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

On the error term in a mixed moment of LL-functions

Rizwanur Khan and Zeyuan Zhang Department of Mathematics
University of Mississippi
University, MS 38677
[email protected], [email protected]
Abstract.

There has recently been some interest in optimizing the error term in the asymptotic for the fourth moment of Dirichlet LL-functions and a closely related mixed moment of LL-functions involving automorphic LL-functions twisted by Dirichlet characters. We obtain an improvement for the error term of the latter.

Key words and phrases:
Dirichlet LL-functions, modular forms, moments.
2020 Mathematics Subject Classification:
11M06, 11M41, 11F11, 11F12
The authors were supported by NSF grant DMS-2001183. The first author was also supported by NSF grant DMS-2140604 and the Simons Foundation (award 630985).

1. Introduction

Heath-Brown [5] was the first to prove an asymptotic formula for the fourth moment of the Riemann Zeta function on the critical line with a power saving error term. The analogous asymptotic for Dirichlet LL-functions required new ideas and came much later. Young [14] proved that

(1.1) 1φ(q)χmodq|L(12,χ)|4=P4(logq)+O(q5512+ϵ),\displaystyle\frac{1}{\varphi^{*}(q)}\mathop{{\sum}^{\star}}_{\chi\bmod q}|L(\tfrac{1}{2},\chi)|^{4}=P_{4}(\log q)+O(q^{-\frac{5}{512}+\epsilon}),

where qq is prime, \mathop{{\sum}^{\star}} restricts the sum to primitive Dirichlet characters, φ(q)\varphi^{*}(q) is the number of primitive Dirichlet characters, and P4P_{4} is a degree four polynomial. Wu [13] obtained a power saving asymptotic for general moduli qq with an error term of O(q11448)O(q^{-\frac{11}{448}}). Blomer, Fouvry, Kowalski, Michel, and Milićević [1] revisited the problem, and introducing new ideas, they significantly sharpened Young’s error term for qq prime to O(q132+ϵ)O(q^{-\frac{1}{32}+\epsilon}). Later Blomer et al. [2] returned to this problem once more and used an idea of Shparlinski and T. Zhang [12] to reduce the error term even further to O(q120+ϵ)O(q^{-\frac{1}{20}+\epsilon}), and to O(q114+ϵ)O(q^{-\frac{1}{14}+\epsilon}) assuming the Ramanujan Conjecture.111Actually the weaker bound O(q116+ϵ)O(q^{-\frac{1}{16}+\epsilon}) is claimed on the Ramanujan conjecture at the end of section 4 in [2], but this seems to be a calculation mistake. This has also been observed in [15, Theorem 1.1] with θ=0\theta=0 and L=1L=1.

In [1], Blomer et al. also considered a related problem which is one step greater in difficulty. They established the mixed moment asymptotic

(1.2) 1φ(q)χmodqL(12,fχ)L(12,χ)¯2=L(1,f)2ζ(2)+O(q168+ϵ),\displaystyle\frac{1}{\varphi^{*}(q)}\mathop{{\sum}^{\star}}_{\chi\bmod q}L(\tfrac{1}{2},f\otimes\chi)\overline{L(\tfrac{1}{2},\chi)}^{2}=\frac{L(1,f)^{2}}{\zeta(2)}+O(q^{-\frac{1}{68}+\epsilon}),

for prime qq, where ff is a fixed (holomorphic or non-holomorphic) Hecke eigenform for SL2()SL_{2}(\mathbb{Z}). There has been some interest in improving the error term in this asymptotic as well. Shparlinski [11] improved the error term in (1.2) slightly to O(q164+ϵ)O(q^{-\frac{1}{64}+\epsilon}). The goal of this paper is to make a more significant improvement to this error term. In fact when ff is holomorphic, the quality of our error term is almost as good as the error term in the fourth moment of Dirichlet LL-functions. This is achieved by an idea which simplifies previous approaches to the problem in a critical range.

Theorem 1.1.

Let qq be an odd prime. Let ηf=122\eta_{f}=\frac{1}{22} for ff a holomorphic Hecke-cusp form, and ηf=5152\eta_{f}=\frac{5}{152} for ff a Hecke Maass-cusp form, for SL2()SL_{2}(\mathbb{Z}). We have

(1.3) 1φ(q)χmodqL(12,fχ)L(12,χ)¯2=L(1,f)2ζ(2)+O(qηf+ϵ).\displaystyle\frac{1}{\varphi^{*}(q)}\mathop{{\sum}^{\star}}_{\chi\bmod q}L(\tfrac{1}{2},f\otimes\chi)\overline{L(\tfrac{1}{2},\chi)}^{2}=\frac{L(1,f)^{2}}{\zeta(2)}+O(q^{-\eta_{f}+\epsilon}).

We also mention that the more difficult related asymptotic

(1.4) 1φ(q)χmodqL(12,fχ)L(12,gχ)¯={2L(1,fg)ζ(2)+O(qδ) if fg,2L(1,sym2f)ζ(2)logq+O(qδ) if f=g,\displaystyle\frac{1}{\varphi^{*}(q)}\mathop{{\sum}^{\star}}_{\chi\bmod q}L(\tfrac{1}{2},f\otimes\chi)\overline{L(\tfrac{1}{2},g\otimes\chi)}=\begin{cases}\frac{2L(1,f\otimes g)}{\zeta(2)}+O(q^{-\delta})&\text{ if }f\neq g,\\ \frac{2L(1,{\rm sym}^{2}f)}{\zeta(2)}\log q+O(q^{-\delta})&\text{ if }f=g,\end{cases}

where ff and gg are Hecke eigenforms and δ>0\delta>0, has been obtained by Kowalski, Michel, and Sawin [10] for prime qq, though no serious attempt has been made to optimize the error term. We refer the reader to [3] for applications of this result.

Acknowledgment

We thank the anonymous referee for their valuable comments.

