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

The mean square of the error term in the prime number theorem

Richard P. Brent111Australian National University, Canberra, Australia <[email protected]>,  David J. Platt222School of Mathematics, University of Bristol, Bristol, UK <[email protected]>  and Timothy S. Trudgian333School of Science, UNSW Canberra at ADFA, Australia <[email protected]>
Abstract

We show that, on the Riemann hypothesis, lim supXI(X)/X20.8603\limsup_{X\to\infty}I(X)/X^{2}\leqslant 0.8603, where I(X)=X2X(ψ(x)x)2𝑑x.I(X)=\int_{X}^{2X}(\psi(x)-x)^{2}\,dx. This proves (and improves on) a claim by Pintz from 1982. We also show unconditionally that 15 374I(X)/X2\frac{1}{5\,374}\leqslant I(X)/X^{2} for sufficiently large XX, and that the I(X)/X2I(X)/X^{2} has no limit as XX\rightarrow\infty.

1 Introduction

Let ψ(x)=nxΛ(n)\psi(x)=\sum_{n\leqslant x}\Lambda(n) where Λ(n)\Lambda(n) is the von Mangoldt function. By the prime number theorem we have ψ(x)x\psi(x)\sim x. Littlewood (see [7, Thm. 15.11]) showed that ψ(x)x=Ω±(x1/2logloglogx)\psi(x)-x=\Omega_{\pm}(x^{1/2}\log\log\log x) as xx\to\infty. In view of Littlewood’s result, it is of interest that, assuming the Riemann hypothesis (RH), the mean square of (ψ(x)x)/x1/2(\psi(x)-x)/x^{1/2} is bounded. Under RH we have

ψ(x)xx1/2log2x,X2X(ψ(x)x)2𝑑xX2.\psi(x)-x\ll x^{1/2}\log^{2}x,\quad\int_{X}^{2X}(\psi(x)-x)^{2}{\,d}x\ll X^{2}. (1)

Note that using the first bound in (1) does not yield the second bound. Define

I(X):=X2X(ψ(x)x)2𝑑x.I(X):=\int_{X}^{2X}(\psi(x)-x)^{2}{\,d}x. (2)

Unconditionally, it is known that I(X)X2I(X)\gg X^{2}. Indeed Popov and Stechkin [12, Thms. 6–7] showed that

X2X|ψ(x)x|𝑑x>X3/2200,\int_{X}^{2X}|\psi(x)-x|{\,d}x>\frac{X^{3/2}}{200}, (3)

where XX is sufficiently large. On using Cauchy–Schwarz, this shows that I(X)/X21/(40 000)I(X)/X^{2}\geqslant 1/(40\,000).

Pintz wrote a series of papers giving bounds on the constant in (3): [8] has an ineffective constant, [10, Cor. 1] has (22000)1(22000)^{-1} and [9, Cor. 1] has 4001400^{-1}. Under RH, Cramér [3] proved that I(X)cX2I(X)\leqslant cX^{2} for sufficiently large XX. Pintz [10, 9] claims that one may take c=1c=1 for all XX sufficiently large. We are unaware of a proof of this, or of any similar results in the literature.

It follows from the above discussion that there exist positive constants A1A_{1} and A2A_{2} for which A1I(X)X2A2A_{1}\leqslant I(X)X^{-2}\leqslant A_{2}, for sufficiently large XX. Actually the upper bound is conditional on RH whereas the lower bound is unconditional. The purpose of this article is give what we believe to be the best known bounds on A1A_{1} and A2A_{2}.

Theorem 1.

Assume the Riemann hypothesis and let I(X)I(X) be defined in (2)(\ref{turnip}). Then, for XX sufficiently large we have 15 374I(X)X20.8603\frac{1}{5\,374}\leqslant I(X)X^{-2}\leqslant 0.8603.

Presumably, both bounds in Theorem 1 could be improved. We computed I(X)I(X) for XX at every integer [1,1011]\in[1,10^{11}] and include two plots showing its short term behaviour as Figures 1 and 2.

Refer to caption

Figure 1: Plot of I(X)/X2{I(X)}/{X^{2}} vs XX for X[1,100]X\in[1,100]

Refer to caption

Figure 2: Plot of I(X)/X2{I(X)}/{X^{2}} vs XX sampled every 10510^{5}

We are not aware of any conjectured results on the limiting behaviour of I(x)x2I(x)x^{-2}, and so prove the following.

Theorem 2.

With I(X)I(X) defined by (2)(\ref{turnip}), we have that limXI(X)/X2\lim_{X\rightarrow\infty}I(X)/X^{2} does not exist.

If RH is false, then I(X)/X2I(X)/X^{2} is unbounded. Hence, we assume RH except where noted (e.g. RH is not necessary in §2). Let

B:=ρ1,ρ2|22+i(γ1γ2)1ρ1ρ2¯(2+i(γ1γ2))|,B:=\sum_{\rho_{1},\rho_{2}}\left|\frac{2^{2+i(\gamma_{1}-\gamma_{2})}-1}{\rho_{1}{\overline{\rho_{2}}}(2+i(\gamma_{1}-\gamma_{2}))}\right|\,, (4)

where ρj=12+iγj\rho_{j}=\frac{1}{2}+i\gamma_{j} denotes a nontrivial zero of ζ(s)\zeta(s). Following along the lines of [7, Thm. 13.5], one can show that

lim supXI(X)X2B.\limsup_{X\to\infty}\frac{I(X)}{X^{2}}\leqslant B\,.

Corollary 2 shows that B0.8603B\leqslant 0.8603. This proves the upper bound in Theorem 1, which proves Pintz’s claim and provides a significant improvement.

In §2 we give some variations on a well-known lemma of Lehman that is useful for estimating bounds on sums over nontrivial zeros of ζ(s)\zeta(s). We then give several such bounds that are used in the proof of Theorem 3. In §3 we prove Theorem 3, which bounds the tail of the sum in (4), and in Corollary 2 we deduce bounds on BB. In §4 we prove the lower bound in Theorem 1. Finally, in §5 we prove Theorem 2.

Throughout this paper we write ϑ\vartheta to denote a complex number with modulus at most unity. Also, expressions such as T/2πT/2\pi should be interpreted as T/(2π)T/(2\pi), and logkx\log^{k}x as (logx)k(\log x)^{k}. The symbols γ,γ1,γ2\gamma,\gamma_{1},\gamma_{2} denote the ordinates of generic nontrivial zeros β+iγ\beta+i\gamma of ζ(s)\zeta(s). If we wish to refer to the kk-th such γ>0\gamma>0 we denote it by γk^\widehat{\gamma_{k}}. For example, γ1^=14.13472514\widehat{\gamma_{1}}=14.13472514\cdots. Finally, we define L=logTL=\log T and L^=log(T/2π)\widehat{L}=\log(T/2\pi).

2 Preliminary results

The results in this section are unconditional.

We state a well-known result due to Backlund [1], with the constants improved by several authors, most recently by Trudgian [14, Thm. 1, Cor. 1], and Platt and Trudgian [11, Cor. 1].

Lemma 1 (Backlund–Platt–Trudgian).

For all T2πeT\geqslant 2\pi e,

N(T)=T2πlogT2πT2π+78+Q(T),N(T)=\frac{T}{2\pi}\log\frac{T}{2\pi}-\frac{T}{2\pi}+\frac{7}{8}+Q(T),

where

|Q(T)|0.11logT+0.29loglogT+2.29+0.2/T.|Q(T)|\leqslant 0.11\log T+0.29\log\log T+2.29+0.2/T\,.

On RH we have Q(T)=O(logT/loglogT)Q(T)=O(\log T/\log\log T), see [7, Cor. 14.4], but we do not use this result.

Corollary 1.

For all T2πT\geqslant 2\pi,

N(T)=T2πlogT2πT2π+78+(0.28ϑ)logT.N(T)=\frac{T}{2\pi}\log\frac{T}{2\pi}-\frac{T}{2\pi}+\frac{7}{8}+(0.28\vartheta)\log T.
Proof.

By Lemma 1, the result holds for all TT1:=1.03108T\geqslant T_{1}:=1.03\cdot 10^{8}. For T[2π,T1)T\in[2\pi,T_{1}), it has been verified by an interval-arithmetic computation, using the nontrivial zeros β+iγ\beta+i\gamma of ζ(s)\zeta(s) with γ(0,T1)\gamma\in(0,T_{1}). ∎

Let AA be a constant such that

N(T)=T2πlogT2πT2π+78+(ϑA)logTN(T)=\frac{T}{2\pi}\log\frac{T}{2\pi}-\frac{T}{2\pi}+\frac{7}{8}+(\vartheta A)\log T

holds for all T2πT\geqslant 2\pi. By Corollary 1, we can assume that A0.28A\leqslant 0.28.

We state a lemma of Lehman [6, Lem. 1]. We have generalised Lehman’s wording, but the original proof still applies.

Lemma 2 (Lehman-decreasing).

If 2πeT1T22\pi e\leqslant T_{1}\leqslant T_{2} and ϕ:[T1,T2][0,)\phi:[T_{1},T_{2}]\mapsto[0,\infty) is monotone non-increasing on [T1,T2][T_{1},T_{2}], then

T1<γT2ϕ(γ)=12πT1T2ϕ(t)log(t/2π)𝑑t+Aϑ(2ϕ(T1)logT1+T1T2ϕ(t)t𝑑t).\sum_{T_{1}<\gamma\leqslant T_{2}}\!\!\phi(\gamma)=\frac{1}{2\pi}\int_{T_{1}}^{T_{2}}\!\phi(t)\log(t/2\pi){\,d}t\,+\,A\vartheta\left(\!2\phi(T_{1})\log T_{1}+\int_{T_{1}}^{T_{2}}\frac{\phi(t)}{t}{\,d}t\right)\!.

