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

The fourth moment of central values of quadratic Hecke LL-functions in the Gaussian field

Peng Gao School of Mathematical Sciences, Beihang University, Beijing 100191, P. R. China [email protected]
Abstract.

We obtain an asymptotic formula for the fourth moment of central values of a family of quadratic Hecke LL-functions in the Gaussian field under the generalized Riemann hypothesis (GRH). We also establish lower bounds unconditionally and upper bounds under GRH for higher moments of the same family.

Mathematics Subject Classification (2010): 11M06, 11M41, 11M50

Keywords: central values, Hecke LL-functions, quadratic Hecke characters, moments

1. Introduction

Moments of central values of families of LL-functions have been intensively studied in the literature in order to understand important arithmetic information they carry. Although much progress has been made towards establishing asymptotic formulas for the first few moments for various families of LL-functions, little is known for the higher moments. In connection with random matrix theory, conjectures on the order of magnitude for the moments were made by J. P. Keating and N. C. Snaith in [Keating-Snaith02]. A simple and powerful method towards establishing lower bounds of these conjectured results was developed by Z. Rudnick and K. Soundararajan in [R&Sound] and [R&Sound1].

For the family of quadratic Dirichlet LL-functions, the above method of Rudnick and Soundararajan allows them to show in [R&Sound1] that for every even natural number kk,

(1.1) d𝒟|d|XL(12,χd)kkX(logX)k(k+1)2,\displaystyle\sum_{\begin{subarray}{c}d\in\mathcal{D}\\ |d|\leq X\end{subarray}}L(\tfrac{1}{2},\chi_{d})^{k}\gg_{k}X(\log X)^{\frac{k(k+1)}{2}},

where χd=(d)\chi_{d}=\left(\frac{d}{\cdot}\right) is the Kronecker symbol and 𝒟\mathcal{D} denotes the set of fundamental discriminants.

In the other direction, Soundararajan [Sound01] proved that, assuming the generalized Riemann hypothesis (GRH), for all real k0k\geq 0,

(1.2) d𝒟|d|XL(12,χd)kk,εX(logX)k(k+1)2+ε.\displaystyle\sum_{\begin{subarray}{c}d\in\mathcal{D}\\ |d|\leq X\end{subarray}}L(\tfrac{1}{2},\chi_{d})^{k}\ll_{k,\varepsilon}X(\log X)^{\frac{k(k+1)}{2}+\varepsilon}.

The above result was further sharpened by A. J. Harper in [Harper] to remove the ε\varepsilon power.

Besides the lower and upper bounds for the moments given in (1.1) and (1.2), asymptotic formulas are known for integers 1k41\leq k\leq 4 in the form

(1.3) 0<dX(d,2)=1d square-freeL(12,χ8d)k=XPk(k+1)2(logX)+Ek(X),\displaystyle\sum_{\begin{subarray}{c}0<d\leq X\\ (d,2)=1\\ \text{$d$ square-free}\end{subarray}}L(\tfrac{1}{2},\chi_{8d})^{k}=XP_{\frac{k(k+1)}{2}}(\log X)+E_{k}(X),

where Pn(x)P_{n}(x) is an explicit linear polynomials of degree nn and Ek(X)E_{k}(X) is the error term. Here we note that for odd, square-free d>0d>0, the character χ8d\chi_{8d} is primitive modulo 8d8d satisfying χ8d(1)=1\chi_{8d}(-1)=1. Note also that we choose to present the asymptotic formulas for a family which is preferred by most recent studies due to technical reasons instead of the family appearing in (1.1) and (1.2).

Evaluation of the first two moments (k=1,2k=1,2) in (1.3) was initiated by M. Jutila in [Jutila]. The error terms in Jutila’s results were subsequently improved in [ViTa, DoHo, Young1] for the first moment and in [sound1, Sono] for the second moment. For smoothed first moment, a result of D. Goldfeld and J. Hoffstein in [DoHo] implies that one can take E1(x)=O(X1/2+ε)E_{1}(x)=O(X^{1/2+\varepsilon}). An error term of the same size was later obtained by M. P. Young in [Young1] using a recursive approach. The same approach was then adapted by K. Sono [Sono] to show that one can take E2(x)=O(X1/2+ε)E_{2}(x)=O(X^{1/2+\varepsilon}) for smoothed second moment. The sizes of these error terms then make the expressions given in (1.3) in agreement with a conjecture made by J. B. Conrey, D. W. Farmer, J. P. Keating, M. O. Rubinstein and N. C. Snaith in [CFKRS] on asymptotic behaviours of these moments of the family of quadratic Dirichlet LL-functions.

The third moment given in (1.3) was originally obtained by Soundararajan in [sound1]. The error term in Soundararajan’s result was later improved by Young in [Young2] to be E3(X)=O(X34+ε)E_{3}(X)=O(X^{\frac{3}{4}+\varepsilon}) for the smoothed version. Under GRH, Q. Shen [Shen] established an asymptotic formula for the fourth moment given in (1.3) with some savings on the powers of logX\log X in the error term. The result of Shen is achieved by combining lower and upper bounds for the fourth moment, with a lower bound obtained unconditionally using the method of Rudnick and Soundararajan in [R&Sound, R&Sound1] (note that such a lower bound is stated in [R&Sound1] without proof). An upper bound for the fourth moment is obtained under GRH by making use of upper bounds for the shifted moments of quadratic Dirichlet LL-functions. This approach originated from earlier work of Soundararajan [Sound01] as well as Soundararajan and Young [S&Y], especially the treatment in [S&Y] on the second moment of quadratic twists of modular LL-functions.

The result of Soundararajan and Young in [S&Y] supplies another example on moments of quadratic families of LL-functions. In [Gao1], the author considered the analogue case of evaluating moments of central values of a family of quadratic Hecke LL-functions in the Gaussian field to obtain the smoothed first three moments of such family with power saving error terms.

To describe the family studied in [Gao1], we let K=(i)K=\mathbb{Q}(i) be the Gaussian field throughout the paper and 𝒪K=[i]\mathcal{O}_{K}=\mathbb{Z}[i] be the ring of integers of KK. We denote χ\chi for a Hecke character of KK and we recall that a Hecke character χ\chi of KK is said to be of trivial infinite type if its component at infinite places of KK is trivial. We write L(s,χ)L(s,\chi) for the LL-function associated to χ\chi and ζK(s)\zeta_{K}(s) for the Dedekind zeta function of KK. In the rest of the paper, we reserve the expression χc\chi_{c} for the the quadratic residue symbol (c)\left(\frac{c}{\cdot}\right) defined in Section 2.1. We also denote N(n)N(n) for the norm of any nKn\in K.

We then consider the family of LL-functions:

(1.4) ={L(12,χ(1+i)5d):d𝒪Kodd and square-free}.\displaystyle\mathcal{F}=\big{\{}L(\tfrac{1}{2},\chi_{(1+i)^{5}d}):d\in\mathcal{O}_{K}\hskip 3.61371pt\text{odd and square-free}\big{\}}.

Here we say that any d𝒪Kd\in\mathcal{O}_{K} 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. Note also that (see Section 2.1 below) the symbol 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 when d𝒪Kd\in\mathcal{O}_{K} is odd and square-free.

The first three moments of the above family \mathcal{F} is studied in [Gao1]. In this paper, our main purpose is to establish the fourth moment of the same family under GRH. In this process, we shall need to make use of results on upper bounds for shifted moments of the family under GRH. Therefore, we begin by considering lower and upper bounds for higher moments of this family. Unconditionally, we have the following result concerning the lower bounds.

Theorem 1.1.

For every even natural number kk, we have

(1.5) (d,2)=1N(d)XL(12,χ(1+i)5d)kkX(logX)k(k+1)/2.\displaystyle\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ N(d)\leq X\end{subarray}}L(\tfrac{1}{2},\chi_{(1+i)^{5}d})^{k}\gg_{k}X(\log X)^{k(k+1)/2}.

Here the “*” on the sum over dd means that the sum is restricted to square-free elements dd in 𝒪K\mathcal{O}_{K}.

The proof of Theorem 1.1 is given in Section 3 and it follows the arguments in the proof of [R&Sound1, Theorem 2] for the case of Dirichlet LL-functions. Also similar to the remarks given below [R&Sound1, Theorem 2], the proof of Theorem 1.1 can be applied to give lower bounds for the moments for all rational numbers kk, provided that we replace L(12,χd)kL(\tfrac{1}{2},\chi_{d})^{k} by |L(12,χd)|k|L(\tfrac{1}{2},\chi_{d})|^{k}.

Before we state a corresponding result on the upper bounds, we need to introduce some notations. Let xx\in\mathbb{R} such that x10x\geq 10 and zz\in\mathbb{C}, we define

(1.9) (z,x)={loglogx|z|(logx)1,log|z|(logx)1<|z|1,0|z|>1.\displaystyle\mathcal{L}(z,x)=\left\{\begin{array}[c]{ll}\operatorname{log}\operatorname{log}x&|z|\leq(\operatorname{log}x)^{-1},\\ -\operatorname{log}|z|&(\operatorname{log}x)^{-1}<|z|\leq 1,\\ 0&|z|>1.\end{array}\right.

We further define for z1,z2z_{1},z_{2}\in\mathbb{C},

(1.10) (z1,z2,x)=\displaystyle\mathcal{M}(z_{1},z_{2},x)= 12((z1,x)+(z2,x)),\displaystyle\frac{1}{2}\left(\mathcal{L}(z_{1},x)+\mathcal{L}(z_{2},x)\right),
(1.11) 𝒱(z1,z2,x)=\displaystyle\mathcal{V}(z_{1},z_{2},x)= 12((2z1,x)+(2z2,x)+(2(z1),x)+(2(z2),x)+2(z1+z2,x)+2(z1+z2¯,x)).\displaystyle\frac{1}{2}\left(\mathcal{L}(2z_{1},x)+\mathcal{L}(2z_{2},x)+\mathcal{L}(2\Re(z_{1}),x)+\mathcal{L}(2\Re(z_{2}),x)+2\mathcal{L}(z_{1}+z_{2},x)+2\mathcal{L}(z_{1}+\overline{z_{2}},x)\right).

Then we have the following result on upper bounds for shifted moments of the family \mathcal{F} given in (1.4) under GRH.

Theorem 1.2.

Assume GRH for ζK(s)\zeta_{K}(s) and L(s,χ(1+i)5d)L(s,\chi_{(1+i)^{5}d}) for all odd, square-free dd. Let XX be large and let z1,z2z_{1},z_{2}\in\mathbb{C} with 0(z1),(z2)1logX0\leq\Re(z_{1}),\Re(z_{2})\leq\frac{1}{\log X}, and |(z1)|,|(z2)|X|\Im(z_{1})|,|\Im(z_{2})|\leq X. Then for any positive real number kk and any ε>0\varepsilon>0, we have

(d,2)=1N(d)X|L(12+z1,χ(1+i)5d)L(12+z2,χ(1+i)5d)|kk,εX(logX)εexp(k(z1,z2,X)+k22𝒱(z1,z2,X)).\displaystyle\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ N(d)\leq X\end{subarray}}|L(\tfrac{1}{2}+z_{1},\chi_{(1+i)^{5}d})L(\tfrac{1}{2}+z_{2},\chi_{(1+i)^{5}d})|^{k}\ll_{k,\varepsilon}X(\operatorname{log}X)^{\varepsilon}\operatorname{exp}\left(k\mathcal{M}(z_{1},z_{2},X)+\frac{k^{2}}{2}\mathcal{V}(z_{1},z_{2},X)\right).

The proof of Theorem 1.2 is given in Section 4 and our proof follows closely the approaches in [Sound01, S&Y, Shen]. We note that similar results to Theorem 1.2 were obtained for the moments of the Riemann zeta function by V. Chandee [Chandee] and for the moments of all Dirichlet LL-functions modulo qq by M. Munsch [Munsch].

We now give two consequences of Theorem 1.2. First, by setting z1=z2=0z_{1}=z_{2}=0 in Theorem 1.2, we deduce immediately the following upper bounds for moments of the central values of the family \mathcal{F} given in (1.4).

Corollary 1.3.

Assume GRH for ζK(s)\zeta_{K}(s) and L(s,χ(1+i)5d)L(s,\chi_{(1+i)^{5}d}) for all odd, square-free dd. For any positive real number kk and any ε>0\varepsilon>0, we have for large XX,

(d,2)=1N(d)X|L(12,χ(1+i)5d)|kk,εX(logX)k(k+1)/2+ε.\displaystyle\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ N(d)\leq X\end{subarray}}|L(\tfrac{1}{2},\chi_{(1+i)^{5}d})|^{k}\ll_{k,\varepsilon}X(\operatorname{log}X)^{k(k+1)/2+\varepsilon}.

The second consequence concerns upper bounds for shifted fourth moment of the family \mathcal{F} given in (1.4), which is what we need in a refined study in Section 5 on the fourth moment under GRH.

Corollary 1.4.

Assume GRH for ζK(s)\zeta_{K}(s) and L(s,χ(1+i)5d)L(s,\chi_{(1+i)^{5}d}) for all odd, square-free dd. Let XX be large and let z1,z2z_{1},z_{2}\in\mathbb{C} with 0(z1),(z2)1logX0\leq\Re(z_{1}),\Re(z_{2})\leq\frac{1}{\log X} and |(z1)|,|(z2)|X|\Im(z_{1})|,|\Im(z_{2})|\leq X. Then

(d,2)=1N(d)X|L(12+z1,χ(1+i)5d)|2|L(12+z2,χ(1+i)5d)|2X(logX)4+ε(1+min{(logX)6,1(|(z1)||(z2)|)6}).\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ N(d)\leq X\end{subarray}}|L(\tfrac{1}{2}+z_{1},\chi_{(1+i)^{5}d})|^{2}|L(\tfrac{1}{2}+z_{2},\chi_{(1+i)^{5}d})|^{2}\ll X(\operatorname{log}X)^{4+\varepsilon}\left(1+\min\left\{(\log X)^{6},\frac{1}{(|\Im(z_{1})|-|\Im(z_{2})|)^{6}}\right\}\right).

We shall omit the proof of the above corollary as it is analogous to the proof [Shen, Theorem 2.4]. With the aide of Corollary 1.4, we are able to obtain a more precise expression for the fourth moment. To present our result, we define constants aka_{k} for any real number k>0k>0 by

(1.12) ak=2k(k+2)2ϖ1mod(1+i)3(11N(ϖ))k(k+1)21+1N(ϖ)((1+1N(ϖ))k+(11N(ϖ))k2+1N(ϖ)).\displaystyle a_{k}=2^{-\frac{k(k+2)}{2}}\prod_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\end{subarray}}\frac{(1-\frac{1}{N(\varpi)})^{\frac{k(k+1)}{2}}}{1+\frac{1}{N(\varpi)}}\left(\frac{(1+\frac{1}{\sqrt{N(\varpi)}})^{-k}+(1-\frac{1}{\sqrt{N(\varpi)}})^{-k}}{2}+\frac{1}{N(\varpi)}\right).

Here and in what follows, we denote ω\omega for a prime number in 𝒪K\mathcal{O}_{K}, by which we mean that the ideal (ω)(\omega) generated by ω\omega is a prime ideal. We also note that the expression ϖ1mod(1+i)3\varpi\equiv 1\bmod{(1+i)^{3}} indicates that ω\omega is a primary element in 𝒪K\mathcal{O}_{K} defined in Section 2.1.

Now, we state our result on a conditional asymptotic evaluation on the fourth moment.

Theorem 1.5.

Assume GRH for ζK(s)\zeta_{K}(s) and L(s,χ(1+i)5d)L(s,\chi_{(1+i)^{5}d}) for all odd, square-free dd. Then for any ε>0\varepsilon>0, we have

(d,2)=1N(d)X|L(12,χ(1+i)5d)|4=πa42734527ζK(2)(π4)10X(logX)10+O(X(logX)9.5+ε),\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ N(d)\leq X\end{subarray}}|L(\tfrac{1}{2},\chi_{(1+i)^{5}d})|^{4}=\frac{\pi a_{4}}{2^{7}\cdot 3^{4}\cdot 5^{2}\cdot 7\cdot\zeta_{K}(2)}\big{(}\frac{\pi}{4}\big{)}^{10}X{(\log X)}^{10}+O\left(X(\log X)^{9.5+\varepsilon}\right),

Here the “*” on the sum over dd means that the sum is restricted to square-free elements dd in 𝒪K\mathcal{O}_{K} and a4a_{4} is defined as in (1.12).

Without assuming GRH, we also have the following lower bound for the fourth moment.

Theorem 1.6.

Unconditionally, we have

(d,2)=1L(12,χ(1+i)5d)4Φ(N(d)X)(πa42734527ζK(2)(π4)10+o(1))X(logX)10.\displaystyle\sideset{}{{}^{*}}{\sum}_{(d,2)=1}L(\tfrac{1}{2},\chi_{(1+i)^{5}d})^{4}\Phi\left(\frac{N(d)}{X}\right)\geq\left(\frac{\pi a_{4}}{2^{7}\cdot 3^{4}\cdot 5^{2}\cdot 7\cdot\zeta_{K}(2)}\big{(}\frac{\pi}{4}\big{)}^{10}+o(1)\right)X(\log X)^{10}.

Here the “*” on the sum over dd means that the sum is restricted to square-free elements dd in 𝒪K\mathcal{O}_{K} and a4a_{4} is defined as in (1.12).

Theorems 1.5 and 1.6 are similar to those of Shen given in [Shen, Theorems 1.1-1.2] on the fourth moment of the family of quadratic Dirichlet LL-functions as well as the result of Soundararajan and Young given in [S&Y, Theorem 1.1] on the second moment of the family of quadratic twists of modular LL-functions. The proof for Theorems 1.5 and 1.6 given in Section 5, also proceed along the same lines of the proofs of [Shen, Theorems 1.1-1.2] and [S&Y, Theorem 1.1].

2. Preliminaries

As a preparation, we first include some auxiliary results needed in the proofs of our theorems.

2.1. Quadratic residue symbol and quadratic Gauss sum

Recall that K=(i)K=\mathbb{Q}(i) and it is well-known that KK have class number one. We denote UK={±1,±i}U_{K}=\{\pm 1,\pm i\} and DK=4D_{K}=-4 for the group of units in 𝒪K\mathcal{O}_{K} and the discriminant of KK, respectively.

Every ideal in 𝒪K\mathcal{O}_{K} co-prime to 22 has a unique generator congruent to 11 modulo (1+i)3(1+i)^{3} which is called primary. It follows from Lemma 6 on [I&R, p. 121] that an element n=a+bi𝒪Kn=a+bi\in\mathcal{O}_{K} with a,ba,b\in\mathbb{Z} is primary if and only if a1(mod4),b0(mod4)a\equiv 1\pmod{4},b\equiv 0\pmod{4} or a3(mod4),b2(mod4)a\equiv 3\pmod{4},b\equiv 2\pmod{4}.

For n𝒪K,(n,2)=1n\in\mathcal{O}_{K},(n,2)=1, we denote the symbol (n)\left(\frac{\cdot}{n}\right) for the quadratic residue symbol modulo nn in KK. For a prime ϖ[i]\varpi\in\mathbb{Z}[i] with N(ϖ)2N(\varpi)\neq 2, the quadratic symbol is defined for a𝒪Ka\in\mathcal{O}_{K}, (a,ϖ)=1(a,\varpi)=1 by (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 ϖ|a\varpi|a, we define (aϖ)=0\left(\frac{a}{\varpi}\right)=0. Then the quadratic symbol is extended to any composite nn with (N(n),2)=1(N(n),2)=1 multiplicatively. We further define (c)=1\left(\frac{\cdot}{c}\right)=1 for cUKc\in U_{K}.

The following quadratic reciprocity law (see [G&Zhao4, (2.1)]) holds for two co-prime primary elements m,n𝒪Km,n\in\mathcal{O}_{K}:

(2.1) (mn)=(nm).\displaystyle\left(\frac{m}{n}\right)=\left(\frac{n}{m}\right).

Moreover, we deduce from Lemma 8.2.1 and Theorem 8.2.4 in [BEW] that the following supplementary laws hold for primary n=a+bin=a+bi with a,ba,b\in\mathbb{Z}:

(2.2) (in)=(1)(1a)/2and(1+in)=(1)(ab1b2)/4.\displaystyle\left(\frac{i}{n}\right)=(-1)^{(1-a)/2}\qquad\mbox{and}\qquad\hskip 7.22743pt\left(\frac{1+i}{n}\right)=(-1)^{(a-b-1-b^{2})/4}.

For any complex number zz, we define

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).

For any r𝒪Kr\in\mathcal{O}_{K}, we define the quadratic Gauss sum g(r,χ)g(r,\chi) associated to any quadric Hecke character χ\chi modulo qq of trivial infinite type and the quadratic Gauss sum g(r,n)g(r,n) associated to the quadratic residue symbol (n)\left(\frac{\cdot}{n}\right) for any (n,2)=1(n,2)=1 by

(2.3) g(r,χ)=xmodqχ(x)e~(rxq),g(r,n)=xmodn(xn)e~(rxn).\displaystyle g(r,\chi)=\sum_{x\bmod{q}}\chi(x)\widetilde{e}\left(\frac{rx}{q}\right),\quad g(r,n)=\sum_{x\bmod{n}}\left(\frac{x}{n}\right)\widetilde{e}\left(\frac{rx}{n}\right).

When r=1r=1, we shall denote g(χ)g(\chi) for g(1,χ)g(1,\chi) and g(n)g(n) for g(1,n)g(1,n). Recall from [G&Zhao3, (2.2)] that for primary nn, we have

(2.4) g(n)=(in)N(n)1/2.\displaystyle g(n)=\left(\frac{i}{n}\right)N(n)^{1/2}.

A Hecke character χ\chi is said to be 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). Recall from Section 1 that we denote χc\chi_{c} for the the quadratic residue symbol (c)\left(\frac{c}{\cdot}\right) and we define χc(n)\chi_{c}(n) to be 0 when 1+i|n1+i|n. In Section 2.1 of [G&Zhao4], it is shown that the symbol χi(1+i)5d\chi_{i(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 and square-free d𝒪Kd\in\mathcal{O}_{K}. When replacing dd by i3di^{3}d, we see that the symbol χ(1+i)5d\chi_{(1+i)^{5}d} also defines a primitive quadratic Hecke character modulo (1+i)5d(1+i)^{5}d of trivial infinite type for any odd and square-free d𝒪Kd\in\mathcal{O}_{K}. Our next lemma evaluates g(χ(1+i)5d)g(\chi_{(1+i)^{5}d}) exactly.

Lemma 2.2.

For any odd, square-free d𝒪Kd\in\mathcal{O}_{K}, we have

(2.5) g(χ(1+i)5d)=N((1+i)5d)1/2.\displaystyle g(\chi_{(1+i)^{5}d})=N((1+i)^{5}d)^{1/2}.
Proof.

It suffices to prove (2.5) with dd replaced by jdjd, where j=1j=1 or ii and dd is primary and square-free. It follows from the Chinese remainder theorem that x=j(1+i)5y+dzx=j(1+i)^{5}y+dz varies over the residue class modulo (1+i)3d(1+i)^{3}d when yy and zz vary over the residue class modulo dd and (1+i)5(1+i)^{5}, respectively. We then deduce that

g(χj(1+i)5d)=\displaystyle g(\chi_{j(1+i)^{5}d})= zmod(1+i)5ymodd(j(1+i)5j(1+i)5y+dz)(dj(1+i)5y+dz)e~(jyd)e~(z(1+i)5).\displaystyle\sum_{z\bmod{(1+i)^{5}}}\sum_{y\bmod{d}}\left(\frac{j(1+i)^{5}}{j(1+i)^{5}y+dz}\right)\left(\frac{d}{j(1+i)^{5}y+dz}\right)\widetilde{e}\left(\frac{jy}{d}\right)\widetilde{e}\left(\frac{z}{(1+i)^{5}}\right).

As χj(1+i)5\chi_{j(1+i)^{5}} is a Hecke character of trivial infinite type modulo (1+i)5(1+i)^{5}, we deduce that

(2.6) (j(1+i)5j(1+i)5y+dz)=χj(1+i)5(j(1+i)5y+dz)=χj(1+i)5(dz).\displaystyle\left(\frac{j(1+i)^{5}}{j(1+i)^{5}y+dz}\right)=\chi_{j(1+i)^{5}}(j(1+i)^{5}y+dz)=\chi_{j(1+i)^{5}}(dz).

On the other hand, we denote s(z)s(z) to be the unique element in UKU_{K} such that s(z)zs(z)z is primary for any (z,2)=1(z,2)=1. It follows from the quadratic reciprocity law (2.1) that

(2.7) (dj(1+i)5y+dz)=(s(z)(1+i)5jyd).\displaystyle\left(\frac{d}{j(1+i)^{5}y+dz}\right)=\left(\frac{s(z)(1+i)^{5}jy}{d}\right).

We then conclude from (2.6) and (2.7) that

(2.8) g(χ(1+i)5d)=zmod(1+i)5ymodd(j(1+i)5dz)(s(z)(1+i)jyd)e~(jyd)e~(z(1+i)5)=zmod(1+i)5(j(1+i)5z)(s(z)jd)e~(z(1+i)5)ymodd(jyd)e~(jyd)=N(d)1/2zmod(1+i)5(j(1+i)5z)(s(z)ijd)e~(z(1+i)5),\displaystyle\begin{split}g(\chi_{(1+i)^{5}d})=&\sum_{z\bmod{(1+i)^{5}}}\sum_{y\bmod{d}}\left(\frac{j(1+i)^{5}}{dz}\right)\left(\frac{s(z)(1+i)jy}{d}\right)\widetilde{e}\left(\frac{jy}{d}\right)\widetilde{e}\left(\frac{z}{(1+i)^{5}}\right)\\ =&\sum_{z\bmod{(1+i)^{5}}}\left(\frac{j(1+i)^{5}}{z}\right)\left(\frac{s(z)j}{d}\right)\widetilde{e}\left(\frac{z}{(1+i)^{5}}\right)\sum_{y\bmod{d}}\left(\frac{jy}{d}\right)\widetilde{e}\left(\frac{jy}{d}\right)\\ =&N(d)^{1/2}\sum_{z\bmod{(1+i)^{5}}}\left(\frac{j(1+i)^{5}}{z}\right)\left(\frac{s(z)ij}{d}\right)\widetilde{e}\left(\frac{z}{(1+i)^{5}}\right),\end{split}

where the last equality above follows from (2.4).

In order to evaluate the last sum in (2.8), we note that it suffices to take zz to vary over the reduced residue class modulo (1+i)5(1+i)^{5}. One representation of such class consists of the following 1616 elements (note that ±1,±i\pm 1,\pm i consists of the reduced residue class modulo (1+i)3(1+i)^{3} and 0,10,1 consists of the residue class modulo 1+i1+i):

{±1,±i}+l(1+i)3+k(1+i)4,l{0,1},k{0,1}.\displaystyle\{\pm 1,\pm i\}+l(1+i)^{3}+k(1+i)^{4},\quad l\in\{0,1\},k\in\{0,-1\}.

We further write d=a+bid=a+bi with a,ba,b\in\mathbb{Z} (recall that a1(mod4),b0(mod4)a\equiv 1\pmod{4},b\equiv 0\pmod{4} or a3(mod4),b2(mod4)a\equiv 3\pmod{4},b\equiv 2\pmod{4}) and check by direct calculations using (2.2) to see that (2.5) is valid with dd replaced by jdjd, where j=1j=1 or ii and dd is primary and square-free. This completes the proof of the lemma. ∎

Let φ[i](n)\varphi_{[i]}(n) denote the number of elements in the reduced residue class of 𝒪K/(n)\mathcal{O}_{K}/(n), we recall from [G&Zhao4, Lemma 2.2] the following explicitly evaluations of g(r,n)g(r,n) for primary nn.

Lemma 2.3.
  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,0,iflh+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}

2.4. The approximate functional equation

Let χ\chi be a primitive quadratic Hecke character modulo mm of trivial infinite type of KK. Let

(2.9) 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).

