Exponential mixing and essential spectral gaps for Anosov subgroups
Abstract.
Let be a Zariski dense -Anosov subgroup of a connected semisimple real algebraic group for some nonempty subset of simple roots . In the Anosov setting, there is a natural compact metric space equipped with a family of translation flows , parameterized by vectors in the interior of the -limit cone of , and they are Hölder conjugate to Hölder reparametrizations of the Gromov geodesic flow. We prove that for all outside an exceptional cone , which is a smooth image of the walls of the Weyl chamber, the translation flow is exponentially mixing with respect to the Bowen–Margulis–Sullivan measure associated to . Moreover, the exponential rate is uniform for a compact set of such . We also obtain an essential spectral gap formulated in terms of the Selberg zeta function and a prime orbit theorem with a power saving error term. Our proof relies on Lie theoretic techniques to prove the crucial local non-integrability condition (LNIC) for the translation flows and thereby implement Dolgopyat’s method in a uniform fashion. The exceptional cone arises from the failure of LNIC for those vectors.
1. Introduction
The main result of this paper is exponential mixing of translation flows associated to Anosov subgroups. Anosov subgroups, first introduced by Labourie [36] and later generalized by Guichard–Wienhard [25], includes many interesting geometric examples such as the images of Hitchin representations [36] into , strongly convex cocompact projective representations into [19], maximal representations into and [10, 11], and Barbot representations into [2]. In recent times, there has been a lot activity in understanding the dynamics related to Anosov subgroups (see for instance [60, 9, 39, 40, 23, 12, 18, 61, 31], etc.). Mixing properties are of particular interest because existing techniques allow one to readily derive various counting and equidistribution results. What is more, exponential mixing allows one to effectivize these results, i.e., obtain precise error terms.
Let us first provide some basic setup for the main theorem. Let be a connected semisimple real algebraic group with , be a maximal real split torus of and fix a closed positive Weyl chamber . Let denote the set of all simple roots for and let denote the opposition involution preserving and acting on by precomposition. Fix a nonempty subset . Let denote the standard parabolic subgroup of corresponding to and be its Langlands decomposition where centralizes and . The -Furstenberg boundary is defined as and we denote the unique open -orbit in by .
Let be a torsion-free Zariski dense -Anosov subgroup, that is, is a torsion-free Gromov hyperbolic group with Gromov boundary and continuous -equivariant maps and such that for all distinct . The images and are the unique -minimal subsets of and . Let . Unless (in which case is compact), does not necessarily act on properly discontinuously (see [24, §1.5] for an extensive discussion). However, the -action restricted to the -invariant subset is always properly discontinuous.
Given , the translation action of on the -coordinate of induces the one-parameter diagonal flow on which we denote by . We further assume that where denotes -limit cone of (see Definition 2.6). In this case, the dynamical system is known to be locally mixing where is the Bowen–Margulis–Sullivan measure associated to . That is, there exists such that for all , we have
When , Thirion [69] first proved the above for Schottky subgroups of and this was later generalized by Sambarino [60] to fundamental groups of compact negatively curved manifolds. Generalizing their work, the authors [18] proved local mixing on (which is a principal -bundle over ) for general -Anosov subgroups. The arguments in [18] easily generalize for arbitrary without the -valued holonomy (cf. [61, Appendix B] for a general theorem and a proof sketch amending the one in [60] using strategies similar to [18]). It is an interesting (and still open) question as to what the rate of mixing above should be in general. It is known to the authors that Dolgopyat’s method cannot be carried out as the non-concentration property or its generalization (cf. [64, 17]) fails miserably whenever is of higher rank; and hence, the rate is expected to be subexponential when . See the results below for the case .
On the other hand, the situation is completely different for the dynamical system . Here, is a compact metric space over which is a trivial -bundle, is a probability measure on such that we have the product structure , and is the translation flow on induced by via the bundle projection (see Subsection 2.4 for details). So far in the literature on Anosov subgroups, translation flows have been studied more often as an auxiliary tool rather than a dynamical system of intrinsic interest. However, translation flows are natural dynamical systems which mimic (and in fact generalize) the geodesic flow for convex cocompact hyperbolic manifolds and they are Hölder conjugate to Hölder reparametrizations of the Gromov geodesic flow associated to . In this vein, we prove that translation flows are exponentially mixing for generic vectors . Indeed, we obtain a detailed characterization of the exponential mixing property for the whole family of translation flows.
The set of generic for which we can prove exponential mixing can be described in terms of a fundamental object associated to called its -growth indicator (see Definition 2.7) which is the higher rank generalization of critical exponent in rank one. We say is -regular if and we define the exceptional cone
It follows from the properties of (see Theorem 2.11(2)(4) and Proposition B.1) that is the image of under a diffeomorphism on and bounds an open strictly convex cone.
We are now ready to state the main theorem regarding exponential mixing of the translation flows associated to -regular . We remark that the translation flows are homothety equivariant, i.e., for all and . For , we denote by the space of -Hölder continuous functions on endowed with the -Hölder norm .
Theorem 1.1.
Let . There exist
-
—
a continuous piecewise-smooth function which is positive except possibly on ,
-
—
a smooth positive function (independent of ),
which are both homothety-invariant, such that for all , and , and , we have
Remark 1.2.
An imprecise version of Theorem 1.1 had been claimed in a talk by the second named author at Institut des Hautes Études Scientifiques in June 2023 as it was well-known to the authors and much of this paper had been written at the time. However, a sticking point delayed the authors in the distribution of the manuscript. See Remark 4.9.
We obtain the following simple corollary when because in this case and so the family of translation flows collapses to a trivial one-dimensional family coinciding with rescalings of the one-parameter diagonal flow. This theorem already generalizes the case that is a projective Anosov subgroup of which now follows from the combination of the works of Delarue–Montclair–Sanders [20] and Stoyanov [67] (see the discussion below) and strengthens the local mixing result above. It also generalizes the case that is of rank one which gives the geodesic flow for convex cocompact locally symmetric spaces and is contained in the work of Stoyanov [67] (see [17] of the authors for the frame flow).
Theorem 1.3.
Suppose . Let . Then, there exist and (independent of ) such that for all and , we have
We also obtain the following corollary regarding uniform exponential mixing.
Theorem 1.4.
Let . Suppose is a compact subset. Then, there exist and (independent of ) such that for all , and , and , we have
Using the spectral bounds of transfer operators in Theorem 4.7, we actually prove the more detailed theorem below from which the above theorems follow. To describe the behavior of the correlation function especially near the boundary of the limit cone , we introduce the map . Equip with the inner product and norm induced by the Killing form on . Define by
where is the bounded elementary function (in the formal sense) with vanishing limit at provided by Theorem 4.1 which can be computed explicitly from the proofs in the paper. Note that is homogeneous of degree because so is . The complex numbers which appear are the Pollicott–Ruelle resonances for the translation flow associated to . The concept of Pollicott–Ruelle resonances originated in [47, 48, 56, 55].
Theorem 1.5.
Let . There exist
-
—
a continuous piecewise-smooth function which is positive except possibly on ,
-
—
a smooth positive function (independent of ),
such that for all , there exist and
-
—
a finite set which come in conjugate pairs,
-
—
a finite set of finite-rank positive bilinear forms ,
which are all homothety-invariant, such that for all and , we have
1.1. Applications
Again using the powerful Theorem 4.7, we also obtain the following applications.
For all and , define
where denotes the primitive period of . We may drop the subscript for the former and write since the set of closed orbits coincide for all translation flows. Moreover, there is a canonical bijective correspondence
where the latter is the set of conjugacy classes of corresponding to the subset of primitive loxodromic elements . Note that in our setting, all nontorsion elements of are loxodromic (see Theorem 2.11). Under the above bijective correspondence, we also have
where we have used the notation and , which is the topological entropy of the translation flow , from Subsection 2.4.
For all , recall the Selberg zeta function [65] defined by
which converges absolutely using the noneffective prime orbit theorem in [16, Eq. (1.8)] (cf. [26, Chapter 2, Definition 4.1] and the subsequent remark). By the work of Pollicott [48], it then extends meromorphically to the domain for some with a simple zero at . He also gives a simple explicit formula for in terms of the expansion/contraction rate from the Anosov property and the topological entropy from which we see that on some conical neighborhood of .
Theorem 1.6.
There exist a homothety-invariant continuous piecewise-smooth function which is positive except possibly on such that for all , the Selberg zeta function has only finitely many zeros on .
We say that the dynamical system has an essential spectral gap of . We emphasize that is an explicit function which dictates the behavior of the decay of the essential spectral gap near the boundary of the limit cone whereas we have no information on the zeros of the Selberg zeta function .
Remark 1.7.
The above could also be formulated equivalently in terms of the Ruelle zeta function defined by for all and then extended meromorphically as above.
Recall the offset logarithmic integral function defined by for all . We have the following prime orbit theorem with a power saving error term.
Theorem 1.8.
There exist
-
—
a continuous piecewise-smooth function which is positive except possibly on ,
-
—
a smooth positive function ,
which are both homothety-invariant, such that for all and , we have
Both the theorems above generalize those in [51, 20] for the case that is a projective Anosov subgroup of (see the discussion below). The above theorem also effectivizes the prime orbit theorem of Sambarino [59, Theorem 7.8] and its generalization by Chow–Fromm [16, Eq. (1.8)] (cf. [16, Corollary 1.4]). In fact, another approach to proving it is by effectivizing the work of [16] with Theorem 1.1 as input.
1.2. Related works
We discuss the related works of Pollicott–Sharp [51] and Delarue–Montclair–Sanders [20]. Both of these works are in the projective Anosov setting.