In Lemma 2, we can let T2T_{2}\to\infty if the first integral converges. Lemma 2 does not apply if ϕ(t)\phi(t) is increasing. In this case, Lemma 3 provides an alternative.

Lemma 3 (Lehman-increasing).

If 2πeT1T22\pi e\leqslant T_{1}\leqslant T_{2} and ϕ:[T1,T2][0,)\phi:[T_{1},T_{2}]\mapsto[0,\infty) is monotone non-decreasing on [T1,T2][T_{1},T_{2}], then

T1<γT2ϕ(γ)=12πT1T2ϕ(t)log(t/2π)𝑑t+Aϑ(2ϕ(T2)logT2+T1T2ϕ(t)t𝑑t).\sum_{T_{1}<\gamma\leqslant T_{2}}\!\!\phi(\gamma)=\frac{1}{2\pi}\int_{T_{1}}^{T_{2}}\!\phi(t)\log(t/2\pi){\,d}t+A\vartheta\left(\!2\phi(T_{2})\log T_{2}+\int_{T_{1}}^{T_{2}}\frac{\phi(t)}{t}{\,d}t\right)\!.
Proof.

We follow the proof of [6, Lem. 1] with appropriate modifications. ∎

We need to apply a Lehman-like lemma to a function ϕ(t)\phi(t) which decreases and then increases. Hence we state the following lemma.

Lemma 4 (Lehman-unimodal).

Suppose that 2πeT1T22\pi e\leqslant T_{1}\leqslant T_{2}, and that ϕ:[T1,T2][0,)\phi:[T_{1},T_{2}]\mapsto[0,\infty). If there exists θ[T1,T2]\theta\in[T_{1},T_{2}] such that ϕ\phi is non-increasing on [T1,θ][T_{1},\theta] and non-decreasing on [θ,T2][\theta,T_{2}], then

T1<γT2ϕ(γ)\displaystyle\sum_{T_{1}<\gamma\leqslant T_{2}}\phi(\gamma) =12πT1T2ϕ(t)log(t/2π)𝑑t\displaystyle=\frac{1}{2\pi}\int_{T_{1}}^{T_{2}}\phi(t)\log(t/2\pi){\,d}t
+Aϑ(2ϕ(T1)logT1+2ϕ(T2)logT2+T1T2ϕ(t)t𝑑t).\displaystyle\;\;\;\;+A\vartheta\left(2\phi(T_{1})\log T_{1}+2\phi(T_{2})\log T_{2}+\int_{T_{1}}^{T_{2}}\frac{\phi(t)}{t}{\,d}t\right).
Proof.

Apply Lemma 2 on [T1,θ][T_{1},\theta] and Lemma 3 on [θ,T2][\theta,T_{2}]. ∎

We need some elementary integrals. For k0k\geqslant 0, T1T\geqslant 1 let

Ik:=TTlogktt2𝑑t.I_{k}:=T\int_{T}^{\infty}\frac{\log^{k}t}{t^{2}}{\,d}t.

Then I0=1I_{0}=1 and IkI_{k} satisfies the recurrence Ik=Lk+kIk1I_{k}=L^{k}+kI_{k-1} for k1k\geqslant 1. Thus I1=L+1I_{1}=L+1, I2=L2+2L+2I_{2}=L^{2}+2L+2, I3=L3+3L2+6L+6I_{3}=L^{3}+3L^{2}+6L+6, etc.

We also need

T2Tlogtt3𝑑t=2L+14T^{2}\int_{T}^{\infty}\frac{\log t}{t^{3}}{\,d}t=\frac{2L+1}{4} (5)

and

T2Tlog2tt3𝑑t=2L2+2L+14,T^{2}\int_{T}^{\infty}\frac{\log^{2}t}{t^{3}}{\,d}t=\frac{2L^{2}+2L+1}{4}\,, (6)

which may be found in a similar fashion to I1I_{1} and I2I_{2} respectively.

We now state some lemmas that will be used in §3. Lemmas 58 are applications of Lemma 2.

Lemma 5.

If T2πeT\geqslant 2\pi e, then

γ>T1γ2L2πT.\sum_{\gamma>T}\frac{1}{\gamma^{2}}\leqslant\frac{L}{2\pi T}.
Proof.

We apply Lemma 2 with ϕ(t)=1/t2\phi(t)=1/t^{2}, T1=TT_{1}=T, and let the upper limit T2T_{2}\to\infty. Using the integral I1I_{1} above, this gives

γ>T1γ2\displaystyle\sum_{\gamma>T}\frac{1}{\gamma^{2}} =12πTlog(t/2π)t2𝑑t+Aϑ(2LT2+Tdtt3)\displaystyle=\frac{1}{2\pi}\int_{T}^{\infty}\frac{\log(t/2\pi)}{t^{2}}{\,d}t+A\vartheta\left(\frac{2L}{T^{2}}+\int_{T}^{\infty}\frac{dt}{t^{3}}\right)
=L+1log(2π)2πT+Aϑ(4L+12T2)\displaystyle=\frac{L+1-\log(2\pi)}{2\pi T}+A\vartheta\left(\frac{4L+1}{2T^{2}}\right)
L2πT,\displaystyle\leqslant\frac{L}{2\pi T}\,,

where the final inequality uses T2πeT\geqslant 2\pi e and A0.28A\leqslant 0.28. ∎

Lemma 6.

If T4πeT\geqslant 4\pi e, then

γ>Tlog(γ/2π)γ2L2L2πT.\sum_{\gamma>T}\frac{\log(\gamma/2\pi)}{\gamma^{2}}\leqslant\frac{L^{2}-L}{2\pi T}\,.
Proof.

We apply Lemma 2 with ϕ(t)=log(t/2π)/t2\phi(t)=\log(t/2\pi)/t^{2}, T1=TT_{1}=T, and let the upper limit T2T_{2}\to\infty. Since log(t/2π)/t2\log(t/2\pi)/t^{2} is decreasing on [4πe,)[4\pi e,\infty), Lemma 2 is applicable. Making use of the integrals I2I_{2} and (5) above, we obtain

γ>Tlog(γ/2π)γ2\displaystyle\sum_{\gamma>T}\frac{\log(\gamma/2\pi)}{\gamma^{2}} =12πTlog2(t/2π)t2𝑑t\displaystyle=\frac{1}{2\pi}\int_{T}^{\infty}\frac{\log^{2}(t/2\pi)}{t^{2}}{\,d}t
+Aϑ(2log(T/2π)logTT2+Tlog(t/2π)t3𝑑t)\displaystyle\;\;\;\;+A\vartheta\left(\frac{2\log(T/2\pi)\log T}{T^{2}}+\int_{T}^{\infty}\frac{\log(t/2\pi)}{t^{3}}{\,d}t\right)
=L^2+2L^+22πT+Aϑ(2LL^T2+2L^+14T2)L2L2πT,\displaystyle=\frac{\widehat{L}^{2}+2\widehat{L}+2}{2\pi T}+A\vartheta\left(\frac{2L\widehat{L}}{T^{2}}+\frac{2\widehat{L}+1}{4T^{2}}\right)\leqslant\frac{L^{2}-L}{2\pi T}\,,

where the final inequality uses T4πeT\geqslant 4\pi e and A0.28A\leqslant 0.28. ∎

Lemma 7.

If T100T\geqslant 100, then

γ>Tlog2(γ/2π)γ2L31.39L22πT.\sum_{\gamma>T}\frac{\log^{2}(\gamma/2\pi)}{\gamma^{2}}\leqslant\frac{L^{3}-1.39L^{2}}{2\pi T}.
Proof.

We apply Lemma 2 with ϕ(t)=log2(t/2π)/t2\phi(t)=\log^{2}(t/2\pi)/t^{2}, T1=TT_{1}=T, and T2T_{2}\to\infty. Since ϕ(t)\phi(t) is monotonic decreasing on [100,)[100,\infty), Lemma 2 is applicable. Using the integrals I3I_{3} and (6) above, we obtain

γ>Tlog2(γ/2π)γ2\displaystyle\sum_{\gamma>T}\frac{\log^{2}(\gamma/2\pi)}{\gamma^{2}} =12πTlog3(t/2π)t2𝑑t\displaystyle=\frac{1}{2\pi}\int_{T}^{\infty}\frac{\log^{3}(t/2\pi)}{t^{2}}{\,d}t
+Aϑ(2log2(T/2π)logTT2+Tlog2(t/2π)t3𝑑t)\displaystyle\;\;\;\;+A\vartheta\left(\frac{2\log^{2}(T/2\pi)\log T}{T^{2}}+\int_{T}^{\infty}\frac{\log^{2}(t/2\pi)}{t^{3}}{\,d}t\right)
=L^3+3L^2+6L^+62πT+Aϑ(8LL^2+2L^2+2L^+14T2)\displaystyle=\frac{\widehat{L}^{3}+3\widehat{L}^{2}+6\widehat{L}+6}{2\pi T}+A\vartheta\left(\frac{8L\widehat{L}^{2}+2\widehat{L}^{2}+2\widehat{L}+1}{4T^{2}}\right)
L31.39L22πT,\displaystyle\leqslant\frac{L^{3}-1.39L^{2}}{2\pi T}\,,

where the final inequality uses T100T\geqslant 100 and A0.28A\leqslant 0.28. ∎

The following lemma improves on the upper bound of [4, Lem. 2.10].

Lemma 8.

If T4πeT\geqslant 4\pi e, then

0<γT1γL^24π.\sum_{0<\gamma\leqslant T}\frac{1}{\gamma}\leqslant\frac{\widehat{L}^{2}}{4\pi}\,. (7)
Proof.

