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

First moment of central values of quadratic Dirichlet LL-functions

Peng Gao School of Mathematical Sciences, Beihang University, Beijing 100191, China penggao@buaa.edu.cn  and  Liangyi Zhao School of Mathematics and Statistics, University of New South Wales, Sydney NSW 2052, Australia l.zhao@unsw.edu.au
Abstract.

We evaluate the first moment of central values of the family of quadratic Dirichlet LL-functions using the method of double Dirichlet series. Under the generalized Riemann hypothesis, we prove an asymptotic formula with an error term of size that is the fourth root of that of the primary main term.

Mathematics Subject Classification (2010): 11M06, 11M41

Keywords: quadratic Dirichlet LL-functions, first moment, double Dirichlet series

1. Introduction

Moments of central values of families of LL-functions have been widely studied in the literature as they have many important applications. In this paper, we are interested in the first moment of central values of the family of quadratic Dirichlet LL-functions. For this family, an asymptotic formula for the first moment was initially obtained by M. Jutila [Jutila] with the main term of size XlogXX\log X and an error term of size O(X3/4+ε)O(X^{3/4+\varepsilon}) for any ε>0\varepsilon>0. An error term of the same size was later given by A. I. Vinogradov and L. A. Takhtadzhyan in [ViTa]. Using the method of double Dirichlet series, D. Goldfeld and J. Hoffstein [DoHo] improved the error term to O(X19/32+ε)O(X^{19/32+\varepsilon}). It is also implicit in [DoHo] that one may obtain an error term of size O(X1/2+ε)O(X^{1/2+\varepsilon}) for the smoothed first moment, a result that is achieved via a different approach by M. P. Young [Young1] who utlized a recursive argument. The optimal error term is conjectured to be O(X1/4+ε)O(X^{1/4+\varepsilon}) in [DoHo] and this been observed in a numerical study conducted by M. W. Alderson and M. O. Rubinstein in [AR12]. In the function field setting, owing partially to the established truth of the Riemann hypothesis there, the analogous asymptotic formula with an error term of the conjectured size was obtained by A. M. Florea [Florea17].

The method of multiple Dirichlet series is a powerful tool when studying moments of LL-functions. The success of such method relies heavily on the analytic properties of these series. In [DoHo], D. Goldfeld and J. Hoffstein used a double Dirichlet series in their work by treating the variables separately. They applied the theory of Eisenstein series of metaplectic type to obtain analytic continuation of the series in one variable. It was later pointed out by A. Diaconu, D. Goldfeld and J. Hoffstein in [DGH] that there are many advantages in viewing multiple Dirichlet series as functions of several complex variables. From this point of view, much progress has been made towards understanding analytic properties of various multiple Dirichlet series in [DGH], including a result on third moment of central values of the family of quadratic Dirichlet LL-functions.

For any integer m0,1(mod4)m\equiv 0,1\pmod{4}, let χ(m)=(m)\chi^{(m)}=\left(\frac{m}{\cdot}\right) be the Kronecker symbol defined on [iwakow, p. 52]. As usual, ζ(s)\zeta(s) is the Riemann zeta function. For any LL-function, we write L(c)L^{(c)} (resp. L(c)L_{(c)}) for the function given by the Euler product defining LL but omitting those primes dividing (resp. not dividing) cc. We reserve the letter pp for a prime throughout the paper and we write LpL_{p} for L(p)L_{(p)} for simplicity. In [Blomer11], V. Blomer obtained meromorphic continuation to the whole 2\mathbb{C}^{2} for the double Dirichlet series given by

ζ(2)(2s+2w1)(d,2)=1L(s,χ(4d)ψ)ψ(d)dw.\displaystyle\zeta^{(2)}(2s+2w-1)\sum_{(d,2)=1}\frac{L(s,\chi^{(4d)}\psi)\psi^{\prime}(d)}{d^{w}}.

The primary goal of [Blomer11] is to establish a subconvexity bound of the above double series. Its analytic properties actually also allow one to evaluate the smoothed first moment of L(1/2,χ(4d))L(1/2,\chi^{(4d)}) given by

(1.1) (d,2)=1L(12,χ(4d))w(dX),\displaystyle\sum_{(d,2)=1}L(\tfrac{1}{2},\chi^{(4d)})w\left(\frac{d}{X}\right),

where w(t)w(t) is a non-negative Schwartz function. Under generalized Riemann hypothesis (GRH) and arguing in a manner similar to (3.42) below, one is able to obtain an asymptotical formula for the expression above with the error term being the conjectured size O(X1/4+ε)O(X^{1/4+\varepsilon}).

The success of Blomer’s method relies on obtaining enough functional equations for the underlying double Dirichlet series. However, it is generally a challenging task to establish meromorphic continuation of other multiple Dirichlet series to the entire complex space this way in order to study the corresponding first moment. Thus, it is desirable to seek for an alternative approach to circumvent this difficulty. For this, we note that M. Čech [Cech1] investigated the LL-functions ratios conjecture for the case of quadratic Dirichlet LL-functions using multiple Dirichlet series. The advantage of this method is that instead of pursuing meromorphic continuation to the entire complex space for the multiple Dirichlet series involved as done in other works, one makes a crucial use of the functional equation of a general (not necessarily primitive) quadratic Dirichlet LL-function [Cech1, Proposition 2.3] to extend the multiple Dirichlet series under consideration to a suitable large region for the purpose of the investigation.

Motivated by the work of Čech, we adapt the approach in [Cech1] to assess the first moment of central values of a family of quadratic Dirichlet LL-functions. To state our result, we write χn\chi_{n} for the quadratic character (n)\left(\frac{\cdot}{n}\right) for an odd, positive integer nn. By the quadratic reciprocity law, L(2)(s,χn)=L(s,χ(4n))L^{(2)}(s,\chi_{n})=L(s,\chi^{(4n)}) (resp. L(s,χ(4n))L(s,\chi^{(-4n)})) if n1(mod4)n\equiv 1\pmod{4} (resp. n1(mod4)n\equiv-1\pmod{4}). Notice that one can factor every such mm uniquely into m=dl2m=dl^{2} so that dd is a fundamental discriminant, i.e. dd is either square-free and d1(mod4)d\equiv 1\pmod{4} or d=4nd=4n with n2,3(mod4)n\equiv 2,3\pmod{4} and square-free. It is known (see [MVa1, Theorem 9.13]) that every primitive quadratic Dirichlet character is of the form χ(d)\chi^{(d)} for some fundamental discriminant dd. For such dd, it follows from [iwakow, Theorem 4.15] that the function L(s,χ(d))L(s,\chi^{(d)}) has an analytic continuation to the entirety of \mathbb{C}. Thus the same can be said of L(2)(s,χn)L^{(2)}(s,\chi_{n}).

In this paper, we evaluate asymptotically the first moment of the family of quadratic Dirichlet LL-functions L(2)(s,χn)L^{(2)}(s,\chi_{n}) averaged over all odd, positive nn. Our main result is as follows.

Theorem 1.1.

Under the notation as above and the truth of GRH, suppose that w(t)w(t) is a non-negative Schwartz function and w^(s)\widehat{w}(s) is its Mellin transform. For 1/2>(α)>01/2>\Re(\alpha)>0 and any ε>0\varepsilon>0, we have

(1.2) (n,2)=1L(2)(12+α,χn)w(nX)=Xw^(1)ζ(1+2α)ζ(2+2α)1212α2(1222α)+X1αw^(1α)παΓ(1/2α)Γ(α2)Γ(1α2)Γ(α)ζ(12α)ζ(2)22α6+O((1+|α|)5+εX1/4+ε).\displaystyle\begin{split}\sum_{\begin{subarray}{c}(n,2)=1\end{subarray}}L^{(2)}(\tfrac{1}{2}+\alpha,\chi_{n})w\left(\frac{n}{X}\right)=&X\widehat{w}(1)\frac{\zeta(1+2\alpha)}{\zeta(2+2\alpha)}\frac{1-2^{-1-2\alpha}}{2(1-2^{-2-2\alpha})}+X^{1-\alpha}\widehat{w}(1-\alpha)\frac{\pi^{\alpha}\Gamma(1/2-\alpha)\Gamma(\frac{\alpha}{2})}{\Gamma(\frac{1-\alpha}{2})\Gamma(\alpha)}\frac{\zeta(1-2\alpha)}{\zeta(2)}\frac{2^{2\alpha}}{6}\\ &\hskip 85.35826pt+O\left((1+|\alpha|)^{5+\varepsilon}X^{1/4+\varepsilon}\right).\end{split}

Notice that the error term in (1.2) is uniform for α\alpha, we can therefore take the limit α0+\alpha\rightarrow 0^{+} to deduce the following asymptotic formula for the smoothed first moment of central values of the family of quadratic Dirichlet LL-functions under consideration.

Corollary 1.2.

With the notation as above and assuming the truth of GRH, we have, for any ε>0\varepsilon>0,

(1.3) (n,2)=1L(2)(12,χn)w(nX)=XQ(logX)+O(X1/4+ε).\displaystyle\begin{split}&\sum_{\begin{subarray}{c}(n,2)=1\end{subarray}}L^{(2)}(\tfrac{1}{2},\chi_{n})w\left(\frac{n}{X}\right)=XQ(\log X)+O\left(X^{1/4+\varepsilon}\right).\end{split}

where QQ is a linear polynomial whose coefficients depend only on the absolute constants and w^(1)\widehat{w}(1) and w^(1)\widehat{w}^{\prime}(1).

Note that our error term above is consistent with the conjecture size given in [DoHo]. The explicit expression of QQ is omitted here as our main focus is the error term. The proof of Theorem 1.1 requires one to obtain meromorphic continuation of certain double Dirichlet series, which we get by making a crucial use of the functional equation of a general quadratic Dirichlet LL-function in [Cech1, Proposition 2.3] to convert the original double Dirichlet series to its dual series which we carefully analyze using the ideas of K. Soundararajan and M. P. Young in [S&Y].

