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

On the asymptotic support of Plancherel measures for homogeneous spaces

Benjamin Harris The MITRE Corporation. Approved for Public Release; Distribution Unlimited. Public Release Case Number 21-3406. Affiliation with the MITRE Corporation is for identification purposes only and is not intended to convey or imply MITRE’s concurrent with, or support for, the positions, opinions, or viewpoints expressed by the author. [email protected]  and  Yoshiki Oshima Department of Pure and Applied Mathematics, Graduate School of Information Science and Technology, Osaka University, 1-5 Yamadaoka, Suita, Osaka 565-0871, Japan [email protected]
Abstract.

Let GG be a real linear reductive group and let HH be a unimodular, locally algebraic subgroup. Let suppL2(G/H)\operatorname{supp}L^{2}(G/H) be the set of irreducible unitary representations of GG contributing to the decomposition of L2(G/H)L^{2}(G/H), namely the support of the Plancherel measure. In this paper, we will relate suppL2(G/H)\operatorname{supp}L^{2}(G/H) with the image of moment map from the cotangent bundle T(G/H)𝔤T^{*}(G/H)\to\mathfrak{g}^{*}.

For the homogeneous space X=G/HX=G/H, we attach a complex Levi subgroup LXL_{X} of the complexification of GG and we show that in some sense “most” of representations in suppL2(G/H)\operatorname{supp}L^{2}(G/H) are obtained as quantizations of coadjoint orbits 𝒪\mathcal{O} such that 𝒪G/L\mathcal{O}\simeq G/L and that the complexification of LL is conjugate to LXL_{X}. Moreover, the union of such coadjoint orbits 𝒪\mathcal{O} coincides asymptotically with the moment map image. As a corollary, we show that L2(G/H)L^{2}(G/H) has a discrete series if the moment map image contains a nonempty subset of elliptic elements.

Key words and phrases:
Plancherel Measure, Homogeneous Space, The Orbit Method, Coadjoint Orbit, Harmonic Analysis, Reductive Group, Discrete Series
2010 Mathematics Subject Classification:
22E46

1. Introduction

Let GG be a connected, complex reductive group, let σ\sigma be an antiholomorphic involution of GG, and let

(Gσ)eGGσ(G^{\sigma})_{e}\subset G_{\mathbb{R}}\subset G^{\sigma}

be a real form of GG. Let HGH\subset G be a (Zariski) closed, complex algebraic subgroup, and let X=G/HX=G/H be the corresponding algebraic homogeneous space for GG. Assume HH is σ\sigma stable with real points H:=HσGHH_{\mathbb{R}}:=H^{\sigma}\cap G_{\mathbb{R}}\subset H. Let 𝔤\mathfrak{g} (resp. 𝔤\mathfrak{g}_{\mathbb{R}}, 𝔥\mathfrak{h}, 𝔥\mathfrak{h}_{\mathbb{R}}) denote the Lie algebra of GG (resp. GG_{\mathbb{R}}, HH, HH_{\mathbb{R}}). Let H0GH_{0}\subset G_{\mathbb{R}} be a closed (not necessarily algebraic) subgroup for which the Lie algebra 𝔥0\mathfrak{h}_{0} of H0H_{0} is equal to the Lie algebra 𝔥\mathfrak{h}_{\mathbb{R}} of HH_{\mathbb{R}}. In this case, we say that the corresponding homogeneous space X0:=G/H0X_{0}:=G_{\mathbb{R}}/H_{0} is locally algebraic.

Next, we assume that X0X_{0} admits a nonzero, GG_{\mathbb{R}}-invariant density ν\nu. Recall GG_{\mathbb{R}} acts continuously on the Hilbert space

L2(X0):={f:X0measurable|X0|f(x)|2dν<}L^{2}(X_{0}):=\left\{f\colon X_{0}\rightarrow\mathbb{C}\ \text{measurable}\,\Bigl{|}\,\int_{X_{0}}|f(x)|^{2}d\nu<\infty\right\}

and it preserves the unitary structure on L2(X0)L^{2}(X_{0}). The theory of direct integrals yields a decomposition of L2(X0)L^{2}(X_{0}) into irreducible unitary representations of GG_{\mathbb{R}}. To be more precise, let G^\widehat{G}_{\mathbb{R}} be the unitary dual of GG_{\mathbb{R}}, that is, the set of all isomorphism classes of irreducible unitary representations, equipped with the Fell topology and the corresponding Borel structure. Then there exist a finite Borel measure mm on G^\widehat{G}_{\mathbb{R}} and a measurable function n():G^>0{}n(\cdot):\widehat{G}_{\mathbb{R}}\to\mathbb{Z}_{>0}\cup\{\infty\} such that

L2(X0)G^πn(π)𝑑m.L^{2}(X_{0})\simeq\int_{\widehat{G}_{\mathbb{R}}}^{\oplus}\pi^{\oplus n(\pi)}dm.

The measure mm is unique up to equivalence because GG_{\mathbb{R}} is of type I. See [Dix77, Paragraphe VIII], [Fol16, § 7.4], [Mac76] and [Wal92, Chapter 14] for this theory. The support of L2(X0)L^{2}(X_{0}), denoted suppL2(X0)\operatorname{supp}L^{2}(X_{0}), is defined to be the support of the measure mm. Therefore, suppL2(X0)G^\operatorname{supp}L^{2}(X_{0})\subset\widehat{G}_{\mathbb{R}} is the smallest closed subset satisfying m(G^suppL2(X0))=0m\bigl{(}\widehat{G}_{\mathbb{R}}\setminus\operatorname{supp}L^{2}(X_{0})\bigr{)}=0.

The explicit form of the above decomposition of L2(X0)L^{2}(X_{0}) is called the Plancherel formula. It has been studied for a long time in several settings after the pioneering work of Gelfand. Among them we note that:

  • Harish-Chandra obtained the Plancherel formula for Riemannian symmetric spaces X0=G/KX_{0}=G_{\mathbb{R}}/K_{\mathbb{R}} and the group case X0=(G×G)/Δ(G)X_{0}=(G^{\prime}_{\mathbb{R}}\times G^{\prime}_{\mathbb{R}})/\Delta(G^{\prime}_{\mathbb{R}}).

  • The Plancherel formula for symmetric spaces was established by works of T.Oshima, Delorme [Del98], and van den Ban-Schlichtkrull [BS05].

  • Delorme-Knop-Krötz-Schlichtkrull [DKKS21] is a recent study toward the Plancherel formula for real spherical spaces.

  • When H0H_{0} is an arithmetic subgroup, the study of irreducible decomposition of L2(X0)L^{2}(X_{0}) is a vast subject in connection with automorphic representations.

Our setting that H0H_{0} is unimodular and locally algebraic include these settings. The aim of this paper is to study the asymptotic behavior of suppL2(X0)\operatorname{supp}L^{2}(X_{0}). As far as the authors know, this is the first result about the spectrum of L2(X0)L^{2}(X_{0}) in this generality.

We would also like to note two general results on the space of functions on X0X_{0} when H0H_{0} has finitely many connected components. Kobayashi-Oshima [KO13] proved that the finiteness of multiplicities on the space of functions on X0X_{0} (or more generally, induced representations) is characterized by the real sphericity. Recently, Benoist-Kobayashi [BK15, BKa, BK21, BKb] obtained a simple criterion for L2(X0)L^{2}(X_{0}) to be a tempered representation. A relationship between Benoist-Kobayashi’s result and our theorem will be discussed at the end of introduction.

Our study is motivated by the orbit method [Kir04], [Ver83]. Let us briefly explain. For a Lie group GG, we write G^\widehat{G} for the unitary dual of GG, that is, the set of equivalence classes of the irreducible unitary representations of GG. When GG is a connected, simply connected nilpotent Lie group, Kirillov [Kir62] establishes a bijective correspondence between G^\widehat{G} and the coadjoint orbits of GG. Moreover, characters, inductions, and restrictions of representations can be simply described in terms of the corresponding coadjoint orbit geometry. For example, when HH is a connected closed subgroup of GG, the following equivalence holds for πG^\pi\in\widehat{G}:

(1.1) πsuppL2(G/H)𝒪Im(μ:T(G/H)𝔤),\pi\in\operatorname{supp}L^{2}(G/H)\Longleftrightarrow\mathcal{O}\subset\operatorname{Im}(\mu\colon T^{*}(G/H)\to\mathfrak{g}^{*}),

where 𝒪\mathcal{O} denotes the coadjoint orbit for GG corresponding to π\pi and μ\mu denotes the moment map. See [Kir04] for the details.

For a reductive Lie group GG_{\mathbb{R}}, most irreducible, unitary representations naturally arise from coadjoint orbits. However, some do not. For instance, complementary series of SL(2,)\operatorname{SL}(2,\mathbb{R}) are not naturally associated to coadjoint orbits. Nevertheless, the set of coadjoint orbits is a good approximation of G^\widehat{G}_{\mathbb{R}}. In particular, we can define an irreducible, unitary representation from a semisimple orbital parameter (see Definition 1.1). Our main result Theorem 1.4 shows that the equivalence (1.1) is “asymptotically true” in our setting.

To be more precise, we need some notation and terminology. For ξ𝔤\xi\in\mathfrak{g}^{*}, let G(ξ)G(\xi) denote the stabilizer subgroup of ξ\xi for the coadjoint action of GG and let 𝔤(ξ)\mathfrak{g}(\xi) denote its Lie algebra, namely,

G(ξ)={gGAd(g)(ξ)=ξ},𝔤(ξ)={Y𝔤ad(Y)(ξ)=0}.G(\xi)=\{g\in G\mid\operatorname{Ad}^{*}(g)(\xi)=\xi\},\quad\mathfrak{g}(\xi)=\{Y\in\mathfrak{g}\mid\operatorname{ad}^{*}(Y)(\xi)=0\}.

Similarly, for ξ𝔤\xi\in\mathfrak{g}_{\mathbb{R}}^{*} or ξ1𝔤\xi\in\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*}, define

G(ξ)={gGAd(g)(ξ)=ξ},𝔤(ξ)={Y𝔤ad(Y)(ξ)=0}.G_{\mathbb{R}}(\xi)=\{g\in G_{\mathbb{R}}\mid\operatorname{Ad}^{*}(g)(\xi)=\xi\},\quad\mathfrak{g}_{\mathbb{R}}(\xi)=\{Y\in\mathfrak{g}_{\mathbb{R}}\mid\operatorname{ad}^{*}(Y)(\xi)=0\}.

When ξ\xi is semisimple, i.e. the coadjoint orbit through ξ\xi is closed, 𝔤(ξ)\mathfrak{g}(\xi) (resp. 𝔤(ξ)\mathfrak{g}_{\mathbb{R}}(\xi)) is called a Levi subalgebra of 𝔤\mathfrak{g} (resp. 𝔤\mathfrak{g}_{\mathbb{R}}). In the following, we often abbreviate the coadjoint action Ad(g)(ξ)\operatorname{Ad}^{*}(g)(\xi) to gξg\cdot\xi.

Let 𝔩𝔤\mathfrak{l}\subset\mathfrak{g} be a Levi subalgebra. Write Z(𝔩)Z(\mathfrak{l}) for the center of 𝔩\mathfrak{l} and define

Z(𝔩)reg:={λZ(𝔩)𝔤(λ)=𝔩},Z(\mathfrak{l})^{*}_{\text{reg}}:=\left\{\lambda\in Z(\mathfrak{l})^{*}\mid\mathfrak{g}(\lambda)=\mathfrak{l}\right\},

namely, Z(𝔩)regZ(\mathfrak{l})^{*}_{\text{reg}} is the set of \mathbb{C}-linear functionals on the center of 𝔩\mathfrak{l} with (minimal possible) stabilizer 𝔩\mathfrak{l}. Fix a Cartan subalgebra 𝔧𝔩\mathfrak{j}\subset\mathfrak{l}. Let Δ(𝔤,𝔧)\Delta(\mathfrak{g},\mathfrak{j}) (resp. Δ(𝔩,𝔧)\Delta(\mathfrak{l},\mathfrak{j})) be the roots of 𝔤\mathfrak{g} with respect to 𝔧\mathfrak{j} (resp. 𝔩\mathfrak{l} with respect to 𝔧\mathfrak{j}), and let Δ+(𝔩,𝔧)Δ(𝔩,𝔧)\Delta^{+}(\mathfrak{l},\mathfrak{j})\subset\Delta(\mathfrak{l},\mathfrak{j}) be a choice of positive roots. We say λZ(𝔩)reg\lambda\in Z(\mathfrak{l})^{*}_{\text{reg}} is in the good range if

αΔ(𝔤,𝔧)&λ,α>0λ+ρ𝔩,α>0.\alpha\in\Delta(\mathfrak{g},\mathfrak{j})\ \&\ \langle\lambda,\alpha^{\vee}\rangle\in\mathbb{R}_{>0}\Longrightarrow\langle\lambda+\rho_{\mathfrak{l}},\alpha^{\vee}\rangle\in\mathbb{R}_{>0}.

This definition is independent of the choices of 𝔧𝔩\mathfrak{j}\subset\mathfrak{l} and Δ+(𝔩,𝔧)\Delta^{+}(\mathfrak{l},\mathfrak{j}). Denote by Z(𝔩)grZ(\mathfrak{l})^{*}_{\text{gr}} the collection of Z(𝔩)regZ(\mathfrak{l})^{*}_{\text{reg}} that lie in the good range. Suppose moreover that 𝔩\mathfrak{l} is σ\sigma-stable and let 𝔩:=𝔩σ\mathfrak{l}_{\mathbb{R}}:=\mathfrak{l}^{\sigma}. Let 1Z(𝔩)\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*} denote the set of purely imaginary valued linear functionals on the center of 𝔩\mathfrak{l}_{\mathbb{R}}. Then 1Z(𝔩)\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*} is naturally viewed as a real form of Z(𝔩)Z(\mathfrak{l})^{*}. Let

1Z(𝔩)reg:=Z(𝔩)reg1Z(𝔩),1Z(𝔩)gr:=Z(𝔩)gr1Z(𝔩).\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\text{reg}}:=Z(\mathfrak{l})^{*}_{\text{reg}}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*},\quad\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\text{gr}}:=Z(\mathfrak{l})^{*}_{\text{gr}}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}.

Then 1Z(𝔩)reg\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\text{reg}} is a complement of a finite union of coroot subspaces with codimension one or two in 1Z(𝔩)\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}.

If λ𝒪1𝔤\lambda\in\mathcal{O}\subset\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*} is a point within a coadjoint orbit, then we define the Duflo double cover of G(λ)G_{\mathbb{R}}(\lambda) by G~(λ)=G(λ)×Sp(Tλ𝒪)Mp(Tλ𝒪)\widetilde{G}_{\mathbb{R}}(\lambda)=G_{\mathbb{R}}(\lambda)\times_{\operatorname{Sp}(T_{\lambda}\mathcal{O})}\operatorname{Mp}(T_{\lambda}\mathcal{O}). See [HO20, §2.1] for a more detailed explanation about this double cover.

Definition 1.1.

A semisimple orbital parameter for GG_{\mathbb{R}} is a pair (𝒪,Γ)(\mathcal{O},\Gamma) where

  1. (a)

    𝒪1𝔤\mathcal{O}\subset\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*} is a semisimple (i.e. closed) coadjoint orbit

  2. (b)

    for every λ𝒪\lambda\in\mathcal{O}, Γλ\Gamma_{\lambda} is a genuine one-dimensional unitary representation of G~(λ)\widetilde{G_{\mathbb{R}}}(\lambda).

In addition, this pair must satisfy

  1. (i)

    gΓλ=Γgλg\cdot\Gamma_{\lambda}=\Gamma_{g\cdot\lambda} for every gGg\in G_{\mathbb{R}}, λ𝒪\lambda\in\mathcal{O}

  2. (ii)

    dΓλ=λ|𝔤(λ)d\Gamma_{\lambda}=\lambda|_{\mathfrak{g}_{\mathbb{R}}(\lambda)} for every λ𝒪\lambda\in\mathcal{O}.

Let (𝒪,Γ)(\mathcal{O},\Gamma) be a semisimple orbital parameter. Take λ𝒪\lambda\in\mathcal{O} and put 𝔩:=𝔤(λ)\mathfrak{l}:=\mathfrak{g}(\lambda). Then we can regard λ\lambda as an element of 1Z(𝔩)reg\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\text{reg}} by restriction. Assume λ1Z(𝔩)reg\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\text{reg}} is in the good range. This assumption only depends on 𝒪\mathcal{O} and not on the choice of λ𝒪\lambda\in\mathcal{O}; hence, in this case, we say 𝒪\mathcal{O} is in the good range. Then we can construct an irreducible unitary representation π(𝒪,Γ)\pi(\mathcal{O},\Gamma) by using cohomological induction. See Section 2 for the definition. If we take λ𝒪\lambda\in\mathcal{O} and put 𝔩:=𝔤(λ)\mathfrak{l}_{\mathbb{R}}:=\mathfrak{g}_{\mathbb{R}}(\lambda), then we also write π(𝔩,Γλ)\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda}) for π(𝒪,Γ)\pi(\mathcal{O},\Gamma).

Let μ:TX𝔤\mu\colon T^{*}X\rightarrow\mathfrak{g}^{*} denote the moment map defined by

(x,ξ)ξTxX(𝔤/𝔤x)𝔤.(x,\xi)\mapsto\xi\in T_{x}^{*}X\simeq(\mathfrak{g}/\mathfrak{g}_{x})^{*}\hookrightarrow\mathfrak{g}^{*}.

The following theorem is a consequence of [Kno94, 3.3 Corollary].

Theorem 1.2 (cf. [Kno94]).

Let XX be an algebraic homogeneous space for a connected, complex reductive group GG admitting a nonzero GG-invariant density. Then there exists a complex Levi subalgebra 𝔩X𝔤\mathfrak{l}_{X}\subset\mathfrak{g} and a complex subspace 𝔞XZ(𝔩X)\mathfrak{a}_{X}^{*}\subset Z(\mathfrak{l}_{X})^{*} satisfying Z𝔤(𝔞X)=𝔩XZ_{\mathfrak{g}}(\mathfrak{a}_{X}^{*})=\mathfrak{l}_{X}, both unique up to GG-conjugacy, such that μ(TX)¯=G𝔞X¯\overline{\mu(T^{*}X)}=\overline{G\cdot\mathfrak{a}_{X}^{*}}.

To state our main result we introduce some notation.

Let ZZ be a finite dimensional real vector space and let SZS\subset Z be a subset. We define the asymptotic cone of SS in ZZ to be

AC(S):={ξZ| SC is unbounded for any open conic neighborhood C of ξ }{0}.\operatorname{AC}(S):=\biggl{\{}\xi\in Z\;\biggl{|}\;\begin{aligned} &\text{ $S\cap C$ is unbounded for}\\ &\text{ any open conic neighborhood $C$ of $\xi$ }\end{aligned}\biggr{\}}\cup\{0\}.

If 𝔩1,𝔩2𝔤\mathfrak{l}_{1},\mathfrak{l}_{2}\subset\mathfrak{g} are subalgebras, we write 𝔩1𝔩2\mathfrak{l}_{1}\sim\mathfrak{l}_{2} if there exists gGg\in G such that Ad(g)𝔩1=𝔩2\operatorname{Ad}(g)\mathfrak{l}_{1}=\mathfrak{l}_{2}.

Theorem 1.3.

Let 𝔩\mathfrak{l}_{\mathbb{R}} be a Lie subalgebra of 𝔤\mathfrak{g}_{\mathbb{R}} such that 𝔩𝔩X\mathfrak{l}_{\mathbb{R}}\otimes\mathbb{C}\sim\mathfrak{l}_{X}. Then

(1.2) AC({λ1Z(𝔩)gr|π(𝔩,Γλ)suppL2(X0)and (Gλ)𝔞X})1Z(𝔩)reg=AC({λ1Z(𝔩)grπ(𝔩,Γλ)suppL2(X0)})1Z(𝔩)reg=1μ(TX0)¯1Z(𝔩)reg.\begin{split}&\phantom{=}\operatorname{AC}\Bigl{(}\Bigl{\{}\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\mathrm{gr}}\;\Bigl{|}\;\begin{aligned} &\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\in\operatorname{supp}L^{2}(X_{0})\\ &\text{and }(G\cdot\lambda)\cap\mathfrak{a}_{X}^{*}\neq\emptyset\end{aligned}\Bigr{\}}\Bigr{)}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})_{\rm reg}^{*}\\ &=\operatorname{AC}\bigl{(}\bigl{\{}\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\mathrm{gr}}\mid\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\in\operatorname{supp}L^{2}(X_{0})\bigr{\}}\bigr{)}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})_{\rm reg}^{*}\\ &=\overline{\sqrt{-1}\mu(T^{*}X_{0})}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})_{\mathrm{reg}}^{*}.\end{split}

Further, we have either

(1.3) dim(1μ(TX0)¯1Z(𝔩)reg)=dim𝔞X\dim\bigl{(}\overline{\sqrt{-1}\mu(T^{*}X_{0})}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})_{\mathrm{reg}}^{*}\bigr{)}=\dim_{\mathbb{C}}\mathfrak{a}_{X}^{*}

or

1μ(TX0)¯1Z(𝔩)reg=.\overline{\sqrt{-1}\mu(T^{*}X_{0})}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})_{\mathrm{reg}}^{*}=\emptyset.

The intersection is always nonempty for some 𝔩\mathfrak{l}_{\mathbb{R}}.

Note that 1μ(TX0)¯1Z(𝔩)reg\overline{\sqrt{-1}\mu(T^{*}X_{0})}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})_{\mathrm{reg}}^{*} is a semialgebraic set and its dimension is well-defined. Theorem 1.3 says there exist 𝔩\mathfrak{l}_{\mathbb{R}} and infinitely many representations of the form π(𝔩,Γλ)\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda}) in suppL2(X0)\operatorname{supp}L^{2}(X_{0}).

Following the spirit of orbit method, we can restate Theorem 1.3 as follows.

Theorem 1.4.

In the above setting, we have

AC(π(𝒪,Γ)suppL2(X0)(G𝒪)𝔞X𝒪)(GZ(𝔩X)reg)\displaystyle\phantom{=}\operatorname{AC}\Biggl{(}\bigcup_{\begin{subarray}{c}\pi(\mathcal{O},\Gamma)\in\operatorname{supp}L^{2}(X_{0})\\ (G\cdot\mathcal{O})\cap\mathfrak{a}_{X}^{*}\neq\emptyset\end{subarray}}\mathcal{O}\Biggr{)}\cap(G\cdot Z(\mathfrak{l}_{X})^{*}_{\mathrm{reg}})
=AC(π(𝒪,Γ)suppL2(X0)𝒪)(GZ(𝔩X)reg)\displaystyle=\operatorname{AC}\Biggl{(}\bigcup_{\begin{subarray}{c}\pi(\mathcal{O},\Gamma)\in\operatorname{supp}L^{2}(X_{0})\end{subarray}}\mathcal{O}\Biggr{)}\cap(G\cdot Z(\mathfrak{l}_{X})^{*}_{\mathrm{reg}})
=1μ(TX0)¯(GZ(𝔩X)reg).\displaystyle=\overline{\sqrt{-1}\mu(T^{*}X_{0})}\cap(G\cdot Z(\mathfrak{l}_{X})^{*}_{\mathrm{reg}}).

Here, we assume 𝒪GZ(𝔩X)reg\mathcal{O}\subset G\cdot Z(\mathfrak{l}_{X})^{*}_{\mathrm{reg}} and 𝒪\mathcal{O} is in the good range for the first two lines of above equations.

We remark that 1μ(TX0)¯(GZ(𝔩X)reg)\overline{\sqrt{-1}\mu(T^{*}X_{0})}\cap(G\cdot Z(\mathfrak{l}_{X})^{*}_{\mathrm{reg}}) is an open dense subset of 1μ(TX0)¯\overline{\sqrt{-1}\mu(T^{*}X_{0})} by Theorem 1.2.

The significance of Theorem 1.3 and Theorem 1.4 is that in some sense “most” of the representations in suppL2(X0)\operatorname{supp}L^{2}(X_{0}) are of the form π(𝔩,Γλ)\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda}) where 𝔩\mathfrak{l}_{\mathbb{R}} is a real form of 𝔩\mathfrak{l} and 𝔩𝔩X\mathfrak{l}\sim\mathfrak{l}_{X}. Next, we give a precise statement along these lines.

If 𝔩𝔤\mathfrak{l}\subset\mathfrak{g} is a Levi subalgebra, denote by G^𝔩\widehat{G}_{\mathbb{R}}^{\mathfrak{l}} the collection of irreducible, unitary representations of GG_{\mathbb{R}} of the form π(𝔩,Γλ)\pi(\mathfrak{l}_{\mathbb{R}}^{\prime},\Gamma_{\lambda}) such that the complexification of 𝔩\mathfrak{l}^{\prime}_{\mathbb{R}} is GG-conjugate to 𝔩\mathfrak{l} and λ\lambda is in the good range.

Let 𝔧\mathfrak{j} be a Cartan subalgebra of 𝔤\mathfrak{g} and let W=W(𝔤,𝔧)W=W(\mathfrak{g},\mathfrak{j}) be the Weyl group. An irreducible unitary representation π\pi of GG_{\mathbb{R}} has an infinitesimal character, which is regarded as a WW-orbit in 𝔧\mathfrak{j}^{*} via the Harish-Chandra isomorphism. We write χπ𝔧/W\chi_{\pi}\in\mathfrak{j}^{*}/W for this. By taking a conjugation, we may assume 𝔧𝔩X\mathfrak{j}\subset\mathfrak{l}_{X} and then we have inclusions 𝔞XZ(𝔩X)𝔧\mathfrak{a}_{X}^{*}\subset Z(\mathfrak{l}_{X})^{*}\subset\mathfrak{j}^{*}. Write ρ𝔩X𝔧\rho_{\mathfrak{l}_{X}}\in\mathfrak{j}^{*} for the half sum of positive roots in 𝔩X\mathfrak{l}_{X}.

The following theorem is essentially same as Theorem 4.3.

Theorem 1.5.
  1. (i)

    If πsuppL2(G/H0)\pi\in\operatorname{supp}L^{2}(G_{\mathbb{R}}/H_{0}), then χπ\chi_{\pi} has a representative ξ𝔞X+ρ𝔩X\xi\in\mathfrak{a}_{X}^{*}+\rho_{\mathfrak{l}_{X}}, namely, χπW(𝔞X+ρ𝔩X)\chi_{\pi}\subset W\cdot(\mathfrak{a}_{X}^{*}+\rho_{\mathfrak{l}_{X}}).

  2. (ii)

    There exists a constant d>0d>0 which only depends on GG such that the following holds: if πsuppL2(G/H0)G^𝔩X\pi\in\operatorname{supp}L^{2}(G_{\mathbb{R}}/H_{0})\setminus\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{X}}, then there exist a representative ξχπ(𝔧)\xi\in\chi_{\pi}(\subset\mathfrak{j}^{*}) and a root αΔ(𝔤,𝔧)Δ(𝔩X,𝔧)\alpha\in\Delta(\mathfrak{g},\mathfrak{j})\setminus\Delta(\mathfrak{l}_{X},\mathfrak{j}) such that ξ𝔞X+ρ𝔩X\xi\in\mathfrak{a}_{X}^{*}+\rho_{\mathfrak{l}_{X}} and |ξ,α|<d|\langle\xi,\alpha^{\vee}\rangle|<d.

The conclusion of (ii) means that the distance between ξ\xi and Z(𝔩X)Z(𝔩X)regZ(\mathfrak{l}_{X})^{*}\setminus Z(\mathfrak{l}_{X})^{*}_{\rm reg} is bounded by a constant.

As a corollary to Theorems 1.3 and 1.5, we obtain the following. The proof is given in Section 4.

Corollary 1.6.
  1. (i)

    The asymptotic cone

    AC(πsuppL2(G/H0)G^𝔩Xχπ)\operatorname{AC}\Bigl{(}\bigcup_{\pi\in\operatorname{supp}L^{2}(G_{\mathbb{R}}/H_{0})\cap\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{X}}}\chi_{\pi}\Bigr{)}

    in 𝔧\mathfrak{j}^{*} contains a real semialgebraic variety with real dimension dim𝔞X\dim_{\mathbb{C}}\mathfrak{a}_{X}^{*}.

  2. (ii)

    The asymptotic cone

    AC(πsuppL2(G/H0)G^𝔩Xχπ)\operatorname{AC}\Bigl{(}\bigcup_{\pi\in\operatorname{supp}L^{2}(G_{\mathbb{R}}/H_{0})\setminus\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{X}}}\chi_{\pi}\Bigr{)}

    in 𝔧\mathfrak{j}^{*} is contained in a real algebraic variety with real dimension less than dim𝔞X\dim_{\mathbb{C}}\mathfrak{a}_{X}^{*}.

Finally, we can show that, under certain additional assumptions, some of the subfamilies of representations of the form π(𝔩,Γλ)\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda}) occurring in suppL2(X0)\operatorname{supp}L^{2}(X_{0}) must be discrete. An element ξ𝔤\xi\in\mathfrak{g}_{\mathbb{R}}^{*} is said to be elliptic if there exists a Cartan involution θ\theta such that θ(ξ)=ξ\theta(\xi)=\xi. A coadjoint orbit 𝒪𝔤\mathcal{O}\subset\mathfrak{g}_{\mathbb{R}}^{*} is said to be elliptic if one of (or equivalently, every) element in 𝒪\mathcal{O} is elliptic. Let (𝔤)ell𝔤(\mathfrak{g}_{\mathbb{R}})^{*}_{\text{ell}}\subset\mathfrak{g}_{\mathbb{R}}^{*} denote the subset of all elliptic elements.

Theorem 1.7.

Assume 𝔤𝔥\mathfrak{g}\neq\mathfrak{h}. If μ(TX0)(𝔤)ell\mu(T^{*}X_{0})\cap(\mathfrak{g}_{\mathbb{R}}^{*})_{\mathrm{ell}} contains a nonempty open subset of μ(TX0)\mu(T^{*}X_{0}), then there exist infinitely many distinct irreducible, unitary representations (π,V)(\pi,V) such that

HomG(V,L2(X0)){0}.\operatorname{Hom}_{G_{\mathbb{R}}}(V,L^{2}(X_{0}))\neq\{0\}.

In particular, X0X_{0} has a discrete series.

We have μ(TX0)=G𝔥\mu(T^{*}X_{0})=G_{\mathbb{R}}\cdot\mathfrak{h}_{\mathbb{R}}^{\perp}, where 𝔥:=(𝔤/𝔥)𝔤\mathfrak{h}_{\mathbb{R}}^{\perp}:=(\mathfrak{g}_{\mathbb{R}}/\mathfrak{h}_{\mathbb{R}})^{*}\subset\mathfrak{g}_{\mathbb{R}}^{*}. Hence the condition of Theorem 1.7 is equivalent to that 𝔥(𝔤)ell\mathfrak{h}_{\mathbb{R}}^{\perp}\cap(\mathfrak{g}_{\mathbb{R}}^{*})_{\mathrm{ell}} contains a nonempty open subset of 𝔥\mathfrak{h}_{\mathbb{R}}^{\perp}.

Remark 1.8.

Here are some remarks about Theorem 1.7.

  1. (1)

    It follows from the proof of Theorem 1.7 that if the condition of Theorem 1.7 holds, then we can find a θ\theta-stable parabolic subalgebra 𝔮𝔤\mathfrak{q}\subset\mathfrak{g} such that A𝔮(λ)A_{\mathfrak{q}}(\lambda) occurs as a discrete spectrum in L2(X0)L^{2}(X_{0}) for infinitely many parameters λ\lambda in the good range. We will find such 𝔮\mathfrak{q} explicitly for an example in §9.2.

  2. (2)

    For symmetric spaces, the existence of discrete series is equivalent to the rank condition rankG/H=rankK/(HK)\operatorname{rank}G/H=\operatorname{rank}K/(H\cap K) by [FJ80], [MO84]. This rank condition is equivalent to the condition in Theorem 1.7 for symmetric spaces.

  3. (3)

    When X0X_{0} is a spherical space, Theorem 1.7 is proved in [DKKS21, §12].

For some XX, the Levi subalgebra 𝔩X\mathfrak{l}_{X} becomes a Cartan subalgebra. In that case, Theorem 1.4 was proved as [HW17, Theorem 1.1]. For a Cartan subalgebra 𝔧\mathfrak{j}, the set G^𝔧\widehat{G}_{\mathbb{R}}^{\mathfrak{j}} consists of all tempered representations with regular infinitesimal characters. If we take a closure of G^𝔧\widehat{G}_{\mathbb{R}}^{\mathfrak{j}} with respect to the Fell topology of G^\widehat{G}_{\mathbb{R}}, then we get the set of all tempered representations.

We remark that it may happen that G^𝔩G^𝔩\widehat{G}_{\mathbb{R}}^{\mathfrak{l}}\cap\widehat{G}_{\mathbb{R}}^{\mathfrak{l}^{\prime}}\neq\emptyset even if 𝔩\mathfrak{l} and 𝔩\mathfrak{l}^{\prime} are not conjugate. When GG_{\mathbb{R}} is compact for example, we have G^𝔩G^𝔩\widehat{G}_{\mathbb{R}}^{\mathfrak{l}}\subset\widehat{G}_{\mathbb{R}}^{\mathfrak{l}^{\prime}} if 𝔩𝔩\mathfrak{l}\supset\mathfrak{l}^{\prime} and G^𝔧=G^\widehat{G}_{\mathbb{R}}^{\mathfrak{j}}=\widehat{G}_{\mathbb{R}} for a Cartan subalgebra 𝔧\mathfrak{j}.

Our proof can be divided into two parts: the first part (§3, §4) is algebraic and the second part (§5–§8) is analytic.

In the first part, we prove Theorem 1.5. Thanks to the local structure theorem for complex algebraic homogeneous spaces, we show that a certain ideal J𝔞XJ_{\mathfrak{a}_{X}} of the enveloping algebra 𝒰(𝔤)\mathcal{U}(\mathfrak{g}) annihilates all functions on G/H0G_{\mathbb{R}}/H_{0}. Hence for πsuppL2(G/H0)\pi\in\operatorname{supp}L^{2}(G_{\mathbb{R}}/H_{0}), the annihilator of π\pi contains J𝔞XJ_{\mathfrak{a}_{X}}. This information together with the unitarity of π\pi is enough to get the conclusion of Theorem 1.5. In the course of proof, we utilize the Beilinson-Bernstein localization and realize representations as the global sections of twisted 𝒟\mathscr{D}-modules on partial flag varieties.

In the second part, the wave front set of representations plays a central role. Our argument is partly similar to [HHO16, Har18, HW17], but requires some new ingredients. It was proved in [HW17, Theorem 2.1] that the wave front set of L2(G/H0)L^{2}(G_{\mathbb{R}}/H_{0}) equals the image of moment map. By the first part of our proof, we can show that the contribution from suppL2(G/H0)G^𝔩X\operatorname{supp}L^{2}(G_{\mathbb{R}}/H_{0})\setminus\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{X}} to the wave front set is small. Then we have a relationship between suppL2(G/H0)G^𝔩X\operatorname{supp}L^{2}(G_{\mathbb{R}}/H_{0})\cap\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{X}} and the image of moment map. To obtain Theorem 1.3, we need a calculation of the wave front set of a direct integral of representations in G^𝔩X\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{X}} (Theorem 5.1). §5–§7 is devoted to the proof of Theorem 5.1. For this, we use a formula for the distribution character of πG^𝔩\pi\in\widehat{G}_{\mathbb{R}}^{\mathfrak{l}} in [HO20]. This formula is a consequence of Schmid-Vilonen’s formula [SV98] which gives characters of representations in terms of characteristic cycles of sheaves on the flag variety.

In the end of introduction we would like to pose some questions concerning theorems above, for which the authors do not know the answer. The first one is about the converse of Theorem 1.7.

Question 1.

Assume H0H_{0} has only finitely many connected components and X0X_{0} has a discrete series. Then does μ(TX0)(𝔤)ell\mu(T^{*}X_{0})\cap(\mathfrak{g}_{\mathbb{R}}^{*})_{\mathrm{ell}} contain a nonempty open subset of μ(TX0)\mu(T^{*}X_{0})?

When H0H_{0} is a cocompact discrete subgroup of GG_{\mathbb{R}} and if GG_{\mathbb{R}} does not have a discrete series, then the statement of Question 1 does not hold. Thus, we require the assumption that H0H_{0} has finitely many connected components.

When X0X_{0} is a symmetric space, Question 1 is known to be true as mentioned in Remark 1.8 (2).

The existence of discrete series for non-symmetric spaces was considered in [Kob94, Kob98c]. The results there are compatible with the statement of Question 1. For (generalized) Stiefel manifolds, discrete series were studied in [Kob92, Li93]. For spherical spaces, recent results are in [DKKS21, §13] and [KKOS20].

To state the second question, we will enlarge the set of representations G^𝔩X\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{X}}. If we drop the condition that 𝒪\mathcal{O} is in the good range, π(𝒪,Γ)\pi(\mathcal{O},\Gamma) is still unitary, but it may be reducible or zero (see Remark 2.1). We include all irreducible components of such π(𝒪,Γ)\pi(\mathcal{O},\Gamma) and also include limits for these representations with respect to the Fell topology. Write G^,e𝔩X\widehat{G}_{\mathbb{R},\mathrm{e}}^{\mathfrak{l}_{X}} for this enlarged set.

Question 2.

When H0H_{0} has only finitely many connected components, do we have suppL2(X0)G^,e𝔩X\operatorname{supp}L^{2}(X_{0})\subset\widehat{G}_{\mathbb{R},\mathrm{e}}^{\mathfrak{l}_{X}}?

Again, Question 2 does not hold when H0H_{0} is an infinite discrete group in general.

For symmetric spaces, Question 2 is true by the Plancherel formula. Question 2 is also true when H0H_{0} is algebraic and 𝔩X\mathfrak{l}_{X} is a Cartan subalgebra because in that case L2(X0)L^{2}(X_{0}) is tempered and G^,e𝔩X\widehat{G}_{\mathbb{R},\mathrm{e}}^{\mathfrak{l}_{X}} is the set of all irreducible tempered representations. This follows from Benoist-Kobayashi’s results [BKa, Corollary 5.6 (i)] and [BKb, Theorem 1.1].


Acknowledgments

The authors thank Professor Bernhard Krötz for discussions about the relationship between this paper and Krötz’s work on spherical spaces. They are grateful to Professor Toshiyuki Kobayashi for constant encouragement and kind explanations about his studies which inspires us. B. Harris was supported by an AMS-Simons Travel Grant during the early part of this work. Y. Oshima was partially supported by JSPS KAKENHI Grant Number JP20K14325.

2. Quantization of semisimple coadjoint orbits

In this section we recall from [Duf82], [Vog00] and [HO20, §2] the definition of representations which correspond to semisimple coadjoint orbits, or more precisely semisimple orbital parameters (𝒪,Γ)(\mathcal{O},\Gamma). We follow notation and terminology of [HO20, §2].

Let (𝒪,Γ)(\mathcal{O},\Gamma) be a semisimple orbital parameter in the sense of Definition 1.1. Fix λ𝒪\lambda\in\mathcal{O} and let L:=G(λ)L_{\mathbb{R}}:=G_{\mathbb{R}}(\lambda) and 𝔩:=𝔤(λ)\mathfrak{l}_{\mathbb{R}}:=\mathfrak{g}_{\mathbb{R}}(\lambda). The Duflo double cover of LL_{\mathbb{R}} is defined as L~:=L×Sp(Tλ𝒪)Mp(Tλ𝒪)\widetilde{L_{\mathbb{R}}}:=L_{\mathbb{R}}\times_{\operatorname{Sp}(T_{\lambda}\mathcal{O})}\operatorname{Mp}(T_{\lambda}\mathcal{O}). Then

Γλ:L~×\Gamma_{\lambda}\colon\widetilde{L_{\mathbb{R}}}\rightarrow\mathbb{C}^{\times}

is a unitary one-dimensional representation satisfying dΓλ=λd\Gamma_{\lambda}=\lambda. Let 𝔧\mathfrak{j}_{\mathbb{R}} be a Cartan subalgebra of 𝔩\mathfrak{l}_{\mathbb{R}}. We can regard λ1𝔧\lambda\in\sqrt{-1}\mathfrak{j}_{\mathbb{R}}^{*} by extending λ\lambda by zero on 𝔧[𝔩,𝔩]\mathfrak{j}_{\mathbb{R}}\cap[\mathfrak{l}_{\mathbb{R}},\mathfrak{l}_{\mathbb{R}}].

In order to define the representation π(𝒪,Γ)\pi(\mathcal{O},\Gamma) of GG_{\mathbb{R}} we need to choose a complex parabolic subalgebra 𝔮𝔤\mathfrak{q}\subset\mathfrak{g} with Levi factor 𝔩=𝔤(λ)\mathfrak{l}=\mathfrak{g}(\lambda), which we call a polarization for λ\lambda. We say a polarization 𝔮\mathfrak{q} with nilradical 𝔫\mathfrak{n} is admissible if

λ,α>0αΔ(𝔫,𝔧).\langle\lambda,\alpha^{\vee}\rangle\in\mathbb{R}_{>0}\ \Longrightarrow\ \alpha\in\Delta(\mathfrak{n},\mathfrak{j}).

Moreover, we say an admissible polarization 𝔮\mathfrak{q} is maximally real if dim(𝔮σ(𝔮))\dim(\mathfrak{q}\cap\sigma(\mathfrak{q})) is maximal among all admissible polarizations for λ\lambda.

Fix a maximally real, admissible polarization 𝔮𝔤\mathfrak{q}\subset\mathfrak{g} with nilradical 𝔫\mathfrak{n}. In addition, fix a maximal compact subgroup KGK_{\mathbb{R}}\subset G_{\mathbb{R}} with Cartan involution θ\theta such that KLLK_{\mathbb{R}}\cap L_{\mathbb{R}}\subset L_{\mathbb{R}} is maximal compact. We decompose λ=λc+λn\lambda=\lambda_{c}+\lambda_{n} where λc(1Z(𝔩))θ\lambda_{c}\in(\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*})^{\theta} and λn(1Z(𝔩))θ\lambda_{n}\in(\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*})^{-\theta}. Define Δ(𝔫𝔭,𝔧)\Delta(\mathfrak{n}_{\mathfrak{p}},\mathfrak{j}) to be the collection of roots αΔ(𝔫,𝔧)\alpha\in\Delta(\mathfrak{n},\mathfrak{j}) with λn,α0\langle\lambda_{n},\alpha^{\vee}\rangle\neq 0. As in [HO20, §2.2], one checks that

𝔭=𝔤(λn)+𝔫𝔭where𝔫𝔭=αΔ(𝔫𝔭,𝔧)𝔤α\mathfrak{p}=\mathfrak{g}(\lambda_{n})+\mathfrak{n}_{\mathfrak{p}}\ \text{where}\ \mathfrak{n}_{\mathfrak{p}}=\sum_{\alpha\in\Delta(\mathfrak{n}_{\mathfrak{p}},\mathfrak{j})}\mathfrak{g}_{\alpha}

is a σ\sigma-stable parabolic subalgebra of 𝔤\mathfrak{g} with real form 𝔭\mathfrak{p}_{\mathbb{R}}. Define P:=NG(𝔭)P_{\mathbb{R}}:=N_{G_{\mathbb{R}}}(\mathfrak{p}_{\mathbb{R}}) to be the corresponding parabolic subgroup, and let P=MA(NP)P_{\mathbb{R}}=M_{\mathbb{R}}A_{\mathbb{R}}(N_{P})_{\mathbb{R}} be the Langlands decomposition of PP_{\mathbb{R}} with G(λn)=MAG_{\mathbb{R}}(\lambda_{n})=M_{\mathbb{R}}A_{\mathbb{R}}.

Following [HO20, §2.2], we define an elliptic coadjoint orbit 𝒪M:=Mλc\mathcal{O}^{M_{\mathbb{R}}}:=M_{\mathbb{R}}\cdot\lambda_{c}. Further, we obtain a genuine, one-dimensional, unitary representation ΓλcM\Gamma^{M_{\mathbb{R}}}_{\lambda_{c}} of M~(λ)\widetilde{M_{\mathbb{R}}}(\lambda) from Γλ\Gamma_{\lambda} by the formula [HO20, (2.13)]. The coadjoint orbit 𝒪M\mathcal{O}^{M_{\mathbb{R}}} and the one-dimensional representation ΓλcM\Gamma^{M_{\mathbb{R}}}_{\lambda_{c}} give rise to an elliptic orbital parameter (𝒪M,ΓM)(\mathcal{O}^{M_{\mathbb{R}}},\Gamma^{M_{\mathbb{R}}}) for MM_{\mathbb{R}}.

In [HO20, §2.3 and §2.4], we give a unitary representation π(𝒪M,ΓM)\pi(\mathcal{O}^{M_{\mathbb{R}}},\Gamma^{M_{\mathbb{R}}}) of MM_{\mathbb{R}} associated to (𝒪M,ΓM)(\mathcal{O}^{M_{\mathbb{R}}},\Gamma^{M_{\mathbb{R}}}). Then a unitary representation π(𝒪,Γ)\pi(\mathcal{O},\Gamma) is defined by the normalized parabolic induction

π(𝒪,Γ):=IndPG(π(𝒪M,ΓM)).\pi(\mathcal{O},\Gamma):=\operatorname{Ind}^{G_{\mathbb{R}}}_{P_{\mathbb{R}}}(\pi(\mathcal{O}^{M_{\mathbb{R}}},\Gamma^{M_{\mathbb{R}}})).

We also denote the same representation by π(𝔩,Γλ)\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda}). This representation does not depend on the choices of λ\lambda, 𝔮\mathfrak{q} or KK_{\mathbb{R}}.

Remark 2.1.

The construction of π(𝒪,Γ)\pi(\mathcal{O},\Gamma) here can be extended to the case where 𝒪\mathcal{O} is not necessarily in the good range. The admissibility of the polarization implies the elliptic orbital parameter above is in the fair range in the sense of [KV95]. In general, we still obtain unitary representations but they can be reducible or zero. In this paper, we only consider π(𝒪,Γ)\pi(\mathcal{O},\Gamma) for parameters in the good range as it is enough for our purpose and it makes our treatment easier.

