Hostname: page-component-7bb8b95d7b-dtkg6 Total loading time: 0 Render date: 2024-09-27T02:38:07.480Z Has data issue: false hasContentIssue false

An approximation formula for the shifted cubic moment of automorphic L-functions in the weight aspect

Published online by Cambridge University Press:  06 September 2023

Olga Balkanova*
Affiliation:
Steklov Mathematical Institute of Russian Academy of Sciences, 8 Gubkina Street, Moscow 119991, Russia
John Brian Conrey
Affiliation:
American Institute of Mathematics, Caltech 8-32, 1200 E California Boulevard, Pasadena, CA 91125, USA e-mail: conrey@aimath.org
Dmitry Frolenkov
Affiliation:
HSE University and Steklov Mathematical Institute of Russian Academy of Sciences, 8 Gubkina Street, Moscow 119991, Russia e-mail: frolenkov@mi-ras.ru
Rights & Permissions [Opens in a new window]

Abstract

Consider the family of automorphic L-functions associated with primitive cusp forms of level one, ordered by weight k. Assuming that k tends to infinity, we prove a new approximation formula for the cubic moment of shifted L-values over this family which relates it to the fourth moment of the Riemann zeta function. More precisely, the formula includes a conjectural main term, the fourth moment of the Riemann zeta function and error terms of size smaller than that predicted by the recipe conjectures.

Type
Article
Copyright
© The Author(s), 2023. Published by Cambridge University Press on behalf of The Canadian Mathematical Society

1 Introduction

This paper is a continuation and generalization of [Reference Frolenkov6], where the cubic moment of central values $L(f,1/2)$ of automorphic L-functions associated with primitive cusp forms of level one and large weight has been considered. In the present work, we study the shifted cubic moment, namely,

(1.1) $$ \begin{align} \mathcal{M}(\alpha_1,\alpha_2,\alpha_3):=\sum_{f\in H_{2k}} \omega(f)L(f,1/2+\alpha_1)L(f,1/2+\alpha_3)L(f,1/2+\alpha_3), \end{align} $$

where $H_{2k}$ is the normalized Hecke basis of the space of holomorphic cusp forms of weight $2k \geq 2$ and level one. The harmonic weight $\omega (f)$ is defined in a standard way (see [Reference Frolenkov6, (1.6)]).

One of the reasons why we are interested in investigating the shifted cubic moment is that its asymptotic behavior can be conjectured using the “recipe” by Conrey, Farmer, Keating, Rubinstein, and Snaith [Reference Conrey, Farmer, Keating, Rubinstein and Snaith2]. These conjectures predict that the main term of (1.1) is given by

(1.2) $$ \begin{align} \mathcal{MT}_3(\alpha_1,\alpha_2,\alpha_3)&= \sum_{\epsilon_1=\pm1}\sum_{\epsilon_2=\pm1}\sum_{\epsilon_3=\pm1} \mathcal{C}(\epsilon_1,\epsilon_2,\epsilon_3) \nonumber\\ &\quad \times \zeta(1+\epsilon_1\alpha_1+\epsilon_2\alpha_2) \zeta(1+\epsilon_1\alpha_1+\epsilon_3\alpha_3) \zeta(1+\epsilon_2\alpha_2+\epsilon_3\alpha_3), \end{align} $$

where

(1.3) $$ \begin{align} \mathcal{C}(1,1,1)=1,\quad \mathcal{C}(1,1,-1)=(-1)^kX_{k}(\alpha_3), \end{align} $$
(1.4) $$ \begin{align} \mathcal{C}(1,-1,1)=(-1)^kX_{k}(\alpha_2),\quad \mathcal{C}(-1,1,1)=(-1)^kX_{k}(\alpha_1), \end{align} $$
(1.5) $$ \begin{align} \mathcal{C}(1,-1,-1)=X_{k}(\alpha_2)X_{k}(\alpha_3),\quad \mathcal{C}(-1,1,-1)=X_{k}(\alpha_1)X_{k}(\alpha_3), \end{align} $$
(1.6) $$ \begin{align} \mathcal{C}(-1,-1,1)=X_{k}(\alpha_1)X_{k}(\alpha_2),\quad \mathcal{C}(-1,-1,-1)=(-1)^kX_{k}(\alpha_1)X_{k}(\alpha_2)X_{k}(\alpha_3), \end{align} $$

and

(1.7) $$ \begin{align} X_{k}(\alpha):=(2\pi)^{2\alpha}\frac{\Gamma(k-\alpha)}{\Gamma(k+\alpha)}. \end{align} $$

Unfortunately, an unconditional proof of an asymptotic formula for (1.1) is out of current technology. Nevertheless, we are able to investigate the structure and different properties of (1.1). In general, shifts reveal more clearly the combinatorial structure of moments, while the case of central values is just the limit case when all of the shifts are zero. For this reason, it is quite useful to introduce shifts. Accordingly, in [Reference Frolenkov6], the majority of computations was performed for the shifted cubic moment, and only at the last step shifts are taken to be zero. Unfortunately, the method used in [Reference Frolenkov6] has a disadvantage of not being fully symmetric in terms of shifts. Without any doubts, (1.1) is symmetric in $\alpha _1,\alpha _2,\alpha _3$ , and therefore, the asymptotic formula for this moment should have the same property. The method of studying the cubic moment in [Reference Frolenkov6] is based on the decomposition $3=2+1$ . This means that we first use the series representation for one of the L-functions and reduce the problem to the study of the second moment. This process surely shuffles a bit the structure of the cubic moment; and in this paper, we try to overcome this issue.

We introduce some notation in order to formulate our main result. First, let $\bar {\alpha }=(\alpha _1,\alpha _2,\alpha _3)$ and

(1.8) $$ \begin{align} \zeta_4(\bar{\alpha};w) & = \zeta\left(\frac{1+\alpha_1+\alpha_2-\alpha_3}{2}-w\right) \zeta\left(\frac{1+\alpha_1-\alpha_2+\alpha_3}{2}-w\right)\nonumber\\& \quad \times \zeta\left(\frac{1-\alpha_1+\alpha_2+\alpha_3}{2}-w\right) \zeta\left(\frac{1+\alpha_1+\alpha_2+\alpha_3}{2}+w\right). \end{align} $$

Note that the product (1.8) is symmetric in $\alpha _1,\alpha _2,\alpha _3$ . To shorten formulas, we also introduce the following notation ${}_p\mathrm {I}_q$ for the generalized hypergeometric function multiplied by Gamma factors

(1.9) $$ \begin{align} {}_{p}\mathrm{I}_{q} \left( \begin{matrix} a_1,\ldots,a_p \\ b_1,\ldots,b_q \end{matrix}; z \right) := \frac{\Gamma(a_1)\ldots \Gamma(a_p)}{\Gamma(b_1)\ldots \Gamma(b_q)} {}_{p}F_{q} &\left( \begin{matrix} a_1,\ldots,a_p \\ b_1,\ldots,b_q \end{matrix}; z \right) =\nonumber\\ &= \sum_{j=0}^{\infty}\frac{\Gamma(a_1+j)\ldots \Gamma(a_p+j)}{\Gamma(b_1+j)\ldots \Gamma(b_q+j)}\frac{z^j}{j!}. \end{align} $$

Finally, let

(1.10) $$ \begin{align} \mathcal{H}_1(\bar{\alpha},k;w)= {}_{3}\mathrm{I}_{2} \left( \begin{matrix} \frac{1+\alpha_1+\alpha_2-\alpha_3}{2}-w, \frac{1+\alpha_1-\alpha_2+\alpha_3}{2}-w, \frac{1-\alpha_1+\alpha_2+\alpha_3}{2}-w \\ k+\frac{1+\alpha_1+\alpha_2+\alpha_3}{2}-w, \frac{3+\alpha_1+\alpha_2+\alpha_3}{2}-w-k \end{matrix}; 1 \right) \end{align} $$

and

(1.11) $$ \begin{align} l_1(\bar{\alpha},k;w)= (2\pi)^{\frac{1+\alpha_1+\alpha_2+\alpha_3}{2}+w}\exp\left(\frac{\pi|\Im w|}{2}-\pi i\text{sgn}(\Im w)\frac{1+\alpha_1+\alpha_2+\alpha_3}{4}\right). \end{align} $$

Note that the expressions (1.10) and (1.11) are also symmetric in $\alpha _1,\alpha _2,\alpha _3$ . The main result of the paper is contained in the following theorem.

Theorem 1.1 For any $\epsilon>0,$ there exists $\delta (\epsilon )$ such that for $|\alpha _j|\ll \delta (\epsilon )$ the following formula holds:

(1.12) $$ \begin{align} \mathcal{M}(\alpha_1,\alpha_2,\alpha_3)& = \mathcal{MT}_3(\alpha_1,\alpha_2,\alpha_3)+ \nonumber\\& \quad + \frac{(-1)^k}{2\pi i}\int_{(0)}\zeta_4(\bar{\alpha};w)\mathcal{H}_1(\bar{\alpha},k;w)l_1(\bar{\alpha},k;w)dw+ O\left(\frac{k^{\epsilon}}{k}\right). \end{align} $$

The requirement of $\alpha _1,\alpha _2,\alpha _3$ being small is more or less of a technical nature and allows us to simplify the proof of several estimates on special functions.

Note that the error term in (1.12) is smaller than $O(k^{-1/2})$ predicted by the recipe conjectures.

The structure of the paper is as follows: First, in Section 2, we provide a statement of the formula for the shifted cubic moment proved in [Reference Frolenkov6]. Then, in Section 3, we prove various results on special functions that will be used later to obtain a symmetric version of the formula for the cubic moment. Finally, in Section 4, we prove Theorem 1.1.

2 Reciprocity type formula for the cubic moment

In this section, we state the formula proved in [Reference Frolenkov6], which relates the considered cubic moment to the fourth moment of the Riemann zeta function. Let

(2.1) $$ \begin{align} \rho=\alpha_1,\quad u=\frac{\alpha_2+\alpha_3}{2},\quad v=\frac{\alpha_2-\alpha_3}{2}. \end{align} $$

For such $u,v,\rho $ , the cubic moment in [Reference Frolenkov6, (1.9)] coincides with (1.1). Let

(2.2) $$ \begin{align} \mathcal{MT}_D(\bar{\alpha})& = \zeta(1+\alpha_1+\alpha_2)\zeta(1+\alpha_1+\alpha_3)\zeta(1+\alpha_2+\alpha_3)\mathcal{C}(1,1,1)+ \nonumber\\& \quad + \zeta(1+\alpha_1+\alpha_2)\zeta(1+\alpha_1-\alpha_3)\zeta(1+\alpha_2-\alpha_3)\mathcal{C}(1,1,-1)+ \nonumber\\& \quad + \zeta(1+\alpha_1-\alpha_2)\zeta(1+\alpha_1+\alpha_3)\zeta(1-\alpha_2+\alpha_3)\mathcal{C}(1,-1,1)+\nonumber \\& \quad + \zeta(1+\alpha_1-\alpha_2)\zeta(1+\alpha_1-\alpha_3)\zeta(1-\alpha_2-\alpha_3)\mathcal{C}(1,-1,-1), \end{align} $$

where $\mathcal {C}(\pm 1,\pm 1,\pm 1)$ are defined in (1.3)–(1.6). We also introduce the functions

(2.3) $$ \begin{align} Z_{1}(u,v,\rho,k;w)&= \zeta\left(\frac{1+\rho}{2}+v+w\right)\zeta\left(\frac{1+\rho}{2}-v+w\right) \nonumber\\& \quad \times \zeta\left(\frac{1+\rho}{2}+u-w\right)\zeta\left(\frac{1+\rho}{2}-u-w\right) \hat{g}_{1}(u,v,\rho,k;w), \end{align} $$
(2.4) $$ \begin{align} \mathcal{ET}_1(u,v,\rho) &= (-1)^k\left(R_{1,1}+R_{1,2}-R_{1,3}- R_{1,4}\right)(u,v,\rho,k) \nonumber\\& \quad + \frac{(-1)^k}{2\pi i}\int_{(0)} Z_{1}(u,v,\rho,k;w)dw, \end{align} $$

where $\hat {g}_{1}(u,v,\rho ,k;w)$ is given in [Reference Frolenkov6, Lemma 4.7] and

(2.5) $$ \begin{align} R_{1,1}(u,v,\rho,k)=\zeta(1-2v)\zeta(\rho+u+v)\zeta(\rho-u+v)\hat{g}_{1}(u,v,\rho,k;1/2-\rho/2-v), \end{align} $$
(2.6) $$ \begin{align} R_{1,2}(u,v,\rho,k)=\zeta(1+2v)\zeta(\rho+u-v)\zeta(\rho-u-v)\hat{g}_{1}(u,v,\rho,k;1/2-\rho/2+v), \end{align} $$
(2.7) $$ \begin{align} R_{1,3}(u,v,\rho,k)=-\zeta(1-2u)\zeta(\rho+u+v)\zeta(\rho+u-v)\hat{g}_{1}(u,v,\rho,k;-1/2+\rho/2+u), \end{align} $$
(2.8) $$ \begin{align} R_{1,4}(u,v,\rho,k)=-\zeta(1+2u)\zeta(\rho-u+v)\zeta(\rho-u-v)\hat{g}_{1}(u,v,\rho,k;-1/2+\rho/2-u). \end{align} $$