A well-known result of E. Hecke shows that L(s,χ)L(s,\chi) has an analytic continuation to the whole complex plane and satisfies the functional equation (see [iwakow, Theorem 3.8])

(2.10) 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 |W(χ)|=(N(m))1/2|W(\chi)|=(N(m))^{1/2}.

When we take χ=χ(1+i)5d\chi=\chi_{(1+i)^{5}d} for any odd, square-free d𝒪Kd\in\mathcal{O}_{K}, it follows from [iwakow, Theorem 3.8] that we have W(χ(1+i)5d)=g(χ(1+i)5d)W(\chi_{(1+i)^{5}d})=g(\chi_{(1+i)^{5}d}), so that Lemma 2.2 implies that in this case the functional equation becomes

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}).

For n𝒪Kn\in\mathcal{O}_{K} and rational integer k1k\geq 1, we let d[i],k(n)d_{[i],k}(n) denote the analogue on 𝒪K\mathcal{O}_{K} of the usual function dkd_{k} on \mathbb{Z}. Thus d[i],k(n)d_{[i],k}(n) equals the coefficient of N(n)sN(n)^{-s} in the Dirichlet series expansion of the kk-th power of ζK(s)\zeta_{K}(s). We shall also write d[i](n)d_{[i]}(n) for d[i],2(n)d_{[i],2}(n). In particular, when nn is primary, we have d[i],1(n)=1d_{[i],1}(n)=1 and

d[i](n)=d1mod(1+i)3d|n1.\displaystyle d_{[i]}(n)=\sum_{\begin{subarray}{c}d\equiv 1\bmod{(1+i)^{3}}\\ d|n\end{subarray}}1.

We denote also for rational integer j1j\geq 1 and any real number t>0t>0,

(2.11) Vj(t)=12πi(2)wj(s)tsdss,wj(s)=(25/2π)js(Γ(12+s)Γ(12))j.\displaystyle V_{j}(t)=\frac{1}{2\pi i}\int\limits\limits_{(2)}w_{j}(s)t^{-s}\frac{ds}{s},\quad w_{j}(s)=\left(\frac{2^{5/2}}{\pi}\right)^{js}\left(\frac{\Gamma(\frac{1}{2}+s)}{\Gamma(\frac{1}{2})}\right)^{j}.

We shall also write V(t),w(t)V(t),w(t) for V2(t),w2(t)V_{2}(t),w_{2}(t) respectively in the rest of the paper.

By setting j=1,2,α1=α2=0,Gj(s)=1j=1,2,\alpha_{1}=\alpha_{2}=0,G_{j}(s)=1 in [Gao1, Lemma 2.6], we obtain the following approximate functional equation for L(12,χ(1+i)5d)jL(\tfrac{1}{2},\chi_{(1+i)^{5}d})^{j}. Note here that the derivation of [Gao1, Lemma 2.6] assumes a rapid decay of Gj(s)G_{j}(s) but the proof also carries over to the case Gj(s)=1G_{j}(s)=1 due to the rapid decay of wj(s)w_{j}(s).

Lemma 2.5 (Approximate functional equation).

For any odd, square-free d𝒪Kd\in\mathcal{O}_{K}, we have for j=1,2j=1,2,

(2.12) L(12,χ(1+i)5d)j=2n1mod(1+i)3χ(1+i)5d(n)d[i],j(n)N(n)12Vj(N(n)N(d)j/2).\displaystyle\begin{split}L(\tfrac{1}{2},\chi_{(1+i)^{5}d})^{j}=&2\sum_{\begin{subarray}{c}n\equiv 1\bmod{(1+i)^{3}}\end{subarray}}\frac{\chi_{(1+i)^{5}d}(n)d_{[i],j}(n)}{N(n)^{\frac{1}{2}}}V_{j}\left(\frac{N(n)}{N(d)^{j/2}}\right).\end{split}

2.6. Poisson summation

In this section we gather some Poisson summation formulas over KK. We first recall that the Mellin transform f^\hat{f} for any function ff is defined to be

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

We now state a formula for smoothed character sums over all elements in 𝒪K\mathcal{O}_{K}.

Lemma 2.7.

Let n𝒪Kn\in\mathcal{O}_{K} and let χ\chi be a Hecke character (modn)\pmod{n} of trivial infinite type. For any smooth function W:+W:\mathbb{R}^{+}\rightarrow\mathbb{R} of compact support, we have for X>0X>0,

(2.13) m𝒪Kχ(m)W(N(m)X)=XN(n)k𝒪Kg(k,χ)W~(N(k)XNK(n)).\displaystyle\sum_{m\in\mathcal{O}_{K}}\chi(m)W\left(\frac{N(m)}{X}\right)=\frac{X}{N(n)}\sum_{k\in\mathcal{O}_{K}}g(k,\chi)\widetilde{W}\left(\sqrt{\frac{N(k)X}{N_{K}(n)}}\right).

The above is also valid when we replace χ\chi by (n)\left(\frac{\cdot}{n}\right) and g(k,χ)g(k,\chi) by g(k,n)g(k,n). Here g(k,χ),g(k,n)g(k,\chi),g(k,n) are defined in (2.3) and

(2.14) 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.

Moreover, the function W~(t)\widetilde{W}(t) is real-valued for all t0t\geq 0 and when t>0t>0, we have for cs=ε>0c_{s}=\varepsilon>0,

(2.15) W~(t)=\displaystyle\widetilde{W}(t)= π2πi(cs)W^(1s)(πt)2sΓ(s)Γ(1s)𝑑s.\displaystyle\frac{\pi}{2\pi i}\int\limits\limits_{(c_{s})}\widehat{W}(1-s)(\pi t)^{-2s}\frac{\Gamma(s)}{\Gamma(1-s)}ds.
Proof.

This lemma is essentially [G&Zhao4, Lemma 2.7] except for the last assertion. To establish it, we evaluate (2.14) using polar coordinates to see that

(2.16) W~(t)=2cos(2πty)W(x2+y2)dxdy=40π/20cos(2πtrsinθ)W(r2)rdrdθ=20π/20cos(2πtr1/2sinθ)W(r)drdθ.\displaystyle\widetilde{W}(t)=\int\limits_{\mathbb{R}^{2}}\cos(2\pi ty)W(x^{2}+y^{2})\ \mathrm{d}x\mathrm{d}y=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=2\int\limits^{\pi/2}_{0}\int\limits^{\infty}_{0}\cos(2\pi tr^{1/2}\sin\theta)W(r)\ \mathrm{d}r\mathrm{d}\theta.

The first equality above shows that W~(t)\widetilde{W}(t)\in\mathbb{R} for all t0t\geq 0.

We now apply inverse Mellin transform to write W(t)W(t) as

W(t)=12πi(cu)W^(u)tu𝑑u,\displaystyle W\left(t\right)=\frac{1}{2\pi i}\int\limits\limits_{(c_{u})}\widehat{W}(u)t^{-u}du,

where cu=ε>0c_{u}=\varepsilon>0. It follows from this and (2.16) that

W~(t)=\displaystyle\widetilde{W}(t)= 20π/20cos(2πtr1/2sinθ)12πi(cu)W^(1+u)rudrrdθ.\displaystyle 2\int\limits^{\pi/2}_{0}\int\limits^{\infty}_{0}\cos(2\pi tr^{1/2}\sin\theta)\frac{1}{2\pi i}\int\limits\limits_{(c_{u})}\widehat{W}(1+u)r^{-u}\frac{\mathrm{d}r}{r}\mathrm{d}\theta.

We make some changes of variables (first r1/2rr^{1/2}\to r, then 2πtrsinθr2\pi tr\sin\theta\to r) to see that

(2.17) W~(t)=42πi(cu)W^(1+u)0π/20cos(r)(r2πtsinθ)2udrrdθ𝑑u=42πi(cu)W^(1+u)(2πt)2u(0π/2(sinθ)2udθ0cos(r)r2udrr)𝑑u=π2πi(cu)W^(1+u)(πt)2uΓ(u)Γ(1+u)𝑑u,\displaystyle\begin{split}\widetilde{W}(t)=&\frac{4}{2\pi i}\int\limits\limits_{(c_{u})}\widehat{W}(1+u)\int\limits^{\pi/2}_{0}\int\limits^{\infty}_{0}\cos(r)\left(\frac{r}{2\pi t\sin\theta}\right)^{-2u}\frac{\mathrm{d}r}{r}\mathrm{d}\theta du\\ =&\frac{4}{2\pi i}\int\limits\limits_{(c_{u})}\widehat{W}(1+u)(2\pi t)^{2u}\left(\int\limits^{\pi/2}_{0}(\sin\theta)^{2u}\mathrm{d}\theta\int\limits^{\infty}_{0}\cos(r)r^{-2u}\frac{\mathrm{d}r}{r}\right)du\\ =&\frac{\pi}{2\pi i}\int\limits\limits_{(c_{u})}\widehat{W}(1+u)(\pi t)^{2u}\frac{\Gamma(-u)}{\Gamma(1+u)}du,\end{split}

where the last line above follows from the relation (see [Gao1, Section 2.4]) that

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)}.

By a further change of variable usu\rightarrow-s in the last integral of (2.17), we can recast W~(t)\widetilde{W}(t) as

(2.18) W~(t)=π2πi(cs)W^(1s)(πt)2sΓ(s)Γ(1s)𝑑s,\displaystyle\begin{split}\widetilde{W}(t)=&\frac{\pi}{2\pi i}\int\limits\limits_{(c_{s})}\widehat{W}(1-s)(\pi t)^{-2s}\frac{\Gamma(s)}{\Gamma(1-s)}ds,\end{split}

where we can retake cs=εc_{s}=\varepsilon as well and this completes the proof. ∎

We remark that when χ\chi is a primitive Hecke character, we have

g(r,χ)=χ¯(r)g(χ).\displaystyle g(r,\chi)=\overline{\chi}(r)g(\chi).

It follows from this and the expression given in (2.15) that the formula given (2.13) is equivalent to a version of the Poisson summation formula over number fields by L. Goldmakher and B. Louvel in [G&L, Lemma 3.2] for the case of the Gaussian field.

In the proof of Theorems 1.5 and 1.6, we need to consider a smoothed character sum over odd algebraic integers in 𝒪K\mathcal{O}_{K}. For this, we quote the following Poisson summation formula from [G&Zhao4, Corollary 2.8], which is a consequence of Lemma 2.7 above.

Lemma 2.8.

Let n𝒪Kn\in\mathcal{O}_{K} be primary and (n)\left(\frac{\cdot}{n}\right) be the quadratic residue symbol (modn)\pmod{n}. 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).

2.9. Analytical behaviors of certain Dirichlet series

In this section, we discuss the analytical behaviors of several Dirichlet series that are needed in the proof of Theorems 1.5 and 1.6. We first define for (α),(β)>12\Re(\alpha),\Re(\beta)>\frac{1}{2},

(2.19) Z(α,β)=n1,n21mod(1+i)3n1n2=d[i](n1)d[i](n2)N(n1)αN(n2)β𝒫(n1n2),\displaystyle Z(\alpha,\beta)=\sum_{\begin{subarray}{c}n_{1},n_{2}\equiv 1\bmod{(1+i)^{3}}\\ n_{1}n_{2}=\square\end{subarray}}\frac{d_{[i]}(n_{1})d_{[i]}(n_{2})}{N(n_{1})^{\alpha}N(n_{2})^{\beta}}\mathcal{P}(n_{1}n_{2}),

where we define 𝒫(n)\mathcal{P}(n) for primary nn by

(2.20) 𝒫(n)=ϖ1mod(1+i)3ϖ|n(N(ϖ)N(ϖ)+1).\displaystyle\mathcal{P}(n)=\prod_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ \varpi|n\end{subarray}}\left(\frac{N(\varpi)}{N(\varpi)+1}\right).

Our first result concerns the analytical behaviors of Z(α,β)Z(\alpha,\beta). The proof is similar to [Shen, Lemma 4.1], so we omit it here.

Lemma 2.10.

For (α),(β)>12\Re(\alpha),\Re(\beta)>\frac{1}{2}, we have

Z(α,β)=ζK3(2α)ζK3(2β)ζK4(α+β)Z1(α,β),\displaystyle Z(\alpha,\beta)=\zeta_{K}^{3}(2\alpha)\zeta_{K}^{3}(2\beta)\zeta_{K}^{4}(\alpha+\beta)Z_{1}(\alpha,\beta),

where

Z1(α,β)=(114α)3(114β)3(112α+β)4ϖ1mod(1+i)3ϖ|n1n2Z1,ϖ(α,β),\displaystyle Z_{1}(\alpha,\beta)=\left(1-\frac{1}{4^{\alpha}}\right)^{3}\left(1-\frac{1}{4^{\beta}}\right)^{3}\left(1-\frac{1}{2^{\alpha+\beta}}\right)^{4}\prod_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ \varpi|n_{1}n_{2}\end{subarray}}Z_{1,\varpi}(\alpha,\beta),

and where for primary ϖ\varpi,

Z1,ϖ(α,β)\displaystyle Z_{1,\varpi}(\alpha,\beta)
=\displaystyle= (11N(ϖ)2α)(11N(ϖ)2β)(11N(ϖ)α+β)4(1+4N(ϖ)α+β+1N(ϖ)2α+1N(ϖ)2β+1N(ϖ)2α+2β\displaystyle\left(1-\frac{1}{N(\varpi)^{2\alpha}}\right)\left(1-\frac{1}{N(\varpi)^{2\beta}}\right)\left(1-\frac{1}{N(\varpi)^{\alpha+\beta}}\right)^{4}\Bigg{(}1+\frac{4}{N(\varpi)^{\alpha+\beta}}+\frac{1}{N(\varpi)^{2\alpha}}+\frac{1}{N(\varpi)^{2\beta}}+\frac{1}{N(\varpi)^{2\alpha+2\beta}}
1N(ϖ)+1(3N(ϖ)2α+3N(ϖ)2β+4N(ϖ)α+β1N(ϖ)4α1N(ϖ)4β\displaystyle-\frac{1}{N(\varpi)+1}\Big{(}\frac{3}{N(\varpi)^{2\alpha}}+\frac{3}{N(\varpi)^{2\beta}}+\frac{4}{N(\varpi)^{\alpha+\beta}}-\frac{1}{N(\varpi)^{4\alpha}}-\frac{1}{N(\varpi)^{4\beta}}
3N(ϖ)2α+2β+2N(ϖ)2α+4β+2N(ϖ)4α+2β1N(ϖ)4α+4β)).\displaystyle-\frac{3}{N(\varpi)^{2\alpha+2\beta}}+\frac{2}{N(\varpi)^{2\alpha+4\beta}}+\frac{2}{N(\varpi)^{4\alpha+2\beta}}-\frac{1}{N(\varpi)^{4\alpha+4\beta}}\Big{)}\Bigg{)}.

Moreover, Z1(α,β)Z_{1}(\alpha,\beta) is analytic and uniformly bounded in the region (α),(β)14+ε\Re(\alpha),\Re(\beta)\geq\frac{1}{4}+\varepsilon.

Let g(k,n)g(k,n) be defined as in (2.3). We now fix a generator for every prime ideal (ϖ)𝒪K(\varpi)\in\mathcal{O}_{K} by taking ϖ\varpi to be primary if (ϖ,1+i)=1(\varpi,1+i)=1 and 1+i1+i for the ideal (1+i)(1+i) (noting that (1+i)(1+i) is the only prime ideal in 𝒪K\mathcal{O}_{K} that lies above the integral ideal (2)(2)\in\mathbb{Z}). We also fix 11 as the generator for the ring [i]\mathbb{Z}[i] itself and extend the choice of the generator for any ideal of 𝒪K\mathcal{O}_{K} multiplicatively. We denote the set of such generators by GG. For any k𝒪Kk\in\mathcal{O}_{K}, we shall hence denote k1,k2k_{1},k_{2} to be the unique pair of elements in 𝒪K\mathcal{O}_{K} such that k=k1k22k=k_{1}k^{2}_{2} with k1k_{1} being square-free and k2Gk_{2}\in G. In this way, we define

Z(α,β,a,k)\displaystyle Z(\alpha,\beta,a,k) =n1,n21mod(1+i)3(n1n2,a)=1d[i](n1)d[i](n2)N(n1)αN(n2)βg(k1k22,n1n2)N(n1n2).\displaystyle=\sum_{\begin{subarray}{c}n_{1},n_{2}\equiv 1\bmod{(1+i)^{3}}\\ (n_{1}n_{2},a)=1\end{subarray}}\frac{d_{[i]}(n_{1})d_{[i]}(n_{2})}{N(n_{1})^{\alpha}N(n_{2})^{\beta}}\frac{g(k_{1}k^{2}_{2},n_{1}n_{2})}{N(n_{1}n_{2})}.

Our next lemma gives the analytic properties of Z(α,β,a,k)Z(\alpha,\beta,a,k), we omit the proof here since it is similar to [Shen, Lemma 5.2].

Lemma 2.11.

For any k𝒪Kk\in\mathcal{O}_{K}, let k=k1k22k=k_{1}k_{2}^{2} with k1k_{1} square-free and k2Gk_{2}\in G. Then for (α),(β)>12\Re(\alpha),\Re(\beta)>\frac{1}{2}, we have

Z(α,β,a,k)=L2(12+α,χik1)L2(12+β,χik1)Z2(α,β,a,k),\displaystyle Z(\alpha,\beta,a,k)=L^{2}(\tfrac{1}{2}+\alpha,\chi_{ik_{1}})L^{2}(\tfrac{1}{2}+\beta,\chi_{ik_{1}})Z_{2}(\alpha,\beta,a,k),

where

Z2(α,β,a,k)=Z2,1+i(α,β,a,k)ϖ1mod(1+i)3Z2,ϖ(α,β,a,k).\displaystyle Z_{2}(\alpha,\beta,a,k)=Z_{2,1+i}(\alpha,\beta,a,k)\prod_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\end{subarray}}Z_{2,\varpi}(\alpha,\beta,a,k).

Here for ϖ|2a\varpi|2a,

Z2,ϖ(α,β,a,k)=(1χik1(ϖ)N(ϖ)12+α)2(1χik1(ϖ)N(ϖ)12+β)2,\displaystyle Z_{2,\varpi}(\alpha,\beta,a,k)=\left(1-\frac{\chi_{ik_{1}}(\varpi)}{N(\varpi)^{\frac{1}{2}+\alpha}}\right)^{2}\left(1-\frac{\chi_{ik_{1}}(\varpi)}{N(\varpi)^{\frac{1}{2}+\beta}}\right)^{2},

and for ϖ2a\varpi\nmid 2a,

Z2,ϖ(α,β,a,k)=(1χik1(ϖ)N(ϖ)12+α)2(1χik1(ϖ)N(ϖ)12+β)2n1=0n2=0d[i](ϖn1)d[i](ϖn2)N(ϖ)n1α+n2βg(k,ϖn1+n2)N(ϖ)n1+n2.\displaystyle Z_{2,\varpi}(\alpha,\beta,a,k)=\left(1-\frac{\chi_{ik_{1}}(\varpi)}{N(\varpi)^{\frac{1}{2}+\alpha}}\right)^{2}\left(1-\frac{\chi_{ik_{1}}(\varpi)}{N(\varpi)^{\frac{1}{2}+\beta}}\right)^{2}\sum_{n_{1}=0}^{\infty}\sum_{n_{2}=0}^{\infty}\frac{d_{[i]}(\varpi^{n_{1}})d_{[i]}(\varpi^{n_{2}})}{N(\varpi)^{n_{1}\alpha+n_{2}\beta}}\frac{g(k,\varpi^{n_{1}+n_{2}})}{N(\varpi)^{n_{1}+n_{2}}}.

Moreover, Z2(α,β,a,k)Z_{2}(\alpha,\beta,a,k) is analytic in the region (α),(β)>0\Re(\alpha),\Re(\beta)>0 and for (α),(β)1logX\Re(\alpha),\Re(\beta)\geq\frac{1}{\log X}, we have

(2.21) Z2(α,β,a,k)d[i]4(a)d[i]12(k)(logX)10,\displaystyle Z_{2}(\alpha,\beta,a,k)\ll d^{4}_{[i]}(a)d^{12}_{[i]}(k)(\log X)^{10},

where the implied constant is absolute.

We define for a prime ω𝒪K\omega\in\mathcal{O}_{K}, α,β,γ\alpha,\beta,\gamma\in\mathbb{C},

K1(α,β,γ;ϖ)=\displaystyle K_{1}(\alpha,\beta,\gamma;\varpi)= (11N(ϖ)12+α)2(11N(ϖ)12+β)2(11N(ϖ)2α+2γ)2(11N(ϖ)2β+2γ)2,\displaystyle\left(1-\frac{1}{N(\varpi)^{\frac{1}{2}+\alpha}}\right)^{2}\left(1-\frac{1}{N(\varpi)^{\frac{1}{2}+\beta}}\right)^{2}\left(1-\frac{1}{N(\varpi)^{2\alpha+2\gamma}}\right)^{2}\left(1-\frac{1}{N(\varpi)^{2\beta+2\gamma}}\right)^{2},
K2(α,β,γ;ϖ)=(11N(ϖ)12+α)2(11N(ϖ)12+β)2((11N(ϖ))(1+1N(ϖ)2α+2γ)(1+1N(ϖ)2β+2γ)+1N(ϖ)(11N(ϖ)2α+2γ)2(11N(ϖ)2β+2γ)2+(11N(ϖ))4N(ϖ)α+β+2γ+2(11N(ϖ)2γ)(1N(ϖ)12+α+1N(ϖ)12+β+1N(ϖ)12+2α+β+2γ+1N(ϖ)12+α+2β+2γ)).\displaystyle\begin{split}K_{2}(\alpha,\beta,\gamma;\varpi)=&\left(1-\frac{1}{N(\varpi)^{\frac{1}{2}+\alpha}}\right)^{2}\left(1-\frac{1}{N(\varpi)^{\frac{1}{2}+\beta}}\right)^{2}\Bigg{(}\left(1-\frac{1}{N(\varpi)}\right)\left(1+\frac{1}{N(\varpi)^{2\alpha+2\gamma}}\right)\left(1+\frac{1}{N(\varpi)^{2\beta+2\gamma}}\right)\\ &+\frac{1}{N(\varpi)}\left(1-\frac{1}{N(\varpi)^{2\alpha+2\gamma}}\right)^{2}\left(1-\frac{1}{N(\varpi)^{2\beta+2\gamma}}\right)^{2}+\left(1-\frac{1}{N(\varpi)}\right)\frac{4}{N(\varpi)^{\alpha+\beta+2\gamma}}\\ &+2\left(1-\frac{1}{N(\varpi)^{2\gamma}}\right)\left(\frac{1}{N(\varpi)^{\frac{1}{2}+\alpha}}+\frac{1}{N(\varpi)^{\frac{1}{2}+\beta}}+\frac{1}{N(\varpi)^{\frac{1}{2}+2\alpha+\beta+2\gamma}}+\frac{1}{N(\varpi)^{\frac{1}{2}+\alpha+2\beta+2\gamma}}\right)\Bigg{)}.\end{split}

With the above notations, we note the following result concerning a Dirichlet series related to Z2(α,β,a,k)Z_{2}(\alpha,\beta,a,k).

Lemma 2.12.

For (α),(β)>0,(γ)>12\Re(\alpha),\Re(\beta)>0,\Re(\gamma)>\tfrac{1}{2}, we have

(2.22) kG(1)N(k)N(k)2γZ2(α,β,a,±ik2)=(212γ1)ζK(2γ)ζK2(2α+2γ)ζK2(2β+2γ)Z3(α,β,γ,a),\displaystyle\sum_{k\in G}\frac{(-1)^{N(k)}}{N(k)^{2\gamma}}Z_{2}(\alpha,\beta,a,\pm ik^{2})=(2^{1-2\gamma}-1)\zeta_{K}(2\gamma)\zeta_{K}^{2}(2\alpha+2\gamma)\zeta_{K}^{2}(2\beta+2\gamma)Z_{3}(\alpha,\beta,\gamma,a),

where

Z3(α,β,γ,a)=ϖ1mod(1+i)3Z3,ϖ(α,β,γ,a),\displaystyle Z_{3}(\alpha,\beta,\gamma,a)=\prod_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\end{subarray}}Z_{3,\varpi}(\alpha,\beta,\gamma,a),

Here for ϖ|2a\varpi|2a, Z3,ϖ(α,β,γ,a)=K1(α,β,γ;ϖ)Z_{3,\varpi}(\alpha,\beta,\gamma,a)=K_{1}(\alpha,\beta,\gamma;\varpi) and for ϖ2a\varpi\nmid 2a, Z3,ϖ(α,β,γ,a)=K2(α,β,γ;ϖ)Z_{3,\varpi}(\alpha,\beta,\gamma,a)=K_{2}(\alpha,\beta,\gamma;\varpi).

In addition, we have

  • (1)

    Z3(α,β,γ,a)Z_{3}(\alpha,\beta,\gamma,a) is analytic and uniformly bounded for (α),(β)12+ε,(γ)2ε\Re(\alpha),\Re(\beta)\geq\tfrac{1}{2}+\varepsilon,\Re(\gamma)\geq 2\varepsilon.

  • (2)

    Z3(α,β,γ,a)(logX)10Z_{3}(\alpha,\beta,\gamma,a)\ll(\log X)^{10} for (α),(β)12+1logX,(γ)2logX\Re(\alpha),\Re(\beta)\geq\tfrac{1}{2}+\frac{1}{\log X},\Re(\gamma)\geq\frac{2}{\log X} with the implied constant being absolute.

Proof.

We let f(k)=g(k,n)/N(k)sf(k)=g(k,n)/N(k)^{s} and we divide the sum over kk in (2.22) into two sums, according to (k,1+i)=1(k,1+i)=1 or not, to see that

kG(1)N(k)f(±ik2)=(2kGf(±2ik2)kGf(±ik2)),\displaystyle\sum_{k\in G}(-1)^{N(k)}f(\pm ik^{2})=\left(2\sum_{k\in G}f(\pm 2ik^{2})-\sum_{k\in G}f(\pm ik^{2})\right),

Note that when (n,1+i)=1(n,1+i)=1, g(k,n)=g(2k,n)g(k,n)=g(2k,n) by Lemma 2.3. It follows that we have f(±2ik2)=4sf(±ik2)f(\pm 2ik^{2})=4^{-s}f(\pm ik^{2}) so that

kG(1)N(k)f(±ik2)=(212s1)kGf(±ik2).\displaystyle\sum_{k\in G}(-1)^{N(k)}f(\pm ik^{2})=(2^{1-2s}-1)\sum_{k\in G}f(\pm ik^{2}).

We then deduce from this and the definition of Z2Z_{2} given in Lemma 2.11 that

kG(1)N(k)N(k)2γZ2(α,β,a,±ik2)=\displaystyle\sum_{k\in G}\frac{(-1)^{N(k)}}{N(k)^{2\gamma}}Z_{2}(\alpha,\beta,a,\pm ik^{2})= (212γ1)kG1N(k)2γZ2(α,β,a,k2)=ϖGb=0Z2,ϖ(α,β,a,ϖ2b)N(ϖ)2bγ,\displaystyle(2^{1-2\gamma}-1)\sum_{k\in G}\frac{1}{N(k)^{2\gamma}}Z_{2}(\alpha,\beta,a,k^{2})=\prod_{\begin{subarray}{c}\varpi\in G\end{subarray}}\sum_{b=0}^{\infty}\frac{Z_{2,\varpi}(\alpha,\beta,a,\varpi^{2b})}{N(\varpi)^{2b\gamma}},

where the last equality above follows from the observation that g(±ik2,n)g(\pm ik^{2},n) is multiplicative with respect to kk and that we have g(±ik2,n)=g(k2,n)g(\pm ik^{2},n)=g(k^{2},n) when nn is primary from Lemma 2.3. The assertions of Lemma 2.12 now follows by arguing similarly to the proof of [Shen, Lemma 5.3]. ∎

Lastly, we note the following result which can be established similar to the proof of [Shen, Lemma 5.4].