In the above construction, π(𝒪M,ΓM)\pi(\mathcal{O}^{M_{\mathbb{R}}},\Gamma^{M_{\mathbb{R}}}) can be defined as the cohomological induction for a θ\theta-stable parabolic subalgebra treated in [KV95, Chapter V]. On the (𝔤,K)(\mathfrak{g},K)-module level, the induction IndPG\operatorname{Ind}^{G_{\mathbb{R}}}_{P_{\mathbb{R}}} can be also defined in terms of cohomological induction for a σ\sigma-stable parabolic subalgebra as in [KV95, Proposition 11.47]. Following [KV95, (11.71)], we define functors (𝔮,LK𝔤,Ku)j()(\prescript{u}{}{\mathcal{R}}_{\mathfrak{q},L_{\mathbb{R}}\cap K_{\mathbb{R}}}^{\mathfrak{g},K_{\mathbb{R}}})^{j}(\cdot) and (𝔮,LK𝔤,Ku)j()(\prescript{u}{}{\mathcal{L}}_{\mathfrak{q},L_{\mathbb{R}}\cap K_{\mathbb{R}}}^{\mathfrak{g},K_{\mathbb{R}}})^{j}(\cdot) from the category of (𝔩,LK)(\mathfrak{l},L_{\mathbb{R}}\cap K_{\mathbb{R}})-modules to that of (𝔤,K)(\mathfrak{g},K_{\mathbb{R}})-modules as

(𝔮,LK𝔤,Ku)j(Z)=(Γ𝔤,LK𝔤,K)j(Hom𝔮(𝒰(𝔤),Z)LK),\displaystyle(\prescript{u}{}{\mathcal{R}}_{\mathfrak{q},L_{\mathbb{R}}\cap K_{\mathbb{R}}}^{\mathfrak{g},K_{\mathbb{R}}})^{j}(Z)=(\Gamma_{\mathfrak{g},L_{\mathbb{R}}\cap K_{\mathbb{R}}}^{\mathfrak{g},K_{\mathbb{R}}})^{j}\bigl{(}\operatorname{Hom}_{\mathfrak{q}}(\mathcal{U}(\mathfrak{g}),Z)_{L_{\mathbb{R}}\cap K_{\mathbb{R}}}\bigr{)},
(𝔮,LK𝔤,Ku)j(Z)=(Π𝔤,LK𝔤,K)j(𝒰(𝔤)𝒰(𝔮)Z)\displaystyle(\prescript{u}{}{\mathcal{L}}_{\mathfrak{q},L_{\mathbb{R}}\cap K_{\mathbb{R}}}^{\mathfrak{g},K_{\mathbb{R}}})_{j}(Z)=(\Pi_{\mathfrak{g},L_{\mathbb{R}}\cap K_{\mathbb{R}}}^{\mathfrak{g},K_{\mathbb{R}}})_{j}\bigl{(}\mathcal{U}(\mathfrak{g})\otimes_{\mathcal{U}(\mathfrak{q})}Z\bigr{)}

for jj\in\mathbb{N}. Here, an (𝔩,LK)(\mathfrak{l},L_{\mathbb{R}}\cap K_{\mathbb{R}})-module ZZ is regarded as a (𝔮,LK)(\mathfrak{q},L_{\mathbb{R}}\cap K_{\mathbb{R}})-module by the trivial 𝔫\mathfrak{n}-action, (Γ𝔤,LK𝔤,K)j(\Gamma_{\mathfrak{g},L_{\mathbb{R}}\cap K_{\mathbb{R}}}^{\mathfrak{g},K_{\mathbb{R}}})^{j} is the jj-th derived Zuckerman functor, and (Π𝔤,LK𝔤,K)j(\Pi_{\mathfrak{g},L_{\mathbb{R}}\cap K_{\mathbb{R}}}^{\mathfrak{g},K_{\mathbb{R}}})_{j} is its dual version. Then by induction in stages, we have an isomorphism on the (𝔤,K)(\mathfrak{g},K)-module level

π(𝔩,Γλ)=π(𝒪,Γ)(𝔮,LK𝔤,Ku)s(Γλeρ(𝔫)),\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})=\pi(\mathcal{O},\Gamma)\simeq(\prescript{u}{}{\mathcal{R}}_{\mathfrak{q},L_{\mathbb{R}}\cap K_{\mathbb{R}}}^{\mathfrak{g},K_{\mathbb{R}}})^{s}(\Gamma_{\lambda}\otimes e^{\rho(\mathfrak{n})}),

where s=dim(𝔫𝔨)s=\dim_{\mathbb{C}}(\mathfrak{n}\cap\mathfrak{k}) and eρ(𝔫)e^{\rho(\mathfrak{n})} denotes the genuine character of L~\widetilde{L}_{\mathbb{R}} associated with the Lagrangian subspace 𝔫Tλ𝒪\mathfrak{n}\subset T_{\lambda}{\mathcal{O}} (see [Duf82, Chapitre I] for the definition). In fact, by [KV95, Theorem 5.99 and Proposition 11.52],

(𝔮,LK𝔭,MKu)j(Γλeρ(𝔫))=0 for js,\displaystyle(\prescript{u}{}{\mathcal{R}}_{\mathfrak{q},L_{\mathbb{R}}\cap K_{\mathbb{R}}}^{\mathfrak{p},M_{\mathbb{R}}\cap K_{\mathbb{R}}})^{j}(\Gamma_{\lambda}\otimes e^{\rho(\mathfrak{n})})=0\ \text{ for $j\neq s$},
(𝔭,MK𝔤,Ku)j=0 for j0,\displaystyle(\prescript{u}{}{\mathcal{R}}_{\mathfrak{p},M_{\mathbb{R}}\cap K_{\mathbb{R}}}^{\mathfrak{g},K_{\mathbb{R}}})^{j}=0\ \text{ for $j\neq 0$},
(𝔮,LK𝔤,Ku)j(Γλeρ(𝔫)){(𝔭,MK𝔤,Ku)0(𝔮,LK𝔭,MKu)s(Γλeρ(𝔫)) for j=s,0 for js.\displaystyle(\prescript{u}{}{\mathcal{R}}_{\mathfrak{q},L_{\mathbb{R}}\cap K_{\mathbb{R}}}^{\mathfrak{g},K_{\mathbb{R}}})^{j}(\Gamma_{\lambda}\otimes e^{\rho(\mathfrak{n})})\simeq\begin{cases}(\prescript{u}{}{\mathcal{R}}_{\mathfrak{p},M_{\mathbb{R}}\cap K_{\mathbb{R}}}^{\mathfrak{g},K_{\mathbb{R}}})^{0}(\prescript{u}{}{\mathcal{R}}_{\mathfrak{q},L_{\mathbb{R}}\cap K_{\mathbb{R}}}^{\mathfrak{p},M_{\mathbb{R}}\cap K_{\mathbb{R}}})^{s}(\Gamma_{\lambda}\otimes e^{\rho(\mathfrak{n})})\ \text{ for $j=s$},\\ 0\ \text{ for $j\neq s$}.\end{cases}

Note that π(𝔩,Γλ)\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda}) has infinitesimal character λ+ρ𝔩\lambda+\rho_{\mathfrak{l}}, where we choose positive roots Δ+(𝔩,𝔧)Δ(𝔩,𝔧)\Delta^{+}(\mathfrak{l},\mathfrak{j})\subset\Delta(\mathfrak{l},\mathfrak{j}) and write ρ𝔩=12Δ+(𝔩,𝔧)α\rho_{\mathfrak{l}}=\frac{1}{2}\sum_{\Delta^{+}(\mathfrak{l},\mathfrak{j})}\alpha. By [KV95, Theorem 5.99 and Proposition 11.65], π(𝔩,Γλ)\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda}) can be also constructed by the functor u\prescript{u}{}{\mathcal{L}}:

(𝔮,LK𝔤,Ku)j(Γλeρ(𝔫))(σ(𝔮),LK𝔤,Ku)j(Γλeρ(θ(𝔫))).\displaystyle(\prescript{u}{}{\mathcal{R}}_{\mathfrak{q},L_{\mathbb{R}}\cap K_{\mathbb{R}}}^{\mathfrak{g},K_{\mathbb{R}}})^{j}(\Gamma_{\lambda}\otimes e^{\rho(\mathfrak{n})})\simeq(\prescript{u}{}{\mathcal{L}}_{\sigma(\mathfrak{q}),L_{\mathbb{R}}\cap K_{\mathbb{R}}}^{\mathfrak{g},K_{\mathbb{R}}})_{j}(\Gamma_{\lambda}\otimes e^{\rho(\theta(\mathfrak{n}))}).

Here, eρ(θ(𝔫))e^{\rho(\theta(\mathfrak{n}))} is the character defined in [Duf82, Chapitre I] associated with the Lagrangian subspace θ(𝔫)Tλ𝒪\theta(\mathfrak{n})\subset T_{\lambda}{\mathcal{O}}.

Following [HO20, Appendix A] (cf. also [Mat04, Theorem 2.2.3]), we define a virtual (𝔤,K)(\mathfrak{g},K_{\mathbb{R}})-module

π(𝒪,Γ,𝔮):=j(1)j(𝔮,LK𝔤,Ku)s+j(Γλeρ(𝔫))\pi(\mathcal{O},\Gamma,\mathfrak{q}):=\sum_{j}(-1)^{j}(\prescript{u}{}{\mathcal{R}}_{\mathfrak{q},L_{\mathbb{R}}\cap K_{\mathbb{R}}}^{\mathfrak{g},K_{\mathbb{R}}})^{s+j}(\Gamma_{\lambda}\otimes e^{\rho(\mathfrak{n})})

for any polarization 𝔮\mathfrak{q}. Note that the functor u\prescript{u}{}{\mathcal{R}} here is denoted by II in [HO20]. Then [HO20, Theorem A.1] says π(𝒪,Γ,𝔮)\pi(\mathcal{O},\Gamma,\mathfrak{q}) does not depend on the choice of polarization 𝔮\mathfrak{q} as long as 𝔮\mathfrak{q} is admissible. In the same way, we can prove that a virtual module

π(𝒪,Γ,𝔮):=j(1)j(σ(𝔮),LK𝔤,Ku)sj(Γλeρ(θ(𝔫)))\pi^{\prime}(\mathcal{O},\Gamma,\mathfrak{q}):=\sum_{j}(-1)^{j}(\prescript{u}{}{\mathcal{L}}_{\sigma(\mathfrak{q}),L_{\mathbb{R}}\cap K_{\mathbb{R}}}^{\mathfrak{g},K_{\mathbb{R}}})_{s-j}(\Gamma_{\lambda}\otimes e^{\rho(\theta(\mathfrak{n}))})

does not depend on the choice of admissible polarization 𝔮\mathfrak{q}. Since π(𝒪,Γ,𝔮)=π(𝒪,Γ,𝔮)\pi(\mathcal{O},\Gamma,\mathfrak{q})=\pi^{\prime}(\mathcal{O},\Gamma,\mathfrak{q}) for a maximally real admissible polarization 𝔮\mathfrak{q}, the same is true for any admissible polarization, namely we have

π(𝒪,Γ,𝔮)=π(𝒪,Γ,𝔮)=π(𝒪,Γ)\pi(\mathcal{O},\Gamma,\mathfrak{q})=\pi^{\prime}(\mathcal{O},\Gamma,\mathfrak{q})=\pi(\mathcal{O},\Gamma)

as a virtual (𝔤,K)(\mathfrak{g},K_{\mathbb{R}})-module for any admissible polarization 𝔮\mathfrak{q}.

By the Beilinson-Bernstein localization, this representation can be also realized as global sections on the flag variety. For an admissible polarization 𝔮\mathfrak{q}, let QQ be the parabolic subgroup of GG with Lie algebra 𝔮\mathfrak{q}, let Y:=G/σ(Q)Y:=G/\sigma(Q) be the partial flag variety, the collection of all parabolic subgroups which are conjugate to σ(Q)\sigma(Q) and let S=K/(σ(Q)K)S=K/(\sigma(Q)\cap K) be the KK-orbit through the base point in YY. Let 𝒟Y,λ\mathscr{D}_{Y,\lambda} be the sheaf of rings of twisted differential operators on YY corresponding to the parameter λ\lambda (see e.g. [Bie90]). Then we have a spectral sequence of (𝔤,K)(\mathfrak{g},K_{\mathbb{R}})-modules (see e.g. [Kit12, Theorem 5.4], [Osh13, (6.3)])

Hp(Y,Rqi+)(σ(𝔮),LK𝔤,Ku)spq(Γλeρ(θ(𝔫))).H^{p}(Y,R^{q}i_{+}\mathcal{L})\Rightarrow(\prescript{u}{}{\mathcal{L}}_{\sigma(\mathfrak{q}),L_{\mathbb{R}}\cap K_{\mathbb{R}}}^{\mathfrak{g},K_{\mathbb{R}}})_{s-p-q}(\Gamma_{\lambda}\otimes e^{\rho(\theta(\mathfrak{n}))}).

Here, i:SYi\colon S\to Y is the natural immersion. \mathcal{L} is the KK-equivariant line bundle (i.e. invertible 𝒪\mathcal{O}-module) on SS given by K×(σ(Q)K)τK\times_{(\sigma(Q)\cap K)}\tau for an algebraic character τ\tau of σ(Q)K\sigma(Q)\cap K whose restriction to LKL_{\mathbb{R}}\cap K_{\mathbb{R}} is

Γλeρ(θ(𝔫))top(𝔨/(𝔩𝔨)).\Gamma_{\lambda}\otimes e^{-\rho(\theta(\mathfrak{n}))}\otimes\bigwedge^{\rm top}(\mathfrak{k}/(\mathfrak{l}\cap\mathfrak{k})).

Then \mathcal{L} can be viewed as a twisted 𝒟\mathscr{D}-module on SS and its (higher) direct images Rqi+R^{q}i_{+}\mathcal{L} are defined as 𝒟Y,λ\mathscr{D}_{Y,\lambda}-modules. Our assumption on λ\lambda implies YY is 𝒟Y,λ\mathscr{D}_{Y,\lambda}-affine so that Hp(Y,Rqi+)=0H^{p}(Y,R^{q}i_{+}\mathcal{L})=0 for p>0p>0. Hence the above spectral sequence collapses and we have

Γ(Y,Rqi+)(σ(𝔮),LK𝔤,Ku)sq(Γλeρ(θ(𝔫))).\Gamma(Y,R^{q}i_{+}\mathcal{L})\simeq(\prescript{u}{}{\mathcal{L}}_{\sigma(\mathfrak{q}),L_{\mathbb{R}}\cap K_{\mathbb{R}}}^{\mathfrak{g},K_{\mathbb{R}}})_{s-q}(\Gamma_{\lambda}\otimes e^{\rho(\theta(\mathfrak{n}))}).

We therefore have

(2.1) q(1)qΓ(Y,Rqi+)=π(𝒪,Γ).\sum_{q}(-1)^{q}\Gamma(Y,R^{q}i_{+}\mathcal{L})=\pi(\mathcal{O},\Gamma).

We end this section by giving the Langlands parameter of π(𝒪,Γ)\pi(\mathcal{O},\Gamma) when 𝒪\mathcal{O} is in the good range. In order to do this, we need to write a one-dimensional representation of LL_{\mathbb{R}} as a quotient of standard module. Let JJ_{\mathbb{R}} be the maximally noncompact Cartan subalgebra of LL_{\mathbb{R}} and let J=TA1J_{\mathbb{R}}=T_{\mathbb{R}}A^{1}_{\mathbb{R}} be its Cartan decomposition with respect to θ\theta, namely, T=JθT_{\mathbb{R}}=J_{\mathbb{R}}^{\theta} and A1A^{1}_{\mathbb{R}} is the connected subgroup of LL_{\mathbb{R}} with Lie algebra 𝔞1=𝔧θ\mathfrak{a}^{1}_{\mathbb{R}}=\mathfrak{j}_{\mathbb{R}}^{-\theta}. Take a Borel subalgebra 𝔟𝔩\mathfrak{b}_{\mathfrak{l}} of 𝔩\mathfrak{l} such that 𝔟𝔩𝔧\mathfrak{b}_{\mathfrak{l}}\supset\mathfrak{j} and 𝔟𝔩+(𝔩𝔨)=𝔩\mathfrak{b}_{\mathfrak{l}}+(\mathfrak{l}\cap\mathfrak{k})=\mathfrak{l}. Write 𝔫𝔩\mathfrak{n}_{\mathfrak{l}} for the nilradical of 𝔟𝔩\mathfrak{b}_{\mathfrak{l}}. Define a character e2ρ(𝔫𝔩)e^{2\rho(\mathfrak{n}_{\mathfrak{l}})^{\prime}} of JJ_{\mathbb{R}} by

e2ρ(𝔫𝔩)(ta)=det(Ad(t)|𝔫𝔩𝔨)det(Ad(a)|𝔫𝔩)e^{2\rho(\mathfrak{n}_{\mathfrak{l}})^{\prime}}(ta)=\det(\operatorname{Ad}(t)|_{\mathfrak{n}_{\mathfrak{l}}\cap\mathfrak{k}})\cdot\det(\operatorname{Ad}(a)|_{\mathfrak{n}_{\mathfrak{l}}})

for tTt\in T_{\mathbb{R}} and aA1a\in A^{1}_{\mathbb{R}}, which is the same as the character 2ρ(𝔫𝔩)\mathbb{C}_{2\rho(\mathfrak{n}_{\mathfrak{l}})^{\prime}} defined in [KV95, (11.111)]. The differential of e2ρ(𝔫𝔩)e^{2\rho(\mathfrak{n}_{\mathfrak{l}})^{\prime}} equals 2ρ(𝔫𝔩)2\rho(\mathfrak{n}_{\mathfrak{l}}), but it may not be equal to det(Ad(ta)|𝔫𝔩)\det(\operatorname{Ad}(ta)|_{\mathfrak{n}_{\mathfrak{l}}}) when TT_{\mathbb{R}} is disconnected. The trivial representation of LL_{\mathbb{R}} is the irreducible quotient of the standard module (I𝔟𝔩,T𝔩,LK)sL(e2ρ(𝔫𝔩))(I^{\mathfrak{l},L_{\mathbb{R}}\cap K_{\mathbb{R}}}_{\mathfrak{b}_{\mathfrak{l}},T_{\mathbb{R}}})^{s_{L}}(e^{2\rho(\mathfrak{n}_{\mathfrak{l}})^{\prime}}), where sL:=dim(𝔫𝔩𝔨)s_{L}:=\dim_{\mathbb{C}}(\mathfrak{n}_{\mathfrak{l}}\cap\mathfrak{k}). By induction in stages, it turns out that π(𝒪,Γ)\pi(\mathcal{O},\Gamma) is the unique irreducible quotient of the standard module

(𝔟𝔩+𝔫,T𝔤,Ku)s+sL(Γλeρ(𝔫)e2ρ(𝔫𝔩)).(\prescript{u}{}{\mathcal{R}}^{\mathfrak{g},K_{\mathbb{R}}}_{\mathfrak{b}_{\mathfrak{l}}+\mathfrak{n},T_{\mathbb{R}}})^{s+s_{L}}(\Gamma_{\lambda}\otimes e^{\rho(\mathfrak{n})}\otimes e^{2\rho(\mathfrak{n}_{\mathfrak{l}})^{\prime}}).

In the notation of [AvLTV20] (cf. also [KV95, §XI.9]), the irreducible admissible representations of GG_{\mathbb{R}} are parametrized by data (J,γ,Δi+)(J_{\mathbb{R}},\gamma,\Delta_{i\mathbb{R}}^{+}), where JGJ_{\mathbb{R}}\subset G_{\mathbb{R}} is a Cartan subgroup with Lie algebra 𝔧\mathfrak{j}_{\mathbb{R}}, γ\gamma is a level one character of the ρabs\rho_{\text{abs}} double cover of JJ_{\mathbb{R}} (see Section 5 of [AvLTV20] for an explanation), and Δi+\Delta_{i\mathbb{R}}^{+} is a choice of positive roots among the set of imaginary roots for 𝔧\mathfrak{j}_{\mathbb{R}} in 𝔤\mathfrak{g}_{\mathbb{R}} for which dγ𝔧d\gamma\in\mathfrak{j}^{*} is weakly dominant. This triple must satisfy a couple of other technical assumptions (see Theorem 6.1 of [AvLTV20]). The above argument shows that the irreducible representation π(𝒪,Γ)\pi(\mathcal{O},\Gamma) corresponds to the parameter (J,γ,Δi+)(J_{\mathbb{R}},\gamma,\Delta_{i\mathbb{R}}^{+}), where γ\gamma is the character of ρabs\rho_{\rm abs}-cover of JJ_{\mathbb{R}} such that

γρabsΓλeρ(𝔫)e2ρ(𝔫𝔩).\gamma\otimes\rho_{\rm abs}\simeq\Gamma_{\lambda}\otimes e^{\rho(\mathfrak{n})}\otimes e^{2\rho(\mathfrak{n}_{\mathfrak{l}})^{\prime}}.

Δi+\Delta_{i\mathbb{R}}^{+} and ρabs\rho_{\rm abs} are defined by the positive system for the Borel subalgebra 𝔟𝔩+𝔫\mathfrak{b}_{\mathfrak{l}}+\mathfrak{n}.

3. Annihilator ideas of induced representations

In this section we will study annihilator ideals of irreducible subrepresentations of C(G/H0)C^{\infty}(G_{\mathbb{R}}/H_{0}).

First, we need the following fact on algebraic subgroups. See [BBHM63, Theorems 4 and 8].

Fact 3.1.

Let GG be a complex algebraic group and HH an algebraic subgroup. The following three conditions are equivalent.

  1. (1)

    G/HG/H is quasi-affine.

  2. (2)

    Every finite-dimensional rational HH-module is a HH-submodule of a finite-dimensional rational GG-module.

  3. (3)

    There exists a vector ww in a rational GG-module such that HH is the stabilizer subgroup of ww.

When one (or all) of the conditions in Fact 3.1 is satisfied, HH is said to be observable in GG.

Let GG be a connected, complex reductive group with real form (Gσ)eGGσ(G^{\sigma})_{e}\subset G_{\mathbb{R}}\subset G^{\sigma} for an antiholomorphic involution σ\sigma of GG. Suppose that a connected, complex algebraic subgroup HH of GG is defined over \mathbb{R}, namely σ(H)=H\sigma(H)=H. Write 𝔥=𝔥σ\mathfrak{h}_{\mathbb{R}}=\mathfrak{h}^{\sigma} for the real form of 𝔥\mathfrak{h}. Let H0GH_{0}\subset G_{\mathbb{R}} be a closed subgroup whose Lie algebra 𝔥0\mathfrak{h}_{0} is equal to the Lie algebra 𝔥\mathfrak{h}_{\mathbb{R}}. Here, the closedness of H0H_{0} in GG_{\mathbb{R}} is considered in the classical topology and H0H_{0} is not necessarily algebraic. In particular, we allow H0H_{0} to have infinitely many connected components.

Lemma 3.2.

If H0GH_{0}\subset G_{\mathbb{R}} is a unimodular subgroup, then HH is an observable subgroup of GG.

Proof.

Let d:=dim𝔥d:=\dim\mathfrak{h}. If H0GH_{0}\subset G_{\mathbb{R}} is a unimodular subgroup of GG_{\mathbb{R}}, then the identity component (H0)e(H_{0})_{e} of H0H_{0} acts trivially on d𝔥0\bigwedge^{d}\mathfrak{h}_{0}. Since 𝔥0=𝔥\mathfrak{h}_{0}=\mathfrak{h}_{\mathbb{R}}, the complexification 𝔥\mathfrak{h} annihilates d𝔥\bigwedge^{d}\mathfrak{h}. This implies that HGH\subset G is a unimodular subgroup.

Let W:=d𝔤W:=\bigwedge^{d}\mathfrak{g} with the GG-action dAd\bigwedge^{d}\operatorname{Ad}. Take a nonzero vector ww in d𝔥d𝔤\bigwedge^{d}\mathfrak{h}\subset\bigwedge^{d}\mathfrak{g}. Define SS to be the stabilizer subgroup of ww in GG. By definition of SS and Fact 3.1 (3), SS is observable in GG. Since HH is unimodular, HSH\subset S. Moreover, SS normalizes HH and hence HH is observable in SS by [BBHM63, Theorem 2]. The transitivity of the condition (2) in Fact 3.1 implies that HH is observable in GG. ∎

In the following we assume that H0H_{0} is unimodular.

We now use the local structure theorem for X=G/HX=G/H (see [Kno94, Theorem 2.3, Proposition 2.4, Lemma 3.1]). The theorem states that there exist a parabolic subgroup QXQ_{X} of GG with Levi factor LXL_{X} and an LXL_{X}-stable subvariety ZXZ\subset X such that

  • the natural map QX×LXZXQ_{X}\times^{L_{X}}Z\to X is an open immersion, and

  • if LX0L_{X}^{0} denotes the kernel of LXAut(Z)L_{X}\to\operatorname{Aut}(Z), then LX0L_{X}^{0} contains a commutator subgroup [LX,LX][L_{X},L_{X}].

Let AX=LX/LX0A_{X}=L_{X}/L_{X}^{0} with Lie algebra 𝔞X\mathfrak{a}_{X}, which is a torus. It follows from the proof of [Kno94, Theorem 2.3, Proposition 2.4, Lemma 3.1] that 𝔞X\mathfrak{a}_{X}^{*} intersects Z(𝔩X)regZ(\mathfrak{l}_{X})^{*}_{\mathrm{reg}}. Hence 𝔩X={Y𝔤ad(Y)(𝔞X)=0}\mathfrak{l}_{X}=\{Y\in\mathfrak{g}\mid\operatorname{ad}^{*}(Y)(\mathfrak{a}_{X}^{*})=0\}.

Next, fix a Cartan subgroup JLXJ\subset L_{X} and a Borel subgroup BB of GG such that JBQXJ\subset B\subset Q_{X}. Note that there are natural inclusions 𝔞X(𝔩X/[𝔩X,𝔩X])=Z(𝔩X)𝔧\mathfrak{a}_{X}^{*}\subset(\mathfrak{l}_{X}/[\mathfrak{l}_{X},\mathfrak{l}_{X}])^{*}=Z(\mathfrak{l}_{X})^{*}\subset\mathfrak{j}^{*}. Fix a positive system Δ+(𝔤,𝔧)\Delta^{+}(\mathfrak{g},\mathfrak{j}) as the roots for BB, and let FλF_{\lambda} denote the irreducible, finite-dimensional representation of GG with highest weight λ𝔧\lambda\in\mathfrak{j}^{*}. Let R(G/H)R(G/H) denote the space of regular functions on G/HG/H.

Lemma 3.3.

If λ𝔧\lambda\in\mathfrak{j}^{*} is a dominant integral weight and FλF_{\lambda} occurs in the irreducible decomposition of R(G/H)R(G/H), then λ𝔞X\lambda\in\mathfrak{a}_{X}^{*}.

Proof.

Suppose FλR(G/H)F_{\lambda}\subset R(G/H). If fFλR(G/H)f\in F_{\lambda}\subset R(G/H) is a highest weight vector, then f(b1x)=bλf(x)f(b^{-1}x)=b^{\lambda}f(x) for bBb\in B, xXx\in X. Observe that QX×LXZB×BLXZQ_{X}\times^{L_{X}}Z\simeq B\times^{B\cap L_{X}}Z, which can be identified with an open subvariety of XX. Therefore, f|Z0f|_{Z}\not\equiv 0. Since JLX0J\cap L_{X}^{0} acts trivially on ZZ, λ=0\lambda=0 on 𝔧𝔩X0\mathfrak{j}\cap\mathfrak{l}_{X}^{0}, namely, λ𝔞X\lambda\in\mathfrak{a}_{X}^{*}. ∎

Differentiating the action of GG_{\mathbb{R}} on G/H0G_{\mathbb{R}}/H_{0} and the action of GG on G/HG/H we obtain maps

𝒰(𝔤)Φ0Diff(G/H0),𝒰(𝔤)ΦDiff(G/H)\mathcal{U}(\mathfrak{g})\stackrel{{\scriptstyle\Phi_{0}}}{{\longrightarrow}}\operatorname{Diff}(G_{\mathbb{R}}/H_{0}),\quad\mathcal{U}(\mathfrak{g})\stackrel{{\scriptstyle\Phi}}{{\longrightarrow}}\operatorname{Diff}(G/H)

of the universal enveloping algebra into the algebras of differential operators. Here, Diff(G/H0)\operatorname{Diff}(G_{\mathbb{R}}/H_{0}) (resp. Diff(G/H)\operatorname{Diff}(G/H)) denotes the algebra of \mathbb{C}-valued real analytic differential operators on G/H0G_{\mathbb{R}}/H_{0} (resp. complex algebraic differential operators on G/HG/H). Since the complexificiation of 𝔥0\mathfrak{h}_{0} is 𝔥\mathfrak{h}, the map G/H0gH0gHG/HG_{\mathbb{R}}/H_{0}\ni gH_{0}\mapsto gH\in G/H is locally well-defined and the image of this map is a totally real submanifold of G/HG/H. The differential operators in ImΦ\operatorname{Im}\Phi can be viewed as holomorphic differential operators on the connected complex manifold G/HG/H. Hence such operators are zero if and only if their restrictions to a totally real submanifold are zero. This implies KerΦ=KerΦ0\operatorname{Ker}\Phi=\operatorname{Ker}\Phi_{0}.

Finally, we have the composition

𝒰(𝔤)ΦDiff(G/H)ψEndR(G/H).\mathcal{U}(\mathfrak{g})\stackrel{{\scriptstyle\Phi}}{{\rightarrow}}\operatorname{Diff}(G/H)\stackrel{{\scriptstyle\psi}}{{\rightarrow}}\operatorname{End}R(G/H).

Recall that HGH\subset G observable means that G/HG/H is quasi-affine, i.e. G/HG/H is isomorphic to an open subset of an affine variety. Since no nonzero differential operator on an affine variety annihilates all regular functions on that space, the map ψ\psi is injective. Therefore, KerΦ=Ker(ψΦ)\operatorname{Ker}\Phi=\operatorname{Ker}(\psi\circ\Phi). Now, we may decompose by the Peter-Weyl theorem

R(G/H)=FλR(G/H)Fλ(Fλ)HR(G/H)=\bigoplus_{{F_{\lambda}\subset R(G/H)}}F_{\lambda}\otimes(F_{\lambda}^{*})^{H}

and we note

Ann𝒰(𝔤)R(G/H)=FλR(G/H)Ann𝒰(𝔤)(Fλ).\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}R(G/H)=\bigcap_{F_{\lambda}\subset R(G/H)}\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}(F_{\lambda}).

Therefore, we have

(3.1) KerΦ0=KerΦ=FλR(G/H)Ann𝒰(𝔤)(Fλ)λ𝔞XAnn𝒰(𝔤)(Fλ)\operatorname{Ker}\Phi_{0}=\operatorname{Ker}\Phi=\bigcap_{F_{\lambda}\subset R(G/H)}\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}(F_{\lambda})\supset\bigcap_{\lambda\in\mathfrak{a}_{X}^{*}}\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}(F_{\lambda})

where the last inclusion follows from Lemma 3.3. Here and in what follows, we assume λ\lambda is dominant and integral whenever we write FλF_{\lambda}.

The cotangent bundle of XX is TX{(gH,ξ)ξ(𝔤/Ad(g)𝔥)}T^{*}X\simeq\{(gH,\xi)\mid\xi\in(\mathfrak{g}/\operatorname{Ad}(g)\mathfrak{h})^{*}\} and the moment map is given by

μ:TX𝔤,(x,ξ)ξ𝔤.\mu\colon T^{*}X\rightarrow\mathfrak{g}^{*},\quad(x,\xi)\mapsto\xi\in\mathfrak{g}^{*}.

As we stated in Theorem 1.2, [Kno94, Lemma 3.1 and Corollary 3.3] give the image of the moment map in terms of 𝔞X\mathfrak{a}_{X}^{*}:

μ(TX)¯=G𝔞X¯.\overline{\mu(T^{*}X)}=\overline{G\cdot\mathfrak{a}_{X}^{*}}.

In particular, the image of the moment map contains a dense subset of semisimple elements.

Let 𝔮X𝔤\mathfrak{q}_{X}\subset\mathfrak{g} be the Lie algebra of QXQ_{X} with Levi decomposition

QX=LXNX,𝔮X=𝔩X𝔫X,𝔫X=αΔ(𝔫X,𝔧)𝔤α.Q_{X}=L_{X}N_{X},\qquad\mathfrak{q}_{X}=\mathfrak{l}_{X}\oplus\mathfrak{n}_{X},\qquad\mathfrak{n}_{X}=\bigoplus_{\alpha\in\Delta(\mathfrak{n}_{X},\mathfrak{j})}\mathfrak{g}_{\alpha}.

Define

QX0:=LX0NX,J𝔞X:=Ker(𝒰(𝔤)Diff(G/QX0)).Q_{X}^{0}:=L_{X}^{0}N_{X},\qquad J_{\mathfrak{a}_{X}}:=\operatorname{Ker}\bigl{(}\mathcal{U}(\mathfrak{g})\rightarrow\operatorname{Diff}(G/Q_{X}^{0})\bigr{)}.

The following fact is the Corollary on page 453 of [BB82].

Fact 3.4 (Borho-Brylinski).

We have

(3.2) J𝔞X=Ann𝒰(𝔤)(𝒰(𝔤)𝒰(𝔮X0))=λ𝔞XAnn𝒰(𝔤)(𝒰(𝔤)𝒰(𝔮X)λ).\displaystyle J_{\mathfrak{a}_{X}}=\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}(\mathcal{U}(\mathfrak{g})\otimes_{\mathcal{U}(\mathfrak{q}_{X}^{0})}\mathbb{C})=\bigcap_{\lambda\in\mathfrak{a}_{X}^{*}}\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}(\mathcal{U}(\mathfrak{g})\otimes_{\mathcal{U}(\mathfrak{q}_{X})}\mathbb{C}_{\lambda}).

Here, \mathbb{C} is the trivial 𝒰(𝔮X0)\mathcal{U}(\mathfrak{q}_{X}^{0})-module, and λ\mathbb{C}_{\lambda} is the one-dimensional 𝒰(𝔮X)\mathcal{U}(\mathfrak{q}_{X})-module on which Z(𝔩X)Z(\mathfrak{l}_{X}) acts by λ\lambda.

Since each FλF_{\lambda} for λ𝔞X\lambda\in\mathfrak{a}_{X}^{*} is a quotient of 𝒰(𝔤)𝒰(𝔮X)λ\mathcal{U}(\mathfrak{g})\otimes_{\mathcal{U}(\mathfrak{q}_{X})}\mathbb{C}_{\lambda}, we deduce

Ann𝒰(𝔤)(𝒰(𝔤)𝒰(𝔮X)λ)Ann𝒰(𝔤)Fλ.\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}(\mathcal{U}(\mathfrak{g})\otimes_{\mathcal{U}(\mathfrak{q}_{X})}\mathbb{C}_{\lambda})\subset\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}F_{\lambda}.

Together with (3.1), and (3.2), this implies

(3.3) J𝔞Xλ𝔞XAnn𝒰(𝔤)(Fλ)KerΦ0.J_{\mathfrak{a}_{X}}\subset\bigcap_{\lambda\in\mathfrak{a}_{X}^{*}}\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}(F_{\lambda})\subset\operatorname{Ker}\Phi_{0}.

The following lemma simplifies the statement of our later result:

Lemma 3.5.

ρ(𝔫X)𝔞X\rho(\mathfrak{n}_{X})\in\mathfrak{a}_{X}^{*}.

Proof.

Since HH is unimodular, X=G/HX=G/H has a GG-invariant differential form of top degree. By restriction, it gives a QXQ_{X}-invariant form on QX×LXZQ_{X}\times^{L_{X}}Z. Therefore, the line bundle

(dimXTX)|ZdimZTZdim𝔫XTZ(QX×LXZ)(\bigwedge^{\dim X}T^{*}X)|_{Z}\simeq\bigwedge^{\dim Z}T^{*}Z\otimes\bigwedge^{\dim\mathfrak{n}_{X}}T^{*}_{Z}(Q_{X}\times^{L_{X}}Z)

has a nonzero LXL_{X}-invariant section, and hence in particular an LX0L_{X}^{0}-invariant section. Recall that LX0L_{X}^{0} acts trivially on ZZ and on TZT^{*}Z. On the other hand, the fibers of TZ(QX×LXZ)T^{*}_{Z}(Q_{X}\times^{L_{X}}Z) are identified with (𝔮X/𝔩X)(\mathfrak{q}_{X}/\mathfrak{l}_{X})^{*}. As a result, LX0L_{X}^{0} must act trivially on dim𝔫X(𝔮X/𝔩X)\bigwedge^{\dim\mathfrak{n}_{X}}(\mathfrak{q}_{X}/\mathfrak{l}_{X})^{*}, which implies ρ(𝔫X)\rho(\mathfrak{n}_{X}) is zero on 𝔩X0\mathfrak{l}_{X}^{0} and ρ(𝔫X)𝔞X\rho(\mathfrak{n}_{X})\in\mathfrak{a}_{X}^{*}. ∎

Suppose that VV is an irreducible (𝔤,K)(\mathfrak{g},K)-module and suppose there exists an injective linear map

VC(G/H0)V\hookrightarrow C^{\infty}(G_{\mathbb{R}}/H_{0})

which respects actions of 𝔤\mathfrak{g} and KK_{\mathbb{R}}. The enveloping algebra 𝒰(𝔤)\mathcal{U}(\mathfrak{g}) acts on VV via the map Φ0\Phi_{0} together with the restriction of the action of Diff(G/H0)\operatorname{Diff}(G_{\mathbb{R}}/H_{0}) on C(G/H0)C^{\infty}(G_{\mathbb{R}}/H_{0}) to VV. In particular, we have Ann𝒰(𝔤)(V)KerΦ0\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}(V)\supset\operatorname{Ker}\Phi_{0}. By (3.3), we obtain the following proposition.

Proposition 3.6.

If VV is an irreducible (𝔤,K)(\mathfrak{g},K)-module and there exists an injective linear map VC(G/H0)V\hookrightarrow C^{\infty}(G_{\mathbb{R}}/H_{0}) which respects actions of 𝔤\mathfrak{g} and KK_{\mathbb{R}}, then

Ann𝒰(𝔤)(V)J𝔞X.\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}(V)\supset J_{\mathfrak{a}_{X}}.

For an infinitesimal character ξ:Z(𝒰(𝔤))\xi\colon Z(\mathcal{U}(\mathfrak{g}))\rightarrow\mathbb{C}, define

Iξ:=𝒰(𝔤)Ker(Z(𝒰(𝔤))ξ).I_{\xi}:=\mathcal{U}(\mathfrak{g})\cdot\operatorname{Ker}(Z(\mathcal{U}(\mathfrak{g}))\stackrel{{\scriptstyle\xi}}{{\rightarrow}}\mathbb{C}).

Let WW be the Weyl group for Δ(𝔤,𝔧)\Delta(\mathfrak{g},\mathfrak{j}). Recall that there exists a natural algebra isomorphism (so-called the Harish-Chandra isomorphism) γ:Z(𝒰(𝔤))S(𝔧)W\gamma\colon Z(\mathcal{U}(\mathfrak{g}))\simeq S(\mathfrak{j})^{W}. If ξ:Z(𝒰(𝔤))\xi\colon Z(\mathcal{U}(\mathfrak{g}))\rightarrow\mathbb{C} is the infinitesimal character of VV, then we may compose with γ1\gamma^{-1} to give an element of 𝔧/W\mathfrak{j}^{*}/W or a representative ξ𝔧\xi\in\mathfrak{j}^{*}.

Lemma 3.7.

Suppose that VV is an irreducible (𝔤,K)(\mathfrak{g},K)-module with infinitesimal character ξ𝔧\xi\in\mathfrak{j}^{*} and Ann𝒰(𝔤)(V)J𝔞X\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}(V)\supset J_{\mathfrak{a}_{X}}. Then

(Wξ)(𝔞X+ρ𝔩X),(W\cdot\xi)\cap(\mathfrak{a}_{X}^{*}+\rho_{\mathfrak{l}_{X}})\neq\emptyset,

where we put

ρ𝔩X:=12αΔ(𝔩X,𝔧)Δ+(𝔤,𝔧)α.\rho_{\mathfrak{l}_{X}}:=\frac{1}{2}\sum_{\alpha\in\Delta(\mathfrak{l}_{X},\mathfrak{j})\cap\Delta^{+}(\mathfrak{g},\mathfrak{j})}\alpha.
Proof.

Suppose zZ(𝒰(𝔤))z\in Z(\mathcal{U}(\mathfrak{g})) with

γ(z)|𝔞X+ρ𝔩X=0.\gamma(z)|_{\mathfrak{a}_{X}^{*}+\rho_{\mathfrak{l}_{X}}}=0.

Recall that FλF_{\lambda} has infinitesimal character λ+ρ=λ+ρ𝔩X+ρ(𝔫X)\lambda+\rho=\lambda+\rho_{\mathfrak{l}_{X}}+\rho(\mathfrak{n}_{X}). In view of Lemma 3.5, zAnn𝒰(𝔤)(Fλ)z\in\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}(F_{\lambda}) for all λ𝔞X\lambda\in\mathfrak{a}_{X}^{*}, and by (3.3), zJ𝔞Xz\in J_{\mathfrak{a}_{X}}.

Now, assume that the conclusion of Lemma 3.7 is false. That is, assume that (Wξ)(𝔞X+ρ𝔩X)=(W\cdot\xi)\cap(\mathfrak{a}_{X}^{*}+\rho_{\mathfrak{l}_{X}})=\emptyset. Then we may choose a polynomial pPol(𝔧)Wp\in\operatorname{Pol}(\mathfrak{j}^{*})^{W} such that p(wξ)0p(w\cdot\xi)\neq 0 for all wWw\in W but

p|𝔞X+ρ𝔩X=0.p|_{\mathfrak{a}_{X}^{*}+\rho_{\mathfrak{l}_{X}}}=0.

Identify Pol(𝔧)WS(𝔧)W\operatorname{Pol}(\mathfrak{j}^{*})^{W}\simeq S(\mathfrak{j})^{W} in the usual way and write z=γ1(p)Z(𝒰(𝔤))z=\gamma^{-1}(p)\in Z(\mathcal{U}(\mathfrak{g})). Then zJ𝔞Xz\in J_{\mathfrak{a}_{X}} by the above argument. Since zγ(z)(ξ)Iξz-\gamma(z)(\xi)\in I_{\xi} by the definition of IξI_{\xi},

γ(z)(ξ)=z(zγ(z)(ξ))J𝔞X+Iξ.\gamma(z)(\xi)=z-(z-\gamma(z)(\xi))\in J_{\mathfrak{a}_{X}}+I_{\xi}.

But, then γ(z)(ξ)0\gamma(z)(\xi)\neq 0 implies 1J𝔞X+Iξ1\in J_{\mathfrak{a}_{X}}+I_{\xi}. On the other hand, Ann𝒰(𝔤)(V)J𝔞X+Iξ\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}(V)\supset J_{\mathfrak{a}_{X}}+I_{\xi} by our assumption. Hence we must have V=0V=0, which is a contradiction. ∎

For λZ(𝔩X)\lambda\in Z(\mathfrak{l}_{X})^{*} define the two-sided ideal

Jλ:=Ann𝒰(𝔤)(𝒰(𝔤)𝒰(𝔮X)λρ(𝔫X)).J_{\lambda}:=\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}(\mathcal{U}(\mathfrak{g})\otimes_{\mathcal{U}(\mathfrak{q}_{X})}\mathbb{C}_{\lambda-\rho(\mathfrak{n}_{X})}).

Note that JλIλ+ρ𝔩XJ_{\lambda}\supset I_{\lambda+\rho_{\mathfrak{l}_{X}}}, or equivalently, the generalized Verma module 𝒰(𝔤)𝒰(𝔮X)λρ(𝔫X)\mathcal{U}(\mathfrak{g})\otimes_{\mathcal{U}(\mathfrak{q}_{X})}\mathbb{C}_{\lambda-\rho(\mathfrak{n}_{X})} has the infinitesimal character λ+ρ𝔩X\lambda+\rho_{\mathfrak{l}_{X}}.

Lemma 3.8.

Suppose that VV is an irreducible (𝔤,K)(\mathfrak{g},K_{\mathbb{R}})-module and Ann𝒰(𝔤)(V)J𝔞X\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}(V)\supset J_{\mathfrak{a}_{X}}. Then there exists λ𝔞X\lambda\in\mathfrak{a}_{X}^{*} such that

Ann𝒰(𝔤)(V)Jλ.\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}(V)\supset J_{\lambda}.
Proof.

Let ξ𝔧\xi\in\mathfrak{j}^{*} be the infinitesimal character of VV. By Lemma 3.7, there exists a finite, nonempty collection {λ1,,λm}𝔞X\{\lambda_{1},\ldots,\lambda_{m}\}\subset\mathfrak{a}_{X}^{*} for which

(Wξ)(𝔞X+ρ𝔩X)={λ1+ρ𝔩X,,λm+ρ𝔩X}.(W\cdot\xi)\cap(\mathfrak{a}_{X}^{*}+\rho_{\mathfrak{l}_{X}})=\{\lambda_{1}+\rho_{\mathfrak{l}_{X}},\ldots,\lambda_{m}+\rho_{\mathfrak{l}_{X}}\}.

By an argument similar to the proof of [Soe89, Theorem 25], we obtain

i=1mJλiNJ𝔞X+Iξ\prod_{i=1}^{m}J_{\lambda_{i}}^{N}\subset J_{\mathfrak{a}_{X}}+I_{\xi}

for some large integer NN. Since VV is irreducible, our assumption Ann𝒰(𝔤)(V)J𝔞X+Iξ\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}(V)\supset J_{\mathfrak{a}_{X}}+I_{\xi} implies that

Ann𝒰(𝔤)(V)Jλi\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}(V)\supset J_{\lambda_{i}}

for some i{1,,m}i\in\{1,\dots,m\}. ∎

4. Reduction to quantizations of semisimple orbits

In the previous section we saw that the annihilators of irreducible subrepresentations of C(G/H0)C^{\infty}(G_{\mathbb{R}}/H_{0}) contain JλJ_{\lambda}, the annihilator of a generalized Verma module.

We will show in Proposition 4.1 that this statement of annihilators implies that representations are realized as global sections of 𝒟\mathscr{D}-modules on a partial flag variety unless the infinitesimal character is close to certain root hyperplanes.

Fix a holomorphic involution θ\theta of GG that commutes with σ\sigma and restricts to a Cartan involution on GG_{\mathbb{R}}. Let K=GθK=G^{\theta}.

If 𝔮=𝔩+𝔫\mathfrak{q}=\mathfrak{l}+\mathfrak{n} is a parabolic subalgebra of 𝔤\mathfrak{g} and Y:=G/QY:=G/Q is the corresponding partial flag variety, we write 𝒟Y,λ\mathscr{D}_{Y,\lambda} for the sheaf of twisted differential operators on YY with parameter λZ(𝔩)\lambda\in Z(\mathfrak{l})^{*} (see e.g. [Bie90]). Our normalization is that λ=ρ(𝔫)\lambda=\rho(\mathfrak{n}) corresponds to ordinary (untwisted) differential operators.

We retain the notation of the previous section. Recall that we defined a Levi subalgebra 𝔩X\mathfrak{l}_{X} and an ideal J𝔞X𝒰(𝔤)J_{\mathfrak{a}_{X}}\subset\mathcal{U}(\mathfrak{g}) for a homogeneous space X=G/HX=G/H.

Proposition 4.1.

There exists a constant d>0d>0 which depends only on GG such that if VV is an irreducible (𝔤,K)(\mathfrak{g},K)-module and if Ann𝒰(𝔤)(V)J𝔞X\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}(V)\supset J_{\mathfrak{a}_{X}}, then at least one of the following holds:

  1. (1)

    There exist a parabolic subalgebra 𝔮=𝔩X+𝔫\mathfrak{q}=\mathfrak{l}_{X}+\mathfrak{n}, a parameter λ𝔞X\lambda\in\mathfrak{a}_{X}^{*} in the good range, and a KK-equivariant 𝒟Y,λ\mathscr{D}_{Y,\lambda}-module \mathcal{M} on Y:=G/QY:=G/Q such that VΓ(Y,)V\simeq\Gamma(Y,\mathcal{M}). Here, we say λ\lambda is in the good range if λ+ρ𝔩X,α0\langle\lambda+\rho_{\mathfrak{l}_{X}},\alpha^{\vee}\rangle\not\in\mathbb{R}_{\geq 0} for every αΔ(𝔫,𝔧)\alpha\in\Delta(\mathfrak{n},\mathfrak{j}).

  2. (2)

    There exist a representative ξ𝔞X+ρ𝔩X\xi\in\mathfrak{a}_{X}^{*}+\rho_{\mathfrak{l}_{X}} of the infinitesimal character of VV and a root αΔ(𝔤,𝔧)Δ(𝔩X,𝔧)\alpha\in\Delta(\mathfrak{g},\mathfrak{j})\setminus\Delta(\mathfrak{l}_{X},\mathfrak{j}) such that |ξ,α|<d|\langle\xi,\alpha^{\vee}\rangle|<d.

Proof.

Take dd such that d>maxαΔ(𝔤,𝔧)|ρ𝔩X,α|d>\max_{\alpha\in\Delta(\mathfrak{g},\mathfrak{j})}|\langle\rho_{\mathfrak{l}_{X}},\alpha^{\vee}\rangle| and suppose the condition (2) in Proposition 4.1 does not hold.

By Lemma 3.8, there exists λ𝔞X\lambda\in\mathfrak{a}_{X}^{*} such that Ann𝒰(𝔤)(V)Jλ\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}(V)\supset J_{\lambda}. Then, take a parabolic subalgebra 𝔮=𝔩X+𝔫\mathfrak{q}=\mathfrak{l}_{X}+\mathfrak{n} of 𝔤\mathfrak{g} such that λ,α>0\langle\lambda,\alpha^{\vee}\rangle\not\in\mathbb{R}_{>0} for αΔ(𝔫,𝔧)\alpha\in\Delta(\mathfrak{n},\mathfrak{j}). For example, we may choose

Δ(𝔫,𝔧)={αΔ(𝔤,𝔧)Reλ,α<0}{αΔ(𝔫X,𝔧)Reλ,α=0}.\Delta(\mathfrak{n},\mathfrak{j})=\{\alpha\in\Delta(\mathfrak{g},\mathfrak{j})\mid\mathrm{Re}\,\langle\lambda,\alpha^{\vee}\rangle<0\}\cup\{\alpha\in\Delta(\mathfrak{n}_{X},\mathfrak{j})\mid\mathrm{Re}\,\langle\lambda,\alpha^{\vee}\rangle=0\}.

