Hostname: page-component-848d4c4894-ttngx Total loading time: 0 Render date: 2024-05-23T15:46:19.451Z Has data issue: false hasContentIssue false

A solution to the Erdős–Sárközy–Sós problem on asymptotic Sidon bases of order 3

Published online by Cambridge University Press:  10 May 2024

Cédric Pilatte*
Affiliation:
Mathematical Institute, University of Oxford, Andrew Wiles Building, Oxford OX2 6GG, UK cedric.pilatte@maths.ox.ac.uk
Rights & Permissions [Opens in a new window]

Abstract

A set $S\subset {\mathbb {N}}$ is a Sidon set if all pairwise sums $s_1+s_2$ (for $s_1, s_2\in S$, $s_1\leqslant s_2$) are distinct. A set $S\subset {\mathbb {N}}$ is an asymptotic basis of order 3 if every sufficiently large integer $n$ can be written as the sum of three elements of $S$. In 1993, Erdős, Sárközy and Sós asked whether there exists a set $S$ with both properties. We answer this question in the affirmative. Our proof relies on a deep result of Sawin on the $\mathbb {F}_q[t]$-analogue of Montgomery's conjecture for convolutions of the von Mangoldt function.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BYCreative Common License - NC
This is an Open Access article, distributed under the terms of the Creative Commons Attribution-NonCommercial licence (https://creativecommons.org/licenses/by-nc/4.0), which permits non-commercial re-use, distribution, and reproduction in any medium, provided the original work is properly cited. Written permission must be obtained prior to any commercial use. Compositio Mathematica is © Foundation Compositio Mathematica.
Copyright
© 2024 The Author(s)

1. Introduction

A set $S$ of natural numbers is called a Sidon set, or a Sidon sequence, if all the sums $s_1+s_2$, for $s_1, s_2\in S$, $s_1\leqslant s_2$, are distinct. Sidon sequences are named after Simon Sidon, who in 1932 asked Erdős about the possible growth rate of such sequences. We write

\[ S(x) := \big| \{n\leqslant x: n\in S\} \big|. \]

Erdős observed that the greedy algorithm generates a Sidon sequence $S$ with $S(x) \gg x^{1/3}$. He also conjectured that, for every $\varepsilon >0$, there is a Sidon sequence $S$ with $S(x) \gg x^{1/2-\varepsilon }$ (see [Reference ErdősErd81]). In some sense, this would be best possible, as Erdős showed that any Sidon sequence $S$ satisfies

\[ S(x) \ll {x^{1/2}}{(\log x)^{-1/2}} \]

for infinitely many $x$ (see [Reference Halberstam and RothHR83, Theorem 8]). The lower bound was improved in 1981 by Ajtai, Komlós and Szemerédi [Reference Ajtai, Komlós and SzemerédiAKS81], who proved the existence of a Sidon sequence $S$ with

\[ S(x) \gg x^{1/3} \log x^{1/3} \]

using graph-theoretic tools. In a 1998 landmark paper, Ruzsa [Reference RuzsaRuz98] used the fact that the primes form a ‘multiplicative Sidon set’ to construct a Sidon sequence $S$ with

\[ S(x) \gg x^{\sqrt{2}-1-o(1)}. \]

This is still the best-known lower bound: improving the exponent $\sqrt {2}-1 \approx 0.4142\ldots$ would be a major achievement.

It is worth mentioning that much more is known about finite Sidon sets. In particular, the maximal size of a Sidon subset of $\{1,2,\ldots, n\}$ is $n^{1/2}+O(n^{1/4})$ (see Halberstam and Roth [Reference Halberstam and RothHR83, Chapter I, § 3] for detailed references).

Sidon sequences have become classical objects of interest in arithmetic combinatorics. It is natural to study the properties of the sumset $S+S := \{s_1+s_2:s_1, s_2\in S\}$, or more generally of the $k$th iterated sumset ${kS := \underbrace {S+S+\cdots +S}_{k \text { times}}}$, when $S$ is a Sidon sequence.

A subset $A\subset {\mathbb {N}}$ is called an asymptotic basis of order $k$ if $ {\mathbb {N}} \setminus kA$ is finite. In other words, $A$ is an asymptotic basis of order $k$ if every sufficiently large integer can be written as the sum of $k$ elements of $A$. Many well-known problems can be restated as follows: for a given set $A$, what is the smallest $k$ such that $A$ is an asymptotic basis of order $k$? If $A$ is the set of primes, this is essentially Goldbach's conjecture, whereas if $A$ is the set of perfect $d$th powers, this is a variant of Waring's problem.

In [Reference Erdős, Sárközy and SósESS94a, Problem 14] and [Reference Erdős, Sárközy and SósESS94b, Problem 8], Erdős, Sárközy and Sós asked the following question.

Problem 1.1 Does there exist a Sidon sequence $S$ which is an asymptotic basis of order 3?

This problem also appears in Erdős [Reference ErdősErd98, p. 212] and Sárközy [Reference SárközySár01, Problem 32].Footnote 1 In his paper [Reference ErdősErd98], Erdős describes Problem 1.1 as ‘an old problem of Nathanson and myself’.

It is easy to see that no Sidon sequence can be an asymptotic basis of order $2$. In fact, Erdős, Sárközy and Sós [Reference Erdős, Sárközy and SósESS95] proved that for any Sidon set $S\subset \{1, 2, \ldots, n\}$, the sumset $S+S$ cannot contain $\geqslant Cn^{1/2}$ consecutive integers, where $C$ is an absolute constant. Thus, the constant ‘$3$’ in Problem 1.1 can certainly not be improved.

In this article, we settle Problem 1.1 by proving the existence of a Sidon sequence $S$ which is also an asymptotic basis of order $3$.

A number of partial results have been previously obtained in the direction of Problem 1.1.

To make sense of the last row, we recall that a set $S\subset {\mathbb {N}}$ is an asymptotic basis of order $3+\varepsilon$ if, for every $\varepsilon >0$, every sufficiently large integer $n$ can be written as the sum of four elements of $S$, with one of them $\leqslant n^{\varepsilon }$. Our solution to Problem 1.1 also generalises another result of Cilleruelo [Reference CillerueloCil15, Theorem 1.2], which states that there is an asymptotic basis $S$ of order $3$ such that every integer $n$ can be represented in at most two different ways as a sum of two elements of $S$. Finally, Kiss and Sándor [Reference Kiss and SándorKS23] showed the existence of a Sidon sequence $S\subset {\mathbb {N}}$ whose $3$-fold sumset $S+S+S$ has lower asymptotic density $>0.0064$.Footnote 2

The results in [Reference CillerueloCil15, Reference KissKis10, Reference Kiss, Rozgonyi and SándorKRS14, Reference Kiss and SándorKS23] quoted above all use some variant of the probabilistic method. The starting point of all these approaches is to consider a random subset $S$ of $ {\mathbb {N}}$ where each $n\in {\mathbb {N}}$ is chosen to be in $S$, independently, with some probability $p(n)$: a typical choice is $p(n) = n^{-\gamma }$ for some $\gamma >0$. In order for $S$ to resemble a Sidon sequence (i.e. to be able to use the alteration method), it is necessary to have something roughly like $p(n) \ll n^{-2/3}$. In [Reference CillerueloCil15, Reference KissKis10, Reference Kiss, Rozgonyi and SándorKRS14], the authors choose $p(n) = n^{-\gamma }$ for some $\gamma <2/3$, while for [Reference Kiss and SándorKS23] they choose $p(n) = c n^{-2/3}$ for some optimised constant $c>0$. However, for such choices of $p(n)$, the set $S$ will almost surely not be an asymptotic basis of order $3$ (this follows from [Reference GoguelGog75, Satz B]). Hence, it seems that purely probabilistic approaches are of little use for addressing Problem 1.1.

In order to obtain an asymptotic basis of order $3$, it is more promising to start from the Sidon sequence of Ruzsa mentioned previously [Reference RuzsaRuz98], as it is much denser than the Sidon sets obtained by the probabilistic method.