Let us also define

(2.9) $$ \begin{align} Z_{2}(u,v,\rho,k;w) &= \zeta\left(\frac{1+\rho}{2}+v-w\right)\zeta\left(\frac{1+\rho}{2}-v-w\right) \nonumber\\ & \quad \times \zeta\left(\frac{1+\rho}{2}+u+w\right)\zeta\left(\frac{1+\rho}{2}-u+w\right) \hat{g}_{2}(u,v,\rho,k;w), \end{align} $$
(2.10) $$ \begin{align} \mathcal{ET}_2(u,v,\rho) & = \left(R_{2,1}+R_{2,2}-R_{2,3}-R_{2,4}\right)(u,v,\rho,k) \nonumber\\ & \quad + \frac{1}{2\pi i}\int_{(0)}Z_{2}(u,v,\rho,k;w)dw, \end{align} $$

where $\hat {g}_{2}(u,v,\rho ,k;w)$ is given in [Reference Frolenkov6, Lemma 4.3] and

(2.11) $$ \begin{align} R_{2,1}(u,v,\rho,k)=\zeta(1-2u)\zeta(\rho+u+v)\zeta(\rho+u-v)\hat{g}_{2}(u,v,\rho,k;1/2-\rho/2-u), \end{align} $$
(2.12) $$ \begin{align} R_{2,2}(u,v,\rho,k)=\zeta(1+2u)\zeta(\rho-u+v)\zeta(\rho-u-v)\hat{g}_{2}(u,v,\rho,k;1/2-\rho/2+u), \end{align} $$
(2.13) $$ \begin{align} R_{2,3}(u,v,\rho,k)=-\zeta(1+2v)\zeta(\rho+u-v)\zeta(\rho-u-v)\hat{g}_{2}(u,v,\rho,k;-1/2+\rho/2-v), \end{align} $$
(2.14) $$ \begin{align} R_{2,4}(u,v,\rho,k)=-\zeta(1-2v)\zeta(\rho+u+v)\zeta(\rho-u+v)\hat{g}_{2}(u,v,\rho,k;-1/2+\rho/2+v). \end{align} $$

Finally, let

(2.15) $$ \begin{align} \mathcal{ET}_3(u,v,\rho)=(-1)^k\frac{\Gamma(k-u+v)}{\Gamma(k+u-v)}(2\pi)^{2u-2v} \mathcal{ET}_2(v,u,\rho). \end{align} $$

Theorem 2.1 The following formula holds:

(2.16) $$ \begin{align} \mathcal{M}(\alpha_1,\alpha_2,\alpha_3)=\mathcal{MT}_D(\bar{\alpha})+\mathcal{ET}_1(u,v,\rho)+\mathcal{ET}_2(u,v,\rho)+\mathcal{ET}_3(u,v,\rho). \end{align} $$

Proof Formula (2.16) is a modification of [Reference Frolenkov6, (5.1)]. The term $\mathcal {MT}_D(\bar {\alpha })$ corresponds to $\mathcal {MT}_1(u,v,\rho )$ defined in [Reference Frolenkov6, (5.2)]. Formula (2.2) follows immediately from [Reference Frolenkov6, (3.8) and (5.2)] together with (2.1). We are left to prove (2.15). It follows from [Reference Frolenkov6, (3.5)] and [Reference Olver, Lozier, Boisvert and Clarke8, 15.8.1] that

(2.17) $$ \begin{align} \psi_k(x;u,v) &= 2(2\pi)^{2u}\frac{\Gamma(k-u+v)\Gamma(k-u-v)}{\Gamma(2k)} \nonumber\\ & \quad \times\sin{(\pi(1/2+v))}x^k(1+x)^{-k+v} {}_{2}F_{1} \left( \begin{matrix} k+u-v,k-u-v \\ 2k \end{matrix}; \frac{x}{1+x} \right). \end{align} $$

Comparing (2.17) with [Reference Frolenkov6, (3.4)], we conclude

(2.18) $$ \begin{align} \psi_k(x;u,v)=\frac{\Gamma(k-u+v)}{\Gamma(k+u-v)}(2\pi)^{2u-2v} \Phi_k\left(\frac{x}{1+x};v,u\right). \end{align} $$

Now, (2.15) follows from [Reference Frolenkov6, (5.24) and (5.5)] and (2.18). Note that there are several typos in [Reference Frolenkov6, (4.18) and (4.19)]. According to (2.18) and [Reference Frolenkov6, (4.14) and (4.17)], we have

(2.19) $$ \begin{align} \hat{g}_{3}(u,v,\rho,k;w)=\frac{\Gamma(k-u+v)}{\Gamma(k+u-v)}(2\pi)^{2u-2v} \hat{g}_{2}(v,u,\rho,k;w). \end{align} $$

Therefore, the correct versions of [Reference Frolenkov6, (4.18) and (4.19)] could be obtained from [Reference Frolenkov6, (4.15) and (4.16)] with the use of (2.19). Typos in [Reference Frolenkov6, (4.18) and (4.19)] do not affect other formulas or results in [Reference Frolenkov6] since in [Reference Frolenkov6], we deal with the case $u=v=0$ . In that case [Reference Frolenkov6, (4.18) and (4.19)] are correct.

We are going to deduce Theorem 1.1 from Theorem 2.1. As one can see from (2.2) and (1.2), the term $\mathcal {MT}_D(\bar {\alpha })$ in (2.16) contains four of eight summands in (1.2). The remaining four terms are hidden in the expressions $R_{i,j}(u,v,\rho ,k)$ . To evaluate $R_{1,1},R_{1,2}$ and to estimate $R_{2,1},R_{2,2}$ , we will prove some new results on the behavior of functions $\hat {g}_{j}(v,u,\rho ,k;w)$ , $j=1,2$ . This is done in the next section.

3 Special functions

In this section, we first recall some definitions and properties of special functions appearing in [Reference Frolenkov6, Section 4]. Then, we will obtain some new results concerning these functions required to prove Theorem 1.1. First, it follows from (2.1) and the statement of Theorem 1.1 that

(3.1) $$ \begin{align} |v|+|u|+|\rho|<\delta \end{align} $$

for some small $\delta>0.$ Also throughout the section, we will frequently use the Stirling bound [Reference Olver, Lozier, Boisvert and Clarke8, 5.11.9] on Gamma factors

(3.2) $$ \begin{align} |\Gamma(x+iy)|\ll|y|^{x-1/2}e^{-\pi|y|/2}. \end{align} $$

3.1 Properties of $g_{2}(u,v,\rho ,k;x)$

Let (see [Reference Frolenkov6, (3.4)])

(3.3) $$ \begin{align} \Phi_k(x;u,v) &:= 2(2\pi)^{2u}\frac{\Gamma(k-u+v)\Gamma(k-u-v)}{\Gamma(2k)} \nonumber\\ & \quad \times\sin{(\pi(1/2+u))}x^k(1-x)^{-u}{}_2F_{1}(k-u+v,k-u-v,2k;x), \end{align} $$

and for $\Re {w}>-1/2-\Re {\rho }/2+\Re {u}$ (see [Reference Frolenkov6, (4.14)]),

(3.4) $$ \begin{align} \hat{g}_{2}(u,v,\rho,k;w)= \int_0^1(1-x)^{w+\rho/2-1/2}x^{-1-\rho} \Phi_{k}\left(x;u,v\right)dx. \end{align} $$

Lemma 3.1 For $|v|+|u|+|\rho |<\delta $ and

$$\begin{align*}\max(-1/2+\Re{(\rho/2)}+\Re{v},-1/2+|\Re{u}|-\Re{(\rho/2)})<\Re{w},\end{align*}$$

we have

(3.5) $$ \begin{align} \hat{g}_{2}(u,v,\rho,k;w)&=2(2\pi)^{2u}\cos(\pi u) \frac{\Gamma(k-u- v)}{\Gamma(k+u+v)} \frac{1}{2\pi i} \int_{(\Delta)}\frac{\Gamma(k-z)}{\Gamma(k+z)} \nonumber\\ & \quad \times \frac{\Gamma(1/2+\rho/2-v+w-z)}{\Gamma(1/2-\rho/2-v+w)} \Gamma\left(z-u+v\right)\Gamma\left(z+u+v\right)\Gamma\left(z-\rho\right) dz, \end{align} $$

where

$$\begin{align*}\max(\Re{(\rho)},|\Re{u}|-\Re{(v)})<\Delta<\min(1/2+\Re{(\rho/2-v+w)},k).\end{align*}$$

Proof Substituting (3.3) to (3.4), applying [Reference Olver, Lozier, Boisvert and Clarke8, 15.8.1] and making the change of variable $x=(1+t)^{-1}$ , we obtain

(3.6) $$ \begin{align} \frac{\hat{g}_{2}(u,v,\rho,k;w)}{2(2\pi)^{2u}\cos(\pi u)}&= \int_0^{\infty}t^{w-v+\rho/2-1/2-k}(1+t)^{v-w+\rho/2-1/2} \nonumber\\& \quad \times {}_{2}F_{1} \left( \begin{matrix} k-u+v,k+u+v \\ 2k \end{matrix}; -t^{-1} \right) dt. \end{align} $$

Using the Mellin integral representation [Reference Olver, Lozier, Boisvert and Clarke8, 15.6.6] of the hypergeometric function in (3.6), changing the orders of integration and evaluating the integral over t with the help of [Reference Olver, Lozier, Boisvert and Clarke8, 5.12.3], we find

(3.7) $$ \begin{align} \frac{\hat{g}_{2}(u,v,\rho,k;w)}{2(2\pi)^{2u}\cos(\pi u)} & = \frac{\Gamma(k-u- v)}{\Gamma(k+u+v)} \frac{1}{2\pi i} \int_{(\Delta)}\frac{\Gamma(k-u+v+s)\Gamma(k+u+v+s)}{\Gamma(2k+s)} \nonumber\\ &\quad\times \frac{\Gamma(1/2+\rho/2-v+w-k-s)}{\Gamma(1/2-\rho/2-v+w)} \Gamma\left(-s\right)\Gamma\left(k-\rho+s\right) ds, \end{align} $$

where

$$\begin{align*}\max(\Re{(\rho)}-k,|\Re{u}|-\Re{(v)}-k)<\Delta<\min(1/2+\Re{(\rho/2-v+w)}-k,0).\end{align*}$$

Finally, making the change of variable $s=z-k$ , we obtain (3.5).

Lemma 3.2 For $\epsilon>0$ , there exists $\delta (\epsilon )$ such that for $|u|,|v|,|\rho |\ll \delta (\epsilon )$ and any fixed A such that $1<A<K$ we have

(3.8) $$ \begin{align} \hat{g}_{2}(u,v,\rho,k;ir)\ll \frac{(k(|r|+1))^{\epsilon}}{k(1+|r|)^A}. \end{align} $$

Proof Moving the line of integration in (3.5) to the right on $\Re {z}=\sigma <k$ , we cross poles at the points

(3.9) $$ \begin{align} z_p(j)=1/2+\rho/2-v+ir+j \quad\text{for}\quad j=0,1,2,\ldots. \end{align} $$

The residues are estimated in the following way:

(3.10) $$ \begin{align} \frac{\Gamma(k-u- v)}{\Gamma(k+u+v)} \frac{\Gamma(k-1/2-\rho/2+v-ir-j)}{\Gamma(k+1/2+\rho/2-v+ir+j)} &\frac{\Gamma\left(1/2-\rho/2-v+ir+j\right)}{\Gamma(1/2-\rho/2-v+ir)}\nonumber\\\times\ \Gamma\left(1/2+\rho/2+ir+ j-u\right)&\Gamma\left(1/2+\rho/2+ir+j+u\right)\ll \nonumber\\ \ll \frac{k^{4\delta}(1+|r|)^{3j+\delta}e^{-\pi|r|}}{(k+|r|)^{1+2j-3\delta}} &\ll \frac{(k(|r|+1))^{\epsilon}}{k(1+|r|)^A}. \end{align} $$

Therefore,

(3.11) $$ \begin{align} \hat{g}_{2}&(u,v,\rho,k;ir)\ll\frac{(k(|r|+1))^{\epsilon}}{k(1+|r|)^A}+ \bigg|\frac{k^{4\delta}}{2\pi i} \int_{(\sigma)}\frac{\Gamma(k-z)}{\Gamma(k+z)} \nonumber\\ &\times \frac{\Gamma(1/2+\rho/2-v+ir-z)}{\Gamma(1/2-\rho/2-v+ir)} \Gamma\left(z-u+v\right)\Gamma\left(z+u+v\right)\Gamma\left(z-\rho\right) dz\bigg|. \end{align} $$

Let $z=\sigma +iy$ and $y_0=C\log ^2(k(1+|r|))$ , where the constant C is chosen in such way that $|r-y_0|>r/2$ . To estimate the integral (3.11), we apply (3.2) and split it in two parts, the first one over $|y|>y_0$ and the second over $|y|\le y_0$ . The first integral is smaller than

(3.12) $$ \begin{align} \int_{|y|>y_0}\frac{|y|^{3\sigma+3\delta-3/2}e^{-\pi(3|y|+|y-r|-|r|)/2}} {(|y-r|+1)^{\sigma-3\delta/2}(|r|+1)^{-3\delta/2}(k+|y|)^{2\sigma}}dy\ll \frac{1}{(k(1+|r|))^A}. \end{align} $$

The second integral is smaller than

(3.13) $$ \begin{align} &\int_{|y|\le y_0} \frac{|y|^{3\sigma+3\delta-3/2}(|y-r|+1)^{-\sigma+3\delta/2}} {(|r|+1)^{-3\delta/2}(k+|y|)^{2\sigma}}dy\ll \nonumber\\ &\quad\ll \int_{|y|\le y_0} \frac{(k(1+|r|))^{\epsilon}|y|^{3\sigma-3/2}dy} {(k+|y|)^{2\sigma}(|y-r|+1)^{\sigma}}\ll \frac{(k(1+|r|))^{\epsilon}}{(k^2(1+|r|))^{\sigma}}. \end{align} $$

Substituting (3.12) and (3.13) to (3.11), we prove (3.8).