As we assumed that (2) does not hold, λ,α>d\langle\lambda,\alpha^{\vee}\rangle\not\in\mathbb{R}_{>-d} for αΔ(𝔫,𝔧)\alpha\in\Delta(\mathfrak{n},\mathfrak{j}) and then our choice of dd shows λ+ρ𝔩X,α0\langle\lambda+\rho_{\mathfrak{l}_{X}},\alpha^{\vee}\rangle\not\in\mathbb{R}_{\geq 0} namely, λ\lambda is in the good range with respect to 𝔮\mathfrak{q}.

We require the following fact which tells that annihilators of generalized Verma modules do not depend on the choice of polarizations.

Fact 4.2 ([Jan83, Corollar 15.27]).

Let 𝔮=𝔩+𝔫\mathfrak{q}=\mathfrak{l}+\mathfrak{n} and 𝔮=𝔩+𝔫\mathfrak{q}^{\prime}=\mathfrak{l}+\mathfrak{n}^{\prime} be two parabolic subalgebras with the same Levi factor. Then we have

Ann(𝒰(𝔤)𝒰(𝔮)λρ(𝔫))=Ann(𝒰(𝔤)𝒰(𝔮)λρ(𝔫))\operatorname{Ann}(\mathcal{U}(\mathfrak{g})\otimes_{\mathcal{U}(\mathfrak{q})}\mathbb{C}_{\lambda-\rho(\mathfrak{n})})=\operatorname{Ann}(\mathcal{U}(\mathfrak{g})\otimes_{\mathcal{U}(\mathfrak{q}^{\prime})}\mathbb{C}_{\lambda-\rho(\mathfrak{n}^{\prime})})

for λZ(𝔩)\lambda\in Z(\mathfrak{l})^{*}.

Let Y:=G/QY:=G/Q and let 𝒟Y,λ\mathscr{D}_{Y,\lambda} be the ring of twisted differential operators. We have a natural homomorphism

ϕ:𝒰(𝔤)Γ(Y,𝒟Y,λ).\phi\colon\mathcal{U}(\mathfrak{g})\to\Gamma(Y,\mathscr{D}_{Y,\lambda}).

The kernel of ϕ\phi is Ann(𝒰(𝔤)𝒰(𝔮)λρ(𝔫))\operatorname{Ann}(\mathcal{U}(\mathfrak{g})\otimes_{\mathcal{U}(\mathfrak{q})}\mathbb{C}_{\lambda-\rho(\mathfrak{n})}) (see [BB82, §3.6, Corollary] or [Soe89, Corollar 7]), which also equals JλJ_{\lambda} by Fact 4.2 above. Since λ\lambda is in the good range, ϕ\phi is surjective (see [Bie90, I.5.6 Proposition] for a proof). Hence ϕ\phi induces an isomorphism of algebras

𝒰(𝔤)/JλΓ(Y,𝒟Y,λ).\mathcal{U}(\mathfrak{g})/J_{\lambda}\simeq\Gamma(Y,\mathscr{D}_{Y,\lambda}).

Moreover, by [Bie90, I.6.3 Theorem],

VV𝒰(𝔤)/Jλ𝒟Y,λV\mapsto V\otimes_{\mathcal{U}(\mathfrak{g})/J_{\lambda}}\mathscr{D}_{Y,\lambda}

gives an equivalence of categories between (𝒰(𝔤)/Jλ)(\mathcal{U}(\mathfrak{g})/J_{\lambda})-modules and 𝒟Y,λ\mathscr{D}_{Y,\lambda}-modules, whose inverse is given by taking the space of global sections. Therefore, :=V𝒰(𝔤)/Jλ𝒟Y,λ\mathcal{M}:=V\otimes_{\mathcal{U}(\mathfrak{g})/J_{\lambda}}\mathscr{D}_{Y,\lambda} satisfies the condition (1) of Proposition 4.1.

For dd to be independent of VV or H0H_{0}, we may take the maximum of the above definition of dd for 𝔩X\mathfrak{l}_{X} running over all Levi subalgebras of 𝔤\mathfrak{g}. ∎

Let G^\widehat{G}_{\mathbb{R}} denote the set consisting of irreducible, unitary representations of GG_{\mathbb{R}}. Let X0=G/H0X_{0}=G_{\mathbb{R}}/H_{0}. Recall that we defined suppL2(X0)\operatorname{supp}L^{2}(X_{0}) to be the support of the Plancherel measure. Then by [Ber88, §2.3], for almost every (π,Vπ)(\pi,V_{\pi}) in suppL2(X0)\operatorname{supp}L^{2}(X_{0}), there exists an injective map

(Vπ)KC(X0)(V_{\pi})_{K}\hookrightarrow C^{\infty}(X_{0})

which respects actions of 𝔤\mathfrak{g} and KK_{\mathbb{R}}. Here, (Vπ)K(V_{\pi})_{K} denotes the underlying (𝔤,K)(\mathfrak{g},K)-module of VπV_{\pi}. Then by Proposition 3.6, we have Ann𝒰(𝔤)((Vπ)K)J𝔞X\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}\bigl{(}(V_{\pi})_{K}\bigr{)}\supset J_{\mathfrak{a}_{X}}.

In a way similar to [BD60, Théorèm 1], we can show that the set of irreducible unitarizable (𝔤,K)(\mathfrak{g},K)-modules VV satisfying

(4.1) Ann𝒰(𝔤)(V)J𝔞X\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}(V)\supset J_{\mathfrak{a}_{X}}

is closed in G^\widehat{G}_{\mathbb{R}}. That is, (4.1) is a closed condition in G^\widehat{G}_{\mathbb{R}}. Therefore, (4.1) is satisfied for every irreducible representation in suppL2(X0)\operatorname{supp}L^{2}(X_{0}).

Here is the main theorem in this section.

Theorem 4.3.

There exists a constant d>0d>0 which depends only on GG such that if (π,Vπ)suppL2(X0)(\pi,V_{\pi})\in\operatorname{supp}L^{2}(X_{0}), then at least one of the following holds:

  1. (1)

    There exist 𝔩𝔤\mathfrak{l}_{\mathbb{R}}\subset\mathfrak{g}_{\mathbb{R}} and (𝒪,Γ)(\mathcal{O},\Gamma) such that

    • The complexification 𝔩\mathfrak{l} of 𝔩\mathfrak{l}_{\mathbb{R}} is GG-conjugate to 𝔩X\mathfrak{l}_{X},

    • (𝒪,Γ)(\mathcal{O},\Gamma) is a semisimple orbital parameter such that ππ(𝒪,Γ)\pi\simeq\pi(\mathcal{O},\Gamma), and

    • 𝒪\mathcal{O} intersects 𝔞X1Z(𝔩)gr\mathfrak{a}_{X}^{*}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm gr}.

  2. (2)

    We can take a representative ξ𝔞X+ρ𝔩\xi\in\mathfrak{a}_{X}^{*}+\rho_{\mathfrak{l}} of the infinitesimal character of π\pi and a root αΔ(𝔤,𝔧)Δ(𝔩,𝔧)\alpha\in\Delta(\mathfrak{g},\mathfrak{j})\setminus\Delta(\mathfrak{l},\mathfrak{j}) such that |ξ,α|<d|\langle\xi,\alpha^{\vee}\rangle|<d.

Proof.

As mentioned above, we have Ann𝒰(𝔤)((Vπ)K)J𝔞X\operatorname{Ann}_{\mathcal{U}(\mathfrak{g})}\bigl{(}(V_{\pi})_{K}\bigr{)}\supset J_{\mathfrak{a}_{X}}. By Proposition 4.1, we may assume that Proposition 4.1 (1) holds, namely, there exist a parabolic subalgebra 𝔮=𝔩X+𝔫\mathfrak{q}=\mathfrak{l}_{X}+\mathfrak{n}, a parameter λ𝔞X\lambda\in\mathfrak{a}_{X}^{*} in the good range, and a KK-equivariant 𝒟Y,λ\mathscr{D}_{Y,\lambda}-module \mathcal{M} on Y:=G/QY:=G/Q such that (Vπ)KΓ(Y,)(V_{\pi})_{K}\simeq\Gamma(Y,\mathcal{M}).

Let Y~\tilde{Y} be the complete flag variety for GG and let p:Y~Yp\colon\tilde{Y}\twoheadrightarrow Y be the natural projection. Then we have natural isomorphisms

pppp𝒪Y𝒪Y,p_{*}p^{*}\mathcal{M}\simeq p_{*}p^{*}\mathcal{O}_{Y}\otimes_{\mathcal{O}_{Y}}\mathcal{M}\simeq\mathcal{M},

where pp_{*} denotes the direct image of 𝒪\mathcal{O}-modules. Hence

(Vπ)KΓ(Y~,p).(V_{\pi})_{K}\simeq\Gamma(\tilde{Y},p^{*}\mathcal{M}).

It is easy to see that this isomorphism respects (𝔤,K)(\mathfrak{g},K)-actions.

The pull-back pp^{*}\mathcal{M} is a twisted 𝒟Y~\mathscr{D}_{\tilde{Y}}-module. More precisely, it is a KK-equivariant 𝒟Y~,λ+ρ𝔩\mathscr{D}_{\tilde{Y},\lambda+\rho_{\mathfrak{l}}}-module. Let S~\tilde{S} be a dense KK-orbit in p1(supp)=supppp^{-1}(\operatorname{supp}\mathcal{M})=\operatorname{supp}p^{*}\mathcal{M}. Since Y~\tilde{Y} is 𝒟Y~,λ+ρ𝔩\mathscr{D}_{\tilde{Y},\lambda+\rho_{\mathfrak{l}}}-affine and Γ(Y~,p)\Gamma(\tilde{Y},p^{*}\mathcal{M}) is an irreducible (𝔤,K)(\mathfrak{g},K)-module, pp^{*}\mathcal{M} is a minimal extension of a KK-equivariant line bundle with connection ~\tilde{\mathcal{L}} on S~\tilde{S}. Fix a point oS~o\in\tilde{S} and let BB be the stabilizer of oo in GG. We may assume that BB contains a θ\theta-stable and σ\sigma-stable Cartan subgroup JJ. Write J=TAJ_{\mathbb{R}}=T_{\mathbb{R}}A_{\mathbb{R}} for the Cartan decomposition of the real form of JJ. By replacing QQ with its GG-conjugate, we may assume that QQ is the stabilizer of the point p(o)Yp(o)\in Y. Let Q=LNQ=LN be the Levi decomposition such that LJL\supset J. Note that LL is GG-conjugate to LXL_{X}.

Then by the correspondence between the Langlands classification and the Beilinson-Bernstein classification of (𝔤,K)(\mathfrak{g},K)-modules (see [Sch91], [KV95, Chapter XI]), the Langlands parameter of (Vπ)K(V_{\pi})_{K} is given as (J,γ,Δi+)(J_{\mathbb{R}},\gamma,\Delta_{i\mathbb{R}}^{+}) in the notation of [AvLTV20] such that dγ=λ+ρ𝔩(𝔧)d\gamma=\lambda+\rho_{\mathfrak{l}}(\in\mathfrak{j}^{*}) and Δi+\Delta_{i\mathbb{R}}^{+} is the set of imaginary roots which are not the roots in 𝔟\mathfrak{b}. We note that λ+ρ𝔩\lambda+\rho_{\mathfrak{l}} is regular as we assumed Proposition 4.1 (1).

Write dγ=Re(dγ)+1Im(dγ)d\gamma=\mathrm{Re}\,(d\gamma)+\sqrt{-1}\mathrm{Im}\,(d\gamma), where Re(dγ),Im(dγ)Hom(𝔧,)𝔧\mathrm{Re}\,(d\gamma),\mathrm{Im}\,(d\gamma)\in\operatorname{Hom}(\mathfrak{j}_{\mathbb{R}},\mathbb{R})\subset\mathfrak{j}^{*} and write dγ¯h=Re(dγ)+1Im(dγ)\overline{d\gamma}^{h}=-\mathrm{Re}\,(d\gamma)+\sqrt{-1}\mathrm{Im}\,(d\gamma) for the Hermitian dual. We use similar notation for any vector in 𝔧\mathfrak{j}^{*}.

We want to prove that Re(λ)=0\mathrm{Re}\,(\lambda)=0. Since the ρ\rho-shift of dγ|𝔱d\gamma|_{\mathfrak{t}_{\mathbb{R}}} is a differential of a character of TT_{\mathbb{R}}, the compact part of the Cartan subgroup, we have Re(dγ)|𝔱=0\mathrm{Re}\,(d\gamma)|_{\mathfrak{t}}=0. Moreover, since (π,Vπ)(\pi,V_{\pi}) is unitary, π\pi is isomorphic to its Hermitian dual. Hence by uniqueness in the Langlands classification, dγd\gamma and dγ¯h\overline{d\gamma}^{h} lie in the same Weyl group orbit. Let wW(Δ(𝔤,𝔧))w\in W(\Delta(\mathfrak{g},\mathfrak{j})) such that wdγ=dγ¯hw\cdot d\gamma=\overline{d\gamma}^{h}. The Weyl group W(Δ(𝔤,𝔧))W(\Delta(\mathfrak{g},\mathfrak{j})) preserves the real span of roots so it preserves 1𝔱𝔞\sqrt{-1}\mathfrak{t}_{\mathbb{R}}\oplus\mathfrak{a}_{\mathbb{R}}. Hence wIm(dγ)|𝔞=Im(dγ)|𝔞w\cdot\mathrm{Im}\,(d\gamma)|_{\mathfrak{a}}=\mathrm{Im}\,(d\gamma)|_{\mathfrak{a}}. Put

Δ1:={αΔ(𝔤,𝔧)Im(dγ)|𝔞,α=0}.\displaystyle\Delta_{1}:=\{\alpha\in\Delta(\mathfrak{g},\mathfrak{j})\mid\langle\mathrm{Im}\,(d\gamma)|_{\mathfrak{a}},\alpha^{\vee}\rangle=0\}.

Then wW(Δ1)w\in W(\Delta_{1}).

We note that

dγ=(Re(dγ)|𝔞+1Im(dγ)|𝔱)+1Im(dγ)|𝔞,\displaystyle d\gamma=\bigl{(}\mathrm{Re}\,(d\gamma)|_{\mathfrak{a}}+\sqrt{-1}\mathrm{Im}\,(d\gamma)|_{\mathfrak{t}}\bigr{)}+\sqrt{-1}\mathrm{Im}\,(d\gamma)|_{\mathfrak{a}},
λ=(Re(dγ)|𝔞+1Im(dγ)|𝔱ρ𝔩)+1Im(dγ)|𝔞,\displaystyle\lambda=\bigl{(}\mathrm{Re}\,(d\gamma)|_{\mathfrak{a}}+\sqrt{-1}\mathrm{Im}\,(d\gamma)|_{\mathfrak{t}}-\rho_{\mathfrak{l}}\bigr{)}+\sqrt{-1}\mathrm{Im}\,(d\gamma)|_{\mathfrak{a}},

and we have

Re(dγ)|𝔞+1Im(dγ)|𝔱,α,ρ𝔩,α,\displaystyle\langle\mathrm{Re}\,(d\gamma)|_{\mathfrak{a}}+\sqrt{-1}\mathrm{Im}\,(d\gamma)|_{\mathfrak{t}},\,\alpha^{\vee}\rangle\in\mathbb{R},\quad\langle\rho_{\mathfrak{l}},\alpha^{\vee}\rangle\in\mathbb{R},
1Im(dγ)|𝔞,α1.\displaystyle\langle\sqrt{-1}\mathrm{Im}\,(d\gamma)|_{\mathfrak{a}},\,\alpha^{\vee}\rangle\in\sqrt{-1}\mathbb{R}.

Hence for αΔ(𝔤,𝔧)\alpha\in\Delta(\mathfrak{g},\mathfrak{j}),

αΔ1dγ,αλ,α.\alpha\in\Delta_{1}\Leftrightarrow\langle d\gamma,\alpha^{\vee}\rangle\in\mathbb{R}\Leftrightarrow\langle\lambda,\alpha^{\vee}\rangle\in\mathbb{R}.

Since λ,α=0\langle\lambda,\alpha^{\vee}\rangle=0 for αΔ(𝔩)\alpha\in\Delta(\mathfrak{l}), we have Δ(𝔩)Δ1\Delta(\mathfrak{l})\subset\Delta_{1}. As we assumed Proposition 4.1 (1), λ\lambda is in the good range. Hence for αΔ1\alpha\in\Delta_{1},

(4.2) λ,α<0(resp.=0,>0) if αΔ(𝔫)(resp.αΔ(𝔩),Δ(𝔫)).\langle\lambda,\alpha^{\vee}\rangle<0\ (\text{resp.}\ =0,\ >0)\text{ if }\alpha\in\Delta(\mathfrak{n})\ (\text{resp.}\ \alpha\in\Delta(\mathfrak{l}),\ -\Delta(\mathfrak{n})).

Since (π,Vπ)(\pi,V_{\pi}) is unitary, Re(dγ)\mathrm{Re}\,(d\gamma) lies in a certain bounded region (see [Kna86, Chapter XVI, §5]). Suppose that the condition (2) of Theorem 4.3 does not hold for the constant dd greater than

max{ρ𝔩,ββΔ1}+max{2Re(dγ),ββΔ1}.\max\{\langle\rho_{\mathfrak{l}},\beta^{\vee}\rangle\mid\beta\in\Delta_{1}\}+\max\{\langle 2\mathrm{Re}\,(d\gamma),\beta^{\vee}\rangle\mid\beta\in\Delta_{1}\}.

Combining with (4.2), we have for αΔ1\alpha\in\Delta_{1},

(4.3) αΔ(𝔫)dγ,αd,αΔ(𝔫)dγ,αd,αΔ(𝔩)|dγ,α|max{ρ𝔩,ββΔ1}.\begin{split}&\alpha\in\Delta(\mathfrak{n})\Leftrightarrow\langle d\gamma,\alpha^{\vee}\rangle\leq-d,\qquad\alpha\in-\Delta(\mathfrak{n})\Leftrightarrow\langle d\gamma,\alpha^{\vee}\rangle\geq d,\\ &\alpha\in\Delta(\mathfrak{l})\Leftrightarrow|\langle d\gamma,\alpha^{\vee}\rangle|\leq\max\{\langle\rho_{\mathfrak{l}},\beta^{\vee}\rangle\mid\beta\in\Delta_{1}\}.\end{split}

If αΔ1Δ(𝔫)\alpha\in\Delta_{1}\cap\Delta(\mathfrak{n}), then

dγ,w1α\displaystyle\langle d\gamma,w^{-1}\cdot\alpha^{\vee}\rangle =wdγ,α\displaystyle=\langle w\cdot d\gamma,\alpha^{\vee}\rangle
=dγ¯h,α\displaystyle=\langle\overline{d\gamma}^{h},\alpha^{\vee}\rangle
=dγ2Re(dγ),α\displaystyle=\langle d\gamma-2\mathrm{Re}\,(d\gamma),\alpha^{\vee}\rangle
<max{ρ𝔩,ββΔ1},\displaystyle<-\max\{\langle\rho_{\mathfrak{l}},\beta^{\vee}\rangle\mid\beta\in\Delta_{1}\},

where the last inequality follows from our choice of dd and dγ,αd\langle d\gamma,\alpha^{\vee}\rangle\leq-d. Therefore, w(Δ1Δ(𝔫))=Δ1Δ(𝔫)w\cdot(\Delta_{1}\cap\Delta(\mathfrak{n}))=\Delta_{1}\cap\Delta(\mathfrak{n}) by (4.3) and hence wW(Δ(𝔩))w\in W(\Delta(\mathfrak{l})). If αΔ(𝔩)\alpha\in\Delta(\mathfrak{l}), then

|dγ,(α¯h)|=|dγ¯h,α|=|wdγ,α|=|dγ,(w1α)|<d|\langle d\gamma,(\overline{\alpha}^{h})^{\vee}\rangle|=|\langle\overline{d\gamma}^{h},\alpha^{\vee}\rangle|=|\langle w\cdot d\gamma,\alpha^{\vee}\rangle|=|\langle d\gamma,(w^{-1}\cdot\alpha)^{\vee}\rangle|<d

and hence α¯hΔ(𝔩)\overline{\alpha}^{h}\in\Delta(\mathfrak{l}), namely, Δ(𝔩)\Delta(\mathfrak{l}) is preserved by the Hermitian dual. In addition, wΔ(𝔩)w\in\Delta(\mathfrak{l}) implies that Re(dγ)=12(dγdγ¯h)=12(dγwdγ)\mathrm{Re}\,(d\gamma)=\frac{1}{2}(d\gamma-\overline{d\gamma}^{h})=\frac{1}{2}(d\gamma-w\cdot d\gamma) is a linear combination of Δ(𝔩)\Delta(\mathfrak{l}). Therefore, in view of the decomposition

Re(dγ)=Re(λ)+Re(ρ𝔩)Z(𝔩)([𝔩,𝔩]𝔧),\mathrm{Re}\,(d\gamma)=\mathrm{Re}\,(\lambda)+\mathrm{Re}\,(\rho_{\mathfrak{l}})\in Z(\mathfrak{l})^{*}\oplus([\mathfrak{l},\mathfrak{l}]\cap\mathfrak{j})^{*},

we obtain Re(λ)=0\mathrm{Re}\,(\lambda)=0. Since 𝔩:=𝔩𝔤\mathfrak{l}_{\mathbb{R}}:=\mathfrak{l}\cap\mathfrak{g}_{\mathbb{R}} is a real form of 𝔩\mathfrak{l}, we have proved that λ1Z(𝔩)\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}. As we assumed (1) of Proposition 4.1, we have λ𝔞X1Z(𝔩)gr\lambda\in\mathfrak{a}_{X}^{*}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm gr}.

Recall that \mathcal{M} is an irreducible KK-equivariant 𝒟Y,λ\mathscr{D}_{Y,\lambda}-module on Y=G/QY=G/Q such that (Vπ)KΓ(Y,)(V_{\pi})_{K}\simeq\Gamma(Y,\mathcal{M}). Let SS be the KK-orbit in YY containing p(o)p(o). Then S=p(S~)S=p(\tilde{S}) and supp=S¯\operatorname{supp}\mathcal{M}=\overline{S}, the closure of SS. Let i:SYi\colon S\hookrightarrow Y and i~:p1(S)Y~\tilde{i}\colon p^{-1}(S)\hookrightarrow\tilde{Y} denote the natural inclusion maps. We have the following commutative diagram:

Let i:=Li[dimSdimY]i^{\dagger}:=Li^{*}[\dim S-\dim Y] denotes the shifted inverse image functor for 𝒟\mathscr{D}-modules as in [HTT08]. Since supp=S¯\operatorname{supp}\mathcal{M}=\overline{S}, the complex ii^{\dagger}\mathcal{M} is concentrated in one degree, namely, Hq(i)=0H^{q}(i^{\dagger}\mathcal{M})=0 for q0q\neq 0. Let :=H0(i)\mathcal{L}:=H^{0}(i^{\dagger}\mathcal{M}), which is a KK-equivariant twisted 𝒟\mathscr{D}-module on SS. By an isomorphism pii~pp^{*}i^{\dagger}\mathcal{M}\simeq\tilde{i}^{\dagger}p^{*}\mathcal{M}, we have p|S~~p^{*}\mathcal{L}|_{\tilde{S}}\simeq\tilde{\mathcal{L}}. Hence \mathcal{L} must be a KK-equivariant line bundle.

Next, decompose the map ii into i=jki=j\circ k:

S𝑘Y(S¯S)𝑗YS\xrightarrow{k}Y\setminus(\overline{S}\setminus S)\xrightarrow{j}Y

so jj is an open immersion and kk is a closed immersion. By the definition of \mathcal{L}, we have k(j1)k^{\dagger}(j^{-1}\mathcal{M})\simeq\mathcal{L}. Since j1j^{-1}\mathcal{M} is supported on SS, there is an isomorphism j1k+j^{-1}\mathcal{M}\simeq k_{+}\mathcal{L} by Kashiwara’s equivalence. Then we get a nonzero element in

Hom(j1,k+)Hom(,jk+).\operatorname{Hom}(j^{-1}\mathcal{M},k_{+}\mathcal{L})\simeq\operatorname{Hom}(\mathcal{M},j_{*}k_{+}\mathcal{L}).

Hence Hom(,i+)0\operatorname{Hom}(\mathcal{M},i_{+}\mathcal{L})\neq 0. Write the KK-equivariant line bundle \mathcal{L} on SS as =K×(QK)τ\mathcal{L}=K\times_{(Q\cap K)}\tau for a character of QKQ\cap K. As in Section 2, we define a unitary character Γλ\Gamma_{\lambda} of L~\widetilde{L_{\mathbb{R}}} such that dΓλ=λ1Z(𝔩)d\Gamma_{\lambda}=\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*} and

(Γλeρ(θσ(𝔫)))|LKtop(𝔨/(𝔩𝔨))τ|LK.(\Gamma_{\lambda}\otimes e^{-\rho(\theta\sigma(\mathfrak{n}))})|_{L_{\mathbb{R}}\cap K_{\mathbb{R}}}\otimes\wedge^{\rm top}(\mathfrak{k}/(\mathfrak{l}\cap\mathfrak{k}))\simeq\tau|_{L_{\mathbb{R}}\cap K_{\mathbb{R}}}.

Notice that the roles of QQ and σ(Q)\sigma(Q) here are interchanged from Section 2. Then by (2.1)

q(1)qΓ(Y,Rqi+)=π(𝒪,Γ).\sum_{q}(-1)^{q}\Gamma(Y,R^{q}i_{+}\mathcal{L})=\pi(\mathcal{O},\Gamma).

We saw above that Hom(,R0i+)0\operatorname{Hom}(\mathcal{M},R^{0}i_{+}\mathcal{L})\neq 0 and hence Hom𝔤,K((Vπ)K,Γ(Y,R0i+))0\operatorname{Hom}_{\mathfrak{g},K}((V_{\pi})_{K},\Gamma(Y,R^{0}i_{+}\mathcal{L}))\neq 0. On the other hand, Rqi+R^{q}i_{+}\mathcal{L} for q>0q>0 is supported on S¯S\overline{S}\setminus S. Hence the irreducible (𝔤,K)(\mathfrak{g},K)-module (Vπ)K(V_{\pi})_{K} does not appear in the composition series of Γ(Y,Rqi+)\Gamma(Y,R^{q}i_{+}\mathcal{L}) for q>0q>0. Since π(𝒪,Γ)\pi(\mathcal{O},\Gamma) is irreducible, we conclude that π(𝒪,Γ)(Vπ)K\pi(\mathcal{O},\Gamma)\simeq(V_{\pi})_{K}. Thus, the condition (1) in Theorem 4.3 holds. ∎

Note that if π\pi satisfies (1) in Theorem 4.3, then πG^𝔩X\pi\in\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{X}} in the notation of Section 1.

Now we prove Corollary 1.6.

Proof of Corollary 1.6.

(i) is a direct consequence of Theorem 1.3, which will be proved in Section 8.

To prove (ii), recall the Langlands classification of irreducible admissible representations of GG_{\mathbb{R}}. In the notation of [AvLTV20], they are parametrized by triples (J,γ,Δi+)(J_{\mathbb{R}},\gamma,\Delta_{i\mathbb{R}}^{+}). Write π(J,γ,Δi+)\pi(J_{\mathbb{R}},\gamma,\Delta_{i\mathbb{R}}^{+}) for the irreducible representation of GG_{\mathbb{R}} corresponding to (J,γ,Δi+)(J_{\mathbb{R}},\gamma,\Delta_{i\mathbb{R}}^{+}). Then the infinitesimal character of π(J,γ,Δi+)\pi(J_{\mathbb{R}},\gamma,\Delta_{i\mathbb{R}}^{+}) is given by the WW-orbit through dγd\gamma.

Since there are finitely many Cartan subgroups JJ_{\mathbb{R}} up to conjugation and the asymptotic cone commutes with finite union, we may fix JJ_{\mathbb{R}} and treat only representations of the form π(J,γ,Δi+)\pi(J_{\mathbb{R}},\gamma,\Delta_{i\mathbb{R}}^{+}). By replacing 𝔧\mathfrak{j} in the statement of Corollary 1.6 with its conjugation, we may moreover assume that the complexified Lie algebra of our fixed JJ_{\mathbb{R}} is the same as 𝔧\mathfrak{j} in the statement.

Suppose that π(J,γ,Δi+)suppL2(X0)G^𝔩X\pi(J_{\mathbb{R}},\gamma,\Delta_{i\mathbb{R}}^{+})\in\operatorname{supp}L^{2}(X_{0})\setminus\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{X}}. Then by Theorem 4.3, it satisfies (2) in the theorem. Hence there exist wWw\in W and ξ𝔧\xi\in\mathfrak{j}^{*} such that dγ=wξd\gamma=w\cdot\xi, ξ𝔞X+ρ𝔩X\xi\in\mathfrak{a}_{X}^{*}+\rho_{\mathfrak{l}_{X}} and |ξ,α|<d|\langle\xi,\alpha^{\vee}\rangle|<d for some αΔ(𝔤,𝔧)Δ(𝔩X,𝔧)\alpha\in\Delta(\mathfrak{g},\mathfrak{j})\setminus\Delta(\mathfrak{l}_{X},\mathfrak{j}). Therefore,

(4.4) AC({dγπ(J,γ,Δi+)suppL2(X0)G^𝔩X})wWαΔ(𝔤,𝔧)Δ(𝔩X,𝔧)wAC({ξ𝔞X+ρ𝔩X:|ξ,α|<d})=wWαΔ(𝔤,𝔧)Δ(𝔩X,𝔧)w(𝔞Xα).\begin{split}&\operatorname{AC}\bigl{(}\bigl{\{}d\gamma\mid\pi(J_{\mathbb{R}},\gamma,\Delta_{i\mathbb{R}}^{+})\in\operatorname{supp}L^{2}(X_{0})\setminus\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{X}}\bigr{\}}\bigr{)}\\ &\subset\bigcup_{w\in W}\bigcup_{\alpha\in\Delta(\mathfrak{g},\mathfrak{j})\setminus\Delta(\mathfrak{l}_{X},\mathfrak{j})}w\cdot\operatorname{AC}(\{\xi\in\mathfrak{a}_{X}^{*}+\rho_{\mathfrak{l}_{X}}:|\langle\xi,\alpha^{\vee}\rangle|<d\})\\ &=\bigcup_{w\in W}\bigcup_{\alpha\in\Delta(\mathfrak{g},\mathfrak{j})\setminus\Delta(\mathfrak{l}_{X},\mathfrak{j})}w\cdot(\mathfrak{a}_{X}^{*}\cap\alpha^{\perp}).\end{split}

Consider the decomposition dγ=Re(dγ)+1Im(dγ)d\gamma=\mathrm{Re}\,(d\gamma)+\sqrt{-1}\mathrm{Im}\,(d\gamma) with Re(dγ),Im(dγ)𝔧\mathrm{Re}\,(d\gamma),\mathrm{Im}\,(d\gamma)\in\mathfrak{j}_{\mathbb{R}}^{*}. If π(J,γ,Δi+)\pi(J_{\mathbb{R}},\gamma,\Delta_{i\mathbb{R}}^{+}) is unitary, then Re(dγ)\mathrm{Re}\,(d\gamma) is bounded. Hence the left hand side of (4.4) is contained in 1𝔧\sqrt{-1}\mathfrak{j}_{\mathbb{R}}^{*}. Define

1𝔧,X,sing:=1𝔧wWαΔ(𝔤,𝔧)Δ(𝔩X,𝔧)w(𝔞Xα).\sqrt{-1}\mathfrak{j}_{\mathbb{R},X,\mathrm{sing}}^{*}:=\sqrt{-1}\mathfrak{j}_{\mathbb{R}}^{*}\cap\bigcup_{w\in W}\bigcup_{\alpha\in\Delta(\mathfrak{g},\mathfrak{j})\setminus\Delta(\mathfrak{l}_{X},\mathfrak{j})}w\cdot(\mathfrak{a}_{X}^{*}\cap\alpha^{\perp}).

Then

AC({dγπ(J,γ,Δi+)suppL2(X0)G^𝔩X})1𝔧,X,sing\operatorname{AC}\bigl{(}\bigl{\{}d\gamma\mid\pi(J_{\mathbb{R}},\gamma,\Delta_{i\mathbb{R}}^{+})\in\operatorname{supp}L^{2}(X_{0})\setminus\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{X}}\bigr{\}}\bigr{)}\subset\sqrt{-1}\mathfrak{j}_{\mathbb{R},X,\mathrm{sing}}^{*}

and it is easy to see that

dim1𝔧,X,sing<dim𝔞X.\dim_{\mathbb{R}}\sqrt{-1}\mathfrak{j}_{\mathbb{R},X,\mathrm{sing}}^{*}<\dim_{\mathbb{C}}\mathfrak{a}_{X}.

Since

AC(πsuppL2(G/H0)G^𝔩Xχπ)\displaystyle\operatorname{AC}\Bigl{(}\bigcup_{\pi\in\operatorname{supp}L^{2}(G_{\mathbb{R}}/H_{0})\setminus\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{X}}}\chi_{\pi}\Bigr{)}
=WAC({dγπ(J,γ,Δi+)suppL2(X0)G^𝔩X}),\displaystyle=W\cdot\operatorname{AC}\bigl{(}\bigl{\{}d\gamma\mid\pi(J_{\mathbb{R}},\gamma,\Delta_{i\mathbb{R}}^{+})\in\operatorname{supp}L^{2}(X_{0})\setminus\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{X}}\bigr{\}}\bigr{)},

Corollary 1.6 (ii) is proved. ∎

To describe representations of type (2) in Theorem 4.3, we introduce some notation. For a Levi subalgebra 𝔩𝔤\mathfrak{l}\subset\mathfrak{g}, its Cartan subalgebra 𝔧𝔩\mathfrak{j}\subset\mathfrak{l} and a constant d>0d>0, define subsets Ξ(𝔩,d)𝔧\Xi(\mathfrak{l},d)\subset\mathfrak{j}^{*} and G^(𝔩,d)G^\widehat{G}_{\mathbb{R}}(\mathfrak{l},d)\subset\widehat{G}_{\mathbb{R}} by

Ξ(𝔩,d):={ξZ(𝔩)+ρ𝔩αΔ(𝔤,𝔧)Δ(𝔩,𝔧) such that |ξ,α|<d},\displaystyle\Xi(\mathfrak{l},d):=\{\xi\in Z(\mathfrak{l})^{*}+\rho_{\mathfrak{l}}\mid\exists\alpha\in\Delta(\mathfrak{g},\mathfrak{j})\setminus\Delta(\mathfrak{l},\mathfrak{j})\text{ such that }|\langle\xi,\alpha^{\vee}\rangle|<d\},
G^(𝔩,d):={πG^The infinitesimal character of π has a representative in Ξ(𝔩,d)}.\displaystyle\widehat{G}_{\mathbb{R}}(\mathfrak{l},d):=\{\pi\in\widehat{G}_{\mathbb{R}}\mid\text{The infinitesimal character of $\pi$ has a representative in $\Xi(\mathfrak{l},d)$}\}.

For the proof of main theorems in §8, we need Lemma 4.4, which states that the contribution to singular spectrum from representations of type (2) in Theorem 4.3 is small.

For a unitary representation (Π,VΠ)(\Pi,V_{\Pi}) of GG_{\mathbb{R}}, define the wave front set and the singular spectrum of Π\Pi by

WF(Π)=u,vVΠWFe(π(g)u,v)¯,SS(Π)=u,vVΠSSe(π(g)u,v)¯.\operatorname{WF}(\Pi)=\overline{\bigcup_{u,v\in V_{\Pi}}\operatorname{WF}_{e}(\pi(g)u,v)},\quad\operatorname{SS}(\Pi)=\overline{\bigcup_{u,v\in V_{\Pi}}\operatorname{SS}_{e}(\pi(g)u,v)}.

Here, WFe(Π(g)u,v)\operatorname{WF}_{e}(\Pi(g)u,v) is the wave front set of the matrix coefficient function (Π(g)u,v)(\Pi(g)u,v) at eGe\in G, Similarly, SSe(Π(g)u,v)\operatorname{SS}_{e}(\Pi(g)u,v) is the singular spectrum (or the analytic wave front set) of (Π(g)u,v)(\Pi(g)u,v) at ee. Both WF(Π)\operatorname{WF}(\Pi) and SS(Π)\operatorname{SS}(\Pi) are closed GG-invariant subset of 𝔤(TeG)\mathfrak{g}^{*}(\simeq T^{*}_{e}G). We always have WF(Π)SS(Π)\operatorname{WF}(\Pi)\subset\operatorname{SS}(\Pi). See [HHO16] for the equivalence with Howe’s original definition [How81] of the wave front set. We note that a relationship between the singular spectrum of functions and the spectrum of representations was studied in Kashiwara-Vergne [KV79]. Such a microlocal point of view also appeared in Kobayashi’s theory [Kob98a, Kob98b] on the admissibility of restrictions of representations.

Lemma 4.4.

Let Π\Pi be a unitary representation of GG_{\mathbb{R}} and suppΠG^(𝔩,d)\operatorname{supp}\Pi\subset\widehat{G}_{\mathbb{R}}(\mathfrak{l},d). Then WF(Π)(GZ(𝔩)reg)=SS(Π)(GZ(𝔩)reg)=\operatorname{WF}(\Pi)\cap(G\cdot Z(\mathfrak{l})^{*}_{\mathrm{reg}})=\operatorname{SS}(\Pi)\cap(G\cdot Z(\mathfrak{l})^{*}_{\mathrm{reg}})=\emptyset.

Proof.

The proof follows the same line of arguments as in the proof of [Har18, Theorem 1.1].

To each Langlands parameter Γ\Gamma, one defines the Langlands quotient J(Γ)J(\Gamma), which is an irreducible representation of GG_{\mathbb{R}}. In [Har18, Section 2], we associate a contour C(Γ)𝔤C(\Gamma)\subset\mathfrak{g}^{*}. By [Har18, Lemma 3.4, Lemma 3.5], it is enough to show:

AC(J(Γ)G^(𝔩,d)C(Γ))(GZ(𝔩)reg)=.\operatorname{AC}\Bigl{(}\bigcup_{J(\Gamma)\in\widehat{G}_{\mathbb{R}}(\mathfrak{l},d)}C(\Gamma)\Bigr{)}\cap(G\cdot Z(\mathfrak{l})^{*}_{\mathrm{reg}})=\emptyset.

If J(Γ)J(\Gamma) has infinitesimal character ξ𝔧\xi\in\mathfrak{j}^{*}, then C(Γ)GξC(\Gamma)\subset G\cdot\xi by the definition of C(Γ)C(\Gamma). Hence

AC(J(Γ)G^(𝔩,d)C(Γ))AC(GΞ(𝔩,d)).\operatorname{AC}\Bigl{(}\bigcup_{J(\Gamma)\in\widehat{G}_{\mathbb{R}}(\mathfrak{l},d)}C(\Gamma)\Bigr{)}\subset\operatorname{AC}(G\cdot\Xi(\mathfrak{l},d)).

Since AC(GΞ(𝔩,d))\operatorname{AC}(G\cdot\Xi(\mathfrak{l},d)) is GG-stable, it is enough to show that

(4.5) AC(GΞ(𝔩,d))Z(𝔩)reg=.\operatorname{AC}(G\cdot\Xi(\mathfrak{l},d))\cap Z(\mathfrak{l})^{*}_{\mathrm{reg}}=\emptyset.

Let W=W(𝔤,𝔧)W=W(\mathfrak{g},\mathfrak{j}) be the Weyl group which acts on 𝔧\mathfrak{j}^{*}. We claim that

(4.6) AC(GS)𝔧=WAC(S)\displaystyle\operatorname{AC}(G\cdot S)\cap\mathfrak{j}^{*}=W\cdot\operatorname{AC}(S)

for any subset S𝔧S\subset\mathfrak{j}^{*}. Indeed, we have

AC(GS)AC(WS)=WAC(S).\displaystyle\operatorname{AC}(G\cdot S)\supset\operatorname{AC}(W\cdot S)=W\cdot\operatorname{AC}(S).

For the other inclusion, let ξAC(GS)𝔧\xi\in\operatorname{AC}(G\cdot S)\cap\mathfrak{j}^{*}. Then there exist giGg_{i}\in G, siSs_{i}\in S, and ti>0t_{i}\in\mathbb{R}_{>0} for ii\in\mathbb{N} such that ti+t_{i}\to+\infty and ti1(gisi)ξt_{i}^{-1}(g_{i}\cdot s_{i})\to\xi when ii\to\infty. Let p:𝔤𝔧/Wp\colon\mathfrak{g}^{*}\to\mathfrak{j}^{*}/W be the map induced from the isomorphism S(𝔤)GS(𝔧)WS(\mathfrak{g})^{G}\simeq S(\mathfrak{j})^{W}. By applying pp to the convergent sequence, we obtain ti1p(si)ξt_{i}^{-1}p(s_{i})\to\xi in 𝔧/W\mathfrak{j}^{*}/W, which implies ξAC(WS)\xi\in\operatorname{AC}(W\cdot S).

Plugging S=Ξ(𝔩,d)S=\Xi(\mathfrak{l},d) into (4.6), we get

AC(GΞ(𝔩,d))Z(𝔩)reg=(WAC(Ξ(𝔩,d)))Z(𝔩)reg.\operatorname{AC}(G\cdot\Xi(\mathfrak{l},d))\cap Z(\mathfrak{l})^{*}_{\mathrm{reg}}=\bigl{(}W\cdot\operatorname{AC}(\Xi(\mathfrak{l},d))\bigr{)}\cap Z(\mathfrak{l})^{*}_{\mathrm{reg}}.

If λAC(Ξ(𝔩,d))\lambda\in\operatorname{AC}(\Xi(\mathfrak{l},d)), then 𝔤(λ)𝔩\mathfrak{g}(\lambda)\supsetneq\mathfrak{l}. Hence λWAC(Ξ(𝔩,d))\lambda\in W\cdot\operatorname{AC}(\Xi(\mathfrak{l},d)) implies dim𝔤(λ)>dim𝔩\dim\mathfrak{g}(\lambda)>\dim\mathfrak{l}. Therefore, (WAC(Ξ(𝔩,d)))Z(𝔩)reg=\bigl{(}W\cdot\operatorname{AC}(\Xi(\mathfrak{l},d))\bigr{)}\cap Z(\mathfrak{l})^{*}_{\mathrm{reg}}=\emptyset and (4.5) is proved. ∎

5. Wave front sets of direct integrals for a Levi, part 1

Let 𝔩\mathfrak{l}_{\mathbb{R}} be a Levi subalgebra of 𝔤\mathfrak{g}_{\mathbb{R}}. Define a subset G^𝔩G^\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{\mathbb{R}}}\subset\widehat{G}_{\mathbb{R}} as

G^𝔩={πG^Γλ such that ππ(𝔩,Γλ) and λ1Z(𝔩)gr}.\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{\mathbb{R}}}=\{\pi\in\widehat{G}_{\mathbb{R}}\mid\exists\Gamma_{\lambda}\text{ such that }\pi\simeq\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\text{ and }\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\text{gr}}\}.

The definition of π(𝔩,Γλ)\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda}) was given in §2. For a complex Levi subalgebra 𝔩𝔤\mathfrak{l}^{\prime}\subset\mathfrak{g}, we defined G^𝔩\widehat{G}_{\mathbb{R}}^{\mathfrak{l}^{\prime}} in Section 1. By these definitions,

G^𝔩=𝔩G^𝔩,\widehat{G}_{\mathbb{R}}^{\mathfrak{l}^{\prime}}=\bigcup_{\mathfrak{l}_{\mathbb{R}}}\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{\mathbb{R}}},

where 𝔩\mathfrak{l}_{\mathbb{R}} runs over all Levi subalgebras of 𝔤\mathfrak{g}_{\mathbb{R}} such that 𝔩𝔩\mathfrak{l}\sim\mathfrak{l}^{\prime}.

We want to prove the following theorem on the wave front set and the singular spectrum:

Theorem 5.1.

Let 𝔩\mathfrak{l}_{\mathbb{R}} be a Levi subalgebra of 𝔤\mathfrak{g}_{\mathbb{R}}. Suppose that (Π,VΠ)(\Pi,V_{\Pi}) is a unitary representation of GG_{\mathbb{R}} which is isomorphic to a direct integral of representations in G^𝔩\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{\mathbb{R}}}:

ΠπG^𝔩πn(π)𝑑mΠ.\Pi\simeq\int^{\oplus}_{\pi\in\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{\mathbb{R}}}}\pi^{\oplus n(\pi)}dm_{\Pi}.

Then

WF(Π)(GZ(𝔩)reg)\displaystyle\operatorname{WF}(\Pi)\cap(G\cdot Z(\mathfrak{l})^{*}_{\rm reg}) =SS(Π)(GZ(𝔩)reg)\displaystyle=\operatorname{SS}(\Pi)\cap(G\cdot Z(\mathfrak{l})^{*}_{\rm reg})
=AC(π(𝔩,Γλ)supp(mΠ)Gλ)(GZ(𝔩)reg).\displaystyle=\operatorname{AC}\Biggl{(}\bigcup_{\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\in\operatorname{supp}(m_{\Pi})}G_{\mathbb{R}}\cdot\lambda\Biggr{)}\cap(G\cdot Z(\mathfrak{l})^{*}_{\rm reg}).

In this section, we prove the following inclusion.

Lemma 5.2.

In the setting of Theorem 5.1,

SS(Π)(GZ(𝔩)reg)AC(π(𝔩,Γλ)supp(mΠ)Gλ).\operatorname{SS}(\Pi)\cap(G\cdot Z(\mathfrak{l})^{*}_{\rm reg})\subset\operatorname{AC}\Biggl{(}\bigcup_{\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\in\operatorname{supp}(m_{\Pi})}G_{\mathbb{R}}\cdot\lambda\Biggr{)}.

The proof of Theorem 5.1 will be completed in the subsequent two sections.

Before starting the proof of Lemma 5.2, we see that G^𝔩\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{\mathbb{R}}} is a locally closed subset of G^\widehat{G}_{\mathbb{R}} with respect to the Fell topology. Let {πj}j\{\pi^{j}\}_{j} be a sequence in G^𝔩\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{\mathbb{R}}} which converges to πG^\pi\in\widehat{G}_{\mathbb{R}}. Let πj=π(𝒪j,Γj)\pi^{j}=\pi(\mathcal{O}^{j},\Gamma^{j}) and 𝒪j=Gλj\mathcal{O}^{j}=G_{\mathbb{R}}\cdot\lambda^{j}. Recall from Section 2 and [HO20, §2] that π(𝒪j,Γj)\pi(\mathcal{O}^{j},\Gamma^{j}) is defined as a unitary parabolic induction for a parabolic subgroup P=MA(NP)P_{\mathbb{R}}=M_{\mathbb{R}}A_{\mathbb{R}}(N_{P})_{\mathbb{R}}. Since there are only finitely many possibilities for PP_{\mathbb{R}}, we may assume that PP_{\mathbb{R}} does not depend on jj by passing to a subsequence. We have a decomposition λj=λcj+λnj\lambda^{j}=\lambda^{j}_{c}+\lambda^{j}_{n} and let (𝒪j)M=Mλcj(\mathcal{O}^{j})^{M_{\mathbb{R}}}=M_{\mathbb{R}}\cdot\lambda^{j}_{c}. Then we can define a semisimple orbital parameter ((𝒪j)M,(Γj)M)\bigl{(}(\mathcal{O}^{j})^{M_{\mathbb{R}}},(\Gamma^{j})^{M_{\mathbb{R}}}\bigr{)} for MM_{\mathbb{R}} such that π(𝒪j,Γj)\pi(\mathcal{O}^{j},\Gamma^{j}) is induced from π((𝒪j)M,(Γj)M)\pi((\mathcal{O}^{j})^{M_{\mathbb{R}}},(\Gamma^{j})^{M_{\mathbb{R}}}). By [BD60], the map G^𝔧/W\widehat{G}_{\mathbb{R}}\to\mathfrak{j}^{*}/W sending an irreducible unitary representation to its infinitesimal character is continuous. Therefore, the infinitesimal character of πj\pi^{j} converges to that of π\pi. This implies that λcj\lambda_{c}^{j} is bounded and hence there are only finitely many possibilities for ((𝒪j)M,(Γj)M)\bigl{(}(\mathcal{O}^{j})^{M_{\mathbb{R}}},(\Gamma^{j})^{M_{\mathbb{R}}}\bigr{)}. Passing to a subsequence, we may assume all parameters ((𝒪j)M,(Γj)M)\bigl{(}(\mathcal{O}^{j})^{M_{\mathbb{R}}},(\Gamma^{j})^{M_{\mathbb{R}}}\bigr{)} are the same so let (𝒪M,ΓM)=((𝒪j)M,(Γj)M)(\mathcal{O}^{M_{\mathbb{R}}},\Gamma^{M_{\mathbb{R}}})=\bigl{(}(\mathcal{O}^{j})^{M_{\mathbb{R}}},(\Gamma^{j})^{M_{\mathbb{R}}}\bigr{)} and λc=λcj\lambda_{c}=\lambda^{j}_{c}. We may also assume that λnj\lambda_{n}^{j} converges to λn1𝔞\lambda_{n}\in\sqrt{-1}\mathfrak{a}_{\mathbb{R}}^{*}. Then as noted in the proof of [SRV98, Corollary 8.9], π\pi is isomorphic to an irreducible constituent of IndPG((𝒪M,ΓM)eλn)\operatorname{Ind}_{P_{\mathbb{R}}}^{G_{\mathbb{R}}}\bigl{(}(\mathcal{O}^{M_{\mathbb{R}}},\Gamma^{M_{\mathbb{R}}})\boxtimes e^{\lambda_{n}}\bigr{)}. If λc+λn\lambda_{c}+\lambda_{n} is in the good range, then the induced representation is irreducible and πG^𝔩\pi\in\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{\mathbb{R}}}. Otherwise, λc+λn+ρ𝔩\lambda_{c}+\lambda_{n}+\rho_{\mathfrak{l}} is singular and π\pi has the singular infinitesimal character. Since the set of representations with singular infinitesimal characters is closed in G^\widehat{G}_{\mathbb{R}}, the above argument proves that G^𝔩\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{\mathbb{R}}} is locally closed.