In [51], the results are for the case that is a Hitchin surface subgroup of , i.e., the image of a Hitchin representation of a surface group into , viewed as a projective Anosov subgroup (in which case ). The authors use a very different kind of coding from that of ours to introduce transfer operators. They then use it to study the Selberg zeta function and obtain both an essential spectral gap and a prime orbit theorem with a power saving error term. This was generalized in [20] to the general projective Anosov setting.
In [20], the results are for the case that is a projective Anosov subgroup of (in which case ). The authors obtain a nice result which shows the existence of a -invariant and flow-invariant open subset containing where the -action is properly discontinuous and the translation flow is a contact Axiom A flow on with as its basic set [20, Theorem A]. With this construction in hand, exponential mixing follows at once via Stoyanov’s work [67]. They also obtain related theorems already mentioned above. The main common difficulty to both [20] and our paper is that the actions which naturally appear are not necessarily properly discontinuous (see Appendix A). This poses an obstruction to using the smooth structure on to prove a local non-integrability condition (LNIC) and also to invoking [67]. Delarue–Montclair–Sanders overcome this by constructing . We take a more direct approach and use weaker properties (which is often advantageous) and do not invoke [67] (see Subsection 1.3 for more details).
Thus, a key point in our paper compared to the aforementioned works is that we handle the vastly more general setting of arbitrary -Anosov subgroups which turns out to have major differences compared to the setting of projective Anosov subgroups. In the general setting, there is a nontrivial family of translation flows parametrized by vectors in and we discover that we must delete an exceptional set for the expected results, i.e., exponential mixing, essential spectral gap, and prime orbit theorem with a power saving error term hold for the generic vectors in . We also obtain the precise nature of the dependence on the vectors in .
1.3. On the proof of Theorem 1.1
For general -Anosov subgroups, we have a family of translation flows on and we begin by proving that they are metric Anosov. We note that when , the space does not naturally inherit a -invariant metric from the Riemannian metric on and so the metric on needs careful treatment. Moreover, we also show that the strong (un)stable foliations in of a translation flow are projections of the horospherical foliations of (Theorem 3.2). Our proofs of these properties, following [18], use the fact that one can always construct a projective Anosov representation of using the Plücker representation and the results of Bridgeman–Canary–Labourie–Sambarino [9] on the metric Anosov property in this case. It is likely that the arguments in [9] can be extended to general Anosov subgroups to give more direct proofs of these properties but we do not do so for expedition reasons. An immediate consequence of the metric Anosov property is that there exist Markov sections for the translation flows which are compatible with the corresponding strong (un)stable foliations. The constructed framework is an important tool which do not appear in previous works and for which reason our approach is fruitful in the general -Anosov setting.
Thanks to the Markov sections, we may freely introduce symbolic dynamics and thermodynamic formalism, and use transfer operators. More specifically, we define transfer operators defined by
where is the first return time map associated to the translation flow on . We wish to then execute Dolgopyat’s method. Stoyanov [67] has done related work for Axiom A flows but we do not prove this or other strong properties and rely on weaker properties instead. In any case, we need to prove the crucial local non-integrability condition (LNIC) (Proposition 5.11). We wish to use Lie theoretic arguments for this. However, as alluded to in Subsection 1.2, we need to carefully deal with some topological issues due to the fractal nature of the limit set and the -action on not being properly discontinuous in many cases. Following [64], the relationship between the strong (un)stable foliations and horospherical foliations allows us to lift and extend the rectangles of the Markov section to open sets and use auxiliary smooth extensions of the Poincaré map and the first return time map. We then generalize the techniques in [64, 18] to prove first a smooth version and then a reverse Lipschitz version of LNIC on the original Markov section which is of fractal nature. In the process, it becomes clear that the aforementioned derivation holds for all but those in a certain exceptional set . Fortunately, Dolgopyat’s method is robust enough that the above ingredients suffice to be able to run the classical arguments (for all ).
Throughout the paper, we also take a more unified viewpoint and carefully investigate the dependence of the family of translation flows on the parameter . This requires additional important properties. The first one is the existence of a compatible family of Markov sections for all . This allows us to identify the corresponding rectangles with each other and therefore use the same domain to define the transfer operators and only vary the first return time maps which is smooth in . We also show some bounds for which is used to carry through the dependence on throughout the proofs.
1.4. Organization of the paper
Section 2 covers background on semisimple real algebraic groups, Anosov subgroups, and translation flows. In Section 3, we prove that the translation flow is metric Anosov for a suitable metric and describe the foliations in terms of the -horospherical subgroups to establish Markov sections compatible with the Lie structure. Section 4 contains the formulation of a key theorem for a Lipschitz version of Dolgopyat’s method from which exponential mixing for Lipschitz continuous functions can be derived. In Section 5 we prove the crucial LNIC. Finally, in Section 6, we prove the aforementioned key theorem. We also include some appendices after that which contain proofs of some basic facts.
Acknowledgements
We thank Hee Oh for her encouragements. We thank Rafael Potrie for useful correspondence regarding metric Anosov flows. We also thank Institut des Hautes Études Scientifiques and the Fields Institute for their hospitality and facilitating conversations. Sarkar acknowledges support by an AMS-Simons Travel Grant.
2. Anosov subgroups and their translation flows
2.1. Lie theoretic preliminaries
Let be a connected semisimple real algebraic group with identity element and Lie algebra . Let denote the Killing form and be a Cartan involution, i.e., the symmetric bilinear form defined by for all is positive definite. We write for the inner product and for the induced norm. Let be the associated eigenspace decomposition corresponding to the eigenvalues and of respectively. Then is a maximal compact subgroup. Let be a maximal abelian subalgebra and let . Identifying via the Killing form, let be the restricted root system of and be a choice of sets of positive and negative roots. Let be the corresponding closed positive Weyl chamber. We have the associated restricted root space decomposition
where is the centralizer of in .
Let denote the set of all simple roots. We denote by the root subsystem generated by any subset (it is empty if so is ). Fix a nonempty subset . Define Lie subalgebras and of by
where for the former, we take the empty intersection to be . Let , and . We denote by the interior of in the topology of . Let be the Weyl group which is a finite group. The adjoint action induces an action which we also denote by . Let denote the orthogonal projection map which is in fact the unique projection map invariant under precomposition by all elements of the Weyl group which act trivially on . Let be the longest element with respect to the generating set consisting of root reflections. Then, and the map is called the opposition involution of . The opposition involution preserves and acts on by precomposition.
For a fixed , the expanding/contracting -horospherical subgroups are
The definition is independent of the choice of and .
Let denote the standard parabolic subgroup associated to , i.e., is the normalizer of in . The Lie algebra of is given by . We denote the -Furstenberg boundary of by
Note that we also have an action induced by the left translation action. For all , we denote
There is a unique open -orbit of given by
The Levi subgroup is the centralizer of in . It satisfies , , and where is an almost direct product of a connected semisimple real algebraic group and a compact center. In all of the above and below notations, when , we omit the subscript/superscript .
Definition 2.1 (Iwasawa cocycle).
The Iwasawa cocycle is the map which assigns to a pair the unique element such that where such that .
Definition 2.2 (-Busemann function).
The -Busemann function is the map defined by
for all and , where is any lift of under the natural projection . This is well-defined, i.e., independent of the choice of [52, Lemma 6.1], -equivariant, and satisfies the identity
for all and .
Define a left -action on by
for all and . Then the stabilizer of is and we have the following diffeomorphism.
Definition 2.3 (-Hopf parameterization).
The -Hopf parameterization is the left -equivariant diffeomorphism defined by
We note that the right -action on corresponds to the -translation on the -coordinate of .
Definition 2.4 (-Gromov product).
For , the -Gromov product is defined as
which is independent of choice of .
2.2. Zariski dense discrete subgroups
Let be a Zariski dense discrete subgroup.
Definition 2.5 (-limit set).
The Jordan projection of an element is the unique element such that the hyperbolic component of the Jordan decomposition of is conjugate to .
Definition 2.6 (-limit cone).
The -limit cone of is the unique minimal closed cone in containing . It is a convex cone with nonempty interior [3].
The Cartan projection of an element is the unique element such that for all .
Definition 2.7 (-growth indicator).
The growth indicator of is defined by
with the convention that , where is any Euclidean norm on and is the abscissa of convergence of the Poincaré series
This definition does not depend on the choice of norm (though we have fixed the one induced by the Killing form). Quint [53] showed that is homogeneous of degree , concave, upper semicontinuous, and satisfies , , and . Denote by any maximal growth direction of .
The following generalization of Patterson–Sullivan measures [45, 68] is also due to Quint [52] (see also [1]).
Definition 2.8 (-conformal measure).
Let . A Borel probability measure on is called a -conformal measure if
We say that is tangent to at if and . We also denote the dual limit cone
(1) |
Theorem 2.9 (53, Theorem 8.4).
If is tangent to at , then there exists a -conformal measure.
2.3. Anosov subgroups
The notation introduced in the remainder of this section will be fixed throughout the paper. The following definition of Anosov subgroup is due to Guichard–Wienhard [25, Corollary 4.16] (see also [36, 24, 27] for other equivalent characterizations).
Definition 2.10 (Anosov subgroup).
We call a Zariski dense discrete subgroup a (-)Anosov subgroup if it is a finitely generated Gromov hyperbolic group with Gromov boundary and it admits continuous -equivariant boundary maps and such that for all .
Let be a Zariski dense -Anosov subgroup with boundary maps and . We record some basic properties in the following theorem. The first 3 properties are [9, Theorem 6.1], [46, Proposition 4.6], [25, Lemma 3.1], respectively, and the last three properties follow from [61, Theorem A] and [58, Lemma 4.8] (cf. [58, Corollary 4.9] and [60, Theorem 4.20]).
Theorem 2.11.
