Hostname: page-component-7bb8b95d7b-cx56b Total loading time: 0 Render date: 2024-09-25T06:32:43.211Z Has data issue: false hasContentIssue false

The error term in the truncated Perron formula for the logarithm of an L-function

Published online by Cambridge University Press:  09 March 2023

Stephan Ramon Garcia
Affiliation:
Department of Mathematics and Statistics, Pomona College, 610 North College Avenue, Claremont, CA 91711, USA e-mail: stephan.garcia@pomona.edu
Jeffrey Lagarias
Affiliation:
Department of Mathematics, University of Michigan, 530 Church Street, Ann Arbor, MI 48109, USA e-mail: lagarias@umich.edu
Ethan Simpson Lee*
Affiliation:
School of Mathematics, University of Bristol, Fry Building, Woodland Road, Bristol BS8 1UG, UK
Rights & Permissions [Opens in a new window]

Abstract

We improve upon the traditional error term in the truncated Perron formula for the logarithm of an L-function. All our constants are explicit.

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

1 Introduction

The truncated Perron formula relates the summatory function of an arithmetic function to a contour integral that may be estimated using techniques from complex analysis. Let $F(s) = \sum _{n=1}^{\infty } f(n) n^{-s}$ be absolutely convergent on $\operatorname {Re} s> c_F$ ; examples include the Riemann zeta function, Dirichlet L-functions, the Dedekind zeta function associated with a number field, and Artin L-functions. The truncated Perron formula tells us that if $x>0$ is not an integer, $T\geq 1$ , and $c> c_F$ , then

(1.1) $$ \begin{align} \sum_{n \leq x} f(n) = \frac{1}{2\pi i }\int_{c-iT}^{c+iT} F(s) \frac{x^s}{s}\,ds + O^*\bigg( \sum_{n=1}^{\infty} \bigg( \frac{x}{n} \bigg)^{c} |f(n)| \min\left\{ 1, \frac{1}{T |\log \frac{x}{n}|}\right\}\bigg), \end{align} $$

in which $O^*(g(x)) = h(x)$ means $|h(x)|\leq g(x)$ (see [Reference Koukoulopoulos8, Chapter 7], [Reference Montgomery and Vaughan10, Section 5.1], [Reference Murty11, Example 4.4.15], and [Reference Tenenbaum15, Section II.2]). We let T depend on x, and let $c = c_F + 1/\log {x}$ , so that $x^c = ex^{c_F}$ . A variation of (1.1) improves the order of the error term by truncating the integral at $\pm T^*$ for an unknown $T^*\in [T,O(T)]$ [Reference Cully-Hugill and Johnston3], although this is inconvenient if one must avoid $T^*$ that correspond to the ordinates of nontrivial zeros of $F(s)$ . The authors of [Reference Cully-Hugill and Johnston3] have also informed us in a personal communication that their paper inherited an unfortunate typo from another paper, so the error term in their variation of the truncated Perron formula could be worse by a factor of $\log {x}$ ; this means that our main result (Theorem 1.1) will be comparable in strength and more straightforward to apply when compared against the outcome of their result.

For $\operatorname {Re}{s}> 1$ , the logarithm of the Riemann zeta function $\zeta (s) = \sum _{n=1}^{\infty } n^{-s}$ is $\log \zeta (s) = \sum _{n=1}^{\infty } \Lambda (n)(\log n)^{-1}n^{-s}$ , in which $\Lambda (n)$ is the von Mangoldt function. The logarithm of a typical L-function is of the form $\sum _{n=1}^{\infty } \Lambda (n)a_n(\log n)^{-1}n^{-s}$ , in which the $a_n$ are easily controlled. For example, $|a_n| \leq 1$ for Dirichlet L-functions and $|a_n|\leq d$ for Artin L-functions of degree d (see [Reference Iwaniec and Kowalski7, Chapter 5]. In these cases, the error term in (1.1) with $c = 1/\log {x}$ is on the order of

(1.2) $$ \begin{align} \sum_{n=2}^{\infty} \Big( \frac{x}{n}\Big)^{\frac{1}{\log x}} \frac{ \Lambda(n)}{n \log n} \min\left\{ 1 , \frac{1}{T | \log \frac{x}{n} |} \right\} = O\bigg( \frac{\log x}{T} \bigg). \end{align} $$

Granville and Soundararajan used ( 1.2 ) with Dirichlet L-functions to study large character sums [Reference Granville and Soundararajan6, equation (8.1)]. Cho and Kim applied it to Artin L-functions to obtain asymptotic bounds on Dedekind zeta residues [Reference Cho and Kim1, Proposition 3.1]. A bilinear relative of (1.2) appears in Selberg’s work on primes in short intervals [Reference Selberg14, Lemma 4]. Analogous sums arise with the logarithmic derivative of an L-function in [Reference Davenport4, p. 106] and [Reference Patterson12, p. 44].

We improve upon (1.2) asymptotically and explicitly in the following result.

Theorem 1.1 If $x \geq 3.5$ is a half integer and $T \geq (\log \frac {3}{2})^{-1}> 2.46$ , then

(1.3) $$ \begin{align} \sum_{n=2}^{\infty} \Big( \frac{x}{n}\Big)^{\frac{1}{\log x}} \frac{ \Lambda(n)}{n \log n} \min\left\{ 1 , \frac{1}{T | \log \frac{x}{n} |} \right\} \leq \frac{R(x)}{T}, \end{align} $$

in which

(1.4) $$ \begin{align} R(x) = 40.23 \log \log x +58.12 +\frac{3.87}{\log x} +\frac{5.22 \log x}{\sqrt{x}} -\frac{1.84}{\sqrt{x}}. \end{align} $$

Our result has a wide and explicit range of applicability. For example, the following corollary employs (1.1) with $T=x$ and $c = 1/\log {x}$ . Since one can use analytic techniques to see the integral below is asymptotic to $\log {L(1,\chi )}$ , one can relate $\log {L(1,\chi )}$ to a short sum. We hope to do so explicitly in the future.