The underlying idea behind Ruzsa's construction is to consider an infinite set of primes $\mathcal {P}$, and for each $p\in \mathcal {P}$, to define a natural number $n_p$ that ‘behaves like a logarithm of $p$’, so that an equality $n_{p_1} + n_{p_2} = n_{p_3} + n_{p_4}$ only holds if $p_1p_2 = p_3p_4$. Since the primes form a multiplicative Sidon set, we get $\{p_1, p_2\} = \{p_3, p_4\}$, so the set $S := \{n_p : p\in \mathcal {P}\}$ has the Sidon property. In his original paper, Ruzsa defined $n_p$ in terms of the binary expansion of the real number $\log _b{p}$ (for some $b\in {\mathbb {R}}^{>1}$) by concatenating certain blocks of digits of $\log _b{p}$ to obtain a natural number.

Cilleruelo [Reference CillerueloCil14] proposed a neat variant of Ruzsa's construction, where the real logarithm is replaced by a family of discrete logarithms. Fix an increasing sequence $\ell _1, \ell _2, \ldots$ of primes $\ell _i\not \in \mathcal {P}$ and choose for each $i$ a primitive root $\omega _i$ modulo $\ell _i$. Write $\log _{\ell _i, \omega _i}(p)$ for the unique $e\in \{0, 1, \ldots, \ell _i-2\}$ such that $\omega _i^e \equiv p \pmod {\ell _i}$. Cilleruelo defined $n_p$ by encoding many of these discrete logarithms $\log _{\ell _i, \omega _i}(p)$ into a single natural number, using a suitable base expansion. Now, an equality of the form $n_{p_1} + n_{p_2} = n_{p_3} + n_{p_4}$ implies that

\[ \log_{\ell_i, \omega_i}(p_1) + \log_{\ell_i, \omega_i}(p_2) = \log_{\ell_i, \omega_i}(p_3) + \log_{\ell_i, \omega_i}(p_4) \]

for many values of $i$, and thus $p_1p_2 \equiv p_3p_4 \pmod {\ell _i}$ for many $i$. With the appropriate quantification, this is only possible if $p_1p_2 = p_3p_4$, which forces $\{p_1, p_2\} = \{p_3, p_4\}$ as before.

Neither Ruzsa's nor Cilleruelo's construction constitutes an asymptotic basis of order $3$. There is an obvious obstruction, which is intrinsically tied to the way the information about $p$ is encoded in $n_p$. For the readers familiar with Ruzsa's construction, the reason is due to the presence of zeros between the blocks of digits of $\log _b{p}$ within $n_p$. These zeros cannot be removed, for this would destroy the Sidon property of $S$. Cilleruelo's construction suffers from a similar drawback.

Our point of departure is Cilleruelo's construction. To avoid the issue raised in the previous paragraph, we replace the zeros in the encoding of $n_p$ with random numbers chosen from a carefully designed set $A$ (see Lemma 3.1 and Definition 3.3). The remaining task is to prove that our modified Cilleruelo construction produces an asymptotic basis of order $3$. Eventually, this reduces to information about the distribution of products of three primes in arithmetic progressions.

However, the kind of information we would need is not remotely within reach of our current knowledge of prime numbers. We would need something along the lines of Montgomery's conjecture, a far-reaching generalisation of the generalised Riemann hypothesis. Montgomery's conjecture [Reference Iwaniec and KowalskiIK04, Eq. (17.5), p. 419] states that, for every $\varepsilon >0$,

(1)\begin{equation} \psi(x;q,a) := \sum_{\substack{n\leqslant x\\ n\equiv a\,(\mathrm{mod}\ q)}} \Lambda(n) = \frac{x}{\varphi(q)} + O_{\varepsilon}\big(q^{-1/2}x^{1/2+\varepsilon}\big) \end{equation}

uniformly in $x, q\geqslant 1$ and $(a, q)=1$. This implies the asymptotic formula $\psi (x;q,a) \sim x/\varphi (q)$ uniformly in the residue class and the modulus, provided that $q\leqslant x^{1-\varepsilon }$. We would need a similar (suitably normalised) statement, with the triple convolution $\Lambda *\Lambda *\Lambda$ in place of the von Mangoldt function $\Lambda$.

Cilleruelo [Reference CillerueloCil14, § 4] observed that his construction works equally well over $\mathbb {F}_q[t]$. Let $\mathcal {F}$ be a set of irreducible monic polynomials in $\mathbb {F}_q[t]$. Fix a sequence $(g_i)$ of irreducible monic polynomials not in $\mathcal {F}$, and for each $g_i$, choose a generator $\omega _i$ of $(\mathbb {F}_q[t]/(g_i))^{\times }$. For $f\in \mathcal {F}$, write $\log _{g_i, \omega _i}(f)$ for the unique integer $e\in \{0, 1, \ldots, q^{\deg g_i}-1\}$ such that $\omega _i^{e} \equiv f\pmod {g_i}$. As before, one can define an integer $n_f$ in terms of many such integers $\log _{g_i, \omega _i}(f)$, so that $S := \{n_f : f\in \mathcal {F}\}$ is a Sidon sequence. It is important to remember that $S$ is a Sidon sequence of integers, even if $\mathcal {F}$ is a subset of $\mathbb {F}_q[t]$.

The final and crucial ingredient to our proof is the remarkable work of Sawin [Reference SawinSaw21], who proved Montgomery-type results for ‘factorisation functions’ in $\mathbb {F}_q[t]$, a class of arithmetic functions that encompasses (the $\mathbb {F}_q[t]$-version of) the von Mangoldt function and its convolutions. For example, [Reference SawinSaw21, Theorem 1.2] implies that, for all $\varepsilon >0$, there is some $q_0(\varepsilon )$ such that, for all prime powers $q\geqslant q_0(\varepsilon )$, the analogue of Montgomery's conjecture (1) holds over $\mathbb {F}_q[t]$ (for squarefree moduli), i.e.

(2)\begin{equation} \sum_{\substack{f\in \mathbb{F}_q[t]\\ f\text{ monic of degree }n\\ f\equiv a\,(\mathrm{mod}\ g)}} \Lambda(f) = \frac{q^n}{\varphi(g)} + O\big((q^{\deg g})^{-1/2}(q^n)^{1/2+\varepsilon}\big) \end{equation}

uniformly in the modulus $g$ (squarefree monic polynomial) and in the residue class $a$, with ${(a,g)=1}$. Sawin's work relies on deep algebraic geometry methods, including sheaf cohomology, the characteristic cycle, vanishing cycles theory and perverse sheaves. We use a bound similar to (2) for the triple convolution $\Lambda *\Lambda *\Lambda$ (see Lemma 5.2). With this powerful result, we are able to prove that our modified Cilleruelo sequence is an asymptotic basis of order $3$, thereby solving Problem 1.1.

To conclude this introduction, we mention an application of our work to the study of $B_h[1]$ sequences. A set $S\subset {\mathbb {N}}$ is said to be $B_h[g]$ set if every positive integer can be written as the sum of $h$ terms from $S$ at most $g$ different ways. For any $h\geqslant 2$, with the probabilistic method, Kiss and Sándor successively showed the existence of a $B_h[1]$ set which is an asymptotic basis of order $2h+1$ (see [Reference Kiss and SándorKS21]), then $2h$ (see [Reference Kiss and SándorKS22]). A simple modification of our proof should establish the existence of a $B_h[1]$ set which is an asymptotic basis of order $2h-1$, for $h\geqslant 2$, thus answering a question of Kiss and Sándor [Reference Kiss and SándorKS21, Reference Kiss and SándorKS22].

2. Notation

The sumset of two sets $A,B\subset {\mathbb {Z}}$ is $A + B := \{a+b : a\in A, \, b\in B\}$. We write $f \ll g$ or $f = O(g)$ if $\big |f\big | \leqslant C g$ for some absolute constant $C>0$. If instead $C$ depends on a parameter $\theta$, we write $f \ll _{\theta } g$ or $f = O_{\theta }(g)$.

For $b\geqslant 1$ and $a\in {\mathbb {Z}}$, we write $\big [a \ \mathrm {mod}\ b \big ]$ for the unique integer $n$ satisfying $0\leqslant n < b$ and ${n\equiv a \pmod b}$.