2. Sketch.

We briefly indicate the main idea behind the improved error term in Theorem 1.1 before proceeding to the proof. After using an approximate functional equation and the orthogonality of characters, the error term is roughly

(2.1) max1NMq2+ϵ|1MNnmmodqnN,mMnmλf(m)d(n)|,\displaystyle\max_{1\leq NM\leq q^{2+\epsilon}}\Bigg{|}\frac{1}{\sqrt{MN}}\sum_{\begin{subarray}{c}n\equiv m\bmod q\\ n\sim N,m\sim M\\ n\neq m\end{subarray}}\lambda_{f}(m)d(n)\Bigg{|},

where λf(m)\lambda_{f}(m) are the Hecke eigenvalues associated to ff and d(n)d(n) is the divisor function. This sum is treated using different techniques depending on the the sizes of NN and MM.

Suppose that NMN\geq M. When N/MN/M is bounded by a reasonably small power of qq, this is treated by solving a ‘shifted convolution problem’, entailing the spectral theory of automorphic forms. Now suppose N/MN/M is somewhat large, the most difficult case being N/Mq12N/M\sim q^{\frac{1}{2}}, or equivalently Nq32N\sim q^{\frac{3}{2}} and Mq12M\sim q^{\frac{1}{2}} when NMq2NM\sim q^{2}. The congruence in (2.1) is detected using additive characters and then Voronoi summation is applied to the nn sum (for this step we need a long nn sum, hence the assumption NMN\geq M). Writing N=N1N2N=N_{1}N_{2} for N1N2N_{1}\leq N_{2}, this leads to a sum of the shape (after opening up the divisor function)

(2.2) h1q/N1,h2q/N2,mMλf(m)S(m,h1h2,q).\displaystyle\sum_{\begin{subarray}{c}h_{1}\sim q/N_{1},h_{2}\sim q/N_{2},m\sim M\end{subarray}}\lambda_{f}(m)S(m,h_{1}h_{2},q).

An effective way to treat the most difficult case N/Mq12N/M\sim q^{\frac{1}{2}} is to use Fouvry, Kowalksi, and Michel’s [4] bounds for bilinear sums of Kloosterman sums, based on their work on algebraic trace functions. To use this, one groups together two of the three variables h1,h2,mh_{1},h_{2},m to form a new variable, which gives a bilinear sum in this new variable and the remaining variable. The bound is non-trivial when one variable is longer than qϵq^{\epsilon} and one is longer than q12+ϵq^{\frac{1}{2}+\epsilon} (the so called Pólya-Vinogradov range). However this strategy fails when we have N1q12N_{1}\sim q^{\frac{1}{2}} and N2qN_{2}\sim q, for then h1Mq12,h21h_{1}\sim M\sim q^{\frac{1}{2}},h_{2}\sim 1, and there is no grouping that satisfies the requirements. This troublesome range was treated by a new bound for bilinear sums of Kloosterman sums obtained by Blomer et al. [1, Theorem 5.1, equation (5.3)]. Later, Shparlinski gave a better bound using his own result on sums of Kloosterman sums [11, Theorem 2.1]. The aforementioned bounds for sums of Kloosterman sums are very general because they can be used for arbitrary coefficients in place of λf(m)\lambda_{f}(m). On the other hand, we make gains by using the automorphic nature of these coefficients. We transform the sum of Kloosterman sums using Poisson summation to an additively twisted sum of λf(m)\lambda_{f}(m), and then employ a Wilton type bound giving square-root cancelation. This strategy yields a bound which essentially forces N1N_{1} to be close to N2N_{2}, thus giving a superior treatment of the critical range N/Mq12N/M\sim q^{\frac{1}{2}} and N1q12N_{1}\sim q^{\frac{1}{2}} and N2qN_{2}\sim q discussed above. Moreover, this approach is simple in that it completely avoids the consideration of bilinear sums of Kloosterman sums.

The case M>NM>N is somewhat similar after the ‘switching trick’ of Blomer et al. (Voronoi summation) which reduces to sums like (2.1) with NMN^{\prime}\geq M^{\prime} typically, but now without the restriction NMq2+ϵN^{\prime}M^{\prime}\leq q^{2+\epsilon}. The lack of this restriction in this case weakens the final bound a bit.

3. Proof of Theorem 1.1

3.1. Preliminaries

Throughout, qq will denote an odd prime number, and ff will denote a holomorphic Hecke cusp form or non-holomorphic Hecke-Maass cusp form for SL2()SL_{2}(\mathbb{Z}), with Hecke eigenvalues λf(n)\lambda_{f}(n). We will follow the ϵ\epsilon convention: that is, ϵ\epsilon will always represent an arbitrarily small positive constant, but not necessarily the same one from one occurrence to another. All implied constants may depend implicitly on ϵ\epsilon and ff. By a ‘negligible’ quantity, we will mean one which is O(qA)O(q^{-A}) for any A>0A>0, where the implied constant depends on AA. Throughout, we will use WW, possibly with a subscript, to denote a smooth function compactly supported on the interval [12,2][\frac{1}{2},2], with derivatives satisfying

W(j)(x)(qϵ)j\displaystyle W^{(j)}(x)\ll(q^{\epsilon})^{j}

for any j0j\geq 0, where the implied constant depends on jj. Such a weight function may differ from one occurrence to the next even when we use the same notation.

We will write e(x)e(x) to denote e2πixe^{2\pi ix}. We will use the following elementary orthogonality relation:

(3.1) 1qamodqe(anq)={1 if q|n,0 if qn.\displaystyle\frac{1}{q}\sum_{a\bmod q}e\Big{(}\frac{an}{q}\Big{)}=\begin{cases}1&\text{ if }q|n,\\ 0&\text{ if }q\nmid n.\end{cases}

We have the average bound [6, Lemma 1]

nx|λf(n)|2x1+ϵ.\sum_{n\leq x}|\lambda_{f}(n)|^{2}\ll x^{1+\epsilon}.

For individual bounds, we have

|λf(n)|nθf+ϵ,|\lambda_{f}(n)|\ll n^{\theta_{f}+\epsilon},

where θf=764\theta_{f}=\frac{7}{64} if ff is Maass (see [9]) and θf=0\theta_{f}=0 if ff is holomorphic.

