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

on the Rankin-Selberg problem in families

Jiseong Kim University at Buffalo, Department of Mathematics 244 Mathematics Building Buffalo, NY 14260-2900 [email protected]
Abstract.

In this paper, we investigate the Rankin-Selberg problem over short intervals in families of holomorphic modular forms and Hecke-Maass cusp forms. Our investigation assumes a Lindelöf-on-average bound for holomorphic modular forms, and for Hecke-Maass cusp forms, we make no assumptions.

Keywords. Automorphic form; Hecke eigenvalue; Rankin-Selberg problem; Short intervals.

1. Introduction

Let ={z=x+iy|x,y(0,)},\mathbb{H}=\{z=x+iy|x\in\mathbb{R},y\in(0,\infty)\}, and let G=SL(2,).G=SL(2,\mathbb{Z}). Define jγ(z)=(cz+d)1,j_{\gamma}(z)=(cz+d)^{-1}, where γ=(abcd)G.\gamma=\left(\begin{matrix}a&b\\ c&d\end{matrix}\right)\in G. A holomorphic function f:f:\mathbb{H}\rightarrow\mathbb{C} satisfying the equation

f(γz)=jγ(z)kf(z)for all γSL(2,),f(\gamma z)=j_{\gamma}(z)^{-k}f(z)\;\;\;\textrm{for all }\;\;\gamma\in SL(2,\mathbb{Z}),

is called a modular form of weight k.k. It is well known that any such modular form f(z)f(z) has a Fourier expansion at the cusp \infty given by

(1.1) f(z)=n=0bf(n)e(nz)f(z)=\sum_{n=0}^{\infty}b_{f}(n)e(nz)

where e(z)=e2πiz,e(z)=e^{2\pi iz}, and the normalized Fourier coefficient af(n)a_{f}(n) of f(z)f(z) is defined as

(1.2) af(n):=bf(n)n(k1)/2.a_{f}(n):=b_{f}(n)n^{-(k-1)/2}.

The set of all modular forms of a fixed weight kk forms a vector space, and we denote it by Mk.M_{k}. Furthermore, we focus on a subspace of this space, denoted by CkC_{k}, which comprises of all modular forms of a weight kk that have a zero constant term. Let us consider the nnth Hecke operator, denoted by TnT_{n}, which is defined as

(Tnf)(z)=1nad=nb(modd)f(az+bd)(T_{n}f)(z)=\frac{1}{\sqrt{n}}\sum_{ad=n}\sum_{b(\rm{mod}\thinspace\it{d})}f(\frac{az+b}{d})

for all fCk.f\in C_{k}.

It is well established that there exists an orthonormal basis of eigenfunctions, called Hecke cusp forms, for all Hecke operators TnT_{n} in the space of modular forms CkC_{k}. In this paper, we will use SkS_{k} to denote this orthonormal basis of CkC_{k}.

When ff is a Hecke cusp form, eigenvalues λf(n)\lambda_{f}(n) of the nnth Hecke operator are real numbers that satisfy the following properties:

af(n)=af(1)λf(n), and λf(m)λf(n)=d|(m,n)λf(mnd2).a_{f}(n)=a_{f}(1)\lambda_{f}(n),\;\;\textrm{ and }\;\;\lambda_{f}(m)\lambda_{f}(n)=\sum_{d|(m,n)}\lambda_{f}\left(\frac{mn}{d^{2}}\right).

For details, see [11, Chapter 14]. A function u:Gu:G\setminus\mathbb{H}\rightarrow\mathbb{C} is called a Maass form of type vv\in\mathbb{C} if uu satisfies the following properties:

  • G\|u(z)|2dxdyy2<,\displaystyle\iint_{G\textbackslash\mathbb{H}}|u(z)|^{2}\frac{dxdy}{y^{2}}<\infty,

  • u(γz)=u(z)u(\gamma z)=u(z) for all γG,\gamma\in G,

  • Δu=v(1v)u,\Delta u=v(1-v)u, where Δ=y2(2x2+2y2),\displaystyle\Delta=-y^{2}\left(\frac{\partial^{2}}{\partial x^{2}}+\frac{\partial^{2}}{\partial y^{2}}\right),

  • 01u(z)𝑑x=0.\displaystyle\int_{0}^{1}u(z)dx=0.

When uu is a non-constant Maass form of type v,v, we can express it using the Whittaker expansion given by

u(z)=2πyn0au(n)Kv12(2π|n|y)e(nx).u(z)=\sqrt{2\pi y}\sum_{n\neq 0}a_{u}(n)K_{v-\frac{1}{2}}(2\pi|n|y)e(nx).

For further details, see [2, Chapter 3]. We are interested in a particular type of Maass forms called Hecke-Maass forms, which are eigenfunctions of the Hecke operators TnT_{n} for all nn.

There have been several results on the asymptotic evaluations of sums related to Hecke eigenvalues. One of the most notable results is the Rankin-Selberg formula that was first established by Rankin [18] and Selberg [19]. The formula states that for a Hecke cusp form ff,

(1.3) 1mXλf(m)2=Ress=1L(sym2f,s)ζ(2)X+Of(X3/5)\sum_{1\leq m\leq X}\lambda_{f}(m)^{2}=\frac{\textrm{Res}_{s=1}L(\textrm{sym}^{2}f,s)}{\zeta(2)}X+O_{f}(X^{3/5})

where L(sym2f,s)L\left(\textrm{sym}^{2}f,s\right) is the symmetric square LL-function of f,f, ζ(s)\zeta(s) is the Riemann zeta function (for the details, see [11, 5.12]). The Rankin-Selberg problem is to improve the exponent 3/5. Recently, Huang [8] improved the exponent 3/53/5 to 3/51/560+o(1)3/5-1/560+o(1). It is worth noting that these results apply not only to the holomorphic SL(2,)SL(2,\mathbb{Z}) Hecke cusp forms but also to the SL(2,)SL(2,\mathbb{Z}) Hecke-Maass cusp forms. For convenience, we denote the coefficient of XX by c1,f.c_{1,f}. Ivic [10] proved that for 1HX,1\leq H\leq X, the following holds:

X2X(x<mx+Hλf(m)2c1,fH)2𝑑xf,kX(9+12α)/(7+4α)H8/(7+4α)\int_{X}^{2X}\left(\sum_{x<m\leq x+H}\lambda_{f}(m)^{2}-c_{1,f}H\right)^{2}dx\ll_{f,k}X^{\left(9+12\alpha\right)/\left(7+4\alpha\right)}H^{8/\left(7+4\alpha\right)}

where α=lim suptlog|ζ(1/2+it)|logt.\alpha=\limsup_{t\rightarrow\infty}\frac{\log\left|\zeta(1/2+it)\right|}{\log t}. Note that under the Lindelöf hypothesis, we have α=0.\alpha=0. In [10], Ivic also proved that, assuming the generalized Lindelöf hypothesis L(sym2f,12+it)f,ϵkϵ(|t|+1)ϵ\displaystyle L(\textrm{sym}^{2}f,\frac{1}{2}+it)\ll_{f,\epsilon}k^{\epsilon}(|t|+1)^{\epsilon}, the following holds:

X2X(x<mx+Hλf(m)2c1,fH)2𝑑xf,k,ϵX1+ϵH4/3(Xϵ<H<X1ϵ).\int_{X}^{2X}\left(\sum_{x<m\leq x+H}\lambda_{f}(m)^{2}-c_{1,f}H\right)^{2}dx\ll_{f,k,\epsilon}X^{1+\epsilon}H^{4/3}\;\;\;\;\;(X^{\epsilon}<H<X^{1-\epsilon}).

Recently, Mangerel [15] proved the following theorem:

Theorem.

Let 10H0X/(10logX)10\leq H_{0}\leq X/(10\log X) and set H:=H0logXH:=H_{0}\log X. Then there exist a constant θ>0\theta>0 such that

1XX2X(x<mx+Hλf(m)2c1,fH)2𝑑xH2(loglogH0logH0+loglogX(logX)θ).\frac{1}{X}\int_{X}^{2X}\left(\sum_{x<m\leq x+H}\lambda_{f}(m)^{2}-c_{1,f}H\right)^{2}dx\ll H^{2}\left(\frac{\log\log H_{0}}{\log H_{0}}+\frac{\log\log X}{(\log X)^{\theta}}\right).

In order to prove Theorem 1.1, we apply the Petersson trace formula (see Lemma 2.3) and assume the following Lindelöf-on-average bound for holomorphic cusp forms:

(1.4) fSkφf|L(sym2f,1/2+it)|2ϵkϵ(|t|+1)ϵfort,for allϵ(0,1),\sum_{f\in S_{k}}\varphi_{f}|L(\textrm{sym}^{2}f,1/2+it)|^{2}\ll_{\epsilon}k^{\epsilon}(|t|+1)^{\epsilon}\;\;\textrm{for}\;\;t\in\mathbb{R},\;\textrm{for all}\;\epsilon\in(0,1),

where

φf:=(f2gSk1g2)1.\varphi_{f}:=\left(\|f\|^{2}\sum_{g\in S_{k}}\frac{1}{\|g\|^{2}}\right)^{-1}.
Theorem 1.1.

Assume the Lindelöf-on-average bound (1.4). Let ϵ\epsilon be a sufficiently small number in (0,1). Let 1logH<logX<(logk)/4.1\ll\log H<\log X<(\log k)/4. Then

(1.5) fSkφf1XX2X(x<mx+Hλf2(m)c1,fH)2𝑑xϵkϵH1+ϵ.\begin{split}\sum_{f\in S_{k}}\varphi_{f}\frac{1}{X}\int_{X}^{2X}\left(\sum_{x<m\leq x+H}\lambda_{f}^{2}(m)-c_{1,f}H\right)^{2}dx\ll_{\epsilon}k^{\epsilon}H^{1+\epsilon}.\end{split}

Note that

fSk1f2(4π)k1Γ(k1)ask,1f2ϵk1+ϵ.\sum_{f\in S_{k}}\frac{1}{\|f\|^{2}}\sim\frac{(4\pi)^{k-1}}{\Gamma(k-1)}\;\textrm{as}\;k\rightarrow\infty,\;\;\frac{1}{\|f\|^{2}}\ll_{\epsilon}k^{-1+\epsilon}.

Let uju_{j} be an SL(2,)SL(2,\mathbb{Z}) Hecke-Maass cusp form with Laplacian eigenvalue 14+tj2,\frac{1}{4}+t_{j}^{2}, and let λuj(n)\lambda_{u_{j}}(n) be the nnth normalized Fourier coefficient of uj.u_{j}. The function λuj(n)\lambda_{u_{j}}(n) also satisfies

(1.6) λuj(m)λuj(n)=d|(m,n)λuj(mnd2).\lambda_{u_{j}}(m)\lambda_{u_{j}}(n)=\sum_{d|(m,n)}\lambda_{u_{j}}\left(\frac{mn}{d^{2}}\right).

As mentioned before, we have

(1.7) 1mXλuj(m)2=Ress=1L(sym2uj,s)ζ(2)X+Ouj(X3/51/560+o(1)).\sum_{1\leq m\leq X}\lambda_{u_{j}}(m)^{2}=\frac{\textrm{Res}_{s=1}L(\textrm{sym}^{2}u_{j},s)}{\zeta(2)}X+O_{u_{j}}(X^{3/5-1/560+o(1)}).

For convenience, we denote the coefficient of XX by c1,uj.c_{1,u_{j}}. Our methods for Theorem 1.1 apply equally well to SL(2,)SL(2,\mathbb{Z}) Hecke-Maass cusp forms uj.u_{j}. For the Hecke-Maass cusp forms uj,u_{j}, one needs to apply the Kuznetsov trace formula (see Lemma 2.4) instead of the Petersson trace formula. According to the result of Khan and Young [12, Theorem 1.1, Theorem 1.3], for any T>1T>1 and 0<ϵ<2,0<\epsilon<2, we have

(1.8) Ttj2T|L(sym2uj,1/2+iU)|2ϵT2+ϵforU[0,(2ϵ)T].\sum_{T\leq t_{j}\leq 2T}\left|L(\textrm{sym}^{2}u_{j},1/2+iU)\right|^{2}\ll_{\epsilon}T^{2+\epsilon}\;\;\textrm{for}\;\;U\in[0,(2-\epsilon)T].

Therefore, we unconditionally prove the following theorem.

Theorem 1.2.

Let ϵ\epsilon be a sufficiently small number in (0,1). Let 1logH<logX,1\ll\log H<\log X, and let 2logX(logH)/2<logT.2\log X-(\log H)/2<\log T. Then

(1.9) Ttj2Tϖj1XX2X(x<mx+Hλuj2(m)c1,ujH)2𝑑xϵTϵH1+ϵ\sum_{T\leq t_{j}\leq 2T}\varpi_{j}\frac{1}{X}\int_{X}^{2X}\left(\sum_{x<m\leq x+H}\lambda_{u_{j}}^{2}(m)-c_{1,u_{j}}H\right)^{2}dx\ll_{\epsilon}T^{\epsilon}H^{1+\epsilon}

where

w(j)=ζ(2)L(sym2uj,1),ϖj=w(j)(Ttj2Tw(j))1.w(j)=\frac{\zeta(2)}{L(\textrm{sym}^{2}u_{j},1)},\;\;\;\varpi_{j}=w(j)\left(\sum_{T\leq t_{j}\leq 2T}w(j)\right)^{-1}.

