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

\DefineSimpleKey

bibarxiv

A note on the Sauvageot density principle

Yugo Takanashi Graduate School of Mathematical Sciences, University of Tokyo, 3-8-1 Komaba, Meguro-Ku, Tokyo 153-8914, Japan [email protected]
Abstract.

In this short note, we address a gap in the proof of Sauvageot’s paper [Sau97] pointed out in [NelsonVenkatesh2021-orbitmethod] and provide a complete proof of its main theorem.

1. Introduction

The Sauvageot density principle for reductive groups over local fields was proved in [Sau97] and it has been applied to prove well known results; see, for example, [Shi12], [FLM15] and [MikrosBergeron2017]. It was noted in [Shi12]*Appendix A that there are some minor errors in the proof of this result. More recently, in [NelsonVenkatesh2021-orbitmethod]*p.159, footnote, it was pointed out that there is a gap in the proof. In this short note, we fill the gap in the proof pointed out in [NelsonVenkatesh2021-orbitmethod]. We also fix some notational and logical errors in the proof, pointed out in [Shi12].

Roughly speaking, to prove the density principle, we need to approximate a function on the unitary dual using the Fourier transform of compactly supported functions. The proof given in [Sau97] utilizes only the functions that factor through the space of Bernstein components. However, the map from the unitary dual to the union of Bernstein components has non-trivial fibers, and the functions used by Sauvageot cannot separate the points within a single fiber. We will explain why this method by Sauvageot actually fails, see Remark 6.1. This appears to be the gap identified in [NelsonVenkatesh2021-orbitmethod]*p.159, footnote.

The main idea to fill the gap is to utilize a module version of the Stone-Weierstrass theorem, see Lemma 6.2, and the measure theoretic analog of this result, see Theorem 6.11. This enables us to distinguish points within each fiber of the map, allowing us to recover Sauvageot’s original result.

In the last part of this paper, we will present two applications of Sauvegeot’s density principle. One is the automorphic Plancherel density theorem by Shin. We will give a somewhat extended version of his result. The other one is an application of the Weyl law by Finis-Matz and Eikemeier, which produces spherical cusp forms of a quasi-split group with prescribed local properties.

1.1. Notation and assumptions

Let FF be a local field or global field of characteristic 0. Let 𝔾\mathbb{G} be a connected reductive group over FF. We denote by GG the set of FF-valued points of 𝔾\mathbb{G}. Unless otherwise stated, we will identify 𝔾\mathbb{G} and G=𝔾(F)G=\mathbb{G}(F). Also, we will only consider the sets of FF-valued points of FF-subgroups of 𝔾\mathbb{G} and we referring to them simply as subgroups.

We use the following standard notation.

  • Let AGA_{G} denote the maximal split torus of the center ZGZ_{G} of GG.

  • Let 𝒫G(M)\mathcal{P}^{G}(M) denote the set of parabolic subgroups of GG with the Levi factor MM.

  • Let 𝒳(G)\mathcal{X}(G) denote the group of unramified characters of GG. Let 𝒳unit(G)\mathcal{X}_{\mathrm{unit}}(G) denote the group of unitary unramified characters of GG.

  • Let X(M)X^{*}(M) be the group of rational characters of MM over FF. We set 𝔞M=X(M)\mathfrak{a}^{*}_{M}=X^{*}(M)\otimes_{\mathbb{Z}}\mathbb{R} and 𝔞M,=X(M)\mathfrak{a}^{*}_{M,\mathbb{C}}=X^{*}(M)\otimes_{\mathbb{Z}}\mathbb{C}. We have the natural map

    𝔞M,𝒳(M):λχλ.\displaystyle\mathfrak{a}^{*}_{M,\mathbb{C}}\to\mathcal{X}(M)\colon\lambda\mapsto\chi_{\lambda}.
  • Let W(G,M)W(G,M) denote the relative Weyl group for MM.

If F=F=\mathbb{R}, let Lie(G)\operatorname{Lie}(G) or 𝔤\mathfrak{g} denote the Lie algebra of GG. We always fix a maximal compact subgroup of GG and we denote it by KK. By a representation of a real reductive group, we mean an admissible (𝔤,K)(\mathfrak{g}\otimes_{\mathbb{R}}\mathbb{C},K) over \mathbb{C}. If we consider a topological representation, we take the Casselman-Wallach globalization. If FF is a pp-adic field, by a representation of pp-adic groups, we mean a smooth admissible representation over \mathbb{C}.

1.2. Acknowledgment

I became aware of Sauvageot’s paper while collaborating with Satoshi Wakatsuki. I would like to thank Satoshi Wakatsuki for his encouragement and for careful reading. I would like to thank Sug Woo Shin for his encouragement, careful reading and helpful comments which greatly improved the readability of this article. Some parts of this paper are based on his work [Shi12]. I would like to thank Masaki Natori and Mao Hoshino for their careful reading and for pointing out some mistakes in definition 6.6 and Lemma 6.7. I would like to thank Yoichi Mieda for his careful reading. This work is supported by JSPS Grant-in-Aid for JSPS Fellows 23KJ0403.

2. The space of infinitesimal characters

2.1. Definition

We first recall the elementary facts about the space of infinitesimal characters, following [BernsteinDeligneKazhdan1986]*2.

Let F,GF,G as in section 1.1.

The notion of infinitesimal characters for reductive groups over \mathbb{R} is well known; they correspond to the characters of the commutative algebra Z(U(𝔤))Z(U(\mathfrak{g}_{\mathbb{C}})). Let Θ(G)\Theta(G) denote the set of infinitesimal characters of GG .

Definition 2.1.

We assume that FF is a pp-adic field. We call a pair (M,σ)(M,\sigma), consisting of a Levi subgroup MM of GG and a supercuspidal representation σ\sigma of MM a cuspidal datum of GG. We refer to a GG-conjugacy class of a cuspidal datum of GG as an infinitesimal character of GG. We write Θ(G)\Theta(G) for the set of infinitesimal characters. We write [M,σ]G[M,\sigma]_{G} for the class of (M,σ)(M,\sigma).

The following subrepresentation theorem is well known.

Theorem 2.2 ([Renard2010-p-adiques]*VI 5.4, Théorème).

We assume that FF is a pp-adic field. Let π\pi be an irreducible representation of GG.

  1. (1)

    There exists a cuspidal datum (M,σ)(M,\sigma) and a parabolic subgroup P𝒫G(M)P\in\mathcal{P}^{G}(M) such that πiPG(σ)\pi\subset i^{G}_{P}(\sigma).

  2. (2)

    Let (M,σ)(M,\sigma) and (M,σ)(M^{\prime},\sigma^{\prime}) be two cuspidal data. We assume that there exist parabolic subgroups P𝒫G(M)P\in\mathcal{P}^{G}(M) and P𝒫G(M)P^{\prime}\in\mathcal{P}^{G}(M) such that the parabolic inductions iPG(σ)i^{G}_{P}(\sigma) and iPG(σ)i^{G}_{P^{\prime}}(\sigma^{\prime}) have a common irreducible component. Then, the infinitesimal characters [M,σ]G[M,\sigma]_{G} and [M,σ]G[M^{\prime},\sigma^{\prime}]_{G} are equal. Also, the converse holds true.

  3. (3)

    In the case of (2)(2), the semisimplifications of iPG(σ)i^{G}_{P}(\sigma) and iPG(σ)i^{G}_{P^{\prime}}(\sigma^{\prime}) are equal.

Construction 2.3.

We assume that FF is a pp-adic field. Let (M,σ)(M,\sigma) be a cuspidal datum of GG. We call the image of the map

(2.1) 𝒳(M)Θ(G):χ[M,σχ]G\displaystyle\mathcal{X}(M)\to\Theta(G)\colon\chi\mapsto[M,\sigma\otimes\chi]_{G}

a connected component of Θ(G)\Theta(G) containing [M,σ]G[M,\sigma]_{G}. We denote it by C=C[M,σ]C=C_{[M,\sigma]}. We denote the stabilizer of σ\sigma in 𝒳(M)\mathcal{X}(M) by 𝒳(M)σ\mathcal{X}(M)_{\sigma}. Also, we denote the stabilizer of C=C[M,σ]C=C_{[M,\sigma]} in W(G,M)W(G,M) by W(G,M)CW(G,M)_{C}.

Through the above map, we can identify the set C[M,σ]C_{[M,\sigma]} with the quotient of the complex algebraic torus 𝒳(M)/𝒳(M)σ\mathcal{X}(M)/\mathcal{X}(M)_{\sigma} by the finite group W(G,M)CW(G,M)_{C}. Thus, we endow the set C=C[M,σ]C=C_{[M,\sigma]} with the structure of affine algebraic variety (𝒳(M)/𝒳(M)σ)//W(G,M)C(\mathcal{X}(M)/\mathcal{X}(M)_{\sigma})//W(G,M)_{C}. We also endow the set Θ(G)\Theta(G) with the structure of a countable disjoint union of complex algebraic varieties as the countable disjoint union of connected components. Note that we can identify the connected component CC with the image of the map

(2.2) 𝔞M,Θ(G):λ[M,σχλ]G.\displaystyle\mathfrak{a}^{*}_{M,\mathbb{C}}\to\Theta(G)\colon\lambda\mapsto[M,\sigma\otimes\chi_{\lambda}]_{G}.

This definition can be used to define the similar notion in the archimedean case.

2.2. A ring of functions and hermitian involutions

Definition 2.4.

We assume that FF is a pp-adic field. Let 𝒜(G)\mathcal{A}(G) denote the ring of regular functions on Θ(G)\Theta(G) supported on finite connected components of Θ(G)\Theta(G). We assume that F=F=\mathbb{R}. Let θ\theta be a Cartan involution of G()G(\mathbb{R}). In this case, let 𝔥Lie(G)\mathfrak{h}\subset\operatorname{Lie}(G) be the Lie algebra of a θ\theta-stable maximally split maximal torus HGH\subset G. Then, the subalgebra decomposes as 𝔞0𝔱0\mathfrak{a}_{0}\oplus\mathfrak{t}_{0} with respect to the Cartan involution. The space 𝔞0\mathfrak{a}_{0} (resp. 𝔱0\mathfrak{t}_{0}) is the 1-1-eigenspace (resp. +1+1-eigenspace) with respect to the action of θ\theta. We set 𝔥=𝔞0i𝔱0\mathfrak{h}_{\mathbb{R}}=\mathfrak{a}_{0}\oplus i\mathfrak{t}_{0}. Then, we have the Fourier-Laplace transform

Cc(𝔥)W(G(),H())𝒫𝒲(𝔥)W(G(),H()),\displaystyle C_{c}^{\infty}(\mathfrak{h}_{\mathbb{R}})^{W(G(\mathbb{C}),H(\mathbb{C}))}\to\mathcal{PW}(\mathfrak{h}_{\mathbb{R}}\otimes_{\mathbb{R}}\mathbb{C})^{W(G(\mathbb{C}),H(\mathbb{C}))},

from the space of compactly supported functions to the space of Paley-Wiener functions. We denote by 𝒜(G)\mathcal{A}(G) the ring of functions on the right hand side.

Definition 2.5.

We assume that FF is a pp-adic field. We define the anti-holomorphic involution on the set Θ(G)\Theta(G) by [M,σ]G[M,σh]G[M,\sigma]_{G}\mapsto[M,\sigma^{h}]_{G}, where σh=σ¯\sigma^{h}=\overline{\sigma}^{\vee}. We denote the image by [M,σ]Gh[M,\sigma]_{G}^{h}. Let Θ(G)herm\Theta(G)_{\mathrm{herm}} denote the set of fixed points of this involution. We assume that F=F=\mathbb{R}. The linear bijection Lie(G)Lie(G):XX\operatorname{Lie}(G)\to\operatorname{Lie}(G)\colon X\mapsto-X extends \mathbb{C}-semilinearly to the semilinear isomorphism Z(U(Lie(G)))Z(U(Lie(G)))Z(U(\operatorname{Lie}(G)_{\mathbb{C}}))\to Z(U(\operatorname{Lie}(G)_{\mathbb{C}})). We denote this map by zzhz\mapsto z^{h}. This involution induces the anti-holomorphic involution ()h:Θ(G)Θ(G)(\cdot)^{h}\colon\Theta(G)\to\Theta(G).

Lemma 2.6.