Definition 2.1 (Generalised base)

Let $\boldsymbol {b} = (b_1, b_2, \ldots )$ be an infinite sequence of integers $\geqslant 2$. For any $n\geqslant 1$ and any $x_1, \ldots, x_n \in {\mathbb {Z}}$, we write

\[ \overline{x_n \ldots x_1}^{\boldsymbol{b}} := x_1 + x_2b_1 + x_3b_1b_2 + \cdots + x_n b_1b_2\cdots b_{n-1}. \]

Any $m\in {\mathbb {Z}}^{\geqslant 1}$ can be uniquely represented as $m = \overline {x_n \ldots x_1}^{\boldsymbol {b}}$ for some $0\leqslant x_i< b_i$ with $x_n\neq 0$. Recall, however, that the notation $\overline {x_n \ldots x_1}^{\boldsymbol {b}}$ is defined for arbitrary integers $x_i$: we do not always require $0\leqslant x_i< b_i$.

3. Construction

The following lemma is a slight strengthening of the statement that, for any sufficiently large $p$, there exists a set $A\subset {\mathbb {Z}}/p {\mathbb {Z}}$ such that $A$ and $A+A$ are disjoint, and moreover $A+A+A = {\mathbb {Z}}/p {\mathbb {Z}}$.

Lemma 3.1 For every sufficiently large prime $p$, there is a set $A \subset \{1, 2, \ldots, \lfloor p/2 \rfloor - 1\}$ such that:

  1. (i) the sets $A$ and $A+A+\{0, 1\}$ are disjoint;

  2. (ii) $A+A+A$ contains $p+2$ consecutive integers.

It turns out that we only need to use the existence of a single pair $(p, A)$ with these properties. Lemma 3.1 can be shown by the alteration method in probabilistic combinatorics. We provide a detailed proof in Appendix A.

For the remainder of this paper, we fix a pair $(p, A)$ satisfying the conclusion of Lemma 3.1. In particular, $p$ should be thought of as an absolute constant. Let $c = 0.35$, the point is that

\[ \frac13 < c < \frac{3-\sqrt{5}}{2}. \]

Let $C_0 > 100p$ be a large absolute constant that will be chosen later.

Definition 3.2 ($\mathcal {P}_{d}$, $g_i$, $\omega _i$, $\mathcal {F}_k$, $\mathcal {F}$)

Let $q\geqslant C_0$ be a prime (or a prime power). Let $\mathcal {P}_{d}$ be the set of irreducible monic polynomials $f\in \mathbb {F}_q[t]$ of degree $d$. Recall the standard formula of Gauss

(3)\begin{equation} |\mathcal{P}_{d}| = \frac{1}{d}\sum_{e \mid d} \mu \biggl(\frac{d}{e}\biggr) q^{e} = \frac{q^d}{d} + O(q^{d/2}). \end{equation}

For $i\geqslant 1$, let $g_i$ be an arbitrary element of $\mathcal {P}_{2i-1}$, and fix an arbitrary generator $\omega _i$ of $(\mathbb {F}_q[t]/(g_i))^{\times }$.

For $k\geqslant C_0$, let

\[ \mathcal{F}_k := \bigcup_{\substack{c k^2\leqslant 2i < c (k+1)^2}} \mathcal{P}_{2i}. \]

Let $\mathcal {F} := \bigcup _{k\geqslant C_0} \mathcal {F}_k$.

We work in the generalised base $\boldsymbol {b} = (b_1, b_2, \ldots )$ where

(4)\begin{equation} b_i = \begin{cases} q^{i}-1 & i\text{ odd},\\ p & i \text{ even.} \end{cases} \end{equation}

Definition 3.3 ($e_i$, $r_i$, $s$, $n_f$, $S$)

Let $k\geqslant C_0$ and let $f\in \mathcal {F}_k$. We associate to $f$ a positive integer $n_f$ as follows.

  • For $1\leqslant i\leqslant k$, let $e_i(f)$ be the unique integer such that $0\leqslant e_i(f)< b_{2i-1} = q^{2 i-1}-1$ and

    \[ \omega_i^{e_i(f)} \equiv f \pmod {g_i}. \]
  • Let $r_1(f), \ldots, r_k(f)$ be a sequence of independent and identically distributed (i.i.d.) random variables, each uniformly distributed on $A$ (in particular, $1\leqslant r_i(f) \leqslant p/2-1$).

  • Let $s(f)$ be an integer chosen uniformly at random in the set $\{1, 2, \ldots, q^{3k}\}$.

It is understood that the family of all random variables $r_i(f)$ and $s(f)$ (over all choices of $f$ and $i$) is independent.

We define $n_f$ to be the integer

\[ n_f := \overline{s\, r_k\, e_k\, \ldots\, r_2\, e_2\, r_1\, e_1}^{\boldsymbol{b}} \]

omitting the dependence of each digit on $f$ for conciseness. To be explicit,

\[ n_f = e_1(f) + r_1(f) b_1 + e_2(f) b_1b_2 + r_2(f) b_1 b_2 b_3 + \cdots + r_k(f) b_1b_2\cdots b_{2k-1} + s(f) b_1b_2\cdots b_{2k}, \]

with $b_i$ as in (4).Footnote 3

Finally, we define $S = \{n_f : f\in \mathcal {F}\}$.

We show that, with probability $1$, the elements of $S$ form a Sidon sequence and an asymptotic basis of order $3$.

Lemma 3.4

  1. (i) Let $f\in \mathcal {F}_k$. Then $n_f = q^{k^2+O(k)}$.

  2. (ii) Let $f_1, f_2\in \mathcal {F}$ be such that $n_{f_1} = n_{f_2}$. Then $f_1 = f_2$.

Proof. (i) On the one hand,

\[ n_f = \overline{s\, r_k\, e_k\, \ldots\, r_2\, e_2\, r_1\, e_1}^{\boldsymbol{b}} \geqslant \overline{1\, \underbrace{0\, 0\, \ldots\, 0\, 0}_{2k\text{ zeros}}}^{\boldsymbol{b}} = \prod_{i=1}^k p(q^{2i-1}-1) > \prod_{i=1}^k q^{2i-1} = q^{k^2}. \]

On the other hand,

\[ n_f = \overline{s\, r_k\, e_k\, \ldots\, r_2\, e_2\, r_1\, e_1}^{\boldsymbol{b}} \leqslant \overline{(q^{3k}+1)\, \underbrace{0\, 0\, \ldots\, 0\, 0}_{2k\text{ zeros}}}^{\boldsymbol{b}} = (q^{3k}+1) \prod_{i=1}^k p(q^{2i-1}-1) = q^{k^2+O(k)}, \]

as $p \leqslant q$. This shows that $n_f = q^{k^2+O(k)}$.

(ii) Let $n = n_{f_1} = n_{f_2}$. Suppose that $f_1\in \mathcal {F}_{k_1}$ and $f_2\in \mathcal {F}_{k_2}$. By part (i), we have

\[ q^{k_1^2 + O(k_1)} = n = q^{k_2^2 + O(k_2)}, \]

from which we infer that ${k_1 = k_2+O(1)}$. Let $k = \min (k_1, k_2)$. Since $n_{f_1} = n_{f_2} = n$, we also have

\[ \overline{r_k(f_1)\, e_k(f_1)\, \ldots\, r_1(f_1)\, e_1(f_1)}^{\boldsymbol{b}} = \big[n \ \mathrm{mod}\ b_1b_2\cdots b_{2k} \big] = \overline{r_k(f_2)\, e_k(f_2)\, \ldots\, r_1(f_2)\, e_1(f_2)}^{\boldsymbol{b}}. \]

The left- and right-hand side are two base-$\boldsymbol {b}$ representations of the same natural number, and since the $i$th digit is between $0$ and $b_i-1$ in both cases, we see that ${e_i(f_1) = e_i(f_2)}$ (and ${r_i(f_1) = r_i(f_2)}$) for all ${1\leqslant i\leqslant k}$. By the definition of $e_i$, this implies that ${f_1 \equiv f_2\pmod {g_i}}$. By the Chinese remainder theorem, we deduce that

(5)\begin{equation} {f_1 \equiv f_2\pmod {g_1\cdots g_k}}. \end{equation}

However, $\deg (f_1-f_2) \leqslant c k^2 + O(k)$ by definition of $\mathcal {F}_k$, whereas $\deg (g_1\cdots g_k) = k^2$. Therefore, $\deg (f_1-f_2) < \deg (g_1\cdots g_k)$ if $C_0$ is sufficiently large, and (5) can only hold if $f_1 = f_2$.

