The Voronoi Summation Formula for and the Godement-Jacquet Kernels
Abstract.
Let be the ring of adeles of a number field and be an irreducible cuspidal automorphic representation of . In [JL22, JL23], the authors introduced -Schwartz space and -Fourier transform with a non-trivial additive character of , proved the associated Poisson summation formula over , based on the Godement-Jacquet theory for the standard -functions , and provided interesting applications. In this paper, in addition to the further development of the local theory, we found two global applications. First, we find a Poisson summation formula proof of the Voronoi summation formula for over a number field, which was first proved by A. Ichino and N. Templier ([IT13, Theorem 1]). Then we introduce the notion of the Godement-Jacquet kernels and their dual kernels for any irreducible cuspidal automorphic representation of and show in Theorems 6.9 and 6.13 that and are related by the nonlinear -Fourier transform if and only if is a zero of , the finite part of the standard automorphic -function , which are the -versions of [Clo22, Theorem 1.1], where the Tate kernel with and the trivial character are considered.
Key words and phrases:
Poisson Summation Formula, Voronoi Summation Formula, Bessel Function, Generalized Schwartz Space, Non-Linear Fourier Transform/Hankel Transform, Global Zeta Integral, Godement-Jacquet Kernel, Automorphic -function2010 Mathematics Subject Classification:
Primary 11F66, 22E50, 43A32; Secondary 11F70, 22E53, 44A201. Introduction
1.1. -Poisson summation formula
Let be a number field and the associated ring of adeles. For any irreducible cuspidal automorphic representation of for any integer , the Godement-Jacquet theory ([GJ72]) establishes the analytic continuation and functional equation for the standard -functions . The key input from the harmonic analysis to the Godement-Jacquet theory is the classical Fourier analysis on the affine space , the space of -matrices, and the associated Poisson summation formula. From the point of view in the Braverman-Kazhdan-Ngô program ([BK00] and [Ngo20]), this classical theory of Fourier analysis should be reformulated on the group . In such a reformulation, the classical (additive) Fourier transform is converted to a convolution integral with a kernel function and the Poisson summation formula is converted to a theta inversion formula, which is a generalization of the classical theta inversion formula. We refer to [JL23, Section 2] for details.
In [JL22, JL23], the Godement-Jacquet theory has been reformulated as harmonic analysis on for as a vast generalization of the 1950 thesis of J. Tate ([Tat67]). More precisely, for any irreducible cuspidal automorphic representation of , the space of -Schwartz functions on is defined, which is denoted by , and the -Foruier transform (or operator) is defined for any given non-trivial additive character of , which takes the -Schwartz space to the -Schwartz space , where is the contragredient of . We refer to Section 2 for details. The -Poisson summation formula as proved in [JL23, Theorem 4.7] (or recalled in Theorem 2.1) takes the following form.
Theorem 1.1 (-Poisson summation formula).
With the notation introduced above, the -theta function converges absolutely and locally uniformly for any and any . Moreover, the following identity
(1.1) |
holds as functions in .
This -reformulation in [JL22, JL23] of the Godement-Jacquet theory has found following nice applications, among others:
-
(1)
The local theory of such a -reformulation as developed in [JL22, JL23] proves that the (nonlinear) Fourier transform that is responsible for the local functional equation is given by a convolution operator with an explicitly defined kernel function on , which will be related to a Bessel function in Section 4 of this paper. As consequences, all the Langlands local gamma factors take the form of those in the theory of I. Gelfand, M. Graev, and I. Piatetski-Shapiro in [GGPS69] and of A. Weil in [Wei95], where the local gamma factors associated with quasi-characters of were considered.
-
(2)
The global theory of such a -reformation as developed in [JL23] gives the adelic formulation of A. Connes’ theorem ([Con99, Theorem III.1]) for , and the complete version of C. Soulé’s theorem ([Sou01, Theorem 2]) that provides a spectral interpretation of the zeros of . We refer to [JL23, Theorem 8.1] for details.
-
(3)
In Section 5 of this paper, the local and global theory of such a -reformulation in [JL22, JL23] provides a Poisson summation formula proof of the Voronoi formula for any irreducible cuspidal automorphic representation of , which was previously proved by S. Miller and W. Schmid in [MS11] and by A. Ichino and N. Templier in [IT13] by using the Rankin-Selberg convolution of H. Jacquet, I. Piatetski-Shapiro and J. Shalika in [JPSS83, CPS04].
-
(4)
In Section 6 of this paper, the local and global theory of such a -reformulation in [JL22, JL23] defines the Godement-Jacquet kernels for and proves Theorems 6.9 and 6.13, which are the -versions of Clozel’s theorem ([Clo22, Theorem 1.1]) for any irreducible cuspidal automorphic representation of . In [Clo22], Clozel formulates and proves such a theorem for the Tate kernel associated with the Dedekind zeta function for any number field .
1.2. Voronoi summation formula
The classical Voronoi summation formula and its recent extension to the -version have been one of the most powerful tools in Number Theory and relevant areas in Analysis. We refer to an enlightening survey paper by S. Miller and W. Schmid ([MS04b]) for a detailed account of the current state of the art of the Voronoi summation formula and its applications to important problems in Number Theory.
The Voronoi summation formula for was first studied by S. Miller and W. Schmid in [MS06] for and in [MS11] for general . They use two approaches. One is based on classical harmonic analysis that has been developed in their earlier paper ([MS04a]), and the other is based on the adelic version of the Rankin-Selberg convolutions for , which was developed by H. Jacquet, I. Piatetski-Shapiro and J. Shalika in [JPSS83] and by J. Cogdell and Piatetski-Shapiro in [CPS04] (and also by Jacquet in [Jac09] for the Archimedean local theory). The classical approach to the Voronoi formula for has also been discussed in [GL06] and [GL08]. A complete treatment of the adelic approach to the Voronoi formula for over a general number field was given by A. Ichino and N. Templier in [IT13]. We recall their general Voromoi formula for below.
For each irreducible cuspidal automorphic representation of , the Voronoi summation formula is an identity of two summations. One side of the identity is given by certain data associated with and the other side is given by certain corresponding data associated with , the contragredient of . Let be a non-trivial additive character on . At each local place of , to a smooth compactly supported function is associated a dual function such that the following functional equation
(1.2) |
holds for all and all unitary characters of . Since any irreducible cuspidal automorphic representation of is generic, i.e. it has a non-zero Whittaker-Fourier coefficient. If we write , where denotes the set of all local places of , then at any local place , the local component is an irreducible admissible and generic representation of . Let be the local Whittaker model of , and be any Whittaker function on that belongs to (see Section 3 for the details).
Let be a finite set of including all Archimedean places and the local places where or is ramified. As usual, we write , where , which is naturally embedded as a subring of , and the subring of adeles with trivial component above . At , we take the unramified Whittaker vector of , which is so normalized that . Denote by , which is the normalized unramified Whittaker function of . Similarly, we define to be the (normalized) unramified Whittaker function of . We recall that the functions and are related by the following
for all , where is the longest Weyl element of as defined in (3.3). The following is the Voronoi formula proved in [IT13, Theorem 1]. The unexplained notation will be defined in Sections 3 and 5.
Theorem 1.2 (Voronoi Summation Formula).
For , let be the set of places such that . At each let . Then:
where and the same for , and is a finite Euler product of the local Kloosterman integrals:
For the place , the local Kloosterman integral is defined by
where
as given in [IT13, Section 2.6].
The proof of Theorem 1.2 in [IT13] is based on the local and global theory of the Rankin-Selberg convolution for ([JPSS83, CPS04, Jac09]). It is important to mention that Theorem 1.2 and its proof has be extended by A. Corbett to cover an even more general situation with more applications in Number Theory ([Cor21, Theorem 3.4]).
