θΏ™ζ˜―indexlocζδΎ›ηš„ζœεŠ‘οΌŒδΈθ¦θΎ“ε…₯任何密码

Selberg zeta functions have second moment at Οƒ=1\sigma=1

RamΕ«nas GarunkΕ‘tis RamΕ«nas GarunkΕ‘tis
Institute of Mathematics
Faculty of Mathematics and Informatics
Vilnius University
Naugarduko 24, 03225 Vilnius, Lithuania
ramunas.garunkstis@mif.vu.lt https://klevas.mif.vu.lt/Β garunkstis/
and JokΕ«bas Putrius JokΕ«bas Putrius
Institute of Mathematics, Faculty of Mathematics and Informatics, Vilnius University
Naugarduko 24, 03225 Vilnius, Lithuania
jokubas.putrius@mif.vu.lt
(Date: November 10, 2025)
Abstract.

In this paper, we demonstrate the existence of the second moment of the Selberg zeta function for a Fuchsian group of the first kind at Οƒ=1\sigma=1. The prime geodesic theorem plays a crucial role in this context. The proof extends to Beurling zeta-functions satisfying a weak form of the Riemann hypothesis and to general Dirichlet series with positive coefficients, the partial sums of which are well-behaved. Note that by employing the recent approach of Broucke and Hilberdink in proving the second moment theorem, we can circumvent the separation condition introduced by Landau for general Dirichlet series.

Key words and phrases:
Selberg zeta-function, second moment
2020 Mathematics Subject Classification:
11M36

1. Introduction

Let s=Οƒ+i​ts=\sigma+it be the complex variable. Let Ξ“βŠ‚PSL2​(ℝ)\Gamma\subset\text{PSL}_{2}(\mathbb{R}) be any Fuchsian group of the first kind. In this paper, we consider the Selberg zeta function for Ξ“,\Gamma, which is defined for Οƒ>1\sigma>1 by the absolutely convergent Euler product (see Iwaniec [iwaniec1995, Section 10.8])

Z​(s)=∏P∏k=0∞(1βˆ’N​(P)βˆ’sβˆ’k)Z(s)=\prod_{P}\prod_{k=0}^{\infty}(1-N(P)^{-s-k})

where PP runs through all the primitive hyperbolic conjugacy classes of Ξ“\Gamma and N​(P)=Ξ±2N(P)=\alpha^{2} if the eigenvalues of PP are Ξ±\alpha and Ξ±βˆ’1\alpha^{-1}, with Ξ±\alpha having the bigger modulus. Note that for N​(P)N(P), the prime geodesic theorem is valid, i.e. there exists 0<ΞΈ<10<\theta<1 such that

(1) βˆ‘N​(P)≀x1=Li⁑(x)+O​(xΞΈ).\sum_{N(P)\leq x}1=\operatorname{Li}(x)+O(x^{\theta}).

The strongest version of the prime geodesic theorem can be found in Iwaniec [iwaniec1995, Theorem 10.5]. The function Z​(s)Z(s) has a meromorphic continuation to the whole complex plane and satisfies the functional equation

Z​(s)=Ψ​(s)​Z​(1βˆ’s),Z(s)=\Psi(s)Z(1-s),

where Ψ​(s)\Psi(s) is a meromorphic function of order 2. More about Z​(s)Z(s) see, for example, Jorgenson and SmajloviΔ‡ [js2017].

The first author of the present paper, together with Drungilas and Novikas, have already investigated the existence of the second moment of Z​(s)Z(s) for PSL2​(β„€)\text{PSL}_{2}(\mathbb{Z}) in [drungilas2021]. There a conditional result for the existence of the second moment of Z​(1+i​t)Z(1+it) and 1/Z​(1+i​t)1/Z(1+it) is proven (see [drungilas2021, Theorem 3]). This paper demonstrates that they exist unconditionally for all Ξ“\Gamma. The second moments of logarithm and the logarithmic derivative of Selberg-zeta functions were investigated in [dgk2013], Aoki [aoki2020], Hashimoto [hashimoto2023], [hashimoto2024].

For Οƒ>1\sigma>1 we can write Z​(s)Z(s) and its multiplicative inverse as a general Dirichlet series (compare to [drungilas2021, equations (11)])

Z​(s)=βˆ‘n=1∞bnyns,1Z​(s)=βˆ‘n=1∞cnyns,Z(s)=\sum_{n=1}^{\infty}\frac{b_{n}}{y^{s}_{n}},\ \frac{1}{Z(s)}=\sum_{n=1}^{\infty}\frac{c_{n}}{y^{s}_{n}},

where 1=y1<y2<…1=y_{1}<y_{2}<\dots and bn,cnβˆˆβ„€.b_{n},c_{n}\in\mathbb{Z}. Our main result is the following theorem.

Theorem 1.

Let Z​(s)Z(s) be the Selberg zeta function for a Fuchsian group of the first kind. Then

(2) limTβ†’βˆž1Tβ€‹βˆ«0T|Z​(1+i​t)|2​𝑑t=βˆ‘n=1∞bn2yn2\lim_{T\rightarrow\infty}\frac{1}{T}\int_{0}^{T}|Z(1+it)|^{2}dt=\sum_{n=1}^{\infty}\frac{b_{n}^{2}}{y_{n}^{2}}

and

(3) limTβ†’βˆž1Tβ€‹βˆ«1T|Z​(1+i​t)|βˆ’2​𝑑t=βˆ‘n=1∞cn2yn2.\lim_{T\rightarrow\infty}\frac{1}{T}\int_{1}^{T}|Z(1+it)|^{-2}dt=\sum_{n=1}^{\infty}\frac{c_{n}^{2}}{y_{n}^{2}}.

Both series on the RHS converge.

For Οƒ>1,\sigma>1, we can write

Z​(s)=Z​(s+1)β€‹βˆP(1βˆ’N​(P)βˆ’s).Z(s)=Z(s+1)\prod_{P}(1-N(P)^{-s}).

TheoremΒ 1 will follow by considering the Ruelle zeta function, which is defined as

(4) Z1​(s)=Z​(s)/Z​(s+1)=∏P(1βˆ’N​(P)βˆ’s).\displaystyle Z_{1}(s)=Z(s)/Z(s+1)=\prod_{P}(1-N(P)^{-s}).

This function has a meromorphic continuation to β„‚\mathbb{C}. For Οƒ>1,\sigma>1, we can write

Z1​(s)=βˆ‘n=1∞bnβ€²xns,1Z1​(s)=βˆ‘n=1∞cnβ€²xns,Z_{1}(s)=\sum_{n=1}^{\infty}\frac{b^{\prime}_{n}}{x^{s}_{n}},\ \frac{1}{Z_{1}(s)}=\sum_{n=1}^{\infty}\frac{c^{\prime}_{n}}{x^{s}_{n}},

where 1=x1<x2<…1=x_{1}<x_{2}<\dots and bnβ€²,cnβ€²βˆˆβ„€.b^{\prime}_{n},c^{\prime}_{n}\in\mathbb{Z}. Note that the absolute value of bnβ€²b^{\prime}_{n} is bounded by how many expressions of the form

N​(P1)​…​N​(PK)N(P_{1})\dots N(P_{K})

the number xnx_{n} has. While cnβ€²c^{\prime}_{n} is positive and equal to how many expressions of the form

N​(P1)Ξ±1​…​N​(PK)Ξ±K,N(P_{1})^{\alpha_{1}}\dots N(P_{K})^{\alpha_{K}},

Ξ±kβˆˆβ„•,\alpha_{k}\in\mathbb{N}, the number xnx_{n} has. Thus we have

(5) |bnβ€²|≀cnβ€².|b^{\prime}_{n}|\leq c^{\prime}_{n}.

