Home Zariski-dense non-tempered subgroups in higher rank of nearly optimal growth
Article Open Access

Zariski-dense non-tempered subgroups in higher rank of nearly optimal growth

  • Mikołaj Frączyk and Hee Oh ORCID logo EMAIL logo
Published/Copyright: July 8, 2025

Abstract

We construct the first example of a Zariski-dense, discrete, non-lattice subgroup Γ 0 of a higher rank simple Lie group 𝐺, which is non-tempered in the sense that the quasi-regular representation L 2 ( Γ 0 \ G ) is non-tempered. More precisely, let n 3 and let Γ be the fundamental group of a closed hyperbolic 𝑛-manifold that contains a properly embedded totally geodesic hyperplane. We show that there exists a non-empty open subset 𝒪 of Hom ( Γ , SO ( n , 2 ) ) such that, for any σ O , the subgroup σ ( Γ ) is a Zariski-dense and non-tempered Anosov subgroup of SO ( n , 2 ) . In addition, the growth indicator of σ ( Γ ) is nearly optimal: it almost realizes the supremum of growth indicators among all non-lattice discrete subgroups, a bound imposed by property (T) of SO ( n , 2 ) .

1 Introduction

Let 𝐺 be a connected semisimple real algebraic group. Let Γ < G be a discrete subgroup of 𝐺. Denote by d x a 𝐺-invariant measure on the homogeneous space Γ \ G . Consider the Hilbert space L 2 ( Γ \ G ) = L 2 ( Γ \ G , d x ) . The right translation action of 𝐺 on Γ \ G induces a unitary representation of 𝐺 on L 2 ( Γ \ G ) , called the quasi-regular representation.

A unitary representation ( π , H ) of 𝐺 is called tempered if it is weakly contained in the (right) regular representation L 2 ( G ) , i.e., any diagonal matrix coefficients of ( π , H ) can be approximated by a convex linear combination of diagonal matrix coefficients of L 2 ( G ) , uniformly on compact subsets of 𝐺. This notion, due to Harish-Chandra, plays a central role in harmonic analysis on semisimple groups.

Definition 1.1

We call a discrete subgroup Γtempered in 𝐺 if its quasi-regular representation L 2 ( Γ \ G ) is tempered.

Temperedness of Γ is equivalent to the statement that all matrix coefficients of L 2 ( Γ \ G ) are L 2 + ε -integrable for any ε > 0 (see [10]). If 𝐺 has Kazhdan’s property (T), that is, all simple factors of 𝐺 have rank at least 2 or are isogenous to Sp ( n , 1 ) or F 4 ( 20 ) , then a quantitative form of property (T) implies the existence of p = p G > 0 such that, for any non-lattice discrete subgroup Γ < G , all matrix coefficients of L 2 ( Γ \ G ) are L p -integrable [9, 31, 26].

In rank-one groups, the situation is quite different, for example, any lattice admits a non-elementary infinite index normal subgroup [11], whereas the Margulis normal subgroup theorem precludes such behavior in higher rank. Moreover, there are also convex cocompact subgroups of SO ( n , 1 ) , n 2 , whose critical exponents can be made arbitrarily close to the volume entropy of the hyperbolic 𝑛-space H n , that is, n 1 (see [29, Section 6]; such examples cannot occur in higher rank because of (1.2)). These high-exponent groups furnish Zariski-dense, non-tempered subgroups by [30, Theorem 1.4] and [8, Theorem 4.2].

For higher rank groups, previously known non-tempered examples were all lattices of a proper algebraic subgroup of 𝐺 (see [7, Example B], [3]). It remained open whether one could find a Zariski-dense, non-lattice, non-tempered subgroup of a higher rank simple group 𝐺. Our main result answers this in the affirmative.

Theorem 1.2

For each n 3 , there exists a Zariski-dense, non-lattice, non-tempered subgroup of SO ( n , 2 ) .

Remark 1.3

For a geometrically finite discrete subgroup Γ < SO ( n , 1 ) , the hyperbolic manifold Γ \ H n possesses a square-integrable base eigenfunction of the Laplacian if and only if Γ is non-tempered [32, 38, 37]. By contrast, a recent result of [15] shows that, for any non-lattice discrete subgroup Γ of a higher rank simple algebraic group 𝐺, the base eigenfunction on the corresponding locally symmetric manifold is never square-integrable. Hence the appearance of a non-tempered subgroup in Theorem 1.2 underscores another sharp distinction in the behavior of infinite-volume locally symmetric manifolds between the higher rank and rank-one cases.

Temperedness of Γ can be characterized in terms of its growth indicator ψ Γ . Fix a Cartan decomposition G = K exp ( a + ) K , where 𝐾 is a maximal compact subgroup and a + is a positive Weyl chamber of a Cartan subalgebra 𝔞. There exists a unique element μ ( g ) a + for g G such that g K exp μ ( g ) K , called the Cartan projection of 𝑔.

For a discrete subgroup Γ of 𝐺, denote by L Γ a + its limit cone, which is defined as the asymptotic cone of μ ( Γ ) . The growth indicator ψ G , Γ = ψ Γ : a + R { } , introduced by Quint [34], is a higher rank version of the critical exponent. It is outside the limit cone L Γ . For each v L Γ , the value ψ Γ ( v ) encodes the exponential growth rate of Γ in the direction 𝑣,

(1.1) ψ Γ ( v ) = v inf v C lim sup T log # { γ Γ : μ ( γ ) C , μ ( γ ) T } T ,

where the infimum is taken over all open cones C a + containing 𝑣. This definition is independent of the choice of a norm on 𝔞.

Denote by ρ = ρ G the half-sum of all positive roots of ( Lie G , a ) counted with multiplicity. The linear form 2 ρ a gives the exponential volume growth rate of 𝐺: for any v a + ,

2 ρ ( v ) = v inf v C lim sup T log Vol { g G : μ ( g ) C , μ ( γ ) T } T ,

where the infimum is taken over all open cones C a + containing 𝑣. We have ψ Γ 2 ρ on a + for any discrete subgroup Γ < G and equality holds for lattices Γ (see [33]). If 𝐺 has Kazhdan’s property (T), there exists a constant η G > 0 such that, for any non-lattice discrete subgroup Γ of 𝐺, we have

(1.2) ψ Γ ( 2 η G ) ρ on a +

(see [9, Theorem 4.4], [35, Theorem 5.1], and also [27, Theorem 7.1]).

Definition 1.4

A discrete subgroup Γ < G has slow growth if ψ Γ ρ on a + .

The slow growth means, informally, that the number of elements of Γ in a ball of radius 𝑅 in 𝐺 is bounded (up to sub-exponential factors) by a constant times the square root of the ball’s volume as R . It turns out that the slow growth property of Γ determines the temperedness: ψ Γ ρ on a + if and only if Γ is tempered. This was shown in [16] for Borel–Anosov subgroups, and in [28] for general discrete subgroups.

Theorem 1.5, which is a more elaborate version of Theorem 1.2, provides the first Zariski-dense, non-lattice subgroups of higher rank simple Lie groups that do not have slow growth. Moreover, these examples have nearly optimal growth. For n 3 , the identity component of the special orthogonal group SO ( n , 2 ) is a simple Lie group of rank two. As discussed in Section 4, we can identify its positive Weyl chamber a + with

a + = { v = ( v 1 , v 2 , 0 , , 0 , v 2 , v 1 ) R n + 2 : v 1 v 2 0 } .

The set of simple roots of SO ( n , 2 ) is given by α 1 ( v ) = v 1 v 2 and α 2 ( v ) = v 2 , and 𝜌 is the following:

ρ ( v ) = 1 2 ( n v 1 + ( n 2 ) v 2 )

for any v = ( v 1 , v 2 , 0 , , 0 , v 2 , v 1 ) a + .

