Character sum, reciprocity and Voronoi formula
Abstract.
We prove a novel four-variable character sum identity which serves as a twisted, non-archimedean counterpart to Weber’s integrals for Bessel functions. Using this identity and ideas from Venkatesh’s thesis, we present a new, spectral proof of the Voronoi formula for classical modular forms.
Key words and phrases:
Modular form, automorphic form, -function, functional equation, Voronoi formula, Petersson trace formula, character sum, Kloosterman sum, Gauss sum, spectral identity, Analytic Number TheoryMathematics Subject Classification:
11F11, 11F12, 11F66, 11F72, 11L051. Introduction
1.1. Character sums and modular forms
Character sums of various forms have been instrumental in the development of number theory. A notable example is Kloosterman’s 1926 work on the circle method and representation of natural numbers by quadratic forms, where the following character sum, now known as the Kloosterman sum, proved fundamental:
(1.1) |
where are integers and . His method can be applied more generally to estimate the sizes of Fourier coefficients of holomorphic modular forms, providing improvements over Hecke’s bound; see [25, Chp. 20.3] and [6, Sect. 3.4]. The relationship between the Kloosterman sum and modularity can also be understood through trace formulae after Petersson and Selberg; see [25, Chp. 14].
Another intriguing connection was revealed by Kuznetsov [27, Thm. 4 and Sect. 7.1] while developing his celebrated trace formula. He proved the remarkable identity
(1.2) |
by applying Hecke’s multiplication rule
(1.3) |
to his formula, where denotes the normalized Hecke eigenvalues of a cuspidal eigenform of level . Historically, (1.2) was discovered by Selberg in the 1930s, but no proof was provided. As noted in [8, p. 2318], a direct proof of (1.2) has resisted straightforward attempts.
Identities of character sums also serve as essential local inputs in the arithmetic and analytic aspects of modular forms (or more generally, automorphic forms) via trace formulae. For instance, Iwaniec [24] obtained an averaged form of Waldspurger/Kohnen–Zagier’s formula through a delicate identity relating Kloosterman sums to Salié sums, which can be considered as a non-archimedean analogue of the integral identities of Bessel functions due to Hardy and Weber (see [24, Lem. 4], [5, Thm. 1.1, Cor. 8.4, Sects. 18 and 22] and [12]). As another example, Duke–Iwaniec [14] discovered a curious identity between the Kloosterman sum and a cubic exponential sum, which plays an important role in the cubic liftings of [30, 17]. This identity is a non-archimedean version of Nicholson’s identity for the Bessel and Airy functions.
In this work, we prove a new reciprocity identity of a four-variable character sum associated with a Dirichlet character . This result is the intrinsic form of an identity of Kloosterman sums (Prop. 1.5) which should be interpreted as a twisted, non-archimedean counterpart of another Weber’s integral identity (eq. (1.15)).
Theorem 1.1.
Let , be integers such that and . 111 The notation refers to for some . Let be a Dirichlet character. Define
(1.4) |
Then we have
(1.5) |
1.2. Global results
There has been considerable interest in deriving the analytic continuation and functional equations of -functions associated with modular forms directly from their Dirichlet series, avoiding the traditional reliance on integral representations as in the cases of Hecke, Rankin and Selberg. In his seminal 1977 proceedings [39, pp. 141–142], Zagier suggested an approach based on the property that the set of holomorphic Poincaré series spans the linear space of holomorphic cusp forms. This idea was later implemented in two different cases by Goldfeld [19] and Mizumoto [32]. The method has since found various arithmetic applications as explained in e.g., Goldfeld–Zhang [20], Nelson [33], and Diamantis–O’Sullivan [13].
In his thesis [35, Chp. 3], Venkatesh gave an elegant new proof of the classical converse theorem for holomorphic modular forms of level , which may be viewed as a limiting variant of [19, Sect. 2] (cf. [11]). Several aspects of his approach are distinctive both conceptually and technically. Venkatesh circumvented the need for the aforementioned property of the holomorphic Poincaré series with a very neat argument. In fact, his method works equally well for non-holomorphic modular forms by utilizing the Kuznetsov trace formula in place of the Petersson trace formula which he employed in the holomorphic case. Another idea from [35] is to begin with Dirichlet polynomials instead of working directly with -functions. Recent works of Booker–Farmer–Lee [11] and Blomer–Leung [10] have yielded more general converse theorems using this approach.
