Lemma 2.4].

Lemma 4.1 Let $V_1, \dots , V_k$ be subsets of an abelian group. Then

$$ \begin{align*} E_{+}\bigg(\bigsqcup_{i=1}^{k}V_i\bigg) \leq \bigg(\sum_{i=1}^k E_{+}(V_i)^{1/4}\bigg)^4. \end{align*} $$

The following lemma is taken from [Reference Mohammadi and Stevens5] and may also be extracted from [Reference Roche-Newton, Shparlinski and Winterhof8, Reference Rudnev, Shkredov and Stevens10]. Lemma 4.2 is slightly different to its analogs over commutative rings as highlighted by the duality of the inequalities (4.5) and (4.6).

Lemma 4.2 Let $X \subseteq GL_2(\mathbb {F}_q)$ . There exist sets $X_*\subset X$ , $D \subset XX$ , as well as numbers $\tau $ and $\kappa $ satisfying

(4.1) $$ \begin{align} \frac{E_{\times}(X)}{2|X|^2}\leq \tau \leq |X|, \end{align} $$
(4.2) $$ \begin{align} \frac{E_{\times}(X)}{\tau^2\cdot \log |X|}\ll |D| \ll (\log|X|)^6 \frac{|X_*|^4}{E_{\times}(X)}, \end{align} $$
(4.3) $$ \begin{align} |X_*|^2 \gg \frac{E_{\times}(X)}{|X| (\log|X|)^{7/2}}, \end{align} $$
(4.4) $$ \begin{align} \kappa \gg \frac{|D|\tau}{|X_*|(\log|X|)^2}, \end{align} $$

such that either

(4.5) $$ \begin{align} r_{DX^{-1}}(x) \geq \kappa \quad \text{for all} \quad x\in X_*, \end{align} $$

or

(4.6) $$ \begin{align} r_{X^{-1}D}(x) \geq \kappa \quad \text{for all} \quad x\in X_*. \end{align} $$

We need a dyadic pigeonhole argument, which can be found in [Reference Murphy and Petridis6, Lemma 18].

Lemma 4.3 For $\Omega \subseteq M_2(\mathbb {F}_q)$ , let $w,f: \Omega \rightarrow \mathbb {R}^+$ with $f(x) \leq M, \;\forall x\in \Omega $ . Let $W = \sum _{x\in \Omega } w(x)$ . If $\sum _{x\in \Omega } f(x) w(x) \geq K,$ then there exists a subset $D \subset \Omega $ and a number $\tau $ such that $\tau \leq f(x) < 2\tau $ for all $x\in D$ and $K/(2W) \leq \tau \leq M\,.$ Moreover,

$$ \begin{align*} \frac{K}{2 + 2\log_2M} \leq \sum_{x\in D} f(x)w(x) \leq 2\tau \sum_{x\in D}w(x) \leq \min\{2\tau W,\, 4\tau^2 |D|\}\,. \end{align*} $$

Proof of Lemma 4.2

We use the identities in (1.2) and apply Lemma 4.3, by taking $\Omega = X X, \, f = w = r_{XX},\, M = |X|,\, K = E_{\times }(X)$ , and $W = |X|^2$ , to find a set $D \subset XX$ and a number $\tau $ , satisfying (4.1), such that $D = \{\, \lambda \in XX: \tau \leq r_{XX}(\lambda ) < 2\tau \,\}\,$ and

(4.7) $$ \begin{align} \tau^2 |D| \gg E_{\times}(X) / \log |X|\,. \end{align} $$

Define $P_1 = \{\,(x,y)\in X\times X: xy \in D\,\}$ and $A_x = \{\,y: (x,y)\in P_1\,\}$ for $x \in X$ . By the definition of D, we know that $\tau |D| \leq |P_1| < 2\tau |D|$ . We can use Lemma 4.3 again with $\Omega = X, f(x) = |A_x|, w = 1, M = W = |X|$ , and $K = |P_1|$ to find a set $V\subset X$ and a number $\kappa _1$ such that $V = \{\,x\in X: \kappa _1 \leq |A_x| < 2\kappa _1\,\}\,$ and

(4.8) $$ \begin{align} |V|\kappa_1 \gg |P_1|/\log|X| \gg \tau |D| / \log |X|\,. \end{align} $$

Now, we split the analysis into two cases based on $|V|$ :

Case 1 ( $|V| \geq \kappa _1 (\log |X|)^{-1/2}$ ): In this case, we simply set $X_* = V$ and $\kappa = \kappa _1$ . For each $x\in V$ , there are at least $\kappa _1$ different y such that $x y \in D$ . Therefore, $r_{DX^{-1}}(x) \geq \kappa \;\forall x \in X_*$ .