Theorem 1.5

Let n 3 and let Γ be the fundamental group of a closed hyperbolic 𝑛-manifold with a properly embedded totally geodesic hyperplane. For any ε > 0 , there exists a non-empty open subset O = O ( ε ) of Hom ( Γ , SO ( n , 2 ) ) such that, for any σ O , the following hold:

  1. σ ( Γ ) is a Zariski-dense, { α 1 } -Anosov[1], and non-tempered subgroup of SO ( n , 2 ) without slow growth;

  2. for all v a + , we have

    ψ σ ( Γ ) ( v ) ( 2 ( n 1 ) n + ε ) ρ ( v ) ;

  3. there exists a unit vector v σ a + such that

    (1.3) ψ σ ( Γ ) ( v σ ) ( 2 ( n 1 ) n ε ) ρ ( v σ ) .

Moreover, σ ( Γ ) has nearly optimal growth in the sense that

(1.4) ψ σ ( Γ ) ( v σ ) sup Λ ψ Λ ( v σ ) ε ,

where the supremum is taken over all non-lattice discrete subgroups Λ < SO ( n , 2 ) .

We have an upper bound on the growth of arbitrary non-lattice discrete subgroups coming from the effective property (T) of 𝐺 (see [31] and Proposition 4.1). Inequalities (1.3) and (1.4) show that our examples almost saturate this bound. At least inside SO ( n , 2 ) , this means that one cannot hope to improve existing growth-gap theorems (e.g. [27]) by merely imposing Zariski-density. It remains an intriguing question whether such an improvement is possible in other higher rank groups, for example in SL n ( R ) , n 3 .

Remark 1.6

There are many examples of Zariski-dense discrete subgroups in higher rank that are tempered, for instance, the image of any Hitchin representation of a surface group into a real split simple algebraic group of higher rank [16, 12].

Our construction of a non-tempered Zariski-dense subgroup of SO ( n , 2 ) goes as follows. We begin with a uniform lattice Γ in SO ( n , 1 ) that decomposes as an amalgamated product of two subgroups over a uniform lattice in SO ( n 1 , 1 ) . For n 3 , any lattice of SO ( n , 1 ) is non-tempered, when viewed inside SO ( n , 2 ) (Corollary 4.6). The inclusion id Γ : Γ SO ( n , 2 ) can be deformed via the bending construction [22, 24], yielding a discrete Zariski-dense subgroup Γ 1 of SO ( n , 2 ) . The heart of the paper is to show that Γ 1 is non-tempered. We present two proofs. In the first, we consider the Chabauty topology on the space of closed subgroups of SO ( n , 2 ) and show that the property of being non-tempered is open, by studying the convergence of the matrix coefficients of quasi-regular representations[2]. As a consequence, all sufficiently small (discrete) deformations of SO ( n , 1 ) remain non-tempered, so Γ 1 satisfies Theorem 1.2. For the second proof, we track how the growth indicator of Γ evolves under the deformation, using the property that Γ is an Anosov subgroup. The limit cone of the deformation is known to vary continuously in this setting [24] (see also [13]) and a certain critical exponent of Γ 1 varies continuously as well [6]. Hence, for small deformations, the growth indicator of Γ 1 can be controlled by the growth indicator of Γ and hence it is not smaller than the half-sum of positive roots 𝜌, proving Theorem 1.5.

2 Convergence of matrix coefficients and Chabauty topology

Let 𝐺 be a locally compact second countable group. Let C = C G denote the space of all closed subgroups of 𝐺 equipped with the Chabauty topology, that is, a sequence of closed subgroups H n converges to 𝐻 as n if, for any element h H , there exists a sequence h n H n with h n h and the limit points of any sequence g n H n belong to 𝐻. The space ℭ is a compact space. When a sequence H i converges to a closed subgroup 𝐻, we say that 𝐻 is the Chabauty limit of H i . Note that the Chabauty limit of a sequence of discrete subgroups is not necessarily a discrete subgroup.

For a unimodular closed subgroup 𝐻 of 𝐺, denote by ν H a Haar measure on 𝐻. For s C c ( G ) and any locally finite measure 𝜈 on 𝐻, we write

ν ( s ) : = H s ( h ) d ν ( h ) .

Note that, for a non-negative function s C c ( G ) with ν H ( s ) 0 , the normalized measure ν H ( s ) 1 ν H is independent of the choice of a Haar measure ν H . Let M ( G ) be the space of all locally finite Borel measures on 𝐺, equipped with the weak-* topology. Throughout the paper, 𝑒 denotes the identity element of a relevant group.

Proposition 2.1

Let Γ n be a sequence of discrete subgroups of 𝐺 converging to a closed subgroup 𝐻 in the Chabauty topology. Then 𝐻 is unimodular, and for any non-negative function s C c ( G ) with s ( e ) > 0 , we have

(2.1) lim n ν Γ n ( s ) 1 ν Γ n = ν H ( s ) 1 ν H in M ( G ) .

Proof

Consider a non-negative function s C c ( G ) with s ( e ) > 0 . For simplicity, set ν n = ν Γ n and ν n : = ν n ( s ) 1 ν n . Then ν n ( s ) = 1 .

First we show that the sequence ν n is relatively compact in M ( G ) . Since s ( e ) > 0 , it follows from the continuity of 𝑠 that there exists a symmetric neighborhood 𝑈 of 𝑒 such that κ : = inf g U 2 s ( g ) > 0 . Fix any compact subset 𝐶 of 𝐺. Let

m C : = max { # F : F C , g 1 U g 2 U = for all g 1 g 2 F } .

Note that

m C ν G ( C U ) ν G ( U ) .

For any n N , choose a maximal subset F n Γ n C such that g 1 U g 2 U = for all g 1 g 2 F n . Then Γ n C F n U 2 , so

ν n ( C ) # F n ν n ( U 2 ) m C κ s ( g ) d ν n ( g ) .

Therefore, for all n N , we have

ν n ( C ) m C κ .

Since 𝐶 is an arbitrary compact subset of 𝐺, it follows that the sequence ν n , n N , forms a relatively compact subset of M ( G ) .

Let ν M ( G ) be a weak-* limit of the sequence ν n . By construction, 𝜈 is a locally finite measure supported on 𝐻 and ν ( s ) = 1 . It remains to show that 𝜈 is a Haar measure on 𝐻. Let φ C c ( G ) and h H . Let γ n Γ n be a sequence with lim n γ n = h . Then, since ν n is a Haar measure of Γ n , we get

| φ ( g ) φ ( h g ) d ν ( g ) | | φ ( g ) d ν ( g ) φ ( g ) d ν n ( g ) | + | φ ( γ n g ) d ν n ( g ) φ ( h g ) d ν n ( g ) | + | φ ( h g ) d ν n ( g ) φ ( h g ) d ν ( g ) | .

The first and the third term converge to zero since ν n weakly converges to 𝜈. The middle term goes to zero because φ ( γ n ) converges uniformly to φ ( ) . Hence, the right-hand side converges to 0 as n , so 𝜈 is indeed left 𝐻-invariant. Similarly, we can show that 𝜈 is also a right 𝐻-invariant. This proves that 𝐻 is unimodular. Since ν ( s ) = 1 , we have ν = ν H ( s ) 1 ν H and thus the desired convergence (2.1) follows from ν n ν . ∎

Remark 2.2

The normalization of measures by the integral of 𝑠 is necessary in the above proposition. For example, if G = SL 2 ( F p ( ( t ) ) ) and

Γ n : = { ( 1 f ( t ) 0 1 ) : f ( t ) = a n t n + a n + 1 t n + 1 + + a 2 n t 2 n F p [ t ] } ,

then, as n , Γ n converges to the trivial subgroup { e } in the Chabauty topology, but the sequence ν Γ n of counting measures on Γ n fails to converge on the account of mass near identity blowing up to infinity.

On the other hand, we can skip the normalization if the group 𝐺 has the no-small-subgroup property. We say that a locally compact group 𝐺 has no small subgroup if there exists a neighborhood of 𝑒 in 𝐺 which does not contain any non-trivial subgroup of 𝐺; this notion was first introduced in [19]. It is a well-known fact that a real Lie group 𝐺 has no small subgroup; this can be easily seen, using the fact that the exponential map is a diffeomorphism of a neighborhood of 0 in 𝔤 onto a neighborhood of the 𝑒 in 𝐺.