For smooth sums of Hecke eigenvalues, we note a simple consequence of Voronoi summation: for MqϵM\geq q^{\epsilon}, we have

(3.2) m1λf(m)W(mM)qA\displaystyle\sum_{m\geq 1}\lambda_{f}(m)W\Big{(}\frac{m}{M}\Big{)}\ll q^{-A}

for any A>0A>0. See for example [1, Lemmas 2.3, 2.4], where we can take ϑ=0\vartheta=0 since the Selberg eigenvalue conjecture is known in level 1.

For (not necessarily smooth) sums of Hecke eigenvalues twisted by additive characters, we have

Lemma 3.1.

Let α\alpha\in\mathbb{R}. We have

nNλf(n)e(nα)N12+ϵ.\sum_{n\leq N}\lambda_{f}(n)e(n\alpha)\ll N^{\frac{1}{2}+\epsilon}.

The implied constant is uniform in α\alpha, but depends on ff.

Proof.

See [8, Theorem 5.3], [7, Theorem 8.1]. ∎

3.2. Some known bounds

For N,M1N,M\geq 1, define

(3.3) E±(M,N):=S1+S2\displaystyle E^{\pm}(M,N):=S_{1}+S_{2}
=1MNm±nmodqmnλf(m)d(n)W1(mM)W2(nN)1qMNm,n1λf(m)d(n)W1(mM)W2(nN).\displaystyle=\frac{1}{\sqrt{MN}}\sum_{\begin{subarray}{c}m\equiv\pm n\bmod q\\ m\neq n\end{subarray}}\lambda_{f}(m)d(n)W_{1}\Big{(}\frac{m}{M}\Big{)}W_{2}\Big{(}\frac{n}{N}\Big{)}-\frac{1}{q\sqrt{MN}}\sum_{m,n\geq 1}\lambda_{f}(m)d(n)W_{1}\Big{(}\frac{m}{M}\Big{)}W_{2}\Big{(}\frac{n}{N}\Big{)}.

Note that E±(M,N)E^{\pm}(M,N) is the same as the quantity Ef,E±(M,N)=Bf,E±(M,N)E_{f,E}^{\pm}(M,N)=B_{f,E}^{\pm}(M,N) defined on [1, page 771]. To establish Theorem 1.1, we need to prove

(3.4) maxMNq2+ϵ|E±(M,N)|qηf+ϵ.\displaystyle\max_{MN\leq q^{2+\epsilon}}|E^{\pm}(M,N)|\ll q^{-\eta_{f}+\epsilon}.

We begin by recalling some established bounds for E±(M,N)E^{\pm}(M,N), but for later use it will be important to not restrict to MNq2+ϵMN\leq q^{2+\epsilon}.

Lemma 3.2.

Let MNq4+ϵMN\leq q^{4+\epsilon}. We have

(3.5) E±(M,N){qϵMNq if NMqϵMθfMNq if M>N,\displaystyle E^{\pm}(M,N)\ll\begin{cases}q^{\epsilon}\frac{\sqrt{MN}}{q}&\text{ if }N\geq M\\ q^{\epsilon}M^{\theta_{f}}\frac{\sqrt{MN}}{q}&\text{ if }M>N,\end{cases}
(3.6) E±(M,N)qϵ(1+(NMq2)14)(max(N,M)qmin(N,M))14(1+(max(N,M)qmin(N,M))14).\displaystyle E^{\pm}(M,N)\ll q^{\epsilon}\Big{(}1+\Big{(}\frac{NM}{q^{2}}\Big{)}^{\frac{1}{4}}\Big{)}\Big{(}\frac{\max(N,M)}{q\min(N,M)}\Big{)}^{\frac{1}{4}}\Big{(}1+\Big{(}\frac{\max(N,M)}{q\min(N,M)}\Big{)}^{\frac{1}{4}}\Big{)}.

If max(N,M)min(N,M)>10\frac{\max(N,M)}{\min(N,M)}>10, then

(3.7) E±(M,N)qϵ(qmin(N,M)max(N,M))12+qϵMθf(NM)12q2.\displaystyle E^{\pm}(M,N)\ll q^{\epsilon}\Big{(}\frac{q\min(N,M)}{\max(N,M)}\Big{)}^{\frac{1}{2}}+q^{\epsilon}M^{\theta_{f}}\frac{(NM)^{\frac{1}{2}}}{q^{2}}.
Proof.

The first bound (3.5) is the trivial bound given in [1, Proposition 3.1].

The second bound is obtained in [1, Theorem 3.2] by solving a shifted convolution problem using the spectral theory of automorphic forms, with an additional idea required to avoid dependence on the Ramanujan conjecture. However, just reading the statement of that theorem is not enough because it assumes NMq2+ϵNM\leq q^{2+\epsilon}. We look at the proof of [1, Theorem 3.2] and find that the bound (3.6) is contained in the bounds given by [1, equations (3.12), (3.16)], after dividing by the (MN)12(MN)^{\frac{1}{2}} factor in [1, equation (3.4)]. This then gives exactly what is seen in the statement of [1, Theorem 3.2], apart from an extra factor O(1+(NMq2)14)O(1+(\frac{NM}{q^{2}})^{\frac{1}{4}}) which we retain in (3.6) and which would have been redundant had we been assuming MNq2+ϵMN\leq q^{2+\epsilon}, and a missing O(q12+θf+ϵ)O(q^{-\frac{1}{2}+\theta_{f}+\epsilon}) term, which is specific to the case of ff and gg both cuspidal (see the bottom of [1, page 726]) and therefore does not appear in (3.6).