Lemma 2.13.

For (α),(β)>12,0<(γ)<12\Re(\alpha),\Re(\beta)>\tfrac{1}{2},0<\Re(\gamma)<\tfrac{1}{2}, we have

a1mod(1+i)3μ[i](a)N(a)22γZ3(α,β,γ,a)=ζK(2α+2γ)ζK(2β+2γ)ζK4(α+β+2γ)ζK2(12+α+2γ)ζK2(12+β+2γ)Z4(α,β,γ),\displaystyle\sum_{\begin{subarray}{c}a\equiv 1\bmod{(1+i)^{3}}\end{subarray}}\frac{\mu_{[i]}(a)}{N(a)^{2-2\gamma}}Z_{3}(\alpha,\beta,\gamma,a)=\frac{\zeta_{K}(2\alpha+2\gamma)\zeta_{K}(2\beta+2\gamma)\zeta_{K}^{4}(\alpha+\beta+2\gamma)}{\zeta_{K}^{2}(\frac{1}{2}+\alpha+2\gamma)\zeta_{K}^{2}(\frac{1}{2}+\beta+2\gamma)}Z_{4}(\alpha,\beta,\gamma),

where

Z4(α,β,γ)=\displaystyle Z_{4}(\alpha,\beta,\gamma)= K1(α,β,γ;1+i)ϖ1mod(1+i)3(K2(α,β,γ;ϖ)1N(ϖ)22γK1(α,β,γ;ϖ))\displaystyle K_{1}(\alpha,\beta,\gamma;1+i)\prod_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\end{subarray}}\left(K_{2}(\alpha,\beta,\gamma;\varpi)-\frac{1}{N(\varpi)^{2-2\gamma}}K_{1}(\alpha,\beta,\gamma;\varpi)\right)
×ϖ1mod(1+i)3(11N(ϖ)2α+2γ)(11N(ϖ)2β+2γ)(11N(ϖ)α+β+2γ)4(11N(ϖ)12+α+2γ)2(11N(ϖ)12+β+2γ)2.\displaystyle\times\prod_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\end{subarray}}\frac{(1-\frac{1}{N(\varpi)^{2\alpha+2\gamma}})(1-\frac{1}{N(\varpi)^{2\beta+2\gamma}})(1-\frac{1}{N(\varpi)^{\alpha+\beta+2\gamma}})^{4}}{(1-\frac{1}{N(\varpi)^{\frac{1}{2}+\alpha+2\gamma}})^{2}(1-\frac{1}{N(\varpi)^{\frac{1}{2}+\beta+2\gamma}})^{2}}.

Moreover, Z4(α,β,γ)Z_{4}(\alpha,\beta,\gamma) is analytic and uniformly bounded for (α),(β)38,116(γ)18\Re(\alpha),\Re(\beta)\geq\frac{3}{8},-\frac{1}{16}\leq\Re(\gamma)\leq\frac{1}{8}.

2.14. A mean value estimate for quadratic Hecke LL-functions

In [DRHB, Theorem 1], D. R. Heath-Brown established a powerful quadratic large sieve result for Dirichlet characters. Such result was extended by K. Onodera in [Onodera] to quadratic residue symbols in the Gaussian field. Applying Onodera’s result in a similar fashion as in the derivation of [DRHB, Theorem 2] by Heath-Brown to obtain a mean value estimation for the fourth moment of the family of primitive quadratic Dirichlet LL-functions, we have the following upper bound for the fourth moment of quadratic Hecke LL-functions unconditionally.

Lemma 2.15.

Suppose σ+it\sigma+it is a complex number with σ12\sigma\geq\frac{1}{2}. Then

(d,2)=1N(d)X|L(σ+it,χ(1+i)5d)|4X1+ε(1+|t|2)1+ε.\displaystyle\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ N(d)\leq X\end{subarray}}|L(\sigma+it,\chi_{(1+i)^{5}d})|^{4}\ll X^{1+\varepsilon}(1+|t|^{2})^{1+\varepsilon}.

3. Proof of Theorem 1.1

We recall a result of Landau [Landau] implies that for an algebraic number field FF of degree n=2n=2 and any primitive ideal character χ\chi of FF with conductor qq, we have for X>1X>1,

(3.1) NF(I)Xχ(I)|NF(q)DF|1/3log2(|NF(q)DF|)X1/3,\displaystyle\sum_{N_{F}(I)\leq X}\chi(I)\ll|N_{F}(q)\cdot D_{F}|^{1/3}\log^{2}(|N_{F}(q)\cdot D_{F}|)X^{1/3},

where NF(q),NF(I)N_{F}(q),N_{F}(I) denotes the norm of qq and II respectively, DFD_{F} denotes the discriminant of FF and II runs over integral ideas of FF.

We deduce from (3.1) by partial summation that for odd, square-free dd, the series

n1mod(1+i)3χ(1+i)5d(n)N(n)\displaystyle\sum_{\begin{subarray}{c}n\equiv 1\bmod{(1+i)^{3}}\end{subarray}}\frac{\chi_{(1+i)^{5}d}(n)}{\sqrt{N(n)}}

is convergent and equals to L(12,χ(1+i)5d)L(\tfrac{1}{2},\chi_{(1+i)^{5}d}). In particular, this implies that L(12,χ(1+i)5d)L(\tfrac{1}{2},\chi_{(1+i)^{5}d})\in\mathbb{R}. It follows from this and the observation that kk is an even natural number that we may further restrict the sum over dd in (1.5) to satisfy X/2<N(d)XX/2<N(d)\leq X.

We set x=X110kx=X^{\frac{1}{10k}} and apply Hölder’s inequality to see that

(3.2) (d,2)=1X/2<N(d)XL(12,χ(1+i)5d)k𝒮1k𝒮2k1,\displaystyle\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ X/2<N(d)\leq X\end{subarray}}L(\tfrac{1}{2},\chi_{(1+i)^{5}d})^{k}\geq\frac{\mathcal{S}_{1}^{k}}{\mathcal{S}_{2}^{k-1}},

where

𝒮1=(d,2)=1X/2<N(d)XL(12,χ(1+i)5d)A(d)k1,𝒮2=(d,2)=1X/2<N(d)XA(d)k.\displaystyle\mathcal{S}_{1}=\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ X/2<N(d)\leq X\end{subarray}}L(\tfrac{1}{2},\chi_{(1+i)^{5}d})A({d})^{k-1},\ \ \ \ \mathcal{S}_{2}=\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ X/2<N(d)\leq X\end{subarray}}A({d})^{k}.

and

A(d)=n1mod(1+i)3N(n)xχ(1+i)5d(n)N(n).A(d)=\sum_{\begin{subarray}{c}n\equiv 1\bmod{(1+i)^{3}}\\ N(n)\leq x\end{subarray}}\frac{\chi_{(1+i)^{5}d}(n)}{\sqrt{N(n)}}.

In the remaining of the proof, it thus suffices to bound 𝒮1\mathcal{S}_{1} and 𝒮2\mathcal{S}_{2}. We bound 𝒮2\mathcal{S}_{2} first by noting that

𝒮2(d,2)=1A(d)kW(N(d)X),\displaystyle\mathcal{S}_{2}\ll\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\end{subarray}}A({d})^{k}W(\frac{N(d)}{X}),

where W(t)W(t) is any non-negative smooth function that is supported on (12ε,1+ε)(\frac{1}{2}-\varepsilon,1+\varepsilon) for some fixed small 0<ε<1/20<\varepsilon<1/2 such that W(t)1W(t)\gg 1 for t(12,1)t\in(\tfrac{1}{2},1).

We now expand A(d)kA(d)^{k} to see that

𝒮2=n1,,nk1mod(1+i)3N(n1),,N(nk)x1N(n1nk)(d,2)=1((1+i)5dn1nk)W(N(d)X).\mathcal{S}_{2}=\sum_{\begin{subarray}{c}n_{1},\ldots,n_{k}\equiv 1\bmod{(1+i)^{3}}\\ N(n_{1}),\ldots,N(n_{k})\leq x\end{subarray}}\frac{1}{\sqrt{N(n_{1}\cdots n_{k})}}\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\end{subarray}}\left(\frac{(1+i)^{5}d}{n_{1}\cdots n_{k}}\right)W(\frac{N(d)}{X}).

We consider the inner sum above by setting n=n1nkn=n_{1}\cdots n_{k}. Using Möbius function to express the condition that dd is square-free, we see that

(3.3) (d,2)=1((1+i)5dn)W(N(d)X)=α1mod(1+i)3N(α)2X1/2((1+i)5α2n)(d,2)=1(dn)W(N(α2d)X).\displaystyle\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\end{subarray}}\left(\frac{(1+i)^{5}d}{n}\right)W(\frac{N(d)}{X})=\sum_{\begin{subarray}{c}\alpha\equiv 1\bmod(1+i)^{3}\\ N(\alpha)\leq 2X^{1/2}\end{subarray}}\left(\frac{(1+i)^{5}\alpha^{2}}{n}\right)\sum_{\begin{subarray}{c}(d,2)=1\end{subarray}}\left(\frac{d}{n}\right)W(\frac{N(\alpha^{2}d)}{X}).

Note that for smoothed sums involving any non-principal Hecke character modulo aa of trivial infinite type, we have (see [G&Zhao2019, (1.4)]) that for y1y\geq 1

(3.4) c1mod(1+i)3χ(c)Φ(N(c)y)εN(a)(1+ε)/2.\displaystyle\sum_{c\equiv 1\bmod{(1+i)^{3}}}\chi(c)\Phi\left(\frac{N(c)}{y}\right)\ll_{\varepsilon}N(a)^{(1+\varepsilon)/2}.

Now, we write d=jdd=jd^{\prime} with jUKj\in U_{K} and dd^{\prime} being primary and apply the quadratic reciprocity law (2.1) to see that

(3.5) (dn)=(jn)(nd).\displaystyle\left(\frac{d}{n}\right)=\left(\frac{j}{n}\right)\left(\frac{n}{d^{\prime}}\right).

Note that if nn is not a square then the symbol (n)\left(\frac{n}{\cdot}\right) can be regarded as a non-principal Hecke character (mod16n)\pmod{16n} of trivial infinite type. By decomposing the sum over dd in (3.3) into sums over j,dj,d^{\prime} and apply (3.4) to the sum over dd^{\prime}, we deduce that the sum over dd in (3.3) is N(n)1/2+ε\ll N(n)^{1/2+\varepsilon}. On the other hand, the sum over dd is trivially X/N(α)2\ll X/N(\alpha)^{2}. We then conclude that if nn is not a square,

(3.6) (d,2)=1((1+i)5dn)W(N(α2d)X)α1mod(1+i)3N(α)2X1/2((1+i)5α2n)min(N(n)1/2+ε,XN(α)2)X12N(n)14+ε.\displaystyle\sum_{\begin{subarray}{c}(d,2)=1\end{subarray}}\left(\frac{(1+i)^{5}d}{n}\right)W(\frac{N(\alpha^{2}d)}{X})\ll\sum_{\begin{subarray}{c}\alpha\equiv 1\bmod(1+i)^{3}\\ N(\alpha)\leq 2X^{1/2}\end{subarray}}\left(\frac{(1+i)^{5}\alpha^{2}}{n}\right)\min\Big{(}N(n)^{1/2+\varepsilon},\frac{X}{N(\alpha)^{2}}\Big{)}\ll X^{\frac{1}{2}}N(n)^{\frac{1}{4}+\varepsilon}.

If nn is a perfect square, then by applying the following result for the Gauss circle problem (see [Huxley1]) with θ=131/416\theta=131/416,

(3.7) N(a)x1=πx+O(xθ)\displaystyle\sum_{N(a)\leq x}1=\pi x+O(x^{\theta})

together with a routine argument, we see that

(3.8) (d,2)=1((1+i)5dn)W(N(d)X)N(d)X(d,2n)=11=2πX3ζK(2)𝒫(n)+O(X12+εN(n)ε),\displaystyle\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\end{subarray}}\left(\frac{(1+i)^{5}d}{n}\right)W(\frac{N(d)}{X})\ll\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}N(d)\leq X\\ (d,2n)=1\end{subarray}}1=\frac{2\pi X}{3\zeta_{K}(2)}\mathcal{P}(n)+O(X^{\frac{1}{2}+\varepsilon}N(n)^{\varepsilon}),

where 𝒫(n)\mathcal{P}(n) is defined in (2.20).

Using (3.6) and (3.8), we see that

(3.9) 𝒮2Xn1,,nk1mod(1+i)3N(n1),,N(nk)xn1nk=n2𝒫(n)N(n)+O(X12+ϵxk(34+ϵ)).\displaystyle\mathcal{S}_{2}\ll X\sum_{\begin{subarray}{c}n_{1},\ldots,n_{k}\equiv 1\bmod{(1+i)^{3}}\\ N(n_{1}),\ldots,N(n_{k})\leq x\\ n_{1}\cdots n_{k}=n^{2}\end{subarray}}\frac{\mathcal{P}(n)}{N(n)}+O\Big{(}X^{\frac{1}{2}+\epsilon}x^{k(\frac{3}{4}+\epsilon)}\Big{)}.

We note that

(3.10) n1mod(1+i)3N(n)2xd[i],k(n2)N(n)𝒫(n)n1,,nk1mod(1+i)3N(n1),,N(nk)xn1nk=n2𝒫(n)N(n)n1mod(1+i)3N(n)2xkd[i],k(n2)N(n)𝒫(n).\displaystyle\sum_{\begin{subarray}{c}n\equiv 1\bmod{(1+i)^{3}}\\ N(n)^{2}\leq x\end{subarray}}\frac{d_{[i],k}(n^{2})}{N(n)}\mathcal{P}(n)\ll\sum_{\begin{subarray}{c}n_{1},\ldots,n_{k}\equiv 1\bmod{(1+i)^{3}}\\ N(n_{1}),\ldots,N(n_{k})\leq x\\ n_{1}\cdots n_{k}=n^{2}\end{subarray}}\frac{\mathcal{P}(n)}{N(n)}\ll\sum_{\begin{subarray}{c}n\equiv 1\bmod{(1+i)^{3}}\\ N(n)^{2}\leq x^{k}\end{subarray}}\frac{d_{[i],k}(n^{2})}{N(n)}\mathcal{P}(n).

Similar to [Selberg, Theorem 2], we have that for a positive constant C(k)C(k),

(3.11) n1mod(1+i)3N(n)zd[i],k(n2)N(n)𝒫(n)C(k)(logz)k(k+1)/2.\displaystyle\sum_{\begin{subarray}{c}n\equiv 1\bmod{(1+i)^{3}}\\ N(n)\leq z\end{subarray}}\frac{d_{[i],k}(n^{2})}{N(n)}\mathcal{P}(n)\sim C(k)(\log z)^{k(k+1)/2}.

Applying this in (3.10) and setting x=X110kx=X^{\frac{1}{10k}} in (3.9), we deduce that

(3.12) 𝒮2X(logX)k(k+1)/2.\displaystyle\mathcal{S}_{2}\ll X(\log X)^{k(k+1)/2}.

We evaluate 𝒮1\mathcal{S}_{1} next. By applying the approximate functional equation (2.12) for L(12,χ(1+i)5d)L(\tfrac{1}{2},\chi_{(1+i)^{5}d}), we see that

(3.13) 𝒮1=2n1mod(1+i)31N(n)n1,,nk1mod(1+i)3N(n1),,N(nk1)x1N(n1nk1)(d,2)=1X/2<N(d)X((1+i)5dnn1nk1)V1(N(n)N(d)).\displaystyle\begin{split}\mathcal{S}_{1}=2\sum_{\begin{subarray}{c}n\equiv 1\bmod{(1+i)^{3}}\end{subarray}}\frac{1}{\sqrt{N(n)}}\sum_{\begin{subarray}{c}n_{1},\ldots,n_{k}\equiv 1\bmod{(1+i)^{3}}\\ N(n_{1}),\ldots,N(n_{k-1})\leq x\end{subarray}}\frac{1}{\sqrt{N(n_{1}\cdots n_{k-1})}}\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ X/2<N(d)\leq X\end{subarray}}\left(\frac{(1+i)^{5}d}{nn_{1}\cdots n_{k-1}}\right)V_{1}\Big{(}\frac{N(n)}{\sqrt{N(d)}}\Big{)}.\end{split}

Here we note that (see [sound1, Lemma 2.1]) V1(x)V_{1}(x) is real-valued and smooth on [0,)[0,\infty) and the jj-th derivative of V1(x)V_{1}(x) satisfies

(3.14) V1(x)=1+O(ξ1/2ϵ)for 0<ξ<1andV1(j)(ξ)=O(eξ)forξ>0,j0.\displaystyle V_{1}\left(x\right)=1+O(\xi^{1/2-\epsilon})\;\mbox{for}\;0<\xi<1\quad\mbox{and}\quad V^{(j)}_{1}\left(\xi\right)=O(e^{-\xi})\;\mbox{for}\;\xi>0,\;j\geq 0.

If nn1nk1nn_{1}\cdots n_{k-1} is not a square, then using (3.4), (3.14) and partial summation we see that

(3.15) (d,2)=1X/2<N(d)X((1+i)5dn1nk1)V1(N(n)N(d))X12N(nn1nk1)14+ϵeN(n)/X.\displaystyle\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ X/2<N(d)\leq X\end{subarray}}\left(\frac{(1+i)^{5}d}{n_{1}\cdots n_{k-1}}\right)V_{1}\Big{(}\frac{N(n)}{\sqrt{N(d)}}\Big{)}\ll X^{\frac{1}{2}}N(nn_{1}\cdots n_{k-1})^{\frac{1}{4}+\epsilon}e^{-N(n)/\sqrt{X}}.

If nn1nk1nn_{1}\cdots n_{k-1} is an square, then using the right-hand side expression in (3.8) together with (3.14) and partial summation implies that the sum over dd in (3.13) is

(3.16) (d,2)=1X/2<N(d)X((1+i)5dn1nk1)V1(N(n)N(d))=X3ζK(2)𝒫(nn1nk1)12V1(2N(n)Xt)𝑑t+O(X12+ϵN(nn1nk1)εeN(n)/X).\displaystyle\begin{split}&\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ X/2<N(d)\leq X\end{subarray}}\left(\frac{(1+i)^{5}d}{n_{1}\cdots n_{k-1}}\right)V_{1}\Big{(}\frac{N(n)}{\sqrt{N(d)}}\Big{)}\\ =&\frac{X}{3\zeta_{K}(2)}\mathcal{P}(nn_{1}\cdots n_{k-1})\int_{1}^{2}V_{1}\Big{(}\frac{\sqrt{2}N(n)}{\sqrt{Xt}}\Big{)}dt+O(X^{\frac{1}{2}+\epsilon}N(nn_{1}\cdots n_{k-1})^{\varepsilon}e^{-N(n)/\sqrt{X}}).\end{split}

Applying (3.15) and (3.16) in (3.13), we see that the error terms in (3.15) and (3.16) contribute to (3.13) an amount X12+ϵx(k1)(34+ϵ)X38+ϵX3940\ll X^{\frac{1}{2}+\epsilon}x^{(k-1)(\frac{3}{4}+\epsilon)}X^{\frac{3}{8}+\epsilon}\ll X^{\frac{39}{40}}.

To estimate contribution of the main term in (3.16) to (3.13), we write n1nk1n_{1}\cdots n_{k-1} as rs2rs^{2} where rr and ss are primary and rr is square-free. Then nn must be of the form r2r\ell^{2} where \ell is primary. It follows that the contribution of the main term in (3.16) to (3.13) is

2X3ζK(2)r,s1mod(1+i)3rs2=n1nk1N(n1),,N(nk1)x1N(rs)1mod(1+i)31N()12𝒫(rs)V1(2N(r2)Xt)𝑑t.\frac{2X}{3\zeta_{K}(2)}\sum_{\begin{subarray}{c}r,s\equiv 1\bmod{(1+i)^{3}}\\ rs^{2}=n_{1}\cdots n_{k-1}\\ N(n_{1}),\ldots,N(n_{k-1})\leq x\end{subarray}}\frac{1}{N(rs)}\sum_{\begin{subarray}{c}\ell\equiv 1\bmod{(1+i)^{3}}\end{subarray}}\frac{1}{N(\ell)}\int_{1}^{2}\mathcal{P}(rs\ell)V_{1}\Big{(}\frac{\sqrt{2}N(r\ell^{2})}{\sqrt{Xt}}\Big{)}dt.

Note that N(r)xk1<X110N(r)\leq x^{k-1}<X^{\frac{1}{10}}, and by a standard argument, we see that the sum over \ell above is

=23𝒫(rs)ϖ1mod(1+i)3ϖrs(11N(ϖ)(N(ϖ+1))14π4logXN(r)+O(1).=\frac{2}{3}\mathcal{P}(rs)\prod_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ \varpi\nmid rs\end{subarray}}\Big{(}1-\frac{1}{N(\varpi)(N(\varpi+1)}\Big{)}\frac{1}{4}\cdot\frac{\pi}{4}\log\frac{\sqrt{X}}{N(r)}+O(1).

It follows that the contribution of the main term in (3.16) to (3.13) is

XlogXr,s1mod(1+i)3rs2=n1nk1N(n1),,N(nk1)x1N(rs)𝒫(rs)\displaystyle\gg X\log X\sum_{\begin{subarray}{c}r,s\equiv 1\bmod{(1+i)^{3}}\\ rs^{2}=n_{1}\cdots n_{k-1}\\ N(n_{1}),\ldots,N(n_{k-1})\leq x\end{subarray}}\frac{1}{N(rs)}\mathcal{P}(rs)
XlogXr primary and square-frees primaryN(rs2)xd[i],k1(rs2)N(rs)𝒫(rs)X(logX)(logx)k1+k(k1)/2.\displaystyle\gg X\log X\sum_{\begin{subarray}{c}r\text{ primary and square-free}\\ s\text{ primary}\\ N(rs^{2})\leq x\end{subarray}}\frac{d_{[i],k-1}(rs^{2})}{N(rs)}\mathcal{P}(rs)\gg X(\log X)(\log x)^{k-1+k(k-1)/2}.

We then deduce that

𝒮1X(logX)k(k+1)/2.\mathcal{S}_{1}\gg X(\log X)^{k(k+1)/2}.

The assertion of Theorem 1.1 now follows from (3.2), (3.12) and the above bound.

4. Proof of Theorem 1.2

4.1. A few lemmas

We first include a few lemmas needed in the proof of Theorem 1.2. We denote Λ[i](n)\Lambda_{[i]}(n) for the von Mangoldt function on KK. Thus Λ[i](n)\Lambda_{[i]}(n) equals the coefficient of N(n)sN(n)^{-s} in the Dirichlet series expansion of ζK(s)/ζK(s)\zeta^{{}^{\prime}}_{K}(s)/\zeta_{K}(s). Our first lemma provides an upper bound of log|L(s,χ)|\log|L(s,\chi)| in terms of a sum involving prime powers.

Lemma 4.2.

Let χ\chi be a non-principal primitive quadratic Hecke character modulo mm of trivial infinite type. Assume GRH for ζK(s)\zeta_{K}(s) and L(s,χ)L(s,\chi). Let TT be a large number and x2x\geq 2. Let λ0=0.56\lambda_{0}=0.56\ldots denote the unique positive real number satisfying eλ0=λ0e^{-\lambda_{0}}=\lambda_{0}. For all λ0λ5\lambda_{0}\leq\lambda\leq 5 we have uniformly for |t|T|t|\leq T and 12σ12+λlogx\tfrac{1}{2}\leq\sigma\leq\tfrac{1}{2}+\frac{\lambda}{\log x} that

(4.1) log|L(σ+it,χ)|(n)N(n)xΛ[i](n)N(n)12+λlogx+itlogN(n)log(x/N(n))logx+(logT+logN(m)2)(12σ+1+λlogx)+O(1logx),\displaystyle\log|L(\sigma+it,\chi)|\leq\Re{\sum_{\begin{subarray}{c}(n)\\ N(n)\leq x\end{subarray}}\frac{\Lambda_{[i]}(n)}{N(n)^{\frac{1}{2}+\frac{\lambda}{\log x}+it}\log N(n)}\frac{\log(x/N(n))}{\log x}}+(\log T+\frac{\log N(m)}{2})(\tfrac{1}{2}-\sigma+\frac{1+\lambda}{\log x})+O\Big{(}\frac{1}{\log x}\Big{)},

where the sum (n)\displaystyle\sum_{(n)} means that the sum is over integral ideals of 𝒪K\mathcal{O}_{K}.

Proof.

We denote ss for σ+it\sigma+it and we interpret log|L(s,χ)|\log|L(s,\chi)| as -\infty when L(s,χ)=0L(s,\chi)=0. Thus we may suppose L(σ+it,χ)0L(\sigma+it,\chi)\neq 0 in the rest of the proof. Recall the associated function Λ(s,χ)\Lambda(s,\chi) defined as in (2.9). Here Λ(s,χ)\Lambda(s,\chi) is analytical in the entire complex plane since χ\chi is non-principal. As Γ(s)\Gamma(s) has simple poles at the non-positive rational integers (see [Da, §10]), we see from the expression of Λ(s,χ)\Lambda(s,\chi) from (2.9) that L(s,χ)L(s,\chi) has simple zeros at s=0,1,2,s=0,-1,-2,\ldots, these are called the trivial zeros of L(s,χ)L(s,\chi). Since we are assuming GRH, we know that the non-trivial zeros of L(s,χ)L(s,\chi) are precisely the zeros of Λ(s,χ)\Lambda(s,\chi).

Let ρ=12+iγ\rho=\tfrac{1}{2}+i\gamma run over the non-trivial zeros of L(s,χ)L(s,\chi). We then deduce from [HIEK, Theorem 5.6] and the observation that L(s,χ)L(s,\chi) is analytic at s=1s=1 since χ\chi is non-principal and primitive that

(4.2) Λ(s,χ)=eA+Bsρ(1sρ)esρ,\displaystyle\Lambda(s,\chi)=e^{A+Bs}\prod_{\rho}\left(1-\frac{s}{\rho}\right)e^{\frac{s}{\rho}},

where A,B=B(χ)A,B=B(\chi) are constants.

Taking the logarithmic derivative on both sides of (4.2) and making use of (2.9), we obtain that

LL(s,χ)=ΓΓ(s)logπ+12logN(m)Bρ(1sρ+1ρ).-\frac{L^{\prime}}{L}(s,\chi)=\frac{\Gamma^{\prime}}{\Gamma}(s)-\log\pi+\frac{1}{2}\log N(m)-B-\sum_{\rho}\left(\frac{1}{s-\rho}+\frac{1}{\rho}\right).

This implies that

(4.3) LL(s,χ)=ΓΓ(s)logπ+12logN(m)(B)ρ(1sρ+1ρ).\displaystyle-\Re{\frac{L^{\prime}}{L}(s,\chi)}=\Re{\frac{\Gamma^{\prime}}{\Gamma}(s)}-\log\pi+\frac{1}{2}\log N(m)-\Re(B)-\sum_{\rho}\Re{\left(\frac{1}{s-\rho}+\frac{1}{\rho}\right)}.

On the other hand, combining the functional equation (2.10) with (4.2), we see that

eA+Bsρ(1sρ)esρ=W(χ)(N(m))1/2eA+B(1s)ρ(11sρ)e1sρ.\displaystyle e^{A+Bs}\prod_{\rho}\left(1-\frac{s}{\rho}\right)e^{\frac{s}{\rho}}=W(\chi)(N(m))^{-1/2}e^{A+B(1-s)}\prod_{\rho}\left(1-\frac{1-s}{\rho}\right)e^{\frac{1-s}{\rho}}.

Taking logarithmic derivative on both sides of the above expression, we obtain that

2B=ρ(1sρ+11sρ+1ρ+1ρ)=2ρ1ρ.2B=-\sum_{\rho}\left(\frac{1}{s-\rho}+\frac{1}{1-s-\rho}+\frac{1}{\rho}+\frac{1}{\rho}\right)=-2\sum_{\rho}\frac{1}{\rho}.