We have the following properties:
-
((1))
the boundary maps and are Hölder homeomorphisms onto the limit sets and , respectively;
-
((2))
the limit cone is contained in ;
-
((3))
every nontorsion element is loxodromic;
-
((4))
the growth indicator is analytic and strictly concave except along radial directions on and vertically tangent on ;
-
((5))
if is tangent to , then ;
-
((6))
if , then there exists a unique
(2) where denotes the conjugacy class of , such that is tangent to .
For , let be the projection map defined by
(3) |
The -action on descends to an action on via the projection given explicitly by
(4) |
The left -action on is not necessarily properly discontinuous. For example, when and , the -action is not properly discontinuous (see [4, 35, 19] for a general criterion) but on the other hand, when for some , there are examples where the -action is properly discontinuous (see [24, Corollaries 1.10 and 1.11]). The -action on is also not necessarily properly discontinuous (see Theorem A.2 for a self-contained construction of numerous examples). However, restricting the -action to where , we have the following theorem.
Theorem 2.12.
If , then the -action on induced by is properly discontinuous and cocompact.
In light of the above theorem, for , let
where the -action is the one induced by . The space is equipped with a Bowen–Margulis–Sullivan (BMS) measure . Let be as in Theorem 2.11(6). By [33, Theorem 1.11], there exists a unique -conformal measure (resp. -conformal measure ) on (resp. ) and moreover, (resp. ) is supported on (resp. ). We define a locally finite Borel measure on by
(5) |
where denotes the Lebesgue measure on . Observe that is left -invariant, so it descends to a probability measure on (after renormalization). By [61, Proposition 3.3.2], is the measure of maximal entropy for the translation flows on .
A consequence of Theorem 2.12 is that the restriction of the -action on to is properly discontinuous. Moreover, is a trivial -vector bundle over by a standard argument.
2.4. Translation flows
Given a vector , we have a flow on which is given by translation by in the -coordinate. The flow descends via to what we call a translation flow on given explicitly by
By [61], the topological entropy of the flow is (see Eq. 2). Note that only the value of is required to determine the translation flow on and furthermore, rescaling simply rescales the -coordinate of . We introduce the following notation to fix a family of translation flows without the redundacies from scaling or . For , set
(6) |
Note that is the unique linear form tangent to at . Then, the aforementioned redundancies are removed if is chosen so that and the topological entropy is then by Theorem 2.11(6).
We recall the relation between the translation flow and the (unique up to Hölder reparametrization (see Definition 3.9)) Gromov geodesic flow associated to .
Theorem 2.13.
For all , there exists a Hölder homemorphism conjugating to a Hölder reparametrization of .
Henceforth, fix , say . The above theorem in particular shows that for any the dynamical system is conjugate to a reparametrization of . In fact, we will see that the conjugating homeomorphism and reparametrizations are bi-Lipschitz (see Theorem 3.13). In light of this fact, we take a unified viewpoint where we have a single compact metric space
equipped with a family of pairs of BMS measures and translation flows
To distinguish the family for , we set
The translation flows are then homothety equivariant, i.e., we have
(7) |
When , Theorems 2.12 and 2.13 follow from the authors’ prequel work [18, Theorem 4.15]. Such theorems first appeared in the work of Sambarino [59, Theorem 3.2] for fundamental groups of compact negatively curved manifolds and in [9, Proposition 4.2] for projective Anosov subgroups whose generalization was outline in [13, Theorem A.2]. The proofs can be generalized for arbitrary (cf. [61] and see [33, Theorem 9.1] for a complete alternative proof).
3. Anosov property of the translation flow
The purpose of this section is to prove that the translation flow on is metric Anosov (see Definition 3.7). Before stating the precise theorem, we first define the metric on we will be using. Fix a bi-invariant Riemannian metric on the compact subgroup . Since and are quotients of , the metric on along with Euclidean metric on induces a product metric on which we restrict to . We call any metric on which is locally bi-Lipschitz equivalent to the product metric on a locally product-like metric. Construction of a -invariant locally product-like metric on can be done as in [14, Lemma 4.11] (cf. [9, Lemma 5.2]). We equip (resp. ) with any -invariant locally product-like metric (resp. ).
Remark 3.1.
When , we have the diffeomorphism where is a subgroup of . Then using a -equivariant continuous section , any left -invariant and right -invariant Riemannian metric on descends to a metric on pulls back to a locally product-like metric on .
This section is devoted to proving the following theorem.
Theorem 3.2.
The translation flow on is a metric Anosov flow with respect to the pair of foliations induced by the -horospherical foliations (see Subsection 3.2).
We obtain the following as a consequence of Theorems 3.2 and 3.8.
Corollary 3.3.
The translation flow on has a Markov section with respect to (see Definition 3.6).
3.1. Metric Anosov flows, Markov sections, and reparametrizations
In this subsection, we briefly recall the definition of metric Anosov flows and the fact that they admit Markov sections due to Pollicott [49]. The reader can also refer to [9, §3.2] and [18, §4.1] for details omitted in this subsection. We also discuss reparametrizations and give a Lipschitz criterion for when a reparametrization of a metric Anosov flow is also metric Anosov.
For this subsection, let be an arbitrary continuous flow on an arbitrary metric space . Given a foliation of , a point and , let denote the leaf through and . Given a foliation transverse to , we define the corresponding central foliation such that for all , we have if and only if for some .
Definition 3.4 (Local product structure).
A pair of foliations of is said to have local product structure if there exists such that for all , there exists a homeomorphism where is a neighborhood of such that is a chart for both and .
Let be a pair of foliations transverse to . Denote and and suppose and have local product structures given by defined on for some sufficiently small so that the image sets are always contained in a unit ball. For subsets and such that and , we call
a rectangle of size if for some , and the center of . For any rectangle , we can extend the map to defined by for all and .
Definition 3.5 (Complete set of rectangles).
A set for some consisting of rectangles of size with respect to in is called a complete set of rectangles of size if:
-
((1))
for all with ;
-
((2))
.
Let be a complete set of rectangles of size in . We introduce some notation related to . Let
Define the first return time map by
Define
(8) |
Define the Poincaré first return map by
Let be its projection where is the projection defined by for all . We define the cores
Definition 3.6 (Markov section).
We say the complete set of rectangles is a Markov section (with respect to ) if is satisfies the Markov property: and for all such that , for all .
Observe that if is a Markov section, then is constant on for all and .
Definition 3.7 (Metric Anosov flow).
The flow is said to be metric Anosov (with respect to ) if there exist , , and such that
-
((1))
and have local product structures;
-
((2))
it satisfies the Anosov property: for all , , and , we have
for all .
The existence of Markov sections for metric Anosov flows is due to Pollicott [49] and generalizes results of Bowen and Ratner [5, 54].
Theorem 3.8.
If is a metric Anosov flow with respect to , then there exists a Markov section with respect to (see Definition 3.6).
We now recall the definition of reparametrization of a flow and give a Lipschitz criterion to verify if a given reparametrization of a metric Anosov flow is also metric Anosov with respect to a given pair of foliations.
Definition 3.9 (Reparametrization).
A flow on is called a reparametrization of if it is of the form for all and , where is a continuous map satisfying
-
((1))
positivity: for all and ;
-
((2))
the cocycle condition: for all and .
In that case, is itself a reparametrization of for some continuous map which we call the inverse of and satisfies
-
((1))
,
-
((2))
for all .
We say that a reparametrization is Lipschitz (resp. Hölder) if is Lipschitz (resp. Hölder) continuous for all .
Proposition 3.10.
Let be a metric Anosov flow with respect to and with constants . Suppose that is a reparametrization of and is a pair of foliations transverse to such that and leafwise. Suppose there exists , , and such that
-
((1))
for all , , and , if and , then
-
((2))
for all and .
Then, is also metric Anosov with respect to and with constants .
Proof.
Let , , , , and be as in the proposition. It suffices to check the Anosov property for . We check it for the stable foliation and the unstable foliation is done similarly. We may assume is small enough so that by the Anosov property of , there exist and such that for all , , and we have
Then, for all , , , and we have
∎
3.2. Stable and unstable foliations for the translation flow
In this subsection, we introduce a pair of natural foliations on transverse to the translation flow. The right -orbits in give the -horospherical foliations. Recall that is the normalizer of in . In particular, normalizes both and . Then the -horospherical foliations descend to foliations of which we denote by , and its restriction to by . More explicitly,
(see [32, Lemma 7.4] for the equalities) for all . We denote the image foliations induced by and under the projection (see Eq. 3) by and and call them the strong unstable and strong stable foliations, respectively. Finally, we use the same notation for the image foliation under the vector bundle projection . The following lemma ensures that the above procedure produces well-defined foliations on , , and .
Lemma 3.11.
Let such that
Then, for all with , we have
Proof.
Let as in the lemma and suppose with . Since , there exists such that . Then, by an elementary computation, we have
Since by hypothesis, we have
Since , we also have for some . Combining with , we obtain . Since normalizes and is a direct product, it follows that
Similar to before, we have
and using the hypothesis , we conclude that . ∎
Hence, we obtain (weak/strong) (stable/unstable) foliations on as follows: for all , we have
(9) |
for any choice of such that . It is easy to see that and are transverse to the translation flow, and and have local product structures.
3.3. Projective Anosov representations
In this subsection, we recall from [9] facts about projective Anosov representations that we will need for the proof of Theorem 3.2. Let denote the Lie algebra of . Then the adjoint representation induces the representation . Define the vector space (one can see that is a vector space using the Iwasawa decomposition and the fact that is -invariant). Then, we have the induced irreducible representation , called the Plücker representation, and it induces diffeomorphisms and [38, Theorem 7.25]. By [25, Proposition 4.3]), the restriction is a projective Anosov representation in the sense of [9, Definition 2.2] and in particular, the previous diffeomorphisms restrict to -equivariant bi-Lipschitz maps and .
