Uniform bounds for Kloosterman sums of half-integral weight, same-sign case
Abstract.
In the previous paper [Sun23], the author proved a uniform bound for sums of half-integral weight Kloosterman sums. This bound was applied to prove an exact formula for partitions of rank modulo 3. That uniform estimate provides a more precise bound for a certain class of multipliers compared to the 1983 result by Goldfeld and Sarnak and generalizes the 2009 result from Sarnak and Tsimerman to the half-integral weight case. However, the author only considered the case when the parameters satisfied . In this paper, we prove the same uniform bound when for further applications.
Key words and phrases:
Kloosterman sum; Maass form; Kuznetsov trace formula1. Introduction
For positive integers , and , we define the generalized Kloosterman sums with a multiplier system as
where is a congruence subgroup of , is a weight multiplier system on , and as defined in (2.6). These Kloosterman sums have been studied by Goldfeld and Sarnak [GS83] and Pribitkin [Pri00]. In a previous paper [Sun23], the author proved a uniform bound for the sums of such Kloosterman sums and applied the bound to estimate the partial sums of Rademacher-type exact formulas. For example, the Fourier coefficient of , which is a sixth order mock theta function defined in [AH91, (0.17)], can be written as [Sun23, Theorem 2.2]
(1.1) |
where is the multiplier system of weight for Dedekind’s eta-function (see (2.4)) and is the -Bessel function. The author bounded the error
(1.2) |
with [Sun23, Theorem 1.6] to get
Our estimate can be applied to sums of Kloosterman sum with a class of multiplier systems. However, in the former paper [Sun23] we only focus on the case . In this paper we add the complementary case for further applications like Corollary 1.6.
To state the results, we need to classify our half-integral weight multipliers first:
Definition 1.1.
Let and where is some even fundamental discriminant and is the multiplier for the theta function. We say a weight multiplier on is admissible if it satisfies the following two conditions:
-
(1)
Level lifting: there exist positive integers and such that the map gives:
-
(i)
an injection from weight automorphic eigenforms of the hyperbolic Laplacian on to those on and keeps the eigenvalue;
-
(ii)
an injection from weight holomorphic cusp forms on to weight holomorphic cusp forms on .
Here is a multiple of and depends on .
-
(i)
-
(2)
Average Weil bound: for and , we have
The author proved [Sun23, Proposition 5.1] that the following class of multipliers satisfy both the conditions:
Lemma 1.2.
Let denote the -th Fourier coefficient of an orthonormal basis of (the space of square-integrable eigenforms of the weight Laplacian with respect to , see Section 2 for details). Here are our theorems:
Theorem 1.3.
Remark.
Based on Theorem 1.3 and [Sun23, Theorem 1.4], we are able to reformulate the Linnik-Selberg conjecture in the half-integral weight case:
Conjecture 1.4.
Suppose and is a weight multiplier system on . If there is no eigenvalue for the hyperbolic Laplacian in , then
(1.4) |
where the second sum runs over normalized eigenforms of with eigenvalue .
We also get the following bound by the properties of Bessel functions.
Theorem 1.5.
With the same setting as Theorem 1.3, for or , we have
(1.5) |
except the case when , and are all of the same sign such that . Here is the Bessel function or . Note that when and are .
1.1. Application
In [BO12], Bringmann and Ono constructed Maass-Poincaré series to prove that the Fourier coefficients of weight , harmonic Maass forms have exact formulas, i.e. they equal infinite sums of Kloosterman sums (up to Fourier coefficients from holomorphic theta functions when the weight is ). The specific case, weight , is crucial as it is related to the rank of partitions. Readers may also refer to [BO06] or [Sun23, Section 4] for further introduction.
We can denote the weight Maass-Poincaré series as for cusp and multiplier of as [BO12, (3.4)]. The exact formulas are derived from at , while we only know the convergence at (see [BO12, Lemma 3.1, (3.7)]). Bringmann and Ono called a weight harmonic Maass form on is good if those Maass-Poincaré series corresponding to nontrivial terms in the principal parts of are individually convergent. They conclude the exact formula for its Fourier coefficients only when is good.