4. Sidon property

Lemma 4.1 Let $f_1, f_2, f_3, f_4\in \mathcal {F}$ be such that $n_{f_1} + n_{f_2} = n_{f_3} + n_{f_4}$. Then $\{f_1, f_2\} = \{f_3, f_4\}$.

Proof. Without loss of generality, suppose that $f_i\in \mathcal {F}_{k_i}$ for $i\in \{1,2,3,4\}$, where $k_1\geqslant k_2$ and ${k_3\geqslant k_4}$. By part (i) of Lemma 3.4, we have $q^{k_1^2+O(k_1)} = n_{f_1} + n_{f_2} = n_{f_3} + n_{f_4} = q^{k_3^2 + O(k_3)}$ and, thus, $k_1 = k_3 + O(1)$.

We claim that $k_2 = k_4 + O(1)$. Write the integer $n := n_{f_1} + n_{f_2} = n_{f_3} + n_{f_4}$ in base $\boldsymbol {b}$, as

\[ n = \overline{y_l\, x_l\, \ldots \, y_2\, x_2\, y_1\, x_1}^{\boldsymbol{b}}, \]

for some $0\leqslant x_i < q^{2i-1}-1$ and $0\leqslant y_1 < p$, where $x_l$ and $y_l$ are not both zero. Let us align the base-$\boldsymbol {b}$ digits of $n_{f_1}$, $n_{f_2}$ and $n$:

Since $n_{f_1} + n_{f_2} = n$ and $r_i(f_j) \leqslant p/2-1$ for all $i$ and $j$, we can easily express the digits $x_i$ and $y_i$ in terms of the digits of $n_{f_1}$ and $n_{f_2}$. The computation gets slightly more complicated when $i$ is close to $k_2$ (or larger than $k_1$) because $s(f_2)$ can be as large as $q^{3k_2}$, which exceeds $b_{2k_2+1} = q^{2k_2+1}-1$.

On the one hand, for $1\leqslant i\leqslant k_2$, we have:

  • $x_i = \big [e_i(f_1) + e_i(f_2) \ \mathrm {mod}\ q^{2i-1}-1 \big ]$; and

  • $y_i = r_i(f_1) + r_i(f_2) + \mathbf {1}_{e_i(f_1) + e_i(f_2) \geqslant q^{2i-1}-1}$.

On the other hand, for $k_2+3\leqslant i\leqslant k_1$ we have:Footnote 4

  • $x_i = e_i(f_1)$; and

  • $y_i = r_i(f_1)$.

In particular, for $1\leqslant i\leqslant k_2$ we have $y_i \in A+A+\{0, 1\}$, whereas for $k_2+3 \leqslant i\leqslant k_1$ we have $y_i\in A$. Let $i_0$ be the largest integer $\leqslant l$ such that $y_1, \ldots, y_{i_0} \in A+A+\{0, 1\}$. If $k_2+3>k_1$, then $i_0 = k_1+O(1) = k_2 + O(1)$. Otherwise, since $A$ and $A+A+\{0, 1\}$ are disjoint by Lemma 3.1(i), we have $k_2 \leqslant i_0 < k_2+3$. Hence, $i_0 = k_2 + O(1)$ in both cases.

Repeating the whole argument with $f_3$ and $f_4$ in place of $f_1$ and $f_2$, we obtain that $i_0 = k_4 + O(1)$, whence $k_2 = k_4 + O(1)$ as claimed.

Our previous computations have shown that

\[ x_i = \big[e_i(f_1) + e_i(f_2) \ \mathrm{mod}\ q^{2i-1}-1 \big] = \big[e_i(f_3) + e_i(f_4) \ \mathrm{mod}\ q^{2i-1}-1 \big] \]

for $1\leqslant i \leqslant \min (k_2, k_4)$, and

\[ x_i = e_i(f_1) = e_i(f_3) \]

for $\max (k_2, k_4)+3 \leqslant i\leqslant \min (k_1, k_3)$.

By the definition of $e_i(f)$ and the Chinese remainder theorem, this implies that

(6)\begin{equation} f_1f_2 \equiv f_3f_4 \biggl( \mathrm{mod}{\prod_{i=1}^{\min(k_2, k_4)} g_i}\biggr) \end{equation}

and

(7)\begin{equation} f_1 \equiv f_3 \biggl( \mathrm{mod}{\prod_{i = \max(k_2, k_4)+3}^{\min(k_1, k_3)} g_i}\biggr). \end{equation}

Suppose first that $f_1 = f_3$. Then $n_{f_2} = n - n_{f_1} = n - n_{f_3} = n_{f_4}$ and, hence, $f_2 = f_4$ by Lemma 3.4(ii), which is what we needed to prove.

If $f_1f_2 = f_3f_4$, we immediately conclude that $\{f_1, f_2\} = \{f_3, f_4\}$, since all $f_i$ are irreducible monic polynomials.

Hence, we may suppose that $f_1 \neq f_3$ and $f_1f_2 \neq f_3f_4$. By (7) we must have

\[ c k_1^2+O(k_1) \geqslant \deg(f_1-f_3) \geqslant \deg \bigg( \prod_{i = k_2+O(1)}^{k_1-O(1)} g_i \bigg) = k_1^2-k_2^2 - O(k_1). \]

Similarly, by (6), we have

\[ c k_1^2 + c k_2^2 + O(k_1) \geqslant \deg(f_1f_2 - f_3f_4) \geqslant \deg \bigg( \prod_{i=1}^{k_2 - O(1)} g_i \bigg) = k_2^2 - O(k_2). \]

Therefore, we have

\[ \begin{cases} k_1^2 - k_2^2 \leqslant c k_1^2 + O(k_1),\\ k_2^2 \leqslant c k_1^2 + c k_2^2 + O(k_1). \end{cases} \]

Hence,

\[ (1-c) k_1^2 - O(k_1) \leqslant k_2^2 \leqslant \frac{c}{1-c} k_1^2 + O(k_1), \]

which is impossible since $1-c > {c}/({1-c})$, if $C_0$ is sufficiently large.

5. Asymptotic basis of order 3

Lemma 5.1 Let $m \geqslant 3$ be an integer. We can write $m = \overline {z\, y_k\, x_k\, \ldots \, y_2\, x_2\, y_1\, x_1}^{\boldsymbol {b}}$, for some $k\geqslant 0$ and some integers $x_i, y_i, z$ satisfying:

  • $0\leqslant x_i < q^{2i-1}-1$ for all $1\leqslant i\leqslant k$;

  • $0\leqslant y_i < 2p$ for all $1\leqslant i\leqslant k$;

  • $\{y_i-2, y_i-1, y_i\} \subset A+A+A$ for all $1\leqslant i\leqslant k$;

  • $3\leqslant z \leqslant 6p q^{2k+1}$.

Proof. We proceed in a similar way to the standard algorithm that generates the base-$b$ expansion of an integer. Let us inductively define a sequence $(m_i)$ of integers $\geqslant 3$. Let $m_1 := m$. Suppose we have defined $m_1, \ldots, m_{2l-1}$ for some $l\geqslant 1$.

If $m_{2l-1} > 6p q^{2l-1}$, we define

\[ x_l := \big[m_{2l-1} \ \mathrm{mod}\ q^{2l-1}-1 \big] \]

and we set $m_{2l} := ({m_{2l-1} - x_l})/({q^{2l-1}-1})$.

Let $0\leqslant y_l < 2p$ be an integer with ${y_l \equiv m_{2l} \pmod p}$ and such that

\[ {\{y_l-2, y_l-1, y_l\} \subset A+A+A}. \]

Such an integer exists by Lemma 3.1(ii) (note that $A+A+A \subset [0, 3p/2]$). We define $m_{2l+1} := (m_{2l} - y_l)/{p}$. Since $m_{2l-1} > 6p q^{2l-1}$, we have

\[ m_{2l+1} \geqslant \frac{m_{2l}}{p} - 2 \geqslant \frac{m_{2l-1}}{p q^{2l-1}} - 3 \geqslant 3. \]

We have thus constructed $x_l$, $y_l$, $m_{2l}$ and $m_{2l+1}$.

Otherwise, if $m_{2l-1} \leqslant 6p q^{2l-1}$, so we can define $z := m_{2l-1}$ and stop the construction (thus, $k=l-1$).