The -bundle
over , where is equipped with a -action and a flow that commute with each other which are given as follows: for all , , and , let
Let . Then descends to a flow on called the -geodesic flow.
Any Euclidean metric on induces a metric on and this further induces a metric on in a natural way. Any metric on obtained in this fashion is called a linear metric.
Theorem 3.12 (9, Proposition 5.7).
There exists a -invariant metric on , which is bi-Lipschitz equivalent to any linear metric, such that it descends to a metric on for which the flow is metric Anosov with respect to the pair of foliations which are defined as follows: for all ,
3.4. Proof of Theorem 3.2
Theorem 3.13.
There exists a bi-Lipschitz homeomorphism which conjugates the translation flow to a Lipschitz reparametrization of the -geodesic flow .
Proof.
We proceed in the following two steps.
Step 1: Construction of . We only give an overview (omitting details) of the construction of as it is similar to the full construction in [18, Theorem 4.15] which deals with the case . Consider the trivial -bundle
equipped with a left -action defined by
and a flow defined by
which commutes with the -action.
Let for some be a locally finite open cover of such that for all distinct and . Using a partition of unity , for some index set , subordinate to such that is smooth along the -geodesic flow for all , we can construct a -equivariant section of the form . Explicitly, we can take
Let denote the image of under the exponential map on the -coordinate. Equip with the -action and flow it inherits from and the unique -invariant bundle norm satisfying for all . Then, is an -bundle over , descends to a flow on , and descends to a norm on . Using the Morse property [21, Theorem 4.13] of Kapovich–Leeb–Porti [27, Proposition 5.16] and an analogue of Sullivan’s Shadow Lemma [32, Lemma 3.1], one can show that is discretely contracting with respect to , i.e., there exists and such that for all . By [9, Lemma 4.3], the norm on descends to a norm on such that is uniformly contracting with respect to . Note that where
Let be the -equivariant map defined by
Then, it can be shown as in [18, Theorem 4.15] that is locally Lipschitz and descends to a Lipschitz homeomorphism and the reparametrization is Lipschitz. (Here, is Lipschitz instead of Hölder as in [18] since the induced maps and are Lipschitz.)
Step 2: is bi-Lipschitz. By the above, it suffices to show that is locally Lipschitz with respect to and . Recall that is a locally bi-Lipschitz equivalent to the product metric on and is locally bi-Lipschitz equivalent to any linear metric on . Thus, it suffices to show that is locally Lipschitz with respect to the product metric on and any linear metric on . Fix a Euclidean metric on . We also use to denote the induced metrics on and . Fix a compact subset that is “radially symmetric” in the sense that if for , then as well. We need to show that there exists such that for any for , we have
For convenience, let so that
Observe that since is compact, there exists constants such that
and
where the last inequality uses the observation that for , the linear form which satisfies is smoothly determined by and the unit vector (or equivalently, ). The triangle inequality gives
Hence, it suffices to show that there exists a constant such that for all , we have
Since is uniformly contracting with respect to , there exists such that
i.e., . Then, for all , we have
Lastly, since is locally Lipschitz with respect to any linear metric, there exists such that for all and this completes the proof. ∎
The following is an immediate consequence of Theorems 3.12 and 3.13.
Proposition 3.14.
The reparametrization of is a metric Anosov flow with respect to .
Proof of Theorem 3.2.
In view of Proposition 3.14, by considering as a reparametrization of , it suffices to check Properties (1) and (2) in Proposition 3.10.
Proof of Property (1). Fix a compact fundamental domain for the -action. Fix sufficiently small so that the closed -balls in injectively project into and that the projection of the closed -neighborhood of is compact. Since is compact and is locally bi-Lipschitz equivalent to the product metric, there exists such that for all and with , we have
(10) |
Let which will be specified later. Fix , , and such that . By compactness of , we can uniformly choose sufficiently small so that . We want to show that there exists a uniform constant such that
(11) |
By (10), for each , we have
By (9), we have for a uniform constant by smoothness of the -Gromov product restricted to . Then
This establishes the right hand inequality of (11).
For the left hand inequality of (11), we observe that compactness of and the Lipschitz properties of the maps and in the proof of Theorem 3.13 implies that there exists a uniform constant such that
This establishes Property (1) for the strong unstable foliations. The argument for the strong stable foliations is similar.
Proof of Property (2). Property (2) is immediate from the fact that for all for some positive continuous function (see [18, Remark 4.11]). ∎
Remark 3.15.
Although we have only shown that the distinguished translation flow on is metric Anosov, it is not difficult to see that the above work can be adapted to show that any translation flow with on is a Lipschitz reparametrization of and also metric Anosov. Indeed, the proof of Theorems 3.13 and 3.14 is easily modified to show that for all there exists a bi-Lipschitz homeomorphism conjugating to a Lipschitz reparametrization of and that is metric Anosov. In fact, such a homeomorphism can be explicitly described as follows. Let be the projection map determined by the decomposition , be a locally finite open cover of such that for all distinct and and for some index set be a partition of unity subordinate to such that is smooth along the -geodesic flow for all and be a partition of unity subordinate to the cover . Define the map by where is given by
for all , for some . Then, it descends to the desired map .
4. Transfer operators, Dolgopyat’s method, and the proof of Theorem 1.5 and applications
In this section, we cover the necessary background for transfer operators in order to state Theorem 4.7 which is the main technical theorem regarding spectral bounds. We also outline how to derive the main theorem stated in Theorem 1.5 from Theorem 4.7. A key point in this section is that all the theorems include the precise dependence on a parameter associated to .
4.1. Reduction of Theorem 1.5 by rescaling
Due to the homothety equivariance property for the family of translation flows (see Eq. 7), it is convenient to fix a scaling for each direction in for the purpose of proving Theorem 1.5. It turns out that there is a particular scaling so that we have additional properties. Using the isomorphism induced by the inner product on , we abuse notation and identify the dual limit cone with a dual limit cone (see Eq. 1). Note that (see Theorem 2.11(2)) and Proposition B.1 implies that . For all with , we define to be the unique vector such that . Then, Eq. 6 gives
(12) |
Therefore, Theorem 1.5 follows from the following exponential mixing theorem. Theorem 1.4 also follows either directly from the following or via Theorem 1.5. Note that one can repeat a standard convolution argument as in [34, Appendix] using Lipschitz continuous bump functions to handle general -Hölder functions (see also [42, Corollary 5.2] and [62, Theorem 3.1.4]).
Theorem 4.1.
There exists a bounded elementary function with vanishing limit at such that for all open neighborhoods and with and , there exist , , a finite set which come in conjugate pairs, and a finite set of finite-rank positive bilinear forms , such that for all and , we have
The rest of the paper is devoted to the proof of the above theorem. To this end, we fix an open neighborhood henceforth and define the subset of unit vectors
Note that we have a corresponding compact subset of unit vectors
4.2. Compatible family of Markov sections
Corollary 3.3 gives a Markov section for any fixed translation flow. Recalling that we have a family of translation flows , we fix a corresponding family of Markov sections on . They have several other corresponding objects (see Subsection 3.1) all of which we denote using a superscript ‘’ on the same notations. By the same observation as in [18, §4.2] regarding Markov sections of reparametrized flows, we deduce the remarkable property that the family of Markov sections can be chosen such that for different unit vectors , the Markov sections and consist of corresponding rectangles which project to each other along the translation flow foliation while fixing their centers. Consequently, for some fixed , we denote
and the family comes equipped with a family
of compatibility maps which are Lipschitz homeomorphisms and vary smoothly in . More precisely, there exists a family of Lipschitz continuous functions which vary smoothly in such that at the centers for all and
All the properties are clear save injectivity which we now justify. For all , the map moves the points in along the translation flow foliation to which the strong stable and unstable foliations corresponding to both are transverse and hence preserves the time ordering along the translation flow foliation. Now, if for some distinct , then the time ordering would be violated since and must lie on the same translation flow orbit; implying injectivity.
We now use the compatibility maps to identify the Markov sections for all with a fixed Markov section of some size , where corresponds to , i.e., . We need to address a subtle point. For any outside of some open neighborhood of , the corresponding Markov sections need not be of size . However, apart from that, they are legitimate Markov sections and come from the local product structures on the image , where are the centers for all . Thus, we can drop all superscripts henceforth, except for which we view as a family of first return time maps on the same set ; since they still vary in , they need to be distinguished. A useful property is that they remain positive, again due to preservation of time ordering. We record this and other useful properties in the following lemma (recall Eq. 8).
Lemma 4.2.
The family is smooth in and there exists such that
-
((1))
for all and ,
-
((2))
for all .
Consequently, is bounded above by .
Proof.
Positivity of for all , as mentioned above, is clear and so we prove the other properties. We use notations from Subsection 5.1 and terminology from Subsection 4.3 to shorten the proof. Namely, we use the first return vector map descended from the one in Subsection 5.1 which is Lipschitz continuous on cylinders of length . Then, we have the identity for all . Thus, the lemma follows from this and the property that . ∎
Due to the above lemma, we conclude that the smooth family can be extended to a smooth family of uniformly essentially Lipschitz (i.e., with a uniform Lipschitz constant on cylinders of length ) nonnegative first return time maps .
Using a compatible family of Markov sections as constructed in this subsection is necessary to execute Dolgopyat’s method in a uniform fashion for the family of translation flows. Although we cannot make sense of a translation flow on or even transfer operators associated to any , the above uniform bounds and the extended family of first return time maps is used to obtain other Lipschitz bounds and a uniform version of LNIC to derive the precise form of the exponential decay rate in Theorem 4.7 for the family of transfer operators in .