Corollary 1.2 Let $L(s, \chi )$ be an entire Artin L-function of degree d such that

$$ \begin{align*} L(s,\chi) = \prod_{p}\prod_{i=1}^{d} \left(1 - \frac{\alpha_i(p)}{p^s}\right)^{-1} \quad \text{for }\operatorname{Re} s> 1 \end{align*} $$

with $a(p^k) = \alpha _1(p)^k + \cdots + \alpha _d(p)^k$ for prime p. Then, with $R(x)$ as in ( 1.4 ), we have

$$ \begin{align*} \sum_{1<n < x} \frac{\Lambda(n) a(n)}{n \log{n}} = \frac{1}{2\pi i }\int_{\tfrac{1}{\log{x}}-ix}^{\tfrac{1}{\log{x}}+ix} \frac{x^s}{s} \log{L(1+s,\chi)}\,ds + O^*\!\left( \frac{d\,R(x)}{x} \right). \end{align*} $$

2 Preliminaries

Here, we establish several lemmas needed for the proof of Theorem 1.1.

Lemma 2.1 If $\sigma>0$ , then $\log \zeta (1+\sigma ) \leq - \log \sigma + \gamma \sigma $ .

Proof For $s> 1$ , we have $\zeta (s) \leq e^{\gamma (s-1)}/(s-1)$ [Reference Ramaré13, Lemma 5.4]. Let $s = 1+\sigma $ and take logarithms to obtain the desired result.

For real $z,w$ , the equation $z = we^w$ can be solved for w if and only if $z \geq -e^{-1}$ . There are two branches for $-e^{-1} \leq z < 0$ . The lower branch defines the Lambert $W_{-1}(z)$ function [Reference Corless, Gonnet, Hare, Jeffrey and Knuth2], which decreases to $-\infty $ as $z\to 0^-$ (see Figure 1a). For $n \geq 6>2e$ , we define the strictly increasing sequence

(2.1) $$ \begin{align}\qquad y_n = \frac{-n}{2} W_{-1}\Big( \frac{-2}{n} \Big) \qquad \text{for }n \geq 6. \end{align} $$

Figure 1 Graphs relevant to the construction of the sequence $y_n$ .

Lemma 2.2 For $n \geq 8$ , we have $\frac {2y_n}{\log y_n} = n$ and $ y_n \geq \frac {n}{2} \log n$ .

Proof For $n \geq 6$ , the definition of $W_{-1}$ and (2.1) confirm that $\frac {2y_n}{\log y_n} = n$ . Thus, the desired inequality is equivalent to $W_{-1}(\frac {-2}{n} ) \leq -\log n$ . Since $f(w) = we^w$ decreases on $(-\infty ,-1]$ (Figure 1b) and $-\frac {1}{e} < -\frac {2}{n} <0$ , the desired inequality is equivalent to

$$ \begin{align*} -\frac{2}{n} \geq f(-\log n) = (-\log n)e^{-\log n} = -\frac{\log n}{n}, \end{align*} $$

which holds whenever $\log n \geq 2$ . This occurs for $n \geq e^2 \approx 7.38906$ .

Remark 2.3 For all $-e^{-1} \leq x < 0$ , the bound $W_{-1}(x) \leq \log (-x) - \log (-\log (-x))$ is valid (see [Reference Lóczi9, equations (8) and (39)]). It follows from this observation and (2.1) that

$$ \begin{align*} y_n \geq \frac{n}{2} \left(\log\left(\frac{1}{2}\log\frac{n}{2}\right) + \log{n} \right), \end{align*} $$

which also implies Lemma 2.2 for $n\geq 15$ .

The next lemma is needed later to handle a few exceptional primes.

Lemma 2.4 Let $x>1$ be a half integer, and let $C = \frac {1284699552}{444215525}= 2.89206\ldots $ .

  1. 1. Let $p_{-8} < p_{-7} < \cdots < p_{-1} < x$ denote the largest eight primes (if they exist) in the interval $(\frac {x}{2},x)$ . We have the sharp bound

    (2.2) $$ \begin{align} F_1(x) = \sum_{1 \leq n \leq 8} \frac{1}{ x-p_{-n} } \leq C \end{align} $$
    (see Figure 2a). The corresponding summand in ( 2.2 ) is zero if $p_{-n}$ does not exist.
  2. 2. Let $x<p_1 <p_2<\cdots < p_8$ denote the smallest eight primes (if they exist) in the interval $(x,\frac {3x}{2})$ . We have the sharp bound

    (2.3) $$ \begin{align} F_2(x) = \sum_{1 \leq n \leq 8} \frac{1}{p_n -x} \leq C \end{align} $$
    (see Figure 2a). The corresponding summand in ( 2.3 ) is zero if $p_{n}$ does not exist.

Figure 2 The functions $F_1(x)$ and $F_2(x)$ behave erratically.