Suppose that TT1T\geqslant T_{1}, where T14πeT_{1}\geqslant 4\pi e will be determined later. Using Lemma 2 with ϕ(t)=1/t\phi(t)=1/t, we obtain

T1<γT1γ\displaystyle\sum_{T_{1}<\gamma\leqslant T}\frac{1}{\gamma} =12πT1Tlog(t/2π)t𝑑t+Aϑ(2logT1T1+T1Tdtt2)\displaystyle=\frac{1}{2\pi}\int_{T_{1}}^{T}\frac{\log(t/2\pi)}{t}{\,d}t+A\vartheta\left(\frac{2\log T_{1}}{T_{1}}+\int_{T_{1}}^{T}\frac{dt}{t^{2}}\right)
=14π(L^2log2(T1/2π))+Aϑ(2logT1+1T1).\displaystyle=\frac{1}{4\pi}\left(\widehat{L}^{2}-\log^{2}(T_{1}/2\pi)\right)+A\vartheta\left(\frac{2\log T_{1}+1}{T_{1}}\right)\,. (8)

Thus, including a sum over γT1\gamma\leqslant T_{1}, we have

0<γT1γ\displaystyle\sum_{0<\gamma\leqslant T}\frac{1}{\gamma} L^24π+ε(T1),\displaystyle\leqslant\frac{\widehat{L}^{2}}{4\pi}+\varepsilon(T_{1}),

where

ε(T1)=0<γT11γlog2(T1/2π)4π+A(2logT1+1T1).\displaystyle\varepsilon(T_{1})=\sum_{0<\gamma\leqslant T_{1}}\frac{1}{\gamma}-\frac{\log^{2}(T_{1}/2\pi)}{4\pi}+A\left(\frac{2\log T_{1}+1}{T_{1}}\right)\,.

Using A0.28A\leqslant 0.28, and summing over the first 8080 nontrivial zeros of ζ(s)\zeta(s), shows that ε(202)<0\varepsilon(202)<0. Thus, we take T1=202T_{1}=202, whence (7) holds for TT1=202T\geqslant T_{1}=202. We can verify numerically that (7) also holds for T[4πe,T1)T\in[4\pi e,T_{1}). ∎

Remark 1.

The motivation for our proof of Lemma 8 is as follows. Define

H:=limT(0<γT1γlog2(T/2π)4π).H:=\lim_{T\to\infty}\left(\sum_{0<\gamma\leqslant T}\frac{1}{\gamma}-\frac{\log^{2}(T/2\pi)}{4\pi}\right)\,.

It is easy to show, using (8), that the limit defining HH exists. A computation shows that H0.0171594H\approx-0.0171594. Since HH is negative, we expect that ε(T1)\varepsilon(T_{1}) should be negative for all sufficiently large T1T_{1}. See also [5], and [2, Lem. 3].

3 Bounding the tail in the series for BB

We are now ready to bound the tail of the series (4). Our main result is stated in Theorem 3. Bounds on BB are deduced in Corollary 2.

Theorem 3.

Assume RH. If T100T\geqslant 100, L=logTL=\log T, and BB is defined by (4)\eqref{eq:c6}, then

B|γ1|T,|γ2|T|22+i(γ1γ2)1ρ1ρ2¯(2+i(γ1γ2))|+10L3+11L2π2T.B\leqslant\sum_{|\gamma_{1}|\leqslant T,|\gamma_{2}|\leqslant T}\left|\frac{2^{2+i(\gamma_{1}-\gamma_{2})}-1}{\rho_{1}{\overline{\rho_{2}}}(2+i(\gamma_{1}-\gamma_{2}))}\right|+\frac{10L^{3}+11L^{2}}{\pi^{2}T}\,.
Proof.

Initially, we ignore the numerators |22+i(γ1γ2)1||2^{2+i(\gamma_{1}-\gamma_{2})}-1| in (4), since they are easily bounded. Define

S(T):=|γ1|T,|γ2|T|1ρ1ρ2¯(2+i(γ1γ2))|,S(T):=\sum_{|\gamma_{1}|\leqslant T,|\gamma_{2}|\leqslant T}\left|\frac{1}{\rho_{1}{\overline{\rho_{2}}}(2+i(\gamma_{1}-\gamma_{2}))}\right|\,, (9)

and S:=limTS(T)S_{\infty}:=\lim_{T\to\infty}S(T), with S0.217S_{\infty}\approx 0.217. We refer to E(T):=SS(T)E(T):=S_{\infty}-S(T) as the tail of the series with parameter TT. Thus, the tail is the sum of terms with max(|γ1|,|γ2|)>T\max(|\gamma_{1}|,|\gamma_{2}|)>T. Comparing with (4), and using |22+i(γ1γ2)1|5|2^{2+i(\gamma_{1}-\gamma_{2})}-1|\leqslant 5, we see that the error caused by summing (4) with max(|γ1|,|γ2)T\max(|\gamma_{1}|,|\gamma_{2})\leqslant T is at most 5E(T)5E(T).

We consider bounding sums of the tail terms. By using the symmetry (γ1,γ2)(γ1,γ2)(\gamma_{1},\gamma_{2})\to(-\gamma_{1},-\gamma_{2}), i.e. complex conjugation, we can assume that γ1>0\gamma_{1}>0 (but we must multiply the resulting bound by 22). We can also use the symmetry (γ1,γ2)(γ2,γ1)(\gamma_{1},\gamma_{2})\to(\gamma_{2},\gamma_{1}) if γ2>0\gamma_{2}>0, and (γ1,γ2)(γ2,γ1)(\gamma_{1},\gamma_{2})\to(-\gamma_{2},-\gamma_{1}) if γ2<0\gamma_{2}<0, to reduce to the case that |γ2|γ1|\gamma_{2}|\leqslant\gamma_{1} (again doubling the resulting bound). Terms on the diagonal γ1=γ2\gamma_{1}=\gamma_{2} and anti-diagonal γ1=γ2\gamma_{1}=-\gamma_{2} are given double the necessary weight, but this does not affect the validity of the bound.

For each γ1>0\gamma_{1}>0, possible γ2\gamma_{2} satisfy γ2[γ1,γ1]\gamma_{2}\in[-\gamma_{1},\gamma_{1}]. Since γ2\gamma_{2} is the ordinate of a nontrivial zero of ζ(s)\zeta(s), it is never zero, in fact |γ2|>14|\gamma_{2}|>14.

We now bound the terms 1/|ρ1ρ2¯(2+i(γ1γ2))|1/|\rho_{1}\overline{\rho_{2}}(2+i(\gamma_{1}-\gamma_{2}))| and various sums. Our strategy is to fix γ1\gamma_{1} and sum over all possible γ2\gamma_{2}, then allow γ1\gamma_{1} to vary and sum over all γ1>T\gamma_{1}>T. Since |γ1|<|ρ1||\gamma_{1}|<|\rho_{1}| and |γ2|<|ρ2||\gamma_{2}|<|\rho_{2}|, we actually bound

t(γ1,γ2):=1|γ1γ2(2+i(γ1γ2))|,t(\gamma_{1},\gamma_{2}):=\frac{1}{|\gamma_{1}\gamma_{2}(2+i(\gamma_{1}-\gamma_{2}))|}\,,

which is only slightly larger, since 1|ρj/γj|1+1/8γj21.0011\leqslant|\rho_{j}/\gamma_{j}|\leqslant 1+1/8\gamma_{j}^{2}\leqslant 1.001.

It is useful to define D:=1/t(γ1,γ2)D:=1/t(\gamma_{1},\gamma_{2}). We assume that TT0=100T\geqslant T_{0}=100. Since we eventually sum over γ1>T\gamma_{1}>T, we also assume that γ1T0\gamma_{1}\geqslant T_{0}.

First suppose that γ2\gamma_{2} is positive. In this case, we have 0<γ2γ10<\gamma_{2}\leqslant\gamma_{1} and Dγ1γ2max(2,γ1γ2)D\geqslant\gamma_{1}\gamma_{2}\max(2,\gamma_{1}-\gamma_{2}). Thus the terms t(γ1,γ2)t(\gamma_{1},\gamma_{2}) are bounded by ϕ(γ2)/γ12\phi(\gamma_{2})/\gamma_{1}^{2}, where, writing T=γ1T=\gamma_{1},

ϕ(t):={Tt(Tt)=1t+1Tt if t(0,T2];T/2T2=12+1T2 if t(T2,T].\phi(t):=\begin{cases}\displaystyle\frac{T}{t(T-t)}=\;\frac{1}{t}+\frac{1}{T-t}\;\text{ if }t\in(0,T-2];\\[11.0pt] \displaystyle\;\;\frac{T/2}{T-2}\;\;=\;\frac{1}{2}+\frac{1}{T-2}\text{ if }t\in(T-2,T].\end{cases}

Note that ϕ(t)\phi(t) is positive, decreasing on the interval (0,T/2](0,T/2], increasing on the interval (T/2,T2](T/2,T-2], and constant on the interval [T2,T][T-2,T]. Thus, for summing ϕ(γ2)\phi(\gamma_{2}) over γ2(2πe,T]\gamma_{2}\in(2\pi e,T], Lemma 4 applies with T1=2πeT_{1}=2\pi e, T2=T2T1T_{2}=T\geqslant 2T_{1}, and θ=T/2\theta=T/2.

To apply Lemma 4, we need to bound (1/2π)T1Tϕ(t)log(t/2π)𝑑t(1/2\pi)\int_{T_{1}}^{T}\phi(t)\log(t/2\pi){\,d}t (the main term), and also the error terms AT1T(ϕ(t)/t)𝑑tA\int_{T_{1}}^{T}(\phi(t)/t){\,d}t and 2Aϕ(Tj)log(Tj)2A\phi(T_{j})\log(T_{j}) (j=1,2j=1,2). We consider these in turn.