Let 𝔮𝔤\mathfrak{q}\subset\mathfrak{g} be a parabolic subalgebra with Levi factor 𝔩\mathfrak{l} and nilradical 𝔫\mathfrak{n}. We may define 1Z(𝔩),𝔮\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}} to be the subset of λ1Z(𝔩)reg\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})_{\text{reg}}^{*} such that for all αΔ(𝔫,𝔧)\alpha\in\Delta(\mathfrak{n},\mathfrak{j}), either

Imλ,α>0\operatorname{Im}\langle\lambda,\alpha^{\vee}\rangle>0

or

Imλ,α=0andReλ,α>0.\operatorname{Im}\langle\lambda,\alpha^{\vee}\rangle=0\ \text{and}\ \operatorname{Re}\langle\lambda,\alpha^{\vee}\rangle>0.

As noted in [HO20], in this case, 𝔮\mathfrak{q} defines a maximally real, admissible polarization of the coadjoint orbit 𝒪λ:=Gλ\mathcal{O}_{\lambda}:=G_{\mathbb{R}}\cdot\lambda. Although this assignment of 𝔮\mathfrak{q} to λ\lambda is not canonical, it is convenient for our argument to make such an assignment.

Since there are finitely many parabolic subalgebras 𝔮𝔤\mathfrak{q}\subset\mathfrak{g} with Levi factor 𝔩\mathfrak{l}, we have a finite disjoint union

1Z(𝔩)reg=𝔮𝔤1Z(𝔩),𝔮,1Z(𝔩)gr=𝔮𝔤1Z(𝔩)gr,𝔮,\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\text{reg}}=\bigsqcup_{\mathfrak{q}\subset\mathfrak{g}}\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}},\quad\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\text{gr}}=\bigsqcup_{\mathfrak{q}\subset\mathfrak{g}}\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}_{\text{gr}},

where 1Z(𝔩)gr,𝔮:=1Z(𝔩),𝔮1Z(𝔩)gr\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}_{\text{gr}}:=\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\text{gr}}.

Next, let G^(𝔩,𝔮)\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})} denote the collection of representations π(𝔩,Γλ)\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda}) such that λ1Z(𝔩)gr,𝔮\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}_{\rm gr}. Equivalently, G^(𝔩,𝔮)\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})} consists of π(𝒪,Γ)\pi(\mathcal{O},\Gamma) such that 𝒪1Z(𝔩)gr,𝔮\mathcal{O}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}_{\rm gr}\neq\emptyset. Therefore, we have a finite union

(5.1) G^𝔩=𝔮𝔤G^(𝔩,𝔮).\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{\mathbb{R}}}=\bigcup_{\mathfrak{q}\subset\mathfrak{g}}\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}.

Note that the right hand side of (5.1) may not be disjoint. In the same way as above, we can show that G^(𝔩,𝔮)\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})} is a locally closed subset of G^\widehat{G}_{\mathbb{R}}.

The set G^(𝔩,𝔮)\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})} can be identified with the collection of Γλ\Gamma_{\lambda} with λ1Z(𝔩)gr,𝔮\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}_{\rm gr}. To see this, suppose that π(𝔩,Γλ)π(𝔩,Γλ)\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\simeq\pi(\mathfrak{l}_{\mathbb{R}},\Gamma^{\prime}_{\lambda^{\prime}}) for λ,λ1Z(𝔩)gr,𝔮\lambda,\lambda^{\prime}\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}_{\rm gr}, Then by comparing the infinitesimal characters, λ+ρ𝔩\lambda+\rho_{\mathfrak{l}} and λ+ρ𝔩\lambda^{\prime}+\rho_{\mathfrak{l}} lie in the same Weyl group orbit. By our assumption, λ+ρ𝔩\lambda+\rho_{\mathfrak{l}} and λ+ρ𝔩\lambda^{\prime}+\rho_{\mathfrak{l}} satisfy the same dominance condition imposed by 𝔮\mathfrak{q} and hence λ=λ\lambda=\lambda^{\prime}. In view of the Langlands parameters of two representations (see the discussion at the end of Section 2), we have Γ=Γ\Gamma=\Gamma^{\prime}. Therefore, G^(𝔩,𝔮)\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})} is identified with the set of Γλ\Gamma_{\lambda}, or equivalently, the map

{(𝒪,Γ) : a semisimple orbital parameter𝒪1Z(𝔩)gr,𝔮}G^(𝔩,𝔮)\bigl{\{}\text{$(\mathcal{O},\Gamma)$ : a semisimple orbital parameter}\mid\mathcal{O}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}_{\rm gr}\neq\emptyset\bigr{\}}\to\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}

given by (𝒪,Γ)π(𝒪,Γ)(\mathcal{O},\Gamma)\mapsto\pi(\mathcal{O},\Gamma) is bijective.

By writing the measure mΠm_{\Pi} as a finite sum of measures supported on G^(𝔩,𝔮)\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})} for various 𝔮\mathfrak{q}, it is enough to prove Lemma 5.2 when mΠm_{\Pi} is a measure on G^(𝔩,𝔮)\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})} for one parabolic subalgebra 𝔮\mathfrak{q}. We thus fix 𝔮\mathfrak{q} and suppose Π\Pi is a direct integral of representations in G^(𝔩,𝔮)\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})} in the rest of this section.

Next, we need to define what it means for a measure mm on G^(𝔩,𝔮)\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})} to be of at most polynomial growth. Observe that we have a finite to one map

p:G^(𝔩,𝔮)π(𝔩,Γλ)λ1Z(𝔩)gr,𝔮.\displaystyle p\colon\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}\ni\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\mapsto\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}_{\rm gr}.

For a Borel measure mm on G^(𝔩,𝔮)\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}, let pmp_{*}m denote the pushforward of mm under the above map. Fix a norm |||\cdot| on 1Z(𝔩)\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}. We say that mm is of at most polynomial growth if there exist a constant M0>0M_{0}>0 and a finite measure mfm_{f} on 1Z(𝔩)gr,𝔮\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}_{\rm gr} such that

(5.2) pm(1+|λ|2)M0/2mf.p_{*}m\leq(1+|\lambda|^{2})^{M_{0}/2}m_{f}.

Here, mmm\leq m^{\prime} for measures mm and mm^{\prime} means that m(E)m(E)m(E)\leq m^{\prime}(E) for all measurable sets EE.

Our proof of Lemma 5.2 involves the Harish-Chandra distribution character of π(𝒪,Γ)\pi(\mathcal{O},\Gamma). Let Θ(𝒪,Γ)\Theta(\mathcal{O},\Gamma) denote the Harish-Chandra character of the representation π(𝒪,Γ)\pi(\mathcal{O},\Gamma). Define the analytic function jGj_{G_{\mathbb{R}}} utilizing the relation

exp(dg)=jG(X)dX\exp^{*}(dg)=j_{G_{\mathbb{R}}}(X)dX

where dgdg denotes a nonzero GG_{\mathbb{R}}-invariant density on GG_{\mathbb{R}} and dXdX denotes a nonzero translation invariant density on 𝔤\mathfrak{g}_{\mathbb{R}}. Normalize dgdg and dXdX so that jG(0)=1j_{G_{\mathbb{R}}}(0)=1, and let jG1/2j_{G_{\mathbb{R}}}^{1/2} be the unique analytic square root of jGj_{G_{\mathbb{R}}} with jG1/2(0)=1j_{G_{\mathbb{R}}}^{1/2}(0)=1. Since Θ(𝒪,Γ)\Theta(\mathcal{O},\Gamma) is an analytic function on the subset of regular, semisimple elements in GG_{\mathbb{R}}, we may define

θ(𝒪,Γ):=jG1/2(X)expΘ(𝒪,Γ)\theta(\mathcal{O},\Gamma):=j_{G_{\mathbb{R}}}^{1/2}(X)\cdot\exp^{*}\Theta(\mathcal{O},\Gamma)

to be the Lie algebra analogue of the character of π(𝒪,Γ)\pi(\mathcal{O},\Gamma). Note θ(𝒪,Γ)\theta(\mathcal{O},\Gamma) is an analytic function on the collection of regular, semisimple elements in 𝔤\mathfrak{g}_{\mathbb{R}}.

Fix a choice of positive roots Δ+(𝔩,𝔧)Δ(𝔩,𝔧)\Delta^{+}(\mathfrak{l},\mathfrak{j})\subset\Delta(\mathfrak{l},\mathfrak{j}), and define ρ𝔩:=12αΔ+(𝔩,𝔧)α\rho_{\mathfrak{l}}:=\frac{1}{2}\sum_{\alpha\in\Delta^{+}(\mathfrak{l},\mathfrak{j})}\alpha. Given a semisimple orbital parameter (𝒪,Γ)(\mathcal{O},\Gamma) with λ𝒪1Z(𝔩)gr,𝔮\lambda\in\mathcal{O}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}_{\rm gr}\neq\emptyset, we define a contour

𝒞(𝒪,𝔮):={gλ+uρ𝔩gG,uU,Ad(g)𝔮=Ad(u)𝔮}\mathcal{C}(\mathcal{O},\mathfrak{q}):=\left\{g\cdot\lambda+u\cdot\rho_{\mathfrak{l}}\mid g\in G_{\mathbb{R}},\ u\in U,\ \operatorname{Ad}(g)\cdot\mathfrak{q}=\operatorname{Ad}(u)\cdot\mathfrak{q}\right\}

in 𝔤\mathfrak{g}^{*}. Here, σc\sigma_{c} is an anti-holomorphic involution on GG which commutes with σ\sigma such that U:=GσcU:=G^{\sigma_{c}} is a compact real form of GG. For a coadjoint GG-orbit Ω𝔤\Omega\subset\mathfrak{g}^{*}, the Kirillov-Kostant-Souriau GG-invariant, holomorphic 22-form ω\omega on Ω\Omega is defined by

ωξ(ad(X)(ξ),ad(Y)(ξ)):=ξ([X,Y]).\omega_{\xi}(\operatorname{ad}^{*}(X)(\xi),\operatorname{ad}^{*}(Y)(\xi)):=\xi([X,Y]).

Suppose that Ω\Omega is the regular, coadjoint GG-orbit through ξ=λ+ρ𝔩𝔧𝔤\xi=\lambda+\rho_{\mathfrak{l}}\in\mathfrak{j}^{*}\subset\mathfrak{g}^{*} and put n:=12dimΩn:=\frac{1}{2}\dim_{\mathbb{C}}\Omega. Then 𝒞(𝒪,𝔮)\mathcal{C}(\mathcal{O},\mathfrak{q}) is a real 2n2n-dimensional closed submanifold of Ω\Omega (see [HO20]). Define the 2n2n-form

ν:=ωn(2π1)nn!.\nu:=\frac{\omega^{\wedge n}}{(2\pi\sqrt{-1})^{n}n!}.

For a function φ\varphi on 𝔤\mathfrak{g}_{\mathbb{R}}, we define the (inverse) Fourier transform as the following functions on 1𝔤\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*}:

φ^(η):=𝔤eη,Xφ(X)𝑑X,φˇ(η):=𝔤eη,Xφ(X)𝑑X.\displaystyle\hat{\varphi}(\eta):=\int_{\mathfrak{g}_{\mathbb{R}}}e^{-\langle\eta,X\rangle}\varphi(X)dX,\quad\check{\varphi}(\eta):=\int_{\mathfrak{g}_{\mathbb{R}}}e^{\langle\eta,X\rangle}\varphi(X)dX.

The main result of [HO20] is

(5.3) θ(𝒪,Γ),φ=𝒞(𝒪,𝔮)φˇν,\langle\theta(\mathcal{O},\Gamma),\varphi\rangle=\int_{\mathcal{C}(\mathcal{O},\mathfrak{q})}\check{\varphi}\,\nu,

where φCc(𝔤)\varphi\in C_{c}^{\infty}(\mathfrak{g}_{\mathbb{R}}) is a smooth, compactly supported function on 𝔤\mathfrak{g}_{\mathbb{R}}. Observe that φˇ\check{\varphi} extends to a holomorphic function on 𝔤\mathfrak{g}^{*}. We remark that for any semisimple orbit 𝒪\mathcal{O} with 𝒪1Z(𝔩)gr\mathcal{O}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm gr}\neq\emptyset, the contour 𝒞(𝒪,𝔮)\mathcal{C}(\mathcal{O},\mathfrak{q}) and the forms ω,ν\omega,\nu are defined in the same way, even if it does not come from a semisimple orbital parameter (𝒪,Γ)(\mathcal{O},\Gamma).

Fix a KK_{\mathbb{R}}-invariant norm |||\cdot| on 𝔤:=Hom(𝔤,)\mathfrak{g}_{\mathbb{R}}^{*}:=\operatorname{Hom}_{\mathbb{R}}(\mathfrak{g}_{\mathbb{R}},\mathbb{R}). If η𝔤:=Hom(𝔤,)\eta\in\mathfrak{g}^{*}:=\operatorname{Hom}_{\mathbb{C}}(\mathfrak{g},\mathbb{C}), write

η=Reη+1Imη\eta=\mathrm{Re}\,\eta+\sqrt{-1}\;\!\mathrm{Im}\,\eta

where Reη,Imη𝔤\mathrm{Re}\,\eta,\,\mathrm{Im}\,\eta\in\mathfrak{g}_{\mathbb{R}}^{*}. Extend |||\cdot| to a norm on 𝔤\mathfrak{g}^{*} by defining |η|2=|Reη|2+|Imη|2|\eta|^{2}=|\mathrm{Re}\,\eta|^{2}+|\mathrm{Im}\,\eta|^{2}.

Fix dd\in\mathbb{R} such that d>maxαΔ(𝔤,𝔧)|ρ𝔩,α|d>\max_{\alpha\in\Delta(\mathfrak{g},\mathfrak{j})}|\langle\rho_{\mathfrak{l}},\alpha^{\vee}\rangle|. Writing mΠm_{\Pi} as a sum of two measures according to the decomposition

G^(𝔩,𝔮)=(G^(𝔩,𝔮)G^(𝔩,d))(G^(𝔩,𝔮)G^(𝔩,d))\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}=\bigl{(}\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}\cap\widehat{G}_{\mathbb{R}}(\mathfrak{l},d)\bigr{)}\cup\bigl{(}\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}\setminus\widehat{G}_{\mathbb{R}}(\mathfrak{l},d)\bigr{)}

and using Lemma 4.4, it is enough to show Lemma 5.2 when suppmΠG^(𝔩,d)=\operatorname{supp}m_{\Pi}\cap\widehat{G}_{\mathbb{R}}(\mathfrak{l},d)=\emptyset. This assumption makes it easier for us to estimate the integral (5.3) as we see below.

Lemma 5.3.

Suppose that mm is a measure on G^(𝔩,𝔮)\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})} with at most polynomial growth and suppmG^(𝔩,d)=\operatorname{supp}m\cap\widehat{G}_{\mathbb{R}}(\mathfrak{l},d)=\emptyset.

  1. (i)

    Let α\alpha be a function on 𝔤\mathfrak{g}^{*}, and assume α|𝒞(𝒪,𝔮)\alpha|_{\mathcal{C}(\mathcal{O},\mathfrak{q})} is measurable for all coadjoint orbits 𝒪\mathcal{O} with 𝒪1Z(𝔩)gr,𝔮\mathcal{O}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})_{\rm gr}^{*,\mathfrak{q}}\neq\emptyset. Assume that for every NN\in\mathbb{N} and every b>0b>0 there exist constants CN,b>0C_{N,b}>0 such that

    (5.4) supη𝔤|Reη|b(1+|Imη|2)N/2|α(η)|CN,b.\sup_{\begin{subarray}{c}\eta\in\mathfrak{g}^{*}\\ |\mathrm{Re}\,\eta|\leq b\end{subarray}}(1+|\mathrm{Im}\,\eta|^{2})^{N/2}|\alpha(\eta)|\leq C_{N,b}.

    Then the integral

    (5.5) C(m),α:=π(𝒪,Γ)G^(𝔩,𝔮)(𝒞(𝒪,𝔮)α(η)ν)𝑑m\langle C(m),\alpha\rangle:=\int_{\pi(\mathcal{O},\Gamma)\in\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}}\left(\int_{\mathcal{C}(\mathcal{O},\mathfrak{q})}\alpha(\eta)\nu\right)dm

    converges absolutely.

  2. (ii)

    If φCc(𝔤)\varphi\in C_{c}^{\infty}(\mathfrak{g}_{\mathbb{R}}), then the integral

    (5.6) C(m),φˇ:=π(𝒪,Γ)G^(𝔩,𝔮)(𝒞(𝒪,𝔮)φˇν)𝑑m\langle C(m),\check{\varphi}\rangle:=\int_{\pi(\mathcal{O},\Gamma)\in\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}}\left(\int_{\mathcal{C}(\mathcal{O},\mathfrak{q})}\check{\varphi}\,\nu\right)dm

    converges absolutely. The functional φC(m),φˇ\varphi\mapsto\langle C(m),\check{\varphi}\rangle is a well-defined distribution on 𝔤\mathfrak{g}_{\mathbb{R}}, which is the integral

    θ(m):=π(𝒪,Γ)G^(𝔩,𝔮)θ(𝒪,Γ)𝑑m.\theta(m):=\int_{\pi(\mathcal{O},\Gamma)\in\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}}\theta(\mathcal{O},\Gamma)dm.
  3. (iii)

    For φCc(𝔤)\varphi\in C_{c}^{\infty}(\mathfrak{g}_{\mathbb{R}}), the Fourier transform of θ(m)φ\theta(m)\varphi is given by

    (θ(m)φ)^(ξ)=C(m)η,φˇ(ηξ).(\theta(m)\varphi)^{^}\,(\xi)=\langle C(m)_{\eta},\check{\varphi}(\eta-\xi)\rangle.

    It is a smooth, polynomially bounded function on 1𝔤\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*}.

  4. (iv)

    We have

    (5.7) SS0(θ(m))AC(π(𝒪,Γ)suppm𝒪).\operatorname{SS}_{0}\bigl{(}\theta(m)\bigr{)}\subset\operatorname{AC}\Biggl{(}\bigcup_{\pi(\mathcal{O},\Gamma)\in\operatorname{supp}m}\mathcal{O}\Biggr{)}.

To prove part (i), we need another lemma. Define

Λ:={λ1Z(𝔩):|λ+ρ𝔩,α|d(αΔ(𝔤,𝔧)Δ(𝔩,𝔧))}.\Lambda:=\bigl{\{}\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}:|\langle\lambda+\rho_{\mathfrak{l}},\alpha^{\vee}\rangle|\geq d\ \bigl{(}\forall\alpha\in\Delta(\mathfrak{g},\mathfrak{j})\setminus\Delta(\mathfrak{l},\mathfrak{j})\bigr{)}\bigr{\}}.

Then Λ\Lambda is a closed subset of 1Z(𝔩)gr\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm gr}.

Lemma 5.4.

For any M>0M>0, there exist constants kM,CM>0k_{M},C_{M}>0 such that

η𝒞(𝒪λ,𝔮)(1+|Imη|2)kM/2|ν|CM(1+|λ|2)M/2\int_{\eta\in\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q})}(1+|\mathrm{Im}\,\eta|^{2})^{-k_{M}/2}|\nu|\leq C_{M}(1+|\lambda|^{2})^{-M/2}

for λΛ\lambda\in\Lambda. Here, we write 𝒪λ:=Gλ\mathcal{O}_{\lambda}:=G_{\mathbb{R}}\cdot\lambda.

proof of Lemma 5.4.

Fix any λ0Λ\lambda_{0}\in\Lambda. The Euclidean metric on 𝔤\mathfrak{g}^{*} induces a Riemannian metric on the submanifold 𝒞(𝒪λ0,𝔮)\mathcal{C}(\mathcal{O}_{\lambda_{0}},\mathfrak{q}). Let νE\nu_{E} be the volume form of this Riemannian manifold 𝒞(𝒪λ0,𝔮)\mathcal{C}(\mathcal{O}_{\lambda_{0}},\mathfrak{q}).

We first claim that

(5.8) ξ𝒞(𝒪λ0,𝔮)(1+|ξ|2)N/2νE<\displaystyle\int_{\xi\in\mathcal{C}(\mathcal{O}_{\lambda_{0}},\mathfrak{q})}(1+|\xi|^{2})^{-N/2}\nu_{E}<\infty

for sufficiently large N>0N>0. To see this, we use an argument similar to [SV98, (3.14)]. Consider the one point compactification of 𝔤\mathfrak{g}, which is a sphere SS. Let 𝒞(𝒪λ0,𝔮)¯\overline{\mathcal{C}(\mathcal{O}_{\lambda_{0}},\mathfrak{q})} be the closure of 𝒞(𝒪λ0,𝔮)\mathcal{C}(\mathcal{O}_{\lambda_{0}},\mathfrak{q}) in SS. With respect to a standard metric on the sphere SS, its compact semialgebraic subset 𝒞(𝒪λ0,𝔮)¯\overline{\mathcal{C}(\mathcal{O}_{\lambda_{0}},\mathfrak{q})} has finite volume (see e.g. [OS17]). By comparing the standard metric on SS and the Euclidean metric on 𝔤\mathfrak{g}, this can be restated as (5.8) for N4nN\geq 4n.

Next, define a semialgebraic set

𝒞(Λ):={(λ,η)Λ×𝔤:η𝒞(𝒪λ,𝔮)}.\mathcal{C}(\Lambda):=\{(\lambda,\eta)\in\Lambda\times\mathfrak{g}^{*}:\eta\in\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q})\}.

For each λΛ\lambda\in\Lambda, there is an isomorphism

i:𝒞(𝒪λ,𝔮)𝒞(𝒪λ0,𝔮),gλ+uρ𝔩gλ0+uρ𝔩.i\colon\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q})\xrightarrow{\sim}\mathcal{C}(\mathcal{O}_{\lambda_{0}},\mathfrak{q}),\quad g\cdot\lambda+u\cdot\rho_{\mathfrak{l}}\mapsto g\cdot\lambda_{0}+u\cdot\rho_{\mathfrak{l}}.

Define the semialgebraic functions f((λ,η)):=1+|Imη|2f((\lambda,\eta)):=1+|\mathrm{Im}\,\eta|^{2} on 𝒞(Λ)\mathcal{C}(\Lambda). In the following, we will compare some other semialgebraic functions on 𝒞(Λ)\mathcal{C}(\Lambda) with ff. On 𝒞(𝒪λ,𝔮)\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q}), we have two volume forms iνEi^{*}\nu_{E} and |ν||\nu|. Define h:=|ν|iνEh:=\frac{|\nu|}{i^{*}\nu_{E}}, which is a semialgebraic function on 𝒞(Λ)\mathcal{C}(\Lambda). It is easy to see that the set {(λ,η):|f(λ,η)|t}\{(\lambda,\eta):|f(\lambda,\eta)|\leq t\} is compact for any t>0t>0. Then the function

h¯(t)=sup{h((λ,η)):|f(λ,η)|t}\overline{h}(t)=\sup\{h((\lambda,\eta)):|f(\lambda,\eta)|\leq t\}

is defined for large t>0t>0 and is semialgebraic. By [Hör83b, Theorem A.2.5], h¯(t)A1tN1\overline{h}(t)\leq A_{1}\cdot t^{N_{1}} for some constants A1,N1>0A_{1},N_{1}>0. Hence we get

(5.9) h((λ,η))A1(1+|Imη|2)N1\displaystyle h((\lambda,\eta))\leq A_{1}\cdot(1+|\mathrm{Im}\,\eta|^{2})^{N_{1}}

for (λ,η)𝒞(Λ)(\lambda,\eta)\in\mathcal{C}(\Lambda).

The functions (λ,η)1+|i(η)|2(\lambda,\eta)\mapsto 1+|i(\eta)|^{2} and (λ,η)1+|λ|2(\lambda,\eta)\mapsto 1+|\lambda|^{2} are also semialgebraic on 𝒞(Λ)\mathcal{C}(\Lambda). Hence we similarly have

(5.10) 1+|i(η)|2A2(1+|Imη|2)N2,1+|λ|2A3(1+|Imη|2)N3\displaystyle 1+|i(\eta)|^{2}\leq A_{2}\cdot(1+|\mathrm{Im}\,\eta|^{2})^{N_{2}},\quad 1+|\lambda|^{2}\leq A_{3}\cdot(1+|\mathrm{Im}\,\eta|^{2})^{N_{3}}

for some constants A2,N2,A3,N3>0A_{2},N_{2},A_{3},N_{3}>0. The lemma follows from an isomorphism i:𝒞(𝒪λ0,𝔮)𝒞(𝒪λ,𝔮)i\colon\mathcal{C}(\mathcal{O}_{\lambda_{0}},\mathfrak{q})\simeq\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q}) and the estimates (5.8), (5.9) and (5.10). ∎

proof of Lemma 5.3.

Since mm is of at most polynomial growth, pm(1+|λ|2)M0/2mfp_{*}m\leq(1+|\lambda|^{2})^{M_{0}/2}m_{f} for a finite measure mfm_{f} and a constant M0>0M_{0}>0. By our assumption on mm, we may assume that suppmf\operatorname{supp}m_{f} is contained in Λ\Lambda. In addition, |Reη||\mathrm{Re}\,\eta| for η𝒞(𝒪,𝔮)\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q}) is bounded by a constant. Hence the absolute convergence of (5.5) follows from Lemma 5.4 and (5.4).

To prove part (ii), recall that for φCc(𝔤)\varphi\in C_{c}^{\infty}(\mathfrak{g}_{\mathbb{R}}), the Paley-Wiener Theorem assures us that there exists a constant B>0B>0 and for every NN\in\mathbb{N}, there exists a constant AN>0A_{N}>0 such that

|φˇ(η)|ANeB|Reη|(1+|Imη|2)N/2.\left|\check{\varphi}(\eta)\right|\leq\frac{A_{N}e^{B|\mathrm{Re}\,\eta|}}{(1+|\mathrm{Im}\,\eta|^{2})^{N/2}}.

Hence, we may plug in φˇ\check{\varphi} for α\alpha and the absolute convergence of (5.6) follows from part (i). Further, the constants that bound this integral can be shown to be bounded by seminorms on the space of smooth compactly supported densities on 𝔤\mathfrak{g}_{\mathbb{R}}. Therefore, the integral the θ(m)\theta(m) defined in part (ii) is given as a well-defined distribution

φC(m),φˇ.\varphi\mapsto\langle C(m),\check{\varphi}\rangle.

By (5.3), this is the integral of θ(𝒪,Γ)\theta(\mathcal{O},\Gamma).

Next, we prove part (iii). Let φCc(𝔤)\varphi\in C_{c}^{\infty}(\mathfrak{g}_{\mathbb{R}}). Then θ(m)φ\theta(m)\varphi is a distribution with compact support. Hence the Fourier transform (θ(m)φ)^(\theta(m)\varphi)^{^} is a smooth, polynomially bounded function on 1𝔤\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*}. The value of (θ(m)φ)^(\theta(m)\varphi)^{^} at ξ1𝔤\xi\in\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*} is given as

(θ(m)φ)^(ξ)=θ(m)φ,eξ,=θ(m),eξ,φ\displaystyle(\theta(m)\varphi)^{^}\,(\xi)=\langle\theta(m)\varphi,e^{-\langle\xi,\cdot\rangle}\rangle=\langle\theta(m),e^{-\langle\xi,\cdot\rangle}\varphi\rangle =C(m),(eξ,φ)\displaystyle=\langle C(m),(e^{-\langle\xi,\cdot\rangle}\varphi)^{\vee}\rangle
=C(m)η,φˇ(ηξ).\displaystyle=\langle C(m)_{\eta},\check{\varphi}(\eta-\xi)\rangle.

Thus, (iii) is proved.

For part (iv), we require some additional notation. Choose a basis {X1,,Xn}\{X_{1},\ldots,X_{n}\} of 𝔤\mathfrak{g}, and define the differential operator

Dα:=X1α1XnαnD^{\alpha}:=\partial_{X_{1}}^{\alpha_{1}}\cdots\partial_{X_{n}}^{\alpha_{n}}

for every multi-index α=(α1,,αN)n\alpha=(\alpha_{1},\ldots,\alpha_{N})\in\mathbb{N}^{n}. In addition, define |α|=α1++αn|\alpha|=\alpha_{1}+\cdots+\alpha_{n}. If 0𝒰1𝒰2𝔤0\in\mathcal{U}_{1}\subset\mathcal{U}_{2}\subset\mathfrak{g} are precompact, open subsets of 𝔤\mathfrak{g} with 𝒰1¯𝒰2\overline{\mathcal{U}_{1}}\subset\mathcal{U}_{2}, then there exists a sequence {φN,𝒰1,𝒰2}\{\varphi_{N,\mathcal{U}_{1},\mathcal{U}_{2}}\} of functions indexed by NN\in\mathbb{N} and satisfying the following properties (see pages 25–26, 282 of [Hör83a]):

  1. (1)

    φN,𝒰1,𝒰2Cc(𝒰2)\varphi_{N,\mathcal{U}_{1},\mathcal{U}_{2}}\in C_{c}^{\infty}(\mathcal{U}_{2}) for all NN\in\mathbb{N}

  2. (2)

    φN,𝒰1,𝒰2(x)=1\varphi_{N,\mathcal{U}_{1},\mathcal{U}_{2}}(x)=1 if x𝒰1x\in\mathcal{U}_{1}

  3. (3)

    There exists a constant Cα>0C_{\alpha}>0 for every multi-index αn\alpha\in\mathbb{N}^{n} such that

    supx𝒰2|(Dα+βφN,𝒰1,𝒰2)(x)|Cα|β|+1(N+1)|β|\sup_{x\in\mathcal{U}_{2}}|(D^{\alpha+\beta}\varphi_{N,\mathcal{U}_{1},\mathcal{U}_{2}})(x)|\leq C_{\alpha}^{|\beta|+1}(N+1)^{|\beta|}

    for every multi-index βn\beta\in\mathbb{N}^{n} with |β|N|\beta|\leq N.

For the sequel, we fix 𝒰1\mathcal{U}_{1}, 𝒰2\mathcal{U}_{2}, and take a sequence of functions {φN,𝒰1,𝒰2}\{\varphi_{N,\mathcal{U}_{1},\mathcal{U}_{2}}\} satisfying (1)–(3). Write φN:=φN,𝒰1,𝒰2\varphi_{N}:=\varphi_{N,\mathcal{U}_{1},\mathcal{U}_{2}}.

Fix

ξAC(π(𝒪,Γ)suppm𝒪).\xi\notin\operatorname{AC}\left(\bigcup_{\pi(\mathcal{O},\Gamma)\in\operatorname{supp}m}\mathcal{O}\right).

In order to prove (5.7), it is enough to show the following by [Hör83a, §8.4]: there exists an open subset ξW1𝔤\xi\in W\subset\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*} and a constant C>0C>0 such that

(5.11) |C(m)η,φˇN(ηtξ)|CN+1(N+1)N(1+t2)N/2\left|\langle C(m)_{\eta},\,\check{\varphi}_{N}(\eta-t\xi^{\prime})\rangle\right|\leq C^{N+1}\frac{(N+1)^{N}}{(1+t^{2})^{N/2}}

for all ξW\xi^{\prime}\in W and t>0t>0.

Choose an open cone Ψ1𝔤\Psi\subset\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*} such that

ξΨΨ¯{0}1𝔤AC(π(𝒪,Γ)suppμ𝒪),\xi\in\Psi\subset\overline{\Psi}\setminus\{0\}\subset\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*}\setminus\operatorname{AC}\left(\bigcup_{\pi(\mathcal{O},\Gamma)\in\operatorname{supp}\mu}\mathcal{O}\right),

and define

W:={ξΨ||ξ|2<|ξ|<2|ξ|}.W:=\left\{\xi^{\prime}\in\Psi\,\Bigl{|}\,\frac{|\xi|}{2}<|\xi^{\prime}|<2|\xi|\right\}.

We require a lemma.

Lemma 5.5.

There exist constants D,ϵ,ϵ>0D,\epsilon,\epsilon^{\prime}>0 such that

(5.12) |1Imηtξ|ϵt|\sqrt{-1}\mathrm{Im}\,\eta-t\xi^{\prime}|\geq\epsilon t

and

(5.13) |1Imηtξ|ϵ|Imη||\sqrt{-1}\mathrm{Im}\,\eta-t\xi^{\prime}|\geq\epsilon^{\prime}|\mathrm{Im}\,\eta|

if

ηπ(𝒪,Γ)suppm𝒞(𝒪,𝔮),ξW¯,andt>D.\eta\in\bigcup_{\pi(\mathcal{O},\Gamma)\in\operatorname{supp}m}\mathcal{C}(\mathcal{O},\mathfrak{q}),\ \ \ \xi^{\prime}\in\overline{W},\ \ \text{and}\ \ t>D.
proof of Lemma 5.5.

Assume that (5.12) does not hold. Then we may find sequences {ξj}W¯\{\xi_{j}\}\subset\overline{W}, {tj}>0\{t_{j}\}\subset\mathbb{R}_{>0}, and {ηj}\{\eta_{j}\} with ηj𝒞(𝒪j,𝔮)\eta_{j}\in\mathcal{C}(\mathcal{O}_{j},\mathfrak{q}) satisfying π(𝒪j,Γj)suppm\pi(\mathcal{O}_{j},\Gamma_{j})\in\operatorname{supp}m such that |tjξj1Imηj|<tjj|t_{j}\xi_{j}-\sqrt{-1}\mathrm{Im}\,\eta_{j}|<\frac{t_{j}}{j} and tj>jt_{j}>j. Further, we may write ηj=ηj+ηj′′\eta_{j}=\eta_{j}^{\prime}+\eta_{j}^{\prime\prime} where ηj𝒪j\eta_{j}^{\prime}\in\mathcal{O}_{j} and ηj′′Uρ𝔩\eta_{j}^{\prime\prime}\in U\cdot\rho_{\mathfrak{l}}. Since Reηj\mathrm{Re}\,\eta_{j} and ηj′′\eta_{j}^{\prime\prime} are bounded, |tjξjηj|<tjj+a|t_{j}\xi_{j}-\eta_{j}^{\prime}|<\frac{t_{j}}{j}+a for a constant a>0a>0. But, then {ηj/tj}\{\eta_{j}^{\prime}/t_{j}\} has a convergent subsequence which must therefore lie in both W¯\overline{W} and AC(π(𝒪,Γ)suppm𝒪)\operatorname{AC}\left(\bigcup_{\pi(\mathcal{O},\Gamma)\in\operatorname{supp}m}\mathcal{O}\right), which is a contradiction. This implies (5.12).

Next, we utilize the triangle inequality to obtain

|tξ||Imη||1Imηtξ|.|t\xi^{\prime}|\geq|\mathrm{Im}\,\eta|-|\sqrt{-1}\mathrm{Im}\,\eta-t\xi^{\prime}|.

Combining with (5.12) yields

|1Imηtξ|ϵtϵ|ξ|(|Imη||1Imηtξ|).|\sqrt{-1}\mathrm{Im}\,\eta-t\xi^{\prime}|\geq\epsilon t\geq\frac{\epsilon}{|\xi^{\prime}|}\left(|\mathrm{Im}\,\eta|-|\sqrt{-1}\mathrm{Im}\,\eta-t\xi^{\prime}|\right).

Recall |ξ|2|ξ||\xi^{\prime}|\leq 2|\xi|, collect the |1Imηtξ||\sqrt{-1}\mathrm{Im}\,\eta-t\xi^{\prime}| terms on one side of the equation, and put ϵ:=(1+ϵ2|ξ|)1ϵ2|ξ|\epsilon^{\prime}:=(1+\frac{\epsilon}{2|\xi|})^{-1}\frac{\epsilon}{2|\xi|}. Then (5.13) follows. ∎

In order to prove (5.11), for each M>0M>0, we will first show the existence of a constant CM>0C_{M}>0 such that

(5.14) |η𝒞(𝒪λ,𝔮)φˇN(ηtξ)ν|CMN+1(1+|λ|2)M/2(N+1)N(1+t2)N/2\left|\int_{\eta\in\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q})}\check{\varphi}_{N}(\eta-t\xi^{\prime})\,\nu\right|\leq\frac{C_{M}^{N+1}}{(1+|\lambda|^{2})^{M/2}}\frac{(N+1)^{N}}{(1+t^{2})^{N/2}}

for all ξW\xi^{\prime}\in W, for all π(𝒪,Γ)suppm\pi(\mathcal{O},\Gamma)\in\operatorname{supp}m, for all NN\in\mathbb{N}, and for t>0t>0. In order to prove (5.14), we need an estimate of φˇN\check{\varphi}_{N}. By the proof of the Paley-Wiener Theorem (see for instance page 181 of [Hör83a]) and part (3) of the definition of {φN,𝒰1,𝒰2}\{\varphi_{N,\mathcal{U}_{1},\mathcal{U}_{2}}\}, there exist constants B,C>0B,C^{\prime}>0 such that

|φˇN(η)|(C)N+1(N+1)NeB|Reη|(1+|Imη|2)N/2.|\check{\varphi}_{N}(\eta)|\leq\frac{(C^{\prime})^{N+1}(N+1)^{N}e^{B\cdot|\mathrm{Re}\,{\eta}|}}{(1+|\mathrm{Im}\,\eta|^{2})^{N/2}}.

Using that |Reη||\mathrm{Re}\,\eta| is bounded by a constant a>0a>0 for η𝒞(𝒪,𝔮)\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q}), and putting C:=CeBaC:=C^{\prime}e^{B\cdot a}, we deduce

(5.15) |φˇN(ηtξ)|CN+1(N+1)N(1+|1Imηtξ|2)N/2|\check{\varphi}_{N}(\eta-t\xi^{\prime})|\leq\frac{C^{N+1}(N+1)^{N}}{(1+|\sqrt{-1}\mathrm{Im}\,\eta-t\xi^{\prime}|^{2})^{N/2}}

whenever ξ1𝔤\xi^{\prime}\in\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*} and η𝒞(𝒪λ,𝔮)\eta\in\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q}) with λ1Z(𝔩)gr,𝔮\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}_{\rm gr}. For fixed MM, define

φNM:=φN+kM,\varphi^{M}_{N}:=\varphi_{N+k_{M}},

where kMk_{M} is the constant in Lemma 5.4. Observe that for every MM, the sequence φNM\varphi^{M}_{N} still satisfies the properties (1)–(3). Therefore, in order to verify part (iv), we may replace φN\varphi_{N} with φNM\varphi^{M}_{N}. Utilizing Lemma 5.4 and (5.15), we obtain for all M,NM,N\in\mathbb{N} and π(𝒪λ,Γ)suppm\pi(\mathcal{O}_{\lambda},\Gamma)\in\operatorname{supp}m,

|η𝒞(𝒪λ,𝔮)(φNM)(ηtξ)ν|\displaystyle\phantom{=}\left|\int_{\eta\in\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q})}(\varphi^{M}_{N})^{\vee}(\eta-t\xi^{\prime})\,\nu\right|
CM(1+|λ|2)M/2supη𝒞(𝒪λ,𝔮)(1+|Imη|2)kM/2|(φNM)(ηtξ)|\displaystyle\leq\frac{C_{M}}{(1+|\lambda|^{2})^{M/2}}\sup_{\eta\in\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q})}(1+|\mathrm{Im}\,\eta|^{2})^{k_{M}/2}|(\varphi^{M}_{N})^{\vee}(\eta-t\xi^{\prime})|
(5.16) CM(1+|λ|2)M/2supη𝒞(𝒪λ,𝔮)CN+kM+1(N+kM+1)N+kM(1+|1Imηtξ|2)(N+kM)/2(1+|Imη|2)kM/2\displaystyle\leq\frac{C_{M}}{(1+|\lambda|^{2})^{M/2}}\sup_{\eta\in\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q})}\frac{C^{N+k_{M}+1}(N+k_{M}+1)^{N+k_{M}}}{(1+|\sqrt{-1}\mathrm{Im}\,\eta-t\xi^{\prime}|^{2})^{(N+k_{M})/2}}\cdot(1+|\mathrm{Im}\,\eta|^{2})^{k_{M}/2}

for some constant CM>0C_{M}>0. For fixed MM, if NN is sufficiently large, we have

(5.17) (N+kM+1)N+kM=(N+kM+1)N(N+kM+1)kM2N(N+1)NkMN+kM+1\begin{split}(N+k_{M}+1)^{N+k_{M}}&=(N+k_{M}+1)^{N}(N+k_{M}+1)^{k_{M}}\\ &\leq 2^{N}(N+1)^{N}k_{M}^{N+k_{M}+1}\end{split}

where we have used that ts>stt^{s}>s^{t} for s>t3s>t\geq 3. For every fixed MM\in\mathbb{N} and sufficiently large NN, we may utilize (5.12), (5.13) and (5.17) to bound (5) by

CMN+1(N+1)N(1+|λ|2)M/2(1+|Imη|2)kM/2(1+(ϵ|Imη|)2)kM/21(1+(ϵt)2)N/2\displaystyle\leq\frac{C_{M}^{N+1}(N+1)^{N}}{(1+|\lambda|^{2})^{M/2}}\cdot\frac{(1+|\mathrm{Im}\,\eta|^{2})^{k_{M}/2}}{(1+(\epsilon^{\prime}|\mathrm{Im}\,\eta|)^{2})^{k_{M}/2}}\cdot\frac{1}{(1+(\epsilon t)^{2})^{N/2}}
CMN+1(N+1)N(1+|λ|2)M/21(1+t2)N/2\displaystyle\leq\frac{C_{M}^{N+1}(N+1)^{N}}{(1+|\lambda|^{2})^{M/2}}\cdot\frac{1}{(1+t^{2})^{N/2}}

for the constant CM>0C_{M}>0 which we increased in each line. Thus, (5.14) is proved. Then

|π(𝒪,Γ)G^(𝔩,𝔮)η𝒞(𝒪,𝔮)(φNM)(ηtξ)ν𝑑m|\displaystyle\phantom{=}\left|\int_{\pi(\mathcal{O},\Gamma)\in\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}}\int_{\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q})}(\varphi_{N}^{M})^{\vee}(\eta-t\xi^{\prime})\,\nu dm\right|
CMN+1(N+1)N(1+t2)N/2π(𝒪λ,Γ)G^(𝔩,𝔮)1(1+|λ|2)M/2𝑑m\displaystyle\leq\frac{C_{M}^{N+1}(N+1)^{N}}{(1+t^{2})^{N/2}}\int_{\pi(\mathcal{O}_{\lambda},\Gamma)\in\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}}\frac{1}{(1+|\lambda|^{2})^{M/2}}dm
CMN+1(N+1)N(1+t2)N/21Z(𝔩)gr,𝔮1(1+|λ|2)M/2(1+|λ|2)M0/2𝑑mf\displaystyle\leq\frac{C_{M}^{N+1}(N+1)^{N}}{(1+t^{2})^{N/2}}\int_{\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}_{\rm gr}}\frac{1}{(1+|\lambda|^{2})^{M/2}}(1+|\lambda|^{2})^{M_{0}/2}dm_{f}
CMN+1(N+1)N(1+t2)N/21Z(𝔩)gr,𝔮1(1+|λ|2)(MM0)/2𝑑mf.\displaystyle\leq\frac{C_{M}^{N+1}(N+1)^{N}}{(1+t^{2})^{N/2}}\int_{\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}_{\rm gr}}\frac{1}{(1+|\lambda|^{2})^{(M-M_{0})/2}}dm_{f}.

where we have increased the constant CMC_{M} as necessary throughout the calculation. Since the final integral converges if MM0M\geq M_{0}, we may absorb the value of the integral into the constant CMC_{M} to bound the entire expression by

CMN+1(N+1)N(1+t2)N/2.\leq\frac{C_{M}^{N+1}(N+1)^{N}}{(1+t^{2})^{N/2}}.

Part (iv) follows. ∎

The proof of Lemma 5.2 now proceeds exactly line by line the same as the proof of [HHO16, Proposition 7.1] except one must substitute (5.7) in for (7.1) of [HHO16]. For this argument, we only need (5.7) for a finite measure mm. Lemma 5.3 was stated more generally for a measure with at most polynomial growth because it will be necessary in the next section.

6. Wave front sets of direct integrals for a Levi, part 2

We retain the notation of the previous section. The purpose of this section is to prove the following lemma using Lemma 6.6 and Lemma 6.8. The proof of these lemmas will be postponed in the next section.

Lemma 6.1.

In the setting of Theorem 5.1,

(6.1) WF(Π)AC(π(𝔩,Γλ)suppmΠGλ)(GZ(𝔩)reg).\operatorname{WF}(\Pi)\supset\operatorname{AC}\Biggl{(}\bigcup_{\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\in\operatorname{supp}m_{\Pi}}G_{\mathbb{R}}\cdot\lambda\Biggr{)}\cap(G\cdot Z(\mathfrak{l})^{*}_{\rm reg}).

Lemma 5.2 and Lemma 6.1 combine to imply Theorem 5.1 since WF(Π)SS(Π)\operatorname{WF}(\Pi)\subset\operatorname{SS}(\Pi) for any unitary representation Π\Pi of GG_{\mathbb{R}}.

We first show the following:

Lemma 6.2.

For any subset S1Z(𝔩)S\subset\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*},

(6.2) AC(GS)(GZ(𝔩)reg)=G(AC(S)1Z(𝔩)reg).\operatorname{AC}(G_{\mathbb{R}}\cdot S)\cap(G\cdot Z(\mathfrak{l})^{*}_{\rm reg})=G_{\mathbb{R}}\cdot(\operatorname{AC}(S)\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}).

In addition, if (GS)1Z(𝔩)=S(G_{\mathbb{R}}\cdot S)\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}=S holds, then

(6.3) AC(GS)1Z(𝔩)reg=AC(S)1Z(𝔩)reg.\operatorname{AC}(G_{\mathbb{R}}\cdot S)\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}=\operatorname{AC}(S)\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}.
Proof.