We remark here that our proof of Theorem 1.1 implies that (1.3) holds with the error term O(X1/2+ε)O\left(X^{1/2+\varepsilon}\right) unconditionally. It may also be applied to study the first moment of the family of quadratic Dirichlet LL-functions given in (1.1). For this reason, we have included a large sieve result (Lemma 2.7 below) which shall be applied to control the size of the LL-values on average without the Lindelöf hypothesis, although the latter may lead to better error terms. We shall discuss this further after the proof of Lemma 3.4.

Lastly, we point out that the novelty of our method in the paper is that instead of seeking for meromorphic continuation of the underlying multiple Dirichlet series to the entire complex space, we only aim to meromorphically continue such series to a region large enough for our purpose. This provides a great degree of flexibility in our treatment and can be easily adapted to investigate other problems. For example, one may study the first moment of families of primitive quadratic Dirichlet LL-functions using the functional equation of the primitive quadratic Dirichlet LL-functions themselves to obtain an asymptotic formula with an error term O(X1/2+ε)O(X^{1/2+\varepsilon}), recovering a result of M.P. Young [Young1]. The case for primitive quadratic Hecke LL-functions over imaginary quadratic number fields of class number one has been implicitly worked out in [G&Zhao2023] and the arguments therein carry over to Dirichlet LL-functions as well. We also note that the analogous result of A. Florea [Florea17] on the first moment of primitive quadratic LL-functions over function fields suggests that there may be a secondary main term of size X1/3X^{1/3} for the number fields case as well. Thus, some new insights may be required for an asymptotic formula for the first moment of primitive quadratic Dirichlet LL-functions, if one desires an error term of size O(X1/4+ε)O(X^{1/4+\varepsilon}).

We end this section by making the convention that, throughout the paper, ε\varepsilon denotes a small positive quantity that may not be the same in each appearance and the implied constants in \ll and OO can depend on ε\varepsilon.

2. Preliminaries

2.1. Gauss sums

We write ψj=χ(4j)\psi_{j}=\chi^{(4j)} for j=±1,±2j=\pm 1,\pm 2 where we recall that χ(d)=(d)\chi^{(d)}=\left(\frac{d}{\cdot}\right) is the Kronecker symbol for integers d0,1(mod4)d\equiv 0,1\pmod{4}. Note that each ψj\psi_{j} is a character modulo 4|j|4|j|. Let ψ0\psi_{0} stand for the primitive principal character.

Given any Dirichlet character χ\chi modulo nn and any integer qq, the Gauss sum τ(χ,q)\tau(\chi,q) is defined to be

τ(χ,q)=j(modn)χ(j)e(jqn),wheree(z)=exp(2πiz).\tau(\chi,q)=\sum_{j\negthickspace\negthickspace\negthickspace\pmod{n}}\chi(j)e\left(\frac{jq}{n}\right),\quad\mbox{where}\quad e(z)=\exp(2\pi iz).

For the evaluation of τ(χ,q)\tau(\chi,q), we cite the following result from [Cech1, Lemma 2.2].

Lemma 2.2.
  1. (1)

    If l1(mod4)l\equiv 1\pmod{4}, then

    τ(χ(4l),q)={0,if (q,2)=1,2τ(χl,q),if q2(mod4),2τ(χl,q),if q0(mod4).\tau\left(\chi^{(4l)},q\right)=\begin{cases}0,&\hbox{if $(q,2)=1$,}\\ -2\tau\left(\chi_{l},q\right),&\hbox{if $q\equiv 2\pmod{4}$,}\\ 2\tau\left(\chi_{l},q\right),&\hbox{if $q\equiv 0\pmod{4}$.}\end{cases}
  2. (2)

    If l3(mod4)l\equiv 3\pmod{4}, then

    τ(χ(4l),q)={0,if 2|q,2iτ(χl,q),if q1(mod4),2iτ(χl,q),if q3(mod4).\tau\left(\chi^{(4l)},q\right)=\begin{cases}0,&\hbox{if $2|q$,}\\ -2i\tau\left(\chi_{l},q\right),&\hbox{if $q\equiv 1\pmod{4}$,}\\ 2i\tau\left(\chi_{l},q\right),&\hbox{if $q\equiv 3\pmod{4}$.}\end{cases}

Recall that for an odd positive integer nn, we define χn=(n)\chi_{n}=\left(\frac{\cdot}{n}\right). We then define an associated Gauss sum G(χn,q)G\left(\chi_{n},q\right) by

G(χn,q)=(1i2+(1n)1+i2)τ(χn,q)={τ(χn,q),if n1(mod4),iτ(χn,q),if n3(mod4).\displaystyle\begin{split}G\left(\chi_{n},q\right)&=\left(\frac{1-i}{2}+\left(\frac{-1}{n}\right)\frac{1+i}{2}\right)\tau\left(\chi_{n},q\right)=\begin{cases}\tau\left(\chi_{n},q\right),&\hbox{if $n\equiv 1\pmod{4}$,}\\ -i\tau\left(\chi_{n},q\right),&\hbox{if $n\equiv 3\pmod{4}$}.\end{cases}\end{split}

The advantage of G(χn,q)G\left(\chi_{n},q\right) over τ(χn,q)\tau\left(\chi_{n},q\right) is that G(χn,q)G\left(\chi_{n},q\right) is now a multiplicative function of nn. In fact, upon denoting φ(m)\varphi(m) for the Euler totient function of mm, we have the following result from [sound1, Lemma 2.3] that evaluates G(χn,q)G\left(\chi_{n},q\right).

Lemma 2.3.

If (m,n)=1(m,n)=1 then G(χmn,q)=G(χm,q)G(χn,q)G(\chi_{mn},q)=G(\chi_{m},q)G(\chi_{n},q). Suppose that pap^{a} is the largest power of pp dividing qq (put a=a=\infty if m=0m=0). Then for k0k\geq 0 we have

G(χpk,q)={φ(pk),if kak even,0,if kak odd,pa,if k=a+1k even,(qpap)pap,if k=a+1k odd,0,if ka+2.G\left(\chi_{p^{k}},q\right)=\begin{cases}\varphi(p^{k}),&\hbox{if $k\leq a$, $k$ even,}\\ 0,&\hbox{if $k\leq a$, $k$ odd,}\\ -p^{a},&\hbox{if $k=a+1$, $k$ even,}\\ \left(\frac{qp^{-a}}{p}\right)p^{a}\sqrt{p},&\hbox{if $k=a+1$, $k$ odd,}\\ 0,&\hbox{if $k\geq a+2$}.\end{cases}

2.4. Functional equations for Dirichlet LL-functions

We quote the following functional equation from [Cech1, Proposition 2.3] concerning all Dirichlet characters χ\chi modulo nn, which plays a key role in our proof of Theorem 1.1.

Lemma 2.5.

Let χ\chi be any Dirichlet character modulo nn\neq\square such that χ(1)=1\chi(-1)=1. Then we have

(2.1) L(s,χ)=πs1/2nsΓ(1s2)Γ(s2)K(1s,χ),whereK(s,χ)=q=1τ(χ,q)qs.L(s,\chi)=\frac{\pi^{s-1/2}}{n^{s}}\frac{\Gamma\left(\frac{1-s}{2}\right)}{\Gamma\left(\frac{s}{2}\right)}K(1-s,\chi),\quad\mbox{where}\quad K(s,\chi)=\sum_{q=1}^{\infty}\frac{\tau(\chi,q)}{q^{s}}.

2.6. Bounding LL-functions

For a fixed quadratic character ψ\psi modulo nn, let ψ^\widehat{\psi} be the primitive character that induces ψ\psi so that we have ψ^=χ(d)\widehat{\psi}=\chi^{(d)} for some fundamental discriminant d|nd|n (see [MVa1, Theorem 9.13]). We gather in this section certain estimations on L(s,ψ)L(s,\psi) that are necessary in the proof of Theorem 1.1. Most of the estimations here are unconditional, except for the following one, which asserts that when (s)1/2+ε\Re(s)\geq 1/2+\varepsilon for any ε>0\varepsilon>0, we have by [iwakow, Theorem 5.19] that under GRH,

(2.2) |L(s,ψ^)|1|sn|ε.\displaystyle\begin{split}&\big{|}L(s,\widehat{\psi})\big{|}^{-1}\ll|sn|^{\varepsilon}.\end{split}

Write n=n1n2n=n_{1}n_{2} uniquely such that (n1,d)=1(n_{1},d)=1 and that p|n2p|dp|n_{2}\Rightarrow p|d. The above notations imply that for any integer qq,

(2.3) L(q)(s,ψ)=L(s,ψ^)p|qn1(1ψ^(p)ps).\displaystyle\begin{split}L^{(q)}(s,\psi)=L(s,\widehat{\psi})\prod_{p|qn_{1}}\left(1-\frac{\widehat{\psi}(p)}{p^{s}}\right).\end{split}

Observe that

|1ψ^(p)ps|2pmax(0,(s)).\displaystyle\Big{|}1-\frac{\widehat{\psi}(p)}{p^{s}}\Big{|}\leq 2p^{\max(0,-\Re(s))}.

We then deduce that

(2.4) p|qn1(1ψ^(p)ps)2ω(q1n)(qn1)max(0,(s))(qn1)max(0,(s))+ε,\displaystyle\begin{split}\prod_{p|qn_{1}}\left(1-\frac{\widehat{\psi}(p)}{p^{s}}\right)\ll 2^{\omega(q_{1}n)}(qn_{1})^{\max(0,-\Re(s))}\ll(qn_{1})^{\max(0,-\Re(s))+\varepsilon},\end{split}

where ω(n)\omega(n) denotes the number of distinct prime factors of nn and the last estimation above follows from the well-known bound (see [MVa1, Theorem 2.10])

ω(h)loghloglogh,forh3.\displaystyle\omega(h)\ll\frac{\log h}{\log\log h},\quad\mbox{for}\quad h\geq 3.

