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

The Voronoi Summation Formula for GLn{\mathrm{GL}}_{n} and the Godement-Jacquet Kernels

Dihua Jiang and Zhaolin Li School of Mathematics, University of Minnesota, 206 Church St. S.E., Minneapolis, MN 55455, USA. [email protected] [email protected]
Abstract.

Let 𝔸{\mathbb{A}} be the ring of adeles of a number field kk and π\pi be an irreducible cuspidal automorphic representation of GLn(𝔸){\mathrm{GL}}_{n}({\mathbb{A}}). In [JL22, JL23], the authors introduced π\pi-Schwartz space 𝒮π(𝔸×){\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}) and π\pi-Fourier transform π,ψ{\mathcal{F}}_{\pi,\psi} with a non-trivial additive character ψ\psi of k\𝔸k\backslash{\mathbb{A}}, proved the associated Poisson summation formula over 𝔸×{\mathbb{A}}^{\times}, based on the Godement-Jacquet theory for the standard LL-functions L(s,π)L(s,\pi), 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 GLn{\mathrm{GL}}_{n} 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 Hπ,sH_{\pi,s} and their dual kernels Kπ,sK_{\pi,s} for any irreducible cuspidal automorphic representation π\pi of GLn(𝔸){\mathrm{GL}}_{n}({\mathbb{A}}) and show in Theorems 6.9 and 6.13 that Hπ,sH_{\pi,s} and Kπ,1sK_{\pi,1-s} are related by the nonlinear π\pi_{\infty}-Fourier transform if and only if ss\in{\mathbb{C}} is a zero of Lf(s,πf)=0L_{f}(s,\pi_{f})=0, the finite part of the standard automorphic LL-function L(s,π)L(s,\pi), which are the (GLn,π)({\mathrm{GL}}_{n},\pi)-versions of [Clo22, Theorem 1.1], where the Tate kernel with n=1n=1 and π\pi 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 LL-function
2010 Mathematics Subject Classification:
Primary 11F66, 22E50, 43A32; Secondary 11F70, 22E53, 44A20
The research of this paper is supported in part by the NSF Grant DMS-2200890.

1. Introduction

1.1. π\pi-Poisson summation formula

Let kk be a number field and 𝔸=𝔸k{\mathbb{A}}={\mathbb{A}}_{k} the associated ring of adeles. For any irreducible cuspidal automorphic representation π\pi of GLn(𝔸){\mathrm{GL}}_{n}({\mathbb{A}}) for any integer n1n\geq 1, the Godement-Jacquet theory ([GJ72]) establishes the analytic continuation and functional equation for the standard LL-functions L(s,π)L(s,\pi). The key input from the harmonic analysis to the Godement-Jacquet theory is the classical Fourier analysis on the affine space Mn{\mathrm{M}}_{n}, the space of n×nn\times n-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 GLn(𝔸){\mathrm{GL}}_{n}({\mathbb{A}}). 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 GL1{\mathrm{GL}}_{1} for L(s,π)L(s,\pi) as a vast generalization of the 1950 thesis of J. Tate ([Tat67]). More precisely, for any irreducible cuspidal automorphic representation π\pi of GLn(𝔸){\mathrm{GL}}_{n}({\mathbb{A}}), the space of π\pi-Schwartz functions on GL1(𝔸)=𝔸×{\mathrm{GL}}_{1}({\mathbb{A}})={\mathbb{A}}^{\times} is defined, which is denoted by 𝒮π(𝔸×){\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}), and the π\pi-Foruier transform (or operator) π,ψ{\mathcal{F}}_{\pi,\psi} is defined for any given non-trivial additive character ψ\psi of k\𝔸k\backslash{\mathbb{A}}, which takes the π\pi-Schwartz space 𝒮π(𝔸×){\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}) to the π~\widetilde{\pi}-Schwartz space 𝒮π~(𝔸×){\mathcal{S}}_{\widetilde{\pi}}({\mathbb{A}}^{\times}), where π~\widetilde{\pi} is the contragredient of π\pi. We refer to Section 2 for details. The π\pi-Poisson summation formula as proved in [JL23, Theorem 4.7] (or recalled in Theorem 2.1) takes the following form.

Theorem 1.1 (π\pi-Poisson summation formula).

With the notation introduced above, the π\pi-theta function Θπ(x,ϕ):=αk×ϕ(αx)\Theta_{\pi}(x,\phi):=\sum_{\alpha\in k^{\times}}\phi(\alpha x) converges absolutely and locally uniformly for any x𝔸×x\in{\mathbb{A}}^{\times} and any ϕ𝒮π(𝔸×)\phi\in{\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}). Moreover, the following identity

(1.1) Θπ(x,ϕ)=Θπ~(x1,π,ψ(ϕ))\displaystyle\Theta_{\pi}(x,\phi)=\Theta_{\widetilde{\pi}}(x^{-1},{\mathcal{F}}_{\pi,\psi}(\phi))

holds as functions in x𝔸×x\in{\mathbb{A}}^{\times}.

This GL1{\mathrm{GL}}_{1}-reformulation in [JL22, JL23] of the Godement-Jacquet theory has found following nice applications, among others:

  1. (1)

    The local theory of such a GL1{\mathrm{GL}}_{1}-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 kπν,ψνk_{\pi_{\nu},\psi_{\nu}} on GL1{\mathrm{GL}}_{1}, 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 GL1{\mathrm{GL}}_{1} were considered.

  2. (2)

    The global theory of such a GL1{\mathrm{GL}}_{1}-reformation as developed in [JL23] gives the adelic formulation of A. Connes’ theorem ([Con99, Theorem III.1]) for L(s,π)L(s,\pi), and the complete version of C. Soulé’s theorem ([Sou01, Theorem 2]) that provides a spectral interpretation of the zeros of L(s,π)L(s,\pi). We refer to [JL23, Theorem 8.1] for details.

  3. (3)

    In Section 5 of this paper, the local and global theory of such a GL1{\mathrm{GL}}_{1}-reformulation in [JL22, JL23] provides a Poisson summation formula proof of the Voronoi formula for any irreducible cuspidal automorphic representation π\pi of GLn{\mathrm{GL}}_{n}, 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. (4)

    In Section 6 of this paper, the local and global theory of such a GL1{\mathrm{GL}}_{1}-reformulation in [JL22, JL23] defines the Godement-Jacquet kernels for L(s,π)L(s,\pi) and proves Theorems 6.9 and 6.13, which are the (GLn,π)({\mathrm{GL}}_{n},\pi)-versions of Clozel’s theorem ([Clo22, Theorem 1.1]) for any irreducible cuspidal automorphic representation π\pi of GLn{\mathrm{GL}}_{n}. In [Clo22], Clozel formulates and proves such a theorem for the Tate kernel associated with the Dedekind zeta function ζk(s)\zeta_{k}(s) for any number field kk.

1.2. Voronoi summation formula

The classical Voronoi summation formula and its recent extension to the GLn{\mathrm{GL}}_{n}-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 GLn{\mathrm{GL}}_{n} was first studied by S. Miller and W. Schmid in [MS06] for n=3n=3 and in [MS11] for general nn. 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 GLn×GL1{\mathrm{GL}}_{n}\times{\mathrm{GL}}_{1}, 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 GLn{\mathrm{GL}}_{n} has also been discussed in [GL06] and [GL08]. A complete treatment of the adelic approach to the Voronoi formula for GLn{\mathrm{GL}}_{n} over a general number field was given by A. Ichino and N. Templier in [IT13]. We recall their general Voromoi formula for GLn{\mathrm{GL}}_{n} below.

For each irreducible cuspidal automorphic representation π\pi of GLn(𝔸){\mathrm{GL}}_{n}({\mathbb{A}}), the Voronoi summation formula is an identity of two summations. One side of the identity is given by certain data associated with π\pi and the other side is given by certain corresponding data associated with π~\widetilde{\pi}, the contragredient of π\pi. Let ψ=νψν\psi=\otimes_{\nu}\psi_{\nu} be a non-trivial additive character on k\𝔸k\backslash{\mathbb{A}}. At each local place ν\nu of kk, to a smooth compactly supported function wν(x)𝒞c(kν×)w_{\nu}(x)\in{\mathcal{C}}_{c}^{\infty}(k_{\nu}^{\times}) is associated a dual function w~ν(x)\widetilde{w}_{\nu}(x) such that the following functional equation

(1.2) kν×w~ν(y)χν(y)1|y|νsn12d×y=γ(1s,πν×χν,ψν)kν×wν(y)χν(y)|y|ν1sn12d×y\displaystyle\int_{k_{\nu}^{\times}}\widetilde{w}_{\nu}(y)\chi_{\nu}(y)^{-1}|y|_{\nu}^{s-\frac{n-1}{2}}\,\mathrm{d}^{\times}y=\gamma(1-s,\pi_{\nu}\times\chi_{\nu},\psi_{\nu})\int_{k_{\nu}^{\times}}w_{\nu}(y)\chi_{\nu}(y)|y|_{\nu}^{1-s-\frac{n-1}{2}}\,\mathrm{d}^{\times}y

holds for all ss\in{\mathbb{C}} and all unitary characters χν\chi_{\nu} of kν×k_{\nu}^{\times}. Since any irreducible cuspidal automorphic representation π\pi of GLn(𝔸){\mathrm{GL}}_{n}({\mathbb{A}}) is generic, i.e. it has a non-zero Whittaker-Fourier coefficient. If we write π=ν|k|πν\pi=\otimes_{\nu\in|k|}\pi_{\nu}, where |k||k| denotes the set of all local places of kk, then at any local place ν\nu, the local component πν\pi_{\nu} is an irreducible admissible and generic representation of GLn(kν){\mathrm{GL}}_{n}(k_{\nu}). Let 𝒲(πν,ψν){\mathcal{W}}(\pi_{\nu},\psi_{\nu}) be the local Whittaker model of πν\pi_{\nu}, and Wν(g)W_{\nu}(g) be any Whittaker function on GLn(kν){\mathrm{GL}}_{n}(k_{\nu}) that belongs to 𝒲(πν,ψν){\mathcal{W}}(\pi_{\nu},\psi_{\nu}) (see Section 3 for the details).

Let SS be a finite set of |k||k| including all Archimedean places and the local places ν\nu where πν\pi_{\nu} or ψν\psi_{\nu} is ramified. As usual, we write 𝔸=𝔸S×𝔸S{\mathbb{A}}={\mathbb{A}}_{S}\times{\mathbb{A}}^{S}, where 𝔸S=νSkν{\mathbb{A}}_{S}=\prod_{\nu\in S}k_{\nu}, which is naturally embedded as a subring of 𝔸{\mathbb{A}}, and 𝔸S{\mathbb{A}}^{S} the subring of adeles 𝔸{\mathbb{A}} with trivial component above SS. At νS\nu\notin S, we take the unramified Whittaker vector Wν{}^{\circ}W_{\nu} of πν\pi_{\nu}, which is so normalized that Wν(In)=1{}^{\circ}W_{\nu}({\mathrm{I}}_{n})=1. Denote by WS:=νSWν{}^{\circ}W^{S}:=\prod_{\nu\notin S}{{}^{\circ}W_{\nu}}, which is the normalized unramified Whittaker function of πS=νSπν\pi^{S}=\otimes_{\nu\notin S}\pi_{\nu}. Similarly, we define W~S=νSW~ν{}^{\circ}\widetilde{W}^{S}=\prod_{\nu\notin S}{{}^{\circ}\widetilde{W}}_{\nu} to be the (normalized) unramified Whittaker function of π~S=νSπ~ν\widetilde{\pi}^{S}=\otimes_{\nu\notin S}\widetilde{\pi}_{\nu}. We recall that the functions WS{{}^{\circ}W^{S}} and W~S{{}^{\circ}\widetilde{W}^{S}} are related by the following

W~S(g)=WS(wng1t){{}^{\circ}\widetilde{W}^{S}(g)}={{}^{\circ}W^{S}}(w_{n}{{}^{t}g^{-1}})

for all gGn(𝔸S)g\in{\mathrm{G}}_{n}({\mathbb{A}}^{S}), where wnw_{n} is the longest Weyl element of Gn=GLn{\mathrm{G}}_{n}={\mathrm{GL}}_{n} 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 ζ𝔸S\zeta\in{\mathbb{A}}^{S}, let R=RζR=R_{\zeta} be the set of places ν\nu such that |ζν|>1|\zeta_{\nu}|>1. At each νS\nu\in S let wν𝒞c(kν×)w_{\nu}\in{\mathcal{C}}_{c}^{\infty}(k_{\nu}^{\times}). Then:

αk×ψ(αζ)WS((αIn1))wS(α)=αk×KlR(α,ζ,W~R)W~RS((αIn1))w~S(α),\sum_{\alpha\in k^{\times}}\psi(\alpha\zeta)\cdot{{}^{\circ}W^{S}}\left(\begin{pmatrix}\alpha&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)w_{S}(\alpha)=\sum_{\alpha\in k^{\times}}{\mathrm{Kl}}_{R}(\alpha,\zeta,{{}^{\circ}\widetilde{W}_{R}})\cdot{{}^{\circ}\widetilde{W}^{R\cup S}}\left(\begin{pmatrix}\alpha&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)\widetilde{w}_{S}(\alpha),

where wS(α):=νSwν(α)w_{S}(\alpha):=\prod_{\nu\in S}w_{\nu}(\alpha) and the same for w~S(α)\widetilde{w}_{S}(\alpha), and KlR(α,ζ,W~R){\mathrm{Kl}}_{R}(\alpha,\zeta,{{}^{\circ}\widetilde{W}_{R}}) is a finite Euler product of the local Kloosterman integrals:

KlR(α,ζ,W~R):=νRKlν(α,ζν,W~ν).{\mathrm{Kl}}_{R}(\alpha,\zeta,{{}^{\circ}{\widetilde{W}_{R}}}):=\prod_{\nu\in R}{\mathrm{Kl}}_{\nu}(\alpha,\zeta_{\nu},{{}^{\circ}\widetilde{W}_{\nu}}).

For the place ν\nu, the local Kloosterman integral Klν(α,ζν,W~ν){\mathrm{Kl}}_{\nu}(\alpha,\zeta_{\nu},{{}^{\circ}\widetilde{W}_{\nu}}) is defined by

Klν(α,ζν,W~ν):=|ζν|νn2Uτ(Fν)ψν(un2,n1)¯W~ν(τu)du,{\mathrm{Kl}}_{\nu}(\alpha,\zeta_{\nu},{{}^{\circ}\widetilde{W}_{\nu}}):=|\zeta_{\nu}|_{\nu}^{n-2}\int_{U^{-}_{\tau}(F_{\nu})}\overline{\psi_{\nu}(u_{n-2,n-1})}{{}^{\circ}\widetilde{W}_{\nu}}(\tau u)\,\mathrm{d}u,

where

Uτ={(In20010001)}andτ=(010In200001)(In2000αζν1000ζν),U^{-}_{\tau}=\left\{\begin{pmatrix}{\mathrm{I}}_{n-2}&*&0\\ 0&1&0\\ 0&0&1\end{pmatrix}\right\}\quad{\rm and}\quad\tau=\begin{pmatrix}0&1&0\\ {\mathrm{I}}_{n-2}&0&0\\ 0&0&1\end{pmatrix}\begin{pmatrix}{\mathrm{I}}_{n-2}&0&0\\ 0&-\alpha\zeta_{\nu}^{-1}&0\\ 0&0&-\zeta_{\nu}\end{pmatrix},

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 GLn×GL1{\mathrm{GL}}_{n}\times{\mathrm{GL}}_{1} ([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 π\pi-Poisson summation formula on GL1{\mathrm{GL}}_{1} 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 LL-functions of GLn×GL1{\mathrm{GL}}_{n}\times{\mathrm{GL}}_{1}.

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 π\pi-Schwartz function on 𝔸×{\mathbb{A}}^{\times} and are related by the π\pi-Fourier transform π,ψ{\mathcal{F}}_{\pi,\psi} in the sense of [JL22, JL23]. More precisely, for any irreducible smooth representation πν\pi_{\nu} of GLn(kν){\mathrm{GL}}_{n}(k_{\nu}), which is of Casselman-Wallach type if ν\nu is an infinite local place of kk, we define the πν\pi_{\nu}-Bessel function 𝔟πν,ψν(x){\mathfrak{b}}_{\pi_{\nu},\psi_{\nu}}(x) on kν×k_{\nu}^{\times} (Definitions 4.2, 4.9 and 4.12) and obtain a series of results on the relations between the πν\pi_{\nu}-Bessel functions 𝔟πν,ψν(x){\mathfrak{b}}_{\pi_{\nu},\psi_{\nu}}(x), the πν\pi_{\nu}-Fourier transforms πν,ψν{\mathcal{F}}_{\pi_{\nu},\psi_{\nu}} and the πν\pi_{\nu}-kernel functions kπν,ψν(x)k_{\pi_{\nu},\psi_{\nu}}(x) as introduced and studied in [JL22, JL23] (see (2.18) and (2.20) for details), and on new formulas for the dual functions wν~(x)\widetilde{w_{\nu}}(x) of wν(x)𝒞c(kν×)w_{\nu}(x)\in{\mathcal{C}}^{\infty}_{c}(k_{\nu}^{\times}). After all the local preparation, we deduce the Voromoi formula in Theorem 1.2 from the π\pi-Poisson summation formula in Theorem 1.1. We summarize those local results as the following theorem.

Theorem 1.3.

For any local place ν\nu of the number field kk, let πν\pi_{\nu} be an irreducible smooth representation πν\pi_{\nu} of GLn(kν){\mathrm{GL}}_{n}(k_{\nu}), which is of Casselman-Wallach type if ν\nu is infinite. For any wν(x)𝒞c(kν×)w_{\nu}(x)\in{\mathcal{C}}^{\infty}_{c}(k_{\nu}^{\times}), wν~(x)\widetilde{w_{\nu}}(x) is the dual function of wν(x)w_{\nu}(x) as (1.2) or in Theorem 1.2. Then the following hold.

  • (1)

    The πν\pi_{\nu}-Fourier transform πν,ψν{\mathcal{F}}_{\pi_{\nu},\psi_{\nu}} realizes the duality between wν(x)w_{\nu}(x) and wν~(x)\widetilde{w_{\nu}}(x), up to normalization,

    πν,ψν(wν()||ν1n2)(x)=wν~(x)|x|ν1n2,xkν×.{\mathcal{F}}_{\pi_{\nu},\psi_{\nu}}(w_{\nu}(\cdot)|\cdot|_{\nu}^{1-\frac{n}{2}})(x)=\widetilde{w_{\nu}}(x)|x|_{\nu}^{1-\frac{n}{2}},\quad\forall x\in k_{\nu}^{\times}.
  • (2)

    The dual function w~(x)\widetilde{w}(x) of wν(x)w_{\nu}(x) enjoys the following formula:

    wν~(x)=|x|νn21(kπν,ψν()(wν()||νn21))(x),xkν×,\widetilde{w_{\nu}}(x)=|x|_{\nu}^{\frac{n}{2}-1}\left(k_{\pi_{\nu},\psi_{\nu}}(\cdot)*(w_{\nu}^{\vee}(\cdot)|\cdot|_{\nu}^{\frac{n}{2}-1})\right)(x),\quad\forall x\in k_{\nu}^{\times},

    where kπν,ψν(x)k_{\pi_{\nu},\psi_{\nu}}(x) is the πν\pi_{\nu}-kernel function of πν\pi_{\nu} as in (2.20) and wν(x)=wν(x1)w_{\nu}^{\vee}(x)=w_{\nu}(x^{-1}).

  • (3)

    As distributions on kν×k_{\nu}^{\times}, the πν\pi_{\nu}-kernel function kπν,ψνk_{\pi_{\nu},\psi_{\nu}} as in (2.20) and the πν\pi_{\nu}-Bessel function are related by the following identity:

    kπν,ψν(x)=𝔟πν,ψν(x)|x|ν12,xkν×.k_{\pi_{\nu},\psi_{\nu}}(x)={\mathfrak{b}}_{\pi_{\nu},\psi_{\nu}}(x)|x|_{\nu}^{\frac{1}{2}},\quad\forall x\in k_{\nu}^{\times}.
  • (4)

    The dual function wν~(x)\widetilde{w_{\nu}}(x) of wν(x)w_{\nu}(x) enjoys the following formula:

    wν~(x)=|x|νn12(𝔟πν,ψν()(wν()||νn32))(x),xkν×.\widetilde{w_{\nu}}(x)=|x|_{\nu}^{\frac{n-1}{2}}\left({\mathfrak{b}}_{\pi_{\nu},\psi_{\nu}}(\cdot)*(w_{\nu}^{\vee}(\cdot)|\cdot|_{\nu}^{\frac{n-3}{2}})\right)(x),\quad\forall x\in k_{\nu}^{\times}.

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 πν\pi_{\nu}-Bessel functions 𝔟πν,ψν(x){\mathfrak{b}}_{\pi_{\nu},\psi_{\nu}}(x) in the pp-adic case is defined by means of the Whittaker model of πν\pi_{\nu} following the general framework of E. Baruch in [Bar05]. Hence we have to assume in the pp-adic case that πν\pi_{\nu} is generic in the definition of the πν\pi_{\nu}-Bessel functions 𝔟πν,ψν(x){\mathfrak{b}}_{\pi_{\nu},\psi_{\nu}}(x). 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 GLn{\mathrm{GL}}_{n} of Casselman-Wallach type. Hence in the real or complex case, the definition of the πν\pi_{\nu}-Bessel functions 𝔟πν,ψν(x){\mathfrak{b}}_{\pi_{\nu},\psi_{\nu}}(x) does not require that πν\pi_{\nu} is generic. Since the πν\pi_{\nu}-kernel function kπν,ψνk_{\pi_{\nu},\psi_{\nu}} as in (2.20) is defined based on the Godement-Jacquet theory, which does not require that πν\pi_{\nu} is generic, the uniform result in Part (3) of Theorem 1.3 suggests that one may define the πν\pi_{\nu}-kernel function kπν,ψνk_{\pi_{\nu},\psi_{\nu}} to be the πν\pi_{\nu}-Bessel functions 𝔟πν,ψν(x){\mathfrak{b}}_{\pi_{\nu},\psi_{\nu}}(x) in the pp-adic case when πν\pi_{\nu} is not generic. Finally, let us mention that the πν\pi_{\nu}-Bessel functions 𝔟πν,ψν(x){\mathfrak{b}}_{\pi_{\nu},\psi_{\nu}}(x) 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 |k|=|k||k|f|k|=|k|_{\infty}\cup|k|_{f}, where |k||k|_{\infty} denotes the subset of |k||k| consisting of all Archimedean local places of kk, and |k|f|k|_{f} denotes the subset of |k||k| consisting of all finite local places of kk. Write 𝔸=ν|k|kν{\mathbb{A}}_{\infty}=\prod_{\nu\in|k|_{\infty}}k_{\nu}. For x=(xν)𝔸x=(x_{\nu})\in{\mathbb{A}}_{\infty}, set |x|:=ν|k||xν|ν|x|_{\infty}:=\prod_{\nu\in|k|_{\infty}}|x_{\nu}|_{\nu}. Let 𝒪=𝒪k{\mathcal{O}}={\mathcal{O}}_{k} be the ring of algebraic integers of kk. L. Clozel defines in [Clo22] the Tate kernel:

(1.3) Hs(x):=Xs1𝔑(𝔞)X𝔑(𝔞)sκ1s\displaystyle H_{s}(x):=X^{s-1}\sum_{{\mathfrak{N}}({\mathfrak{a}})\leq X}{\mathfrak{N}}({\mathfrak{a}})^{-s}-\frac{\kappa}{1-s}

for x𝔸×x\in{\mathbb{A}}_{\infty}^{\times} with X=|x|X=|x|_{\infty}, where 𝔞{\mathfrak{a}} runs over nonzero integral ideals 𝔞𝒪{\mathfrak{a}}\subset{\mathcal{O}}, and κ=Ress=1ζk(s)\kappa={\mathrm{Res}}_{s=1}\zeta_{k}(s). Here ζk(s)=𝔞𝒪𝔑(𝔞)s\zeta_{k}(s)=\sum_{{\mathfrak{a}}\subset{\mathcal{O}}}{\mathfrak{N}}({\mathfrak{a}})^{-s} is the Dedekind zeta function of kk with 𝔑(𝔞):=|Nk/𝔞|{\mathfrak{N}}({\mathfrak{a}}):=|N_{k/{\mathbb{Q}}}{\mathfrak{a}}|, the absolute norm of 𝔞{\mathfrak{a}}; and the dual kernel

(1.4) Ks(x):=D12Xs1𝔞𝒟1,𝔑(𝔞)X𝔑(𝔞)sκD121s,\displaystyle K_{s}(x):=D^{-\frac{1}{2}}X^{s-1}\sum_{{\mathfrak{a}}\subset\mathcal{D}^{-1},{\mathfrak{N}}({\mathfrak{a}})\leq X}{\mathfrak{N}}({\mathfrak{a}})^{-s}-\frac{\kappa\cdot D^{\frac{1}{2}}}{1-s},

where 𝒟=𝒟k\mathcal{D}=\mathcal{D}_{k} is the difference of 𝒪{\mathcal{O}} and 𝒟1\mathcal{D}^{-1} is the inverse difference; and D=𝔑(𝒟)D={\mathfrak{N}}(\mathcal{D}) is the absolute value of the discriminant. Theorem 1.1 of [Clo22] expresses the relation between those tempered distributions Hs(x)H_{s}(x) and Ks(x)K_{s}(x) on 𝔸×{\mathbb{A}}_{\infty}^{\times} in terms of the condition: ζk(s)=0\zeta_{k}(s)=0 with σ=Re(s)(0,1)\sigma={\mathrm{Re}}(s)\in(0,1), which is more precisely stated as follows.

Theorem 1.4 (Clozel).

Assume that σ=Re(s)(0,1)\sigma={\mathrm{Re}}(s)\in(0,1). Then ζk(s)=0\zeta_{k}(s)=0 if and only if

(Hs)=K1s{\mathcal{F}}_{\infty}(H_{s})=-K_{1-s}

where {\mathcal{F}}_{\infty} is the usual Fourier transform over 𝔸{\mathbb{A}}_{\infty} with a suitable normalized measure.

The second global result of this paper is to define the kernel functions for any irreducible cuspidal automorphic representation π\pi of GLn(𝔸){\mathrm{GL}}_{n}({\mathbb{A}}), which will be called the Godement-Jacquet kernels, and prove the analogy of Theorem 1.4 for the Godement-Jacquet kernels and the standard LL-functions L(s,π)L(s,\pi) (see Theorems 6.9 and 6.13 for the exact statements).

Let π\pi be an irreducible cuspidal automorphic representation of GLn(𝔸){\mathrm{GL}}_{n}({\mathbb{A}}) and write π=ππf\pi=\pi_{\infty}\otimes\pi_{f}, where πf:=p<πp\pi_{f}:=\otimes_{p<\infty}\pi_{p}, and write the standard LL-function of π\pi as

(1.5) L(s,π)=L(s,π)Lf(s,πf)\displaystyle L(s,\pi)=L_{\infty}(s,\pi_{\infty})\cdot L_{f}(s,\pi_{f})

for Re(s){\mathrm{Re}}(s) sufficiently positive. As usual, L(s,π)L(s,\pi) is called the complete LL-function associated with π\pi, and Lf(s,πf)L_{f}(s,\pi_{f}) is called the finite part of the LL-function associated with π\pi. The local and global theory of R. Godement and H. Jacquet in [GJ72] introduces the global zeta integrals for L(s,π)L(s,\pi) and proves that L(s,π)L(s,\pi) has analytic continuation to an entire function in ss\in{\mathbb{C}} and satisfies the functional equation

L(s,π)=ϵ(s,π)L(1s,π~).L(s,\pi)=\epsilon(s,\pi)\cdot L(1-s,\widetilde{\pi}).

Following the reformulation as developed in [JL23], for an irreducible cuspidal automorphic representation π\pi of GLn(𝔸){\mathrm{GL}}_{n}({\mathbb{A}}), there exists a π\pi-Schwartz space 𝒮π(𝔸×){\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}) as defined in (2.23), which defines the GL1{\mathrm{GL}}_{1}-zeta integral

(1.6) 𝒵(s,ϕ)=𝔸×ϕ(x)|x|𝔸s12d×x\displaystyle{\mathcal{Z}}(s,\phi)=\int_{{\mathbb{A}}^{\times}}\phi(x)|x|_{\mathbb{A}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x

for any ϕ𝒮π(𝔸×)\phi\in{\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}). By [JL23, Theorem 4.6] the zeta integral 𝒵(s,ϕ){\mathcal{Z}}(s,\phi) converges absolutely for Re(s)>n+12{\mathrm{Re}}(s)>\frac{n+1}{2}, admits analytic continuation to an entire function in ss\in{\mathbb{C}}, and satisfies the functional equation

(1.7) 𝒵(s,ϕ)=𝒵(1s,π,ψ(ϕ)).\displaystyle{\mathcal{Z}}(s,\phi)={\mathcal{Z}}(1-s,{\mathcal{F}}_{\pi,\psi}(\phi)).

where π,ψ{\mathcal{F}}_{\pi,\psi} is the π\pi-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 L(s,π)L(s,\pi) in Definition 6.5, which can be briefly explained as follows.

Write x𝔸×x\in{\mathbb{A}}^{\times} as x=xxfx=x_{\infty}\cdot x_{f} with x𝔸×x_{\infty}\in{\mathbb{A}}_{\infty}^{\times} and xf𝔸f×x_{f}\in{\mathbb{A}}_{f}^{\times}. Set 𝔸>1:={x𝔸×:|x|𝔸>1}{\mathbb{A}}^{>1}:=\{x\in{\mathbb{A}}^{\times}\ \colon\ |x|_{\mathbb{A}}>1\}. For x𝔸>1x\in{\mathbb{A}}^{>1}, we have that |x|=|x|𝔸|xf|𝔸>1|x|=|x_{\infty}|_{{\mathbb{A}}}\cdot|x_{f}|_{{\mathbb{A}}}>1 and |xf|𝔸>|x|𝔸1|x_{f}|_{{\mathbb{A}}}>|x_{\infty}|_{{\mathbb{A}}}^{-1}. Write

𝒮π(𝔸×)=𝒮π(𝔸×)𝒮πf(𝔸f×).{\mathcal{S}}_{\pi}({\mathbb{A}}^{\times})={\mathcal{S}}_{\pi_{\infty}}({\mathbb{A}}_{\infty}^{\times})\otimes{\mathcal{S}}_{\pi_{f}}({\mathbb{A}}_{f}^{\times}).

For ϕ=ϕϕf𝒮π(𝔸×)\phi=\phi_{\infty}\otimes\phi_{f}\in{\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}) with ϕ𝒮π(𝔸×)\phi_{\infty}\in{\mathcal{S}}_{\pi_{\infty}}({\mathbb{A}}_{\infty}^{\times}) and ϕf𝒮πf(𝔸f×)\phi_{f}\in{\mathcal{S}}_{\pi_{f}}({\mathbb{A}}_{f}^{\times}), we write

(1.8) 𝔸>1ϕ(x)|x|𝔸s12d×x=𝔸×ϕ(x)|x|𝔸s12d×x𝔸f×>|x|1ϕf(xf)|xf|𝔸s12d×xf,\displaystyle\int_{{\mathbb{A}}^{>1}}\phi(x)|x|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x=\int_{{\mathbb{A}}_{\infty}^{\times}}\phi_{\infty}(x_{\infty})|x_{\infty}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x_{\infty}\int_{{\mathbb{A}}_{f}^{\times}}^{>|x_{\infty}|^{-1}}\phi_{f}(x_{f})|x_{f}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x_{f},

for any ss\in{\mathbb{C}}, where the inner integral is taken over the domain {xf𝔸f×:|xf|𝔸>|x|𝔸1}\{x_{f}\in{\mathbb{A}}_{f}^{\times}\ \colon\ |x_{f}|_{{\mathbb{A}}}>|x_{\infty}|_{{\mathbb{A}}}^{-1}\}. Proposition 6.1 shows that the integral converges absolutely for any ss\in{\mathbb{C}} and is holomorphic in ss\in{\mathbb{C}}. By the Fubini theorem and the support in 𝔸f×{\mathbb{A}}_{f}^{\times} of ϕf\phi_{f}, which is a fractional ideal of kk (Proposition 6.6) , the inner integral 𝔸f×>|x|𝔸1ϕf(xf)|xf|𝔸s12d×xf\int_{{\mathbb{A}}_{f}^{\times}}^{>|x_{\infty}|_{{\mathbb{A}}}^{-1}}\phi_{f}(x_{f})|x_{f}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x_{f} converges absolutely for any ss\in{\mathbb{C}} and any x𝔸×x_{\infty}\in{\mathbb{A}}_{\infty}^{\times}. The Godement-Jacquet kernel for L(s,π)L(s,\pi) is defined by

(1.9) Hπ,s(x,ϕf):=|x|𝔸s12𝔸f×>|x|𝔸1ϕf(xf)|xf|𝔸s12d×xf,\displaystyle H_{\pi,s}(x_{\infty},\phi_{f}):=|x_{\infty}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\int_{{\mathbb{A}}_{f}^{\times}}^{>|x_{\infty}|_{{\mathbb{A}}}^{-1}}\phi_{f}(x_{f})|x_{f}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x_{f},

for x𝔸×x_{\infty}\in{\mathbb{A}}_{\infty}^{\times} and for all ss\in{\mathbb{C}}. Clozel defines in [Clo22] the dual kernel for the case of GL1{\mathrm{GL}}_{1} with π\pi the trivial character. We define here the dual kernel of the Godement-Jacquet kernel Hπ,s(x,ϕf)H_{\pi,s}(x_{\infty},\phi_{f}) for L(s,π)L(s,\pi) to be

(1.10) Kπ,s(x,ϕf):=|x|𝔸s12𝔸f×>|x|𝔸1πf,ψf(ϕf)(xf)|xf|𝔸s12d×xf,\displaystyle K_{\pi,s}(x_{\infty},\phi_{f}):=|x_{\infty}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\int_{{\mathbb{A}}_{f}^{\times}}^{>|x_{\infty}|_{{\mathbb{A}}}^{-1}}{\mathcal{F}}_{\pi_{f},\psi_{f}}(\phi_{f})(x_{f})|x_{f}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x_{f},

for x𝔸×x_{\infty}\in{\mathbb{A}}_{\infty}^{\times} and for all ss\in{\mathbb{C}}. Proposition 6.6 shows that both kernel functions Hπ,s(x,ϕf)H_{\pi,s}(x_{\infty},\phi_{f}) and Kπ,s(x,ϕf)K_{\pi,s}(x_{\infty},\phi_{f}) on 𝔸×{\mathbb{A}}_{\infty}^{\times} can be extended uniquely to tempered distributions on 𝔸{\mathbb{A}}_{\infty} for any ϕf𝒮πf(𝔸f×)\phi_{f}\in{\mathcal{S}}_{\pi_{f}}({\mathbb{A}}_{f}^{\times}) and for any ss\in{\mathbb{C}}, by using the work of S. Miller and W. Schmid in [MS04a]. We are able to match the kernels Hπ,s(x,ϕf)H_{\pi,s}(x_{\infty},\phi_{f}) and Kπ,s(x,ϕf)K_{\pi,s}(x_{\infty},\phi_{f}) with the Euler product expression or Dirichlet series expression of the finite part LL-function Lf(s,πf)L_{f}(s,\pi_{f}) by specifically choosing the πf\pi_{f}-Schwartz functions ϕf𝒮πf(𝔸f×)\phi_{f}\in{\mathcal{S}}_{\pi_{f}}({\mathbb{A}}_{f}^{\times}), and prove in Section 6 the π\pi-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 π\pi of GLn(𝔸){\mathrm{GL}}_{n}({\mathbb{A}}), with a choice of ϕ=ϕϕf𝒮π(𝔸×)\phi^{\star}=\phi_{\infty}\otimes\phi^{\star}_{f}\in{\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}), the Godement-Jacquet kernel Hπ,s(x)=Hπ,s(x,ϕf)H_{\pi,s}(x)=H_{\pi,s}(x,\phi_{f}^{\star}) and its dual kernel Kπ,s(x)=Kπ,s(x,ϕf)K_{\pi,s}(x)=K_{\pi,s}(x,\phi_{f}^{\star}) enjoy the following identity:

(1.11) Hπ,s(x)=π,ψ(Kπ,1s)(x)=𝔸×kπ,ψ(xy)Kπ,1s(y)d×y.\displaystyle H_{\pi,s}(x)=-{\mathcal{F}}_{\pi_{\infty},\psi_{\infty}}(K_{\pi,1-s})(x)=-\int_{{\mathbb{A}}_{\infty}^{\times}}k_{\pi_{\infty},\psi_{\infty}}(xy)K_{\pi,1-s}(y)\,\mathrm{d}^{\times}y.

as distributions on 𝔸×{\mathbb{A}}_{\infty}^{\times} if and only if ss is a zero of Lf(s,πf)L_{f}(s,\pi_{f}).

