Hostname: page-component-586b7cd67f-t7fkt Total loading time: 0 Render date: 2024-11-21T23:38:12.575Z Has data issue: false hasContentIssue false

On the smallest gap in a sequence with Poisson pair correlations

Published online by Cambridge University Press:  21 November 2024

Daniel Altman*
Affiliation:
Department of Mathematics, Stanford University, Stanford, CA, USA
Zachary Chase
Affiliation:
Department of Computer Science and Engineering, University of California, San Diego, CA, USA
*
Corresponding author: Daniel Altman; Email: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

We prove that any increasing sequence of real numbers with average gap $1$ and Poisson pair correlations has some gap that is at least $3/2+10^{-9}$. This improves upon a result of Aistleitner, Blomer, and Radziwiłł.

Type
Paper
Copyright
© The Author(s), 2024. Published by Cambridge University Press

1. Introduction

Let $\lambda = (\lambda _n)_{n=1}^\infty$ be an increasing sequence of real numbers. Often, for number-theoretic sequences $\lambda$ , the average gap $\lambda _{n+1}-\lambda _n$ is well-understood, while little is known about the distribution function of the gaps. Sometimes, however, statistical information about the collection of gaps $\lambda _{n+k}-\lambda _n$ is of importance.

For example, letting $(\gamma _n)_{n=1}^\infty$ denote the imaginary parts of the zeroes of the Riemann zeta function in the critical strip in increasing order, we know

\begin{equation*}\#\{\gamma _n \le T\} \sim \frac {T\log T}{2\pi }\end{equation*}

as $T \to \infty$ , and Montgomery’s pair-correlation conjecture predicts that

\begin{equation*}\frac {2\pi }{T\log T}\#\left\{(n,m) \,:\, \gamma _n,\gamma _m \le T \text {, } \frac {2\pi a}{\log T} \le \gamma _m-\gamma _n \le \frac {2\pi b}{\log T}\right\} \to \int _a^b \left (1-\textrm {sinc}^2(\pi t)\right )dt \end{equation*}

as $T \to \infty$ , for any fixed $0 \lt a \lt b$ , where $\textrm{sinc}(x)\,:\!=\, \frac{\sin x}{x}$ .

Henceforth, let $\lambda = (\lambda _n)_{n=1}^\infty$ denote an increasing sequence with average gap $1$ :

\begin{equation*}\frac {1}{N}\sum _{n \le N} (\lambda _{n+1}-\lambda _n) \to 1. \end{equation*}

With this normalization, we may define the pair correlation function $R_\lambda$ by

\begin{equation*}R_\lambda (I,N) \,:\!=\, \frac {1}{N}\#\left\{(i,j) \,:\, 1 \le i \not = j \le N \,:\, \lambda _j-\lambda _i \in I\right\},\end{equation*}

where $I \subseteq{\mathbb{R}}$ is a bounded interval and $N$ a positive integer.

For example, up to some normalization technicalities, Montgomery’s pair-correlation conjecture asserts that $R_{(\gamma _n)_n}$ converges (in distribution) to a distribution with cumulative distribution function $1-\textrm{sinc}^2(\pi t)$ .

Motivated by the fact that the pair correlation function of a random sequence generated by a Poisson point process converges in distribution to the uniform distribution, an increasing sequence of real numbers $(\lambda _n)_{n=1}^\infty$ with average gap $1$ is said to have Poisson pair correlations (PPC) if $R_{(\lambda _n)_n}$ converges to the uniform distribution:

(1) \begin{equation} \lim _{N \to \infty } \frac{1}{N}\left|\left\{(i,j) \,:\, 1 \le i \not = j \le N \,:\, \lambda _j-\lambda _i \in I\right\}\right | = |I| \end{equation}

for all intervals $I \subseteq{\mathbb{R}}$ , where $|\cdot |$ denotes the Lebesgue measure.

Despite Montgomery’s pair-correlation conjecture concerning increasing sequences of real numbers, most research on properties of general sequences with Poisson pair correlation concerns sequences in the torus (see, for example, [Reference Aistleitner, Blomer and Radziwiłł1Reference Steinerberger12]), with little investigated about sequences of real numbers.

Some specific number-theoretic sequences of real numbers have been shown to have PPC. Sarnak [Reference Rudnick and Zaharescu10] showed that almost every positive definite binary quadratic form (in a suitable sense) gives rise to a sequence with PPC, by ordering the values it takes on (pairs of) positive integers and appropriately normalizing. Concretely, a consequence of the work of Eskin, Margulis, and Mozes [Reference Eskin, Margulis and Mozes4] is that the ordered sequence of values of $x^2+\sqrt{2}y^2$ for $x,y \in \mathbb{N}$ has PPC.

Aistleitner, Blomer, and Radziwiłł [Reference Aistleitner, Blomer and Radziwiłł1] studied the related triple correlation function of certain number-theoretic sequences, while also initiating a study of general sequences of real numbers with Poisson pair (and triple) correlations. They asked the following.

Question. Let $\lambda _1 \lt \lambda _2 \lt \dots$ be an increasing sequence of real numbers with average gap $1$ and with Poisson pair correlations. How small can $\limsup _{n \to \infty } \lambda _{n+1}-\lambda _n$ be?

Among increasing sequences of real numbers with average gap $1$ and PPC, Aistleitner, Blomer, and Radziwiłł exhibited one with maximum gap $2$ , and proved that any such sequence must have a gap of size at least $3/2-\epsilon$ , for any $\epsilon \gt 0$ . They asked in their paper [Reference Aistleitner, Blomer and Radziwiłł1] as well as at Oberwolfach 2019 (communicated to the authors by Ben Green) to improve either bound. Our main theorem is an improved lower bound.

Theorem 1.1. Let $\lambda _1 \lt \lambda _2 \lt \dots$ be an increasing sequence of real numbers with average gap $1$ and Poisson pair correlations. Then $\limsup _{n \to \infty } \lambda _{n+1}-\lambda _n \gt \frac{3}{2}+10^{-9}$ .

We leave open the question of how small the largest gap can be; in light of Theorem 1.1, it lies between $\frac{3}{2}+10^{-9}$ and $2$ , inclusive.

2. Motivation and proof sketch of Theorem 1.1

In this section we motivate the proof of Theorem 1.1, overlooking some technical complications and emphasizing the main ideas.

Let us begin by sketching the proof given in [Reference Aistleitner, Blomer and Radziwiłł1] that any strictly increasing sequence of real numbers with mean gap $1$ and PPC has a gap at least $\frac{3}{2}-\epsilon$ , for any $\epsilon \gt 0$ . An interested reader may wish to consult the proof sketch of this result in [Reference Aistleitner, Blomer and Radziwiłł1, Section 1.5], or the proof itself in [Reference Aistleitner, Blomer and Radziwiłł1, Section 7].

One begins by observing that for any sequence $(\lambda _n)_n$ that has PPC, the distribution function of the gaps $\lambda _{n+1}-\lambda _n$ , which we will denote by $F$ , can grow at most linearly (indeed, if it grows more than linearly in any small interval, say, this will contradict the PPC condition for this interval). Now, if $\lambda _{n+1}-\lambda _n \le 3/2-\epsilon$ for each $n$ (that is, $F(3/2-\epsilon ) = 1$ ), then the linear growth condition implies that the distribution function, when plotted, lies on or above the straight line between $(1/2-\epsilon, 0)$ and $(3/2-\epsilon, 1)$ . This implies that that the mean gap size is strictly smaller than 1, a contradiction.