Proof (a) If $x \geq 10.5$ , then $2,3,5 \notin (\frac {x}{2},x)$ . Computation confirms that

$$\begin{align*}F_1(x) \leq F_1(3.5) = \frac{8}{3} =2.66\ldots\end{align*}$$

for $x \leq 9.5$ . If $x \geq 10.5$ , then any prime in $(\frac {x}{2},x)$ is congruent to one of $1, 7, 11, 13, 17, 19, 23, 29 \pmod {30}$ . There are finitely many patterns modulo $30$ that the $p_{-8},p_{-7},\ldots ,p_{-1}$ may assume. Among these, computation confirms that $F_1(x)$ is maximized if

$$ \begin{align*} p_{-1} &= \lfloor x \rfloor \equiv 19\pmod{30}, & p_{-5} &= \lfloor x \rfloor-12 \equiv 7\pmod{30},\\ p_{-2} &= \lfloor x \rfloor-2 \equiv 17\pmod{30}, & p_{-6} &= \lfloor x \rfloor-18 \equiv 1\pmod{30},\\ p_{-3} &= \lfloor x \rfloor-6 \equiv 13\pmod{30}, & p_{-7} &= \lfloor x \rfloor-20 \equiv 29\pmod{30},\\ p_{-4} &= \lfloor x \rfloor-8 \equiv 11\pmod{30}, & p_{-8} &= \lfloor x \rfloor-26 \equiv 23\pmod{30}, \end{align*} $$

which yields the desired upper bound C. This prime pattern first occurs for $x = 88,819.5$ (see https://oeis.org/A022013).

(b) If $x \geq 5.5$ , then $2,3,5 \notin (x, \frac {3x}{2})$ . Observe that $F_2(x) \leq 2$ for $x \leq 4.5$ (attained at $x=1.5, 2.5, 4.5$ ). If $x \geq 5.5$ , then (as in (a)), any prime in $(x,\frac {3x}{2})$ is congruent to one of $1, 7, 11, 13, 17, 19, 23, 29 \pmod {30}$ . It follows that $F_2(x)$ is maximized if

$$ \begin{align*} p_{1} &= \lceil x \rceil \equiv 11\pmod{30}, & p_{5} &= \lceil x \rceil+12 \equiv 23\pmod{30},\\ p_{2} &= \lceil x \rceil+2 \equiv 13\pmod{30}, & p_{6} &= \lceil x \rceil+18 \equiv 29\pmod{30},\\ p_{3} &= \lceil x \rceil+6 \equiv 17\pmod{30}, & p_{7} &= \lceil x \rceil+20 \equiv 1\pmod{30}, \\ p_{4} &= \lceil x \rceil+8 \equiv 19\pmod{30}, & p_8 &= \lceil x \rceil + 26 \equiv 7 \pmod{30}, \end{align*} $$

which yields the desired upper bound C. This prime pattern occurs for $x=10.5$ , but not all eight primes lie in $(x,\frac {3}{2}x)$ . Therefore, the first admissible value is $x = 15,760,090.5$ (see https://oeis.org/A022011).

We also need an elementary estimate on kth powers in intervals.

Lemma 2.5 Let $X> 1$ be a noninteger, $h>1$ , and $k\geq 2$ .

  1. (1) There are at most $N_k + 1$ perfect kth powers in $[X,X+h)$ , in which $N_k \leq \frac {h}{k \sqrt {X}}$ .

  2. (2) The shortest gap between kth powers in $[X,X+h)$ (if they exist) is $G_k \geq k \sqrt {X}$ .

Proof We may assume that X is so large that $N_k \geq 1$ . Let $m = \lceil X^{\frac {1}{k}} \rceil $ so that $m^k$ is the first kth power larger than X. Consider the gaps $g_1,g_2,\ldots ,g_{N_k}$ between the $N_k$ consecutive kth powers in $[X,X+h)$ (see Figure 3). Then

$$ \begin{align*} G_k = \min\{ g_1,g_2,\ldots,g_{N_k}\} = g_1 = (m+1)^k - m^k \geq km^{k-1} \geq k X^{\frac{k-1}{k}} \geq k \sqrt{X}. \end{align*} $$

The desired inequality follows since $N_k G_k \leq g_1 + g_2+ \cdots + g_{N_k} \leq h$ .

Figure 3 Proof of Lemma 2.5.

Finally, we need an estimate on the nth harmonic number $H_n = \sum _{j=1}^n \frac {1}{j}$ :

(2.4) $$ \begin{align}\qquad \frac{1}{2n+ \frac{2}{5}} < H_n - \log n - \gamma < \frac{1}{2n + \frac{1}{3}} \leq \frac{3}{7} \quad\text{for}\quad n \geq 1, \end{align} $$

in which $\gamma $ is the Euler–Mascheroni constant [Reference Tóth and Mare16]. We require the upper bound

(2.5) $$ \begin{align} \sum_{\ell=1}^n \frac{1}{2\ell - 1} &= \left(1 + \frac{1}{2} + \frac{1}{3} + \cdots + \frac{1}{2n}\right) - \left( \frac{1}{2} + \frac{1}{4} + \cdots + \frac{1}{2n} \right) = H_{2n} - \frac{1}{2} H_n \nonumber \\ &\leq \bigg(\log 2n +\gamma + \frac{1}{4n + \frac{1}{3}} \bigg) - \frac{1}{2} \bigg( \log n + \gamma + \frac{1}{2n+\frac{2}{5}} \bigg) \nonumber \\ &\leq \frac{1}{2}\log n + \frac{1}{2}\gamma + \log 2 + \frac{1}{4n + \frac{1}{3}}- \frac{1}{4n+\frac{4}{5}} \nonumber \\ &= \frac{1}{2}\log n + \frac{1}{2}\gamma + \log 2 + \frac{7}{240 n^2+68 n+4} \nonumber\\ &\leq \frac{1}{2}\log n + \frac{1}{2}\gamma + \log 2 + \frac{7}{312} \quad \text{for}\quad n \geq 1. \end{align} $$

3 Proof of Theorem 1.1

In what follows, $x\in \mathbb {N}+ \frac {1}{2} = \{ \frac {3}{2}, \frac {5}{2}, \frac {7}{2},\ldots \}$ and $c= \frac {1}{\log x}$ . Minor improvements below are possible; these were eschewed in favor of a final estimate of simple shape.