We denote the pullback of an element a𝒜(G)a\in\mathcal{A}(G) by the map hh by aha^{h}. Then, the map aah¯a\mapsto\overline{a^{h}} induces the \mathbb{C}-linear involution of the algebra 𝒜(G)\mathcal{A}(G).

3. The space of induced representations

We introduce the space of induced representations from essentially discrete series representations, which is similar to the space of infinitesimal characters.

Definition 3.1.

We call a pair (M,σ)(M,\sigma) which consists of a Levi subgroup MM of GG and an essentially discrete series representation σ\sigma of MM a discrete pair of GG. We refer to a GG-conjugacy class of a discrete pair of GG as a discrete datum of GG. We denote the class of (M,σ)(M,\sigma) by {M,σ}G\{M,\sigma\}_{G} and the set of discrete data of GG by Θdisc(G)\Theta_{\mathrm{disc}}(G).

Lemma 3.2.

Let (M,σ)(M,\sigma) and (M,σ)(M^{\prime},\sigma^{\prime}) be discrete data of GG and P𝒫G(M)P\in\mathcal{P}^{G}(M) and P𝒫G(M)P^{\prime}\in\mathcal{P}^{G}(M^{\prime}). Then, the semisimplifications iPG(σ)i_{P}^{G}(\sigma) and iPG(σ)i_{P^{\prime}}^{G}(\sigma^{\prime}) are equal if and only if the discrete data {M,σ}G\{M,\sigma\}_{G} and {M,σ}G\{M^{\prime},\sigma^{\prime}\}_{G} are equal.

Proof.

First, by the character formula for parabolic induction [Lipsman1971]*Theorem 9.1 and [vanDijk1972]*5.3, the character of iPG(σ)i_{P}^{G}(\sigma) is independent from the choice of PP. Hence, we assume the real part of the central character Re(χσ)\operatorname{Re}(\chi_{\sigma}) of σ\sigma is positive with respect to PP and PP is standard. Here, the positivity means that Re(χσ)\operatorname{Re}(\chi_{\sigma}) is in the closure of the cone defined by the fundamental weight associated to PP (that is the relative Weyl chamber associated to PP).

Now, we take the Levi subgroup M1M_{1} which contains MM such that Re(χσ)𝔞M1\operatorname{Re}(\chi_{\sigma})\in\mathfrak{a}_{M_{1}}^{*} and regular in 𝔞M1\mathfrak{a}_{M_{1}}^{*}. Then, we obtain the standard parabolic subgroup P1P_{1} of GG with Levi subgroup M1M_{1} and Re(χσ)\operatorname{Re}(\chi_{\sigma}) is positive with respect to P1P_{1}. Also, we write the representation σ\sigma as σuRe(χσ)\sigma_{u}\otimes\operatorname{Re}(\chi_{\sigma}), where σu\sigma_{u} is a unitary discrete series representation. In this situation, we have iPG(σ)=iP1G(iPM1M1(σu)Re(χσ))i_{P}^{G}(\sigma)=i_{P_{1}}^{G}(i^{M_{1}}_{P\cap M_{1}}(\sigma_{u})\otimes\operatorname{Re}(\chi_{\sigma})). By decomposing iPM1M1(σu)i^{M_{1}}_{P\cap M_{1}}(\sigma_{u}) to the irreducible tempered representations and taking their semisimplifications, we obtain the sum of standard modules. By the remark in the proof of [Clo86]*Proposition 2.1 and [Hecht1979]*Proposition 6.2, the standard representations form a finite basis of the Grothendieck group of representations π\pi^{\prime} of GG with infG(π)=infG(iPG(σ))\inf_{G}(\pi^{\prime})=\inf_{G}(i^{G}_{P}(\sigma)).

If we apply this procedure to the representation iPG(σ)i_{P^{\prime}}^{G}(\sigma^{\prime}) to obtain M1M_{1}^{\prime}, P1P_{1}^{\prime} and so on. By assumption, we see that P1=P1P_{1}=P_{1}^{\prime}, M1=M1M_{1}=M_{1}^{\prime}, Re(χσ)=Re(χσ)\operatorname{Re}(\chi_{\sigma})=\operatorname{Re}(\chi_{\sigma^{\prime}}) and (M,σu)(M,\sigma_{u}) and (M,σu)(M^{\prime},\sigma^{\prime}_{u}) are conjugate by an element wM1(F)w\in M_{1}(F), i.e. σu=w(σu)\sigma^{\prime}_{u}=w(\sigma_{u}) and M=w(M)M^{\prime}=w(M). As the character Re(χσ)\operatorname{Re}(\chi_{\sigma}) is in 𝔞M1\mathfrak{a}_{M_{1}}^{*}, the element ww acts this element trivially and we have w(M,σ)=(M,σ)w(M,\sigma)=(M^{\prime},\sigma^{\prime}). Hence the result. ∎

Remark 3.3.

We identify {M,σ}G\{M,\sigma\}_{G} with the semisimplification iPG(σ)ssi^{G}_{P}(\sigma)_{ss} of the induced representation iPG(σ)i^{G}_{P}(\sigma). Using Lemma 3.2 and 2.3, we can endow the set Θdisc(G)\Theta_{\mathrm{disc}}(G) with the countable union of complex algebraic varieties and we can also define the notion of connected component. We omit the details.

Construction 3.4.

We define the map

infG:Θdisc(G)Θ(G)\displaystyle\mathrm{inf}_{G}\colon\Theta_{\mathrm{disc}}(G)\to\Theta(G)

defined as follows.

Let {M,σ}G\{M,\sigma\}_{G} be a discrete pair of GG.

  • If F=F=\mathbb{R}, the map is just taking the infinitesimal character of iPG(σ)i^{G}_{P}(\sigma) for any P𝒫G(M)P\in\mathcal{P}^{G}(M).

  • If FF is a pp-adic field, we can embed σ\sigma in iQM(τ)i^{M}_{Q}(\tau) for some cuspidal datum (L,τ)Θ(M)(L,\tau)\in\Theta(M) and Q𝒫M(L)Q\in\mathcal{P}^{M}(L). Then, we set infG({M,σ}G)=[L,τ]G\inf_{G}(\{M,\sigma\}_{G})=[L,\tau]_{G}. This map is well-defined by Theorem 2.2 and Lemma 3.2.

The following lemma is obvious.

Lemma 3.5.

The map

infG:Θdisc(G)Θ(G)\displaystyle\mathrm{inf}_{G}\colon\Theta_{\mathrm{disc}}(G)\to\Theta(G)

taking the infinitesimal characters of representations is continuous and proper. Also, each fiber is a finite set.

Proof.

See [NelsonVenkatesh2021-orbitmethod]*p.157, line 1 for the first statement. We prove the second statement. Note that the set of GG-conjugacy classes of Levi subgroups of GG is finite. Thus, the second statement follows from [Wallach1988realreductiveI]*Theorem 5.5.6 if F=F=\mathbb{R}, and [Renard2010-p-adiques]*VI 6.2, Lemme if FF is a pp-adic field and the fact that the parabolic induction of any irreducible representation is of finite length. ∎

Definition 3.6.

Let Θdisc(G)temp\Theta_{\mathrm{disc}}(G)_{\mathrm{temp}} denote the closed subset of Θdisc(G)\Theta_{\mathrm{disc}}(G) which consists of classes {M,σ}G\{M,\sigma\}_{G} with σ\sigma unitary discrete series. This set also endows with the structure of the countable union of algebraic varieties over \mathbb{R}, as the countable union of quotients of compact tori or \mathbb{R}-vector spaces by finite groups.

Remark 3.7.

We can identify the element {M,σ}GΘdisc(G)temp\{M,\sigma\}_{G}\in\Theta_{\mathrm{disc}}(G)_{\mathrm{temp}} with the semisimple representation iPG(σ)i^{G}_{P}(\sigma). As the spaces Θdisc(G),Θ(G)\Theta_{\mathrm{disc}}(G),\Theta(G) and Θdisc(G)temp\Theta_{\mathrm{disc}}(G)_{\mathrm{temp}} are countable unions of quotients of locally Euclidean spaces by finite groups, we have the notion of smooth functions, holomorphic functions on these spaces and so on.

The following is well known, for example, see [Wal03]*Proposition III 4.1 and [Harish-Chandra1972-Eisensteinintegral]*Lemma 12.

Lemma 3.8.

Let {M,σ}G\{M,\sigma\}_{G} and {M,σ}G\{M^{\prime},\sigma^{\prime}\}_{G} be elements of Θdisc(G)temp\Theta_{\mathrm{disc}}(G)_{\mathrm{temp}}. Then, if the parabolic inductions iPG(σ)i^{G}_{P}(\sigma) and iPG(σ)i^{G}_{P^{\prime}}(\sigma^{\prime}) have a common irreducible constituent, then we have the equality {M,σ}G={M,σ}G\{M,\sigma\}_{G}=\{M^{\prime},\sigma^{\prime}\}_{G}.

4. On the generic irreducibility theorem

As noted in [Shi12], the proof of the generic irreducibility theorem [Sau97]*Théorème 3.2 appears to contain an error. Konno [Konno2003]*Theorem 4.2 provided a remedy. However, a stronger hypothesis than [Sau97]*Théorème 3.2 was used in [Konno2003]*p.403, l.24. We will show that this strengthening is unnecessary by proving the following lemma. Furthermore, the proof of [Sau97]*Corollaire 3.3 seems to require the original version of the theorem.

Lemma 4.1.

Let VV be a real vector space. Let SS be a finite subset in VV. Let TT^{\vee} be a finite subset in the linear dual VV^{\vee} of VV. If no element of TT^{\vee} vanishes at an element in SS, then there exist vspan(S)v\in\mathrm{span}_{\mathbb{R}}(S) such that λ(v)0\lambda(v)\neq 0 for all λT\lambda\in T^{\vee}.

Proof.

Let W=span(S)W=\mathrm{span}_{\mathbb{R}}(S). Then, every element of TT^{\vee} has non-zero restriction on WW. Using the standard hyperplane avoidance technique, we can take vWv\in W as in the statement. ∎

Let P=MNP=MN be a parabolic subgroup of GG with a Levi decomposition P=MNP=MN. Let Φ(G,AM)\Phi(G,A_{M}) be the set of weights of the action of AMA_{M} on Lie(G)\mathrm{Lie}(G).

Theorem 4.2.

For every irreducible representation σ\sigma of MM, there exists a neighborhood 𝒰\mathcal{U} of 0(𝔞M)0\in(\mathfrak{a}_{M})^{*}_{\mathbb{C}} such that iPG(σχλ)i_{P}^{G}(\sigma\otimes\chi_{\lambda}) is irreducible for any element λ𝒰\lambda\in\mathcal{U} such that α(λ)0\alpha^{\vee}(\lambda)\neq 0, αΦ(G,AM)\alpha\in\Phi(G,A_{M}).

Proof.

If F=F=\mathbb{R}, this follows from [Renard2024]*Theorem 1.1. We assume that FF is a pp-adic field. Then, the proof is the same as that in [Konno2003]*Theorem 4.2, if we combine the argument at [Konno2003]*p.403, l.24 with the previous lemma. ∎

5. The unitary dual and the Plancherel measure

We will review some definitions and results on the unitary dual and the Plancherel measures for reductive groups over local fields.

Definition 5.1.

We write Irrunit(G)\operatorname{Irr}_{\mathrm{unit}}(G) for the unitary dual of GG. This space is naturally endowed with the Fell topology. Let Irrtemp(G)\operatorname{Irr}_{\mathrm{temp}}(G) denote the closed subset of Irrunit(G)\operatorname{Irr}_{\mathrm{unit}}(G) consisting of tempered representations.

We fix a Haar measure on GG throughout. Also, the space Θdisc(G)\Theta_{\mathrm{disc}}(G) is a countable union of quotients of locally Euclidean spaces by finite groups. Thus, we have the quotient measure on Θdisc(G)\Theta_{\mathrm{disc}}(G), induced from the locally Euclidean measures. This measure naturally restricts to the subspace Θdisc(G)temp\Theta_{\mathrm{disc}}(G)_{\mathrm{temp}}. Let dπd\pi denote this measure, where π=iPG(σ)Θdisc(G)temp\pi=i^{G}_{P}(\sigma)\in\Theta_{\mathrm{disc}}(G)_{\mathrm{temp}}.

Theorem 5.2 ([Wal03]*Théorème VIII 1.1, [Wallach1992RRGII]*Theorem 13.4).