In [drungilas2021], the conditional version of Theorem 1 was proved for Ξ“=PSL2​(β„€)\Gamma=\text{PSL}_{2}(\mathbb{Z}) assuming the conditions for coefficients

βˆ‘xn≀xbnβ€²,βˆ‘xn≀xcnβ€²βˆ’Οβ€‹xβ‰ͺx​exp⁑(βˆ’45​log⁑xlog⁑log⁑x),\displaystyle\sum_{x_{n}\leq x}b^{\prime}_{n},\sum_{x_{n}\leq x}c^{\prime}_{n}-\rho x\ll x\exp\left(-45\frac{\log x}{\log\log x}\right),

where ρ>0\rho>0 is the residue of Z1βˆ’1​(s)Z_{1}^{-1}(s) at s=1s=1, or the separation condition

(6) xn+1βˆ’xn≫exp⁑(βˆ’exp⁑(c​log⁑xn​log⁑log⁑xn))\displaystyle x_{n+1}-{x_{n}}\gg\exp\left(-\exp\left(c\sqrt{\log x_{n}\log\log x_{n}}\right)\right)

are valid for some specific c>0c>0. The separation condition (6) was inherited from Landau [landau1909, Chapter 75], where he investigated the second moment for the general Dirichlet series.

The function Z1Z_{1} is a Beurling zeta-function (see also [dgn2019, Ending notes] and [drungilas2021, Introduction]). Notably, the proof that the second moment of Z1​(1+i​t)Z_{1}(1+it) exists does not rely on specific properties of the Selberg zeta-function, aside from the prime geodesic theorem (1) with ΞΈ<1\theta<1. Therefore, we will derive the moment of Z1​(1+i​t)Z_{1}(1+it) from the following two general propositions, which examine the cases of the Beurling zeta-function and general Dirichlet series.

We consider generalized prime numbers systems 𝒫\mathcal{P}, often called gg-primes,

1<p1≀p2≀…,1<p_{1}\leq p_{2}\leq\dots,

with pjβˆˆβ„p_{j}\in\mathbb{R} tending to infinity and the associated Beurling zeta-function

(7) ΢𝒫​(s)=∏j=1∞(1βˆ’1pjs)βˆ’1.\displaystyle\zeta_{\mathcal{P}}(s)=\prod_{j=1}^{\infty}\left(1-\frac{1}{p_{j}^{s}}\right)^{-1}.

A formal multiplication of the product gives the Dirichlet series

(8) ΢𝒫​(s)=βˆ‘n=1∞dnΞ½ns,\displaystyle\zeta_{\mathcal{P}}(s)=\sum_{n=1}^{\infty}\frac{d_{n}}{\nu_{n}^{s}},

where 1=Ξ½1<Ξ½2<Ξ½3<…1=\nu_{1}<\nu_{2}<\nu_{3}<\dots is the increasing sequence of power-products of gg-primes with corresponding multiplicities 1=d1,d2,d3,…1=d_{1},d_{2},d_{3},\dots. Similarly, there are integers 1=e1,e2,e3,…,1=e_{1},e_{2},e_{3},\dots, such that

(9) 1΢𝒫​(s)=∏j=1∞(1βˆ’1pjs)=βˆ‘n=1∞enΞ½ns.\displaystyle\frac{1}{\zeta_{\mathcal{P}}(s)}=\prod_{j=1}^{\infty}\left(1-\frac{1}{p_{j}^{s}}\right)=\sum_{n=1}^{\infty}\frac{e_{n}}{\nu_{n}^{s}}.

As in (5) we have

(10) |en|≀dn.|e_{n}|\leq d_{n}.

In this paper, we always assume that there is a fixed 0<Ξ±<10<\alpha<1, such that

(11) π𝒫​(x)=βˆ‘pn≀x1=Li⁑(x)+O​(xΞ±).\displaystyle\pi_{\mathcal{P}}(x)=\sum_{p_{n}\leq x}1=\operatorname{Li}(x)+O(x^{\alpha}).

Then Hilberdink and Lapidus [hilberdink2006, Theorem 2.1] obtained that ΢𝒫​(s)\zeta_{\mathcal{P}}(s) has an analytic continuation to the half-plane β„œβ‘s>Ξ±\Re s>\alpha except for a simple pole at s=1s=1 and ΢𝒫​(s)β‰ 0\zeta_{\mathcal{P}}(s)\neq 0 in this region, With these considerations, we formulate the following proposition.

Proposition 2.

Under the assumption (11), we have that

(12) limTβ†’βˆž1Tβ€‹βˆ«1T|΢𝒫​(1+i​t)|2​𝑑t=βˆ‘n=1∞dn2Ξ½n2\lim_{T\rightarrow\infty}\frac{1}{T}\int_{1}^{T}|\zeta_{\mathcal{P}}(1+it)|^{2}dt=\sum_{n=1}^{\infty}\frac{d^{2}_{n}}{\nu_{n}^{2}}

and

(13) limTβ†’βˆž1Tβ€‹βˆ«1T|΢𝒫​(1+i​t)|βˆ’2​𝑑t=βˆ‘n=1∞en2Ξ½n2.\lim_{T\rightarrow\infty}\frac{1}{T}\int_{1}^{T}|\zeta_{\mathcal{P}}(1+it)|^{-2}dt=\sum_{n=1}^{\infty}\frac{e^{2}_{n}}{\nu_{n}^{2}}.

Both series on the RHS converge.

Moreover in [hilberdink2006, Theorem 2.2] it is proved that

(14) βˆ‘xn≀xdn=η​x+D​(x),\displaystyle\sum_{x_{n}\leq x}d_{n}=\eta x+D(x),

where Ξ·>0\eta>0 is a constant and

(15) D​(x)β‰ͺx​exp⁑(βˆ’c​log⁑x​log⁑log⁑x),\displaystyle D(x)\ll x\exp\left(-c\sqrt{\log x\log\log x}\right),

with c>0c>0 fixed. Note that Broucke, Debruyne, and Vindas [bdv20] constructed a Beurling generalized number system which satisfies (11) with Ξ±=1/2\alpha=1/2 and (14) with D​(x)=Ω​(x​exp⁑(βˆ’C​log⁑x​log⁑log⁑x))D(x)=\Omega\left(x\exp\left(-C\sqrt{\log x\log\log x}\right)\right) for some C>0C>0. Proposition 2 allows to remove the separation condition of the type (6) in Theorem 1 in [drungilas2021], where the second moment of ΢𝒫±1​(1+i​t)\zeta_{\mathcal{P}}^{\pm 1}(1+it) was treated.

