on the Rankin-Selberg problem in families
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 and let Define where A holomorphic function satisfying the equation
is called a modular form of weight It is well known that any such modular form has a Fourier expansion at the cusp given by
(1.1) |
where and the normalized Fourier coefficient of is defined as
(1.2) |
The set of all modular forms of a fixed weight forms a vector space, and we denote it by Furthermore, we focus on a subspace of this space, denoted by , which comprises of all modular forms of a weight that have a zero constant term. Let us consider the th Hecke operator, denoted by , which is defined as
for all
It is well established that there exists an orthonormal basis of eigenfunctions, called Hecke cusp forms, for all Hecke operators in the space of modular forms . In this paper, we will use to denote this orthonormal basis of .
When is a Hecke cusp form, eigenvalues of the th Hecke operator are real numbers that satisfy the following properties:
For details, see [11, Chapter 14]. A function is called a Maass form of type if satisfies the following properties:
-
•
-
•
for all
-
•
where
-
•
When is a non-constant Maass form of type we can express it using the Whittaker expansion given by
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 for all .
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 ,
(1.3) |
where is the symmetric square -function of 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 to . It is worth noting that these results apply not only to the holomorphic Hecke cusp forms but also to the Hecke-Maass cusp forms. For convenience, we denote the coefficient of by Ivic [10] proved that for the following holds:
where Note that under the Lindelöf hypothesis, we have In [10], Ivic also proved that, assuming the generalized Lindelöf hypothesis , the following holds:
Recently, Mangerel [15] proved the following theorem:
Theorem.
Let and set . Then there exist a constant such that
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) |
where
Theorem 1.1.
Assume the Lindelöf-on-average bound (1.4). Let be a sufficiently small number in (0,1). Let Then
(1.5) |
Note that
Let be an Hecke-Maass cusp form with Laplacian eigenvalue and let be the th normalized Fourier coefficient of The function also satisfies
(1.6) |
As mentioned before, we have
(1.7) |
For convenience, we denote the coefficient of by Our methods for Theorem 1.1 apply equally well to Hecke-Maass cusp forms For the Hecke-Maass cusp forms 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 and we have
(1.8) |
Therefore, we unconditionally prove the following theorem.
Theorem 1.2.
Let be a sufficiently small number in (0,1). Let and let Then
(1.9) |
where
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 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) |
It should be noted that while the identity for square-free integers involves , the convolution structure (1.10) considers . 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 into and for some We split the sum over into and By (1.10), we have
When is small, it is expected that
This can be demonstrated by using the Fourier expansion of
(see Proposition 3.1). However, when is large (i.e. ), we have
where
As a result, we encounter a double sum
which is difficult to control. To avoid this, we take a different approach by considering the average over all modular forms of weight 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
with a small error term (see Proposition 3.5).
Remark 1.4.
Our approach clearly applies to demonstrating analogous results for a linear space of cusp forms with weight and level Some of the results in our paper still hold if we replace and with the coefficients of the standard -function attached to a automorphic representation and a suitable family of automorphic representations, respectively. For instance, suppose we let denote the th coefficient of an -function associated with a Hecke-Maass cusp form . We can address the corresponding problem for by considering the expression:
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 Hecke-Maass (or holomorphic) forms:
(1.11) |
Assuming the Lindelöf-on-average bound for , we can easily modify our method to show that
is sufficiently small.
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 as . Additionally, for any condition , we define
We use the common notation that the least common multiple of and , and the greatest common divisor of and are denoted as and respectively. Finally, we use the convention that denotes an arbitrarily small positive quantity that may vary from line to line.
Lemma 2.1.
There exist functions and that map from to and satisfy the following properties:
-
•
The Fourier transforms and have support for a sufficiently large absolute constant .
-
•
where is the indicator function over the interval
-
•
Proof.
See [17, Chapter 1, page 6]. ∎
Since is divisor-bounded, we can apply the following lemma.
Lemma 2.2.
Let be a fixed constant and suppose . Then, for any positive integer , we have the inequality
(2.1) |
Proof.
Using Shiu’s Theorem ([16, Lemma 2.3,(i)]), we obtain the bound
Since for any prime , we have
(2.2) |
By Merten’s theorem, we see that
for some constant Therefore,
∎
Lemma 2.3.
(Petersson trace formula) For any two natural numbers and we have
(2.3) |
where the implied constant is absolute, and denotes the Petersson norm of over .
Proof.
See [11, Corollary 14.24, Theorem 16.7]. ∎
Lemma 2.4.
(Kuznetsov trace formula) Let be an orthonormal basis of Hecke-Maass cusp forms for and let be the Laplace eigenvalue of Suppose and let Then, we have
(2.4) |
where is defined in Theorem 1.2, and the summation over is taken over all . Note that
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 we have
3. Propositions and the proof of Theorem 1.1
Let represent from now on. We begin by splitting the integral in (1.5) into three parts:
(3.1) |
(3.2) |
and
(3.3) |
Using the Cauchy-Schwarz inequality, we obtain an upper bound for the integral in (1.5):
Hence, we need to demonstrate that each of the three integrals is bounded by on average over 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 , the range of is restricted to . 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
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 As ,
Proof.
