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

Ratios conjecture for quadratic twist of modular LL-functions

Peng Gao School of Mathematical Sciences, Beihang University, Beijing 100191, China [email protected]  and  Liangyi Zhao School of Mathematics and Statistics, University of New South Wales, Sydney NSW 2052, Australia [email protected]
Abstract.

We develop the LL-functions ratios conjecture with one shift in the numerator and denominator in certain ranges for the family of quadratic twist of modular LL-functions using multiple Dirichlet series under the generalized Riemann hypothesis.

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

Keywords: ratios conjecture, quadratic twist, modular LL-functions

1. Introduction

As a significant conjecture with many important applications, the LL-functions ratios conjecture makes predictions on the asymptotic behaviors of the sum of ratios of products of shifted LL-functions. This conjecture originated from the work of D. W. Farmer [Farmer93] on shifted moments of the Riemann zeta function, and is formulated for general LL-functions by J. B. Conrey, D. W. Farmer and M. R. Zirnbauer in [CFZ, Section 5].

There are now several results available in the literature on the ratios conjecture, all valid for certain ranges of the relevant parameters, starting with the work of H. M. Bui, A. Florea and J. P. Keating in [BFK21] over function fields for quadratic LL-functions. Utilizing the powerful tool of multiple Dirichlet series, M. Čech [Cech1] studied the case of quadratic Dirichlet LL-functions under the assumption of the generalized Riemann hypothesis (GRH). This was the first result of its type over number fields.

Following the approach of M. Čech in [Cech1], the authors studied the ratios conjecture for quadratic Hecke LL-functions over the Gaussian field (i)\mathbb{Q}(i) in [G&Zhao14]. It is the aim of this paper to further apply the method of multiple Dirichlet series to develope the ratios conjecture for quadratic twists of modular LL-functions. To state our result, we fix a holomorphic Hecke eigenform ff of weight κ\kappa of the full modular group SL2()SL_{2}(\mathbb{Z}). The Fourier expansion of ff at infinity can be written as

f(z)=n=1λf(n)n(κ1)/2e(nz),wheree(z)=exp(2πiz).\displaystyle f(z)=\sum_{n=1}^{\infty}\lambda_{f}(n)n^{(\kappa-1)/2}e(nz),\quad\mbox{where}\quad e(z)=\exp(2\pi iz).

We reserve the letter pp for a prime throughout the paper. For any Dirichlet character χ\chi modulo dd, the twisted modular LL-function L(s,fχd)L(s,f\otimes\chi_{d}) is defined for (s)>1\Re(s)>1 by

(1.1) L(s,fχ)\displaystyle L(s,f\otimes\chi) =n=1λf(n)χ(n)ns=pd(1λf(p)χ(p)ps+1p2s)1.\displaystyle=\sum_{n=1}^{\infty}\frac{\lambda_{f}(n)\chi(n)}{n^{s}}=\prod_{p\nmid d}\left(1-\frac{\lambda_{f}(p)\chi(p)}{p^{s}}+\frac{1}{p^{2s}}\right)^{-1}.

We also write L(s,sym2f)L(s,\text{sym}^{2}f) for the symmetric square LL-function of ff defined in (2.9) and note that L(s,sym2f)L(s,\text{sym}^{2}f) is holomorphic for (s)1/2\Re(s)\geq 1/2 (see the discussions given in Section 2.1).

Let χ(m)=(m)\chi^{(m)}=\left(\frac{m}{\cdot}\right) denote the Kronecker symbol defined on [iwakow, p. 52] for any integer m0,1(mod4)m\equiv 0,1\pmod{4}. Note that every such mm factors uniquely into m=dl2m=dl^{2}, where dd is a fundamental discriminant, i.e. dd is square-free, d1(mod4)d\equiv 1\pmod{4} or d=4nd=4n with nn square-free, n2,3(mod4)n\equiv 2,3\pmod{4}. Furthermore, when dd is a fundamental discriminant, the function L(s,fχ(d))L(s,f\otimes\chi^{(d)}) has an analytical continuation to the entire complex plane and satisfies the functional equation (see, for example, [S&Y])

(1.2) Λ(s,fχ(d))=:(|d|2π)sΓ(s+κ12)L(s,fχ(d))=iκϵ(d)Λ(1s,fχ(d)),\displaystyle\Lambda(s,f\otimes\chi^{(d)})=:\left(\frac{|d|}{2\pi}\right)^{s}\Gamma\left(s+\frac{\kappa-1}{2}\right)L(s,f\otimes\chi^{(d)})=i^{\kappa}\epsilon(d)\Lambda(1-s,f\otimes\chi^{(d)}),

where ϵ(d)=1\epsilon(d)=1 if d>0d>0 and ϵ(d)=1\epsilon(d)=-1 if d<0d<0.

For an odd, positive integer nn, we denote χn\chi_{n} for the quadratic character (n)\left(\frac{\cdot}{n}\right). Given 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 also write LpL_{p} for L(p)L_{(p)}. We observe that the quadratic reciprocity law implies that L(2)(s,fχn)=L(s,fχ(4n))L^{(2)}(s,f\otimes\chi_{n})=L(s,f\otimes\chi^{(4n)}) when n1(mod4)n\equiv 1\pmod{4} and that L(2)(s,fχn)=L(s,fχ(4n))L^{(2)}(s,f\otimes\chi_{n})=L(s,f\otimes\chi^{(-4n)}) when n1(mod4)n\equiv-1\pmod{4}. It follows from (1.1), (1.2) and the above discussions that L(2)(s,fχn)L^{(2)}(s,f\otimes\chi_{n}) can also be continued analytically to the entirety of \mathbb{C}.

Our result in this paper investigates the ratios conjecture with one shift in the numerator and denominator for the family of quadratic twist of modular LL-functions L(2)(s,fχn)L^{(2)}(s,f\otimes\chi_{n}) averaged over all odd, positive nn.

Theorem 1.1.

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

(1.3) (n,2)=1L(2)(12+α,fχn)L(2)(12+β,fχn)w(nX)=Xw^(1)L(2)(1+2α,sym2f)P(1,12+α,12+β;f)+O((1+|α|)ε|β|εXN(α,β)+ε),\displaystyle\begin{split}\sum_{\begin{subarray}{c}(n,2)=1\end{subarray}}\frac{L^{(2)}(\tfrac{1}{2}+\alpha,f\otimes\chi_{n})}{L^{(2)}(\tfrac{1}{2}+\beta,f\otimes\chi_{n})}w\left(\frac{n}{X}\right)=&X\widehat{w}(1)L^{(2)}(1+2\alpha,\text{sym}^{2}f)P(1,\tfrac{1}{2}+\alpha,\tfrac{1}{2}+\beta;f)+O\left((1+|\alpha|)^{\varepsilon}|\beta|^{\varepsilon}X^{N(\alpha,\beta)+\varepsilon}\right),\end{split}

where

(1.4) P(s,w,z;f)=(112s)1ζ(2)(2w)×p>2(1+1p2z(11ps)(1+1p2w)+(11ps)1p2wλf2(p)2p2w+s+1p4w+s(11ps)λf2(p)pz+w),\displaystyle\begin{split}P(s,w,z;f)=&\left(1-\frac{1}{2^{s}}\right)\frac{1}{\zeta^{(2)}(2w)}\\ &\times\prod_{p>2}\left(1+\frac{1}{p^{2z}}\left(1-\frac{1}{p^{s}}\right)\left(1+\frac{1}{p^{2w}}\right)+\left(1-\frac{1}{p^{s}}\right)\frac{1}{p^{2w}}-\frac{\lambda^{2}_{f}(p)-2}{p^{2w+s}}+\frac{1}{p^{4w+s}}-\left(1-\frac{1}{p^{s}}\right)\frac{\lambda^{2}_{f}(p)}{p^{z+w}}\right),\end{split}

and

(1.5) N(α,β)=max{12(α),12(β)}.N(\alpha,\beta)=\max\left\{1-2\Re(\alpha),1-2\Re(\beta)\right\}.

It follows from our discussions in Section 3.2 below that the value P(1,12+α,12+β;f)P(1,\tfrac{1}{2}+\alpha,\tfrac{1}{2}+\beta;f) is finite for the ranges of α\alpha and β\beta defined in the statement of Theorem 1.1. The main term in (1.3) is consistent with the ratios conjecture, which can be derived following the treatments given in [G&Zhao2022-2, Section 5]. However, the error term in (1.3) is inferior to the prediction of the ratios conjecture. The latter asserts that (1.3) holds uniformly for |(α)|<1/4|\Re(\alpha)|<1/4, (logX)1(β)<1/4(\log X)^{-1}\ll\Re(\beta)<1/4 and (α),(β)X1ε\Im(\alpha),\;\Im(\beta)\ll X^{1-\varepsilon} with an error term O(X1/2+ε)O(X^{1/2+\varepsilon}). Nevertheless, the advantage of our result in (1.3) is that there is no constraint on imaginary parts of α\alpha or β\beta.

Theorem 1.1 will be established using the method in the proof of [Cech1, Theorem 1.2]. In particular, we note that the functional equation of a general (not necessarily primitive) quadratic Dirichlet LL-function given in [Cech1, Proposition 2.3] plays a crucial role in our proof.

2. Preliminaries

We first include some auxiliary results.

2.1. Quadratic Gauss sums

Recall that χ(d)=(d)\chi^{(d)}=\left(\frac{d}{\cdot}\right) for the Kronecker symbol. Moreover, let ψj,j{±1,±2}\psi_{j},j\in\{\pm 1,\pm 2\} denote the quadratic characters given by ψj=χ(4j)\psi_{j}=\chi^{(4j)}. Note that ψj\psi_{j} is a primitive character modulo 4j4j for each jj. We also denote ψ0\psi_{0} the primitive principle character.

For any integer qq and any Dirichlet character χ\chi modulo nn, we define the associated Gauss sum τ(χ,q)\tau(\chi,q) by

τ(χ,q)=j(modn)χ(j)e(jqn).\tau(\chi,q)=\sum_{j\pmod{n}}\chi(j)e\left(\frac{jq}{n}\right).