When dd is a fundamental discriminant, we recall the convexity bound for L(s,χ(d))L(s,\chi^{(d)}) (see [iwakow, Exercise 3, p. 100]) asserts that

(2.5) L(s,χ(d)){(|d|(1+|s|))(1(s))/2+ε,0(s)1,1,(s)>1.\displaystyle\begin{split}L(s,\chi^{(d)})\ll\begin{cases}&\left(|d|(1+|s|)\right)^{(1-\Re(s))/2+\varepsilon},\quad 0\leq\Re(s)\leq 1,\\ &1,\quad\Re(s)>1.\end{cases}\end{split}

To estimate L(s,χ(d))L(s,\chi^{(d)}) for (s)<0\Re(s)<0, we note the following functional equation (see [sound1, p. 456]) for a primitive even character χ\chi.

(2.6) Λ(s,χ(d)):=(|d|π)s/2Γ(s2)L(s,χ(d))=Λ(1s,χ(d)).\displaystyle\Lambda(s,\chi^{(d)}):=\Big{(}\frac{|d|}{\pi}\Big{)}^{s/2}\Gamma\Big{(}\frac{s}{2}\Big{)}L(s,\chi^{(d)})=\Lambda(1-s,\chi^{(d)}).

Moreover, Stirling’s formula ([iwakow, (5.113)]) implies that, for constants a0a_{0}, b0b_{0},

(2.7) Γ(a0(1s)+b0)Γ(a0s+b0)(1+|s|)a0(12(s)).\displaystyle\frac{\Gamma(a_{0}(1-s)+b_{0})}{\Gamma(a_{0}s+b_{0})}\ll(1+|s|)^{a_{0}(1-2\Re(s))}.

We also conclude from (2.5)–(2.7) that

(2.8) L(s,χ(d)){1(s)>1,(|d|(1+|s|))(1(s))/2+ε0(s)<1,(|d|(1+|s|))1/2(s)+ε(s)<0.\displaystyle\begin{split}L(s,\chi^{(d)})\ll\begin{cases}1\qquad&\Re(s)>1,\\ (|d|(1+|s|))^{(1-\Re(s))/2+\varepsilon}\qquad&0\leq\Re(s)<1,\\ (|d|(1+|s|))^{1/2-\Re(s)+\varepsilon}\qquad&\Re(s)<0.\end{cases}\end{split}

From (2.3), (2.4) and (2.8), we deduce that for all complex numbers ss,

(2.9) L(q)(s,ψ)(qn1)max(0,(s))+ε(n(1+|s|))max{1/2(s),(1(s))/2,0}+ε.\displaystyle\begin{split}L^{(q)}(s,\psi)\ll(qn_{1})^{\max(0,-\Re(s))+\varepsilon}(n(1+|s|))^{\max\{1/2-\Re(s),(1-\Re(s))/2,0\}+\varepsilon}.\end{split}

We conclude this section by including the following large sieve result for quadratic Dirichlet LL-functions, which is a consequence of [DRHB, Theorem 2].

Lemma 2.7.

With the notation as above, let S(X)S(X) denote the set of real, primitive characters χ\chi with conductor not exceeding XX. Then we have, for any complex number ss with (s)1/2\Re(s)\geq 1/2 and any ε>0\varepsilon>0,

(2.10) χS(X)|L(s,χ)|\displaystyle\sum_{\begin{subarray}{c}\chi\in S(X)\end{subarray}}|L(s,\chi)|\ll X1+ε|s|1/4+ε.\displaystyle X^{1+\varepsilon}|s|^{1/4+\varepsilon}.
Proof.

From [DRHB, Theorem 2], we get

χS(X)|L(s,χ)|4\displaystyle\sum_{\begin{subarray}{c}\chi\in S(X)\end{subarray}}|L(s,\chi)|^{4}\ll (X|s|)1+ε.\displaystyle(X|s|)^{1+\varepsilon}.

The lemma now follows from the above and Hölder’s inequality. ∎

2.8. Some results on multivariable complex functions

We gather here some results from multivariable complex analysis. We begin with the notation of a tube domain.

Definition 2.9.

An open set TnT\subset\mathbb{C}^{n} is a tube if there is an open set UnU\subset\mathbb{R}^{n} such that T={zn:(z)U}.T=\{z\in\mathbb{C}^{n}:\ \Re(z)\in U\}.

For a set UnU\subset\mathbb{R}^{n}, we define T(U)=U+innT(U)=U+i\mathbb{R}^{n}\subset\mathbb{C}^{n}. We quote the following Bochner’s Tube Theorem [Boc].

Theorem 2.10.

Let UnU\subset\mathbb{R}^{n} be a connected open set and f(z)f(z) a function holomorphic on T(U)T(U). Then f(z)f(z) has a holomorphic continuation to the convex hull of T(U)T(U).

We denote the convex hull of an open set TnT\subset\mathbb{C}^{n} by T^\widehat{T}. Our next result is [Cech1, Proposition C.5] on the modulus of holomorphic continuations of multivariable complex functions.

Proposition 2.11.

Assume that TnT\subset\mathbb{C}^{n} is a tube domain, g,h:Tg,h:T\rightarrow\mathbb{C} are holomorphic functions, and let g~,h~\tilde{g},\tilde{h} be their holomorphic continuations to T^\widehat{T}. If |g(z)||h(z)||g(z)|\leq|h(z)| for all zTz\in T and h(z)h(z) is nonzero in TT, then also |g~(z)||h~(z)||\tilde{g}(z)|\leq|\tilde{h}(z)| for all zT^z\in\widehat{T}.

3. Proof of Theorem 1.1

For (s)\Re(s), (w)\Re(w) sufficiently large, define

(3.1) A(s,w)=(n,2)=1L(2)(w,χn)ns=(nm,2)=1χn(m)mwns=(m,2)=1L(s,χ(4m))mw.\displaystyle\begin{split}A(s,w)=&\sum_{\begin{subarray}{c}(n,2)=1\end{subarray}}\frac{L^{(2)}(w,\chi_{n})}{n^{s}}=\sum_{\begin{subarray}{c}(nm,2)=1\end{subarray}}\frac{\chi_{n}(m)}{m^{w}n^{s}}=\sum_{\begin{subarray}{c}(m,2)=1\end{subarray}}\frac{L(s,\chi^{(4m)})}{m^{w}}.\end{split}

We shall develop some analytic properties of A(s,w)A(s,w), necessary in establishing Theorem 1.1. Before delving into this analysis, we make the following observation. In [Blomer11, Lemma 2], Blomer developed a relation between 𝐙(s,w)\mathbf{Z}(s,w) (which may be regarded of as an analogue of our A(s,w)A(s,w) in (3.1)) and 𝐙(w,s)\mathbf{Z}(w,s). It is thus a natural question to ask if such a relation is possible for A(s,w)A(s,w) and A(w,s)A(w,s) using quadratic reciprocity. As remarked in the introduction of this paper, that the advantage of considering all characters is that we can convert A(s,w)A(s,w) to a dual sum and make use of the properties of the dual sum to obtain a good error term. If we consider primitive LL-functions, then it would be possible to obtain a functional equation relating A(s,w)A(s,w) and A(w,s)A(w,s), but then the error term would be much larger.

3.1. First region of absolute convergence of A(s,w)A(s,w)

Using the first equality in the definition for A(s,w)A(s,w) in (3.1), we get

(3.2) A(s,w)=(n,2)=1L(2)(w,χn)ns=(h,2)=11h2s(n,2)=1L(2)(w,χn)p|h(1χn(p)pw)ns=ζ(2)(2s)(n,2)=1L(2)(w,χn)nsL(2)(2s+w,χn),\displaystyle\begin{split}A(s,w)=&\sum_{\begin{subarray}{c}(n,2)=1\end{subarray}}\frac{L^{(2)}(w,\chi_{n})}{n^{s}}=\sum_{\begin{subarray}{c}(h,2)=1\end{subarray}}\frac{1}{h^{2s}}\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(n,2)=1\end{subarray}}\frac{L^{(2)}(w,\chi_{n})\prod_{p|h}(1-\chi_{n}(p)p^{-w})}{n^{s}}\\ =&\zeta^{(2)}(2s)\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(n,2)=1\end{subarray}}\frac{L^{(2)}(w,\chi_{n})}{n^{s}L^{(2)}(2s+w,\chi_{n})},\end{split}

where \sum^{*} henceforth denotes the sum over square-free integers. Note that the last equality follows by using the identity (which is obtained by writing the corresponding series into an Euler product)

(h,2)=11h2sp|h(1χn(p)pw)=ζ(2)(2s)L(2)(2s+w,χn).\sum_{\begin{subarray}{c}(h,2)=1\end{subarray}}\frac{1}{h^{2s}}\prod_{p|h}\left(1-\frac{\chi_{n}(p)}{p^{w}}\right)=\frac{\zeta^{(2)}(2s)}{L^{(2)}(2s+w,\chi_{n})}.