In this article, we employ Venkatesh’s ideas to provide new, spectral proofs of the functional equation for the twisted Hecke -function and the Voronoi formula for holomorphic cusp forms 222attributed to Wilton in [31, Sect. 3]. , in line with the philosophy of “beyond endoscopy” (see [36]). We derive these results from a spectral identity for averages of smooth Dirichlet polynomials (Thm. 1.2), where Thm. 1.1 plays a significant role. Our work also uses the circle of techniques under “spectral reciprocity” (e.g., [7, 8]), but with several important variations as we will discuss in Sect. 1.3.
Our main results are described as follows. Let be an integer, , , and be a primitive Dirichlet character. The Gauss sum associated with is defined by
(1.6) |
Suppose is an orthogonal basis of holomorphic cusp forms of level and weight , for which the Fourier expansion of is given by
(1.7) |
where , with and . The harmonic weighting refers to
For , the twisted Hecke -function associated with and is defined by the Dirichlet series
(1.8) |
Denote by the -Bessel function, and let
(1.9) |
Theorem 1.2.
For any and , we have
(1.10) |
In Sect. 7, we isolate a single cusp form from (1.10), which requires a careful discussion on the analytic continuation and polynomial growth of ; see Sect. 1.3.5 and App. A. Hence, we have
Corollary 1.3.
Let and be a primitive Dirichlet character. Then for any ,
(1.11) |
Our argument for Thm. 1.2 applies to additive twists with very minor adjustments.
Corollary 1.4.
Let , , and . Then for , we have
(1.12) |
Alternatively, one may deduce Cor. 1.4 from Cor. 1.3 using [26, Thm. 1.3]. We refer the reader to [34], [31], [4] for other recent approaches to proving the classical Voronoi formula (1.12).
Remark 1.
We are aware of an article by Herman [23]. After a thorough review, we identified major gaps and inaccuracies in his approach, which we rectified in our work [28]. In addition, our work provided generalizations and new observations. Readers are referred to [28, Sect. 1.6] and [29] for a detailed comparison between the two works.
Remark 2.
Generalizing our main results to cusp forms of level with nebentypus (say primitive) should not pose significant difficulty provided that , apart from the need of more involved notations. This can be achieved by combining the arguments for Thm. 1.2 with that for [28, Thm. 1.2].
This work and [28] are both complementary and independent in interest, as they address different types of ramifications which require entirely different treatment of the character sums. Notably, Thm. 1.1 is unique to this paper, and the bulk of our argument, spanning Sect. 3.2 to Sect. 6, is distinct from [28]; see also our comments in Sect. 1.3.3 and Sect. 1.3.4.
Remark 3.
The manipulations of the character sums, which is the focus of this paper, carry over to the non-holomorphic case. In the non-holomorphic case, it suffices to apply the Kuznetsov trace formula and adapt the argument for Prop. A.1 to account for the different weight functions.
1.3. A road map for Theorem 1.2
Although we initially represent the -functions in terms of their Dirichlet series, the essential components of our argument are inherently local; see Thm. 1.1, Prop. 1.5 and eq. (1.15). In Sects. 3.1–3.2, we apply the Petersson formula and Poisson summation to the left-hand side of (1.10), getting
(1.13) |
where , and
(1.14) |
1.3.1. Two dualities
The adelic viewpoint suggests that serves as the archimedean counterpart of . Correspondingly, our argument hinges on two intriguing geometric dualities, one for the Fourier–Hankel transform of and the other for the finite Fourier transform .
The first duality is commonly known as Weber’s second exponential integral identity (Lem. 2.4):
(1.15) |
The Hankel inversion formula (Lem. 2.3) is a limiting form of (1.15). In Sect. 4.1, we will apply Hankel’s formula and Weber’s identity successively. The second duality is a twisted, non-archimedean analogue of (1.15):
Proposition 1.5.
Let with and be a primitive character. Then
(1.16) |
1.3.2. Multiplicative (analytic–arithmetic) cancellation
The next step is to apply Poisson summation to the -sum in (1.13), though an observant reader might question its applicability due to the apparent singularity at . However, the factor of (1.13) is cancelled out upon inserting (1.16) and (1.15) (with appropriately chosen )! Additionally, the factor from (1.16) perfectly cancels the final exponential factor in (1.15), paving the way for the two upcoming steps.