The following result is quoted 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}

We further define 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}

We denote φ(m)\varphi(m) for the Euler totient function of mm. Our next result is taken from [sound1, Lemma 2.3] and evaluates G(χm,q)G\left(\chi_{m},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. Modular LL-functions

For any Dirichlet character χ\chi, the twisted modular LL-function L(s,fχ)L(s,f\otimes\chi) has an Euler product for (s)>1\Re(s)>1 given by

(2.1) L(s,fχ)\displaystyle L(s,f\otimes\chi) =pj=12(1αf(p,j)χ(p)ps)1.\displaystyle=\prod_{p}\prod^{2}_{j=1}(1-\alpha_{f}(p,j)\chi(p)p^{-s})^{-1}.

By Deligne’s proof [D] of the Weil conjecture, we know that

(2.2) |αf(p,1)|=|αf(p,2)|=1andαf(p,1)αf(p,2)=1.\displaystyle|\alpha_{f}(p,1)|=|\alpha_{f}(p,2)|=1\quad\mbox{and}\quad\alpha_{f}(p,1)\alpha_{f}(p,2)=1.

It follows from this and (1.1) that λf(n)\lambda_{f}(n) is multiplicative and for ν1\nu\geq 1, we have

(2.3) λf(pν)=j=0ναfνj(p,1)αfj(p,2).\displaystyle\begin{split}\lambda_{f}(p^{\nu})=&\sum^{\nu}_{j=0}\alpha^{\nu-j}_{f}(p,1)\alpha^{j}_{f}(p,2).\end{split}

The above relation further implies that λf(n)\lambda_{f}(n)\in\mathbb{R}, λf(1)=1\lambda_{f}(1)=1 and

(2.4) |λf(n)|d(n)nε,\displaystyle\begin{split}|\lambda_{f}(n)|\leq d(n)\ll n^{\varepsilon},\end{split}

where d(n)d(n) is the number of divisors of nn and the last estimate above follows from [MVa1, Theorem 2.11].

We also deduce from (2.1) that for (s)>1\Re(s)>1,

(2.5) L1(s,fχ)=pj=12(1αf(p,j)χ(p)ps)=:n=1cf(n)χ(n)ns.\displaystyle\begin{split}L^{-1}(s,f\otimes\chi)=\prod_{p}\prod^{2}_{j=1}(1-\alpha_{f}(p,j)\chi(p)p^{-s})=:\sum^{\infty}_{n=1}\frac{c_{f}(n)\chi(n)}{n^{s}}.\end{split}

It follows from (2.2), (2.3) and the Euler product given in (2.5) that cf(n)c_{f}(n) is a multiplicative function of nn such that

(2.6) cf(pk)={1,k=0,2,λf(p),k=1,0,k>2.\displaystyle\begin{split}c_{f}(p^{k})=\displaystyle\begin{cases}1,\quad k=0,2,\\ -\lambda_{f}(p),\quad k=1,\\ 0,\quad k>2.\end{cases}\end{split}

In particular, we deduce from (2.2) and (2.6) that we have |cf(pk)|2|c_{f}(p^{k})|\leq 2 for all k0k\geq 0, so that for any ε>0\varepsilon>0,

(2.7) cf(n)2ω(n)nε,\displaystyle\begin{split}c_{f}(n)\ll 2^{\omega(n)}\ll n^{\varepsilon},\end{split}

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

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

Recall also that the symmetric square LL-function L(s,sym2f)L(s,\operatorname{sym}^{2}f) of ff is defined for (s)>1\Re(s)>1 by (see [iwakow, p. 137] and [iwakow, (25.73)])

(2.9) L(s,sym2f)=p1ij2(1αf(p,i)αf(p,j)ps)1=p(1αf2(p,1)ps)1(11ps)1(1αf2(p,2)ps)1=ζ(2s)n=1λf(n2)ns=p(1λf(p2)ps+λf(p2)p2s1p3s)1.\displaystyle\begin{split}L(s,\operatorname{sym}^{2}f)=&\prod_{p}\prod_{1\leq i\leq j\leq 2}\left(1-\frac{\alpha_{f}(p,i)\alpha_{f}(p,j)}{p^{s}}\right)^{-1}=\prod_{p}\left(1-\frac{\alpha^{2}_{f}(p,1)}{p^{s}}\right)^{-1}\left(1-\frac{1}{p^{s}}\right)^{-1}\left(1-\frac{\alpha^{2}_{f}(p,2)}{p^{s}}\right)^{-1}\\ =&\zeta(2s)\sum_{n=1}^{\infty}\frac{\lambda_{f}(n^{2})}{n^{s}}=\prod_{p}\left(1-\frac{\lambda_{f}(p^{2})}{p^{s}}+\frac{\lambda_{f}(p^{2})}{p^{2s}}-\frac{1}{p^{3s}}\right)^{-1}.\end{split}

It follows from a result of G. Shimura [Shimura] that the corresponding completed LL-function

(2.10) Λ(s,sym2f)=\displaystyle\Lambda(s,\operatorname{sym}^{2}f)= π3s/2Γ(s+12)Γ(s+κ12)Γ(s+κ2)L(s,sym2f)\displaystyle\pi^{-3s/2}\Gamma\left(\frac{s+1}{2}\right)\Gamma\left(\frac{s+\kappa-1}{2}\right)\Gamma\left(\frac{s+\kappa}{2}\right)L(s,\operatorname{sym}^{2}f)

is entire and satisfies the functional equation Λ(s,sym2f)=Λ(1s,sym2f)\Lambda(s,\operatorname{sym}^{2}f)=\Lambda(1-s,\operatorname{sym}^{2}f).

2.5. Functional equations for Dirichlet LL-functions

A key ingredient needed in our proof of Theorem 1.1 is the following functional equation for all Dirichlet characters χ\chi modulo nn from [Cech1, Proposition 2.3].

Lemma 2.6.

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

(2.11) 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}}.

2.7. Bounding LL-functions

In this section, we gather various estimates on the values of LL-functions under GRH, necessary in the sequal. For any quadratic character ψ\psi modulo nn, we write ψ^\widehat{\psi} for the primitive character that induces ψ\psi. Note that every such ψ^\widehat{\psi} can be written in form ψ^=χ(d)\widehat{\psi}=\chi^{(d)} for some fundamental discriminant d|nd|n (see [MVa1, Theorem 9.13]). We further write nn uniquely as n=n1n2n=n_{1}n_{2} with (n1,d)=1(n_{1},d)=1 and p|n2p|dp|n_{2}\Rightarrow p|d. Using these notations, we get that for any integer qq,

(2.12) L(q)(s,fψ)=L(s,fψ^)p|qn1(1αf(p,1)ψ^(p)ps)(1αf(p,2)ψ^(p)ps).\displaystyle\begin{split}L^{(q)}(s,f\otimes\psi)=L(s,f\otimes\widehat{\psi})\prod_{p|qn_{1}}\left(1-\frac{\alpha_{f}(p,1)\widehat{\psi}(p)}{p^{s}}\right)\left(1-\frac{\alpha_{f}(p,2)\widehat{\psi}(p)}{p^{s}}\right).\end{split}

We now apply (2.2) to see that for i=1,2i=1,2,

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

It follows from the above that

(2.13) p|qn1(1αf(p,1)ψ^(p)ps)(1αf(p,2)ψ^(p)ps)4ω(q1n)(qn1)max(0,(2s))(qn1)max(0,(2s))+ε,\displaystyle\begin{split}\prod_{p|qn_{1}}\left(1-\frac{\alpha_{f}(p,1)\widehat{\psi}(p)}{p^{s}}\right)\left(1-\frac{\alpha_{f}(p,2)\widehat{\psi}(p)}{p^{s}}\right)\ll 4^{\omega(q_{1}n)}(qn_{1})^{\max(0,-\Re(2s))}\ll(qn_{1})^{\max(0,-\Re(2s))+\varepsilon},\end{split}

where the last estimate above follows from (2.8).

It follows from this and the functional equation in (1.2) that, for (s)1/2\Re(s)\leq 1/2,

(2.14) L(s,fψ^)n12(s)Γ(1s+κ12)Γ(s+κ12)L(1s,fψ^)(n(1+|s|))12(s)L(1s,fψ^),\displaystyle\begin{split}L(s,f\otimes\widehat{\psi})\ll n^{1-2\Re(s)}\frac{\Gamma(1-s+\frac{\kappa-1}{2})}{\Gamma(s+\frac{\kappa-1}{2})}L(1-s,f\otimes\widehat{\psi})\ll(n(1+|s|))^{1-2\Re(s)}L(1-s,f\otimes\widehat{\psi}),\end{split}

where the last bound above follows from Stirling’s formula (see [iwakow, (5.113)]), which gives that for constants c0,d0c_{0},d_{0},

(2.15) Γ(c0(1s)+d0)Γ(c0s+d0)(1+|s|)c0(12(s)).\displaystyle\frac{\Gamma(c_{0}(1-s)+d_{0})}{\Gamma(c_{0}s+d_{0})}\ll(1+|s|)^{c_{0}(1-2\Re(s))}.

We further note from [iwakow, Theorem 5.19, Corollary 5.20] that under GRH, we have for (s)1/2\Re(s)\geq 1/2,

(2.16) |L(s,fψ^)|,|L(s,ψ^)||sn|ε,|L(s,sym2f)||s|ε.\displaystyle\begin{split}&\big{|}L\left(s,f\otimes\widehat{\psi}\right)\big{|},\quad\big{|}L\left(s,\widehat{\psi}\right)\big{|}\ll|sn|^{\varepsilon},\quad\big{|}L\left(s,\text{sym}^{2}f\right)\big{|}\ll|s|^{\varepsilon}.\end{split}

The above combined with (2.12)–(2.16) now implies that

(2.17) L(q)(s,fψ)(qn1)max(0,(s))+ε(n(1+|s|))max{12(s),0}+ε.\displaystyle\begin{split}L^{(q)}(s,f\otimes\psi)\ll(qn_{1})^{\max(0,-\Re(s))+\varepsilon}(n(1+|s|))^{\max\{1-2\Re(s),0\}+\varepsilon}.\end{split}

Our discussions above also apply to other LL-functions. For example, for the Dirichlet LL-function L(s,ψ)L(s,\psi). Under our notation above, if dd is a fundamental discriminant, then we have the following functional equation (see [sound1, p. 456]) for L(s,χ(d))L(s,\chi^{(d)}).