Corollary 1.6.
Suppose is a weight harmonic Maass form on and only has non-trivial principal part at cusp . Then is good if is admissible (Definition 1.1).
The paper is organized as follows. In Section 2 we introduce notations on Kloosterman sums and Maass forms. Section 3 revisits the trace formula by Proskurin [Pro79] and discusses certain properties crucial for the subsequent proof. Section 4 is about bounds on translations of the test function. The proof of Theorem 1.3 is presented in Section 5 and is divided into two cases: the weight and . Readers are advised to take particular note of the Remark following Proposition 5.2 and be mindful of the weight context in Section 5.
2. Background
In this section we recall some basic notions on Maass forms with general weight and multiplier. Let for some denote our congruence subgroup and denote the upper half complex plane. Fixing the argument in , for any and , we define the automorphic factor
and the weight slash operator
for . We say that is a multiplier system of weight if
-
(i)
,
-
(ii)
, and
-
(iii)
for all , where
If is a multiplier system of weight , then it is also a multiplier system of weight and its conjugate is a multiplier system of weight .
We are interested in the following multiplier systems of weight and their conjugates of weight . The theta-multiplier on is given by
(2.1) |
where
and is the extended Kronecker symbol such that for odd . The eta-multiplier on is given by
(2.2) |
where
(2.3) |
Explicit formulas for were given by Rademacher [Rad73, (74.11), (74.12)]:
(2.4) |
for all and also given by Knopp [Kno70]:
(2.5) |
for . The properties and for are convenient in computations.
2.1. Kloosterman sums
For any cusp of , let denote its stabilizer in . For example, . Let denote its scaling matrix satisfying and . We define by the condition
(2.6) |
The cusp is called singular if . When , we drop the subscript, denote , and define for . The Kloosterman sum at the cusp pair with respect to the multiplier system is given by
(2.7) |
They have the relationships
(2.8) |
because
(2.9) |
2.2. Maass forms
We call a function automorphic of weight and multiplier on if
Let denote the linear space consisting of all such functions and denote the space of square-integrable functions on with respect to the measure
and the Petersson inner product
for . For , the Laplacian
(2.10) |
can be expressed as
(2.11) | ||||
(2.12) |
where is the Maass raising operator
(2.13) |
and is the Maass lowering operator
(2.14) |
These operators raise and lower the weight of an automorphic form as
and satisfy the commutative relations
(2.15) |
Moreover, commutes with the weight slash operator for all .
We call a real analytic function an eigenfunction of with eigenvalue if
From (2.15), it is clear that an eigenvalue for the weight Laplacian is also an eigenvalue for weight . We call a Maass form if it is an smooth eigenfunction of and satisfies the growth condition
for all and some depending on when . Moreover, if a Maass form satisfies
for all cusps of , then and we call a Maass cusp form. For details see [AA18, §2.3]
Let denote the space of smooth functions such that both and are bounded. One can show that is dense in and is self-adjoint on . If we let , then for ,
i.e. is bounded from below. By the Friedrichs extension theorem, can be extended to a self-adjoint operator on . The spectrum of consists of two parts: the continuous spectrum and a discrete spectrum of finite multiplicity contained in .
Let denote the first eigenvalue larger than in the discrete spectrum with respect to the congruence subgroup , weight and multiplier . For weight 0, Selberg [Sel65] showed that for all and Selberg’s famous eigenvalue conjecture states that for all . We introduce the hypothesis as
(2.16) |
Selberg’s conjecture includes and the best progress known today is by [Kim03]. We denote as when is clear from context.