Proposition 2 will follow from a statement for general Dirichlet series. Denote

f​(s)=βˆ‘j=1∞ajnjs,f(s)=\sum_{j=1}^{\infty}\frac{a_{j}}{n_{j}^{s}},

where (nj)(n_{j}) is a strictly increasing sequence of positive reals which tends to infinity and aj>0a_{j}>0. Then the second moment of f​(s)f(s) at Οƒ=1\sigma=1 exists even for a weaker bound than in (15).

Proposition 3.

Let aj>0a_{j}>0, Ξ΅>0\varepsilon>0 and assume that

A​(x)=βˆ‘nj≀xaj=ϱ​x+R​(x),A(x)=\sum_{n_{j}\leq x}a_{j}=\varrho x+R(x),

where Ο±>0\varrho>0 is a constant and

(16) R​(x)β‰ͺx(log⁑x)3/2+Ξ΅.\displaystyle R(x)\ll\frac{x}{(\log x)^{3/2+\varepsilon}}.

Then the sum defining f​(s)f(s) converges for Οƒ>1\sigma>1 and the function has a continuous extension to Οƒβ‰₯1\sigma\geq 1, sβ‰ 1s\neq 1, and

(17) limTβ†’βˆž1Tβ€‹βˆ«1T|f​(1+i​t)|2​𝑑t=βˆ‘j=1∞aj2nj2.\displaystyle\lim_{T\rightarrow\infty}\frac{1}{T}\int_{1}^{T}|f(1+it)|^{2}dt=\sum_{j=1}^{\infty}\frac{a^{2}_{j}}{n^{2}_{j}}.

The series on the RHS converges.

Proposition 3 extends the recent result of Broucke and Hilberdink [bh], where they obtained that the second moment of f​(s)f(s) exists for Οƒ>(1+Ξ·)/2\sigma>(1+\eta)/2, if βˆ‘nj≀xaj=ϱ​x+O​(xΞ·)\sum_{n_{j}\leq x}a_{j}=\varrho x+O(x^{\eta}) for some fixed Ο±>0\varrho>0 and Ξ·<1\eta<1. The significant feature of their approach is that they do not require any separation condition for njn_{j} in the proof.

We establish Proposition 3 in the second section. The third section is dedicated to proving Theorem 1 and Proposition 2.

We use notation f​(x)β‰ͺg​(x)f(x)\ll g(x) or f​(x)=O​(g​(x))f(x)=O(g(x)) to mean that |f​(x)|≀c​|g​(x)||f(x)|\leq c|g(x)| for some constant c>0,c>0, when xx is big enough. Additionally, we use f​(x)β‰Ίg​(x)f(x)\prec g(x) or f​(x)=o​(g​(x))f(x)=o(g(x)) to signify that f​(x)/g​(x)β†’0,f(x)/g(x)\rightarrow 0, when xβ†’βˆž.x\rightarrow\infty. The notation f​(x)=Ω​(g​(x))f(x)=\Omega(g(x)) means that exists a constant CC such that the inequality |f​(x)|>C​|g​(x)||f(x)|>C|g(x)| holds infinitely often for arbitrarily large values of xx.

2. Proof of Proposition 3

For the proof of Proposition 3 we adapt the ideas of Broucke and Hilberdink [bh]. Let

fN​(s)=βˆ‘nj≀Najnjs.f_{N}(s)=\sum_{n_{j}\leq N}\frac{a_{j}}{n_{j}^{s}}.
Lemma 4.

The following holds

limTβ†’βˆž1Tβ€‹βˆ«0T|fN​(1+i​t)|2​𝑑t=βˆ‘j=1∞aj2nj2\lim_{T\rightarrow\infty}\frac{1}{T}\int_{0}^{T}|f_{N}(1+it)|^{2}dt=\sum_{j=1}^{\infty}\frac{a_{j}^{2}}{n_{j}^{2}}

as N,Tβ†’βˆžN,T\rightarrow\infty such that (log⁑N)2β‰ΊT.(\log N)^{2}\prec T.

Proof.

Assume that N>ee.N>e^{e}. We compute

(18) 1Tβ€‹βˆ«0T|fN​(1+i​t)|2​𝑑t=12​Tβ€‹βˆ«βˆ’TT|βˆ‘nj≀Najnj1+i​t|2​𝑑t=βˆ‘nj,nk≀Naj​ak2​T​nj​nkβ€‹βˆ«βˆ’TT(njnk)i​t​𝑑t=βˆ‘nj≀Naj2nj2+βˆ‘nj<nk≀Naj​ak2​T​nj​nkβ€‹βˆ«βˆ’TT(njnk)i​t+(nknj)i​t​d​t=βˆ‘nj≀Naj2nj2+2β€‹βˆ‘nj<nk≀Naj​aknj​nk​s​(T​log⁑(nk/nj)),\begin{split}&\frac{1}{T}\int_{0}^{T}|f_{N}(1+it)|^{2}dt=\frac{1}{2T}\int_{-T}^{T}\left|\sum_{n_{j}\leq N}\frac{a_{j}}{n^{1+it}_{j}}\right|^{2}dt\\ &=\sum_{n_{j},n_{k}\leq N}\frac{a_{j}a_{k}}{2Tn_{j}n_{k}}\int_{-T}^{T}\left(\frac{n_{j}}{n_{k}}\right)^{it}dt\\ &=\sum_{n_{j}\leq N}\frac{a_{j}^{2}}{n_{j}^{2}}+\sum_{n_{j}<n_{k}\leq N}\frac{a_{j}a_{k}}{2Tn_{j}n_{k}}\int_{-T}^{T}\left(\frac{n_{j}}{n_{k}}\right)^{it}+\left(\frac{n_{k}}{n_{j}}\right)^{it}dt\\ &=\sum_{n_{j}\leq N}\frac{a_{j}^{2}}{n_{j}^{2}}+2\sum_{n_{j}<n_{k}\leq N}\frac{a_{j}a_{k}}{n_{j}n_{k}}s(T\log(n_{k}/n_{j})),\end{split}

where s​(x)=sin⁑(x)/x.s(x)=\sin(x)/x. We need to show that for every Ξ·>0\eta>0 there exists a T0>0T_{0}>0 such that

(19) |βˆ‘nj<nk≀Naj​aknj​nk​s​(T​log⁑(nk/nj))|<Ξ·,\left|\sum_{n_{j}<n_{k}\leq N}\frac{a_{j}a_{k}}{n_{j}n_{k}}s(T\log(n_{k}/n_{j}))\right|<\eta,

when T>T0T>T_{0} and (log⁑N)2β‰ΊT.(\log N)^{2}\prec T.

Denote

Rk=nk(log⁑nk)3/2+Ρ.R_{k}=\frac{n_{k}}{(\log n_{k})^{3/2+\varepsilon}}.

Let n0>0n_{0}>0. We divide the sum in (19) into three parts

(βˆ‘nj<nk≀n0+βˆ‘n0<nj<nk≀Nnj≀nkβˆ’Rk+βˆ‘n0<nj<nk≀Nnj>nkβˆ’Rk)aj​aknj​nks(Tlog(nk/nj))=:S1+S2+S3.\displaystyle\left(\sum_{n_{j}<n_{k}\leq n_{0}}+\sum_{{n_{0}<n_{j}<n_{k}\leq N\atop n_{j}\leq n_{k}-R_{k}}}+\sum_{{n_{0}<n_{j}<n_{k}\leq N\atop n_{j}>n_{k}-R_{k}}}\right)\frac{a_{j}a_{k}}{n_{j}n_{k}}s(T\log(n_{k}/n_{j}))=:S_{1}+S_{2}+S_{3}.