Since AC(GS)AC(S)\operatorname{AC}(G_{\mathbb{R}}\cdot S)\supset\operatorname{AC}(S) and AC(GS)\operatorname{AC}(G_{\mathbb{R}}\cdot S) is GG_{\mathbb{R}}-stable, we have AC(GS)GAC(S)\operatorname{AC}(G_{\mathbb{R}}\cdot S)\supset G_{\mathbb{R}}\cdot\operatorname{AC}(S). The inclusion

AC(GS)(GZ(𝔩)reg)G(AC(S)1Z(𝔩)reg)\operatorname{AC}(G_{\mathbb{R}}\cdot S)\cap(G\cdot Z(\mathfrak{l})^{*}_{\rm reg})\supset G_{\mathbb{R}}\cdot(\operatorname{AC}(S)\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg})

then follows from GZ(𝔩)regG1Z(𝔩)regG\cdot Z(\mathfrak{l})^{*}_{\rm reg}\supset G_{\mathbb{R}}\cdot\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}.

To prove the other inclusion, take a vector ξ\xi in the left hand side of (6.2). Then in particular ξ1𝔤(GZ(𝔩)reg)\xi\in\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*}\cap(G\cdot Z(\mathfrak{l})^{*}_{\rm reg}). Therefore, if 𝔩:=𝔤(ξ)\mathfrak{l}^{\prime}_{\mathbb{R}}:=\mathfrak{g}_{\mathbb{R}}(\xi), then 𝔩\mathfrak{l}^{\prime} is GG-conjugate to 𝔩\mathfrak{l}. Consider the map

a:G×1(𝔩)1𝔤a\colon G_{\mathbb{R}}\times\sqrt{-1}(\mathfrak{l}^{\prime}_{\mathbb{R}})^{*}\to\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*}

given by (g,η)gη(g,\eta)\mapsto g\cdot\eta. Identify 1(𝔩)𝔩\sqrt{-1}(\mathfrak{l}^{\prime}_{\mathbb{R}})^{*}\simeq\mathfrak{l}^{\prime}_{\mathbb{R}} in an LL_{\mathbb{R}}-invariant way and define

1(𝔩),o:={η1(𝔩)𝔩det(ad(η)|𝔤/𝔩)0}.\sqrt{-1}(\mathfrak{l}^{\prime}_{\mathbb{R}})^{*,o}:=\{\eta\in\sqrt{-1}(\mathfrak{l}^{\prime}_{\mathbb{R}})^{*}\simeq\mathfrak{l}^{\prime}_{\mathbb{R}}\mid\det(\operatorname{ad}(\eta)|_{\mathfrak{g}/\mathfrak{l}^{\prime}})\neq 0\}.

Then aa is submersive on the open set G×1(𝔩),oG_{\mathbb{R}}\times\sqrt{-1}(\mathfrak{l}^{\prime}_{\mathbb{R}})^{*,o}. We see that ξ1(𝔩),o\xi\in\sqrt{-1}(\mathfrak{l}^{\prime}_{\mathbb{R}})^{*,o}. Take an open cone C1(𝔩),oC\subset\sqrt{-1}(\mathfrak{l}^{\prime}_{\mathbb{R}})^{*,o} containing ξ\xi and take a small neighborhood eVGe\in V\subset G_{\mathbb{R}}. Then VCV\cdot C is an open cone in 1𝔤\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*} containing ξ\xi. By ξAC(GS)\xi\in\operatorname{AC}(G_{\mathbb{R}}\cdot S) and the definition of the asymptotic cone,

(GS)(VC) is unbounded.(G_{\mathbb{R}}\cdot S)\cap(V\cdot C)\text{ is unbounded.}

Since (GS)(VC)V((GS)C)(G_{\mathbb{R}}\cdot S)\cap(V\cdot C)\supset V\cdot((G_{\mathbb{R}}\cdot S)\cap C) and VV is bounded,

(GS)C is unbounded.(G_{\mathbb{R}}\cdot S)\cap C\text{ is unbounded.}

Hence there exists λS1Z(𝔩)reg\lambda\in S\subset\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg} and gGg\in G_{\mathbb{R}} such that such that gλ1(𝔩),og\cdot\lambda\in\sqrt{-1}(\mathfrak{l}^{\prime}_{\mathbb{R}})^{*,o}. Since η1(𝔩),o\eta\in\sqrt{-1}(\mathfrak{l}^{\prime}_{\mathbb{R}})^{*,o} implies 𝔤(η)𝔩\mathfrak{g}(\eta)\supset\mathfrak{l}^{\prime}, we have g𝔩𝔩g\cdot\mathfrak{l}\supset\mathfrak{l}^{\prime}. Combining with 𝔩𝔩\mathfrak{l}\sim\mathfrak{l}^{\prime}, we have g𝔩=𝔩g\cdot\mathfrak{l}_{\mathbb{R}}=\mathfrak{l}^{\prime}_{\mathbb{R}}.

Replacing ξ\xi by g1ξg^{-1}\cdot\xi, we have ξ1Z(𝔩)reg\xi\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg} and 𝔩=𝔩\mathfrak{l}_{\mathbb{R}}=\mathfrak{l}^{\prime}_{\mathbb{R}}. Take a Cartan subalgebra 𝔧𝔩\mathfrak{j}_{\mathbb{R}}\subset\mathfrak{l}_{\mathbb{R}}. If two elements in 1Z(𝔩)(1𝔧)\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}(\subset\sqrt{-1}\mathfrak{j}_{\mathbb{R}}^{*}) are GG_{\mathbb{R}}-conjugate, they lie in the same orbit for the Weyl group W=NG(𝔧)/ZG(𝔧)W_{\mathbb{R}}=N_{G_{\mathbb{R}}}(\mathfrak{j}_{\mathbb{R}})/Z_{G_{\mathbb{R}}}(\mathfrak{j}_{\mathbb{R}}). Hence

(GS)C=(WS)C.(G_{\mathbb{R}}\cdot S)\cap C=(W_{\mathbb{R}}\cdot S)\cap C.

Since WW_{\mathbb{R}} is finite, there exists wWw\in W_{\mathbb{R}} such that (wS)C(w\cdot S)\cap C, or equivalently, S(w1C)S\cap(w^{-1}\cdot C) is unbounded for any CC. This shows w1ξAC(S)w^{-1}\cdot\xi\in\operatorname{AC}(S) and hence ξGAC(S)\xi\in G_{\mathbb{R}}\cdot\operatorname{AC}(S), which implies the desired inclusion in (6.2).

To prove (6.3), take a vector ξAC(GS)1Z(𝔩)reg\xi\in\operatorname{AC}(G_{\mathbb{R}}\cdot S)\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}. Then by (6.2), we may write ξ=gξ\xi=g\cdot\xi^{\prime} such that gGg\in G_{\mathbb{R}} and ξAC(S)\xi^{\prime}\in\operatorname{AC}(S). Since 𝔤(ξ)=𝔤(ξ)=𝔩\mathfrak{g}_{\mathbb{R}}(\xi)=\mathfrak{g}_{\mathbb{R}}(\xi^{\prime})=\mathfrak{l}_{\mathbb{R}}, gg normalizes 𝔩\mathfrak{l}_{\mathbb{R}}. By our assumption, gS=Sg\cdot S=S and gAC(S)=AC(S)g\cdot\operatorname{AC}(S)=\operatorname{AC}(S). Hence ξAC(S)\xi\in\operatorname{AC}(S). This proves (6.3). ∎

By applying Lemma 6.2 to S={λ1Z(𝔩)grπ(𝔩,Γλ)suppmΠ}S=\{\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm gr}\mid\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\in\operatorname{supp}m_{\Pi}\}, the right hand side of (6.1) equals

G(AC({λ1Z(𝔩)grπ(𝔩,Γλ)suppmΠ})1Z(𝔩)reg).G_{\mathbb{R}}\cdot\left(\operatorname{AC}\bigl{(}\{\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm gr}\mid\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\in\operatorname{supp}m_{\Pi}\}\bigr{)}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}\right).

Since the wave front set WF(Π)\operatorname{WF}(\Pi) is GG_{\mathbb{R}}-stable, it is enough to show

(6.4) WF(Π)AC({λ1Z(𝔩)grπ(𝔩,Γλ)suppmΠ})1Z(𝔩)reg.\operatorname{WF}(\Pi)\supset\operatorname{AC}\bigl{(}\{\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm gr}\mid\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\in\operatorname{supp}m_{\Pi}\}\bigr{)}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}.

Recall the decompositions

1Z(𝔩)gr=𝔮1Z(𝔩)gr,𝔮, and G^𝔩=𝔮G^(𝔩,𝔮)\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm gr}=\bigsqcup_{\mathfrak{q}}\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})_{\rm gr}^{*,\mathfrak{q}},\quad\text{ and }\quad\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{\mathbb{R}}}=\bigcup_{\mathfrak{q}}\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}

defined in the previous section. Then since the asymptotic cone commutes with finite union, it is enough to show (6.4) when mΠm_{\Pi} is a measure on G^(𝔩,𝔮)\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})} for one parabolic subalgebra 𝔮\mathfrak{q}. Moreover, fix dd\in\mathbb{R} such that d>maxαΔ(𝔤,𝔧)|ρ𝔩,α|d>\max_{\alpha\in\Delta(\mathfrak{g},\mathfrak{j})}|\langle\rho_{\mathfrak{l}},\alpha^{\vee}\rangle| and write mΠm_{\Pi} as a sum of two measures according to the decomposition

G^(𝔩,𝔮)=(G^(𝔩,𝔮)G^(𝔩,d))(G^(𝔩,𝔮)G^(𝔩,d)).\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}=\bigl{(}\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}\cap\widehat{G}_{\mathbb{R}}(\mathfrak{l},d)\bigr{)}\cup\bigl{(}\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}\setminus\widehat{G}_{\mathbb{R}}(\mathfrak{l},d)\bigr{)}.

Since

AC({λ1Z(𝔩)grπ(𝔩,Γλ)G^(𝔩,d)})Z(𝔩)reg=,\operatorname{AC}\bigl{(}\{\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm gr}\mid\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\in\widehat{G}_{\mathbb{R}}(\mathfrak{l},d)\}\bigr{)}\cap Z(\mathfrak{l})^{*}_{\rm reg}=\emptyset,

it is enough to show (6.4) when suppmΠG^(𝔩,d)=\operatorname{supp}m_{\Pi}\cap\widehat{G}_{\mathbb{R}}(\mathfrak{l},d)=\emptyset. We thus assume mΠm_{\Pi} is a measure on G^(𝔩,𝔮)\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})} and suppmΠG^(𝔩,d)=\operatorname{supp}m_{\Pi}\cap\widehat{G}_{\mathbb{R}}(\mathfrak{l},d)=\emptyset.

In order to prove (6.4), we first show

(6.5) WF(Π)WF0θ(m)\operatorname{WF}(\Pi)\supset\operatorname{WF}_{0}\theta(m)

if mm is a measure on G^(𝔩,𝔮)\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})} which is equivalent to mΠm_{\Pi} and satisfies the condition (6.7) given below. We will see later that (6.7) implies mΠm_{\Pi} is of at most polynomial growth and hence θ(m)\theta(m) is defined as in Lemma 5.3.

We next take ξ\xi in the right hand side of (6.4), and define a measure mm depending on ξ\xi, which is equivalent to mΠm_{\Pi} and satisfies the condition (6.7). Then prove that

(6.6) WF0θ(m)ξ.\operatorname{WF}_{0}\theta(m)\ni\xi.

In the next few pages, we prove (6.5) for mm with the condition (6.7). Let (𝒪,Γ)(\mathcal{O},\Gamma) be a semisimple orbital parameter with 𝒪=Gλ\mathcal{O}=G_{\mathbb{R}}\cdot\lambda and λ1Z(𝔩)gr,𝔮\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}_{\rm gr}. We decompose the unitary representation (π(𝒪,Γ),V(𝒪,Γ))G^(𝔩,𝔮)(\pi(\mathcal{O},\Gamma),V_{(\mathcal{O},\Gamma)})\in\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})} as

V(𝒪,Γ)=σK^^V(𝒪,Γ)(σ)V_{(\mathcal{O},\Gamma)}=\widehat{\bigoplus_{\sigma\in\widehat{K}_{\mathbb{R}}}}V_{(\mathcal{O},\Gamma)}(\sigma)

where K:=GθGK_{\mathbb{R}}:=G_{\mathbb{R}}^{\theta}\subset G_{\mathbb{R}} is a maximal compact subgroup. We wish to choose an orthonormal basis {eσ,j(𝒪,Γ)}j\{e_{\sigma,j}(\mathcal{O},\Gamma)\}_{j} of V(𝒪,Γ)(σ)V_{(\mathcal{O},\Gamma)}(\sigma) for each π(𝒪,Γ)G^(𝔩,𝔮)\pi(\mathcal{O},\Gamma)\in\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})} and each σK^\sigma\in\widehat{K}_{\mathbb{R}}. However, we must be careful to choose these bases in a consistent way across parameters (𝒪,Γ)(\mathcal{O},\Gamma). To write down this condition correctly, we require additional notation.

Following Section 2 or [HO20, Section 2], define a parabolic subgroup P=MA(NP)P_{\mathbb{R}}=M_{\mathbb{R}}A_{\mathbb{R}}(N_{P})_{\mathbb{R}}. For each semisimple orbital parameter (𝒪,Γ)(\mathcal{O},\Gamma) with 𝒪=Gλ\mathcal{O}=G_{\mathbb{R}}\cdot\lambda, we decompose λ=λc+λn\lambda=\lambda_{c}+\lambda_{n} and define an elliptic orbital parameter (𝒪M,ΓM)(\mathcal{O}^{M_{\mathbb{R}}},\Gamma^{M_{\mathbb{R}}}) for MM_{\mathbb{R}}.

For an elliptic orbital parameter (𝒪0,Γ0)(\mathcal{O}_{0},\Gamma_{0}) for MM_{\mathbb{R}}, define

G^(𝔩,𝔮)(𝒪0,Γ0)={π(𝒪,Γ)G^(𝔩,𝔮)(𝒪M,ΓM)=(𝒪0,Γ0)}.\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}(\mathcal{O}_{0},\Gamma_{0})=\bigl{\{}\pi(\mathcal{O},\Gamma)\in\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}\mid(\mathcal{O}^{M_{\mathbb{R}}},\Gamma^{M_{\mathbb{R}}})=(\mathcal{O}_{0},\Gamma_{0})\bigr{\}}.

Then G^(𝔩,𝔮)\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})} is the disjoint union of G^(𝔩,𝔮)(𝒪0,Γ0)\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}(\mathcal{O}_{0},\Gamma_{0}) for various (𝒪0,Γ0)(\mathcal{O}_{0},\Gamma_{0}). In [HO20, Sections 2.3 and 2.4], we give a unitary representation (π(𝒪M,ΓM),V(𝒪M,ΓM))(\pi(\mathcal{O}^{M_{\mathbb{R}}},\Gamma^{M_{\mathbb{R}}}),V_{(\mathcal{O}^{M_{\mathbb{R}}},\Gamma^{M_{\mathbb{R}}})}) of MM_{\mathbb{R}} associated to (𝒪M,ΓM)(\mathcal{O}^{M_{\mathbb{R}}},\Gamma^{M_{\mathbb{R}}}). Then we form the bundle

𝒱:=G×P(V(𝒪M,ΓM)eλn+ρ(𝔫𝔭)),\mathcal{V}:=G_{\mathbb{R}}\times_{P_{\mathbb{R}}}\bigl{(}V_{(\mathcal{O}^{M_{\mathbb{R}}},\Gamma^{M_{\mathbb{R}}})}\boxtimes e^{\lambda_{n}+\rho(\mathfrak{n}_{\mathfrak{p}})}\bigr{)},

and we define

V(𝒪,Γ):=L2(G/P,𝒱).V_{(\mathcal{O},\Gamma)}:=L^{2}(G_{\mathbb{R}}/P_{\mathbb{R}},\mathcal{V}).

In order to study the action of KK_{\mathbb{R}} on V(𝒪,Γ)V_{(\mathcal{O},\Gamma)}, it is convenient to use the compact model for the induced representation (see e.g. [Kna86, Chapter 7]) obtained by restricting the sections on G/PG_{\mathbb{R}}/P_{\mathbb{R}} to sections on K/(KM)K_{\mathbb{R}}/(K_{\mathbb{R}}\cap M_{\mathbb{R}}). This gives us an identification

L2(G/P,𝒱)L2(K/(KM),𝒱|K/(KM))L^{2}(G_{\mathbb{R}}/P_{\mathbb{R}},\mathcal{V})\stackrel{{\scriptstyle\sim}}{{\rightarrow}}L^{2}(K_{\mathbb{R}}/(K_{\mathbb{R}}\cap M_{\mathbb{R}}),\mathcal{V}|_{K_{\mathbb{R}}/(K_{\mathbb{R}}\cap M_{\mathbb{R}})})

as unitary KK_{\mathbb{R}} representations. Notice that this compact picture only depends on the elliptic orbital parameter (𝒪M,ΓM)(\mathcal{O}^{M_{\mathbb{R}}},\Gamma^{M_{\mathbb{R}}}) since 𝒱|K/(KM)\mathcal{V}|_{K_{\mathbb{R}}/(K_{\mathbb{R}}\cap M_{\mathbb{R}})} is independent of λn\lambda_{n}. Now, for every σK^\sigma\in\widehat{K}_{\mathbb{R}}, we may fix an orthonormal basis for L2(K/(KM),𝒱|K/(KM))(σ)L^{2}(K_{\mathbb{R}}/(K_{\mathbb{R}}\cap M_{\mathbb{R}}),\mathcal{V}|_{K_{\mathbb{R}}/(K_{\mathbb{R}}\cap M_{\mathbb{R}})})(\sigma), and we may pull this basis back to an orthonormal basis {eσ,j(𝒪M,ΓM,λn)}\{e_{\sigma,j}(\mathcal{O}^{M_{\mathbb{R}}},\Gamma^{M_{\mathbb{R}}},\lambda_{n})\} of V(𝒪,Γ)(σ)V_{(\mathcal{O},\Gamma)}(\sigma). Since the compact model for π(𝒪,Γ)\pi(\mathcal{O},\Gamma) agrees with the compact model for π(𝒪,Γ)\pi(\mathcal{O}^{\prime},\Gamma^{\prime}) whenever (𝒪M,ΓM)=((𝒪)M,(Γ)M)(\mathcal{O}^{M_{\mathbb{R}}},\Gamma^{M_{\mathbb{R}}})=((\mathcal{O}^{\prime})^{M_{\mathbb{R}}},(\Gamma^{\prime})^{M_{\mathbb{R}}}), we note that the basis {eσ,j(𝒪M,ΓM,λn)}\{e_{\sigma,j}(\mathcal{O}^{M_{\mathbb{R}}},\Gamma^{M_{\mathbb{R}}},\lambda_{n})\} depends continuously on the parameter (𝒪,Γ)(\mathcal{O},\Gamma).

Let (M^)ellΠ(\widehat{M}_{\mathbb{R}})_{\text{ell}}^{\Pi} denote the collection of all elliptic semisimple orbital parameters (𝒪0,Γ0)(\mathcal{O}_{0},\Gamma_{0}) for MM_{\mathbb{R}} with mΠ(G^(𝔩,𝔮)(𝒪0,Γ0))0m_{\Pi}(\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}(\mathcal{O}_{0},\Gamma_{0}))\neq 0. Fix a measure mm on G^(𝔩,𝔮)\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})} equivalent to mΠm_{\Pi} such that

(6.7) m(G^(𝔩,𝔮)(𝒪0,Γ0))=1m(\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}(\mathcal{O}_{0},\Gamma_{0}))=1

for every (𝒪0,Γ0)(M^)ellΠ(\mathcal{O}_{0},\Gamma_{0})\in(\widehat{M}_{\mathbb{R}})_{\text{ell}}^{\Pi}. (6.7) implies that mm is of at most polynomial growth. Indeed, we have

λ1Z(𝔩)gr,𝔮pdm(1+|λ|2)Nλ1Z(𝔩)gr,𝔮pdm(1+|λc|2)N=(𝒪0,Γ0)(M^)ellΠG^(𝔩,𝔮)(𝒪0,Γ0)dm(1+|λc|2)N=(𝒪0,Γ0)(M^)ellΠ1(1+|λc|2)N,\begin{split}&\int_{\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})_{\rm gr}^{*,\mathfrak{q}}}\frac{p_{*}dm}{(1+|\lambda|^{2})^{N}}\\ &\leq\int_{\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})_{\rm gr}^{*,\mathfrak{q}}}\frac{p_{*}dm}{(1+|\lambda_{c}|^{2})^{N}}\\ &=\sum_{(\mathcal{O}_{0},\Gamma_{0})\in(\widehat{M}_{\mathbb{R}})_{\text{ell}}^{\Pi}}\int_{\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}(\mathcal{O}_{0},\Gamma_{0})}\frac{dm}{(1+|\lambda_{c}|^{2})^{N}}\\ &=\sum_{(\mathcal{O}_{0},\Gamma_{0})\in(\widehat{M}_{\mathbb{R}})_{\text{ell}}^{\Pi}}\frac{1}{(1+|\lambda_{c}|^{2})^{N}},\end{split}

where 𝒪0=Mλc\mathcal{O}_{0}=M_{\mathbb{R}}\cdot\lambda_{c}. Since the last expression is a sum of over a lattice with uniformly bounded finite multiplicities, it converges for a sufficiently large NN, showing that mm is of at most polynomial growth.

We now fix a multiplicity free subrepresentation

π(𝒪,Γ)G^(𝔩,𝔮)(𝒪0,Γ0)V(𝒪,Γ)𝑑mV(𝒪0,Γ0)VΠ\int^{\oplus}_{\pi(\mathcal{O},\Gamma)\in\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}(\mathcal{O}_{0},\Gamma_{0})}V_{(\mathcal{O},\Gamma)}dm\simeq V^{(\mathcal{O}_{0},\Gamma_{0})}\subset V_{\Pi}

for every (𝒪0,Γ0)(M^)ellΠ(\mathcal{O}_{0},\Gamma_{0})\in(\widehat{M}_{\mathbb{R}})_{\text{ell}}^{\Pi}. Here, VΠV_{\Pi} denotes the representation space of Π\Pi. We may then view λneσ,j(𝒪0,Γ0,λn)\lambda_{n}\mapsto e_{\sigma,j}(\mathcal{O}_{0},\Gamma_{0},\lambda_{n}) as a vector in V(𝒪0,Γ0)V^{(\mathcal{O}_{0},\Gamma_{0})} which we will denote by eσ,j(𝒪0,Γ0)e_{\sigma,j}(\mathcal{O}_{0},\Gamma_{0}). Now, since |eσ,j(𝒪0,Γ0,λn)|=1|e_{\sigma,j}(\mathcal{O}_{0},\Gamma_{0},\lambda_{n})|=1 for all λn\lambda_{n} and m(G^(𝔩,𝔮)(𝒪0,Γ0))=1m(\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}(\mathcal{O}_{0},\Gamma_{0}))=1, we deduce |eσ,j(𝒪0,Γ0)|=1|e_{\sigma,j}(\mathcal{O}_{0},\Gamma_{0})|=1.

Define

V:=eσ,j(𝒪0,Γ0)σK^,(𝒪0,Γ0)(M^)ellΠ¯V^{\prime}:=\overline{\langle e_{\sigma,j}(\mathcal{O}_{0},\Gamma_{0})\mid\sigma\in\widehat{K}_{\mathbb{R}},\ (\mathcal{O}_{0},\Gamma_{0})\in(\widehat{M}_{\mathbb{R}})_{\text{ell}}^{\Pi}\rangle}

to be the closure of the span of the eσ,j(𝒪0,Γ0)e_{\sigma,j}(\mathcal{O}_{0},\Gamma_{0}), and let P:VΠVP\colon V_{\Pi}\rightarrow V^{\prime} denote the orthogonal projection onto VV^{\prime}. Observe that for gg in a small neighborhood of eGe\in G_{\mathbb{R}}, we have

(6.8) Tr(Π(g)P)=π(𝒪0,Γ0)(M^)ellΠσK^j(Π(g)eσ,j(𝒪0,Γ0),eσ,j(𝒪0,Γ0))=π(𝒪0,Γ0)(M^)ellΠσK^jλn(Π(g)eσ,j(𝒪0,Γ0,λn),eσ,j(𝒪0,Γ0,λn))𝑑m=π(𝒪0,Γ0)(M^)ellΠλnσK^j(Π(g)eσ,j(𝒪0,Γ0,λn),eσ,j(𝒪0,Γ0,λn))dm=π(𝒪0,Γ0)(M^)ellΠλnΘπ(𝒪,Γ)𝑑m=π(𝒪,Γ)G^(𝔩,𝔮)Θπ(𝒪,Γ)𝑑m=Θ(m).\begin{split}&\phantom{=}\operatorname{Tr}(\Pi(g)P)\\ &=\sum_{\pi(\mathcal{O}_{0},\Gamma_{0})\in(\widehat{M}_{\mathbb{R}})^{\Pi}_{\text{ell}}}\sum_{\sigma\in\widehat{K}_{\mathbb{R}}}\sum_{j}(\Pi(g)e_{\sigma,j}(\mathcal{O}_{0},\Gamma_{0}),e_{\sigma,j}(\mathcal{O}_{0},\Gamma_{0}))\\ &=\sum_{\pi(\mathcal{O}_{0},\Gamma_{0})\in(\widehat{M}_{\mathbb{R}})^{\Pi}_{\text{ell}}}\sum_{\sigma\in\widehat{K}_{\mathbb{R}}}\sum_{j}\int_{\lambda_{n}}(\Pi(g)e_{\sigma,j}(\mathcal{O}_{0},\Gamma_{0},\lambda_{n}),e_{\sigma,j}(\mathcal{O}_{0},\Gamma_{0},\lambda_{n}))dm\\ &=\sum_{\pi(\mathcal{O}_{0},\Gamma_{0})\in(\widehat{M}_{\mathbb{R}})^{\Pi}_{\text{ell}}}\int_{\lambda_{n}}\sum_{\sigma\in\widehat{K}_{\mathbb{R}}}\sum_{j}(\Pi(g)e_{\sigma,j}(\mathcal{O}_{0},\Gamma_{0},\lambda_{n}),e_{\sigma,j}(\mathcal{O}_{0},\Gamma_{0},\lambda_{n}))dm\\ &=\sum_{\pi(\mathcal{O}_{0},\Gamma_{0})\in(\widehat{M}_{\mathbb{R}})^{\Pi}_{\text{ell}}}\int_{\lambda_{n}}\Theta_{\pi(\mathcal{O},\Gamma)}dm\\ &=\int_{\pi(\mathcal{O},\Gamma)\in\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}}\Theta_{\pi(\mathcal{O},\Gamma)}dm\\ &=\Theta(m).\end{split}

We have defined Θ(m)\Theta(m) on the group in the same way that we defined θ(m)\theta(m) on the Lie algebra. It is a well-defined distribution in a sufficiently small neighborhood of the identity by Lemma 5.3 and the fact that exp\exp restricts to a diffeomorphism of a neighborhood of zero onto a neighborhood of eGe\in G_{\mathbb{R}}.

Next, let ΩK𝒰(𝔨)𝒰(𝔤)\Omega_{K}\in\mathcal{U}(\mathfrak{k})\subset\mathcal{U}(\mathfrak{g}) denote the Casimir operator for KK. We wish to show that (I+ΩK)NP(I+\Omega_{K})^{-N}P is a trace class operator on VΠV_{\Pi} for sufficiently large NN. Let TKT_{\mathbb{R}}\subset K_{\mathbb{R}} be a maximal torus with Lie algebra 𝔱\mathfrak{t}_{\mathbb{R}}, and let 𝒞1𝔱\mathcal{C}\subset\sqrt{-1}\mathfrak{t}_{\mathbb{R}}^{*} be a closed Weyl chamber in 1𝔱\sqrt{-1}\mathfrak{t}_{\mathbb{R}}^{*}. For each (σ,Wσ)K^(\sigma,W_{\sigma})\in\widehat{K}_{\mathbb{R}}, let λσ𝒞\lambda_{\sigma}\in\mathcal{C} be the corresponding highest weight. Then there exists a norm |||\cdot| on the vector space 1𝔱\sqrt{-1}\mathfrak{t}_{\mathbb{R}}^{*} such that ΩKv=|λσ|2v\Omega_{K}\cdot v=|\lambda_{\sigma}|^{2}v for all vWσv\in W_{\sigma}. We calculate

(6.9) Tr((I+ΩK)NP)=π(𝒪0,Γ0)(M^)ellΠσK^j((I+ΩK)Neσ,j(𝒪0,Γ0),eσ,j(𝒪0,Γ0))=π(𝒪0,Γ0)(M^)ellΠσK^n(𝒪0,Γ0,σ)(1+|λσ|2)N,\begin{split}&\operatorname{Tr}((I+\Omega_{K})^{-N}P)\\ =&\sum_{\pi(\mathcal{O}_{0},\Gamma_{0})\in(\widehat{M}_{\mathbb{R}})^{\Pi}_{\text{ell}}}\sum_{\sigma\in\widehat{K}_{\mathbb{R}}}\sum_{j}((I+\Omega_{K})^{-N}e_{\sigma,j}(\mathcal{O}_{0},\Gamma_{0}),e_{\sigma,j}(\mathcal{O}_{0},\Gamma_{0}))\\ =&\sum_{\pi(\mathcal{O}_{0},\Gamma_{0})\in(\widehat{M}_{\mathbb{R}})^{\Pi}_{\text{ell}}}\sum_{\sigma\in\widehat{K}_{\mathbb{R}}}\frac{n(\mathcal{O}_{0},\Gamma_{0},\sigma)}{(1+|\lambda_{\sigma}|^{2})^{N}},\end{split}

where n(𝒪0,Γ0,σ)n(\mathcal{O}_{0},\Gamma_{0},\sigma) denotes the multiplicity of σK^\sigma\in\widehat{K}_{\mathbb{R}} in π(𝒪,Γ)G^(𝔩,𝔮)(𝒪0,Γ0)\pi(\mathcal{O},\Gamma)\in\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}(\mathcal{O}_{0},\Gamma_{0}). Recall (see page 205 of [Kna86]) that

(6.10) n(𝒪0,Γ0,σ)dimσn(\mathcal{O}_{0},\Gamma_{0},\sigma)\leq\dim\sigma

for all π(𝒪0,Γ0)(M^)ell\pi(\mathcal{O}_{0},\Gamma_{0})\in(\widehat{M}_{\mathbb{R}})_{\text{ell}} and all σK^\sigma\in\widehat{K}_{\mathbb{R}}. Further, there exists a natural number rr\in\mathbb{N} and a constant C>0C>0 such that

(6.11) dimσC(1+|λσ|2)r.\dim\sigma\leq C(1+|\lambda_{\sigma}|^{2})^{r}.

Therefore, utilizing (6.10) and (6.11), we have that (6.9) is bounded by

π(𝒪0,Γ0)(M^)ellσK^dimσ(1+|λσ|2)N\displaystyle\leq\sum_{\pi(\mathcal{O}_{0},\Gamma_{0})\in(\widehat{M}_{\mathbb{R}})_{\text{ell}}}\sum_{\sigma\in\widehat{K}_{\mathbb{R}}}\frac{\dim\sigma}{(1+|\lambda_{\sigma}|^{2})^{N}}
Cπ(𝒪0,Γ0)(M^)ellσK^1(1+|λσ|2)Nr.\displaystyle\leq C\sum_{\pi(\mathcal{O}_{0},\Gamma_{0})\in(\widehat{M}_{\mathbb{R}})_{\text{ell}}}\sum_{\sigma\in\widehat{K}_{\mathbb{R}}}\frac{1}{(1+|\lambda_{\sigma}|^{2})^{N-r}}.

Since {λσ}σK^\{\lambda_{\sigma}\}_{\sigma\in\widehat{K}_{\mathbb{R}}} form a subset of a lattice in 1𝔱\sqrt{-1}\mathfrak{t}_{\mathbb{R}}^{*}, we obtain the bound

σK^1(1+|λσ|2)NrCmaxL2(K/M,𝒱)(σ)01(1+|λσ|2)Nr\sum_{\sigma\in\widehat{K}_{\mathbb{R}}}\frac{1}{(1+|\lambda_{\sigma}|^{2})^{N-r}}\leq C\max_{L^{2}(K_{\mathbb{R}}/M_{\mathbb{R}},\mathcal{V})(\sigma)\neq 0}\frac{1}{(1+|\lambda_{\sigma}|^{2})^{N-r}}

for Nr+dim𝔱+1N\geq r+\dim\mathfrak{t}_{\mathbb{R}}+1 and for some C>0C>0. Let ξ(𝒪0,Γ0)1𝔱\xi_{(\mathcal{O}_{0},\Gamma_{0})}\in\sqrt{-1}\mathfrak{t}_{\mathbb{R}}^{*} denote the highest weight of the minimal KK-type of L2(K/M,𝒱(𝒪0,Γ0))L^{2}(K_{\mathbb{R}}/M_{\mathbb{R}},\mathcal{V}_{(\mathcal{O}_{0},\Gamma_{0})}). By Theorem 10.44 of [KV95] and the definition of π(𝒪0,Γ0)\pi(\mathcal{O}_{0},\Gamma_{0}) (see [HO20, §2.3]), we have ξ(𝒪0,Γ0)=λcρ(𝔫𝔨)+ρ(𝔫𝔤θ)\xi_{(\mathcal{O}_{0},\Gamma_{0})}=\lambda_{c}-\rho_{(\mathfrak{n}\cap\mathfrak{k})}+\rho_{(\mathfrak{n}\cap\mathfrak{g}^{-\theta})} when 𝒪0=Mλc\mathcal{O}_{0}=M_{\mathbb{R}}\cdot\lambda_{c} (λc1𝔱\lambda_{c}\in\sqrt{-1}\mathfrak{t}_{\mathbb{R}}^{*}). We observe

π(𝒪0,Γ0)(M^)ell1(1+|ξ(𝒪0,Γ0)|2)Nr\displaystyle\sum_{\pi(\mathcal{O}_{0},\Gamma_{0})\in(\widehat{M}_{\mathbb{R}})_{\text{ell}}}\frac{1}{(1+|\xi_{(\mathcal{O}_{0},\Gamma_{0})}|^{2})^{N-r}}
=\displaystyle= π(𝒪0,Γ0)(M^)ell1(1+|λcρ(𝔫𝔨)+ρ(𝔫𝔤θ)|2)Nr\displaystyle\sum_{\pi(\mathcal{O}_{0},\Gamma_{0})\in(\widehat{M}_{\mathbb{R}})_{\text{ell}}}\frac{1}{(1+|\lambda_{c}-\rho_{(\mathfrak{n}\cap\mathfrak{k})}+\rho_{(\mathfrak{n}\cap\mathfrak{g}^{-\theta})}|^{2})^{N-r}}

is a sum over a lattice, and we observe that each term occurs with uniformly bounded, finite multiplicity. By standard calculus arguments, we deduce that the sum converges for sufficiently large NN. It follows that (I+ΩK)NP(I+\Omega_{K})^{-N}P is of trace class for sufficiently large NN. Utilizing Howe’s original definition of the wave front set of a Lie group representation ([How81], see also [HHO16, §2] for an exposition), we have

(6.12) WFe(Tr(Π(g)(I+ΩK)NP))WF(Π)\operatorname{WF}_{e}(\operatorname{Tr}(\Pi(g)(I+\Omega_{K})^{-N}P))\subset\operatorname{WF}(\Pi)

for sufficiently large NN. Next, utilizing (6.8), we compute for φCc(G)\varphi\in C_{c}^{\infty}(G_{\mathbb{R}})

Θ(m),φ\displaystyle\langle\Theta(m),\varphi\rangle =Tr(Π(φ)P)\displaystyle=\operatorname{Tr}(\Pi(\varphi)P)
=Tr(Π(φ)(I+ΩK)N(I+ΩK)NP)\displaystyle=\operatorname{Tr}(\Pi(\varphi)(I+\Omega_{K})^{N}(I+\Omega_{K})^{-N}P)
=Tr(Π(L(I+ΩK)Nφ)(I+ΩK)NP)\displaystyle=\operatorname{Tr}(\Pi(L_{(I+\Omega_{K})^{N}}\varphi)(I+\Omega_{K})^{-N}P)
=L(I+ΩK)NTr(Π(φ)(I+ΩK)NP).\displaystyle=L_{(I+\Omega_{K})^{N}}\operatorname{Tr}(\Pi(\varphi)(I+\Omega_{K})^{-N}P).

Since applying the differential operator L(I+ΩK)NL_{(I+\Omega_{K})^{N}} can only decrease the wave front set of the distribution Tr(Π(φ)(I+ΩK)NP)\operatorname{Tr}(\Pi(\varphi)(I+\Omega_{K})^{-N}P), we conclude

WFe(Θ(m))WFe(Tr(Π(φ)(I+ΩK)NP)).\operatorname{WF}_{e}(\Theta(m))\subset\operatorname{WF}_{e}(\operatorname{Tr}(\Pi(\varphi)(I+\Omega_{K})^{-N}P)).

Combining with (6.12), we have

WFe(Θ(m))WF(Π).\operatorname{WF}_{e}(\Theta(m))\subset\operatorname{WF}(\Pi).

Finally, since θ(m)\theta(m) differs from expΘ(m)\exp^{*}\Theta(m) only by multiplication with a real analytic function, we conclude (6.5).

Next, we will define a measure mm which is equivalent to mΠm_{\Pi} and satisfies (6.6) and (6.7). We fix a positive definite, KK_{\mathbb{R}}-invariant bilinear form (,)(\cdot,\cdot) on 𝔤\mathfrak{g}_{\mathbb{R}}, which is extended by complex linearity to 𝔤\mathfrak{g}. We may then use (,)(\cdot,\cdot) to give an isomorphism 𝔤𝔤\mathfrak{g}_{\mathbb{R}}\simeq\mathfrak{g}_{\mathbb{R}}^{*}, and we write (,)(\cdot,\cdot) for the corresponding bilinear form on 𝔤\mathfrak{g}^{*}, which is positive definite on 𝔤\mathfrak{g}_{\mathbb{R}}^{*} and negative definite on 1𝔤\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*}. For ξ𝔤\xi\in\mathfrak{g}^{*}, write ξ=Reξ+1Imξ\xi=\mathrm{Re}\,\xi+\sqrt{-1}\mathrm{Im}\,\xi with Reξ,Imξ𝔤\mathrm{Re}\,\xi,\mathrm{Im}\,\xi\in\mathfrak{g}_{\mathbb{R}}^{*}. We write |ξ|:=((Reξ,Reξ)+(Imξ,Imξ))1/2|\xi|:=((\mathrm{Re}\,\xi,\mathrm{Re}\,\xi)+(\mathrm{Im}\,\xi,\mathrm{Im}\,\xi))^{1/2} for ξ𝔤\xi\in\mathfrak{g}^{*}.

Fix

ξAC({λ1Z(𝔩)grπ(𝔩,Γλ)suppmΠ})1Z(𝔩)reg.\xi\in\operatorname{AC}\bigl{(}\bigl{\{}\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm gr}\mid\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\in\operatorname{supp}m_{\Pi}\bigr{\}}\bigr{)}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}.

Replacing ξ\xi by |ξ|1ξ|\xi|^{-1}\cdot\xi we may assume |ξ|=1|\xi|=1. Write pmΠp_{*}m_{\Pi} for the pushforward of mΠm_{\Pi} by the map

p:G^(𝔩,𝔮)π(𝔩,Γλ)λ1Z(𝔩)gr,𝔮.p\colon\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})}\ni\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\mapsto\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}_{\rm gr}.

Then we can take a sequence {ζi}i>01Z(𝔩),𝔮\{\zeta_{i}\}_{i\in\mathbb{Z}_{>0}}\subset\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}} and {ti}i>0>0\{t_{i}\}_{i\in\mathbb{Z}_{>0}}\subset\mathbb{R}_{>0} such that

|ζi|=1,limiζi=ξ,ti>2i+1, and tiζisupppmΠ.|\zeta_{i}|=1,\ \ \lim_{i\to\infty}\zeta_{i}=\xi,\ \ t_{i}>2^{i+1},\text{ and }\ t_{i}\zeta_{i}\in\operatorname{supp}p_{*}m_{\Pi}.

We now want a measure mm on G^(𝔩,𝔮)\widehat{G}_{\mathbb{R}}^{(\mathfrak{l}_{\mathbb{R}},\mathfrak{q})} satisfying

(6.13) pm(B1(tiζi))2i1p_{*}m(B_{1}(t_{i}\zeta_{i}))\geq 2^{-i-1}

for all ii. Here, B1(tiζi)B_{1}(t_{i}\zeta_{i}) is the open ball in 1Z(𝔩)\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*} with radius 11 and center tiζit_{i}\zeta_{i}. It is easy to see that there exists a measure mm which is equivalent to mΠm_{\Pi} and satisfies (6.7) and (6.13). We fix such mm.

In order to prove (6.6), we require a lemma. Suppose WW is a finite-dimensional, real vector space with a positive definite inner product. Let

𝒢(x)=e|x|2/2,𝒢t(x)=et|x|2/2\mathcal{G}(x)=e^{-|x|^{2}/2},\quad\mathcal{G}_{t}(x)=e^{-t|x|^{2}/2}

denote the corresponding Gaussian and family of Gaussians on WW for t>0t>0.

Lemma 6.3 ([Fol89]).

Suppose uu is a tempered distribution on a finite-dimensional, real vector space WW. Then a vector ξW\xi\in W^{*} belongs to WF0(u)\operatorname{WF}_{0}(u) if there exists a sequence ζiW\zeta_{i}\in W^{*} and ti>0t_{i}>0 such that

limiζi=ξ,limiti=\lim_{i\to\infty}\zeta_{i}=\xi,\qquad\lim_{i\to\infty}t_{i}=\infty

and there exist NN\in\mathbb{N} and C>0C>0 such that

(6.14) |(u𝒢ti)^(tiζi)|C(1+ti2)N/2|(u\cdot\mathcal{G}_{t_{i}})^{^}\,(t_{i}\zeta_{i})|\geq C\cdot(1+t_{i}^{2})^{-N/2}

for sufficiently large ii.

Lemma 6.3 is half of [Fol89, Theorem 3.22] with ff replaced by uu and ϕ\phi replaced by 𝒢\mathcal{G}. We will apply Lemma 6.3 in the case W=1𝔤W=\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*} and u=θ(m)𝒢u=\theta(m)\cdot\mathcal{G}. The bilinear form (,)(\cdot,\cdot) we fixed above is negative definite on 1𝔤\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*}. For ζ1Z(𝔩)reg\zeta\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg} and t>0t>0, it follows from Lemma 5.3 that

(6.15) (θ(m)𝒢t+1)^(tζ)=π(𝒪,Γ)suppmη𝒞(𝒪,𝔮)(𝒢t+1)(ηtζ)ν𝑑m=cπ(𝒪,Γ)suppmη𝒞(𝒪,𝔮)1t+1e(tζη,tζη)/2(t+1)ν𝑑m.\begin{split}&(\theta(m)\cdot\mathcal{G}_{t+1})^{^}\,(t\zeta)\\ &=\int_{\pi(\mathcal{O},\Gamma)\in\operatorname{supp}m}\int_{\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q})}(\mathcal{G}_{t+1})^{\vee}(\eta-t\zeta)\nu dm\\ &=c\int_{\pi(\mathcal{O},\Gamma)\in\operatorname{supp}m}\int_{\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q})}\frac{1}{\sqrt{t+1}}e^{(t\zeta-\eta,t\zeta-\eta)/2(t+1)}\nu dm.\end{split}

where the constant c0c\neq 0 depends only on the bilinear form (,)(\cdot,\cdot). We will estimate this integral and set ζ=ζi\zeta=\zeta_{i} and t=tit=t_{i} to prove the inequality (6.14). Note that this integral converges absolutely by part (i) of Lemma 5.3. In addition, since we wish to bound this integral as in (6.14), we may safely ignore the constant cc and the factor 1t+1\frac{1}{\sqrt{t+1}} in what follows.

We estimate the integral as tt\to\infty uniformly when ζ\zeta varies in a compact subset of 1Z(𝔩)reg\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}. Fix a compact set V1Z(𝔩)regV\subset\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg} and suppose ζV\zeta\in V. We break up the integral (6.15) into two pieces

(6.16) π(𝒪,Γ)suppmη𝒞(𝒪,𝔮)|tζη|δte(tζη,tζη)/2(t+1)ν𝑑m\displaystyle\int_{\pi(\mathcal{O},\Gamma)\in\operatorname{supp}m}\int_{\begin{subarray}{c}\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q})\\ |t\zeta-\eta|\leq\delta t\end{subarray}}e^{(t\zeta-\eta,t\zeta-\eta)/2(t+1)}\nu dm
(6.17) +\displaystyle+ π(𝒪,Γ)suppmη𝒞(𝒪,𝔮)|tζη|>δte(tζη,tζη)/2(t+1)ν𝑑m.\displaystyle\int_{\pi(\mathcal{O},\Gamma)\in\operatorname{supp}m}\int_{\begin{subarray}{c}\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q})\\ |t\zeta-\eta|>\delta t\end{subarray}}e^{(t\zeta-\eta,t\zeta-\eta)/2(t+1)}\nu dm.

for some δ>0\delta>0. First, we wish to show that for every δ>0\delta>0, the size of the integral (6.17) decays faster than any rational function of tt as tt\rightarrow\infty. Then we will show that for sufficiently small δ>0\delta>0 and sufficiently large tt, the imaginary part of the integral (6.16) is small relative to the real part of the integral (6.16). Finally, we will show that the real part of the integral (6.16) is positive and bounded below by a rational function of tt.

To analyze these integrals, we put η:=1Imη1𝔤\eta^{\prime}:=\sqrt{-1}\mathrm{Im}\,\eta\in\sqrt{-1}\mathfrak{g}^{*}_{\mathbb{R}}, and we expand

(6.18) e(tζη,tζη)/2(t+1)=e(tζη,tζη)/2(t+1)e(tζη,Reη)/(t+1)e(Reη,Reη)/2(t+1).e^{(t\zeta-\eta,t\zeta-\eta)/2(t+1)}=e^{(t\zeta-\eta^{\prime},t\zeta-\eta^{\prime})/2(t+1)}\cdot e^{-(t\zeta-\eta^{\prime},\mathrm{Re}\,\eta)/(t+1)}\cdot e^{(\mathrm{Re}\,\eta,\mathrm{Re}\,\eta)/2(t+1)}.

Now, we consider the integral (6.17). We observe (tζη,Reη)/(t+1)(t\zeta-\eta^{\prime},\mathrm{Re}\,\eta)/(t+1) is an imaginary number. Hence

(6.19) |e(tζη,Reη)/(t+1)|=1.|e^{(t\zeta-\eta^{\prime},\mathrm{Re}\,\eta)/(t+1)}|=1.

In addition, there exists a constant B>0B>0 such that |Reη|B|\mathrm{Re}\,\eta|\leq B for all η𝒞(𝒪,𝔮)\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q}) and all (𝒪,Γ)(\mathcal{O},\Gamma). Therefore,

(6.20) |e(Reη,Reη)/2(t+1)|eB2/2(t+1)eB2.|e^{(\mathrm{Re}\,\eta,\mathrm{Re}\,\eta)/2(t+1)}|\leq e^{B^{2}/2(t+1)}\leq e^{B^{2}}.

Plugging (6.18), (6.19), and (6.20) into the integral (6.17), we obtain

|π(𝒪,Γ)suppmη𝒞(𝒪,𝔮)|tζη|>δte(tζη,tζη)/2(t+1)ν𝑑m|\displaystyle\left|\int_{\pi(\mathcal{O},\Gamma)\in\operatorname{supp}m}\int_{\begin{subarray}{c}\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q})\\ |t\zeta-\eta|>\delta t\end{subarray}}e^{(t\zeta-\eta,t\zeta-\eta)/2(t+1)}\nu dm\right|
(6.21) c1π(𝒪,Γ)suppmη𝒞(𝒪,𝔮)|tζη|>δt|e|tζη|2/2(t+1)ν|𝑑m\displaystyle\leq c_{1}\int_{\pi(\mathcal{O},\Gamma)\in\operatorname{supp}m}\int_{\begin{subarray}{c}\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q})\\ |t\zeta-\eta|>\delta t\end{subarray}}|e^{-|t\zeta-\eta^{\prime}|^{2}/2(t+1)}\nu|dm

for some constant c1>0c_{1}>0 independent of ζ\zeta, δ\delta, and tt. To bound this latter integral, we will apply part (i) of Lemma 5.3 with

α(η)=e|tζη|2/2(t+1).\alpha(\eta)=e^{-|t\zeta-\eta^{\prime}|^{2}/2(t+1)}.

In order to apply part (i) of Lemma 5.3, we need a lemma bounding the growth of our α(η)\alpha(\eta) as a function of tt.