Next, running the same argument now for the situation when the maximum gap is equal to $3/2$ yields that in this situation we must have

(2) \begin{equation} F(x) = 0 \text{ for } x \le \frac{1}{2} \end{equation}
(3) \begin{equation} F(x) = x-\frac{1}{2} \text{ for } x \in \left[\frac{1}{2},\frac{3}{2}\right]. \end{equation}

Our goal is to show that this is impossible (and to then obtain a small quantitative improvement over $3/2$ ). Suppose that (2) and (3) hold. Then (3) yields that $\text{PPC}\left(\frac{1}{2},\frac{3}{2}\right)$ $\big($ i.e., (1) for $I = \big[\frac{1}{2},\frac{3}{2}\big]\big)$ is already satisfied by the single gaps $\lambda _{n+1}-\lambda _n$ , so there cannot be a nontrivial contribution coming from larger gaps $\lambda _{n+m}-\lambda _n$ , $m \ge 2$ . On the other hand, (2) implies that $\text{PPC}\left(0,\frac{1}{2}\right)$ must come entirely from a $0$ density part of the sequence, and more specifically only from blocks $[n_1,n_2]$ contained in that $0$ density part. Furthermore, for the union of such blocks to nontrivially contribute to the $\text{PPC}$ count, the length of the blocks must grow with $N$ .

The natural question then is whether such blocks can satisfy the $\text{PPC}$ condition on all subintervals of $\left[0,\frac{1}{2}\right]$ . We show that the answer is no. The key is to establish a “bias near $0$ ” of the $\text{PPC}$ count on long blocks whose total gap is at most $1/2$ . A bit more precisely, if $\lambda _1 \lt \dots \lt \lambda _k$ have $\lambda _k -\lambda _1 \le 1/2$ , then $\frac{1}{|J|}\sum _{1 \le i \lt j \le k} 1_{\lambda _j-\lambda _i \in J}$ is larger for intervals $J \subseteq [0,1/2]$ concentrated near $0$ , with the bias becoming more pronounced as $k \to \infty$ . This would contradict the PPC condition on subintervals of $\left[0,\frac{1}{2}\right]$ . A difficulty we encounter in the proof, however, is that the blocks forming the relevant $0$ density part of the sequence need not have total gap at most $1/2$ . We overcome this by suitably partitioning the $0$ density part of the sequence. This is implemented in Section 4 where we use a suitable greedy algorithm to decompose.

3. Proof of Theorem 1.1

Let $\epsilon = 10^{-9}$ . For this section, we fix an increasing sequence of real numbers $\lambda _1 \lt \lambda _2 \lt \dots$ with average gap $1$ and $\text{PPC}$ , that has $\lambda _{n+1}-\lambda _n \le 3/2+\epsilon$ for sufficiently large $n$ . By truncating the sequence, we may assume that

\begin{equation*}g_n \,:\!=\, \lambda _{n+1}-\lambda _n\end{equation*}

satisfies $g_n \le 3/2+\epsilon$ for each $n \ge 1$ . As in Section 2, we write $\text{PPC}(a,b)$ to denote equation (1) for $I = [a,b]$ . We note that a sequence satisfying the PPC condition for all such $I$ necessarily satisfies the same condition for all open or indeed half-open intervals. We may therefore also use $\text{PPC}(a,b)$ to refer to equation (1) for the half-open interval $[a,b)$ , for example.

Recall that the lower bound of $\frac{3}{2}$ was established in [Reference Aistleitner, Blomer and Radziwiłł1], whose proof we sketched in Section 2. We start the proof of Theorem 1.1 by making these arguments quantitative.

We begin with a quantitative extension of equation (2), which said that if the maximum gap is $3/2$ then, in the limit $N\to \infty$ , the proportion of gaps of size at most $1/2$ is $0$ . We obtain the following when the maximum gap is of size at most $3/2 + \epsilon$ .

Proposition 3.1. For all $N$ sufficiently large, we have

\begin{equation*}\frac {1}{N}\#\left\{n \le N \,:\, g_n \le \frac {1}{2}\right\} \le 2\sqrt {\epsilon }.\end{equation*}

Proof. We begin by relating the mean gap size to the gap distribution function under the given restriction on the maximum gap size. To this end, note that

(4) \begin{align} \frac{1}{N}\sum _{n \le N} g_n &= \int _0^{\frac{3}{2}+\epsilon } \frac{1}{N}\#\left\{n \le N \,:\, g_n \gt x\right\}dx \nonumber\\ &= \int _0^{\frac{1}{2}+\sqrt{\epsilon }} \frac{1}{N}\#\left\{n \le N \,:\, g_n \gt x\right\}dx + \int _{\frac{1}{2}+\sqrt{\epsilon }}^{\frac{3}{2}+\epsilon } \frac{1}{N}\#\left\{n \le N \,:\, g_n \gt x\right\}dx \nonumber\\ &= \frac{1}{2}+\sqrt{\epsilon }-\int _0^{\frac{1}{2}+\sqrt{\epsilon }} \frac{1}{N}\#\left\{n \le N \,:\, g_n \le x\right\}dx \nonumber\\ &\quad + \int _{\frac{1}{2}+\sqrt{\epsilon }}^{\frac{3}{2}+\epsilon } \frac{1}{N}\#\left\{n \le N \,:\, g_n \in \left(x,\frac{3}{2}+\epsilon \right)\right\}dx. \end{align}

Now, we claim that

(5) \begin{equation} \limsup _{N \to \infty } \int _{\frac{1}{2}+\sqrt{\epsilon }}^{\frac{3}{2}+\epsilon } \frac{1}{N}\#\left\{n \le N \,:\, g_n \in \left(x,\frac{3}{2}+\epsilon \right)\right\}dx \le \int _{\frac{1}{2}+\sqrt{\epsilon }}^{\frac{3}{2}+\epsilon } \left(\frac{3}{2}+\epsilon -x\right)dx. \end{equation}

Indeed, the pointwise upper bound

\begin{equation*}\frac {1}{N}\#\left\{n \le N \,:\, g_n \in \left(x,\frac {3}{2}+\epsilon \right)\right\} \le \min \left(1,\frac {1}{N}\sum _{n \le N} \sum _{m \le N-n+1} 1_{g_n+\dots +g_{n+m-1} \in (x,\frac {3}{2}+\epsilon )}\right)\end{equation*}

together with the dominated convergence theorem and $\text{PPC}\left(x, \frac{3}{2}+\epsilon \right)$ for $x \in \left(\frac{1}{2}+\sqrt{\epsilon },\frac{3}{2}+\epsilon \right)$ , namely,

\begin{equation*}\lim _{N \to \infty } \frac {1}{N}\sum _{n \le N} \sum _{m \le N-n+1} 1_{g_n+\dots +g_{n+m-1} \in \left(x,\frac {3}{2}+\epsilon \right)} = \frac {3}{2}+\epsilon -x,\end{equation*}

gives (5). That the average gap of $(\lambda _n)_n$ is $1$ corresponds to

(6) \begin{equation} \lim _{N \to \infty } \frac{1}{N}\sum _{n \le N} g_n = 1. \end{equation}

Rearranging (4), taking $N \to \infty$ , and using (5) and (6) gives

\begin{align*} \limsup _{N \to \infty } \int _0^{\frac{1}{2}+\sqrt{\epsilon }} \frac{1}{N}\#\{n \le N \,:\, g_n \le x\}dx &\le -1+\frac{1}{2}+\sqrt{\epsilon }+\int _{\frac{1}{2}+\sqrt{\epsilon }}^{\frac{3}{2}+\epsilon } \left(\frac{3}{2}+\epsilon -x\right)dx \\ &= \frac{3}{2}\epsilon +\frac{1}{2}\epsilon ^2-\epsilon ^{3/2}. \end{align*}

Thus, using the trivial

\begin{equation*}\frac {1}{N} \#\left\{n \le N \,:\, g_n \le x\right\} \ge \frac {1}{N}\#\left\{n \le N \,:\, g_n \le \frac {1}{2}\right\}\end{equation*}

for $x \in \left[\frac{1}{2},\frac{1}{2}+\sqrt{\epsilon }\right]$ and the even more trivial lower bound of $0$ when $x \in \left[0,\frac{1}{2}\right]$ , yields

\begin{equation*}\limsup _{N \to \infty } \sqrt {\epsilon }\cdot \frac {1}{N}\#\left\{n \le N \,:\, g_n \le \frac {1}{2}\right\} \le \frac {3}{2}\epsilon +\frac {1}{2}\epsilon ^2-\epsilon ^{3/2}.\end{equation*}

Dividing by $\sqrt{\epsilon }$ , Proposition3.1 follows.