Here the second equality above follows by noting that the terms containing 1sρ1-s-\rho and sρs-\rho cancel as both ρ,1ρ\rho,1-\rho are zeros from the functional equation (2.10). We note here that

ρ1ρ\sum_{\rho}\frac{1}{\rho}

is convergent since if ρ\rho is a zero, so is ρ¯\overline{\rho} and we have

ρ1+ρ¯1|ρ|2\displaystyle\rho^{-1}+\overline{\rho}^{-1}\ll|\rho|^{-2}

and we know that ρ1|ρ|2\displaystyle\sum_{\rho}\frac{1}{|\rho|^{2}} is convergent by [iwakow, Lemma 5.5].

We then deduce that

(B)=ρ(ρ1).\Re(B)=-\sum_{\rho}\Re(\rho^{-1}).

Combining the above with (4.3), we see that

(4.4) LL(s,χ)=ΓΓ(s)logπ+12logN(m)ρ(1sρ)=ΓΓ(s)logπ+12logN(m)F(s)=log(|t|+1)+12logN(m)+O(1)F(s)logT+12logN(m)+O(1)F(s).\displaystyle\begin{split}-\Re{\frac{L^{\prime}}{L}(s,\chi)}&=\Re{\frac{\Gamma^{\prime}}{\Gamma}(s)}-\log\pi+\frac{1}{2}\log N(m)-\Re{\sum_{\rho}\left(\frac{1}{s-\rho}\right)}\\ &=\Re{\frac{\Gamma^{\prime}}{\Gamma}(s)}-\log\pi+\frac{1}{2}\log N(m)-F(s)\\ &=\log(|t|+1)+\frac{1}{2}\log N(m)+O(1)-F(s)\\ &\leq\log T+\frac{1}{2}\log N(m)+O(1)-F(s).\end{split}

where the third line above follows from ΓΓ(s)=logs+O(|s|1)\frac{\Gamma^{\prime}}{\Gamma}(s)=\log s+O(|s|^{-1}) by (6) of [Da, §10] and where we define

F(s)=ρ1sρ=ρσ1/2(σ1/2)2+(tγ)2.\displaystyle F(s)=\Re{\sum_{\rho}\frac{1}{s-\rho}}=\sum_{\rho}\frac{\sigma-1/2}{(\sigma-1/2)^{2}+(t-\gamma)^{2}}.

Integrating the last expression given for LL(s,χ)-\Re{\frac{L^{\prime}}{L}(s,\chi)} in (4.4) from σ=(s)\sigma=\Re(s) to σ0>12\sigma_{0}>\frac{1}{2}, we obtain by setting s0=σ0+its_{0}=\sigma_{0}+it that

(4.5) log|L(s,χ)|log|L(s0,χ)|=(logT+logN(m)2+O(1))(σ0σ)σσ0F(u+it)𝑑u(σ0σ)(logT+logN(m)2+O(1)),\displaystyle\begin{split}&\log|L(s,\chi)|-\log|L(s_{0},\chi)|\\ &=\Big{(}\log T+\frac{\log N(m)}{2}+O(1)\Big{)}(\sigma_{0}-\sigma)-\int_{\sigma}^{\sigma_{0}}F(u+it)du\\ &\leq(\sigma_{0}-\sigma)\Big{(}\log T+\frac{\log N(m)}{2}+O(1)\Big{)},\end{split}

where the last inequality above follows from the observation that F(s)0F(s)\geq 0 for all ss satisfying (s)12\Re(s)\geq\tfrac{1}{2}.

Next, we deduce upon integrating term by term using the Dirichlet series expansion of LL(s+w,χ)-\frac{L^{\prime}}{L}(s+w,\chi) that

12πi(c)LL(s+w,χ)xww2dw=(n)N(n)xΛ[i](n)χ(n)N(n)slog(xN(n)),\frac{1}{2\pi i}\int\limits_{(c)}-\frac{L^{\prime}}{L}(s+w,\chi)\frac{x^{w}}{w^{2}}dw=\sum_{\begin{subarray}{c}(n)\\ N(n)\leq x\end{subarray}}\frac{\Lambda_{[i]}(n)\chi(n)}{N(n)^{s}}\log\left(\frac{x}{N(n)}\right),

where c>2c>2 is a large number. Now moving the line of integration in the above expression to the left and calculating residues, we see also that

12πi(c)LL(s+w,χ)xww2dw=LL(s,χ)logx(LL(s,χ))ρxρs(ρs)2k=0xks(k+s)2.\frac{1}{2\pi i}\int\limits_{(c)}-\frac{L^{\prime}}{L}(s+w,\chi)\frac{x^{w}}{w^{2}}dw=-\frac{L^{\prime}}{L}(s,\chi)\log x-\Big{(}\frac{L^{\prime}}{L}(s,\chi)\Big{)}^{\prime}-\sum_{\rho}\frac{x^{\rho-s}}{(\rho-s)^{2}}-\sum_{k=0}^{\infty}\frac{x^{-k-s}}{(k+s)^{2}}.

Comparing the above two expressions, we deduce that unconditionally, for any x2x\geq 2, we have

(4.6) LL(s,χ)=(n)N(n)xΛ[i](n)N(n)slog(x/N(n))logx+1logx(LL(s,χ))+1logxρxρs(ρs)2+1logxk=0xks(k+s)2.\displaystyle\begin{split}-\frac{L^{\prime}}{L}(s,\chi)=\sum_{\begin{subarray}{c}(n)\\ N(n)\leq x\end{subarray}}\frac{\Lambda_{[i]}(n)}{N(n)^{s}}\frac{\log(x/N(n))}{\log x}&+\frac{1}{\log x}\Big{(}\frac{L^{\prime}}{L}(s,\chi)\Big{)}^{\prime}+\frac{1}{\log x}\sum_{\rho}\frac{x^{\rho-s}}{(\rho-s)^{2}}+\frac{1}{\log x}\sum_{k=0}^{\infty}\frac{x^{-k-s}}{(k+s)^{2}}.\end{split}

We now integrate the real parts on both sides of (4.6) over σ=(s)\sigma=\Re(s) from σ0\sigma_{0} to \infty to see that for x2x\geq 2,

(4.7) log|L(s0,χ)|=((n)N(n)xΛ[i](n)N(n)s0logN(n)log(x/N(n))logx1logxLL(s0,χ)+1logxρσ0xρs(ρs)2dσ+O(1logx)).\displaystyle\begin{split}\log|L(s_{0},\chi)|=\Re\Big{(}\sum_{\begin{subarray}{c}(n)\\ N(n)\leq x\end{subarray}}\frac{\Lambda_{[i]}(n)}{N(n)^{s_{0}}\log N(n)}\frac{\log(x/N(n))}{\log x}&-\frac{1}{\log x}\frac{L^{\prime}}{L}(s_{0},\chi)\\ &+\frac{1}{\log x}\sum_{\rho}\int_{\sigma_{0}}^{\infty}\frac{x^{\rho-s}}{(\rho-s)^{2}}d\sigma+O\Big{(}\frac{1}{\log x}\Big{)}\Big{)}.\end{split}

Observe that

ρ|σ0xρs(ρs)2𝑑σ|ρσ0x12σ|s0ρ|2𝑑σ=ρx12σ0|s0ρ|2logx=x12σ0F(s0)(σ012)logx.\sum_{\rho}\Big{|}\int_{\sigma_{0}}^{\infty}\frac{x^{\rho-s}}{(\rho-s)^{2}}d\sigma\Big{|}\leq\sum_{\rho}\int_{\sigma_{0}}^{\infty}\frac{x^{\frac{1}{2}-\sigma}}{|s_{0}-\rho|^{2}}d\sigma=\sum_{\rho}\frac{x^{\frac{1}{2}-\sigma_{0}}}{|s_{0}-\rho|^{2}\log x}=\frac{x^{\frac{1}{2}-\sigma_{0}}F(s_{0})}{(\sigma_{0}-\frac{1}{2})\log x}.

Applying this and (4.4) in (4.7), we see that

(4.8) log|L(s0,χ)|(n)N(n)xΛ[i](n)N(n)s0logN(n)log(x/N(n))logx+1logx(logT+logN(m)2)+F(s0)(x12σ0(σ012)log2x1logx)+O(1logx).\displaystyle\begin{split}\log|L(s_{0},\chi)|&\leq\Re\sum_{\begin{subarray}{c}(n)\\ N(n)\leq x\end{subarray}}\frac{\Lambda_{[i]}(n)}{N(n)^{s_{0}}\log N(n)}\frac{\log(x/N(n))}{\log x}+\frac{1}{\log x}(\log T+\frac{\log N(m)}{2})\\ &\hskip 36.135pt+F(s_{0})\Big{(}\frac{x^{\frac{1}{2}-\sigma_{0}}}{(\sigma_{0}-\tfrac{1}{2})\log^{2}x}-\frac{1}{\log x}\Big{)}+O\Big{(}\frac{1}{\log x}\Big{)}.\end{split}

Adding inequalities (4.5) and (4.8), we deduce that

log|L(s,χ)|\displaystyle\log|L(s,\chi)| (n)N(n)xΛ[i](n)N(n)s0logN(n)log(x/N(n))logx+(logT+logN(m)2)(σ0σ+1logx)\displaystyle\leq\Re\sum_{\begin{subarray}{c}(n)\\ N(n)\leq x\end{subarray}}\frac{\Lambda_{[i]}(n)}{N(n)^{s_{0}}\log N(n)}\frac{\log(x/N(n))}{\log x}+(\log T+\frac{\log N(m)}{2})\Big{(}\sigma_{0}-\sigma+\frac{1}{\log x}\Big{)}
+F(s0)(x12σ0(σ012)log2x1logx)+O(1logx)+O(σ0σ).\displaystyle\hskip 36.135pt+F(s_{0})\Big{(}\frac{x^{\frac{1}{2}-\sigma_{0}}}{(\sigma_{0}-\tfrac{1}{2})\log^{2}x}-\frac{1}{\log x}\Big{)}+O\Big{(}\frac{1}{\log x}\Big{)}+O(\sigma_{0}-\sigma).

The assertion of the proposition now follows by setting σ0=12+λlogx\sigma_{0}=\frac{1}{2}+\frac{\lambda}{\log x} with λλ0\lambda\geq\lambda_{0} and by omitting the term involving F(s0)F(s_{0}) since it is negative. ∎

Our next lemma treats essentially the sum over prime squares in (4.1).

Lemma 4.3.

Assume GRH for ζK(s)\zeta_{K}(s). Let λ0\lambda_{0} be given as in Lemma 4.2 and let zz\in\mathbb{C} with 0(z)1logX0\leq\Re(z)\leq\frac{1}{\log X}. We have for x10x\geq 10,

(4.9) ϖ1mod(1+i)3N(ϖ)x1/21N(ϖ)1+2λ0logx+2zlog(xN(ϖ)2)logx=(z,x)+O(1),\displaystyle\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq x^{1/2}\end{subarray}}\frac{1}{N(\varpi)^{1+\frac{2\lambda_{0}}{\log x}+2z}}\frac{\log(\frac{x}{N(\varpi)^{2}})}{\log x}=\mathcal{L}(z,x)+O(1),

where (z,x)\mathcal{L}(z,x) is given in (1.9).

Proof.

We first note that under GRH for ζK(s)\zeta_{K}(s), the prime ideal theorem has the following form (see [iwakow, Theorem 5.15])

(4.10) ϖ1mod(1+i)3N(ϖ)ylogN(ϖ)=y+O(y(logXy)2).\displaystyle\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq y\end{subarray}}\log N(\varpi)=y+O(\sqrt{y}\left(\log Xy)^{2}\right).

It follows from this and (z)0\Re(z)\geq 0 that we have

ϖ1mod(1+i)3N(ϖ)x1/21N(ϖ)1+2λ0logx+2zlogN(ϖ)logxϖ1mod(1+i)3N(ϖ)x1N(ϖ)logN(ϖ)logx=O(1).\displaystyle\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq x^{1/2}\end{subarray}}\frac{1}{N(\varpi)^{1+\frac{2\lambda_{0}}{\log x}+2z}}\frac{\log N(\varpi)}{\log x}\ll\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq\sqrt{x}\end{subarray}}\frac{1}{N(\varpi)}\frac{\log N(\varpi)}{\log x}=O(1).

We then deduce from this that it suffices to establish (4.9) with the left-hand side expression in (4.9) being replaced by f(2λ0logx+2z,x)f(\frac{2\lambda_{0}}{\log x}+2z,x), where

f(z,x)=ϖ1mod(1+i)3N(ϖ)x1/21N(ϖ)1+z\displaystyle f(z,x)=\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq x^{1/2}\end{subarray}}\frac{1}{N(\varpi)^{1+z}}

It is easily seen that f(2λ0logx+2z,x)=O(1)f(\frac{2\lambda_{0}}{\log x}+2z,x)=O(1) when |z|1|z|\geq 1 using (4.10) and partial summation. The same procedure also implies that

(4.11) ϖ1mod(1+i)3N(ϖ)x1/21N(ϖ)=loglogx+O(1).\displaystyle\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq x^{1/2}\end{subarray}}\frac{1}{N(\varpi)}=\log\log x+O(1).

It follows that when |z|(logx)1|z|\leq(\log x)^{-1}, we have

f(2λ0logx+2z,x)=\displaystyle f(\frac{2\lambda_{0}}{\log x}+2z,x)= ϖ1mod(1+i)3N(ϖ)x1/21N(ϖ)+ϖ1mod(1+i)3N(ϖ)x1/2(1N(ϖ)1+2λ0logx+2z1N(ϖ))\displaystyle\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq x^{1/2}\end{subarray}}\frac{1}{N(\varpi)}+\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq x^{1/2}\end{subarray}}(\frac{1}{N(\varpi)^{1+\frac{2\lambda_{0}}{\log x}+2z}}-\frac{1}{N(\varpi)})
=\displaystyle= loglogx+ϖ1mod(1+i)3N(ϖ)x1/2logN(ϖ)N(ϖ)02λ0logx+2zN(ϖ)u𝑑u+O(1)\displaystyle\log\log x+\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq x^{1/2}\end{subarray}}\frac{\log N(\varpi)}{N(\varpi)}\int^{\frac{2\lambda_{0}}{\log x}+2z}_{0}N(\varpi)^{u}du+O(1)
=\displaystyle= loglogx+O(1logxϖ1mod(1+i)3N(ϖ)x1/2logN(ϖ)N(ϖ))+O(1)\displaystyle\log\log x+O(\frac{1}{\log x}\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq x^{1/2}\end{subarray}}\frac{\log N(\varpi)}{N(\varpi)})+O(1)
=\displaystyle= loglogx+O(1),\displaystyle\log\log x+O(1),

where the integral is along the line segment connecting the origin and the point 2λ0logx+2z\frac{2\lambda_{0}}{\log x}+2z on the complex plane.

In the remaining case when (logx)1|z|1(\log x)^{-1}\leq|z|\leq 1, we note that by (4.10) and partial summation,

fz=ϖ1mod(1+i)3N(ϖ)x1/2logN(ϖ)N(ϖ)1+z=1x1u1+zd(u+O(u1/2(logu)2))=1z+xz/2z+O(1).\displaystyle\frac{\partial f}{\partial z}=-\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq x^{1/2}\end{subarray}}\frac{\log N(\varpi)}{N(\varpi)^{1+z}}=-\int^{\sqrt{x}}_{1}\frac{1}{u^{1+z}}d(u+O(u^{1/2}(\log u)^{2}))=-\frac{1}{z}+\frac{x^{-z/2}}{z}+O(1).

Now we assume that 0(w)0\leq\Re(w) and |w|(logx)1|w|\geq(\log x)^{-1} and we integrate f/z\partial f/\partial z from 11 to ww to see that

f(w,x)=\displaystyle f(w,x)= logw+1wxu/2u𝑑u+O(1)=log|w|2xu/2ulogx|1w21wxu/2u2logx𝑑u+O(1)\displaystyle-\log w+\int^{w}_{1}\frac{x^{-u/2}}{u}du+O(1)=-\log|w|-\frac{2x^{-u/2}}{u\log x}\Big{|}^{w}_{1}-2\int^{w}_{1}\frac{x^{-u/2}}{u^{2}\log x}du+O(1)
=\displaystyle= log|w|21wxu/2u2logx𝑑u+O(1).\displaystyle-\log|w|-2\int^{w}_{1}\frac{x^{-u/2}}{u^{2}\log x}du+O(1).

We break the integration 1wxu/2u2logx𝑑u\int^{w}_{1}\frac{x^{-u/2}}{u^{2}\log x}du into two parts, one horizontal integration along the xx-axis from 11 to (w)(logx)1\Re(w)\gg(\log x)^{-1}, and the other vertical integration from (w)\Re(w) to zz. The horizontal integration is

(logx)11u2logx𝑑u=O(1).\displaystyle\ll\int^{\infty}_{(\log x)^{-1}}\frac{1}{u^{2}\log x}du=O(1).

If we write w=σ+itw=\sigma+it, then the vertical integration can be evaluated by breaking the integral over tt for |t|(z)|t|\leq\Re(z) and |t|>(z)|t|>\Re(z). We obtain this way that the vertical integration is

(z)(z)1(z)2logx𝑑t+(z)1t2logx𝑑t1(z)logx=O(1).\displaystyle\ll\int^{\Re(z)}_{-\Re(z)}\frac{1}{\Re(z)^{2}\log x}dt+\int^{\infty}_{\Re(z)}\frac{1}{t^{2}\log x}dt\ll\frac{1}{\Re(z)\log x}=O(1).

It follows that we have f(w,x)=log|w|+O(1)f(w,x)=-\log|w|+O(1) when 0(w)0\leq\Re(w) and |w|(logx)1|w|\geq(\log x)^{-1}. In particular, this applies to the case when w=2λ0logx+2zw=\frac{2\lambda_{0}}{\log x}+2z, thus completes the proof. ∎

Lastly, we present a mean value estimation which will be applied to estimate the sum over primes in (4.1) in our proof of Theorem 1.2.

Lemma 4.4.

Let XX and yy be real numbers. For fixed 0<ε<10<\varepsilon<1, let kk be a natural number with ykX1/2εy^{k}\leq X^{1/2-\varepsilon}. Then for any complex numbers a(ϖ)a(\varpi), we have

(d,2)=1N(d)X|ϖ1mod(1+i)3N(ϖ)ya(ϖ)χ(1+i)5d(ϖ)N(ϖ)12|2kεX(2k)!k!2k(ϖ1mod(1+i)3N(ϖ)y|a(ϖ)|2N(ϖ))k.\displaystyle\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ N(d)\leq X\end{subarray}}\left|\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq y\end{subarray}}\frac{a(\varpi)\chi_{(1+i)^{5}d}(\varpi)}{N(\varpi)^{\frac{1}{2}}}\right|^{2k}\ll_{\varepsilon}X\frac{(2k)!}{k!2^{k}}\left(\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq y\end{subarray}}\frac{|a(\varpi)|^{2}}{N(\varpi)}\right)^{k}.
Proof.

Let W(t)W(t) be any non-negative smooth function that is supported on (12ε1,1+ε1)(\frac{1}{2}-\varepsilon_{1},1+\varepsilon_{1}) for some fixed small 0<ε1<1/20<\varepsilon_{1}<1/2 such that W(t)1W(t)\gg 1 for t(12,1)t\in(\tfrac{1}{2},1). We have that

(d,2)=1N(d)X|ϖ1mod(1+i)3N(ϖ)ya(ϖ)χ(1+i)5d(ϖ)N(ϖ)12|2k(d,2)=1|ϖ1mod(1+i)3N(ϖ)ya(ϖ)χ(1+i)5d(ϖ)N(ϖ)12|2kW(N(d)X)\displaystyle\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ N(d)\leq X\end{subarray}}\left|\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq y\end{subarray}}\frac{a(\varpi)\chi_{(1+i)^{5}d}(\varpi)}{N(\varpi)^{\frac{1}{2}}}\right|^{2k}\ll\sum_{\begin{subarray}{c}(d,2)=1\end{subarray}}\left|\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq y\end{subarray}}\frac{a(\varpi)\chi_{(1+i)^{5}d}(\varpi)}{N(\varpi)^{\frac{1}{2}}}\right|^{2k}W(\frac{N(d)}{X})
\displaystyle\ll (d,2)=1|ϖ1,,ϖk1mod(1+i)3N(ϖ1),,N(ϖk)ya(ϖ1)a(ϖk)N(ϖ1ϖk)((1+i)5dϖ1ϖk)|2W(N(d)X).\displaystyle\sum_{\begin{subarray}{c}(d,2)=1\end{subarray}}\Big{|}\sum_{\begin{subarray}{c}\varpi_{1},\dots,\varpi_{k}\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi_{1}),\dots,N(\varpi_{k})\leq y\end{subarray}}\frac{a(\varpi_{1})\dots a(\varpi_{k})\ }{\sqrt{N(\varpi_{1}\dots\varpi_{k})}}\left(\frac{(1+i)^{5}d}{\varpi_{1}\cdots\varpi_{k}}\right)\Big{|}^{2}W(\frac{N(d)}{X}).

We further expand out the square in the last sum above and treat the sum over dd by applying (3.5) first and then using the smoothed version of Pólya-Vinogradov inequality (3.4) for number fields when the product of the primes involved is not a perfect square to see that we have

(d,2)=1|ϖ1,,ϖk1mod(1+i)3N(ϖ1),,N(ϖk)ya(ϖ1)a(ϖk)N(ϖ1ϖk)((1+i)5dϖ1ϖk)|2W(N(d)X)\displaystyle\sum_{\begin{subarray}{c}(d,2)=1\end{subarray}}\Big{|}\sum_{\begin{subarray}{c}\varpi_{1},\dots,\varpi_{k}\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi_{1}),\dots,N(\varpi_{k})\leq y\end{subarray}}\frac{a(\varpi_{1})\dots a(\varpi_{k})\ }{\sqrt{N(\varpi_{1}\dots\varpi_{k})}}\left(\frac{(1+i)^{5}d}{\varpi_{1}\cdots\varpi_{k}}\right)\Big{|}^{2}W(\frac{N(d)}{X})
\displaystyle\ll Xϖ1,,ϖ2k1mod(1+i)3N(ϖ1),,N(ϖ2k)yϖ1ϖ2k=|a(ϖ1)a(ϖ2k)|N(ϖ1ϖ2k)+O(ϖ1,,ϖ2k1mod(1+i)3N(ϖ1),,N(ϖ2k)y|a(ϖ1)a(ϖ2k)|y2kε),\displaystyle X\sum_{\begin{subarray}{c}\varpi_{1},\dots,\varpi_{2k}\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi_{1}),\dots,N(\varpi_{2k})\leq y\\ \varpi_{1}\dots\varpi_{2k}=\square\end{subarray}}\frac{|a(\varpi_{1})\dots a(\varpi_{2k})|}{\sqrt{N(\varpi_{1}\dots\varpi_{2k})}}+O\left(\sum_{\begin{subarray}{c}\varpi_{1},\dots,\varpi_{2k}\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi_{1}),\dots,N(\varpi_{2k})\leq y\end{subarray}}|a(\varpi_{1})\dots a(\varpi_{2k})|y^{2k\varepsilon^{\prime}}\right),

where we write \square for a square of an element in 𝒪K\mathcal{O}_{K}.

We take ε\varepsilon^{\prime} small enough so that y2ky2kεXy^{2k}y^{2k\varepsilon^{\prime}}\ll X. Then an argument similar to that in the proof of [S&Y, Lemma 6.3] leads to the assertion of the lemma. ∎

4.5. Completion of the proof

With lemmas 4.2-4.4 now available, we proceed to establish an upper bound for the frequency of large values of log|L(12+z1,χ(1+i)5d)L(12+z2,χ(1+i)5d)|\log|L(\tfrac{1}{2}+z_{1},\chi_{(1+i)^{5}d})L(\tfrac{1}{2}+z_{2},\chi_{(1+i)^{5}d})|.

Proposition 4.6.

Assume GRH for ζK(s)\zeta_{K}(s) and L(s,χ(1+i)5d)L(s,\chi_{(1+i)^{5}d}) for all odd, square-free d𝒪Kd\in\mathcal{O}_{K}. Let XX be large and let z1,z2z_{1},z_{2}\in\mathbb{C} with 0(z1),(z2)1logX0\leq\Re(z_{1}),\Re(z_{2})\leq\frac{1}{\log X}, and |(z1)|,|(z2)|X|\Im(z_{1})|,|\Im(z_{2})|\leq X. Let 𝒩(V;z1,z2,X)\mathcal{N}(V;z_{1},z_{2},X) denote the number of odd, square-free d𝒪Kd\in\mathcal{O}_{K} such that N(d)XN(d)\leq X and

log|L(12+z1,χ(1+i)5d)L(12+z2,χ(1+i)5d)|V+(z1,z2,X).\displaystyle\log|L(\tfrac{1}{2}+z_{1},\chi_{(1+i)^{5}d})L(\tfrac{1}{2}+z_{2},\chi_{(1+i)^{5}d})|\geq V+\mathcal{M}(z_{1},z_{2},X).

Then for 10loglogXV𝒱(z1,z2,X)10\sqrt{\operatorname{log}\operatorname{log}X}\leq V\leq\mathcal{V}(z_{1},z_{2},X), we have

𝒩(V;z1,z2,X)Xexp(V22𝒱(z1,z2,X)(125logloglogX));\displaystyle\mathcal{N}(V;z_{1},z_{2},X)\ll X\operatorname{exp}\left(-\frac{V^{2}}{2\mathcal{V}(z_{1},z_{2},X)}\left(1-\frac{25}{\operatorname{log}\operatorname{log}\operatorname{log}X}\right)\right);

for 𝒱(z1,z2,X)<V116𝒱(z1,z2,X)logloglogX\mathcal{V}(z_{1},z_{2},X)<V\leq\frac{1}{16}\mathcal{V}(z_{1},z_{2},X)\operatorname{log}\operatorname{log}\operatorname{log}X, we have

𝒩(V;z1,z2,X)Xexp(V22𝒱(z1,z2,X)(115V𝒱(z1,z2,X)logloglogX)2);\displaystyle\mathcal{N}(V;z_{1},z_{2},X)\ll X\operatorname{exp}\left(-\frac{V^{2}}{2\mathcal{V}(z_{1},z_{2},X)}\left(1-\frac{15V}{\mathcal{V}(z_{1},z_{2},X)\operatorname{log}\operatorname{log}\operatorname{log}X}\right)^{2}\right);

for 116𝒱(z1,z2,X)logloglogX<V\frac{1}{16}\mathcal{V}(z_{1},z_{2},X)\operatorname{log}\operatorname{log}\operatorname{log}X<V, we have

𝒩(V;z1,z2,X)Xexp(11025VlogV).\displaystyle\mathcal{N}(V;z_{1},z_{2},X)\ll X\operatorname{exp}\left(-\frac{1}{1025}V\operatorname{log}V\right).
Proof.

Let λ0\lambda_{0} be the constant defined in Lemma 4.2 and apply this Lemma with λ=λ0+(zi)logx,i=1,2\lambda=\lambda_{0}+\Re(z_{i})\log x,i=1,2 and T=XT=X to L(12+zi,χ(1+i)5d)L(\tfrac{1}{2}+z_{i},\chi_{(1+i)^{5}d}) for N(d)XN(d)\leq X, we see that for 2xX2\leq x\leq X,

log|L(12+zi,χ(1+i)5d)|\displaystyle\log|L(\tfrac{1}{2}+z_{i},\chi_{(1+i)^{5}d})| (n1mod(1+i)3N(n)xΛ[i](n)χ(1+i)5d(n)N(n)12+λ0logx+zilogN(n)log(xN(n))logx)+32(1+λ0)logXlogx+O(1logx),i=1,2.\displaystyle\leq\Re\left(\sum_{\begin{subarray}{c}n\equiv 1\bmod{(1+i)^{3}}\\ N(n)\leq x\end{subarray}}\frac{\Lambda_{[i]}(n)\chi_{(1+i)^{5}d}(n)}{N(n)^{\frac{1}{2}+\frac{\lambda_{0}}{\log x}+z_{i}}\log N(n)}\frac{\log(\frac{x}{N(n)})}{\log x}\right)+\frac{3}{2}(1+\lambda_{0})\frac{\log X}{\log x}+O\left(\frac{1}{\log x}\right),\ i=1,2.

