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

On the error term in the explicit formula of Riemann–von Mangoldt

Michaela Cully-Hugill and Daniel R. Johnston School of Science, UNSW Canberra, Australia m.cully-hugill@adfa.edu.au School of Science, UNSW Canberra, Australia daniel.johnston@adfa.edu.au
Abstract.

We provide an explicit O(xlogx/T)O(x\log x/T) error term for the Riemann–von Mangoldt formula by making results of Wolke (1983) and Ramaré (2016) explicit. We also include applications to primes between consecutive powers, the error term in the prime number theorem and an inequality of Ramanujan.

Key words and phrases:

1. Introduction

1.1. Background

Let

ψ(x)=nxΛ(n)=pkxlogp\psi(x)=\sum_{n\leq x}\Lambda(n)=\sum_{p^{k}\leq x}\log p

be the Chebyshev prime-counting function over prime powers pkp^{k}, with integers k1k\geq 1. The (truncated) Riemann–von Mangoldt formula is written as

ψ(x)=xρ=β+iγ|γ|Txρρ+E(x,T),\psi(x)=x-\sum_{\begin{subarray}{c}\rho=\beta+i\gamma\\ |\gamma|\leq T\end{subarray}}\frac{x^{\rho}}{\rho}+E(x,T), (1.1)

where the sum is over all non-trivial zeros ρ=β+iγ\rho=\beta+i\gamma of the Riemann-zeta function ζ(s)\zeta(s) that have |γ|T|\gamma|\leq T, and E(x,T)E(x,T) is an error term.

Several authors have provided estimates for E(x,T)E(x,T). With reasonable conditions on xx and TT, the standard bound given in the literature (e.g. [11, Chapter 17]) is

E(x,T)=O(xlog2xT),E(x,T)=O\left(\frac{x\log^{2}x}{T}\right), (1.2)

which in 1983 was improved by Goldston [18] to

E(x,T)=O(xlogxloglogxT).E(x,T)=O\left(\frac{x\log x\log\log x}{T}\right). (1.3)

Under the Riemann hypothesis, one can do even better. Namely, Littlewood [25] proves that with the Riemann hypothesis one gets

E(x,T)=O(xlogxT),E(x,T)=O\left(\frac{x\log x}{T}\right), (1.4)

which was later improved by Goldston [17] to O(x/T)O(x/T). Wolke [38] and Ramaré [34] also claimed to prove averaged versions of the Riemann-von Mangoldt formula with unconditional O(x/T)O(x/T) error terms. However, through correspondence with Ramaré, several errors in these works have been uncovered.

In 2016, Dudek [13, Theorem 1.3] gave an explicit version of (1.2). Namely, he proved the following theorem.

Theorem 1.1 (Dudek).

Take 50<T<x50<T<x for half an odd integer x>e60x>e^{60}. Then

E(x,T)=O(2xlog2xT).E(x,T)=O^{*}\left(\frac{2x\log^{2}x}{T}\right).

The first author [7, Theorem 2] recently improved on Dudek’s result by making (1.3) explicit. In this paper, we give an explicit O(xlogx/T)O(x\log x/T) bound for E(x,T)E(x,T), thereby making (1.4) unconditional and explicit. This is done by reworking the papers of Wolke [38] and Ramaré [34]: we optimise parts of their proofs and avoid the errors related to their averaging arguments. This approach gives significantly better explicit bounds for E(x,T)E(x,T) than in previous works.

Our overarching approach is to split E(x,T)E(x,T) into two smaller error terms, say E1(x,T)E_{1}(x,T) and E2(x,T)E_{2}(x,T). We rework [34, Theorem 1.1] to bound E1(x,T)E_{1}(x,T) in a general way that can be applied to other arithmetic functions besides ψ(x)\psi(x). Then, we rework the proof of [38, Theorem 2] and use explicit zero-free regions for ζ(s)\zeta(s) to reach an explicit estimate of the form E2(x,T)=O(x/T)E_{2}(x,T)=O(x/T). Such an estimate for E2(x,T)E_{2}(x,T) becomes insignificant compared to that for E1(x,T)E_{1}(x,T) for large xx.

To demonstrate the usefulness of our results, we use our bounds for E(x,T)E(x,T) to improve the main theorems in [7] and [32].

1.2. Statement of main results

Our main result is as follows.

Theorem 1.2.

For any α(0,1/2]\alpha\in(0,1/2] there exist constants MM and xMx_{M} such that for max{51,logx}<T<(xα2)/2\max\{51,\log x\}<T<(x^{\alpha}-2)/2,

ψ(x)=x|γ|Txρρ+O(MxlogxT)\psi(x)=x-\sum_{\begin{subarray}{c}|\gamma|\leq T\end{subarray}}\frac{x^{\rho}}{\rho}+O^{*}\left(M\frac{x\log x}{T}\right) (1.5)

for all xxMx\geq x_{M}. Some admissible values of xMx_{M}, α\alpha and MM are (40,1/2,5.03)(40,1/2,5.03) and (103,1/100,0.5597)(10^{3},1/100,0.5597), with more given in Table 4 in the appendix.

Using Theorem 1.2 we are able to obtain the following result.

Theorem 1.3.

There is at least one prime between n140n^{140} and (n+1)140(n+1)^{140} for all n1n\geq 1.

This improves upon [7, Thm. 1] by the first author, which asserts that there is always a prime between consecutive 155th155^{\text{th}} powers.

Theorem 1.2, combined with other recent results, also allows us to improve the error term in the prime number theorem for large xx (cf. [32, Theorem 1]).

Theorem 1.4.

Let R=5.5666305R=5.5666305. For each {X,A,B,C,ϵ0}\{X,A,B,C,\epsilon_{0}\} in Table 1 we have

|ψ(x)xx|A(logxR)Bexp(ClogxR),\left|\frac{\psi(x)-x}{x}\right|\leq A\left(\frac{\log x}{R}\right)^{B}\exp\left(-C\sqrt{\frac{\log x}{R}}\right),

and for all logxX\log x\geq X,

|ψ(x)x|ϵ0x.|\psi(x)-x|\leq\epsilon_{0}x.
Table 1. Values of XX, AA, BB, CC and ϵ0\epsilon_{0} for Theorem 1.4. Here, σ\sigma is a parameter that appears in the proof of Theorem 1.4. The entry for X=3600X=3600 is specifically included for Corollary 1.5.
XX σ\sigma AA BB CC ϵ0\epsilon_{0}
10001000 0.9800.980 269.1269.1 1.5201.520 1.8931.893 6.891066.89\cdot 10^{-6}
20002000 0.9840.984 264.8264.8 1.5161.516 1.9141.914 3.4810103.48\cdot 10^{-10}
30003000 0.9860.986 264.3264.3 1.5141.514 1.9251.925 1.4210131.42\cdot 10^{-13}
36003600 0.9880.988 275.2275.2 1.5121.512 1.9361.936 2.0410152.04\cdot 10^{-15}
40004000 0.9880.988 266.5266.5 1.52121.5212 1.9361.936 1.6110161.61\cdot 10^{-16}
50005000 0.9900.990 350.4350.4 1.5101.510 1.9461.946 4.7410194.74\cdot 10^{-19}
60006000 0.9900.990 267.8267.8 1.5101.510 1.9461.946 1.8310211.83\cdot 10^{-21}
70007000 0.9900.990 266.9266.9 1.5101.510 1.9461.946 1.3810231.38\cdot 10^{-23}
80008000 0.9900.990 266.9266.9 1.5101.510 1.9461.946 1.4410251.44\cdot 10^{-25}
90009000 0.9920.992 280.5280.5 1.5081.508 1.9571.957 1.3010271.30\cdot 10^{-27}
1000010000 0.9920.992 268.6268.6 1.5081.508 1.9571.957 2.0610292.06\cdot 10^{-29}

The values of ϵ0\epsilon_{0} in Table 1 are 40–80%\% smaller than those in [32, Table 1]. Replacing AA with A1=A+0.1A_{1}=A+0.1 gives a corresponding expression for θ(x)\theta(x) (see Corollary 7.2). It should be noted that the methods used to prove Theorem 1.4 have been expanded on in recent preprints [15, 23].

By repeating the argument on pages 877–880 of [32], we obtain the following application to an inequality of Ramanujan on the prime counting function π(x)\pi(x).

Corollary 1.5.

For all xexp(3604)x\geq\exp(3604) we have

π(x)2<exlogxπ(xe).\pi(x)^{2}<\frac{ex}{\log x}\pi\left(\frac{x}{e}\right).

Corollary 1.5 improves [32, Theorem 2] by a factor of exp(311)\exp(311).

1.3. Outline of paper

We begin in Section 2 by making an explicit and simplified version of a truncated Perron formula due to Ramaré [34]. In Section 3 we state a number of zero-free regions from the literature that will be required throughout the paper. In Section 4, we make explicit an argument of Wolke [38]. These results are combined in Section 5 to prove Theorem 1.2. Applications of these results (Theorems 1.3 and 1.4) are in Sections 6 and 7.

2. An explicit truncated Perron formula

In this section we prove the following error estimate for the truncated Perron formula. This result is a self-contained variant of [34, Theorem 1.1], which we have simplified and optimised for our purposes.

Theorem 2.1.

Let F(s)=n1an/nsF(s)=\sum_{n\geq 1}a_{n}/n^{s} be a Dirichlet series over complex ss, and κ>0\kappa>0 be a real parameter chosen larger than the abscissa of absolute convergence of F(s)F(s). Also let θ=2/(π2+4+π)\theta^{\prime}=2/(\sqrt{\pi^{2}+4}+\pi). For any T>0T>0, x1x\geq 1, κ>κa\kappa>\kappa_{a}, and λθ/T\lambda\geq\theta^{\prime}/T,

nxan=12πiκiTκ+iTF(s)xssds\displaystyle\sum_{n\leq x}a_{n}=\frac{1}{2\pi i}\int_{\kappa-iT}^{\kappa+iT}F(s)\frac{x^{s}}{s}\mathrm{d}s
+O(xκπλTn1|an|nκ+1πTθ/Tλ|log(x/n)|u|an|eκuu2du).\displaystyle\qquad\qquad\qquad\qquad+O^{*}\left(\frac{x^{\kappa}}{\pi\lambda T}\sum_{n\geq 1}\frac{|a_{n}|}{n^{\kappa}}+\frac{1}{\pi T}\int_{\theta^{\prime}/T}^{\lambda}\sum_{|\log(x/n)|\leq u}|a_{n}|\frac{e^{\kappa u}}{u^{2}}\mathrm{d}u\right).

Theorem 2.1 is proven using Lemma 2.2, which is a specific case of [34, Lem. 2.2]. For both proofs, we will use the step function

v(y)={1y00y<0.v(y)=\begin{cases}1\quad y\geq 0\\ 0\quad y<0.\end{cases}
Lemma 2.2.

For κ>0\kappa^{\prime}>0 and yy\in\mathbb{R} we have

|v(y)12πiκiκ+ieyssds|min{eyκπ|y|,|v(y)eyκπarctan(1/κ)|+|y|eyκπ}.\left|v(y)-\frac{1}{2\pi i}\int_{\kappa^{\prime}-i}^{\kappa^{\prime}+i}\frac{e^{ys}}{s}\mathrm{d}s\right|\leq\min\left\{\frac{e^{y\kappa^{\prime}}}{\pi|y|},\left|v(y)-\frac{e^{y\kappa^{\prime}}}{\pi}\arctan(1/\kappa^{\prime})\right|+\frac{|y|e^{y\kappa^{\prime}}}{\pi}\right\}. (2.1)
Proof.