There exists a positive smooth function μG(π)\mu^{G}(\pi) of moderate growth on Θdisc(G)temp\Theta_{\mathrm{disc}}(G)_{\mathrm{temp}} such that we have

f(1)=πΘdisc(G)temptrπ(f)μG(π)dπ\displaystyle f(1)=\int_{\pi\in\Theta_{\mathrm{disc}}(G)_{\mathrm{temp}}}\operatorname{tr}\pi(f)\mu^{G}(\pi)\,\mathrm{d}\pi

for any smooth function fCc(G)f\in C_{c}^{\infty}(G).

Definition 5.3.

We call the measure μG(π)dπ\mu^{G}(\pi)d\pi the Plancherel measure of GG and we denote it by μG\mu^{G}. Let f^\widehat{f} denote the function πtrπ(f)\pi\mapsto\operatorname{tr}\pi(f) on Irrunit(G)\operatorname{Irr}_{\mathrm{unit}}(G) or Θdisc(G)\Theta_{\mathrm{disc}}(G). We refer to it as the Fourier transform of fCc(G)f\in C_{c}^{\infty}(G). We denote the set of Fourier transforms for KK-finite elements in Cc(G)C_{c}^{\infty}(G) by 𝒯(G)\mathcal{FT}(G).

Definition 5.4.

Let fCc(G)f\in C_{c}^{\infty}(G) be a smooth function. We set

f(g)=f(g1)¯.\displaystyle f^{*}(g)=\overline{f(g^{-1})}.
Lemma 5.5.

Let π\pi be a unitary admissible representation of GG of finite length. Let fCc(G)f\in C_{c}^{\infty}(G) be a smooth KK-finite function. Then, we have

trπ(f)=trπ(f)¯.\displaystyle\operatorname{tr}\pi(f^{*})=\overline{\operatorname{tr}\pi(f)}.
Proof.

We fix an orthonormal basis \mathcal{B} of VV. Then, we have

trπ(f)\displaystyle\operatorname{tr}\pi(f^{*}) =v(π(f)v,v)\displaystyle=\sum_{v\in\mathcal{B}}(\pi(f^{*})v,v)
=vGf(g1)¯(π(g)v,v)dg\displaystyle=\sum_{v\in\mathcal{B}}\int_{G}\overline{f(g^{-1})}(\pi(g)v,v)\,\mathrm{d}g
=vGf(g)(π(g)v,v)¯dg\displaystyle=\sum_{v\in\mathcal{B}}\int_{G}\overline{f(g)(\pi(g)v,v)}\,\mathrm{d}g
=trπ(f)¯.\displaystyle=\overline{\operatorname{tr}\pi(f)}.

Note that by admissibility, the above sums are finite sums. Hence the result. ∎

Lemma 5.6.

The subset of Θdisc(G)temp\Theta_{\mathrm{disc}}(G)_{\mathrm{temp}} which consists of parabolic inductions of regular discrete series representations in the sense of [Sau97]*p.172, line 30 is a dense open subset in Θdisc(G)temp\Theta_{\mathrm{disc}}(G)_{\mathrm{temp}} which is conull with respect to dπd\pi. Thus, we can identify Θdisc(G)temp\Theta_{\mathrm{disc}}(G)_{\mathrm{temp}} with Irrtemp(G)\operatorname{Irr}_{\mathrm{temp}}(G) up to a null set with respect to dπd\pi.

On this subset, we have an equality of induced topologies from the spaces Θdisc(G)temp\Theta_{\mathrm{disc}}(G)_{\mathrm{temp}} and Irrtemp(G)\operatorname{Irr}_{\mathrm{temp}}(G).

Proof.

The first statement follows from Theorem 4.2. The second statement follows from the argument in [Beuzart-Plessis2021-Plancherel-GLnE-GLnF]*p.245, line 33. ∎

Definition 5.7.

By the previous lemma, the measure μG\mu^{G} defines a Borel measure on Irrtemp(G)\operatorname{Irr}_{\mathrm{temp}}(G) and also on Irrunit(G)\operatorname{Irr}_{\mathrm{unit}}(G). The measures are also denoted by μG\mu^{G}.

6. A variant of the Stone-Weierstrass theorem

In this section, we prove the main approximation theorem Theorem 6.11 which will be used to complete the proof of the Sauvageot density principle.

Before proving the approximation theorem, we recall the gap pointed out in [NelsonVenkatesh2021-orbitmethod]*p.159, footnote.

Remark 6.1.

In [Sau97]*p.181, line 4, for any function ff as [Sau97]*p.180, (1) and any positive number ϵ>0\epsilon>0, it is stated that there exist functions g,h𝒜(G)g,h\in\mathcal{A}(G) such that

|fginfG|\displaystyle\lvert f-g\circ\mathrm{inf}_{G}\rvert hinfG\displaystyle\leq h\circ\mathrm{inf}_{G}
μG(hinfG)\displaystyle\mu^{G}(h\circ\mathrm{inf}_{G}) <ϵ.\displaystyle<\epsilon.

We assume that GG has a pair of discrete series representations π1π2\pi_{1}\neq\pi_{2} such that infG(π1)=infG(π2)\mathrm{inf}_{G}(\pi_{1})=\inf_{G}(\pi_{2}). Such a pair exists if G=G2G=G_{2}, see [GS23theta]*Proposition 3.2, (ii). We take the characteristic function of {π1}\{\pi_{1}\} as ff and this satisfies the condition (1)(1) of [Sau97]*p.180. If there exist functions g,h𝒜(G)g,h\in\mathcal{A}(G) such that

|fginfG|\displaystyle\lvert f-g\circ\mathrm{inf}_{G}\rvert hinfG\displaystyle\leq h\circ\mathrm{inf}_{G}
μG(hinfG)\displaystyle\mu^{G}(h\circ\mathrm{inf}_{G}) <ϵ,\displaystyle<\epsilon,

then, we have

|1ginfG(π1)|\displaystyle\lvert 1-g\circ\mathrm{inf}_{G}(\pi_{1})\rvert hinfG(π1),\displaystyle\leq h\circ\mathrm{inf}_{G}(\pi_{1}),
|ginfG(π2)|\displaystyle\lvert g\circ\mathrm{inf}_{G}(\pi_{2})\rvert hinfG(π2).\displaystyle\leq h\circ\mathrm{inf}_{G}(\pi_{2}).

This implies that we have hinfG(π1)12h\circ\inf_{G}(\pi_{1})\geq\frac{1}{2}. Finally, we obtain μG(hμG)12d(π1)\mu^{G}(h\circ\mu^{G})\geq\frac{1}{2}d(\pi_{1}), where d(π1)d(\pi_{1}) is the formal degree of π1\pi_{1} and this gives a contradiction if ϵ<12d(π1)\epsilon<\frac{1}{2}d(\pi_{1}).

Thus, to approximate the functions in [Sau97]*p.180, (1) in the previous sense, we have to utilize more functions.

6.1. Approximation on locally compact Hausdorff spaces

Let X,YX,Y be locally compact Hausdorff spaces and p:XYp\colon X\to Y be a proper map. Let X=X{X}X^{\infty}=X\coprod\{\infty_{X}\} denote the one-point compactification of XX. Then, we can extend pp to a continuous map p:XYp^{\infty}\colon X^{\infty}\to Y^{\infty}. We have the identification of the function spaces:

C0(X){fC(X)|f(X)=0}:ffC_{0}(X)\overset{\cong}{\to}\{f\in C(X^{\infty})\,|\,f(\infty_{X})=0\}\colon f\mapsto f

by the extension f(X)=0f(\infty_{X})=0. This isomorphism preserves the norms ||||||\cdot||_{\infty}. The space C0(X)C_{0}(X) has the norm, so it defines a metric on this space. Also we can take the distance

dX(f,S)=infgSdX(f,g)d_{X}(f,S)=\inf_{g\in S}{d_{X}(f,g)}

for any element fC0(X)f\in C_{0}(X) and subset SC0(X)S\subset C_{0}(X).

We have the Stone-Weierstrass theorem for subalgebras of C0(Y)C_{0}(Y), for example, see [Rud91]*p.122. We utilize the following module version of the Stone-Weierstrass theorem, which is a special case of theorems proved by Nachbin and Prolla. We give a proof of the result to make this paper as self-contained as possible.

Lemma 6.2 (c.f. [Nachbin1965]*p.54, Theorem 1, [Prolla1994]*Theorem 9).

Let AA be a dense \mathbb{C}-subalgebra of C0(Y)C_{0}(Y). Let pAp^{*}A be a \mathbb{C}-subalgebra of C0(X)C_{0}(X) which consists of the pullback of elements in AA by pp. Let MM be a subspace of C0(X)C_{0}(X) which is a pAp^{*}A-module with respect to pointwise multiplication. Then, for any fC0(X)f\in C_{0}(X) we have

dX(f,M)=supyYdp1(y)(f|p1(y),M|p1(y)),d_{X}(f,M)=\sup_{y\in Y}{d_{p^{-1}(y)}(f|p^{-1}(y),M|p^{-1}(y))},

where M|p1(y)M|p^{-1}(y) denotes the set of functions obtained by restriction of elements in MM to the compact set p1(y)p^{-1}(y) and the distance in the right hand side is taken in C(p1(y))C(p^{-1}(y)).

Proof.

By one-point compactification, we can assume that XX and YY are compact and the space AA and MM contain constant functions. Also, by replacing YY by the image of XX, we can assume that pp is surjective.

We obviously have the inequality

dX(f,M)supyYdp1(y)(f|p1(y),M|p1(y)),d_{X}(f,M)\geq\sup_{y\in Y}{d_{p^{-1}(y)}(f|p^{-1}(y),M|p^{-1}(y))},

so we will prove the converse. Let DD denote the right hand side of this inequality and take ϵ>0\epsilon>0. It suffices to show that left hand side is less than D+ϵD+\epsilon.

By the definition of supremum, for each yYy\in Y, there exists an element gyMg_{y}\in M such that

dp1(y)(f|p1(y),gy|p1(y))<D+ϵ.d_{p^{-1}(y)}(f|p^{-1}(y),g_{y}|p^{-1}(y))<D+\epsilon.

By a simple argument using pp is a closed map, we can take an open neighborhood yVyy\in V_{y} such that on Uy=p1(Vy)U_{y}=p^{-1}(V_{y}) we have

dUy(f|Uy,gy|Uy)<D+ϵ.d_{U_{y}}(f|U_{y},g_{y}|U_{y})<D+\epsilon.

Since the open subsets VyV_{y} cover the space YY, we can take a finite open subcover {Vi=Vyi|i=1,2,,n}\{V_{i}=V_{y_{i}}\,|\,i=1,2,...,n\}. We can also take a partition of unity ρi\rho_{i} associated with this cover.

Set gi=gyig_{i}=g_{y_{i}} and

g=ip(ρi)giC0(X).g=\sum_{i}p^{*}(\rho_{i})g_{i}\in C_{0}(X).

Then, we have

|f(x)g(x)|\displaystyle\lvert f(x)-g(x)\rvert =|iρi(p(x))(f(x)gi(x))|\displaystyle=\lvert\sum_{i}\rho_{i}(p(x))(f(x)-g_{i}(x))\rvert
iρi(p(x))|f(x)gi(x)|\displaystyle\leq\sum_{i}\rho_{i}(p(x))\lvert f(x)-g_{i}(x)\rvert
<(iρi)(p(x))(D+ϵ)\displaystyle<(\sum_{i}\rho_{i})(p(x))(D+\epsilon)
=D+ϵ.\displaystyle=D+\epsilon.

By usual Stone–Weierstrass’ theorem, we can approximate each ρi\rho_{i} by an element σi\sigma_{i} of AA. As MM is a pAp^{*}A-module by multiplication, the function

h=ip(σi)gih=\sum_{i}p^{*}(\sigma_{i})g_{i}

belongs to MM and approximates gg. Hence the result. ∎

Remark 6.3.

The results by Nachbin and Prolla can be applied to more general settings, see [Nachbin1965]*p.54, Theorem 1 and [Prolla1994]*Definition 2, Theorem 9.

6.2. Approximation on measure spaces

Let X,YX,Y be locally compact Hausdorff spaces and p:XYp\colon X\to Y be a proper map. Let μX\mu_{X} be a Radon measure on XX and we denote the pushforward measure pμXp_{*}\mu_{X} by μY\mu_{Y}. Then, μY\mu_{Y} is also a Radon measure on YY.

Definition 6.4.