Λ(s,χ(d))=:(|d|π)s2Γ(s2)L(s,χ(d))=Λ(1s,χ(d)).\displaystyle\Lambda(s,\chi^{(d)})=:\Big{(}\frac{|d|}{\pi}\Big{)}^{\frac{s}{2}}\Gamma\Big{(}\frac{s}{2}\Big{)}L(s,\chi^{(d)})=\Lambda(1-s,\chi^{(d)}).

It follows from this and (2.15) that

(2.18) L(q)(s,ψ)(qn1)max(0,(s))+ε(n(1+|s|))max{1/2(s),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),0\}+\varepsilon}.\end{split}

Similarly, by (2.10) and (2.15), we get

(2.19) L(q)(s,sym2f)qmax(0,(s))+ε(1+|s|)max{3(1/2(s)),0}+ε.\displaystyle\begin{split}L^{(q)}(s,\text{sym}^{2}f)\ll q^{\max(0,-\Re(s))+\varepsilon}(1+|s|)^{\max\{3(1/2-\Re(s)),0\}+\varepsilon}.\end{split}

Lastly, from [iwakow, Theorem 5.19, Corollary 5.20], we have, under GRH, for (s)1/2+ε\Re(s)\geq 1/2+\varepsilon,

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

2.8. Some results on multivariable complex functions

We include in this section some results from multivariable complex analysis. First we need 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 have 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) be 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).

The convex hull of an open set TnT\subset\mathbb{C}^{n} is denoted by T^\widehat{T}. Then we quote the result from [Cech1, Proposition C.5] on the modulus of holomorphic continuations of functions in multiple variables.

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

Let μ\mu denote the Möbius function. Using the notations defined in Section 2.4, we define for (s),(w),(z)\Re(s),\Re(w),\Re(z) large enough,

(3.1) A(s,w,z;f)=(n,2)=1L(2)(w,fχn)L(2)(z,fχn)ns=(nmk,2)=1λf(m)cf(k)χn(k)χn(m)kzmwns=(mk,2)=1λf(m)cf(k)L(s,χ(4mk))mwkz.\displaystyle\begin{split}A(s,w,z;f)=\sum_{\begin{subarray}{c}(n,2)=1\end{subarray}}\frac{L^{(2)}(w,f\otimes\chi_{n})}{L^{(2)}(z,f\otimes\chi_{n})n^{s}}=\sum_{\begin{subarray}{c}(nmk,2)=1\end{subarray}}\frac{\lambda_{f}(m)c_{f}(k)\chi_{n}(k)\chi_{n}(m)}{k^{z}m^{w}n^{s}}=\sum_{\begin{subarray}{c}(mk,2)=1\end{subarray}}\frac{\lambda_{f}(m)c_{f}(k)L\left(s,\chi^{(4mk)}\right)}{m^{w}k^{z}}.\end{split}

We shall devote the next few sections to articulating the analytical properties of A(s,w,z;f)A(s,w,z;f) as the proof of Theorem 1.1 relies crucially on them.

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

We start with the series representation for A(s,w,z;f)A(s,w,z;f) given by the first equality in (3.1). This gives that if (z)>1/2\Re(z)>1/2,

(3.2) A(s,w,z;f)=(n,2)=1L(2)(w,fχn)L(2)(z,fχn)ns=(h,2)=11h2s(n,2)=1L(2)(w,fχn)p|h(1αf(p,1)χn(p)pw)(1αf(p,2)χn(p)pw)nsL(2)(z,fχn)p|h(1αf(p,1)χn(p)pz)(1αf(p,2)χn(p)pz)(h,2)=1hmax(0,2(w))+εh2s(n,2)=1|L(2)(w,fχn)L(2)(z,fχn)ns|,\displaystyle\begin{split}A(s,w,z;f)=&\sum_{\begin{subarray}{c}(n,2)=1\end{subarray}}\frac{L^{(2)}(w,f\otimes\chi_{n})}{L^{(2)}(z,f\otimes\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,f\otimes\chi_{n})\prod_{p|h}(1-\alpha_{f}(p,1)\chi_{n}(p)p^{-w})(1-\alpha_{f}(p,2)\chi_{n}(p)p^{-w})}{n^{s}L^{(2)}(z,f\otimes\chi_{n})\prod_{p|h}(1-\alpha_{f}(p,1)\chi_{n}(p)p^{-z})(1-\alpha_{f}(p,2)\chi_{n}(p)p^{-z})}\\ \ll&\sum_{\begin{subarray}{c}(h,2)=1\end{subarray}}\frac{h^{\max(0,-2\Re(w))+\varepsilon}}{h^{2s}}\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(n,2)=1\end{subarray}}\Big{|}\frac{L^{(2)}(w,f\otimes\chi_{n})}{L^{(2)}(z,f\otimes\chi_{n})n^{s}}\Big{|},\end{split}

where \sum^{*} henceforth denotes the sum over square-free integers and the last estimate above is obtained using (2.13) and a similar observation that if (z)>1/2\Re(z)>1/2,

(3.3) p|h(1αf(p,1)χn(p)pz)1(1αf(p,2)χn(p)pz)1hε.\displaystyle\begin{split}\prod_{p|h}(1-\alpha_{f}(p,1)\chi_{n}(p)p^{-z})^{-1}(1-\alpha_{f}(p,2)\chi_{n}(p)p^{-z})^{-1}\ll h^{\varepsilon}.\end{split}

Recall that we have L(2)(w,fχn)=L(w,fχ(±4n))L^{(2)}(w,f\otimes\chi_{n})=L(w,f\otimes\chi^{(\pm 4n)}) for n±1(mod4)n\equiv\pm 1\pmod{4}. We write χ~n\widetilde{\chi}_{n} for the primitive Dirichlet character that induces χ(±4n)\chi^{(\pm 4n)}. For a square-free nn, we see that χ~n=χ(n)\widetilde{\chi}_{n}=\chi^{(n)} is a primitive character modulo nn if n1(mod4)n\equiv 1\pmod{4} and χ~n=χ(4n)\widetilde{\chi}_{n}=\chi^{(4n)} is a primitive character modulo 4n4n if n1(mod4)n\equiv-1\pmod{4}. For n1(mod4)n\equiv 1\pmod{4}, we have

|L(2)(w,fχn)|=|(1αf(2,1)χ~n(2)2w)(1αf(2,2)χ~n(2)2w)L(w,fχ~n)||L(w,fχ~n)|.\displaystyle\begin{split}\big{|}L^{(2)}(w,f\otimes\chi_{n})\big{|}=&\big{|}(1-\alpha_{f}(2,1)\widetilde{\chi}_{n}(2)2^{-w})(1-\alpha_{f}(2,2)\widetilde{\chi}_{n}(2)2^{-w})L(w,f\otimes\widetilde{\chi}_{n})\big{|}\ll\big{|}L(w,f\otimes\widetilde{\chi}_{n})\big{|}.\end{split}

The above bound also holds for n1(mod4)n\equiv-1\pmod{4}.

Similarly, if (z)>1/2\Re(z)>1/2, we have, by (2.20) and an estimation analogous to (3.3), that under GRH,

|L(2)(z,fχn)|1nε|L(z,fχ~n)|1(|z|n)ε.\displaystyle\begin{split}|L^{(2)}(z,f\otimes\chi_{n})|^{-1}\ll n^{\varepsilon}|L(z,f\otimes\widetilde{\chi}_{n})|^{-1}\ll(|z|n)^{\varepsilon}.\end{split}

The above, together with (3.2), allows us to deduce that if (z)>1/2\Re(z)>1/2, then

(3.4) A(s,w,z;f)(h,2)=1hmax(0,2(w))+εh2s(n,2)=1|L(w,fχ~n)||nsε|.\displaystyle\begin{split}A(s,w,z;f)\ll&\sum_{\begin{subarray}{c}(h,2)=1\end{subarray}}\frac{h^{\max(0,-2\Re(w))+\varepsilon}}{h^{2s}}\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(n,2)=1\end{subarray}}\frac{|L(w,f\otimes\widetilde{\chi}_{n})|}{|n^{s-\varepsilon}|}.\end{split}

We now apply (2.17) to deduce from the above that under GRH, except for a simple pole at w=1w=1, both sums of the right-hand side expression in (3.4) are convergent for (s)>1\Re(s)>1, (w)1/2\Re(w)\geq 1/2, (z)>1/2\Re(z)>1/2 as well as for (2s)>1\Re(2s)>1, (2s+2w)>1\Re(2s+2w)>1, (s+2w)>2\Re(s+2w)>2, (w)<1/2\Re(w)<1/2, (z)>1/2\Re(z)>1/2.

We thus conclude that the function A(s,w,z;f)A(s,w,z;f) converges absolutely in the region

S0={(s,w,z):(s)>1,(2s+2w)>1,(s+2w)>2,(z)>12}.S_{0}=\{(s,w,z):\Re(s)>1,\ \Re(2s+2w)>1,\ \Re(s+2w)>2,\ \Re(z)>\tfrac{1}{2}\}.

Note that the condition (2s+2w)>1\Re(2s+2w)>1 is implied by the other conditions so that we have

S0={(s,w,z):(s)>1,(s+2w)>2,(z)>12}.S_{0}=\{(s,w,z):\Re(s)>1,\ \Re(s+2w)>2,\ \Re(z)>\tfrac{1}{2}\}.

Next, we deduce from the last expression of (3.1) that A(s,w,z;f)A(s,w,z;f) is given by the series

(3.5) A(s,w,z;f)=\displaystyle A(s,w,z;f)= (mk,2)=1λf(m)cf(k)L(s,χ(4mk))mwkz.\displaystyle\sum_{\begin{subarray}{c}(mk,2)=1\end{subarray}}\frac{\lambda_{f}(m)c_{f}(k)L(s,\chi^{(4mk)})}{m^{w}k^{z}}.