Lemma 6.4.

For every N,kN,k\in\mathbb{N} and every δ>0\delta>0, there exist constants BN,k,δ>0B_{N,k,\delta}>0 and t0>0t_{0}>0 such that

supη𝒞(𝒪,𝔮)|tζη|δt(1+|η|2)N/2e|tζη|2/2(t+1)BN,k,δ(1+t2)k/2\sup_{\begin{subarray}{c}\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q})\\ |t\zeta-\eta|\geq\delta t\end{subarray}}(1+|\eta^{\prime}|^{2})^{N/2}e^{-|t\zeta-\eta^{\prime}|^{2}/2(t+1)}\leq\frac{B_{N,k,\delta}}{(1+t^{2})^{k/2}}

for t>t0t>t_{0}. The constants BN,k,δB_{N,k,\delta} and t0t_{0} do not depend on ζV\zeta\in V or 𝒪\mathcal{O}.

Proof.

Since ζ\zeta and ηη\eta-\eta^{\prime} lies in a bounded set, |tζη|δt|t\zeta-\eta|\geq\delta t implies that |η||\eta^{\prime}| is at most of order tt when tt\to\infty. On the other hand, |tζη||t\zeta-\eta^{\prime}| is at least of order tt. Hence e|tζη|2/2(t+1)e^{-|t\zeta-\eta^{\prime}|^{2}/2(t+1)} decays exponentially when tt\to\infty. This shows the existence of the constant BN,k,δB_{N,k,\delta} as in the lemma. ∎

Now, to bound (6), we apply the bound in Lemma 6.4 to Lemma 5.4, where we set MM in Lemma 5.4 to be the exponent M0M_{0} in the polynomial growth bound on the measure mm (see (5.2)). We deduce that for every δ>0\delta>0 and kk\in\mathbb{N}, there exists a constant Bk,δ>0B_{k,\delta}>0 such that

π(𝒪,Γ)suppmη𝒞(𝒪,𝔮)|tζη|>δt|e|tζη|2/2(t+1)ν|dmBk,δ(1+t2)k/2.\int_{\pi(\mathcal{O},\Gamma)\in\operatorname{supp}m}\int_{\begin{subarray}{l}\,\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q})\\ |t\zeta-\eta|>\delta t\end{subarray}}\bigl{|}e^{-|t\zeta-\eta^{\prime}|^{2}/2(t+1)}\nu\bigr{|}dm\leq\frac{B_{k,\delta}}{(1+t^{2})^{k/2}}.

Combining with (6), we obtain

(6.22) |π(𝒪,Γ)suppmη𝒞(𝒪,𝔮)|tζη|>δte(tζη,tζη)/2(t+1)ν𝑑m|c1Bk,δ(1+t2)k/2.\left|\int_{\pi(\mathcal{O},\Gamma)\in\operatorname{supp}m}\int_{\begin{subarray}{l}\,\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q})\\ |t\zeta-\eta|>\delta t\end{subarray}}e^{(t\zeta-\eta,t\zeta-\eta)/2(t+1)}\nu dm\right|\leq\frac{c_{1}B_{k,\delta}}{(1+t^{2})^{k/2}}.

The constant c1Bk,δc_{1}B_{k,\delta} does not depend on ζV\zeta\in V.

Next, we focus on the integral (6.16)

π(𝒪,Γ)suppmη𝒞(𝒪,𝔮)|tζη|δte(tζη,tζη)/2(t+1)ν𝑑m.\int_{\pi(\mathcal{O},\Gamma)\in\operatorname{supp}m}\int_{\begin{subarray}{l}\,\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q})\\ |t\zeta-\eta|\leq\delta t\end{subarray}}e^{(t\zeta-\eta,t\zeta-\eta)/2(t+1)}\nu dm.

There are two parts to the integral, the function e(tζη,tζη)/2(t+1)e^{(t\zeta-\eta,t\zeta-\eta)/2(t+1)} and the differential form ν\nu. We must analyze both separately. We begin to analyze the function e(tζη,tζη)/2(t+1)e^{(t\zeta-\eta,t\zeta-\eta)/2(t+1)} by expanding it into three terms as in (6.18). Since Reη\mathrm{Re}\,\eta is bounded, we see that given ϵ>0\epsilon^{\prime}>0, there exists t0>0t_{0}>0 such that whenever t>t0t>t_{0}, we have

(6.23) |e(Reη,Reη)/2(t+1)1|<ϵ|e^{(\mathrm{Re}\,\eta,\mathrm{Re}\,\eta)/2(t+1)}-1|<\epsilon^{\prime}

for all η𝒞(𝒪,𝔮)\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q}). This bounds the third term in the expansion (6.18). Choose B>0B>0 such that |Reη|B|\mathrm{Re}\,\eta|\leq B for all η𝒞(𝒪,𝔮)\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q}). If |tζη|δt|t\zeta-\eta|\leq\delta t, then

|(tζη,Reη)|2(t+1)|tζη||Reη|2(t+1)δtB2(t+1)δB.\frac{|(t\zeta-\eta^{\prime},\mathrm{Re}\,\eta)|}{2(t+1)}\leq\frac{|t\zeta-\eta^{\prime}||\mathrm{Re}\,\eta|}{2(t+1)}\leq\frac{\delta tB}{2(t+1)}\leq\delta B.

Therefore, given ϵ>0\epsilon^{\prime}>0, we may choose δ>0\delta>0 sufficiently small such that we have

(6.24) |e(tζη,Reη)/2(t+1)1|<ϵ|e^{(t\zeta-\eta^{\prime},\mathrm{Re}\,\eta)/2(t+1)}-1|<\epsilon^{\prime}

whenever |tζη|δt|t\zeta-\eta^{\prime}|\leq\delta t and η𝒞(𝒪,𝔮)\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q}). This bounds the second term in the expansion (6.18). Since |tζη|2|t\zeta-\eta^{\prime}|^{2}\in\mathbb{R}, we note

(6.25) e|tζη|2/2(t+1)>0.e^{-|t\zeta-\eta^{\prime}|^{2}/2(t+1)}\in\mathbb{R}_{>0}.

Define

ftζ(η):=e(tζη,tζη)/2(t+1).f_{t\zeta}(\eta):=e^{(t\zeta-\eta,t\zeta-\eta)/2(t+1)}.

Write ftζ=Reftζ+1Imftζf_{t\zeta}=\mathrm{Re}\,f_{t\zeta}+\sqrt{-1}\mathrm{Im}\,f_{t\zeta} with Reftζ,Imftζ\mathrm{Re}\,f_{t\zeta},\mathrm{Im}\,f_{t\zeta}\in\mathbb{R}.

Lemma 6.5.

There exist t0>0t_{0}>0 and δ0>0\delta_{0}>0 such that whenever t>t0t>t_{0}, δ0>δ>0\delta_{0}>\delta>0, ζV\zeta\in V, η𝒞(𝒪,𝔮)\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q}) and |tζη|δt|t\zeta-\eta^{\prime}|\leq\delta t, we have

|Imftζ|<15Reftζ.|\mathrm{Im}\,f_{t\zeta}|<\frac{1}{5}\mathrm{Re}\,f_{t\zeta}.

Lemma 6.5 follows from the expansion (6.18) together with (6.23), (6.24), (6.25). Lemma 6.5 is half of our analysis of the integral (6.16). The other half involves analyzing the differential form ν\nu. In the next section, we define a new real-valued differential form νO\nu^{\operatorname{O}} on 𝒞(𝒪,𝔮)\mathcal{C}(\mathcal{O},\mathfrak{q}). Then we bound the size of the differential form ννO\nu-\nu^{\operatorname{O}} and prove the following lemma.

Lemma 6.6.

There exist t0>0t_{0}>0 and δ0>0\delta_{0}>0 such that for t>t0t>t_{0}, δ0>δ>0\delta_{0}>\delta>0, ζV\zeta\in V, we have

|ννO|15|νO||\nu-\nu^{\operatorname{O}}|\leq\frac{1}{5}|\nu^{\operatorname{O}}|

on 𝒞(𝒪,𝔮)Bδt(tζ)\mathcal{C}(\mathcal{O},\mathfrak{q})\cap B_{\delta t}(t\zeta).

In the above lemma, the inequality |ννO|15|νO||\nu-\nu^{\operatorname{O}}|\leq\frac{1}{5}|\nu^{\operatorname{O}}| means

|(ννO)(Z1,,Z2n)|15|νO(Z1,,Z2n)||(\nu-\nu^{\operatorname{O}})(Z_{1},\ldots,Z_{2n})|\leq\frac{1}{5}|\nu^{\operatorname{O}}(Z_{1},\ldots,Z_{2n})|

for all bases {Z1,,Z2n}\{Z_{1},\ldots,Z_{2n}\} of Tη𝒞(𝒪,𝔮)T_{\eta}\mathcal{C}(\mathcal{O},\mathfrak{q}). Now, we combine Lemma 6.5 and Lemma 6.6 to estimate the integral (6.16). Define

Iζ,δ,tO:=π(𝒪,Γ)suppmη𝒞(𝒪,𝔮)|tζη|δt(Reftζ)νOdm.I_{\zeta,\delta,t}^{\operatorname{O}}:=\int_{\pi(\mathcal{O},\Gamma)\in\operatorname{supp}m}\int_{\begin{subarray}{l}\,\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q})\\ |t\zeta-\eta|\leq\delta t\end{subarray}}(\mathrm{Re}\,f_{t\zeta})\nu^{\operatorname{O}}dm.
Lemma 6.7.

There exist t0>0t_{0}>0 and δ0>0\delta_{0}>0 such that whenever δ0>δ>0\delta_{0}>\delta>0 and t>t0t>t_{0}, we have

|Iζ,δ,tIζ,δ,tO|12Iζ,δ,tO.|I_{\zeta,\delta,t}-I_{\zeta,\delta,t}^{\operatorname{O}}|\leq\frac{1}{2}I_{\zeta,\delta,t}^{\operatorname{O}}.

In the next section, we will see that νO\nu^{\operatorname{O}} is positive with respect to the given orientation of 𝒞(𝒪,𝔮)\mathcal{C}(\mathcal{O},\mathfrak{q}). Using Lemma 6.5 and Lemma 6.6, we have the pointwise estimate

|ftζν(Reftζ)νO|\displaystyle|f_{t\zeta}\nu-(\mathrm{Re}\,f_{t\zeta})\nu^{\operatorname{O}}| |Imftζ||ν|+|Reftζ(ννO)|\displaystyle\leq|\mathrm{Im}\,f_{t\zeta}||\nu|+|\mathrm{Re}\,f_{t\zeta}(\nu-\nu^{\operatorname{O}})|
15|Reftζ|65|νO|+|Reftζ|15|νO|\displaystyle\leq\frac{1}{5}|\mathrm{Re}\,f_{t\zeta}|\cdot\frac{6}{5}|\nu^{\operatorname{O}}|+|\mathrm{Re}\,f_{t\zeta}|\cdot\frac{1}{5}|\nu^{\operatorname{O}}|
12|Reftζ||νO|.\displaystyle\leq\frac{1}{2}|\mathrm{Re}\,f_{t\zeta}||\nu^{\operatorname{O}}|.

Combining this pointwise estimate with the positivity of (Reftζ)νO(\mathrm{Re}\,f_{t\zeta})\nu^{\operatorname{O}} yields Lemma 6.7.

Next, define

Iζ,δ,tO,12:=π(𝒪,Γ)suppmη𝒞(𝒪,𝔮)|tζη|δt1/2(Reftζ)νO𝑑m.I_{\zeta,\delta,t}^{\operatorname{O},\frac{1}{2}}:=\int_{\pi(\mathcal{O},\Gamma)\in\operatorname{supp}m}\int_{\begin{subarray}{l}\,\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q})\\ |t\zeta-\eta|\leq\delta t^{1/2}\end{subarray}}(\mathrm{Re}\,f_{t\zeta})\nu^{\operatorname{O}}dm.

The following lemma will be proved in the next section.

Lemma 6.8.

For any positive numbers δ>δ>0\delta>\delta^{\prime}>0, there exist t0t_{0} and C>0C>0 such that

η𝒞(𝒪λ,𝔮)|tζη|δt1/2νOC\int_{\begin{subarray}{l}\,\eta\in\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q})\\ |t\zeta-\eta|\leq\delta t^{1/2}\end{subarray}}\nu^{\operatorname{O}}\geq C

if ζV\zeta\in V, λ1Z(𝔩)gr,𝔮\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}_{\rm gr}, |tζλ|<δt1/2|t\zeta-\lambda|<\delta^{\prime}t^{1/2} and t>t0t>t_{0}.

We now complete the proof of Lemma 6.1. Let VV be a compact neighborhood of ξ\xi in 1Z(𝔩)reg\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}. Then ζiV\zeta_{i}\in V for sufficiently large ii. Take δ>0\delta>0 sufficiently small so it satisfies δ<δ0\delta<\delta_{0} in Lemma 6.7. To estimate Iδ,tO,12I_{\delta,t}^{\operatorname{O},\frac{1}{2}}, we see that ReftζCδ\mathrm{Re}\,f_{t\zeta}\geq C_{\delta} if |tζη|δt1/2|t\zeta-\eta|\leq\delta t^{1/2} for a constant CδC_{\delta}. Hence

Iδ,tO,12Cδπ(𝒪,Γ)suppmη𝒞(𝒪,𝔮)|tζη|δt1/2νO𝑑m.I_{\delta,t}^{\operatorname{O},\frac{1}{2}}\geq C_{\delta}\int_{\pi(\mathcal{O},\Gamma)\in\operatorname{supp}m}\int_{\begin{subarray}{l}\,\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q})\\ |t\zeta-\eta|\leq\delta t^{1/2}\end{subarray}}\nu^{\operatorname{O}}dm.

Then by applying Lemma 6.8 to ζ=ζi\zeta=\zeta_{i} and t=tit=t_{i}, we have

π(𝒪,Γ)suppmη𝒞(𝒪,𝔮)|tiζiη|δti1/2νO𝑑mCπ(𝒪λ,Γ)suppm|tiζiλ|<δti1/2𝑑m.\int_{\pi(\mathcal{O},\Gamma)\in\operatorname{supp}m}\int_{\begin{subarray}{l}\,\eta\in\mathcal{C}(\mathcal{O},\mathfrak{q})\\ |t_{i}\zeta_{i}-\eta|\leq\delta t_{i}^{1/2}\end{subarray}}\nu^{\operatorname{O}}dm\geq C\int_{\begin{subarray}{l}\,\pi(\mathcal{O}_{\lambda},\Gamma)\in\operatorname{supp}m\\ |t_{i}\zeta_{i}-\lambda|<\delta^{\prime}t_{i}^{1/2}\end{subarray}}dm.

When ii is sufficiently large, we have δti1/2>1\delta^{\prime}t_{i}^{1/2}>1. Hence we have

π(𝒪λ,Γ)suppm|tiζiλ|<δti1/2𝑑mπ(𝒪λ,Γ)suppm|tiζiλ|<1𝑑m=pm(B1(tiζi))2i1>ti1\int_{\begin{subarray}{l}\,\pi(\mathcal{O}_{\lambda},\Gamma)\in\operatorname{supp}m\\ |t_{i}\zeta_{i}-\lambda|<\delta^{\prime}t_{i}^{1/2}\end{subarray}}dm\geq\int_{\begin{subarray}{l}\,\pi(\mathcal{O}_{\lambda},\Gamma)\in\operatorname{supp}m\\ |t_{i}\zeta_{i}-\lambda|<1\end{subarray}}dm=p_{*}m(B_{1}(t_{i}\zeta_{i}))\geq 2^{-i-1}>t_{i}^{-1}

by (6.13). Since (Reftiζi)νO(\mathrm{Re}\,f_{t_{i}\zeta_{i}})\nu^{\operatorname{O}} is positive, we have Iζi,δ,tiO,12Iζi,δ,tiOI_{\zeta_{i},\delta,t_{i}}^{\operatorname{O},\frac{1}{2}}\leq I_{\zeta_{i},\delta,t_{i}}^{\operatorname{O}}. Therefore,

Iζi,δ,tiOCδCti1I_{\zeta_{i},\delta,t_{i}}^{\operatorname{O}}\geq C_{\delta}C\cdot t_{i}^{-1}

for sufficiently large ii. Combining with (6.15), (6.22) and Lemma 6.7, we deduce that there exists a constant C>0C>0 such that

|(θ(m)𝒢ti+1)^(tiζi)|Cti3/2|(\theta(m)\cdot\mathcal{G}_{t_{i}+1})^{^}\,(t_{i}\zeta_{i})|\geq Ct_{i}^{-3/2}

for sufficiently large ii. By Lemma 6.3, we have ξWF0(θ(m))\xi\in\operatorname{WF}_{0}(\theta(m)). Therefore, we obtain (6.4) and then Lemma 6.1.

7. Estimate of Kirillov-Kostant-Souriau form

The purpose of this section is to estimate the volume form on the contour 𝒞(𝒪,𝔮)\mathcal{C}(\mathcal{O},\mathfrak{q}) defined by the Kirillov-Kostant-Souriau symplectic form and to prove Lemma 6.6 and Lemma 6.8.

Recall that for an coadjoint orbit 𝒪λ=Gλ\mathcal{O}_{\lambda}=G_{\mathbb{R}}\cdot\lambda with λ1Z(𝔩)gr\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm gr}, and a polarization 𝔮\mathfrak{q}, the contour 𝒞(𝒪λ,𝔮)\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q}) is defined as

𝒞(𝒪λ,𝔮)={gλ+uρ𝔩gG,uU,g𝔮=u𝔮},\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q})=\{g\cdot\lambda+u\cdot\rho_{\mathfrak{l}}\mid g\in G_{\mathbb{R}},\ u\in U,\ g\cdot\mathfrak{q}=u\cdot\mathfrak{q}\},

which is a closed submanifold of the complex coadjoint orbit G(λ+ρ𝔩)G\cdot(\lambda+\rho_{\mathfrak{l}}). The tangent space of 𝒞(𝒪λ,𝔮)\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q}) is given as

Tgλ+uρ𝔩𝒞(𝒪λ,𝔮)={ad(X)(gλ)+ad(Y)(uρ𝔩)X𝔤,Y𝔲,XYg𝔮}.T_{g\cdot\lambda+u\cdot\rho_{\mathfrak{l}}}\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q})=\{\operatorname{ad}^{*}(X)(g\cdot\lambda)+\operatorname{ad}^{*}(Y)(u\cdot\rho_{\mathfrak{l}})\mid X\in\mathfrak{g}_{\mathbb{R}},\ Y\in\mathfrak{u},\ X-Y\in g\cdot\mathfrak{q}\}.

Then for each such XX and YY, there exists Z𝔤Z\in\mathfrak{g} such that

ad(X)(gλ)+ad(Y)(uρ𝔩)=ad(Z)(gλ+uρ𝔩).\operatorname{ad}^{*}(X)(g\cdot\lambda)+\operatorname{ad}^{*}(Y)(u\cdot\rho_{\mathfrak{l}})=\operatorname{ad}^{*}(Z)(g\cdot\lambda+u\cdot\rho_{\mathfrak{l}}).

Recall that the Kirillov-Kostant-Souriau symplectic form ω\omega on the complex coadjoint orbit G(λ+ρ𝔩)G\cdot(\lambda+\rho_{\mathfrak{l}}) is defined by

ωη(ad(Z)(η),ad(Z)(η)):=η([Z,Z])\omega_{\eta}(\operatorname{ad}^{*}(Z)(\eta),\operatorname{ad}^{*}(Z^{\prime})(\eta)):=\eta([Z,Z^{\prime}])

and then we defined a complex-valued 2n2n-form ν:=(2π1)n(n!)1ωn\nu:=(2\pi\sqrt{-1})^{-n}(n!)^{-1}\omega^{\wedge n}, where 2n2n is the dimension of the orbit G(λ+ρ𝔩)G\cdot(\lambda+\rho_{\mathfrak{l}}).

Let us define another 22-form ωO\omega^{O} on 𝒞(𝒪λ,𝔮)\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q}). Recall from [HO20] that we have a fiber bundle structure

(7.1) ϖ:𝒞(𝒪λ,𝔮)gλ+uρ𝔩gλ𝒪λ.\varpi\colon\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q})\ni g\cdot\lambda+u\cdot\rho_{\mathfrak{l}}\mapsto g\cdot\lambda\in\mathcal{O}_{\lambda}.

The fiber over λ\lambda is identified with (UL)ρ𝔩(UL)/(UJ)(U\cap L)\cdot\rho_{\mathfrak{l}}\simeq(U\cap L)/(U\cap J). For any g0Gg_{0}\in G_{\mathbb{R}}, there exists u0Uu_{0}\in U such that g0𝔮=u0𝔮g_{0}\cdot\mathfrak{q}=u_{0}\cdot\mathfrak{q}. Then the fiber ϖ1(g0λ)\varpi^{-1}(g_{0}\cdot\lambda) is identified with (u0(UL))ρ𝔩(u_{0}(U\cap L))\cdot\rho_{\mathfrak{l}} and then with (UL)ρ𝔩(U\cap L)\cdot\rho_{\mathfrak{l}} by the action of u01u_{0}^{-1}.

Let ωλG\omega^{G_{\mathbb{R}}}_{\lambda} (resp. ωρ𝔩UL\omega^{U\cap L}_{\rho_{\mathfrak{l}}}) denote the Kirillov-Kostant-Souriau form on the real coadjoint orbit 𝒪λ=Gλ\mathcal{O}_{\lambda}=G_{\mathbb{R}}\cdot\lambda (resp. (UL)ρ𝔩(U\cap L)\cdot\rho_{\mathfrak{l}}). To define ωO\omega^{O}, we will decompose the tangent space Tη𝒞(𝒪λ,𝔮)T_{\eta}\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q}) at η=gλ+uρ𝔩\eta=g\cdot\lambda+u\cdot\rho_{\mathfrak{l}} as

Tη𝒞(𝒪λ,𝔮)=Tηb𝒞Tηf𝒞.T_{\eta}\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q})=T_{\eta}^{b}\mathcal{C}\oplus T_{\eta}^{f}\mathcal{C}.

We define Tηf𝒞T_{\eta}^{f}\mathcal{C} as the vectors that are tangent to the fiber of ϖ\varpi. In other words,

Tηf𝒞={ad(Y)(uρ𝔩)Y𝔲(u𝔮)}.T_{\eta}^{f}\mathcal{C}=\{\operatorname{ad}^{*}(Y)(u\cdot\rho_{\mathfrak{l}})\mid Y\in\mathfrak{u}\cap(u\cdot\mathfrak{q})\}.

To define Tηb𝒞T_{\eta}^{b}\mathcal{C}, consider the natural maps

𝔤𝔤𝔤/(g𝔮)𝔲/(𝔲(g𝔮))(𝔲(g𝔮)),\mathfrak{g}_{\mathbb{R}}\to\mathfrak{g}\to\mathfrak{g}/(g\cdot\mathfrak{q})\simeq\mathfrak{u}/(\mathfrak{u}\cap(g\cdot\mathfrak{q}))\simeq(\mathfrak{u}\cap(g\cdot\mathfrak{q}))^{\perp},

where (𝔲(g𝔮))(\mathfrak{u}\cap(g\cdot\mathfrak{q}))^{\perp} is the orthogonal complement of 𝔲(g𝔮)\mathfrak{u}\cap(g\cdot\mathfrak{q}) in 𝔲\mathfrak{u} with respect to an invariant form on 𝔲\mathfrak{u}, which we fix now. Write

φ:𝔤(𝔲(g𝔮))\varphi\colon\mathfrak{g}_{\mathbb{R}}\to(\mathfrak{u}\cap(g\cdot\mathfrak{q}))^{\perp}

for the composite map. Then Xφ(X)g𝔮X-\varphi(X)\in g\cdot\mathfrak{q} for any X𝔤X\in\mathfrak{g}_{\mathbb{R}}. Define

Tηb𝒞={ad(X)(gλ)+ad(φ(X))(uρ𝔩)X𝔤}.T_{\eta}^{b}\mathcal{C}=\{\operatorname{ad}^{*}(X)(g\cdot\lambda)+\operatorname{ad}^{*}(\varphi(X))(u\cdot\rho_{\mathfrak{l}})\mid X\in\mathfrak{g}_{\mathbb{R}}\}.

Tηb𝒞T_{\eta}^{b}\mathcal{C} can be identified with Tgλ𝒪λT_{g\cdot\lambda}\mathcal{O}_{\lambda} via ϖ\varpi. Define ωO\omega^{O} as the 22-form on 𝒞(𝒪λ,𝔮)\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q}) as

ωO|Tηb𝒞=ωλG,ωO|Tηf𝒞=ωρ𝔩UL,ωO(Tηb𝒞,Tηf𝒞)=0.\omega^{O}|_{T_{\eta}^{b}\mathcal{C}}=\omega^{G_{\mathbb{R}}}_{\lambda},\quad\omega^{O}|_{T_{\eta}^{f}\mathcal{C}}=\omega^{U\cap L}_{\rho_{\mathfrak{l}}},\quad\omega^{O}(T_{\eta}^{b}\mathcal{C},T_{\eta}^{f}\mathcal{C})=0.

Here, we use the identifications Tηb𝒞Tgλ𝒪λT_{\eta}^{b}\mathcal{C}\simeq T_{g\cdot\lambda}\mathcal{O}_{\lambda} and ϖ1(gλ)ϖ1(λ)\varpi^{-1}(g\cdot\lambda)\simeq\varpi^{-1}(\lambda). Since λ1𝔤\lambda\in\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*} and ρ𝔩1(𝔲𝔩)\rho_{\mathfrak{l}}\in\sqrt{-1}(\mathfrak{u}\cap\mathfrak{l}_{\mathbb{R}})^{*}, the 22-form ωO\omega^{O} is purely imaginary. Then define a real-valued 2n2n-form νO\nu^{O} on 𝒞(𝒪λ,𝔮)\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q}) as

νO:=(ωO)n(2π1)nn!.\nu^{O}:=\frac{(\omega^{O})^{\wedge n}}{(2\pi\sqrt{-1})^{n}n!}.

In [HO20, Section 3.1] an orientation on 𝒞(𝒪λ,𝔮)\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q}) is defined in terms of symplectic forms ωλG\omega^{G_{\mathbb{R}}}_{\lambda} and ωρ𝔩UL\omega^{U\cap L}_{\rho_{\mathfrak{l}}} and the fiber bundle structure ϖ\varpi. Then it directly follows from definition that νO\nu^{O} is positive with respect to that orientation.

In the following, we estimate the differences ωωO\omega-\omega^{O} and ννO\nu-\nu^{O} to prove Lemma 6.6.

As in the previous section, we fix an inner product on 𝔤\mathfrak{g} and let |||\cdot| denote the corresponding norm on 𝔤\mathfrak{g} and on 𝔤\mathfrak{g}^{*}. For AEnd(𝔤)A\in\operatorname{End}(\mathfrak{g}^{*}) let A\|A\| denote the corresponding operator norm.

We fix a compact set V1Z(𝔩),𝔮V\subset\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}} throughout this section. We will estimate ννO\nu-\nu^{O} on 𝒞(𝒪,𝔮)Bδt(tζ)\mathcal{C}(\mathcal{O},\mathfrak{q})\cap B_{\delta t}(t\zeta), which is an open subset of 𝒞(𝒪,𝔮)\mathcal{C}(\mathcal{O},\mathfrak{q}), for any ζV\zeta\in V and any 𝒞(𝒪,𝔮)\mathcal{C}(\mathcal{O},\mathfrak{q}) when δ\delta is sufficiently small and tt is sufficiently large. Here, Bδt(tζ)B_{\delta t}(t\zeta) denotes the open ball with radius δt\delta t and center tζt\zeta in 𝔤\mathfrak{g} with respect to our fixed norm on 𝔤\mathfrak{g}. For ϵ>0\epsilon>0, let

BϵG:={gGAd(g)𝑖𝑑𝔤<ϵ},\displaystyle B_{\epsilon}^{G_{\mathbb{R}}}:=\left\{g\in G_{\mathbb{R}}\mid\|\operatorname{Ad}^{*}(g)-\it{id}_{\mathfrak{g}^{*}}\|<\epsilon\right\},
BϵU:={uUAd(u)𝑖𝑑𝔤<ϵ},\displaystyle B_{\epsilon}^{U}:=\left\{u\in U\mid\|\operatorname{Ad}^{*}(u)-\it{id}_{\mathfrak{g}^{*}}\|<\epsilon\right\},
BϵG:={gGAd(g)𝑖𝑑𝔤<ϵ}.\displaystyle B_{\epsilon}^{G}:=\left\{g\in G\mid\|\operatorname{Ad}^{*}(g)-\it{id}_{\mathfrak{g}^{*}}\|<\epsilon\right\}.

We need lemmas:

Lemma 7.1.

Given any ϵ>0\epsilon>0, there exist δ>0\delta>0 and t0>0t_{0}>0 such that the following holds: if t>t0t>t_{0}, ζV\zeta\in V, λ1Z(𝔩)gr,𝔮\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}_{\rm gr}, and

η𝒞(𝒪λ,𝔮)Bδt(tζ),\eta\in\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q})\cap B_{\delta t}(t\zeta),

then |λtζ|<ϵt|\lambda-t\zeta|<\epsilon t and there exist gBϵGg\in B_{\epsilon}^{G_{\mathbb{R}}}, uBϵUu^{\prime}\in B_{\epsilon}^{U} and uLULu_{L}\in U\cap L such that

g𝔮=(uuL)𝔮 and η=gλ+(uuL)ρ𝔩.g\cdot\mathfrak{q}=(u^{\prime}u_{L})\cdot\mathfrak{q}\ \text{ and }\ \eta=g\cdot\lambda+(u^{\prime}u_{L})\cdot\rho_{\mathfrak{l}}.
Proof.

Consider the map

G×1𝔩1𝔤,(g,η)gη,G_{\mathbb{R}}\times\sqrt{-1}\mathfrak{l}_{\mathbb{R}}^{*}\rightarrow\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*},\quad(g,\eta)\mapsto g\cdot\eta,

which is a submersion at (e,tζ)(e,t\zeta). Define 1𝔩,o\sqrt{-1}\mathfrak{l}_{\mathbb{R}}^{*,o} as in the proof of Lemma 6.2.

Take an open set V~1𝔩,o\widetilde{V}\subset\sqrt{-1}\mathfrak{l}_{\mathbb{R}}^{*,o} which contains VV. We claim that when V~\widetilde{V} is sufficiently small, we have the following: if λ1Z(𝔩),𝔮\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}, λV~1Z(𝔩)\lambda^{\prime}\in\widetilde{V}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*} and gλ=λg\cdot\lambda=\lambda^{\prime} for some gGg\in G_{\mathbb{R}}, then λ=λ\lambda=\lambda^{\prime}. Indeed, if this is not the case, we may find sequences λj1Z(𝔩),𝔮\lambda_{j}\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}, λj1Z(𝔩)\lambda^{\prime}_{j}\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*} and wjWw_{j}\in W_{\mathbb{R}} such that wjλj=λjw_{j}\cdot\lambda_{j}=\lambda^{\prime}_{j}, wj|Z(𝔩)1w_{j}|_{Z(\mathfrak{l}_{\mathbb{R}})^{*}}\neq 1 and λjλV\lambda^{\prime}_{j}\to\lambda^{\prime}\in V. Here, W=NG(𝔧)/ZG(𝔧)W_{\mathbb{R}}=N_{G_{\mathbb{R}}}(\mathfrak{j}_{\mathbb{R}})/Z_{G_{\mathbb{R}}}(\mathfrak{j}_{\mathbb{R}}) denotes the real Weyl group. By taking a subsequence, we may assume λj\lambda_{j} has a limit λ\lambda and that wj=ww_{j}=w for all jj. Then we have wλ=λw\cdot\lambda=\lambda^{\prime} with λ1Z(𝔩),𝔮¯\lambda\in\overline{\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}}, λ1Z(𝔩),𝔮\lambda^{\prime}\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}, and w|Z(𝔩)1w|_{Z(\mathfrak{l}_{\mathbb{R}})^{*}}\neq 1. It is easy to see from the definition of 1Z(𝔩),𝔮\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}} that this is not possible. Thus, the claim is proved.

Take V~\widetilde{V} that satisfies above claim. For any ϵ>0\epsilon^{\prime}>0 with ϵ>ϵ\epsilon>\epsilon^{\prime}, there exists δ>0\delta^{\prime}>0 such that

BϵGV~ζVBδ(ζ).B_{\epsilon^{\prime}}^{G_{\mathbb{R}}}\cdot\widetilde{V}\supset\bigcup_{\zeta\in V}B_{\delta^{\prime}}(\zeta).

Scaling everything by tt yields

BϵG(tV~)ζVBδt(tζ).B_{\epsilon^{\prime}}^{G_{\mathbb{R}}}\cdot(t\widetilde{V})\supset\bigcup_{\zeta\in V}B_{\delta^{\prime}t}(t\zeta).

Let c=supuU|uρ𝔩|c=\sup_{u\in U}|u\cdot\rho_{\mathfrak{l}}|, and fix 0<δ<δ0<\delta<\delta^{\prime}. Then we may find t0>0t_{0}>0 sufficiently large such that c+δt<δtc+\delta t<\delta^{\prime}t if t>t0t>t_{0}.

Now, suppose that t>t0t>t_{0}, ζV\zeta\in V, λ1Z(𝔩)gr,𝔮\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}_{\rm gr}, and η𝒞(𝒪λ,𝔮)Bδt(tζ)\eta\in\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q})\cap B_{\delta t}(t\zeta). Then by the definition of 𝒞(𝒪λ,𝔮)\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q}), we may write η=gλ+uρ𝔩\eta=g\cdot\lambda+u\cdot\rho_{\mathfrak{l}} such that gGg\in G_{\mathbb{R}}, uUu\in U, and g𝔮=u𝔮g\cdot\mathfrak{q}=u\cdot\mathfrak{q}. We have

|gλtζ||uρ𝔩|+|ηtζ|<δt.|g\cdot\lambda-t\zeta|\leq|u\cdot\rho_{\mathfrak{l}}|+|\eta-t\zeta|<\delta^{\prime}t.

Hence gλBϵG(tV~)g\cdot\lambda\in B_{\epsilon^{\prime}}^{G_{\mathbb{R}}}\cdot(t\widetilde{V}) and we can write gλ=gλg\cdot\lambda=g^{\prime}\cdot\lambda^{\prime} with gBϵGg^{\prime}\in B_{\epsilon^{\prime}}^{G_{\mathbb{R}}} and λtV~\lambda^{\prime}\in t\widetilde{V}. Then 𝔤(λ)(=𝔩)\mathfrak{g}(\lambda)(=\mathfrak{l}) and 𝔤(λ)\mathfrak{g}(\lambda^{\prime}) are conjugate and 𝔤(λ)𝔩\mathfrak{g}(\lambda^{\prime})\subset\mathfrak{l}. Therefore, 𝔤(λ)=𝔩\mathfrak{g}(\lambda^{\prime})=\mathfrak{l} and λ1Z(𝔩)\lambda^{\prime}\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}. Since λ1Z(𝔩),𝔮,λtV~1Z(𝔩)\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}},\lambda^{\prime}\in t\widetilde{V}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*} and they are in the same GG_{\mathbb{R}}-orbit, the claim at the beginning of the proof implies λ=λ\lambda=\lambda^{\prime}. Then g1gLg^{-1}g^{\prime}\in L_{\mathbb{R}} and we may replace gg by gg^{\prime}. We may thus assume gBϵGg\in B_{\epsilon^{\prime}}^{G_{\mathbb{R}}}.

Let YY denote the partial flag variety, the collection of all parabolic subalgebras of 𝔤\mathfrak{g} that are GG-conjugate to 𝔮\mathfrak{q}. Then BϵG𝔮B_{\epsilon^{\prime}}^{G_{\mathbb{R}}}\cdot\mathfrak{q} is a small open neighborhood of 𝔮\mathfrak{q} in YY. If ϵ\epsilon^{\prime} is small enough, then we can take uBϵUu^{\prime}\in B_{\epsilon}^{U} such that g𝔮=u𝔮g\cdot\mathfrak{q}=u^{\prime}\cdot\mathfrak{q}. Then uL:=(u)1uu_{L}:=(u^{\prime})^{-1}\cdot u satisfies uL𝔮=𝔮u_{L}\cdot\mathfrak{q}=\mathfrak{q} and hence uLULu_{L}\in U\cap L.

Moreover,

|λtζ||λgλ|+|gλtζ|<ϵ|λ|+δtϵ|λtζ|+ϵt|ζ|+δt.|\lambda-t\zeta|\leq|\lambda-g\cdot\lambda|+|g\cdot\lambda-t\zeta|<\epsilon^{\prime}|\lambda|+\delta^{\prime}t\leq\epsilon^{\prime}|\lambda-t\zeta|+\epsilon^{\prime}t|\zeta|+\delta^{\prime}t.

By decreasing ϵ\epsilon^{\prime} and δ\delta^{\prime} if necessary we deduce that |λtζ|<ϵt|\lambda-t\zeta|<\epsilon t. ∎

Note that there exists d>0d>0 such that if δ\delta is sufficiently small and tt is sufficiently large, then λBδt(tζ)\lambda\in B_{\delta t}(t\zeta) with ζV\zeta\in V and λ1Z(𝔩)\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*} implies that

(7.2) |λ+ρ𝔩,α|d|λ|(αΔ(𝔫,𝔧)) and |λ|2|ρ𝔩|.|\langle\lambda+\rho_{\mathfrak{l}},\alpha^{\vee}\rangle|\geq d|\lambda|\ (\forall\alpha\in\Delta(\mathfrak{n},\mathfrak{j}))\ \text{ and }\ |\lambda|\geq 2|\rho_{\mathfrak{l}}|.

Here, 𝔫\mathfrak{n} is the nilradical of 𝔮\mathfrak{q}. We fix such dd.

Lemma 7.2.

Let 0<ϵ<10<\epsilon<1 and let λ1Z(𝔩)\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*} which satisfies (7.2). Let gBϵGg\in B_{\epsilon}^{G_{\mathbb{R}}}, uBϵUu^{\prime}\in B_{\epsilon}^{U} and uLULu_{L}\in U\cap L such that

g𝔮=(uuL)𝔮 and η=gλ+(uuL)ρ𝔩.g\cdot\mathfrak{q}=(u^{\prime}u_{L})\cdot\mathfrak{q}\ \text{ and }\ \eta=g\cdot\lambda+(u^{\prime}u_{L})\cdot\rho_{\mathfrak{l}}.

Then there exist d>0d^{\prime}>0 and gcBdϵGg_{c}\in B_{d^{\prime}\epsilon}^{G} such that

η=(gcuL)(λ+ρ𝔩).\eta=(g_{c}u_{L})\cdot(\lambda+\rho_{\mathfrak{l}}).

The constant dd^{\prime} depends only on dd and does not depend on λ\lambda or ϵ\epsilon.

Proof.

Let QQ be the parabolic subgroup of GG with Lie algebra 𝔮\mathfrak{q}, or equivalently the normalizer of 𝔮\mathfrak{q}. By the assumption g𝔮=(uuL)𝔮g\cdot\mathfrak{q}=(u^{\prime}u_{L})\cdot\mathfrak{q}, we have (uuL)1gQ(u^{\prime}u_{L})^{-1}g\in Q. Then (uuL)1gλλ𝔫(u^{\prime}u_{L})^{-1}g\cdot\lambda-\lambda\in\mathfrak{n} with the identification 𝔤𝔤\mathfrak{g}\simeq\mathfrak{g}^{*}. By our assumption on gg and uu^{\prime}, we have (uuL)1guL=uL1(u)1guLBc1ϵG(u^{\prime}u_{L})^{-1}gu_{L}=u_{L}^{-1}(u^{\prime})^{-1}gu_{L}\in B_{c_{1}\epsilon}^{G} for some constant c1>0c_{1}>0. Then

|(uuL)1gλλ|=|(uuL)1guLλλ|<c1ϵ|λ|.|(u^{\prime}u_{L})^{-1}g\cdot\lambda-\lambda|=|(u^{\prime}u_{L})^{-1}gu_{L}\cdot\lambda-\lambda|<c_{1}\epsilon|\lambda|.

Decompose 𝔫\mathfrak{n} into root spaces

𝔫=i=1k𝔫i and put 𝔫>j:=i>j𝔫i.\mathfrak{n}=\bigoplus_{i=1}^{k}\mathfrak{n}_{i}\ \text{ and put }\mathfrak{n}_{>j}:=\bigoplus_{i>j}\mathfrak{n}_{i}.

The ordering is chosen to satisfy [𝔫i,𝔫]𝔫>i[\mathfrak{n}_{i},\mathfrak{n}]\subset\mathfrak{n}_{>i}. We claim that for any 1ik1\leq i\leq k, there exist a constant di>0d_{i}>0 and gciBdiϵGg_{c}^{i}\in B_{d_{i}\epsilon}^{G} such that

(7.3) (gci(λ+ρ𝔩)(λ+ρ𝔩))((uuL)1gλλ)𝔫>i.\bigl{(}g_{c}^{i}\cdot(\lambda+\rho_{\mathfrak{l}})-(\lambda+\rho_{\mathfrak{l}}\bigr{)})-\bigl{(}(u^{\prime}u_{L})^{-1}g\cdot\lambda-\lambda\bigr{)}\in\mathfrak{n}_{>i}.

This can be seen by induction on ii. Given gci1g_{c}^{i-1}, we can find gci=exp(Ni)gci1g_{c}^{i}=\exp(N_{i})g_{c}^{i-1} with Ni𝔫iN_{i}\in\mathfrak{n}_{i} which satisfies (7.3). Moreover, it follows from (7.2) that |Ni||N_{i}| is bounded by the product of ϵ\epsilon and a constant. Hence we get gciBdiϵGg_{c}^{i}\in B_{d_{i}\epsilon}^{G} for some constant did_{i}.

The claim for i=ki=k yields

gck(λ+ρ𝔩)(λ+ρ𝔩)=(uuL)1gλλ.g_{c}^{k}\cdot(\lambda+\rho_{\mathfrak{l}})-(\lambda+\rho_{\mathfrak{l}})=(u^{\prime}u_{L})^{-1}g\cdot\lambda-\lambda.

Then putting gc:=uuLgckuL1g_{c}:=u^{\prime}u_{L}g_{c}^{k}u_{L}^{-1} we get

gcuL(λ+ρ𝔩)=gλ+(uuL)ρ𝔩.g_{c}u_{L}\cdot(\lambda+\rho_{\mathfrak{l}})=g\cdot\lambda+(u^{\prime}u_{L})\cdot\rho_{\mathfrak{l}}.

By uBϵUu^{\prime}\in B_{\epsilon}^{U} and gckBdkϵGg_{c}^{k}\in B_{d_{k}\epsilon}^{G} we can choose a constant dd^{\prime} such that gcBdϵGg_{c}\in B_{d^{\prime}\epsilon}^{G}. ∎

Fix vectors X1o,,X2koX_{1}^{o},\dots,X_{2k}^{o} in 𝔤\mathfrak{g}_{\mathbb{R}} which form a basis of 𝔤/𝔩\mathfrak{g}_{\mathbb{R}}/\mathfrak{l}_{\mathbb{R}}. We have

ad(gXio)(gλ)+ad(φ(gXio))(uρ𝔩)Tηb𝒞,\operatorname{ad}^{*}(g\cdot X_{i}^{o})(g\cdot\lambda)+\operatorname{ad}^{*}(\varphi(g\cdot X_{i}^{o}))(u\cdot\rho_{\mathfrak{l}})\in T_{\eta}^{b}\mathcal{C},

where η=gλ+uρ𝔩\eta=g\cdot\lambda+u\cdot\rho_{\mathfrak{l}}. We take Xi𝔤X_{i}\in\mathfrak{g} such that

(7.4) ad(gXio)(gλ)+ad(φ(gXio))(uρ𝔩)=ad(Xi)(gλ+uρ𝔩)\operatorname{ad}^{*}(g\cdot X_{i}^{o})(g\cdot\lambda)+\operatorname{ad}^{*}(\varphi(g\cdot X_{i}^{o}))(u\cdot\rho_{\mathfrak{l}})=\operatorname{ad}^{*}(X_{i})(g\cdot\lambda+u\cdot\rho_{\mathfrak{l}})

for 1i2k1\leq i\leq 2k.

Next, fix vectors Y1o,,Y2loY_{1}^{o},\dots,Y_{2l}^{o} in 𝔲𝔩\mathfrak{u}\cap\mathfrak{l} which form a basis in (𝔲𝔩)/(𝔲𝔧)(\mathfrak{u}\cap\mathfrak{l})/(\mathfrak{u}\cap\mathfrak{j}). Then

ad(uYio)(uρ𝔩)Tηf𝒞.\operatorname{ad}^{*}(u\cdot Y_{i}^{o})(u\cdot\rho_{\mathfrak{l}})\in T_{\eta}^{f}\mathcal{C}.

We take Yi𝔤Y_{i}\in\mathfrak{g} such that

(7.5) ad(uYio)(uρ𝔩)=ad(Yi)(gλ+uρ𝔩)\operatorname{ad}^{*}(u\cdot Y_{i}^{o})(u\cdot\rho_{\mathfrak{l}})=\operatorname{ad}^{*}(Y_{i})(g\cdot\lambda+u\cdot\rho_{\mathfrak{l}})

for 1i2l1\leq i\leq 2l.

Define Zi𝔤Z_{i}\in\mathfrak{g} for 1i2k+2l=2n1\leq i\leq 2k+2l=2n as

Zi:=Xi(1i2k),Z2k+i:=Yi(1i2l).Z_{i}:=X_{i}\ (1\leq i\leq 2k),\quad Z_{2k+i}:=Y_{i}\ (1\leq i\leq 2l).

The vectors ad(Zi)(gλ+uρ𝔩)\operatorname{ad}^{*}(Z_{i})(g\cdot\lambda+u\cdot\rho_{\mathfrak{l}}) form a basis of the tangent space Tη𝒞(𝒪,𝔮)T_{\eta}\mathcal{C}(\mathcal{O},\mathfrak{q}). Let AA be a 2n2n by 2n2n matrix whose (i,j)(i,j) entry is ωη(ad(Zi)(η),ad(Zj)(η))=η([Zi,Zj])\omega_{\eta}(\operatorname{ad}^{*}(Z_{i})(\eta),\operatorname{ad}^{*}(Z_{j})(\eta))=\eta([Z_{i},Z_{j}]). Then AA is skew symmetric and the 2n2n-form ν=(2π1)n(n!)1ωn\nu=(2\pi\sqrt{-1})^{-n}(n!)^{-1}\omega^{\wedge n} is given by

ν(Z1,,Z2n)=(2π1)nPf(A),\nu(Z_{1},\dots,Z_{2n})=(2\pi\sqrt{-1})^{-n}\operatorname{Pf}(A),

where Pf(A)\operatorname{Pf}(A) denotes the Pfaffian of AA.

We now estimate each entry of AA:

Lemma 7.3.

Let ϵ,d>0\epsilon,d>0. Suppose that λ1Z(𝔩)gr,𝔮\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*,\mathfrak{q}}_{\rm gr} satisfying (7.2), gBϵGg\in B_{\epsilon}^{G_{\mathbb{R}}}, uBϵUu^{\prime}\in B_{\epsilon}^{U}, and uLULu_{L}\in U\cap L such that g𝔮=uuL𝔮g\cdot\mathfrak{q}=u^{\prime}u_{L}\cdot\mathfrak{q}. Define XiX_{i} and YiY_{i} as above for gg and u:=uuLu:=u^{\prime}u_{L}. Then we have

  1. (1)

    |(gλ+uρ𝔩)([Xi,Xj])λ([Xio,Xjo])|C|(g\cdot\lambda+u\cdot\rho_{\mathfrak{l}})([X_{i},X_{j}])-\lambda([X_{i}^{o},X_{j}^{o}])|\leq C,

  2. (2)

    |(gλ+uρ𝔩)([Xi,Yj])|C|(g\cdot\lambda+u\cdot\rho_{\mathfrak{l}})([X_{i},Y_{j}])|\leq C,

  3. (3)

    |(gλ+uρ𝔩)([Yi,Yj])ρ𝔩([Yio,Yjo])|ϵC|(g\cdot\lambda+u\cdot\rho_{\mathfrak{l}})([Y_{i},Y_{j}])-\rho_{\mathfrak{l}}([Y_{i}^{o},Y_{j}^{o}])|\leq\epsilon C

for some constant C>0C>0. Here, CC depends on ϵ\epsilon and dd, but does not depend on λ\lambda, gg, uu^{\prime} or uLu_{L}.