First consider the main term:

12πT1Tϕ(t)log(t/2π)𝑑t\displaystyle\frac{1}{2\pi}\int_{T_{1}}^{T}\phi(t)\log(t/2\pi){\,d}t
=\displaystyle= 12π(T1T2(1t+1Tt)log(t/2π)𝑑t+ϕ(T)T2Tlog(t/2π)𝑑t)\displaystyle\;\frac{1}{2\pi}\left(\int_{T_{1}}^{T-2}\left(\frac{1}{t}+\frac{1}{T-t}\right)\log(t/2\pi){\,d}t+\phi(T)\int_{T-2}^{T}\log(t/2\pi){\,d}t\right)
\displaystyle\leqslant 12π(T1Tlog(t/2π)t𝑑t+L^0T2dtTt+L^(1+2T2))\displaystyle\;\frac{1}{2\pi}\left(\int_{T_{1}}^{T}\frac{\log(t/2\pi)}{t}{\,d}t+\widehat{L}\int_{0}^{T-2}\frac{dt}{T-t}+\widehat{L}\left(1+\frac{2}{T-2}\right)\right)
\displaystyle\leqslant 14π(L^21+2L^log(T/2)+2L^+4L^T2)\displaystyle\;\frac{1}{4\pi}\left(\widehat{L}^{2}-1+2\widehat{L}\log(T/2)+2\widehat{L}+\frac{4\widehat{L}}{T-2}\right)
\displaystyle\leqslant 14π(3L^2+2L^(2+logπ)0.88).\displaystyle\;\frac{1}{4\pi}\left(3\widehat{L}^{2}+2\widehat{L}(2+\log\pi)-0.88\right)\,.

Now consider the error terms. We have

T1Tϕ(t)t𝑑t\displaystyle\int_{T_{1}}^{T}\frac{\phi(t)}{t}{\,d}t =T1T2ϕ(t)t𝑑t+ϕ(T)T2Tdtt\displaystyle=\int_{T_{1}}^{T-2}\frac{\phi(t)}{t}{\,d}t+\phi(T)\int_{T-2}^{T}\frac{dt}{t}
=T1T2(1t2+1T(1t+1Tt))𝑑t+ϕ(T)T2Tdtt\displaystyle=\int_{T_{1}}^{T-2}\left(\frac{1}{t^{2}}+\frac{1}{T}\left(\frac{1}{t}+\frac{1}{T-t}\right)\right){\,d}t+\phi(T)\int_{T-2}^{T}\frac{dt}{t}
1T11T+log(T/T1)+log(T/2)T+T(T2)20.12.\displaystyle\leqslant\frac{1}{T_{1}}-\frac{1}{T}+\frac{\log(T/T_{1})+\log(T/2)}{T}+\frac{T}{(T-2)^{2}}\leqslant 0.12\,.

Also,

2ϕ(T1)logT1=2logT1T1(TTT1)0.41,2\phi(T_{1})\log T_{1}=\frac{2\log T_{1}}{T_{1}}\left(\frac{T}{T-T_{1}}\right)\leqslant 0.41,

and

2ϕ(T2)logT2(1+2T2)logTL^+log(2π)+2logTT2L^+1.94.2\phi(T_{2})\log T_{2}\leqslant\left(1+\frac{2}{T-2}\right)\log T\leqslant\widehat{L}+\log(2\pi)+\frac{2\log T}{T-2}\leqslant\widehat{L}+1.94\,.

Thus, Lemma 4 gives

T1<γTϕ(γ)\displaystyle\sum_{T_{1}<\gamma\leqslant T}\!\!\phi(\gamma) 3L^2+2L^(2+logπ)0.884π+Aϑ(0.41+L^+1.94+0.12)\displaystyle\leqslant\frac{3\widehat{L}^{2}+2\widehat{L}(2+\log\pi)-0.88}{4\pi}+A\vartheta\left(0.41+\widehat{L}+1.94+0.12\right)
3L^2+9.81L^+7.824π.\displaystyle\leqslant\frac{3\widehat{L}^{2}+9.81\widehat{L}+7.82}{4\pi}\,.

Since γ1^<T1<γ2^\widehat{\gamma_{1}}<T_{1}<\widehat{\gamma_{2}}, we have to treat ϕ(γ1^)\phi(\widehat{\gamma_{1}}) separately. We have

ϕ(γ1^)=Tγ1^(Tγ1^)<0.083,\phi(\widehat{\gamma_{1}})=\frac{T}{\widehat{\gamma_{1}}(T-\widehat{\gamma_{1}})}<0.083\,,

and thus

0γTϕ(γ)3L^2+9.81L^+8.874π.\sum_{0\leqslant\gamma\leqslant T}\phi(\gamma)\leqslant\frac{3\widehat{L}^{2}+9.81\widehat{L}+8.87}{4\pi}\,.

Hence, we have shown that

0<γ2γ1t(γ1,γ2)3log2(γ1/2π)+9.81log(γ1/2π)+8.874πγ12.\sum_{0<\gamma_{2}\leqslant\gamma_{1}}t(\gamma_{1},\gamma_{2})\leqslant\frac{3\log^{2}(\gamma_{1}/2\pi)+9.81\log(\gamma_{1}/2\pi)+8.87}{4\pi\gamma_{1}^{2}}\,. (10)

We now consider the case that γ2\gamma_{2} is negative, whence 0<γ2γ10<-\gamma_{2}\leqslant\gamma_{1}. We could use Lemma 2, but we adopt a simpler approach that gives the same leading term.444This is not surprising, since we use Lemma 8, whose proof depends on Lemma 2.

Assuming that γ2<0\gamma_{2}<0, we have Dγ1|γ2|(γ1+|γ2|)γ12|γ2|D\geqslant\gamma_{1}|\gamma_{2}|(\gamma_{1}+|\gamma_{2}|)\geqslant\gamma_{1}^{2}|\gamma_{2}|, and the terms are bounded by

t(γ1,γ2)1γ12|γ2|.t(\gamma_{1},\gamma_{2})\leqslant\frac{1}{\gamma_{1}^{2}|\gamma_{2}|}\,.

Summing over γ2\gamma_{2} satisfying 0<γ2γ10<-\gamma_{2}\leqslant\gamma_{1}, using Lemma 8, gives the bound

γ1γ2<0t(γ1,γ2)log2(γ1/2π)4πγ12.\sum_{-\gamma_{1}\leqslant\gamma_{2}<0}t(\gamma_{1},\gamma_{2})\leqslant\frac{\log^{2}(\gamma_{1}/2\pi)}{4\pi\gamma_{1}^{2}}\,. (11)

We now combine the results for positive and negative γ2\gamma_{2}. Adding the bounds (10) and (11) gives

γ1γ2γ1t(γ1,γ2)log2(γ1/2π)+2.46log(γ1/2π)+2.22πγ12.\sum_{-\gamma_{1}\leqslant\gamma_{2}\leqslant\gamma_{1}}\!\!t(\gamma_{1},\gamma_{2})\leqslant\frac{\log^{2}(\gamma_{1}/2\pi)+2.46\log(\gamma_{1}/2\pi)+2.22}{\pi\gamma_{1}^{2}}\,. (12)

Finally, we sum (12) over all γ1>T\gamma_{1}>T and use Lemmas 57, giving

γ1>T,|γ2|γ1t(γ1,γ2)\displaystyle\sum_{\gamma_{1}>T,\;|\gamma_{2}|\leqslant\gamma_{1}}\!\!t(\gamma_{1},\gamma_{2}) (L31.39L2)+2.46(L2L)+2.22L2π2T\displaystyle\leqslant\frac{(L^{3}-1.39L^{2})+2.46(L^{2}-L)+2.22L}{2\pi^{2}T}
L3+1.1L22π2T.\displaystyle\leqslant\frac{L^{3}+1.1L^{2}}{2\pi^{2}T}\,. (13)

Allowing a factor of 44 for symmetry, and a factor of 55 to allow for the numerator in (4), the tail bound 5E(T)5E(T) is 2020 times the bound (13), so

5E(T)10L3+11L2π2T,5E(T)\leqslant\frac{10L^{3}+11L^{2}}{\pi^{2}T}\,, (14)

which proves the theorem. ∎

It is possible to avoid the use of Lemma 4 in the proof of Theorem 3, by summing the tail terms in a different order, so that the terms in the inner sums are monotonic decreasing and Lemma 2 applies. However, the resulting integrals are more difficult to bound than those occurring in our proof of Theorem 3. Both methods give the same leading term.

Corollary 2.

With the notation of Theorem 3, 0.8520B0.86030.8520\leqslant B\leqslant 0.8603.

Proof.

The bounds on BB follow from Theorem 3 by taking T=260877T=260877 and evaluating the finite double sum, which requires the first 41054\cdot 10^{5} nontrivial zeros of ζ(s)\zeta(s). The evaluation, using interval arithmetic, shows that the finite sum is in the interval [0.852089,0.852098][0.852089,0.852098], so the lower bound 0.85200.8520 stated in the corollary is correct. The tail bound (14) is 0.008199\leqslant 0.008199, and 0.852098+0.008199=0.8602970.852098+0.008199=0.860297. This proves the stated upper bound. ∎

Remark 2.

