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

Real zeros of quadratic Hecke LL-functions

Peng Gao School of Mathematical Sciences, Beihang University, Beijing 100191, P. R. China penggao@buaa.edu.cn
Abstract.

We study real zeros of a family of quadratic Hecke LL-functions in the Gaussian field to show that more than twenty percent of the members have no zeros on the interval (0,1](0,1].

Mathematics Subject Classification (2010): 11M06, 11M20, 11M41, 11R42

Keywords: Hecke LL-functions, quadratic Hecke characters, mollifier, real zeros

1. Introduction

The non-vanishing issue of central values of LL-functions has been studied intensively in the literature due to its deep arithmetic implications. For the classical case, it is expected that the corresponding L(s,χ)L(s,\chi) never vanishes at s=1/2s=1/2 for any Dirichlet character χ\chi. This statement first appeared as a conjecture of S. Chowla [Chow], who concerned with the special case of primitive real characters.

Partial resolutions to Chowla’s conjecture was first given by M. Jutila [Jutila], who evaluated the first two moments of the family of quadratic Dirichlet LL-functions to show that infinitely many such LL-functions do not vanish at the central value. By evaluating mollified moments instead, K. Soundararajan [sound1] improved the result of Jutila to show that unconditionally, at least 87.5%87.5\% of the members of the quadratic family have non-vanishing central values. This percentage can be further improved to be 94.27%94.27\% if one assumes the generalized Riemann hypothesis (GRH). In fact, this follows from a result of A. E. Özluk and C. Snyder [O&S] on the one-level density of low-lying zeros of the family of quadratic Dirichlet LL-functions, replacing the test function used in [O&S] by an optimal one given in a paper of H. Iwaniec, W. Luo and P. Sarnak [ILS, Appendix A] .

The statements of GRH and Chowla’s conjecture imply that L(σ,χ)0L(\sigma,\chi)\neq 0 for every primitive quadratic Dirichlet character and all 0σ10\leq\sigma\leq 1. However, it was not even known previously whether every such Dirichlet LL-function of sufficiently large conductor has a non-trivial real zero or not until J. B. Conrey and K. Soundararajan [C&S] proved in 2002 that for at least 20%20\% of the odd square-free integers d0d\geq 0 we have L(σ,χ8d)>L(\sigma,\chi_{-8d})> 0 for 0σ10\leq\sigma\leq 1, where χ8d=(8d)\chi_{-8d}=\left(\frac{-8d}{\cdot}\right) is the Kronecker symbol.

Motivated by the above result of Conrey and Soundararajan, we investigate in this paper the real zeros of a family of quadratic Hecke LL-functions in the Gaussian number field (i)\mathbb{Q}(i) . To state our result, we first introduce some notations. Throughout the paper, we denote K=(i)K=\mathbb{Q}(i) and 𝒪K\mathcal{O}_{K} for the ring of integers [i]\mathbb{Z}[i] of KK. For any cKc\in K, we write N(c)N(c) for its norm. We denote L(s,χ),ζK(s)L(s,\chi),\zeta_{K}(s) for the LL-function associated to a Hecke character χ\chi of KK and the Dedekind zeta function of KK, respectively.

We say a Hecke character χ\chi is primitive modulo qq if it does not factor through (𝒪K/(q))×\left(\mathcal{O}_{K}/(q^{\prime})\right)^{\times} for any divisor qq^{\prime} of qq such that N(q)<N(q)N(q^{\prime})<N(q). We also say that χ\chi is of trivial infinite type if its component at infinite places of KK is trivial. For any c𝒪Kc\in\mathcal{O}_{K}, we reserve in this paper the symbol χc\chi_{c} for the quadratic residue symbol (c)\left(\frac{c}{\cdot}\right) defined in Section 2.1. It is shown in [Gao2, Section 2.1] that χ(1+i)5d\chi_{(1+i)^{5}d} defines a primitive quadratic Hecke character modulo (1+i)5d(1+i)^{5}d of trivial infinite type for any odd, square-free d𝒪Kd\in\mathcal{O}_{K}. Here, we say that dd is odd if (d,2)=1(d,2)=1 and dd is square-free if the ideal (d)(d) is not divisible by the square of any prime ideal. The family of LL-functions that we consider in this paper is given by

={L(s,χ(1+i)5d):d𝒪Kodd, square-free}.\displaystyle\mathcal{F}=\big{\{}L(s,\chi_{(1+i)^{5}d}):d\in\mathcal{O}_{K}\hskip 3.61371pt\text{odd, square-free}\big{\}}.

Notice that every LL-function in the above family satisfies a functional equation given by (2.4) which implies that Γ(s)L(s,χ(1+i)5d)\Gamma(s)L(s,\chi_{(1+i)^{5}d}) is entire on the whole complex plane. As Γ(s)\Gamma(s) has a simple pole at s=0s=0, this implies that L(0,χ(1+i)5d)=0L(0,\chi_{(1+i)^{5}d})=0. Thus, if one assumes GRH for the above family of LL-functions as well as the non-vanishing of these LL-functions at the central point 1/21/2, then one can deduce that L(σ,(1+i)5d)0L(\sigma,(1+i)^{5}d)\neq 0 for all 0<σ10<\sigma\leq 1. Consequently, we have L(σ,(1+i)5d)>0L(\sigma,(1+i)^{5}d)>0 for all 0<σ10<\sigma\leq 1 by continuity and the fact that L(2,(1+i)5d)>0L(2,(1+i)^{5}d)>0. In this paper, we show that unconditionally, at least a positive percent of these LL-functions do not vanish on (0,1](0,1]. By further keeping in mind that the number of odd, square-free d𝒪Kd\in\mathcal{O}_{K} such that N(d)xN(d)\leq x is asymptotically 2πx3ζK(2)\frac{2\pi x}{3\zeta_{K}(2)} (see [G&Zhao4, Section 3.1]), our result is given as follows.

Theorem 1.1.

For at least 20%20\% of the members of LL-functions in \mathcal{F}, we have L(σ,χ(1+i)5d)>0L(\sigma,\chi_{(1+i)^{5}d})>0 for 0<σ10<\sigma\leq 1. More precisely, the number of LL-functions in \mathcal{F} such that N(d)xN(d)\leq x and L(σ,χ(1+i)5d)>0L(\sigma,\chi_{(1+i)^{5}d})>0 for all 0<σ10<\sigma\leq 1 exceeds 15(2πx3ζK(2))\frac{1}{5}(\frac{2\pi x}{3\zeta_{K}(2)}) for all large xx.

Our proof of Theorem 1.1 is mainly based on the approach of Conrey and Soundararajan in their above mentioned work [C&S] concerning the non-vanishing of quadratic Dirichlet LL-functions on the real line. We also note here that similar to what is pointed out in [C&S], the approach in our paper implies that for all large xx, the number of odd, square-free d𝒪Kd\in\mathcal{O}_{K} with N(d)xN(d)\leq x such that L(s,χ(1+i)5d)L(s,\chi_{(1+i)^{5}d}) has a zero in [σ,1][\sigma,1] is x1(1ε)(σ12)\ll x^{1-(1-\varepsilon)(\sigma-\frac{1}{2})} for any fixed σ1/2\sigma\geq 1/2 and ε>0\varepsilon>0. Another outcome of our approach is that there are many LL-functions having no non-trivial zeros in a thin rectangle containing the real line. More precisely, there exists a constant c>0c>0 such that for at least 20%20\% of the members of \mathcal{F} with N(d)xN(d)\leq x, the corresponding L(s,χ(1+i)5d)L(s,\chi_{(1+i)^{5}d}) has no zeros in the rectangle {σ+it:σ(0,1],|t|c/logx}\{\sigma+it:\ \ \sigma\in(0,1],\ \ |t|\leq c/\log x\} for all large xx.

2. Preliminaries

The proof of our result requires many tools as well as auxiliary results, which we gather them here first.

2.1. Quadratic residue symbol and Gauss sum

It is well-known that K=(i)K=\mathbb{Q}(i) has class number one and that every ideal in 𝒪K\mathcal{O}_{K} co-prime to 22 has a unique generator which is 1mod(1+i)3\equiv 1\bmod(1+i)^{3}. Such a generator is called primary. Further notice that (1+i)(1+i) is the only prime ideal in 𝒪K\mathcal{O}_{K} that lies above the ideal (2)(2)\in\mathbb{Z}. We now fix a generator nn for each ideal (n)(n) in 𝒪K\mathcal{O}_{K} by taking nn to be of the form (1+i)mn(1+i)^{m}n^{\prime} with m1m\geq 1 and nn^{\prime} primary. We denote the set of these generators by GG throughout the paper. For a,b𝒪Ka,b\in\mathcal{O}_{K}, we denote (a,b)(a,b) for their greatest common divisor such that (a,b)G(a,b)\in G.

We denote ϖ\varpi for a prime element in 𝒪K\mathcal{O}_{K}, which means that the ideal generated by ϖ\varpi is a prime ideal. We denote the group of units in 𝒪K\mathcal{O}_{K} by UKU_{K} and the discriminant of KK by DKD_{K}. Thus, we have UK={±1,±i}U_{K}=\{\pm 1,\pm i\} and DK=4D_{K}=-4. We write d=d=\square for a perfect square in 𝒪K\mathcal{O}_{K}, by which we mean that d=n2d=n^{2} for some n𝒪Kn\in\mathcal{O}_{K}.

The quadratic residue symbol (ϖ)\left(\frac{\cdot}{\varpi}\right) is defined for any odd prime ϖ\varpi such that for any a𝒪Ka\in\mathcal{O}_{K}, (aϖ)=0\left(\frac{a}{\varpi}\right)=0 when ϖ|a\varpi|a and (aϖ)a(N(ϖ)1)/2(modϖ)\left(\frac{a}{\varpi}\right)\equiv a^{(N(\varpi)-1)/2}\pmod{\varpi} with (aϖ){±1}\left(\frac{a}{\varpi}\right)\in\{\pm 1\} when (a,ϖ)=1(a,\varpi)=1. The definition is then extended to (n)\left(\frac{\cdot}{n}\right) multiplicatively for any odd n𝒪Kn\in\mathcal{O}_{K}. Here, we define (n)=1\left(\frac{\cdot}{n}\right)=1 for nUKn\in U_{K}.

As mentioned in Section 1, we denote χc\chi_{c} for the quadratic residue symbol (c)\left(\frac{c}{\cdot}\right). In this paper, we regard χ±1\chi_{\pm 1} as a principal character modulo 11, so that χ±1(a)=1\chi_{\pm 1}(a)=1 for all a𝒪Ka\in\mathcal{O}_{K}. Note that this implies that L(s,χ±1)=ζK(s)L(s,\chi_{\pm 1})=\zeta_{K}(s). For other values of cc, we shall regard χc\chi_{c} as a Hecke character of trivial infinite type modulo 16c16c as this is justified in [G&Zhao2019, Section 2.1]. In particular, we notice that (ca)=0\left(\frac{c}{a}\right)=0 when 1+i|a1+i|a and c±1c\neq\pm 1.

We define the quadratic Gauss sum g(r,n)g(r,n) for r,n𝒪Kr,n\in\mathcal{O}_{K} with nn being odd, by

g(r,n)=xmodn(xn)e~(rxn),\displaystyle g(r,n)=\sum_{x\bmod{n}}\left(\frac{x}{n}\right)\widetilde{e}\left(\frac{rx}{n}\right),

where we define for any complex number zz,

e~(z)=exp(2πi(z2iz¯2i)).\displaystyle\widetilde{e}(z)=\exp\left(2\pi i\left(\frac{z}{2i}-\frac{\bar{z}}{2i}\right)\right).

We denote μ[i]\mu_{[i]} for the analogue on 𝒪K\mathcal{O}_{K} of the usual Möbius function on \mathbb{Z} and φ[i](n)\varphi_{[i]}(n) for the number of elements in the reduced residue class of 𝒪K/(n)\mathcal{O}_{K}/(n). We recall the following explicitly evaluations of g(r,n)g(r,n) given in [G&Zhao4, Lemma 2.2].

Lemma 2.2.
  1. (i)

    We have

    g(rs,n)\displaystyle g(rs,n) =(sn)¯g(r,n),(s,n)=1,\displaystyle=\overline{\left(\frac{s}{n}\right)}g(r,n),\qquad(s,n)=1,
    g(k,mn)\displaystyle g(k,mn) =g(k,m)g(k,n),m,n primary and (m,n)=1.\displaystyle=g(k,m)g(k,n),\qquad m,n\text{ primary and }(m,n)=1.
  2. (ii)

    Let ϖ\varpi be a primary prime in 𝒪K\mathcal{O}_{K}. Suppose ϖh\varpi^{h} is the largest power of ϖ\varpi dividing kk. (If k=0k=0 then set h=h=\infty.) Then for l1l\geq 1,

    g(k,ϖl)\displaystyle g(k,\varpi^{l}) ={0iflhis odd,φ[i](ϖl)iflhis even,N(ϖ)l1ifl=h+1is even,(ikϖhϖ)N(ϖ)l1/2ifl=h+1is odd,0iflh+2.\displaystyle=\begin{cases}0\qquad&\text{if}\qquad l\leq h\qquad\text{is odd},\\ \varphi_{[i]}(\varpi^{l})\qquad&\text{if}\qquad l\leq h\qquad\text{is even},\\ -N(\varpi)^{l-1}&\text{if}\qquad l=h+1\qquad\text{is even},\\ \left(\frac{ik\varpi^{-h}}{\varpi}\right)N(\varpi)^{l-1/2}\qquad&\text{if}\qquad l=h+1\qquad\text{is odd},\\ 0\qquad&\text{if}\qquad l\geq h+2.\end{cases}

2.3. The approximate functional equation

For any primitive quadratic Hecke character χ\chi of KK of trivial infinite type, a well-known result of E. Hecke asserts that L(s,χ)L(s,\chi) has an analytic continuation to the entire complex plane with a simple pole at s=1s=1 only when χ\chi is principal. Moreover, L(s,χ)L(s,\chi) satisfies the following functional equation (see [iwakow, Theorem 3.8])

(2.1) Λ(s,χ)=W(χ)(N(m))1/2Λ(1s,χ),\displaystyle\Lambda(s,\chi)=W(\chi)(N(m))^{-1/2}\Lambda(1-s,\chi),

where mm is the conductor of χ\chi, |W(χ)|=(N(m))1/2|W(\chi)|=(N(m))^{1/2} and

(2.2) Λ(s,χ)=(|DK|N(m))s/2(2π)sΓ(s)L(s,χ).\displaystyle\Lambda(s,\chi)=(|D_{K}|N(m))^{s/2}(2\pi)^{-s}\Gamma(s)L(s,\chi).

In particular, we have the following functional equation for ζK(s)\zeta_{K}(s):

(2.3) πsΓ(s)ζK(s)=π(1s)Γ(1s)ζK(1s).\displaystyle\pi^{-s}\Gamma(s)\zeta_{K}(s)=\pi^{-(1-s)}\Gamma(1-s)\zeta_{K}(1-s).

When χ=χ(1+i)5d\chi=\chi_{(1+i)^{5}d} for any odd, square-free d𝒪Kd\in\mathcal{O}_{K}, we combine [iwakow, Theorem 3.8] and [Gao2, Lemma 2.2] to see that W(χ(1+i)5d)=g(χ(1+i)5d)=N((1+i)5d)W(\chi_{(1+i)^{5}d})=g(\chi_{(1+i)^{5}d})=\sqrt{N((1+i)^{5}d)}. Thus, the functional equation (2.1) in this case becomes

(2.4) Λ(s,χ(1+i)5d)=Λ(1s,χ(1+i)5d).\displaystyle\Lambda(s,\chi_{(1+i)^{5}d})=\Lambda(1-s,\chi_{(1+i)^{5}d}).

We now set

(2.5) ξ(s,χ(1+i)5d)=(25N(d)π2)s214Γ(s)L(s,χ(1+i)5d).\displaystyle\xi(s,\chi_{(1+i)^{5}d})=\biggl{(}\frac{2^{5}N(d)}{\pi^{2}}\biggr{)}^{\frac{s}{2}-\frac{1}{4}}\Gamma(s)L(s,\chi_{(1+i)^{5}d}).

It follows from (2.2) and (2.4) that we have the following functional equation

(2.6) ξ(s,χ(1+i)5d)=ξ(1s,χ(1+i)5d).\displaystyle\xi(s,\chi_{(1+i)^{5}d})=\xi(1-s,\chi_{(1+i)^{5}d}).

Throughout the paper, we fix a positive real number κ1/100\kappa\leq 1/100 and let δ1\delta_{1} and δ2\delta_{2} be two complex numbers satisfying max(|δ1|,|δ2)|)κ\max(|\delta_{1}|,|\delta_{2})|)\leq\kappa. We further denote τ=δ1+δ22\tau=\frac{\delta_{1}+\delta_{2}}{2}, δ=δ1δ22\delta=\frac{\delta_{1}-\delta_{2}}{2} so that we have max(|τ|,|δ)|)κ\max(|\tau|,|\delta)|)\leq\kappa as well. We define for real numbers x>0,c>|(τ)|x>0,c>|\Re(\tau)|,

(2.7) Wδ,τ(x)=12πi(c)Γδ(s)xs2ss2τ2𝑑s,\displaystyle W_{\delta,\tau}(x)=\frac{1}{2\pi i}\int\limits_{(c)}\Gamma_{\delta}(s)x^{-s}\frac{2s}{s^{2}-\tau^{2}}ds,

where

Γδ(s)=Γ(12+s+δ)Γ(12+sδ).\Gamma_{\delta}(s)=\Gamma\Big{(}\frac{1}{2}+s+\delta\Big{)}\Gamma\Big{(}\frac{1}{2}+s-\delta\Big{)}.

The following Lemma establishes some analytical properties concerning Wδ,τ(x)W_{\delta,\tau}(x).

Lemma 2.4.

The function Wδ,τ(x)W_{\delta,\tau}(x) is a smooth complex-valued function for x(0,)x\in(0,\infty). For xx near 0, we have

(2.8) Wδ,τ(x)=Γδ(τ)xτ+Γδ(τ)xτ+O(x1/4ϵ).\displaystyle W_{\delta,\tau}(x)=\Gamma_{\delta}(\tau)x^{-\tau}+\Gamma_{\delta}(-\tau)x^{\tau}+O(x^{1/4-\epsilon}).

For large xx and any integer ν\nu, we have

(2.9) Wδ,τ(ν)(x)νxν+3e2x1/2νex1/2.\displaystyle W_{\delta,\tau}^{(\nu)}(x)\ll_{\nu}x^{\nu+3}e^{-2x^{1/2}}\ll_{\nu}e^{-x^{1/2}}.
Proof.

We move the line of integration in (2.7) to (s)=1/4+ϵ\Re(s)=-1/4+\epsilon to deduce readily (2.8). On the other hand, for any c>|(τ)|c>|\Re(\tau)|, we have

Wδ,τ(ν)(x)=(1)ν2πi(c)Γδ(s)s(s+1)(s+ν1)xsν2ss2τ2𝑑s.W_{\delta,\tau}^{(\nu)}(x)=\frac{(-1)^{\nu}}{2\pi i}\int\limits_{(c)}\Gamma_{\delta}(s)s(s+1)\cdots(s+\nu-1)x^{-s-\nu}\frac{2s}{s^{2}-\tau^{2}}ds.

This implies that Wδ,τ(x)W_{\delta,\tau}(x) is smooth. To prove (2.9), we may suppose that x>ν+4x>\nu+4. Using the facts that |Γ(x+iy)|Γ(x)|\Gamma(x+iy)|\leq\Gamma(x) for x1x\geq 1 and sΓ(s)=Γ(s+1)s\Gamma(s)=\Gamma(s+1), we see that

Wδ,τ(ν)(x)ν|Γ(c+ν+3)|2xc(c)|ds||s2τ2|νΓ(c+ν+3)2xcc|(τ)|ν(c+ν+3e)2c+2ν+6xcc|(τ)|,W_{\delta,\tau}^{(\nu)}(x)\ll_{\nu}|\Gamma(c+\nu+3)|^{2}x^{-c}\int\limits_{(c)}\frac{|ds|}{|s^{2}-\tau^{2}|}\ll_{\nu}\Gamma(c+\nu+3)^{2}\frac{x^{-c}}{c-|\Re(\tau)|}\ll_{\nu}\Big{(}\frac{c+\nu+3}{e}\Big{)}^{2c+2\nu+6}\frac{x^{-c}}{c-|\Re(\tau)|},

where the last estimation above follows from Stirling’s formula (see [iwakow, (5.113)]). The last assertion of the lemma now follows by taking c=x1/2ν3(2)c=x^{1/2}-\nu-3(\geq 2) above. ∎

For any complex number ss and any primary n𝒪Kn\in\mathcal{O}_{K}, we define

rs(n)=ab=na,b1mod(1+i)3(N(a)N(b))s.r_{s}(n)=\sum_{\begin{subarray}{c}ab=n\\ a,b\equiv 1\bmod(1+i)^{3}\end{subarray}}\Big{(}\frac{N(a)}{N(b)}\Big{)}^{s}.

Note that rs(n)r_{s}(n) is easily seen to be an even function of ss. We also define for any odd, square-free d𝒪Kd\in\mathcal{O}_{K},

Aδ,τ(d)=n1mod(1+i)3rδ(n)N(n)((1+i)5dn)Wδ,τ(π2N(n)25N(d)).A_{\delta,\tau}(d)=\sum_{n\equiv 1\bmod(1+i)^{3}}\frac{r_{\delta}(n)}{\sqrt{N(n)}}\left(\frac{(1+i)^{5}d}{n}\right)W_{\delta,\tau}\left(\frac{\pi^{2}N(n)}{2^{5}N(d)}\right).

Now, we present an approximate functional equation for ξ(12+δ1,χ(1+i)5d)ξ(12+δ2,χ(1+i)5d)\xi(\frac{1}{2}+\delta_{1},\chi_{(1+i)^{5}d})\xi(\frac{1}{2}+\delta_{2},\chi_{(1+i)^{5}d}).

Lemma 2.5.

With the notations above for δ1\delta_{1} and δ2\delta_{2}. We have for any odd, square-free d𝒪Kd\in\mathcal{O}_{K},

ξ(12+δ1,χ(1+i)5d)ξ(12+δ2,χ(1+i)5d)=Aδ,τ(d).\xi(\tfrac{1}{2}+\delta_{1},\chi_{(1+i)^{5}d})\xi(\tfrac{1}{2}+\delta_{2},\chi_{(1+i)^{5}d})=A_{\delta,\tau}(d).
Proof.

We begin with the following integral for some 3/2|(δ)|>c>1/2+|(δ)|3/2-|\Re(\delta)|>c>1/2+|\Re(\delta)|,

12πi(c)ξ(12+δ+s,χ(1+i)5d)ξ(12δ+s,χ(1+i)5d)2ss2τ2𝑑s.\frac{1}{2\pi i}\int\limits_{(c)}\xi(\tfrac{1}{2}+\delta+s,\chi_{(1+i)^{5}d})\xi(\tfrac{1}{2}-\delta+s,\chi_{(1+i)^{5}d})\frac{2s}{s^{2}-\tau^{2}}ds.

We evaluate the above integral by first writing L(12+δ+s,χ(1+i)5d)L(12δ+s,χ(1+i)5d)L(\tfrac{1}{2}+\delta+s,\chi_{(1+i)^{5}d})L(\tfrac{1}{2}-\delta+s,\chi_{(1+i)^{5}d}) as a Dirichlet series n1mod(1+i)3rδ(n)N(n)12+s((1+i)5dn)\displaystyle{\sum_{n\equiv 1\bmod(1+i)^{3}}}\frac{r_{\delta}(n)}{N(n)^{\frac{1}{2}+s}}\left(\frac{(1+i)^{5}d}{n}\right) to see that the above expression equals Aδ,τ(d)A_{\delta,\tau}(d) upon integrating term by term. On the other hand, we move the line of integration to (s)=c\Re(s)=-c to encounter poles at s=τs=\tau, τ-\tau. The residues contribute ξ(12+δ+τ,χ(1+i)5d)ξ(12δ+τ,χ(1+i)5d)+ξ(12+δτ,χ(1+i)5d)ξ(12δτ,χ(1+i)5d)=2ξ(12+δ1,χ(1+i)5d)ξ(12+δ2,χ(1+i)5d)\xi(\frac{1}{2}+\delta+\tau,\chi_{(1+i)^{5}d})\xi(\frac{1}{2}-\delta+\tau,\chi_{(1+i)^{5}d})+\xi(\frac{1}{2}+\delta-\tau,\chi_{(1+i)^{5}d})\xi(\frac{1}{2}-\delta-\tau,\chi_{(1+i)^{5}d})=2\xi(\frac{1}{2}+\delta_{1},\chi_{(1+i)^{5}d})\xi(\frac{1}{2}+\delta_{2},\chi_{(1+i)^{5}d}) by the functional equation (2.6). Using (2.6) again, we see that the remaining integral on the c-c line equals to Aδ,τ(d)-A_{\delta,\tau}(d) via a change of variable sss\to-s. This leads to the desired result. ∎

2.6. Quadratic large sieves

We include in this section two large sieve results concerning quadratic Hecke characters. They are generalizations in KK of the well-known large sieve results due to D. R. Heath-Brown [DRHB] on quadratic Dirichlet characters. The first lemma can be obtained by applying a large sieve result of K. Onodera [Onodera] on quadratic residue symbols in KK in the proof of [DRHB, Corollary 2] and [sound1, Lemma 2.4].

Lemma 2.7.

Let N,QN,Q be positive integers, and let a1,,ana_{1},\cdots,a_{n} be arbitrary complex numbers. Let S(Q)S(Q) denote the set of χm\chi_{m} for square-free mm satisfying N(m)QN(m)\leq Q. Then for any ϵ>0\epsilon>0,

χS(Q)|n1mod(1+i)3N(n)Nanχ(n)|2ϵ(QN)ϵ(Q+N)n1,n21mod(1+i)3N(n1),N(n2)Nn1n2=|an1an2|.\displaystyle\sum_{\chi\in S(Q)}\Big{|}\sum_{\begin{subarray}{c}n\equiv 1\bmod{(1+i)^{3}}\\ N(n)\leq N\end{subarray}}a_{n}\chi(n)\Big{|}^{2}\ll_{\epsilon}(QN)^{\epsilon}(Q+N)\sum_{\begin{subarray}{c}n_{1},n_{2}\equiv 1\bmod{(1+i)^{3}}\\ N(n_{1}),N(n_{2})\leq N\\ n_{1}n_{2}=\square\end{subarray}}|a_{n_{1}}a_{n_{2}}|.

Let MM be a positive integer, and for each m𝒪Km\in\mathcal{O}_{K} satisfying N(m)MN(m)\leq M, we write m=m1m22m=m_{1}m_{2}^{2} with m1m_{1} square-free and m2Gm_{2}\in G. Suppose the sequence ana_{n} satisfies |an|N(n)ε|a_{n}|\ll N(n)^{\varepsilon}, then

N(m)M1N(m2)|N(n)Nan(mn)|2(MN)εN(M+N).\displaystyle\sum_{N(m)\leq M}\frac{1}{N(m_{2})}\left|\sum_{N(n)\leq N}a_{n}\left(\frac{m}{n}\right)\right|^{2}\ll(MN)^{\varepsilon}N(M+N).

Similarly, combining the above result of Onodera with the derivation of [DRHB, Theorem 2] and [sound1, Lemma 2.5], we have the following result.

Lemma 2.8.

Let S(Q)S(Q) be as in Lemma 2.7. For any complex number σ+it\sigma+it with σ12\sigma\geq\frac{1}{2}, we have

χS(Q)|L(σ+it,χ(1+i)5d)|4\displaystyle\sum_{\chi\in S(Q)}|L(\sigma+it,\chi_{(1+i)^{5}d})|^{4}\ll Q1+ε(1+|t|2)1+ε\displaystyle Q^{1+\varepsilon}(1+|t|^{2})^{1+\varepsilon}
χS(Q)|L(σ+it,χ(1+i)5d)|2\displaystyle\sum_{\chi\in S(Q)}|L(\sigma+it,\chi_{(1+i)^{5}d})|^{2}\ll Q1+ε(1+|t|2)1/2+ε.\displaystyle Q^{1+\varepsilon}(1+|t|^{2})^{1/2+\varepsilon}.

2.9. Poisson summation

We recall the following two dimensional Poisson summation formula, which follows from [G&Zhao4, Lemma 2.7, Corollary 2.8].

Lemma 2.10.

Let n𝒪Kn\in\mathcal{O}_{K} be primary and (n)\left(\frac{\cdot}{n}\right) be the quadratic residue symbol modulo nn. For any smooth function W:+W:\mathbb{R}^{+}\rightarrow\mathbb{R} of compact support, we have for X>0X>0,

m𝒪K(m,1+i)=1(mn)W(N(m)X)=X2N(n)(1+in)k𝒪K(1)N(k)g(k,n)W~(N(k)X2N(n)),\displaystyle\sum_{\begin{subarray}{c}m\in\mathcal{O}_{K}\\ (m,1+i)=1\end{subarray}}\left(\frac{m}{n}\right)W\left(\frac{N(m)}{X}\right)=\frac{X}{2N(n)}\left(\frac{1+i}{n}\right)\sum_{k\in\mathcal{O}_{K}}(-1)^{N(k)}g(k,n)\widetilde{W}\left(\sqrt{\frac{N(k)X}{2N(n)}}\right),

where

(2.10) W~(t)=\displaystyle\widetilde{W}(t)= W(N(x+yi))e~(t(x+yi))dxdy,t0.\displaystyle\int\limits^{\infty}_{-\infty}\int\limits^{\infty}_{-\infty}W(N(x+yi))\widetilde{e}\left(-t(x+yi)\right)\mathrm{d}x\mathrm{d}y,\quad t\geq 0.

When applying the above lemma in the proof of our result, we are led to consider the behaviors of a particular function. In the rest of this section, we include a result on this. We begin by defining

(2.11) Fy(t)=Ψ(t)Wδ,τ(π2y25Xt),F_{y}(t)=\Psi(t)W_{\delta,\tau}\left(\frac{\pi^{2}y}{2^{5}Xt}\right),

where Ψ(t)\Psi(t) is a smooth function compacted in [1,2][1,2] and Wδ,τW_{\delta,\tau} is given in (2.7).

The function that we are interested is then defined for ξ>0\xi>0 and (w)>0\Re(w)>0 by

(2.12) h(ξ,w)=0F~t((ξt)1/2)tw1𝑑t.h(\xi,w)=\int_{0}^{\infty}\widetilde{F}_{t}\left(\left(\frac{\xi}{t}\right)^{1/2}\right)t^{w-1}\,dt.

Thus, h(ξ,w)h(\xi,w) is the Mellin transform of the function F~t((ξ/t)1/2)\widetilde{F}_{t}\left((\xi/t)^{1/2}\right). Here we recall that the Mellin transform g^(s)\widehat{g}(s) of a function gg is given by

g^(s)=0g(t)ts1𝑑t.\displaystyle\widehat{g}(s)=\int_{0}^{\infty}g(t)t^{s-1}dt.

For further reference, we note that if we further assume that gg is support in [1,2][1,2], then integration by parts implies that for (s)>0\Re(s)>0, we have

(2.13) |g^(s)|2(s)|s|ng(n),\displaystyle|\widehat{g}(s)|\ll\frac{2^{\Re(s)}}{|s|^{n}}g_{(n)},

where we define for integers n0n\geq 0,

g(n)=max0jn12|g(j)(t)|𝑑t.g_{(n)}=\max_{0\leq j\leq n}\int_{1}^{2}|g^{(j)}(t)|dt.

Now, we have the following result concerning analytical properties of h(ξ,w)h(\xi,w).

Lemma 2.11.