Let denote the subspace spanned by eigenfunctions of . For each eigenvalue , we write
So corresponds to and any such eigenvalue is called an exceptional eigenvalue. Set
(2.17) |
Let denote the subspace corresponding to the spectral parameter . Complex conjugation gives an isometry
between normed spaces. For each , we have the Fourier expansion
(2.18) |
where is the Whittaker function and
Using the fact that is a real function when is real and [DLMF, (13.4.4), (13.14.3), (13.14.31)], if we denote the Fourier coefficient of as , then
(2.19) |
2.3. Holomorphic cusp forms of half-integral weight
For positive integers and a weight multiplier on , we know that is also a weight multiplier system on . For simplicity we denote . Let (resp. ) be the space of holomorphic cusp forms (resp. modular forms) on which satisfy the transformation law
The inner product on is defined as
It is known that (see e.g. [Ran77, §5]) is a finite-dimensional Hilbert space under the inner product . If we take an orthonormal basis of and write the Fourier expansion of as
then the sum
(2.20) |
is independent of the choice of the basis.
There is a one-to-one correspondence between all with eigenvalue and weight holomorphic modular forms by
(2.21) |
If we write the Fourier expansion of as , then
(2.22) |
3. Trace formula
Let , be a positive integer, and be a singular cusp for the weight multiplier system on . Define the Eisenstein series associated to by
(3.1) |
We also define the Poincaré series for by
Both of them are automorphic functions of weight . The properties of these two series can be found in [Pro05]. The Fourier expansion of is
(3.2) |
where the function is in [AA18, (4.5)]. When , . The Fourier expansion of at for is
(3.3) | ||||||
Suppose and are positive integers and recall the definition of in §2.1. The following notations are very important in the remaining part of this paper:
Setting 3.1.
Let and with where will be finally taken as .
Setting 3.2.
The test function is a four times continuously differentiable function satisfying
-
(1)
and () as .
-
(2)
for .
-
(3)
for and .
-
(4)
.
-
(5)
and are piecewise monotone on a fixed number of intevals.
This test function was used in [ST09] and [AA18]. We also define the following transformations of :
(3.4) |
and for ,
(3.5) |
with the corrected version of [Blo08, (2.12)]
(3.6) |
For an integer , let denote an orthonormal basis for the space of holomorphic cusp forms and
Suppose that the Fourier expansion of each is given by
(3.7) |
Let denote the weight of . Here is the trace formula:
Theorem 3.3 ([Pro05, §6 Theorem]).
Suppose is a multiplier system of weight or on . Let be an orthonormal basis of and be the Eisenstein series associated to a singular cusp . Let denote the -th Fourier coefficient of . Let or denote the -th Fourier coefficient of as in (3.3). Let and be defined as in (3.7). Then for and we have
(3.8) |
where
and
Remark.
We clarify two points in the theorem.
-
(1)
In the term corresponding to holomorphic cusp forms, each function has weight .
-
(2)
The equality of the two expressions in is by (3.3):
3.1. Properties of admissible multipliers
Suppose is a weight admissible multiplier system on (Definition 1.1) with parameters , and . Then we have the following propositions.
Proposition 3.4.
Proposition 3.5.
Suppose that satisfies condition (1) of Definition 1.1 with . For , let . Suppose is an orthonormal basis of and is an orthonormal basis of . Denote as the Fourier coefficient of and as the Fourier coefficient of . Then we have
(3.9) |
Proof.
Now we start to prove a bound for the right hand side of (3.9). First we have
Proposition 3.6 ([Wai17, Theorem 1]).
For , and a quadratic character modulo , suppose
is an orthonormal basis of . For , write with square-free, and . Then we have
Note that the implied constant in Waibel’s bound depends on when expressing Bessel functions (see [Wai17, after (8)] and [Iwa87, Theorem 1 and p. 400]). For our proof, it’s essential that the bound remains uniform across the weights. We modify the estimate and get the following proposition.
Proposition 3.7.
With the same setting as Proposition 3.6,
Proof.
We do the same preparation as [Wai17, around (8)] to estimate the right hand side of (2.20). Let (finally chosen to be ) and define the set of prime numbers
Here we have .
For a orthonormal basis of , the set is an orthonormal subset of . Recall (2.20) and we have
(3.11) |
For those , . Summing (3.11) on and dividing we get
(3.12) |
The average estimate of
can be found in [Wai17, (19)] that for ,
(3.13) |
We break the sum on at the right hand side of (3.12) into and to estimate. The uniform bound of -Bessel functions is given by [Lan00]
(3.14) |
where .