Lemma 3.3 For any $\epsilon>0$ , there exists $\delta (\epsilon )$ such that for $|u|,|v|,|\rho |\ll \delta (\epsilon ),$ we have

(3.14) $$ \begin{align} \hat{g}_{2}(u,v,\rho,k;1/2-\rho/2\pm u)\ll \frac{k^{\epsilon}}{k^2}. \end{align} $$

Proof Using (3.5) with $\Delta =1-\epsilon _0$ ( $\epsilon _0$ is chosen in such way that all poles of $\Gamma (1-v\pm u-z)$ are located to the right of the line $\Re {z}=\Delta $ ), writing $z=\Delta +iy$ and applying (3.2), we obtain

(3.15) $$ \begin{align} \hat{g}_{2}(u,v,\rho,k;1/2-\rho/2\pm u)\ll k^{4\delta} \int_{-\infty}^{\infty}\frac{(1+|y|)^{2\Delta+3\delta-1}e^{-2\pi|y|}} {(k+|y|)^{2\Delta}}dy\ll \frac{k^{\epsilon}}{k^2}. \end{align} $$

Note that estimating the integral representation [Reference Frolenkov6, (4.15)] of $\hat {g}_{2}(\ldots )$ by absolute value with the use of the Stirling bound (3.2) on Gamma factors, we are able to obtain only a weaker estimate $k^{-1+\epsilon }$ . This estimate is sufficient for our further computations.

3.2 Properties of $g_{1}(u,v,\rho ,k;x)$

Let

$$ \begin{align*} F(v,u;y):=(1-y)^vy^{u} {}_{2}F_{1} \left( \begin{matrix} k+u+v,1-k+u+v \\ 1+2u \end{matrix}; y \right). \end{align*} $$

For $0<y<1,$ let (see [Reference Balkanova and Frolenkov1, Lemma 5.1])

(3.16) $$ \begin{align} \phi_k(1-y;u,v)=\tilde{\phi}_k(1-y;u,v)+\tilde{\phi}_k(1-y;u,-v), \end{align} $$

where

(3.17) $$ \begin{align} \tilde{\phi}_k(1-y; u,v)&=(-1)^k\frac{(2\pi)^{2u}\pi}{\sin{(\pi v)}}\left( \frac{\Gamma(k+v-u)\Gamma(k-v-u)}{\Gamma(k+v+u)\Gamma(k-v+u)} \right. \nonumber\\& \quad\left. \times\ \frac{\sin{(\pi(v+u))}}{\Gamma(1-2u)\sin{(2\pi u)}} F(v,-u;y)+ \frac{\sin{(\pi(v-u))}}{\Gamma(1+2u)\sin{(-2\pi u)}} F(v,u;y)\! \right). \end{align} $$

For $-1/2-\Re {\rho }/2+|\Re {v}|<\Re {w}<1/2+\Re {\rho }/2-|\Re {u}|,$ let (see [Reference Frolenkov6, (4.27)])

(3.18) $$ \begin{align} \hat{g}_{1}(u,v,\rho,k;w)= \int_0^1(y(1-y))^{\rho/2-1/2}\frac{(1-y)^{w}}{y^{w}} \phi_{k}\left(1-y;u,v\right)dy. \end{align} $$

We are going to prove an analog of [Reference Frolenkov6, Lemma 4.7] which will be used later to deduce a version of [Reference Frolenkov6, Lemma 4.8] for $\hat {g}_{1}(u,v,\rho ,k;1/2-\rho /2\pm v)$ .

Lemma 3.4 For $|v|+|u|+|\rho |<\delta $ and

$$\begin{align*}-1/2-\Re{\rho}/2+|\Re{v}|<\Re{w}<2+\Re{\rho}/2-|\Re{u}|,\end{align*}$$

we have

(3.19) $$ \begin{align} \hat{g}_{1}(u,v,\rho,k;w)&=\sum_{\pm}\frac{(-1)^k(2\pi)^{2u}}{2\sin(\pm\pi v)}\frac{\Gamma(k-u\mp v)}{\Gamma(k+u\pm v)}\Gamma\left(\frac{1+\rho}{2}\pm v+w\right) \nonumber\\& \quad \times \frac{1}{2\pi i}\int_{(\Delta_{\pm})}\frac{\Gamma(k-1/2+s/2)}{\Gamma(k+1/2-s/2)} \Gamma\left(\frac{1-s}{2}+u\pm v\right)\Gamma\left(\frac{1-s}{2}-u\pm v\right) \nonumber\\& \quad \times \cos\left( \pi(\pm2v-s/2)\right)\frac{\Gamma(\rho/2\mp v-w+s/2)}{\Gamma(1/2+\rho+s/2)}ds, \end{align} $$

where $\max (1-2k,\Re {(-\rho +2w)}\pm 2\Re {v})<\Delta _{\pm }<1-|2\Re {u}|\pm 2\Re {v}$ .

Proof It follows from [Reference Gradshteyn, Ryzhik, Jeffrey and Zwillinger7, 6.574.1] that for $-\Re {(k+u)}<\Re {(v)}<1/2,$ we have

(3.20) $$ \begin{align} F(v,u;y)=(1-y)^v2^{-2v}\frac{\Gamma(k-u-v)\Gamma(1+2u)}{\Gamma(k+u+v)} \int_{0}^{\infty}J_{2u}(t\sqrt{y})J_{2k-1}(t)t^{2v}dt. \end{align} $$

Substituting (3.20) to (3.17), we obtain, for $-\Re {(k-|u|)}<\Re {(v)}<1/2,$

(3.21) $$ \begin{align} \tilde{\phi}_k(1-y;u,v)&=(-1)^k\frac{(2\pi)^{2u}\pi(1-y)^{v}2^{-2v}} {\sin{(\pi v)}\sin{(2\pi u)}}\frac{\Gamma(k-v-u)}{\Gamma(k+v+u)} \nonumber\\& \quad \times \sum_{\pm}\pm\sin{\pi(v\pm u)} \int_{0}^{\infty}J_{\pm2u}(t\sqrt{y})J_{2k-1}(t)t^{2v}dt. \end{align} $$

Applying [Reference Olver, Lozier, Boisvert and Clarke8, 1.14.36] together with [Reference Erdelyi, Magnus, Oberhettinger and Tricomi3, p. 326 f.(1)], we prove that

(3.22) $$ \begin{align} \int_{0}^{\infty}J_{2u}(t\sqrt{y})J_{2k-1}(t)t^{2v}dt= \frac{1}{2\pi i}\int_{(c)} \frac{\Gamma(k+v-s/2)\Gamma(u+s/2)}{\Gamma(k-v+s/2)\Gamma(1+u-s/2)} \frac{2^{2v-1}ds}{y^{s/2}}. \end{align} $$

Next, we change the variable s to $1+2v-s$ , and then apply [Reference Olver, Lozier, Boisvert and Clarke8, 5.5.3] to transform $\Gamma (1/2+u-v+s/2)$ . As a result,

(3.23) $$ \begin{align} \int_{0}^{\infty}&J_{2u}(t\sqrt{y})J_{2k-1}(t)t^{2v}dt= \frac{1}{2\pi i}\int_{(c)} \frac{\Gamma(k-1/2+s/2)}{\Gamma(k+1/2-s/2)} \nonumber\\ &\times \Gamma(1/2+u+v-s/2)\Gamma(1/2-u+v-s/2)\cos\pi(u-v+s/2) \frac{2^{2v-1}ds}{\pi y^{1/2+v-s/2}}. \end{align} $$

Substituting (3.23) to (3.21), we have

(3.24) $$ \begin{align} \tilde{\phi}_k(1-y;u,v)&= (-1)^k \frac{(2\pi)^{2u}(1-y)^{v}}{2\sin{(\pi v)}y^{1/2+v}} \frac{\Gamma(k-v-u)}{\Gamma(k+v+u)}\nonumber\\ & \quad \times \frac{1}{2\pi i}\int_{(c)} \frac{\Gamma(k-1/2+s/2)}{\Gamma(k+1/2-s/2)}\Gamma(1/2+u+v-s/2) \nonumber\\ &\quad\times \Gamma(1/2-u+v-s/2)\cos\pi(2v-s/2) y^{s/2}ds. \end{align} $$

According to [Reference Olver, Lozier, Boisvert and Clarke8, 5.12.1] for $\Re {(-\rho /2-1/2-v)}<\Re {w}<\Re {(\rho /2+s/2-v)},$ we have

(3.25) $$ \begin{align} &\int_0^1(y(1-y))^{\rho/2-1/2} (1-y) {}^{w+v}y^{s/2-1/2-v-w}dy = \frac{\Gamma(\rho/2+s/2-v-w)\Gamma(\rho/2+1/2+v+w)}{\Gamma(\rho+s/2+1/2)}. \end{align} $$

Using (3.16), (3.18), (3.24) and evaluating the resulting integral over y using (3.25), we prove (3.19).

Lemma 3.5 For any $\epsilon>0,$ there exists $\delta (\epsilon )>0$ such that for $|v|+|u|+|\rho |<\delta (\epsilon )$ the following formula holds:

(3.26) $$ \begin{align} & \hat{g}_{1}(u,v,\rho,k;1/2-\rho/2\pm v)= \frac{(-1)^k(2\pi)^{2u}}{\sin(\pm\pi v)} \Gamma(\rho+u\mp v)\Gamma(\rho-u\mp v)\nonumber\\ & \quad \times\left( \sin(\pi\rho) \frac{\Gamma(k-u\mp v)}{\Gamma(k+u\pm v)} \frac{\Gamma(k-\rho\pm2v)}{\Gamma(k+\rho\mp2v)} - \sin\pi(\rho\mp2v) \frac{\Gamma(k-u\pm v)}{\Gamma(k+u\mp v)} \frac{\Gamma(k-\rho)}{\Gamma(k+\rho)} \right) \nonumber\\ & \quad + J_{1}(u,\pm v,\rho,k)+J_{2}(u,\pm v,\rho,k),\nonumber\\ \end{align} $$

where

(3.27) $$ \begin{align} J_{1}(u,\pm v,\rho,k)& = \frac{(-1)^k(2\pi)^{2u}}{2\sin(\pm\pi v)} \frac{\Gamma(k-u\mp v)}{\Gamma(k+u\pm v)}\Gamma(1\pm2v) \nonumber\\ &\quad\times \frac{1}{2\pi i}\int_{(\Delta_1)}\frac{\Gamma(k-1/2+s/2)}{\Gamma(k+1/2-s/2)} \cos\left( \pi(\pm2v-s/2)\right) \nonumber\\ & \quad\times \Gamma\left(\frac{1-s}{2}+u\pm v\right)\Gamma\left(\frac{1-s}{2}-u\pm v\right) \frac{\Gamma(-1/2+\rho\mp2v+s/2)}{\Gamma(1/2+\rho+s/2)}ds, \end{align} $$
(3.28) $$ \begin{align} J_{2}(u,\pm v,\rho,k)& = \frac{(-1)^k(2\pi)^{2u}}{2\sin(\mp\pi v)} \frac{\Gamma(k-u\pm v)}{\Gamma(k+u\mp v)} \nonumber\\ & \quad \times \frac{1}{2\pi i}\int_{(\Delta_2)}\frac{\Gamma(k-1/2+s/2)}{\Gamma(k+1/2-s/2)} \cos\left( \pi(\mp2v-s/2)\right) \nonumber\\ & \quad \times \Gamma\left(\frac{1-s}{2}+u\mp v\right)\Gamma\left(\frac{1-s}{2}-u\mp v\right) \frac{\Gamma(-1/2+\rho+s/2)}{\Gamma(1/2+\rho+s/2)}ds, \end{align} $$

and

$$\begin{align*}\max(1-2k,-1-2\Re{(\rho\pm 4v)})<\Delta_1<1-2|\Re{u}|-2|\Re{v}|,\end{align*}$$
$$\begin{align*}\max(1-2k,-1-2\Re{(\rho)})<\Delta_2<1-2|\Re{u}|-2|\Re{v}|.\end{align*}$$

Moreover, the following estimates hold:

(3.29) $$ \begin{align} J_{1}(u,\pm v,\rho,k)\ll\frac{k^{\epsilon}}{k},\quad J_{2}(u,\pm v,\rho,k) \ll\frac{k^{\epsilon}}{k}. \end{align} $$

Proof The proof is similar to the one of [Reference Frolenkov6, Lemma 4.8]. We move the line of integration in (3.19) to the left, crossing the pole at $s=\pm 2v+2w-\rho $ . Evaluating the residues and letting $w=1/2-\rho /2\pm v,$ we obtain (3.26). To prove (3.29), we estimate Gamma factors in (3.27) and (3.28) by means of the Stirling formula (3.2). Taking into account that $|v|+|u|+|\rho |<\epsilon $ and choosing $\Delta _1=\Delta _2=0,$ we have

$$ \begin{align*} J_{1}(u,\pm v,\rho,k)\ll k^{\epsilon} \int_{-\infty}^{\infty}\frac{(k+|y|)^{\Delta_1-1}}{(1+|y|)^{\Delta_1+1}}dy\ll \frac{k^{\epsilon}}{k}, \end{align*} $$
$$ \begin{align*} J_{2}(u,\pm v,\rho,k)\ll k^{\epsilon} \int_{-\infty}^{\infty}\frac{(k+|y|)^{\Delta_2-1}}{(1+|y|)^{\Delta_2+1\pm2\Re{v}}}dy\ll \frac{k^{\epsilon}}{k}, \end{align*} $$

thus proving (3.29).

