Hostname: page-component-745bb68f8f-g4j75 Total loading time: 0 Render date: 2025-01-21T04:18:46.128Z Has data issue: false hasContentIssue false

ON THE CONVERGENCE OF ZETA FUNCTIONS OF PREHOMOGENEOUS VECTOR SPACES

Published online by Cambridge University Press:  20 January 2025

TOBIAS FINIS
Affiliation:
Institute of Mathematics University of Leipzig Leipzig Germany Tobias.Finis@math.uni-leipzig.de
EREZ LAPID*
Affiliation:
Department of Mathematics Weizmann Institute of Science Rehovot Israel
Rights & Permissions [Opens in a new window]

Abstract

We prove a general convergence result for zeta functions of prehomogeneous vector spaces extending results of H. Saito, F. Sato and Yukie. Our analysis points to certain subspaces which yield boundary terms. We study it further in the setup arising from nilpotent orbits. In certain cases we determine the residue at the rightmost pole of the zeta function.

Type
Article
Creative Commons
Creative Common License - CCCreative Common License - BYCreative Common License - NCCreative Common License - SA
This is an Open Access article, distributed under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike licence (https://creativecommons.org/licenses/by-nc-sa/4.0/), which permits non-commercial re-use, distribution, and reproduction in any medium, provided the same Creative Commons licence is included and the original work is properly cited. The written permission of Cambridge University Press must be obtained for commercial re-use.
Copyright
© The Author(s), 2025. Published by Cambridge University Press on behalf of Foundation Nagoya Mathematical Journal

1 Introduction

The theory of prehomogeneous vector spaces (PVSs) and their zeta functions was initiated by M. Sato and Shintani [Reference Sato41], [Reference Sato and Shintani47]. (See [Reference Kimura23] for history and references.) One of the early motivations was to study asymptotics of arithmetic invariants [Reference Shintani43], an endeavor followed up by a number of authors (e.g., [Reference Datskovsky and Wright9], [Reference Kable and Yukie24], [Reference Wright and Yukie51], [Reference Yukie53]). For these applications, the alternative approach of Bhargava using innovative geometry of numbers techniques was very successful (see [Reference Bhargava1] and the references therein). Nonetheless, zeta functions of PVSs are still useful for these applications [Reference Bhargava, Taniguchi and Thorne5]. Our motivation for the study of the zeta functions comes from their appearance in Arthur’s trace formula [Reference Chaudouard7], [Reference Finis, Hoffmann and Wakatsuki10], [Reference Hoffmann16], [Reference Hoffmann17], [Reference Hoffmann and Wakatsuki19], although this will not play a direct role in the current paper.

The purpose of this paper is to provide a general basic convergence result for zeta functions of regular PVSs over numbers fields. It generalizes earlier results in this direction by F. Satō, H. Saito, Yukie, and others. More importantly, we hope that the method will enhance our understanding of the singularities of the zeta function and ultimately will be incorporated to Arthur’s trace formula. As a first result in this direction, we determine the structure of the first pole (at the edge of the domain of absolute convergence) for a restricted, but natural class of PVSs.

In order to state our main result, let V be a reductive PVS over a number field F, i.e, an F-rational representation $\rho $ of a reductive group G over F on V with an open orbit X. Let $H=G_v$ be the generic stabilizer and let $\chi _1,\dots ,\chi _r\in X^*(G)$ be the fundamental F-rational characters of G. Let $\mathbb {A}$ be the ring of adeles of F. For any rapidly decreasing continuous function $\phi $ on $V(\mathbb {A})$ let $\theta _\phi $ be the automorphic function

$$\begin{align*}\theta_\phi(g)=\sum_{v\in X(F)}\phi(\rho(g)^{-1}v),\ \ g\in G(F)\backslash G(\mathbb{A}). \end{align*}$$

Roughly speaking, the zeta function is the Mellin transform of $\theta _\phi $ . In the simplest case, our result is the following.

Theorem 1.1. Suppose that the connected component $H^\circ $ of H is semisimple. Then, the zeta function

$$\begin{align*}Z(\phi,(s_1,\dots,s_r))=\int_{G(F)\backslash G(\mathbb{A})}\theta_\phi(g)\left|{\det\rho(g)}\right|{}^{-1}\prod_{i=1}^r\left|{\chi_i(g)}\right|{}^{-s_i}\ dg, \end{align*}$$

converges absolutely for $\Re s_i>0$ , $i=1,\dots ,r$ .

Theorem 1.1 cannot hold as stated for a general PVS. The first issue is that assuming convergence, as a function of $x\in G(\mathbb {A})$ the integral

(1.1) $$ \begin{align} \int_{G(F)\backslash G(\mathbb{A})^1}\theta_\phi(gx)\ dg, \end{align} $$

is invariant not just under $G(\mathbb {A})^1$ but also under the adelic points of the group ${G}^\star =HG^{\operatorname {der}}$ . Therefore, in the case where $X^*(H)$ is infinite (or, equivalently $r<\dim X^*(G)$ ), we need to modify the definition of the zeta function as follows:

(1.2) $$ \begin{align} Z(\phi,(s_1,\dots,s_r))=\int_{G(\mathbb{A})^1{G}^\star(\mathbb{A})\backslash G(\mathbb{A})}\int_{G(F)\backslash G(\mathbb{A})^1}\theta_\phi(ga)\ dg\ \left|{\det\rho(a)}\right|{}^{-1}\prod_{i=1}^r\left|{\chi_i(a)}\right|{}^{-s_i}\ da. \end{align} $$

We will show that this integral converges absolutely for $\Re s_i>0$ , $i=1,\dots ,r$ provided that H is connected reductive.

In the non-connected case, there is yet a more fundamental difficulty, namely that already the integral (1.1) may diverge. This happens if $\operatorname {vol}(G_v(F)\backslash (G_v(\mathbb {A})\cap G(\mathbb {A})^1))=\infty $ , or equivalently, if the group of connected components $G_v/G_v^\circ $ acts nontrivially on $X^*(G_v^\circ )$ for some $v\in X(F)$ . We call such elements v isotropic. For instance, this is the case for the PVS of binary quadratic forms (with respect to the action of $\operatorname {GL}_2$ ) in which the centralizer (an orthogonal group in dimension $2$ ) may split over F.

To remedy the situation, we modify $\theta _\phi $ , and consequently the zeta function, by simply removing the sum over isotropic $v\in X(F)$ . Our main result is that the modified zeta function converges for $\Re s_i>0$ , $i=1,\dots ,r$ provided that $H^\circ $ is reductive (see Theorem 4.3 below).

In fact, using different methods, the convergence of this modified zeta function in some orthant $\Re s_i\gg 0$ was proved by H. Saito under the condition that $X^*(H)$ is finite [Reference Saito38].

We also remark that meromorphic continuation to $\mathbb {C}^r$ and functional equations for the zeta function (1.2) were established by F. Satō [Reference Satō39] under the assumptions that $H^\circ $ is reductive, $X^*(H)$ is finite and that there are no isotropic elements in $X(F)$ , extending the classical work by Sato and Shintani [Reference Sato and Shintani48]. For the PVS of binary quadratic forms, meromorphic continuation of (1.2) to $\mathbb {C}$ and the functional equation of a modified zeta function, which includes a contribution from the isotropic elements, were obtained by Yukie [Reference Yukie52] (following the work of Shintani [Reference Shintani45]). We will not deal with these questions here.

The contents of the paper are as follows. We start with recalling general facts, mostly well known, about PVSs (§2). We review examples of PVSs, especially those arising from nilpotent orbits, in §3. The main result is stated in §4. In §5, we give a general bound for the function $\theta _\phi $ (Proposition 5.1). The analysis is based on the examination of the support sets of elements in $X(F)$ with respect to the weights of a maximal F-split torus of G, an idea which goes back to Yukie [Reference Yukie53], [Reference Yukie54]. The linchpin of the argument is a certain nonnegativity statement pertaining to these support sets (Corollary 5.7). Further analysis is carried out in §6 where the key notion of special subspaces of a PVS is introduced. In §7, we prove the main convergence result. An important aspect in the proof is that the existence of isotropic elements is closely related to the existence of special subspaces with additional properties, which we call exceptional (Proposition 7.4). The nonexceptional subspaces exhibit a crucial positivity property (Proposition 7.9), which strengthens Corollary 5.7 supra and guarantees convergence. In §8, we improve the convergence result for PVSs without nontrivial special subspaces (under an additional technical condition which is satisfied in the F-irreducible case) and establish meromorphic continuation to the left of the polar hyperplanes $s_i = 0$ , $i = 1, \ldots , r$ together with an explicit description of the residues. Finally, in §9, we give a general construction of special subspaces for regular PVSs coming from nilpotent orbits.

2 Prehomogeneous vector spaces

We start with basic facts about PVSs. (See [Reference Kimura23, Chapter 2] and [Reference Satō39, §1] for more details.)

2.1. General notation

Throughout the paper F is a field of characteristic $0$ with algebraic closure $\bar F$ . (From §4 onward, F will be a number field.)

We denote by $\mathbf {G}_m$ the multiplicative group.

Whenever a group G acts on a set A, we denote by $A^G$ the set of fixed points of G in A.

We denote by $X_*(T)$ the lattice of F-rational cocharacters of an F-torus T.

For a linear algebraic group G over F, we use the following notation.

  • $G^\circ $ —the identity component of G.

  • $Z(G)$ —the center of G.

  • $G^{\operatorname {der}}$ —the derived group of G.

  • $G^{\operatorname {ab}}=G/G^{\operatorname {der}}$ —the abelianization of G.

  • $N_{G}$ —the unipotent radical of G.

  • $G^{\operatorname {lev}}=G/N_{G}$ —the Levi part of G, so that $G^{\operatorname {red}}:=(G^{\operatorname {lev}})^{\circ }=(G^\circ )^{\operatorname {lev}}$ is reductive.

  • $G^{\operatorname {ss}}=(G^{\operatorname {red}})^{\operatorname {der}}$ —the semisimple part of $G^\circ $ .

  • $G^{\operatorname {tor}}=(G^{\operatorname {red}})^{\operatorname {ab}}$ —the largest toric quotient of $G^\circ $ .

  • $G^{\operatorname {ssu}}=\operatorname {Ker} (G^\circ \rightarrow G^{\operatorname {tor}})$ —an extension of $G^{\operatorname {ss}}$ by $N_{G}$ .

  • —the Lie algebra of G (and similarly for other groups).

  • $\operatorname {Ad}=\operatorname {Ad}_G$ —the adjoint representation of G on .

  • $X^*(G)=X^*(G^{\operatorname {ab}})=X^*(G^{\operatorname {lev}})$ —the finitely generated abelian group of F-rational characters of G.

  • $\Delta _G=\det \operatorname {Ad}\in X^*(G)$ —the modular character of G.

  • $T_G$ —the maximal F-split torus of $Z(G^{\operatorname {red}})$ .

  • $\mathfrak {a}_G=X_*(T_G)\otimes _{\mathbb {Z}}\mathbb {R}$ —the $\mathbb {R}$ -vector space generated by $X_*(T_G)$ .

We often view $G^{\operatorname {lev}}$ as an F-subgroup (the Levi subgroup) of G, uniquely defined up to conjugation by $N_{G}(F)$ ([Reference Mostow27] or [Reference Hochschild15, Chapter VIII, Theorem 4.3]).

If G is connected, then $X^*(G)=X^*(G^{\operatorname {tor}})$ is torsion-free, i.e., it is a lattice.

If G is reductive, then $G^{\operatorname {ssu}}=G^{\operatorname {ss}}=G^{\operatorname {der}}$ .

If T is an F-split torus, then $X^*(T)$ can be identified with the dual lattice of $X_*(T)$ . In general, if $T'$ is the maximal F-split subtorus of T, then we can identify $X_*(T)$ with $X_*(T')$ , while the restriction map $X^*(T)\rightarrow X^*(T')$ is injective with finite cokernel. Thus, we have isomorphisms

$$\begin{align*}X^*(T)\otimes_{\mathbb{Z}}\mathbb{R}\simeq X^*(T')\otimes_{\mathbb{Z}}\mathbb{R}\simeq \mathfrak{a}_{T'}^*= \mathfrak{a}_T^*. \end{align*}$$

The F-split rank of T, i.e. the dimension of $T'$ , is the rank of $X^*(T)$ (and $X_*(T)$ ).

More generally, if G is connected, then the restriction map $X^*(G)=X^*(G^{\operatorname {red}})\rightarrow X^*(T_G)$ is injective with finite cokernel and hence yields an isomorphism

$$\begin{align*}X^*(G)\otimes_{\mathbb{Z}}\mathbb{R}\simeq X^*(T_G)\otimes\mathbb{R}=\mathfrak{a}_G^*. \end{align*}$$

For any finite subset A of a real vector space we denote by $\mathbb {R}_{\ge 0}A$ the polyhedral cone generated by A, i.e., the set of linear combinations $\sum _{a\in A}x_aa$ , where $x_a\ge 0$ for all $a\in A$ . Similarly, we use the notation $\mathbb {R}_{>0}A$ , $\mathbb {Z}_{\ge 0}A$ , $\mathbb {Z}_{>0}A$ where we require the $x_a$ to be strictly positive (resp., nonnegative integers, strictly positive integers) for all a.

Let G be an algebraic F-group and let H be an F-subgroup of G. We record a few basic results.

Lemma 2.1. Suppose that G is connected. Then, the following two conditions are equivalent.

  1. 1. The restriction map $X^*(G)\rightarrow X^*(H^\circ )$ has finite cokernel.

  2. 2. The torus $(Z(H^{\operatorname {red}})\cap G^{\operatorname {ssu}})^\circ $ is F-anisotropic.

Proof. We may assume that H is connected. Fix a Levi subgroup $L\simeq H^{\operatorname {red}}$ of H. Then, the restriction map $X^*(H)\simeq X^*(L)\rightarrow X^*(T_L)$ is injective with finite cokernel. Therefore, we may rephrase condition 1 by saying that the restriction map $X^*(G)=X^*(G^{\operatorname {tor}})\rightarrow X^*(T_L)$ has finite cokernel, or equivalently, that $X_*(T_L)\rightarrow X_*(G/G^{\operatorname {ssu}})$ is injective. In turn, this is equivalent to the triviality of $(T_L\cap G^{\operatorname {ssu}})^\circ $ , which amounts to condition 2.

Lemma 2.2. Suppose that $H=G^\circ $ and let $\Gamma =G/H$ . Then, the kernel of the restriction map $X^*(G)\xrightarrow {r} X^*(H)$ is finite and its image is a finite index sublattice of $X^*(H)^G=X^*(H)^\Gamma $ . Thus, the cokernel of r is finite if and only if G acts trivially on $X^*(H)$ .

Proof. The first part follows from the fact that the homomorphism $H/[G,H]\rightarrow G^{\operatorname {ab}}$ has finite kernel and cokernel. In fact we have (a $\operatorname {Gal}(\bar F/F)$ -equivariant) exact sequence

$$\begin{align*}H_2(\Gamma,\mathbb{Z})\rightarrow H/[G,H]\rightarrow G^{\operatorname{ab}}\rightarrow\Gamma^{\operatorname{ab}}\rightarrow1, \end{align*}$$

[Reference Hilton and Stammbach18, Corollary VI.8.2]. The second part follows from the first part and the fact that if a finite group $\Gamma $ acts nontrivially on a lattice L, then $L^\Gamma $ is of infinite index in L.

We denote the Picard group of a variety X by $\operatorname {Pic} X=H^1(X,\mathcal {O}_X^*)$ . If X is smooth, then we can identify $\operatorname {Pic} X$ with the (Weil) divisor class group $\operatorname {Cl} X$ of X. (See [Reference Hartshorne14, §II.6] for basic facts about divisors and the Picard group.)

Lemma 2.3 ([Reference Raynaud33], Proposition VII.1.5; cf. also the proof of [Reference Popov31], Theorem 4)

Suppose that G is connected. Then, there is an exact sequence

$$\begin{align*}X^*(G)\rightarrow X^*(H)\rightarrow \operatorname{Pic} (G/H). \end{align*}$$

2.2

Recall that a prehomogeneous vector space (PVS) over F is a finite-dimensional F-representation

$$\begin{align*}\rho:G\rightarrow\operatorname{GL}(V), \end{align*}$$

of a connected linear algebraic F-group G on a finite-dimensional F-vector space V, for which there exists a (necessarily unique) open orbit X. Note that X is defined over F and $X(F)$ is $G(F)$ -invariant. If F is algebraically closed, then $X(F)$ is a $G(F)$ -orbit and the stabilizers $G_v$ , $v\in X(F)$ are conjugate by $G(F)$ . In general, given $u\in X(F)$ , the $G(F)$ -orbits in $X(F)$ are in bijection with the elements of the kernel of $H^1(F,G_u)\rightarrow H^1(F,G)$ (a map of pointed sets) [Reference Serre42, I §5.4, Corollary 1 of Proposition 36]. Correspondingly, the F-groups $G_v$ , $v\in X(F)$ are pure inner forms of $G_u$ .

Clearly, a quotient of a PVS by an invariant subspace is also a PVS.

We will write $H=G_v$ (the “generic stabilizer”) when $v\in X(F)$ is immaterial, e.g., when we refer to the group $X^*(H)$ or to properties pertaining to $H(\bar F)$ . Thus, we can identify X with $G/H$ . The F-subgroup ${G}^\star =G^{\operatorname {der}} H$ of G is well defined. Note that in general, H may be disconnected and the groups $G_v^\circ $ , $v\in X(F)$ are not necessarily in the same inner class. In fact, the lattice $X^*(G_v^\circ )$ may depend on the choice of $v\in X(F)$ . For this reason, we will refrain from using the notation $X^*(H^\circ )$ .

For $v\in X(F)$ , the linear map

(2.1)

is surjective, since it is the differential of the dominant map $G\rightarrow V$ , $g\mapsto \rho (g)v$ . The kernel of $D_v$ is the Lie algebra of $G_v$ [Reference Borel2, Lemma II.7.4]. We thus get a $G_v$ -equivariant isomorphism . In particular,

(2.2) $$ \begin{align} \Delta_G\det\rho^{-1}\big|_H=\Delta_H. \end{align} $$

A relative invariant with respect to a character $\chi \in X^*(G)$ is an F-rational function $f\not \equiv 0$ on V such that $f\circ \rho (g)=\chi (g)f$ for all $g\in G$ . We will write $f=\operatorname {RI}_\chi $ .

The following basic properties hold.

  1. 1. $\operatorname {RI}_\chi $ , if it exists, is unique up to a nonzero scalar and it is regular and nonvanishing on X.

  2. 2. The sublattice $X^*(G/{G}^\star )$ , which is the kernel of the restriction map $X^*(G)\rightarrow X^*(H)$ , consists of the characters with respect to which there exists a relative invariant.

  3. 3. Let $\chi _1,\dots ,\chi _r$ be a basis of $X^*(G/{G}^\star )$ . Then $\operatorname {RI}_{\chi _1},\dots ,\operatorname {RI}_{\chi _r}$ are algebraically independent.

  4. 4. The F-irreducible polynomials that are nowhere vanishing on X (if any) are relative invariants. They are called the fundamental relative invariants. The corresponding characters are called fundamental characters. They form a basis for $X^*(G/{G}^\star )$ . Denote the set of all fundamental characters by $\mathfrak {X}$ .

If G is reductive, then the following conditions are equivalent.

  1. 1. $H^\circ $ is reductive.

  2. 2. X is the complement of a hypersurface.

  3. 3. X is the complement of the hypersurface defined by $\operatorname {RI}_{\chi _1}\cdots \operatorname {RI}_{\chi _r}$ where $\mathfrak {X}=\{\chi _1,\dots ,\chi _r\}$ .

  4. 4. X is an affine variety.

  5. 5. V is regular (see [Reference Kimura23, Definition 2.14]).

In this case, $(\det \rho )^2\in X^*(G/{G}^\star )$ . (As far as we know, it is unknown whether $\operatorname {RI}_{\det \rho ^2}$ is a polynomial in general.)

We give some more basic properties of PVSs.

Lemma 2.4. There is a short exact sequence

$$\begin{align*}0 \rightarrow X^*(G/{G}^\star) \rightarrow X^* (G)\rightarrow X^* (H) \rightarrow 0. \end{align*}$$

Thus, $\operatorname {rk} X^*(G)-r=\operatorname {rk} X^*(H)$ .

Proof. Clearly, $X^*(G/{G}^\star )$ is the kernel of the restriction map $X^*(G)\rightarrow X^*(H)$ . The surjectivity of this map follows from Lemma 2.3 and the vanishing of $\operatorname {Pic}X$ . More generally, this is true for any open subvariety X of $Y=\operatorname {Spec} A$ where A is a Noetherian UFD since $\operatorname {Cl} Y=1$ and $\operatorname {Cl} Y\rightarrow \operatorname {Cl} X$ is onto.

Let $\overline {\mathfrak {a}}_G^*$ be the subspace $X^*(G/{G}^\star )\otimes \mathbb {R}$ of $\mathfrak {a}_G^*$ . Thus, $\mathfrak {X}$ forms a basis for $\overline {\mathfrak {a}}_G^*$ .

For completeness, we also note the following fact, which can also be proved by a topological argument for $F=\mathbb {C}$ .

Lemma 2.5. Suppose that V has no relative invariants. Then H is connected.

Proof. We may assume that F is algebraically closed. The condition on V means that $V\setminus X$ has codimension $>1$ in V. It follows from purity of the branch locus that X is simply connected [28, Exp. X, Corollaire 3.3], i.e., there is no nontrivial étale covering of X. On the other hand,

$$\begin{align*}G/H^\circ\rightarrow X, \end{align*}$$

is an étale covering. Thus, H is connected.

Lemma 2.6. The following conditions on $v\in X(F)$ are equivalent.

  1. 1. The cokernel of the restriction map $X^*(G)\rightarrow X^*(G_v^\circ )$ is infinite.

  2. 2. The cokernel of the restriction map $X^*(G_v)\rightarrow X^*(G_v^\circ )$ is infinite.

  3. 3. There is a non-trivial F-split torus in $Z(G_v^{\operatorname {red}})\cap G^{\operatorname {ssu}}$ .

  4. 4. The action of $G_v$ (via the component group $G_v/G_v^\circ $ ) on $X^*(G_v^\circ )$ is nontrivial.

Proof. 1 $\iff $ 2: Lemma 2.4.

1 $\iff $ 3: Lemma 2.1.

2 $\iff $ 4: Lemma 2.2.

Definition 2.7. An element $v\in X(F)$ is isotropic if it satisfies the conditions of Lemma 2.6. Otherwise, v is nonisotropic.

We denote the subset of nonisotropic elements in $X(F)$ by $X(F)^{\operatorname {niso}}$ . It is invariant under $G(F)$ .

A PVS V is isotropic if it contains isotropic elements. Otherwise, V is nonisotropic.

By Lemma 2.6, if H is connected, then V is nonisotropic.

The basic example of an isotropic PVS is the space V of binary quadratic forms with respect to the action of $\operatorname {GL}_2$ . Here, H is an orthogonal group in dimension $2$ ; $v\in X(F)$ is isotropic if and only it is isotropic in the usual sense of quadratic forms. We caution however, that the PVS of quadratic forms in $n>2$ variables (with respect to the $\operatorname {GL}_n$ -action) is nonisotropic.

In general, it is possible that all elements of $X(F)$ are isotropic (for instance, if V is isotropic and $F=\bar F$ ), although we expect that this is never the case if F is a number field.

Assume now that G is reductive, so that every representation of G is completely reducible. We single out two important classes of PVSs in the opposite extremes.

Definition 2.8. Let V be a reductive PVS. We say that

  1. 1. V is basic if for every $v\in X(F)$ , $G_v^\circ $ is not contained in any proper parabolic F-subgroup of G.Footnote 1

  2. 2. V is distinguished if $H/\operatorname {Ker}\rho $ is finite.Footnote 2

Note that every basic PVS is regular, since any connected nonreductive F-subgroup of G is contained in a proper parabolic F-subgroup of G [Reference Borel and Tits4]. Moreover, a basic PVS is necessarily nonisotropic. It is also clear that every subrepresentation of a basic PVS is also basic.

A distinguished PVS is regular and nonisotropic.

3 PVSs of DK-type, examples

A main source of PVSs are gradations on reductive Lie algebras and nilpotent orbits.

3.1

Let $G'$ be a reductive group over F and let $\lambda :\mathbf {G}_m\rightarrow G'$ be a one-parameter subgroup. Let . Decompose

This is a $\mathbb {Z}$ -gradation of . (If $G'$ is semisimple and adjoint, then every gradation of is of this form.) Let , $i\in \mathbb {Z}$ be the corresponding (decreasing) filtration of . The stabilizer P of $\mathcal {F}$ in G under $\operatorname {Ad}$ is the parabolic F-subgroup of G whose Lie algebra is $\mathcal {F}_0$ . The Lie algebra of the unipotent radical of P is $\mathcal {F}_1$ . The centralizer G of $\lambda $ is a Levi subgroup of P whose Lie algebra is . For every $i\ne 0$ , the vector space , which consists of nilpotent elements, comprises finitely many (geometric) orbits with respect to the adjoint action of G [Reference Vinberg49]. In particular, it is a PVS. In the terminology of [Reference Gyoja13], these are PVSs of V-type (for Vinberg). They are not necessarily regular. We also say that $\mathcal {F}$ is a filtration of of V-type. (Note that we may restrict here to the case $i=1$ , since the general case follows from it by considering the graded Lie subalgebra , which is reductive.)

As a special case, let P be a parabolic subgroup of $G'$ with Levi factor G and . Then, is a PVS with respect to the adjoint action of G. In fact, is a PVS with respect to the adjoint action of P by a well-known result of Richardson. We say that V is a PVS of R-type. We denote the nilpotent orbit containing the open orbit of P on (the Richardson orbit with respect to P) by $\operatorname {Rich}(P)$ . The number of irreducible components of V is the F-corank of P in G.

