Hostname: page-component-7bb8b95d7b-s9k8s Total loading time: 0 Render date: 2024-09-20T09:19:41.394Z Has data issue: false hasContentIssue false

Non-concentration property of Patterson–Sullivan measures for Anosov subgroups

Published online by Cambridge University Press:  18 September 2024

DONGRYUL M. KIM
Affiliation:
Department of Mathematics, Yale University, New Haven, CT 06511, USA (e-mail: dongryul.kim@yale.edu)
HEE OH*
Affiliation:
Department of Mathematics, Yale University, New Haven, CT 06511, USA (e-mail: dongryul.kim@yale.edu)
*
Rights & Permissions [Opens in a new window]

Abstract

Let G be a connected semisimple real algebraic group. For a Zariski dense Anosov subgroup $\Gamma <G$ with respect to a parabolic subgroup $P_\theta $, we prove that any $\Gamma $-Patterson–Sullivan measure charges no mass on any proper subvariety of $G/P_\theta $. More generally, we prove that for a Zariski dense $\theta $-transverse subgroup $\Gamma <G$, any $(\Gamma , \psi )$-Patterson–Sullivan measure charges no mass on any proper subvariety of $G/P_\theta $, provided the $\psi $-Poincaré series of $\Gamma $ diverges at its abscissa of convergence. In particular, our result also applies to relatively Anosov subgroups.

Type
Original Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2024. Published by Cambridge University Press

1. Introduction

Let G be a connected semisimple real algebraic group and $\mathfrak g=\operatorname {Lie} G$ . Let A be a maximal real split torus of G and set $\mathfrak {a} = \operatorname {Lie} A$ . Fix a positive Weyl chamber $\mathfrak a^+<\mathfrak a$ and a maximal compact subgroup $K< G$ such that the Cartan decomposition $G=K (\exp \mathfrak a^+) K$ holds. We denote by $\mu (g)\in \mathfrak a$ the Cartan projection of $g\in G$ , that is, the unique element of $\mathfrak a^+$ such that $g\in K \exp (\mu (g))K$ . Let $\Pi $ be the set of all simple roots for $(\mathfrak g, \mathfrak a^+)$ and fix a non-empty subset $\theta \subset \Pi $ . Let $P_\theta $ be the standard parabolic subgroup of G corresponding to $\theta $ and set

$$ \begin{align*}\mathcal F_\theta=G/P_\theta.\end{align*} $$

Let $\Gamma <G$ be a Zariski dense discrete subgroup. Denote by $\Lambda _\theta \subset \mathcal F_{\theta }$ the limit set of $\Gamma $ , which is the unique $\Gamma $ -minimal subset of $\mathcal F_\theta $ [Reference Benoist1]. Let $\mathfrak {a}_\theta =\bigcap _{\alpha \in \Pi - \theta } \ker \alpha $ . For a linear form $\psi \in \mathfrak a_\theta ^*$ , a Borel probability measure $\nu $ on $\mathcal {F}_\theta $ is called a $(\Gamma , \psi )$ -conformal measure if

$$ \begin{align*}\frac{d \gamma_*\nu}{d\nu}(\xi)=e^{\psi(\beta_\xi^\theta(e,\gamma))} \quad \text{for all}\ \gamma \in \Gamma\ \text{and}\ \xi \in \mathcal{F}_\theta, \end{align*} $$

where $\gamma _* \nu (B) = \nu (\gamma ^{-1}B)$ for any Borel subset $B\subset \mathcal F_\theta $ and $\beta _\xi ^\theta $ denotes the $\mathfrak a_\theta $ -valued Busemann map defined in equation (2.2). By a $\Gamma $ -Patterson–Sullivan measure on $\mathcal F_\theta $ , we mean a $(\Gamma ,\psi )$ -conformal measure supported on $\Lambda _\theta $ for some $\psi \in \mathfrak a_{\theta }^*$ .

Patterson–Sullivan measures play a fundamental role in the study of geometry and dynamics for $\Gamma $ -actions. For G of rank one, they were constructed by Patterson and Sullivan for any non-elementary discrete subgroup $\Gamma $ of G [Reference Patterson17, Reference Sullivan22], and hence the name. Their construction was generalized by Quint for any Zariski dense subgroup of a semisimple real algebraic group [Reference Quint19].

A finitely generated subgroup $\Gamma <G$ is called a $\theta $ -Anosov subgroup if there exist $C_1, C_2>0$ such that for all $\gamma \in \Gamma $ and $\alpha \in \theta $ ,

$$ \begin{align*}\alpha(\mu(\gamma))\ge C_1|\gamma| -C_2,\end{align*} $$

where $|\gamma |$ denotes the word length of $\gamma $ with respect to a fixed finite generating set of $\Gamma $ . A $\theta $ -Anosov subgroup is necessarily a word hyperbolic group [Reference Kapovich, Leeb and Porti11, Theorem 1.5, Corollary 1.6]. The notion of Anosov subgroups was first introduced by Labourie for surface groups [Reference Labourie15], and was extended to general word hyperbolic groups by Guichard and Wienhard [Reference Guichard and Wienhard8]. Several equivalent characterizations have been established, one of which is the above definition (see [Reference Guéritaud, Guichard, Kassel and Wienhard7, Reference Kapovich and Leeb9Reference Kapovich, Leeb and Porti11]). Anosov subgroups are regarded as natural generalizations of convex cocompact subgroups of rank one groups, and include the images of Hitchin representations and of maximal representations as well as higher rank Schottky subgroups; see [Reference Kassel12, Reference Wienhard23].

A special case of our main theorem is the following non-concentration property of Patterson–Sullivan measures for $\theta $ -Anosov subgroups.