Proposition 2.3

Suppose that 𝐺 has the no-small-subgroup property (e.g., a real Lie group). Let Γ n be a sequence of discrete subgroups of 𝐺 which converges to a discrete subgroup Γ in the Chabauty topology. Then, as n ,

lim n γ Γ n δ γ = γ Γ δ γ in M ( G ) ,

where δ γ denotes the Dirac measure at { γ } .

Proof

Let ν n : = γ Γ n δ γ and ν : = γ Γ δ γ . Let φ C c ( G ) . We need to show that

lim n φ d ν n = φ d ν .

Let ε > 0 be arbitrary. Fix a compact subset C G and φ C c ( G ) supported on 𝐶. Enlarging 𝐶 if needed, we may assume that Γ C = . By the hypothesis that 𝐺 has the no-small-subgroup property, there is an open neighborhood U G of the identity 𝑒 which contains no non-trivial subgroup of 𝐺. We choose an open symmetric neighborhood U 1 G of 𝑒 such that

  1. U 1 2 U ;

  2. γ U 1 5 C for all γ Γ C ;

  3. the collection γ U 1 5 , γ Γ C , are pairwise disjoint;

  4. for all γ C Γ and u U 1 ,

    | φ ( γ ) φ ( γ u ) | ε # ( Γ C ) .

Consider the following compact subset:

C 1 : = C \ γ Γ C γ U 1 .

Note that Γ C 1 = . Since the sequence Γ n converges to Γ in the Chabauty topology, we have Γ n C 1 = for all 𝑛 large enough. For each fixed γ Γ C , there exists n 0 = n 0 ( γ ) 1 such that Γ n γ U 1 and Γ n C 1 = for all n n 0 . Since Γ C is finite, we have n 0 : = max { n 0 ( γ ) : γ Γ C } < .

On the other hand, we claim that, for any γ C Γ and n 1 , # ( Γ n γ U 1 ) 1 . Indeed, suppose there exists some element γ n Γ n γ U 1 . Then

γ n 1 ( Γ n γ U 1 ) = Γ n ( γ n 1 γ U 1 ) Γ n U 1 2 .

By the no-small-subgroup property of 𝐺, we have either Γ n U 1 2 = { e } or there is some element γ n Γ n ( U 1 4 \ U 1 2 ) ; otherwise, Γ n U 1 2 would be a non-trivial subgroup. In the second case, we would have

γ n γ n γ n ( U 1 4 \ U 1 2 ) γ U 1 5 \ γ U 1 C \ γ U 1 .

Using property (3), we get γ n γ n C 1 , contradicting the fact that Γ n C 1 = . Therefore, we must have Γ n U 1 2 = { e } . This implies that Γ n γ U 1 = { γ n } , proving the claim.

Therefore, for all γ Γ C and n n 0 , we have a unique element γ n Γ n such that Γ n γ U 1 = { γ n } , and γ n γ as n . Since

φ d ν n = γ Γ C φ ( γ n ) for all n n 0 ,

we get from (4) that, for all n n 0 ,

| φ d ν φ d ν n | γ Γ C | φ ( γ ) φ ( γ n ) | ε .

This finishes the proof. ∎

Let 𝐺 be unimodular and d g a Haar measure on 𝐺. For a closed unimodular subgroup 𝐻 of 𝐺, there exists a unique 𝐺-invariant measure d H \ G on H \ G such that, for all ψ C c ( G ) ,

G ψ d g = H \ G H ψ ( h g ) d ν H ( h ) d H \ G ( H g ) .

We then have a unitary representation of 𝐺 on the Hilbert space

L 2 ( H \ G ) = { f : H \ G R : H \ G | f | 2 d H \ G < }

by right translations: g . f ( H g ) : = f ( H g g ) for g , g G and f L 2 ( H \ G ) .

Proposition 2.4

Let Γ n be a sequence of discrete subgroups of 𝐺 which converges to a closed unimodular subgroup 𝐻 in the Chabauty topology. Let K < G be a compact subgroup of 𝐺. For any vectors v , w L 2 ( H \ G ) , there exist sequences v n , w n L 2 ( Γ n \ G ) , n N , such that

  1. for all g G ,

    lim n v n , g . w n L 2 ( Γ n \ G ) = v , g . w L 2 ( H \ G ) ,

    and the convergence is uniform on compact subsets of 𝐺;

  2. we have

    lim n v n L 2 ( Γ n \ G ) = v L 2 ( H \ G ) & lim n w n L 2 ( Γ n \ G ) = w L 2 ( H \ G ) ;

  3. we have that, for all n N ,

    dim K . v n dim K . v and dim K . w n dim K . w .

Proof

Since C c ( H \ G ) is dense in L 2 ( H \ G ) , the matrix coefficient

g v , g . w L 2 ( H \ G )

can be approximated by the matrix coefficients for continuous compactly supported functions, uniformly on compact subsets of 𝐺. This approximation can be done without increasing the dimensions of the spaces spanned by the 𝐾-orbits of 𝑣 and 𝑤. In fact, let u m be a sequence of compactly supported right 𝐾-invariant functions on H \ G converging to the constant function 1 uniformly on compact subsets of H \ G . Since the multiplication by u m is 𝐾-equivariant, we have dim K . ( u m v ) dim K . v , similarly for 𝑤. The matrix coefficients g u m v , g . u m w L 2 ( H \ G ) converge to g v , g . w L 2 ( H \ G ) uniformly on 𝐺. Thus we have shown that v , w can be replaced by compactly supported functions, spanning 𝐾-invariant subspaces of equal or smaller dimension. We need one more step to replace them by continuous functions.

Let ϕ ̃ m C c ( G ) be a sequence of non-negative continuous functions with

ϕ ̃ m ( g ) d g = 1

and support contained in some neighborhood U ̃ m of 𝑒 such that U ̃ m { e } as m . Define ϕ m C c ( G ) by

ϕ m ( g ) = K ϕ ̃ m ( k 1 g k ) d k for g G ,

where d k is the probability Haar measure on 𝐾. Clearly, ϕ m is non-negative, continuous and ϕ m ( g ) d g = 1 . The support of ϕ m is contained in U m : = { k U ̃ m k 1 : k K } . Note that U m { e } as m ; otherwise, we have, by passing to a subsequence, k m g m k m 1 g for some k m K converging to k 0 K , g m U ̃ m and g e . Since g m e as m , this is a contradiction.

Consider the convolution v ϕ m ,

v ϕ m ( H g ) = G v ( H g x ) ϕ m ( x 1 ) d x for H g H \ G ,

and similarly for w ϕ m . The functions v ϕ m and w ϕ m are continuous compactly supported functions on H \ G .

Since the sequence ϕ m is an approximate identity, the matrix coefficient

g v ϕ m , g . w ϕ m L 2 ( H \ G )

converges to g v , g . w L 2 ( H \ G ) , uniformly on compact sets. Furthermore, because ϕ m is 𝐾-conjugation invariant, the map v v ϕ m commutes with the action of 𝐾, i.e.,

k . ( v ϕ m ) = ( k . v ) ϕ m for all k K .

It follows that dim K . ( v ϕ m ) dim K . v , and similarly for 𝑤. Therefore, we may assume without loss of generality that v , w C c ( H \ G ) .

First, let v ̃ 0 C ( G ) be the lift of 𝑣 to 𝐺, i.e., for all g G , v ̃ 0 ( g ) : = v ( H g ) . We note that dim K . v = dim K . v ̃ 0 .

Now, we choose a right 𝐾-invariant non-negative function φ C c ( G ) such that

H φ ( h g ) d ν H ( h ) = 1 for every g H supp v H supp w .

Define v ̃ C c ( G ) by v ̃ ( g ) : = φ ( g ) v ̃ 0 ( g ) for all g G . Then, for each g G , we have

H v ̃ ( h g ) d ν H ( h ) = v ( g ) .

Moreover, dim K . v ̃ dim K . v ̃ 0 = dim K . v .