From the historical development of the Voronoi summation formula, one expects that there should be a proof of the Voronoi formula via a certain kind of Poisson summation formula. In other words, the two sides of the Voronoi formula should be related by a certain kind of Fourier transform and the identity should be deduced from the corresponding Poisson summation formula. In the current proof of Theorem 1.2, such important ingredients from the harmonic analysis were missing, although there were discussions in [IT13] and [Cor21] on the local Bessel transform with the kernels deduced from the local functional equation in the local theory of the Rankin-Selberg convolution in [JPSS83] and [Jac09], and the identity was deduced from explicit computations from the global zeta integrals of the Rankin-Selberg convolution ([JPSS83] and [CPS04]). Over the Archimedean local fields, Z. Qi has developed in [Qi20] a theory of fundamental Bessel functions of high rank and formulated those Bessel transforms in the framework of general Hankel transforms that are integral transforms with Bessel functions as the kernel functions.
The first global result of this paper is to show that Theorem 1.2 is a special case of Theorem 1.1. From our proof, it will be clear that any variant of Theorem 1.2 (see [Cor21, Theorem 3.4] for instance) is also a special case of Theorem 1.1. Note that the -Poisson summation formula on in Theorem 1.1) relies heavily on the work of R. Godement and H. Jacquet ([GJ72]). Hence our proof of Theorem 1.2 is in principle based on the local and global theory of the Godement-Jacquet integrals for the standard -functions of .
In order to carry out such a proof, we have to understand the nature of the functions occurring on the both side of the Voromoi formula in Theorem 1.2, locally and globally, and show that they are -Schwartz function on and are related by the -Fourier transform in the sense of [JL22, JL23]. More precisely, for any irreducible smooth representation of , which is of Casselman-Wallach type if is an infinite local place of , we define the -Bessel function on (Definitions 4.2, 4.9 and 4.12) and obtain a series of results on the relations between the -Bessel functions , the -Fourier transforms and the -kernel functions as introduced and studied in [JL22, JL23] (see (2.18) and (2.20) for details), and on new formulas for the dual functions of . After all the local preparation, we deduce the Voromoi formula in Theorem 1.2 from the -Poisson summation formula in Theorem 1.1. We summarize those local results as the following theorem.
Theorem 1.3.
For any local place of the number field , let be an irreducible smooth representation of , which is of Casselman-Wallach type if is infinite. For any , is the dual function of as (1.2) or in Theorem 1.2. Then the following hold.
-
(1)
The -Fourier transform realizes the duality between and , up to normalization,
-
(2)
The dual function of enjoys the following formula:
where is the -kernel function of as in (2.20) and .
-
(3)
As distributions on , the -kernel function as in (2.20) and the -Bessel function are related by the following identity:
-
(4)
The dual function of enjoys the following formula:
The proof of Theorem 1.3 is given in Sections 3 and 4. More precisely, Part (1) of Theorem 1.3 is Proposition 3.5. Part (2) is Corollary 3.6. Part (3) is a combination of Propositions 4.3, 4.8, and 4.13. Part (4) is an easy consequence of Parts (2) and (3), which is Corollary 4.15.
It is important to point out that the -Bessel functions in the -adic case is defined by means of the Whittaker model of following the general framework of E. Baruch in [Bar05]. Hence we have to assume in the -adic case that is generic in the definition of the -Bessel functions . However, in the real or complex case, we follow the general theory of Z. Qi in [Qi20] on Bessel functions of high rank, which works for general irreducible smooth representations of of Casselman-Wallach type. Hence in the real or complex case, the definition of the -Bessel functions does not require that is generic. Since the -kernel function as in (2.20) is defined based on the Godement-Jacquet theory, which does not require that is generic, the uniform result in Part (3) of Theorem 1.3 suggests that one may define the -kernel function to be the -Bessel functions in the -adic case when is not generic. Finally, let us mention that the -Bessel functions in the real case as given in Definition 4.12 is more general than the one defined in [Qi20], and refer to Remark 4.14 for details.
1.3. Godement-Jacquet kernels and Fourier transform
Write , where denotes the subset of consisting of all Archimedean local places of , and denotes the subset of consisting of all finite local places of . Write . For , set . Let be the ring of algebraic integers of . L. Clozel defines in [Clo22] the Tate kernel:
(1.3) |
for with , where runs over nonzero integral ideals , and . Here is the Dedekind zeta function of with , the absolute norm of ; and the dual kernel
(1.4) |
where is the difference of and is the inverse difference; and is the absolute value of the discriminant. Theorem 1.1 of [Clo22] expresses the relation between those tempered distributions and on in terms of the condition: with , which is more precisely stated as follows.
Theorem 1.4 (Clozel).
Assume that . Then if and only if
where is the usual Fourier transform over with a suitable normalized measure.
The second global result of this paper is to define the kernel functions for any irreducible cuspidal automorphic representation of , which will be called the Godement-Jacquet kernels, and prove the analogy of Theorem 1.4 for the Godement-Jacquet kernels and the standard -functions (see Theorems 6.9 and 6.13 for the exact statements).
Let be an irreducible cuspidal automorphic representation of and write , where , and write the standard -function of as
(1.5) |
for sufficiently positive. As usual, is called the complete -function associated with , and is called the finite part of the -function associated with . The local and global theory of R. Godement and H. Jacquet in [GJ72] introduces the global zeta integrals for and proves that has analytic continuation to an entire function in and satisfies the functional equation
Following the reformulation as developed in [JL23], for an irreducible cuspidal automorphic representation of , there exists a -Schwartz space as defined in (2.23), which defines the -zeta integral
(1.6) |
for any . By [JL23, Theorem 4.6] the zeta integral converges absolutely for , admits analytic continuation to an entire function in , and satisfies the functional equation
(1.7) |
where is the -Fourier transform as defined in (2.25). From the global functional equation in (1.7), we introduce the notion of the Godement-Jacquet kernels for in Definition 6.5, which can be briefly explained as follows.
Write as with and . Set . For , we have that and . Write
For with and , we write
(1.8) |
for any , where the inner integral is taken over the domain . Proposition 6.1 shows that the integral converges absolutely for any and is holomorphic in . By the Fubini theorem and the support in of , which is a fractional ideal of (Proposition 6.6) , the inner integral converges absolutely for any and any . The Godement-Jacquet kernel for is defined by
(1.9) |
for and for all . Clozel defines in [Clo22] the dual kernel for the case of with the trivial character. We define here the dual kernel of the Godement-Jacquet kernel for to be
(1.10) |
for and for all . Proposition 6.6 shows that both kernel functions and on can be extended uniquely to tempered distributions on for any and for any , by using the work of S. Miller and W. Schmid in [MS04a]. We are able to match the kernels and with the Euler product expression or Dirichlet series expression of the finite part -function by specifically choosing the -Schwartz functions , and prove in Section 6 the -versions of the Clozel theorem (Theorem 1.4). Here is an overly simplified version of Theorem 6.9, to which we refer the details.
Theorem 1.5.
For any irreducible cuspidal automorphic representation of , with a choice of , the Godement-Jacquet kernel and its dual kernel enjoy the following identity:
(1.11) |
as distributions on if and only if is a zero of .
1.4. Organiztion of the paper
We recall the -Poisson summation formula on developed by Z. Luo and the first named author of this paper in [JL23] in Section 2. Sections 2.1 and 2.2 are to review briefly the local -Schwartz spaces and local -Fourier operators developed in [JL22] and [JL23]. Based on their work as well as the work of [GJ72], we recall the formulation of the -Poission summation formula on in [JL23] in Section 2.3.
Sections 3 and 4 are devoted to understand the duality between the function and the function by means of the harmonic analysis on as developed in [JL22] and [JL23], and to prove our main local results (Theorem 1.3). By comparing the Godement-Jacquet theory with the Rankin-Selberg convolution, we are able to express the dual function of in terms of the -Fourier transform up to certain normalization (Proposition 3.5), based on Proposition 3.1 that identifies the -Schwartz space , as introduced in [JL23] and recalled in (2.5), with the -Whittaker-Schwartz space as defined in (3.1). As a consequence, we obtain a formula that express the dual function of as a convolution of the -kernel function with , up to certain normalization (Corollary 3.6). In Section 4, we introduce the notion of -Bessel functions (Definitions 4.2, 4.9 and 4.12) and prove the precise relation between the -Bessel functions and -kernel functions as defined in (2.20) (Propositions 4.3, 4.8, and 4.13). In the -adic case, the -Bessel function on is introduced following the work of E. Baruch in [Bar05], which is recalled in Section 4.1. In the real or complex case, we introduce the -Bessel function on by following the general theory of Bessel functions of high rank by Z. Qi in [Qi20]. It should be mentioned that the -Bessel function on in the real case is more general than the one considered in [Qi20] (Remark 4.14). As expected, when our results recover the previous known results as discussed by J. Cogdell in [Cog14] and by D. Soudry in [Sou84].