Theorem 1.1. Let $\Gamma <G$ be a Zariski dense $\theta $ -Anosov subgroup. For any $\Gamma $ -Patterson–Sullivan measure $\nu $ on $\mathcal F_\theta $ , we have

$$ \begin{align*}\nu(S)=0\end{align*} $$

for any proper subvariety S of $ \mathcal F_\theta $ .

Remark 1.2. This was proved by Flaminio and Spatzier [Reference Flaminio and Spatzier6] for $G=\operatorname {SO}(n,1)$ , $n\ge 2$ , and by Edwards, Lee, and Oh [Reference Edwards, Lee and Oh5] when $\theta =\Pi $ and the opposition involution of G is trivial in equation (2.1).

Indeed, we work with a more general class of discrete subgroups, called $\theta $ -transverse subgroups. Denote by $\operatorname {i}$ the opposition involution of G (see equation (2.1)).

Definition 1.3. A discrete subgroup $\Gamma < G$ is called $\theta $ -transverse if:

  • it is $\theta $ -regular, that is, $ \liminf _{\gamma \in \Gamma } \alpha (\mu ({\gamma }))=\infty $ for all $\alpha \in \theta $ ; and

  • it is $\theta $ -antipodal, that is, any two distinct $\xi , \eta \in \Lambda _{\theta \cup \operatorname {i}(\theta )}$ are in general position.

Since $\operatorname {i} (\mu ( g))=\mu (g^{-1}) $ for all $g\in G$ , it follows that $\Gamma $ is $\theta $ -transverse if and only if $\Gamma $ is $\operatorname {i}(\theta )$ -transverse. The class of $\theta $ -transverse subgroups includes all discrete subgroups of rank one Lie groups, $\theta $ -Anosov subgroups, and relatively $\theta $ -Anosov subgroups.

Let $p_{\theta } : \mathfrak a \to \mathfrak a_{\theta }$ be the projection which is invariant under all Weyl elements fixing $\mathfrak a_{\theta }$ pointwise, and set $\mu _\theta =p_\theta \circ \mu $ . A linear form $\psi \in \mathfrak a_{\theta }^*$ is said to be $(\Gamma , \theta )$ -proper if the composition $\psi \circ \mu _\theta : \Gamma \to [-\varepsilon , \infty )$ is a proper map for some $\varepsilon> 0$ . The following is our main theorem from which Theorem 1.1 is deduced by applying Selberg’s lemma [Reference Selberg21].

Theorem 1.4. Let $\Gamma <G$ be a Zariski dense virtually torsion-free $\theta $ -transverse subgroup. Let $\psi \in \mathfrak a_{\theta }^*$ be a $(\Gamma , \theta )$ -proper linear form such that $\sum _{\gamma \in \Gamma } e^{-\psi (\mu _{\theta }(\gamma ))} = \infty $ . For any $(\Gamma , \psi )$ -Patterson–Sullivan measure $\nu $ on $\mathcal F_{\theta }$ , we have

$$ \begin{align*}\nu(S)=0\end{align*} $$

for any proper subvariety S of $ \mathcal F_\theta $ .

For a $\theta $ -Anosov $\Gamma $ , the existence of a $(\Gamma , \psi )$ -Patterson–Sullivan measure implies that $\psi $ is $(\Gamma , \theta )$ -proper and $\sum _{\gamma \in \Gamma } e^{-\psi (\mu _{\theta }(\gamma ))} = \infty $ [Reference Lee and Oh16, Reference Sambarino20]. Therefore, Theorem 1.1 is a special case of Theorem 1.4.

The following is added to the proof. The growth indicator $\psi _\Gamma ^\theta $ is a higher rank version of the classical critical exponent of $\Gamma $ [Reference Kim, Oh and Wang13, Reference Quint18]. For a Zariski dense $\theta $ -transverse subgroup and a $(\Gamma , \theta )$ -proper $\psi \in \mathfrak a_{\theta }^*$ , the existence of a $(\Gamma , \psi )$ -conformal measure implies that $\psi $ is bounded from below by $\psi _\Gamma ^\theta $ ([Reference Quint19, Theorem 8.1] for $\theta = \Pi $ , [Reference Kim, Oh and Wang13, Theorem 1.4] for a general $\theta $ ). When $\Gamma $ is relatively $\theta $ -Anosov and $\psi $ is tangent to $\psi _\Gamma ^\theta $ , the abscissa of convergence of the series $s \mapsto \sum _{\gamma \in \Gamma } e^{-s\psi (\mu _{\theta }(\gamma ))}$ is equal to $1$ , and a recent work [Reference Canary, Zhang and Zimmer3, Theorem 1.1] shows that $\sum _{\gamma \in \Gamma } e^{-\psi (\mu _{\theta }(\gamma ))} = \infty $ . Therefore, Theorem 1.4 also applies in this setting.

2. Ergodic properties of Patterson–Sullivan measures

Let G be a connected semisimple real algebraic group. Let $P < G$ be a minimal parabolic subgroup with a fixed Langlands decomposition $P=MAN, $ where A is a maximal real split torus of G, M is a maximal compact subgroup commuting with A, and N is the unipotent radical of P. We fix a positive Weyl chamber $\mathfrak {a}^+\subset \mathfrak {a} = \operatorname {Lie}A$ so that $\log N$ consists of positive root subspaces. Recall that $K< G$ denotes a maximal compact subgroup such that the Cartan decomposition $G=K(\exp \mathfrak a^+) K$ holds and denote by $\mu :G\to \mathfrak {a}^+$ the Cartan projection, that is, $\mu (g) \in \mathfrak {a}^+$ is the unique element such that $g\in K\exp (\mu (g))K$ for $g\in G$ . Let $w_0\in K$ be an element of the normalizer of A such that $\operatorname {Ad}_{w_0}\mathfrak a^+= -\mathfrak a^+$ . The opposition involution $\operatorname {i}:\mathfrak a \to \mathfrak a$ is defined by

(2.1) $$ \begin{align} \operatorname{i} (u)= -\operatorname{Ad}_{w_0} (u) \quad\text{for}\ u\in \mathfrak{a}. \end{align} $$

Note that $\mu (g^{-1})=\operatorname {i} (\mu (g))$ for all $g\in G$ .