We write mk=(mk)0(mk)12mk=(mk)_{0}(mk)^{2}_{1} with (mk)0(mk)_{0} odd and square-free. Note that χ((mk)0)\chi^{((mk)_{0})} is a primitive character modulo (mk)0(mk)_{0} when (mk)01(mod4)(mk)_{0}\equiv 1\pmod{4}. We now apply (2.4), (2.7) and (2.18) to see that, except for a simple pole at s=1s=1 arising from the summands with mk=mk=\square, the sums over m,km,k such that (mk)01(mod4)(mk)_{0}\equiv 1\pmod{4} in (3.5) converge absolutely in the region

S1=\displaystyle S_{1}= {(s,w,z):(w)>1,(z)>1,(s)12}{(s,w,z):0(s)<12,(s+w)>32,(s+z)>32}\displaystyle\{(s,w,z):\Re(w)>1,\ \Re(z)>1,\ \Re(s)\geq\tfrac{1}{2}\}\bigcup\{(s,w,z):0\leq\Re(s)<\tfrac{1}{2},\ \Re(s+w)>\tfrac{3}{2},\ \Re(s+z)>\tfrac{3}{2}\}
{(s,w,z):(s)<0,(2s+w)>32,(2s+z)>32}.\displaystyle\hskip 85.35826pt\bigcup\{(s,w,z):\Re(s)<0,\ \Re(2s+w)>\tfrac{3}{2},\ \Re(2s+z)>\tfrac{3}{2}\}.

Note that similar estimations hold when (mk)02,3(mod4)(mk)_{0}\equiv 2,3\pmod{4}, in which case χ(4(mk)0)\chi^{(4(mk)_{0})} is a primitive character modulo 4(mk)04(mk)_{0}. We thus conclude that, except for a simple pole at s=1s=1 arising from the summands with mk=mk=\square, the function A(s,w,z;f)A(s,w,z;f) converges absolutely in the region S1S_{1}.

To determine the convex hull of S0S_{0} and S1S_{1}, we first note that for a fixed (z0)>1\Re(z_{0})>1, the points (1/2,1,z0)(1/2,1,z_{0}) and (1,1/2,z0)(1,1/2,z_{0}) are in the closures of S0S_{0} and S1S_{1}, respectively. These two points determine a line segment: (s+w)=3/2\Re(s+w)=3/2 with 1/2(s)11/2\leq\Re(s)\leq 1 on the plane (z)=(z0)\Re(z)=\Re(z_{0}). Note further that when (s)>1/2\Re(s)>1/2 and (z)>1\Re(z)>1, the conditions (s+z)>3/2\Re(s+z)>3/2 and (2s+z)>3/2\Re(2s+z)>3/2 are automatically satisfied. We then deduce that the convex hull of S0S_{0} and S1S_{1} contains points (s,w,z)(s,w,z) satisfying

{(z)>1,(s)>1/2,(s+2w)>2,(s+z)>32,(s+w)>32,(2s+z)>32}.\{\Re(z)>1,\ \Re(s)>1/2,\ \Re(s+2w)>2,\ \Re(s+z)>\tfrac{3}{2},\ \Re(s+w)>\tfrac{3}{2},\ \Re(2s+z)>\tfrac{3}{2}\}.

Combining the above set with the subsets of S1S_{1} containing points with (s)<1/2\Re(s)<1/2, we get that the convex hull of S0S_{0} and S1S_{1} contains points (s,w,z)(s,w,z) satisfying

(3.6) {(z)>1,(s+2w)>2,(s+z)>32,(s+w)>32,(2s+w)>32,(2s+z)>32}.\{\Re(z)>1,\Re(s+2w)>2,\ \Re(s+z)>\tfrac{3}{2},\ \Re(s+w)>\tfrac{3}{2},\ \Re(2s+w)>\tfrac{3}{2},\ \Re(2s+z)>\tfrac{3}{2}\}.

On the other hand, if 1/2<(z)<11/2<\Re(z)<1, the points in S0S_{0} are certainly contained in the convex hull of S0S_{0} and S1S_{1}. We may thus focus on the case (s)<1\Re(s)<1. In this case, we note that the points (1,1/2,1/2)(1,1/2,1/2), (1,1/2,1)(1,1/2,1) are in the closure of S0S_{0} and (1/2,1,1)(1/2,1,1) the closure of S1S_{1}. These three points determine a triangular region which can be regarded as the base of the region \mathcal{R} enclosed by the four planes: (s+w)=3/2\Re(s+w)=3/2, (s)=1\Re(s)=1, (z)=1\Re(z)=1, (s+z)=3/2\Re(s+z)=3/2. It follows that the points in this triangle region are all in the convex hulls of S0S_{0} and S1S_{1}. Further note that the points on the boundary {(z)=1}\mathcal{R}\cap\{\Re(z)=1\} of \mathcal{R} are all in the convex hulls of S0S_{0} and S1S_{1} since they can be identified with the set {(s,w,z):(z)=1,1/2(s)1,(s+w)3/2}\{(s,w,z):\Re(z)=1,1/2\leq\Re(s)\leq 1,\Re(s+w)\geq 3/2\} and hence is contained in the convex hull of the set given in (3.6). Hence the entire region \mathcal{R} lies in the convex hull of S0S_{0} and S1S_{1}. Next, if 1/2<(z)<11/2<\Re(z)<1, the condition (s+z)>3/2\Re(s+z)>3/2 implies that (s)>1/2\Re(s)>1/2 so that one also has (2s+z)>3/2\Re(2s+z)>3/2. Similarly, when 1/2<(z)<11/2<\Re(z)<1, the condition (s+w)>3/2\Re(s+w)>3/2 implies that (2s+w)>3/2\Re(2s+w)>3/2. Lastly, the condition (s+2w)>2\Re(s+2w)>2 implies (s+w)>3/2\Re(s+w)>3/2 and (2s+w)>3/2\Re(2s+w)>3/2 for (s)>1\Re(s)>1. It follows from the discussions here that the intersection of the convex hulls of S0S_{0} and S1S_{1} thus equals

(3.7) S2={(s,w,z):(z)>12,(s+2w)>2,(s+z)>32,(s+w)>32,(2s+w)>32,(2s+z)>32}.S_{2}=\{(s,w,z):\Re(z)>\tfrac{1}{2},\ \Re(s+2w)>2,\ \Re(s+z)>\tfrac{3}{2},\ \Re(s+w)>\tfrac{3}{2},\ \Re(2s+w)>\tfrac{3}{2},\ \Re(2s+z)>\tfrac{3}{2}\}.

The above, together with Theorem 2.10, implies that (s1)(w1)A(s,w,z;f)(s-1)(w-1)A(s,w,z;f) converges absolutely in the region S2S_{2}.

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

It follows from (3.5) that A(s,w,z;f)A(s,w,z;f) has a pole at s=1s=1 arising from the terms with mk=mk=\square. To compute the corresponding residue and for our treatments later in the proof, we introduce the sum

A1(s,w,z;f)=:(mk,2)=1mk=λf(m)cf(k)L(s,χ(4mk))mwkz=(mk,2)=1mk=λf(m)cf(k)ζ(s)p|2mk(1ps)mwkz.\displaystyle\begin{split}A_{1}(s,w,z;f)=:\sum_{\begin{subarray}{c}(mk,2)=1\\ mk=\square\end{subarray}}\frac{\lambda_{f}(m)c_{f}(k)L\left(s,\chi^{(4mk)}\right)}{m^{w}k^{z}}=\sum_{\begin{subarray}{c}(mk,2)=1\\ mk=\square\end{subarray}}\frac{\lambda_{f}(m)c_{f}(k)\zeta(s)\prod_{p|2mk}(1-p^{-s})}{m^{w}k^{z}}.\end{split}

We further denote by at(n)a_{t}(n) for any tt\in\mathbb{C} the multiplicative function such that at(pk)=11/pta_{t}(p^{k})=1-1/p^{t} for any prime pp. Thus, we recast A1(s,w,z;f)A_{1}(s,w,z;f) as

A1(s,w,z;f)=ζ(2)(s)(mk,2)=1mk=λf(m)cf(k)as(mk)mwkz.\displaystyle\begin{split}A_{1}(s,w,z;f)=\zeta^{(2)}(s)\sum_{\begin{subarray}{c}(mk,2)=1\\ mk=\square\end{subarray}}\frac{\lambda_{f}(m)c_{f}(k)a_{s}(mk)}{m^{w}k^{z}}.\end{split}

We now write the last sum above as an Euler product. Slightly abusing notation by writing pkp^{k^{\prime}} with kk^{\prime}\in\mathbb{Z} for the highest power of pp dividing kk, and similarly for mm. Thus, we obtain, utilizing (2.6), that