We then deduce that

log|L(12+z1,χ(1+i)5d)||L(12+z2,χ(1+i)5d)|\displaystyle\log|L(\tfrac{1}{2}+z_{1},\chi_{(1+i)^{5}d})||L(\tfrac{1}{2}+z_{2},\chi_{(1+i)^{5}d})|
(4.12) (ϖ1mod(1+i)3N(ϖl)xl1χ(1+i)5d(ϖl)lN(ϖ)l(12+λ0logx)(N(ϖ)lz1+N(ϖ)lz2)log(xN(ϖ)l)logx)+3(1+λ0)logXlogx+O(1logx).\displaystyle\leq\Re\left(\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi^{l})\leq x\\ l\geq 1\end{subarray}}\frac{\chi_{(1+i)^{5}d}(\varpi^{l})}{lN(\varpi)^{l(\tfrac{1}{2}+\frac{\lambda_{0}}{\log x})}}(N(\varpi)^{-lz_{1}}+N(\varpi)^{-lz_{2}})\frac{\log(\frac{x}{N(\varpi)^{l}})}{\log x}\right)+3(1+\lambda_{0})\frac{\log X}{\log x}+O\left(\frac{1}{\log x}\right).

The terms with l3l\geq 3 in the the above sum contribute O(1)O(1). Using the fact ϖ|d1N(ϖ)logloglogN(d)\displaystyle\sum_{\varpi|d}\frac{1}{N(\varpi)}\ll\log\log\log N(d), we deduce from Lemma 4.3 that

(ϖ1mod(1+i)3N(ϖ)x1/2χ(1+i)5d(ϖ2)2N(ϖ)1+2λ0logx(N(ϖ)2z1+N(ϖ)2z2)log(xN(ϖ)2)logx)=(ϖ1mod(1+i)3N(ϖ)x1/212N(ϖ)1+2λ0logx(N(ϖ)2z1+N(ϖ)2z2)log(xN(ϖ)2)logx)=(z1,z2,x)+O(logloglogX)(z1,z2,X)+O(logloglogX),\displaystyle\begin{split}&\Re\left(\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq x^{1/2}\end{subarray}}\frac{\chi_{(1+i)^{5}d}(\varpi^{2})}{2N(\varpi)^{1+\frac{2\lambda_{0}}{\log x}}}(N(\varpi)^{-2z_{1}}+N(\varpi)^{-2z_{2}})\frac{\log(\frac{x}{N(\varpi)^{2}})}{\log x}\right)\\ =&\Re\left(\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq x^{1/2}\end{subarray}}\frac{1}{2N(\varpi)^{1+\frac{2\lambda_{0}}{\log x}}}(N(\varpi)^{-2z_{1}}+N(\varpi)^{-2z_{2}})\frac{\log(\frac{x}{N(\varpi)^{2}})}{\log x}\right)\\ =&\mathcal{M}(z_{1},z_{2},x)+O(\log\log\log X)\\ \leq&\mathcal{M}(z_{1},z_{2},X)+O(\log\log\log X),\end{split}

where (z1,z2,x)\mathcal{M}(z_{1},z_{2},x) is defined as in (1.10).

Applying the above estimation in (4.12), we obtain that

(4.13) log|L(12+z1,χ(1+i)5d)||L(12+z2,χ(1+i)5d)|(ϖ1mod(1+i)3N(ϖ)xχ(1+i)5d(ϖ)N(ϖ)12+λ0logx(N(ϖ)z1+N(ϖ)z2)log(xN(ϖ))logx)+(z1,z2,X)+5logXlogx+O(logloglogX).\displaystyle\begin{split}&\log|L(\tfrac{1}{2}+z_{1},\chi_{(1+i)^{5}d})||L(\tfrac{1}{2}+z_{2},\chi_{(1+i)^{5}d})|\\ \leq&\Re\left(\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq x\end{subarray}}\frac{\chi_{(1+i)^{5}d}(\varpi)}{N(\varpi)^{\frac{1}{2}+\frac{\lambda_{0}}{\log x}}}(N(\varpi)^{-z_{1}}+N(\varpi)^{-z_{2}})\frac{\log(\frac{x}{N(\varpi)})}{\log x}\right)+\mathcal{M}(z_{1},z_{2},X)+\frac{5\log X}{\log x}+O(\log\log\log X).\end{split}

By taking x=logXx=\log X in (4.13) and bounding the sum over ϖ\varpi in (4.13) trivially (with the help of (4.10)), we see that 𝒩(V;z1,z2,X)=0\mathcal{N}(V;z_{1},z_{2},X)=0 for V>6logXloglogXV>\frac{6\log X}{\log\log X}. Thus, we can assume V6logXloglogXV\leq\frac{6\log X}{\log\log X}.

In what follows, we shall denote 𝒱\mathcal{V} for 𝒱(z1,z2,X)\mathcal{V}(z_{1},z_{2},X) defined in (1.11) and we note that loglogX+O(1)𝒱(z1,z2,x)4loglogX+O(1)\log\log X+O(1)\leq\mathcal{V}(z_{1},z_{2},x)\leq 4\log\log X+O(1). We now set x=XA/Vx=X^{A/V} with

A={12logloglogX10loglogXV𝒱,𝒱2VlogloglogX𝒱<V116𝒱logloglogX,8V>116𝒱logloglogX.\displaystyle A=\left\{\begin{array}[c]{ll}\frac{1}{2}\log\log\log X&10\sqrt{\log\log X}\leq V\leq\mathcal{V},\\ \frac{\mathcal{V}}{2V}\log\log\log X&\mathcal{V}<V\leq\frac{1}{16}\mathcal{V}\log\log\log X,\\ 8&V>\frac{1}{16}\mathcal{V}\log\log\log X.\end{array}\right.

We further denote z=x1/loglogXz=x^{1/\log\log X}, M1M_{1} for the real part of the sum in (4.13) truncated to N(ϖ)zN(\varpi)\leq z, M2M_{2} for the real part of the sum in (4.13) over z<N(ϖ)xz<N(\varpi)\leq x. We then deduce that

log|L(12+z1,χ(1+i)5d)||L(12+z2,χ(1+i)5d)|M1+M2+(z1,z2,X)+5VA.\log|L(\tfrac{1}{2}+z_{1},\chi_{(1+i)^{5}d})||L(\tfrac{1}{2}+z_{2},\chi_{(1+i)^{5}d})|\leq M_{1}+M_{2}+\mathcal{M}(z_{1},z_{2},X)+\frac{5V}{A}.

It follows from this that if log|L(12+z1,χ(1+i)5d)||L(12+z2,χ(1+i)5d)|V+(z1,z2,X)\log|L(\tfrac{1}{2}+z_{1},\chi_{(1+i)^{5}d})||L(\tfrac{1}{2}+z_{2},\chi_{(1+i)^{5}d})|\geq V+\mathcal{M}(z_{1},z_{2},X), then we have either

M2VA, or M1V1:=V(16A).M_{2}\geq\tfrac{V}{A},\text{ or }M_{1}\geq V_{1}:=V(1-\tfrac{6}{A}).

Now, we define

meas(X;M1)\displaystyle\operatorname{meas}(X;M_{1}) =#{N(d)X:d odd and square-free, M1V1},\displaystyle=\#\{N(d)\leq X\ :\ d\text{ odd and square-free, }M_{1}\geq V_{1}\},
meas(X;M2)\displaystyle\operatorname{meas}(X;M_{2}) =#{N(d)X:d odd and square-free, M2VA}.\displaystyle=\#\{N(d)\leq X\ :\ d\text{ odd and square-free, }M_{2}\geq\tfrac{V}{A}\}.

We then take m=[0.9(V2A)]m=[0.9(\frac{V}{2A})] to see that by Lemma 4.4, we have

(V/A)2mmeas(X;S2)(d,2)=1N(d)X|M2|2mX(2m)!m!2m(ϖ1mod(1+i)3z<N(ϖ)x4N(ϖ))mX(3mlogloglogX)m,\displaystyle(V/A)^{2m}\operatorname{meas}(X;S_{2})\leq\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ N(d)\leq X\end{subarray}}|M_{2}|^{2m}\ll X\frac{(2m)!}{m!2^{m}}\left(\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ z<N(\varpi)\leq x\end{subarray}}\frac{4}{N(\varpi)}\right)^{m}\ll X(3m\log\log\log X)^{m},

where the last estimation above follows from (4.11) and Stirling’s formula (see [iwakow, (5.112)]), which implies that

(2m)!m!2m(2me)m.\displaystyle\frac{(2m)!}{m!2^{m}}\ll(\frac{2m}{e})^{m}.

We then deduce that

(4.14) meas(X;M2)Xexp(V3AlogV).\operatorname{meas}(X;M_{2})\ll X\operatorname{exp}\left(-\frac{V}{3A}\log V\right).

Next, we estimate meas(X;M1)\operatorname{meas}(X;M_{1}). For any m(12ε)logXlogzm\leq\frac{(\frac{1}{2}-\varepsilon)\log X}{\log z}, we obtain using Lemma 4.4 that

(4.15) V12mmeas(X;M1)(d,2)=1N(d)X|M1|2mX(2m)!m!2m(ϖ1mod(1+i)3N(ϖ)z|a(ϖ)|2N(ϖ))m,\displaystyle V^{2m}_{1}\operatorname{meas}(X;M_{1})\leq\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ N(d)\leq X\end{subarray}}|M_{1}|^{2m}\ll X\frac{(2m)!}{m!2^{m}}\left(\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq z\end{subarray}}\frac{|a(\varpi)|^{2}}{N(\varpi)}\right)^{m},

where

a(ϖ)=(N(ϖ)z1+N(ϖ)z2)log(xN(ϖ))N(ϖ)λ0logxlogx.\displaystyle a(\varpi)=\frac{\Re(N(\varpi)^{-z_{1}}+N(\varpi)^{-z_{2}})\log(\frac{x}{N(\varpi)})}{N(\varpi)^{\frac{\lambda_{0}}{\log x}}\log x}.

By arguing as in the proof of Lemma 4.3, we see that

ϖ1mod(1+i)3N(ϖ)z|a(ϖ)|2N(ϖ)14ϖ1mod(1+i)3N(ϖ)z1N(ϖ)(N(ϖ)z1+N(ϖ)z1¯+N(ϖ)z2+N(ϖ)z2¯)2=𝒱+O(1).\displaystyle\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq z\end{subarray}}\frac{|a(\varpi)|^{2}}{N(\varpi)}\leq\frac{1}{4}\sum_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ N(\varpi)\leq z\end{subarray}}\frac{1}{N(\varpi)}(N(\varpi)^{-z_{1}}+N(\varpi)^{-\overline{z_{1}}}+N(\varpi)^{-z_{2}}+N(\varpi)^{-\overline{z_{2}}})^{2}=\mathcal{V}+O(1).

Combining with (4.15), this implies that

meas(X;M1)XV12m(2m)!m!2m(𝒱+O(1))mX(2me𝒱+O(1)V12)m.\displaystyle\operatorname{meas}(X;M_{1})\ll XV_{1}^{-2m}\frac{(2m)!}{m!2^{m}}(\mathcal{V}+O(1))^{m}\ll X\left(\frac{2m}{e}\cdot\frac{\mathcal{V}+O(1)}{V_{1}^{2}}\right)^{m}.

We now take m=[V122𝒱]m=[\frac{V_{1}^{2}}{2\mathcal{V}}] when V(loglogX)2V\leq(\log\log X)^{2} and m=[10V]m=[10V] otherwise to see that in either case, we have for XX large,

m(120.1)logXlogz.\displaystyle m\leq\frac{(\frac{1}{2}-0.1)\log X}{\log z}.

A little calculation then shows that

meas(X;M1)Xexp(V122𝒱(1+O(1loglogX)))+Xexp(VlogV).\displaystyle\operatorname{meas}(X;M_{1})\ll X\operatorname{exp}\left(-\frac{V_{1}^{2}}{2\mathcal{V}}\left(1+O\left(\frac{1}{\log\log X}\right)\right)\right)+X\operatorname{exp}\left(-V\log V\right).

We then deduce from the above and (4.14) that

𝒩(V;z1,z2,X)Xexp(V122𝒱(1+O(1loglogX)))+Xexp(VlogV)+Xexp(V3AlogV).\displaystyle\mathcal{N}(V;z_{1},z_{2},X)\ll X\operatorname{exp}\left(-\frac{V_{1}^{2}}{2\mathcal{V}}\left(1+O\left(\frac{1}{\log\log X}\right)\right)\right)+X\operatorname{exp}\left(-V\log V\right)+X\operatorname{exp}\left(-\frac{V}{3A}\log V\right).

It is then easy to check that this leads to the assertion of the proposition. ∎

We now return to the proof of Theorem 1.2 by noting that for kk given as in the theorem, Proposition 4.6 implies that for all V10loglogXV\geq 10\sqrt{\log\log X},

(4.16) 𝒩(V;z1,z2,X){X,V<10loglogX,X(logX)o(1)exp(V22𝒱(z1,z2,X)),10loglogXV4k𝒱(z1,z2,X),X(logX)o(1)exp(4kV),V>4k𝒱(z1,z2,X).\displaystyle\mathcal{N}(V;z_{1},z_{2},X)\ll\begin{cases}X,\quad V<10\sqrt{\log\log X},\\ X(\operatorname{log}X)^{o(1)}\operatorname{exp}\left(-\frac{V^{2}}{2\mathcal{V}(z_{1},z_{2},X)}\right),\quad 10\sqrt{\log\log X}\leq V\leq 4k\mathcal{V}(z_{1},z_{2},X),\\ X(\operatorname{log}X)^{o(1)}\operatorname{exp}(-4kV),\quad V>4k\mathcal{V}(z_{1},z_{2},X).\end{cases}

Note further that we have

(d,2)=1N(d)X|L(12+z1,χ(1+i)5d)||L(12+z2,χ(1+i)5d)|k=\displaystyle\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ N(d)\leq X\end{subarray}}|L(\tfrac{1}{2}+z_{1},\chi_{(1+i)^{5}d})||L(\tfrac{1}{2}+z_{2},\chi_{(1+i)^{5}d})|^{k}= +exp(kV+k(z1,z2,X))𝑑𝒩(V;z1,z2,X)\displaystyle-\int_{-\infty}^{+\infty}\operatorname{exp}(kV+k\mathcal{M}(z_{1},z_{2},X))d\mathcal{N}(V;z_{1},z_{2},X)
=\displaystyle= k+exp(kV+k(z1,z2,X))𝒩(V;z1,z2,X)𝑑V.\displaystyle k\int_{-\infty}^{+\infty}\operatorname{exp}(kV+k\mathcal{M}(z_{1},z_{2},X))\mathcal{N}(V;z_{1},z_{2},X)dV.

Now the assertion of Theorem 1.2 follows by supplying the bound given in (4.16) to evaluate the integration above.

5. Proof of Theorems 1.5 and 1.6

5.1. Initial treatment

Let Φ\Phi be a smooth Schwartz class function which is compactly supported on [12,52][\frac{1}{2},\frac{5}{2}] satisfying 0Φ(t)10\leq\Phi(t)\leq 1 for all tt. We apply the approximate functional equation (2.12) to see that

(d,2)=1L(12,χ(1+i)5d)4Φ(N(d)X)=(d,2)=1AN(d)(d)2Φ(N(d)X),\displaystyle\sideset{}{{}^{*}}{\sum}_{(d,2)=1}L(\tfrac{1}{2},\chi_{(1+i)^{5}d})^{4}\Phi\left(\frac{N(d)}{X}\right)=\sideset{}{{}^{*}}{\sum}_{(d,2)=1}A_{N(d)}(d)^{2}\Phi\left(\frac{N(d)}{X}\right),

where

(5.1) At(d)=2n1mod(1+i)3χ(1+i)5d(n)d[i](n)N(n)12V(N(n)t).\displaystyle A_{t}(d)=2\sum_{\begin{subarray}{c}n\equiv 1\bmod{(1+i)^{3}}\end{subarray}}\frac{\chi_{(1+i)^{5}d}(n)d_{[i]}(n)}{N(n)^{\frac{1}{2}}}V\left(\frac{N(n)}{t}\right).

For two parameters U1,U2U_{1},U_{2} satisfying X910U1U2XX^{\frac{9}{10}}\leq U_{1}\leq U_{2}\leq X, we define

(5.2) S(U1,U2)=(d,2)=1AU1(d)AU2(d)Φ(N(d)X).\displaystyle S(U_{1},U_{2})=\sideset{}{{}^{*}}{\sum}_{(d,2)=1}A_{U_{1}}(d)A_{U_{2}}(d)\Phi\left(\frac{N(d)}{X}\right).

We let

(5.3) h(x,y,z)=Φ(N(x)X)V(N(y)U1)V(N(z)U2).\displaystyle h(x,y,z)=\Phi\left(\frac{N(x)}{X}\right)V\left(\frac{N(y)}{U_{1}}\right)V\left(\frac{N(z)}{U_{2}}\right).

Then applying (5.1) to (5.2) and using the Möbius inversion to remove the square-free condition in (5.2), we obtain that

S(U1,U2)=4(d,2)=1n1,n21mod(1+i)3χ(1+i)5d(n1n2)d[i](n1)d[i](n2)N(n1n2)12h(d,n1,n2)=4(d,2)=1a1mod(1+i)3a2|dμ[i](a)n1,n21mod(1+i)3χ(1+i)5d(n1n2)d[i](n1)d[i](n2)N(n1n2)12h(d,n1,n2)=4a1mod(1+i)3μ[i](a)(d,2)=1n1,n21mod(1+i)3(n1n2,a)=1χ(1+i)5d(n1n2)d[i](n1)d[i](n2)N(n1n2)12h(a2d,n1,n2).\displaystyle\begin{split}S(U_{1},U_{2})=&4\sideset{}{{}^{*}}{\sum}_{(d,2)=1}\sum_{\begin{subarray}{c}n_{1},n_{2}\equiv 1\bmod{(1+i)^{3}}\end{subarray}}\frac{\chi_{(1+i)^{5}d}(n_{1}n_{2})d_{[i]}(n_{1})d_{[i]}(n_{2})}{N(n_{1}n_{2})^{\frac{1}{2}}}h(d,n_{1},n_{2})\\ =&4\sum_{(d,2)=1}\sum_{\begin{subarray}{c}a\equiv 1\bmod{(1+i)^{3}}\\ a^{2}|d\end{subarray}}\mu_{[i]}(a)\sum_{\begin{subarray}{c}n_{1},n_{2}\equiv 1\bmod{(1+i)^{3}}\end{subarray}}\frac{\chi_{(1+i)^{5}d}(n_{1}n_{2})d_{[i]}(n_{1})d_{[i]}(n_{2})}{N(n_{1}n_{2})^{\frac{1}{2}}}h(d,n_{1},n_{2})\\ =&4\sum_{\begin{subarray}{c}a\equiv 1\bmod{(1+i)^{3}}\end{subarray}}\mu_{[i]}(a)\sum_{(d,2)=1}\sum_{\begin{subarray}{c}n_{1},n_{2}\equiv 1\bmod{(1+i)^{3}}\\ (n_{1}n_{2},a)=1\end{subarray}}\frac{\chi_{(1+i)^{5}d}(n_{1}n_{2})d_{[i]}(n_{1})d_{[i]}(n_{2})}{N(n_{1}n_{2})^{\frac{1}{2}}}h(a^{2}d,n_{1},n_{2}).\end{split}

Now we separate the terms with N(a)YN(a)\leq Y and with N(a)>YN(a)>Y for some YXY\leq X to be chosen later, writing S(U1,U2)=S1+S2S(U_{1},U_{2})=S_{1}+S_{2}, respectively. We bound S2S_{2} first in the following result.

Lemma 5.2.

Unconditionally, we have S2X1+εY1S_{2}\ll X^{1+\varepsilon}Y^{-1}. Under GRH, we have S2XY1(logX)46S_{2}\ll XY^{-1}(\log X)^{46}.

Proof.

We first write d=lb2d=lb^{2} with ll square-free and bb primary. We then let c=abc=ab and apply the definition of h(x,y,z)h(x,y,z) in (5.3) to see that

(5.4) S2=4c1mod(1+i)3a1mod(1+i)3N(a)>Ya|cμ[i](a)(l,2)=1n1,n21mod(1+i)3(n1n2,c)=1χ(1+i)5l(n1n2)d[i](n1)d[i](n2)N(n1n2)12h(c2l,n1,n2)=4(2πi)2c1mod(1+i)3a1mod(1+i)3N(a)>Ya|cμ[i](a)(12+ε)(12+ε)w(u)w(v)uvU1uU2v×(l,2)=1Φ(N(c2l)X)Lc(12+u,χ(1+i)5l)2Lc(12+v,χ(1+i)5l)2dudv,\displaystyle\begin{split}S_{2}&=4\sum_{\begin{subarray}{c}c\equiv 1\bmod{(1+i)^{3}}\end{subarray}}\sum_{\begin{subarray}{c}a\equiv 1\bmod{(1+i)^{3}}\\ N(a)>Y\\ a|c\end{subarray}}\mu_{[i]}(a)\sideset{}{{}^{*}}{\sum}_{(l,2)=1}\sum_{\begin{subarray}{c}n_{1},n_{2}\equiv 1\bmod{(1+i)^{3}}\\ (n_{1}n_{2},c)=1\end{subarray}}\frac{\chi_{(1+i)^{5}l}(n_{1}n_{2})d_{[i]}(n_{1})d_{[i]}(n_{2})}{N(n_{1}n_{2})^{\frac{1}{2}}}h(c^{2}l,n_{1},n_{2})\\ &=\frac{4}{(2\pi i)^{2}}\sum_{\begin{subarray}{c}c\equiv 1\bmod{(1+i)^{3}}\end{subarray}}\sum_{\begin{subarray}{c}a\equiv 1\bmod{(1+i)^{3}}\\ N(a)>Y\\ a|c\end{subarray}}\mu_{[i]}(a)\int\limits_{(\frac{1}{2}+\varepsilon)}\int\limits_{(\frac{1}{2}+\varepsilon)}\frac{w(u)w(v)}{uv}U_{1}^{u}U_{2}^{v}\\ &\quad\times\sideset{}{{}^{*}}{\sum}_{(l,2)=1}\Phi\left(\frac{N(c^{2}l)}{X}\right)L_{c}(\tfrac{1}{2}+u,\chi_{(1+i)^{5}l})^{2}L_{c}(\tfrac{1}{2}+v,\chi_{(1+i)^{5}l})^{2}dudv,\end{split}

where we define Lc(s,χ)L_{c}(s,\chi) to be the function by removing the Euler factors from L(s,χ)L(s,\chi) at prime ideals dividing cc.

Applying the estimation

|Lc(12+u,χ(1+i)5l)2Lc(12+v,χ(1+i)5l)2||Lc(12+u,χ(1+i)5l)|4+|Lc(12+u,χ(1+i)5l)|4\displaystyle|L_{c}(\tfrac{1}{2}+u,\chi_{(1+i)^{5}l})^{2}L_{c}(\tfrac{1}{2}+v,\chi_{(1+i)^{5}l})^{2}|\ll|L_{c}(\tfrac{1}{2}+u,\chi_{(1+i)^{5}l})|^{4}+|L_{c}(\tfrac{1}{2}+u,\chi_{(1+i)^{5}l})|^{4}
\displaystyle\ll d[i]2(c)(|L(12+u,χ(1+i)5l)|4+|L(12+u,χ(1+i)5l)|4),\displaystyle d_{[i]}^{2}(c)\left(|L(\tfrac{1}{2}+u,\chi_{(1+i)^{5}l})|^{4}+|L(\tfrac{1}{2}+u,\chi_{(1+i)^{5}l})|^{4}\right),

we can bound S2S_{2} by moving the lines of the integrations in (5.4) to (u)=(v)=1logX\Re(u)=\Re(v)=\frac{1}{\log X} to see that

(5.5) S2(logX)2(c,2)=1d[i](c)4N(a)>Ya|c(1logX)(1logX)|w(u)w(v)|(l,2)=1N(l)5X2N(c)2|L(12+u,χ(1+i)5l)|4|du||dv|.\displaystyle S_{2}\ll(\log X)^{2}\sum_{(c,2)=1}d_{[i]}(c)^{4}\sum_{\begin{subarray}{c}N(a)>Y\\ a|c\end{subarray}}\int\limits_{(\frac{1}{\log X})}\int\limits_{(\frac{1}{\log X})}|w(u)w(v)|\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(l,2)=1\\ N(l)\leq\frac{5X}{2N(c)^{2}}\end{subarray}}|L(\tfrac{1}{2}+u,\chi_{(1+i)^{5}l})|^{4}\ |du||dv|.

By Corollary 1.4, we see that for |(u)|5X(logX)22N(c)2|\Im(u)|\leq\frac{5X(\log X)^{2}}{2N(c)^{2}},

(5.6) (l,2)=1N(l)5X2N(c)2|L(12+u,χ(1+i)5l)|4XN(c)2(logX)13.\displaystyle\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(l,2)=1\\ N(l)\leq\frac{5X}{2N(c)^{2}}\end{subarray}}|L(\tfrac{1}{2}+u,\chi_{(1+i)^{5}l})|^{4}\ll\frac{X}{N(c)^{2}}(\log X)^{13}.

Note that Lemma 2.15 implies that

(5.7) (l,2)=1N(l)5X2N(c)2|L(12+u,χ(1+i)5l)|4(XN(c)2)1+ε(1+|(u)|2)1+ε.\displaystyle\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(l,2)=1\\ N(l)\leq\frac{5X}{2N(c)^{2}}\end{subarray}}|L(\tfrac{1}{2}+u,\chi_{(1+i)^{5}l})|^{4}\ll\left(\frac{X}{N(c)^{2}}\right)^{1+\varepsilon}(1+|\Im(u)|^{2})^{1+\varepsilon}.

Applying (5.6) in (5.5) when |(u)|5X(logX)22N(c)2|\Im(u)|\leq\frac{5X(\log X)^{2}}{2N(c)^{2}} and (5.7) in (5.5) otherwise, together with the observation that w(u)w(u) decreases exponentially in (u)\Im(u), we deduce that

S2X(logX)15(c,2)=1d[i]4(c)N(c)2N(a)>Ya|c1X(logX)15N(c)>Yd[i]5(c)N(c)2XY1(logX)46,\displaystyle S_{2}\ll X(\log X)^{15}\sum_{(c,2)=1}\frac{d^{4}_{[i]}(c)}{N(c)^{2}}\sum_{\begin{subarray}{c}N(a)>Y\\ a|c\end{subarray}}1\ll X(\log X)^{15}\sum_{N(c)>Y}\frac{d^{5}_{[i]}(c)}{N(c)^{2}}\ll XY^{-1}(\log X)^{46},

where the last estimation above follows from partial summation and the following estimation for x>2x>2:

(5.8) N(c)xd[i]5(c)x(logx)31.\displaystyle\sum_{N(c)\leq x}d^{5}_{[i]}(c)\ll x(\log x)^{31}.

