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

Uniform bounds for Kloosterman sums of half-integral weight, same-sign case

Qihang Sun Department of Mathematics, University of Illinois, Urbana, IL 61801 [email protected]
Abstract.

In the previous paper [Sun23], the author proved a uniform bound for sums of half-integral weight Kloosterman sums. This bound was applied to prove an exact formula for partitions of rank modulo 3. That uniform estimate provides a more precise bound for a certain class of multipliers compared to the 1983 result by Goldfeld and Sarnak and generalizes the 2009 result from Sarnak and Tsimerman to the half-integral weight case. However, the author only considered the case when the parameters satisfied m~n~<0\tilde{m}\tilde{n}<0. In this paper, we prove the same uniform bound when m~n~>0\tilde{m}\tilde{n}>0 for further applications.

Key words and phrases:
Kloosterman sum; Maass form; Kuznetsov trace formula

1. Introduction

For positive integers NN, cc and m,nm,n\in{\mathbb{Z}}, we define the generalized Kloosterman sums with a multiplier system ν\nu as

S(m,n,c,ν)=0a,d<cγ=(abcd)Γν¯(γ)(m~a+n~dc)S(m,n,c,\nu)=\sum_{\begin{subarray}{c}0\leq a,d<c\\ \gamma=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\Gamma\end{subarray}}\overline{\nu}(\gamma)\left(\frac{\tilde{m}a+\tilde{n}d}{c}\right)

where Γ=Γ0(N)\Gamma=\Gamma_{0}(N) is a congruence subgroup of SL2(){\rm SL}_{2}({\mathbb{Z}}), ν\nu is a weight kk\in{\mathbb{R}} multiplier system on Γ\Gamma, and n~=nαν,\tilde{n}=n-\alpha_{\nu,\infty} as defined in (2.6). These Kloosterman sums have been studied by Goldfeld and Sarnak [GS83] and Pribitkin [Pri00]. In a previous paper [Sun23], the author proved a uniform bound for the sums of such Kloosterman sums and applied the bound to estimate the partial sums of Rademacher-type exact formulas. For example, the Fourier coefficient G(n)G(n) of γ(q)\gamma(q), which is a sixth order mock theta function defined in [AH91, (0.17)], can be written as [Sun23, Theorem 2.2]

G(n)=A(13;n)=2πe(18)(24n1)143|c>0S(0,n,c,(3)νη¯)cI12(π24n16c),G(n)=A\left(\frac{1}{3};n\right)=\frac{2\pi\,e(-\frac{1}{8})}{(24n-1)^{\frac{1}{4}}}\sum_{3|c>0}\frac{S(0,n,c,\,(\frac{\cdot}{3})\overline{\nu_{\eta}})}{c}I_{\frac{1}{2}}\left(\frac{\pi\sqrt{24n-1}}{6c}\right), (1.1)

where νη\nu_{\eta} is the multiplier system of weight 12\frac{1}{2} for Dedekind’s eta-function (see (2.4)) and IκI_{\kappa} is the II-Bessel function. The author bounded the error

R3(n,x)=2πe(18)(24n1)143|c>xS(0,n,c,(3)νη¯)cI12(π24n16c)R_{3}(n,x)=\frac{2\pi\,e(-\frac{1}{8})}{(24n-1)^{\frac{1}{4}}}\sum_{3|c>x}\frac{S(0,n,c,\,(\frac{\cdot}{3})\overline{\nu_{\eta}})}{c}I_{\frac{1}{2}}\left(\frac{\pi\sqrt{24n-1}}{6c}\right) (1.2)

with [Sun23, Theorem 1.6] to get

R3(n,αn)α,εn1147+ε.R_{3}(n,\alpha\sqrt{n})\ll_{\alpha,\varepsilon}n^{-\frac{1}{147}+\varepsilon}.

Our estimate can be applied to sums of Kloosterman sum with a class of multiplier systems. However, in the former paper [Sun23] we only focus on the case m~n~<0\tilde{m}\tilde{n}<0. In this paper we add the complementary case m~n~>0\tilde{m}\tilde{n}>0 for further applications like Corollary 1.6.

To state the results, we need to classify our half-integral weight multipliers first:

Definition 1.1.

Let k=±12k=\pm\frac{1}{2} and ν=(|D|)νθ2k\nu^{\prime}=(\frac{|D|}{\cdot})\nu_{\theta}^{2k} where DD is some even fundamental discriminant and νθ\nu_{\theta} is the multiplier for the theta function. We say a weight kk multiplier ν\nu on Γ=Γ0(N)\Gamma=\Gamma_{0}(N) is admissible if it satisfies the following two conditions:

  • (1)

    Level lifting: there exist positive integers BB and MM such that the map :(f)(z)=f(Bz)\mathscr{L}:(\mathscr{L}f)(z)=f(Bz) gives:

    • (i)

      an injection from weight kk automorphic eigenforms of the hyperbolic Laplacian Δk\Delta_{k} on (Γ0(N),ν)(\Gamma_{0}(N),\nu) to those on (Γ0(M),ν)(\Gamma_{0}(M),\nu^{\prime}) and keeps the eigenvalue;

    • (ii)

      an injection from weight kk holomorphic cusp forms on (Γ0(N),ν)(\Gamma_{0}(N),\nu) to weight kk holomorphic cusp forms on (Γ0(M),ν)(\Gamma_{0}(M),\nu^{\prime}).

    Here MM is a multiple of 44 and MM depends on BB.

  • (2)

    Average Weil bound: for x>y>0x>y>0 and xyx23x-y\gg x^{\frac{2}{3}}, we have

    N|c[x,y]|S(m,n,c,ν)|cN,ν,ε(xy)(m~n~x)ε.\sum_{N|c\,\in[x,y]}\frac{|S(m,n,c,\nu)|}{c}\ll_{N,\nu,\varepsilon}(\sqrt{x}-\sqrt{y})(\tilde{m}\tilde{n}x)^{\varepsilon}.

The author proved [Sun23, Proposition 5.1] that the following class of multipliers satisfy both the conditions:

Lemma 1.2.

Let ν=(|D|)νθ\nu=(\frac{|D|}{\cdot})\nu_{\theta} or ν=(|D|)νη\nu=(\frac{|D|}{\cdot})\nu_{\eta} where DD is a fundamental discriminant and νθ\nu_{\theta} and νη\nu_{\eta} are the multiplier system for the standard theta function and Dedekind’s eta function, respectively (see (2.1) and (2.4)). Then both ν\nu and its conjugate are admissible.

Let ρj(n)\rho_{j}(n) denote the nn-th Fourier coefficient of an orthonormal basis {vj()}j\{v_{j}(\cdot)\}_{j} of ~k(N,ν){\tilde{\mathcal{L}}}_{k}(N,\nu) (the space of square-integrable eigenforms of the weight kk Laplacian with respect to (Γ0(N),ν)(\Gamma_{0}(N),\nu), see Section 2 for details). Here are our theorems:

Theorem 1.3.

Suppose m~>0\tilde{m}>0, n~>0\tilde{n}>0 and ν\nu is a weight k=±12k=\pm\frac{1}{2} admissible multiplier on Γ0(N)\Gamma_{0}(N). We have

N|cXS(m,n,c,ν)c=rji(0,14]τj(m,n)X2sj12sj1+Oν,ε((Au(m,n)+X16)(m~n~X)ε),\sum_{N|c\leq X}\frac{S(m,n,c,\nu)}{c}=\!\!\!\sum_{r_{j}\in i(0,\frac{1}{4}]}\!\!\!\tau_{j}(m,n)\frac{X^{2s_{j}-1}}{2s_{j}-1}+O_{\nu,\varepsilon}\left(\left(A_{u}(m,n)+X^{\frac{1}{6}}\right)(\tilde{m}\tilde{n}X)^{\varepsilon}\right), (1.3)

where for BB and MM in Definition 1.1, we factor B~=tu2w2B\tilde{\ell}=t_{\ell}u_{\ell}^{2}w_{\ell}^{2} with tt_{\ell} square-free, u|Mu_{\ell}|M^{\infty} positive and (w,M)=1(w_{\ell},M)=1 for {m,n}\ell\in\{m,n\}. Here sj=Imrj+12s_{j}=\operatorname{Im}r_{j}+\frac{1}{2},

τj(m,n)=2ikρj(m)¯ρj(n)π12sj(4m~n~)1sjΓ(sj+k2)Γ(2sj1)Γ(sjk2)\tau_{j}(m,n)=2i^{k}\overline{\rho_{j}(m)}\rho_{j}(n)\pi^{1-2s_{j}}(4\tilde{m}\tilde{n})^{1-s_{j}}\frac{\Gamma(s_{j}+\frac{k}{2})\Gamma(2s_{j}-1)}{\Gamma(s_{j}-\frac{k}{2})}

are the coefficients in [GS83] (as corrected by [AA16, Proposition 7]), and

Au(m,n)\displaystyle A_{u}(m,n) :=(m~131294+um)18(n~131294+un)18(m~n~)316\displaystyle\vcentcolon=\left(\tilde{m}^{\frac{131}{294}}+u_{m}\right)^{\frac{1}{8}}\left(\tilde{n}^{\frac{131}{294}}+u_{n}\right)^{\frac{1}{8}}(\tilde{m}\tilde{n})^{\frac{3}{16}}
(m~n~)143588+m~143588n~316un18+m~316n~143588um18+(m~n~)316(umun)18.\displaystyle\ll(\tilde{m}\tilde{n})^{\frac{143}{588}}+\tilde{m}^{\frac{143}{588}}\tilde{n}^{\frac{3}{16}}\,u_{n}^{\frac{1}{8}}+\tilde{m}^{\frac{3}{16}}\tilde{n}^{\frac{143}{588}}\,u_{m}^{\frac{1}{8}}+(\tilde{m}\tilde{n})^{\frac{3}{16}}(u_{m}u_{n})^{\frac{1}{8}}.
Remark.

We have the following notes for this theorem:

  • The notation u|Mu|M^{\infty} means u|MCu|M^{C} for some positive integer CC.

  • When umu_{m} and unu_{n} are both ON,ν(1)O_{N,\nu}(1), Au(m,n)N,ν(m~n~)143588A_{u}(m,n)\ll_{N,\nu}(\tilde{m}\tilde{n})^{\frac{143}{588}}.

  • In general, Au(m,n)N,ν(m~n~)14A_{u}(m,n)\ll_{N,\nu}(\tilde{m}\tilde{n})^{\frac{1}{4}}.

  • When k=12k=-\frac{1}{2} and rj=i4r_{j}=\frac{i}{4}, we have τj(m,n)=0\tau_{j}(m,n)=0 (see (2.21) and (2.22)).

  • The theorem also applies to the case m~<0\tilde{m}<0 and n~<0\tilde{n}<0 because of (2.8) by conjugation.

Based on Theorem 1.3 and [Sun23, Theorem 1.4], we are able to reformulate the Linnik-Selberg conjecture in the half-integral weight case:

Conjecture 1.4.

Suppose k=±12k=\pm\frac{1}{2} and ν\nu is a weight kk multiplier system on Γ0(N)\Gamma_{0}(N). If there is no eigenvalue for the hyperbolic Laplacian Δk\Delta_{k} in (316,14)(\frac{3}{16},\frac{1}{4}), then

N|cXS(m,n,c,ν)c2X12rj=i4τj(m,n)N,ν,ε|m~n~x|ε,\sum_{N|c\leq X}\frac{S(m,n,c,\nu)}{c}-2X^{\frac{1}{2}}\sum_{r_{j}=\frac{i}{4}}\tau_{j}(m,n)\ll_{N,\nu,\varepsilon}|\tilde{m}\tilde{n}x|^{\varepsilon}, (1.4)

where the second sum runs over normalized eigenforms of Δk\Delta_{k} with eigenvalue 14+rj2=316\frac{1}{4}+r_{j}^{2}=\frac{3}{16}.

We also get the following bound by the properties of Bessel functions.

Theorem 1.5.

With the same setting as Theorem 1.3, for β=12\beta=\frac{1}{2} or 32\frac{3}{2}, we have

N|c>αm~n~S(m,n,c,ν)cβ(4πm~n~c)α,ν,εAu(m,n)(m~n~)ε\sum_{N|c>\alpha\sqrt{\tilde{m}\tilde{n}}}\frac{S(m,n,c,\nu)}{c}\mathscr{B}_{\beta}\left(\frac{4\pi\sqrt{\tilde{m}\tilde{n}}}{c}\right)\ll_{\alpha,\nu,\varepsilon}A_{u}(m,n)(\tilde{m}\tilde{n})^{\varepsilon} (1.5)

except the case when m~\tilde{m}, n~\tilde{n} and k=±12k=\pm\frac{1}{2} are all of the same sign such that rj=i4τj(m,n)0\sum_{r_{j}=\frac{i}{4}}\tau_{j}(m,n)\neq 0. Here β\mathscr{B}_{\beta} is the Bessel function IβI_{\beta} or JβJ_{\beta}. Note that Au(m,n)N,ν(m~n~)143588A_{u}(m,n)\ll_{N,\nu}(\tilde{m}\tilde{n})^{\frac{143}{588}} when umu_{m} and unu_{n} are ON,ν(1)O_{N,\nu}(1).

1.1. Application

In [BO12], Bringmann and Ono constructed Maass-Poincaré series to prove that the Fourier coefficients of weight k+12k\in{\mathbb{Z}}+\frac{1}{2}, k12k\leq\frac{1}{2} harmonic Maass forms have exact formulas, i.e. they equal infinite sums of Kloosterman sums (up to Fourier coefficients from holomorphic theta functions when the weight is 12\frac{1}{2}). The specific case, weight 12\frac{1}{2}, is crucial as it is related to the rank of partitions. Readers may also refer to [BO06] or [Sun23, Section 4] for further introduction.

We can denote the weight 12\frac{1}{2} Maass-Poincaré series as P𝔞(z,m,Γ0(N),12,s,ν)P_{\mathfrak{a}}(z,m,\Gamma_{0}(N),\frac{1}{2},s,\nu) for cusp 𝔞\mathfrak{a} and multiplier ν\nu of Γ0(N)\Gamma_{0}(N) as [BO12, (3.4)]. The exact formulas are derived from P𝔞P_{\mathfrak{a}} at s=34s=\frac{3}{4}, while we only know the convergence at Res>1\operatorname{Re}s>1 (see [BO12, Lemma 3.1, (3.7)]). Bringmann and Ono called a weight 12\frac{1}{2} harmonic Maass form on (Γ0(N),ν)(\Gamma_{0}(N),\nu) is good if those Maass-Poincaré series corresponding to nontrivial terms in the principal parts of ff are individually convergent. They conclude the exact formula for its Fourier coefficients only when ff is good.

With Theorem 1.5 and [BO12, Remark (1) after Theorem 3.2], we have

Corollary 1.6.

Suppose ff is a weight 12\frac{1}{2} harmonic Maass form on (Γ0(N),ν)(\Gamma_{0}(N),\nu) and ff only has non-trivial principal part at cusp \infty. Then ff is good if ν\nu is admissible (Definition 1.1).

The paper is organized as follows. In Section 2 we introduce notations on Kloosterman sums and Maass forms. Section 3 revisits the trace formula by Proskurin [Pro79] and discusses certain properties crucial for the subsequent proof. Section 4 is about bounds on translations of the test function. The proof of Theorem 1.3 is presented in Section 5 and is divided into two cases: the weight k=12k=\frac{1}{2} and k=12k=-\frac{1}{2}. Readers are advised to take particular note of the Remark following Proposition 5.2 and be mindful of the weight context in Section 5.

2. Background

In this section we recall some basic notions on Maass forms with general weight and multiplier. Let Γ=Γ0(N)\Gamma=\Gamma_{0}(N) for some N1N\geq 1 denote our congruence subgroup and {\mathbb{H}} denote the upper half complex plane. Fixing the argument in (π,π](-\pi,\pi], for any γSL2()\gamma\in{\rm SL}_{2}({\mathbb{R}}) and τ=x+iy\tau=x+iy\in{\mathbb{H}}, we define the automorphic factor

j(γ,τ):=cτ+d|cτ+d|=eiarg(cτ+d)j(\gamma,\tau)\vcentcolon=\frac{c\tau+d}{|c\tau+d|}=e^{i\arg(c\tau+d)}

and the weight kk slash operator

f|kγ:=j(γ,τ)kf(γτ)f|_{k}\gamma\vcentcolon=j(\gamma,\tau)^{-k}f(\gamma\tau)

for kk\in{\mathbb{R}}. We say that ν:Γ×\nu:\Gamma\to{\mathbb{C}}^{\times} is a multiplier system of weight kk if

  1. (i)

    |ν|=1|\nu|=1,

  2. (ii)

    ν(I)=eπik\nu(-I)=e^{-\pi ik}, and

  3. (iii)

    ν(γ1γ2)=wk(γ1,γ2)ν(γ1)ν(γ2)\nu(\gamma_{1}\gamma_{2})=w_{k}(\gamma_{1},\gamma_{2})\nu(\gamma_{1})\nu(\gamma_{2}) for all γ1,γ2Γ\gamma_{1},\gamma_{2}\in\Gamma, where

    wk(γ1,γ2):=j(γ2,τ)kj(γ1,γ2τ)kj(γ1γ2,τ)k.w_{k}(\gamma_{1},\gamma_{2})\vcentcolon=j(\gamma_{2},\tau)^{k}j(\gamma_{1},\gamma_{2}\tau)^{k}j(\gamma_{1}\gamma_{2},\tau)^{-k}.

If ν\nu is a multiplier system of weight kk, then it is also a multiplier system of weight k±2k\pm 2 and its conjugate ν¯\overline{\nu} is a multiplier system of weight k-k.

We are interested in the following multiplier systems of weight 12\frac{1}{2} and their conjugates of weight 12-\frac{1}{2}. The theta-multiplier νθ\nu_{\theta} on Γ0(4)\Gamma_{0}(4) is given by

θ(γz)=νθ(γ)cz+dθ(z),γ=(abcd)Γ0(4)\theta(\gamma z)=\nu_{\theta}(\gamma)\sqrt{cz+d}\;\theta(z),\quad\gamma=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\Gamma_{0}(4) (2.1)

where

θ(z):=ne(n2z),νθ(γ)=(cd)εd1,εd={1d1(mod 4),id3(mod 4),\theta(z)\vcentcolon=\sum_{n\in{\mathbb{Z}}}e(n^{2}z),\quad\nu_{\theta}(\gamma)=\left(\frac{c}{d}\right)\varepsilon_{d}^{-1},\quad\varepsilon_{d}=\left\{\begin{array}[]{ll}1&d\equiv 1\ (\mathrm{mod}\ 4),\\ i&d\equiv 3\ (\mathrm{mod}\ 4),\end{array}\right.

and ()\left(\frac{\cdot}{\cdot}\right) is the extended Kronecker symbol such that (1d)=(1)d12=εd2\left(\tfrac{-1}{d}\right)=(-1)^{\frac{d-1}{2}}=\varepsilon_{d}^{2} for odd dd\in{\mathbb{Z}}. The eta-multiplier νη\nu_{\eta} on SL2(){\rm SL}_{2}({\mathbb{Z}}) is given by

η(γz)=νη(γ)cz+dη(z),γ=(abcd)SL2()\eta(\gamma z)=\nu_{\eta}(\gamma)\sqrt{cz+d}\;\eta(z),\quad\gamma=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in{\rm SL}_{2}({\mathbb{Z}}) (2.2)

where

η(z):=q124n=1(1qn)andq=e(z):=e2πiz.\eta(z)\vcentcolon=q^{\frac{1}{24}}\prod_{n=1}^{\infty}(1-q^{n})\quad\text{and}\quad q=e(z)\vcentcolon=e^{2\pi iz}. (2.3)

Explicit formulas for νη\nu_{\eta} were given by Rademacher [Rad73, (74.11), (74.12)]:

νη(γ)=e(18)eπis(d,c)e(a+d24c),s(d,c):=r=1c1rc(drcdrc1),\nu_{\eta}(\gamma)=e\left(-\frac{1}{8}\right)e^{-\pi is(d,c)}e\left(\frac{a+d}{24c}\right),\quad s(d,c)\vcentcolon=\sum_{r=1}^{c-1}\frac{r}{c}\left(\frac{dr}{c}-\left\lfloor\frac{dr}{c}\right\rfloor-1\right), (2.4)

for all cc\in{\mathbb{Z}} and also given by Knopp [Kno70]:

νη(γ)={(dc)e(124((a+d)cbd(c21)3c))if c is odd,(cd)e(124((a+d)cbd(c21)+3d33cd))if c is even,\nu_{\eta}(\gamma)=\left\{\begin{array}[]{ll}\left(\dfrac{d}{c}\right)e\left(\dfrac{1}{24}\Big{(}(a+d)c-bd(c^{2}-1)-3c\Big{)}\right)&\text{if }c\text{ is odd,}\\ \left(\dfrac{c}{d}\right)e\left(\dfrac{1}{24}\Big{(}(a+d)c-bd(c^{2}-1)+3d-3-3cd\Big{)}\right)&\text{if }c\text{ is even,}\end{array}\right. (2.5)

for c>0c>0. The properties νη(±(1b01))=e(b24)\nu_{\eta}\left(\pm\begin{pmatrix}1&b\\ 0&1\end{pmatrix}\right)=e\left(\tfrac{b}{24}\right) and νη(γ)=iνη(γ)\nu_{\eta}(-\gamma)=i\nu_{\eta}(\gamma) for c>0c>0 are convenient in computations.