Theorem 6.13 is a more precise version of Theorem 1.5 (and Theorem 6.9), which is the exact π\pi-analogy of Theorem 1.4.

1.4. Organiztion of the paper

We recall the π\pi-Poisson summation formula on GL1{\mathrm{GL}}_{1} 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 π\pi-Schwartz spaces and local π\pi-Fourier operators developed in [JL22] and [JL23]. Based on their work as well as the work of [GJ72], we recall the formulation of the π\pi-Poission summation formula on GL1{\mathrm{GL}}_{1} in [JL23] in Section 2.3.

Sections 3 and 4 are devoted to understand the duality between the function wν(x)w_{\nu}(x) and the function wν~(x)\widetilde{w_{\nu}}(x) by means of the harmonic analysis on GL1{\mathrm{GL}}_{1} as developed in [JL22] and [JL23], and to prove our main local results (Theorem 1.3). By comparing the Godement-Jacquet theory with the GLn×GL1{\mathrm{GL}}_{n}\times{\mathrm{GL}}_{1} Rankin-Selberg convolution, we are able to express the dual function wν~(x)\widetilde{w_{\nu}}(x) of wν(x)w_{\nu}(x) in terms of the πν\pi_{\nu}-Fourier transform πν,ψν{\mathcal{F}}_{\pi_{\nu},\psi_{\nu}} up to certain normalization (Proposition 3.5), based on Proposition 3.1 that identifies the πν\pi_{\nu}-Schwartz space 𝒮πν(kν×){\mathcal{S}}_{\pi_{\nu}}(k_{\nu}^{\times}), as introduced in [JL23] and recalled in (2.5), with the πν\pi_{\nu}-Whittaker-Schwartz space 𝒲πν(kν×){\mathcal{W}}_{\pi_{\nu}}(k_{\nu}^{\times}) as defined in (3.1). As a consequence, we obtain a formula that express the dual function wν~(x)\widetilde{w_{\nu}}(x) of wν(x)w_{\nu}(x) as a convolution of the πν\pi_{\nu}-kernel function kπν,ψν(x)k_{\pi_{\nu},\psi_{\nu}}(x) with wν(x)w_{\nu}(x), up to certain normalization (Corollary 3.6). In Section 4, we introduce the notion of πν\pi_{\nu}-Bessel functions 𝔟πν,ψν(x){\mathfrak{b}}_{\pi_{\nu},\psi_{\nu}}(x) (Definitions 4.2, 4.9 and 4.12) and prove the precise relation between the πν\pi_{\nu}-Bessel functions and πν\pi_{\nu}-kernel functions as defined in (2.20) (Propositions 4.3, 4.8, and 4.13). In the pp-adic case, the πν\pi_{\nu}-Bessel function 𝔟πν,ψν(x){\mathfrak{b}}_{\pi_{\nu},\psi_{\nu}}(x) on kν×k_{\nu}^{\times} 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 πν\pi_{\nu}-Bessel function 𝔟πν,ψν(x){\mathfrak{b}}_{\pi_{\nu},\psi_{\nu}}(x) on kν×k_{\nu}^{\times} by following the general theory of Bessel functions of high rank by Z. Qi in [Qi20]. It should be mentioned that the πν\pi_{\nu}-Bessel function 𝔟πν,ψν(x){\mathfrak{b}}_{\pi_{\nu},\psi_{\nu}}(x) on ×{\mathbb{R}}^{\times} in the real case is more general than the one considered in [Qi20] (Remark 4.14). As expected, when n=2n=2 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 GLn{\mathrm{GL}}_{n} (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 GLn×GL1{\mathrm{GL}}_{n}\times{\mathrm{GL}}_{1} 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 GLn{\mathrm{GL}}_{n} as in Theorem 1.2 is a special case of the π\pi-Poission summation formula on GL1{\mathrm{GL}}_{1} 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 π\pi-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 Hπ,s(x,ϕf)H_{\pi,s}(x,\phi_{f}) and their dual kernels Kπ,s(x,ϕf)K_{\pi,s}(x,\phi_{f}) (Definition 6.5) for any irreducible cuspidal automorphic representation π\pi of GLn(𝔸){\mathrm{GL}}_{n}({\mathbb{A}}), we develop further properties (Propositions 6.1 and 6.3, and Corollary 6.4) of the global zeta integrals 𝒵(s,ϕ){\mathcal{Z}}(s,\phi), as defined in (1.6), by using the π\pi-Fourier transform π,ψ{\mathcal{F}}_{\pi,\psi} and the associated π\pi-Poisson summation formula as developed in [JL23]. In Proposition 6.6, we show that both kernel functions Hπ,s(x,ϕf)H_{\pi,s}(x_{\infty},\phi_{f}) and Kπ,s(x,ϕf)K_{\pi,s}(x_{\infty},\phi_{f}) on 𝔸×{\mathbb{A}}_{\infty}^{\times} can be extended uniquely to tempered distributions on 𝔸{\mathbb{A}}_{\infty} for any ϕf𝒮πf(𝔸f×)\phi_{f}\in{\mathcal{S}}_{\pi_{f}}({\mathbb{A}}_{f}^{\times}) and for any ss\in{\mathbb{C}}. In Section 6.2, guided by Theorem 1.4, we prove in Proposition 6.7 that if ss\in{\mathbb{C}} is a zero of Lf(s,πf)L_{f}(s,\pi_{f}), then the kernel Hπ,s(x,ϕf)H_{\pi,s}(x_{\infty},\phi_{f}) is equal to the negative of π\pi_{\infty}-Fourier transform of the dual kernel Kπ,1s(x,ϕf)K_{\pi,1-s}(x_{\infty},\phi_{f}). For any ϕ𝒮π(𝔸×)\phi_{\infty}\in{\mathcal{S}}_{\pi_{\infty}}({\mathbb{A}}_{\infty}^{\times}), take ϕ=ϕϕf\phi^{\star}=\phi_{\infty}\otimes\phi_{f}^{\star}, where ϕf:=νϕν\phi_{f}^{\star}:=\otimes_{\nu}\phi_{\nu} with ϕν\phi_{\nu} as given in Proposition 6.8. Theorem 6.9 proves the π\pi-version of Theorem 1.4 for the Euler product expression of Lf(s,πf)L_{f}(s,\pi_{f}). 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 π\pi-version of Theorem 1.4 for the Dirichlet series expression of Lf(s,πf)L_{f}(s,\pi_{f}).

2. GL1{\mathrm{GL}}_{1}-Reformulation of the Godement-Jacquet Theory

We recall from [JL22, JL23] the GL1{\mathrm{GL}}_{1}-reformulation of the Godement-Jacquet theory for the standard LL-functions L(s,π)L(s,\pi) associated with any irreducible cuspidal automorphic representation π\pi of GLn{\mathrm{GL}}_{n} (for n1n\geq 1), where 𝔸{\mathbb{A}} is the ring of adeles of a number field kk. More precisely, we recall the π\pi-Schwartz spaces on GL1{\mathrm{GL}}_{1} and the π\pi-Fourier operators over GL1{\mathrm{GL}}_{1} both for the local and global cases, and the π\pi-Poisson summation formula on GL1{\mathrm{GL}}_{1}.

2.1. π\pi-Schwartz functions

Let |k||k| be the set of all local places of kk. For any local place ν\nu, we denote by F=kνF=k_{\nu}, the local field of kk at ν\nu. If FF is non-Archimedean, we denote by 𝔬=𝔬F{\mathfrak{o}}={\mathfrak{o}}_{F} the ring of integers and by 𝔭=𝔭F{\mathfrak{p}}={\mathfrak{p}}_{F} the maximal ideal of 𝔬{\mathfrak{o}}. Let Gn=GLn{\mathrm{G}}_{n}={\mathrm{GL}}_{n} be the general linear group defined over FF. Fix the maximal compact subgroups KK of Gn(F){\mathrm{G}}_{n}(F), where K=GLn(𝔬F)K={\mathrm{GL}}_{n}({\mathfrak{o}}_{F}) if FF is non-Archimedean, K=OnK={\mathrm{O}}_{n} if F=F={\mathbb{R}}, and K=UnK={\mathrm{U}}_{n} if F=F={\mathbb{C}}.

Let Mn(F){\mathrm{M}}_{n}(F) be the space of all n×nn\times n matrices over FF and 𝒮(Mn(F)){\mathcal{S}}({\mathrm{M}}_{n}(F)) be the space of Schwartz functions on Mn(F){\mathrm{M}}_{n}(F). When FF is Archimedean, it is the space of usual Schwartz functions on the affine space Mn(F){\mathrm{M}}_{n}(F), and when FF is pp-adic, it consists of all locally constant, compactly supported functions on Mn(F){\mathrm{M}}_{n}(F). Let ||F|\cdot|_{F} be the normalized absolute value on the local field FF, which is the modular function of the multiplication of F×F^{\times} on FF with respect to the self-dual additive Haar measure d+x\,\mathrm{d}^{+}x on FF. As a reformulation of the local Godement-Jacquet theory in [JL23, Section 2.2], the (standard) Schwartz space on Gn(F){\mathrm{G}}_{n}(F) is defined to be

(2.1) 𝒮std(Gn(F)):={ξ𝒞(Gn(F)):|detg|Fn2ξ(g)𝒮(Mn(F))},\displaystyle{\mathcal{S}}_{\rm std}({\mathrm{G}}_{n}(F)):=\{\xi\in{\mathcal{C}}^{\infty}({\mathrm{G}}_{n}(F))\colon|\det g|_{F}^{-\frac{n}{2}}\cdot\xi(g)\in{\mathcal{S}}({\mathrm{M}}_{n}(F))\},

where 𝒞(Gn(F)){\mathcal{C}}^{\infty}({\mathrm{G}}_{n}(F)) denotes the space of all smooth functions on Gn(F){\mathrm{G}}_{n}(F). By [JL23, Prposition 2.5], the Schwartz space 𝒮std(Gn(F)){\mathcal{S}}_{\rm std}({\mathrm{G}}_{n}(F)) is a subspace of L2(Gn(F),dg)L^{2}({\mathrm{G}}_{n}(F),\,\mathrm{d}g), which is the space of square-integrable functions on Gn(F){\mathrm{G}}_{n}(F).

Consider the determinant map det=detF:Mn(F)=GLn(F)F.\det={\det}_{F}\colon{\mathrm{M}}_{n}(F)={\mathrm{GL}}_{n}(F)\to F. When restricted to F×F^{\times}, we obtain that

(2.2) det=detF:Gn(F)=GLn(F)F×\displaystyle\det={\det}_{F}\colon{\mathrm{G}}_{n}(F)={\mathrm{GL}}_{n}(F)\to F^{\times}

and the fibers of the determinant map det\det are of the form:

(2.3) Gn(F)x:={gGn(F):detg=x}.\displaystyle{\mathrm{G}}_{n}(F)_{x}:=\{g\in{\mathrm{G}}_{n}(F)\colon\det g=x\}.

When x=1x=1, the fiber is the kernel of the map, i.e. ker(det)=SLn(F)\ker(\det)={\mathrm{SL}}_{n}(F). In general, each fiber Gn(F)x{\mathrm{G}}_{n}(F)_{x} is an SLn(F){\mathrm{SL}}_{n}(F)-torsor. Let d+g\,\mathrm{d}^{+}g be the self-dual Haar measure on Mn(F){\mathrm{M}}_{n}(F) with respect to the standard Fourier transform defined by (2.7) below. On Gn(F){\mathrm{G}}_{n}(F), we fix the Haar measure dg=|detg|Fnd+g\,\mathrm{d}g=|\det g|_{F}^{-n}\cdot\,\mathrm{d}^{+}g. Let d1g\,\mathrm{d}_{1}g be the induced Haar measure dg\,\mathrm{d}g from Gn(F){\mathrm{G}}_{n}(F) to SLn(F){\mathrm{SL}}_{n}(F). It follows that the Haar measure d1g\,\mathrm{d}_{1}g induces an SLn(F){\mathrm{SL}}_{n}(F)-invariant measure dxg\,\mathrm{d}_{x}g on each fiber Gn(F)x{\mathrm{G}}_{n}(F)_{x}.

Let ΠF(Gn)\Pi_{F}({\mathrm{G}}_{n}) be the set of equivalence classes of irreducible smooth representations of Gn(F){\mathrm{G}}_{n}(F) when FF is non-Archimedean; and of irreducible Casselman-Wallach representations of Gn(F){\mathrm{G}}_{n}(F) when FF is Archimedean. For πΠF(Gn)\pi\in\Pi_{F}({\mathrm{G}}_{n}), we denote by 𝒞(π){\mathcal{C}}(\pi) the space of all matrix coefficients of π\pi. Write ξ=|detg|Fn2f(g)𝒮std(Gn(F))\xi=|\det g|_{F}^{\frac{n}{2}}\cdot f(g)\in{\mathcal{S}}_{\rm std}({\mathrm{G}}_{n}(F)) with some f𝒮(Mn(F))f\in{\mathcal{S}}({\mathrm{M}}_{n}(F)) as in (2.1). For φπ𝒞(π)\varphi_{\pi}\in{\mathcal{C}}(\pi), as in [JL23, Section 3.1], we define

(2.4) ϕξ,φπ(x):=Gn(F)xξ(g)φπ(g)dxg=|x|Fn2Gn(F)xf(g)φπ(g)dxg.\displaystyle\phi_{\xi,\varphi_{\pi}}(x):=\int_{{\mathrm{G}}_{n}(F)_{x}}\xi(g)\varphi_{\pi}(g)\,\mathrm{d}_{x}g=|x|_{F}^{\frac{n}{2}}\int_{{\mathrm{G}}_{n}(F)_{x}}f(g)\varphi_{\pi}(g)\,\mathrm{d}_{x}g.

By [JL23, Proposition 3.2], the function ϕξ,φπ(x)\phi_{\xi,\varphi_{\pi}}(x) is absolutely convergent for all xF×x\in F^{\times} and is smooth over F×F^{\times}. As in [JL23, Definition 3.3], for any πΠF(Gn)\pi\in\Pi_{F}({\mathrm{G}}_{n}), the space of π\pi-Schwartz functions is defined as

(2.5) 𝒮π(F×)=Span{ϕ=ϕξ,φπ𝒞(F×):ξ𝒮std(Gn(F)),φπ𝒞(π)}.\displaystyle{\mathcal{S}}_{\pi}(F^{\times})={\mathrm{Span}}\{\phi=\phi_{\xi,\varphi_{\pi}}\in{\mathcal{C}}^{\infty}(F^{\times})\colon\xi\in{\mathcal{S}}_{\rm std}({\mathrm{G}}_{n}(F)),\varphi_{\pi}\in{\mathcal{C}}(\pi)\}.

By [JL23, Corollary 3.8], we have

(2.6) 𝒞c(F×)𝒮π(F×)𝒞(F×).\displaystyle{\mathcal{C}}_{c}^{\infty}(F^{\times})\subset{\mathcal{S}}_{\pi}(F^{\times})\subset{\mathcal{C}}^{\infty}(F^{\times}).

2.2. π\pi-Fourier transform

Let ψ=ψF\psi=\psi_{F} be a fixed non-trivial additive character of FF. The (standard) Fourier transform ψ{\mathcal{F}}_{\psi} on 𝒮(Mn(F)){\mathcal{S}}({\mathrm{M}}_{n}(F)) is defined as follows,

(2.7) ψ(f)(x)=Mn(F)ψ(tr(xy))f(y)d+y.{\mathcal{F}}_{\psi}(f)(x)=\int_{{\mathrm{M}}_{n}(F)}\psi({\mathrm{tr}}(xy))f(y)\,\mathrm{d}^{+}y.

It is well-known that the Fourier transform ψ{\mathcal{F}}_{\psi} extends to a unitary operator on the space L2(Mn(F),d+x)L^{2}({\mathrm{M}}_{n}(F),\,\mathrm{d}^{+}x) and satisfies the following identity:

(2.8) ψψ1=Id.{\mathcal{F}}_{\psi}\circ{\mathcal{F}}_{\psi^{-1}}={\mathrm{Id}}.

Following the reformulation of the local Godement-Jacquet theory in [JL23, Section 2.3], the Fourier transform ψ{\mathcal{F}}_{\psi} on 𝒮(Mn(F)){\mathcal{S}}({\mathrm{M}}_{n}(F)) yields a (nonlinear) Fourier transform GJ{\mathcal{F}}_{{\mathrm{GJ}}} on 𝒮std(Gn(F)){\mathcal{S}}_{\rm std}({\mathrm{G}}_{n}(F)), which is a convolution operator with the distribution kernel:

(2.9) ΦGJ(g):=ψ(trg)|detg|Fn2.\displaystyle\Phi_{{\mathrm{GJ}}}(g):=\psi({\mathrm{tr}}g)\cdot|\det g|_{F}^{\frac{n}{2}}.

More precisely, the Fourier transform GJ{\mathcal{F}}_{{\mathrm{GJ}}} is defined to be

(2.10) GJ(ξ)(g):=(ΦGJξ)(g)\displaystyle{\mathcal{F}}_{{\mathrm{GJ}}}(\xi)(g):=\left(\Phi_{{\mathrm{GJ}}}*\xi^{\vee}\right)(g)

for any ξ𝒮std(Gn(F))\xi\in{\mathcal{S}}_{\rm std}({\mathrm{G}}_{n}(F)), where ξ(g):=ξ(g1)\xi^{\vee}(g):=\xi(g^{-1}). From [JL23, Proposition 2.6], a relation between the (nonlinear) Fourier operator GJ{\mathcal{F}}_{{\mathrm{GJ}}} and the (classical or linear) Fourier transform ψ{\mathcal{F}}_{\psi} is given by

(2.11) GJ(ξ)(g)=(ΦGJξ)(g)=|detg|Fn2ψ(|detg|Fn2ξ)(g).\displaystyle{\mathcal{F}}_{{\mathrm{GJ}}}(\xi)(g)=\left(\Phi_{{\mathrm{GJ}}}*\xi^{\vee}\right)(g)=|\det g|_{F}^{\frac{n}{2}}\cdot{\mathcal{F}}_{\psi}(|\det g|_{F}^{-\frac{n}{2}}\xi)(g).

From the proof of [JL23, Proposition 2.6], it is easy to obtain that

(2.12) (ΦGJξ)(g)=|detg|Fn2(ψ(tr())(|det()|Fn2ξ))(g)\displaystyle\left(\Phi_{{\mathrm{GJ}}}*\xi^{\vee}\right)(g)=|\det g|_{F}^{\frac{n}{2}}\left(\psi({\mathrm{tr}}(\cdot))*(|\det(\cdot)|_{F}^{\frac{n}{2}}\xi)^{\vee}\right)(g)

for any ξ𝒮std(Gn(F))\xi\in{\mathcal{S}}_{\rm std}({\mathrm{G}}_{n}(F)).

As in [JL23, Section 3.2], the π\pi-Fourier transform π,ψ{\mathcal{F}}_{\pi,\psi} is defined through the following diagram:

(2.17)

More precisely, for ϕ=ϕξ,φπ𝒮π(F×)\phi=\phi_{\xi,\varphi_{\pi}}\in{\mathcal{S}}_{\pi}(F^{\times}) with a ξ𝒮std(Gn(F))\xi\in{\mathcal{S}}_{\rm std}({\mathrm{G}}_{n}(F)) and a φπ𝒞(π)\varphi_{\pi}\in{\mathcal{C}}(\pi), the π\pi-Fourier transform π,ψ{\mathcal{F}}_{\pi,\psi} is defined by

(2.18) π,ψ(ϕ)=π,ψ(ϕξ,φπ):=ϕGJ(ξ),φπ,\displaystyle{\mathcal{F}}_{\pi,\psi}(\phi)={\mathcal{F}}_{\pi,\psi}(\phi_{\xi,\varphi_{\pi}}):=\phi_{{\mathcal{F}}_{{\mathrm{GJ}}}(\xi),\varphi_{\pi}^{\vee}},

where φπ(g)=φπ(g1)𝒞(π~)\varphi_{\pi}^{\vee}(g)=\varphi_{\pi}(g^{-1})\in{\mathcal{C}}(\widetilde{\pi}). It was verified in [JL23, Proposition 3.9] that the descending π\pi-Fourier transform π,ψ{\mathcal{F}}_{\pi,\psi} is well defined. From [JL22, Theorem 5.1], the π\pi-Fourier transform π,ψ{\mathcal{F}}_{\pi,\psi} can also be represented as a convolution operator by some kernel function kπ,ψk_{\pi,\psi}, which is explicitly given as follows.

We fix a φπ~𝒞(π~)\varphi_{\widetilde{\pi}}\in{\mathcal{C}}(\widetilde{\pi}) with φπ~(In)=1\varphi_{\widetilde{\pi}}({\mathrm{I}}_{n})=1. We also choose a sequence of test functions {𝔠}=1𝒞c(Gn(F))\{{\mathfrak{c}}_{\ell}\}_{\ell=1}^{\infty}\subset{\mathcal{C}}^{\infty}_{c}({\mathrm{G}}_{n}(F)), such that for any h𝒞c(Gn(F))h\in{\mathcal{C}}_{c}^{\infty}({\mathrm{G}}_{n}(F)),

(2.19) limGn(F)𝔠(g)h(g)dg=h(In).\displaystyle\lim_{\ell\to\infty}\int_{{\mathrm{G}}_{n}(F)}{\mathfrak{c}}_{\ell}(g)h(g)\,\mathrm{d}g=h({\mathrm{I}}_{n}).

In other words, the sequence {𝔠}=1\{{\mathfrak{c}}_{\ell}\}_{\ell=1}^{\infty} tends to the delta mass supported at the identity In{\mathrm{I}}_{n} as \ell\rightarrow\infty. The π\pi-kernel function kπ,ψ(x)k_{\pi,\psi}(x) is defined as

(2.20) kπ,ψ(x):=detg=xregΦGJ(g)φπ~(g)dxg=|x|Fn2detg=xregψ(tr(g))φπ~(g)dxgk_{\pi,\psi}(x):=\int^{\mathrm{reg}}_{\det g=x}\Phi_{{\mathrm{GJ}}}(g)\varphi_{\widetilde{\pi}}(g)\,\mathrm{d}_{x}g=|x|^{\frac{n}{2}}_{F}\int^{\mathrm{reg}}_{\det g=x}\psi({\mathrm{tr}}(g))\varphi_{\widetilde{\pi}}(g)\,\mathrm{d}_{x}g

where ΦGJ\Phi_{{\mathrm{GJ}}} is the kernel function as defined in (2.9) and the integral is regularized as follows:

(2.21) detg=xregΦGJ(g)φπ~(g)dxg:=limdetg=x(ΦGJ𝔠)(g)φπ~(g)dxg.\displaystyle\int^{\mathrm{reg}}_{\det g=x}\Phi_{{\mathrm{GJ}}}(g)\varphi_{\widetilde{\pi}}(g)\,\mathrm{d}_{x}g:=\lim_{\ell\to\infty}\int_{\det g=x}\left(\Phi_{{\mathrm{GJ}}}*{\mathfrak{c}}_{\ell}^{\vee}\right)(g)\varphi_{\widetilde{\pi}}(g)\,\mathrm{d}_{x}g.

It is shown in [JL22, Proposition 3.5, Corollary 3.7, Corollary 4.5 and Theorem 4.6] that kπ,ψk_{\pi,\psi} is a smooth function on F×F^{\times} and is independent of the choice of the matrix coefficient φπ~\varphi_{\widetilde{\pi}} and the chosen sequence {𝔠}=1\{{\mathfrak{c}}_{\ell}\}_{\ell=1}^{\infty} that tends to the delta mass supported at In{\mathrm{I}}_{n}. By [JL22, Theorem 5.1], we have that for any ϕ𝒞c(F×)\phi\in{\mathcal{C}}_{c}^{\infty}(F^{\times})

(2.22) π,ψ(ϕ)(x)=(kπ,ψϕ)(x).\displaystyle{\mathcal{F}}_{\pi,\psi}(\phi)(x)=(k_{\pi,\psi}*\phi^{\vee})(x).

Following [Ngo20], one may call the π\pi-Fourier transform π,ψ{\mathcal{F}}_{\pi,\psi} a generalized Hankel transform or the π\pi-Hankel transform.

2.3. π\pi-Poisson summation formula on GL1{\mathrm{GL}}_{1}

Recall that |k||k| is the set of all local places of the number field kk. Let |k||k|_{\infty} be the subset of |k||k| consisting of all Archimedean local places of kk. We may write |k|=|k||k|f|k|=|k|_{\infty}\cup|k|_{f}, where |k|f|k|_{f} is the set of non-Archimedean local places of kk. Let Π𝔸(Gn)\Pi_{\mathbb{A}}({\mathrm{G}}_{n}) be the set of equivalence classes of irreducible admissible representations of Gn(𝔸){\mathrm{G}}_{n}({\mathbb{A}}). We write π=ν|k|πν\pi=\otimes_{\nu\in|k|}\pi_{\nu} and assume that πνΠkν(Gn)\pi_{\nu}\in\Pi_{k_{\nu}}({\mathrm{G}}_{n}) and at almost all finite local places ν\nu the local representations πν\pi_{\nu} are unramified. This means that when ν|k|f\nu\in|k|_{f}, πν\pi_{\nu} is an irreducible admissible representation of Gn(kν){\mathrm{G}}_{n}(k_{\nu}), and when ν|k|\nu\in|k|_{\infty}, πν\pi_{\nu} is of Casselman-Wallach type as a representation of Gn(kν){\mathrm{G}}_{n}(k_{\nu}). Let 𝒜(Gn)Π𝔸(Gn){\mathcal{A}}({\mathrm{G}}_{n})\subset\Pi_{\mathbb{A}}({\mathrm{G}}_{n}) be the subset consisting of equivalence classes of irreducible admissible automorphic representations of Gn(𝔸){\mathrm{G}}_{n}({\mathbb{A}}), and 𝒜cusp(Gn){\mathcal{A}}_{\mathrm{cusp}}({\mathrm{G}}_{n}) be the subset of cuspidal members in 𝒜(Gn){\mathcal{A}}({\mathrm{G}}_{n}). We refer to [Art13, Chepter 1] or [JL23] for the notation and definition of automorphic representations.

Take any π=ν|k|πνΠ𝔸(Gn)\displaystyle{\pi=\otimes_{\nu\in|k|}}\pi_{\nu}\in\Pi_{{\mathbb{A}}}({\mathrm{G}}_{n}). For each local place ν|k|\nu\in|k|, the πν\pi_{\nu}-Schwartz space 𝒮πν(kν×){\mathcal{S}}_{\pi_{\nu}}(k_{\nu}^{\times}) is defined as in (2.5). Recall from [JL23, Theorem 3.4] that the basic function 𝕃πν𝒮πν(kν×){\mathbb{L}}_{\pi_{\nu}}\in{\mathcal{S}}_{\pi_{\nu}}(k_{\nu}^{\times}) is defined when the local component πν\pi_{\nu} of π\pi is unramified. Then the π\pi-Schwartz space on 𝔸×{\mathbb{A}}^{\times} is defined to be

(2.23) 𝒮π(𝔸×):=ν|k|𝒮πν(kν×),\displaystyle{\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}):=\otimes_{\nu\in|k|}{\mathcal{S}}_{\pi_{\nu}}(k_{\nu}^{\times}),

which is the restricted tensor product of the local πν\pi_{\nu}-Schwartz space 𝒮πν(kν×){\mathcal{S}}_{\pi_{\nu}}(k_{\nu}^{\times}) with respect to the family of the basic functions 𝕃πν{\mathbb{L}}_{\pi_{\nu}} for all the local places ν\nu at which πν\pi_{\nu} are unramified. The factorizable vectors ϕ=νϕν\phi=\otimes_{\nu}\phi_{\nu} in 𝒮π(𝔸×){\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}) can be written as

(2.24) ϕ(x)=ν|k|ϕν(xν),x=(xν)ν.\displaystyle\phi(x)=\prod_{\nu\in|k|}\phi_{\nu}(x_{\nu}),\;x=(x_{\nu})_{\nu}.

Here at almost all finite local places ν\nu, ϕν(xν)=𝕃πν(xν)\phi_{\nu}(x_{\nu})={\mathbb{L}}_{\pi_{\nu}}(x_{\nu}). According to the normalization ([JL23, Theorem 3.4]), we have that 𝕃πν(xν)=1{\mathbb{L}}_{\pi_{\nu}}(x_{\nu})=1 when xν𝔬ν×x_{\nu}\in{\mathfrak{o}}_{\nu}^{\times}, the unit group of the ring 𝔬ν{\mathfrak{o}}_{\nu} of integers at ν\nu. Hence for any given x𝔸×x\in{\mathbb{A}}^{\times}, the product in (2.24) is a finite product.

For any factorizable vectors ϕ=νϕν\phi=\otimes_{\nu}\phi_{\nu} in 𝒮π(𝔸×){\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}), we define the π\pi-Fourier transform (or operator):

(2.25) π,ψ(ϕ):=ν|k|πν,ψν(ϕν).\displaystyle{\mathcal{F}}_{\pi,\psi}(\phi):=\otimes_{\nu\in|k|}{\mathcal{F}}_{\pi_{\nu},\psi_{\nu}}(\phi_{\nu}).

Here at each ν|k|\nu\in|k|, πν,ψν{\mathcal{F}}_{\pi_{\nu},\psi_{\nu}} is the local πν\pi_{\nu}-Fourier transform as defined in (2.17) and (2.18), which takes the πν\pi_{\nu}-Schwartz space 𝒮πν(kν×){\mathcal{S}}_{\pi_{\nu}}(k_{\nu}^{\times}) to the π~ν\widetilde{\pi}_{\nu}-Schwartz space 𝒮π~ν(kν×){\mathcal{S}}_{\widetilde{\pi}_{\nu}}(k_{\nu}^{\times}), and has the property that πν,ψ(𝕃πν)=𝕃π~ν{\mathcal{F}}_{\pi_{\nu},\psi}({\mathbb{L}}_{\pi_{\nu}})={\mathbb{L}}_{\widetilde{\pi}_{\nu}}, when the data are unramified at ν\nu (see [JL23, Theorem 3.10]). Hence the Fourier transform π,ψ{\mathcal{F}}_{\pi,\psi} as defined in (2.25) maps the π\pi-Schwartz space 𝒮π(𝔸×){\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}) to the π~\widetilde{\pi}-Schwartz space 𝒮π~(𝔸×){\mathcal{S}}_{\widetilde{\pi}}({\mathbb{A}}^{\times}), where π~Π𝔸(Gn)\widetilde{\pi}\in\Pi_{\mathbb{A}}({\mathrm{G}}_{n}) is the contragredient of π\pi. The π\pi-Poisson summation formula ([JL23, Theorem 4.7]) can be stated as below.

Theorem 2.1 (π\pi-Poisson summation formula).

For any π𝒜cusp(Gn)\pi\in{\mathcal{A}}_{\mathrm{cusp}}({\mathrm{G}}_{n}), the π\pi-theta function Θπ(x,ϕ):=αk×ϕ(αx)\Theta_{\pi}(x,\phi):=\sum_{\alpha\in k^{\times}}\phi(\alpha x) converges absolutely and locally uniformly for any x𝔸×x\in{\mathbb{A}}^{\times} and any ϕ𝒮π(𝔸×)\phi\in{\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}). Let π~𝒜cusp(Gn)\widetilde{\pi}\in{\mathcal{A}}_{\mathrm{cusp}}({\mathrm{G}}_{n}) be the contragredient of π\pi. Then the following identity

Θπ(x,ϕ)=Θπ~(x1,π,ψ(ϕ)),\Theta_{\pi}(x,\phi)=\Theta_{\widetilde{\pi}}(x^{-1},{\mathcal{F}}_{\pi,\psi}(\phi)),

holds as functions in x𝔸×x\in{\mathbb{A}}^{\times}, where π,ψ{\mathcal{F}}_{\pi,\psi} is the π\pi-Fourier transform as defined in (2.25).

3. Local Harmonic Analysis

In this section, we take F=kνF=k_{\nu} to be a local field of characteristic zero and fix a non-trivial additive character ψ=ψF\psi=\psi_{F} of FF. Since the representations πΠF(Gn)\pi\in\Pi_{F}({\mathrm{G}}_{n}) considered in this section are the local components of irreducible cuspidal automorphic representations of Gn(𝔸){\mathrm{G}}_{n}({\mathbb{A}}), we may only consider generic πΠF(Gn)\pi\in\Pi_{F}({\mathrm{G}}_{n}) without loss of generality.

Let Bn=TnNnB_{n}=T_{n}N_{n} be the Borel subgroup of Gn{\mathrm{G}}_{n}, which consisting of all upper-triangular matrices of Gn{\mathrm{G}}_{n}, where TnT_{n} is the maximal torus consisting of all diagonal matrices of Gn{\mathrm{G}}_{n}, and NnN_{n} is the unipotent radical of BnB_{n}, which consists of matrices n=(ni,j)n=(n_{i,j}) with ni,j=0n_{i,j}=0 if 1j<in1\leq j<i\leq n, and ni,i=1n_{i,i}=1 for i=1,2,,ni=1,2,\dots,n. Without loss of generality, we may take a generic character as

ψ(n)=ψNn(n)=ψF(n1,2+n2,3++nn1,n).\psi(n)=\psi_{N_{n}}(n)=\psi_{F}(n_{1,2}+n_{2,3}+\cdots+n_{n-1,n}).