Write χ~n\widetilde{\chi}_{n} for the primitive Dirichlet character that induces χ(n)\chi^{(n)} (resp. χ(n)\chi^{(-n)}) for n1(mod4)n\equiv 1\pmod{4} (resp. for n1(mod4)n\equiv-1\pmod{4}). Recall that we have L(2)(w,χn)=L(w,χ(±4n))L^{(2)}(w,\chi_{n})=L(w,\chi^{(\pm 4n)}) for n±1(mod4)n\equiv\pm 1\pmod{4}. For a square-free integer nn, we see that χ~n=χ(n)\widetilde{\chi}_{n}=\chi^{(n)} (resp. χ~n=χ(4n)\widetilde{\chi}_{n}=\chi^{(4n)} ) is a primitive character modulo nn for n1(mod4)n\equiv 1\pmod{4} (resp. n1(mod4)n\equiv-1\pmod{4}). In either case, for square-free integers nn,

|L(2)(w,χn)|=|(1χ~n(2)2w)L(w,χ~n)|.\displaystyle\begin{split}\big{|}L^{(2)}(w,\chi_{n})\big{|}=&\big{|}(1-\widetilde{\chi}_{n}(2)2^{-w})L(w,\widetilde{\chi}_{n})\big{|}.\end{split}

It follows from (3.2) and the above that,

(3.3) A(s,w)|ζ(2)(2s)|(n,2)=1|(1χ~n(2)2w)L(w,χ~n)||nsL(2)(2s+w,χn)|.\displaystyle\begin{split}A(s,w)\ll&|\zeta^{(2)}(2s)|\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(n,2)=1\end{subarray}}\frac{|(1-\widetilde{\chi}_{n}(2)2^{-w})L(w,\widetilde{\chi}_{n})|}{|n^{s}L^{(2)}(2s+w,\chi_{n})|}.\end{split}

Now (2.10) and partial summation implies that, except for a simple pole at w=1w=1, both sums of the right-hand side expression in (3.3) are convergent for (s)>1\Re(s)>1, (w)1/2\Re(w)\geq 1/2 as well as for (2s)>1\Re(2s)>1, (2s+w)>1\Re(2s+w)>1, (s+w)>3/2\Re(s+w)>3/2, (w)<1/2\Re(w)<1/2. Hence, A(s,w)A(s,w) converges absolutely in the region

S0={(s,w):(s)>1,(2s+w)>1,(s+w)>3/2}.S_{0}=\{(s,w):\Re(s)>1,\ \Re(2s+w)>1,\ \Re(s+w)>3/2\}.

As the condition (2s+w)>1\Re(2s+w)>1 is contained in the other conditions, the description of S0S_{0} simplifies to

S0={(s,w):(s)>1,(s+w)>3/2}.S_{0}=\{(s,w):\Re(s)>1,\ \Re(s+w)>3/2\}.

Next, upon writing m=m0m12m=m_{0}m^{2}_{1} with m0m_{0} odd and square-free, we recast the last expression of (3.1) as

(3.4) A(s,w)=\displaystyle A(s,w)= (m,2)=1L(s,χ(4m))mw=(m1,2)=11m12w(m0,2)=1L(s,χ(4m0))p|m1(1χ(4m0)(p)ps)m0w.\displaystyle\sum_{\begin{subarray}{c}(m,2)=1\end{subarray}}\frac{L(s,\chi^{(4m)})}{m^{w}}=\sum_{\begin{subarray}{c}(m_{1},2)=1\end{subarray}}\frac{1}{m_{1}^{2w}}\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(m_{0},2)=1\end{subarray}}\frac{L(s,\chi^{(4m_{0})})\prod_{p|m_{1}}(1-\chi^{(4m_{0})}(p)p^{-s})}{m_{0}^{w}}.

Note that χ(4m0)\chi^{(4m_{0})} is a primitive character modulo 4m04m_{0} for m01(mod4)m_{0}\equiv-1\pmod{4}. Arguing as above by making use of (2.10) and partial summation again, the sum over mm such that m01(mod4)m_{0}\equiv-1\pmod{4} in (3.4) converges absolutely in the region

S1=\displaystyle S_{1}= {(s,w):(w)>1,(s+w)32,(2s+w)>32}.\displaystyle\{(s,w):\hbox{$\Re(w)>1,\ \Re(s+w)\geq\tfrac{3}{2},\ \Re(2s+w)>\tfrac{3}{2}$}\}.

Note that in the case m01(mod4)m_{0}\equiv-1\pmod{4}, mm is never a square.

Similarly, the sum over mm with m01(mod4)m_{0}\equiv 1\pmod{4} in (3.4) also converges absolutely. In this case, χ(m0)\chi^{(m_{0})} is a primitive character modulo m0m_{0}. This allows us to deduce that the function A(s,w)A(s,w) converges absolutely in the region S1S_{1}, except for a simple pole at s=1s=1 arising from the summands with m=m=\square.

Notice that the convex hull of S0S_{0} and S1S_{1} is

(3.5) S2={(s,w):(s+w)>32,(2s+w)>32}.S_{2}=\{(s,w):\ \Re(s+w)>\tfrac{3}{2},\ \Re(2s+w)>\tfrac{3}{2}\}.

Hence, Theorem 2.10 implies that (s1)(w1)A(s,w)(s-1)(w-1)A(s,w) converges absolutely in the region S2S_{2}.

3.2. Residue of A(s,w)A(s,w) at s=1s=1

We see that A(s,w)A(s,w) has a pole at s=1s=1 arising from the terms with m=m=\square from (3.4). In order to compute the corresponding residue and for later use, we define the sum

A1(s,w):=(m,2)=1m=L(s,χ(4m))mw=(m,2)=1m=ζ(s)p|2m(1ps)mw.\displaystyle\begin{split}A_{1}(s,w):=\sum_{\begin{subarray}{c}(m,2)=1\\ m=\square\end{subarray}}\frac{L\left(s,\chi^{(4m)}\right)}{m^{w}}=\sum_{\begin{subarray}{c}(m,2)=1\\ m=\square\end{subarray}}\frac{\zeta(s)\prod_{p|2m}(1-p^{-s})}{m^{w}}.\end{split}

For any tt\in\mathbb{C}, let at(n)a_{t}(n) be the multiplicative function such that at(pk)=11/pta_{t}(p^{k})=1-1/p^{t} for any prime pp. This notation renders

A1(s,w)=ζ(2)(s)(m,2)=1m=as(m)mw.\displaystyle\begin{split}A_{1}(s,w)=\zeta^{(2)}(s)\sum_{\begin{subarray}{c}(m,2)=1\\ m=\square\end{subarray}}\frac{a_{s}(m)}{m^{w}}.\end{split}

Recasting the last sum above as an Euler product,

(3.6) A1(s,w)=ζ(2)(s)p>2m0m evenas(pm)pmw=ζ(2)(s)p>2(1+(11ps)1p2w(1p2w)1)=ζ(2)(s)ζ(2)(2w)p>2(11ps+2w)=ζ(s)ζ(2w)P(s,w),\displaystyle\begin{split}A_{1}(s,w)=&\zeta^{(2)}(s)\prod_{p>2}\sum_{\begin{subarray}{c}m\geq 0\\ m\text{ even}\end{subarray}}\frac{a_{s}(p^{m})}{p^{mw}}=\zeta^{(2)}(s)\prod_{p>2}\left(1+\left(1-\frac{1}{p^{s}}\right)\frac{1}{p^{2w}}(1-p^{-2w})^{-1}\right)\\ =&\zeta^{(2)}(s)\zeta^{(2)}(2w)\prod_{p>2}\left(1-\frac{1}{p^{s+2w}}\right)=\ \zeta(s)\zeta(2w)P(s,w),\end{split}

where

(3.7) P(s,w)=(112s)(1122w)p>2(11ps+2w).\displaystyle\begin{split}P(s,w)=&\left(1-\frac{1}{2^{s}}\right)\left(1-\frac{1}{2^{2w}}\right)\prod_{p>2}\left(1-\frac{1}{p^{s+2w}}\right).\end{split}

It follows from (3.6) and (3.7) that except for a simple pole at s=1s=1, the functions P(s,w)P(s,w) and A1(s,w)A_{1}(s,w) are holomorphic in the region

(3.8) S3={(s,w):\displaystyle S_{3}=\Big{\{}(s,w):\ (s+2w)>1,(w)>12}.\displaystyle\Re(s+2w)>1,\Re(w)>\tfrac{1}{2}\Big{\}}.

Note that it may appear from (3.6) that A1(s,w)A_{1}(s,w) has a pole when w=1/2w=1/2. However, as in the region S3S_{3} we have (w)>1/2\Re(w)>1/2, the function (s1)(w1)(s+w3/2)A(s,w)(s-1)(w-1)(s+w-3/2)A(s,w) is holomorphic in S3S_{3}.

As the residue of ζ(s)\zeta(s) at s=1s=1 equals 11, we deduce that

(3.9) Ress=1\displaystyle\mathrm{Res}_{s=1} A(s,12+α)=Ress=1A1(s,12+α)=ζ(1+2α)P(1,12+α).\displaystyle A(s,\tfrac{1}{2}+\alpha)=\mathrm{Res}_{s=1}A_{1}(s,\tfrac{1}{2}+\alpha)=\zeta(1+2\alpha)P(1,\tfrac{1}{2}+\alpha).

3.3. Second region of absolute convergence of A(s,w)A(s,w)

We infer from (3.4) that

(3.10) A(s,w)=(m,2)=1m=L(s,χ(4m))mw+(m,2)=1mL(s,χ(4m))mw=(m,2)=1m=ζ(s)p|2m(1ps)mw+(m,2)=1mL(s,χ(4m))mw:=A1(s,w)+A2(s,w).\displaystyle\begin{split}A(s,w)=&\sum_{\begin{subarray}{c}(m,2)=1\\ m=\square\end{subarray}}\frac{L(s,\chi^{(4m)})}{m^{w}}+\sum_{\begin{subarray}{c}(m,2)=1\\ m\neq\square\end{subarray}}\frac{L(s,\chi^{(4m)})}{m^{w}}\\ =&\sum_{\begin{subarray}{c}(m,2)=1\\ m=\square\end{subarray}}\frac{\zeta(s)\prod_{p|2m}(1-p^{-s})}{m^{w}}+\sum_{\begin{subarray}{c}(m,2)=1\\ m\neq\square\end{subarray}}\frac{L(s,\chi^{(4m)})}{m^{w}}:=\ A_{1}(s,w)+A_{2}(s,w).\end{split}

Here we recall from our discussions in the previous section that A1(s,w)A_{1}(s,w) is holomorphic in the region S3S_{3}, except for a simple pole at s=1s=1.