For yy\in\mathbb{R} and K>κK>\kappa^{\prime} we can evaluate the contour integral

(κiκ+i+κ+iK+i+K+iKi+Kiκi)eyssds=0.\left(\int_{\kappa^{\prime}-i}^{\kappa^{\prime}+i}+\int_{\kappa^{\prime}+i}^{K+i}+\int_{K+i}^{K-i}+\int_{K-i}^{\kappa^{\prime}-i}\right)\frac{e^{ys}}{s}\mathrm{d}s=0.

Consider the case y<0y<0: as KK approaches infinity the third integral approaches zero, and the two horizontal integrals are bounded by eyκ/|y|e^{y\kappa^{\prime}}/|y|. Hence,

|κiκ+ieyssds|2eyκ|y|.\left|\int_{\kappa^{\prime}-i}^{\kappa^{\prime}+i}\frac{e^{ys}}{s}\mathrm{d}s\right|\leq\frac{2e^{y\kappa^{\prime}}}{|y|}.

Therefore, we have

|v(y)12πiκiκ+ieyssds|\displaystyle\left|v(y)-\frac{1}{2\pi i}\int_{\kappa^{\prime}-i}^{\kappa^{\prime}+i}\frac{e^{ys}}{s}\mathrm{d}s\right| eyκπ|y|.\displaystyle\leq\frac{e^{y\kappa^{\prime}}}{\pi|y|}.

The case y>0y>0 is similar, but we use a contour extended to the left, so for K<0K<0,

(κiκ+i+κ+iK+i+K+iKi+Kiκi)eyssds=2πi.\left(\int_{\kappa^{\prime}-i}^{\kappa^{\prime}+i}+\int_{\kappa^{\prime}+i}^{K+i}+\int_{K+i}^{K-i}+\int_{K-i}^{\kappa^{\prime}-i}\right)\frac{e^{ys}}{s}\mathrm{d}s=2\pi i.

Taking KK\rightarrow-\infty brings the third integral to zero, so

|2πiκiκ+ieyssds|2eyκy,\left|2\pi i-\int_{\kappa^{\prime}-i}^{\kappa^{\prime}+i}\frac{e^{ys}}{s}\mathrm{d}s\right|\leq\frac{2e^{y\kappa^{\prime}}}{y},

and thus, for y>0y>0

|v(y)12πiκiκ+ieyssds|\displaystyle\left|v(y)-\frac{1}{2\pi i}\int_{\kappa^{\prime}-i}^{\kappa^{\prime}+i}\frac{e^{ys}}{s}\mathrm{d}s\right| eyκπy.\displaystyle\leq\frac{e^{y\kappa^{\prime}}}{\pi y}.

The above bounds are most useful for large yy. For small yy we can use

κiκ+ieyssds=eyκκiκ+idss+ieyκ11eiyt1κ+itdt.\int_{\kappa^{\prime}-i}^{\kappa^{\prime}+i}\frac{e^{ys}}{s}\mathrm{d}s=e^{y\kappa^{\prime}}\int_{\kappa^{\prime}-i}^{\kappa^{\prime}+i}\frac{\mathrm{d}s}{s}+ie^{y\kappa^{\prime}}\int_{-1}^{1}\frac{e^{iyt}-1}{\kappa^{\prime}+it}\mathrm{d}t.

The first integral is equivalent to 2iarctan(1/κ)2i\text{arctan}(1/\kappa^{\prime}). The second integral can be bounded using the identity

|eiyt1iyt|=|01eiytudu|1.\left|\frac{e^{iyt}-1}{iyt}\right|=\left|\int_{0}^{1}e^{iytu}\mathrm{d}u\right|\leq 1.

Hence, for all yy\in\mathbb{R},

|v(y)12πiκiκ+ieyssds|\displaystyle\left|v(y)-\frac{1}{2\pi i}\int_{\kappa^{\prime}-i}^{\kappa^{\prime}+i}\frac{e^{ys}}{s}\mathrm{d}s\right| |v(y)eyκπarctan(1/κ)|+|y|eyκπ.\displaystyle\leq\left|v(y)-\frac{e^{y\kappa^{\prime}}}{\pi}\text{arctan}(1/\kappa^{\prime})\right|+\frac{|y|e^{y\kappa^{\prime}}}{\pi}.\qed
Proof of Theorem 2.1.

We aim to bound

|n1anv(Tlog(x/n))12πiκiTκ+iTF(s)xssds|xκ\left|\sum_{n\geq 1}a_{n}v(T\log(x/n))-\frac{1}{2\pi i}\int_{\kappa-iT}^{\kappa+iT}F(s)\frac{x^{s}}{s}\mathrm{d}s\right|x^{-\kappa}

for any x1x\geq 1, T>0T>0, and κ>κa>0\kappa>\kappa_{a}>0, where κa\kappa_{a} is the abscissa of absolute convergence of F(s)F(s). First, we take κ=κT\kappa=\kappa^{\prime}T to reach

|n1anv(Tlog(x/n))12πiκiTκ+iTF(s)xssds|xκ\displaystyle\left|\sum_{n\geq 1}a_{n}v(T\log(x/n))-\frac{1}{2\pi i}\int_{\kappa-iT}^{\kappa+iT}F(s)\frac{x^{s}}{s}\mathrm{d}s\right|x^{-\kappa}
n1|an|nκ|v(Tlog(x/n))12πiκiTκ+iT(xn)sdss|(nx)κ\displaystyle\qquad\leq\sum_{n\geq 1}\frac{|a_{n}|}{n^{\kappa}}\left|v(T\log(x/n))-\frac{1}{2\pi i}\int_{\kappa-iT}^{\kappa+iT}\left(\frac{x}{n}\right)^{s}\frac{\mathrm{d}s}{s}\right|\left(\frac{n}{x}\right)^{\kappa}
=n1|an|nκ|v(Tlog(x/n))12πiκiκ+ieTlog(x/n)wdww|eκTlog(x/n).\displaystyle\qquad=\sum_{n\geq 1}\frac{|a_{n}|}{n^{\kappa}}\left|v(T\log(x/n))-\frac{1}{2\pi i}\int_{\kappa^{\prime}-i}^{\kappa^{\prime}+i}e^{T\log(x/n)w}\frac{\mathrm{d}w}{w}\right|e^{-\kappa^{\prime}T\log(x/n)}. (2.2)

Next, we apply Lemma 2.2. Suppose for some θ>0\theta>0 there exists a positive constant cc such that

max|y|<θ(min(1π|y|,|v(y)eyκ1πarctan(1/κ)|+|y|π))c.\displaystyle\max_{|y|<\theta}\left(\min\left(\frac{1}{\pi|y|},\left|\frac{v(y)}{e^{y\kappa^{\prime}}}-\frac{1}{\pi}\arctan(1/\kappa^{\prime})\right|+\frac{|y|}{\pi}\right)\right)\leq c. (2.3)

Splitting the sum in (2.2) at θ\theta and taking y=Tlog(x/n)y=T\log(x/n) in Lemma 2.2 gives

|n1anv(Tlog(x/n))12πiκiTκ+iTF(s)xssds|xκ\displaystyle\left|\sum_{n\geq 1}a_{n}v(T\log(x/n))-\frac{1}{2\pi i}\int_{\kappa-iT}^{\kappa+iT}F(s)\frac{x^{s}}{s}\mathrm{d}s\right|x^{-\kappa}
cT|log(x/n)|<θ|an|nκ+1πTT|log(x/n)|θ|an|nκ|log(x/n)|,\displaystyle\qquad\leq c\sum_{T|\log(x/n)|<\theta}\frac{|a_{n}|}{n^{\kappa}}+\frac{1}{\pi T}\sum_{T|\log(x/n)|\geq\theta}\frac{|a_{n}|}{n^{\kappa}|\log(x/n)|}, (2.4)

and note that the first bound in (2.1) was used to obtain the second term in (2.4). We then have

T|log(x/n)|θ|an|nκ|log(x/n)|\displaystyle\sum_{T|\log(x/n)|\geq\theta}\frac{|a_{n}|}{n^{\kappa}|\log(x/n)|} =T|log(x/n)|θ|an|nκ|log(x/n)|duu2\displaystyle=\sum_{T|\log(x/n)|\geq\theta}\frac{|a_{n}|}{n^{\kappa}}\int_{|\log(x/n)|}^{\infty}\frac{\mathrm{d}u}{u^{2}}
=θ/Tθ/T|log(x/n)|u|an|nκduu2\displaystyle=\int_{\theta/T}^{\infty}\sum_{\theta/T\leq|\log(x/n)|\leq u}\frac{|a_{n}|}{n^{\kappa}}\frac{\mathrm{d}u}{u^{2}}
=θ/T|log(x/n)|u|an|nκduu2TθT|log(x/n)|<θ|an|nκ.\displaystyle=\int_{\theta/T}^{\infty}\sum_{|\log(x/n)|\leq u}\frac{|a_{n}|}{n^{\kappa}}\frac{\mathrm{d}u}{u^{2}}-\frac{T}{\theta}\sum_{T|\log(x/n)|<\theta}\frac{|a_{n}|}{n^{\kappa}}.

Hence,

|n1anv(Tlog(x/n))12πiκiTκ+iTF(s)xssds|xκ\displaystyle\left|\sum_{n\geq 1}a_{n}v(T\log(x/n))-\frac{1}{2\pi i}\int_{\kappa-iT}^{\kappa+iT}F(s)\frac{x^{s}}{s}\mathrm{d}s\right|x^{-\kappa}
1πTθ/T|log(x/n)|u|an|nκduu2+(c1πθ)T|log(x/n)|<θ|an|nκ.\displaystyle\qquad\leq\frac{1}{\pi T}\int_{\theta/T}^{\infty}\sum_{|\log(x/n)|\leq u}\frac{|a_{n}|}{n^{\kappa}}\frac{\mathrm{d}u}{u^{2}}+\left(c-\frac{1}{\pi\theta}\right)\sum_{T|\log(x/n)|<\theta}\frac{|a_{n}|}{n^{\kappa}}. (2.5)

Over T|log(x/n)|<θT|\log(x/n)|<\theta we have |v(y)eyκarctan(1/κ)/π|<1\left|v(y)e^{-y\kappa^{\prime}}-\arctan(1/\kappa^{\prime})/\pi\right|<1, so we can take

c=max|y|<θ(min(1π|y|,1+|y|π))=π2+42π+12.c=\max_{|y|<\theta}\left(\min\left(\frac{1}{\pi|y|},1+\frac{|y|}{\pi}\right)\right)=\frac{\sqrt{\pi^{2}+4}}{2\pi}+\frac{1}{2}.

This implies that if we choose θθ:=2/(π2+4+π)\theta\leq\theta^{\prime}:=2/(\sqrt{\pi^{2}+4}+\pi) then the second term of (2.5) vanishes. Also, the integral in the first term of (2.5) will be minimised for larger θ\theta. Hence, we will take θ=θ\theta=\theta^{\prime}. This leaves us with

|n1anv(Tlog(x/n))12πiκiTκ+iTF(s)xssds|xκπTθ/T|log(x/n)|u|an|nκduu2.\displaystyle\left|\sum_{n\geq 1}a_{n}v(T\log(x/n))-\frac{1}{2\pi i}\int_{\kappa-iT}^{\kappa+iT}F(s)\frac{x^{s}}{s}\mathrm{d}s\right|\leq\frac{x^{\kappa}}{\pi T}\int_{\theta^{\prime}/T}^{\infty}\sum_{|\log(x/n)|\leq u}\frac{|a_{n}|}{n^{\kappa}}\frac{\mathrm{d}u}{u^{2}}.