Note that

T2ϵϵTtj2Tw(j)T2,w(j)ϵ|tj+1|ϵT^{2-\epsilon}\ll_{\epsilon}\sum_{T\leq t_{j}\leq 2T}w(j)\ll T^{2},\;\;w(j)\ll_{\epsilon}|t_{j}+1|^{\epsilon}

(see [7]).

Remark 1.3.

As in [5], one could consider problems related to the variance over arithmetic progressions. However, since we are considering families, these problems are relatively straightforward to handle without resorting to the methods described in [5]. For readers who are interested, we provide a simple example in the Appendix to illustrate this point.

1.1. Sketch of the proof

Because the proofs of Theorem 1.1 and Theorem 1.2 share many similarities, we will provide a sketch of the proof of Theorem 1.1. In [5], Gorodetsky, Matömaki, Radziwiłł and Rodgers used the identity μ2(m)=d2mμ(d)\mu^{2}(m)=\sum_{d^{2}\mid m}\mu(d) to evaluate the variance of the number of square-free integers in short intervals. The squares of Hecke eigenvalues also exhibit a similar convolution structure, as shown by the following equation:

(1.10) λf(m)2=dmλf(d2).\lambda_{f}\left(m\right)^{2}=\sum_{d\mid m}\lambda_{f}\left(d^{2}\right).

It should be noted that while the identity for square-free integers involves d2d^{2}, the convolution structure (1.10) considers dd. As a result, when dealing with square-free integers, we are essentially working with a sparse set of squares. However, when working with Hecke eigenvalues, we are considering a set of natural numbers that is not sparse. This phenomenon is explained well in the paper [4] by Gorodetsky, Mangerel, and Rogers. The authors of [5] split the sum over d2d^{2} into d2zd^{2}\leq z and d2>zd^{2}>z for some z.z. We split the sum over dd into dHd\leq H and d>H.d>H. By (1.10), we have

x<nx+Hλf(n)2=1dx+Hλf(d2)xd<nx+Hd1.\sum_{x<n\leq x+H}\lambda_{f}(n)^{2}=\sum_{1\leq d\leq x+H}\lambda_{f}\left(d^{2}\right)\sum_{\frac{x}{d}<n\leq\frac{x+H}{d}}1.

When dd is small, it is expected that

x<nx+Hλf(n)2Hdx+Hλf(d2)d.\sum_{x<n\leq x+H}\lambda_{f}(n)^{2}\sim H\sum_{d\leq x+H}\frac{\lambda_{f}\left(d^{2}\right)}{d}.

This can be demonstrated by using the Fourier expansion of

xxd<nx+Hd1x\mapsto\sum_{\frac{x}{d}<n\leq\frac{x+H}{d}}1

(see Proposition 3.1). However, when dd is large (i.e. d>Hd>H), we have

xd<nx+Hd1=δ[xd]+1=[x+Hd]\sum_{\frac{x}{d}<n\leq\frac{x+H}{d}}1=\delta_{[\frac{x}{d}]+1=[\frac{x+H}{d}]}

where

δ[xd]+1=[x+Hd]:={1when[xd]+1=[x+Hd]0otherwise}.\delta_{{[\frac{x}{d}]+1=[\frac{x+H}{d}]}}:=\left.\begin{cases}1&\;\textrm{when}\;\;[\frac{x}{d}]+1=[\frac{x+H}{d}]\\ 0&\;\textrm{otherwise}\end{cases}\right\}.

As a result, we encounter a double sum

H<d1x+HH<d2x+Hλf(d12)λf(d22)δ[xd1]+1=[x+Hd1]δ[xd2]+1=[x+Hd2],\sum_{H<d_{1}\leq x+H}\sum_{H<d_{2}\leq x+H}\lambda_{f}\left(d_{1}^{2}\right)\lambda_{f}\left(d_{2}^{2}\right)\delta_{[\frac{x}{d_{1}}]+1=[\frac{x+H}{d_{1}}]}\delta_{[\frac{x}{d_{2}}]+1=[\frac{x+H}{d_{2}}]},

which is difficult to control. To avoid this, we take a different approach by considering the average over all modular forms ff of weight kk and applying the Petersson trace formula (see Lemma 2.3). For the remaining parts, we utilize the Lindelöf-on-average bound to obtain an approximation of the form

c1,fmH2λf(m2)mc_{1,f}\sim\sum_{m\leq H^{2}}\frac{\lambda_{f}\left(m^{2}\right)}{m}

with a small error term (see Proposition 3.5).

Remark 1.4.

Our approach clearly applies to demonstrating analogous results for a linear space Sk(N)S_{k}(N) of cusp forms with weight kk and level N.N. Some of the results in our paper still hold if we replace λf\lambda_{f} and SkS_{k} with the coefficients of the standard LL-function attached to a GL(n)GL(n) automorphic representation and a suitable family of GL(n)GL(n) automorphic representations, respectively. For instance, suppose we let Ag(n,1,1,,1)A_{g}(n,1,1,\dots,1) denote the nnth coefficient of an LL-function L(g,s)L(g,s) associated with a SL(n,)SL(n,\mathbb{Z}) Hecke-Maass cusp form gg. We can address the corresponding problem for gg by considering the expression:

x<nd<x+HAg(d,1,,1)L(g,1)H.\sum_{x<nd<x+H}A_{g}(d,1,\dots,1)-L(g,1)H.

Thus, using orthogonality (for instance, see [3]), we can easily modify our method to establish that the averages of the following expressions over suitable families are bounded by the same upper bound as that of for the SL(2,)SL(2,\mathbb{Z}) Hecke-Maass (or holomorphic) forms:

(1.11) X2X(x<ndx+HdzAg(d,1,,1)HdzAg(d,1,,1)d)2𝑑x,X2X(x<ndx+Hz<dAg(d,1,,1))2𝑑x\begin{split}&\int_{X}^{2X}\left(\sum_{\begin{subarray}{c}x<nd\leq x+H\\ d\leq z\end{subarray}}A_{g}\left(d,1,\dots,1\right)-H\sum_{d\leq z}\frac{A_{g}\left(d,1,\dots,1\right)}{d}\right)^{2}dx,\\ &\int_{X}^{2X}\left(\sum_{\begin{subarray}{c}x<nd\leq x+H\\ z<d\end{subarray}}A_{g}\left(d,1,\dots,1\right)\right)^{2}dx\end{split}

Assuming the Lindelöf-on-average bound for L(g,s)L(g,s), we can easily modify our method to show that

HL(g,1)HdzAg(d,1,,1)dHL(g,1)-H\sum_{d\leq z}\frac{A_{g}(d,1,\dots,1)}{d}

is sufficiently small.

Remark 1.5.

It is conjectured (see [9]) that

1mXλf2(m)c1,fX=Of,k(X38+o(1)).\sum_{1\leq m\leq X}\lambda_{f}^{2}(m)-c_{1,f}X=O_{f,k}\left(X^{\frac{3}{8}+o(1)}\right).

By Theorem 1.1, when H=X3/8+ϵH=X^{3/8+\epsilon}, we have that for almost all fSkf\in S_{k} and x[X,2X],x\in[X,2X],

(x<mx+Hλf2(m)c1,fH)2ϵkϵH1+ϵfor sufficiently largek.\left(\sum_{x<m\leq x+H}\lambda_{f}^{2}(m)-c_{1,f}H\right)^{2}\ll_{\epsilon}k^{\epsilon}H^{1+\epsilon}\;\;\;\textrm{for sufficiently large}\;k.

This leads to the conclusion that

x<mx+X3/8+ϵλf2(m)c1,fX3/8+ϵϵkϵX3/16\sum_{x<m\leq x+X^{3/8+\epsilon}}\lambda_{f}^{2}(m)-c_{1,f}X^{3/8+\epsilon}\ll_{\epsilon}k^{\epsilon}X^{3/16}

for almost all fSkf\in S_{k} and x[X,2X].x\in[X,2X]. It is worth noting that the Lindelöf hypothesis implies that when H=X1/2+ϵ,H=X^{1/2+\epsilon},

x<mx+X1/2+ϵλf2(m)c1,fX1/2+ϵ=o(X1/2+ϵ).\sum_{x<m\leq x+X^{1/2+\epsilon}}\lambda_{f}^{2}(m)-c_{1,f}X^{1/2+\epsilon}=o(X^{1/2+\epsilon}).

2. Lemma

For convenience, we need to smooth the integrals in equations (1.5) and (1.9). We use the Beurling-Selberg polynomial from [5]. To simplify notation, we denote the number of the divisors of mm as τ(m)\tau(m). Additionally, for any condition 𝒜\mathcal{A}, we define

δ𝒜:={1when𝒜is satisfied0otherwise}.\delta_{\mathcal{A}}:=\left.\begin{cases}1&\;\textrm{when}\;\;\mathcal{A}\;\;\textrm{is satisfied}\\ 0&\;\textrm{otherwise}\end{cases}\right\}.

We use the common notation that the least common multiple of aa and bb, and the greatest common divisor of aa and bb are denoted as [a,b][a,b] and (a,b),(a,b), respectively. Finally, we use the convention that ϵ\epsilon denotes an arbitrarily small positive quantity that may vary from line to line.

Lemma 2.1.

There exist functions σ\sigma_{-} and σ+\sigma_{+} that map from \mathbb{R} to \mathbb{R} and satisfy the following properties:

  • The Fourier transforms σ^(x):=σ(y)e(xy)𝑑y\hat{\sigma}_{-}(x):=\int_{\mathbb{R}}\sigma_{-}(y)e(-xy)dy and σ^+(x):=σ+(y)e(xy)𝑑y\hat{\sigma}_{+}(x):=\int_{\mathbb{R}}\sigma_{+}(y)e(-xy)dy have support [BHϵ2,BHϵ2][-BH^{\frac{\epsilon}{2}},BH^{\frac{\epsilon}{2}}] for a sufficiently large absolute constant BB.

  • σ1[1,2]σ+,\sigma_{-}\leq 1_{[1,2]}\leq\sigma_{+}, where 1[1.2]1_{[1.2]} is the indicator function over the interval [1,2].[1,2].

  • 1Hϵ2σ(x)𝑑x,σ+(x)𝑑x1+Hϵ2.\displaystyle 1-H^{-\frac{\epsilon}{2}}\leq\int_{\mathbb{R}}\sigma_{-}(x)dx,\int_{\mathbb{R}}\sigma_{+}(x)dx\leq 1+H^{-\frac{\epsilon}{2}}.

Proof.

See [17, Chapter 1, page 6]. ∎

Since λf(n)\lambda_{f}(n) is divisor-bounded, we can apply the following lemma.

Lemma 2.2.

Let δ(0,1)\delta\in(0,1) be a fixed constant and suppose XYXδ2X\geq Y\geq X^{\delta}\geq 2. Then, for any positive integer ll, we have the inequality

(2.1) 1YX<nX+Yτ(n)lδ(logX)2l1.\frac{1}{Y}\sum_{X<n\leq X+Y}\tau\left(n\right)^{l}\ll_{\delta}\left(\log X\right)^{2^{l}-1}.
Proof.

Using Shiu’s Theorem ([16, Lemma 2.3,(i)]), we obtain the bound

1YX<nX+Yτ(n)lδpX(1+τ(p)l1p).\frac{1}{Y}\sum_{X<n\leq X+Y}\tau\left(n\right)^{l}\ll_{\delta}\prod_{p\leq X}\left(1+\frac{\tau\left(p\right)^{l}-1}{p}\right).

Since τ(p)=2\tau(p)=2 for any prime pp, we have

(2.2) pX(1+τ(p)l1p)=pX(1+2l1p)exp(pX2l1p).\begin{split}\prod_{p\leq X}\left(1+\frac{\tau\left(p\right)^{l}-1}{p}\right)&=\prod_{p\leq X}\left(1+\frac{2^{l}-1}{p}\right)\\ &\leq exp\left(\sum_{p\leq X}\frac{2^{l}-1}{p}\right).\end{split}

By Merten’s theorem, we see that

pX1p=loglogX+C+o(1)\sum_{p\leq X}\frac{1}{p}=\log\log X+C+o(1)

for some constant C.C. Therefore,

exp(pX2l1p)(logX)2l1.exp\left(\sum_{p\leq X}\frac{2^{l}-1}{p}\right)\ll\left(\log X\right)^{2^{l}-1}.

Lemma 2.3.