This technical feature is also crucial in previous works on beyond endoscopy, where the Selberg trace formula was analyzed and the singularities from the archimedean orbital integrals are more subtle (see [3, Sect. 4], [21], [16, Sect. 6]). Once again, smoothing of these singularities is necessary for Poisson summation, but this requires delicate use of an approximate functional equation and the class number formula.
1.3.3. Three-fold reciprocity and local analysis
In (1.14) (cf. Sect. 3.3), we apply a first additive reciprocity:
(1.17) |
In Sect. 4.2, we split the - and -sums in (1.13) appropriately. One roughly gets:
(1.18) |
where is the character sum defined in Thm. 1.1. See (4.2) for the precise expressions.
The character sum (1.4) exhibits twisted multiplicativity (Lem. 4.1), enabling a local analysis. Thm. 1.1 follows from the method of -adic stationary phase ([25, Chp. 12.3]) to be carried out in Sect. 5, along with a second use of (1.17); see (5.2). Our proof of (1.5), and hence Thm. 1.2, is non-trivial even when is a prime.
A third use of (1.17) allows us to combine the exponential phase of (1.5) with that of (1.18), then a second copy of arises upon opening up by its definition (1.4) and gluing variables back together; see (6.5) of Sect. 6. Now, Prop. 1.5 follows from this. The factor from (1.16) will become the root number of (1.11).
1.3.4. Additive cancellations and back to the global aspect
There are two subtle yet crucial cancellations of terms when proving Thm. 1.2. Namely, the diagonal term and dual zeroth frequency from the first applications of Petersson formula and Poisson summation cancel perfectly with the zeroth term (on the physical side) and diagonal term from the second applications of Poisson summation and Petersson formula, respectively. The first cancellation uses the Hankel inversion formula and the fact that .
1.3.5. Remarks on analytic continuation
Obtaining analytic continuation to , for some absolute , is crucial as a prerequisite for discussing the functional equation. However, this is non-trivial for methods based on Dirichlet series. For example, using his refined analysis of the Eichler–Selberg trace formula [3, 1], Altuğ established holomorphy for the standard -function of Ramanujan’s -function only in ; see [2, Cor. 1.2]. Other methods also fall short of reaching , as illustrated in [38], [18], [35, Prop. 10], [22], and [15].
In [35, Chp. 2.6], Venkatesh had the idea of embedding the holomorphic cusp forms of weight and level into the full spectrum of -automorphic forms of level , consisting of Maass cusp forms of weight , holomorphic cusp forms of all weights and the Eisenstein series. Thus, an arbitrary smooth compactly test function on can be put on the geometric side of the trace formula, i.e., using the arithmetic Kuznetsov (or Petersson–Kuznetsov) trace formula (see [25, Thm. 16.5]). This simplifies the analysis, reducing it to handling a basic Fourier integral. For some spectral test functions , one has
(1.19) |
The goal is to deduce an entire continuation of from (1.19).
However, the space of admissible spectral test functions is somewhat limited and the technical challenge lies in constructing a test function in this space that “effectively” isolates a specific part of the spectrum. This problem was addressed in [35, Chp. 6.3] by an elaborate analysis of the Sears–Titchmarsh transform. In [35, p. 50], an extra technical assumption concerning the growth in spectral parameters (Laplace eigenvalues/weights) is needed (cf. [35, eq. (2.35)] and [38, Rmk. 3.17]).
1.4. Acknowledgement
The research is supported by the EPSRC grant: EP/W009838/1.
2. Preliminaries
Let and . The Fourier transform of and the Mellin transform of are
respectively. Their inversion formulae, which hold whenever the integrals converge absolutely, are
(2.1) |
For , the -Bessel functions can be defined by the series
(2.2) |
which converges pointwise absolutely and uniformly on compact subsets of . From [37, Sect. p. 192], they admit the following absolutely convergent Mellin–Barnes representation:
Lemma 2.1.
Let and be defined in (1.9). Then
(2.3) |
Let and . Then the Hankel transform of is defined by
(2.4) |
The rapid decay of can be deduced by integrating-by-parts many times in (2.4) using
(2.5) |
where is a smooth function satisfying for any ; see [37, p. 206].
Lemma 2.2.
For and we have
(2.6) |
This follows directly from the recurrence , and the estimate
(2.7) |
Lemma 2.3 (Hankel inversion formula).
For any , we have
(2.8) |
Lemma 2.4.
Let , , and . Then
(2.9) |
Remark 4.