4.3. Symbolic dynamics
Let be the alphabet of . Define the transition matrix by
The transition matrix is topologically mixing as a consequence of [18, Theorem 8.1], i.e., there exists such that consists only of positive entries. This was assumed in [18, §4.3] as well which is a minor inaccuracy; see Appendix C for the correction. Define the spaces of bi-infinite and infinite admissible sequences
respectively. We will use the term admissible sequences for finite sequences as well in the natural way. For any fixed , we endow with the metric defined by for all . We similarly endow with a metric which we also denote by .
Definition 4.3 (Cylinder).
For all admissible sequences of length , we define the cylinder of length to be
We denote cylinders simply by (or other typewriter style letters) when we do not need to specify the admissible sequence.
By a slight abuse of notation, let also denote the shift map on or . There exist natural continuous surjections and defined by for all and for all . Define and . Then the restrictions and are bijective and satisfy and .
We can take sufficiently close to so that and are Lipschitz continuous [6, Lemma 2.2]. Let denote the space of Lipschitz continuous functions . We use similar notations for Lipschitz function spaces with domain and other codomains. For all function spaces, we suppress the codomain when it is .
Let . Since the horospherical foliations on are smooth, we conclude that is Lipschitz continuous on cylinders of length . Then the maps and are Lipschitz continuous and hence there exist unique Lipschitz extensions which, by abuse of notation, we also denote by and , respectively.
4.4. Thermodynamics
Definition 4.4 (Pressure).
For all , called the potential, the pressure is defined by
where is the set of -invariant Borel probability measures on and is the measure theoretic entropy of with respect to .
For all , there exists a unique -invariant Borel probability measure on which attains the supremum in Definition 4.4, called the -equilibrium state [7, Theorems 2.17 and 2.20]. It satisfies [15, Corollary 3.2].
Let . Associated to , we relabel (see Eq. 2) and recall from Subsection 2.4 that where we have used Eq. 12. Thanks to Corollary 3.3, we can use [13, Theorem A.2] which states that is a measure of maximal entropy for the translation flow which attains the maximal entropy of . We will consider in particular the probability measure on with corresponding pressure [8, Proposition 3.1] (cf. [15, Theorem 4.4]), which we will denote simply by . Define and . Note that . We refer the reader to [18, §4.4] for the relation between and , which we do not require directly in this paper.
4.5. Transfer operators
Throughout the paper, we will use the notation for the complex parameter for the transfer operators.
Definition 4.5 (Transfer operators).
For all and , the transfer operator is defined by
for all and .
We recall the Ruelle–Perron–Frobenius (RPF) theorem along with the theory of Gibbs measures in this setting [7, 44]. For all and , there exist a unique positive function and a unique Borel probability measure on such that and
where is the maximal simple eigenvalue of and the rest of the spectrum of is contained in a disk of radius strictly less than . Moreover, and (see Subsection 4.4).
Let . As usual, it is convenient to normalize the transfer operators. For all define the map
For all , we define the Lipschitz continuous maps and by
and for all . For all , we define and its th iterates for all by
for all and . With this normalization, for all , the maximal simple eigenvalue of is with eigenvector . Moreover, we have .
We fix some related constants. Let . By perturbation theory of operators as in [28, Chapter 7] and [44, Proposition 4.6], we can fix such that the map defined by and the map defined by are Lipschitz uniformly in . To see uniformity in , a standard calculation using the eigenvalue equation and Lemma 4.2 gives . Similar estimates in the construction of the family of eigenvectors (see [44, Theorem 2.2]) gives the latter. Fix such that it is greater that the aforementioned Lipschitz constants and so that for all and . Again by similar estimates, we also have which tends to as tends to the boundary of . Fix
which can be chosen as some elementary function evaluated at , for all (see also [46, Lemma 4.1] for taking the supremum over ).
4.6. Dolgopyat’s Method and the proof of Theorem 4.1 and applications
Recall that denotes the space of complex Lipschitz continuous functions on . It is a Banach space with the Lipschitz seminorm and norm
for all , where denotes the usual -norm. For all , we also generalize the above norm to
for all , which will be required later.
Following Stoyanov [67, §5], we define the convenient new metric on by
We denote to be the space of -Lipschitz continuous functions and to be the -Lipschitz constant for all . We define the cones
Remark 4.6.
If , then using the convexity of , we can derive that is --Lipschitz, i.e., for all . It follows that , however the reverse containment is not true.
We now state Theorem 4.7 regarding spectral bounds of transfer operators. Such bounds are the key for obtaining exponential mixing results.
Theorem 4.7.
There exist an elementary function with vanishing limit at , , and such that for all , there exists such that for all with and , , and , we have
By a standard inductive argument (see the proof after [64, Theorem 5.4]) which goes back to Dolgopyat [22], Theorem 4.7 is a consequence of the following theorem which records the mechanism of Dolgopyat’s method.
Theorem 4.8.
There exist , an elementary function which tends to at , , , , and a set of operators , where is some finite set for all , such that
-
((1))
for all , , , and ;
-
((2))
for all , , , , and ;
-
((3))
for all , , and , if and satisfy
-
(1a)
for all ;
-
(1b)
for all ;
then there exists such that
-
(2a)
for all ;
-
(2b)
for all .
-
(1a)
Theorem 4.1 for Lipschitz continuous functions is derived from the spectral bounds of transfer operators in Theorem 4.7 using arguments of Pollicott and Paley–Wiener. The derivation is similar to that of [64, §10] and so we omit it and describe the differences. One minor difference is that in our case the derivation would be a Lipschitz version of loc. cit. (cf. [41, §9]). Another difference is that we include the Pollicott–Ruelle resonances since we cannot control simple eigenvalue of the transfer operators for the complex parameter in a fixed neighborhood of uniformly in . For this, we also refer the reader to [47]; more specifically, [47, Proposition 4], the proof of [47, Theorem 1], and the remark after [47, Corollary 1].
Let us give some more details for integrating out the stable direction in the derivation mentioned above. The constants and which come from the metric Anosov property vary smoothly in . However, the decay rate could, a priori, get worse as tends to . The exact behavior comes from the lower bound for the derivative of the reparametrization in . But this comes exactly from the upper bound for which, by construction, is bounded above by uniformly in . Thus, one can choose a decay rate uniformly in .
Remark 4.9.
We warn the reader of a subtle point. The derivation in [64, §10] crucially uses the Anosov property of the frame flow; the spectral bounds of transfer operators on their own are not enough. Analogously, although the proof of Theorem 4.7 does not require the metric Anosov property, the proof of Theorem 4.1 would be incomplete without it. Though the metric Anosov property of the translation flow is claimed in the literature on Anosov subgroups, it is far from a triviality. In our paper, the metric Anosov property was stated in Theorem 3.2 and a complete proof was provided in Subsection 3.4, thus, justifying the described derivation above of Theorem 4.1 for Lipschitz continuous functions.
Let us now describe the derivations of Theorems 1.6 and 1.8. The former is proved using the relationship between the zeros of the Selberg zeta function and the transfer operators stated in [47, Proposition 4], and of course the spectral bounds in Theorem 4.7 (cf. proof of [47, Theorem 1]). The latter is proved using the number theoretic techniques in [50] where the weaker spectral bounds in [50, Eq. (2.1)] (cf. [50, Proposition 2]) is replaced again by the stronger spectral bounds in Theorem 4.7.
As a result of our discussion, it suffices to focus only on the proof of Theorem 4.8 for the rest of the paper.
5. Local non-integrability condition
This section is devoted to a crucial ingredient for Dolgopyat’s method stated in Proposition 5.11 called the local non-integrability condition (LNIC) as in [67], which is then upgraded to Proposition 5.12. Although Proposition 5.12 is stated in reverse Lipschitz form on , we obtain it via Proposition 5.11 which requires Lie theory and hence the smooth structure on . Thus, we first resolve technical issues related to the incompatibility of with the smooth structure on .
5.1. Smooth extensions of coding maps
Fix throughout this subsection. As alluded to above, due to the fractal nature of , the Markov section does not inherit the smooth structure on . In order to circumvent this issue, we need to enlarge to an open set and smoothly extend the maps associated to the coding, , , and . This type of construction was done in [64, §5.1] based on [57, Lemma 1.2]. However, even this procedure has many more technical issues in our setting. One problem is that is not naturally situated in a larger smooth manifold with a natural extension of the translation flow. Note that in general the action is not properly discontinuous (see Theorem A.2). To avoid this problem, we work on the cover instead and use the inclusion
which is a Lipschitz embedding into a smooth manifold where the translation flow is still defined. Another problem is that it is difficult to gain control on the -orbit of the open neighborhood of in (e.g., diameter estimates, mutual disjointness, seperation distance) since we do not have a -invariant metric on and may not be properly discontinuous. To avoid these topological obstructions, we allow dependence of the open neighborhoods on the arbitrarily large length of sections of which will appear in Proposition 5.11. We do not follow [57, Lemma 1.2] because, although we believe that it is true, it would take significantly more work to properly formulate and establish the eventually contracting property of the smooth extension of on some fixed open neighborhood of . See the recent related work of Delarue–Montclair–Sanders [20] for the case that is a projective Anosov subgroup of which gives a construction of a flow-invariant smooth manifold on which the translation flow is a contact Axiom A flow.
In light of the above discussion, let us first fix a connected fundamental domain for , unique isometric lifts of for all , and write and . Denote by the unique lift of . Abusing notation, we denote the lift of cylinders by the same symbol. Since the translation flow is defined on , the maps , , and have natural lifts , , and , abusing notation. For all we define to be the unique element such that . Similarly, for all admissible pairs , we write for the unique element such that
Then, is a generating subset. For all admissible sequences for some , we extend the notation and write
(13) |
in ascending order if , and if , and .