Proposition 3.8.
With the same setting as Proposition 3.5, we factor with square-free, and . Then
4. Bounds on and
In this section, all of the implied constants among the estimates for and depend on and the multiplier system unless specified. Recall the definitions (3.4) and (3.5). To deal with the -function in the denominator of , we need [DLMF, (5.6.6-7)]
(4.1) |
Recall (3.4) and (3.5) that we define for . We also have
(4.2) |
Like [AD20, after (5.3)], we define as
(4.3) |
Then
(4.4) |
We refer to [Dun90] for estimates on -Bessel functions. Denote
As a result of the relationship for by [DLMF, (10.11.9)], we have
Moreover, for and ,
(4.5) |
and for because and .
Lemma 4.1.
For , uniformly and with absolute implied constants we have
(4.6) |
Proof.
First we deal with the range . The series expansion of is given by [Dun90, (3.9), (3.16)]:
where . The implied constant in the second equation is absolute. As a function of , and . Then and
For the range , we check with [Dun90, (5.16)] where for are fixed polynomials of whose lowest degree term is :
Our claimed bound is clear as . The implied constant above is absolute due to [Dun90, (3.3)] and [Olv97, Chapter 8, §13] or by [Olv97, Chapter 10, (3.04)].
∎
Lemma 4.2.
For , we have
(4.7) |
Proof.
When we focus on the exceptional eigenvalues of , recall that for . By Proposition 3.4, if we write , assuming (2.16) we have an upper bound when . Moreover, since the exceptional eigenvalues are discrete, we also have a largest eigenvalue less than , hence a lower bound (depending on and ) such that .
Lemma 4.3.
Proof.
For the range we have
Lemma 4.4.
[AD20, Lemma 6.3] Let or . Then
(4.10) |
Remark.
4.1. A special test function
Here we choose a special test function satisfying Setting 3.2 to compute the terms corresponding to the exceptional spectrum in Theorem 1.3.
For or , let be an exceptional eigenvalue of on , we set for and
Since the exceptional spectrum is discrete, let the lower bound for be depending on and . Recall Setting 3.1. Let be .
Setting 4.5.
In addition to the requirement in Setting 3.2, when , we pick as a smoothed function of this piecewise linear one
where
(4.11) |
The above choice for is possible because there is no requirement for when but for in Setting 3.2.
Now we take . When , by the series expansion [DLMF, (10.2.2)]:
we have
(4.12) |
The implied constant is absolute. Now we compute the bound for and .
Proof.
When , we get and by Lemma 4.3, so the lemma is true in this case. When , we have and with (4.12),
Recall that we always have the lower bound for . A bound for and follows from the same process as [Sun23, Proof of Lemma 7.2]:
We also get
and finish the proof.
∎
Lemma 4.7.
Proof.
5. Proof of Theorem 1.3 and Theorem 1.5
The proof depends on the following two propositions for the Fourier coefficients of Maass forms, which were originally obtained for the discrete spectrum in [AD22, Theorem 4.1] and [AD19, Theorem 4.3]. The author proved the generalized propositions in [Sun23, §8] to include the continuous spectrum. Recall our notations in Settings 3.1 and 3.2. Suppose that for some ,
(5.1) |
Then we have the following proposition:
Proposition 5.1 ([Sun23, Proposition 8.1]).
Proposition 5.2 ([Sun23, Proposition 8.2]).
Suppose that is a weight admissible multiplier on with , and given in Definition 1.1. Let denote the Fourier coefficients of an orthonormal basis of . For each singular cusp of , let be the associated Eisenstein series. Let be defined as in (3.3). Suppose . For we factor where is square-free, is positive and . Then we have
Remark.
We make some remarks about the weight :
-
•
The trace formula (Theorem 3.3) works for and .
-
•
The estimates on and in the previous section work for and .
-
•
The above two propositions work for and .
Therefore, in this section, we separate the proof of Theorem 1.3 into two cases and . In the second case we will apply the Maass lowering operator (2.14) to connect the eigenforms of weight and weight .
