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

On the Resolution of Reductive Monoids and Multiplicativity of γ\gamma-Factors

Freydoon Shahidi and William Sokurski
Abstract.

In this article, we give a proof of multiplicativity for γ\gamma-factors, an equality of parabolically induced and inducing factors, in the context of the Braverman-Kazhdan/Ngo program, under the assumption of commutativity of the corresponding Fourier transforms and a certain generalized Harish-Chandra transform. Within our proof, we define a suitable space of Schwartz functions which we prove to contain the basic function. We also discuss the resolution of singularities and their rationality for reductive monoids, which are among the basic objects in the program.

2010 Mathematics Subject Classification:
11F70, 11F66, and 14E15

Introduction

Every theory of LL-functions must satisfy the axiom of multiplicativity/inductivity, which simply requires that γ\gamma-factors for induced representations are equal to those of the inducing representations. This axiom is a theorem for Artin LL-functions and those obtained from the Langlands-Shahidi method [Sha10], and is a main tool in computing γ\gamma-factors, root numbers, and LL-functions. On the other hand, its proof in the cases obtained from Rankin-Selberg methods are quite involved and complicated. It is also central in proving equality of these factors when they are defined by different methods and in establishing the local Langlands correspondence (LLC) [Sha12, Sha17, HT01, Hen00, GT11, CST17] . Its importance as a technical tool in proving certain cases of functoriality [CKPSS04, Kim03, KS02] is now well established.

In this paper we will provide a proof of multiplicativity for γ\gamma-factors defined by the method of Braverman-Kazhdan/Ngo [BK02, BK10, BNS16, Ngô20] and L. Lafforgue [Laf14] in general under the assumption that the ρ\rho-Fourier transforms on the group GG and the inducing Levi subgroup LL commute with the ρ\rho-Harish-Chandra transform, a generalized Satake transform sending Cc(G(k))Cc(L(k))C_{c}^{\infty}(G(k))\to C_{c}^{\infty}(L(k)), where ρ\rho is a finite dimensional representation of the LL-group of GG by means of which the γ\gamma-factors are defined.

Within our proof, we define a space 𝒮ρ(G)\mathcal{S}^{\rho}(G) of ρ\rho-Schwartz functions for every ρ\rho as

(0.1) 𝒮ρ(G):=Cc(G(k))+Jρ(Cc(G(k)))C(G(k)).\mathcal{S}^{\rho}(G):=C_{c}^{\infty}(G(k))+J^{\rho}(C^{\infty}_{c}(G(k)))\subset C^{\infty}(G(k)).

This definition is crucial since the ρ\rho-Schwartz functions defined in this way will be uniformly bi-KK-finite (see equation (5.16) and Lemma 5.55.5), making the descent to the inducing level possible, an important step in the proof of multiplicativity. While the γ\gamma-factor can be defined as the kernel of the Fourier transform, it is the full functional equation that allows our descent to the inducing level in a transparent fashion, using our definition of ρ\rho-Schwartz functions.

In [BK10], Braverman and Kazhdan defined their Schwartz space as a ”saturation” of ours. But our Schwartz space, which is denoted by VρV_{\rho} in [BK10], covers a significant part of theirs and in particular, contains the ρ\rho-basic function as we prove in Proposition 5.35.3 This is done using the extended Satake transform to almost compact functions [Li17] and the fact that it commutes with the Fourier transform induced from tori which is now defined in general, cf. Section 66 and in partiular diagram (6.8)(6.8).

The commutativity assumption allows us to extend the ρ\rho-Harish-Chandra transform to 𝒮ρ(G)\mathcal{S}^{\rho}(G), commuting with JρJ^{\rho} and JρLJ^{\rho_{L}}, respectively, where ρL\rho_{L} is the restriction of ρ\rho to the LL-group of LL. This construction of 𝒮ρ(G)\mathcal{S}^{\rho}(G) agrees with that of Braverman-Kazhdan in the case of doubling method [BK02, GPSR87, Li18, LR05, PSR86, Sha18, JLZ20, GL20], since GG being the interior of the defining monoid embeds as a unique open orbit into the Braverman-Kazhdan space (cf. [Li18]). Our proof is a generalization of Godement-Jacquet for GLnGL_{n}, Theorem 3.4 of [GJ72].

One expects 𝒮ρ(G)Cc(Mρ(k))\mathcal{S}^{\rho}(G)\subset C_{c}(M^{\rho}(k)), where the latter is defined as the space of functions of compact support on Mρ(k)M^{\rho}(k), the monoid attached to ρ\rho whose restriction to G(k)G(k) are smooth (locally constant) and sending Cc(G(k))C_{c}^{\infty}(G(k)) to itself inside Cc(Mρ(k))C_{c}(M^{\rho}(k)). The group GG being smooth as a variety, the singularities of the monoid MρM^{\rho} are outside of GG. Renner’s construction (Section 2) realizes MρLMρM^{\rho_{L}}\subset M^{\rho} as a closed subvariety for any Levi subgroup LL that contains TT, where TT is a fixed maximal torus of GG, used in the construction of MρM^{\rho}.

The ρ\rho-Harish-Chandra transform cannot be defined on Cc(Mρ(k))C_{c}(M^{\rho}(k)) since the modulus character δP\delta_{P} may vanish outside L(k)MρL(k)L(k)\subset M^{\rho_{L}}(k), but it can be defined on the image of 𝒮ρ(G)\mathcal{S}^{\rho}(G) inside Cc(Mρ(k))C(G(k))C_{c}(M^{\rho}(k))\cap C^{\infty}(G(k)), that lands in the image of 𝒮ρL(L)\mathcal{S}^{\rho_{L}}(L) in Cc(MρL(k))C_{c}(M^{\rho_{L}}(k)) by the above discussion.

Our commutativity axiom, which implies multiplicativity and multiplicativity itself give rise to an inductive scheme that allows for a definition of Fourier transform JρJ^{\rho} by building from the case of conjugacy classes of Levi subgroups LL of GG. In fact, Theorem 5.3 gives the γ\gamma factors γ(s,π,ρ,ψ)\gamma(s,\pi,\rho,\psi), π\pi an irreducible constituent of Ind(σ)\mathrm{Ind}(\sigma), equal to the inducing γ\gamma-factor γ(s,σ,ρL,ψ)\gamma(s,\sigma,\rho_{L},\psi), which in turn is defined through convolution by JρLJ^{\rho_{L}}. For example, for GL2GL_{2}, the Levi subgroups consist of split tori for which a canonical Fourier transform exists (c.f. [Ngô20]; see (6.2) here) and GL2GL_{2} itself, which is equivalent to understanding supercuspidal γ\gamma-factors. We refer to section 5.4 for a more detailed discussion of this inductive construction.

In the case of GL2GL_{2}, Laurent Lafforgue [Laf14] has defined a candidate distrubution which is shown formally to commute with the Harish-Chandra transform and evidence exists that it may give the correct supercuspial factors as observed by Jacquet, but it is still unknown if this is the right distribution. Work in this direction for tamely ramified representations is being pursued by the second author.

Although our definition of the space 𝒮ρ(G)\mathcal{S}^{\rho}(G) depends on the knowledge of how JρJ^{\rho} acts on Cc(G(k))C_{c}^{\infty}(G(k)), this seems to be the most efficient way of defining 𝒮ρ(G)\mathcal{S}^{\rho}(G) at present and sufficient for our purposes as a working definition, allowing us to begin making some initial steps toward understanding the general theory, and as observed earlier after equation (0.1), essential in proving the uniform KK-finiteness of ρ\rho-Schwartz functions.

One hopes that the geometry of MρM^{\rho} will provide some insight into what this Fourier transform ought to be. In fact, the geometric techniques used to study the basic functions on reductive monoids via arc spaces in the function field setting [BNS16] tells us that the nature of the singularities of the monoid very much controls the asymptotics of the basic function. Taking cue from this, it is natural to consider the geometry of the singularities in the pp-adic case as well. As a first step, we may classify the singularities of our monoids via the theory of spherical varieties and we find that there is a good and explicit choice of GG-equivariant resolution of singularities [Bri89, Per14]. The resolution is moreover rational and so we may pass without trouble between differential forms on the monoid and its resolution. The geometric aspects of this theory are discussed in part in Section 3 of the present paper. Since our Schwartz spaces are, at least tentatively, linked by the definition of the Fourier transform JρJ^{\rho} via 𝒮ρ(G)=Cc(G(k))+Jρ(Cc(G(k))\mathcal{S}^{\rho}(G)=C_{c}^{\infty}(G(k))+J^{\rho}(C_{c}^{\infty}(G(k)), we are able, at least speculatively, to unite the themes of this paper. Here is the outline of the paper.

Section 1 is a quick review of the method for GL(n)GL(n) as developed in [GJ72]. Renner’s construction of reductive monoids is briefly discussed in Section 2 which concludes with a treatment of the cases of symmetric powers for GL(2)GL(2), describing all the objects involved in those cases. Section 33 covers the geometric aspects studied in the paper. This includes the resolution of the singularities of reductive monoids, leading to a proof of rationality of these singularities. This allows a transfer of measures from the monoid to its resolution as discussed in Section 44 and can be applied to the integration of basic functions on corresponding toric varieties in Example 4.14.1. Multiplicativity is stated and proved in Section 55, concluding with the example of GL(n)GL(n) in 5.35.3 and a discussion of the inductive nature of Fourier transforms in 5.45.4. In proving multiplicativity, we have found it easier to work with the full functional equation rather than the definition given by convolutions. The cases of a tori and unramified data are addressed in Section 66. The paper is concluded with a brief discussion of the doubling construction of Piatetski-Shapiro and Rallis with relevant references cited.

Acknowledgements

The authors would like to thank J. Getz, D. Jiang, and B.C. Ngo for helpful conversations. A part of this paper was presented by the first author during the month long program “On the Langlands Program: Endoscopy and Beyond” at the Institute for Mathematical Sciences, National University of Singapore, December 17, 2018-January 18, 2019. He would like to thank the Institute and the organizers: W. Casselman, P.-H. Chaudouard, W.T. Gan, D. Jiang, L. Zhang, and C. Zhu, for their invitation and hospitality. Finally, we would like to thank Jayce Getz, Chun-Hsien Hsu, and Michel Brion for their comments after the paper was posted on arXiv. Both authors were partially supported by NSF grants DMS 1801273 and DMS 2135021

1. The case of standard representation for GLnGL_{n}

We recall that the Godeement–Jacquet [GJ72] theory for standard LL-functions of GLnGL_{n}, which this method aims to generalize, can be presented briefly through the definition of corresponding γ\gamma-factors.

Let FF be a pp-adic field and G=GLnG=GL_{n}. Let π\pi be an irreducible admissible representation of GLn(F)GL_{n}(F). Given a Schwartz function ϕ\phi on Mn(F)M_{n}(F), i.e., ϕCc(Mn(F))\phi\in C^{\infty}_{c}(M_{n}(F)), a smooth function of compact support on Mn(F)M_{n}(F), one can define a zeta-function

Z(ϕ,f,s)=GLn(F)ϕ(x)f(x)|detx|s𝑑x,Z(\phi,f,s)=\int\limits_{GL_{n}(F)}\phi(x)f(x)|\det x|^{s}dx,

where f(x)=π(x)v,v~,v(π)f(x)=\langle\pi(x)v,\widetilde{v}\rangle,\ v\in\mathcal{H}(\pi) and v~(π~)\widetilde{v}\in\mathcal{H}(\widetilde{\pi}), is a matrix coefficient and ss\in\mathbb{C}. Here π~\widetilde{\pi} is the contragredient of π\pi. Let

ϕ^(x):=Mn(F)ϕ(y)ψ(tr(xy))𝑑y\hat{\phi}(x):=\int\limits_{M_{n}(F)}\phi(y)\psi(tr(xy))dy

be the Fourier transform of ϕ\phi with respect to the (additive) character ψ1\psi\neq 1 of FF.

If fˇ(g)=f(g1)\check{f}(g)=f(g^{-1}), gGLn(F)g\in GL_{n}(F), then we can consider Z(ϕ^,fˇ,s)Z(\hat{\phi},\check{f},s). The Godement–Jacquet theory defines a γ\gamma-factor γstd(π,s)\gamma^{\text{std}}(\pi,s) which depends only on π\pi and ss and is a rational function of qsq^{-s}, satisfying

(1.1) Z(ϕ^,fˇ,(1s)+n12)=γstd(π,s)Z(ϕ,f,s+n12)Z(\hat{\phi},\check{f},(1-s)+\frac{n-1}{2})=\gamma^{\text{std}}(\pi,s)Z(\phi,f,s+\frac{n-1}{2})

for all ϕ\phi and ff.

It is not hard to see that if we introduce the Int(G)(G)-invariant kernel, G=GLnG=GL_{n},

Φψ(g)=ψ(tr(g))|detg|ndg\Phi_{\psi}(g)=\psi(tr(g))|\det g|^{n}dg

of the Fourier transform, then

(1.2) Φψf|det|s+n12=γstd(π,s)f|det|s+n12\Phi_{\psi}\star f|\det|^{s+\frac{n-1}{2}}=\gamma^{\text{std}}(\pi,s)f|\det|^{s+\frac{n-1}{2}}

by virtu of irreducibility of π\pi and the Schur’s lemma.

This formulation for the γ\gamma-factor is a quick and convenient way of introducing them which is amenable to generalization. We can therefore write

γstd(π,s)=Φψ(π)=GLn(F)Φψ(g)π(g)𝑑g,\gamma^{\text{std}}(\pi,s)=\Phi_{\psi}(\pi)=\int\limits_{GL_{n}(F)}\Phi_{\psi}(g)\pi(g)dg,

pointing to the significance of the kernel Φψ\Phi_{\psi} in defining the γ\gamma-factors.

2. The general case; monoids and Renner’s construction

To treat the general case we need to generalize Mn(F)M_{n}(F). Let kk be an algebraically closed field of characteristic zero. A monoid MM is an affine algebraic variety over kk with an associative multiplication and an identity 1. For our purposes, we also need MM to be normal, i.e., k[M]k[M] is integrally closed in k(M)k(M). We can always find a normalization in case MM is not normal, i.e., an epimorphism M~M\widetilde{M}\to M such that integral closure of k[M]k[M] in k(M)k(M) equals k[M~]k[\widetilde{M}] as we realize k[M]k[M~]k[M]\hookrightarrow k[\widetilde{M}].

We thus let MM be a normal monoid and let G=G(M)=MG=G(M)=M^{*}, be the units of MM. We say MM is reductive if GG is. We now like to attach a monoid to a finite dimensional representation ρ\rho of G^=GL\hat{G}={{}^{L}}G, LL-group of GG, ρ:G^GL(Vρ)\rho:\hat{G}\to GL(V_{\rho}), where GG is a split reductive group. Let TGT\subset G be a maximal torus and write

ρ|T^=λW(ρ)λ,\rho\ |\ \hat{T}=\mathop{\bigoplus_{\lambda\in W(\rho)}}\lambda,

where W(ρ)W(\rho) is the set of weights of ρ\rho. Let Λ=Hom(𝔾m,T)\Lambda=\text{Hom}(\mathbb{G}_{m},T) be the set of cocharacters of TT or characters of T^\hat{T} and set Λ=Λ\Lambda_{\mathbb{R}}=\Lambda\otimes{{}_{\mathbb{Z}}}\mathbb{R}. Next, denote by Ω(ρ)\Omega(\rho) the convex span in Λ\Lambda_{\mathbb{R}} of weights of ρ\rho and let ξ(ρ)\xi(\rho) be the cone in Λ\Lambda_{\mathbb{R}} generated by rays through Ω(ρ)\Omega(\rho).