2.1. Kloosterman sums

For any cusp 𝔞\mathfrak{a} of Γ=Γ0(N)\Gamma=\Gamma_{0}(N), let Γ𝔞\Gamma_{\mathfrak{a}} denote its stabilizer in Γ\Gamma. For example, Γ={±(1b01):b}\Gamma_{\infty}=\{\pm\begin{pmatrix}1&b\\ 0&1\end{pmatrix}:b\in{\mathbb{Z}}\}. Let σ𝔞SL2()\sigma_{\mathfrak{a}}\in{\rm SL}_{2}({\mathbb{R}}) denote its scaling matrix satisfying σ𝔞=𝔞\sigma_{\mathfrak{a}}\infty=\mathfrak{a} and σ𝔞1Γ𝔞σ𝔞=Γ\sigma_{\mathfrak{a}}^{-1}\Gamma_{\mathfrak{a}}\sigma_{\mathfrak{a}}=\Gamma_{\infty}. We define αν,𝔞[0,1)\alpha_{\nu,\mathfrak{a}}\in[0,1) by the condition

ν(σ𝔞(1101)σ𝔞1)=e(αν,𝔞).\nu\left(\sigma_{\mathfrak{a}}\begin{pmatrix}1&1\\ 0&1\end{pmatrix}\sigma_{\mathfrak{a}}^{-1}\right)=e(-\alpha_{\nu,\mathfrak{a}}). (2.6)

The cusp 𝔞\mathfrak{a} is called singular if αν,𝔞=0\alpha_{\nu,\mathfrak{a}}=0. When 𝔞=\mathfrak{a}=\infty, we drop the subscript, denote αν:=αν,\alpha_{\nu}\vcentcolon=\alpha_{\nu,\infty}, and define n~:=nαν\tilde{n}\vcentcolon=n-\alpha_{\nu} for nn\in{\mathbb{Z}}. The Kloosterman sum at the cusp pair (,)(\infty,\infty) with respect to the multiplier system ν\nu is given by

S(m,n,c,ν):=0a,d<cγ=(abcd)Γν¯(γ)e(m~a+n~dc)=γΓΓ/Γγ=(abcd)ν¯(γ)e(m~a+n~dc).S(m,n,c,\nu):=\!\!\!\sum_{\begin{subarray}{c}0\leq a,d<c\\ \gamma=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\Gamma\end{subarray}}\!\!\!\overline{\nu}(\gamma)e\left(\frac{\tilde{m}a+\tilde{n}d}{c}\right)=\!\!\!\sum_{\begin{subarray}{c}\gamma\in\Gamma_{\infty}\setminus\Gamma/\Gamma_{\infty}\\ \gamma=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\end{subarray}}\!\!\!\overline{\nu}(\gamma)e\left(\frac{\tilde{m}a+\tilde{n}d}{c}\right). (2.7)

They have the relationships

S(m,n,c,ν)¯={S(1m,1n,c,ν¯)if αν>0,S(m,n,c,ν¯)if αν=0,\overline{S(m,n,c,\nu)}=\left\{\begin{array}[]{ll}S(1-m,1-n,c,\overline{\nu})&\ \ \text{if\ }\alpha_{\nu}>0,\\ S(-m,-n,c,\overline{\nu})&\ \ \text{if\ }\alpha_{\nu}=0,\end{array}\right. (2.8)

because

nν¯={(1n)νif αν>0,nif αν=0.n_{\overline{\nu}}=\left\{\begin{array}[]{ll}-(1-n)_{\nu}&\ \ \text{if\ }\alpha_{\nu}>0,\\ n&\ \ \text{if\ }\alpha_{\nu}=0.\end{array}\right. (2.9)

2.2. Maass forms

We call a function f:f:{\mathbb{H}}\rightarrow{\mathbb{C}} automorphic of weight kk and multiplier ν\nu on Γ\Gamma if

f|kγ=ν(γ)ffor all γΓ.f|_{k}\gamma=\nu(\gamma)f\quad\text{for all }\gamma\in\Gamma.

Let 𝒜k(N,ν)\mathcal{A}_{k}(N,\nu) denote the linear space consisting of all such functions and k(N,ν)𝒜k(N,ν){\mathcal{L}}_{k}(N,\nu)\subset\mathcal{A}_{k}(N,\nu) denote the space of square-integrable functions on Γ\Gamma\setminus{\mathbb{H}} with respect to the measure

dμ(z)=dxdyy2d\mu(z)=\frac{dxdy}{y^{2}}

and the Petersson inner product

f,g:=Γf(z)g(z)¯dxdyy2\langle f,g\rangle\vcentcolon=\int_{\Gamma\setminus{\mathbb{H}}}f(z)\overline{g(z)}\frac{dxdy}{y^{2}}

for f,gk(N,ν)f,g\in{\mathcal{L}}_{k}(N,\nu). For kk\in{\mathbb{R}}, the Laplacian

Δk:=y2(2x2+2y2)ikyx\Delta_{k}\vcentcolon=y^{2}\left(\frac{\partial^{2}}{\partial x^{2}}+\frac{\partial^{2}}{\partial y^{2}}\right)-iky\frac{\partial}{\partial x} (2.10)

can be expressed as

Δk\displaystyle\Delta_{k} =Rk2Lkk2(1k2)\displaystyle=-R_{k-2}L_{k}-\frac{k}{2}\left(1-\frac{k}{2}\right) (2.11)
=Lk+2Rk+k2(1+k2)\displaystyle=-L_{k+2}R_{k}+\frac{k}{2}\left(1+\frac{k}{2}\right) (2.12)

where RkR_{k} is the Maass raising operator

Rk:=k2+2iyz=k2+iy(xiy)R_{k}\vcentcolon=\frac{k}{2}+2iy\frac{\partial}{\partial z}=\frac{k}{2}+iy\left(\frac{\partial}{\partial x}-i\frac{\partial}{\partial y}\right) (2.13)

and LkL_{k} is the Maass lowering operator

Lk:=k2+2iyz¯=k2+iy(x+iy).L_{k}\vcentcolon=\frac{k}{2}+2iy\frac{\partial}{\partial\bar{z}}=\frac{k}{2}+iy\left(\frac{\partial}{\partial x}+i\frac{\partial}{\partial y}\right). (2.14)

These operators raise and lower the weight of an automorphic form as

(Rkf)|k+2γ=Rk(f|kγ),(Lkf)|k2γ=Lk(f|kγ),for f𝒜k(N,ν)(R_{k}f)|_{k+2}\;\gamma=R_{k}(f|_{k}\gamma),\quad(L_{k}f)|_{k-2}\;\gamma=L_{k}(f|_{k}\gamma),\quad\text{for\ }f\in\mathcal{A}_{k}(N,\nu)

and satisfy the commutative relations

RkΔk=Δk+2Rk,LkΔk=Δk2Lk.R_{k}\Delta_{k}=\Delta_{k+2}R_{k},\quad L_{k}\Delta_{k}=\Delta_{k-2}L_{k}. (2.15)

Moreover, Δk\Delta_{k} commutes with the weight kk slash operator for all γSL2()\gamma\in{\rm SL}_{2}({\mathbb{R}}).

We call a real analytic function f:f:{\mathbb{H}}\rightarrow{\mathbb{C}} an eigenfunction of Δk\Delta_{k} with eigenvalue λ\lambda\in{\mathbb{C}} if

Δkf+λf=0.\Delta_{k}f+\lambda f=0.

From (2.15), it is clear that an eigenvalue λ\lambda for the weight kk Laplacian is also an eigenvalue for weight k±2k\pm 2. We call f𝒜k(N,ν)f\in\mathcal{A}_{k}(N,\nu) a Maass form if it is an smooth eigenfunction of Δk\Delta_{k} and satisfies the growth condition

(f|kγ)(x+iy)yσ+y1σ(f|_{k}\gamma)(x+iy)\ll y^{\sigma}+y^{1-\sigma}

for all γSL2()\gamma\in{\rm SL}_{2}({\mathbb{Z}}) and some σ\sigma depending on γ\gamma when y+y\rightarrow+\infty. Moreover, if a Maass form ff satisfies

01(f|kσ𝔞)(x+iy)e(αν,𝔞x)𝑑x=0\int_{0}^{1}(f|_{k}\sigma_{\mathfrak{a}})(x+iy)\;e(\alpha_{\nu,\mathfrak{a}}x)dx=0

for all cusps 𝔞\mathfrak{a} of Γ\Gamma, then fk(N,ν)f\in{\mathcal{L}}_{k}(N,\nu) and we call ff a Maass cusp form. For details see [AA18, §2.3]

Let k(N,ν)k(N,ν)\mathcal{B}_{k}(N,\nu)\subset{\mathcal{L}}_{k}(N,\nu) denote the space of smooth functions ff such that both ff and Δkf\Delta_{k}f are bounded. One can show that k(N,ν)\mathcal{B}_{k}(N,\nu) is dense in k(N,ν){\mathcal{L}}_{k}(N,\nu) and Δk\Delta_{k} is self-adjoint on k(N,ν)\mathcal{B}_{k}(N,\nu). If we let λ0:=λ0(k)=|k|2(1|k|2)\lambda_{0}\vcentcolon=\lambda_{0}(k)=\tfrac{|k|}{2}(1-\tfrac{|k|}{2}), then for fk(N,ν)f\in\mathcal{B}_{k}(N,\nu),

f,Δkfλ0f,f,\langle f,-\Delta_{k}f\rangle\geq\lambda_{0}\langle f,f\rangle,

i.e. Δk-\Delta_{k} is bounded from below. By the Friedrichs extension theorem, Δk-\Delta_{k} can be extended to a self-adjoint operator on k(N,ν){\mathcal{L}}_{k}(N,\nu). The spectrum of Δk\Delta_{k} consists of two parts: the continuous spectrum λ[14,)\lambda\in[\frac{1}{4},\infty) and a discrete spectrum of finite multiplicity contained in [λ0,)[\lambda_{0},\infty).

Let λΔ(G,ν,k)\lambda_{\Delta}(G,\nu,k) denote the first eigenvalue larger than λ0\lambda_{0} in the discrete spectrum with respect to the congruence subgroup GG, weight kk and multiplier ν\nu. For weight 0, Selberg [Sel65] showed that λΔ(Γ(N),𝟏,0)316\lambda_{\Delta}(\Gamma(N),\boldsymbol{1},0)\geq\frac{3}{16} for all NN and Selberg’s famous eigenvalue conjecture states that λΔ(G,𝟏,0)14\lambda_{\Delta}(G,\boldsymbol{1},0)\geq\frac{1}{4} for all GG. We introduce the hypothesis HθH_{\theta} as

Hθ:λΔ(Γ0(N),𝟏,0)14θ2for all N.H_{\theta}:\quad\lambda_{\Delta}(\Gamma_{0}(N),\boldsymbol{1},0)\geq\tfrac{1}{4}-\theta^{2}\ \ \text{for all\ }N. (2.16)

Selberg’s conjecture includes H0H_{0} and the best progress known today is H764H_{\frac{7}{64}} by [Kim03]. We denote λΔ(G,ν,k)\lambda_{\Delta}(G,\nu,k) as λΔ\lambda_{\Delta} when (G,ν,k)(G,\nu,k) is clear from context.

Let ~k(N,ν)k(N,ν){\tilde{\mathcal{L}}}_{k}(N,\nu)\subset{\mathcal{L}}_{k}(N,\nu) denote the subspace spanned by eigenfunctions of Δk\Delta_{k}. For each eigenvalue λ\lambda, we write

λ=14+r2=s(1s),s=12+ir,ri(0,14][0,).\lambda=\tfrac{1}{4}+r^{2}=s(1-s),\quad s=\tfrac{1}{2}+ir,\quad r\in i(0,\tfrac{1}{4}]\cup[0,\infty).

So rir\in i{\mathbb{R}} corresponds to λ<14\lambda<\frac{1}{4} and any such eigenvalue λ(λ0,14)\lambda\in(\lambda_{0},\frac{1}{4}) is called an exceptional eigenvalue. Set

rΔ(N,ν,k):=i14λΔ(Γ0(N),ν,k).r_{\Delta}(N,\nu,k)\vcentcolon=i\cdot\sqrt{\tfrac{1}{4}-\lambda_{\Delta}(\Gamma_{0}(N),\nu,k)}. (2.17)

Let ~k(N,ν,r)~k(N,ν){\tilde{\mathcal{L}}}_{k}(N,\nu,r)\subset{\tilde{\mathcal{L}}}_{k}(N,\nu) denote the subspace corresponding to the spectral parameter rr. Complex conjugation gives an isometry

~k(N,ν,r)~k(N,ν¯,r){\tilde{\mathcal{L}}}_{k}(N,\nu,r)\longleftrightarrow{\tilde{\mathcal{L}}}_{-k}(N,\overline{\nu},r)

between normed spaces. For each v~k(n,ν,r)v\in{\tilde{\mathcal{L}}}_{k}(n,\nu,r), we have the Fourier expansion

v(z)=v(x+iy)=c0(y)+n~0ρ(n)Wk2sgnn~,ir(4π|n~|y)e(n~x)v(z)=v(x+iy)=c_{0}(y)+\sum_{\tilde{n}\neq 0}\rho(n)W_{\frac{k}{2}\operatorname{sgn}\tilde{n},\;ir}(4\pi|\tilde{n}|y)e(\tilde{n}x) (2.18)

where Wκ,μW_{\kappa,\mu} is the Whittaker function and

c0(y)={0αν0,0αν=0 and r0,ρ(0)y12+ir+ρ(0)y12irαν=0 and ri(0,14].c_{0}(y)=\left\{\begin{array}[]{ll}0&\alpha_{\nu}\neq 0,\\ 0&\alpha_{\nu}=0\text{ and }r\geq 0,\\ \rho(0)y^{\frac{1}{2}+ir}+\rho^{\prime}(0)y^{\frac{1}{2}-ir}&\alpha_{\nu}=0\text{ and }r\in i(0,\frac{1}{4}].\end{array}\right.

Using the fact that Wκ,μW_{\kappa,\mu} is a real function when κ\kappa is real and μi\mu\in{\mathbb{R}}\cup i{\mathbb{R}} [DLMF, (13.4.4), (13.14.3), (13.14.31)], if we denote the Fourier coefficient of fc:=f¯f_{c}\vcentcolon=\bar{f} as ρc(n)\rho_{c}(n), then

ρc(n)={ρ(1n)¯,αν>0,n0ρ(n)¯,αν=0.\rho_{c}(n)=\left\{\begin{array}[]{ll}\overline{\rho(1-n)},&\alpha_{\nu}>0,\ n\neq 0\\ \overline{\rho(-n)},&\alpha_{\nu}=0.\end{array}\right. (2.19)

2.3. Holomorphic cusp forms of half-integral weight

For positive integers N,lN,l and a weight k+12k\in{\mathbb{Z}}+\frac{1}{2} multiplier ν\nu on Γ0(N)\Gamma_{0}(N), we know that ν\nu is also a weight k+2lk+2l multiplier system on Γ0(N)\Gamma_{0}(N). For simplicity we denote K=k+2l+12K=k+2l\in{\mathbb{Z}}+\frac{1}{2}. Let SK(N,ν)S_{K}(N,\nu) (resp. MK(N,ν)M_{K}(N,\nu)) be the space of holomorphic cusp forms (resp. modular forms) on Γ0(N)\Gamma_{0}(N) which satisfy the transformation law

F(γz)=ν(γ)(cz+d)KF(z),γΓ0(N).F(\gamma z)=\nu(\gamma)(cz+d)^{K}F(z),\quad\gamma\in\Gamma_{0}(N).

The inner product on SK(N,ν)S_{K}(N,\nu) is defined as

F,GH:=Γ0(N)F(z)G(z)¯yKdxdyy2,f,gSK(N,ν).\langle F,G\rangle_{H}\vcentcolon=\int_{\Gamma_{0}(N)\setminus{\mathbb{H}}}F(z)\overline{G(z)}y^{K}\frac{dxdy}{y^{2}},\quad f,g\in S_{K}(N,\nu).

It is known that (see e.g. [Ran77, §5]) SK(N,ν)S_{K}(N,\nu) is a finite-dimensional Hilbert space under the inner product ,H\langle\cdot,\cdot\rangle_{H}. If we take an orthonormal basis {Fj(): 1jd:=dimSK(N,ν)}\{F_{j}(\cdot):\ 1\leq j\leq d\vcentcolon=\dim S_{K}(N,\nu)\} of SK(N,ν)S_{K}(N,\nu) and write the Fourier expansion of FjF_{j} as

Fj(z)=n=1aj(n)e(n~z),F_{j}(z)=\sum_{n=1}^{\infty}a_{j}(n)e(\tilde{n}z),

then the sum

Γ(K1)(4πn~)K1j=1d|aj(n)|2=1+2πiKN|cS(n,n,c,ν)cJK1(4πn~c)\frac{\Gamma(K-1)}{(4\pi\tilde{n})^{K-1}}\sum_{j=1}^{d}|a_{j}(n)|^{2}=1+2\pi i^{-K}\sum_{N|c}\frac{S(n,n,c,\nu)}{c}J_{K-1}\left(\frac{4\pi\tilde{n}}{c}\right) (2.20)

is independent of the choice of the basis.

There is a one-to-one correspondence between all fk(N,ν)f\in{\mathcal{L}}_{k}(N,\nu) with eigenvalue λ0\lambda_{0} and weight kk holomorphic modular forms FF by

f(z)={yk2F(z)k0,FMk(N,ν),yk2F(z)¯k<0,FMk(N,ν¯).f(z)=\left\{\begin{array}[]{ll}y^{\frac{k}{2}}F(z)&k\geq 0,\;\ F\in M_{k}(N,\nu),\\ y^{-\frac{k}{2}}\overline{F(z)}&k<0,\;\ F\in M_{-k}(N,\overline{\nu}).\end{array}\right. (2.21)

If we write the Fourier expansion of ff as f(z)=nay(n)e(n~x)f(z)=\sum_{n\in{\mathbb{Z}}}a_{y}(n)e(\tilde{n}x), then

{k0ay(n)=0 for n~<0,k<0ay(n)=0 for n~>0.\left\{\begin{array}[]{lll}k\geq 0&\Rightarrow&a_{y}(n)=0\text{\ for\ }\tilde{n}<0,\\ k<0&\Rightarrow&a_{y}(n)=0\text{\ for\ }\tilde{n}>0.\end{array}\right. (2.22)

3. Trace formula

Let k+12k\in{\mathbb{Z}}+\frac{1}{2}, NN be a positive integer, and 𝔞\mathfrak{a} be a singular cusp for the weight kk multiplier system ν\nu on Γ=Γ0(N)\Gamma=\Gamma_{0}(N). Define the Eisenstein series associated to 𝔞\mathfrak{a} by

E𝔞(τ,s):=γΓ𝔞Γν(γ)w(σ𝔞1,γ)¯(Imσ𝔞1γτ)sj(σ𝔞1γ,τ)k.E_{\mathfrak{a}}(\tau,s)\vcentcolon=\sum_{\gamma\in\Gamma_{\mathfrak{a}}\setminus\Gamma}\overline{\nu(\gamma)w(\sigma_{\mathfrak{a}}^{-1},\gamma)}(\operatorname{Im}\sigma_{\mathfrak{a}}^{-1}\gamma\tau)^{s}j(\sigma_{\mathfrak{a}}^{-1}\gamma,\tau)^{-k}. (3.1)

We also define the Poincaré series for m>0m>0 by

𝒰m(τ,s):=γΓΓν¯(γ)(Imγτ)sj(γ,τ)ke(m~γτ).{\mathcal{U}}_{m}(\tau,s)\vcentcolon=\sum_{\gamma\in\Gamma_{\infty}\setminus\Gamma}\overline{\nu}(\gamma)(\operatorname{Im}\gamma\tau)^{s}j(\gamma,\tau)^{-k}e(\tilde{m}\gamma\tau).

Both of them are automorphic functions of weight kk. The properties of these two series can be found in [Pro05]. The Fourier expansion of 𝒰m{\mathcal{U}}_{m} is

𝒰m(x+iy,s)=yse(m~τ)+ysc>0S(m,,c,ν)c2sB(c,m~,~,y,s,k)e(~x){\mathcal{U}}_{m}(x+iy,s)=y^{s}e(\tilde{m}\tau)+y^{s}\sum_{\ell\in{\mathbb{Z}}}\sum_{c>0}\frac{S(m,\ell,c,\nu)}{c^{2s}}B(c,\tilde{m},\tilde{\ell},y,s,k)e(\tilde{\ell}x) (3.2)

where the function BB is in [AA18, (4.5)]. When Res>1\operatorname{Re}s>1, 𝒰m(,s)k(N,ν){\mathcal{U}}_{m}(\cdot,s)\in{\mathcal{L}}_{k}(N,\nu). The Fourier expansion of E𝔞E_{\mathfrak{a}} at s=12+irs=\frac{1}{2}+ir for rr\in{\mathbb{R}} is

E𝔞(x+iy,s)\displaystyle E_{\mathfrak{a}}(x+iy,s) =δ𝔞ys\displaystyle=\delta_{\mathfrak{a}\infty}y^{s} +ρ𝔞(0,r)y1s+0ρ𝔞(,r)Wk2sgn~,12s(4π|~|y)e(~x)\displaystyle+\rho_{\mathfrak{a}}(0,r)y^{1-s}+\sum_{\ell\neq 0}\rho_{\mathfrak{a}}(\ell,r)W_{\frac{k}{2}\operatorname{sgn}\tilde{\ell},\;\frac{1}{2}-s}(4\pi|\tilde{\ell}|y)e(\tilde{\ell}x) (3.3)
=δ𝔞ys\displaystyle=\delta_{\mathfrak{a}\infty}y^{s} +δαν041sΓ(2s1)eπik/2Γ(s+k2)Γ(sk2)y1sφ𝔞0(s)\displaystyle+\frac{\delta_{\alpha_{\nu}0}\cdot 4^{1-s}\Gamma(2s-1)}{e^{\pi ik/2}\Gamma(s+\frac{k}{2})\Gamma(s-\frac{k}{2})}y^{1-s}\varphi_{\mathfrak{a}0}(s)
+0|~|s1πsWk2sgn~,12s(4π|~|y)eπik/2Γ(s+k2sgn~)φ𝔞(s)e(~x).\displaystyle+\sum_{\ell\neq 0}|\tilde{\ell}|^{s-1}\frac{\pi^{s}W_{\frac{k}{2}\operatorname{sgn}\tilde{\ell},\;\frac{1}{2}-s}(4\pi|\tilde{\ell}|y)}{e^{\pi ik/2}\Gamma(s+\frac{k}{2}\operatorname{sgn}\tilde{\ell})}\varphi_{\mathfrak{a}\ell}(s)e(\tilde{\ell}x).