Next, observe that χ(4m)\chi^{(4m)} is a Dirichlet character modulo 4m4m for any m1m\geq 1 such that χ(4m)(1)=1\chi^{(4m)}(-1)=1. We thus apply the functional equation (2.1) given in Lemma 2.5 for L(s,χ(4m))L\left(s,\chi^{(4m)}\right) in the case mm\neq\square, arriving at

(3.11) A2(s,w)=πs1/24sΓ(1s2)Γ(s2)C(1s,s+w),\displaystyle\begin{split}A_{2}(s,w)=\frac{\pi^{s-1/2}}{4^{s}}\frac{\Gamma(\frac{1-s}{2})}{\Gamma(\frac{s}{2})}C(1-s,s+w),\end{split}

where C(s,w)C(s,w) is given by the double Dirichlet series

C(s,w)=\displaystyle C(s,w)= q,m(m,2)=1mτ(χ(4m),q)qsmw=q,m(m,2)=1τ(χ(4m),q)qsmwq,m(m,2)=1m=τ(χ(4m),q)qsmw.\displaystyle\sum_{\begin{subarray}{c}q,m\\ (m,2)=1\\ m\neq\square\end{subarray}}\frac{\tau(\chi^{(4m)},q)}{q^{s}m^{w}}=\sum_{\begin{subarray}{c}q,m\\ \begin{subarray}{c}(m,2)=1\end{subarray}\end{subarray}}\frac{\tau(\chi^{(4m)},q)}{q^{s}m^{w}}-\sum_{\begin{subarray}{c}q,m\\ (m,2)=1\\ m=\square\end{subarray}}\frac{\tau(\chi^{(4m)},q)}{q^{s}m^{w}}.

Note that C(s,w)C(s,w) is initially convergent for (s)\Re(s), (w)\Re(w) large enough by (3.5), (3.8) and the functional equation (3.11). To extend this region, we recast C(s,w)C(s,w) as

(3.12) C(s,w)=q=11qs(m,2)=1τ(χ(4m),q)mwq=11qs(m,2)=1m=τ(χ(4m),q)mw:=C1(s,w)C2(s,w).\displaystyle\begin{split}C(s,w)=&\sum^{\infty}_{q=1}\frac{1}{q^{s}}\sum_{\begin{subarray}{c}(m,2)=1\end{subarray}}\frac{\tau\left(\chi^{(4m)},q\right)}{m^{w}}-\sum^{\infty}_{q=1}\frac{1}{q^{s}}\sum_{\begin{subarray}{c}(m,2)=1\\ m=\square\end{subarray}}\frac{\tau\left(\chi^{(4m)},q\right)}{m^{w}}:=\ C_{1}(s,w)-C_{2}(s,w).\end{split}

For two Dirichlet characters ψ,ψ\psi,\psi^{\prime} whose conductors divide 88, we define

(3.13) C1(s,w;ψ,ψ):=l,q1G(χl,q)ψ(l)ψ(q)lwqsandC2(s,w;ψ,ψ):=l,q1G(χl2,q)ψ(l)ψ(q)l2wqs.\displaystyle\begin{split}C_{1}(s,w;\psi,\psi^{\prime}):=\sum_{l,q\geq 1}\frac{G\left(\chi_{l},q\right)\psi(l)\psi^{\prime}(q)}{l^{w}q^{s}}\quad\mbox{and}\quad C_{2}(s,w;\psi,\psi^{\prime}):=\sum_{l,q\geq 1}\frac{G\left(\chi_{l^{2}},q\right)\psi(l)\psi^{\prime}(q)}{l^{2w}q^{s}}.\end{split}

We follow the arguments contained in [Cech1, §6.4] and apply Lemma 2.2 to obtain that

(3.14) C1(s,w)=2s(C1(s,w;ψ2,ψ1)+C1(s,w;ψ2,ψ1))+4s(C1(s,w;ψ1,ψ0)+C1(s,w;ψ1,ψ0))+C1(s,w;ψ1,ψ1)C1(s,w;ψ1,ψ1),C2(s,w)=21sC2(s,w;ψ1,ψ1)+212sC2(s,w;ψ1,ψ0).\displaystyle\begin{split}C_{1}(s,w)=&-2^{-s}\big{(}C_{1}(s,w;\psi_{2},\psi_{1})+C_{1}(s,w;\psi_{-2},\psi_{1})\big{)}+4^{-s}\big{(}C_{1}(s,w;\psi_{1},\psi_{0})+C_{1}(s,w;\psi_{-1},\psi_{0})\big{)}\\ &\hskip 56.9055pt+C_{1}(s,w;\psi_{1},\psi_{-1})-C_{1}(s,w;\psi_{-1},\psi_{-1}),\\ C_{2}(s,w)=&-2^{1-s}C_{2}(s,w;\psi_{1},\psi_{1})+2^{1-2s}C_{2}(s,w;\psi_{1},\psi_{0}).\end{split}

We now follow the approach by K. Soundararajan and M. P. Young in [S&Y, §3.3] to write every integer q1q\geq 1 uniquely as q=q1q22q=q_{1}q^{2}_{2} with q1q_{1} square-free to derive that

(3.15) Ci(s,w;ψ,ψ)=q1ψ(q1)q1sDi(s,w;q1,ψ,ψ),i=1,2,C_{i}(s,w;\psi,\psi^{\prime})=\sideset{}{{}^{*}}{\sum}_{q_{1}}\frac{\psi^{\prime}(q_{1})}{q_{1}^{s}}\cdot D_{i}(s,w;q_{1},\psi,\psi^{\prime}),\quad i=1,2,

where

(3.16) D1(s,w;q1,ψ,ψ)=l,q2=1G(χl,q1q22)ψ(l)ψ(q22)lwq22sandD2(s,w;q1,ψ,ψ)=l,q2=1G(χl2,q1q22)ψ(l)ψ(q22)l2wq22s.\displaystyle\begin{split}D_{1}(s,w;q_{1},\psi,\psi^{\prime})=\sum_{l,q_{2}=1}^{\infty}\frac{G\left(\chi_{l},q_{1}q^{2}_{2}\right)\psi(l)\psi^{\prime}(q^{2}_{2})}{l^{w}q^{2s}_{2}}\quad\mbox{and}\quad D_{2}(s,w;q_{1},\psi,\psi^{\prime})=\sum_{l,q_{2}=1}^{\infty}\frac{G\left(\chi_{l^{2}},q_{1}q^{2}_{2}\right)\psi(l)\psi^{\prime}(q^{2}_{2})}{l^{2w}q^{2s}_{2}}.\end{split}

The following result develops the required analytic properties of Di(s,w;q1,ψ,ψ)D_{i}(s,w;q_{1},\psi,\psi^{\prime}).

Lemma 3.4.

With the notation as above and assuming the truth of GRH, for ψψ0\psi\neq\psi_{0}, the functions Di(s,w;q1,ψ,ψ)D_{i}(s,w;q_{1},\psi,\psi^{\prime}), with i=1i=1, 22 have meromorphic continuations to the region

(3.17) {(s,w):(s)>1,(w)>34}.\displaystyle\{(s,w):\Re(s)>1,\ \Re(w)>\tfrac{3}{4}\}.

Moreover, the only poles in this region occur for D1(s,w;q1,ψ,ψ)D_{1}(s,w;q_{1},\psi,\psi^{\prime}) at w=3/2w=3/2 when q1=1,ψ=ψ1q_{1}=1,\psi=\psi_{1} or when q1=2,ψ=ψ2q_{1}=2,\psi=\psi_{2} and the corresponding pole is simple in either case. For (s)1+ε\Re(s)\geq 1+\varepsilon, (w)3/4+ε\Re(w)\geq 3/4+\varepsilon, away from the possible poles, we have

(3.18) |Di(s,w;q1,ψ,ψ)|(q1(1+|w|))max{(3/2(w))/2,0}+ε.\displaystyle|D_{i}(s,w;q_{1},\psi,\psi^{\prime})|\ll(q_{1}(1+|w|))^{\max\{(3/2-\Re(w))/2,0\}+\varepsilon}.
Proof.

We focus on D1(s,w;q1,ψ,ψ)D_{1}(s,w;q_{1},\psi,\psi^{\prime}) here since the proof is similar for D2(s,w;q1,ψ,ψ)D_{2}(s,w;q_{1},\psi,\psi^{\prime}) which gives no pole due to the exponent of 2w2w of ll in its defintion. By Lemma 2.3, in the double sum in (3.16) defining D1(s,w;q1,ψ,ψ)D_{1}(s,w;q_{1},\psi,\psi^{\prime}), the summands there are jointly multiplicative functions of l,q2l,q_{2}(here we say S(l,q)S(l,q) is jointly multiplicative in ll and qq if S(l1l2,q1q2)=S(l1,q1)S(l2,q2)S(l_{1}l_{2},q_{1}q_{2})=S(l_{1},q_{1})S(l_{2},q_{2}) for (l1,l2)=(q1,q2)=(l1,q2)=(l2,q1)=1(l_{1},l_{2})=(q_{1},q_{2})=(l_{1},q_{2})=(l_{2},q_{1})=1). Moreover, we may assume that ll is odd in (3.16) since ψψ0\psi\neq\psi_{0}. These observations enable us to recast D1(s,w;q1,ψ,ψ)D_{1}(s,w;q_{1},\psi,\psi^{\prime}) as an Euler product such that