Since the proof of Corollary 2 uses T=260877T=260877, but Theorem 3 and Lemma 7 assume only that T100T\geqslant 100, it is natural to ask if the bounds can be improved if we assume that TT is sufficiently large. This is indeed the case. For T80000T\geqslant 80000, the bound (13) can be improved to (L3+0.4L2)/(2π2T)(L^{3}+0.4L^{2})/(2\pi^{2}T), and it follows that the upper bound in Corollary 2 can be improved to B0.8599B\leqslant 0.8599. The coefficient of L2L^{2} in the bound (13) can be replaced by c(T)=43log252logπ+πA+O(1/L)0.06+O(1/L)c(T)=4-3\log 2-\frac{5}{2}\log\pi+\pi A+O(1/L)\leqslant-0.06+O(1/L), and a bound on the O(1/L)O(1/L) term shows that c(T)0c(T)\leqslant 0 for T1042T\geqslant 10^{42}. The coefficient of L3L^{3} is, however, the best that can be attained by our method.

4 Lower bound on I(X)I(X)

Stechkin and Popov [12, Thm. 7] showed that, if RH were false, then lim infXI(X)/X2=\liminf_{X\to\infty}I(X)/X^{2}=\infty. Given this, we may as well assume RH in this section. Stechkin and Popov [12, Thm. 6] showed that we have for XX large enough

X2X|ψ(u)u|𝑑u>X32200,\int\limits_{X}^{2X}\left|\psi(u)-u\right|{\,d}u>\frac{X^{\frac{3}{2}}}{200}, (15)

which by Cauchy–Schwarz leads immediately to I(X)/X2(40 000)1I(X)/X^{2}\geqslant(40\,000)^{-1}. The bound in (15) follows from showing under the same assumptions that

H(X):=Xlog22X+log22|n0exp(iγnt)ρn|𝑑t>X32200,H(X):=\int\limits_{X-\frac{\log 2}{2}}^{X+\frac{\log 2}{2}}\left|\sum\limits_{n\neq 0}\frac{\exp(i\gamma_{n}t)}{\rho_{n}}\right|{\,d}t>\frac{X^{\frac{3}{2}}}{200}, (16)

where, throughout this section only, for k1k\geqslant 1 we define γk\gamma_{k} (resp. γk\gamma_{-k}) to be the ordinate of the kkth non-trivial zero of ζ(s)\zeta(s), above (resp. below) the real axis. We interpret the sum in (16), which is not absolutely convergent, as

limNn=1N(exp(iγnt)ρn+exp(iγnt)ρn).\lim_{N\to\infty}\sum_{n=1}^{N}\left(\frac{\exp(i\gamma_{n}t)}{\rho_{n}}+\frac{\exp(i\gamma_{-n}t)}{\rho_{-n}}\right).

The key result we need is the following.

Lemma 9.

Let g(z)g(z) be such that g(0)=1g(0)=1 and

δ=1ρ1n2|g(γnγ1)ρn|n1|g(γnγ1)ρn|\delta=\frac{1}{\rho_{1}}-\sum\limits_{n\geqslant 2}\left|\frac{g(\gamma_{n}-\gamma_{1})}{\rho_{n}}\right|-\sum\limits_{n\geqslant 1}\left|\frac{g(-\gamma_{n}-\gamma_{1})}{\rho_{n}}\right|

exists and is finite. Additionally, assume that

g^(y)=12πg(z)exp(izy)𝑑z\widehat{g}(y)=\frac{1}{2\pi}\int\limits_{\mathbb{R}}g(z)\exp(-izy){\,d}z

exists and is supported on [12log2,12log2][-\frac{1}{2}\log 2,\frac{1}{2}\log 2]. Then we have

|H(X)|δmaxyg^(y).|H(X)|\geqslant\frac{\delta}{\max\limits_{y\in\mathbb{R}}\widehat{g}(y)}.
Proof.

This follows from displays (15.4) to (17.4) of [12, Sec. 4]. ∎

Lemma 10.

Let α=log26\alpha=\frac{\log 2}{6} and λ>0\lambda>0. Define

g(z)=(sin(αz)αz)3(1zλ)g(z)=\left(\frac{\sin(\alpha z)}{\alpha z}\right)^{3}\left(1-\frac{z}{\lambda}\right)

and

g^(y)=12πg(z)exp(izy)𝑑z.\widehat{g}(y)=\frac{1}{2\pi}\int\limits_{\mathbb{R}}g(z)\exp(-izy){\,d}z.

Then g(0)=1g(0)=1 and g^(y)\widehat{g}(y) is supported on [12log2,12log2][-\frac{1}{2}\log 2,\frac{1}{2}\log 2]. Furthermore, for real yy, |g^(y)||\widehat{g}(y)| attains its maximum of 94log2\frac{9}{4\log 2}at y=0y=0.

We note that Stechkin and Popov used the fourth power of the sinc function in place of our cube. Almost certainly better choices of the function g(z)g(z) are possible: we leave this to future researchers, in the hope that they can thereby improve the lower bound in Theorem 1.

Lemma 11.

Let gg be as defined in Lemma 10. For T>max(γ1+λ,2πe)T>\max(\gamma_{1}+\lambda,2\pi{\textrm{e}}) not the ordinate of a zero of ζ\zeta set

δT,λ=γ>T|g(γγ1)|+|g(γγ1)|ρ.\delta_{T,\lambda}=\sum\limits_{\gamma>T}\frac{|g(\gamma-\gamma_{1})|+|g(-\gamma-\gamma_{1})|}{\rho}.

Then

δT,λThλ(t)logt2πdt+0.56hλ(T)logT+0.28Thλ(t)t𝑑t\delta_{T,\lambda}\leqslant\int\limits_{T}^{\infty}h_{\lambda}(t)\log\frac{t}{2\pi}{\,d}t+0.56h_{\lambda}(T)\log T+0.28\int\limits_{T}^{\infty}\frac{h_{\lambda}(t)}{t}{\,d}t

where

hλ(t)=tλγ1t(α(tγ1))3+t+λ+γ1t(α(t+γ1))3.h_{\lambda}(t)=\frac{t-{\lambda}-\gamma_{1}}{t(\alpha(t-\gamma_{1}))^{3}}+\frac{t+\lambda+\gamma_{1}}{t(\alpha(t+\gamma_{1}))^{3}}.
Proof.

This is a straightforward application of Corollary 1 and Lemma 2. ∎

Corollary 3.

Let δT,λ\delta_{T,\lambda} be as in Lemma 11, with T=446 000T=446\,000 and λ=10.876\lambda=10.876. Then

δT,λ3.5109.\delta_{T,\lambda}\leqslant 3.5\cdot 10^{-9}.

We can now compute the contribution to δ\delta from the 721 913721\,913 nontrivial zeros with imaginary part less than 446 000446\,000, using λ=10.876\lambda=10.876. We find

1|ρ1|n=2721 913g(γnγ1)ρnn=1721 913g(γnγ1)ρn4.428 225 55102,\frac{1}{|\rho_{1}|}-\sum\limits_{n=2}^{721\,913}\frac{g(\gamma_{n}-\gamma_{1})}{\rho_{n}}-\sum\limits_{n=1}^{721\,913}\frac{g(-\gamma_{n}-\gamma_{1})}{\rho_{n}}\geqslant 4.428\,225\,55\cdot 10^{-2},

so we have δ0.044 282 252.\delta\geqslant 0.044\,282\,252.

Appealing to Lemmas 9 and 10 we can now claim

|H(X)|0.044 282 2524log290.013 641 83,|H(X)|\geqslant 0.044\,282\,252\frac{4\log 2}{9}\geqslant 0.013\,641\,83,

and the lower bound of Theorem 1 results.

5 Non-convergence of I(X)/X2I(X)/X^{2}

Our aim now is to show that I(X)/X2I(X)/X^{2} does not tend to a limit as XX\to\infty. It is more convenient to work with

J(X):=0X(ψ(x)x)2𝑑x,J(X):=\int_{0}^{X}(\psi(x)-x)^{2}{\,d}x, (17)

and deduce results for I(X)I(X). In Theorems 4 and 5 we show that there exist effectively computable constants c1c_{1} and c2c_{2}, satisfying c1<c2c_{1}<c_{2}, such that

lim supX2X2J(X)c2,lim infX2X2J(X)c1.\limsup_{X\to\infty}\frac{2}{X^{2}}\,J(X)\geqslant c_{2},\quad\liminf_{X\to\infty}\frac{2}{X^{2}}\,J(X)\leqslant c_{1}.

Hence J(X)/X2J(X)/X^{2} cannot tend to a limit as XX\to\infty. In Theorem 2 we deduce that I(X)/X2I(X)/X^{2} cannot tend to a limit XX\to\infty.

5.1 Some constants

In sums over zeros, each zero ρ\rho is counted according to its multiplicity mρm_{\rho}. More precisely, a term involving ρ\rho is given a weight mρm_{\rho}. In double sums, a term involving ρ1\rho_{1} and ρ2\rho_{2} is given a weight mρ1mρ2m_{\rho_{1}}m_{\rho_{2}}.

We now define three real constants that are needed later. First, a constant that appears in [7, Thm. 13.6 and Ex. 13.1.1.3] and our Theorem 5:

c1:=ρmρ|ρ|20.046.c_{1}:=\sum_{\rho}\frac{m_{\rho}}{|\rho|^{2}}\approx 0.046. (18)

Second, we define a constant that occurs in Theorem 4:

c2:=ρ1,ρ22ρ1ρ2¯(1+ρ1+ρ2¯)0.104.c_{2}:=\sum_{\rho_{1},\rho_{2}}\frac{2}{\rho_{1}\overline{\rho_{2}}(1+\rho_{1}+\overline{\rho_{2}})}\approx 0.104\,. (19)

Observe that, assuming RH, the “diagonal terms” (i.e. those with ρ1=ρ2\rho_{1}=\rho_{2}) in (19) sum to c1c_{1}.