Let $\Pi $ denote the set of all simple roots for $(\mathfrak g, \mathfrak a^+)$ . Fix a non-empty subset $\theta \subset \Pi $ . Let $ P_\theta ^-$ and $P_\theta ^+$ be a pair of opposite standard parabolic subgroups of G corresponding to $\theta $ ; here, $P_\theta :=P_\theta ^-$ is chosen to contain P. We set

$$ \begin{align*}\mathcal F_\theta^-=G/P_{\theta}^-\quad \text{and} \quad \mathcal F_\theta^{+}=G/P_{\theta}^{+} .\end{align*} $$

We also write $\mathcal F_\theta =\mathcal F_\theta ^-$ for simplicity. We set $P=P_\Pi $ and $\mathcal F=\mathcal F_\Pi $ . Since $P_\theta ^+$ is conjugate to $P_{\operatorname {i}(\theta )}$ , we have $\mathcal F_{\operatorname {i}(\theta )}=\mathcal F_\theta ^+$ . We say $\xi \in \mathcal F_\theta $ and $\eta \in \mathcal F_{\operatorname {i}(\theta )}$ are in general position if $(\xi , \eta )\in G (P_\theta ^-, P_\theta ^+)$ under the diagonal G-action on $\mathcal F_\theta \times \mathcal F_{\operatorname {i}(\theta )}$ . We write

$$ \begin{align*}\mathcal F_\theta^{(2)}= G (P_\theta^-, P_\theta^+),\end{align*} $$

which is the unique open G-orbit in $\mathcal F_\theta \times \mathcal F_{\operatorname {i}(\theta )}$ .

Let $\mathfrak {a}_\theta =\bigcap _{\alpha \in \Pi - \theta } \ker \alpha $ and denote by $\mathfrak a_{\theta }^*$ the space of all linear forms on $\mathfrak a_{\theta }$ . We set $p_\theta :\mathfrak {a}\to \mathfrak {a}_\theta $ as the unique projection invariant under the subgroup of the Weyl group fixing $\mathfrak a_\theta $ pointwise. Set $\mu _{\theta } := p_{\theta } \circ \mu $ .

The $\mathfrak a$ -valued Busemann map $\beta : \mathcal F\times G \times G \to \mathfrak a $ is defined as follows: for $\xi \in \mathcal F$ and $g, h\in G$ ,

$$ \begin{align*} \beta_\xi ( g, h):=\sigma (g^{-1}, \xi)-\sigma(h^{-1}, \xi),\end{align*} $$

where $\sigma (g^{-1},\xi )\in \mathfrak a$ is the unique element such that $g^{-1}k \in K \exp (\sigma (g^{-1}, \xi )) N$ for any $k\in K$ with $\xi =kP$ . For $\xi =kP_\theta \in \mathcal F_\theta $ for $k\in K$ , we define the $\mathfrak a_{\theta }$ -valued Busemann map $\beta ^{\theta } : \mathcal F_{\theta } \times G \times G \to \mathfrak a_{\theta }$ as

(2.2) $$ \begin{align} \beta_{\xi}^\theta (g, h): = p_\theta ( \beta_{kP} (g, h))\in \mathfrak a_\theta;\end{align} $$

this is well defined [Reference Quint19, §6].

In the rest of this section, let $\Gamma < G$ be a Zariski dense $\theta $ -transverse subgroup as in Definition 1.3. For a $(\Gamma , \theta )$ -proper linear form $\psi \in \mathfrak a_\theta ^*$ , we denote by $\delta _\psi \in (0, \infty ]$ the abscissa of convergence of the series $\mathcal P_{\psi }(s) := \sum _{\gamma \in \Gamma }e^{-s \psi (\mu _{\theta }(\gamma ))}$ ; this is well defined [Reference Kim, Oh and Wang13, Lemma 4.2]. We set

$$ \begin{align*}\mathcal D_{\Gamma}^{\theta}:=\{\psi\in \mathfrak a_\theta^*: (\Gamma, \theta)\text{-proper, } \delta_\psi=1, \text{ and } \mathcal P_{\psi}(1)= \infty\}.\end{align*} $$

Note that $\psi \circ \operatorname {i}$ can be regarded as a linear form on $\mathfrak a_{\operatorname {i}(\theta )}$ . Using the property that $\operatorname {i} (\mu (g))=\mu (g^{-1})$ for all $g\in G$ , we deduce that $\mathcal P_{\psi } = \mathcal P_{\psi \circ \operatorname {i}}$ and hence $\psi \in \mathcal D_{\Gamma }^{\theta }$ if and only if ${\psi \circ \operatorname {i} \in \mathcal D_{\Gamma }^{\operatorname {i}(\theta )}}$ .

The $\theta $ -limit set $\Lambda _{\theta }$ of $\Gamma $ is the unique $\Gamma $ -minimal subset of $\mathcal F_{\theta }$ [Reference Benoist1]. We also write

(2.3) $$ \begin{align} \Lambda_\theta^{(2)} := \{(\xi, \eta)\in \mathcal F_\theta^{(2)}:\xi\in \Lambda_\theta,\eta\in \Lambda_{\operatorname{i}(\theta)}\}.\end{align} $$

The following ergodic property of Patterson–Sullivan measures was obtained by Canary, Zhang, and Zimmer [Reference Canary, Zhang and Zimmer4] (see also [Reference Kim, Oh and Wang13, Reference Kim, Oh and Wang14]).