(3.19) D1(s,w;q1,ψ,ψ)=pD1,p(s,w;q1,ψ,ψ),\displaystyle\begin{split}&D_{1}(s,w;q_{1},\psi,\psi^{\prime})=\prod_{p}D_{1,p}(s,w;q_{1},\psi,\psi^{\prime}),\end{split}

where

(3.20) D1,p(s,w;q1,ψ,ψ)={k=0ψ(22k)22ks,p=2,l,k=0ψ(pl)ψ(p2k)G(χpl,q1p2k)plw+2ks,p>2.\displaystyle\begin{split}&D_{1,p}(s,w;q_{1},\psi,\psi^{\prime})=\displaystyle\begin{cases}\displaystyle\sum_{k=0}^{\infty}\frac{\psi^{\prime}(2^{2k})}{2^{2ks}},\quad p=2,\\ \displaystyle\sum_{l,k=0}^{\infty}\frac{\psi(p^{l})\psi^{\prime}(p^{2k})G\left(\chi_{p^{l}},q_{1}p^{2k}\right)}{p^{lw+2ks}},\quad p>2.\end{cases}\end{split}

Now, for any fixed p>2p>2,

(3.21) l,k=0ψ(pl)ψ(p2k)G(χpl,q1p2k)plw+2ks=l=0ψ(pl)G(χpl,q1)plw+l0,k1ψ(pl)ψ(p2k)G(χpl,q1p2k)plw+2ks.\displaystyle\begin{split}&\sum_{l,k=0}^{\infty}\frac{\psi(p^{l})\psi^{\prime}(p^{2k})G\left(\chi_{p^{l}},q_{1}p^{2k}\right)}{p^{lw+2ks}}=\sum_{l=0}^{\infty}\frac{\psi(p^{l})G\left(\chi_{p^{l}},q_{1}\right)}{p^{lw}}+\sum_{l\geq 0,k\geq 1}\frac{\psi(p^{l})\psi^{\prime}(p^{2k})G\left(\chi_{p^{l}},q_{1}p^{2k}\right)}{p^{lw+2ks}}.\end{split}

Remembering that q1q_{1} is square-free, we deduce from Lemma 2.3 that

|G(χpl,q1p2k)|pl,G(χpl,q1p2k)=0,l2k+3.\displaystyle\begin{split}|G(\chi_{p^{l}},q_{1}p^{2k})|\ll p^{l},\quad G(\chi_{p^{l}},q_{1}p^{2k})=0,\quad l\geq 2k+3.\end{split}

The above estimations allow us to see that when (s)>1\Re(s)>1, (w)>3/4\Re(w)>3/4,

(3.22) l0,k1ψ(pl)ψ(p2k)G(χpl,q1p2k)plw+2ks=k1ψ(p2k)G(χ1,q1p2k)p2ks+l,k1ψ(pl)ψ(p2k)G(χpl,q1p2k)plw+2ksp2(s)+|k=11l2k+21pl(w1)+2ks|p2(s)+|k=12k+2p2ks(1pw1+1p(2k+2)(w1))|p2(s)+p2(s)(w)+1+p2(s)4(w)+4.\displaystyle\begin{split}\sum_{l\geq 0,k\geq 1}\frac{\psi(p^{l})\psi^{\prime}(p^{2k})G\left(\chi_{p^{l}},q_{1}p^{2k}\right)}{p^{lw+2ks}}=&\sum_{k\geq 1}\frac{\psi^{\prime}(p^{2k})G\left(\chi_{1},q_{1}p^{2k}\right)}{p^{2ks}}+\sum_{l,k\geq 1}\frac{\psi(p^{l})\psi^{\prime}(p^{2k})G\left(\chi_{p^{l}},q_{1}p^{2k}\right)}{p^{lw+2ks}}\\ \ll&p^{-2\Re(s)}+\Bigg{|}\sum^{\infty}_{k=1}\sum_{1\leq l\leq 2k+2}\frac{1}{p^{l(w-1)+2ks}}\Bigg{|}\\ \ll&p^{-2\Re(s)}+\Bigg{|}\sum^{\infty}_{k=1}\frac{2k+2}{p^{2ks}}\Big{(}\frac{1}{p^{w-1}}+\frac{1}{p^{(2k+2)(w-1)}}\Big{)}\Bigg{|}\\ \ll&p^{-2\Re(s)}+p^{-2\Re(s)-\Re(w)+1}+p^{-2\Re(s)-4\Re(w)+4}.\end{split}

Further applying Lemma 2.3 yields that if p2q1p\nmid 2q_{1} and (w)>3/4\Re(w)>3/4,

(3.23) l=0ψ(pl)G(χpl,q1)plw=1+ψ(p)χ(q1)(p)pw1/2=Lp(w12,χ(q1)ψ)(11p2w1)=Lp(w12,χ(q1)ψ)ζp(2w1).\displaystyle\begin{split}&\sum_{l=0}^{\infty}\frac{\psi(p^{l})G\left(\chi_{p^{l}},q_{1}\right)}{p^{lw}}=1+\frac{\psi(p)\chi^{(q_{1})}(p)}{p^{w-1/2}}=L_{p}\left(w-\tfrac{1}{2},\chi^{(q_{1})}\psi\right)\left(1-\frac{1}{p^{2w-1}}\right)=\frac{L_{p}\left(w-\tfrac{1}{2},\chi^{(q_{1})}\psi\right)}{\zeta_{p}(2w-1)}.\end{split}

We derive from (3.20)–(3.23) that for p2q1p\nmid 2q_{1}, (s)>1\Re(s)>1, (w)>3/4\Re(w)>3/4,

(3.24) D1,p(s,w;q1,ψ,ψ)=Lp(w12,χ(q1)ψ)ζp(2w1)(1+O(p2(s)+p2(s)(w)+1+p2(s)4(w)+4)).\displaystyle\begin{split}&D_{1,p}(s,w;q_{1},\psi,\psi^{\prime})=\frac{L_{p}\left(w-\tfrac{1}{2},\chi^{(q_{1})}\psi\right)}{\zeta_{p}(2w-1)}\left(1+O\Big{(}p^{-2\Re(s)}+p^{-2\Re(s)-\Re(w)+1}+p^{-2\Re(s)-4\Re(w)+4}\Big{)}\right).\end{split}

Now, we deduce the first assertion of the lemma from (3.19), (3.20) and the above. We also see this way that the only poles in the region given in (3.17) are at w=3/2w=3/2 and this occurs when q1=1q_{1}=1, ψ=ψ1\psi=\psi_{1} or when q1=2,ψ=ψ2q_{1}=2,\psi=\psi_{2}. In either case, the corresponding pole is simple.

We further note that Lemma 2.3 implies that when p|q1,p2p|q_{1},p\neq 2,

(3.25) l=0ψ(pl)G(χpl,q1)plw=1ψ(p2)p2w1=1+O(p2(w)+1).\displaystyle\begin{split}&\sum_{l=0}^{\infty}\frac{\psi(p^{l})G\left(\chi_{p^{l}},q_{1}\right)}{p^{lw}}=1-\frac{\psi(p^{2})}{p^{2w-1}}=1+O(p^{-2\Re(w)+1}).\end{split}

It follows from (3.20), (3.22) and (3.25) that for p|q1,p2p|q_{1},p\neq 2, (s)>1,(w)>3/4\Re(s)>1,\Re(w)>3/4,

(3.26) D1,p(s,w;q1,ψ,ψ)=1+O(p2(w)+1+p2(s)+p2(s)(w)+1+p2(s)4(w)+4).\displaystyle\begin{split}&D_{1,p}(s,w;q_{1},\psi,\psi^{\prime})=1+O\Big{(}p^{-2\Re(w)+1}+p^{-2\Re(s)}+p^{-2\Re(s)-\Re(w)+1}+p^{-2\Re(s)-4\Re(w)+4}\Big{)}.\end{split}

We conclude from (3.19), (3.20), (3.24) and (3.26) that for (s)1+ε\Re(s)\geq 1+\varepsilon and (w)3/4+ε\Re(w)\geq 3/4+\varepsilon,

D1(s,w;q1,ψ,ψ)q1ε|L(2q1)(w12,χ(q1)ψ)ζ(2q1)(2w1)|q1ε|L(w12,χ(q1)ψ)ζ(2w1)|(q1(1+|w|))max{(3/2(w))/2,0}+ε,\displaystyle\begin{split}D_{1}(s,w;q_{1},\psi,\psi^{\prime})\ll&q_{1}^{\varepsilon}\Big{|}\frac{L^{(2q_{1})}\left(w-\tfrac{1}{2},\chi^{(q_{1})}\psi\right)}{\zeta^{(2q_{1})}(2w-1)}\Big{|}\ll q_{1}^{\varepsilon}\Big{|}\frac{L\left(w-\tfrac{1}{2},\chi^{(q_{1})}\psi\right)}{\zeta(2w-1)}\Big{|}\ll(q_{1}(1+|w|))^{\max\{(3/2-\Re(w))/2,0\}+\varepsilon},\end{split}

where the last bound follows from (2.2) (by taking ψ^=ψ0\widehat{\psi}=\psi_{0} to be the primitive principal character) and (2.9). This leads to the estimate in (3.18) and completes the proof of the lemma. ∎

We remark here that the use of the Lindelöf hypothesis, a consequence of GRH, in the proof of Lemma 3.4 will lead to the improvement of the exponent of (1+|α|)(1+|\alpha|) in the OO-term of (1.2) to 1/4+ε1/4+\varepsilon. We keep our computations as unconditional as possible, so that one can deduce an unconditional version of our main result from the arguments.

Now applying Lemma 3.4 with (3.14) and (3.15), we infer that (w3/2)C(s,w)(w-3/2)C(s,w) is defined in the region

{(s,w):(s)>1,(w)>3/4,(s+w/2)>7/4}.\{(s,w):\ \Re(s)>1,\ \Re(w)>3/4,\ \Re(s+w/2)>7/4\}.