The unconditionally estimation for S2S_{2} is obtained similarly by applying (5.7) in (5.5) for all uu and this completes the proof of the lemma. ∎

Next, we treat S1S_{1} by applying the Poisson summation formula given in Lemma 2.8 to recast it as

(5.9) S1=4a1mod(1+i)3N(a)Yμ[i](a)n1,n21mod(1+i)3(n1n2,a)=1((1+i)5n1n2)d[i](n1)d[i](n2)N(n1n2)12(d,2)=1(dn1n2)h(a2d,n1,n2)=2Xa1mod(1+i)3N(a)Yμ[i](a)N(a)2k𝒪K(1)N(k)n1,n21mod(1+i)3(n1n2,a)=1d[i](n1)d[i](n2)N(n1n2)12g(k,n1n2)N(n1n2)h~(N(k)X2N(a2n1n2),n1,n2),\displaystyle\begin{split}S_{1}=&4\sum_{\begin{subarray}{c}a\equiv 1\bmod{(1+i)^{3}}\\ N(a)\leq Y\end{subarray}}\mu_{[i]}(a)\sum_{\begin{subarray}{c}n_{1},n_{2}\equiv 1\bmod{(1+i)^{3}}\\ (n_{1}n_{2},a)=1\end{subarray}}\left(\frac{(1+i)^{5}}{n_{1}n_{2}}\right)\frac{d_{[i]}(n_{1})d_{[i]}(n_{2})}{N(n_{1}n_{2})^{\frac{1}{2}}}\sum_{(d,2)=1}\left(\frac{d}{n_{1}n_{2}}\right)h(a^{2}d,n_{1},n_{2})\\ =&2X\sum_{\begin{subarray}{c}a\equiv 1\bmod{(1+i)^{3}}\\ N(a)\leq Y\end{subarray}}\frac{\mu_{[i]}(a)}{N(a)^{2}}\sum_{k\in\mathcal{O}_{K}}(-1)^{N(k)}\sum_{\begin{subarray}{c}n_{1},n_{2}\equiv 1\bmod{(1+i)^{3}}\\ (n_{1}n_{2},a)=1\end{subarray}}\frac{d_{[i]}(n_{1})d_{[i]}(n_{2})}{N(n_{1}n_{2})^{\frac{1}{2}}}\frac{g(k,n_{1}n_{2})}{N(n_{1}n_{2})}\widetilde{h}\Big{(}\sqrt{\frac{N(k)X}{2N(a^{2}n_{1}n_{2})}},n_{1},n_{2}\Big{)},\end{split}

where

h~(t,y,z)=Φ~(t)V(N(y)U1)V(N(z)U2).\displaystyle\widetilde{h}(t,y,z)=\widetilde{\Phi}(t)V\left(\frac{N(y)}{U_{1}}\right)V\left(\frac{N(z)}{U_{2}}\right).

Now we write S1=S1(k=0)+S1(k0)S_{1}=S_{1}(k=0)+S_{1}(k\neq 0), where S1(k=0)S_{1}(k=0) corresponds to the term with k=0k=0. By applying (2.11) and (2.18), we see that when k0k\neq 0,

(5.10) h~(N(k)X2N(a2n1n2),n1,n2)=π(2πi)3(ε)(ε)(ε)(N(a)2N(k))s𝒥(s)w(u)w(v)1N(n1)usN(n2)vsU1uU2vXsuv𝑑s𝑑u𝑑v,=π(2πi)3(ε)(ε)(12+ε)(N(a)2N(k))s𝒥(s)w(u+s)w(v+s)1N(n1)uN(n2)vU1u+sU2v+sXs(u+s)(v+s)𝑑s𝑑u𝑑v,\displaystyle\begin{split}&\widetilde{h}\Big{(}\sqrt{\frac{N(k)X}{2N(a^{2}n_{1}n_{2})}},n_{1},n_{2}\Big{)}\\ =&\frac{\pi}{(2\pi i)^{3}}\int\limits_{(\varepsilon)}\int\limits_{(\varepsilon)}\int\limits_{(\varepsilon)}\left(\frac{N(a)^{2}}{N(k)}\right)^{s}\mathcal{J}(s)w(u)w(v)\frac{1}{N(n_{1})^{u-s}N(n_{2})^{v-s}}\frac{U_{1}^{u}U_{2}^{v}X^{-s}}{uv}\ ds\ du\ dv,\\ =&\frac{\pi}{(2\pi i)^{3}}\int\limits_{(\varepsilon)}\int\limits_{(\varepsilon)}\int\limits_{(\tfrac{1}{2}+\varepsilon)}\left(\frac{N(a)^{2}}{N(k)}\right)^{s}\mathcal{J}(s)w(u+s)w(v+s)\frac{1}{N(n_{1})^{u}N(n_{2})^{v}}\frac{U_{1}^{u+s}U_{2}^{v+s}X^{-s}}{(u+s)(v+s)}\ ds\ du\ dv,\end{split}

where

𝒥(s)=Φ^(1s)(π22)sΓ(s)Γ(1s)\displaystyle\mathcal{J}(s)=\widehat{\Phi}(1-s)(\frac{\pi^{2}}{2})^{-s}\frac{\Gamma(s)}{\Gamma(1-s)}

and where the last expression in (5.10) follows from moving the lines of the first triple integral in (5.10) to (s)=12+ε,(u)=(v)=12+2ε\Re(s)=\frac{1}{2}+\varepsilon,\Re(u)=\Re(v)=\frac{1}{2}+2\varepsilon, and a change the variables u=us,v=vsu^{\prime}=u-s,v^{\prime}=v-s.

Substituting the last expression in (5.10) to (5.9), we see by using our notation for k1,k2k_{1},k_{2} given in Section 2.9 that

(5.11) S1(k0)=2Xa1mod(1+i)3N(a)Yμ[i](a)N(a)2k𝒪Kk0(1)N(k)π(2πi)3(ε)(ε)(12+ε)(N(a)2N(k))s𝒥(s)w(u+s)w(v+s)×U1u+sU2v+sXs(u+s)(v+s)Z(12+u,12+v,a,k)dsdudv=2πX(2πi)3a1mod(1+i)3N(a)Yμ[i](a)N(a)2k𝒪Kk0(1)N(k)N(k)s(ε)(ε)(12+ε)N(a)2s𝒥(s)w(u+s)w(v+s)×U1u+sU2v+sXs(u+s)(v+s)L2(1+u,χik1)L2(1+v,χik1)Z2(12+u,12+v,a,k)dsdudv.\displaystyle\begin{split}S_{1}(k\neq 0)=&2X\sum_{\begin{subarray}{c}a\equiv 1\bmod{(1+i)^{3}}\\ N(a)\leq Y\end{subarray}}\frac{\mu_{[i]}(a)}{N(a)^{2}}\sum_{\begin{subarray}{c}k\in\mathcal{O}_{K}\\ k\neq 0\end{subarray}}(-1)^{N(k)}\frac{\pi}{(2\pi i)^{3}}\int\limits_{(\varepsilon)}\int\limits_{(\varepsilon)}\int\limits_{(\tfrac{1}{2}+\varepsilon)}\left(\frac{N(a)^{2}}{N(k)}\right)^{s}\mathcal{J}(s)w(u+s)w(v+s)\\ &\times\frac{U_{1}^{u+s}U_{2}^{v+s}X^{-s}}{(u+s)(v+s)}Z(\tfrac{1}{2}+u,\tfrac{1}{2}+v,a,k)\ ds\ du\ dv\\ =&\frac{2\pi X}{(2\pi i)^{3}}\sum_{\begin{subarray}{c}a\equiv 1\bmod{(1+i)^{3}}\\ N(a)\leq Y\end{subarray}}\frac{\mu_{[i]}(a)}{N(a)^{2}}\sum_{\begin{subarray}{c}k\in\mathcal{O}_{K}\\ k\neq 0\end{subarray}}\frac{(-1)^{N(k)}}{N(k)^{s}}\int\limits_{(\varepsilon)}\int\limits_{(\varepsilon)}\int\limits_{(\tfrac{1}{2}+\varepsilon)}N(a)^{2s}\mathcal{J}(s)w(u+s)w(v+s)\\ &\quad\times\frac{U_{1}^{u+s}U_{2}^{v+s}X^{-s}}{(u+s)(v+s)}L^{2}(1+u,\chi_{ik_{1}})L^{2}(1+v,\chi_{ik_{1}})Z_{2}(\tfrac{1}{2}+u,\tfrac{1}{2}+v,a,k)\ ds\ du\ dv.\end{split}

We observe that if we move the lines of integrations over u,vu,v in the last expression in (5.11) to the left, then we encounter poles at u=v=0u=v=0 only when k1=±ik_{1}=\pm i. For this reason, we further write S1(k0)=S1(k1=±i)+S1(k1±i)S_{1}(k\neq 0)=S_{1}(k_{1}=\pm i)+S_{1}(k_{1}\neq\pm i), where

(5.12) S1(k1=±i)=2πX(2πi)3a1mod(1+i)3N(a)Yμ[i](a)N(a)2k2G(1)N(k2)N(k2)2s(ε)(ε)(12+ε)N(a)2s𝒥(s)w(u+s)w(v+s)×U1u+sU2v+sXs(u+s)(v+s)ζK2(1+u)ζK2(1+v)Z2(12+u,12+v,a,±ik22)dsdudv,\displaystyle\begin{split}S_{1}(k_{1}=\pm i)=&\frac{2\pi X}{(2\pi i)^{3}}\sum_{\begin{subarray}{c}a\equiv 1\bmod{(1+i)^{3}}\\ N(a)\leq Y\end{subarray}}\frac{\mu_{[i]}(a)}{N(a)^{2}}\sum_{\begin{subarray}{c}k_{2}\in G\end{subarray}}\frac{(-1)^{N(k_{2})}}{N(k_{2})^{2s}}\int\limits_{(\varepsilon)}\int\limits_{(\varepsilon)}\int\limits_{(\tfrac{1}{2}+\varepsilon)}N(a)^{2s}\mathcal{J}(s)w(u+s)w(v+s)\\ &\times\frac{U_{1}^{u+s}U_{2}^{v+s}X^{-s}}{(u+s)(v+s)}\zeta^{2}_{K}(1+u)\zeta_{K}^{2}(1+v)Z_{2}(\tfrac{1}{2}+u,\tfrac{1}{2}+v,a,\pm ik_{2}^{2})\ ds\ du\ dv,\end{split}

and

(5.13) S1(k1±i)=2πX(2πi)3a1mod(1+i)3N(a)Yμ[i](a)N(a)2k𝒪Kk0,k1±i(1)N(k)N(k)s(ε)(ε)(12+ε)N(a)2s𝒥(s)w(u+s)w(v+s)×U1u+sU2v+sXs(u+s)(v+s)L2(1+u,χik1)L2(1+v,χik1)Z2(12+u,12+v,a,k)dsdudv.\displaystyle\begin{split}S_{1}(k_{1}\neq\pm i)=&\frac{2\pi X}{(2\pi i)^{3}}\sum_{\begin{subarray}{c}a\equiv 1\bmod{(1+i)^{3}}\\ N(a)\leq Y\end{subarray}}\frac{\mu_{[i]}(a)}{N(a)^{2}}\sum_{\begin{subarray}{c}k\in\mathcal{O}_{K}\\ k\neq 0,k_{1}\neq\pm i\end{subarray}}\frac{(-1)^{N(k)}}{N(k)^{s}}\int\limits_{(\varepsilon)}\int\limits_{(\varepsilon)}\int\limits_{(\tfrac{1}{2}+\varepsilon)}N(a)^{2s}\mathcal{J}(s)w(u+s)w(v+s)\\ &\quad\times\frac{U_{1}^{u+s}U_{2}^{v+s}X^{-s}}{(u+s)(v+s)}L^{2}(1+u,\chi_{ik_{1}})L^{2}(1+v,\chi_{ik_{1}})Z_{2}(\tfrac{1}{2}+u,\tfrac{1}{2}+v,a,k)\ ds\ du\ dv.\end{split}

5.3. Computing S1S_{1}: the term S1(k=0)S_{1}(k=0)

Note that by Lemma 2.3 we have g(0,n)=φ[i](n)g(0,n)=\varphi_{[i]}(n) if n=n=\square, and 0 otherwise. Thus we get

(5.14) S1(k=0)=\displaystyle S_{1}(k=0)= 2Xn1,n21mod(1+i)3n1n2=d[i](n1)d[i](n2)N(n1n2)12φ[i](n1n2)N(n1n2)a1mod(1+i)3(a,n1n2)=1N(a)Yμ[i](a)N(a)2h~(0,n1,n2).\displaystyle 2X\sum_{\begin{subarray}{c}n_{1},n_{2}\equiv 1\bmod{(1+i)^{3}}\\ n_{1}n_{2}=\square\end{subarray}}\frac{d_{[i]}(n_{1})d_{[i]}(n_{2})}{N(n_{1}n_{2})^{\frac{1}{2}}}\frac{\varphi_{[i]}(n_{1}n_{2})}{N(n_{1}n_{2})}\sum_{\begin{subarray}{c}a\equiv 1\bmod{(1+i)^{3}}\\ (a,n_{1}n_{2})=1\\ N(a)\leq Y\end{subarray}}\frac{\mu_{[i]}(a)}{N(a)^{2}}\widetilde{h}\Big{(}0,n_{1},n_{2}\Big{)}.

We note that

a1mod(1+i)3(a,n1n2)=1N(a)Yμ[i](a)N(a)2=43ζK(2)ϖ1mod(1+i)3ϖ|n1n2(11N(ϖ)2)1+O(Y1).\displaystyle\sum_{\begin{subarray}{c}a\equiv 1\bmod{(1+i)^{3}}\\ (a,n_{1}n_{2})=1\\ N(a)\leq Y\end{subarray}}\frac{\mu_{[i]}(a)}{N(a)^{2}}=\frac{4}{3\zeta_{K}(2)}\prod_{\begin{subarray}{c}\varpi\equiv 1\bmod{(1+i)^{3}}\\ \varpi|n_{1}n_{2}\end{subarray}}\left(1-\frac{1}{N(\varpi)^{2}}\right)^{-1}+O\left(Y^{-1}\right).

Applying this to (5.14), we see that

S1(k=0)=8X3ζK(2)n1,n21mod(1+i)3n1n2=d[i](n1)d[i](n2)N(n1n2)12𝒫(n1n2)h~(0,n1,n2)+O(XYn1,n21mod(1+i)3n1n2=d[i](n1)d[i](n2)N(n1n2)12h~(0,n1,n2)),\displaystyle\begin{split}S_{1}(k=0)=\frac{8X}{3\zeta_{K}(2)}\sum_{\begin{subarray}{c}n_{1},n_{2}\equiv 1\bmod{(1+i)^{3}}\\ n_{1}n_{2}=\square\end{subarray}}\frac{d_{[i]}(n_{1})d_{[i]}(n_{2})}{N(n_{1}n_{2})^{\frac{1}{2}}}\mathcal{P}(n_{1}n_{2})\widetilde{h}\Big{(}0,n_{1},n_{2}\Big{)}\\ +O\Big{(}\frac{X}{Y}\sum_{\begin{subarray}{c}n_{1},n_{2}\equiv 1\bmod{(1+i)^{3}}\\ n_{1}n_{2}=\square\end{subarray}}\frac{d_{[i]}(n_{1})d_{[i]}(n_{2})}{N(n_{1}n_{2})^{\frac{1}{2}}}\widetilde{h}\Big{(}0,n_{1},n_{2}\Big{)}\Big{)},\end{split}

where we recall that 𝒫(n1n2)\mathcal{P}(n_{1}n_{2}) is defined in (2.20).

Due to the rapid decay of h~\widetilde{h}, we can estimate the error term above as

n1,n21mod(1+i)3n1n2=d[i](n1)d[i](n2)N(n1n2)12h~(0,n1,n2)n1,n21mod(1+i)3n1n2=N(n1),N(n2)X1+εd[i](n1)d[i](n2)N(n1n2)12.\displaystyle\sum_{\begin{subarray}{c}n_{1},n_{2}\equiv 1\bmod{(1+i)^{3}}\\ n_{1}n_{2}=\square\end{subarray}}\frac{d_{[i]}(n_{1})d_{[i]}(n_{2})}{N(n_{1}n_{2})^{\frac{1}{2}}}\widetilde{h}\Big{(}0,n_{1},n_{2}\Big{)}\ll\sum_{\begin{subarray}{c}n_{1},n_{2}\equiv 1\bmod{(1+i)^{3}}\\ n_{1}n_{2}=\square\\ N(n_{1}),N(n_{2})\ll X^{1+\varepsilon}\end{subarray}}\frac{d_{[i]}(n_{1})d_{[i]}(n_{2})}{N(n_{1}n_{2})^{\frac{1}{2}}}.

We now denote d=(n1,n2)d=(n_{1},n_{2}) with dd being primary together with a change of variables: njdnj,j=1,2n_{j}\rightarrow dn_{j},j=1,2 to see that for the new variables n1,n2n_{1},n_{2}, we have (n1,n2)=1(n_{1},n_{2})=1 and the condition that d2n1n2=d^{2}n_{1}n_{2}=\square further implies that both nj,j=1,2n_{j},j=1,2 are squares now so that we have

n1,n21mod(1+i)3n1n2=N(n1),N(n2)X1+εd[i](n1)d[i](n2)N(n1n2)12d1mod(1+i)3N(d)X1+εn1,n21mod(1+i)3n1,n2=N(n1),N(n2)X1+ε/N(d)d[i](dn1)d[i](dn2)N(d2n1n2)12\displaystyle\sum_{\begin{subarray}{c}n_{1},n_{2}\equiv 1\bmod{(1+i)^{3}}\\ n_{1}n_{2}=\square\\ N(n_{1}),N(n_{2})\ll X^{1+\varepsilon}\end{subarray}}\frac{d_{[i]}(n_{1})d_{[i]}(n_{2})}{N(n_{1}n_{2})^{\frac{1}{2}}}\ll\sum_{\begin{subarray}{c}d\equiv 1\bmod{(1+i)^{3}}\\ N(d)\ll X^{1+\varepsilon}\end{subarray}}\sum_{\begin{subarray}{c}n_{1},n_{2}\equiv 1\bmod{(1+i)^{3}}\\ n_{1},n_{2}=\square\\ N(n_{1}),N(n_{2})\ll X^{1+\varepsilon}/N(d)\end{subarray}}\frac{d_{[i]}(dn_{1})d_{[i]}(dn_{2})}{N(d^{2}n_{1}n_{2})^{\frac{1}{2}}}
\displaystyle\ll d1mod(1+i)3N(d)X1+εd[i](d)2N(d)n1,n21mod(1+i)3n1,n2=N(n1),N(n2)X1+ε/N(d)d[i](n1)d[i](n2)N(n1n2)12\displaystyle\sum_{\begin{subarray}{c}d\equiv 1\bmod{(1+i)^{3}}\\ N(d)\ll X^{1+\varepsilon}\end{subarray}}\frac{d_{[i]}(d)^{2}}{N(d)}\sum_{\begin{subarray}{c}n_{1},n_{2}\equiv 1\bmod{(1+i)^{3}}\\ n_{1},n_{2}=\square\\ N(n_{1}),N(n_{2})\ll X^{1+\varepsilon}/N(d)\end{subarray}}\frac{d_{[i]}(n_{1})d_{[i]}(n_{2})}{N(n_{1}n_{2})^{\frac{1}{2}}}
\displaystyle\ll d1mod(1+i)3N(d)X1+εd[i](d)2N(d)(n1mod(1+i)3N(n)2X1+ε/N(d)d[i](n2)N(n))2(logX)6d1mod(1+i)3N(d)X1+εd[i](d)2N(d)(logX)10.\displaystyle\sum_{\begin{subarray}{c}d\equiv 1\bmod{(1+i)^{3}}\\ N(d)\ll X^{1+\varepsilon}\end{subarray}}\frac{d_{[i]}(d)^{2}}{N(d)}\Big{(}\sum_{\begin{subarray}{c}n\equiv 1\bmod{(1+i)^{3}}\\ N(n)^{2}\ll X^{1+\varepsilon}/N(d)\end{subarray}}\frac{d_{[i]}(n^{2})}{N(n)}\Big{)}^{2}\ll(\log X)^{6}\sum_{\begin{subarray}{c}d\equiv 1\bmod{(1+i)^{3}}\\ N(d)\ll X^{1+\varepsilon}\end{subarray}}\frac{d_{[i]}(d)^{2}}{N(d)}\ll(\log X)^{10}.

where the last line follows by applying estimations similar to that given in (5.8).

Further note that we have

Φ(N(x+yi))dxdy=02π0Φ(r2)r𝑑r𝑑θ=πΦ^(1).\displaystyle\int\limits^{\infty}_{-\infty}\int\limits^{\infty}_{-\infty}\Phi\left(N(x+yi)\right)\mathrm{d}x\mathrm{d}y=\int^{2\pi}_{0}\int^{\infty}_{0}\Phi(r^{2})rdrd\theta=\pi\widehat{\Phi}(1).

We now conclude from the above discussions that

(5.15) S1(k=0)=8X3ζK(2)n1,n21mod(1+i)3n1n2=d[i](n1)d[i](n2)N(n1n2)12𝒫(n1n2)h~(0,n1,n2)+O(XY(logX)10)=8πΦ^(1)X3ζK(2)1(2πi)2(2)(2)w(u)w(v)uvU1uU2vZ(12+u,12+v)𝑑u𝑑v+O(XY(logX)10),\displaystyle\begin{split}S_{1}(k=0)=&\frac{8X}{3\zeta_{K}(2)}\sum_{\begin{subarray}{c}n_{1},n_{2}\equiv 1\bmod{(1+i)^{3}}\\ n_{1}n_{2}=\square\end{subarray}}\frac{d_{[i]}(n_{1})d_{[i]}(n_{2})}{N(n_{1}n_{2})^{\frac{1}{2}}}\mathcal{P}(n_{1}n_{2})\widetilde{h}\Big{(}0,n_{1},n_{2}\Big{)}+O\left(\frac{X}{Y}(\log X)^{10}\right)\\ =&\frac{8\pi\widehat{\Phi}(1)X}{3\zeta_{K}(2)}\frac{1}{(2\pi i)^{2}}\int\limits\limits_{(2)}\int\limits\limits_{(2)}\frac{w(u)w(v)}{uv}U_{1}^{u}U_{2}^{v}Z(\tfrac{1}{2}+u,\tfrac{1}{2}+v)dudv+O\left(\frac{X}{Y}(\log X)^{10}\right),\end{split}

where Z(12+u,12+v)Z(\tfrac{1}{2}+u,\tfrac{1}{2}+v) is defined as in (2.19).

It follows from (5.15) and Lemma 2.10 that

(5.16) S1(k=0)=8πΦ^(1)X3ζK(2)1(2πi)2(2)(2)U1uU2vuv(2u)3(2v)3(u+v)4(u,v)𝑑u𝑑v+O(XY1(logX)10),\displaystyle S_{1}(k=0)=\frac{8\pi\widehat{\Phi}(1)X}{3\zeta_{K}(2)}\frac{1}{(2\pi i)^{2}}\int\limits\limits_{(2)}\int\limits\limits_{(2)}\frac{U_{1}^{u}U_{2}^{v}}{uv(2u)^{3}(2v)^{3}(u+v)^{4}}\mathcal{E}(u,v)\ du\ dv+O\left(XY^{-1}(\log X)^{10}\right),

where

(u,v)=w(u)w(v)ζK3(1+2u)(2u)3ζK3(1+2v)(2v)3ζK4(1+u+v)(u+v)4Z1(12+u,12+v).\mathcal{E}(u,v)=w(u)w(v)\zeta_{K}^{3}(1+2u)(2u)^{3}\zeta_{K}^{3}(1+2v)(2v)^{3}\zeta_{K}^{4}(1+u+v)(u+v)^{4}Z_{1}(\tfrac{1}{2}+u,\tfrac{1}{2}+v).

Applying Lemma 2.10 again, we see that \mathcal{E} is analytic for (u),(v)>14+ε\Re(u),\Re(v)>-\frac{1}{4}+\varepsilon.

We first move the lines of the integrations in (5.16) to (u)=(v)=110\Re(u)=\Re(v)=\frac{1}{10} by noting that we encounter no poles. We then move the line of the integration over vv to (v)=15\Re(v)=-\frac{1}{5} to see that we encounter two poles of order 44 at v=0v=0 and v=uv=-u in the process. It follows that

(5.17) 1(2πi)2(110)(110)U1uU2vuv(2u)3(2v)3(u+v)4(u,v)𝑑u𝑑v=12πi(110)(Resv=0+Resv=u)(U1uU2vuv(2u)3(2v)3(u+v)4(u,v))𝑑u+O(U1110U215).\displaystyle\begin{split}&\frac{1}{(2\pi i)^{2}}\int\limits_{(\frac{1}{10})}\int\limits_{(\frac{1}{10})}\frac{U_{1}^{u}U_{2}^{v}}{uv(2u)^{3}(2v)^{3}(u+v)^{4}}\mathcal{E}(u,v)\ du\ dv\\ &=\frac{1}{2\pi i}\int\limits_{(\frac{1}{10})}\left(\underset{v=0}{\operatorname{Res}}+\underset{v=-u}{\operatorname{Res}}\right)\left(\frac{U_{1}^{u}U_{2}^{v}}{uv(2u)^{3}(2v)^{3}(u+v)^{4}}\mathcal{E}(u,v)\right)du+O\left(U_{1}^{\frac{1}{10}}U_{2}^{-\frac{1}{5}}\right).\end{split}

We treat the contribution from the residue at v=0v=0 in (5.17) to see that

I1(u)\displaystyle I_{1}(u) =Resv=0(U1uU2vuv(2u)3(2v)3(u+v)4(u,v))\displaystyle=\underset{v=0}{\operatorname{Res}}\left(\frac{U_{1}^{u}U_{2}^{v}}{uv(2u)^{3}(2v)^{3}(u+v)^{4}}\mathcal{E}(u,v)\right)
=U1u263!u11((u,0)(u3(logU2)312u2(logU2)2+60ulogU2120)\displaystyle=\frac{U_{1}^{u}}{2^{6}\cdot 3!u^{11}}\Big{(}\mathcal{E}(u,0)(u^{3}(\log U_{2})^{3}-12u^{2}(\log U_{2})^{2}+60u\log U_{2}-120)
+(0,1)(u,0)(3u3(logU2)224u2logU2+60u)+(0,2)(u,0)(3u3logU212u2)+(0,3)(u,0)u3),\displaystyle\quad+\mathcal{E}^{(0,1)}(u,0)(3u^{3}(\log U_{2})^{2}-24u^{2}\log U_{2}+60u)+\mathcal{E}^{(0,2)}(u,0)(3u^{3}\log U_{2}-12u^{2})+\mathcal{E}^{(0,3)}(u,0)u^{3}\Big{)},

where (i,j)(u,v)=i+juivj(u,v)\mathcal{E}^{(i,j)}(u,v)=\frac{\partial^{i+j}\mathcal{E}}{\partial u^{i}\partial v^{j}}(u,v).

It follows by moving the line of the integration over uu from (u)=110\Re(u)=\frac{1}{10} to (u)=110\Re(u)=-\frac{1}{10} that we have

12πi(110)I1(u)𝑑u=Resu=0I1(u)+O(U1110log3X)=5!(0,0)263!10!((logU1)10+5(logU1)9logU29(logU1)8(logU2)2+6(logU1)7(logU2)3)+O((logX)9+U1110(logX)3).\displaystyle\begin{split}\frac{1}{2\pi i}\int\limits_{(\frac{1}{10})}I_{1}(u)du&=\underset{u=0}{\operatorname{Res}}\ I_{1}(u)+O(U_{1}^{-\frac{1}{10}}\log^{3}X)\\ &=\frac{5!\mathcal{E}(0,0)}{2^{6}\cdot 3!\cdot 10!}\left(-(\log U_{1})^{10}+5(\log U_{1})^{9}\log U_{2}-9(\log U_{1})^{8}(\log U_{2})^{2}+6(\log U_{1})^{7}(\log U_{2})^{3}\right)\\ &\quad+O\left((\log X)^{9}+U_{1}^{-\frac{1}{10}}(\log X)^{3}\right).\end{split}

Similarly, we have that

12πi(110)Resv=u(U1uU2vuv(2u)3(2v)3(u+v)4(u,v))𝑑uU1110U2110(logX)3.\displaystyle\frac{1}{2\pi i}\int\limits_{(\frac{1}{10})}\underset{v=-u}{\operatorname{Res}}\left(\frac{U_{1}^{u}U_{2}^{v}}{uv(2u)^{3}(2v)^{3}(u+v)^{4}}\mathcal{E}(u,v)\right)du\ll U_{1}^{\frac{1}{10}}U_{2}^{-\frac{1}{10}}(\log X)^{3}.