Let , be the center, and . We have a metric on (resp. ) which is induced by (resp. ) using coordinate pullbacks and denoted by the same notation. Then, has a local product structure such that contains , for some . Now, fix open neighborhoods of and define such that consists of mutually disjoint rectangles. Define , , , and , and similarly define , , , and . We similarly omit the superscript ‘’ and use the superscript ‘’ for other sets.
Let . Recall that the translation flow is defined on and smooth. Thus, for all and , we obtain compactly contained open neighborhoods of which are decreasing in and the natural smooth injective extensions of with cylinders ; and we define and , for all admissible sequences .
There are also natural smooth extensions for all admissible pairs . We define the smooth maps by
(14) |
for all admissible sequences .
Following [18, §5] we can construct a smooth section
in a similar fashion such that:
-
((1))
for all and , and , there exists unique such that ;
-
((2))
for all and , , and , there exists unique such that .
Observe that can indeed be constructed such that it does not depend on due to the properties of the compatible Markov sections.
We introduce the first return vector map and holonomy. Although we only need the first return time map, since it does not require excess work, we provide the general definitions and the subsequent fundamental lemmas in anticipation that it will turn out useful elsewhere.
Definition 5.1 (First return vector map, Holonomy).
The first return -vector map and -holonomy are -invariant maps and respectively that associate to each for some the unique elements and which satisfy
We often drop the prefix “-”.
5.2. Local non-integrability condition
We are now ready to prove Propositions 5.11 and 5.12. The proof follows the techniques developed in [64, 17]. The version here is actually more simplified in some parts and we obtain a stronger LNIC because we are dealing with the “geodesic flow” (see Remark 5.10).
We start with a slight generalization of [64, Definition 6.1]. Recall the identifications from Subsection 4.2.
Definition 5.2 (Associated sequence in ).
Let with . Let be the center. Consider a sequence such that , , and . We define an associated sequence in to be the unique sequence where
Remark 5.3.
In the above definition, the associated sequence corresponding to each is independent of .
We continue using the notation in the above definition. Define the subsets
where the first is open and the second is compact. Now, if the above sequence corresponds to some and some such that and respectively, then we can define the map by
To view it as a function of the first coordinate for a fixed , we write .
Let be the center. Let and be an admissible sequence. Then, there exists an element which we denote by such that
This is well-defined because .
Let with . Let and be orthogonal projection maps. Let
denote the orthogonal projection map, i.e., the projection with respect to the decomposition . Similarly, let
be the Cartesian projection map of onto the factor. Note that .
Also define where is as in Subsection 5.1 and is the center.
In order to derive the LNIC in Proposition 5.11, we need the following two lemmas regarding . We omit the proofs since they are almost a verbatim repetition of that of [64, Lemmas 6.2 and 6.3].
Lemma 5.4.
Let , be an admissible sequence, and . Let and such that where is the center. Then, we have
In particular, for all with , we have
Lemma 5.5.
For all , we have
and in particular, for all with , we have
where is a diffeomorphism onto its image which is also smooth in and satisfies and .
For all , let be the corresponding root direction, i.e., . The following lemma is more or less a standard Lie algebra fact. We give a proof for the sake of completeness and for comparison with the analogous statement in [17, Proposition 4.4]. Using loc. cit. more directions can be produced (including those in ) whereas our lemma below gives the stronger quantifier “for all and ” (cf. Remark 5.10).
Lemma 5.6.
Let . For all nonzero , there exists a nonzero such that .
Proof.
Let and be nonzero. We will show that satisfies the lemma. Indeed because for all , we have
Thus, . Moreover, we have
which implies . Hence, . We use the symbol to mean proportionality with a positive constant. Recalling that , for all , we have
which implies since is positive definite. ∎
Repeating an analogous proof for arbitrary , we have the following generalization of [23, Lemma 2.11] which is itself a generalization of [70, Proposition 3.12]. Alternatively, [29, Theorem 1.1] (which is a stronger measure theoretic statement but for subvarieties) also suffices for our purposes in the proof of Lemma 5.8.
Lemma 5.7.
For any open subset with , the intersection is not contained in any smooth submanifold of of lower dimension.
We write the next lemma and its corollary in its most general form, though we apply it for . Let with . We define
Since is always nonempty, if , then we trivially have . Morevoer, implies and .
Lemma 5.8.
There exists such that the following holds. Let . For all nonzero , there exists such that
(15) |
Proof.
By compactness of and continuity of in for all , it suffices to show that for all and nonzero , there exists such that . Let be nonzero. Let and suppose for the sake of contradiction that . Without loss of generality, we may assume that at the center . Consider the smooth map
Then, we have for all .
Note that for all . Let for some be a nonzero component in the decomposition of according to . By Lemma 5.6, there exists such that . Using the definition of , we have . Then so there is a neighborhood containing such that is a smooth submanifold (in fact a subvariety) of of strictly lower dimension. However, , but on the other hand contains for some open set . It follows via the diffeomorphism that is contained in a smooth submanifold of of strictly lower dimension, contradicting Lemma 5.7. ∎
For any normed vector space , let denote the unit sphere in centered at . Since Eq. 15 is an open condition and is compact, a straightforward compactness argument gives the following corollary.
Corollary 5.9.
There exists such that the following holds. Let . There exist nontrivial for some and such that if with for all , then for all , there exists such that
Remark 5.10.
It is important to note that although Corollary 5.9 is analogous to [64, Lemma 6.4], when projected to for both results, the former has a stronger quantifier “for all ”. As a result, Proposition 5.11 also has a stronger quantifier and is hence stronger than the projection of [64, Proposition 6.5]. Note that the stronger proposition is possible only for the “geodesic flow” and is ultimately rooted in Lemma 5.6. Due to this stronger proposition, we do not require the non-concentration property as in [64]. At a conceptual level, this is similar to what happens in [67].
With Lemmas 5.5 and 5.9 in hand, the proof of Proposition 5.11 is nearly the same as that of [64, Proposition 6.5]. Note that in the proposition, the open neighborhood can be chosen to be convex simply by shrinking it if necessary to an appropriate open ball. Also, its dependence on is not an issue since the proposition can be upgraded as discussed below.
Proposition 5.11 (LNIC).
There exist , , and such that for all , there exists a convex open neighborhood of the center such that there exist sections for some mutually distinct admissible sequences for all integers such that for all , , and , there exist such that
where we define by
and we denote for all and .
We immediately derive the following corollary from Proposition 5.11 first on the open neighborhood of the center for a given since is convex and then upgrade it to using the topological mixing property of the transition matrix (see Subsection 4.3). The mutual disjointness in Proposition 5.12 follows from mutual distinctness of the admissible sequences in Proposition 5.11.
Proposition 5.12.
There exist , an unbounded subset , and such that for all , there exists a set of Lipschitz sections of such that for all and , there exist such that
Moreover, are mutually disjoint.
Fix , , and to be the ones provided by Proposition 5.12 henceforth.
6. Dolgopyat operators and the proof of Theorem 4.8
In this section we carry out Dolgopyat’s method to prove Theorem 4.8. Many of the techniques used here go all the way back to Dolgopyat [22]. Since it is sensitive to the setting at hand, some of the arguments require careful treatment in each instance. Over the years, it has been made cleaner and more efficient in some settings such as by Stoyanov [67].
Thanks to Remark 5.10, we are able to follow Stoyanov’s strategy using cylinders and the new distance . One especially important advantage of Stoyanov’s version of Dolgopyat’s method is that we do not require the Federer/doubling property. This is not just a convenience—there is no appropriate metric to use on which is also compatible with and a doubling property is not known yet for PS measures in higher rank to the best of the authors’ knowledge. The reader may also find it useful to consult [43, 62, 63] where many full proofs are provided in a similar framework.
We fix for the rest of the section. Recall that this implies and hence, Proposition 5.12 applies in this section.
6.1. Preliminary lemmas and constants
We obtain the following eventually contracting property, which also implies the same for the new distance function , from the metric Anosov property of the translation flow in Theorem 3.2, and a compactness argument for the lower bound. Another lemma follows from it as in [67, Proposition 3.3].
Lemma 6.1.
There exist and such that for all cylinders with , and , we have both
Lemma 6.2.
There exist and such that for all cylinders with and subcylinders with and , we have
We fix constants , , , , and provided by Lemmas 6.1 and 6.2 henceforth. We use these bounds extensively without comments. We also fix such that
(16) |
The following is a Lasota–Yorke [37] type of lemma whose proof is similar to that of [67, Lemma 5.4] and [43, Lemma 3.9]. Its proof gives the explicit constant which we fix henceforth.
Lemma 6.3.
There exists such that for all , , and , we have
-
((1))
if (resp. ) for some , then (resp. ) for ;
-
((2))
if and satisfy
for all , for all , for some , then for all , for all , we have
for all .
Now, we fix and some other significant positive constants
(17) | ||||
(18) | ||||
(19) | ||||
(20) |
Having fixed , we also fix the corresponding set of Lipschitz sections of provided by Proposition 5.12.
6.2. Construction of Dolgopyat operators
Let . We define the set for some consisting of maximal cylinders with so that . We define the set for some consisting of subcylinders with and . We define the index set . We define which satisfies for all distinct . For all , we define the function
We record a number of basic facts derived from Lemmas 6.1 and 6.2.
-
—
For all and , we have the diameter bounds:
(21) (22) (23) -
—
For all , we have with Lipschitz constant (cf. [67, Lemma 5.2]):
(24) -
—
For all , if , then
(25)
Definition 6.4.
For all with and , and , we define the Dolgopyat operators by
for all .
Definition 6.5.
For all , a subset is said to be dense if for all , there exists such that . Denote to be the set of all dense subsets of .