Case 2 ( $|V| < \kappa _1 (\log |X|)^{-1/2}$ ): In this case, we find another pair $U, \kappa _2$ that satisfies $|U| \gg \kappa _2 (\log |X|)^{-1/2}$ and set $X_* = U$ and $\kappa = \kappa _2$ . Let $P_2 = \{\,(x,y)\in P_1: x\in V\,\}$ and $B_y = \{\,x: (x,y)\in P_2\,\}$ . By definition, we have $|P_2| \geq |V|\kappa _1$ . We apply Lemma 4.3 again, with $\Omega = X, f(y) = |B_y|, w = 1, K = |P_2|$ and $W = M =|X|$ to get $U \subset X$ and a number $\kappa _2$ such that $U = \{\,y\in X: \kappa _2 \leq |B_y| < 2\kappa _2\,\}\,$ and

(4.9) $$ \begin{align} |U|\kappa_2 \gg |P_2|/\log|X| \geq \kappa_1 |V| /\log|X| \,. \end{align} $$

Combining this inequality with the assumption of this case ( $\kappa _1 \geq |V|(\log |X|)^{1/2}$ ) and $|V| \geq \kappa _2$ , we conclude $|U| \gg \kappa _2 (\log |X|)^{-1/2}$ . We can then argue similarly as in Case 1 to conclude $r_{X^{-1}D}(x) \geq \kappa \;\forall x \in X_*$ .

Now, (4.4) follows from either of (4.8) or (4.9). To prove (4.3), we first note that in either of the cases above we have $|X_*|\gg \kappa (\log |X|)^{-1/2}$ . Then using the lower bound on $\kappa $ , (4.7) and (4.1), we have $|X_*|^2\gg |D|\tau (\log |X|)^{-5/2}\gg E_{\times }(X)/(|X|\log |X|)^{7/2}$ as required. Finally, to deduce the required upper bound on $|D|$ in (4.2) note that, as shown above, $|D|\tau \ll |X_*|^2(\log |X|)^{5/2}$ , which with (4.7) implies $|D|E_{\times }(X)(\log |X|)^{-1}\ll (|D|\tau )^2\ll |X_*|^4(\log |X|)^5$ .

Lemma 4.4 Let $X\subseteq GL_2({\mathbb F}_q)$ . Then there exists $X_*\subseteq X$ , with

$$\begin{align*}|X_*| \gg \frac{E_{\times}(X)^{1/2}}{|X|^{1/2}(\log |X|)^{7/4}}, \end{align*}$$

such that

(4.10) $$ \begin{align} E_{+}(X_*) \ll \frac{|X_*|^4|X|^6(\log|X|)^2}{q^4 E_{\times}(X)^2} + \frac{q^{13/2}|X_*|^3|X|(\log|X|)^5}{E_{\times}(X)}. \end{align} $$

Proof We apply Lemma 4.2 to the set X and henceforth assume its full statement, keeping the same notation. Without loss of generality, assume $r_{X^{-1}D}(x) \geq \kappa \;\forall x\in X_*$ . Thus,

$$ \begin{align*} E_{+}(X_*) &= |\{(x_1, x_2, x_3, x_4)\in X_*^4: x_1 + x_2 = x_3 + x_4\}|\\ &\leq \kappa^{-2}|\{(d_1, d_2, x_1, x_2, y_1, y_2)\in D^2\times X_*^2 \times X^2: x_1 + y_1^{-1} d_1 = x_2 + y_2^{-1} d_2 \}|\\ &= \kappa^{-2} \mathcal{I}(X^{-1}, D, -X_*, -X^{-1}, D, X_*). \end{align*} $$

Then applying Proposition 3.1 and (4.4), we obtain

$$ \begin{align*} E_{+}(X_*) &\ll \kappa^{-2}\cdot\bigg( \frac{(|D||X||X_*|)^{2}}{q^4} + q^{13/2}|D||X||X_*|\bigg)\\ &\ll\frac{|X_*|^4|X|^2(\log|X|)^2}{q^4\tau^2} + \frac{q^{13/2}|X_*|^3|X|(\log|X|)^4}{|D|\tau^2}\,. \end{align*} $$

Finally, applying (4.1) and (4.2), we obtain the required bound in (4.10) for $E_{+}(X_*)$ .