(3.8) A1(s,w,z;f)=ζ(2)(s)p>2m,k0m+k evenλf(pm)cf(pk)as(pm+k)pmw+kz=ζ(2)(s)p>2(m02mλf(pm)as(pm)pmw+m0(m,2)=1cf(p)λf(pm)as(pm+1)pz+mw+m02mcf(p2)λf(pm)as(pm+2)p2z+mw)=ζ(2)(s)p>2(1+(11ps)1p2z+(11ps)(1+1p2z)m=1λf(p2m)p2mw(11ps)λf(p)pzm=0λf(p2m+1)p(2m+1)w).\displaystyle\begin{split}A_{1}&(s,w,z;f)\\ =&\zeta^{(2)}(s)\prod_{p>2}\sum_{\begin{subarray}{c}m,k\geq 0\\ m+k\text{ even}\end{subarray}}\frac{\lambda_{f}(p^{m})c_{f}(p^{k})a_{s}(p^{m+k})}{p^{mw+kz}}\\ =&\zeta^{(2)}(s)\prod_{p>2}\left(\sum_{\begin{subarray}{c}m\geq 0\\ 2\mid m\end{subarray}}\frac{\lambda_{f}(p^{m})a_{s}(p^{m})}{p^{mw}}+\sum_{\begin{subarray}{c}m\geq 0\\ (m,2)=1\end{subarray}}\frac{c_{f}(p)\lambda_{f}(p^{m})a_{s}(p^{m+1})}{p^{z+mw}}+\sum_{\begin{subarray}{c}m\geq 0\\ 2\mid m\end{subarray}}\frac{c_{f}(p^{2})\lambda_{f}(p^{m})a_{s}(p^{m+2})}{p^{2z+mw}}\right)\\ =&\zeta^{(2)}(s)\prod_{p>2}\left(1+\left(1-\frac{1}{p^{s}}\right)\frac{1}{p^{2z}}+\left(1-\frac{1}{p^{s}}\right)\left(1+\frac{1}{p^{2z}}\right)\sum^{\infty}_{m=1}\frac{\lambda_{f}(p^{2m})}{p^{2mw}}-\left(1-\frac{1}{p^{s}}\right)\frac{\lambda_{f}(p)}{p^{z}}\sum^{\infty}_{m=0}\frac{\lambda_{f}(p^{2m+1})}{p^{(2m+1)w}}\right).\end{split}

We now use (2.3) to infer that (we may assume that αf(p,1)αf(p,2)\alpha_{f}(p,1)\neq\alpha_{f}(p,2) as the other case follows from continuity)

(3.9) i=0λf(p2i)piu=i=01piuαf2i+1(p,1)αf2i+1(p,2)αf(p,1)αf(p,2)=(1αf2(p,1)pu)1(1αf2(p,2)pu)1(1+1pu).\displaystyle\begin{split}\sum^{\infty}_{i=0}\frac{\lambda_{f}(p^{2i})}{p^{iu}}=&\sum^{\infty}_{i=0}\frac{1}{p^{iu}}\frac{\alpha^{2i+1}_{f}(p,1)-\alpha^{2i+1}_{f}(p,2)}{\alpha_{f}(p,1)-\alpha_{f}(p,2)}=\left(1-\frac{\alpha^{2}_{f}(p,1)}{p^{u}}\right)^{-1}\left(1-\frac{\alpha^{2}_{f}(p,2)}{p^{u}}\right)^{-1}\left(1+\frac{1}{p^{u}}\right).\end{split}

Similarly, we have

(3.10) i=0λf(p2i+1)p(i+1)u=(1αf2(p,1)pu)1(1αf2(p,2)pu)1λf(p)pu.\displaystyle\begin{split}\sum^{\infty}_{i=0}\frac{\lambda_{f}(p^{2i+1})}{p^{(i+1)u}}=&\left(1-\frac{\alpha^{2}_{f}(p,1)}{p^{u}}\right)^{-1}\left(1-\frac{\alpha^{2}_{f}(p,2)}{p^{u}}\right)^{-1}\frac{\lambda_{f}(p)}{p^{u}}.\end{split}

Inserting (3.9) and (3.10) into (3.8) yields

(3.11) A1(s,w,z;f)=ζ(2)(s)p>2m,k0m+k evenλf(pm)cf(pk)as(pm+k)pmw+kz=ζ(2)(s)p>2(1αf2(p,1)p2w)1(1αf2(p,2)p2w)1×p>2((11ps)(1+1p2z)(1+1p2w)+1ps(1αf2(p,1)p2w)(1αf2(p,2)p2w)(11ps)λf2(p)pz+w)=ζ(2)(s)L(2)(2w,sym2f)ζ(2)(2w)×p>2((11ps)(1+1p2z)(1+1p2w)+1ps(1αf2(p,1)p2w)(1αf2(p,2)p2w)(11ps)λf2(p)pz+w)=ζ(2)(s)L(2)(2w,sym2f)ζ(2)(2w)×p>2(1+1p2z(11ps)(1+1p2w)+(11ps)1p2wαf2(p,1)+αf2(p,2)p2w+s+1p4w+s(11ps)λf2(p)pz+w)=ζ(s)L(2)(2w,sym2f)P(s,w,z;f),\displaystyle\begin{split}A_{1}&(s,w,z;f)=\zeta^{(2)}(s)\prod_{p>2}\sum_{\begin{subarray}{c}m,k\geq 0\\ m+k\text{ even}\end{subarray}}\frac{\lambda_{f}(p^{m})c_{f}(p^{k})a_{s}(p^{m+k})}{p^{mw+kz}}\\ =&\zeta^{(2)}(s)\prod_{p>2}\left(1-\frac{\alpha^{2}_{f}(p,1)}{p^{2w}}\right)^{-1}\left(1-\frac{\alpha^{2}_{f}(p,2)}{p^{2w}}\right)^{-1}\\ &\times\prod_{p>2}\left(\left(1-\frac{1}{p^{s}}\right)\left(1+\frac{1}{p^{2z}}\right)\left(1+\frac{1}{p^{2w}}\right)+\frac{1}{p^{s}}\left(1-\frac{\alpha^{2}_{f}(p,1)}{p^{2w}}\right)\left(1-\frac{\alpha^{2}_{f}(p,2)}{p^{2w}}\right)-\left(1-\frac{1}{p^{s}}\right)\frac{\lambda^{2}_{f}(p)}{p^{z+w}}\right)\\ =&\frac{\zeta^{(2)}(s)L^{(2)}(2w,\text{sym}^{2}f)}{\zeta^{(2)}(2w)}\\ &\times\prod_{p>2}\left(\left(1-\frac{1}{p^{s}}\right)\left(1+\frac{1}{p^{2z}}\right)\left(1+\frac{1}{p^{2w}}\right)+\frac{1}{p^{s}}\left(1-\frac{\alpha^{2}_{f}(p,1)}{p^{2w}}\right)\left(1-\frac{\alpha^{2}_{f}(p,2)}{p^{2w}}\right)-\left(1-\frac{1}{p^{s}}\right)\frac{\lambda^{2}_{f}(p)}{p^{z+w}}\right)\\ =&\frac{\zeta^{(2)}(s)L^{(2)}(2w,\text{sym}^{2}f)}{\zeta^{(2)}(2w)}\\ &\times\prod_{p>2}\left(1+\frac{1}{p^{2z}}\left(1-\frac{1}{p^{s}}\right)\left(1+\frac{1}{p^{2w}}\right)+\left(1-\frac{1}{p^{s}}\right)\frac{1}{p^{2w}}-\frac{\alpha^{2}_{f}(p,1)+\alpha^{2}_{f}(p,2)}{p^{2w+s}}+\frac{1}{p^{4w+s}}-\left(1-\frac{1}{p^{s}}\right)\frac{\lambda^{2}_{f}(p)}{p^{z+w}}\right)\\ =&\zeta(s)L^{(2)}(2w,\text{sym}^{2}f)P(s,w,z;f),\end{split}

where P(s,w,z;f)P(s,w,z;f) is defined in (1.4).

We deduce from (1.4) and (3.11) that except for a simple pole at s=1s=1, P(s,w,z;f)P(s,w,z;f) and A1(s,w,z;f)A_{1}(s,w,z;f) are holomorphic in the region

(3.12) S3={(s,w,z):\displaystyle S_{3}=\{(s,w,z):\ (2z)>1,(s+2z)>1,(s+2w)>1,(w+z)>1,(4w)>1,(s+w+z)>1}.\displaystyle\Re(2z)>1,\ \Re(s+2z)>1,\ \Re(s+2w)>1,\ \Re(w+z)>1,\ \Re(4w)>1,\ \Re(s+w+z)>1\}.

Recalling that the residue of ζ(s)\zeta(s) at s=1s=1 is 11, we arrive at

(3.13) Ress=1\displaystyle\mathrm{Res}_{s=1} A(s,12+α,12+β;f)=Ress=1A1(s,12+α,12+β;f)=L(2)(1+2α,sym2f)P(1,12+α,12+β;f).\displaystyle A(s,\tfrac{1}{2}+\alpha,\tfrac{1}{2}+\beta;f)=\mathrm{Res}_{s=1}A_{1}(s,\tfrac{1}{2}+\alpha,\tfrac{1}{2}+\beta;f)=L^{(2)}(1+2\alpha,\text{sym}^{2}f)P(1,\tfrac{1}{2}+\alpha,\tfrac{1}{2}+\beta;f).

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

From (3.5), we get

(3.14) A(s,w,z;f)=(mk,2)=1mk=λf(m)cf(k)L(s,χ(4mk))mwkz+(mk,2)=1mkλf(m)cf(k)L(s,χ(4mk))mwkz=(mk,2)=1mk=λf(m)cf(k)ζ(s)p|2mk(1ps)mwkz+(mk,2)=1mkλf(m)cf(k)L(s,χ(4mk))mwkz=:A1(s,w,z;f)+A2(s,w,z;f).\displaystyle\begin{split}A(s,w,z;f)=&\sum_{\begin{subarray}{c}(mk,2)=1\\ mk=\square\end{subarray}}\frac{\lambda_{f}(m)c_{f}(k)L\left(s,\chi^{(4mk)}\right)}{m^{w}k^{z}}+\sum_{\begin{subarray}{c}(mk,2)=1\\ mk\neq\square\end{subarray}}\frac{\lambda_{f}(m)c_{f}(k)L\left(s,\chi^{(4mk)}\right)}{m^{w}k^{z}}\\ =&\sum_{\begin{subarray}{c}(mk,2)=1\\ mk=\square\end{subarray}}\frac{\lambda_{f}(m)c_{f}(k)\zeta(s)\prod_{p|2mk}(1-p^{-s})}{m^{w}k^{z}}+\sum_{\begin{subarray}{c}(mk,2)=1\\ mk\neq\square\end{subarray}}\frac{\lambda_{f}(m)c_{f}(k)L\left(s,\chi^{(4mk)}\right)}{m^{w}k^{z}}\\ =:\ &A_{1}(s,w,z;f)+A_{2}(s,w,z;f).\end{split}