The procedure must stop since $(m_i)$ is decreasing. When the procedure terminates, we end up with integers $x_1, \ldots, x_k$, $y_1, \ldots, y_k$ and $z$ that satisfy the required properties, by construction.

Lemma 5.2 Let $d\geqslant 1$ and let $g\in \mathbb {F}_q[t]$ be a squarefree monic polynomial of degree

\[ 2\leqslant \deg(g) \leqslant 3\theta d \]

for some $0 < \theta < 1$. Let $a \in (\mathbb {F}_q[t]/(g))^{\times }$. Then, with $\phi (g) := \big |(\mathbb {F}_q[t]/(g))^{\times }\big |$, we have

\[ \biggl| \sum_{\substack{f_1, f_2, f_3 \in \mathcal{P}_{d}\\ f_i~\mathrm{distinct}\\ f_1f_2f_3 \equiv a\ (\mathrm{mod}(g))}} \hspace{-0.3cm} 1 - \frac{1}{\phi(g)} { \big|\mathcal{P}_{d}\big|\choose 3} \biggr| \ll e^{O_{\theta}(d)} q^{({3d - \deg(g)})/{2}}. \]

Proof. We use [Reference SawinSaw21, Lemma 9.14] with $\omega = 3$, $n_1 = n_2 = n_3 = d$, $n = 3d$ and $h = 1$ (meaning that $c=0$). The author writes $\mathbb {F}_q[t]^{+}$ and $\mathcal {M}_n$ for the set of monic polynomials in $\mathbb {F}_q[t]$ and the set of monic polynomials of degree $n$ in $\mathbb {F}_q[t]$, respectively. The definition of $H^{h}_{n_1, \ldots, n_{\omega }}(f)$ can be found at the bottom of [Reference SawinSaw21, p. 89]. In particular,

\[ H^{1}_{d, d, d}(f) = d^3 \sum_{\substack{f_1, f_2, f_3 \in \mathcal{P}_{d}\\ f_i \text{ distinct}\\ f_1f_2f_3 = f}} 1, \]

and, thus,

\[ \sum_{f\in \mathcal{M}_{3d}}H^{1}_{d, d, d}(f) = d^3 { \big|\mathcal{P}_{d}\big|\choose 3}. \]

Since $c = 0$, the conclusion of [Reference SawinSaw21, Lemma 9.14] simplifies to

\[ d^3 \biggl| \sum_{\substack{f_1, f_2, f_3 \in \mathcal{P}_{d}\\ f_i \text{ distinct}\\ f_1f_2f_3 \equiv a (\mathrm{mod}(g))}} \hspace{-0.3cm} 1 - \frac{1}{\phi(g)} { \big|\mathcal{P}_{d}\big|\choose 3} \biggr| \ll \frac{(3d)!}{(d!)^3} (O_{\theta} (q))^{({3d - \deg(g)})/{2}}. \]

By Stirling's formula, ${(3d)!}/{(d!)^3} \ll e^{O(d)}$, and we obtain Lemma 5.2.

Lemma 5.3 Let $m\geqslant q^{(C_0+2)^2}$. Then $\mathbb {P}\big (m\not \in S+S+S\big ) \leqslant \exp (-m^{3c-1 - o(1)})$ as $m\to +\infty$.Footnote 5

Proof. Let $m \geqslant q^{(C_0+2)^2}$. We write $m = \overline {z\, y_k\, x_k\, \ldots \, y_2\, x_2\, y_1\, x_1}^{\boldsymbol {b}}$ with $x_i$, $y_i$ and z satisfying the conclusion of Lemma 5.1. By a computation similar to that in Lemma 3.4(i), we have

\[ q^{k^2} \leqslant m \leqslant q^{(k+2)^2}. \]

In particular, $k \geqslant C_0$ and $m = q^{k^2+O(k)}$.

We show that $m$ is very likely to be expressible as $n_{f_1}+n_{f_2}+n_{f_3}$ for some $f_i\in \mathcal {F}_k$.

We first focus on the digits $x_i$. By the Chinese remainder theorem, there is some $a\in \mathbb {F}_q[t]$ such that

\[ a \equiv \omega_i^{x_i} \pmod{g_i} \]

for all $1\leqslant i\leqslant k$. Let $g = \prod _{1\leqslant i\leqslant k} g_i$; it is a squarefree polynomial of degree $\deg (g) = k^2$. Let $d$ be an even integer with $c k^2 \leqslant d < c(k+1)^2$. Let $\mathcal {E}$ be the set of all triples $(f_1, f_2, f_3) \in (\mathcal {P}_{d})^3$ of distinct polynomials such that $f_1f_2f_3 \equiv a \pmod {g}$. By Lemma 5.2 with $\theta = k^2/(3d) \leqslant 1/(3c) < 1$, we have

(8)\begin{align} \left|\big|\mathcal{E}\big| - \frac{1}{\phi(g)} { \big|\mathcal{P}_{d}\big|\choose 3}\right| \ll e^{O(d)} q^{({3d-k^2})/{2}}. \end{align}

Using the fact that $\phi (g) = \prod _{i=1}^k (q^{2i-1}-1) = q^{k^2-O(k)}$ and the bound (3) on the size of $\mathcal {P}_{d}$, we can simplify (8) to get

\[ \big|\mathcal{E}\big| = q^{3d-k^2+O(k+\log d)} + O\bigl(q^{({3d-k^2})/{2}+O({d}/{\log q})}\bigr) = q^{(3c - 1)k^2+O(k)} + O\bigl(q^{(({3c -1})/{2}+O({1}/{\log q}))k^2}\bigr), \]

using $d = c k^2 + O(k)$ for the last step. If $C_0$ is sufficiently large, the error term is negligible and we obtain

(9)\begin{equation} \big|\mathcal{E}\big| = q^{(3c - 1)k^2+O(k)}. \end{equation}

Let us estimate, for a fixed triple $(f_1, f_2, f_3)\in \mathcal {E}$, the probability that $n_{f_1}+n_{f_2}+n_{f_3} = m$. By the definition of $n_{f_i}$, we can write

\[ n_{f_1}+n_{f_2}+n_{f_3} = \overline{s'\, r_k'\, e_k'\, \ldots\, r_2'\, e_2'\, r_1'\, e_1'}^{\boldsymbol{b}} \]

where $e_i'$, $r_i'$ and $s'$ are defined by:

  • $e_i' = \big [e_i(f_1) + e_i(f_2) + e_i(f_3) \ \mathrm {mod}\ q^{2i-1}-1 \big ]$ for $1\leqslant i\leqslant k$;

  • $r_i' = r_i(f_1) + r_i(f_2) + r_i(f_3) + \kappa _i$ for $1\leqslant i\leqslant k$, with

    \[ \kappa_i := \lfloor \frac{e_i(f_1) + e_i(f_2) + e_i(f_3)}{q^{2i-1}-1} \rfloor \in \{0, 1, 2\}; \]
  • $s' = s(f_1) + s(f_2) + s(f_3)$.

The assumption that $(f_1, f_2, f_3)\in \mathcal {E}$ ensures that $e_i' = x_i$ for all $1\leqslant i \leqslant k$. Indeed, both $e_i'$ and $x_i$ are between $0$ and $q^{2i-1}-2$, and

\[ \omega_i^{e_i'} \equiv \omega_i^{e_i(f_1) + e_i(f_2) + e_i(f_3)} \equiv f_1f_2f_3 \equiv a \equiv \omega_i^{x_i} \pmod{g_i}. \]

Hence,

\[ \mathbb{P}\big(n_{f_1}+n_{f_2}+n_{f_3} = m\big) \geqslant \mathbb{P}\big(r_1' = y_1,\,r_2' = y_2,\,\ldots,r_k' = y_k,\, s' = z\big) = \mathbb{P}\big(s' = z\big) \prod_{1\leqslant i\leqslant k}\mathbb{P}\big(r_i' = y_i\big), \]

where we used the independence of the family of all random variables $r_i(f_j)$ and $s(f_j)$ in the last step.