With all these ingredients, we are able to give a new proof of the Voronoi formula for (Theorem 1.2) as proved in [IT13] in Section 5. In fact, the proof of the Voronoi formula in [IT13] is based on the Rankin-Selberg convolution for in [JPSS83], [CPS04] and [Jac09]. And our proof is based on the Godement-Jacquet theory in [GJ72] and its reformulation in [JL22, JL23]. The main idea is that the Voronoi summation formula for as in Theorem 1.2 is a special case of the -Poission summation formula on as in Theorem 2.1 ([JL23, Theorem 4.7]), after the long computations carried out in Section 3 of this paper and in [JL22, JL23]. Those computations enable us to express the summands on the dual side (the right-hand side) of the Voronoi formula in Theorem 1.2 as the global -Fourier transform of the summands on the given side (the left-hand side), which is Proposition 5.4.
In Section 6.1, in order to define the Godement-Jacquet kernels and their dual kernels (Definition 6.5) for any irreducible cuspidal automorphic representation of , we develop further properties (Propositions 6.1 and 6.3, and Corollary 6.4) of the global zeta integrals , as defined in (1.6), by using the -Fourier transform and the associated -Poisson summation formula as developed in [JL23]. In Proposition 6.6, we show that both kernel functions and on can be extended uniquely to tempered distributions on for any and for any . In Section 6.2, guided by Theorem 1.4, we prove in Proposition 6.7 that if is a zero of , then the kernel is equal to the negative of -Fourier transform of the dual kernel . For any , take , where with as given in Proposition 6.8. Theorem 6.9 proves the -version of Theorem 1.4 for the Euler product expression of . With the help of Lemma 6.10, we obtain the Dirichlet series expression of the kernels in Propositions 6.11 and 6.12. Finally, Theorem 6.13 establishes the -version of Theorem 1.4 for the Dirichlet series expression of .
2. -Reformulation of the Godement-Jacquet Theory
We recall from [JL22, JL23] the -reformulation of the Godement-Jacquet theory for the standard -functions associated with any irreducible cuspidal automorphic representation of (for ), where is the ring of adeles of a number field . More precisely, we recall the -Schwartz spaces on and the -Fourier operators over both for the local and global cases, and the -Poisson summation formula on .
2.1. -Schwartz functions
Let be the set of all local places of . For any local place , we denote by , the local field of at . If is non-Archimedean, we denote by the ring of integers and by the maximal ideal of . Let be the general linear group defined over . Fix the maximal compact subgroups of , where if is non-Archimedean, if , and if .
Let be the space of all matrices over and be the space of Schwartz functions on . When is Archimedean, it is the space of usual Schwartz functions on the affine space , and when is -adic, it consists of all locally constant, compactly supported functions on . Let be the normalized absolute value on the local field , which is the modular function of the multiplication of on with respect to the self-dual additive Haar measure on . As a reformulation of the local Godement-Jacquet theory in [JL23, Section 2.2], the (standard) Schwartz space on is defined to be
(2.1) |
where denotes the space of all smooth functions on . By [JL23, Prposition 2.5], the Schwartz space is a subspace of , which is the space of square-integrable functions on .
Consider the determinant map When restricted to , we obtain that
(2.2) |
and the fibers of the determinant map are of the form:
(2.3) |
When , the fiber is the kernel of the map, i.e. . In general, each fiber is an -torsor. Let be the self-dual Haar measure on with respect to the standard Fourier transform defined by (2.7) below. On , we fix the Haar measure . Let be the induced Haar measure from to . It follows that the Haar measure induces an -invariant measure on each fiber .
Let be the set of equivalence classes of irreducible smooth representations of when is non-Archimedean; and of irreducible Casselman-Wallach representations of when is Archimedean. For , we denote by the space of all matrix coefficients of . Write with some as in (2.1). For , as in [JL23, Section 3.1], we define
(2.4) |
By [JL23, Proposition 3.2], the function is absolutely convergent for all and is smooth over . As in [JL23, Definition 3.3], for any , the space of -Schwartz functions is defined as
(2.5) |
By [JL23, Corollary 3.8], we have
(2.6) |
2.2. -Fourier transform
Let be a fixed non-trivial additive character of . The (standard) Fourier transform on is defined as follows,
(2.7) |
It is well-known that the Fourier transform extends to a unitary operator on the space and satisfies the following identity:
(2.8) |
Following the reformulation of the local Godement-Jacquet theory in [JL23, Section 2.3], the Fourier transform on yields a (nonlinear) Fourier transform on , which is a convolution operator with the distribution kernel:
(2.9) |
More precisely, the Fourier transform is defined to be
(2.10) |
for any , where . From [JL23, Proposition 2.6], a relation between the (nonlinear) Fourier operator and the (classical or linear) Fourier transform is given by
(2.11) |
From the proof of [JL23, Proposition 2.6], it is easy to obtain that
(2.12) |
for any .
As in [JL23, Section 3.2], the -Fourier transform is defined through the following diagram:
(2.17) |
More precisely, for with a and a , the -Fourier transform is defined by
(2.18) |
where . It was verified in [JL23, Proposition 3.9] that the descending -Fourier transform is well defined. From [JL22, Theorem 5.1], the -Fourier transform can also be represented as a convolution operator by some kernel function , which is explicitly given as follows.
We fix a with . We also choose a sequence of test functions , such that for any ,
(2.19) |
In other words, the sequence tends to the delta mass supported at the identity as . The -kernel function is defined as
(2.20) |
where is the kernel function as defined in (2.9) and the integral is regularized as follows:
(2.21) |
It is shown in [JL22, Proposition 3.5, Corollary 3.7, Corollary 4.5 and Theorem 4.6] that is a smooth function on and is independent of the choice of the matrix coefficient and the chosen sequence that tends to the delta mass supported at . By [JL22, Theorem 5.1], we have that for any
(2.22) |
Following [Ngo20], one may call the -Fourier transform a generalized Hankel transform or the -Hankel transform.
2.3. -Poisson summation formula on
Recall that is the set of all local places of the number field . Let be the subset of consisting of all Archimedean local places of . We may write , where is the set of non-Archimedean local places of . Let be the set of equivalence classes of irreducible admissible representations of . We write and assume that and at almost all finite local places the local representations are unramified. This means that when , is an irreducible admissible representation of , and when , is of Casselman-Wallach type as a representation of . Let be the subset consisting of equivalence classes of irreducible admissible automorphic representations of , and be the subset of cuspidal members in . We refer to [Art13, Chepter 1] or [JL23] for the notation and definition of automorphic representations.
Take any . For each local place , the -Schwartz space is defined as in (2.5). Recall from [JL23, Theorem 3.4] that the basic function is defined when the local component of is unramified. Then the -Schwartz space on is defined to be
(2.23) |
which is the restricted tensor product of the local -Schwartz space with respect to the family of the basic functions for all the local places at which are unramified. The factorizable vectors in can be written as
(2.24) |
Here at almost all finite local places , . According to the normalization ([JL23, Theorem 3.4]), we have that when , the unit group of the ring of integers at . Hence for any given , the product in (2.24) is a finite product.
For any factorizable vectors in , we define the -Fourier transform (or operator):
(2.25) |
Here at each , is the local -Fourier transform as defined in (2.17) and (2.18), which takes the -Schwartz space to the -Schwartz space , and has the property that , when the data are unramified at (see [JL23, Theorem 3.10]). Hence the Fourier transform as defined in (2.25) maps the -Schwartz space to the -Schwartz space , where is the contragredient of . The -Poisson summation formula ([JL23, Theorem 4.7]) can be stated as below.