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

Next, we apply the functional equation given in Lemma 2.6 for L(s,χ(4mk))L\left(s,\chi^{(4mk)}\right) in the case mkmk\neq\square by observing that χ(4mk)\chi^{(4mk)} is a Dirichlet character modulo 4mk4mk for any m,k1m,k\geq 1 with χ(4mk)(1)=1\chi^{(4mk)}(-1)=1. We obtain from (2.11) that

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

where C(s,w,z;f)C(s,w,z;f) is given by the triple Dirichlet series

C(s,w,z;f)=\displaystyle C(s,w,z;f)= q,m,k(mk,2)=1mkλf(m)cf(k)τ(χ(4mk),q)qsmwkz\displaystyle\sum_{\begin{subarray}{c}q,m,k\\ (mk,2)=1\\ mk\neq\square\end{subarray}}\frac{\lambda_{f}(m)c_{f}(k)\tau(\chi^{(4mk)},q)}{q^{s}m^{w}k^{z}}
=\displaystyle= q,m,k(mk,2)=1λf(m)cf(k)τ(χ(4mk),q)qsmwkzq,m,k(mk,2)=1mk=λf(m)cf(k)τ(χ(4mk),q)qsmwkz.\displaystyle\sum_{\begin{subarray}{c}q,m,k\\ \begin{subarray}{c}(mk,2)=1\end{subarray}\end{subarray}}\frac{\lambda_{f}(m)c_{f}(k)\tau(\chi^{(4mk)},q)}{q^{s}m^{w}k^{z}}-\sum_{\begin{subarray}{c}q,m,k\\ (mk,2)=1\\ mk=\square\end{subarray}}\frac{\lambda_{f}(m)c_{f}(k)\tau(\chi^{(4mk)},q)}{q^{s}m^{w}k^{z}}.

By (3.7), (3.12) and the functional equation (3.15), we see that C(s,w,z;f)C(s,w,z;f) is initially defined for (s)\Re(s), (z)\Re(z) and (w)\Re(w) sufficiently large. To extend this region, we exchange the summations in C(s,w,z;f)C(s,w,z;f) and set mk=lmk=l to obtain that

(3.16) C(s,w,z;f)=q=11qs(l,2)=1τ(χ(4l),q)r(l,zw)lwq=11qs(l,2)=1l=τ(χ(4l),q)r(l,zw)lw=:C1(s,w,z;f)C2(s,w,z;f),\displaystyle\begin{split}C(s,w,z;f)=&\sum^{\infty}_{q=1}\frac{1}{q^{s}}\sum_{\begin{subarray}{c}(l,2)=1\end{subarray}}\frac{\tau\left(\chi^{(4l)},q\right)r(l,z-w)}{l^{w}}-\sum^{\infty}_{q=1}\frac{1}{q^{s}}\sum_{\begin{subarray}{c}(l,2)=1\\ l=\square\end{subarray}}\frac{\tau\left(\chi^{(4l)},q\right)r(l,z-w)}{l^{w}}\\ =:\ &C_{1}(s,w,z;f)-C_{2}(s,w,z;f),\end{split}

where

(3.17) r(l,t)=k|lλf(l/k)cf(k)kt.\displaystyle\begin{split}r(l,t)=\sum_{\begin{subarray}{c}k|l\end{subarray}}\frac{\lambda_{f}(l/k)c_{f}(k)}{k^{t}}.\end{split}

We now define, for two Dirichlet characters ψ\psi and ψ\psi^{\prime} whose conductors divide 88,

(3.18) C1(s,w,z;ψ,ψ,f)=:l,q1G(χl,q)ψ(l)ψ(q)r(l,zw)lwqs,andC2(s,w,z;ψ,ψ,f)=:l,q1G(χl2,q)ψ(l)ψ(q)r(l2,zw)l2wqs.\displaystyle\begin{split}C_{1}(s,w,z;\psi,\psi^{\prime},f)=:\ &\sum_{l,q\geq 1}\frac{G\left(\chi_{l},q\right)\psi(l)\psi^{\prime}(q)r(l,z-w)}{l^{w}q^{s}},\quad\mbox{and}\\ C_{2}(s,w,z;\psi,\psi^{\prime},f)=:\ &\sum_{l,q\geq 1}\frac{G\left(\chi_{l^{2}},q\right)\psi(l)\psi^{\prime}(q)r(l^{2},z-w)}{l^{2w}q^{s}}.\end{split}

Following the arguments contained in [Cech1, §6.4] and making use of Lemma 2.2, we see that

(3.19) C1(s,w,z;f)=2s(C1(s,w,z;ψ2,ψ1,f)+C1(s,w,z;ψ2,ψ1,f))+4s(C1(s,w,z;ψ1,ψ0,f)+C1(s,w,z;ψ1,ψ0,f))+C1(s,w,z;ψ1,ψ1,f)C1(s,w,z;ψ1,ψ1,f),andC2(s,w,z;f)=21sC2(s,w,z;ψ1,ψ1,f)+212sC2(s,w,z;ψ1,ψ0,f).\displaystyle\begin{split}C_{1}(s,w,z;f)=&-2^{-s}\big{(}C_{1}(s,w,z;\psi_{2},\psi_{1},f)+C_{1}(s,w,z;\psi_{-2},\psi_{1},f)\big{)}+4^{-s}\big{(}C_{1}(s,w,z;\psi_{1},\psi_{0},f)\\ &\hskip 42.67912pt+C_{1}(s,w,z;\psi_{-1},\psi_{0},f)\big{)}+C_{1}(s,w,z;\psi_{1},\psi_{-1},f)-C_{1}(s,w,z;\psi_{-1},\psi_{-1},f),\quad\mbox{and}\\ C_{2}(s,w,z;f)=&-2^{1-s}C_{2}(s,w,z;\psi_{1},\psi_{1},f)+2^{1-2s}C_{2}(s,w,z;\psi_{1},\psi_{0},f).\end{split}

Note that every integer q1q\geq 1 can be written uniquely as q=q1q22q=q_{1}q^{2}_{2} with q1q_{1} square-free. We may thus write

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

where

(3.21) D1(s,w,t;q1,ψ,ψ,f)=l,q2=1G(χl,q1q22)ψ(l)ψ(q22)r(l,t)lwq22s,andD2(s,w,t;q1,ψ,ψ,f)=l,q2=1G(χl2,q1q22)ψ(l)ψ(q22)r(l2,t)l2wq22s.\displaystyle\begin{split}D_{1}(s,w,t;q_{1},\psi,\psi^{\prime},f)=&\sum_{l,q_{2}=1}^{\infty}\frac{G\left(\chi_{l},q_{1}q^{2}_{2}\right)\psi(l)\psi^{\prime}(q^{2}_{2})r(l,t)}{l^{w}q^{2s}_{2}},\quad\mbox{and}\\ D_{2}(s,w,t;q_{1},\psi,\psi^{\prime},f)=&\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})r(l^{2},t)}{l^{2w}q^{2s}_{2}}.\end{split}

We have the following result for the analytic properties of Di(s,w,t;q1,ψ,ψ,f)D_{i}(s,w,t;q_{1},\psi,\psi^{\prime},f).

Lemma 3.4.

With the notation as above, for ψψ0\psi\neq\psi_{0}, the functions Di(s,w,t;q1,ψ,ψ,f),i=1,2D_{i}(s,w,t;q_{1},\psi,\psi^{\prime},f),i=1,2 have meromorphic continuations to the region

{(s,w,t):(s)>1,(w)>1,(w+t)>1}.\{(s,w,t):\Re(s)>1,\ \Re(w)>1,\ \Re(w+t)>1\}.

For (s)>1+ε,(w)>1+ε\Re(s)>1+\varepsilon,\Re(w)>1+\varepsilon and (w+t)>1+ε\Re(w+t)>1+\varepsilon, we have

(3.22) |Di(s,w,t;q1,ψ,ψ,f)||q1w(t+w)|ε.\displaystyle|D_{i}(s,w,t;q_{1},\psi,\psi^{\prime},f)|\ll|q_{1}w(t+w)|^{\varepsilon}.
Proof.

As the proofs are similar, we consider only the case for D1(s,w,t;q1,ψ,ψ,f)D_{1}(s,w,t;q_{1},\psi,\psi^{\prime},f) here. First D1(s,w,t;q1,ψ,ψ,f)D_{1}(s,w,t;q_{1},\psi,\psi^{\prime},f) are jointly multiplicative functions of l,q2l,q_{2} by Lemma 2.3 in the double sum defining D1D_{1} in (3.21). Moreover, as ψψ0\psi\neq\psi_{0}, we may assume that ll is odd. We write D1(s,w,z;q1,ψ,ψ,f)D_{1}(s,w,z;q_{1},\psi,\psi^{\prime},f) using (2.6) into an Euler product so that

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

Then we have

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

Note that for a fixed p>2p>2,

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

Observe that

(3.26) r(pl,t)=λf(pl)λf(pl1)λf(p)pt+λf(pl2)p2t,\displaystyle\begin{split}r(p^{l},t)=\lambda_{f}(p^{l})-\frac{\lambda_{f}(p^{l-1})\lambda_{f}(p)}{p^{t}}+\frac{\lambda_{f}(p^{l-2})}{p^{2t}},\end{split}

where we denote λf(pi)=0\lambda_{f}(p^{-i})=0 for integers i0i\geq 0.