Let FtF_{t} be defined by (2.11) and let ξ>0\xi>0. The function h(ξ,w)h(\xi,w) defined in (2.12) is an entire function of ww in 1(w)>1+|(τ)|1\geq\Re(w)>-1+|\Re(\tau)| such that for any cc with 1+(w)>c>max(|(τ)|,(w))1+\Re(w)>c>\max(|\Re(\tau)|,\Re(w)),

(2.14) h(ξ,w)=Ψ^(1+w)ξwπ2πi(c)(25/2π)2sΓδ(s)(Xξ)sπ2s+2wΓ(sw)Γ(1s+w)2ss2τ2𝑑s.\displaystyle\begin{split}h(\xi,w)=&\widehat{\Psi}(1+w)\xi^{w}\frac{\pi}{2\pi i}\int\limits\limits_{(c)}\left(\frac{2^{5/2}}{\pi}\right)^{2s}\Gamma_{\delta}(s)\left(\frac{X}{\xi}\right)^{s}\pi^{-2s+2w}\frac{\Gamma(s-w)}{\Gamma(1-s+w)}\frac{2s}{s^{2}-\tau^{2}}ds.\end{split}

Moreover, when 1(w)>1+|(τ)|1\geq\Re(w)>-1+|\Re(\tau)|, we have

(2.15) h(ξ,w)(1+|w|)32(w)exp(110ξ1/4X1/4(|w|+1)1/2)ξ(w)|Ψ^(1+w)|.\displaystyle h(\xi,w)\ll(1+|w|)^{3-2\Re(w)}\exp\Bigg{(}-\frac{1}{10}\frac{\xi^{1/4}}{X^{1/4}(|w|+1)^{1/2}}\Bigg{)}\xi^{\Re(w)}|\widehat{\Psi}(1+w)|.
Proof.

Notice that for any smooth function WW, we can evaluate the function W~(t)\widetilde{W}(t) defined in (2.10) in polar coordinates as

W~(t)=\displaystyle\widetilde{W}(t)= 40π/20cos(2πtrsinθ)W(r2)rdrdθ.\displaystyle 4\int\limits^{\pi/2}_{0}\int\limits^{\infty}_{0}\cos(2\pi tr\sin\theta)W(r^{2})\ r\mathrm{d}r\mathrm{d}\theta.

Using this and the definition of Wδ,τ(t)W_{\delta,\tau}(t) in (2.7), we deduce that, for 1+(w)>c>max(|(τ)|,(w))1+\Re(w)>c>\max(|\Re(\tau)|,\Re(w)),

h(ξ,w)\displaystyle h(\xi,w)
=\displaystyle= 400π/20cos(2π(ξt)1/2rsinθ)Ψ(r2)Wδ,τ(π2t25Xr2)rdrdθtw1𝑑t\displaystyle 4\int\limits_{0}^{\infty}\int\limits^{\pi/2}_{0}\int\limits^{\infty}_{0}\cos(2\pi\left(\frac{\xi}{t}\right)^{1/2}r\sin\theta)\Psi(r^{2})W_{\delta,\tau}\left(\frac{\pi^{2}t}{2^{5}Xr^{2}}\right)\ r\mathrm{d}r\mathrm{d}\theta t^{w-1}\,dt
=\displaystyle= 200π/20cos(r)Ψ((rt1/22πξ1/2sinθ)2)Wδ,τ((π2t25X)(rt1/22πξ1/2sinθ)2)d(rt1/22πξ1/2sinθ)2dθtw1𝑑t\displaystyle 2\int\limits_{0}^{\infty}\int\limits^{\pi/2}_{0}\int\limits^{\infty}_{0}\cos(r)\Psi((\frac{rt^{1/2}}{2\pi\xi^{1/2}\sin\theta})^{2})W_{\delta,\tau}\left(\left(\frac{\pi^{2}t}{2^{5}X}\right)(\frac{rt^{1/2}}{2\pi\xi^{1/2}\sin\theta})^{-2}\right)\ \mathrm{d}(\frac{rt^{1/2}}{2\pi\xi^{1/2}\sin\theta})^{2}\mathrm{d}\theta t^{w-1}\,dt
=\displaystyle= 20π/20cos(r)Wδ,τ((π225X)(r2πξ1/2sinθ)2)d(r2πξ1/2sinθ)2dθ0Ψ((rt1/22πξ1/2sinθ)2)tw+1dtt\displaystyle 2\int\limits^{\pi/2}_{0}\int\limits^{\infty}_{0}\cos(r)W_{\delta,\tau}\left(\left(\frac{\pi^{2}}{2^{5}X}\right)(\frac{r}{2\pi\xi^{1/2}\sin\theta})^{-2}\right)\ \mathrm{d}(\frac{r}{2\pi\xi^{1/2}\sin\theta})^{2}\mathrm{d}\theta\int\limits_{0}^{\infty}\Psi((\frac{rt^{1/2}}{2\pi\xi^{1/2}\sin\theta})^{2})t^{w+1}\,\frac{dt}{t}
=\displaystyle= 2Ψ^(1+w)0π/20cos(r)Wδ,τ((π225X)(r2πξ1/2sinθ)2)(r2πξ1/2sinθ)2w2d(r2πξ1/2sinθ)2dθ\displaystyle 2\widehat{\Psi}(1+w)\int\limits^{\pi/2}_{0}\int\limits^{\infty}_{0}\cos(r)W_{\delta,\tau}\left(\left(\frac{\pi^{2}}{2^{5}X}\right)(\frac{r}{2\pi\xi^{1/2}\sin\theta})^{-2}\right)(\frac{r}{2\pi\xi^{1/2}\sin\theta})^{-2w-2}\ \mathrm{d}(\frac{r}{2\pi\xi^{1/2}\sin\theta})^{2}\mathrm{d}\theta
=\displaystyle= 2Ψ^(1+w)ξw0π/20cos(r)12πi(cs)(25/2π)2sΓδ(s)(1X(r2πξ1/2sinθ)2)s(r2πsinθ)2w22ss2τ2𝑑sd(r2πsinθ)2dθ\displaystyle 2\widehat{\Psi}(1+w)\xi^{w}\int\limits^{\pi/2}_{0}\int\limits^{\infty}_{0}\cos(r)\frac{1}{2\pi i}\int\limits\limits_{(c_{s})}\left(\frac{2^{5/2}}{\pi}\right)^{2s}\Gamma_{\delta}(s)\left(\frac{1}{X}(\frac{r}{2\pi\xi^{1/2}\sin\theta})^{-2}\right)^{-s}(\frac{r}{2\pi\sin\theta})^{-2w-2}\ \frac{2s}{s^{2}-\tau^{2}}ds\ \mathrm{d}(\frac{r}{2\pi\sin\theta})^{2}\mathrm{d}\theta
=\displaystyle= 4Ψ^(1+w)ξw12πi(cs)(25/2π)2sΓδ(s)(Xξ)s(2π)2s+2w0π/20cos(r)r2s2w1(sinθ)2s+2w2ss2τ2𝑑sdrdθ.\displaystyle 4\widehat{\Psi}(1+w)\xi^{w}\frac{1}{2\pi i}\int\limits\limits_{(c_{s})}\left(\frac{2^{5/2}}{\pi}\right)^{2s}\Gamma_{\delta}(s)\left(\frac{X}{\xi}\right)^{s}(2\pi)^{-2s+2w}\int\limits^{\pi/2}_{0}\int\limits^{\infty}_{0}\cos(r)r^{2s-2w-1}\left(\sin\theta\right)^{-2s+2w}\frac{2s}{s^{2}-\tau^{2}}ds\ \mathrm{d}r\mathrm{d}\theta.

We apply the relation (see [Gao1, Section 2.4])

0π/2(sinθ)udθ0cos(r)rudrr=π22u1Γ(u2)Γ(2u2),\displaystyle\int\limits^{\pi/2}_{0}(\sin\theta)^{-u}\mathrm{d}\theta\int\limits^{\infty}_{0}\cos(r)r^{u}\frac{\mathrm{d}r}{r}=\frac{\pi}{2}2^{u-1}\frac{\Gamma\left(\frac{u}{2}\right)}{\Gamma\left(\frac{2-u}{2}\right)},

to see that

0π/2(sinθ)(2s2w)dθ0cos(r)r2s2wdrr=π222s2w1Γ(sw)Γ(1s+w).\displaystyle\int\limits^{\pi/2}_{0}(\sin\theta)^{-(2s-2w)}\mathrm{d}\theta\int\limits^{\infty}_{0}\cos(r)r^{2s-2w}\frac{\mathrm{d}r}{r}=\frac{\pi}{2}2^{2s-2w-1}\frac{\Gamma(s-w)}{\Gamma(1-s+w)}.

Substituting this into the above expression for h(ξ,w)h(\xi,w), we readily derive (2.14) by noticing that we can now take cs>max(|(τ)|,(w))c_{s}>\max(|\Re(\tau)|,\Re(w)). This also implies that h(ξ,w)h(\xi,w) is an entire function of ww for 1(w)>1+|(τ)|1\geq\Re(w)>-1+|\Re(\tau)|.

It remains to establish (2.15). For this, we set c=(s)c=\Re(s) and we may assume that c2c\geq 2 here. By apply Stirling’s formula given in [iwakow, (5.112)], we see that

(2.16) (25/2π)2sΓδ(s)(π)2s+2wΓ(sw)Γ(1s+w)2ss2τ2(25/2π2)2c(1+|s|)2c1eπ|(s)|(1+|sw|)2c2(w)1(|s|e1/2)2c1eπ|(s)|(1+|s|)2c2(w)1(1+|w|)2c2(w)1.\displaystyle\begin{split}&\left(\frac{2^{5/2}}{\pi}\right)^{2s}\Gamma_{\delta}(s)(\pi)^{-2s+2w}\frac{\Gamma(s-w)}{\Gamma(1-s+w)}\frac{2s}{s^{2}-\tau^{2}}\\ \ll&\left(\frac{2^{5/2}}{\pi^{2}}\right)^{2c}(1+|s|)^{2c-1}e^{-\pi|\Im(s)|}(1+|s-w|)^{2c-2\Re(w)-1}\\ \ll&(\frac{|s|}{e^{1/2}})^{2c-1}e^{-\pi|\Im(s)|}(1+|s|)^{2c-2\Re(w)-1}(1+|w|)^{2c-2\Re(w)-1}.\end{split}

This implies that

(c)(25/2π)2sΓδ(s)(π)2s+2wΓ(sw)Γ(1s+w)2sdss2τ2ec|c|4c2(w)2(1+|w|)2c2(w)1.\displaystyle\begin{split}&\int\limits\limits_{(c)}\left(\frac{2^{5/2}}{\pi}\right)^{2s}\Gamma_{\delta}(s)(\pi)^{-2s+2w}\frac{\Gamma(s-w)}{\Gamma(1-s+w)}\frac{2sds}{s^{2}-\tau^{2}}\ll e^{-c}|c|^{4c-2\Re(w)-2}(1+|w|)^{2c-2\Re(w)-1}.\end{split}

By taking

c=max(2,ξ1/4X1/4(1+|w|)1/2),\displaystyle\begin{split}c=\max(2,\frac{\xi^{1/4}}{X^{1/4}(1+|w|)^{1/2}}),\end{split}

we see that the bound given (2.15) follows. ∎

2.12. Analytical behaviors of certain functions

Besides the function h(ξ,w)h(\xi,w) considered in the previous section, we also need to know analytical behaviors of a few other functions that are needed in the paper. We include several results in this section. First, we note that the following result can be established similar to [sound1, Lemma 5.3].

Lemma 2.13.

Let α,l𝒪K\alpha,l\in\mathcal{O}_{K} be primary and for each k𝒪K,k0k\in\mathcal{O}_{K},k\neq 0, we write kk uniquely by

(2.17) k=k1k22,k=k_{1}k_{2}^{2},

with k1k_{1} square-free and k2Gk_{2}\in G. For (s)>1+|(δ)|\Re(s)>1+|\Re(\delta)|, we have

n1mod(1+i)3(n,α)=1rδ(n)N(n)sg(k,ln)N(n)1/2=L(sδ,χik1)L(s+δ,χik1)ϖG𝒢δ;ϖ(s;k,l,α)\displaystyle\sum_{\begin{subarray}{c}n\equiv 1\bmod{(1+i)^{3}}\\ (n,\alpha)=1\end{subarray}}\frac{r_{\delta}(n)}{N(n)^{s}}\cdot\frac{g(k,ln)}{N(n)^{1/2}}\ =\ L(s-\delta,\chi_{ik_{1}})L(s+\delta,\chi_{ik_{1}})\prod_{\varpi\in G}\mathcal{G}_{\delta;\varpi}(s;k,l,\alpha)
=:\displaystyle=: L(sδ,χik1)L(s+δ,χik1)𝒢δ(s;k,l,α).\displaystyle\ L(s-\delta,\chi_{ik_{1}})L(s+\delta,\chi_{ik_{1}})\mathcal{G}_{\delta}(s;k,l,\alpha).

Here 𝒢δ;ϖ(s;k,l,α)\mathcal{G}_{\delta;\varpi}(s;k,l,\alpha) is defined by

𝒢δ;ϖ(s;k,l,α)\displaystyle\mathcal{G}_{\delta;\varpi}(s;k,l,\alpha) ={(11N(ϖ)sδ(ik1ϖ))(11N(ϖ)s+δ(ik1ϖ))if ϖ|2α,(11N(ϖ)sδ(ik1ϖ))(11N(ϖ)s+δ(ik1ϖ))r=0rδ(ϖr)N(ϖ)rsg(k,ϖr+ordϖ(l))N(ϖ)r/2if ϖ2α.\displaystyle=\begin{cases}\biggl{(}1-\frac{1}{N(\varpi)^{s-\delta}}\left(\frac{ik_{1}}{\varpi}\right)\biggr{)}\biggl{(}1-\frac{1}{N(\varpi)^{s+\delta}}\left(\frac{ik_{1}}{\varpi}\right)\biggr{)}&\ \ \ \ \text{if }\varpi|2\alpha,\\ \biggl{(}1-\frac{1}{N(\varpi)^{s-\delta}}\left(\frac{ik_{1}}{\varpi}\right)\biggr{)}\biggl{(}1-\frac{1}{N(\varpi)^{s+\delta}}\left(\frac{ik_{1}}{\varpi}\right)\biggr{)}\displaystyle\sum_{r=0}^{\infty}\frac{r_{\delta}(\varpi^{r})}{N(\varpi)^{rs}}\frac{g(k,\varpi^{r+\text{{ord}}_{\varpi}(l)})}{N(\varpi)^{r/2}}&\ \ \ \ \text{if }\varpi\nmid 2\alpha.\end{cases}

Moreover, the function 𝒢δ(s;k,l,α)\mathcal{G}_{\delta}(s;k,l,\alpha) is holomorphic for (s)>12+|(δ)|\Re(s)>\frac{1}{2}+|\Re(\delta)| and satisfies the bound that uniformly for (s)12+|(δ)|+ϵ\Re(s)\geq\frac{1}{2}+|\Re(\delta)|+\epsilon,

|𝒢δ(s;k,l,α)|N(αk)ϵN(l)12+ϵN((l,k22))12.|{\mathcal{G}}_{\delta}(s;k,l,\alpha)|\ll N(\alpha k)^{\epsilon}N(l)^{\frac{1}{2}+\epsilon}N((l,k_{2}^{2}))^{\frac{1}{2}}.

Our next two lemmas provide bounds for certain dyadic sums involving 𝒢δ\mathcal{G}_{\delta} and h(ξ,w)h(\xi,w).

Lemma 2.14.

Let K,L1K,L\geq 1 be two integers and let k2k_{2} be defined in (2.17). For any sequence of complex numbers δl\delta_{l} satisfying |δl|N()ε|\delta_{l}|\ll N(\ell)^{\varepsilon} and δl=0\delta_{l}=0 when (l,2α)1(l,2\alpha)\neq 1, we have for (w)=12+|(δ)|+ε\Re(w)=-\frac{1}{2}+|\Re(\delta)|+\varepsilon,

KN(k)<2K1N(k2)|N(l)=L2L1δlN(l)𝒢δ(1+w;k,l,α)|2ε(N(α)LK)εL(L+K).\sum_{K\leq N(k)<2K}\frac{1}{N(k_{2})}\left|\sum_{\begin{subarray}{c}N(l)=L\end{subarray}}^{2L-1}\frac{\delta_{l}}{\sqrt{N(l)}}\mathcal{G}_{\delta}(1+w;k,l,\alpha)\right|^{2}\ll_{\varepsilon}(N(\alpha)LK)^{\varepsilon}L(L+K).
Proof.

We write for any k0,k𝒪Kk\neq 0,k\in\mathcal{O}_{K} as k=ukϖG,ai1ϖiaik=u_{k}\displaystyle\prod_{\varpi\in G,\ a_{i}\geq 1}\varpi_{i}^{a_{i}} with ukUKu_{k}\in U_{K}. We define for the ϖi\varpi_{i} appearing in this product,

(2.18) a(k)=iϖiai+1andb(k)=ai=1ϖiai2ϖiai1.a(k)=\prod_{i}\varpi_{i}^{a_{i}+1}\ \ \ \ \text{and}\ \ \ \ b(k)=\prod_{a_{i}=1}\varpi_{i}\prod_{a_{i}\geq 2}\varpi_{i}^{a_{i}-1}.

It follows from the definition of 𝒢δ\mathcal{G}_{\delta} in Lemma 2.13 and part (ii) of Lemma 2.2 that we may write l=gml=gm with g|a(k),gGg|a(k),g\in G, (m,k)=1(m,k)=1 and mm square-free, for otherwise we have 𝒢δ(1+w;k,l,α)=0\mathcal{G}_{\delta}(1+w;k,l,\alpha)=0. Further, we see that when (l,2α)=1(l,2\alpha)=1,

𝒢δ(1+w;k,l,α)=N(m)(ikm)ϖGϖ|m(1+rδ(ϖ)N(ϖ)1+w(ik1ϖ))1𝒢δ(1+w;k,g,α).\mathcal{G}_{\delta}(1+w;k,l,\alpha)=\sqrt{N(m)}\left(\frac{ik}{m}\right)\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|m\end{subarray}}\left(1+\frac{r_{\delta}(\varpi)}{N(\varpi)^{1+w}}\left(\frac{ik_{1}}{\varpi}\right)\right)^{-1}\mathcal{G}_{\delta}(1+w;k,g,\alpha).

Applying the above and the Cauchy-Schwarz inequality, we deduce that

KN(k)<2K1N(k2)|N(l)=L2L1δlN(l)𝒢δ(1+w;k,l,α)|2εKεKN(k)<2K1N(k2)g|a(k)N(g)<2LΨ(k,g),\sum_{K\leq N(k)<2K}\frac{1}{N(k_{2})}\left|\sum_{\begin{subarray}{c}N(l)=L\end{subarray}}^{2L-1}\frac{\delta_{l}}{\sqrt{N(l)}}\mathcal{G}_{\delta}(1+w;k,l,\alpha)\right|^{2}\ll_{\varepsilon}K^{\varepsilon}\sum_{K\leq N(k)<2K}\frac{1}{N(k_{2})}\sum_{\begin{subarray}{c}g|a(k)\\ N(g)<2L\end{subarray}}\Psi(k,g),

where

Ψ(k,g)=|LN(g)N(m)<2LN(g)μ[i]2(m)δgmN(g)𝒢δ(1+w;k,g,α)(ikm)ϖGϖ|m(1+rδ(ϖ)N(ϖ)1+w(ik1ϖ))1|2.\Psi(k,g)=\Bigg{|}\sum_{\begin{subarray}{c}\frac{L}{N(g)}\leq N(m)<\frac{2L}{N(g)}\end{subarray}}\frac{\mu^{2}_{[i]}(m)\delta_{gm}}{\sqrt{N(g)}}\mathcal{G}_{\delta}(1+w;k,g,\alpha)\left(\frac{ik}{m}\right)\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|m\end{subarray}}\left(1+\frac{r_{\delta}(\varpi)}{N(\varpi)^{1+w}}\left(\frac{ik_{1}}{\varpi}\right)\right)^{-1}\Bigg{|}^{2}.

We use the bound for 𝒢δ\mathcal{G}_{\delta} given in Lemma 2.13 in the above expression to see that

(2.19) Ψ(k,g)ε(N(α)K)εN(g)1+ε|LN(g)N(m)<2LN(g)μ[i]2(m)δgm(ikm)ϖGϖ|m(1+rδ(ϖ)N(ϖ)1+w(ik1ϖ))1|2.\Psi(k,g)\ll_{\varepsilon}(N(\alpha)K)^{\varepsilon}N(g)^{1+\varepsilon}\Bigg{|}\sum_{\begin{subarray}{c}\frac{L}{N(g)}\leq N(m)<\frac{2L}{N(g)}\end{subarray}}\mu^{2}_{[i]}(m)\delta_{gm}\left(\frac{ik}{m}\right)\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|m\end{subarray}}\left(1+\frac{r_{\delta}(\varpi)}{N(\varpi)^{1+w}}\left(\frac{ik_{1}}{\varpi}\right)\right)^{-1}\Bigg{|}^{2}.

Note that as (ikm)0\left(\frac{ik}{m}\right)\neq 0,

ϖGϖ|m(1+rδ(ϖ)N(ϖ)1+w(ik1ϖ))1\displaystyle\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|m\end{subarray}}\left(1+\frac{r_{\delta}(\varpi)}{N(\varpi)^{1+w}}\left(\frac{ik_{1}}{\varpi}\right)\right)^{-1} =ϖGϖ|m(1rδ(ϖ)2N(ϖ)2+2w)1ϖGϖ|m(1rδ(ϖ)N(ϖ)1+w(ik1ϖ))\displaystyle=\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|m\end{subarray}}\left(1-\frac{r_{\delta}(\varpi)^{2}}{N(\varpi)^{2+2w}}\right)^{-1}\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|m\end{subarray}}\left(1-\frac{r_{\delta}(\varpi)}{N(\varpi)^{1+w}}\left(\frac{ik_{1}}{\varpi}\right)\right)
=ϖGϖ|m(1rδ(ϖ)2N(ϖ)2+2w)1j|mj1mod(1+i)3μ[i](j)rδ(j)N(j)1+w(ik1j).\displaystyle=\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|m\end{subarray}}\left(1-\frac{r_{\delta}(\varpi)^{2}}{N(\varpi)^{2+2w}}\right)^{-1}\sum_{\begin{subarray}{c}j|m\\ j\equiv 1\bmod{(1+i)^{3}}\end{subarray}}\frac{\mu_{[i]}(j)r_{\delta}(j)}{N(j)^{1+w}}\left(\frac{ik_{1}}{j}\right).

Using this in (2.19), we obtain via another application of the Cauchy-Schwarz inequality that

Ψ(k,g)ε(N(α)K)εN(g)1+εN(j)<2LN(g)|LN(g)N(m)<2LN(g)j|mμ[i]2(m)δgm(ikm)ϖGϖ|m(1rδ(ϖ)2N(ϖ)2+2w)1|2.\Psi(k,g)\ll_{\varepsilon}(N(\alpha)K)^{\varepsilon}N(g)^{1+\varepsilon}\sum_{N(j)<\frac{2L}{N(g)}}\Bigg{|}\sum_{\begin{subarray}{c}\frac{L}{N(g)}\leq N(m)<\frac{2L}{N(g)}\\ j|m\end{subarray}}\mu^{2}_{[i]}(m)\delta_{gm}\left(\frac{ik}{m}\right)\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|m\end{subarray}}\left(1-\frac{r_{\delta}(\varpi)^{2}}{N(\varpi)^{2+2w}}\right)^{-1}\Bigg{|}^{2}.

We relabel mm by jmjm and note that for all ϖ|m\varpi|m and (w)=12+|(δ)|+ε\Re(w)=-\frac{1}{2}+|\Re(\delta)|+\varepsilon, we have

μ[i]2(j)(ikj)ϖGϖ|j(1rδ(ϖ)2N(ϖ)2+2w)1N(j)ε.\displaystyle\mu^{2}_{[i]}(j)\left(\frac{ik}{j}\right)\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|j\end{subarray}}\left(1-\frac{r_{\delta}(\varpi)^{2}}{N(\varpi)^{2+2w}}\right)^{-1}\ll N(j)^{\varepsilon}.

This implies that

(2.20) Ψ(k,g)ε(N(α)LK)εN(g)1+εN(j)<2LN(g)|LN(gj)N(m)<2LN(gj)(m,2αj)=1μ[i]2(m)δgjm(ikm)ϖGϖ|m(1rδ(ϖ)2N(ϖ)2+2w)1|2.\Psi(k,g)\ll_{\varepsilon}(N(\alpha)LK)^{\varepsilon}N(g)^{1+\varepsilon}\sum_{N(j)<\frac{2L}{N(g)}}\Bigg{|}\sum_{\begin{subarray}{c}\frac{L}{N(gj)}\leq N(m)<\frac{2L}{N(gj)}\\ (m,2\alpha j)=1\end{subarray}}\mu^{2}_{[i]}(m)\delta_{gjm}\left(\frac{ik}{m}\right)\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|m\end{subarray}}\left(1-\frac{r_{\delta}(\varpi)^{2}}{N(\varpi)^{2+2w}}\right)^{-1}\Bigg{|}^{2}.

Notice that g|a(k)g|a(k) implies b(g)|kb(g)|k by (2.18). We may thus relabel such kk by fb(g)fb(g) to deduce from (2.20) that

(2.21) KN(k)<2K1N(k2)g|a(k)N(g)<2LΨ(k,g)N(g)<2LKN(k)<2Kb(g)|k1N(k2)Ψ(k,g)=N(g)<2LKN(b(g))N(f)<2KN(b(g))1N(k2)Ψ(fb(g),g)ε(N(α)LK)εN(g)<2LN(g)1+εKN(b(g))N(f)<2KN(b(g))1N(k2)×N(j)<2LN(g)|LN(gj)N(m)<2LN(gj)μ[i]2(m)δgjm(ifb(g)m)ϖGϖ|m(1rδ(ϖ)2N(ϖ)2+2w)1|2.\displaystyle\begin{split}&\sum_{K\leq N(k)<2K}\frac{1}{N(k_{2})}\sum_{\begin{subarray}{c}g|a(k)\\ N(g)<2L\end{subarray}}\Psi(k,g)\\ \leq&\sum_{N(g)<2L}\sum_{\begin{subarray}{c}K\leq N(k)<2K\\ b(g)|k\end{subarray}}\frac{1}{N(k_{2})}\Psi(k,g)=\sum_{N(g)<2L}\sum_{\frac{K}{N(b(g))}\leq N(f)<\frac{2K}{N(b(g))}}\frac{1}{N(k_{2})}\Psi(fb(g),g)\\ \ll_{\varepsilon}&(N(\alpha)LK)^{\varepsilon}\sum_{N(g)<2L}N(g)^{1+\varepsilon}\sum_{\frac{K}{N(b(g))}\leq N(f)<\frac{2K}{N(b(g))}}\frac{1}{N(k_{2})}\\ &\times\sum_{N(j)<\frac{2L}{N(g)}}\Bigg{|}\sum_{\begin{subarray}{c}\frac{L}{N(gj)}\leq N(m)<\frac{2L}{N(gj)}\end{subarray}}\mu^{2}_{[i]}(m)\delta_{gjm}\left(\frac{ifb(g)}{m}\right)\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|m\end{subarray}}\left(1-\frac{r_{\delta}(\varpi)^{2}}{N(\varpi)^{2+2w}}\right)^{-1}\Bigg{|}^{2}.\end{split}

On writing f=f1f22f=f_{1}f_{2}^{2} with f1f_{1} square-free and f2Gf_{2}\in G, we observe that the relation fb(g)=k1k22fb(g)=k_{1}k_{2}^{2} implies that f2|k2f_{2}|k_{2}, so that N(k2)1N(f2)1N(k_{2})^{-1}\ll N(f_{2})^{-1}. Applying this in (2.21), we see that the assertion of our lemma follows from Lemma 2.7. ∎

Lemma 2.15.

Let K,L1K,L\geq 1 be two integers and let N(α)X,X>0N(\alpha)\leq X,X>0. For (w)=12+|(δ)|+ε\Re(w)=-\frac{1}{2}+|\Re(\delta)|+\varepsilon and any sequence of complex numbers γl\gamma_{l} satisfying |γl|1|\gamma_{l}|\leq 1, the expression

(2.22) KN(k)<2K1N(k2)|N(l)=L(l,2α)=12L1γlN(l)𝒢δ(1+w;k,l,α)h(N(k)X2N(α2l),w)|2\displaystyle\sum_{\begin{subarray}{c}K\leq N(k)<2K\end{subarray}}\frac{1}{N(k_{2})}\Bigg{|}\sum_{\begin{subarray}{c}N(l)=L\\ (l,2\alpha)=1\end{subarray}}^{2L-1}\frac{\gamma_{l}}{N(l)}\mathcal{G}_{\delta}(1+w;k,l,\alpha)h(\frac{N(k)X}{2N(\alpha^{2}l)},w)\Bigg{|}^{2}

is bounded by

ε|Ψ^(1+w)|2(1+|w|)84|(δ)|+εN(α)24|(δ)|+εL22|(δ)|+εK2|(δ)|+εX12|(δ)|εexp(120K4N(α)2L(1+|w|2)4),\displaystyle\ll_{\varepsilon}|\widehat{\Psi}(1+w)|^{2}(1+|w|)^{8-4|\Re(\delta)|+\varepsilon}\frac{N(\alpha)^{2-4|\Re(\delta)|+\varepsilon}L^{2-2|\Re(\delta)|+\varepsilon}K^{2|\Re(\delta)|+\varepsilon}}{X^{1-2|\Re(\delta)|-\varepsilon}}\exp\Bigg{(}-\frac{1}{20}\frac{\sqrt[4]{K}}{\sqrt[4]{N(\alpha)^{2}L(1+|w|^{2})}}\Bigg{)},

and also by

ε((1+|w|)N(α)LKX)ε|Ψ^(1+w)|2(N(α)2L(1+|w|2)K)2|(τ)|2|(δ)|N(α)2LKX12|(δ)|(K+L).\displaystyle\ll_{\varepsilon}((1+|w|)N(\alpha)LKX)^{\varepsilon}|\widehat{\Psi}(1+w)|^{2}\Big{(}\frac{N(\alpha)^{2}L(1+|w|^{2})}{K}\Big{)}^{2|\Re(\tau)|-2|\Re(\delta)|}\frac{N(\alpha)^{2}L}{KX^{1-2|\Re(\delta)|}}(K+L).
Proof.

We apply Lemma 2.11, Lemma 2.13 to bound respectively h(ξ,w)h(\xi,w) and 𝒢δ\mathcal{G}_{\delta} to see that the expression in (2.22) is

\displaystyle\ll |Ψ^(1+w)|2(1+|w|)84|(δ)|+εN(α)24|(δ)|+εL2|(δ)|+εK2|(δ)|+εX12|(δ)|εexp(120K4N(α)2L(1+|w|2)4)\displaystyle|\widehat{\Psi}(1+w)|^{2}(1+|w|)^{8-4|\Re(\delta)|+\varepsilon}\frac{N(\alpha)^{2-4|\Re(\delta)|+\varepsilon}L^{-2|\Re(\delta)|+\varepsilon}K^{2|\Re(\delta)|+\varepsilon}}{X^{1-2|\Re(\delta)|-\varepsilon}}\exp\Bigg{(}-\frac{1}{20}\frac{\sqrt[4]{K}}{\sqrt[4]{N(\alpha)^{2}L(1+|w|^{2})}}\Bigg{)}
×KN(k)<2K1N(k)N(k2)(N(l)=L(l,2α)=12L1N((l,k22))12)2\displaystyle\times\sum_{\begin{subarray}{c}K\leq N(k)<2K\end{subarray}}\frac{1}{N(k)N(k_{2})}\Bigg{(}\sum_{\begin{subarray}{c}N(l)=L\\ (l,2\alpha)=1\end{subarray}}^{2L-1}N((l,k_{2}^{2}))^{\frac{1}{2}}\Bigg{)}^{2}
\displaystyle\ll |Ψ^(1+w)|2(1+|w|)84|(δ)|+εN(α)24|(δ)|+εL2|(δ)|+εK2|(δ)|+εX12|(δ)|εexp(120K4N(α)2L(1+|w|2)4)\displaystyle|\widehat{\Psi}(1+w)|^{2}(1+|w|)^{8-4|\Re(\delta)|+\varepsilon}\frac{N(\alpha)^{2-4|\Re(\delta)|+\varepsilon}L^{-2|\Re(\delta)|+\varepsilon}K^{2|\Re(\delta)|+\varepsilon}}{X^{1-2|\Re(\delta)|-\varepsilon}}\exp\Bigg{(}-\frac{1}{20}\frac{\sqrt[4]{K}}{\sqrt[4]{N(\alpha)^{2}L(1+|w|^{2})}}\Bigg{)}
×KN(k)<2K1N(k)N(k2)(N(l)=L(l,2α)=12L1N(k2))2.\displaystyle\times\sum_{\begin{subarray}{c}K\leq N(k)<2K\end{subarray}}\frac{1}{N(k)N(k_{2})}\Bigg{(}\sum_{\begin{subarray}{c}N(l)=L\\ (l,2\alpha)=1\end{subarray}}^{2L-1}N(k_{2})\Bigg{)}^{2}.