Theorem 2.1 (-Poisson summation formula).
For any , the -theta function converges absolutely and locally uniformly for any and any . Let be the contragredient of . Then the following identity
holds as functions in , where is the -Fourier transform as defined in (2.25).
3. Local Harmonic Analysis
In this section, we take to be a local field of characteristic zero and fix a non-trivial additive character of . Since the representations considered in this section are the local components of irreducible cuspidal automorphic representations of , we may only consider generic without loss of generality.
Let be the Borel subgroup of , which consisting of all upper-triangular matrices of , where is the maximal torus consisting of all diagonal matrices of , and is the unipotent radical of , which consists of matrices with if , and for . Without loss of generality, we may take a generic character as
Let be a non-zero member in , which is one-dimensional if is generic. For any , define the Whittaker function by . Let be the Whittaker model of , which consisting of Whittaker functions with runs through the space of . Let be the subspace of consisting of all smooth vectors of . We define the -Whittaker-Schwartz space on to be
(3.1) |
where is the normalized absolute value on .
Proposition 3.1.
For any , which is generic, the -Schwartz space and the -Whittaker-Schwartz space coincide with each other: .
Proof.
We first show that . For any unitary character of and , the local Rankin-Selberg integral for
where as in Proposition 3.2, is absolutely convergent when is sufficiently positive and the fractional ideal generated by all such integrals is by [JPSS83, Theorem 2.7] for the non-Archimedean case and a holomorphic multiple of , bounded at infinity in vertical strips due to [Jac09, Theorem 2.1] for Archimedean case.
According to [JL23, Theorem 3.4], there is some such that
when is sufficiently positive. In particular, fix a sufficiently positive such that both functions and belong to , the space of -functions on . It follows that
for all unitary character of . From the general theory about absolutely continuous measures on local compact abelian groups (See [HR79, Theorem 23.11] for instance), we must have that for a.e. , which implies for all since both functions are smooth. Hence we obtain that .
From Proposition 3.1, the following assertion is clear, since the -Schwartz space is independent of the choice of the character .
Corollary 3.2.
The space of -Whittaker-Schwartz functions defined in (3.1) is independent of the choice of the character .
By Corollary 3.2, we may denote by the -Whittaker-Schwartz space on as defined in (3.1). After identifying the -Schwartz space with the -Whittaker-Schwartz space , we are going to understand the -Fourier transform in terms of the structure of Whittaker models.
Proposition 3.3.
For , we may write as in (3.1) that
for some . Then the -Fourier transform can be expressed by the following formula:
where for any is a Whiitaker function in and
(3.2) |
Here we denote by the longest Weyl element of , which is defined inductively by
(3.3) |
Proof.
From the functional equation for the local zeta integrals as proved in [JL23, Theorem 3.10], we have that
On the other hand, from the functional equation for the local zeta integrals as proved in [JPSS83, Theorem 2.7] for the non-Archimedean case and in [Jac09, Theorem 2.1] for the Archimedean case, we have that
From the absolute convergence of the local zeta integrals and , we may choose and fix a with sufficiently negative, such that both functions
belong to . It follows that
for any unitary character . Now we use the same argument as in the proof of Proposition 3.1 to deduce that
for any , as they are smooth in . ∎
In particular, in the case , we have a much simpler formula.
Corollary 3.4.
When , the action of the longest Weyl group element of on the Kirillov model of is given by the (non-linear) Fourier transform :
where the -Schwartz space and the -Whittaker-Schwartz space can be identified with the Kirillov model of by Proposition 3.1 and is the central character of .
Proof.
According to [IT13, Lemma 5.2], for any , there is a unique smooth function on of rapid decay at infinity and with at most polynomial growth at zero such that the local functional equation (1.2) holds as meromorphic functions in . The map is called the Bessel transform in [IT13]. Some more discussions and explicit formulas related to this map were given in [Cor21, Section 4] based on the local functional equation of the Rankin-Selberg convolution for from [JPSS83] and [Jac09]. Over the Archimedean local fields, the map has been studied in [Qi20] in the framework of Hankel transforms with the Bessel functions of high rank as the kernel functions. The following result says that the map is given by the -Fourier transform up to certain normalization.
Proposition 3.5.
The dual function of as defined by (1.2) can be expressed in terms of -Fourier transforms:
Proof.
Since , we have as well. By [JL23, Theorem 3.4] and as in the proof of Proposition 3.1, the right-hand side of (1.2) can be written as
By the local functional equation in [JL23, Theorem 3.10], we have
It follows that the left-hand side of (1.2) can be written as
as meromorphic functions in . Since the integrals on both sides of the above equation converge absolutely for sufficiently negative, we choose one of such and fix it such that the two smooth functions and belong to . Again, by the general theory as in [HR79, Theorem 23.11], we obtain that
which implies that , as functions on , is smooth, of rapid decay at infinity, and with at most polynomial growth at zero. ∎
Corollary 3.6.
For any , the dual function associated with any is given by the following formula:
where is the -kernel function associated with as in (2.20) and .
4. -Bessel functions
The -Fourier transform can be expressed as a convolution operator with the -kernel function as in (2.20) using the structures of the -Schwartz space and the -Schwartz space . When consider the -Fourier transform as a transformation from the -Whittaker-Schwartz space to -Whittaker-Schwartz space , we are able to show that the -Fourier transform can be expressed as a convolution operator with certain Bessel functions as the kernel functions. We do this for the Archimedean case and non-Archimedean case, separately.
4.1. -Bessel functions: -adic case
Assume that is non-Archimedean. In this case, a basic theory of Bessel functions was developed by E. Baruch in [Bar05], from which we recall some relevant definitions and results on Bessel functions in order to understand the -Fourier transform.
Let be the roots of with respect to the -split maximal torus , be the set of positive roots with respect to and be the corresponding set of negative roots. Let be the set of simple roots. Let be the Weyl group of . For every , denote
where is the longest Weyl element of as in Proposition 3.3. We also write
Let ( resp.) be the unipotent subgroup associated to ( resp.). Let
For every , where is the character group of , define
Recall from Section 2.1 that is the maximal open compact subgroup of . With the Iwasawa decomposition , for any , we set . It is easy to check that this is well defined.
Let be generic and be the space of Whittaker functions. Following [Bar05, Definition 5.1], we denote by the set of functions such that for every and every , there exist positive constants such that if then implies that . For a positive integer , we denote the congruence subgroup given by . Write , where is a fixed uniformizer of . Let . For any , denote
According to [Bar05, Theorem 7.3], for all sufficiently large . Due to [Bar05, Proposition 8.1], for large enough, the integral converges and is independent of for . Moreover, by the uniqueness of Whittaker functionals, it follows that there exists a function, which we denote by such that
for . This function was called the Bessel function of attached to the Weyl group element in [Bar05, Section 8]. Moreover, if , then the integral converges absolutely for and the Bessel function has the following integral representation:
(4.1) |
according to [Bar05, Theorem 5.7 and Theorem 8.1].
Lemma 4.1.
Let be a non-Archimedean local field. Define
Then this space can be identified with the space : . In particular, the space is independent of the choice of the character .
Proof.
We first prove that
Take . Then we have that . It follows that there are positive constants and such that implies that
which implies that
On the other hand, take , for any positive integer , we have
and since for sufficiently large, we get
According to [Bar05, Corollary 5.5], is invariant under right translations by , in particular, for , and , where , we have
According to [JL70, Lemma 2.9.1], is an irreducible representation under the above action. Hence we obtain that
Note that is a bijection from to itself. Therefore we obtain that . ∎
In order to understand the -Fourier transform and the associated -kernel function as in (2.20) in terms of the -Whittaker-Schwartz space to -Whittaker-Schwartz space , we define the -Bessel function of on , which is related to the one attached to the particular Weyl element , up to normalization.
Definition 4.2.
Let be a non-Archimedean local field of characteristic zero. For any , which is generic, the associated -Bessel function on is defined by
(4.2) |
where is a Weyl group element of
Proposition 4.3.