Recall that $s'$ is the sum of three independent random numbers chosen uniformly in $\{1, 2, \ldots, q^{3k}\}$, so $s'$ can be equal to any integer in the set $\{3, 4, \ldots, 3q^{3k}\}$, each of which occurs with probability at least $q^{-9k}$. Since $z$ is a fixed element of $\{3, 4, \ldots, 6p q^{2k+1}\}$ and $6p q^{2k+1} \leqslant 3q^{3k}$, we get

\[ \mathbb{P}\big(s' = z\big) \geqslant q^{-9k}. \]

Similarly, $r_i'$ is the sum of three elements of $A$ chosen uniformly and independently at random, plus a fixed carry $\kappa _i \in \{0, 1, 2\}$. Since $\{y_i-2, y_i-1, y_i\} \subset A+A+A$ by Lemma 5.1, there is at least one triple $(a_1, a_2, a_3)\in A^3$ such that $a_1+a_2+a_3+\kappa _i = y_i$. Thus, we see that $\mathbb {P}\big (r_i' = y_i\big ) \geqslant |A|^{-3} \geqslant p^{-3}$. We have thus shown that

\[ \mathbb{P}\big(n_{f_1}+n_{f_2}+n_{f_3} = m\big) \geqslant q^{-9k}p^{-3k} \geqslant q^{-O(k)}. \]

To conclude the argument, we wish to restrict to a large subset $\mathcal {E}_0$ of $\mathcal {E}$ such that no polynomial appears in more than one triple of $\mathcal {E}_0$. Observe that, if two triples $(f, f_1, f_2)$ and $(f, f_3, f_4)$ are both in $\mathcal {E}$, then $\{f_1, f_2\} = \{f_3, f_4\}$. Indeed, two such triples satisfy $f_1f_2 \equiv f^{-1}a \equiv f_3f_4 \pmod {g}$, which implies that $f_1f_2 = f_3f_4$ in $\mathbb {F}_q[t]$ since

\[ \deg(f_1f_2 - f_3f_4) \leqslant 2d < k^2 = \deg(g), \]

and thus $\{f_1, f_2\} = \{f_3, f_4\}$. In addition, if $(f_1, f_2, f_3)\in \mathcal {E}$ then so does $(f_{\sigma (1)}, f_{\sigma (2)}, f_{\sigma (3)})$ for every permutation $\sigma$. Therefore, any given polynomial $f\in \mathcal {P}_{d}$ appears at most once in a triple of $\mathcal {E}$, up to permutations. This shows that there is a suitable subset $\mathcal {E}_0$ with $\big |\mathcal {E}_0\big | = \tfrac 16 \big |\mathcal {E}\big |$.

Therefore,

\begin{align*} \mathbb{P}\big(m\not\in S+S+S\big) &\leqslant \mathbb{P} \biggl(\bigcap_{(f_1, f_2, f_3)\in \mathcal{E}_0} \{n_{f_1}+n_{f_2}+n_{f_3} \neq m\} \biggr)\\ &= \prod_{(f_1, f_2, f_3)\in \mathcal{E}_0} \mathbb{P}\big(n_{f_1}+n_{f_2}+n_{f_3} \neq m\big) \leqslant \big(1-q^{-O(k)}\big)^{|\mathcal{E}_0|}, \end{align*}

using the independence of the family of random variables $(n_{f_1}+n_{f_2}+n_{f_3})_{(f_1, f_2, f_3)\in \mathcal {E}_0}$. By (9) and the simple inequality $1-x\leqslant \exp (-x)$, we get

\[ \mathbb{P}\big(m\not\in S+S+S\big) \leqslant \exp \big({-}q^{-O(k)} q^{(3c - 1)k^2-O(k)} \big) \leqslant \exp \big({-}m^{(3c-1+o(1))} \big), \]

as $m = q^{k^2+O(k)}$.

Theorem 5.4 With probability $1$, the elements of $S$ form a Sidon sequence and an asymptotic basis of order $3$.

Proof. By Lemma 4.1, $S$ always has the Sidon property. By Lemma 5.3,

\[ \mathbb{P}\big(m\not\in S+S+S\big) \leqslant \exp(-m^{3c-1 - o(1)}) \ll m^{-2} \]

for sufficiently large $m$, so

\[ \mathbb{P}\big(\big|{\mathbb{N}} \setminus S+S+S\big| = +\infty\big) = 0 \]

by the Borel–Cantelli lemma. We conclude that $S$ is almost surely an asymptotic basis of order $3$.

Acknowledgements

I would like to thank my advisors, Ben Green and James Maynard, for their expert guidance and continuing encouragement. I am very grateful to Oliver Riordan for pointing out an error in the appendix of an earlier version of this paper, and for suggesting the method used here to fix it. The author is supported by the Oxford Mathematical Institute and a Saven European Scholarship.

Conflicts of Interest

None.

Appendix A. Auxiliary set

In this section, we use the notation $f*g$ for the additive convolution

\[ f*g (n) = (f*g) (n) := \sum_{\substack{a, b\in {\mathbb{Z}}\\ a+b=n}} f(a)g(b). \]

Proof of Lemma 3.1 Let $I = \{1, 2, \ldots, \lfloor p/2 \rfloor - 1\} \subset {\mathbb {Z}}$.

Let $R$ be the random set obtained by selecting each element of $I$, independently, with probability $Kp^{-2/3}$, where $K = \lceil \log p \rceil$. In particular, $\mathbb {E}\big [\big |R\big |\big ] \asymp Kp^{1/3}$.

Note that $R$ almost satisfies property (i) of Lemma 3.1, in the sense that

\[ \mathbb{E} \big[\big|R\cap (R+R+\{0, 1\})\big| \big] \leqslant \sum_{\substack{a,b,c \in I\\ a-b-c \in \{0, 1\}}} \mathbb{P}\big(a,b,c\in R\big). \]

The probability $\mathbb {P}\big (a,b,c\in R\big )$ is $(Kp^{-2/3})^k$ where $k = |\{a,b,c\}|$. Therefore,

(A.1)\begin{equation} \mathbb{E} \big[\big|R\cap (R+R+\{0, 1\})\big| \big] \ll p^2 (Kp^{-2/3})^3 + p (Kp^{-2/3})^2 + 1 (Kp^{-2/3})^1 \ll K^3. \end{equation}

It will be useful to know the bound $\big \|\mathbf {1}_{R} *\mathbf {1}_{R}\big \|_{\infty } \ll 1$, which holds with high probability. Indeed, if $\mathbf {1}_{R}*\mathbf {1}_{R}(n) \geqslant 9$ for some $n\in {\mathbb {N}}$, then $R$ contains elements $a_1< a_2<\cdots < a_8$ such that $a_i+a_{9-i} = n$ for all $i$ and, thus,

(A.2)\begin{equation} \mathbb{P}\big(\exists n, \ \mathbf{1}_{R}*\mathbf{1}_{R}(n) \geqslant 9\big) \leqslant \sum_{n=1}^{p} \sum_{\substack{a_1, \ldots, a_8\in I\\ a_i\text{ distinct}\\ a_i+a_{9-i} = n}} \mathbb{P}\big(\forall i,\, a_i\in R\big) \ll p^5 (Kp^{-2/3})^8 \ll K^8p^{-1/3} \ll K^{-1}. \end{equation}

Similarly, if $\mathbf {1}_{R}*\mathbf {1}_{-R}(n) \geqslant 8$ for some $n\in {\mathbb {Z}}^{\neq 0}$, then $R$ contains distinct elements $a_1,\ldots, a_8$ such that $a_{2i}-a_{2i-1} = {n}$ for $i\in \{1,2,3,4\}$, and as before we get

(A.3)\begin{equation} \mathbb{P}\big(\exists n\in {\mathbb{Z}}^{\neq 0}, \ \mathbf{1}_{R}*\mathbf{1}_{-R}(n) \geqslant 8\big) \ll K^{-1}. \end{equation}

We now turn to property (ii) of Lemma 3.1. Let $n$ be an integer with $p/5 \leqslant n \leqslant 7p/5$. We show that $n$ can be written in relatively many ways as the sum of three distinct elements of $R$, with high probability. Let $\mathcal {T}(n)$ be the collection of all sets $\{a, b, c\}$ of three pairwise distinct elements of $I$ such that $a+b+c = n$. We show that

(A.4)\begin{equation} \mathbb{P}\biggl( \sum_{\substack{T\in \mathcal{T}(n)}} \mathbf{1}_{T\subset R} \leqslant K^{2}\biggr) \ll e^{-K^2}. \end{equation}