We now use the quantitative version of (3) to argue that the PPC( $\frac{1}{2},\frac{3}{2}+\epsilon$ ) contribution comes nearly entirely from single gaps $g_n$ .

Proposition 3.2. For all $N$ sufficiently large, we have

\begin{equation*}\frac {1}{N} \sum _{n \le N} \sum _{2 \le m \le N-n+1} 1_{g_n+\dots +g_{n+m-1} \in (\frac {1}{2},\frac {3}{2}+\epsilon )} \le 2\sqrt {\epsilon }.\end{equation*}

Proof. As in the proof of Proposition3.1,

\begin{align*}\frac {1}{N}\sum _{n \le N} g_n & = \frac {1}{2}-\sqrt {\epsilon }-\int _0^{\frac {1}{2}-\sqrt {\epsilon }} \frac {1}{N}\#\left\{n \le N \,:\, g_n \le x\right\}dx\nonumber\\& \quad + \int _{\frac {1}{2}-\sqrt {\epsilon }}^{\frac {3}{2}+\epsilon } \frac {1}{N}\#\left\{n \le N \,:\, g_n \in \left(x,\frac {3}{2}+\epsilon \right)\right\}dx,\end{align*}

which, by merely dropping a (negative) term, gives

(7) \begin{equation} \frac{1}{N}\sum _{n \le N} g_n \le \frac{1}{2}-\sqrt{\epsilon }+\int _{\frac{1}{2}-\sqrt{\epsilon }}^{\frac{3}{2}+\epsilon } \frac{1}{N}\#\left\{n \le N \,:\, g_n \in \left(x,\frac{3}{2}+\epsilon \right)\right\}dx. \end{equation}

We write

\begin{align*}\frac {1}{N}\#\left\{n \le N \,:\, g_n \in \left(x,\frac {3}{2}+\epsilon \right)\right\} & = \frac {1}{N}\sum _{n \le N}\sum _{m \le N-n+1} 1_{g_n+\dots +g_{n+m-1} \in \left(x,\frac {3}{2}+\epsilon \right)}\nonumber\\[4pt]& \quad -\frac {1}{N}\sum _{n \le N}\sum _{2 \le m \le N-n+1} 1_{g_n+\dots +g_{n+m-1} \in \left(x,\frac {3}{2}+\epsilon \right)}\end{align*}

and use the same dominated convergence theorem argument as in the proof of Proposition3.1 to obtain, from (7), that

\begin{align*} &\limsup _{N \to \infty } \int _{\frac{1}{2}-\sqrt{\epsilon }}^{\frac{3}{2}+\epsilon } \frac{1}{N}\sum _{n \leq N}\sum _{2 \le m \le N-n+1} 1_{g_n+\dots +g_{n+m-1} \in \left(x,\frac{3}{2}+\epsilon \right)}dx \\[3pt] &\qquad \qquad \le -1+\frac{1}{2}-\sqrt{\epsilon }+\int _{\frac{1}{2}-\sqrt{\epsilon }}^{\frac{3}{2}+\epsilon } \left(\frac{3}{2}+\epsilon -x\right)dx \\[3pt] &\qquad \qquad = \frac{3}{2}\epsilon +\frac{1}{2}\epsilon ^2+\epsilon ^{3/2}, \end{align*}

and thus

\begin{equation*}\limsup _{N \to \infty } \sqrt {\epsilon }\cdot \frac {1}{N}\sum _{n \leq N}\sum _{2 \le m \le N-n+1} 1_{g_n+\dots +g_{n+m-1} \in \left(\frac {1}{2},\frac {3}{2}+\epsilon \right)} \le \frac {3}{2}\epsilon +\frac {1}{2}\epsilon ^2+\epsilon ^{3/2}.\end{equation*}

Dividing by $\sqrt{\epsilon }$ , the proposition follows.

We now exploit the aforementioned “bias” towards $0$ exhibited by large intervals with sum of gaps at most $1/2$ . We will need a technical lemma, proven in the appendix but assumed for now.

Lemma 3.3. For positive integers $a,b,c,L$ satisfying $1 \le a \le b \le c \le L$ , we have

\begin{align*} & (a-1)a+(b-a)(b-a+1)+(c-b)(c-b+1)+(L-c)(L-c+1)\\& +(a-1)(b-a)+(b-a)(c-b)+(c-b)(L-c) \ge \frac {5}{12}L^2+\frac {1}{6}L-\frac {7}{12}.\end{align*}

Lemma 3.3 allows us to show that, instead of getting the desired $\frac{1}{4}+\frac{1}{8} = \frac{3}{8}$ for $\text{PPC}\left(0,\frac{1}{4}\right)+\text{PPC}\left(0,\frac{1}{8}\right)$ , we get at least $\frac{5}{12} = \frac{3}{8}+\frac{1}{24}$ , asymptotically for large intervals.

Proposition 3.4. Let $L \ge 1$ be a positive integer and $g_1,\dots, g_L$ be positive reals with $\sum _{i=1}^L g_i \le \frac{1}{2}$ . Then,

\begin{equation*}\sum _{n \le L}\sum _{m \le L-n+1} 1_{g_n+\dots +g_{n+m-1} \le \frac {2}{8}}+\sum _{n \le L}\sum _{m \le L-n+1} 1_{g_n+\dots +g_{n+m-1} \le \frac {1}{8}} \ge \frac {5}{6}\binom {L+1}{2} - \frac {5}{6}L.\end{equation*}

Proof. By scaling, it suffices to prove the proposition when $\sum _{i=1}^L g_i = \frac{1}{2}$ . Suppose $\sum _{i=1}^L g_i = \frac{1}{2}$ . Let

\begin{equation*}a = \min \left\{j \le L \,:\, g_1+\dots +g_j \ge \frac {1}{8}\right\},\end{equation*}
\begin{equation*}b = \min \left\{j \le L \,:\, g_1+\dots +g_j \ge \frac {2}{8}\right\},\end{equation*}
\begin{equation*}c = \min \left\{j \le L \,:\, g_1+\dots +g_j \ge \frac {3}{8}\right\},\end{equation*}

and note

\begin{align*}\sum _{n \le L}\sum _{m \le L-n+1} 1_{g_n+\dots +g_{n+m-1} \le \frac {1}{8}} & \ge \frac {(a-1)a}{2}+\frac {(b-a)(b-a+1)}{2}+\frac {(c-b)(c-b+1)}{2}\\& \quad +\frac {(L-c+1)(L-c+2)}{2}.\end{align*}

by doing casework in which intervals $n$ and $n+m-1$ lie (the different intervals are $[1,a),[a,b),[b,c),[c,L]$ ). Similarly,

\begin{align*}\sum _{n \le L} \sum _{m \le L-n+1} 1_{g_n+\dots +g_{n+m-1} \le \frac {2}{8}} & \ge \frac {(a-1)a}{2}+\frac {(b-a)(b-a+1)}{2}+\frac {(c-b)(c-b+1)}{2}\\& \quad +\frac {(L-c+1)(L-c+2)}{2}+(a-1)(b-a)\\[5pt]& \quad +(b-a)(c-b)+(c-b)(L-c+1),\end{align*}