We declare that all implicit constants for the bounds in this section depend on , and , and we drop the subscripts unless specified.
Since the exceptional spectral parameter of Laplacian on is discrete, has a positive lower bound denoted as depending on and . We also have assuming (2.16) by Theorem 3.4. For simplicity let
and
(5.2) |
Recall the notations in Setting 3.1 and Setting 3.2. The following inequalities will be used later in the proof:
(5.3) |
(5.4) |
5.1. On the case
Let denote the coefficients of an orthonormal basis of . For each singular cusp of , let be defined as in (3.3). Recall the definition of in Theorem 1.3 and the notations in Settings 3.1 and 3.2. We claim the following proposition:
Proposition 5.3.
With the same setting as Theorem 1.3 for , when , we have
(5.5) |
We first show that Proposition 5.3 implies Theorem 1.3 in the case , which follows from a similar process as [Sun23, after Proposition 9.1]. Recall that for and that the corresponding exceptional eigenvalue . The sum to be estimated is
(5.6) |
where
Since and , the quantity
is bounded from above and below. By Proposition 5.2,
(5.7) |
When , since by (5.3),
(5.8) |
So in this case we get Theorem 1.3 where the terms are absorbed in the errors.
When , the segment for summing Kloosterman sums on contributes a by condition (2) of Definition 1.1. The segment for can be broken into no more than dyadic intervals with and we use Proposition 5.3 for both the Kloosterman sum and the terms. In summing dyadic intervals, for each , we get
The difference between the above quantity and the quantity in (5.6) is
(5.9) |
by (5.7). In conclusion, for we get
where the second equality follows from (5.9) and the last inequality is by Proposition 5.3. Theorem 1.3 follows in the case .
The proof of Proposition 5.3 takes the rest of this subsection. For , by Proposition 5.2 we have
Recall that and in Setting 3.1. Thanks to (2.16) and Proposition 3.4, when we have . Since by hypothesis, it follows from (5.4) that
Applying Lemma 4.7 where and recalling the definition of in Theorem 1.3, we get
(5.10) |
When and , Lemma 4.7 and (5.4) give
(5.11) |
With the help of (5.10) and (5.11) we break up the left hand side of (5.5) to obtain the following analogue to [Sun23, (9.8)]:
(5.12) |
The first sum above can be estimated by condition (2) of Definition 1.1 as
(5.13) |
We then prove a bound for . By Theorem 3.3, we have
(5.14) |
5.1.1. Contribution from holomorphic forms
For or , recall the notation before Theorem 3.3. For , let be an orthonormal basis of with Fourier coefficient . By Proposition 3.8, uniformly for every with , we have and
We also have
by [Dun18, Lemma 5.1 and proof of Lemma 7.1] and Lemma 4.3. Note that [Dun18, Lemma 5.1] is only for , while the same process works for . Then the contribution from is
Recall and (5.2) for the definition of . Since
we get . Moreover, since
and by hypothesis, we also get
Finally we conclude
(5.15) |
5.1.2. Contribution from Maass cusp forms and Eisenstein series.
We combine the two propositions at the beginning of this section and bounds on in Section 4 to estimate the contribution from the remaining part of (5.14) other than . The process is the same as [Sun23, §9.1] for as shares the same bound as there. We record the bounds in the following equations.
Fix . In the following estimations we focus on the discrete spectrum because each bound for for any interval is the same as the bound for in the continuous spectrum. This is a direct result from Proposition 5.1 and Proposition 5.2. Recall that in the assumption of Proposition 5.3.
Let
Divide into two parts: and . We apply Proposition 5.2 on the first range and from (4.10) to get
(5.19) |
by partial summation as in (5.18). We divide the second range into dyadic intervals . Applying Proposition 5.1 with and from (4.10), we get
(5.20) |
Next we sum over dyadic intervals. For the first term , when
and when
So after summing up from (5.20), recalling in Setting 3.1, using and (5.3), we have
(5.21) |
Combining (5.19) and (5.21) we have
(5.22) |
5.2. On the case
Recall the remark after Proposition 5.2. Let denote the Fourier coefficients of an orthonormal basis of . For each singular cusp of , let be the associated Eisenstein series in weight . Let be defined as in (3.3) associated with for .