Let σ=ξ(ρ)X(T)\sigma^{\vee}=\xi(\rho)^{\vee}\cap X^{*}(T), be the “rational” dual cone to ξ(ρ)X(T)\xi(\rho)\cap X_{*}(T) and k[σ]k[\sigma^{\vee}] the group algebra of σ\sigma^{\vee}. One can then identify σ\sigma^{\vee} as a subset of k[σ]k[\sigma^{\vee}] by μσ\mu\in\sigma^{\vee} going to χμ\chi_{\mu}, defind by

χμ(η)=0unless η=μ,ησ,\chi_{\mu}(\eta)=0\quad\text{unless }\ \eta=\mu,\ \eta\in\sigma^{\vee},

χμ(μ)=1\chi_{\mu}(\mu)=1, and χμ1χμ2=χμ1+μ2\chi_{\mu_{1}}\cdot\chi_{\mu_{2}}=\chi_{\mu_{1}+\mu_{2}}, where the sum is the one on the semigroup σ\sigma^{\vee}. We note that this is valid for any semigroup SS and

k[S]=χs|sS.k[S]=\langle\chi_{s}|s\in S\rangle.

Now, assume GG has a character

ν:G𝔾m\nu:G\longrightarrow\mathbb{G}_{m}

such that

νG^𝜌GL(Vρ)\mathbb{C}^{*}\overset{\nu^{\vee}}{\longrightarrow}\hat{G}\overset{\rho}{\longrightarrow}GL(V_{\rho})

sends zz\in\mathbb{C}^{*} to zIdz\cdot\text{Id}. This means that ν,ω=1\langle\nu,\omega\rangle=1 for any weight ω\omega of ρ\rho. In fact, for zz\in\mathbb{C}^{*},

zν,ω=ω(ν(z))=ρ(ν(z))=zz^{\langle\nu,\omega\rangle}=\omega(\nu^{\vee}(z))=\rho(\nu^{\vee}(z))=z

and thus ν,ω=1\langle\nu,\omega\rangle=1. Then νσ\nu\in\sigma^{\vee} and its existence implies that ξ(ρ)\xi(\rho) is strictly convex, i.e., has no lines in it. In fact, the cone ξ(ρ)\xi(\rho) is contained in the open half–space of vectors xΛx\in\Lambda_{\mathbb{R}}, Λ=Λ\Lambda_{\mathbb{R}}=\Lambda\otimes_{\mathbb{Z}}\mathbb{R}, Λ=Hom(𝔾m,T)\Lambda=\text{Hom}(\mathbb{G}_{m},T), satisfying ν,x>0\langle\nu,x\rangle>0. It is therefore strictly convex. (cf. [Ngô20],Proposition 5.1).

By the theory of toric varieties [CLS11], the strictly convex cone ξ(ρ)\xi(\rho) determines (uniquely) a normal toric variety, i.e., a normal affine torus embedding j:TMTj:T\subset M_{T}. Here MTM_{T} is the monoid for TT attached to ρ|T^\rho|\hat{T}. More precisely, MT=Spec(k[σ])M_{T}=\text{Spec}(k[\sigma^{\vee}]) by Theorem 1.3.8, pg. 39 of [CLS11]. By definition 3.19 of [Ren05], k[σ]k[\sigma^{\vee}] is generated by X(MT)X(M_{T}) and thus X(MT)=σX(M_{T})=\sigma^{\vee}, the semigroup defining MTM_{T}. The embedding j:TMTj:T\subset M_{T}, defines j:X(MT)X(T)j^{*}:X(M_{T})\hookrightarrow X(T), a morphism of semigroups, into the character group of TT.

The dominant characters in X(T)X(T) all lie in X(MT)X(M_{T}) and are those that extend to semigroup morphism MT𝔸1=𝔾aM_{T}\to\mathbb{A}^{1}=\mathbb{G}_{a} (Proposition 3.20 of [Ren05]).

Finally we observe that ν\nu is integral and dominant and thus νX(MT)\nu\in X(M_{T}).


To proceed, we remark that the Weyl group W=W(G,T)W=W(G,T) acts on T,MT,X(T)T,M_{T},X(T) and X(MT)X(M_{T}) in the usual manner. Thus the dual rational cone σ\sigma^{\vee} may be identified with X(MT)X(M_{T}), both semigroups, since its group algebra generated by elements of X(MT)X(M_{T}) or σ\sigma^{\vee}, is k[MT]k[M_{T}] as we discussed earlier.

Let λX(T)\lambda\in X(T) be a dominant (and integral) character. Then λ|Tder\lambda|T_{\text{der}} defines an irreducible finite dimensional (rational) representation μλ\mu^{\circ}_{\lambda} of GderG_{\text{der}}, Tder=TGderT_{\text{der}}=T\cap G_{\text{der}}, of highest weight λ|Tder\lambda|T_{\text{der}}. Since

μλ|Z(G)Gder=λ|Z(G)Gder,\mu^{\circ}_{\lambda}|Z(G)\cap G_{\text{der}}=\lambda|Z(G)\cap G_{\text{der}},

we can extend μλ\mu^{\circ}_{\lambda} to an irreducible rational representation μλ=μλ(λ|Z(G))\mu_{\lambda}=\mu^{\circ}_{\lambda}\otimes(\lambda|Z(G)) of

G=(Gder×Z(G))/GderZ(G).G=(G_{\text{der}}\times Z(G))/G_{\text{der}}\cap Z(G).
Definition 2.1.

μλ\mu_{\lambda} is called the irreducible representation of GG of highest weight λ\lambda, where λ\lambda is a dominant rational character of TT.

This in particular is valid for dominant elements in X(MT)X(M_{T}). We note that νX(MT)\nu\in X(M_{T}) is one such.


Now choose {λi}i=1s\{\lambda_{i}\}^{s}_{i=1} so that i=1sWλiX(MT)\bigcup^{s}_{i=1}W\!\cdot\!\lambda_{i}\subset X(M_{T}) generates X(MT)X(M_{T}). Let (μλi,Vλi)(\mu_{\lambda_{i}},V_{\lambda_{i}}) be the representation attached to λi\lambda_{i}. Set μ=i=1sμλi\mu=\bigoplus\limits^{s}_{i=1}\mu_{\lambda_{i}} and V=i=1sVλiV=\bigoplus\limits^{s}_{i=1}V_{\lambda_{i}}. The character ν\nu will be among these λi\lambda_{i}. We may assume λ1=ν\lambda_{1}=\nu. Define M1=μ(G)¯End(V)M_{1}=\overline{\mu(G)}\subset\text{End}(V). We let MM be a normalization of M1M_{1}.

We note that we may take λi|Tder\lambda_{i}|T_{\text{der}} to be among the fundamental weights of GderG_{\text{der}}, with μλ1=ν\mu_{\lambda_{1}}=\nu extending the trivial representation of GderG_{\text{der}} since ν\nu is a representation (character) of G/GderZ(G)/GderZ(G)G/G_{\text{der}}\simeq Z(G)/G_{\text{der}}\cap Z(G).


2.1. The case of symmetric powers of GL2GL_{2}

As an example in this section we consider the symmetric power representations of GL2()GL_{2}(\mathbb{C}) and describe these objects in this case.

Let G=GL2G=GL_{2} and ρ=Symn:GL2()GLn+1()\rho=\text{Sym}^{n}:GL_{2}(\mathbb{C})\to GL_{n+1}(\mathbb{C}), the nn-th symmestric power of the standard representation of GL2()GL_{2}(\mathbb{C}). Write n+1=e1,,en+1\mathbb{C}^{n+1}=\langle e_{1},\dots,e_{n+1}\rangle with the basis e1,,en+1e_{1},\dots,e_{n+1}. Let {μi}\{\mu_{i}\} denote the weights of Symn\text{Sym}^{n}. Then we can order them as

μi((x00y))=xiyni((x,y)()2),\mu_{i}(\left(\begin{array}[]{cc}x&0\\ 0&y\end{array}\right))=x^{i}y^{n-i}\quad((x,y)\in(\mathbb{C}^{*})^{2}),

i=0,,ni=0,\dots,n. We have

ξ(Symn)X(T)=0span {(nk,k)|k=0,,n}\xi(\text{Sym}^{n})\cap X_{*}(T)=\mathbb{Z}_{\geq 0}-\text{span }\{(n-k,k)|k=0,\dots,n\}

inside 2\mathbb{R}^{2} which equals

0span {(m,l)|m+ln}.\mathbb{Z}_{\geq 0}-\text{span }\{(m,l)|m+l\in n\mathbb{Z}\}.

The dual cone to {(m,l)|m+ln}\{(m,l)|m+l\in n\mathbb{Z}\} is

{(a,b)1n×1n|ab}.\{(a,b)\in\frac{1}{n}\mathbb{Z}\times\frac{1}{n}\mathbb{Z}\ |\ a-b\in\mathbb{Z}\}.

Thus the dual to ξ(Symn)X(T)\xi(\text{Sym}^{n})\cap X_{*}(T) is the 0span\mathbb{Z}_{\geq 0}-\text{span} of {(1,0),(0,1),(1n,1n)}\{(1,0),(0,1),(\frac{1}{n},\frac{1}{n})\}. It is a lattice in the shaded area, corresponding to σ\sigma^{\vee}

xxyy(0,1)(0,1)(1n,1n)(\frac{1}{n},\frac{1}{n})(1,0)(1,0)

We use x,yx,y, and zz to denote (1,0),(0,1)(1,0),(0,1) and (1n,1n)(\frac{1}{n},\frac{1}{n}) in the semigroup algebra k[σ]k[\sigma^{\vee}] as before, i.e., x=χ(1,0)x=\chi_{(1,0)} and so on, then

k(x,y,z)=k[X,Y,Z]/(XYZn).k(x,y,z)=k[X,Y,Z]/(XY-Z^{n}).

The corresponding toric variety is

MT=Speck[X,Y,Z]/(XYZn)k3,M_{T}=\text{Spec}\,k[X,Y,Z]/(XY-Z^{n})\subseteq k^{3},

the variety defined by the zeros of XYZn=0XY-Z^{n}=0, and

T\displaystyle T =\displaystyle\!\!\!=\hskip-7.22743pt MT(k)3\displaystyle M_{T}\cap(k^{*})^{3}
=\displaystyle\!\!\!=\hskip-7.22743pt {(t1,t2nt11,t2)|tik,i=1,2}\displaystyle\{(t_{1},t^{n}_{2}t^{-1}_{1},t_{2})|t_{i}\in k^{*},\ i=1,2\}

The monoid MM for Symn\text{Sym}^{n} (Renner’s construction): The dual cone in

X(T)=X(T^)X^{*}(T)\otimes_{\mathbb{Z}}\mathbb{Q}=X_{*}(\hat{T})\otimes_{\mathbb{Z}}\mathbb{Q}

is generated by (1,0)(1n,1n)(1,0)(\frac{1}{n},\frac{1}{n}) and (0,1)(0,1). The vectors (1,0)(1,0) and (0,1)(0,1) are WW-conjugate and therefore we have as our dominant weights λi={(1,0),(1n,1n)}\lambda_{i}=\{(1,0),(\frac{1}{n},\frac{1}{n})\}. They correspond, respectively, to std, the standard representation, and ν=det1/n\nu=\det^{1/n} (to be explained) and thus

μ:G\displaystyle\mu:G\! \displaystyle\longrightarrow End(VstdVν)=M2×𝔸1\displaystyle\!\!\text{End}(V_{\text{std}}\oplus V_{\nu})=M_{2}\times\mathbb{A}^{1}
g\displaystyle g \displaystyle\!\longmapsto (g,(detg)1/n).\displaystyle\!\!(g,(\det g)^{1/n}).

Then

M\displaystyle M =\displaystyle= μ(G)¯\displaystyle\!\!\overline{\mu(G)}
=\displaystyle= {(g,a)|degg=an¯}\displaystyle\!\!\overline{\{(g,a)\ |\ \deg g=a^{n}}\}
\displaystyle\simeq Speck[X1,,X5]/(X1X4X2X3=X5n)\displaystyle\!\!\text{Spec}\,k[X_{1},\dots,X_{5}]\ /\ (X_{1}X_{4}-X_{2}X_{3}=X^{n}_{5})
=\displaystyle= Var(X1X4X2X3=X5n).\displaystyle\!\!Var(X_{1}X_{4}-X_{2}X_{3}=X^{n}_{5}).

The character ν\nu for Symn\text{Sym}^{n}:

Recall that the fibered product of GL2GL_{2} and 𝔾m\mathbb{G}_{m} giving the units of the monoid for Symn\text{Sym}^{n} is (cf. [Sha17])

G=GL2×𝔾m𝔾m={(g,a)|detg=an}={GL2n=oddSL2×GL1n=even.G=GL_{2}\times_{\mathbb{G}_{m}}\mathbb{G}_{m}=\{(g,a)\ |\ \det g=a^{n}\}=\left\{\begin{array}[]{cc}GL_{2}&n=\text{odd}\\ SL_{2}\times GL_{1}&n=\text{even}\end{array}\right..

We then have the commuting diagram

G=GL2×𝔾m𝔾mProj1GL2Proj2det𝔾m𝔾m\begin{CD}G=GL_{2}\times_{\mathbb{G}_{m}}\mathbb{G}_{m}@>{{\text{Proj}}_{1}}>{}>GL_{2}\\ \hskip 54.2025pt@V{{\text{Proj}}_{2}}V{}V@V{}V{\det}V\\ \hskip 54.2025pt\mathbb{G}_{m}@>{}>{}>{\mathbb{G}_{m}}\end{CD}
xxn.\hskip 54.2025ptx\longmapsto x^{n}.

Thus

(g,a)g\downmapsto\downmapstoadetg=an\begin{array}[]{ccc}(g,a)&{\longmapsto}&\hskip-14.45377ptg\\ \downmapsto&&\hskip-14.45377pt\downmapsto\\ a&\longmapsto&{\det g=a^{n}}\end{array}

and the left vertical arrow, the Proj2, gives

ν:(g,a)(detg)1n=a\nu:(g,a)\longrightarrow(\det g)^{\frac{1}{n}}=a

for which

zν diag (z1/n,,z1/n)SymnzIn+1.z\overset{\nu^{\vee}}{\longrightarrow}\text{ diag }(z^{1/n},\dots,z^{1/n})\overset{\text{Sym}^{n}}{\longrightarrow}z\cdot I_{n+1}.

3. Some geometry of reductive monoids as spherical varieties

Renner’s classification of reductive monoids uses the “extension principle” [Ren05]. The extension principle follows in the spirit of many similar classification results for spherical varieties that rely on the existence of an open B×BopB\times B^{\textrm{op}}-orbit where BB is a Borel subgroup of GG, that is in Renner’s case adapted to account for the monoid structure. By Renner’s classification, the category of Reductive monoids is equivalent to the category of tuples (G,T,T¯)(G,T,\overline{T}), where TT is any maximal torus in GG and T¯\overline{T} is a Weyl-group stable toric variety. A morphism of data (G,T,T¯)(G,T,T¯)(G,T,\overline{T})\to(G^{\prime},T^{\prime},\overline{T^{\prime}}) are given by a pair (φ,τ)(\varphi,\tau) where φ:GG\varphi:G\to G^{\prime} is a morphism of reductive groups and τ:T¯T¯\tau:\overline{T}\to\overline{T^{\prime}} a morphism of toric varieties such that the restriction of each morphism to the maximal torus agree φ|T=τ|T\varphi|T=\tau|T. In the following, we reframe these results in terms of the theory of spherical varieties, in order to state the existence of a GG-equivariant resolution of singularities [BK07], [Rit03].

Let GG be a split reductive group defined over a characteristic zero field kk. Let XX be a variety defined over kk with an rational action α:G×XX\alpha:G\times X\to X. In this case we say XX is a GG-variety. Let 𝒪X\mathscr{O}_{X} be the sheaf of regular on XX. If XX is affine, we will identify 𝒪X\mathscr{O}_{X} with the coordinate algebra k[X]k[X]. In this case α\alpha induces as usual a co-action map α:k[X]k[G]k[X]\alpha^{*}:k[X]\to k[G]\otimes k[X] by (αf)(x,g)=f(g1x)=ihi(g)fi(x)(\alpha^{*}f)(x,g)=f(g^{-1}x)=\sum_{i}h_{i}(g)f_{i}(x) with hik[G]h_{i}\in k[G], where the latter is a finite sum. Thus each ff determines a finite dimensional GG-module. Because GG is reductive and we are in characteristic 0, each finite dimensional GG-module decomposes as a finite sum of irreducible representations indexed by their highest weight vector with weight λ\lambda. As such we may decompose k[X]=k[X]λk[X]=\bigoplus k[X]_{\lambda}, indexed by the λ\lambda that appear in k[X]k[X].