While Bessel functions with positive arguments are more common in the analytic theory of automorphic forms (e.g., (1.12) or (2.11)), it will be technically convenient to invoke those with complex arguments as intermediates in our case. Also, the integral identity (2.9) is a variant of (1.15), where the latter only converges conditionally.
Lemma 2.5.
[9, Lem. 8.1] Let be a real-valued function and . Suppose there exist such that for any , we have for any , for any , and . Then for any , we have
(2.10) |
Lemma 2.6 (Petersson trace formula).
Let be an orthogonal basis of holomorphic cuspidal Hecke eigenforms of level and weight . For any , we have 333 In this article, we use to denote the indicator function with respect to the condition .
(2.11) |
Lemma 2.7 (Poisson summation).
Let and . For , and be -periodic,
(2.12) |
3. Setting the stage: Petersson–Poisson–Reciprocity
Sects. 3–6 are devoted to proving Thm. 1.2. Let be a given test function. Define
(3.1) |
Our analysis of this article applies to all even integers , though it is well-known by the Riemann–Roch theorem that there is no non-zero holomorphic cusp form of level and weight .
3.1. Step 1: Petersson trace formula
3.2. Step 2: Poisson summation
We apply Lem. 2.7 to the -sum of (3.2), which gives
The -sum can be decomposed via with and , i.e.,
(3.3) |
where the last line follows from the primitivity of ; see [25, eq. (3.12)]. Hence, we get
(3.4) |
We split the -sum of (3.4) by , where and . In particular, we have
(3.5) |
Notice that the dual zeroth frequency (i.e., the term ) of (3.4) is non-vanishing only when and . In this case, the -sum is given by
We extract the zeroth frequency from the rest of the frequencies, resulting in the expression:
(3.6) |
where
(3.7) |
3.3. Step 3: Additive reciprocity.
4. Step 4: Preparation on the -sums
To prepare for Poisson summation in the -sums and for swapping the roles of and at a later stage, two crucial components are needed:
-
•
(Analytic) Transform the -sum with in (3.3) to a sum over all integers, and remove the singularity at by suitable analytic manipulations;
-
•
(Arithmetic) Transform the character sum in (3.3) into a suitable form that facilitates the eventual combination of - and -sum.
The second point requires a fairly intricate analysis of character sums, see Sects. 4.2–6.1.
4.1. Step 4.1: Analytic preparation
The -sum of (3.3) can be rewritten to sum over all non-zero integers. Indeed, observe that the contribution from in (3.3) is given by
(4.1) |
upon making the changes of variables and . The change of variables and the fact that (follows from (2.2) and ) allow us to rewrite (4.1) as:
Hence, the expression (3.3) becomes:
(4.2) |
The last integral converges absolutely since . By the dominated convergence theorem,
The Hankel inversion formula (Lem. 2.3) with the change of variables give
(4.3) |
where the interchange of the order of integration is permitted by absolute convergence and the decay of the Hankel transform of . Now, we are in a position to apply Lem.2.4. In other words,
(4.4) |
By the decay of and dominated convergence again, it follows from continuity that
(4.5) |
Substitute the last expression back into (4.1) and observe the cancellations in (1). a pair of exponential phases, and (2). a pair of factors . Thus, we obtain
(4.6) |
4.2. Step 4.2: Arithmetic preparation
We now analyze the arithmetic component of (4.1), i.e.,
(4.8) |
For technical convenience, we consider a more general character sum define as follows. Let be integers such that and . Let be a primitive character mod . Define
(4.9) |
We decouple the -sum of (4.1) by (as before), where and . Then (4.8) is equal to and
(4.10) |
We have the following twisted multiplicativity for the character sum (4.9).
Lemma 4.1.
Let be integers such that , , and . Let and be primitive characters mod and respectively. Then
Proof.
By the Chinese remainder theorem, we have
Applying the change of variables and yields the first equality. To get the second equality, apply the change of variables and instead. ∎
Remark 5.
This lemma explains why the factor is purposefully left out when defining . Otherwise, no multiplicative relation would hold as the factor is not multiplicative in , .
5. Local analysis of character sums: proof of Theorem 1.1
Restatement of Theorem 1.1.
Let be integers such that and . Let be a character mod . Then we have
Proof.