Upon writing N(k)=N(k1k22)N(k)=N(k_{1}k^{2}_{2}), we readily deduce the first bound of the lemma from the above estimation.

To derive the second bound, we set c=|(τ)|+εc=|\Re(\tau)|+\varepsilon to recast the integral in (2.14) as

12πi(|(τ)|+ε)g(s,w)(Xξ)s𝑑s.\displaystyle\frac{1}{2\pi i}\int\limits_{(|\Re(\tau)|+\varepsilon)}g(s,w)\left(\frac{X}{\xi}\right)^{s}\,ds.

This implies that

|N(l)=L(l,2α)=12L1γlN(l)𝒢δ(1+w;k,l,α)h(N(k)X2N(α2l),w)|\displaystyle\Bigg{|}\sum_{\begin{subarray}{c}N(l)=L\\ (l,2\alpha)=1\end{subarray}}^{2L-1}\frac{\gamma_{l}}{N(l)}\mathcal{G}_{\delta}(1+w;k,l,\alpha)h(\frac{N(k)X}{2N(\alpha^{2}l)},w)\Bigg{|}
\displaystyle\ll |Ψ^(1+w)|(N(α)1+2|(τ)|2|(δ)|+εK12+|(τ)||(δ)|εX12|(δ)|ε)(|(τ)|+ε)|g(s,w)N(l)=L(l,2α)=12L1γlN(l)1+ws𝒢δ(1+w;k,l,α)||ds|.\displaystyle|\widehat{\Psi}(1+w)|\left(\frac{N(\alpha)^{1+2|\Re(\tau)|-2|\Re(\delta)|+\varepsilon}}{K^{\frac{1}{2}+|\Re(\tau)|-|\Re(\delta)|-\varepsilon}X^{\frac{1}{2}-|\Re(\delta)|-\varepsilon}}\right)\int\limits_{(|\Re(\tau)|+\varepsilon)}\Bigg{|}g(s,w)\sum_{\begin{subarray}{c}N(l)=L\\ (l,2\alpha)=1\end{subarray}}^{2L-1}\frac{\gamma_{l}}{N(l)^{1+w-s}}\mathcal{G}_{\delta}(1+w;k,l,\alpha)\Bigg{|}\,|ds|.

Notice that (2.16) is still valid with c=|(τ)|+εc=|\Re(\tau)|+\varepsilon and (w)=12+|(δ)|+ε\Re(w)=-\frac{1}{2}+|\Re(\delta)|+\varepsilon so that it implies that g(s,w)ε(1+|w|)2|(τ)|2|(δ)|+εexp(π2|(s)|)g(s,w)\ll_{\varepsilon}(1+|w|)^{2|\Re(\tau)|-2|\Re(\delta)|+\varepsilon}\exp(-\frac{\pi}{2}|\Im(s)|). We apply this estimation and the Cauchy-Schwarz inequality to deduce that

|N(l)=L(l,2α)=12L1γlN(l)𝒢δ(1+w;k,l,α)h(N(k)X2N(α2l),w)|(1+|w|)2|(τ)|2|(δ)|+ε|Ψ^(1+w)|(N(α)1+2|(τ)|2|(δ)|+εK12+|(τ)||(δ)|εX12|(δ)|ε)×(|(τ)|+ε)(exp(π2|(s)|)|N(l)=L(l,2α)=12L1γlN(l)1+ws𝒢δ(1+w;k,l,α)|2|ds|)1/2.\displaystyle\begin{split}&\Bigg{|}\sum_{\begin{subarray}{c}N(l)=L\\ (l,2\alpha)=1\end{subarray}}^{2L-1}\frac{\gamma_{l}}{N(l)}\mathcal{G}_{\delta}(1+w;k,l,\alpha)h(\frac{N(k)X}{2N(\alpha^{2}l)},w)\Bigg{|}\\ \ll&(1+|w|)^{2|\Re(\tau)|-2|\Re(\delta)|+\varepsilon}|\widehat{\Psi}(1+w)|\left(\frac{N(\alpha)^{1+2|\Re(\tau)|-2|\Re(\delta)|+\varepsilon}}{K^{\frac{1}{2}+|\Re(\tau)|-|\Re(\delta)|-\varepsilon}X^{\frac{1}{2}-|\Re(\delta)|-\varepsilon}}\right)\\ &\times\int\limits_{(|\Re(\tau)|+\varepsilon)}\Bigg{(}\exp(-\tfrac{\pi}{2}|\Im(s)|)\Bigg{|}\sum_{\begin{subarray}{c}N(l)=L\\ (l,2\alpha)=1\end{subarray}}^{2L-1}\frac{\gamma_{l}}{N(l)^{1+w-s}}\mathcal{G}_{\delta}(1+w;k,l,\alpha)\Bigg{|}^{2}\,|ds|\Bigg{)}^{1/2}.\end{split}

By inserting the above bound into (2.22) and applying Lemma 2.14, we obtain the second bound of the lemma. ∎

In the remaining of the section, we include two more results concerning various functions studied in this paper.

Lemma 2.16.

Let RR be a polynomial satisfying R(0)=R(0)=0R(0)=R^{\prime}(0)=0. Let gg be a multiplicative function satisfying g(ϖ)=1+O(N(ϖ)ν)g(\varpi)=1+O(N(\varpi)^{-\nu}) for some fixed ν>0\nu>0. Let uu and vv be two bounded complex numbers such that (u+v)\Re(u+v) and (uv)\Re(u-v) are D/logy\geq-D/\log y for an absolute positive constant DD and a large real number yy. Then we have for (s)>1+D/logy\Re(s)>1+D/\log y,

(2.23) n1mod(1+i)3rv(n)μ[i](nc)N(n)s+ug(n)=μ[i](c)G(s,c;u,v)ζK(s+u+v)ζK(s+uv),\displaystyle\sum_{\begin{subarray}{c}n\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{r_{v}(n)\mu_{[i]}(nc)}{N(n)^{s+u}}g(n)=\frac{\mu_{[i]}(c)G(s,c;u,v)}{\zeta_{K}(s+u+v)\zeta_{K}(s+u-v)},

where G(s,c;u,v)=ϖGGϖ(s,c;u,v)G(s,c;u,v)=\prod_{\begin{subarray}{c}\varpi\in G\end{subarray}}G_{\varpi}(s,c;u,v) is a holomorphic function in (s)>max(12,1ν)+D/logy\Re(s)>\max(\frac{1}{2},1-\nu)+D/\log y defined by

Gϖ(s,c;u,v):={(11N(ϖ)s+u+v)1(11N(ϖ)s+uv)1if ϖ|2c,(11N(ϖ)s+u+v)1(11N(ϖ)s+uv)1(1g(ϖ)rv(ϖ)N(ϖ)s+u)otherwise.G_{\varpi}(s,c;u,v):=\begin{cases}(1-\frac{1}{N(\varpi)^{s+u+v}})^{-1}(1-\frac{1}{N(\varpi)^{s+u-v}})^{-1}&\text{if }\varpi|2c,\\ (1-\frac{1}{N(\varpi)^{s+u+v}})^{-1}(1-\frac{1}{N(\varpi)^{s+u-v}})^{-1}(1-\frac{g(\varpi)r_{v}(\varpi)}{N(\varpi)^{s+u}})&\text{otherwise}.\end{cases}

Moreover, for any odd c𝒪Kc\in\mathcal{O}_{K} with N(c)yN(c)\leq y, we have

(2.24) N(n)y/N(c)n1mod(1+i)3rv(n)μ[i](nc)N(n)1+ug(n)R(log(y/N(cn))logy)=O(E(c)log2y(yN(c))(u)+|(v)|exp(A0log(y/N(c))))+Ress=0μ[i](c)G(s+1,c;u,v)sζK(1+s+u+v)ζK(1+s+uv)k=01(slogy)kR(k)(log(y/N(c))logy),\displaystyle\begin{split}\sum_{\begin{subarray}{c}N(n)\leq y/N(c)\\ n\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{r_{v}(n)\mu_{[i]}(nc)}{N(n)^{1+u}}&g(n)R\left(\frac{\log(y/N(cn))}{\log y}\right)=O\Big{(}\frac{E(c)}{\log^{2}y}\left(\frac{y}{N(c)}\right)^{-\Re(u)+|\Re(v)|}\exp(-A_{0}\sqrt{\log(y/N(c))})\Big{)}\\ &+\mathop{\text{Res}}_{s=0}\frac{\mu_{[i]}(c)G(s+1,c;u,v)}{s\zeta_{K}(1+s+u+v)\zeta_{K}(1+s+u-v)}\sum_{k=0}^{\infty}\frac{1}{(s\log y)^{k}}R^{(k)}\Big{(}\frac{\log(y/N(c))}{\log y}\Big{)},\end{split}

where A0>0A_{0}>0 is an absolute constant and E(c)=ϖGϖ|c(1+1/N(ϖ))E(c)=\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|c\end{subarray}}(1+1/\sqrt{N(\varpi)}).

Proof.

First note that we can establish (2.23) by considering Euler products. To prove (2.24), we may assume that N(c)y/2N(c)\leq y/2. We then apply the Taylor expansion R(x)=j=0R(j)(0)j!xj=j=2R(j)(0)j!xjR(x)=\sum_{j=0}^{\infty}\frac{R^{(j)}(0)}{j!}x^{j}=\sum_{j=2}^{\infty}\frac{R^{(j)}(0)}{j!}x^{j} to write the sum in (2.24) as

j=2R(j)(0)(logy)j1j!N(n)y/N(c)n1mod(1+i)3rv(n)μ[i](nc)N(n)1+ug(n)logj(yN(cn)).\displaystyle\sum_{j=2}^{\infty}\frac{R^{(j)}(0)}{(\log y)^{j}}\frac{1}{j!}\sum_{\begin{subarray}{c}N(n)\leq y/N(c)\\ n\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{r_{v}(n)\mu_{[i]}(nc)}{N(n)^{1+u}}g(n)\log^{j}\left(\frac{y}{N(cn)}\right).

We note that the inner sum above can be regarded as a Riesz type means so that we can apply the treatment given in [MVa1, Sect 5.1] to further write the above sums as

(2.25) j=2R(j)(0)(logy)j12πi(D+1log(y/N(c)))μ[i](c)G(s+1,c;u,v)ζK(1+s+u+v)ζK(1+s+uv)(yN(c))sdssj+1.\displaystyle\sum_{j=2}^{\infty}\frac{R^{(j)}(0)}{(\log y)^{j}}\frac{1}{2\pi i}\int\limits_{(\frac{D+1}{\log(y/N(c))})}\frac{\mu_{[i]}(c)G(s+1,c;u,v)}{\zeta_{K}(1+s+u+v)\zeta_{K}(1+s+u-v)}\Big{(}\frac{y}{N(c)}\Big{)}^{s}\frac{ds}{s^{j+1}}.

To evaluate the integral above, we set T=exp(log(y/N(c)))T=\exp(\sqrt{\log(y/N(c))}) and apply the zero free region for ζK(s)\zeta_{K}(s) (see [MVa1, Section 8.4]) to choose a positive constant A1A_{1} such that ζK(1+s+u+v)ζK(1+s+uv)\zeta_{K}(1+s+u+v)\zeta_{K}(1+s+u-v) has no zeros in the region (s)(u)+|(v)|A1/logT,(s)T\Re(s)\geq-\Re(u)+|\Re(v)|-A_{1}/\log T,\Im(s)\leq T. We then notice that we have G(s+1,c;u,v)E(c)G(s+1,c;u,v)\ll E(c) uniformly for (s)(u)+|(v)|A1/logT\Re(s)\geq-\Re(u)+|\Re(v)|-A_{1}/\log T. Moreover, similar to the bound given for the Riemann zeta function in [MVa1, Theorem 6.7], we have the following estimation for 1/ζK(1+s)1/\zeta_{K}(1+s) which asserts that for (s)>A1/logT\Re(s)>-A_{1}/\log T,

(2.26) 1ζK(1+s)min(1,log2(s)).\displaystyle\frac{1}{\zeta_{K}(1+s)}\ll\min(1,\log^{2}\Im(s)).

We now truncate the integral in (2.25) to the line segment D+1log(y/N(c))iT\frac{D+1}{\log(y/N(c))}-iT to D+1log(y/N(c))+iT\frac{D+1}{\log(y/N(c))}+iT with T=exp(log(y/N(c)))T=\exp(\sqrt{\log(y/N(c))}). By applying the above estimations for G(s+1,c;u,v)G(s+1,c;u,v) and 1/ζK(1+s)1/\zeta_{K}(1+s), we see that the error introduced by doing so is

(2.27) E(c)(logy/N(c))2/T2.\displaystyle\ll E(c)(\log y/N(c))^{2}/T^{2}.

We then shift the remaining integral to the line segment (u)+|(v)|A1/logT-\Re(u)+|\Re(v)|-A_{1}/\log T. Again by the above estimations for G(s+1,c;u,v)G(s+1,c;u,v) and 1/ζK(1+s)1/\zeta_{K}(1+s), we see that the two horizontal integrals are

(2.28) E(c)(log(y/N(c)))21T3.\displaystyle\ll E(c)(\log(y/N(c)))^{2}\frac{1}{T^{3}}.

For the vertical integral, we note that as min((u+v),(uv))D/logy\min(\Re(u+v),\Re(u-v))\geq-D/\log y, we have

(u)+|(v)|A1/logT1/log(y/N(c)).\displaystyle-\Re(u)+|\Re(v)|-A_{1}/\log T\ll-1/\sqrt{\log(y/N(c))}.

We can thus divide the vertical integral into two parts, one over the segment |(s)|1/log(y/N(c))|\Im(s)|\leq 1/\sqrt{\log(y/N(c))} and one over the rest. We apply (2.26) and the bound G(s+1,c;u,v)E(c)G(s+1,c;u,v)\ll E(c) to see that the vertical integral is

\displaystyle\ll E(c)((log(y/N(c)))j/2+(log(y/N(c)))2(log(y/N(c)))j/2)(y/N(c))(u)+|(v)|A1/logT.\displaystyle E(c)((\log(y/N(c)))^{j/2}+(\log(y/N(c)))^{2}(\log(y/N(c)))^{j/2})(y/N(c))^{-\Re(u)+|\Re(v)|-A_{1}/\log T}.

As RR is a polynomial, we know that R(j)(0)0R^{(j)}(0)\neq 0 only for finitely many jj so that we may assume that jj is bounded. It follows that for an appropriate positive constant A0A_{0}, the above is

(2.29) E(c)(yN(c))(u)+|(v)|exp(A0log(y/N(c))).\displaystyle\ll E(c)\left(\frac{y}{N(c)}\right)^{-\Re(u)+|\Re(v)|}\exp(-A_{0}\sqrt{\log(y/N(c))}).

Furthermore, we note that we encounter a multiple pole at s=0s=0 in the above process. Thus, by combining (2.27), (2.28) and (2.29), we conclude that the expression given in (2.25) is

(2.30) =Ress=0μ[i](c)G(s+1,c;u,v)sζK(1+s+u+v)ζK(1+s+uv)j=2R(j)(0)(y/N(c))ssj(logy)j+O(E(c)(logy)2(yN(c))(u)+|(v)|exp(A0log(y/N(c)))).\displaystyle\begin{split}=&\mathop{\text{R}es}_{s=0}\frac{\mu_{[i]}(c)G(s+1,c;u,v)}{s\zeta_{K}(1+s+u+v)\zeta_{K}(1+s+u-v)}\sum_{j=2}^{\infty}\frac{R^{(j)}(0)(y/N(c))^{s}}{s^{j}(\log y)^{j}}\\ &+O\Big{(}\frac{E(c)}{(\log y)^{2}}\left(\frac{y}{N(c)}\right)^{-\Re(u)+|\Re(v)|}\exp(-A_{0}\sqrt{\log(y/N(c))})\Big{)}.\end{split}

We now calculate the residue above using the Taylor expansion of (y/N(c))s(y/N(c))^{s} around s=0s=0 and discarding the powers of ss that are j+1\geq j+1 since they make no contributions. This way, we see that the residue equals

j=2R(j)(0)sj(logy)j(ljsll!(log(y/N(c))l)=k=0sk(logy)kl=0R(k+l)(0)l!(log(y/N(c))logy)l=k=0sk(logy)kR(k)(log(y/N(c))logy),\sum_{j=2}^{\infty}\frac{R^{(j)}(0)}{s^{j}(\log y)^{j}}\Big{(}\sum_{l\leq j}\frac{s^{l}}{l!}(\log(y/N(c))^{l}\Big{)}=\sum_{k=0}^{\infty}\frac{s^{-k}}{(\log y)^{k}}\sum_{l=0}^{\infty}\frac{R^{(k+l)}(0)}{l!}\Big{(}\frac{\log(y/N(c))}{\log y}\Big{)}^{l}=\sum_{k=0}^{\infty}\frac{s^{-k}}{(\log y)^{k}}R^{(k)}\Big{(}\frac{\log(y/N(c))}{\log y}\Big{)},

where the first equality above following by setting k=jlk=j-l while noting that R(0)=R(0)=0R(0)=R^{\prime}(0)=0. Applying this in (2.30) allows us to deduce (2.24) and this completes the proof of the lemma. ∎

For our next result, we define for odd primary primes ϖ\varpi,

hw(ϖ)=\displaystyle h_{w}(\varpi)= (1+1N(ϖ)+1N(ϖ)1+2wN(ϖ)2δ+N(ϖ)2δN(ϖ)2+2w+1N(ϖ)3+4w),\displaystyle\Big{(}1+\frac{1}{N(\varpi)}+\frac{1}{N(\varpi)^{1+2w}}-\frac{N(\varpi)^{-2\delta}+N(\varpi)^{2\delta}}{N(\varpi)^{2+2w}}+\frac{1}{N(\varpi)^{3+4w}}\Big{)},
Hw(ϖ)=\displaystyle H_{w}(\varpi)= 1+1N(ϖ)1+2wrδ(ϖ)2N(ϖ)1+2whw(ϖ),\displaystyle 1+\frac{1}{N(\varpi)^{1+2w}}-\frac{r_{\delta}(\varpi)^{2}}{N(\varpi)^{1+2w}h_{w}(\varpi)},

and extend the above definitions multiplicatively to functions hw(n),Hw(n)h_{w}(n),H_{w}(n) on primary odd, square-free algebraic numbers n𝒪Kn\in\mathcal{O}_{K}.

Moreover, we write any odd l𝒪Kl\in\mathcal{O}_{K} as l=l1l22l=l_{1}l_{2}^{2}, with l1l_{1} being square-free and l2Gl_{2}\in G. We define an absolutely convergent function ηw(s;l)=ϖGηϖ;w(s;l)\eta_{w}(s;l)=\prod_{\begin{subarray}{c}\varpi\in G\end{subarray}}\eta_{\varpi;w}(s;l) for complex numbers w,sw,s in the region |(w)|14|\Re(w)|\leq\frac{1}{4} and (s)>12\Re(s)>\frac{1}{2} such that η1+i;w(s;l)=(12s2w)(12s)(1ws+2w)\eta_{1+i;w}(s;l)=(1-2^{-s-2w})(1-2^{-s})(1-w^{-s+2w}) and for primes (ϖ,2)=1(\varpi,2)=1,

(2.31) ηϖ;w(s;l)={(N(ϖ)N(ϖ)+1)(11N(ϖ)s)(1+1N(ϖ)+1N(ϖ)sN(ϖ)2w+N(ϖ)2wN(ϖ)s+1+1N(ϖ)2s+1)if ϖl,(N(ϖ)N(ϖ)+1)(11N(ϖ)s)if ϖ|l1,(N(ϖ)N(ϖ)+1)(11N(ϖ)2s)otherwise.\displaystyle\begin{split}\eta_{\varpi;w}(s;l)=\begin{cases}\left(\frac{N(\varpi)}{N(\varpi)+1}\right)\Big{(}1-\frac{1}{N(\varpi)^{s}}\Big{)}\Big{(}1+\frac{1}{N(\varpi)}+\frac{1}{N(\varpi)^{s}}-\frac{N(\varpi)^{2w}+N(\varpi)^{-2w}}{N(\varpi)^{s+1}}+\frac{1}{N(\varpi)^{2s+1}}\Big{)}&\text{if }\varpi\nmid l,\\ \left(\frac{N(\varpi)}{N(\varpi)+1}\right)\Big{(}1-\frac{1}{N(\varpi)^{s}}\Big{)}&\text{if }\varpi|l_{1},\\ \left(\frac{N(\varpi)}{N(\varpi)+1}\right)\Big{(}1-\frac{1}{N(\varpi)^{2s}}\Big{)}&\text{otherwise}.\end{cases}\end{split}

Lastly, we denote 𝒞{\mathcal{C}} for a closed contour (oriented counter-clockwise) containing the points ±τ\pm\tau with perimeter length being |δ1|\ll|\delta_{1}|. Also, for w𝒞w\in{\mathcal{C}} we have |(w)||τ|+C/logX|\Re(w)|\leq|\tau|+C/\log X, |(w)|C/logX|\Im(w)|\leq C/\log X for some absolute constant CC and such that min(|w2τ2|,|w2δ2|)ϵ2/(3log2X)\min(|w^{2}-\tau^{2}|,|w^{2}-\delta^{2}|)\geq\epsilon^{2}/(3\log^{2}X) . We then have the following result.

Lemma 2.17.

With notations as above, we have for ww on the contour 𝒞{\mathcal{C}} and x2x\geq 2,

(2.32) ηδ(1+2w;1)N(γ)xγ1mod(1+i)3μ[i]2(γ)Hw(γ)N(γ)1+2τhw(γ)G(1,γ;δ1+w,δ)G(1,γ;δ2+w,δ)=ζK(1+2τ)(1x2τ)(1+O(|wτ|))+O(x2τ).\displaystyle\begin{split}&\eta_{\delta}(1+2w;1)\sum_{\begin{subarray}{c}N(\gamma)\leq x\\ \gamma\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{\mu^{2}_{[i]}(\gamma)H_{w}(\gamma)}{N(\gamma)^{1+2\tau}h_{w}(\gamma)}G(1,\gamma;\delta_{1}+w,\delta)G(1,\gamma;\delta_{2}+w,\delta)\\ =&\zeta_{K}(1+2\tau)(1-x^{-2\tau})(1+O(|w-\tau|))+O(x^{-2\tau}).\end{split}

Also, for any smooth function RR on [0,1][0,1] and 1yx1\leq y\leq x, we have

(2.33) ηδ(1+2w;1)yN(γ)xγ1mod(1+i)3μ[i]2(γ)Hw(γ)N(γ)1+2τhw(γ)G(1,γ;δ1+w,δ)G(1,γ;δ2+w,δ)R(logN(γ)logx)=π4(1+O(|δ1|))yxR(logtlogx)dtt1+2τ.\displaystyle\begin{split}&\eta_{\delta}(1+2w;1)\sum_{\begin{subarray}{c}y\leq N(\gamma)\leq x\\ \gamma\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{\mu^{2}_{[i]}(\gamma)H_{w}(\gamma)}{N(\gamma)^{1+2\tau}h_{w}(\gamma)}G(1,\gamma;\delta_{1}+w,\delta)G(1,\gamma;\delta_{2}+w,\delta)R\Big{(}\frac{\log N(\gamma)}{\log x}\Big{)}\\ =&\frac{\pi}{4}\cdot(1+O(|\delta_{1}|))\int_{y}^{x}R\Big{(}\frac{\log t}{\log x}\Big{)}\frac{dt}{t^{1+2\tau}}.\end{split}
Proof.

Applying the definition of G(s,γ;u,v)G(s,\gamma;u,v) given in Lemma 2.16 allows us to write the expression given in (2.32) as

ηδ(1+2w;1)G(1,1;δ1+w,δ)G(1,1;δ2+w,δ)N(γ)xγ1mod(1+i)3fw(γ)N(γ)1+2τ,\eta_{\delta}(1+2w;1)G(1,1;\delta_{1}+w,\delta)G(1,1;\delta_{2}+w,\delta)\sum_{\begin{subarray}{c}N(\gamma)\leq x\\ \gamma\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{f_{w}(\gamma)}{N(\gamma)^{1+2\tau}},

where

fw(γ)=μ[i]2(γ)Hw(γ)hw(γ)ϖGϖ|γ(1rδ(ϖ)N(ϖ)1+δ1+whw(ϖ))1(1rδ(ϖ)N(ϖ)1+δ2+whw(ϖ))1.f_{w}(\gamma)=\mu^{2}_{[i]}(\gamma)\frac{H_{w}(\gamma)}{h_{w}(\gamma)}\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|\gamma\end{subarray}}\Big{(}1-\frac{r_{\delta}(\varpi)}{N(\varpi)^{1+\delta_{1}+w}h_{w}(\varpi)}\Big{)}^{-1}\Big{(}1-\frac{r_{\delta}(\varpi)}{N(\varpi)^{1+\delta_{2}+w}h_{w}(\varpi)}\Big{)}^{-1}.

Note here that fw(γ)f_{w}(\gamma) is a multiplicative function such that fw(ϖ)=1+O(1/N(ϖ))f_{w}(\varpi)=1+O(1/\sqrt{N(\varpi)}). This allows us to write γ1mod(1+i)3fw(γ)/N(γ)s=ζK(s)Fw(s)\sum_{\gamma\equiv 1\bmod(1+i)^{3}}f_{w}(\gamma)/N(\gamma)^{s}=\zeta_{K}(s)F_{w}(s) so that FF is holomorphic in (s)>12\Re(s)>\frac{1}{2}.

Now we apply Perron’s formula as given in [MVa1, Theorem 5.2, Corollary 5.3] to obtain that for σ=2τ+ε\sigma=-2\tau+\varepsilon, T=x1+2τ+εT=x^{1+2\tau+\varepsilon},

N(γ)xγ1mod(1+i)3fw(γ)N(γ)1+2τ\displaystyle\sum_{\begin{subarray}{c}N(\gamma)\leq x\\ \gamma\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{f_{w}(\gamma)}{N(\gamma)^{1+2\tau}} =12πiσiTσ+iTζK(1+2τ+s)Fw(1+2τ+s)xss𝑑s+R,\displaystyle=\frac{1}{2\pi i}\int^{\sigma+iT}_{\sigma-iT}\zeta_{K}(1+2\tau+s)F_{w}(1+2\tau+s)\frac{x^{s}}{s}ds+R,

where

Rx/2<N(n)<xN(n)xfw(n)N(n)1+2τmin(1,xT|xN(n)|)+4σ+xσTn1mod(1+i)3|fw(n)|N(n)1+2τN(n)σ.\displaystyle R\ll\sum_{\begin{subarray}{c}x/2<N(n)<x\\ N(n)\neq x\end{subarray}}\frac{f_{w}(n)}{N(n)^{1+2\tau}}\min(1,\frac{x}{T|x-N(n)|})+\frac{4^{\sigma}+x^{\sigma}}{T}\sum_{n\equiv 1\bmod(1+i)^{3}}\frac{\frac{|f_{w}(n)|}{N(n)^{1+2\tau}}}{N(n)^{\sigma}}.

It is then easy to see that

(2.34) R1k<xxTk+x2τ+εTxlogxTx2τ.\displaystyle R\ll\sum_{\begin{subarray}{c}1\leq k<x\end{subarray}}\frac{x}{Tk}+\frac{x^{-2\tau+\varepsilon}}{T}\ll\frac{x\log x}{T}\ll x^{-2\tau}.

We now shift the contour of integration to σ1=2τ1/2ε\sigma_{1}=-2\tau-1/2-\varepsilon to see that we have

(2.35) 12πiσ±iTσ1±iTζK(1+2τ+s)Fw(1+2τ+s)xss𝑑sx2τ.\displaystyle\frac{1}{2\pi i}\int^{\sigma_{1}\pm iT}_{\sigma\pm iT}\zeta_{K}(1+2\tau+s)F_{w}(1+2\tau+s)\frac{x^{s}}{s}ds\ll x^{-2\tau}.

Moreover, by applying the following subconvexity bound for ζk(s)\zeta_{k}(s) on the critical line given in [HB1988],

ζK(s)(1+|s|)1/3+ε,(s)=1/2,\displaystyle\zeta_{K}(s)\ll(1+|s|)^{1/3+\varepsilon},\quad\Re(s)=1/2,

we deduce that

(2.36) 12πiσ1iTσ1+iTζK(1+2τ+s)Fw(1+2τ+s)xss𝑑sTT(1+|t|)1/3+εx2τ1/2+ε1+|t||dt|x2τ.\displaystyle\frac{1}{2\pi i}\int^{\sigma_{1}+iT}_{\sigma_{1}-iT}\zeta_{K}(1+2\tau+s)F_{w}(1+2\tau+s)\frac{x^{s}}{s}ds\ll\int^{T}_{-T}(1+|t|)^{1/3+\varepsilon}\frac{x^{-2\tau-1/2+\varepsilon}}{1+|t|}|dt|\ll x^{-2\tau}.

By combining (2.34), (2.35) and (2.36), together with the observation that ζK(1+2τ)=π41/(2τ)+O(1)\zeta_{K}(1+2\tau)=\frac{\pi}{4}\cdot 1/(2\tau)+O(1), Fw(1+2τ)=Fw(1)+O(τ)F_{w}(1+2\tau)=F_{w}(1)+O(\tau), we then conclude that

N(γ)xγ1mod(1+i)3fw(γ)N(γ)1+2τ\displaystyle\sum_{\begin{subarray}{c}N(\gamma)\leq x\\ \gamma\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{f_{w}(\gamma)}{N(\gamma)^{1+2\tau}} =ζK(1+2τ)(1x2τ)Fw(1+2τ)+O(x2τ).\displaystyle=\zeta_{K}(1+2\tau)(1-x^{-2\tau})F_{w}(1+2\tau)+O(x^{-2\tau}).

If we set for brevity that F(w)=ηδ(1+2w;1)G(1,1;δ1+w,δ)G(1,1;δ2+w,δ)Fw(1+2τ)F(w)=\eta_{\delta}(1+2w;1)G(1,1;\delta_{1}+w,\delta)G(1,1;\delta_{2}+w,\delta)F_{w}(1+2\tau), then a little calculation implies that F(τ)=1F(\tau)=1 and F(w)=F(τ)+O(|wτ|)F(w)=F(\tau)+O(|w-\tau|). This implies (2.32), which in turn implies (2.33) by partial summation and this completes the proof of the lemma. ∎

3. Plan of the proof

We define for any large number XX and any smooth function Φ\Phi supported in [1,2][1,2],

𝒩(X,Φ)=(d,2)=1L(β,χ(1+i)5d)=0,0<β1μ[i]2(d)Φ(N(d)X),\displaystyle{\mathcal{N}}(X,\Phi)=\sum_{\begin{subarray}{c}(d,2)=1\\ L(\beta,\chi_{(1+i)^{5}d})=0,0<\beta\leq 1\end{subarray}}\mu^{2}_{[i]}(d)\Phi(\frac{N(d)}{X}),

so that 𝒩(X,Φ){\mathcal{N}}(X,\Phi) is a weighted count on the number of odd, square-free algebraic integers d𝒪Kd\in\mathcal{O}_{K} such that XN(d)2XX\leq N(d)\leq 2X and that L(s,χ(1+i)5d)L(s,\chi_{(1+i)^{5}d}) has a non-trivial real zero. The proof of Theorem 1.1 is based on an estimation of 𝒩(X,Φ){\mathcal{N}}(X,\Phi). To achieve this, we apply the following argument principle given in [C&S, Lemma 2.1], which is originally due to A. Selberg [Selberg46].

Lemma 3.1.