Proof.

By Lemma 7.2, there exists gcBdϵGg_{c}\in B_{d^{\prime}\epsilon}^{G} such that

gcuL(λ+ρ𝔩)=gλ+uρ𝔩.g_{c}u_{L}\cdot(\lambda+\rho_{\mathfrak{l}})=g\cdot\lambda+u\cdot\rho_{\mathfrak{l}}.

In the following proof, we say a vector in 𝔤\mathfrak{g} or an element in GG is bounded if it lies in a compact set which depends only on ϵ\epsilon and dd. For instance gg, uu^{\prime}, and uLu_{L} are bounded, but λ\lambda is not bounded.

Consider the equation

(7.6) ad(gXio)(uρ𝔩)+ad(φ(Xio))(uρ𝔩)\displaystyle-\operatorname{ad}^{*}(g\cdot X_{i}^{o})(u\cdot\rho_{\mathfrak{l}})+\operatorname{ad}^{*}(\varphi(X_{i}^{o}))(u\cdot\rho_{\mathfrak{l}}) =ad(Xi)(gλ+uρ𝔩)\displaystyle=\operatorname{ad}^{*}(X^{\prime}_{i})(g\cdot\lambda+u\cdot\rho_{\mathfrak{l}})
(\displaystyle\bigl{(} =ad(Xi)(gcuL(λ+ρ𝔩))).\displaystyle=\operatorname{ad}^{*}(X^{\prime}_{i})(g_{c}u_{L}\cdot(\lambda+\rho_{\mathfrak{l}}))\bigr{)}.

If we put Xi:=XigXioX^{\prime}_{i}:=X_{i}-g\cdot X_{i}^{o}, then this is equivalent to (7.4). In particular, (7.6) is satisfied for at least one XiX^{\prime}_{i} and hence the left hand side of (7.6) is contained in gcuL[𝔤,𝔧]g_{c}u_{L}\cdot[\mathfrak{g},\mathfrak{j}] with the identification 𝔤𝔤\mathfrak{g}\simeq\mathfrak{g}^{*}. Since the left hand side of (7.6) is bounded and gcuLg_{c}u_{L} is bounded, the first condition of (7.2) implies that there exists a bounded vector XiX^{\prime}_{i} which satisfies (7.6). Then by putting Xi=Xi+gXioX_{i}=X^{\prime}_{i}+g\cdot X_{i}^{o}, we find a bounded vector XiX_{i} which satisfies (7.4). Note that by (7.6) again, ad(Xi)(gλ)\operatorname{ad}^{*}(X^{\prime}_{i})(g\cdot\lambda) is also bounded.

We may thus assume that XiX_{i} are bounded vectors. To prove (1), it is enough to show that (gλ)([Xi,Xj])λ([Xio,Xjo])(g\cdot\lambda)([X_{i},X_{j}])-\lambda([X_{i}^{o},X_{j}^{o}]) is bounded. We calculate

(gλ)([Xi,Xj])λ([Xio,Xjo])\displaystyle(g\cdot\lambda)([X_{i},X_{j}])-\lambda([X_{i}^{o},X_{j}^{o}])
=(gλ)([gXio+Xi,gXjo+Xj])λ([Xio,Xjo])\displaystyle=(g\cdot\lambda)([g\cdot X_{i}^{o}+X^{\prime}_{i},\>g\cdot X_{j}^{o}+X^{\prime}_{j}])-\lambda([X_{i}^{o},X_{j}^{o}])
=(gλ)([Xi,gXjo])+(gλ)([gXio,Xj])+(gλ)([Xi,Xj])\displaystyle=(g\cdot\lambda)([X^{\prime}_{i},\>g\cdot X_{j}^{o}])+(g\cdot\lambda)([g\cdot X_{i}^{o},X^{\prime}_{j}])+(g\cdot\lambda)([X^{\prime}_{i},X^{\prime}_{j}])
=ad(Xi)(gλ),gXjo+ad(Xj)(gλ),gXioad(Xi)(gλ),gXj.\displaystyle=-\langle\operatorname{ad}^{*}(X^{\prime}_{i})(g\cdot\lambda),\>g\cdot X_{j}^{o}\rangle+\langle\operatorname{ad}^{*}(X^{\prime}_{j})(g\cdot\lambda),\>g\cdot X_{i}^{o}\rangle-\langle\operatorname{ad}^{*}(X^{\prime}_{i})(g\cdot\lambda),\>g\cdot X^{\prime}_{j}\rangle.

The last three terms are all bounded and (1) is proved.

Since the left hand side of (7.5) is bounded, we may assume that YiY_{i} is also bounded. For example, if we take YiY_{i} from gcuL[𝔤,𝔧]g_{c}u_{L}\cdot[\mathfrak{g},\mathfrak{j}], then by (7.2) YiY_{i} is bounded. Moreover, we claim that

(7.7) ϵ1|ad(Yi(gcuL)Yio)(gλ+uρ𝔩)|\epsilon^{-1}|\operatorname{ad}^{*}(Y_{i}-(g_{c}u_{L})\cdot Y_{i}^{o})(g\cdot\lambda+u\cdot\rho_{\mathfrak{l}})|

is bounded. Indeed,

ad(Yi)(gλ+uρ𝔩)ad((gcuL)Yio)(gλ+uρ𝔩)\displaystyle\operatorname{ad}^{*}(Y_{i})(g\cdot\lambda+u\cdot\rho_{\mathfrak{l}})-\operatorname{ad}^{*}((g_{c}u_{L})\cdot Y_{i}^{o})(g\cdot\lambda+u\cdot\rho_{\mathfrak{l}})
=ad(uYio)(uρ𝔩)(gcuL)(ad(Yio)(λ+ρ𝔩))\displaystyle=\operatorname{ad}^{*}(u\cdot Y_{i}^{o})(u\cdot\rho_{\mathfrak{l}})-(g_{c}u_{L})\cdot\bigl{(}\operatorname{ad}^{*}(Y_{i}^{o})(\lambda+\rho_{\mathfrak{l}})\bigr{)}
=(uuL)(ad(Yio)(ρ𝔩))(gcuL)(ad(Yio)(ρ𝔩)).\displaystyle=(u^{\prime}u_{L})\cdot\bigl{(}\operatorname{ad}^{*}(Y_{i}^{o})(\rho_{\mathfrak{l}})\bigr{)}-(g_{c}u_{L})\cdot\bigl{(}\operatorname{ad}^{*}(Y_{i}^{o})(\rho_{\mathfrak{l}})\bigr{)}.

Here, we used ad(Yio)(λ)=0\operatorname{ad}^{*}(Y_{i}^{o})(\lambda)=0 which follows from Yio𝔩Y_{i}^{o}\in\mathfrak{l}. Then the claim follows from gcBc1ϵGg_{c}\in B_{c_{1}\epsilon}^{G} and uBϵUu^{\prime}\in B_{\epsilon}^{U}.

(2) follows from

(gλ+uρ𝔩)([Xi,Yj])\displaystyle(g\cdot\lambda+u\cdot\rho_{\mathfrak{l}})([X_{i},Y_{j}]) =Xi,ad(Yj)(gλ+uρ𝔩)\displaystyle=\langle X_{i},\operatorname{ad}^{*}(Y_{j})(g\cdot\lambda+u\cdot\rho_{\mathfrak{l}})\rangle
=Xi,ad(uYjo)(uρ𝔩).\displaystyle=\langle X_{i},\operatorname{ad}^{*}(u\cdot Y_{j}^{o})(u\cdot\rho_{\mathfrak{l}})\rangle.

For (3), put Yi:=Yi(gcuL)YioY^{\prime}_{i}:=Y_{i}-(g_{c}u_{L})\cdot Y_{i}^{o}. Then

(gλ+uρ𝔩)([Yi,Yj])\displaystyle(g\cdot\lambda+u\cdot\rho_{\mathfrak{l}})([Y_{i},Y_{j}])
=(gλ+uρ𝔩)([Yi+(gcuL)Yio,Yj+(gcuL)Yjo])\displaystyle=(g\cdot\lambda+u\cdot\rho_{\mathfrak{l}})([Y^{\prime}_{i}+(g_{c}u_{L})\cdot Y_{i}^{o},\>Y^{\prime}_{j}+(g_{c}u_{L})\cdot Y_{j}^{o}])
=(gλ+uρ𝔩)([(gcuL)Yio,(gcuL)Yjo])(gcuL)Yjo,ad(Yi)(gλ+uρ𝔩)\displaystyle=(g\cdot\lambda+u\cdot\rho_{\mathfrak{l}})([(g_{c}u_{L})\cdot Y_{i}^{o},\>(g_{c}u_{L})\cdot Y_{j}^{o}])-\langle(g_{c}u_{L})\cdot Y_{j}^{o},\>\operatorname{ad}^{*}(Y^{\prime}_{i})(g\cdot\lambda+u\cdot\rho_{\mathfrak{l}})\rangle
+(gcuL)Yio,ad(Yj)(gλ+uρ𝔩)+Yi,ad(Yj)(gλ+uρ𝔩).\displaystyle\qquad\qquad+\langle(g_{c}u_{L})\cdot Y_{i}^{o},\>\operatorname{ad}^{*}(Y^{\prime}_{j})(g\cdot\lambda+u\cdot\rho_{\mathfrak{l}})\rangle+\langle Y^{\prime}_{i},\>\operatorname{ad}^{*}(Y^{\prime}_{j})(g\cdot\lambda+u\cdot\rho_{\mathfrak{l}})\rangle.

Since (7.7) is bounded, the last three terms are all bounded by ϵC\epsilon C for some constant CC. The first term is calculated as

(gλ+uρ𝔩)([(gcuL)Yio,(gcuL)Yjo])\displaystyle(g\cdot\lambda+u\cdot\rho_{\mathfrak{l}})([(g_{c}u_{L})\cdot Y_{i}^{o},\>(g_{c}u_{L})\cdot Y_{j}^{o}])
=((gcuL)(λ+ρ𝔩))([(gcuL)Yio,(gcuL)Yjo])\displaystyle=((g_{c}u_{L})\cdot(\lambda+\rho_{\mathfrak{l}}))([(g_{c}u_{L})\cdot Y_{i}^{o},\>(g_{c}u_{L})\cdot Y_{j}^{o}])
=(λ+ρ𝔩)([Yio,Yjo])\displaystyle=(\lambda+\rho_{\mathfrak{l}})([Y_{i}^{o},Y_{j}^{o}])
=ρ𝔩([Yio,Yjo]).\displaystyle=\rho_{\mathfrak{l}}([Y_{i}^{o},Y_{j}^{o}]).

(3) is thus proved. ∎

We now prove Lemma 6.6, namely, we prove

|(νηνηO)(ad(Z1)(η),,ad(Z2n)(η))|15|νηO(ad(Z1)(η),,ad(Z2n)(η))|\bigl{|}(\nu_{\eta}-\nu^{O}_{\eta})\bigl{(}\operatorname{ad}^{*}(Z_{1})(\eta),\dots,\operatorname{ad}^{*}(Z_{2n})(\eta)\bigr{)}\bigr{|}\leq\frac{1}{5}\bigl{|}\nu^{O}_{\eta}\bigl{(}\operatorname{ad}^{*}(Z_{1})(\eta),\dots,\operatorname{ad}^{*}(Z_{2n})(\eta)\bigr{)}\bigr{|}

on 𝒞(𝒪,𝔮)Bδt(tζ)\mathcal{C}(\mathcal{O},\mathfrak{q})\cap B_{\delta t}(t\zeta) when δ\delta is sufficiently small and tt is sufficiently large, or equivalently, |λ||\lambda| is sufficiently large. Since ν\nu and νO\nu^{O} are differential forms of top degree, it is enough to prove the inequality for our particular basis ad(Z1)(η),,ad(Z2n)(η)\operatorname{ad}^{*}(Z_{1})(\eta),\dots,\operatorname{ad}^{*}(Z_{2n})(\eta) of the tangent space chosen above.

Similarly to the matrix AA, let AOA^{O} be a 2n2n by 2n2n matrix whose (i,j)(i,j) entry is ωηO(ad(Zi)(η),ad(Zj)(η))\omega^{O}_{\eta}(\operatorname{ad}^{*}(Z_{i})(\eta),\operatorname{ad}^{*}(Z_{j})(\eta)). We have

νηO(ad(Z1)(η),,ad(Z2n)(η))=(2π1)nPf(AO).\nu^{O}_{\eta}\bigl{(}\operatorname{ad}^{*}(Z_{1})(\eta),\dots,\operatorname{ad}^{*}(Z_{2n})(\eta)\bigr{)}=(2\pi\sqrt{-1})^{-n}\operatorname{Pf}(A^{O}).

Hence it is enough to prove

(7.8) |Pf(A)Pf(AO)|15|Pf(AO)|.|\operatorname{Pf}(A)-\operatorname{Pf}(A^{O})|\leq\frac{1}{5}|\operatorname{Pf}(A^{O})|.

By definition of ωO\omega^{O}, the matrix AOA^{O} is block diagonal and each entry does not depend on η\eta. The upper left 2k2k by 2k2k part of AOA^{O} is λ([Xio,Xjo])\lambda([X_{i}^{o},X_{j}^{o}]). The lower right 2l2l by 2l2l part is ρ𝔩([Yio,Yjo])\rho_{\mathfrak{l}}([Y_{i}^{o},Y_{j}^{o}]). Since the the Kirillov-Kostant-Souriau form is nondegenerate, the Pfaffian of AOA^{O} does not vanish. Assuming (7.2), the Pf(AO)\operatorname{Pf}(A^{O}) grows exactly of order |λ|k|\lambda|^{k}, namely, there exist constants C1,C2>0C_{1},C_{2}>0 such that

C1|λ|k|Pf(AO)|C2|λ|k.C_{1}|\lambda|^{k}\leq|\operatorname{Pf}(A^{O})|\leq C_{2}|\lambda|^{k}.

In light of the estimate of the entries of AAOA-A^{O} given in Lemma 7.3, there exist C3,C4>0C_{3},C_{4}>0 such that

|Pf(A)Pf(AO)|C3|λ|k1+C4ϵ|λ|k.|\operatorname{Pf}(A)-\operatorname{Pf}(A^{O})|\leq C_{3}|\lambda|^{k-1}+C_{4}\epsilon|\lambda|^{k}.

Therefore, for sufficiently small ϵ\epsilon and sufficiently large |λ||\lambda| we obtain (7.8). We fix such sufficiently small ϵ>0\epsilon>0 and then Lemma 7.1 gives δ>0\delta>0. By decreasing δ\delta if necessary to have (7.2), we conclude that the inequality in Lemma 6.6 holds for sufficiently large tt.

It remains to prove Lemma 6.8. For this we use the fiber bundle structure ϖ:𝒞(𝒪λ,𝔮)𝒪λ\varpi\colon\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q})\to\mathcal{O}_{\lambda} as (7.1). We have a canonical volume form

νρ𝔩UL:=(ωρ𝔩UL)l(2π1)ll!\nu^{U\cap L}_{\rho_{\mathfrak{l}}}:=\frac{(\omega^{U\cap L}_{\rho_{\mathfrak{l}}})^{\wedge l}}{(2\pi\sqrt{-1})^{l}l!}

on the fiber ϖ1(λ)\varpi^{-1}(\lambda) and then on any fiber by an isomorphism ϖ1(g0λ)ϖ1(λ)\varpi^{-1}(g_{0}\cdot\lambda)\simeq\varpi^{-1}(\lambda). The volume of the fiber with respect to this form is a constant, which we denote by cc. Then for an open subset B𝒪λB\subset\mathcal{O}_{\lambda}, we have

(7.9) ϖ1(B)νO=cBνλG,\int_{\varpi^{-1}(B)}\nu^{O}=c\int_{B}\nu^{G_{\mathbb{R}}}_{\lambda},

where we put

νλG:=(ωλG)k(2π1)kk!.\nu^{G_{\mathbb{R}}}_{\lambda}:=\frac{(\omega^{G_{\mathbb{R}}}_{\lambda})^{\wedge k}}{(2\pi\sqrt{-1})^{k}k!}.

To study the volume form on the base 𝒪λ\mathcal{O}_{\lambda}, we fix a constant d>0d>0 and assume

(7.10) |λ,α|d|λ|(αΔ(𝔫,𝔧)) and |λ|2|ρ𝔩|.|\langle\lambda,\alpha^{\vee}\rangle|\geq d|\lambda|\ (\forall\alpha\in\Delta(\mathfrak{n},\mathfrak{j}))\ \text{ and }\ |\lambda|\geq 2|\rho_{\mathfrak{l}}|.

Let 𝔩\mathfrak{l}_{\mathbb{R}}^{\perp} be the orthogonal complement of 𝔩\mathfrak{l}_{\mathbb{R}} in 𝔤\mathfrak{g}_{\mathbb{R}} and fix a basis X1o,,X2koX_{1}^{o},\dots,X_{2k}^{o} of 𝔩\mathfrak{l}_{\mathbb{R}}^{\perp}. Let x1,,x2kx_{1},\dots,x_{2k} be linear coordinate functions on 𝔩\mathfrak{l}_{\mathbb{R}}^{\perp} with respect to this basis. Then we have a natural map

ψ:𝔩𝒪λ,Xexp(X)λ.\psi\colon\mathfrak{l}_{\mathbb{R}}^{\perp}\to\mathcal{O}_{\lambda},\quad X\mapsto\exp(X)\cdot\lambda.

Under the assumption (7.10), there exists 0<ϵ<10<\epsilon^{\prime}<1 which does not depend on λ\lambda such that ψ:Bϵψ(Bϵ)\psi\colon B_{\epsilon^{\prime}}\to\psi(B_{\epsilon^{\prime}}) is a diffeomorphism, where BϵB_{\epsilon^{\prime}} is the open ball in 𝔩\mathfrak{l}_{\mathbb{R}}^{\perp} with center at origin and radius ϵ\epsilon^{\prime} with respect to our linear coordinate. Decreasing ϵ\epsilon^{\prime} if necessary, we may further assume that ψ\psi restricted to some open set containing the closure of BϵB_{\epsilon^{\prime}} is a diffeomorphism onto its image. Moreover, ψ(Bϵ)BC1ϵ|λ|(λ)\psi(B_{\epsilon^{\prime}})\subset B_{C_{1}\epsilon^{\prime}|\lambda|}(\lambda) for some constant C1C_{1}. We claim that

|ψνλG|C2|λ|k|dx1dx2k||\psi^{*}\nu^{G_{\mathbb{R}}}_{\lambda}|\geq C_{2}|\lambda|^{k}|dx_{1}\wedge\cdots\wedge dx_{2k}|

on ψ(Bϵ)\psi(B_{\epsilon^{\prime}}) for some constant C2>0C_{2}>0. Indeed, we can find such C2C_{2} when |λ||\lambda| is bounded. Then the claim follows because |λ|k|ψνλG||\lambda|^{-k}|\psi^{*}\nu^{G_{\mathbb{R}}}_{\lambda}| is invariant under the scaling λaλ(a>0)\lambda\to a\lambda\ (a>0). Therefore, we have

BC1ϵ|λ|(λ)νλGC2|λ|kBϵ|dx1dx2k|C3(ϵ)2k|λ|k\int_{B_{C_{1}\epsilon^{\prime}|\lambda|}(\lambda)}\nu^{G_{\mathbb{R}}}_{\lambda}\geq C_{2}|\lambda|^{k}\int_{B_{\epsilon^{\prime}}}|dx_{1}\wedge\cdots\wedge dx_{2k}|\geq C_{3}(\epsilon^{\prime})^{2k}|\lambda|^{k}

for some constant C3>0C_{3}>0.

Combining with (7.9), we obtain the following.

Lemma 7.4.

There exist positive numbers ϵ0\epsilon_{0} and CC such that

ϖ1(Bϵ|λ|(λ))νOCϵ2k|λ|k\int_{\varpi^{-1}(B_{\epsilon|\lambda|}(\lambda))}\nu^{O}\geq C\epsilon^{2k}|\lambda|^{k}

for 0<ϵ<ϵ00<\epsilon<\epsilon_{0} and any λ\lambda satisfying (7.10).

To prove Lemma 6.8, fix positive numbers δ>δ>0\delta>\delta^{\prime}>0. If tt is sufficiently large, then ζV\zeta\in V and |tζλ|<δt1/2|t\zeta-\lambda|<\delta^{\prime}t^{1/2} imply that λ\lambda satisfies (7.10). Moreover, t1|λ|t^{-1}|\lambda| is bounded from below and from above by positive constants. Define ϵ\epsilon by the equation

δt12=ϵ|λ|+maxuU|uρ𝔩|+δt12.\delta t^{\frac{1}{2}}=\epsilon|\lambda|+\max_{u\in U}|u\cdot\rho_{\mathfrak{l}}|+\delta^{\prime}t^{\frac{1}{2}}.

When tt becomes larger, |λ||\lambda| is of order tt and ϵ\epsilon is of order t12t^{-\frac{1}{2}}. Hence if tt is sufficiently large, then ϵ\epsilon becomes arbitrarily small positive number. By the inclusion ϖ1(Bϵ|λ|(λ))Bδt1/2(tζ)\varpi^{-1}(B_{\epsilon|\lambda|}(\lambda))\subset B_{\delta t^{1/2}}(t\zeta) and by Lemma 7.4, we have

η𝒞(𝒪λ,𝔮)|ηtζ|δt1/2νOϖ1(Bϵ|λ|(λ))νOCϵ2k|λ|k.\int_{\begin{subarray}{l}\,\eta\in\mathcal{C}(\mathcal{O}_{\lambda},\mathfrak{q})\\ |\eta-t\zeta|\leq\delta t^{1/2}\end{subarray}}\nu^{O}\geq\int_{\varpi^{-1}(B_{\epsilon|\lambda|}(\lambda))}\nu^{O}\geq C\epsilon^{2k}|\lambda|^{k}.

Since ϵ2|λ|\epsilon^{2}|\lambda| is bounded from below by a positive constant, we obtain Lemma 6.8.

8. Proof of main theorems

In this section, we prove Theorem 1.3, Theorem 1.4 and Theorem 1.7.

Suppose that X0=G/H0X_{0}=G_{\mathbb{R}}/H_{0} is a locally algebraic homogeneous space with GG_{\mathbb{R}}-invariant density. Our proof depends on the following result of the wave front set of induced representation:

Theorem 8.1 ([HW17, Theorem 2.1]).

Let μ:TX0𝔤\mu\colon T^{*}X_{0}\to\mathfrak{g}_{\mathbb{R}}^{*} be the moment map. Then

WF(L2(X0))=1μ(TX0)¯.\operatorname{WF}(L^{2}(X_{0}))=\overline{\sqrt{-1}\mu(T^{*}X_{0})}.

First, we prove Theorem 1.4. According to Theorem 4.3, we can divide the set suppL2(X0)\operatorname{supp}L^{2}(X_{0}) as

(8.1) suppL2(X0)G^(𝔩X,d)𝔩G^𝔩\operatorname{supp}L^{2}(X_{0})\subset\widehat{G}_{\mathbb{R}}(\mathfrak{l}_{X},d)\cup\bigcup_{\mathfrak{l}_{\mathbb{R}}}\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{\mathbb{R}}}

for some constant dd, where 𝔩\mathfrak{l}_{\mathbb{R}} runs over representatives of all GG_{\mathbb{R}}-conjugacy classes such that 𝔩(=𝔩)\mathfrak{l}(=\mathfrak{l}_{\mathbb{R}}\otimes\mathbb{C}) is GG-conjugate to 𝔩X\mathfrak{l}_{X}. If dd is large enough, π(𝔩,Γλ)G^𝔩G^(𝔩X,d)\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\in\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{\mathbb{R}}}\setminus\widehat{G}_{\mathbb{R}}(\mathfrak{l}_{X},d) implies λ\lambda is far from Z(𝔩)Z(𝔩)regZ(\mathfrak{l}_{\mathbb{R}})^{*}\setminus Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}. In view of the Langlands parameter of π(𝔩,Γλ)\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda}) in Section 2, we have

(8.2) (G^𝔩G^(𝔩X,d))(G^𝔩G^(𝔩X,d))=\bigl{(}\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{\mathbb{R}}}\setminus\widehat{G}_{\mathbb{R}}(\mathfrak{l}_{X},d)\bigr{)}\cap\bigl{(}\widehat{G}_{\mathbb{R}}^{\mathfrak{l}^{\prime}_{\mathbb{R}}}\setminus\widehat{G}_{\mathbb{R}}(\mathfrak{l}_{X},d)\bigr{)}=\emptyset

if 𝔩\mathfrak{l}_{\mathbb{R}} and 𝔩\mathfrak{l}^{\prime}_{\mathbb{R}} are not GG_{\mathbb{R}}-conjugate and if dd is sufficiently large. We fix dd satisfying (8.1) and (8.2). Then we obtain the decomposition of suppL2(X0)\operatorname{supp}L^{2}(X_{0}):

suppL2(X0)=(suppL2(X0)G^(𝔩X,d))𝔩((suppL2(X0)G^𝔩)G^(𝔩X,d))\operatorname{supp}L^{2}(X_{0})=\bigl{(}\operatorname{supp}L^{2}(X_{0})\cap\widehat{G}_{\mathbb{R}}(\mathfrak{l}_{X},d)\bigr{)}\sqcup\bigsqcup_{\mathfrak{l}_{\mathbb{R}}}\bigl{(}(\operatorname{supp}L^{2}(X_{0})\cap\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{\mathbb{R}}})\setminus\widehat{G}_{\mathbb{R}}(\mathfrak{l}_{X},d)\bigr{)}

In this decomposition, we note that suppL2(X0)G^(𝔩X,d)\operatorname{supp}L^{2}(X_{0})\cap\widehat{G}_{\mathbb{R}}(\mathfrak{l}_{X},d) is open in suppL2(X0)\operatorname{supp}L^{2}(X_{0}) and (suppL2(X0)G^𝔩)G^(𝔩X,d)(\operatorname{supp}L^{2}(X_{0})\cap\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{\mathbb{R}}})\setminus\widehat{G}_{\mathbb{R}}(\mathfrak{l}_{X},d) is closed in suppL2(X0)\operatorname{supp}L^{2}(X_{0}).

Let

L2(X0)G^πn(π)𝑑mL^{2}(X_{0})\simeq\int_{\widehat{G}_{\mathbb{R}}}^{\oplus}\pi^{\oplus n(\pi)}dm

be the irreducible decomposition. Define

V=G^(𝔩X,d)πn(π)𝑑m,V𝔩=G^𝔩G^(𝔩X,d)πn(π)𝑑m,V^{\prime}=\int_{\widehat{G}_{\mathbb{R}}(\mathfrak{l}_{X},d)}^{\oplus}\pi^{\oplus n(\pi)}dm,\qquad V_{\mathfrak{l}_{\mathbb{R}}}=\int_{\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{\mathbb{R}}}\setminus\widehat{G}_{\mathbb{R}}(\mathfrak{l}_{X},d)}^{\oplus}\pi^{\oplus n(\pi)}dm,

and regard them as subrepresentations of L2(X0)L^{2}(X_{0}) so we have

L2(X0)=V𝔩V𝔩,WF(L2(X0))=WF(V)𝔩WF(V𝔩).L^{2}(X_{0})=V^{\prime}\oplus\bigoplus_{\mathfrak{l}_{\mathbb{R}}}V_{\mathfrak{l}_{\mathbb{R}}},\qquad\operatorname{WF}(L^{2}(X_{0}))=\operatorname{WF}(V^{\prime})\cup\bigcup_{\mathfrak{l}_{\mathbb{R}}}\operatorname{WF}(V_{\mathfrak{l}_{\mathbb{R}}}).

By Lemma 4.4, we have

WF(L2(X0))GZ(𝔩X)reg=𝔩(WF(V𝔩)GZ(𝔩X)reg).\operatorname{WF}(L^{2}(X_{0}))\cap G\cdot Z(\mathfrak{l}_{X})^{*}_{\rm reg}=\bigcup_{\mathfrak{l}_{\mathbb{R}}}\bigl{(}\operatorname{WF}(V_{\mathfrak{l}_{\mathbb{R}}})\cap G\cdot Z(\mathfrak{l}_{X})^{*}_{\rm reg}\bigr{)}.

Hence Theorem 5.1 and Theorem 8.1 imply

1μ(TX0)¯GZ(𝔩X)reg=𝔩AC(π(𝔩,Γλ)suppV𝔩Gλ)GZ(𝔩X)reg.\overline{\sqrt{-1}\mu(T^{*}X_{0})}\cap G\cdot Z(\mathfrak{l}_{X})^{*}_{\rm reg}=\bigcup_{\mathfrak{l}_{\mathbb{R}}}\operatorname{AC}\Biggl{(}\bigcup_{\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\in\operatorname{supp}V_{\mathfrak{l}_{\mathbb{R}}}}G_{\mathbb{R}}\cdot\lambda\Biggr{)}\cap G\cdot Z(\mathfrak{l}_{X})^{*}_{\rm reg}.

Since (suppL2(X0)G^𝔩)G^(𝔩X,d)(\operatorname{supp}L^{2}(X_{0})\cap\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{\mathbb{R}}})\setminus\widehat{G}_{\mathbb{R}}(\mathfrak{l}_{X},d) is closed in suppL2(X0)\operatorname{supp}L^{2}(X_{0}), we have

suppV𝔩=(suppL2(X0)G^𝔩)G^(𝔩X,d).\operatorname{supp}V_{\mathfrak{l}_{\mathbb{R}}}=(\operatorname{supp}L^{2}(X_{0})\cap\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{\mathbb{R}}})\setminus\widehat{G}_{\mathbb{R}}(\mathfrak{l}_{X},d).

As in (4.5), we can easily show that

AC(π(𝔩,Γλ)G^(𝔩X,d)Gλ)GZ(𝔩X)reg=.\displaystyle\operatorname{AC}\Biggl{(}\bigcup_{\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\in\widehat{G}_{\mathbb{R}}(\mathfrak{l}_{X},d)}G_{\mathbb{R}}\cdot\lambda\Biggr{)}\cap G\cdot Z(\mathfrak{l}_{X})^{*}_{\rm reg}=\emptyset.

Hence

𝔩AC(π(𝔩,Γλ)suppV𝔩Gλ)GZ(𝔩X)reg\displaystyle\bigcup_{\mathfrak{l}_{\mathbb{R}}}\operatorname{AC}\Biggl{(}\bigcup_{\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\in\operatorname{supp}V_{\mathfrak{l}_{\mathbb{R}}}}G_{\mathbb{R}}\cdot\lambda\Biggr{)}\cap G\cdot Z(\mathfrak{l}_{X})^{*}_{\rm reg}
=AC(π(𝔩,Γλ)suppL2(X0)Gλ)GZ(𝔩X)reg.\displaystyle=\operatorname{AC}\Biggl{(}\bigcup_{\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\in\operatorname{supp}L^{2}(X_{0})}G_{\mathbb{R}}\cdot\lambda\Biggr{)}\cap G\cdot Z(\mathfrak{l}_{X})^{*}_{\rm reg}.

Therefore, putting

S𝔩:={λ1Z(𝔩)grΓλ such that π(𝔩,Γλ)suppL2(X0)},S_{\mathfrak{l}_{\mathbb{R}}}:=\{\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm gr}\mid\exists\Gamma_{\lambda}\text{ such that }\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\in\operatorname{supp}L^{2}(X_{0})\},

we have

(8.3) 1μ(TX0)¯GZ(𝔩X)reg=𝔩(AC(GS𝔩)GZ(𝔩X)reg).\overline{\sqrt{-1}\mu(T^{*}X_{0})}\cap G\cdot Z(\mathfrak{l}_{X})^{*}_{\rm reg}=\bigcup_{\mathfrak{l}_{\mathbb{R}}}\bigl{(}\operatorname{AC}(G_{\mathbb{R}}\cdot S_{\mathfrak{l}_{\mathbb{R}}})\cap G\cdot Z(\mathfrak{l}_{X})^{*}_{\rm reg}\bigr{)}.

This proves the equation

1μ(TX0)¯GZ(𝔩X)reg=AC(π(𝒪,Γ)suppL2(X0)𝒪)GZ(𝔩X)reg\overline{\sqrt{-1}\mu(T^{*}X_{0})}\cap G\cdot Z(\mathfrak{l}_{X})^{*}_{\rm reg}=\operatorname{AC}\Biggl{(}\bigcup_{\pi(\mathcal{O},\Gamma)\in\operatorname{supp}L^{2}(X_{0})}\mathcal{O}\Biggr{)}\cap G\cdot Z(\mathfrak{l}_{X})^{*}_{\rm reg}

in Theorem 1.4. To show the remaining equation in Theorem 1.4, we replace (8.1) by

suppL2(X0)G^(𝔩X,d)𝔩{π(𝔩,Γλ)G^𝔩λ𝔞X},\operatorname{supp}L^{2}(X_{0})\subset\widehat{G}_{\mathbb{R}}(\mathfrak{l}_{X},d)\cup\bigcup_{\mathfrak{l}_{\mathbb{R}}}\{\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\in\widehat{G}_{\mathbb{R}}^{\mathfrak{l}_{\mathbb{R}}}\mid\lambda\in\mathfrak{a}_{X}^{*}\},

which was proved in Theorem 4.3. Then the same argument shows

1μ(TX0)¯GZ(𝔩X)reg=AC(π(𝒪,Γ)suppL2(X0)(G𝒪)𝔞X𝒪)GZ(𝔩X)reg.\overline{\sqrt{-1}\mu(T^{*}X_{0})}\cap G\cdot Z(\mathfrak{l}_{X})^{*}_{\rm reg}=\operatorname{AC}\Biggl{(}\bigcup_{\begin{subarray}{c}\pi(\mathcal{O},\Gamma)\in\operatorname{supp}L^{2}(X_{0})\\ (G\cdot\mathcal{O})\cap\mathfrak{a}_{X}\neq\emptyset\end{subarray}}\mathcal{O}\Biggr{)}\cap G\cdot Z(\mathfrak{l}_{X})^{*}_{\rm reg}.

This completes the proof of Theorem 1.4.


Next, we prove (LABEL:eq:main) in Theorem 1.3. Fix a Levi subalgebra 𝔩\mathfrak{l}_{\mathbb{R}} with 𝔩𝔩X\mathfrak{l}\sim\mathfrak{l}_{X}. Taking the intersection of 1Z(𝔩)reg\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg} and (8.3), we have

(8.4) 1μ(TX0)¯1Z(𝔩)reg=𝔩AC(GS𝔩)1Z(𝔩)reg.\overline{\sqrt{-1}\mu(T^{*}X_{0})}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}=\bigcup_{\mathfrak{l}^{\prime}_{\mathbb{R}}}\operatorname{AC}(G_{\mathbb{R}}\cdot S_{\mathfrak{l}^{\prime}_{\mathbb{R}}})\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}.

If 𝔩=𝔩\mathfrak{l}_{\mathbb{R}}=\mathfrak{l}^{\prime}_{\mathbb{R}}, then

AC(GS𝔩)1Z(𝔩)reg=AC(S𝔩)1Z(𝔩)reg\operatorname{AC}(G_{\mathbb{R}}\cdot S_{\mathfrak{l}_{\mathbb{R}}})\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}=\operatorname{AC}(S_{\mathfrak{l}_{\mathbb{R}}})\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}

by applying (6.3). If 𝔩\mathfrak{l}_{\mathbb{R}} and 𝔩\mathfrak{l}^{\prime}_{\mathbb{R}} are not GG_{\mathbb{R}}-conjugate, then (6.2) gives

AC(GS𝔩)1Z(𝔩)reg=G(AC(S𝔩)1Z(𝔩)reg)1Z(𝔩)reg=\operatorname{AC}(G_{\mathbb{R}}\cdot S_{\mathfrak{l}^{\prime}_{\mathbb{R}}})\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}=G_{\mathbb{R}}\cdot\bigl{(}\operatorname{AC}(S_{\mathfrak{l}^{\prime}_{\mathbb{R}}})\cap\sqrt{-1}Z(\mathfrak{l}^{\prime}_{\mathbb{R}})^{*}_{\rm reg}\bigr{)}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}=\emptyset

because G1Z(𝔩)regG_{\mathbb{R}}\cdot\sqrt{-1}Z(\mathfrak{l}^{\prime}_{\mathbb{R}})^{*}_{\rm reg} does not intersect 1Z(𝔩)reg\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}. Therefore, the right hand side of (8.4) equals

AC(S𝔩)1Z(𝔩)reg\displaystyle\operatorname{AC}(S_{\mathfrak{l}_{\mathbb{R}}})\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}
=AC({λ1Z(𝔩)grπ(𝔩,Γλ)suppL2(X0)})1Z(𝔩)reg.\displaystyle=\operatorname{AC}\bigl{(}\bigl{\{}\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\mathrm{gr}}\mid\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\in\operatorname{supp}L^{2}(X_{0})\bigr{\}}\bigr{)}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})_{\mathrm{reg}}^{*}.

This prove the second equation of (LABEL:eq:main). The other equation of (LABEL:eq:main) can be proved in the same way.

To prove the remaining assertion of Theorem 1.3, we may replace μ(TX0)\mu(T^{*}X_{0}) by μ(TX)\mu(T^{*}X_{\mathbb{R}}), where X:=G/HX_{\mathbb{R}}:=G_{\mathbb{R}}/H_{\mathbb{R}}. Indeed, if 𝔥:={ξ𝔤ξ|𝔥=0}\mathfrak{h}_{\mathbb{R}}^{\perp}:=\{\xi\in\mathfrak{g}_{\mathbb{R}}^{*}\mid\xi|_{\mathfrak{h}_{\mathbb{R}}}=0\}, then μ(TX0)=G𝔥=μ(TX)\mu(T^{*}X_{0})=G_{\mathbb{R}}\cdot\mathfrak{h}_{\mathbb{R}}^{\perp}=\mu(T^{*}X_{\mathbb{R}}). The manifold XX_{\mathbb{R}} may not be an algebraic variety but a union of connected components of the \mathbb{R}-valued points of G/HG/H. We have XXX_{\mathbb{R}}\subset X and for xXx\in X_{\mathbb{R}} there is a natural decomposition TxX=TxX1TxXT_{x}X=T_{x}X_{\mathbb{R}}\oplus\sqrt{-1}T_{x}X_{\mathbb{R}}. Hence there exists a natural inclusion TXTXT^{*}X_{\mathbb{R}}\subset T^{*}X. Put n:=dimμ(TX)n:=\dim_{\mathbb{C}}\mu(T^{*}X). By Theorem 1.2,

n=dimG𝔞X=dim𝔞X+dim𝔤/𝔩.n=\dim_{\mathbb{C}}G\cdot\mathfrak{a}_{X}^{*}=\dim_{\mathbb{C}}\mathfrak{a}_{X}^{*}+\dim_{\mathbb{C}}\mathfrak{g}/\mathfrak{l}.

Define

(TX)o:={(x,ξ)TXξGZ(𝔩X)reg and rankdμ(x,ξ)=n},\displaystyle(T^{*}X)^{o}:=\{(x,\xi)\in T^{*}X\mid\xi\in G\cdot Z(\mathfrak{l}_{X})^{*}_{\rm reg}\text{ and }\operatorname{rank}d\mu_{(x,\xi)}=n\},
(TX)o:=TX(TX)o.\displaystyle(T^{*}X_{\mathbb{R}})^{o}:=T^{*}X_{\mathbb{R}}\cap(T^{*}X)^{o}.

Then (TX)o(T^{*}X)^{o} is a Zariski open dense set in TXT^{*}X. Therefore, (TX)o(T^{*}X_{\mathbb{R}})^{o} is open and dense in TXT^{*}X_{\mathbb{R}}.

Observe that

(8.5) (GZ(𝔩X)reg)𝔤=𝔩GZ(𝔩)reg.\bigl{(}G\cdot Z(\mathfrak{l}_{X})^{*}_{\rm reg}\bigr{)}\cap\mathfrak{g}_{\mathbb{R}}^{*}=\bigsqcup_{\mathfrak{l}_{\mathbb{R}}}G_{\mathbb{R}}\cdot Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}.

Here, as in (8.1), 𝔩\mathfrak{l}_{\mathbb{R}} runs over representatives of all GG_{\mathbb{R}}-conjugacy classes of Levi subalgebras of 𝔤\mathfrak{g}_{\mathbb{R}} such that 𝔩𝔩X\mathfrak{l}\sim\mathfrak{l}_{X}. Indeed, if ξ\xi is in the left hand side of (8.5), then 𝔤(ξ)\mathfrak{g}_{\mathbb{R}}(\xi) is GG_{\mathbb{R}}-conjugate to exactly one of 𝔩\mathfrak{l}_{\mathbb{R}} in the right hand side of (8.5). Then ξZ(𝔩)reg\xi\in Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg} for this 𝔩\mathfrak{l}_{\mathbb{R}}. Let S=1Z(𝔩)S=\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*} and apply Lemma 6.2. Since GSG_{\mathbb{R}}\cdot S is a cone, AC(GS)=GS¯\operatorname{AC}(G_{\mathbb{R}}\cdot S)=\overline{G_{\mathbb{R}}\cdot S}. Then (6.2) multiplied by 1\sqrt{-1} becomes

GZ(𝔩)¯(GZ(𝔩X)reg)=GZ(𝔩)reg.\overline{G_{\mathbb{R}}\cdot Z(\mathfrak{l}_{\mathbb{R}})^{*}}\cap\bigl{(}G\cdot Z(\mathfrak{l}_{X})^{*}_{\rm reg}\bigr{)}=G_{\mathbb{R}}\cdot Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}.

This shows each GZ(𝔩)regG_{\mathbb{R}}\cdot Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg} is closed and hence also open in (GZ(𝔩X)reg)𝔤\bigl{(}G\cdot Z(\mathfrak{l}_{X})^{*}_{\rm reg}\bigr{)}\cap\mathfrak{g}_{\mathbb{R}}^{*}.

Fix 𝔩\mathfrak{l}_{\mathbb{R}}. Suppose first that μ((TX)o)\mu((T^{*}X_{\mathbb{R}})^{o}) intersects GZ(𝔩)G_{\mathbb{R}}\cdot Z(\mathfrak{l}_{\mathbb{R}})^{*}, Then since the rank of μ\mu equals nn everywhere on TX(TX)oT^{*}X_{\mathbb{R}}\cap(T^{*}X)^{o}, we have

dim(μ((TX)o)GZ(𝔩)reg)=n.\dim_{\mathbb{R}}\bigl{(}\mu((T^{*}X_{\mathbb{R}})^{o})\cap G_{\mathbb{R}}\cdot Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}\bigr{)}=n.

By μ((TX)o)¯=μ(TX)¯\overline{\mu((T^{*}X_{\mathbb{R}})^{o})}=\overline{\mu(T^{*}X_{\mathbb{R}})}, we have dim(μ(TX)¯GZ(𝔩)reg)=n.\dim_{\mathbb{R}}\bigl{(}\overline{\mu(T^{*}X_{\mathbb{R}})}\cap G_{\mathbb{R}}\cdot Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}\bigr{)}=n. Since μ(TX)¯\overline{\mu(T^{*}X_{\mathbb{R}})} is GG_{\mathbb{R}}-stable, dim(μ(TX)¯Z(𝔩)reg)=dim𝔞X\dim_{\mathbb{R}}\bigl{(}\overline{\mu(T^{*}X_{\mathbb{R}})}\cap Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}\bigr{)}=\dim_{\mathbb{C}}\mathfrak{a}_{X}^{*}. Hence (1.3) follows.

Suppose next that μ((TX)o)GZ(𝔩)=\mu((T^{*}X_{\mathbb{R}})^{o})\cap G_{\mathbb{R}}\cdot Z(\mathfrak{l}_{\mathbb{R}})^{*}=\emptyset. Then since μ((TX)o)¯=μ(TX)¯\overline{\mu((T^{*}X_{\mathbb{R}})^{o})}=\overline{\mu(T^{*}X_{\mathbb{R}})} and since GZ(𝔩)regG_{\mathbb{R}}\cdot Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg} is open in (8.5), we have

μ(TX)¯GZ(𝔩)reg=μ(TX)¯Z(𝔩)reg=.\overline{\mu(T^{*}X_{\mathbb{R}})}\cap G_{\mathbb{R}}\cdot Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}=\overline{\mu(T^{*}X_{\mathbb{R}})}\cap Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg}=\emptyset.

Finally, as μ((TX)o)\mu((T^{*}X_{\mathbb{R}})^{o}) is nonempty and contained in the set (8.5), μ(TX)¯\overline{\mu(T^{*}X_{\mathbb{R}})} intersects GZ(𝔩)regG_{\mathbb{R}}\cdot Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\rm reg} for at least one 𝔩\mathfrak{l}_{\mathbb{R}}. We finish the proof of Theorem 1.3.


Let us prove Theorem 1.7. There exists a local isomorphism between TX0T^{*}X_{0} and TXT^{*}X_{\mathbb{R}} so we may replace the assumption of Theorem 1.7 by

μ(TX)(𝔤)ell\mu(T^{*}X_{\mathbb{R}})\cap(\mathfrak{g}_{\mathbb{R}}^{*})_{\mathrm{ell}} contains a nonempty open subset of μ(TX)\mu(T^{*}X_{\mathbb{R}}).

Let us assume this. Take a nonempty open subset UTXU\subset T^{*}X_{\mathbb{R}} such that μ(U)(𝔤)ell\mu(U)\subset(\mathfrak{g}_{\mathbb{R}}^{*})_{\mathrm{ell}}. Define (TX)o(T^{*}X)^{o} and (TX)o(T^{*}X_{\mathbb{R}})^{o} as in the proof of Theorem 1.3 above. Since (TX)o(T^{*}X_{\mathbb{R}})^{o} is open and dense in TXT^{*}X_{\mathbb{R}}, we may assume U(TX)oU\subset(T^{*}X_{\mathbb{R}})^{o}. Then by shrinking UU if necessary, we may further assume that μ(U)\mu(U) is a real submanifold of (GZ(𝔩X)reg)𝔤\bigl{(}G\cdot Z(\mathfrak{l}_{X})^{*}_{\rm reg}\bigr{)}\cap\mathfrak{g}_{\mathbb{R}}^{*} of dimension n:=dimμ(TX)n:=\dim_{\mathbb{C}}\mu(T^{*}X). On the other hand, μ(U)G𝔞X¯𝔤\mu(U)\subset\overline{G\cdot\mathfrak{a}_{X}^{*}}\cap\mathfrak{g}_{\mathbb{R}}^{*} by Theorem 1.2. Since (G𝔞X)𝔤(G\cdot\mathfrak{a}_{X}^{*})\cap\mathfrak{g}_{\mathbb{R}}^{*} is a semialgebraic set of (real) dimension nn, we can find a vector ξ(0)μ(U)\xi(\neq 0)\in\mu(U) and an open neighborhood VV of ξ\xi in 𝔤\mathfrak{g}_{\mathbb{R}}^{*} such that

Vμ(U)=V(G𝔞X)𝔤.V\cap\mu(U)=V\cap(G\cdot\mathfrak{a}_{X}^{*})\cap\mathfrak{g}_{\mathbb{R}}^{*}.

By our assumption, the left hand side is contained in (𝔤)ell(\mathfrak{g}_{\mathbb{R}}^{*})_{\mathrm{ell}}. If we put V~:=1>0V\tilde{V}:=\sqrt{-1}\mathbb{R}_{>0}\cdot V, then V~\tilde{V} is an open cone containing 1ξ\sqrt{-1}\xi and

V~(G𝔞X)1𝔤1(𝔤)ell.\tilde{V}\cap(G\cdot\mathfrak{a}_{X}^{*})\cap\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*}\subset\sqrt{-1}(\mathfrak{g}_{\mathbb{R}}^{*})_{\mathrm{ell}}.

Since 1ξ1μ(TX0)¯(GZ(𝔩X)reg)\sqrt{-1}\xi\in\overline{\sqrt{-1}\mu(T^{*}X_{0})}\cap(G\cdot Z(\mathfrak{l}_{X})^{*}_{\mathrm{reg}}), Theorem 1.4 yields

1ξAC(π(𝒪,Γ)suppL2(X0)(G𝒪)𝔞X𝒪).\sqrt{-1}\xi\in\operatorname{AC}\Biggl{(}\bigcup_{\begin{subarray}{c}\pi(\mathcal{O},\Gamma)\in\operatorname{supp}L^{2}(X_{0})\\ (G\cdot\mathcal{O})\cap\mathfrak{a}_{X}^{*}\neq\emptyset\end{subarray}}\mathcal{O}\Biggr{)}.