The third bound (3.7) is obtained in [1, section 6.4] by applying Voronoi summation to both nn and mm sums and then bounding trivially using Weil’s bound for the Kloosterman sum. The assumption max(N,M)min(N,M)>10\frac{\max(N,M)}{\min(N,M)}>10 implies that the condition nmn\neq m is vacuous in the sum S1S_{1} in the definition (3.3) of E±(M,N)E^{\pm}(M,N). Then E±(M,N)E^{\pm}(M,N) may be expressed as [1, equation (6.21)], for which we have the bound given in [1, equation (6.23)] plus the trivial bound for the expression in [1, equation (6.22)]. The latter bound gives the term qϵMθf(NM)12q2q^{\epsilon}M^{\theta_{f}}\frac{(NM)^{\frac{1}{2}}}{q^{2}} in (3.7) which does not appear in [1, equation (6.23)] because it is assumed there that MNq2+ϵMN\leq q^{2+\epsilon}, while we do not make this assumption. ∎

3.3. A new bound

Define the sum (closely following the notation in [1, equation (6.26)]) for M,N1,N21M,N_{1},N_{2}\geq 1 with N=N1N2N=N_{1}N_{2} and N1N2N_{1}\leq N_{2}:

C±(M,N,N1,N2):=1NMn1,n2,m1n1n2±mmodqλf(m)W1(n1N1)W2(n2N2)W3(mM).C^{\pm}(M,N,N_{1},N_{2}):=\frac{1}{\sqrt{NM}}\sum_{\begin{subarray}{c}n_{1},n_{2},m\geq 1\\ n_{1}n_{2}\equiv\pm m\bmod q\end{subarray}}\lambda_{f}(m)W_{1}\Big{(}\frac{n_{1}}{N_{1}}\Big{)}W_{2}\Big{(}\frac{n_{2}}{N_{2}}\Big{)}W_{3}\Big{(}\frac{m}{M}\Big{)}.

We will prove the following simple bound which was not observed in [1] or [11].

Lemma 3.3.

Let MN<18q3MN<\frac{1}{8}q^{3} and MqϵM\geq q^{\epsilon}. We have

(3.8) C±(M,N,N1,N2)qϵ(N1N2)12+qA,\displaystyle C^{\pm}(M,N,N_{1},N_{2})\ll q^{\epsilon}\Big{(}\frac{N_{1}}{N_{2}}\Big{)}^{\frac{1}{2}}+q^{-A},

for any A>0A>0.

Proof.

Detecting the congruence n1n2±mmodqn_{1}n_{2}\equiv\pm m\bmod q using additive characters, we have

C±(M,N,N1,N2)\displaystyle C^{\pm}(M,N,N_{1},N_{2}) =1qNMamodqn1,n2,m1λf(m)W1(n1N1)W2(n2N2)W3(mM)e(a(n1n2m)q)\displaystyle=\frac{1}{q\sqrt{NM}}\sum_{a\bmod q}\ \sum_{\begin{subarray}{c}n_{1},n_{2},m\geq 1\end{subarray}}\lambda_{f}(m)W_{1}\Big{(}\frac{n_{1}}{N_{1}}\Big{)}W_{2}\Big{(}\frac{n_{2}}{N_{2}}\Big{)}W_{3}\Big{(}\frac{m}{M}\Big{)}e\Big{(}\frac{a(n_{1}n_{2}\mp m)}{q}\Big{)}
=1qNMamodqbmodqn1,n2,m1n2bmodqλf(m)W1(n1N1)W2(n2N2)W3(mM)e(a(n1bm)q).\displaystyle=\frac{1}{q\sqrt{NM}}\sum_{\begin{subarray}{c}a\bmod q\\ b\bmod q\end{subarray}}\ \sum_{\begin{subarray}{c}n_{1},n_{2},m\geq 1\\ n_{2}\equiv b\bmod q\end{subarray}}\lambda_{f}(m)W_{1}\Big{(}\frac{n_{1}}{N_{1}}\Big{)}W_{2}\Big{(}\frac{n_{2}}{N_{2}}\Big{)}W_{3}\Big{(}\frac{m}{M}\Big{)}e\Big{(}\frac{a(n_{1}b\mp m)}{q}\Big{)}.

Now we apply Poisson summation to the n2n_{2} sum (more precisely, we write n2=q+bn_{2}=\ell q+b and apply Poisson summation to the \ell sum) to get

(3.9) C±(M,N,N1,N2)=1qNMN2qamodqbmodqn1,m1<k<λf(m)e(bkq)e(a(n1bm)q)W1(n1N1)W2^(kN2q)W3(mM),C^{\pm}(M,N,N_{1},N_{2})\\ =\frac{1}{q\sqrt{NM}}\frac{N_{2}}{q}\sum_{\begin{subarray}{c}a\bmod q\\ b\bmod q\end{subarray}}\ \sum_{\begin{subarray}{c}n_{1},m\geq 1\\ -\infty<k<\infty\end{subarray}}\lambda_{f}(m)e\Big{(}\frac{bk}{q}\Big{)}e\Big{(}\frac{a(n_{1}b\mp m)}{q}\Big{)}W_{1}\Big{(}\frac{n_{1}}{N_{1}}\Big{)}\widehat{W_{2}}\Big{(}\frac{kN_{2}}{q}\Big{)}W_{3}\Big{(}\frac{m}{M}\Big{)},

where W2^(y)=W2(x)e(yx)𝑑x\widehat{W_{2}}(y)=\int_{\mathbb{R}}W_{2}(x)e(-yx)dx is the Fourier transform of W2W_{2}. By repeated integration by parts, we may restrict the kk sum to |k|<q1+ϵN2|k|<\frac{q^{1+\epsilon}}{N_{2}}, up to a negligible error.

Consider first the contribution to (3.9) of the terms with q|n1q|n_{1}, a condition which implies that N112qN_{1}\geq\frac{1}{2}q. This contribution is

(3.10) 1qNMN2qamodqbmodqn1,m10|k|<q1+ϵN2q|n1λf(m)e(bkq)e(a(m)q)W1(n1N1)W2^(kN2q)W3(mM)+O(qA)\displaystyle\frac{1}{q\sqrt{NM}}\frac{N_{2}}{q}\sum_{\begin{subarray}{c}a\bmod q\\ b\bmod q\end{subarray}}\ \sum_{\begin{subarray}{c}n_{1},m\geq 1\\ 0\leq|k|<\frac{q^{1+\epsilon}}{N_{2}}\\ q|n_{1}\end{subarray}}\lambda_{f}(m)e\Big{(}\frac{bk}{q}\Big{)}e\Big{(}\frac{a(\mp m)}{q}\Big{)}W_{1}\Big{(}\frac{n_{1}}{N_{1}}\Big{)}\widehat{W_{2}}\Big{(}\frac{kN_{2}}{q}\Big{)}W_{3}\Big{(}\frac{m}{M}\Big{)}+O(q^{-A})

for any A>0A>0. Using (3.1), we see that the aa sum vanishes unless qq divides mm, in which case we must have M12qM\geq\frac{1}{2}q. But since we are assuming MN<18q3MN<\frac{1}{8}q^{3}, we cannot have M12qM\geq\frac{1}{2}q and N2N112qN_{2}\geq N_{1}\geq\frac{1}{2}q, so it follows that (3.10) is O(qA)O(q^{-A}).