We are now ready to prove Theorem 2.1.

Proof of Theorem 2.1

We begin by describing an algorithm, which constructs two sequences of sets $A = S_1 \supseteq S_2 \supseteq \cdots \supseteq S_{k+1}$ and $\emptyset = T_0 \subseteq T_1 \subseteq \cdots \subseteq T_{k}$ such that $ S_{i} \sqcup T_{i-1} = A$ , for $i = 1, \dots , k+1$ .

Let $1\leq M\leq |A|$ be a parameter. At any step $i\geq 1$ , if $E_{\times }(S_i)\leq |A|^3/M$ the algorithm halts. Otherwise if

(4.11) $$ \begin{align} E_{\times}(S_i)> \frac{|A|^{3}}{M}, \end{align} $$

through a use of Lemma 4.4, with $X=S_i$ , we identify a set $V_i:= X_* \subseteq S_i$ , with

(4.12) $$ \begin{align} |V_i| \gg \frac{E_{\times}(S_i)^{1/2}}{|S_i|^{1/2}(\log|A|)^{7/4}}> \frac{|A|}{M^{1/2}(\log|A|)^{7/4}} \end{align} $$

and

(4.13) $$ \begin{align} E_{+}(V_i)\ll \frac{|V_i|^4|S_i|^6(\log|S_i|)^2}{q^4 E_{\times}(S_i)^2} + \frac{q^{13/2}|V_i|^3|S_i|(\log|S_i|)^5}{E_{\times}(S_i)}. \end{align} $$

We then set $S_{i+1} = S_i \setminus V_i$ , $T_{i+1} = T_i \sqcup V_i$ and repeat this process for the step $i+1$ . From (4.12), we deduce $|V_i|\gg |A|^{1/2}(\log |A|)^{-7/4}$ and so the cardinality of each $S_i$ monotonically decreases. This in turn implies that this process indeed terminates after a finite number of iterations k. We set $B = S_{k+1}$ and $C = T_k$ , noting that $A = B\sqcup C$ and that

(4.14) $$ \begin{align} E_{\times}(B) \leq \frac{|A|^{3}}{M}. \end{align} $$

We apply the inequalities (4.11), (4.12) and $|S_i| \leq |A|$ , to (4.13), to get

$$ \begin{align*} E_{+}(V_i) &\ll M^2|V_i|^4q^{-4}(\log|A|)^2 + M|A|^{-2}|V_i|^3 q^{13/2} (\log|A|)^5 \\ &\ll \big(M^2 q^{-4}(\log|A|)^2 + M^{3/2}|A|^{-3}q^{13/2}(\log|A|)^{27/4}\big)\cdot|V_i|^4. \end{align*} $$

Then, observing that

$$ \begin{align*} C = T_k = \bigsqcup_{i=1}^{k} V_i \subseteq A, \end{align*} $$

we use Lemma 4.1 to obtain

$$ \begin{align*} E_{+}(C) &\ll (M^2q^{-4}(\log|A|)^2 + M^{3/2}|A|^{-3}q^{13/2}(\log|A|)^{27/4}) \bigg(\sum_{i=1}^k|V_i|\bigg)^4 \\ &\leq M^{2}|A|^4q^{-4}(\log|A|)^{2} + M^{3/2}|A|q^{13/2}(\log|A|)^{27/4}. \end{align*} $$

Note that Lemma 4.1 is applicable because $M_2({\mathbb F}_q)$ is an abelian group under addition. Comparing this with (4.14), we see the choice $M = M(|A|)$ , given by (2.1) is optimal.

5 Proofs of Theorem 2.2 and Corollary 2.3

Proof of Theorem 2.2

We proceed similarly to the proof of [Reference Roche-Newton, Rudnev and Shkredov7, Theorem 6]. Note that

$$ \begin{align*} E_+(A, B) &= |C|^{-2}|\{\,(a, a^{\prime}, b, b^{\prime}, c, c^{\prime})\in A^2\times B^2\times C^2: a + b c c^{-1} = a^{\prime} + b^{\prime} c^{\prime} (c^{\prime})^{-1}\,\}|\\ &\leq |C|^{-2}|\{\,(a, a^{\prime}, s, s^{\prime}, c, c^{\prime})\in A^2\times (BC)^2\times (C^{-1})^2: a + s c = a^{\prime} + s^{\prime} c^{\prime}\,\}|. \end{align*} $$

The required result then follows by applying Proposition 3.1.