We first consider the case when , where is a prime and . Since , we may write and for some . In this, we have
Claim: Unless one of the following holds: (1) , (2) , or (3) , we have
Indeed, the congruence condition for implies that it is equal to unless one of the following holds: , (2) or (3). If , then
and hence unless (1) or (2) or (3) holds. Similarly, the congruence condition implies that unless or (2) or (3), and the presence of implies that it is also if . This proves our claim.
We apply a change of variable to obtain
Case 1: . In this case, we have
Notice that since , we have . Hence, the change of variables is admissible. When , the quantity is well-defined due to the presence of . With the above change of variables , it follows that
Hence,
With another change of variables , we get
Case 2: . In this case, we have
Apply the change of variables and observe the fact that , we have
Another change of variables , which is admissible as , yields
Using
we arrive at
A final change of variables implies
Case 3: . In this final case, we make use of Case 2 to deduce the answer. For , we have
and this yields
Combining all three cases together, we have proved that for for some prime and ,
(5.1) |
for any with , and any character mod . Finally, applying the first and second equality of Lem. 4.1 to the left-hand and right-hand side of (5.1) respectively, with the observation:
(5.2) |
we see that both sides satisfy the same multiplicative relations. This yields the result for general . ∎
6. Backward maneuver: Finishing the Proof of Theorem 1.2
6.1. Step 5: Reciprocity and recombining sums
6.2. Step 6: Second Poisson
The treatment of is simpler. Observe that
A change of variables and the Hankel inversion formula (Lem. 2.3) imply that
(6.4) |
For , the character sum can be expressed in terms of Kloosterman sums via (3.5) and (3.2):
(6.5) |
Note: . Plugging (6.5) into and making a change of variables , we arrive at
We now readily observe the role reversal of the -sum and -sum as discussed in Sect. 1.3.1!
The bounds for the Hankel transform and Bessel functions recorded in Sect. 2 allow us to apply the dominated convergence theorem and obtain
where is a smooth function on such that on , on , and on . Inside the limit, we apply Poisson summation to the -sum. By dominated convergence again,
(6.6) |
upon taking the limit . Inserting (6.4)–(6.6) into (6.1)–(6.3), we readily observe that
(6.7) |
by combining the -sum and -sum.
6.3. Step 7: Petersson in reverse
Inserting (6.7) back into (3.6) yields
(6.8) |
Applying the Petersson formula (2.11) in reverse to the -sum, Thm. 1.2 follows.
Remark 6.
Readers should note that the diagonal term of (3.2) from the initial application of the Petersson trace formula conveniently cancelled with the dual zeroth frequency from the Poisson summation. Only after this does the crucial “role-reversal” of sums takes place.
7. Proof of Corollary 1.3
In App. A, we show that the -series (1.8) can be analytically continued to , and
(7.1) |
has polynomial growth on the vertical strip and as . Granting this (i.e., Prop. A.1), we prove Cor. 1.3 as follows. Pick any . Then its Mellin transform is entire and it follows from repeated integration by parts that for any ,
(7.2) |
Let , and . In (3.1), apply (2.1) and rearrange sums and integrals, we have
By Prop. A.1 and (7.2), we may shift the line of integration above to .
Next, it follows from (2.3) and (2.1) that
(7.3) |
Using (7.2), (A.3) and the holomorphy of , we may shift the line of integration to in (7.3). Inserting the resultant into the right-hand side of (1.10), we deduce that
Upon exchanging the order of sums and integrals, we find that the -sum converges absolutely and is precisely the Dirichlet series . In other words,
(7.4) |
By Prop. A.1, (7.2) and (A.3), we may shift the line of integration for (7.4) back to . As a result, we have
(7.5) |
Let . 444 Here, we use the finite dimensionality of the linear space of holomorphic cusp forms of a given weight and level. The vectors for are linearly independent over , and thus, there exists such that the submatrix is invertible. By Prop. A.1, the function admits a holomorphic continuation (say ) to the region for :
(7.6) |
Hence, each of admits a holomorphic continuation to the same region.
Since (1.10) holds for any and any , we have, on the vertical strip ,
Since is invertible, the functional equation (1.11) for on follows immediately.
From the holomorphic continuation of to and the functional equation just proved, also extends holomorphically to . As a result, admits an entire continuation and now the functional equation holds for all . This completes the proof.
Appendix A Analytic continuation and polynomial growth
The proof of Prop. A.1 below is similar to that in [28], but requires several adjustments to accommodate our current setting. For completeness and convenience of the reader, we supply a short argument as follows.
Proposition A.1.