Let f(z)f(z) be a holomorphic function that does not vanish in the region (z)W\Re(z)\geq W. Let {\mathcal{B}} be the rectangular box with vertices W0±iHW_{0}\pm iH, W1±iHW_{1}\pm iH where H>0H>0 and W0,W1W_{0},W_{1} are two real numbers such that W0<W<W1W_{0}<W<W_{1}. Then

4Hβ+iγf(β+iγ)=0cos(πγ2H)sinh(π(βW0)2H)=HHcos(πt2H)log|f(W0+it)|dt+W0W1sinh(π(αW0)2H)log|f(α+iH)f(αiH)|dαHHcos(πW1W0+it2iH)logf(W1+it)𝑑t.\displaystyle\begin{split}&4H\sum_{\begin{subarray}{c}\beta+i\gamma\in{\mathcal{B}}\\ f(\beta+i\gamma)=0\end{subarray}}\cos\Big{(}\frac{\pi\gamma}{2H}\Big{)}\sinh\Big{(}\frac{\pi(\beta-W_{0})}{2H}\Big{)}\\ =&\int_{-H}^{H}\cos\Big{(}\frac{\pi t}{2H}\Big{)}\log|f(W_{0}+it)|dt+\int_{W_{0}}^{W_{1}}\sinh\Big{(}\frac{\pi(\alpha-W_{0})}{2H}\Big{)}\log|f(\alpha+iH)f(\alpha-iH)|d\alpha\\ &-\Re\int_{-H}^{H}\cos\Big{(}\pi\frac{W_{1}-W_{0}+it}{2iH}\Big{)}\log f(W_{1}+it)dt.\end{split}

To proceed further, we define for any sequence {an}n𝒪K\{a_{n}\}_{n\in\mathcal{O}_{K}} of complex numbers and any smooth function Φ\Phi supported in [1,2][1,2],

𝒮(ad;Φ)=𝒮(ad;Φ,X)=1X(d,2)=1μ[i]2(d)adΦ(N(d)X).{\mathcal{S}}(a_{d};\Phi)={\mathcal{S}}(a_{d};\Phi,X)=\frac{1}{X}\sum_{\begin{subarray}{c}(d,2)=1\end{subarray}}\mu^{2}_{[i]}(d)a_{d}\Phi(\frac{N(d)}{X}).

For a real parameter 1<Y2X1<Y\leq\sqrt{2X} to be determined later, we write

μ[i]2(n)=MY(n)+RY(n),\mu_{[i]}^{2}(n)=M_{Y}(n)+R_{Y}(n),

with

MY(n)=G2nN()Yμ[i](),RY(n)=G2nN()>Yμ[i]().M_{Y}(n)=\sum_{\begin{subarray}{c}\ell\in G\\ \ell^{2}\mid n\\ N(\ell)\leq Y\end{subarray}}\mu_{[i]}(\ell),\ \ \ \ \ R_{Y}(n)=\sum_{\begin{subarray}{c}\ell\in G\\ \ell^{2}\mid n\\ N(\ell)>Y\end{subarray}}\mu_{[i]}(\ell).

We then deduce that 𝒮(ad;Φ)=𝒮M(ad;Φ)+O(𝒮R(ad;Φ)){\mathcal{S}}(a_{d};\Phi)={\mathcal{S}}_{M}(a_{d};\Phi)+O({\mathcal{S}}_{R}(a_{d};\Phi)), where

𝒮M(ad;Φ)=\displaystyle{\mathcal{S}}_{M}(a_{d};\Phi)= 𝒮M,X,Y(ad;Φ)=1X(d,2)=1MY(d)adΦ(N(d)X),\displaystyle{\mathcal{S}}_{M,X,Y}(a_{d};\Phi)=\frac{1}{X}\sum_{(d,2)=1}M_{Y}(d)a_{d}\Phi\biggl{(}\frac{N(d)}{X}\biggr{)},
𝒮R(ad;Φ)=\displaystyle{\mathcal{S}}_{R}(a_{d};\Phi)= 𝒮R,X,Y(ad;Φ)=1X(d,2)=1|RY(d)adΦ(N(d)X)|.\displaystyle{\mathcal{S}}_{R,X,Y}(a_{d};\Phi)=\frac{1}{X}\sum_{(d,2)=1}\Big{|}R_{Y}(d)a_{d}\Phi\Big{(}\frac{N(d)}{X}\Big{)}\Big{|}.

Let ε>0\varepsilon>0 be a small number and let b,Mb,M be two parameters such that b[ϵ,1ϵ]b\in[\epsilon,1-\epsilon] and XϵMXX^{\epsilon}\leq M\leq X. Also, let P(x)P(x) be a polynomial satisfying P(0)=P(0)=0P(0)=P^{\prime}(0)=0 and P(b)=1P(b)=1, P(b)=0P^{\prime}(b)=0. We define a sequence {λ(n)}n𝒪K\{\lambda(n)\}_{n\in\mathcal{O}_{K}} such that when n1mod(1+i)3n\equiv 1\bmod(1+i)^{3} and N(n)MN(n)\leq M, we have

(3.1) λ(n):=μ[i](n)Q(log(M/N(n))logM):={μ[i](n)if N(n)M1b,μ[i](n)P(log(M/N(n))logM)if M1bN(n)M.\displaystyle\lambda(n):=\mu_{[i]}(n)Q\Big{(}\frac{\log(M/N(n))}{\log M}\Big{)}:=\begin{cases}\mu_{[i]}(n)&\text{if }N(n)\leq M^{1-b},\\ \mu_{[i]}(n)P(\frac{\log(M/N(n))}{\log M})&\text{if }M^{1-b}\leq N(n)\leq M.\\ \end{cases}

For other values of nn, we define λ(n)=0\lambda(n)=0. The definition above then implies that λ(n)N(n)ϵ\lambda(n)\ll N(n)^{\epsilon} for all nn. We use the λ(n)\lambda(n) to define for any odd d𝒪Kd\in\mathcal{O}_{K}, the following mollifier function

(3.2) M(s,d)=N(n)Mn1mod(1+i)3λ(n)N(n)sχ(1+i)5d(n).\displaystyle M(s,d)=\sum_{\begin{subarray}{c}N(n)\leq M\\ n\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{\lambda(n)}{N(n)^{s}}\chi_{(1+i)^{5}d}(n).

Further, we define for any complex number δ1\delta_{1},

𝒲(δ1,Φ)=𝒮(|L(12+δ1,χ(1+i)5d)M(12+δ1,d)|2;Φ)𝒮(1;Φ).{\mathcal{W}}(\delta_{1},\Phi)=\frac{{\mathcal{S}}(|L(\tfrac{1}{2}+\delta_{1},\chi_{(1+i)^{5}d})M(\tfrac{1}{2}+\delta_{1},d)|^{2};\Phi)}{{\mathcal{S}}(1;\Phi)}.

We now apply Lemma 3.1 to the function f(s,d):=L(s,χ(1+i)5d)M(s,d)f(s,d):=L(s,\chi_{(1+i)^{5}d})M(s,d) with W0=12RlogXW_{0}=\frac{1}{2}-\frac{R}{\log X}, H=SlogXH=\frac{S}{\log X}, and W1=σ0:=1+3loglogM/logMW_{1}=\sigma_{0}:=1+3\log\log M/\log M and R,S(ϵ,1/ϵ)R,S\in(\epsilon,1/\epsilon) to be chosen later, such that f(s,d)f(s,d) has no zeros in (s)>σ0\Re(s)>\sigma_{0}. Arguing as in [C&S, Section 2], we deduce that

(3.3) 𝒩(X,Φ)X𝒮(1;Φ)8Ssinh(πR2S)(J1(X;Φ)+J2(X;Φ))+X8Ssinh(πR2S)𝒮(I(d);Φ),\displaystyle{\mathcal{N}}(X,\Phi)\leq\frac{X{\mathcal{S}}(1;\Phi)}{8S\sinh(\frac{\pi R}{2S})}\Big{(}J_{1}(X;\Phi)+J_{2}(X;\Phi)\Big{)}+\frac{X}{8S\sinh(\frac{\pi R}{2S})}{\mathcal{S}}(I(d);\Phi),

where

(3.4) J1(X;Φ)=0Scos(πt2S)log𝒲(RlogX+itlogX;Φ)𝑑t,J2(X;Φ)=R(σ012)logXsinh(π(x+R)2S)log𝒲(xlogX+iSlogX;Φ)𝑑x,I(d)=SScos(π(σ01/2)logX+R+it2iS)logf(σ0+itlogX,d)𝑑t,\displaystyle\begin{split}J_{1}(X;\Phi)=&\int_{0}^{S}\cos\Big{(}\frac{\pi t}{2S}\Big{)}\log{\mathcal{W}}\Big{(}-\frac{R}{\log X}+i\frac{t}{\log X};\Phi\Big{)}dt,\\ J_{2}(X;\Phi)=&\int_{-R}^{(\sigma_{0}-\frac{1}{2})\log X}\sinh\Big{(}\frac{\pi(x+R)}{2S}\Big{)}\log{\mathcal{W}}\Big{(}\frac{x}{\log X}+i\frac{S}{\log X};\Phi\Big{)}dx,\\ I(d)=&-\Re\int_{-S}^{S}\cos\Big{(}\pi\frac{(\sigma_{0}-1/2)\log X+R+it}{2iS}\Big{)}\log f\Big{(}\sigma_{0}+i\frac{t}{\log X},d\Big{)}dt,\end{split}

It remains to estimate the quantities on the right-hand side of (3.3). We first note the following result concerning 𝒮(I(d);Φ){\mathcal{S}}(I(d);\Phi), whose proof is given in Section 4.1.

Proposition 3.2.

Let Φ\Phi be a smooth function supported on [1,2][1,2] such that 0Φ(t)10\leq\Phi(t)\ll 1 and 12Φ(t)𝑑t1\int_{1}^{2}\Phi(t)dt\gg 1. For large X>0X>0, MX,σ0=1+3loglogM/logMM\leq\sqrt{X},\sigma_{0}=1+3\log\log M/\log M and ϵ(δ1)<3/4\epsilon\leq\Re(\delta_{1})<3/4, we have

𝒲(δ1,Φ)=1+O(Φ(2)Xϵ(M2(δ1)(1b)+M(12(δ1))(1b)X12)).{\mathcal{W}}(\delta_{1},\Phi)=1+O(\Phi_{(2)}X^{\epsilon}(M^{-2\Re(\delta_{1})(1-b)}+M^{(\frac{1}{2}-\Re(\delta_{1}))(1-b)}X^{-\frac{1}{2}})).

Moreover, the function f(s,d)=L(s,χ(1+i)5d)M(s,d)f(s,d)=L(s,\chi_{(1+i)^{5}d})M(s,d) has no zeros in (s)>σ0\Re(s)>\sigma_{0}, and we have

𝒮(I(d);Φ)exp(π(1/2+ϵ)logX2S)M(1b)Xϵ.{\mathcal{S}}(I(d);\Phi)\ll\exp\Big{(}\pi\frac{(1/2+\epsilon)\log X}{2S}\Big{)}M^{-(1-b)}X^{\epsilon}.

Next, notice that an evaluation on the terms in (3.3) involving with J1(X;Φ)J_{1}(X;\Phi) and J2(X;Φ)J_{2}(X;\Phi) requires one to study 𝒲(δ1,Φ){\mathcal{W}}(\delta_{1},\Phi). As Proposition 3.2 enables us to treat 𝒲(δ1,Φ){\mathcal{W}}(\delta_{1},\Phi) when (δ1)\Re(\delta_{1}) is slightly away from 0, we may focus on the case when δ1\delta_{1} is near 0. For this purpose, we shall evaluate more generally the following expression:

(3.5) 𝒮(ξ(12+δ1,χ(1+i)5d)ξ(12+δ2,χ(1+i)5d)M(12+δ1,d)M(12+δ2,d);Ψ),\displaystyle{\mathcal{S}}(\xi(\tfrac{1}{2}+\delta_{1},\chi_{(1+i)^{5}d})\xi(\tfrac{1}{2}+\delta_{2},\chi_{(1+i)^{5}d})M(\tfrac{1}{2}+\delta_{1},d)M(\tfrac{1}{2}+\delta_{2},d);\Psi),

where ξ\xi is given in (2.5), Ψ\Psi is a smooth function supported on [1,2][1,2] and δ1,δ2\delta_{1},\delta_{2} satisfy the conditions given in Section 2.3. Later, we shall set δ2=δ1¯\delta_{2}=\overline{\delta_{1}} and Ψ(t)=Φ(t)tτ\Psi(t)=\Phi(t)t^{-\tau} in (3.5) to retrieve the expression 𝒮(1;Φ)(25X/π2)τΓδ(τ)𝒲(δ1,Φ){\mathcal{S}}(1;\Phi)(2^{5}X/\pi^{2})^{\tau}\Gamma_{\delta}(\tau){\mathcal{W}}(\delta_{1},\Phi).

In order to evaluate (3.5), we apply the approximate functional equation for ξ(12+δ1,χ(1+i)5d)ξ(12+δ2,χ(1+i)5d)\xi(\frac{1}{2}+\delta_{1},\chi_{(1+i)^{5}d})\xi(\frac{1}{2}+\delta_{2},\chi_{(1+i)^{5}d}) to see that we may recast the expression in (3.5) as

𝒮M(Aδ,τ(d)M(12+δ1,d)M(12+δ2,d);Ψ)+𝒮R(Aδ,τ(d)M(12+δ1,d)M(12+δ2,d);Ψ).\displaystyle{\mathcal{S}}_{M}(A_{\delta,\tau}(d)M(\tfrac{1}{2}+\delta_{1},d)M(\tfrac{1}{2}+\delta_{2},d);\Psi)+{\mathcal{S}}_{R}(A_{\delta,\tau}(d)M(\tfrac{1}{2}+\delta_{1},d)M(\tfrac{1}{2}+\delta_{2},d);\Psi).

In Section 4.2, we obtain the following estimation for 𝒮R{\mathcal{S}}_{R}.

Proposition 3.3.

With the above notations, we have for MXM\leq\sqrt{X},

𝒮R(Aδ,τ(d)M(12+δ1,d)M(12+δ2,d);Ψ)Xκ+ϵ(1Y+M(δ1)+M(δ2)+M2(τ)Y12+M12(τ)X12).{\mathcal{S}}_{R}(A_{\delta,\tau}(d)M(\tfrac{1}{2}+\delta_{1},d)M(\tfrac{1}{2}+\delta_{2},d);\Psi)\ll X^{\kappa+\epsilon}\Big{(}\frac{1}{Y}+\frac{M^{-\Re(\delta_{1})}+M^{-\Re(\delta_{2})}+M^{-2\Re(\tau)}}{Y^{\frac{1}{2}}}+\frac{M^{1-2\Re(\tau)}}{X^{\frac{1}{2}}}\Big{)}.

The treatment on 𝒮M(Aδ,τ(d)M(12+δ1,d)M(12+δ2,d);Ψ){\mathcal{S}}_{M}(A_{\delta,\tau}(d)M(\tfrac{1}{2}+\delta_{1},d)M(\frac{1}{2}+\delta_{2},d);\Psi) is more involved. Our proof in fact requires us to evaluate more generally 𝒮M(Aδ,τ(d)((1+i)5dl);Ψ){\mathcal{S}}_{M}(A_{\delta,\tau}(d)\left(\frac{(1+i)^{5}d}{l}\right);\Psi) for any odd primary l𝒪Kl\in\mathcal{O}_{K}. To state our result, we introduce a few notations. We define for any two complex numbers ss and ww,

(3.6) Z(s;w)=ζK(s2w)ζK(s)ζK(s+2w).\displaystyle Z(s;w)=\zeta_{K}(s-2w)\zeta_{K}(s)\zeta_{K}(s+2w).

Our evaluation on SMS_{M} is given in the following result.

Proposition 3.4.

With the above notations and writing any odd primary l𝒪Kl\in\mathcal{O}_{K} as l=l1l22l=l_{1}l_{2}^{2}, with l1l_{1} being primary, square-free and l2Gl_{2}\in G, we have

(3.7) 𝒮M(((1+i)5dl)Aδ,τ(d);Ψ)=23ζK(2)N(l1)μ=±(rδ(l1)Γδ(μτ)(25XN(l1)π2)μτΨ^(μτ+1)Z(1+2μτ;δ)ηδ(1+2μτ;l)+rτ(l1)Γτ(μδ)(25XN(l1)π2)μδΨ^(μδ+1)Z(1+2μδ;τ)ητ(1+2μδ;l))+(l)+O(|rδ(l1)|Xϵ(XN(l1))14+N(l)ϵXκ+ϵN(l1)2κ12Y14κ).\displaystyle\begin{split}{\mathcal{S}}_{M}\Big{(}\left(\frac{(1+i)^{5}d}{l}\right)A_{\delta,\tau}(d);\Psi\Big{)}=&\frac{2}{3\zeta_{K}(2)\sqrt{N(l_{1})}}\sum_{\mu=\pm}\Big{(}r_{\delta}(l_{1})\Gamma_{\delta}(\mu\tau)\left(\frac{2^{5}X}{N(l_{1})\pi^{2}}\right)^{\mu\tau}{\widehat{\Psi}}(\mu\tau+1)Z(1+2\mu\tau;\delta)\eta_{\delta}(1+2\mu\tau;l)\\ &+r_{\tau}(l_{1})\Gamma_{\tau}(\mu\delta)\left(\frac{2^{5}X}{N(l_{1})\pi^{2}}\right)^{\mu\delta}{\widehat{\Psi}}(\mu\delta+1)Z(1+2\mu\delta;\tau)\eta_{\tau}(1+2\mu\delta;l)\Big{)}\\ &+{\mathcal{R}}(l)+O\Big{(}\frac{|r_{\delta}(l_{1})|X^{\epsilon}}{(XN(l_{1}))^{\frac{1}{4}}}+\frac{N(l)^{\epsilon}X^{\kappa+\epsilon}N(l_{1})^{2\kappa-\frac{1}{2}}}{Y^{1-4\kappa}}\Big{)}.\end{split}

Here (l){\mathcal{R}}(l) is a remainder term bounded on average by

(3.8) l1mod(1+i)3LN(l)2L1|(l)|(L1+ϵY1+ϵX12|(δ)|ϵ+L1+κ+ϵY2κ+ϵX12|(δ)|ϵ)Ψ(3)Ψ(4)ϵ.\displaystyle\sum_{\begin{subarray}{c}l\equiv 1\bmod(1+i)^{3}\\ L\leq N(l)\leq 2L-1\end{subarray}}|{\mathcal{R}}(l)|\ll\Big{(}\frac{L^{1+\epsilon}Y^{1+\epsilon}}{X^{\frac{1}{2}-|\Re(\delta)|-\epsilon}}+\frac{L^{1+\kappa+\epsilon}Y^{2\kappa+\epsilon}}{X^{\frac{1}{2}-|\Re(\delta)|-\epsilon}}\Big{)}\Psi_{(3)}\Psi_{(4)}^{\epsilon}.

With Propositions 3.2-3.4 available, we are able to obtain the following result concerning 𝒲(δ1,Φ){\mathcal{W}}(\delta_{1},\Phi) for |δ1||\delta_{1}| being small.

Proposition 3.5.

Let Φ\Phi be a non-negative smooth function on [1,2][1,2] satisfying Φ(t)1\Phi(t)\ll 1 and 12Φ(t)𝑑t1\int_{1}^{2}\Phi(t)dt\gg 1. Let δ1\delta_{1} be a complex number such that (δ1)1ϵlogX\Re(\delta_{1})\geq-\frac{1}{\epsilon\log X}, κ|δ1|ϵlogX\kappa\geq|\delta_{1}|\geq\frac{\epsilon}{\log X}. We take δ2=δ1¯\delta_{2}=\overline{\delta_{1}} so that τ=(δ1)\tau=\Re(\delta_{1}), and δ=i(δ1)\delta=i\Im(\delta_{1}). Then with the mollifier function being given in (3.2), we have for M=X125κM=X^{\frac{1}{2}-5\kappa},

𝒲(δ1,Φ)=\displaystyle{\mathcal{W}}(\delta_{1},\Phi)= 1+(1(25X/π2)2τ2τlogM(25Xπ2)τ(25X/π2)δ(25X/π2)δ2δlogM)0bM2τ(1x)|Q(x)+Q′′(x)2δ1logM|2𝑑x\displaystyle 1+\Big{(}\frac{1-(2^{5}X/\pi^{2})^{-2\tau}}{2\tau\log M}-\left(\frac{2^{5}X}{\pi^{2}}\right)^{-\tau}\frac{(2^{5}X/\pi^{2})^{\delta}-(2^{5}X/\pi^{2})^{-\delta}}{2\delta\log M}\Big{)}\int_{0}^{b}M^{-2\tau(1-x)}\Big{|}Q^{\prime}(x)+\frac{Q^{\prime\prime}(x)}{2\delta_{1}\log M}\Big{|}^{2}dx
+O(XκϵΦ(3)Φ(4)ϵ+M2τ(1b)|δ1|6log5X).\displaystyle+O(X^{-\kappa-\epsilon}\Phi_{(3)}\Phi_{(4)}^{\epsilon}+M^{-2\tau(1-b)}|\delta_{1}|^{6}\log^{5}X).

The proofs of Propositions 3.2-3.5 will occupy major parts in the rest of the paper. In the end, the proof of Theorem 1.1 follows by gathering these results together with suitable choices for the parameters involved.

4. Proofs of Propositions 3.2 and 3.3

4.1. Proof of Proposition 3.2

In this section we estimate 𝒮(I(d);Φ){\mathcal{S}}(I(d);\Phi) by proving Proposition 3.2. We note first that similar to the bounds given in [C&S, (4.1), (4.2)], it follows from Lemma 2.7 that for any odd l𝒪Kl\in\mathcal{O}_{K} such that N(l)2XN(l)\leq\sqrt{2X}, we have

(4.1) XN(d)2Xμ[i]2((1+i)d)|M(s,d)|4Xϵ(X+XM2(12(s))+M4(1(s))),X/N(l)2N(m)2X/N(l)2μ[i]2((1+i)m)|M(s,l2m)|4Xϵ(XN(l)2+XN(l)2M2(12(s))+M4(1(s))).\displaystyle\begin{split}\sum_{X\leq N(d)\leq 2X}\mu^{2}_{[i]}((1+i)d)|M(s,d)|^{4}\ll&X^{\epsilon}(X+XM^{2(1-2\Re(s))}+M^{4(1-\Re(s))}),\\ \sum_{X/N(l)^{2}\leq N(m)\leq 2X/N(l)^{2}}\mu^{2}_{[i]}((1+i)m)|M(s,l^{2}m)|^{4}\ll&X^{\epsilon}\Big{(}\frac{X}{N(l)^{2}}+\frac{X}{N(l)^{2}}M^{2(1-2\Re(s))}+M^{4(1-\Re(s))}\Big{)}.\end{split}

Now, we write B(s,d)=L(s,χ(1+i)5d)M(s,d)1B(s,d)=L(s,\chi_{(1+i)^{5}d})M(s,d)-1 and we observe that it follows from [G&Zhao4, Section 3.1] and partial summation that we have

(4.2) 𝒮(1;Φ)=2πΦ^(1)3ζK(2)+O(X1/2)1.\displaystyle\begin{split}{\mathcal{S}}(1;\Phi)=\frac{2\pi{\widehat{\Phi}}(1)}{3\zeta_{K}(2)}+O(X^{-1/2})\gg 1.\end{split}

Using the above, we see that

(4.3) 𝒲(δ1,Φ)=1+O(𝒮(B(12+δ1,d);Φ)+𝒮(|B(12+δ1,d)|2;Φ)).\displaystyle{\mathcal{W}}(\delta_{1},\Phi)=1+O({\mathcal{S}}(B(\tfrac{1}{2}+\delta_{1},d);\Phi)+{\mathcal{S}}(|B(\tfrac{1}{2}+\delta_{1},d)|^{2};\Phi)).\

To estimate the error terms above, for any real number c>12(δ1)c>\frac{1}{2}-\Re(\delta_{1}), consider the integral

12πi(c)Γ(s)B(12+δ1+s,d)Xs𝑑s.\frac{1}{2\pi i}\int\limits_{(c)}\Gamma(s)B(\tfrac{1}{2}+\delta_{1}+s,d)X^{s}ds.

By moving the line of integration above to (s)=(δ1)\Re(s)=-\Re(\delta_{1}), we see that the pole at s=0s=0 contributes B(12+δ1;d)B(\frac{1}{2}+\delta_{1};d). This implies that

B(12+δ1,d)=\displaystyle B(\tfrac{1}{2}+\delta_{1},d)= 12πi(c)Γ(s)B(12+δ1+s,d)Xs𝑑s12πi((δ1))Γ(s)B(12+δ1+s,d)Xs𝑑s\displaystyle\frac{1}{2\pi i}\int\limits_{(c)}\Gamma(s)B(\tfrac{1}{2}+\delta_{1}+s,d)X^{s}ds-\frac{1}{2\pi i}\int\limits_{(-\Re(\delta_{1}))}\Gamma(s)B(\tfrac{1}{2}+\delta_{1}+s,d)X^{s}ds
:=\displaystyle:= T1(12+δ1,d)T2(12+δ1,d).\displaystyle T_{1}(\frac{1}{2}+\delta_{1},d)-T_{2}(\frac{1}{2}+\delta_{1},d).

We apply the Cauchy-Schwarz inequality to see that

|T2(12+δ1,d)|2X2(δ1)(((δ1))|Γ(s)B(12+δ1+s,d)2ds|)(((δ1))|Γ(s)ds|),|T_{2}(\tfrac{1}{2}+\delta_{1},d)|^{2}\ll X^{-2\Re(\delta_{1})}\Big{(}\int\limits_{(-\Re(\delta_{1}))}|\Gamma(s)B(\tfrac{1}{2}+\delta_{1}+s,d)^{2}ds|\Big{)}\Big{(}\int\limits_{(-\Re(\delta_{1}))}|\Gamma(s)ds|\Big{)},

It follows from this and the rapid decay of |Γ(s)||\Gamma(s)| when |(s)||\Im(s)|\to\infty that we have

|T2(12+δ1,d)|2X2(δ1)(1+((δ1))|Γ(s)||L(12+δ1+s,χ8d)M(12+δ1+s,d)|2|ds|).|T_{2}(\tfrac{1}{2}+\delta_{1},d)|^{2}\ll X^{-2\Re(\delta_{1})}\Big{(}1+\int\limits_{(-\Re(\delta_{1}))}|\Gamma(s)||L(\tfrac{1}{2}+\delta_{1}+s,\chi_{-8d})M(\tfrac{1}{2}+\delta_{1}+s,d)|^{2}|ds|\Big{)}.

Applying the Cauchy-Schwarz inequality again, we see that

𝒮(|T2(12+δ1,d)|2;Φ)X2(δ1)(1+((δ1))|Γ(s)|𝒮(|L(12+δ1+s,χ8d)|4;Φ)12𝒮(|M(12+δ1+s,d)|4;Φ)12|ds|),{\mathcal{S}}(|T_{2}(\frac{1}{2}+\delta_{1},d)|^{2};\Phi)\ll X^{-2\Re(\delta_{1})}\Big{(}1+\int\limits_{(-\Re(\delta_{1}))}|\Gamma(s)|{\mathcal{S}}(|L(\tfrac{1}{2}+\delta_{1}+s,\chi_{-8d})|^{4};\Phi)^{\frac{1}{2}}{\mathcal{S}}(|M(\tfrac{1}{2}+\delta_{1}+s,d)|^{4};\Phi)^{\frac{1}{2}}|ds|\Big{)},

Applying Lemma 2.8 and (4.1) in the above estimation, we see that for MXM\leq\sqrt{X},

(4.4) 𝒮(|T2(12+δ1,d)|2;Φ)X2(δ1)+ϵ.\displaystyle{\mathcal{S}}(|T_{2}(\tfrac{1}{2}+\delta_{1},d)|^{2};\Phi)\ll X^{-2\Re(\delta_{1})+\epsilon}.

With one more application of the Cauchy-Schwarz inequality, we deduce from the above that

(4.5) 𝒮(|T2(12+δ1,d)|;Φ)X(δ1)+ϵ.\displaystyle{\mathcal{S}}(|T_{2}(\tfrac{1}{2}+\delta_{1},d)|;\Phi)\ll X^{-\Re(\delta_{1})+\epsilon}.

Now, we consider the contribution from T1T_{1}. For (s)>1\Re(s)>1, we write

B(s,d)=n1mod(1+i)3b(n)N(n)s((1+i)5dn).B(s,d)=\sum_{\begin{subarray}{c}n\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{b(n)}{N(n)^{s}}\left(\frac{(1+i)^{5}d}{n}\right).

By (3.1) and (3.2), we see that b(n)=0b(n)=0 for all N(n)M1bN(n)\leq M^{1-b} and that |b(n)|N(n)ϵd[i](n)N(n)ϵ|b(n)|\ll N(n)^{\epsilon}d_{[i]}(n)\ll N(n)^{\epsilon} for all nn, where we denote d[i](n)d_{[i]}(n) for the analogue on 𝒪K\mathcal{O}_{K} of the usual divisor function dd on \mathbb{Z}. Moreover, as λ\lambda is supported on square-free numbers, we have b(m2)=d|m2d1mod(1+i)3λ(d)=d|md1mod(1+i)3λ(d)=b(m)b(m^{2})=\sum_{\begin{subarray}{c}d|m^{2}\\ d\equiv 1\bmod(1+i)^{3}\end{subarray}}\lambda(d)=\sum_{\begin{subarray}{c}d|m\\ d\equiv 1\bmod(1+i)^{3}\end{subarray}}\lambda(d)=b(m). It follows that we have b(n)=0b(n)=0 for all primary square values N(n)M2(1b)N(n)\leq M^{2(1-b)}. Now, we have

(4.6) T1(12+δ1,d)=12πi(c)Γ(s)n1mod(1+i)3M1bN(n)Xlog2Xb(n)N(n)12+δ1+s((1+i)5dn)Xsds+n1mod(1+i)3N(n)>Xlog2Xb(n)N(n)12+δ1((1+i)5dn)(12πi(c)Γ(s)(XN(n))s𝑑s).\displaystyle\begin{split}T_{1}(\tfrac{1}{2}+\delta_{1},d)=&\frac{1}{2\pi i}\int\limits_{(c)}\Gamma(s)\sum_{\begin{subarray}{c}n\equiv 1\bmod(1+i)^{3}\\ M^{1-b}\leq N(n)\leq X\log^{2}X\end{subarray}}\frac{b(n)}{N(n)^{\frac{1}{2}+\delta_{1}+s}}\left(\frac{(1+i)^{5}d}{n}\right)X^{s}ds\\ &+\sum_{\begin{subarray}{c}n\equiv 1\bmod(1+i)^{3}\\ N(n)>X\log^{2}X\end{subarray}}\frac{b(n)}{N(n)^{\frac{1}{2}+\delta_{1}}}\left(\frac{(1+i)^{5}d}{n}\right)\Big{(}\frac{1}{2\pi i}\int\limits_{(c)}\Gamma(s)\left(\frac{X}{N(n)}\right)^{s}ds\Big{)}.\end{split}

Using the result that

12πi(c)Γ(s)(XN(n))s𝑑s=exp(N(n)/(20X)),\frac{1}{2\pi i}\int\limits_{(c)}\Gamma(s)\left(\frac{X}{N(n)}\right)^{s}ds=\exp(-N(n)/(20X)),

we see that the second term on the right side of (4.6) contributes X5\ll X^{-5}. We then move the line of integration in the first term on the right side of (4.6)to (s)=1logX\Re(s)=\frac{1}{\log X} to see that

(4.7) T1(12+δ1,d)=12πi(1logX)Γ(s)n1mod(1+i)3M1bN(n)Xlog2Xb(n)N(n)12+δ1+s((1+i)5dn)Xsds+O(X5).\displaystyle T_{1}(\tfrac{1}{2}+\delta_{1},d)=\frac{1}{2\pi i}\int\limits_{(\frac{1}{\log X})}\Gamma(s)\sum_{\begin{subarray}{c}n\equiv 1\bmod(1+i)^{3}\\ M^{1-b}\leq N(n)\leq X\log^{2}X\end{subarray}}\frac{b(n)}{N(n)^{\frac{1}{2}+\delta_{1}+s}}\left(\frac{(1+i)^{5}d}{n}\right)X^{s}ds+O(X^{-5}).