Third, a constant that will be used in §5.3:

c3:=γ>01γ20.023 105,c_{3}:=\sum_{\gamma>0}\frac{1}{\gamma^{2}}\leqslant 0.023\;105, (20)

where this estimate has been computed to high accuracy previously (see, e.g. [4]). We can replicate this result by summing numerically over zeros below 3.721461083.72146\cdot 10^{8} and using Lemma 5 for the tail.

5.2 The limsup result

We use the explicit formula for ψ(x)\psi(x) (see, e.g., [7, Thm. 12.5]) in the form

ψ(x)x=|γ|Txρρ+O(xlog2xT)\psi(x)-x=-\sum_{|\gamma|\leqslant T}\frac{x^{\rho}}{\rho}+O\left(\frac{x\log^{2}x}{T}\right)

for TT0T\geqslant T_{0}, xX0x\geqslant X_{0}, and xTx\geqslant T.

Theorem 4.

With J(X)J(X) as in (17) and c2c_{2} as in (19),

lim supX2J(X)X2c2.\limsup_{X\to\infty}\frac{2J(X)}{X^{2}}\geqslant c_{2}.
Proof.

Fix some small ε>0\varepsilon>0. We can assume RH, since otherwise J(X)/X2J(X)/X^{2} is unbounded. Proceeding as in the proof of  [7, Thm. 13.5], but with the integral over [T,X][T,X] instead of [X,2X][X,2X], and using the Cauchy–Schwartz inequality for the error term, we obtain

TX(ψ(x)x)2𝑑x=TX|γ1|T,|γ2|Tx1+i(γ1γ2)ρ1ρ2¯dx+O(X5/2log2XT),\int_{T}^{X}(\psi(x)-x)^{2}{\,d}x=\int_{T}^{X}\!\sum_{|\gamma_{1}|\leqslant T,\,|\gamma_{2}|\leqslant T}\frac{x^{1+i(\gamma_{1}-\gamma_{2})}}{\rho_{1}\overline{\rho_{2}}}{\,d}x+O\left(\frac{X^{5/2}\log^{2}X}{T}\right),

provided XTmax(T0,X0)X\geqslant T\geqslant\max(T_{0},X_{0}). We also have, from [7, Thm. 13.5],

0T(ψ(x)x)2𝑑xT2.\int_{0}^{T}(\psi(x)-x)^{2}{\,d}x\ll T^{2}.

Thus

0X(ψ(x)x)2𝑑x=TX\displaystyle\int_{0}^{X}(\psi(x)-x)^{2}{\,d}x=\int_{T}^{X}\! |γ1|T,|γ2|Tx1+i(γ1γ2)ρ1ρ2¯dx\displaystyle\sum_{|\gamma_{1}|\leqslant T,\,|\gamma_{2}|\leqslant T}\frac{x^{1+i(\gamma_{1}-\gamma_{2})}}{\rho_{1}\overline{\rho_{2}}}{\,d}x
+O(T2+X5/2(logX)2/T).\displaystyle+O\left(T^{2}+X^{5/2}(\log X)^{2}/T\right).

Now, from [7, (13.16)], ρ1,ρ2|1ρ1ρ2¯(2+i(γ1γ2)|1\displaystyle\;\sum_{\rho_{1},\rho_{2}}\left|\frac{1}{\rho_{1}\overline{\rho_{2}}(2+i(\gamma_{1}-\gamma_{2})}\right|\ll 1.
Thus, if we exchange the order of integration and summation (valid since the sum is finite), and normalise by X2X^{2}, we obtain

J(X)X2=|γ1|T,|γ2|TXi(γ1γ2)ρ1ρ2¯(2+i(γ1γ2))+O(T2X2+X1/2log2XT).\frac{J(X)}{X^{2}}=\!\sum_{|\gamma_{1}|\leqslant T,\,|\gamma_{2}|\leqslant T}\frac{X^{i(\gamma_{1}-\gamma_{2})}}{\rho_{1}\overline{\rho_{2}}(2+i(\gamma_{1}-\gamma_{2}))}+O\left(\frac{T^{2}}{X^{2}}+\frac{X^{1/2}\log^{2}X}{T}\right).

Choosing T=X5/6T=X^{5/6}, and assuming that XX06/5X\geqslant X_{0}^{6/5} so TX0T\geqslant X_{0}, the error term becomes O(X1/3(logX)2)O(X^{-1/3}(\log X)^{2}). Now, choosing Xlog6(1/ε)/ε3X\geqslant\log^{6}(1/\varepsilon)/\varepsilon^{3}, the error term is O(ε)O(\varepsilon). To summarise, we obtain error O(ε)O(\varepsilon) provided that T=X5/6T=X^{5/6} and XX1X\geqslant X_{1}, where X1=max(X06/5,T06/5,log6(1/ε)/ε3)X_{1}=\max(X_{0}^{6/5},T_{0}^{6/5},\log^{6}(1/\varepsilon)/\varepsilon^{3}).

We shall need another parameter Y=log3(1/ε)/εY=\log^{3}(1/\varepsilon)/\varepsilon. Note that, by the conditions on TT and XX, we necessarily have YTY\leqslant T for ε(0,1/e)\varepsilon\in(0,1/e), since T=X5/6log(1/ε)5/ε5/2log3(1/ε)/ε=YT=X^{5/6}\geqslant\log(1/\varepsilon)^{5}/\varepsilon^{5/2}\geqslant\log^{3}(1/\varepsilon)/\varepsilon=Y.

It remains to consider the main sum over pairs (1/2+iγ1,1/2iγ2)(1/2+i\gamma_{1},1/2-i\gamma_{2}) of zeros with |γ1|,|γ2|T|\gamma_{1}|,|\gamma_{2}|\leqslant T. Observe that the sum is real, as we can see by grouping the term for (1/2+iγ1,1/2iγ2)(1/2+i\gamma_{1},1/2-i\gamma_{2}) with the conjugate term for (1/2iγ1,1/2+iγ2)(1/2-i\gamma_{1},1/2+i\gamma_{2}). Using Dirichlet’s theorem [13, §8.2], we can find some tlogX1t\geqslant\log X_{1}, such that |{tγ/(2π)}|ε|\{t\gamma/(2\pi)\}|\leqslant\varepsilon for all zeros 1/2+iγ1/2+i\gamma with 0<γY0<\gamma\leqslant Y, where YTY\leqslant T is as above.555Here {x}\{x\} denotes the fractional part of xx. Set X=exp(t)X=\exp(t). Then, for all the (γ1,γ2)(\gamma_{1},\gamma_{2}) occurring in the main sum with max(|γ1|,|γ2|)Y\max(|\gamma_{1}|,|\gamma_{2}|)\leqslant Y, we have Xi(γ1γ2)=1+O(ε)X^{i(\gamma_{1}-\gamma_{2})}=1+O(\varepsilon). Hence, for this choice of XX, we have

J(X)X2=|γ1|Y,|γ2|Y1ρ1ρ2¯(2+i(γ1γ2))+R(Y)+O(ε),\frac{J(X)}{X^{2}}=\!\sum_{|\gamma_{1}|\leqslant Y,\,|\gamma_{2}|\leqslant Y}\frac{1}{\rho_{1}\overline{\rho_{2}}(2+i(\gamma_{1}-\gamma_{2}))}+R(Y)+O(\varepsilon),

where

|R(Y)|max(|γ1|,|γ2|)>Y|1ρ1ρ2¯(2+i(γ1γ2))|log3YY|R(Y)|\leqslant\sum_{\max(|\gamma_{1}|,|\gamma_{2}|)>Y}\left|\frac{1}{\rho_{1}\overline{\rho_{2}}(2+i(\gamma_{1}-\gamma_{2}))}\right|\ll\frac{\log^{3}Y}{Y}

is the tail of an absolutely convergent double sum, see (9) and [7, p. 424]. Thus, with our choice Y=log3(1/ε)/εY=\log^{3}(1/\varepsilon)/\varepsilon, we have R(Y)=O(ε)R(Y)=O(\varepsilon).

Recalling the definition of the constant c2c_{2} in (19), we have shown that, for any sufficiently small ε>0\varepsilon>0, there exists X=X(ε)X=X(\varepsilon) such that

2J(X)X2c2O(ε).\frac{2J(X)}{X^{2}}\geqslant c_{2}-O(\varepsilon). (21)

Since ε\varepsilon can be arbitrarily small, this proves the result. ∎

Remark 3.

The least XX satisfying (21) may be bounded using [13, (8.2.1)]. The result is doubly exponential in 1/ε1/\varepsilon. More precisely,

X(ε)exp(exp((1/ε)1+o(1))) as ε0.X(\varepsilon)\leqslant\exp(\exp((1/\varepsilon)^{1+o(1)}))\text{ as $\varepsilon\to 0$}.

5.3 A lower bound on c2c_{2}

The constants c1c_{1} and c2c_{2} are of little interest, so far as the theory of ψ(x)\psi(x) goes, if RH is false. Hence, we assume RH. In Corollary 5 we show that c1<c2c_{1}<c_{2}. Although computations of c2c_{2} suggest this, they do not provide a proof unless they come with a (possibly one-sided) error bound. Here we show how rigorous lower bounds on c2c_{2} can be computed. This provides a way of proving rigorously, without extensive computation, that c1<c2c_{1}<c_{2}.

First we extract the real part of the expression (19). This leads to sharper bounds on the terms than if we included the imaginary parts, which must ultimately cancel.

Lemma 12.