To get the bound (19) we will prove the following three statements.

(i) There is n0=n0​(Ξ·)n_{0}=n_{0}(\eta) such that |S3|<Ξ·/3|S_{3}|<\eta/3 for all TT;

(ii) there is T1=T1​(Ξ·)T_{1}=T_{1}(\eta) such that |S2|≀η/3|S_{2}|\leq\eta/3 for T>T1T>T_{1} and all n0n_{0};

(iii) for any given n0n_{0} there is T2=T2​(Ξ·,n0)T_{2}=T_{2}(\eta,n_{0}) such that |S1|≀η/3|S_{1}|\leq\eta/3 for T>T2T>T_{2}.

Proof of (i). For S3S_{3} use |s​(x)|≀1.|s(x)|\leq 1. Then, for all TT,

|S3|β‰€βˆ‘n0≀nk≀Naknkβ€‹βˆ‘nkβˆ’Rk≀nj<nkajnjβ‰ͺβˆ‘n0≀nk<Nak​(A​(nk)βˆ’A​(nkβˆ’Rk))nk​(nkβˆ’Rk)β‰ͺβˆ‘n0≀nkak(nkβˆ’Rk)​(log⁑nk)3/2+Ξ΅.\begin{split}&|S_{3}|\leq\sum_{n_{0}\leq n_{k}\leq N}\frac{a_{k}}{n_{k}}\sum_{n_{k}-R_{k}\leq n_{j}<n_{k}}\frac{a_{j}}{n_{j}}\\ &\ll\sum_{n_{0}\leq n_{k}<N}\frac{a_{k}(A(n_{k})-A(n_{k}-R_{k}))}{n_{k}(n_{k}-R_{k})}\\ &\ll\sum_{n_{0}\leq n_{k}}\frac{a_{k}}{(n_{k}-R_{k})(\log n_{k})^{3/2+\varepsilon}}.\end{split}

The last sum converges if and only if the sum

(20) βˆ‘n0≀nkaknk​(log⁑nk)3/2+Ξ΅\sum_{n_{0}\leq n_{k}}\frac{a_{k}}{n_{k}(\log n_{k})^{3/2+\varepsilon}}

converges. Using Abel’s summation formula we obtain

βˆ‘n0≀nk≀Naknk​(log⁑nk)3/2+Ξ΅β‰ͺ1(log⁑N)3/2+Ξ΅+1(log⁑n0)3/2+Ξ΅+∫n0N1u​(log⁑(u))3/2+Ρ​𝑑uβ‰ͺ1.\begin{split}&\sum_{n_{0}\leq n_{k}\leq N}\frac{a_{k}}{n_{k}(\log n_{k})^{3/2+\varepsilon}}\ll\frac{1}{(\log N)^{3/2+\varepsilon}}+\frac{1}{(\log n_{0})^{3/2+\varepsilon}}\\ &+\int_{n_{0}}^{N}\frac{1}{u(\log(u))^{3/2+\varepsilon}}du\ll 1.\end{split}

Thus, the sum (20) converges. Hence, we can pick n0n_{0} large enough such that, for all TT,

|S3|<Ξ·3.\displaystyle|S_{3}|<\frac{\eta}{3}.

This proves (i).

Proof of (ii). For S2S_{2} use |s​(x)|≀xβˆ’1|s(x)|\leq x^{-1} and log⁑(nk/nj)β‰₯(nkβˆ’nj)/nk.\log(n_{k}/n_{j})\geq(n_{k}-n_{j})/n_{k}. Then

|S2|≀1Tβ€‹βˆ‘nk≀Naknkβ€‹βˆ‘nj≀nkβˆ’Rkaj​nknj​(nkβˆ’nj)β‰ͺ1Tβ€‹βˆ‘nk≀Naknk​(βˆ‘nj≀nk/2ajnj+βˆ‘nk/2<nj≀nkβˆ’Rkajnkβˆ’nj).\begin{split}&|S_{2}|\leq\frac{1}{T}\sum_{n_{k}\leq N}\frac{a_{k}}{n_{k}}\sum_{n_{j}\leq n_{k}-R_{k}}\frac{a_{j}n_{k}}{n_{j}(n_{k}-n_{j})}\\ &\ll\frac{1}{T}\sum_{n_{k}\leq N}\frac{a_{k}}{n_{k}}\left(\sum_{n_{j}\leq n_{k}/2}\frac{a_{j}}{n_{j}}+\sum_{n_{k}/2<n_{j}\leq n_{k}-R_{k}}\frac{a_{j}}{n_{k}-n_{j}}\right).\end{split}

Using Abel’s summation formula twice we obtain

1Tβ€‹βˆ‘nk≀Naknkβ€‹βˆ‘nj≀nk/2ajnjβ‰ͺ1Tβ€‹βˆ‘nk≀Naknk​(2​A​(nk/2)nk+∫1nkA​(u)u2​𝑑u)β‰ͺ1Tβ€‹βˆ‘nk≀Nak​log⁑nknkβ‰ͺlog⁑NT​(A​(N)N+∫1NA​(u)u2​𝑑u)β‰ͺ(log⁑N)2T.\begin{split}&\frac{1}{T}\sum_{n_{k}\leq N}\frac{a_{k}}{n_{k}}\sum_{n_{j}\leq n_{k}/2}\frac{a_{j}}{n_{j}}\ll\frac{1}{T}\sum_{n_{k}\leq N}\frac{a_{k}}{n_{k}}\left(\frac{2A(n_{k}/2)}{n_{k}}+\int_{1}^{n_{k}}\frac{A(u)}{u^{2}}du\right)\\ &\ll\frac{1}{T}\sum_{n_{k}\leq N}\frac{a_{k}\log n_{k}}{n_{k}}\ll\frac{\log N}{T}\left(\frac{A(N)}{N}+\int_{1}^{N}\frac{A(u)}{u^{2}}du\right)\ll\frac{(\log N)^{2}}{T}.\end{split}

Now we estimate

βˆ‘nk/2<nj≀nkβˆ’Rkajnkβˆ’nj.\sum_{n_{k}/2<n_{j}\leq n_{k}-R_{k}}\frac{a_{j}}{n_{k}-n_{j}}.

Let K​(k)K(k) be the unique integer such that nk/4≀2K​(k)​Rk<nk/2.n_{k}/4\leq 2^{K(k)}R_{k}<n_{k}/2. Thus

K​(k)=3/2+Ξ΅log⁑2​log⁑log⁑nk+O​(1)β‰ͺlog⁑log⁑nk.K(k)=\frac{3/2+\varepsilon}{\log 2}\log\log n_{k}+O(1)\ll\log\log n_{k}.