To bound the left-hand side, we use Janson's inequality. This inequality involves the quantities

\[ \lambda := \mathbb{E} \biggl[ \sum_{\substack{T\in \mathcal{T}(n)}} \mathbf{1}_{T\subset R} \biggr] = \sum_{\substack{T\in \mathcal{T}(n)}} (Kp^{-2/3})^3 \gg p^2(Kp^{-2/3})^3 = K^3 \]

and

\[ \Delta := \sum_{\substack{T_1, T_2\in \mathcal{T}(n)\\ T_1\cap T_2\neq \emptyset \\ T_1\neq T_2}} \mathbb{P}\big(T_1\cup T_2\subset R\big). \]

Janson's inequality [Reference Janson, Rucinski and LuczakJRL11, Theorem 2.14] gives

(A.5)\begin{equation} \mathbb{P}\biggl( \sum_{\substack{T\in \mathcal{T}(n)}} \mathbf{1}_{T\subset R} \leqslant \tfrac12 \lambda\biggr) \leqslant \exp \left( -\frac{\lambda^2}{8(\lambda+\Delta)}\right). \end{equation}

Observe that, if $T_1, T_2\in \mathcal {T}(n)$ are distinct but not disjoint, we must have $\big |T_1 \cap T_2\big | = 1$, so there are pairwise distinct elements $a, b_1, b_2, c_1, c_2\in I$ such that $T_1 = \{a, b_1, c_1\}$ and $T_2 = \{a, b_2, c_2\}$. In particular,

\[ \mathbb{P}\big(T_1\cup T_2\subset R\big) = (K p^{-2/3})^5, \]

and there are $\ll p^3$ such pairs $(T_1, T_2)$. Hence,

\[ \Delta \ll p^3 (Kp^{-2/3})^5 \ll 1 \ll \lambda. \]

Provided $p$ is larger than some absolute constant, we have $K^{2} \leqslant \tfrac 12 \lambda$, so (A.4) follows from (A.5), using our estimates for $\lambda$ and $\Delta$. By the union bound, we get

(A.6) \begin{align} \mathbb{P} \biggl( \exists n\in [p/5, 7p/5]\cap {\mathbb{Z}},\ \sum_{\substack{T\in \mathcal{T}(n)}} \mathbf{1}_{T\subset R} \leqslant K^{2} \biggr) \ll p e^{-K^2} \ll K^{-1}. \end{align}

We have shown that, typically, $R$ satisfies property (ii) of Lemma 3.1 in a robust sense (by (A.6)), but $R$ just fails to satisfy property (i) (see (A.1)).

We define $X := R\cap (R+R+\{0, 1\})$ and $A := R\setminus X$. This set $A$ satisfies property (i) of Lemma 3.1 by construction. We must show that $A$ still satisfies property (ii) of Lemma 3.1, with high probability.

Let $n\in [p/5, 7p/5]\cap {\mathbb {Z}}$. We know that, with high probability, $n$ is the sum of three distinct elements of $R$ in many different ways. Thus, having $n\not \in A+A+A$ means that, whenever $n$ is written as $n = a+b+c$ with distinct $a,b,c\in R$, at least one of $a,b,c$ is in $X$.

Define

\[ \mathcal{Q}(n) := \bigl\{(a,b,c,d,e) \in I^5: a,b,c \text{ pairwise distinct}, \ a+b+c = n, \ c-d-e\in \{0, 1\}\bigr\}. \]

Suppose that $n$ is such that

\[ \big|\mathcal{Q}(n) \cap R^5\big|\geqslant K. \]

We claim that either $\max _{m\in {\mathbb {N}}} \mathbf {1}_{R} *\mathbf {1}_{R}(m) \geqslant 10$, $\max _{m\in {\mathbb {N}}} \mathbf {1}_{R} *\mathbf {1}_{-R}(m) \geqslant 10$, or there are four tuples $(a_i, b_i, c_i, d_i, e_i)_{1\leqslant i \leqslant 4} \in \mathcal {Q}(n) \cap R^5$ such that, for all $1\leqslant i< j\leqslant 4$,

(A.7)\begin{equation} \{a_i, b_i, c_i, d_i, e_i\} \cap \{a_j, b_j, c_j, d_j, e_j\} = \emptyset. \end{equation}

Suppose first that there is some $a\in R$ appearing as the first coordinate of $200$ tuples

\[ {(a, b_i, c_i, d_i, e_i)_{1\leqslant i\leqslant 200} \in \mathcal{Q}(n) \cap R^5}. \]

Then $n-a$ can be written as $b_i+c_i$ for all $1\leqslant i\leqslant 200$. If $\mathbf {1}_{R} *\mathbf {1}_{R}(n-a) < 10$, there are $b,c\in R$ such that $b_i = b$ and $c_i = c$ for at least $20$ values of $i\in \{1, \ldots, 200\}$, say for $i \in J$ where $|J| \geqslant 20$. In turn, this implies that $d_i + e_i \in \{c-1, c\}$ for all $i\in J$, but if $\mathbf {1}_{R} *\mathbf {1}_{R}(c-1) < 10$ and $\mathbf {1}_{R} *\mathbf {1}_{R}(c) < 10$ there must exist two distinct indices $i, j\in J$ such that $(d_i, e_i) = (d_j, e_j)$. This is impossible since $(a, b_i, c_i, d_i, e_i)$ and $(a, b_j, c_j, d_j, e_j)$ are distinct tuples.

Similar reasoning shows that no $r\in R$ can appear as the second or third coordinate of $200$ tuples in $\mathcal {Q}(n) \cap R^5$, unless $\max _{m\in {\mathbb {N}}} \mathbf {1}_{R} *\mathbf {1}_{R}(m) \geqslant 10$.

Suppose that there is some $d\in R$ that appears as the fourth coordinate of $200$ tuples

\[ {(a_i, b_i, c_i, d, e_i)_{1\leqslant i\leqslant 200} \in \mathcal{Q}(n) \cap R^5}. \]

Then $c_i-e_i \in \{d, d+1\}$ for all $i$. If $\max _{m\in {\mathbb {N}}} \mathbf {1}_{R} *\mathbf {1}_{-R}(m) < 10$, there are $c, e\in R$ and at least $10$ values of $i$ such that $c_i = c$ and $e_i = e$. Thus, for these $\geqslant 10$ values of $i$, we have $a_i + b_i = n-c$. If $\mathbf {1}_{R} *\mathbf {1}_{R}(n-c) < 10$, there are two distinct indices $i,j$ such that $(a_i, b_i, c_i, d, e_i) = (a_j, b_j, c_j, d, e_j)$, a contradiction.

The same reasoning shows that no $e\in R$ appears in $200$ distinct elements of $\mathcal {Q}(n) \cap R^5$, unless $\big \|\mathbf {1}_{R} *\mathbf {1}_{R}\big \|_{\infty } \geqslant 10$ or $\max _{m\in {\mathbb {N}}} \mathbf {1}_{R} *\mathbf {1}_{-R}(m) \geqslant 10$.

The claim follows easily from these observations. Indeed, suppose that $\big \|\mathbf {1}_{R} *\mathbf {1}_{R}\big \|_{\infty } < 10$, $\max _{m\in {\mathbb {N}}} \mathbf {1}_{R} *\mathbf {1}_{-R}(m) < 10$ and $\big |\mathcal {Q}(n) \cap R^5\big |\geqslant K$. Let $(a_1, b_1, c_1, d_1, e_1)$ be an arbitrary element of $\mathcal {Q}(n) \cap R^5$. The previous observations show that the number of tuples in $\mathcal {Q}(n) \cap R^5$ having a coordinate in $\{a_1, \ldots, e_1\}$ is bounded above by an absolute constant. Since $K$ is assumed to be sufficiently large, we can find another tuple $(a_2, b_2, c_2, d_2, e_2)$ such that $\{a_1, \ldots, e_1\} \cap \{a_2, \ldots, e_2\} = \emptyset$. Repeating, we find four tuples $(a_i, b_i, c_i, d_i, e_i)_{1\leqslant i\leqslant 4}\in \mathcal {Q}(n) \cap R^5$ satisfying (A.7).