Proof.
For any , we know from Proposition 4.1 that there is some such that
for any . According to Proposition 3.3, we have
as . According to [Bar05, Theorem 5.7], the function
is compactly supported once we fix . Hence the function
belongs to the space , as a function in with fixed, and its Fourier transform along at
exists. For the Weyl group element , it is easy to check that
from which we deduce the following formula for :
where . By (4.1), we obtain that
From the definition of the Bessel function in (4.2), we obtain that
as functions in . Now we calculate for a fixed , the Fourier transform with ,
According to [Bar05, Theorem 5.7], we can apply the Fourier inversion formula to obtain
Hence we obtain from the above calculation that
On the other hand, we know from [JL22, Theorem 5.2] that
for any . Therefore, as distributions on , we obtain that for any . Since both functions are smooth, the identity holds as functions in . ∎
Recall that in the case, D. Soudry defined in [Sou84] the Bessel function on by the following equation
(4.3) |
for all , where the integral converges in the sense that it stabilizes for large compacts as in [Sou84, Lemma 4.1]. By an elementary computation, we see the relation between these two Bessel functions is
In [Sou84], Soudry computes the Mellin transform of the product of two Bessel functions instead of showing the gamma factor is the Mellin transform of . In fact, we have
Corollary 4.4.
Proof.
We refer to [Cog14] for further discussion of the -Bessel functions and related topics.
4.2. -Bessel functions: complex case
If , let us first recall from [Kna94] the classification of irreducible admissible representations of . For , let and , where is the complex conjugate of . For any and , let be the representation of given by , which we write . For each with , let be the representation of . Then defines a one-dimensional representation of the diagonal maximal torus of , which can be extended trivially to a one-dimensional representation of the upper triangular Borel subgroup . We set
which is the unitary induction as in [Kna01, Chapter VII]. According to [Zel75, ZN66], we have
Theorem 4.5 (Classification).
The irreducible admissible representations of can be classified as follows.
-
(1)
If the parameters of satisfies , then has a unique irreducible quotient .
-
(2)
the representations exhaust the irreducible admissible representations of , up to infinitesimal equivalence.
-
(3)
Two such representations and are infinitesimally equivalent if and only if there exists a permutation of such that for .
According to [Jac79], the associated local factors can be expressed as follows.
Theorem 4.6 (Local Factors).
Let be an irreducible admissible representation of with , where and for every . The local -factor and local -factor associated with are given by
For any , the local -factor associated with is given by
(4.4) |
Remark 4.7.
In [Qi20], Z. Qi defines a Bessel kernel function for any by the following Mellin-Barnes type integral,
where
and is any contour such that
-
•
is upward directed from to , where ,
-
•
all the set lie on the left side of , and
-
•
if and large enough, then .
For more details, we refer to [Qi20, Definition 3.2]. Then [Qi20] defines
(4.5) |
and [Qi20, Lemma 3.10] secures the absolute convergence of this series. The following is the analogy in the complex case of Proposition 4.3.
Proposition 4.8.
For any , which is parameterized by as in Theorem 4.5, as distributions on , the identity: holds for any .
Proof.
According to [Qi20, Theorem 3.15], for any , there is a unique function , which is contained in the space as defined in [JL23, Definition 2.1], such that
It follows that according to [JL23, Theorem 2.3, Proposition 3.7, and Corollary 3.8]. From [Qi20, Proposition 3.17], we have that
On the other hand, we have that
due to [JL22, Theorem 5.1]. The -kernel function is a smooth function on according to [JL22, Corollary 4.5], while the function is real analytic on due to [Qi20, Proposition 3.17]. Since is arbitrary, we thus deduce that for any , as functions on . ∎
As in Definition 4.2, we introduce the -Beesel function on .
4.3. -Bessel functions: real case
If , we recall from [Kna94] the classification of irreducible admissible representations of . For any , let be the discrete series of , that is, the representation space consists of analytic functions in the upper half-plane with
finite, and the action of is given by
Let be the subgroup of elements in with and
be the induced representation of , where we still use the unitary induction as in [Kna01, Chapter VII]. For each pair , let be the representation of obtained by tensoring the above representation on with the quasi-character , that is, , where . For a pair , let be the representation of : .
For any partition of : with each equal to or and with , we associate the block diagonal subgroup . For each , let be the representation of of the form or as defined above. We extend the tensor product of these representations to the corresponding block upper triangular subgroup by making it the identity on the block strictly upper triangular subgroup. We set
Theorem 4.10 (Classification).
The irreducible admissible representations of can be classified as follows.
-
(1)
If the parameters of satisfy
then has a unique irreducible quotient .
-
(2)
The representations exhaust the irreducible admissible representations of , up to infinitesimal equivalence.
-
(3)
Two such representations and are infinitesimally equivalent if and only if and there exists a permutation such that for each .
According to [Jac79] again, the local factors can be expressed as follows:
Theorem 4.11 (Local Factors).
For a representation of or as defined above, denote
then for , we have . Similarly denote
then the -factor of is given by . Finally, the local -factor associated with is given by
where is the contragredient of .
For any , according to [JL23, Theorem 3.10], there is some function such that such that
Due to [Igu78, Theorem 4.2], for large enough, we have
We choose a contour with the following three properties:
-
(1)
is upward directed from to , where is small enough, say
-
(2)
The sets for and for , all lie on the left side of , and
-
(3)
If , then for large enough, .
Then for large enough, we have that for fixed and ,
for some constant for all s with , and the constant only depends on , , , and , and is independent of . It follows that
for some other constant according to [Qi20, Lemma 1.3] and Property (1) of the contour . Hence, as , the above integral goes zero, and we are able to change the integral from to according to the Cauchy residue theorem and Property (2) of the contour , that is
According to Property (1) of the contour and [Qi20, Lemma 1.3] again, we have that
Hence we can change the order of integration using Fubini’s theorem to obtain that
As in Definitions 4.2 and 4.9, we define the -Bessel function on as follows
Definition 4.12.
For any , which is generic, the -Beesel function on is given as
The integral in Definition 4.12 is absolutely convergent to a smooth function in because of Property (1) of the contour and [Qi20, Lemma 1.3]. Moreover we prove the following proposition, which is the analogy in the real case of Propositions 4.3 and 4.8.
Proposition 4.13.
For any , which is generic, the -kernel function and the -Bessel function are related by the following identity as functions on , i.e.
Proof.
4.4. -Bessel functions and dual functions
From Definitions 4.2, 4.9 and 4.12, for a given , we define the (normalized) -Bessel function on for every local field of characteristic zero. In Propositions 4.3, 4.8 and 4.13, we obtain the relation between the -kernel function and the -Bessel function . As a record, we state the corresponding formula for the dual function of following Corollary 3.6
Corollary 4.15.
For any , the dual function associated with any is given by the following formula:
for all , where .
5. A New Proof of the Voronoi Summation Formula
In this section, we give a new proof of the Voronoi summation formula based on the -Poisson summation formula ([JL23, Theorem 4.7]), which was recalled in Theorem 2.1. Let be a number field, the notations are all as in Section 2.
Lemma 5.1.
At any local place of , for any and , the function
belongs to the space . If and is unramified, let be the normalized unramifield Whittaker function associated with , then the function
belongs to the space .
Proof.
The first claim is trivial because for any , we have that
As for the second claim, since , we observe that for any given , if is small enough. Hence we have that
shares the same asymptotic behavior as with the function , which belongs to the space by Proposition 3.1. Hence we must have the function
belonging to the space . ∎
Lemma 5.2.
Let be a finite place such that both and are unramified, the -basic function as defined in [JL23, Theorem 3.4] enjoys the following formula:
Proof.
Lemma 5.3.
For the finite places where and are unramified, .
Proof.
According to [JL23, Lemma 5.3], the -basic function is supported in . The assertion follows clearly. ∎
Now we are ready to prove Theorem 1.2 by using Theorem 2.1. Recall that is the finite set of local places of that contains all the Archimedean places and those local places where either or is ramified. For any , we take
and
(5.1) |
Then the function belongs to the space according to Lemmas 5.1, 5.2 and 5.3. It is clear that the function is factorizable: . In order to use Theorem 2.1 in the proof, we calculate its local -Fourier transform of at each place . Let be as in Theorem 1.2.