It remains to consider the contribution to (3.9) of the terms with qn1q\nmid n_{1}. Evaluating the bb sum using (3.1) enforces akn1¯modqa\equiv k\overline{n_{1}}\bmod q, where n1¯\overline{n_{1}} denotes the multiplicative inverse of n1n_{1} modulo qq. Thus we have that (3.9) equals

N2qNMn1,m10|k|<q1+ϵN2qn1λf(m)e(mkn1¯)q)W1(n1N1)W2^(kN2q)W3(mM)+O(qA).\frac{N_{2}}{q\sqrt{NM}}\sum_{\begin{subarray}{c}n_{1},m\geq 1\\ 0\leq|k|<\frac{q^{1+\epsilon}}{N_{2}}\\ q\nmid n_{1}\end{subarray}}\lambda_{f}(m)e\Big{(}\frac{\mp mk\overline{n_{1}})}{q}\Big{)}W_{1}\Big{(}\frac{n_{1}}{N_{1}}\Big{)}\widehat{W_{2}}\Big{(}\frac{kN_{2}}{q}\Big{)}W_{3}\Big{(}\frac{m}{M}\Big{)}+O(q^{-A}).

The contribution to the sum of the term k=0k=0 is negligible by using (3.2). We estimate the contribution of the terms k0k\neq 0 by bounding the mm sum using Lemma 3.1 and the rest trivially, to get

N2qNMq1+ϵN2N1Mqϵ(N1N2)12.\frac{N_{2}}{q\sqrt{NM}}\frac{q^{1+\epsilon}}{N_{2}}N_{1}\sqrt{M}\ll q^{\epsilon}\Big{(}\frac{N_{1}}{N_{2}}\Big{)}^{\frac{1}{2}}.

In order to use the new bound towards our goal (3.4), we need

Lemma 3.4.

If max(N,M)min(N,M)>10\frac{\max(N,M)}{\min(N,M)}>10 and MqϵM\geq q^{\epsilon}, then

(3.11) |E±(M,N)|maxN=N1N2N1N2|C±(M,N,N1,N2)|+O(qA).\displaystyle|E^{\pm}(M,N)|\ll\max_{\begin{subarray}{c}N=N_{1}N_{2}\\ N_{1}\leq N_{2}\end{subarray}}|C^{\pm}(M,N,N_{1},N_{2})|+O(q^{-A}).

for any A>0A>0.

Proof.

The assumption MqϵM\geq q^{\epsilon} implies by (3.2) that the sum S2S_{2} in the definition (3.3) of E±(M,N)E^{\pm}(M,N) is negligible. The assumption max(N,M)min(N,M)>10\frac{\max(N,M)}{\min(N,M)}>10 implies that the condition nmn\neq m is vacuous in the sum S1S_{1} in the definition (3.3) of E±(M,N)E^{\pm}(M,N). After removing this condition in S1S_{1}, we may open the divisor function and apply smooth partitions of unity, as explained in [1, section 6.4.1], to obtain (3.11). ∎

3.4. The case M>NM>N

Recall our goal (3.4). We need to start with the ‘switching trick’ of [1].

Lemma 3.5.

Let MNq2+ϵMN\leq q^{2+\epsilon} and M>NM>N. Define

M=q2M,N=q2N.M^{*}=\frac{q^{2}}{M},\ \ N^{*}=\frac{q^{2}}{N}.

We have

(3.12) E±(M,N)qϵmax1NqϵN1MqϵM|E±(M,N)|+O(q16+ϵ).\displaystyle E^{\pm}(M,N)\ll q^{\epsilon}\max_{\begin{subarray}{c}1\leq N^{\prime}\leq q^{\epsilon}N^{*}\\ 1\leq M^{\prime}\leq q^{\epsilon}M^{*}\end{subarray}}|E^{\pm}(M^{\prime},N^{\prime})|+O(q^{-\frac{1}{6}+\epsilon}).
Proof.

By applying Voronoi summation to the nn and mm sums in (3.3), Blomer et al. [1, page 765] showed222They actually have an additional factor (M/M)2θf(M^{*}/M^{\prime})^{2\theta_{f}}, where θf=7/64\theta_{f}=7/64, which arises from the use of their Lemma 2.4, but this factor is not needed because we work with Hecke eigenforms of level 1. In this setting, the Selberg eigenvalue conjecture is known, so that we may take ϑ=0\vartheta=0 in their Lemma 2.4.

(3.13) E±(M,N)\displaystyle E^{\pm}(M,N) qϵmax1NqϵN1MqϵM|1MNm,n1m±nmodqλf(m)d(n)W1(mM)W2(nN)|\displaystyle\ll q^{\epsilon}\max_{\begin{subarray}{c}1\leq N^{\prime}\leq q^{\epsilon}N^{*}\\ 1\leq M^{\prime}\leq q^{\epsilon}M^{*}\end{subarray}}\Bigg{|}\frac{1}{\sqrt{M^{*}N^{*}}}\sum_{\begin{subarray}{c}m,n\geq 1\\ m\equiv\pm n\bmod q\end{subarray}}\lambda_{f}(m)d(n)W_{1}\Big{(}\frac{m}{M^{\prime}}\Big{)}W_{2}\Big{(}\frac{n}{N^{\prime}}\Big{)}\Bigg{|}
+qϵmax1NqϵN1MqϵM|1qMNm,n1λf(m)d(n)W1(mM)W2(nN)|+O(q1+ϵ),\displaystyle+q^{\epsilon}\max_{\begin{subarray}{c}1\leq N^{\prime}\leq q^{\epsilon}N^{*}\\ 1\leq M^{\prime}\leq q^{\epsilon}M^{*}\end{subarray}}\Bigg{|}\frac{1}{q\sqrt{M^{*}N^{*}}}\sum_{m,n\geq 1}\lambda_{f}(m)d(n)W_{1}\Big{(}\frac{m}{M^{\prime}}\Big{)}W_{2}\Big{(}\frac{n}{N^{\prime}}\Big{)}\Bigg{|}+O(q^{-1+\epsilon}),

where

M=q2M,N=q2N.M^{*}=\frac{q^{2}}{M},\ \ N^{*}=\frac{q^{2}}{N}.