3.1 When n is very far from x

Suppose that $n \leq \frac {x}{2}$ or $\frac {3x}{2} \geq n$ . Then $\log \frac {x}{n} \leq -\log \frac {3}{2}$ or $\log \frac {3}{2} <\log 2 \leq \log \frac {x}{n}$ , so $|\log \frac {x}{n}| \geq \log \frac {3}{2}$ . If $T \geq (\log \frac {3}{2})^{-1}> 2.46$ , then

$$ \begin{align*} \min\left\{ 1 , \frac{1}{T | \log \frac{x}{n}| } \right\} \leq \frac{1}{T |\log \frac{x}{n}|} \leq \frac{1}{T \log \frac{3}{2}}. \end{align*} $$

For such T, the previous inequality and Lemma 2.1 imply (recall that $c = \frac {1}{\log x}$ )

(3.1) $$ \begin{align} \sum_{\substack{n \leq \frac{x}{2}\, \text{or}\\ n \geq \frac{3x}{2}}} \Big( \frac{x}{n}\Big)^c \frac{ \Lambda(n) }{n \log n} \min\Big\{ 1 , \frac{1}{T | \log \frac{x}{n} |} \Big\} &\leq x^c \sum_{\substack{n \leq \frac{x}{2}\, \text{or}\\ n \geq \frac{3x}{2} }} \frac{1}{n^{1+c}}\bigg( \frac{\Lambda(n)}{\log n}\bigg) \bigg(\frac{1}{T \log \frac{3}{2} } \bigg) \nonumber \\ &\leq \frac{ x^c}{T\log \frac{3}{2}} \log \zeta(1+c) \nonumber \\ &\leq \frac{ x^c}{T \log \frac{3}{2}} ( - \log c + \gamma c) \nonumber \\ &= \frac{e}{T \log \frac{3}{2}}\bigg(\log \log x + \frac{\gamma}{\log x} \bigg). \end{align} $$

3.2 Reduction to a sum over prime powers

Suppose that $\frac {x}{2} < n < \frac {3x}{2}$ . Let $z = 1-\frac {n}{x}$ and observe that $|z| < \frac {1}{2}$ . Then

$$ \begin{align*} \log \frac{x}{n} = - \log(1-z) = z \left(- \frac{\log(1-z)}{z} \right), \end{align*} $$

in which the function in parentheses is positive and achieves its minimum value $2 \log \frac {3}{2} = 0.81093\ldots $ on $|z| < \frac {1}{2}$ at its left endpoint $-\frac {1}{2}$ (see Figure 4a). Then

(3.2) $$ \begin{align} | \log(1-z) |> \bigg(2 \log \frac{3}{2} \bigg) |z| \quad\text{for}\quad |z| < \frac{1}{2}, \end{align} $$

whose validity is illustrated in Figure 4b. Therefore,

(3.3) $$ \begin{align} \bigg| \log \frac{x}{n} \bigg|> \bigg( 2 \log \frac{3}{2}\bigg) \bigg|1 - \frac{n}{x}\bigg| \quad\text{for}\quad \frac{x}{2} < n < \frac{3x}{2}, \end{align} $$

Figure 4 Graphs relevant to the derivation of (3.2).

and hence

(3.4) $$ \begin{align} & \sum_{\frac{x}{2} < n < \frac{3x}{2}} \Big( \frac{x}{n}\Big)^c \frac{ \Lambda(n) }{n \log n} \min\left\{ 1 , \frac{1}{T | \log \frac{x}{n} |} \right\} \nonumber \\&\qquad\leq \frac{x^c}{T} \sum_{ \frac{x}{2} < n < \frac{3x}{2}} \bigg( \frac{ \Lambda(n) }{n^{1+c} \log n}\bigg) \bigg( \frac{1}{ | \log \frac{x}{n} |} \bigg) \nonumber \\&\qquad\leq \frac{x^c}{T} \sum_{\frac{x}{2} < n < \frac{3x}{2}} \bigg( \frac{ \Lambda(n) }{n^{1+c} \log n}\bigg) \frac{1}{ (2\log \frac{3}{2}) | 1 - \frac{n}{x} |} && (\text{by (3.3)}) \nonumber \\ &\qquad\leq \frac{ x^c}{T (2\log\frac{3}{2}) } \sum_{\frac{x}{2} < n < \frac{3x}{2}} \frac{2}{x} \bigg( \frac{ \Lambda(n) }{n^c \log n}\bigg)\frac{1}{ | 1 - \frac{n}{x} |} && (\text{since }\tfrac{x}{2} < n) \nonumber \\ &\qquad\leq \frac{x^c }{T \log \frac{3}{2}} \sum_{\frac{x}{2} < n < \frac{3x}{2}} \bigg( \frac{ \Lambda(n) }{n^c \log n}\bigg)\frac{1}{ | x-n |} \nonumber \\ &\qquad\leq \frac{x^c }{T \log\frac{3}{2}} \sum_{\frac{x}{2} < p^k < \frac{3 x}{2}} \bigg( \frac{ \log p }{(p^k)^c k\log p}\bigg)\frac{1}{ | x-p^k |} && (\text{def. of }\Lambda)\nonumber \\ &\qquad\leq \frac{e }{T \log\frac{3}{2}} \sum_{\frac{x}{2} < p^k < \frac{3x}{2}} \frac{1}{k | x-p^k |} &&(\text{since }c=\tfrac{1}{\log x}), \end{align} $$

in which the final two sums run over all prime powers $p^k$ in the stated interval.