For the unramified paces , by Lemma 5.2, we obtain that
By [JL23, Lemma 5.3] (or Lemma 5.3), if , then . Since when , we obtain that if . It follows that . Applying the -Fourier transform to the both sides, we obtain that
(5.2) |
according to [JL23, Theorem 3.10]. Note that the basic function in the -Schwartz space and , the Whittaker model of .
At , the function takes the following form . By Proposition 3.5, we obtain that
(5.3) |
Finally, at the local places , since is disjoint to , the function takes the following form
with . Recall from Section 4.1 that be the simple root for the root system with respect to . The one-parameter subgroups associated with and are given by
(5.4) |
Then the function can be written as
where . It is clear that . By Proposition 3.3, the -Fourier transform of is give by
Since
where with , we obtain that
Hence the -Fourier transform of can be written as
By the explicit computation of the last integral in [IT13, Section 2.6], we obtain that
(5.5) |
Thus, by (5.2), (5.3), and (5.5), we obtain a formula for the -Fourier transform of , which is the product of the local -Fourier transform of at all local places .
Proposition 5.4.
Let be the function as defined in (5.1). The -Fourier transform of can be explicitly written as
where the Kloosterman integral is given by .
Finally we write the summation on the one side as
and that on the other side as
because for every . By the -Poisson summation formula in Theorem 2.1, which is
we deduce the Voronoi formula in Theorem 1.2:
We deduce the Voronoi summation formula for from the -Poisson summation formula in Theorem 1.2.
Remark 5.5.
In [Cor21, Theorem 3.4], Corbett extends the Voronoi formula in Theorem 1.2 to a more general situation by allowing the local component at to be more general functions in . More precisely, if one take for and for , where , and , , are as in [Cor21, Theorem 3.4], then according to Lemmas 5.1, 5.2 and 5.3, the function . It is clear that the proof of Proposition 5.4 works for such special choices of functions as well. In particular, we obtain from Proposition 3.3 that at each local place , the Fourier transform is equal to the function in [Cor21, proof of Theorem 3.4]. The extended Voronoi formula for in [Cor21, Theorem 3.4] by using the Rankin-Selberg convolution for , can be deduced by the same argument as in our proof of Theorem 1.2 from the -Poisson summation formula in [JL23, Theorem 4.7]. We omit further details.
6. On the Godement-Jacquet Kernels
For any , the goal of this section is to define the Godement-Jacquet kernels for and their dual kernels, and to prove the -versions of [Clo22, Theorem 1.1], which can be viewed as the case of and is recalled in Theorem 1.4.
6.1. Godement-Jacquet kernel and its dual
We recall from [JL23, Section 4.2] the global zeta integral for the standard -function as stated in (1.6) is
(6.1) |
for any . By [JL23, Theorem 4.6] the zeta integral converges absolutely for , admits analytic continuation to an entire function in , and satisfies the functional equation
(6.2) |
where is the -Fourier transform as defined in (2.18). As explained in [JL23], this is a reformulation of the Godement-Jacquet theory for the standard -functions .
Consider the fibration through the idele norm map :
where and . One can have a suitable Haar measure on that is compatible with the Haar measures on and the Haar measure on . Write , where , and is the subset of consisting of elements with for all .
When , the absolutely convergent zeta integral as in (6.1) can be written as
(6.3) |
Proposition 6.1.
The first integral on the right-hand side of (6.3):
(6.4) |
converges absolutely at any and is holomorphic as a function in , for any .
Proof.
Let be a factorizable -Schwartz function. Let be a finite subset of (the set of all local places of ) that contains and such that for any both and are unramified and , the basic function in as in [JL23, Theorem 3.4]. Write . According to [JL23, Proposition 5.5 and Lemma 5.2], there is a positive real number , which depends only on the given , such that for any real number , the limit holds for every and for every . From the definition of the -Schwartz space in (2.5) and [JL23, Proposition 3.7], we know that when is large enough for all . Hence for every , there is a constant such that . By [JL23, Lemma 5.3], there is a positive real number , which also depends only on the given , such that for any , we have that holds for every . It is clear that for any constant and constant with , we must have that the inequality:
(6.5) |
holds for every .
We first estimate the inner integral, which can be written as
(6.6) |
Fix a and a section of the norm map and view as the -component of . Define
where , and with being the number of real places and the number of complex ones. Let be a basis for the group of units in the ring of integers in modulo the group of roots of unity in , and set
where is the class number of . We choose representatives of idele classes, and define . Then is a fundamental domain of according to [Tat67, Theorem 4.3.2], which is compact. Hence we can write (6.6) as
(6.7) |
Without loss of generality, we may take . Write . By (6.5), we have that
for any constant . Since , we must have that
Hence we obtain
(6.8) |
Since belongs to a compact set , the Archimedean part of belongs to a compact subset of . Hence there is a constant such that
For , we know from [JL23, Proposition 3.7] that for any constant is of rapid decay as . From the choice of the fundamental domain , we must have that . Due to [JL23, Lemma 5.3], there are integers such that for , if , then . According to [Neu99, Proposition 5.2], the image of in is a lattice, and there is a constant such that the (partial) theta series
Thus we obtain that and there is a constant such that . It follows that the integral converges absolutely as long as for any . Since is arbitrarily large with , we obtain that the integral
converges absolutely for any and hence is holomorphic as a function in .
Since a general element in is a finite linear combination of the factorizable functions, it is clear that the above statement for the integrals hold for general . ∎
From the above proof, we also obtain
Corollary 6.2.
For any , the inner integral
(6.9) |
always converges absolutely for any .
By using the -Poisson summation formula (Theorem 2.1), we obtain
Proposition 6.3.
For any , the following identity
holds for any .
Proof.
Applying Proposition 6.3 to the second integral on the right-hand side of (6.3), we obtain that for ,
(6.10) |
Corollary 6.4.
The second integral in (6.3): converges absolutely for and has analytic continuation to an entire function in . Moreover, the following identity
holds by analytic continuation for , where the integral on the right-hand side converges absolutely for all .
Set . By combining (6.3) with (6.10), we obtain that when
(6.11) |
which holds for all by analytic continuation. From the proof of Proposition 6.1, both integrals on the right-hand side converge absolutely when belongs to the vertical strip for any constant with . Hence they converge absolutely at any .
We are going to calculate the integral in another way. For , we write with and . For , we have that and . For , we write
(6.12) |
for any , where the inner integral is taken over . By the Fubini theorem, we know (from the proof of Proposition 6.6) that the inner integral
converges absolutely for any and any .
Definition 6.5 (Godement-Jacquet Kernels).
For any , take any , the Godement-Jacquet kernels associated with are defined to be
for and for all .
From (6.12), we obtain that
(6.13) |
In the spirit of [Clo22], to each , we define the dual kernel of the Godement-Jacquet kernel associated with to be
(6.14) |
for and for all .
With a suitable choice of the functions , the kernel functions and may have simple expressions. We refer to Proposition 6.11 for details. We establish the distribution property for and .
Proposition 6.6.
Set and write . For any and for any , the Godement-Jacquet kernel function and its dual kernel function on enjoy the following properties.
-
(1)
Both and vanish to infinity order along .
-
(2)
Both and have unique canonical extension across to the whole space .
-
(3)
Both and are tempered distributions on .
Proof.
By definition, we have that . It is enough to show that Properties (1), (2), and (3) hold for the kernel function . We prove (1) and (2) by using the work of S. Miller and W. Schmid in [MS04a] (in particular [MS04a, Definition 2.4, Lemma 2.8, Definition 2.6]). Then we prove (3) by showing that is of polynomial growth as the Eucilidean norm of tends to ([Tre67, Theorem 25.4]).