Assume RH. If c2c_{2} is defined by (19), then

c2=γ1>0,γ2T(γ1,γ2),c_{2}=\sum_{\gamma_{1}>0,\;\gamma_{2}}T(\gamma_{1},\gamma_{2}),

where

T(γ1,γ2)=2(1+6γ1γ2γ12γ22)(14+γ12)(14+γ22)(4+(γ1γ2)2).T(\gamma_{1},\gamma_{2})=\frac{2(1+6\gamma_{1}\gamma_{2}-\gamma_{1}^{2}-\gamma_{2}^{2})}{(\frac{1}{4}+\gamma_{1}^{2})(\frac{1}{4}+\gamma_{2}^{2})(4+(\gamma_{1}-\gamma_{2})^{2})}\,. (22)
Proof.

We expand (19), using ρj=12+iγj\rho_{j}=\frac{1}{2}+i\gamma_{j} (this is where RH is required), omit the imaginary parts since the final result is real, and use symmetry to reduce to the case γ1>0\gamma_{1}>0 (so in the resulting sum, γ1\gamma_{1} is positive but γ2\gamma_{2} may have either sign). ∎

Lemma 13 gives a region in which the terms occurring in (22) are positive.

Lemma 13.

If T(γ1,γ2)T(\gamma_{1},\gamma_{2}) is as in (22), and γ2/γ1[38,3+8]\gamma_{2}/\gamma_{1}\in[3-\sqrt{8},3+\sqrt{8}], then T(γ1,γ2)>0T(\gamma_{1},\gamma_{2})>0.

Proof.

Since the denominator of T(γ1,γ2)T(\gamma_{1},\gamma_{2}) is positive, it is sufficient to consider the numerator, which we write as 2P(γ1,γ2)2P(\gamma_{1},\gamma_{2}), where

P(x,y)=1+6xyx2y2.P(x,y)=1+6xy-x^{2}-y^{2}.

Let r=y/xr=y/x, so P(x,y)=1(r26r+1)x2P(x,y)=1-(r^{2}-6r+1)x^{2}. Now r26r+1=(r3)28r^{2}-6r+1=(r-3)^{2}-8 vanishes at r=3±8r=3\pm\sqrt{8}, and is negative iff r(38,3+8)r\in(3-\sqrt{8},3+\sqrt{8}). Thus P(x,y)P(x,y) is positive for r[38,3+8]r\in[3-\sqrt{8},3+\sqrt{8}]. Taking x=γ1,y=γ2x=\gamma_{1},y=\gamma_{2} proves the lemma. ∎

Define

S(Y)=   0<γ1YYγ2YT(γ1,γ2).S(Y)=\sum_{{\;\;\;0<\gamma_{1}\leqslant Y}\atop{-Y\leqslant\gamma_{2}\leqslant Y}}T(\gamma_{1},\gamma_{2}).

Then c2=limYS(Y)c_{2}=\lim_{Y\to\infty}S(Y). Clearly S(Y)S(Y) is constant between ordinates of nontrivial zeros of ζ(s)\zeta(s), and has jumps

J(γ)=limε0(S(γ+ε)S(γε))J(\gamma)=\lim_{\varepsilon\to 0}(S(\gamma+\varepsilon)-S(\gamma-\varepsilon))

at positive ordinates γ\gamma of zeros of ζ(s)\zeta(s). We shall show that all these jumps are positive, so S(Y)S(Y) is monotonic non-decreasing, and c2>S(Y)c_{2}>S(Y) for all Y>0Y>0. This allows us to prove that c2>c1c_{2}>c_{1} by computing S(Y)S(Y) for sufficiently large YY (see Corollary 5).

If γ>0\gamma>0 is the ordinate of a simple zero666For simplicity we assume here that all zeros of ζ(s)\zeta(s) are simple, but one can modify the proofs in an obvious way to account for multiple zeros, if they exist. of ζ(s)\zeta(s), then

J(γ)\displaystyle J(\gamma) =0<γ1γT(γ1,γ)+0<γ1γT(γ1,γ)+γ<γ2<γT(γ,γ2)\displaystyle=\sum_{0<\gamma_{1}\leqslant\gamma}T(\gamma_{1},\gamma)+\!\!\sum_{0<\gamma_{1}\leqslant\gamma}T(\gamma_{1},-\gamma)+\!\!\!\sum_{-\gamma<\gamma_{2}<\gamma}T(\gamma,\gamma_{2})
=T(γ,γ)+T(γ,γ)+2γ<γ2<γT(γ,γ2).\displaystyle=T(\gamma,\gamma)+T(\gamma,-\gamma)+2\!\!\!\!\sum_{-\gamma<\gamma_{2}<\gamma}T(\gamma,\gamma_{2})\,. (23)

This may be seen by drawing a rectangle with vertices at (0,γ)(0,\gamma), (γ,γ)(\gamma,\gamma), (γ,γ)(\gamma,-\gamma), (0,γ)(0,-\gamma), following the north, east and south edges, and using the symmetry T(x,y)=T(y,x)T(x,y)=T(y,x).

To show that J(γ)>0J(\gamma)>0, we split the last sum in (23) into three pieces, A:=(γ,0]A:=(-\gamma,0], B:=(0,(38)γ)B:=(0,(3-\sqrt{8})\gamma), and C:=[(38)γ,γ)C:=[(3-\sqrt{8})\gamma,\gamma). This gives

J(γ)=\displaystyle J(\gamma)= T(γ,γ)+T(γ,γ)\displaystyle\;\;T(\gamma,\gamma)+T(\gamma,-\gamma)
+2γ2AT(γ,γ2)+2γ2BT(γ,γ2)+2γ2CT(γ,γ2).\displaystyle+2\sum_{\gamma_{2}\in A}T(\gamma,\gamma_{2})+2\sum_{\gamma_{2}\in B}T(\gamma,\gamma_{2})+2\sum_{\gamma_{2}\in C}T(\gamma,\gamma_{2}).

By Lemma 13, the sum with γ2C\gamma_{2}\in C consists only of positive terms, so

J(γ)T(γ,γ)+T(γ,γ)+2γ2AT(γ,γ2)+2γ2BT(γ,γ2).J(\gamma)\geqslant T(\gamma,\gamma)+T(\gamma,-\gamma)+2\sum_{\gamma_{2}\in A}T(\gamma,\gamma_{2})+2\sum_{\gamma_{2}\in B}T(\gamma,\gamma_{2}). (24)

We now show that the diagonal term T(γ,γ)T(\gamma,\gamma) in (24) is positive, and sufficiently large to dominate the anti-diagonal term T(γ,γ)T(\gamma,-\gamma) and the sums over AA and BB.

Lemma 14 (diagonal term).

We have T(γ,γ)1.99/γ2.T(\gamma,\gamma)\geqslant 1.99/\gamma^{2}\,.

Proof.

Since γ>0\gamma>0 is the ordinate of a nontrivial zero of ζ(s)\zeta(s), we have γ>14\gamma>14. Thus, using (22), we have T(γ,γ)=2/(14+γ2)>1.99/γ2.T(\gamma,\gamma)={2}/{(\frac{1}{4}+\gamma^{2})}>{1.99}/{\gamma^{2}}.

Lemma 15 (anti-diagonal term and interval AA).

If c3c_{3} is as in (20), then

|T(γ,γ)|2+γ<γ2<0|T(γ,γ2)|16c3γ2<0.37γ2.\frac{|T(\gamma,-\gamma)|}{2}+\sum_{-\gamma<\gamma_{2}<0}|T(\gamma,\gamma_{2})|\leqslant\frac{16c_{3}}{\gamma^{2}}<\frac{0.37}{\gamma^{2}}\,.
Proof.

Write (22) as T(γ,γ2)=N/DT(\gamma,\gamma_{2})=N/D, where the numerator is

N=2(1+6γγ2γ2γ22),N=2(1+6\gamma\gamma_{2}-\gamma^{2}-\gamma_{2}^{2}), (25)

and the denominator is

D=(14+γ2)(14+γ22)(4+(γγ2)2)>γ2γ22(γγ2)2.D=(\textstyle\frac{1}{4}+\gamma^{2})(\frac{1}{4}+\gamma_{2}^{2})(4+(\gamma-\gamma_{2})^{2})>\gamma^{2}\gamma_{2}^{2}(\gamma-\gamma_{2})^{2}. (26)

Thus, N/2=1(r26r+1)γ2N/2=1-(r^{2}-6r+1)\gamma^{2}, where r=γ2/γr=\gamma_{2}/\gamma. Now r26r+1[1,8]r^{2}-6r+1\in[1,8] for r[1,0]r\in[-1,0]. Thus N/2[18γ2,1γ2]N/2\in[1-8\gamma^{2},1-\gamma^{2}], and |N|<16γ2|N|<16\gamma^{2}.

For the denominator, we have D>γ4γ22(1r)2[γ4γ22,4γ4γ22]D>\gamma^{4}\gamma_{2}^{2}(1-r)^{2}\in[\gamma^{4}\gamma_{2}^{2},4\gamma^{4}\gamma_{2}^{2}], so D>γ4γ22D>\gamma^{4}\gamma_{2}^{2}. Combining the inequalities for NN and DD gives

|T(γ,γ2)|<16γ2γ22.|T(\gamma,\gamma_{2})|<\frac{16}{\gamma^{2}\gamma_{2}^{2}}\,.

Now, summing over γ2<0\gamma_{2}<0, and recalling the definition of c3c_{3} in (20), gives the result. ∎

Lemma 16 (interval BB).

We have

0<γ2<(38)γ|T(γ,γ2)|(3+8)c32γ2<0.068γ2.\sum_{0<\gamma_{2}<(3-\sqrt{8})\gamma}|T(\gamma,\gamma_{2})|\leqslant\frac{(3+\sqrt{8})c_{3}}{2\gamma^{2}}<\frac{0.068}{\gamma^{2}}\,.
Proof.