Definition 3.1.

A GG-variety XX is spherical if XX has an open BB-orbit for some (hence any) Borel BB in GG.

Suppose XX is spherical. Then as above, by highest weight theory, each dominant integral character λ\lambda of T(k)T(k) that appears in k[X]k[X] has a highest weight vector fλf_{\lambda}. The line kfλk\cdot f_{\lambda} is the unique line stabilized by BB on which BB acts through the character λ:fλ(bx)=λ(b)fλ\lambda:f_{\lambda}(bx)=\lambda(b)f_{\lambda}. In other words fλf_{\lambda} is a semi-invariant. Suppose f1f_{1} and f2f_{2} are semi-invariants that are λ\lambda-eigenfunctions appearing in k[X]k[X]. Then the rational function f1/f2f_{1}/f_{2} is BB-invariant. As the BB-orbit in GG is dense, this implies f1/f2f_{1}/f_{2} is constant. Hence for spherical varieties, each λ\lambda that appears can only appear with multiplicity one. For general reductive groups actions, even the “naive” (categorical) quotient is reasonably well behaved.

Theorem 3.2.

Let XX be an affine GG-variety for a reductive group GG. Then the ring of GG-invariants k[X]Gk[X]^{G} is finitely generated, say k[X]G=k[f1,,fn]k[X]^{G}=k[f_{1},\ldots,f_{n}]. Then k[X]Gk[X]k[X]^{G}\hookrightarrow k[X] defines a good surjective quotient which is moreoever a categorical quotient q:XX//Gq:X\to X//G. Each fiber of qq contains a unique closed GG-orbit in XX, and X//GX//G is normal if XX is.

Definition 3.3.

A spherical variety XX is simple if it has a unique closed GG-orbit.

We are interested in reductive monoids, which have open GG-orbit and are spherical with respect to G×GG\times G with an open dense borel =B×Bop=B\times B^{\mathrm{op}} orbit.

Proposition 3.4.

Suppose XX has an open GG-orbit. Then XX has a unique closed orbit.

Proof.

The reductive quotient q:XX//Gq:X\to X//G is constant on orbits, in particular on the open orbit. Hence X//G={pt}X//G=\{\mathrm{pt}\}. The fiber q1(pt)q^{-1}(\mathrm{pt}) contains a unique closed orbit by Theorem 3.2 ∎

Therefore such XX are simple. Once again let us consider an affine simple XX as a G×GG\times G variety. Decomposing λk[X]λVλ\bigoplus_{\lambda}k[X]_{\lambda}\cong V_{\lambda} where VλV_{\lambda} is the highest weight module for λ\lambda. The (B,λ)(B,\lambda) eigenfunction fλf_{\lambda} is UU-invariant. Thus one may consider taking U×UopU\times U^{\mathrm{op}}-invariants k[X]U×Uopk[X]^{U\times U^{\textrm{op}}} are therefore generated as a vector space by the fλf_{\lambda}. Using the following

Theorem 3.5.

Let GG be reductive with maximal unipotent subgroup UU, and let XX be a GG-variety. Then k[X]Uk[X]^{U} is finitely generated. Moreover X/U=speck[X]UX/U=\operatorname{spec}k[X]^{U} is normal if XX is.

Proof.

One first establishes the theorem for G/UG/U i.e. k[G]Uk[G]^{U} is finitely generated and in fact G/UG/U is a geometric quotient (a so called horospherical variety). One has a map Φ:X/UX×GG/U\Phi:X/U\cong X\times^{G}G/U where the quotient is by the diagonal action. On coordinate rings: a UU-invariant ff defines a GG-invariant functions (Φf)(x,gU)=f(gx)(\Phi^{*}f)(x,gU)=f(gx). Thus by Theorem 1.2 k[X×G/U]Gk[X]Uk[X\times G/U]^{G}\cong k[X]^{U} is finitely generated. ∎

We can conclude that the variety X/(U×Uop)X/(U\times U^{\mathrm{op}}) is a TU\BBop/UopUop\G/UT\cong U\backslash BB^{\mathrm{op}}/U^{\mathrm{op}}\hookrightarrow U^{\mathrm{op}}\backslash G/U variety, on which TT acts on fλf_{\lambda} through the character λ\lambda. In other words, we have a ring (VλVλ)(U×Uop)\bigoplus(V_{\lambda}\otimes V_{\lambda}^{*})^{(U\times U^{\textrm{op}})} graded by ΛX={λX(T):k[X]λ0}\Lambda_{X}=\{\lambda\in X^{*}(T):\ k[X]_{\lambda}\neq 0\}. By Theorem 3.53.5, this is a finitely generated monoid. Each summand has a diagonalizable action by the torus TT, giving an equivalent characterization of toric varieties: hence defines an affine TT embedding. If XX is a reductive monoid, this must therefore be equal to Renner’s cone. Moreover, if XX is normal the associated toric variety is normal, hence the cone of weights of XX defining the toric variety is saturated.

Remark 3.6.

We have XX normal, and the map XX//UX\to X//U is faithfully flat. It is sometimes called a toric degeneration or contraction, see [Pop87]. This is used to show that XX has rational singularities if and only if T¯\overline{T} does, because the argument relies on a flat base change argument [Elk78].

Recall that a GG-variety is simple if it contains a unique closed orbit. When XX is affine, it is enough that GG embeds as an open subvariety to imply XX is simple. We state without proof the following:

Proposition 3.7.

Any GG-variety can be covered by simple GG-varieties.

To classify a general spherical variety one needs the following additional data.

Definition 3.8.

Let V(X)V(X) denote the GG-stable discrete valuations on k[G]k[G].

Definition 3.9.

Let DivB(X)\text{Div}_{B}(X) denote the set of BB-stable prime divisors of XX.

Definition 3.10.

Let ZZ be a GG-orbit in XX. Then DivB(X:Z)\text{Div}_{B}(X:Z) is the set of BB-stable prime divisors containing ZZ.

Definition 3.11.

Let (X)\mathcal{B}(X) denote the set of irreducible GG-stable divisors.

Proposition 3.12.

For a divisor DDivB(X)D\in\text{Div}_{B}(X), either

  1. (1)

    The BB-orbit BDB\cdot D is open and dense in the open orbit of XX.

  2. (2)

    DD is GG-stable.

Briefly (although see [Kno91] for details) a simple spherical variety XX is determined by its weight monoid, plus the data of which BB-stable boundary divisors containing the unique closed orbit ZZ are GG-stable and which are not. More precisely, each divisor DD defines a valuation by first restricting DGD\cap G which defines a valuation on the multiplicative group of rational functions k(G)×k(G)^{\times} on GG (the valuation vDGv_{D}\cap G is the order of vanishing of a rational function f/gf/g on DGD\cap G). This defines a so-called colored cone, defined by the 0\mathbb{Q}_{\geq 0} span of the finite number of valuations vDv_{D} as above in which the colors are the valuations that are BB-stable but not GG-stable (or dually, their corresponding divisors have a BB-open-orbit).

Thus the set D(X:Z)D(X:Z) gives the set of colors of the simple spherical variety XX. The cone generated by (X)\mathcal{B}(X) and the natural image of D(X:Z)D(X:Z) in the set of K[G]K[G] valuations is the colored cone 𝒞\mathcal{C} determined by the data (V(X),(X))(V(X),\mathcal{B}(X)) that determines up to isomorphism the spherical variety XX. For reductive monoids, this cone is equivalent to the one constructed in the earlier section via highest weight theory.

Example 3.13.

For a reductive monoid MM, there is a beautiful description of the boundary M=M\G\partial M=M\backslash G in terms of B×BopB\times B^{\text{op}}-stable boundary divisors in the form of an extended Bruhat decomposition: Let R=NG(T)¯MR=\overline{N_{G}(T)}\subset M be the Zariski closure of the normalizer of a maximal torus TT in GG. Let I(M)I(M) be the set of idempotents in MM, and note that reductive monoids are regular (in Renner’s sense), that is, we can decompose M=GI(M)M=G\cdot I(M). Then we can construct the Renner monoid (sometimes called the Rook monoid) :=R/T\mathcal{R}:=R/T. Because reductive monoids are regular, \mathcal{R} makes sense as a finite monoid whose unit group is the Weyl group WW, and having the property that

M=xBxB.M=\coprod_{x\in\mathcal{R}}BxB.

From this description, it may be deduced that the set D(M:Z)D(M:Z), with Z={0}Z=\{0\} the unique closed orbit in MM, is given by the codimension one orbits BsαBop¯\overline{Bs_{\alpha}B^{\text{op}}} for sαs_{\alpha}\in\mathcal{R} the simple reflection in the Weyl group determined by the simple root α\alpha.

Thus, for a monoid with reductive group GG embedded as its unit group, the colors of MM are all the B×BopB\times B^{\text{op}} stable irreducible divisors of GG, and thus the monoid is determined purely by the data (M)\mathcal{B}(M) or equivalently 𝒞(M)\mathcal{C}(M). We state for convenience this form of the classification.

Theorem 3.14.

Let GG be a reductive group. The irreducible, normal algebraic monoids MM with unit group GG are the strictly convex polyhedral cones in X(T)X_{*}(T)\otimes\mathbb{Q} generated by D(M)D(M) and a finite set of elements in V(G)V(G).

The theory of reductive monoids affords us an explicit description of (M)\mathcal{B}(M) purely in monoid-theoretic terms.

Definition 3.15.

A spherical variety XX is toroidal if D(X)D(X) is empty.

Proposition 3.16.

Suppose the spherical variety XX is toroidal and let X=XG\partial X=X\setminus G. Let PXP_{X} be the G×GG\times G stabilizer of X\partial X. Then PP is a parabolic and moreover satisfies the local structure theorem, i.e. there is a Levi LPL\subset P, depending only on GG and a closed LL-variety ZZ such that

Pu×ZXXP_{u}\times Z\to X\setminus\partial X

is an isomorphism. Moreover, ZZ is a toric variety under L/[L,L].

As a consequence of the above isomorphism, the LL orbits of ZZ correspond to GG orbits in XX. Note that when X=MX=M is a reductive monoid, this is precisely Renner’s extension theorem [Ren05] with PM=B×BopP_{M}=B\times B^{\text{op}} and Z=T¯Z=\overline{T}. The proposition implies that the singularities of XX are those determined by the cone of the toric variety ZZ.

Let us recall Renner’s extension theorem for normal reductive monoids. It states that a morphism of reductive monoids MMM\to M^{\prime} is given by the data (G,T,T¯)(G,T,\overline{T}) and (G,T,T¯)(G^{\prime},T^{\prime},\overline{T^{\prime}}) and a morphism φ:MM\varphi:M\to M^{\prime} is equivalent to φ|GG\varphi|G\to G^{\prime} and τ|T¯:T¯T¯\tau|\overline{T}:\overline{T}\to\overline{T^{\prime}}. Briefly, in MM one has an analogue of the open cell which is the image of an open embedding Uop×T¯×UUopT¯UU^{\mathrm{op}}\times\overline{T}\times U\to U^{\mathrm{op}}\overline{T}U which has codimension 2\geq 2 in MM. Thus one gets an equivariant morphism (u,t,u)φ(u)τ(t)φ(u)M(u^{\prime},t,u)\mapsto\varphi(u^{\prime})\tau(t)\varphi(u)\in M^{\prime}. By normality of MM, the codimension 2\geq 2 condition extends the map uniquely to MMM\to M^{\prime}, and one verifies this is in fact a morphism of monoids.

Remark 3.17.

The above map in Renner uses an open Bruhat cell analog in the context of monoids, which yields a structure theorem parallel to Proposition 3.163.16. More generally, GG-equivariant dominant morphisms φ:YY\varphi:Y\to Y^{\prime} of spherical varieties are in bijection with linear maps φ:X(Y)X(Y)\varphi_{*}:X_{*}(Y)\otimes\mathbb{Q}\to X_{*}(Y^{\prime})\otimes\mathbb{Q}, where X(Y)X_{*}(Y) and X(Y)X_{*}(Y^{\prime}) are the lattices of co-weights of the underlying group GG, such that the image of the colored fan 𝒞Y\mathcal{C}_{Y} is contained within 𝒞Y\mathcal{C}_{Y^{\prime}}.

In view of the above structure theory, the singularities of a spherical GG variety XX are determined by those of its associated toric variety X//UT¯X//U\cong\overline{T}. Smooth toric varieties are classified as follows.

Theorem 3.18.

An affine toric variety T¯\overline{T} is smooth if and only if the extremal rays of the rational polyhedral cone generated by its weight lattice is a basis for for the character lattice X(T)X^{*}(T).

As a general toric variety is glued from affine toric varities, it is given by a fan consisting of rational polyhedral cones. Thus a toric variety is smooth if and only if its fan consists of rational polyhedral cones whose extremal rays generate (as a \mathbb{Z}-module) X(T)X^{*}(T). Moreover any peicewise linear morphism of rational polyhedral cones 𝒞𝒞\mathcal{C}\to\mathcal{C}^{\prime} defines a TT-equivariant morphism of toric varietie, thus, by Theorem 3.183.18, one obtains an algorithm giving a resolution of singularities of a toric variety T¯\overline{T}.

Theorem 3.19.

Let σ=𝒞(T¯)X(T)=\sigma^{\vee}=\mathcal{C}(\overline{T})\cap X^{*}(T)= the monoid of weights of T¯\overline{T}. Starting from an extremal ray 𝒞(T¯)\mathcal{C}(\overline{T}), successively subdividing the cone such that each resulting cone is smooth, defines a smooth fan consisting of smooth cones σ~\tilde{\sigma}^{\vee} such that the inclusion map σ~σ\tilde{\sigma}^{\vee}\hookrightarrow\sigma^{\vee} defines a canonical TT-equivariant resolution of singularities T~T¯\widetilde{T}\to\overline{T}.

Finally, one may define a GG-equivariant resolution from Remark 3.173.17 and Theorems 3.183.18 and 3.193.19:

Theorem 3.20.

Let MM be a reductive monoid. Then there exists a smooth spherical GG-variety M~\tilde{M} that is toroidal, and a proper GG-equivariant morphism φ:M~M\varphi:\tilde{M}\to M

Proof.

Take the colored cone 𝒞M\mathcal{C}_{M} determed by MM. Deleting all colors, construct a fan 𝒞~M\tilde{\mathcal{C}}_{M} generated by a subdivision of 𝒞M\mathcal{C}_{M} into smooth colorless cones (which always exists, see [CLS11]). Then T~T¯M\tilde{T}\to\overline{T}\subset M is a resolution of toric varieties, and M~\tilde{M} has an affine chart given by the open cell Uop×T~×UU^{\textrm{op}}\times\tilde{T}\times U by Proposition 3.163.16, and therefore is smooth. The natural inclusion of 𝒞~M\tilde{\mathcal{C}}_{M} into 𝒞M\mathcal{C}_{M} defines a dominant GG-equivariant morphism M~M\tilde{M}\to M, giving us our resolution. ∎

Example 3.21.

The case of G=GL2G=GL_{2} and ρ=Symn\rho=\mathrm{Sym}^{n} monoids are determined by their respective toric varieties {xyzn=0}𝔸2/Cn\{xy-z^{n}=0\}\cong\mathbb{A}^{2}/C_{n}, which are realized finite quotients of 𝔸2\mathbb{A}^{2} by cyclic group CnC_{n} of order nn. These are the well known finite quotient surface singularities of type AnA_{n} and their resolution are give by [n2][\frac{n}{2}] blow-ups at the origin. Likewise, the resolution of MρM^{\rho} is given by [n2][\frac{n}{2}]-blow-ups at 0 of MρM^{\rho}.