Without loss of generality, we may assume that is factorizable. Let be a finite set such that for , both and are unramified and , the basic function in . According to [JL23, Lemma 5.3], there are integers such that the support of is contained in . According to (6.5), for any , there is a constant such that for any . Write
where is the local uniformizer in and runs over the algebraic direct sum . Then for in the component, we have and the inequality: is equivalent to the inequality: . We may write a fractional ideal in as . We set , which is the fractional ideal depending on the support of .
According to the normalization of our Haar measure on , we have that
where the last summation runs over all fractional ideals of that are contained in with absolute norm less than or equal to . Write and obtain that
Let be the number of ideals with . According to the Wiener-Ikehara theorem ([MV07, Corollary 8.8]), there is a constant such that for all , and in particular . We obtain that
For a fixed , and any fixed , we have that
When , we deduce that
Hence we obtain that
(6.15) |
for any and , where is a constant depending on , , , , and is independent of . Moreover, from the above calculation, we obtain that
(6.16) |
if . From (6.16), it is clear that the kernel function vanishes at some neighborhood for any point . By [MS04a, Definition 2.4, Lemma 2.8, Definition 2.6], vanishes of infinity order at and has a unique canonical extension across to the whole , which we still denote by . We establish Properties (1) and (2).
6.2. -Fourier transform
From (6.14), we obtain for that
(6.17) |
which converges absolutely for all . The following is the duality relation of the Godement-Jacquet kernels and via the -Fourier transform when is such that , which is part of [Clo22, Theorem 1.1] for .
Proposition 6.7.
Proof.
For , we have
By the reformulation of the Godement-Jacquet local theory in [JL23, Theorem 3.4], we obtain that
where at almost all finite local places with equal to the basic function , we have that , for the remaining finite local places , where is holomorphic in , and is holomorphic in if . Hence we obtain that is a finite product of holomorphic functions. From (6.11), we have
for all by analytic continuation. Hence we obtain that if is such that , then we must have that
(6.18) |
Note that by Proposition 6.1 both integrals converges absolutely for any . From (6.17), we have that
is absolutely convergent according to (6.15). By [JL22, Theorem 5.1], which is recalled in (2.22), there is a -kernel function , such that for any
Since , By using the Fubini’s theorem and Proposition 6.1 again, we obtain that
By definition as in (2.22), we write the -Fourier transform of the dual kernel , viewed as a distribution on , to be
Hence we obtain that
By combining (6.18) with (6.13), we obtain the following identity as distributions on
for all . Therefore, as distributions on , we have that
∎
For any , we write that when is sufficiently positive. By [JL23, Corollary 3.8] and the theory of Mellin transforms, we obtain
Proposition 6.8.
For any , there is a function such that
holds as functions in by meromorphic continuation.
For any , take , where with as given in Proposition 6.8 and , the basic function, for almost all . It is clear that such a function belongs to the -Schwartz space . As in (6.1), the zeta integral
converges absolutely when and can be written as
(6.19) |
where when . We set
(6.20) |
and call the Godement-Jacquet kernel associated with the Euler product , and its dual kernel.
Theorem 6.9.
For any , take with where is as given in Proposition 6.8. Then the Godement-Jacquet kernel associated with the Euler product and its dual kernel enjoy the following identity:
(6.21) |
as distributions on if and only if is a zero of .
Proof.
6.3. Clozel’s theorem for
In [Clo22, Section 1], Clozel defines the Tate kernel and its dual kernel associated with the Dirichlet series expression of the Dedekind zeta function of the ground number field and prove Theorem 1.1 of [Clo22] by two methods, one is an approach from the Tate functional equation and the other is a more classical approach from analytic number theory. For a general , we define in Definition 6.5 and (6.14) the Godement-Jacquet kernels and their dual kernels via the -Fourier transform associated with the global functional equation in the reformulation of the Godement-Jacquet theory. By using the testing functions for the local zeta integrals and the local -factors at all finite local places (Proposition 6.8), we obtain the -version of [Clo22, Theorem 1.1] when the kernel functions are related to the -function with the Euler product expression (Theorem 6.9). In order to obtain the -version of [Clo22, Theorem 1.1] when the kernel functions are related to the -function with its Dirichlet series expression, we are going to refine the structure of the testing functions in Proposition 6.8 by using the construction in [Hum21].
Lemma 6.10.
For , assume that is generic. Then there exists a function such that
the support of is contained in , and is invariant under the action of .
Proof.
If , then is a quasi-character of . If is unramified, it is well-known that one takes with the characteristic function of , and has the following identity
which holds for all by meromorphic continuation , where is the uniformizer of . It is clear that in this case is supported on and is invariant under . If is ramified, then we know that . We can take a function such that if and otherwise. Then according to our normalization of the Haar measure, we obtain by an easy computation that
It is clear that in this case, is supported in and is invariant under the action of .
In the following, we assume that . For each non-negative integer , we define the congruence subgroup as in [Hum21] to be
According to the classification of irreducible generic representations and [JPSS81, Theorem 5], there is a minimal positive integer for which the vector space
is non-trivial and in fact of dimension one. Choose and , respectively, such that the matrix coefficient has value at . Since , we may take a Schwartz-Bruhat function of the form:
Then by [Hum21, Theorem 1.2], when is sufficiently positive, one has that
According to [JL23, Propositon 3.2, Theorem 3.4], the fiber integration as defined in (2.4) yields that
(6.22) |
where is the fiber at of the determinant map as in (2.3), is well defined and when is sufficiently positive, we have that
(6.23) |
It remains to verify the invariance property for this function . If , we must have that Thus for any with , we must have that . By the fiber integration in (6.22), we have that when . Moreover, for any , we have
where Since , one can see at once that and . Therefore we obtain that for any . ∎
Let be the particularly chosen function such that for ramified places we take as given in Lemma 6.10 since each local component of is irreducible and generic when . Denote by
(6.24) |
the Godement-Jacquet kernel as in (6.20), with replaced by the particularly chosen .
Proposition 6.11.
For any , the Godement-Jacquet kernel as defined in (6.24) enjoys the following expression:
as a function in for all .
Proof.
Write
(6.25) |
where is the local uniformizer in and runs over the algebraic direct sum . Consider the integral
(6.26) |
where is as given in (6.24). We know is supported on the the ring of -integers according to Lemma 6.10 and [JL23, Lemma 5.3]. It follows that we may assume for all . If belongs to the -component of (6.25), then , where is the cardinality of the residue field of . The range of the integral in (6.26) is equivalent to the condition that . Since the integrand in (6.26) is invariant under , we obtain that
(6.27) |
with . We may write any fractional ideal in , in a unique way, as with , and regard the function as
Then the function is supported on the set of integral ideals and (6.27) can be written as
for any fractional ideal . According to the normalization of the Haar measure, the integral (6.26) is equal to
where the summation runs over all the integral ideals of .
On the other hand, for the particularly given Schwartz function , we have that
for is sufficiently positive. By using the uniqueness of the coefficients of the Dirichlet series (see [MV07, Theorem 1.6]), we obtain that , and hence
∎
In order to define the dual kernel, we consider the local functional equation in the reformulation of the local Godement-Jacquet theory in [JL23, Theorem 3.10]:
(6.28) |
By Proposition 6.8 and [JL23, Theorem 3.4], we have that
Hence for sufficiently negative, we obtain that
(6.29) |
If we write
(6.30) |
with , then following the argument as in the proof of Proposition 6.11, we can obtain a Dirichlet series expression for the dual kernel of the Godement-Jacquet kernel .
Proposition 6.12.
Proof.
We first claim that each local component of is invariant under . In fact, if , the claim is clear because the classical Fourier transform of an -invariant functions is still -invariant by changing variables. If , at ramified places, since is as given by the fiber integration of and as in Lemma 6.10, we know from [JL23, Equation (3.17)] that
where is the classical Fourier transform given by (2.7). If for any , then we already know . Since
and by the definition of , we see that , we know is invariant under . At the remaining unramified places where , we know and by [JL23, Lemma 5.3] we know is invariant under . Let be as in Proposition 6.1 for . Then there are integers such that the support of is contained in
(6.31) |
Write . It is clear that is constant on each -component and supported on the -component with for and for . We may write any fraction ideal in in a unique way as and regard the function as a function on the set of fractional ideals sending to Then we obtain that
for in the -component, where . Write , where for ’s are defined from (6.31) and for we define . Then by the same argument, we obtain that
where the summation runs over all fractional ideal of that are contained in with norm and ∎
Therefore we obtain a -version of [Clo22, Theorem 1.1] when the kernel functions and are given in terms of the -function with its Dirichlet series expression.