Suppose mm and nn are positive integers and recall the definition of n~\tilde{n} in §2.1. The following notations are very important in the remaining part of this paper:

Setting 3.1.

Let a=4πm~n~>0a=4\pi\sqrt{\tilde{m}\tilde{n}}>0 and 0<Tx/30<T\leq x/3 with Tx1δT\asymp x^{1-\delta} where δ(0,12)\delta\in(0,\frac{1}{2}) will be finally taken as 13\frac{1}{3}.

Setting 3.2.

The test function ϕ:=ϕa,x,T:[0,)\phi\vcentcolon=\phi_{a,x,T}:[0,\infty)\rightarrow{\mathbb{R}} is a four times continuously differentiable function satisfying

  1. (1)

    ϕ(0)=ϕ(0)=0\phi(0)=\phi^{\prime}(0)=0 and ϕ(j)(x)εx2ε\phi^{(j)}(x)\ll_{\varepsilon}x^{-2-\varepsilon} (j=0,,4j=0,\cdots,4) as xx\rightarrow\infty.

  2. (2)

    ϕ(t)=1\phi(t)=1 for a2xtax\frac{a}{2x}\leq t\leq\frac{a}{x}.

  3. (3)

    ϕ(t)=0\phi(t)=0 for ta2x+2Tt\leq\frac{a}{2x+2T} and taxTt\geq\frac{a}{x-T}.

  4. (4)

    ϕ(t)(axTax)1x2aT\phi^{\prime}(t)\ll\left(\frac{a}{x-T}-\frac{a}{x}\right)^{-1}\ll\frac{x^{2}}{aT}.

  5. (5)

    ϕ\phi and ϕ\phi^{\prime} are piecewise monotone on a fixed number of intevals.

This test function was used in [ST09] and [AA18]. We also define the following transformations of ϕ\phi:

ϕ~(r)=0Jr1(u)ϕ(u)duu\widetilde{\phi}(r)=\int_{0}^{\infty}J_{r-1}(u)\phi(u)\frac{du}{u} (3.4)

and for k0k\geq 0,

ϕ^(r):=π2e(1+k)πi20(cos(kπ2+πir)J2ir(u)cos(kπ2πir)J2ir(u))ϕ(u)duush(πr)(ch(2πr)+cosπk)Γ(12k2+ir)Γ(12k2ir)\widehat{\phi}(r)\vcentcolon=\pi^{2}e^{\frac{(1+k)\pi i}{2}}\frac{\int_{0}^{\infty}\left(\cos(\frac{k\pi}{2}+\pi ir)J_{2ir}(u)-\cos(\frac{k\pi}{2}-\pi ir)J_{-2ir}(u)\right)\phi(u)\frac{du}{u}}{\operatorname{sh}(\pi r)(\operatorname{ch}(2\pi r)+\cos\pi k)\Gamma(\frac{1}{2}-\frac{k}{2}+ir)\Gamma(\frac{1}{2}-\frac{k}{2}-ir)} (3.5)

with the corrected version of [Blo08, (2.12)]

ϕ^(i4)={eπi40cos(u)ϕ(u)u32𝑑uk=12,12e3πi40sin(u)ϕ(u)u32𝑑uk=32.\widehat{\phi}\left(\tfrac{i}{4}\right)=\left\{\begin{array}[]{ll}e^{\frac{\pi i}{4}}\int_{0}^{\infty}\cos(u)\phi(u)u^{-\frac{3}{2}}du&\ \ k=\frac{1}{2},\vspace{4px}\\ \frac{1}{2}e^{\frac{3\pi i}{4}}\int_{0}^{\infty}\sin(u)\phi(u)u^{-\frac{3}{2}}du&\ \ k=\frac{3}{2}.\end{array}\right. (3.6)

For an integer l1l\geq 1, let BlB_{l} denote an orthonormal basis for the space of holomorphic cusp forms Sk+2l(N,ν)S_{k+2l}(N,\nu) and

k:=l=1Bl.\mathscr{B}_{k}\vcentcolon=\bigcup_{l=1}^{\infty}B_{l}.

Suppose that the Fourier expansion of each FkF\in\mathscr{B}_{k} is given by

F(z):=n=1aF(n)e(n~z).F(z)\vcentcolon=\sum_{n=1}^{\infty}a_{F}(n)e(\tilde{n}z). (3.7)

Let wFw_{F} denote the weight of FkF\in\mathscr{B}_{k}. Here is the trace formula:

Theorem 3.3 ([Pro05, §6 Theorem]).

Suppose ν\nu is a multiplier system of weight k=12k=\frac{1}{2} or 32\frac{3}{2} on Γ\Gamma. Let {vj()}\{v_{j}(\cdot)\} be an orthonormal basis of ~k(N,ν){\tilde{\mathcal{L}}}_{k}(N,\nu) and E𝔞(,s)E_{\mathfrak{a}}(\cdot,s) be the Eisenstein series associated to a singular cusp 𝔞\mathfrak{a}. Let ρj(n)\rho_{j}(n) denote the nn-th Fourier coefficient of vjv_{j}. Let φ𝔞n(12+ir)\varphi_{\mathfrak{a}n}(\frac{1}{2}+ir) or ρ𝔞(n,r)\rho_{\mathfrak{a}}(n,r) denote the nn-th Fourier coefficient of E𝔞(,12+ir)E_{\mathfrak{a}}(\cdot,\frac{1}{2}+ir) as in (3.3). Let k\mathscr{B}_{k} and aF(n)a_{F}(n) be defined as in (3.7). Then for m~>0\tilde{m}>0 and n~>0\tilde{n}>0 we have

c>0S(m,n,c,ν)cϕ(4πm~n~c)=𝒰k+𝒲+singular𝔞𝔞,\displaystyle\sum_{c>0}\frac{S(m,n,c,\nu)}{c}\phi\left(\frac{4\pi\sqrt{\tilde{m}\tilde{n}}}{c}\right)=\mathcal{U}_{k}+\mathcal{W}+\!\!\!\sum_{\mathrm{singular\ }\mathfrak{a}}\!\!\!\mathcal{E}_{\mathfrak{a}}, (3.8)

where

𝒰k=Fk4Γ(wF)eπiwF/2(4π)wF(m~n~)(wF1)/2aF(m)¯aF(n)ϕ~(wF),\mathcal{U}_{k}=\sum_{F\in\mathscr{B}_{k}}\frac{4\Gamma(w_{F})e^{\pi iw_{F}/2}}{(4\pi)^{w_{F}}(\tilde{m}\tilde{n})^{(w_{F}-1)/2}}\overline{a_{F}(m)}a_{F}(n)\widetilde{\phi}(w_{F}),
𝒲=4m~n~rjρj(m)¯ρj(n)chπrjϕ^(rj),\mathcal{W}=4\sqrt{\tilde{m}\tilde{n}}\sum_{r_{j}}\frac{\overline{\rho_{j}(m)}\rho_{j}(n)}{\operatorname{ch}\pi r_{j}}\widehat{\phi}(r_{j}),

and

𝔞\displaystyle\mathcal{E}_{\mathfrak{a}} =(m~n~)irφ𝔞m(12+ir)¯φ𝔞n(12+ir)ϕ^(r)drchπr|Γ(12+k2+ir)|2\displaystyle=\int_{-\infty}^{\infty}\left(\frac{\tilde{m}}{\tilde{n}}\right)^{-ir}\frac{\overline{\varphi_{\mathfrak{a}m}\left(\frac{1}{2}+ir\right)}\varphi_{\mathfrak{a}n}\left(\frac{1}{2}+ir\right)\widehat{\phi}(r)dr}{\operatorname{ch}\pi r\;|\Gamma(\frac{1}{2}+\frac{k}{2}+ir)|^{2}}
=4m~n~14πρ𝔞(m,r)¯ρ𝔞(n,r)ϕ^(r)chπr𝑑r.\displaystyle=4\sqrt{\tilde{m}\tilde{n}}\cdot\frac{1}{4\pi}\int_{-\infty}^{\infty}\overline{\rho_{\mathfrak{a}}\left(m,r\right)}\rho_{\mathfrak{a}}\left(n,r\right)\frac{\widehat{\phi}(r)}{\operatorname{ch}\pi r}dr.
Remark.

We clarify two points in the theorem.

  • (1)

    In the term 𝒰k\mathcal{U}_{k} corresponding to holomorphic cusp forms, each function FkF\in\mathscr{B}_{k} has weight wF=k+2l52w_{F}=k+2l\geq\frac{5}{2}.

  • (2)

    The equality of the two expressions in 𝔞\mathcal{E}_{\mathfrak{a}} is by (3.3):

    n~πρ𝔞(n,r)=eπik2πirn~irΓ(12+ir+k2sgnn~)φ𝔞n(12+ir).\sqrt{\frac{\tilde{n}}{\pi}}\,\rho_{\mathfrak{a}}(n,r)=\frac{e^{-\frac{\pi ik}{2}}\pi^{ir}\tilde{n}^{ir}}{\Gamma\left(\frac{1}{2}+ir+\frac{k}{2}\operatorname{sgn}\tilde{n}\right)}\varphi_{\mathfrak{a}n}\left(\frac{1}{2}+ir\right).

3.1. Properties of admissible multipliers

Suppose ν\nu is a weight kk admissible multiplier system on Γ=Γ0(N)\Gamma=\Gamma_{0}(N) (Definition 1.1) with parameters BB, MM and DD. Then we have the following propositions.

Proposition 3.4.

[Sun23, Proposition 5.7] Suppose that ν\nu satisfies condition (1) of Definition 1.1 and assume HθH_{\theta} (2.16). Then we have 2ImrΔ(N,ν,k)θ.2\operatorname{Im}r_{\Delta}\left(N,\nu,k\right)\leq\theta.

Proposition 3.5.

Suppose that ν\nu satisfies condition (1) of Definition 1.1 with ν=(|D|)νθ2k\nu^{\prime}=(\frac{|D|}{\cdot})\nu_{\theta}^{2k}. For ll\in{\mathbb{Z}}, let K=k+2l52K=k+2l\geq\frac{5}{2}. Suppose {Fj,l()}j\{F_{j,l}(\cdot)\}_{j} is an orthonormal basis of SK(N,ν)S_{K}(N,\nu) and {Gj,l()}j\{G_{j,l}(\cdot)\}_{j} is an orthonormal basis of SK(M,ν)S_{K}(M,\nu^{\prime}). Denote aF,j,l(n)a_{F,j,l}(n) as the Fourier coefficient of Fj,lF_{j,l} and aG,j,l(n)a_{G,j,l}(n) as the Fourier coefficient of Gj,lG_{j,l}. Then we have

j=1dimSK(N,ν)|aF,j,l(n)|2N,νj=1dimSK(M,ν)|aG,j,l(Bn~)|2.\sum_{j=1}^{\dim S_{K}(N,\nu)}|a_{F,j,l}(n)|^{2}\ll_{N,\nu}\sum_{j=1}^{\dim S_{K}(M,\nu^{\prime})}|a_{G,j,l}(B\tilde{n})|^{2}. (3.9)
Proof.

By condition (1) of Definition 1.1, we know that

{[Γ0(N):Γ0(M)]12Fj,l(Bz): 1jdimSK(N,ν)}\left\{[\Gamma_{0}(N):\Gamma_{0}(M)]^{-\frac{1}{2}}F_{j,l}(Bz):\ 1\leq j\leq\dim S_{K}(N,\nu)\right\} (3.10)

is an orthonormal subset of SK(M,ν)S_{K}(M,\nu^{\prime}). Since the left hand side of (2.20) is independent from the choice of basis, we expand (3.10) to an orthonormal basis of SK(M,ν)S_{K}(M,\nu^{\prime}) and get the result. ∎

Now we start to prove a bound for the right hand side of (3.9). First we have

Proposition 3.6 ([Wai17, Theorem 1]).

For K+12K\in{\mathbb{Z}}+\frac{1}{2}, K52K\geq\frac{5}{2} and a quadratic character χ\chi modulo MM, suppose

{Φj=n=1aj(n)e(nz): 1jd:=dimSK(M,χνθ2K)}\left\{\Phi_{j}=\sum_{n=1}^{\infty}a_{j}(n)e(nz):\ 1\leq j\leq d\vcentcolon=\dim S_{K}(M,\chi\nu_{\theta}^{2K})\right\}

is an orthonormal basis of SK(M,χνθ2K)S_{K}(M,\chi\nu_{\theta}^{2K}). For n1n\geq 1, write n=tu2w2n=tu^{2}w^{2} with tt square-free, u|Mu|M^{\infty} and (w,M)=1(w,M)=1. Then we have

Γ(K1)(4πn)K1j=1d|aj(n)|2K,M,ε(n37+u)nε.\frac{\Gamma(K-1)}{(4\pi n)^{K-1}}\sum_{j=1}^{d}|a_{j}(n)|^{2}\ll_{K,M,\varepsilon}\left(n^{\frac{3}{7}}+u\right)n^{\varepsilon}.

Note that the implied constant in Waibel’s bound depends on KK when expressing Bessel functions (see [Wai17, after (8)] and [Iwa87, Theorem 1 and p. 400]). For our proof, it’s essential that the bound remains uniform across the weights. We modify the estimate and get the following proposition.

Proposition 3.7.

With the same setting as Proposition 3.6,

Γ(K1)(4πn)K1j=1d|aj(n)|2M,ε(n1942+u)nε.\frac{\Gamma(K-1)}{(4\pi n)^{K-1}}\sum_{j=1}^{d}|a_{j}(n)|^{2}\ll_{M,\varepsilon}(n^{\frac{19}{42}}+u)n^{\varepsilon}.
Proof.

We do the same preparation as [Wai17, around (8)] to estimate the right hand side of (2.20). Let P>1+(log2nM)2P>1+(\log 2nM)^{2} (finally chosen to be Mn17\asymp_{M}n^{\frac{1}{7}}) and define the set of prime numbers

𝒫:={p prime:P<p2P,p2nM}.\mathcal{P}\vcentcolon=\ \{p\text{ prime}:P<p\leq 2P,\ p\nmid 2nM\}.

Here we have #𝒫P/logP\#\mathcal{P}\asymp P/\log P.

For {Φj}j\{\Phi_{j}\}_{j} a orthonormal basis of SK(M,χνθ2K)S_{K}(M,\chi\nu_{\theta}^{2K}), the set {[Γ0(M):Γ0(pM)]12Φj}j\{[\Gamma_{0}(M):\Gamma_{0}(pM)]^{-\frac{1}{2}}\Phi_{j}\}_{j} is an orthonormal subset of SK(pM,χνθ2K)S_{K}(pM,\chi\nu_{\theta}^{2K}). Recall (2.20) and we have

Γ(K1)(4πn)K1j=1d|aj(n)|2[Γ0(M):Γ0(pM)]1+2πiKpM|cS(n,n,c,χνθ2K)cJK1(4πnc).\frac{\Gamma(K-1)}{(4\pi n)^{K-1}}\sum_{j=1}^{d}\frac{|a_{j}(n)|^{2}}{[\Gamma_{0}(M):\Gamma_{0}(pM)]}\leq 1+2\pi i^{-K}\sum_{pM|c}\frac{S(n,n,c,\chi\nu_{\theta}^{2K})}{c}J_{K-1}\left(\frac{4\pi n}{c}\right). (3.11)

For those p𝒫p\in\mathcal{P}, [Γ0(M):Γ0(pM)]p+1[\Gamma_{0}(M):\Gamma_{0}(pM)]\leq p+1. Summing (3.11) on p𝒫p\in\mathcal{P} and dividing #𝒫\#\mathcal{P} we get

Γ(K1)(4πn)K1j=1d|aj(n)|2P+(logP)p𝒫|pM|cS(n,n,c,χνθ2K)cJK1(4πnc)|.\frac{\Gamma(K-1)}{(4\pi n)^{K-1}}\sum_{j=1}^{d}|a_{j}(n)|^{2}\ll P+(\log P)\sum_{p\in\mathcal{P}}\left|\sum_{pM|c}\frac{S(n,n,c,\chi\nu_{\theta}^{2K})}{c}J_{K-1}\left(\frac{4\pi n}{c}\right)\right|. (3.12)

The average estimate of

KpM(μ)(x):=pM|cxS(n,n,c,χνθ2K)ce(2μnc),μ{1,0,1}K_{pM}^{(\mu)}(x)\vcentcolon=\sum_{pM|c\leq x}\frac{S(n,n,c,\chi\nu_{\theta}^{2K})}{\sqrt{c}}e\left(\frac{2\mu n}{c}\right),\quad\mu\in\{-1,0,1\}

can be found in [Wai17, (19)] that for μ{1,0,1}\mu\in\{-1,0,1\},

p𝒫|KpM(μ)(x)|M,ε(xun12+xP12+(x+n)58(x14P38+n18x18P14))(nx)ε.\sum_{p\in\mathcal{P}}|K_{pM}^{(\mu)}(x)|\ll_{M,\varepsilon}\left(xun^{-\frac{1}{2}}+xP^{-\frac{1}{2}}+(x+n)^{\frac{5}{8}}\left(x^{\frac{1}{4}}P^{\frac{3}{8}}+n^{\frac{1}{8}}x^{\frac{1}{8}}P^{\frac{1}{4}}\right)\right)(nx)^{\varepsilon}. (3.13)

We break the sum on c0(modpM)c\equiv 0\ (\mathrm{mod}\ pM) at the right hand side of (3.12) into cnc\leq n and cnc\geq n to estimate. The uniform bound of JJ-Bessel functions is given by [Lan00]

|Jβ(x)|c0x13for all β>0 and x>0,|J_{\beta}(x)|\leq c_{0}x^{-\frac{1}{3}}\quad\text{for all }\beta>0\text{ and }x>0, (3.14)

where c0=0.7857c_{0}=0.7857\cdots.

When cnc\leq n, using (3.14) and [DLMF, (10.6.1)]

2Jβ1(x)=Jβ2(x)Jβ(x),2J_{\beta-1}^{\prime}(x)=J_{\beta-2}(x)-J_{\beta}(x),

we find that

(x12JK1(4πnx))n13x76+n23x136.\left(x^{-\frac{1}{2}}J_{K-1}\left(\frac{4\pi n}{x}\right)\right)^{\prime}\ll n^{-\frac{1}{3}}x^{-\frac{7}{6}}+n^{\frac{2}{3}}x^{-\frac{13}{6}}. (3.15)

Then a partial summation using (3.13), (3.15) and (3.14) yields

p𝒫|pM|cnS(n,n,c,χνθ2K)cJK1(4πnc)|M,ε(n1942+u)nε.\sum_{p\in\mathcal{P}}\left|\sum_{pM|c\leq n}\frac{S(n,n,c,\chi\nu_{\theta}^{2K})}{c}J_{K-1}\left(\frac{4\pi n}{c}\right)\right|\ll_{M,\varepsilon}(n^{\frac{19}{42}}+u)n^{\varepsilon}. (3.16)

When cnc\geq n, we get another bound

(x12JK1(4πnx))nx52for xn\left(x^{-\frac{1}{2}}J_{K-1}\left(\frac{4\pi n}{x}\right)\right)^{\prime}\ll nx^{-\frac{5}{2}}\quad\text{for }x\geq n (3.17)

by |Jβ1(x)|(x/2)β1Γ(β)|J_{\beta-1}(x)|\leq\frac{(x/2)^{\beta-1}}{\Gamma(\beta)} [DLMF, (10.14.4)] and |Jβ(x)|1|J_{\beta}(x)|\leq 1 [DLMF, (10.14.1)]. Remember K52K\geq\frac{5}{2} here. We do a partial summation again using (3.13) and (3.17) and get

p𝒫|pM|cnS(n,n,c,χνθ2K)cJK1(4πnc)|M,ε(n37+u)nε.\sum_{p\in\mathcal{P}}\left|\sum_{pM|c\geq n}\frac{S(n,n,c,\chi\nu_{\theta}^{2K})}{c}J_{K-1}\left(\frac{4\pi n}{c}\right)\right|\ll_{M,\varepsilon}(n^{\frac{3}{7}}+u)n^{\varepsilon}. (3.18)

From (3.16), (3.18), (3.12) and PMn17P\asymp_{M}n^{\frac{1}{7}}, we finish the proof. ∎

Combining Proposition 3.5 and Proposition 3.7, one observes the following bound:

Proposition 3.8.