Notice that by (2.2) and (2.3), we have

(3.27) |λf(pl)|l+1,l0.\displaystyle\begin{split}|\lambda_{f}(p^{l})|\leq l+1,\quad l\geq 0.\end{split}

It follows from this and (3.26) that for l1l\geq 1,

|r(pl,t)|{4(l+1)(1+pt),l=1,4(l+1)(1+p2t),l2.\displaystyle\begin{split}|r(p^{l},t)|\leq\displaystyle\begin{cases}4(l+1)(1+p^{-t}),&l=1,\\ 4(l+1)(1+p^{-2t}),&l\geq 2.\end{cases}\end{split}

Also, note that Lemma 2.3 implies that

|G(χpl,q1p2k)|pl.\displaystyle\begin{split}|G(\chi_{p^{l}},q_{1}p^{2k})|\ll p^{l}.\end{split}

We apply the above estimations to see that when (s)>1/2\Re(s)>1/2, (w)>1\Re(w)>1, (w+t)>1\Re(w+t)>1,

(3.28) l0,k1ψ(pl)ψ(p2k)G(χpl,q1p2k)r(pl,t)plw+2ks=k1ψ(p2k)G(χ1,q1p2k)p2ks+l,k1ψ(pl)ψ(p2k)G(χpl,q1p2k)r(pl,t)plw+2ksp2s+p2s(1pw14(l+1)(1+pt)+l21pl(w1)4(l+1)(1+p2t))p2s+p2sw+1+p2s2w+2+p2swt+1+p2s2w2t+2p2s+p2swt+1.\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)r(p^{l},t)}{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)r(p^{l},t)}{p^{lw+2ks}}\\ &\ll p^{-2s}+p^{-2s}\left(\frac{1}{p^{w-1}}4(l+1)(1+p^{-t})+\sum_{l\geq 2}\frac{1}{p^{l(w-1)}}4(l+1)(1+p^{-2t})\right)\\ &\ll p^{-2s}+p^{-2s-w+1}+p^{-2s-2w+2}+p^{-2s-w-t+1}+p^{-2s-2w-2t+2}\ll p^{-2s}+p^{-2s-w-t+1}.\end{split}

We now apply Lemma 2.3 and (3.26) to see that when p2q1p\nmid 2q_{1} and (w)>1\Re(w)>1,

(3.29) l=0ψ(pl)G(χpl,q1)r(pl,t)plw=1+ψ(p)λf(p)χ(q1)(p)pw1/2(11pt)=Lp(w12,fχ(q1)ψ)(1λf2(p)p2w1χ(q1)(p)ψ(p)λf(p)pw+t1/2+λf2(p)p2w+t1)=Lp(w12,fχ(q1)ψ)Lp(2w1,sym2f)ζp(2w1)×(1χ(q1)(p)ψ(p)λf(p)pw+t1/2+O(1p2w+t1+1pw+t1/2+2w1+1p2(2w1)))=Lp(w12,fχ(q1)ψ)Lp(2w1,sym2f)ζp(2w1)Lp(t+w1/2,fχ(q1)ψ)×(1+O(1p2w+2t1+1p2w+t1+1pw+t1/2+2w1+1p2(2w1))).\displaystyle\begin{split}\sum_{l=0}^{\infty}&\frac{\psi(p^{l})G\left(\chi_{p^{l}},q_{1}\right)r(p^{l},t)}{p^{lw}}=1+\frac{\psi(p)\lambda_{f}(p)\chi^{(q_{1})}(p)}{p^{w-1/2}}\left(1-\frac{1}{p^{t}}\right)\\ =&L_{p}\left(w-\tfrac{1}{2},f\otimes\chi^{(q_{1})}\psi\right)\left(1-\frac{\lambda_{f}^{2}(p)}{p^{2w-1}}-\frac{\chi^{(q_{1})}(p)\psi(p)\lambda_{f}(p)}{p^{w+t-1/2}}+\frac{\lambda_{f}^{2}(p)}{p^{2w+t-1}}\right)\\ =&\frac{L_{p}\left(w-\tfrac{1}{2},f\otimes\chi^{(q_{1})}\psi\right)}{L_{p}(2w-1,\text{sym}^{2}f)\zeta_{p}(2w-1)}\\ &\hskip 85.35826pt\times\left(1-\frac{\chi^{(q_{1})}(p)\psi(p)\lambda_{f}(p)}{p^{w+t-1/2}}+O\Big{(}\frac{1}{p^{2w+t-1}}+\frac{1}{p^{w+t-1/2+2w-1}}+\frac{1}{p^{2(2w-1)}}\Big{)}\right)\\ =&\frac{L_{p}\left(w-\tfrac{1}{2},f\otimes\chi^{(q_{1})}\psi\right)}{L_{p}(2w-1,\text{sym}^{2}f)\zeta_{p}(2w-1)L_{p}\left(t+w-1/2,f\otimes\chi^{(q_{1})}\psi\right)}\\ &\hskip 85.35826pt\times\left(1+O\Big{(}\frac{1}{p^{2w+2t-1}}+\frac{1}{p^{2w+t-1}}+\frac{1}{p^{w+t-1/2+2w-1}}+\frac{1}{p^{2(2w-1)}}\Big{)}\right).\end{split}

We deduce from (3.24), (3.25), (3.28) and (3.29) that for p2q1p\nmid 2q_{1}, (s)>12,(w)>1,(w+t)>1\Re(s)>\frac{1}{2},\Re(w)>1,\Re(w+t)>1,

(3.30) D1,p(s,w,t;q1,ψ,ψ,f)=Lp(w12,fχ(q1)ψ)Lp(2w1,sym2f)ζp(2w1)Lp(t+w1/2,fχ(q1)ψ)×(1+O(p(2w+2t1)+p(2w+t1)+p(w+t1/2+2w1)+p2(2w1)+p2s+p2swt+1)).\displaystyle\begin{split}D_{1,p}&(s,w,t;q_{1},\psi,\psi^{\prime},f)\\ =&\frac{L_{p}\left(w-\tfrac{1}{2},f\otimes\chi^{(q_{1})}\psi\right)}{L_{p}(2w-1,\text{sym}^{2}f)\zeta_{p}(2w-1)L_{p}\left(t+w-1/2,f\otimes\chi^{(q_{1})}\psi\right)}\\ &\hskip 28.45274pt\times\left(1+O\Big{(}p^{-(2w+2t-1)}+p^{-(2w+t-1)}+p^{-(w+t-1/2+2w-1)}+p^{-2(2w-1)}+p^{-2s}+p^{-2s-w-t+1}\Big{)}\right).\end{split}

The first assertion of the lemma now follows from (3.23), (3.24) and (3.30).

We next note that Lemma 2.3, (3.17) and (3.27) implies that when p|q1,p2p|q_{1},p\neq 2,

(3.31) l=0ψ(pl)G(χpl,q1)r(pl,t)plw=1ψ(p2)p2w1(λf(p2)λf2(p)pt+1p2t)=1+O(p2w1+p2w2t1).\displaystyle\begin{split}&\sum_{l=0}^{\infty}\frac{\psi(p^{l})G\left(\chi_{p^{l}},q_{1}\right)r(p^{l},t)}{p^{lw}}=1-\frac{\psi(p^{2})}{p^{2w-1}}\left(\lambda_{f}(p^{2})-\frac{\lambda^{2}_{f}(p)}{p^{t}}+\frac{1}{p^{2t}}\right)=1+O(p^{-2w-1}+p^{-2w-2t-1}).\end{split}

We thus deduce from (3.24), (3.28) and (3.31) that for p|q1p|q_{1}, p2p\neq 2, (s)>1/2\Re(s)>1/2, (w)>1\Re(w)>1, (w+t)>1\Re(w+t)>1,

(3.32) D1,p(s,w,t;q1,ψ,ψ,f)=1+O(p2w1+p2w2t1+p2s+p2swt+1).\displaystyle\begin{split}&D_{1,p}(s,w,t;q_{1},\psi,\psi^{\prime},f)=1+O\Big{(}p^{-2w-1}+p^{-2w-2t-1}+p^{-2s}+p^{-2s-w-t+1}\Big{)}.\end{split}

We conclude from (3.23), (3.24), (3.30) and (3.32) that for (s)>1+ε\Re(s)>1+\varepsilon, (w)>1+ε\Re(w)>1+\varepsilon and (w+t)>1+ε\Re(w+t)>1+\varepsilon,

D1(s,w,t;q1,ψ,ψ,f)q1ε|L(2q1)(w12,fχ(q1)ψ)L(2q1)(2w1,sym2f)ζ(2q1)(2w1)L(2q1)(t+w1/2,fχ(q1)ψ)|q1ε|L(w12,fχ(q1)ψ)L(2w1,sym2f)ζ(2w1)L(t+w1/2,fχ(q1)ψ)||q1w(w+t)|ε,\displaystyle\begin{split}D_{1}(s,w,t;q_{1},\psi,\psi^{\prime},f)\ll&q_{1}^{\varepsilon}\Big{|}\frac{L^{(2q_{1})}\left(w-\tfrac{1}{2},f\otimes\chi^{(q_{1})}\psi\right)}{L^{(2q_{1})}(2w-1,\text{sym}^{2}f)\zeta^{(2q_{1})}(2w-1)L^{(2q_{1})}\left(t+w-1/2,f\otimes\chi^{(q_{1})}\psi\right)}\Big{|}\\ \ll&q_{1}^{\varepsilon}\Big{|}\frac{L\left(w-\tfrac{1}{2},f\otimes\chi^{(q_{1})}\psi\right)}{L(2w-1,\text{sym}^{2}f)\zeta(2w-1)L\left(t+w-1/2,f\otimes\chi^{(q_{1})}\psi\right)}\Big{|}\ll|q_{1}w(w+t)|^{\varepsilon},\end{split}

where the last estimation above follows (2.16). This implies (3.22) and hence completes the proof of the lemma. ∎

The above lemma now allows us to extend C(s,w,z;f)C(s,w,z;f) to the region

{(s,w,z):(s)>1,(w)>1,(z)>1}.\{(s,w,z):\ \Re(s)>1,\ \Re(w)>1,\ \Re(z)>1\}.

Using (3.12), (3.14) and the above, we can extend (s1)(w1)A(s,w,z;f)(s-1)(w-1)A(s,w,z;f) to the region

S4=\displaystyle S_{4}= {(s,w,z):(2z)>1,(s+2z)>1,(s+2w)>1,(w+z)>1,(4w)>1,(s+w+z)>1,\displaystyle\{(s,w,z):\Re(2z)>1,\ \Re(s+2z)>1,\ \Re(s+2w)>1,\ \Re(w+z)>1,\ \Re(4w)>1,\ \Re(s+w+z)>1,
(s+w)>1,(s+z)>1,(1s)>1}.\displaystyle\hskip 72.26999pt\ \Re(s+w)>1,\ \Re(s+z)>1,\ \Re(1-s)>1\}.