Therefore,

\begin{align*}\sum _{n \le L} \sum _{m \le L-n+1} 1_{g_n+\dots +g_{n+m-1} \le \frac {2}{8}} & + \sum _{n \le L}\sum _{m \le L-n+1} 1_{g_n+\dots +g_{n+m-1} \le \frac {1}{8}} \ge (a-1)a+(b-a)(b-a+1)\\& +(c-b)(c-b+1)+(L-c+1)(L-c+2)+(a-1)(b-a)\\& \quad +(b-a)(c-b)+(c-b)(L-c+1).\end{align*}

Lower bounding $L-c+1$ and $L-c+2$ by $L-c$ and $L-c+1$ , respectively, Lemma 3.3 finishes the proof of Proposition3.4, since $\frac{5}{12}L^2+\frac{1}{6}L-\frac{7}{12} \ge \frac{5}{6}{\left(\begin{smallmatrix}L + 1\\2\end{smallmatrix}\right)}-\frac{5}{6}L$ for $L \ge 1$ .

We now proceed to isolate the relevant “ $0$ density” parts of the sequence on which the gaps are at most $1/2$ , in order to exploit the bias that Proposition3.4 illustrates.

Here and henceforth, we let $[N]\,:\!=\,\{1,2,\ldots, N\}$ .

Definition 3.5. For a nonempty interval $J \subseteq [N]$ , let $L(J),R(J)$ denote the left and right endpoints of $J$ , respectively, and let $\textrm{sum}(J) = \sum _{n \in J} g_n$ .

Definition 3.6. For a interval $J \subseteq [N]$ , we denote

\begin{equation*}\text {PPC}^J(0,a) \,:\!=\, \sum _{n \le n^{\prime} \in J} 1_{g_n+\dots +g_{n^{\prime}} \lt a}.\end{equation*}

For intervals $J_1, J_2 \subseteq [N]$ , with $R(J_1) \lt L(J_2)$ , we denote

\begin{equation*}\text {PPC}^{J_1,J_2}(0,a) \,:\!=\, \sum _{(n,n^{\prime}) \in J_1\times J_2} 1_{g_n+\dots +g_{n^{\prime}} \lt a}.\end{equation*}

Take a large $N$ . Let $\mathcal{I}^N$ denote the collection of all maximal intervals on which the gaps $g_n$ are at most $\frac{1}{2}$ . More formally, we define $\mathcal{I}^N$ to be the collection of all intervals $I \subseteq [N]$ such that (1) $g_n \le \frac{1}{2}$ for each $n \in I$ , (2) $L(I) = 1$ or $g_{L(I)-1} \gt \frac{1}{2}$ , and (3) $R(I) = N$ or $g_{R(I)+1} \gt \frac{1}{2}$ .

We begin by noting the following.

Lemma 3.7. For all large $N$ , we have

\begin{equation*}\sum _{I \in \mathcal {I}^N} |I| \leq 2 \sqrt {\epsilon } N.\end{equation*}

Furthermore, as $N \to \infty$ , we have

\begin{equation*} \sum _{I \in \mathcal {I}^N} \binom {|I|+1}{2} \geq N\left(\frac {1}{2}+o(1)\right).\end{equation*}

Proof. By definition we have

\begin{equation*}\sum _{I \in \mathcal {I}^N} |I| = \frac {1}{N} \# \left\{n \leq N \,:\, g_n \leq \frac {1}{2}\right\}.\end{equation*}

Proposition3.1 then gives the first inequality. For the second inequality, note, by the maximality of the intervals comprising $\mathcal{I}^N$ , that

\begin{equation*} \left(\frac {1}{2} +o(1)\right)N = \text {PPC}^{[N]}\left(0,\frac {1}{2}\right) = \sum _{I \in \mathcal {I}^N}\text {PPC}^I\left(0,\frac {1}{2}\right) \leq \sum _{I \in \mathcal {I}^N} \binom {|I|+1}{2}.\end{equation*}

We provide a quick remark on motivation.

Remark 3.8. Observe that, if it were the case that $\textrm{sum}(I) \le \frac{1}{2}$ for each $I \in \mathcal{I}^N$ , then we could conclude the proof of Theorem 1.1 as follows. By Proposition 3.4, one has

\begin{align*} \left(\frac{3}{8} + o(1)\right)N &= \text{PPC}^{[N]}\left(0,\frac{1}{8}\right) + \text{PPC}^{[N]}\left(0,\frac{1}{4}\right) \\ &\geq \frac{5}{6}\sum _I \binom{|I|+1}{2} - \frac{5}{6}\sum _I |I| \\ &\geq \left(\frac{5}{12} - \frac{5}{3}\sqrt \epsilon +o(1)\right)N \end{align*}

as $N \to \infty$ , which would give our desired contradiction by taking $N$ sufficiently large.

However, it need not be the case that $\textrm{sum}(I) \le \frac{1}{2}$ for each $I \in \mathcal{I}^N$ . In light of Remark 3.8, therefore, the strategy is to partition each $I \in \mathcal{I}^N$ into subintervals $\{J_k^I\}_k$ with $\textrm{sum}\left(J_k^I\right) \leq 1/2$ for each $k$ and such that, for all $a\in (0,1/2]$ and all $k$ , the contribution to PPC $(0,a)$ from windows that overlap with $J_k^I$ comes nearly entirely from windows that lie entirely inside $J_k^I$ . The existence of such a partition is not at all immediate. We provide in Proposition3.9 below a precise statement of what is needed.

Proposition 3.9. Let $\mathcal{I}^N$ be as above. There exists a partition of each $I \in \mathcal{I}^N$ into subintervals $\{J_k^I\}_{k=1}^{r_I}$ such that the following two hold.

  1. 1. $\textrm{sum}\left(J_k^I\right)\leq \frac{1}{2}$ for each $I \in \mathcal{I}^N$ and $k \in \{1,\dots, r_I\}$ , and

  2. 2. $\sum _{I \in \mathcal{I}^N}\sum _{k=1}^{r_I} \binom{|J^I_k|+1}{2} \geq \left(\frac{1}{2} - 4\sqrt{2}\epsilon ^{1/4}\right) N$ .

To quickly conclude the proof of Theorem 1.1, we postpone the proof of Proposition3.9 (and the description of the partition) to the following section, and assume it for now.

Proof of Theorem 1.1. We will proceed along the lines of Remark 3.8, but now use Proposition3.9 to partition into subintervals on which we can apply Proposition3.4. This yields the following computations, where Proposition3.4 is used in the penultimate line, and we use Lemma 3.7 in the final line:

\begin{align*} \left(\frac{3}{8}+o(1)\right)N &= \text{PPC}^{[N]}\left(0,\frac{1}{8}\right)+\text{PPC}^{[N]}\left(0,\frac{1}{4}\right) \\ &\ge \sum _{I \in \mathcal{I}^N} \sum _k \text{PPC}^{J_k^I}\left(0,\frac{1}{8}\right)+\text{PPC}^{J_k^I}\left(0,\frac{1}{4}\right)\\ &\ge \frac{5}{6}\sum _{I \in \mathcal{I}^N} \sum _k \binom{|J_k^I|+1}{2} - \frac{5}{6}\sum _{I \in \mathcal{I}^N} \sum _k |J_k^I| \\ &\ge \frac{5}{6}\left(\frac{1}{2}-4\sqrt{2}\epsilon ^{1/4}\right)N-\frac{5}{3}\sqrt{\epsilon }N. \end{align*}

Rearranging, dividing by $N$ , and sending $N \to \infty$ , we obtain

\begin{equation*}\frac {10\sqrt {2}}{3}\epsilon ^{1/4} + \frac {5}{3}\sqrt \epsilon - \frac {1}{24} \ge 0,\end{equation*}

which is indeed false for $\epsilon = 10^{-9}$ (but not for $\epsilon = 10^{-8}$ ). This gives the desired contradiction to our assumption that a sequence $(\lambda _n)_n$ with PPC, average gap $1$ , and maximum gap $3/2+\epsilon$ exists.