Splitting the integral from θ/T\theta^{\prime}/T to \infty at λ\lambda then gives the further bound

xκθ/T|log(x/n)|u|an|nκduu2\displaystyle x^{\kappa}\int_{\theta^{\prime}/T}^{\infty}\sum_{|\log(x/n)|\leq u}\frac{|a_{n}|}{n^{\kappa}}\frac{\mathrm{d}u}{u^{2}} xκn1|an|nκλduu2+θ/Tλ|log(x/n)|u|an|(xn)κduu2\displaystyle\leq x^{\kappa}\sum_{n\geq 1}\frac{|a_{n}|}{n^{\kappa}}\int_{\lambda}^{\infty}\frac{\mathrm{d}u}{u^{2}}+\int_{\theta^{\prime}/T}^{\lambda}\sum_{|\log(x/n)|\leq u}|a_{n}|\left(\frac{x}{n}\right)^{\kappa}\frac{\mathrm{d}u}{u^{2}}
xκλn1|an|nκ+θ/Tλ|log(x/n)|u|an|eκuu2du,\displaystyle\leq\frac{x^{\kappa}}{\lambda}\sum_{n\geq 1}\frac{|a_{n}|}{n^{\kappa}}+\int_{\theta^{\prime}/T}^{\lambda}\sum_{|\log(x/n)|\leq u}|a_{n}|\frac{e^{\kappa u}}{u^{2}}\mathrm{d}u,

from which Theorem 2.1 follows. ∎

3. Zero-free regions

There are a range of explicit zero-free regions for ζ(s)\zeta(s) in the literature. We will combine several existing results to have an optimal zero-free region for different heights in the complex plane. There is no need for zero-free regions for small (s)\Im(s), however, as the Riemann hypothesis has been verified up to a height HH. The most recent computation of HH is from Platt and Trudgian [31].

Lemma 3.1 ([31]).

If ζ(β+it)=0\zeta(\beta+it)=0 for any |t|31012|t|\leq 3\cdot 10^{12} then β=12\beta=\frac{1}{2}.

The most recent explicit version of the “classical” zero-free region is from Mossinghoff and Trudgian [27].

Lemma 3.2 ([27]).

For |t|2|t|\geq 2 there are no zeros of ζ(β+it)\zeta(\beta+it) in the region β1ν1(t)\beta\geq 1-\nu_{1}(t), where

ν1(t)=1R0log|t|\nu_{1}(t)=\frac{1}{R_{0}\log|t|}

andaaaThis value of R0R_{0} is lower than that appearing in [27, Theorem 1]. However, since the Riemann hypothesis has now been verified to a higher height (Lemma 3.1), we can take R0=5.5666305R_{0}=5.5666305 as discussed in [27, Section 6.1]. R0=5.5666305R_{0}=5.5666305.

Ford [16] gave a version of the classical result which improves on Mossinghoff and Trudgian’s result for large |t||t|.

Lemma 3.3.

For |t|5.45108|t|\geq 5.45\cdot 10^{8} there are no zeros of ζ(β+it)\zeta(\beta+it) in the region β1ν2(t)\beta\geq 1-\nu_{2}(t), where

ν2(t)=1R(|t|)log|t|,\nu_{2}(t)=\frac{1}{R(|t|)\log|t|},

with

R(t)=J(t)+0.685+0.155loglogtlogt(0.049620.0196J(t)+1.15),R(t)=\frac{J(t)+0.685+0.155\log\log t}{\log t\left(0.04962-\frac{0.0196}{J(t)+1.15}\right)},

and

J(t)=16logt+loglogt+log(0.63).J(t)=\frac{1}{6}\log t+\log\log t+\log(0.63).
Proof.

This result is almost identical to [16, Theorem 3], except we use an improved constant in J(t)J(t). Ford’s (1.6) can be replaced with a more recent result due to Hiary [19, Theorem 1.1]. We note however, that an error was recently discovered in [19] (see [29, §2.2.1]). Despite this, Hiary’s result can be recovered (and in fact improved) as discussed in the preprint [20]. ∎

For very large tt, we use an explicit Vinogradov–Korobov zero-free region, also due to Ford [16].

Lemma 3.4 ([16, Theorem 5]).

For |t|3|t|\geq 3 there are no zeros of ζ(β+it)\zeta(\beta+it) in the region β1ν3(t)\beta\geq 1-\nu_{3}(t) where

ν3(t)=1clog2/3|t|(loglog|t|)1/3\nu_{3}(t)=\frac{1}{c\log^{2/3}|t|(\log\log|t|)^{1/3}} (3.1)

and c=57.54c=57.54.

To use the widest zero-free region we set

ν(t)={12,if |t|31012,max{ν1(t),ν2(t),ν3(t)},otherwise,\nu(t)=\begin{cases}\frac{1}{2},&\text{if $|t|\leq 3\cdot 10^{12}$},\\ \max\{\nu_{1}(t),\nu_{2}(t),\nu_{3}(t)\},&\text{otherwise,}\end{cases} (3.2)

noting that ν2(t)ν1(t)\nu_{2}(t)\geq\nu_{1}(t) for texp(46.3)t\geq\exp(46.3) and ν3(t)ν2(t)\nu_{3}(t)\geq\nu_{2}(t) for texp(54599)t\geq\exp(54599).

4. Wolke’s method

In this section we prove an explicit bound for the integral in Theorem 2.1,

12πiκiTκ+iTF(s)xssds,\frac{1}{2\pi i}\int_{\kappa-iT}^{\kappa+iT}F(s)\frac{x^{s}}{s}\mathrm{d}s,

for the case κ=1+1/logx\kappa=1+1/\log x and F(s)=n1Λ(n)ns=(ζ/ζ)(s)F(s)=\sum_{n\geq 1}\Lambda(n)n^{-s}=-(\zeta^{\prime}/\zeta)(s).

Theorem 4.1.

Let α(0,1]\alpha\in(0,1] and ω[0,1]\omega\in[0,1]. There exists constants KK and xKx_{K} such that if xxKx\geq x_{K} and max{51,logx}<T<xα21\max\{51,\log x\}<T<\frac{x^{\alpha}}{2}-1,

12πi1+εiT1+ε+iT(ζζ(s))xssds=xρ=β+iγ|γ|Txρρ+O(KxT(logx)1ω),\frac{1}{2\pi i}\int_{1+\varepsilon-iT}^{1+\varepsilon+iT}\left(-\frac{\zeta^{\prime}}{\zeta}(s)\right)\frac{x^{s}}{s}\mathrm{d}s=x-\sum_{\begin{subarray}{c}\rho=\beta+i\gamma\\ |\gamma|\leq T\end{subarray}}\frac{x^{\rho}}{\rho}+O^{*}\left(\frac{Kx}{T}(\log x)^{1-\omega}\right), (4.1)

where ε=1/logx\varepsilon=1/\log x. Some corresponding values of α\alpha, ω\omega, KK and xKx_{K} are given in Table 3 in the appendix.

Although setting ω=1\omega=1 gives the best result asymptotically, the resulting constant will be quite large unless xKx_{K} is very large. For the best results we take ω\omega closer to 1 for large xx and small TT, and ω\omega closer to 0 for small xx and large TT.

To prove Theorem 4.1 we make [38, Theorem 2] explicit. The restriction T>51T>51 is so we can directly use some results from [13], which contains a proof of an expression similar to (4.1) albeit with a weaker error term. The restriction T>logxT>\log x is not necessary but allows us to obtain a better value of KK for large xx.

In what follows, let ω¯0\overline{\omega}\geq 0 be a parameter to be optimised in our computations. The value ω\omega appearing in Theorem 4.1 will be given by

ω={ω¯,if ω¯<1,1,if ω¯1.\omega=\begin{cases}\overline{\omega},&\text{if $\overline{\omega}<1$},\\ 1,&\text{if $\overline{\omega}\geq 1$}.\end{cases}

We also make use of explicit zero-free regions, and set ν(t)\nu(t) as in (3.2).

We now prove a couple of preliminary lemmas. In what follows, N(T)N(T) denotes the number of zeros of ζ(s)\zeta(s) with imaginary part up to height TT.

Lemma 4.2.

Let t>1t>1. Then N(t+1)N(t1)<logtN(t+1)-N(t-1)<\log t.

Proof.

Dudek [13, Lemma 2.6] proves this lemma for t>50t>50. For 1<t501<t\leq 50 we used Odlyzko’s tables [28] to verify the lemma manually. ∎

Lemma 4.3.

Let T>51T>51 and x>ex>e. There exists a τ(T1/(logx)ω¯,T+1/(logx)ω¯)\tau\in(T-1/(\log x)^{\overline{\omega}},T+1/(\log x)^{\overline{\omega}}) such that when s=σ+iτs=\sigma+i\tau with σ>1\sigma>-1, we have

|ζζ(s)|<(logx)ω¯(log2T+logT)+20logT.\left|\frac{\zeta^{\prime}}{\zeta}(s)\right|<(\log x)^{\overline{\omega}}(\log^{2}T+\log T)+20\log T.
Proof.

By Lemma 4.2,

N(T+1/(logx)ω¯)N(T1/(logx)ω¯)N(T+1)N(T1)<logT.N(T+1/(\log x)^{\overline{\omega}})-N(T-1/(\log x)^{\overline{\omega}})\leq N(T+1)-N(T-1)<\log T. (4.2)

As there are at most logT\log T zeros of ζ(s)\zeta(s) with imaginary part in the interval (T1/(logx)ω¯,T+1/(logx)ω¯)(T-1/(\log x)^{\overline{\omega}},T+1/(\log x)^{\overline{\omega}}), we can split (T1/(logx)ω¯,T+1/(logx)ω¯)(T-1/(\log x)^{\overline{\omega}},T+1/(\log x)^{\overline{\omega}}) into at most logT+1\lfloor\log T\rfloor+1 zero-free regions. At least one of these regions will have height greater than

2(logx)ω¯(logT+1).\frac{2}{(\log x)^{\overline{\omega}}(\log T+1)}.

We define τ\tau to be the midpoint of such a region so that

|τγ|1(logx)ω¯(logT+1)|\tau-\gamma|\geq\frac{1}{(\log x)^{\overline{\omega}}(\log T+1)} (4.3)

for all zeros ρ=β+iγ\rho=\beta+i\gamma. Next we use the following result which holds for all σ>1\sigma>-1 and τ>50\tau>50 [13, p. 187]

ζζ(s)=|γτ|<11sρ+O(19logτ).\frac{\zeta^{\prime}}{\zeta}(s)=\sum_{|\gamma-\tau|<1}\frac{1}{s-\rho}+O^{*}(19\log\tau). (4.4)

Using (4.2) and (4.3),

||γτ|<11sρ||γτ|<11|τγ|<(logx)ω¯(logT+1)logT\left|\sum_{|\gamma-\tau|<1}\frac{1}{s-\rho}\right|\leq\sum_{|\gamma-\tau|<1}\frac{1}{|\tau-\gamma|}<(\log x)^{\overline{\omega}}(\log T+1)\log T

which completes the proof of the lemma upon noting that 19logτ20logT19\log\tau\leq 20\log T. ∎

We are now in a position to prove Theorem 4.1.

Proof of Theorem 4.1.

Letting τ\tau be as in Lemma 4.3, we define the contour C=C1C2C3C4C=C_{1}\cup C_{2}\cup C_{3}\cup C_{4}, where

C1=[1+εiτ,1+ε+iτ],\displaystyle C_{1}=[1+\varepsilon-i\tau,1+\varepsilon+i\tau], C2=[1+ε+iτ,1+iτ],\displaystyle C_{2}=[1+\varepsilon+i\tau,-1+i\tau],
C3=[1+iτ,1iτ],\displaystyle C_{3}=[-1+i\tau,-1-i\tau], C4=[1iτ,1+εiτ].\displaystyle C_{4}=[-1-i\tau,1+\varepsilon-i\tau].