The function $\hat {g}_{1}(\ldots )$ can be expressed in terms of $\mathcal {H}_1(\bar {\alpha },k;w)$ (see (1.10)) and $\hat {g}_{2}(\ldots )$ (see [Reference Frolenkov6, (4.16)]).

Lemma 3.6 The following equality holds:

(3.30) $$ \begin{align} \hat{g}_{1}(u,v,\rho,k;-w)&= 2(-1)^k(2\pi)^{\alpha_2+\alpha_3} \Gamma\left(\frac{1+\alpha_1-\alpha_2-\alpha_3}{2}+w\right) \nonumber\\& \quad\times \frac{\sin\pi(\alpha_1/2-w)\cos\pi\left(\frac{\alpha_1-\alpha_2-\alpha_3}{2}+w\right)} {\cos\pi\left(\frac{\alpha_1+\alpha_2+\alpha_3}{2}-w\right)}\mathcal{H}_1(\bar{\alpha},k;w)- \nonumber\\&\quad -\frac{(-1)^k\sin(\pi\alpha_1)} {\cos\pi\left(\frac{\alpha_1+\alpha_2+\alpha_3}{2}-w\right)} \hat{g}_{2}(u,v,\rho,k;w). \end{align} $$

Proof Moving the line of integration in [Reference Frolenkov6, (4.28)] to the left and evaluating residues at the points

$$ \begin{align*} s_1(j)=1-2k-2j,\quad s_2(j)=-\rho-2u-2w-2j\quad j=0,1,2\ldots, \end{align*} $$

we obtain (see also [Reference Frolenkov6, Lemma 4.11])

(3.31) $$ \begin{align} &\hat{g}_{1}(u,v,\rho,k;w)=-2(-1)^k(2\pi)^{2u} \Gamma\left(\frac{1+\rho}{2}-u-w\right) \frac{\sin(\pi\rho)\cos(\pi u)} {\cos\pi(\rho/2+u+w)}\nonumber\\& \quad\times {}_{3}\mathrm{I}_{2} \left( \begin{matrix} k-u+v,k-u-v, k-\rho \\ 2k, k+\frac{1-\rho}{2}-u-w \end{matrix}; 1 \right) + 2(-1)^k(2\pi)^{2u} \Gamma\left(\frac{1+\rho}{2}-u-w\right) \nonumber\\& \quad \times \frac{\sin\pi(\rho/2+w)\cos\pi\left(\rho/2-u-w\right)} {\cos\pi\left(\rho/2+u+w\right)} {}_{3}\mathrm{I}_{2} \left( \begin{matrix} \frac{1+\rho}{2}+v+w, \frac{1+\rho}{2}-v+w, \frac{1-\rho}{2}+u+w, \\ k+\frac{1+\rho}{2}+u+w, \frac{3+\rho}{2}+u+w-k \end{matrix}; 1 \right).\nonumber\\ \end{align} $$

Rewriting (3.31) in terms of $\alpha _1,\alpha _2,\alpha _3$ using (2.1) and applying (1.10), [Reference Frolenkov6, (4.16)], we prove (3.30).

3.3 Properties of $\mathcal {H}_1(\bar {\alpha },k;w)$

The following statement is the core part of the paper.

Proposition 3.7 For any $\epsilon>0,$ there exists $\delta (\epsilon )$ such that for $|\alpha _j|\ll \delta (\epsilon ),$ we have

(3.32) $$ \begin{align} \mathcal{H}_1(\bar{\alpha},k;ir)\ll e^{-\pi|r|/2}(1+|r|)^{\epsilon}\frac{k^{\epsilon}}{\sqrt{k}}, \end{align} $$

and for $|r|\ll \log ^2k$ , we have

(3.33) $$ \begin{align} \mathcal{H}_1(\bar{\alpha},k;ir)\ll\frac{k^{\epsilon}}{k}. \end{align} $$

The proof of the estimates (3.32) and (3.33) is based on the use of the following integral representation for $\mathcal {H}_1(\bar {\alpha },k;w)$ .

Lemma 3.8 We have

(3.34) $$ \begin{align} \mathcal{H}_1(\bar{\alpha},k;w)&= \frac{\Gamma\left(\frac{1-\alpha_1+\alpha_2-\alpha_3}{2}-w\right)}{ \Gamma\left(1+\alpha_2+\alpha_3\right)} \int_0^{1} (1-x)^{\frac{-1+\alpha_1+\alpha_2-\alpha_3}{2}-w} \nonumber\\& \quad\times x^{\frac{-1+\alpha_1+\alpha_2+\alpha_3}{2}+w} {}_{2}F_{1} \left( \begin{matrix} 1+\alpha_2-k,k+\alpha_2 \\ \end{matrix}; 1+\alpha_2+\alpha_3 \right) {x}dx. \end{align} $$

Proof Applying [Reference Prudnikov, Brychkov and Marichev9, (7.4.4.2)] with

$$ \begin{align*} a=\frac{1+\alpha_1-\alpha_2+\alpha_3}{2}-w,\, b=\frac{1+\alpha_1+\alpha_2-\alpha_3}{2}-w,\, c=\frac{1-\alpha_1+\alpha_2+\alpha_3}{2}-w, \end{align*} $$
$$ \begin{align*} d=\frac{3+\alpha_1+\alpha_2+\alpha_3}{2}-w-k,\, e=k+\frac{1+\alpha_1+\alpha_2+\alpha_3}{2}-w,\, s=\frac{1+\alpha_1+\alpha_2+\alpha_3}{2}+w, \end{align*} $$

we deduce

(3.35) $$ \begin{align} \mathcal{H}_1(\bar{\alpha},k;w)& = \frac{\Gamma\left(\frac{1+\alpha_1+\alpha_2-\alpha_3}{2}-w\right) \Gamma\left(\frac{1-\alpha_1+\alpha_2+\alpha_3}{2}-w\right)}{ \Gamma\left(1+\alpha_1+\alpha_2\right)\Gamma\left(1+\alpha_2+\alpha_3\right)} \nonumber\\& \quad \times \Gamma\left(\frac{1+\alpha_1+\alpha_2+\alpha_3}{2}+w\right) {}_{3}F_{2} \left( \begin{matrix} 1+\alpha_2-k,k+\alpha_2, \frac{1+\alpha_1+\alpha_2+\alpha_3}{2}+w \\ 1+\alpha_1+\alpha_2,1+\alpha_2+\alpha_3 \end{matrix}; 1 \right). \end{align} $$

This formula can be viewed as an analog of [Reference Frolenkov6, (4.41)]. For the hypergeometric function on the right-hand side of (3.35), we use the integral representation [Reference Olver, Lozier, Boisvert and Clarke8, 16.5.2] with

$$ \begin{align*} a_0=\frac{1+\alpha_1+\alpha_2+\alpha_3}{2}+w,\quad b_0=1+\alpha_1+\alpha_2, \end{align*} $$

thus proving (3.34).

Therefore, we have reduced the problem of estimating $\mathcal {H}_1(\bar {\alpha },k;w)$ to the problem of finding an asymptotic formula uniform in x for $ {}_{2}F_{1} \left ( \begin {matrix} \\ \end {matrix}; x \right ) $ on the right-hand side of (3.34). We will independently consider the case of x being small, x being close to $1$ and the intermediate case $\varepsilon <x<1-\varepsilon $ . In the last case, one can apply the following result of Farid Khwaja and Olde Daalhuis [Reference Farid Khwaja and Olde Daalhuis4, Theorem 3.2].

Lemma 3.9 For some small fixed $\varepsilon>0$ such that $\varepsilon <y<\pi /2-\varepsilon $ and $k\rightarrow \infty ,$ we have

(3.36) $$ \begin{align}& {}_{2}F_{1} \left( \begin{matrix} 1+\alpha_2-k,k+\alpha_2 \\ 1+\alpha_2+\alpha_3 \end{matrix}; \cos^2y \right) = (-1)^k\frac{\Gamma(1+\alpha_2+\alpha_3)\Gamma(k-\alpha_3)}{\Gamma(k+\alpha_2)} \nonumber\\&\quad\times \left(O(\Phi_n(k,y))+ y^{\alpha_3-\alpha_2} \left(J_{\alpha_2-\alpha_3}(2y\lambda)\cos(\pi\alpha_2)- Y_{\alpha_2-\alpha_3}(2y\lambda)\sin(\pi\alpha_2)\right) \sum_{j=0}^{n-1}\frac{c_j(y)}{k^j} \right. \nonumber \\& \left.\quad\qquad - iy^{\alpha_3-\alpha_2-1} \left(J_{1+\alpha_2-\alpha_3}(2y\lambda)\sin(\pi\alpha_2)+ Y_{1+\alpha_2-\alpha_3}(2y\lambda)\cos(\pi\alpha_2)\right) \sum_{j=1}^{n-1}\frac{d_j(y)}{k^j} \right),\nonumber\\ \end{align} $$

where $\lambda =k-\frac {1}{2}$ , $|c_j(y)|\ll _{\epsilon }1$ , $|d_j(y)|\ll _{\epsilon }1$ , and

(3.37) $$ \begin{align} c_0(y)=-\frac{y^{1/2+\alpha_2-\alpha_3}}{\sin^{1/2+\alpha_2-\alpha_3}y \cos^{1/2+\alpha_2+\alpha_3}y},\quad d_0(y)=0, \end{align} $$
(3.38) $$ \begin{align} \Phi_n(k,y)& =\frac{y^{\Re{(\alpha_3-\alpha_2)}}}{k^n} \left( |J_{\alpha_2-\alpha_3}(2y\lambda)|+ |Y_{\alpha_2-\alpha_3}(2y\lambda)|+ \right.\nonumber\\ & \left.\quad + y^{-1}|Y_{1+\alpha_2-\alpha_3}(2y\lambda)|+ y^{-1}|J_{1+\alpha_2-\alpha_3}(2y\lambda)| \right). \end{align} $$

Proof We apply [Reference Farid Khwaja and Olde Daalhuis4, Theorem 3.2] with

$$ \begin{align*} z=-\cos(2y),\,\lambda=k-\frac{1}{2},\, a=\frac{1}{2}+\alpha_2,\,c=1+\alpha_2+\alpha_3. \end{align*} $$

In this case, [Reference Farid Khwaja and Olde Daalhuis4, (3.9)] becomes

$$ \begin{align*} \xi=\log\left(-z-i\sqrt{1-z^2} \right)=-2iy, \end{align*} $$

and [Reference Farid Khwaja and Olde Daalhuis4, (3.10)] proves (3.37). Finally, the K-Bessel functions of purely imaginary argument in [Reference Farid Khwaja and Olde Daalhuis4, (3.8)] can be transformed into combination of Y and J-Bessel functions by means of [Reference Olver, Lozier, Boisvert and Clarke8, 10.27.8 and 10.4.3]

$$ \begin{align*} K_{\nu}(2iy\lambda)=\frac{-\pi i}{2}e^{-\nu\pi i/2}H_{\nu}^{(2)}(2y\lambda)=\frac{-\pi i}{2}e^{-\nu\pi i/2} \left( J_{\nu}(2y\lambda)-iY_{\nu}(2y\lambda)\right), \end{align*} $$
$$ \begin{align*} K_{\nu}(-2iy\lambda)=\frac{\pi i}{2}e^{\nu\pi i/2}H_{\nu}^{(1)}(2y\lambda)=\frac{\pi i}{2}e^{\nu\pi i/2}\left( J_{\nu}(2y\lambda)+iY_{\nu}(2y\lambda) \right). \end{align*} $$

After some straightforward computations, we obtain (3.36).

It is explained in [Reference Farid Khwaja and Olde Daalhuis5, Section 4], why [Reference Farid Khwaja and Olde Daalhuis4, Theorem 3.2] does not provide an adequate asymptotic formula either for y close to 0 or to $\pi /2$ . The case y close to $\pi /2$ is studied in [Reference Farid Khwaja and Olde Daalhuis5, Section 4.1], and the case of y close to $0$ is considered in [Reference Farid Khwaja and Olde Daalhuis5, Section 4.4]. Both asymptotic formulas are given not in terms of Bessel functions, but in terms of Kummer functions. See [Reference Olver, Lozier, Boisvert and Clarke8, Section13] for the definition and properties of these functions.