As in the proof of Lemma 15, write (22) as T(γ,γ2)=N/DT(\gamma,\gamma_{2})=N/D, where NN and DD are as in (25)–(26). Now γ2/γ<38\gamma_{2}/\gamma<3-\sqrt{8}, so 1γ2/γ>821-\gamma_{2}/\gamma>\sqrt{8}-2, and (γγ2)2>4(38)γ2(\gamma-\gamma_{2})^{2}>4(3-\sqrt{8})\gamma^{2}. This gives

D>4(38)γ4γ22.D>4(3-\sqrt{8})\gamma^{4}\gamma_{2}^{2}.

Also, N/2=1(r26r+1)γ2N/2=1-(r^{2}-6r+1)\gamma^{2}, where r=γ2/γ[0,38]r=\gamma_{2}/\gamma\in[0,3-\sqrt{8}]. Thus 0r26r+110\leqslant r^{2}-6r+1\leqslant 1 and |N|2γ2|N|\leqslant 2\gamma^{2}. The inequalities for DD and NN give

|T(γ,γ2)|<2γ24(38)γ4γ22=3+82γ2γ22.|T(\gamma,\gamma_{2})|<\frac{2\gamma^{2}}{4(3-\sqrt{8})\gamma^{4}\gamma_{2}^{2}}=\frac{3+\sqrt{8}}{2\gamma^{2}\gamma_{2}^{2}}\,.

Now, summing over γ2>0\gamma_{2}>0 gives the result. ∎

Lemma 17.

S(Y)S(Y) is monotonic non-decreasing for Y[0,)Y\in[0,\infty), with jumps of at least 1.11/γ21.11/\gamma^{2} at ordinates γ>0\gamma>0 of ζ(s)\zeta(s).

Proof.

Using the inequality (24) and Lemmas 1416, we have

J(γ)1.9920.3720.068γ2>1.11γ2.J(\gamma)\geqslant\frac{1.99-2\cdot 0.37-2\cdot 0.068}{\gamma^{2}}>\frac{1.11}{\gamma^{2}}\,.

Thus, S(Y)S(Y) has positive jumps at ordinates γ>0\gamma>0 of zeros of ζ(s)\zeta(s), and is constant between these ordinates. ∎

Corollary 4.

Assume RH. For all Y>0Y>0, we have c2>S(Y)c_{2}>S(Y).

Proof.

This follows as S(Y)S(Y) is monotonic non-decreasing with limit c2c_{2}, and has positive jumps at arbitrarily large YY. ∎

Corollary 5.

Assume RH. Then c1<c2c_{1}<c_{2}.

Proof.

Take Y=70Y=70 in Corollary 4. Computing S(70)S(70), which involves a double sum over first 1717 nontrivial zeros in the upper half-plane, gives a lower bound c2>S(70)>0.0466c_{2}>S(70)>0.0466. Since c1<0.0462c_{1}<0.0462, the result follows. ∎

Remark 4.

RH is probably not necessary for Corollary 5. Any exceptional zeros off the critical line must have large height, and consequently they would make little difference to the numerical values of c1c_{1} and c2c_{2}.

Remark 5.

Taking Y=74 920.83Y=74\,920.83 in Corollary 4, and using the first 10510^{5} zeros of ζ(s)\zeta(s), we obtain

c2>S(Y)>0.104004 and c2c1>0.0578.c_{2}>S(Y)>0.104004\text{ and }c_{2}-c_{1}>0.0578\,.

This is much stronger than the bound used in the proof of Corollary 5, though at the expense of more computation. Our best estimate, using an integral approximation for the higher zeros, is c20.10446c_{2}\approx 0.10446 .

5.4 Non-existence of a limit

First we prove a result analogous to Theorem 4, but with lim sup\limsup replaced by lim inf\liminf. Then we deduce that neither I(X)/X2I(X)/X^{2} nor J(X)/X2J(X)/X^{2} has a limit as XX\to\infty.

Theorem 5.

Assume RH. With J(X)J(X) as in (17) and c1c_{1} as in (18),

lim infX2J(X)X2c1.\liminf_{X\to\infty}\frac{2J(X)}{X^{2}}\leqslant c_{1}.
Proof.

Define

F(X):=\displaystyle F(X):= 1X(ψ(x)x)2𝑑x=J(X)J(1), and\displaystyle\int_{1}^{X}(\psi(x)-x)^{2}{\,d}x=J(X)-J(1),\text{ and}
G(X):=\displaystyle G(X):= 1X(ψ(x)x)2dxx2c1logX.\displaystyle\int_{1}^{X}(\psi(x)-x)^{2}\,\frac{dx}{x^{2}}\sim c_{1}\log X.

Here the asymptotic result is given in [7, Ex. 13.1.1.3], which follows from [7, Thm. 13.6] after a change of variables x=exp(u)x=\exp(u). Using integration by parts, we obtain

G(X)=F(X)X2+21XF(x)dxx3.G(X)=\frac{F(X)}{X^{2}}+2\int_{1}^{X}F(x)\,\frac{dx}{x^{3}}\,.

Now F(X)X2F(X)\ll X^{2}, so

21XF(x)dxx3G(X)c1logX as X.2\int_{1}^{X}F(x)\,\frac{dx}{x^{3}}\sim G(X)\sim c_{1}\log X\text{ as $X\to\infty$}.

Dividing by 2logX2\log X gives

1XF(x)x2dxx/1Xdxxc12 as X.\int_{1}^{X}\frac{F(x)}{x^{2}}\,\frac{dx}{x}\left/\int_{1}^{X}\frac{dx}{x}\right.\sim\frac{c_{1}}{2}\text{ as $X\to\infty$}. (27)

Now, if F(x)/x2c1/2+εF(x)/x^{2}\geqslant c_{1}/2+\varepsilon for some positive ε\varepsilon and all sufficiently large xx, we get a contradiction to (27). Thus, letting ε0\varepsilon\to 0, we obtain the result. ∎

Corollary 6.

With J(X)J(X) as in (17), limXJ(X)X2\displaystyle\lim_{X\to\infty}\frac{J(X)}{X^{2}} does not exist.

Proof.

The result holds if RH is false. Hence, assume RH. From Corollary 5, c1<c2c_{1}<c_{2}, so the result is implied by Theorems 4 and 5. ∎

We conclude by showing the non-existence of limXI(X)X2\lim_{X\rightarrow\infty}I(X)X^{-2}, thereby proving Theorem 2. Suppose, on the contrary, that the limit exists. Now, from the definitions (2) and (17), we have

J(X)X2=k=1I(X/2k)X2=k=14kI(X/2k)(X/2k)2,\frac{J(X)}{X^{2}}=\sum_{k=1}^{\infty}\frac{I(X/2^{k})}{X^{2}}=\sum_{k=1}^{\infty}4^{-k}\frac{I(X/2^{k})}{(X/2^{k})^{2}}\,,

and the series converge since the kk-th terms are O(4k)O(4^{-k}). Hence there exists limXJ(X)/X2\lim_{X\to\infty}J(X)/X^{2}, but this contradicts Corollary 6. Thus, our original assumption is false, and the result follows.

References

  • [1] R. J. Backlund. Über die Nullstellen der Riemannschen Zetafunktion. Acta Math.  41:345–375, 1918.
  • [2] J. Büthe. Estimating π(x)\pi(x) and related functions under partial RH assumptions. Math. Comp., 85(301):2483–2498, 2016.
  • [3] H. Cramér. Ein Mittelwertsatz in der Primzahltheorie. Math. Z., 12:147–153, 1922.
  • [4] P. Demichel, Y. Saouter, and T. Trudgian. A still sharper region where π(x)li(x)\pi(x)-\textrm{li}(x) is positive. Math. Comp., 84(295):2433–2446, 2015.
  • [5] M. Hassani. Explicit approximation of the sums over the imaginary part of the non-trivial zeros of the Riemann zeta function. Appl. Math. E-Notes, 16:109–116, 2016.
  • [6] R. S. Lehman. On the difference π(x)li(x)\pi(x)-\textrm{li}(x). Acta Arith., 11:397–410, 1966.
  • [7] H. Montgomery and R. C. Vaughan. Multiplicative Number Theory. I. Classical Theory. Cambridge Studies in Advanced Mathematics, 97. Cambridge University Press, Cambridge, 2007.
  • [8] J. Pintz. On the remainder term of the prime number formula VI. Ineffective mean value theorems. Studia Sci. Math. Hungar., 15:225–230, 1980.
  • [9] J. Pintz. On the remainder term of the prime number formula and the zeros of Riemann’s zeta-function. Number theory, Noordwijkerhout 1983, Lecture Notes in Mathematics, 1068, Springer-Verlag, Berlin, 1984.
  • [10] J. Pintz. On the mean value of the remainder term of the prime number formula. In Elementary and Analytic Theory of Numbers (Warsaw, 1982), 411–417, Banach Center Publ., 17, PWN, Warsaw, 1985.
  • [11] D. J. Platt and T. S. Trudgian. An improved explicit bound on |ζ(12+it)||\zeta(\frac{1}{2}+it)|. J. Number Theory, 147:842–851, 2015.
  • [12] S. B. Stechkin and A. Yu. Popov. Asymptotic distribution of prime numbers in the mean. Russian Math. Surveys, 51(6):1025–1092, 1996.
  • [13] E. C. Titchmarsh, edited and with a preface by D. R. Heath-Brown. The Theory of the Riemann Zeta-Function, 2nd edition. Oxford Univ. Press, New York, 1986.
  • [14] T. S. Trudgian. An improved upper bound for the argument of the Riemann zeta-function on the critical line, II. J. Number Theory, 134:280–292, 2014.