It follows from this and the Cauchy-Schwarz inequality that we have

|T1(12+δ1,d)|2\displaystyle|T_{1}(\tfrac{1}{2}+\delta_{1},d)|^{2} X10+((1logX)|Γ(s)||n1mod(1+i)3M1bN(n)Xlog2Xb(n)N(n)12+δ1+s((1+i)5dn)|2|ds|)((1logX)|Γ(s)ds|)\displaystyle\ll X^{-10}+\Big{(}\int\limits_{(\frac{1}{\log X})}|\Gamma(s)|\Big{|}\sum_{\begin{subarray}{c}n\equiv 1\bmod(1+i)^{3}\\ M^{1-b}\leq N(n)\leq X\log^{2}X\end{subarray}}\frac{b(n)}{N(n)^{\frac{1}{2}+\delta_{1}+s}}\left(\frac{(1+i)^{5}d}{n}\right)\Big{|}^{2}|ds|\Big{)}\Big{(}\int\limits_{(\frac{1}{\log X})}|\Gamma(s)ds|\Big{)}
X10+Xϵ(1logX)|Γ(s)||n1mod(1+i)3M1bN(n)Xlog2Xb(n)N(n)12+δ1+s((1+i)5dn)|2|ds|.\displaystyle\ll X^{-10}+X^{\epsilon}\int\limits_{(\frac{1}{\log X})}|\Gamma(s)|\Big{|}\sum_{\begin{subarray}{c}n\equiv 1\bmod(1+i)^{3}\\ M^{1-b}\leq N(n)\leq X\log^{2}X\end{subarray}}\frac{b(n)}{N(n)^{\frac{1}{2}+\delta_{1}+s}}\left(\frac{(1+i)^{5}d}{n}\right)\Big{|}^{2}|ds|.

We split the sum over nn above into dyadic blocks and apply Lemma 2.7 to see that

𝒮(|T1(12+δ1,d)|2;Φ)M2(δ1)(1b)Xϵ.{\mathcal{S}}(|T_{1}(\tfrac{1}{2}+\delta_{1},d)|^{2};\Phi)\ll M^{-2\Re(\delta_{1})(1-b)}X^{\epsilon}.

Combining this with (4.4), we deduce that

(4.8) 𝒮(|B(12+δ1,d)|2;Φ)M2(δ1)(1b)Xϵ.\displaystyle{\mathcal{S}}(|B(\tfrac{1}{2}+\delta_{1},d)|^{2};\Phi)\ll M^{-2\Re(\delta_{1})(1-b)}X^{\epsilon}.

Next, we bound 𝒮(T1(12+δ1,d);Φ){\mathcal{S}}(T_{1}(\tfrac{1}{2}+\delta_{1},d);\Phi) by applying (4.7) to see that

(4.9) 𝒮(T1(12+δ1,d);Φ)X5+Xϵn1mod(1+i)3M1bN(n)Xlog2X|b(n)|N(n)12+(δ1)|𝒮(χ(1+i)5d(n);Φ)|.\displaystyle{\mathcal{S}}(T_{1}(\tfrac{1}{2}+\delta_{1},d);\Phi)\ll X^{-5}+X^{\epsilon}\sum_{\begin{subarray}{c}n\equiv 1\bmod(1+i)^{3}\\ M^{1-b}\leq N(n)\leq X\log^{2}X\end{subarray}}\frac{|b(n)|}{N(n)^{\frac{1}{2}+\Re(\delta_{1})}}\Big{|}{\mathcal{S}}\big{(}\chi_{(1+i)^{5}d}(n);\Phi\big{)}\Big{|}.

We now apply the Mellin transform to see that for any c>1c>1,

𝒮(χ(1+i)5d(n);Φ)=\displaystyle{\mathcal{S}}\big{(}\chi_{(1+i)^{5}d}(n);\Phi\big{)}= ((1+i)5n)2πi(c)d1mod(1+i)3μ[i]2((1+i)d)χn(d)N(d)wXw1Φ^(w)dw\displaystyle\frac{\left(\frac{(1+i)^{5}}{n}\right)}{2\pi i}\int\limits_{(c)}\sum_{d\equiv 1\bmod(1+i)^{3}}\frac{\mu^{2}_{[i]}((1+i)d)\chi_{n}(d)}{N(d)^{w}}X^{w-1}{\widehat{\Phi}}(w)dw
=\displaystyle= ((1+i)5n)2πi(c)L(w,χn)L(2w,χn)(1+χn(1+i)2w)1Xw1Φ^(w)𝑑w.\displaystyle\frac{\left(\frac{(1+i)^{5}}{n}\right)}{2\pi i}\int\limits_{(c)}\frac{L(w,\chi_{n})}{L(2w,\chi_{n})}(1+\frac{\chi_{n}(1+i)}{2^{w}})^{-1}X^{w-1}{\widehat{\Phi}}(w)dw.

Moving the line of integration above to (w)=12+1logX\Re(w)=\frac{1}{2}+\frac{1}{\log X}, we see that we encounter a pole at w=1w=1 if and only if n=n=\square and the corresponding contribution of the residue is 1\ll 1. It follows that

|𝒮(χ(1+i)5d(n);Φ)|δ(n=)+X12+ϵ(12+1logX)|L(w,χn)||Φ^(w)||dw|,|{\mathcal{S}}\big{(}\chi_{(1+i)^{5}d}(n);\Phi\big{)}|\ll\delta(n=\square)+X^{-\frac{1}{2}+\epsilon}\int\limits_{(\frac{1}{2}+\frac{1}{\log X})}|L(w,\chi_{n})||{\widehat{\Phi}}(w)||dw|,

where δ(n=)\delta(n=\square) is 11 if n=n=\square and 0 otherwise. As b(n)=0b(n)=0 for all perfect squares with norm M2(1b)\leq M^{2(1-b)}, we deduce that

(4.10) n1mod(1+i)3M1bN(n)Xlog2X|b(n)|N(n)12+(δ1)|𝒮(χ(1+i)5d(n);Φ)|XϵM2(δ1)(1b)+X12+ϵ(12+1logX)n1mod(1+i)3M1bN(n)Xlog2X1N(n)12+(δ1)|L(w,χn)||Φ^(w)||dw|.\displaystyle\begin{split}&\sum_{\begin{subarray}{c}n\equiv 1\bmod(1+i)^{3}\\ M^{1-b}\leq N(n)\leq X\log^{2}X\end{subarray}}\frac{|b(n)|}{N(n)^{\frac{1}{2}+\Re(\delta_{1})}}\Big{|}{\mathcal{S}}\big{(}\chi_{(1+i)^{5}d}(n);\Phi\big{)}\Big{|}\\ \ll&X^{\epsilon}M^{-2\Re(\delta_{1})(1-b)}+X^{-\frac{1}{2}+\epsilon}\int\limits_{(\frac{1}{2}+\frac{1}{\log X})}\sum_{\begin{subarray}{c}n\equiv 1\bmod(1+i)^{3}\\ M^{1-b}\leq N(n)\leq X\log^{2}X\end{subarray}}\frac{1}{N(n)^{\frac{1}{2}+\Re(\delta_{1})}}|L(w,\chi_{n})||{\widehat{\Phi}}(w)||dw|.\end{split}

Note that Lemma 2.8 implies that

n1mod(1+i)3NN(n)2N|L(w,χn)|N1+ϵ(1+|w|2)14+ϵ.\sum_{\begin{subarray}{c}n\equiv 1\bmod(1+i)^{3}\\ N\leq N(n)\leq 2N\end{subarray}}|L(w,\chi_{n})|\ll N^{1+\epsilon}(1+|w|^{2})^{\frac{1}{4}+\epsilon}.

Applying this together with (2.13) by setting g=Φg=\Phi and ν=2\nu=2 there, we obtain from (4.10) that

n1mod(1+i)3M1bN(n)Xlog2X|b(n)|N(n)12+(δ1)|𝒮(χ(1+i)5d(n);Φ)|XϵΦ(2)(M2(δ1)(1b)+X(δ1)+M(12(δ1))(1b)X12).\displaystyle\begin{split}&\sum_{\begin{subarray}{c}n\equiv 1\bmod(1+i)^{3}\\ M^{1-b}\leq N(n)\leq X\log^{2}X\end{subarray}}\frac{|b(n)|}{N(n)^{\frac{1}{2}+\Re(\delta_{1})}}\Big{|}{\mathcal{S}}\big{(}\chi_{(1+i)^{5}d}(n);\Phi\big{)}\Big{|}\ll X^{\epsilon}\Phi_{(2)}\Big{(}M^{-2\Re(\delta_{1})(1-b)}+X^{-\Re(\delta_{1})}+M^{(\frac{1}{2}-\Re(\delta_{1}))(1-b)}X^{-\frac{1}{2}}\Big{)}.\end{split}

We insert the above in (4.9) and further combine the result with (4.5) to deduce that for MXM\leq\sqrt{X},

(4.11) 𝒮(B(12+δ1,d);Φ)XϵΦ(2)(M2(δ1)(1b)+M(12(δ1))(1b)X12).\displaystyle{\mathcal{S}}(B(\frac{1}{2}+\delta_{1},d);\Phi)\ll X^{\epsilon}\Phi_{(2)}\Big{(}M^{-2\Re(\delta_{1})(1-b)}+M^{(\frac{1}{2}-\Re(\delta_{1}))(1-b)}X^{-\frac{1}{2}}\Big{)}.

The first statement of Proposition 3.2 now follows by applying (4.8) and (4.11) in (4.3).

To prove the second assertion, we note that

f(s,d)=1+B(s,d)=1+O(n1mod(1+i)3N(n)M1bd[i](n)N(n)(s)).f(s,d)=1+B(s,d)=1+O\Big{(}\sum_{\begin{subarray}{c}n\equiv 1\bmod(1+i)^{3}\\ N(n)\geq M^{1-b}\end{subarray}}\frac{d_{[i]}(n)}{N(n)^{\Re(s)}}\Big{)}.

We deduce readily from this that f(s,d)f(s,d) has no zeros to the right of σ0\sigma_{0}. Moreover, we have logf(s,d)=B(s,d)+O(|B(s,d)|2)\log f(s,d)=B(s,d)+O(|B(s,d)|^{2}) for (s)>σ0\Re(s)>\sigma_{0}. Thus, we obtain that

𝒮(I(d);Φ)exp(π(12+ϵ)logX2S)(|𝒮(B(s,d);Φ)|+𝒮(|B(s,d)|2;Φ)).{\mathcal{S}}(I(d);\Phi)\ll\exp\Big{(}\pi\frac{(\frac{1}{2}+\epsilon)\log X}{2S}\Big{)}\Big{(}|{\mathcal{S}}(B(s,d);\Phi)|+{\mathcal{S}}(|B(s,d)|^{2};\Phi)\Big{)}.

By applying (4.8) and (4.11), we see that the second statement of Proposition 3.2 now follows.

4.2. Proof of Proposition 3.3

In this section, we estimate 𝒮R(Aδ,τ(d)M(12+δ1,d)M(12+δ2,d);Ψ){\mathcal{S}}_{R}(A_{\delta,\tau}(d)M(\tfrac{1}{2}+\delta_{1},d)M(\tfrac{1}{2}+\delta_{2},d);\Psi) by proving Proposition 3.3. We first notice that |RY(d)|k|d1N(d)ϵ|R_{Y}(d)|\leq\sum_{k|d}1\ll N(d)^{\epsilon} and that RY(d)=0R_{Y}(d)=0 unless d=l2md=l^{2}m where mm is square-free and N(l)>YN(l)>Y. It follows that

(4.12) 𝒮R(Aδ,τ(d)M(12+δ1,d)M(12+δ2,d);Ψ)X1+ϵN(l)>Y(l,2)=1X/N(l)2N(m)2X/N(l)2|Aδ,τ(l2m)M(12+δ1,l2m)M(12+δ2,l2m)|,\displaystyle\begin{split}&{\mathcal{S}}_{R}(A_{\delta,\tau}(d)M(\tfrac{1}{2}+\delta_{1},d)M(\tfrac{1}{2}+\delta_{2},d);\Psi)\\ \ll&X^{-1+\epsilon}\sum_{\begin{subarray}{c}N(l)>Y\\ (l,2)=1\end{subarray}}\sideset{}{{}^{\flat}}{\sum}_{X/N(l)^{2}\leq N(m)\leq 2X/N(l)^{2}}|A_{\delta,\tau}(l^{2}m)M(\tfrac{1}{2}+\delta_{1},l^{2}m)M(\tfrac{1}{2}+\delta_{2},l^{2}m)|,\end{split}

where \sum^{\flat} means that the sum is over odd and square-free m𝒪Km\in\mathcal{O}_{K}. Now, applying twice the Cauchy-Schwarz inequality, we see that the sum over mm above is

(4.13) (m|M(12+δ1,l2m)|4)14(m|M(12+δ2,l2m)|4)14(m|Aδ,τ(l2m)|2)12.\displaystyle\ll\biggl{(}\sideset{}{{}^{\flat}}{\sum}_{m}|M(\tfrac{1}{2}+\delta_{1},l^{2}m)|^{4}\biggr{)}^{\frac{1}{4}}\biggl{(}\sideset{}{{}^{\flat}}{\sum}_{m}|M(\tfrac{1}{2}+\delta_{2},l^{2}m)|^{4}\biggr{)}^{\frac{1}{4}}\biggl{(}\sideset{}{{}^{\flat}}{\sum}_{m}|A_{\delta,\tau}(l^{2}m)|^{2}\biggr{)}^{\frac{1}{2}}.

Note that for any c>12+|(δ)|c>\frac{1}{2}+|\Re(\delta)|, we have

(4.14) Aδ,τ(l2m)=\displaystyle A_{\delta,\tau}(l^{2}m)= 12πi(c)Γδ(s)(25N(l2m)π2)s2ss2τ2n1mod(1+i)3rδ(n)N(n)s+12((1+i)5l2mn)ds.\displaystyle\frac{1}{2\pi i}\int\limits_{(c)}\Gamma_{\delta}(s)\left(\frac{2^{5}N(l^{2}m)}{\pi^{2}}\right)^{s}\frac{2s}{s^{2}-\tau^{2}}\sum_{n\equiv 1\bmod(1+i)^{3}}\frac{r_{\delta}(n)}{N(n)^{s+\frac{1}{2}}}\left(\frac{(1+i)^{5}l^{2}m}{n}\right)ds.

We write the sum above as

(4.15) n1mod(1+i)3rδ(n)N(n)s+12((1+i)5l2mn)=L(12+s+δ,χ(1+i)5m)L(12+sδ,χ(1+i)5m)(s,l),\displaystyle\sum_{n\equiv 1\bmod(1+i)^{3}}\frac{r_{\delta}(n)}{N(n)^{s+\frac{1}{2}}}\left(\frac{(1+i)^{5}l^{2}m}{n}\right)=L(\tfrac{1}{2}+s+\delta,\chi_{(1+i)^{5}m})L(\tfrac{1}{2}+s-\delta,\chi_{(1+i)^{5}m}){\mathcal{E}}(s,l),

with

(s,l)=ϖGϖ|l(11N(ϖ)s+12+δ((1+i)5mϖ))(11N(ϖ)s+12δ((1+i)5mϖ)).{\mathcal{E}}(s,l)=\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|l\end{subarray}}\biggl{(}1-\frac{1}{N(\varpi)^{s+\frac{1}{2}+\delta}}\left(\frac{(1+i)^{5}m}{\varpi}\right)\biggr{)}\biggl{(}1-\frac{1}{N(\varpi)^{s+\frac{1}{2}-\delta}}\left(\frac{(1+i)^{5}m}{\varpi}\right)\biggr{)}.

Observe that the left side of (4.15) is analytic for all ss since χ(1+i)5m\chi_{(1+i)^{5}m} is non-principal. Thus, we we may move the line of integration in (4.14) to (s)=κ+1/logX\Re(s)=\kappa+1/\log X without encountering any pole. Using the estimations that |(s,l)|ϖGϖ|l(1+1/N(ϖ))2N(l)ϵXϵ|{\mathcal{E}}(s,l)|\leq\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|l\end{subarray}}(1+1/\sqrt{N(\varpi)})^{2}\ll N(l)^{\epsilon}\ll X^{\epsilon}, 2s/(s2τ2)Xϵ2s/(s^{2}-\tau^{2})\ll X^{\epsilon}, and the rapid decay of |Γδ(s)||\Gamma_{\delta}(s)| when |(s)||\Im(s)|\rightarrow\infty, we apply the Cauchy-Schwarz inequality to see that

|Aδ,τ(l2m)|2X2κ+ϵ(κ+1logX)|Γδ(s)||L(12+s+δ,χ(1+i)5m)L(12+sδ,χ(1+i)5m)|2|ds|.|A_{\delta,\tau}(l^{2}m)|^{2}\ll X^{2\kappa+\epsilon}\int\limits_{(\kappa+\frac{1}{\log X})}|\Gamma_{\delta}(s)||L(\tfrac{1}{2}+s+\delta,\chi_{(1+i)^{5}m})L(\tfrac{1}{2}+s-\delta,\chi_{(1+i)^{5}m})|^{2}|ds|.

It follows from this and Lemma 2.8 that we have

X/N(l)2N(m)2X/N(l)2|Aδ,τ(l2m)|2X1+2κ+ϵN(l)2(κ+1logX)|Γδ(s)|(1+|s|2)1+ϵ|ds|X1+2κ+ϵN(l)2.\sideset{}{{}^{\flat}}{\sum}_{X/N(l)^{2}\leq N(m)\leq 2X/N(l)^{2}}|A_{\delta,\tau}(l^{2}m)|^{2}\ll\frac{X^{1+2\kappa+\epsilon}}{N(l)^{2}}\int\limits_{(\kappa+\frac{1}{\log X})}|\Gamma_{\delta}(s)|(1+|s|^{2})^{1+\epsilon}|ds|\ll\frac{X^{1+2\kappa+\epsilon}}{N(l)^{2}}.

We apply this with (4.1) to see that the expression in (4.13) is bounded by

X1/2+κ+ϵN(l)(X1/4N(l)1/2+X1/4N(l)1/2M(δ1)+M1/2(δ1))(X1/4N(l)1/2+X1/4N(l)1/2M(δ2)+M1/2(δ2)).\ll\frac{X^{1/2+\kappa+\epsilon}}{N(l)}\Big{(}\frac{X^{1/4}}{N(l)^{1/2}}+\frac{X^{1/4}}{N(l)^{1/2}}M^{-\Re(\delta_{1})}+M^{1/2-\Re(\delta_{1})}\Big{)}\Big{(}\frac{X^{1/4}}{N(l)^{1/2}}+\frac{X^{1/4}}{N(l)^{1/2}}M^{-\Re(\delta_{2})}+M^{1/2-\Re(\delta_{2})}\Big{)}.

Inserting the above in (4.12) and keeping in mind that MXM\leq\sqrt{X}, we see that the assertion of Proposition 3.3 follows.

5. Proof of Proposition 3.4

5.1. A first decomposition

Note that we have

(5.1) 𝒮M(((1+i)5dl)Aδ,τ(d);Ψ)=n1mod(1+i)3rδ(n)N(n)𝒮M(((1+i)5dln);FN(n)),\displaystyle{\mathcal{S}}_{M}\biggl{(}\left(\frac{(1+i)^{5}d}{l}\right)A_{\delta,\tau}(d);\Psi\biggr{)}=\sum_{n\equiv 1\bmod(1+i)^{3}}\frac{r_{\delta}(n)}{\sqrt{N(n)}}{\mathcal{S}}_{M}\biggl{(}\left(\frac{(1+i)^{5}d}{ln}\right);F_{N(n)}\biggr{)},

where FN(n)(t)F_{N(n)}(t) is defined as in (2.11).

By definition, we have

𝒮M(((1+i)5dln);FN(n))=\displaystyle{\mathcal{S}}_{M}\biggl{(}\left(\frac{(1+i)^{5}d}{ln}\right);F_{N(n)}\biggr{)}= 1X(d,2)=1(α1mod(1+i)3N(α)Yα2|dμ[i](α))((1+i)5dln)FN(n)(N(d)X)\displaystyle\frac{1}{X}\sum_{(d,2)=1}\Big{(}\sum_{\begin{subarray}{c}\alpha\equiv 1\bmod(1+i)^{3}\\ N(\alpha)\leq Y\\ \alpha^{2}|d\end{subarray}}\mu_{[i]}(\alpha)\Big{)}\left(\frac{(1+i)^{5}d}{ln}\right)F_{N(n)}\left(\frac{N(d)}{X}\right)
=\displaystyle= 1Xα1mod(1+i)3N(α)Yμ[i](α)(d,2)=1((1+i)5α2dln)FN(n)(N(α2d)X).\displaystyle\frac{1}{X}\sum_{\begin{subarray}{c}\alpha\equiv 1\bmod(1+i)^{3}\\ N(\alpha)\leq Y\end{subarray}}\mu_{[i]}(\alpha)\sum_{(d,2)=1}\left(\frac{(1+i)^{5}\alpha^{2}d}{ln}\right)F_{N(n)}\left(\frac{N(\alpha^{2}d)}{X}\right).

We apply the Poisson summation formula given in Lemma 2.10 to treat the last sum above to obtain that

(5.2) 𝒮M(((1+i)5dln);FN(n))=12N(ln)((1+i)6ln)α1mod(1+i)3N(α)Y(α,ln)=1μ[i](α)N(α)2k𝒪K(1)N(k)g(k,ln)F~N(n)(N(k)X2N(α2ln)).\displaystyle{\mathcal{S}}_{M}\biggl{(}\left(\frac{(1+i)^{5}d}{ln}\right);F_{N(n)}\biggr{)}=\frac{1}{2N(ln)}\left(\frac{(1+i)^{6}}{ln}\right)\sum_{\begin{subarray}{c}\alpha\equiv 1\bmod(1+i)^{3}\\ N(\alpha)\leq Y\\ (\alpha,ln)=1\end{subarray}}\frac{\mu_{[i]}(\alpha)}{N(\alpha)^{2}}\sum_{k\in\mathcal{O}_{K}}(-1)^{N(k)}g(k,ln){\widetilde{F}}_{N(n)}\Big{(}\sqrt{\frac{N(k)X}{2N(\alpha^{2}ln)}}\Big{)}.

The above allows us to deduce from (5.1) that

(5.3) 𝒮M(((1+i)5dl)Aδ,τ(d);Ψ)=𝒫(l)+0(l),\displaystyle{\mathcal{S}}_{M}\biggl{(}\left(\frac{(1+i)^{5}d}{l}\right)A_{\delta,\tau}(d);\Psi\biggr{)}={\mathcal{P}}(l)+{\mathcal{R}}_{0}(l),

where 𝒫(l){\mathcal{P}}(l) arises from the k=0k=0 term in (5.2) and 0(l){\mathcal{R}}_{0}(l) includes the remaining non-zero terms kk in (5.2). Hence

(5.4) 𝒫(l)=12N(l)n1mod(1+i)3rδ(n)N(n)32((1+i)6ln)α1mod(1+i)3N(α)Y(α,ln)=1μ[i](α)N(α)2g(0,ln)F~N(n)(0),0(l)=12N(l)n1mod(1+i)3rδ(n)N(n)32((1+i)6ln)α1mod(1+i)3N(α)Y(α,ln)=1μ[i](α)N(α)2k𝒪Kk0(1)N(k)g(k,ln)F~N(n)(N(k)X2N(α2ln)).\displaystyle\begin{split}{\mathcal{P}}(l)=&\frac{1}{2N(l)}\sum_{n\equiv 1\bmod(1+i)^{3}}\frac{r_{\delta}(n)}{N(n)^{\frac{3}{2}}}\left(\frac{(1+i)^{6}}{ln}\right)\sum_{\begin{subarray}{c}\alpha\equiv 1\bmod(1+i)^{3}\\ N(\alpha)\leq Y\\ (\alpha,ln)=1\end{subarray}}\frac{\mu_{[i]}(\alpha)}{N(\alpha)^{2}}g(0,ln){\widetilde{F}_{N(n)}}(0),\\ {\mathcal{R}}_{0}(l)=&\frac{1}{2N(l)}\sum_{n\equiv 1\bmod(1+i)^{3}}\frac{r_{\delta}(n)}{N(n)^{\frac{3}{2}}}\left(\frac{(1+i)^{6}}{ln}\right)\sum_{\begin{subarray}{c}\alpha\equiv 1\bmod(1+i)^{3}\\ N(\alpha)\leq Y\\ (\alpha,ln)=1\end{subarray}}\frac{\mu_{[i]}(\alpha)}{N(\alpha)^{2}}\sum_{\begin{subarray}{c}k\in\mathcal{O}_{K}\\ k\neq 0\end{subarray}}(-1)^{N(k)}g(k,ln){\widetilde{F}}_{N(n)}\Big{(}\sqrt{\frac{N(k)X}{2N(\alpha^{2}ln)}}\Big{)}.\end{split}

We show in what follows that 𝒫(l){\mathcal{P}}(l) contributes to a main term and 0(l){\mathcal{R}}_{0}(l) also contributes to secondary main terms.

5.2. The principal term 𝒫(l){\mathcal{P}}(l)

Note that g(0,ln)=φ[i](ln)g(0,ln)=\varphi_{[i]}(ln) if ln=ln=\square and g(0,ln)=0g(0,ln)=0 otherwise. Also, we have

α1mod(1+i)3N(α)Y(α,ln)=1μ[i](α)N(α)2=1ζK(2)ϖGϖ|2ln(11N(ϖ)2)1(1+O(1Y)).\sum_{\begin{subarray}{c}\alpha\equiv 1\bmod(1+i)^{3}\\ N(\alpha)\leq Y\\ (\alpha,ln)=1\end{subarray}}\frac{\mu_{[i]}(\alpha)}{N(\alpha)^{2}}=\frac{1}{\zeta_{K}(2)}\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|2ln\end{subarray}}\Big{(}1-\frac{1}{N(\varpi)^{2}}\Big{)}^{-1}\Big{(}1+O\Big{(}\frac{1}{Y}\Big{)}\Big{)}.

Using the above observations, we see that

𝒫(l)=1+O(Y1)ζK(2)n1mod(1+i)3ln=rδ(n)N(n)12((1+i)6ln)ϖGϖ|2ln(N(ϖ)N(ϖ)+1)F~N(n)(0).{\mathcal{P}}(l)=\frac{1+O(Y^{-1})}{\zeta_{K}(2)}\sum_{\begin{subarray}{c}n\equiv 1\bmod(1+i)^{3}\\ ln=\square\end{subarray}}\frac{r_{\delta}(n)}{N(n)^{\frac{1}{2}}}\left(\frac{(1+i)^{6}}{ln}\right)\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|2ln\end{subarray}}\left(\frac{N(\varpi)}{N(\varpi)+1}\right){\widetilde{F}}_{N(n)}(0).

As l=l1l22l=l_{1}l_{2}^{2} with l1l_{1} primary and square-free, we see that ln=ln=\square is equivalent to n=l1m2n=l_{1}m^{2} for some primary mm. Thus

𝒫(l)\displaystyle{\mathcal{P}}(l) =1+O(Y1)ζK(2)N(l1)m1mod(1+i)3rδ(l1m2)N(m)ϖGϖ|2lm(N(ϖ)N(ϖ)+1)F~N(l1m2)(0).\displaystyle=\frac{1+O(Y^{-1})}{\zeta_{K}(2)\sqrt{N(l_{1})}}\sum_{\begin{subarray}{c}m\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{r_{\delta}(l_{1}m^{2})}{N(m)}\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|2lm\end{subarray}}\left(\frac{N(\varpi)}{N(\varpi)+1}\right){\widetilde{F}}_{N(l_{1}m^{2})}(0).

Note that we have for any c>|(τ)|c>|\Re(\tau)|,

F~N(l1m2)(0)=Ψ(N(x+yi))Wδ,τ(π2N(l1m2)25XN(x+yi))dxdy=12πi(c)Γδ(s)(25/2π)2s(XN(l1m2))s(Ψ(N(x+yi))N(x+yi)sdxdy)2ss2τ2𝑑s=π2πi(c)Γδ(s)(25/2π)2s(XN(l1m2))sΨ^(1+s)2ss2τ2𝑑s,\displaystyle\begin{split}\widetilde{F}_{N(l_{1}m^{2})}\left(0\right)=&\int\limits^{\infty}_{-\infty}\int\limits^{\infty}_{-\infty}\Psi\left(N(x+yi)\right)W_{\delta,\tau}\left(\frac{\pi^{2}N(l_{1}m^{2})}{2^{5}XN(x+yi)}\right)\mathrm{d}x\mathrm{d}y\\ =&\frac{1}{2\pi i}\int\limits\limits_{(c)}\Gamma_{\delta}(s)\left(\frac{2^{5/2}}{\pi}\right)^{2s}\Big{(}\frac{X}{N(l_{1}m^{2})}\Big{)}^{s}\left(\int\limits^{\infty}_{-\infty}\int\limits^{\infty}_{-\infty}\Psi\left(N(x+yi)\right)N(x+yi)^{s}\mathrm{d}x\mathrm{d}y\right)\frac{2s}{s^{2}-\tau^{2}}ds\\ =&\frac{\pi}{2\pi i}\int\limits\limits\limits_{(c)}\Gamma_{\delta}(s)\left(\frac{2^{5/2}}{\pi}\right)^{2s}\Big{(}\frac{X}{N(l_{1}m^{2})}\Big{)}^{s}\widehat{\Psi}(1+s)\frac{2s}{s^{2}-\tau^{2}}ds,\end{split}

where we deduce the last equality above by noticing that

Ψ(N(x+yi))N(x+yi)sdxdy=02π0Ψ(r2)r2sr𝑑r𝑑θ=πΨ^(1+s).\displaystyle\int\limits^{\infty}_{-\infty}\int\limits^{\infty}_{-\infty}\Psi\left(N(x+yi)\right)N(x+yi)^{s}\mathrm{d}x\mathrm{d}y=\int^{2\pi}_{0}\int^{\infty}_{0}\Psi(r^{2})r^{2s}rdrd\theta=\pi\widehat{\Psi}(1+s).

We then conclude that for any c>κc>\kappa,

(5.5) 𝒫(l)=2π31+O(Y1)ζK(2)N(l1)I(l),\displaystyle{\mathcal{P}}(l)=\frac{2\pi}{3}\cdot\frac{1+O(Y^{-1})}{\zeta_{K}(2)\sqrt{N(l_{1})}}I(l),

where

(5.6) I(l)=12πi(c)Γδ(s)(25/2π)2s(XN(l1))sΨ^(1+s)2ss2τ2m1mod(1+i)3rδ(l1m2)N(m)1+2sϖGϖ|lm(N(ϖ)N(ϖ)+1)ds.\displaystyle I(l)=\frac{1}{2\pi i}\int\limits_{(c)}\Gamma_{\delta}(s)\left(\frac{2^{5/2}}{\pi}\right)^{2s}\Big{(}\frac{X}{N(l_{1})}\Big{)}^{s}\widehat{\Psi}(1+s)\frac{2s}{s^{2}-\tau^{2}}\sum_{\begin{subarray}{c}m\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{r_{\delta}(l_{1}m^{2})}{N(m)^{1+2s}}\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|lm\end{subarray}}\left(\frac{N(\varpi)}{N(\varpi)+1}\right)ds.

Now, by comparing Euler factors, we see that for (s)>1+2|(δ)|\Re(s)>1+2|\Re(\delta)|,

m1mod(1+i)3rδ(l1m2)N(m)sϖGϖ|lm(N(ϖ)N(ϖ)+1)=rδ(l1)Z(s;δ)ηδ(s;l),\sum_{\begin{subarray}{c}m\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{r_{\delta}(l_{1}m^{2})}{N(m)^{s}}\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|lm\end{subarray}}\left(\frac{N(\varpi)}{N(\varpi)+1}\right)=r_{\delta}(l_{1})Z(s;\delta)\eta_{\delta}(s;l),

where ZZ and η\eta are as defined in (3.6) and (2.31). Using this in (5.6), we see that

(5.7) I(l)\displaystyle I(l) =rδ(l1)2πi(c)Γδ(s)(25/2π)2s(XN(l1))sΨ^(1+s)2ss2τ2Z(1+2s;δ)ηδ(1+2s;l)𝑑s.\displaystyle=\frac{r_{\delta}(l_{1})}{2\pi i}\int\limits_{(c)}\Gamma_{\delta}(s)\left(\frac{2^{5/2}}{\pi}\right)^{2s}\Big{(}\frac{X}{N(l_{1})}\Big{)}^{s}\widehat{\Psi}(1+s)\frac{2s}{s^{2}-\tau^{2}}Z(1+2s;\delta)\eta_{\delta}(1+2s;l)ds.