Let (Z,μ)(Z,\mu) be a topological space with a positive Borel measure. A function ϕ:Z\phi\colon Z\to\mathbb{C} is said to be μ\mu-Riemann integrable i it is bounded, compactly supported and its set of discontinuous points is μ\mu-null set. We denote the set of μ\mu-Riemann integrable functions by (Z,μ)\mathcal{RI}(Z,\mu). We assume that ZZ is a locally compact Hausdorff space and μ\mu is a positive Radon measure on ZZ. Let 1(Z,μ)\mathcal{L}^{1}(Z,\mu) denote the set of μ\mu-integrable functions on ZZ.

Example 6.5.

The characteristic function 𝟏U\mathbf{1}_{U} of an open set UU is μ\mu-Riemann integrable if and only if UU is relatively compact and its boundary U\partial U is a μ\mu-null set.

Definition 6.6.

Let MM be a subspace of 1(X,μX)\mathcal{L}^{1}(X,\mu_{X}) such that any element mMm\in M is μX\mu_{X}-integrable and stable under pointwise complex conjugation. Let τM\tau_{M} denote the linear topology on 1(X,μX)\mathcal{L}^{1}(X,\mu_{X}) whose fundamental system of neighborhoods of 0 is generated by

Uϵ={h1(X,μX)|There exists mM such that |h|m,μX(m)<ϵ},U_{\epsilon}=\{h\in\mathcal{L}^{1}(X,\mu_{X})\,|\,\\ \text{There exists $m\in M$ such that }\lvert h\rvert\leq m,\mu_{X}(m)<\epsilon\},

for ϵ>0\epsilon>0.

Lemma 6.7.

Let M¯τM\overline{M}^{\tau_{M}} denote the closure of MM with respect to τM\tau_{M}. Then, we have τM=τM¯τM\tau_{M}=\tau_{\overline{M}^{\tau_{M}}}.

Proof.

As MM is stable under complex conjugation, it suffices to show the lemma after replacing 1(X,μX)\mathcal{L}^{1}(X,\mu_{X}) with the subspace 1(X,μX,)\mathcal{L}^{1}(X,\mu_{X},\mathbb{R}) of real valued functions and MM with the subspace MM_{\mathbb{R}} of real valued functions in MM. If we consider the topology defined by M¯τM\overline{M}^{\tau_{M}}, we write the open subset UϵU_{\epsilon} as UϵU^{\prime}_{\epsilon}, i.e.

Uϵ={h1(X,μX)|There exists mM¯τM such that |h|m,μX(m)<ϵ}.U^{\prime}_{\epsilon}=\{h\in\mathcal{L}^{1}(X,\mu_{X})\,|\,\\ \text{There exists $m^{\prime}\in\overline{M}^{\tau_{M}}$ such that }\lvert h\rvert\leq m^{\prime},\mu_{X}(m^{\prime})<\epsilon\}.

Then, it is trivial that we have UϵUϵU_{\epsilon}\subset U^{\prime}_{\epsilon}. We claim that if ϵ<ϵ\epsilon^{\prime}<\epsilon, then we have UϵUϵU^{\prime}_{\epsilon^{\prime}}\subset U_{\epsilon}. We assume that a function h1(X,μX)h\in\mathcal{L}^{1}(X,\mu_{X}) is in UϵU^{\prime}_{\epsilon^{\prime}}. Then, we can find mM¯τMm^{\prime}\in\overline{M}^{\tau_{M}} and we have

|h|m,\displaystyle\lvert h\rvert\leq m^{\prime},
μX(m)<ϵ.\displaystyle\mu_{X}(m^{\prime})<\epsilon^{\prime}.

We set δ=ϵϵ\delta=\epsilon-\epsilon^{\prime}. Then, we have δ>0\delta>0. As the function mm^{\prime} is in M¯τM\overline{M}^{\tau_{M}}, there exist functions n1Mn_{1}\in M and n2Mn_{2}\in M such that |mn1|n2\lvert m^{\prime}-n_{1}\rvert\leq n_{2} and μX(n2)<12δ\mu_{X}(n_{2})<\frac{1}{2}\delta. Then, we have

|h|n1+n2\displaystyle\lvert h\rvert\leq n_{1}+n_{2}

and

μX(n1+n2)<ϵ+2×12δ=ϵ.\displaystyle\mu_{X}(n_{1}+n_{2})<\epsilon^{\prime}+2\times\frac{1}{2}\delta=\epsilon.

Thus, we have UϵUϵU_{\epsilon^{\prime}}^{\prime}\subset U_{\epsilon}. ∎

Corollary 6.8.

We have M¯τM¯𝒯M¯τM=M¯τM\overline{\overline{M}^{\tau_{M}}}^{\mathcal{T}_{\overline{M}^{\tau_{M}}}}=\overline{M}^{\tau_{M}}.

The following lemma is [Sau97]*Lemme 2.3.

Lemma 6.9.

Let AC0(Y)A\subset C_{0}(Y) be a dense \mathbb{C}-subalgebra. Let M1(X,μX)M\subset\mathcal{L}^{1}(X,\mu_{X}) be a pAp^{*}A-module with respect to pointwise multiplication and stable under complex conjugation, such that any element mMm\in M is μX\mu_{X}-integrable. We assume that MM also satisfies the following condition: For any function mMm\in M, there exists a function nMn\in M such that |m|n\lvert m\rvert\leq n.

Then, for any function mMm\in M and cC0(Y)c\in C_{0}(Y), the function (pc)m(p^{*}c)m is in M¯τM\overline{M}^{\tau_{M}}.

Proof.

We take a function nMn\in M as in the hypothesis of the lemma. There exists a function aAa\in A such that |ca|ϵ\lvert c-a\rvert\leq\epsilon with ϵ\epsilon sufficiently small with respect to μX(n)\mu_{X}(n). Then, we have

|(pc)m(pa)m|ϵn.\displaystyle\lvert(p^{*}c)m-(p^{*}a)m\rvert\leq\epsilon n.

Hence the result. ∎

Remark 6.10.

Note that the proof of Lemma 6.9 and Corollary 6.8 does not use the Hausdorff property of XX.

Theorem 6.11.

Let AA be a dense \mathbb{C}-subalgebra of C0(Y)C_{0}(Y). Let MC0(X)1(X,μX)M\subset C_{0}(X)\cap\mathcal{L}^{1}(X,\mu_{X}) be a pAp^{*}A-module with respect to pointwise multiplication. We assume the following conditions:

  1. (1)

    MM is stable under pointwise complex conjugation.

  2. (2)

    The restriction of the elements of MM to p1(y)p^{-1}(y) is dense in C(p1(y))C(p^{-1}(y)) for every yYy\in Y.

  3. (3)

    For any function mMm\in M, there exists a function nMn\in M such that |m|n\lvert m\rvert\leq n.

Then, we have (X,μX)M¯τM\mathcal{RI}(X,\mu_{X})\subset\overline{M}^{\tau_{M}}.

Proof.

It suffices to show that we have (X,μX)M¯τM¯𝒯M¯τM\mathcal{RI}(X,\mu_{X})\subset\overline{\overline{M}^{\tau_{M}}}^{\mathcal{T}_{\overline{M}^{\tau_{M}}}} by Corollary 6.8. We will write down this condition. Take any function ϕ(X,μX)\phi\in\mathcal{RI}(X,\mu_{X}) and any positive number ϵ>0\epsilon>0. We want to show that there exist functions m1,m2M¯τMm^{\prime}_{1},m^{\prime}_{2}\in\overline{M}^{\tau_{M}} such that we have

|ϕm1|\displaystyle\lvert\phi-m^{\prime}_{1}\rvert m2\displaystyle\leq m^{\prime}_{2}
μX(m2)\displaystyle\mu_{X}(m^{\prime}_{2}) <ϵ.\displaystyle<\epsilon.

We fix ϵ>0\epsilon>0. By assumption, the support of ϕ\phi is compact. Thus, we can take a positive function cCc(Y)c\in C_{c}(Y) such that on the support of ϕ\phi, we have pc=1p^{*}c=1. By the assumptions (2),(3)(2),(3) for MM, it is easy to find a positive function mcMm_{c}\in M such that we have pcmcp^{*}c\leq m_{c} on XX. Indeed, we claim that for any point xXx\in X, we can find mxMm_{x}\in M such that mx(x)>0m_{x}(x)>0. This claim clearly proves the existence of the function mcm_{c}. By the assumption (2)(2), we can take a function nxMn_{x}\in M such that we have nx(x)0n_{x}(x)\neq 0. Then, the assumption (3)(3) implies that there exists mxMm_{x}\in M such that we have |nx|mx\lvert n_{x}\rvert\leq m_{x}. We have mx(x)|nx(x)|>0m_{x}(x)\geq\lvert n_{x}(x)\rvert>0 and this proves the claim.

We take a positive number ϵ1>0\epsilon_{1}>0 such that

ϵ1<13mc+μX(mc)ϵ.\displaystyle\epsilon_{1}<\frac{1}{3||m_{c}||_{\infty}+\mu_{X}(m_{c})}\epsilon.

By [Bou04]*IV 12 Lemma 5, we obtain functions c1,c2Cc(X)c_{1},c_{2}\in C_{c}(X) such that

|ϕc1|\displaystyle\lvert\phi-c_{1}\rvert c2\displaystyle\leq c_{2}
μX(c2)\displaystyle\mu_{X}(c_{2}) ϵ1.\displaystyle\leq\epsilon_{1}.

We apply Lemma 6.2 to obtain functions m1,m2Mm_{1},m_{2}\in M such that

|c1m1|\displaystyle\lvert c_{1}-m_{1}\rvert <ϵ1\displaystyle<\epsilon_{1}
|c2m2|\displaystyle\lvert c_{2}-m_{2}\rvert <ϵ1.\displaystyle<\epsilon_{1}.

At this stage, we have

|ϕ(pc)m1|\displaystyle\lvert\phi-(p^{*}c)m_{1}\rvert (pc)|ϕc1|+(pc)|c1m1|\displaystyle\leq(p^{*}c)\lvert\phi-c_{1}\rvert+(p^{*}c)\lvert c_{1}-m_{1}\rvert
(pc)m2+(pc)|c1m1|+(pc)|c2m2|\displaystyle\leq(p^{*}c)m_{2}+(p^{*}c)\lvert c_{1}-m_{1}\rvert+(p^{*}c)\lvert c_{2}-m_{2}\rvert
(pc)m2+ϵ1mc+ϵ1mc\displaystyle\leq(p^{*}c)m_{2}+\epsilon_{1}m_{c}+\epsilon_{1}m_{c}
=(pc)m2+2ϵ1mc.\displaystyle=(p^{*}c)m_{2}+2\epsilon_{1}m_{c}.

Here, we have

μX((pc)m2)\displaystyle\mu_{X}((p^{*}c)m_{2}) μX((pc)|m2c2|)+μX((pc)c2)\displaystyle\leq\mu_{X}((p^{*}c)\lvert m_{2}-c_{2}\rvert)+\mu_{X}((p^{*}c)c_{2})
ϵ1μX(pc)+ϵ1c.\displaystyle\leq\epsilon_{1}\mu_{X}(p^{*}c)+\epsilon_{1}||c||_{\infty}.

Thus, we have

μX((pc)m2+2ϵ1mc)(3mc+μX(mc))ϵ1<ϵ\displaystyle\mu_{X}((p^{*}c)m_{2}+2\epsilon_{1}m_{c})\leq(3||m_{c}||_{\infty}+\mu_{X}(m_{c}))\epsilon_{1}<\epsilon

and (pc)m1,(pc)m2+2ϵ1mcM¯τM(p^{*}c)m_{1},(p^{*}c)m_{2}+2\epsilon_{1}m_{c}\in\overline{M}^{\tau_{M}} by Lemma 6.9. Hence the result. ∎

Example 6.12.

We take

  • the space Θdisc(G)temp\Theta_{\mathrm{disc}}(G)_{\mathrm{temp}} as XX,

  • the space Θ(G)herm\Theta(G)_{\mathrm{herm}} as YY,

  • the map infG\inf_{G} as pp,

  • the Radon measure μG\mu^{G} as μX\mu_{X},

  • the algebra 𝒜(G)\mathcal{A}(G) as AA and

  • the module 𝒯(G)\mathcal{FT}(G) as MM, see definition 5.3.