4. Partitioning, and a proof of Proposition3.9

The following examples are helpful to keep in mind to explain the need for care when choosing the partition of a given $I \in \mathcal{I}^N$ , and to help motivate the partition we will use.

\begin{equation*}I_1 = \left \{\frac {2}{5},0,0,0,\dots, 0,0,0,\frac {1}{3},0,0,0,\dots, 0,0,0,\frac {2}{5}\right \}\end{equation*}
\begin{equation*} I_2 = \left \{\frac {1}{4},0,0,0,\dots, 0,0,0,\frac {1}{2},\frac {1}{2},\frac {1}{2},\frac {1}{2},\frac {1}{2},0,0,0,\dots, 0,0,0,\frac {1}{4}\right \}\end{equation*}
\begin{equation*} I_3 = \left \{0,0,\dots, 0,0,\frac {1}{3},0,0,\dots, 0,0,\frac {1}{3},0,0,\dots, 0,0\right \}.\end{equation*}

The first example $I_1$ shows that a “greedy division”, in which one goes from left to right, dividing immediately before the sum first exceeds $1/2$ , will not work. Indeed, that division is

\begin{equation*}{ \left \{\frac {2}{5},0,0,0,\dots, 0,0,0\right \},\left \{\frac {1}{3},0,0,0,\dots, 0,0,0\right \},\left \{\frac {2}{5}\right \}},\end{equation*}

which is problematic as there will be much contribution to the PPC( $0,\frac{1}{2}$ ) count coming from different subintervals; specifically, any index besides the first in the first subinterval and any index in the second subinterval would prove a nontrivial contribution.

Similarly, another “greedy division” in which the largest numbers successively “claim” the largest subinterval they can, will not work. For the case of $I_1$ , the division is

\begin{equation*} \left \{\frac {2}{5},0,0,0,\dots, 0,0,0\right \},\left \{\frac {1}{3}\right \},\left \{0,0,0,\dots, 0,0,0,\frac {2}{5}\right \},\end{equation*}

which has a large contribution to $\text{PPC}\left(0,\frac{1}{2}\right)$ coming from any index in the first subinterval besides the first and any index in the last subinterval besides the last. Note thus that even “non-adjacent” subintervals can cause issues.

A division that does work for $I_1$ is

\begin{equation*}{ \left \{\frac {2}{5}\right \},\left \{0,0,0,\dots, 0,0,0,\frac {1}{3},0,0,0,\dots, 0,0,0\right \},\left \{\frac {2}{5}\right \},} \end{equation*}

as there is only a minor contribution (namely, linear in the size of the interval rather than quadratic) coming from different subintervals.

For $I_2$ , essentially any reasonable division is permissible, but we draw attention to it as it shows that sometimes the reason for negligible contribution from different subintervals are subintervals in between. For example, if we decompose as

\begin{equation*} \left \{\frac {1}{4},0,0,0,\dots, 0,0,0\right \},\left \{\frac {1}{2}\right \},\left \{\frac {1}{2}\right \},\left \{\frac {1}{2}\right \},\left \{\frac {1}{2}\right \},\left \{\frac {1}{2}\right \},\left \{0,0,0,\dots, 0,0,0,\frac {1}{4}\right \},\end{equation*}

then the reason that there is no contribution to $\text{PPC}\left(0,\frac{1}{2}\right)$ from the subintervals $\{\frac{1}{4},0,0,\dots, 0,0\}$ and $\{0,0,\dots, 0,0,\frac{1}{4}\}$ are the five subintervals $\{\frac{1}{2}\}$ in between.

For $I_3$ , any division will admit a large PPC contribution from different subintervals. This would of course be harmful, but we make use of the fact that it won’t exist often in our situation, since it provides a nontrivial contribution to $\text{PPC}(\frac{1}{2},1)$ (which we already know comes nearly entirely from single gaps).

With the above examples in mind, we now choose the partition we use to prove Proposition3.9.

Fix $I \in \mathcal{I}^N$ . In the following definition, ties may be broken arbitrarily. Let $J_1$ be the largest subinterval of $I$ with $\textrm{sum}(J_1) \le \frac{1}{2}$ . With $J_1,\dots, J_r$ already defined, if $\cup _{k=1}^r J_k \not = I$ , let $J_{r+1}$ be the largest subinterval of $I\setminus \cup _{k=1}^r J_k$ with $\textrm{sum}(J_{r+1}) \le \frac{1}{2}$ . Let $J_1,\dots, J_s$ be all the subintervals resulting from this process. Of course $s \le |I| \lt +\infty$ .

Clearly $I = \sqcup _{k=1}^s J_k$ and $\textrm{sum}(J_k) \leq 1/2$ for each $k$ , establishing the first requirement of Proposition3.9. We now begin to proceed to establish the second.

Hopefully not confusing the reader, we renumber now so that $J_1$ is the leftmost interval, with $J_2$ to the immediate right of $J_1$ , $J_3$ to the immediate right of $J_2$ , etc.. For $1 \le k \le s-1$ , let $g_1(k) \in \{k,k+1\}$ and $b_1(k) \in \{k,k+1\}\setminus \{g_1(k)\}$ be such that $J_{g_1(k)}$ was chosen before $J_{b_1(k)}$ . Note, in particular, that $|J_{g_1(k)}| \ge |J_{b_1(k)}|$ .

We quickly pin down exactly which different subintervals need to be considered with regards to their contribution to $\text{PPC}(0,a)$ , for $a \leq 1/2$ .

Definition 4.1. Call $k \in [2,s-1]$ sandwiched if it was chosen after each of its neighbouring subintervals, i.e., if $b_1(k-1) = k$ and $b_1(k) = k$ . For a sandwiched $k$ , let $g_2(k) \in \{k-1,k+1\}$ and $b_2(k) \in \{k-1,k+1\}\setminus \{g_2(k)\}$ be such that $J_{g_2(k)}$ was chosen before $J_{b_2(k)}$ . In particular, $|J_{g_2(k)}| \ge |J_{b_2(k)}|$ .

Lemma 4.2. If $n \le n^{\prime} \in I$ have $g_n+\dots +g_{n^{\prime}} \le \frac{1}{2}$ , then either

  1. 1. $(n,n^{\prime}) \in J_k\times J_k$ for some $k$ ,

  2. 2. $(n,n^{\prime}) \in J_k\times J_{k+1}$ for some $k$ , or

  3. 3. $(n,n^{\prime}) \in J_{k-1}\times J_{k+1}$ for some sandwiched $k$ .