Combining (5.16)-(5.17), together with the observation that lims0ζK(1+s)s=π/4\displaystyle\lim_{s\rightarrow 0}\zeta_{K}(1+s)s=\pi/4 and Z1(12,12)=4a4Z_{1}(\frac{1}{2},\frac{1}{2})=4a_{4}, we obtain that

(5.18) S1(k=0)=8πX3ζK(2)Φ^(1)(π4)1045!a4263!10!((logU1)10+5(logU1)9logU29(logU1)8(logU2)2+6(logU1)7(logU2)3)+O(X(logX)9+XY1(logX)10).\displaystyle\begin{split}&S_{1}(k=0)\\ =&\frac{8\pi X}{3\zeta_{K}(2)}\widehat{\Phi}(1)\cdot\left(\frac{\pi}{4}\right)^{10}\frac{4\cdot 5!a_{4}}{2^{6}\cdot 3!\cdot 10!}\left(-(\log U_{1})^{10}+5(\log U_{1})^{9}\log U_{2}-9(\log U_{1})^{8}(\log U_{2})^{2}+6(\log U_{1})^{7}(\log U_{2})^{3}\right)\\ &+O\left(X(\log X)^{9}+XY^{-1}(\log X)^{10}\right).\end{split}

5.4. Computing S1S_{1}: the term S1(k1=±i)S_{1}(k_{1}=\pm i)

In this section, we evaluate S1(k1=±i)S_{1}(k_{1}=\pm i) given in (5.12). Applying Lemma 2.12, we see that

S1(k1=±i)=\displaystyle S_{1}(k_{1}=\pm i)= 2πX(2πi)3a1mod(1+i)3N(a)Yμ[i](a)N(a)2k2[i](1)N(k2)(ε)(ε)(12+ε)N(a)2s𝒥(s)w(u+s)w(v+s)U1u+sU2v+sXs(u+s)(v+s)\displaystyle\frac{2\pi X}{(2\pi i)^{3}}\sum_{\begin{subarray}{c}a\equiv 1\bmod{(1+i)^{3}}\\ N(a)\leq Y\end{subarray}}\frac{\mu_{[i]}(a)}{N(a)^{2}}\sum_{\begin{subarray}{c}k_{2}\in\mathbb{Z}[i]\end{subarray}}(-1)^{N(k_{2})}\int\limits_{(\varepsilon)}\int\limits_{(\varepsilon)}\int\limits_{(\tfrac{1}{2}+\varepsilon)}N(a)^{2s}\mathcal{J}(s)w(u+s)w(v+s)\frac{U_{1}^{u+s}U_{2}^{v+s}X^{-s}}{(u+s)(v+s)}
×ζK2(1+u)ζK2(1+v)(212s1)ζK(2s)ζK2(1+2u+2s)ζK2(1+2v+2s)Z3(12+u,12+v,s,a)dsdudv.\displaystyle\times\zeta^{2}_{K}(1+u)\zeta^{2}_{K}(1+v)(2^{1-2s}-1)\zeta_{K}(2s)\zeta_{K}^{2}(1+2u+2s)\zeta_{K}^{2}(1+2v+2s)Z_{3}(\tfrac{1}{2}+u,\tfrac{1}{2}+v,s,a)\ ds\ du\ dv.

As Z3(12+u,12+v,s)Z_{3}(\tfrac{1}{2}+u,\tfrac{1}{2}+v,s) is analytic in the region (u),(v)ε,(s)2ε\Re(u),\Re(v)\geq\varepsilon,\Re(s)\geq 2\varepsilon by (1) of Lemma 2.12 and (212s1)ζK(2s)(2^{1-2s}-1)\zeta_{K}(2s) is analytic at s=12s=\frac{1}{2}, we move the lines of the integration above to (u)=(v)=1,(s)=110\Re(u)=\Re(v)=1,\Re(s)=\frac{1}{10} without encountering any poles to see that

(5.19) S1(k1=±i)=2πX(2πi)3a1mod(1+i)3N(a)Yμ[i](a)N(a)2(1)(1)(110)N(a)2s𝒥(s)w(u+s)w(v+s)U1u+sU2v+sXs(u+s)(v+s)×ζK2(1+u)ζK2(1+v)(212s1)ζK(2s)ζK2(1+2u+2s)ζK2(1+2v+2s)Z3(12+u,12+v,s,a)dsdudv.\displaystyle\begin{split}S_{1}(k_{1}=\pm i)=&\frac{2\pi X}{(2\pi i)^{3}}\sum_{\begin{subarray}{c}a\equiv 1\bmod{(1+i)^{3}}\\ N(a)\leq Y\end{subarray}}\frac{\mu_{[i]}(a)}{N(a)^{2}}\int\limits_{(1)}\int\limits_{(1)}\int\limits_{(\frac{1}{10})}N(a)^{2s}\mathcal{J}(s)w(u+s)w(v+s)\frac{U_{1}^{u+s}U_{2}^{v+s}X^{-s}}{(u+s)(v+s)}\\ &\times\zeta^{2}_{K}(1+u)\zeta^{2}_{K}(1+v)(2^{1-2s}-1)\zeta_{K}(2s)\zeta_{K}^{2}(1+2u+2s)\zeta_{K}^{2}(1+2v+2s)Z_{3}(\tfrac{1}{2}+u,\tfrac{1}{2}+v,s,a)\ ds\ du\ dv.\end{split}

We extend the sum over aa in (5.19) to include all primary elements in 𝒪K\mathcal{O}_{K}, introducing an error term

(5.20) 2πX(2πi)3a1mod(1+i)3N(a)>Yμ[i](a)N(a)2(1)(1)(110)N(a)2s𝒥(s)w(u+s)w(v+s)U1u+sU2v+sXs(u+s)(v+s)×ζK2(1+u)ζK2(1+v)(212s1)ζK(2s)ζK2(1+2u+2s)ζK2(1+2v+2s)Z3(12+u,12+v,s,a)dsdudv.\displaystyle\begin{split}&\frac{2\pi X}{(2\pi i)^{3}}\sum_{\begin{subarray}{c}a\equiv 1\bmod{(1+i)^{3}}\\ N(a)>Y\end{subarray}}\frac{\mu_{[i]}(a)}{N(a)^{2}}\int\limits_{(1)}\int\limits_{(1)}\int\limits_{(\frac{1}{10})}N(a)^{2s}\mathcal{J}(s)w(u+s)w(v+s)\frac{U_{1}^{u+s}U_{2}^{v+s}X^{-s}}{(u+s)(v+s)}\\ \times&\zeta^{2}_{K}(1+u)\zeta^{2}_{K}(1+v)(2^{1-2s}-1)\zeta_{K}(2s)\zeta_{K}^{2}(1+2u+2s)\zeta_{K}^{2}(1+2v+2s)Z_{3}(\tfrac{1}{2}+u,\tfrac{1}{2}+v,s,a)\ ds\ du\ dv.\end{split}

To facilitate our estimation of the triple integral in the above expression and other similar integrals in what follows, we gather here a few bounds on ζK(s)\zeta_{K}(s) that hold uniformly in specified regions. On write s=σ+its=\sigma+it, we have

(5.21) ζK(s)(1+(|t|+4)1σ)min(1|σ1|,log(|t|+4)),σ>1,ζK(s)(1+|s|2)1σ/2+ε,0σ1,1ζK(s)log(|t|+4),1σ2.\displaystyle\begin{split}\zeta_{K}(s)\ll&\big{(}1+(|t|+4)^{1-\sigma}\big{)}\min\big{(}\frac{1}{|\sigma-1|},\log(|t|+4)\big{)},\quad\sigma>1,\\ \zeta_{K}(s)\ll&\left(1+|s|^{2}\right)^{1-\sigma/2+\varepsilon},\quad 0\leq\sigma\leq 1,\\ \frac{1}{\zeta_{K}(s)}\ll&\log(|t|+4),\quad 1\leq\sigma\leq 2.\end{split}

The first and third estimation above can be established similar to the proofs of [MVa1, Corollary 1.17] and [MVa1, Lemma 6.7], respectively. The second estimation above is the convexity bound for ζK(s)\zeta_{K}(s) (see [iwakow, Exercise 3, p. 100]).

Also, by applying (3.7), we see that for the jj-th derivative of ζK(s)\zeta_{K}(s) with j1j\geq 1, we have for (s)>1\Re(s)>1,

ζK(j)(s)=\displaystyle\zeta^{(j)}_{K}(s)= (n)N(n)>1(logN(n))jN(n)s=1(logu)jusd((n)N(n)u1)\displaystyle\sum_{\begin{subarray}{c}(n)\\ N(n)>1\end{subarray}}(-\log N(n))^{j}N(n)^{-s}=\int^{\infty}_{1}(-\log u)^{j}u^{-s}d(\sum_{\begin{subarray}{c}(n)\\ N(n)\leq u\end{subarray}}1)
=\displaystyle= 1(logu)jusd(π4u+O(uθ))\displaystyle\int^{\infty}_{1}(-\log u)^{j}u^{-s}d(\frac{\pi}{4}u+O(u^{\theta}))
=\displaystyle= π4j!(1s)j+O(1(s+jlogu)(logu)jus1+θ𝑑u).\displaystyle\frac{\pi}{4}\frac{j!}{(1-s)^{j}}+O\big{(}\int^{\infty}_{1}(s+\frac{j}{\log u})(-\log u)^{j}u^{-s-1+\theta}du\big{)}.

We deduce readily from the above that for j1j\geq 1 and (s)>0.4\Re(s)>0.4, we have

(5.22) ζK(j)(s)1+|s|.\displaystyle\zeta^{(j)}_{K}(s)\ll 1+|s|.

We further note that integrating by parts implies that for (s)<1\Re(s)<1 and any integer ν1\nu\geq 1,

(5.23) Φ^(1s)ν3|(s)||s||s1|ν1Φ(ν),\displaystyle\widehat{\Phi}(1-s)\ll_{\nu}\frac{3^{|\Re(s)|}}{|s||s-1|^{\nu-1}}\Phi_{(\nu)},

where

Φ(ν)=max0jν1252|Φ(j)(t)|𝑑t.\Phi_{(\nu)}=\underset{0\leq j\leq\nu}{\operatorname{max}}\int_{\frac{1}{2}}^{\frac{5}{2}}|\Phi^{(j)}(t)|dt.

We move the lines of the integrations in (5.20) over u,vu,v to (u)=(v)=1logX,(s)=2logX\Re(u)=\Re(v)=\frac{1}{\log X},\Re(s)=\frac{2}{\log X} without encountering any poles. Then by (2) of Lemma 2.12 and the estimations given in (5.21) and (5.23), we see that on the new lines of integrations, the expression in (5.20) is

X(logX)20a>Y(a,2)=11a24logX(1logX)(1logX)(2logX)(1+|2s|)1+ε|𝒥(s)||Γ(u+s+12)2Γ(v+s+12)2||ds||du||dv|\displaystyle\ll X(\log X)^{20}\sum_{\begin{subarray}{c}a>Y\\ (a,2)=1\end{subarray}}\frac{1}{a^{2-\frac{4}{\log X}}}\int\limits_{(\frac{1}{\log X})}\int\limits_{(\frac{1}{\log X})}\int\limits_{(\frac{2}{\log X})}\ (1+|2s|)^{1+\varepsilon}|\mathcal{J}(s)|\left|\Gamma(u+s+\tfrac{1}{2})^{2}\Gamma(v+s+\tfrac{1}{2})^{2}\right|\ |ds|\ |du|\ |dv|
X(logX)20Y1(2logX)(1+|2s|)1+ε|Γ(s)Γ(1s)||Φ^(1s)||ds|XY1(logX)22Φ(4),\displaystyle\ll X(\log X)^{20}Y^{-1}\int_{(\frac{2}{\log X})}(1+|2s|)^{1+\varepsilon}|\frac{\Gamma(s)}{\Gamma(1-s)}||\widehat{\Phi}(1-s)||ds|\ll XY^{-1}(\log X)^{22}\Phi_{(4)},

where the last estimation above follows from (5.23) with ν=4\nu=4 and the bound Γ(s)Γ(1s)|s|2(s)1\frac{\Gamma(s)}{\Gamma(1-s)}\ll|s|^{2\Re(s)-1}.

We conclude from the above discussions and (5.19) that

(5.24) S1(k1=±i)=2πX(2πi)3a1mod(1+i)3μ[i](a)N(a)2(ε)(ε)(12+ε)N(a)2s𝒥(s)w(u+s)w(v+s)U1u+sU2v+sXs(u+s)(v+s)×ζK2(1+u)ζK2(1+v)(212s1)ζK(2s)ζK2(1+2u+2s)ζK2(1+2v+2s)Z3(12+u,12+v,s,a)dsdudv+O(XY1(logX)22Φ(4)).\displaystyle\begin{split}&S_{1}(k_{1}=\pm i)\\ =&\frac{2\pi X}{(2\pi i)^{3}}\sum_{\begin{subarray}{c}a\equiv 1\bmod{(1+i)^{3}}\end{subarray}}\frac{\mu_{[i]}(a)}{N(a)^{2}}\int\limits_{(\varepsilon)}\int\limits_{(\varepsilon)}\int\limits_{(\tfrac{1}{2}+\varepsilon)}N(a)^{2s}\mathcal{J}(s)w(u+s)w(v+s)\frac{U_{1}^{u+s}U_{2}^{v+s}X^{-s}}{(u+s)(v+s)}\\ &\times\zeta^{2}_{K}(1+u)\zeta^{2}_{K}(1+v)(2^{1-2s}-1)\zeta_{K}(2s)\zeta_{K}^{2}(1+2u+2s)\zeta_{K}^{2}(1+2v+2s)Z_{3}(\tfrac{1}{2}+u,\tfrac{1}{2}+v,s,a)\ ds\ du\ dv\\ &+O\left(XY^{-1}(\log X)^{22}\Phi_{(4)}\right).\end{split}

We further apply Lemma 2.13 to deduce from (5.24) that

(5.25) S1(k1=±i)=2πX(2πi)3(ε)(ε)(12+ε)𝒥(s)(212s1)ζK(2s)w(u+s)w(v+s)U1u+sU2v+sXs(u+s)(v+s)×ζK2(1+u)ζK2(1+v)ζK3(1+2u+2s)ζK3(1+2v+2s)ζK4(1+u+v+2s)ζK2(1+u+2s)ζK2(1+v+2s)×Z4(12+u,12+v,s)dsdudv+O(XY1(logX)22Φ(4)).\displaystyle\begin{split}&S_{1}(k_{1}=\pm i)\\ =&\frac{2\pi X}{(2\pi i)^{3}}\int\limits_{(\varepsilon)}\int\limits_{(\varepsilon)}\int\limits_{(\tfrac{1}{2}+\varepsilon)}\mathcal{J}(s)(2^{1-2s}-1)\zeta_{K}(2s)w(u+s)w(v+s)\frac{U_{1}^{u+s}U_{2}^{v+s}X^{-s}}{(u+s)(v+s)}\\ &\times\zeta_{K}^{2}(1+u)\zeta_{K}^{2}(1+v)\frac{\zeta_{K}^{3}(1+2u+2s)\zeta_{K}^{3}(1+2v+2s)\zeta_{K}^{4}(1+u+v+2s)}{\zeta_{K}^{2}(1+u+2s)\zeta_{K}^{2}(1+v+2s)}\\ &\times Z_{4}(\tfrac{1}{2}+u,\tfrac{1}{2}+v,s)\ ds\ du\ dv+O\left(XY^{-1}(\log X)^{22}\Phi_{(4)}\right).\end{split}

We now move the lines of the integrations in (5.25) to (u)=(v)=(s)=1100\Re(u)=\Re(v)=\Re(s)=\frac{1}{100} without encountering any poles, as Z4(12+u,12+v,s)Z_{4}(\tfrac{1}{2}+u,\tfrac{1}{2}+v,s) is analytic and uniformly bounded in the region (u),(v)18,116(s)18\Re(u),\Re(v)\geq-\frac{1}{8},-\frac{1}{16}\leq\Re(s)\leq\frac{1}{8} by Lemma 2.13. Next, we move the line of the integration over vv to (v)=150+1logX\Re(v)=-\frac{1}{50}+\frac{1}{\log X} to encounter a pole of order 22 at v=0v=0 and a pole of order 44 at v=sv=-s to see that triple integral in (5.25) is

(5.26) 1(2πi)2(1100)(1100)(I2(u,s)+I3(u,s))𝑑u𝑑s+O(U1150U21100X1100Φ(4)),\displaystyle\frac{1}{(2\pi i)^{2}}\int\limits_{(\frac{1}{100})}\int\limits_{(\frac{1}{100})}(I_{2}(u,s)+I_{3}(u,s))\ du\ ds+O\left(U_{1}^{\frac{1}{50}}U_{2}^{-\frac{1}{100}}X^{-\frac{1}{100}}\Phi_{(4)}\right),

where the error term above follows from using estimations given in (5.21) and (5.23) and where we denote I2(u,s),I3(u,s)I_{2}(u,s),I_{3}(u,s) for the residues of the integrand in (5.25) at v=0v=0 and v=sv=-s, respectively.

We evaluate the double integral of I2(u,s)I_{2}(u,s) in (5.26) by writing the integrand in (5.25) as

U1u+sU2v+sXs(u+s)(v+s)1u2v2(u+2s)2(v+2s)2s(2u+2s)3(2v+2s)3(u+v+2s)4(u,v,s),\frac{U_{1}^{u+s}U_{2}^{v+s}X^{-s}}{(u+s)(v+s)}\frac{1}{u^{2}v^{2}}\frac{(u+2s)^{2}(v+2s)^{2}}{s(2u+2s)^{3}(2v+2s)^{3}(u+v+2s)^{4}}\mathcal{F}(u,v,s),

where

(u,v,s)=\displaystyle\mathcal{F}(u,v,s)= 𝒥(s)(212s1)ζK(2s)(s)w(u+s)w(v+s)\displaystyle\mathcal{J}(s)(2^{1-2s}-1)\zeta_{K}(2s)(s)w(u+s)w(v+s)
×(ζK(1+u)u)2(ζK(1+v)v)2\displaystyle\quad\times(\zeta_{K}(1+u)u)^{2}(\zeta_{K}(1+v)v)^{2}
×(ζK(1+2u+2s)(2u+2s))3(ζK(1+2v+2s)(2v+2s))3(ζK(1+u+v+2s)(u+v+2s))4(ζK(1+u+2s)(u+2s))2(ζK(1+v+2s)(v+2s))2\displaystyle\times\frac{(\zeta_{K}(1+2u+2s)(2u+2s))^{3}(\zeta_{K}(1+2v+2s)(2v+2s))^{3}(\zeta_{K}(1+u+v+2s)(u+v+2s))^{4}}{(\zeta_{K}(1+u+2s)(u+2s))^{2}(\zeta_{K}(1+v+2s)(v+2s))^{2}}
×Z4(12+u,12+v,s).\displaystyle\quad\times Z_{4}(\tfrac{1}{2}+u,\tfrac{1}{2}+v,s).

It is easy to see that (u,v,s)\mathcal{F}(u,v,s) is analytic for (u+2s),(v+2s)>0\Re(u+2s),\Re(v+2s)>0 by Lemma 2.13 and that

I2(u,s)=\displaystyle I_{2}(u,s)= U1u+sU2sXs16(u+2s)3(u+s)4s4u2((u,0,s)(s(u+2s)logU210s3u)+(0,1,0)(u,0,s)s(u+2s)).\displaystyle\frac{U_{1}^{u+s}U_{2}^{s}X^{-s}}{16(u+2s)^{3}(u+s)^{4}s^{4}u^{2}}\left(\mathcal{F}(u,0,s)(s(u+2s)\log U_{2}-10s-3u)+\mathcal{F}^{(0,1,0)}(u,0,s)s(u+2s)\right).

It follows from this that by moving the line of the double integral in (5.26) involving I2(u,s)I_{2}(u,s) from (u)=1100\Re(u)=\frac{1}{100} to (u)=1100+1logX\Re(u)=-\frac{1}{100}+\frac{1}{\log X} and applying (5.21), (5.22), we have

(5.27) 1(2πi)2(1100)(1100)I2(u,s)𝑑u𝑑s=12πi(1100)Resu=0(I2(u,s))𝑑s+O(U21100X1100(logX)5).\displaystyle\frac{1}{(2\pi i)^{2}}\int\limits_{(\frac{1}{100})}\int\limits_{(\frac{1}{100})}I_{2}(u,s)\ du\ ds=\frac{1}{2\pi i}\int\limits_{(\frac{1}{100})}\underset{u=0}{\operatorname{Res}}\left(I_{2}(u,s)\right)ds+O\left(U_{2}^{\frac{1}{100}}X^{-\frac{1}{100}}(\log X)^{5}\right).

Note that

Resu=0I2(u,s)=\displaystyle\underset{u=0}{\operatorname{Res}}\ I_{2}(u,s)= U1sU2sXs64s11((0,0,s)(s2logU1logU25slogU15slogU2+26)\displaystyle\frac{U_{1}^{s}U_{2}^{s}X^{-s}}{64s^{11}}\Big{(}\mathcal{F}(0,0,s)(s^{2}\log U_{1}\log U_{2}-5s\log U_{1}-5s\log U_{2}+26)
+(1,0,0)(0,0,s)(s2logU25s)+(0,1,0)(0,0,s)(s2logU15s)+(1,1,0)(0,0,s)s2).\displaystyle+\mathcal{F}^{(1,0,0)}(0,0,s)(s^{2}\log U_{2}-5s)+\mathcal{F}^{(0,1,0)}(0,0,s)(s^{2}\log U_{1}-5s)+\mathcal{F}^{(1,1,0)}(0,0,s)s^{2}\Big{)}.

As one checks that the expression in the parenthesis above is analytic for 116(s)18-\frac{1}{16}\leq\Re(s)\leq\frac{1}{8}, we can move the line of the integral in (5.27) involving Res u=0I2(u,s)\displaystyle\underset{u=0}{\text{Res }}I_{2}(u,s) to (s)=1100\Re(s)=-\frac{1}{100}. In this process, we encounter a pole at s=0s=0 so that we have

(5.28) 12πi(1100)Resu=0(I2(u,s))𝑑s=(0,0,0)64j1+j2+j3+j4=10j1,j2,j3,j40(1)j3B(j4)j1!j2!j3!j4!(logj1U1)(logj2U2)(logj3X)+O(U11100U21100X1100+(logX)9),\displaystyle\begin{split}\frac{1}{2\pi i}\int\limits_{(\frac{1}{100})}\underset{u=0}{\operatorname{Res}}\left(I_{2}(u,s)\right)ds=&\frac{\mathcal{F}(0,0,0)}{64}\sum_{\begin{subarray}{c}j_{1}+j_{2}+j_{3}+j_{4}=10\\ j_{1},j_{2},j_{3},j_{4}\geq 0\end{subarray}}\frac{(-1)^{j_{3}}B(j_{4})}{j_{1}!j_{2}!j_{3}!j_{4}!}(\log^{j_{1}}U_{1})(\log^{j_{2}}U_{2})(\log^{j_{3}}X)\\ &+O\left(U_{1}^{-\frac{1}{100}}U_{2}^{-\frac{1}{100}}X^{\frac{1}{100}}+(\log X)^{9}\right),\end{split}

where

B(j)={26if j=0,5(logU1+logU2)if j=1,2logU1logU2if j=2,0if j3.\displaystyle B(j)=\left\{\begin{array}[c]{ll}26&\text{if }j=0,\\ -5(\log U_{1}+\log U_{2})&\text{if }j=1,\\ 2\log U_{1}\log U_{2}&\text{if }j=2,\\ 0&\text{if }j\geq 3.\end{array}\right.

To evaluate (0,0,0)\mathcal{F}(0,0,0), we use the fact that sΓ(s)=1s\Gamma(s)=1 when s=0s=0 and the functional equation (2.10) for ζK(s)\zeta_{K}(s):

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)

to obtain that ζK(0)=14\zeta_{K}(0)=-\frac{1}{4}. On the other hand, a direct calculation shows that Z4(12,12,0)=163ζK(2)a4Z_{4}(\tfrac{1}{2},\tfrac{1}{2},0)=\frac{16}{3\zeta_{K}(2)}a_{4}. We then deduce that

(0,0,0)=Φ^(1)ζK(0)(π4)10Z4(12,12,0)=4Φ^(1)a43ζK(2).\displaystyle\mathcal{F}(0,0,0)=\widehat{\Phi}(1)\zeta_{K}(0)(\frac{\pi}{4})^{10}Z_{4}(\tfrac{1}{2},\tfrac{1}{2},0)=-\frac{4\widehat{\Phi}(1)a_{4}}{3\zeta_{K}(2)}.

Similarly, by moving the line of the integration over uu from (u)=1100\Re(u)=\frac{1}{100} to (u)=1logX\Re(u)=\frac{1}{\log X} in the double integral of I3(u,s)I_{3}(u,s) in (5.26) and applying estimations given in (5.21)-(5.23), we see that

(5.29) 1(2πi)2(1100)(1100)I3(u,s)𝑑u𝑑sU11100X1100(logX)5Φ(5).\displaystyle\frac{1}{(2\pi i)^{2}}\int\limits_{(\frac{1}{100})}\int\limits_{(\frac{1}{100})}I_{3}(u,s)\ du\ ds\ll U_{1}^{\frac{1}{100}}X^{-\frac{1}{100}}(\log X)^{5}\Phi_{(5)}.

We now summarize our result on S1(k1=±i)S_{1}(k_{1}=\pm i) in the following lemma, by combining the estimations from (5.25)-(5.29).

Lemma 5.5.

We have

S1(k1=±i)=\displaystyle S_{1}(k_{1}=\pm i)= 8πa4Φ^(1)X364ζK(2)(π4)10j1+j2+j3+j4=10(1)j3B(j4)j1!j2!j3!j4!(logj1U1)(logj2U2)(logj3X)\displaystyle-\frac{8\pi a_{4}\widehat{\Phi}(1)X}{3\cdot 64\zeta_{K}(2)}(\frac{\pi}{4})^{10}\sum_{j_{1}+j_{2}+j_{3}+j_{4}=10}\frac{(-1)^{j_{3}}B(j_{4})}{j_{1}!j_{2}!j_{3}!j_{4}!}(\log^{j_{1}}U_{1})(\log^{j_{2}}U_{2})(\log^{j_{3}}X)
+XO((logX)9+U11100U21100X1100+U21100X1100(logX)5)\displaystyle+X\cdot O\Big{(}(\log X)^{9}+U_{1}^{-\frac{1}{100}}U_{2}^{-\frac{1}{100}}X^{\frac{1}{100}}+U_{2}^{\frac{1}{100}}X^{-\frac{1}{100}}(\log X)^{5}\Big{)}
+XO(U1150U21100X1100Φ(4)+Y1(logX)22Φ(4)+U11100X1100(logX)5Φ(5)).\displaystyle+X\cdot O\Big{(}U_{1}^{\frac{1}{50}}U_{2}^{-\frac{1}{100}}X^{-\frac{1}{100}}\Phi_{(4)}+Y^{-1}(\log X)^{22}\Phi_{(4)}+U_{1}^{\frac{1}{100}}X^{-\frac{1}{100}}(\log X)^{5}\Phi_{(5)}\Big{)}.