References
- [Art13] J. Arthur. The endoscopic classification of representations: Orthogonal and symplectic groups, volume 61 of American Mathematical Society Colloquium Publications. AMS, Providence, RI, 2013.
- [Bar05] E. Baruch. Bessel functions for over a -adic field. In Automorphic representations, -functions and applications: progress and prospects, volume 11 of Ohio State Univ. Math. Res. Inst. Publ., pages 1–40. de Gruyter, Berlin, 2005.
- [BK00] A. Braverman and D. Kazhdan. -functions of representations and lifting. Geom. Funct. Anal., pages 237–278, 2000. With an appendix by V. Vologodsky, GAFA 2000 (Tel Aviv, 1999).
- [Clo22] Laurent Clozel. A universal lower bound for certain quadratic integrals of automorphic -functions. with an appendix by l. clozel and peter sarnak, 2022.
- [Cog14] J. Cogdell. Bessel functions for . Indian J. Pure Appl. Math., 45(5):557–582, 2014.
- [Con99] A. Connes. Trace formula in noncommutative geometry and the zeros of the Riemann zeta function. Selecta Math. (N.S.), 5(1):29–106, 1999.
- [Cor21] A. Corbett. Voronoï summation for : collusion between level and modulus. Amer. J. Math., 143(5):1361–1395, 2021.
- [CPS04] J. Cogdell and I. Piatetski-Shapiro. Remarks on Rankin-Selberg convolutions. In Contributions to automorphic forms, geometry, and number theory, pages 255–278. Johns Hopkins Univ. Press, Baltimore, MD, 2004.
- [GGPS69] I. Gelfand, M. Graev, and I. Pyatetskii-Shapiro. Representation theory and automorphic functions. W. B. Saunders Co., Philadelphia, Pa.-London-Toronto, Ont.,, 1969.
- [GJ72] R. Godement and H. Jacquet. Zeta functions of simple algebras. Lecture Notes in Mathematics, Vol. 260. Springer-Verlag, Berlin-New York, 1972.
- [GL06] D. Goldfeld and X. Li. Voronoi formulas on . Int. Math. Res. Not., pages Art. ID 86295, 25, 2006.
- [GL08] D. Goldfeld and X. Li. The Voronoi formula for . Int. Math. Res. Not. IMRN, (2):Art. ID rnm144, 39, 2008.
- [HR79] E. Hewitt and K. Ross. Abstract harmonic analysis. Vol. I, volume 115 of Grundlehren der Mathematischen Wissenschaften. Springer-Verlag, Berlin-New York, second edition, 1979.
- [Hum21] P. Humphries. Test vectors for non-Archimedean Godement-Jacquet zeta integrals. Bull. Lond. Math. Soc., 53(1):92–99, 2021.
- [Igu78] J.-I. Igusa. Forms of higher degree, volume 59 of Tata Institute of Fundamental Research Lectures on Mathematics and Physics. Tata Institute of Fundamental Research, Bombay; New Delhi, 1978.
- [IT13] A. Ichino and N. Templier. On the Voronoĭ formula for . Amer. J. Math., 135(1):65–101, 2013.
- [Jac79] H. Jacquet. Principal -functions of the linear group. In Automorphic forms, representations and -functions, Part 2, Proc. Sympos. Pure Math., XXXIII, pages 63–86. Amer. Math. Soc., Providence, R.I., 1979.
- [Jac09] H. Jacquet. Archimedean Rankin-Selberg integrals. In Automorphic forms and -functions II. Local aspects, volume 489 of Contemp. Math., pages 57–172. Amer. Math. Soc., Providence, RI, 2009.
- [JL70] H. Jacquet and R. Langlands. Automorphic forms on . Lecture Notes in Mathematics, Vol. 114. Springer-Verlag, Berlin-New York, 1970.
- [JL22] D. Jiang and Z. Luo. Certain Fourier operators on and local Langlands gamma functions. Pacific J. Math., 318(2):339–374, 2022.
- [JL23] D. Jiang and Z. Luo. Certain Fourier operators and their associated Poisson summation formulae on . Pacific J. Math., accepted, 2023.
- [JPSS81] H. Jacquet, I. Piatetski-Shapiro, and J. Shalika. Conducteur des représentations génériques du groupe linéaire. C. R. Acad. Sci. Paris Sér. I Math., 292(13):611–616, 1981.
- [JPSS83] H. Jacquet, I. Piatetskii-Shapiro, and J. Shalika. Rankin-Selberg convolutions. Amer. J. Math., 105(2):367–464, 1983.
- [Kna94] A. Knapp. Local Langlands correspondence: the Archimedean case. In Motives, volume 55 of Proc. Sympos. Pure Math., pages 393–410. Amer. Math. Soc., Providence, RI, 1994.
- [Kna01] A. Knapp. Representation theory of semisimple groups. Princeton Landmarks in Mathematics. Princeton University Press, Princeton, NJ, 2001.
- [MS04a] S. Miller and W. Schmid. Distributions and analytic continuation of Dirichlet series. J. Funct. Anal., 214(1):155–220, 2004.
- [MS04b] S. Miller and W. Schmid. Summation formulas, from Poisson and Voronoi to the present. In Noncommutative harmonic analysis, volume 220 of Progr. Math., pages 419–440. Birkhäuser Boston, MA, 2004.
- [MS06] S. Miller and W. Schmid. Automorphic distributions, -functions, and Voronoi summation for . Ann. of Math. (2), 164(2):423–488, 2006.
- [MS11] S. Miller and W. Schmid. A general Voronoi summation formula for . In Geometry and analysis. No. 2, volume 18 of Adv. Lect. Math. (ALM), pages 173–224. Int. Press, Somerville, MA, 2011.
- [MV07] H. Montgomery and R. Vaughan. Multiplicative number theory. I. Classical theory, volume 97 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2007.
- [Neu99] J. Neukirch. Algebraic number theory, volume 322 of Grundlehren der mathematischen Wissenschaften. Springer-Verlag, Berlin, 1999.
- [Ngo20] B. C. Ngo. Hankel transform, Langlands functoriality and functional equation of automorphic -functions. Jpn. J. Math., 15(1):121–167, 2020.
- [Qi20] Z. Qi. Theory of fundamental Bessel functions of high rank. Mem. Amer. Math. Soc., 267(1303), 2020.
- [Sou84] D. Soudry. The and factors for generic representations of over a local non-Archimedean field . Duke Math. J., 51(2):355–394, 1984.
- [Sou01] C. Soulé. On zeroes of automorphic -functions. In Dynamical, spectral, and arithmetic zeta functions, volume 290 of Contemp. Math., pages 167–179. Amer. Math. Soc., Providence, RI, 2001.
- [Tat67] J. Tate. Fourier analysis in number fields, and Hecke’s zeta-functions. In Algebraic Number Theory, pages 305–347. Thompson, Washington, D.C., 1967.
- [Tre67] F. Treves. Topological vector spaces, distributions and kernels. Academic Press, New York-London,, 1967.
- [Wei95] A. Weil. Fonction zêta et distributions. In Séminaire Bourbaki, Vol. 9, pages Exp. No. 312, 523–531. Soc. Math. France, Paris, 1995.
- [Zel75] D. Zelobenko. Complex harmonic analysis on semisimple Lie groups. In Proceedings of the International Congress of Mathematicians, Vol. 2, pages 129–134. Canad. Math. Congress, Montreal, Que., 1975.
- [ZN66] D. Zelobenko and M. Naimark. Description of completely irreducible representations of a semi-simple complex Lie group. Dokl. Akad. Nauk SSSR, 171:25–28, 1966.