With the same setting as Proposition 3.5, we factor Bn~=tnun2wn2B\tilde{n}=t_{n}u_{n}^{2}w_{n}^{2} with tnt_{n} square-free, un|Mu_{n}|M^{\infty} and (wn,M)=1(w_{n},M)=1. Then

Γ(K1)(4πn~)K1j=1dimSK(N,ν)|aF,j,l(n)|2N,ν,ε(n~1942+un)n~ε.\frac{\Gamma(K-1)}{(4\pi\tilde{n})^{K-1}}\sum_{j=1}^{\dim S_{K}(N,\nu)}|a_{F,j,l}(n)|^{2}\ll_{N,\nu,\varepsilon}(\tilde{n}^{\frac{19}{42}}+u_{n})\tilde{n}^{\varepsilon}.

4. Bounds on ϕ~\widetilde{\phi} and ϕ^\widehat{\phi}

In this section, all of the implied constants among the estimates for ϕ~\widetilde{\phi} and ϕ^\widehat{\phi} depend on NN and the multiplier system ν\nu unless specified. Recall the definitions (3.4) and (3.5). To deal with the Γ\Gamma-function in the denominator of ϕ^\widehat{\phi}, we need [DLMF, (5.6.6-7)]

Γ(x)2ch(πr)|Γ(x+ir)|2Γ(x)2for x0 and r.\frac{\Gamma(x)^{2}}{\operatorname{ch}(\pi r)}\leq|\Gamma(x+ir)|^{2}\leq\Gamma(x)^{2}\quad\text{for }x\geq 0\text{ and }r\in{\mathbb{R}}. (4.1)

Recall (3.4) and (3.5) that we define ϕ^\widehat{\phi} for k0k\geq 0. We also have

ϕ^(r)=π2e1+k2πish(πr)(ch(2πr)+cos(πk))Γ(12k2+ir)Γ(12k2ir){coskπ2ch(πr)(ϕ~(1+2ir)ϕ~(12ir))isinkπ2sh(πr)(ϕ~(1+2ir)+ϕ~(12ir))}.\displaystyle\begin{split}&\widehat{\phi}(r)=\frac{\pi^{2}e^{\frac{1+k}{2}\pi i}}{\operatorname{sh}(\pi r)(\operatorname{ch}(2\pi r)+\cos(\pi k))\Gamma(\frac{1}{2}-\frac{k}{2}+ir)\Gamma(\frac{1}{2}-\frac{k}{2}-ir)}\\ &\cdot\left\{\cos\tfrac{k\pi}{2}\operatorname{ch}(\pi r)\left(\widetilde{\phi}(1+2ir)-\widetilde{\phi}(1-2ir)\right)-i\sin\tfrac{k\pi}{2}\operatorname{sh}(\pi r)\left(\widetilde{\phi}(1+2ir)+\widetilde{\phi}(1-2ir)\right)\right\}.\end{split} (4.2)

Like [AD20, after (5.3)], we define ξk\xi_{k} as

ξk(r):=2iπ2e1+k2πiΓ(12k2+ir)Γ(12k2ir).\xi_{k}(r)\vcentcolon=\frac{2i\pi^{2}e^{\frac{1+k}{2}\pi i}}{\Gamma(\frac{1}{2}-\frac{k}{2}+ir)\Gamma(\frac{1}{2}-\frac{k}{2}-ir)}. (4.3)

Then

ξk(r){1for r[1,1],|r|keπ|r|for r(,1][1,).\xi_{k}(r)\left\{\begin{array}[]{ll}\ll 1&\quad\text{for }r\in[-1,1],\\ \asymp|r|^{k}e^{\pi|r|}&\quad\text{for }r\in(-\infty,-1]\cup[1,\infty).\end{array}\right. (4.4)

We refer to [Dun90] for estimates on JJ-Bessel functions. Denote

Fμ(z):=Jμ(z)+Jμ(z)2cos(μπ/2),Gμ(z):=Jμ(z)Jμ(z)2sin(μπ/2).F_{\mu}(z)\vcentcolon=\frac{J_{\mu}(z)+J_{-\mu}(z)}{2\cos(\mu\pi/2)},\quad G_{\mu}(z)\vcentcolon=\frac{J_{\mu}(z)-J_{-\mu}(z)}{2\sin(\mu\pi/2)}.

As a result of the relationship J2ir(u)¯=J2ir(u)\overline{J_{2ir}(u)}=J_{-2ir}(u) for r,ur,u\in{\mathbb{R}} by [DLMF, (10.11.9)], we have

F2ir(u)=ReJ2ir(u)ch(πr),G2ir(u)=ImJ2ir(u)sh(πr).F_{2ir}(u)=\frac{\operatorname{Re}J_{2ir}(u)}{\operatorname{ch}(\pi r)}\in{\mathbb{R}},\quad G_{2ir}(u)=\frac{\operatorname{Im}J_{2ir}(u)}{\operatorname{sh}(\pi r)}\in{\mathbb{R}}.

Moreover, for k+12k\in{\mathbb{Z}}+\frac{1}{2} and k0k\geq 0,

ϕ^(r)=ξk(r)ch(πr)ch(2πr)0(G2ir(u)coskπ2F2ir(u)sinkπ2)ϕ(u)u𝑑u\widehat{\phi}(r)=\frac{\xi_{k}(r)\operatorname{ch}(\pi r)}{\operatorname{ch}(2\pi r)}\int_{0}^{\infty}\left(G_{2ir}(u)\cos\frac{k\pi}{2}-F_{2ir}(u)\sin\frac{k\pi}{2}\right)\frac{\phi(u)}{u}du (4.5)

and ϕ^(r)=ϕ^(r)\widehat{\phi}(r)=\widehat{\phi}(-r) for rr\in{\mathbb{R}} because Fμ(z)=Fμ(z)F_{\mu}(z)=F_{-\mu}(z) and Gμ(z)=Gμ(z)G_{\mu}(z)=G_{-\mu}(z).

Lemma 4.1.

For r[1,1]r\in[-1,1], uniformly and with absolute implied constants we have

G2ir(u){ln(u2),u[0,32],u32,u[32,).\displaystyle\begin{split}G_{2ir}(u)\ll\left\{\begin{array}[]{ll}\ln(\tfrac{u}{2}),&u\in[0,\frac{3}{2}],\vspace{4px}\\ u^{-\frac{3}{2}},&u\in[\frac{3}{2},\infty).\end{array}\right.\end{split} (4.6)
Proof.

First we deal with the range u[0,32]u\in[0,\frac{3}{2}]. The series expansion of GirG_{ir} is given by [Dun90, (3.9), (3.16)]:

G2ir(u)\displaystyle G_{2ir}(u) =(4rch(πr)πsh(πr))12=0(1)(u2/4)sin(2rln(u/2)ϕ2r,)!j=02(j+4r2)1/2\displaystyle=\left(\frac{4r\operatorname{ch}(\pi r)}{\pi\operatorname{sh}(\pi r)}\right)^{\frac{1}{2}}\sum_{\ell=0}^{\infty}\frac{(-1)^{\ell}(u^{2}/4)^{\ell}\sin(2r\ln(u/2)-\phi_{2r,\ell})}{\ell!\prod_{j=0}^{\ell^{2}}(j+4r^{2})^{1/2}}
=(ch(πr)πrsh(πr))12sin(2rln(u2)ϕ2r,0)+O((u2)2).\displaystyle=\left(\frac{\operatorname{ch}(\pi r)}{\pi r\operatorname{sh}(\pi r)}\right)^{\frac{1}{2}}\sin\left(2r\ln\left(\frac{u}{2}\right)-\phi_{2r,0}\right)+O\left(\left(\frac{u}{2}\right)^{2}\right).

where ϕr,=argΓ(1++ir)\phi_{r,\ell}=\arg\Gamma(1+\ell+ir). The implied constant in the second equation is absolute. As a function of rr, ϕ2r,0C[0,1]\phi_{2r,0}\in C^{\infty}[0,1] and limr0ϕ2r,0=0\lim_{r\rightarrow 0}\phi_{2r,0}=0. Then ϕ2r,0/r=O(1)\phi_{2r,0}/r=O(1) and

G2ir(u)\displaystyle G_{2ir}(u) r1O(|2rln(u2)|+|ϕ2r,0|)+O((u2)2)\displaystyle\ll r^{-1}O\left(\left|2r\ln\left(\frac{u}{2}\right)\right|+|\phi_{2r,0}|\right)+O\left(\left(\frac{u}{2}\right)^{2}\right)
ln(u2)+O(1).\displaystyle\ll\ln\left(\frac{u}{2}\right)+O(1).

For the range u32u\geq\frac{3}{2}, we check with [Dun90, (5.16)] where Us(p)U_{s}(p) for s0s\geq 0 are fixed polynomials of pp whose lowest degree term is psp^{s}:

G2ir(u)\displaystyle G_{2ir}(u) =(4/π24r2+u2)14(C4r2+u2+O(14r2+u2))\displaystyle=\left(\frac{4/\pi^{2}}{4r^{2}+u^{2}}\right)^{\frac{1}{4}}\left(\frac{C}{\sqrt{4r^{2}+u^{2}}}+O\left(\frac{1}{4r^{2}+u^{2}}\right)\right)
(r2+u24)34+O((r2+u24)54).\displaystyle\ll\left(r^{2}+\frac{u^{2}}{4}\right)^{-\frac{3}{4}}+O\left(\left(r^{2}+\frac{u^{2}}{4}\right)^{-\frac{5}{4}}\right).

Our claimed bound is clear as r20r^{2}\geq 0. The implied constant above is absolute due to [Dun90, (3.3)] and [Olv97, Chapter 8, §13] or by [Olv97, Chapter 10, (3.04)].

Lemma 4.2.

For r[1,1]r\in[-1,1], we have

|ϕ~(1+2ir)|1,|ϕ^(r)|ε(ax)ε.|\widetilde{\phi}(1+2ir)|\ll 1,\quad|\widehat{\phi}(r)|\ll_{\varepsilon}(ax)^{\varepsilon}. (4.7)
Proof.

A trivial bound of J2irJ_{2ir} is given by the integral representation [DLMF, (10.9.4)]:

Jν(z)=(z/2)νπΓ(ν+12)0πcos(zcosθ)(sinθ)2ν𝑑θ,Reν>12.J_{\nu}(z)=\frac{(z/2)^{\nu}}{\sqrt{\pi}\Gamma(\nu+\frac{1}{2})}\int_{0}^{\pi}\cos(z\cos\theta)(\sin\theta)^{2\nu}d\theta,\quad\operatorname{Re}\nu>-\frac{1}{2}.

Then we have |J2ir(u)|π|Γ(12+2ir)||J_{2ir}(u)|\leq\frac{\sqrt{\pi}}{|\Gamma(\frac{1}{2}+2ir)|} and

|ϕ~(1+2ir)|3a8x3a2xduuln4for r[1,1].|\widetilde{\phi}(1+2ir)|\ll\int_{\frac{3a}{8x}}^{\frac{3a}{2x}}\frac{du}{u}\leq\ln 4\quad\text{for }r\in[-1,1].

This implies that

ch(πr)0F2ir(u)ϕ(u)u𝑑u1for r[1,1].\operatorname{ch}(\pi r)\int_{0}^{\infty}F_{2ir}(u)\frac{\phi(u)}{u}du\ll 1\quad\text{for }r\in[-1,1].

Let the closed interval [α,β]=[\alpha,\beta]=\emptyset when α>β\alpha>\beta. With the help of (4.5), (4.4) and Lemma 4.1 we get

|ϕ^(r)|\displaystyle|\widehat{\phi}(r)| ch(πr)ch(2πr)(|0G2ir(u)ϕ(u)duu|+|0F2ir(u)ϕ(u)duu|)\displaystyle\ll\frac{\operatorname{ch}(\pi r)}{\operatorname{ch}(2\pi r)}\left(\left|\int_{0}^{\infty}G_{2ir}(u)\phi(u)\frac{du}{u}\right|+\left|\int_{0}^{\infty}F_{2ir}(u)\phi(u)\frac{du}{u}\right|\right)
[3a8x,32]ln(u2)duu+[32,3a2x]u52𝑑u+O(1)\displaystyle\ll\int_{[\frac{3a}{8x},\frac{3}{2}]}\ln\left(\frac{u}{2}\right)\frac{du}{u}+\int_{[\frac{3}{2},\frac{3a}{2x}]}u^{-\frac{5}{2}}du+O(1)
(ln3a16x)2+O(1)(ax)ε.\displaystyle\ll\left(\ln\frac{3a}{16x}\right)^{2}+O(1)\ll(ax)^{\varepsilon}.

The last inequality is because a=4πm~n~>0a=4\pi\sqrt{\tilde{m}\tilde{n}}>0 has a lower bound depending on ν\nu. ∎

When we focus on the exceptional eigenvalues λj[316,14)\lambda_{j}\in[\frac{3}{16},\frac{1}{4}) of Δk\Delta_{k}, recall that λj=14+rj2\lambda_{j}=\frac{1}{4}+r_{j}^{2} for rji(0,14]r_{j}\in i(0,\frac{1}{4}]. By Proposition 3.4, if we write tj=Imrjt_{j}=\operatorname{Im}r_{j}, assuming HθH_{\theta} (2.16) we have an upper bound tjθ2t_{j}\leq\frac{\theta}{2} when rji4r_{j}\neq\frac{i}{4}. Moreover, since the exceptional eigenvalues are discrete, we also have a largest eigenvalue less than 14\frac{1}{4}, hence a lower bound t¯>0\underline{t}>0 (depending on NN and ν\nu) such that tjt¯t_{j}\geq\underline{t}.

Lemma 4.3.

With the hypothesis HθH_{\theta} (2.16) for θ16\theta\leq\frac{1}{6}, when r=itr=it and t[t¯,θ2]t\in[\underline{t},\frac{\theta}{2}], we have

ϕ~(1±2t)(ax)±2tandϕ^(r)(ax)2t+(xa)2t(ax)θ+(xa)θ.\widetilde{\phi}(1\pm 2t)\ll\left(\frac{a}{x}\right)^{\pm 2t}\quad\text{and}\quad\widehat{\phi}(r)\ll\left(\frac{a}{x}\right)^{2t}+\left(\frac{x}{a}\right)^{2t}\ll\left(\frac{a}{x}\right)^{\theta}+\left(\frac{x}{a}\right)^{\theta}. (4.8)

Moreover, for r=i4r=\frac{i}{4} we have

ϕ~(1±12)(ax)±12andϕ^(i4){(xa)12,k=12,(ax)12,k=32.\widetilde{\phi}\left(1\pm\frac{1}{2}\right)\ll\left(\frac{a}{x}\right)^{\pm\frac{1}{2}}\quad\text{and}\quad\widehat{\phi}\left(\frac{i}{4}\right)\ll\left\{\begin{array}[]{ll}(\frac{x}{a})^{\frac{1}{2}},&k=\frac{1}{2},\vspace{4px}\\ (\frac{a}{x})^{\frac{1}{2}},&k=\frac{3}{2}.\end{array}\right. (4.9)
Proof.

As in the previous lemma, when t[t¯,θ2]t\in[\underline{t},\frac{\theta}{2}], the bound [DLMF, (10.9.4)] gives

|J±2t(u)|u±2tΓ(12θ)and|ϕ~(1±2t)|3a8x3a2xu±2tduu(ax)±2t.|J_{\pm 2t}(u)|\ll\frac{u^{\pm 2t}}{\Gamma(\frac{1}{2}-\theta)}\quad\text{and}\quad|\widetilde{\phi}(1\pm 2t)|\ll\int_{\frac{3a}{8x}}^{\frac{3a}{2x}}u^{\pm 2t}\;\frac{du}{u}\ll\left(\frac{a}{x}\right)^{\pm 2t}.

The bound for ϕ^\widehat{\phi} follows from (4.2). When r=i4r=\frac{i}{4}, by [DLMF, (10.16.1)] we have

J12(u)u12andJ12(u)u12sinuu12.J_{-\frac{1}{2}}(u)\ll u^{-\frac{1}{2}}\quad\text{and}\quad J_{\frac{1}{2}}(u)\ll u^{-\frac{1}{2}}\sin u\leq u^{\frac{1}{2}}.

The bounds for ϕ~(1±12)\widetilde{\phi}(1\pm\frac{1}{2}) and ϕ^(i4)\widehat{\phi}(\frac{i}{4}) follow from the same process above with (3.4) and (3.6). ∎

For the range |r|1|r|\geq 1 we have

Lemma 4.4.

[AD20, Lemma 6.3] Let k=12k=\frac{1}{2} or 32\frac{3}{2}. Then

ϕ^(r){rk32,r1,rkmin(r32,r52xT),rmax(ax,1).\widehat{\phi}(r)\ll\left\{\begin{array}[]{ll}r^{k-\frac{3}{2}},&\ r\geq 1,\\ r^{k}\min(r^{-\frac{3}{2}},r^{-\frac{5}{2}}\frac{x}{T}),&\ r\geq\max(\frac{a}{x},1).\end{array}\right. (4.10)
Remark.

In the original paper they stated the result for k=±12k=\pm\frac{1}{2}. However, the power rkr^{k} in the estimate above only arises from ξk(r)eπ|r|\xi_{k}(r)e^{-\pi|r|} (4.3) and by (4.4) we get the above lemma on weight k=32k=\frac{3}{2}.

4.1. A special test function

Here we choose a special test function ϕ\phi satisfying Setting 3.2 to compute the terms corresponding to the exceptional spectrum ri(0,14]r\in i(0,\frac{1}{4}] in Theorem 1.3.

For k=12k=\frac{1}{2} or 32\frac{3}{2}, let λ[316,14)\lambda\in[\frac{3}{16},\frac{1}{4}) be an exceptional eigenvalue of Δk\Delta_{k} on Γ0(N)\Gamma_{0}(N), we set λ=s(1s)\lambda=s(1-s) for s(12,34]s\in(\frac{1}{2},\frac{3}{4}] and

t=Imr=14λ=14s(1s)=s12.t=\operatorname{Im}r=\sqrt{\tfrac{1}{4}-\lambda}=\sqrt{\tfrac{1}{4}-s(1-s)}=s-\tfrac{1}{2}.

Since the exceptional spectrum is discrete, let the lower bound for t>0t>0 be t¯\underline{t} depending on NN and ν\nu. Recall Setting 3.1. Let 0<TTx30<T^{\prime}\leq T\leq\frac{x}{3} be T:=Txδx12δT^{\prime}\vcentcolon=Tx^{-\delta}\asymp x^{1-2\delta}.

Setting 4.5.

In addition to the requirement in Setting 3.2, when axT1.999\frac{a}{x-T}\leq 1.999, we pick ϕ\phi as a smoothed function of this piecewise linear one

xxyya2x+2T\frac{a}{2x+2T}a2x\frac{a}{2x}ax\frac{a}{x}axT\frac{a}{x-T}11

where

{ϕ(u)=2x(x+T)aTu(a2x+2T2T,a2x+2T),ϕ(u)=x(xT)aTu(axT,axT+T),0ϕ(u)4x(x+T)aTu(a2x+2T,a2x+2T2T)(a2x+2T,a2x),0ϕ(u)2x(xT)aTu(ax,axT)(axT+T,axT),ϕ(u)=0otherwise.\left\{\begin{array}[]{ll}\phi^{\prime}(u)=\tfrac{2x(x+T)}{aT}&u\in(\tfrac{a}{2x+2T-2T^{\prime}},\;\tfrac{a}{2x+2T^{\prime}}),\\ \\ \phi^{\prime}(u)=-\tfrac{x(x-T)}{aT}&u\in(\tfrac{a}{x-T^{\prime}},\;\tfrac{a}{x-T+T^{\prime}}),\\ \\ 0\leq\phi^{\prime}(u)\leq\tfrac{4x(x+T)}{aT}&u\in(\tfrac{a}{2x+2T},\;\tfrac{a}{2x+2T-2T^{\prime}})\cup(\tfrac{a}{2x+2T^{\prime}},\;\tfrac{a}{2x}),\\ \\ 0\geq\phi^{\prime}(u)\geq-\tfrac{2x(x-T)}{aT}&u\in(\tfrac{a}{x},\;\tfrac{a}{x-T^{\prime}})\cup(\tfrac{a}{x-T+T^{\prime}},\;\tfrac{a}{x-T}),\vspace{10px}\\ \phi^{\prime}(u)=0&\text{otherwise}.\tfrac{}{}\end{array}\right. (4.11)

The above choice for ϕ\phi^{\prime} is possible because there is no requirement for ϕ′′(u)\phi^{\prime\prime}(u) when u2u\leq 2 but for uu\rightarrow\infty in Setting 3.2.

Now we take r=iti(0,14]r=it\in i(0,\frac{1}{4}]. When u1.999u\leq 1.999, by the series expansion [DLMF, (10.2.2)]:

Jν(z)=(z2)νj=0(1)jj!Γ(j+1+ν)(z2)2j,J_{\nu}(z)=\left(\frac{z}{2}\right)^{\nu}\sum_{j=0}^{\infty}\frac{(-1)^{j}}{j!\Gamma(j+1+\nu)}\left(\frac{z}{2}\right)^{2j},

we have

J±2t(u)=(u/2)±2tΓ(1±2t)+O((u2)2±2t),0<u1.999.J_{\pm 2t}(u)=\frac{(u/2)^{\pm 2t}}{\Gamma(1\pm 2t)}+O\left(\left(\frac{u}{2}\right)^{2\pm 2t}\right),\qquad 0<u\leq 1.999. (4.12)

The implied constant is absolute. Now we compute the bound for ϕ~\widetilde{\phi} and ϕ^\widehat{\phi}.

Lemma 4.6.