Let ψ\ell_{\psi} be a non-zero member in HomNn(F)(π,ψ){\mathrm{Hom}}_{N_{n}(F)}(\pi,\psi), which is one-dimensional if πΠF(Gn)\pi\in\Pi_{F}({\mathrm{G}}_{n}) is generic. For any vVπv\in V_{\pi}, define the Whittaker function by Wv(g):=ψ(π(g)v)W_{v}(g):=\ell_{\psi}(\pi(g)v). Let 𝒲(π,ψ){\mathcal{W}}(\pi,\psi) be the Whittaker model of π\pi, which consisting of Whittaker functions Wv(g)W_{v}(g) with vv runs through the space VπV_{\pi} of π\pi. Let VπV_{\pi}^{\infty} be the subspace of VπV_{\pi} consisting of all smooth vectors of VπV_{\pi}. We define the π\pi-Whittaker-Schwartz space on F×F^{\times} to be

(3.1) 𝒲π,ψ(F×):={ω(x):=|x|1n2Wv((xIn1)):vVπ},\displaystyle{\mathcal{W}}_{\pi,\psi}(F^{\times}):=\{\omega(x):=|x|^{1-\frac{n}{2}}\cdot W_{v}\left(\begin{pmatrix}x&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)\colon v\in V_{\pi}^{\infty}\},

where ||=||F|\cdot|=|\cdot|_{F} is the normalized absolute value on FF.

Proposition 3.1.

For any πΠF(Gn)\pi\in\Pi_{F}({\mathrm{G}}_{n}), which is generic, the π\pi-Schwartz space and the π\pi-Whittaker-Schwartz space coincide with each other: 𝒮π(F×)=𝒲π,ψ(F×){\mathcal{S}}_{\pi}(F^{\times})={\mathcal{W}}_{\pi,\psi}(F^{\times}).

Proof.

We first show that 𝒲π,ψ(F×)𝒮π(F×){\mathcal{W}}_{\pi,\psi}(F^{\times})\subset{\mathcal{S}}_{\pi}(F^{\times}). For any unitary character χ\chi of F×F^{\times} and W𝒲(π,ψ)W\in{\mathcal{W}}(\pi,\psi), the local Rankin-Selberg integral for GLn×GL1{\mathrm{GL}}_{n}\times{\mathrm{GL}}_{1}

Ψ(s,W,χ):=F×W((xIn1))χ(x)|x|sn12d×x=F×ω(x)χ(x)|x|s12d×x,\displaystyle\Psi(s,W,\chi):=\int_{F^{\times}}W\left(\begin{pmatrix}x&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)\chi(x)|x|^{s-\frac{n-1}{2}}\,\mathrm{d}^{\times}x=\int_{F^{\times}}\omega(x)\chi(x)|x|^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x,

where ω(x)𝒲π,ψ(F×)\omega(x)\in{\mathcal{W}}_{\pi,\psi}(F^{\times}) as in Proposition 3.2, is absolutely convergent when Re(s){\mathrm{Re}}(s) is sufficiently positive and the fractional ideal generated by all such integrals is [qs,qs]L(s,π×χ){\mathbb{C}}[q^{-s},q^{s}]L(s,\pi\times\chi) by [JPSS83, Theorem 2.7] for the non-Archimedean case and a holomorphic multiple of L(s,π×χ)L(s,\pi\times\chi), bounded at infinity in vertical strips due to [Jac09, Theorem 2.1] for Archimedean case.

According to [JL23, Theorem 3.4], there is some ϕ𝒮π(F×)\phi\in{\mathcal{S}}_{\pi}(F^{\times}) such that

Ψ(s,W,χ)=𝒵(s,ϕ,χ):=F×ϕ(x)χ(x)|x|s12d×x\displaystyle\Psi(s,W,\chi)={\mathcal{Z}}(s,\phi,\chi):=\int_{F^{\times}}\phi(x)\chi(x)|x|^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x

when Re(s){\mathrm{Re}}(s) is sufficiently positive. In particular, fix a s0s_{0}\in{\mathbb{R}} sufficiently positive such that both functions ϕ()||s012\phi(\cdot)|\cdot|^{s_{0}-\frac{1}{2}} and ω()||s012\omega(\cdot)|\cdot|^{s_{0}-\frac{1}{2}} belong to L1(F×)L^{1}(F^{\times}), the space of L1L^{1}-functions on F×F^{\times}. It follows that

F×(ϕ(x)|x|s012ω(x)|x|s012)χ(x)d×x=0\int_{F^{\times}}\left(\phi(x)|x|^{s_{0}-\frac{1}{2}}-\omega(x)|x|^{s_{0}-\frac{1}{2}}\right)\chi(x)\,\mathrm{d}^{\times}x=0

for all unitary character χ\chi of F×F^{\times}. From the general theory about absolutely continuous measures on local compact abelian groups (See [HR79, Theorem 23.11] for instance), we must have that ϕ(x)|x|s012ω(x)|x|s012=0\phi(x)|x|^{s_{0}-\frac{1}{2}}-\omega(x)|x|^{s_{0}-\frac{1}{2}}=0 for a.e. xF×x\in F^{\times}, which implies ω(x)=ϕ(x)\omega(x)=\phi(x) for all xF×x\in F^{\times} since both functions are smooth. Hence we obtain that 𝒲π,ψ(F×)𝒮π(F×){\mathcal{W}}_{\pi,\psi}(F^{\times})\subset{\mathcal{S}}_{\pi}(F^{\times}).

Again, by [JL23, Theorem 3.4] and the local theory of the Rankin-Selberg convolution of GLn×GL1{\mathrm{GL}}_{n}\times{\mathrm{GL}}_{1} as in [JPSS83, Theorem 2.7] for the non-Archimedean case and in [Jac09, Theorem 2.1] for Archimedean case, we can repeat the above discussion to prove that 𝒮π(F×)𝒲π,ψ(F×){\mathcal{S}}_{\pi}(F^{\times})\subset{\mathcal{W}}_{\pi,\psi}(F^{\times}). Hence we get that 𝒮π(F×)=𝒲π,ψ(F×){\mathcal{S}}_{\pi}(F^{\times})={\mathcal{W}}_{\pi,\psi}(F^{\times}). ∎

From Proposition 3.1, the following assertion is clear, since the π\pi-Schwartz space 𝒮π(F×){\mathcal{S}}_{\pi}(F^{\times}) is independent of the choice of the character ψ\psi.

Corollary 3.2.

The space of π\pi-Whittaker-Schwartz functions 𝒲π,ψ(F×){\mathcal{W}}_{\pi,\psi}(F^{\times}) defined in (3.1) is independent of the choice of the character ψ\psi.

By Corollary 3.2, we may denote by 𝒲π(F×){\mathcal{W}}_{\pi}(F^{\times}) the π\pi-Whittaker-Schwartz space on F×F^{\times} as defined in (3.1). After identifying the π\pi-Schwartz space 𝒮π(F×){\mathcal{S}}_{\pi}(F^{\times}) with the π\pi-Whittaker-Schwartz space 𝒲π(F×){\mathcal{W}}_{\pi}(F^{\times}), we are going to understand the π\pi-Fourier transform π,ψ:𝒮π(F×)𝒮π~(F×){\mathcal{F}}_{\pi,\psi}\colon{\mathcal{S}}_{\pi}(F^{\times})\to{\mathcal{S}}_{\widetilde{\pi}}(F^{\times}) in terms of the structure of Whittaker models.

Proposition 3.3.

For ϕ𝒮π(F×)\phi\in{\mathcal{S}}_{\pi}(F^{\times}), we may write as in (3.1) that

ϕ(x)=ω(x)=W((xIn1))|x|1n2\displaystyle\phi(x)=\omega(x)=W\left(\begin{pmatrix}x&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)|x|^{1-\frac{n}{2}}

for some W𝒲(π,ψ)W\in{\mathcal{W}}(\pi,\psi). Then the π\pi-Fourier transform can be expressed by the following formula:

π,ψ(ϕ)(x)=π,ψ(ω)(x)=|x|1n2Fn2(π(wn,1)W~)(xyIn21)dy,\displaystyle{\mathcal{F}}_{\pi,\psi}(\phi)(x)={\mathcal{F}}_{\pi,\psi}(\omega)(x)=|x|^{1-\frac{n}{2}}\int_{F^{n-2}}\left(\pi(w_{n,1})\widetilde{W}\right)\begin{pmatrix}x&&\\ y&{\mathrm{I}}_{n-2}&\\ &&1\end{pmatrix}\,\mathrm{d}y,

where W~(g):=W(w0tg1)\widetilde{W}(g):=W(w_{0}\,^{t}g^{-1}) for any gGLn(F)g\in{\mathrm{GL}}_{n}(F) is a Whiitaker function in 𝒲(π~,ψ1){\mathcal{W}}(\widetilde{\pi},\psi^{-1}) and

(3.2) wn,1=(1wn1).\displaystyle w_{n,1}=\begin{pmatrix}1&\\ &w_{n-1}\end{pmatrix}.

Here we denote by wmw_{m} the longest Weyl element of Gm=GLm{\mathrm{G}}_{m}={\mathrm{GL}}_{m}, which is defined inductively by

(3.3) wm=(1wm1),withw2=(11).\displaystyle w_{m}=\begin{pmatrix}&1\\ w_{m-1}&\end{pmatrix},\quad{\rm with}\quad w_{2}=\begin{pmatrix}&1\\ 1&\end{pmatrix}.
Proof.

From the functional equation for the local zeta integrals 𝒵(s,ϕ,χ){\mathcal{Z}}(s,\phi,\chi) as proved in [JL23, Theorem 3.10], we have that

Ψ(s,W,χ)=𝒵(s,ϕ,χ)=𝒵(1s,π,ψ(ϕ),χ1)γ(s,π×χ,ψ)1.\displaystyle\Psi(s,W,\chi)={\mathcal{Z}}(s,\phi,\chi)={\mathcal{Z}}(1-s,{\mathcal{F}}_{\pi,\psi}(\phi),\chi^{-1})\gamma(s,\pi\times\chi,\psi)^{-1}.

On the other hand, from the functional equation for the local zeta integrals Ψ(s,W,χ)\Psi(s,W,\chi) 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

Ψ(s,W,χ)=γ(s,π×χ,ψ)1F×Fn2(π(wn,1)W~)(xyIn21)dyχ1(x)|x|3n2sd×x.\displaystyle\Psi(s,W,\chi)=\gamma(s,\pi\times\chi,\psi)^{-1}\int_{F^{\times}}\int_{F^{n-2}}\left(\pi(w_{n,1})\widetilde{W}\right)\begin{pmatrix}x&&\\ y&{\mathrm{I}}_{n-2}&\\ &&1\end{pmatrix}\,\mathrm{d}y\;\chi^{-1}(x)|x|^{\frac{3-n}{2}-s}\,\mathrm{d}^{\times}x.

From the absolute convergence of the local zeta integrals 𝒵(s,ϕ,χ){\mathcal{Z}}(s,\phi,\chi) and Ψ(s,W,χ)\Psi(s,W,\chi), we may choose and fix a s0s_{0}\in{\mathbb{C}} with Re(s0){\mathrm{Re}}(s_{0}) sufficiently negative, such that both functions

π,ψ(ϕ)()||12s0and||3n2s0Fn2(π(wn,1)W~)(yIn21)dy\displaystyle{\mathcal{F}}_{\pi,\psi}(\phi)(\cdot)|\cdot|^{\frac{1}{2}-s_{0}}\quad{\rm and}\quad|\cdot|^{\frac{3-n}{2}-s_{0}}\int_{F^{n-2}}\left(\pi(w_{n,1})\widetilde{W}\right)\begin{pmatrix}\cdot&&\\ y&{\mathrm{I}}_{n-2}&\\ &&1\end{pmatrix}\,\mathrm{d}y

belong to L1(F×)L^{1}(F^{\times}). It follows that

F×(π,ψ(ϕ)(x)|x|12s0Fn2(π(wn,1)W~)(xyIn21)dy|x|3n2s0)χ1(x)d×x=0\displaystyle\int_{F^{\times}}\left({\mathcal{F}}_{\pi,\psi}(\phi)(x)|x|^{\frac{1}{2}-s_{0}}-\int_{F^{n-2}}\left(\pi(w_{n,1})\widetilde{W}\right)\begin{pmatrix}x&&\\ y&{\mathrm{I}}_{n-2}&\\ &&1\end{pmatrix}\,\mathrm{d}y|x|^{\frac{3-n}{2}-s_{0}}\right)\chi^{-1}(x)\,\mathrm{d}^{\times}x=0

for any unitary character χ\chi. Now we use the same argument as in the proof of Proposition 3.1 to deduce that

π,ψ(ϕ)(x)=|x|1n2Fn2(π(wn,1)W~)(xyIn21)dy{\mathcal{F}}_{\pi,\psi}(\phi)(x)=|x|^{1-\frac{n}{2}}\int_{F^{n-2}}\left(\pi(w_{n,1})\widetilde{W}\right)\begin{pmatrix}x&&\\ y&{\mathrm{I}}_{n-2}&\\ &&1\end{pmatrix}\,\mathrm{d}y

for any xF×x\in F^{\times}, as they are smooth in xx. ∎

In particular, in the case n=2n=2, we have a much simpler formula.

Corollary 3.4.

When n=2n=2, the action of the longest Weyl group element w2w_{2} of G2{\mathrm{G}}_{2} on the Kirillov model of π\pi is given by the (non-linear) Fourier transform π,ψ{\mathcal{F}}_{\pi,\psi}:

π,ψ(ϕ)=π(w2)(ϕ)ωπ1,forϕ𝒮π(F×),\displaystyle{\mathcal{F}}_{\pi,\psi}(\phi)=\pi(w_{2})(\phi)\cdot\omega_{\pi}^{-1},\quad{\rm for}\quad\phi\in{\mathcal{S}}_{\pi}(F^{\times}),

where the π\pi-Schwartz space 𝒮π(F×){\mathcal{S}}_{\pi}(F^{\times}) and the π\pi-Whittaker-Schwartz space 𝒲π(F×){\mathcal{W}}_{\pi}(F^{\times}) can be identified with the Kirillov model of π\pi by Proposition 3.1 and ωπ\omega_{\pi} is the central character of π\pi.

Proof.

According to Proposition 3.3, let ϕ(x)=W((x1))\phi(x)=W\left(\begin{pmatrix}x&\\ &1\end{pmatrix}\right), we have

π,ψ(ϕ)(x)=W(w2(x11))=W((x1x1)(x1)w2)=ωπ(x1)(π(w0)ϕ)(x).\displaystyle{\mathcal{F}}_{\pi,\psi}(\phi)(x)=W\left(w_{2}\begin{pmatrix}x^{-1}&\\ &1\end{pmatrix}\right)=W\left(\begin{pmatrix}x^{-1}&\\ &x^{-1}\end{pmatrix}\begin{pmatrix}x&\\ &1\end{pmatrix}w_{2}\right)=\omega_{\pi}(x^{-1})(\pi(w_{0})\phi)(x).

According to [IT13, Lemma 5.2], for any w(x)𝒞c(F×)w(x)\in{\mathcal{C}}_{c}^{\infty}(F^{\times}), there is a unique smooth function w~(x)\widetilde{w}(x) on F×F^{\times} 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 ss\in{\mathbb{C}}. The map w(x)w~(x)w(x)\mapsto\widetilde{w}(x) 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 GLn×GL1{\mathrm{GL}}_{n}\times{\mathrm{GL}}_{1} from [JPSS83] and [Jac09]. Over the Archimedean local fields, the map w(x)w~(x)w(x)\mapsto\widetilde{w}(x) 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 w(x)w~(x)w(x)\mapsto\widetilde{w}(x) is given by the π\pi-Fourier transform up to certain normalization.

Proposition 3.5.

The dual function w~(x)\widetilde{w}(x) of w(x)𝒞c(F×)w(x)\in{\mathcal{C}}_{c}^{\infty}(F^{\times}) as defined by (1.2) can be expressed in terms of π\pi-Fourier transforms:

π,ψ(w()||1n2)=w~()||1n2.{\mathcal{F}}_{\pi,\psi}(w(\cdot)|\cdot|^{1-\frac{n}{2}})=\widetilde{w}(\cdot)|\cdot|^{1-\frac{n}{2}}.
Proof.

Since w(x)𝒞c(F×)w(x)\in{\mathcal{C}}_{c}^{\infty}(F^{\times}), we have w(x)|x|1n2𝒞c(F×)w(x)|x|^{1-\frac{n}{2}}\in{\mathcal{C}}_{c}^{\infty}(F^{\times}) 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

γ(1s,π×χ,ψ)F×w(y)χ(y)|y|1sn12d×y=γ(1s,π×χ,ψ)𝒵(1s,w()||1n2,χ).\displaystyle\gamma(1-s,\pi\times\chi,\psi)\int_{F^{\times}}w(y)\chi(y)|y|^{1-s-\frac{n-1}{2}}\,\mathrm{d}^{\times}y=\gamma(1-s,\pi\times\chi,\psi){\mathcal{Z}}(1-s,w(\cdot)|\cdot|^{1-\frac{n}{2}},\chi).

By the local functional equation in [JL23, Theorem 3.10], we have

γ(1s,π×χ,ψ)𝒵(1s,w()||1n2,χ)=𝒵(s,π,ψ(w()||1n2),χ1)\gamma(1-s,\pi\times\chi,\psi){\mathcal{Z}}(1-s,w(\cdot)|\cdot|^{1-\frac{n}{2}},\chi)={\mathcal{Z}}(s,{\mathcal{F}}_{\pi,\psi}(w(\cdot)|\cdot|^{1-\frac{n}{2}}),\chi^{-1})

It follows that the left-hand side of (1.2) can be written as

F×w~(y)χ1(y)|y|sn12d×y=F×π,ψ(w()||1n2)(y)χ1(y)|y|s12d×y\int_{F^{\times}}\widetilde{w}(y)\chi^{-1}(y)|y|^{s-\frac{n-1}{2}}\,\mathrm{d}^{\times}y=\int_{F^{\times}}{\mathcal{F}}_{\pi,\psi}(w(\cdot)|\cdot|^{1-\frac{n}{2}})(y)\chi^{-1}(y)|y|^{s-\frac{1}{2}}\,\mathrm{d}^{\times}y

as meromorphic functions in ss\in{\mathbb{C}}. Since the integrals on both sides of the above equation converge absolutely for Re(s){\mathrm{Re}}(s) sufficiently negative, we choose one of such s0s_{0}\in{\mathbb{C}} and fix it such that the two smooth functions w~(y)|y|s0n12\widetilde{w}(y)|y|^{s_{0}-\frac{n-1}{2}} and π,ψ(w()||1n2)(y)|y|s012{\mathcal{F}}_{\pi,\psi}(w(\cdot)|\cdot|^{1-\frac{n}{2}})(y)|y|^{s_{0}-\frac{1}{2}} belong to L1(F×)L^{1}(F^{\times}). Again, by the general theory as in [HR79, Theorem 23.11], we obtain that

w~(y)|y|s0n12=π,ψ(w()||1n2)(y)|y|s012,\widetilde{w}(y)|y|^{s_{0}-\frac{n-1}{2}}={\mathcal{F}}_{\pi,\psi}(w(\cdot)|\cdot|^{1-\frac{n}{2}})(y)|y|^{s_{0}-\frac{1}{2}},

which implies that π,ψ(w()||1n2)(x)=w~(x)|x|1n2{\mathcal{F}}_{\pi,\psi}(w(\cdot)|\cdot|^{1-\frac{n}{2}})(x)=\widetilde{w}(x)|x|^{1-\frac{n}{2}}, as functions on F×F^{\times}, is smooth, of rapid decay at infinity, and with at most polynomial growth at zero. ∎

Combining Proposition 3.5 with the formula in (2.22), we obtain a formula for w~(x)\widetilde{w}(x) for any w𝒞c(F×)w\in{\mathcal{C}}^{\infty}_{c}(F^{\times}).

Corollary 3.6.

For any πΠF(Gn)\pi\in\Pi_{F}({\mathrm{G}}_{n}), the dual function w~(x)\widetilde{w}(x) associated with any w𝒞c(F×)w\in{\mathcal{C}}^{\infty}_{c}(F^{\times}) is given by the following formula:

w~(x)=|x|Fn21(kπ,ψ()(w()||Fn21))(x),\widetilde{w}(x)=|x|_{F}^{\frac{n}{2}-1}\left(k_{\pi,\psi}(\cdot)*(w^{\vee}(\cdot)|\cdot|_{F}^{\frac{n}{2}-1})\right)(x),

where kπ,ψ(x)k_{\pi,\psi}(x) is the π\pi-kernel function associated with π\pi as in (2.20) and w(x)=w(x1)w^{\vee}(x)=w(x^{-1}).

4. π\pi-Bessel functions

The π\pi-Fourier transform π,ψ{\mathcal{F}}_{\pi,\psi} can be expressed as a convolution operator with the π\pi-kernel function kπ,ψk_{\pi,\psi} as in (2.20) using the structures of the π\pi-Schwartz space 𝒮π(F×){\mathcal{S}}_{\pi}(F^{\times}) and the π~\widetilde{\pi}-Schwartz space 𝒮π~(F×){\mathcal{S}}_{\widetilde{\pi}}(F^{\times}). When consider the π\pi-Fourier transform π,ψ{\mathcal{F}}_{\pi,\psi} as a transformation from the π\pi-Whittaker-Schwartz space 𝒲π(F×){\mathcal{W}}_{\pi}(F^{\times}) to π~\widetilde{\pi}-Whittaker-Schwartz space 𝒲π~(F×){\mathcal{W}}_{\widetilde{\pi}}(F^{\times}), we are able to show that the π\pi-Fourier transform π,ψ{\mathcal{F}}_{\pi,\psi} 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. π\pi-Bessel functions: pp-adic case

Assume that FF 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 π\pi-Fourier transform.

Let Φ={αi,j=eiej1i<jn}\Phi=\{\alpha_{i,j}=e_{i}-e_{j}\mid 1\leq i<j\leq n\} be the roots of Gn{\mathrm{G}}_{n} with respect to the FF-split maximal torus TnT_{n}, Φ+={αi,ji<j}\Phi^{+}=\{\alpha_{i,j}\mid i<j\} be the set of positive roots with respect to BnB_{n} and Φ={αi,ji>j}\Phi^{-}=\{\alpha_{i,j}\mid i>j\} be the corresponding set of negative roots. Let Δ={αi,i+11in1}\Delta=\{\alpha_{i,i+1}\mid 1\leq i\leq n-1\} be the set of simple roots. Let 𝕎{\mathbb{W}} be the Weyl group of Gn{\mathrm{G}}_{n}. For every w𝕎w\in{\mathbb{W}}, denote

S(w)={αΔw(α)<0}andS(w)=S(wwn),\displaystyle S(w)=\{\alpha\in\Delta\mid w(\alpha)<0\}\quad\mathrm{and}\quad S^{\circ}(w)=S(ww_{n}),

where wnw_{n} is the longest Weyl element of Gn{\mathrm{G}}_{n} as in Proposition 3.3. We also write

S(w)={αΦ+w(α)<0}andS+(w)={αΦw(α)>0}.\displaystyle S^{-}(w)=\{\alpha\in\Phi^{+}\mid w(\alpha)<0\}\quad\mathrm{and}\quad S^{+}(w)=\{\alpha\in\Phi\mid w(\alpha)>0\}.

Let NwN_{w}^{-} (Nw+N_{w}^{+} resp.) be the unipotent subgroup associated to S(w)S^{-}(w) (S+(w)S^{+}(w) resp.). Let

Tw={tTnψ(u)=ψ(w(t)uw(t)1),uNw}.\displaystyle T_{w}=\{t\in T_{n}\mid\psi(u)=\psi(w(t)uw(t)^{-1}),\;\forall u\in N_{w}^{-}\}.

For every λX(Tn)\lambda\in X(T_{n})\otimes_{{\mathbb{Z}}}{\mathbb{R}}, where X(Tn)X(T_{n}) is the character group of TnT_{n}, define

|λ|(t):=|λ(t)|F,tTn.\displaystyle|\lambda|(t):=|\lambda(t)|_{F},\;\forall t\in T_{n}.

Recall from Section 2.1 that K=GLn(𝔬)K={\mathrm{GL}}_{n}({\mathfrak{o}}) is the maximal open compact subgroup of Gn{\mathrm{G}}_{n}. With the Iwasawa decomposition Gn=NnTnK{\mathrm{G}}_{n}=N_{n}T_{n}K, for any g=utkg=utk, we set |λ|(g):=|λ|(t)|\lambda|(g):=|\lambda|(t). It is easy to check that this is well defined.

Let πΠF(Gn)\pi\in\Pi_{F}({\mathrm{G}}_{n}) be generic and 𝒲(π,ψ){\mathcal{W}}(\pi,\psi) be the space of Whittaker functions. Following [Bar05, Definition 5.1], we denote by 𝒲(π,ψ){\mathcal{W}}^{\circ}(\pi,\psi) the set of functions W𝒲(π,ψ)W\in{\mathcal{W}}(\pi,\psi) such that for every w𝕎w\in{\mathbb{W}} and every αS(w)\alpha\in S^{\circ}(w), there exist positive constants Dα<EαD_{\alpha}<E_{\alpha} such that if gBnwBng\in B_{n}wB_{n} then W(g)0W(g)\neq 0 implies that Dα<|α|(g)<EαD_{\alpha}<|\alpha|(g)<E_{\alpha}. For a positive integer mm, we denote KmK_{m} the congruence subgroup given by Km=In+Mn(𝔭m)K_{m}={\mathrm{I}}_{n}+\mathrm{M}_{n}({\mathfrak{p}}^{m}). Write d=diag(1,ϖ2,ϖ4,,ϖ2n2)Gn(F)d={\rm diag}(1,\varpi^{2},\varpi^{4},\cdots,\varpi^{2n-2})\in{\mathrm{G}}_{n}(F), where ϖ\varpi is a fixed uniformizer of FF. Let Nn(m)=Nn(dmKmdm)N_{n}(m)=N_{n}\cap(d^{m}K_{m}d^{-m}). For any W𝒲(π,ψ)W\in{\mathcal{W}}(\pi,\psi), denote

Wm(g)=Nn(m)W(gn)ψ1(n)dn.\displaystyle W_{m}(g)=\int_{N_{n}(m)}W(gn)\psi^{-1}(n)\,\mathrm{d}n.

According to [Bar05, Theorem 7.3], Wm𝒲(π,ψ)W_{m}\in{\mathcal{W}}^{\circ}(\pi,\psi) for all sufficiently large mm. Due to [Bar05, Proposition 8.1], for mm large enough, the integral NwWm(gn)ψ1(n)dn\int_{N_{w}^{-}}W_{m}\left(gn\right)\psi^{-1}(n)\,\mathrm{d}n converges and is independent of mm for gNnTwwNwg\in N_{n}T_{w}wN_{w}^{-}. Moreover, by the uniqueness of Whittaker functionals, it follows that there exists a function, which we denote by jπ,ψ,w(g)j_{\pi,\psi,w}(g) such that

1vol(Nm)NwWm(gn)ψ1(n)dn=jπ,ψ,w(g)W(In)\displaystyle\frac{1}{\mathrm{vol}(N_{m})}\int_{N_{w}^{-}}W_{m}\left(gn\right)\psi^{-1}(n)\,\mathrm{d}n=j_{\pi,\psi,w}(g)W({\mathrm{I}}_{n})

for gNnTwwNwg\in N_{n}T_{w}wN_{w}^{-}. This function jπ,ψ,w(g)j_{\pi,\psi,w}(g) was called the Bessel function of π\pi attached to the Weyl group element ww in [Bar05, Section 8]. Moreover, if W𝒲(π,ψ)W\in{\mathcal{W}}^{\circ}(\pi,\psi), then the integral NwW(gn)ψ1(n)dn\displaystyle{\int_{N_{w}^{-}}W(gn)\psi^{-1}(n)\,\mathrm{d}n} converges absolutely for gNnTwwNwg\in N_{n}T_{w}wN_{w}^{-} and the Bessel function jπ,ψ,w(g)j_{\pi,\psi,w}(g) has the following integral representation:

(4.1) jπ,ψ,w(g)W(In)=NwW(gn)ψ1(n)dn\displaystyle j_{\pi,\psi,w}(g)\cdot W({\mathrm{I}}_{n})=\int_{N_{w}^{-}}W(gn)\psi^{-1}(n)\,\mathrm{d}n

according to [Bar05, Theorem 5.7 and Theorem 8.1].

Lemma 4.1.

Let FF be a non-Archimedean local field. Define

𝒲π,ψ(F×):={ω(x)=|x|1n2W((xIn1)):W𝒲(π,ψ)}.\displaystyle{\mathcal{W}}^{\circ}_{\pi,\psi}(F^{\times}):=\{\omega(x)=|x|^{1-\frac{n}{2}}\cdot W\left(\begin{pmatrix}x&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)\ \colon W\in{\mathcal{W}}^{\circ}(\pi,\psi)\}.

Then this space can be identified with the space 𝒞c(F×){\mathcal{C}}_{c}^{\infty}(F^{\times}): 𝒲π,ψ(π,ψ)=𝒞c(F×){\mathcal{W}}^{\circ}_{\pi,\psi}(\pi,\psi)={\mathcal{C}}_{c}^{\infty}(F^{\times}). In particular, the space 𝒲π,ψ(π,ψ){\mathcal{W}}^{\circ}_{\pi,\psi}(\pi,\psi) is independent of the choice of the character ψ\psi.

Proof.

We first prove that

{W((In1)):W𝒲(π,ψ)}=𝒞c(F×).\displaystyle\left\{W\left(\begin{pmatrix}\cdot&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right):\;W\in{\mathcal{W}}^{\circ}(\pi,\psi)\right\}={\mathcal{C}}_{c}^{\infty}(F^{\times}).

Take w=(In11)𝕎w_{*}=\begin{pmatrix}&{\mathrm{I}}_{n-1}\\ 1&\end{pmatrix}\in{\mathbb{W}}. Then we have that α1,2=e1e2S(wwn)\alpha_{1,2}=e_{1}-e_{2}\in S(w_{*}w_{n}). It follows that there are positive constants DD and EE such that W((xIn1))0W\left(\begin{pmatrix}x&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)\neq 0 implies that

D<|α1,2|((xIn1))=|x|<E,\displaystyle D<|\alpha_{1,2}|\left(\begin{pmatrix}x&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)=|x|<E,

which implies that

{W((In1)):W𝒲(π,ψ)}𝒞c(F×).\displaystyle\left\{W\left(\begin{pmatrix}\cdot&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right):\;W\in{\mathcal{W}}^{\circ}(\pi,\psi)\right\}\subset{\mathcal{C}}_{c}^{\infty}(F^{\times}).

On the other hand, take W𝒲(π,ψ)0W\in{\mathcal{W}}(\pi,\psi)\neq 0, for any positive integer mm, we have

Wm(In)=NmW(n)ψ1(n)dn=Vol(Nm)0,\displaystyle W_{m}({\mathrm{I}}_{n})=\int_{N_{m}}W(n)\psi^{-1}(n)\,\mathrm{d}n=\mathrm{Vol}(N_{m})\neq 0,

and since Wm𝒲(π,ψ)W_{m}\in{\mathcal{W}}^{\circ}(\pi,\psi) for mm sufficiently large, we get

{W((In1)):W𝒲(π,ψ)}0.\displaystyle\left\{W\left(\begin{pmatrix}\cdot&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)\colon W\in{\mathcal{W}}^{\circ}(\pi,\psi)\right\}\neq 0.

According to [Bar05, Corollary 5.5], 𝒲(π,ψ){\mathcal{W}}^{\circ}(\pi,\psi) is invariant under right translations by BnB_{n}, in particular, for b=(bIn2)Bb^{\prime}=\begin{pmatrix}b&\\ &{\mathrm{I}}_{n-2}\end{pmatrix}\in B, and W𝒲(π,ψ)W\in{\mathcal{W}}^{\circ}(\pi,\psi), where b=(tn1)b=\begin{pmatrix}t&n\\ &1\end{pmatrix}, we have

(bW)((xIn1))=ψ(nx)W((txIn1)).\displaystyle(bW)\left(\begin{pmatrix}x&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)=\psi(nx)W\left(\begin{pmatrix}tx&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right).

According to [JL70, Lemma 2.9.1], 𝒞c(F×){\mathcal{C}}_{c}^{\infty}(F^{\times}) is an irreducible representation under the above action. Hence we obtain that

{W((In1)):W𝒲(π,ψ)}=𝒞c(F×).\displaystyle\left\{W\left(\begin{pmatrix}\cdot&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)\colon W\in{\mathcal{W}}^{\circ}(\pi,\psi)\right\}={\mathcal{C}}_{c}^{\infty}(F^{\times}).

Note that ff||1n2f\mapsto f|\cdot|^{1-\frac{n}{2}} is a bijection from 𝒞c(F×){\mathcal{C}}_{c}^{\infty}(F^{\times}) to itself. Therefore we obtain that 𝒲π,ψ(F×)=𝒞c(F×){\mathcal{W}}^{\circ}_{\pi,\psi}(F^{\times})={\mathcal{C}}_{c}^{\infty}(F^{\times}). ∎

In order to understand the π\pi-Fourier transform π,ψ{\mathcal{F}}_{\pi,\psi} and the associated π\pi-kernel function kπ,ψk_{\pi,\psi} as in (2.20) in terms of the π\pi-Whittaker-Schwartz space 𝒲π(F×){\mathcal{W}}_{\pi}(F^{\times}) to π~\widetilde{\pi}-Whittaker-Schwartz space 𝒲π~(F×){\mathcal{W}}_{\widetilde{\pi}}(F^{\times}), we define the π\pi-Bessel function of π\pi on F×F^{\times}, which is related to the one attached to the particular Weyl element w=(In11)𝕎w_{*}=\begin{pmatrix}&{\mathrm{I}}_{n-1}\\ 1&\end{pmatrix}\in{\mathbb{W}}, up to normalization.

Definition 4.2.

Let FF be a non-Archimedean local field of characteristic zero. For any πΠF(Gn)\pi\in\Pi_{F}({\mathrm{G}}_{n}), which is generic, the associated π\pi-Bessel function 𝔟π,ψ(x){\mathfrak{b}}_{\pi,\psi}(x) on F×F^{\times} is defined by

(4.2) 𝔟π,ψ(x)=|x|F1n2jπ,ψ,w((In1x1))\displaystyle{\mathfrak{b}}_{\pi,\psi}(x)=|x|_{F}^{\frac{1-n}{2}}\cdot j_{\pi,\psi,w_{*}}\left(\begin{pmatrix}&{\mathrm{I}}_{n-1}\\ x^{-1}&\end{pmatrix}\right)

where w=(In11)𝕎w_{*}=\begin{pmatrix}&{\mathrm{I}}_{n-1}\\ 1&\end{pmatrix}\in{\mathbb{W}} is a Weyl group element of Gn{\mathrm{G}}_{n}

Proposition 4.3.

For any πΠF(Gn)\pi\in\Pi_{F}({\mathrm{G}}_{n}), which is generic, as functions on F×F^{\times}, the π\pi-kernel function kπ,ψk_{\pi,\psi} as in (2.20) and the π\pi-Bessel function as defined in (4.2) are related by the following identity:

kπ,ψ(x)=𝔟π,ψ(x)|x|12,xF×.k_{\pi,\psi}(x)={\mathfrak{b}}_{\pi,\psi}(x)|x|^{\frac{1}{2}},\quad\forall x\in F^{\times}.
Proof.