Proof of Corollary 2.3

Since $|A|\gg q^3$ , we may assume $A\subseteq GL_2({\mathbb F}_q)$ . We use Theorem 2.2, with $A=B=C$ and apply the lower bound on $E_+(A)$ given by (1.3) to obtain (2.3). To prove (2.4), we follow the same process and apply the assumption $|AA|\ll |A|$ , to obtain

(5.1) $$ \begin{align} |A+A| \gg \min\{\,q^4,\, |A|^3/q^{13/2}\,\}, \end{align} $$

which gives the required result.

To prove (2.5), we use Theorem 2.2, to get

$$\begin{align*}\frac{|A+A|^2|A|^2}{|A+A+A|}\leq E_{+}(A+A, A) \ll \frac{|A+A|^2|A|^2}{q^4} + q^{13/2}|A+A|. \end{align*}$$

Recalling (5.1), this rearranges to

$$\begin{align*}|A+A+A|\gg \min\left\{q^4, \frac{|A+A||A|^2}{q^{13/2}}\right\}\gg \min\left\{q^4, \frac{|A|^2}{q^{5/2}}, \frac{|A|^5}{q^{13}}\right\}. \end{align*}$$

The required result then easily follows.

6 Proofs of Theorem 2.4, Corollary 2.5, and Theorem 2.6

Proof of Theorem 2.4

For $\lambda \in AB +C$ , write

$$\begin{align*}t(\lambda) = |\{\,(a, b, c)\in A\times B \times C: a b + c = \lambda\,\}|. \end{align*}$$

By the Cauchy–Schwarz inequality, we have

$$\begin{align*}(|A||B||C|)^2 = \left(\sum_{\lambda\in AB+C} t(\lambda)\right)^2 \leq |AB +C|\sum_{\lambda\in AB+C} t(\lambda)^2. \end{align*}$$

Further noting that

$$\begin{align*}\sum_{\lambda\in AB+C} t(\lambda)^2 = \mathcal{I}(A, B, -C, -A, B, C). \end{align*}$$

We apply Proposition 3.1 to obtain

$$\begin{align*}|AB + C|\gg \min\left\{\,q^4,\, \frac{|A||B||C|}{q^{13/2}}\,\right\}. \end{align*}$$

This immediately implies the required result.

For the set $(A+B)C$ , as above we have

$$\begin{align*}|(A+B)C|\ge \frac{|A|^2|B|^2|C|^2}{|\{\,(a, b, c, a', b', c')\in (A\times B\times C)^2\colon (a+b)c=(a'+b')c'\,\}|}. \end{align*}$$

To estimate the denominator, we follow the argument in the proof of Proposition 3.1. In particular, we first define a graph G with the vertex set $V=M_2(\mathbb {F}_q)\times M_2(\mathbb {F}_q)\times M_2(\mathbb {F}_q)$ , and there is a direct edge going from $(a, e, c)$ to $(b, f, d)$ if $b a+e f=c+d$ . The only difference here compared to that graph in Section 3 is that we switch between $b a$ and $a b$ . By using a similar argument as in Section 3, we have this graph is a $(q^{12}, q^8,c q^{13/2})$ -digraph, where c is a positive constant.

To bound the denominator, we observe that the equation

$$ \begin{align*} (a+b)c=(a'+b')c' \end{align*} $$

gives us a direct edge from $(c, -b', -ac )$ to $(b, c', a'c')$ . So, let $U:=\{(c, -b', -ac)\colon a\in A, c\in C, b'\in B\}$ and $W=\{(b, c', a'c')\colon b\in B, c'\in C, a'\in A\}$ . Since $C\subseteq GL_2(\mathbb {F}_q)$ , we have $|U|=|W|=|A||B||C|$ . So applying Lemma 3.2, the number of edges from U to W is at most

$$\begin{align*}\frac{|A|^2|B|^2|C|^2}{q^4}+q^{13/2}|A||B||C|. \end{align*}$$

In other words,

$$\begin{align*}|\{\,(a, b, c, a', b', c')\in (A\times B\times C)^2\colon (a+b)c=(a'+b')c'\,\}|\ll \frac{|A|^2|B|^2|C|^2}{q^4}+q^{13/2}|A||B||C|, \end{align*}$$

and we get the desired estimate.