(Petersson trace formula) For any two natural numbers mm and n,n, we have

(2.3) fSkλf(n)λf(m)f2=(4π)k1Γ(k1)δm=n+O((4π)k1Γ(k1)(log(3mn))2τ((m,n))(mn)1/4k1/2),\begin{split}\sum_{f\in S_{k}}\frac{\lambda_{f}(n)\lambda_{f}(m)}{\|f\|^{2}}&=\frac{(4\pi)^{k-1}}{\Gamma(k-1)}\delta_{m=n}\\ &+O\left(\frac{(4\pi)^{k-1}}{\Gamma(k-1)}(\log(3mn))^{2}\frac{\tau\big{(}(m,n)\big{)}(mn)^{1/4}}{k^{1/2}}\right),\end{split}

where the implied constant is absolute, and f\|f\| denotes the Petersson norm of ff over CkC_{k}.

Proof.

See [11, Corollary 14.24, Theorem 16.7]. ∎

Lemma 2.4.

(Kuznetsov trace formula) Let {uj}\{u_{j}\} be an orthonormal basis of Hecke-Maass cusp forms for SL(2,),SL(2,\mathbb{Z}), and let 14+tj2\frac{1}{4}+t_{j}^{2} be the Laplace eigenvalue of uj.u_{j}. Suppose T>1,T>1, and let m,n.m,n\in\mathbb{N}. Then, we have

(2.4) jw(j)etj/Tλuj(m)λuj(n)=T26δm=n+Oϵ(T1+ϵ(mn)ϵ+(mn)1/2+ϵ)\sum_{j}w(j)e^{-t_{j}/T}\lambda_{u_{j}}(m)\lambda_{u_{j}}(n)=\frac{T^{2}}{6}\delta_{m=n}+O_{\epsilon}\left(T^{1+\epsilon}(mn)^{\epsilon}+(mn)^{1/2+\epsilon}\right)

where wjw_{j} is defined in Theorem 1.2, and the summation over jj is taken over all uju_{j}. Note that

jw(j)etj/TT26.\sum_{j}w(j)e^{-t_{j}/T}\sim\frac{T^{2}}{6}.
Proof.

See [1, Lemma 1]. ∎

Unlike holomorphic cusp forms, we do not have a suitable bound such as Deligne’s bound for our purposes. Consequently, we cannot apply Lemma 2.2. Nevertheless, since we are dealing with families of Hecke-Maass forms, we can employ the Kuznetsov trace formula instead of Lemma 2.2.

Lemma 2.5.

For any d0,δ1,δ2,d_{0},\delta_{1},\delta_{2}\in\mathbb{N}, we have

jw(j)etj/T|λuj((d0δ1)2)λuj((d0δ2)2)|ϵT2(d02δ1δ2)ϵ+(d02δ1δ2)1/2+ϵ.\sum_{j}w(j)e^{-t_{j}/T}\left|\lambda_{u_{j}}\left(\left(d_{0}\delta_{1}\right)^{2}\right)\lambda_{u_{j}}\left(\left(d_{0}\delta_{2}\right)^{2}\right)\right|\ll_{\epsilon}T^{2}\left(d_{0}^{2}\delta_{1}\delta_{2}\right)^{\epsilon}+\left(d_{0}^{2}\delta_{1}\delta_{2}\right)^{1/2+\epsilon}.
Proof.

Using (1.6), we obtain

λuj((d0δi)2)=r|d0δiμ(ri)λuj2(d0δi/ri)=r|d0δiμ(ri)ki|d0δiriλuj(ki2)\begin{split}\lambda_{u_{j}}\left(\left(d_{0}\delta_{i}\right)^{2}\right)&=\sum_{r|d_{0}\delta_{i}}\mu(r_{i})\lambda_{u_{j}}^{2}\left(d_{0}\delta_{i}/r_{i}\right)\\ &=\sum_{r|d_{0}\delta_{i}}\mu(r_{i})\sum_{k_{i}|\frac{d_{0}\delta_{i}}{r_{i}}}\lambda_{u_{j}}(k_{i}^{2})\end{split}

for i=1,2.i=1,2. Note that the term

ki|d0δiriλuj(ki2)\sum_{k_{i}|\frac{d_{0}\delta_{i}}{r_{i}}}\lambda_{u_{j}}(k_{i}^{2})

is positive. Therefore, by the triangle inequality, we see that

|λuj((d0δ1)2)λuj((d0δ2)2)|r1|d0δ1|μ(r1)|r2|d0δ2|μ(r2)|k1|d0δ1r1k2|d0δ2r2λuj(k12)λuj(k22).\left|\lambda_{u_{j}}\left(\left(d_{0}\delta_{1}\right)^{2}\right)\lambda_{u_{j}}\left(\left(d_{0}\delta_{2}\right)^{2}\right)\right|\leq\sum_{r_{1}|d_{0}\delta_{1}}|\mu(r_{1})|\sum_{r_{2}|d_{0}\delta_{2}}|\mu(r_{2})|\sum_{k_{1}|\frac{d_{0}\delta_{1}}{r_{1}}}\sum_{k_{2}|\frac{d_{0}\delta_{2}}{r_{2}}}\lambda_{u_{j}}\left(k_{1}^{2}\right)\lambda_{u_{j}}\left(k_{2}^{2}\right).

By Lemma 2.4, we have

(2.5) jw(j)etj/Tr1|d0δ1|μ(r1)|r2|d0δ2|μ(r2)|k1|d0δ1r1k2|d0δ2r2λuj(k12)λuj(k22)ϵT2r1|d0δ1|μ(r1)|r2|d0δ2|μ(r2)|(k1|d0δ1r1k2|d0δ2r2k1=k21+k1|d0δ1r1k2|d0δ2r2((k1k2)ϵT1ϵ+(k1k2)1/2+ϵT2))ϵT2r1|d0δ1|μ(r1)|r2|d0δ2|μ(r2)|(τ(d0δ1r1)τ(d0δ2r2)(1+(d02δ1δ2)ϵT1ϵ+(d02δ1δ2)1/2+ϵT2)).\begin{split}&\sum_{j}w(j)e^{-t_{j}/T}\sum_{r_{1}|d_{0}\delta_{1}}|\mu(r_{1})|\sum_{r_{2}|d_{0}\delta_{2}}|\mu(r_{2})|\sum_{k_{1}|\frac{d_{0}\delta_{1}}{r_{1}}}\sum_{k_{2}|\frac{d_{0}\delta_{2}}{r_{2}}}\lambda_{u_{j}}\left(k_{1}^{2}\right)\lambda_{u_{j}}\left(k_{2}^{2}\right)\\ &\ll_{\epsilon}T^{2}\sum_{r_{1}|d_{0}\delta_{1}}|\mu(r_{1})|\sum_{r_{2}|d_{0}\delta_{2}}|\mu(r_{2})|\left(\mathop{\sum_{k_{1}|\frac{d_{0}\delta_{1}}{r_{1}}}\sum_{k_{2}|\frac{d_{0}\delta_{2}}{r_{2}}}}_{k_{1}=k_{2}}1+\sum_{k_{1}|\frac{d_{0}\delta_{1}}{r_{1}}}\sum_{\begin{subarray}{c}k_{2}|\frac{d_{0}\delta_{2}}{r_{2}}\end{subarray}}\left(\frac{(k_{1}k_{2})^{\epsilon}}{T^{1-\epsilon}}+\frac{(k_{1}k_{2})^{1/2+\epsilon}}{T^{2}}\right)\right)\\ &\ll_{\epsilon}T^{2}\sum_{r_{1}|d_{0}\delta_{1}}|\mu(r_{1})|\sum_{r_{2}|d_{0}\delta_{2}}|\mu(r_{2})|\left(\tau\left(\frac{d_{0}\delta_{1}}{r_{1}}\right)\tau\left(\frac{d_{0}\delta_{2}}{r_{2}}\right)\left(1+\frac{\left(d_{0}^{2}\delta_{1}\delta_{2}\right)^{\epsilon}}{T^{1-\epsilon}}+\frac{\left(d_{0}^{2}\delta_{1}\delta_{2}\right)^{1/2+\epsilon}}{T^{2}}\right)\right).\end{split}

Using the divisor bound τ(n)ϵnϵ,\tau(n)\ll_{\epsilon}n^{\epsilon}, we can bound the above expression as follows:

ϵT2(d02δ1δ2)ϵ+(d02δ1δ2)1/2+ϵ.\ll_{\epsilon}T^{2}\left(d_{0}^{2}\delta_{1}\delta_{2}\right)^{\epsilon}+\left(d_{0}^{2}\delta_{1}\delta_{2}\right)^{1/2+\epsilon}.

3. Propositions and the proof of Theorem 1.1

Let zz represent H1+ϵH^{1+\epsilon} from now on. We begin by splitting the integral in (1.5) into three parts:

(3.1) X2X(x<ndx+Hdzλf(d2)Hdzλf(d2)d)2𝑑x,\int_{X}^{2X}\left(\sum_{\begin{subarray}{c}x<nd\leq x+H\\ d\leq z\end{subarray}}\lambda_{f}(d^{2})-H\sum_{d\leq z}\frac{\lambda_{f}\left(d^{2}\right)}{d}\right)^{2}dx,
(3.2) X2X(x<ndx+Hz<dλf(d2))2𝑑x,\int_{X}^{2X}\left(\sum_{\begin{subarray}{c}x<nd\leq x+H\\ z<d\end{subarray}}\lambda_{f}(d^{2})\right)^{2}dx,

and

(3.3) X2X(Hc1,fHdzλf(d2)d)2𝑑x.\int_{X}^{2X}\left(Hc_{1,f}-H\sum_{d\leq z}\frac{\lambda_{f}\left(d^{2}\right)}{d}\right)^{2}dx.

Using the Cauchy-Schwarz inequality, we obtain an upper bound for the integral in (1.5):

1XfSkφf((3.1)+(3.2)+(3.3)).\frac{1}{X}\sum_{f\in S_{k}}\varphi_{f}\left(\eqref{1}+\eqref{2}+\eqref{3}\right).

Hence, we need to demonstrate that each of the three integrals is bounded by kϵH1+ϵk^{\epsilon}H^{1+\epsilon} on average over fSk.f\in S_{k}. We analyze the contribution of equation (3.1) in Propositions 3.1 and 3.2 by utilizing the Petersson trace formula alongside the proof presented in [5]. Although it is possible to obtain Proposition 3.2 for individual ff, the range of HH is restricted to logH<(logX/4)\log H<(\log X/4). In Proposition 3.3, we examine the contribution of equation (3.2) by applying the Petersson trace formula. In Proposition 3.5, we obtain a pointwise bound for

Hc1,fHdzλf(d2)dHc_{1,f}-H\sum_{d\leq z}\frac{\lambda_{f}\left(d^{2}\right)}{d}

using the Lindelöf-on-average bound to handle the contribution of equation (3.3).

3.1. Propositions

The following proposition employs the same method as used in Proposition 5 of [5]. Like Proposition 5, we anticipate that the integral (3.1) will provide the dominant term.

Proposition 3.1.

Let 1logH<logX.1\ll\log H<\log X. As XX\rightarrow\infty,

fSkφf1XX2X(x<ndx+Hdzλf(d2)Hdzλf(d2)d)2𝑑x=(1π2+O(Hϵ/2))dz0<nX51n2sin2(πnHd)+Oϵ(z3+2ϵ(logX)2k12)+O((logX)63).\begin{split}\sum_{f\in S_{k}}&\varphi_{f}\frac{1}{X}\int_{X}^{2X}\left(\sum_{\begin{subarray}{c}x<nd\leq x+H\\ d\leq z\end{subarray}}\lambda_{f}(d^{2})-H\sum_{d\leq z}\frac{\lambda_{f}\left(d^{2}\right)}{d}\right)^{2}dx\\ &=\left(\frac{1}{\pi^{2}}+O\left(H^{-\epsilon/2}\right)\right)\sum_{d\leq z}\sum_{0<n\leq X^{5}}\frac{1}{n^{2}}\sin^{2}\left(\frac{\pi nH}{d}\right)+O_{\epsilon}\left(z^{3+2\epsilon}(\log X)^{2}k^{-\frac{1}{2}}\right)+O\left((\log X)^{63}\right).\end{split}
Proof.