Cauchy’s residue theorem gives

12πi1+εiτ1+ε+iτ\displaystyle\frac{1}{2\pi i}\int_{1+\varepsilon-i\tau}^{1+\varepsilon+i\tau} (ζζ(s))xssds\displaystyle\left(-\frac{\zeta^{\prime}}{\zeta}(s)\right)\frac{x^{s}}{s}\mathrm{d}s
=x|γ|τxρρlog(2π)12πiC2C3C4(ζζ(s))xssds,\displaystyle=x-\sum_{|\gamma|\leq\tau}\frac{x^{\rho}}{\rho}-\log(2\pi)-\frac{1}{2\pi i}\int_{C_{2}\cup C_{3}\cup C_{4}}\left(-\frac{\zeta^{\prime}}{\zeta}(s)\right)\frac{x^{s}}{s}\mathrm{d}s, (4.5)

noting that (ζ/ζ)(0)=log(2π)(\zeta^{\prime}/\zeta)(0)=\log(2\pi). We begin by converting the left-hand side of (4) into an integral involving TT as opposed to τ\tau. To do this, we note that when (s)=1+ε\Re(s)=1+\varepsilon, the main theorem in [12] gives

|ζζ(s)||ζζ(1+ε)|<logx.\left|\frac{\zeta^{\prime}}{\zeta}(s)\right|\leq\left|\frac{\zeta^{\prime}}{\zeta}(1+\varepsilon)\right|<\log x.

Hence,

|12πi1+ε+iT1+ε+iτ(ζζ(s))xssds|\displaystyle\left|\frac{1}{2\pi i}\int_{1+\varepsilon+iT}^{1+\varepsilon+i\tau}\left(-\frac{\zeta^{\prime}}{\zeta}(s)\right)\frac{x^{s}}{s}\mathrm{d}s\right| ex2π(T1)|τT|logx\displaystyle\leq\frac{ex}{2\pi(T-1)}|\tau-T|\log x
ex2π(T1)(logx)1ω¯\displaystyle\leq\frac{ex}{2\pi(T-1)}(\log x)^{1-\overline{\omega}} (4.6)

and identically for the integral from 1+εiτ1+\varepsilon-i\tau to 1+εiT1+\varepsilon-iT. Therefore,

12πi1+εiτ1+ε+iτ(ζζ(s))xssds=12πi1+εiT1+ε+iT\displaystyle\frac{1}{2\pi i}\int_{1+\varepsilon-i\tau}^{1+\varepsilon+i\tau}\left(-\frac{\zeta^{\prime}}{\zeta}(s)\right)\frac{x^{s}}{s}\mathrm{d}s=\frac{1}{2\pi i}\int_{1+\varepsilon-iT}^{1+\varepsilon+iT} (ζζ(s))xssds\displaystyle\left(-\frac{\zeta^{\prime}}{\zeta}(s)\right)\frac{x^{s}}{s}\mathrm{d}s
+O(exπ(T1)(logx)1ω¯).\displaystyle+O^{*}\left(\frac{ex}{\pi(T-1)}(\log x)^{1-\overline{\omega}}\right).

Next we consider the sum on the right-hand side of (4). In the case T<τT<\tau,

|γ|τxρρ=|γ|Txρρ+T<|γ|τxρρ.\sum_{|\gamma|\leq\tau}\frac{x^{\rho}}{\rho}=\sum_{|\gamma|\leq T}\frac{x^{\rho}}{\rho}+\sum_{T<|\gamma|\leq\tau}\frac{x^{\rho}}{\rho}.

By (4.2) we then have

|T<|γ|τxρρ|2logτx1ν(τ)T2xlog(T+1)xν(T+1)T.\left|\sum_{T<|\gamma|\leq\tau}\frac{x^{\rho}}{\rho}\right|\leq 2\log\tau\frac{x^{1-\nu(\tau)}}{T}\leq\frac{2x\log(T+1)}{x^{\nu(T+1)}T}.

The case T>τT>\tau is similar, with an error bounded by 2xlogT/xν(T)(T1)2x\log T/x^{\nu(T)}(T-1). Hence,

|γ|τxρρ=|γ|Txρρ+O(2xlog(T+1)xν(T+1)(T1)).\sum_{|\gamma|\leq\tau}\frac{x^{\rho}}{\rho}=\sum_{|\gamma|\leq T}\frac{x^{\rho}}{\rho}+O^{*}\left(\frac{2x\log(T+1)}{x^{\nu(T+1)}(T-1)}\right). (4.7)

Finally we deal with the integral on the right-hand side of (4). For the integral over C3C_{3} we use the inequality [13, Lemma 2.3]

|ζζ(s)|<9+log|s|,\left|\frac{\zeta^{\prime}}{\zeta}(s)\right|<9+\log|s|,

which holds when (s)=1\Re(s)=-1. In particular,

|12πiC3(ζζ(s))xssds|\displaystyle\left|\frac{1}{2\pi i}\int_{C_{3}}\left(-\frac{\zeta^{\prime}}{\zeta}(s)\right)\frac{x^{s}}{s}\mathrm{d}s\right| 2τ2π9+log(τ2+1)x\displaystyle\leq\frac{2\tau}{2\pi}\frac{9+\log(\sqrt{\tau^{2}+1})}{x}
(T+1)(9+log((T+1)2+1)πx.\displaystyle\leq\frac{(T+1)(9+\log(\sqrt{(T+1)^{2}+1})}{\pi x}.

Next we bound the integral over C2C_{2}. The same bound will also hold for the integral over C4C_{4} by symmetry.

Case 1: ω=0\omega=0.

When ω=ω¯=0\omega=\overline{\omega}=0 we bound the integral over C2C_{2} using Lemma 4.3. That is,

|12πiC2(ζζ(s))xssds|\displaystyle\left|\frac{1}{2\pi i}\int_{C_{2}}\left(-\frac{\zeta^{\prime}}{\zeta}(s)\right)\frac{x^{s}}{s}\mathrm{d}s\right| log2T+21logT2πτ11+εxσdσ\displaystyle\leq\frac{\log^{2}T+21\log T}{2\pi\tau}\int_{-1}^{1+\varepsilon}x^{\sigma}\mathrm{d}\sigma
ex(log2T+21logT)2πlogx(T1).\displaystyle\leq\frac{ex(\log^{2}T+21\log T)}{2\pi\log x(T-1)}. (4.8)

Case 2: ω>0\omega>0.

In the case where ω>0\omega>0 we need to be more careful since the expression in (4) is not O(x/T)O(x/T) (unless further restrictions are placed on TT). To overcome this issue, we let D(0,1)D\in(0,1) be a parameter that we will optimise for different entries in Table 3 and let σ0=σ0(x,α):=1Dν(xα)\sigma_{0}=\sigma_{0}(x,\alpha):=1-D\nu(x^{\alpha}). We then split C2C_{2} into C2=C21C22C_{2}=C_{21}\cup C_{22} where

C21=[1+ε+iτ,σ0+iτ],C22=[σ0+iτ,1+iτ].C_{21}=[1+\varepsilon+i\tau,\sigma_{0}+i\tau],\quad C_{22}=[\sigma_{0}+i\tau,-1+i\tau].

By Lemma 4.3

|12πiC22(ζζ(s))xssds|\displaystyle\left|\frac{1}{2\pi i}\int_{C_{22}}\left(-\frac{\zeta^{\prime}}{\zeta}(s)\right)\frac{x^{s}}{s}\mathrm{d}s\right| (logx)ω¯(log2T+logT)+20logT2πτ1σ0xσdσ\displaystyle\leq\frac{(\log x)^{\overline{\omega}}(\log^{2}T+\log T)+20\log T}{2\pi\tau}\int_{-1}^{\sigma_{0}}x^{\sigma}\mathrm{d}\sigma
x(logx)ω¯1(log2T+logT)+20logT/logx2πxDν(xα)(T1).\displaystyle\leq x\frac{(\log x)^{\overline{\omega}-1}(\log^{2}T+\log T)+20\log T/\log x}{2\pi x^{D\nu(x^{\alpha})}(T-1)}. (4.9)

For the integral over C21C_{21} we follow Wolke and use the following formula for ζ/ζ\zeta^{\prime}/\zeta [35, Lemma 2]. Let, for y=x1/4>1y=x^{1/4}>1,

Λy(n)={Λ(n),1nyΛ(n)log(y2/n)logy,yny2.\Lambda_{y}(n)=\begin{cases}\Lambda(n),&1\leq n\leq y\\ \Lambda(n)\frac{\log(y^{2}/n)}{\log y},&y\leq n\leq y^{2}.\end{cases}

Then if s1s\neq 1 and ss is not a zero of ζ\zeta,

ζζ(s)\displaystyle-\frac{\zeta^{\prime}}{\zeta}(s) =ny2Λy(n)ns+y1sy2(1s)(1s)2logy1logyq=1y2qsy2(2q+s)(2q+s)2\displaystyle=\sum_{n\leq y^{2}}\frac{\Lambda_{y}(n)}{n^{s}}+\frac{y^{1-s}-y^{2(1-s)}}{(1-s)^{2}\log y}-\frac{1}{\log y}\sum_{q=1}^{\infty}\frac{y^{-2q-s}-y^{-2(2q+s)}}{(2q+s)^{2}}
1logyρyρsy2(ρs)(sρ)2\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad-\frac{1}{\log y}\sum_{\rho}\frac{y^{\rho-s}-y^{2(\rho-s)}}{(s-\rho)^{2}}
=Z1(s)+Z2(s)+Z3(s)+Z4(s),say.\displaystyle=Z_{1}(s)+Z_{2}(s)+Z_{3}(s)+Z_{4}(s),\quad\text{say}.

For Z1Z_{1},

|12πiC21Z1(s)xssds|\displaystyle\left|\frac{1}{2\pi i}\int_{C_{21}}Z_{1}(s)\frac{x^{s}}{s}\mathrm{d}s\right| 12πτny2Λy(n)σ01+ε(xn)σdσ\displaystyle\leq\frac{1}{2\pi\tau}\sum_{n\leq y^{2}}\Lambda_{y}(n)\int_{\sigma_{0}}^{1+\varepsilon}\left(\frac{x}{n}\right)^{\sigma}\mathrm{d}\sigma
12π(T1)ny2Λy(n)(x/n)1+εlog(x/n)\displaystyle\leq\frac{1}{2\pi(T-1)}\sum_{n\leq y^{2}}\Lambda_{y}(n)\frac{(x/n)^{1+\varepsilon}}{\log(x/n)}
ex2π(T1)nx1/2Λ(n)nclog(x/n)\displaystyle\leq\frac{ex}{2\pi(T-1)}\sum_{n\leq x^{1/2}}\frac{\Lambda(n)}{n^{c}\log(x/n)}
exπlogx(T1)nx1/2Λ(n)nc.\displaystyle\leq\frac{ex}{\pi\log x(T-1)}\sum_{n\leq x^{1/2}}\frac{\Lambda(n)}{n^{c}}. (4.10)

Then by the corollary of the main theorem in [33],

nx1/2Λ(n)n12logxγ+8log2x12logx,\sum_{n\leq x^{1/2}}\frac{\Lambda(n)}{n}\leq\frac{1}{2}\log x-\gamma+\frac{8}{\log^{2}x}\leq\frac{1}{2}\log x, (4.11)

for x>50x>50. By partial summation, we then deduce from (4.11) that

nx1/2Λ(n)nc1elogx.\sum_{n\geq x^{1/2}}\frac{\Lambda(n)}{n^{c}}\leq\frac{1}{e}\log x.

This gives an O(x/T)O(x/T) bound for (4).