Hence there exist infinitely many semisimple orbital parameter (𝒪j,Γj)(\mathcal{O}_{j},\Gamma_{j}) and λj𝒪j\lambda_{j}\in\mathcal{O}_{j} such that

π(𝒪j,Γj)suppL2(X0),λj|λj|1ξ|1ξ|(j) and\displaystyle\pi(\mathcal{O}_{j},\Gamma_{j})\in\operatorname{supp}L^{2}(X_{0}),\quad\frac{\lambda_{j}}{|\lambda_{j}|}\to\frac{\sqrt{-1}\xi}{|\sqrt{-1}\xi|}\ (j\to\infty)\text{ and }
(G𝒪j)𝔞XZ(𝔩X)reg.\displaystyle(G\cdot\mathcal{O}_{j})\cap\mathfrak{a}_{X}^{*}\cap Z(\mathfrak{l}_{X})^{*}_{\rm reg}\neq\emptyset.

For large enough jj, we have λjV~\lambda_{j}\in\tilde{V} and then λj1(𝔤)ell\lambda_{j}\in\sqrt{-1}(\mathfrak{g}_{\mathbb{R}})_{\mathrm{ell}}. Moreover, it is easy to see that π(𝒪j,Γj)\pi(\mathcal{O}_{j},\Gamma_{j}) is an isolated point in the set

{π(𝒪,Γ)(G𝒪)𝔞X}\{\pi(\mathcal{O},\Gamma)\mid(G\cdot\mathcal{O})\cap\mathfrak{a}_{X}^{*}\neq\emptyset\}

with respect to the Fell topology. Hence it is an isolated point in suppL2(X0)\operatorname{supp}L^{2}(X_{0}). As a consequence, π(𝒪j,Γj)\pi(\mathcal{O}_{j},\Gamma_{j}) appears in the discrete spectrum of the decomposition in L2(X0)L^{2}(X_{0}) for large jj and therefore L2(X0)L^{2}(X_{0}) has infinitely many discrete series. This completes the proof of Theorem 1.7.

9. Examples

9.1. GL(n,)/(GL(m,)×GL(k,))\operatorname{GL}(n,\mathbb{R})/(\operatorname{GL}(m,\mathbb{R})\times\operatorname{GL}(k,\mathbb{Z}))

Let X0=GL(n,)/(GL(m,)×GL(k,))X_{0}=\operatorname{GL}(n,\mathbb{R})/(\operatorname{GL}(m,\mathbb{R})\times\operatorname{GL}(k,\mathbb{Z})) for m+knm+k\leq n, where GL(m,)×GL(k,)\operatorname{GL}(m,\mathbb{R})\times\operatorname{GL}(k,\mathbb{Z}) is embedded as a subgroup of GL(n,)\operatorname{GL}(n,\mathbb{R}) in a standard way. Below, we compute the image of the real moment map μ(1TX0)\mu(\sqrt{-1}T^{*}X_{0}) for every n,m,kn,m,k. Combining this calculation with Theorem 1.3, we obtain the asymptotic support of Plancherel measure for X0X_{0}. The discrete group part GL(k,)\operatorname{GL}(k,\mathbb{Z}) does not affect the moment map image or the asymptotic support of Plancherel measure.

Proposition 9.1.

Let X0=GL(n,)/(GL(m,)×GL(k,))X_{0}=\operatorname{GL}(n,\mathbb{R})/(\operatorname{GL}(m,\mathbb{R})\times\operatorname{GL}(k,\mathbb{Z})).

  1. (i)

    If 2mn2m\leq n, then μ(1TX0)\mu(\sqrt{-1}T^{*}X_{0}) contains a Zariski open dense subset of 1𝔤𝔩(n,)\sqrt{-1}\mathfrak{gl}(n,\mathbb{R})^{*}. If 𝔧𝔤𝔩(n,)\mathfrak{j}_{\mathbb{R}}\subset\mathfrak{gl}(n,\mathbb{R}) is a Cartan subalgebra, then

    AC({λ1(𝔧)regπ(𝔧,Γλ)suppL2(X0)})=1𝔧.\operatorname{AC}\bigl{(}\bigl{\{}\lambda\in\sqrt{-1}(\mathfrak{j}_{\mathbb{R}})_{\operatorname{reg}}^{*}\mid\pi(\mathfrak{j}_{\mathbb{R}},\Gamma_{\lambda})\in\operatorname{supp}L^{2}(X_{0})\bigr{\}}\bigr{)}=\sqrt{-1}\mathfrak{j}_{\mathbb{R}}^{*}.

    In particular, suppL2(X0)\operatorname{supp}L^{2}(X_{0}) “asymptotically contains the entire tempered dual of GL(n,)\operatorname{GL}(n,\mathbb{R})”.

  2. (ii)

    If 2m>n2m>n, form the Levi subgroup

    L:=GL(1,)×(2n2m)×GL(2mn,)L:=\operatorname{GL}(1,\mathbb{C})^{\times(2n-2m)}\times\operatorname{GL}(2m-n,\mathbb{C})

    with Lie algebra 𝔩\mathfrak{l}. Let 𝔩𝔩\mathfrak{l}_{\mathbb{R}}\subset\mathfrak{l} be a real form contained in 𝔤𝔩(n,)\mathfrak{gl}(n,\mathbb{R}), and identify 1Z(𝔩)Z(𝔩)\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}\simeq Z(\mathfrak{l}_{\mathbb{R}}) by dividing by 1\sqrt{-1} and using the trace form. Let Z(𝔩)0Z(\mathfrak{l}_{\mathbb{R}})_{0} denote the set of matrices X0Z(𝔩)X_{0}\in Z(\mathfrak{l}_{\mathbb{R}}) with

    rankX02n2m,\operatorname{rank}X_{0}\leq 2n-2m,

    and let

    1Z(𝔩)0,reg1Z(𝔩)\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{0,\operatorname{reg}}\subset\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}

    denote the set of regular elements in the corresponding subset of 1Z(𝔩)\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}. Then

    AC({λ1Z(𝔩)grπ(𝔩,Γλ)suppL2(X0)})1Z(𝔩)reg=1Z(𝔩)0,reg.\operatorname{AC}\bigl{(}\bigl{\{}\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})_{\operatorname{gr}}^{*}\mid\pi(\mathfrak{l}_{\mathbb{R}},\Gamma_{\lambda})\in\operatorname{supp}L^{2}(X_{0})\bigr{\}}\bigr{)}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{\operatorname{reg}}\\ =\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}})^{*}_{0,\operatorname{reg}}.
Remark 9.2.

In the case (ii), real forms of LL are of the form

Ls=GL(1,)×s×GL(1,)×2(nms)×GL(2mn,)L_{\mathbb{R}}^{s}=\operatorname{GL}(1,\mathbb{C})^{\times s}\times\operatorname{GL}(1,\mathbb{R})^{\times 2(n-m-s)}\times\operatorname{GL}(2m-n,\mathbb{R})

with 0snm0\leq s\leq n-m. For fixed ss, we may form the larger real Levi subgroup

L~s=GL(2,)×s×GL(1,)×2(nms)×GL(2mn,).\widetilde{L_{\mathbb{R}}}^{s}=\operatorname{GL}(2,\mathbb{R})^{\times s}\times\operatorname{GL}(1,\mathbb{R})^{\times 2(n-m-s)}\times\operatorname{GL}(2m-n,\mathbb{R}).

Take a representation of the form

(9.1) σ1σsτ1τ2(nms)τν\sigma_{1}\boxtimes\cdots\boxtimes\sigma_{s}\boxtimes\tau_{1}\boxtimes\cdots\boxtimes\tau_{2(n-m-s)}\boxtimes\tau_{\nu}

where τi\tau_{i} (1i2(nms))(1\leq i\leq 2(n-m-s)) and τν\tau_{\nu} are one-dimensional unitary representations and σi\sigma_{i} are relative discrete series representations. If PsP_{\mathbb{R}}^{s} is a real parabolic with Levi factor L~s\widetilde{L}_{\mathbb{R}}^{s}, then the representations π(𝔩s,Γλ)\pi(\mathfrak{l}_{\mathbb{R}}^{s},\Gamma_{\lambda}) with λZ(𝔩s)gr\lambda\in Z(\mathfrak{l}_{\mathbb{R}}^{s})^{*}_{\operatorname{gr}} are obtained by unitary parabolic induction from PsP_{\mathbb{R}}^{s}-representations of the form (9.1) to GL(n,)\operatorname{GL}(n,\mathbb{R}).

When 2mn>12m-n>1, the condition λ1Z(𝔩s)0,reg\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}}^{s})^{*}_{0,\operatorname{reg}} implies that λ\lambda vanishes on the last component 𝔤𝔩(2mn,)\mathfrak{gl}(2m-n,\mathbb{R}) of 𝔩s\mathfrak{l}_{\mathbb{R}}^{s} and hence τν\tau_{\nu} is trivial on the identity component of GL(2mn,)\operatorname{GL}(2m-n,\mathbb{R}).

We remark that when k=0k=0, according to a result of Benoist-Kobayashi [BK15], L2(X0)L^{2}(X_{0}) is tempered if and only if 2mn+12m\leq n+1.

Proof.

First, we prove part (i). Assume nn even, and put p:=n2p:=\frac{n}{2}. Consider the set p\mathcal{F}_{p} consisting of all matrices of the following form:

A=(a100b1000a200b2000ap00bp100000010000001000).A=\begin{pmatrix}a_{1}&0&\cdots&0&b_{1}&0&\cdots&0\\ 0&a_{2}&\cdots&0&0&b_{2}&\cdots&0\\ \vdots&\vdots&\ddots&\vdots&\vdots&\vdots&\ddots&\vdots\\ 0&0&\cdots&a_{p}&0&0&\cdots&b_{p}\\ 1&0&\cdots&0&0&0&\cdots&0\\ 0&1&\cdots&0&0&0&\cdots&0\\ \vdots&\vdots&\ddots&\vdots&\vdots&\vdots&\ddots&\vdots\\ 0&0&\cdots&1&0&0&\cdots&0\end{pmatrix}.

Each matrix AA has ss submatrices of the form

Aj=(ajbj10).A_{j}=\begin{pmatrix}a_{j}&b_{j}\\ 1&0\end{pmatrix}.

We note that the 2n2n eigenvalues of a matrix ApA\in\mathcal{F}_{p} is simply the union of the eigenvalues of these nn two by two matrices AjA_{j}. Now, for fixed eigenvalues λ1\lambda_{1}, λ2\lambda_{2} with either (a) λ1\lambda_{1}, λ2\lambda_{2} both real or (b) λ1¯=λ2\overline{\lambda_{1}}=\lambda_{2}, we may choose aja_{j} and bjb_{j} such that AjA_{j} has the eigenvalues λ1\lambda_{1}, λ2\lambda_{2} by setting aj:=λ1+λ2a_{j}:=\lambda_{1}+\lambda_{2} and bj:=λ1λ2b_{j}:=-\lambda_{1}\lambda_{2}. After identifying 1𝔤𝔩(n,)𝔤𝔩(n,)\sqrt{-1}\mathfrak{gl}(n,\mathbb{R})^{*}\simeq\mathfrak{gl}(n,\mathbb{R}), notice that all of the above matrices ApA\in\mathcal{F}_{p} lie in 𝔤𝔩(m,)μ(1L2(X0))\mathfrak{gl}(m,\mathbb{R})^{\perp}\subset\mu(\sqrt{-1}L^{2}(X_{0})). Since two \mathbb{C}-diagonalizable matrices in 𝔤𝔩(n,)\mathfrak{gl}(n,\mathbb{R}) are GL(n,)\operatorname{GL}(n,\mathbb{R})-conjugate if, and only if they have the same eigenvalues, part (i) follows in the case nn even. When nn odd, add an extra row and column to every ApA\in\mathcal{F}_{p} with all zeroes except for the desired nnth (real) eigenvalue in the diagonal entry.

For part (ii), put p:=nmp:=n-m and note 0p<n/20\leq p<n/2. Take complex numbers {λ11,λ12,λ21,λ22,,λp1,λp2}\{\lambda_{11},\lambda_{12},\lambda_{21},\lambda_{22},\ldots,\lambda_{p1},\lambda_{p2}\} such that either (a) λk1\lambda_{k1}, λk2\lambda_{k2} both real or (b) λ¯k1=λk2\overline{\lambda}_{k1}=\lambda_{k2}. As before, we can find a 2p2p by 2p2p matrix in p\mathcal{F}_{p} with these specified eigenvalues. Adding extra rows and columns to get a 2n2n by 2n2n matrix AA. When the 2p2p eigenvalues are distinct, we see that the stabilizer of AA for the adjoint action of GL(n,)\operatorname{GL}(n,\mathbb{C}) is isomorphic to

L:=GL(1,)×(2n2m)×GL(2mn,).L:=\operatorname{GL}(1,\mathbb{C})^{\times(2n-2m)}\times\operatorname{GL}(2m-n,\mathbb{C}).

Further, we see that for every real form 𝔩\mathfrak{l}_{\mathbb{R}} of 𝔩\mathfrak{l} (as described in the remark above), every element of Z(𝔩)0Z(\mathfrak{l}_{\mathbb{R}})_{0} is a conjugate of a matrix of the form AA as above.

Next, recall μ(1TX0)=Ad(GL(n,))𝔤𝔩(m,)\mu(\sqrt{-1}T^{*}X_{0})=\operatorname{Ad}^{*}(\operatorname{GL}(n,\mathbb{R}))\cdot\mathfrak{gl}(m,\mathbb{R})^{\perp}. We see that every B𝔤𝔩(m,)𝔤𝔩(n,)B\in\mathfrak{gl}(m,\mathbb{R})^{\perp}\subset\mathfrak{gl}(n,\mathbb{R}) with zeroes in an m×mm\times m block in the bottom right is a sum of a matrix B1B_{1} with mm zero columns and a matrix B2B_{2} with mm zero rows. In particular, rankB2n2m\operatorname{rank}B\leq 2n-2m. It follows that 𝔩\mathfrak{l} is the Levi subalgebra 𝔩X\mathfrak{l}_{X} in Theorem 1.2 for X=GL(2n)/GL(2m)X=\operatorname{GL}(2n)/\operatorname{GL}(2m) and that the closure of the conjugates of the matrices of the form BB intersected with Z(𝔩)Z(\mathfrak{l}_{\mathbb{R}}) constitute Z(𝔩)0Z(\mathfrak{l}_{\mathbb{R}})_{0}. Part (ii) follows. ∎

9.2. Sp(2n,)/(Sp(2m,)×Sp(2k,))\operatorname{Sp}(2n,\mathbb{R})/(\operatorname{Sp}(2m,\mathbb{R})\times\operatorname{Sp}(2k,\mathbb{Z}))

Similarly to the previous subsection, we calculate the moment map image for X0=Sp(2n,)/(Sp(2m,)×Sp(2k,))X_{0}=\operatorname{Sp}(2n,\mathbb{R})/(\operatorname{Sp}(2m,\mathbb{R})\times\operatorname{Sp}(2k,\mathbb{Z})) with m+knm+k\leq n, where Sp(2m,)×Sp(2k,)\operatorname{Sp}(2m,\mathbb{R})\times\operatorname{Sp}(2k,\mathbb{Z}) is embedded as a subgroup of Sp(2n,)\operatorname{Sp}(2n,\mathbb{R}) in a standard way.

Let G=Sp(2n,)G_{\mathbb{R}}=\operatorname{Sp}(2n,\mathbb{R}) and H=Sp(2m,)H_{\mathbb{R}}=\operatorname{Sp}(2m,\mathbb{R}). Let V=2nV=\mathbb{R}^{2n} with a symplectic form (,)(\cdot,\cdot). Then we identify GG_{\mathbb{R}} with the automorphism group of (V,(,))(V,(\cdot,\cdot)). The Lie algebra 𝔤\mathfrak{g}_{\mathbb{R}} consists of A𝔤𝔩(V)A\in\mathfrak{gl}(V) satisfying

Av1,v2+v1,Av2=0.\langle Av_{1},v_{2}\rangle+\langle v_{1},Av_{2}\rangle=0.

For A𝔤A\in\mathfrak{g}_{\mathbb{R}}, define a bilinear form (,)A(\cdot,\cdot)_{A} on VV by

(v1,v2)A:=Av1,v2.(v_{1},v_{2})_{A}:=\langle Av_{1},v_{2}\rangle.

This form is symmetric and hence its signature (p,q)=sign(,)A(p,q)=\operatorname{sign}(\cdot,\cdot)_{A} is defined. Write sign(A):=sign(,)A\operatorname{sign}(A):=\operatorname{sign}(\cdot,\cdot)_{A}.

Let V=WWV=W\oplus W^{\prime} be an orthogonal decomposition into symplectic vector spaces with dimW=2m\dim W=2m. Let

𝔥:={A𝔤A(W)W,A(W)=0}𝔰𝔭(2m,).\mathfrak{h}_{\mathbb{R}}:=\{A\in\mathfrak{g}_{\mathbb{R}}\mid A(W)\subset W,\ A(W^{\prime})=0\}\simeq\mathfrak{sp}(2m,\mathbb{R}).

Then

𝔥={A𝔤A(W),W=0}.\mathfrak{h}_{\mathbb{R}}^{\perp}=\{A\in\mathfrak{g}_{\mathbb{R}}\mid\langle A(W),W\rangle=0\}.

Here and in what follows, we identify 𝔤\mathfrak{g}_{\mathbb{R}} with 𝔤\mathfrak{g}_{\mathbb{R}}^{*} by an invariant form.

Lemma 9.3.

Let A𝔤A\in\mathfrak{g}_{\mathbb{R}}. Then AG𝔥A\in G_{\mathbb{R}}\cdot\mathfrak{h}^{\perp}_{\mathbb{R}} if and only if there exists a 2m2m-dimensional subspace W1VW_{1}\subset V such that ,|W1\langle\cdot,\cdot\rangle|_{W_{1}} is nondegenerate and (,)A|W1=0(\cdot,\cdot)_{A}|_{W_{1}}=0.

Proof.

If Ag𝔥A\in g\cdot\mathfrak{h}_{\mathbb{R}}^{\perp}, then W1=gWW_{1}=g\cdot W satisfies the condition.

Conversely, suppose W1W_{1} satisfies the condition in the lemma. Then standard symplectic bases of W1W_{1} and WW can be extended to a standard symplectic basis of VV, respectively. Hence we can find gGg\in G_{\mathbb{R}} such that gW=W1g\cdot W=W_{1} and then we have Ag𝔥A\in g\cdot\mathfrak{h}_{\mathbb{R}}^{\perp}. ∎

For semisimple AA, this condition is characterized by sign(A)\operatorname{sign}(A).

Lemma 9.4.

Suppose that A𝔤A\in\mathfrak{g}_{\mathbb{R}} is semisimple and let sign(A)=(p,q)\operatorname{sign}(A)=(p,q). Then the following two conditions are equivalent.

  1. (1)

    There exists a 2m2m-dimensional subspace W1VW_{1}\subset V such that ,|W1\langle\cdot,\cdot\rangle|_{W_{1}} is nondegenerate and (,)A|W1=0(\cdot,\cdot)_{A}|_{W_{1}}=0.

  2. (2)

    max{p,q}2n2m\max\{p,q\}\leq 2n-2m.

Proof.

It is easy to see that the maximal isotropic subspace of VV with respect to the symmetric form (,)A(\cdot,\cdot)_{A}, which has signature (p,q)(p,q), is 2nmax{p,q}2n-\max\{p,q\}. Hence (1) implies (2).

We now prove the other implication. Since V=Im(A)Ker(A)V=\operatorname{Im}(A)\oplus\operatorname{Ker}(A), by considering A|Im(A)A|_{\operatorname{Im}(A)}, our claim is reduced to the case when Im(A)=V\operatorname{Im}(A)=V. Thus we assume rankA=2n\operatorname{rank}A=2n.

Since AA is semisimple, we can find an orthogonal decomposition V=iViV=\bigoplus_{i}V_{i} as a symplectic vector space such that A(Vi)=ViA(V_{i})=V_{i} and dimVi=2\dim V_{i}=2 or 44. This follows from the classification of Cartan subalgebras of 𝔰𝔭(2n,)\mathfrak{sp}(2n,\mathbb{R}). See [Sug59, §3, Type (CI)] for such a classification result. Let Ai:=A|ViA_{i}:=A|_{V_{i}} so that AiA_{i} is regarded as an element in 𝔰𝔭(2,)\mathfrak{sp}(2,\mathbb{R}) or 𝔰𝔭(4,)\mathfrak{sp}(4,\mathbb{R}).

When dimVi=4\dim V_{i}=4, we may assume that it cannot decompose into two AiA_{i}-stable 22-dimensional symplectic vector spaces. Then sign(Ai)=(2,2)\operatorname{sign}(A_{i})=(2,2). In this case, there exists a 22-dimensional subspace WiViW_{i}\subset V_{i} such that ,\langle\cdot,\cdot\rangle is nondegenerate and (,)A=0(\cdot,\cdot)_{A}=0 on WiW_{i}.

When dimVi=dimVi=2\dim V_{i}=\dim V_{i^{\prime}}=2 and sign(Ai)=sign(Ai)=(1,1)\operatorname{sign}(A_{i})=\operatorname{sign}(A_{i^{\prime}})=(1,1) with iii\neq i^{\prime}, there exists a 22-dimensional subspace WiViViW_{i}\subset V_{i}\oplus V_{i^{\prime}} such that ,\langle\cdot,\cdot\rangle is nondegenerate and (,)A=0(\cdot,\cdot)_{A}=0 on WiW_{i}.

Similarly, when dimVi=dimVi=2\dim V_{i}=\dim V_{i^{\prime}}=2, sign(Ai)=(2,0)\operatorname{sign}(A_{i})=(2,0) and sign(Ai)=(0,2)\operatorname{sign}(A_{i^{\prime}})=(0,2), there exists a 22-dimensional subspace WiViViW_{i}\subset V_{i}\oplus V_{i^{\prime}} satisfying the same conditions.

Making appropriate pairs among ViV_{i} and taking sum of above WiW_{i}, we obtain WW in (1). ∎

For complex Lie algebras 𝔤𝔥\mathfrak{g}\supset\mathfrak{h} analogues of Lemmas 9.3 and 9.4 are proved in a similar and easier way. We have for a semisimple element A𝔤A\in\mathfrak{g}

(9.2) AG𝔥rankA4n4m,A\in G\cdot\mathfrak{h}^{\perp}\Leftrightarrow\operatorname{rank}A\leq 4n-4m,

where rankA\operatorname{rank}A is the rank of AA viewed as a 2n2n by 2n2n matrix with complex entries.

For 0rn0\leq r\leq n, let

Lr:=GL(1,)×r×Sp(2(nr),),L^{r}:=\operatorname{GL}(1,\mathbb{C})^{\times r}\times\operatorname{Sp}(2(n-r),\mathbb{C}),

the Levi subgroup of Sp(2n,)\operatorname{Sp}(2n,\mathbb{C}). By (9.2), the Levi subalgebra 𝔩X\mathfrak{l}_{X} in Theorem 1.2 for X=G/HX=G/H is a Cartan algebra if 2mn2m\leq n; and 𝔩2(nm)\mathfrak{l}^{2(n-m)} if 2m>n2m>n. For s,t,u0s,t,u\geq 0 with s+2t+uns+2t+u\leq n, let

Ls,t,u=U(1)×s×GL(1,)×t×GL(1,)×u×Sp(2(ns2tu),).L_{\mathbb{R}}^{s,t,u}=\operatorname{U}(1)^{\times s}\times\operatorname{GL}(1,\mathbb{C})^{\times t}\times\operatorname{GL}(1,\mathbb{R})^{\times u}\times\operatorname{Sp}(2(n-s-2t-u),\mathbb{R}).

Then Ls,t,uL_{\mathbb{R}}^{s,t,u} with s+2t+u=rs+2t+u=r are all the real Levi subgroups of Sp(2n,)\operatorname{Sp}(2n,\mathbb{R}) whose complexifications are conjugate to LrL^{r}. In particular, Ls,t,uL_{\mathbb{R}}^{s,t,u} for s+2t+u=ns+2t+u=n are all the Cartan subgroups of Sp(2n,)\operatorname{Sp}(2n,\mathbb{R}) up to conjugation.

For fixed s,t,us,t,u, we may form the larger real Levi subgroup

L~s,t,u=GL(2,)×t×GL(1,)×u×Sp(2(n2tu),).\widetilde{L}_{\mathbb{R}}^{s,t,u}=\operatorname{GL}(2,\mathbb{R})^{\times t}\times\operatorname{GL}(1,\mathbb{R})^{\times u}\times\operatorname{Sp}(2(n-2t-u),\mathbb{R}).

Take a representation of the form

(9.3) σ1σtτ1τ2(nms)κ\sigma_{1}\boxtimes\cdots\boxtimes\sigma_{t}\boxtimes\tau_{1}\boxtimes\cdots\boxtimes\tau_{2(n-m-s)}\boxtimes\kappa

where τi\tau_{i} are one-dimensional unitary representations, σi\sigma_{i} are relative discrete series representations, and κ\kappa is (a Hilbert completion of) A𝔮(λ)A_{\mathfrak{q}}(\lambda) such that the Levi factor of 𝔮\mathfrak{q} is the complexification of U(1)×s×Sp(2(ns2tu),)\operatorname{U}(1)^{\times s}\times\operatorname{Sp}(2(n-s-2t-u),\mathbb{R}). If Ps,t,uP_{\mathbb{R}}^{s,t,u} is a real parabolic with Levi factor L~s,t,u\widetilde{L}_{\mathbb{R}}^{s,t,u}, then the representations π(𝔩s,t,u,Γλ)\pi(\mathfrak{l}_{\mathbb{R}}^{s,t,u},\Gamma_{\lambda}) with λ1Z(𝔩s,t,u)gr\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}}^{s,t,u})^{*}_{\operatorname{gr}} are obtained by unitary parabolic induction from Ps,t,uP_{\mathbb{R}}^{s,t,u}-representations of the form (9.3) to Sp(2n,)\operatorname{Sp}(2n,\mathbb{R}).

Let λ1Z(𝔩s,t,u)\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}}^{s,t,u})^{*}. It has ss parameters corresponding to the first component U(1)×s\operatorname{U}(1)^{\times s}, which we denote by (a1,,as)s(a_{1},\dots,a_{s})\in\mathbb{R}^{s}. If λ1Z(𝔩s,t,u)reg\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}}^{s,t,u})^{*}_{\operatorname{reg}}, then a1,,asa_{1},\dots,a_{s} are nonzero; and if one has a representation π(𝔩s,t,u,Γλ)\pi(\mathfrak{l}_{\mathbb{R}}^{s,t,u},\Gamma_{\lambda}), then a1,,asa_{1},\dots,a_{s} are integers. For nonnegative integers s1,s2s_{1},s_{2}, write

1Z(𝔩(s1,s2),t,u)\displaystyle\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}}^{(s_{1},s_{2}),t,u})^{*}
:={λ1Z(𝔩s,t,u)#{iai>0}=s1 and #{iai<0}=s2}\displaystyle:=\bigl{\{}\lambda\in\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}}^{s,t,u})^{*}\mid\#\{i\mid a_{i}>0\}=s_{1}\text{ and }\#\{i\mid a_{i}<0\}=s_{2}\bigr{\}}

Suppose that among ss parameters (a1,,as)(a_{1},\dots,a_{s}), s1s_{1} of them are positive and s2s_{2} of them are negative. If we regard 1λZ(𝔩(s1,s2),t,u)𝔤\sqrt{-1}\lambda\in Z(\mathfrak{l}_{\mathbb{R}}^{(s_{1},s_{2}),t,u})^{*}\subset\mathfrak{g}_{\mathbb{R}}^{*} as an element in 𝔤\mathfrak{g}_{\mathbb{R}}, the signature of 1λ\sqrt{-1}\lambda defined above is (2s1+2t+u,2s2+2t+u)(2s_{1}+2t+u,2s_{2}+2t+u) when we suitably fix a parameterization of characters of U(1)\operatorname{U}(1). We have a decomposition

1Z(𝔩s,t,u)reg=s1+s2=s1Z(𝔩(s1,s2),t,u)1Z(𝔩s,t,u)reg.\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}}^{s,t,u})^{*}_{\operatorname{reg}}=\bigcup_{s_{1}+s_{2}=s}\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}}^{(s_{1},s_{2}),t,u})^{*}\cap\sqrt{-1}Z(\mathfrak{l}_{\mathbb{R}}^{s,t,u})^{*}_{\operatorname{reg}}.

Summing up above arguments and by Theorem 1.3, we have the following.

Proposition 9.5.

Let X0=Sp(2n,)/(Sp(2m,)×Sp(2k,))X_{0}=\operatorname{Sp}(2n,\mathbb{R})/(\operatorname{Sp}(2m,\mathbb{R})\times\operatorname{Sp}(2k,\mathbb{Z})).

  1. (i)

    If 2mn2m\leq n, then μ(1TX0)\mu(\sqrt{-1}T^{*}X_{0}) intersects the set of regular semisimple elements in 1𝔤\sqrt{-1}\mathfrak{g}_{\mathbb{R}}^{*}. Take a Cartan subalgebra

    𝔧=𝔩s,t,u=𝔲(1)s𝔤𝔩(1,)t𝔤𝔩(1,)u,\mathfrak{j}_{\mathbb{R}}=\mathfrak{l}^{s,t,u}_{\mathbb{R}}=\mathfrak{u}(1)^{\oplus s}\oplus\mathfrak{gl}(1,\mathbb{C})^{\oplus t}\oplus\mathfrak{gl}(1,\mathbb{R})^{\oplus u},

    where s+2t+u=ns+2t+u=n. Then

    AC({λ1(𝔧)regπ(𝔧,Γλ)suppL2(X0)})1(𝔧)reg=s11(𝔩(s1,ss1),t,u)reg,\operatorname{AC}\bigl{(}\bigl{\{}\lambda\in\sqrt{-1}(\mathfrak{j}_{\mathbb{R}})_{\operatorname{reg}}^{*}\mid\pi(\mathfrak{j}_{\mathbb{R}},\Gamma_{\lambda})\in\operatorname{supp}L^{2}(X_{0})\bigr{\}}\bigr{)}\cap\sqrt{-1}(\mathfrak{j}_{\mathbb{R}})_{\operatorname{reg}}^{*}\\ =\bigcup_{s_{1}}\sqrt{-1}(\mathfrak{l}^{(s_{1},s-s_{1}),t,u}_{\mathbb{R}})^{*}_{\operatorname{reg}},

    where s1s_{1} runs over nonnegative integers satisfying

    2mn+s2s1n2m+s2.\frac{2m-n+s}{2}\leq s_{1}\leq\frac{n-2m+s}{2}.
  2. (ii)

    If 2m>n2m>n, then take a Levi subalgebra

    𝔩s,t,u=𝔲(1)s𝔤𝔩(1,)t𝔤𝔩(1,)u𝔰𝔭(2(ns2tu),)\mathfrak{l}_{\mathbb{R}}^{s,t,u}=\mathfrak{u}(1)^{\oplus s}\oplus\mathfrak{gl}(1,\mathbb{C})^{\oplus t}\oplus\mathfrak{gl}(1,\mathbb{R})^{\oplus u}\oplus\mathfrak{sp}(2(n-s-2t-u),\mathbb{R})

    for nonnegative integers s,t,us,t,u such that s+2t+u=2(nm)s+2t+u=2(n-m). Then

    AC({λ1(𝔩s,t,u)regπ(𝔩s,t,u,Γλ)suppL2(X0)})1(𝔩s,t,u)reg\operatorname{AC}\bigl{(}\bigl{\{}\lambda\in\sqrt{-1}(\mathfrak{l}_{\mathbb{R}}^{s,t,u})_{\operatorname{reg}}^{*}\mid\pi(\mathfrak{l}_{\mathbb{R}}^{s,t,u},\Gamma_{\lambda})\in\operatorname{supp}L^{2}(X_{0})\bigr{\}}\bigr{)}\cap\sqrt{-1}(\mathfrak{l}_{\mathbb{R}}^{s,t,u})_{\operatorname{reg}}^{*}

    equals 1(𝔩(s2,s2),t,u)reg\sqrt{-1}(\mathfrak{l}^{(\frac{s}{2},\frac{s}{2}),t,u}_{\mathbb{R}})^{*}_{\operatorname{reg}} if ss is even; and empty if ss is odd.

Note that when k=0k=0, L2(X0)L^{2}(X_{0}) is tempered if and only if 2mn2m\leq n by [BK15].

We now deduce which elliptic orbits appear in the image of moment map. Let 𝔱\mathfrak{t} be a Cartan subalgebra of K(GL(n,))K(\simeq\operatorname{GL}(n,\mathbb{C})) and let ϵ1,,ϵn\epsilon_{1},\dots,\epsilon_{n} be a standard basis of 𝔱\mathfrak{t}^{*}. The roots in 𝔨\mathfrak{k} and 𝔤\mathfrak{g} are as follows

Δ(𝔨,𝔱)={ϵiϵj:1i,jn,ij},\displaystyle\Delta(\mathfrak{k},\mathfrak{t})=\{\epsilon_{i}-\epsilon_{j}:1\leq i,j\leq n,\ i\neq j\},
Δ(𝔤,𝔱)={±2ϵi:1in}{±ϵi±ϵj:1i,jn,ij}.\displaystyle\Delta(\mathfrak{g},\mathfrak{t})=\{\pm 2\epsilon_{i}:1\leq i\leq n\}\cup\{\pm\epsilon_{i}\pm\epsilon_{j}:1\leq i,j\leq n,\ i\neq j\}.

Suppose first that n2mn\geq 2m. This case was previously studied in [HW17, Example 7.5]. Then the moment map image μ(TX0)\mu(T^{*}X_{0}) contains a regular semisimple orbit of 𝔤\mathfrak{g}_{\mathbb{R}}^{*}. Suppose AA is regular so that sign(A)=(p,2np)\operatorname{sign}(A)=(p,2n-p) for some pp. By Lemma 9.3 and Lemma 9.4, Aμ(TX0)A\in\mu(T^{*}X_{0}) if and only if 2mp2n2m2m\leq p\leq 2n-2m. If n=2mn=2m, then sign(A)=(n,n)\operatorname{sign}(A)=(n,n) is the only possibility. The Harish-Chandra parameters for discrete series of Sp(2n,)\operatorname{Sp}(2n,\mathbb{R}) are given in terms of standard coordinates as follows:

(9.4) i=1naiϵi with ai and |a1|>|a2|>>|an|>0.\sum_{i=1}^{n}a_{i}\epsilon_{i}\text{ with }a_{i}\in\mathbb{Z}\text{ and }|a_{1}|>|a_{2}|>\cdots>|a_{n}|>0.

If (p,q)(p,q) is the signature for the corresponding orbit, then pp is the number of positives in {a1,,an}\{a_{1},\dots,a_{n}\} and qq is the number of negatives. As a consequence of Theorem 1.7, for any given subset SS of {1,2,,n}\{1,2,\dots,n\} with 2m#S2n2m2m\leq\#S\leq 2n-2m, there exist infinitely many distinct discrete series representations of Sp(2n,)\operatorname{Sp}(2n,\mathbb{R}) which are isomorphic to subrepresentations of L2(X0)L^{2}(X_{0}) and has the Harish-Chandra parameters as (9.4) satisfying {i:ai>0}=S\{i:a_{i}>0\}=S.

Suppose next that n<2mn<2m. Then the maximal rank of AA is 4n4m4n-4m. If rankA=4n4m\operatorname{rank}A=4n-4m, then Aμ(TX0)A\in\mu(T^{*}X_{0}) if and only if (p,q)=(2n2m,2n2m)(p,q)=(2n-2m,2n-2m). Let SS be a subset of {1,,2n2m}\{1,\dots,2n-2m\} such that #S=nm\#S=n-m. Let S:={1,,2n2m}SS^{\prime}:=\{1,\dots,2n-2m\}\setminus S. Let 𝔮S\mathfrak{q}_{S} be a parabolic subalgebra of 𝔤\mathfrak{g} such that the roots of its nilradical 𝔫S\mathfrak{n}_{S} are

Δ(𝔫S,𝔱)\displaystyle\Delta(\mathfrak{n}_{S},\mathfrak{t}) ={ϵi±ϵj:iS,i<j}{2ϵi:iS}\displaystyle=\{\epsilon_{i}\pm\epsilon_{j}:i\in S,\ i<j\}\cup\{2\epsilon_{i}:i\in S\}
{ϵi±ϵj:iS,i<j}{2ϵi:iS}.\displaystyle\cup\{-\epsilon_{i}\pm\epsilon_{j}:i\in S^{\prime},\ i<j\}\cup\{-2\epsilon_{i}:i\in S^{\prime}\}.

The real Levi factor for 𝔮\mathfrak{q} is isomorphic to 𝔲(1)(2n2m)𝔰𝔭(4m2n,)\mathfrak{u}(1)^{\oplus(2n-2m)}\oplus\mathfrak{sp}(4m-2n,\mathbb{R}). The elliptic coadjoint orbits with signature (p,q)=(2n2m,2n2m)(p,q)=(2n-2m,2n-2m) correspond to A𝔮S(λ)A_{\mathfrak{q}_{S}}(\lambda) for some SS as above. Therefore, for any given S{1,,2n2m}S\subset\{1,\dots,2n-2m\} with #S=nm\#S=n-m, there exist infinitely many parameters λ\lambda in the good range such that (Hilbert completions of) A𝔮S(λ)A_{\mathfrak{q}_{S}}(\lambda) occurs as a discrete spectrum of L2(X0)L^{2}(X_{0}).

In particular, we have

Corollary 9.6.

Sp(2n,)/(Sp(2m,)×Sp(2k,))\operatorname{Sp}(2n,\mathbb{R})/(\operatorname{Sp}(2m,\mathbb{R})\times\operatorname{Sp}(2k,\mathbb{Z})) has discrete series for any n,m,kn,m,k with m+knm+k\leq n.

References

  • [AvLTV20] J. Adams, M. van Leeuwen, P. Trapa, and D. Vogan, Unitary representations of real reductive groups, Astérisque (2020), no. 417.
  • [BB82] W. Borho and J.-L. Brylinski, Differential operators on homogeneous spaces I: Irreducibility of the associated variety for annihilators of induced modules, Invent. Math. 69 (1982), 437–472.
  • [BBHM63] A. Białynicki-Birula, G. Hochschild, and G. D. Mostow, Extensions of representations of algebraic linear groups, Amer. J. Math. 85 (1963), 131–144.
  • [BD60] P. Bernat and J. Dixmier, Sur le dual d’un groupe de Lie, C. R. Acad. Sci. Paris 250 (1960), 1778–1779.
  • [Ber88] J. Bernstein, On the support of Plancherel measure, J. Geom. Phys. 5 (1988), no. 4, 663–710.
  • [Bie90] F. Bien, 𝒟\mathscr{D}-modules and spherical representations, Mathematical Notes, vol. 39, Princeton University Press, Princeton, NJ, 1990.
  • [BKa] Y. Benoist and T. Kobayashi, Tempered homogeneous spaces II, To be published in: Dynamics, Geometry, Number Theory: The Impact of Margulis on Modern Mathematics, available at arXiv:1706.10131.
  • [BKb] by same author, Tempered homogeneous spaces IV, arXiv:2009.10391.
  • [BK15] by same author, Tempered reductive homogeneous spaces, J. Eur. Math. Soc. (JEMS) 17 (2015), no. 12, 3015–3036.
  • [BK21] by same author, Tempered homogeneous spaces III, J. Lie Theory 31 (2021), no. 3, 833–869.
  • [BS05] E. van den Ban and H. Schlichtkrull, The Plancherel decomposition for a reductive symmetric space. II. representation theory, Inventiones Mathematicae 161 (2005), no. 3, 567–628.
  • [Del98] P. Delorme, Formule de Plancherel pour les espaces symetriques reductifs, Annals of Mathematics Series 2 147 (1998), 417–452.
  • [Dix77] J. Dixmier, CC^{*}-algebras, North-Holland Mathematical Library, Vol. 15, North-Holland Publishing Co., Amsterdam-New York-Oxford, 1977.
  • [DKKS21] P. Delorme, F. Knop, B. Krötz, and H. Schlichtkrull, Plancherel theory for real spherical spaces: Construction of the Bernstein morphisms, J. Amer. Math. Soc. 34 (2021), no. 3, 815–908.
  • [Duf82] M. Duflo, Construction de représentations unitaires d’un groupe de Lie, Harmonic Analysis and Group Representations, Liguori, Naples, 1982, pp. 129–221.
  • [FJ80] M. Flensted-Jensen, Discrete series for semisimple symmetric spaces, Ann. of Math. (2) 11 (1980), no. 2, 253–311.
  • [Fol89] G.B. Folland, Harmonic analysis in phase space, Princeton University Press, Princeton, 1989.
  • [Fol16] by same author, A course in abstract harmonic analysis, second ed., Textbooks in Mathematics, CRC Press, Boca Raton, FL, 2016.
  • [Har18] B. Harris, Wave front sets of reductive Lie group representations II, Trans. Amer. Math. Soc. 370 (2018), no. 8, 5931–5962.
  • [HHO16] B. Harris, H. He, and G. Ólafsson, Wave front sets of reductive Lie group representations, Duke Math. J. 165 (2016), no. 5, 793–846.
  • [HO20] B. Harris and Y. Oshima, Irreducible characters and semisimple coadjoint orbits, J. Lie Theory 30 (2020), no. 3, 715–765.
  • [Hör83a] L. Hörmander, The analysis of linear partial differential operators I, Grundlehren der mathematischen Wissenschaften, vol. 256, Springer-Verlag, Berlin, 1983.
  • [Hör83b] by same author, The analysis of linear partial differential operators II, Grundlehren der mathematischen Wissenschaften, vol. 257, Springer-Verlag, Berlin, 1983.
  • [How81] R. Howe, Wave front sets of representations of Lie groups, Automorphic Forms, Representation Theory, and Arithmetic, Tata Inst. Fund. Res. Studies in Math., vol. 10, Tata Institute of Fundamental Ressearch, Bombay, 1981, pp. 117–140.
  • [HTT08] R. Hotta, K. Takeuchi, and T. Tanisaki, D-modules, perverse sheaves, and representation theory, Progress in Mathematics, vol. 236, Birkhäuser, Boston, 2008.
  • [HW17] B. Harris and T. Weich, Wave front sets of reductive Lie group representations III, Adv. Math. 313 (2017), 176–236.
  • [Jan83] J.C. Jantzen, Einhüllende Algebren halbeinfacher Lie-Algebren, Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)], vol. 3, Springer-Verlag, Berlin, 1983.
  • [Kir62] A. Kirillov, Unitary representations of nilpotent Lie groups, Uspehi Mat. Nauk 17 (1962), no. 4 (106), 57–110.
  • [Kir04] A. Kirillov, Lectures on the orbit method, vol. 64, American Mathematical Society, Providence, RI, 2004.
  • [Kit12] S. Kitchen, Cohomology of standard modules on partial flag varieties, Represent. Theory 16 (2012), 317–344.
  • [KKOS20] B. Krötz, J. Kuit, E. Opdam, and H. Schlichtkrull, The infinitesimal characters of discrete series for real spherical spaces, Geom. Funct. Anal. 30 (2020), no. 3, 804–857.
  • [Kna86] A. Knapp, Representation theory of semisimple Lie groups, Princeton University Press, Princeton, NJ, 1986.
  • [Kno94] F. Knop, The asymptotic behavior of invariant collective motion, Invent. Math. 116 (1994), 309–328.
  • [KO13] T. Kobayashi and T. Oshima, Finite multiplicity theorems for induction and restriction, Adv. Math. 248 (2013), 921–944.
  • [Kob92] T. Kobayashi, Singular unitary representations and discrete series for indefinite Stiefel manifolds U(p,q;𝔽)/U(pm,q;𝔽)\operatorname{U}(p,q;\mathbb{F})/\operatorname{U}(p-m,q;\mathbb{F}), Mem. Amer. Math. Soc. 95 (1992), no. 462.
  • [Kob94] by same author, Discrete decomposability of the restriction of A𝔮(λ)\operatorname{A}_{\mathfrak{q}}(\lambda) with respect to reductive subgroups and its applications, Invent. Math. 117 (1994), no. 2, 181–205.
  • [Kob98a] by same author, Discrete decomposability of the restriction of A𝔮(λ){A}_{\mathfrak{q}}(\lambda) with respect to reductive subgroups. II. micro-local analysis and asymptotic KK-support, Ann. of Math. (2) 147 (1998), no. 3, 709–729.
  • [Kob98b] by same author, Discrete decomposability of the restriction of A𝔮(λ){A}_{\mathfrak{q}}(\lambda) with respect to reductive subgroups. III. restriction of Harish-Chandra modules and associated varieties, Invent. Math. 131 (1998), no. 2, 229–256.
  • [Kob98c] by same author, Discrete series representations for the orbit space arising from two involutions of real reductive Lie groups, J. Funct. Anal. 152 (1998), no. 1, 100–135.
  • [KV79] M. Kashiwara and M. Vergne, KK-types and singular spectrum, Noncommutative Harmonic Analysis (Proc. Third Colloq., Marseilles-Luminy, 1978), Lecture Notes in Mathematics., vol. 728, Springer Verlag, 1979, pp. 177–200.
  • [KV95] A. Knapp and D. Vogan, Cohomological induction and unitary representations, Princeton Mathematical Series, vol. 45, Princeton University Press, Princeton, 1995.
  • [Li93] J.-S. Li, On the discrete series of generalized stiefel manifolds, Trans. Amer. Math. Soc. 340 (1993), no. 2, 753–766.
  • [Mac76] G. Mackey, The theory of unitary group representations, University of Chicago Press, Chicago, Ill.-London, 1976.
  • [Mat04] H. Matumoto, On the representations of Sp(p,q){\rm Sp}(p,q) and SO(2n){\rm SO}^{*}(2n) unitarily induced from derived functor modules, Compos. Math. 140 (2004), no. 4, 1059–1096.
  • [MO84] T. Matsuki and T. Oshima, A description of discrete series for semisimple symmetric spaces, Group Representations and Systems of Differential Equations (Tokyo, 1982), Adv. Stud. Pure Math., vol. 4, North-Holland, 1984, pp. 331–390.
  • [OS17] T. Ohmoto and M. Shiota, C1C^{1}-triangulations of semi-algebraic sets, J. Topol. 10 (2017), no. 3, 765–775.
  • [Osh13] Y. Oshima, Localization of cohomological induction, Publ. Res. Inst. Math. Sci. 49 (2013), no. 2, 361–391.
  • [Sch91] W. Schmid, Construction and classification of irreducible Harish-Chandra modules, Harmonic analysis on reductive groups (Brunswick, ME, 1989), Progr. Math., vol. 101, Birkhäuser Boston, Boston, MA, 1991, pp. 235–275.
  • [Soe89] W. Soergel, Universelle versus relative Einhüllende: Eine geometrische Untersuchung von Quotienten von universellen Einhüllenden halbeinfacher Lie-Algebren, Math. Ann. 284 (1989), no. 2, 177–198.
  • [SRV98] S. Salamanca-Riba and D. Vogan, On the classification of unitary representations of reductive Lie groups, Ann. of Math. (2) 148 (1998), no. 3, 1067–1133.
  • [Sug59] M. Sugiura, Conjugate classes of cartan subalgebras in real semisimple Lie algberas, J. Math. Soc. Japan 11 (1959), no. 4, 374–434.
  • [SV98] W. Schmid and K. Vilonen, Two geometric character formulas for reductive Lie groups, J. Amer. Math. Soc. 11 (1998), no. 4, 799–867.
  • [Ver83] M. Vergne, Representations of Lie groups and the orbit method, Emmy Noether in Bryn Mawr (Bryn Mawr, Pa., 1982) (New York), Springer, 1983, pp. 59–101.
  • [Vog00] D. Vogan, The method of coadjoint orbits for real reductive groups, Representation Theory of Lie Groups (Providence), IAS/Park City Mathematical Series, vol. 8, American Mathematical Society, 2000, pp. 179–238.
  • [Wal92] N. Wallach, Real reductive groups II, Academic Press, 1992.