The remainder of the proof uses ideas from [Reference Goldston5, Lemma 2] to estimate

(3.5) $$ \begin{align} \sum_{\frac{x}{2} < p^k < \frac{3x}{2}} \frac{1}{ k | x-p^k |} = \underbrace{\sum_{\frac{x}{2} < p < \frac{3x}{2}} \frac{1}{ | x-p |} }_{S_{\mathrm{prime}}(x)} + \underbrace{ \sum_{ \substack{\frac{x}{2} < p^k < \frac{3x}{2}\\ k\geq 2}} \frac{1}{ k | x-p^k |} }_{S_{\mathrm{power}}(x)}. \end{align} $$

3.3 The sum over primes

First observe that

$$ \begin{align*} S_{\mathrm{prime}}(x) \leq \underbrace{\sum_{\frac{x}{2} < p < x} \frac{1}{ x-p } }_{S_{\mathrm{prime}}^-(x)} + \underbrace{\sum_{x < p < \frac{3x}{2}} \frac{1}{ p-x} }_{S_{\mathrm{prime}}^+(x)}. \end{align*} $$

We require the Brun–Titchmarsh theorem (see [Reference Montgomery and Vaughan10, Corollary 2]):

(3.6) $$ \begin{align} \pi(X+Y) - \pi(X) \leq \frac{2Y}{\log{Y}},\quad\text{where }\pi(x) = \sum_{p\leq x}1, X> 0\text{, and }Y > 1. \end{align} $$

3.3.1 The lower sum over primes

Let $p_{-k} < p_{-(k-1)} < \cdots < p_{-2} < p_{-1}$ be the primes in $(\frac {x}{2}, x)$ ; note that $k \leq \frac {x}{2}$ . Apply (3.6) with $X = x-y_n$ and $Y = y_n$ to get

$$ \begin{align*}\qquad 0 \leq \pi(x) - \pi(x-y_n) \leq \frac{2y_n}{\log y_n} = n \qquad \text{for }6 \leq n \leq k \end{align*} $$

by Lemma 2.2, so $(x-y_n,x]$ contains at most n primes. Thus, $p_{-(n+1)} \leq x - y_n$ and

(3.7) $$ \begin{align} \frac{1}{x-p_{-(n+1)}} \leq \frac{1}{y_n} \qquad \text{for }6 \leq n \leq k-1. \end{align} $$

Then Lemma 2.2, which requires $k \geq 8$ , and the integral test provide

$$ \begin{align*} \sum_{ \frac{x}{2} < p < x} \frac{1}{ x-p } &=\sum_{ 1 \leq n\leq 8} \frac{1}{ x-p_{-n} } + \sum_{9 \leq n \leq k} \frac{1}{ x-p_{-n} } \\&\leq F_1(x) + \sum_{8\leq n \leq k-1} \frac{1}{y_n} && (\text{by (2.2) and (3.7)})\\&\leq C + 2 \sum_{7 < n \leq \frac{x}{2}} \frac{1}{n \log n} && (\text{by Lemma 2.2})\\&\leq C + 2 \log \log x, \end{align*} $$

which is valid for $k \leq 7$ since Lemma 2.4a shows that the sum is majorized by C.

3.3.2 The upper sum over primes

Let $p_1<p_2< \cdots < p_k$ denote the primes in $(x, \frac {3x}{2})$ and note that $k \leq \frac {x}{2}$ . Then (3.6) with $X = x$ and $Y = y_n$ ensures that

$$ \begin{align*} 0 \leq \pi(x+y_n) - \pi(x) \leq \frac{2y_n}{\log y_n} = n \qquad \text{for}\qquad 6 \leq n \leq k \end{align*} $$

by Lemma 2.2, so $(x, x+y_n]$ contains at most n primes. Thus, $p_{n+1} \geq x + y_n$ and

(3.8) $$ \begin{align} \frac{1}{p_{n+1} - x} \leq \frac{1}{y_n} \qquad \text{for}\qquad 6 \leq n \leq k. \end{align} $$

An argument similar to that above reveals that

$$ \begin{align*} \sum_{x < p < \frac{3x}{2} } \frac{1}{p-x} \leq \sum_{1\leq n \leq 8} \frac{1}{p_n -x} + \sum_{9\leq n \leq k} \frac{1}{p_n-x } \leq C +2 \log\log x. \end{align*} $$

3.3.3 Final bound over primes

For $x \in \mathbb {N} + \frac {1}{2}$ , the previous inequalities yield

(3.9) $$ \begin{align} S_{\mathrm{prime}}(x) = \sum_{ \frac{x}{2} < p < \frac{3x}{2}} \frac{1}{ | x-p |} \leq 2C +4 \log\log x. \end{align} $$

3.4 The sum over prime powers

We now majorize

$$ \begin{align*} S_{\mathrm{power}}(x) = \sum_{ \substack{ \frac{x}{2} < p^k <\frac{3x}{2} \\ k\geq 2}} \frac{1}{ k | x-p^k |}. \end{align*} $$

3.4.1 Initial reduction

To bound $S_{\mathrm {power}}(x)$ it suffices to majorize

(3.10) $$ \begin{align} S_{\mathrm{sqf}}(x) = \sum_{ \substack{ \frac{x}{2} < n^k <\frac{3x}{2} \\ k\geq 2 \\ n \geq 2 \text{sq.~free} }} \frac{1}{ k | x-n^k |}, \end{align} $$

in which the prime powers $p^k$ are replaced with the powers $n^k$ of square free $n \geq 2$ . The square-free restriction ensures that powers such as $2^6 = (2^2)^3 = (2^3)^2$ are not counted multiple times in (3.10). If $\frac {x}{2} < n^k < \frac {3x}{2}$ and $k \geq 2$ , then (since $n \geq 2$ )

(3.11) $$ \begin{align} k \leq \frac{\log \frac{3x}{2} }{\log 2} \leq \lfloor 2.4 \log x \rfloor \qquad \text{for}\qquad x \geq 3.5. \end{align} $$

3.4.2 Nearest-power sets

The largest contributions to $S_{\mathrm {sqf}}(x)$ come from the powers closest to x. We handle those summands separately and split the sum (3.10) accordingly. For each $k \geq 2$ , the inequalities $\lfloor x^{\frac {1}{k}} \rfloor ^k < x < \lceil x^{\frac {1}{k}} \rceil ^k$ exhibit the two kth powers nearest to x. Define

(3.12) $$ \begin{align} \mathcal{N}_k \subseteq \big\{ \lfloor x^{\frac{1}{k}} \rfloor^k , \lceil x^{\frac{1}{k}} \rceil^k \big\} \end{align} $$

according to the following rules:

  • $\mathcal {N}_k$ contains $\lfloor x^{\frac {1}{k}} \rfloor ^k$ if it is square free and belongs to $( \frac {x}{2}, \frac {3x}{2} )$ .