Other important special cases, this time of regular PVSs, arise from the Dynkin–Kostant classification of nilpotent orbits. In more detail, let $(e,h,f)$ be an -triple in . (By the Jacobson–Morozov theorem, every nilpotent is a part of an -triple.) It corresponds to an F-homomorphism $\operatorname {SL}_2\rightarrow G^{\prime \operatorname {der}}$ . Let be the grading induced by h. The corresponding (decreasing) filtration , $i\in \mathbb {Z}$ of depends only on e and in fact only on the P-orbit of e, where $P=P(e)$ is the stabilizer of $\mathcal {F}=(\mathcal {F}_i)_{i\in \mathbb {Z}}$ in $G'$ under the adjoint action. Recall that P is the parabolic subgroup of $G'$ with Lie algebra $\mathcal {F}_0$ . The Lie algebra of the unipotent radical of P is $\mathcal {F}_1$ . (That $\mathcal {F}$ depends only on e follows from the fact that two -triples $(e,h,f)$ and $(e,h',f')$ with the same first element e are conjugate by an element of $N_P(F) \cap G^{\prime }_e (F)$ [Reference Bourbaki3, Lemma VIII.11.1.4].)

Let G be the Levi quotient of P, so that . Let with the G-action $\rho $ induced by the adjoint representation. Let $\mathfrak {o}$ be the geometric $G'$ -orbit of e and let $\mathfrak {o}_{\mathcal {F}}=\mathfrak {o}\cap \mathcal {F}_2$ .

Theorem 3.1. With the notation above:

  1. 1. V is a regular PVS with respect to $\rho $ .

  2. 2. $\mathfrak {o}_{\mathcal {F}}=X+\mathcal {F}_3$ .

  3. 3. $\mathfrak {o}_{\mathcal {F}}$ is a P-orbit.

  4. 4. For every $g\in G'$ , $\operatorname {Ad}(g)\mathfrak {o}_{\mathcal {F}}\cap \mathfrak {o}_{\mathcal {F}}\ne \emptyset $ if and only if $g\in P$ . In particular, $G^{\prime }_v\subset P$ for any $v\in \mathfrak {o}_{\mathcal {F}}$ .

These are the PVSs of DK-type (Dynkin–Kostant) in the language of [Reference Gyoja13]. (We quickly sketch how to derive these assertions from the literature. Note first that is the open G-orbit X of V by [Reference Bourbaki3, Proposition VIII.11.3.6]. V is regular by [Reference Kac22], see below. In addition, the $N_P$ -orbit of e is $e + \mathcal {F}_3$ [Reference Ranga Rao34, Lemma 1]. This shows the second and third assertions. The fourth assertion easily reduces to the special case $G^{\prime }_e\subset P$ . Let $g \in G^{\prime }_e$ . Then $(e, \operatorname {Ad} (g) h, \operatorname {Ad} (g) f)$ is an -triple with the same first element as $(e,h,f)$ , and therefore conjugate to $(e,h,f)$ by an element $n \in N_P \cap G^{\prime }_e$ . Since $ng$ centralizes h, we have $ng \in G$ , and therefore $g \in P$ .)

If we fix in addition to e the -triple $(e,h,f)$ , then we can identify G with the centralizer of h in $G'$ and V with the subspace of . We have then $X=V\cap \mathfrak {o}$ . We may identify with the dual of via the Killing form. Fixing a vector space isomorphism between and , $\operatorname {RI}_{(\det \rho )^2}$ is given by the determinant of the linear map , a polynomial in $e\in V$ whose nonzero locus is X [Reference Kac22, Proposition 1.1, 1.2, 2.1].

Fix a minimal parabolic F-subgroup $P_0'$ of $G'$ and a maximal split F-torus $T_0'$ contained in $P_0'$ . Then, we may choose $(e,h,f)$ in its $G'(F)$ -orbit so that is dominant and $P=P(e)$ is standard. This is the canonical parabolic subgroup pertaining to $\mathfrak {o}$ . G is then identified with the standard Levi subgroup of P.

An orbit $\mathfrak {o}$ is called even if for all odd i (or equivalently, if ). For even orbits clearly $\mathfrak {o}=\operatorname {Rich}(P)$ . A particular case is given by distinguished orbits, where the connected stabilizer $(G^{\prime }_e)^{\circ }$ , $e\in \mathfrak {o}$ is modulo the center of $G'$ a unipotent group. In this case, the PVS V is distinguished (and the converse holds for even orbits $\mathfrak {o}$ ). In general, not every standard parabolic subgroup is the canonical parabolic subgroup of some nilpotent orbit, let alone an even one. (By [Reference Kac22, Proposition 2.1] and [Reference Gyoja13, Theorem 2.4], the Richardson orbit of a standard parabolic subgroup P has P as its canonical parabolic subgroup if and only if the associated PVS is regular and for all $i \ge 1$ , where denotes the descending central series of .)

Any PVS of DK-type arises from an even nilpotent orbit for the reductive Lie algebra . Hence, any PVS of DK-type is also of R-type, although not necessarily with the same $G'$ .

3.2

Let us give some concrete examples of DK-type arising from even nilpotent orbits. For each case we provide the following data:

  • The numbering in the Sato–Kimura classification (appendix of [Reference Kimura23]) in the irreducible case,

  • the group $G'$ ,

  • the (even) nilpotent orbit $\mathfrak {o}$ in (either the corresponding partition for classical groups or the notation in the Bala–Carter classification),

  • the canonical Levi G,

  • the regular PVS ,

  • the representation $\rho $ of G on V,

  • the fundamental relative invariant polynomials (FRIPs),

  • the generic stabilizer H.

  • If V is basic, distinguished, or isotropic, we indicate that.

For simplicity we only consider split $G'$ .

In the following $m,n,k$ are positive integers (possibly with some restrictions). We denote by $\operatorname {Mat}_{m,n}$ the space of $m\times n$ matrices and by $\operatorname {Mat}_n$ (resp., $\operatorname {Sym}_n$ , $\operatorname {Skew}_n$ ) the spaces of square (resp., symmetric, skew-symmetric) $n\times n$ -matrices. We denote by $\operatorname {Pf}$ the Pfaffian of a skew-symmetric matrix.

Example 3.2 (#1 for $k=2$ )

  • $G'=\operatorname {GL}_{n_1+\dots +n_k}$ , $k>1$ , $n_i=n_{k+1-i}$ for all i, $n_i\le n_{i+1}$ for $i\le l:=\lfloor \frac {k-1}{2}\rfloor $ .

  • $\mathfrak {o}$ corresponds to the partition

    $$\begin{align*}(\overbrace{k,\dots,k}^{n_1},\overbrace{k-2,\dots,k-2}^{n_2-n_1},\dots,\overbrace{k-2l}^{n_{l+1}-n_l}). \end{align*}$$
  • $G=\operatorname {GL}_{n_1}\times \dots \times \operatorname {GL}_{n_k}$ .

  • $V=\operatorname {Mat}_{n_1,n_2}\oplus \dots \oplus \operatorname {Mat}_{n_{k-1},n_k}$ .

  • $\rho (g_1,\dots ,g_k)(x_1,\dots ,x_{k-1})=(g_1x_1g_2^{-1},\dots ,g_{k-1}x_{k-1}g_k^{-1})$ .

  • FRIPs: $\begin {cases}\det x_i&\text {if }n_i=n_{i+1}\\\det (x_i\dots x_{k-i})&\text {if }n_i<n_{i+1}\end {cases}$

  • $H=\operatorname {GL}_{n_1}\times \operatorname {GL}_{n_2-n_1}\times \dots \times \operatorname {GL}_{n_{l+1}-n_l}$ embedded as

    $$\begin{align*}(g_1,\dots,g_{l+1})\mapsto (g_1,\operatorname{diag}(g_1,g_2),\dots,\operatorname{diag}(g_1,\dots,g_{l+1}),\dots,\operatorname{diag}(g_1,g_2),g_1). \end{align*}$$
  • Identifying $X^*(G)$ with $\mathbb {Z}^k$ (and consequently $\mathfrak {a}_G^*$ with $\mathbb {R}^k$ ) by $(\lambda _1,\dots ,\lambda _k)\mapsto \prod _{i=1}^k\det g_i^{\lambda _i}$ , the subspace $\overline {\mathfrak {a}}_G^*$ is given by the equations $\lambda _{i+1}+\dots +\lambda _{k-i}=0$ whenever $n_i<n_{i+1}$ or $i=0$ .

  • V is basic if and only if $n_1=\dots =n_k$ . V is distinguished if and only if $n_1=\dots =n_k = 1$ .

Example 3.3 (#13 for $k=1$ , up to castling)

  • $G'=\operatorname {Sp}_{2(n_1+\dots +n_k)+n_{k+1}}$ , $n_1\le \dots \le n_{k+1}$ even. Set $m_i=n_i-n_{i-1}$ (with $n_0=0$ ).

  • $\mathfrak {o}$ corresponds to the partition

    $$\begin{align*}(\overbrace{2k+1,\dots,2k+1}^{m_1},\overbrace{2k-1,\dots,2k-1}^{m_2},\dots,\dots,\overbrace{1,\dots,1}^{m_{k+1}}). \end{align*}$$
  • $G=\operatorname {GL}_{n_1}\times \dots \times \operatorname {GL}_{n_k}\times \operatorname {Sp}_{n_{k+1}}$ .

  • $V=\operatorname {Mat}_{n_1,n_2}\oplus \dots \oplus \operatorname {Mat}_{n_k,n_{k+1}}$ .

  • $\rho (g_1,\dots ,g_{k+1})(x_1,\dots ,x_k)=(g_1x_1g_2^{-1},\dots ,g_kx_kg_{k+1}^{-1})$ .

  • FRIPs: $\begin {cases}\det x_i&\text {if }n_i=n_{i+1},\\ \operatorname {Pf}(x_i\dots x_{k+1}J_{n_{k+1}}x_{k+1}^t\dots x_i^t)&\text {otherwise},\end {cases}$ , $i=1,\dots ,k+1$ where $J_{n_{k+1}}\in \operatorname {Skew}_{n_{k+1}}$ defines $\operatorname {Sp}_{n_{k+1}}$ .

  • $H=\operatorname {Sp}_{m_1}\times \dots \times \operatorname {Sp}_{m_{k+1}}$ embedded as

    $$\begin{align*}(h_1,\dots,h_{k+1})\mapsto (h_1,\iota_2(h_1,h_2),\dots,\iota_{k+1}(h_1,\dots,h_{k+1})) \end{align*}$$
    where $\iota _j$ is an embedding $\operatorname {Sp}_{m_1}\times \dots \times \operatorname {Sp}_{m_j}\hookrightarrow \operatorname {Sp}_{n_j}$ .
  • V is basic if and only if $n_1=\dots =n_k$ .

Example 3.4 (#2 for $k=1$ )

  • $G'=\operatorname {Sp}_{2(n_1+\dots +n_k)}$ , $n_1\le \dots \le n_k$ . Set $m_i=n_i-n_{i-1}$ (with $n_0=0$ ).

  • $\mathfrak {o}$ corresponds to the partition

    $$\begin{align*}(\overbrace{2k,\dots,2k}^{m_1},\overbrace{2k-2,\dots,2k-2}^{m_2},\dots,\dots,\overbrace{2,\dots,2}^{m_k}). \end{align*}$$
  • $G=\operatorname {GL}_{n_1}\times \dots \times \operatorname {GL}_{n_k}$ .

  • $V=\operatorname {Mat}_{n_1,n_2}\oplus \dots \oplus \operatorname {Mat}_{n_{k-1},n_k}\oplus \operatorname {Sym}_{n_k}$ .

  • $\rho (g_1,\dots ,g_{k+1})(x_1,\dots ,x_k)=(g_1x_1g_2^{-1},\dots ,g_{k-1}x_{k-1}g_k^{-1},g_kx_kg_k^t)$ .

  • FRIPs: $\begin {cases}\det x_i&\text {if } n_i=n_{i+1}\text { or }i=k,\\ \det x_i\cdots x_kx_{k-1}^t\cdots x_i^t&\text {otherwise},\end {cases}$ , $i=1,\dots ,k$ .

  • $H=\operatorname {O}_{m_1}\times \dots \times \operatorname {O}_{m_k}$ embedded as

    $$\begin{align*}(h_1,\dots,h_k)\mapsto (h_1,\iota_2(h_1,h_2),\dots,\iota_k(h_1,\dots,h_k)), \end{align*}$$
    where $\iota _j$ is an embedding $\operatorname {O}_{m_1}\times \dots \times \operatorname {O}_{m_j}\hookrightarrow \operatorname {O}_{n_j}$ .
  • V is basic if and only if $n_1=\dots =n_k\ne 2$ . V is isotropic if and only if $m_i=2$ for some i. V is distinguished if and only if $m_i=0$ or $1$ for all i.

Example 3.5 (#15 for $k=1$ , up to castling)

  • $G'=\operatorname {SO}_{2(n_1+\dots +n_k)+n_{k+1}}$ , $n_1\le \dots \le n_{k+1}$ . Set $m_i=n_i-n_{i-1}$ (with $n_0=0$ ).

  • $\mathfrak {o}$ corresponds to the partition

    $$\begin{align*}(\overbrace{2k+1,\dots,2k+1}^{m_1},\overbrace{2k-1,\dots,2k-1}^{m_2},\dots,\dots,\overbrace{1,\dots,1}^{m_{k+1}}). \end{align*}$$
  • $G=\operatorname {GL}_{n_1}\times \dots \times \operatorname {GL}_{n_k}\times \operatorname {SO}_{n_{k+1}}$ .

  • $V=\operatorname {Mat}_{n_1,n_2}\oplus \dots \oplus \operatorname {Mat}_{n_k,n_{k+1}}$ .

  • $\rho (g_1,\dots ,g_{k+1})(x_1,\dots ,x_k)=(g_1x_1g_2^{-1},\dots ,g_kx_kg_{k+1}^{-1})$ .

  • FRIPs: $\begin {cases}\det x_i&\text {if }n_i=n_{i+1},\\ \det (x_i\dots x_{k+1}J_{n_{k+1}}x_{k+1}^t\dots x_i^t)&\text {otherwise},\end {cases}$ , $i=1,\dots ,k+1$ where $J_{n_{k+1}}\in \operatorname {Sym}_{n_{k+1}}$ defines $\operatorname {SO}_{n_{k+1}}$ .

  • $H=S(\operatorname {O}_{m_1}\times \dots \times \operatorname {O}_{m_{k+1}})$ embedded as

    $$\begin{align*}(h_1,\dots,h_{k+1})\mapsto (h_1,\iota_2(h_1,h_2),\dots,\iota_{k+1}(h_1,\dots,h_{k+1})), \end{align*}$$
    where $\iota _j$ is an embedding $\operatorname {O}_{m_1}\times \dots \times \operatorname {O}_{m_j}\hookrightarrow \operatorname {O}_{n_j}$ .
  • V is basic if and only if $n_1=\dots =n_k\ne 2$ and $m_{k+1}\ne 2$ . V is isotropic if and only if $m_i=2$ for some i. V is distinguished if and only if $m_i=0$ or $1$ for all i.

Example 3.6 (#3 for $k=1$ )

  • $G'=\operatorname {SO}_{2(n_1+\dots +n_k)}$ , $n_1\le \dots \le n_k$ even. Set $m_i=n_i-n_{i-1}$ (with $n_0=0$ ).

  • $\mathfrak {o}$ corresponds to the partition

    $$\begin{align*}(\overbrace{2k,\dots,2k}^{m_1},\overbrace{2k-2,\dots,2k-2}^{m_2},\dots,\dots,\overbrace{2,\dots,2}^{m_k}). \end{align*}$$
  • $G=\operatorname {GL}_{n_1}\times \dots \times \operatorname {GL}_{n_k}$ .

  • $V=\operatorname {Mat}_{n_1,n_2}\oplus \dots \oplus \operatorname {Mat}_{n_{k-1},n_k}\oplus \operatorname {Skew}_{n_k}$ .

  • $\rho (g_1,\dots ,g_{k+1})(x_1,\dots ,x_k)=(g_1x_1g_2^{-1},\dots ,g_{k-1}x_{k-1}g_k^{-1},g_kx_kg_k^t)$ .

  • FRIPs: $\begin {cases}\det x_i&\text {if } n_i=n_{i+1},\\ \operatorname {Pf}(x_i\cdots x_kx_{k-1}^t\cdots x_i^t)&\text {otherwise},\end {cases}$ , $i=1,\dots ,k$ .

  • $H=\operatorname {Sp}_{m_1}\times \dots \times \operatorname {Sp}_{m_k}$ embedded as

    $$\begin{align*}(h_1,\dots,h_k)\mapsto (h_1,\iota_2(h_1,h_2),\dots,\iota_k(h_1,\dots,h_k)), \end{align*}$$
    where $\iota _j$ is an embedding $\operatorname {Sp}_{m_1}\times \dots \times \operatorname {Sp}_{m_j}\hookrightarrow \operatorname {Sp}_{n_j}$ .
  • V is basic if and only if $n_1=\dots =n_k$ .

Example 3.7 (#4)

  • $G'=G_2$ .

  • $\mathfrak {o}$ is the subregular (10-dimensional) orbit.

  • $G=\operatorname {GL}_2$ , with $\Delta _0^G$ consisting of the short simple root.

  • V is the space of binary cubic forms

  • $\rho $ is the symmetric cube representation twisted by $\det ^{-1}$ .

  • FRIP: the discriminant.

  • H is (an F-form of) the symmetric group $S_3$ (with its faithful two-dimensional representation).

  • V is distinguished.

Example 3.8 (#8)

  • $G'=F_4$ , $\Delta _0=\{\alpha _1,\alpha _2,\alpha _3,\alpha _4\}$ , $\alpha _1,\alpha _2$ long, $\alpha _3,\alpha _4$ short.

  • $\mathfrak {o}=F_4(a_3)$ (40-dimensional)

  • $G=(\operatorname {GL}_2\times \operatorname {GL}_3)/\{(\lambda ^2 I_2,\lambda ^{-1}I_3)\}$ , $\Delta _0^G=\{\alpha _1,\alpha _3,\alpha _4\}$ .

  • $V=\operatorname {Hom}_F(F^2,\operatorname {Sym}_3)$ .

  • $\rho (g_1,g_2)(A)(\xi )=g_2A(\xi g_1)g_2^t$

  • FRIP: the discriminant of the binary cubic form $\xi \mapsto \det A(\xi )$ .

  • H is (an F-form of) of the symmetric group $S_4$ , embedded in G by its faithful three-dimensional representation on the $\operatorname {GL}_3$ -factor and by its irreducible two-dimensional representation (which factors through the surjection $S_4 \to S_3$ ) on the $\operatorname {GL}_2$ -factor.

  • V is distinguished.

Example 3.9 (#12)

  • $G'=E_6$ .

  • $\mathfrak {o}=D_4(a_1)$ (58-dimensional)

  • $G=\operatorname {SL}_3\times \operatorname {SL}_3\times \operatorname {GL}_2$ .

  • $V=\operatorname {Hom}_F(F^2,\operatorname {Mat}_3)$ .

  • $\rho (g_1,g_2,g_3)(A)(\xi )=g_1A(\xi g_3)g_2^{-1}$ , $\xi \in F^2$ .

  • FRIP: the discriminant of the binary cubic form $\xi \mapsto \det A(\xi )$ .

  • H is the normalizer of a maximal torus in $\operatorname {SL}_3$ , embedded diagonally into the $\operatorname {SL}_3$ -factors and mapped to the $\operatorname {GL}_2$ -factor via the faithful two-dimensional representation of its component group $S_3$ .

  • V is isotropic.

Other examples of basic PVSs of DK-type are summarized in Table 1 which refers to the numbering in the list in the appendix of [Reference Kimura23] whenever $\rho $ is reduced. (See also [Reference Igusa21].)

Table 1 Irreducible basic PVSs of DK-type for exceptional groups

Here, $\pi $ is the embedding $G_2\hookrightarrow \operatorname {SO}_7\hookrightarrow \operatorname {GL}_7$ and $\phi $ is its lifting to an embedding $G_2\hookrightarrow \operatorname {Spin}_7$ . Also, $\sigma $ is the embedding of $F_4$ as the fixed point subgroup of the outer involution of $E_6$ . Finally,

$$ \begin{align*} \iota_1(x,y)&=(\psi_{7,3}(\phi(x),y),y),\\ \iota_2(x,y)&=(\psi_{6,4}(\operatorname{Sym}^3(y),(x,y)),\operatorname{Sym}^2(y)),\\ \iota_3(x,y)&=\psi_{7,7}(\phi(x),\phi(y)), \end{align*} $$

where $\psi _{m,n}$ is the homomorphism $\operatorname {Spin}_m\times \operatorname {Spin}_n\rightarrow \operatorname {Spin}_{m+n}$ and we identify $\operatorname {Spin}_3\simeq \operatorname {SL}_2$ , $\operatorname {Spin}_6\simeq \operatorname {SL}_4$ , $\operatorname {Spin}_4\simeq \operatorname {SL}_2\times \operatorname {SL}_2$ , $\operatorname {SO}_4\simeq (\operatorname {SL}_2\times \operatorname {SL}_2)/\{\pm 1\}$ .

3.3

Another interesting set of examples of irreducible PVSs of DK-type arises from the Freudenthal-Tits magic square (see Table 2). In these cases, G is a maximal Levi subgroup of $G'$ corresponding to a simple root $\alpha $ which is a leaf in the Dynkin diagram (it is the unique simple root which is not orthogonal to the highest root $\tilde \alpha $ of $G'$ ). The highest weight $\mu $ of $\rho $ is $\tilde \alpha -\alpha $ and its restriction to $G^{\operatorname {der}}$ is the fundamental weight corresponding to the simple root $\beta $ adjacent to $\alpha $ (and of the same length) in the Dynkin diagram. Let $Q=LN'$ be the maximal parabolic subgroup of $G^{\operatorname {der}}$ corresponding to $\beta $ . Then, there exists a nontrivial element $w_0$ in the normalizer of L in $G^{\operatorname {der}}$ . We have $\alpha =w_0\mu $ , which is the lowest weight of V. Clearly, L stabilizes the root spaces $V_\alpha $ and $V_\mu $ . Remarkably, any $v\in V_\alpha +V_\mu $ outside $V_\alpha \cup V_\mu $ is regular. We have $H^\circ =L^{\operatorname {der}}$ and $[H:H^\circ ]=2$ with H containing a representative of $w_0$ . In particular, V is not basic. Since as a representation of H, we see that the restriction of $\rho $ to H is $\operatorname {Ind}_{H^\circ }^H(\rho '\oplus 1)$ where $\rho '$ is the representation of L on . It turns out that $N'$ is abelian. Therefore, is a PVS and in fact it is basic: the generic stabilizer (whose identity component is denoted by K in the table) is the centralizer of (a suitable choice of) $w_0$ in L, i.e., the stabilizer of an outer involution of L corresponding to $w_0$ . See drawings below.

Table 2 PVSs of DK-type pertaining to the Freudenthal–Tits magic square

The fundamental relative invariant polynomials in these cases (which are of degree 4) are called Freudenthal quartics (see [Reference Freudenthal12], [Reference Igusa20], [Reference Saito37]). They arise by considering $\operatorname {ad} (x)^4$ , as a map between the one-dimensional spaces and (cf. [Reference Gyoja13, Table II]). The orbit structure of V is described in [Reference Röhrle32].

What is more, the Richardson orbit with respect to the maximal parabolic subgroup corresponding to $\beta $ is also even. These orbits give rise to the PVSs #8 (Example 3.8), #12 (Example 3.9), #9 and #28, respectively, in the list of the appendix of [Reference Kimura23]. They all have a fundamental relative invariant polynomial of degree 12, obtained as the determinant of , $x \in V$ , where is two-dimensional. Alternatively, the PVS mentioned above has a cubic fundamental relative invariant polynomial, and the PVS V associated to $\beta $ is the direct sum of and of its isomorphic image under the simple reflection $w_\alpha $ , which enables one to construct the fundamental relative invariant of V as the discriminant of the cubic form associated to pairs of elements of .

Sato and Kimura classified the irreducible PVSs over $\bar F$ up to a suitable notion of equivalence weaker than isomorphism [Reference Sato and Kimura46].

There are 29 essential cases of regular PVSs, listed in the appendix of [Reference Kimura23].Footnote 3 All these examples are either of DK-type or restrictions thereof to suitable subgroups.Footnote 4 Most of these examples are non-isotropic. The exceptions are

  • The symmetric square representation of $\operatorname {GL}_2$ (Example 3.4 for $k=1$ , $n_1=2$ ).

  • Example 3.9.

  • The representation $\operatorname {GL}_2\times \operatorname {SO}_{k+2}$ on $\operatorname {Mat}_{2,k+2}$ (Example 3.5 for $k=1$ , $n_1=2$ ).

  • The restriction of the latter to $\operatorname {GL}_2\times \operatorname {Spin}_7$ (resp., $\operatorname {GL}_2\times G_2$ ) for $k=6$ (resp., $k=5$ ) via the Spin representation $\operatorname {Spin}_7\hookrightarrow \operatorname {SO}_8$ (resp., the $7$ -dimensional representation $G_2\hookrightarrow \operatorname {SO}_7$ ).

The non-regular irreducible PVSs over $\bar F$ are (up to castling equivalence) mostly of R-type or restrictions thereof. The notable exception is the representation $\rho $ of $\operatorname {SL}_2\times \operatorname {SL}_n$ on $\operatorname {Hom}_F(F^2,\operatorname {Skew}_n)$ given by $\rho (g_1,g_2)A(\xi )=g_2A(\xi g_1)g_2^t$ , which is a PVS for odd n but not of R-type for $n>7$ . Another curious example is the restriction of the representation of $\operatorname {GL}_3\times \operatorname {Sp}_{2n}$ (of R-type) on $\operatorname {Mat}_{3,2n}$ , $n>1$ given by $\rho (g_1,g_2)A=g_1Ag_2^{-1}$ to the image of the symmetric square representation of $\operatorname {GL}_2$ times $\operatorname {Sp}_{2n}$ . This irreducible, non-regular PVS has a relative invariant.