Without loss of generality, we assume that H.H\in\mathbb{N}. We have

(3.4) xndx+Hdzλf(d2)=1dx+Hdzλf(d2)xd<nx+Hd1=1dx+Hdzλf(d2)(Hd+ψ(xd)ψ(x+Hd))\begin{split}\sum_{\begin{subarray}{c}x\leq nd\leq x+H\\ d\leq z\end{subarray}}\lambda_{f}\left(d^{2}\right)&=\sum_{\begin{subarray}{c}1\leq d\leq x+H\\ d\leq z\end{subarray}}\lambda_{f}\left(d^{2}\right)\sum_{\frac{x}{d}<n\leq\frac{x+H}{d}}1\\ &=\sum_{\begin{subarray}{c}1\leq d\leq x+H\\ d\leq z\end{subarray}}\lambda_{f}\left(d^{2}\right)\left(\frac{H}{d}+\psi\left(\frac{x}{d}\right)-\psi\left(\frac{x+H}{d}\right)\right)\end{split}

where ψ(x)=x[x]1/2.\psi(x)=x-[x]-1/2. It is known that for any NN\in\mathbb{N} and xx\in\mathbb{R}, we have

ψ(x)=12πi0<|n|<N1ne(nx)+O(min(1,1Nx))\psi(x)=-\frac{1}{2\pi i}\sum_{0<|n|<N}\frac{1}{n}e(nx)+O\left(\min\left(1,\frac{1}{N\|x\|}\right)\right)

(see [11][(4.18)]). We will choose the value of NN later. Therefore, we see that

(3.5) x<ndx+Hdzλf(d2)Hdzλf(d2)d=12πi1dx+Hdzλf(d2)(0<|n|N1n(e(nxd)e(n(x+H)d)))+O((x,z)),\begin{split}&\sum_{\begin{subarray}{c}x<nd\leq x+H\\ d\leq z\end{subarray}}\lambda_{f}(d^{2})-H\sum_{d\leq z}\frac{\lambda_{f}\left(d^{2}\right)}{d}\\ &=-\frac{1}{2\pi i}\sum_{\begin{subarray}{c}1\leq d\leq x+H\\ d\leq z\end{subarray}}\lambda_{f}\left(d^{2}\right)\left(\sum_{0<|n|\leq N}\frac{1}{n}\left(e\left(\frac{nx}{d}\right)-e\left(\frac{n\left(x+H\right)}{d}\right)\right)\right)+O(\mathcal{M}(x,z)),\end{split}

where

(3.6) (x,z)=1dx+Hdz|λf(d2)|(min(1,1Nx+Hd)+min(1,1Nxd)).\begin{split}\mathcal{M}(x,z)&=\sum_{\begin{subarray}{c}1\leq d\leq x+H\\ d\leq z\end{subarray}}\left|\lambda_{f}\left(d^{2}\right)\right|\left(\min\left(1,\frac{1}{N\|\frac{x+H}{d}\|}\right)+\min\left(1,\frac{1}{N\|\frac{x}{d}\|}\right)\right).\end{split}

Using the Cauchy-Schwarz inequality, we have

(3.7) 1XXx2X(1dz|λf(d2)|min(1,1Nx+Hd))21XXx2X(dx+Hdz|λf(d2)|+dx+Hdzd|λf(d2)|N)21XXx2X((dx+Hdz|λf(d2)|)2+(dx+Hdzd|λf(d2)|N)2).\begin{split}\frac{1}{X}&\sum_{X\leq x\leq 2X}\left(\sum_{1\leq d\leq z}\left|\lambda_{f}\left(d^{2}\right)\right|\min\left(1,\frac{1}{N\|\frac{x+H}{d}\|}\right)\right)^{2}\\ &\ll\frac{1}{X}\sum_{X\leq x\leq 2X}\left(\sum_{\begin{subarray}{c}d\mid x+H\\ d\leq z\end{subarray}}\left|\lambda_{f}\left(d^{2}\right)\right|+\sum_{\begin{subarray}{c}d\nmid x+H\\ d\leq z\end{subarray}}\frac{d\left|\lambda_{f}\left(d^{2}\right)\right|}{N}\right)^{2}\\ &\ll\frac{1}{X}\sum_{X\leq x\leq 2X}\left(\left(\sum_{\begin{subarray}{c}d\mid x+H\\ d\leq z\end{subarray}}\left|\lambda_{f}\left(d^{2}\right)\right|\right)^{2}+\left(\sum_{\begin{subarray}{c}d\nmid x+H\\ d\leq z\end{subarray}}\frac{d\left|\lambda_{f}\left(d^{2}\right)\right|}{N}\right)^{2}\right).\end{split}

By Deligne’s bound and using (2.1), we estimate the first term in the last inequality of (3.7) as

(3.8) 1XXx2Xτ(x+H)6(logX)63.\begin{split}&\ll\frac{1}{X}\sum_{X\leq x\leq 2X}\tau(x+H)^{6}\\ &\ll(\log X)^{63}.\end{split}

Similarly, we estimate the second term as

(3.9) 1XXx2X(dzdτ(d)2N)2(z2(logz)3N)2.\begin{split}&\ll\frac{1}{X}\sum_{X\leq x\leq 2X}\left(\sum_{d\leq z}\frac{d\tau(d)^{2}}{N}\right)^{2}\\ &\ll\left(\frac{z^{2}\left(\log z\right)^{3}}{N}\right)^{2}.\end{split}

Using the same approach as in (3.7), we have

1XXx2X(1dz|λf(d2)|min(1,1Nxd))2(logX)63+(z2(logz)3N)2.\frac{1}{X}\sum_{X\leq x\leq 2X}\left(\sum_{1\leq d\leq z}\left|\lambda_{f}\left(d^{2}\right)\right|\min\left(1,\frac{1}{N\|\frac{x}{d}\|}\right)\right)^{2}\ll(\log X)^{63}+\left(\frac{z^{2}\left(\log z\right)^{3}}{N}\right)^{2}.

In order to proceed, we choose N=X5.N=X^{5}. Then, we obtain

1X1[1,2](xX)|(x,z)|2𝑑x(logX)63.\frac{1}{X}\int 1_{[1,2]}\left(\frac{x}{X}\right)\left|\mathcal{M}(x,z)\right|^{2}dx\ll(\log X)^{63}.

Next, we consider the contribution of the main term in (3.5), which is given by

(3.10) 14π2X1[1,2](xX)|dzλf(d2)0<|n|<X51ne(nxd)(1e(nHd))|2𝑑x.\frac{1}{4\pi^{2}X}\int_{-\infty}^{\infty}1_{[1,2]}\left(\frac{x}{X}\right)\left|\sum_{d\leq z}\lambda_{f}\left(d^{2}\right)\sum_{0<|n|<X^{5}}\frac{1}{n}e\left(\frac{nx}{d}\right)\left(1-e\left(\frac{nH}{d}\right)\right)\right|^{2}dx.

Since σ1[1,2]σ+,\sigma_{-}\leq 1_{[1,2]}\leq\sigma_{+}, (3.10) lies between

(3.11) 14π2d1zd2zλf(d12)λf(d22)0<|n1|<X51n10<|n2|<X51n2×(1e(n1Hd1))(1e(n2Hd2))¯σ^(X(n1d1n2d2))\begin{split}\frac{1}{4\pi^{2}}&\sum_{d_{1}\leq z}\sum_{d_{2}\leq z}\lambda_{f}\left(d_{1}^{2}\right)\lambda_{f}\left(d_{2}^{2}\right)\sum_{0<|n_{1}|<X^{5}}\frac{1}{n_{1}}\sum_{0<|n_{2}|<X^{5}}\frac{1}{n_{2}}\\ &\times\left(1-e\left(\frac{n_{1}H}{d_{1}}\right)\right)\overline{\left(1-e\left(\frac{n_{2}H}{d_{2}}\right)\right)}\hat{\sigma}_{-}\left(-X\left(\frac{n_{1}}{d_{1}}-\frac{n_{2}}{d_{2}}\right)\right)\end{split}

and

(3.12) 14π2d1zd2zλf(d12)λf(d22)0<|n1|<X51n10<|n2|<X51n2×(1e(n1Hd1))(1e(n2Hd2))¯σ^+(X(n1d1n2d2)).\begin{split}\frac{1}{4\pi^{2}}&\sum_{d_{1}\leq z}\sum_{d_{2}\leq z}\lambda_{f}\left(d_{1}^{2}\right)\lambda_{f}\left(d_{2}^{2}\right)\sum_{0<|n_{1}|<X^{5}}\frac{1}{n_{1}}\sum_{0<|n_{2}|<X^{5}}\frac{1}{n_{2}}\\ &\times\left(1-e\left(\frac{n_{1}H}{d_{1}}\right)\right)\overline{\left(1-e\left(\frac{n_{2}H}{d_{2}}\right)\right)}\hat{\sigma}_{+}\left(-X\left(\frac{n_{1}}{d_{1}}-\frac{n_{2}}{d_{2}}\right)\right).\end{split}

By applying the Petersson trace formula, we have that

(3.13) fSkφfd1zd2zλf(d12)λf(d22)0<|n1|<X51n10<|n2|<X51n2×(1e(n1Hd1))(1e(n2Hd2))¯σ^±(X(n1d1n2d2))=dz10<|n1|<X51n10<|n2|<X51n2(1e(n1Hd))(1e(n2Hd))¯σ^±(X(n1n2d))+Oϵ(d1zd2z(d1d2)12+ϵ0<|n1|<X51|n1|0<|n2|<X51|n2||σ^±(X(n1d1n2d2))|k12).\begin{split}&\sum_{f\in S_{k}}\varphi_{f}\sum_{d_{1}\leq z}\sum_{d_{2}\leq z}\lambda_{f}\left(d_{1}^{2}\right)\lambda_{f}\left(d_{2}^{2}\right)\sum_{0<|n_{1}|<X^{5}}\frac{1}{n_{1}}\sum_{0<|n_{2}|<X^{5}}\frac{1}{n_{2}}\\ &\;\;\;\;\;\;\times\left(1-e\left(\frac{n_{1}H}{d_{1}}\right)\right)\overline{\left(1-e\left(\frac{n_{2}H}{d_{2}}\right)\right)}\hat{\sigma}_{\pm}\left(-X\left(\frac{n_{1}}{d_{1}}-\frac{n_{2}}{d_{2}}\right)\right)\\ &=\sum_{d\leq z}1\sum_{0<|n_{1}|<X^{5}}\frac{1}{n_{1}}\sum_{0<|n_{2}|<X^{5}}\frac{1}{n_{2}}\left(1-e\left(\frac{n_{1}H}{d}\right)\right)\overline{\left(1-e\left(\frac{n_{2}H}{d}\right)\right)}\hat{\sigma}_{\pm}\left(-X\left(\frac{n_{1}-n_{2}}{d}\right)\right)\\ &\;\;+O_{\epsilon}\left(\sum_{d_{1}\leq z}\sum_{d_{2}\leq z}\left(d_{1}d_{2}\right)^{\frac{1}{2}+\epsilon}\sum_{0<|n_{1}|<X^{5}}\frac{1}{|n_{1}|}\sum_{0<|n_{2}|<X^{5}}\frac{1}{|n_{2}|}\left|\hat{\sigma}_{\pm}\left(-X\left(\frac{n_{1}}{d_{1}}-\frac{n_{2}}{d_{2}}\right)\right)\right|k^{-\frac{1}{2}}\right).\end{split}

By applying the condition on the supports of σ^\hat{\sigma}_{-} and σ^+\hat{\sigma}_{+}, we see that

(3.14) σ^±(X(n1n2d))0|n1n2d|BXHϵ2|n1n2|BHϵ2dX.\begin{split}\hat{\sigma}_{\pm}\left(-X\left(\frac{n_{1}-n_{2}}{d}\right)\right)\neq 0&\Rightarrow\left|\frac{n_{1}-n_{2}}{d}\right|\leq\frac{B}{X}H^{\frac{\epsilon}{2}}\\ &\Rightarrow\left|n_{1}-n_{2}\right|\leq\frac{BH^{\frac{\epsilon}{2}}d}{X}.\end{split}

Assuming that ϵ\epsilon is sufficiently small and XX is sufficiently large so that d<zϵX1ϵd<z\ll_{\epsilon}X^{1-\epsilon}, we can imply the following:

σ^±(X(n1n2d))0n1=n2.\hat{\sigma}_{\pm}\left(-X\left(\frac{n_{1}-n_{2}}{d}\right)\right)\neq 0\Rightarrow n_{1}=n_{2}.

Therefore, equation (3.13) can be written as

(3.15) 4σ^±(0)dz10<|n|X51n2sin2(πnHd)+Oϵ(z3+2ϵ(logX)2k12).4\hat{\sigma}_{\pm}(0)\sum_{d\leq z}1\sum_{0<|n|\leq X^{5}}\frac{1}{n^{2}}\sin^{2}\left(\frac{\pi nH}{d}\right)+O_{\epsilon}\left(z^{3+2\epsilon}(\log X)^{2}k^{-\frac{1}{2}}\right).

Note that the main term is non-negative. By applying the third property of σ±\sigma_{\pm} in Lemma 2.1, the above term can be written as

(3.16) 4(1+O(Hϵ/2))dz10<nX51n2sin2(πnHd)+Oϵ(z3+2ϵ(logX)2k12)4\left(1+O\left(H^{-\epsilon/2}\right)\right)\sum_{d\leq z}1\sum_{0<n\leq X^{5}}\frac{1}{n^{2}}\sin^{2}\left(\frac{\pi nH}{d}\right)+O_{\epsilon}\left(z^{3+2\epsilon}(\log X)^{2}k^{-\frac{1}{2}}\right)

The following proposition employs a similar method as in [5, Lemma 9].

Proposition 3.2.

Assume the condition in Proposition 3.1. As XX\rightarrow\infty, we have

(3.17) 1π2dz10<nX51n2sin2(πnHd)ϵHzϵ\begin{split}\frac{1}{\pi^{2}}\sum_{d\leq z}1\sum_{0<n\leq X^{5}}\frac{1}{n^{2}}\sin^{2}\left(\frac{\pi nH}{d}\right)\ll_{\epsilon}Hz^{\epsilon}\end{split}
Proof.