Then

βˆ‘nk/2<nj≀nkβˆ’Rkajnkβˆ’nj=βˆ‘r=1K​(k)βˆ‘2rβˆ’1​Rk≀nkβˆ’nj<2r​Rkajnkβˆ’nj+βˆ‘2K​(k)​Rk≀nkβˆ’nj≀nk/2ajnkβˆ’njβ‰ͺβˆ‘r=1K​(k)A​(nkβˆ’2rβˆ’1​Rk)βˆ’A​(nkβˆ’2r​Rk)2rβˆ’1​Rk+A​(3​nk/4)βˆ’A​(nk/2)2K​(k)​Rkβ‰ͺβˆ‘r=1K​(k)2r​Rk2rβˆ’1​Rk+1β‰ͺK​(k)β‰ͺlog⁑log⁑nk.\begin{split}&\sum_{n_{k}/2<n_{j}\leq n_{k}-R_{k}}\frac{a_{j}}{n_{k}-n_{j}}\\ &=\sum_{r=1}^{K(k)}\sum_{2^{r-1}R_{k}\leq n_{k}-n_{j}<2^{r}R_{k}}\frac{a_{j}}{n_{k}-n_{j}}+\sum_{2^{K(k)}R_{k}\leq n_{k}-n_{j}\leq n_{k}/2}\frac{a_{j}}{n_{k}-n_{j}}\\ &\ll\sum_{r=1}^{K(k)}\frac{A(n_{k}-2^{r-1}R_{k})-A(n_{k}-2^{r}R_{k})}{2^{r-1}R_{k}}+\frac{A(3n_{k}/4)-A(n_{k}/2)}{2^{K(k)}R_{k}}\\ &\ll\sum_{r=1}^{K(k)}\frac{2^{r}R_{k}}{2^{r-1}R_{k}}+1\ll K(k)\ll\log\log n_{k}.\end{split}

Thus,

(21) |S2|β‰ͺ(log⁑N)2T+1Tβ€‹βˆ‘nk≀Nak​log⁑log⁑nknkβ‰ͺ(log⁑N)2T\displaystyle|S_{2}|\ll\frac{(\log N)^{2}}{T}+\frac{1}{T}\sum_{n_{k}\leq N}\frac{a_{k}\log\log n_{k}}{n_{k}}\ll\frac{(\log N)^{2}}{T}

uniformly in n0n_{0}. This proves the statement (ii), since (log⁑N)2β‰ΊT(\log N)^{2}\prec T.

The statement (iii) follows by the bound S1β‰ͺn0Tβˆ’1.S_{1}\ll_{n_{0}}T^{-1}.

We have proven (19) and this proves the lemma. ∎

Lemma 5.

We have

1Tβ€‹βˆ«1T|f​(1+i​t)βˆ’fN​(1+i​t)|2​𝑑tβ‰ͺ1T2+1(log⁑N)3+2​Ρ+T(log⁑N)2​(1+Ξ΅)\frac{1}{T}\int_{1}^{T}|f(1+it)-f_{N}(1+it)|^{2}dt\ll\frac{1}{T^{2}}+\frac{1}{(\log N)^{3+2\varepsilon}}+\frac{T}{(\log N)^{2(1+\varepsilon)}}

uniformly in N>eeN>e^{e}.

Proof.

Assume that N>ee.N>e^{e}. For Οƒ>1\sigma>1

(22) f​(s)βˆ’fN​(s)=∫N∞1xs​𝑑A​(x)=βˆ’Ο±β€‹N1βˆ’s1βˆ’sβˆ’R​(N)Ns+sβ€‹βˆ«N∞R​(x)xs+1​𝑑x.f(s)-f_{N}(s)=\int_{N}^{\infty}\frac{1}{x^{s}}dA(x)=-\frac{\varrho N^{1-s}}{1-s}-\frac{R(N)}{N^{s}}+s\int_{N}^{\infty}\frac{R(x)}{x^{s+1}}dx.

The integral

∫N∞d​xx​(log⁑x)3/2+Ξ΅β‰ͺ1(log⁑N)1/2+Ξ΅\int_{N}^{\infty}\frac{dx}{x(\log x)^{3/2+\varepsilon}}\ll\frac{1}{(\log N)^{1/2+\varepsilon}}

converges. Thus, the equation (22) holds when Οƒ=1\sigma=1 and gives the continuous extension of f​(s)f(s) to Οƒβ‰₯1\sigma\geq 1. Then, for t∈[1,2​T]t\in[1,2T] we have

(23) |f​(1+i​t)βˆ’fN​(1+i​t)|β‰ͺ1t+1(log⁑N)3/2+Ξ΅+T​|∫N∞R​(x)x2+i​t​𝑑x|.\displaystyle|f(1+it)-f_{N}(1+it)|\ll\frac{1}{t}+\frac{1}{(\log N)^{3/2+\varepsilon}}+T\left|\int_{N}^{\infty}\frac{R(x)}{x^{2+it}}dx\right|.

Using the inequality (a+b+c)2≀3​(a2+b2+c2)(a+b+c)^{2}\leq 3(a^{2}+b^{2}+c^{2}) we get

1Tβ€‹βˆ«1T|f​(1+i​t)βˆ’fN​(1+i​t)|2​𝑑tβ‰ͺ1T2+1(log⁑N)3+2​Ρ+Tβ€‹βˆ«1T|∫N∞R​(x)x2+i​t​𝑑x|2​𝑑t.\frac{1}{T}\int_{1}^{T}|f(1+it)-f_{N}(1+it)|^{2}dt\ll\frac{1}{T^{2}}+\frac{1}{(\log N)^{3+2\varepsilon}}+T\int_{1}^{T}\left|\int_{N}^{\infty}\frac{R(x)}{x^{2+it}}dx\right|^{2}dt.

Therefore, it is enough to evaluate

I​(T)=∫12​T|∫N∞R​(x)x2+i​t​𝑑x|2​𝑑t.I(T)=\int_{1}^{2T}\left|\int_{N}^{\infty}\frac{R(x)}{x^{2+it}}dx\right|^{2}dt.

To do so we use the observation that the Fourier transform of

14​T​(1βˆ’|t|4​T)+=14​T​max⁑{1βˆ’|t|4​T,0}\displaystyle\frac{1}{4T}\left(1-\frac{|t|}{4T}\right)_{+}=\frac{1}{4T}\max\left\{1-\frac{|t|}{4T},0\right\}

is

(24) s2​(2​T​u)=(sin⁑(2​T​u)2​T​u)2.\displaystyle s^{2}(2Tu)=\left(\frac{\sin(2Tu)}{2Tu}\right)^{2}.

Define

J​(T)=14​Tβ€‹βˆ«βˆ’4​T4​T(1βˆ’|t|4​T)+​|∫N∞R​(x)x2+i​t​𝑑x|2​𝑑t.J(T)=\frac{1}{4T}\int_{-4T}^{4T}\left(1-\frac{|t|}{4T}\right)_{+}\left|\int_{N}^{\infty}\frac{R(x)}{x^{2+it}}dx\right|^{2}dt.

Note that I​(T)≀4​T​J​(T).I(T)\leq 4TJ(T). Now square out the integral in J​(T),J(T), interchange the order of integration and use the observation to see that

J​(T)β‰ͺ∫N∞∫N∞|R​(x)​R​(y)|(x​y)2​s2​(2​T​log⁑xy)​𝑑y​𝑑x.J(T)\ll\int_{N}^{\infty}\int_{N}^{\infty}\frac{|R(x)R(y)|}{(xy)^{2}}s^{2}(2T\log\frac{x}{y})dydx.