4 Statement of main result

From now on assume that F is a number field. Let $\mathbb {A} = F_\infty \times \mathbb {A}_{\operatorname {fin}}$ be the ring of adeles of F. For any connected F-group G we write

$$\begin{align*}G(\mathbb{A})^1=\cap_{\chi\in X^*(G)}\operatorname{Ker}\left|{\chi}\right|, \end{align*}$$

where $\left |{\chi }\right |:G(\mathbb {A})\rightarrow \mathbb {R}_{>0}$ is the composition $G(\mathbb {A})\xrightarrow {\chi }\mathbb {A}^*\xrightarrow {\left |{\cdot }\right |}\mathbb {R}_{>0}$ . The automorphic space $G(F)\backslash G(\mathbb {A})^1$ is of finite volume.

Let

$$\begin{align*}\mathfrak{a}_{G,\mathbb{C}}^*=\mathfrak{a}_G^*\otimes_{\mathbb{R}}\mathbb{C}=X^*(G)\otimes_{\mathbb{Z}}\mathbb{C}. \end{align*}$$

Every $\lambda \in \mathfrak {a}_{G,\mathbb {C}}^*$ defines a (topological) quasi-character $g\mapsto g^\lambda $ of $G(\mathbb {A})$ by

$$\begin{align*}g^{\sum c_i\chi_i}=\prod_i\left|{\chi_i(g)}\right|{}^{c_i},\ \ \chi_i\in X^*(G),\ c_i\in\mathbb{C}. \end{align*}$$

All quasi-characters of $G(\mathbb {A})/G(\mathbb {A})^1$ are obtained this way. In particular, we get an isomorphism of topological groups

$$\begin{align*}\mathrm{i}\mathfrak{a}_G^*\xrightarrow{\operatorname{Exp}_G}\operatorname{PD}(G(\mathbb{A})/G(\mathbb{A})^1),\ \ \lambda\mapsto (g\mapsto g^\lambda), \end{align*}$$

where $\operatorname {PD}$ denotes the Pontryagin dual.

The following is a supplement to Lemma 2.1.

Lemma 4.1. Suppose that G is connected and H is an F-subgroup of G. Write $H(\mathbb {A})^{(1)}=H(\mathbb {A})\cap G(\mathbb {A})^1$ . Then, the following two conditions are equivalent.Footnote 5

  1. 1. The restriction map $X^*(G)\xrightarrow {r} X^*(H^\circ )$ has finite cokernel.

  2. 2. The volume of the quotient $H(F)\backslash H(\mathbb {A})^{(1)}$ is finite.

Proof. We may assume without loss of generality that H is connected, since $[H(F):H^\circ (F)]$ is finite and $H(\mathbb {A})/H^\circ (\mathbb {A})$ is compact. Now, $\operatorname {vol}(H(F)\backslash H(\mathbb {A})^1)<\infty $ . Therefore, $\operatorname {vol}(H(F)\backslash H(\mathbb {A})^{(1)})<\infty $ if and only if $Y=H(\mathbb {A})^{(1)}/H(\mathbb {A})^1$ is compact. Note that Y is the kernel of the homomorphism

$$\begin{align*}H(\mathbb{A})/H(\mathbb{A})^1\rightarrow G(\mathbb{A})/G(\mathbb{A})^1. \end{align*}$$

Passing to the Pontryagin dual, the lemma follows from the commutative diagram

In fact, Y is compact if and only if $Y=1$ if and only if the restriction map $\mathfrak {a}_G^*\rightarrow \mathfrak {a}_H^*$ is surjective.

Henceforth, $(G,V,\rho )$ is a reductive PVS (i.e., G is reductive). We apply the criterion of Lemma 4.1 to the stabilizer $G_v$ of a vector $v \in X (F)$ . Recall Definition 2.7.

Lemma 4.2. Let $v \in X (F)$ . Then,

  1. 1. The group $G_v(\mathbb {A})^{(1)}=G_v(\mathbb {A})\cap G(\mathbb {A})^1$ is unimodular.

  2. 2. The volume $\kappa _v=\operatorname {vol}(G_v(F)\backslash G_v(\mathbb {A})^{(1)})$ is finite if and only if v is non-isotropic.

Proof. The first assertion follows from the fact that modular character of $G_v$ is the restriction of a character of G by (2.2). The second assertion follows from Lemma 4.1 and Lemma 2.6.

We say that a continuous function $\phi $ on $V (\mathbb {A})$ is rapidly decreasing, if it is supported in $V (F_\infty ) K$ for some compact subgroup K of $V(\mathbb {A}_{\operatorname {fin}})$ and ${\mathrm{sup}}_{v \in V (\mathbb {A})}\left |{f (v)}\right |(1+\lVert v_\infty \rVert )^N < \infty $ for every nonnegative integer N, where $\lVert \cdot \rVert $ denotes a norm on the real vector space $V(F_\infty )$ . Let $C_{\operatorname {rd}}(V(\mathbb {A}))$ be the space of continuous rapidly decreasing functions on $V(\mathbb {A})$ . It is a (strict) LF-space, i.e., the strict (locally convex) inductive limit of Fréchet spaces.

For any $\phi \in C_{\operatorname {rd}}(V(\mathbb {A}))$ we form the theta function

$$\begin{align*}\theta_\phi(g)=\theta_\phi^V(g)=\sum_{v\in X(F)}\phi(\rho(g)^{-1}v),\ \ g\in G(\mathbb{A}). \end{align*}$$

The sum is locally uniformly absolutely convergent and defines a continuous function on $G(F)\backslash G(\mathbb {A})$ . We would like to study this function and the corresponding zeta function which is roughly speaking the Mellin transform of $\theta _\phi $ .

Let $\overline {\mathfrak {a}}_{G,\mathbb {C}}^* = \overline {\mathfrak {a}}_G^* \otimes \mathbb {C}=X^*(G/{G}^\star )\otimes \mathbb {C}$ . Ideally, we would like to study the double integral

(4.1) $$ \begin{align} \int_{G(\mathbb{A})^1{G}^\star(\mathbb{A})\backslash G(\mathbb{A})}\int_{G(F)\backslash G(\mathbb{A})^1}\theta_\phi(ga)\ dg\ a^{-\lambda}\left|{\det\rho(a)}\right|{}^{-1}\ da, \end{align} $$

for $\lambda $ in a suitable region of convergence in $\overline {\mathfrak {a}}_{G,\mathbb {C}}^*$ . Note that the outer integral formally makes sense, since the inner integral can be written as the sum over the $G(F)$ -orbits in $X(F)$ , and the contribution of an orbit $\rho (G(F))v$ is

$$\begin{align*}\int_{G_v(F)\backslash G(\mathbb{A})^1}\phi(\rho(ga)^{-1}v)\ dg= \kappa_v\int_{(G_v(\mathbb{A})\cap G(\mathbb{A})^1)\backslash G(\mathbb{A})^1}\phi(\rho(ga)^{-1}v)\ dg, \end{align*}$$

where $\kappa _v=\operatorname {vol}(G_v(F)\backslash (G_v(\mathbb {A})\cap G(\mathbb {A})^1))$ as above. Note that the images of $G_v(\mathbb {A})$ and ${G}^\star (\mathbb {A})$ in $G(\mathbb {A})^1\backslash G(\mathbb {A})=\mathfrak {a}_G=X_*(G^{\operatorname {ab}})\otimes \mathbb {R}$ coincide—they are both equal to the image of $X_*({G}^\star /G^{\operatorname {der}})\otimes \mathbb {R}$ in $\mathfrak {a}_G$ . Therefore, we can write the above as

$$\begin{align*}\kappa_v\int_{G_v(\mathbb{A})\backslash G(\mathbb{A})^1{G}^\star(\mathbb{A})}\phi(\rho(ga)^{-1}v)\left|{\det\rho(g)}\right|{}^{-1}\ dg. \end{align*}$$

This also explains why the more straightforward integral $\int _{G(F)\backslash G(\mathbb {A})}\theta _\phi (g)g^{-\lambda }\ dg$ would not converge (for any $\lambda $ ) if $\overline {\mathfrak {a}}_G^* \neq \mathfrak {a}_G^*$ (i.e., if $\operatorname {rk} X^*(H)>0$ by Lemma 2.4).

A necessary condition for the convergence of this orbital integral for all $\phi $ is that $\kappa _v$ is finite, which by Lemma 4.2 is equivalent to v being non-isotropic. Define

$$\begin{align*}\theta_\phi^{\operatorname{niso}}(g)=\sum_{v\in X(F)^{\operatorname{niso}}}\phi(\rho(g)^{-1}v),\ \ g\in G(\mathbb{A}). \end{align*}$$

Thus, $\theta _\phi ^{\operatorname {niso}}$ omits the $G(F)$ -orbits in $X(F)$ whose contributions to (4.1) clearly diverge.

Our main result, which will be ultimately proved in §7, is the following.

Theorem 4.3. Assume that $(G,V,\rho )$ is a regular PVS. Then, the double integral

$$\begin{align*}Z(\phi,\lambda)=\int_{G(\mathbb{A})^1{G}^\star(\mathbb{A})\backslash G(\mathbb{A})}\int_{G(F) \backslash G (\mathbb{A})^1} \theta_\phi^{\operatorname{niso}} (ga) \, dg\ a^{-\lambda}\left|{\det\rho(a)}\right|{}^{-1} \, da, \end{align*}$$

is absolutely convergent and holomorphic for any $\lambda \in \overline {\mathfrak {a}}_{G,\mathbb {C}}^*$ such that $\Re \lambda \in \mathbb {R}_{>0}\mathfrak {X}$ . Moreover, the integral is continuous in $\phi \in C_{\operatorname {rd}}(V(\mathbb {A}))$ .

Roughly speaking, Theorem 4.3 means that the only problematic terms in (4.1) are those arising from the orbits $\rho (G(F))v$ where $\kappa _v=\infty $ .

The proof of Theorem 4.3 is based on a direct estimation of the theta function $\theta _\phi ^{\operatorname {niso}}$ using standard lattice-point counting methods. In fact, this idea has already been used previously by Yukie [Reference Yukie53], [Reference Yukie54] to obtain partial results in the direction of Theorem 4.3, mostly via a case-by-case analysis, for absolutely irreducible representations $\rho $ .Footnote 6 Our approach is completely general and avoids any classification result. Also, the possibility to define a zeta function in the case where $\overline {\mathfrak {a}}_G^* \neq \mathfrak {a}_G^*$ does not seem to have been noticed in the literature.

An alternative classification-free method was developed by Sato [Reference Satō40] and H. Saito [Reference Saito38], among others, using the decomposition of $X(F)$ into $G(F)$ -orbits (Galois cohomology) and an analysis of the resulting Euler products. Saito proves Theorem 4.3 under the additional assumption that $H^\circ $ is semisimple and the component group $H/H^\circ $ is abelian. Under the weaker assumption $\overline {\mathfrak {a}}_G^* = \mathfrak {a}_G^*$ , i.e., $X^*(H)$ finite, he proves convergence for all $\Re \lambda $ in some cone of the form $\lambda _0+\mathbb {R}_{>0}\mathfrak {X}$ .Footnote 7

We expect Theorem 4.3 to hold without the regularity assumption on V. In the case where there are no relative invariants (in which V is called special), the set $G(\mathbb {A})X(F)$ is open in $V(\mathbb {A})$ , and the integral $\int _{G(F)\backslash G(\mathbb {A})^1}\theta _\phi (g)\ dg$ converges to a constant multiple of $\int _{G(\mathbb {A})X(F)}\phi (x)\ dx$ . (See [Reference Morishita and Watanabe29], [Reference Ono30] and the discussion in [Reference Hoffmann17, §2.4]. The set $G(\mathbb {A})X(F)$ is described in [Reference Saito36, Proposition 1.7].)

5 Bounding the theta function

We start with a general bound on $\theta _\phi (g)$ (Proposition 5.1 below) which is valid without the regularity assumption, but is by itself insufficient to prove convergence of the zeta integral of $\theta _\phi ^{\operatorname {niso}}(g)$ . To state it, we need to set some more notation which will be used throughout the paper.

5.1

Recall that G is a reductive group over a number field F. Fix a maximal F-split torus $T_0$ of G and a minimal parabolic F-subgroup $P_0$ of G containing $T_0$ . We have a Levi decomposition $P_0=M_0\ltimes N_0$ where $M_0$ is the centralizer of $T_0$ in G and $N_0=N_{P_0}$ , which is a maximal unipotent F-subgroup of G.

If P is a standard parabolic subgroup of G, then we denote by $M_P$ its standard Levi subgroup.

For simplicity we write $\mathfrak {a}_0=\mathfrak {a}_{P_0}=\mathfrak {a}_{M_0}\simeq \mathfrak {a}_{T_0}$ . Let

$$\begin{align*}H_0:M_0(\mathbb{A})\rightarrow\mathfrak{a}_0, \end{align*}$$

be the group homomorphism defined by

$$\begin{align*}e^{\left\langle{\chi},{H_0(m)}\right\rangle}=\left|{\chi(m)}\right|,\ \ m\in M_0(\mathbb{A}),\ \chi\in X^*(M_0). \end{align*}$$

Let $\Delta _0\subset X^*(T_0)$ be the set of simple roots of $T_0$ on $\operatorname {Lie} N_0$ .

Fix a maximal compact subgroup K of $G(\mathbb {A})$ which is in good position with respect to $P_0$ .

For any $g\in G(\mathbb {A})$ , we write $m_0(g)\in M_0(\mathbb {A})/(M_0(\mathbb {A})\cap K)$ for the $M_0(\mathbb {A})$ -part in the Iwasawa decomposition $G(\mathbb {A})=N_0(\mathbb {A})M_0(\mathbb {A})K$ of g.

We extend $H_0$ to a left- $N_0(\mathbb {A})$ and right-K-invariant function on $G(\mathbb {A})$ . Thus,

$$\begin{align*}H_0(g)=H_0(m_0(g)). \end{align*}$$

Fix a Siegel set $\mathfrak {S}$ in $G(\mathbb {A})$ of the form

$$\begin{align*}\mathfrak{S}=\{g\in G(\mathbb{A})\mid\left\langle{\alpha},{H_0(g)}\right\rangle>c_0\ \forall\alpha\in\Delta_0\}, \end{align*}$$

for some fixed constant $c_0$ . Thus,

  1. 1. $\mathfrak {S}$ is left $P_0(F)N_0(\mathbb {A})$ -invariant.

  2. 2. For any $g\in G(\mathbb {A})$ , the set $\{\gamma \in P_0(F)\backslash G(F)\mid \gamma g\in \mathfrak {S}\}$ is finite.

  3. 3. There exists a compact subset $\Omega \subset G(\mathbb {A})$ such that

    (5.1) $$ \begin{align} \mathfrak{S}\subset P_0(F)(T_0(\mathbb{A})\cap\mathfrak{S})\Omega. \end{align} $$
  4. 4. $G(F)\mathfrak {S}=G(\mathbb {A})$ provided that $c_0$ is sufficiently negative.

Let $\delta _{P_0}$ be the modulus function of $P_0(\mathbb {A})$ .

5.2

Recall that V is a PVS. (For now, we do not assume that V is regular.) The first order of business is the following result, which will be proved in the rest of this section.

Proposition 5.1. There exists a continuous seminorm $\nu $ on $C_{\operatorname {rd}}(V(\mathbb {A}))$ such that

$$\begin{align*}\left|{\theta_\phi(g)}\right|\le\delta_{P_0}(m_0(g))\left|{\det\rho(g)}\right|\nu(\phi), \end{align*}$$

for all $\phi \in C_{\operatorname {rd}}(V(\mathbb {A}))$ and $g\in \mathfrak {S}$ .

Note that $\left |{\det \rho }\right |$ is the modulus function for the action of $G(\mathbb {A})$ on $V(\mathbb {A})$ , that is,

$$\begin{align*}\int_{V(\mathbb{A})}f(\rho(g)^{-1}v)\ dv=\left|{\det\rho(g)}\right|\int_{V(\mathbb{A})}f(v)\ dv, \end{align*}$$

for any $f\in L^1(V(\mathbb {A}))$ and $g\in G(\mathbb {A})$ .

5.3

Let $\Phi _G$ be the set of roots of $T_0$ on G and let $\Psi _V$ be the set of weights of $T_0$ on V. The set $\Phi _G$ is a (not necessarily reduced) root system and the co-roots are in $X_*(T_0)$ . We decompose and V according to the roots and weights:

(5.2a)
(5.2b) $$ \begin{align} V&=\oplus_{\beta\in\Psi_V}V_\beta. \end{align} $$

The vector spaces , $\alpha \in \Phi _G$ , are not necessarily one-dimensional unless G is split. Moreover, is not necessarily a commutative Lie subalgebra (namely, if $2\alpha \in \Phi _G$ ). The spaces and $V_\beta $ are invariant under $M_0$ . We write

$$\begin{align*}n_\beta=\dim V_\beta,\ \ \beta\in\Psi_V, \end{align*}$$

and

$$\begin{align*}\delta_V=\det\rho\big|_{T_0}=\sum_{\beta\in\Psi_V}n_\beta\beta\in X^*(T_0). \end{align*}$$

By convention, $V_\gamma =0$ if $\gamma \notin \Psi _V$ . Note that

(5.3)

where is the Lie algebra representation obtained by differentiating $\rho $ .

Let

where $\Phi _G^+\subset \Phi _G$ is the subset of positive roots with respect to $P_0$ , so that

$$\begin{align*}\delta_{P_0}(m)=e^{\left\langle{\delta_0},{H_0(m)}\right\rangle},\ \ m\in M_0(\mathbb{A}). \end{align*}$$

We enumerate the roots $\Phi _G\cup \{0\}$ (including multiplicities) as $\alpha _i$ , $i\in \mathcal {I}_G$ for some index set $\mathcal {I}_G$ , and fix a basis $\mathcal {B}_G=(x_i)_{i\in \mathcal {I}_G}$ for such that for all i. Similarly, we enumerate the weights $\Psi _V$ including multiplicities as $\beta _j$ , $j\in \mathcal {J}_V$ , and fix a basis $\mathcal {B}_V=(y_j)_{j\in \mathcal {J}_V}$ for V such that $y_j\in V_{\beta _j}$ for all j.

Let $\Gamma (V)$ be the (finite) set of subspaces of V that are sums of weight subspaces. Thus, every $U\in \Gamma (V)$ is of the form

$$\begin{align*}U=\oplus_{\beta\in\Psi_U}V_\beta, \end{align*}$$

where $\Psi _U$ is a subset of $\Psi _V$ . We will also write

(5.4) $$ \begin{align} \mathcal{J}_{V/U}=\{j\in\mathcal{J}_V\mid\beta_j\notin\Psi_U\}. \end{align} $$

Lemma 5.2. Let $\psi \in \mathbb {Z}_{\ge 0}\mathfrak {X}$ , so that $f=\operatorname {RI}_\psi $ is a polynomial. Let $U\in \Gamma (V)$ be such that $f\big |_U\not \equiv 0$ . Then,

  1. 1. $\psi \in \mathbb {Z}_{\ge 0}\Psi _U$ .

  2. 2. There exists $U'\in \Gamma (V)$ , $U'\subset U$ such that $f\big |_{U'}\not \equiv 0$ and $\psi \in \mathbb {R}_{>0}\Psi _{U'}$ .

Proof. Let $\mathcal {B}_{V^*}=(y_j^*)_{j\in \mathcal {J}_V}$ be the basis of $V^*$ dual to $\mathcal {B}_V$ and let $\mathcal {B}_{F[V]}$ be the basis of $F[V]$ consisting of monomials in $\mathcal {B}_{V^*}$ . By definition, the weight of a monomial $y_{j_1}^*\cdots y_{j_k}^*$ in $\mathcal {B}_{F[V]}$ is the sum $\beta _{j_1}+\dots +\beta _{j_k}$ . We view $F[U]$ as a subalgebra of $F[V]$ using the unique projection $V\rightarrow U$ with kernel in $\Gamma (V)$ . The basis $\mathcal {B}_{F[V]}$ contains $\mathcal {B}_{F[U]}$ (defined in the obvious way).

Let $h=f\big |_U$ . Expand h in the basis $\mathcal {B}_{F[U]}$ . Since $f=\operatorname {RI}_\psi $ , the weight of any basis element with nonzero coefficient is $\psi $ . The first part follows.

Let $U'$ be the smallest subspace in $\Gamma (V)$ such that $h\in F[U']$ (viewed as a subspace of $F[U]$ via the projection $U\rightarrow U'$ with kernel in $\Gamma (V)$ ). Then, $\Psi _{U'}$ is the union of $\{\beta _{j_1},\dots ,\beta _{j_k}\}$ as we range over all the monomials $y_{j_1}^*\dots y_{j_k}^*$ in $\mathcal {B}_{F(U)}$ with nonzero coefficients in the expansion of h. Once again, since $f=\operatorname {RI}_\psi $ , for any basis element with nonzero coefficient as above we have $\beta _{j_1}+\dots +\beta _{j_k}=\psi $ . Averaging over all these expressions of $\psi $ , we deduce that $\psi \in \mathbb {R}_{>0}\Psi _{U'}$ .

Consider the set $\Gamma _{\operatorname {reg}}(V)$ of subspaces in $\Gamma (V)$ that contain a regular element:

$$\begin{align*}\Gamma_{\operatorname{reg}}(V)=\{U\in\Gamma(V)\mid U\cap X\ne\emptyset\}. \end{align*}$$

Corollary 5.3. Let $U\in \Gamma _{\operatorname {reg}}(V)$ . Then,

  1. 1. $\mathfrak {X}\subset \mathbb {Z}_{\ge 0}\Psi _U$ .

  2. 2. Assume that V is regular. Let $\psi \in \mathbb {Z}_{>0}\mathfrak {X}$ . Then there exists $U'\in \Gamma _{\operatorname {reg}}(V)$ , $U'\subset U$ such that $\psi \in \mathbb {R}_{>0}\Psi _{U'}$ .

Proof. Let $\psi \in \mathbb {Z}_{\ge 0}\mathfrak {X}$ and let $f=\operatorname {RI}_\psi $ . Since $U\in \Gamma _{\operatorname {reg}}(V)$ , the restriction of f to U is not identically zero. By Lemma 5.2, $\psi \in \mathbb {Z}_{\ge 0}\Psi _U$ , which implies the first part, and there exists $U'\subset U$ , $U'\in \Gamma (V)$ such that $f\big |_{U'}\not \equiv 0$ and $\psi \in \mathbb {R}_{>0}\Psi _{U'}$ .

If V is regular and $\psi \in \mathbb {Z}_{>0}\mathfrak {X}$ , then $X=\{v\in V\mid f(v)\ne 0\}$ . Hence, $U'\cap X\ne \emptyset $ . The second part follows.

A connected algebraic F-subgroup R of G is called $T_0$ -saturated if its Lie algebra is of the form for some subset S of $\Phi _G\cup \{0\}$ . In this case, we write $\Phi _R = S \setminus \{0\}$ . Examples of $T_0$ -saturated subgroups are parabolic F-subgroups containing $T_0$ and their unipotent radicals, and semistandard Levi subgroups. If R is $T_0$ -saturated, then we write

(5.5)

For any subspace $U\subset V$ denote by $\operatorname {Stab}(U)=\{g\in G\mid \rho (g)U=U\}$ the stabilizer of U. The Lie algebra of $\operatorname {Stab}(U)$ is [Reference Borel2, Lemma II.7.4].

Let $U\in \Gamma (V)$ and let R be a $T_0$ -saturated subgroup of G. We denote by $R\star U$ the smallest R-stable space in $\Gamma (V)$ containing U.

The following is an immediate consequence of (5.3).

Lemma 5.4. Suppose that S is a subset of $\Psi _V$ containing $\Psi _U$ such that for all $\beta \in S$ and $\alpha \in \Phi _R$ with $\beta +\alpha \in \Psi _V$ , we have $\beta +\alpha \in S$ . Then, $\Psi _{R\star U}\subset S$ . In particular, for every $\beta \in \Psi _{R\star U}$ there exist $\gamma \in \Psi _U$ and $\alpha _1,\dots ,\alpha _k\in \Phi _R$ such that $\beta =\gamma +\alpha _1+\dots +\alpha _k$ .

Remark 5.5. It will follow from Lemma 6.9 below that $\Psi _{R\star U}$ is the smallest subset S of $\Psi _V$ containing $\Psi _U$ with the property that for every $\beta \in S$ and $\alpha \in \Phi _R$ such that $\beta +\alpha \in \Psi _V$ , we have $\beta +\alpha \in S$ .

5.4

For any $U\in \Gamma (V)$ define the gauge

(5.6) $$ \begin{align} \lambda(U)=\delta_0+\sum_{\beta\in\Psi_V\setminus\Psi_U}n_\beta\beta=\delta_0+\delta_V-\sum_{\beta\in\Psi_U}n_\beta\beta\in X^*(T_0). \end{align} $$

It will turn out to be a key parameter governing the growth of the subsum of $\theta _\phi $ pertaining to U.

The following combinatorial-geometric lemma is crucial.

Lemma 5.6. Let $U\in \Gamma _{\operatorname {reg}}(V)$ . Then, there exists a one-to-one function

$$\begin{align*}\iota:\mathcal{J}_V\rightarrow\mathcal{I}_G, \end{align*}$$

such that

$$\begin{align*}\beta_j+\alpha_{\iota(j)}\in\Psi_U, \end{align*}$$

for all $j\in \mathcal {J}_V$ .

Proof. Let $v_0\in U\cap X$ . Recall the surjection $D_{v_0}$ (2.1). Since $v_0\in U$ , for all $\alpha \in \Phi _G\cup \{0\}$ by (5.3). Writing the matrix of $D_{v_0}$ with respect to the bases $\mathcal {B}_G$ and $\mathcal {B}_V$ , a nonzero entry at a position $(j,i)$ , $j\in \mathcal {J}_V$ , $i\in \mathcal {I}_G$ , can only occur if $\beta _j - \alpha _i \in \Psi _U$ . A non-zero minor of order $\dim V$ therefore gives rise to a function $\iota $ as required.

Note that we used that G is reductive in the proof, namely that $-\Phi _G=\Phi _G$ .