Assuming HθH_{\theta} (2.16) for θ16\theta\leq\frac{1}{6} and with the choice of ϕ\phi in Setting 4.5, when r=iti(0,14]r=it\in i(0,\frac{1}{4}],

ϕ~(12t)=1Γ(12t)a2xax(u2)2tϕ(u)u𝑑u+O(a2tx2tδ+1)=22t(22t1)2tΓ(12t)(xa)2t+O(a2tx2tδ+1),\displaystyle\begin{split}\widetilde{\phi}(1-2t)&=\frac{1}{\Gamma(1-2t)}\int_{\frac{a}{2x}}^{\frac{a}{x}}\left(\frac{u}{2}\right)^{-2t}\frac{\phi(u)}{u}du+O\left(a^{-2t}x^{2t-\delta}+1\right)\\ &=\frac{2^{2t}(2^{2t}-1)}{2t\Gamma(1-2t)}\left(\frac{x}{a}\right)^{2t}+O\left(a^{-2t}x^{2t-\delta}+1\right),\end{split} (4.13)
Proof.

When 1.999<axT3a2x1.999<\frac{a}{x-T}\leq\frac{3a}{2x}, we get xax\ll a and ϕ~(12t)=O(1)\widetilde{\phi}(1-2t)=O(1) by Lemma 4.3, so the lemma is true in this case. When axT1.999\frac{a}{x-T}\leq 1.999, we have axa\ll x and with (4.12),

ϕ~(12t)\displaystyle\widetilde{\phi}(1-2t) =0(u/2)2tΓ(12t)ϕ(u)u𝑑u+O(0(u2)22tϕ(u)u𝑑u)\displaystyle=\int_{0}^{\infty}\frac{(u/2)^{-2t}}{\Gamma(1-2t)}\frac{\phi(u)}{u}du+O\left(\int_{0}^{\infty}\left(\frac{u}{2}\right)^{2-2t}\frac{\phi(u)}{u}du\right)
=22tΓ(12t)a2xaxu2t1𝑑u+22tΓ(12t)(a2x+2Ta2x+axaxT)u2t1ϕ(u)du\displaystyle=\frac{2^{2t}}{\Gamma(1-2t)}\int_{\frac{a}{2x}}^{\frac{a}{x}}u^{-2t-1}du+\frac{2^{2t}}{\Gamma(1-2t)}\left(\int_{\frac{a}{2x+2T}}^{\frac{a}{2x}}+\int_{\frac{a}{x}}^{\frac{a}{x-T}}\right)u^{-2t-1}\phi(u)du
+O(0u12tϕ(u)𝑑u)\displaystyle+O\left(\int_{0}^{\infty}u^{1-2t}\phi(u)du\right)
=:22t(22t1)2tΓ(12t)(xa)2t+(I1+I2)+O(I3).\displaystyle=:\frac{2^{2t}(2^{2t}-1)}{2t\Gamma(1-2t)}\left(\frac{x}{a}\right)^{2t}+(I_{1}+I_{2})+O(I_{3}).

Recall that we always have the lower bound t¯>0\underline{t}>0 for t=Imrt=\operatorname{Im}r. A bound for I1I_{1} and I2I_{2} follows from the same process as [Sun23, Proof of Lemma 7.2]:

I1+I2(a2x+2Ta2x+axaxT)u2t1ϕ(u)dua2tx2tδ.I_{1}+I_{2}\ll\left(\int_{\frac{a}{2x+2T}}^{\frac{a}{2x}}+\int_{\frac{a}{x}}^{\frac{a}{x-T}}\right)u^{-2t-1}\phi(u)du\ll a^{-2t}x^{2t-\delta}.

We also get

I33a8x3a2xu12t𝑑u(ax)22t1I_{3}\ll\int_{\frac{3a}{8x}}^{\frac{3a}{2x}}u^{1-2t}du\ll\left(\frac{a}{x}\right)^{2-2t}\ll 1

and finish the proof.

Lemma 4.7.

Assume HθH_{\theta} (2.16) for θ16\theta\leq\frac{1}{6}. For r=iti(0,θ2]r=it\in i(0,\frac{\theta}{2}] we have

ϕ^(r)=ekπi2cos(πt)Γ(12+k2+t)Γ(2t)Γ(12k2+t)22tπ2t(m~n~)t(22t1)x2t2t+O(x2tδa2t+a2tx2t+1).\widehat{\phi}(r)=\frac{e^{\frac{k\pi i}{2}}\cos(\pi t)\Gamma(\frac{1}{2}+\frac{k}{2}+t)\Gamma(2t)}{\Gamma(\frac{1}{2}-\frac{k}{2}+t)2^{2t}\pi^{2t}(\tilde{m}\tilde{n})^{t}}\cdot\frac{(2^{2t}-1)x^{2t}}{2t}+O\left(\frac{x^{2t-\delta}}{a^{2t}}+\frac{a^{2t}}{x^{2t}}+1\right).

Moreover,

ϕ^(i4)={2eπi4(21)(xa)12+O(xδ(xa)12+1)for k=12,e3πi4(112)(ax)12+O(xδ(ax)12+1)for k=32.\widehat{\phi}(\tfrac{i}{4})=\left\{\begin{array}[]{lr}2e^{\frac{\pi i}{4}}(\sqrt{2}-1)(\frac{x}{a})^{\frac{1}{2}}+O(x^{-\delta}(\frac{x}{a})^{\frac{1}{2}}+1)&\text{for\ }k=\frac{1}{2},\vspace{4px}\\ e^{\frac{3\pi i}{4}}(1-\tfrac{1}{\sqrt{2}})(\frac{a}{x})^{\frac{1}{2}}+O(x^{-\delta}(\frac{a}{x})^{\frac{1}{2}}+1)&\text{for\ }k=\frac{3}{2}.\end{array}\right.

The implied constants only depend on NN and ν\nu.

Proof.

When t[t¯,θ2]t\in[\underline{t},\frac{\theta}{2}], we substitute Lemma 4.6 into (3.5) and use Lemma 4.3 to get

ϕ^(it)\displaystyle\widehat{\phi}(it) =iπ2ekπi2(cos(kπ2πt)ϕ~(12t)cos(kπ2+πt)ϕ~(1+2t))isin(πt)cos(2πt)Γ(12k2t)Γ(12k2+t)\displaystyle=\frac{i\pi^{2}e^{\frac{k\pi i}{2}}\left(\cos(\frac{k\pi}{2}-\pi t)\widetilde{\phi}(1-2t)-\cos(\frac{k\pi}{2}+\pi t)\widetilde{\phi}(1+2t)\right)}{i\sin(\pi t)\cos(2\pi t)\Gamma(\frac{1}{2}-\frac{k}{2}-t)\Gamma(\frac{1}{2}-\frac{k}{2}+t)}
=π2ekπi2cos(kπ2πt)22t(22t1)(x/a)2tsin(πt)cos(2πt)Γ(12k2t)Γ(12k2+t)2tΓ(12t)+O(x2tδa2t+a2tx2t+1).\displaystyle=\frac{\pi^{2}e^{\frac{k\pi i}{2}}\cos(\frac{k\pi}{2}-\pi t)2^{2t}(2^{2t}-1)(x/a)^{2t}}{\sin(\pi t)\cos(2\pi t)\Gamma(\frac{1}{2}-\frac{k}{2}-t)\Gamma(\frac{1}{2}-\frac{k}{2}+t)2t\Gamma(1-2t)}+O\left(\frac{x^{2t-\delta}}{a^{2t}}+\frac{a^{2t}}{x^{2t}}+1\right).

With the help of the functional equation of the Γ\Gamma function

Γ(z)Γ(1z)=πsin(πz)for z\Gamma(z)\Gamma(1-z)=\frac{\pi}{\sin(\pi z)}\quad\text{for\ }z\in{\mathbb{C}}\setminus{\mathbb{Z}}

and the trigonometric identities

sin(π2x)=cosx,2cosxcosy=cos(x+y)+cos(xy)for x,y,\sin(\tfrac{\pi}{2}-x)=\cos x,\quad 2\cos x\cos y=\cos(x+y)+\cos(x-y)\quad\text{for\ }x,y\in{\mathbb{R}},

we have

πsin(πt)Γ(12t)=2cos(πt)Γ(2t),\displaystyle\frac{\pi}{\sin(\pi t)\Gamma(1-2t)}=2\cos(\pi t)\Gamma(2t),
πΓ(12k2t)=Γ(12+k2+t)cos(kπ2+πt),\displaystyle\frac{\pi}{\Gamma(\frac{1}{2}-\frac{k}{2}-t)}=\Gamma(\tfrac{1}{2}+\tfrac{k}{2}+t)\cos(\tfrac{k\pi}{2}+\pi t),
and 2cos(kπ2πt)cos(kπ2+πt)=cos(2πt).\displaystyle 2\cos(\tfrac{k\pi}{2}-\pi t)\cos(\tfrac{k\pi}{2}+\pi t)=\cos(2\pi t).

Then the first part of the lemma follows. The implied constant only depends on NN and ν\nu because t[t¯,θ2]t\in[\underline{t},\frac{\theta}{2}] is bounded above and below away from 0.

When t=14t=\frac{1}{4}, the process is similar to the proof of Lemma 4.6 with the help of (3.6). First we deal with the case k=12k=\frac{1}{2} with cosu=1+O(u2)\cos u=1+O(u^{2}) for u[0,π2]u\in[0,\frac{\pi}{2}]. Thus, when axT>π2\frac{a}{x-T}>\frac{\pi}{2}, we have xax\ll a and ϕ^(i4)=O(1)\widehat{\phi}(\frac{i}{4})=O(1) in this case. When axTπ2\frac{a}{x-T}\leq\frac{\pi}{2}, we have axa\ll x and

ϕ^(i4)\displaystyle\widehat{\phi}(\tfrac{i}{4}) =eπi40ϕ(u)u32𝑑u+O(0ϕ(u)u12𝑑u)\displaystyle=e^{\frac{\pi i}{4}}\int_{0}^{\infty}\phi(u)u^{-\frac{3}{2}}du+O\left(\int_{0}^{\infty}\phi(u)u^{\frac{1}{2}}du\right)
=eπi4a2xaxu32𝑑u+eπi4(a2x+2Ta2x+axaxT)u32ϕ(u)du+O(1)\displaystyle=e^{\frac{\pi i}{4}}\int_{\frac{a}{2x}}^{\frac{a}{x}}u^{-\frac{3}{2}}du+e^{\frac{\pi i}{4}}\left(\int_{\frac{a}{2x+2T}}^{\frac{a}{2x}}+\int_{\frac{a}{x}}^{\frac{a}{x-T}}\right)u^{-\frac{3}{2}}\phi(u)du+O(1)
=eπi4(222)(xa)12+O(xδ(xa)12)+O(1).\displaystyle=e^{\frac{\pi i}{4}}(2\sqrt{2}-2)\left(\frac{x}{a}\right)^{\frac{1}{2}}+O\left(x^{-\delta}\left(\frac{x}{a}\right)^{\frac{1}{2}}\right)+O(1).

The case for k=32k=\frac{3}{2} is similar using sinu=u+O(u3)\sin u=u+O(u^{3}) for u[0,π2]u\in[0,\frac{\pi}{2}]. ∎

5. Proof of Theorem 1.3 and Theorem 1.5

The proof depends on the following two propositions for the Fourier coefficients of Maass forms, which were originally obtained for the discrete spectrum in [AD22, Theorem 4.1] and [AD19, Theorem 4.3]. The author proved the generalized propositions in [Sun23, §8] to include the continuous spectrum. Recall our notations in Settings 3.1 and 3.2. Suppose that for some β(12,1)\beta\in(\frac{1}{2},1),

N|c>0|S(n,n,c,ν)|c1+βN,ν,ε|n~|ε\sum_{N|c>0}\frac{|S(n,n,c,\nu)|}{c^{1+\beta}}\ll_{N,\nu,\varepsilon}|\tilde{n}|^{\varepsilon} (5.1)

Then we have the following proposition:

Proposition 5.1 ([Sun23, Proposition 8.1]).

Suppose that ν\nu is a multiplier on Γ=Γ0(N)\Gamma=\Gamma_{0}(N) of weight k=±12k=\pm\frac{1}{2} which satisfies (5.1) for some β(12,1)\beta\in(\frac{1}{2},1). Let ρj(n)\rho_{j}(n) denote the Fourier coefficients of an orthonormal basis {vj()}\{v_{j}(\cdot)\} of ~k(N,ν){\tilde{\mathcal{L}}}_{k}(N,\nu). For each singular cusp 𝔞\mathfrak{a} of (Γ,ν)(\Gamma,\nu), let E𝔞(,s)E_{\mathfrak{a}}(\cdot,s) be the associated Eisenstein series. Let ρ𝔞(n,r)\rho_{\mathfrak{a}}(n,r) be defined as in (3.3). Then for x>0x>0 we have

xksgnn~|n~|(xrj2x|ρj(n)|2eπrj+singular𝔞\displaystyle x^{k\operatorname{sgn}\tilde{n}}|\tilde{n}|\left(\sum_{x\leq r_{j}\leq 2x}|\rho_{j}(n)|^{2}e^{-\pi r_{j}}+\right.\sum_{\mathrm{singular\ }\mathfrak{a}} |r|[x,2x]|ρ𝔞(n,r)|2eπ|r|dr)\displaystyle\left.\int_{|r|\in[x,2x]}|\rho_{\mathfrak{a}}(n,r)|^{2}e^{-\pi|r|}dr\right)
N,ν,εx2+|n~|β+εx12βlogβx.\displaystyle\ll_{N,\nu,\varepsilon}x^{2}+|\tilde{n}|^{\beta+\varepsilon}x^{1-2\beta}\log^{\beta}x.
Remark.

In Definition 1.1, an admissible multiplier satisfies (5.1) with β=12+ε\beta=\frac{1}{2}+\varepsilon for any ε\varepsilon.

Proposition 5.2 ([Sun23, Proposition 8.2]).

Suppose that ν\nu is a weight k=±12k=\pm\frac{1}{2} admissible multiplier on Γ=Γ0(N)\Gamma=\Gamma_{0}(N) with MM, DD and BB given in Definition 1.1. Let ρj(n)\rho_{j}(n) denote the Fourier coefficients of an orthonormal basis {vj()}\{v_{j}(\cdot)\} of ~k(N,ν){\tilde{\mathcal{L}}}_{k}(N,\nu). For each singular cusp 𝔞\mathfrak{a} of (Γ,ν)(\Gamma,\nu), let E𝔞(,s)E_{\mathfrak{a}}(\cdot,s) be the associated Eisenstein series. Let ρ𝔞(n,r)\rho_{\mathfrak{a}}(n,r) be defined as in (3.3). Suppose x1x\geq 1. For n0n\neq 0 we factor Bn~=tnun2wn2B\tilde{n}=t_{n}u_{n}^{2}w_{n}^{2} where tnt_{n} is square-free, un|Mu_{n}|M^{\infty} is positive and (wn,M)=1{(w_{n},M)=1}. Then we have

xksgnn~|n~|(|rj|x|ρj(n)|2chπrj+singular𝔞xx|ρ𝔞(n,r)|2chπr𝑑r)N,ν,ε(|n~|131294+un)x3|n~|ε.x^{k\operatorname{sgn}{\tilde{n}}}|{\tilde{n}}|\left(\sum_{|r_{j}|\leq x}\frac{|\rho_{j}(n)|^{2}}{\operatorname{ch}\pi r_{j}}+\sum_{\mathrm{singular\ }\mathfrak{a}}\int_{-x}^{x}\frac{|\rho_{\mathfrak{a}}(n,r)|^{2}}{\operatorname{ch}\pi r}dr\right)\ll_{N,\nu,\varepsilon}\left(|{\tilde{n}}|^{\frac{131}{294}}+u_{n}\right)x^{3}|\tilde{n}|^{\varepsilon}.
Remark.

We make some remarks about the weight kk:

  • The trace formula (Theorem 3.3) works for k=12k=\frac{1}{2} and 32\frac{3}{2}.

  • The estimates on ϕ^\widehat{\phi} and ϕ~\widetilde{\phi} in the previous section work for k=12k=\frac{1}{2} and 32\frac{3}{2}.

  • The above two propositions work for k=12k=\frac{1}{2} and 12{-\tfrac{1}{2}}.

Therefore, in this section, we separate the proof of Theorem 1.3 into two cases k=12k=\frac{1}{2} and 12{-\tfrac{1}{2}}. In the second case we will apply the Maass lowering operator L32L_{\frac{3}{2}} (2.14) to connect the eigenforms of weight 32\frac{3}{2} and weight 12-\frac{1}{2}.

We declare that all implicit constants for the bounds in this section depend on NN, ν\nu and ε\varepsilon, and we drop the subscripts unless specified.

Since the exceptional spectral parameter rji(0,14]r_{j}\in i(0,\frac{1}{4}] of Laplacian Δk\Delta_{k} on Γ=Γ0(N)\Gamma=\Gamma_{0}(N) is discrete, tj=Imrjt_{j}=\operatorname{Im}r_{j} has a positive lower bound denoted as t¯>0\underline{t}>0 depending on NN and ν\nu. We also have 2ImrΔθ2\operatorname{Im}r_{\Delta}\leq\theta assuming HθH_{\theta} (2.16) by Theorem 3.4. For simplicity let

A(m,n):=(m~131294+um)12(n~131294+un)12(m~n~)131588+m~131588un12+n~131588um12+(umun)12A(m,n)\vcentcolon=(\tilde{m}^{\frac{131}{294}}+u_{m})^{\frac{1}{2}}(\tilde{n}^{\frac{131}{294}}+u_{n})^{\frac{1}{2}}\ll(\tilde{m}\tilde{n})^{\frac{131}{588}}+\tilde{m}^{\frac{131}{588}}u_{n}^{\frac{1}{2}}+\tilde{n}^{\frac{131}{588}}u_{m}^{\frac{1}{2}}+(u_{m}u_{n})^{\frac{1}{2}}

and

Au(m,n):=A(m,n)14(m~n~)316(m~n~)143588+m~143588n~316un18+m~316um18n~143588+(m~n~)316(umun)18.\displaystyle\begin{split}A_{u}(m,n)&\vcentcolon=A(m,n)^{\frac{1}{4}}(\tilde{m}\tilde{n})^{\frac{3}{16}}\\ &\ll(\tilde{m}\tilde{n})^{\frac{143}{588}}+\tilde{m}^{\frac{143}{588}}\tilde{n}^{\frac{3}{16}}\,u_{n}^{\frac{1}{8}}+\tilde{m}^{\frac{3}{16}}u_{m}^{\frac{1}{8}}\tilde{n}^{\frac{143}{588}}+(\tilde{m}\tilde{n})^{\frac{3}{16}}(u_{m}u_{n})^{\frac{1}{8}}.\end{split} (5.2)

Recall the notations in Setting 3.1 and Setting 3.2. The following inequalities will be used later in the proof:

A(m,n)Au(m,n)(m~n~)14;A(m,n)\ll A_{u}(m,n)\ll(\tilde{m}\tilde{n})^{\frac{1}{4}}; (5.3)
(ax)βA(m,n)Au(m,n)for 0β32,when xAu(m,n)2.\left(\frac{a}{x}\right)^{\beta}A(m,n)\ll A_{u}(m,n)\quad\text{for }0\leq\beta\leq\frac{3}{2},\quad\text{when }x\gg A_{u}(m,n)^{2}. (5.4)

5.1. On the case k=12k=\frac{1}{2}

Let ρj(n)\rho_{j}(n) denote the coefficients of an orthonormal basis {vj()}\{v_{j}(\cdot)\} of ~12(N,ν){\tilde{\mathcal{L}}}_{\frac{1}{2}}(N,\nu). For each singular cusp 𝔞\mathfrak{a} of Γ=Γ0(N)\Gamma=\Gamma_{0}(N), let ρ𝔞(n,r)\rho_{\mathfrak{a}}(n,r) be defined as in (3.3). Recall the definition of τj(m,n)\tau_{j}(m,n) in Theorem 1.3 and the notations in Settings 3.1 and 3.2. We claim the following proposition:

Proposition 5.3.

With the same setting as Theorem 1.3 for k=12k=\frac{1}{2}, when 2xAu(m,n)22x\geq A_{u}(m,n)^{2}, we have

x<c2xN|cS(m,n,c,ν)crji(0,14](22sj11)τj(m,n)x2sj12sj1(x16+Au(m,n))(m~n~x)ε.\displaystyle\begin{split}\sum_{\begin{subarray}{c}x<c\leq 2x\\ N|c\end{subarray}}\frac{S(m,n,c,\nu)}{c}-\sum_{r_{j}\in i(0,\frac{1}{4}]}(2^{2s_{j}-1}-1)\tau_{j}(m,n)\frac{x^{2s_{j}-1}}{2s_{j}-1}\ll\left(x^{\frac{1}{6}}+A_{u}(m,n)\right)(\tilde{m}\tilde{n}x)^{\varepsilon}.\end{split} (5.5)

We first show that Proposition 5.3 implies Theorem 1.3 in the case k=12k=\frac{1}{2}, which follows from a similar process as [Sun23, after Proposition 9.1]. Recall that 2Imrj=2sj12\operatorname{Im}r_{j}=2s_{j}-1 for rji(0,14]r_{j}\in i(0,\frac{1}{4}] and that the corresponding exceptional eigenvalue λj=14+rj2=sj(1sj)\lambda_{j}=\frac{1}{4}+r_{j}^{2}=s_{j}(1-s_{j}). The sum to be estimated is

N|cXS(m,n,c,ν)crji(0,14]τj(m,n)X2Imrj2Imrj,\sum_{N|c\leq X}\frac{S(m,n,c,\nu)}{c}-\sum_{r_{j}\in i(0,\frac{1}{4}]}\tau_{j}(m,n)\frac{X^{2\operatorname{Im}r_{j}}}{2\operatorname{Im}r_{j}}, (5.6)

where

τj(m,n)=2i12ρj(m)¯ρj(n)π12sj(4m~n~)1sjΓ(sj+14)Γ(2sj1)Γ(sj14).\tau_{j}(m,n)=2i^{\frac{1}{2}}\overline{\rho_{j}(m)}\rho_{j}(n)\pi^{1-2s_{j}}(4\tilde{m}\tilde{n})^{1-s_{j}}\frac{\Gamma(s_{j}+\frac{1}{4})\Gamma(2s_{j}-1)}{\Gamma(s_{j}-\frac{1}{4})}.