The above together with (3.8) and (3.10) now implies that (s1)(w1)(s+w3/2)A(s,w)(s-1)(w-1)(s+w-3/2)A(s,w) can be extended to the region

(3.27) S4=\displaystyle S_{4}= {(s,w):(s+2w)>1,(s+w)>3/4,(ws)>3/2,(s)<0}.\displaystyle\{(s,w):\ \Re(s+2w)>1,\ \Re(s+w)>3/4,\ \Re(w-s)>3/2,\ \Re(s)<0\}.

Note that the condition (w+s)>3/4\Re(w+s)>3/4 and (s)<0\Re(s)<0 imply that (w)>3/4\Re(w)>3/4 in S4S_{4}, so that we have S4S3S_{4}\subset S_{3}. Additionally, the condition (s+2w)>1\Re(s+2w)>1 is redundant. So

S4={(s,w):(s+w)>3/4,(ws)>3/2,(s)<0}.S_{4}=\{(s,w):\ \Re(s+w)>3/4,\ \Re(w-s)>3/2,\ \Re(s)<0\}.

We remark here that it is possible to repeat the argument of [Cech1, Section 6.4] to study the analytical property of C(s,w)C(s,w) without relying on the joint multiplicativity of G(χl,q1q2)G(\chi_{l},q_{1}q_{2}) in ll and q2q_{2}. However, the analogue of S4S_{4} obtained this way would be more complicated to describe than what we have above.

As noted below (3.8), the function (s1)(w1)(s+w3/2)A(s,w)(s-1)(w-1)(s+w-3/2)A(s,w) is holomorphic in S3S_{3} and hence in S4S_{4}. Similarly, (s1)(w1)(s+w3/2)A(s,w)(s-1)(w-1)(s+w-3/2)A(s,w) is holomorphic in S2S_{2}. Then Theorem 2.10 implies that (s1)(w1)(s+w3/2)A(s,w)(s-1)(w-1)(s+w-3/2)A(s,w) is holomorphic in the convex hull of S2S4S_{2}\cup S_{4}. We denoted this convex hull by S5S_{5} and

S5={(s,w):(s+w)>3/4}.S_{5}=\{(s,w):\ \Re(s+w)>3/4\}.

Now Theorem 2.10 renders that (s1)(w1)(s+w3/2)A(s,w)(s-1)(w-1)(s+w-3/2)A(s,w) admits analytic continuation to the region S5S_{5}.

3.5. Residue of A(s,w)A(s,w) at s=3/2ws=3/2-w

We keep the notation from the proof of Lemma 3.4. We deduce from (3.12), (3.14), (3.15) and Lemma 3.4, that C(s,w)C(s,w) has a pole at w=3/2w=3/2 and

(3.28) Resw=3/2C(s,w)=\displaystyle\mathrm{Res}_{w=3/2}C(s,w)= 4sResw=3/2D1(s,w;1,ψ1,ψ0)+Resw=3/2D1(s,w;1,ψ1,ψ1).\displaystyle 4^{-s}\mathrm{Res}_{w=3/2}D_{1}(s,w;1,\psi_{1},\psi_{0})+\mathrm{Res}_{w=3/2}D_{1}(s,w;1,\psi_{1},\psi_{-1}).

Here we remark that the pole at w=3/2w=3/2 for D1(s,w;q1,ψ,ψ)D_{1}(s,w;q_{1},\psi,\psi^{\prime}) when q1=2,ψ=ψ2q_{1}=2,\psi=\psi_{2} does not lead to any pole in C(s,w)C(s,w) since this only affects C1(s,w;ψ2,ψ1)C_{1}(s,w;\psi_{2},\psi_{1}) and then we have ψ1(q1)=0\psi_{1}(q_{1})=0 when q1=2q_{1}=2.

We apply Lemma 2.3 to see that for p2p\neq 2,

(3.29) l0,k1ψ(pk)G(χpl,p2k)p3l/2+2ks=k1G(χ1,p2k)p2ks+l,k1G(χpl,p2k)p3l/2+2ks=p2s(1p2s)1+k11p2ks(l=1kφ(p2l)p3l+p2kpp3(2k+1)/2)=p2s(1p2s)1+1pk11p2ks=(1+1p)p2s(1p2s)1.\displaystyle\begin{split}\sum_{l\geq 0,k\geq 1}\frac{\psi^{\prime}(p^{k})G\left(\chi_{p^{l}},p^{2k}\right)}{p^{3l/2+2ks}}=&\sum_{k\geq 1}\frac{G\left(\chi_{1},p^{2k}\right)}{p^{2ks}}+\sum_{l,k\geq 1}\frac{G\left(\chi_{p^{l}},p^{2k}\right)}{p^{3l/2+2ks}}\\ =&p^{-2s}(1-p^{-2s})^{-1}+\sum_{k\geq 1}\frac{1}{p^{2ks}}\Big{(}\sum^{k}_{l=1}\frac{\varphi(p^{2l})}{p^{3l}}+\frac{p^{2k}\sqrt{p}}{p^{3(2k+1)/2}}\Big{)}\\ =&p^{-2s}(1-p^{-2s})^{-1}+\frac{1}{p}\sum_{k\geq 1}\frac{1}{p^{2ks}}=\Big{(}1+\frac{1}{p}\Big{)}p^{-2s}(1-p^{-2s})^{-1}.\end{split}

We derive from (3.20), (3.21) and (3.23) that for p2p\neq 2,

(3.30) D1,p(s,w;1,ψ1,ψ0)=D1,p(s,w;1,ψ1,ψ1)=1+1pw1/2+l0,k1ψ(pk)G(χpl,p2k)plw+2ks=ζp(w1/2)Qp(s,w),\displaystyle\begin{split}D_{1,p}(s,w;1,\psi_{1},\psi_{0})=D_{1,p}(s,w;1,\psi_{1},\psi_{-1})=&1+\frac{1}{p^{w-1/2}}+\sum_{l\geq 0,k\geq 1}\frac{\psi^{\prime}(p^{k})G\left(\chi_{p^{l}},p^{2k}\right)}{p^{lw+2ks}}=\zeta_{p}(w-1/2)Q_{p}(s,w),\end{split}

where, using (3.29),

(3.31) Qp(s,w)|w=3/2=(11p2)(1p2s)1.\displaystyle\begin{split}Q_{p}(s,w)\Big{|}_{w=3/2}=&\Big{(}1-\frac{1}{p^{2}}\Big{)}(1-p^{-2s})^{-1}.\end{split}

It follows from (3.19), (3.20), (3.30) and (3.31) that we have

(3.32) D1(s,w;1,ψ1,ψ0)=ζ(w1/2)Q(s,w),D1(s,w;1,ψ1,ψ1)=(122s)ζ(w1/2)Q(s,w),\displaystyle\begin{split}D_{1}(s,w;1,\psi_{1},\psi_{0})=\zeta(w-1/2)Q(s,w),\quad D_{1}(s,w;1,\psi_{1},\psi_{-1})=(1-2^{-2s})\zeta(w-1/2)Q(s,w),\end{split}

with

(3.33) Q(s,w)|w=3/2=2ζ(2s)3ζ(2).\displaystyle\begin{split}Q(s,w)\Big{|}_{w=3/2}=\frac{2\zeta(2s)}{3\zeta(2)}.\end{split}

As the residue of ζ(s)\zeta(s) at s=1s=1 equals 11, we deduce from (3.28), (3.32) and (3.33) that

Resw=3/2C(s,w)=\displaystyle\mathrm{Res}_{w=3/2}C(s,w)= 23ζ(2s)ζ(2).\displaystyle\frac{2}{3}\frac{\zeta(2s)}{\zeta(2)}.

Now (3.10), the functional equation (3.11) and the above lead to

Ress=3/2wA(s,w)=Ress=3/2wA2(s,w)=2π1w343/2wΓ(w1/22)Γ(3/2w2)ζ(2w1)ζ(2).\displaystyle\begin{split}\mathrm{Res}_{s=3/2-w}A(s,w)=\mathrm{Res}_{s=3/2-w}A_{2}(s,w)=\frac{2\cdot\pi^{1-w}}{3\cdot 4^{3/2-w}}\frac{\Gamma(\frac{w-1/2}{2})}{\Gamma(\frac{3/2-w}{2})}\frac{\zeta(2w-1)}{\zeta(2)}.\end{split}

Setting w=1/2+αw=1/2+\alpha in the above gives

(3.34) Ress=1αA(s,12+α)=22α1π1/2αΓ(α2)3Γ(1α2)ζ(2α)ζ(2).\displaystyle\begin{split}&\mathrm{Res}_{s=1-\alpha}A(s,\tfrac{1}{2}+\alpha)=\frac{2^{2\alpha-1}\pi^{1/2-\alpha}\Gamma(\frac{\alpha}{2})}{3\Gamma(\frac{1-\alpha}{2})}\frac{\zeta(2\alpha)}{\zeta(2)}.\end{split}

Note that the functional equation (2.6) for d=1d=1 implies that

ζ(2α)=π2α1/2Γ(12α)Γ(α)ζ(12α).\displaystyle\zeta(2\alpha)=\pi^{2\alpha-1/2}\frac{\Gamma(\tfrac{1}{2}-\alpha)}{\Gamma(\alpha)}\zeta(1-2\alpha).

The above allows us to recast the expression in (3.34) as

(3.35) Ress=1αA(s,12+α)=παΓ(12α)Γ(α2)Γ(1α2)Γ(α)ζ(12α)ζ(2)22α6.\displaystyle\begin{split}&\mathrm{Res}_{s=1-\alpha}A(s,\tfrac{1}{2}+\alpha)=\frac{\pi^{\alpha}\Gamma(\tfrac{1}{2}-\alpha)\Gamma(\frac{\alpha}{2})}{\Gamma(\frac{1-\alpha}{2})\Gamma(\alpha)}\cdot\frac{\zeta(1-2\alpha)}{\zeta(2)}\cdot\frac{2^{2\alpha}}{6}.\end{split}

3.6. Bounding A(s,w)A(s,w) in vertical strips

We shall estimate |A(s,w)||A(s,w)| in vertical strips, which is necessary in the proof of Theorem 1.1.