For Z2Z_{2},

|12πiC21Z2(s)xssds|\displaystyle\left|\frac{1}{2\pi i}\int_{C_{21}}Z_{2}(s)\frac{x^{s}}{s}\mathrm{d}s\right| 12πτ3logyσ01+ε(x1σ4+x1σ2)xσdσ\displaystyle\leq\frac{1}{2\pi\tau^{3}\log y}\int_{\sigma_{0}}^{1+\varepsilon}(x^{\frac{1-\sigma}{4}}+x^{\frac{1-\sigma}{2}})x^{\sigma}\mathrm{d}\sigma
2π(T1)3logx(43x1+3(1+ε)4logx+2x1+(1+ε)2logx)\displaystyle\leq\frac{2}{\pi(T-1)^{3}\log x}\left(\frac{4}{3}\frac{x^{\frac{1+3(1+\varepsilon)}{4}}}{\log x}+2\frac{x^{\frac{1+(1+\varepsilon)}{2}}}{\log x}\right)
4x(23e3/4+e1/2)π(T1)3log2x.\displaystyle\leq\frac{4x\left(\frac{2}{3}e^{3/4}+e^{1/2}\right)}{\pi(T-1)^{3}\log^{2}x}.

For Z3Z_{3},

|12πiC21Z3(s)xssds|\displaystyle\left|\frac{1}{2\pi i}\int_{C_{21}}Z_{3}(s)\frac{x^{s}}{s}\mathrm{d}s\right| 12πτ3logyσ01+ε(yσq=1y2q+y2σq=1y4q)xσdσ\displaystyle\leq\frac{1}{2\pi\tau^{3}\log y}\int_{\sigma_{0}}^{1+\varepsilon}\left(y^{-\sigma}\sum_{q=1}^{\infty}y^{-2q}+y^{-2\sigma}\sum_{q=1}^{\infty}y^{-4q}\right)x^{\sigma}\mathrm{d}\sigma
2π(T1)3logx((ex)3/4log(x3/4)1y21+(ex)1/2log(x1/2)1y41)\displaystyle\leq\frac{2}{\pi(T-1)^{3}\log x}\left(\frac{(ex)^{3/4}}{\log(x^{3/4})}\frac{1}{y^{2}-1}+\frac{(ex)^{1/2}}{\log(x^{1/2})}\frac{1}{y^{4}-1}\right)
=2xπ(T1)3log2x(43e3/4x3/4x1/4+2e1/2x3/2x1/2).\displaystyle=\frac{2x}{\pi(T-1)^{3}\log^{2}x}\left(\frac{4}{3}\frac{e^{3/4}}{x^{3/4}-x^{1/4}}+\frac{2e^{1/2}}{x^{3/2}-x^{1/2}}\right).

For Z4Z_{4}, we first note that

1+εσ0=1logx+Dν(xα)=dα(x)ν(xα),1+\varepsilon-\sigma_{0}=\frac{1}{\log x}+D\nu(x^{\alpha})=d_{\alpha}(x)\nu(x^{\alpha}), (4.12)

where dα(x):=D+1/ν(xα)logxd_{\alpha}(x):=D+1/\nu(x^{\alpha})\log x is non-increasing. Moreover, for a zero ρ=β+iγ\rho=\beta+i\gamma with |γ|2T+2<xα|\gamma|\leq 2T+2<x^{\alpha},

σ0β>(1D)ν(xα)=:να(x).\sigma_{0}-\beta>(1-D)\nu(x^{\alpha})=:\nu_{\alpha}(x). (4.13)

Now,

|12πiC21Z4(s)xssds|\displaystyle\left|\frac{1}{2\pi i}\int_{C_{21}}Z_{4}(s)\frac{x^{s}}{s}\mathrm{d}s\right| (4.14)
12πlogyρ|γ|2T+2exτσ01+εyβσ+y2(βσ)|σ+iτρ|2dσ\displaystyle\qquad\leq\frac{1}{2\pi\log y}\sum_{\begin{subarray}{c}\rho\\ |\gamma|\leq 2T+2\end{subarray}}\frac{ex}{\tau}\int_{\sigma_{0}}^{1+\varepsilon}\frac{y^{\beta-\sigma}+y^{2(\beta-\sigma)}}{|\sigma+i\tau-\rho|^{2}}\mathrm{d}\sigma
+12πlogyρ|γ|>2T+2xτσ01+ε(xy)σ1yβ1+(xy2)σ1y2(β1)|σ+iτρ|2dσ\displaystyle\qquad\qquad+\frac{1}{2\pi\log y}\sum_{\begin{subarray}{c}\rho\\ |\gamma|>2T+2\end{subarray}}\frac{x}{\tau}\int_{\sigma_{0}}^{1+\varepsilon}\frac{\left(\frac{x}{y}\right)^{\sigma-1}y^{\beta-1}+\left(\frac{x}{y^{2}}\right)^{\sigma-1}y^{2(\beta-1)}}{|\sigma+i\tau-\rho|^{2}}\mathrm{d}\sigma
=2xπτlogx(eF1(x,τ)+F2(x,τ)),say.\displaystyle\qquad=\frac{2x}{\pi\tau\log x}\left(eF_{1}(x,\tau)+F_{2}(x,\tau)\right),\quad\text{say.} (4.15)

To bound F1(x,τ)F_{1}(x,\tau) we will use the estimate

σ01+εdσ|σ+iτρ|21|τγ|2dα(x)v(xα).\int_{\sigma_{0}}^{1+\varepsilon}\frac{\mathrm{d}\sigma}{|\sigma+i\tau-\rho|^{2}}\leq\frac{1}{|\tau-\gamma|^{2}}d_{\alpha}(x)v(x^{\alpha}). (4.16)

whenever |τγ|>2|\tau-\gamma|>2. On the other hand, when |τγ|2|\tau-\gamma|\leq 2, we use the estimate

σ01+εdσ|σ+iτρ|2σ01+εdσ(σβ)2\displaystyle\int_{\sigma_{0}}^{1+\varepsilon}\frac{\mathrm{d}\sigma}{|\sigma+i\tau-\rho|^{2}}\leq\int_{\sigma_{0}}^{1+\varepsilon}\frac{\mathrm{d}\sigma}{(\sigma-\beta)^{2}} =1σ0β11+εβ\displaystyle=\frac{1}{\sigma_{0}-\beta}-\frac{1}{1+\varepsilon-\beta}
<dα(x)(dα(x)+1D)vα(x).\displaystyle<\frac{d_{\alpha}(x)}{(d_{\alpha}(x)+1-D)v_{\alpha}(x)}. (4.17)

Using (4.16), (4) and Lemma 4.2 we have that F1(x,τ)F_{1}(x,\tau) is bounded above by

1xνα(x)/4+xνα(x)/2(|γ|2T+2|τγ|2+h=2|γ|2T+22(h1)<|τγ|2h)σ01+εdσ|σ+iτρ|2\displaystyle\frac{1}{x^{\nu_{\alpha}(x)/4}+x^{\nu_{\alpha}(x)/2}}\left(\sum_{\begin{subarray}{c}|\gamma|\leq 2T+2\\ |\tau-\gamma|\leq 2\end{subarray}}+\sum_{h=2}^{\infty}\sum_{\begin{subarray}{c}|\gamma|\leq 2T+2\\ 2(h-1)<|\tau-\gamma|\leq 2h\end{subarray}}\right)\int_{\sigma_{0}}^{1+\varepsilon}\frac{d\sigma}{|\sigma+i\tau-\rho|^{2}}
1xνα(x)/4+xνα(x)/2(2dα(x)logτ(dα(x)+1D)vα(x)\displaystyle\leq\frac{1}{x^{\nu_{\alpha}(x)/4}+x^{\nu_{\alpha}(x)/2}}\left(\frac{2d_{\alpha}(x)\log\tau}{(d_{\alpha}(x)+1-D)v_{\alpha}(x)}\right.
+2dα(x)ν(xα)log(2T+1)m=11(2m)2)\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\left.+2d_{\alpha}(x)\nu(x^{\alpha})\log(2T+1)\sum_{m=1}^{\infty}\frac{1}{(2m)^{2}}\right)
2dα(x)log(xα)xνα(x)/4+xνα(x)/2(1(dα(x)+1D)vα(x)+v(xα)π224).\displaystyle\leq\frac{2d_{\alpha}(x)\log(x^{\alpha})}{x^{\nu_{\alpha}(x)/4}+x^{\nu_{\alpha}(x)/2}}\left(\frac{1}{(d_{\alpha}(x)+1-D)v_{\alpha}(x)}+\frac{v(x^{\alpha})\pi^{2}}{24}\right). (4.18)

Then,

F2(x,τ)\displaystyle F_{2}(x,\tau) ρ|γ|>2T+21|γ2|2σ01+ε(x3/4)σ1+(x1/2)σ1dσ\displaystyle\leq\sum_{\begin{subarray}{c}\rho\\ |\gamma|>2T+2\end{subarray}}\frac{1}{\left|\frac{\gamma}{2}\right|^{2}}\int_{\sigma_{0}}^{1+\varepsilon}(x^{3/4})^{\sigma-1}+(x^{1/2})^{\sigma-1}\mathrm{d}\sigma
4(e3/4log(x3/4)+e1/2log(x1/2))ρ1|γ|2\displaystyle\leq 4\left(\frac{e^{3/4}}{\log(x^{3/4})}+\frac{e^{1/2}}{\log(x^{1/2})}\right)\sum_{\rho}\frac{1}{|\gamma|^{2}}
=4logx(43e3/4+2e1/2)ρ1|γ|2,\displaystyle=\frac{4}{\log x}\left(\frac{4}{3}e^{3/4}+2e^{1/2}\right)\sum_{\rho}\frac{1}{|\gamma|^{2}},

where ρ1/|γ|20.04621\sum_{\rho}1/|\gamma|^{2}\leq 0.04621 by [3, Example 1].

Combining all of our estimates we see that

12πi1+εiT1+ε+iT(ζζ(s))xssds=x|γ|Txρρ+E(x,T),\frac{1}{2\pi i}\int_{1+\varepsilon-iT}^{1+\varepsilon+iT}\left(-\frac{\zeta^{\prime}}{\zeta}(s)\right)\frac{x^{s}}{s}\mathrm{d}s=x-\sum_{|\gamma|\leq T}\frac{x^{\rho}}{\rho}+E(x,T),

where |E(x,T)||E(x,T)| is bounded above by

log(2π)+exπ(T1)logx+2xlog(T+1)xν(T+1)T\displaystyle\log(2\pi)+\frac{ex}{\pi(T-1)}\log x+\frac{2x\log(T+1)}{x^{\nu(T+1)}T} +(T+1)(9+log((T+1)2+1))πx\displaystyle+\frac{(T+1)(9+\log(\sqrt{(T+1)^{2}+1}))}{\pi x}
+ex(log2T+21logT)πlogx(T1)\displaystyle\qquad+\frac{ex(\log^{2}T+21\log T)}{\pi\log x(T-1)} (4.19)