The contribution of the diagonal terms n=mn=m in the first sum on the right hand side of (3.13) is bounded absolutely by

qϵ1MNMqϵ1MNMqϵMN=qϵNM.q^{\epsilon}\frac{1}{\sqrt{M^{*}N^{*}}}M^{\prime}\ll q^{\epsilon}\frac{1}{\sqrt{M^{*}N^{*}}}M^{*}\ll q^{\epsilon}\sqrt{\frac{M^{*}}{N^{*}}}=q^{\epsilon}\sqrt{\frac{N}{M}}.

We can assume this is O(q16+ϵ)O(q^{-\frac{1}{6}+\epsilon}) or else by (3.6) we have E±(M,N)q16+ϵE^{\pm}(M,N)\ll q^{-\frac{1}{6}+\epsilon}. Thus at the cost of an admissible error term, the diagonal terms can be removed. By (3.2), the second sum on the right hand side of (3.13) is negligible unless M<qϵM^{\prime}<q^{\epsilon}. Thus this sum is trivially bounded by

qϵ1qMNNqϵqMN.q^{\epsilon}\frac{1}{q\sqrt{M^{*}N^{*}}}N^{*}\ll\frac{q^{\epsilon}}{q}\sqrt{\frac{M}{N}}.

We can assume this is O(q14+ϵ)O(q^{-\frac{1}{4}+\epsilon}) or else by (3.7) we have E±(M,N)q14+ϵE^{\pm}(M,N)\ll q^{-\frac{1}{4}+\epsilon}. The lemma follows. ∎

We will apply Lemma 3.3 and Lemma 3.4 to bound E±(M,N)E^{\pm}(M^{\prime},N^{\prime}). For this, we need to verify the conditions of these lemmas. Suppose that the condition max(N,M)min(N,M)>10\frac{\max(N^{\prime},M^{\prime})}{\min(N^{\prime},M^{\prime})}>10 does not hold. Then by (3.5) and (3.6), we would have

E±(M,N)qϵmin((M)θf(NMq2)12,(NMq2)14q14)qϵmin(q732(NMq2)12,(NMq2)14q14)q332+ϵ,E^{\pm}(M^{\prime},N^{\prime})\ll q^{\epsilon}\min\Bigg{(}(M^{\prime})^{\theta_{f}}\Big{(}\frac{N^{\prime}M^{\prime}}{q^{2}}\Big{)}^{-\frac{1}{2}},\Big{(}\frac{N^{\prime}M^{\prime}}{q^{2}}\Big{)}^{\frac{1}{4}}q^{-\frac{1}{4}}\Bigg{)}\\ \ll q^{\epsilon}\min\Bigg{(}q^{\frac{7}{32}}\Big{(}\frac{N^{\prime}M^{\prime}}{q^{2}}\Big{)}^{-\frac{1}{2}},\Big{(}\frac{N^{\prime}M^{\prime}}{q^{2}}\Big{)}^{\frac{1}{4}}q^{-\frac{1}{4}}\Bigg{)}\ll q^{-\frac{3}{32}+\epsilon},

and we would be done by (3.12). So we can assume that max(N,M)min(N,M)>10\frac{\max(N^{\prime},M^{\prime})}{\min(N^{\prime},M^{\prime})}>10 holds. Now suppose that the condition MqϵM^{\prime}\geq q^{\epsilon} does not hold. Then by (3.5) and (3.7), we would have

E±(M,N)qϵmin(Nq,(qN)12+Nq2)(Nq,(qN)12+q1)q14+ϵ,E^{\pm}(M^{\prime},N^{\prime})\ll q^{\epsilon}\min\Bigg{(}\frac{\sqrt{N^{\prime}}}{q},\Big{(}\frac{q}{N^{\prime}}\Big{)}^{\frac{1}{2}}+\frac{\sqrt{N^{\prime}}}{q^{2}}\Bigg{)}\ll\Bigg{(}\frac{\sqrt{N^{\prime}}}{q},\Big{(}\frac{q}{N^{\prime}}\Big{)}^{\frac{1}{2}}+q^{-1}\Bigg{)}\ll q^{-\frac{1}{4}+\epsilon},

and we would be done by (3.12). So we can assume that MqϵM^{\prime}\geq q^{\epsilon} holds. Finally suppose that the condition MN<18q3M^{\prime}N^{\prime}<\frac{1}{8}q^{3} does not hold. Then by (3.5), we would have MNq3M^{*}N^{*}\gg q^{3}, which implies MNqMN\ll q. Then by (3.5), we would have E±(M,N)(M)θfq12+ϵq73212+ϵE^{\pm}(M^{\prime},N^{\prime})\ll(M^{\prime})^{\theta_{f}}q^{-\frac{1}{2}+\epsilon}\ll q^{\frac{7}{32}-\frac{1}{2}+\epsilon} and we would be done by (3.12). So we can assume the condition MN<18q3M^{\prime}N^{\prime}<\frac{1}{8}q^{3} as well.

Now we are ready to prove

Lemma 3.6.

We have

maxMNq2+ϵM>N|E±(M,N)|qηf+ϵ.\max_{\begin{subarray}{c}MN\leq q^{2+\epsilon}\\ M>N\end{subarray}}|E^{\pm}(M,N)|\ll q^{-\eta_{f}+\epsilon}.
Proof.