Choose a non-negative function s C c ( G ) such that s ( e ) > 0 and

H s ( h ) d ν H ( h ) = 1 .

Set α n : = γ Γ n s ( γ ) , and define v n C c ( Γ n \ G ) as follows: for all g G ,

v n ( g ) : = α n 1 / 2 γ Γ n v ̃ ( γ g ) .

Then dim K . v n dim K . v ̃ dim K . v .

Let w ̃ C c ( G ) and w n C c ( Γ n \ G ) be functions constructed in the same way for the vector 𝑤. We claim that, for all g G ,

v n , g . w n L 2 ( Γ n \ G ) v , g . w L 2 ( H \ G ) ,

uniformly on compact subsets of 𝐺. Indeed,

v n , g . w n L 2 ( Γ n \ G ) = α n 1 Γ n \ G ( γ Γ n v ̃ ( γ x ) γ Γ n w ̃ ( γ x g ) ) d x = G v ̃ ( x ) ( α n 1 γ Γ n w ̃ ( γ x g ) ) d x .

Proposition 2.1 yields the weak-* convergence of measures

α n 1 γ Γ n δ γ d ν H .

It follows that

lim n α n 1 γ Γ n w ̃ ( γ x g ) = H w ̃ ( h x g ) d ν H ( h )

and the convergence is uniform for all 𝑔 and 𝑥 in a given compact subset of 𝐺. Indeed, for x , g C , 𝐶 compact, the family of functions w ̃ ( x g ) is equicontinuous and supported in a single compact set, so the integrals converge uniformly for any weak-* convergent sequence of measures. Since v ̃ is compactly supported, we get

lim n v n , g . w n L 2 ( Γ n \ G ) = G v ̃ ( x ) Γ w ̃ ( h x g ) d ν H ( h ) d x ,

and the convergence is uniform for all 𝑔 in a given compact subset of 𝐺. Since

G v ̃ ( x ) Γ w ̃ ( h x g ) d ν H ( h ) d g = G v ̃ ( x ) w ( H x g ) d x = H \ G v ( H x ) w ( H x g ) d H \ G ( H x ) = v , g . w L 2 ( H \ G ) ,

this finishes the proof of (1) and (3). Claim (2) follows since the above argument applies when v = w and g = e and hence gives v n , v n L 2 ( Γ n \ G ) v , v L 2 ( H \ G ) and similarly for w n and 𝑤. ∎

Remark 2.5

This proposition implies that if Γ n converges to 𝐻 in the Chabauty topology, then L 2 ( H \ G ) is weakly contained in n = n 0 L 2 ( Γ n \ G ) for all n 0 1 .[3]

3 Temperedness is a closed condition in Hom ( Γ , G )

Let 𝐺 be a connected semisimple real algebraic group. Let 𝑃 be a minimal parabolic subgroup of 𝐺 with a fixed Langlands decomposition P = M A N , where 𝐴 is a maximal real split torus of 𝐺, 𝑀 is the maximal compact subgroup of 𝑃, which commutes with 𝐴, and 𝑁 is the unipotent radical of 𝑃. We denote by 𝔤 and 𝔞 the Lie algebras of 𝐺 and 𝐴 respectively. We fix a positive Weyl chamber a + a so that Lie N consists of positive root subspaces. Let Σ + = Σ + ( g , a ) denote the set of all positive roots for ( g , a + ) . For each α Σ + , let m ( α ) be its multiplicity. We also write Π Σ + for the set of all simple roots. We denote by

(3.1) ρ = 1 2 α Σ + m ( α ) α

the half-sum of the positive roots for ( g , a + ) , counted with multiplicity.

We fix a maximal compact subgroup 𝐾 of 𝐺 so that the Cartan decomposition

G = K ( exp a + ) K

holds, that is, for any g G , there exists a unique element μ ( g ) a + such that

g K exp μ ( g ) K .

Let d g be a Haar measure on 𝐺. The right translation action of 𝐺 on itself induces the regular representation L 2 ( G ) = L 2 ( G , d g ) .

Following Harish-Chandra, we call a unitary representation ( π , H ) of 𝐺tempered if 𝜋 is weakly contained in the regular representation L 2 ( G ) .

For any p > 0 , a unitary representation ( π , H ) of 𝐺 is said to be almost L p -integrable if all of its matrix coefficients are L p + ε -integrable for any ε > 0 .

Denote by Ξ = Ξ G the Harish-Chandra function of 𝐺. It is a bi-𝐾-invariant function satisfying that, for any ε > 0 , there exist c , c ε > 0 such that

c e ρ ( v ) Ξ ( exp v ) c ε e ( 1 ε ) ρ ( v ) for all v a + .

We will use the following characterization of a tempered representation of 𝐺 given by Cowling, Haggerup and Howe.

Theorem 3.1

Theorem 3.1 ([10])

For a unitary representation ( π , H ) of 𝐺, the following are equivalent:

  1. 𝜋 is tempered;

  2. 𝜋 is almost L 2 -integrable;

  3. for any 𝐾-finite unit vectors v 1 , v 2 H and any g G ,

    | π ( g ) v 1 , v 2 | ( dim π ( K ) v 1 dim π ( K ) v 2 ) 1 / 2 Ξ G ( g ) .

Definition 3.2

We say that a unimodular subgroup 𝐻 is a tempered subgroup of 𝐺 (or 𝐺-tempered) if the quasi-regular representation L 2 ( H \ G ) is a tempered representation of 𝐺.

Lemma 3.3

Lemma 3.3 ([3, Proposition 3.1])

Let 𝐻 be a unimodular closed subgroup of 𝐺. If 𝐻 is 𝐺-tempered, then any unimodular closed subgroup H < H is also 𝐺-tempered.

We show that temperedness is a closed condition both for the Chabauty topology and the algebraic topology (Theorems 3.4 and 3.7).

Theorem 3.4

The Chabauty limit of a sequence of tempered discrete subgroups of 𝐺 is unimodular and tempered.

Proof

Suppose that Γ n is a sequence of tempered discrete subgroups converging to a closed subgroup 𝐻 in the Chabauty topology. We have that 𝐻 is unimodular by Proposition 2.1. We claim that L 2 ( H \ G ) is tempered. Suppose not. By Theorem 3.1, there exist 𝐾-finite unit vectors v , w L 2 ( H \ G ) and g G such that

| v , g . w L 2 ( H \ G ) | > Ξ ( g ) dim K . v 1 / 2 dim K . w 1 / 2 .

By Proposition 2.4, there exist vectors v n , w n L 2 ( Γ n \ G ) such that

v n v , w n w as n , dim K . v n dim K . v , dim K . w n dim K . w , v , g . w L 2 ( H \ G ) = lim n v n , g . w n L 2 ( Γ n \ G ) .

We can normalize v n , w n to be unit vectors without affecting the above properties. We deduce that, for all 𝑛 large enough,

| v n , g . w n L 2 ( Γ n \ G ) | > Ξ ( g ) dim K . v n 1 / 2 dim K . w n 1 / 2 .

This is a contradiction since L 2 ( Γ n \ G ) is tempered.

Alternatively, one can use [18, Theorem 4.2], which shows that L 2 ( H \ G ) is weakly contained in the direct sum n = 1 L 2 ( Γ n \ G ) . If Γ n were all tempered, we would deduce that L 2 ( H \ G ) is weakly contained in n = 1 L 2 ( G ) , hence in L 2 ( G ) , which then implies that 𝐻 is tempered. ∎

Definition 3.5

We say that a sequence of discrete subgroups Γ i of 𝐺 converges to a discrete subgroup Γ algebraically if there exists a sequence of isomorphisms χ i : Γ Γ i such that, for all γ Γ , χ i ( γ ) converges to 𝛾 as i . In other words, χ i converges to the natural inclusion id Γ in Hom ( Γ , G ) , where the space Hom ( Γ , G ) is endowed with the topology of pointwise convergence. In this case, Γ is called the algebraic limit of Γ i

Remark 3.6

We refer the readers to [4] for a comparison of algebraic and Chabauty convergence; in particular, each notion fails to imply the other in general.