Theorem 2.1. [Reference Canary, Zhang and Zimmer4, Proposition 9.1, Corollary 11.1]

Suppose that $\theta =\operatorname {i}(\theta )$ . Let $\Gamma <G$ be a Zariski dense $\theta $ -transverse subgroup. For any $\psi \in \mathcal D_{\Gamma }^{\theta }$ , there exists a unique $(\Gamma , \psi )$ -Patterson–Sullivan measure $\nu _\psi $ on $\Lambda _{\theta }$ and $\nu _\psi $ is non-atomic. Moreover, the diagonal $\Gamma $ -action on $(\Lambda _{\theta }^{(2)}, (\nu _\psi \times \nu _{\psi \circ \operatorname {i}})|_{\Lambda _{\theta }^{(2)}})$ is ergodic.

3. A property of convergence group actions

In this section, we prove a certain property of convergence group actions which we will need in the proof of our main theorem in the next section. We refer to [Reference Bowditch2] for basic properties of convergence group actions. Let $\Gamma $ be a countable group acting on a compact metrizable space X (with $\# X\ge 3$ ) by homeomorphisms. This action is called a convergence group action if for any sequence of distinct elements $\gamma _n \in \Gamma $ , there exist a subsequence $\gamma _{n_k}$ and $a, b \in X$ such that as $k \to \infty $ , $\gamma _{n_k}(x) $ converges to $ a $ for all $x\in X-\{b\}$ , uniformly on compact subsets. In this case, we say $\Gamma $ acts on X as a convergence group, which we suppose in the following. Any element $\gamma \in \Gamma $ of infinite order fixes precisely one or two points of X, and $\gamma $ is called parabolic or loxodromic accordingly. In that case, there exist $a_{\gamma }, b_{\gamma } \in X$ , fixed by $\gamma $ , such that $\gamma ^n|_{X- \{b_{\gamma }\}} \to a_\gamma $ uniformly on compact subsets as $n \to \infty $ . We have $\gamma $ loxodromic if and only if $a_{\gamma }\ne b_{\gamma }$ , in which case $a_{\gamma }$ and $b_\gamma $ are called the attracting and repelling fixed points of $\gamma $ , respectively.

We will use the following lemma in the next section.

Lemma 3.1. Let $\Gamma $ be a torsion-free countable group acting on a compact metric space X as a convergence group. For any compact subset W of X with at least two points, the subgroup $\Gamma _W = \{\gamma \in \Gamma : \gamma W = W\}$ acts on $X- W$ properly discontinuously, that is, for any $\eta \in X - W$ , there exists an open neighborhood U of $\eta $ such that $\gamma U \cap U \neq \emptyset $ for $\gamma \in \Gamma _W$ implies $\gamma = e$ .

Proof. Suppose not. Then there exist $\eta \in X - W$ , a decreasing sequence of open neighborhoods $U_n$ of $\eta $ in X with $\bigcap _n U_n=\{\eta \}$ , and a sequence $e\ne \gamma _n\in \Gamma $ such that $\gamma _n W=W$ and $\gamma _n U_n \cap U_n \neq \emptyset $ for each $n\in \mathbb N$ . Hence, there exists a sequence $\eta _n \in U_n\cap \gamma _n^{-1}U_n$ ; so $\eta _n \to \eta $ and $\gamma _n \eta _n \to \eta $ as $n \to \infty $ .

We claim that the elements $\gamma _n$ are all pairwise distinct, possibly after passing to a subsequence. Otherwise, it would mean that, after passing to a subsequence, $\gamma _n$ terms are a constant sequence, say $\gamma _n = \gamma \ne e$ . Since $\gamma \eta = \lim _n \gamma _n \eta _n = \eta $ , $\eta $ must be a fixed point of $\gamma $ . Since $\Gamma $ is torsion-free, $\gamma $ is either parabolic or loxodromic, and in particular it has at most two fixed points in X, including $\eta $ . Since $\eta \not \in W$ and W has at least two points, we can take $w \in W$ which is not fixed by $\gamma $ . Then, as $n\to +\infty $ , $\gamma ^n w\to \eta $ or $\gamma ^{-n}w\to \eta $ . Since W is a compact subset such that $\gamma W=W$ and $\eta \notin W$ , this yields a contradiction.

Therefore, we may assume that $\{\gamma _n\}$ is an infinite sequence of distinct elements. Since the action of $\Gamma $ on X is a convergence group action, there exist a subsequence $\gamma _{n_k}$ and $a, b \in X$ such that as $k\to \infty $ , $\gamma _{n_k}(x)$ converges to $ a$ for all $x\in X - \{b\}$ , uniformly on compact subsets. There are two cases to consider. Suppose that $b=\eta $ . Then $W \subset X - \{b\}$ , and hence $\gamma _{n_k}W \to a$ uniformly as $k \to \infty $ . Since $\gamma _{n_k} W = W$ and W is a compact subset, it follows that $W=\{a\}$ , which contradicts the hypothesis that W consists of at least two elements. Now suppose that $b \neq \eta $ . Since $\eta _{n_k}$ converges to $ \eta $ , we may assume that $\eta _{n_k} \neq b$ for all k. Noting that $\# W \ge 2$ , we can take $w_0 \in W - \{b\}$ . If we now consider the following compact subset:

$$ \begin{align*}W_0 := \{\eta_{n_k} : k \in \mathbb N \} \cup \{\eta, w_0\} \subset X - \{b\},\end{align*} $$

we then have $\gamma _{n_k} W_0 \to a$ uniformly as $k \to \infty $ . Since $\eta _{n_k} \in W_0$ for each k and ${\gamma _{n_k}\eta _{n_k} \to \eta} $ as $k \to \infty $ , we must have

$$ \begin{align*}a = \eta.\end{align*} $$

However, since $w_0 \in W_0\cap W$ , $\gamma _{n_k} w_0 \to \eta $ as $k \to \infty $ . This implies $\eta \in W$ since W is compact and $\gamma _{n_k}w_0\in W$ , yielding a contradiction to the hypothesis $\eta \notin W$ . This completes the proof.