The space 𝒯(G)\mathcal{FT}(G) is an 𝒜(G)\mathcal{A}(G)-module by the Paley-Wiener theorem [ClozelDelorme1990]*Théorème 1 if F=F=\mathbb{R} and [Renard2010-p-adiques]*VI.10.3 Théorème if FF is a pp-adic field. Then, Lemmas in [Sau97]*Lemme 3.1, 3.4, 3.5, Lemma 2.6, Lemma 3.5, Lemma 3.8, Lemma 5.5, and the linear independence of irreducible characters imply that this sextuple satisfies the condition of Theorem 6.11. Also, the similar result holds for any reductive group schemes over a product of local fields.

Definition 6.13.

Let Irrunit(G)good\operatorname{Irr}_{\mathrm{unit}}(G)_{\mathrm{good}} denote the subset of Irrunit(G)\operatorname{Irr}_{\mathrm{unit}}(G) which consists of the parabolic inductions associated to a discrete pair (M,σ)(M,\sigma) with σ\sigma unitary and regular in the sense of [Sau97]*p.172, line 30. Let Irrunit(G)bad\operatorname{Irr}_{\mathrm{unit}}(G)_{\mathrm{bad}} denote the complement in Irrunit(G)\operatorname{Irr}_{\mathrm{unit}}(G) of Irrunit(G)good\operatorname{Irr}_{\mathrm{unit}}(G)_{\mathrm{good}}.

Remark 6.14.

We have μG(Irrunit(G)bad)=0\mu^{G}(\operatorname{Irr}_{\mathrm{unit}}(G)_{\mathrm{bad}})=0 by Theorem 4.2.

Corollary 6.15.

Let GG be a reductive group scheme over a product of local fields of characteristic 0. Let ϕ\phi be a function in (Irrtemp(G),μG)\mathcal{RI}(\operatorname{Irr}_{\mathrm{temp}}(G),\mu^{G}) and ϵ\epsilon be a positive real number. Then, there exist functions h1,h2Cc(G)h_{1},h_{2}\in C_{c}^{\infty}(G) such that we have

|ϕh1^|\displaystyle\lvert\phi-\widehat{h_{1}}\rvert h2^,\displaystyle\leq\widehat{h_{2}},
μG(h2^)\displaystyle\mu^{G}(\widehat{h_{2}}) <ϵ,\displaystyle<\epsilon,

on the set Irrunit(G)good\operatorname{Irr}_{\mathrm{unit}}(G)_{\mathrm{good}}.

Proof.

We take a function ϕ(Irrtemp(G),μG)\phi\in\mathcal{RI}(\operatorname{Irr}_{\mathrm{temp}}(G),\mu^{G}). Then, by Theorem 6.11 and Example 6.12, we can find functions h1,h2Cc(G)h_{1},h_{2}\in C_{c}^{\infty}(G) such that we have

|ϕh1^|\displaystyle\lvert\phi-\widehat{h_{1}}\rvert h2^,\displaystyle\leq\widehat{h_{2}},
μG(h2^)\displaystyle\mu^{G}(\widehat{h_{2}}) <ϵ,\displaystyle<\epsilon,

on the subset Irrunit(G)good{\operatorname{Irr}_{\mathrm{unit}}(G)}_{\mathrm{good}} in Irrunit(G)\operatorname{Irr}_{\mathrm{unit}}(G), through the identification explained in Lemma 5.6. Hence the result. ∎

7. The Sauvageot density principle

In this section, we will prove the following result.

Proposition 7.1 ([Sau97]*Théorème 5.4).

Let GG be a reductive group schemes over a product of local fields. Let CC be a compact subset of Θ(G)\Theta(G) and let ϵ>0\epsilon>0 be a positive real number. Then, there exists a function hCc(G)h\in C_{c}^{\infty}(G) which satisfies the following conditions.

  1. (1)

    We have h^0\widehat{h}\geq 0 on Irrunit(G)\operatorname{Irr}_{\mathrm{unit}}(G).

  2. (2)

    We have μG(h^)ϵ\mu^{G}(\widehat{h})\leq\epsilon.

  3. (3)

    For any πIrrunit(G)bad\pi\in\operatorname{Irr}_{\mathrm{unit}}(G)_{\mathrm{bad}} with infG(π)C\inf_{G}(\pi)\in C, we have

    h^(π)1.\displaystyle\widehat{h}(\pi)\geq 1.

Before proving this theorem, we first show that this theorem and Theorem 6.11 imply the following Sauvageot density principle.

Theorem 7.2 (Sauvageot’s density principle, [Sau97]*Théorème 7.3).

Let ϕ\phi be a bounded, Borel measurable, and compactly supported function on Irrtemp(G)Irrunit(G)\operatorname{Irr}_{\mathrm{temp}}(G)\subset\operatorname{Irr}_{\mathrm{unit}}(G) such that the set of discontinuity of ϕ\phi is a μG\mu^{G}-null set. Let ϵ>0\epsilon>0 be a positive real number. Then, there exist functions h1,h2Cc(G)h_{1},h_{2}\in C_{c}^{\infty}(G) such that we have

|ϕh1^|\displaystyle\lvert\phi-\widehat{h_{1}}\rvert h2^\displaystyle\leq\widehat{h_{2}}
μG(h2^)\displaystyle\mu^{G}(\widehat{h_{2}}) <ϵ\displaystyle<\epsilon

on Irrunit(G)\operatorname{Irr}_{\mathrm{unit}}(G).

Proof that Proposition 7.1 implies Theorem 7.2.

We take a function cCc(Θ(G))c\in C_{c}(\Theta(G)) such that c=infGcc^{\prime}=\inf_{G}^{*}c is equal to 11 on the support of ϕ\phi. We take a positive number ϵ1\epsilon_{1} such that

ϵ1<12(c+1)ϵ.\displaystyle\epsilon_{1}<\frac{1}{2(||c||_{\infty}+1)}\epsilon.

We take the functions h1,h2Cc(G)h^{\prime}_{1},h^{\prime}_{2}\in C_{c}^{\infty}(G) satisfying the condition of Corollary 6.15 for ϕ\phi and ϵ1\epsilon_{1}. Then, we have

|ϕh1^|\displaystyle\lvert\phi-\widehat{h^{\prime}_{1}}\rvert h2^\displaystyle\leq\widehat{h^{\prime}_{2}}
μG(h2^)\displaystyle\mu^{G}(\widehat{h^{\prime}_{2}}) <ϵ1\displaystyle<\epsilon_{1}

on the complement of Irrunit(G)bad\operatorname{Irr}_{\mathrm{unit}}(G)_{\mathrm{bad}}. Then, we have

|ϕch1^|ch2^.\displaystyle\lvert\phi-c^{\prime}\widehat{h^{\prime}_{1}}\rvert\leq c^{\prime}\widehat{h^{\prime}_{2}}.

By Proposition 7.1, we can take an element h3Cc(G)h^{\prime}_{3}\in C_{c}^{\infty}(G) such that we have

|ϕch1^|+ch2^\displaystyle\lvert\phi-c^{\prime}\widehat{h^{\prime}_{1}}\rvert+c^{\prime}\widehat{h^{\prime}_{2}} h3^\displaystyle\leq\widehat{h^{\prime}_{3}}
μG(h3^)\displaystyle\mu^{G}(\widehat{h^{\prime}_{3}}) <ϵ1\displaystyle<\epsilon_{1}

on the set Irrunit(G)bad\operatorname{Irr}_{\mathrm{unit}}(G)_{\mathrm{bad}}. Then, we have

|ϕch1^|\displaystyle\lvert\phi-c^{\prime}\widehat{h^{\prime}_{1}}\rvert ch2^+h3^\displaystyle\leq c^{\prime}\widehat{h^{\prime}_{2}}+\widehat{h^{\prime}_{3}}
μG(ch2^+h3^)\displaystyle\mu^{G}(c^{\prime}\widehat{h^{\prime}_{2}}+\widehat{h^{\prime}_{3}}) <(c+1)ϵ1<12ϵ.\displaystyle<(||c||_{\infty}+1)\epsilon_{1}<\frac{1}{2}\epsilon.

By applying Corollary 6.8, Lemma 6.9 and Remark 6.10, we can find functions h1,h2Cc(G)h_{1},h_{2}\in C_{c}^{\infty}(G) such that we have

|ϕh1^|\displaystyle\lvert\phi-\widehat{h_{1}}\rvert h2^\displaystyle\leq\widehat{h_{2}}
μG(h2^)\displaystyle\mu^{G}(\widehat{h_{2}}) <ϵ.\displaystyle<\epsilon.

Hence the result. ∎

Lemma 7.3.

[Sau97]*Corollaire 6.2 Proposition 7.1 for reductive groups over local fields implies the general case of Proposition 7.1.

Thus, we return to the setting where GG is a reductive group over a local field FF.

Lemma 7.4 (c.f. [Sau97]*Lemme 2.6).

Let VV be a finite dimensional \mathbb{R}-vector space and let Γ\Gamma be a discrete subgroup of VV. We set X=V/ΓX=V/\Gamma. For any compact set CC of XX and a family of open neighborhood UxU_{x} for each point xCx\in C, there exists a finite covering {Wi}iI\{W_{i}\}_{i\in I} of CC by relatively compact open subsets WiW_{i} which satisfy the following property.

  1. (1)

    For any iIi\in I, there exists a point xCx\in C such that Wi¯Vx\overline{W_{i}}\subset V_{x}.

  2. (2)

    Each point of Wi¯\overline{W_{i}} is contained in at most 2dimX2^{\dim X} elements of {Wj¯}jI\{\overline{W_{j}}\}_{j\in I}.

Proof.

We may assume that V=nV=\mathbb{R}^{n} and Γ=m\Gamma=\mathbb{Z}^{m} for mnm\leq n. Let {pi}i=1,,n\{p_{i}\}_{i=1,...,n} be the standard coordinate functions on n\mathbb{R}^{n}. We set 𝒰={Ux}xK\mathcal{U}=\{U_{x}\}_{x\in K}. Let KϵK_{\epsilon} denote the closed ϵ\epsilon-neighborhood of KK. Then, as KK is compact, there exists an ϵ>0\epsilon>0 such that

KK=Kϵ𝒰.\displaystyle K\subset K^{\prime}=K_{\epsilon}\subset\bigcup\mathcal{U}.

We consider the subgroup 12kn/m\frac{1}{2^{k}}\mathbb{Z}^{n}/\mathbb{Z}^{m} for sufficiently large kk. This defines a family of boxes in XX with width 12k\frac{1}{2^{k}}, where a box in XX with width rr is the projection of the sets in n\mathbb{R}^{n} of the form

{yX||pi(y)pi(c)|<r/2,i=1,2,,n}\displaystyle\{y\in X\,|\,\lvert p_{i}(y)-p_{i}(c)\rvert<r/2,i=1,2,...,n\}

for some cXc\in X and r>0r>0. For each element BB in the family above, we take a box BB^{\prime} with width 111012k\frac{11}{10}\frac{1}{2^{k}}, roughly speaking, the thickened box BB^{\prime} obtained from BB. We will denote the family of the thickened boxes by k\mathcal{B}_{k}.

We set k(K)={Bk|B¯K}\mathcal{B}_{k}(K)=\{B\in\mathcal{B}_{k}\,|\,\overline{B}\cap K\neq\emptyset\}. This set is obviously finite, as the set KK is compact. By enlarging kk, we may assume that

KBk(K)BBk(K)B¯K𝒰.\displaystyle K\subset\bigcup_{B\in\mathcal{B}_{k}(K)}B\subset\bigcup_{B\in\mathcal{B}_{k}(K)}\overline{B}\subset K^{\prime}\subset\bigcup\mathcal{U}.

As KK^{\prime} is compact, the set Bk(K)B¯\bigcup_{B\subset\mathcal{B}_{k}(K)}\overline{B} is compact in XX. Thus, applying the argument using the Lebesgue number for this set and the covering induced by 𝒰\mathcal{U}, we divide each box in k(K)\mathcal{B}_{k}(K) into 2n2^{n} boxes several times to obtain a new set of boxes \mathcal{B} in XX, such that for any element BB\in\mathcal{B}, there exists an element U𝒰U\in\mathcal{U} such that B¯U\overline{B}\subset U. We take the set \mathcal{B} as the set {Wi}i\{W_{i}\}_{i} in the statement. Then, the first property is satisfied by definition. By noting that the projection VV/ΓV\twoheadrightarrow V/\Gamma is a local isomorphism and counting locally, it is easy to see that the second property holds true for this set. ∎

We recall the notion of vertical bands with compact real part, which are used in the proof.