Theorem 3.7

The algebraic limit of a sequence of tempered discrete subgroups of 𝐺 is tempered.

Proof

Let Γ i be a sequence of tempered discrete subgroups of 𝐺 which converges to a discrete subgroup Γ algebraically. By passing to a subsequence if necessary, we may assume that Γ i converges to a closed subgroup 𝐻 in the Chabauty topology. Since Γ is the algebraic limit of Γ i , we have Γ < H .

By Theorem 3.4, 𝐻 is unimodular and tempered. Since any closed unimodular subgroup of a tempered subgroup is tempered by Lemma 3.3, Γ is tempered as desired. ∎

The following is an equivalent formulation of Theorem 3.7.

Theorem 3.8

If a discrete subgroup Γ is a non-tempered subgroup of 𝐺, there exists an open neighborhood 𝒪 of id Γ in Hom ( Γ , G ) such that, for any σ O , σ ( Γ ) is non-tempered.

4 Growth indicator of a lattice of SO ( n , 1 ) as a subgroup of SO ( n , 2 )

Let G = SO ( n , 2 ) for n 2 . Consider the quadratic form

Q ( x 1 , , x n + 2 ) = x 1 x n + 2 + x 2 x n + 1 + i = 3 n x i 2 .

We realize 𝐺 as the identity component of the following special orthogonal group:

SO ( Q ) = { g SL n + 2 ( R ) : Q ( g X ) = Q ( X ) for all X R n + 2 } .

Consider the diagonal subgroup

A = { diag ( e t 1 , e t 2 , 1 , , 1 , e t 2 , e t 1 ) : t 1 , t 2 R } ,

which is a maximal real split torus of 𝐺. We denote by 𝔤 the Lie algebra of 𝐺 and set

a = { v = diag ( v 1 , v 2 , 0 , , 0 , v 2 , v 1 ) : v 1 , v 2 R } = log A .

For simplicity, we write v = ( v 1 , v 2 , 0 , , 0 , v 2 , v 1 ) for an element of 𝔞. Choose a positive Weyl chamber

a + = { v = ( v 1 , v 2 , 0 , , 0 , v 2 , v 1 ) : v 1 v 2 0 } .

Since 𝐺 is invariant under the Cartan involution g g T ,

K = { g G : g g T = e } = G SO ( n + 2 )

is a maximal compact subgroup of 𝐺 and we have the Cartan decomposition G = K ( exp a + ) K . We denote by μ : G a + the Cartan projection of 𝐺.

We then have two simple (restricted) roots α 1 and α 2 for ( g , a ) given by

α 1 ( v ) = v 1 v 2 and α 2 ( v ) = v 2 for all v a .

By explicit computation of 𝔤, we can see that the set of all positive roots of 𝔤 is given by

Σ + ( g , a ) = { α 1 , α 2 , α 1 + α 2 , α 1 + 2 α 2 } .

The direct sum of root subspaces is given by

{ ( 0 x Y 1 z 0 0 0 Y 2 0 z Y 1 t Y 2 t 0 x 0 0 ) : x , z R , Y 1 , Y 2 R n 2 } ,

where the subspaces corresponding to x R , Y 1 R n 2 , Y 2 R n 2 , and z R are root subspaces for α 1 , α 1 + α 2 , α 2 , and α 1 + 2 α 2 respectively. Hence the multiplicities are given by

m ( α 1 ) = m ( α 1 + 2 α 2 ) = 1 , m ( α 1 + α 2 ) = m ( α 2 ) = n 2 .

Since ( α 1 + α 2 ) ( v ) = v 1 and ( α 1 + 2 α 2 ) ( v ) = v 1 + v 2 , the half-sum of all positive roots counted with multiplicity is

ρ ( v ) = α Σ + m ( α ) α ( v ) = 1 2 ( n v 1 + ( n 2 ) v 2 ) for v a + .

Bound on growth indicator for general non-lattice subgroups

Recall the definition of the growth indicator of a discrete subgroup of 𝐺 from (1.1). For any discrete subgroup Γ of 𝐺, the growth indicator ψ Γ is concave and upper-semicontinuous [33, I.1 Théorème]. Since dim a + = 2 , it follows that ψ Γ is continuous on the limit cone L Γ .

The quantitative Kazhdan’s property (T) of the group 𝐺 obtained in [31] yields the following explicit upper bound.

Proposition 4.1

For any non-lattice discrete subgroup Γ of 𝐺, we have

ψ Γ ( v ) ( n 1 ) v 1 + ( n 2 ) v 2 for all v a + .

Proof

By [27, Theorem 7.1], we have

ψ Γ ( v ) ( 2 ρ Θ ) ( v ) for all v a + ,

where Θ is the half-sum of all roots in a maximal strongly orthogonal system of Σ + ( g , a ) . Since { α 1 , α 1 + 2 α 2 } is a maximal strongly orthogonal system, we have Θ ( v ) = v 1 for all v a + . Therefore,

( 2 ρ Θ ) ( v ) = ( n 1 ) v 1 + ( n 2 ) v 2 ,

proving the claim. ∎

Growth indicator for discrete subgroups of 𝐺 that are lattices of 𝐻

Let

H = SO ( n , 1 ) .

The restriction of the quadratic form 𝑄 to the hyperplane V : = { x 1 = x n + 2 } yields a quadratic form Q 0 = Q | V in ( n + 1 ) variables. We identify

H = SO ( n , 1 ) = { g G : g ( V ) = V } = SO ( Q 0 ) .

Since 𝐻 is invariant under the Cartan involution g g T , the intersection K H is a maximal compact subgroup of 𝐻. Denoting by 𝔥 the Lie algebra of 𝐻, we have

h a = { diag ( 0 , v 2 , 0 , , 0 , v 2 , 0 ) : v 2 R } .

Note that the Cartan projection μ ( H ) is equal to a + ker α 2 , i.e.,

μ ( H ) = { v = ( v 1 , 0 , , 0 , v 1 ) : v 1 0 } .

To see that, apply the Weyl element switching the first two rows (and hence the last two rows) to h a , resulting in { ( v 2 , 0 , , 0 , v 2 ) : v 2 R } = ker α 2 .

Proposition 4.2

Let Γ < G be a discrete subgroup such that Γ is a lattice of 𝐻. Then

ψ Γ ( v ) = { ( n 1 ) v 1 for v = ( v 1 , 0 , , 0 , v 1 ) , v 1 0 , for v μ ( H ) .

In other words,

(4.1) ψ Γ 2 ( n 1 ) n ρ on a +

with the equality on μ ( H ) .

Proof

Since Γ is a lattice of 𝐻, the limit cone of Γ satisfies

L Γ = μ ( H ) = a + ker α 2 .

Hence, for v μ ( H ) , ψ Γ ( v ) = . Let denote the norm on 𝔞 induced from the Riemannian metric on G / K . Since H / ( H K ) G / K is an isometric embedding, we have that, for all h H , μ ( h ) is equal to the Riemannian distance d H / ( H K ) ( h o , o ) in H / ( H K ) . Since ψ Γ is independent of the choice of a norm, we may assume that, for all h H , μ ( h ) is equal to the hyperbolic distance d H n ( h o , o ) by identifying H / ( H K ) H n , which is equivalent to ( v 1 , 0 , , 0 , v 1 ) = | v 1 | . Since Γ < H is a lattice, we have

# { γ Γ : d H n ( γ o , o ) < T } C e ( n 1 ) T as T

(cf. [14, 17]). Hence, for v = ( v 1 , 0 , , 0 , v 1 ) with v 1 0 ,

ψ Γ ( v ) = v lim sup T log # { γ Γ : μ ( γ ) T } T = ( n 1 ) v 1 .

Since ρ ( v 1 , 0 , , 0 , v 1 ) = n 2 v 1 by (3.1), the claim follows. ∎

Remark 4.3

Note that the upper bound (4.1) already follows from Proposition 4.1 once we know that L Γ a α 1 . The above proposition shows that this upper bound is optimal for the case at hand.