Lemma 3.10 Assume that $|\alpha _j|\ll \delta (\epsilon )$ . For some small fixed $\varepsilon>0$ such that $0<x<\varepsilon $ and $k\rightarrow \infty ,$ we have

(3.39) $$ \begin{align}\qquad &{}_{2}F_{1} \left( \begin{matrix} 1+\alpha_2-k,k+\alpha_2 \\ 1+\alpha_2+\alpha_3 \end{matrix}; x \right) = \frac{\Gamma(1+\alpha_2+\alpha_3)\Gamma(k-\alpha_3)}{\Gamma(k+\alpha_2)}e^{-k\zeta_0} \nonumber\\&\quad\times \left(O(\Phi^M_n(k,x))+ \frac{k^{\alpha_2+\alpha_3}}{\Gamma(1+\alpha_2+\alpha_3)} M(1+\alpha_2-k,1+\alpha_2+\alpha_3,2k\zeta_0) \sum_{j=0}^{n-1}\frac{a_{j,0}}{k^j} \right.\nonumber\\&\left.\quad\qquad + \frac{k^{\alpha_2+\alpha_3-1}}{\Gamma(\alpha_2+\alpha_3)} M(1+\alpha_2-k,\alpha_2+\alpha_3,2k\zeta_0) \sum_{j=0}^{n-1}\frac{a_{j,1}}{k^j} \right),\nonumber\\[-24pt] \end{align} $$

where $|a_{j,0}|,|a_{j,1}|\ll 1$ and

(3.40) $$ \begin{align} \kern32pt\cos\theta_0=1-2x,\quad \sigma_0+\sin\sigma_0=\theta_0,\quad \zeta_0=1-\cos\sigma_0, \end{align} $$
(3.41) $$ \begin{align} \Phi^M_n(k,x)& = k^{\epsilon-n}|M(1+\alpha_2-k,1+\alpha_2+\alpha_3,2k\zeta_0)| \nonumber\\ &\quad +k^{\epsilon-1-n}|M(1+\alpha_2-k,\alpha_2+\alpha_3,2k\zeta_0)|. \end{align} $$

Proof This is [Reference Farid Khwaja and Olde Daalhuis5, (48)]. The estimate (3.41) of the error term can be obtained in the same way as in [Reference Farid Khwaja and Olde Daalhuis4, Theorem 3.2].

Lemma 3.11 Assume that $|\alpha _j|\ll \delta (\epsilon )$ . For some small fixed $\varepsilon>0$ such that $1-\varepsilon <x<1$ and $k\rightarrow \infty ,$ we have

(3.42) $$ \begin{align}& {}_{2}F_{1} \left( \begin{matrix} 1+\alpha_2-k,k+\alpha_2 \\ 1+\alpha_2+\alpha_3 \end{matrix}; x \right) = \frac{\Gamma(1+\alpha_2+\alpha_3)}{\Gamma(k+\alpha_2)}e^{-k\zeta_1} \nonumber\\&\quad \times \left(O(\Phi^U_n(k,x))+ U(1+\alpha_2-k,1+\alpha_2-\alpha_3,2k\zeta_1) \sum_{j=0}^{n-1}\frac{b_{j,0}}{k^{j+\alpha_3-\alpha_2}}\nonumber \right.\\&\left. \quad\qquad - (k-1-\alpha_3) U(1+\alpha_2-k,\alpha_2-\alpha_3,2k\zeta_1) \sum_{j=0}^{n-1}\frac{b_{j,1}}{k^{j+1+\alpha_3-\alpha_2}} \right), \end{align} $$

where $|b_{j,0}|,|b_{j,1}|\ll 1$ and

(3.43) $$ \begin{align} \kern37pt\cos\theta_1=2x-1,\quad \sigma_1+\sin\sigma_1=\theta_1,\quad \zeta_1=1-\cos\sigma_1, \end{align} $$
(3.44) $$ \begin{align} \Phi^U_n(k,x)&= k^{\epsilon-n}|U(1+\alpha_2-k,1+\alpha_2-\alpha_3,2k\zeta)|\nonumber\\ &\quad +k^{\epsilon-n}|U(1+\alpha_2-k,\alpha_2-\alpha_3,2k\zeta)|. \end{align} $$

Proof This is [Reference Farid Khwaja and Olde Daalhuis5, (74)]. The estimate (3.44) of the error term can be obtained in the same way as in [Reference Farid Khwaja and Olde Daalhuis4, Theorem 3.2].

We are left to obtain asymptotic formulas for the Kummer functions of the form

(3.45) $$ \begin{align} M(\alpha-a,c,az),\, U(\alpha-a,c,az)\text{ as }a\rightarrow+\infty, \end{align} $$

which appears on the right-hand sides of (3.39) and (3.42). To do this, we generalize the results of Temme [Reference Temme10, Sections 27.4.4 and 27.4.5]. Note that the M-Kummer function is in fact $_{1}F_{1}$ -hypergeometric function, i.e., by [Reference Olver, Lozier, Boisvert and Clarke8, 13.2.2 and 16.2.1], we have

(3.46) $$ \begin{align} M(a,b,z)=\frac{\Gamma(b)}{\Gamma(a)}\sum_{j=0}^{\infty}\frac{\Gamma(a+j)}{\Gamma(b+j)}\frac{z^j}{j!}= {}_{1}F_{1} \left( \begin{matrix} a \\ b \end{matrix}; z \right). \end{align} $$

So, we are left to obtain an asymptotic formula for

(3.47) $$ \begin{align} {}_{1}F_{1} \left( \begin{matrix} \alpha-a \\ c \end{matrix}; az \right). \end{align} $$

The next Lemma is a generalization of [Reference Temme10, (27.4.71)].

Lemma 3.12 For $0\le z<4$ , fixed values of $\alpha $ and c, and $a\rightarrow \infty ,$ we have

(3.48) $$ \begin{align} & {}_{1}F_{1} \left( \begin{matrix} \alpha-a \\ c \end{matrix}; az \right) = \frac{\Gamma(1-\alpha+a)\Gamma(c)}{\Gamma(c+a-\alpha)}e^{az/2}\gamma^{1-c} \nonumber\\&\quad \times \left(O(\Phi^J_n(a,z))+ J_{c-1}(2\gamma a) \sum_{j=0}^{n-1}\frac{\tilde{A}_{j}(a,z)}{a^{j}}+ \gamma J_{c-2}(2\gamma a) \sum_{j=0}^{n-1}\frac{\tilde{B}_{j}(a,z)}{a^{j}} \right), \end{align} $$

where $|\tilde {A}_j(a,z)|,|\tilde {B}_j(a,z)|\ll 1$ and

(3.49) $$ \begin{align} \theta=\arcsin\frac{\sqrt{z}}{2},\quad \gamma=\theta+\frac{1}{2}\sin(2\theta), \end{align} $$
(3.50) $$ \begin{align} \Phi^J_n(a,z)= a^{-n}|J_{c-1}(2\gamma a)| +\gamma a^{-n}|J_{c-2}(2\gamma a)|. \end{align} $$

Proof Using [Reference Temme10, (10.1.9) and (10.3.46)], we first show that

(3.51) $$ \begin{align} {}_{1}F_{1} \left( \begin{matrix} \alpha-a \\ c \end{matrix}; az \right) = \frac{\Gamma(1-\alpha+a)\Gamma(c)}{\Gamma(c+a-\alpha)}\frac{e^{az/2}}{2\pi i}\int_{L}e^{(c+a-\alpha)s-az/s}f(-az,-s)\frac{ds}{s^c}, \end{align} $$

where L is the Hankel contour (see [Reference Temme10, Figure 2.2]), and (see [Reference Temme10, (10.3.28)])

(3.52) $$ \begin{align} f(z,s)=e^{zg_0(s)}\left(\frac{s}{1-e^{-s}}\right)^{c},\quad g_0(s)=\frac{1}{s}-\frac{1}{e^s-1}-\frac{1}{2}. \end{align} $$

Note that

(3.53) $$ \begin{align} e^{cs}f(-z,-s)=f(az,s), \end{align} $$

and therefore,

(3.54) $$ \begin{align} e^{az/2}e^{(c+a)s-az/s}f(-az,-s)=e^{as-az/(e^s-1)}\left(\frac{s}{1-e^{-s}}\right)^{c}. \end{align} $$

Substituting (3.54) to (3.51), we obtain an analog of [Reference Temme10, (27.4.65) and (27.4.66)]:

(3.55) $$ \begin{align} {}_{1}F_{1} \left( \begin{matrix} \alpha-a \\ c \end{matrix}; az \right) = \frac{\Gamma(1-\alpha+a)\Gamma(c)}{\Gamma(c+a-\alpha)}\frac{1}{2\pi i}\int_{L}e^{a\psi(w)}\tilde{g}(w)\frac{dw}{w^c}, \end{align} $$

where

(3.56) $$ \begin{align} \psi(w)=w-\frac{z}{e^w-1},\quad \tilde{g}(w)=e^{-\alpha w}\left(\frac{w}{1-e^{-w}}\right)^{c}. \end{align} $$

The only difference with [Reference Temme10, (27.4.66)] is the presence of the multiple $e^{-\alpha w}$ in (3.56). Proceeding as in [Reference Temme10], we obtain [Reference Temme10, (27.4.69)] with $p(t)$ being replaced by

(3.57) $$ \begin{align} \tilde{p}(t)=\left(\frac{w}{t}\right)^{-c}\tilde{g}(w)\frac{dw}{dt}= e^{-\alpha w}\left(\frac{1-e^{-w}}{t}\right)^{-c+2}\frac{t^2+\gamma^2}{(1-e^{-w})^2+ze^{-w}}, \end{align} $$

and [Reference Temme10, (27.4.71)] with $A_k,B_k$ being replaced by

(3.58) $$ \begin{align} \tilde{p}_j(t)=-t^c\frac{d}{dt}(t^{1-c}\tilde{q}_{j-1}(t))=\tilde{A}_j(a,z)+\tilde{B}_j(a,z)t+(t+\gamma^2/t)\tilde{q}_{j}(t), \end{align} $$
(3.59) $$ \begin{align} \tilde{A}_j(a,z)=\frac{\tilde{p}_j(i\gamma)+\tilde{p}_j(i\gamma)}{2},\quad \tilde{B}_j(a,z)=\frac{\tilde{p}_j(i\gamma)-\tilde{p}_j(i\gamma)}{2i\gamma}, \end{align} $$

where $\tilde {p}_0(t)=\tilde {p}(t).$ The closed formulas for the first coefficients are (see [Reference Temme10, (27.4.74)]):

(3.60) $$ \begin{align} \tilde{A}_0(a,z)=\left(\frac{\gamma}{2\sin\theta}\right)^{c}\sqrt{\frac{2\tan\theta}{\gamma}}\cos(c-2\alpha)\theta, \end{align} $$
(3.61) $$ \begin{align} \tilde{B}_0(a,z)=\left(\frac{\gamma}{2\sin\theta}\right)^{c}\sqrt{\frac{2\tan\theta}{\gamma}}\frac{\sin(c-2\alpha)\theta}{\gamma}. \end{align} $$

Finally, the estimate (3.50) of the error term can be obtained in the same way as in [Reference Farid Khwaja and Olde Daalhuis4, Theorem 3.2].

Now, we consider the case of the U-Kummer function [Reference Temme10, Section 27.4.5].