We denote by $\Lambda _X$ the set of all accumulation points of a $\Gamma $ -orbit in X. If $\#\Lambda _X\ge 3$ , the $\Gamma $ -action is called non-elementary and $\Lambda _X$ is the unique $\Gamma $ -minimal subset [Reference Bowditch2].

A well-known example of a convergence group action is given by a word hyperbolic group $\Gamma $ . Fix a finite symmetric generating subset $S_{\Gamma }$ of $\Gamma $ . A geodesic ray in $\Gamma $ is an infinite sequence $(\gamma _i)_{i=0}^{\infty }$ of elements of $\Gamma $ such that $\gamma _i^{-1}\gamma _{i+1}\in S_{\Gamma }$ for all $i\ge 0$ . The Gromov boundary $\partial \Gamma $ is the set of equivalence classes of geodesic rays, where two rays are equivalent to each other if and only if their Hausdorff distance is finite. The group $\Gamma $ acts on $\partial \Gamma $ by $\gamma \cdot [(\gamma _i)]=[(\gamma \gamma _i)]$ . This action is known to be a convergence group action [Reference Bowditch2, Lemma 1.11].

Another important example of a convergence group action is the action of a $\theta $ -transverse subgroup $\Gamma $ on $\Lambda _{\theta \cup \operatorname {i}(\theta )}$ .

Proposition 3.2. [Reference Kapovich, Leeb and Porti10, Theorem 4.21]

For a $\theta $ -transverse subgroup $\Gamma $ , the action of $\Gamma $ on $\Lambda _{\theta \cup \operatorname {i}(\theta )}$ is a convergence group action.

4. Non-concentration property

We fix a non-empty subset $\theta \subset \Pi $ . We first prove the following proposition from which we will deduce Theorem 1.4.

Proposition 4.1. Let $\Gamma < G$ be a torsion-free Zariski dense discrete subgroup admitting a convergence group action on a compact metrizable space X. We assume that this action is $\theta $ -antipodal in the sense that there exist $\Gamma $ -equivariant homeomorphisms $f_{\theta } : \Lambda _X \to \Lambda _{\theta }$ and $f_{\operatorname {i}(\theta )} : \Lambda _X \to \Lambda _{\operatorname {i}(\theta )}$ such that for any $\xi \ne \eta $ in $\Lambda _X$ ,

$$ \begin{align*}(f_{\theta}(\xi), f_{\operatorname{i}(\theta)}(\eta))\in \Lambda_\theta^{(2)}.\end{align*} $$

Let $\nu $ be a $\Gamma $ -quasi-invariant measure on $\Lambda _{\theta }$ such that:

  1. (1) $\nu $ is non-atomic;

  2. (2) $\Gamma $ acts ergodically on $(\Lambda _{\theta }^{(2)}, (\nu \times \nu _{\operatorname {i}})|_{\Lambda _\theta ^{(2)}})$ for some $\Gamma $ -quasi-invariant measure $\nu _{\operatorname {i}}$ on $\Lambda _{\operatorname {i}(\theta )}$ .

Then, for any proper algebraic subset S of $\mathcal F_{\theta }$ , we have

$$ \begin{align*}\nu(S) = 0.\end{align*} $$

Proof. We first claim that the $\Gamma $ -action on $(\Lambda _\theta \times \Lambda _{\operatorname {i}(\theta )}, \nu \times \nu _{{\operatorname {i}}})$ is ergodic. Set $R:=(\Lambda _\theta \times \Lambda _{\operatorname {i}(\theta )} ) - \Lambda _\theta ^{(2)}$ . Since the $\Gamma $ -action on $(\Lambda _{\theta }^{(2)}, (\nu \times \nu _{\operatorname {i}})|_{\Lambda _\theta ^{(2)}})$ is ergodic, it suffices to show that

$$ \begin{align*}(\nu\times\nu_{{\operatorname{i}}})(R)=0.\end{align*} $$

For $y\in \Lambda _{\operatorname {i}(\theta )}$ , let $R(y):=\{x\in \Lambda _{\theta } : (x,y) \in R\}$ . By the antipodal property of the pair $(f_\theta , f_{\operatorname {i}(\theta )})$ , we have that for each $y\in \Lambda _{\operatorname {i}(\theta )}$ , we have $R(y)=\emptyset $ or $R(y)= \{(f_{\theta } \circ f_{\operatorname {i}(\theta )}^{-1})(y)\}$ and hence $\nu ( R(y))=0$ by the non-atomicity of $\nu $ .

Therefore,

(4.1) $$ \begin{align} (\nu\times\nu_{{\operatorname{i}}})(R)=\int_{y\in \Lambda_{\operatorname{i}(\theta)}}\nu (R(y))\,d\nu_{\operatorname{i}}(y)=0, \end{align} $$

proving the claim.

Now suppose that $\nu (S)> 0$ for some proper algebraic subset $S \subset \mathcal F_{\theta }$ . We may assume that S is irreducible and of minimal dimension among all such algebraic subsets of $\mathcal F_\theta $ . Let $W = f_{\theta }^{-1}(S\cap \Lambda _\theta ) \subset \Lambda _X$ . Since $\nu $ is non-atomic and $\nu (S)> 0$ , we have $\# W = \infty> 2$ . This implies $\#\Lambda _X\ge 3$ . By the property of a non-elementary convergence group action, $\Lambda _X$ is the unique $\Gamma $ -minimal subset of X and there always exists a loxodromic element of $\Gamma $ [Reference Bowditch2].