Define the function W(y)W(y) by

W(y)=sin(πy)πyh(y/Hε/4),W(y)=\frac{\sin(\pi y)}{\pi y}h\left(y/H^{\varepsilon/4}\right),

where hh is a smooth bump function such that h(x)=1h(x)=1 for |x|1|x|\leq 1 and h(x)=0h(x)=0 for |x|2|x|\geq 2. Then, we have

(3.18) H2dz1d20<nX5(sin2(πnHd)(πnHd)2|W(nHd)|2)\displaystyle H^{2}\sum_{d\leq z}\frac{1}{d^{2}}\sum_{0<n\leq X^{5}}\left(\frac{\sin^{2}\left(\frac{\pi nH}{d}\right)}{\left(\frac{\pi nH}{d}\right)^{2}}-\left|W\left(\frac{nH}{d}\right)\right|^{2}\right)
dz10<nX5δnHdHϵ/4n2\displaystyle\ll\sum_{d\leq z}1\sum_{0<n\leq X^{5}}\frac{\delta_{\frac{nH}{d}\geq H^{\epsilon/4}}}{n^{2}}
dH1ϵ/41+H1ϵ/4<dzH1ϵ/4d\displaystyle\ll\sum_{d\leq H^{1-\epsilon/4}}1+\sum_{H^{1-\epsilon/4}<d\leq z}\frac{H^{1-\epsilon/4}}{d}
H1ϵ/4logz.\displaystyle\ll H^{1-\epsilon/4}\log z.

Let us consider

H2dz1d20<nX5|W(nHd)|2.H^{2}\sum_{d\leq z}\frac{1}{d^{2}}\sum_{0<n\leq X^{5}}\left|W\left(\frac{nH}{d}\right)\right|^{2}.

Let g(x)=|W(ex)|2ex.g(x)=\left|W(e^{x})\right|^{2}e^{x}. Then we have the following expression for the Fourier transform of g(x)g(x):

g^(x)=0|W(y)|2y2πix𝑑y,\hat{g}(x)=\int_{0}^{\infty}\left|W(y)\right|^{2}y^{-2\pi ix}dy,

where g^(ξ)\hat{g}(\xi) is entire and satisfies g^(ξ)=O(H2ε/(|ξ|+1)3)\hat{g}(\xi)=O\left(H^{2\varepsilon}/(|\xi|+1)^{3}\right) uniformly for |(ξ)|<1/2π|\Im(\xi)|<1/2\pi (see [5, (16)]). By the Mellin inversion, we get

(3.19) |W(r)|2=r112πi(c)rsg^(s2πi)𝑑s|W(r)|^{2}=r^{-1}\frac{1}{2\pi i}\int_{(c)}r^{s}\hat{g}\left(\frac{s}{2\pi i}\right)ds

where c(1,1).c\in(-1,1). By using (3.19), we have

(3.20) H2dz1d2n1|W(nHd)|2=H2πidz1dn11n(2ϵ)(nHd)sg^(s2πi)𝑑s=H2πidz1d(2ϵ)ζ(1s)(Hd)sg^(s2πi)𝑑s.\begin{split}&H^{2}\sum_{d\leq z}\frac{1}{d^{2}}\sum_{n\geq 1}\left|W\left(\frac{nH}{d}\right)\right|^{2}\\ &=\frac{H}{2\pi i}\sum_{d\leq z}\frac{1}{d}\sum_{n\geq 1}\frac{1}{n}\int_{(-2\epsilon)}\left(\frac{nH}{d}\right)^{s}\hat{g}\left(\frac{s}{2\pi i}\right)ds\\ &=\frac{H}{2\pi i}\sum_{d\leq z}\frac{1}{d}\int_{(-2\epsilon)}\zeta(1-s)\left(\frac{H}{d}\right)^{s}\hat{g}\left(\frac{s}{2\pi i}\right)ds.\end{split}

By using the fact that |ζ(1+2ϵ+it)|ζ(1+2ϵ)ϵ1for anyt,|\zeta(1+2\epsilon+it)|\leq\zeta(1+2\epsilon)\ll_{\epsilon}1\;\;\textrm{for any}\;\;t\in\mathbb{R}, we see that

(2ϵ)ζ(1s)(Hd)sg^(s2πi)𝑑sϵd2ϵ.\int_{(-2\epsilon)}\zeta(1-s)\left(\frac{H}{d}\right)^{s}\hat{g}\left(\frac{s}{2\pi i}\right)ds\ll_{\epsilon}d^{2\epsilon}.

Therefore, (3.20) is bounded by

(3.21) ϵHdz1dd2ϵϵHz2ϵ\begin{split}&\ll_{\epsilon}H\sum_{d\leq z}\frac{1}{d}d^{2\epsilon}\\ &\ll_{\epsilon}Hz^{2\epsilon}\end{split}

Remark 3.3.

In contrast to Lemma 9 in [5], we are unable to obtain an asymptotic estimate in this case because the exponent of dd in the second line of equation (3.20) is 11. However, we can anticipate that the order of HH is 11.

Proposition 3.4.

Assume the condition in Proposition 3.1. As XX\rightarrow\infty,

(3.22) fSkφf1XX2X(x<ndx+Hd>zλf(d2))2𝑑xϵ(Hlog2X+Hz+HX2+ϵk1/2).\begin{split}\sum_{f\in S_{k}}\varphi_{f}\frac{1}{X}&\int_{X}^{2X}\left(\sum_{\begin{subarray}{c}x<nd\leq x+H\\ d>z\end{subarray}}\lambda_{f}(d^{2})\right)^{2}dx\ll_{\epsilon}\left(H\log\frac{2X+H}{z}+H\frac{X^{2+\epsilon}}{k^{1/2}}\right).\end{split}
Proof.

Since z>H,z>H, for each x[X,2X],x\in[X,2X], we have

(3.23) (x<ndx+Hd>zλf(d2))2=(z<dx+Hλf(d2)δ[xd]+1=[x+Hd])2.\begin{split}\left(\sum_{\begin{subarray}{c}x<nd\leq x+H\\ d>z\end{subarray}}\lambda_{f}(d^{2})\right)^{2}=\left(\sum_{z<d\leq x+H}\lambda_{f}\left(d^{2}\right)\delta_{[\frac{x}{d}]+1=[\frac{x+H}{d}]}\right)^{2}.\end{split}

By squaring out the sum, we express the above quantity as

(3.24) z<d1x+Hz<d2x+Hλf(d12)λf(d22)δ[xd1]+1=[x+Hd1]δ[xd2]+1=[x+Hd2].\sum_{z<d_{1}\leq x+H}\sum_{z<d_{2}\leq x+H}\lambda_{f}\left(d_{1}^{2}\right)\lambda_{f}\left(d_{2}^{2}\right)\delta_{[\frac{x}{d_{1}}]+1=[\frac{x+H}{d_{1}}]}\delta_{[\frac{x}{d_{2}}]+1=[\frac{x+H}{d_{2}}]}.

By Lemma 2.3, we see that

(3.25) fSk1f2z<d1x+Hz<d2x+Hλf(d12)λf(d22)δ[xd1]+1=[x+Hd1]δ[xd2]+1=[x+Hd2]ϵfSk1f2(z<dx+Hδ[xd]+1=[x+Hd]+z<d1x+Hz<d2x+H(d1d2)1/2+ϵk1/2δ[xd1]+1=[x+Hd1]δ[xd2]+1=[x+Hd2]).\begin{split}&\sum_{f\in S_{k}}\frac{1}{\|f\|^{2}}\sum_{z<d_{1}\leq x+H}\sum_{z<d_{2}\leq x+H}\lambda_{f}\left(d_{1}^{2}\right)\lambda_{f}\left(d_{2}^{2}\right)\delta_{[\frac{x}{d_{1}}]+1=[\frac{x+H}{d_{1}}]}\delta_{[\frac{x}{d_{2}}]+1=[\frac{x+H}{d_{2}}]}\\ &\ll_{\epsilon}\sum_{f\in S_{k}}\frac{1}{\|f\|^{2}}\left(\sum_{z<d\leq x+H}\delta_{[\frac{x}{d}]+1=[\frac{x+H}{d}]}+\sum_{z<d_{1}\leq x+H}\sum_{z<d_{2}\leq x+H}\frac{\left(d_{1}d_{2}\right)^{1/2+\epsilon}}{k^{1/2}}\delta_{[\frac{x}{d_{1}}]+1=[\frac{x+H}{d_{1}}]}\delta_{[\frac{x}{d_{2}}]+1=[\frac{x+H}{d_{2}}]}\right).\end{split}

For each interval [a,a+d)[X,2X+H],[a,a+d)\subset[X,2X+H], there are HH terms b[a,a+d)b\in[a,a+d) such that

(3.26) δ[bd]+1=[b+Hd]=1,\delta_{[\frac{b}{d}]+1=[\frac{b+H}{d}]}=1,

and 0 for other terms. Therefore, by interchanging the order of summations, we have

(3.27) fSk1X1f2X2X(x<ndx+Hd>zλf(d2))2𝑑xϵ1XfSk1f2(z<d2X+HHXd+H+HX3+2ϵk1/2)ϵfSk1f2(Hlog2X+Hz+HX2+2ϵk1/2).\begin{split}&\sum_{f\in S_{k}}\frac{1}{X}\frac{1}{\|f\|^{2}}\int_{X}^{2X}\left(\sum_{\begin{subarray}{c}x<nd\leq x+H\\ d>z\end{subarray}}\lambda_{f}(d^{2})\right)^{2}dx\\ &\ll_{\epsilon}\frac{1}{X}\sum_{f\in S_{k}}\frac{1}{\|f\|^{2}}\left(\sum_{z<d\leq 2X+H}H\frac{X}{d}+H+H\frac{X^{3+2\epsilon}}{k^{1/2}}\right)\\ &\ll_{\epsilon}\sum_{f\in S_{k}}\frac{1}{\|f\|^{2}}\left(H\log\frac{2X+H}{z}+\frac{HX^{2+2\epsilon}}{k^{1/2}}\right).\end{split}

Proposition 3.5.

Assume the condition in Proposition 3.1 and the Lindelöf-on-average bound (1.4). Then

fSkφf1XX2X|Hc1,fHdzλf(d2)d|2𝑑x=Oϵ(H2kϵz1+ϵ).\sum_{f\in S_{k}}\varphi_{f}\frac{1}{X}\int_{X}^{2X}\left|Hc_{1,f}-H\sum_{d\leq z}\frac{\lambda_{f}\left(d^{2}\right)}{d}\right|^{2}dx=O_{\epsilon}\left(H^{2}k^{\epsilon}z^{-1+\epsilon}\right).
Proof.

Let T(1,X].T\in(1,X]. Using Perron’s formula (see [6, Lemma 1.1]) and Deligne’s bound, we get

(3.28) dzλf(d2)d=12πi1logziT1logz+iTL(sym2f,w+1)ζ(2w+2)zww𝑑w+Oϵ(zϵT+zϵ1).\sum_{d\leq z}\frac{\lambda_{f}\left(d^{2}\right)}{d}=\frac{1}{2\pi i}\int_{\frac{1}{\log z}-iT}^{\frac{1}{\log z}+iT}\frac{L\left(\textrm{sym}^{2}f,w+1\right)}{\zeta(2w+2)}\frac{z^{w}}{w}dw+O_{\epsilon}\left(\frac{z^{\epsilon}}{T}+z^{\epsilon-1}\right).