Since tj=Imrj[t¯,14]t_{j}=\operatorname{Im}r_{j}\in[\underline{t},\frac{1}{4}] and sj=Imrj+12[t¯+12,34]s_{j}=\operatorname{Im}r_{j}+\frac{1}{2}\in[\underline{t}+\frac{1}{2},\frac{3}{4}], the quantity

π12sj41sjΓ(sj+14)Γ(2sj1)Γ(sj14)\pi^{1-2s_{j}}4^{1-s_{j}}\frac{\Gamma(s_{j}+\frac{1}{4})\Gamma(2s_{j}-1)}{\Gamma(s_{j}-\frac{1}{4})}

is bounded from above and below. By Proposition 5.2,

τj(m,n)2sj1|ρj(m)ρj(n)|(m~n~)1sjA(m,n)(m~n~)12sj+ε.\frac{\tau_{j}(m,n)}{2s_{j}-1}\ll|\rho_{j}(m)\rho_{j}(n)|(\tilde{m}\tilde{n})^{1-s_{j}}\ll A(m,n)(\tilde{m}\tilde{n})^{\frac{1}{2}-s_{j}+\varepsilon}. (5.7)

When XAu(m,n)2X\ll A_{u}(m,n)^{2}, since A(m,n)(m~n~)14A(m,n)\ll(\tilde{m}\tilde{n})^{\frac{1}{4}} by (5.3),

τj(m,n)X2sj12sj1A(m,n)|m~n~|12sj+εAu(m,n)4sj2=A(m,n)sj+12|m~n~|1814sj+εAu(m,n)(m~n~)ε.\displaystyle\begin{split}\tau_{j}(m,n)\frac{X^{2s_{j}-1}}{2s_{j}-1}&\ll A(m,n)|\tilde{m}\tilde{n}|^{\frac{1}{2}-s_{j}+\varepsilon}A_{u}(m,n)^{4s_{j}-2}\\ &=A(m,n)^{s_{j}+\frac{1}{2}}|\tilde{m}\tilde{n}|^{\frac{1}{8}-\frac{1}{4}s_{j}+\varepsilon}\ll A_{u}(m,n)(\tilde{m}\tilde{n})^{\varepsilon}.\end{split} (5.8)

So in this case we get Theorem 1.3 where the τj\tau_{j} terms are absorbed in the errors.

When XAu(m,n)2X\geq A_{u}(m,n)^{2}, the segment for summing Kloosterman sums on 1cAu(m,n)21\leq c\leq A_{u}(m,n)^{2} contributes a Oν,ε(Au(m,n)|m~n~|ε)O_{\nu,\varepsilon}(A_{u}(m,n)|\tilde{m}\tilde{n}|^{\varepsilon}) by condition (2) of Definition 1.1. The segment for Au(m,n)2cXA_{u}(m,n)^{2}\leq c\leq X can be broken into no more than O(logX)O(\log X) dyadic intervals x<c2xx<c\leq 2x with Au(m,n)2xX2A_{u}(m,n)^{2}\leq x\leq\frac{X}{2} and we use Proposition 5.3 for both the Kloosterman sum and the τj\tau_{j} terms. In summing dyadic intervals, for each rji(0,14]r_{j}\in i(0,\frac{1}{4}], we get

=1log2(X/Au(m,n)2)\displaystyle\sum_{\ell=1}^{\left\lceil\log_{2}\left(X/A_{u}(m,n)^{2}\right)\right\rceil} (22sj11)τj(m,n)2sj1(X2)2sj1\displaystyle\frac{(2^{2s_{j}-1}-1)\tau_{j}(m,n)}{2s_{j}-1}\left(\frac{X}{2^{\ell}}\right)^{2s_{j}-1}
=τj(m,n)2sj1X2sj1(12(12sj)log2(X/Au(m,n)2)).\displaystyle\qquad=\frac{\tau_{j}(m,n)}{2s_{j}-1}X^{2s_{j}-1}\left(1-2^{(1-2s_{j})\left\lceil\log_{2}\left(X/A_{u}(m,n)^{2}\right)\right\rceil}\right).

The difference between the above quantity and the quantity τj(m,n)X2sj12sj1\tau_{j}(m,n)\dfrac{X^{2s_{j}-1}}{2s_{j}-1} in (5.6) is

τj(m,n)X2sj12sj12(12sj)log2(X/Au(m,n)2)τj(m,n)2sj1Au(m,n)4sj2Au(m,n){\tau_{j}(m,n)}\frac{X^{2s_{j}-1}}{2s_{j}-1}\cdot 2^{(1-2s_{j})\left\lceil\log_{2}\left(X/A_{u}(m,n)^{2}\right)\right\rceil}\ll\frac{\tau_{j}(m,n)}{2s_{j}-1}A_{u}(m,n)^{4s_{j}-2}\ll A_{u}(m,n) (5.9)

by (5.7). In conclusion, for XAu(m,n)2X\geq A_{u}(m,n)^{2} we get

N|cXS(m,n,c,ν)crji(0,14]τj(m,n)X2sj12sj1\displaystyle\sum_{\begin{subarray}{c}N|c\leq X\end{subarray}}\frac{S(m,n,c,\nu)}{c}-\sum_{r_{j}\in i(0,\frac{1}{4}]}\tau_{j}(m,n)\frac{X^{2s_{j}-1}}{2s_{j}-1}
=Au(m,n)2<cXS(m,n,c,ν)crji(0,14]τj(m,n)X2sj12sj1+O(Au(m,n)|m~n~|ε)\displaystyle=\sum_{\begin{subarray}{c}A_{u}(m,n)^{2}<c\leq X\end{subarray}}\frac{S(m,n,c,\nu)}{c}-\sum_{r_{j}\in i(0,\frac{1}{4}]}\tau_{j}(m,n)\frac{X^{2s_{j}-1}}{2s_{j}-1}+O(A_{u}(m,n)|\tilde{m}\tilde{n}|^{\varepsilon})
==1log2(X/Au(m,n)2)(X2<cX21S(m,n,c,ν)crji(0,14](22sj11)τj(m,n)2sj1(X2)2sj1)\displaystyle=\sum_{\ell=1}^{\left\lceil\log_{2}\left(X/A_{u}(m,n)^{2}\right)\right\rceil}\left(\sum_{\begin{subarray}{c}\frac{X}{2^{\ell}}<c\leq\frac{X}{2^{\ell-1}}\end{subarray}}\frac{S(m,n,c,\nu)}{c}-\sum_{r_{j}\in i(0,\frac{1}{4}]}\frac{(2^{2s_{j}-1}-1)\tau_{j}(m,n)}{2s_{j}-1}\left(\frac{X}{2^{\ell}}\right)^{2s_{j}-1}\right)
+O(Au(m,n)|m~n~|ε)\displaystyle\quad+O(A_{u}(m,n)|\tilde{m}\tilde{n}|^{\varepsilon})
(X16+Au(m,n))|m~n~X|ε\displaystyle\ll\left(X^{\frac{1}{6}}+A_{u}(m,n)\right)|\tilde{m}\tilde{n}X|^{\varepsilon}

where the second equality follows from (5.9) and the last inequality is by Proposition 5.3. Theorem 1.3 follows in the case k=12k=\frac{1}{2}.

The proof of Proposition 5.3 takes the rest of this subsection. For rji(0,14]r_{j}\in i(0,\frac{1}{4}], by Proposition 5.2 we have

m~n~ρj(m)¯ρj(n)A(m,n)(m~n~)ε.\sqrt{\tilde{m}\tilde{n}}\;\overline{\rho_{j}(m)}\rho_{j}(n)\ll A(m,n)(\tilde{m}\tilde{n})^{\varepsilon}.

Recall that a=4πm~n~a=4\pi\sqrt{\tilde{m}\tilde{n}} and δ=13\delta=\frac{1}{3} in Setting 3.1. Thanks to H764H_{\frac{7}{64}} (2.16) and Proposition 3.4, when rj=itji(0,θ2]r_{j}=it_{j}\in i(0,\frac{\theta}{2}] we have 2tj<δ=132t_{j}<\delta=\frac{1}{3}. Since 2xAu(m,n)22x\geq A_{u}(m,n)^{2} by hypothesis, it follows from (5.4) that

m~n~ρj(m)¯ρj(n)(x2tjδa2tj+a2tjx2tj+1)Au(m,n)(m~n~)ε.\displaystyle\sqrt{\tilde{m}\tilde{n}}\;\overline{\rho_{j}(m)}\rho_{j}(n)\left(\frac{x^{2t_{j}-\delta}}{a^{2t_{j}}}+\frac{a^{2t_{j}}}{x^{2t_{j}}}+1\right)\ll A_{u}(m,n)(\tilde{m}\tilde{n})^{\varepsilon}.

Applying Lemma 4.7 where tj[t¯,θ2]t_{j}\in[\underline{t},\frac{\theta}{2}] and recalling the definition of τj\tau_{j} in Theorem 1.3, we get

4m~n~ρj(m)¯ρj(n)chπrjϕ^(rj)=(22sj11)τj(m,n)x2sj12sj1+O(Au(m,n)(m~n~)ε).\displaystyle\begin{split}4\sqrt{\tilde{m}\tilde{n}}&\;\frac{\overline{\rho_{j}(m)}\rho_{j}(n)}{\operatorname{ch}\pi r_{j}}\widehat{\phi}(r_{j})\\ &=(2^{2s_{j}-1}-1)\tau_{j}(m,n)\frac{x^{2s_{j}-1}}{2s_{j}-1}+O\left(A_{u}(m,n)(\tilde{m}\tilde{n})^{\varepsilon}\right).\end{split} (5.10)

When rj=i4r_{j}=\frac{i}{4} and k=12k=\frac{1}{2}, Lemma 4.7 and (5.4) give

4m~n~ρj(m)¯ρj(n)cosπ4ϕ^(i4)=2(21)τj(m,n)x12+O(x12δ(m~n~)ε).\displaystyle\begin{split}4\sqrt{\tilde{m}\tilde{n}}\;\frac{\overline{\rho_{j}(m)}\rho_{j}(n)}{\cos\frac{\pi}{4}}\widehat{\phi}(\tfrac{i}{4})=2(\sqrt{2}-1)\tau_{j}(m,n)x^{\frac{1}{2}}+O\left(x^{\frac{1}{2}-\delta}(\tilde{m}\tilde{n})^{\varepsilon}\right).\end{split} (5.11)

With the help of (5.10) and (5.11) we break up the left hand side of (5.5) to obtain the following analogue to [Sun23, (9.8)]:

|x<c2xN|cS(m,n,c,ν)crji(0,14](22sj11)τj(m,n)x2sj12sj1||x<c2xN|cS(m,n,c,ν)cN|c>0S(m,n,c,ν)cϕ(ac)|+O((x12δ+Au(m,n))(m~n~)ε)+|N|c>0S(m,n,c,ν)cϕ(ac)4m~n~rji(0,14]ρj(m)¯ρj(n)chπrjϕ^(rj)|=:S1+O((x12δ+Au(m,n))(m~n~)ε)+S2.\displaystyle\begin{split}&\left|\sum_{\begin{subarray}{c}x<c\leq 2x\\ N|c\end{subarray}}\frac{S(m,n,c,\nu)}{c}-\sum_{r_{j}\in i(0,\frac{1}{4}]}(2^{2s_{j}-1}-1)\tau_{j}(m,n)\frac{x^{2s_{j}-1}}{2s_{j}-1}\right|\\ \leq&\left|\sum_{\begin{subarray}{c}x<c\leq 2x\\ N|c\end{subarray}}\frac{S(m,n,c,\nu)}{c}-\sum_{N|c>0}\frac{S(m,n,c,\nu)}{c}\phi\left(\frac{a}{c}\right)\right|+O\left(\left(x^{\frac{1}{2}-\delta}+A_{u}(m,n)\right)(\tilde{m}\tilde{n})^{\varepsilon}\right)\\ &+\left|\sum_{N|c>0}\frac{S(m,n,c,\nu)}{c}\phi\left(\frac{a}{c}\right)-4\sqrt{\tilde{m}\tilde{n}}\sum_{r_{j}\in i(0,\frac{1}{4}]}\frac{\overline{\rho_{j}(m)}\rho_{j}(n)}{\operatorname{ch}\pi r_{j}}\widehat{\phi}(r_{j})\right|\\ =:&\;S_{1}+O\left(\left(x^{\frac{1}{2}-\delta}+A_{u}(m,n)\right)(\tilde{m}\tilde{n})^{\varepsilon}\right)+S_{2}.\end{split} (5.12)

The first sum S1S_{1} above can be estimated by condition (2) of Definition 1.1 as

S1xTcx2xc2x+2TN|c|S(m,n,c,ν)|cN,ν,δ,εx12δ(m~n~x)ε.\displaystyle\begin{split}S_{1}\leq\sum_{\begin{subarray}{c}x-T\leq c\leq x\\ 2x\leq c\leq 2x+2T\\ N|c\end{subarray}}\frac{|S(m,n,c,\nu)|}{c}\ll_{N,\nu,\delta,\varepsilon}x^{\frac{1}{2}-\delta}(\tilde{m}\tilde{n}x)^{\varepsilon}.\end{split} (5.13)

We then prove a bound for S2S_{2}. By Theorem 3.3, we have

S2|𝒰12|+|m~n~rj0ρj(m)¯ρj(n)chπrjϕ^(rj)+m~n~singular𝔞ρ𝔞(m,r)¯ρ𝔞(n,r)ϕ^(r)chπr𝑑r|.S_{2}\ll|\mathcal{U}_{\frac{1}{2}}|+\left|\sqrt{\tilde{m}\tilde{n}}\sum_{r_{j}\geq 0}\frac{\overline{\rho_{j}(m)}\rho_{j}(n)}{\operatorname{ch}\pi r_{j}}\widehat{\phi}(r_{j})+\sqrt{\tilde{m}\tilde{n}}\sum_{\mathrm{singular\ }\mathfrak{a}}\int_{-\infty}^{\infty}\overline{\rho_{\mathfrak{a}}(m,r)}\rho_{\mathfrak{a}}(n,r)\frac{\widehat{\phi}(r)}{\operatorname{ch}\pi r}dr\right|. (5.14)

5.1.1. Contribution from holomorphic forms

For k=12k=\frac{1}{2} or 32\frac{3}{2}, recall the notation k\mathscr{B}_{k} before Theorem 3.3. For l1l\geq 1, let {Fj,l()}j\{F_{j,l}(\cdot)\}_{j} be an orthonormal basis of Sk+2l(N,ν)S_{k+2l}(N,\nu) with Fourier coefficient aF,j,la_{F,j,l}. By Proposition 3.8, uniformly for every l1l\geq 1 with dl:=dimSk+2l(N,ν)d_{l}\vcentcolon=\dim S_{k+2l}(N,\nu), we have k+2l52k+2l\geq\frac{5}{2} and

Γ(k+2l1)(4π)k+2l1(m~n~)k+2l12j=1dlaF,j,l(m)¯aF,j,l(n)\displaystyle\frac{\Gamma(k+2l-1)}{(4\pi)^{k+2l-1}(\tilde{m}\tilde{n})^{\frac{k+2l-1}{2}}}\sum_{j=1}^{d_{l}}\overline{a_{F,j,l}(m)}a_{F,j,l}(n)
(Γ(k+2l1)(4πn~)k+2l1j=1dl|aF,j,l(m)|2)12(Γ(k+2l1)(4πm~)k+2l1j=1dl|aF,j,l(n)|2)12\displaystyle\leq\left(\frac{\Gamma(k+2l-1)}{(4\pi\tilde{n})^{k+2l-1}}\sum_{j=1}^{d_{l}}|a_{F,j,l}(m)|^{2}\right)^{\frac{1}{2}}\left(\frac{\Gamma(k+2l-1)}{(4\pi\tilde{m})^{k+2l-1}}\sum_{j=1}^{d_{l}}|a_{F,j,l}(n)|^{2}\right)^{\frac{1}{2}}
(m~1942+um)12(n~1942+un)12(m~n~)ε.\displaystyle\ll(\tilde{m}^{\frac{19}{42}}+u_{m})^{\frac{1}{2}}(\tilde{n}^{\frac{19}{42}}+u_{n})^{\frac{1}{2}}(\tilde{m}\tilde{n})^{\varepsilon}.

We also have

l=1(k+2l1)|ϕ~(k+2l)|1+ax\sum_{l=1}^{\infty}(k+2l-1)\,|\widetilde{\phi}(k+2l)|\ll 1+\frac{a}{x}

by [Dun18, Lemma 5.1 and proof of Lemma 7.1] and Lemma 4.3. Note that [Dun18, Lemma 5.1] is only for k=12k=\frac{1}{2}, while the same process works for k=32k=\frac{3}{2}. Then the contribution from 𝒰k\mathcal{U}_{k} is

𝒰k\displaystyle\mathcal{U}_{k} =l=1k+2l14πϕ~(k+2l)Γ(k+2l1)(4π)k+2l1(m~n~)k+2l12j=1dlaF,j,l(m)¯aF,j,l(n)\displaystyle=\sum_{l=1}^{\infty}\frac{k+2l-1}{4\pi}\,\widetilde{\phi}\,(k+2l)\frac{\Gamma(k+2l-1)}{(4\pi)^{k+2l-1}(\tilde{m}\tilde{n})^{\frac{k+2l-1}{2}}}\sum_{j=1}^{d_{l}}\overline{a_{F,j,l}(m)}a_{F,j,l}(n)
(1+ax)(m~1942+um)12(n~1942+un)12(m~n~)ε.\displaystyle\ll\left(1+\frac{a}{x}\right)(\tilde{m}^{\frac{19}{42}}+u_{m})^{\frac{1}{2}}(\tilde{n}^{\frac{19}{42}}+u_{n})^{\frac{1}{2}}(\tilde{m}\tilde{n})^{\varepsilon}.

Recall a=4πm~n~a=4\pi\sqrt{\tilde{m}\tilde{n}} and (5.2) for the definition of Au(m,n)A_{u}(m,n). Since

n~1942n~13129414+38(n~131294+un)14n~38andunun14n~38(n~131294+un)14n~38,\tilde{n}^{\frac{19}{42}}\ll\tilde{n}^{\frac{131}{294}\cdot\frac{1}{4}+\frac{3}{8}}\leq(\tilde{n}^{\frac{131}{294}}+u_{n})^{\frac{1}{4}}\tilde{n}^{\frac{3}{8}}\quad\text{and}\quad u_{n}\ll u_{n}^{\frac{1}{4}}\tilde{n}^{\frac{3}{8}}\leq(\tilde{n}^{\frac{131}{294}}+u_{n})^{\frac{1}{4}}\tilde{n}^{\frac{3}{8}},

we get (m~1942+um)12(n~1942+un)12Au(m,n)(\tilde{m}^{\frac{19}{42}}+u_{m})^{\frac{1}{2}}(\tilde{n}^{\frac{19}{42}}+u_{n})^{\frac{1}{2}}\ll A_{u}(m,n). Moreover, since

n~1942+1n~13129434+98(n~131294+un)34n~98,nunun34n~98(n~131294+un)34n~98,\tilde{n}^{\frac{19}{42}+1}\ll\tilde{n}^{\frac{131}{294}\cdot\frac{3}{4}+\frac{9}{8}}\leq(\tilde{n}^{\frac{131}{294}}+u_{n})^{\frac{3}{4}}\tilde{n}^{\frac{9}{8}},\ \quad nu_{n}\ll u_{n}^{\frac{3}{4}}\tilde{n}^{\frac{9}{8}}\leq(\tilde{n}^{\frac{131}{294}}+u_{n})^{\frac{3}{4}}\tilde{n}^{\frac{9}{8}},

and 2xAu(m,n)22x\geq A_{u}(m,n)^{2} by hypothesis, we also get

(m~1942+um)12(n~1942+un)12ax(m~1942+um)12(n~1942+un)12aAu(m,n)2Au(m,n).(\tilde{m}^{\frac{19}{42}}+u_{m})^{\frac{1}{2}}(\tilde{n}^{\frac{19}{42}}+u_{n})^{\frac{1}{2}}\cdot\frac{a}{x}\ll(\tilde{m}^{\frac{19}{42}}+u_{m})^{\frac{1}{2}}(\tilde{n}^{\frac{19}{42}}+u_{n})^{\frac{1}{2}}\cdot\frac{a}{A_{u}(m,n)^{2}}\ll A_{u}(m,n).

Finally we conclude

𝒰kAu(m,n)(m~n~)εfor k=12 or 32.\mathcal{U}_{k}\ll A_{u}(m,n)(\tilde{m}\tilde{n})^{\varepsilon}\qquad\text{for }k=\tfrac{1}{2}\text{ or }\tfrac{3}{2}. (5.15)

5.1.2. Contribution from Maass cusp forms and Eisenstein series.

We combine the two propositions at the beginning of this section and bounds on ϕ^\widehat{\phi} in Section 4 to estimate the contribution from the remaining part of S2S_{2} (5.14) other than 𝒰k\mathcal{U}_{k}. The process is the same as [Sun23, §9.1] for |r|1|r|\leq 1 as ϕ^\widehat{\phi} shares the same bound as ϕˇ\check{\phi} there. We record the bounds in the following equations.