Taking c=κ+ϵc=\kappa+\epsilon above implies that I(l)|rδ(l1)|(X/N(l1))κ+ϵI(l)\ll|r_{\delta}(l_{1})|(X/N(l_{1}))^{\kappa+\epsilon}. We now move the line of integration in (5.7) to s=14+ϵ\Re s=-\frac{1}{4}+\epsilon to encounter simple poles at s=±τs=\pm\tau, ±δ\pm\delta in the process. We note the following convexity bounds for ζK(s)\zeta_{K}(s) from [iwakow, Exercise 3, p. 100]:

(5.8) ζK(s)(1+|s|2)1(s)2+ε,0<(s)<1.\displaystyle\begin{split}\zeta_{K}(s)\ll(1+|s|^{2})^{\frac{1-\Re(s)}{2}+\varepsilon},\quad 0<\Re(s)<1.\end{split}

It follows that |Z(1+2s;δ)|(1+|s|2)3|Z(1+2s;\delta)|\ll(1+|s|^{2})^{3} on the new line. Moreover, we have

|ηδ(1+2s;l)|ϖGϖ|l1(1+O(1N(ϖ)))ϖGϖl1(1+O(1N(ϖ)1+ϵ))N(l1)ϵ.|\eta_{\delta}(1+2s;l)|\ll\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|l_{1}\end{subarray}}(1+O(\frac{1}{\sqrt{N(\varpi)}}))\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi\nmid l_{1}\end{subarray}}(1+O(\frac{1}{N(\varpi)^{1+\epsilon}}))\ll N(l_{1})^{\epsilon}.

Thus, we deduce that the integral on the new line is

|rδ(l1)|N(l1)14+ϵX14ϵ(14+ϵ)|s|5|Ψ^(1+s)|Γδ(s)||ds||rδ(l1)|N(l1)14+ϵX14ϵ.\ll\frac{|r_{\delta}(l_{1})|N(l_{1})^{\frac{1}{4}+\epsilon}}{X^{\frac{1}{4}-\epsilon}}\int\limits_{(-\frac{1}{4}+\epsilon)}|s|^{5}|{\widehat{\Psi}}(1+s)|\Gamma_{\delta}(s)||ds|\ll\frac{|r_{\delta}(l_{1})|N(l_{1})^{\frac{1}{4}+\epsilon}}{X^{\frac{1}{4}-\epsilon}}.

It follows that

I(l)\displaystyle I(l) =rδ(l1)Ress=±δ,±τ{Γδ(s)(25/2π)2s(XN(l1))sΨ^(1+s)2ss2τ2Z(1+2s;δ)ηδ(1+2s;l)}+O(|rδ(l1)|N(l1)14+ϵX14ϵ).\displaystyle=r_{\delta}(l_{1})\mathop{\text{Res}}_{s=\pm\delta,\pm\tau}\biggl{\{}\Gamma_{\delta}(s)\left(\frac{2^{5/2}}{\pi}\right)^{2s}\Big{(}\frac{X}{N(l_{1})}\Big{)}^{s}\widehat{\Psi}(1+s)\frac{2s}{s^{2}-\tau^{2}}Z(1+2s;\delta)\eta_{\delta}(1+2s;l)\biggr{\}}+O\biggl{(}\frac{|r_{\delta}(l_{1})|N(l_{1})^{\frac{1}{4}+\epsilon}}{X^{\frac{1}{4}-\epsilon}}\biggr{)}.

Applying this in (5.5), we obtain that

(5.9) 𝒫(l)=2πrδ(l1)3ζK(2)N(l1)Ress=±δs=±τ{Γδ(s)(25/2π)2s(XN(l1))sΨ^(1+s)2ss2τ2Z(1+2s;δ)ηδ(1+2s;l)}+O(|rδ(l1)|Xκ+ϵYN(l1)12+κ+|rδ(l1)|Xϵ(XN(l1))14).\displaystyle\begin{split}{\mathcal{P}}(l)=&\frac{2\pi r_{\delta}(l_{1})}{3\zeta_{K}(2)\sqrt{N(l_{1})}}\mathop{\text{Res}}_{\begin{subarray}{c}s=\pm\delta\\ s=\pm\tau\end{subarray}}\biggl{\{}\Gamma_{\delta}(s)\left(\frac{2^{5/2}}{\pi}\right)^{2s}\Big{(}\frac{X}{N(l_{1})}\Big{)}^{s}\widehat{\Psi}(1+s)\frac{2s}{s^{2}-\tau^{2}}Z(1+2s;\delta)\eta_{\delta}(1+2s;l)\biggr{\}}\\ &+O\biggl{(}\frac{|r_{\delta}(l_{1})|X^{\kappa+\epsilon}}{YN(l_{1})^{\frac{1}{2}+\kappa}}+\frac{|r_{\delta}(l_{1})|X^{\epsilon}}{(XN(l_{1}))^{\frac{1}{4}}}\biggr{)}.\end{split}

5.3. The secondary main terms

We apply the Mellin transform to recast 0(l){\mathcal{R}}_{0}(l) given in (5.4) as

(5.10) 0(l)=\displaystyle{\mathcal{R}}_{0}(l)= 12N(l)α1mod(1+i)3N(α)Y(α,l)=1μ[i](α)N(α)2k𝒪Kk0(1)N(k)2πi(c)n1mod(1+i)3(n,α)=1rδ(n)N(n)32+wg(k,ln)h(N(k)X2N(α2l),w)dw,\displaystyle\frac{1}{2N(l)}\sum_{\begin{subarray}{c}\alpha\equiv 1\bmod(1+i)^{3}\\ N(\alpha)\leq Y\\ (\alpha,l)=1\end{subarray}}\frac{\mu_{[i]}(\alpha)}{N(\alpha)^{2}}\sum_{\begin{subarray}{c}k\in\mathcal{O}_{K}\\ k\neq 0\end{subarray}}\frac{(-1)^{N(k)}}{2\pi i}\int\limits_{(c)}\sum_{\begin{subarray}{c}n\equiv 1\bmod(1+i)^{3}\\ (n,\alpha)=1\end{subarray}}\frac{r_{\delta}(n)}{N(n)^{\frac{3}{2}+w}}g(k,ln)h\Big{(}\frac{N(k)X}{2N(\alpha^{2}l)},w\Big{)}dw,

where c>|(δ)|c>|\Re(\delta)| and hh is defined as in (2.12).

We apply Lemma 2.13 in (5.10) and move the line of integration to (w)=12+|(δ)|+ϵ\Re(w)=-\frac{1}{2}+|\Re(\delta)|+\epsilon. In this process, we encounter poles only when k1=±ik_{1}=\pm i, with simple poles at w=±δw=\pm\delta. The residues of these poles contribute secondary main terms. We therefore write 0(l)=(l)+𝒫+(l)+𝒫(l){\mathcal{R}}_{0}(l)={\mathcal{R}}(l)+{\mathcal{P}}_{+}(l)+{\mathcal{P}}_{-}(l) where

(5.11) (l)=12N(l)α1mod(1+i)3N(α)Y(α,l)=1μ[i](α)N(α)2k𝒪Kk0(1)N(k)2πi(12+|(δ)|+ϵ)L(1+w+δ,χik1)L(1+wδ,χik1)×𝒢δ(1+w;k,l,α)h(N(k)X2N(α2l),w)dw,\displaystyle\begin{split}{\mathcal{R}}(l)=&\frac{1}{2N(l)}\sum_{\begin{subarray}{c}\alpha\equiv 1\bmod(1+i)^{3}\\ N(\alpha)\leq Y\\ (\alpha,l)=1\end{subarray}}\frac{\mu_{[i]}(\alpha)}{N(\alpha)^{2}}\sum_{\begin{subarray}{c}k\in\mathcal{O}_{K}\\ k\neq 0\end{subarray}}\frac{(-1)^{N(k)}}{2\pi i}\int\limits_{(-\frac{1}{2}+|\Re(\delta)|+\epsilon)}L(1+w+\delta,\chi_{ik_{1}})L(1+w-\delta,\chi_{ik_{1}})\\ &\times{\mathcal{G}}_{\delta}(1+w;k,l,\alpha)h\Big{(}\frac{N(k)X}{2N(\alpha^{2}l)},w\Big{)}dw,\end{split}

and (after replacing kk by ±ik2\pm ik^{2})

(5.12) 𝒫+(l)=π412N(l)α1mod(1+i)3N(α)Y(α,l)=1μ[i](α)N(α)2μ=±ζK(1+2μδ)kG(1)N(k)𝒢δ(1+μδ;ik2,l,α)h(N(k)2X2N(α2l),μδ),𝒫(l)=π412N(l)α1mod(1+i)3N(α)Y(α,l)=1μ[i](α)N(α)2μ=±ζK(1+2μδ)kG(1)N(k)𝒢δ(1+μδ;ik2,l,α)h(N(k)2X2N(α2l),μδ).\displaystyle\begin{split}{\mathcal{P}}_{+}(l)=&\frac{\pi}{4}\cdot\frac{1}{2N(l)}\sum_{\begin{subarray}{c}\alpha\equiv 1\bmod(1+i)^{3}\\ N(\alpha)\leq Y\\ (\alpha,l)=1\end{subarray}}\frac{\mu_{[i]}(\alpha)}{N(\alpha)^{2}}\sum_{\mu=\pm}\zeta_{K}(1+2\mu\delta)\sum_{k\in G}(-1)^{N(k)}{\mathcal{G}}_{\delta}(1+\mu\delta;ik^{2},l,\alpha)h\Big{(}\frac{N(k)^{2}X}{2N(\alpha^{2}l)},\mu\delta\Big{)},\\ {\mathcal{P}}_{-}(l)=&\frac{\pi}{4}\cdot\frac{1}{2N(l)}\sum_{\begin{subarray}{c}\alpha\equiv 1\bmod(1+i)^{3}\\ N(\alpha)\leq Y\\ (\alpha,l)=1\end{subarray}}\frac{\mu_{[i]}(\alpha)}{N(\alpha)^{2}}\sum_{\mu=\pm}\zeta_{K}(1+2\mu\delta)\sum_{k\in G}(-1)^{N(k)}{\mathcal{G}}_{\delta}(1+\mu\delta;-ik^{2},l,\alpha)h\Big{(}\frac{N(k)^{2}X}{2N(\alpha^{2}l)},\mu\delta\Big{)}.\end{split}

Note that it follows from Lemma 2.2 and Lemma 2.13 that we have 𝒢δ(1+μδ;ik2,l,α)=𝒢δ(1+μδ;ik2,l,α){\mathcal{G}}_{\delta}(1+\mu\delta;ik^{2},l,\alpha)={\mathcal{G}}_{\delta}(1+\mu\delta;-ik^{2},l,\alpha). Note also that we have

(5.13) jG(1)N(j)N(j)2u𝒢v(s;ij2,,α)=jG(j,1+i)=1N(j)2u𝒢v(s;ij2,,α)+jG1+i|jN(j)2u𝒢v(s;ij2,,α)=jGN(j)2u𝒢v(s;ij2,,α)+2jG1+i|jN(j)2u𝒢v(s;ij2,,α)=jGN(j)2u𝒢v(s;ij2,,α)+212ujGN(j)2u𝒢v(s;i(1+i)2j2,,α)=(1212u)jGN(j)2u𝒢v(s;ij2,,α),\displaystyle\begin{split}&\sum_{\begin{subarray}{c}j\in G\end{subarray}}(-1)^{N(j)}N(j)^{-2u}\mathcal{G}_{v}(s;ij^{2},\ell,\alpha)\\ =&-\sum_{\begin{subarray}{c}j\in G\\ (j,1+i)=1\end{subarray}}N(j)^{-2u}\mathcal{G}_{v}(s;ij^{2},\ell,\alpha)+\sum_{\begin{subarray}{c}j\in G\\ 1+i|j\end{subarray}}N(j)^{-2u}\mathcal{G}_{v}(s;ij^{2},\ell,\alpha)\\ =&-\sum_{\begin{subarray}{c}j\in G\end{subarray}}N(j)^{-2u}\mathcal{G}_{v}(s;ij^{2},\ell,\alpha)+2\sum_{\begin{subarray}{c}j\in G\\ 1+i|j\end{subarray}}N(j)^{-2u}\mathcal{G}_{v}(s;ij^{2},\ell,\alpha)\\ =&-\sum_{\begin{subarray}{c}j\in G\end{subarray}}N(j)^{-2u}\mathcal{G}_{v}(s;ij^{2},\ell,\alpha)+2^{1-2u}\sum_{\begin{subarray}{c}j\in G\end{subarray}}N(j)^{-2u}\mathcal{G}_{v}(s;i(1+i)^{2}j^{2},\ell,\alpha)\\ =&-(1-2^{1-2u})\sum_{\begin{subarray}{c}j\in G\end{subarray}}N(j)^{-2u}\mathcal{G}_{v}(s;ij^{2},\ell,\alpha),\end{split}

where the last equality above follows from the observation that we have 𝒢v(s;i(1+i)2j2,,α)=𝒢v(s;ij2,,α)\mathcal{G}_{v}(s;i(1+i)^{2}j^{2},\ell,\alpha)=\mathcal{G}_{v}(s;ij^{2},\ell,\alpha) by Lemma 2.2 and Lemma 2.13.

We now define for μ=±\mu=\pm and any complex number uu with (u)>12\Re(u)>\frac{1}{2},

(u,v;l,α)=\displaystyle{\mathcal{H}}(u,v;l,\alpha)= N(l)ukG(1)N(k)N(k)2u𝒢v(1+v;ik2,l,α).\displaystyle N(l)^{u}\sum_{k\in G}\frac{(-1)^{N(k)}}{N(k)^{2u}}{\mathcal{G}}_{v}(1+v;ik^{2},l,\alpha).

It follows from Lemma 2.13 that the above series converges absolutely when (u)>12\Re(u)>\frac{1}{2}. We further apply (5.13) to recast (u,v;l,α){\mathcal{H}}(u,v;l,\alpha) as

(u,v;l,α)=N(l)u(1212u)ϖGb=0𝒢v;ϖ(1+v;iϖ2b,l,α)N(ϖ)2bu.\displaystyle{\mathcal{H}}(u,v;l,\alpha)=-N(l)^{u}(1-2^{1-2u})\prod_{\varpi\in G}\sum_{b=0}^{\infty}\frac{{\mathcal{G}}_{v;\varpi}(1+v;i\varpi^{2b},l,\alpha)}{N(\varpi)^{2bu}}.

We now apply Lemma 2.2 to evaluate 𝒢v;ϖ{\mathcal{G}}_{v;\varpi} defined in Lemma 2.13 to see that

(5.14) (u,v;l,α)=N(l)(1212u)N(l1)u12ζK(2u)ζK(2u+1+4v)1(u,v;l,α),\displaystyle{\mathcal{H}}(u,v;l,\alpha)=-N(l)(1-2^{1-2u})N(l_{1})^{u-\frac{1}{2}}\zeta_{K}(2u)\zeta_{K}(2u+1+4v){\mathcal{H}}_{1}(u,v;l,\alpha),

where 1=ϖG1;ϖ{\mathcal{H}}_{1}=\prod_{\varpi\in G}{\mathcal{H}}_{1;\varpi} with

1;ϖ={(11N(ϖ))(11N(ϖ)1+2v)(11N(ϖ)2u+1+4v)if ϖ|2α,(11N(ϖ))(11N(ϖ)1+2v)(1+1N(ϖ)+1N(ϖ)1+2v1N(ϖ)2u+2+4v)if ϖ2αl,(11N(ϖ))(11N(ϖ)1+2v)(1+1N(ϖ)2u+2v)if ϖ|l1,(11N(ϖ))(11N(ϖ)1+2v)(1+1N(ϖ)1+2v)if ϖ|l,ϖl1.{\mathcal{H}}_{1;\varpi}=\begin{cases}\Big{(}1-\frac{1}{N(\varpi)}\Big{)}\Big{(}1-\frac{1}{N(\varpi)^{1+2v}}\Big{)}\Big{(}1-\frac{1}{N(\varpi)^{2u+1+4v}}\Big{)}&\text{if }\varpi|2\alpha,\\ \Big{(}1-\frac{1}{N(\varpi)}\Big{)}\Big{(}1-\frac{1}{N(\varpi)^{1+2v}}\Big{)}\Big{(}1+\frac{1}{N(\varpi)}+\frac{1}{N(\varpi)^{1+2v}}-\frac{1}{N(\varpi)^{2u+2+4v}}\Big{)}&\text{if }\varpi\nmid 2\alpha l,\\ \Big{(}1-\frac{1}{N(\varpi)}\Big{)}\Big{(}1-\frac{1}{N(\varpi)^{1+2v}}\Big{)}\Big{(}1+\frac{1}{N(\varpi)^{2u+2v}}\Big{)}&\text{if }\varpi|l_{1},\\ \Big{(}1-\frac{1}{N(\varpi)}\Big{)}\Big{(}1-\frac{1}{N(\varpi)^{1+2v}}\Big{)}\Big{(}1+\frac{1}{N(\varpi)^{1+2v}}\Big{)}&\text{if }\varpi|l,\varpi\nmid l_{1}.\end{cases}

It then follows from this and (5.14) that (u,v;l,α){\mathcal{H}}(u,v;l,\alpha) is analytic in the domain (u)>12+|(v)|\Re(u)>-\frac{1}{2}+|\Re(v)|.

We now apply Lemma 2.11 and the observation that 𝒢v(1+v;ik2,l,α){\mathcal{G}}_{v}(1+v;ik^{2},l,\alpha) is an even function of vv to deduce that, for c>max(12+|(δ)|,|(τ)|)=12+|(δ)|c>\max(\frac{1}{2}+|\Re(\delta)|,|\Re(\tau)|)=\frac{1}{2}+|\Re(\delta)|,

(5.15) kG(1)N(k)𝒢δ(1+μδ;±ik2,l,α)h(N(k)2X2N(α2l),μδ)=π2πi(c)Γδ(s)(26N(α)2π2)sΨ^(1+μδ)(X2N(α)2)μδΓ(sμδ)Γ(1s+μδ)π2s+2μδ2ss2τ2(sμδ,μδ;l,α)𝑑s.\displaystyle\begin{split}&\sum_{k\in G}(-1)^{N(k)}{\mathcal{G}}_{\delta}(1+\mu\delta;\pm ik^{2},l,\alpha)h\Big{(}\frac{N(k)^{2}X}{2N(\alpha^{2}l)},\mu\delta\Big{)}\\ =&\frac{\pi}{2\pi i}\int\limits_{(c)}\Gamma_{\delta}(s)\left(\frac{2^{6}N(\alpha)^{2}}{\pi^{2}}\right)^{s}{\widehat{\Psi}}(1+\mu\delta)\left(\frac{X}{2N(\alpha)^{2}}\right)^{\mu\delta}\frac{\Gamma(s-\mu\delta)}{\Gamma(1-s+\mu\delta)}\pi^{-2s+2\mu\delta}\frac{2s}{s^{2}-\tau^{2}}{\mathcal{H}}(s-\mu\delta,\mu\delta;l,\alpha)ds.\end{split}

The analytical properties of HH discussed above allow us to move the line of integration in (LABEL:Hint) to (s)=κ+1logX\Re(s)=\kappa+\frac{1}{\log X} without encountering any pole. We then deduce from this and (5.12) that

(5.16) 𝒫±(l)=π4π2πi(κ+1logX)Γδ(s)(26π2)s2ss2τ2μ=±Ψ^(1+μδ)(X2)μδζK(1+2μδ)Γ(sμδ)Γ(1s+μδ)π2s+2μδ×12N(l)α1mod(1+i)3N(α)Y(α,l)=1μ[i](α)N(α)22s+2μδ(sμδ,μδ;l,α)ds.\displaystyle\begin{split}{\mathcal{P}}_{\pm}(l)=&\frac{\pi}{4}\cdot\frac{\pi}{2\pi i}\int\limits_{(\kappa+\frac{1}{\log X})}\Gamma_{\delta}(s)\left(\frac{2^{6}}{\pi^{2}}\right)^{s}\frac{2s}{s^{2}-\tau^{2}}\sum_{\mu=\pm}{\widehat{\Psi}}(1+\mu\delta)\left(\frac{X}{2}\right)^{\mu\delta}\zeta_{K}(1+2\mu\delta)\frac{\Gamma(s-\mu\delta)}{\Gamma(1-s+\mu\delta)}\pi^{-2s+2\mu\delta}\\ &\times\frac{1}{2N(l)}\sum_{\begin{subarray}{c}\alpha\equiv 1\bmod(1+i)^{3}\\ N(\alpha)\leq Y\\ (\alpha,l)=1\end{subarray}}\frac{\mu_{[i]}(\alpha)}{N(\alpha)^{2-2s+2\mu\delta}}{\mathcal{H}}(s-\mu\delta,\mu\delta;l,\alpha)ds.\end{split}

We now extend the sum over α\alpha in (5.16) to infinity. To estimate the error introduced, we let 𝒞{\mathcal{C}} be the circle centred at 0, with radius κ+1/(2logX)\kappa+1/(2\log X), and oriented counter-clockwise. Notice that, for any complex number ss with (s)=κ+1logX\Re(s)=\kappa+\frac{1}{\log X}, the function 2zΨ^(1+z)(X2N(α)2)zζK(1+2z)Γ(sμδ)Γ(1s+μδ)(π)2s+2μδ(sz,z;l,α)2z{\widehat{\Psi}}(1+z)\left(\frac{X}{2N(\alpha)^{2}}\right)^{z}\zeta_{K}(1+2z)\frac{\Gamma(s-\mu\delta)}{\Gamma(1-s+\mu\delta)}(\pi)^{-2s+2\mu\delta}{\mathcal{H}}(s-z,z;l,\alpha) is analytic for zz inside 𝒞{\mathcal{C}}. Thus, we deduce from Cauchy’s theorem that

(5.17) μ=±Ψ^(1+μδ)(X2N(α)2)μδζK(1+2μδ)Γ(sμδ)Γ(1s+μδ)π2s+2μδ(sμδ,μδ;l,α)=12πi𝒞Ψ^(1+z)(X2N(α)2)zζK(1+2z)Γ(sz)Γ(1s+z)π2s+2z(sz,z;l,α)2zz2δ2𝑑z.\displaystyle\begin{split}&\sum_{\mu=\pm}{\widehat{\Psi}}(1+\mu\delta)\left(\frac{X}{2N(\alpha)^{2}}\right)^{\mu\delta}\zeta_{K}(1+2\mu\delta)\frac{\Gamma(s-\mu\delta)}{\Gamma(1-s+\mu\delta)}\pi^{-2s+2\mu\delta}{\mathcal{H}}(s-\mu\delta,\mu\delta;l,\alpha)\\ =&\frac{1}{2\pi i}\int_{{\mathcal{C}}}{\widehat{\Psi}}(1+z)\left(\frac{X}{2N(\alpha)^{2}}\right)^{z}\zeta_{K}(1+2z)\frac{\Gamma(s-z)}{\Gamma(1-s+z)}\pi^{-2s+2z}{\mathcal{H}}(s-z,z;l,\alpha)\frac{2z}{z^{2}-\delta^{2}}dz.\end{split}

We observe that 2κ+3/(2logX)(sz)1/(2logX)2\kappa+3/(2\log X)\geq\Re(s-z)\geq 1/(2\log X) for zz on 𝒞{\mathcal{C}} and that |2zz2δ2Ψ^(1+z)ζK(1+2z)(X2N(α)2)z|(log2X)(XN(α)2)κ+1/logX|\frac{2z}{z^{2}-\delta^{2}}{\widehat{\Psi}}(1+z)\zeta_{K}(1+2z)\left(\frac{X}{2N(\alpha)^{2}}\right)^{z}|\ll(\log^{2}X)(XN(\alpha)^{2})^{\kappa+1/\log X}. Also, by Stirling’s formula (see [iwakow, (5.112)]), we have

|Γ(sz)Γ(1s+z)π2s+2z|(1+|(s)|)4κ+3/logX11.\Big{|}\frac{\Gamma(s-z)}{\Gamma(1-s+z)}\pi^{-2s+2z}\Big{|}\ll(1+|\Im(s)|)^{4\kappa+3/\log X-1}\ll 1.

Moreover, applying the convexity bounds for ζK(s)\zeta_{K}(s) given in (5.8), we deduce from (5.14) that |(sz,z;l,α)|N(l)1+ϵN(l1)2κ+3/(2logX)12N(α)ϵ(1+|s|)logX|{\mathcal{H}}(s-z,z;l,\alpha)|\ll N(l)^{1+\epsilon}N(l_{1})^{2\kappa+3/(2\log X)-\frac{1}{2}}N(\alpha)^{\epsilon}(1+|s|)\log X. These estimates allow us to bound the quantities in (5.17) by

N(l)1+ϵN(l1)2κ12(XN(α)2)κ+ϵ(1+|s|).\ll N(l)^{1+\epsilon}N(l_{1})^{2\kappa-\frac{1}{2}}(XN(\alpha)^{2})^{\kappa+\epsilon}(1+|s|).

It follows that

12N(l)α1mod(1+i)3N(α)>Y(α,l)=1μ[i](α)N(α)22sμ=±Ψ^(1+μδ)(X2N(α)2)μδζK(1+2μδ)Γ(sμδ)Γ(1s+μδ)π2s+2z(sμδ,μδ;l,α)\displaystyle\frac{1}{2N(l)}\sum_{\begin{subarray}{c}\alpha\equiv 1\bmod(1+i)^{3}\\ N(\alpha)>Y\\ (\alpha,l)=1\end{subarray}}\frac{\mu_{[i]}(\alpha)}{N(\alpha)^{2-2s}}\sum_{\mu=\pm}{\widehat{\Psi}}(1+\mu\delta)\left(\frac{X}{2N(\alpha)^{2}}\right)^{\mu\delta}\zeta_{K}(1+2\mu\delta)\frac{\Gamma(s-\mu\delta)}{\Gamma(1-s+\mu\delta)}\pi^{-2s+2z}{\mathcal{H}}(s-\mu\delta,\mu\delta;l,\alpha)
\displaystyle\ll N(l)ϵN(l1)2κ12Xκ+ϵ(1+|s|)Y1+4κ.\displaystyle N(l)^{\epsilon}N(l_{1})^{2\kappa-\frac{1}{2}}X^{\kappa+\epsilon}(1+|s|)Y^{-1+4\kappa}.

Applying the above in (5.16), we see that the error introduced by extending the sum over α\alpha to infinity is

N(l)ϵN(l1)2κ12Xκ+ϵY1+4κ(κ+1logX)|Γδ(s)|(1+|s|)|s||s2τ2||ds|N(l)ϵN(l1)2κ12Xκ+ϵY1+4κ.\displaystyle\ll N(l)^{\epsilon}N(l_{1})^{2\kappa-\frac{1}{2}}X^{\kappa+\epsilon}Y^{-1+4\kappa}\int\limits_{(\kappa+\frac{1}{\log X})}|\Gamma_{\delta}(s)|(1+|s|)\frac{|s|}{|s^{2}-\tau^{2}|}|ds|\ll N(l)^{\epsilon}N(l_{1})^{2\kappa-\frac{1}{2}}X^{\kappa+\epsilon}Y^{-1+4\kappa}.

We then conclude that we have

(5.18) 𝒫±(l)=π24μ=±Ψ^(1+μδ)(X2)μδζK(1+2μδ)12πi(κ+1logX)𝒥(s,μδ;l)2ss2τ2𝑑s+O(N(l)ϵN(l1)2κ12Xκ+ϵY1+4κ),\displaystyle{\mathcal{P}}_{\pm}(l)=\frac{\pi^{2}}{4}\cdot\sum_{\mu=\pm}{\widehat{\Psi}}(1+\mu\delta)\left(\frac{X}{2}\right)^{\mu\delta}\zeta_{K}(1+2\mu\delta)\frac{1}{2\pi i}\int\limits_{(\kappa+\frac{1}{\log X})}{\mathcal{J}}(s,\mu\delta;l)\frac{2s}{s^{2}-\tau^{2}}ds+O(N(l)^{\epsilon}N(l_{1})^{2\kappa-\frac{1}{2}}X^{\kappa+\epsilon}Y^{-1+4\kappa}),

where

𝒥(s,v;l):=\displaystyle{\mathcal{J}}(s,v;l):= Γv(s)(26π2)sΓ(sv)Γ(1s+v)π2s+2z𝒦(s,v;l),\displaystyle\Gamma_{v}(s)\left(\frac{2^{6}}{\pi^{2}}\right)^{s}\frac{\Gamma(s-v)}{\Gamma(1-s+v)}\pi^{-2s+2z}{\mathcal{K}}(s,v;l),
𝒦(s,v;l):=\displaystyle{\mathcal{K}}(s,v;l):= 12N(l)α1mod(1+i)3(α,l)=1μ[i](α)N(α)22s+2μδ(sv,v;l,α).\displaystyle\frac{1}{2N(l)}\sum_{\begin{subarray}{c}\alpha\equiv 1\bmod(1+i)^{3}\\ (\alpha,l)=1\end{subarray}}\frac{\mu_{[i]}(\alpha)}{N(\alpha)^{2-2s+2\mu\delta}}{\mathcal{H}}(s-v,v;l,\alpha).

A direct calculation using the expression for {\mathcal{H}} given in (5.14) yields

𝒦(s,v;l)=\displaystyle{\mathcal{K}}(s,v;l)= 14N(l1)12+vφ[i](l)N(l)ϖGϖ|2l(11N(ϖ)1+2v)ϖGϖ|lϖl1(1+1N(ϖ)1+2v)rs(l1)\displaystyle-\frac{1}{4N(l_{1})^{\frac{1}{2}+v}}\frac{\varphi_{[i]}(l)}{N(l)}\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|2l\end{subarray}}\biggl{(}1-\frac{1}{N(\varpi)^{1+2v}}\biggr{)}\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|l\\ \varpi\nmid l_{1}\end{subarray}}\biggl{(}1+\frac{1}{N(\varpi)^{1+2v}}\biggr{)}r_{s}(l_{1})
×(4s+4s212v21+2v4s)ζK(2s2v)ζK(2s+1+2v)\displaystyle\times\biggl{(}\frac{4^{s}+4^{-s}-2^{-1-2v}-2^{1+2v}}{4^{s}}\biggr{)}\zeta_{K}(2s-2v)\zeta_{K}(2s+1+2v)
×ϖGϖ2l(11N(ϖ))(11N(ϖ)1+2v)(1+1N(ϖ)+1N(ϖ)1+2v+1N(ϖ)3+4v(N(ϖ)2s+N(ϖ)2s)N(ϖ)2+2v).\displaystyle\times\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi\nmid 2l\end{subarray}}\biggl{(}1-\frac{1}{N(\varpi)}\biggr{)}\biggl{(}1-\frac{1}{N(\varpi)^{1+2v}}\biggr{)}\biggl{(}1+\frac{1}{N(\varpi)}+\frac{1}{N(\varpi)^{1+2v}}+\frac{1}{N(\varpi)^{3+4v}}-\frac{(N(\varpi)^{2s}+N(\varpi)^{-2s})}{N(\varpi)^{2+2v}}\biggr{)}.

Using this together with the functional equation for ζK(s)\zeta_{K}(s) given in (2.3), and the identity Γ(z)Γ(z+12)=π12212zΓ(2z)\Gamma(z)\Gamma(z+\frac{1}{2})=\pi^{\frac{1}{2}}2^{1-2z}\Gamma(2z), we obtain the following identity

(5.19) ζK(1+2v)𝒥(s,v;l)=\displaystyle\zeta_{K}(1+2v){\mathcal{J}}(s,v;l)= 1π4rs(l1)3ζK(2)N(l1)(26π2N(l1))vΓs(v)Z(1+2v;s)ηs(1+2v;l).\displaystyle\frac{1}{\pi}\cdot\frac{4r_{s}(l_{1})}{3\zeta_{K}(2)\sqrt{N(l_{1})}}\left(\frac{2^{6}}{\pi^{2}N(l_{1})}\right)^{v}\Gamma_{s}(v)Z(1+2v;s)\eta_{s}(1+2v;l).