Proof. First note that if $n \in J_k$ and $n^{\prime} \in \cup _{\Delta \ge 3} J_{k+\Delta }$ , then $g_n+\dots +g_{n^{\prime}} \gt \frac{1}{2}$ , since $\textrm{sum}(J_{k+1}\cup J_{k+2}) \gt \frac{1}{2}$ (for otherwise whichever of $J_{k+1},J_{k+2}$ was chosen first would have “engulfed” the other). Now suppose $n \in J_{k-1}$ and $n^{\prime} \in J_{k+1}$ for some $k$ . If $b_1(k-1) = k-1$ , then $\textrm{sum}(R(J_{k-1})\cup J_k) \gt \frac{1}{2}$ , so $g_n+\dots +g_{n^{\prime}} \gt \frac{1}{2}$ . Similarly, if $b_1(k) = k+1$ , then $\textrm{sum}(J_k\cup L(J_{k+1})) \gt \frac{1}{2}$ also yields $g_n+\dots +g_{n^{\prime}} \gt \frac{1}{2}$ . Hence, $k$ is sandwiched.

We now proceed to argue that the PPC $\left(0,\frac{1}{2}\right)$ contribution coming from cases (2) or (3) in Lemma 4.2 is small. We begin with case (2).

We shall argue that the $\text{PPC}\left(0,\frac{1}{2}\right)$ contribution coming from adjacent subintervals $J_k,J_{k+1}$ is small by arguing that $|J_{b_1(k)}|$ is (usually) small. We do this by arguing that we would otherwise have too large of a contribution to $\text{PPC}\left(\frac{1}{2},\frac{3}{2}\right)$ coming from gaps $\lambda _{n+m}-\lambda _n$ with $m \ge 2$ (contradicting Proposition3.2).

Proposition 4.3. For any $k \in [s-1]$ , one has

\begin{equation*}\sum _{(n,n^{\prime}) \in J_k\times J_{k+1}} 1_{g_n+\dots +g_{n^{\prime}} \gt \frac {1}{2}} \ge \frac {1}{2}|J_{b_1(k)}|^2.\end{equation*}

Proof. Without loss of generality, by symmetry we may assume $b_1(k) = k+1$ .

For $y \in J_{k+1}$ and $x \in J_k$ , if $y-x+1 \gt |J_k|$ , then $g_x+\dots +g_y \gt \frac{1}{2}$ , since otherwise $[x,y]$ would have been chosen as $J_k$ (in the greedy process defining the partition) instead of $J_k$ . Therefore,

\begin{align*} \sum _{(n,n^{\prime}) \in J_k\times J_{k+1}} 1_{g_n+\dots +g_{n^{\prime}} \gt \frac{1}{2}} &\ge \sum _{y=L(J_{k+1})}^{R(J_{k+1})} \sum _{x = L(J_k)}^{y-R(J_k)+L(J_k)-1} 1 \\[4pt] &= \left(\frac{L(J_{k+1})+R(J_{k+1})}{2}\right)|J_{k+1}|-R(J_k)|J_{k+1}| \\[4pt] &= \frac{1}{2}|J_{k+1}|^2+\frac{1}{2}|J_{k+1}|, \end{align*}

with the last equality using $R(J_k) = L(J_{k+1})-1$ .

Next we proceed to bound the contribution from intervals $J_{k-1}$ , $J_{k+1}$ for $k$ sandwiched. For such $k$ , we argue that $|J_{b_2(k)}|$ is (usually) small.

Proposition 4.4. For a sandwiched $k$ , one has

\begin{equation*}\sum _{(n,n^{\prime}) \in J_{k-1}\times J_{k+1}} 1_{g_n+\dots +g_{n^{\prime}} \gt \frac {1}{2}} \ge \frac {1}{2}|J_{b_2(k)}|^2.\end{equation*}

Proof. Without loss of generality, by symmetry we may assume $b_2(k) = k+1$ .

For $y \in J_{k+1}$ and $x \in J_{k-1}$ , if $y-x+1 \gt |J_{k-1}|$ , then $g_x+\dots +g_y \gt \frac{1}{2}$ , since otherwise $[x,y]$ would have been chosen instead of $J_{k-1}$ (recall $J_k$ was also chosen after $J_{k-1}$ , since $k$ is sandwiched). Therefore,

\begin{equation*}\sum _{(n,n^{\prime}) \in J_{k-1}\times J_{k+1}} 1_{g_n+\dots +g_{n^{\prime}} \gt \frac {1}{2}} \ge \sum _{y=L(J_{k+1})}^{R(J_{k+1})} \sum _{x = L(J_{k-1})}^{\min (R(J_{k-1}),y-|J_{k-1}|)} 1.\end{equation*}

If $R(J_{k+1})-|J_{k-1}| \le R(J_{k-1})$ , then we obtain

\begin{align*} \sum _{(n,n^{\prime}) \in J_{k-1}\times J_{k+1}} 1_{g_n+\dots +g_{n^{\prime}} \gt \frac{1}{2}} &\ge \sum _{y = L(J_{k+1})}^{R(J_{k+1})} \sum _{x = L(J_{k-1})}^{y-|J_{k-1}|} 1 \\ &= |J_{k+1}|\left (\frac{R(J_{k+1})+L(J_{k+1})}{2}-|J_{k-1}|-L(J_{k-1})+1\right ) \\ &= |J_{k+1}|\left (\frac{R(J_{k+1})+L(J_{k+1})}{2}-R(J_{k-1})\right ) \\ &= |J_{k+1}|\left (\frac{R(J_{k+1})+L(J_{k+1})}{2}-L(J_{k+1})+|J_k|+1\right ) \\ &= |J_{k+1}|\left (\frac{|J_{k+1}|-1}{2}+|J_k|+1\right ), \end{align*}

and conclude by observing that $|J_k| \ge 0$ and $\frac{|J_{k+1}|-1}{2}+1 \ge \frac{1}{2}|J_{k+1}|$ .

If, instead, $R(J_{k+1})-|J_{k-1}| \ge R(J_{k-1})$ , we obtain

(8) \begin{equation} \sum _{(n,n^{\prime}) \in J_{k-1}\times J_{k+1}} 1_{g_n+\dots +g_{n^{\prime}} \gt \frac{1}{2}} \ge \sum _{y=L(J_{k+1})}^{|J_{k-1}|+R(J_{k-1})} \sum _{x=L(J_{k-1})}^{y-|J_{k-1}|} 1 +\sum _{y=|J_{k-1}|+R(J_{k-1})+1}^{R(J_{k+1})} \sum _{x=L(J_{k-1})}^{R(J_{k-1})} 1. \end{equation}

The first double sum on the RHS of (8) is equal to

\begin{equation*}(|J_{k-1}|+R(J_{k-1})-L(J_{k+1})+1)\left (\frac {|J_{k-1}|+R(J_{k-1})+L(J_{k+1})}{2}-R(J_{k-1})\right ),\end{equation*}

while the second double sum is equal to

\begin{equation*}\left(R(J_{k+1})-|J_{k-1}|-R(J_{k-1})\right)|J_{k-1}|.\end{equation*}

Adding these two sums and simplifying we obtain,

\begin{align*} (|J_{k-1}| - |J_k|)\left(\frac {|J_{k-1}| + |J_k| + 1}{2}\right) & + |J_{k-1}|(|J_k| + |J_{k+1}|-|J_{k-1}|)\\& = -\frac {1}{2}(|J_{k-1}|-|J_k|)^2 + \frac {1}{2}(|J_{k-1}| - |J_k|) + |J_{k-1}||J_{k+1}|.\end{align*}

Now recall that we have $|J_k| \leq |J_{k+1}| \leq |J_{k-1}|$ and furthermore, by our assumption in the second case, we have $|J_{k-1}| \leq |J_k| + |J_{k+1}|$ . Thus we may obtain

\begin{equation*} -\frac {1}{2}(|J_{k-1}|-|J_k|)^2 + \frac {1}{2}(|J_{k-1}| - |J_k|) + |J_{k-1}||J_{k+1}| \geq -\frac {1}{2}|J_{k+1}|^2 + 0 + |J_{k+1}|^2 = \frac {1}{2}|J_{k+1}|^2.\end{equation*}

This completes the proof.