Since $\Gamma < G$ is Zariski dense, $\Lambda _\theta $ is Zariski dense in $\mathcal F_\theta $ as well, and hence $\Lambda _{\theta } \not \subset S$ . Therefore, $X - W$ is a non-empty open subset intersecting $\Lambda _X$ . Since $\Gamma $ acts minimally on $\Lambda _X$ and the set of attracting fixed points of loxodromic elements of $\Gamma $ is a non-empty $\Gamma $ -invariant subset, there exists a loxodromic element $\gamma _0 \in \Gamma $ whose attracting fixed point $a_{\gamma _0}$ is contained in $\Lambda _X- W$ . Hence, applying Lemma 3.1 to $\eta =a_{\gamma _0}$ , we have an open neighborhood $U $ of $a_{\gamma _0}$ in $\Lambda _X$ such that

(4.2) $$ \begin{align} \gamma U \cap U = \emptyset \end{align} $$

for all non-trivial $\gamma \in \Gamma $ with $\gamma W = W$ .

Since $\gamma _0^m|_{\Lambda _X - \{b_{\gamma _0} \}} \to a_{\gamma _0}$ uniformly on compact subsets as $m \to +\infty $ and $\#\Lambda _X\ge 3$ , U contains a point $\xi \in \Lambda _X-\{a_{\gamma _0}, b_{\gamma _0}\}$ . By replacing $\gamma _0$ by a large power $\gamma _0^m$ if necessary, we can find an open neighborhood V of $\xi $ contained in $U-\{a_{\gamma _0}\}$ such that $\gamma _0 V \subset U$ and $\gamma _0 V \cap V = \emptyset $ .

We now consider the subset

$$ \begin{align*}S \times f_{\operatorname{i}(\theta)}(V)\end{align*} $$

of $\mathcal F_{\theta } \times \mathcal F_{\operatorname {i}(\theta )}$ . Since $\nu (S)> 0$ and $\nu _{\operatorname {i}}(f_{\operatorname {i}(\theta )}(V))> 0$ , we have that $\Gamma (S \times f_{\operatorname {i}(\theta )}(V))$ has full $\nu \times \nu _{\operatorname {i}}$ -measure by the ergodicity of the $\Gamma $ -action on $(\Lambda _{\theta } \times \Lambda _{\operatorname {i}(\theta )}, \nu \times \nu _{\operatorname {i}})$ . Since $(\nu \times \nu _{\operatorname {i}})(S \times \gamma _0 f_{\operatorname {i}(\theta )}(V))> 0$ , there exists $\gamma \in \Gamma $ such that

$$ \begin{align*}(\nu \times \nu_{\operatorname{i}})( (S \times \gamma_0 f_{\operatorname{i}(\theta)}(V)) \cap (\gamma S \times \gamma f_{\operatorname{i}(\theta)}(V)))> 0.\end{align*} $$

In particular, we have

$$ \begin{align*}\nu(S \cap \gamma S)> 0 \quad \text{and} \quad \nu_{\operatorname{i}}(\gamma_0 f_{\operatorname{i}(\theta)}(V) \cap \gamma f_{\operatorname{i}(\theta)}(V)) > 0.\end{align*} $$

Since S was chosen to be of minimal dimension and irreducible among proper algebraic sets with positive $\nu $ -measure, we must have $S = \gamma S$ . It follows from the $\Gamma $ -invariance of $\Lambda _{\theta }$ that $W = \gamma W$ .

The $\Gamma $ -equivariance of $f_{\operatorname {i}(\theta )}$ implies that

(4.3) $$ \begin{align} \nu_{\operatorname{i}}(f_{\operatorname{i}(\theta)}(\gamma_0 V \cap \gamma V))> 0. \end{align} $$

Since $\gamma _0 V \cap V = \emptyset $ , we have $\gamma \neq e$ . Hence, it follows from $V \subset U$ , $\gamma _0 V \subset U$ , and the choice in equation (4.2) of U that

$$ \begin{align*}\gamma_0 V \cap \gamma V \subset U \cap \gamma U = \emptyset,\end{align*} $$

which gives a contradiction to equation (4.3). This finishes the proof.

4.1. Proof of Theorem 1.4

Let $\Gamma < G$ be a Zariski dense $\theta $ -transverse subgroup and $\nu $ a $(\Gamma , \psi )$ -Patterson–Sullivan measure for a $(\Gamma , \theta )$ -proper linear form $\psi \in \mathfrak a_{\theta }^*$ such that $\sum _{\gamma \in \Gamma } e^{-\psi (\mu _{\theta }(\gamma ))} = \infty $ . We may assume without loss of generality that $\Gamma $ is torsion-free. Indeed, let $\Gamma _0 < \Gamma $ be a torsion-free subgroup of finite index. Then $\Gamma _0$ is also a Zariski dense $\theta $ -transverse subgroup of G. Moreover, $\nu $ is a $(\Gamma _0, \psi )$ -Patterson–Sullivan measure since the limit sets for $\Gamma $ and $\Gamma _0$ are the same. Write $\Gamma = \bigcup _{i = 1}^n \gamma _i \Gamma _0$ for some $\gamma _1, \ldots , \gamma _n \in \Gamma $ . By [Reference Benoist1, Lemma 4.6], there exists $C> 0$ such that $\| \mu (\gamma _i \gamma ) - \mu (\gamma )\| \le C$ for all $\gamma \in \Gamma _0$ and $i = 1, \ldots , n$ . Hence, we have that $\psi $ is $(\Gamma _0, \theta )$ -proper as well and