Lemma 3.13 For $0\le z<4$ , fixed values of $\alpha $ and c, and $a\rightarrow \infty ,$ we have

(3.62) $$ \begin{align} &U(\alpha-a,c,az)= \Gamma(1-\alpha+a)e^{az/2}\gamma^{1-c} \nonumber\\&\quad \times \left(O(\Phi^C_n(a,z))+ C_{c-1}(2\gamma a) \sum_{j=0}^{n-1}\frac{\tilde{A}_{j}(a,z)}{a^{j}}+ \gamma C_{c-2}(2\gamma a) \sum_{j=0}^{n-1}\frac{\tilde{B}_{j}(a,z)}{a^{j}} \right), \end{align} $$

where $\tilde {A}_j(a,z),\tilde {B}_j(a,z)$ are the same as in (3.48),

(3.63) $$ \begin{align} C_{\nu}(z)=\cos\pi(a-\alpha)J_{\nu}(z)+\sin\pi(a-\alpha)Y_{\nu}(z), \end{align} $$

and

(3.64) $$ \begin{align} \Phi^C_n(a,z)=a^{-n}\left( |J_{c-1}(2\gamma a)|+|J_{c-2}(2\gamma a)|+|Y_{c-1}(2\gamma a)|+|Y_{c-2}(2\gamma a)|\right). \end{align} $$

Proof Using [Reference Temme10, (10.1.11) and (10.3.61)], we prove an analog of [Reference Temme10, (27.4.80)]:

(3.65) $$ \begin{align} \frac{U(\alpha-a,c,az)}{\Gamma(c+a-\alpha)}&= \frac{e^{\mp\pi i(a-\alpha)}}{\Gamma(c)} {}_{1}F_{1} \left( \begin{matrix} \alpha-a \\ c \end{matrix}; az \right) \nonumber\\ &\quad - \frac{e^{az}e^{\pm\pi ic}}{\Gamma(\alpha-a)}U(c+a-\alpha,c,aze^{\pm\pi i}). \end{align} $$

For the $_{1}F_{1}$ function in (3.65), we can apply (3.48). Therefore, we are left to consider $U(c+a-\alpha ,c,az)$ (at the end, we will replace $az$ by $aze^{\pm \pi i}$ ). Applying [Reference Temme10, (10.3.27)] and (3.53), we prove an analog of [Reference Temme10, (10.3.62)]:

(3.66) $$ \begin{align} U(c+a-\alpha,c,az)=\frac{e^{az/2}}{\Gamma(c-\alpha+a)}\int_0^{\infty}e^{-as-az/s}e^{\alpha s}f(-az,-s)\frac{ds}{s^c}, \end{align} $$

where the function $f(\cdot ,\cdot )$ is defined by (3.52). Simplifying the function under the integral in (3.66), we have the following analog of [Reference Temme10, (27.4.81)]:

(3.67) $$ \begin{align} U(c+a-\alpha,c,az)=\frac{1}{\Gamma(c-\alpha+a)}\int_0^{\infty}e^{-a\varphi(w)} \tilde{g}(-w)\frac{dw}{w^c}, \end{align} $$

where

(3.68) $$ \begin{align} \varphi(w)=w+\frac{z}{e^w-1},\quad \tilde{g}(-w)=e^{\alpha w}\left(\frac{e^{w}-1}{w}\right)^{-c}. \end{align} $$

The only difference with [Reference Temme10, (27.4.81)] is the presence of the factor $e^{\alpha w}$ . Arguing in the same way as in [Reference Temme10, Section 27.4.1], making the change of variable [Reference Temme10, (27.4.32)], we obtain an analog of [Reference Temme10, (27.4.38) and (27.4.82)], namely,

(3.69) $$ \begin{align} U(c+a-\alpha,c,az)=\frac{e^{az/2}}{\Gamma(c-\alpha+a)}\int_0^{\infty}e^{-a(t+\beta^2/t)}\tilde{f}(-t)\frac{dt}{t^c}, \end{align} $$

where

(3.70) $$ \begin{align} \beta=\frac{w_0+\sinh w_0}{2},\quad w_0=2\operatorname{\mathrm{arcsinh}}\frac{\sqrt{z}}{2} \end{align} $$

and

(3.71) $$ \begin{align} \tilde{f}(-t)=e^{\alpha w(t)}\left(\frac{e^{w(t)}-1}{t}\right)^{-c+2}\frac{t^2-\beta^2}{(e^{w(t)}-1)^2-ze^{w(t)}} \end{align} $$

is an analog of [Reference Temme10, (27.4.41)]. Note that $\tilde {f}(t)$ coincides with $\tilde {p}(t)$ defined in (3.57) after changing z to $-z$ . Therefore, we show the following analog of [Reference Temme10, (27.4.84)]:

(3.72) $$ \begin{align} U(\alpha-a,c,aze^{\pm\pi i})& = 2\gamma^{1-c}e^{\pm \pi i(1-c)/2}\frac{e^{-az/2}}{\Gamma(c-\alpha+a)} \nonumber\\ &\quad \times\left( K_{1-c}(\pm 2i\gamma a) \sum_{j=0}^{\infty}\frac{\tilde{A}_{j}(a,z)}{a^{j}}\mp i\gamma K_{c-2}(\pm2i\gamma a) \sum_{j=0}^{\infty}\frac{\tilde{B}_{j}(a,z)}{a^{j}} \right). \end{align} $$

Using the relation [Reference Olver, Lozier, Boisvert and Clarke8, (10.27.3)], i.e., $K_{-\nu }(z)=K_{\nu }(z)$ , and [Reference Temme10, (10.3.64)], we infer

(3.73) $$ \begin{align} e^{\pm \pi i(c+1)/2}K_{c-1}(\pm 2i\gamma a)=\frac{\pi}{2}\left( Y_{c-1}(2\gamma a)\pm iJ_{c-1}(2\gamma a)\right) \end{align} $$
(3.74) $$ \begin{align} e^{\pm \pi i(c+1)/2}K_{c-2}(\pm 2i\gamma a)=\frac{\pi}{2}e^{\pm \pi i/2}\left( Y_{c-2}(2\gamma a)\pm iJ_{c-2}(2\gamma a)\right). \end{align} $$

Truncating the series (3.72) at the point $j=n$ with the error (3.64) (see the proof of [Reference Farid Khwaja and Olde Daalhuis4, Theorem 3.2]), we obtain

(3.75) $$ \begin{align} &U(\alpha-a,c,aze^{\pm\pi i}) = \pi\gamma^{1-c}e^{\mp\pi ic}\frac{e^{-az/2}}{\Gamma(c-\alpha+a)}\nonumber\\& \quad \times \left( O(\Phi^C_n(a,z)) + \left( Y_{c-1}(2\gamma a)\pm iJ_{c-1}(2\gamma a)\right) \sum_{j=0}^{n-1}\frac{\tilde{A}_{j}(a,z)}{a^{j}}\right.\nonumber\\& \left.\qquad\quad +\ \gamma \left( Y_{c-2}(2\gamma a)\pm iJ_{c-2}(2\gamma a)\right) \sum_{j=0}^{n-1}\frac{\tilde{B}_{j}(a,z)}{a^{j}} \right). \end{align} $$

Substituting (3.48) and (3.75) into (3.65), we have

(3.76) $$ \begin{align} &\frac{U(\alpha-a,c,az)}{\Gamma(c+a-\alpha)} = \frac{\Gamma(1-\alpha+a)}{\Gamma(c+a-\alpha)}e^{az/2}\gamma^{1-c} \nonumber\\&\quad\times \left(O(\Phi^C_n(a,z))+ C_{c-1}(2\gamma a) \sum_{j=0}^{n-1}\frac{\tilde{A}_{j}(a,z)}{a^{j}}+ \gamma C_{c-2}(2\gamma a) \sum_{j=0}^{n-1}\frac{\tilde{B}_{j}(a,z)}{a^{j}} \right), \end{align} $$

where

(3.77) $$ \begin{align} C_{\nu}(2\gamma a)=e^{\mp\pi i(a-\alpha)}J_{\nu}(2\gamma a)-\sin\pi(\alpha-a)\left( Y_{\nu}(2\gamma a)\pm iJ_{\nu}(2\gamma a)\right). \end{align} $$

Opening the brackets in (3.77), we complete the proof of (3.63).

Proof of Proposition 3.7

First, we decompose the integral in (3.34) smoothly (say by inserting functions $\chi _0(x),\chi _{1/2}(x),\chi _1(x)$ ) into three ranges:

(3.78) $$ \begin{align} 0<x<\delta_0,\quad \delta_0<x<1-\delta_1,\quad 1-\delta_1<x<1, \end{align} $$

where $\delta _0,\delta _1$ are some small constants.

Consider first the part over $0<x<\delta _0$ . In this case, we apply Lemma 3.10 followed by Lemma 3.12. Taking n in (3.39) and (3.48) sufficiently large, we obtain that the contribution of the error terms is negligible, and thus, it is enough to investigate only the contribution of the main term (all other terms will have the same structure and will be smaller) which is bounded by the sum of the integrals of the following type:

(3.79) $$ \begin{align} &e^{-\pi|\Im{w}|/2}(1+|w|)^{\epsilon}k^{\epsilon} \sum_{j=0}^2\int_0^{1}\chi_0(x) (1-x)^{\frac{-1+\alpha_1+\alpha_2-\alpha_3}{2}-w} \nonumber\\ &\quad\times x^{\frac{-1+\alpha_1+\alpha_2+\alpha_3}{2}+w}A_j(x)\gamma(x)^{-\nu_j} J_{\nu_j}(2\gamma(x) k)dx, \end{align} $$

where $\nu =-j+\alpha _2+\alpha _3$ for $j=0,1,2$ and $A_j(x)$ are products of coefficients in (3.39) and (3.48). Note that

(3.80) $$ \begin{align} \gamma(x)=\frac{\arccos(1-2x)}{2}=\sqrt{x}+O(x^{3/2}), \end{align} $$

since we first perform transformations (3.40) and then (3.49) with $z=2\zeta _0$ . Now, we decompose the integral (3.79) smoothly at the point $x=k^{-2+\epsilon }$ . For $0<x<k^{-2+\epsilon }$ using [Reference Olver, Lozier, Boisvert and Clarke8, (10.7.3)] and (3.80), we show that $ \gamma (x)^{-\nu _j}J_{\nu _j}(2\gamma (x) k)\ll k^{\nu _j}$ , and therefore,

(3.81) $$ \begin{align} \int_0^{k^{-2+\epsilon}} (1-x)^{\frac{-1+\alpha_1+\alpha_2-\alpha_3}{2}-w} x^{\frac{-1+\alpha_1+\alpha_2+\alpha_3}{2}+w}A_j(x)\gamma(x)^{-\nu_j} J_{\nu_j}(2\gamma(x) k)dx\ll k^{-1+\epsilon}. \end{align} $$

In the case of $k^{-2+\epsilon }<x<\delta _0$ , we apply the asymptotic formula [Reference Gradshteyn, Ryzhik, Jeffrey and Zwillinger7, 8.451.1] for the Bessel function. The error term is again negligible and we are left to estimate the integral of the following type (here, we set $w=ir$ ):

(3.82) $$ \begin{align} \int_{k^{-2+\epsilon}}^{1}\chi_0(x)W(x) x^{\frac{-1+\alpha_1+\alpha_2+\alpha_3}{2}}\gamma(x)^{\alpha_2+\alpha_3} e^{ig_{\pm}(x)}\frac{dx}{\sqrt{\gamma(x) k}}, \end{align} $$

where $W(x)=(1-x)^{\frac {-1+\alpha _1+\alpha _2-\alpha _3}{2}}A_j(x)$ and the phase function is equal to

(3.83) $$ \begin{align} g_{\pm}(x)=\pm2\gamma(x) k+r\log x-r\log(1-x). \end{align} $$

Estimating it by absolute value using (3.79), we prove (3.32). For $|r|\ll \log ^2k$ , we have $g^{\prime }_{\pm }(x)\gg k/\sqrt {x}$ , and therefore, using the first derivative test, we estimate (3.82) as

(3.84) $$ \begin{align} \max_{k^{-2+\epsilon}<x<\delta_0} \frac{x^{\frac{\alpha_1+\alpha_2+\alpha_3}{2}}\gamma(x)^{\alpha_2+\alpha_3} }{k\sqrt{\gamma(x) k}}\ll\frac{k^{\epsilon}}{k}. \end{align} $$

As a result, in view of (3.79), we prove (3.33).

Consider the third part of the integral (3.34) over $1-\delta _1<x<1$ . In this case, we apply Lemma 3.11 followed by Lemma 3.13. Again the contribution of the error terms is negligible, and thus, it is enough to investigate only the contribution of the main term (all other terms will have the same structure and will be smaller), which is the sum of the integrals of the following type:

(3.85) $$ \begin{align} & e^{-\pi|\Im{w}|/2}(1+|w|)^{\epsilon}k^{\epsilon} \sum_{j=0}^2\int_0^{1}\chi_1(x) (1-x)^{\frac{-1+\alpha_1+\alpha_2-\alpha_3}{2}-w} \nonumber\\ & \qquad\qquad\qquad\qquad\qquad \times x^{\frac{-1+\alpha_1+\alpha_2+\alpha_3}{2}+w}D_j(x)\gamma(x)^{-\nu_j} C_{\nu_j}(2\gamma_1(x) k)dx, \end{align} $$

where $\nu _j=-j+\alpha _2-\alpha _3$ for $j=0,1,2$ , $D_j(x)$ are products of coefficients in (3.42) and (3.62), and by (3.63), we have

(3.86) $$ \begin{align} C_{\nu_j}(2\gamma_1(x) k)= (-1)^{k-1}\cos(\pi\alpha_2)J_{\nu_j}(2\gamma_1(x) k)+(-1)^k\sin(\pi\alpha_2)Y_{\nu_j}(2\gamma_1(x) k). \end{align} $$

Note that now since we first perform transformations (3.43) and then (3.49) with $z=2\zeta _1$ , we have

(3.87) $$ \begin{align} \gamma_1(x)=\frac{\arccos(2x-1)}{2}=\sqrt{1-x}+O((1-x)^{3/2}). \end{align} $$

Now, we decompose the integral in (3.85) smoothly $x=1-k^{-2+\epsilon }$ . For $1-k^{-2+\epsilon }<x<1$ using [Reference Olver, Lozier, Boisvert and Clarke8, (10.7.3)–(10.7.5)] and (3.87), we have