6.3. Proof of Theorem 4.8
Properties (1) and (3)(3)(2b) in Theorem 4.8 are derived from Lemma 6.3 using estimates from Eqs. 17, 19, and 20. We omit the proofs of since they are almost identical to those in [67, 43, 62, 63].
Similarly, we also omit the proof of Property (2) in Theorem 4.8 and refer the reader to [67, 62, 63]. However, we mention that the proof of Property (2) uses the Gibbs property of the measure which is automatically satisfied since it comes from a equilibrium state. This replaces the role of the Federer/doubling property.
For the sake of completeness, we include the proof of Property (3)(3)(2a) in Theorem 4.8 where the crucial LNIC from Proposition 5.12 is used. Here we use .
Lemma 6.6.
Let . Suppose for some and such that for some and . Then, there exists such that
for all and .
Proof.
Let , , , and be as in the lemma. The upper bound always holds on by Eq. 18. Recall that and hence in this section. Thus, the lemma follows from Proposition 5.12 and the estimate
using Eq. 16. ∎
Now, for all with and , for all integers , for all and for all , we define the functions by
for all . We need another lemma before proving Property (3)(3)(2a) in Theorem 4.8. See [67, 62, 63, 43] for a proof.
Lemma 6.7.
Let . Suppose and satisfy Properties (3)(3)(1a) and (3)(3)(1b) in Theorem 4.8. Then for all , we have
and also either of the alternatives
-
(1)
for all ,
-
(2)
for all .
For any , let denote the angle between and viewed as vectors in .
Lemma 6.8.
Suppose such that and for some and . Then we have
Lemma 6.9.
Let with and . Suppose and satisfy Properties (3)(3)(1a) and (3)(3)(1b) in Theorem 4.8. For all integers , there exists such that and such that or for all .
Proof.
Let , , , and be as in the lemma. Suppose Alternative (1) in Lemma 6.7 holds for some . Then it is a straightforward calculation to check that for all , using Eq. 20. Otherwise, Alternative (2) in Lemma 6.7 holds for all . Choose satisfying the hypotheses in Lemma 6.6 and satisfying the conclusion of Lemma 6.6. Let and . Note that for all . We would like to apply Lemma 6.8 but first we need to establish bounds on relative angle and relative size. We start with the former. For all , let such that . Then recalling the hypotheses for and , and Eqs. 25 and 19, for all , we have
Using some elementary geometry, the above shows that with , for all and hence,
for all . For notational convenience, we define by
By Lemma 6.6, we have . The second inequality is to ensure that we take the correct branch of angle in the following calculations. We will use these bounds to obtain a lower bound for or where we define
Using the triangle inequality and previously computed bounds, we have
and hence,
for all and . Without loss of generality, we may assume that for all , which establishes the required bound on relative angle. For the bound on relative size, let such that for some . Then by Lemma 6.7, we have
for all , which establishes the required bound on relative size. Now applying Lemmas 6.8 and 20 and on gives or for all . ∎
Lemma 6.10.
There exists such that for all with and , if and satisfy Properties (3)(3)(1a) and (3)(3)(1b) in Theorem 4.8, then there exists such that
for all .
Appendix A Failure of proper discontinuity of
In Theorem A.2, we exhibit numerous instances where the action is not properly discontinuous. We remind the reader that the action is of course properly discontinuous whenever is of rank one, in which case , is compact, and is convex cocompact.
We will use the following lemma. Recall the Weyl group . Note that for any loxodromic element for some , its set of fixed points is among which the attracting and repelling ones are and .
Lemma A.1.
Let be a semisimple real algebraic group. Let be a loxodromic element so that for some . Let be an element in the Weyl group. Then
Proof.
Let and be as in the lemma. By -equivariance, it suffices to show that for all . Indeed, we calculate that
and so by definition, . ∎
Theorem A.2.
There exist infinitely many triples where is a connected semisimple real algebraic group of rank at least , is a Zariski dense -Anosov subgroup for some subset with , and , such that the action defined in Eq. 4 is not properly discontinuous.
Proof.
Let be any rank one connected simple real algebraic group (of which there are infinitely many), and be any convex cocompact subgroup. We simply write for the representation obtained by restricting the identity map . Take a different convex cocompact discrete faithful representation and consider the induced representation
Take the rank two connected semisimple real algebraic group . Denote by and the Weyl group and the Furstenberg boundary associated to , and by and the same for , respectively. Note that . Take the discrete subgroup which is the self-joining of via . It is known that such discrete subgroups are -Anosov either by checking that any orbit of gives a quasiisometric embedding of or by checking the antipodality condition (in the current setting, ). Hence, is also -Anosov for any nonempty subset . We also assume that is Zariski dense (cf. [30, Lemma 2.2] for an equivalent criterion). Denote by the limit set of and let be the induced (continuous) boundary map.
Now, take any loxodromic element for some , so that for some . Write for the attracting and repelling fixed points of , respectively, which are distinct. Then, the set of fixed points of is
among which the attracting and repelling ones are . Take any Weyl group element , say for the sake of concreteness, and also the fixed point , where since . By Lemma A.1, we have
Now, simply choose with so that
(26) |
Take any such that , say . Now, by compactness of , we could pass to subsequences of to get such that converges; however, for our choice of , the original sequence is already just a constant sequence. This imples the easier property that is not properly discontinuous. Using Eq. 26 as well, we have for all , so we similarly conclude that is a constant sequence and hence converges. Thus, is not properly discontinuous. ∎
Appendix B Non-obtuseness of Weyl chambers
We give a quick proof of a fundamental proposition below regarding angles in Weyl chambers. Although a stronger statement holds regarding roots, the proposition below suffices for our purposes. We refer the reader to the concise book of Serre [66, Chapter V] for the basics of root systems used here. In fact the proof given here is along the lines of [66, Chapter V, §7, Proposition 3].
We prefer to work on and so we endow it with the inner product induced by . Let be the closed positive Weyl chamber corresponding to . We write for the symmetry associated to the root . It follows from the definition of root systems that for all , the number is an integer such that is also a root.
Proposition B.1.
We have for all . Equivalently, if , then for all .
Proof.
Suppose that . Then, there exist two distinct simple roots and for the sake of contradiction, suppose that . Then and . It is a standard fact of root systems that . Now, by our initial assumption, the product cannot vanish. Next, by the Cauchy–Schwarz inequality, if the product is , then and are collinear in which case or by integrality of and , contradicting simplicity of and . Thus, . In any case, integrality again implies that or . Therefore, either or , both of which contradict simplicity of and once again. ∎
Appendix C A minor correction
We take the opportunity to make a minor correction in our prequel [18]. In [18, §4.3], the transition matrix is not topologically mixing a priori. However, it is topologically transitive (also called irreducible) as a consequence of the density result in [18, Proposition 6.3] which only uses ergodicity from [18, Proposition 3.7] due to Lee–Oh. One can then arrange the Markov section, by inserting extra rectangles if necessary as in the beginning of the proof of [18, Theorem 6.10], so that coprime primitive periods exist for the Poincaré first return map, implying that is aperiodic. Then, towards the end of the proof of [18, Theorem 6.10], “for any ” should be replaced with “for some coprime ” and again we conclude .
With the above correction, the topological mixing result in [18, Theorem 8.1] is valid. Only now, one derives as a consequence that is indeed topologically mixing.
References
- Alb [99] P. Albuquerque. Patterson-Sullivan theory in higher rank symmetric spaces. Geom. Funct. Anal., 9(1):1–28, 1999.
- Bar [10] Thierry Barbot. Three-dimensional Anosov flag manifolds. Geom. Topol., 14(1):153–191, 2010.
- Ben [97] Y. Benoist. Propriétés asymptotiques des groupes linéaires. Geom. Funct. Anal., 7(1):1–47, 1997.
- Ben [96] Yves Benoist. Actions propres sur les espaces homogènes réductifs. Ann. of Math. (2), 144(2):315–347, 1996.
- Bow [70] Rufus Bowen. Markov partitions for Axiom diffeomorphisms. Amer. J. Math., 92:725–747, 1970.
- Bow [73] Rufus Bowen. Symbolic dynamics for hyperbolic flows. Amer. J. Math., 95:429–460, 1973.
- Bow [08] Rufus Bowen. Equilibrium states and the ergodic theory of Anosov diffeomorphisms, volume 470 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, revised edition, 2008. With a preface by David Ruelle, Edited by Jean-René Chazottes.
- BR [75] Rufus Bowen and David Ruelle. The ergodic theory of Axiom A flows. Invent. Math., 29(3):181–202, 1975.
- BCLS [15] Martin Bridgeman, Richard Canary, François Labourie, and Andres Sambarino. The pressure metric for Anosov representations. Geom. Funct. Anal., 25(4):1089–1179, 2015.
- BILW [05] Marc Burger, Alessandra Iozzi, François Labourie, and Anna Wienhard. Maximal representations of surface groups: symplectic Anosov structures. Pure Appl. Math. Q., 1(3):543–590, 2005.
- BIW [24] Marc Burger, Alessandra Iozzi, and Anna. Wienhard. Maximal representations and Anosov structures. 2024. In preparation.
- BLLO [23] Marc Burger, Or Landesberg, Minju Lee, and Hee Oh. The Hopf-Tsuji-Sullivan dichotomy in higher rank and applications to Anosov subgroups. J. Mod. Dyn., 19:301–330, 2023.
- Car [23] León Carvajales. Growth of quadratic forms under Anosov subgroups. Int. Math. Res. Not. IMRN, (1):785–854, 2023.
- Cha [94] Christophe Champetier. Petite simplification dans les groupes hyperboliques. Ann. Fac. Sci. Toulouse Math. (6), 3(2):161–221, 1994.
- Che [02] N. Chernov. Invariant measures for hyperbolic dynamical systems. In Handbook of dynamical systems, Vol. 1A, pages 321–407. North-Holland, Amsterdam, 2002.