We now cease referring to a specific $I \in \mathcal{I}^N$ . To denote dependence on $I \in \mathcal{I}^N$ , we denote $I = \sqcup _{k=1}^r J_k^I$ its decomposition.

Proposition 4.5. For any $a \in \left(0,\frac{1}{2}\right)$ , it holds that

\begin{equation*}\sum _{I \in \mathcal {I}^N} \sum _k \text {PPC}^{J_k^I,J_{k+1}^I}(0,a) \le 2\sqrt {2}\epsilon ^{1/4}\left (\sum _{I \in \mathcal {I}^N} \sum _k |J_k^I|^2\right )^{1/2}\sqrt {N}\end{equation*}

and

\begin{equation*}\sum _{I \in \mathcal {I}^N} \sum _{k { \text { sandwiched}}} \text {PPC}^{J_{k-1}^I,J_{k+1}^I}(0,a) \le 2\sqrt {2}\epsilon ^{1/4}\left (\sum _{I \in \mathcal {I}^N} \sum _k |J_k^I|^2\right )^{1/2}\sqrt {N}.\end{equation*}

Proof. By trivially bounding $\text{PPC}^{J_k^I,J_{k+1}^I}(0,a) \le |J_{g_1(k)}^I|\hspace{1mm} |J_{b_1(k)}^I|$ and Cauchy-Schwarz,

\begin{align*} \sum _{I \in \mathcal{I}^N} \sum _k \text{PPC}^{J_k^I,J_{k+1}^I}(0,a) &\le \sum _{I \in \mathcal{I}^N} \sum _k |J_{g_1(k)}^I|\hspace{1mm} |J_{b_1(k)}^I| \\ &\le \left (\sum _{I,k} |J_{g_1(k)}^I|^2\right )^{1/2}\left (\sum _{I,k} |J_{b_1(k)}^I|^2\right )^{1/2} \\ &\le \left (2\sum _{I \in \mathcal{I}^N} \sum _k |J_k^I|^2\right )^{1/2}\left (\sum _{I \in \mathcal{I}^N} \sum _k 2 \text{PPC}^{J_k^I,J_{k+1}^I}\left(\frac{1}{2},\frac{3}{2}\right)\right )^{1/2}, \end{align*}

where the last inequality used Proposition4.3 together with the fact that $\textrm{sum}\left(J_k^I \cup J_{k+1}^I\right) \le 1 \le \frac{3}{2}$ . Now just observe

\begin{equation*}\sum _{I \in \mathcal {I}^N} \sum _k 2 \text {PPC}^{J_k^I,J_{k+1}^I}\left(\frac {1}{2},\frac {3}{2}\right) \le 2\sum _{n \le N} \sum _{2 \le m \le N-n+1} 1_{g_n+\dots +g_{n+m-1} \in (\frac {1}{2},\frac {3}{2})},\end{equation*}

which, by Proposition3.2, is at most $4\sqrt{\epsilon } N$ . The first inequality of the lemma follows.

For the second inequality of the lemma, we argue as above, except this time using Proposition4.4:

\begin{align*} &\sum _{I \in \mathcal{I}^N} \sum _{k \text{ sandwiched}} \text{PPC}^{J_{k-1}^I,J_{k+1}^I}(0,a)\\ &\qquad \qquad \qquad \le \sum _{I \in \mathcal{I}^N} \sum _{k \text{ sandwiched}} |J_{g_2(k)}^I|\hspace{1mm} |J_{b_2(k)}^I| \\ &\qquad \qquad \qquad \le \left (\sum _{I \in \mathcal{I}^N} \sum _{k \text{ sandwiched}} |J_{g_2(k)}^I|^2\right )^{1/2}\left (\sum _{I \in \mathcal{I}^N} \sum _{k \text{ sandwiched}} |J_{b_2(k)}^I|^2\right )^{1/2} \\ &\qquad \qquad \qquad \le \left (2\sum _{I \in \mathcal{I}^N} \sum _k |J_k^I|^2\right )^{1/2}\left (\sum _{I \in \mathcal{I}^N} \sum _k 2 \text{PPC}^{J_{k-1}^I,J_{k+1}^I}\left(\frac{1}{2},\frac{3}{2}\right)\right )^{1/2}, \end{align*}

where the last inequality used $\textrm{sum}(J_{k-1}\cup J_k\cup J_{k+1}) \le \frac{3}{2}$ . Now just observe

\begin{equation*}\sum _{I \in \mathcal {I}^N} \sum _k 2 \text {PPC}^{J_k^I,J_{k+1}^I}\left(\frac {1}{2},\frac {3}{2}\right) \le 2\sum _{n \le N} \sum _{2 \le m \le N-n+1} 1_{g_n+\dots +g_{n+m-1} \in (\frac {1}{2},\frac {3}{2})},\end{equation*}

which, by Proposition3.2, is at most $4\sqrt{\epsilon } N$ . The second inequality of the lemma follows.

We are ready to complete the proof of Proposition3.9.

Proposition 4.6. For all $N$ large,

\begin{equation*}\sum _{I \in \mathcal {I}^N}\sum _k \binom {|J_k^I|+1}{2} \geq \left(\frac {1}{2}-4\sqrt {2}\epsilon ^{1/4}\right)N.\end{equation*}

Proof. Using Lemma 4.2 we have that

\begin{align*}\left(\frac {1}{2}+o(1)\right)N & = \sum _{I\in \mathcal {I}^N}\text {PPC}^I\left(0,\frac {1}{2}\right) = \sum _I\sum _{k=1}^{r_I} \text {PPC}^{J^I_k}\left(0,\frac {1}{2}\right) +\sum _{k=1}^{r_I-1}\text {PPC}^{J^I_k,J^I_{k+1}}\left(0, \frac {1}{2}\right)\\& + \sum _{k \text { sandwiched}}\text {PPC}^{J_{k-1},J_{k+1}}\left(0,\frac {1}{2}\right).\end{align*}

Invoking Proposition4.5 in the first line we have:

\begin{align*} \left(\frac{1}{2}+o(1)\right)N-\sum _{I \in \mathcal{I}^N}\sum _k \binom{|J_k^I|+1}{2} &\le 4\sqrt 2 \epsilon ^{1/4}\left (\sum _{I \in \mathcal{I}^N} \sum _k |J_k^I|^2\right )^{1/2}\sqrt{N} \\ & \le 8\epsilon ^{1/4}\left (\sum _{I \in \mathcal{I}^N} \sum _k \binom{|J_k^I|+1}{2}\right )^{1/2}\sqrt{N}. \end{align*}

Writing

\begin{equation*}\sum _{I \in \mathcal {I}^N} \sum _k \binom {|J_k^I|+1}{2} = \left(\frac {1}{2}-\delta \right) N,\end{equation*}

we see

\begin{equation*}\delta +o(1)\le 8 \epsilon ^{1/4}\left(\frac {1}{2} - \delta \right)^{1/2},\end{equation*}

which yields (after a little computation) $\delta \le 4\sqrt{2}\epsilon ^{1/4}$ for $N$ large enough.

5. Appendix: proof of Lemma 3.3

We restate Lemma 3.3 for the reader’s convenience.

Lemma 3.3 (Lemma 3.3). For positive integers $a,b,c,L$ satisfying $1 \le a \le b \le c \le L$ , we have

\begin{align*}(a-1)a & + (b-a)(b-a+1)+(c-b)(c-b+1)+(L-c)(L-c+1)\\& +(a-1)(b-a)+(b-a)(c-b)+(c-b)(L-c) \ge \frac {5}{12}L^2+\frac {1}{6}L-\frac {7}{12}.\end{align*}

Proof. Fix $L \ge 1$ . It clearly suffices to prove the inequality for all real numbers $a,b,c$ satisfying $1 \le a \le b \le c \le L$ . By compactness, we may work with a triple $(a,b,c)$ that achieves the minimum value of the left hand side (which we denote $\textrm{LHS}$ ) minus the right hand side (which we denote $\textrm{RHS}$ ). We divide into three cases.