We use Lemma 3.3 and Lemma 3.4 to bound E±(M,N)E^{\pm}(M^{\prime},N^{\prime}), which implies a bound for E±(M,N)E^{\pm}(M,N) via Lemma 3.5. We add our bound to the collection of bounds given in [1]. We need to find the worst case (maximum) of the minimum of all these bounds for MNq2+ϵMN\leq q^{2+\epsilon} and M>NM>N. This leads to a linear optimization problem to maximize the minimum of the exponents of qq from each bound. Such a problem was solved in [1] by a computer search, and we do the same by simply inserting the exponent of qq from our bound for C±(M,N,N1,N2)C^{\pm}(M^{\prime},N^{\prime},N_{1},N_{2}), given in (3.8), into the existing Mathematica code from [1, section 7.4.2]. The underlined portion below indicates the new addition to this code. This returns an exponent of 5152-\frac{5}{152} for ff Maass and 122-\frac{1}{22} for ff holomorphic (in the latter case we also omit the term 7m/647m/64 from the first entry in the existing code, corresponding to the bound (3.5), since θf=0\theta_{f}=0).

In[1] := Maximize[{Min[(m + n)/2 - 1 + 7m/64, Max[(m - n - 1)/2, (m - n - 1)/4],
(1 + n - m)/2, 7/64(mstar-mprime) + Max[(2 mprime - mstar - nstar + 1)/ 2,
(2 n1 + 2 mprime - 2 - mstar - nstar)/2], (n1 - n2)/2,
7/64(mstar-mprime) + Max[(mprime + n2 + 1/2 - nstar - mstar)/ 2,
(2 mprime + n1 - mstar - nstar)/ 2, (2 n1 + 2 mprime - 2 - mstar - nstar)/2],
7/64(mstar-mprime) + mprime + n1 - (mstar + nstar)/2,
7/64(mstar-mprime) + Max[Min[2 n1 - (mstar + nstar)/2,
2/3 nprime + n1 + 1/2 - 1/6 mprime - n2 - (mstar + nstar)/2],
(2 mprime - mstar - nstar)/ 2, (2 mprime + 2 n1 - 2 - mstar - nstar)/2,
2 mprime - n2 - (mstar + nstar)/2]], 0 <= n, n <= m, m + n <= 2,
nstar == 2 - n, mstar == 2 - m, 0 <= nprime, nprime <= nstar,
0 <= mprime, mprime <= mstar, n1 + n2 == nprime, 0 <= n1, n1 <= n2},
{m, n, n1, n2, nprime, nstar, mprime, mstar}]
Out[1] := {5152,{𝚖2419,𝚗1538,𝚗𝟷6176,𝚗𝟸6176,𝚗𝚙𝚛𝚒𝚖𝚎6138,𝚗𝚜𝚝𝚊𝚛6138\Bigl{\{}-\frac{5}{152},\Bigl{\{}{\tt m}\rightarrow\frac{24}{19},\,\,{\tt n}\rightarrow\frac{15}{38},\,\,{\tt n1}\rightarrow\frac{61}{76},\,\,{\tt n2}\rightarrow\frac{61}{76},\,\,{\tt nprime}\rightarrow\frac{61}{38},\,\,{\tt nstar}\rightarrow\frac{61}{38},
𝚖𝚙𝚛𝚒𝚖𝚎1419,𝚖𝚜𝚝𝚊𝚛1419}}{\tt mprime}\rightarrow\frac{14}{19},\,\,{\tt mstar}\rightarrow\frac{14}{19}\Bigr{\}}\Bigr{\}}
In[2] := Maximize[{Min[(m + n)/2 - 1, Max[(m - n - 1)/2, (m - n - 1)/4],
(1 + n - m)/2, 7/64(mstar-mprime) + Max[(2 mprime - mstar - nstar + 1)/ 2,
(2 n1 + 2 mprime - 2 - mstar - nstar)/2], (n1 - n2)/2,
7/64(mstar-mprime) + Max[(mprime + n2 + 1/2 - nstar - mstar)/ 2,
(2 mprime + n1 - mstar - nstar)/ 2, (2 n1 + 2 mprime - 2 - mstar - nstar)/2],
7/64(mstar-mprime) + mprime + n1 - (mstar + nstar)/2,
7/64(mstar-mprime) + Max[Min[2 n1 - (mstar + nstar)/2,
2/3 nprime + n1 + 1/2 - 1/6 mprime - n2 - (mstar + nstar)/2],
(2 mprime - mstar - nstar)/ 2, (2 mprime + 2 n1 - 2 - mstar - nstar)/2,
2 mprime - n2 - (mstar + nstar)/2]], 0 <= n, n <= m, m + n <= 2,
nstar == 2 - n, mstar == 2 - m, 0 <= nprime, nprime <= nstar,
0 <= mprime, mprime <= mstar, n1 + n2 == nprime, 0 <= n1, n1 <= n2},
{m, n, n1, n2, nprime, nstar, mprime, mstar}]
Out[2] := {122,{𝚖1511,𝚗611,𝚗𝟷811,𝚗𝟸811,𝚗𝚙𝚛𝚒𝚖𝚎1611,𝚗𝚜𝚝𝚊𝚛1611\Bigl{\{}-\frac{1}{22},\Bigl{\{}{\tt m}\rightarrow\frac{15}{11},\,\,{\tt n}\rightarrow\frac{6}{11},\,\,{\tt n1}\rightarrow\frac{8}{11},\,\,{\tt n2}\rightarrow\frac{8}{11},\,\,{\tt nprime}\rightarrow\frac{16}{11},\,\,{\tt nstar}\rightarrow\frac{16}{11},
𝚖𝚙𝚛𝚒𝚖𝚎711,𝚖𝚜𝚝𝚊𝚛711}}{\tt mprime}\rightarrow\frac{7}{11},\,\,{\tt mstar}\rightarrow\frac{7}{11}\Bigr{\}}\Bigr{\}}

3.5. The case NMN\geq M

Recall our goal (3.4). Assuming MNq2+ϵMN\leq q^{2+\epsilon} and NMN\geq M, we will apply Lemma 3.3 and Lemma 3.4 to bound E±(M,N)E^{\pm}(M,N). For this purpose, we need to verify the conditions max(N,M)min(N,M)>10,Mqϵ,\frac{\max(N,M)}{\min(N,M)}>10,M\geq q^{\epsilon}, and MN<18q3MN<\frac{1}{8}q^{3} of these lemmas. This is done just as in section 3.4 and is in fact easier. This is because the factor MθfM^{\theta_{f}} does not occur when using (3.5) because we assume that NMN\geq M, and the condition MN<18q3MN<\frac{1}{8}q^{3} is automatic for qq large enough by the assumption MNq2+ϵMN\leq q^{2+\epsilon}.