As it is easy to see that the left side above is invariant under sss\to-s, we deduce that 𝒥(s,v;l)=𝒥(s,v;l){\mathcal{J}}(s,v;l)={\mathcal{J}}(-s,v;l).

We now evaluate the integral in (5.18) for μ=±\mu=\pm by moving the line of integration to (s)=κ1logX\Re(s)=-\kappa-\frac{1}{\log X} to encounter simple poles at s=±δ,±τs=\pm\delta,\pm\tau in the process. It follows that

12πi(κ+1logX)𝒥(s,μδ;l)2ss2τ2𝑑s=\displaystyle\frac{1}{2\pi i}\int\limits_{(\kappa+\frac{1}{\log X})}{\mathcal{J}}(s,\mu\delta;l)\frac{2s}{s^{2}-\tau^{2}}ds= Ress=±δ,±τ𝒥(s,μδ;l)2ss2τ2+12πi(κ1logX)𝒥(s,μδ;l)2ss2τ2𝑑s\displaystyle\mathop{\text{Res}}_{s=\pm\delta,\pm\tau}{\mathcal{J}}(s,\mu\delta;l)\frac{2s}{s^{2}-\tau^{2}}+\frac{1}{2\pi i}\int\limits_{(-\kappa-\frac{1}{\log X})}{\mathcal{J}}(s,\mu\delta;l)\frac{2s}{s^{2}-\tau^{2}}ds
=\displaystyle= Ress=±δ,±τ𝒥(s,μδ;l)2ss2τ212πi(κ+1logX)𝒥(s,μδ;l)2ss2τ2𝑑s,\displaystyle\mathop{\text{Res}}_{s=\pm\delta,\pm\tau}{\mathcal{J}}(s,\mu\delta;l)\frac{2s}{s^{2}-\tau^{2}}-\frac{1}{2\pi i}\int\limits_{(\kappa+\frac{1}{\log X})}{\mathcal{J}}(s,\mu\delta;l)\frac{2s}{s^{2}-\tau^{2}}ds,

where the last integral above follows from the previous one upon a change of variable sss\rightarrow-s while noticing that 𝒥(s,μδ;l)=𝒥(s,μδ;l){\mathcal{J}}(s,\mu\delta;l)={\mathcal{J}}(-s,\mu\delta;l). Thus we deduce that

12πi(κ+1logX)𝒥(s,μδ;l)2ss2τ2𝑑s=12(Ress=±δ,±τ𝒥(s,μδ;l)2ss2τ2)=Ress=μδ𝒥(s,μδ;l)2ss2τ2+𝒥(τ,μδ;l).\frac{1}{2\pi i}\int\limits_{(\kappa+\frac{1}{\log X})}{\mathcal{J}}(s,\mu\delta;l)\frac{2s}{s^{2}-\tau^{2}}ds=\frac{1}{2}\biggl{(}\mathop{\text{Res}}_{s=\pm\delta,\pm\tau}{\mathcal{J}}(s,\mu\delta;l)\frac{2s}{s^{2}-\tau^{2}}\biggr{)}=\mathop{\text{Res}}_{s=\mu\delta}{\mathcal{J}}(s,\mu\delta;l)\frac{2s}{s^{2}-\tau^{2}}+{\mathcal{J}}(\tau,\mu\delta;l).

Applying the above in (5.18), we obtain that

(5.20) 𝒫+(l)=𝒫(l)=π24μ=±Ψ^(1+μδ)(X2)μδζK(1+2μδ)(Ress=μδ𝒥(s,μδ;l)2ss2τ2+𝒥(τ,μδ;l))+O(N(l)ϵXκ+ϵN(l1)2κ12Y14κ).\displaystyle\begin{split}{\mathcal{P}}_{+}(l)={\mathcal{P}}_{-}(l)=&\frac{\pi^{2}}{4}\cdot\sum_{\mu=\pm}{\widehat{\Psi}}(1+\mu\delta)\left(\frac{X}{2}\right)^{\mu\delta}\zeta_{K}(1+2\mu\delta)\left(\mathop{\text{Res}}_{s=\mu\delta}{\mathcal{J}}(s,\mu\delta;l)\frac{2s}{s^{2}-\tau^{2}}+{\mathcal{J}}(\tau,\mu\delta;l)\right)\\ &+O\Big{(}\frac{N(l)^{\epsilon}X^{\kappa+\epsilon}N(l_{1})^{2\kappa-\frac{1}{2}}}{Y^{1-4\kappa}}\Big{)}.\end{split}

5.4. Estimation of (l){\mathcal{R}}(l)

In this section, we estimate (l){\mathcal{R}}(l) given in (5.11) on average by deriving the bound given in (3.8). We denote βl=(l)¯|(l)|\beta_{l}=\frac{\overline{{\mathcal{R}}(l)}}{|{\mathcal{R}}(l)|} if (l)0{\mathcal{R}}(l)\neq 0, and βl=1\beta_{l}=1 otherwise. We deduce from (5.11) that

LN(l)2L1|(l)|=LN(l)2L1βl(l)\displaystyle\sum_{L\leq N(l)\leq 2L-1}|{\mathcal{R}}(l)|=\sum_{L\leq N(l)\leq 2L-1}\beta_{l}{\mathcal{R}}(l)\ll α1mod(1+i)3N(α)Y1N(α)2(12+|(δ)|+ϵ)k𝒪Kk0Uδ(α,k,w)|dw|,\displaystyle\sum_{\begin{subarray}{c}\alpha\equiv 1\bmod(1+i)^{3}\\ N(\alpha)\leq Y\end{subarray}}\frac{1}{N(\alpha)^{2}}\int\limits_{(-\frac{1}{2}+|\Re(\delta)|+\epsilon)}\sum_{\begin{subarray}{c}k\in\mathcal{O}_{K}\\ k\neq 0\end{subarray}}U_{\delta}(\alpha,k,w)|dw|,

where

Uδ(α,k,w)=|L(1+w+δ,χik1)L(1+wδ,χik1)||LN(l)2L1(l,2α)=1βlN(l)𝒢δ(1+w;k,l,α)h(N(k)X2N(α2l),w)|.U_{\delta}(\alpha,k,w)=|L(1+w+\delta,\chi_{ik_{1}})L(1+w-\delta,\chi_{ik_{1}})|\biggl{|}\sum_{\begin{subarray}{c}L\leq N(l)\leq 2L-1\\ (l,2\alpha)=1\end{subarray}}\frac{\beta_{l}}{N(l)}{\mathcal{G}}_{\delta}(1+w;k,l,\alpha)h\biggl{(}\frac{N(k)X}{2N(\alpha^{2}l)},w\biggr{)}\biggr{|}.

Recall the definition of k2k_{2} in (2.17). We apply the Cauchy-Schwarz inequality to see that for any integer K1K\geq 1,

KN(k)<2KUδ(α,k,w)(N(k)=K2K1N(k2)|L(1+w+δ,χik1)L(1+wδ,χik1)|2)12×(N(k)=K2K11N(k2)|N(l)=L(l,2α)=12L1βlN(l)𝒢δ(1+w;k,l,α)h(N(k)X2N(α2l),w)|2)12.\displaystyle\begin{split}\sum_{\begin{subarray}{c}K\leq N(k)<2K\end{subarray}}U_{\delta}(\alpha,k,w)\ll&\Big{(}\sum_{N(k)=K}^{2K-1}N(k_{2})|L(1+w+\delta,\chi_{ik_{1}})L(1+w-\delta,\chi_{ik_{1}})|^{2}\Big{)}^{\frac{1}{2}}\\ &\times\biggl{(}\sum_{N(k)=K}^{2K-1}\frac{1}{N(k_{2})}\biggl{|}\sum_{\begin{subarray}{c}N(l)=L\\ (l,2\alpha)=1\end{subarray}}^{2L-1}\frac{\beta_{l}}{N(l)}{\mathcal{G}}_{\delta}(1+w;k,l,\alpha)h\biggl{(}\frac{N(k)X}{2N(\alpha^{2}l)},w\biggr{)}\biggr{|}^{2}\ \biggr{)}^{\frac{1}{2}}.\end{split}

Using the Cauchy-Schwarz inequality again together with Lemma 2.8, we see that

(N(k)=K2K1N(k2)|L(1+w+δ,χik1)L(1+wδ,χik1)|2)12(K(1+|w|2))12+ϵ.\Big{(}\sum_{N(k)=K}^{2K-1}N(k_{2})|L(1+w+\delta,\chi_{ik_{1}})L(1+w-\delta,\chi_{ik_{1}})|^{2}\Big{)}^{\frac{1}{2}}\ll(K(1+|w|^{2}))^{\frac{1}{2}+\epsilon}.

This implies that

(5.21) KN(k)<2KU(α,k,w)(K(1+|w|2))12+ϵ(N(k)=K2K11N(k2)|N(l)=L(l,2α)=12L1βlN(l)𝒢δ(1+w;k,l,α)h(N(k)X2N(α2l),w)|2)12.\displaystyle\begin{split}&\sum_{\begin{subarray}{c}K\leq N(k)<2K\end{subarray}}U(\alpha,k,w)\ll(K(1+|w|^{2}))^{\frac{1}{2}+\epsilon}\biggl{(}\sum_{N(k)=K}^{2K-1}\frac{1}{N(k_{2})}\biggl{|}\sum_{\begin{subarray}{c}N(l)=L\\ (l,2\alpha)=1\end{subarray}}^{2L-1}\frac{\beta_{l}}{N(l)}{\mathcal{G}}_{\delta}(1+w;k,l,\alpha)h\biggl{(}\frac{N(k)X}{2N(\alpha^{2}l)},w\biggr{)}\biggr{|}^{2}\ \biggr{)}^{\frac{1}{2}}.\end{split}

Now, applying the first bound of Lemma 2.15 as well as the estimation given in (2.13) for Ψ(1+w)\Psi(1+w), we see that we may restrict the sum of the right side of (5.21) to K=2jN(α)2L(1+|w|2)(logX)4K=2^{j}\leq N(\alpha)^{2}L(1+|w|^{2})(\log X)^{4}. We then apply the second bound in Lemma 2.15 to (5.21) to see that

KN(k)<2KU(α,k,w)(K(1+|w|2))12+ϵ×((1+|w|)N(α)LX)ε|Φ^(1+w)|(N(α)2L(1+|w|2)K)|(τ)||(δ)|×N(α)L1/2K1/2X1/2|(δ)|(K1/2+L1/2)(1+|w|2)12+ϵ×(N(α)LX)ε|Φ^(1+w)|(N(α)2L(1+|w|2)K)|(τ)||(δ)|×N(α)L1/2X1/2|(δ)|(K1/2+L1/2).\displaystyle\begin{split}\sum_{\begin{subarray}{c}K\leq N(k)<2K\end{subarray}}U(\alpha,k,w)\ll&(K(1+|w|^{2}))^{\frac{1}{2}+\epsilon}\times((1+|w|)N(\alpha)LX)^{\varepsilon}|\widehat{\Phi}(1+w)|\Big{(}\frac{N(\alpha)^{2}L(1+|w|^{2})}{K}\Big{)}^{|\Re(\tau)|-|\Re(\delta)|}\\ &\times\frac{N(\alpha)L^{1/2}}{K^{1/2}X^{1/2-|\Re(\delta)|}}(K^{1/2}+L^{1/2})\\ \ll&(1+|w|^{2})^{\frac{1}{2}+\epsilon}\times(N(\alpha)LX)^{\varepsilon}|\widehat{\Phi}(1+w)|\Big{(}\frac{N(\alpha)^{2}L(1+|w|^{2})}{K}\Big{)}^{|\Re(\tau)|-|\Re(\delta)|}\\ &\times\frac{N(\alpha)L^{1/2}}{X^{1/2-|\Re(\delta)|}}(K^{1/2}+L^{1/2}).\end{split}

Summing over K=2j,KN(α)2L(1+|w|2)(logX)4K=2^{j},K\leq N(\alpha)^{2}L(1+|w|^{2})(\log X)^{4}, we deduce from the above that

k𝒪Kk0U(α,k,w)(1+|w|2)1+ϵ(N(α)L)ε|Φ^(1+w)|(N(α)2LX12|(δ)|ε+N(α)1+2κL1+κX12|(δ)|ε).\sum_{\begin{subarray}{c}k\in\mathcal{O}_{K}\\ k\neq 0\end{subarray}}U(\alpha,k,w)\ll(1+|w|^{2})^{1+\epsilon}(N(\alpha)L)^{\varepsilon}|\widehat{\Phi}(1+w)|\Big{(}\frac{N(\alpha)^{2}L}{X^{\frac{1}{2}-|\Re(\delta)|-\varepsilon}}+\frac{N(\alpha)^{1+2\kappa}L^{1+\kappa}}{X^{\frac{1}{2}-|\Re(\delta)|-\varepsilon}}\Big{)}.

It follows that

l=L2L1|(l)|\displaystyle\sum_{l=L}^{2L-1}|{\mathcal{R}}(l)|\ll L1+ϵX12|(δ)|ϵN(α)YN(α)ϵ(1+N(α)1+2κLκ)(12+|(δ)|+ϵ)|Ψ^(1+w)|(1+|w|2)1+ϵ|dw|.\displaystyle\frac{L^{1+\epsilon}}{X^{\frac{1}{2}-|\Re(\delta)|-\epsilon}}\sum_{N(\alpha)\leq Y}N(\alpha)^{\epsilon}(1+N(\alpha)^{-1+2\kappa}L^{\kappa})\int\limits_{(-\frac{1}{2}+|\Re(\delta)|+\epsilon)}|{\widehat{\Psi}}(1+w)|(1+|w|^{2})^{1+\epsilon}|dw|.

Now, the bound given in (3.8) follows from this by using (2.13) for Ψ(1+w)\Psi(1+w) with n=3n=3 when |w|Ψ(4)/Ψ(3)|w|\leq\Psi_{(4)}/\Psi_{(3)} and n=4n=4 otherwise to evaluate the integral above.

5.5. Conclusion

Combining (5.3), (5.9), (5.20) and our result for the average size of (l){\mathcal{R}}(l) given in the previous section, we see that in order to prove Proposition 3.4, it remains to simply the expression 𝒫(l)+𝒫+(l)+𝒫(l){\mathcal{P}}(l)+{\mathcal{P}}_{+}(l)+{\mathcal{P}}_{-}(l) given in (5.9) and (5.20). We now employ the identity given in (5.19) to see that the contribution from the poles at μδ\mu\delta to (5.9) and (5.20) cancel precisely each other, while the contribution from the poles at ±τ\pm\tau in both these expressions gives rise to the main term on the right side of (3.7), This completes the proof of Proposition 3.4.

6. Proof of Theorem 1.1

We now proceed to complete our proof of Theorem 1.1. We begin by evaluating 𝒲(δ1,Φ){\mathcal{W}}(\delta_{1},\Phi) when δ1\delta_{1} is near 0.

6.1. Proof of Proposition 3.5

We apply Lemma 2.5 together with the definition of ξ(s,χ(1+i)5d)\xi(s,\chi_{(1+i)^{5}d}) given in (2.5) to see that, by setting Φτ(t)=tτΦ(t)\Phi_{-\tau}(t)=t^{-\tau}\Phi(t),

(6.1) 𝒮(1;Φ)𝒲(δ1,Φ)=(25X/π2)τΓδ(τ)𝒮(Aδ,τ(d)|M(12+δ1,d)|2;Φτ).\displaystyle{\mathcal{S}}(1;\Phi){\mathcal{W}}(\delta_{1},\Phi)=\frac{(2^{5}X/\pi^{2})^{-\tau}}{\Gamma_{\delta}(\tau)}{\mathcal{S}}(A_{\delta,\tau}(d)|M(\tfrac{1}{2}+\delta_{1},d)|^{2};\Phi_{-\tau}).

We further set Y=X4κY=X^{4\kappa} to rewrite the above as

(6.2) 𝒮(1;Φ)𝒲(δ1,Φ)=\displaystyle{\mathcal{S}}(1;\Phi){\mathcal{W}}(\delta_{1},\Phi)= (25X/π2)τΓδ(τ){𝒮M(Aδ,τ(d)|M(12+δ1,d)|2;Φτ)+O(𝒮R(Aδ,τ(d)|M(12+δ1,d)|2;Φτ))}\displaystyle\frac{(2^{5}X/\pi^{2})^{-\tau}}{\Gamma_{\delta}(\tau)}\Big{\{}{\mathcal{S}}_{M}(A_{\delta,\tau}(d)|M(\tfrac{1}{2}+\delta_{1},d)|^{2};\Phi_{-\tau})+O({\mathcal{S}}_{R}(A_{\delta,\tau}(d)|M(\tfrac{1}{2}+\delta_{1},d)|^{2};\Phi_{-\tau}))\Big{\}}
(6.3) =\displaystyle= (25X/π2)τΓδ(τ)𝒮M(Aδ,τ(d)|M(12+δ1,d)|2;Φτ)+O(Xκ+ε),\displaystyle\frac{(2^{5}X/\pi^{2})^{-\tau}}{\Gamma_{\delta}(\tau)}{\mathcal{S}}_{M}(A_{\delta,\tau}(d)|M(\tfrac{1}{2}+\delta_{1},d)|^{2};\Phi_{-\tau})+O\Big{(}X^{-\kappa+\varepsilon}\Big{)},

where the last estimation above follows from Proposition 3.3.

Note that we have

(6.4) 𝒮M(Aδ,τ(d)|M(12+δ1,d)|2;Φτ)\displaystyle{\mathcal{S}}_{M}(A_{\delta,\tau}(d)|M(\tfrac{1}{2}+\delta_{1},d)|^{2};\Phi_{-\tau})
(6.5) =\displaystyle= l1mod(1+i)3(r,s1mod(1+i)3rs=lλ(r)λ(s)N(r)12+δ1N(s)12+δ2)𝒮M(Aδ,τ(d)((1+i)5dl);Φτ),\displaystyle\sum_{\begin{subarray}{c}l\equiv 1\bmod(1+i)^{3}\end{subarray}}\Big{(}\sum_{\begin{subarray}{c}r,s\equiv 1\bmod(1+i)^{3}\\ rs=l\end{subarray}}\frac{\lambda(r)\lambda(s)}{N(r)^{\frac{1}{2}+\delta_{1}}N(s)^{\frac{1}{2}+\delta_{2}}}\Big{)}{\mathcal{S}}_{M}\Big{(}A_{\delta,\tau}(d)\left(\frac{(1+i)^{5}d}{l}\right);\Phi_{-\tau}\Big{)},

We now apply Proposition 3.4 to evaluate 𝒮M(Aδ,τ(d)|M(12+δ1,d)|2;Φτ){\mathcal{S}}_{M}(A_{\delta,\tau}(d)|M(\tfrac{1}{2}+\delta_{1},d)|^{2};\Phi_{-\tau}). Using the estimations that |λ(n)|N(n)ϵ|\lambda(n)|\ll N(n)^{\epsilon}, rδ(n)N(n)ϵr_{\delta}(n)\ll N(n)^{\epsilon}, d[i](n)N(n)ϵd_{[i]}(n)\ll N(n)^{\epsilon}, and τ1/εlogX\tau\geq-1/\varepsilon\log X, we deduce that various remainder terms in Proposition 3.4 contribute

(6.6) N(l)M2N(l)εN(l)12+τ(Xε(XN(l1))14+N(l)εXκ+εN(l1)2κ12Y14κ+|(l)|)Xκ+εΦ(3)Φ(4)ϵ.\displaystyle\ll\sum_{N(l)\leq M^{2}}\frac{N(l)^{\varepsilon}}{N(l)^{\frac{1}{2}+\tau}}\Big{(}\frac{X^{\varepsilon}}{(XN(l_{1}))^{\frac{1}{4}}}+\frac{N(l)^{\varepsilon}X^{\kappa+\varepsilon}N(l_{1})^{2\kappa-\frac{1}{2}}}{Y^{1-4\kappa}}+|{\mathcal{R}}(l)|\Big{)}\ll X^{-\kappa+\varepsilon}\Phi_{(3)}\Phi_{(4)}^{\epsilon}.

We obtain from (6.1)-(6.6) that

(6.7) 𝒮(1;Φ)𝒲(δ1,Φ)=(25X/π2)τΓδ(τ)l1mod(1+i)3(r,s1mod(1+i)3rs=lλ(r)λ(s)N(r)12+δ1N(s)12+δ2)(l)+O(XκϵΦ(3)Φ(4)ϵ),\displaystyle{\mathcal{S}}(1;\Phi){\mathcal{W}}(\delta_{1},\Phi)=\frac{(2^{5}X/\pi^{2})^{-\tau}}{\Gamma_{\delta}(\tau)}\sum_{\begin{subarray}{c}l\equiv 1\bmod(1+i)^{3}\end{subarray}}\Big{(}\sum_{\begin{subarray}{c}r,s\equiv 1\bmod(1+i)^{3}\\ rs=l\end{subarray}}\frac{\lambda(r)\lambda(s)}{N(r)^{\frac{1}{2}+\delta_{1}}N(s)^{\frac{1}{2}+\delta_{2}}}\Big{)}{\mathcal{M}}(l)+O(X^{-\kappa-\epsilon}\Phi_{(3)}\Phi_{(4)}^{\epsilon}),

where (l)=1(l)+2(l){\mathcal{M}}(l)={\mathcal{M}}_{1}(l)+{\mathcal{M}}_{2}(l) with

1(l)=\displaystyle{\mathcal{M}}_{1}(l)= 2π3ζK(2)N(l1)μ=±rδ(l1)Γδ(μτ)(25XN(l1)π2)μτΦ^(1+μττ)Z(1+2μτ;δ)ηδ(1+2μτ;l),\displaystyle\frac{2\pi}{3\zeta_{K}(2)\sqrt{N(l_{1})}}\sum_{\mu=\pm}r_{\delta}(l_{1})\Gamma_{\delta}(\mu\tau)\left(\frac{2^{5}X}{N(l_{1})\pi^{2}}\right)^{\mu\tau}{\widehat{\Phi}}(1+\mu\tau-\tau)Z(1+2\mu\tau;\delta)\eta_{\delta}(1+2\mu\tau;l),
2(l)=\displaystyle{\mathcal{M}}_{2}(l)= 2π3ζK(2)N(l1)μ=±rτ(l1)Γτ(μδ)(25XN(l1)π2)μδΦ^(1+μδτ)Z(1+2μδ;τ)ητ(1+2μδ;l).\displaystyle\frac{2\pi}{3\zeta_{K}(2)\sqrt{N(l_{1})}}\sum_{\mu=\pm}r_{\tau}(l_{1})\Gamma_{\tau}(\mu\delta)\left(\frac{2^{5}X}{N(l_{1})\pi^{2}}\right)^{\mu\delta}{\widehat{\Phi}}(1+\mu\delta-\tau)Z(1+2\mu\delta;\tau)\eta_{\tau}(1+2\mu\delta;l).

By taking 𝒞{\mathcal{C}} to be the closed contour described in the paragraph above Lemma 2.17, we can evaluate the contribution of 1(l){\mathcal{M}}_{1}(l) to (6.7) by

(6.8) 12πi𝒞2πΦ^(1+wτ)3ζK(2)Γδ(τ)(25Xπ2)wτZ(1+2w;δ)Γδ(w)2ww2τ2×{l1mod(1+i)3rδ(l1)N(l1)1/2+w(r,s1mod(1+i)3rs=lλ(r)λ(s)N(r)12+δ1N(s)12+δ2)ηδ(1+2w;l)}dw.\displaystyle\begin{split}&\frac{1}{2\pi i}\int\limits_{{\mathcal{C}}}\frac{2\pi{\widehat{\Phi}}(1+w-\tau)}{3\zeta_{K}(2)\Gamma_{\delta}(\tau)}\left(\frac{2^{5}X}{\pi^{2}}\right)^{w-\tau}Z(1+2w;\delta)\Gamma_{\delta}(w)\frac{2w}{w^{2}-\tau^{2}}\\ &\times\biggl{\{}\sum_{\begin{subarray}{c}l\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{r_{\delta}(l_{1})}{N(l_{1})^{1/2+w}}\Big{(}\sum_{\begin{subarray}{c}r,s\equiv 1\bmod(1+i)^{3}\\ rs=l\end{subarray}}\frac{\lambda(r)\lambda(s)}{N(r)^{\frac{1}{2}+\delta_{1}}N(s)^{\frac{1}{2}+\delta_{2}}}\Big{)}\eta_{\delta}(1+2w;l)\biggr{\}}dw.\end{split}

Using the fact that λ\lambda is supported on square-free elements in 𝒪K\mathcal{O}_{K}, we now write r=αar=\alpha a, s=αbs=\alpha b with α\alpha, aa, bb being primary, square-free and (a,b)=1(a,b)=1. This implies that l=α2abl=\alpha^{2}ab, l1=abl_{1}=ab, and l2=αl_{2}=\alpha so that the sum over ll in (6.8) can be written as

(6.9) α1mod(1+i)3a,b1mod(1+i)3(a,b)=1rδ(ab)N(ab)12+wλ(αa)λ(αb)N(α)1+δ1+δ2N(a)12+δ1N(b)12+δ2ηδ(1+2w;α2ab)=α1mod(1+i)31N(α)1+2τa,b1mod(1+i)3(a,b)=1rδ(a)rδ(b)N(a)1+δ1+wN(b)1+δ2+wλ(αa)λ(αb)ηδ(1+2w;α2ab).\displaystyle\begin{split}&\sum_{\alpha\equiv 1\bmod(1+i)^{3}}\sum_{\begin{subarray}{c}a,b\equiv 1\bmod(1+i)^{3}\\ (a,b)=1\end{subarray}}\frac{r_{\delta}(ab)}{N(ab)^{\frac{1}{2}+w}}\frac{\lambda(\alpha a)\lambda(\alpha b)}{N(\alpha)^{1+\delta_{1}+\delta_{2}}N(a)^{\frac{1}{2}+\delta_{1}}N(b)^{\frac{1}{2}+\delta_{2}}}\eta_{\delta}(1+2w;\alpha^{2}ab)\\ =&\sum_{\alpha\equiv 1\bmod(1+i)^{3}}\frac{1}{N(\alpha)^{1+2\tau}}\sum_{\begin{subarray}{c}a,b\equiv 1\bmod(1+i)^{3}\\ (a,b)=1\end{subarray}}\frac{r_{\delta}(a)r_{\delta}(b)}{N(a)^{1+\delta_{1}+w}N(b)^{1+\delta_{2}+w}}\lambda(\alpha a)\lambda(\alpha b)\eta_{\delta}(1+2w;\alpha^{2}ab).\end{split}

We then apply the relation

ηδ(1+2w;α2ab)=ηδ(1+2w;1)hw(α)hw(a)hw(b)ϖGϖ|α(1+1N(ϖ)1+2w)\eta_{\delta}(1+2w;\alpha^{2}ab)=\frac{\eta_{\delta}(1+2w;1)}{h_{w}(\alpha)h_{w}(a)h_{w}(b)}\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|\alpha\end{subarray}}\Big{(}1+\frac{1}{N(\varpi)^{1+2w}}\Big{)}

to recast the expressions given in (6.9) as

ηδ(1+2w;1)α1mod(1+i)31N(α)1+2τhw(α)ϖGϖ|α(1+1N(ϖ)1+2w)a,b1mod(1+i)3(a,b)=1rδ(a)λ(αa)N(a)1+δ1+whw(a)rδ(b)λ(αb)N(b)1+δ2+whw(b).\eta_{\delta}(1+2w;1)\sum_{\alpha\equiv 1\bmod(1+i)^{3}}\frac{1}{N(\alpha)^{1+2\tau}h_{w}(\alpha)}\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|\alpha\end{subarray}}\Big{(}1+\frac{1}{N(\varpi)^{1+2w}}\Big{)}\sum_{\begin{subarray}{c}a,b\equiv 1\bmod(1+i)^{3}\\ (a,b)=1\end{subarray}}\frac{r_{\delta}(a)\lambda(\alpha a)}{N(a)^{1+\delta_{1}+w}h_{w}(a)}\frac{r_{\delta}(b)\lambda(\alpha b)}{N(b)^{1+\delta_{2}+w}h_{w}(b)}.

Next, we use the Möbius function to detect the condition (a,b)=1(a,b)=1 in the above expression to see that it equals to

(6.10) ηδ(1+2w;1)α1mod(1+i)31N(α)1+2τhw(α)ϖGϖ|α(1+1N(ϖ)1+2w)β1mod(1+i)3rδ(β)2μ[i](β)N(β)2+2τ+2whw(β)2×a,b1mod(1+i)3rδ(a)λ(aαβ)N(a)1+δ1+whw(a)rδ(b)λ(bαβ)N(b)1+δ2+whw(b).\displaystyle\begin{split}&\eta_{\delta}(1+2w;1)\sum_{\alpha\equiv 1\bmod(1+i)^{3}}\frac{1}{N(\alpha)^{1+2\tau}h_{w}(\alpha)}\prod_{\begin{subarray}{c}\varpi\in G\\ \varpi|\alpha\end{subarray}}\Big{(}1+\frac{1}{N(\varpi)^{1+2w}}\Big{)}\sum_{\begin{subarray}{c}\beta\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{r_{\delta}(\beta)^{2}\mu_{[i]}(\beta)}{N(\beta)^{2+2\tau+2w}h_{w}(\beta)^{2}}\\ &\times\sum_{\begin{subarray}{c}a,b\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{r_{\delta}(a)\lambda(a\alpha\beta)}{N(a)^{1+\delta_{1}+w}h_{w}(a)}\frac{r_{\delta}(b)\lambda(b\alpha\beta)}{N(b)^{1+\delta_{2}+w}h_{w}(b)}.\end{split}

We further group terms according to γ=αβ\gamma=\alpha\beta to see that (6.10) becomes

(6.11) ηδ(1+2w;1)γ1mod(1+i)3Hw(γ)N(γ)1+2τhw(γ)(a1mod(1+i)3rδ(a)λ(aγ)N(a)1+δ1+whw(a))(b1mod(1+i)3rδ(b)λ(bγ)N(b)1+δ2+whw(b)).\displaystyle\begin{split}\eta_{\delta}(1+2w;1)\sum_{\gamma\equiv 1\bmod(1+i)^{3}}\frac{H_{w}(\gamma)}{N(\gamma)^{1+2\tau}h_{w}(\gamma)}\Big{(}\sum_{\begin{subarray}{c}a\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{r_{\delta}(a)\lambda(a\gamma)}{N(a)^{1+\delta_{1}+w}h_{w}(a)}\Big{)}\Big{(}\sum_{\begin{subarray}{c}b\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{r_{\delta}(b)\lambda(b\gamma)}{N(b)^{1+\delta_{2}+w}h_{w}(b)}\Big{)}.\end{split}

We now apply Lemma 2.16 to evaluate the sum over aa given in (6.11) by first considering the case N(γ)M1bN(\gamma)\leq M^{1-b}. We set u=δ1+wu=\delta_{1}+w, v=δv=\delta, g(n)=1/hw(n)g(n)=1/h_{w}(n) and we denote G(s,γ;u,v)G(s,\gamma;u,v) by Gw(s,γ;u,v)G_{w}(s,\gamma;u,v) in Lemma 2.16 to see that by applying it with y=My=M, R(x)=P(x)R(x)=P(x) and then with y=M1by=M^{1-b}, R(x)=(1P(b+x(1b)))R(x)=(1-P(b+x(1-b))), the sum of these two applications yields

(6.12) N(a)M/N(γ)a1mod(1+i)3rδ(a)μ[i](aγ)N(a)1+δ1+whw(a)Q(log(M/N(aγ))logM)=μ[i](γ)Gw(1,γ;δ1+w,δ)ζK(1+δ1+w+δ)ζK(1+δ1+wδ)+O(E(γ)log2M(M1bN(γ))τ(w)exp(A0logM1b/N(γ)).\displaystyle\begin{split}&\sum_{\begin{subarray}{c}N(a)\leq M/N(\gamma)\\ a\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{r_{\delta}(a)\mu_{[i]}(a\gamma)}{N(a)^{1+\delta_{1}+w}h_{w}(a)}Q\Big{(}\frac{\log(M/N(a\gamma))}{\log M}\Big{)}\\ =&\frac{\mu_{[i]}(\gamma)G_{w}(1,\gamma;\delta_{1}+w,\delta)}{\zeta_{K}(1+\delta_{1}+w+\delta)\zeta_{K}(1+\delta_{1}+w-\delta)}+O\Big{(}\frac{E(\gamma)}{\log^{2}M}\left(\frac{M^{1-b}}{N(\gamma)}\right)^{-\tau-\Re(w)}\exp(-A_{0}\sqrt{\log M^{1-b}/N(\gamma)}\Big{)}.\end{split}

Here one checks that the sum of the k1k\geq 1 terms of the above two applications of Lemma 2.16 cancel each other. Thus the main term above comes from the sum of the k=0k=0 terms only. Note also that the main term above is |δ1|2\ll|\delta_{1}|^{2} due to the choice of the contour 𝒞{\mathcal{C}}.