  • $\mathcal {N}_k$ contains $\lceil x^{\frac {1}{k}} \rceil ^k$ if it is square free and belongs to $( \frac {x}{2},\frac {3x}{2} )$ .

Consequently, $\mathcal {N}_k$ , if nonempty, contains only powers that satisfy the restrictions in (3.10). The square-free condition ensures that $\mathcal {N}_j \cap \mathcal {N}_k = \varnothing $ for $j \neq k$ .

Write $S_{\mathrm {sqf}}(x) = S_{\text {near}}(x) + S_{\text {far}}(x)$ , in which

(3.13) $$ \begin{align} S_{\text{near}}(x) \,\,= \!\!\!\! \sum_{ \substack{ \frac{x}{2} < n^k < \frac{3x}{2} \\ k \geq 2 \\ n \geq 2\text{ sq.~free} \\ n^k \in \mathcal{N}_k}} \frac{1}{ k | x-n^k |} \quad \text{and} \quad S_{\text{far}}(x) \,\,= \!\!\!\! \sum_{ \substack{ \frac{x}{2} < n^k < \frac{3x}{2} \\ k \geq 2 \\ n \geq 2\text{ sq.~free} \\ n^k \notin \mathcal{N}_k}} \frac{1}{ k | x-n^k |}. \end{align} $$

3.4.3 Near Sum

For $x\geq 3.5$ , a nearest-neighbor overestimate provides

(3.14) $$ \begin{align} S_{\text{near}}(x) &= \sum_{ \substack{ \frac{x}{2} < n^k < \frac{3x}{2} \\ k \geq 2 \\ n \geq 2\text{ sq.~free} \\ n^k \in \mathcal{N}_k}} \frac{1}{ k | x-n^k |} && (\text{by 3.13}) \nonumber\\ &= \sum_{k=2}^{\lfloor 2.4\log x \rfloor } \sum_{ m \in \mathcal{N}_k } \frac{1}{ k | x-m |} && (\text{by 3.11}) \nonumber\\ &\leq \frac{1}{2} \sum_{k=2}^{\lfloor 2.4\log x \rfloor } \sum_{ m \in \mathcal{N}_k } \frac{1}{ | x-m |} && ( \text{since }k \geq 2 ) \nonumber\\ &\leq \frac{1}{2} \sum_{j=0}^{\lfloor 2.4\log x \rfloor-2 } \left( \frac{1}{ x - (\lfloor x \rfloor-j)} + \frac{1}{ (\lceil x \rceil +j) - x} \right) && (\text{see below}) \nonumber\\ &\leq \frac{1}{2} \sum_{\ell=1}^{\lfloor 2.4\log x \rfloor-1 } \frac{2}{\ell - \frac{1}{2}} \quad<\quad 2 \!\!\!\!\sum_{\ell=1}^{\lfloor 2.4\log x \rfloor } \frac{1}{2\ell - 1} \nonumber\\ &\leq \log( \lfloor 2.4\log x \rfloor) + \gamma + 2\log 2 + \frac{14}{312}&& ( \text{by 2.5}) \nonumber\\ &< \log \log x + \gamma + 2\log 2 + \log 2.4 + \frac{7}{156}. \end{align} $$

Let us elaborate on a crucial step above. Consider the at most $\lfloor 2 \log x \rfloor - 1$ pairs of values $|x-m|$ that arise as m ranges over each $\mathcal {N}_k$ with $2 \leq k \leq \lfloor 2 \log x \rfloor $ (since $\mathcal {N}_j(x) \cap \mathcal {N}_k = \varnothing $ for $j \neq k$ , no m appears more than once). Replace these values with the absolute deviations of x from its $2 \times (\lfloor 2 \log x \rfloor - 1)$ nearest neighbors $\lfloor x \rfloor -j$ (to the left) and $\lceil x \rceil +j$ (to the right), in which $0 \leq j \leq \lfloor 2 \log x \rfloor -2$ . Since $x \in \mathbb {N}+\frac {1}{2}$ , these deviations are of the form $\ell - \frac {1}{2}$ for $1 \leq \ell \leq \lfloor 2\log x \rfloor -1$ .

3.4.4 Splitting the second sum

From (3.13), the second sum in question is

$$ \begin{align*} S_{\text{far}}(x) = \!\!\!\! \sum_{ \substack{ \frac{x}{2} < n^k < \frac{3x}{2} \\ k \geq 2 \\ n \geq 2\text{ sq.~free} \\ n^k \notin \mathcal{N}_k}} \frac{1}{ k | x-n^k |} \leq \!\!\!\! \sum_{ \substack{ \frac{x}{2} < n^k < \frac{3x}{2} \\ n,k \geq 2 \\ n^k \notin \mathcal{N}_k}} \frac{1}{ k | x-n^k |} = S_{\text{far}}^-(x) + S_{\text{far}}^+(x), \end{align*} $$

in which

(3.15) $$ \begin{align} S_{\text{far}}^-(x) = \sum_{k\geq 2} \sum_{ \substack{ \frac{x}{2} < n^k < x \\ n \geq 2, n^k \notin \mathcal{N}_k}} \frac{1}{ k | x-n^k |} \quad \text{and} \quad S_{\text{far}}^+(x) = \sum_{k\geq 2} \sum_{ \substack{ x < n^k < \frac{3x}{2} \\ n \geq 2, n^k \notin \mathcal{N}_k}} \frac{1}{ k | x-n^k |}. \end{align} $$