Remark 4.4

We remark that Proposition 4.2 holds in a more general setting: let 𝐺 be a connected semisimple real algebraic subgroup with Cartan decomposition G = K A + K and H < G a connected reductive real algebraic subgroup such that

H = ( K H ) ( A + H ) ( K H ) .

Let Γ be a lattice of 𝐻. Then ψ Γ ( v ) = 2 ρ H ( v ) if v log ( H A + ) and otherwise, where 2 ρ H is the sum of all positive roots of ( Lie ( H ) , log ( H A + ) ) .

We recall the following criterion on the temperedness of L 2 ( Γ \ G ) .

Theorem 4.5

Theorem 4.5 ([16], [28, Theorem 5.1])

For any discrete subgroup Γ of a connected semisimple real algebraic group 𝐺, we have ψ Γ ρ if and only if Γ is a tempered subgroup of 𝐺. Moreover, if ψ Γ ( 1 + η ) ρ , then L 2 ( Γ \ G ) is almost L p for p 2 1 η .

That L 2 ( Γ \ G ) is almost L p means that every matrix coefficient of the quasi-regular representation L 2 ( Γ \ G ) is L p + ε -integrable for any ε > 0 . By Theorem 3.1, a discrete subgroup Γ is 𝐺-tempered if and only if L 2 ( Γ \ G ) is almost L 2 .

Since ψ Γ = 2 ( n 1 ) n ρ on μ ( H ) by Proposition 4.2, we obtain the following examples of non-tempered subgroups of 𝐺.

Corollary 4.6

Let Γ be a lattice of H = SO ( n , 1 ) , considered as a subgroup of G = SO ( n , 2 ) . Then Γ is 𝐺-tempered if and only if n = 2 . Moreover, for each n 2 , L 2 ( Γ \ G ) is almost L n .

5 Deformations and non-tempered Zariski-dense examples

Let G = SO ( n , 2 ) and H = SO ( n , 1 ) = Isom + ( H n ) . Let Γ be a torsion-free uniform lattice of 𝐻 such that M = Γ \ H n is a closed hyperbolic 𝑛-manifold with a properly embedded totally geodesic hyperplane 𝑆.

Remark 5.1

For any n 2 , such a Γ exists, for instance, consider a quadratic form

Q 0 ( x 1 , , x n + 1 ) = i = 1 n x i 2 d x n + 1 2

for a square-free integer 𝑑. Let Γ < SO ( Q 0 ) SL n + 1 ( Z d ) be a torsion-free subgroup of finite index. Then Γ is a uniform lattice of SO ( Q 0 ) (see [5]). Considering SL n as a subgroup of SL n + 1 embedded as the lower diagonal block subgroup, the intersection Δ = Γ SL n is a uniform lattice of SO ( Q 0 ) SL n SO ( n 1 , 1 ) . Now M = Γ \ H n is a closed hyperbolic 𝑛-manifold with a properly embedded geodesic hyperplane S = Δ \ H n 1 .

We may assume that Γ SO ( n 1 , 1 ) = Δ is a uniform lattice of SO ( n 1 , 1 ) by replacing Γ by a conjugate if necessary.

We briefly recall the bending construction of Johnson–Millson [22]. Their bending was constructed with the ambient group SL n + 2 ( R ) . We use a modification by Kassel [24, Section 6] where the bending was done inside G = SO ( n , 2 ) . There exists a one-parameter subgroup a t G which centralizes SO ( n 1 , 1 ) . If 𝑆 is separating, i.e., M S is the disjoint union of two connected components M 1 and M 2 , then Γ = Γ 1 Δ Γ 2 . Consider the homomorphism σ t : Γ G given by

σ t ( γ ) = { γ for γ Γ 1 , a t γ a t for γ Γ 2 .

Since a t commutes with Δ, σ t is well-defined. If 𝑆 does not separate 𝑀, then Γ is an HNN extension of Δ, and we have a homomorphism σ t defined similarly (cf. [24, Section 6.3]).

The following Zariski-density and discreteness results were obtained in [24] and [20] respectively.

Proposition 5.2

For all sufficiently small t 0 , σ t ( Γ ) is discrete and Zariski-dense in G = SO ( n , 2 ) .

We now give a proof of Theorem 1.2.

Theorem 5.3

Let n 3 . For all sufficiently small t 0 , the subgroup σ t ( Γ ) is a non-tempered, Zariski-dense and discrete subgroup of G = SO ( n , 2 ) .

Proof

The subgroup Γ is a non-tempered subgroup of 𝐺 for n 3 by Corollary 4.6. Hence the claim follows from Theorem 3.8 and Proposition 5.2. ∎

6 Anosov representations and non-temperedness

In this section, we prove a stronger result than Theorem 1.2 using the theory of Anosov representations. We keep the notation for G = SO ( n , 2 ) , H = SO ( n , 1 ) , 𝔞, α 1 , α 2 , etc. from Section 4. Let Γ be a torsion-free uniform lattice of 𝐻 such that the closed hyperbolic manifold Γ \ H n has a properly embedded totally geodesic hyperplane as in Section 5.

Definition 6.1

For a non-empty subset θ Π = { α 1 , α 2 } , a finitely generated subgroup Γ 0 of 𝐺 is called 𝜃-Anosov if there exists C > 0 such that, for all γ Γ 0 and α θ , we have α ( μ ( γ ) ) C 1 | γ | C , where | γ | denotes the word length of 𝛾 with respect to a fixed finite generating subset of Γ 0 . A Π-Anosov subgroup is called Borel–Anosov.

Lemma 6.2

The subgroup Γ is an { α 1 } -Anosov subgroup of 𝐺.

Proof

Note that β 1 : = α 1 restricted to h a is a simple root of ( h , h a ) with respect to the choice of a positive Weyl chamber

( h a ) + = { v = ( 0 , v 2 , 0 , , 0 , v 2 , 0 ) : v 2 0 } .

Since Γ is a uniform lattice of 𝐻, it is in particular a convex cocompact subgroup of 𝐻, and hence a { β 1 } -Anosov subgroup of 𝐻 (see [21]). Therefore, there exists C 1 such that, for all γ Γ , β 1 ( μ H ( γ ) ) C 1 | γ | C , where μ H denotes the Cartan projection map of 𝐻. Since β 1 μ H = α 1 μ | H , it follows that α 1 ( μ ( γ ) ) C 1 | γ | C for all γ Γ . This proves the claim. ∎

Theorem 6.3

Let n 3 , and G = SO ( n , 2 ) . There exists a non-empty open subset 𝒪 of Hom ( Γ , G ) such that, for any σ O , we have

  1. 𝜎 is injective and discrete;

  2. σ ( Γ ) is a Zariski-dense { α 1 } -Anosov subgroup of 𝐺;

  3. σ ( Γ ) is not 𝐺-tempered.

By [1, Proposition 8.2], the set of Zariski-dense representations of Γ forms an open subset of Hom ( Γ , G ) , which we know is non-empty by Proposition 5.2. Moreover, all Anosov representations are discrete with finite kernel and the set of all { α 1 } -Anosov representations forms an open subset in Hom ( Γ , G ) by [21, 23]. Since Γ is assumed to be torsion-free, Theorem 6.3 follows from Theorem 3.8 and non-temperedness of Γ.

In the rest of this section, we will give a different proof of Theorem 6.3 (3) using the continuity of limit cones under a small deformation of Γ and the Anosov property of Γ.

For any discrete subgroup Γ 0 of 𝐺 and any linear form ψ a such that ψ > 0 on L Γ 0 { 0 } , denote by δ ψ , Γ 0 the abscissa of convergence of the series s γ Γ 0 e s ψ ( μ ( γ ) ) . This is well-defined and 0 δ ψ , Γ 0 < . Since ρ > 0 on a + { 0 } , δ ρ , Γ 0 is well-defined for any discrete subgroup Γ 0 < G . Theorem 4.5 can be reformulated as follows.