Now, replacing aa by bb and δ1\delta_{1} by δ2\delta_{2} in (6.12), we see that a similar expression for the sum over bb in (6.11) holds. It follows that the case N(γ)M1bN(\gamma)\leq M^{1-b} in (6.11) equals

ηδ(1+2w;1)N(γ)M1bγ1mod(1+i)3Hw(γ)N(γ)1+2τhw(γ)μ2[i](γ)Gw(1,γ;δ1+w,δ)Gw(1,γ;δ2+w,δ)μ=±ζK(1+δ1+w+μδ)ζK(1+δ2+w+μδ)+O(N(γ)M1bγ1mod(1+i)3Hw(γ)N(γ)1+2τhw(γ)E(γ)|δ1|2log2M(M1bN(γ))τ(w)exp(A0log(M1b/N(γ)))=ηδ(1+2w;1)N(γ)M1bγ1mod(1+i)3Hw(γ)N(γ)1+2τhw(γ)μ2[i](γ)Gw(1,γ;δ1+w,δ)Gw(1,γ;δ2+w,δ)μ=±ζK(1+δ1+w+μδ)ζK(1+δ2+w+μδ)+O(|δ1|2log2MM(1b)(τ(w))).\displaystyle\begin{split}&\eta_{\delta}(1+2w;1)\sum_{\begin{subarray}{c}N(\gamma)\leq M^{1-b}\\ \gamma\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{H_{w}(\gamma)}{N(\gamma)^{1+2\tau}h_{w}(\gamma)}\frac{\mu^{2}_{[i]}(\gamma)G_{w}(1,\gamma;\delta_{1}+w,\delta)G_{w}(1,\gamma;\delta_{2}+w,\delta)}{\prod_{\mu=\pm}\zeta_{K}(1+\delta_{1}+w+\mu\delta)\zeta_{K}(1+\delta_{2}+w+\mu\delta)}\\ &+O\Big{(}\sum_{\begin{subarray}{c}N(\gamma)\leq M^{1-b}\\ \gamma\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{H_{w}(\gamma)}{N(\gamma)^{1+2\tau}h_{w}(\gamma)}E(\gamma)\frac{|\delta_{1}|^{2}}{\log^{2}M}\left(\frac{M^{1-b}}{N(\gamma)}\right)^{-\tau-\Re(w)}\exp(-A_{0}\sqrt{\log(M^{1-b}/N(\gamma))}\Big{)}\\ =&\eta_{\delta}(1+2w;1)\sum_{\begin{subarray}{c}N(\gamma)\leq M^{1-b}\\ \gamma\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{H_{w}(\gamma)}{N(\gamma)^{1+2\tau}h_{w}(\gamma)}\frac{\mu^{2}_{[i]}(\gamma)G_{w}(1,\gamma;\delta_{1}+w,\delta)G_{w}(1,\gamma;\delta_{2}+w,\delta)}{\prod_{\mu=\pm}\zeta_{K}(1+\delta_{1}+w+\mu\delta)\zeta_{K}(1+\delta_{2}+w+\mu\delta)}\\ &+O\Big{(}\frac{|\delta_{1}|^{2}}{\log^{2}M}M^{(1-b)(-\tau-\Re(w))}\Big{)}.\end{split}

We apply the last expression above in (6.8) to see that, based on our choice of 𝒞{\mathcal{C}} and MM, the contribution of the error term above is

(6.13) O(log2X|δ1|3M2τ(1b)).\displaystyle\begin{split}O\Big{(}\log^{2}X|\delta_{1}|^{3}M^{-2\tau(1-b)}\Big{)}.\end{split}

On the other hand, the main term coming from the contribution of the terms N(γ)M1bN(\gamma)\leq M^{1-b} to (6.8) equals to

(6.14) 12πi𝒞2πΦ^(1+wτ)3ζK(2)Γδ(τ)(25Xπ2)wτZ(1+2w;δ)Γδ(w)2ww2τ2ηδ(1+2w;1)×N(γ)M1bγ1mod(1+i)3Hw(γ)N(γ)1+2τhw(γ)μ2[i](γ)Gw(1,γ;δ1+w,δ)Gw(1,γ;δ2+w,δ)μ=±ζK(1+δ1+w+μδ)ζK(1+δ2+w+μδ)dw.\displaystyle\begin{split}\frac{1}{2\pi i}\int\limits_{{\mathcal{C}}}&\frac{2\pi{\widehat{\Phi}}(1+w-\tau)}{3\zeta_{K}(2)\Gamma_{\delta}(\tau)}\left(\frac{2^{5}X}{\pi^{2}}\right)^{w-\tau}Z(1+2w;\delta)\Gamma_{\delta}(w)\frac{2w}{w^{2}-\tau^{2}}\eta_{\delta}(1+2w;1)\\ &\times\sum_{\begin{subarray}{c}N(\gamma)\leq M^{1-b}\\ \gamma\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{H_{w}(\gamma)}{N(\gamma)^{1+2\tau}h_{w}(\gamma)}\frac{\mu^{2}_{[i]}(\gamma)G_{w}(1,\gamma;\delta_{1}+w,\delta)G_{w}(1,\gamma;\delta_{2}+w,\delta)}{\prod_{\mu=\pm}\zeta_{K}(1+\delta_{1}+w+\mu\delta)\zeta_{K}(1+\delta_{2}+w+\mu\delta)}dw.\end{split}

Note that as μ=±ζK(1+δ1+w+μδ)1ζK(1+δ2+w+μδ)1\prod_{\mu=\pm}\zeta_{K}(1+\delta_{1}+w+\mu\delta)^{-1}\zeta_{K}(1+\delta_{2}+w+\mu\delta)^{-1} vanishes at w=τw=-\tau, the integrand in (6.14) has only a simple pole at w=τw=\tau inside 𝒞{\mathcal{C}}. It then follows from Cauchy’s theorem that the expression in (6.14) equals

(6.15) 2πΦ^(1)3ζK(2)ηδ(1+2τ;1)ζK(1+2τ)N(γ)M1bγ1mod(1+i)3μ2[i](γ)Hτ(γ)N(γ)1+2τhτ(γ)Gτ(1,γ;δ1+τ,δ)Gτ(1,γ;δ2+τ,δ).\displaystyle\begin{split}\frac{2\pi{\widehat{\Phi}}(1)}{3\zeta_{K}(2)}\frac{\eta_{\delta}(1+2\tau;1)}{\zeta_{K}(1+2\tau)}\sum_{\begin{subarray}{c}N(\gamma)\leq M^{1-b}\\ \gamma\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{\mu^{2}_{[i]}(\gamma)H_{\tau}(\gamma)}{N(\gamma)^{1+2\tau}h_{\tau}(\gamma)}G_{\tau}(1,\gamma;\delta_{1}+\tau,\delta)G_{\tau}(1,\gamma;\delta_{2}+\tau,\delta).\end{split}

We apply Lemma 2.17 to evaluate (6.15) and then combine the result together with the error term given in (6.13) to see that the contribution to 𝒮(1;Φ)𝒲(δ1,Φ){\mathcal{S}}(1;\Phi){\mathcal{W}}(\delta_{1},\Phi) from the part of 1(l){\mathcal{M}}_{1}(l) restricting to the case N(γ)M1bN(\gamma)\leq M^{1-b} equals

(6.16) 2πΦ^(1)3ζK(2)(1M2τ(1b))+O(log2X|δ1|3M2τ(1b)).\displaystyle\frac{2\pi{\widehat{\Phi}}(1)}{3\zeta_{K}(2)}(1-M^{-2\tau(1-b)})+O(\log^{2}X|\delta_{1}|^{3}M^{-2\tau(1-b)}).

We now evaluate (6.11) for the case N(γ)>M1bN(\gamma)>M^{1-b} by applying Lemma 2.16 with uu, vv, g(n)g(n), Gw(s,γ;u,v)G_{w}(s,\gamma;u,v) as before, and with R(x)=P(x)R(x)=P(x), y=My=M this time. We deduce then that for any odd M1b<N(γ)<MM^{1-b}<N(\gamma)<M,

(6.17) N(a)M/N(γ)a1mod(1+i)3rδ(a)μ[i](aγ)N(a)1+δ1+whw(a)Q(log(M/N(aγ))logM)=O(E(γ)log2M(MN(γ))τ(w)exp(A0log(M/N(γ)))+Ress=0μ[i](γ)Gw(1+s,γ;δ1+w,δ)sζK(1+s+δ1+w+δ)ζK(1+s+δ1+wδ)k=01(slogM)kQ(k)(log(M/N(γ))logM).\displaystyle\begin{split}\sum_{\begin{subarray}{c}N(a)\leq M/N(\gamma)\\ a\equiv 1\bmod(1+i)^{3}\end{subarray}}&\frac{r_{\delta}(a)\mu_{[i]}(a\gamma)}{N(a)^{1+\delta_{1}+w}h_{w}(a)}Q\Big{(}\frac{\log(M/N(a\gamma))}{\log M}\Big{)}=O\Big{(}\frac{E(\gamma)}{\log^{2}M}\left(\frac{M}{N(\gamma)}\right)^{-\tau-\Re(w)}\exp(-A_{0}\sqrt{\log(M/N(\gamma)})\Big{)}\\ &+\mathop{\text{Res}}_{s=0}\frac{\mu_{[i]}(\gamma)G_{w}(1+s,\gamma;\delta_{1}+w,\delta)}{s\zeta_{K}(1+s+\delta_{1}+w+\delta)\zeta_{K}(1+s+\delta_{1}+w-\delta)}\sum_{k=0}^{\infty}\frac{1}{(s\log M)^{k}}Q^{(k)}\Big{(}\frac{\log(M/N(\gamma))}{\log M}\Big{)}.\end{split}

To evaluate the residue above, we write the Taylor expansion of Gw(1+s,γ;u,v)/(ζK(1+s+δ1+w+δ)ζK(1+s+δ1+wδ))G_{w}(1+s,\gamma;u,v)/(\zeta_{K}(1+s+\delta_{1}+w+\delta)\zeta_{K}(1+s+\delta_{1}+w-\delta)) as a0+a1s+a2s2+a_{0}+a_{1}s+a_{2}s^{2}+\ldots to see that a0=42π2(δ1+w+δ)(δ1+wδ)Gw(1,γ;δ1+w,δ)+O((|δ1|+|w|)3)a_{0}=\frac{4^{2}}{\pi^{2}}\cdot(\delta_{1}+w+\delta)(\delta_{1}+w-\delta)G_{w}(1,\gamma;\delta_{1}+w,\delta)+O((|\delta_{1}|+|w|)^{3}), a1=242π2(δ1+w)Gw(1,γ;δ1+w,δ)+O((|δ1|+|w|)2)a_{1}=2\cdot\frac{4^{2}}{\pi^{2}}(\delta_{1}+w)G_{w}(1,\gamma;\delta_{1}+w,\delta)+O((|\delta_{1}|+|w|)^{2}), a2=42π2Gw(1,γ;δ1+w,δ)+O(|δ1|+|w|)a_{2}=\frac{4^{2}}{\pi^{2}}\cdot G_{w}(1,\gamma;\delta_{1}+w,\delta)+O(|\delta_{1}|+|w|), and that ann1a_{n}\ll_{n}1 for n3n\geq 3. It follows from this that the residue term in (6.17) equals

(6.18) 42π2μ[i](γ)Gw(1,γ;δ1+w,δ)((δ1+w+δ)(δ1+wδ)Q(log(M/N(γ))logM)+2δ1+wlogMQ(log(M/N(γ))logM)+1log2MQ(log(M/N(γ))logM))+O(|δ1|3).\displaystyle\begin{split}\frac{4^{2}}{\pi^{2}}\cdot\mu_{[i]}(\gamma)G_{w}(1,\gamma;\delta_{1}+w,\delta)\Big{(}(\delta_{1}+w+\delta)(\delta_{1}+w-\delta)&Q\Big{(}\frac{\log(M/N(\gamma))}{\log M}\Big{)}+2\frac{\delta_{1}+w}{\log M}Q^{\prime}\Big{(}\frac{\log(M/N(\gamma))}{\log M}\Big{)}\\ &+\frac{1}{\log^{2}M}Q^{\prime\prime}\Big{(}\frac{\log(M/N(\gamma))}{\log M}\Big{)}\Big{)}+O(|\delta_{1}|^{3}).\end{split}

Again, replacing aa by bb and δ1\delta_{1} by δ2\delta_{2} in (6.18) yields an expression for the sum over bb in (6.11) when M1b<N(γ)<MM^{1-b}<N(\gamma)<M. We can thus apply these expressions to evaluate (6.11) for the case N(γ)>M1bN(\gamma)>M^{1-b}. By doing so, we see that the contribution of the remainder terms is

(6.19) M1bN(γ)M1N(γ)1+2τ(E(γ)log2M|δ1|2(MN(γ))τ(w)exp(A0log(M/N(γ)))+|δ1|5)|δ1|2log2MMτ(w)+M2τ(1b)|δ1|5logM.\displaystyle\begin{split}&\ll\sum_{M^{1-b}\leq N(\gamma)\leq M}\frac{1}{N(\gamma)^{1+2\tau}}\Big{(}\frac{E(\gamma)}{\log^{2}M}|\delta_{1}|^{2}\Big{(}\frac{M}{N(\gamma)}\Big{)}^{-\tau-\Re(w)}\exp(-A_{0}\sqrt{\log(M/N(\gamma))})+|\delta_{1}|^{5}\Big{)}\\ &\ll\frac{|\delta_{1}|^{2}}{\log^{2}M}M^{-\tau-\Re(w)}+M^{-2\tau(1-b)}|\delta_{1}|^{5}\log M.\end{split}

On the other hand, by writing Q(j)γQ^{(j)}_{\gamma} for Q(j)(log(M/N(γ))/logM)Q^{(j)}(\log(M/N(\gamma))/\log M), we obtain that the main term is

(4π)4ηδ(1+2w;1)M1b<N(γ)Mγ1mod(1+i)3μ2[i](γ)Hw(γ)N(γ)1+2τhw(γ)Gw(1,γ;δ1+w,δ)Gw(1,γ;δ2+w,γ)×((δ1+w+δ)(δ1+wδ)Qγ+2δ1+wlogMQγ+1log2MQγ)×((δ2+w+δ)(δ2+wδ)Qγ+2δ2+wlogMQγ+1log2MQγ).\displaystyle\begin{split}&(\frac{4}{\pi})^{4}\cdot\eta_{\delta}(1+2w;1)\sum_{\begin{subarray}{c}M^{1-b}<N(\gamma)\leq M\\ \gamma\equiv 1\bmod(1+i)^{3}\end{subarray}}\frac{\mu^{2}_{[i]}(\gamma)H_{w}(\gamma)}{N(\gamma)^{1+2\tau}h_{w}(\gamma)}G_{w}(1,\gamma;\delta_{1}+w,\delta)G_{w}(1,\gamma;\delta_{2}+w,\gamma)\\ &\times\Big{(}(\delta_{1}+w+\delta)(\delta_{1}+w-\delta)Q_{\gamma}+2\frac{\delta_{1}+w}{\log M}Q^{\prime}_{\gamma}+\frac{1}{\log^{2}M}Q^{\prime\prime}_{\gamma}\Big{)}\\ &\times\Big{(}(\delta_{2}+w+\delta)(\delta_{2}+w-\delta)Q_{\gamma}+2\frac{\delta_{2}+w}{\log M}Q^{\prime}_{\gamma}+\frac{1}{\log^{2}M}Q^{\prime\prime}_{\gamma}\Big{)}.\end{split}

We apply Lemma 2.17 to compute the expression above. By a suitable change of variables, we see that it equals

(6.20) N(w)+O(M2τ(1b)|δ1|5logM),\displaystyle\begin{split}N(w)+O(M^{-2\tau(1-b)}|\delta_{1}|^{5}\log M),\end{split}

where

N(w)=(4π)3logM0bM2τ(1x)((δ1+w+δ)(δ1+wδ)Q(x)+2δ1+wlogMQ(x)+Q(x)log2M)×((δ2+w+δ)(δ2+wδ)Q(x)+2δ2+wlogMQ(x)+Q(x)log2M)dx.\displaystyle\begin{split}N(w)=&(\frac{4}{\pi})^{3}\cdot\log M\int_{0}^{b}M^{-2\tau(1-x)}\Big{(}(\delta_{1}+w+\delta)(\delta_{1}+w-\delta)Q(x)+2\frac{\delta_{1}+w}{\log M}Q^{\prime}(x)+\frac{Q^{\prime\prime}(x)}{\log^{2}M}\Big{)}\\ &\times\Big{(}(\delta_{2}+w+\delta)(\delta_{2}+w-\delta)Q(x)+2\frac{\delta_{2}+w}{\log M}Q^{\prime}(x)+\frac{Q^{\prime\prime}(x)}{\log^{2}M}\Big{)}dx.\end{split}

We now apply (6.19) and (6.20) to evaluate the expression in (6.8) when N(γ)>M1bN(\gamma)>M^{1-b}. Based on our choices for MM and 𝒞{\mathcal{C}}, we see that the error terms in (6.19) and (6.20) contribute

(6.21) O(|δ1|3M2τlog2X+|δ1|6M2τ(1b)log5X).\displaystyle O\Big{(}|\delta_{1}|^{3}M^{-2\tau}\log^{2}X+|\delta_{1}|^{6}M^{-2\tau(1-b)}\log^{5}X\Big{)}.

Moreover, we see that the main term in (6.8) equals

(6.22) 12πi𝒞2πΦ^(1+wτ)3ζK(2)Γδ(τ)(25Xπ2)wτZ(1+2w;δ)Γδ(w)N(w)2ww2τ2dw.\displaystyle\frac{1}{2\pi i}\int\limits_{\mathcal{C}}\frac{2\pi{\widehat{\Phi}}(1+w-\tau)}{3\zeta_{K}(2)\Gamma_{\delta}(\tau)}\left(\frac{2^{5}X}{\pi^{2}}\right)^{w-\tau}Z(1+2w;\delta)\Gamma_{\delta}(w)N(w)\frac{2w}{w^{2}-\tau^{2}}dw.

Applying the relations

Φ^(1+wτ)Γδ(w)/Γδ(τ)=\displaystyle{\widehat{\Phi}}(1+w-\tau)\Gamma_{\delta}(w)/\Gamma_{\delta}(\tau)= Φ^(1)+O(|δ1|),\displaystyle{\widehat{\Phi}}(1)+O(|\delta_{1}|),
2wZ(1+2w;δ)=\displaystyle 2wZ(1+2w;\delta)= (π4)314(w2δ2)+O(|δ1|log2X),\displaystyle(\frac{\pi}{4})^{3}\cdot\frac{1}{4(w^{2}-\delta^{2})}+O(|\delta_{1}|\log^{2}X),
N(w)\displaystyle N(w)\ll M2τ(1b)|δ1|4logM,\displaystyle M^{-2\tau(1-b)}|\delta_{1}|^{4}\log M,

we see that the expression in (6.22) equals

(6.23) 2πΦ^(1)3ζK(2)12πi𝒞(25Xπ2)wτ(π4)3N(w)14(w2δ2)1w2τ2dw+O(|δ1|6M2τ(1b)log5X)=2πΦ^(1)3ζK(2)18δ1δ2τ((π4)3N(τ)(25Xπ2)2τ(π4)3N(τ))+O(|δ1|6M2τ(1b)log5X).\displaystyle\begin{split}&\frac{2\pi{\widehat{\Phi}}(1)}{3\zeta_{K}(2)}\frac{1}{2\pi i}\int\limits_{\mathcal{C}}\left(\frac{2^{5}X}{\pi^{2}}\right)^{w-\tau}(\frac{\pi}{4})^{3}N(w)\frac{1}{4(w^{2}-\delta^{2})}\frac{1}{w^{2}-\tau^{2}}dw+O(|\delta_{1}|^{6}M^{-2\tau(1-b)}\log^{5}X)\\ =&\frac{2\pi{\widehat{\Phi}}(1)}{3\zeta_{K}(2)}\frac{1}{8\delta_{1}\delta_{2}\tau}\Big{(}(\frac{\pi}{4})^{3}N(\tau)-\left(\frac{2^{5}X}{\pi^{2}}\right)^{-2\tau}(\frac{\pi}{4})^{3}N(-\tau)\Big{)}+O(|\delta_{1}|^{6}M^{-2\tau(1-b)}\log^{5}X).\end{split}

Now, using integration by parts together with the observations that Q(0)=Q(0)=0Q(0)=Q^{\prime}(0)=0, and Q(b)=1Q(b)=1, Q(b)=0Q^{\prime}(b)=0, we obtain after a little calculation that

(π4)3N(τ)=8δ1δ2τM2τ(1b)+4δ1δ2logM0bM2τ(1x)|Q(x)+Q(x)2δ1logM|2dx(π4)3N(τ)=4π4δ1δ2logM0bM2τ(1x)|Q(x)+Q(x)2δ1logM|2dx.\displaystyle\begin{split}(\frac{\pi}{4})^{3}N(\tau)=&8\delta_{1}\delta_{2}\tau M^{-2\tau(1-b)}+\frac{4\delta_{1}\delta_{2}}{\log M}\int_{0}^{b}M^{-2\tau(1-x)}\Big{|}Q^{\prime}(x)+\frac{Q^{\prime\prime}(x)}{2\delta_{1}\log M}\Big{|}^{2}dx\\ (\frac{\pi}{4})^{3}N(-\tau)=&\frac{4}{\pi}\cdot\frac{4\delta_{1}\delta_{2}}{\log M}\int_{0}^{b}M^{-2\tau(1-x)}\Big{|}Q^{\prime}(x)+\frac{Q^{\prime\prime}(x)}{2\delta_{1}\log M}\Big{|}^{2}dx.\end{split}

The above expressions allow us to recast (6.23) as

2πΦ^(1)3ζK(2)(M2τ(1b)+1(25X/π2)2τ2τlogM0bM2τ(1x)|Q(x)+Q(x)2δ1logM|2dx)+O(M2τ(1b)|δ1|6log5X).\displaystyle\begin{split}\frac{2\pi{\widehat{\Phi}}(1)}{3\zeta_{K}(2)}\Big{(}M^{-2\tau(1-b)}+\frac{1-(2^{5}X/\pi^{2})^{-2\tau}}{2\tau\log M}&\int_{0}^{b}M^{-2\tau(1-x)}\Big{|}Q^{\prime}(x)+\frac{Q^{\prime\prime}(x)}{2\delta_{1}\log M}\Big{|}^{2}dx\Big{)}+O(M^{-2\tau(1-b)}|\delta_{1}|^{6}\log^{5}X).\end{split}

It follows from this and (6.21) that the contribution of 1(l){\mathcal{M}}_{1}(l) to 𝒮(1;Φ)𝒲(δ1,Φ){\mathcal{S}}(1;\Phi){\mathcal{W}}(\delta_{1},\Phi) from the terms M1b<N(γ)MM^{1-b}<N(\gamma)\leq M equals

2πΦ^(1)3ζK(2)(M2τ(1b)+1(25X/π2)2τ2τlogM0bM2τ(1x)|Q(x)+Q(x)2δ1logM|2dx)+O(|δ1|3M2τlog2X+M2τ(1b)|δ1|6log5X).\displaystyle\begin{split}&\frac{2\pi{\widehat{\Phi}}(1)}{3\zeta_{K}(2)}\Big{(}M^{-2\tau(1-b)}+\frac{1-(2^{5}X/\pi^{2})^{-2\tau}}{2\tau\log M}\int_{0}^{b}M^{-2\tau(1-x)}\Big{|}Q^{\prime}(x)+\frac{Q^{\prime\prime}(x)}{2\delta_{1}\log M}\Big{|}^{2}dx\Big{)}\\ +&O(|\delta_{1}|^{3}M^{-2\tau}\log^{2}X+M^{-2\tau(1-b)}|\delta_{1}|^{6}\log^{5}X).\end{split}

Combining the above with (6.16), we conclude that the 1(l){\mathcal{M}_{1}}(l) contribution to 𝒮(1;Φ)𝒲(δ1,Φ){\mathcal{S}}(1;\Phi){\mathcal{W}}(\delta_{1},\Phi) equals

(6.24) 2πΦ^(1)3ζK(2)(1+1(25X/π2)2τ2τlogM0bM2τ(1x)|Q(x)+Q(x)2δ1logM|2dx)+O(M2τ(1b)|δ1|6log5X).\displaystyle\begin{split}&\frac{2\pi{\widehat{\Phi}}(1)}{3\zeta_{K}(2)}\Big{(}1+\frac{1-(2^{5}X/\pi^{2})^{-2\tau}}{2\tau\log M}\int_{0}^{b}M^{-2\tau(1-x)}\Big{|}Q^{\prime}(x)+\frac{Q^{\prime\prime}(x)}{2\delta_{1}\log M}\Big{|}^{2}dx\Big{)}+O(M^{-2\tau(1-b)}|\delta_{1}|^{6}\log^{5}X).\end{split}

We obtain the 2(l){\mathcal{M}}_{2}(l) contribution to (6.7) in a similar way to see that it equals

(6.25) 2πΦ^(1)3ζK(2)(25Xπ2)τ(25X/π2)δ(25X/π2)δ2δlogM0bM2τ(1x)|Q(x)+Q(x)2δ1logM|2dx+O(XτM2τ(1b)|δ1|6log5X).\displaystyle\begin{split}&-\frac{2\pi{\widehat{\Phi}}(1)}{3\zeta_{K}(2)}\left(\frac{2^{5}X}{\pi^{2}}\right)^{-\tau}\frac{(2^{5}X/\pi^{2})^{\delta}-(2^{5}X/\pi^{2})^{-\delta}}{2\delta\log M}\int_{0}^{b}M^{-2\tau(1-x)}\Big{|}Q^{\prime}(x)+\frac{Q^{\prime\prime}(x)}{2\delta_{1}\log M}\Big{|}^{2}dx\\ +&O(X^{-\tau}M^{-2\tau(1-b)}|\delta_{1}|^{6}\log^{5}X).\end{split}

By applying (4.2), (6.24) and (6.25) in (6.7), we deduce readily the statement of Proposition 3.5.

6.2. Completion of the proof

We are now able to complete our proof of Theorem 1.1. For this, we take our parameters exactly as those used by Conrey and Soundararajan in their proof of [C&S, Theorem 1]. Precisely, we take κ=1010,M=X125κ\kappa=10^{-10},M=X^{\frac{1}{2}-5\kappa} in (3.2) and S=π/(2(1b))(120κ))S=\pi/(2(1-b))(1-20\kappa)) with b=0.64b=0.64. We further take R=6.8R=6.8 and P(x)=3(x/b)22(x/b)3P(x)=3(x/b)^{2}-2(x/b)^{3}. Let σ0\sigma_{0} be given as in Proposition 3.2 and let Φ\Phi to be a smooth function supported in (1,2)(1,2) satisfying 0Φ(t)10\leq\Phi(t)\leq 1 for all tt, Φ(t)=1\Phi(t)=1 for t(1+ϵ,2ϵ)t\in(1+\epsilon,2-\epsilon), and |Φ(ν)(t)|ν,ε1|\Phi^{(\nu)}(t)|\ll_{\nu,\varepsilon}1. We apply Proposition 3.2 in (3.3) to deduce that

(6.26) 𝒩(X,Φ)X𝒮(1;Φ)8Ssinh(πR2S)(J1(X;Φ)+J2(X;Φ))+o(X),\displaystyle{\mathcal{N}}(X,\Phi)\leq\frac{X{\mathcal{S}}(1;\Phi)}{8S\sinh(\frac{\pi R}{2S})}(J_{1}(X;\Phi)+J_{2}(X;\Phi))+o(X),

where J1J_{1} and J2J_{2} are given in (3.4).

Now, Proposition 3.5 implies that for real numbers u,vu,v satisfying κlogX|u+iv|1/ϵ\kappa\log X\geq|u+iv|\geq-1/\epsilon and u1/ϵu\geq-1/\epsilon, we have

𝒲(u+ivlogX,Φ)=𝒱(u,v)+O(M2u(1b)/logX(1+|u|+|v|)6logX),{\mathcal{W}}\Big{(}\frac{u+iv}{\log X},\Phi\Big{)}={\mathcal{V}}(u,v)+O\Big{(}M^{-2u(1-b)/\log X}\frac{(1+|u|+|v|)^{6}}{\log X}\Big{)},

where

𝒱(u,v)=1+eulogX2logM(sinhuusinvv)0bM2u(1x)/logX|Q(x)+Q(x)logX2(x+iy)logM|2dx.{\mathcal{V}}(u,v)=1+\frac{e^{-u}\log X}{2\log M}\Big{(}\frac{\sinh u}{u}-\frac{\sin v}{v}\Big{)}\int_{0}^{b}M^{-2u(1-x)/\log X}\Big{|}Q^{\prime}(x)+\frac{Q^{\prime\prime}(x)\log X}{2(x+iy)\log M}\Big{|}^{2}dx.

The above expression implies that 𝒱(u,v)1{\mathcal{V}}(u,v)\geq 1, from which we deduce that

(6.27) J1(X;Φ)=0Scos(πt2S)log𝒱(R,t)dt+O(1logX).\displaystyle J_{1}(X;\Phi)=\int_{0}^{S}\cos\left(\frac{\pi t}{2S}\right)\log{\mathcal{V}}(-R,t)dt+O\Big{(}\frac{1}{\log X}\Big{)}.

Moreover, based on Proposition 3.2 and our choice for SS, we can extend the integral in the definition of J2(X;Φ)J_{2}(X;\Phi) in (3.4) to infinity with an negligible error. This way, we obtain that

(6.28) J2(X;Φ)=0sinh(πu2S)log𝒱(uR,S)du+o(1).\displaystyle J_{2}(X;\Phi)=\int_{0}^{\infty}\sinh\left(\frac{\pi u}{2S}\right)\log{\mathcal{V}}(u-R,S)du+o(1).

We conclude from (6.26)-(6.28) that

𝒩(X,Φ)\displaystyle{\mathcal{N}}(X,\Phi) X𝒮(1;Φ)8Ssinh(πR2S)(0Scos(πt2S)log𝒱(R,t)dt+0sinh(πu2S)log𝒱(uR,S)du)+o(X).\displaystyle\leq\frac{X{\mathcal{S}}(1;\Phi)}{8S\sinh\left(\frac{\pi R}{2S}\right)}\Big{(}\int_{0}^{S}\cos\left(\frac{\pi t}{2S}\right)\log{\mathcal{V}}(-R,t)dt+\int_{0}^{\infty}\sinh\left(\frac{\pi u}{2S}\right)\log{\mathcal{V}}(u-R,S)du\Big{)}+o(X).

As shown in [C&S, p. 10], the above further implies that 𝒩(X,Φ)0.79X𝒮(1;Φ)+o(X){\mathcal{N}}(X,\Phi)\leq 0.79X{\mathcal{S}}(1;\Phi)+o(X). By summing over X=x/2X=x/2, x/4x/4, \ldots, we obtain the proof of Theorem 1.1.

Acknowledgments. P. G. is supported in part by NSFC grant 11871082.

References