when ω=ω¯=0\omega=\overline{\omega}=0 and

log(2π)+exπ(T1)(logx)1ω¯+2xlog(T+1)xν(T+1)T+(T+1)(9+log((T+1)2+1))πx\displaystyle\log(2\pi)+\frac{ex}{\pi(T-1)}(\log x)^{1-\overline{\omega}}+\frac{2x\log(T+1)}{x^{\nu(T+1)}T}+\frac{(T+1)(9+\log(\sqrt{(T+1)^{2}+1}))}{\pi x}
+2(x(logx)ω¯1(log2T+logT)+20logT/logx2πxDν(xα)(T1)+xπ(T1)\displaystyle\ +2\left(x\frac{(\log x)^{\overline{\omega}-1}(\log^{2}T+\log T)+20\log T/\log x}{2\pi x^{D\nu(x^{\alpha})}(T-1)}+\frac{x}{\pi(T-1)}\right.
+4x(23e3/4+e1/2)π(T1)3log2x+2xπ(T1)3log2x(43e3/4x3/4x1/4+2e1/2x3/2x1/2)\displaystyle\qquad\quad+\frac{4x\left(\frac{2}{3}e^{3/4}+e^{1/2}\right)}{\pi(T-1)^{3}\log^{2}x}+\frac{2x}{\pi(T-1)^{3}\log^{2}x}\left(\frac{4}{3}\frac{e^{3/4}}{x^{3/4}-x^{1/4}}+\frac{2e^{1/2}}{x^{3/2}-x^{1/2}}\right)
+4αexdα(x)π(T1)(xνα(x)/4+xνα(x)/2)(1(dα(x)+1D)vα(x)+v(xα)π224)\displaystyle\qquad\quad+\frac{4\alpha exd_{\alpha}(x)}{\pi(T-1)(x^{\nu_{\alpha}(x)/4}+x^{\nu_{\alpha}(x)/2})}\left(\frac{1}{(d_{\alpha}(x)+1-D)v_{\alpha}(x)}+\frac{v(x^{\alpha})\pi^{2}}{24}\right)
+8xπ(T1)log2x(43e3/4+2e1/2)ρ1|γ|2)\displaystyle\left.\qquad\quad+\frac{8x}{\pi(T-1)\log^{2}x}\left(\frac{4}{3}e^{3/4}+2e^{1/2}\right)\sum_{\rho}\frac{1}{|\gamma|^{2}}\right) (4.20)

otherwise. Note that the terms multiplied by 2 in (4) are those that occur when bounding both the integral over C2C_{2} and C4C_{4}. Thus, we can write E(x,T)=O(Kx(logx)1ω/T)E(x,T)=O^{*}(Kx(\log x)^{1-\omega}/T) where KK can be computed by evaluating an upper bound for each error term in (4) or (4) in the range xxKx\geq x_{K}. ∎

Extra care must be taken when bounding some of the error terms in the above proof due to ν(t)\nu(t) being defined as a composite function. Most notably, we need to be careful of the behaviour at the crossover point λ=54598.16\lambda=54598.16\ldots such that ν3(t)ν2(t)\nu_{3}(t)\geq\nu_{2}(t) for all texp(λ)t\geq\exp(\lambda) or equivalently ν3(tα)ν2(tα)\nu_{3}(t^{\alpha})\geq\nu_{2}(t^{\alpha}) for all texp(λ/α)t\geq\exp(\lambda/\alpha).

For example, the error in (4) can be written as H(x,T)(x(logx)1ω/T)H(x,T)(x(\log x)^{1-\omega}/T) where HH decreases in xx whilst ν(T)=1/2\nu(T)=1/2, and increases in xx whilst ν(T)=ν1(T)\nu(T)=\nu_{1}(T) or ν(T)=ν2(T)\nu(T)=\nu_{2}(T). Thus, in most cases, an upper bound for H(x,T)H(x,T) occurs at or beyond x=exp(λ/α)x=\exp(\lambda/\alpha) corresponding to the upper bound T<xα24<xαT<\frac{x^{\alpha}-2}{4}<x^{\alpha} for TT. As a result, in the cases where xK<exp(λ/α)x_{K}<\exp(\lambda/\alpha) we evaluated (4) at x=exp(λ/α)x=\exp(\lambda/\alpha) to obtain an upper bound. In such cases we also had to verify that (4) was decreasing for xexp(λ/α)x\geq\exp(\lambda/\alpha). A similar treatment is required for the terms in (4.7) and (4).

We also remark that the error term

2xπ(T1)\frac{2x}{\pi(T-1)}

in (4) is bounded below by 2x/πT2x/\pi T independent of the choice of α\alpha or ω¯\overline{\omega}. Thus, KK is bounded below by 2/π=0.636612/\pi=0.63661\ldots when ω=1\omega=1.

Since our computations rely heavily on zero-free regions, we can obtain significant improvements by assuming the Riemann hypothesis. In particular, we have the following result.

Theorem 4.4.

Assuming the Riemann hypothesis, we can take K=0.6373K=0.6373 in (4.1) for ω=1\omega=1, α=1/2\alpha=1/2 and logx1000\log x\geq 1000.

Proof.

Redefine ν(t)=12\nu(t)=\frac{1}{2} in the proof of Theorem 4.1. The parameters we used to obtain K=0.6373K=0.6373 were D=0.4D=0.4 and ω¯=3\overline{\omega}=3. ∎

Remark.

With more work, the value of KK in Theorem 4.4 can be lowered and in fact made arbitrarily small as xx\to\infty. In particular, one could bound the integral over C2C_{2} using conditional estimates on ζ/ζ\zeta^{\prime}/\zeta (e.g. [37, Corollary 1]).

5. Proof of Theorem 1.2

Using the results of Sections 2 and 4, we now prove Theorem 1.2. To begin with, we let xxMe40x\geq x_{M}\geq e^{40}, max{51,logx}<T<xα21\max\{51,\log x\}<T<\frac{x^{\alpha}}{2}-1, κ=1+1/logx\kappa=1+1/\log x and set an=Λ(n)a_{n}=\Lambda(n) in Theorem 2.1 to obtain

ψ(x)=12πiκiTκ+iT(ζζ(s))xssds\displaystyle\psi(x)=\frac{1}{2\pi i}\int_{\kappa-iT}^{\kappa+iT}\left(-\frac{\zeta^{\prime}}{\zeta}(s)\right)\frac{x^{s}}{s}\mathrm{d}s
+O(xκπλTn1Λ(n)nκ+1πTθ/Tλ|log(x/n)|uΛ(n)eκuu2du).\displaystyle\qquad\qquad\qquad+O^{*}\left(\frac{x^{\kappa}}{\pi\lambda T}\sum_{n\geq 1}\frac{\Lambda(n)}{n^{\kappa}}+\frac{1}{\pi T}\int_{\theta^{\prime}/T}^{\lambda}\sum_{|\log(x/n)|\leq u}\Lambda(n)\frac{e^{\kappa u}}{u^{2}}\mathrm{d}u\right). (5.1)

For the first term, we have by Theorem 4.1,

12πiκiTκ+iT(ζζ(s))xssds=x+|γ|Txρρ+O(KxT(logx)1ω),\frac{1}{2\pi i}\int_{\kappa-iT}^{\kappa+iT}\left(-\frac{\zeta^{\prime}}{\zeta}(s)\right)\frac{x^{s}}{s}\mathrm{d}s=x+\sum_{|\gamma|\leq T}\frac{x^{\rho}}{\rho}+O^{*}\left(K\frac{x}{T}(\log x)^{1-\omega}\right), (5.2)

for some constant KK corresponding to xK=xMx_{K}=x_{M}. Then, the first sum in (5) is bounded by

xκλπTn1Λ(n)nκexlogxλπT\frac{x^{\kappa}}{\lambda\pi T}\sum_{n\geq 1}\frac{\Lambda(n)}{n^{\kappa}}\leq\frac{ex\log x}{\lambda\pi T} (5.3)

by the main theorem in [12]. For the second term in (5), the condition |log(x/n)|u|\log(x/n)|\leq u is equivalent to xeunxeuxe^{-u}\leq n\leq xe^{u}. For all u[0,λ]u\in[0,\lambda] we have eu1ue^{-u}\geq 1-u and eu1+(c01)ue^{u}\leq 1+(c_{0}-1)u with

c0=c0(λ)=eλ1λ+1.c_{0}=c_{0}(\lambda)=\frac{e^{\lambda}-1}{\lambda}+1.

Hence, the sum simplifies to

1πTθ/Tλ|log(x/n)|uΛ(n)eκuu2du\displaystyle\frac{1}{\pi T}\int_{\theta^{\prime}/T}^{\lambda}\sum_{|\log(x/n)|\leq u}\Lambda(n)\frac{e^{\kappa u}}{u^{2}}\mathrm{d}u eκλπTθ/Tλ1u2I(x,u)Λ(n)du,\displaystyle\leq\frac{e^{\kappa\lambda}}{\pi T}\int^{\lambda}_{\theta^{\prime}/T}\frac{1}{u^{2}}\sum_{I(x,u)}\Lambda(n)\mathrm{d}u, (5.4)

where

I(x,u)={n1:xuxnx+(c01)ux}.I(x,u)=\left\{n\geq 1\>:\>x-ux\leq n\leq x+(c_{0}-1)ux\right\}.

Let x=max{xux,0}x_{-}=\max\{x-ux,0\} and x+=x+(c01)uxx_{+}=x+(c_{0}-1)ux for θ/Tuλ\theta^{\prime}/T\leq u\leq\lambda. Defining

θ(x)=pxlogp,\theta(x)=\sum_{p\leq x}\log p,

we have by an explicit form of the Brun–Titchmarsh theorem [26, Theorem 2],

θ(x+)θ(x)\displaystyle\theta(x_{+})-\theta(x_{-}) 2logx+log(x+x)(x+x)\displaystyle\leq\frac{2\log x_{+}}{\log(x_{+}-x_{-})}(x_{+}-x_{-})
2log(x+(c01)ux)log(c0ux)c0ux\displaystyle\leq\frac{2\log(x+(c_{0}-1)ux)}{\log(c_{0}ux)}c_{0}ux
c0ux1(x,T),\displaystyle\leq c_{0}ux\cdot\mathscr{E}_{1}(x,T),

where

1(x,T)=2log(x+(c01)λx)log(c0θx/T).\mathscr{E}_{1}(x,T)=\frac{2\log(x+(c_{0}-1)\lambda x)}{\log(c_{0}\theta^{\prime}x/T)}.

To obtain the corresponding inequality for ψ\psi, we use [4, Corollary 5.1], which states that for all xe40x\geq e^{40},

0<ψ(x)θ(x)<a1x1/2+a2x1/3,0<\psi(x)-\theta(x)<a_{1}x^{1/2}+a_{2}x^{1/3},

with a1=1+1.93378108a_{1}=1+1.93378\cdot 10^{-8} and a2=1.0432a_{2}=1.0432. Thus,

nI(x,u)Λ(n)\displaystyle\sum_{n\in I(x,u)}\Lambda(n) ψ(x+)ψ(x)+logx\displaystyle\leq\psi(x_{+})-\psi(x_{-})+\log x
<c0ux1(x,T)+a1x++a2x+3+logx\displaystyle<c_{0}ux\cdot\mathscr{E}_{1}(x,T)+a_{1}\sqrt{x_{+}}+a_{2}\sqrt[3]{x_{+}}+\log x
c0ux1(x,T)+2(x),\displaystyle\leq c_{0}ux\cdot\mathscr{E}_{1}(x,T)+\mathscr{E}_{2}(x),

where

2(x)=a1(x+(c01)λx)1/2+a2(x+(c01)λx)1/3+logx.\mathscr{E}_{2}(x)=a_{1}(x+(c_{0}-1)\lambda x)^{1/2}+a_{2}(x+(c_{0}-1)\lambda x)^{1/3}+\log x.