For $k \geq 2$ , Lemma 2.5 with $X =h= \frac {x}{2}$ , then with $X = x$ and $h = \frac {x}{2}$ , implies that

(3.16) $$ \begin{align} G_k^- \geq \frac{k \sqrt{x}}{\sqrt{2}}, \quad N_k^- \leq \frac{\sqrt{x}}{2 \sqrt{2}}, \qquad \text{and} \qquad G_k^+ \geq k \sqrt{x}, \quad N_k^+ \leq \frac{ \sqrt{x}}{4} \end{align} $$

are admissible in Figure 5. For $1\leq j\leq N^-_k$ and $1\leq j\leq N^+_k$ , respectively,

Figure 5 Analysis of kth powers in $[\tfrac {x}{2}, \tfrac {3x}{2}]$ , in which $m_- = \lfloor x^{1/k} \rfloor $ and $m_+ = \lceil x^{1/k} \rceil $ are excluded from consideration. There are at most $N_k^-$ admissible kth powers in $[\frac {x}{2},x)$ , with minimal gap size $G_k^-$ , and at most $N_k^+$ admissible kth powers in $[x,\frac {3x}{2})$ , with minimal gap size $G_k^+$ .

$$ \begin{align*} |x - (m_- - j)^k| \geq \frac{j k \sqrt{x}}{\sqrt{2}} \qquad\text{and}\qquad |x - (m_+ + j)^k| \geq j k \sqrt{x}. \end{align*} $$

Let $N_k^{\pm } \geq 1$ , since otherwise the corresponding sum estimated below is zero. Then

(3.17) $$ \begin{align} \sum_{ \substack{\frac{x}{2} < n^k < x \\ n \geq 2, n^k \notin \mathcal{N}_k}} \frac{1}{ k | x-n^k |} = \sum_{j=1}^{N^-_k} \frac{1}{k |x-(m_- - j)^k|} \leq \frac{\sqrt{2}H_{N^-_k} }{k^2 \sqrt{x}} \end{align} $$

and

(3.18) $$ \begin{align} \sum_{\substack{x < n^k < \tfrac{3x}{2} \\ n \geq 2, n^k \notin \mathcal{N}_k}} \frac{1}{ k | x-n^k |} = \sum_{j=1}^{N^+_k} \frac{1}{k |x-(m_+ + j)^k|} \leq \frac{H_{N^+_k}}{k^2 \sqrt{x}}. \end{align} $$

Therefore,

(3.19) $$ \begin{align} S_{\text{far}}^-(x) &=\sum_{k\geq 2} \sum_{ \substack{ \frac{x}{2} < n^k < x \\ n \geq 2, n^k \notin \mathcal{N}_k}} \frac{1}{ k | x-n^k |} \leq \frac{\sqrt{2}}{\sqrt{x}} \sum_{k\geq 2}\frac{H_{N^-_k} }{k^2 } && (\text{by (3.15) and (3.17)}) \nonumber\\ &\leq \frac{\sqrt{2}}{\sqrt{x}} \bigg(\frac{1}{2}\log{x} - \frac{3}{2}\log{2} + \gamma + \frac{3}{7} \bigg) \sum_{k\geq 2}\frac{1}{k^2} && (\text{by (2.4) and (3.16)}) \nonumber \\ &=\frac{\pi^2-6}{3\sqrt{2 x}} \bigg(\frac{1}{2}\log{x} - \frac{3}{2}\log{2} + \gamma + \frac{3}{7} \bigg) &&(\text{since }\zeta(2)-1 = \tfrac{\pi^2-6}{6}) \end{align} $$

and

(3.20) $$ \begin{align} S_{\text{far}}^+(x) &=\sum_{k\geq 2} \sum_{ \substack{ x < n^k < \frac{3x}{2} \\ n \geq 2, n^k \notin \mathcal{N}_k}} \frac{1}{ k | x-n^k |} \leq \frac{1}{ \sqrt{x}} \sum_{k \geq 2} \frac{H_{N^+_k}}{k^2} && (\text{by (3.15) and (3.18)}) \nonumber\\ &\leq \frac{1}{ \sqrt{x}} \bigg(\frac{1}{2}\log{x} - 2\log{2} + \gamma + \frac{3}{7} \bigg) \sum_{k \geq 2} \frac{1}{k^2} && (\text{by (2.4) and (3.16)}) \nonumber \\ &= \frac{\pi^2-6}{6\sqrt{x}} \bigg(\frac{1}{2}\log{x} - 2\log{2} + \gamma + \frac{3}{7} \bigg) &&(\text{since }\zeta(2)-1 = \tfrac{\pi^2-6}{6}). \end{align} $$

3.4.5 Final prime-power estimate

Using (3.14), (3.19), and (3.20), we can bound

$$ \begin{align*} S_{\mathrm{power}}(x) \leq S_{\mathrm{sqf}} (x) = S_{\mathrm{near}}(x) + S_{\mathrm{far}}^-(x) + S_{\mathrm{far}}^+(x). \end{align*} $$

We postpone doing this explicitly until the finale below.