Definition 7.5.

First, assume that F=F=\mathbb{R}. In this case, by a vertical band with compact real part in 𝔥\mathfrak{h}^{*}\otimes_{\mathbb{R}}\mathbb{C}, we mean a subset BB of 𝔥\mathfrak{h}^{*}\otimes_{\mathbb{R}}\mathbb{C} with compact real part with respect to 𝔥\mathfrak{h}^{*}_{\mathbb{R}}. We call the projection of a vertical band with compact real part in 𝔥\mathfrak{h}^{*}\otimes_{\mathbb{R}}\mathbb{C} onto Θ(G)\Theta(G), a vertical band with compact real part in Θ(G)\Theta(G).

We assume that FF is a pp-adic field. By a vertical band with compact real part in 𝔞M,\mathfrak{a}^{*}_{M,\mathbb{C}}, we mean a subset of 𝔞M,\mathfrak{a}^{*}_{M,\mathbb{C}} with compact real part with respect to X(M)X^{*}(M)\otimes_{\mathbb{Z}}\mathbb{R}. We call the finite union of images of vertical bands with compact real part in 𝔞M,\mathfrak{a}^{*}_{M,\mathbb{C}} under the map (2.2)(2.2), a vertical band with compact real part in Θ(G)\Theta(G).

The proof of Proposition 7.1.

The proof we give is essentially the same as that in [Sau97]*Théorème 5.4, but we fix minor notational errors pointed out in [Shi12]*Appendix A.

We take arbitrary positive numbers ϵ1,ϵ2,ϵ3\epsilon_{1},\epsilon_{2},\epsilon_{3}. Also, we fix a compact neighborhood C1C_{1} of CC.

By [Sau97]*Lemme 5.1, for any point θC\theta\in C, we can find an open neighborhood Vθ,ϵ1C1V_{\theta,\epsilon_{1}}\subset C_{1} of θ\theta and a function hθ,ϵ1Cc(G)h_{\theta,\epsilon_{1}}\in C_{c}^{\infty}(G), such that the following condition holds: For any irreducible representation π\pi of GG with infG(π)Vθ,ϵ1\inf_{G}(\pi)\in V_{\theta,\epsilon_{1}}, we have

0\displaystyle 0\leq hθ,ϵ1^(π)ϵ1,πIrrunit(G)good,\displaystyle\widehat{h_{\theta,\epsilon_{1}}}(\pi)\leq\epsilon_{1},\quad\pi\in\operatorname{Irr}_{\mathrm{unit}}(G)_{\mathrm{good}},
1\displaystyle 1\leq hθ,ϵ1^(π)cG,πIrrunit(G)bad,\displaystyle\widehat{h_{\theta,\epsilon_{1}}}(\pi)\leq c_{G},\quad\pi\in\operatorname{Irr}_{\mathrm{unit}}(G)_{\mathrm{bad}},

where cGc_{G} is a positive constant depending only on GG.

Then, by Lemma 7.4, there exists a subset {θi}i=1,2,,N(ϵ1)\{\theta_{i}\}_{i=1,2,...,N(\epsilon_{1})} of CC and an open covering {Wi}i=1,,N(ϵ1)\{W_{i}\}_{i=1,...,N(\epsilon_{1})} of CC such that

i=1,,N(ϵ1)𝟏Wi¯\displaystyle\sum_{i=1,...,N(\epsilon_{1})}\mathbf{1}_{\overline{W_{i}}} cG𝟏C1\displaystyle\leq c^{\prime}_{G}\mathbf{1}_{C_{1}}
Wi¯\displaystyle\overline{W_{i}} Vθi,ϵ1,\displaystyle\subset V_{\theta_{i},\epsilon_{1}},

where 𝟏Z\mathbf{1}_{Z} denotes the characteristic function of ZZ and cGc^{\prime}_{G} is a positive constant which only depends on GG.

We set

M(ϵ1)=supi=1,,N(ϵ1),πIrrunit(G)|hθi,ϵ1^(π)|.\displaystyle M(\epsilon_{1})=\sup_{i=1,...,N(\epsilon_{1}),\pi\in\operatorname{Irr}_{\mathrm{unit}}(G)}\lvert\widehat{h_{\theta_{i},\epsilon_{1}}}(\pi)\rvert.

By [Sau97]*Lemme 3.4, there exists a vertical band BB with compact real part in Θ(G)\Theta(G) containing C1C_{1} such that the functions ϕθi,ϵ1^(π)\widehat{\phi_{\theta_{i},\epsilon_{1}}}(\pi) are zero if πIrrunit(G)\pi\in\operatorname{Irr}_{\mathrm{unit}}(G) and infG(π)B\inf_{G}(\pi)\not\in B. We will choose an open neighborhood WiW^{\prime}_{i} of Wi¯\overline{W_{i}} in Vθ,ϵ1V_{\theta,\epsilon_{1}} such that we have

infG(μG)(i=1,,N(ϵ1)(WiWi¯))ϵ2.\displaystyle{\mathrm{inf}_{G}}_{*}(\mu^{G})(\bigcup_{i=1,...,N(\epsilon_{1})}(W_{i}^{\prime}\setminus\overline{W_{i}})\,)\leq\epsilon_{2}.

This follows from the outer regularity of the measure infG(μG){\mathrm{inf}_{G}}_{*}(\mu^{G}), see [Rud87]*Theorem 2.18.

We choose a function ai𝒜(G)a_{i}\in\mathcal{A}(G) by applying [Sau97]*Lemme 5.2. Then, the function aia_{i} is real on Θ(G)herm\Theta(G)_{\mathrm{herm}} and we have

0ai(θ)\displaystyle 0\leq a_{i}(\theta) 1+ϵ3,θBΘ(G)herm,\displaystyle\leq 1+\epsilon_{3},\quad\theta\in B\cap\Theta(G)_{\mathrm{herm}},
ai(θ)\displaystyle a_{i}(\theta) 1,θWi¯Θ(G)herm,\displaystyle\geq 1,\quad\theta\in\overline{W_{i}}\cap\Theta(G)_{\mathrm{herm}},
ai(θ)\displaystyle a_{i}(\theta) ϵ3,θ(BWi)Θ(G)herm.\displaystyle\leq\epsilon_{3},\quad\theta\in(B\setminus W^{\prime}_{i})\cap\Theta(G)_{\mathrm{herm}}.

We consider the function

i=1,,N(ϵ1)(infGai)hθi,ϵ1^.\displaystyle\sum_{i=1,...,N(\epsilon_{1})}(\mathrm{inf}_{G}^{*}a_{i})\widehat{h_{\theta_{i},\epsilon_{1}}}.

First, this function is of the form h1^(G)\widehat{h_{1}}\in\mathcal{F}(G). We consider the value of this function on each set:

  • On the set infG1(iWi¯)Irrunit(G)good\inf_{G}^{-1}(\bigcup_{i}\overline{W_{i}})\cap\operatorname{Irr}_{\mathrm{unit}}(G)_{\mathrm{good}}, we have

    |h1^|(1+ϵ3)ϵ1cG(𝟏C1infG).\displaystyle\lvert\widehat{h_{1}}\rvert\leq(1+\epsilon_{3})\epsilon_{1}c^{\prime}_{G}(\mathbf{1}_{C_{1}}\circ\mathrm{inf}_{G}).
  • On the set infG1(iWi¯)Irrunit(G)bad\inf_{G}^{-1}(\bigcup_{i}\overline{W_{i}})\cap\operatorname{Irr}_{\mathrm{unit}}(G)_{\mathrm{bad}}, we have

    |h1^|(1+ϵ3)cGcG(𝟏C1infG)(𝟏Irrunit(G)bad).\displaystyle\lvert\widehat{h_{1}}\rvert\leq(1+\epsilon_{3})c_{G}c^{\prime}_{G}(\mathbf{1}_{C_{1}}\circ\mathrm{inf}_{G})(\mathbf{1}_{\operatorname{Irr}_{\mathrm{unit}}(G)_{\mathrm{bad}}}).
  • On the set infG1(i(WiWi¯))Irrunit(G)\inf_{G}^{-1}(\bigcup_{i}(W^{\prime}_{i}\setminus\overline{W_{i}}))\cap\operatorname{Irr}_{\mathrm{unit}}(G), we have

    |h1^|(1+ϵ3)M(ϵ1)N(ϵ1)(𝟏i=1,,N(ϵ1)(WiWi¯)infG).\displaystyle\lvert\widehat{h_{1}}\rvert\leq(1+\epsilon_{3})M(\epsilon_{1})N(\epsilon_{1})(\mathbf{1}_{\bigcup_{i=1,...,N(\epsilon_{1})}(W_{i}^{\prime}\setminus\overline{W_{i}})}\circ\mathrm{inf}_{G}).
  • Otherwise, we have

    |h1^|ϵ3i|hθi,ϵ1^|.\displaystyle\lvert\widehat{h_{1}}\rvert\leq\epsilon_{3}\sum_{i}\lvert\widehat{h_{\theta_{i},\epsilon_{1}}}\rvert.

We take a positive function h2^\widehat{h_{2}} with h2^1\widehat{h_{2}}\geq 1 on C1C_{1} by applying [Sau97]*Lemme 5.3. We set

h^=h1^+ϵ3M(ϵ1)N(ϵ1)h2^.\displaystyle\widehat{h^{\prime}}=\widehat{h_{1}}+\epsilon_{3}M(\epsilon_{1})N(\epsilon_{1})\widehat{h_{2}}.

By considering whether πinfG1(Vθi,ϵ1)\pi\in\inf_{G}^{-1}(V_{\theta_{i},\epsilon_{1}}) or not, it is easy to see that we have h^0\widehat{h^{\prime}}\geq 0 on C1C_{1}. Similarly, on the set infG1(C1)Irrunit(G)bad\inf_{G}^{-1}(C_{1})\cap\operatorname{Irr}_{\mathrm{unit}}(G)_{\mathrm{bad}}, we have h^1\widehat{h^{\prime}}\geq 1. By the estimate above, if we choose sufficiently small numbers ϵ1,ϵ2,ϵ3\epsilon_{1},\epsilon_{2},\epsilon_{3}, we have

μG(h^)<ϵ/2.\displaystyle\mu^{G}(\widehat{h^{\prime}})<\epsilon/2.

By the argument starting from [Sau97]*p.178, line 12, we can replace this function hh^{\prime} to obtain a new function hCc(G)h\in C_{c}^{\infty}(G) which satisfies the statement of our proposition. As this procedure is explained completely in [Sau97], we do not repeat the argument. ∎

8. Applications

8.1. Some lemmas

Let 𝔾\mathbb{G} be a reductive group scheme over a product AA of local fields of characteristic 0. We set G=𝔾(A)G=\mathbb{G}(A). Let μG\mu^{G} denote the Plancherel measure of GG. Let {μnG}n1\{\mu^{G}_{n}\}_{n\geq 1} be a sequence of positive Borel measures on Irrunit(G)\operatorname{Irr}_{\mathrm{unit}}(G).

Lemma 8.1 ([Sau97]*Proposition 1.3).

Let μnG\mu^{G}_{n} be a sequence of positive Borel measures on Irrunit(G)\operatorname{Irr}_{\mathrm{unit}}(G). We assume that for any hCc(G)h\in C_{c}^{\infty}(G), we have

limnμnG(h^)=μG(h^).\displaystyle\lim_{n\to\infty}\mu^{G}_{n}(\widehat{h})=\mu^{G}(\widehat{h}).

Then, it follows that

limnμnG(ϕ)=μG(ϕ)\displaystyle\lim_{n\to\infty}\mu^{G}_{n}(\phi)=\mu^{G}(\phi)

for any ϕ(Irrtemp(G),μG)\phi\in\mathcal{RI}(\operatorname{Irr}_{\mathrm{temp}}(G),\mu^{G}).

Proof.

Let ϵ\epsilon be any positive real number. By Theorem 7.2, we can find functions h1,h2Cc(G)h_{1},h_{2}\in C_{c}^{\infty}(G) such that

|ϕh1^|h2^,\displaystyle\lvert\phi-\widehat{h_{1}}\rvert\leq\widehat{h_{2}},
μG(h2^)<ϵ.\displaystyle\mu^{G}(\widehat{h_{2}})<\epsilon.