(3.88) $$ \begin{align} \gamma_1(x)^{-\nu_j}C_{\nu_j}(2\gamma_1(x) k)\ll \gamma_1(x)^{-\nu_j}(\gamma_1(x) k)^{-|\nu_j|}\ll (k(1-x))^{\epsilon}, \end{align} $$

and therefore,

(3.89) $$ \begin{align} \int_{1-k^{-2+\epsilon}}^1 (1-x)^{\frac{-1+\alpha_1+\alpha_2-\alpha_3}{2}-w} x^{\frac{-1+\alpha_1+\alpha_2+\alpha_3}{2}+w}D_j(x)\gamma_1(x)^{-\nu_j} C_{\nu_j}(2\gamma_1(x) k)dx\ll k^{-1+\epsilon}. \end{align} $$

To handle the case of $1-\delta _1<x<1-k^{-2+\epsilon }$ , we apply an asymptotic formulas [Reference Gradshteyn, Ryzhik, Jeffrey and Zwillinger7, 8.451.1 and 8.451.2] for the Bessel functions. Using them, one can write $C_{\nu _j}(2\gamma _1(x) k)$ as

(3.90) $$ \begin{align} C_{\nu_j}(y)=\sum_{\pm}\frac{e^{\pm iy}}{\sqrt{y}}V_{\pm}(y), \end{align} $$

where $V_{\pm }(y)$ smoothed non-oscillating functions uniformly bounded by a constant. So we are left to estimate integral of the following type (here, we set $w=ir$ ):

(3.91) $$ \begin{align} \int_0^{1-k^{-2+\epsilon}}\chi_1(x)W(x) (1-x)^{\frac{-1+\alpha_1+\alpha_2-\alpha_3}{2}} \gamma(x)^{-\alpha_2+\alpha_3} e^{ig_{\pm}(x)}\frac{dx}{\sqrt{\gamma_1(x) k}}, \end{align} $$

where $W(x)=x^{\frac {-1+\alpha _1+\alpha _2+\alpha _3}{2}}V_{\pm }(2\gamma (x) k)D_j(x)$ and the phase function

(3.92) $$ \begin{align} g_{\pm}(x)=\pm2\gamma_1(x) k+r\log x-r\log(1-x). \end{align} $$

Estimating the integral by absolute value and in view of (3.85), we obtain the estimate (3.32). For $|r|\ll \log ^2k,$ we have $g^{\prime }_{\pm }(x)\gg k/\sqrt {1-x}$ and thus using the first derivative test, we estimate (3.91) as

(3.93) $$ \begin{align} \max_{1-\delta_1<x<1-k^{-2+\epsilon}} \frac{(1-x)^{\frac{\alpha_1+\alpha_2-\alpha_3}{2}}\gamma_1(x)^{-\alpha_2+\alpha_3} }{k\sqrt{\gamma_1(x) k}}\ll\frac{k^{\epsilon}}{k}. \end{align} $$

Finally, in view of (3.85), we prove the estimate (3.33).

Consider the second part of the integral (3.34) over $\delta _0<x<1-\delta _1$ . In this case, we make the change of variable $x=\cos ^2y$ and apply Lemma 3.9. Again the contribution of the error terms is negligible and thus it is enough to investigate only the contribution of the main term (all other terms will have the same structure and will be smaller). Note that now y is bounded away from 0 and $\pi /2$ . So the arguments of Y and J-Bessel functions in (3.36) are comparable with k and we can apply for their combination an asymptotic formula similar to (3.90). Finally, we obtain the integrals

(3.94) $$ \begin{align} \frac{e^{-\pi|\Im{w}|/2}(1+|w|)^{\epsilon}}{k^{1/2+\alpha_2+\alpha_3}}\int_0^{1}\chi_{1/2}(\cos^2y)W_{\pm}(y) e^{if_{\pm}(y)}dy, \end{align} $$

where $W_{\pm }(y)$ are smoothed non-oscillating functions uniformly bounded by a constant and

(3.95) $$ \begin{align} f_{\pm}(y)=\pm(2k-1)y-2r\log(\tan y). \end{align} $$

Estimating the integral (3.94) by absolute value, we obtain the estimate (3.32). For $|r|\ll \log ^2k$ , we have $f^{\prime }_{\pm }(y)\gg k$ , and thus using the first derivative test, we estimate (3.94) as

(3.96) $$ \begin{align} \frac{e^{-\pi|\Im{w}|/2}(1+|w|)^{\epsilon}}{k^{1/2+\alpha_2+\alpha_3}}\max_{0<y<\pi/2} \frac{\chi_{1/2}(\cos^2y)W_{\pm}(y)}{k}\ll\frac{e^{-\pi|\Im{w}|/2}(1+|w|)^{\epsilon}k^{\epsilon}}{k}, \end{align} $$

completing the proof of (3.33).

4 The proof of Theorem 1.1

As one can see from (2.2) and (1.2), the term $\mathcal {MT}_D(\bar {\alpha })$ in (2.16) contains four of eight summands appearing in (1.2). The remaining four terms are hidden in the expressions $R_{i,j}(u,v,\rho ,k)$ . Using (2.11), (2.12), and (3.14), we obtain that $R_{2,1}$ and $R_{2,2}$ are bounded by $k^{-2+\epsilon }$ . It follows from [Reference Frolenkov6, (5.18)] and [Reference Frolenkov6, (5.19)] that

(4.1) $$ \begin{align} -R_{2,3}(u,v,\rho,k)&= \zeta(1-\alpha_1+\alpha_2)\zeta(1-\alpha_1-\alpha_3)\zeta(1+\alpha_2-\alpha_3) \nonumber\\ &\quad \times \mathcal{C}(-1,1,-1)\frac{\cos\frac{\pi(\alpha_2+\alpha_3)}{2}} {2\cos\frac{\pi(\alpha_1+\alpha_3)}{2}\cos\frac{\pi(\alpha_1-\alpha_2)}{2}}, \end{align} $$
(4.2) $$ \begin{align} -R_{2,4}(u,v,\rho,k)&=\zeta(1-\alpha_1-\alpha_2)\zeta(1-\alpha_1+\alpha_3)\zeta(1-\alpha_2+\alpha_3) \nonumber\\ &\quad \times \mathcal{C}(-1,-1,1)\frac{\cos\frac{\pi(\alpha_2+\alpha_3)}{2}} {2\cos\frac{\pi(\alpha_1+\alpha_2)}{2}\cos\frac{\pi(\alpha_1-\alpha_3)}{2}}. \end{align} $$

Now using (2.15), (4.1), and (4.2), the term $\mathcal {ET}_3(u,v,\rho )$ produces the following parts of the main term $\mathcal {MT}_3(\alpha _1,\alpha _2,\alpha _3)$ :

(4.3) $$ \begin{align} -R_{3,3}(u,v,\rho,k)&= \zeta(1-\alpha_1+\alpha_2)\zeta(1-\alpha_1+\alpha_3)\zeta(1+\alpha_2+\alpha_3)\nonumber\\ & \quad \times \mathcal{C}(-1,1,1)\frac{\cos\frac{\pi(\alpha_2-\alpha_3)}{2}} {2\cos\frac{\pi(\alpha_1-\alpha_3)}{2}\cos\frac{\pi(\alpha_1-\alpha_2)}{2}}, \end{align} $$

and

(4.4) $$ \begin{align} -R_{3,4}(u,v,\rho,k)&=\zeta(1-\alpha_1-\alpha_2)\zeta(1-\alpha_1-\alpha_3)\zeta(1-\alpha_2-\alpha_3) \nonumber\\ & \quad \times \mathcal{C}(-1,-1,-1)\frac{\cos\frac{\pi(\alpha_2-\alpha_3)}{2}} {2\cos\frac{\pi(\alpha_1+\alpha_2)}{2}\cos\frac{\pi(\alpha_1+\alpha_3)}{2}}. \end{align} $$

Substituting [Reference Frolenkov6, (4.29)] into [Reference Frolenkov6, (5.37)] and [Reference Frolenkov6, (5.38)], we obtain (the estimate [Reference Frolenkov6, (4.31)] could be easily generalized in case of small values of $u,v,\rho $ )

(4.5) $$ \begin{align} R_{1,3}(u,v,\rho,k)& =-\zeta(1-\alpha_1-\alpha_2)\zeta(1-\alpha_1-\alpha_3)\zeta(1-\alpha_2-\alpha_3) \nonumber\\ & \quad \times \frac{\Gamma(k-\alpha_1-\alpha_2-\alpha_3)}{\Gamma(k+\alpha_1+\alpha_2+\alpha_3)} \frac{(2\pi)^{2\alpha_1+2\alpha_2+2\alpha_3}\cos\frac{\pi(2\alpha_1+\alpha_2+\alpha_3)}{2}} {2\cos\frac{\pi(\alpha_1+\alpha_2)}{2}\cos\frac{\pi(\alpha_1+\alpha_3)}{2}} +O(k^{-1+\epsilon}), \end{align} $$
(4.6) $$ \begin{align} R_{1,4}(u,v,\rho,k)& =-\zeta(1-\alpha_1+\alpha_2)\zeta(1-\alpha_1+\alpha_3)\zeta(1+\alpha_2+\alpha_3) \nonumber\\ & \quad \times (-1)^k\mathcal{C}(-1,1,1) \frac{\cos\frac{\pi(2\alpha_1-\alpha_2-\alpha_3)}{2}} {2\cos\frac{\pi(\alpha_1-\alpha_2)}{2}\cos\frac{\pi(\alpha_1-\alpha_3)}{2}} +O(k^{-1+\epsilon}). \end{align} $$

Substituting (3.26) into (2.5), (2.6) and using the functional equation for the Riemann zeta functions:

$$ \begin{align*} &\zeta(\rho+u\pm v)\zeta(\rho-u\pm v)\Gamma(\rho+u\pm v)\Gamma(\rho-u\pm v)\\ & \quad = \frac{(2\pi)^{2\rho\pm2v}}{4\cos\frac{\pi(\rho+u\pm v)}{2} \cos\frac{\pi(\rho-u\pm v)}{2}} \zeta(1-\rho-u\mp v)\zeta(1-\rho+u\mp v), \end{align*} $$

we obtain

(4.7) $$ \begin{align} &R_{1,1}(u,v,\rho,k)=\zeta(1-\alpha_1-\alpha_2)\zeta(1-\alpha_1+\alpha_3)\zeta(1-\alpha_2+\alpha_3) \nonumber\\& \quad \times \frac{(-1)^k(2\pi)^{2\alpha_1+2\alpha_2}}{4\cos\frac{\pi(\alpha_1+\alpha_2)}{2}\cos\frac{\pi(\alpha_1-\alpha_3)}{2}} \left( \frac{\Gamma(k-\alpha_3)\Gamma(k-\alpha_1-\alpha_2+\alpha_3)}{\Gamma(k+\alpha_3)\Gamma(k+\alpha_1+\alpha_2-\alpha_3)} \frac{\sin(\pi\alpha_1)}{\sin\frac{\pi(\alpha_3-\alpha_2)}{2}}\right.\nonumber\\& \left. \qquad\qquad\qquad\qquad\qquad\quad\qquad + \frac{\Gamma(k-\alpha_1)\Gamma(k-\alpha_2)}{\Gamma(k+\alpha_1)\Gamma(k+\alpha_2)} \frac{\sin\pi(\alpha_1+\alpha_2-\alpha_3)}{\sin\frac{\pi(\alpha_2-\alpha_3)}{2}} \right)+O(k^{-1+\epsilon}),\nonumber\\[-24pt] \end{align} $$
(4.8) $$ \begin{align} &R_{1,2}(u,v,\rho,k) = \zeta(1-\alpha_1+\alpha_2)\zeta(1-\alpha_1-\alpha_3)\zeta(1+\alpha_2-\alpha_3) \nonumber\\& \quad \times \frac{(-1)^k(2\pi)^{2\alpha_1+2\alpha_3}}{4\cos\frac{\pi(\alpha_1+\alpha_3)}{2}\cos\frac{\pi(\alpha_1-\alpha_2)}{2}} \left(\frac{\Gamma(k-\alpha_2)\Gamma(k-\alpha_1+\alpha_2-\alpha_3)}{\Gamma(k+\alpha_2)\Gamma(k+\alpha_1-\alpha_2+\alpha_3)} \frac{\sin(\pi\alpha_1)}{\sin\frac{\pi(\alpha_2-\alpha_3)}{2}}\right.\nonumber\\&\left. \qquad\qquad\qquad\qquad\qquad\qquad\quad + \frac{\Gamma(k-\alpha_1)\Gamma(k-\alpha_3)}{\Gamma(k+\alpha_1)\Gamma(k+\alpha_3)} \frac{\sin\pi(\alpha_1-\alpha_2+\alpha_3)}{\sin\frac{\pi(\alpha_3-\alpha_2)}{2}} \right)+O(k^{-1+\epsilon}).\nonumber\\[-24pt] \end{align} $$

Therefore, we have considered all $R_{i,j}(u,v,\rho ,k)$ . Applying (4.3), (4.6) (note that there is a factor $(-1)^{k+1}$ in (2.4)), and

$$ \begin{align*} \cos\frac{\pi(\alpha_2-\alpha_3)}{2}+\cos\frac{\pi(2\alpha_1-\alpha_2-\alpha_3)}{2}= 2\cos\frac{\pi(\alpha_1-\alpha_3)}{2}\cos\frac{\pi(\alpha_1-\alpha_2)}{2}, \end{align*} $$

we find that

(4.9) $$ \begin{align} &-(-1)^kR_{1,4}(u,v,\rho,k)-R_{3,3}(u,v,\rho,k)\nonumber\\ & \quad= \zeta(1-\alpha_1+\alpha_2)\zeta(1-\alpha_1+\alpha_3)\zeta(1+\alpha_2+\alpha_3) \mathcal{C}(-1,1,1) +O(k^{-1+\epsilon}). \end{align} $$

To transform the remaining $R_{i,j}(\cdot )$ functions, we will apply the following asymptotic formula for (1.7):

(4.10) $$ \begin{align} X_{k}(\alpha)=\frac{(2\pi)^{2\alpha}}{k^{2\alpha}}\left(1+O\left(\frac{1}{k}\right)\right), \end{align} $$

which is a consequence of the Stirling formula [Reference Olver, Lozier, Boisvert and Clarke8, 5.11.3 and 5.11.13].