For any ϕ𝒞c(F×)𝒮π(F×)\phi\in{\mathcal{C}}_{c}^{\infty}(F^{\times})\subset{\mathcal{S}}_{\pi}(F^{\times}), we know from Proposition 4.1 that there is some W𝒲(π,ψ)W\in{\mathcal{W}}^{\circ}(\pi,\psi) such that

ϕ(x)=W((xIn1))|x|1n2\displaystyle\phi(x)=W\left(\begin{pmatrix}x&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)|x|^{1-\frac{n}{2}}

for any xF×x\in F^{\times}. According to Proposition 3.3, we have

π,ψ(ϕ)(x)\displaystyle{\mathcal{F}}_{\pi,\psi}(\phi)(x) =|x|1n2Fn2(π(wn,1)W~)((xyIn21))dy\displaystyle=|x|^{1-\frac{n}{2}}\int_{F^{n-2}}\left(\pi(w_{n,1})\widetilde{W}\right)\left(\begin{pmatrix}x&&\\ y&{\mathrm{I}}_{n-2}&\\ &&1\end{pmatrix}\right)\,\mathrm{d}y
=|x|1n2Fn2W(w(x1In1)(10y1yn2010000100001))dy\displaystyle=|x|^{1-\frac{n}{2}}\int_{F^{n-2}}W\left(w_{*}\begin{pmatrix}x^{-1}&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\begin{pmatrix}1&0&y_{1}&\cdots&y_{n-2}\\ 0&1&\cdots&0&0\\ \rotatebox{90.0}{$\cdots$}&\rotatebox{90.0}{$\cdots$}&\rotatebox{135.0}{$\cdots$}&\rotatebox{90.0}{$\cdots$}&\rotatebox{90.0}{$\cdots$}\\ 0&0&\cdots&1&0\\ 0&0&\cdots&0&1\end{pmatrix}\right)\,\mathrm{d}y

as w=wnwn,1w_{*}=w_{n}w_{n,1}. According to [Bar05, Theorem 5.7], the function

(z,y1,y2,,yn2)W(w(x1In1)(1zy1yn2010000100001))\displaystyle(z,y_{1},y_{2},\cdots,y_{n-2})\mapsto W\left(w_{*}\begin{pmatrix}x^{-1}&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\begin{pmatrix}1&z&y_{1}&\cdots&y_{n-2}\\ 0&1&\cdots&0&0\\ \rotatebox{90.0}{$\cdots$}&\rotatebox{90.0}{$\cdots$}&\rotatebox{135.0}{$\cdots$}&\rotatebox{90.0}{$\cdots$}&\rotatebox{90.0}{$\cdots$}\\ 0&0&\cdots&1&0\\ 0&0&\cdots&0&1\end{pmatrix}\right)

is compactly supported once we fix xx. Hence the function

f(z,x):=Fn2W(w(x1In1)(1zy1yn2010000100001))dy.\displaystyle f(z,x):=\int_{F^{n-2}}W\left(w_{*}\begin{pmatrix}x^{-1}&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\begin{pmatrix}1&z&y_{1}&\cdots&y_{n-2}\\ 0&1&\cdots&0&0\\ \rotatebox{90.0}{$\cdots$}&\rotatebox{90.0}{$\cdots$}&\rotatebox{135.0}{$\cdots$}&\rotatebox{90.0}{$\cdots$}&\rotatebox{90.0}{$\cdots$}\\ 0&0&\cdots&1&0\\ 0&0&\cdots&0&1\end{pmatrix}\right)\,\mathrm{d}y.

belongs to the space 𝒞c(F){\mathcal{C}}_{c}^{\infty}(F), as a function in zz with xx fixed, and its Fourier transform f^\widehat{f} along zz at 11

f^(1,x)=Ff(z,x)ψ1(z)dz\displaystyle\widehat{f}(1,x)=\int_{F}f(z,x)\psi^{-1}(z)\,\mathrm{d}z

exists. For the Weyl group element ww_{*}, it is easy to check that

Nw={(1zy1yn2010000100001)z,y1,,yn2F},N_{w_{*}}^{-}=\left\{\begin{pmatrix}1&z&y_{1}&\cdots&y_{n-2}\\ 0&1&\cdots&0&0\\ \rotatebox{90.0}{$\cdots$}&\rotatebox{90.0}{$\cdots$}&\rotatebox{135.0}{$\cdots$}&\rotatebox{90.0}{$\cdots$}&\rotatebox{90.0}{$\cdots$}\\ 0&0&\cdots&1&0\\ 0&0&\cdots&0&1\end{pmatrix}\mid z,y_{1},\cdots,y_{n-2}\in F\right\},

from which we deduce the following formula for f^(1,x)\widehat{f}(1,x):

f^(1,x)=NwW(w(x1In1)n)ψ1(n)dn\widehat{f}(1,x)=\int_{N_{w_{*}}^{-}}W\left(w_{*}\begin{pmatrix}x^{-1}&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}n\right)\psi^{-1}(n)\,\mathrm{d}n

where ψ(n)=ψ(z)\psi(n)=\psi(z). By (4.1), we obtain that

f^(1,x)=jπ,ψ,w(w(x1In1))W(In)=jπ,ψ,w((In1x1))W(In).\widehat{f}(1,x)=j_{\pi,\psi,w_{*}}\left(w_{*}\begin{pmatrix}x^{-1}&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)\cdot W({\mathrm{I}}_{n})=j_{\pi,\psi,w_{*}}\left(\begin{pmatrix}&{\mathrm{I}}_{n-1}\\ x^{-1}&\end{pmatrix}\right)\cdot W({\mathrm{I}}_{n}).

From the definition of the Bessel function jπ,ψ(x)j_{\pi,\psi}(x) in (4.2), we obtain that

f^(1,x)=𝔟π,ψ(x)|x|n12W(In)\widehat{f}(1,x)={\mathfrak{b}}_{\pi,\psi}(x)|x|^{\frac{n-1}{2}}W({\mathrm{I}}_{n})

as functions in xF×x\in F^{\times}. Now we calculate for a fixed xF×x\in F^{\times}, the Fourier transform f^(t,x)\widehat{f}(t,x) with tF×t\in F^{\times},

f^(t,x)\displaystyle\widehat{f}(t,x) =Ff(z,x)ψ1(tz)dz\displaystyle=\int_{F}f(z,x)\psi^{-1}(tz)\,\mathrm{d}z
=NwW(w((xt)1In1)n(tIn1))ψ1(z)dn|t|1n\displaystyle=\int_{N_{w_{*}}^{-}}W\left(w_{*}\begin{pmatrix}(xt)^{-1}&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}n\begin{pmatrix}t&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)\psi^{-1}(z)\,\mathrm{d}n\cdot|t|^{1-n}
=|t|1n𝔟π,ψ(xt)|xt|n12π((tIn1))W(In).\displaystyle=|t|^{1-n}{\mathfrak{b}}_{\pi,\psi}(xt)|xt|^{\frac{n-1}{2}}\pi\left(\begin{pmatrix}t&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)W({\mathrm{I}}_{n}).

According to [Bar05, Theorem 5.7], we can apply the Fourier inversion formula to obtain

f(0,x)\displaystyle f(0,x) =Ff^(t,x)dt\displaystyle=\int_{F}\widehat{f}(t,x)\,\mathrm{d}t
=F𝔟π,ψ(xt)|xt|n12W((tIn1))|t|1ndt\displaystyle=\int_{F}{\mathfrak{b}}_{\pi,\psi}(xt)|xt|^{\frac{n-1}{2}}W\left(\begin{pmatrix}t&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)|t|^{1-n}\,\mathrm{d}t
=F×𝔟π,ψ(xt)|xt|n12ϕ(t)|t|1n2d×t.\displaystyle=\int_{F^{\times}}{\mathfrak{b}}_{\pi,\psi}(xt)|xt|^{\frac{n-1}{2}}\phi(t)|t|^{1-\frac{n}{2}}\,\mathrm{d}^{\times}t.

Hence we obtain from the above calculation that

π,ψ(ϕ)(x)=|x|1n2f(0,x)=F×𝔟π,ψ(xt)|tx|12ϕ(t)d×t.\displaystyle{\mathcal{F}}_{\pi,\psi}(\phi)(x)=|x|^{1-\frac{n}{2}}f(0,x)=\int_{F^{\times}}{\mathfrak{b}}_{\pi,\psi}(xt)|tx|^{\frac{1}{2}}\phi(t)\,\mathrm{d}^{\times}t.

On the other hand, we know from [JL22, Theorem 5.2] that

π,ψ(ϕ)(x)=Fkπ,ψ(xt)ϕ(t)d×t\displaystyle{\mathcal{F}}_{\pi,\psi}(\phi)(x)=\int_{F}k_{\pi,\psi}(xt)\phi(t)\,\mathrm{d}^{\times}t

for any ϕ𝒞c(F×)\phi\in{\mathcal{C}}_{c}^{\infty}(F^{\times}). Therefore, as distributions on F×F^{\times}, we obtain that kπ,ψ(x)=𝔟π,ψ(x)|x|12k_{\pi,\psi}(x)={\mathfrak{b}}_{\pi,\psi}(x)|x|^{\frac{1}{2}} for any xF×x\in F^{\times}. Since both functions are smooth, the identity holds as functions in xF×x\in F^{\times}. ∎

Recall that in the GL2{\mathrm{GL}}_{2} case, D. Soudry defined in [Sou84] the Bessel function Jπ(x)J_{\pi}(x) on F×F^{\times} by the following equation

(4.3) FW((x1)(1y1))ψ1(y)dy=Jπ(x)W(I2)\displaystyle\int_{F}W\left(\begin{pmatrix}&x\\ -1&\end{pmatrix}\begin{pmatrix}1&y\\ &1\end{pmatrix}\right)\psi^{-1}(y)\,\mathrm{d}y=J_{\pi}(x)W({\mathrm{I}}_{2})

for all W𝒲(π,ψ)W\in{\mathcal{W}}(\pi,\psi), 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

Jπ(x)=ωπ(x)𝔟π,ψ(x)|x|12\displaystyle J_{\pi}(x)=\omega_{\pi}(x){\mathfrak{b}}_{\pi,\psi}(-x)|x|^{\frac{1}{2}}

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 JπJ_{\pi}. In fact, we have

Corollary 4.4.
F×pvJπ(y)χ1(y)|y|sd×y=ωπ(1)χ(1)γ(12,π×ωπ1χs,ψ).\int_{F^{\times}}^{\mathrm{pv}}J_{\pi}(y)\chi^{-1}(y)|y|^{-s}\,\mathrm{d}^{\times}y=\omega_{\pi}(-1)\chi(-1)\gamma(\frac{1}{2},\pi\times\omega_{\pi}^{-1}\chi_{s},\psi).
Proof.

[JL22, Theorem 5.2] tells us

F×pvkπ,ψ(y)χ1(y)|y|sd×y=γ(12,π×χs,ψ).\int_{F^{\times}}^{\mathrm{pv}}k_{\pi,\psi}(y)\chi^{-1}(y)|y|^{-s}\,\mathrm{d}^{\times}y=\gamma(\frac{1}{2},\pi\times\chi_{s},\psi).

Taking into account their relations, we can obtain what we want. ∎

We refer to [Cog14] for further discussion of the GL2{\mathrm{GL}}_{2}-Bessel functions and related topics.

4.2. π\pi-Bessel functions: complex case

If F=F={\mathbb{C}}, let us first recall from [Kna94] the classification of irreducible admissible representations of Gn=GLn(){\mathrm{G}}_{n}={\mathrm{GL}}_{n}({\mathbb{C}}). For zz\in{\mathbb{C}}, let [z]=z/zz¯[z]=z/\sqrt{z\overline{z}} and |z|=zz¯|z|_{{\mathbb{C}}}=z\overline{z}, where z¯\overline{z} is the complex conjugate of zz. For any ll\in{\mathbb{Z}} and tt\in{\mathbb{C}}, let σ=σ(l,t)\sigma=\sigma(l,t) be the representation of GL1(){\mathrm{GL}}_{1}({\mathbb{C}}) given by z[z]l|z|tz\mapsto[z]^{l}|z|_{{\mathbb{C}}}^{t}, which we write []l||t[\cdot]^{l}\otimes|\cdot|_{{\mathbb{C}}}^{t}. For each jj with 1jn1\leq j\leq n, let σj\sigma_{j} be the representation []lj||tj[\cdot]^{l_{j}}\otimes|\cdot|_{{\mathbb{C}}}^{t_{j}} of GL1(){\mathrm{GL}}_{1}({\mathbb{C}}). Then (σ1,,σn)(\sigma_{1},\cdots,\sigma_{n}) defines a one-dimensional representation of the diagonal maximal torus TnT_{n} of Gn{\mathrm{G}}_{n}, which can be extended trivially to a one-dimensional representation of the upper triangular Borel subgroup BnB_{n}. We set

I(σ1,,σn)=indBnGn(σ1,,σn),\displaystyle{\mathrm{I}}(\sigma_{1},\cdots,\sigma_{n})=\mathrm{ind}^{{\mathrm{G}}_{n}}_{B_{n}}(\sigma_{1},\cdots,\sigma_{n}),

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 Gn=GLn(){\mathrm{G}}_{n}={\mathrm{GL}}_{n}({\mathbb{C}}) can be classified as follows.

  • (1)

    If the parameters tjt_{j} of (σ1,,σn)(\sigma_{1},\cdots,\sigma_{n}) satisfies Ret1Ret2Retn{\mathrm{Re}}\;t_{1}\geq{\mathrm{Re}}\;t_{2}\geq\cdots\geq{\mathrm{Re}}\;t_{n}, then I(σ1,,σn){\mathrm{I}}(\sigma_{1},\cdots,\sigma_{n}) has a unique irreducible quotient J(σ1,,σn){\mathrm{J}}(\sigma_{1},\cdots,\sigma_{n}).

  • (2)

    the representations J(σ1,,σn){\mathrm{J}}(\sigma_{1},\cdots,\sigma_{n}) exhaust the irreducible admissible representations of GnG_{n}, up to infinitesimal equivalence.

  • (3)

    Two such representations J(σ1,,σn){\mathrm{J}}(\sigma_{1},\cdots,\sigma_{n}) and J(σ1,,σn){\mathrm{J}}(\sigma_{1}^{\prime},\cdots,\sigma_{n}^{\prime}) are infinitesimally equivalent if and only if there exists a permutation jj of {1,,n}\{1,\cdots,n\} such that σi=σj(i)\sigma_{i}^{\prime}=\sigma_{j(i)} for 1in1\leq i\leq n.

According to [Jac79], the associated local factors can be expressed as follows.

Theorem 4.6 (Local Factors).

Let π=J(σ1,,σn)\pi={\mathrm{J}}(\sigma_{1},\cdots,\sigma_{n}) be an irreducible admissible representation of Gn=GLn(){\mathrm{G}}_{n}={\mathrm{GL}}_{n}({\mathbb{C}}) with σj=[]lj||tj\sigma_{j}=[\cdot]^{l_{j}}\otimes|\cdot|_{{\mathbb{C}}}^{t_{j}}, where ljl_{j}\in{\mathbb{Z}} and tjt_{j}\in{\mathbb{C}} for every 1jn1\leq j\leq n. The local LL-factor and local ϵ\epsilon-factor associated with π\pi are given by

L(s,π)=j=1n2(2π)(s+ti+|lj|2)Γ(s+t+|lj|2)andϵ(s,π,ψ)=j=1ni|lj|.\displaystyle L(s,\pi)=\prod_{j=1}^{n}2(2\pi)^{-(s+t_{i}+\frac{|l_{j}|}{2})}\Gamma(s+t+\frac{|l_{j}|}{2})\quad{\rm and}\quad\epsilon(s,\pi,\psi)=\prod_{j=1}^{n}i^{|l_{j}|}.

For any mm\in{\mathbb{Z}}, the local γ\gamma-factor associated with π\pi is given by

γ(1s,π×[]m,ψ)\displaystyle\gamma(1-s,\pi\times[\cdot]^{m},\psi) =ϵ(1s,π×[]m,ψ)L(s,π~×[]m)L(1s,π×[]m)\displaystyle=\epsilon(1-s,\pi\times[\cdot]^{m},\psi)\frac{L(s,\widetilde{\pi}\times[\cdot]^{-m})}{L(1-s,\pi\times[\cdot]^{m})}
(4.4) =j=1ni|lj+m|(2π)12(stj)Γ(stj+|lj+m|2)Γ(1s+tj+|lj+m|2).\displaystyle=\prod_{j=1}^{n}i^{|l_{j}+m|}\cdot(2\pi)^{1-2(s-t_{j})}\cdot\frac{\Gamma(s-t_{j}+\frac{|l_{j}+m|}{2})}{\Gamma(1-s+t_{j}+\frac{|l_{j}+m|}{2})}.
Remark 4.7.

Using notations in [Qi20], we have that

γ(1s,π×[]m,ψ)=G(𝐭,𝐥+m𝐞𝐧)(s),\gamma(1-s,\pi\times[\cdot]^{m},\psi)=G_{(\mathbf{t},\mathbf{l}+m\mathbf{e^{n}})}(s),

where 𝐭=(t1,,tn)n\mathbf{t}=(t_{1},\cdots,t_{n})\in{\mathbb{C}}^{n}, 𝐥=(l1,,ln)n\mathbf{l}=(l_{1},\cdots,l_{n})\in{\mathbb{Z}}^{n}, and 𝐞𝐧=(1,,1)\mathbf{e^{n}}=(1,\cdots,1).

In [Qi20], Z. Qi defines a Bessel kernel function j𝐭,𝐥j_{\mathbf{t},{\mathbf{l}}} for any (𝐭,𝐥)n×n({\mathbf{t}},{\mathbf{l}})\in{\mathbb{C}}^{n}\times{\mathbb{Z}}^{n} by the following Mellin-Barnes type integral,

j𝐭,𝐥(x)=12πi𝒞(𝐭,𝐥)G(𝐭,𝐥)(s)x2sds,\displaystyle j_{{\mathbf{t}},{\mathbf{l}}}(x)=\frac{1}{2\pi i}\int_{{\mathcal{C}}_{({\mathbf{t}},{\mathbf{l}})}}G_{({\mathbf{t}},{\mathbf{l}})}(s)x^{-2s}\,\mathrm{d}s,

where

G𝐭,𝐥(s):=j=1ni|lj|(2π)12(stj)Γ(stj+|lj|2)Γ(1(stj)+|lj|2)\displaystyle G_{{\mathbf{t}},{\mathbf{l}}}(s):=\prod_{j=1}^{n}i^{|l_{j}|}(2\pi)^{1-2(s-t_{j})}\frac{\Gamma(s-t_{j}+\frac{|l_{j}|}{2})}{\Gamma(1-(s-t_{j})+\frac{|l_{j}|}{2})}

and 𝒞(𝐭,𝐥){\mathcal{C}}_{({\mathbf{t}},{\mathbf{l}})} is any contour such that

  • 2𝒞𝐭,𝐥2\cdot{\mathcal{C}}_{{\mathbf{t}},{\mathbf{l}}} is upward directed from σ\sigma-\infty to σ+\sigma+\infty, where σ<1+1n(Re(j=1ntj)1)\sigma<1+\frac{1}{n}\left({\mathrm{Re}}\;\displaystyle{\left(\sum_{j=1}^{n}t_{j}\right)-1}\right),

  • all the set tj|lj|t_{j}-|l_{j}|-{\mathbb{N}} lie on the left side of 2𝒞(𝐭,𝐥)2\cdot{\mathcal{C}}_{({\mathbf{t}},{\mathbf{l}})}, and

  • if s2𝒞(𝐭,𝐥)s\in 2\cdot{\mathcal{C}}_{({\mathbf{t}},{\mathbf{l}})} and |Ims||{\mathrm{Im}}\;s| large enough, then Res=σ{\mathrm{Re}}\;s=\sigma.

For more details, we refer to [Qi20, Definition 3.2]. Then [Qi20] defines

(4.5) J𝐭,𝐥(z)=12πmj(𝐭,𝐥+m𝐞𝐧)(|z|1/2)[z]m,\displaystyle J_{{\mathbf{t}},{\mathbf{l}}}(z)=\frac{1}{2\pi}\sum_{m\in{\mathbb{Z}}}j_{({\mathbf{t}},{\mathbf{l}}+m\mathbf{e^{n}})}(|z|_{{\mathbb{C}}}^{1/2})[z]^{m},

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 πΠ(Gn)\pi\in\Pi_{\mathbb{C}}({\mathrm{G}}_{n}), which is parameterized by π=π(𝐭,𝐥)\pi=\pi(\mathbf{t},\mathbf{l}) as in Theorem 4.5, as distributions on ×{\mathbb{C}}^{\times}, the identity: kπ,ψ(z)=J(𝐭,𝐥)(z)|z|12k_{\pi,\psi}(z)=J_{(\mathbf{t},\mathbf{l})}(z)|z|_{{\mathbb{C}}}^{\frac{1}{2}} holds for any z×z\in{\mathbb{C}}^{\times}.

Proof.

According to [Qi20, Theorem 3.15], for any ϕ𝒞c(×)\phi\in{\mathcal{C}}_{c}^{\infty}({\mathbb{C}}^{\times}), there is a unique function Υ(z)|z|12𝒮sis(𝐭,𝐥)(×)\Upsilon(z)|z|_{{\mathbb{C}}}^{\frac{1}{2}}\in\mathcal{S}_{\mathrm{sis}}^{(-{\mathbf{t}},-{\mathbf{l}})}({\mathbb{C}}^{\times}), which is contained in the space (×){\mathcal{F}}({\mathbb{C}}^{\times}) as defined in [JL23, Definition 2.1], such that

𝒵(1s,ϕ||12,[]m)γ(1s,π×[]m,ψ)=𝒵(s,Υ||12,[]m).\displaystyle{\mathcal{Z}}(1-s,\phi|\cdot|_{{\mathbb{C}}}^{\frac{1}{2}},[\cdot]^{m})\gamma(1-s,\pi\times[\cdot]^{m},\psi)={\mathcal{Z}}(s,\Upsilon|\cdot|_{{\mathbb{C}}}^{\frac{1}{2}},[\cdot]^{-m}).

It follows that Υ||12=π,ψ(ϕ||12)\Upsilon|\cdot|_{{\mathbb{C}}}^{\frac{1}{2}}={\mathcal{F}}_{\pi,\psi}(\phi|\cdot|_{{\mathbb{C}}}^{\frac{1}{2}}) according to [JL23, Theorem 2.3, Proposition 3.7, and Corollary 3.8]. From [Qi20, Proposition 3.17], we have that

Υ(z)=×ϕ(y)J(𝐭,𝐥)(zy)dy.\displaystyle\Upsilon(z)=\int_{{\mathbb{C}}^{\times}}\phi(y)J_{(\mathbf{t},\mathbf{l})}(zy)\,\mathrm{d}y.

On the other hand, we have that

Υ(z)=×ϕ(y)kπ,ψ(yz)|yz|12dy\displaystyle\Upsilon(z)=\int_{{\mathbb{C}}^{\times}}\phi(y)k_{\pi,\psi}(yz)|yz|_{{\mathbb{C}}}^{-\frac{1}{2}}\,\mathrm{d}y

due to [JL22, Theorem 5.1]. The π\pi-kernel function kπ,ψk_{\pi,\psi} is a smooth function on ×{\mathbb{C}}^{\times} according to [JL22, Corollary 4.5], while the function J𝐭,𝐥J_{{\mathbf{t}},{\mathbf{l}}} is real analytic on ×{\mathbb{C}}^{\times} due to [Qi20, Proposition 3.17]. Since ϕ𝒞c(×)\phi\in{\mathcal{C}}_{c}^{\infty}({\mathbb{C}}^{\times}) is arbitrary, we thus deduce that kπ,ψ(z)=J(𝐭,𝐥)(z)|z|12k_{\pi,\psi}(z)=J_{(\mathbf{t},\mathbf{l})}(z)|z|_{{\mathbb{C}}}^{\frac{1}{2}} for any z×z\in{\mathbb{C}}^{\times}, as functions on ×{\mathbb{C}}^{\times}. ∎

As in Definition 4.2, we introduce the π\pi-Beesel function on ×{\mathbb{C}}^{\times}.

Definition 4.9.

For any πΠ(Gn)\pi\in\Pi_{\mathbb{C}}({\mathrm{G}}_{n}), which is generic, the π\pi-Beesel function 𝔟π,ψ(x){\mathfrak{b}}_{\pi,\psi}(x) on ×{\mathbb{C}}^{\times} is given as

𝔟π,ψ(x)=J(𝐭,𝐥)(x){\mathfrak{b}}_{\pi,\psi}(x)=J_{(\mathbf{t},\mathbf{l})}(x)

for any x×x\in{\mathbb{C}}^{\times}, where π=π(𝐭,𝐥)\pi=\pi(\mathbf{t},\mathbf{l}) is given by the classification in Theorem 4.5, and J(𝐭,𝐥)(x)J_{(\mathbf{t},\mathbf{l})}(x) is given in (4.5) and was originally defined in [Qi20].

4.3. π\pi-Bessel functions: real case

If F=F={\mathbb{R}}, we recall from [Kna94] the classification of irreducible admissible representations of Gn=GLn(){\mathrm{G}}_{n}={\mathrm{GL}}_{n}({\mathbb{R}}). For any l1l\geq 1, let Dl+D_{l}^{+} be the discrete series of SL2(){\mathrm{SL}}_{2}({\mathbb{R}}), that is, the representation space consists of analytic functions ff in the upper half-plane with

f2:=|f(z)|2yl1dxdy\displaystyle\|f\|^{2}:=\iint|f(z)|^{2}y^{l-1}\,\mathrm{d}x\,\mathrm{d}y

finite, and the action of g=(abcd)g=\begin{pmatrix}a&b\\ c&d\end{pmatrix} is given by

Dl+(g)f(z):=(bz+d)(l+1)f(az+cbz+d).\displaystyle D_{l}^{+}(g)f(z):=(bz+d)^{-(l+1)}f\left(\frac{az+c}{bz+d}\right).

Let SL2±(){\mathrm{SL}}_{2}^{\pm}({\mathbb{R}}) be the subgroup of elements gg in GL2(){\mathrm{GL}}_{2}({\mathbb{R}}) with |detg|=1|\det g|=1 and

Dl:=indSL2()SL2±()(Dl+)\displaystyle D_{l}:=\mathrm{ind}_{{\mathrm{SL}}_{2}({\mathbb{R}})}^{{\mathrm{SL}}_{2}^{\pm}({\mathbb{R}})}(D_{l}^{+})

be the induced representation of SL2±(){\mathrm{SL}}_{2}^{\pm}({\mathbb{R}}), where we still use the unitary induction as in [Kna01, Chapter VII]. For each pair (l,t)1×(l,t)\in{\mathbb{Z}}_{\geq 1}\times{\mathbb{C}}, let σ=σ(l,t)\sigma=\sigma(l,t) be the representation of GL2(){\mathrm{GL}}_{2}({\mathbb{R}}) obtained by tensoring the above representation on SL±(){\mathrm{SL}}^{\pm}({\mathbb{R}}) with the quasi-character g|detg|tg\mapsto|\det g|^{t}, that is, σ=Dl|det()|t\sigma=D_{l}\otimes|\det(\cdot)|^{t}, where ||=|||\cdot|=|\cdot|_{\mathbb{R}}. For a pair (δ,t)/2×(\delta,t)\in\mathbb{Z}/2\mathbb{Z}\times{\mathbb{C}}, let σ=σ(δ,t)\sigma=\sigma(\delta,t) be the representation of GL1()=×{\mathrm{GL}}_{1}({\mathbb{R}})={\mathbb{R}}^{\times}: σ=sgnδ||t\sigma={\mathrm{sgn}}^{\delta}\otimes|\cdot|^{t}.

For any partition of nn: (n1,,nr)(n_{1},\cdots,n_{r}) with each njn_{j} equal to 11 or 22 and with j=1rnj=n\displaystyle{\sum_{j=1}^{r}n_{j}=n}, we associate the block diagonal subgroup M=GLn1()××GLnr()M={\mathrm{GL}}_{n_{1}}({\mathbb{R}})\times\cdots\times{\mathrm{GL}}_{n_{r}}({\mathbb{R}}). For each 1jr1\leq j\leq r, let σj\sigma_{j} be the representation of GLnj(){\mathrm{GL}}_{n_{j}}({\mathbb{R}}) of the form σ(lj,tj)\sigma(l_{j},t_{j}) or σ(δj,tj)\sigma(\delta_{j},t_{j}) as defined above. We extend the tensor product of these representations to the corresponding block upper triangular subgroup QQ by making it the identity on the block strictly upper triangular subgroup. We set

I(σ1,,σr):=indQGn(σ1,,σr).\displaystyle{\mathrm{I}}(\sigma_{1},\cdots,\sigma_{r}):=\mathrm{ind}_{Q}^{{\mathrm{G}}_{n}}(\sigma_{1},\cdots,\sigma_{r}).
Theorem 4.10 (Classification).

The irreducible admissible representations of Gn=GLn(){\mathrm{G}}_{n}={\mathrm{GL}}_{n}({\mathbb{R}}) can be classified as follows.

  • (1)

    If the parameters nj1tjn_{j}^{-1}t_{j} of (σ1,,σr)(\sigma_{1},\cdots,\sigma_{r}) satisfy

    n11Ret1n21Ret2nr1Retr,\displaystyle n_{1}^{-1}{\mathrm{Re}}\;t_{1}\geq n_{2}^{-1}{\mathrm{Re}}\;t_{2}\geq\cdots\geq n_{r}^{-1}{\mathrm{Re}}\;t_{r},

    then I(σ1,,σr){\mathrm{I}}(\sigma_{1},\cdots,\sigma_{r}) has a unique irreducible quotient J(σ1,,σr){\mathrm{J}}(\sigma_{1},\cdots,\sigma_{r}).

  • (2)

    The representations J(σ1,,σr){\mathrm{J}}(\sigma_{1},\cdots,\sigma_{r}) exhaust the irreducible admissible representations of Gn{\mathrm{G}}_{n}, up to infinitesimal equivalence.

  • (3)

    Two such representations J(σ1,,σr){\mathrm{J}}(\sigma_{1},\cdots,\sigma_{r}) and J(σ1,,σr){\mathrm{J}}(\sigma_{1}^{\prime},\cdots,\sigma_{r}^{\prime}) are infinitesimally equivalent if and only if r=rr^{\prime}=r and there exists a permutation j(i)j(i) such that σi=σj(i)\sigma_{i}^{\prime}=\sigma_{j(i)} for each 1ir1\leq i\leq r.

According to [Jac79] again, the local factors can be expressed as follows:

Theorem 4.11 (Local Factors).

For a representation σ\sigma of GL1(){\mathrm{GL}}_{1}({\mathbb{R}}) or GL2(){\mathrm{GL}}_{2}({\mathbb{R}}) as defined above, denote

L(s,σ)={πs+t+δ2Γ(s+t+δ2)ifn=1,σ=sgnδ||t,2(2π)(s+t+l2)Γ(s+t+l2)ifn=2,σ=Dl|det()|t.\displaystyle L(s,\sigma)=\begin{cases}\pi^{-\frac{s+t+\delta}{2}}\Gamma(\frac{s+t+\delta}{2})&{\rm if}\ n=1,\;\sigma={\mathrm{sgn}}^{\delta}\otimes|\cdot|^{t},\\ 2(2\pi)^{-(s+t+\frac{l}{2})}\Gamma(s+t+\frac{l}{2})&{\rm if}\ n=2,\;\sigma=D_{l}\otimes|\det(\cdot)|^{t}.\end{cases}

then for π=J(σ1,,σr)\pi={\mathrm{J}}(\sigma_{1},\cdots,\sigma_{r}), we have L(s,π)=j=1rL(s,σj)L(s,\pi)=\prod_{j=1}^{r}L(s,\sigma_{j}). Similarly denote

ϵ(s,σ,ψ)={iδifn=1,σ=sgnδ||t,il+1ifn=2,σ=Dl|det()|t.\displaystyle\epsilon(s,\sigma,\psi)=\begin{cases}i^{\delta}&{\rm if}\ n=1,\;\sigma={\mathrm{sgn}}^{\delta}\otimes|\cdot|^{t},\\ i^{l+1}&{\rm if}\ n=2,\;\sigma=D_{l}\otimes|\det(\cdot)|^{t}.\end{cases}

then the ϵ\epsilon-factor of π=J(σ1,,σr)\pi={\mathrm{J}}(\sigma_{1},\cdots,\sigma_{r}) is given by ϵ(s,π,ψ)=j=1lϵ(s,σj,ψ)\epsilon(s,\pi,\psi)=\prod_{j=1}^{l}\epsilon(s,\sigma_{j},\psi). Finally, the local γ\gamma-factor associated with π=J(σ1,,σr)\pi={\mathrm{J}}(\sigma_{1},\cdots,\sigma_{r}) is given by

γ(s,π×sgnδ,ψ)=ϵ(s,π×sgnδ,ψ)L(1s,π~×sgnδ)L(s,π×sgnδ),\displaystyle\gamma(s,\pi\times{\mathrm{sgn}}^{\delta},\psi)=\epsilon(s,\pi\times{\mathrm{sgn}}^{\delta},\psi)\frac{L(1-s,\widetilde{\pi}\times{\mathrm{sgn}}^{\delta})}{L(s,\pi\times{\mathrm{sgn}}^{\delta})},

where π~\widetilde{\pi} is the contragredient of π\pi.

For any ϕ(x)𝒞c(×)\phi(x)\in{\mathcal{C}}_{c}^{\infty}({\mathbb{R}}^{\times}), according to [JL23, Theorem 3.10], there is some function Υ\Upsilon such that Υ||12=π~,ψ(ϕ||12)\Upsilon|\cdot|^{\frac{1}{2}}={\mathcal{F}}_{\widetilde{\pi},\psi}(\phi|\cdot|^{\frac{1}{2}}) such that

𝒵(s,Υ||12,sgnδ)=γ(1s,π×δ,ψ)𝒵(1s,ϕ||12,sgnδ).\displaystyle{\mathcal{Z}}(s,\Upsilon|\cdot|^{\frac{1}{2}},{\mathrm{sgn}}^{\delta})=\gamma(1-s,\pi\times\delta,\psi)\cdot{\mathcal{Z}}(1-s,\phi|\cdot|^{\frac{1}{2}},{\mathrm{sgn}}^{\delta}).

Due to [Igu78, Theorem 4.2], for Res=σ0{\mathrm{Re}}\;s=\sigma_{0} large enough, we have

Υ(x)\displaystyle\Upsilon(x) =12δ/2(12πiσ0iσ0+iγ(1s,π×sgnδ,ψ)𝒵(1s,ϕ||12,sgnδ)|x|sds)(sgnx)δ\displaystyle=\frac{1}{2}\sum_{\delta\in{\mathbb{Z}}/2{\mathbb{Z}}}\left(\frac{1}{2\pi i}\int_{\sigma_{0}-i\infty}^{\sigma_{0}+i\infty}\gamma(1-s,\pi\times{\mathrm{sgn}}^{\delta},\psi)\cdot{\mathcal{Z}}(1-s,\phi|\cdot|^{\frac{1}{2}},{\mathrm{sgn}}^{\delta})|x|^{-s}\,\mathrm{d}s\right)({\mathrm{sgn}}x)^{\delta}
=12δ/2(12πiσ0iσ0+iγ(1s,π×sgnδ,ψ)×ϕ(y)|y|sdy|x|sds)(sgnx)δ.\displaystyle=\frac{1}{2}\sum_{\delta\in{\mathbb{Z}}/2{\mathbb{Z}}}\left(\frac{1}{2\pi i}\int_{\sigma_{0}-i\infty}^{\sigma_{0}+i\infty}\gamma(1-s,\pi\times{\mathrm{sgn}}^{\delta},\psi)\cdot\int_{{\mathbb{R}}^{\times}}\phi(y)|y|^{-s}\,\mathrm{d}y|x|^{-s}\,\mathrm{d}s\right)({\mathrm{sgn}}x)^{\delta}.