5.6. Computing S1S_{1}: the term S1(k±i)S_{1}(k\neq\pm i)

In this section, we estimate S1(k±i)S_{1}(k\neq\pm i). We first deduce from (5.13) that

(5.30) S1(k1±i)Xa1mod(1+i)3N(a)Y1N(a)2k𝒪Kk0,k1±i|(ε)(ε)(12+ε)N(a)2s𝒥(s)w(u+s)w(v+s)×U1u+sU2v+sXs(u+s)(v+s)1N(k)sL2(1+u,χik1)L2(1+v,χik1)Z2(12+u,12+v,a,k)dsdudv|.\displaystyle\begin{split}S_{1}(k_{1}\neq\pm i)\ll&X\sum_{\begin{subarray}{c}a\equiv 1\bmod{(1+i)^{3}}\\ N(a)\leq Y\end{subarray}}\frac{1}{N(a)^{2}}\sum_{\begin{subarray}{c}k\in\mathcal{O}_{K}\\ k\neq 0,k_{1}\neq\pm i\end{subarray}}\Bigg{|}\int\limits_{(\varepsilon)}\int\limits_{(\varepsilon)}\int\limits_{(\tfrac{1}{2}+\varepsilon)}N(a)^{2s}\mathcal{J}(s)w(u+s)w(v+s)\\ &\quad\times\frac{U_{1}^{u+s}U_{2}^{v+s}X^{-s}}{(u+s)(v+s)}\frac{1}{N(k)^{s}}L^{2}(1+u,\chi_{ik_{1}})L^{2}(1+v,\chi_{ik_{1}})Z_{2}(\tfrac{1}{2}+u,\tfrac{1}{2}+v,a,k)\ ds\ du\ dv\Bigg{|}.\end{split}

Let ZZ be a parameter to be chosen later. We denote S1,1(k1±i)S_{1,1}(k_{1}\neq\pm i) for the right-hand side expression above truncated to N(k1)ZN(k_{1})\leq Z and S1,2(k1±i)S_{1,2}(k_{1}\neq\pm i) for the right-hand side expression above over N(k1)>ZN(k_{1})>Z. For S1,1(k1±i)S_{1,1}(k_{1}\neq\pm i), we shift the the lines of the integrations to (u)=(v)=12+1logX,(s)=34\Re(u)=\Re(v)=-\frac{1}{2}+\frac{1}{\log X},\Re(s)=\frac{3}{4}. For S1,2(k1±i)S_{1,2}(k_{1}\neq\pm i), we shift the the lines of the integrations to (u)=(v)=12+1logX,(s)=54\Re(u)=\Re(v)=-\frac{1}{2}+\frac{1}{\log X},\Re(s)=\frac{5}{4}. Thus we obtain via (2.21) that

(5.31) S1,1(k1±i)X14U114U214(logX)10N(a)Y(a,2)=1d[i]4(a)N(a)(12+1logX)(12+1logX)(34)|𝒥(s)w(u+s)w(v+s)|×k1±iN(k1)Zd[i]12(k1)N(k1)34|L(1+u,χik1)|4|ds||du||dv|,\displaystyle\begin{split}S_{1,1}(k_{1}\neq\pm i)\ll&X^{\frac{1}{4}}U_{1}^{\frac{1}{4}}U_{2}^{\frac{1}{4}}(\log X)^{10}\sum_{\begin{subarray}{c}N(a)\leq Y\\ (a,2)=1\end{subarray}}\frac{d^{4}_{[i]}(a)}{\sqrt{N(a)}}\int\limits_{(-\frac{1}{2}+\frac{1}{\log X})}\int\limits_{(-\frac{1}{2}+\frac{1}{\log X})}\int\limits_{(\frac{3}{4})}|\mathcal{J}(s)w(u+s)w(v+s)|\\ &\times\sideset{}{{}^{\flat}}{\sum}_{\begin{subarray}{c}k_{1}\neq\pm i\\ N(k_{1})\leq Z\end{subarray}}\frac{d^{12}_{[i]}(k_{1})}{N(k_{1})^{\frac{3}{4}}}|L(1+u,\chi_{ik_{1}})|^{4}\ |ds|\ |du|\ |dv|,\end{split}

where \sideset{}{{}^{\flat}}{\sum} denotes the sum over all square-free elements in 𝒪K\mathcal{O}_{K}.

By the Cauchy-Schwarz inequality, we have that

(5.32) k1±iN(k1)Zd[i]12(k1)N(k1)34|L(1+u,χik1)|4(k1±iN(k1)Zd[i]24(k1)N(k1))2(k1±iN(k1)Z1N(k1)12|L(1+u,χik1)|8)1/2(logZ)223(k1±iN(k1)Z1N(k1)12|L(1+u,χik1)|8)1/2.\displaystyle\begin{split}&\sideset{}{{}^{\flat}}{\sum}_{\begin{subarray}{c}k_{1}\neq\pm i\\ N(k_{1})\leq Z\end{subarray}}\frac{d^{12}_{[i]}(k_{1})}{N(k_{1})^{\frac{3}{4}}}|L(1+u,\chi_{ik_{1}})|^{4}\ll\Big{(}\sideset{}{{}^{\flat}}{\sum}_{\begin{subarray}{c}k_{1}\neq\pm i\\ N(k_{1})\leq Z\end{subarray}}\frac{d^{24}_{[i]}(k_{1})}{N(k_{1})}\Big{)}^{2}\Big{(}\sideset{}{{}^{\flat}}{\sum}_{\begin{subarray}{c}k_{1}\neq\pm i\\ N(k_{1})\leq Z\end{subarray}}\frac{1}{N(k_{1})^{\frac{1}{2}}}|L(1+u,\chi_{ik_{1}})|^{8}\Big{)}^{1/2}\\ \ll&(\log Z)^{2^{23}}\Big{(}\sideset{}{{}^{\flat}}{\sum}_{\begin{subarray}{c}k_{1}\neq\pm i\\ N(k_{1})\leq Z\end{subarray}}\frac{1}{N(k_{1})^{\frac{1}{2}}}|L(1+u,\chi_{ik_{1}})|^{8}\Big{)}^{1/2}.\end{split}

Similar to our proof of Theorem 1.2 (and hence Corollary 1.3), we can show that under GRH, for |(u)|T|\Im(u)|\leq T,

k1±iN(k1)Z|L(1+u,χik1)|8Z(logZ)37.\displaystyle\sideset{}{{}^{\flat}}{\sum}_{\begin{subarray}{c}k_{1}\neq\pm i\\ N(k_{1})\leq Z\end{subarray}}|L(1+u,\chi_{ik_{1}})|^{8}\ll Z(\log Z)^{37}.

Applying the above estimation in (5.32), we deduce that for |(u)|Z|\Im(u)|\leq Z,

(5.33) k1±iN(k1)Zd[i]12(k1)N(k1)34|L(1+u,χik1)|4Z14(logZ)224.\displaystyle\sideset{}{{}^{\flat}}{\sum}_{\begin{subarray}{c}k_{1}\neq\pm i\\ N(k_{1})\leq Z\end{subarray}}\frac{d^{12}_{[i]}(k_{1})}{N(k_{1})^{\frac{3}{4}}}|L(1+u,\chi_{ik_{1}})|^{4}\ll Z^{\frac{1}{4}}(\log Z)^{2^{24}}.

Unconditionally, we have similar to Lemma 2.15 that

k1±iN(k1)Z|L(1+u,χik1)|4Z1+ε(1+|(u)|2)1+ε.\displaystyle\sideset{}{{}^{\flat}}{\sum}_{\begin{subarray}{c}k_{1}\neq\pm i\\ N(k_{1})\leq Z\end{subarray}}|L(1+u,\chi_{ik_{1}})|^{4}\ll Z^{1+\varepsilon}(1+|\Im(u)|^{2})^{1+\varepsilon}.

It follows from this and partial summation that we have

(5.34) k1±iN(k1)Zd[i]12(k1)N(k1)34|L(1+u,χik1)|4k1±iN(k1)Z1N(k1)34ε|L(1+u,χik1)|4Z14+ε(1+|(u)|2)1+ε.\displaystyle\sideset{}{{}^{\flat}}{\sum}_{\begin{subarray}{c}k_{1}\neq\pm i\\ N(k_{1})\leq Z\end{subarray}}\frac{d^{12}_{[i]}(k_{1})}{N(k_{1})^{\frac{3}{4}}}|L(1+u,\chi_{ik_{1}})|^{4}\ll\sideset{}{{}^{\flat}}{\sum}_{\begin{subarray}{c}k_{1}\neq\pm i\\ N(k_{1})\leq Z\end{subarray}}\frac{1}{N(k_{1})^{\frac{3}{4}-\varepsilon}}|L(1+u,\chi_{ik_{1}})|^{4}\ll Z^{\frac{1}{4}+\varepsilon}(1+|\Im(u)|^{2})^{1+\varepsilon}.

Applying the estimation given (5.33) to (5.31) for |(u)|Z|\Im(u)|\leq Z and using (5.34) otherwise, we obtain by noting the exponential decay of ww that

(5.35) S1,1(k1±i)\displaystyle S_{1,1}(k_{1}\neq\pm i)\ll X14U114U214(logX)10Y(logY)24Z14(logZ)224Φ(4).\displaystyle X^{\frac{1}{4}}U_{1}^{\frac{1}{4}}U_{2}^{\frac{1}{4}}(\log X)^{10}Y(\log Y)^{2^{4}}Z^{\frac{1}{4}}(\log Z)^{2^{24}}\Phi_{(4)}.

Similarly, we have

(5.36) S1,2(k1±i)X14U134U234Y3/2(logY)24Z14(logZ)224Φ(4).\displaystyle S_{1,2}(k_{1}\neq\pm i)\ll X^{-\frac{1}{4}}U_{1}^{\frac{3}{4}}U_{2}^{\frac{3}{4}}Y^{3/2}(\log Y)^{2^{4}}Z^{-\frac{1}{4}}(\log Z)^{2^{24}}\Phi_{(4)}.

We further observe that if instead we apply (5.34) to (5.31) for all (u)\Im(u), then we obtain corresponding estimations for S1,1(k1±i),S1,2(k1±i)S_{1,1}(k_{1}\neq\pm i),S_{1,2}(k_{1}\neq\pm i) by replacing (logZ)224(\log Z)^{2^{24}} with ZεZ^{\varepsilon} in both (5.35) and (5.36). On setting Z=U1U2Y2X1Z=U_{1}U_{2}Y^{2}X^{-1} and keeping in mind that our choices of YY and ZZ will be at most powers of XX (see Section 5.8), we immediately derive from (5.30), (5.35) and (5.36) the following result.

Lemma 5.7.

Unconditionally, we have

S1(k±i)U112U212YXεΦ(4).S_{1}(k\neq\pm i)\ll U_{1}^{\frac{1}{2}}U_{2}^{\frac{1}{2}}YX^{\varepsilon}\Phi_{(4)}.

Under GRH, we have

S1(k±i)U112U212Y(logX)225Φ(4).S_{1}(k\neq\pm i)\ll U_{1}^{\frac{1}{2}}U_{2}^{\frac{1}{2}}Y(\log X)^{2^{25}}\Phi_{(4)}.

5.8. Proof of Theorem 1.5

We now complete the proof for Theorem 1.5 in this section. By setting U1=U2=U=X(logX)250U_{1}=U_{2}=U=\frac{X}{(\log X)^{2^{50}}} and Y=X12U114U214Y=X^{\frac{1}{2}}U_{1}^{-\frac{1}{4}}U_{2}^{-\frac{1}{4}}, we deduce from (5.2), (5.18), Lemma 5.2, Lemma 5.5 and Lemma 5.7 that under GRH,

(5.37) S(U1,U2)\displaystyle S(U_{1},U_{2}) =(d,2)=1|AU(d)|2Φ(N(d)X)=πa4Φ^(1)2734527ζK(2)(π4)10X(logX)10+O(X(logX)9+ε+X(logX)20Φ(5)).\displaystyle=\sideset{}{{}^{*}}{\sum}_{(d,2)=1}\left|A_{U}(d)\right|^{2}\Phi\left(\tfrac{N(d)}{X}\right)=\frac{\pi a_{4}\widehat{\Phi}(1)}{2^{7}\cdot 3^{4}\cdot 5^{2}\cdot 7\cdot\zeta_{K}(2)}(\frac{\pi}{4})^{10}X(\log X)^{10}+O\left(X(\log X)^{9+\varepsilon}+X(\log X)^{-20}\Phi_{(5)}\right).

Here we note that, similar to [sound1, Lemma 2.1], we can show that the function VV appearing in the definition of AU(d)A_{U}(d) given in (5.1) is a real-valued function, so that we have AU2(d))=|AU(d)|2A^{2}_{U}(d))=|A_{U}(d)|^{2}.

We define BU(d)=L(12,χ(1+i)5d)2AU(d)B_{U}(d)=L(\tfrac{1}{2},\chi_{(1+i)^{5}d})^{2}-A_{U}(d) to see that

BU(d)=1πi(c)w(s)L(12+s,χ(1+i)5d)2N(d)sUss𝑑s.\displaystyle B_{U}(d)=\frac{1}{\pi i}\int\limits_{(c)}w(s)L(\tfrac{1}{2}+s,\chi_{(1+i)^{5}d})^{2}\frac{N(d)^{s}-U^{s}}{s}ds.

We move the line of the integral in the above expression for BU(d)B_{U}(d) to (s)=0\Re(s)=0 by realizing that N(d)sUss\frac{N(d)^{s}-U^{s}}{s} is entire there. Further applying the bound that |N(d)itUitit|log(N(d)U)|\frac{N(d)^{it}-U^{it}}{it}|\ll\log(\frac{N(d)}{U}) for tt\in\mathbb{R}, we see that

BU(d)log(N(d)U)|w(it)||L(12+it,χ(1+i)5d)|2𝑑t.B_{U}(d)\ll\log\left(\frac{N(d)}{U}\right)\int_{-\infty}^{\infty}|w(it)||L(\tfrac{1}{2}+it,\chi_{(1+i)^{5}d})|^{2}dt.

It follows that

(d,2)=1|BU(d)|2Φ(N(d)X)(logXU)2++|w(it1)||w(it2)|(d,2)=1|L(12+it1,χ8d)|2|L(12+it2,χ8d)|2Φ(N(d)X)𝑑t1𝑑t2.\displaystyle\sideset{}{{}^{*}}{\sum}_{(d,2)=1}|B_{U}(d)|^{2}\Phi\left(\tfrac{N(d)}{X}\right)\ll\left(\log\frac{X}{U}\right)^{2}\int_{-\infty}^{+\infty}\int_{-\infty}^{+\infty}|w(it_{1})||w(it_{2})|\sideset{}{{}^{*}}{\sum}_{(d,2)=1}|L(\tfrac{1}{2}+it_{1},\chi_{8d})|^{2}|L(\tfrac{1}{2}+it_{2},\chi_{8d})|^{2}\Phi(\tfrac{N(d)}{X})dt_{1}dt_{2}.

We estimation the sum over dd above using Corollary 1.4 when |t1|,|t2|X|t_{1}|,|t_{2}|\leq X and Lemma 2.15 otherwise. We then deduce by the exponential decay of w(it)w(it) in tt that

(5.38) (d,2)=1|BU(d)|2Φ(N(d)X)X(logX)9+ε.\displaystyle\sideset{}{{}^{*}}{\sum}_{(d,2)=1}|B_{U}(d)|^{2}\Phi\left(\tfrac{N(d)}{X}\right)\ll X(\log X)^{9+\varepsilon}.

Combining (5.37) and (5.38), we see that

(5.39) (d,2)=1L(12,χ(1+i)5d)4Φ(N(d)X)=(d,2)=1(AU(d)+BU(d))2Φ(N(d)X)=(d,2)=1AU(d)2Φ(N(d)X)+O((d,2)=1|BU(d)|2Φ(N(d)X)+2(d,2)=1|AU(d)||BU(d)|Φ(N(d)X))=πa4Φ^(1)2734527ζK(2)(π4)10X(logX)10+O(X(logX)9.5+ε+X(logX)5Φ(5)),\displaystyle\begin{split}&\sideset{}{{}^{*}}{\sum}_{(d,2)=1}L(\tfrac{1}{2},\chi_{(1+i)^{5}d})^{4}\Phi\left(\tfrac{N(d)}{X}\right)\\ =&\sideset{}{{}^{*}}{\sum}_{(d,2)=1}(A_{U}(d)+B_{U}(d))^{2}\Phi\left(\tfrac{N(d)}{X}\right)\\ =&\sideset{}{{}^{*}}{\sum}_{(d,2)=1}A_{U}(d)^{2}\Phi\left(\tfrac{N(d)}{X}\right)+O\Big{(}\sideset{}{{}^{*}}{\sum}_{(d,2)=1}|B_{U}(d)|^{2}\Phi\left(\tfrac{N(d)}{X}\right)+2\sideset{}{{}^{*}}{\sum}_{(d,2)=1}|A_{U}(d)||B_{U}(d)|\Phi\left(\tfrac{N(d)}{X}\right)\Big{)}\\ =&\frac{\pi a_{4}\widehat{\Phi}(1)}{2^{7}\cdot 3^{4}\cdot 5^{2}\cdot 7\cdot\zeta_{K}(2)}(\frac{\pi}{4})^{10}X(\log X)^{10}+O\left(X(\log X)^{9.5+\varepsilon}+X(\log X)^{-5}\Phi_{(5)}\right),\end{split}

where the last expression above follows from an application of the Cauchy-Schwarz inequality to estimate the sum involving the product of |AU(d)||A_{U}(d)| and |BU(d)||B_{U}(d)|.

We now take Φ(t)\Phi(t) to be supported on [1,2][1,2] satisfying Φ(t)=1\Phi(t)=1 for t(1+𝒵1,2𝒵1)t\in(1+\mathcal{Z}^{-1},2-\mathcal{Z}^{-1}) and Φ(ν)(t)ν𝒵ν\Phi^{(\nu)}(t)\ll_{\nu}\mathcal{Z}^{\nu} for all rational integer ν0\nu\geq 0. We then deduce that Φ(ν)ν𝒵ν\Phi_{(\nu)}\ll_{\nu}\mathcal{Z}^{\nu}, and that Φ^(1)=1+O(𝒵1)\widehat{\Phi}(1)=1+O(\mathcal{Z}^{-1}). Thus (5.39) implies that

(d,2)=1L(12,χ(1+i)5d)4Φ(N(d)X)\displaystyle\sideset{}{{}^{*}}{\sum}_{(d,2)=1}L(\tfrac{1}{2},\chi_{(1+i)^{5}d})^{4}\Phi(\tfrac{N(d)}{X})
=\displaystyle= πa4Φ^(1)2734527ζK(2)(π4)10X(logX)10+O(X(logX)10𝒵1+X(logX)9.5+ε+X(logX)5𝒵5).\displaystyle\frac{\pi a_{4}\widehat{\Phi}(1)}{2^{7}\cdot 3^{4}\cdot 5^{2}\cdot 7\cdot\zeta_{K}(2)}(\frac{\pi}{4})^{10}X(\log X)^{10}+O\left(X(\log X)^{10}\mathcal{Z}^{-1}+X(\log X)^{9.5+\varepsilon}+X(\log X)^{-5}\mathcal{Z}^{5}\right).

We then deduce by taking 𝒵=logX\mathcal{Z}=\log X that

(5.40) (d,2)=1X<N(d)2XL(12,χ(1+i)5d)4(d,2)=1L(12,χ(1+i)5d)4Φ(N(d)X)=πa4Φ^(1)2734527ζK(2)(π4)10X(logX)10+O(X(logX)9.5+ε).\displaystyle\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ X<N(d)\leq 2X\end{subarray}}L(\tfrac{1}{2},\chi_{(1+i)^{5}d})^{4}\geq\sideset{}{{}^{*}}{\sum}_{(d,2)=1}L(\tfrac{1}{2},\chi_{(1+i)^{5}d})^{4}\Phi(\tfrac{N(d)}{X})=\frac{\pi a_{4}\widehat{\Phi}(1)}{2^{7}\cdot 3^{4}\cdot 5^{2}\cdot 7\cdot\zeta_{K}(2)}(\frac{\pi}{4})^{10}X(\log X)^{10}+O\left(X(\log X)^{9.5+\varepsilon}\right).

Similarly, we can choose Φ(t)\Phi(t) in (5.39) such that Φ(t)=1\Phi(t)=1 for t[1,2]t\in[1,2], Φ(t)=0\Phi(t)=0 for all t(1𝒵1,2+𝒵1)t\notin(1-\mathcal{Z}^{-1},2+\mathcal{Z}^{-1}), and Φ(ν)(t)ν𝒵ν\Phi^{(\nu)}(t)\ll_{\nu}\mathcal{Z}^{\nu} for all ν0\nu\geq 0. Taking 𝒵=logX\mathcal{Z}=\log X, we can deduce that

(5.41) (d,2)=1X<N(d)2XL(12,χ(1+i)5d)4(d,2)=1L(12,χ(1+i)5d)4Φ(N(d)X)=πa4Φ^(1)2734527ζK(2)(π4)10X(logX)10+O(X(logX)9.5+ε).\displaystyle\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ X<N(d)\leq 2X\end{subarray}}L(\tfrac{1}{2},\chi_{(1+i)^{5}d})^{4}\leq\sideset{}{{}^{*}}{\sum}_{(d,2)=1}L(\tfrac{1}{2},\chi_{(1+i)^{5}d})^{4}\Phi(\tfrac{N(d)}{X})=\frac{\pi a_{4}\widehat{\Phi}(1)}{2^{7}\cdot 3^{4}\cdot 5^{2}\cdot 7\cdot\zeta_{K}(2)}(\frac{\pi}{4})^{10}X(\log X)^{10}+O\left(X(\log X)^{9.5+\varepsilon}\right).

Combining (5.40) and (5.41), we obtain that

(d,2)=1X<N(d)2XL(12,χ(1+i)5d)4=πa4Φ^(1)2734527ζK(2)(π4)10X(logX)10+O(X(logX)9.5+ε).\displaystyle\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ X<N(d)\leq 2X\end{subarray}}L(\tfrac{1}{2},\chi_{(1+i)^{5}d})^{4}=\frac{\pi a_{4}\widehat{\Phi}(1)}{2^{7}\cdot 3^{4}\cdot 5^{2}\cdot 7\cdot\zeta_{K}(2)}(\frac{\pi}{4})^{10}X(\log X)^{10}+O\left(X(\log X)^{9.5+\varepsilon}\right).

The assertion of Theorem 1.5 now follows by summing the above over X=x2X=\frac{x}{2}, X=x4X=\frac{x}{4}, \dots and then resetting xx to be XX.

5.9. Proof of Theorem 1.6

In this section, we complete the proof for Theorem 1.6. We first apply the Cauchy-Schwartz inequality to see that

(5.42) (d,2)=1L(12,χ(1+i)5d)4Φ(dX)𝒜2,\displaystyle\sideset{}{{}^{*}}{\sum}_{(d,2)=1}L(\tfrac{1}{2},\chi_{(1+i)^{5}d})^{4}\Phi\left(\tfrac{d}{X}\right)\geq\frac{\mathcal{A}^{2}}{\mathcal{B}},

where

𝒜=\displaystyle\mathcal{A}= (d,2)=1AU(d)L(12,χ(1+i)5d)2Φ(N(d)X),\displaystyle\sideset{}{{}^{*}}{\sum}_{(d,2)=1}A_{U}(d)L(\tfrac{1}{2},\chi_{(1+i)^{5}d})^{2}\Phi\left(\frac{N(d)}{X}\right),
=\displaystyle\mathcal{B}= (d,2)=1AU(d)2Φ(N(d)X).\displaystyle\sideset{}{{}^{*}}{\sum}_{(d,2)=1}A_{U}(d)^{2}\Phi\left(\frac{N(d)}{X}\right).

Here we recall that AU(d)A_{U}(d) is defined as in (5.1).

We can evaluate \mathcal{B} similar to our evaluation of S(U1,U2)S(U_{1},U_{2}) in Section 5.8, except that we now set U1=U2=U=X14εU_{1}=U_{2}=U=X^{1-4\varepsilon}. By setting Y=X12U114U214Y=X^{\frac{1}{2}}U_{1}^{-\frac{1}{4}}U_{2}^{-\frac{1}{4}} again, we deduce from (5.18), Lemma 5.2, Lemma 5.5 and Lemma 5.7 that unconditionally,

=πa4(1+O(ε))2734527ζK(2)(π4)10Φ^(1)X(logX)10+O(X(logX)9+XΦ(5)),\displaystyle\mathcal{B}=\frac{\pi a_{4}\left(1+O(\varepsilon)\right)}{2^{7}\cdot 3^{4}\cdot 5^{2}\cdot 7\cdot\zeta_{K}(2)}(\frac{\pi}{4})^{10}\widehat{\Phi}(1)X(\log X)^{10}+O\left(X(\log X)^{9}+X\Phi_{(5)}\right),

with the implied constant in O(ε)O(\varepsilon) being absolute.

To evaluate 𝒜\mathcal{A}, we recast it as

𝒜=4(d,2)=1n1,n21mod(1+i)3χ(1+i)5d(n1n2)d[i](n1)d[i](n2)N(n1n2)12h1(d,n1,n2),\mathcal{A}=4\sideset{}{{}^{*}}{\sum}_{(d,2)=1}\sum_{\begin{subarray}{c}n_{1},n_{2}\equiv 1\bmod{(1+i)^{3}}\end{subarray}}\frac{\chi_{(1+i)^{5}d}(n_{1}n_{2})d_{[i]}(n_{1})d_{[i]}(n_{2})}{N(n_{1}n_{2})^{\frac{1}{2}}}h_{1}(d,n_{1},n_{2}),

where

h1(x,y,z)=Φ(N(x)X)V(N(y)U)V(N(z)N(x)).h_{1}(x,y,z)=\Phi\left(\frac{N(x)}{X}\right)V\left(\frac{N(y)}{U}\right)V\left(\frac{N(z)}{N(x)}\right).

A similar argument to our evaluation of \mathcal{B} above implies that, by taking Y=X12U14X14Y=X^{\frac{1}{2}}U^{-\frac{1}{4}}X^{-\frac{1}{4}}, we have unconditionally,

𝒜=πa4(1+O(ε))2734527ζK(2)(π4)10Φ^(1)X(logX)10+O(X(logX)9+XΦ(5)),\displaystyle\mathcal{A}=\frac{\pi a_{4}\left(1+O(\varepsilon)\right)}{2^{7}\cdot 3^{4}\cdot 5^{2}\cdot 7\cdot\zeta_{K}(2)}(\frac{\pi}{4})^{10}\widehat{\Phi}(1)X(\log X)^{10}+O\left(X(\log X)^{9}+X\Phi_{(5)}\right),

with the implied constant in O(ε)O(\varepsilon) being absolute.

We now take 𝒵=logX\mathcal{Z}=\log X and take Φ\Phi so that Φ(t)=1\Phi(t)=1 for t(1+𝒵1,2𝒵1)t\in(1+\mathcal{Z}^{-1},2-\mathcal{Z}^{-1}), Φ(t)=0\Phi(t)=0 for all t(1,2)t\notin(1,2), and Φ(ν)(t)ν𝒵ν\Phi^{(\nu)}(t)\ll_{\nu}\mathcal{Z}^{\nu} for all ν0\nu\geq 0. Applying our estimations for 𝒜\mathcal{A} and \mathcal{B} in (5.42), we deduce that

(d,2)=1X<N(d)2XL(12,χ(1+i)5d)4(1+O(ε))πa42734527ζK(2)(π4)10Φ^(1)X(logX)10.\displaystyle\sideset{}{{}^{*}}{\sum}_{\begin{subarray}{c}(d,2)=1\\ X<N(d)\leq 2X\end{subarray}}L(\tfrac{1}{2},\chi_{(1+i)^{5}d})^{4}\geq\left(1+O(\varepsilon)\right)\frac{\pi a_{4}}{2^{7}\cdot 3^{4}\cdot 5^{2}\cdot 7\cdot\zeta_{K}(2)}(\frac{\pi}{4})^{10}\widehat{\Phi}(1)X(\log X)^{10}.

The assertion of Theorem 1.6 now follows by summing the above over X=x2X=\frac{x}{2}, X=x4X=\frac{x}{4}, \dots and then resetting xx to be XX.

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

References