Corollary 5.7. Let $U\in \Gamma _{\operatorname {reg}}(V)$ . Then, for any $J\subset \mathcal {J}_V$ we have

$$\begin{align*}\delta_0+\sum_{j\in J}\beta_j\in\mathbb{R}_{\ge0}(\Psi_U\cup\Delta_0). \end{align*}$$

In particular, we have

(5.7) $$ \begin{align} \lambda(U)\in\mathbb{R}_{\ge0}(\Psi_U\cup\Delta_0). \end{align} $$

Proof. We write

$$\begin{align*}\delta_0+\sum_{j\in J}\beta_j=\delta_0-\sum_{j\in J}\alpha_{\iota(j)}+\sum_{j\in J} (\beta_j+\alpha_{\iota(j)}), \end{align*}$$

and observe that $\delta _0-\sum _{j\in J}\alpha _{\iota (j)}\in \mathbb {R}_{\ge 0}\Delta _0$ since $\iota $ is injective.

Taking $J=\mathcal {J}_{V/U}$ (cf. (5.4) and (5.6)) we obtain (5.7).

5.5

The proof of Proposition 5.1 is based on a standard bound for lattice sums.

For any $\xi \in V$ denote by $\xi _\beta \in V_\beta $ , $\beta \in \Psi _V$ the components of $\xi $ with respect to the decomposition (5.2b).

We write $x_\pm =\max (0,\pm x)$ for the positive and negative parts of a real number x, so that $x=x_+-x_-$ .

For simplicity, we write $A(\cdot ,\phi )\ll _{N,\phi }^{\operatorname {rd}} B_N(\cdot )$ if for every $N\ge 0$ there exists a continuous seminorm $\nu _N$ on $C_{\operatorname {rd}}(V(\mathbb {A}))$ such that $\left |{A(\cdot ,\phi )}\right |\le B_N(\cdot )\nu _N(\phi )$ for every $\phi \in C_{\operatorname {rd}}(V(\mathbb {A}))$ .

Lemma 5.8. Let $\Omega $ be a compact subset of $G(\mathbb {A})$ . Let $U\in \Gamma (V)$ . Then,

$$\begin{align*}\sum_{\xi\in U: \, \xi_{\beta} \neq 0\, \forall\beta \in\Psi_U} \left|{\phi(\rho(g)^{-1}\xi)}\right| \ll_{N,\phi}^{\operatorname{rd}} e^{\sum_{\beta\in\Psi_U} (n_\beta\left\langle{\beta},{H_0(g)}\right\rangle_+ - N\left\langle{\beta},{H_0(g)}\right\rangle_-)},\ \ g \in T_0(\mathbb{A})\Omega. \end{align*}$$

Equivalently,

$$\begin{align*}e^{\left\langle{\lambda(U)},{H_0(g)}\right\rangle}\left|{\det\rho(g)}\right|{}^{-1}\sum_{\xi\in U: \, \xi_{\beta} \neq 0\, \forall\beta \in\Psi_U} \left|{\phi(\rho(g)^{-1}\xi)}\right| \ll_{N,\phi}^{\operatorname{rd}} \delta_{P_0}(m_0(g))e^{-N\sum_{\beta\in\Psi_U}\left\langle{\beta},{H_0(g)}\right\rangle_-}. \end{align*}$$

In particular, for $U=V$

$$\begin{align*}\left|{\det\rho(g)}\right|{}^{-1}\sum_{\xi\in V: \, \xi_{\beta} \neq 0\, \forall\beta \in\Psi_V} \left|{\phi(\rho(g)^{-1}\xi)}\right| \ll_{N,\phi}^{\operatorname{rd}} \min(1,\left|{\det\rho(g)}\right|)^N. \end{align*}$$

Proof. The first two statements are clearly equivalent by the definition (5.6) of $\lambda (U)$ . Since $\rho $ acts continuously on $C_{\operatorname {rd}}(V(\mathbb {A}))$ , we may assume without loss of generality that $g\in T_0(\mathbb {A})$ . (In fact, since both sides are left $T_0(F)$ -invariant, we may assume that $g\in T_0(\mathbb {R})$ where $\mathbb {R}$ is viewed as a subring of $\mathbb {A}=\mathbb {A}_{\mathbb {Q}}\otimes F$ via $x\mapsto x\otimes 1$ .)

Upon replacing $\phi $ by $\prod _{\beta \in \Psi _V}(1+\lVert v_{\beta ,\infty }\rVert )^{-n}\mathbf {1}_{K_\beta }(v_{\beta ,\operatorname {fin}})$ , $n\gg 0$ where $K_\beta $ is a compact open subgroup of $V_\beta (\mathbb {A}_{\operatorname {fin}})$ for every $\beta \in \Psi _V$ , we may also assume that $\phi \big |_{U (\mathbb {A})}$ factors as a product $\phi (v) = \prod _{\beta \in \Psi _U} \phi _\beta (v_\beta )$ , $v \in U (\mathbb {A})$ , where each factor $\phi _\beta \in C_{\operatorname {rd}}(V_\beta (\mathbb {A}))$ is nonnegative real. Since $T_0$ acts by scalars on each space $V_\beta $ , we are reduced to the assertion (for $t\in \mathbb {A}^\times $ )

$$\begin{align*}\sum_{\xi_{\beta} \neq 0} \phi_\beta (t^{-1}\xi_{\beta})\ll_{N,\phi_\beta}^{\operatorname{rd}} \begin{cases}\left|{t}\right|{}^{n_\beta}&\left|{t}\right| \ge 1,\\ \left|{t}\right|{}^{N}&\left|{t}\right| \le 1.\end{cases} \end{align*}$$

In turn, this immediately reduces to the following elementary statement for $\varphi \in C_{\operatorname {rd}}(\mathbb {A})$ :

$$\begin{align*}\sum_{\xi\in F^\times}\varphi(t^{-1}\xi)\ll_{N,\varphi}^{\operatorname{rd}}\begin{cases}\left|{t}\right|&\left|{t}\right|\ge1\\\left|{t}\right|{}^N&\left|{t}\right|\le1.\end{cases} \end{align*}$$

To prove this, note that given a compact subset K of $\mathbb {A}$ , we have

$$\begin{align*}\#(F^\times\cap tK)\ll\left|{t}\right|. \end{align*}$$

In particular, $F^\times \cap tK=\emptyset $ if $\left |{t}\right |$ is sufficiently small. Thus, given a compact subset $K_{\operatorname {fin}}$ of $\mathbb {A}_{\operatorname {fin}}$ ,

$$\begin{align*}\#\{\xi\in F^\times\mid (t^{-1}\xi)_{\operatorname{fin}}\in K_{\operatorname{fin}}\text{ and }\lVert(t^{-1}\xi)_\infty\rVert\le R\}\ll R^{[F:\mathbb{Q}]}\left|{t}\right|,\ \ R\ge0. \end{align*}$$

The statement above follows easily from this estimate and the fact that $\varphi \in C_{\operatorname {rd}}(\mathbb {A})$ .

Proof of Proposition 5.1

Define the support of an element $\xi \in V$ to be

$$\begin{align*}\operatorname{supp}\xi=\{\beta\in\Psi_V\mid\xi_\beta\ne0\}, \end{align*}$$

so that for any $U\in \Gamma (V)$ , $\xi \in U\iff \operatorname {supp}\xi \subset \Psi _U$ .

We decompose $\theta _\phi $ according to the support of the summands, i.e., we write

(5.8) $$ \begin{align} \theta_\phi(g)=\sum_{U\in\Gamma(V)}\Theta_U(g)\text{ where } \Theta_U(g)=\sum_{\xi\in X(F):\,\operatorname{supp}\xi=\Psi_U}\phi(\rho(g)^{-1}\xi). \end{align} $$

Of course, $\Theta _U$ is no longer $G(F)$ -invariant or even $P_0(F)$ -invariant, but it is $T_0(F)$ -invariant.

At any rate, for the proof of Proposition 5.1 we may assume by (5.1) that $g\in (T_0(\mathbb {A})\cap \mathfrak {S})\Omega $ , which we will do throughout.

Fix $U\in \Gamma (V)$ . By Lemma 5.8, we have

(5.9) $$ \begin{align} \Theta_U(g)\ll_{N,\phi}^{\operatorname{rd}} \delta_{P_0}(m_0(g))\left|{\det\rho(g)}\right|e^{-\left\langle{\lambda(U)},{H_0(g)}\right\rangle - N \sum_{\beta\in\Psi_U} \left\langle{\beta},{H_0(g)}\right\rangle_-}. \end{align} $$

Moreover, we may assume that $U\in \Gamma _{\operatorname {reg}}(V)$ , for otherwise $\Theta _U(g)=0$ . By (5.7), we have

$$\begin{align*}\lambda(U)=\sum_{\beta\in\Psi_U}c_\beta\beta+\sum_{\alpha\in\Delta_0}d_\alpha\alpha, \end{align*}$$

with suitable coefficients $c_\beta ,d_\alpha \ge 0$ . Since $g\in (T_0(\mathbb {A})\cap \mathfrak {S})\Omega $ , we deduce that

$$\begin{align*}\Theta_U(g)\ll_{N,\phi}^{\operatorname{rd}} \delta_{P_0}(m_0(g))\left|{\det\rho(g)}\right|e^{-\sum_{\beta\in\Psi_U}(c_\beta \left\langle{\beta},{H_0(g)}\right\rangle_+ +N\left\langle{\beta},{H_0(g)}\right\rangle_-)}. \end{align*}$$

In particular,

(5.10) $$ \begin{align} \Theta_U(g)\ll_\phi^{\operatorname{rd}} \delta_{P_0}(m_0(g))\left|{\det\rho(g)}\right|. \end{align} $$

The proposition follows.

6 Special subspaces

In this section, we provide some geometric prerequisites to analyze further the behavior of $\theta _\phi $ on the Siegel domain. A crucial role will be played by certain subspaces of V that are PVSs for parabolic subgroups of G (cf. Definition 6.5 below).

6.1

We start with the following elementary fact from convex geometry.

Lemma 6.1. Let W be a finite-dimensional real vector space and let $v_i$ , $i\in I$ be finitely many vectors in W. Given $\lambda \in \mathbb {R}_{\ge 0}\{v_i, i\in I\}$ we write $\lambda =\sum _{i\in I}\lambda _i v_i$ with $\lambda _i\ge 0$ for all i and with maximal number of nonzero (i.e., positive) coefficients. Then,

  1. 1. $I':=\{i\in I\mid \lambda _i>0\}$ is uniquely determined by $\lambda $ .

  2. 2. Suppose that $\lambda = \sum _{i\in I}\lambda ^{\prime }_i v_i$ with $\lambda ^{\prime }_i \in \mathbb {R}$ , $i \in I$ , such that $\lambda ^{\prime }_i\ge 0$ for all $i\notin I'$ . Then, $\lambda ^{\prime }_i=0$ for all $i\notin I'$ .

  3. 3. In particular, for any $i\in I$ , if $v_i$ is in the linear span of $\{v_j\}_{j\in I'}$ then $i\in I'$ .

In the situation of Lemma 6.1, we will say that $v_i$ , $i\in I'$ is the positive envelope of $\lambda $ with respect to $v_i$ , $i\in I$ .

Fix $U\in \Gamma _{\operatorname {reg}}(V)$ . Recall that $\lambda (U) \in X^*(T_0)$ was defined in (5.6) and that $\lambda (U)\in \mathbb {R}_{\ge 0}(\Delta _0\cup \Psi _U)$ by (5.7). Let $\mathcal {C}_U \subset \mathfrak {a}_0$ be the cone dual to $\mathbb {R}_{\ge 0}(\Delta _0\cup \Psi _U)$ , that is,

(6.1) $$ \begin{align} \mathcal{C}_U=\{x\in\mathfrak{a}_{0,+}\mid\left\langle{\beta},{x}\right\rangle\ge0\ \forall\beta\in\Psi_U\}. \end{align} $$

Let $\mathcal {F}_U$ be the face of $\mathcal {C}_U$ defined by $\lambda (U)$ , i.e.,

$$\begin{align*}\mathcal{F}_U=\{x\in\mathcal{C}_U\mid\left\langle{\lambda(U)},{x}\right\rangle=0\}. \end{align*}$$

Roughly speaking, $\mathcal {F}_U$ describes the directions in $T_0(F)\backslash (T_0(\mathbb {A})\cap \mathfrak {S})$ where the bounded function $\delta _{P_0}^{-1}\left |{\det \rho }\right |{}^{-1}\Theta _U$ (see (5.8)) does not decay. To analyze $\mathcal {F}_U$ , let $A\subset \Delta _0\cup \Psi _U$ be the positive envelope of $\lambda (U)$ with respect to $\Delta _0\cup \Psi _U$ . Let $I=A\cap \Delta _0$ and $J=A\cap \Psi _U$ .

(Note that we do not assume at the outset that $\Delta _0$ and $\Psi _U$ are disjoint.)

Definition 6.2. Let $\mathbf {E}(U)$ (the enveloping group of U) be the standard parabolic subgroup of G such that $I=\Delta _0^{\mathbf {E}(U)}$ .

Then,

$$\begin{align*}\mathcal{F}_U=\{x\in\mathcal{C}_U\mid\left\langle{\alpha},{x}\right\rangle=\left\langle{\beta},{x}\right\rangle=0\ \forall\alpha\in\Delta_0^{\mathbf{E}(U)},\beta\in J\}. \end{align*}$$

Remark 6.3. Let $U\in \Gamma _{\operatorname {reg}}(V)$ and let $\iota $ be as in Lemma 5.6. Then, the image under $\iota $ of $\mathcal {J}_{V/U}$ contains $\mathcal {I}_{N_{\mathbf {E}(U)}}$ and is contained in $\mathcal {I}_{\mathbf {E}(U)}$ .

Indeed, as in the proof of Corollary 5.7 we write

$$\begin{align*}\lambda (U) =\sum_{j\in \mathcal{J}_{V/U}} (\beta_j+\alpha_{\iota(j)}) + \delta_0-\sum_{j\in \mathcal{J}_{V/U}}\alpha_{\iota(j)}. \end{align*}$$

It follows from the definition of $\mathbf {E}(U)$ that $\delta _0-\sum _{j\in \mathcal {J}_{V/U}}\alpha _{\iota (j)}$ can be written as a linear combination of $\Delta _0^{\mathbf {E}(U)}$ . This precisely means that $\mathcal {I}_{N_{\mathbf {E}(U)}}\subset \iota (\mathcal {J}_{V/U})\subset \mathcal {I}_{\mathbf {E}(U)}$ .

6.2

To go further, we need a follow-up of Lemma 5.6. It will play a key role.

Proposition 6.4. Let $U\in \Gamma _{\operatorname {reg}}(V)$ . Assume that $P_0$ stabilizes U, so that the stabilizer $S=\operatorname {Stab}(U)$ of U in G is a standard F-parabolic subgroup of G. Then,

  1. 1. There exists a one-to-one function

    $$\begin{align*}\iota:\mathcal{J}_{V/U}\rightarrow\mathcal{I}_{N_S}, \end{align*}$$
    such that $\beta _j+\alpha _{\iota (j)}\in \Psi _U$ for all $j\in \mathcal {J}_{V/U}$ . In particular, $\dim V/U\le \dim G/S$ .
  2. 2. $S\subset \mathbf {E}(U)$ .

  3. 3. The following conditions are equivalent.

    1. (a) $\dim V/U=\dim G/S$ .

    2. (b) $\iota $ is a bijection.

    3. (c) $\lambda (U)=\sum _{j\in \mathcal {J}_{V/U}}(\beta _j+\alpha _{\iota (j)})+\delta _0^S$ .

    4. (d) For some (or every) $v\in U\cap X$ , $\rho (S)v$ is dense in U and $G_v^\circ \subset S$ .

    5. (e) For some (or every) $v\in U\cap X$ , the variety

      $$\begin{align*}G_{v,U}=\{g\in G\mid\rho(g)^{-1}v\in U\}, \end{align*}$$
      is the union of finitely many left cosets of S in G.
  4. 4. If the conditions in part 4 are satisfied, then U is a PVS with respect to S with open orbit $U^{\operatorname {reg}}=U\cap X$ . Moreover, $G_{v,U}=G_vS$ for any $v\in U^{\operatorname {reg}}$ .

    On the other hand, if the conditions in part 4 are not satisfied, then $\mathbf {E}(U) \neq S$ .

Proof. Write for the Lie algebra of S. Let $N=N_S$ , and let $\bar {N}$ be the unipotent radical of the parabolic subgroup of G opposite to S. Denote the corresponding Lie algebras by and .

Let $v_0\in U\cap X$ . The composition of the surjection $D_{v_0}$ (2.1) with the projection $V\rightarrow V/U$ factors as a surjective map

We have for all $\alpha \in \Phi _{\bar {N}}=-\Phi _N$ . The existence of $\iota $ follows as in the proof of Lemma 5.6. Moreover,

$$\begin{align*}\lambda(U)=\sum_{\beta\in\Psi_V\setminus\Psi_U}n_\beta\beta+\delta_0= \sum_{j\in\mathcal{J}_{V/U}}(\beta_j+\alpha_{\iota(j)})+\delta_0-\sum_{j\in\mathcal{J}_{V/U}}\alpha_{\iota(j)}. \end{align*}$$

Clearly,

where $\mu =\sum _{i\in \mathcal {I}_N\setminus \iota (\mathcal {J}_{V/U})}\alpha _i$ is a (possibly empty) sum of roots in N. Therefore, $S\subset \mathbf {E}(U)$ , and the inclusion is proper if $\mu \ne 0$ .

Note that the following statements are equivalent: $\mu =0$ ; $\iota $ is a bijection; $\dim V/U=\dim G/S$ ; $\tilde D_{v_0}$ is an isomorphism of vector spaces.

Observe that the kernel of $\tilde {D}_{v_0}$ is and we have a short exact sequence

Thus, $\tilde {D}_{v_0}$ is an isomorphism if and only if and . Since $\operatorname {Ker} D_{v_0}=\operatorname {Lie} G_{v_0}$ , the first condition means that $G_{v_0}^\circ \subset S$ , while the second condition means that the S-orbit of $v_0$ is dense in U. This proves that conditions (33a)–(33d) are equivalent and that the validity of (33d) does not depend on the choice of $v\in U\cap X$ . Therefore, under these conditions, $U\cap X$ is a single S-orbit, since $U\cap X$ is irreducible. Since $U\cap X=\rho (G_{v_0,U})^{-1}v_0$ , this also implies that $G_{v_0,U}=G_{v_0}S$ , which is a finite union of left S-cosets since $G_{v_0}^\circ \subset S$ .

Finally, if $G_{v_0,U}=\bigcup _{i=1}^n \eta _i S$ is the union of finitely many left S-cosets, then

$$\begin{align*}U\cap X=\rho(G_{v_0,U})^{-1}v_0=\bigcup_{i=1}^n \rho(S)\rho(\eta_i)^{-1}v_0, \end{align*}$$

is the union of finitely many S-orbits, at least one of which must be dense, while $G_{\rho (\eta _i)^{-1}v_0}^\circ \subset S$ for every $i = 1, \ldots , n$ , since $G_{\rho (\eta _i)^{-1}v_0}=\eta _i^{-1}G_{v_0}\eta _i\subset \eta _i^{-1}G_{v_0,U}$ is contained in a finite union of left S-cosets. Thus, condition (e) is also equivalent to the other ones. This concludes the proof of the proposition.

As we shall see, the spaces considered in the proposition above deserve special attention.

Definition 6.5. We denote by $\Gamma _{\operatorname {spcl}}(V)$ the set of spaces satisfying the conditions of Proposition 6.4 part 4. Thus, $\Gamma _{\operatorname {spcl}}(V)$ consists of the subspaces $U\in \Gamma _{\operatorname {reg}}(V)$ that are stabilized by $P_0$ and satisfy $\dim V/U=\dim G/\operatorname {Stab}(U)$ .

A subspace in $\Gamma _{\operatorname {spcl}}(V)$ will be called special.

Remark 6.6. Trivially, $V\in \Gamma _{\operatorname {spcl}}(V)$ . It follows from Proposition 6.4 that if V is basic (see Definition 2.8), then $\Gamma _{\operatorname {spcl}}(V)=\{V\}$ .

The converse is not true: $V=\operatorname {Mat}_n$ is a PVS with respect to $G=\operatorname {GL}_n$ acting by matrix multiplication. Here $H=1$ , but it is easy to see that $\Gamma _{\operatorname {spcl}}(V)=\{V\}$ .

In fact, the four examples in Table 2 of §3.3 provide counterexamples for irreducible PVS of DK-type. Indeed, by Proposition 6.4, if $U\in \Gamma _{\operatorname {spcl}}(V)$ , then it is stable under the maximal standard subgroup parabolic Q of G whose Levi part L satisfies $H^\circ =L^{\operatorname {der}}$ . In particular, U is stable under L. However, V decomposes into four irreducible subrepresentations of L, $V=U_1\oplus U_2\oplus U_3\oplus U_4$ where $U_1,U_2$ are one-dimensional, $\Psi _{U_1}$ consists of the highest weight of V, and $\dim U_3=\dim U_4=\dim G/Q$ . Since every non-trivial $P_0$ -stable subspace of V contains $U_1$ , we cannot have $\dim V/U=\dim G/Q$ . Hence $U=V$ .

Consider the list of reduced, absolutely irreducible PVSs of Sato and Kimura in the appendix to [Reference Kimura23]. Of the 29 regular cases, 23 are of DK-type.Footnote 8 Among these, we have $\Gamma _{\operatorname {spcl}}(V) \neq \{V \}$ only in the following eight cases: # 2 (Example 3.4 for $k=1$ and $n_1=2$ ), # 4 (Example 3.7), # 8 (Example 3.8), # 9, # 11, # 12 (Example 3.9), # 15 (Example 3.5 for $k=1$ and $n_1=2$ or $n_2-2$ ), # 28. (To verify the assertion $\Gamma _{\operatorname {spcl}}(V) \neq \{V \}$ in the remaining cases # 9, # 11 and # 28, one may use the construction of §9.)

Remark 6.7. In general, we may have $\mathbf {E}(U)\supsetneq \operatorname {Stab}(U)$ even if $U\in \Gamma _{\operatorname {spcl}}(V)$ . See Examples 6.15 and 7.8 below.

In principle, using Proposition 6.4 one can enumerate $\Gamma _{\operatorname {spcl}}(V)$ , at least in small rank cases or when H is contained in only a few parabolic subgroups.

Example 6.8. Consider Example 3.4 for $k=1$ , $n_1=2$ and Example 3.7. Write the simple roots of $G'$ as $\{\alpha ,\beta \}$ where $\alpha $ is the short root. Then $\Delta _0^G=\{\alpha \}$ and $\Psi _V=\{\beta ,\beta +\alpha ,\beta +2\alpha [,\beta +3\alpha ]\}$ where root in square brackets occurs for $G'=G_2$ . It is easy to see that $\Gamma _{\operatorname {spcl}}(V)=\{V,U\}$ where $\Psi _U=\Psi _V\setminus \{\beta \}$ . $\operatorname {Stab}(U)=\mathbf {E}(U)$ is the Borel subgroup of G.

6.3

For $\alpha \in \Phi _G$ denote by $N_{(\alpha )}$ the unipotent subgroup of G with Lie algebra if $2\alpha \in \Phi _G$ and otherwise.

The following general fact about representations is probably well-known.

Lemma 6.9. Let $U\in \Gamma (V)$ and $\alpha \in \Phi _G$ . Suppose that U is stable under $N_{(\alpha )}$ . Then, for every $\beta \in \Psi _U$ such that $\alpha +\beta \in \Psi _V$ we have $\alpha +\beta \in \Psi _U$ .

This follows from the representation theory of $\operatorname {SL}_2$ . To explain this, we need a standard consequence of the Jacobson–Morozov Theorem.

Lemma 6.10. Let $\alpha \in \Phi _G$ and let . Then, there exists a unique F-morphism $\varphi :\operatorname {SL}_2 \rightarrow G$ such that $d\varphi ({\bigl (\begin{smallmatrix}{0}&{1}\\{0}&{0}\end{smallmatrix}\bigr )})=e$ , the image of the diagonal torus under $\varphi $ is contained in $T_0$ and the co-character $t\mapsto \varphi ({\bigl (\begin{smallmatrix}{t}&{}\\{}&{t^{-1}}\end{smallmatrix}\bigr )})$ is $\alpha ^\vee $ .

Proof. This is well-known. For convenience we include the proof.

For the uniqueness of $\varphi $ – see [Reference Collingwood and McGovern8, Lemma 3.4.4].

By passing to the subgroup generated by $N_{(\alpha )}$ and $N_{(-\alpha )}$ , we may assume that G is semisimple of F-rank one. We may also assume without loss of generality that $\alpha>0$ , so that $N_{(\alpha )}\subset N_0$ .

By the Jacobson–Morozov Theorem, there exists an F-morphism $\varphi ':\operatorname {SL}_2 \rightarrow G$ such that $d\varphi '({\bigl (\begin{smallmatrix}{0}&{1}\\{0}&{0}\end{smallmatrix}\bigr )})=e$ . Let a be the restriction of $\varphi '$ to the diagonal torus, i.e., $a(s)=\varphi '({\bigl (\begin{smallmatrix}{s}&{}\\{}&{s^{-1}}\end{smallmatrix}\bigr )})$ , so that $\operatorname {Ad}(a(s))e=s^2 e$ . Let T be the image of a, which is a maximal split torus in G. By the Bruhat decomposition, for any $g\notin P_0$ we have $gN_0g^{-1}\cap N_0=1$ . Thus, $T\subset P_0$ . It follows that $T=nT_0n^{-1}$ for some element n of $N_0(F)$ .

We claim that n centralizes e. Indeed, for every $t\in T_0$ , both $\operatorname {Ad}(t)e$ and $\operatorname {Ad}(ntn^{-1})e$ are scalar multiples of e. Fix $t\in T_0$ such that its centralizer in $N_0$ is trivial. Write $ntn^{-1}=n't$ where $n'=[n,t]\in N_0$ . Then, $\operatorname {Ad}(n')e$ is a scalar multiple of e, and hence $\operatorname {Ad}(n')e=e$ since $n'$ is unipotent. Note that the centralizer $N_{0,e}$ of e in $N_0$ is invariant under conjugation by $T_0$ (since $\operatorname {Ad}(T_0)$ acts by scalars on e). By [Reference Borel2, Proposition III.9.3] the maps $N_0\rightarrow N_0$ and $N_{0,e}\rightarrow N_{0,e}$ given by $n\mapsto n'$ are bijective. Since $n'\in N_{0,e}$ , we infer that $n\in N_{0,e}$ as claimed.