Note that the condition (1s)>1\Re(1-s)>1 is equivalent to (s)<0\Re(s)<0 so that the conditions (s+w)>1,(s+z)>1\Re(s+w)>1,\Re(s+z)>1 is the same as (w)>1\Re(w)>1, (z)>1\Re(z)>1. It follows that the rest of the conditions given in the definition of S4S_{4} are superseded by the above three conditions. Thus

S4={(s,w,z):(s)<0,(s+w)>1,(s+z)>1}.S_{4}=\{(s,w,z):\ \Re(s)<0,\ \Re(s+w)>1,\ \Re(s+z)>1\}.

We further note that the region S2S_{2} contains the subset given by

{(s,w,z):(z)>1,(s)>1,(s+2w)>2}.\{(s,w,z):\Re(z)>1,\ \Re(s)>1,\ \Re(s+2w)>2\}.

As the region S4S_{4} contains points (s,w,z)(s,w,z) such that

{(s)<0, 1<(s+w)<(z),(s)>1(z)},\{\Re(s)<0,\ 1<\Re(s+w)<\Re(z),\ \Re(s)>1-\Re(z)\},

it is then readily seen that the convex hull of the above regions contains S5{(s,w,z):(z)>1}S_{5}\cap\{(s,w,z):\Re(z)>1\}, where

S5={(s,w,z):\displaystyle S_{5}=\{(s,w,z):\ (s+2w)>2,(s+2z)>2,(s+z)>1,(s+w)>1,(w)>14,(z)>12}.\displaystyle\Re(s+2w)>2,\ \Re(s+2z)>2,\ \Re(s+z)>1,\ \Re(s+w)>1,\ \Re(w)>\tfrac{1}{4},\ \Re(z)>\tfrac{1}{2}\}.

On the other hand, when 1/2<(z)<11/2<\Re(z)<1, we note that the points in S2S_{2} are certainly contained in the convex hull of S2S_{2} and S4S_{4} and one checks S2{(s,w,z):(s)>1}=S5{(s,w,z):(s)>1}S_{2}\cap\{(s,w,z):\Re(s)>1\}=S_{5}\cap\{(s,w,z):\Re(s)>1\}. We may thus focus on the case (s)<1\Re(s)<1. In this case, we note that the points (1,1/2,1/2)(1,1/2,1/2), (1,1/2,1)(1,1/2,1) are in the closure of S2S_{2} and the point (0,1,1)(0,1,1) is in the closure of S4S_{4}. These three points determine a triangular region which can be regarded as the base of the region 𝒮\mathcal{S} enclosed by the four planes: (s+2w)=2\Re(s+2w)=2, (s)=1\Re(s)=1, (z)=1\Re(z)=1, (s+2z)=2\Re(s+2z)=2. It follows that the points in this triangular region are all in the convex hulls of S2S_{2} and S4S_{4}. Further note that the points on the boundary 𝒮{(z)=1}\mathcal{S}\cap\{\Re(z)=1\} of 𝒮\mathcal{S} are all in the convex hull of S2S_{2} and S4S_{4} since they can be identified with the set {(s,w,z):(z)=1,0(s)1,(s+2w)2}\{(s,w,z):\Re(z)=1,0\leq\Re(s)\leq 1,\Re(s+2w)\geq 2\} and hence is contained in the convex hull of S5{(s,w,z):(z)>1}S_{5}\cap\{(s,w,z):\Re(z)>1\}. We then deduce that the entire region 𝒮\mathcal{S} lies in the convex hull of S0S_{0} and S1S_{1}. We next note that when 1/2<(z)<11/2<\Re(z)<1, the condition (s+2z)>2\Re(s+2z)>2 implies that (s)>0\Re(s)>0 so that one also has (s+z)>1\Re(s+z)>1 and that the condition (s+2w)>2\Re(s+2w)>2 implies (s+w)>1\Re(s+w)>1. It follows from these discussions we see that the intersection of the convex hull of S2S_{2} and S4S_{4} thus contains S5S_{5}.

We apply Theorem 2.10 again to conclude that (s1)(w1)A(s,w,z;f)(s-1)(w-1)A(s,w,z;f) converges absolutely in the region S5S_{5}.

3.5. Bounding A(s,w,z;f)A(s,w,z;f) in vertical strips

In order to prove Theorem 1.1, we also need to estimate |A(s,w,z;f)||A(s,w,z;f)| in vertical strips.

We set for any fixed 0<δ<1/10000<\delta<1/1000 and the previously defined regions SjS_{j},

S~j=Sj,δ{(s,w,z):(s)>5/2,(w)>1/2δ},\widetilde{S}_{j}=S_{j,\delta}\cap\{(s,w,z):\Re(s)>-5/2,\ \Re(w)>1/2-\delta\},

where Sj,δ={(s,w,z)+δ(1,1,1):(s,w,z)Sj}S_{j,\delta}=\{(s,w,z)+\delta(1,1,1):(s,w,z)\in S_{j}\}. Set

p(s,w)=(s1)(w1),p(s,w)=(s-1)(w-1),

so that p(s,w)A(s,w,z;f)p(s,w)A(s,w,z;f) is an analytic function in the regions under our consideration. We also write p~(s,w)=1+|p(s,w)|\tilde{p}(s,w)=1+|p(s,w)|.

We consider the bound for A(s,w,z;f)A(s,w,z;f) given in (3.4) and apply (2.17) to deduce that, under GRH, in S~0\widetilde{S}_{0},

|p(s,w)A(s,w,z;f)|p~(s,w)|wz|ε(1+|w|)max{12(w),0}+ε.\displaystyle\begin{split}|p(s,w)A(s,w,z;f)|\ll\tilde{p}(s,w)|wz|^{\varepsilon}(1+|w|)^{\max\{1-2\Re(w),0\}+\varepsilon}.\end{split}

Similarly, using the estimates (2.18) in (3.5) renders that, under GRH, in the region S~1\widetilde{S}_{1},

|p(s,w)A(s,w,z;f)|p~(s,w)(1+|s|)max{1/2(s),0}+ε.\displaystyle|p(s,w)A(s,w,z;f)|\ll\tilde{p}(s,w)(1+|s|)^{\max\{1/2-\Re(s),0\}+\varepsilon}.

We then deduce from the above and Proposition 2.11 that in the convex hull S~2\widetilde{S}_{2} of S~0\widetilde{S}_{0} and S~1\widetilde{S}_{1}, we have under GRH,

(3.33) |p(s,w)A(s,w,z;f)|p~(s,w)|wz|ε(1+|w|)max{12(w),0}+ε(1+|s|)3+ε.|p(s,w)A(s,w,z;f)|\ll\tilde{p}(s,w)|wz|^{\varepsilon}(1+|w|)^{\max\{1-2\Re(w),0\}+\varepsilon}(1+|s|)^{3+\varepsilon}.

Moreover, by (3.11) and the estimations given in (2.18) for ζ(s)\zeta(s) (corresponding to the case ψ=ψ0\psi=\psi_{0} being the primitive principal character) and in (2.19) for L(2)(2w,sym2f)L^{(2)}(2w,\text{sym}^{2}f) that in the region S~3\widetilde{S}_{3}, under GRH,

(3.34) |A1(s,w,z;f)||w|ε(1+|s|)max{(1(s))/2,1/2(s),0}+ε.\displaystyle|A_{1}(s,w,z;f)|\ll|w|^{\varepsilon}(1+|s|)^{\max\{(1-\Re(s))/2,1/2-\Re(s),0\}+\varepsilon}.

Also, by (3.16)–(3.21) and Lemma 3.4,

(3.35) |C(s,w,z;f)||wz|ε|C(s,w,z;f)|\ll|wz|^{\varepsilon}

in the region

{(s,w,z):(w)>1+ε,(z)>1+ε,(s)>1+ε}.\{(s,w,z):\Re(w)>1+\varepsilon,\ \Re(z)>1+\varepsilon,\ \Re(s)>1+\varepsilon\}.

Now, applying (3.14), the functional equation (3.15), the bounds given in (3.34), (3.35), together with (2.15), we obtain that in the region S~4\widetilde{S}_{4},

(3.36) |p(s,w)A(s,w,z;f)|p~(s,w)|wz|ε(1+|w|)max{3(1/2(w)),0}+ε(1+|s|)3+ε.|p(s,w)A(s,w,z;f)|\ll\tilde{p}(s,w)|wz|^{\varepsilon}(1+|w|)^{\max\{3(1/2-\Re(w)),0\}+\varepsilon}(1+|s|)^{3+\varepsilon}.

We now conclude from (3.33), (3.36) 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},

(3.37) |p(s,w)A(s,w,z;f)|p~(s,w)|wz|ε(1+|w|)max{3(1/2(w)),0}+ε(1+|s|)3+ε.|p(s,w)A(s,w,z;f)|\ll\tilde{p}(s,w)|wz|^{\varepsilon}(1+|w|)^{\max\{3(1/2-\Re(w)),0\}+\varepsilon}(1+|s|)^{3+\varepsilon}.

3.6. Completion of proof

The Mellin inversion yields that

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

where A(s,w,z;f)A(s,w,z;f) is defined in (3.1) and where we recall that w^\widehat{w} is the Mellin transform of ww defined by

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

Now repeated integration by parts gives that for any integer E0E\geq 0,

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

We evaluate the integral in (3.38) by shifting the line of integration to (s)=N(α,β)+ε\Re(s)=N(\alpha,\beta)+\varepsilon, where N(α,β)N(\alpha,\beta) is given in (1.5). Applying (3.37) and (3.39) gives that the integral on the new line can be absorbed into the OO-term in (1.3). We encounter a simple pole at s=1s=1 in the process whose residue is given in (3.13). This yields the main terms in (1.3) and completes the proof of Theorem 1.1.

Acknowledgments. P. G. is supported in part by NSFC grant 11871082 and L. Z. by the FRG Grant PS43707 at the University of New South Wales.

References