Proposition 6.4

For any discrete subgroup Γ 0 of a connected semisimple real algebraic group G 0 , we have δ ρ , Γ 0 1 if and only if Γ 0 is G 0 -tempered.

Proof

By [25, Theorem 2.5], we have ψ Γ 0 δ ρ , Γ 0 ρ and ψ Γ 0 ( v ) = δ ρ , Γ 0 ρ ( v ) for some non-zero v a + . Therefore, the claim follows from Theorem 4.5. ∎

Set a α 1 = ker α 2 and a α 1 + = a + ker α 2 . Let p α 1 : a a α 1 denote the unique projection invariant under the Weyl element fixing a α 1 pointwise, which is simply the reflection about a α 1 . The space of linear forms a α 1 can be identified with the set of all linear forms in a which are invariant under p α 1 . The following follows by combining [6, Proposition 8.1] and [36, Corollary 5.5.3], both of whose proofs are based on thermodynamic formalism.

Theorem 6.5

For any ψ a α 1 which is positive on a α 1 + { 0 } , the critical exponent δ ψ , σ ( Γ ) varies analytically on any sufficiently small analytic neighborhood of an { α 1 } -Anosov representation of Hom ( Γ , G ) .

Since Γ is a convex cocompact subgroup of 𝐻, the following is a special case of Kassel’s theorem [24, Proposition 5.1] (see also [13, Theorem 1.1] for a recent generalization).

Proposition 6.6

For any η > 0 , we have an open neighborhood 𝒪 of id Γ in Hom ( Γ , G ) such that, for any σ O , the limit cone of σ ( Γ ) is contained in

C η : = { v a + : v a α 1 < η v } .

Remark 6.7

For the bending deformations σ t discussed in Section 5, we always have a non-trivial element of 𝛾 (of infinite order) such that σ t ( γ ) = γ , and hence

μ ( σ t ( γ ) ) μ ( H ) { 0 } .

Therefore, we have the following property: for all sufficiently small t 0 , the limit cone of σ t ( Γ ) contains the ray μ ( H ) . Since σ t ( Γ ) is Zariski-dense, its limit cone is convex and has non-empty interior [2]. Therefore, Proposition 6.6 implies that the limit cone of σ t ( Γ ) is the convex cone given by

L σ t ( Γ ) = { v = ( v 1 , v 2 , 0 , , v 2 , v 1 ) a + : 0 v 2 c σ t v 1 } ,

where c σ t > 0 tends to 0 as t 0 .

Recall from Proposition 4.2. that

δ ρ , Γ = 2 ( n 1 ) n .

The following proposition gives an alternative proof of Theorem 6.3 (3).

Proposition 6.8

For any sufficiently small ε > 0 , there exists an open neighborhood O = O ( ε ) of id Γ in Hom ( Γ , G ) such that, for any σ O ,

| δ ρ , σ ( Γ ) 2 ( n 1 ) n | < ε .

In particular, for n 3 , we have ψ Γ ρ , and hence σ ( Γ ) is non-tempered in 𝐺 for all σ O ( n 2 n )

Proof

Let ρ be the restriction of 𝜌 to a α 1 . We may consider ρ as a linear form on 𝔞 by precomposing with p α 1 . Note that ρ is non-negative on a α 1 + .

Let ε > 0 . We can find η > 0 so that, for any v C η = { v a + : v a α 1 < η v } ,

ε ρ ( v ) ( ρ ρ ) ( v ) ε ρ ( v ) .

We can take a small neighborhood 𝒪 of id Γ so that, for any σ O , the limit cone of σ ( Γ ) is contained in the cone C η by Proposition 6.6. In particular, μ ( σ ( γ ) ) C η for all γ Γ except for some finite subset F σ . Then, for any σ O , we have that, for all s > 0 ,

γ Γ F σ e ( 1 ε ) s ρ ( μ ( σ ( γ ) ) ) γ Γ F σ e s ρ ( μ ( σ ( γ ) ) ) .

It follows that

δ ( 1 ε ) ρ , σ ( Γ ) δ ρ , σ ( Γ ) and hence δ ρ , σ ( Γ ) ( 1 ε ) δ ρ , σ ( Γ ) .

Similarly, we have

γ Γ F σ e ( 1 + ε ) s ρ ( μ ( σ ( γ ) ) ) γ Γ F σ e s ρ ( μ ( σ ( γ ) ) ) , δ ( 1 + ε ) ρ , σ ( Γ ) δ ρ , σ ( Γ ) and hence δ ρ , σ ( Γ ) ( 1 + ε ) δ ρ , σ ( Γ ) .

Therefore,

(6.1) ( 1 ε ) δ ρ , σ ( Γ ) δ ρ , σ ( Γ ) ( 1 + ε ) δ ρ , σ ( Γ ) .

By replacing 𝒪 by a smaller neighborhood of id Γ if necessary, we may assume that

(6.2) | δ ρ , σ ( Γ ) δ ρ , Γ | ε for all σ O

by Theorem 6.5. Hence, using that 1 δ ρ , Γ = 2 ( n 1 ) / n 2 , we deduce from (6.1) and (6.2) that

| δ ρ , σ ( Γ ) δ ρ , Γ | < 5 ε for all σ O .

Since δ ρ , Γ = 2 ( n 1 ) / n , the claim follows. ∎

We can also obtain the following estimates for the growth indicator ψ σ ( Γ ) .

Corollary 6.9

For any sufficiently small ε > 0 , there exists an open neighborhood O = O ( ε ) of id Γ in Hom ( Γ , G ) such that, for any σ O ,

ψ σ ( Γ ) ( v ) ( 2 ( n 1 ) n + ε ) ρ ( v ) for all v a + ,
(6.3) ψ σ ( Γ ) ( v σ ) ( 2 ( n 1 ) n ε ) ρ ( v σ ) for some unit vector v σ a + .
Moreover, v σ converges to a unit vector in a α 1 as σ id Γ .

Proof

Recall that ψ σ ( Γ ) δ ρ , σ ( Γ ) ρ and ψ σ ( Γ ) ( v σ ) = δ ρ , σ ( Γ ) ρ ( v σ ) for some non-zero vector v σ on the limit cone L σ ( Γ ) (see [25, Theorem 2.5]). Hence the inequalities follow from Proposition 6.8. The last claim follows from Proposition 6.6. ∎

Finally, since v σ is of the form ( v σ , 1 , c σ v σ , 1 , 0 , , c σ v σ , 1 , v σ , 1 ) for some v σ , 1 > 0 with c σ 0 , inequality (6.3) and Proposition 4.1 imply inequality (1.4) in Theorem 1.5. Hence, together with Theorem 6.3, Proposition 6.8, and Corollary 6.9, this completes the proof of Theorem 1.5.

Funding statement: Mikołaj Frączyk was supported by the Dioscuri programme initiated by the Max Planck Society, jointly managed with the National Science Centre in Poland, and mutually funded by the Polish Ministry of Education and Science and the German Federal Ministry of Education and Research.

Acknowledgements

We would like to thank Dongryul Kim and Tobias Weich for useful comments on a preliminary version of the article. We would also like to thank Marc Burger for telling us about the reference [18]. We also thank the referees for their careful reading of our paper and their useful comments.