Fix k=12k=\frac{1}{2}. In the following estimations we focus on the discrete spectrum rj0r_{j}\geq 0 because each bound for rj[a,b]r_{j}\in[a,b] for any interval [a,b][a,b]\subset{\mathbb{R}} is the same as the bound for r[a,b][b,a]r\in[a,b]\cup[-b,-a] in the continuous spectrum. This is a direct result from Proposition 5.1 and Proposition 5.2. Recall that 2xAu(m,n)22x\geq A_{u}(m,n)^{2} in the assumption of Proposition 5.3.

For r[0,1)r\in[0,1), we apply Lemma 4.2, Proposition 5.2 and Cauchy-Schwarz to get

m~n~r[0,1)|ρj(m)¯ρj(n)chπrjϕ^(rj)|A(m,n)(m~n~x)ε.\sqrt{\tilde{m}\tilde{n}}\sum_{r\in[0,1)}\left|\frac{\overline{\rho_{j}(m)}\rho_{j}(n)}{\operatorname{ch}\pi r_{j}}\widehat{\phi}(r_{j})\right|\ll A(m,n)(\tilde{m}\tilde{n}x)^{\varepsilon}. (5.16)

For r[1,ax)r\in[1,\frac{a}{x}), we apply Proposition 5.2 and ϕ^(r)r1\widehat{\phi}(r)\ll r^{-1} from (4.10). Since

S(R):=m~n~r[1,R]|ρj(m)¯ρj(n)chπrj|A(m,n)R52(m~n~)ε,\displaystyle\begin{split}S(R)\vcentcolon=\sqrt{\tilde{m}\tilde{n}}\sum_{r\in[1,R]}\left|\frac{\overline{\rho_{j}(m)}\rho_{j}(n)}{\operatorname{ch}\pi r_{j}}\right|\ll A(m,n)R^{\frac{5}{2}}(\tilde{m}\tilde{n})^{\varepsilon},\end{split} (5.17)

with the help of (5.4) we have

m~n~r[1,ax)|ρj(m)¯ρj(n)chπrjϕ^(rj)|r1S(r)|r=1ax+1axS(r)r2𝑑rA(m,n)(ax)32(m~n~x)εAu(m,n)(m~n~x)ε.\displaystyle\begin{split}\sqrt{\tilde{m}\tilde{n}}\sum_{r\in[1,\frac{a}{x})}\left|\frac{\overline{\rho_{j}(m)}\rho_{j}(n)}{\operatorname{ch}\pi r_{j}}\widehat{\phi}(r_{j})\right|&\ll r^{-1}S(r)\Big{|}_{r=1}^{\frac{a}{x}}+\int_{1}^{\frac{a}{x}}S(r)r^{-2}dr\\ &\ll A(m,n)\left(\frac{a}{x}\right)^{\frac{3}{2}}(\tilde{m}\tilde{n}x)^{\varepsilon}\ll A_{u}(m,n)(\tilde{m}\tilde{n}x)^{\varepsilon}.\end{split} (5.18)

Let

P(m,n):=2(m~n~)18A(m,n)121.P(m,n)\vcentcolon=2(\tilde{m}\tilde{n})^{\frac{1}{8}}A(m,n)^{-\frac{1}{2}}\geq 1.

Divide rmax(ax,1)r\geq\max(\frac{a}{x},1) into two parts: max(ax,1)r<P(m,n)\max\left(\frac{a}{x},1\right)\leq r<P(m,n) and rmax(ax,1,P(m,n))r\geq\max\left(\frac{a}{x},1,P(m,n)\right). We apply Proposition 5.2 on the first range and ϕ^(r)r1\widehat{\phi}(r)\ll r^{-1} from (4.10) to get

m~n~max(ax,1)rj<P(m,n)|ρj(m)¯ρj(n)chπrjϕ^(rj)|Au(m,n)(m~n~x)ε\sqrt{\tilde{m}\tilde{n}}\sum_{\max(\frac{a}{x},1)\leq r_{j}<P(m,n)}\left|\frac{\overline{\rho_{j}(m)}\rho_{j}(n)}{\operatorname{ch}\pi r_{j}}\widehat{\phi}(r_{j})\right|\\ \ll A_{u}(m,n)(\tilde{m}\tilde{n}x)^{\varepsilon} (5.19)

by partial summation as in (5.18). We divide the second range into dyadic intervals Crj<2CC\leq r_{j}<2C. Applying Proposition 5.1 with β=12+ε\beta=\frac{1}{2}+\varepsilon and ϕ^(r)min(r1,r2xT)\widehat{\phi}(r)\ll\min(r^{-1},r^{-2}\frac{x}{T}) from (4.10), we get

m~n~Crj<2C|ρj(m)¯ρj(n)chπrjϕ^(rj)|min(C1,C2xT)C12(C2+(m~14+n~14)C+(m~n~)14)(m~n~x)ε(min(C12,C12xT)+(m~14+n~14)C12+(m~n~)14C32)(m~n~x)ε.\displaystyle\begin{split}\sqrt{\tilde{m}\tilde{n}}&\sum_{C\leq r_{j}<2C}\left|\frac{\overline{\rho_{j}(m)}\rho_{j}(n)}{\operatorname{ch}\pi r_{j}}\widehat{\phi}(r_{j})\right|\\ &\ll\min\left(C^{-1},C^{-2}\frac{x}{T}\right)C^{-\frac{1}{2}}\left(C^{2}+(\tilde{m}^{\frac{1}{4}}+\tilde{n}^{\frac{1}{4}})C+(\tilde{m}\tilde{n})^{\frac{1}{4}}\right)(\tilde{m}\tilde{n}x)^{\varepsilon}\\ &\ll\left(\min\left(C^{\frac{1}{2}},C^{-\frac{1}{2}}\frac{x}{T}\right)+(\tilde{m}^{\frac{1}{4}}+\tilde{n}^{\frac{1}{4}})C^{-\frac{1}{2}}+(\tilde{m}\tilde{n})^{\frac{1}{4}}C^{-\frac{3}{2}}\right)(\tilde{m}\tilde{n}x)^{\varepsilon}.\end{split} (5.20)

Next we sum over dyadic intervals. For the first term min(C12,C12xT)\min(C^{\frac{1}{2}},C^{-\frac{1}{2}}\frac{x}{T}), when

min(C12,C12xT)=C12:j1: 2jC=xTCP(m,n)C12j=12j2(xT)12(xT)12,\min\left(C^{\frac{1}{2}},C^{-\frac{1}{2}}\frac{x}{T}\right)=C^{\frac{1}{2}}:\quad\sum_{\begin{subarray}{c}j\geq 1:\ 2^{j}C=\frac{x}{T}\\ C\geq P(m,n)\end{subarray}}C^{\frac{1}{2}}\leq\sum_{j=1}^{\infty}2^{-\frac{j}{2}}\left(\frac{x}{T}\right)^{\frac{1}{2}}\ll\left(\frac{x}{T}\right)^{\frac{1}{2}},

and when

min(C12,C12xT)=C12xT:j0:C=2jxTC12xTj=02j2(xT)12(xT)12.\min\left(C^{\frac{1}{2}},C^{-\frac{1}{2}}\frac{x}{T}\right)=C^{-\frac{1}{2}}\frac{x}{T}:\quad\sum_{j\geq 0:\ C=2^{j}\frac{x}{T}}C^{-\frac{1}{2}}\frac{x}{T}\leq\sum_{j=0}^{\infty}2^{-\frac{j}{2}}\left(\frac{x}{T}\right)^{\frac{1}{2}}\ll\left(\frac{x}{T}\right)^{\frac{1}{2}}.

So after summing up from (5.20), recalling Tx1δT\asymp x^{1-\delta} in Setting 3.1, using CP(m,n)C\geq P(m,n) and (5.3), we have

m~n~rjmax(ax,1,P(m,n))|ρj(m)¯ρj(n)chπrjϕ^(rj)|((xT)12+(m~+n~)14(m~n~)116A(m,n)14+(m~n~)116A(m,n)34)(m~n~x)ε(xδ2+(m~n~)316A(m,n)14)(m~n~x)ε.\displaystyle\begin{split}\sqrt{\tilde{m}\tilde{n}}&\sum_{r_{j}\geq\max(\frac{a}{x},1,P(m,n))}\left|\frac{\overline{\rho_{j}(m)}\rho_{j}(n)}{\operatorname{ch}\pi r_{j}}\widehat{\phi}(r_{j})\right|\\ &\ll\left(\left(\frac{x}{T}\right)^{\frac{1}{2}}+(\tilde{m}+\tilde{n})^{\frac{1}{4}}(\tilde{m}\tilde{n})^{-\frac{1}{16}}A(m,n)^{\frac{1}{4}}+(\tilde{m}\tilde{n})^{\frac{1}{16}}A(m,n)^{\frac{3}{4}}\right)(\tilde{m}\tilde{n}x)^{\varepsilon}\\ &\ll\left(x^{\frac{\delta}{2}}+(\tilde{m}\tilde{n})^{\frac{3}{16}}A(m,n)^{\frac{1}{4}}\right)(\tilde{m}\tilde{n}x)^{\varepsilon}.\end{split} (5.21)

Combining (5.19) and (5.21) we have

m~n~rjmax(ax,1)|ρj(m)¯ρj(n)chπrjϕ^(rj)|(xδ2+Au(m,n))(m~n~x)ε.\sqrt{\tilde{m}\tilde{n}}\sum_{r_{j}\geq\max(\frac{a}{x},1)}\left|\frac{\overline{\rho_{j}(m)}\rho_{j}(n)}{\operatorname{ch}\pi r_{j}}\widehat{\phi}(r_{j})\right|\\ \ll\left(x^{\frac{\delta}{2}}+A_{u}(m,n)\right)(\tilde{m}\tilde{n}x)^{\varepsilon}. (5.22)

From (5.12), (5.13), (5.14), (5.15), (5.16), (5.18), and (5.22), we get

x<c2xN|cS(m,n,c,ν)crji(0,14](22sj11)\displaystyle\sum_{\begin{subarray}{c}x<c\leq 2x\\ N|c\end{subarray}}\frac{S(m,n,c,\nu)}{c}-\sum_{r_{j}\in i(0,\frac{1}{4}]}(2^{2s_{j}-1}-1) τj(m,n)x2sj12sj1\displaystyle\tau_{j}(m,n)\frac{x^{2s_{j}-1}}{2s_{j}-1}
(x12δ+xδ2+Au(m,n))(m~n~x)ε.\displaystyle\ll\left(x^{\frac{1}{2}-\delta}+x^{\frac{\delta}{2}}+A_{u}(m,n)\right)(\tilde{m}\tilde{n}x)^{\varepsilon}.

Proposition 5.3 follows by choosing δ=13\delta=\frac{1}{3}. We finish the proof of Theorem 1.3 in weight 12\frac{1}{2}.

5.2. On the case k=12k=-\frac{1}{2}

Recall the remark after Proposition 5.2. Let ρj(n)\rho_{j}^{\prime}(n) denote the Fourier coefficients of an orthonormal basis {vj()}\{v_{j}^{\prime}(\cdot)\} of ~32(N,ν){\tilde{\mathcal{L}}}_{\frac{3}{2}}(N,\nu). For each singular cusp 𝔞\mathfrak{a} of (Γ,ν)(\Gamma,\nu), let E𝔞(,s)E_{\mathfrak{a}}^{\prime}(\cdot,s) be the associated Eisenstein series in weight 32\frac{3}{2}. Let ρ𝔞(n,r)\rho_{\mathfrak{a}}^{\prime}(n,r) be defined as in (3.3) associated with E𝔞(z,12+ir)E_{\mathfrak{a}}^{\prime}(z,\frac{1}{2}+ir) for rr\in{\mathbb{R}}.

Recall the definition of the Maass lowering operator LkL_{k} in (2.14) and HθH_{\theta} (2.16) for θ=764\theta=\frac{7}{64}. By [DFI02, (4.52)] (where they used Λk\Lambda_{k} for the lowering operator and λ(s)=s(1s)\lambda(s)=s(1-s)), the set

{vj:=(116+rj2)12L32vj:rji4} is an orthonormal basis of rji4~12(N,ν,rj).\left\{v_{j}\vcentcolon=\left(\tfrac{1}{16}+r_{j}^{2}\right)^{-\frac{1}{2}}L_{\frac{3}{2}}v_{j}^{\prime}:\ r_{j}\neq\tfrac{i}{4}\right\}\text{ is an orthonormal basis of }\bigoplus_{r_{j}\neq\frac{i}{4}}{\tilde{\mathcal{L}}}_{-\frac{1}{2}}(N,\nu,r_{j}).

Combining [DFI02, (4.36), (4.27) and the last equation of p. 502], for rji4r_{j}\neq\frac{i}{4} and n~>0\tilde{n}>0, since

L32(W34n~,Imr(4πn~y)e(n~x))=(116+r2)Wn~4,Imr(4πn~y)e(n~x),L_{\frac{3}{2}}\left(W_{\frac{3}{4}\tilde{n},\,\operatorname{Im}r}(4\pi\tilde{n}y)e(\tilde{n}x)\right)=-(\tfrac{1}{16}+r^{2})W_{-\frac{\tilde{n}}{4},\,\operatorname{Im}r}(4\pi\tilde{n}y)e(\tilde{n}x),

the Fourier coefficient ρj(n)\rho_{j}(n) of vjv_{j} satisfies

ρj(n)=(116+r2)12ρj(n)for rji4,n~>0,\rho_{j}(n)=-(\tfrac{1}{16}+r^{2})^{\frac{1}{2}}\rho_{j}^{\prime}(n)\qquad\text{for }r_{j}\neq\tfrac{i}{4},\ \tilde{n}>0, (5.23)

and then

|ρj(n)||ρj(n)|if |rj|1,Imrjθ2and|ρj(n)|r|ρj(n)|if rj1,|\rho_{j}(n)|\asymp|\rho_{j}^{\prime}(n)|\quad\text{if }|r_{j}|\leq 1,\ \operatorname{Im}r_{j}\leq\tfrac{\theta}{2}\qquad\text{and}\quad|\rho_{j}(n)|\asymp r|\rho_{j}^{\prime}(n)|\quad\text{if }r_{j}\geq 1, (5.24)

where the bound 2Imrjθ2\operatorname{Im}r_{j}\leq\theta is from Proposition 3.4.

In the case rj=i4r_{j}=\frac{i}{4}, (2.21) and (2.22) show that ρj(n)=0\rho_{j}(n)=0 and

τj(m,n)=0for n~>0,rj=i4.\tau_{j}(m,n)=0\qquad\text{for\ }\tilde{n}>0,\ r_{j}=\tfrac{i}{4}. (5.25)

Moreover, by [DFI02, (4.48)], if E𝔞(z,s)E_{\mathfrak{a}}(z,s) is the Eisenstein series defined in weight 12-\frac{1}{2}, then

L32E𝔞(z,12+ir)=(14ir)E𝔞(z,s)and(116+r2)12|ρ𝔞(n,r)|=|ρ𝔞(n,r)|.L_{\frac{3}{2}}E_{\mathfrak{a}}^{\prime}(z,\tfrac{1}{2}+ir)=(\tfrac{1}{4}-ir)E_{\mathfrak{a}}(z,s)\quad\text{and}\quad(\tfrac{1}{16}+r^{2})^{\frac{1}{2}}|\rho_{\mathfrak{a}}^{\prime}(n,r)|=|\rho_{\mathfrak{a}}(n,r)|.

We also get

|ρ𝔞(n,r)||ρ𝔞(n,r)|if r[1,1]and|ρ𝔞(n,r)|r|ρ𝔞(n,r)|if |r|1.|\rho_{\mathfrak{a}}(n,r)|\asymp|\rho_{\mathfrak{a}}^{\prime}(n,r)|\quad\text{if }r\in[-1,1]\qquad\text{and}\quad|\rho_{\mathfrak{a}}(n,r)|\asymp r|\rho_{\mathfrak{a}}^{\prime}(n,r)|\quad\text{if }|r|\geq 1. (5.26)

We have the following proposition:

Proposition 5.4.

With the same setting as Theorem 1.3 for k=12k=-\frac{1}{2}, when 2xAu(m,n)22x\geq A_{u}(m,n)^{2}, we have

x<c2xN|cS(m,n,c,ν)crji(0,θ2](22sj11)τj(m,n)x2sj12sj1(x16+Au(m,n))(m~n~x)ε,\displaystyle\begin{split}\sum_{\begin{subarray}{c}x<c\leq 2x\\ N|c\end{subarray}}\frac{S(m,n,c,\nu)}{c}-\sum_{r_{j}\in i(0,\frac{\theta}{2}]}(2^{2s_{j}-1}-1)\tau_{j}(m,n)\frac{x^{2s_{j}-1}}{2s_{j}-1}\ll\left(x^{\frac{1}{6}}+A_{u}(m,n)\right)(\tilde{m}\tilde{n}x)^{\varepsilon},\end{split}

Note that here τj(m,n)\tau_{j}(m,n) is defined in weight 12-\frac{1}{2}, i.e.

τj(m,n)=2i12ρj(m)¯ρj(n)π12sj(4m~n~)1sjΓ(sj14)Γ(2sj1)Γ(sj+14)\tau_{j}(m,n)=2i^{-\frac{1}{2}}\overline{\rho_{j}(m)}\rho_{j}(n)\pi^{1-2s_{j}}(4\tilde{m}\tilde{n})^{1-s_{j}}\frac{\Gamma(s_{j}-\frac{1}{4})\Gamma(2s_{j}-1)}{\Gamma(s_{j}+\frac{1}{4})}

where ρj(n)\rho_{j}(n) is from (5.24) as the Fourier coefficient of vj~12(N,ν,rj)v_{j}\in{\tilde{\mathcal{L}}}_{-\frac{1}{2}}(N,\nu,r_{j}).

The proof that Proposition 5.4 implies Theorem 1.3 in the case k=12k=-\frac{1}{2} is the same as the case of weight 12\frac{1}{2} before. This is because τj(m,n)=0\tau_{j}(m,n)=0 for rj=i4r_{j}=\frac{i}{4} (5.25) and because (5.7), (5.8) and (5.9) still hold for rji(0,θ2]r_{j}\in i(0,\frac{\theta}{2}] (the process only involves estimates on ρj(n)\rho_{j}(n) with some applications of Proposition 5.2 in weight 12-\frac{1}{2}). In the rest of this subsection we prove Proposition 5.4.

First we show that the main terms corresponding to rj=itji(0,θ2]r_{j}=it_{j}\in i(0,\frac{\theta}{2}] are the same when we shift the weight between 12-\frac{1}{2} and 32\frac{3}{2}. Recall sj=12+tjs_{j}=\frac{1}{2}+t_{j}. Let τj(m,n)\tau_{j}^{\prime}(m,n) denote the corresponding coefficients for x2sj1x^{2s_{j}-1} in weight 32\frac{3}{2}:

τj(m,n)=2e3πi4ρj(m)¯ρj(n)π2tj(4m~n~)12tjΓ(54+tj)Γ(2tj)Γ(tj14),\tau_{j}^{\prime}(m,n)=2e^{\frac{3\pi i}{4}}\overline{\rho_{j}^{\prime}(m)}\rho_{j}^{\prime}(n)\pi^{-2t_{j}}(4\tilde{m}\tilde{n})^{\frac{1}{2}-t_{j}}\frac{\Gamma(\frac{5}{4}+t_{j})\Gamma(2t_{j})}{\Gamma(t_{j}-\frac{1}{4})},

where ρj(n)\rho_{j}^{\prime}(n) is defined at the beginning of this subsection.

We claim that

τj(m,n)=τj(m,n),for m~,n~>0 and rji(0,14].\tau_{j}^{\prime}(m,n)=\tau_{j}(m,n),\quad\text{for }\tilde{m},\tilde{n}>0\text{ and }r_{j}\in i(0,\tfrac{1}{4}]. (5.27)

When rj=i4r_{j}=\frac{i}{4}, this is true because both of them equal to zero by (5.25) and Γ(0)=\Gamma(0)=\infty. When rji(0,θ2]r_{j}\in i(0,\frac{\theta}{2}],

τj(m,n)\displaystyle\tau_{j}(m,n) =2eπi4ρj(m)¯ρj(n)π2tj(4m~n~)12tjΓ(14+tj)Γ(2tj)Γ(34+tj)\displaystyle=2e^{-\frac{\pi i}{4}}\overline{\rho_{j}(m)}\rho_{j}(n)\pi^{-2t_{j}}(4\tilde{m}\tilde{n})^{\frac{1}{2}-t_{j}}\frac{\Gamma(\frac{1}{4}+t_{j})\Gamma(2t_{j})}{\Gamma(\frac{3}{4}+t_{j})}
=2e3πi4(116tj2)ρj(m)¯ρj(n)π2tj(4m~n~)12tjΓ(54+tj)/(14+tj)(14+tj)Γ(14+tj)Γ(2tj)\displaystyle=-2e^{\frac{3\pi i}{4}}\left(\frac{1}{16}-t_{j}^{2}\right)\overline{\rho_{j}^{\prime}(m)}\rho_{j}^{\prime}(n)\pi^{-2t_{j}}(4\tilde{m}\tilde{n})^{\frac{1}{2}-t_{j}}\frac{\Gamma(\frac{5}{4}+t_{j})/(\frac{1}{4}+t_{j})}{(-\frac{1}{4}+t_{j})\Gamma(-\frac{1}{4}+t_{j})}\Gamma(2t_{j})
=τj(m,n).\displaystyle=\tau_{j}^{\prime}(m,n).