Proof of Corollary 2.5

It follows from Theorem 2.4 that

(6.1) $$ \begin{align} |AA + A + A| \gg q^4 \quad \text{if} \quad |A|^2 |A+A|\gg q^{10+1/2} \end{align} $$

and

(6.2) $$ \begin{align} |AA(A+A)| \gg q^4 \quad \text{if} \quad |A|^2 |AA|\gg q^{10+1/2}. \end{align} $$

Note that by Corollary 2.3, if $|A|\gg q^{3+7/16}$ , we have

$$\begin{align*}|A|^2\cdot \max\{\,|A+A|,\, |AA|\,\}\gg q^{4/3}|A|^{8/3} \gg q^{10 + 1/2}. \end{align*}$$

Hence, one of the conditions in (6.1) or (6.2) is satisfied, which in turn gives the required estimate.

Proof of Theorem 2.6

By the Cauchy–Schwarz inequality and Proposition 3.1, we have

$$ \begin{align*} \mathcal{J}(A,B,C,D) &= |\{\,(a, b, c, d)\in A\times B \times C\times D: a+b = c d\,\}|\\ &\leq |B|^{1/2}|\{\,(a, a^{\prime}, c, c^{\prime}, d, d^{\prime}) \in A^2 \times C^2 \times D^2: c d - a = c^{\prime} d^{\prime} - a^{\prime}\,\}|^{1/2}\\ &\ll \frac{|A||B|^{1/2}|C||D|}{q^2} + q^{13/4} (|A||B||C||D|)^{1/2}.\\[-42pt] \end{align*} $$

References

Anh, D. N. V., Ham, L. Q., Koh, D., Pham, T., and Vinh, L. A., On a theorem of Hegyvári and Hennecart . Pacific J. Math. 305(2020), no. 2, 407421.CrossRefGoogle Scholar
Balog, A. and Wooley, T. D., A low-energy decomposition theorem . Q. J. Math. 68(2017), 207226.Google Scholar
Hegyvári, N. and Hennecart, F., Expansion for cubes in the Heisenberg group . Forum Math. 30(2018), 227236.CrossRefGoogle Scholar
Karabulut, Y. D., Koh, D., Pham, T., Shen, C.-Y., and Vinh, L. A., Expanding phenomena over matrix rings . Forum Math. 31(2019), 951970.CrossRefGoogle Scholar
Mohammadi, A. and Stevens, S., Low-energy decomposition results over finite fields. Preprint, 2021. arXiv:2102.01655 Google Scholar
Murphy, B. and Petridis, G., Products of differences over arbitrary finite fields. Discrete Anal. 18(2018), 142.Google Scholar
Roche-Newton, O., Rudnev, M., and Shkredov, I. D., New sum-product type estimates over finite fields . Adv. Math. 293(2016), 589605.CrossRefGoogle Scholar
Roche-Newton, O., Shparlinski, I. E., and Winterhof, A., Analogues of the Balog–Wooley decomposition for subsets of finite fields and character sums with convolutions . Ann. Comb. 23(2019), 183205.CrossRefGoogle Scholar
Rudnev, M., On the number of incidences between points and planes in three dimensions . Combinatorica 38(2018), 219254.CrossRefGoogle Scholar
Rudnev, M., Shkredov, I. D., and Stevens, S., On the energy variant of the sum-product conjecture . Rev. Mat. Iberoam. 36(2020), no. 1, 207232.CrossRefGoogle Scholar
Sárközy, A., On sums and products of residues modulo $p$ $.$ Acta Arith. 118(2005), 403409.CrossRefGoogle Scholar
Shkredov, I. D., An application of the sum-product phenomenon to sets avoiding several linear equations. Sbornik: Mathematics, 209(4), 580.CrossRefGoogle Scholar
Shkredov, I. D., A short remark on the multiplicative energy of the spectrum . Math. Notes 105(2019), 449457.CrossRefGoogle Scholar
Shkredov, I. D., A remark on sets with small Wiener norm . In: Raigorodskii, A. and Rassias, M. T. (eds.), Trigonometric sums and their applications, Springer, Cham, 2020, pp. 261272.CrossRefGoogle Scholar
Swaenepoel, C. and Winterhof, A., Additive double character sums over some structured sets and applications . Acta Arith. 199(2021), 135143.CrossRefGoogle Scholar
Vu, V., Sum-product estimates via directed expanders . Math. Res. Lett. 15(2008), 375388.CrossRefGoogle Scholar