References

[1] N. A’Campo and M. Burger, Réseaux arithmétiques et commensurateur d’après G. A. Margulis, Invent. Math. 116 (1994), no. 1–3, 1–25. 10.1007/BF01231555Search in Google Scholar

[2] Y. Benoist, Propriétés asymptotiques des groupes linéaires, Geom. Funct. Anal. 7 (1997), no. 1, 1–47. 10.1007/PL00001613Search in Google Scholar

[3] Y. Benoist and T. Kobayashi, Tempered homogeneous spaces II, Dynamics, geometry, number theory—the impact of Margulis on modern mathematics, University of Chicago, Chicago (2022), 213–245. 10.7208/chicago/9780226804163.003.0006Search in Google Scholar

[4] I. Biringer and J. Souto, Algebraic and geometric convergence of discrete representations into PSL 2 C , Geom. Topol. 14 (2010), no. 4, 2431–2477. 10.2140/gt.2010.14.2431Search in Google Scholar

[5] A. Borel and Harish-Chandra, Arithmetic subgroups of algebraic groups, Ann. of Math. (2) 75 (1962), 485–535. 10.2307/1970210Search in Google Scholar

[6] M. Bridgeman, R. Canary, F. Labourie and A. Sambarino, The pressure metric for Anosov representations, Geom. Funct. Anal. 25 (2015), no. 4, 1089–1179. 10.1007/s00039-015-0333-8Search in Google Scholar

[7] M. Burger, J.-S. Li and P. Sarnak, Ramanujan duals and automorphic spectrum, Bull. Amer. Math. Soc. (N. S.) 26 (1992), no. 2, 253–257. 10.1090/S0273-0979-1992-00267-7Search in Google Scholar

[8] K. Corlette, Hausdorff dimensions of limit sets. I, Invent. Math. 102 (1990), no. 3, 521–541. 10.1007/BF01233439Search in Google Scholar

[9] M. Cowling, Sur les coefficients des représentations unitaires des groupes de Lie simples, Analyse harmonique sur les groupes de Lie II, Lecture Notes in Math. 739, Springer, Berlin (1979), 132–178. 10.1007/BFb0062491Search in Google Scholar

[10] M. Cowling, U. Haagerup and R. Howe, Almost L 2 matrix coefficients, J. reine angew. Math. 387 (1988), 97–110. 10.1515/crll.1988.387.97Search in Google Scholar

[11] T. Delzant, Sous-groupes distingués et quotients des groupes hyperboliques, Duke Math. J. 83 (1996), no. 3, 661–682. 10.1215/S0012-7094-96-08321-0Search in Google Scholar

[12] S. Dey, D. Kim and H. Oh, Ahlfors regularity of Patterson–Sullivan measures of Anosov groups and applications, preprint (2024), https://arxiv.org/abs/2401.12398. Search in Google Scholar

[13] S. Dey and H. Oh, Deformations of Anosov subgroups: Limit cones and growth indicators, preprint (2024), https://arxiv.org/abs/2411.04020. Search in Google Scholar

[14] W. Duke, Z. Rudnick and P. Sarnak, Density of integer points on affine homogeneous varieties, Duke Math. J. 71 (1993), no. 1, 143–179. 10.1215/S0012-7094-93-07107-4Search in Google Scholar

[15] S. Edwards, M. Fraczyk, M. Lee and H. Oh, Infinite volume and atoms at the bottom of the spectrum, C. R. Math. Acad. Sci. Paris 362 (2024), 1873–1880. 10.5802/crmath.586Search in Google Scholar

[16] S. Edwards and H. Oh, Temperedness of L 2 ( Γ G ) and positive eigenfunctions in higher rank, Commun. Am. Math. Soc. 3 (2023), 744–778. 10.1090/cams/25Search in Google Scholar

[17] A. Eskin and C. McMullen, Mixing, counting, and equidistribution in Lie groups, Duke Math. J. 71 (1993), no. 1, 181–209. 10.1215/S0012-7094-93-07108-6Search in Google Scholar

[18] J. M. G. Fell, Weak containment and induced representations of groups. II, Trans. Amer. Math. Soc. 110 (1964), 424–447. 10.1090/S0002-9947-1964-0159898-XSearch in Google Scholar

[19] M. Gotô and H. Yamabe, On some properties of locally compact groups with no small subgroup, Nagoya Math. J. 2 (1951), 29–33. 10.1017/S0027763000010011Search in Google Scholar

[20] O. Guichard, Groupes plongés quasi isométriquement dans un groupe de Lie, Math. Ann. 330 (2004), no. 2, 331–351. 10.1007/s00208-004-0553-ySearch in Google Scholar

[21] O. Guichard and A. Wienhard, Anosov representations: Domains of discontinuity and applications, Invent. Math. 190 (2012), no. 2, 357–438. 10.1007/s00222-012-0382-7Search in Google Scholar

[22] D. Johnson and J. J. Millson, Deformation spaces associated to compact hyperbolic manifolds, Discrete groups in geometry and analysis: Papers in honor of Mostow on his sixtieth birthday, Birkhäuser, Boston (1987), 48–106. 10.1007/978-1-4899-6664-3_3Search in Google Scholar

[23] M. Kapovich, B. Leeb and J. Porti, Anosov subgroups: Dynamical and geometric characterizations, Eur. J. Math. 3 (2017), no. 4, 808–898. 10.1007/s40879-017-0192-ySearch in Google Scholar

[24] F. Kassel, Deformation of proper actions on reductive homogeneous spaces, Math. Ann. 353 (2012), no. 2, 599–632. 10.1007/s00208-011-0672-1Search in Google Scholar

[25] D. M. Kim, Y. N. Minsky and H. Oh, Tent property of the growth indicator functions and applications, Geom. Dedicata 218 (2024), no. 1, Paper No. 14. 10.1007/s10711-023-00846-3Search in Google Scholar

[26] B. Kostant, On the existence and irreducibility of certain series of representations, Bull. Amer. Math. Soc. 75 (1969), 627–642. 10.1090/S0002-9904-1969-12235-4Search in Google Scholar

[27] M. Lee and H. Oh, Dichotomy and measures on limit sets of Anosov groups, Int. Math. Res. Not. IMRN 2024 (2024), no. 7, 5658–5688. 10.1093/imrn/rnad188Search in Google Scholar

[28] C. Lutsko, T. Weich and L. Wolf, Polyhedral bounds on the joint spectrum and temperedness of locally symmetric spaces, preprint (2024), https://arxiv.org/abs/2402.02530. Search in Google Scholar

[29] M. Magee, Quantitative spectral gap for thin groups of hyperbolic isometries, J. Eur. Math. Soc. (JEMS) 17 (2015), no. 1, 151–187. 10.4171/jems/500Search in Google Scholar

[30] K. Matsuzaki, Y. Yabuki and J. Jaerisch, Normalizer, divergence type, and Patterson measure for discrete groups of the Gromov hyperbolic space, Groups Geom. Dyn. 14 (2020), no. 2, 369–411. 10.4171/ggd/548Search in Google Scholar

[31] H. Oh, Uniform pointwise bounds for matrix coefficients of unitary representations and applications to Kazhdan constants, Duke Math. J. 113 (2002), no. 1, 133–192. 10.1215/S0012-7094-02-11314-3Search in Google Scholar

[32] S. J. Patterson, The limit set of a Fuchsian group, Acta Math. 136 (1976), no. 3–4, 241–273. 10.1007/BF02392046Search in Google Scholar

[33] J.-F. Quint, Divergence exponentielle des sous-groupes discrets en rang supérieur, Comment. Math. Helv. 77 (2002), no. 3, 563–608. 10.1007/s00014-002-8352-0Search in Google Scholar

[34] J.-F. Quint, L’indicateur de croissance des groupes de Schottky, Ergodic Theory Dynam. Systems 23 (2003), no. 1, 249–272. 10.1017/S0143385702001268Search in Google Scholar

[35] J.-F. Quint, Propriété de Kazhdan et sous-groupes discrets de covolume infini, Proceedings of the Conference on Harmonic Analysis, Trav. Math. 14, Université du Luxembourg, Luxembourg (2003), 143–151. Search in Google Scholar

[36] A. Sambarino, A report on an ergodic dichotomy, Ergodic Theory Dynam. Systems 44 (2024), no. 1, 236–289. 10.1017/etds.2023.13Search in Google Scholar

[37] D. Sullivan, Entropy, Hausdorff measures old and new, and limit sets of geometrically finite Kleinian groups, Acta Math. 153 (1984), no. 3–4, 259–277. 10.1007/BF02392379Search in Google Scholar

[38] D. Sullivan, Related aspects of positivity in Riemannian geometry, J. Differential Geom. 25 (1987), no. 3, 327–351. 10.4310/jdg/1214440979Search in Google Scholar

Received: 2024-11-04
Published Online: 2025-07-08

© 2025 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 29.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/crelle-2025-0043/html
Scroll to top button