Next we discuss the rationality of singularities. Recall that XX has rational singularities if a (and hence any) resolution r:X~Xr:\widetilde{X}\to X has vanishing higher cohomology: Rir𝒪X~=0R^{i}r_{*}\mathcal{O}_{\widetilde{X}}=0 if i>0i>0. Moreover, recall that for a resolution r:X~Xr:\widetilde{X}\to X the fiber over the singular locus Xsing=r1(Xsing)=EX^{\mathrm{sing}}=r^{-1}(X^{\mathrm{sing}})=E is the exceptional divisor.

Theorem 3.22.

Let XX be a normal variety with rational singularities. Let j:XsmXj:X^{\text{sm}}\hookrightarrow X denote the embedding of the smooth locus of XX into XX. If we define ωX:=j(ωXsm)\omega_{X}:=j_{*}(\omega_{X^{\text{sm}}}), by extending (uniquely, by normality) algebraic top-dimensional differentials form to XX. If r:X~Xr:\widetilde{X}\to X is any resolution of singularities, then rωXr^{*}\omega_{X} extends over EE to an algebraic differential form on X~\widetilde{X}.

Theorem 3.23.

Normal spherical varieties, and hence reductive monoids have rational singularities.

Proof.

XX//UT¯X\to X//U\cong\overline{T} is a flat deformation onto a toric variety, and toric varieties have rational singularities. Rationality “descends” under flat deformation (use flat base change, see [Elk78]), hence XX has rational singularities. ∎

4. Integration on singular varieties

Let ω\omega be a top-differential form on a reductive monoid MM. Taking the canonical resolution as above r:M~Mr:\widetilde{M}\to M we get a well defined differential form rωr^{*}\omega on M~\widetilde{M}. Restricing rr to the open set on M~\widetilde{M} over which rr is an isomorphism, we have upon passing to kk-points

X~(k)|rω|=X~(k)|Jac(r)||ω|=X(k)|ω|,\int_{\widetilde{X}(k)}|r^{*}\omega|=\int_{\widetilde{X}(k)}|\mathrm{Jac}(r)||\omega|=\int_{X(k)}|\omega|,

where Jac(r)\mathrm{Jac}(r) is the Jacobian of rr and |ω||\omega| is the measure constructed by Weil from the top-differential form ω\omega on XX.

Example 4.1.

As a first basic example, we may consider the basic function on toric varieties. Recall that the basic function on a toric variety T¯\overline{T} defined by cone σ\sigma with dual cone σ\sigma^{\vee}. Suppose the generators of the weight monoid of T¯\overline{T} are e1,,ene_{1},\ldots,e_{n}. The eie_{i} are co-characters k×T(k)k^{\times}\to T(k). For each co-character λ\lambda in the weight monoid, λ(ϖ𝒪k)\lambda(\varpi\mathscr{O}_{k}) is an open neighbood on which the value of the basic function fT¯ρf^{\rho}_{\overline{T}} is #{(ai)0n|iaiei=λ}\#\{(a_{i})\in\mathbb{Z}^{n}_{\geq 0}\ |\sum_{i}a_{i}e_{i}=\lambda\}. It is known [Stu95, Cas17] that the function

λ#{(ai)0n|iaiei=λ}:σX(T)0\lambda\mapsto\#\{(a_{i})\in\mathbb{Z}^{n}_{\geq 0}\ |\sum_{i}a_{i}e_{i}=\lambda\}:\sigma\cap X_{*}(T)\to\mathbb{Z}_{\geq 0}

as an integer valued function of the weight monoid is quasi-polynomial, i.e., it is in the sub-algebra of functions on X(T)X_{*}(T) generated by polynomials and periodic functions. In particular, this means one may dominate the basic function with a polynomial function on the lattice.

It is also classical that locally in some coordinate system on T~(k)\widetilde{T}(k), the map r:T~(k)T¯(k)r:\widetilde{T}(k)\to\overline{T}(k) is given by a monomial transformation, i.e., has the form (t1,,tn)(t1d11tndn1,,t1dn1tndnn)(t_{1},\ldots,t_{n})\mapsto(t_{1}^{d_{11}}\cdots t_{n}^{d_{n1}},\ldots,t_{1}^{d_{n1}}\cdots t_{n}^{d_{nn}}). For example, the Sym2\mathrm{Sym}^{2} toric variety is the cone {xyz2=0}\{xy-z^{2}=0\} and is resolved affine-locally by the classical “cylinder” resolution (x,y,z)(xz,yz,z)(x,y,z)\mapsto(xz,yz,z). Therefore the |Jac(r)||\mathrm{Jac}(r)| is in local coordinates a product of the form |ti|di|t_{i}|^{d_{i}}. For the cylinder resolution it is |z|2|z|^{2}. Thus as ϖN0\varpi^{N}\to 0 the Jacobian grows as the inverse of an exponential while the basic function is dominated by polynomial growth. Hence the pullback of the differential form r(fρω)r^{*}(f^{\rho}\omega) makes sense as valdet\mathrm{val}\circ\mathrm{det}\to\infty on T~(k)\widetilde{T}(k).

5. Multiplicativity


Every theory of LL-functions is expected to satisfy “multiplicativity”, i.e., the equality of γ\gamma-factors for parabolically inducing and induced data. Our goal in the rest of the paper is to establish, under natural assumptions on related Fourier transforms, the multiplicativity in the context of this theory in general. Our proof is a generalization of the standard case of Godement–Jacquet [GJ72]. We first need to connect Renner’s construction to parabolic induction.


5.1. Renner’s construction and parabolic induction

We start by observing that Renner’s construction [Ren05] respects parabolic induction. More precisely, let P=LNP\!=\!LN be a parabolic subgroup of GG, with unipotent radical NN and a Levi subgroup LL which we fix by assuming TLT\subset L. Now LL is a reductive group with a maximal torus TT to which Renner’s construction applies. Let ρL=ρ|L^\rho_{L}=\rho|\hat{L}, where L^\hat{L} is the connected component of the LL-group of LL.

Now ρL|T^\rho_{L}|\hat{T} gives the same weights as ρ|T^\rho|\hat{T} does and thus ρL\rho_{L} shares the same toric variety MTM_{T} coming from ρ\rho. Let WL=W(L,T)W_{L}=W(L,T). Then each orbit WλiW\lambda_{i} breaks up to a disjoint union of orbits WLλjW_{L}\lambda_{j} and thus

i=1sVλi|L=i=1sj=1siVλjL,\bigoplus^{s}_{i=1}V_{\lambda_{i}}\ {\big{|}}_{L}=\bigoplus^{s}_{i=1}\bigoplus^{s_{i}}_{j=1}\ V^{L}_{\lambda_{j}},

where

Vλi|L=j=1siVλjL.V_{\lambda_{i}}\ {\big{|}}L=\bigoplus^{s_{i}}_{j=1}\ V^{L}_{\lambda_{j}}.

(5.1) Conclusion: The monoid MρLM^{\rho_{L}} attached to ρL\rho_{L} by Renner’s construction for LL, L=G(MρL)L=G(M^{\rho_{L}}), is the same as the closure of LL as a subgroup of GG upon action on V=i=1sVλiV=\bigoplus^{s}_{i=1}V_{\lambda_{i}}, i.e., MρL=μ(L)¯M^{\rho_{L}}=\overline{\mu(L)}.


We also need to remark that the character ν:Gm\nu:G\to\mathbb{C}_{m} discussed earlier, when restricted to LL maybe considered as the corresponding character νL\nu_{L} of LL, i.e., νL:=ν|L\nu_{L}:=\nu|L. In fact, ν\nu^{\vee} and νL\nu^{\vee}_{L} both take values in the centers of G^\hat{G} and L^\hat{L}, respectively. The natural embedding of L^G^\hat{L}\subset\hat{G} by means of the root data of L^\hat{L} and G^\hat{G} which are dual to root data of LL and GG who share the maximal torus TT, identifies Z(G^)Z(\hat{G}) as a subgroup of Z(L^)Z(\hat{L}). We therefore have the commutative diagram


\textstyle{\mathbb{C}^{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ν\scriptstyle{{\nu^{\vee}}}νL\scriptstyle{\nu_{L}^{\vee}}G^\textstyle{\hat{G}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ\scriptstyle{\!\!\!\!\!\!\!\!\!\!\rho}GLN(Vρ)\textstyle{GL_{N}(V\rho)}L^\textstyle{\hat{L}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρL\scriptstyle{\rho_{{}_{L}}}

and consequently ρLνL(z)=zId\rho_{L}\cdot\nu_{L}^{\vee}(z)=z\cdot\text{Id} as needed.


The shift in general. To get the precise γ\gamma-factor in general one needs to shift ss by ηG,λ\langle\eta_{G},\lambda\rangle or π|ν|s\pi\otimes|\nu|^{s} should change to π|ν|s+ηG,λ\pi\otimes|\nu|^{s+\langle\eta_{G},\lambda\rangle}, with notation as in [Ngô20], where ηG\eta_{G} is half the sum of positive roots in a Borel subgroup of GG and λ\lambda the hightest weight of ρ\rho.

In our setting, we need to deal with the representation ρL\rho_{{}_{L}} of L^\hat{L} as well which is not necessarily irreducible. Let λ1,,λr\lambda_{1},\dots,\lambda_{r} be the hightest weights of ρL\rho_{{}_{L}}. We may assume λ1=λ\lambda_{1}=\lambda. The shift will then be ηL,λ1++λr\langle\eta_{L},\lambda_{1}+\cdots+\lambda_{r}\rangle.


Let us define δG,ρ=|ν|2ηG,λ\delta_{G,\rho}=|\nu|^{\langle 2\eta_{G},\lambda\rangle}, δL,ρL=|νL|2ηL,λ1++λr\delta_{L,\rho_{L}}=|\nu_{{}_{L}}|^{\langle 2\eta_{L},\lambda_{1}+\cdots+\lambda_{r}\rangle} and set

νG/L=νG/L,ρ:=δG,ρ/δL,ρL.\nu_{G/L}=\nu_{G/L},\rho:=\delta_{G,\rho}/\delta_{L,\rho_{{}_{L}}}.

Finally, let δP\delta_{P} be the modulus character of P=NLP=NL.


5.2. The ρ\rho–Harish–Chandra transform

We now recall the ρ\rho-Harish–Chandra transform, a generalization of Satake transform. Given ΦCc(G(k))\Phi\in C^{\infty}_{c}(G(k)), define its Harish–Chandra transform ΦPCc(L(k))\Phi_{P}\in C^{\infty}_{c}(L(k)) by

(5.1) ΦP(l)=δP12(l)N(k)Φ(nl)𝑑n.\Phi_{P}(l)=\delta^{-\frac{1}{2}}_{P}(l)\int\limits_{N(k)}\Phi(nl)dn.

Next define the ρ\rho-Harish–Chandra transform, ρ\rho–HC in short, by

(5.2) ΦPρ(l)=νG/L,ρ12(l)ΦP(l).\Phi^{\rho}_{P}(l)=\nu^{\frac{1}{2}}_{G/L,\rho}(l)\ \Phi_{P}(l).

Fourier transforms


The conjectural Fourier transform (kernel) JρJ^{\rho} is supposed to give the γ\gamma-factor γ(s,π,ρ)\gamma(s,\pi,\rho) for every irreducible admissible representation π\pi of G(k)G(k) through the convolution

(5.3) Jρf|ν|s+ηG,λ=γ(s,π,ρ)f|ν|s+ηG,λ,J^{\rho}*f|\nu|^{s+\langle\eta_{G},\lambda\rangle}=\gamma(s,\pi,\rho)f|\nu|^{s+\langle\eta_{G},\lambda\rangle},

where π\pi is an irreducible admissible representation of G(k)G(k) and f(g)=π(g)v,v~f(g)=\langle\pi(g)v,\widetilde{v}\rangle is a matrix coefficient of π\pi. In the context of parabolic induction from P=NLP=NL we will also have γ(s,σ,ρL)\gamma(s,\sigma,\rho_{{}_{L}}) defined by JρLJ^{\rho_{{}_{L}}}.


Fourier transforms and Schwartz spaces

In this section we define suitable spaces of Schwartz functions on G(k)G(k) and L(k)L(k), assuming how Fourier transforms JρJ^{\rho} and JρLJ^{\rho_{{}_{L}}} act on Cc(G(k))C^{\infty}_{c}(G(k)) and Cc(L(k))C^{\infty}_{c}(L(k)), respectively. Then

Jρ:Cc(G(k))C(G(k))J^{\rho}:C^{\infty}_{c}(G(k))\longrightarrow C^{\infty}(G(k))

and

JρL:Cc(L(k))C(L(k)).J^{\rho_{{}_{L}}}:C^{\infty}_{c}(L(k))\longrightarrow C^{\infty}(L(k)).

Assume JρJ^{\,\rho} and JρLJ^{\,\rho_{{}_{L}}} commute with the ρ\rho–Harish–Chandra transform ΦPρ\Phi^{\rho}_{P}, i.e.,

(5.4) (JρΦ)Pρ=JρLΦPρ(J^{\rho}\Phi)^{\rho}_{P}=J^{\rho_{{}_{L}}}\Phi^{\rho}_{P}

or the following diagram commutes

(5.5) Cc(G(k))JρJρ(Cc(G(k))C(G(k))ρHCρHCCc(L(k))JρLJρL(Cc(L(k))C(L(k)).\begin{CD}C^{\infty}_{c}(G(k))@>{J^{\rho}}>{}>J^{\rho}(C^{\infty}_{c}(G(k))\subset C^{\infty}(G(k))\\ @V{\rho-\rm{HC}}V{}V\hskip-72.26999pt@V{}V{{\rho}-\rm{HC}}V\\ C^{\infty}_{c}(L(k))@>{J^{\rho_{L}}}>{}>J^{\rho_{{}_{L}}}(C^{\infty}_{c}(L(k))\subset C^{\infty}(L(k)).\end{CD}

We now define the Schwartz spaces 𝒮ρ(G)=𝒮ρ(G(k))\mathcal{S}^{\rho}(G)=\mathcal{S}^{\rho}(G(k)) and 𝒮ρL(L)=𝒮ρL(L(k))\mathcal{S}^{\rho_{{}_{L}}}(L)=\mathcal{S}^{\rho_{{}_{L}}}(L(k)) as follows:

(5.6) 𝒮ρ(G):=Cc(G(k))+Jρ(Cc(G(k))C(G(k))\mathcal{S}^{\rho}(G):=C^{\infty}_{c}(G(k))+J^{\rho}(C^{\infty}_{c}(G(k))\subset C^{\infty}(G(k))

and

(5.7) 𝒮ρL(L):=Cc(L(k))+JρL(Cc(L(k))C(L(k)).\mathcal{S}^{\rho_{{}_{L}}}(L):=C^{\infty}_{c}(L(k))+J^{\rho_{{}_{L}}}(C^{\infty}_{c}(L(k))\subset C^{\infty}(L(k)).

As we pointed out in the introduction, these are subspaces of the conjectural ρ\rho-Schwartz spaces and suitable to our purposes. Moreover, as we prove in Proposition 5.3, they contain the ρ\rho and ρL\rho_{L}-basic functions. We recall that the ρ\rho-basic function ϕρ\phi^{\rho} is the unique one for which

Z(ϕρ,fs)=L(s,π,ρ),Z(\phi^{\rho},f_{s})=L(s,\pi,\rho),

where fsf_{s} is the normalized spherical matrix coefficient of π|ν|s\pi\otimes|\nu|^{s} with ZZ defined as in (5.11)(5.11).

Note that ΦΦPρ\Phi\mapsto\Phi^{\rho}_{P} sends Cc(G(k))C^{\infty}_{c}(G(k)) into Cc(L(k))C^{\infty}_{c}(L(k)) and thus (5.5) implies that it also sends

Jρ(Cc(G(k)))JρL(Cc(L(k))).J^{\rho}(C^{\infty}_{c}(G(k)))\longrightarrow J^{\rho_{{}_{L}}}(C^{\infty}_{c}(L(k))).