We choose a contour 𝒞{\mathcal{C}} with the following three properties:

  • (1)

    𝒞{\mathcal{C}} is upward directed from σ0i\sigma_{0}^{\prime}-i\infty to σ0+i\sigma_{0}^{\prime}+i\infty, where σ0\sigma_{0}^{\prime} is small enough, say

    σ0<12+Re(j=1rnjtj)1n,\displaystyle\sigma_{0}^{\prime}<\frac{1}{2}+\frac{{\mathrm{Re}}\left(\sum_{j=1}^{r}n_{j}t_{j}\right)-1}{n},
  • (2)

    The sets tjδjt_{j}-\delta_{j}-{\mathbb{N}} for nj=1n_{j}=1 and tjlj2t_{j}-\frac{l_{j}}{2}-{\mathbb{N}} for nj=2n_{j}=2, 1jr1\leq j\leq r all lie on the left side of 𝒞{\mathcal{C}}, and

  • (3)

    If s𝒞s\in{\mathcal{C}}, then for |Ims||{\mathrm{Im}}\;s| large enough, Res=σ0{\mathrm{Re}}s=\sigma_{0}^{\prime}.

Then for t=|Ims|t=|{\mathrm{Im}}\;s| large enough, we have that for fixed x×x\in{\mathbb{R}}^{\times} and ϕ𝒞c(×)\phi\in{\mathcal{C}}_{c}^{\infty}({\mathbb{R}}^{\times}),

×ϕ(y)|y|sdy|x|sC\displaystyle\int_{{\mathbb{R}}^{\times}}\phi(y)|y|^{-s}\,\mathrm{d}y|x|^{-s}\leq C

for some constant CC for all s with σ0Resσ0\sigma_{0}^{\prime}\leq{\mathrm{Re}}\;s\leq\sigma_{0}, and the constant CC only depends on xx, φ\varphi, σ0\sigma_{0}, and σ0\sigma_{0}^{\prime}, and is independent of t=|Ims|t=|{\mathrm{Im}}\;s|. It follows that

σ0+itσ0+itγ(1s,π×sgnδ,ψ)×ϕ(y)|y|sdy|x|sdsCt1\displaystyle\int_{\sigma_{0}^{\prime}+it}^{\sigma_{0}+it}\gamma(1-s,\pi\times{\mathrm{sgn}}^{\delta},\psi)\int_{{\mathbb{R}}^{\times}}\phi(y)|y|^{-s}\,\mathrm{d}y|x|^{-s}\,\mathrm{d}s\leq C^{\prime}t^{-1}

for some other constant CC^{\prime} according to [Qi20, Lemma 1.3] and Property (1) of the contour 𝒞{\mathcal{C}}. Hence, as tt\rightarrow\infty, the above integral goes zero, and we are able to change the integral from (σ0i,σ0+i)(\sigma_{0}-i\infty,\sigma_{0}+i\infty) to 𝒞{\mathcal{C}} according to the Cauchy residue theorem and Property (2) of the contour 𝒞{\mathcal{C}}, that is

δ/2(12πiσ0iσ0+iγ(1s,π×sgnδ,ψ)×ϕ(y)|y|sdy|x|sds)(sgnx)δ\displaystyle\sum_{\delta\in{\mathbb{Z}}/2{\mathbb{Z}}}\left(\frac{1}{2\pi i}\int_{\sigma_{0}-i\infty}^{\sigma_{0}+i\infty}\gamma(1-s,\pi\times{\mathrm{sgn}}^{\delta},\psi)\cdot\int_{{\mathbb{R}}^{\times}}\phi(y)|y|^{-s}\,\mathrm{d}y|x|^{-s}\,\mathrm{d}s\right)({\mathrm{sgn}}x)^{\delta}
=δ/2(12πi𝒞γ(1s,π×sgnδ,ψ)×ϕ(y)|y|sdy|x|sds)(sgnx)δ.\displaystyle\qquad=\sum_{\delta\in{\mathbb{Z}}/2{\mathbb{Z}}}\left(\frac{1}{2\pi i}\int_{{\mathcal{C}}}\gamma(1-s,\pi\times{\mathrm{sgn}}^{\delta},\psi)\cdot\int_{{\mathbb{R}}^{\times}}\phi(y)|y|^{-s}\,\mathrm{d}y|x|^{-s}\,\mathrm{d}s\right)({\mathrm{sgn}}x)^{\delta}.

According to Property (1) of the contour 𝒞{\mathcal{C}} and [Qi20, Lemma 1.3] again, we have that

𝒞×|γ(1s,π×sgnδ,ψ)ϕ(y)||xy|sdyds<.\displaystyle\int_{{\mathcal{C}}}\int_{{\mathbb{R}}^{\times}}|\gamma(1-s,\pi\times{\mathrm{sgn}}^{\delta},\psi)\phi(y)|\cdot|xy|^{-s}\,\mathrm{d}y\,\mathrm{d}s<\infty.

Hence we can change the order of integration using Fubini’s theorem to obtain that

Υ(x)=×ϕ(y)(12δ/212πi𝒞γ(1s,π×sgnδ,ψ)|xy|ssgn(x)δds)dy.\displaystyle\Upsilon(x)=\int_{{\mathbb{R}}^{\times}}\phi(y)\left(\frac{1}{2}\sum_{\delta\in{\mathbb{Z}}/2{\mathbb{Z}}}\frac{1}{2\pi i}\int_{{\mathcal{C}}}\gamma(1-s,\pi\times{\mathrm{sgn}}^{\delta},\psi)|xy|^{-s}{\mathrm{sgn}}(x)^{\delta}\,\mathrm{d}s\right)\,\mathrm{d}y.

As in Definitions 4.2 and 4.9, we define the π\pi-Bessel function 𝔟π,ψ(x){\mathfrak{b}}_{\pi,\psi}(x) on ×{\mathbb{R}}^{\times} as follows

Definition 4.12.

For any πΠ(Gn)\pi\in\Pi_{\mathbb{R}}({\mathrm{G}}_{n}), which is generic, the π\pi-Beesel function 𝔟π,ψ(x){\mathfrak{b}}_{\pi,\psi}(x) on ×{\mathbb{R}}^{\times} is given as

𝔟π,ψ(±x)=12δ/212πi𝒞γ(1s,π×sgnδ,ψ)|x|s(±)δds,x>0.\displaystyle{\mathfrak{b}}_{\pi,\psi}(\pm x)=\frac{1}{2}\sum_{\delta\in{\mathbb{Z}}/2{\mathbb{Z}}}\frac{1}{2\pi i}\int_{{\mathcal{C}}}\gamma(1-s,\pi\times{\mathrm{sgn}}^{\delta},\psi)|x|^{-s}(\pm)^{\delta}\,\mathrm{d}s,\;x>0.

The integral in Definition 4.12 is absolutely convergent to a smooth function in xx because of Property (1) of the contour 𝒞{\mathcal{C}} 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 πΠ(Gn)\pi\in\Pi_{\mathbb{R}}({\mathrm{G}}_{n}), which is generic, the π\pi-kernel function kπ,ψ(x)k_{\pi,\psi}(x) and the π\pi-Bessel function 𝔟π,ψ(x){\mathfrak{b}}_{\pi,\psi}(x) are related by the following identity as functions on ×{\mathbb{R}}^{\times}, i.e.

kπ,ψ(x)=𝔟π,ψ(x)|x|12,x×.\displaystyle k_{\pi,\psi}(x)={\mathfrak{b}}_{\pi,\psi}(x)|x|_{\mathbb{R}}^{\frac{1}{2}},\quad\forall x\in{\mathbb{R}}^{\times}.
Proof.

Similar to Proposition 4.9, let us compare the integral

Υ(x)=×ϕ(y)𝔟π,ψ(xy)dy\displaystyle\Upsilon(x)=\int_{{\mathbb{R}}^{\times}}\phi(y){\mathfrak{b}}_{\pi,\psi}(xy)\,\mathrm{d}y

with the integral

Υ(x)=×ϕ(y)kπ,ψ(xy)|xy|12dy,\displaystyle\Upsilon(x)=\int_{{\mathbb{R}}^{\times}}\phi(y)k_{\pi,\psi}(xy)|xy|_{\mathbb{R}}^{-\frac{1}{2}}\,\mathrm{d}y,

for any ϕ𝒞c(×)\phi\in{\mathcal{C}}^{\infty}_{c}({\mathbb{R}}^{\times}). It is clear that kπ,ψ=𝔟π,ψ(x)|x|12k_{\pi,\psi}={\mathfrak{b}}_{\pi,\psi}(x)|x|_{\mathbb{R}}^{\frac{1}{2}} because of Definition 4.12 and the smothness of both kπ,ψk_{\pi,\psi} ([JL22, Corollary 4.5]) and 𝔟π,ψ{\mathfrak{b}}_{\pi,\psi} as functions on ×{\mathbb{R}}^{\times}. ∎

Remark 4.14.

In the special case that π=π(𝐭,δ)=J(σ1,,σn)\pi=\pi({\mathbf{t}},\mathbf{\delta})={\mathrm{J}}(\sigma_{1},\cdots,\sigma_{n}) as Theorem 4.10 with all nj=1n_{j}=1 for 1jn1\leq j\leq n, the π\pi-Bessel function 𝔟π,ψ{\mathfrak{b}}_{\pi,\psi} in Definition 4.12 is exactly the Bessel function J(𝐭,δ)J_{({\mathbf{t}},\mathbf{\delta})} defined in [Qi20, Section 3.3.2], where 𝐭=(t1,,tn)n{\mathbf{t}}=(t_{1},\cdots,t_{n})\in{\mathbb{C}}^{n} and δ=(δ1,,δn)(/2)n\mathbf{\delta}=(\delta_{1},\cdots,\delta_{n})\in({\mathbb{Z}}/2{\mathbb{Z}})^{n}.

4.4. π\pi-Bessel functions and dual functions

From Definitions 4.2, 4.9 and 4.12, for a given πΠF(Gn)\pi\in\Pi_{F}({\mathrm{G}}_{n}), we define the (normalized) π\pi-Bessel function 𝔟π,ψ(x){\mathfrak{b}}_{\pi,\psi}(x) on F×F^{\times} for every local field FF of characteristic zero. In Propositions 4.3, 4.8 and 4.13, we obtain the relation between the π\pi-kernel function kπ,ψ(x)k_{\pi,\psi}(x) and the π\pi-Bessel function 𝔟π,ψ(x){\mathfrak{b}}_{\pi,\psi}(x). As a record, we state the corresponding formula for the dual function w~(x)\widetilde{w}(x) of w(x)𝒞c(F×)w(x)\in{\mathcal{C}}^{\infty}_{c}(F^{\times}) following Corollary 3.6

Corollary 4.15.

For any πΠF(Gn)\pi\in\Pi_{F}({\mathrm{G}}_{n}), the dual function w~(x)\widetilde{w}(x) associated with any w𝒞c(F×)w\in{\mathcal{C}}^{\infty}_{c}(F^{\times}) is given by the following formula:

w~(x)=|x|Fn12(𝔟π,ψ()(w()||Fn32))(x),\widetilde{w}(x)=|x|_{F}^{\frac{n-1}{2}}\left({\mathfrak{b}}_{\pi,\psi}(\cdot)*(w^{\vee}(\cdot)|\cdot|_{F}^{\frac{n-3}{2}})\right)(x),

for all xF×x\in F^{\times}, where w(x)=w(x1)w^{\vee}(x)=w(x^{-1}).

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 π\pi-Poisson summation formula ([JL23, Theorem 4.7]), which was recalled in Theorem 2.1. Let kk be a number field, the notations are all as in Section 2.

Lemma 5.1.

At any local place ν\nu of kk, for any wν(x)𝒞c(kν×)w_{\nu}(x)\in{\mathcal{C}}_{c}^{\infty}(k_{\nu}^{\times}) and ζνkν×\zeta_{\nu}\in k_{\nu}^{\times}, the function

ϕν(x)=ψν(xζν)wν(x)|x|ν1n2\phi_{\nu}(x)=\psi_{\nu}(x\zeta_{\nu})\cdot w_{\nu}(x)\cdot|x|_{\nu}^{1-\frac{n}{2}}

belongs to the space 𝒮πν(kν×){\mathcal{S}}_{\pi_{\nu}}(k_{\nu}^{\times}). If ν<\nu<\infty and πν\pi_{\nu} is unramified, let Wν{}^{\circ}W_{\nu} be the normalized unramifield Whittaker function associated with πν\pi_{\nu}, then the function

φν(x)=ψν(xζν)Wν((xIn1))|x|ν1n2\varphi_{\nu}(x)=\psi_{\nu}(x\zeta_{\nu})\cdot{{}^{\circ}W_{\nu}}\left(\begin{pmatrix}x&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)\cdot|x|_{\nu}^{1-\frac{n}{2}}

belongs to the space 𝒮πν(kν×){\mathcal{S}}_{\pi_{\nu}}(k_{\nu}^{\times}).

Proof.

The first claim is trivial because for any wν(x)𝒞c(kν×)w_{\nu}(x)\in{\mathcal{C}}_{c}^{\infty}(k_{\nu}^{\times}), we have that

ϕν(x)=ψν(xζν)wν(x)|x|ν1n2𝒞c(kν×)𝒮πν(kν×).\displaystyle\phi_{\nu}(x)=\psi_{\nu}(x\zeta_{\nu})w_{\nu}(x)|x|_{\nu}^{1-\frac{n}{2}}\in{\mathcal{C}}_{c}^{\infty}(k_{\nu}^{\times})\subset{\mathcal{S}}_{\pi_{\nu}}(k_{\nu}^{\times}).

As for the second claim, since ν<\nu<\infty, we observe that for any given ζνkν×\zeta_{\nu}\in k_{\nu}^{\times}, ψν(xζν)=1\psi_{\nu}(x\zeta_{\nu})=1 if |x|ν|x|_{\nu} is small enough. Hence we have that

φν(x)=ψν(xζν)Wν((xIn1))|x|ν1n2\displaystyle\varphi_{\nu}(x)=\psi_{\nu}(x\zeta_{\nu})\cdot{{}^{\circ}W_{\nu}}\left(\begin{pmatrix}x&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)\cdot|x|_{\nu}^{1-\frac{n}{2}}

shares the same asymptotic behavior as |x|0|x|\rightarrow 0 with the function Wν((xIn1))|x|ν1n2{}^{\circ}W_{\nu}\left(\begin{pmatrix}x&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)|x|_{\nu}^{1-\frac{n}{2}}, which belongs to the space 𝒮πν(kν×){\mathcal{S}}_{\pi_{\nu}}(k_{\nu}^{\times}) by Proposition 3.1. Hence we must have the function

φν(x)=ψν(xζν)Wν((xIn1))|x|ν1n2\varphi_{\nu}(x)=\psi_{\nu}(x\zeta_{\nu})\cdot{{}^{\circ}W_{\nu}}\left(\begin{pmatrix}x&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)\cdot|x|_{\nu}^{1-\frac{n}{2}}

belonging to the space 𝒮πν(kν×){\mathcal{S}}_{\pi_{\nu}}(k_{\nu}^{\times}). ∎

Lemma 5.2.

Let ν\nu be a finite place such that both πν\pi_{\nu} and ψν\psi_{\nu} are unramified, the πν\pi_{\nu}-basic function 𝕃πν(x)𝒮πν(kν×){\mathbb{L}}_{\pi_{\nu}}(x)\in{\mathcal{S}}_{\pi_{\nu}}(k_{\nu}^{\times}) as defined in [JL23, Theorem 3.4] enjoys the following formula:

𝕃πν(x)=Wν((xIn1))|x|ν1n2.{\mathbb{L}}_{\pi_{\nu}}(x)={{}^{\circ}W_{\nu}}\left(\begin{pmatrix}x&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)|x|_{\nu}^{1-\frac{n}{2}}.
Proof.

By [JL23, Theorem 3.4], the Mellin transform of the πν\pi_{\nu}-basic function 𝕃πν(x){\mathbb{L}}_{\pi_{\nu}}(x) equals L(s,π×χ)L(s,\pi\times\chi), and the same happens to the function Wν((xIn1))|x|ν1n2{{}^{\circ}W_{\nu}}\left(\begin{pmatrix}x&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)|x|_{\nu}^{1-\frac{n}{2}} by the Rankin-Selberg convolution for GLn×GL1{\mathrm{GL}}_{n}\times{\mathrm{GL}}_{1} in [JPSS83]. Hence the two functions are equal by the Mellin inversion, following the same argument in the proof of Proposition 3.1. ∎

Lemma 5.3.

For the finite places where ψν\psi_{\nu} and πν\pi_{\nu} are unramified, ψν(x)𝕃πν(x)=𝕃πν(x)\psi_{\nu}(x){\mathbb{L}}_{\pi_{\nu}}(x)={\mathbb{L}}_{\pi_{\nu}}(x).

Proof.

According to [JL23, Lemma 5.3], the πν\pi_{\nu}-basic function 𝕃πν{\mathbb{L}}_{\pi_{\nu}} is supported in 𝔬ν{0}{\mathfrak{o}}_{\nu}\setminus\{0\}. The assertion follows clearly. ∎

Now we are ready to prove Theorem 1.2 by using Theorem 2.1. Recall that SS is the finite set of local places of kk that contains all the Archimedean places and those local places ν\nu where either πν\pi_{\nu} or ψν\psi_{\nu} is ramified. For any ζ𝔸S\zeta\in{\mathbb{A}}^{S}, we take

w():=WS((In1))νSwν()=WS((In1))wS()w(\cdot):={{}^{\circ}W^{S}}\left(\begin{pmatrix}\cdot&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)\prod_{\nu\in S}w_{\nu}(\cdot)={{}^{\circ}W^{S}}\left(\begin{pmatrix}\cdot&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)w_{S}(\cdot)

and

(5.1) ϕ():=ψS(ζ)w()||𝔸1n2.\displaystyle\phi(\cdot):=\psi^{S}(\cdot\zeta)w(\cdot)|\cdot|_{{\mathbb{A}}}^{1-\frac{n}{2}}.

Then the function ϕ(x)\phi(x) belongs to the space 𝒮π(𝔸×){\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}) according to Lemmas 5.1, 5.2 and 5.3. It is clear that the function ϕ(x)\phi(x) is factorizable: ϕ(x)=νϕν(xν)\phi(x)=\prod_{\nu}\phi_{\nu}(x_{\nu}). In order to use Theorem 2.1 in the proof, we calculate its local πν\pi_{\nu}-Fourier transform of ϕν\phi_{\nu} at each place ν\nu. Let R=RζR=R_{\zeta} be as in Theorem 1.2.

For the unramified paces νRS\nu\notin R\cup S, by Lemma 5.2, we obtain that

ϕν(xν)=ψν(xνζν)Wν((xνIn1))|xν|ν1n2=ψν(xνζν)𝕃πν(xν).\phi_{\nu}(x_{\nu})=\psi_{\nu}(x_{\nu}\zeta_{\nu})\cdot{{}^{\circ}W_{\nu}}\left(\begin{pmatrix}x_{\nu}&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)|x_{\nu}|_{\nu}^{1-\frac{n}{2}}=\psi_{\nu}(x_{\nu}\zeta_{\nu}){\mathbb{L}}_{\pi_{\nu}}(x_{\nu}).

By [JL23, Lemma 5.3] (or Lemma 5.3), if 𝕃πν(xν)0{\mathbb{L}}_{\pi_{\nu}}(x_{\nu})\neq 0, then xν𝔬ν{0}x_{\nu}\in{\mathfrak{o}}_{\nu}\setminus\{0\}. Since |ζν|ν1|\zeta_{\nu}|_{\nu}\leq 1 when νR\nu\notin R, we obtain that ψν(xνζν)=1\psi_{\nu}(x_{\nu}\zeta_{\nu})=1 if 𝕃πν(xν)0{\mathbb{L}}_{\pi_{\nu}}(x_{\nu})\neq 0. It follows that ϕν(xν)=𝕃πν(xν)\phi_{\nu}(x_{\nu})={\mathbb{L}}_{\pi_{\nu}}(x_{\nu}). Applying the πν\pi_{\nu}-Fourier transform to the both sides, we obtain that

(5.2) πν,ψν(ϕν)(xν)=πν,ψν(𝕃πν)(xν)=𝕃π~ν(xν)=W~ν((xνIn1))|xν|ν1n2\displaystyle{\mathcal{F}}_{\pi_{\nu},\psi_{\nu}}(\phi_{\nu})(x_{\nu})={\mathcal{F}}_{\pi_{\nu},\psi_{\nu}}({\mathbb{L}}_{\pi_{\nu}})(x_{\nu})={\mathbb{L}}_{\widetilde{\pi}_{\nu}}(x_{\nu})={{}^{\circ}\widetilde{W}_{\nu}}\left(\begin{pmatrix}x_{\nu}&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)|x_{\nu}|_{\nu}^{1-\frac{n}{2}}

according to [JL23, Theorem 3.10]. Note that 𝕃π~ν{\mathbb{L}}_{\widetilde{\pi}_{\nu}} the basic function in the π~ν\widetilde{\pi}_{\nu}-Schwartz space 𝒮π~ν(kν×){\mathcal{S}}_{\widetilde{\pi}_{\nu}}(k_{\nu}^{\times}) and W~ν𝒲(π~ν,ψν1){}^{\circ}\widetilde{W}_{\nu}\in{\mathcal{W}}(\widetilde{\pi}_{\nu},\psi^{-1}_{\nu}), the Whittaker model of π~ν\widetilde{\pi}_{\nu}.

At νS\nu\in S, the function ϕν(xν)\phi_{\nu}(x_{\nu}) takes the following form ϕν(xν)=wν(xν)|xν|ν1n2\phi_{\nu}(x_{\nu})=w_{\nu}(x_{\nu})|x_{\nu}|_{\nu}^{1-\frac{n}{2}}. By Proposition 3.5, we obtain that

(5.3) πν,ψν(ϕν)(xν)=πν,ψν(wν()||1n2)(xν)=w~ν(xν)|xν|ν1n2.\displaystyle{\mathcal{F}}_{\pi_{\nu},\psi_{\nu}}(\phi_{\nu})(x_{\nu})={\mathcal{F}}_{\pi_{\nu},\psi_{\nu}}(w_{\nu}(\cdot)|\cdot|^{1-\frac{n}{2}})(x_{\nu})=\widetilde{w}_{\nu}(x_{\nu})|x_{\nu}|_{\nu}^{1-\frac{n}{2}}.

Finally, at the local places vRv\in R, since RR is disjoint to SS, the function ϕν\phi_{\nu} takes the following form

ϕν(xν)=ψν(xνζν)Wν((xνIn1))|xν|ν1n2\phi_{\nu}(x_{\nu})=\psi_{\nu}(x_{\nu}\zeta_{\nu})\cdot{{}^{\circ}W_{\nu}}\left(\begin{pmatrix}x_{\nu}&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)|x_{\nu}|_{\nu}^{1-\frac{n}{2}}

with |ζν|ν>1|\zeta_{\nu}|_{\nu}>1. Recall from Section 4.1 that αi,i+1\alpha_{i,i+1} be the simple root for the root system Φ\Phi with respect to (Gn,Bn,Tn)({\mathrm{G}}_{n},B_{n},T_{n}). The one-parameter subgroups associated with α1,2\alpha_{1,2} and α2,1\alpha_{2,1} are given by

(5.4) χα1,2(u):=(1u1In2)andχα2,1(u):=(1u1In2).\displaystyle\chi_{\alpha_{1,2}}(u):=\begin{pmatrix}1&u&\\ &1&\\ &&{\mathrm{I}}_{n-2}\end{pmatrix}\quad{\rm and}\quad\chi_{\alpha_{2,1}}(u):=\begin{pmatrix}1&&\\ u&1&\\ &&{\mathrm{I}}_{n-2}\end{pmatrix}.

Then the function ϕν\phi_{\nu} can be written as

ϕν(xν)=Wν((xνIn1)χα1,2(ζν))|xν|ν1n2=Wζν((xνIn1))|xν|ν1n2\phi_{\nu}(x_{\nu})={{}^{\circ}W_{\nu}}\left(\begin{pmatrix}x_{\nu}&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\chi_{\alpha_{1,2}}(\zeta_{\nu})\right)|x_{\nu}|_{\nu}^{1-\frac{n}{2}}=W_{\zeta_{\nu}}\left(\begin{pmatrix}x_{\nu}&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)|x_{\nu}|_{\nu}^{1-\frac{n}{2}}

where Wζν(g):=Wν(gχα1,2(ζν))W_{\zeta_{\nu}}(g):={{}^{\circ}W_{\nu}}(g\chi_{\alpha_{1,2}}(\zeta_{\nu})). It is clear that Wζν𝒲(πν,ψν)W_{\zeta_{\nu}}\in{\mathcal{W}}(\pi_{\nu},\psi_{\nu}). By Proposition 3.3, the πν\pi_{\nu}-Fourier transform of ϕν\phi_{\nu} is give by

πν,ψν(ϕν)(xν)=|xν|ν1n2kνn2(πν(wn.1)Wζν~)((xνyIn21))dy.{\mathcal{F}}_{\pi_{\nu},\psi_{\nu}}(\phi_{\nu})(x_{\nu})=|x_{\nu}|_{\nu}^{1-\frac{n}{2}}\int_{k_{\nu}^{n-2}}\left(\pi_{\nu}(w_{n.1})\widetilde{W_{\zeta_{\nu}}}\right)\left(\begin{pmatrix}x_{\nu}&&\\ y&{\mathrm{I}}_{n-2}&\\ &&1\end{pmatrix}\right)\,\mathrm{d}y.

Since

πν(wn.1)Wζν~((xyIn21))\displaystyle\pi_{\nu}(w_{n.1})\widetilde{W_{\zeta_{\nu}}}\left(\begin{pmatrix}x&&\\ y&{\mathrm{I}}_{n-2}&\\ &&1\end{pmatrix}\right) =Wζν~((xyIn21)wn,1)\displaystyle=\widetilde{W_{\zeta_{\nu}}}\left(\begin{pmatrix}x&&\\ y&{\mathrm{I}}_{n-2}&\\ &&1\end{pmatrix}w_{n,1}\right)
=Wν(wn(xyIn21)twn,1tχα1,2(ζν))\displaystyle={{}^{\circ}W}_{\nu}\left(w_{n}\begin{pmatrix}x&&\\ y&{\mathrm{I}}_{n-2}&\\ &&1\end{pmatrix}^{-t}w_{n,1}^{-t}\chi_{\alpha_{1,2}}(\zeta_{\nu})\right)

where W~(g)=W(wngt)\widetilde{W}(g)=W(w_{n}g^{-t}) with gt:=g1tg^{-t}:={{}^{t}g^{-1}}, we obtain that

πν(wn.1)Wζν~((xyIn21))=Wν~((xyIn21)wn,1χα2,1(ζν)).\pi_{\nu}(w_{n.1})\widetilde{W_{\zeta_{\nu}}}\left(\begin{pmatrix}x&&\\ y&{\mathrm{I}}_{n-2}&\\ &&1\end{pmatrix}\right)=\widetilde{{{}^{\circ}W}_{\nu}}\left(\begin{pmatrix}x&&\\ y&{\mathrm{I}}_{n-2}&\\ &&1\end{pmatrix}w_{n,1}\chi_{\alpha_{2,1}}(-\zeta_{\nu})\right).

Hence the πν\pi_{\nu}-Fourier transform of ϕν\phi_{\nu} can be written as

πν,ψν(ϕν)(xν)\displaystyle{\mathcal{F}}_{\pi_{\nu},\psi_{\nu}}(\phi_{\nu})(x_{\nu}) =|xν|ν1n2kνn2Wν~((xνyIn21)wn,1χα2,1(ζν))dy\displaystyle=|x_{\nu}|_{\nu}^{1-\frac{n}{2}}\int_{k_{\nu}^{n-2}}\widetilde{{{}^{\circ}W}_{\nu}}\left(\begin{pmatrix}x_{\nu}&&\\ y&{\mathrm{I}}_{n-2}&\\ &&1\end{pmatrix}w_{n,1}\chi_{\alpha_{2,1}}(-\zeta_{\nu})\right)\,\mathrm{d}y
=|xν|ν1n2kνn2W~v((xνIn21)(1yIn21)wn,1χα2,1(ζν))dy.\displaystyle=|x_{\nu}|_{\nu}^{1-\frac{n}{2}}\int_{k_{\nu}^{n-2}}\widetilde{{}^{\circ}W}_{v}\left(\begin{pmatrix}x_{\nu}&&\\ &{\mathrm{I}}_{n-2}&\\ &&1\end{pmatrix}\begin{pmatrix}1&&\\ y&{\mathrm{I}}_{n-2}&\\ &&1\end{pmatrix}w_{n,1}\chi_{\alpha_{2,1}}(-\zeta_{\nu})\right)\,\mathrm{d}y.

By the explicit computation of the last integral in [IT13, Section 2.6], we obtain that

(5.5) πν,ψν(ϕν)(xν)=|xν|ν1n2Klν(xν,ζν,W~ν).\displaystyle{\mathcal{F}}_{\pi_{\nu},\psi_{\nu}}(\phi_{\nu})(x_{\nu})=|x_{\nu}|_{\nu}^{1-\frac{n}{2}}{\mathrm{Kl}}_{\nu}(x_{\nu},\zeta_{\nu},{{}^{\circ}\widetilde{W}_{\nu}}).

Thus, by (5.2), (5.3), and (5.5), we obtain a formula for the π\pi-Fourier transform of ϕ\phi, which is the product of the local πν\pi_{\nu}-Fourier transform of ϕν\phi_{\nu} at all local places ν\nu.

Proposition 5.4.

Let ϕ𝒮π(𝔸×)\phi\in{\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}) be the function as defined in (5.1). The π\pi-Fourier transform of ϕ\phi can be explicitly written as

π,ψ(ϕ)(x)=νπν,ψν(ϕν)(xν)=|x|𝔸1n2KlR(x,ζ,W~R)W~SR((xIn1))w~S(x),{\mathcal{F}}_{\pi,\psi}(\phi)(x)=\prod_{\nu}{\mathcal{F}}_{\pi_{\nu},\psi_{\nu}}(\phi_{\nu})(x_{\nu})=|x|_{\mathbb{A}}^{1-\frac{n}{2}}{\mathrm{Kl}}_{R}(x,\zeta,\widetilde{{}^{\circ}W}_{R})\ \widetilde{{}^{\circ}W}^{S\cup R}\left(\begin{pmatrix}x&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)\widetilde{w}_{S}(x),

where the Kloosterman integral is given by KlR(x,ζ,W~R)=νRKlν(xν,ζν,W~ν){\mathrm{Kl}}_{R}(x,\zeta,\widetilde{{}^{\circ}W}_{R})=\prod_{\nu\in R}{\mathrm{Kl}}_{\nu}(x_{\nu},\zeta_{\nu},{{}^{\circ}\widetilde{W}_{\nu}}).

Finally we write the summation on the one side as

αk×ϕ(α)=αk×ψS(αζ)WS((αIn1))wS(α)\sum_{\alpha\in k^{\times}}\phi(\alpha)=\sum_{\alpha\in k^{\times}}\psi^{S}(\alpha\zeta)\ {{}^{\circ}W^{S}}\left(\begin{pmatrix}\alpha&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)w_{S}(\alpha)

and that on the other side as

αk×π,ψ(ϕ)(α)=αk×KlR(α,ζ,W~R)W~SR((αIn1))w~S(α),\sum_{\alpha\in k^{\times}}{\mathcal{F}}_{\pi,\psi}(\phi)(\alpha)=\sum_{\alpha\in k^{\times}}{\mathrm{Kl}}_{R}(\alpha,\zeta,\widetilde{{}^{\circ}W}_{R})\ \widetilde{{}^{\circ}W}^{S\cup R}\left(\begin{pmatrix}\alpha&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)\widetilde{w}_{S}(\alpha),

because |α|𝔸=1|\alpha|_{{\mathbb{A}}}=1 for every αk×\alpha\in k^{\times}. By the π\pi-Poisson summation formula in Theorem 2.1, which is

αk×ϕ(α)=αk×π,ψ(ϕ)(α),\sum_{\alpha\in k^{\times}}\phi(\alpha)=\sum_{\alpha\in k^{\times}}{\mathcal{F}}_{\pi,\psi}(\phi)(\alpha),

we deduce the Voronoi formula in Theorem 1.2:

αk×ψS(αζ)WS((αIn1))wS(α)=αk×KlR(α,ζ,W~R)W~SR((αIn1))w~S(α).\sum_{\alpha\in k^{\times}}\psi^{S}(\alpha\zeta)\ {{}^{\circ}W^{S}}\left(\begin{pmatrix}\alpha&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)w_{S}(\alpha)=\sum_{\alpha\in k^{\times}}{\mathrm{Kl}}_{R}(\alpha,\zeta,\widetilde{{}^{\circ}W}_{R})\ \widetilde{{}^{\circ}W}^{S\cup R}\left(\begin{pmatrix}\alpha&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\right)\widetilde{w}_{S}(\alpha).

We deduce the Voronoi summation formula for GLn{\mathrm{GL}}_{n} from the π\pi-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 ϕν\phi_{\nu} at νR\nu\in R to be more general functions in 𝒲πν(kν×){\mathcal{W}}_{\pi_{\nu}}(k_{\nu}^{\times}). More precisely, if one take ϕν(x)=ψν(xζν)wν(x)|x|ν1n2\phi_{\nu}(x)=\psi_{\nu}(x\zeta_{\nu})\cdot w_{\nu}(x)\cdot|x|_{\nu}^{1-\frac{n}{2}} for νS\nu\in S and ϕν(x)=ψν(xζν)Wν((xIn1)ξ)|x|ν1n2\phi_{\nu}(x)=\psi_{\nu}(x\zeta_{\nu})\cdot W_{\nu}\left(\begin{pmatrix}x&\\ &{\mathrm{I}}_{n-1}\end{pmatrix}\xi\right)\cdot|x|_{\nu}^{1-\frac{n}{2}} for νS\nu\notin S, where wν𝒞c(F×)w_{\nu}\in{\mathcal{C}}_{c}^{\infty}(F^{\times}), Wν𝒲(πν,ψν)W_{\nu}\in{\mathcal{W}}(\pi_{\nu},\psi_{\nu}) and SS, ζ\zeta, ξ\xi are as in [Cor21, Theorem 3.4], then according to Lemmas 5.1, 5.2 and 5.3, the function ϕ:=νϕν𝒮π(𝔸×)\phi:=\otimes_{\nu}\phi_{\nu}\in{\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}). It is clear that the proof of Proposition 5.4 works for such special choices of functions ϕ\phi as well. In particular, we obtain from Proposition 3.3 that at each local place ν\nu, the Fourier transform πν,ψν(ϕν){\mathcal{F}}_{\pi_{\nu},\psi_{\nu}}(\phi_{\nu}) is equal to the function ν(x;ζν,ξν)|x|1n2\mathfrak{H}_{\nu}(x;\zeta_{\nu},\xi_{\nu})|x|^{1-\frac{n}{2}} in [Cor21, proof of Theorem 3.4]. The extended Voronoi formula for GLn{\mathrm{GL}}_{n} in [Cor21, Theorem 3.4] by using the Rankin-Selberg convolution for GLn×GL1{\mathrm{GL}}_{n}\times{\mathrm{GL}}_{1}, can be deduced by the same argument as in our proof of Theorem 1.2 from the π\pi-Poisson summation formula in [JL23, Theorem 4.7]. We omit further details.