By symmetry, it suffices to bound the integral over the domain N≀y≀x.N\leq y\leq x. Split the integral over yy in three parts: N≀y≀max⁑{N,x/2},N\leq y\leq\max\{N,x/2\}, max⁑{N,x/2}≀y≀max⁑{N,xβˆ’x/T}\max\{N,x/2\}\leq y\leq\max\{N,x-x/T\}, and max⁑{N,xβˆ’x/T}≀y≀x.\max\{N,x-x/T\}\leq y\leq x. Use s2​(u)≀min⁑{1,uβˆ’2}s^{2}(u)\leq\min\{1,u^{-2}\} and log⁑(x/y)β‰₯(xβˆ’y)/x.\log(x/y)\geq(x-y)/x. Then the contribution of the first range is bounded by

β‰ͺ1T2β€‹βˆ«N∞1x​(log⁑x)3/2+Ξ΅β€‹βˆ«Nx/2x2y​(xβˆ’y)2​(log⁑y)3/2+Ρ​𝑑y​𝑑xβ‰ͺ1T2β€‹βˆ«N∞1x​(log⁑x)3/2+Ξ΅β€‹βˆ«Nx/21y​(log⁑y)3/2+Ρ​𝑑y​𝑑xβ‰ͺ1T2​(log⁑N)1+2​Ρ.\begin{split}&\ll\frac{1}{T^{2}}\int_{N}^{\infty}\frac{1}{x(\log x)^{3/2+\varepsilon}}\int_{N}^{x/2}\frac{x^{2}}{y(x-y)^{2}(\log y)^{3/2+\varepsilon}}dydx\\ &\ll\frac{1}{T^{2}}\int_{N}^{\infty}\frac{1}{x(\log x)^{3/2+\varepsilon}}\int_{N}^{x/2}\frac{1}{y(\log y)^{3/2+\varepsilon}}dydx\ll\frac{1}{T^{2}(\log N)^{1+2\varepsilon}}.\end{split}

The second range gives

β‰ͺ1T2β€‹βˆ«N∞1x​(log⁑x)3/2+Ξ΅β€‹βˆ«x/2xβˆ’x/Tx2y​(xβˆ’y)2​(log⁑y)3/2+Ρ​𝑑y​𝑑xβ‰ͺ1Tβ€‹βˆ«N∞d​xx​(log⁑x)3+2​Ρβ‰ͺ1T​(log⁑N)2​(1+Ξ΅),\begin{split}&\ll\frac{1}{T^{2}}\int_{N}^{\infty}\frac{1}{x(\log x)^{3/2+\varepsilon}}\int_{x/2}^{x-x/T}\frac{x^{2}}{y(x-y)^{2}(\log y)^{3/2+\varepsilon}}dydx\\ &\ll\frac{1}{T}\int_{N}^{\infty}\frac{dx}{x(\log x)^{3+2\varepsilon}}\ll\frac{1}{T(\log N)^{2(1+\varepsilon)}},\end{split}

here we used the substitution u=x/(xβˆ’y).u=x/(x-y). The last range yields

β‰ͺ∫N∞1x​(log⁑x)3/2+Ξ΅β€‹βˆ«xβˆ’x/Tx1y​(log⁑y)3/2+Ρ​𝑑y​𝑑xβ‰ͺ1Tβ€‹βˆ«N∞d​xx​(log⁑x)3+2​Ρ\displaystyle\ll\int_{N}^{\infty}\frac{1}{x(\log x)^{3/2+\varepsilon}}\int_{x-x/T}^{x}\frac{1}{y(\log y)^{3/2+\varepsilon}}dydx\ll\frac{1}{T}\int_{N}^{\infty}\frac{dx}{x(\log x)^{3+2\varepsilon}}
β‰ͺ1T​(log⁑N)2​(1+Ξ΅).\displaystyle\ll\frac{1}{T(\log N)^{2(1+\varepsilon)}}.

Combining everything we get the desired bound. ∎

Proof of Proposition 3.

The continuous extension of f​(s)f(s) is obtained at the beginning of the proof of Lemma 5.

Next, we rewrite the LHS of (17) as

1Tβ€‹βˆ«0T|f​(1+i​t)|2​𝑑t=\displaystyle\frac{1}{T}\int_{0}^{T}|f(1+it)|^{2}dt= 1Tβ€‹βˆ«0T|fN​(1+i​t)|2​𝑑t+∫0T|f​(1+i​t)βˆ’fN​(1+i​t)|2​𝑑t\displaystyle\frac{1}{T}\int_{0}^{T}|f_{N}(1+it)|^{2}dt+\int_{0}^{T}|f(1+it)-f_{N}(1+it)|^{2}dt
(25) +2β€‹β„œβ‘(1Tβ€‹βˆ«0TfN​(1+i​t)​(f​(1+i​t)βˆ’fN​(1+i​t)Β―)​𝑑t),\displaystyle+2\Re\left(\frac{1}{T}\int_{0}^{T}f_{N}(1+it)(\overline{f(1+it)-f_{N}(1+it)})dt\right),

where the last term is bounded by

2​(1T2β€‹βˆ«0T|fN​(1+i​t)|2​𝑑tβ€‹βˆ«0T|f​(1+i​t)βˆ’fN​(1+i​t)|2​𝑑t)1/2.2\left(\frac{1}{T^{2}}\int_{0}^{T}|f_{N}(1+it)|^{2}dt\int_{0}^{T}|f(1+it)-f_{N}(1+it)|^{2}dt\right)^{1/2}.

Then choosing NN such that (log⁑N)2β‰ΊTβ‰Ί(log⁑N)2​(1+Ξ΅)(\log N)^{2}\prec T\prec(\log N)^{2(1+\varepsilon)} by Lemmas 4 and 5 we obtain the moment (17).

The series on the RHS of (17) converges because

βˆ‘j=1∞aj2nj2β‰ͺβˆ‘j=1∞aj​(A​(nj)βˆ’A​(njβˆ’1))nj2β‰ͺβˆ‘n=1∞ajnj​(log⁑nj)3/2+Ξ΅.\sum_{j=1}^{\infty}\frac{a^{2}_{j}}{n_{j}^{2}}\ll\sum_{j=1}^{\infty}\frac{a_{j}(A(n_{j})-A(n_{j}-1))}{n_{j}^{2}}\ll\sum_{n=1}^{\infty}\frac{a_{j}}{n_{j}(\log n_{j})^{3/2+\varepsilon}}.

This concludes the proposition. ∎

3. Proofs of Theorem 1 and Proposition 2

Proof of Proposition 2.

Equality (12) follows from Proposition 3.

The series on the RHS of (13) converges because, using (10), we get

βˆ‘n=1∞en2xn2β‰€βˆ‘n=1∞dn2xn2.\sum_{n=1}^{\infty}\frac{e^{2}_{n}}{x_{n}^{2}}\leq\sum_{n=1}^{\infty}\frac{d^{2}_{n}}{x_{n}^{2}}.