We thus have:


Proposition 5.1.

The ρ\rho–Harish–Chandra transform ΦPρ\Phi^{\rho}_{P} sends 𝒮ρ(G)\mathcal{S}^{\rho}(G) into 𝒮ρ(L)\displaystyle\mathcal{S}^{\rho}(L). In particular, equation (5.4)(5.4) is valid for our spaces of ρ\rho-Swartz functions on G(k)G(k) and L(k)L(k).

We remark that this definition will be needed in our proof of multiplicativity, Theorem 5.35.3, the discussion after equation (5.16).

Remark 5.2.

This definition of Schwartz spaces agrees with ideas of Braverman–Kazhdan [BK02, GL20] and with the case of standard representation of GLn()GL_{n}(\mathbb{C}). To wit consider G=GL1G=GL_{1}, i.e., the Tate’s setting, and check it for the Φ0=char(Ok)\Phi_{0}=\text{char}(O_{k}), i.e., the corresponding “basic function”. Let Φ=char(Ok)Cc(k)\Phi=\text{char}(O^{*}_{k})\in C^{\infty}_{c}(k^{*}). Now JρJ^{\rho} is just the standard Fourier transform

(5.8) JρΦ(y)=Φ^(y)=kΦ(x)ψ(tr(xy))𝑑x.J^{\rho}\Phi(y)=\hat{\Phi}(y)=\int\limits_{k}\Phi(x)\psi(tr(xy))dx.

It can be easily checked that

Φ0\displaystyle\Phi_{0} =char(Pk1Ok)+Φ^\displaystyle=\rm{char}(P^{-1}_{k}\setminus O_{k})+\hat{\Phi}
Cc(k)+Jρ(Cc(k)),\displaystyle\in C^{\infty}_{c}(k^{*})+J^{\rho}(C_{c}^{\infty}(k^{*})),

where PkP_{k} is the maximal ideal of OkO_{k}.

This simple calculation allows us to prove the following general result:

Proposition 5.3.

The space 𝒮ρ(G)\mathcal{S}^{\rho}(G) contains the ρ\rho-basic function.

Proof.

Note that when L=TL=T is a maximal torus, the ρ\rho-Harish-Chandra transform becomes (a twist of) the Satake transform, and in this case the above diagram (5.5)(5.5) can be extended to the class of almost compact (ac) spherical functions as defined by Wen-Wei Li in [Li17], and we note that the ρ\rho-basic function is amongst this class (see [Sak18]). The above computation for Φ0\Phi_{0} can be extended to show that the function fTstd=char(𝔸n(Ok)Tn(k))f_{T}^{\textrm{std}}=\rm{char}(\mathbb{A}^{n}(O_{k})\cap T_{n}(k)) is also a sum in Cc(Tn(k))+JρT(Cc(Tn(k))C^{\infty}_{c}(T_{n}(k))+J^{\rho_{T}}(C_{c}^{\infty}(T_{n}(k)), where Tn(k)𝔸n(k)T_{n}(k)\hookrightarrow\mathbb{A}^{n}(k) is the standard embedding of a maximal torus Tn𝔾mnT_{n}\cong\mathbb{G}_{m}^{n} of GLnGL_{n} into affine space.

Let Sat:=stdHC\mathrm{Sat}:=\textrm{std}-\rm{HC} be this extended Satake transform. Given a decomposition ϕTnstd=f1+Jstd(f2)\phi_{T_{n}}^{\textrm{std}}=f_{1}+J^{\textrm{std}}(f_{2}), with f1,f2Cc(Tn(k))f_{1},f_{2}\in C^{\infty}_{c}(T_{n}(k)), the commutativity of (5.5) implies that the standard basic function on GLn(k)GL_{n}(k), ϕstd=Sat1(f1)+Sat1(JstdT(f2))=Sat1(f1)+Jstd(Sat1(f2)),\phi^{\rm{std}}=\mathrm{Sat}^{-1}(f_{1})+\mathrm{Sat}^{-1}(J^{\rm{std}_{T}}(f_{2}))=\mathrm{Sat}^{-1}(f_{1})+J^{\rm{std}}(\mathrm{Sat}^{-1}(f_{2})), lies in 𝒮std(GLn(k))\mathcal{S}^{\rm{std}}(GL_{n}(k)) as defined in (5.6)(5.6).

Note that here Sat is an isomorphism of KK-spherical compactly supported functions on G(k)G(k) and the Weyl-invariant compactly supported functions on T(k)/T(𝒪k)T(k)/T(\mathcal{O}_{k}). The basic function on Tn(k)T_{n}(k), and the functions in its decomposition as f1+Jstd(f2)f_{1}+J^{\textrm{std}}(f_{2}) are invariant under permutations of the coordinates, and so the above maps are well-defined in the remarks above.

We can deduce the analogous case for a general ρ\rho from the standard case above as follows: Let TT be a maximal torus in GG with representation ρ\rho of the dual group of GG. One obtains a canonical map ρT~:Tn(k)T(k)\tilde{\rho_{T}}:T_{n}(k)\to T(k) that extends to a map 𝔸n(k)MT(k)\mathbb{A}^{n}(k)\to M_{T}(k), the target of this map being the toric variety constructed in section 22 (see Section 66). The ρ\rho-Schwartz space on T(k)T(k) can be defined as the image of

Cc(𝔸n(k))C(Tn(k))ρ(Cc(𝔸n(k))C(Tn(k))),C^{\infty}_{c}(\mathbb{A}_{n}(k))\cap C^{\infty}(T_{n}(k))\to\rho_{*}(C^{\infty}_{c}(\mathbb{A}_{n}(k))\cap C^{\infty}(T_{n}(k))),

the pushforward by ρT~\tilde{\rho_{T}}. Then the torus basic function ϕTρ\phi^{\rho}_{T} can be expressed as ρ(ϕTnstd).\rho_{*}(\phi^{\rm{std}}_{T_{n}}). Moreover, this pushforward is compatible with the ρT\rho_{T}-Fourier transform on tori, as in diagram (6.4)(6.4). That is,

ϕTρ=ρ(ϕTnstd)\displaystyle\phi^{\rho}_{T}=\rho_{*}(\phi^{\rm{std}}_{T_{n}}) =ρ(f1+JTstd(f2))\displaystyle=\rho_{*}(f_{1}+J_{T}^{\textrm{std}}(f_{2}))
=ρ(f1)+ρ(JTstd(f2))\displaystyle=\rho_{*}(f_{1})+\rho_{*}(J_{T}^{\rm{std}}(f_{2}))
=ρ(f1)+JρT(ρ(f2)),\displaystyle=\rho_{*}(f_{1})+J^{\rho_{T}}(\rho_{*}(f_{2})),

which shows that ϕTρCc(T(k))+JρT(Cc(T(k))\phi^{\rho}_{T}\in C_{c}^{\infty}(T(k))+J^{\rho_{T}}(C^{\infty}_{c}(T(k)). Finally, the commutativity of diagram (6.8)(6.8) allows us to lift this decomposition to a decomposition of the basic function as

ϕρ\displaystyle\phi^{\rho} =Sat1(ϕTρ)\displaystyle=\mathrm{Sat}^{-1}(\phi^{\rho}_{T})
=Sat1(ρ(f1))+Sat1(JρT)(ρ(f2)))\displaystyle=\mathrm{Sat}^{-1}(\rho_{*}(f_{1}))+\mathrm{Sat}^{-1}(J^{\rho_{T}})(\rho_{*}(f_{2})))
=Sat1(ρ(f1))+Jρ(Sat1(ρ(f2)).\displaystyle=\mathrm{Sat}^{-1}(\rho_{*}(f_{1}))+J^{\rho}(\mathrm{Sat}^{-1}(\rho_{*}(f_{2})).


Multiplicativity. As we discussed earlier every theory of LL-functions must satisfy multiplicativity, an axiom that is a theorem for all the Artin LL-functions and is the main tool in computing γ\gamma-factors and LL-functions. To explain, let P=NLP=NL be a parabolic subgroup of GG with a Levi subgroup LL, uniquely fixed such that LTL\supset T, the maximal torus of GG fixed in our construction throughout. Let σ\sigma be an irreducible admissible representation of L(k)L(k) and let ρ\rho be a finite dimensional complex representation of G^\hat{G} and ρL=ρ|L^\rho_{{}_{L}}=\rho|\hat{L} as before. For each irreducible admissible representation σ\sigma of L(k)L(k), we can define the γ\gamma-factors γ(s,σ,ρL)\gamma(s,\sigma,\rho_{{}_{L}}) and γ(s,IndP(k)G(k)σ,ρ)\gamma(s,\text{Ind}_{P(k)}^{G(k)}\sigma,\rho). Multiplicativity states that:

(5.9) γ(s,IndP(k)G(k)σ,ρ)=γ(s,σ,ρL).\gamma(s,\text{Ind}_{P(k)}^{G(k)}\sigma,\rho)=\gamma(s,\sigma,\rho_{{}_{L}}).

Here we suppress the dependence of the factors on the non-trivial additive character of kk.

We note that, since the induced representation IndP(k)G(k)σ\text{Ind}^{G(k)}_{P(k)}\sigma may not be irreducible, the γ\gamma-factor γ(s,IndP(k)G(k)σ,ρ)\gamma(s,\text{Ind}^{G(k)}_{P(k)}\sigma,\rho) is defined to be γ(s,π,ρ)\gamma(s,\pi,\rho), where π\pi is any irreducible constituent of IndP(k)G(k)σ\text{Ind}^{G(k)}_{P(k)}\sigma. The γ\gamma-factor will not depend on the choice of π\pi as the proof below establishes.

A proof of (5.9) is usually fairly hard for γ\gamma-factors defined by Rankin–Selberg method [JPSS83, Sou93]. On the contrary, (5.9) is a general result within the Langlands–Shahidi [Sha10] method with a very natural proof.

Our aim here is to give a general proof of (5.9) within the Braverman–Kazhdan/Ngo and Lafforgue programs using (5.4) and Int(K)(K)-invariance of JρJ^{\rho}. It follows the arguments given in [GJ72]. We now proceed to give a proof of (5.9) which we formally state as:

Theorem 5.4.

Let σ\sigma be an irreducible admissible representation of L(k)L(k) and let

Π=IndP(k)G(k)σ\Pi=\text{Ind}\,^{G(k)}_{P(k)}\sigma. Let ρ\rho be an irreducible finite dimensional complex representation of G^\hat{G} and let ρL=ρ|L^\rho_{{}_{L}}=\rho|\hat{L}. Assume the validity of (5.4) and thus (5.5) for Schwartz functions, and that Jρ(k1xk)=Jρ(x)J^{\,\rho}(k^{-1}xk)=J^{\,\rho}(x), kKk\in K, a maximal compact subgroup of G(k)G(k) satisfying G(k)=P(k)KG(k)=P(k)K, which follows from the expected Int(G)(G)-invariance of the kernel JρJ^{\,\rho}. Then

γ(s,Π,ρ)=γ(s,σ,ρL).\gamma(s,\Pi,\rho)=\gamma(s,\sigma,\rho_{{}_{L}}).
Proof.

Let Π~=IndP(k)G(k)σ~\widetilde{\Pi}=\text{Ind}^{G(k)}_{P(k)}\widetilde{\sigma} be the contragredient of Π\Pi. Choose vΠv\in\Pi and v~Π~\widetilde{v}\in\widetilde{\Pi}. Then a

matrix coefficient for Π\Pi can be written as

(5.10) f(g)\displaystyle f(g) =\displaystyle= Π(g)v,v~\displaystyle\langle\Pi(g)v,\widetilde{v}\rangle
=\displaystyle= Kv(kg),v~(k)0𝑑k,\displaystyle\int\limits_{K}\langle v(kg),\widetilde{v}(k)\rangle_{0}\ dk,

where v()v(\cdot) and v~()\widetilde{v}(\cdot) are values of the functions in Π\Pi and Π~\widetilde{\Pi}, respectively, and σ(l)w,w~0\langle\sigma(l)w,\widetilde{w}\rangle_{0} is a matrix coefficient for σ\sigma, wσw\in\sigma and w~σ~\widetilde{w}\in\widetilde{\sigma}. Let Φ𝒮ρ(G)\Phi\in\mathcal{S}^{\rho}(G) be a ρ\rho-Schwartz function in C(G(k))C^{\infty}(G(k)). We absorb the complex number ss in ff by replacing π\pi by π|ν|s\pi\otimes|\nu|^{s} and then ignoring it throughout the proof.

Now we have the zeta function

(5.11) Z(Φ,f)=Φ(g)f(g)δG,ρ1/2(g)𝑑g.Z(\Phi,f)=\int\Phi(g)f(g)\delta^{1/2}_{G,\rho}(g)dg.

As explained earlier, the shift δG,ρ1/2\delta^{1/2}_{G,\rho} allows us to get the precise LL-function at ss, rather than a shift of ss, when Φ\Phi is the basic function of σ\sigma for a spherical representation σ\sigma. Using (5.10), (5.11) equals

Z(Φ,f)=\displaystyle Z(\Phi,f)= K×G(k)Φ(g)v(kg),v~(k)0δG,ρ1/2(g)𝑑k𝑑g\displaystyle\int\limits_{K\times G(k)}\Phi(g)\langle v(kg),\widetilde{v}(k)\rangle_{0}\ \delta^{1/2}_{G,\rho}(g)dk\,dg
(5.12) =\displaystyle= K×G(k)Φ(k1g)v(g),v~(k)0δG,ρ1/2(g)𝑑k𝑑g.\displaystyle\int\limits_{K\times G(k)}\Phi(k^{-1}g)\langle v(g),\widetilde{v}(k)\rangle_{0}\ \delta^{1/2}_{G,\rho}(g)dk\,dg.

Write g=nlhg=nlh, nN(k)n\in N(k), lL(k)l\in L(k), hKh\in K. Then

dg=δP1(l)dndldh.dg=\delta^{-1}_{P}(l)dn\,dl\,dh.

With notation as in [GJ72], define:

(5.13) (hΦk1)(x):=Φ(k1xh).(h\cdot\Phi\cdot k^{-1})(x):=\Phi(k^{-1}xh).

Thus

(5.14) Z(Φ,f)=N(k)×L(k)×K×K(hΦk1)(nl)δP1/2(l)σ(l)v(h),v~(k)0δG,ρ1/2(l)δP1(l)𝑑n𝑑l𝑑h𝑑k.Z(\Phi,f)=\int\limits_{N(k)\times L(k)\times K\times K}(h\cdot\Phi\cdot k^{-1})(nl)\ \delta^{1/2}_{P}(l)\langle\sigma(l)v(h),\widetilde{v}(k)\rangle_{0}\ \delta^{1/2}_{G,\rho}(l)\delta^{-1}_{P}(l)dn\,dl\,dh\,dk.

Recall the HC–transform (hΦk1)P(h\cdot\Phi\cdot k^{-1})_{P}:

(hΦk1)P(l)=δP1/2(l)N(k)(hΦk1)(nl)𝑑n.(h\cdot\Phi\cdot k^{-1})_{P}(l)=\delta^{-1/2}_{P}(l)\int\limits_{N(k)}(h\cdot\Phi\cdot k^{-1})(nl)dn.

Then (5.14) equals

L(k)×K×KνG/L,ρ1/2(l)(hΦk1)P(l)σ(l)v(h),v~(k)0δL,ρL1/2(l)𝑑l𝑑h𝑑k\displaystyle\int\limits_{L(k)\times K\times K}\nu_{G/L,\rho}^{1/2}(l)(h\cdot\Phi\cdot k^{-1})_{P}(l)\langle\sigma(l)v(h),\widetilde{v}(k)\rangle_{0}\ \delta^{1/2}_{L,\rho_{{}_{L}}}(l)dl\,dh\,dk
(5.15) =L(k)×K×K(hΦk1)Pρ(l)σ(l)v(h),v~(k)0δL,ρL1/2(l)𝑑l𝑑h𝑑k.\displaystyle=\int\limits_{L(k)\times K\times K}(h\cdot\Phi\cdot k^{-1})^{\rho}_{P}(l)\langle\sigma(l)v(h),\widetilde{v}(k)\rangle_{0}\ \delta_{L,\rho_{{}_{L}}}^{1/2}(l)dl\,dh\,dk.