We have

|μnG(ϕ)μG(ϕ)||μnG(ϕ)μnG(h1^)|+|μG(ϕ)μG(h1^)|+|μnG(h1^)μG(h1^)|.\lvert\mu^{G}_{n}(\phi)-\mu^{G}(\phi)\rvert\leq\lvert\mu^{G}_{n}(\phi)-\mu^{G}_{n}(\widehat{h_{1}})\rvert+\lvert\mu^{G}(\phi)-\mu^{G}(\widehat{h_{1}})\rvert\\ +\lvert\mu^{G}_{n}(\widehat{h_{1}})-\mu^{G}(\widehat{h_{1}})\rvert.

By the assumption on h1,h2h_{1},h_{2}, we have

|μnG(ϕ)μnG(h1^)|μnG(h2),\displaystyle\lvert\mu^{G}_{n}(\phi)-\mu^{G}_{n}(\widehat{h_{1}})\rvert\leq\mu^{G}_{n}(h_{2}),
|μG(ϕ)μG(h1^)|μG(h2).\displaystyle\lvert\mu^{G}(\phi)-\mu^{G}(\widehat{h_{1}})\rvert\leq\mu^{G}(h_{2}).

By substituting these inequalities into the previous inequality and considering the limit superior, we obtain

lim supn|μnG(ϕ)μG(ϕ)|2μG(h2^)<2ϵ.\displaystyle\limsup_{n\to\infty}\lvert\mu^{G}_{n}(\phi)-\mu^{G}(\phi)\rvert\leq 2\mu^{G}(\widehat{h_{2}})<2\epsilon.

As we can take ϵ>0\epsilon>0 arbitrarily, we obtain

limnμnG(ϕ)=μG(ϕ).\lim_{n\to\infty}\mu^{G}_{n}(\phi)=\mu^{G}(\phi).\qed

We also give a lemma to globalize local characters. Let Λ=Λ1Λ2\Lambda=\Lambda_{1}\coprod\Lambda_{2} be a set such that Λ2\Lambda_{2} is finite. For λΛ1\lambda\in\Lambda_{1}, let 𝒫λ=(Gλ,Kλ)\mathcal{P}_{\lambda}=(G_{\lambda},K_{\lambda}) be a pair consisting of a locally compact Hausdorff group GλG_{\lambda} and its open compact subgroup KλK_{\lambda}. For λΛ2\lambda\in\Lambda_{2}, let 𝒫λ=Gλ\mathcal{P}_{\lambda}=G_{\lambda} be a locally compact Hausdorff space. Let GG be a restricted product λΛ1(Gλ,Kλ)×λΛ2Gλ\prod_{\lambda\in\Lambda_{1}}^{\prime}(G_{\lambda},K_{\lambda})\times\prod_{\lambda\in\Lambda_{2}}G_{\lambda} of the family 𝒫λ\mathcal{P}_{\lambda}. Then, GG is locally compact. Let Γ\Gamma be a discrete subgroup of GG such that the map ΓGGλ0\Gamma\hookrightarrow G\twoheadrightarrow G_{\lambda_{0}} is injective for an element λ0Λ\lambda_{0}\in\Lambda.

Lemma 8.2.

Let SS be a finite subset of Λ1\Lambda_{1} which does not contain λ0\lambda_{0}. Let χS\chi_{S} be a continuous character of KS=λSKλK_{S}=\prod_{\lambda\in S}K_{\lambda}. Then, there exists a character χ\chi of G/ΓG/\Gamma, such that we have

χKS\displaystyle\chi_{\restriction_{K_{S}}} =χS,\displaystyle=\chi_{S},
χΛ1(S{λ0})Kλ\displaystyle\chi_{\prod_{\Lambda_{1}\setminus(S\cup\{\lambda_{0}\})}K_{\lambda}} =𝟏Λ1(S{λ0})Kλ.\displaystyle=\mathbf{1}_{\prod_{\Lambda_{1}\setminus(S\cup\{\lambda_{0}\})}K_{\lambda}}.
Proof.

By the hypothesis on λ0\lambda_{0}, we have

Γλλ0Kλ={1}.\displaystyle\Gamma\cap\prod_{\lambda\neq\lambda_{0}}K_{\lambda}=\{1\}.

Also, the group λλ0Kλ\prod_{\lambda\neq\lambda_{0}}K_{\lambda} is compact, the image of the map

λλ0KλG/Γ.\displaystyle\prod_{\lambda\neq\lambda_{0}}K_{\lambda}\to G/\Gamma.

is compact and closed. Thus, the subgroup Γλλ0Kλ/Γ\Gamma\prod_{\lambda\neq\lambda_{0}}K_{\lambda}/\Gamma of G/ΓG/\Gamma is closed and isomorphic to λλ0Kλ\prod_{\lambda\neq\lambda_{0}}K_{\lambda}. By the Pontryagin duality, there exists a unitary extension χ\chi of the character χS𝟏Λ1(S{λ0})Kλ\chi_{S}\otimes\mathbf{1}_{\prod_{\Lambda_{1}\setminus(S\cup\{\lambda_{0}\})}K_{\lambda}} to G/ΓG/\Gamma. ∎

Corollary 8.3.

Let TT be a torus over a number field kk, and let SS be a finite set of finite places of kk. Suppose that χS\chi_{S} be a character of the maximal compact subgroup T(kS)cT(k_{S})_{c} of T(kS)T(k_{S}). Then, there exists a unitary character χ\chi^{\prime} of T(𝔸k)/T(k)T(\mathbb{A}_{k})/T(k) which extends χS\chi_{S} on T(kS)cT(k_{S})_{c} and unramified outside SS.

We assume that SS is a singleton {u0}\{u_{0}\} and ATu0=AT×kku0A_{T_{u_{0}}}=A_{T}\times_{k}k_{u_{0}}. Let χu0\chi_{u_{0}} be a unitary character of T(ku0)T(k_{u_{0}}). Then, there exists a unitary character χ\chi of T(𝔸k)/T(k)T(\mathbb{A}_{k})/T(k) which extends χu0\chi_{u_{0}} on T(ku0)T(k_{u_{0}}) and unramified outside {u0}\{u_{0}\}.

Proof.

The first statement follows from the previous lemma by taking the set of places of kk as Λ\Lambda, the set of finite places of kk as Λ1\Lambda_{1}, the set of infinite places of kk as Λ2\Lambda_{2} and an infinite place kk as λ0\lambda_{0}. We now consider the second statement. By applying the first statement to obtain a global character χ\chi^{\prime} which extends χu0T(ku0)c{\chi_{u_{0}}}_{\restriction{T(k_{u_{0}})_{c}}}. Considering the twist by χ\chi^{\prime}, we may assume that χu0T(ku0)c=1{\chi_{u_{0}}}_{\restriction{T(k_{u_{0}})_{c}}}=1. Note that the group T(ku0)cT(k_{u_{0}})_{c} is the intersection of the kernels of unramified characters of T(ku0)T(k_{u_{0}}). Thus, the character χu0\chi_{u_{0}} is unramified. By the assumption ATu0=AT×kku0A_{T_{u_{0}}}=A_{T}\times_{k}k_{u_{0}}, we have the injection

T(ku0)/T(ku0)cT(𝔸k)/T(𝔸k)1,\displaystyle T(k_{u_{0}})/T(k_{u_{0}})_{c}\hookrightarrow T(\mathbb{A}_{k})/T(\mathbb{A}_{k})^{1},

and the unramified character χu0\chi_{u_{0}} can be extended to an automorphic character χ\chi of T(𝔸k)T(\mathbb{A}_{k}) by the Pontryagin duality. Hence the result. ∎

8.2. Shin’s automorphic density theorem

Let kk be a number field with a real place u0u_{0}. Thus, we have ku0k_{u_{0}}\cong\mathbb{R}. Let GG be a connected reductive group over kk. We assume that the center ZGZ_{G} of GG is an induced torus over kk. For a finite set SS of places of kk, we denote the group vSG(kv)\prod_{v\in S}G(k_{v}) by GSG_{S}. Similarly, we set GS=vSG(Fv)G^{S}=\prod_{v\not\in S}^{\prime}G(F_{v}).

Let λ\lambda be an automorphic unitary character of ZG(𝔸k)Z_{G}(\mathbb{A}_{k}). As we have ZG(𝔸k)=(ZG)S×ZGSZ_{G}(\mathbb{A}_{k})=(Z_{G})_{S}\times Z_{G}^{S}, we can decompose this central character as λ=λSλS\lambda=\lambda_{S}\boxtimes\lambda^{S}. By [MoeglinWaldspurger2016-Stablisation2]*VI 2.8, we have the invariant trace formula with the fixed central character λ\lambda.

Definition 8.4.

Let Irrunit(G(FS),λS)\operatorname{Irr}_{\mathrm{unit}}(G(F_{S}),\lambda_{S}) be a set of unitary representations of G(FS)G(F_{S}) with the central character λS\lambda_{S}. This set is also endowed with the restriction of the Fell topology on Irrunit(G(FS))\operatorname{Irr}_{\mathrm{unit}}(G(F_{S})). A Borel subset USU_{S} of Irrunit(GS,λS)\operatorname{Irr}_{\mathrm{unit}}(G_{S},\lambda_{S}) is said to be μGS\mu^{G_{S}}-regular if the measure μGS(US)\mu^{G_{S}}(\partial U_{S}) of the boundary is equal to zero.

Theorem 8.5 (c.f. [Shi12]*Theorem 4.11).

Let kk be a number field with a real place u0u_{0}. Let GG be a connected reductive group over kk. Let λ\lambda be an automorphic unitary character of ZG(𝔸k)Z_{G}(\mathbb{A}_{k}). Let SS be a finite set of places of kk containing all the infinite places and assume that GG and λ\lambda is unramified outside SS. We set S=S{u0}S^{\prime}=S\setminus\{u_{0}\}.

We assume the following.

  • The center ZGZ_{G} of GG is induced.

  • The group G(ku0)G(k_{u_{0}}) has a discrete series representation with character λu0\lambda_{u_{0}} and also we have AG×kku0=AGku0A_{G}\times_{k}k_{u_{0}}=A_{G_{k_{u_{0}}}}.

  • We are given a μGS\mu^{G_{S^{\prime}}}-regular relatively quasi-compact set USIrrunit(G(kS),λS)U_{S^{\prime}}\subset\operatorname{Irr}_{\mathrm{unit}}(G(k_{S^{\prime}}),\lambda_{S^{\prime}}) such that μGS(US)>0\mu^{G_{S^{\prime}}}(U_{S^{\prime}})>0.

Then, there exists a cuspidal automorphic representation Π\Pi of G(𝔸k)G(\mathbb{A}_{k}) with the central character λ\lambda such that

  • we have ΠSUS\Pi_{S^{\prime}}\in U_{S^{\prime}},

  • the representation Πv\Pi_{v} is unramified for vSv\not\in S, and

  • the representation Πu0\Pi_{u_{0}} is a discrete series representation of G(ku0)G(k_{u_{0}}) and the infinitesimal character of Πu0\Pi_{u_{0}} is sufficiently regular.

Proof.

The proof is essentially the same as [Shi12]*Theorem 4.11. We use the invariant trace formula for the fixed central character λ\lambda constructed in [MoeglinWaldspurger2016-Stablisation2]*VI 2.8, instead of the usual Arthur trace formula.

We first recall some definitions which are similar to those in [Shi12]. We set G=G/ZGG^{\natural}=G/Z_{G}. Let Πcusp(G(𝔸k),λ)\Pi_{\mathrm{cusp}}(G(\mathbb{A}_{k}),\lambda) denote the set of cuspidal automorphic representations of G(𝔸k)G(\mathbb{A}_{k}) with the central character λ\lambda and mcusp(π)m_{\mathrm{cusp}}(\pi) be the multiplicity of π\pi in the space of cuspidal automorphic forms on G(𝔸k)G(\mathbb{A}_{k}). Let {ξn}n1\{\xi_{n}\}_{n\geq 1} be a sequence of the finite dimensional algebraic representation of G(ku0×)G(k_{u_{0}}\times_{\mathbb{R}}\mathbb{C}) with the central character λu0\lambda_{u_{0}} on ZG(ku0)Z_{G}(k_{u_{0}}), such that limnξn=\lim_{n\to\infty}\xi_{n}=\infty in the sense of [Shi12]*Definition 3.5. We define the functional μϕS,ξn,λcusp\mu^{\mathrm{cusp}}_{\phi^{S^{\prime}},\xi_{n},\lambda} on Cc(G(FS),λS1)C_{c}^{\infty}(G(F_{S^{\prime}}),\lambda_{S^{\prime}}^{-1}) as

μϕS,ξn,λcusp(f)=(1)q(Gu0)dim(ξn)vol(G(k)\G(𝔸k))πΠcusp(G(𝔸k),λ)mcusp(π)trπ(fϕSϕξn).\mu^{\mathrm{cusp}}_{\phi^{S^{\prime}},\xi_{n},\lambda}(f)\\ =\frac{(-1)^{q(G_{u_{0}})}}{\dim(\xi_{n})\operatorname{vol}(G^{\natural}(k)\backslash G^{\natural}(\mathbb{A}_{k}))}\sum_{\pi\in\Pi_{\mathrm{cusp}}(G(\mathbb{A}_{k}),\lambda)}m_{\mathrm{cusp}}(\pi)\operatorname{tr}\pi(f\otimes\phi^{S}\otimes\phi_{\xi_{n}}).