We can now take $\varphi =\varphi '(n^{-1}\cdot n)$ .

Proof of Lemma 6.9

For any integer k let $V_k\in \Gamma (V)$ be such that

$$\begin{align*}\Psi_{V_k}=\{\beta\in\Psi_V\mid\left\langle{\beta},{\alpha^\vee}\right\rangle=k\}. \end{align*}$$

Fix . Using Lemma 6.10 and basic facts about representation theory of $\operatorname {SL}_2$ , it follows that for any integer k, the map

$$\begin{align*}d\rho(e):V_{k-1}\rightarrow V_{k+1}\text{ is } \begin{cases}\text{surjective}&\text{if }k\ge0,\\\text{injective}&\text{if }k\le0,\end{cases} \end{align*}$$

Thus, for every $\beta \in \Psi _V$ ,

$$\begin{align*}d\rho(e):V_\beta\rightarrow V_{\beta+\alpha}\text{ is } \begin{cases}\text{surjective}&\text{if }\left\langle{\beta},{\alpha^\vee}\right\rangle+1\ge0,\\\text{injective}&\text{if }\left\langle{\beta},{\alpha^\vee}\right\rangle+1\le0.\end{cases} \end{align*}$$

In particular, $d\rho (e)(V_\beta )\ne 0$ if $V_{\beta +\alpha }\ne 0$ . The lemma follows.

Lemma 6.11. Suppose that $U_1,U_2\in \Gamma _{\operatorname {spcl}}(V)$ and $U_1\cap U_2\in \Gamma _{\operatorname {reg}}(V)$ . Then, $U_1\cap U_2\in \Gamma _{\operatorname {spcl}}(V)$ and $\operatorname {Stab}(U_1\cap U_2)=\operatorname {Stab}(U_1)\cap \operatorname {Stab}(U_2)$ .

Proof. Let $U=U_1\cap U_2$ , $S=\operatorname {Stab}(U)$ , $N=N_S$ , $S_k=\operatorname {Stab}(U_k)$ , $N_k=N_{S_k}$ , $M_k=M_{S_k}$ , $k=1,2$ . Clearly, $\Psi _U=\Psi _{U_1}\cap \Psi _{U_2}$ , $P_0$ stabilizes U and $S\supset S_1\cap S_2$ . In particular, $\Phi _N\subset \Phi _{N_1}\cup \Phi _{N_2}$ .

By Proposition 6.4, there exists a one-to-one function

$$\begin{align*}\iota:\mathcal{J}_{V/U}\rightarrow\mathcal{I}_N, \end{align*}$$

such that $\beta _j+\alpha _{\iota (j)}\in \Psi _U$ for all $j\in \mathcal {J}_{V/U}$ .

We claim that if $\beta _j\in \Psi _V\setminus \Psi _{U_1}$ , then $\alpha _{\iota (j)}\in \Phi _{N_1}$ . Indeed, suppose otherwise. Then, $\alpha _{\iota (j)}\in \Phi _{M_1}$ (since $\alpha _{\iota (j)}>0$ as $\alpha _{\iota (j)}\in \Phi _N$ ). Therefore, using Lemma 6.9, $\beta _j=(\beta _j+\alpha _{\iota (j)})-\alpha _{\iota (j)}\in \Psi _{U_1}$ since $U_1$ is $S_1$ -stable, in contradiction to the assumption that $\beta _j\notin \Psi _{U_1}$ .

Thus, the restriction of $\iota $ to $\mathcal {J}_{V/U_1}$ is a bijection with $\mathcal {I}_{N_1}$ (since $U_1\in \Gamma _{\operatorname {spcl}}(V)$ ). Similarly, the restriction of $\iota $ to $\mathcal {J}_{V/U_2}$ is a bijection with $\mathcal {I}_{N_2}$ .

It follows that $\iota $ is onto and $\Phi _N=\Phi _{N_1}\cup \Phi _{N_2}$ . Thus, $U \in \Gamma _{\operatorname {spcl}}(V)$ and $N=N_1N_2$ , which implies that $S=S_1\cap S_2$ .

Lemma 6.12. Let $U\in \Gamma _{\operatorname {reg}}(V)$ and $\tilde U=\mathbf {E}(U)\star U$ . Then, $\tilde U\in \Gamma _{\operatorname {spcl}}(V)$ and $\mathbf {E}(\tilde U)=\operatorname {Stab}(\tilde U)=\mathbf {E}(U)$ .

Proof. Clearly $\tilde U\in \Gamma _{\operatorname {reg}}(V)$ , since $\tilde U\supset U$ , and $\operatorname {Stab}(\tilde U)\supset \mathbf {E}:=\mathbf {E}(U)$ . In view of Proposition 6.4, in order to prove the lemma we only need to show that $\tilde {\mathbf {E}}:=\mathbf {E}(\tilde U)\subset \mathbf {E}$ . We have $\Psi _{\tilde U}\supset \Psi _U$ . Write

$$\begin{align*}\lambda(\tilde U)= \sum_{\beta\in\Psi_{\tilde U}} c_\beta \beta +\sum_{\alpha\in \Delta_0^{\tilde{ \mathbf{E}}}} d_\alpha\alpha, \end{align*}$$

with $c_\beta \ge 0$ for all $\beta \in \Psi _{\tilde U}$ and $d_\alpha>0$ for all $\alpha \in \Delta _0^{\tilde {\mathbf {E}}}$ . Then,

$$\begin{align*}\lambda(U)= \sum_{\beta\in\Psi_U} c_\beta\beta + \sum_{\beta\in\Psi_{\tilde U}\setminus\Psi_U} (c_\beta+n_\beta)\beta + \sum_{\alpha\in \Delta_0^{\tilde {\mathbf{E}}}} d_\alpha \alpha. \end{align*}$$

Using Lemma 5.4 for each $\beta \in \Psi _{\tilde U}\setminus \Psi _U$ , we can rewrite the above relation as

$$\begin{align*}\lambda(U)= \sum_{\beta\in\Psi_U} c^{\prime}_\beta \beta +\sum_{\alpha\in \Delta_0} d^{\prime}_\alpha \alpha, \end{align*}$$

where $c^{\prime }_\beta \ge 0$ for all $\beta \in \Psi _U$ , $d^{\prime }_\alpha \ge 0$ for $\alpha \in \Delta _0\setminus \Delta _0^{\mathbf {E}}$ and $d^{\prime }_\alpha \ge d_\alpha>0$ for all $\alpha \in \Delta _0^{\tilde {\mathbf {E}}}\setminus \Delta _0^{\mathbf {E}}$ . By Lemma 6.1 we deduce that $d^{\prime }_\alpha =0$ for all $\alpha \in \Delta _0\setminus \Delta _0^{\mathbf {E}}$ . This implies that $\tilde {\mathbf {E}}\subset \mathbf {E}$ , as required.

Corollary 6.13.

  1. 1. If $U\in \Gamma _{\operatorname {reg}}(V)$ is stable under $\mathbf {E}(U)$ , then $U\in \Gamma _{\operatorname {spcl}}(V)$ .

  2. 2. If $U\in \Gamma _{\operatorname {reg}}(V)$ and $\mathbf {E}(U)\subsetneq G$ , then there exists $\tilde U\in \Gamma _{\operatorname {spcl}}(V)$ , $\tilde U\ne V$ , containing U.

  3. 3. In particular, if $\Gamma _{\operatorname {spcl}}(V)=\{V\}$ , then $\mathbf {E}(U)=G$ for every $U\in \Gamma _{\operatorname {reg}}(V)$ .

Example 6.14. Consider Example 3.2 with $k=3$ and $n_1<n_2$ . Then, it is easy to see that $\Gamma _{\operatorname {spcl}}(V)=\{V,V_1,V_2\}$ where

$$ \begin{align*} V_1&=\{(x_1,x_2)\mid\text{the first }n_2-n_1\text{ columns of }x_1\text{ vanish}\},\\ V_2&=\{(x_1,x_2)\mid\text{the last }n_2-n_3\text{ rows of }x_2\text{ vanish}\}. \end{align*} $$

The general case of Example 3.2 is far more complicated and will not be considered here.

Example 6.15. Consider Example 3.8 pertaining to $F_4$ . Write for simplicity $(i_1i_2i_3i_4)$ instead of $i_1\alpha _1+i_2\alpha _2+i_3\alpha _3+i_4\alpha _4\in X^*(T_0)$ .

We explicate $\Psi _V$ in a way that demonstrates its realization as pairs of $3\times 3$ -symmetric matrices as follows:

$$\begin{align*}\Psi_V=\left(\begin{smallmatrix}(0100)&(0110)&(0111)\\(0110)&(0120)&(0121)\\(0111)&(0121)&(0122)\end{smallmatrix}\right),\left(\begin{smallmatrix}(1100)&(1110)&(1111)\\(1110)&(1120)&(1121)\\(1111)&(1121)&(1122)\end{smallmatrix}\right). \end{align*}$$

Consider the subspaces $U_1,U_2,U_3$ in $\Gamma (V)$ given by

$$ \begin{align*} \Psi_{U_1}&=\Psi_V\setminus \{(0100),(1100)\},\\ \Psi_{U_2}&=\Psi_V\setminus \{(0100),(0110),(0120)\},\\ \Psi_{U_3}&=\Psi_V\setminus \{(0100),(0110),(0111)\}. \end{align*} $$

In matrices,

$$\begin{align*}U_1=\left(\begin{smallmatrix} 0&*&* \\ *&*&* \\ *&*&* \end{smallmatrix}\right), \left(\begin{smallmatrix} 0&*&* \\ *&*&* \\ *&*&* \end{smallmatrix}\right),\ \ U_2=\left(\begin{smallmatrix} 0&0&* \\ 0&0&* \\ *&*&* \end{smallmatrix}\right), \left(\begin{smallmatrix} *&*&* \\ *&*&* \\ *&*&* \end{smallmatrix}\right),\ \ U_3=\left(\begin{smallmatrix} 0&0&0 \\ 0&*&* \\ 0&*&* \end{smallmatrix}\right), \left(\begin{smallmatrix} *&*&* \\ *&*&* \\ *&*&* \end{smallmatrix}\right). \end{align*}$$

Let $U_4=U_1\cap U_2$ , $U_5=U_2\cap U_3$ so that

$$\begin{align*}U_4=\left(\begin{smallmatrix} 0&0&* \\ 0&0&* \\ *&*&* \end{smallmatrix}\right), \left(\begin{smallmatrix} 0&*&* \\ *&*&* \\ *&*&* \end{smallmatrix}\right),\ \ U_5=\left(\begin{smallmatrix} 0&0&0 \\ 0&0&* \\ 0&*&* \end{smallmatrix}\right), \left(\begin{smallmatrix} *&*&* \\ *&*&* \\ *&*&* \end{smallmatrix}\right). \end{align*}$$

Then, it is not difficult to check that

$$\begin{align*}\Gamma_{\operatorname{spcl}}(V)=\{V,U_1,U_2,U_3,U_4,U_5\}. \end{align*}$$

Moreover,

$$ \begin{align*} &\operatorname{Stab}(U_1)=\mathbf{E}(U_1)=(*00*),\ \operatorname{Stab}(U_2)=\mathbf{E}(U_2)=(00{*}0),\ \operatorname{Stab}(U_3)=\mathbf{E}(U_3)=(000*),\\ &\operatorname{Stab}(U_4)=\operatorname{Stab}(U_5)=\mathbf{E}(U_5)=(0000),\ \ \mathbf{E}(U_4)=(00{*}0). \end{align*} $$

7 Proof of Theorem 4.3

In this section, we prove Theorem 4.3 by providing an estimate on $\Theta _U$ which is finer than (5.10) for spaces $U \in \Gamma _{\operatorname {reg}} (V)$ containing non-isotropic elements (see Definition 2.7).

7.1

Let $\mathcal {C}$ be a closed cone in a finite-dimensional real vector space W.

Let f be a continuous real-valued function on W and let $d>0$ . We say that f is positively-d-homogeneous if $f(rv)=r^df(v)$ for all $r>0$ , $v\in W$ . (In the cases at hand, $d=1$ .) In this case, f is positive on $\mathcal {C}\setminus \{0\}$ if (and of course, only if) it is positive on the intersection of $\mathcal {C}$ with the unit sphere of W with respect to a prescribed norm.

The space of continuous positively-d-homogeneous functions on W is a Banach space with respect to the supremum norm on the unit ball and the positivity on $\mathcal {C}\setminus \{0\}$ is an open condition.

By integration using spherical coordinates, it is clear that if f is positively-d-homogeneous and positive on $\mathcal {C}\setminus \{0\}$ , then the integral of $e^{-f}$ over $\mathcal {C}$ converges.

Suppose that f is positively- $1$ -homogeneous and let $\lambda _j$ , $j\in J$ be a finite set of vectors in $W^*$ . Then, the positively- $1$ -homogeneous function $f+N\sum _{j\in J}\left \langle {\lambda _j},{v}\right \rangle _-$ is positive on $\mathcal {C}\setminus \{0\}$ for $N\gg 0$ if and only if f is positive on the subcone $\{v\in \mathcal {C}:\left \langle {\lambda _j},{v}\right \rangle \ge 0\ \forall j\in J\}\setminus \{0\}$ .

These facts yield the following elementary convergence result.

Lemma 7.1. Let $\chi _i$ , $i\in I$ and $\lambda _j$ , $j\in J$ be finite sets of vectors in $\mathfrak {a}_0^*$ . Let $W \subset \mathfrak {a}_0$ be a subspace and f a continuous positively- $1$ -homogeneous function on W. Assume that f is positive on the cone

$$\begin{align*}\{x\in\mathfrak{a}_{0,+} \cap W \mid\left\langle{\lambda_j},{x}\right\rangle\ge0\ \forall j\in J\}\setminus\{0\}. \end{align*}$$

Then, there exists $c>0$ such that for all $N\gg 0$ the integral

$$\begin{align*}\int_{\mathfrak{a}_{0,+} \cap W} e^{-f(H)-N\sum_{j\in J}\left\langle{\lambda_j},{H}\right\rangle_- - \sum_{i\in I}\sigma_i \left\langle{\chi_i},{H}\right\rangle} \ dH, \end{align*}$$

converges for $\left |{\sigma _i}\right |<c$ , $i\in I$ . Moreover, if $\chi _i\in \mathbb {R}_{\ge 0}\{\lambda _j\}_{j\in J}$ for all $i\in I$ , then for all $\sigma _i>-c$ , $i\in I$ the integral converges for sufficiently large N (depending on the maximum value of the $\sigma _i$ ’s).

7.2

We now turn to the support sets of non-isotropic elements.

Definition 7.2. We denote by $\Gamma _{\operatorname {niso}} (V) \subset \Gamma _{\operatorname {reg}} (V)$ the set of subspaces in $\Gamma (V)$ that contain a non-isotropic element.

Definition 7.3. Let $U\in \Gamma _{\operatorname {spcl}}(V)$ . We say that U is exceptional if there exists $U'\in \Gamma _{\operatorname {reg}}(V)$ , $U'\subset U$ such that the linear span of $X^*(G)\cup \Delta ^{\operatorname {Stab}(U)}_0\cup \Psi _{U'}$ is a proper subspace of $\mathfrak {a}_0^*$ . In this case, we will also say that $(U,U')$ is an exceptional pair.

We denote by $\Gamma _{\operatorname {excp}}(V)$ the set of exceptional subspaces of V.

Proposition 7.4. Let $U\in \Gamma _{\operatorname {spcl}}(V)\cap \Gamma _{\operatorname {niso}}(V)$ . Then U is not exceptional.

Conversely, suppose that V is regular. Then, for any isotropic $v\in X(F)$ there exists an exceptional pair $(U,U')$ such that $v\in \rho (G(F))U'$ .

Thus, if V is regular, then the set of isotropic elements in $X(F)$ is

$$\begin{align*}\cup_{U\in\Gamma_{\operatorname{excp}}(V)}\rho(G(F))U\cap X(F)=\cup_{(U,U')}\rho(G(F))U'\cap X(F), \end{align*}$$

where $(U,U')$ range over the exceptional pairs.

Remark 7.5. The first implication cannot be reversed. There is an example of a regular PVS V and a non-exceptional $U\in \Gamma _{\operatorname {spcl}}(V)$ such that all elements of $U\cap X (F)$ are isotropic (see Example 7.8 below).

Proof. For the first part, let $U\in \Gamma _{\operatorname {spcl}}(V)$ and $S=\operatorname {Stab}(U)$ . Assume that there exists $U'\in \Gamma _{\operatorname {reg}}(V)$ , $U'\subset U$ such that the linear span of $X^*(G)\cup \Delta ^S_0\cup \Psi _{U'}$ is a proper subspace of $\mathfrak {a}_0^*$ . Equivalently, there exists a (split) torus $1\ne T\subset T_0 \cap Z(M_S) \cap G^{\operatorname {der}}$ fixing $U'$ pointwise (namely, the identity component of $\cap _{\chi \in X^*(G)\cup \Delta _0^S\cup \Psi _{U'}}\operatorname {Ker}\chi $ ). We will show that all elements of $U\cap X(F)$ are isotropic.

Fix a regular element u in $U'$ . By Proposition 6.4, we have $G_u^\circ \subset S$ . Since T lies in the center of $M_S$ , $[T,G_u\cap S]\subset [Z(M_S),S]\subset N_S$ . On the other hand, $T\subset G_u^\circ $ (since T stabilizes all elements of $U'$ ). Therefore, T centralizes $(G_u\cap S)/(G_u^\circ \cap N_S)$ . In particular, the image of T in $G_u^\circ /(G_u^\circ \cap N_S)$ is central. Note that $G_u^\circ \cap N_S$ is contained in the unipotent radical of $G_u^\circ $ .

Now, let $v\in U\cap X(F)$ . By Proposition 6.4, v is in the S-orbit (but not necessarily the $S(F)$ -orbit) of u. Let $x\in S (\bar {F})$ be such that $u=\rho (x)v$ . Then, $x^{-1}Tx$ is a torus in $G^{\operatorname {der}}\cap G_v^\circ $ . Moreover, for any $\sigma \in \operatorname {Gal}(\bar {F}/F)$ we have $\sigma (x)x^{-1}\in G_u\cap S$ . Let $T'$ be the image of $x^{-1}Tx$ in $G_v^\circ /(G_v^\circ \cap N_S)$ . Then, $T'$ is central. Moreover, $T'$ is a split torus. In fact, since T centralizes $\sigma (x)x^{-1}(G_u^\circ \cap N_S)$ for all $\sigma \in \operatorname {Gal}(\bar {F}/F)$ , the map $t\mapsto x^{-1}tx (G_v^\circ \cap N_S)$ is an F-isomorphism of T with $T'$ . In particular, the image of $T'$ in $G_v^{\operatorname {red}}$ is a non-trivial F-split central torus. Thus, v is isotropic by Lemma 2.6. This proves the first part.

For the second part, suppose that V is regular and $v\in X (F)$ is isotropic. Let $T\ne 1$ be a split torus contained in $Z(G^\circ _v)\cap G^{\operatorname {der}}$ . The centralizer $L=C_G(T)$ contains $G_v^\circ $ . Upon replacing v by an element in its $G(F)$ -orbit, we can assume that T is contained in $T_0 \cap G^{\operatorname {der}}$ and that L is a standard Levi subgroup of G. Thus, $X_*(T)\subset \mathfrak {a}_L$ and in the notation of [Reference Mœglin and Waldspurger26, I.1.10] we have $X_*(T)\cap \mathfrak {a}_L^{\operatorname {reg}}\ne \emptyset $ . Fix an element $h \in X_*(T)\cap \mathfrak {a}_L^{\operatorname {reg}} \subset \mathfrak {a}^G_0$ . Upon conjugating by an element of the Weyl group of $T_0$ , we may assume that $\left \langle {\alpha },{h}\right \rangle $ is non-negative precisely for the roots $\alpha $ of the (proper) standard parabolic subgroup Q of G with Levi subgroup L [Reference Mœglin and Waldspurger26, Lemma I.1.10].

Consider the subspaces $U,U'\in \Gamma (V)$ given by

$$\begin{align*}\Psi_U = \{ \beta \in \Psi_V \mid\left\langle{\beta},{h}\right\rangle \ge 0 \},\ \ \Psi_{U'}= \{ \beta \in \Psi_V \mid\left\langle{\beta},{h}\right\rangle = 0 \} \subset\Psi_U. \end{align*}$$

Clearly U is stable under Q and $U'\subset U$ . Since T stabilizes v, $\operatorname {supp} v\subset \Psi _{U'}$ so that $v \in U'$ . In particular, $U'\in \Gamma _{\operatorname {reg}} (V)$ , and a fortiori $U \in \Gamma _{\operatorname {reg}} (V)$ . Recall the surjection (2.1). It satisfies

by (5.3). In particular, and where and is the Lie algebra of the unipotent radical of the parabolic subgroup of G opposite to Q. We claim that the preimage of U under $D_v$ is . It suffices to show that . Suppose that and $D_v(x)\in U$ . Then, by the above, $D_v(x)=0$ . Hence, x is in the Lie algebra of $G_v$ , which is contained in , and thus $x=0$ .

Hence, the induced map is an isomorphism. It follows from Proposition 6.4 that $U\in \Gamma _{\operatorname {spcl}}(V)$ and $\operatorname {Stab}(U)=Q$ .

Since the nonzero vector $h \in \mathfrak {a}^G_0$ annihilates $\Psi _{U'}\cup \Delta ^Q_0$ , we obtain that $(U,U')$ is exceptional.

Finally, the last part follows from the first two parts.

Remark 7.6. We do not know whether the second (and hence also the last) part of Proposition 7.4 hold without the regularity assumption on V.

Example 7.7. Proposition 7.4 is best illustrated by the case where V is the PVS of binary quadratic forms with respect to the action of $G=\operatorname {GL}_2$ , i.e., the symmetric square representation. (Example 3.4 with $k=1$ , $n_1=2$ .) Thus, X is the set of binary quadratic forms with nonzero discriminant. In this case, $\Psi _V = \{ \beta , \beta + \alpha , \beta + 2 \alpha \}$ , where $\alpha $ is the simple root of G, and $\beta $ the lowest weight of V. Note that $\beta + \alpha $ is the determinant character of G. Apart from V itself, there is a single space U in $\Gamma _{\operatorname {spcl}}(V)$ , namely $U=V_{\beta +\alpha }+V_{\beta +2\alpha }$ , and $\mathbf {E}(U)=\operatorname {Stab}(U)$ is the standard Borel subgroup of G. The space U is exceptional. More precisely, $(U,U')$ is an exceptional pair where $U'=V_{\beta +\alpha }$ . Let $v\in X$ . Then, $G_v$ is an orthogonal group so that $G_v^\circ $ is a one-dimensional torus. Thus, Proposition 7.4 amounts to the fact that $G_v^\circ $ is split if and only if v can be represented as $xy$ in suitable coordinates.

Example 7.8. Let us revisit Example 3.9 pertaining to the simply connected group $G'=E_6$ .

For brevity we will write for the linear combinations of simple roots with coefficients $a,b,c,d,e,f$ .

We will also write standard parabolic subgroups as where $*$ indicates that the simple root is in the Levi subgroup and $0$ otherwise.

Let $\theta $ be the outer involution of $G'$ acting on the simple roots as indicated.

Let P be the maximal standard parabolic subgroup . The Levi subgroup G has derived group $\operatorname {SL}_2 \times \operatorname {SL}_3 \times \operatorname {SL}_3$ .

The unipotent radical $N_{P}$ , which has dimension $29$ , has a three-step filtration, the first filtration step V (the abelianization) has dimension $18$ . We may write $\Psi _V$ explicitly to reflect the realization of V as the space of pairs of $3\times 3$ -matrices as follows.

We observe that for all proper subspaces $U\in \Gamma _{\operatorname {spcl}}(V)$ , $U \cap X (F)$ consists entirely of isotropic elements. Namely, if $U\in \Gamma _{\operatorname {spcl}}(V)$ contains a non-isotropic element x, then $G^\circ _x$ has to be contained in $Q = \operatorname {Stab} (U)$ . But the projection of $G^\circ _x$ to any of the two $\operatorname {SL}_3$ factors of $G^{\operatorname {der}}$ is an anisotropic maximal torus of this group, which implies that Q contains both $\operatorname {SL}_3$ factors. Therefore, either $Q=G$ and $U=V$ or the intersection of Q with the $\operatorname {SL}_2$ factor is the standard Borel subgroup. However, for this maximal parabolic subgroup Q of G the only non-trivial Q-invariant subspace of V does not intersect X.