Since KK is compact and vv and v~\widetilde{v} are smooth functions, there exist matrix coefficients fiL(l)f^{L}_{i}(l) of σ\sigma and continuous functions λi\lambda_{i} on K×KK\times K such that

σ(l)v(h),v~(k)0=ifiL(l)λi(h,k).\langle\sigma(l)v(h),\widetilde{v}(k)\rangle_{0}=\sum\limits_{i}f^{L}_{i}(l)\lambda_{i}(h,k).

Similarly, there are Schwartz functions Φj\Phi_{j} in 𝒮ρ(G)\mathcal{S}^{\rho}(G) and continuous symmetric functions μj\mu_{j} on K×KK\times K such that

(5.16) hΦk1=jΦjμj(h,k).h\cdot\Phi\cdot k^{-1}=\sum\limits_{j}\Phi_{j}\mu_{j}(h,k).

This is clearly true if ΦCc(G(k))\Phi\in C^{\infty}_{c}(G(k)), since it will then be uniformly smooth. Otherwise, using (5.16) we have

(hΦk1)=jΦ^jμj(h,k),(h\cdot\Phi\cdot k^{-1})^{\wedge}=\sum\limits_{j}\hat{\Phi}_{j}\mu_{j}(h,k),

where Φ^:=JρΦ\hat{\Phi}:=J^{\,\rho}\Phi for simplicity. But Lemma 5.6, proved later, implies

kΦ^h1\displaystyle k\cdot\hat{\Phi}\cdot h^{-1} =\displaystyle\!=\! jΦ^jμj(k,h)\displaystyle\sum\limits_{j}\hat{\Phi}_{j}\mu_{j}(k,h)

for all hh and kk in KK and thus (5.16) holds for all ΦSρ(G)\Phi\in S^{\,{}^{\rho}}(G). Consequently

(hΦk1)Pρ(l)=jΦj,Pρ(l)μj(h,k)(h\cdot\Phi\cdot k^{-1})^{\rho}_{P}(l)=\sum\limits_{j}\Phi_{j,P}^{\rho}(l)\mu_{j}(h,k)

with Φj,Pρ𝒮ρL(L)\Phi_{j,P}^{\rho}\in\mathcal{S}^{\rho_{{}_{L}}}(L) by Proposition 5.1. Let

cij=K×Kλiμj(h,k)𝑑h𝑑k.c_{ij}=\int_{K\times K}\lambda_{i}\mu_{j}(h,k)dhdk.

Then we have

(5.17) Z(Φ,f)=i,jcijZ(Φj,Pρ,fiL)Z(\Phi,f)=\sum\limits_{i,j}c_{ij}Z(\Phi^{\rho}_{j,P},f^{L}_{i})

For simplicity of notation, let for each Φ𝒮ρ(G)\Phi\in\mathcal{S}^{\rho}(G), Φ^:=JρΦ\hat{\Phi}:=J^{{}^{\rho}}\Phi. We now calculate

Z(Φ^,fˇ)=Φ^(g)fˇ(g)δG,ρ1/2(g)|ν(g)|𝑑g.Z(\hat{\Phi},\check{f})=\int\hat{\Phi}(g)\check{f}(g)\delta^{1/2}_{G,\rho}(g)\ |\nu(g)|\ dg.

One needs to be careful since the involution gg1g\mapsto g^{-1} will now play an important role. We should also point out that |ν(g)||\nu(g)| needs to be inserted to take into account the appearance of 1s1-s in the left hand side of the functional equation as it is the case in equation (1.1) for GLnGL_{n}. We note that |ν(g)|=|ν(l)|=|νL(l)||\nu(g)|=|\nu(l)|=|\nu_{L}(l)| and that s-s will appear as π~|ν()|s\widetilde{\pi}\otimes|\nu(\hbox{})|^{-s}, and thus included in π~\widetilde{\pi} as ss did in π\pi as π|ν()|s\pi\otimes|\nu(\hbox{})|^{s}. Note that in the case

of GLnGL_{n} and ρ=std\rho=\text{std}, ν=det\nu=\det, and δG,std1/2=|det|n12\delta_{G,\text{std}}^{1/2}=|\det|^{\frac{n-1}{2}} as reflected in (1.1).

We have

Z(Φ^,fˇ)\displaystyle Z(\hat{\Phi},\check{f}) =\displaystyle= G(k)Φ^(g)f(g1)δG,ρ1/2(g)|ν(g)|𝑑g\displaystyle\int\limits_{G(k)}\hat{\Phi}(g)f(g^{-1})\delta^{1/2}_{G,\rho}(g)\ |\nu(g)|\ dg
=\displaystyle= G(k)×KΦ^(g)v(kg1),v~(k)0δG,ρ1/2(g)|ν(g)|𝑑g𝑑k.\displaystyle\int\limits_{G(k)\times K}\hat{\Phi}(g)\langle v(kg^{-1}),\widetilde{v}(k)\rangle_{0}\ \delta^{1/2}_{G,\rho}(g)\ |\nu(g)|\ dg\ dk.

Changing gg to gkgk, (5.18) equals

(5.19) =G(k)×KΦ^(gk)v(g1),v~(k)0δG,ρ1/2(g)|ν(g)|𝑑g𝑑k.=\int\limits_{G(k)\times K}\hat{\Phi}(gk)\langle v(g^{-1}),\widetilde{v}(k)\rangle_{0}\ \delta^{1/2}_{G,\rho}(g)\ |\nu(g)|\ dg\ dk.

Write g=h1lng=h^{-1}ln, nN(k)n\in N(k), lL(k)l\in L(k), hKh\in K. Then

(5.20) dg=d(g1)=δP(l)dhdldndg=d(g^{-1})=\delta_{P}(l)\ dh\ dl\ dn

and (5.19) equals

(5.21) =K×L(k)×N(k)×KΦ^(h1lnk)v(l1h),v~(k)0δG,ρ1/2(l)δP(l)|ν(l)|𝑑h𝑑l𝑑n𝑑k.=\int\limits_{K\times L(k)\times N(k)\times K}\hat{\Phi}(h^{-1}lnk)\langle v(l^{-1}h),\widetilde{v}(k)\rangle_{0}\ \delta^{1/2}_{G,\rho}(l)\delta_{P}(l)\ |\nu(l)|\ dh\ dl\ dn\ dk.
=K×L(k)×K(δP1/2(l)N(k)(kΦ^h1)(ln)𝑑n)v(l1h),v~(k)0δG,ρ1/2(l)δP1/2(l)|νL(l)|𝑑h𝑑l𝑑k\displaystyle=\int\limits_{K\times L(k)\times K}(\delta^{1/2}_{P}(l)\int\limits_{N(k)}(k\cdot\hat{\Phi}\cdot h^{-1})(ln)dn)\langle v(l^{-1}h),\widetilde{v}(k)\rangle_{0}\ \delta^{1/2}_{G,\rho}(l)\delta^{1/2}_{P}(l)\ |\nu_{L}(l)|\ dh\ dl\ dk
=K×L(k)×K(δP1/2(l)N(k)(kΦ^h1)(nl)𝑑n)v(l1h),v~(k)0νG/L,ρ1/2(l)δP1/2(l)δL,ρL1/2(l)|νL(l)|𝑑h𝑑l𝑑k\displaystyle=\int\limits_{K\times L(k)\times K}(\delta^{-1/2}_{P}(l)\int\limits_{N(k)}(k\cdot\hat{\Phi}\cdot h^{-1})(nl)dn)\langle v(l^{-1}h),\widetilde{v}(k)\rangle_{0}\ \nu^{1/2}_{G/L,\rho}(l)\ \delta^{1/2}_{P}(l)\ \delta^{1/2}_{L,\rho_{L}}(l)\ |\nu_{L}(l)|\ dh\ dl\ dk
=K×L(k)×KνG/L,ρ1/2(l)(kΦ^h1)P(l)v(l1h),v~(k)0δL,ρL1/2(l)|νL(l)|δP1/2(l)𝑑h𝑑l𝑑k,\displaystyle=\int\limits_{K\times L(k)\times K}\nu^{1/2}_{G/L,\rho}(l)(k\cdot\hat{\Phi}\cdot h^{-1})_{P}(l)\langle v(l^{-1}h),\widetilde{v}(k)\rangle_{0}\ \delta^{1/2}_{L,\rho_{L}}(l)\ |\nu_{L}(l)|\ \delta^{1/2}_{P}(l)\ dh\ dl\ dk,

which finally equals

(5.22) K×L(k)×K(kΦ^h1)Pρ(l)δP1/2(l)σ(l1)v(h),v~(k)0δL,ρL1/2(l)|νL(l)|δP1/2(l)𝑑h𝑑l𝑑k.\int\limits_{K\times L(k)\times K}(k\cdot\hat{\Phi}\cdot h^{-1})^{\rho}_{P}(l)\ \delta^{-1/2}_{P}(l)\langle\sigma(l^{-1})v(h),\widetilde{v}(k)\rangle_{0}\ \delta^{1/2}_{L,\rho_{L}}(l)\ |\nu_{L}(l)|\ \delta^{1/2}_{P}(l)\ dh\ dl\ dk.

Again, for simplicity for each Φ𝒮ρ(G)\Phi\in\mathcal{S}^{\rho}(G), let Φ^\hat{\Phi} denote its ρ\rho-Fourier transform

(5.23) Φ^(x)=(JρΦˇ)(x).\hat{\Phi}(x)=(J^{\rho}*\check{\Phi})(x).

To proceed, we need:

Lemma 5.5.

Let kk and hh be in KK. Then

(kΦh1)=hΦ^k1.({k\cdot\Phi\cdot h^{-1}})^{\wedge}=h\cdot\hat{\Phi}\cdot k^{-1}.
Proof.

We have

(kΦh1)(x)\displaystyle({k\cdot\Phi\cdot h^{-1}})^{\wedge}(x) =\displaystyle= Mρ(k)(kΦh1)(y)Jρ(xy)𝑑y\displaystyle\int\limits_{M^{\rho}(k)}(k\cdot\Phi\cdot h^{-1})(y)J^{\rho}(xy)dy
=\displaystyle= Φ(h1yk)Jρ(xy)𝑑y\displaystyle\int\Phi(h^{-1}yk)J^{\rho}(xy)dy
=\displaystyle= Φ(y)Jρ(xhyk1)𝑑y,\displaystyle\int\Phi(y)J^{\rho}(xhyk^{-1})dy,

since hh and kk are in KK whose modulus character is 1. Now using Int(G)(G)-invariance of

JρJ^{{}^{\rho}}, (5.24) equals

=\displaystyle= Φ(y)Jρ(k1xhy)𝑑y\displaystyle\int\Phi(y)J^{\rho}(k^{-1}xhy)dy
=\displaystyle= Φ^(k1xh)\displaystyle\hat{\Phi}(k^{-1}xh)
=\displaystyle= (hΦ^k1)(x),\displaystyle(h\cdot\hat{\Phi}\cdot k^{-1})(x),

completing the proof. ∎

Remark 5.6.

In terms of JρJ^{\rho} we have proved:

(5.25) Jρ(kΦh1)=h(JρΦˇ)k1J^{\rho}*(k\cdot\Phi\cdot h^{-1})^{\vee}=h\cdot(J^{\rho}*\check{\Phi})\cdot k^{-1}

We now apply Lemma 5.5 to equation (5.22) to get:

(5.26) Z(Φ^,fˇ)=L(k)×K×K((hΦk1))Pρ(l)σ(l1)v(h),v~(k)0δL,ρL1/2(l)|νL(l)|𝑑l𝑑h𝑑k.Z(\hat{\Phi},\check{f})=\int\limits_{L(k)\times K\times K}{(({h\cdot\Phi\cdot k^{-1}})^{\wedge}})^{\rho}_{P}(l)\langle\sigma(l^{-1})v(h),\widetilde{v}(k)\rangle_{0}\ \delta^{1/2}_{L,\rho_{{}_{L}}}(l)\ |\nu_{L}(l)|\ dl\,dh\,dk.

But

(hΦk1)\displaystyle({h\cdot\Phi\cdot k^{-1}})^{\wedge} =\displaystyle= (jΦj)μj(h,k)\displaystyle(\sum\limits_{j}\Phi_{j})^{\wedge}\cdot\mu_{j}(h,k)
=\displaystyle= jΦ^jμj(h,k)\displaystyle\sum\limits_{j}\hat{\Phi}_{j}\cdot\mu_{j}(h,k)

and

σ(l1)v(h),v~(k)0\displaystyle\langle\sigma(l^{-1})v(h),\widetilde{v}(k)\rangle_{0} =\displaystyle= ifiL(l1)λi(h,k)\displaystyle\sum\limits_{i}f^{L}_{i}(l^{-1})\cdot\lambda_{i}(h,k)
=\displaystyle= ifˇiL(l)λi(h,k)\displaystyle\sum\limits_{i}\check{f}^{L}_{i}(l)\cdot\lambda_{i}(h,k)

and therefore (5.26) equals

(5.27) i,jcijL(k)(JρΦj)Pρ(l)fˇiL(l)δL,ρL1/2(l)|νL(l)|𝑑l.\sum\limits_{i,j}c_{ij}\int\limits_{L(k)}(J^{\rho}\Phi_{j})^{\rho}_{P}(l)\check{f}^{L}_{i}(l)\delta^{1/2}_{L,\rho_{{}_{L}}}(l)\ |\nu_{L}(l)|\ dl.

We can now apply the commutativity of ρ\rho–Harish–Chandra transform and Fourier transforms JρJ^{\rho} and JρLJ^{\rho_{{}_{L}}}, i.e., equation (5.4) to conclude that (5.27) equals

(5.28) i,jcijL(k)JLρ(Φj,Pρ)(l)fˇiL(l)δL,ρL1/2(l)|νL(l)|𝑑l.\sum\limits_{i,j}c_{ij}\int\limits_{L(k)}J^{{{}^{\rho}_{{}_{L}}}}(\Phi^{\rho}_{j,P})(l)\check{f}_{i}^{L}(l)\delta^{1/2}_{L,\rho_{{}_{L}}}(l)\ |\nu_{L}(l)|\ dl.