Case 1: $c=L$ .

As one may compute, $\frac{\partial }{\partial c}[\textrm{LHS}-\textrm{RHS}] = -a+2c-L = L-a$ .

Subcase 1: $a=L$ . Then $b=L$ , which gives $\textrm{LHS}-\textrm{RHS} = \frac{28L^2-51L+28}{48}$ , which is non-negative, since it is equal to $5$ at $L=1$ and has derivative $56L-51$ , which is positive for $L \ge 1$ .

Subcase 2: $a \not = L$ . Then $\frac{\partial }{\partial c}[\textrm{LHS}-\textrm{RHS}] \gt 0$ , so since $(a,b,c)$ is a minimizer, we must have $b=c$ , for otherwise we can decrease $c$ a bit to decrease $\textrm{LHS}-\textrm{RHS}$ . Thus, $\textrm{LHS}-\textrm{RHS} = a^2-(L+1)a+\frac{28L^2-3L+28}{48}$ , which has minimum occurring at $a = \frac{L+1}{2}$ , which gives $\textrm{LHS}-\textrm{RHS} = \frac{1}{3}L^2-\frac{9}{16}L+\frac{1}{3}$ , which is always non-negative, since it is at $L=1$ and the derivative is $\frac{2}{3}L-\frac{9}{16}$ , which is non-negative for $L \ge 1$ .

Case 2: $c \not = L$ and $b = c$ .

Since $c \not = L$ and $(a,b,c)$ is a minimizer, we must have $\frac{\partial }{\partial c}[\textrm{LHS}-\textrm{RHS}] \ge 0$ , for otherwise we could increase $b,c$ a bit to decrease $\textrm{LHS}-\textrm{RHS}$ . Recall $\frac{\partial }{\partial c}[\textrm{LHS}-\textrm{RHS}] = -a+2c-L$ ; so, $2c-L \ge a$ .

Subcase 1: $a=b$ . In this case, $\textrm{LHS}-\textrm{RHS} = 2c^2-(2L+2)c+\frac{7L^2+10L+7}{12}$ , which has minimum at $c = \frac{L+1}{2}$ , which yields $\textrm{LHS}-\textrm{RHS} = \frac{(L-1)^2}{12}$ , which is non-negative.

Subcase 2: $a \not = b$ . In this case, we must have $\frac{\partial }{\partial b}[\textrm{LHS}-\textrm{RHS}] \le 0$ , since otherwise we could decrease $b$ a bit to decrease $\textrm{LHS}-\textrm{RHS}$ . Note $\frac{\partial }{\partial b}[\textrm{LHS}-\textrm{RHS}] = -1+2b-L$ , so since $2b-L \ge a$ , we must have $a=1$ and $2b-L = 1$ . So, we have $a=1,b=\frac{L+1}{2},c=\frac{L+1}{2}$ , which indeed has $\textrm{LHS}-\textrm{RHS} \ge 0$ .

Case 3: $c \not = L$ and $b \not = c$ .

In this case, $\frac{\partial }{\partial c}[\textrm{LHS}-\textrm{RHS}]$ must be $0$ , for otherwise we could perturb $c$ a bit to decrease $\textrm{LHS}-\textrm{RHS}$ . So, $-a+2c-L = 0$ .

Subcase 1: $a = 1$ . Having $a=1$ and $-a+2c-L=0$ , i.e., $c = \frac{L+1}{2}$ , yields $\textrm{LHS}-\textrm{RHS} = \frac{1}{4}[4b^2-4(L+1)b+3L^2+2L-1]$ , which is minimized as $b = \frac{L+1}{2}$ , which was dealt with in Subcase 2 of Case 2.

Subcase 2: $a = b$ . Then, $48(\textrm{LHS}-\textrm{RHS}) = 84b^2-(72L+96)b+16L^2+45L+28$ , which has minimum at $b = \frac{72L+96}{84}$ , which gives $48(\textrm{LHS}-\textrm{RHS}) = 16L^2+45L+28$ , which is clearly non-negative.

Subcase 3: $a \not \in \{1,b\}$ . Then $\frac{\partial }{\partial a}[\textrm{LHS}-\textrm{RHS}] = 0$ , for otherwise we could perturb $a$ a bit to decrease $\textrm{LHS}-\textrm{RHS}$ . So, $-1+2a-c = 0$ . Together with $-a+2c-L = 0$ yields $a = \frac{L+2}{3}, c = \frac{2L+1}{3}$ , which yields $\textrm{LHS}-\textrm{RHS} = b^2-(L+1)b+\frac{12L^2+29L+12}{48}$ , which is minimized at $b = \frac{L+1}{2}$ , which yields $\frac{5L}{48}$ , which is clearly non-negative.

Acknowledgments

We would like to thank our advisor Ben Green for suggesting this problem to us. We would also like to thank Zach Hunter for providing feedback on an earlier version of this document.

Footnotes

*

Zachary Chase is partially supported by Ben Green’s Simons Investigator Grant 376201 and gratefully acknowledges the support of the Simons Foundation.

References

Aistleitner, C., Blomer, V. and Radziwiłł, M. (2024) Triple correlation and long gaps in the spectrum of flat tori. J. Eur. Math. Soc. 26(1) 4174.CrossRefGoogle Scholar
Aistleitner, C., Lachmann, T. and Pausinger, F. (2018) Pair correlations and equidistribution. J. Num. Theory 182 206220.CrossRefGoogle Scholar
Aistleitner, C., Larcher, G. and Lewko, M. (2017) Additive energy and the Hausdorff dimension of the exceptional set in metric pair correlation problems, with an Appendix by Jean Bourgain. Israel J. Math. 222(1) 463485.CrossRefGoogle Scholar
Eskin, A., Margulis, G. and Mozes, S. (2005) Quadratic forms of signature (2, 2) and eigenvalue spacings on rectangular 2-tori. Ann. Math. 161 679725.CrossRefGoogle Scholar
Heath-Brown, D. R. (2010) Pair correlation for fractional parts of $\alpha n^2$ , Math. Proc. Cambridge Philos. Soc. 148 385407.CrossRefGoogle Scholar
Kuipers, L. and Niederreiter, H. (2006) Uniform distribution of sequences. Dover Publications, Mineola, NY Google Scholar
Rudnick, Z. and Sarnak, P. (1998) The pair correlation function of fractional parts of polynomials. Comm. Math. Phys. 194(1) 6170.CrossRefGoogle Scholar
Rudnick, Z. and Zaharescu, A. (1999) A metric result on the pair correlation of fractional parts of sequences,. Acta Arith. 89(3) 283293.CrossRefGoogle Scholar
Rudnick, Z. and Zaharescu, A. (2002) The distribution of spacings between fractional parts of lacunary sequences. Forum Math. 14(5) 691712.CrossRefGoogle Scholar
Sarnak, P. (1997) Values at integers of binary quadratic forms. Harmonic analysis and number theory (Montreal 1996), 181-203, CMS Conf. Providence, RI.: Amer. Math. Soc., Proc. 21.Google Scholar
Steinerberger, S. (2018) Poissonian pair correlation and discrepancy,. Indag. Math. 29 11671178.CrossRefGoogle Scholar
Walker, A. (2018) The Primes are not metric Poissonian. Mathematika 64, 230236.CrossRefGoogle Scholar