For previously defined regions SjS_{j}, we set

S~j=Sj,δ{(s,w):(s)5/2, 2(w)},\widetilde{S}_{j}=S_{j,\delta}\cap\{(s,w):\Re(s)\geq-5/2,\ 2\geq\Re(w)\},

where δ\delta is a fixed number with 0<δ<1/10000<\delta<1/1000 and where Sj,δ={(s,w)+δ(1,1):(s,w)Sj}S_{j,\delta}=\{(s,w)+\delta(1,1):(s,w)\in S_{j}\} for j4j\neq 4 and S4,δ={(s,w)+δ(1,1):(s,w)S4}S_{4,\delta}=\{(s,w)+\delta(-1,1):(s,w)\in S_{4}\}. We further set

p(s,w)=(s1)(w1)(s+w3/2)Γ(w2).p(s,w)=(s-1)(w-1)(s+w-3/2)\Gamma(\tfrac{w}{2}).

Observe that p(s,w)A(s,w)p(s,w)A(s,w) is analytic in the regions under our consideration.

We note that the function Γ(1w2)(w1)\Gamma(\frac{1-w}{2})(w-1) is also analytic when (w)2\Re(w)\leq 2. When 1/2<(w)21/2<\Re(w)\leq 2, we apply (2.7) to see that in this case we have

(w1)Γ(w2)|Γ(1w2)(w1)(w10)2|.\displaystyle(w-1)\Gamma(\tfrac{w}{2})\ll\Big{|}\Gamma(\tfrac{1-w}{2})(w-1)(w-10)^{2}\Big{|}.

Now, (2.10) and partial summation can be used to bound A(s,w)A(s,w) via (3.3). So in S~0{(s,w):2(w)>1/2}\widetilde{S}_{0}\cap\{(s,w):2\geq\Re(w)>1/2\},

(3.36) p(s,w)A(s,w)|w1/4(1+1002w)(s+5)(w10)2(s+w1/2)Γ(1w2)(w1)|.\displaystyle\begin{split}p(s,w)A(s,w)\ll\Big{|}w^{1/4}(1+100\cdot 2^{-w})(s+5)(w-10)^{2}(s+w-1/2)\Gamma(\tfrac{1-w}{2})(w-1)\Big{|}.\end{split}

For (w)1/2\Re(w)\leq 1/2, we apply (2.6) to estimate A(s,w)A(s,w) from above by way of (3.3) to revert to the case (w)1/2\Re(w)\geq 1/2. This gives that the bound in (3.36) continues to hold in S~0{(s,w):(w)1/2}\widetilde{S}_{0}\cap\{(s,w):\Re(w)\leq 1/2\}. Thus (3.36) is valid in the entire region S~0\widetilde{S}_{0}.

Similarly, we bound the expression for A(s,w)A(s,w) given in (3.4). In S~1\widetilde{S}_{1},

p(s,w)A(s,w)|(1+1002w)(s+5)(w10)2(s+w1/2)Γ(1w2)(w1)||s+5|max{1/2(s),0}+ε.\displaystyle p(s,w)A(s,w)\ll\Big{|}(1+100\cdot 2^{-w})(s+5)(w-10)^{2}(s+w-1/2)\Gamma(\tfrac{1-w}{2})(w-1)\Big{|}\cdot\Big{|}s+5\Big{|}^{\max\{1/2-\Re(s),0\}+\varepsilon}.

From this, we apply Proposition 2.11 by taking g=p(s,w)A(s,w)g=p(s,w)A(s,w), h=(1+1002w)(s+5)5(w10)3(s+w1/2)Γ(1w2)(w1)h=(1+100\cdot 2^{-w})(s+5)^{5}(w-10)^{3}(s+w-1/2)\Gamma(\frac{1-w}{2})(w-1) here and note that h0h\neq 0 in S~0S~1\widetilde{S}_{0}\cup\widetilde{S}_{1}. We thus deduce that in the convex hull S~2\widetilde{S}_{2} of S~0\widetilde{S}_{0} and S~1\widetilde{S}_{1},

(3.37) p(s,w)A(s,w)|(1+1002w)(s+5)5(w10)2(s+w1/2)Γ(1w2)(w1)|.p(s,w)A(s,w)\ll\Big{|}(1+100\cdot 2^{-w})(s+5)^{5}(w-10)^{2}(s+w-1/2)\Gamma(\tfrac{1-w}{2})(w-1)\Big{|}.

Moreover, S~4S~3\widetilde{S}_{4}\subset\widetilde{S}_{3} and the conditions (s+w)>3/4\Re(s+w)>3/4, (s)<0\Re(s)<0 given in (3.27) for the definition of S4S_{4} imply that (w)>3/4\Re(w)>3/4 so that ζ(2w)1\zeta(2w)\ll 1 in S~4\widetilde{S}_{4}. Then (2.9) can be used to bound ζ(s)\zeta(s) (corresponding to the case with ψ=ψ0\psi=\psi_{0} being the primitive principal character). This lead to an estimate for A1(s,w)A_{1}(s,w) in (3.6). Arguing as above reveals that in the region S~4\widetilde{S}_{4}, we have

(3.38) p(s,w)A1(s,w)|(1+1002w)(s+5)5(w10)3(s+w3/2)Γ(1w2)(w1)|.\displaystyle p(s,w)A_{1}(s,w)\ll\Big{|}(1+100\cdot 2^{-w})(s+5)^{5}(w-10)^{3}(s+w-3/2)\Gamma(\tfrac{1-w}{2})(w-1)\Big{|}.

Also, we deduce from (3.12)–(3.16) and Lemma 3.4 that, under GRH,

(3.39) |C(s,w)|(1+|w|)max{(3/2(w))/2,0}+ε|C(s,w)|\ll(1+|w|)^{\max\{(3/2-\Re(w))/2,0\}+\varepsilon}

in the region

{(s,w):(s)1+ε,(w)3/4+ε}.\{(s,w):\Re(s)\geq 1+\varepsilon,\ \Re(w)\geq 3/4+\varepsilon\}.

Now applying (3.10), the functional equation (3.11) together with (2.7) to bound the ratio of the gamma functions appearing there, as well as the bounds in (3.38) and (3.39), we obtain that in the region S~4\widetilde{S}_{4}, under GRH,

(3.40) p(s,w)A(s,w)|(s+5)(w10)2(s+w1/2)Γ(1w2)(w1)|(1+|s+w|)max{(3/2(s+w))/2,0}+ε(1+|s|)3+ε|(s+5)6(w10)4(s+w1/2)Γ(1w2)(w1)|.\displaystyle\begin{split}p(s,w)A(s,w)\ll&\Big{|}(s+5)(w-10)^{2}(s+w-1/2)\Gamma(\tfrac{1-w}{2})(w-1)\Big{|}(1+|s+w|)^{\max\{(3/2-\Re(s+w))/2,0\}+\varepsilon}(1+|s|)^{3+\varepsilon}\\ \ll&\Big{|}(s+5)^{6}(w-10)^{4}(s+w-1/2)\Gamma(\tfrac{1-w}{2})(w-1)\Big{|}.\end{split}

Lastly, we conclude from (3.37), (3.40) and Proposition 2.11 that in the convex hull S~5\widetilde{S}_{5} of S~2\widetilde{S}_{2} and S~4\widetilde{S}_{4}, under GRH,

(3.41) p(s,w)A(s,w)|(1+1002w)(s+5)6(w10)4(s+w1/2)Γ(1w2)(w1)|.p(s,w)A(s,w)\ll\Big{|}(1+100\cdot 2^{-w})(s+5)^{6}(w-10)^{4}(s+w-1/2)\Gamma(\tfrac{1-w}{2})(w-1)\Big{|}.

3.7. Completing the proof

Using the Mellin inversion, we see that for the function A(s,w)A(s,w) defined in (3.1),

(3.42) (n,2)=1L(2)(12+α,χn)w(nX)=12πi(2)A(s,12+α)Xsw^(s)ds,\sum_{\begin{subarray}{c}(n,2)=1\end{subarray}}L^{(2)}(\tfrac{1}{2}+\alpha,\chi_{n})w\left(\frac{n}{X}\right)=\frac{1}{2\pi i}\int\limits_{(2)}A\left(s,\tfrac{1}{2}+\alpha\right)X^{s}\widehat{w}(s)\mathrm{d}s,

where w^\widehat{w} is the Mellin transform of ww given by

w^(s)=0w(t)tsdtt.\displaystyle\widehat{w}(s)=\int\limits^{\infty}_{0}w(t)t^{s}\frac{\mathrm{d}t}{t}.

Integration by parts renders that for any integer E0E\geq 0,

(3.43) w^(s)1(1+|s|)E.\displaystyle\widehat{w}(s)\ll\frac{1}{(1+|s|)^{E}}.

We shift the line of integration in (3.42) to (s)=1/4+ε\Re(s)=1/4+\varepsilon. The integral on the new line can be absorbed into the OO-term in (1.2) upon using (2.7), (3.41) and (3.43). We also encounter two simple poles at s=1s=1 and s=1αs=1-\alpha in the move with the corresponding residues given in (3.9) and (3.35), respectively. Direct computations now lead to the main terms given in (1.2). This completes the proof of Theorem 1.1.

Declarations. We declare that the authors have no competing interests as defined by Springer, or other interests that might be perceived to influence the results and/or discussion reported in this paper. We further declare that there is no data associated with the results in this article.

Acknowledgments. The authors are grateful to M. B. Milinovich for some helpful discussions. P. G. was supported in part by NSFC grant 11871082 and L. Z. by the FRG Grant PS43707 at the University of New South Wales. Moreover, the authors thank the anonymous referee for his/her very careful reading of the paper and many helpful comments and suggestions.

References