$$ \begin{align*}\infty = \sum_{\gamma \in \Gamma} e^{-\psi(\mu_{\theta}(\gamma))} = \sum_{i = 1}^n \sum_{\gamma \in \Gamma_0} e^{-\psi(\mu_{\theta}(\gamma_i \gamma))} \le n e^{\|\psi\| C} \sum_{\gamma \in \Gamma_0} e^{-\psi(\mu_{\theta}(\gamma))},\end{align*} $$

where $\|\psi \|$ denotes the operator norm of $\psi $ . In particular, $\sum _{\gamma \in \Gamma _0} e^{-\psi (\mu _{\theta }(\gamma ))} = \infty $ . Therefore, replacing $\Gamma $ by $\Gamma _0$ , we assume that $\Gamma $ is torsion-free. By Proposition 3.2, the action of $\Gamma $ on $\Lambda _{\theta \cup \operatorname {i}(\theta )}$ is a convergence group action.

Since there exists a $(\Gamma , \psi )$ -conformal measure, we have $\delta _{\psi } \le 1$ by [Reference Kim, Oh and Wang13, Lemma 7.3]. Therefore, the hypothesis $\sum _{\gamma \in \Gamma } e^{-\psi (\mu _{\theta }(\gamma ))} = \infty $ implies that $\psi \in \mathcal D_{\Gamma }^{\theta }$ . Moreover, the $\theta $ -antipodality of $\Gamma $ implies that the canonical projections

$$ \begin{align*}f_{\theta} : \Lambda_{\theta \cup \operatorname{i}(\theta)} \to \Lambda_{\theta} \quad \text{and} \quad f_{\operatorname{i}(\theta)} : \Lambda_{\theta \cup \operatorname{i}(\theta)} \to \Lambda_{\operatorname{i}(\theta)}\end{align*} $$

are $\Gamma $ -equivariant $\theta $ -antipodal homeomorphisms [Reference Kim, Oh and Wang13, Lemma 9.5]. This implies that Theorem 2.1 indeed holds for a general $\theta $ without the hypothesis $\theta =\operatorname {i}(\theta )$ . Hence, $\nu =\nu _\psi $ , $\nu _\psi $ is non-atomic, and the diagonal $\Gamma $ -action on $(\Lambda _{\theta }^{(2)}, (\nu _\psi \times \nu _{\psi \circ \operatorname {i}})|_{\Lambda _{\theta }^{(2)}})$ is ergodic. Since $\nu _{\psi \circ \operatorname {i}}$ is $\Gamma $ -conformal, it is $\Gamma $ -quasi-invariant. Therefore, Theorem 1.4 follows from Proposition 4.1.

We emphasize again that Lemma 3.1 and Proposition 4.1 were introduced to deal with the case when $\operatorname {i}$ is non-trivial. Indeed, when $\operatorname {i}$ is trivial, Theorem 1.4 follows from the following $\theta $ -version of [Reference Edwards, Lee and Oh5, Theorem 9.3].

Theorem 4.2. Let $\Gamma < G$ be a Zariski dense discrete subgroup. Let $\nu $ be a $\Gamma $ -quasi-invariant measure on $\Lambda _{\theta }$ . Suppose that the diagonal $\Gamma $ -action on $(\Lambda _{\theta } \times \Lambda _{\theta }, {\nu \times \nu })$ is ergodic. Then, for any proper algebraic subset S of $\mathcal F_{\theta }$ , we have

$$ \begin{align*}\nu(S) = 0.\end{align*} $$

Proof. The proof is identical to the proof of [Reference Edwards, Lee and Oh5, Theorem 9.3] except that we work with a general $\theta $ . We reproduce it here for the convenience of the readers. Let S be a proper irreducible subvariety of $\mathcal F_\theta $ with $\nu (S)>0$ and of minimal dimension. Since ${(\nu \times \nu ) (S\times S)>0}$ , the $\Gamma $ -ergodicity of $\nu \times \nu $ implies that $(\nu \times \nu )(\Gamma (S\times S))=1$ . It follows that for any $\gamma _0\in \Gamma $ , there exists $\gamma \in \Gamma $ such that $(S \times \gamma _0S)\cap (\gamma S\times \gamma S)$ has positive $\nu \times \nu $ -measure; hence, $\nu (S\cap \gamma S)>0$ and $\nu (\gamma _0 S\cap \gamma S)>0$ . Since S is irreducible and of minimal dimension, it follows that $S=\gamma S=\gamma _0 S$ . Since $\gamma _0\in \Gamma $ was arbitrary, we have $\Gamma S=S$ , which contradicts the Zariski density hypothesis on $\Gamma $ .

We finally mention that the proof of Proposition 4.1 implies the following when the second measure cannot be taken to be the same as the first measure.

Theorem 4.3. Let $\Gamma < G$ be a Zariski dense torsion-free discrete subgroup acting on $\Lambda _\theta $ as a convergence group. Let $\nu $ be a non-atomic $\Gamma $ -quasi-invariant measure on $\Lambda _{\theta }$ . Suppose that the diagonal $\Gamma $ -action on $(\Lambda _{\theta } \times \Lambda _{\theta }, \nu \times \nu ')$ is ergodic for some $\Gamma $ -quasi-invariant measure $\nu '$ on $\Lambda _\theta $ . Then, for any proper algebraic subset S of $\mathcal F_{\theta }$ , we have

$$ \begin{align*}\nu(S) = 0.\end{align*} $$

Proof. Since $\Gamma $ acts ergodically on the entire product space $(\Lambda _{\theta } \times \Lambda _{\theta }, \nu \times \nu ')$ , the first part of the proof of Proposition 4.1 is not relevant. Suppose that S is an irreducible proper subvariety of $\mathcal F_\theta $ and of minimal dimension among all subvarieties with positive $\nu $ -measure. Then, setting $W = S \cap \Lambda _{\theta }$ , as in the proof of Proposition 4.1, we can find non-empty open subsets $V\subset U\subset \Lambda _\theta - W$ such that $ \gamma U \cap U = \emptyset $ for all non-trivial $\gamma \in \Gamma $ with $\gamma W = W$ , and $\gamma _0V\subset U$ and $\gamma _0V\cap V=\emptyset $ for some $\gamma _0\in \Gamma $ . Using $(\nu \times \nu ')(S\times V)>0$ , we then get a contradiction by the same argument as in Proposition 4.1.