Applying Hölder’s inequality, the Phragmen-Lindelöf principle, and (1.4), we obtain

(3.29) fSkφf|1logz±iT12+ϵ±iTL(sym2f,w+1)ζ(2w+2)zww𝑑w|2fSkφf1logz±iT12+ϵ±iT|L(sym2f,w+1)ζ(2w+2)zww|2𝑑wϵkϵT2ϵ.\begin{split}&\sum_{f\in S_{k}}\varphi_{f}\left|\int_{\frac{1}{\log z}\pm iT}^{-\frac{1}{2}+\epsilon\pm iT}\frac{L\left(\textrm{sym}^{2}f,w+1\right)}{\zeta(2w+2)}\frac{z^{w}}{w}dw\right|^{2}\\ &\ll\sum_{f\in S_{k}}\varphi_{f}\int_{\frac{1}{\log z}\pm iT}^{-\frac{1}{2}+\epsilon\pm iT}\left|\frac{L\left(\textrm{sym}^{2}f,w+1\right)}{\zeta(2w+2)}\frac{z^{w}}{w}\right|^{2}dw\\ &\ll_{\epsilon}\frac{k^{\epsilon}}{T^{2-\epsilon}}.\end{split}

Therefore, by moving the line of the integral to the line Re(w)=1/2+ϵ,\textrm{Re}(w)=-1/2+\epsilon, we have

(3.30) fSkφf(dzλf(d2)dc1,f)2ϵfSkφf|1/2+ϵiT1/2+ϵ+iTL(sym2f,w+1)ζ(2w+2)zww𝑑w|2+z2ϵT2+z2ϵ2+kϵT2ϵ.\begin{split}\sum_{f\in S_{k}}\varphi_{f}\left(\sum_{d\leq z}\frac{\lambda_{f}\left(d^{2}\right)}{d}-c_{1,f}\right)^{2}\ll_{\epsilon}&\sum_{f\in S_{k}}\varphi_{f}\left|\int_{-1/2+\epsilon-iT}^{-1/2+\epsilon+iT}\frac{L\left(\textrm{sym}^{2}f,w+1\right)}{\zeta(2w+2)}\frac{z^{w}}{w}dw\right|^{2}\\ &\;\;\;+\frac{z^{2\epsilon}}{T^{2}}+z^{2\epsilon-2}+\frac{k^{\epsilon}}{T^{2-\epsilon}}.\end{split}

By applying Hölder’s inequality and (1.4), we obtain

(3.31) fSkφf|1/2+ϵiT1/2+ϵ+iTL(sym2f,w+1)ζ(2w+2)zww𝑑w|2fSkφf1/2+ϵiT1/2+ϵ+iT|L(sym2f,w+1)ζ(2w+2)zww12ϵ|2𝑑w1/2+ϵiT1/2+ϵ+iT|1w12+ϵ|2𝑑wϵz1+2ϵkϵTϵ.\begin{split}\sum_{f\in S_{k}}&\varphi_{f}\left|\int_{-1/2+\epsilon-iT}^{-1/2+\epsilon+iT}\frac{L\left(\textrm{sym}^{2}f,w+1\right)}{\zeta(2w+2)}\frac{z^{w}}{w}dw\right|^{2}\\ &\ll\sum_{f\in S_{k}}\varphi_{f}\int_{-1/2+\epsilon-iT}^{-1/2+\epsilon+iT}\left|\frac{L\left(\textrm{sym}^{2}f,w+1\right)}{\zeta(2w+2)}\frac{z^{w}}{w^{\frac{1}{2}-\epsilon}}\right|^{2}dw\int_{-1/2+\epsilon-iT}^{-1/2+\epsilon+iT}\left|\frac{1}{w^{\frac{1}{2}+\epsilon}}\right|^{2}dw\\ &\ll_{\epsilon}z^{-1+2\epsilon}k^{\epsilon}T^{\epsilon}.\end{split}

By choosing T=z1ϵT=z^{1-\epsilon} and modifying ϵ,\epsilon, the proof is completed. ∎

3.2. Proof of Theorem 1.1

By Proposition 3.1 and 3.2, the contribution from (3.1) is bounded by

ϵfSkφf(H1+ϵzϵ+z3+2ϵk1/2+(logX)63).\ll_{\epsilon}\sum_{f\in S_{k}}\varphi_{f}\left(H^{1+\epsilon}z^{\epsilon}+z^{3+2\epsilon}k^{-1/2}+(\log X)^{63}\right).

By Proposition 3.3 and the condition logX<(logk)/4,\log X<(\log k)/4, the contribution from (3.2) is bounded by

ϵkϵHlogX.\ll_{\epsilon}k^{\epsilon}H\log X.

By Proposition 3.4, the contribution from (3.3) is bounded by

ϵkϵH1+ϵ.\ll_{\epsilon}k^{\epsilon}H^{1+\epsilon}.

Since logX<(logk)/4,\log X<(\log k)/4, these are bounded by

ϵkϵH1+ϵ.\ll_{\epsilon}k^{\epsilon}H^{1+\epsilon}.

4. Propositions and the proof of Theorem 1.2

The proof of Theorem 1.2 is similar to that of Theorem 1.1, but with some slight modifications. In particular, Propositions 4.1 and 4.2 use the Kuznetsov trace formula instead of the SL(2,)SL(2,\mathbb{Z}) Ramanujan-Petersson conjecture (|λuj(n)|τ(n)|\lambda_{u_{j}}(n)|\leq\tau(n)).

4.1. Propositions

Proposition 4.1.

Let 1logH<logX,1\ll\log H<\log X, and let 2logX(logH)/2<logT.2\log X-(\log H)/2<\log T. As XX\rightarrow\infty,

jϖjetj/T1XX2X(x<ndx+Hdzλuj(d2)Hdzλuj(d2)d)2𝑑x=(1π2+O(Hϵ/2))dz0<nX51n2sin2(πnHd)+Oϵ(z2+4ϵ(logX)2T1+z4+4ϵ(logX)2T2+XϵTϵ)\begin{split}\sum_{j}&\varpi_{j}e^{-t_{j}/T}\frac{1}{X}\int_{X}^{2X}\left(\sum_{\begin{subarray}{c}x<nd\leq x+H\\ d\leq z\end{subarray}}\lambda_{u_{j}}(d^{2})-H\sum_{d\leq z}\frac{\lambda_{u_{j}}\left(d^{2}\right)}{d}\right)^{2}dx\\ &=\left(\frac{1}{\pi^{2}}+O\left(H^{-\epsilon/2}\right)\right)\sum_{d\leq z}\sum_{0<n\leq X^{5}}\frac{1}{n^{2}}\sin^{2}\left(\frac{\pi nH}{d}\right)\\ &+\;\;\;O_{\epsilon}\left(z^{2+4\epsilon}(\log X)^{2}T^{-1}+z^{4+4\epsilon}(\log X)^{2}T^{-2}+X^{\epsilon}T^{\epsilon}\right)\end{split}
Proof.

Without loss of generality, we assume that H.H\in\mathbb{N}. Using the same argument as in the proof of Proposition 3.1, we have

(4.1) x<ndx+Hdzλuj(d2)Hdzλuj(d2)d=12πi1dx+Hdzλuj(d2)(0<|n|N1n(e(nxd)e(n(x+H)d)))+O((x,z)),\begin{split}&\sum_{\begin{subarray}{c}x<nd\leq x+H\\ d\leq z\end{subarray}}\lambda_{u_{j}}(d^{2})-H\sum_{d\leq z}\frac{\lambda_{u_{j}}\left(d^{2}\right)}{d}\\ &=-\frac{1}{2\pi i}\sum_{\begin{subarray}{c}1\leq d\leq x+H\\ d\leq z\end{subarray}}\lambda_{u_{j}}\left(d^{2}\right)\left(\sum_{0<|n|\leq N}\frac{1}{n}\left(e\left(\frac{nx}{d}\right)-e\left(\frac{n\left(x+H\right)}{d}\right)\right)\right)+O(\mathcal{M^{\prime}}(x,z)),\end{split}

where

(4.2) (x,z)=1dx+Hdz|λuj(d2)|(min(1,1Nx+Hd)+min(1,1Nxd)).\begin{split}\mathcal{M^{\prime}}(x,z)&=\sum_{\begin{subarray}{c}1\leq d\leq x+H\\ d\leq z\end{subarray}}\left|\lambda_{u_{j}}\left(d^{2}\right)\right|\left(\min\left(1,\frac{1}{N\|\frac{x+H}{d}\|}\right)+\min\left(1,\frac{1}{N\|\frac{x}{d}\|}\right)\right).\end{split}

Using the Cauchy-Schwarz inequality, we see that

(4.3) 1XXx2X(1dz|λuj(d2)|min(1,1Nx+Hd))21XXx2X(dx+Hdz|λuj(d2)|+dx+Hdzd|λuj(d2)|N)21XXx2X((dx+Hdz|λuj(d2)|)2+(dx+Hdzd|λuj(d2)|N)2).\begin{split}\frac{1}{X}&\sum_{X\leq x\leq 2X}\left(\sum_{1\leq d\leq z}\left|\lambda_{u_{j}}\left(d^{2}\right)\right|\min\left(1,\frac{1}{N\|\frac{x+H}{d}\|}\right)\right)^{2}\\ &\ll\frac{1}{X}\sum_{X\leq x\leq 2X}\left(\sum_{\begin{subarray}{c}d\mid x+H\\ d\leq z\end{subarray}}\left|\lambda_{u_{j}}\left(d^{2}\right)\right|+\sum_{\begin{subarray}{c}d\nmid x+H\\ d\leq z\end{subarray}}\frac{d\left|\lambda_{u_{j}}\left(d^{2}\right)\right|}{N}\right)^{2}\\ &\ll\frac{1}{X}\sum_{X\leq x\leq 2X}\left(\left(\sum_{\begin{subarray}{c}d\mid x+H\\ d\leq z\end{subarray}}\left|\lambda_{u_{j}}\left(d^{2}\right)\right|\right)^{2}+\left(\sum_{\begin{subarray}{c}d\nmid x+H\\ d\leq z\end{subarray}}\frac{d\left|\lambda_{u_{j}}\left(d^{2}\right)\right|}{N}\right)^{2}\right).\end{split}

By squaring out the first term in the last inequality in (4.3) and applying Lemma 2.5 with the divisor bound τ(n)ϵnϵ,\tau(n)\ll_{\epsilon}n^{\epsilon}, we get

(4.4) jw(j)etj/T1XXx2X(dx+Hdz|λuj(d2)|)2jw(j)etj/T1XXx2Xd1x+Hd1zd2x+Hd2z|λuj(d12)λuj(d22)|ϵT2Xϵzϵ+Xϵz1+ϵ.\begin{split}\sum_{j}&w(j)e^{-t_{j}/T}\frac{1}{X}\sum_{X\leq x\leq 2X}\left(\sum_{\begin{subarray}{c}d\mid x+H\\ d\leq z\end{subarray}}\left|\lambda_{u_{j}}\left(d^{2}\right)\right|\right)^{2}\\ &\ll\sum_{j}w(j)e^{-t_{j}/T}\frac{1}{X}\sum_{X\leq x\leq 2X}\sum_{\begin{subarray}{c}d_{1}\mid x+H\\ d_{1}\leq z\end{subarray}}\sum_{\begin{subarray}{c}d_{2}\mid x+H\\ d_{2}\leq z\end{subarray}}\left|\lambda_{u_{j}}\left(d_{1}^{2}\right)\lambda_{u_{j}}\left(d_{2}^{2}\right)\right|\\ &\ll_{\epsilon}T^{2}X^{\epsilon}z^{\epsilon}+X^{\epsilon}z^{1+\epsilon}.\end{split}

Similarly, the contribution from the second term in the last inequality in (4.3) is

(4.5) jw(j)etj/T1XXx2X(dx+Hdzd|λuj(d2)|N)2jw(j)etj/T1XN2Xx2Xd1zd2zd1d2|λuj(d12)λuj(d22)|ϵT2Xϵz4+ϵ+Xϵz5+ϵN2.\begin{split}\sum_{j}&w(j)e^{-t_{j}/T}\frac{1}{X}\sum_{X\leq x\leq 2X}\left(\sum_{\begin{subarray}{c}d\nmid x+H\\ d\leq z\end{subarray}}\frac{d\left|\lambda_{u_{j}}\left(d^{2}\right)\right|}{N}\right)^{2}\\ &\ll\sum_{j}w(j)e^{-t_{j}/T}\frac{1}{XN^{2}}\sum_{X\leq x\leq 2X}\sum_{d_{1}\leq z}\sum_{d_{2}\leq z}d_{1}d_{2}\left|\lambda_{u_{j}}\left(d_{1}^{2}\right)\lambda_{u_{j}}\left(d_{2}^{2}\right)\right|\\ &\ll_{\epsilon}\frac{T^{2}X^{\epsilon}z^{4+\epsilon}+X^{\epsilon}z^{5+\epsilon}}{N^{2}}.\end{split}

Similarly, we have

jw(j)etj/T1XXx2X(1dz|λuj(d2)|min(1,1Nxd))2ϵT2Xϵzϵ+Xϵz1+ϵ+T2Xϵz4+ϵ+Xϵz5+ϵN2.\begin{split}\sum_{j}&w(j)e^{-t_{j}/T}\frac{1}{X}\sum_{X\leq x\leq 2X}\left(\sum_{1\leq d\leq z}\left|\lambda_{u_{j}}\left(d^{2}\right)\right|\min\left(1,\frac{1}{N\|\frac{x}{d}\|}\right)\right)^{2}\\ &\ll_{\epsilon}T^{2}X^{\epsilon}z^{\epsilon}+X^{\epsilon}z^{1+\epsilon}+\frac{T^{2}X^{\epsilon}z^{4+\epsilon}+X^{\epsilon}z^{5+\epsilon}}{N^{2}}.\end{split}

In order to proceed, we choose N=X5.N=X^{5}. By the condition 2logX(logH)/2<logT,2\log X-(\log H)/2<\log T, we have

jw(j)etj/T1X1[1,2](xX)|(x,z)|2𝑑xϵT2Xϵ.\sum_{j}w(j)e^{-t_{j}/T}\frac{1}{X}\int 1_{[1,2]}\left(\frac{x}{X}\right)\left|\mathcal{M^{\prime}}(x,z)\right|^{2}dx\ll_{\epsilon}T^{2}X^{\epsilon}.