Here, the function ϕξnCc(G(ku0),λu01)\phi_{\xi_{n}}\in C_{c}^{\infty}(G(k_{u_{0}}),\lambda_{u_{0}}^{-1}) is the Euler-Poincare function associated with ξn\xi_{n} and ϕS\phi^{S} is the unit of the ring of unramified functions in Cc(G(𝔸kS),(λS)1)C_{c}^{\infty}(G(\mathbb{A}^{S}_{k}),(\lambda^{S})^{-1}). Let 𝒯(G(FS),λS1)\mathcal{FT}(G(F_{S^{\prime}}),\lambda_{S^{\prime}}^{-1}) denote the space of Fourier transforms of functions in Cc(G(kS),λS1)C_{c}^{\infty}(G(k_{S^{\prime}}),\lambda_{S^{\prime}}^{-1}). The last expression is equal to

vol(KSZG(𝔸k)/ZG(𝔸k))dim(ξn)vol(G(k)\G(𝔸k))×πΠcusp(G(𝔸k),λ),πKS0,ξn-cohomologicalmcusp(π)trπS(f)\frac{\operatorname{vol}(K^{S}Z_{G}(\mathbb{A}_{k})/Z_{G}(\mathbb{A}_{k}))}{\dim(\xi_{n})\operatorname{vol}(G^{\natural}(k)\backslash G^{\natural}(\mathbb{A}_{k}))}\\ \times\sum_{\pi\in\Pi_{\mathrm{cusp}}(G(\mathbb{A}_{k}),\lambda),\pi^{K^{S}}\neq 0,\text{$\xi_{n}$-cohomological}}m_{\mathrm{cusp}}(\pi)\operatorname{tr}\pi_{S^{\prime}}(f)

if nn is large.

The functional μϕS,ξn,λcusp\mu^{\mathrm{cusp}}_{\phi^{S^{\prime}},\xi_{n},\lambda} factors through the map Cc(G(kS),λS1)𝒯(G(FS),λS1)C_{c}^{\infty}(G(k_{S^{\prime}}),\lambda_{S^{\prime}}^{-1})\to\mathcal{FT}(G(F_{S^{\prime}}),\lambda_{S^{\prime}}^{-1}) taking the Fourier transform. We denote the resulting functional on 𝒯(G(FS),λS1)\mathcal{FT}(G(F_{S^{\prime}}),\lambda_{S^{\prime}}^{-1}) by μ^ϕnS,ξ,λcusp\widehat{\mu}^{\mathrm{cusp}}_{\phi_{n}^{S^{\prime}},\xi,\lambda}. Applying [Shi12]*Lemma 4.9 and ϕξn(1)=dim(ξn)\phi_{\xi_{n}}(1)=\dim(\xi_{n}), we have

limnμ^ϕS,ξn,λcusp(f^)=f(1)=μS(f^),\displaystyle\lim_{n\to\infty}\widehat{\mu}^{\mathrm{cusp}}_{\phi^{S^{\prime}},\xi_{n},\lambda}(\widehat{f})=f(1)=\mu^{S^{\prime}}(\widehat{f}),

for any fCc(G(FS),λS1)f\in C_{c}^{\infty}(G(F_{S^{\prime}}),\lambda_{S^{\prime}}^{-1}). The assumption AG×kku0=AGu0A_{G}\times_{k}k_{u_{0}}=A_{G_{u_{0}}} is used here to apply the result [GKM1997]*Theorem 5.2. An analog of the Sauvageot density principle for the fixed central character λS1\lambda^{-1}_{S^{\prime}} proved in [Binder2019]*Proposition 3.3.2 by using Theorem 7.2. Hence, the computation above and an analog of Lemma 8.1 implies the result. ∎

Remark 8.6.

Instead of assuming that the center ZGZ_{G} is induced, we can prove a similar result by assuming that GvG_{v} has an anisotropic center over kvk_{v} for vSv\in S, by considering the Poisson summation formula and finiteness of the class number for ZGZ_{G}.

8.3. Globalization to spherical cusp forms

Let kk be a number field. We set G=PGLn(𝔸k)G=\mathrm{PGL}_{n}(\mathbb{A}_{k}). Let SS_{\infty} be the set of infinite places of kk. Let A0A_{0} be a maximal split torus of GG over kk. We denote the group W(G,A0)W(G,A_{0}) by WW_{\mathbb{R}}. Note that the space of spherical tempered representations Irrtemp,sph(GS)\operatorname{Irr}_{\mathrm{temp},\mathrm{sph}}(G_{S_{\infty}}) is an open and closed subset of Irrtemp(GS)\operatorname{Irr}_{\mathrm{temp}}(G_{S_{\infty}}). This space is identified with i𝔞0/Wi\mathfrak{a}^{*}_{0}/W_{\mathbb{R}}. Thus, for any subset Ω\Omega and positive number t>0t\in\mathbb{R}_{>0}, we have the dilatation tΩt\Omega of Ω\Omega.

Theorem 8.7.

For any finite set of finite places SS, μGS\mu^{G_{S}}-regular relatively quasi-compact set USU_{S} of Irrunit(GS)\operatorname{Irr}_{\mathrm{unit}}(G_{S}) with positive measure and any open subset UU_{\infty} of Irrtemp,sph(GS)\operatorname{Irr}_{\mathrm{temp},\mathrm{sph}}(G_{S_{\infty}}) with positive measure, there exists a cuspidal automorphic representation of PGLn(𝔸k)\mathrm{PGL}_{n}(\mathbb{A}_{k}) such that

  • we have ΠSUS\Pi_{S}\in U_{S},

  • πv\pi_{v} is unramified for vSv\not\in S, and

  • ΠS\Pi_{S_{\infty}} is spherical and the representation is in the set tUtU_{\infty} for some t1t\geq 1.

Proof.

This follows from [Eikemeier2022]*Theorem 1.2 and Theorem 7.2 and we will give the details below. The first result is an extension of the result [FinisMatz2021]*Theorem 1.1 to quasi-split groups, see the footnote of [FinisMatz2021]*p.1038.

Now, we give the details. Let T0T_{0} be maximal split torus of GG and 𝔱0\mathfrak{t}_{0} be its Lie algebra. Then, the infinitesimal characters of spherical representations of GSG_{S_{\infty}} are parametrized by the quotient (𝔱0)/W(\mathfrak{t}^{*}_{0}\otimes_{\mathbb{R}}\mathbb{C})/W_{\mathbb{R}}. Those for unitary spherical representations are in the subset i𝔱0/Wi\mathfrak{t}^{*}_{0}/W_{\mathbb{R}}. For any bounded set B(𝔱0)/WB\subset(\mathfrak{t}^{*}_{0}\otimes_{\mathbb{R}}\mathbb{C})/W_{\mathbb{R}} and for any smooth function τCc(G(𝔸k,fin))\tau\in C_{c}^{\infty}(G(\mathbb{A}_{k,\mathrm{fin}})), we set

m(B,τ)=λBπΠdisc(G(𝔸k)),πK0,Wλπ=Wλmdisc(π)trπfin(τ).\displaystyle m(B,\tau)=\sum_{\lambda\in B}\sum_{\pi\in\Pi_{\mathrm{disc}}(G(\mathbb{A}_{k})),\pi_{\infty}^{K_{\infty}}\neq 0,W_{\mathbb{R}}\lambda_{\pi_{\infty}}=W_{\mathbb{R}}\lambda}m_{\mathrm{disc}}(\pi)\operatorname{tr}\pi_{\mathrm{fin}}(\tau).

This counts the contribution of spherical discrete automorphic representations with the infinitesimal character in the WW_{\mathbb{R}}-orbit of BB. Let Ωi𝔱0\Omega\subset i\mathfrak{t}^{*}_{0} be a bounded open subset with the rectifiable boundary in the sense of [FinisMatz2021]*p.1094, line 15. By shrinking Ω\Omega, we can always achieve this condition. For t1t\geq 1, let ΛΩ(t)\Lambda_{\Omega}(t) be the volume of tΩt\Omega with respect to the Plancherel measure, see [Eikemeier2022]*p.14, line 25 for an explicit description. Then, by [Eikemeier2022]*Theorem 1.2, we have

limt1ΛΩ(t)m(tΩ,τ)=τ(1)=μGS(τS^).\displaystyle\lim_{t\to\infty}\frac{1}{\Lambda_{\Omega}(t)}m(t\Omega,\tau)=\tau(1)=\mu^{G_{S}}(\widehat{\tau_{S}}).

We take τ=τSvS𝟏Kv\tau=\tau_{S}\otimes\bigotimes_{v\not\in S}\mathbf{1}_{K_{v}}. Note that if we write m(B,τS)m(B,\tau_{S}) for m(B,τ)m(B,\tau), then, we have

m(B,τS)\displaystyle m(B,\tau_{S}) =λBπΠdisc(G(𝔸k)),πKSK0,Wλπ=Wλmdisc(π)trπS(τS)\displaystyle=\sum_{\lambda\in B}\sum_{\pi\in\Pi_{\mathrm{disc}}(G(\mathbb{A}_{k})),\pi^{K^{S}K_{\infty}}\neq 0,W_{\mathbb{R}}\lambda_{\pi_{\infty}}=W_{\mathbb{R}}\lambda}m_{\mathrm{disc}}(\pi)\operatorname{tr}\pi_{S}(\tau_{S})
=λBπΠdisc(G(𝔸k)),πKSK0,Wλπ=Wλmdisc(π)τS^(πS).\displaystyle=\sum_{\lambda\in B}\sum_{\pi\in\Pi_{\mathrm{disc}}(G(\mathbb{A}_{k})),\pi^{K^{S}K_{\infty}}\neq 0,W_{\mathbb{R}}\lambda_{\pi_{\infty}}=W_{\mathbb{R}}\lambda}m_{\mathrm{disc}}(\pi)\widehat{\tau_{S}}(\pi_{S}).

Thus, if for any F(Irrunit(GS),μGS)F\in\mathcal{RI}(\operatorname{Irr}_{\mathrm{unit}}(G_{S}),\mu^{G_{S}}) we set

μdisc,tGS(F)=1ΛΩ(t)λBπΠdisc(G(𝔸k)),πKSK0,Wλπ=Wλmdisc(π)F(πS),\displaystyle\mu^{G_{S}}_{\mathrm{disc},t}(F)=\frac{1}{\Lambda_{\Omega}(t)}\sum_{\lambda\in B}\sum_{\pi\in\Pi_{\mathrm{disc}}(G(\mathbb{A}_{k})),\pi^{K^{S}K_{\infty}}\neq 0,W_{\mathbb{R}}\lambda_{\pi_{\infty}}=W_{\mathbb{R}}\lambda}m_{\mathrm{disc}}(\pi)F(\pi_{S}),

then, we have

limtμdisc,tGS(τS^)=μGS(τS^).\displaystyle\lim_{t\to\infty}\mu^{G_{S}}_{\mathrm{disc},t}(\widehat{\tau_{S}})=\mu^{G_{S}}(\widehat{\tau_{S}}).

Then, Lemma 8.1 implies the result. Note that we can replace mdisc(π)m_{\mathrm{disc}}(\pi) with mcusp(π)m_{\mathrm{cusp}}(\pi), see [FLM15]*Remark 1.1. ∎

Remark 8.8.

The argument of this type can be applied to more general (not necessarily adjoint) quasi-split reductive groups, see [Eikemeier2022]*Theorem 1.2.

References