Recall that the definition on ϕ^\widehat{\phi} (3.5) is for weight k0k\geq 0 and here we use ϕ^\widehat{\phi} for weight 32\frac{3}{2}. We derive

4m~n~ρj(m)¯ρj(n)chπrjϕ^(rj)=(22sj11)τj(m,n)x2sj12sj1+O(Au(m,n)(m~n~)ε).4\sqrt{\tilde{m}\tilde{n}}\;\frac{\overline{\rho_{j}^{\prime}(m)}\rho_{j}^{\prime}(n)}{\operatorname{ch}\pi r_{j}}\widehat{\phi}(r_{j})=(2^{2s_{j}-1}-1)\tau_{j}^{\prime}(m,n)\frac{x^{2s_{j}-1}}{2s_{j}-1}+O\left(A_{u}(m,n)(\tilde{m}\tilde{n})^{\varepsilon}\right). (5.28)

by the same process as we derive (5.10) above. Since τj(m,n)=0\tau_{j}^{\prime}(m,n)=0 when rj=i4r_{j}=\frac{i}{4}, we have 2tjθ<δ2t_{j}\leq\theta<\delta (with θ=764\theta=\frac{7}{64} (2.16) and δ=13\delta=\frac{1}{3} chosen later) by Proposition 3.4 and still get

|x<c2xN|cS(m,n,c,ν)crji(0,θ2](22sj11)τj(m,n)x2sj12sj1||x<c2xN|cS(m,n,c,ν)cN|c>0S(m,n,c,ν)cϕ(ac)|+O(Au(m,n)(m~n~)ε)+|N|c>0S(m,n,c,ν)cϕ(ac)4m~n~rji(0,θ2]ρj(m)¯ρj(n)chπrjϕ^(rj)|=:S3+O(Au(m,n)(m~n~)ε)+S4.\displaystyle\begin{split}&\left|\sum_{\begin{subarray}{c}x<c\leq 2x\\ N|c\end{subarray}}\frac{S(m,n,c,\nu)}{c}-\sum_{r_{j}\in i(0,\frac{\theta}{2}]}(2^{2s_{j}-1}-1)\tau_{j}^{\prime}(m,n)\frac{x^{2s_{j}-1}}{2s_{j}-1}\right|\\ \leq&\left|\sum_{\begin{subarray}{c}x<c\leq 2x\\ N|c\end{subarray}}\frac{S(m,n,c,\nu)}{c}-\sum_{N|c>0}\frac{S(m,n,c,\nu)}{c}\phi\left(\frac{a}{c}\right)\right|+O\left(A_{u}(m,n)(\tilde{m}\tilde{n})^{\varepsilon}\right)\\ &+\left|\sum_{N|c>0}\frac{S(m,n,c,\nu)}{c}\phi\left(\frac{a}{c}\right)-4\sqrt{\tilde{m}\tilde{n}}\sum_{r_{j}\in i(0,\frac{\theta}{2}]}\frac{\overline{\rho_{j}^{\prime}(m)}\rho_{j}^{\prime}(n)}{\operatorname{ch}\pi r_{j}}\widehat{\phi}(r_{j})\right|\\ =:&\;S_{3}+O\left(A_{u}(m,n)(\tilde{m}\tilde{n})^{\varepsilon}\right)+S_{4}.\end{split} (5.29)

The first sum S3S_{3} above can be estimated similarly by condition (2) of Definition 1.1 as

S3xTcx2xc2x+2TN|c|S(m,n,c,ν)|cN,ν,δ,εx12δ(m~n~x)ε.\displaystyle\begin{split}S_{3}\leq\sum_{\begin{subarray}{c}x-T\leq c\leq x\\ 2x\leq c\leq 2x+2T\\ N|c\end{subarray}}\frac{|S(m,n,c,\nu)|}{c}\ll_{N,\nu,\delta,\varepsilon}x^{\frac{1}{2}-\delta}(\tilde{m}\tilde{n}x)^{\varepsilon}.\end{split} (5.30)

By Theorem 3.3,

S4|𝒰32|+|m~n~rj0ρj(m)¯ρj(n)chπrjϕ^(rj)+m~n~singular𝔞ρ𝔞(m,r)¯ρ𝔞(n,r)ϕ^(r)chπr𝑑r|.S_{4}\ll|\mathcal{U}_{\frac{3}{2}}|+\left|\sqrt{\tilde{m}\tilde{n}}\sum_{r_{j}\geq 0}\frac{\overline{\rho_{j}^{\prime}(m)}\rho_{j}^{\prime}(n)}{\operatorname{ch}\pi r_{j}}\widehat{\phi}(r_{j})+\sqrt{\tilde{m}\tilde{n}}\sum_{\mathrm{singular\ }\mathfrak{a}}\int_{-\infty}^{\infty}\overline{\rho_{\mathfrak{a}}^{\prime}(m,r)}\rho_{\mathfrak{a}}^{\prime}(n,r)\frac{\widehat{\phi}(r)}{\operatorname{ch}\pi r}dr\right|.

The bound for 𝒰32\mathcal{U}_{\frac{3}{2}} is done in (5.15). Estimates for the remaining part of S4S_{4} follow from the same process as §5.1.2 in the case of weight 12\frac{1}{2}, taking (5.24) and (5.26) into account. For the same reason as the beginning of §5.1.2, we just record the bounds with respect to the discrete spectrum here.

For r[0,1)r\in[0,1), we apply Proposition 5.2, (5.24) and (4.2) to get

m~n~r[0,1)|ρj(m)¯ρj(n)chπrjϕ^(rj)|m~n~r[0,1)|ρj(m)¯ρj(n)chπrjϕ^(rj)|A(m,n)(m~n~x)ε.\sqrt{\tilde{m}\tilde{n}}\sum_{r\in[0,1)}\left|\frac{\overline{\rho_{j}^{\prime}(m)}\rho_{j}^{\prime}(n)}{\operatorname{ch}\pi r_{j}}\widehat{\phi}(r_{j})\right|\ll\sqrt{\tilde{m}\tilde{n}}\sum_{r\in[0,1)}\left|\frac{\overline{\rho_{j}(m)}\rho_{j}(n)}{\operatorname{ch}\pi r_{j}}\widehat{\phi}(r_{j})\right|\ll A(m,n)(\tilde{m}\tilde{n}x)^{\varepsilon}. (5.31)

For r[1,ax)r\in[1,\frac{a}{x}), we apply Proposition 5.2, ρj(n)rj1|ρj(n)|\rho_{j}^{\prime}(n)\ll r_{j}^{-1}|\rho_{j}(n)| from (5.24), and ϕ^(r)1\widehat{\phi}(r)\ll 1 from (4.10). Since

s(R):=m~n~r[1,R]|ρj(m)¯ρj(n)chπrj|A(m,n)R72(m~n~)ε\displaystyle\begin{split}s(R)\vcentcolon=\sqrt{\tilde{m}\tilde{n}}\sum_{r\in[1,R]}\left|\frac{\overline{\rho_{j}(m)}\rho_{j}(n)}{\operatorname{ch}\pi r_{j}}\right|\ll A(m,n)R^{\frac{7}{2}}(\tilde{m}\tilde{n})^{\varepsilon}\end{split} (5.32)

by Cauchy-Schwarz, with the help of (5.4) we have

m~n~rj[1,ax)|ρj(m)¯ρj(n)chπrjϕ^(rj)|m~n~rj[1,ax)|ρj(m)¯ρj(n)chπrjrj2|r2s(r)|r=1ax+1axs(r)r3𝑑rA(m,n)(ax)32(m~n~x)εAu(m,n)(m~n~x)ε.\displaystyle\begin{split}\sqrt{\tilde{m}\tilde{n}}\sum_{r_{j}\in[1,\frac{a}{x})}\left|\frac{\overline{\rho_{j}^{\prime}(m)}\rho_{j}^{\prime}(n)}{\operatorname{ch}\pi r_{j}}\widehat{\phi}(r_{j})\right|&\ll\sqrt{\tilde{m}\tilde{n}}\sum_{r_{j}\in[1,\frac{a}{x})}\left|\frac{\overline{\rho_{j}(m)}\rho_{j}(n)}{\operatorname{ch}\pi r_{j}}r_{j}^{-2}\right|\\ &\ll r^{-2}s(r)\Big{|}_{r=1}^{\frac{a}{x}}+\int_{1}^{\frac{a}{x}}s(r)r^{-3}dr\\ &\ll A(m,n)\left(\frac{a}{x}\right)^{\frac{3}{2}}(\tilde{m}\tilde{n}x)^{\varepsilon}\\ &\ll A_{u}(m,n)(\tilde{m}\tilde{n}x)^{\varepsilon}.\end{split} (5.33)

We still let

P(m,n)=2(m~n~)18A(m,n)121P(m,n)=2(\tilde{m}\tilde{n})^{\frac{1}{8}}A(m,n)^{-\frac{1}{2}}\geq 1

and divide rmax(ax,1)r\geq\max(\frac{a}{x},1) into two parts: max(ax,1)r<P(m,n)\max\left(\frac{a}{x},1\right)\leq r<P(m,n) and rmax(ax,1,P(m,n))r\geq\max\left(\frac{a}{x},1,P(m,n)\right). In the first range, we apply Proposition 5.2, (5.24) and ϕ^(r)1\widehat{\phi}(r)\ll 1 from (4.10) to get

m~n~max(ax,1)rj<P(m,n)|ρj(m)¯ρj(n)chπrjϕ^(rj)|Au(m,n)(m~n~x)ε\sqrt{\tilde{m}\tilde{n}}\sum_{\max(\frac{a}{x},1)\leq r_{j}<P(m,n)}\left|\frac{\overline{\rho_{j}^{\prime}(m)}\rho_{j}^{\prime}(n)}{\operatorname{ch}\pi r_{j}}\widehat{\phi}(r_{j})\right|\\ \ll A_{u}(m,n)(\tilde{m}\tilde{n}x)^{\varepsilon} (5.34)

by partial summation similar as (5.33). We divide the second range into dyadic intervals Crj<2CC\leq r_{j}<2C and apply Proposition 5.1, (5.24) and ϕ^(r)min(1,xrT)\widehat{\phi}(r)\ll\min(1,\frac{x}{rT}) from (4.10):

m~n~Crj<2C|ρj(m)¯ρj(n)chπrjϕ^(rj)|m~n~Crj<2C|ρj(m)¯ρj(n)chπrjrj2ϕ^(rj)|min(1,xCT)C2(C52+(m~14+n~14)C32+(m~n~)14C12)(m~n~x)ε(min(C12,C12xT)+(m~14+n~14)C12+(m~n~)14C32)(m~n~x)ε.\displaystyle\begin{split}\sqrt{\tilde{m}\tilde{n}}&\sum_{C\leq r_{j}<2C}\left|\frac{\overline{\rho_{j}^{\prime}(m)}\rho_{j}^{\prime}(n)}{\operatorname{ch}\pi r_{j}}\widehat{\phi}(r_{j})\right|\ll\sqrt{\tilde{m}\tilde{n}}\sum_{C\leq r_{j}<2C}\left|\frac{\overline{\rho_{j}(m)}\rho_{j}(n)}{\operatorname{ch}\pi r_{j}}r_{j}^{-2}\widehat{\phi}(r_{j})\right|\\ &\ll\min\left(1,\frac{x}{CT}\right)C^{-2}\left(C^{\frac{5}{2}}+(\tilde{m}^{\frac{1}{4}}+\tilde{n}^{\frac{1}{4}})C^{\frac{3}{2}}+(\tilde{m}\tilde{n})^{\frac{1}{4}}C^{\frac{1}{2}}\right)(\tilde{m}\tilde{n}x)^{\varepsilon}\\ &\ll\left(\min\left(C^{\frac{1}{2}},C^{-\frac{1}{2}}\frac{x}{T}\right)+(\tilde{m}^{\frac{1}{4}}+\tilde{n}^{\frac{1}{4}})C^{-\frac{1}{2}}+(\tilde{m}\tilde{n})^{\frac{1}{4}}C^{-\frac{3}{2}}\right)(\tilde{m}\tilde{n}x)^{\varepsilon}.\end{split} (5.35)

Summing up from (5.35) similar as we did after (5.20) and recalling Tx1δT\asymp x^{1-\delta} in Setting 3.1, we have

m~n~rjmax(ax,1,P(m,n))|ρj(m)¯ρj(n)chπrjϕ^(rj)|((xT)12+(m~+n~)14(m~n~)116A(m,n)14+(m~n~)116A(m,n)34)(m~n~x)ε(xδ2+(m~n~)316A(m,n)14)(m~n~x)ε.\displaystyle\begin{split}\sqrt{\tilde{m}\tilde{n}}&\sum_{r_{j}\geq\max(\frac{a}{x},1,P(m,n))}\left|\frac{\overline{\rho_{j}^{\prime}(m)}\rho_{j}^{\prime}(n)}{\operatorname{ch}\pi r_{j}}\widehat{\phi}(r_{j})\right|\\ &\ll\left(\left(\frac{x}{T}\right)^{\frac{1}{2}}+(\tilde{m}+\tilde{n})^{\frac{1}{4}}(\tilde{m}\tilde{n})^{-\frac{1}{16}}A(m,n)^{\frac{1}{4}}+(\tilde{m}\tilde{n})^{\frac{1}{16}}A(m,n)^{\frac{3}{4}}\right)(\tilde{m}\tilde{n}x)^{\varepsilon}\\ &\ll\left(x^{\frac{\delta}{2}}+(\tilde{m}\tilde{n})^{\frac{3}{16}}A(m,n)^{\frac{1}{4}}\right)(\tilde{m}\tilde{n}x)^{\varepsilon}.\end{split} (5.36)

From (5.34) and (5.36) we have

m~n~rjmax(ax,1)|ρj(m)¯ρj(n)chπrjϕ^(rj)|(xδ2+Au(m,n))(m~n~x)ε.\sqrt{\tilde{m}\tilde{n}}\sum_{r_{j}\geq\max(\frac{a}{x},1)}\left|\frac{\overline{\rho_{j}^{\prime}(m)}\rho_{j}^{\prime}(n)}{\operatorname{ch}\pi r_{j}}\widehat{\phi}(r_{j})\right|\\ \ll\left(x^{\frac{\delta}{2}}+A_{u}(m,n)\right)(\tilde{m}\tilde{n}x)^{\varepsilon}. (5.37)

Combining (5.29), (5.30), (5.15), (5.31), (5.33), and (5.37), we get

x<c2xN|cS(m,n,c,ν)crji(0,14](22sj11)\displaystyle\sum_{\begin{subarray}{c}x<c\leq 2x\\ N|c\end{subarray}}\frac{S(m,n,c,\nu)}{c}-\sum_{r_{j}\in i(0,\frac{1}{4}]}(2^{2s_{j}-1}-1) τj(m,n)x2sj12sj1\displaystyle\tau_{j}(m,n)\frac{x^{2s_{j}-1}}{2s_{j}-1}
(x12δ+xδ2+Au(m,n))(m~n~x)ε.\displaystyle\ll\left(x^{\frac{1}{2}-\delta}+x^{\frac{\delta}{2}}+A_{u}(m,n)\right)(\tilde{m}\tilde{n}x)^{\varepsilon}.

Proposition 5.4 follows by choosing δ=13\delta=\frac{1}{3} and we finish the proof of Theorem 1.3.

Proof of Theorem 1.5.

The proof follows from the same process as [Sun23, §9.2]. Note that we need to restrict rj=i4τj(m,n)=0\sum_{r_{j}=\frac{i}{4}}\tau_{j}(m,n)=0 when m~>0\tilde{m}>0, n~>0\tilde{n}>0 and k=12k=\frac{1}{2} (and the conjugate case m~<0\tilde{m}<0, n~<0\tilde{n}<0 and k=12k=-\frac{1}{2} by (2.8)), otherwise the sum may not converge. ∎

Acknowledgement

The author thanks Professor Scott Ahlgren for his plenty of helpful discussions and suggestions.

References

  • [AA16] Scott Ahlgren and Nickolas Andersen. Algebraic and transcendental formulas for the smallest parts function. Adv. Math., 289:411–437, 2016.
  • [AA18] Scott Ahlgren and Nickolas Andersen. Kloosterman sums and Maass cusp forms of half integral weight for the modular group. Int. Math. Res. Not. IMRN, 2018(2):492–570, 2018.
  • [AD19] Scott Ahlgren and Alexander Dunn. Maass forms and the mock theta function f(q)f(q). Math. Ann., 374(3-4):1681–1718, 2019.
  • [AD20] Nickolas Andersen and William D. Duke. Modular invariants for real quadratic fields and Kloosterman sums. Algebra Number Theory, 14(6):1537–1575, 2020.
  • [AD22] Nickolas Andersen and William Duke. Asymptotic distribution of traces of singular moduli. Discrete Anal., 4:1–14, 2022.
  • [AH91] George E. Andrews and Dean Hickerson. Ramanujan’s “lost” notebook VII: The sixth order mock theta functions. Adv. Math., 89:60–105, 1991.
  • [Blo08] Valentin Blomer. Sums of Hecke eigenvalues over values of quadratic polynomials. Int. Math. Res. Not. IMRN, 2008:rnn 059, 2008.
  • [BO06] Kathrin Bringmann and Ken Ono. The f(q)f(q) mock theta function conjecture and partition ranks. Invent. Math., 165(2):243–266, 2006.
  • [BO12] Kathrin Bringmann and Ken Ono. Coefficients of harmonic Maass forms. In Krishnaswami Alladi and Frank Garvan, editors, Partitions, q-Series, and Modular Forms, Proceedings of the Conference Held at the University of Florida, Gainesville, FL, March 2008. Developments in Mathematics, vol. 23, pages 23–38. Springer New York, NY, 2012.
  • [DFI02] W. Duke, J.B. Friedlander, and H. Iwaniec. The subconvexity problem for Artin LL-functions. Invent. Math., 149(3):489–577, September 2002.
  • [DLMF] NIST Digital Library of Mathematical Functions. http://dlmf.nist.gov/, Release 1.0.25 of 2019-12-15. F. W. J. Olver, A. B. Olde Daalhuis, D. W. Lozier, B. I. Schneider, R. F. Boisvert, C. W. Clark, B. R. Miller, B. V. Saunders, H. S. Cohl, and M. A. McClain, eds.
  • [Dun90] T. Mark Dunster. Bessel functions of purely imaginary order, with an application to second-order linear differential equations having a large parameter. SIAM J. Math. Anal., 21(4):995–1018, 1990.
  • [Dun18] Alexander Dunn. Uniform bounds for sums of Kloosterman sums of half integral weight. Res. Number Theory, 4(4):Art. 45, 21, 2018.
  • [GS83] Dorian Goldfeld and Peter Sarnak. Sums of Kloosterman sums. Invent. Math., 71(2):243–250, 1983.
  • [Iwa87] Henryk Iwaniec. Fourier coefficients of modular forms of half-integral weight. Invent. Math., 87:385–401, 1987.
  • [Kim03] Henry H. Kim. Functoriality for the exterior square of GL4\mathrm{GL}_{4} and the symmetric fourth of GL2\mathrm{GL}_{2}. J. Amer. Math. Soc., 16(1):139–183, 2003. With appendix 1 by D. Ramakrishnan and appendix 2 by Kim and P. Sarnak.
  • [Kno70] Marvin I. Knopp. Modular functions in analytic number theory. Markham Publishing Co., Chicago, 1970.
  • [Lan00] L. J. Landau. Bessel functions: monotonicity and bounds. J. London Math. Soc., 61(2):197–215, 2000.
  • [Olv97] Frank W. J. Olver. Asymptotics and Special Functions. AKP classics. A K Peters, Wellesley, Massachusetts, 1997.
  • [Pri00] Wladimir de Azevedo Pribitkin. A generalization of the Goldfeld-Sarnak estimate on Selberg’s Kloosterman zeta-function. Forum Math., 12(4):449–459, 2000.
  • [Pro79] N. V. Proskurin. The summation formulas for general Kloosterman sums. Zap. Nauchn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI), 82:103–135, 1979.
  • [Pro05] N. V. Proskurin. On the general Kloosterman sums. J. Math. Sci., 129(3):3874–3889, 2005.
  • [Rad73] Hans Rademacher. Topics in analytic number theory. Springer-Verlag, New York-Heidelberg, 1973. Edited by E. Grosswald, J. Lehner and M. Newman, Die Grundlehren der mathematischen Wissenschaftern, Band 169.
  • [Ran77] Robert A. Rankin. Modular Forms and Functions. Cambridge University Press, 1977.
  • [Sel65] Atle Selberg. On the estimation of Fourier coefficients of modular forms. In Proc. Sympos. Pure Math., Vol. VIII, pages 1–15. Amer. Math. Soc., Providence, R.I., 1965.
  • [ST09] Peter Sarnak and Jacob Tsimerman. On Linnik and Selberg’s conjecture about sums of Kloosterman sums. In Yuri Tschinkel and Yuri Zarhin, editors, Algebra, Arithmetic, and Geometry: Volume II: In Honor of Yu. I. Manin, pages 619–635, Boston, 2009. Birkhäuser Boston.
  • [Sun23] Qihang Sun. Uniform bounds for Kloosterman bounds of half-integral weight with applications. https://arxiv.org/abs/2305.19651, 2023.
  • [Wai17] Fabian Waibel. Fourier coefficients of half-integral weight cusp forms and Waring’s problem. Ramanujan J., 47:185–200, 2017.