6. On the Godement-Jacquet Kernels

For any π𝒜cusp(Gn)\pi\in{\mathcal{A}}_{\mathrm{cusp}}({\mathrm{G}}_{n}), the goal of this section is to define the Godement-Jacquet kernels for Lf(s,π)L_{f}(s,\pi) and their dual kernels, and to prove the π\pi-versions of [Clo22, Theorem 1.1], which can be viewed as the case of n=1n=1 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 LL-function L(s,π)L(s,\pi) as stated in (1.6) is

(6.1) 𝒵(s,ϕ)=𝔸×ϕ(x)|x|𝔸s12d×x\displaystyle{\mathcal{Z}}(s,\phi)=\int_{{\mathbb{A}}^{\times}}\phi(x)|x|_{\mathbb{A}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x

for any ϕ𝒮π(𝔸×)\phi\in{\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}). By [JL23, Theorem 4.6] the zeta integral 𝒵(s,ϕ){\mathcal{Z}}(s,\phi) converges absolutely for Re(s)>n+12{\mathrm{Re}}(s)>\frac{n+1}{2}, admits analytic continuation to an entire function in ss\in{\mathbb{C}}, and satisfies the functional equation

(6.2) 𝒵(s,ϕ)=𝒵(1s,π,ψ(ϕ))\displaystyle{\mathcal{Z}}(s,\phi)={\mathcal{Z}}(1-s,{\mathcal{F}}_{\pi,\psi}(\phi))

where π,ψ{\mathcal{F}}_{\pi,\psi} is the π\pi-Fourier transform as defined in (2.18). As explained in [JL23], this is a reformulation of the Godement-Jacquet theory for the standard LL-functions L(s,π)L(s,\pi).

Consider the fibration through the idele norm map ||𝔸|\cdot|_{\mathbb{A}}:

1𝔸1𝔸×+×11\to{\mathbb{A}}^{1}\to{\mathbb{A}}^{\times}\to{\mathbb{R}}^{\times}_{+}\to 1

where +×={x×:x>0}{\mathbb{R}}^{\times}_{+}=\{x\in{\mathbb{R}}^{\times}\ \colon\ x>0\} and 𝔸1={x𝔸×:|x|𝔸=1}{\mathbb{A}}^{1}=\{x\in{\mathbb{A}}^{\times}\ \colon\ |x|_{\mathbb{A}}=1\}. One can have a suitable Haar measure d×𝔞\,\mathrm{d}^{\times}{\mathfrak{a}} on 𝔸1{\mathbb{A}}^{1} that is compatible with the Haar measures d×x\,\mathrm{d}^{\times}x on 𝔸×{\mathbb{A}}^{\times} and the Haar measure d×t\,\mathrm{d}^{\times}t on +×=𝔸×/𝔸1{\mathbb{R}}^{\times}_{+}={\mathbb{A}}^{\times}/{\mathbb{A}}^{1}. Write 𝔸×=𝔸××𝔸f×{\mathbb{A}}^{\times}={\mathbb{A}}_{\infty}^{\times}\times{\mathbb{A}}_{f}^{\times}, where 𝔸×=ν|k|kν×{\mathbb{A}}_{\infty}^{\times}=\prod_{\nu\in|k|_{\infty}}k_{\nu}^{\times}, and 𝔸f×{\mathbb{A}}_{f}^{\times} is the subset of 𝔸×{\mathbb{A}}^{\times} consisting of elements (xν)𝔸×(x_{\nu})\in{\mathbb{A}}^{\times} with xν=1x_{\nu}=1 for all ν|k|\nu\in|k|_{\infty}.

When Re(s)>n+12{\mathrm{Re}}(s)>\frac{n+1}{2}, the absolutely convergent zeta integral 𝒵(s,ϕ){\mathcal{Z}}(s,\phi) as in (6.1) can be written as

(6.3) 𝔸×ϕ(x)|x|𝔸s12d×x=1𝔸1ϕ(t𝔞)ts12d×𝔞d×t+01𝔸1ϕ(t𝔞)ts12d×𝔞d×t.\displaystyle\int_{{\mathbb{A}}^{\times}}\phi(x)|x|_{\mathbb{A}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x=\int_{1}^{\infty}\int_{{\mathbb{A}}^{1}}\phi(t{\mathfrak{a}})t^{s-\frac{1}{2}}\,\mathrm{d}^{\times}{\mathfrak{a}}\,\mathrm{d}^{\times}t+\int_{0}^{1}\int_{{\mathbb{A}}^{1}}\phi(t{\mathfrak{a}})t^{s-\frac{1}{2}}\,\mathrm{d}^{\times}{\mathfrak{a}}\,\mathrm{d}^{\times}t.
Proposition 6.1.

The first integral on the right-hand side of (6.3):

(6.4) 1𝔸1ϕ(t𝔞)ts12d×𝔞d×t\displaystyle\int_{1}^{\infty}\int_{{\mathbb{A}}^{1}}\phi(t{\mathfrak{a}})t^{s-\frac{1}{2}}\,\mathrm{d}^{\times}{\mathfrak{a}}\,\mathrm{d}^{\times}t

converges absolutely at any ss\in{\mathbb{C}} and is holomorphic as a function in ss\in{\mathbb{C}}, for any ϕ𝒮π(𝔸×)\phi\in{\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}).

Proof.

Let ϕf=νϕν𝒮πf(𝔸f×)\phi_{f}=\otimes_{\nu}\phi_{\nu}\in{\mathcal{S}}_{\pi_{f}}({\mathbb{A}}_{f}^{\times}) be a factorizable π\pi-Schwartz function. Let S=S(π,ψ,ϕf)S=S(\pi,\psi,\phi_{f}) be a finite subset SS of |k|=|k||k|f|k|=|k|_{\infty}\cup|k|_{f} (the set of all local places of kk) that contains |k||k|_{\infty} and such that for any νS\nu\notin S both πν\pi_{\nu} and ψν\psi_{\nu} are unramified and ϕν=𝕃πν\phi_{\nu}={\mathbb{L}}_{\pi_{\nu}}, the basic function in 𝒮πν(kν×){\mathcal{S}}_{\pi_{\nu}}(k_{\nu}^{\times}) as in [JL23, Theorem 3.4]. Write Sf=S|k|f={ν1,ν2,,νκ}S_{f}=S\cap|k|_{f}=\{\nu_{1},\nu_{2},\cdots,\nu_{\kappa}\}. According to [JL23, Proposition 5.5 and Lemma 5.2], there is a positive real number sπs_{\pi}, which depends only on the given π𝒜cusp(Gn)\pi\in{\mathcal{A}}_{\mathrm{cusp}}({\mathrm{G}}_{n}), such that for any real number a0>sπa_{0}>s_{\pi}, the limit lim|x|ν0ϕν(x)|x|νa0=0\lim_{|x|_{\nu}\rightarrow 0}\phi_{\nu}(x)|x|_{\nu}^{a_{0}}=0 holds for every ϕν𝒮πν(kν×)\phi_{\nu}\in{\mathcal{S}}_{\pi_{\nu}}(k_{\nu}^{\times}) and for every ν|k|\nu\in|k|. From the definition of the πν\pi_{\nu}-Schwartz space 𝒮πν(kν×){\mathcal{S}}_{\pi_{\nu}}(k_{\nu}^{\times}) in (2.5) and [JL23, Proposition 3.7], we know that ϕν(x)=0\phi_{\nu}(x)=0 when |x|ν|x|_{\nu} is large enough for all ν<\nu<\infty. Hence for every νSf\nu\in S_{f}, there is a constant Cν>0C_{\nu}>0 such that |ϕν(x)|Cν|x|νa0|\phi_{\nu}(x)|\leq C_{\nu}|x|_{\nu}^{-a_{0}}. By [JL23, Lemma 5.3], there is a positive real number bπ>sπ>0b_{\pi}>s_{\pi}>0, which also depends only on the given π\pi, such that for any b0>bπb_{0}>b_{\pi}, we have that |𝕃πν(x)||x|νb0|{\mathbb{L}}_{\pi_{\nu}}(x)|\leq|x|_{\nu}^{-b_{0}} holds for every νS\nu\notin S. It is clear that for any constant c>bπc>b_{\pi} and constant C1C_{1} with maxνSf{Cν,1}C1\displaystyle{\max_{\nu\in S_{f}}\{C_{\nu},1\}\leq C_{1}}, we must have that the inequality:

(6.5) |ϕf(xf)|C1|xf|𝔸c\displaystyle|\phi_{f}(x_{f})|\leq C_{1}|x_{f}|_{{\mathbb{A}}}^{-c}

holds for every xf𝔸f×x_{f}\in{\mathbb{A}}^{\times}_{f}.

We first estimate the inner integral, which can be written as

(6.6) 𝔸1|ϕ(t𝔞)|d×𝔞=𝔸1/k×γk×|ϕ(γt𝔞)|d×𝔞.\displaystyle\int_{{\mathbb{A}}^{1}}|\phi(t{\mathfrak{a}})|\,\mathrm{d}^{\times}{\mathfrak{a}}=\int_{{\mathbb{A}}^{1}/k^{\times}}\sum_{\gamma\in k^{\times}}|\phi(\gamma t{\mathfrak{a}})|\,\mathrm{d}^{\times}{\mathfrak{a}}.

Fix a ν0|k|\nu_{0}\in|k|_{\infty} and a section +×kν0×𝔸×{\mathbb{R}}_{+}^{\times}\rightarrow k_{\nu_{0}}^{\times}\hookrightarrow{\mathbb{A}}^{\times} of the norm map 𝔸×+×{\mathbb{A}}^{\times}\rightarrow{\mathbb{R}}^{\times}_{+} and view t+×t\in{\mathbb{R}}_{+}^{\times} as the ν0\nu_{0}-component of 𝔸×{\mathbb{A}}^{\times}. Define

:𝔸1(𝔸××𝔬f×)r,𝔞(,log|𝔞|ν,)ν|k|{ν0}\ell:{\mathbb{A}}^{1}\cap({\mathbb{A}}_{\infty}^{\times}\times{\mathfrak{o}}_{f}^{\times})\rightarrow{\mathbb{R}}^{r},\quad{\mathfrak{a}}\mapsto(\cdots,\log|{\mathfrak{a}}|_{\nu},\cdots)_{\nu\in|k|_{\infty}-\{\nu_{0}\}}

where 𝔬f×=ν<𝔬ν×{\mathfrak{o}}_{f}^{\times}=\prod_{\nu<\infty}{\mathfrak{o}}_{\nu}^{\times}, and r:=r1+r21r:=r_{1}+r_{2}-1 with r1r_{1} being the number of real places and r2r_{2} the number of complex ones. Let {ϵi}1ir\{\epsilon_{i}\}_{1\leq i\leq r} be a basis for the group of units in the ring of integers in kk modulo the group of roots of unity in kk, and set

P={i=1rxi(ϵi):0xi<1,1ir}andE0={𝔞1(P):0arg𝔞ν0<2πk}\displaystyle P=\left\{\sum_{i=1}^{r}x_{i}\ell(\epsilon_{i})\ \colon 0\leq x_{i}<1,\;\forall 1\leq i\leq r\right\}\quad{\rm and}\quad E_{0}=\left\{{\mathfrak{a}}\in\ell^{-1}(P)\ \colon 0\leq\arg{\mathfrak{a}}_{\nu_{0}}<\frac{2\pi}{\hbar_{k}}\right\}

where k\hbar_{k} is the class number of kk. We choose representatives 𝔞(1),,𝔞(k){\mathfrak{a}}^{(1)},\cdots,{\mathfrak{a}}^{(\hbar_{k})} of idele classes, and define E:=i=1kE0𝔞(i)\displaystyle{E:=\cup_{i=1}^{\hbar_{k}}E_{0}{\mathfrak{a}}^{(i)}}. Then EE is a fundamental domain of k×\𝔸1k^{\times}\backslash{\mathbb{A}}^{1} according to [Tat67, Theorem 4.3.2], which is compact. Hence we can write (6.6) as

(6.7) 𝔸1|ϕ(t𝔞)|d×𝔞=Eγk×|ϕ(γt𝔞)|d×𝔞.\displaystyle\int_{{\mathbb{A}}^{1}}|\phi(t{\mathfrak{a}})|\,\mathrm{d}^{\times}{\mathfrak{a}}=\int_{E}\sum_{\gamma\in k^{\times}}|\phi(\gamma t{\mathfrak{a}})|\,\mathrm{d}^{\times}{\mathfrak{a}}.

Without loss of generality, we may take ϕ=ϕϕf𝒮π(𝔸×)=𝒮π(𝔸×)𝒮πf(𝔸f×)\phi=\phi_{\infty}\otimes\phi_{f}\in{\mathcal{S}}_{\pi}({\mathbb{A}}^{\times})={\mathcal{S}}_{\pi_{\infty}}({\mathbb{A}}_{\infty}^{\times})\otimes{\mathcal{S}}_{\pi_{f}}({\mathbb{A}}_{f}^{\times}). Write t𝔞=(α,αf)𝔸×=𝔸××𝔸f×t{\mathfrak{a}}=(\alpha_{\infty},\alpha_{f})\in{\mathbb{A}}^{\times}={\mathbb{A}}_{\infty}^{\times}\times{\mathbb{A}}_{f}^{\times}. By (6.5), we have that

|ϕ(γt𝔞)|=|ϕ(γα)ϕf(γαf)|C1|ϕ(γα)||γαf|fc|\phi(\gamma t{\mathfrak{a}})|=|\phi_{\infty}(\gamma\alpha_{\infty})\cdot\phi_{f}(\gamma\alpha_{f})|\leq C_{1}|\phi_{\infty}(\gamma\alpha_{\infty})|\cdot|\gamma\alpha_{f}|_{f}^{-c}

for any constant c>bπc>b_{\pi}. Since 𝔞𝔸1{\mathfrak{a}}\in{\mathbb{A}}^{1}, we must have that

|γαf|fc=|γα|c|γ(α,αf)|𝔸c=|γα|c|γt𝔞|𝔸c=|γα|c|t|𝔸c=|γα|ctc.|\gamma\alpha_{f}|_{f}^{-c}=|\gamma\alpha_{\infty}|_{\infty}^{c}\cdot|\gamma(\alpha_{\infty},\alpha_{f})|_{\mathbb{A}}^{-c}=|\gamma\alpha_{\infty}|_{\infty}^{c}\cdot|\gamma t{\mathfrak{a}}|_{\mathbb{A}}^{-c}=|\gamma\alpha_{\infty}|_{\infty}^{c}\cdot|t|_{\mathbb{A}}^{-c}=|\gamma\alpha_{\infty}|_{\infty}^{c}\cdot t^{-c}.

Hence we obtain

(6.8) |ϕ(γt𝔞)|C1|ϕ(γα)||γα|ctc.\displaystyle|\phi(\gamma t{\mathfrak{a}})|\leq C_{1}|\phi_{\infty}(\gamma\alpha_{\infty})|\cdot|\gamma\alpha_{\infty}|_{\infty}^{c}\cdot t^{-c}.

Since 𝔞{\mathfrak{a}} belongs to a compact set EE, the Archimedean part of 𝔞{\mathfrak{a}} belongs to a compact subset of 𝔸×{\mathbb{A}}_{\infty}^{\times}. Hence there is a constant C2C_{2} such that

γk×|ϕ(γt𝔞)|C2tcγk×|ϕ(γt)||γt|c.\sum_{\gamma\in k^{\times}}|\phi(\gamma t{\mathfrak{a}})|\leq C_{2}\cdot t^{-c}\cdot\sum_{\gamma\in k^{\times}}|\phi_{\infty}(\gamma t)|\cdot|\gamma t|_{\infty}^{c}.

For ϕ𝒮π(𝔸×)\phi_{\infty}\in{\mathcal{S}}_{\pi_{\infty}}({\mathbb{A}}_{\infty}^{\times}), we know from [JL23, Proposition 3.7] that ϕ(x)|x|c\phi_{\infty}(x)|x|_{\infty}^{c} for any constant cc is of rapid decay as |x||x|_{\infty}\to\infty. From the choice of the fundamental domain EE, we must have that αf𝔬f×\alpha_{f}\in{\mathfrak{o}}_{f}^{\times}. Due to [JL23, Lemma 5.3], there are integers e1,,eκe_{1},\cdots,e_{\kappa} such that for γk×\gamma\in k^{\times}, if ϕ(γt𝔞)0\phi(\gamma t{\mathfrak{a}})\neq 0, then γ𝔪:=𝔭1e1𝔭κeκ\gamma\in{\mathfrak{m}}:={\mathfrak{p}}_{1}^{e_{1}}\cdots{\mathfrak{p}}_{\kappa}^{e_{\kappa}}. According to [Neu99, Proposition 5.2], the image of 𝔪{\mathfrak{m}} in 𝔸×{\mathbb{A}}^{\times}_{\infty} is a lattice, and there is a constant C3C_{3} such that the (partial) theta series

γk×|ϕ(γt)||γt|cC3.\sum_{\gamma\in k^{\times}}|\phi_{\infty}(\gamma t)|\cdot|\gamma t|_{\infty}^{c}\leq C_{3}.

Thus we obtain that γk×|ϕ(γt𝔞)|C2C3tc\sum_{\gamma\in k^{\times}}|\phi(\gamma t{\mathfrak{a}})|\leq C_{2}C_{3}t^{-c} and there is a constant C4C_{4} such that 𝔸1|ϕ(t𝔞)|d𝔞C4tc\int_{{\mathbb{A}}^{1}}|\phi(t{\mathfrak{a}})|\,\mathrm{d}{\mathfrak{a}}\leq C_{4}t^{-c}. It follows that the integral 1𝔸1ϕ(t𝔞)ts12d×𝔞d×t\int_{1}^{\infty}\int_{{\mathbb{A}}^{1}}\phi(t{\mathfrak{a}})t^{s-\frac{1}{2}}\,\mathrm{d}^{\times}{\mathfrak{a}}\,\mathrm{d}^{\times}t converges absolutely as long as Re(s)<c+12{\mathrm{Re}}(s)<c+\frac{1}{2} for any c>bπc>b_{\pi}. Since cc is arbitrarily large with c>bπc>b_{\pi}, we obtain that the integral

1𝔸1ϕ(t𝔞)ts12d×𝔞d×t\int_{1}^{\infty}\int_{{\mathbb{A}}^{1}}\phi(t{\mathfrak{a}})t^{s-\frac{1}{2}}\,\mathrm{d}^{\times}{\mathfrak{a}}\,\mathrm{d}^{\times}t

converges absolutely for any ss\in{\mathbb{C}} and hence is holomorphic as a function in ss\in{\mathbb{C}}.

Since a general element in 𝒮π(𝔸×){\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}) is a finite linear combination of the factorizable functions, it is clear that the above statement for the integrals hold for general ϕ𝒮π(𝔸×)\phi\in{\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}). ∎

From the above proof, we also obtain

Corollary 6.2.

For any t+×t\in{\mathbb{R}}^{\times}_{+}, the inner integral

(6.9) 𝔸1ϕ(t𝔞)d×𝔞\displaystyle\int_{{\mathbb{A}}^{1}}\phi(t{\mathfrak{a}})\,\mathrm{d}^{\times}{\mathfrak{a}}

always converges absolutely for any ϕ𝒮π(𝔸×)\phi\in{\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}).

By using the π\pi-Poisson summation formula (Theorem 2.1), we obtain

Proposition 6.3.

For any t+×t\in{\mathbb{R}}^{\times}_{+}, the following identity

𝔸1ϕ(t𝔞)tsd×𝔞=𝔸1π,ψ(ϕ)(t1𝔞)tsd×𝔞\int_{{\mathbb{A}}^{1}}\phi(t{\mathfrak{a}})t^{s}\,\mathrm{d}^{\times}{\mathfrak{a}}=\int_{{\mathbb{A}}^{1}}{\mathcal{F}}_{\pi,\psi}(\phi)(t^{-1}{\mathfrak{a}})t^{s}\,\mathrm{d}^{\times}{\mathfrak{a}}

holds for any ϕ𝒮π(𝔸×)\phi\in{\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}).

Proof.

From (6.9), the integral 𝔸1ϕ(t𝔞)tsd×𝔞\int_{{\mathbb{A}}^{1}}\phi(t{\mathfrak{a}})t^{s}\,\mathrm{d}^{\times}{\mathfrak{a}} converges absolutely for any t+×t\in{\mathbb{R}}^{\times}_{+} and for any ϕ𝒮π(𝔸×)\phi\in{\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}). We write

𝔸1ϕ(t𝔞)tsd×𝔞\displaystyle\int_{{\mathbb{A}}^{1}}\phi(t{\mathfrak{a}})t^{s}\,\mathrm{d}^{\times}{\mathfrak{a}} =αk×αEϕ(t𝔞)tsd×𝔞\displaystyle=\sum_{\alpha\in k^{\times}}\int_{\alpha E}\phi(t{\mathfrak{a}})t^{s}\,\mathrm{d}^{\times}{\mathfrak{a}}
=αk×Eϕ(αt𝔞)tsd×𝔞=E(αk×ϕ(αt𝔞))tsd×𝔞\displaystyle=\sum_{\alpha\in k^{\times}}\int_{E}\phi(\alpha t{\mathfrak{a}})t^{s}\,\mathrm{d}^{\times}{\mathfrak{a}}=\int_{E}\left(\sum_{\alpha\in k^{\times}}\phi(\alpha t{\mathfrak{a}})\right)t^{s}\,\mathrm{d}^{\times}{\mathfrak{a}}

where EE is the fundamental domain of k×k^{\times} in 𝔸1{\mathbb{A}}^{1} as above, which is compact. By Theorem 2.1:

αk×ϕ(αt𝔞)=αk×π,ψ(ϕ)(αt𝔞),\sum_{\alpha\in k^{\times}}\phi(\alpha t{\mathfrak{a}})=\sum_{\alpha\in k^{\times}}{\mathcal{F}}_{\pi,\psi}(\phi)(\frac{\alpha}{t{\mathfrak{a}}}),

we obtain that

𝔸1ϕ(t𝔞)tsd×𝔞=E(αk×π,ψ(ϕ)(αt𝔞))tsd×𝔞=𝔸1π,ψ(ϕ)(t1𝔞)tsd×𝔞,\displaystyle\int_{{\mathbb{A}}^{1}}\phi(t{\mathfrak{a}})t^{s}\,\mathrm{d}^{\times}{\mathfrak{a}}=\int_{E}\left(\sum_{\alpha\in k^{\times}}{\mathcal{F}}_{\pi,\psi}(\phi)(\frac{\alpha}{t{\mathfrak{a}}})\right)t^{s}\,\mathrm{d}^{\times}{\mathfrak{a}}=\int_{{\mathbb{A}}^{1}}{\mathcal{F}}_{\pi,\psi}(\phi)(t^{-1}{\mathfrak{a}})t^{s}\,\mathrm{d}^{\times}{\mathfrak{a}},

where all changes of the order of integrations are verified due to the absolute convergence. ∎

Applying Proposition 6.3 to the second integral on the right-hand side of (6.3), we obtain that for Re(s)>n+12{\mathrm{Re}}(s)>\frac{n+1}{2},

(6.10) 01𝔸1ϕ(t𝔞)ts12d×𝔞d×t\displaystyle\int_{0}^{1}\int_{{\mathbb{A}}^{1}}\phi(t{\mathfrak{a}})t^{s-\frac{1}{2}}\,\mathrm{d}^{\times}{\mathfrak{a}}\,\mathrm{d}^{\times}t =01𝔸1π,ψ(ϕ)(t1𝔞)ts12d×𝔞d×t=1𝔸1π,ψ(ϕ)(t𝔞)t12sd×𝔞d×t.\displaystyle=\int_{0}^{1}\int_{{\mathbb{A}}^{1}}{\mathcal{F}}_{\pi,\psi}(\phi)(t^{-1}{\mathfrak{a}})t^{s-\frac{1}{2}}\,\mathrm{d}^{\times}{\mathfrak{a}}\,\mathrm{d}^{\times}t=\int_{1}^{\infty}\int_{{\mathbb{A}}^{1}}{\mathcal{F}}_{\pi,\psi}(\phi)(t{\mathfrak{a}})t^{\frac{1}{2}-s}\,\mathrm{d}^{\times}{\mathfrak{a}}\,\mathrm{d}^{\times}t.

By Proposition 6.1 and (6.10), we obtain the following

Corollary 6.4.

The second integral in (6.3): 01𝔸1ϕ(t𝔞)ts12d×𝔞d×t\int_{0}^{1}\int_{{\mathbb{A}}^{1}}\phi(t{\mathfrak{a}})t^{s-\frac{1}{2}}\,\mathrm{d}^{\times}{\mathfrak{a}}\,\mathrm{d}^{\times}t converges absolutely for Re(s)>n+12{\mathrm{Re}}(s)>\frac{n+1}{2} and has analytic continuation to an entire function in ss\in{\mathbb{C}}. Moreover, the following identity

01𝔸1ϕ(t𝔞)ts12d×𝔞d×t=1𝔸1π,ψ(ϕ)(t𝔞)t12sd×𝔞d×t\int_{0}^{1}\int_{{\mathbb{A}}^{1}}\phi(t{\mathfrak{a}})t^{s-\frac{1}{2}}\,\mathrm{d}^{\times}{\mathfrak{a}}\,\mathrm{d}^{\times}t=\int_{1}^{\infty}\int_{{\mathbb{A}}^{1}}{\mathcal{F}}_{\pi,\psi}(\phi)(t{\mathfrak{a}})t^{\frac{1}{2}-s}\,\mathrm{d}^{\times}{\mathfrak{a}}\,\mathrm{d}^{\times}t

holds by analytic continuation for ss\in{\mathbb{C}}, where the integral on the right-hand side converges absolutely for all ss\in{\mathbb{C}}.

Set 𝔸>1:={x𝔸×:|x|𝔸>1}{\mathbb{A}}^{>1}:=\{x\in{\mathbb{A}}^{\times}\ \colon\ |x|_{\mathbb{A}}>1\}. By combining (6.3) with (6.10), we obtain that when Re(s)>n+12{\mathrm{Re}}(s)>\frac{n+1}{2}

(6.11) 𝔸×ϕ(x)|x|𝔸s12d×x=𝔸>1ϕ(x)|x|𝔸s12d×x+𝔸>1π,ψ(ϕ)(x)|x|𝔸12sd×x,\displaystyle\int_{{\mathbb{A}}^{\times}}\phi(x)|x|_{\mathbb{A}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x=\int_{{\mathbb{A}}^{>1}}\phi(x)|x|_{\mathbb{A}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x+\int_{{\mathbb{A}}^{>1}}{\mathcal{F}}_{\pi,\psi}(\phi)(x)|x|_{\mathbb{A}}^{\frac{1}{2}-s}\,\mathrm{d}^{\times}x,

which holds for all ss\in{\mathbb{C}} by analytic continuation. From the proof of Proposition 6.1, both integrals on the right-hand side converge absolutely when ss\in{\mathbb{C}} belongs to the vertical strip 12c<Re(s)<12+c\frac{1}{2}-c<{\mathrm{Re}}(s)<\frac{1}{2}+c for any constant cc with c>max{bπ,bπ~}c>\max\{b_{\pi},b_{\widetilde{\pi}}\}. Hence they converge absolutely at any ss\in{\mathbb{C}}.

We are going to calculate the integral 𝔸>1ϕ(x)|x|𝔸s12d×x\int_{{\mathbb{A}}^{>1}}\phi(x)|x|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x in another way. For x=(xν)𝔸×x=(x_{\nu})\in{\mathbb{A}}^{\times}, we write x=xxfx=x_{\infty}\cdot x_{f} with x𝔸×x_{\infty}\in{\mathbb{A}}_{\infty}^{\times} and xf𝔸f×x_{f}\in{\mathbb{A}}_{f}^{\times}. For x𝔸>1x\in{\mathbb{A}}^{>1}, we have that |x|=|x|𝔸|xf|𝔸>1|x|=|x_{\infty}|_{{\mathbb{A}}}\cdot|x_{f}|_{{\mathbb{A}}}>1 and |xf|𝔸>|x|𝔸1|x_{f}|_{{\mathbb{A}}}>|x_{\infty}|_{{\mathbb{A}}}^{-1}. For ϕ=ϕϕf𝒮π(𝔸×)=𝒮π(𝔸×)𝒮πf(𝔸f×)\phi=\phi_{\infty}\otimes\phi_{f}\in{\mathcal{S}}_{\pi}({\mathbb{A}}^{\times})={\mathcal{S}}_{\pi_{\infty}}({\mathbb{A}}_{\infty}^{\times})\otimes{\mathcal{S}}_{\pi_{f}}({\mathbb{A}}_{f}^{\times}), we write

(6.12) 𝔸>1ϕ(x)|x|𝔸s12d×x=𝔸×ϕ(x)|x|𝔸s12d×x𝔸f×>|x|1ϕf(xf)|xf|𝔸s12d×xf,\displaystyle\int_{{\mathbb{A}}^{>1}}\phi(x)|x|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x=\int_{{\mathbb{A}}_{\infty}^{\times}}\phi_{\infty}(x_{\infty})|x_{\infty}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x_{\infty}\int_{{\mathbb{A}}_{f}^{\times}}^{>|x_{\infty}|^{-1}}\phi_{f}(x_{f})|x_{f}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x_{f},

for any ss\in{\mathbb{C}}, where the inner integral is taken over {xf𝔸f×:|xf|𝔸>|x|𝔸1}\{x_{f}\in{\mathbb{A}}_{f}^{\times}\ \colon\ |x_{f}|_{{\mathbb{A}}}>|x_{\infty}|_{{\mathbb{A}}}^{-1}\}. By the Fubini theorem, we know (from the proof of Proposition 6.6) that the inner integral

𝔸f×>|x|𝔸1ϕf(xf)|xf|𝔸s12d×xf\int_{{\mathbb{A}}_{f}^{\times}}^{>|x_{\infty}|_{{\mathbb{A}}}^{-1}}\phi_{f}(x_{f})|x_{f}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x_{f}

converges absolutely for any ss\in{\mathbb{C}} and any x𝔸×x_{\infty}\in{\mathbb{A}}_{\infty}^{\times}.

Definition 6.5 (Godement-Jacquet Kernels).

For any π=ππf𝒜cusp(Gn)\pi=\pi_{\infty}\otimes\pi_{f}\in{\mathcal{A}}_{\mathrm{cusp}}({\mathrm{G}}_{n}), take any ϕf𝒮πf(𝔸×)\phi_{f}\in{\mathcal{S}}_{\pi_{f}}({\mathbb{A}}^{\times}), the Godement-Jacquet kernels associated with π\pi are defined to be

Hπ,s(x,ϕf):=|x|𝔸s12𝔸f×>|x|𝔸1ϕf(xf)|xf|𝔸s12d×xf,H_{\pi,s}(x_{\infty},\phi_{f}):=|x_{\infty}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\int_{{\mathbb{A}}_{f}^{\times}}^{>|x_{\infty}|_{{\mathbb{A}}}^{-1}}\phi_{f}(x_{f})|x_{f}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x_{f},

for x𝔸×x_{\infty}\in{\mathbb{A}}_{\infty}^{\times} and for all ss\in{\mathbb{C}}.

From (6.12), we obtain that

(6.13) 𝔸>1ϕ(x)|x|𝔸s12d×x=𝔸×ϕ(x)Hπ,s(x,ϕf)d×x.\displaystyle\int_{{\mathbb{A}}^{>1}}\phi(x)|x|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x=\int_{{\mathbb{A}}_{\infty}^{\times}}\phi_{\infty}(x_{\infty})H_{\pi,s}(x_{\infty},\phi_{f})\,\mathrm{d}^{\times}x_{\infty}.

In the spirit of [Clo22], to each π𝒜cusp(Gn)\pi\in{\mathcal{A}}_{\mathrm{cusp}}({\mathrm{G}}_{n}), we define the dual kernel of the Godement-Jacquet kernel Hπ,s(x,ϕf)H_{\pi,s}(x_{\infty},\phi_{f}) associated with π\pi to be

(6.14) Kπ,s(x,ϕf):=|x|𝔸s12𝔸f×>|x|𝔸1πf,ψf(ϕf)(xf)|xf|𝔸s12d×xf,\displaystyle K_{\pi,s}(x_{\infty},\phi_{f}):=|x_{\infty}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\int_{{\mathbb{A}}_{f}^{\times}}^{>|x_{\infty}|_{{\mathbb{A}}}^{-1}}{\mathcal{F}}_{\pi_{f},\psi_{f}}(\phi_{f})(x_{f})|x_{f}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x_{f},

for x𝔸×x_{\infty}\in{\mathbb{A}}_{\infty}^{\times} and for all ss\in{\mathbb{C}}.

With a suitable choice of the functions ϕf\phi_{f}, the kernel functions Hπ,s(x,ϕf)H_{\pi,s}(x_{\infty},\phi_{f}) and Kπ,s(x,ϕf)K_{\pi,s}(x_{\infty},\phi_{f}) may have simple expressions. We refer to Proposition 6.11 for details. We establish the distribution property for Hπ,s(x,ϕf)H_{\pi,s}(x_{\infty},\phi_{f}) and Kπ,s(x,ϕf)K_{\pi,s}(x_{\infty},\phi_{f}).

Proposition 6.6.

Set 𝔅:={x𝔸:|x|𝔸=0}{\mathfrak{B}}_{\infty}:=\{x_{\infty}\in{\mathbb{A}}_{\infty}\ \colon\ |x_{\infty}|_{{\mathbb{A}}}=0\} and write 𝔸=𝔸×𝔅{\mathbb{A}}_{\infty}={\mathbb{A}}_{\infty}^{\times}\cup{\mathfrak{B}}_{\infty}. For any ϕf𝒮πf(𝔸f×)\phi_{f}\in{\mathcal{S}}_{\pi_{f}}({\mathbb{A}}_{f}^{\times}) and for any ss\in{\mathbb{C}}, the Godement-Jacquet kernel function Hπ,s(x,ϕf)H_{\pi,s}(x_{\infty},\phi_{f}) and its dual kernel function Kπ,s(x,ϕf)K_{\pi,s}(x_{\infty},\phi_{f}) on 𝔸×{\mathbb{A}}_{\infty}^{\times} enjoy the following properties.

  1. (1)

    Both Hπ,s(x,ϕf)H_{\pi,s}(x_{\infty},\phi_{f}) and Kπ,s(x,ϕf)K_{\pi,s}(x_{\infty},\phi_{f}) vanish to infinity order along 𝔅{\mathfrak{B}}_{\infty}.