Next, we prove equality (13). Let

΢𝒫​N​(s)=βˆ‘xn≀Ndnxns,(ΞΆπ’«βˆ’1)N​(s)=βˆ‘xn≀Nenxns,\zeta_{\mathcal{P}N}(s)=\sum_{x_{n}\leq N}\frac{d_{n}}{x^{s}_{n}},\ (\zeta_{\mathcal{P}}^{-1})_{N}(s)=\sum_{x_{n}\leq N}\frac{e_{n}}{x^{s}_{n}},

where (ΞΆπ’«βˆ’1)N​(s)(\zeta_{\mathcal{P}}^{-1})_{N}(s) denotes the partial Dirichlet sum of 1/΢𝒫​(s)1/\zeta_{\mathcal{P}}(s). To prove (13) we will show that for (ΞΆπ’«βˆ’1)N​(1+i​t)(\zeta_{\mathcal{P}}^{-1})_{N}(1+it) and 1/΢𝒫​(1+i​t)βˆ’(ΞΆπ’«βˆ’1)N​(1+i​t)1/\zeta_{\mathcal{P}}(1+it)-(\zeta_{\mathcal{P}}^{-1})_{N}(1+it) the statements analogous to Lemmas 4 and 5 hold.

To get the analog of Lemma 4 we use the equality

1Tβ€‹βˆ«0T|(ΞΆπ’«βˆ’1)N​(1+i​t)|2​𝑑t=βˆ‘Ξ½j≀Nej2Ξ½j2+2β€‹βˆ‘Ξ½j<Ξ½k≀Nej​ekΞ½j​νk​s​(T​log⁑(Ξ½k/Ξ½j))\frac{1}{T}\int_{0}^{T}|(\zeta_{\mathcal{P}}^{-1})_{N}(1+it)|^{2}dt=\sum_{\nu_{j}\leq N}\frac{e^{2}_{j}}{\nu_{j}^{2}}+2\sum_{\nu_{j}<\nu_{k}\leq N}\frac{e_{j}e_{k}}{\nu_{j}\nu_{k}}s(T\log(\nu_{k}/\nu_{j}))

and by (10) we get

|βˆ‘Ξ½j<Ξ½k≀Nej​ekΞ½j​νk​s​(T​log⁑(Ξ½k/Ξ½j))|β‰€βˆ‘xj<xk≀Ndj​dkΞ½j​νk​|s​(T​log⁑(Ξ½k/Ξ½j))|.\left|\sum_{\nu_{j}<\nu_{k}\leq N}\frac{e_{j}e_{k}}{\nu_{j}\nu_{k}}s(T\log(\nu_{k}/\nu_{j}))\right|\leq\sum_{x_{j}<x_{k}\leq N}\frac{d_{j}d_{k}}{\nu_{j}\nu_{k}}|s(T\log(\nu_{k}/\nu_{j}))|.

Then in the same way as in Lemma 4 we get that

βˆ‘xj<xk≀Ndj​dkΞ½j​νk​|s​(T​log⁑(Ξ½k/Ξ½j))|=βˆ‘xj<xk≀Ndj​dkΞ½j​νk​s​(T​log⁑(Ξ½k/Ξ½j))β†’0,\displaystyle\sum_{x_{j}<x_{k}\leq N}\frac{d_{j}d_{k}}{\nu_{j}\nu_{k}}|s(T\log(\nu_{k}/\nu_{j}))|=\sum_{x_{j}<x_{k}\leq N}\frac{d_{j}d_{k}}{\nu_{j}\nu_{k}}s(T\log(\nu_{k}/\nu_{j}))\rightarrow 0,

if (log⁑N)2β‰ΊT.(\log N)^{2}\prec T. Hence,

(26) limTβ†’βˆž1Tβ€‹βˆ«0T|(ΞΆπ’«βˆ’1)N​(1+i​t)|2​𝑑t=βˆ‘n=1∞en2Ξ½n2\displaystyle\lim_{T\rightarrow\infty}\frac{1}{T}\int_{0}^{T}|(\zeta_{\mathcal{P}}^{-1})_{N}(1+it)|^{2}dt=\sum_{n=1}^{\infty}\frac{e^{2}_{n}}{\nu_{n}^{2}}

as N,Tβ†’βˆžN,T\rightarrow\infty such that (log⁑N)2β‰ΊT.(\log N)^{2}\prec T.

We consider the statement analogous to Lemma 5. By the discussion below Theorem 1 in [dgn2019] we have that

(27) E​(x):=βˆ‘Ξ½n≀xenβ‰ͺx​exp⁑(βˆ’c​log⁑x​log⁑log⁑x)(c>0).\displaystyle E(x):=\sum_{\nu_{n}\leq x}e_{n}\ll x\exp\left(-c\sqrt{\log x\log\log x}\right)\quad(c>0).

Clearly, x​exp⁑(βˆ’c​log⁑x​log⁑log⁑x)β‰ͺx​(log⁑x)βˆ’3/2βˆ’Ξ΅x\exp\left(-c\sqrt{\log x\log\log x}\right)\ll x(\log x)^{-3/2-\varepsilon}. Similarly to (23) we obtain

(28) |΢𝒫​(1+i​t)βˆ’1βˆ’(ΞΆπ’«βˆ’1)N​(1+i​t)|=βˆ’E​(N)N1+i​t+T​|∫N∞E​(x)x2+i​t​𝑑x|.\displaystyle\left|\zeta_{\mathcal{P}}(1+it)^{-1}-(\zeta_{\mathcal{P}}^{-1})_{N}(1+it)\right|=-\frac{E(N)}{N^{1+it}}+T\left|\int_{N}^{\infty}\frac{E(x)}{x^{2+it}}dx\right|.

Further, following the proof of Lemma 5 (replace R​(x)R(x) by E​(x)E(x)) we arrive at

(29) 1Tβ€‹βˆ«1T|΢𝒫​(1+i​t)βˆ’1βˆ’(ΞΆπ’«βˆ’1)N​(1+i​t)|2​𝑑tβ‰ͺT(log⁑N)2​(1+Ξ΅).\displaystyle\frac{1}{T}\int_{1}^{T}|\zeta_{\mathcal{P}}(1+it)^{-1}-(\zeta_{\mathcal{P}}^{-1})_{N}(1+it)|^{2}dt\ll\frac{T}{(\log N)^{2(1+\varepsilon)}}.

By (26) and (29) similarly as in the proof of Theorem 3 we derive (13).

Before concluding the proof, we note that the sharper bound (27) (compared to (16)) enables us to derive a bound of the type (29) without utilizing the Fourier transform (24). Indeed, from (28) by (27) we obtain

|΢𝒫​(1+i​t)βˆ’1βˆ’(ΞΆπ’«βˆ’1)N​(1+i​t)|β‰ͺT​exp⁑(βˆ’c​log⁑N)​log⁑N\displaystyle|\zeta_{\mathcal{P}}(1+it)^{-1}-(\zeta_{\mathcal{P}}^{-1})_{N}(1+it)|\ll T\exp\left(-c\sqrt{\log N}\right)\sqrt{\log N}

and

1Tβ€‹βˆ«1T|΢𝒫​(1+i​t)βˆ’1βˆ’(ΞΆπ’«βˆ’1)N​(1+i​t)|2​𝑑tβ‰ͺT2​exp⁑(βˆ’2​c​log⁑N)​log⁑N.\displaystyle\frac{1}{T}\int_{1}^{T}|\zeta_{\mathcal{P}}(1+it)^{-1}-(\zeta_{\mathcal{P}}^{-1})_{N}(1+it)|^{2}dt\ll T^{2}\exp\left(-2c\sqrt{\log N}\right)\log N.