Using the same argument as in the proof of Proposition 3.1, we can express the contribution of the main term in (4.1) as

(4.6) (1π2+O(Hϵ/2))dz10<nX51n2sin2(πnHd)+Oϵ(z2+4ϵ(logX)2T1+z4+4ϵ(logX)2T2).\left(\frac{1}{\pi^{2}}+O\left(H^{-\epsilon/2}\right)\right)\sum_{d\leq z}1\sum_{0<n\leq X^{5}}\frac{1}{n^{2}}\sin^{2}\left(\frac{\pi nH}{d}\right)+O_{\epsilon}\left(z^{2+4\epsilon}(\log X)^{2}T^{-1}+z^{4+4\epsilon}(\log X)^{2}T^{-2}\right).

Proposition 4.2.

Assume the conditions in Proposition 4.1. As XX\rightarrow\infty,

(4.7) jϖjetj/T1XX2X(x<ndx+Hd>zλuj(d2))2𝑑xϵTϵH1+ϵ.\begin{split}\sum_{j}&\varpi_{j}e^{-t_{j}/T}\frac{1}{X}\int_{X}^{2X}\left(\sum_{\begin{subarray}{c}x<nd\leq x+H\\ d>z\end{subarray}}\lambda_{u_{j}}(d^{2})\right)^{2}dx\\ &\ll_{\epsilon}T^{\epsilon}H^{1+\epsilon}.\end{split}
Proof.

Since z>H,z>H, for each x[X,2X],x\in[X,2X], we have

(4.8) (x<ndx+Hd>zλuj(d2))2=(z<dx+Hλuj(d2)δ[xd]+1=[x+Hd])2.\begin{split}\left(\sum_{\begin{subarray}{c}x<nd\leq x+H\\ d>z\end{subarray}}\lambda_{u_{j}}(d^{2})\right)^{2}=\left(\sum_{z<d\leq x+H}\lambda_{u_{j}}\left(d^{2}\right)\delta_{[\frac{x}{d}]+1=[\frac{x+H}{d}]}\right)^{2}.\end{split}

By squaring out the sum, we express the above quantity as

(4.9) z<d1x+Hz<d2x+Hλuj(d12)λuj(d22)δ[xd1]+1=[x+Hd1]δ[xd2]+1=[x+Hd2].\sum_{z<d_{1}\leq x+H}\sum_{z<d_{2}\leq x+H}\lambda_{u_{j}}\left(d_{1}^{2}\right)\lambda_{u_{j}}\left(d_{2}^{2}\right)\delta_{[\frac{x}{d_{1}}]+1=[\frac{x+H}{d_{1}}]}\delta_{[\frac{x}{d_{2}}]+1=[\frac{x+H}{d_{2}}]}.

By Lemma 2.4, we observe that

(4.10) jϖjetj/Tz<d1x+Hz<d2x+Hλuj(d12)λuj(d22)δ[xd1]+1=[x+Hd1]δ[xd2]+1=[x+Hd2]ϵ(z<dx+Hδ[xd]+1=[x+Hd]+z<d1x+Hz<d2x+H((d1d2)ϵT1ϵ+(d1d2)1/2+ϵT2)δ[xd]+1=[x+Hd]).\begin{split}&\sum_{j}\varpi_{j}e^{-t_{j}/T}\sum_{z<d_{1}\leq x+H}\sum_{z<d_{2}\leq x+H}\lambda_{u_{j}}\left(d_{1}^{2}\right)\lambda_{u_{j}}\left(d_{2}^{2}\right)\delta_{[\frac{x}{d_{1}}]+1=[\frac{x+H}{d_{1}}]}\delta_{[\frac{x}{d_{2}}]+1=[\frac{x+H}{d_{2}}]}\\ &\ll_{\epsilon}\left(\sum_{z<d\leq x+H}\delta_{[\frac{x}{d}]+1=[\frac{x+H}{d}]}+\sum_{z<d_{1}\leq x+H}\sum_{z<d_{2}\leq x+H}\left(\frac{\left(d_{1}d_{2}\right)^{\epsilon}}{T^{1-\epsilon}}+\frac{\left(d_{1}d_{2}\right)^{1/2+\epsilon}}{T^{2}}\right)\delta_{[\frac{x}{d}]+1=[\frac{x+H}{d}]}\right).\end{split}

For each interval [a,a+d)[X,2X+H],[a,a+d)\subset[X,2X+H], there are HH terms b[a,a+d)b\in[a,a+d) such that

(4.11) δ[bd]+1=[b+Hd]=1,\delta_{[\frac{b}{d}]+1=[\frac{b+H}{d}]}=1,

and 0 for other terms. Therefore, by interchanging the order of summations, we have

(4.12) jϖjetj/T1XX2X(x<ndx+Hd>zλuj(d2))2𝑑xϵTϵ1X(z<d2X+HHXd+H+HX2+ϵT1ϵ+HX3+ϵT2)ϵTϵ(Hlog2X+Hz+HX1+ϵT1ϵ+HX2+ϵT2).\begin{split}&\sum_{j}\varpi_{j}e^{-t_{j}/T}\frac{1}{X}\int_{X}^{2X}\left(\sum_{\begin{subarray}{c}x<nd\leq x+H\\ d>z\end{subarray}}\lambda_{u_{j}}(d^{2})\right)^{2}dx\\ &\ll_{\epsilon}T^{\epsilon}\frac{1}{X}\left(\sum_{z<d\leq 2X+H}H\frac{X}{d}+H+H\frac{X^{2+\epsilon}}{T^{1-\epsilon}}+H\frac{X^{3+\epsilon}}{T^{2}}\right)\\ &\ll_{\epsilon}T^{\epsilon}\left(H\log\frac{2X+H}{z}+H\frac{X^{1+\epsilon}}{T^{1-\epsilon}}+H\frac{X^{2+\epsilon}}{T^{2}}\right).\end{split}

Since 2logX(logH)/2<logT,2\log X-(\log H)/2<\log T, the proof is completed. ∎

Proposition 4.3.

Assume the conditions in Proposition 4.1. Then

jϖjetj/T1XX2X|Hc1,ujHdzλuj(d2)d|2𝑑x=Oϵ(TϵH2z1+ϵ).\sum_{j}\varpi_{j}e^{-t_{j}/T}\frac{1}{X}\int_{X}^{2X}\left|Hc_{1,u_{j}}-H\sum_{d\leq z}\frac{\lambda_{u_{j}}\left(d^{2}\right)}{d}\right|^{2}dx=O_{\epsilon}\left(T^{\epsilon}H^{2}z^{-1+\epsilon}\right).
Proof.

Let U(1,X].U\in(1,X]. Using Perron’s formula (see [6, Lemma 1.1]) and the Kim-Sarnak bound |λuj(n)|n7/64|\lambda_{u_{j}}(n)|\ll n^{7/64} ([13]), we see that

(4.13) dzλuj(d2)d=12πi1logziU1logz+iUL(sym2uj,w+1)ζ(2w+2)zww𝑑w+Oϵ(z7/32+ϵU+z7/32+ϵ1),\sum_{d\leq z}\frac{\lambda_{u_{j}}\left(d^{2}\right)}{d}=\frac{1}{2\pi i}\int_{\frac{1}{\log z}-iU}^{\frac{1}{\log z}+iU}\frac{L\left(\textrm{sym}^{2}u_{j},w+1\right)}{\zeta(2w+2)}\frac{z^{w}}{w}dw+O_{\epsilon}\left(\frac{z^{7/32+\epsilon}}{U}+z^{7/32+\epsilon-1}\right),

where w(j)ϵ|tj+1|ϵw(j)\ll_{\epsilon}|t_{j}+1|^{\epsilon} (see [7]). Using Hölder’s inequality, the Phragmen-Lindelöf principle, and (1.8), we have

(4.14) jw(j)etj/T|1logz±iU12+ϵ±iUL(sym2uj,w+1)ζ(2w+2)zww𝑑w|2jw(j)etj/T1logz±iU12+ϵ±iU|L(sym2uj,w+1)ζ(2w+2)zww|2𝑑wϵT2+ϵU2.\begin{split}&\sum_{j}w(j)e^{-t_{j}/T}\left|\int_{\frac{1}{\log z}\pm iU}^{-\frac{1}{2}+\epsilon\pm iU}\frac{L\left(\textrm{sym}^{2}u_{j},w+1\right)}{\zeta(2w+2)}\frac{z^{w}}{w}dw\right|^{2}\\ &\ll\sum_{j}w(j)e^{-t_{j}/T}\int_{\frac{1}{\log z}\pm iU}^{-\frac{1}{2}+\epsilon\pm iU}\left|\frac{L\left(\textrm{sym}^{2}u_{j},w+1\right)}{\zeta(2w+2)}\frac{z^{w}}{w}\right|^{2}dw\\ &\ll_{\epsilon}\frac{T^{2+\epsilon}}{U^{2}}.\end{split}

Therefore, by shifting the line of the integral to the line Re(w)=1/2+ϵ,\textrm{Re}(w)=-1/2+\epsilon, we have

(4.15) jw(j)etj/T(dzλuj(d2)dc1,uj)2ϵjw(j)etj/T|1/2+ϵiU1/2+ϵ+iUL(sym2uj,w+1)ζ(2w+2)zww𝑑w|2+T2(z7/16+ϵU2+z7/16+ϵ2+TϵU2).\begin{split}\sum_{j}&w(j)e^{-t_{j}/T}\left(\sum_{d\leq z}\frac{\lambda_{u_{j}}\left(d^{2}\right)}{d}-c_{1,u_{j}}\right)^{2}\\ &\ll_{\epsilon}\sum_{j}w(j)e^{-t_{j}/T}\left|\int_{-1/2+\epsilon-iU}^{-1/2+\epsilon+iU}\frac{L\left(\textrm{sym}^{2}u_{j},w+1\right)}{\zeta(2w+2)}\frac{z^{w}}{w}dw\right|^{2}\\ &\;\;\;+T^{2}\left(\frac{z^{7/16+\epsilon}}{U^{2}}+z^{7/16+\epsilon-2}+\frac{T^{\epsilon}}{U^{2}}\right).\end{split}

Using Hölder’s inequality and (1.8), we see that

(4.16) jw(j)etj/T|1/2+ϵiU1/2+ϵ+iUL(sym2uj,w+1)ζ(2w+2)zww𝑑w|2jw(j)etj/T1/2+ϵiU1/2+ϵ+iU|L(sym2uj,w+1)ζ(2w+2)zww12ϵ|2𝑑w1/2+ϵiU1/2+ϵ+iU|1w12+ϵ|2𝑑wϵz1+ϵUϵT2+ϵ.\begin{split}\sum_{j}&w(j)e^{-t_{j}/T}\left|\int_{-1/2+\epsilon-iU}^{-1/2+\epsilon+iU}\frac{L\left(\textrm{sym}^{2}u_{j},w+1\right)}{\zeta(2w+2)}\frac{z^{w}}{w}dw\right|^{2}\\ &\ll\sum_{j}w(j)e^{-t_{j}/T}\int_{-1/2+\epsilon-iU}^{-1/2+\epsilon+iU}\left|\frac{L\left(\textrm{sym}^{2}u_{j},w+1\right)}{\zeta(2w+2)}\frac{z^{w}}{w^{\frac{1}{2}-\epsilon}}\right|^{2}dw\int_{-1/2+\epsilon-iU}^{-1/2+\epsilon+iU}\left|\frac{1}{w^{\frac{1}{2}+\epsilon}}\right|^{2}dw\\ &\ll_{\epsilon}z^{-1+\epsilon}U^{\epsilon}T^{2+\epsilon}.\end{split}

By choosing U=z1+ϵU=z^{1+\epsilon} and modifying ϵ,\epsilon, the proof is completed. ∎

4.2. Proof of Theorem 1.2

By combining the bounds from Proposition 3.2, 4.1, 4.2, and 4.3, we get

Ttj2Tϖjetj/T1XX2X(x<mx+Hλuj2(m)c1,ujH)2𝑑xϵTϵH1+ϵ.\sum_{T\leq t_{j}\leq 2T}\varpi_{j}e^{-t_{j}/T}\frac{1}{X}\int_{X}^{2X}\left(\sum_{x<m\leq x+H}\lambda_{u_{j}}^{2}(m)-c_{1,u_{j}}H\right)^{2}dx\ll_{\epsilon}T^{\epsilon}H^{1+\epsilon}.

Since etj/T1e^{-t_{j}/T}\asymp 1 for tj[T,2T],t_{j}\in[T,2T], the proof is completed.

Acknowledgements

The author would like to express gratitude to Professor Xiaoqing Li for her unwavering support. Additionally, we extend our appreciation to the anonymous referee for providing us with a multitude of valuable suggestions, which have significantly enhanced the quality of our results. The author is supported by the Doctoral Dissertation Fellowship from the Department of Mathematics at the State University of New York at Buffalo.

5. Data Availability

Data sharing not applicable to this article as no data sets were generated or analysed during the current study.

6. Declarations

No funding was received for conducting this study. The authors have no relevant financial or non-financial interests to disclose.