4 Conclusion

For $x \geq 3.5$ , with $T \geq (\log \frac {3}{2})^{-1}$ , the sum (1.3) is bounded by

$$ \begin{align*} &\underbrace{\frac{e}{T \log \frac{3}{2}}\bigg(\log \log x + \frac{\gamma}{\log x} \bigg)}_{\text{by (3.1)}} + \underbrace{\frac{e }{T \log\frac{3}{2}} \sum_{\frac{x}{2} < p^k < \frac{3x}{2}} \frac{1}{k | x-p^k |} }_{\text{by (3.4)}} \\ &\quad \leq \frac{e}{T \log \frac{3}{2}} \bigg(\log \log x + \frac{\gamma}{\log x} \bigg) + \frac{e}{T \log \frac{3}{2}} \underbrace{ \big(S_{\text{prime}}(x) + S_{\text{sqf}}(x) \big)}_{\text{by (3.5) and (3.10)}} \\ &\quad \leq \frac{e}{T \log \frac{3}{2}} \bigg(\log \log x + \frac{\gamma}{\log x} \bigg) + \frac{e}{T \log \frac{3}{2}}\bigg[ \underbrace{2C +4 \log\log x}_{S_{\text{prime}}(x)\text{ bounded by (3.9)}} \\ &\qquad + \underbrace{ \bigg(\log \log x + \gamma + 2\log 2 + \log 2.4 +\frac{7}{156} \bigg) }_{S_{\text{near}}(x)\text{ bounded by (3.14)}} + \underbrace{\frac{\pi^2-6}{3\sqrt{2 x}} \bigg(\frac{1}{2}\log{x} - \frac{3}{2}\log{2} + \gamma + \frac{3}{7} \bigg)}_{S_{\text{far}}^-\text{ bounded by (3.19)}} \\ &\qquad + \underbrace{\frac{\pi^2-6}{6\sqrt{x}} \bigg(\frac{1}{2}\log{x} - 2\log{2} + \gamma + \frac{3}{7} \bigg)}_{S_{\text{far}}^+(x)\text{ bounded by (3.20)}} \bigg] \\ &\quad < \frac{1}{T}\bigg( 40.22465 \log \log x +58.11106 +\frac{3.86972}{\log x} +\frac{5.21918 \log x}{\sqrt{x}} -\frac{1.85268}{\sqrt{x}} \bigg). \end{align*} $$

Acknowledgment

E.S.L. thanks the Heilbronn Institute for Mathematical Research for their support. We also thank the referee for their suggestions.

Footnotes

S.R.G. was supported by NSF Grant DMS-2054002.

References

Cho, P. J. and Kim, H. H., Extreme residues of Dedekind zeta functions. Math. Proc. Cambridge Philos. Soc. 163(2017), no. 2, 369380.CrossRefGoogle Scholar
Corless, R. M., Gonnet, G. H., Hare, D. E. G., Jeffrey, D. J., and Knuth, D. E., On the Lambert $W$ function. Adv. Comput. Math. 5(1996), no. 4, 329359.CrossRefGoogle Scholar
Cully-Hugill, M. and Johnston, D. R., On the error term in the explicit formula of Riemann–von Mangoldt. Int. J. Number Theory (2022).Google Scholar
Davenport, H., Multiplicative number theory, 3rd ed., Graduate Texts in Mathematics, 74, Springer, New York, 2000.Google Scholar
Goldston, D. A., On a result of Littlewood concerning prime numbers. II . Acta Arith. 43(1983), no. 1, 4951.CrossRefGoogle Scholar
Granville, A. and Soundararajan, K., Large character sums . J. Amer. Math. Soc. 14(2001), no. 2, 365397.CrossRefGoogle Scholar
Iwaniec, H. and Kowalski, E., Analytic number theory, American Mathematical Society Colloquium Publications, 53, American Mathematical Society, Providence, RI, 2004.Google Scholar
Koukoulopoulos, D., The distribution of prime numbers, Graduate Studies in Mathematics, 203, American Mathematical Society, Providence, RI, 2019.CrossRefGoogle Scholar
Lóczi, L., Guaranteed- and high-precision evaluation of the Lambert W function . Appl. Math. Comput.. 433 (2022), 127406.Google Scholar
Montgomery, H. L. and Vaughan, R. C., The large sieve . Mathematika 20(1973), 119134.CrossRefGoogle Scholar
Murty, M. R., Problems in analytic number theory, Graduate Texts in Mathematics, Springer, New York, 2008.Google Scholar
Patterson, S. J., An introduction to the theory of the Riemann zeta-function, Cambridge Studies in Advanced Mathematics, 14, Cambridge University Press, Cambridge, 1988.CrossRefGoogle Scholar
Ramaré, O., An explicit density estimate for Dirichlet $L$ -series . Math. Comp. 85(2016), no. 297, 325356.CrossRefGoogle Scholar
Selberg, A., On the normal density of primes in small intervals, and the difference between consecutive primes . Arch. Math. Naturvid. 47(1943), no. 6, 87105.Google Scholar
Tenenbaum, G., Introduction to analytic and probabilistic number theory, Cambridge Studies in Advanced Mathematics, 46, Cambridge University Press, Cambridge, 1995, Translated from the second French edition (1995) by C. B. Thomas.Google Scholar
Tóth, L. and Mare, S., E 3432. Amer. Math. Monthly. 98(1991), no. 3, 264.Google Scholar
Figure 0

Figure 1 Graphs relevant to the construction of the sequence $y_n$.

Figure 1

Figure 2 The functions $F_1(x)$ and $F_2(x)$ behave erratically.

Figure 2

Figure 3 Proof of Lemma 2.5.

Figure 3

Figure 4 Graphs relevant to the derivation of (3.2).

Figure 4

Figure 5 Analysis of kth powers in $[\tfrac {x}{2}, \tfrac {3x}{2}]$, in which $m_- = \lfloor x^{1/k} \rfloor $ and $m_+ = \lceil x^{1/k} \rceil $ are excluded from consideration. There are at most $N_k^-$ admissible kth powers in $[\frac {x}{2},x)$, with minimal gap size $G_k^-$, and at most $N_k^+$ admissible kth powers in $[x,\frac {3x}{2})$, with minimal gap size $G_k^+$.