Let and be integers. Then the function defined in (7.1) admits a holomorphic continuation to the half-plane and satisfies the estimate
(A.1) |
for any with and . The implicit constant depends only on .
Proof.
Take such that for any . Inserting this into (1.8) for , observe that (7.1) can be written as
(A.2) |
and . For , we have , and from (3.6)–(3.2), it follows that
Bounding the second summand on the right-hand side is harder, and this will be our focus.
We split the -sum above into two parts, according to the conditions and . These two parts are denoted by and respectively.
By Lem. 2.1, we have
where . Then integrating by parts times, it follows that the above expression is
Let and with . Then Stirling’s formula gives
(A.3) |
For any and , we have
(A.4) |
Hence, if and such that , then
(A.5) |
(A.6) |
Next, we estimate with . By (2.5), it suffices to estimate the oscillatory integral:
To this end, we make use of Lem. 2.5 with the functions
Suppose and . We observe the following bounds:
(A.7) |
where Leibniz’s rule and Faà di Bruno’s formula were used. Apply Lem. 2.5 with the parameters:
(A.8) |
we conclude that:
(A.9) |
for , , , where the implicit constants depend on . When , we have
(A.10) |
As a result, we obtain the estimate:
provided that , with , , and . By taking , , with , we have
(A.11) |
On the strip , the same estimates hold with the implicit constants depend only on .
References
- [1] S. Altuğ “Beyond endoscopy via the trace formula, II: Asymptotic expansions of Fourier transforms and bounds towards the Ramanujan conjecture” In Amer. J. Math. 139.4, 2017, pp. 863–913
- [2] S. Altuğ “Beyond endoscopy via the trace formula—III The standard representation” In J. Inst. Math. Jussieu 19.4, 2020, pp. 1349–1387
- [3] Salim Ali Altuğ “Beyond endoscopy via the trace formula: 1. Poisson summation and isolation of special representations” In Compos. Math. 151.10, 2015, pp. 1791–1820
- [4] Edgar Assing and Andrew Corbett “Voronoï summation via switching cusps” In Monatsh. Math. 194.4, 2021, pp. 657–685
- [5] Ehud Moshe Baruch and Zhengyu Mao “Bessel identities in the Waldspurger correspondence over the real numbers” In Israel J. Math. 145, 2005, pp. 1–81
- [6] Valentin Blomer and Farrell Brumley “The role of the Ramanujan conjecture in analytic number theory” In Bull. Amer. Math. Soc. (N.S.) 50.2, 2013, pp. 267–320
- [7] Valentin Blomer and Rizwanur Khan “Twisted moments of -functions and spectral reciprocity” In Duke Math. J. 168.6, 2019, pp. 1109–1177
- [8] Valentin Blomer and Rizwanur Khan “Uniform subconvexity and symmetry breaking reciprocity” In J. Funct. Anal. 276.7, 2019, pp. 2315–2358
- [9] Valentin Blomer, Rizwanur Khan and Matthew Young “Distribution of mass of holomorphic cusp forms” In Duke Math. J. 162.14, 2013, pp. 2609–2644
- [10] Valentin Blomer and Wing Hong Leung “A GL(3) converse theorem via a "beyond endoscopy” approach”, 2024 arXiv:2401.04037 [math.NT]
- [11] Andrew R. Booker, Michael Farmer and Min Lee “An extension of Venkatesh’s converse theorem to the Selberg class” In Forum Math. Sigma 11, 2023, pp. Paper No. e26\bibrangessep10
- [12] Jingsong Chai and Zhi Qi “Bessel identities in the Waldspurger correspondence over the complex numbers” In Israel J. Math. 235.1, 2020, pp. 439–463
- [13] Nikolaos Diamantis and Cormac O’Sullivan “Kernels of -functions of cusp forms” In Math. Ann. 346.4, 2010, pp. 897–929
- [14] W. Duke and H. Iwaniec “A relation between cubic exponential and Kloosterman sums” In A tribute to Emil Grosswald: number theory and related analysis 143, Contemp. Math. Amer. Math. Soc., Providence, RI, 1993, pp. 255–258
- [15] W. Duke and H. Iwaniec “Convolution -series” In Compos. Math. 91.2, 1994, pp. 145–158