Substituting this into (5.4), we obtain

eκλπTθ/Tλ1u2I(x,u)Λ(n)du\displaystyle\frac{e^{\kappa\lambda}}{\pi T}\int^{\lambda}_{\theta^{\prime}/T}\frac{1}{u^{2}}\sum_{I(x,u)}\Lambda(n)\mathrm{d}u eκλπTθ/Tλ(c0x1(x,T)u+2(x)u2)du\displaystyle\leq\frac{e^{\kappa\lambda}}{\pi T}\int_{\theta^{\prime}/T}^{\lambda}\left(\frac{c_{0}x\cdot\mathscr{E}_{1}(x,T)}{u}+\frac{\mathscr{E}_{2}(x)}{u^{2}}\right)\mathrm{d}u
eκλπ(c0x1(x,T)Tlog(λθT)+2(x)θ).\displaystyle\leq\frac{e^{\kappa\lambda}}{\pi}\left(\frac{c_{0}x\cdot\mathscr{E}_{1}(x,T)}{T}\log\left(\frac{\lambda}{\theta^{\prime}}T\right)+\frac{\mathscr{E}_{2}(x)}{\theta^{\prime}}\right). (5.5)

Combining (5.2), (5.3) and (5), the proof of Theorem 1.2 is complete upon optimising over λ\lambda to compute values of xMx_{M}, α(0,1/2]\alpha\in(0,1/2] and MM such that

ψ(x)=xρ=β+iγ|γ|Txρρ+O(MxlogxT),\psi(x)=x-\sum_{\begin{subarray}{c}\rho=\beta+i\gamma\\ |\gamma|\leq T\end{subarray}}\frac{x^{\rho}}{\rho}+O^{*}\left(\frac{Mx\log x}{T}\right),

for all xxMx\geq x_{M} and max{51,logx}<T<xα21\max\{51,\log x\}<T<\frac{x^{\alpha}}{2}-1. Note we cannot take α>1/2\alpha>1/2, as 2(x)\mathscr{E}_{2}(x) would not be O(xlogx/T)O\left(x\log x/T\right) for values of TT asymptotically larger than x\sqrt{x}.

Under assumption of the Riemann hypothesis, we also have the following.

Theorem 5.1.

Assuming the Riemann hypothesis, we can take M=4.150M=4.150 in (1.5) with logx1000\log x\geq 1000 and α=1/2\alpha=1/2.

Proof.

Calculate MM as above using λ=0.52\lambda=0.52 and the value of K=0.6373K=0.6373 from Theorem 4.4. ∎

6. Application: primes between consecutive powers

For sufficiently large xx, the best unconditional short-interval result for primes is [x,x+x0.525][x,x+x^{0.525}], from Baker, Harman, and Pintz [1]. This can be narrowed to (x,x+Cxlogx](x,x+C\sqrt{x}\log x] for some constant CC if assuming the Riemann hypothesis, as per Cramér’s result [6]. The latest explicit version of this is [5, Thm. 1.5], with C=22/25C=22/25 for all x4x\geq 4. Legendre conjectured that something just better than this should be true: that there should be a prime between n2n^{2} and (n+1)2(n+1)^{2} for all positive integers nn. This is approximately equivalent to primes in intervals of the form (x,x+2x](x,x+2\sqrt{x}]. Although proving Legendre’s conjecture is out of reach even under RH, we do know there are primes between higher consecutive powers. Using Ingham’s method from [21], it was shown in [7] that there are primes between consecutive cubes, n3n^{3} and (n+1)3(n+1)^{3}, for all nexp(exp(32.892))n\geq\exp(\exp(32.892)).

It is also possible to find primes between higher consecutive powers for all positive nn. In [7], this is done using an explicit version of Goldston’s estimate estimate for the error in the prime number theorem [18]. We can now use Theorem 1.2 in place of Goldston’s result, and prove Theorem 1.3.

To consider the interval (nm,(n+1)m)(n^{m},(n+1)^{m}), we set n=x1mn=x^{\frac{1}{m}} and look at the slightly smaller interval (x,x+h](x,x+h] with h=mx11/mh=mx^{1-1/m}. Theorem 1.2 implies

ψ(x+h)ψ(x)hρ=β+iγ|γ|T|(x+h)ρxρρ|MG(x,h)T\psi(x+h)-\psi(x)\geq\ h-\sum_{\begin{subarray}{c}\rho=\beta+i\gamma\\ |\gamma|\leq T\end{subarray}}\left|\frac{(x+h)^{\rho}-x^{\rho}}{\rho}\right|-M\frac{G(x,h)}{T} (6.1)

for all xxMx\geq x_{M} and max{51,logx}<T<(xα2)/2\max\{51,\log x\}<T<(x^{\alpha}-2)/2, where G(x,h)=(x+h)log(x+h)+xlogxG(x,h)=(x+h)\log(x+h)+x\log x, and values for xMx_{M}, α\alpha and MM are given in Table 4. In order to find at least one prime in (x,x+h](x,x+h], (6.1) needs to be positive. Fixing an mm, we want to maximise (6.1) with respect to TT, and solve for xx. This process follows the same steps as in Section 4 of [7], and we set T=xμT=x^{\mu} for some μ(0,1)\mu\in(0,1).bbbNote that μ\mu here is denoted α\alpha in [7]. We then arrive at a similar condition to equation (21) in [7]: that there are primes between consecutive mthm^{\text{th}} powers for all xx satisfying

1F(x)MG(x,h)xμh+E(x)h>01-F(x)-M\frac{G(x,h)}{x^{\mu}h}+\frac{E(x)}{h}>0 (6.2)

where F(x)F(x) and E(x)E(x) are defined as in equation (19) and (20), respectively, of [7]. A zero-free region is used in this definition of F(x)F(x), so we use the ν(t)\nu(t) defined in Section 3 of this paper, instead of that used in [7].

For logxM=103\log x_{M}=10^{3} we can take M=0.6651M=0.6651, so for m=140m=140 we can take μ=0.0080155\mu=0.0080155 to have (6.2) hold for all logx4242\log x\geq 4242. Note that with these values the condition on TT in Theorem 1.2 is satisfied for all logx375\log x\geq 375. The interval estimates for primes in [8] (corrected from the original paper [9]) for x41018x\geq 4\cdot 10^{18} and xe600x\geq e^{600} verify that there are primes between consecutive 140th140^{\text{th}} powers for log(41018)logx4367\log(4\cdot 10^{18})\leq\log x\leq 4367. The calculations of [36] verify the remaining smaller values of xx.

7. Application: the error term in the prime number theorem

The prime number theorem is equivalent to the statement ψ(x)x\psi(x)\sim x. For large xx, the best unconditional estimates on the error |ψ(x)x||\psi(x)-x| are from Platt and Trudgian [32, Theorem 1]. For R=5.5734125R=5.5734125 they give values for XX, AA, BB and CC such that

|ψ(x)xx|A(logxR)Bexp(ClogxR)\left|\frac{\psi(x)-x}{x}\right|\leq A\left(\frac{\log x}{R}\right)^{B}\exp\left(-C\sqrt{\frac{\log x}{R}}\right) (7.1)

for all logxX\log x\geq X. To obtain (7.1), Platt and Trudgian employ a method of Pintz [30] and use Dudek’s error term for the Riemann–von Mangoldt formula (Theorem 1.1).

In this section we prove Theorem 1.4, which is a non-trivial improvement on Platt and Trudgian’s result. Most of the improvement comes from using Theorem 1.2 in place of Dudek’s error term. We will also incorporate some other recent results.

Firstly, we use the smaller value R=5.5666305R=5.5666305 which is the same as R0R_{0} appearing in the zero-free region in see Lemma 3.2. In the following lemma, we also make small improvements to some of the zero-density estimates in [24].

Lemma 7.1.

Let N(σ,T)N(\sigma,T) be the number of zeros ρ=β+iγ\rho=\beta+i\gamma of the Riemann zeta-function with σ<β1\sigma<\beta\leq 1 and 0γT0\leq\gamma\leq T. Then, for the values of C1(σ)C_{1}(\sigma) and C2(σ)C_{2}(\sigma) in Table 2, we have

N(σ,T)C1(σ)T8(1σ)/3(logT)52σ+C2(σ)log2T.N(\sigma,T)\leq C_{1}(\sigma)T^{8(1-\sigma)/3}(\log T)^{5-2\sigma}+C_{2}(\sigma)\log^{2}T.
Proof.

Using Platt and Trudgian’s verification of the Riemann hypothesis up to H0=31012H_{0}=3\cdot 10^{12} [31], and the divisor function estimate in Theorem 2 of [10] to replace (3.13) of [24], we can recalculate the constants in Lemma 4.14 of [24]. Using the notation from [24], we want to optimise over kk, μ\mu, α\alpha, δ\delta, dd, HH, and η\eta. We chose H=H01H=H_{0}-1, η=0.2535\eta=0.2535, k=1k=1, μ=1.237\mu=1.237, δ=0.313\delta=0.313 and optimised over the other parameters for each σ\sigma. ∎

Table 2. Some values of C1(σ)C_{1}(\sigma) and C2(σ)C_{2}(\sigma) for Lemma 7.1.
σ\sigma α\alpha dd C1(σ)C_{1}(\sigma) C2(σ)C_{2}(\sigma)
0.980 0.063 0.336 15.743 2.214
0.982 0.063 0.336 15.878 2.204
0.984 0.061 0.336 16.013 2.187
0.986 0.061 0.336 16.148 2.171
0.988 0.060 0.337 16.284 2.148
0.990 0.060 0.337 16.421 2.132
0.992 0.058 0.337 16.558 2.115
Proof of Theorem 1.4.

We only require a slight modification of the proof of [32, Theorem 1]. First we assume logx1000\log x\geq 1000 so that we may take xM=exp(1000)x_{M}=\exp(1000), α=1/10\alpha=1/10 and M=2.045M=2.045 in Theorem 1.2. The argument in [32, pp. 874–875] then follows through with minimal modification so that if C1(σ)C_{1}(\sigma) and C2(σ)C_{2}(\sigma) are as in Lemma 7.1 and

k(σ,x0)\displaystyle k(\sigma,x_{0}) =[exp(1016σ3logx0R)(logx0R)52σ]1\displaystyle=\left[\exp\left(\frac{10-16\sigma}{3}\sqrt{\frac{\log x_{0}}{R}}\right)\left(\sqrt{\frac{\log x_{0}}{R}}\right)^{5-2\sigma}\right]^{-1}
C3(σ,x0)\displaystyle C_{3}(\sigma,x_{0}) =M(logx0)exp(2logx0R)k(σ,x0)\displaystyle=M(\log x_{0})\exp\left(-2\sqrt{\frac{\log x_{0}}{R}}\right)k(\sigma,x_{0})
C4(σ,x0)\displaystyle C_{4}(\sigma,x_{0}) =x0σ1(2πlogx0R+1.8642)k(σ,x0)\displaystyle=x_{0}^{\sigma-1}\left(\frac{2}{\pi}\frac{\log x_{0}}{R}+1.8642\right)k(\sigma,x_{0})
C5(σ,x0)\displaystyle C_{5}(\sigma,x_{0}) =8.01C2(σ)exp(2logx0R)logx0Rk(σ,x0)\displaystyle=8.01\cdot C_{2}(\sigma)\exp\left(-2\sqrt{\frac{\log x_{0}}{R}}\right)\frac{\log x_{0}}{R}k(\sigma,x_{0})
A(σ,x0)\displaystyle A(\sigma,x_{0}) =2.0025252σC1(σ)+C3(σ,x0)+C4(σ,x0)+C5(σ,x0),\displaystyle=2.0025\cdot 2^{5-2\sigma}\cdot C_{1}(\sigma)+C_{3}(\sigma,x_{0})+C_{4}(\sigma,x_{0})+C_{5}(\sigma,x_{0}),

then

|ψ(x)xx|A(σ,x0)(logxR)52σ2exp(1016σ3logxR)\left|\frac{\psi(x)-x}{x}\right|\leq A(\sigma,x_{0})\left(\frac{\log x}{R}\right)^{\frac{5-2\sigma}{2}}\exp\left(\frac{10-16\sigma}{3}\sqrt{\frac{\log x}{R}}\right)

for all σ[0.75,1)\sigma\in[0.75,1) and x>x0x>x_{0} provided A(σ,x)A(\sigma,x) is decreasing for x>x0x>x_{0}. The only difference between our formulae and those on page 875 of [32] is the expression for C3(σ,x0)C_{3}(\sigma,x_{0}). Taking x0=exp(X)x_{0}=\exp(X) it is then possible to compute the values of AA in Table 1. The values of σ\sigma in Table 1 were chosen to optimise the value of ϵ0\epsilon_{0}. ∎

Remark.