Define $U_1,U_2,U_3\in \Gamma (V)$ by

In terms of matrices, these spaces are described as follows.

$$\begin{align*}U_1=\left(\begin{smallmatrix}***\\ 0**\\ 0**\end{smallmatrix}\right),\left(\begin{smallmatrix}***\\ 0**\\ 0**\end{smallmatrix}\right),\ \ U_2=\left(\begin{smallmatrix}0**\\ 0**\\ 0**\end{smallmatrix}\right),\left(\begin{smallmatrix}***\\ ***\\ ***\end{smallmatrix}\right),\ \ U_3=\left(\begin{smallmatrix}***\\ 00*\\ 00*\end{smallmatrix}\right),\left(\begin{smallmatrix}***\\ 0**\\ 0**\end{smallmatrix}\right). \end{align*}$$

Let $U_i'=\theta (U_i)$ , $i=1,2,3$ and $U_i"=U_i\cap U_i'$ so that

$$\begin{align*}U_1'=\left(\begin{smallmatrix}***\\ ***\\ 00*\end{smallmatrix}\right),\left(\begin{smallmatrix}***\\ ***\\ 00*\end{smallmatrix}\right),\ \ U_2'=\left(\begin{smallmatrix}***\\ ***\\ 000\end{smallmatrix}\right),\left(\begin{smallmatrix}***\\ ***\\ ***\end{smallmatrix}\right),\ \ U_3'=\left(\begin{smallmatrix}***\\ 00*\\ 00*\end{smallmatrix}\right),\left(\begin{smallmatrix}***\\ ***\\ 00*\end{smallmatrix}\right), \end{align*}$$

and

$$\begin{align*}U_1"=\left(\begin{smallmatrix}***\\ 0**\\ 00*\end{smallmatrix}\right),\left(\begin{smallmatrix}***\\ 0**\\ 00*\end{smallmatrix}\right),\ \ U_2"=\left(\begin{smallmatrix}0**\\ 0**\\ 000\end{smallmatrix}\right),\left(\begin{smallmatrix}***\\ ***\\ ***\end{smallmatrix}\right),\ \ U_3"=\left(\begin{smallmatrix}***\\ 00*\\ 00*\end{smallmatrix}\right),\left(\begin{smallmatrix}***\\ 0**\\ 00*\end{smallmatrix}\right). \end{align*}$$

We write $U_{i,j}=U_i\cap U_j$ , $U_{i',j}=U_i'\cap U_j$ , $U_{i',j'}=U_i'\cap U_j'$ , $U_{i",j}=U_i"\cap U_j$ and $U_{i",j'}=U_i"\cap U_j'$ . Then, a straightforward but tedious computation gives the following Hasse diagram for $\Gamma _{\operatorname {spcl}}(V)$ :

Moreover,

The stabilizers of the rest of the elements of $\Gamma _{\operatorname {spcl}}(V)$ are given by Lemma 6.11. All spaces except $U_2$ , $U^{\prime }_2$ and V are exceptional.

The spaces $U \in \Gamma _{\operatorname {spcl}}(V)$ with $\mathbf {E}(U) = \operatorname {Stab}(U)$ are $V, U_1, U_1', U_2, U_2', U_{1,2}, U_{1',2'}, U_1", U_2"$ . For $U = U_3, U_{3,2'}, U_{1,2'}$ we have $\mathbf {E}(U) \star U = U_1$ . For $U = U_3', U_{3',2}, U_{1',2}$ we have $\mathbf {E}(U) \star U = U_1'$ . For $U = U_3", U_{1",2}, U_{1",2'}$ we have $\mathbf {E}(U) \star U = U_1"$ .

7.3

We can now formulate and prove the main geometric statement about $\lambda (U)$ where $U\in \Gamma _{\operatorname {niso}} (V)$ .

Consider the natural surjection

$$\begin{align*}q:\mathfrak{a}_G=X_*(G^{\operatorname{ab}})\otimes\mathbb{R}\longrightarrow\overline{\mathfrak{a}}_G=X_*(G/{G}^\star)\otimes\mathbb{R}. \end{align*}$$

The kernel of q is the image of $X_*({G}^\star /G^{\operatorname {der}})\otimes \mathbb {R}$ in $\mathfrak {a}_G$ . The dual space $\overline {\mathfrak {a}}_G^*=X^*(G/{G}^\star )\otimes \mathbb {R}\hookrightarrow \mathfrak {a}_G^*$ of $\overline {\mathfrak {a}}_G$ has basis $\mathfrak {X}$ . Fix a section for q, i.e., a subspace $\mathfrak {b}$ of $\mathfrak {a}_G$ such that the restriction of q to $\mathfrak {b}$ is an isomorphism. In other words, $\mathfrak {b}$ is a complement of $\operatorname {Ker} q$ in $\mathfrak {a}_G$ . Let

$$\begin{align*}W=(\mathfrak{a}_0^G)\oplus\mathfrak{b}\hookrightarrow(\mathfrak{a}_0^G)\oplus\mathfrak{a}_G=\mathfrak{a}_0. \end{align*}$$

Recall the cone $\mathcal {C}_U$ defined in (6.1).

Proposition 7.9. Suppose that V is regular. Let $U \in \Gamma _{\operatorname {niso}} (V)$ . Then, for every $\mu \in \mathbb {R}_{>0}\mathfrak {X}$ , the restriction of the linear form $\lambda (U)+\mu $ to W is positive on the cone $\mathcal {C}_U\cap W\setminus \{0\}$ .

Proof. Suppose that $x\in \mathcal {C}_U$ is such that $\left \langle {\lambda (U)+\mu },{x}\right \rangle \le 0$ . By Corollary 5.3 part 1, we have $\left \langle {\chi },{x}\right \rangle \ge 0$ for all $\chi \in \mathfrak {X}$ . By (5.7), we obtain $\left \langle {\lambda (U)},{x}\right \rangle =0$ and $\left \langle {\chi },{x}\right \rangle =0$ for all $\chi \in \mathfrak {X}$ . Also, by the definition of $\mathbf {E}=\mathbf {E}(U)$ , we have $\left \langle {\alpha },{x}\right \rangle =0$ for all $\alpha \in \Delta _0^{\mathbf {E}}$ .

Let $U'\in \Gamma _{\operatorname {reg}}(V)$ be as in Corollary 5.3 part 2. Then, $\left \langle {\beta },{x}\right \rangle =0$ for all $\beta \in \Psi _{U'}$ .

Let $\tilde U=\mathbf {E}(U)\star U$ . By Lemma 6.12, $\tilde U\in \Gamma _{\operatorname {spcl}}(V)$ and $\mathbf {E}(\tilde U)=\operatorname {Stab}(\tilde U)=\mathbf {E}$ .

Since $\tilde U\supset U$ , and $U\in \Gamma _{\operatorname {niso}}(V)$ we may apply Proposition 7.4 to conclude that $\tilde U$ is not exceptional. Thus, the set $\Psi _{U'}\cup \Delta _0^{\mathbf {E}}\cup \mathfrak {a}_G^*$ spans $\mathfrak {a}_0^*$ . Let

$$\begin{align*}p:\mathfrak{a}_0^*\rightarrow W^*, \end{align*}$$

be the restriction map, which is surjective. We have $\mathfrak {a}_G^*=\operatorname {Ker} p\oplus \overline {\mathfrak {a}}_G^*$ . Since $\mathfrak {X}$ spans $\overline {\mathfrak {a}}_G^*$ , we infer that $p(\Psi _{U'}\cup \Delta _0^{\mathbf {E}}\cup \mathfrak {X})$ spans $W^*$ . Restricting to $x \in W$ , we deduce $x=0$ .

7.4

We can now finish the proof of Theorem 4.3. As we have already noted in the discussion following (4.1), assuming convergence, the integral

$$\begin{align*}a\mapsto\int_{G(F) \backslash G (\mathbb{A})^1} \theta^{\operatorname{niso}} (ha) \, dh, \end{align*}$$

is ${G}^\star (\mathbb {A})$ -invariant. It therefore remains to prove the convergence of

$$\begin{align*}\int_{\mathfrak{b}}\int_{G(F) \backslash G (\mathbb{A})^1} \theta^{\operatorname{niso}} (ha) \, dh\ e^{-\left\langle{\delta_V+\mu},{a}\right\rangle}\ da, \end{align*}$$

where $\mathfrak {b}$ is as in Proposition 7.9 and $\mu \in \mathbb {R}_{>0}\mathfrak {X}$ .

Choose a Siegel set as in §5.1 such that $G(F)\mathfrak {S}=G(\mathbb {A})$ . For any non-negative function f on $G(F)\backslash G(\mathbb {A})^1$ we have

(7.1) $$ \begin{align} \begin{aligned} \int_{G(F)\backslash G(\mathbb{A})^1}f(g)\ dg&\le \int_K\int_{\mathfrak{a}_0^G}\int_{P_0(F)\backslash P_0(\mathbb{A})^1}f(pe^x k)\delta_{P_0}(e^x)^{-1}\tau_{c_0}(x)\ dp\ dx\ dk\\&\le \int_{\mathfrak{a}_0^G}{\mathrm{sup}}_{y\in\Omega} f(e^x y)\delta_{P_0}(e^x)^{-1}\tau_{c_0}(x)\ dx, \end{aligned} \end{align} $$

where $\tau _{c_0}$ is the characteristic function of $\{x\in \mathfrak {a}_0^G\mid \left \langle {\alpha },{x}\right \rangle>c_0\ \forall \alpha \in \Delta _0\}$ .

It is therefore enough to show the convergence of

(7.2) $$ \begin{align} \int_W\delta_{P_0}(e^x)^{-1}{\mathrm{sup}}_{y\in\Omega}\left|{\theta_\phi^{\operatorname{niso}} (e^x y)}\right| \ e^{-\left\langle{\delta_V+\mu},{x}\right\rangle}\tau_{c_0}(x)\ dx, \end{align} $$

where $W=\mathfrak {a}_0^G\oplus \mathfrak {b}\hookrightarrow \mathfrak {a}_0$ and we extend $\tau _{c_0}$ trivially on $\mathfrak {b}$ .

As in the proof of Proposition 5.1 we bound $\theta ^{\operatorname {niso}} (g)$ by splitting the sum according to support sets, i.e., we write

$$\begin{align*}\theta^{\operatorname{niso}}_\phi(g)=\sum_{U\in\Gamma(V)}\Theta_U^{\operatorname{niso}}(g), \end{align*}$$

with

$$\begin{align*}\Theta_U^{\operatorname{niso}}(g)=\sum_{\xi\in X(F)^{\operatorname{niso}}: \,\operatorname{supp}\xi=\Psi_U}\phi(\rho(g)^{-1}\xi). \end{align*}$$

The integral (7.2) is then majorized by the sum over $U\in \Gamma (V)$ of

(7.3) $$ \begin{align} \int_W\delta_{P_0}(e^x)^{-1}{\mathrm{sup}}_{y\in\Omega}\left|{\Theta_U^{\operatorname{niso}} (e^x y)}\right| \ e^{-\left\langle{\delta_V+\mu},{x}\right\rangle}\tau_{c_0}(x)\ dx. \end{align} $$

Moreover, by Lemma 5.8 for any $N\ge 0$ we have

$$\begin{align*}\Theta_U^{\operatorname{niso}}(g)\ll^{\operatorname{rd}}_{N,\phi} \delta_{P_0}(m_0(g))\left|{\det\rho(g)}\right|e^{-\left\langle{\lambda(U)},{H_0(g)}\right\rangle - N \sum_{\beta\in\Psi_U} \left\langle{\beta},{H_0(g)}\right\rangle_-}, \ \ g\in T_0(\mathbb{A})\Omega. \end{align*}$$

Therefore, (7.3) is majorized by

(7.4) $$ \begin{align} \nu_N (\phi) \int_W e^{-\left\langle{\lambda (U)+\mu},{x}\right\rangle}e^{-N \sum_{\beta\in\Psi_U} \left\langle{\beta},{x}\right\rangle_-}\tau_{c_0}(x)\ dx, \end{align} $$

for suitable seminorms $\nu _N$ on $C_{\operatorname {rd}}(V(\mathbb {A}))$ . We may assume of course that $U \in \Gamma _{\operatorname {niso}} (V)$ , for otherwise $\Theta _U^{\operatorname {niso}}\equiv 0$ . By Proposition 7.9, the restriction of the linear form $\lambda (U)+\mu $ to W is then positive on the cone $\mathcal {C}_U\cap W\setminus \{0\}$ . Therefore, the integral (7.4) converges by Lemma 7.1.

This finishes the proof of Theorem 4.3.

8 PVSs without non-trivial special subspaces

In the section, we analyze the case where a regular PVS V has regular F-irreducible components and no non-trivial special subspaces. Under these conditions, we obtain a strengthening of our main result, namely meromorphic continuation of the zeta function to the left of the point $\lambda = 0$ with at most simple poles along the hyperplanes defined by the fundamental characters.

We continue to assume that G is reductive. In general, a subrepresentation of a regular PVS is not necessarily regular. We will say that a PVS V is completely factorizable (CF) if all its F-irreducible subrepresentations are regular. It is clear that every subrepresentation of a CF PVS is again CF.

For example, every basic PVS (see Definition 2.8) is CF, since any subrepresentation is basic, hence regular. Examples 3.23.6 are CF precisely when $n_1=\dots =n_k$ . (However, $n_{k+1}$ may be bigger than $n_k$ in cases 3.3 and 3.5.)

Note that in general, an F-irreducible representation $\pi $ of G does not admit a central character. However, the restriction of $\pi $ to the torus $T_G$ acts by a character, which we denote by $\omega _\pi $ .

Lemma 8.1 (cf. [Reference Rubenthaler35], Proposition 2.13)

Let $(G,V,\rho )$ be a CF PVS and let $V=\oplus _{i\in I}(V_i,\rho _i)$ be a decomposition into irreducible subrepresentations over F. Let $\chi _i=\omega _{\rho _i}$ , $i\in I$ . Then,

  1. 1. V is regular.

  2. 2. $X=\prod _{i\in I}X_i$ , where $X_i$ is the regular part of $V_i$ .

  3. 3. The fundamental invariant polynomials of V are those of the $V_i$ ’s.

  4. 4. The restrictions of the fundamental characters of V to $T_G$ are $n_i\chi _i$ for some positive integers $n_i$ , $i\in I$ .

  5. 5. The characters $\chi _i$ are linearly independent over $\mathbb {Z}$ .

  6. 6. The union $\Psi _V=\cup \Psi _{V_i}$ is disjoint.

  7. 7. The irreducible components $V_i$ are the isotypic components of V.

  8. 8. The set $\Gamma _{\operatorname {reg}}(V)$ consists of the direct sums $U=\oplus _{i\in I} U_i$ with $U_i \in \Gamma _{\operatorname {reg}}(V_i)$ , $i \in I$ .

  9. 9. The group ${G}^\star /G^{\operatorname {der}}\operatorname {Ker}\rho $ is F-anisotropic.

Proof. First note that every F-irreducible regular PVS admits a unique fundamental invariant polynomial (over F).

Parts 1, 2, and 3 are contained in [Reference Rubenthaler35, Proposition 2.13]. (Note that part 3 over $\bar F$ easily implies it over F.) Part 4 immediately follows. Part 5 follows since the fundamental characters of a PVS are always linearly independent. Parts 6, 7 and 8 are easy consequences of part 5.

Finally, part 9 is equivalent to the statement that $T_G\cap \operatorname {Ker}\rho $ is of finite index in $T_G\cap {G}^\star $ . Note that $T_G\cap \operatorname {Ker}\rho $ is the intersection of $\operatorname {Ker}\chi _i$ , $i\in I$ and by part 4, this is of finite index in the common kernel of the restrictions of the characters in $X^*(G/{G}^\star )$ to $T_G$ . The lemma follows.

The following easy consequence about the special subspaces of V and its subrepresentations will be used repeatedly.

Lemma 8.2. Let V be a CF PVS and $V'$ a subrepresentation of V (defined over F).

  1. 1. Let $V"$ be the (unique) complement to $V'$ in V. Then the map $U \mapsto U \oplus V"$ defines an injection of $\Gamma _{\operatorname {spcl}} (V')$ into $\Gamma _{\operatorname {spcl}} (V)$ .

  2. 2. If $\Gamma _{\operatorname {spcl}}(V)=\{V\}$ , then $\Gamma _{\operatorname {spcl}}(V')=\{V'\}$ .

Proof. By Lemma 8.1 part 7, the map $U \mapsto U \oplus V"$ defines an injection of $\Gamma _{\operatorname {reg}} (V')$ into $\Gamma _{\operatorname {reg}} (V)$ . The first part follows now from the definition of a special subspace. The second part is clear.

Let $\mathcal {S}(V(\mathbb {A}))$ be the space of Schwartz–Bruhat functions on $V(\mathbb {A})$ . As a locally convex topological vector space, $\mathcal {S}(V(\mathbb {A}))$ can be identified with $\mathcal {S}(V(F_\infty ))\otimes \mathcal {S}(V(\mathbb {A}_{\operatorname {fin}}))$ where $\mathcal {S}(V(F_\infty ))$ is the usual Schwartz space and $\mathcal {S}(V(\mathbb {A}_{\operatorname {fin}}))$ is the countable-dimensional space of locally constant compactly supported functions on $V(\mathbb {A}_{\operatorname {fin}})$ with the finite topology. In other words, $\mathcal {S}(V(\mathbb {A}_{\operatorname {fin}}))=\varinjlim U$ with the inductive limit topology, where U ranges over the finite-dimensional subspaces of $\mathcal {S}(V(\mathbb {A}_{\operatorname {fin}}))$ with their canonical Hausdorff topology. Equivalently, $\mathcal {S}(V(\mathbb {A}))=\varinjlim \mathcal {S}(V(F_\infty ))\otimes U$ with the locally convex inductive limit topology where U is as before. The space $\mathcal {S}(V(\mathbb {A}))$ is Hausdorff, complete and nuclear. The embedding $\mathcal {S}(V(\mathbb {A}))\rightarrow C_{\operatorname {rd}}(V(\mathbb {A}))$ is continuous.

Assume that $(G,V,\rho )$ is CF with irreducible components $(\rho _i,V_i)$ , $i\in I$ over F and let $\chi _i=\omega _{\rho _i}$ , $i\in I$ . For any $J\subset I$ let $V^J=\oplus _{j\in J}V_j$ and $V_J=V/V^J=\oplus _{j\notin J}\rho _j$ . Let $\rho _J$ be the representation of G on $V_J$ . These are again CF PVSs.

As usual, let $A_G = T_G (\mathbb {R})^\circ $ be the complement to $G(\mathbb {A})^1$ in $G (\mathbb {A})$ obtained by embedding $\mathbb {R}$ into $F_\infty = \mathbb {R} \otimes F$ via $x \mapsto x \otimes 1$ , and let $A_{\operatorname {Ker} \rho } = A_G \cap \operatorname {Ker} \rho $ . Since ${G}^\star /G^{\operatorname {der}}\operatorname {Ker}\rho $ is F-anisotropic, we have $G(\mathbb {A})^1{G}^\star (\mathbb {A})\backslash G(\mathbb {A}) \simeq A_{\operatorname {Ker}\rho } \backslash A_G$ and we can write the zeta function of $\phi \in \mathcal {S}(V(\mathbb {A}))$ as

$$\begin{align*}Z(\phi,\lambda)=\int_{G(F) A_{\operatorname{Ker}\rho} \backslash G (\mathbb{A})} \theta_\phi^{\operatorname{niso}} (g) g^{-\lambda}\left|{\det\rho(g)}\right|{}^{-1} \, dg, \end{align*}$$

for $\lambda \in (\mathfrak {a}_{G,\mathbb {C}}/\mathfrak {a}_{\operatorname {Ker} \rho ,\mathbb {C}})^* = \mathbb {C} \mathfrak {X}$ . We know from Theorem 4.3 that the integral converges absolutely for $\Re \lambda \in \mathbb {R}_{>0}\mathfrak {X}$ . In the following, we will use the isomorphism $A_{\operatorname {Ker}\rho } \backslash A_G \simeq \mathbb {R}^I, a \mapsto (\log \left |{\chi _i (a)}\right |)_{i \in I}$ and the Lebesgue measure on the vector space $\mathbb {R}^I$ to normalize the measure on the quotient $G(F) A_{\operatorname {Ker}\rho } \backslash G (\mathbb {A})$ .

For any $\phi \in \mathcal {S}(V(\mathbb {A}))$ , let $\phi _J\in \mathcal {S}(V_J(\mathbb {A}))$ be given by

$$\begin{align*}\phi_J(v)=\int_{V^J(\mathbb{A})}\phi(u+v)\ du. \end{align*}$$

For $\lambda =\sum _{i\in I}\lambda _i\chi _i$ we write $\lambda _J=\sum _{j\notin J}\lambda _j\chi _j$ . We will consider the zeta functions $Z^{V_J}(\phi _J,\lambda _J)$ simultaneously. In the limiting case $J = I$ , where $V_I = 0$ , the term $Z^{V_I}(\phi _I,\lambda _I)$ is of course interpreted as the constant $\operatorname {vol} (G(F) \backslash G (\mathbb {A})^1) \hat \phi (0)$ .

The main result of this section is the following.

Theorem 8.3. Let $(G,V,\rho )$ be a CF PVS such that $\Gamma _{\operatorname {spcl}}(V)=\{V\}$ . Then, there exists $c>0$ such that for any $\phi \in \mathcal {S}(V(\mathbb {A}))$ the zeta function $Z(\phi ,\lambda )$ admits a meromorphic continuation to the region

$$\begin{align*}\mathcal{D}=\{\lambda=\sum_{i\in I}\lambda_i\chi_i\mid\Re\lambda_i>-c\}, \end{align*}$$

and $Z (\phi , \lambda ) \prod _{i\in I}\lambda _i$ is holomorphic in $\mathcal {D}$ . Moreover, with the notation above,

(8.1) $$ \begin{align} \sum_{J\subset I}(-1)^{\# J}\frac{Z^{V_J}(\phi_J,\lambda_J)}{\prod_{j\in J}\lambda_j}, \end{align} $$

is holomorphic in $\mathcal {D}$ .

Theorem 8.3 will be proved in the remainder of this section. Applying it to $V_J$ for every $J\subset I$ and using inclusion–exclusion we obtain

Corollary 8.4. Under the above conditions, for every subset $J\subset I$ the function

$$\begin{align*}f_J (\lambda_J) = \sum_{K\subset I \setminus J}(-1)^{\# K}\frac{Z^{V_{J \cup K}}(\phi_{J \cup K},\lambda_{J \cup K})}{\prod_{j\in K}\lambda_j}, \end{align*}$$

is holomorphic in $\{\lambda _J=\sum _{j\notin J}\lambda _j\chi _j\mid \Re \lambda _j>-c\}$ and we have

$$\begin{align*}Z (\phi, \lambda) = \sum_{J\subset I}\frac{f_J (\lambda_J)}{\prod_{j\in J}\lambda_j}, \end{align*}$$

in $\mathcal {D}$ .

Theorem 8.3 and its corollary apply in particular to any basic PVS, as well as to the “magic square” examples of §3.3 – see Remark 6.6. Recall that the basic PVSs obtained from nilpotent orbits for the classical groups are the following:

  1. 1. $\operatorname {GL}_n^k$ acting on $\operatorname {Mat}_n^{k-1}$ (Example 3.2).

  2. 2. $\operatorname {GL}_n^k \times \operatorname {Sp}_{n+m}$ acting on $\operatorname {Mat}_n^{k-1} \oplus \operatorname {Mat}_{n,n+m}$ for n, m even (Example 3.3).

  3. 3. $\operatorname {GL}_n^k$ acting on $\operatorname {Mat}_n^{k-1} \oplus \operatorname {Sym}_n$ for $n \neq 2$ (Example 3.4).

  4. 4. $\operatorname {GL}_n^k \times \operatorname {SO}_{n+m}$ acting on $\operatorname {Mat}_n^{k-1} \oplus \operatorname {Mat}_{n,n+m}$ for n, $m \neq 2$ (Example 3.5).

  5. 5. $\operatorname {GL}_n^k$ acting on $\operatorname {Mat}_n^{k-1} \oplus \operatorname {Skew}_n$ (Example 3.6).

For F-irreducible V such that $\Gamma _{\operatorname {spcl}}(V)=\{V\}$ , we obtain that the zeta function has a simple pole at $\lambda = 0$ with residue $\operatorname {vol} (G(F) \backslash G (\mathbb {A})^1) \hat \phi (0)$ , as in Tate’s thesis (cf. Proposition 8.5 below). Example 3.7 (treated in [Reference Wright50], building on [Reference Shintani44]) and Example 3.8 (treated in [Reference Yukie53]) show that without the assumption $\Gamma _{\operatorname {spcl}}(V)=\{V\}$ , the structure of the pole of the zeta function at $\lambda = 0$ may be more complicated.

Note that the assumption $\Gamma _{\operatorname {spcl}}(V)=\{V\}$ implies (by Lemma 8.2 and Proposition 7.4) that all $V_J$ are non-isotropic, i.e. that $\theta _{\phi _J}^{\operatorname {niso}} (g) = \theta _{\phi _J} (g)$ . (In our proof below we will simply work with $\theta _{\phi _J} (g)$ throughout.)

We first consider the case where $\rho $ is irreducible over F. For simplicity we write $x_{\le 1}=\min (x,1)$ and $x_{\ge 1}=\max (1,x)$ for any $x\in \mathbb {R}_{>0}$ . Denote by $\mathbf {1}_A$ the characteristic function of a set A.

For brevity, we will use the notation $A(\cdot ,\phi )\ll _{N,\phi } B_N(\cdot )$ to mean that for every $N\ge 0$ there exists a continuous seminorm $\nu _N$ on $\mathcal {S}(V(\mathbb {A}))$ such that $\left |{A(\cdot ,\phi )}\right |\le B_N(\cdot )\nu _N(\phi )$ for every $\phi \in \mathcal {S}(V(\mathbb {A}))$ . Recall that we have already used the notation $\ll _{N,\phi }^{\operatorname {rd}}$ in §5.5 with respect to the space $C_{\operatorname {rd}}(V(\mathbb {A}))$ . Clearly $\ll _{N,\phi }^{\operatorname {rd}}$ implies $\ll _{N,\phi }$ .

Proposition 8.5. Suppose that $(G,V,\rho )$ is an F-irreducible PVS such that $\Gamma _{\operatorname {spcl}}(V)=\{V\}$ and let $\chi =\omega _\rho $ . For any $\phi \in \mathcal {S}(V(\mathbb {A}))$ let

$$\begin{align*}\gimel_\phi(g)=\left|{\det\rho(g)}\right|{}^{-1}\theta_\phi(g)-\mathbf{1}_{\mathbb{R}_{>1}}(\left|{\chi(g)}\right|)\hat\phi(0),\ \ g\in G(F)\backslash G(\mathbb{A}). \end{align*}$$

Let $\beta _0$ be the lowest weight of V. Then,

(8.2) $$ \begin{align} \gimel_\phi(g)\ll_{N,\phi}\delta_{P_0}(m_0(g))e^{-\left\langle{\mu},{H_0(g)}\right\rangle-\left\langle{\beta_0},{H_0(g)}\right\rangle_+-N\left\langle{\chi},{H_0(g)}\right\rangle_-},\ \ g\in\mathfrak{S,} \end{align} $$

for a suitable $\mu \in \mathbb {R}_{>0}(\Delta _0)$ . Therefore, there exists $c>0$ such that the integral

(8.3) $$ \begin{align} \int_{G(F) A_{\operatorname{Ker} \rho} \backslash G(\mathbb{A})}\gimel_\phi(g)\left|{\chi(g)}\right|{}^{-s} \, dg = Z (\phi, s \chi) - \operatorname{vol} (G(F) \backslash G (\mathbb{A})^1) \frac{\hat\phi(0)}{s}, \end{align} $$

converges absolutely for $\Re s>-c$ .