Applying (4.1), (4.8) (note that there is a factor $(-1)^{k}$ in (2.4)), and the relation (see (4.10))

$$ \begin{align*} (2\pi)^{2\alpha_1+2\alpha_3} \frac{\Gamma(k-\alpha_2)\Gamma(k-\alpha_1+\alpha_2-\alpha_3)}{\Gamma(k+\alpha_2)\Gamma(k+\alpha_1-\alpha_2+\alpha_3)} &=X_{k}(\alpha_2)X_{k}(\alpha_1-\alpha_2+\alpha_3)+O(k^{-1+\epsilon})\\ & =\mathcal{C}(-1,1,-1)+O(k^{-1+\epsilon}), \end{align*} $$

we find that

(4.11) $$ \begin{align} &(-1)^kR_{1,2}(u,v,\rho,k)-R_{2,3}(u,v,\rho,k) = \zeta(1-\alpha_1+\alpha_2)\zeta(1-\alpha_1-\alpha_3)\zeta(1+\alpha_2-\alpha_3) \nonumber\\& \qquad \times \frac{\mathcal{C}(-1,1,-1)} {4\cos\frac{\pi(\alpha_1+\alpha_3)}{2}\cos\frac{\pi(\alpha_1-\alpha_2)}{2}\sin\frac{\pi(\alpha_2-\alpha_3)}{2}}\nonumber\\& \qquad \times \left(2\cos\frac{\pi(\alpha_2+\alpha_3)}{2}\sin\frac{\pi(\alpha_2-\alpha_3)}{2} + \sin(\pi\alpha_1)- \sin\pi(\alpha_1-\alpha_2+\alpha_3) \vphantom{\frac{\pi(\alpha_2+\alpha_3)}{2}}\right)+O(k^{-1+\epsilon}) \nonumber\\& \quad = \zeta(1-\alpha_1+\alpha_2)\zeta(1-\alpha_1-\alpha_3)\zeta(1+\alpha_2-\alpha_3) \mathcal{C}(-1,1,-1)+O(k^{-1+\epsilon}).\nonumber\\ \end{align} $$

Using (4.2), (4.7), and the relation (see (4.10)):

$$ \begin{align*} (2\pi)^{2\alpha_1+2\alpha_2} \frac{\Gamma(k-\alpha_3)\Gamma(k-\alpha_1-\alpha_2+\alpha_3)}{\Gamma(k+\alpha_3)\Gamma(k+\alpha_1+\alpha_2-\alpha_3)} &=X_{k}(\alpha_3)X_{k}(\alpha_1+\alpha_2-\alpha_3)+O(k^{-1+\epsilon})\\ & = \mathcal{C}(-1,-1,1)+O(k^{-1+\epsilon}), \end{align*} $$

we prove that

(4.12) $$ \begin{align} & (-1)^kR_{1,1}(u,v,\rho,k)-R_{2,4}(u,v,\rho,k) \nonumber\\& \quad = \frac{\zeta(1-\alpha_1-\alpha_2)\zeta(1-\alpha_1+\alpha_3)\zeta(1-\alpha_2+\alpha_3)\mathcal{C}(-1,-1,1)} {4\cos\frac{\pi(\alpha_1+\alpha_2)}{2}\cos\frac{\pi(\alpha_1-\alpha_3)}{2}\sin\frac{\pi(\alpha_3-\alpha_2)}{2}} \nonumber\\& \qquad \times\left( 2\cos\frac{\pi(\alpha_2+\alpha_3)}{2}\sin\frac{\pi(\alpha_3-\alpha_2)}{2} +\sin(\pi\alpha_1)-\sin\pi(\alpha_1+\alpha_2-\alpha_3) \right) \nonumber\\& \qquad +O(k^{-1+\epsilon}) =\zeta(1-\alpha_1-\alpha_2)\zeta(1-\alpha_1+\alpha_3)\zeta(1-\alpha_2+\alpha_3)\mathcal{C}(-1,-1,1)+O(k^{-1+\epsilon}).\nonumber\\ \end{align} $$

Combining (4.4), (4.5), and the relation (see (4.10)):

$$ \begin{align*} (2\pi)^{2\alpha_1+2\alpha_2+2\alpha_3} \frac{\Gamma(k-\alpha_1-\alpha_2-\alpha_3)}{\Gamma(k+\alpha_1+\alpha_2+\alpha_3)} &=X_{k}(\alpha_1+\alpha_2+\alpha_3)+O(k^{-1+\epsilon})\\ & = (-1)^k\mathcal{C}(-1,-1,-1)+O(k^{-1+\epsilon}), \end{align*} $$

we infer that

(4.13) $$ \begin{align} &-(-1)^kR_{1,3}(u,v,\rho,k)-R_{3,4}(u,v,\rho,k) = \zeta(1-\alpha_1-\alpha_2)\zeta(1-\alpha_1-\alpha_3)\zeta(1-\alpha_2-\alpha_3) \nonumber\\& \qquad \times \frac{\mathcal{C}(-1,-1,-1)}{2\cos\frac{\pi(\alpha_1+\alpha_2)}{2}\cos\frac{\pi(\alpha_1+\alpha_3)}{2}} \left(\cos\frac{\pi(\alpha_2-\alpha_3)}{2}+ \cos\frac{\pi(2\alpha_1+\alpha_2+\alpha_3)}{2} \right)+O(k^{-1+\epsilon}) \nonumber\\& \quad = \zeta(1-\alpha_1-\alpha_2)\zeta(1-\alpha_1-\alpha_3)\zeta(1-\alpha_2-\alpha_3) \mathcal{C}(-1,-1,-1)+O(k^{-1+\epsilon}).\nonumber\\ \end{align} $$

Substituting (2.4), (2.10), (2.15) into (2.16) and applying (2.2), (4.9), (4.11), (4.12), (4.13), we obtain

(4.14) $$ \begin{align} &\mathcal{M}(\alpha_1,\alpha_2,\alpha_3) = \mathcal{MT}_3(\alpha_1,\alpha_2,\alpha_3)+ \frac{(-1)^k}{2\pi i}\int_{(0)} Z_{1}(u,v,\rho,k;w)dw \nonumber \\& \quad +\frac{1}{2\pi i}\int_{(0)} Z_{2}(u,v,\rho,k;w)+ (-1)^k\frac{\Gamma(k-u+v)}{\Gamma(k+u-v)}(2\pi)^{2u-2v}Z_{2}(v,u,\rho,k;w) dw\nonumber\\ & \quad +O(k^{-1+\epsilon}).\nonumber\\ \end{align} $$

Using (2.9) and (3.8), we show that

(4.15) $$ \begin{align} \mathcal{M}(\alpha_1,\alpha_2,\alpha_3) = \mathcal{MT}_3(\alpha_1,\alpha_2,\alpha_3)+ \frac{(-1)^k}{2\pi i}\int_{(0)} Z_{1}(u,v,\rho,k;-w)dw+O(k^{-1+\epsilon}). \end{align} $$

Applying the functional equation [Reference Olver, Lozier, Boisvert and Clarke8, 25.4.1] to the $\zeta \left (\frac {1+\rho }{2}-u-w\right )$ in (2.3), rewriting the obtained expression in terms of $\bar {\alpha }$ and using (1.8), we have

(4.16) $$ \begin{align} Z_{1}(u,v,\rho,k;-w)= \frac{\zeta_4(\bar{\alpha};w)(2\pi)^{\frac{1+\alpha_1-\alpha_2-\alpha_3}{2}+w} \hat{g}_{1}(u,v,\rho,k;-w)}{2\cos\frac{\pi}{2}\left(\frac{1+\alpha_1-\alpha_2-\alpha_3}{2}+w\right) \Gamma\left(\frac{1+\alpha_1-\alpha_2-\alpha_3}{2}+w\right)}. \end{align} $$

Substituting (3.30) into (4.16) and using (3.8), we find

(4.17) $$ \begin{align} &\frac{(-1)^k}{2\pi i}\int_{(0)} Z_{1}(u,v,\rho,k;-w)dw \nonumber\\& \quad = \frac{1}{2\pi i}\int_{(0)} \frac{\sin\pi(\alpha_1/2-w)\cos\pi\left(\frac{\alpha_1-\alpha_2-\alpha_3}{2}+w\right)} {\cos\frac{\pi}{2}\left(\frac{1+\alpha_1-\alpha_2-\alpha_3}{2}+w\right) \cos\pi\left(\frac{\alpha_1+\alpha_2+\alpha_3}{2}-w\right)} \nonumber\\& \qquad \times \mathcal{H}_1(\bar{\alpha},k;w)\zeta_4(\bar{\alpha};w)(2\pi)^{\frac{1+\alpha_1+\alpha_2+\alpha_3}{2}+w}dw +O(k^{-1+\epsilon}). \end{align} $$

Let $w=ir$ , then

(4.18) $$ \begin{align} &\frac{\sin\pi(\alpha_1/2-w)\cos\pi\left(\frac{\alpha_1-\alpha_2-\alpha_3}{2}+w\right)} {\cos\frac{\pi}{2}\left(\frac{1+\alpha_1-\alpha_2-\alpha_3}{2}+w\right) \cos\pi\left(\frac{\alpha_1+\alpha_2+\alpha_3}{2}-w\right)} \nonumber\\& \quad = \exp\left(\frac{\pi|r|}{2}-\pi i\text{sgn}(r) \frac{1+\alpha_1+\alpha_2+\alpha_3}{4}\right)+O(\exp(-\pi|r|/2)). \end{align} $$

Substituting (4.18) into (4.17) and using (3.33), we obtain

(4.19) $$ \begin{align} &\frac{(-1)^k}{2\pi i}\int_{(0)} Z_{1}(u,v,\rho,k;-w)dw \nonumber\\ & \quad = \frac{1}{2\pi i}\int_{(0)} l_1(\bar{\alpha},k;w) \mathcal{H}_1(\bar{\alpha},k;w)\zeta_4(\bar{\alpha};w)dw +O(k^{-1+\epsilon}). \end{align} $$

Finally, (1.12) follows from (4.15) and (4.19).

Acknowledgments

We thank the referee for helpful comments.

Footnotes

Research of Olga Balkanova and Dmitry Frolenkov was supported by the Theoretical Physics and Mathematics Advancement Foundation “BASIS.” Research of Brian Conrey was supported in part by an FRG grant from the NSF.

References

Balkanova, O. and Frolenkov, D., Moments of $L$ -functions and Liouville–Green method . J. Eur. Math. Soc. 23(2021), no. 4, 13331380.CrossRefGoogle Scholar
Conrey, J. B., Farmer, D. W., Keating, J. P., Rubinstein, M. O., and Snaith, N. C., Integral moments of L-functions . Proc. Lond. Math. Soc. (3) 91(2005), no. 1, 33104.CrossRefGoogle Scholar
Erdelyi, A., Magnus, W., Oberhettinger, F., and Tricomi, F. G., Higher transcendental functions. Vol. I, based, in part, on notes left by Harry Bateman, McGraw-Hill, New York, 1953.Google Scholar
Farid Khwaja, S. and Olde Daalhuis, A. B., Uniform asymptotic expansions for hypergeometric functions with large parameters IV . Anal. Appl. (Singap.) 12(2014), 667710.CrossRefGoogle Scholar
Farid Khwaja, S. and Olde Daalhuis, A. B., Computation of the coefficients appearing in the uniform asymptotic expansions of integrals . Stud. Appl. Math. 139(2017), 551567.CrossRefGoogle Scholar
Frolenkov, D., The cubic moment of automorphic $L$ -functions in the weight aspect . J. Number Theory 207(2020), 247281.CrossRefGoogle Scholar
Gradshteyn, I. S. and Ryzhik, I. M., Table of integrals, series, and products. 7th ed., Jeffrey, A. and Zwillinger, D. (eds.), Academic Press, New York, 2007.Google Scholar
Olver, F. W. J., Lozier, D. W., Boisvert, R. F., and Clarke, C. W., NIST handbook of mathematical functions, Cambridge University Press, Cambridge, 2010.Google Scholar
Prudnikov, A. P., Brychkov, Y. A., and Marichev, O. I., Integral and series, More Special Functions, 3, translated by G. G. Gould, Gordon and Breach Science Publishers, New York, 1986.Google Scholar
Temme, N. M., Asymptotic methods for integrals, Series in Analysis, 6, World Scientific, Hackensack, 2015.Google Scholar