Then choosing T=(log⁑N)3T=(\log N)^{3} we again derive (13). Similarly, the moment (12) can be proved without using the Fourier transform. ∎

Proof of Theorem 1.

Theorem 1 follows from Proposition 2. The proof is similar to that of Theorem 3 in [drungilas2021], but we include a shortened version here for the reader’s convenience.

We write (see (4))

Z​(s)=Z1​(s)​Z2​(s),Z(s)=Z_{1}(s)Z_{2}(s),

where Z1Z_{1} is the Ruelle zeta function and Z2​(s)=Z​(s+1).Z_{2}(s)=Z(s+1).

The function Z2Z_{2} is bounded on the line Οƒ=1,\sigma=1, thus there exists a constant C>0C>0 such that

|Z2​(1+i​t)|Β±2≀C,|Z_{2}(1+it)|^{\pm 2}\leq C,

for all tβˆˆβ„.t\in\mathbb{R}. Also, the Dirichlet series representation of this function is absolutely convergent at Οƒ=1.\sigma=1.

We are going to show that the series

βˆ‘n=1∞bn2yn2andβˆ‘n=1∞cn2yn2\sum_{n=1}^{\infty}\frac{b^{2}_{n}}{y_{n}^{2}}\quad\text{and}\quad\sum_{n=1}^{\infty}\frac{c^{2}_{n}}{y_{n}^{2}}

on the right-hand side of (2) and (3) converge. We start with the second one.

For Οƒ>1,\sigma>1, write

1Z1​(s)=βˆ‘n=1∞cnβ€²xnsand1Z2​(s)=βˆ‘n=1∞cnβ€²β€²yns,\frac{1}{Z_{1}(s)}=\sum_{n=1}^{\infty}\frac{c^{\prime}_{n}}{x_{n}^{s}}\quad\text{and}\quad\frac{1}{Z_{2}(s)}=\sum_{n=1}^{\infty}\frac{c^{\prime\prime}_{n}}{y_{n}^{s}},

Also let

(Z1βˆ’1)N​(s)=βˆ‘xn≀Ncnβ€²xns(Z^{-1}_{1})_{N}(s)=\sum_{x_{n}\leq N}\frac{c^{\prime}_{n}}{x_{n}^{s}}

be the partial sum of 1/Z1​(s)1/Z_{1}(s).

Let

cn​(N)=βˆ‘xj​yk=ynxj≀Ncj′​ckβ€²β€².c_{n}(N)=\sum_{x_{j}y_{k}=y_{n}\atop x_{j}\leq N}c^{\prime}_{j}c^{\prime\prime}_{k}.

Then,

0≀cn​(N)β‰€βˆ‘xj​yk=yncj′​ckβ€²β€²=cn.0\leq c_{n}(N)\leq\sum_{x_{j}y_{k}=y_{n}}c^{\prime}_{j}c^{\prime\prime}_{k}=c_{n}.

In view of [landau1909, Satz 29 in Β§221], we obtain

(30) βˆ‘yn≀Ncn2yn2+βˆ‘yn>Ncn2​(N)yn2=limTβ†’βˆž1Tβ€‹βˆ«1T|Z1​N​(1+i​t)​Z2​(1+i​t)|βˆ’2​𝑑t≀Cβ€‹βˆ‘n=1∞cn′⁣2xn2.\begin{split}&\sum_{y_{n}\leq N}\frac{c^{2}_{n}}{y_{n}^{2}}+\sum_{y_{n}>N}\frac{c^{2}_{n}(N)}{y_{n}^{2}}=\lim_{T\rightarrow\infty}\frac{1}{T}\int_{1}^{T}|Z_{1N}(1+it)Z_{2}(1+it)|^{-2}dt\leq\\ &C\sum_{n=1}^{\infty}\frac{c^{\prime 2}_{n}}{x_{n}^{2}}.\end{split}

By Proposition 2, the series on the far right-hand side above converges, therefore the series on the left-hand side also converges. Thus

limNβ†’βˆžlimTβ†’βˆž1Tβ€‹βˆ«1T|Z1​N​(1+i​t)​Z2​(1+i​t)|βˆ’2​𝑑t=βˆ‘n=1∞cn2yn2\lim_{N\rightarrow\infty}\lim_{T\rightarrow\infty}\frac{1}{T}\int_{1}^{T}|Z_{1N}(1+it)Z_{2}(1+it)|^{-2}dt=\sum_{n=1}^{\infty}\frac{c^{2}_{n}}{y_{n}^{2}}

and the series on the right-hand side converges. Using inequalities (5), |bn|≀cn|b_{n}|\leq c_{n} and reasoning as in (30) we have that

limNβ†’βˆžlimTβ†’βˆž1Tβ€‹βˆ«1T|Z1​N​(1+i​t)​Z2​(1+i​t)|2​𝑑t=βˆ‘n=1∞bn2yn2\lim_{N\rightarrow\infty}\lim_{T\rightarrow\infty}\frac{1}{T}\int_{1}^{T}|Z_{1N}(1+it)Z_{2}(1+it)|^{2}dt=\sum_{n=1}^{\infty}\frac{b^{2}_{n}}{y_{n}^{2}}

and that the series on the right-hand side converges.

Similarly as in (2), we get

1T​(∫1T|Z​(1+i​t)|βˆ’2​𝑑tβˆ’βˆ«1T|Z1​N​(1+i​t)​Z2​(1+i​t)|βˆ’2​𝑑t)≀CTβ€‹βˆ«1T|Z​(1+i​t)βˆ’1βˆ’(Z1βˆ’1)N​(1+i​t)|​𝑑t+2​CTβ„œβˆ«1T(Z(1+it)βˆ’1βˆ’(Z1βˆ’1)N(1+it))(Z1βˆ’1)N​(1+i​t)Β―)dt.\begin{split}&\frac{1}{T}\left(\int_{1}^{T}|Z(1+it)|^{-2}dt-\int_{1}^{T}|Z_{1N}(1+it)Z_{2}(1+it)|^{-2}dt\right)\leq\\ &\frac{C}{T}\int_{1}^{T}|Z(1+it)^{-1}-(Z^{-1}_{1})_{N}(1+it)|dt+\\ &\frac{2C}{T}\Re\int_{1}^{T}\left(Z(1+it)^{-1}-(Z^{-1}_{1})_{N}(1+it))\overline{(Z^{-1}_{1})_{N}(1+it)}\right)dt.\end{split}

Using the Cauchy-Schwarz inequality, Lemmas 4 and 5, and choosing NN such that (log⁑N)2β‰ΊTβ‰Ί(log⁑N)2​(1+Ξ΅)(\log N)^{2}\prec T\prec(\log N)^{2(1+\varepsilon)}, we see that the left-hand side of the above inequality goes to 0 as Tβ†’βˆž.T\rightarrow\infty. Therefore, (3) is proven. To prove (2), we argue likewise, using (26) and (29) instead of Lemmas 4 and 5. ∎

Acknowledgment. This work is funded by the Research Council of Lithuania (LMTLT), agreement No. S-MIP-22-81.