Recall the definition of the Maass lowering operator in (2.14) and (2.16) for . By [DFI02, (4.52)] (where they used for the lowering operator and ), the set
Combining [DFI02, (4.36), (4.27) and the last equation of p. 502], for and , since
the Fourier coefficient of satisfies
(5.23) |
and then
(5.24) |
where the bound is from Proposition 3.4.
Moreover, by [DFI02, (4.48)], if is the Eisenstein series defined in weight , then
We also get
(5.26) |
We have the following proposition:
Proposition 5.4.
With the same setting as Theorem 1.3 for , when , we have
The proof that Proposition 5.4 implies Theorem 1.3 in the case is the same as the case of weight before. This is because for (5.25) and because (5.7), (5.8) and (5.9) still hold for (the process only involves estimates on with some applications of Proposition 5.2 in weight ). In the rest of this subsection we prove Proposition 5.4.
First we show that the main terms corresponding to are the same when we shift the weight between and . Recall . Let denote the corresponding coefficients for in weight :
where is defined at the beginning of this subsection.
Recall that the definition on (3.5) is for weight and here we use for weight . We derive
(5.28) |
by the same process as we derive (5.10) above. Since when , we have (with (2.16) and chosen later) by Proposition 3.4 and still get
(5.29) |
The first sum above can be estimated similarly by condition (2) of Definition 1.1 as
(5.30) |
By Theorem 3.3,
The bound for is done in (5.15). Estimates for the remaining part of follow from the same process as §5.1.2 in the case of weight , taking (5.24) and (5.26) into account. For the same reason as the beginning of §5.1.2, we just record the bounds with respect to the discrete spectrum here.
For , we apply Proposition 5.2, from (5.24), and from (4.10). Since
(5.32) |
by Cauchy-Schwarz, with the help of (5.4) we have
(5.33) |
We still let
and divide into two parts: and . In the first range, we apply Proposition 5.2, (5.24) and from (4.10) to get
(5.34) |
by partial summation similar as (5.33). We divide the second range into dyadic intervals and apply Proposition 5.1, (5.24) and from (4.10):
(5.35) |
Summing up from (5.35) similar as we did after (5.20) and recalling in Setting 3.1, we have
(5.36) |
From (5.34) and (5.36) we have
(5.37) |
Acknowledgement
The author thanks Professor Scott Ahlgren for his plenty of helpful discussions and suggestions.
References
- [AA16] Scott Ahlgren and Nickolas Andersen. Algebraic and transcendental formulas for the smallest parts function. Adv. Math., 289:411–437, 2016.
- [AA18] Scott Ahlgren and Nickolas Andersen. Kloosterman sums and Maass cusp forms of half integral weight for the modular group. Int. Math. Res. Not. IMRN, 2018(2):492–570, 2018.
- [AD19] Scott Ahlgren and Alexander Dunn. Maass forms and the mock theta function . Math. Ann., 374(3-4):1681–1718, 2019.
- [AD20] Nickolas Andersen and William D. Duke. Modular invariants for real quadratic fields and Kloosterman sums. Algebra Number Theory, 14(6):1537–1575, 2020.
- [AD22] Nickolas Andersen and William Duke. Asymptotic distribution of traces of singular moduli. Discrete Anal., 4:1–14, 2022.
- [AH91] George E. Andrews and Dean Hickerson. Ramanujan’s “lost” notebook VII: The sixth order mock theta functions. Adv. Math., 89:60–105, 1991.
- [Blo08] Valentin Blomer. Sums of Hecke eigenvalues over values of quadratic polynomials. Int. Math. Res. Not. IMRN, 2008:rnn 059, 2008.
- [BO06] Kathrin Bringmann and Ken Ono. The mock theta function conjecture and partition ranks. Invent. Math., 165(2):243–266, 2006.