Proof. We will estimate $\gimel _\phi $ using the decomposition (5.8) of $\theta _\phi (g)$ . We write

$$\begin{align*}\gimel_\phi (g) = \left|{\det\rho(g)}\right|{}^{-1} \sum_{U\in\Gamma_{\operatorname{reg}}(V), \, U \neq V} \Theta_U(g) + \left|{\det\rho(g)}\right|{}^{-1}\Theta_V (g)-\mathbf{1}_{\mathbb{R}_{>1}}(\left|{\chi(g)}\right|)\hat\phi(0), \end{align*}$$

and separately bound the summands $\Theta _U$ , $U\in \Gamma _{\operatorname {reg}}(V)$ , $U\ne V$ and $\left |{\det \rho }\right |{}^{-1}\Theta _V-\mathbf {1}_{\mathbb {R}_{>1}}(\left |{\chi }\right |)\hat \phi (0)$ on $\mathfrak {S}$ .

Consider the contribution from $U\in \Gamma _{\operatorname {reg}}(V)$ , $U\ne V$ . By Lemma 5.8 we have

$$\begin{align*}\left|{\det\rho(g)}\right|{}^{-1}\Theta_U(g)\ll_{N,\phi}^{\operatorname{rd}}\delta_{P_0}(m_0(g))e^{-\left\langle{\lambda(U)},{H_0(g)}\right\rangle -N\sum_{\beta\in\Psi_U}\left\langle{\beta},{H_0(g)}\right\rangle_-},\ \ g\in\mathfrak{S}. \end{align*}$$

Recall that $\lambda (U)=\delta _0+\sum _{\beta \in \Psi _V\setminus \Psi _U}n_\beta \beta \in \mathbb {R}_{\ge 0}(\Psi _U\cup \Delta _0)$ by (5.6) and (5.7). By Corollary 6.13 part 3 and our assumption on V we have $\mathbf {E}(U)=G$ . Thus, the positive envelope of $\lambda (U)$ with respect to $\Delta _0\cup \Psi _U$ (see §6.1) contains $\Delta _0$ and we have

$$\begin{align*}\lambda(U)=\sum_{\beta\in\Psi_U}c_\beta\beta+\sum_{\alpha\in\Delta_0}d_\alpha\alpha, \end{align*}$$

with suitable coefficients $c_\beta \ge 0$ and $d_\alpha> 0$ . Moreover, by considering $\left \langle {\lambda (U)},{H}\right \rangle $ for $H \in \mathfrak {a}_G$ we see that $\sum _{\beta \in \Psi _U}c_\beta = \sum _{\beta \in \Psi _V\setminus \Psi _U}n_\beta \ge 1$ since $U\ne V$ . Using that on the Siegel domain $\left \langle {\beta _0},{H_0(g)}\right \rangle -\left \langle {\beta },{H_0(g)}\right \rangle $ is bounded from above for all $\beta \in \Phi _{V}$ , we obtain

$$\begin{align*}\left|{\det\rho(g)}\right|{}^{-1}\Theta_U(g)\ll_{N,\phi}^{\operatorname{rd}}\delta_{P_0}(m_0(g))e^{-\left\langle{\mu},{H_0(g)}\right\rangle - \left\langle{\beta_0},{H_0(g)}\right\rangle_+ - N\sum_{\beta\in\Psi_U}\left\langle{\beta},{H_0(g)}\right\rangle_-},\ \ g\in\mathfrak{S}. \end{align*}$$

By Corollary 5.3 part 1, $\chi \neq 0$ is a non-negative linear combination of $\Psi _U$ . Therefore, we can replace here $\sum _{\beta \in \Psi _U}\left \langle {\beta },{H_0(g)}\right \rangle _-$ by a suitable positive multiple of $\left \langle {\chi },{H_0(g)}\right \rangle _-$ , proving the required bound for the contribution of any $U \neq V$ .

As for $U=V$ , we claim that

(8.4) $$ \begin{align} \left|{\det\rho(g)}\right|{}^{-1}\Theta_V(g)-\mathbf{1}_{\mathbb{R}_{>1}}(\left|{\chi(g)}\right|)\hat\phi(0)\ll_{N,\phi}e^{-\left\langle{\beta_0},{H_0(g)}\right\rangle_+-N\left\langle{\chi},{H_0(g)}\right\rangle_-},\ \ g\in\mathfrak{S}. \end{align} $$

We may assume that $g\in T_0(\mathbb {A})\cap \mathfrak {S}$ , or even that $g\in T_0(\mathbb {R})\cap \mathfrak {S}$ . We split into cases according to the size of $\left \langle {\beta _0},{H_0(g)}\right \rangle $ . Fix a constant $c_1$ .

If $\left \langle {\beta _0},{H_0(g)}\right \rangle \le c_1$ , then (8.4) holds because $\left |{\det \rho (g)}\right |{}^{-1}\Theta _V(g)\ll _{N,\phi }^{\operatorname {rd}}e^{-N\left \langle {\chi },{H_0(g)}\right \rangle _-}$ by Lemma 5.8.

Assume that $\left \langle {\beta _0},{H_0(g)}\right \rangle>c_1$ . Then $\left \langle {\beta },{H_0(g)}\right \rangle $ is bounded away from $0$ for every $\beta \in \Psi _V$ , since $g\in \mathfrak {S}$ . By a standard argument, e.g., using the Poisson summation formula, we have

(8.5) $$ \begin{align} \left|{\det\rho(g)}\right|{}^{-1}\sum_{v\in V(F)}\phi(\rho(g)^{-1}v)-\hat\phi(0)\ll_\phi e^{-\left\langle{\beta_0},{H_0(g)}\right\rangle}. \end{align} $$

(Here we cannot replace $\ll _\phi $ by $\ll _\phi ^{\operatorname {rd}}$ .) On the other hand,

$$\begin{align*}\left|{\Theta_V(g)-\sum_{v\in V(F)}\phi(\rho(g)^{-1}v)}\right|\le\sum_{v\in V(F)\mid P(v)=0}\left|{\phi(\rho(g)^{-1}v)}\right|, \end{align*}$$

for some non-zero polynomial P on V. (For instance, we can take the relative invariant times the product over $\beta \in \Psi _V$ of arbitrary nonzero linear forms on $V_\beta $ .) It follows from [Reference Finis and Lapid11, Lemma 6.4] that

(8.6) $$ \begin{align} \left|{\det\rho(g)}\right|{}^{-1}(\Theta_V(g)-\sum_{v\in V(F)}\phi(\rho(g)^{-1}v))\ll_\phi e^{-\left\langle{\beta_0},{H_0(g)}\right\rangle}. \end{align} $$

(In fact, the argument gives the stronger assertion with $\ll _\phi ^{\operatorname {rd}}$ .)

Since $\left \langle {\chi },{H_0(g)}\right \rangle -\left \langle {\beta _0},{H_0(g)}\right \rangle $ is bounded from below for $g\in \mathfrak {S}$ , we may choose $c_1$ such that $\left |{\chi (g)}\right |>1$ whenever $\left \langle {\beta _0},{H_0(g)}\right \rangle>c_1$ . Thus, (8.4) for $\left \langle {\beta _0},{H_0(g)}\right \rangle>c_1$ follows from (8.5) and (8.6). This finishes the proof of (8.2).

The convergence of

$$\begin{align*}\int_{P_0 (F) A_{\operatorname{Ker} \rho} \backslash \mathfrak{S}}\left|{\gimel_\phi(g)}\right| \left|{\chi(g)}\right|{}^{-\sigma} \, dg, \end{align*}$$

for $\sigma>-c$ for a suitable $c>0$ follows from (8.2) and Lemma 7.1, since for any complement W of $\mathfrak {a}_{\operatorname {Ker} \rho }$ in $\mathfrak {a}_0$ the positively- $1$ -homogeneous function $\left \langle {\mu },{x}\right \rangle +\left \langle {\beta _0},{x}\right \rangle _+$ on W is positive on the cone $\{x\in \mathfrak {a}_{0,+} \cap W \mid \left \langle {\chi },{x}\right \rangle \ge 0\}\setminus \{0\}$ .

This implies the absolute convergence of (8.3) for $\Re s>-c$ .

Proof of Theorem 8.3

By induction on the cardinality of $I \setminus J$ , starting with the trivial case $J=I$ (and, if we like, Proposition 8.5 for $\# (I \setminus J) = 1$ ), we may assume holomorphic continuation of the functions $Z^{V_J}(\phi _J,\lambda _J) \prod _{j\notin J}\lambda _j$ to $\{\lambda _J=\sum _{j\notin J}\lambda _j\chi _j\mid \Re \lambda _j>-c\}$ for all $J\neq \emptyset $ . To prove the theorem, we need to show that (8.1) is holomorphic on $\mathcal {D}$ .

For every $i\in I$ let $\beta _i$ be the lowest weight of $V_i$ . For any $J\subset I$ let $\chi _J=\prod _{j\notin J}\chi _j$ and let $G^J$ be the subset of $G(\mathbb {A})$ defined by the conditions $\left |{\chi _j(g)}\right |>1$ for all $j\in J$ .

Consider

$$\begin{align*}\daleth_\phi(g)=\sum_{J\subset I}(-1)^{\# J}\left|{\det\rho_J(g)}\right|{}^{-1}\theta_{\phi_J}^{V_J}(g)\mathbf{1}_{G^J}(g). \end{align*}$$

We claim that there exists $\mu \in \mathbb {R}_{>0} \Delta _0$ such that

(8.7) $$ \begin{align} \daleth_\phi(g)\ll_{N,\phi} \delta_{P_0}(m_0(g))e^{-\left\langle{\mu},{H_0(g)}\right\rangle-\sum_{i\in I}\left\langle{\beta_i},{H_0(g)}\right\rangle_+-N\sum_{i\in I}\left\langle{\chi_i},{H_0(g)}\right\rangle_-},\ \ g\in\mathfrak{S}. \end{align} $$

We may identify $\mathcal {S}(V(\mathbb {A}))$ with the completed tensor product $\hat \otimes _{i\in I}\mathcal {S}(V_i(\mathbb {A}))$ with respect to, say, the projective tensor product topology (the spaces $\mathcal {S}(V_i(\mathbb {A}))$ are nuclear). Therefore, it is enough to prove (8.7) for $\phi $ of the form $\phi =\otimes \psi _i$ where $\psi _i\in \mathcal {S}(V_i(\mathbb {A}))$ . We then have

$$\begin{align*}\daleth_\phi(g)=\prod_{i\in I}\big(\left|{\det\rho_i(g)}\right|{}^{-1}\theta_{\psi_i}^{V_i}(g)-\mathbf{1}_{\mathbb{R}_{>1}}(\left|{\chi_i(g)}\right|)\hat\psi_i(0)\big). \end{align*}$$

If we simply apply (8.2) to each factor, the bound that we get will not be good enough. This is not surprising since we have not yet used that V itself is a PVS, only the individual factors. Instead, we write for each $i\in I$

$$\begin{align*}\theta_{\psi_i}^{V_i}=\Theta_{\psi_i}^{V_i,\operatorname{fs}}+\Theta_{\psi_i}^{V_i,\operatorname{nfs}}, \end{align*}$$

where

$$\begin{align*}\Theta_{\psi_i}^{V_i,\operatorname{fs}} = \sum_{\xi\in X_i (F):\,\operatorname{supp}\xi=\Psi_{V_i}}\phi(\rho_i(g)^{-1}\xi), \end{align*}$$

is the summand corresponding to $U=V_i$ (full support) in the decomposition (5.8) for $\theta _{\psi _i}^{V_i}$ . Then, we can decompose

(8.8) $$ \begin{align} \daleth_\phi=\sum_{J\subset I}\prod_{j\in J}\left|{\det\rho_j}\right|{}^{-1}\Theta_{\psi_j}^{V_j,\operatorname{nfs}}\prod_{j\notin J} \big(\left|{\det\rho_j}\right|{}^{-1}\Theta_{\psi_j}^{V_j,\operatorname{fs}}-\mathbf{1}_{\mathbb{R}_{>1}}(\left|{\chi_j}\right|)\hat\psi_j(0)\big). \end{align} $$

For any subset $J\subset I$ let $\phi ^J=\otimes _{j\in J}\psi _j$ . Then,

(8.9) $$ \begin{align} \prod_{j\in J}\Theta_{\psi_j}^{V_j,\operatorname{nfs}} = \sum_{U \in\Gamma_{\operatorname{reg}}(V^J), \, U \not\supset V_j \, \forall j\in J} \Theta^{V^J}_{\phi^J;U}, \end{align} $$

is the contribution to $\Theta ^{V^J}_{\phi ^J}$ with respect to the decomposition (5.8) of the subspaces U not containing any $V_j$ , $j \in J$ . As in the proof of Proposition 8.5, there exists $\mu \in \mathbb {R}_{>0}\Delta _0$ (which we may take to be independent of J) such that

(8.10) $$ \begin{align} \prod_{j\in J}\left|{\det\rho_j(g)}\right|{}^{-1}\Theta_{\psi_j}^{V_j,\operatorname{nfs}}(g)\ll_{N,\phi^J}\delta_{P_0}(m_0(g)) e^{-\left\langle{\mu},{H_0(g)}\right\rangle-\sum_{j\in J}\left\langle{\beta_j},{H_0(g)}\right\rangle_+-N\sum_{j\in J}\left\langle{\chi_j},{H_0(g)}\right\rangle_-}, \end{align} $$

on $\mathfrak {S}$ . Indeed, this estimate holds for each individual contribution $\Theta ^{V^J}_{\phi ^J;U}$ to (8.9). Since $\mathbf {E}(U)=G$ by Lemma 8.2 and Corollary 6.13 part 3, the positive envelope of $\lambda (U)$ with respect to $\Psi _U\cup \Delta _0$ contains $\Delta _0$ . Writing

$$\begin{align*}\lambda(U)=\sum_{\beta\in\Psi_U}c_\beta\beta+\sum_{\alpha\in\Delta_0}d_\alpha\alpha, \end{align*}$$

with $c_\beta \ge 0$ and $d_\alpha> 0$ , and considering $\left \langle {\lambda (U)},{H}\right \rangle $ for $H \in \mathfrak {a}_G$ , we obtain

$$\begin{align*}\sum_{j \in J} \sum_{\beta\in\Psi_U \cap \Psi_{V_j}} c_\beta \chi_j = \sum_{j \in J} \sum_{\beta\in\Psi_{V_j} \setminus\Psi_U}n_\beta \chi_j, \end{align*}$$

and therefore $\sum _{\beta \in \Psi _U \cap \Psi _{V_j}} c_\beta = \sum _{\beta \in \Psi _{V_j} \setminus \Psi _U}n_\beta \ge 1$ for all $j \in J$ , since the $\chi _j$ are linearly independent over $\mathbb {Z}$ and $U\not \supset V_j$ for all $j\in J$ .

Using Lemma 5.8 together with Corollary 5.3 part 1 and the fact that $\left \langle {\beta _j},{H_0(g)}\right \rangle -\left \langle {\beta },{H_0(g)}\right \rangle $ is bounded from above for any $\beta \in \Phi _{V_j}$ and $g\in \mathfrak {S}$ , we conclude (8.10) as in the irreducible case.

Combining (8.10) with the estimate (8.4) for every $j\notin J$ and using the decomposition (8.8) we obtain the required estimate (8.7).

Note that the expression (8.1) is equal to the integral

$$\begin{align*}\int_{A_{\operatorname{Ker} \rho} G(F)\backslash G(\mathbb{A})}\daleth_\phi(g)e^{-\left\langle{\lambda},{H_0(g)}\right\rangle}\ dg. \end{align*}$$

Therefore, Theorem 8.3 follows from (8.7) and Lemma 7.1 by noting that for any complement W of $\mathfrak {a}_{\operatorname {Ker} \rho }$ in $\mathfrak {a}_0$ , the positively- $1$ -homogeneous function $\left \langle {\mu },{x}\right \rangle +\sum _{i\in I}\left \langle {\beta _i},{x}\right \rangle _+$ on W is positive on the cone $\{x\in \mathfrak {a}_{0,+} \cap W \mid \left \langle {\chi _i},{x}\right \rangle \ge 0\ \forall i\in I\}\setminus \{0\}$ .

Remark 8.6. Let V be a regular PVS with $\Gamma _{\operatorname {spcl}}(V)=\{V\}$ . The example of $V=\operatorname {Mat}_n$ with $G=\operatorname {GL}_n$ acting by matrix multiplication shows that V is not necessarily CF. One may ask whether disjointness of the sets of weights $\Psi _{V_i}$ of the irreducible components $V_i$ of V is sufficient to imply that V is CF. This would follow if one could show that an F-irreducible PVS V such that $\Gamma _{\operatorname {spcl}}(V)=\{V\}$ is necessarily regular.Footnote 9

In the case of PVSs V of DK-type, the construction of §9 below shows that the condition $\Gamma _{\operatorname {spcl}}(V)=\{V\}$ implies that V is CF.

9 A construction of special subspaces for PVSs of DK-type

We finish the paper by providing a general construction of special subspaces for regular PVSs coming from nilpotent orbits (see §3).

Let $G'$ be a reductive group over F. Recall that any nilpotent gives rise to a filtration $\mathcal {F}(e)=(\mathcal {F}_i(e))_{i\in \mathbb {Z}}$ of whose stabilizer $\operatorname {Stab}(\mathcal {F}(e))$ in $G'$ is a parabolic subgroup of $G'$ with Lie algebra $\mathcal {F}_0(e)$ .

Clearly, $\mathcal {F}(\operatorname {Ad}(g)e)=\operatorname {Ad}(g)\mathcal {F}(e)$ for all $g\in G'$ . In particular, e and $\operatorname {Ad}(g)e$ define the same filtration if and only if $g\in \operatorname {Stab}(\mathcal {F}(e))$ .

If $\mathfrak {o}$ is the nilpotent orbit of e, then we say that $\mathcal {F}(e)$ is an $\mathfrak {o}$ -filtration. Thus, the variety of $\mathfrak {o}$ -filtrations is a $G'$ -orbit whose stabilizer is the canonical parabolic subgroup pertaining to $\mathfrak {o}$ .

9.1

We recall the notion of induction of nilpotent orbits [Reference Lusztig and Spaltenstein25].

Let Q be a parabolic F-subgroup of $G'$ with Lie algebra . Let L be the Levi quotient of Q, with Lie algebra . Let $\operatorname {pr}_Q:Q\rightarrow L$ (resp., ) be the canonical projections. Let $\mathfrak {o}^L$ be a nilpotent orbit in and let $\tilde {\mathfrak {o}}^L$ be its inverse image under . It is an irreducible subvariety of consisting of nilpotent elements. By definition, the induced orbit $\operatorname {Ind}_Q\mathfrak {o}^L$ is the nilpotent orbit in intersecting $\tilde {\mathfrak {o}}^L$ in an open subset. In fact, it depends only on L (and $\mathfrak {o}^L$ ), and not on Q, but it will be important for us to keep track of Q.

Definition 9.1. Suppose that $\mathfrak {o}=\operatorname {Ind}_Q\mathfrak {o}^L$ where Q is a parabolic F-subgroup of $G'$ with Levi quotient L and $\mathfrak {o}^L$ is a nilpotent orbit of . Let $\mathcal {F}^L$ be an $\mathfrak {o}^L$ -filtration of . We then say that the triple $\mathfrak {d}=(Q,\mathfrak {o}^L,\mathcal {F}^L)$ is an induced filtration data (IFD) for $\mathfrak {o}$ .

For instance, $(G',\mathfrak {o},\mathcal {F})$ is a (trivial) IFD for any $\mathfrak {o}$ -filtration $\mathcal {F}$ of .

The variety $\mathfrak {D}=\mathfrak {D}_{\mathfrak {o}}$ of IFDs for $\mathfrak {o}$ comprises finitely many $G'$ -orbits. (If $\mathfrak {D}_{\mathfrak {o}}$ is a single orbit, i.e., if $\mathfrak {o}$ is not induced non-trivially, then $\mathfrak {o}$ is called rigid.)

Lemma 9.2. Let $\mathfrak {d}=(Q,\mathfrak {o}^L,\mathcal {F}^L)\in \mathfrak {D}$ . Define inductively

and

Then $\tilde {\mathcal {F}} = (\tilde {\mathcal {F}}_i)_i$ is a filtration of of V-type.

Note that for all i but that the preimage strictly contains $\tilde {\mathcal {F}}_i$ for $i>2$ .

Proof. Without loss of generality, we may assume that Q is a standard parabolic subgroup of $G'$ and $\mathcal {F}^L$ the standard $\mathfrak {o}^L$ -filtration defined by a dominant element $h^L \in \mathfrak {a}_0^L$ , i.e. .

Let $h = h^L + h_L \in \mathfrak {a}_0^{G'}$ be the unique vector with $h_L \in \mathfrak {a}^{G'}_L$ and $\left \langle {\alpha },{h}\right \rangle = 2$ for all $\alpha \in \Delta _0^{G'} \setminus \Delta _0^Q$ . We claim that the inductively defined spaces $\tilde {\mathcal {F}}_i$ coincide with the filtration obtained from the gradation of defined by h: .

To see this, note that we have $\left \langle {\alpha },{h}\right \rangle = \left \langle {\alpha },{h^L}\right \rangle \in \{ 0, 1, 2 \}$ for $\alpha \in \Delta _0^Q$ (and therefore $\left \langle {\alpha },{h}\right \rangle \in \{ 0, 1, 2 \}$ for all $\alpha \in \Delta _0^{G'}$ ). This implies the claim for $i = 0$ , $1$ , $2$ . The claim for $i \ge 3$ follows, since any positive root $\alpha \in \Phi _{G'}$ with $\left \langle {\alpha },{h}\right \rangle \ge 3$ can be written as $\alpha = \beta + \gamma $ with positive roots $\beta $ and $\gamma $ and $\left \langle {\beta },{h}\right \rangle \in \{ 1, 2 \}$ .

The relation between $\tilde {\mathcal {F}}_i$ and $\tilde {\mathcal {F}}_{-i}$ holds for any filtration obtained from a gradation.

Definition 9.3. We call the filtration $(\tilde {\mathcal {F}}_i)_i$ of Lemma 9.2 the I-filtration associated to $\mathfrak {d}$ .

We say that a filtration $\tilde {\mathcal {F}}$ of is an I-filtration compatible with $\mathfrak {o}$ if it arises from some $\mathfrak {d} \in \mathfrak {D}$ by this construction.

The stabilizer $\operatorname {Stab}(\mathfrak {d})$ of $\mathfrak {d}=(Q,\mathfrak {o}^L,\mathcal {F}^L)\in \mathfrak {D}$ is the parabolic subgroup of Q whose image in L is $\operatorname {Stab}_L (\mathcal {F}^L)$ . It is also the stabilizer of the associated I-filtration $\tilde {\mathcal {F}}$ . Its Lie algebra is $\tilde {\mathcal {F}}_0$ .

$\mathfrak {o}$ -filtrations are I-filtrations compatible with $\mathfrak {o}$ , but in general there will be others. For example, if $\mathfrak {o}=\operatorname {Rich}(Q)$ , there is an I-filtration compatible with $\mathfrak {o}$ with , which is an $\mathfrak {o}$ -filtration only if $\mathfrak {o}$ is even and Q is conjugate to the canonical parabolic subgroup of $\mathfrak {o}$ . On the other hand, different IFDs may give rise to the same I-filtration.

Lemma 9.4. Let $\mathfrak {d}=(Q,\mathfrak {o}^L,\mathcal {F}^L)$ be an IFD for $\mathfrak {o}$ . Let $P'=\operatorname {Stab}(\mathfrak {d})$ and $(\tilde {\mathcal {F}}_i)_i$ be the associated I-filtration. Then,

  1. 1. $\mathfrak {o}_{\tilde {\mathcal {F}}} = \mathfrak {o}\cap \tilde {\mathcal {F}}_2$ is open in $\tilde {\mathcal {F}}_2$ .

  2. 2. .

  3. 3. $\mathfrak {o}_{\tilde {\mathcal {F}}}$ is a $P'$ -orbit.

  4. 4. Let $v\in \mathfrak {o}_{\tilde {\mathcal {F}}}$ . Then,

    $$\begin{align*}\{g\in G'\mid\operatorname{Ad}(g)^{-1}v\in\mathfrak{o}_{\tilde{\mathcal{F}}}\}=G^{\prime}_v P', \end{align*}$$
    and $(G^{\prime }_v)^\circ \subset P'$ .

Proof. For simplicity, we sometimes write $\bar {\phantom {x}}$ for the image under $\operatorname {pr}_Q$ or .

The first part follows since is open in for every $u\in \mathfrak {o}^L$ and $\mathfrak {o}^L\cap \mathcal {F}^L_2$ is open in $\mathcal {F}_2^L$ .

For the second part, suppose that with $\bar u\in \mathcal {F}_2^L\setminus \mathfrak {o}^L$ . Then, $\bar u$ lies in a nilpotent orbit $\mathfrak {o}'\ne \mathfrak {o}^L$ of in the closure of $\mathfrak {o}^L$ and u lies in the closure of $\operatorname {Ind}_Q\mathfrak {o}'$ . The latter has smaller dimension than $\mathfrak {o}$ . Thus, $u\notin \mathfrak {o}$ .

For the last part, suppose that $u=\operatorname {Ad}(g)^{-1}v\in \mathfrak {o}_{\tilde {\mathcal {F}}}$ . Then, by the second part $\bar u\in \mathfrak {o}^L$ . By [Reference Lusztig and Spaltenstein25, Theorem 1.3(d)], $g\in G^{\prime }_vQ$ . Write $g=hq$ where $h\in G^{\prime }_v$ , $q\in Q$ . Then, $\bar u=\operatorname {Ad}(\bar {q})^{-1}\bar v$ . Hence, $\operatorname {Ad}(\bar q)^{-1}\bar v\in \mathfrak {o}^L \cap \mathcal {F}^L_2$ . By Theorem 3.1 part 4 (applied to $\mathfrak {o}^L$ ), we infer that $\bar q\in \operatorname {pr}_Q(P')$ , so that $q\in P'$ . Hence $g\in G^{\prime }_v P'$ .