  2. (2)

    Both Hπ,s(x,ϕf)H_{\pi,s}(x_{\infty},\phi_{f}) and Kπ,s(x,ϕf)K_{\pi,s}(x_{\infty},\phi_{f}) have unique canonical extension across 𝔅{\mathfrak{B}}_{\infty} to the whole space 𝔸{\mathbb{A}}_{\infty}.

  3. (3)

    Both Hπ,s(x,ϕf)H_{\pi,s}(x_{\infty},\phi_{f}) and Kπ,s(x,ϕf)K_{\pi,s}(x_{\infty},\phi_{f}) are tempered distributions on 𝔸{\mathbb{A}}_{\infty}.

Proof.

By definition, we have that Kπ,s(x,ϕf)=Hπ~,s(x,π,ψ(ϕf))K_{\pi,s}(x_{\infty},\phi_{f})=H_{\widetilde{\pi},s}(x_{\infty},{\mathcal{F}}_{\pi,\psi}(\phi_{f})). It is enough to show that Properties (1), (2), and (3) hold for the kernel function Hπ,s(x,ϕf)H_{\pi,s}(x_{\infty},\phi_{f}). 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 Hπ,s(x,ϕf)H_{\pi,s}(x_{\infty},\phi_{f}) is of polynomial growth as the Eucilidean norm of xx_{\infty} tends to \infty ([Tre67, Theorem 25.4]).

Without loss of generality, we may assume that ϕf=νϕν𝒮πf(𝔸f×)\phi_{f}=\otimes_{\nu}\phi_{\nu}\in{\mathcal{S}}_{\pi_{f}}({\mathbb{A}}_{f}^{\times}) is factorizable. Let T|k|fT\subset|k|_{f} be a finite set such that for νT\nu\notin T, both ψν\psi_{\nu} and πν\pi_{\nu} are unramified and ϕν(x)=𝕃kν(x)\phi_{\nu}(x)={\mathbb{L}}_{k_{\nu}}(x), the basic function in 𝒮πν(kν×){\mathcal{S}}_{\pi_{\nu}}(k_{\nu}^{\times}). According to [JL23, Lemma 5.3], there are integers {eν}νT\{e_{\nu}\}_{\nu\in T} such that the support of ϕf\phi_{f} is contained in (νT(𝔭νeν{0})×νT(𝔬ν{0}))𝔸f×\left(\prod_{\nu\in T}\left({\mathfrak{p}}_{\nu}^{e_{\nu}}\setminus\{0\}\right)\times\prod_{\nu\notin T}\left({\mathfrak{o}}_{\nu}\setminus\{0\}\right)\right)\bigcap{\mathbb{A}}_{f}^{\times}. According to (6.5), for any c>bπc>b_{\pi}, there is a constant C1C_{1} such that |ϕf(xf)|C1|xf|𝔸c|\phi_{f}(x_{f})|\leq C_{1}|x_{f}|_{{\mathbb{A}}}^{-c} for any xf𝔸f×x_{f}\in{\mathbb{A}}_{f}^{\times}. Write

𝔸f×=α=(αν)(ν|k|fϖναν𝔬ν×),{\mathbb{A}}_{f}^{\times}=\bigsqcup_{\alpha=(\alpha_{\nu})}\left(\prod_{\nu\in|k|_{f}}\varpi_{\nu}^{\alpha_{\nu}}{\mathfrak{o}}_{\nu}^{\times}\right),

where ϖν\varpi_{\nu} is the local uniformizer in kνk_{\nu} and α\alpha runs over the algebraic direct sum ν|k|f\oplus_{\nu\in|k|_{f}}{\mathbb{Z}}. Then for xx in the α=(αν)\alpha=(\alpha_{\nu}) component, we have |x|f=ν|k|fqναν|x|_{f}=\prod_{\nu\in|k|_{f}}q_{\nu}^{-\alpha_{\nu}} and the inequality: |x|f|x|𝔸|x|_{f}\leq|x_{\infty}|_{{\mathbb{A}}} is equivalent to the inequality: ν|k|fqναν<|x|𝔸\prod_{\nu\in|k|_{f}}q_{\nu}^{\alpha_{\nu}}<|x_{\infty}|_{{\mathbb{A}}}. We may write a fractional ideal 𝔩{\mathfrak{l}} in kk as ν𝔭ναν\prod_{\nu}{\mathfrak{p}}_{\nu}^{\alpha_{\nu}}. We set 𝔪=𝔪T:=νT𝔭νeν{\mathfrak{m}}={\mathfrak{m}}_{T}:=\prod_{\nu\in T}{\mathfrak{p}}_{\nu}^{e_{\nu}}, which is the fractional ideal depending on the support of ϕf\phi_{f}.

According to the normalization of our Haar measure on 𝔸f×{\mathbb{A}}_{f}^{\times}, we have that

𝔸f×>|x|𝔸1|ϕf(xf)|xf|𝔸s12|d×x\displaystyle\int_{{\mathbb{A}}_{f}^{\times}}^{>|x_{\infty}|_{{\mathbb{A}}}^{-1}}\left|\phi_{f}(x_{f})|x_{f}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\right|\,\mathrm{d}^{\times}x C1𝔸f×>|x|𝔸1|xf|𝔸c+Re(s)12d×xfC1𝔩𝔪,𝔑(𝔩)<|x|𝔸𝔑(𝔩)c+12Re(s),\displaystyle\leq C_{1}\cdot\int_{{\mathbb{A}}_{f}^{\times}}^{>|x_{\infty}|_{{\mathbb{A}}}^{-1}}|x_{f}|_{{\mathbb{A}}}^{-c+{\mathrm{Re}}(s)-\frac{1}{2}}\,\mathrm{d}^{\times}x_{f}\leq C_{1}\cdot\sum_{{\mathfrak{l}}\subset{\mathfrak{m}},{\mathfrak{N}}({\mathfrak{l}})<|x_{\infty}|_{{\mathbb{A}}}}{\mathfrak{N}}({\mathfrak{l}})^{c+\frac{1}{2}-{\mathrm{Re}}(s)},

where the last summation runs over all fractional ideals of kk that are contained in 𝔪{\mathfrak{m}} with absolute norm less than or equal to |x|𝔸|x_{\infty}|_{{\mathbb{A}}}. Write 𝔧=𝔪1𝔩{\mathfrak{j}}={\mathfrak{m}}^{-1}{\mathfrak{l}} and obtain that

𝔩𝔪,𝔑(𝔩)<|x|𝔸𝔑(𝔩)c+12Re(s)=𝔧𝔬,𝔑(𝔧)<|x|𝔸𝔑(𝔪)𝔑(𝔪)c+12Re(s)𝔑(𝔧)c+12Re(s)\sum_{{\mathfrak{l}}\subset{\mathfrak{m}},{\mathfrak{N}}({\mathfrak{l}})<|x_{\infty}|_{{\mathbb{A}}}}{\mathfrak{N}}({\mathfrak{l}})^{c+\frac{1}{2}-{\mathrm{Re}}(s)}=\sum_{{\mathfrak{j}}\subset{\mathfrak{o}},{\mathfrak{N}}({\mathfrak{j}})<\frac{|x_{\infty}|_{{\mathbb{A}}}}{{\mathfrak{N}}({\mathfrak{m}})}}{\mathfrak{N}}({\mathfrak{m}})^{c+\frac{1}{2}-{\mathrm{Re}}(s)}\cdot{\mathfrak{N}}({\mathfrak{j}})^{c+\frac{1}{2}-{\mathrm{Re}}(s)}

Let a(n)a(n) be the number of ideals 𝔧𝔬{\mathfrak{j}}\subset{\mathfrak{o}} with 𝔑(𝔧)=n{\mathfrak{N}}({\mathfrak{j}})=n. According to the Wiener-Ikehara theorem ([MV07, Corollary 8.8]), there is a constant CC^{\prime} such that nxa(n)Cx\sum_{n\leq x}a(n)\leq C^{\prime}x for all x0x\geq 0, and in particular a(n)Cna(n)\leq C^{\prime}n. We obtain that

𝔧𝔬,𝔑(𝔧)<|x|𝔸𝔑(𝔪)𝔑(𝔪)c+12Re(s)𝔑(𝔧)c+12Re(s)\displaystyle\sum_{{\mathfrak{j}}\subset{\mathfrak{o}},{\mathfrak{N}}({\mathfrak{j}})<\frac{|x_{\infty}|_{{\mathbb{A}}}}{{\mathfrak{N}}({\mathfrak{m}})}}{\mathfrak{N}}({\mathfrak{m}})^{c+\frac{1}{2}-{\mathrm{Re}}(s)}\cdot{\mathfrak{N}}({\mathfrak{j}})^{c+\frac{1}{2}-{\mathrm{Re}}(s)} =𝔑(𝔪)c+12Re(s)n|x|𝔸𝔑(𝔪)a(n)nc+12Re(s)\displaystyle={\mathfrak{N}}({\mathfrak{m}})^{c+\frac{1}{2}-{\mathrm{Re}}(s)}\sum_{n\leq\frac{|x_{\infty}|_{{\mathbb{A}}}}{{\mathfrak{N}}({\mathfrak{m}})}}a(n)n^{c+\frac{1}{2}-{\mathrm{Re}}(s)}
𝔑(𝔪)c+12Re(s)Cn|x|𝔸𝔑(𝔪)nc+32Re(s).\displaystyle\leq{\mathfrak{N}}({\mathfrak{m}})^{c+\frac{1}{2}-{\mathrm{Re}}(s)}C^{\prime}\sum_{n\leq\frac{|x_{\infty}|_{{\mathbb{A}}}}{{\mathfrak{N}}({\mathfrak{m}})}}n^{c+\frac{3}{2}-{\mathrm{Re}}(s)}.

For a fixed ss\in{\mathbb{C}}, and any fixed c>max{bπ,Re(s)32}c>\max\{b_{\pi},{\mathrm{Re}}(s)-\frac{3}{2}\}, we have that

n|x|𝔸𝔑(𝔪)nc+32Re(s)1|x|𝔸𝔑(𝔪)+1xc+32Re(s)dx(|x|𝔸𝔑(𝔪)+1)c+52Re(s)c+52Re(s).\sum_{n\leq\frac{|x_{\infty}|_{{\mathbb{A}}}}{{\mathfrak{N}}({\mathfrak{m}})}}n^{c+\frac{3}{2}-{\mathrm{Re}}(s)}\leq\int_{1}^{\frac{|x_{\infty}|_{{\mathbb{A}}}}{{\mathfrak{N}}({\mathfrak{m}})}+1}x^{c+\frac{3}{2}-{\mathrm{Re}}(s)}\,\mathrm{d}x\leq\frac{\left(\frac{|x_{\infty}|_{{\mathbb{A}}}}{{\mathfrak{N}}({\mathfrak{m}})}+1\right)^{c+\frac{5}{2}-{\mathrm{Re}}(s)}}{c+\frac{5}{2}-{\mathrm{Re}}(s)}.

When |x|𝔸𝔑(𝔪)|x_{\infty}|_{{\mathbb{A}}}\geq{\mathfrak{N}}({\mathfrak{m}}), we deduce that

(|x|𝔸𝔑(𝔪)+1)c+52Re(s)c+52Re(s)2c+52Re(s)𝔑(𝔪)Re(s)c52|x|𝔸c+52Re(s)c+52Re(s).\frac{\left(\frac{|x_{\infty}|_{{\mathbb{A}}}}{{\mathfrak{N}}({\mathfrak{m}})}+1\right)^{c+\frac{5}{2}-{\mathrm{Re}}(s)}}{c+\frac{5}{2}-{\mathrm{Re}}(s)}\leq\frac{2^{c+\frac{5}{2}-{\mathrm{Re}}(s)}{\mathfrak{N}}({\mathfrak{m}})^{{\mathrm{Re}}(s)-c-\frac{5}{2}}|x_{\infty}|_{{\mathbb{A}}}^{c+\frac{5}{2}-{\mathrm{Re}}(s)}}{c+\frac{5}{2}-{\mathrm{Re}}(s)}.

Hence we obtain that

(6.15) 𝔸f×>|x|𝔸1|ϕf(xf)|xf|𝔸s12d×xf|2c+52Re(s)C1C(c+52Re(s))𝔑(𝔪)2|x|𝔸c+52Re(s)=C~|x|𝔸c+52Re(s),\displaystyle\int_{{\mathbb{A}}_{f}^{\times}}^{>|x_{\infty}|_{{\mathbb{A}}}^{-1}}\left|\phi_{f}(x_{f})|x_{f}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x_{f}\right|\leq\frac{2^{c+\frac{5}{2}-{\mathrm{Re}}(s)}C_{1}C^{\prime}}{(c+\frac{5}{2}-{\mathrm{Re}}(s)){\mathfrak{N}}({\mathfrak{m}})^{2}}|x_{\infty}|_{{\mathbb{A}}}^{c+\frac{5}{2}-{\mathrm{Re}}(s)}=\widetilde{C}|x_{\infty}|_{{\mathbb{A}}}^{c+\frac{5}{2}-{\mathrm{Re}}(s)},

for any c>max{bπ,Re(s)32}c>\max\{b_{\pi},{\mathrm{Re}}(s)-\frac{3}{2}\} and |x|𝔸>𝔑(𝔪)|x_{\infty}|_{{\mathbb{A}}}>{\mathfrak{N}}({\mathfrak{m}}), where C~=2c+52Re(s)C1C(c+52Re(s))𝔑(𝔪)2\widetilde{C}=\frac{2^{c+\frac{5}{2}-{\mathrm{Re}}(s)}C_{1}C^{\prime}}{(c+\frac{5}{2}-{\mathrm{Re}}(s)){\mathfrak{N}}({\mathfrak{m}})^{2}} is a constant depending on kk, ϕf\phi_{f}, ss, cc, and is independent of |x|𝔸|x_{\infty}|_{{\mathbb{A}}}. Moreover, from the above calculation, we obtain that

(6.16) 𝔸f×>|x|𝔸1|ϕf(xf)|xf|𝔸s12d×xf|=0\displaystyle\int_{{\mathbb{A}}_{f}^{\times}}^{>|x_{\infty}|_{{\mathbb{A}}}^{-1}}\left|\phi_{f}(x_{f})|x_{f}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x_{f}\right|=0

if |x|𝔸𝔑(𝔪)|x_{\infty}|_{{\mathbb{A}}}\leq{\mathfrak{N}}({\mathfrak{m}}). From (6.16), it is clear that the kernel function Hs,π(x,ϕf)H_{s,\pi}(x_{\infty},\phi_{f}) vanishes at some neighborhood for any point x𝔅x_{\infty}\in{\mathfrak{B}}_{\infty}. By [MS04a, Definition 2.4, Lemma 2.8, Definition 2.6], Hs,π(x,ϕf)H_{s,\pi}(x_{\infty},\phi_{f}) vanishes of infinity order at 𝔅{\mathfrak{B}}_{\infty} and has a unique canonical extension across 𝔅{\mathfrak{B}}_{\infty} to the whole 𝔸{\mathbb{A}}_{\infty}, which we still denote by Hπ,s(x,ϕf)H_{\pi,s}(x_{\infty},\phi_{f}). We establish Properties (1) and (2).

For Property (3), because of the estimate in (6.15), the kernel Hπ,s(x,ϕf)H_{\pi,s}(x_{\infty},\phi_{f}) is of polynomial growth as the Eucilidean norm of xx_{\infty} tends to \infty. Hence Hπ,s(x,ϕF)H_{\pi,s}(x_{\infty},\phi_{F}) is tempered as a distribution on 𝔸{\mathbb{A}}_{\infty} according to [Tre67, Theorem 25.4]. ∎

6.2. π\pi_{\infty}-Fourier transform

From (6.14), we obtain for ϕ=ϕϕf𝒮π(𝔸×)\phi=\phi_{\infty}\otimes\phi_{f}\in{\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}) that

(6.17) 𝔸>1π,ψ(ϕ)(x)|x|12sd×x=𝔸×π,ψ(ϕ)(x)Kπ,1s(x,ϕf)d×x,\displaystyle\int_{{\mathbb{A}}^{>1}}{\mathcal{F}}_{\pi,\psi}(\phi)(x)|x|^{\frac{1}{2}-s}\,\mathrm{d}^{\times}x=\int_{{\mathbb{A}}_{\infty}^{\times}}{\mathcal{F}}_{\pi_{\infty},\psi_{\infty}}(\phi_{\infty})(x_{\infty})K_{\pi,1-s}(x_{\infty},\phi_{f})\,\mathrm{d}^{\times}x_{\infty},

which converges absolutely for all ss\in{\mathbb{C}}. The following is the duality relation of the Godement-Jacquet kernels Hπ,s(x,ϕf)H_{\pi,s}(x_{\infty},\phi_{f}) and Kπ,s(x,ϕf)K_{\pi,s}(x,\phi_{f}) via the π\pi_{\infty}-Fourier transform when ss\in{\mathbb{C}} is such that Lf(s,πf)=0L_{f}(s,\pi_{f})=0, which is part of [Clo22, Theorem 1.1] for π𝒜cusp(Gn)\pi\in{\mathcal{A}}_{\mathrm{cusp}}({\mathrm{G}}_{n}).

Proposition 6.7.

For any π𝒜cusp(Gn)\pi\in{\mathcal{A}}_{\mathrm{cusp}}({\mathrm{G}}_{n}), take ϕ=ϕϕf𝒮π(𝔸×)\phi=\phi_{\infty}\otimes\phi_{f}\in{\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}). Then the Godement-Jacquet kernel Hπ,s(x,ϕf)H_{\pi,s}(x,\phi_{f}) associated with π\pi and its dual kernel Kπ,s(x,ϕf)K_{\pi,s}(x,\phi_{f}) enjoy the following identity:

Hπ,s(x,ϕf)=π,ψ(Kπ,1s(,ϕf))(x)=𝔸×kπ,ψ(xy)Kπ,1s(y,ϕf)d×y.H_{\pi,s}(x,\phi_{f})=-{\mathcal{F}}_{\pi_{\infty},\psi_{\infty}}(K_{\pi,1-s}(\cdot,\phi_{f}))(x)=-\int_{{\mathbb{A}}_{\infty}^{\times}}k_{\pi_{\infty},\psi_{\infty}}(xy)K_{\pi,1-s}(y,\phi_{f})\,\mathrm{d}^{\times}y.

as distributions on 𝔸×{\mathbb{A}}_{\infty}^{\times} if ss is a zero of Lf(s,πf)L_{f}(s,\pi_{f}), where kπ,ψk_{\pi_{\infty},\psi_{\infty}} is the π\pi_{\infty}-kernel function as given in (2.20) that gives the π\pi_{\infty}-Fourier transform as a convolution integral operator as in (2.22).

Proof.

For Re(s)>n+12{\mathrm{Re}}(s)>\frac{n+1}{2}, we have

𝒵(s,ϕ)=𝔸×ϕ(x)|x|𝔸s12d×x=ν|k|𝒵ν(s,ϕν).{\mathcal{Z}}(s,\phi)=\int_{{\mathbb{A}}^{\times}}\phi(x)|x|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x=\prod_{\nu\in|k|}{\mathcal{Z}}_{\nu}(s,\phi_{\nu}).

By the reformulation of the Godement-Jacquet local theory in [JL23, Theorem 3.4], we obtain that

𝔸×ϕ(x)|x|𝔸s12d×x=𝒵(s,ϕ)Lf(s,πf)ν|k|𝒵ν(s,ϕν)\int_{{\mathbb{A}}^{\times}}\phi(x)|x|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x={\mathcal{Z}}_{\infty}(s,\phi_{\infty})\cdot L_{f}(s,\pi_{f})\cdot\prod_{\nu\in|k|}{\mathcal{Z}}^{*}_{\nu}(s,\phi_{\nu})

where at almost all finite local places ν\nu with ϕν\phi_{\nu} equal to the basic function 𝕃πν{\mathbb{L}}_{\pi_{\nu}}, we have that 𝒵(s,ϕν)=1{\mathcal{Z}}^{*}(s,\phi_{\nu})=1, for the remaining finite local places ν\nu, where 𝒵ν(s,ϕν):=L(s,πν)1𝒵ν(s,ϕν){\mathcal{Z}}^{*}_{\nu}(s,\phi_{\nu}):=L(s,\pi_{\nu})^{-1}{\mathcal{Z}}_{\nu}(s,\phi_{\nu}) is holomorphic in ss\in{\mathbb{C}}, and 𝒵(s,ϕ):=ν|k|𝒵ν(s,ϕν){\mathcal{Z}}_{\infty}(s,\phi_{\infty}):=\prod_{\nu\in|k|_{\infty}}{\mathcal{Z}}_{\nu}(s,\phi_{\nu}) is holomorphic in ss\in{\mathbb{C}} if ϕ𝒞c(𝔸×)\phi_{\infty}\in{\mathcal{C}}_{c}^{\infty}({\mathbb{A}}_{\infty}^{\times}). Hence we obtain that ν|k|𝒵ν(s,ϕν)\prod_{\nu\in|k|}{\mathcal{Z}}^{*}_{\nu}(s,\phi_{\nu}) is a finite product of holomorphic functions. From (6.11), we have

Z(s,ϕ)=𝔸>1ϕ(x)|x|𝔸s12d×x+𝔸>1π,ψ(ϕ)(x)|x|𝔸12sd×xZ(s,\phi)=\int_{{\mathbb{A}}^{>1}}\phi(x)|x|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x+\int_{{\mathbb{A}}^{>1}}{\mathcal{F}}_{\pi,\psi}(\phi)(x)|x|_{{\mathbb{A}}}^{\frac{1}{2}-s}\,\mathrm{d}^{\times}x

for all ss\in{\mathbb{C}} by analytic continuation. Hence we obtain that if ss\in{\mathbb{C}} is such that Lf(s,πf)=0L_{f}(s,\pi_{f})=0, then we must have that

(6.18) 𝔸>1ϕ(x)|x|𝔸s12d×x=𝔸>1π,ψ(ϕ)(x)|x|𝔸12sd×x.\displaystyle\int_{{\mathbb{A}}^{>1}}\phi(x)|x|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x=-\int_{{\mathbb{A}}^{>1}}{\mathcal{F}}_{\pi,\psi}(\phi)(x)|x|_{{\mathbb{A}}}^{\frac{1}{2}-s}\,\mathrm{d}^{\times}x.

Note that by Proposition 6.1 both integrals converges absolutely for any ss\in{\mathbb{C}}. From (6.17), we have that

𝔸>1π,ψ(ϕ)(x)|x|12sd×x=𝔸×π,ψ(ϕ)(x)Kπ,1s(x,ϕf)d×x\int_{{\mathbb{A}}^{>1}}{\mathcal{F}}_{\pi,\psi}(\phi)(x)|x|^{\frac{1}{2}-s}\,\mathrm{d}^{\times}x=\int_{{\mathbb{A}}_{\infty}^{\times}}{\mathcal{F}}_{\pi_{\infty},\psi_{\infty}}(\phi_{\infty})(x_{\infty})K_{\pi,1-s}(x_{\infty},\phi_{f})\,\mathrm{d}^{\times}x_{\infty}

is absolutely convergent according to (6.15). By [JL22, Theorem 5.1], which is recalled in (2.22), there is a π\pi_{\infty}-kernel function kπ,ψk_{\pi_{\infty},\psi_{\infty}}, such that for any ϕ𝒞c(𝔸×)\phi_{\infty}\in{\mathcal{C}}_{c}^{\infty}({\mathbb{A}}_{\infty}^{\times})

π,ψ(ϕ)(x)=(kπ,ψϕ)(x)=𝔸×kπ,ψ(xy)ϕ(y)d×y.{\mathcal{F}}_{\pi_{\infty},\psi_{\infty}}(\phi_{\infty})(x_{\infty})=(k_{\pi_{\infty},\psi_{\infty}}*\phi_{\infty}^{\vee})(x_{\infty})=\int_{{\mathbb{A}}_{\infty}^{\times}}k_{\pi_{\infty},\psi_{\infty}}(x_{\infty}y_{\infty})\phi_{\infty}(y_{\infty})\,\mathrm{d}^{\times}y_{\infty}.

Since π,ψ(ϕ)𝒮π,ψ(𝔸×){\mathcal{F}}_{\pi_{\infty},\psi_{\infty}}(\phi_{\infty})\in{\mathcal{S}}_{\pi_{\infty},\psi_{\infty}}({\mathbb{A}}^{\times}_{\infty}), By using the Fubini’s theorem and Proposition 6.1 again, we obtain that

𝔸>1π,ψ(ϕ)(x)|x|12sd×x\displaystyle\int_{{\mathbb{A}}^{>1}}{\mathcal{F}}_{\pi,\psi}(\phi)(x)|x|^{\frac{1}{2}-s}\,\mathrm{d}^{\times}x =𝔸×𝔸×kπ,ψ(xy)ϕ(y)d×yKπ,1s(x,ϕf)d×x\displaystyle=\int_{{\mathbb{A}}_{\infty}^{\times}}\int_{{\mathbb{A}}_{\infty}^{\times}}k_{\pi_{\infty},\psi_{\infty}}(x_{\infty}y_{\infty})\phi_{\infty}(y_{\infty})\,\mathrm{d}^{\times}y_{\infty}K_{\pi,1-s}(x_{\infty},\phi_{f})\,\mathrm{d}^{\times}x_{\infty}
=𝔸×ϕ(y)𝔸×kπ,ψ(xy)Kπ,1s(x,ϕf)d×xd×y.\displaystyle=\int_{{\mathbb{A}}_{\infty}^{\times}}\phi_{\infty}(y_{\infty})\int_{{\mathbb{A}}_{\infty}^{\times}}k_{\pi_{\infty},\psi_{\infty}}(x_{\infty}y_{\infty})K_{\pi,1-s}(x_{\infty},\phi_{f})\,\mathrm{d}^{\times}x_{\infty}\,\mathrm{d}^{\times}y_{\infty}.

By definition as in (2.22), we write the π\pi_{\infty}-Fourier transform of the dual kernel Kπ,1s(x,ϕf)K_{\pi,1-s}(x_{\infty},\phi_{f}), viewed as a distribution on 𝔸×{\mathbb{A}}_{\infty}^{\times}, to be

π,ψ(Kπ,1s(,ϕf))(y)=𝔸×kπ,ψ(xy)Kπ,1s(x,ϕf)d×x.{\mathcal{F}}_{\pi_{\infty},\psi_{\infty}}(K_{\pi,1-s}(\cdot,\phi_{f}))(y_{\infty})=\int_{{\mathbb{A}}_{\infty}^{\times}}k_{\pi_{\infty},\psi_{\infty}}(x_{\infty}y_{\infty})K_{\pi,1-s}(x_{\infty},\phi_{f})\,\mathrm{d}^{\times}x_{\infty}.

Hence we obtain that

𝔸>1π,ψ(ϕ)(x)|x|12sd×x=𝔸×ϕ(y)π,ψ(Kπ,1s(,ϕf))(y)d×y.\int_{{\mathbb{A}}^{>1}}{\mathcal{F}}_{\pi,\psi}(\phi)(x)|x|^{\frac{1}{2}-s}\,\mathrm{d}^{\times}x=\int_{{\mathbb{A}}_{\infty}^{\times}}\phi_{\infty}(y_{\infty}){\mathcal{F}}_{\pi_{\infty},\psi_{\infty}}(K_{\pi,1-s}(\cdot,\phi_{f}))(y_{\infty})\,\mathrm{d}^{\times}y_{\infty}.

By combining (6.18) with (6.13), we obtain the following identity as distributions on 𝔸×{\mathbb{A}}_{\infty}^{\times}

𝔸×ϕ(y)π,ψ(Kπ,1s(,ϕf))(y)d×y=𝔸×ϕ(x)Hπ,s(x,ϕf)d×x\int_{{\mathbb{A}}_{\infty}^{\times}}\phi_{\infty}(y_{\infty}){\mathcal{F}}_{\pi_{\infty},\psi_{\infty}}(K_{\pi,1-s}(\cdot,\phi_{f}))(y_{\infty})\,\mathrm{d}^{\times}y_{\infty}=-\int_{{\mathbb{A}}_{\infty}^{\times}}\phi_{\infty}(x_{\infty})H_{\pi,s}(x_{\infty},\phi_{f})\,\mathrm{d}^{\times}x_{\infty}

for all ϕ𝒞c(𝔸×)\phi_{\infty}\in{\mathcal{C}}_{c}^{\infty}({\mathbb{A}}_{\infty}^{\times}). Therefore, as distributions on 𝔸×{\mathbb{A}}_{\infty}^{\times}, we have that

π,ψ(Kπ,1s(,ϕf))(x)=Hπ,s(x,ϕf).{\mathcal{F}}_{\pi_{\infty},\psi_{\infty}}(K_{\pi,1-s}(\cdot,\phi_{f}))(x_{\infty})=-H_{\pi,s}(x_{\infty},\phi_{f}).

For any π=ππf𝒜cusp(Gn)\pi=\pi_{\infty}\otimes\pi_{f}\in{\mathcal{A}}_{\mathrm{cusp}}({\mathrm{G}}_{n}), we write that Lf(s,πf)=ν<L(s,πν)L_{f}(s,\pi_{f})=\prod_{\nu<\infty}L(s,\pi_{\nu}) when Re(s){\mathrm{Re}}(s) is sufficiently positive. By [JL23, Corollary 3.8] and the theory of Mellin transforms, we obtain

Proposition 6.8.

For any ν|k|f\nu\in|k|_{f}, there is a function ϕν𝒮πν(kν×)\phi_{\nu}\in{\mathcal{S}}_{\pi_{\nu}}(k_{\nu}^{\times}) such that

kν×ϕν(x)|x|νs12d×x=L(s,πν)\int_{k_{\nu}^{\times}}\phi_{\nu}(x)|x|_{\nu}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x=L(s,\pi_{\nu})

holds as functions in ss\in{\mathbb{C}} by meromorphic continuation.

For any ϕ𝒮π(𝔸×)\phi_{\infty}\in{\mathcal{S}}_{\pi_{\infty}}({\mathbb{A}}_{\infty}^{\times}), take ϕ=ϕϕf\phi^{\star}=\phi_{\infty}\otimes\phi_{f}^{\star}, where ϕf:=νϕν\phi_{f}^{\star}:=\otimes_{\nu}\phi_{\nu} with ϕν\phi_{\nu} as given in Proposition 6.8 and ϕν=𝕃ν\phi_{\nu}={\mathbb{L}}_{\nu}, the basic function, for almost all ν\nu. It is clear that such a function ϕ\phi belongs to the π\pi-Schwartz space 𝒮π(𝔸×){\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}). As in (6.1), the zeta integral

𝒵(s,ϕ)=𝔸×ϕ(x)|x|𝔸s12d×x{\mathcal{Z}}(s,\phi^{\star})=\int_{{\mathbb{A}}^{\times}}\phi^{\star}(x)|x|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x

converges absolutely when Re(s)>n+12{\mathrm{Re}}(s)>\frac{n+1}{2} and can be written as

(6.19) 𝒵(s,ϕ)=𝒵(s,ϕ)𝒵(s,ϕf)=𝒵(s,ϕ)Lf(s,ϕf),\displaystyle{\mathcal{Z}}(s,\phi^{\star})={\mathcal{Z}}(s,\phi_{\infty})\cdot{\mathcal{Z}}(s,\phi^{\star}_{f})={\mathcal{Z}}(s,\phi_{\infty})\cdot L_{f}(s,\phi_{f}),

where 𝒵(s,ϕf)=ν|k|f𝒵(s,ϕν)=ν|k|fL(s,πν){\mathcal{Z}}(s,\phi^{\star}_{f})=\prod_{\nu\in|k|_{f}}{\mathcal{Z}}(s,\phi_{\nu})=\prod_{\nu\in|k|_{f}}L(s,\pi_{\nu}) when Re(s)>n+12{\mathrm{Re}}(s)>\frac{n+1}{2}. We set

(6.20) Hπ,s(x):=Hπ,s(x,ϕf)andKπ,s(x):=Kπ,s(x,ϕf),\displaystyle H_{\pi,s}(x):=H_{\pi,s}(x,\phi^{\star}_{f})\quad{\rm and}\quad K_{\pi,s}(x):=K_{\pi,s}(x,\phi^{\star}_{f}),

and call Hπ,s(x)H_{\pi,s}(x) the Godement-Jacquet kernel associated with the Euler product Lf(s,πf)L_{f}(s,\pi_{f}), and Kπ,s(x)K_{\pi,s}(x) its dual kernel.

Theorem 6.9.

For any π𝒜cusp(Gn)\pi\in{\mathcal{A}}_{\mathrm{cusp}}({\mathrm{G}}_{n}), take ϕ=ϕϕf𝒮π(𝔸×)\phi^{\star}=\phi_{\infty}\otimes\phi^{\star}_{f}\in{\mathcal{S}}_{\pi}({\mathbb{A}}^{\times}) with ϕf:=ν|k|fϕν\phi_{f}^{\star}:=\otimes_{\nu\in|k|_{f}}\phi_{\nu} where ϕν\phi_{\nu} is as given in Proposition 6.8. Then the Godement-Jacquet kernel Hπ,s(x)H_{\pi,s}(x) associated with the Euler product Lf(s,π)L_{f}(s,\pi) and its dual kernel Kπ,s(x)K_{\pi,s}(x) enjoy the following identity:

(6.21) Hπ,s(x)=π,ψ(Kπ,1s)(x)=𝔸×kπ,ψ(xy)Kπ,1s(y)d×y.\displaystyle H_{\pi,s}(x)=-{\mathcal{F}}_{\pi_{\infty},\psi_{\infty}}(K_{\pi,1-s})(x)=-\int_{{\mathbb{A}}_{\infty}^{\times}}k_{\pi_{\infty},\psi_{\infty}}(xy)K_{\pi,1-s}(y)\,\mathrm{d}^{\times}y.

as distributions on 𝔸×{\mathbb{A}}_{\infty}^{\times} if and only if ss is a zero of Lf(s,πf)L_{f}(s,\pi_{f}).

Proof.

By Proposition 6.7, we only need to consider that if (6.21) holds, then ss\in{\mathbb{C}} is such that Lf(s,πf)=0.L_{f}(s,\pi_{f})=0.

By the choice of ϕ=ϕϕf\phi^{\star}=\phi_{\infty}\otimes\phi^{\star}_{f}, we have from Proposition 6.8 that

𝔸×ϕ(x)|x|s12d×x=𝒵(s,ϕ)Lf(s,πf).\int_{{\mathbb{A}}^{\times}}\phi(x)|x|^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x={\mathcal{Z}}_{\infty}(s,\phi_{\infty})\cdot L_{f}(s,\pi_{f}).