Using Theorem 1.2 meant that the C3(σ,x0)C_{3}(\sigma,x_{0}) term in the above proof became essentially negligible for the values of x0x_{0} considered. Thus, reducing MM is unlikely to substantially improve Theorem 1.4 using the above method.

We also have the following corollary (cf. [32, Corollary 1]).

Corollary 7.2.

For each row {X,A,B,C}\{X,A,B,C\} in Table 1 we have

|θ(x)xx|A1(logxR)Bexp(ClogxR),for alllogxX\left|\frac{\theta(x)-x}{x}\right|\leq A_{1}\left(\frac{\log x}{R}\right)^{B}\exp\left(-C\sqrt{\frac{\log x}{R}}\right),\quad\text{for all}\ \log x\geq X

where A1=A+0.1A_{1}=A+0.1.

Proof.

By [4, Corollary 5.1], we have for xexp(1000)x\geq\exp(1000) that

ψ(x)θ(x)<a1x1/2+a2x1/3,\displaystyle\psi(x)-\theta(x)<a_{1}x^{1/2}+a_{2}x^{1/3},

with a1=1+1.999861012a_{1}=1+1.99986\cdot 10^{-12} and a2=1+1.936108a_{2}=1+1.936\cdot 10^{-8}. The result then follows by a straightforward computation since

|θ(x)xx|ψ(x)θ(x)x+|ψ(x)xx|.\left|\frac{\theta(x)-x}{x}\right|\leq\frac{\psi(x)-\theta(x)}{x}+\left|\frac{\psi(x)-x}{x}\right|.\qed

Corollary 7.2 allows us to improve on existing knowledge of an inequality due to Ramanujan. Namely, in one of his notebooks, Ramanujan proved that

π(x)2<exlogxπ(xe)\pi(x)^{2}<\frac{ex}{\log x}\pi\left(\frac{x}{e}\right) (7.2)

holds for sufficiently large xx [2, pp. 112–114].

It is still an open problem to determine the largest value of xx for which (7.2) holds. However, it is widely believed that the last integer counterexample occurs at x=38,358,837,682x=38,358,837,682. In fact, this follows under assumption of the Riemann hypothesis [14, Theorem 1.3].

Using their expression for |θ(x)x||\theta(x)-x|, Platt and Trudgian [32, Theorem 2] were able to show (unconditionally) that (7.2) holds for xexp(3915)x\geq\exp(3915). Substituting our results for Corollary 7.2 when X=3600X=3600 into the formulae on page 879 of [32] gives a small improvement. In particular, we get that Ramanujan’s inequality (7.2) holds for xexp(3604)x\geq\exp(3604) (Corollary 1.5).

We also note that the second author recently showed that Ramanujan’s inequality (7.2) holds for 38,358,837,683xexp(103)38,358,837,683\leq x\leq\exp(103) [22, Theorem 5.1]. However, we are not able to improve on this result here as Theorem 1.4 and Corollary 7.2 only give good bounds on prime counting functions for large xexp(1000)x\geq\exp(1000).

Acknowledgements

Thanks to all our colleagues at UNSW Canberra. Particularly our supervisor Tim Trudgian for his constant support, Aleks Simonič for helping us with those pesky semicircles, and Ethan Lee for the long discussions over tea. We also thank Ramaré for his insights and correspondence.

Appendix: Tables 3 and 4

Table 3. Some corresponding values of xKx_{K}, α\alpha, ω\omega, DD and KK for Theorem 4.1.
log(xK)\log(x_{K}) α\alpha ω\omega ω¯\overline{\omega} DD KK
4040 1/21/2 0 0 2.0532.053
10310^{3} 1/21/2 0 0 1.6731.673
101010^{10} 1/21/2 0.30.3 0.30.3 0.540.54 3.1913.191
101310^{13} 1/21/2 11 1.41.4 0.500.50 0.63670.6367
10310^{3} 1/101/10 0.20.2 0.20.2 0.450.45 2.5962.596
101010^{10} 1/101/10 0.90.9 0.90.9 0.540.54 11.7711.77
10310^{3} 1/1001/100 0.80.8 0.80.8 0.520.52 2.1862.186
101010^{10} 1/1001/100 11 1.51.5 0.500.50 0.63670.6367
Table 4. Some corresponding values of xMx_{M}, α\alpha, λ\lambda and MM for Theorem 1.2.
log(xM)\log(x_{M}) α\alpha λ\lambda MM
4040 1/21/2 0.480.48 6.4316.431
10310^{3} 1/21/2 0.520.52 5.8235.823
101010^{10} 1/21/2 0.520.52 4.1434.143
101310^{13} 1/21/2 0.520.52 4.1404.140
10310^{3} 1/101/10 1.051.05 2.0452.045
101010^{10} 1/101/10 1.061.06 1.3841.384
10310^{3} 1/1001/100 1.801.80 0.66510.6651
101010^{10} 1/1001/100 1.881.88 0.62690.6269

Note that although we could have considered larger values of α(0,1]\alpha\in(0,1] in Table 3, the restriction α1/2\alpha\leq 1/2 is required for Theorem 1.2.

References

  • [1] R. C. Baker, G. Harman and J. Pintz “The difference between consecutive primes. II” In Proc. London Math. Soc. 83.3 Cambridge University Press, 2001, pp. 532–562
  • [2] B. C. Berndt “Ramanujan’s Notebooks: Part IV” Springer-Verlag, 1994
  • [3] R. Brent, D. Platt and T. Trudgian “Accurate estimation of sums over zeros of the Riemann zeta-function” In Math. Comp., 2021, pp. 2923–2935
  • [4] S. Broadbent, H. Kadiri, A. Lumley, N. Ng and K. Wilk “Sharper bounds for the Chebyshev function θ(x)\theta(x) In Math. Comp. 90.331, 2021, pp. 2281–2315
  • [5] E. Carneiro, M. B. Milinovich and K. Soundararajan “Fourier optimization and prime gaps” In Comment. Math. Helv. 94.3, 2019, pp. 533–568
  • [6] H. Cramér “On the order of magnitude of the difference between consecutive prime numbers” In Acta Arith. 2 Instytut Matematyczny Polskiej Akademii Nauk, 1936, pp. 23–46
  • [7] M. Cully-Hugill “Primes between consecutive powers” In J. Number Theory 247, 2023, pp. 100–117
  • [8] M. Cully-Hugill and E. S. Lee “Explicit interval estimates for prime numbers” In Preprint on arXiv:2103.05986, 2022
  • [9] M. Cully-Hugill and E. S. Lee “Explicit interval estimates for prime numbers” In Math. Comp. 91.336, 2022, pp. 1955–1970
  • [10] M. Cully-Hugill and T. Trudgian “Two explicit divisor sums” In Ramanujan J. 56 Springer, 2021, pp. 141––149
  • [11] H. Davenport “Multiplicative Number Theory” Second edition New York: Springer-Verlag, 1980
  • [12] H. Delange “Une remarque sur la dérivée logarithmique de la fonction zêta de Riemann” In Colloq. Math. 53.2, 1987, pp. 333–335
  • [13] A. W. Dudek “An explicit result for primes between cubes” In Funct. Approx. Comment. Math. 55.2, 2016, pp. 177–197
  • [14] A. W. Dudek and D. J. Platt “On solving a curious inequality of Ramanujan” In Exp. Math. 24.3, 2015, pp. 289–294
  • [15] A. Fiori, H Kadiri and J. Swidinsky “Sharper bounds for the error term in the prime number theorem” In Preprint on arXiv:2206.12557, 2022
  • [16] K. Ford “Zero-free regions for the Riemann zeta function” In Number theory for the millennium, II (Urbana, IL, 2000) A K Peters, Natick, MA, 2002, pp. 25–56
  • [17] D. Goldston “On a result of Littlewood concerning prime numbers” In Acta Arith. 40 Instytut Matematyczny Polskiej Akademii Nauk, 1982, pp. 263–271
  • [18] D. A. Goldston “On a result of Littlewood concerning prime numbers. II” In Acta. Arith. 43.1, 1983, pp. 49–51
  • [19] G. A. Hiary “An explicit van der Corput estimate for ζ(1/2+it)\zeta(1/2+it) In Indag. Math. 27.2 Elsevier, 2016, pp. 524–533
  • [20] G. A Hiary, Patel D. and A. Yang “An improved explicit estimate for ζ(1/2+it)\zeta(1/2+it)” Submitted In Preprint on arXiv:2207.02366, 2022
  • [21] A. E. Ingham “On the difference between consecutive primes” In Q. J. Math 8.1 Oxford University Press, 1937, pp. 255–266
  • [22] D. R. Johnston “Improving bounds on prime counting functions by partial verification of the Riemann hypothesis” In Ramanujan J., 2022
  • [23] D. R. Johnston and A. Yang “Some explicit estimates for the error term in the prime number theorem” In Preprint on arXiv:2204.01980, 2022
  • [24] H. Kadiri, A. Lumley and N. Ng “Explicit zero density for the Riemann zeta function” In J. Math. Anal. Appl. 465.1 Elsevier, 2018, pp. 22–46
  • [25] J. E. Little-Wood “Two notes on the Riemann Zeta-function” In Math. Proc. Cambridge Philos. Soc. 22.3, 1924, pp. 234–242 Cambridge University Press
  • [26] H. L. Montgomery and R. C. Vaughan “The large sieve” In Mathematika 20, 1973, pp. 119–134
  • [27] M. J. Mossinghoff and T. S. Trudgian “Nonnegative trigonometric polynomials and a zero-free region for the Riemann zeta-function” In J. Number Theory 157 Elsevier, 2015, pp. 329–349
  • [28] A. Odlyzko “Tables of zeros of the Riemann zeta function” Visited on 01-Sep-2021 <http://www.dtc.umn.edu/~odlyzko/zeta_tables/index.html>, 2021
  • [29] D. Patel “Explicit sub-Weyl bound for the Riemann zeta function”, 2021
  • [30] J. Pintz “On the remainder term of the prime number formula. II. On a theorem of Ingham” In Acta Arith. 37, 1980, pp. 209–220
  • [31] D. Platt and T. Trudgian “The Riemann hypothesis is true up to 310123\cdot 10^{12} In Bull. Lond. Math. Soc. 53.3, 2021, pp. 792–797
  • [32] D. J. Platt and T. S. Trudgian “The error term in the prime number theorem” In Math. Comp. 90.328, 2021, pp. 871–881
  • [33] O. Ramaré “Explicit estimates for the summatory function of Λ(n)/n\Lambda(n)/n from the one of Λ(n)\Lambda(n) In Acta Arith. 159.2, 2013, pp. 113–122
  • [34] O. Ramaré “Modified truncated Perron formulae” In Ann. Math. Blaise Pascal 23.1, 2016, pp. 109–128
  • [35] A. Selberg “On the normal density of primes in small intervals, and the difference between consecutive primes” In Arch. Math. Naturvid. 47, 1943, pp. 87–105
  • [36] T. Silva, S. Herzog and S. Pardi “Empirical verification of the even Goldbach conjecture and computation of prime gaps up to 410184\cdot 10^{18} In Math. Comp. 83.288, 2014, pp. 2033–2060
  • [37] A. Simonič “Estimates for LL-functions in the critical strip under GRH with effective applications” In Preprint on arXiv:2201.10771, 2022
  • [38] D. Wolke “On the explicit formula of Riemann–von Mangoldt, II” In J. London Math. Soc. 2.3 Oxford University Press, 1983, pp. 406–416