Also, $(G^{\prime }_v)^\circ \subset Q$ again by [Reference Lusztig and Spaltenstein25, Theorem 1.3(d)], since $\bar v\in \mathfrak {o}^L$ by the second part. Thus, $(G^{\prime }_v)^\circ =Q_v^\circ $ . Since $\operatorname {pr}_Q(Q_v)\subset L_{\bar v}\subset \operatorname {pr}_Q(P')$ , by Theorem 3.1 we deduce that $(G^{\prime }_v)^\circ \subset P'$ .

Finally, the third part follows from the last part.

9.2

We fix $P_0'$ , $T_0'$ as in §3. The following standard fact is an easy consequence of the Bruhat decomposition.

Remark 9.5. Let P be a standard parabolic F-subgroup of $G'$ with standard Levi subgroup M. Then, for any parabolic F-subgroup Q of $G'$ there exists a unique parabolic subgroup $Q'$ of the form $Q'=Q^p$ , $p\in P(F)$ such that $Q'\cap M$ is a standard parabolic subgroup of M.

Namely, if Q is conjugate to the standard parabolic F-subgroup $Q"$ of G, then we can write $Q = (Q")^{n p}$ with $n \in N_{G'(F)} (T_0')$ and $p \in P(F)$ by the Bruhat decomposition, and moreover the Weyl group element $\bar {n} \in W = N_{G'(F)} (T_0') / T_0' (F)$ defined by n can be taken of minimal length in its $(W_{Q"}, W_P)$ -double coset. Then $Q'=(Q")^n$ works by [Reference Carter6, Proposition 2.8.9], and uniqueness follows from [Reference Carter6, Proposition 2.3.3, 2.8.1].

Now let $\mathfrak {o}$ be a nilpotent orbit in and let $(G,V,\rho )$ be the regular PVS of DK-type attached to $\mathfrak {o}$ (cf. §3). Thus, G is the standard Levi subgroup of the (standard) canonical parabolic subgroup P of $\mathfrak {o}$ . We write $\mathfrak {o}_{\mathcal {F}}=\mathfrak {o}\cap \mathcal {F}_2$ , where $\mathcal {F}$ is the standard $\mathfrak {o}$ -filtration (i.e., with $\operatorname {Stab}(\mathcal {F})=P$ ).

Denote by $[\mathfrak {F}]$ the (finite) set of $G'$ -orbits of I-filtrations compatible with $\mathfrak {o}$ . The elements of $[\mathfrak {F}]$ give rise to special subspaces of V. More precisely,

Proposition 9.6. For any $G'$ -orbit $\mathfrak {C}\in [\mathfrak {F}]$ there exists a unique I-filtration $\tilde {\mathcal {F}} \in \mathfrak {C}$ such that $\operatorname {Stab}(\tilde {\mathcal {F}})\cap G$ is a standard parabolic subgroup of G and $\mathfrak {o} \cap \mathcal {F}_2 \cap \tilde {\mathcal {F}}_2 \ne \emptyset $ . Moreover, $U:=\tilde {\mathcal {F}}_2\cap V\in \Gamma _{\operatorname {spcl}}(V)$ and $\operatorname {Stab}(U)=\operatorname {Stab}(\tilde {\mathcal {F}})\cap G$ .

Proof. We first show that the set

$$\begin{align*}\mathfrak{C}'=\{\tilde{\mathcal{F}}\in\mathfrak{C}\mid \mathfrak{o} \cap \mathcal{F}_2 \cap \tilde{\mathcal{F}}_2 \ne\emptyset\}, \end{align*}$$

is a P-orbit. Clearly, $\mathfrak {C}'$ is stable under P, and it is non-empty since $\mathfrak {o} \cap \tilde {\mathcal {F}}_2 \ne \emptyset $ for all $\tilde {\mathcal {F}} \in \mathfrak {C}$ . Assume that $\tilde {\mathcal {F}}\in \mathfrak {C}'$ and $\operatorname {Ad}(g)\tilde {\mathcal {F}}\in \mathfrak {C}'$ for some $g\in G'$ . We need to show that $g\in PP'$ for $P'=\operatorname {Stab}(\tilde {\mathcal {F}})$ . By assumption, $\mathfrak {o}_{\tilde {\mathcal {F}}}$ intersects both $\mathfrak {o}_{\mathcal {F}}$ and $\operatorname {Ad}(g)^{-1}\mathfrak {o}_{\mathcal {F}}$ . Since $\mathfrak {o}_{\tilde {\mathcal {F}}}$ is a $P'$ -orbit by Lemma 9.4 part 3, there exists $p'\in P'$ such that

$$\begin{align*}\operatorname{Ad}(p')(\mathfrak{o}_{\mathcal{F}})\cap\operatorname{Ad}(g)^{-1}(\mathfrak{o}_{\mathcal{F}})\ne\emptyset. \end{align*}$$

By Theorem 3.1 we conclude that $gp'\in P$ , as required.

The first assertion follows now from Remark 9.5. Let $\tilde {\mathcal {F}} = (\tilde {\mathcal {F}}_i)_i \in \mathfrak {C}'$ be the unique I-filtration provided by this construction. As before, let $P'=\operatorname {Stab}(\tilde {\mathcal {F}})$ . Since $P'$ contains $T_0'$ , we have

$$\begin{align*}\tilde{\mathcal{F}}_2\cap\mathcal{F}_2=(\tilde{\mathcal{F}}_2\cap V)+(\tilde{\mathcal{F}}_2\cap\mathcal{F}_3). \end{align*}$$

Since $\mathfrak {o}_{\mathcal {F}}=X+\mathcal {F}_3$ by Theorem 3.1, we conclude that

$$\begin{align*}\mathfrak{o}_{\tilde{\mathcal{F}}}\cap\mathfrak{o}_{\mathcal{F}}=\mathfrak{o}\cap\tilde{\mathcal{F}}_2\cap\mathcal{F}_2=(\mathfrak{o}_{\tilde{\mathcal{F}}}\cap X)+(\tilde{\mathcal{F}}_2\cap\mathcal{F}_3). \end{align*}$$

In particular, $\mathfrak {o}_{\tilde {\mathcal {F}}}\cap X\ne \emptyset $ , so that $U\in \Gamma _{\operatorname {reg}}(V)$ .

Clearly, $\operatorname {Stab}(U)$ contains the standard parabolic subgroup $P'\cap G$ of G. Let $v_0\in \mathfrak {o}\cap U$ . It follows from Lemma 9.4 part 4 that there exist $\eta _1,\dots ,\eta _k\in G'$ such that

$$\begin{align*}\{g\in G\mid \operatorname{Ad}(g)^{-1}v_0\in U\}=\{g\in G\mid \operatorname{Ad}(g)^{-1}v_0\in\mathfrak{o}_{\tilde{\mathcal{F}}}\}=\bigcup_{1 \le i \le k} \eta_iP'\cap G. \end{align*}$$

Each non-empty intersection $\eta _iP'\cap G$ is a left coset of $P'\cap G$ in G. The second assertion of the proposition now follows from Proposition 6.4 part 4.

For instance, if $\mathfrak {o}=\operatorname {Rich}(Q)$ , where Q is a parabolic F-subgroup of $G'$ , then Proposition 9.6 says that there is a unique $G(F)$ -conjugate $Q'$ of Q such that and $Q'\cap G$ is a standard parabolic subgroup of G. In general, $Q'\cap P$ is not a parabolic subgroup of $G'$ . In other words, if Q is standard, it is not necessary that $Q'=Q$ .

Consider the following examples.

  1. 1. Rank two cases (see Example 6.8).

    We may write the subregular orbit $\mathfrak {o}$ as $\operatorname {Rich}(Q)$ where $\Delta _0^Q=\{\beta \}$ . This IFD gives rise to , the unique non-trivial special subspace.

  2. 2. Example 6.14 (= Example 3.2 with $k=3$ and $n_1<n_2$ ).

    We have $\mathfrak {o}=\operatorname {Rich}(Q_1)=\operatorname {Rich}(Q_2)$ where $Q_1$ and $Q_2$ are the standard parabolic subgroups of type $(n_2,n_1,n_1)$ and $(n_1,n_1,n_2)$ . These IFDs give rise to the spaces in $\Gamma _{\operatorname {spcl}}(V)$ .

    On the other hand, if $k=5$ and $n_1<n_2<n_3$ and Q is the standard parabolic subgroup of type $(n_3,n_2,n_2,n_1,n_1)$ , then $\mathfrak {o}=\operatorname {Rich}(Q)$ but .

  3. 3. Example 6.15 (= Example 3.8). In the following two examples, for standard parabolic subgroups $Q_1 \supset Q_2$ we will use the notation $\operatorname {Rich}^{Q_1}(Q_2)$ for the Richardson orbit of the standard parabolic subgroup $L_1 \cap Q_2$ inside the standard Levi subgroup $L_1$ of $Q_1$ .

    We have $\mathfrak {o}=\operatorname {Rich}(({*}{*}0{*}))=\operatorname {Rich}((0{*}{*}0))$ . This will give rise to the subspaces $U_1$ and $U_2$ , respectively. Using induction by stages, we may also write $\mathfrak {o}=\operatorname {Ind}_Q\mathfrak {o}'$ where $Q=(0{*}{*}{*})$ and $\mathfrak {o}'=\operatorname {Rich}^Q((0{*}{*}0))$ .

    The weighted Dynkin diagram of $\mathfrak {o}'$ is The standard $\mathfrak {o}'$ -filtration satisfies $\Phi _{\mathcal {F}^{\prime }_2}=\{(0120),(0121),(0122)\}$ , and the associated I-filtration $\Phi _{\tilde {\mathcal {F}}_2}= \Phi ^+_{G'} \setminus \{(0100),(0010),(0001),(0110),(0011),(0111)\}$ . Therefore, this IFD gives rise to the space  $U_3$ .

    Note that writing $\mathfrak {o}=\operatorname {Ind}_{({*}{*}{*}0)}\mathfrak {o}"$ where $\mathfrak {o}"=\operatorname {Rich}^{({*}{*}{*}0)}((0{*}{*}0))$ will give rise to $U_2$ again, since the weighted Dynkin diagram of $\mathfrak {o}"$ is and therefore again.

  4. 4. Example 7.8 (= Example 3.9)

    We have $\mathfrak {o}=\operatorname {Rich}(Q^{\prime }_1)=\operatorname {Rich}(Q^{\prime }_2)$ , where , (which are associate). These give rise to $U_1$ and $U_2$ , respectively.

    They also give three presentations $\mathfrak {o}=\operatorname {Ind}_{Q_i}\mathfrak {o}^{\prime }_i$ , $i=1,2,3$ where

    Note that $L_1^{\operatorname {der}}=\operatorname {SL}_5\times \operatorname {SL}_2$ and $\mathfrak {o}^{\prime }_1$ is the minimal orbit of the $\operatorname {SL}_5$ -factor. The weighted Dynkin diagram is . We have . We get the space $U_1\cap U_2$ .

    The weighted Dynkin diagram of $\mathfrak {o}^{\prime }_2$ is and

    We get the space $U_2"$ .

    The weighted Dynkin diagram of $\mathfrak {o}^{\prime }_3$ is , and we get $U_1$ again.

    Of course, we can apply the involution $\theta $ to all of the above, obtaining in addition $U_1'$ , $U_2'$ and $U_1' \cap U_2'$ .

    Finally, we also have $\mathfrak {o}=\operatorname {Ind}_Q\mathfrak {o}'$ where and $\mathfrak {o}'$ is the minimal orbit of the Levi subgroup of Q (whose derived group is $\operatorname {Spin}_8$ ). The weighted Dynkin diagram of $\mathfrak {o}'$ is and . Here, $\tilde {\mathcal {F}}_2\cap \mathfrak {o}_{\mathcal {F}}=\emptyset $ for the standard I-filtration $\tilde {\mathcal {F}}$ . However, it is easy to see that if w is the simple reflection with respect to (the simple root defining G), then $\operatorname {Stab}(w\tilde {\mathcal {F}}) \cap G$ is the standard Borel subgroup of G and $w\tilde {\mathcal {F}}_2\cap \mathfrak {o}_{\mathcal {F}}\ne \emptyset $ . The intersection $w\tilde {\mathcal {F}}_2\cap V$ is $U_3"$ .

    Note that $U_3$ is not obtained from this construction (either directly or by intersection).

Acknowledgment

The first-named author would like to thank the Weizmann Institute of Science for its hospitality.

Footnotes

1 This condition depends on v. A strictly stronger condition would be that $H^\circ $ is not contained in any proper parabolic subgroup of G (not necessarily defined over F).

2 The terminology comes from nilpotent orbits—see §3 below.

3 The general case is obtained from these by castling transforms, see [Reference Kimura23, §7.1]. These preserve irreducibility and the generic stabilizer but may yield PVSs which are far removed from nilpotent orbits.

4 The examples not of DK-type are # 17, # 18, # 19, # 22, # 25 and # 26. Example # 1 can be of DK-type or not.

5 Note that both conditions imply that $H(\mathbb {A})\cap G(\mathbb {A})^1$ is unimodular.

6 Yukie proved a general qualitative convergence result for distinguished, absolutely irreducible PVSs.

7 Strictly speaking, Saito assumes that there exists $v\in X(F)$ such that $X^*(G_v^\circ )=1$ . By Lemma 2.2 and Lemma 2.6 this implies that $v\in X(F)^{\operatorname {niso}}$ and $X^*(H)$ is finite. Conversely, if $X^*(H)$ is finite, then $X(F)^{\operatorname {niso}}=\{v\in X(F)\mid X^*(G_v^\circ )=1\}$ . Also, he considers a more general situation where G is a connected linear algebraic group and X the complement of a hypersurface.

8 Case # 1 can be of DK-type or not, but it is always basic.

9 This can be easily checked when F is algebraically closed by classification.

References

Bhargava, M., Higher composition laws and applications , Int. Congr. Math., Eur. Math. Soc., Zürich II (2006), 271294. MR 2275598Google Scholar
Borel, A., Linear algebraic groups, second ed., Graduate Texts in Mathematics, Vol. 126, Springer-Verlag, New York, 1991. MR1102012 (92d:20001)CrossRefGoogle Scholar
Bourbaki, N., Lie groups and Lie algebras. Chapters 7–9, Elements of Mathematics (Berlin), Springer-Verlag, Berlin, 2005, Translated from the 1975 and 1982 French originals by Andrew Pressley. MR2109105 (2005h:17001)Google Scholar
Borel, A. and Tits, J., Éléments unipotents et sous-groupes paraboliques de groupes réductifs. I , Invent. Math. 12 (1971), 95104. MR0294349 (45 #3419)CrossRefGoogle Scholar
Bhargava, M., Taniguchi, T. and Thorne, F., Improved error estimates for the Davenport-Heilbronn theorems , Math. Ann. 389 (2024), 34713512.CrossRefGoogle Scholar
Carter, R. W., Finite groups of Lie type, Pure and Applied Mathematics (New York), John Wiley & Sons Inc., New York, 1985, Conjugacy classes and complex characters, A Wiley-Interscience Publication. MR794307 (87d:20060)Google Scholar
Chaudouard, P.-H., Sur la contribution unipotente dans la formule des traces d’Arthur pour les groupes généraux linéaires , Israel J. Math. 218 (2017), no. 1, 175271. MR 3625130CrossRefGoogle Scholar
Collingwood, D. H. and McGovern, W. M., Nilpotent orbits in semisimple Lie algebras, Van Nostrand Reinhold Mathematics Series, Van Nostrand Reinhold Co., New York, 1993. MR1251060 (94j:17001)Google Scholar
Datskovsky, B. and Wright, D. J., Density of discriminants of cubic extensions , J. Reine Angew. Math. 386 (1988), 116138. MR 936994Google Scholar
Finis, T., Hoffmann, W. and Wakatsuki, S., The subregular unipotent contribution to the geometric side of the Arthur trace formula for the split exceptional group ${G}_2$ , Geometric aspects of the trace formula, Simons Symp., Springer, Cham, 2018, pp. 163182. MR 3969874CrossRefGoogle Scholar
Finis, T. and Lapid, E., On the continuity of the geometric side of the trace formula , Acta Math. Vietnam. 41 (2016), no. 3, 425455. MR 3534542CrossRefGoogle Scholar
Freudenthal, H., Beziehungen der ${E}_7$ und ${E}_8$ zur Oktavenebene. I, Indag. Math. 16 (1954), 218230, Nederl. Akad. Wetensch. Proc. Ser. A 57. MR 63358CrossRefGoogle Scholar
Gyoja, A., Invariants, nilpotent orbits, and prehomogeneous vector spaces , J. Algebra 142 (1991), no. 1, 210232. MR 1125214CrossRefGoogle Scholar
Hartshorne, R., Algebraic geometry, Graduate Texts in Mathematics, Vol. 52, Springer-Verlag, New York-Heidelberg, 1977. MR 463157CrossRefGoogle Scholar
Hochschild, G. P., Basic theory of algebraic groups and Lie algebras, Graduate Texts in Mathematics, Vol. 75, Springer-Verlag, New York-Berlin, 1981. MR 620024CrossRefGoogle Scholar
Hoffmann, W., The nonsemisimple term in the trace formula for rank one lattices , J. Reine Angew. Math. 379 (1987), 121. MR 903631Google Scholar
Hoffmann, W., The trace formula and prehomogeneous vector spaces , Families of automorphic forms and the trace formula, Simons Symp., Springer, Cham, 2016, pp. 175215. MR 3675167CrossRefGoogle Scholar
Hilton, P. J. and Stammbach, U., A course in homological algebra, second ed.,, Graduate Texts in Mathematics, Vol. 4, Springer-Verlag, New York, 1997. MR 1438546CrossRefGoogle Scholar
Hoffmann, W. and Wakatsuki, S., On the geometric side of the Arthur trace formula for the symplectic group of rank 2 , Mem. Amer. Math. Soc. 255 (2018), no. 1224, v+88. MR 3843151Google Scholar
Igusa, J.-i., Exponential sums associated with a Freudenthal quartic , J. Fac. Sci. Univ. Tokyo Sect. IA Math. 24 (1977), no. 1, 231245. MR 0447140 (56 #5455)Google Scholar
Igusa, J.-i., On a certain class of prehomogeneous vector spaces , J. Pure Appl. Algebra 47 (1987), no. 3, 265282. MR 910424 (88m:20089)CrossRefGoogle Scholar
Kac, V. G., Some remarks on nilpotent orbits , J. Algebra 64 (1980), no. 1, 190213. MR 575790 (81i:17005)CrossRefGoogle Scholar
Kimura, T., Introduction to prehomogeneous vector spaces, Translations of Mathematical Monographs, Vol. 215, American Mathematical Society, Providence, RI, 2003, Translated from the 1998 Japanese original by Makoto Nagura and Tsuyoshi Niitani and revised by the author. MR 1944442Google Scholar
Kable, A. C. and Yukie, A., On the number of quintic fields , Invent. Math. 160 (2005), no. 2, 217259. MR 2138067CrossRefGoogle Scholar
Lusztig, G. and Spaltenstein, N., Induced unipotent classes , J. London Math. Soc. (2) 19 (1979), no. 1, 4152. MR 527733CrossRefGoogle Scholar
Mœglin, C. and Waldspurger, J.-L., Spectral decomposition and Eisenstein series, Cambridge Tracts in Mathematics, Vol. 113, Cambridge University Press, Cambridge, 1995, Une paraphrase de l’Écriture [A paraphrase of Scripture]. MR 1361168CrossRefGoogle Scholar
Mostow, G. D., Fully reducible subgroups of algebraic groups , Amer. J. Math. 78 (1956), 200221. MR0092928 (19,1181f)CrossRefGoogle Scholar
Revêtements étales et groupe fondamental (SGA 1) , Documents Mathématiques (Paris) [Mathematical Documents (Paris)], vol. 3, Société Mathématique de France, Paris, 2003, Séminaire de géométrie algébrique du Bois Marie 1960–61. [Algebraic Geometry Seminar of Bois Marie 1960-61], Directed by A. Grothendieck, With two papers by M. Raynaud, Updated and annotated reprint of the 1971 original [Lecture Notes in Math., 224, Springer, Berlin; MR0354651 (50 #7129)]. MR 2017446Google Scholar
Morishita, M. and Watanabe, T., A note on the mean value theorem for special homogeneous spaces , Nagoya Math. J. 143 (1996), 111117. MR 1413009 (97h:11042)CrossRefGoogle Scholar
Ono, T., A mean value theorem in adele geometry , J. Math. Soc. Japan 20 (1968), 275288. MR 0230726CrossRefGoogle Scholar
Popov, V. L., Picard groups of homogeneous spaces of linear algebraic groups and one-dimensional homogeneous vector fiberings , Izv. Akad. Nauk SSSR Ser. Mat. 38 (1974), 294322. MR 357399Google Scholar
Röhrle, G. E., On extraspecial parabolic subgroups , Linear algebraic groups and their representations (Los Angeles, CA, 1992), Contemp. Math., Vol. 153, Amer. Math. Soc., Providence, RI, 1993, pp. 143155. MR 1247502Google Scholar
Raynaud, M., Faisceaux amples sur les schémas en groupes et les espaces homogènes, Lecture Notes in Mathematics, Vol. 119, Springer-Verlag, Berlin-New York, 1970. MR 260758Google Scholar
Ranga Rao, R., Orbital integrals in reductive groups , Ann. Math. (2) 96 (1972), 505510. MR0320232 (47 #8771)CrossRefGoogle Scholar
Rubenthaler, H., Decomposition of reductive regular prehomogeneous vector spaces , Ann. Inst. Fourier (Grenoble) 61 (2011), no. 5, 21832218 (2012). MR 2961852CrossRefGoogle Scholar
Saito, H., Explicit form of the zeta functions of prehomogeneous vector spaces , Math. Ann. 315 (1999), no. 4, 587615. MR 1731463 (2001b:11111)CrossRefGoogle Scholar
Saito, H., Global zeta functions of Freudenthal quartics , Internat. J. Math. 13 (2002), no. 8, 797820. MR 1928799 (2003i:11051)CrossRefGoogle Scholar
Saito, H., Convergence of the zeta functions of prehomogeneous vector spaces , Nagoya Math. J. 170 (2003), 131. MR 1994885 (2005d:11170)CrossRefGoogle Scholar
Satō, F., Zeta functions in several variables associated with prehomogeneous vector spaces. I. Functional equations , Tohoku Math. J. (2) 34 (1982), no. 3, 437483. MR 676121CrossRefGoogle Scholar
Satō, F., Zeta functions in several variables associated with prehomogeneous vector spaces. II. A convergence criterion , Tohoku Math. J. (2) 35 (1983), no. 1, 7799. MR 695661CrossRefGoogle Scholar
Sato, M., Theory of prehomogeneous vector spaces (algebraic part)—the English translation of Sato’s lecture from Shintani’s note , Nagoya Math. J. 120 (1990), 134, Notes by Takuro Shintani, Translated from the Japanese by Masakazu Muro. MR 1086566CrossRefGoogle Scholar
Serre, J.-P., Galois cohomology , Springer-Verlag, Berlin, 1997, Translated from the French by Patrick Ion and revised by the author. MR 1466966 (98g:12007)Google Scholar
Shintani, T., On Dirichlet series whose coefficients are class numbers of integral binary cubic forms , Proc. Japan Acad. 46 (1970), 909911. MR0292755 (45 #1837)Google Scholar
Shintani, T., On Dirichlet series whose coefficients are class numbers of integral binary cubic forms , J. Math. Soc. Japan 24 (1972), 132188. MR0289428 (44 #6619)CrossRefGoogle Scholar
Shintani, T., On zeta-functions associated with the vector space of quadratic forms , J. Fac. Sci. Univ. Tokyo Sect. I A Math. 22 (1975), 2565. MR0384717 (52 #5590)Google Scholar
Sato, M. and Kimura, T., A classification of irreducible prehomogeneous vector spaces and their relative invariants , Nagoya Math. J. 65 (1977), 1155. MR 430336CrossRefGoogle Scholar
Sato, M. and Shintani, T., On zeta functions associated with prehomogeneous vector spaces , Proc. Nat. Acad. Sci. U.S.A. 69 (1972), 10811082. MR 296079CrossRefGoogle ScholarPubMed
Sato, M. and Shintani, T., On zeta functions associated with prehomogeneous vector spaces , Ann. of Math. (2) 100 (1974), 131170. MR0344230 (49 #8969)CrossRefGoogle Scholar
Vinberg, È. B., The Weyl group of a graded Lie algebra , Izv. Akad. Nauk SSSR Ser. Mat. 40 (1976), no. 3, 488526, 709. MR 0430168 (55 #3175)Google Scholar
Wright, D. J., The adelic zeta function associated to the space of binary cubic forms. I. Global theory , Math. Ann. 270 (1985), no. 4, 503534. MR 776169CrossRefGoogle Scholar
Wright, D. J. and Yukie, Akihiko, Prehomogeneous vector spaces and field extensions , Invent. Math. 110 (1992), no. 2, 283314. MR 1185585CrossRefGoogle Scholar
Yukie, A., On the Shintani zeta function for the space of binary quadratic forms , Math. Ann. 292 (1992), no. 2, 355374. MR 1149040CrossRefGoogle Scholar
Yukie, Akihiko, Shintani zeta functions , London Mathematical Society Lecture Note Series, Vol. 183, Cambridge University Press, Cambridge, 1993. MR 1267735Google Scholar
Yukie, A., On the convergence of the zeta function for certain prehomogeneous vector spaces , Nagoya Math. J. 140 (1995), 131. MR 1369477CrossRefGoogle Scholar
Figure 0

Table 1 Irreducible basic PVSs of DK-type for exceptional groups

Figure 1

Table 2 PVSs of DK-type pertaining to the Freudenthal–Tits magic square