Without loss of generality, we assume that We have
(3.4) |
where It is known that for any and , we have
(see [11][(4.18)]). We will choose the value of later. Therefore, we see that
(3.5) |
where
(3.6) |
Using the Cauchy-Schwarz inequality, we have
(3.7) |
By Deligne’s bound and using (2.1), we estimate the first term in the last inequality of (3.7) as
(3.8) |
Similarly, we estimate the second term as
(3.9) |
Using the same approach as in (3.7), we have
In order to proceed, we choose Then, we obtain
Next, we consider the contribution of the main term in (3.5), which is given by
(3.10) |
Since (3.10) lies between
(3.11) |
and
(3.12) |
By applying the Petersson trace formula, we have that
(3.13) |
By applying the condition on the supports of and , we see that
(3.14) |
Assuming that is sufficiently small and is sufficiently large so that , we can imply the following:
Therefore, equation (3.13) can be written as
(3.15) |
Note that the main term is non-negative. By applying the third property of in Lemma 2.1, the above term can be written as
(3.16) |
∎
The following proposition employs a similar method as in [5, Lemma 9].
Proposition 3.2.
Assume the condition in Proposition 3.1. As , we have
(3.17) |
Proof.
Define the function by
where is a smooth bump function such that for and for . Then, we have
(3.18) | ||||
Let us consider
Let Then we have the following expression for the Fourier transform of :
where is entire and satisfies uniformly for (see [5, (16)]). By the Mellin inversion, we get
(3.19) |
where By using (3.19), we have
(3.20) |
By using the fact that we see that
Therefore, (3.20) is bounded by
(3.21) |
∎
Remark 3.3.
Proposition 3.4.
Assume the condition in Proposition 3.1. As ,
(3.22) |
Proof.
Since for each we have
(3.23) |
By squaring out the sum, we express the above quantity as
(3.24) |
By Lemma 2.3, we see that
(3.25) |
For each interval there are terms such that
(3.26) |
and for other terms. Therefore, by interchanging the order of summations, we have
(3.27) |
∎
Proposition 3.5.
Assume the condition in Proposition 3.1 and the Lindelöf-on-average bound (1.4). Then
Proof.
Let Using Perron’s formula (see [6, Lemma 1.1]) and Deligne’s bound, we get
(3.28) |
Applying Hölder’s inequality, the Phragmen-Lindelöf principle, and (1.4), we obtain
(3.29) |
Therefore, by moving the line of the integral to the line we have
(3.30) |
By applying Hölder’s inequality and (1.4), we obtain
(3.31) |
By choosing and modifying the proof is completed. ∎
3.2. Proof of Theorem 1.1
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 Ramanujan-Petersson conjecture ().
4.1. Propositions
Proposition 4.1.
Let and let As ,
Proof.
Without loss of generality, we assume that Using the same argument as in the proof of Proposition 3.1, we have
(4.1) |
where
(4.2) |
Using the Cauchy-Schwarz inequality, we see that
(4.3) |
By squaring out the first term in the last inequality in (4.3) and applying Lemma 2.5 with the divisor bound we get
(4.4) |
Similarly, the contribution from the second term in the last inequality in (4.3) is
(4.5) |
Similarly, we have
In order to proceed, we choose By the condition we have
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) |
∎
Proposition 4.2.
Assume the conditions in Proposition 4.1. As ,
(4.7) |
Proof.
Since for each we have
(4.8) |
By squaring out the sum, we express the above quantity as
(4.9) |
By Lemma 2.4, we observe that
(4.10) |
For each interval there are terms such that
(4.11) |
and for other terms. Therefore, by interchanging the order of summations, we have
(4.12) |
Since the proof is completed. ∎
Proposition 4.3.
Assume the conditions in Proposition 4.1. Then
Proof.
Let Using Perron’s formula (see [6, Lemma 1.1]) and the Kim-Sarnak bound ([13]), we see that
(4.13) |
where (see [7]). Using Hölder’s inequality, the Phragmen-Lindelöf principle, and (1.8), we have
(4.14) |
Therefore, by shifting the line of the integral to the line we have
(4.15) |
Using Hölder’s inequality and (1.8), we see that
(4.16) |
By choosing and modifying 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
Since for 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 is a prime number and , we can use the orthogonality of Dirichlet characters to rewrite the sum
as
where the sum over runs over all Dirichlet characters modulo We can treat the sum over nontrivial characters as an error term. Therefore, we have
By modifying the argument in [14, Lemma 3.4, Remark 3.5], we can estimate the sum over coprime to as
(7.1) |
where
and for some constant . Therefore, we have
Expanding the square, we obtain
(7.2) |
Therefore, we see that
(7.3) |
We define as follows: if divides , then ; otherwise, . By expanding and squaring out the last term, we obtain
Using the Petersson trace formula, the average of the right-hand side of the above inequality is given by
(7.4) |
By (1.4), we see that
Therefore,
(7.5) |
Thus, the average variance of Hecke eigenvalues of holomorphic cusp forms over arithmetic progressions is bounded by for families , provided that is less than .
References
- [1] V. Blomer, J. Buttcane, and N. Raulf, A Sato-Tate law for , Comment. Math. Helv., 89 (2014), pp. 895–919.
- [2] D. Goldfeld, Automorphic forms and L-functions for the group , 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 , 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 -functions, Journal of the Institute of Mathematics of Jussieu, (2021), pp. 1–45.
- [13] H. H. Kim, Functoriality for the exterior square of and the symmetric fourth of , 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 and similar arithmetical functions. I. The zeros of the function on the line . 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.