But

Z(Φ^,fˇ)=γ(Π,ρ)Z(Φ,f)Z(\hat{\Phi},\check{f})=\gamma(\Pi,\rho)Z(\Phi,f)

by the functional equation for GG. On the other hand the functional equation for LL gives (5.28) as

γ(σ,ρL)i,jcijL(k)Φj,Pρ(l)fiL(l)δL,ρL1/2(l)𝑑l\gamma(\sigma,\rho_{{}_{L}})\sum\limits_{i,j}c_{ij}\int\limits_{L(k)}\Phi_{j,P}^{\rho}(l)f^{L}_{i}(l)\delta^{1/2}_{L,\rho_{{}_{L}}}(l)dl

which equals

=\displaystyle= γ(σ,ρL)i,jcijZ(Φj,Pρ,fiL)\displaystyle\!\!\!\gamma(\sigma,\rho_{{}_{L}})\sum\limits_{i,j}c_{ij}Z(\Phi_{j,P}^{\rho},f^{L}_{i})
=\displaystyle= γ(σ,ρL)Z(Φ,f)\displaystyle\!\!\!\gamma(\sigma,\rho_{{}_{L}})Z(\Phi,f)

by (5.17). The equality

γ(IndP(k)G(k)σ,ρ)=γ(σ,ρL)\gamma(\text{Ind}^{G(k)}_{P(k)}\sigma,\rho)=\gamma(\sigma,\rho_{{}_{L}})

is now immediate. ∎


5.3. The case of GLnGL_{n}.

We now determine νG/L\nu_{{}_{G/L}} in the case G=GLnG=GL_{n} and ρ=std\rho=\text{std}, i.e., that of Godement–Jacquet [GJ72] and show that it agrees with calculations in Lemma 3.4.0 of loc. cit., after a suitable normalization. We thus assume P=NLP=NL is the standard maximal parabolic subgroup of GLnGL_{n}, containing the subgroup of upper triangular elements BB, NBN\subset B, with L=GLn×GLn′′,n=n+n′′L=GL_{n^{\prime}}\times GL_{n^{{}^{\prime\prime}}},n=n^{\prime}+n^{{}^{\prime\prime}}. Recall that we need to determine ν,G/Lstd=δG,std/δL,std\nu_{{}_{G/L},\text{std}}=\delta_{G,\text{std}}\,/\,\delta_{L,\text{std}}. But for g=diag(g,g′′)Lg=\text{diag}(g^{\prime},g^{{}^{\prime\prime}})\in L

|detg|n′′δG1/2(g,g′′)\displaystyle|\det{g^{\prime}}|^{-n^{{}^{\prime\prime}}}\cdot\,\delta_{G}^{1/2}(g^{\prime},g^{{}^{\prime\prime}}) =\displaystyle= |detg|n′′|detg|12(n+n′′1)\displaystyle|\det{g^{\prime}}|^{-n^{{}^{\prime\prime}}}\,|\det g|^{\frac{1}{2}(n^{\prime}+n^{{}^{\prime\prime}}-1)}
=\displaystyle= |detg|12(n1)|detg′′|12(n′′1)|detg|n′′/2|detg′′|n/2\displaystyle|\det g^{\prime}|^{\frac{1}{2}(n^{\prime}-1)}\,|\det g^{{}^{\prime\prime}}|^{\frac{1}{2}(n^{{}^{\prime\prime}}-1)}\cdot|\det g^{\prime}|^{-n^{{}^{\prime\prime}}/2}\ |\det g^{{}^{\prime\prime}}|^{n^{\prime}/2}
=\displaystyle= δL1/2(g,g′′)|detg|n′′/2|detg′′|n/2.\displaystyle\delta_{L}^{1/2}(g^{\prime},g^{{}^{\prime\prime}})\cdot\,|\det g^{\prime}|^{-n^{{}^{\prime\prime}}/2}\ |\det g^{{}^{\prime\prime}}|^{n^{\prime}/2}.

Thus

(5.29) νG/L1/2(g,g′′)=|detg|n′′/2|detg′′|n/2.\nu^{1/2}_{{}_{G/L}}(g^{\prime},g^{{}^{\prime\prime}})=|\det g^{\prime}|^{n^{{}^{\prime\prime}}\!/2}\,\,|\det g^{{}^{\prime\prime}}|^{n^{\prime}/2}.

Moreover

(5.30) δP(g,g′′)=|detg|n′′|detg′′|n\delta_{P}(g^{\prime},g^{{}^{\prime\prime}})=|\det g^{\prime}|^{n^{{}^{\prime\prime}}}\cdot\,|\det g^{{}^{\prime\prime}}|^{-n^{\prime}}

and thus

(5.31) νG/L1/2δP1/2(g,g′′)=|detg′′|n.\nu^{1/2}_{{}_{G/L}}\,\delta^{-1/2}_{P}(g^{\prime},g^{{}^{\prime\prime}})=|\det g^{{}^{\prime\prime}}|^{n^{\prime}}.

We now verify that Lemma 3.4.0 of [GJ72] is equivalent to our commutative diagram (5.5).


Let J=JstdJ=J^{\text{std}} be the standard Fourier transform for Cc(Mn(k))C^{\infty}_{c}(M_{n}(k)) and JLJ_{L} its restriction to Cc(Mn(k)×Mn′′(k))C^{\infty}_{c}(M_{n^{\prime}}(k)\times M_{n^{{}^{\prime\prime}}}(k)). With notation as in pages 37–38 of [GJ72],

(5.32) φΦ(x,y)=knn′′Φ((xu0y))𝑑u,\varphi_{{\Phi}}(x,y)=\int_{k^{n^{\prime}n^{{}^{\prime\prime}}}}\Phi(\left(\begin{array}[]{cc}x&u\\ 0&y\end{array}\right))\ du,

where uMn×n′′(k)u\in M_{n^{\prime}\times n^{{}^{\prime\prime}}}(k) and ΦCc(Mn(k))\Phi\in C^{\infty}_{c}(M_{n}(k)), is the analogue of our HC–transform. In fact, (5.32) can be written as

(5.37) φΦ(x,y)\displaystyle\varphi_{\Phi}(x,y) =\displaystyle= Φ((Iuy10I)(x00y))𝑑u\displaystyle\int\Phi(\left(\begin{array}[]{cc}I&uy^{-1}\\ 0&I\end{array}\right)\left(\begin{array}[]{cc}x&0\\ 0&y\end{array}\right))\ du
(5.38) =\displaystyle= |dety|nN(k)Φ(nl)𝑑n,\displaystyle|\det y|^{n^{\prime}}\int\limits_{N(k)}\Phi(nl)dn,

=νG/L1/2(l)δP1/2(l)N(k)Φ(nl)𝑑n=\nu^{1/2}_{G/L}(l)\ \delta^{-1/2}_{P}(l)\int\limits_{N(k)}\Phi(nl)\ dn

by (5.31), where N=Mn×n′′N=M_{n^{\prime}\times n^{{}^{\prime\prime}}} and l=diag(x,y)l=\text{diag}(x,y).

In the notation of [GJ72], Lemma 3.4.0 of [GJ72] states that

(5.39) φΦ^(x,y)=φΦ^(x,y),\varphi_{\hat{\Phi}}(x,y)=\widehat{\varphi_{\Phi}}(x,y),

for ΦCc(Mn(F))\Phi\in C^{\infty}_{c}(M_{n}(F)) with Φ^\hat{\Phi} its standard Fourier transform.

Then by (5.33) the right hand side of (5.34) equals

(5.42) φΦ^(x,y)\displaystyle\widehat{\varphi_{\Phi}}(x,y) =\displaystyle= JL(Φ((xu0y))𝑑u)\displaystyle J_{L}(\int\Phi(\left(\begin{array}[]{cc}x&u\\ 0&y\end{array}\right))\ du)
=\displaystyle= JL(|dety|nN(k)Φ(nl)𝑑n)\displaystyle J_{L}(|\det y|^{n^{\prime}}\int\limits_{N(k)}\Phi(nl)\ dn)
=\displaystyle= JL(νG/L1/2(l)ΦP(l))\displaystyle J_{L}(\nu^{1/2}_{G/L}(l)\Phi_{P}(l))
=\displaystyle= JL(ΦPstd(l))\displaystyle J_{L}(\Phi^{\text{std}}_{P}(l))

by (5.31), where

ΦP(l)=δP1/2(l)N(k)Φ(nl)𝑑n\Phi_{P}(l)=\delta^{-1/2}_{P}(l)\int\limits_{N(k)}\Phi(nl)\ dn

as defined by (5.1).

Similarly from the left hand side of (5.34), using a change of variables as in (5.33), we have

(5.46) φΦ^(x,y)\displaystyle\varphi_{\hat{\Phi}}(x,y) =\displaystyle= Φ^((xu0y))𝑑u\displaystyle\int\hat{\Phi}(\left(\begin{array}[]{cc}x&u\\ 0&y\end{array}\right))\ du
=\displaystyle= |dety|nN(k)Φ^(nl)𝑑n\displaystyle|\det y|^{n^{\prime}}\int\limits_{N(k)}\hat{\Phi}(nl)\ dn
=\displaystyle= νG/L1/2(l)(Φ^)P(l)\displaystyle\nu^{1/2}_{G/L}(l)\ (\hat{\Phi})_{P}(l)
=\displaystyle= (JΦ)Pstd(l).\displaystyle(J\Phi)^{\text{std}}_{P}(l).

Thus (5.34) is equivalent to (5.4) for GLnGL_{n} and ρ=std\rho=\text{std}.

5.4. Inductive definition of JρJ^{\rho}

In the introduction we mentioned that multiplicativity plus a definition of Fourier transform that acts through the correct scalar factors equal to the gamma factors on supercuspidal representations/characters, is enough to characterize the full Fourier transform. Indeed, if we assume that JρJ^{\rho} is a good distribution in the sense of Braverman-Kazhan [BK10], then we can identify JρJ^{\rho} with a rational, scalar valued function πγ(ρ,π)\pi\mapsto\gamma(\rho,\pi), where γ(ρ,π)\gamma(\rho,\pi) is defined by Jρπ=γ(ρ,π)πJ^{\rho}\star\pi=\gamma(\rho,\pi)\pi.

Our results on mulitplicativity allow in principle for us to construct in an inductive fashion a distribtion JρJ^{\rho} on GG by formally inducing from JρLJ^{\rho_{L}} for each conjugacy class of Levi subgroup LGL\subset G. In fact, our setup and definitions, culminating in Theorem 5.35.3, are normalized so as to make induction of representations adjoint to our ρ\rho-Harish-Chandra transform, that is, we have an equality

Jρ,IndL(θ)=JρL,θ.\langle J^{\rho},\text{Ind}_{L}(\theta)\rangle=\langle J^{\rho_{L}},\theta\rangle.

Here θ\theta is a supercuspidal character of a representation on LL. The ajdunction allows us to identify the JρJ^{\rho} and JρLJ^{\rho_{L}} actions on the Bernstein components of IndL(σ)=π\text{Ind}_{L}(\sigma)=\pi on G(k)G(k) and the Bernstein component of σ\sigma on L(k)L(k), respectively. In 5.35.3, we started with an assumption of knowledge of JρJ^{\rho} and JρLJ^{\rho_{L}} and we showed that this is equivalent to an equality of gamma factors. However, the gamma factors determine the distribution uniquely, and so one can in principle characterize completely a distribution JρJ^{\rho} by specifying its action on supercuspidal representations on G(k)G(k), and postulating multiplicativity as an axiom. More concretely, if we inductively know JρLJ^{\rho_{L}} for conjugacy classes of parabolic subgroups LL, we may formally induce to provide a definition of JρJ^{\rho} with a correct action, at least on functions whose spectral decomposition consists solely of induced data from LL:

IndL(JρL),f=Jρ,HC(f).\langle\text{Ind}_{L}(J^{\rho_{L}}),f\rangle=\langle J^{\rho},HC(f)\rangle.

The distribution IndL(JρL)\text{Ind}_{L}(J^{\rho_{L}}) can be a priori defined by the above in order to meet this adjunction, and in fact JρLJ^{\rho_{L}} will then be represented by the conjugation-invariant function

IndL(JρL):x|DG(x)|12y|DL(y)|12JρL(y)\text{Ind}_{L}(J^{\rho_{L}}):x\mapsto|D_{G}(x)|^{\frac{-1}{2}}\sum_{y}|D_{L}(y)|^{\frac{1}{2}}J^{\rho_{L}}(y)

where the yy are chosen representatives of L(k)L(k)-conjugacy classes of elements that are G(k)G(k)-conjugate to xx, and DGD_{G} and DLD_{L} are the respective discriminant functions on GG and LL. (Here we are identifying JρLJ^{\rho_{L}} with the invariant function representing it).

That IndL(JρL)\text{Ind}_{L}(J^{\rho_{L}}) satisfies the first adjunction, and therefore multiplicativity, follows from the formula for the trace, and the expression of the distribution character Θπ=IndL(Θσ)\Theta_{\pi}=\text{Ind}_{L}(\Theta_{\sigma}) in terms of Θσ\Theta_{\sigma}, adapted to the ρ\rho-setting.

6. Example: The case of Tori and unramified data

We now consider the case of tori, which for present purposes we assume are split. Let TT be a split torus over kk. When TT is a maximal split torus in a reductive group GG, the upcoming discussion gives the first term of the inductive construction defining the Fourier transform for L=TL=T, with minimal parabolic P0=P=LN=TNP_{0}=P=LN=TN which is a Borel subgroup. Let ρ=ρT\rho=\rho_{T} be a finite dimensional represenation of T^\hat{T}. Our notation is justified if we assume ρT=ρ|T^\rho_{T}=\rho|\hat{T}, where ρ\rho is a representation of G^\hat{G}. Let n=dim ρTn=\text{dim }\rho_{T}. Then

ρT:T^GLn().\rho_{T}:\hat{T}\to GL_{n}(\mathbb{C}).

Write

(6.1) ρT=μ1μn,\rho_{T}=\mu_{1}\oplus\cdots\mu_{n},

where the μi,1in\mu_{i},1\leq i\leq n, are the weights of ρT\rho_{T}. We note that they are not necessarily distinct. If we realize these weights of T^\hat{T} as co-characters of TT, we get a map ρT~:𝔾mnT\tilde{\rho_{T}}:\mathbb{G}_{m}^{n}\to T (defined over kk, as TT is split), which being dual to ρT\rho_{T}, is given by (c.f. [Ngô20])

ρT~(x1,,xn)=μ1(x1)μn(xn).\tilde{\rho_{T}}(x_{1},\ldots,x_{n})=\mu_{1}(x_{1})\cdots\mu_{n}(x_{n}).

We can extend this to a monoid homomorphism

ρT~:𝔸nMρT,\tilde{\rho_{T}}:\mathbb{A}^{n}\to M^{\rho_{T}},

where MρTM^{\rho_{T}} is the corresponding toric variety. As in [Ngô20], define the trace function h:𝔸n𝔸h:\mathbb{A}^{n}\to\mathbb{A} by

h((xi))=ixih((x_{i}))=\sum_{i}x_{i}

and set

hψ:knh_{\psi}:k^{n}\to\mathbb{C}^{*}

by xψ(h(x))x\mapsto\psi(h(x)), where ψ\psi is our fixed non-trivial character of kk.


Denote by JstdJ^{\text{std}} the kernel

Jstd(g)=ψ(tr(g))|detg|ndgJ^{\text{std}}(g)=\psi(\text{tr}(g))|\text{det}g|^{n}dg

for gGLn(k)g\in GL_{n}(k) as defined in Section 1, i.e., the standard Fourier transform on Mn(k)M_{n}(k). We use again JstdJ^{\text{std}} for its restrition to 𝔸n\mathbb{A}^{n}, the monoid for Tn=𝔾mnT_{n}=\mathbb{G}_{m}^{n}.


In [Ngô20] Ngo defines the kernel JρTJ^{\rho_{T}} for the Fourier transform on TT by

(6.2) JρT(t)=ρT~1(t)hψ(x)𝑑x,J^{\rho_{T}}(t)=\int_{\widetilde{\rho_{T}}^{-1}(t)}h_{\psi}(x)dx,

which equals to

(6.3) JρT(t)=xU(k)hψ(xt)𝑑xJ^{\rho_{T}}(t)=\int_{x\in U(k)}h_{\psi}(xt)dx

where UU is the kernel of ρT~\tilde{\rho_{T}}. In Proposition 6 of [Ngô20], Ngo regularizes this integration into a principal value integral.

The space of Schwartz functions on knk^{n} are compactly supported functions in knk^{n} that are restrictions of standard Schwartz functions on Mn(k)M_{n}(k) to knk^{n}. Their further restriction to Tn(k)T_{n}(k) is 𝒮std(Tn)\mathcal{S}^{\text{std}}(T_{n}) in our notation.


Let ρ\rho_{*} be the push-forward of ρT~\tilde{\rho_{T}}. We will verify that the diagram

(6.4) 𝒮std(Tn)ρ𝒮ρT(T)JstdJρT𝒮std(Tn)ρ𝒮ρT(T).\begin{CD}\mathcal{S}^{\text{std}}(T_{n})@>{\rho_{*}}>{}>\mathcal{S}^{\rho_{T}}(T)\\ @V{J^{\text{std}}}V{}V@V{}V{J^{\rho_{T}}}V\\ \mathcal{S}^{\text{std}}(T_{n})@>{\rho_{*}}>{}>\mathcal{S}^{\rho_{T}}(T).\end{CD}

commutes, where 𝒮ρT(T)\mathcal{S}^{\rho_{T}}(T) is the image of 𝒮std(Tn)\mathcal{S}^{\text{std}}(T_{n}) under ρ\rho_{*}.