- [BO12] Kathrin Bringmann and Ken Ono. Coefficients of harmonic Maass forms. In Krishnaswami Alladi and Frank Garvan, editors, Partitions, q-Series, and Modular Forms, Proceedings of the Conference Held at the University of Florida, Gainesville, FL, March 2008. Developments in Mathematics, vol. 23, pages 23–38. Springer New York, NY, 2012.
- [DFI02] W. Duke, J.B. Friedlander, and H. Iwaniec. The subconvexity problem for Artin -functions. Invent. Math., 149(3):489–577, September 2002.
- [DLMF] NIST Digital Library of Mathematical Functions. http://dlmf.nist.gov/, Release 1.0.25 of 2019-12-15. F. W. J. Olver, A. B. Olde Daalhuis, D. W. Lozier, B. I. Schneider, R. F. Boisvert, C. W. Clark, B. R. Miller, B. V. Saunders, H. S. Cohl, and M. A. McClain, eds.
- [Dun90] T. Mark Dunster. Bessel functions of purely imaginary order, with an application to second-order linear differential equations having a large parameter. SIAM J. Math. Anal., 21(4):995–1018, 1990.
- [Dun18] Alexander Dunn. Uniform bounds for sums of Kloosterman sums of half integral weight. Res. Number Theory, 4(4):Art. 45, 21, 2018.
- [GS83] Dorian Goldfeld and Peter Sarnak. Sums of Kloosterman sums. Invent. Math., 71(2):243–250, 1983.
- [Iwa87] Henryk Iwaniec. Fourier coefficients of modular forms of half-integral weight. Invent. Math., 87:385–401, 1987.
- [Kim03] Henry H. Kim. Functoriality for the exterior square of and the symmetric fourth of . J. Amer. Math. Soc., 16(1):139–183, 2003. With appendix 1 by D. Ramakrishnan and appendix 2 by Kim and P. Sarnak.
- [Kno70] Marvin I. Knopp. Modular functions in analytic number theory. Markham Publishing Co., Chicago, 1970.
- [Lan00] L. J. Landau. Bessel functions: monotonicity and bounds. J. London Math. Soc., 61(2):197–215, 2000.
- [Olv97] Frank W. J. Olver. Asymptotics and Special Functions. AKP classics. A K Peters, Wellesley, Massachusetts, 1997.
- [Pri00] Wladimir de Azevedo Pribitkin. A generalization of the Goldfeld-Sarnak estimate on Selberg’s Kloosterman zeta-function. Forum Math., 12(4):449–459, 2000.
- [Pro79] N. V. Proskurin. The summation formulas for general Kloosterman sums. Zap. Nauchn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI), 82:103–135, 1979.
- [Pro05] N. V. Proskurin. On the general Kloosterman sums. J. Math. Sci., 129(3):3874–3889, 2005.
- [Rad73] Hans Rademacher. Topics in analytic number theory. Springer-Verlag, New York-Heidelberg, 1973. Edited by E. Grosswald, J. Lehner and M. Newman, Die Grundlehren der mathematischen Wissenschaftern, Band 169.
- [Ran77] Robert A. Rankin. Modular Forms and Functions. Cambridge University Press, 1977.
- [Sel65] Atle Selberg. On the estimation of Fourier coefficients of modular forms. In Proc. Sympos. Pure Math., Vol. VIII, pages 1–15. Amer. Math. Soc., Providence, R.I., 1965.
- [ST09] Peter Sarnak and Jacob Tsimerman. On Linnik and Selberg’s conjecture about sums of Kloosterman sums. In Yuri Tschinkel and Yuri Zarhin, editors, Algebra, Arithmetic, and Geometry: Volume II: In Honor of Yu. I. Manin, pages 619–635, Boston, 2009. Birkhäuser Boston.
- [Sun23] Qihang Sun. Uniform bounds for Kloosterman bounds of half-integral weight with applications. https://arxiv.org/abs/2305.19651, 2023.
- [Wai17] Fabian Waibel. Fourier coefficients of half-integral weight cusp forms and Waring’s problem. Ramanujan J., 47:185–200, 2017.