- CF [23] Michael Chow and Elijah Fromm. Joint equidistribution of maximal flat cylinders and holonomies for anosov homogeneous spaces. 2023. Preprint, arXiv:2305.03590.
- CS [22] Michael Chow and Pratyush Sarkar. Exponential mixing of frame flows for convex cocompact locally symmetric spaces. 2022. Preprint, arXiv:2211.14737.
- CS [23] Michael Chow and Pratyush Sarkar. Local mixing of one-parameter diagonal flows on Anosov homogeneous spaces. Int. Math. Res. Not. IMRN, (18):15834–15895, 2023.
- DGK [17] Jeffrey Danciger, François Guéritaud, and Fanny Kassel. Convex cocompact actions in real projective geometry. Ann. Sci. École Norm. Sup., to appear, 2017. Preprint, arXiv:1704.08711.
- DMS [24] Benjamin Delarue, Daniel Monclair, and Andrew Sanders. Locally homogeneous Axiom A flows I: projective Anosov subgroups and exponential mixing. 2024. Preprint, arXiv:2403.14257.
- DKO [24] Subhadip Dey, Dongryul M. Kim, and Hee Oh. Ahlfors regularity of patterson-sullivan measures of anosov groups and applications. 2024. Preprint, arXiv:2401.12398.
- Dol [98] Dmitry Dolgopyat. On decay of correlations in Anosov flows. Ann. of Math. (2), 147(2):357–390, 1998.
- ELO [23] Samuel Edwards, Minju Lee, and Hee Oh. Anosov groups: local mixing, counting and equidistribution. Geom. Topol., 27(2):513–573, 2023.
- GGKW [17] François Guéritaud, Olivier Guichard, Fanny Kassel, and Anna Wienhard. Anosov representations and proper actions. Geom. Topol., 21(1):485–584, 2017.
- GW [12] Olivier Guichard and Anna Wienhard. Anosov representations: domains of discontinuity and applications. Invent. Math., 190(2):357–438, 2012.
- Hej [76] Dennis A. Hejhal. The Selberg trace formula for . Vol. I, volume Vol. 548 of Lecture Notes in Mathematics. Springer-Verlag, Berlin-New York, 1976.
- KLP [17] Michael Kapovich, Bernhard Leeb, and Joan Porti. Anosov subgroups: dynamical and geometric characterizations. Eur. J. Math., 3(4):808–898, 2017.
- Kat [95] Tosio Kato. Perturbation theory for linear operators. Classics in Mathematics. Springer-Verlag, Berlin, 1995. Reprint of the 1980 edition.
- [29] Dongryul M. Kim and Hee Oh. Non-concentration property of patterson-sullivan measures for anosov subgroups. 2023. Preprint, arXiv:2304.14911.
- [30] Dongryul M. Kim and Hee Oh. Rigidity of kleinian groups via self-joinings: measure theoretic criterion. 2023. Preprint, arXiv:2302.03552.
- KO [24] Dongryul M. Kim and Hee Oh. Relatively anosov groups: finiteness, measure of maximal entropy, and reparameterization. 2024. Preprint, arXiv:2404.09745.
- [32] Dongryul M. Kim, Hee Oh, and Yahui Wang. Ergodic dichotomoy for subspace flows in higher rank. 2023. Preprint, arXiv:2310.19976.
- [33] Dongryul M. Kim, Hee Oh, and Yahui Wang. Properly discontinuous actions, growth indicators and conformal measures for transverse subgroups. 2023. Preprint, arXiv:2306.06846.
- KM [96] D. Y. Kleinbock and G. A. Margulis. Bounded orbits of nonquasiunipotent flows on homogeneous spaces. In Sinaĭ’s Moscow Seminar on Dynamical Systems, volume 171 of Amer. Math. Soc. Transl. Ser. 2, pages 141–172. Amer. Math. Soc., Providence, RI, 1996.
- Kob [96] Toshiyuki Kobayashi. Criterion for proper actions on homogeneous spaces of reductive groups. J. Lie Theory, 6(2):147–163, 1996.
- Lab [06] François Labourie. Anosov flows, surface groups and curves in projective space. Invent. Math., 165(1):51–114, 2006.
- LY [73] A. Lasota and James A. Yorke. On the existence of invariant measures for piecewise monotonic transformations. Trans. Amer. Math. Soc., 186:481–488 (1974), 1973.
- Lee [13] John M. Lee. Introduction to smooth manifolds, volume 218 of Graduate Texts in Mathematics. Springer, New York, second edition, 2013.
- [39] Minju Lee and Hee Oh. Ergodic decompositions of geometric measures on anosov homogeneous spaces. arXiv:2010.11337, 2020.
- [40] Minju Lee and Hee Oh. Invariant measures for horospherical actions and anosov groups. arXiv:2008.05296, 2020.
- LPS [23] Jialun Li, Wenyu Pan, and Pratyush Sarkar. Exponential mixing of frame flows for geometrically finite hyperbolic manifolds. 2023. Preprint, arXiv:2302.03798.
- MW [12] François Maucourant and Barak Weiss. Lattice actions on the plane revisited. Geom. Dedicata, 157:1–21, 2012.
- OW [16] Hee Oh and Dale Winter. Uniform exponential mixing and resonance free regions for convex cocompact congruence subgroups of . J. Amer. Math. Soc., 29(4):1069–1115, 2016.
- PP [90] William Parry and Mark Pollicott. Zeta functions and the periodic orbit structure of hyperbolic dynamics. Astérisque, (187-188):268, 1990.
- Pat [76] S. J. Patterson. The limit set of a Fuchsian group. Acta Math., 136(3-4):241–273, 1976.
- PS [16] Vesselin Petkov and Luchezar Stoyanov. Ruelle transfer operators with two complex parameters and applications. Discrete Contin. Dyn. Syst., 36(11):6413–6451, 2016.
- Pol [85] Mark Pollicott. On the rate of mixing of Axiom A flows. Invent. Math., 81(3):413–426, 1985.
- Pol [86] Mark Pollicott. Meromorphic extensions of generalised zeta functions. Invent. Math., 85(1):147–164, 1986.
- Pol [87] Mark Pollicott. Symbolic dynamics for Smale flows. Amer. J. Math., 109(1):183–200, 1987.
- PS [01] Mark Pollicott and Richard Sharp. Error terms for closed orbits of hyperbolic flows. Ergodic Theory Dynam. Systems, 21(2):545–562, 2001.
- PS [24] Mark Pollicott and Richard Sharp. Zeta functions in higher Teichmüller theory. Math. Z., 306(3):Paper No. 37, 20, 2024.
- [52] J.-F. Quint. Mesures de Patterson-Sullivan en rang supérieur. Geom. Funct. Anal., 12(4):776–809, 2002.
- [53] Jean-François Quint. Divergence exponentielle des sous-groupes discrets en rang supérieur. Comment. Math. Helv., 77(3):563–608, 2002.
- Rat [73] M. Ratner. Markov partitions for Anosov flows on -dimensional manifolds. Israel J. Math., 15:92–114, 1973.
- Rue [87] D. Ruelle. Resonances for Axiom flows. J. Differential Geom., 25(1):99–116, 1987.
- Rue [86] David Ruelle. Resonances of chaotic dynamical systems. Phys. Rev. Lett., 56(5):405–407, 1986.
- Rue [89] David Ruelle. The thermodynamic formalism for expanding maps. Comm. Math. Phys., 125(2):239–262, 1989.
- [58] A. Sambarino. Hyperconvex representations and exponential growth. Ergodic Theory Dynam. Systems, 34(3):986–1010, 2014.
- [59] Andrés Sambarino. Quantitative properties of convex representations. Comment. Math. Helv., 89(2):443–488, 2014.
- Sam [15] Andrés Sambarino. The orbital counting problem for hyperconvex representations. Ann. Inst. Fourier (Grenoble), 65(4):1755–1797, 2015.
- Sam [24] Andrés Sambarino. A report on an ergodic dichotomy. Ergodic Theory Dynam. Systems, 44(1):236–289, 2024.
- Sar [19] Pratyush Sarkar. Uniform exponential mixing for congruence covers of convex cocompact hyperbolic manifolds. 2019. Preprint, arXiv:1903.00825.
- Sar [23] Pratyush Sarkar. Congruence counting in schottky and continued fractions semigroups of . 2023. Preprint, arXiv:2108.00545.
- SW [21] Pratyush Sarkar and Dale Winter. Exponential mixing of frame flows for convex cocompact hyperbolic manifolds. Compos. Math., 157(12):2585–2634, 2021.
- Sel [56] A. Selberg. Harmonic analysis and discontinuous groups in weakly symmetric Riemannian spaces with applications to Dirichlet series. J. Indian Math. Soc. (N.S.), 20:47–87, 1956.
- Ser [01] Jean-Pierre Serre. Complex semisimple Lie algebras. Springer Monographs in Mathematics. Springer-Verlag, Berlin, 2001. Translated from the French by G. A. Jones, Reprint of the 1987 edition.
- Sto [11] Luchezar Stoyanov. Spectra of Ruelle transfer operators for axiom A flows. Nonlinearity, 24(4):1089–1120, 2011.
- Sul [79] Dennis Sullivan. The density at infinity of a discrete group of hyperbolic motions. Inst. Hautes Études Sci. Publ. Math., (50):171–202, 1979.
- Thi [09] Xavier Thirion. Propriétés de mélange du flot des chambres de Weyl des groupes de ping-pong. Bull. Soc. Math. France, 137(3):387–421, 2009.
- Win [15] Dale Winter. Mixing of frame flow for rank one locally symmetric spaces and measure classification. Israel J. Math., 210(1):467–507, 2015.