- [16] Melissa Emory, Malors Espinosa-Lara, Debanjana Kundu and Tian An Wong “Beyond Endoscopy via Poisson Summation for GL(2,K)”, 2024 arXiv:2404.10139 [math.NT]
- [17] Solomon Friedberg and Omer Offen “On the Cubic Shimura lift to : The Fundamental Lemma”, 2022 arXiv:2202.01247 [math.NT]
- [18] Satadal Ganguly and Ramdin Mawia “Rankin-Selberg -functions and “beyond endoscopy”” In Math. Z. 296.1-2, 2020, pp. 175–184 DOI: 10.1007/s00209-019-02431-5
- [19] Dorian Goldfeld “Analytic and arithmetic theory of Poincaré series” Luminy Conference on Arithmetic In Astérisque, 1979, pp. 95–107
- [20] Dorian Goldfeld and Shouwu Zhang “The holomorphic kernel of the Rankin-Selberg convolution” In Surveys in differential geometry 7, Surv. Differ. Geom. Int. Press, Somerville, MA, 2000, pp. 195–219
- [21] Oscar E. González et al. “On smoothing singularities of elliptic orbital integrals on and beyond endoscopy” In J. Number Theory 183, 2018, pp. 407–427
- [22] P. Herman “Beyond endoscopy for the Rankin-Selberg -function” In J. Number Theory 131.9, 2011, pp. 1691–1722
- [23] P. Herman “The functional equation and beyond endoscopy” In Pacific J. Math. (Rogawski Memorial Volume) 260.2, 2012, pp. 497–513
- [24] H. Iwaniec “On Waldspurger’s theorem” In Acta Arith. 49.2, 1987, pp. 205–212
- [25] Henryk Iwaniec and Emmanuel Kowalski “Analytic number theory” 53, American Mathematical Society Colloquium Publications American Mathematical Society, Providence, RI, 2004
- [26] Eren Kıral and Fan Zhou “The Voronoi formula and double Dirichlet series” In Algebra & Number Theory 10.10 Mathematical Sciences Publishers, 2016, pp. 2267–2286
- [27] N.. Kuznecov “The Petersson conjecture for cusp forms of weight zero and the Linnik conjecture. Sums of Kloosterman sums” In Mat. Sb. (N.S.) 111(153).3, 1980, pp. 334–383
- [28] Chung-Hang Kwan and Wing Hong Leung “Trace formula and functional equation”, 2024 arXiv:2410.16921 [math.NT]
- [29] Chung-Hang Kwan and Wing Hong Leung “Discussions of Herman’s “The functional equation and Beyond Endoscopy”” URL: https://drive.google.com/file/d/1SymUkyOuMHhQ1CMVAgkQEKHLCTvsd9gi/view?usp=sharing
- [30] Zhengyu Mao and Stephen Rallis “On a cubic lifting” In Israel J. Math. 113, 1999, pp. 95–124
- [31] Stephen D. Miller and Wilfried Schmid “Summation formulas, from Poisson and Voronoi to the present” In Noncommutative harmonic analysis, Progr. Math. 220 Birkhäuser, 2004, pp. 419–440
- [32] Shin-ichiro Mizumoto “On the second -functions attached to Hilbert modular forms” In Math. Ann. 269.2, 1984, pp. 191–216
- [33] Paul D. Nelson “Stable averages of central values of Rankin-Selberg -functions: some new variants” In J. Number Theory 133.8, 2013, pp. 2588–2615
- [34] Nicolas Templier “Voronoĭ summation for ” In Representation theory, automorphic forms & complex geometry Int. Press, Somerville, MA, [2019] ©2019, pp. 163–196
- [35] Akshay Venkatesh “Limiting forms of the trace formula” Thesis (Ph.D.)–Princeton University ProQuest LLC, Ann Arbor, MI, 2002, pp. 168
- [36] Akshay Venkatesh ““Beyond endoscopy” and special forms on GL(2)” In J. Reine Angew. Math. 577, 2004, pp. 23–80
- [37] G.. Watson “A treatise on the theory of Bessel functions” Reprint of the second (1944) edition, Cambridge Mathematical Library Cambridge University Press, Cambridge, 1995
- [38] Paul-James White “The base change -function for modular forms and Beyond Endoscopy” In J. Number Theory 140, 2014, pp. 13–37
- [39] D. Zagier “Modular forms whose Fourier coefficients involve zeta-functions of quadratic fields” In Modular functions of one variable, VI (Proc. Second Internat. Conf., Univ. Bonn, Bonn, 1976) 627, Lecture Notes in Math. Springer, Berlin-New York, 1977, pp. 105–169