From the proof of Proposition 6.7, we deduce that if (6.21) holds, then we must have that 𝒵(s,ϕ)Lf(s,πf)=0{\mathcal{Z}}_{\infty}(s,\phi_{\infty})\cdot L_{f}(s,\pi_{f})=0 for any ϕ𝒞c(𝔸×)\phi_{\infty}\in{\mathcal{C}}_{c}^{\infty}({\mathbb{A}}_{\infty}^{\times}). It is clear that one is able to choose a particular test function ϕ\phi_{\infty} such that 𝒵(s,ϕ)0{\mathcal{Z}}_{\infty}(s,\phi_{\infty})\neq 0. Hence we get that Lf(s,πf)=0L_{f}(s,\pi_{f})=0. ∎

6.3. Clozel’s theorem for π\pi

In [Clo22, Section 1], Clozel defines the Tate kernel and its dual kernel associated with the Dirichlet series expression of the Dedekind zeta function ζk(s)\zeta_{k}(s) of the ground number field kk 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 π𝒜cusp(Gn)\pi\in{\mathcal{A}}_{\mathrm{cusp}}({\mathrm{G}}_{n}), we define in Definition 6.5 and (6.14) the Godement-Jacquet kernels Hπ,s(x,ϕf)H_{\pi,s}(x,\phi_{f}) and their dual kernels Kπ,s(x,ϕf)K_{\pi,s}(x,\phi_{f}) via the πf\pi_{f}-Fourier transform πf,ψf{\mathcal{F}}_{\pi_{f},\psi_{f}} 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 LL-factors (L(s,πν)(L(s,\pi_{\nu}) at all finite local places ν|k|f\nu\in|k|_{f} (Proposition 6.8), we obtain the π\pi-version of [Clo22, Theorem 1.1] when the kernel functions are related to the LL-function L(s,π)L(s,\pi) with the Euler product expression (Theorem 6.9). In order to obtain the π\pi-version of [Clo22, Theorem 1.1] when the kernel functions are related to the LL-function L(s,π)L(s,\pi) 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 ν|k|f\nu\in|k|_{f}, assume that (πν,Vπν)Πkν(Gn)(\pi_{\nu},V_{\pi_{\nu}})\in\Pi_{k_{\nu}}({\mathrm{G}}_{n}) is generic. Then there exists a function ϕν𝒮πν(kν×)\phi_{\nu}\in{\mathcal{S}}_{\pi_{\nu}}(k_{\nu}^{\times}) such that

𝒵ν(s,ϕν):=kν×ϕν(x)|x|νs12d×x=L(s,πν),{\mathcal{Z}}_{\nu}(s,\phi_{\nu}):=\int_{k_{\nu}^{\times}}\phi_{\nu}(x)|x|_{\nu}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x=L(s,\pi_{\nu}),

the support of ϕν\phi_{\nu} is contained in 𝔬ν{0}{\mathfrak{o}}_{\nu}\setminus\{0\}, and ϕν\phi_{\nu} is invariant under the action of 𝔬ν×{\mathfrak{o}}_{\nu}^{\times}.

Proof.

If n=1n=1, then πν\pi_{\nu} is a quasi-character of kν×k_{\nu}^{\times}. If πν\pi_{\nu} is unramified, it is well-known that one takes ϕν(x)=|x|ν121𝔬ν(x)\phi_{\nu}(x)=|x|_{\nu}^{\frac{1}{2}}1_{{\mathfrak{o}}_{\nu}}(x) with 1𝔬ν1_{{\mathfrak{o}}_{\nu}} the characteristic function of 𝔬ν{\mathfrak{o}}_{\nu}, and has the following identity

𝒵ν(s,ϕν)=kν×ϕν(x)|x|νs12d×x=11πν(ϖν)qs=L(s,πν),{\mathcal{Z}}_{\nu}(s,\phi_{\nu})=\int_{k_{\nu}^{\times}}\phi_{\nu}(x)|x|_{\nu}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x=\frac{1}{1-\pi_{\nu}(\varpi_{\nu})q^{-s}}=L(s,\pi_{\nu}),

which holds for all ss\in{\mathbb{C}} by meromorphic continuation , where ϖν\varpi_{\nu} is the uniformizer of kνk_{\nu}. It is clear that in this case ϕν\phi_{\nu} is supported on kν×𝔬ν=𝔬ν{0}k_{\nu}^{\times}\cap{\mathfrak{o}}_{\nu}={\mathfrak{o}}_{\nu}\setminus\{0\} and is invariant under 𝔬ν×{\mathfrak{o}}_{\nu}^{\times}. If πν\pi_{\nu} is ramified, then we know that L(s,πν)=1L(s,\pi_{\nu})=1. We can take a function ϕν\phi_{\nu} such that ϕν(x)=1\phi_{\nu}(x)=1 if x𝔬νx\in{\mathfrak{o}}_{\nu} and ϕν(x)=0\phi_{\nu}(x)=0 otherwise. Then according to our normalization of the Haar measure, we obtain by an easy computation that

𝒵ν(s,ϕν)=kν×ϕν(x)|x|νs12d×x=1=L(s,πν).{\mathcal{Z}}_{\nu}(s,\phi_{\nu})=\int_{k_{\nu}^{\times}}\phi_{\nu}(x)|x|_{\nu}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x=1=L(s,\pi_{\nu}).

It is clear that in this case, ϕν\phi_{\nu} is supported in 𝔬ν{0}{\mathfrak{o}}_{\nu}\setminus\{0\} and is invariant under the action of 𝔬ν×{\mathfrak{o}}_{\nu}^{\times}.

In the following, we assume that n2n\geq 2. For each non-negative integer mm, we define the congruence subgroup K0(𝔭νm)K_{0}({\mathfrak{p}}_{\nu}^{m}) as in [Hum21] to be

K0(𝔭νm):={x=(xij)Gn(𝔬ν):xn,1,,xn,n1𝔭νm}.K_{0}({\mathfrak{p}}_{\nu}^{m}):=\{x=(x_{ij})\in{\mathrm{G}}_{n}({\mathfrak{o}}_{\nu})\ \colon\ x_{n,1},\cdots,x_{n,n-1}\in{\mathfrak{p}}_{\nu}^{m}\}.

According to the classification of irreducible generic representations and [JPSS81, Theorem 5], there is a minimal positive integer c(πν)c(\pi_{\nu}) for which the vector space

VπνK0(𝔭νc(πν)):={vVπν:πν(x)v=ωπν(xn,n)v,xK0(𝔭νc(πν))}V_{\pi_{\nu}}^{K_{0}({\mathfrak{p}}_{\nu}^{c(\pi_{\nu})})}:=\{v\in V_{\pi_{\nu}}\ \colon\ \pi_{\nu}(x)v=\omega_{\pi_{\nu}}(x_{n,n})v,\ \forall x\in K_{0}({\mathfrak{p}}_{\nu}^{c(\pi_{\nu})})\}

is non-trivial and in fact of dimension one. Choose vVπνK0(𝔭νc(πν))v^{\circ}\in V_{\pi_{\nu}}^{K_{0}({\mathfrak{p}}_{\nu}^{c(\pi_{\nu})})} and v~Vπν~K0(𝔭νc(πν~))\widetilde{v^{\circ}}\in V_{\widetilde{\pi_{\nu}}}^{K_{0}({\mathfrak{p}}_{\nu}^{c(\widetilde{\pi_{\nu}})})}, respectively, such that the matrix coefficient φπν(g):=πν(g)v,v~\varphi_{\pi_{\nu}}(g):=\langle\pi_{\nu}(g)v^{\circ},\widetilde{v^{\circ}}\rangle has value 11 at In{\mathrm{I}}_{n}. Since n2n\geq 2, we may take a Schwartz-Bruhat function fν𝒮(Mn(kν))f_{\nu}\in{\mathcal{S}}({\mathrm{M}}_{n}(k_{\nu})) of the form:

fν(x)={ωπν1(xn,n)vol(K0(𝔭νc(πν)))ifxMn(𝔬ν)withxn,1,,xn,n1𝔭νc(πν)andxn,n𝔬ν×,0otherwise.\displaystyle f_{\nu}(x)=\begin{cases}\frac{\omega_{\pi_{\nu}}^{-1}(x_{n,n})}{\mathrm{vol}(K_{0}({\mathfrak{p}}_{\nu}^{c(\pi_{\nu})}))}&\mathrm{if}\;x\in{\mathrm{M}}_{n}({\mathfrak{o}}_{\nu})\;\mathrm{with}\;x_{n,1},\cdots,x_{n,n-1}\in{\mathfrak{p}}_{\nu}^{c(\pi_{\nu})}\;\mathrm{and}\;x_{n,n}\in{\mathfrak{o}}_{\nu}^{\times},\\ 0&\mathrm{otherwise}.\end{cases}

Then by [Hum21, Theorem 1.2], when Res{\mathrm{Re}}s is sufficiently positive, one has that

Gn(kν)fν(g)φπν(g)|detg|s+n12dg=L(s,πν).\int_{{\mathrm{G}}_{n}(k_{\nu})}f_{\nu}(g)\varphi_{\pi_{\nu}}(g)|\det g|^{s+\frac{n-1}{2}}\,\mathrm{d}g=L(s,\pi_{\nu}).

According to [JL23, Propositon 3.2, Theorem 3.4], the fiber integration as defined in (2.4) yields that

(6.22) ϕπν(x)=|x|νn2Gn(F)xfν(g)φπν(g)dxg,\displaystyle\phi_{\pi_{\nu}}(x)=|x|_{\nu}^{\frac{n}{2}}\int_{{\mathrm{G}}_{n}(F)_{x}}f_{\nu}(g)\varphi_{\pi_{\nu}}(g)\,\mathrm{d}_{x}g,

where Gn(kν)x{\mathrm{G}}_{n}(k_{\nu})_{x} is the fiber at xx of the determinant map as in (2.3), is well defined and when Re(s){\mathrm{Re}}(s) is sufficiently positive, we have that

(6.23) kν×ϕπν(x)|x|νs12d×x=L(s,πν).\displaystyle\int_{k_{\nu}^{\times}}\phi_{\pi_{\nu}}(x)|x|_{\nu}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x=L(s,\pi_{\nu}).

It remains to verify the invariance property for this function ϕν\phi_{\nu}. If x𝔬νx\notin{\mathfrak{o}}_{\nu}, we must have that Gn(kν)xMn(𝔬ν)=.{\mathrm{G}}_{n}(k_{\nu})_{x}\cap{\mathrm{M}}_{n}({\mathfrak{o}}_{\nu})=\emptyset. Thus for any gGn(kν)xg\in{\mathrm{G}}_{n}(k_{\nu})_{x} with x𝔬νx\notin{\mathfrak{o}}_{\nu}, we must have that fν(g)=0f_{\nu}(g)=0. By the fiber integration in (6.22), we have that ϕπν(x)=0\phi_{\pi_{\nu}}(x)=0 when x𝔬νx\notin{\mathfrak{o}}_{\nu}. Moreover, for any u𝔬ν×u\in{\mathfrak{o}}_{\nu}^{\times}, we have

ϕπν(xu)=|xu|νn2Gn(kν)xufν(g)φπν(g)dxg=|x|νn2Gn(kν)xfν(hu)φπν(hu)dxh,\phi_{\pi_{\nu}}(xu)=|xu|_{\nu}^{\frac{n}{2}}\int_{{\mathrm{G}}_{n}(k_{\nu})_{xu}}f_{\nu}(g)\varphi_{\pi_{\nu}}(g)\,\mathrm{d}_{x}g=|x|_{\nu}^{\frac{n}{2}}\int_{{\mathrm{G}}_{n}(k_{\nu})_{x}}f_{\nu}(hu^{*})\varphi_{\pi_{\nu}}(hu^{*})\,\mathrm{d}_{x}h,

where u:=diag(u,1,1,,1).u^{*}:=\mathrm{diag}(u,1,1,\cdots,1). Since uK0(𝔭νc(πν))u^{*}\in K_{0}({\mathfrak{p}}_{\nu}^{c(\pi_{\nu})}), one can see at once that fν(hu)=fν(h)f_{\nu}(hu^{*})=f_{\nu}(h) and φπν(hu)=φπν(h)\varphi_{\pi_{\nu}}(hu^{*})=\varphi_{\pi_{\nu}}(h). Therefore we obtain that ϕπν(xu)=ϕπν(x)\phi_{\pi_{\nu}}(xu)=\phi_{\pi_{\nu}}(x) for any u𝔬ν×u\in{\mathfrak{o}}_{\nu}^{\times}. ∎

Let ϕf=νϕν\phi_{f}^{\circ}=\otimes_{\nu}\phi_{\nu} be the particularly chosen function such that for ramified places we take ϕν\phi_{\nu} as given in Lemma 6.10 since each local component πν\pi_{\nu} of π\pi is irreducible and generic when n2n\geq 2. Denote by

(6.24) π,s(x):=Hπ,s(x,ϕf)\displaystyle{\mathcal{H}}_{\pi,s}(x):=H_{\pi,s}(x,\phi_{f}^{\circ})

the Godement-Jacquet kernel as in (6.20), with ϕf\phi_{f}^{\star} replaced by the particularly chosen ϕf\phi_{f}^{\circ}.

Proposition 6.11.

For any π𝒜cusp(Gn)\pi\in{\mathcal{A}}_{\mathrm{cusp}}({\mathrm{G}}_{n}), the Godement-Jacquet kernel π,s(x){\mathcal{H}}_{\pi,s}(x) as defined in (6.24) enjoys the following expression:

Hπ,s(x)=|x|s12n|x|anns,H_{\pi,s}(x)=|x|^{s-\frac{1}{2}}\sum_{n\leq|x|}a_{n}n^{-s},

as a function in x𝔸×x\in{\mathbb{A}}_{\infty}^{\times} for all ss\in{\mathbb{C}}.

Proof.

Write

(6.25) 𝔸f×=α=(αν)(νϖναν𝔬ν×),\displaystyle{\mathbb{A}}^{\times}_{f}=\bigsqcup_{\alpha=(\alpha_{\nu})}\left(\prod_{\nu}\varpi_{\nu}^{\alpha_{\nu}}{\mathfrak{o}}_{\nu}^{\times}\right),

where ϖν\varpi_{\nu} is the local uniformizer in kνk_{\nu} and α\alpha runs over the algebraic direct sum ν|k|f\oplus_{\nu\in|k|_{f}}{\mathbb{Z}}. Consider the integral

(6.26) 𝔸f×|x|𝔸1ϕf(xf)|xf|𝔸s12d×xf,\displaystyle\int_{{\mathbb{A}}_{f}^{\times}}^{\geq|x_{\infty}|_{{\mathbb{A}}}^{-1}}\phi^{\circ}_{f}(x_{f})|x_{f}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x_{f},

where ϕf=νϕν\phi_{f}^{\circ}=\otimes_{\nu}\phi_{\nu} is as given in (6.24). We know ϕν\phi_{\nu} is supported on the the ring 𝔬ν{\mathfrak{o}}_{\nu} of ν\nu-integers according to Lemma 6.10 and [JL23, Lemma 5.3]. It follows that we may assume αν0\alpha_{\nu}\geq 0 for all ν<\nu<\infty. If xfx_{f} belongs to the α=(αν)\alpha=(\alpha_{\nu})-component of (6.25), then |xf|=νqναν|x_{f}|=\prod_{\nu}q_{\nu}^{-\alpha_{\nu}}, where qνq_{\nu} is the cardinality of the residue field of kνk_{\nu}. The range |xf|𝔸|x|𝔸1|x_{f}|_{{\mathbb{A}}}\geq|x_{\infty}|_{{\mathbb{A}}}^{-1} of the integral in (6.26) is equivalent to the condition that νqναν|x|𝔸\prod_{\nu}q_{\nu}^{\alpha_{\nu}}\leq|x_{\infty}|_{{\mathbb{A}}}. Since the integrand in (6.26) is invariant under ν𝔬ν×\prod_{\nu}{\mathfrak{o}}_{\nu}^{\times}, we obtain that

(6.27) ϕf(xf)|xf|𝔸s12=ϕf((ϖναν))(νqναν)12(νqναν)s\displaystyle\phi^{\circ}_{f}(x_{f})|x_{f}|_{{\mathbb{A}}}^{s-\frac{1}{2}}=\phi^{\circ}_{f}((\varpi_{\nu}^{\alpha_{\nu}}))(\prod_{\nu}q_{\nu}^{\alpha_{\nu}})^{\frac{1}{2}}(\prod_{\nu}q_{\nu}^{-\alpha_{\nu}})^{s}

with (ϖναν)𝔸f×(\varpi_{\nu}^{\alpha_{\nu}})\in{\mathbb{A}}_{f}^{\times}. We may write any fractional ideal 𝔩{\mathfrak{l}} in kk, in a unique way, as 𝔩=ν𝔭νeν{\mathfrak{l}}=\prod_{\nu}{\mathfrak{p}}_{\nu}^{e_{\nu}} with (eν)ν|k|f(e_{\nu})\in\oplus_{\nu\in|k|_{f}}{\mathbb{Z}}, and regard the function ϕf\phi_{f}^{\circ} as

ϕf:𝔩=ν𝔭νeνϕf((ϖνeν))=ν|k|fϕν(ϖνeν)\phi_{f}^{\circ}\ \colon\ {\mathfrak{l}}=\prod_{\nu}{\mathfrak{p}}_{\nu}^{e_{\nu}}\mapsto\phi_{f}^{\circ}((\varpi_{\nu}^{e_{\nu}}))=\prod_{\nu\in|k|_{f}}\phi_{\nu}(\varpi_{\nu}^{e_{\nu}})

Then the function ϕf\phi_{f}^{\circ} is supported on the set of integral ideals and (6.27) can be written as

ϕf(xf)|xf|𝔸s12=ϕf(𝔩)𝔑(𝔩)12𝔑(𝔩)s,\phi^{\circ}_{f}(x_{f})|x_{f}|_{{\mathbb{A}}}^{s-\frac{1}{2}}=\phi_{f}^{\circ}({\mathfrak{l}})\cdot{\mathfrak{N}}({\mathfrak{l}})^{\frac{1}{2}}\cdot{\mathfrak{N}}({\mathfrak{l}})^{-s},

for any fractional ideal 𝔩=ν|k|f𝔭ναν{\mathfrak{l}}=\prod_{\nu\in|k|_{f}}{\mathfrak{p}}_{\nu}^{\alpha_{\nu}}. According to the normalization of the Haar measure, the integral (6.26) is equal to

𝔑(𝔩)|x|𝔸ϕf(𝔩)𝔑(𝔩)12𝔑(𝔩)s=n|x|𝔸(𝔑(𝔩)=nϕf(𝔩))n12ns.\sum_{{\mathfrak{N}}({\mathfrak{l}})\leq|x_{\infty}|_{{\mathbb{A}}}}\phi_{f}^{\circ}({\mathfrak{l}})\cdot{\mathfrak{N}}({\mathfrak{l}})^{\frac{1}{2}}\cdot{\mathfrak{N}}({\mathfrak{l}})^{-s}=\sum_{n\leq|x_{\infty}|_{{\mathbb{A}}}}\left(\sum_{{\mathfrak{N}}({\mathfrak{l}})=n}\phi^{\circ}_{f}({\mathfrak{l}})\right)n^{\frac{1}{2}}n^{-s}.

where the summation runs over all the integral ideals 𝔩{\mathfrak{l}} of kk.

On the other hand, for the particularly given Schwartz function ϕf𝒮πf(𝔸f×)\phi^{\circ}_{f}\in{\mathcal{S}}_{\pi_{f}}({\mathbb{A}}_{f}^{\times}), we have that

Lf(s,πf)\displaystyle L_{f}(s,\pi_{f}) =n=1anns=𝔸f×ϕf(xf)|xf|𝔸s12d×xf=lim|x|𝔸f×|x|𝔸1ϕf(xf)|xf|𝔸s12d×xf\displaystyle=\sum_{n=1}^{\infty}a_{n}n^{-s}=\int_{{\mathbb{A}}^{\times}_{f}}\phi^{\circ}_{f}(x_{f})|x_{f}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x_{f}=\lim_{|x_{\infty}|\rightarrow\infty}\int_{{\mathbb{A}}_{f}^{\times}}^{\geq|x_{\infty}|_{{\mathbb{A}}}^{-1}}\phi^{\circ}_{f}(x_{f})|x_{f}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x_{f}
=lim|x|𝔸𝔑(𝔩)|x|𝔸ϕf(𝔩)𝔑(𝔩)12𝔑(𝔩)s=n=1(𝔑(𝔩)=nϕf(𝔩))n12ns\displaystyle=\lim_{|x_{\infty}|_{{\mathbb{A}}}\rightarrow\infty}\sum_{{\mathfrak{N}}({\mathfrak{l}})\leq|x_{\infty}|_{{\mathbb{A}}}}\phi_{f}^{\circ}({\mathfrak{l}})\cdot{\mathfrak{N}}({\mathfrak{l}})^{\frac{1}{2}}\cdot{\mathfrak{N}}({\mathfrak{l}})^{-s}=\sum_{n=1}^{\infty}\left(\sum_{{\mathfrak{N}}({\mathfrak{l}})=n}\phi^{\circ}_{f}({\mathfrak{l}})\right)n^{\frac{1}{2}}n^{-s}

for Re(s){\mathrm{Re}}(s) is sufficiently positive. By using the uniqueness of the coefficients of the Dirichlet series (see [MV07, Theorem 1.6]), we obtain that an=𝔑(𝔩)=n(ϕf(𝔩))n12a_{n}=\sum_{{\mathfrak{N}}({\mathfrak{l}})=n}\left(\phi^{\circ}_{f}({\mathfrak{l}})\right)n^{\frac{1}{2}}, and hence

𝔸f×|x|𝔸1ϕf(xf)|xf|𝔸s12d×xf=n|x|𝔸anns.\displaystyle\int_{{\mathbb{A}}_{f}^{\times}}^{\geq|x_{\infty}|_{{\mathbb{A}}}^{-1}}\phi^{\circ}_{f}(x_{f})|x_{f}|_{{\mathbb{A}}}^{s-\frac{1}{2}}\,\mathrm{d}^{\times}x_{f}=\sum_{n\leq|x_{\infty}|_{{\mathbb{A}}}}a_{n}n^{-s}.

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) 𝒵(1s,πν,ψν(ϕν))=γ(s,πν,ψν)𝒵(s,ϕν).\displaystyle{\mathcal{Z}}(1-s,{\mathcal{F}}_{\pi_{\nu},\psi_{\nu}}(\phi_{\nu}))=\gamma(s,\pi_{\nu},\psi_{\nu})\cdot{\mathcal{Z}}(s,\phi_{\nu}).

By Proposition 6.8 and [JL23, Theorem 3.4], we have that

γ(s,πν,ψν)L(s,πν)=ϵ(s,πν,ψν)L(1s,π~ν).\gamma(s,\pi_{\nu},\psi_{\nu})\cdot L(s,\pi_{\nu})=\epsilon(s,\pi_{\nu},\psi_{\nu})\cdot L(1-s,\widetilde{\pi}_{\nu}).

Hence for Re(s){\mathrm{Re}}(s) sufficiently negative, we obtain that

(6.29) 𝒵f(1s,πf,ψf(ϕf))\displaystyle{\mathcal{Z}}_{f}(1-s,{\mathcal{F}}_{\pi_{f},\psi_{f}}(\phi_{f})) =𝔸f×πf,ψf(ϕf)(xf)|xf|𝔸12sd×xf=(ν<ϵ(s,πν,ψν))Lf(1s,π~f).\displaystyle=\int_{{\mathbb{A}}^{\times}_{f}}{\mathcal{F}}_{\pi_{f},\psi_{f}}(\phi_{f})(x_{f})|x_{f}|_{{\mathbb{A}}}^{\frac{1}{2}-s}\,\mathrm{d}^{\times}x_{f}=\left(\prod_{\nu<\infty}\epsilon(s,\pi_{\nu},\psi_{\nu})\right)\cdot L_{f}(1-s,\widetilde{\pi}_{f}).

If we write

(6.30) (ν<ϵ(1s,πν,ψν))L(s,π~)=n=1anns,\displaystyle\left(\prod_{\nu<\infty}\epsilon(1-s,\pi_{\nu},\psi_{\nu})\right)L(s,\widetilde{\pi})=\sum_{n=1}^{\infty}a_{n}^{*}n^{-s},

with ana_{n}^{*}\in{\mathbb{C}}, 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 Hπ,s(x,ϕf)H_{\pi,s}(x,\phi^{\circ}_{f}).

Proposition 6.12.

With ϕf\phi_{f}^{\circ} as in Proposition 6.11, the dual kernel of the Godement-Jacquet kernel π,s(x){\mathcal{H}}_{\pi,s}(x) as in (6.24) is given by

𝒦π,s(x):=Kπ,s(x,ϕf)=|x|s12n|x|anns,{\mathcal{K}}_{\pi,s}(x):=K_{\pi,s}(x,\phi_{f}^{\circ})=|x|^{s-\frac{1}{2}}\sum_{n\leq|x|}a_{n}^{*}n^{-s},

where {an}\{a_{n}^{*}\} is defined via (6.30).

Proof.

We first claim that each local component of πf,ψf(ϕf){\mathcal{F}}_{\pi_{f},\psi_{f}}(\phi_{f}^{\circ}) is invariant under 𝔬ν×{\mathfrak{o}}_{\nu}^{\times}. In fact, if n=1n=1, the claim is clear because the classical Fourier transform of an 𝔬ν×{\mathfrak{o}}_{\nu}^{\times}-invariant functions is still 𝔬ν×{\mathfrak{o}}_{\nu}^{\times}-invariant by changing variables. If n2n\geq 2, at ramified places, since ϕν\phi_{\nu} is as given by the fiber integration of fνf_{\nu} and φν\varphi_{\nu} as in Lemma 6.10, we know from [JL23, Equation (3.17)] that

πν,ψν(ϕν)(x)=|x|νn2GLn(kν)xψν(fν)(g)φπν(g1)dg,{\mathcal{F}}_{\pi_{\nu},\psi_{\nu}}(\phi_{\nu})(x)=|x|_{\nu}^{\frac{n}{2}}\int_{{\mathrm{GL}}_{n}(k_{\nu})_{x}}{\mathcal{F}}_{\psi_{\nu}}(f_{\nu})(g)\varphi_{\pi_{\nu}}(g^{-1})\,\mathrm{d}g,

where ψν{\mathcal{F}}_{\psi_{\nu}} is the classical Fourier transform given by (2.7). If u:=diag(u,1,,1)u^{*}:=\mathrm{diag}(u,1,\cdots,1) for any u𝔬ν×u\in{\mathfrak{o}}_{\nu}^{\times}, then we already know φπν((gu)1)=φπν(g1)\varphi_{\pi_{\nu}}\left((gu^{*})^{-1}\right)=\varphi_{\pi_{\nu}}(g^{-1}). Since

ψν(fν)(gu)=Mn(kν)ψν(tr(xuy))fν(y)d+y=Mn(kν)ψν(tr(xy))fν((u)1y)d+y,{\mathcal{F}}_{\psi_{\nu}}(f_{\nu})(gu^{*})=\int_{\mathrm{M}_{n}(k_{\nu})}\psi_{\nu}({\mathrm{tr}}(xu^{*}y))f_{\nu}(y)\,\mathrm{d}^{+}y=\int_{\mathrm{M}_{n}(k_{\nu})}\psi_{\nu}({\mathrm{tr}}(xy))f_{\nu}((u^{*})^{-1}y)\,\mathrm{d}^{+}y,

and by the definition of fνf_{\nu}, we see that fν((u)1y)=fν(y)f_{\nu}((u^{*})^{-1}y)=f_{\nu}(y), we know πν,ψν(ϕν){\mathcal{F}}_{\pi_{\nu},\psi_{\nu}}(\phi_{\nu}) is invariant under 𝔬ν×{\mathfrak{o}}_{\nu}^{\times}. At the remaining unramified places where ϕν=𝕃πν\phi_{\nu}={\mathbb{L}}_{\pi_{\nu}}, we know πν,ψν(𝕃πν)=𝕃πν~{\mathcal{F}}_{\pi_{\nu},\psi_{\nu}}({\mathbb{L}}_{\pi_{\nu}})={\mathbb{L}}_{\widetilde{\pi_{\nu}}} and by [JL23, Lemma 5.3] we know πν,ψν(ϕν){\mathcal{F}}_{\pi_{\nu},\psi_{\nu}}(\phi_{\nu}) is invariant under 𝔬ν×{\mathfrak{o}}_{\nu}^{\times}. Let SfS_{f} be as in Proposition 6.1 for ϕf\phi_{f}^{\circ}. Then there are integers a1,,aκa_{1},\cdots,a_{\kappa} such that the support of πf,ψf{\mathcal{F}}_{\pi_{f},\psi_{f}} is contained in

(6.31) (νSf(𝔭νaν{0})×νSf(𝔬ν{0}))𝔸f×.\displaystyle\left(\prod_{\nu\in S_{f}}({\mathfrak{p}}_{\nu}^{a_{\nu}}\setminus\{0\})\times\prod_{\nu\notin S_{f}}({\mathfrak{o}}_{\nu}\setminus\{0\})\right)\cap{\mathbb{A}}_{f}^{\times}.

Write 𝔸f×=α=(αν)(ν|k|fϖναν𝔬ν×){\mathbb{A}}_{f}^{\times}=\bigsqcup_{\alpha=(\alpha_{\nu})}\left(\prod_{\nu\in|k|_{f}}\varpi_{\nu}^{\alpha_{\nu}}{\mathfrak{o}}_{\nu}^{\times}\right). It is clear that πf,ψf(ϕf){\mathcal{F}}_{\pi_{f},\psi_{f}}(\phi_{f}^{\circ}) is constant on each α\alpha-component and supported on the α=(αν)ν\alpha=(\alpha_{\nu})_{\nu}-component with ανaν\alpha_{\nu}\geq a_{\nu} for νSf\nu\in S_{f} and αν0\alpha_{\nu}\geq 0 for νSf\nu\notin S_{f}. We may write any fraction ideal 𝔩{\mathfrak{l}} in kk in a unique way as 𝔩=ν𝔭νeν{\mathfrak{l}}=\prod_{\nu}{\mathfrak{p}}_{\nu}^{e_{\nu}} and regard the function πf,ψf(ϕf){\mathcal{F}}_{\pi_{f},\psi_{f}}(\phi_{f}^{\circ}) as a function on the set of fractional ideals sending 𝔩{\mathfrak{l}} to ν|k|fπν,ψν(ϕν)(ϖνeν).\prod_{\nu\in|k|_{f}}{\mathcal{F}}_{\pi_{\nu},\psi_{\nu}}(\phi_{\nu})(\varpi_{\nu}^{e_{\nu}}). Then we obtain that

πf,ψf(ϕf)(xf)|xf|𝔸s12=ϕf(𝔩)𝔑(𝔩)12𝔑(𝔩)s{\mathcal{F}}_{\pi_{f},\psi_{f}}(\phi^{\circ}_{f})(x_{f})|x_{f}|_{{\mathbb{A}}}^{s-\frac{1}{2}}=\phi_{f}^{\circ}({\mathfrak{l}})\cdot{\mathfrak{N}}({\mathfrak{l}})^{\frac{1}{2}}\cdot{\mathfrak{N}}({\mathfrak{l}})^{-s}

for xx in the α=(αν)ν\alpha=(\alpha_{\nu})_{\nu}-component, where 𝔩=ν|k|f𝔭ναν{\mathfrak{l}}=\prod_{\nu\in|k|_{f}}{\mathfrak{p}}_{\nu}^{\alpha_{\nu}}. Write 𝔫=ν|k|f𝔭aν{\mathfrak{n}}=\prod_{\nu\in|k|_{f}}{\mathfrak{p}}^{a_{\nu}}, where for νSf\nu\in S_{f} aνa_{\nu}’s are defined from (6.31) and for νSf\nu\notin S_{f} we define aν=0a_{\nu}=0. Then by the same argument, we obtain that

an=𝔩𝔫,𝒩(𝔩)=nπf,ψf(ϕf)(𝔩)𝔑(𝔩)12𝔑(𝔩)s,a_{n}^{*}=\sum_{{\mathfrak{l}}\subset{\mathfrak{n}},{\mathcal{N}}({\mathfrak{l}})=n}{\mathcal{F}}_{\pi_{f},\psi_{f}}(\phi_{f}^{\circ})({\mathfrak{l}}){\mathfrak{N}}({\mathfrak{l}})^{\frac{1}{2}}{\mathfrak{N}}({\mathfrak{l}})^{-s},

where the summation runs over all fractional ideal 𝔩{\mathfrak{l}} of kk that are contained in 𝔫{\mathfrak{n}} with norm nn and 𝒦π,s(x)=|x|s12n|x|anns.{\mathcal{K}}_{\pi,s}(x)=|x|^{s-\frac{1}{2}}\sum_{n\leq|x|}a_{n}^{*}n^{-s}.

Therefore we obtain a π\pi-version of [Clo22, Theorem 1.1] when the kernel functions π,s{\mathcal{H}}_{\pi,s} and 𝒦π,s{\mathcal{K}}_{\pi,s} are given in terms of the LL-function Lf(s,πf)L_{f}(s,\pi_{f}) with its Dirichlet series expression.

Theorem 6.13.

For any π𝒜cusp(Gn)\pi\in{\mathcal{A}}_{\mathrm{cusp}}({\mathrm{G}}_{n}), if the Godement-Jacquet kernel π,s{\mathcal{H}}_{\pi,s} and its dual 𝒦π,s{\mathcal{K}}_{\pi,s} are defined as in Propositions 6.11 and 6.12, respectively, then

π,s(x)=π,ψ(𝒦π,1s)(x)=𝔸×kπ,ψ(xy)𝒦π,1s(y)d×y{\mathcal{H}}_{\pi,s}(x)=-{\mathcal{F}}_{\pi_{\infty},\psi_{\infty}}({\mathcal{K}}_{\pi,1-s})(x)=-\int_{{\mathbb{A}}_{\infty}^{\times}}k_{\pi_{\infty},\psi_{\infty}}(xy){\mathcal{K}}_{\pi,1-s}(y)\,\mathrm{d}^{\times}y

as distributions on 𝔸×{\mathbb{A}}_{\infty}^{\times} if and only if ss is a zero of Lf(s,πf)L_{f}(s,\pi_{f}). Any unexplained notation is the same as in Theorem 6.9.

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 GL(n){\rm GL}(n) over a pp-adic field. In Automorphic representations, LL-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. γ\gamma-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 ll-functions. with an appendix by l. clozel and peter sarnak, 2022.
  • [Cog14] J. Cogdell. Bessel functions for GL2GL_{2}. 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 GLn{\rm GL}_{n}: 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 GL(n){\rm GL}(n). Int. Math. Res. Not., pages Art. ID 86295, 25, 2006.
  • [GL08] D. Goldfeld and X. Li. The Voronoi formula for GL(n,){\rm GL}(n,\mathbb{R}). 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 GL(n){\rm GL}(n). Amer. J. Math., 135(1):65–101, 2013.
  • [Jac79] H. Jacquet. Principal LL-functions of the linear group. In Automorphic forms, representations and LL-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 LL-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 GL(2){\rm GL}(2). Lecture Notes in Mathematics, Vol. 114. Springer-Verlag, Berlin-New York, 1970.
  • [JL22] D. Jiang and Z. Luo. Certain Fourier operators on GL1{\rm GL}_{1} 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 GL1\mathrm{GL}_{1}. 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, LL-functions, and Voronoi summation for GL(3){\rm GL}(3). Ann. of Math. (2), 164(2):423–488, 2006.
  • [MS11] S. Miller and W. Schmid. A general Voronoi summation formula for GL(n,)GL(n,\mathbb{Z}). 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 LL-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 LL and γ\gamma factors for generic representations of GSp(4,k)×GL(2,k){\rm GSp}(4,\,k)\times{\rm GL}(2,\,k) over a local non-Archimedean field kk. Duke Math. J., 51(2):355–394, 1984.
  • [Sou01] C. Soulé. On zeroes of automorphic LL-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.