7. Appendix

In this section, we only consider the variance of Hecke eigenvalues of holomorphic cusp forms over arithmetic progressions modulo prime numbers. However, it is straightforward to extend these results to Hecke-Maass cusp forms and to any arithmetic progression. Assuming qq is a prime number and (a,q)=1(a,q)=1, we can use the orthogonality of Dirichlet characters to rewrite the sum

n=1namodqXλf(n)2\sum_{n=1\atop n\equiv a\;\textrm{mod}\;q}^{X}\lambda_{f}(n)^{2}

as

1ϕ(q)χχ¯(a)n=1Xλf(n)2χ(n),\frac{1}{\phi(q)}\sum_{\chi}\bar{\chi}(a)\sum_{n=1}^{X}\lambda_{f}(n)^{2}\chi(n),

where the sum over χ\chi runs over all Dirichlet characters modulo q.q. We can treat the sum over nontrivial characters as an error term. Therefore, we have

n=1namodqXλf(n)21ϕ(q)n=1(n,q)=1Xλf(n)2=1ϕ(q)χχ0χ(a)n=1Xλf(n)2χ(n).\sum_{n=1\atop n\equiv a\;\textrm{mod}\;q}^{X}\lambda_{f}(n)^{2}-\frac{1}{\phi(q)}\sum_{n=1\atop(n,q)=1}^{X}\lambda_{f}(n)^{2}=\frac{1}{\phi(q)}\sum_{\chi\neq\chi_{0}}\chi(a)\sum_{n=1}^{X}\lambda_{f}(n)^{2}\chi(n).

By modifying the argument in [14, Lemma 3.4, Remark 3.5], we can estimate the sum over nn coprime to qq as

(7.1) 1ϕ(q)n=1(n,q)=1Xλf(n)2=wf,qϕ(q)X+Ef,q,\frac{1}{\phi(q)}\sum_{n=1\atop(n,q)=1}^{X}\lambda_{f}(n)^{2}=\frac{w_{f,q}}{\phi(q)}X+E_{f,q},

where

Ef,q=Oϵ(X1/2logX(logq)7ϕ(q)((1/2)|n=1(n,q)=1λf(n)2ns1s|2𝑑s)1/2+(logq)7ϕ(q)X1/2+ϵ),E_{f,q}=O_{\epsilon}\left(\frac{X^{1/2}\log X(\log q)^{7}}{\phi(q)}\left(\int_{(1/2)}\bigg{|}\sum_{\begin{subarray}{c}n=1\\ \left(n,q\right)=1\end{subarray}}^{\infty}\frac{\lambda_{f}\left(n\right)^{2}}{n^{s}}\frac{1}{s}\bigg{|}^{2}ds\right)^{1/2}+\frac{(\log q)^{7}}{\phi(q)}X^{1/2+\epsilon}\right),

and for some constant wf,qw_{f,q}. Therefore, we have

|n=1namodqXλf(n)2wf,qϕ(q)X|2=|Ef,q+1ϕ(q)χχ0χ¯(a)n=1Xλf(n)2χ(n)|2.\left|\sum_{n=1\atop n\equiv a\;\textrm{mod}\;q}^{X}\lambda_{f}(n)^{2}-\frac{w_{f,q}}{\phi(q)}X\right|^{2}=\left|E_{f,q}+\frac{1}{\phi(q)}\sum_{\chi\neq\chi_{0}}\bar{\chi}(a)\sum_{n=1}^{X}\lambda_{f}(n)^{2}\chi(n)\right|^{2}.

Expanding the square, we obtain

(7.2) |Ef,q+1ϕ(q)χχ0χ¯(a)n=1Xλf(n)2χ(n)|2=|Ef,q|2+2Re(Ef,q1ϕ(q)χχ0χ¯(a)n=1Xλf(n)2χ(n))+|1ϕ(q)χχ0χ¯(a)n=1Xλf(n)2χ(n)|2.\begin{split}&\left|E_{f,q}+\frac{1}{\phi(q)}\sum_{\chi\neq\chi_{0}}\bar{\chi}(a)\sum_{n=1}^{X}\lambda_{f}(n)^{2}\chi(n)\right|^{2}\\ &=|E_{f,q}|^{2}+2Re\left(E_{f,q}\frac{1}{\phi(q)}\sum_{\chi\neq\chi_{0}}\bar{\chi}(a)\sum_{n=1}^{X}\lambda_{f}(n)^{2}\chi(n)\right)+\left|\frac{1}{\phi(q)}\sum_{\chi\neq\chi_{0}}\bar{\chi}(a)\sum_{n=1}^{X}\lambda_{f}(n)^{2}\chi(n)\right|^{2}.\end{split}

Therefore, we see that

(7.3) 1q1a=1q1|n=1namodqXλf(n)2wf,qϕ(q)X|2=Ef,q2+1q1a=1q1|1ϕ(q)χχ0χ¯(a)n=1Xλf(n)2χ(n)|2.\begin{split}&\frac{1}{q-1}\sum_{a=1}^{q-1}\left|\sum_{n=1\atop n\equiv a\;\textrm{mod}\;q}^{X}\lambda_{f}(n)^{2}-\frac{w_{f,q}}{\phi(q)}X\right|^{2}\\ &=E_{f,q}^{2}+\frac{1}{q-1}\sum_{a=1}^{q-1}\left|\frac{1}{\phi(q)}\sum_{\chi\neq\chi_{0}}\bar{\chi}(a)\sum_{n=1}^{X}\lambda_{f}(n)^{2}\chi(n)\right|^{2}.\end{split}

We define G(n,m)G(n,m) as follows: if qq divides nmn-m, then G(n,m)=q2G(n,m)=q-2; otherwise, G(n,m)=1G(n,m)=-1. By expanding and squaring out the last term, we obtain

1q1a=1q1|1ϕ(q)χχ0χ¯(a)n=1(n,q)=1Xλf(n)2χ(n)|2=1ϕ(q)2n=1(n,q)=1Xm=1(m,q)=1Xλf(n)2λf(m)2G(n,m).\frac{1}{q-1}\sum_{a=1}^{q-1}\left|\frac{1}{\phi(q)}\sum_{\chi\neq\chi_{0}}\bar{\chi}(a)\sum_{n=1\atop(n,q)=1}^{X}\lambda_{f}(n)^{2}\chi(n)\right|^{2}=\frac{1}{\phi(q)^{2}}\sum_{n=1\atop(n,q)=1}^{X}\sum_{m=1\atop(m,q)=1}^{X}\lambda_{f}(n)^{2}\lambda_{f}(m)^{2}G(n,m).

Using the Petersson trace formula, the average of the right-hand side of the above inequality is given by

(7.4) 1ϕ(q)2d=1(d,q)=1Xn=1(n,q)=1Xdm=1(m,q)=1XdG(dn,dm)=1ϕ(q)2d=1(d,q)=1Xn=1(n,q)=1Xdm=1(m,q)=1XdG(n,m)1ϕ(q)2d=1(d,q)=1Xn=1(n,q)=1Xdq1q(X(logX)).\begin{split}\frac{1}{\phi(q)^{2}}\sum_{d=1\atop(d,q)=1}^{X}\sum_{n=1\atop(n,q)=1}^{\frac{X}{d}}\sum_{m=1\atop(m,q)=1}^{\frac{X}{d}}G(dn,dm)&=\frac{1}{\phi(q)^{2}}\sum_{d=1\atop(d,q)=1}^{X}\sum_{n=1\atop(n,q)=1}^{\frac{X}{d}}\sum_{m=1\atop(m,q)=1}^{\frac{X}{d}}G(n,m)\\ &\ll\frac{1}{\phi(q)^{2}}\sum_{d=1\atop(d,q)=1}^{X}\sum_{n=1\atop(n,q)=1}^{\frac{X}{d}}q\\ &\ll\frac{1}{q}\left(X(\log X)\right).\end{split}

By (1.4), we see that

fSkφf(1/2)|n=1(n,q1)=1λf(q0n)2ns1s|2𝑑sϵfSkφfkϵ.\sum_{f\in S_{k}}\varphi_{f}\int_{(1/2)}\bigg{|}\sum_{\begin{subarray}{c}n=1\\ \left(n,q_{1}\right)=1\end{subarray}}^{\infty}\frac{\lambda_{f}\left(q_{0}n\right)^{2}}{n^{s}}\frac{1}{s}\bigg{|}^{2}ds\ll_{\epsilon}\sum_{f\in S_{k}}\varphi_{f}k^{\epsilon}.

Therefore,

(7.5) fSkφf|Ef,q|2ϵkϵX1+ϵq2+ϵ.\begin{split}\sum_{f\in S_{k}}\varphi_{f}|E_{f,q}|^{2}\ll_{\epsilon}k^{\epsilon}X^{1+\epsilon}q^{-2+\epsilon}.\end{split}

Thus, the average variance of Hecke eigenvalues of holomorphic cusp forms over arithmetic progressions is bounded by kϵX1+ϵq1+ϵk^{\epsilon}X^{1+\epsilon}q^{-1+\epsilon} for families fSkf\in S_{k}, provided that qq is less than X1ϵX^{1-\epsilon}.

References

  • [1] V. Blomer, J. Buttcane, and N. Raulf, A Sato-Tate law for GL(3)\rm GL(3), Comment. Math. Helv., 89 (2014), pp. 895–919.
  • [2] D. Goldfeld, Automorphic forms and L-functions for the group GL(n,R){\rm GL}(n,\rm R), vol. 99 of Cambridge Studies in Advanced Mathematics, Cambridge University Press, Cambridge, 2015. With an appendix by Kevin A. Broughan, Paperback edition of the 2006 original [ MR2254662].
  • [3] D. Goldfeld, E. Stade, and M. Woodbury, An asymptotic orthogonality relation for gl(n,𝕣){\rm gl}(n,\mathbb{r}), 2022.
  • [4] O. Gorodetsky, A. Mangerel, and B. Rodgers, Squarefrees are Gaussian in short intervals, J. Reine Angew. Math., 795 (2023), pp. 1–44.
  • [5] O. Gorodetsky, K. Matomäki, M. Radziwiłł, and B. Rodgers, On the variance of squarefree integers in short intervals and arithmetic progressions, Geom. Funct. Anal., 31 (2021), pp. 111–149.
  • [6] G. Harman, Prime-detecting sieves, vol. 33 of London Mathematical Society Monographs Series, Princeton University Press, Princeton, NJ, 2007.
  • [7] J. Hoffstein and P. Lockhart, Coefficients of Maass forms and the Siegel zero, Ann. of Math. (2), 140 (1994), pp. 161–181. With an appendix by Dorian Goldfeld, Hoffstein and Daniel Lieman.
  • [8] B. Huang, On the Rankin-Selberg problem, Math. Ann., 381 (2021), pp. 1217–1251.
  • [9] A. Ivić, Large values of certain number-theoretic error terms, Acta Arith., 56 (1990), pp. 135–159.
  • [10]  , On the Rankin-Selberg problem in short intervals, Mosc. J. Comb. Number Theory, 2 (2012), pp. 3–17.
  • [11] H. Iwaniec and E. Kowalski, Analytic number theory, vol. 53 of American Mathematical Society Colloquium Publications, American Mathematical Society, Providence, RI, 2004.
  • [12] R. Khan and M. P. Young, Moments and hybrid subconvexity for symmetric-square LL-functions, Journal of the Institute of Mathematics of Jussieu, (2021), pp. 1–45.
  • [13] H. H. Kim, Functoriality for the exterior square of GL4{\rm GL}_{4} and the symmetric fourth of GL2{\rm GL}_{2}, J. Amer. Math. Soc., 16 (2003), pp. 139–183. With appendix 1 by Dinakar Ramakrishnan and appendix 2 by Kim and Peter Sarnak.
  • [14] J. Kim, On the asymptotics of the shifted sums of hecke eigenvalue squares, Forum Mathematicum, 35 (2023), pp. 297–328.
  • [15] A. P. Mangerel, Divisor-bounded multiplicative functions in short intervals, Res. Math. Sci., 10 (2023), p. Paper No. 12.
  • [16] K. Matomäki, M. Radziwiłł, and T. Tao, Correlations of the von Mangoldt and higher divisor functions II: divisor correlations in short ranges, Math. Ann., 374 (2019), pp. 793–840.
  • [17] H. L. Montgomery, Ten lectures on the interface between analytic number theory and harmonic analysis, vol. 84 of CBMS Regional Conference Series in Mathematics, Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 1994.
  • [18] R. A. Rankin, Contributions to the theory of Ramanujan’s function τ(n)\tau(n) and similar arithmetical functions. I. The zeros of the function n=1τ(n)/ns\sum^{\infty}_{n=1}\tau(n)/n^{s} on the line Rs=13/2{R}s=13/2. II. The order of the Fourier coefficients of integral modular forms, Proc. Cambridge Philos. Soc., 35 (1939), pp. 351–372.
  • [19] A. Selberg, Bemerkungen über eine Dirichletsche Reihe, die mit der Theorie der Modulformen nahe verbunden ist, Arch. Math. Naturvid., 43 (1940), pp. 47–50.