Acknowledgements

We would like to thank Subhadip Dey for helpful conversations. H.O. is partially supported by the NSF grant No. DMS-1900101.

References

Benoist, Y.. Propriétés asymptotiques des groupes linéaires. Geom. Funct. Anal. 7(1) (1997), 147.CrossRefGoogle Scholar
Bowditch, B.. Convergence groups and configuration spaces. Geometric Group Theory Down Under (Canberra, 1996). Ed. J. Cossey, C. F. Miller, W. D. Neumann and M. Shapiro. de Gruyter, Berlin, 1999, pp. 2354.Google Scholar
Canary, R., Zhang, T. and Zimmer, A.. Patterson–Sullivan measures for relatively Anosov groups. Preprint, 2023, arXiv:2308.04023.Google Scholar
Canary, R., Zhang, T. and Zimmer, A.. Patterson–Sullivan measures for transverse subgroups. J. Mod. Dyn. 20 (2024), 319377.CrossRefGoogle Scholar
Edwards, S., Lee, M. and Oh, H.. Torus counting and self-joinings of Kleinian groups. J. Reine Angew. Math. 807 (2024), 151185.Google Scholar
Flaminio, L. and Spatzier, R.. Geometrically finite groups, Patterson–Sullivan measures and Ratner’s rigidity theorem. Invent. Math. 99(3) (1990), 601626.CrossRefGoogle Scholar
Guéritaud, F., Guichard, O., Kassel, F. and Wienhard, A.. Anosov representations and proper actions. Geom. Topol. 21(1) (2017), 485584.CrossRefGoogle Scholar
Guichard, O. and Wienhard, A.. Anosov representations: domains of discontinuity and applications. Invent. Math. 190(2) (2012), 357438.CrossRefGoogle Scholar
Kapovich, M. and Leeb, B.. Discrete isometry groups of symmetric spaces. Handbook of Group Actions. Volume IV (Advanced Lectures in Mathematics (ALM), 41). Ed. L. Ji, A. Papadopoulos and S.-T. Yau. International Press, Somerville, MA, 2018, pp. 191290.Google Scholar
Kapovich, M., Leeb, B. and Porti, J.. Anosov subgroups: dynamical and geometric characterizations. Eur. J. Math. 3(4) (2017), 808898.CrossRefGoogle Scholar
Kapovich, M., Leeb, B. and Porti, J.. A Morse lemma for quasigeodesics in symmetric spaces and Euclidean buildings. Geom. Topol. 22(7) (2018), 38273923.CrossRefGoogle Scholar
Kassel, F.. Geometric structures and representations of discrete groups. Proceedings of the International Congress of Mathematicians—Rio de Janeiro 2018. Volume II. Invited Lectures. Ed. B. Sirakov, P. Souza and M. Viana. World Science Publishing, Hackensack, NJ, 2018, pp. 11151151.Google Scholar
Kim, D. M., Oh, H. and Wang, Y.. Properly discontinuous actions, growth indicators and conformal measures for transverse subgroups. Preprint, 2023, arXiv:2306.06846.Google Scholar
Kim, D. M., Oh, H. and Wang, Y.. Ergodic dichotomy for subspace flows in higher rank. Preprint, 2023, arXiv:2310.19976.Google Scholar
Labourie, F.. Anosov flows, surface groups and curves in projective space. Invent. Math. 165(1) (2006), 51114.CrossRefGoogle Scholar
Lee, M. and Oh, H.. Invariant measures for horospherical actions and Anosov groups. Int. Math. Res. Not. IMRN 19 (2023), 1622616295.CrossRefGoogle Scholar
Patterson, S.. The limit set of a Fuchsian group. Acta Math. 136(3–4) (1976), 241273.CrossRefGoogle Scholar
Quint, J.-F.. Divergence exponentielle des sous-groupes discrets en rang supérieur. Comment. Math. Helv. 77(3) (2002), 563608.CrossRefGoogle Scholar
Quint, J.-F.. Mesures de Patterson–Sullivan en rang supérieur. Geom. Funct. Anal. 12(4) (2002), 776809.CrossRefGoogle Scholar
Sambarino, A.. A report on an ergodic dichotomy. Ergod. Th. & Dynam. Sys. 44(1) (2024), 236289.CrossRefGoogle Scholar
Selberg, A.. On discontinuous groups in higher-dimensional symmetric spaces. Contributions to Function Theory (Internat. Colloq. Function Theory, Bombay, 1960). Tata Institute of Fundamental Research, Bombay, 1960, pp. 147164.Google Scholar
Sullivan, D.. The density at infinity of a discrete group of hyperbolic motions. Publ. Math. Inst. Hautes Études Sci. 50 (1979), 171202.CrossRefGoogle Scholar
Wienhard, A.. An invitation to higher Teichmüller theory. Proceedings of the International Congress of Mathematicians—Rio de Janeiro 2018. Volume II. Invited Lectures. Ed. B. Sirakov, P. Souza and M. Viana. World Science Publishing, Hackensack, NJ, 2018, pp. 10131039.Google Scholar