By this claim and the union bound, we deduce that

\begin{align*} &\mathbb{P} \bigl(\exists n\in [p/5, 7p/5]\cap {\mathbb{Z}}, \big|\mathcal{Q}(n) \cap R^5\big|\geqslant K\bigr)\\ &\quad \leqslant \mathbb{P}\bigl( \big\|\mathbf{1}_{R} *\mathbf{1}_{R}\big\|_{\infty} \geqslant 10\bigr) + \mathbb{P} \bigl(\max_{m\in {\mathbb{N}}} \mathbf{1}_{R} *\mathbf{1}_{-R}(m) \geqslant 10\bigr) \\ &\qquad\quad + \sum_{n\in [p/5, 7p/5]\cap {\mathbb{Z}}}\ \sum_{\substack{(a_i, b_i, c_i, d_i, e_i)_{1\leqslant i\leqslant 4} \in \mathcal{Q}(n) \\ \text{satisfying (A.7)}}} \mathbb{P} \bigl(\forall i,\ a_i, b_i, c_i, d_i, e_i\in R \bigr). \end{align*}

By (A.2) and (A.3), the probabilities $\mathbb {P}\big( \big \|\mathbf {1}_{R} *\mathbf {1}_{R}\big \|_{\infty } \geqslant 10\big)$ and $\mathbb {P} (\max _{m\in {\mathbb {N}}} \mathbf {1}_{R} *\mathbf {1}_{-R}(m) \geqslant 10)$ are $\ll K^{-1}$. By (A.7), the events $\{a_i, \ldots, e_i\in R\}$ and $\{a_j, \ldots, e_j\in R\}$ are independent for $i\neq j$. Thus, the last probability is

\[ \mathbb{P} (\forall i,\ a_i, b_i, c_i, d_i, e_i\in R) = \prod_{i=1}^4 \mathbb{P}\big(a_i, b_i, c_i, d_i, e_i\in R\big) = \prod_{i=1}^4 (Kp^{-2/3})^{|\{a_i, \ldots, e_i\}|}. \]

For $k\in \{3,4,5\}$, let

\[ \mathcal{Q}_k(n) = \bigl\{ (a,b,c,d,e)\in \mathcal{Q}(n) : \big| \{a,b,c,d,e\}\big| = k \bigr\}. \]

It is not hard to see that $\big |\mathcal {Q}_k(n)\big | \ll p^{k-2}$ for $k\in \{3,4,5\}$. Hence, we obtain

\begin{align*} \mathbb{P} \bigl(\exists n\in [p/5, 7p/5]\cap {\mathbb{Z}},\ \big|\mathcal{Q}(n) \cap R^5\big|\geqslant K\bigr) &\ll K^{-1} + \sum_{n\in [p/5, 7p/5]\cap {\mathbb{Z}}} \prod_{i=1}^4 \sum_{k =3}^5 \big|\mathcal{Q}_{k}(n)\big| (Kp^{-2/3})^{k}\\ &\ll K^{-1} + p \biggl( \sum_{k=3}^5 p^{k-2} (Kp^{-2/3})^k \biggr)^4 \\ &\ll K^{-1} + p (K^{5} p^{-1/3})^4 \\ &\ll K^{-1}. \end{align*}

The last computation and (A.6) imply that there is a set $R\subset I$ such that:

  1. (a) for all $n\in [p/5, 7p/5]\cap {\mathbb {Z}}$, there are $>K^2$ sets $\{a,b,c\}$ of three distinct elements of $R$ such that $n = a+b+c$;

  2. (b) for all $n\in [p/5, 7p/5]\cap {\mathbb {Z}}$, there are $\ll K$ sets $\{a,b,c\}$ of three distinct elements of $R$ such that $n = a+b+c$, with one of $a,b,c$ in $R+R+\{0, 1\}$;

since with high probability, both properties hold simultaneously.

As announced earlier, we define $X = R\cap (R+R+\{0, 1\}$ and $A = R\setminus X$. Then $A$ and $A+A+\{0, 1\}$ are disjoint by construction, and for every $n\in [p/5, 7p/5]\cap {\mathbb {Z}}$ there are $K^2 - O(K) \geqslant 1$ ways to write $n$ as the sum of three elements of $R$, none of which being in $X$. This means that we can write $n$ as the sum of three elements of $A$, and we are done.

Footnotes

1 There is a typo in [Reference SárközySár01, Problem 32]: ‘2’ should be replaced with ‘3’.

2 The constant $0.064$ in their paper [Reference Kiss and SándorKS23] is a typo.

3 Note that $s(f)$ may exceed $b_{2k+1}$, whereas for $1\leqslant i\leqslant 2k$ the $i$th digit ($e_{(i+1)/2}$ or $r_{i/2}$) is always between $0$ and $b_i-1$.

4 Note that $s(f_2) \leqslant q^{3k_2} < b_{2k_2+1}b_{2k_2+2}b_{2k_2+3}$, which is why these formulas are valid for $k_2+3 \leqslant i\leqslant k_1$.

5 Recall that $\mathbb {P}$ refers to the probability in the random choice of $r_i(f), s(f)$ (which are the random variables used to define $S$).

References

Ajtai, M., Komlós, J. and Szemerédi, E., A dense infinite Sidon sequence, European J. Combin. 2 (1981), 111.CrossRefGoogle Scholar
Cilleruelo, J., Infinite Sidon sequences, Adv. Math. 255 (2014), 474486.CrossRefGoogle Scholar
Cilleruelo, J., On Sidon sets and asymptotic bases, Proc. Lond. Math. Soc. (3) 111 (2015), 12061230.CrossRefGoogle Scholar
Deshouillers, J.-M. and Plagne, A., A Sidon basis, Acta Math. Hungar. 123 (2009).CrossRefGoogle Scholar
Erdős, P., Solved and unsolved problems in combinatorics and combinatorial number theory, European J. Combin. 2 (1981), 111.Google Scholar
Erdős, P., The probability method: successes and limitations, J. Statist. Plann. Inference 72 (1998), 207213.CrossRefGoogle Scholar
Erdős, P., Sárközy, A. and Sós, V. T., On additive properties of general sequences, Discrete Math. 136 (1994), 7599.CrossRefGoogle Scholar
Erdős, P., Sárközy, A. and Sós, V. T., On sum sets of Sidon sets, I, J. Number Theory 47 (1994), 329347.CrossRefGoogle Scholar
Erdős, P., Sárközy, A. and Sós, V. T., On sum sets of Sidon sets, II, Israel J. Math. 90 (1995), 221233.CrossRefGoogle Scholar
Goguel, J. H., Über Summen von zufälligen Folgen natürlicher Zahlen, J. Reine Angew. Math. 272 (1975), 6377.Google Scholar
Halberstam, H. and Roth, K. F., Sequences (Springer, New York, 1983).CrossRefGoogle Scholar
Iwaniec, H. and Kowalski, E., Analytic number theory, vol. 53 (American Mathematical Society, 2004).Google Scholar
Janson, S., Rucinski, A. and Luczak, T., Random graphs (John Wiley & Sons, 2011).Google Scholar
Kiss, S. Z., On Sidon sets which are asymptotic bases, Acta Math. Hungar. 128 (2010), 4658.CrossRefGoogle Scholar
Kiss, S. Z., Rozgonyi, E. and Sándor, C., On Sidon sets which are asymptotic bases of order 4, Funct. Approx. Comment. Math. 51 (2014), 393413.CrossRefGoogle Scholar
Kiss, S. Z. and Sándor, C., Generalized asymptotic Sidon basis, Discrete Math. 344 (2021), 112208.CrossRefGoogle Scholar
Kiss, S. Z. and Sándor, C., On $B_h[1]$-sets which are asymptotic bases of order 2h, Preprint (2022), arXiv:2202.13841.Google Scholar
Kiss, S. Z. and Sándor, C., Dense sumsets of Sidon sequences, European J. Combin. 107 (2023), 103600.CrossRefGoogle Scholar
Ruzsa, I. Z., An infinite Sidon sequence, J. Number Theory 68 (1998), 6371.CrossRefGoogle Scholar
Sárközy, A., Unsolved problems in number theory, Period. Math. Hungar. 42 (2001), 1735.CrossRefGoogle Scholar
Sawin, W., Square-root cancellation for sums of factorization functions over squarefree progressions in $\mathbb {F}_q[t]$, Acta Math., to appear. Preprint (2021), arXiv:2102.09730.Google Scholar