Now we prove the following, which together with Lemma 3.6 gives Theorem 1.1.

Lemma 3.7.

We have

maxMNq2+ϵNM|E±(M,N)|q120+ϵ.\max_{\begin{subarray}{c}MN\leq q^{2+\epsilon}\\ N\geq M\end{subarray}}|E^{\pm}(M,N)|\ll q^{-\frac{1}{20}+\epsilon}.
Proof.

As before, we simply insert our bound from Lemma 3.3 into the existing Mathematica code, this time from [1, section 7.4.1], and compute.

In[3] := Maximize[{Min[(m + n)/2 - 1, Max[(n - m - 1)/2, (n - m - 1)/4],
(1 + m - n)/2, Max[(m + 1 - n)/2, (2 n1 + m - 2 - n)/2],
Max[1/4 - n1/2, (m - n2)/2, (2 n1 + m - n - 2)/2], (m + 2 n1 - n)/ 2,
Max[Min[2 n1 - (m + n)/2, n/6 + n1 + 1/2 - n2 - 2 m/3], (m - n)/2,
(m + 2 n1 - 2 - n)/2, 3 m/2 - n2 - n/2], If[m + n1prime <= 1,
(2 m + 2 n1prime + 2 n2prime - 1 - n1circ - n2circ - m)/2
+ Max[-n2prime/2, 1/4 - (m + n1prime)/2], 10], (n1 - n2)/2,
If[m + n2prime <= 2 n1prime, (2 m + 2 n1prime + 2 n2prime - 1
- n1circ - n2circ - m)/2 + 1/4 - 5 n1prime/12 - (m + n2prime)/6, 10]],
m >= 0, n >= m, m + n <= 2, n1 + n2 == n, n1 <= n2, n1 >= 0,
n1prime >= 0, n1prime <= n1circ, n1circ == 1 - n1, n2prime >= 0,
n2prime <= n2circ, n2circ == 1 - n2}, {m, n, n1, n2, n1prime,
n2prime, n1circ, n2circ}]
Out[3] := {120,{𝚖35,𝚗75,𝚗𝟷710,𝚗𝟸710,𝚗𝟷𝚙𝚛𝚒𝚖𝚎310,𝚗𝟷𝚙𝚛𝚒𝚖𝚎310\Bigl{\{}-\frac{1}{20},\Bigl{\{}{\tt m}\rightarrow\frac{3}{5},\,\,{\tt n}\rightarrow\frac{7}{5},\,\,{\tt n1}\rightarrow\frac{7}{10},\,\,{\tt n2}\rightarrow\frac{7}{10},\,\,{\tt n1prime}\rightarrow\frac{3}{10},\,\,{\tt n1prime}\rightarrow\frac{3}{10},
𝚗𝟷𝚌𝚒𝚛𝚌310,𝚗𝟷𝚌𝚒𝚛𝚌310}}{\tt n1circ}\rightarrow\frac{3}{10},\,\,{\tt n1circ}\rightarrow\frac{3}{10}\Bigr{\}}\Bigr{\}}

References

  • [1] V. Blomer, É. Fouvry, E. Kowalski, P. Michel, and D. Milićević, On moments of twisted LL-functions, Amer. J. Math. 139 (2017), no. 3, 707–768.
  • [2] by same author, Some applications of smooth bilinear forms with Kloosterman sums, Tr. Mat. Inst. Steklova 296 (2017), no. Analiticheskaya i Kombinatornaya Teoriya Chisel, 24–35, English version published in Proc. Steklov Inst. Math. 296 (2017), no. 1, 18–29.
  • [3] V. Blomer, É. Fouvry, E. Kowalski, P. Michel, D. Milićević, and W. Sawin, The Second Moment Theory of Families of LL-Functions: The Case of Twisted Hecke LL-Functions, Mem. Amer. Math. Soc. 282 (2023), no. 1394.
  • [4] É. Fouvry, Emmanuel K., and P. Michel, Algebraic trace functions over the primes, Duke Math. J. 163 (2014), no. 9, 1683–1736.
  • [5] D. R. Heath-Brown, The fourth power moment of the Riemann zeta function, Proc. London Math. Soc. (3) 38 (1979), no. 3, 385–422.
  • [6] H. Iwaniec, The spectral growth of automorphic LL-functions, J. Reine Angew. Math. 428 (1992), 139–159.
  • [7] by same author, Introduction to the spectral theory of automorphic forms, Biblioteca de la Revista Matemática Iberoamericana. [Library of the Revista Matemática Iberoamericana], Revista Matemática Iberoamericana, Madrid, 1995.
  • [8] by same author, Topics in classical automorphic forms, Graduate Studies in Mathematics, vol. 17, American Mathematical Society, Providence, RI, 1997.
  • [9] 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), no. 1, 139–183, With appendix 1 by D. Ramakrishnan and appendix 2 by H. Kim and P. Sarnak.
  • [10] E. Kowalski, P. Michel, and W. Sawin, Bilinear forms with Kloosterman sums and applications, Ann. of Math. (2) 186 (2017), no. 2, 413–500.
  • [11] I. E. Shparlinski, On sums of Kloosterman and Gauss sums, Trans. Amer. Math. Soc. 371 (2019), no. 12, 8679–8697.
  • [12] I. E. Shparlinski and T. Zhang, Cancellations amongst Kloosterman sums, Acta Arith. 176 (2016), no. 3, 201–210.
  • [13] X. Wu, The fourth moment of Dirichlet LL-functions at the central value, Math. Ann. (doi:10.1007/s00208-022-02483-9).
  • [14] M. P. Young, The fourth moment of Dirichlet LL-functions, Ann. of Math. (2) 173 (2011), no. 1, 1–50.
  • [15] R. Zacharias, Mollification of the fourth moment of Dirichlet LL-functions, Acta Arith. 191 (2019), no. 3, 201–257.