Let ϕ𝒮std(Tn)\phi\in\mathcal{S}^{\text{std}}(T_{n}) and define

(6.5) ϕ~(t~)=U(k)ϕ(ut~)𝑑u\tilde{\phi}(\tilde{t})=\int_{U(k)}\phi(u\tilde{t})du

where t~T(k)\tilde{t}\in T(k). The commutativity of (6.4) is equivalent to

Lemma 6.1.

For ϕ𝒮std(Tn)\phi\in\mathcal{S}^{\text{std}}(T_{n}), define ϕ~\tilde{\phi} by (6.5). Then

ρ(Jstdϕ)=JρTϕ~\rho_{*}(J^{\text{std}}\star\phi^{\vee})=J^{\rho_{T}}\star\tilde{\phi}^{\vee}
Proof.

By definition, for tTt\in T,

(6.6) ρ(Jstdϕ)(t)\displaystyle\rho_{*}(J^{\text{std}}\star\phi^{\vee})(t) =U(k)(Jstdϕ)(ut)𝑑u\displaystyle=\int_{U(k)}(J^{\text{std}}\star\phi^{\vee})(ut)du
=U(k)(Tn(k)hψ(utt~)ϕ(t~)𝑑t~)𝑑u\displaystyle=\int_{U(k)}(\int_{T_{n}(k)}h_{\psi}(ut\tilde{t})\phi(\tilde{t})d\tilde{t})du
=Tn(k)hψ(tt~)(U(k)ϕ(u1t~)𝑑u)𝑑t~\displaystyle=\int_{T_{n}(k)}h_{\psi}(t\tilde{t})(\int_{U(k)}\phi(u^{-1}\tilde{t})du)d\tilde{t}
=Tn(k)hψ(tt~)ϕ~(t~)𝑑t~\displaystyle=\int_{T_{n}(k)}h_{\psi}(t\tilde{t})\tilde{\phi}(\tilde{t})d\tilde{t}
=T(k)(U(k)hψ(utt~)𝑑u)ϕ~(ut~)𝑑t~\displaystyle=\int_{T(k)}(\int_{U(k)}h_{\psi}(ut\tilde{t})du)\tilde{\phi}(u\tilde{t})d\tilde{t}
=T(k)(U(k)hψ(ut~)𝑑u)ϕ~(t1t~)𝑑t~\displaystyle=\int_{T(k)}(\int_{U(k)}h_{\psi}(u\tilde{t})du)\tilde{\phi}(t^{-1}\tilde{t})d\tilde{t}
=(JρTϕ~)(t),\displaystyle=(J^{\rho_{T}}\star\tilde{\phi}^{\vee})(t),

using T=Tn/UT=T_{n}/U in (6.6), then the lemma follows. ∎

The push-forward ρ\rho_{*} can be restricted to

[T^n]Wn(Tn(k),Tn(𝒪k))Wn\mathbb{C}[\hat{T}_{n}]^{W_{n}}\cong\mathcal{H}(T_{n}(k),T_{n}(\mathcal{O}_{k}))^{W_{n}}

leading to

(6.7) (Tn(k),Tn(𝒪k))Wnρ(T(k),T(𝒪k))WJstdJρT(Tn(k),Tn(𝒪))Wnρ(T(k),T(𝒪k))W,\begin{CD}\mathcal{H}(T_{n}(k),T_{n}(\mathcal{O}_{k}))^{W_{n}}@>{\rho_{*}}>{}>\mathcal{H}(T(k),T(\mathcal{O}_{k}))^{W}\\ @V{J^{\text{std}}}V{}V@V{}V{J^{\rho_{T}}}V\\ \mathcal{H}(T_{n}(k),T_{n}(\mathcal{O}))^{W_{n}}@>{\rho_{*}}>{}>\mathcal{H}(T(k),T(\mathcal{O}_{k}))^{W},\end{CD}

where WnW_{n} is the Weyl group Wn=W(Gln,Tn)W_{n}=W(Gl_{n},T_{n}) ,W:=W(G,T)W:=W(G,T) and \mathcal{H} denotes the corresponding Hecke algebra. Identifying, via the corresponding Satake isomorphisms

(GLn(k)):=(GLn(k),GLn(𝒪k))(Tn(k),Tn(𝒪k))Wn\mathcal{H}^{\circ}(GL_{n}(k)):=\mathcal{H}(GL_{n}(k),GL_{n}(\mathcal{O}_{k}))\cong\mathcal{H}(T_{n}(k),T_{n}(\mathcal{O}_{k}))^{W_{n}}

and

(G(k),G(𝒪k))(T(k),T(𝒪k))W\mathcal{H}(G(k),G(\mathcal{O}_{k}))\cong\mathcal{H}(T(k),T(\mathcal{O}_{k}))^{W}
(6.8) (GLn(k))Sat(Tn(k),Tn(𝒪k))Wnρ(T(k),T(𝒪k))WSat1(G(k),G(𝒪k))JstdJstdJρTJρ(GLn(k))Sat(Tn(k),Tn(𝒪k))Wnρ(T(k),T(𝒪k))WSat1(G(k),G(𝒪k)),\begin{CD}\mathcal{H}^{\circ}(GL_{n}(k))@>{Sat}>{}>\mathcal{H}(T_{n}(k),T_{n}(\mathcal{O}_{k}))^{W_{n}}@>{\rho_{*}}>{}>\mathcal{H}(T(k),T(\mathcal{O}_{k}))^{W}@>{Sat^{-1}}>{}>\mathcal{H}(G(k),G(\mathcal{O}_{k}))\\ @V{J^{\text{std}}}V{}V@V{}V{J^{std}}V@V{}V{J^{\rho_{T}}}V@V{J^{\rho}}V{}V\\ \mathcal{H}^{\circ}(GL_{n}(k))@>{Sat}>{}>\mathcal{H}(T_{n}(k),T_{n}(\mathcal{O}_{k}))^{W_{n}}@>{\rho_{*}}>{}>\mathcal{H}(T(k),T(\mathcal{O}_{k}))^{W}@>{Sat^{-1}}>{}>\mathcal{H}(G(k),G(\mathcal{O}_{k})),\end{CD}

in which JρTJ^{\rho_{T}} defines the Fourier transform

Jρ:𝒮(G)𝒮(G)J^{\rho}:\mathcal{S}(G)\to\mathcal{S}(G)

restricted to (G(k),G(𝒪k))\mathcal{H}(G(k),G(\mathcal{O}_{k})). Consequently, at least on (G(k),G(𝒪k))\mathcal{H}(G(k),G(\mathcal{O}_{k})), the Fourier transform JρJ^{\rho} and JρTJ^{\rho_{T}} commute with the Harish-Chandra transform.

7. The case of standard LL-functions for classical groups; the doubling method

We conclude by addressing multiplicativity in the case of standard LL-functions, twisted by a character, for classical groups as developed by Piatetski-Shapiro and Rallis, which has been addressed further by a number of other authors [BK02, JLZ20, Li18, Sha18] within our present context. We refer to the local theory developed by Lapid and Rallis. We will be brief and only mention the relevant statements.

The ρ\rho-Harish-Chandra transform is the one given in Proposition 11 of [LR05] as Ψ(ω,s)\Psi(\omega,s) with notation as in [LR05]. Our commutativity equation (5.45.4) in this case is equation (1717) in Lemma 99 of [LR05] in which Jρ=Mν(ω,A,s)J^{\rho}=M^{*}_{\nu}(\omega,A,s), a normalized intertwining operator, while JρLJ^{\rho_{L}} acts as the operator induced from M𝒲(ω,B,s)M^{*}_{\mathcal{W}}(\omega,B,s) with notation as in [LR05] in the context of doubling construction, or simply put JρL=M𝒲(ω,B,s)J^{\rho_{L}}=M^{*}_{\mathcal{W}}(\omega,B,s).

References

  • [BNS16] Alexis Bouthier, Bao Chau Ngo, and Yiannis Sakellaridis. On the formal arc space of a reductive monoid. American Journal of Mathematics, 138(1):81–108, 2016.
  • [BK10] Alexander Braverman and David Kazhdan. γ\gamma-functions of representations and lifting. In Visions in Mathematics, pages 237–278. Springer, 2010.
  • [BK02] Alexander Braverman and David A Kazhdan. Normalized intertwining operators and nilpotent elements in the Langlands dual group. Moscow Mathematical Journal, 2(3):533–553, 2002.
  • [Bri89] Michel Brion. Spherical varieties; an introduction. In Topological methods in algebraic transformation groups, pages 11–26. Springer, 1989.
  • [BK07] Michel Brion and Shrawan Kumar. Frobenius splitting methods in geometry and representation theory, volume 231. Springer Science & Business Media, 2007.
  • [Cas17] William Casselman. Symmetric powers and the Satake transform. Bulletin of the Iranian Mathematical Society, 43(4 (Special Issue)):17–54, 2017.
  • [CST17] J. W. Cogdell, F. Shahidi, and T.-L. Tsai. Local Langlands correspondence for GLn\mathrm{GL}_{n} and the exterior and symmetric square root numbers for GL(n). Duke Mathematical Journal, 166(11):2053 – 2132, 2017.
  • [CKPSS04] James W. Cogdell, Henry H. Kim, Ilya I. Piatetski-Shapiro, and Freydoon Shahidi. Functoriality for the classical groups. Publications Mathématiques de l’IHÉS, 99:163–233, 2004.
  • [CLS11] David A. Cox, John B. Little, and Henry K. Schenck. Toric Varieties, volume 124 of GSM. American Mathematical Soc., 2011.
  • [Elk78] Renée Elkik. Singularités rationnelles et déformations. Inventiones Mathematicae, 47(2):139–147, 1978.
  • [GT11] Wee Teck Gan and Shuichiro Takeda. The local Langlands conjecture for GSp (4). Annals of Mathematics, pages 1841–1882, 2011.
  • [GPSR87] Stephen Gelbart, Ilya Piatetski-Shapiro, and Stephen Rallis. Explicit constructions of automorphic L-functions, volume 1254 of SLN. Springer, 1987.
  • [GL20] Jayce Robert Getz and Baiying Liu. A refined poisson summation formula for certain braverman-kazhdan spaces. Science China Mathematics, pages 1–30, 2020.
  • [GJ72] Roger Godement and Hervé Jacquet. Zeta functions of simple algebras, volume 260. Springer, 1972.
  • [HT01] Michael Harris and Richard Taylor. The Geometry and Cohomology of Some Simple Shimura Varieties.(AM-151), Volume 151, volume 151. Princeton university press, 2001.
  • [Hen00] Guy Henniart. Une preuve simple des conjectures de Langlands pour GL(n) sur un corps p-adique. Inventiones mathematicae, 139(2):439–455, 2000.
  • [JPSS83] Hervé Jacquet, Ilya I. Piatetski-Shapiro, and Joseph A. Shalika. Rankin-selberg convolutions. American Journal of Mathematics, 105(2):367–464, 1983.
  • [JLZ20] Dihua Jiang, Zhilin Luo, and Lei Zhang. Harmonic analysis and gamma functions on symplectic groups. arXiv preprint arXiv:2006.08126, 2020.
  • [Kim03] Henry Kim. Functoriality for the exterior square of GL4{GL}_{4} and the symmetric fourth of GL2{GL}_{2}. Journal of the American Mathematical Society, 16(1):139–183, 2003.
  • [KS02] Henry H Kim and Freydoon Shahidi. Functorial products for GL2×GL3{GL}_{2}\times{GL}_{3} and functorial symmetric cube for GL2{GL}_{2}, with an appendix by C. Bushnell and G. Henniart. Annals of Math, 155(8):837–893, 2002.
  • [Kno91] Friedrich Knop. The Luna-Vust theory of spherical embeddings. In Proceedings of the Hyderabad conference on Algebraic Groups, pages 225–249, 1991.
  • [Laf14] Laurent Lafforgue. Noyaux du transfert automorphe de Langlands et formules de poisson non linéaires. Japanese Journal of Mathematics, 9(1):1–68, 2014.
  • [LR05] E Lapid and Stephen Rallis. On the local factors of representations of classical groups. In Automorphic representations, L-functions and applications: progress and prospects, volume 11 of Ohio State Univ. Math. Res. Inst. Publ, pages 309–359. de Gruyter, Berlin, 2005.
  • [Li17] Wen-Wei Li. Basic functions and unramified local l-factors for split groups. Science China Mathematics, 60(5):777–812, 2017.
  • [Li18] Wen-Wei Li. Zeta integrals, Schwartz spaces and local functional equations, volume 2228 of SLN. Springer, 2018.
  • [Los09] Ivan Losev. Proof of the Knop conjecture. In Annales de l’Institut Fourier, volume 59, pages 1105–1134, 2009.
  • [Los10] Ivan Losev. Uniqueness properties for spherical varieties. Les cours du CIRM, 1(1):113–120, 2010.
  • [Ngô20] Bao Châu Ngô. Hankel transform, Langlands functoriality and functional equation of automorphic L-functions. Japanese Journal of Mathematics, 15(1):121–167, 2020.
  • [Per14] Nicolas Perrin. Introduction to spherical varieties. http://relaunch.hcm.uni-bonn.de/fileadmin/perrin/spherical.pdf, 2014.
  • [PSR86] Ilya Piatetski-Shapiro and Stephen Rallis. ε\varepsilon factor of representations of classical groups. Proceedings of the National Academy of Sciences, 83(13):4589–4593, 1986.
  • [Pop87] Vladimir L Popov. Contraction of the actions of reductive algebraic groups. Mathematics of the USSR-Sbornik, 58(2):311, 1987.
  • [Ren05] Lex E Renner. Linear algebraic monoids, Encylopedia of Mathematical Sciences, volume 134. Springer, 2005.
  • [Rit03] Alvaro Rittatore. Reductive embeddings are cohen-macaulay. Proceedings of the American Mathematical Society, 131(3):675–684, 2003.
  • [Sak18] Yiannis Sakellaridis. Inverse satake transforms. In Werner Müller, Sug Woo Shin, and Nicolas Templier, editors, Geometric Aspects of the Trace Formula, pages 321–349, Cham, 2018. Springer International Publishing.
  • [Sha10] Freydoon Shahidi. Eisenstein series and automorphic L-functions, volume 58, American Mathematical Society Colloquium Publications. American Mathematical Society, Providence, RI, 2010.
  • [Sha12] Freydoon Shahidi. On equality of arithmetic and analytic factors through local Langlands correspondence. Pacific Journal of Mathematics, 260(2):695–715, 2012.
  • [Sha17] Freydoon Shahidi. Local factors, reciprocity and Vinberg monoids. In “Prime Numbers and Representation Theory”, volume 2 of Lecture Series of Modern Number Theory, pages 200–256. Science Press, 2017.
  • [Sha18] Freydoon Shahidi. On generalized Fourier transforms for standard L-functions. In “Geometric Aspects of the Trace Formula”, Simons Symposium on the Trace Formula, pages 351–404. Springer, 2018.
  • [Sou93] David Soudry. Rankin-Selberg Convolutions for SO2l+1×GLn\mathrm{SO}_{2l+1}\times\mathrm{GL}_{n}: Local Theory. Memoirs of AMS 105, no. 500. American Mathematical Soc., 1993.
  • [Stu95] Bernd Sturmfels. On vector partition functions. Journal of Combinatorial theory, series A, 72(2):302–309, 1995.
  • [Tat50] John Tate. Fourier analysis in number fields and hecke’s zeta functions. Thesis Princeton, 1950.

Purdue University
West Lafayette, IN 47907
[email protected]
[email protected]