diff --git "a/7tE4T4oBgHgl3EQfcwyw/content/tmp_files/2301.05086v1.pdf.txt" "b/7tE4T4oBgHgl3EQfcwyw/content/tmp_files/2301.05086v1.pdf.txt" new file mode 100644--- /dev/null +++ "b/7tE4T4oBgHgl3EQfcwyw/content/tmp_files/2301.05086v1.pdf.txt" @@ -0,0 +1,3933 @@ +arXiv:2301.05086v1 [math.GR] 12 Jan 2023 +Sublinear Rigidity of Lattices in Semisimple Lie Groups +Ido Grayevsky +Abstract +Let G be a real centre-free semisimple Lie group without compact factors. I prove that irreducible +lattices in G are rigid under two types of sublinear distortions. +I show that if Λ ≤ G is a discrete +subgroup that sublinearly covers a lattice, then Λ is itself a lattice. I use this result to prove that the +class of lattices in groups that do not admit R-rank 1 factors is SBE complete: if Λ is an abstract +finitely generated group that is Sublinearly BiLipschitz Equivalent (SBE) to a lattice in G, then Λ can +be homomorphically mapped into G with finite kernel and image a lattice in G. This generalizes the well +known quasi-isometric completeness of lattices in semisimple Lie groups. +1 +Introduction +The quasi-isometric rigidity and classification of irreducible lattices in semisimple Lie groups was established +in the 1990’s by the accumulated work of many authors - Pansu [48], Schwartz [52], Kleiner-Leeb [32], +Eskin [22], Drut¸u [16], to name a few which are closely related to this paper. See [23] for a concise survey of +the following result. +Theorem 1.1 (Quasi-Isometric Completeness, Theorem I in [23]). Let G be a real finite-centre semisimple +Lie group without compact factors, Γ ≤ G an irreducible lattice and Λ an abstract finitely generated group. +If Γ and Λ are quasi-isometric, then there is a group homomorphism Φ : Λ → G with finite kernel whose +image Λ′ := Φ(Λ) is a lattice in G. Put differently, there is a lattice Λ′ ≤ G and a finite subgroup F ≤ Λ +such that the sequence +1 → F → Λ → Λ′ → 1 +is exact. Moreover, Λ′ is uniform if and only if Γ is uniform. +In the standard terminology of metric rigidity, Theorem 1.1 states that the classes of uniform and non- +uniform lattices in a group G as in the statement are quasi-isometrically complete. The main goal of the +current work is to generalize this result to a sublinear setting. Sublinear BiLipschitz Equivalences (SBE) are +a sublinear generalization of quasi-isometries, brought forward by Cornulier [13] in the past decade or so. A +function u : R≥0 → R≥1 is sublinear if limr→∞ +u(r) +r += 0. For a, b ∈ R≥0 denote a ∨ b : max{a, b}, and for a +pointed metric space (X, x0, dX) and x, x1, x2 ∈ X, let |x|X := dX(x, x0), |x1 − x2|X := dX(x1, x2). +Definition 1.2. Let (X, dX, x0), (Y, dY , y0) be pointed metric spaces, L ∈ R>0 a constant, u : R≥0 → R≥1 +a sublinear function. A map f : X → Y is an (L, u)-SBE if the following conditions are satisfied: +1. f(x0) = y0, +2. ∀x1, x2 ∈ X, +1 +L|x1 − x2|X − u +� +|x1|X ∨ |x2|X +� +≤ |f(x1) − f(x2)|Y ≤ L|x1 − x2|X + u +� +|x1|X ∨ |x2|X +� +, +3. ∀y ∈ Y ∃x ∈ X such that |y − f(x)|Y ≤ u(|y|Y ). +1 + +A map is an SBE if it is an (L, u)-SBE for some L and u as above. +I prove SBE-completeness for irreducible lattices in groups without R-rank 1 factors. +Theorem 1.3 (SBE-Completeness). Let G be a real centre-free semisimple Lie group without compact or +R-rank 1 factors. +Let Γ ≤ G be an irreducible lattice and Λ an abstract finitely generated group, both +considered as metric spaces with some word metric. Assume there is an (L, u)-SBE f : Λ → Γ with u a +subadditive sublinear function. Then there is a group homomorphism Φ : Λ → G with finite kernel whose +image Λ′ := Φ(Λ) is a lattice in G. Moreover, Λ′ is uniform if and only if Γ is uniform. +The main ingredients in the proof are of independent interest: the first is geometric rigidity for the +corresponding symmetric space, stating that every self SBE of such a space is sublinearly close to an isom- +etry. This generalizes Kleiner and Leeb’s result [32] on self quasi-isometries of symmetric spaces, as well as +Eskin’s [22] and Drutu’s [16] results on self quasi-isometries of non-uniform lattices. For the definition of the +‘compact core’ of a lattice see Section 2.2.2. +Theorem 1.4 (Sublinear Geometric Rigidity). Let X be a symmetric space of noncompact type without +R-rank 1 factors, Γ ≤ Isom(X) an irreducible lattice and X0 ⊂ X the compact core of Γ in X. For any +(L, u)-SBE map f : X0 → X0 there is a sublinear function v = v(L, u) and an isometry g : X → X such +that d +� +q(x), g(x) +� +≤ v(|x|) for all x ∈ X0. +The proof of Theorem 1.3 actually requires a stronger version of Theorem 1.4, formulated in Lemma 6.24. +Notice that Theorem 1.3 and Theorem 1.4 both hold for uniform as well as non-uniform lattices. I remark +that ‘generalized quasi-isometries’ already appeared in the context of geometric rigidity, as a technical tool +in Eskin and Farb’s work [21] [22] on quasi-isometries. Indeed much of their work is carried for maps which +are even more general than SBE. It seems however that their approach cannot yield a sublinear bound as in +Theorem 1.4, which is necessary for the proof of Theorem 1.3. +The second ingredient is a property I call sublinear rigidity, stating that a discrete subgroup Λ ≤ G which +sublinearly covers a lattice is itself a lattice. Sublinear rigidity holds for groups of any R-rank, and is the +cornerstone of this work. Its proof contains the bulk of the original ideas that appear in this paper. +Definition 1.5. For a function u : R≥0 → R>0 and a subset Y ⊂ X, define the u-neighbourhood of Y to be +Nu(Y ) := {x ∈ X | d(x, Y ) ≤ u(|x|)} +A subset Y ⊂ X is said to sublinearly cover Z ⊂ X if Z ⊂ Nu(Y ) for some sublinear function u. +For the definition of a Q-rank 1 lattice, see Section 2.1 and Theorem 2.20. +Theorem 1.6 (Sublinear Rigidity). Let G be a real centre-free semisimple Lie group without compact factors. +Let Γ ≤ G be an irreducible lattice, Λ ≤ G a discrete subgroup that sublinearly covers Γ. If Γ is of Q-rank 1, +assume further that Λ is irreducible. Then Λ is a lattice in G. +The notion of irreducibility in the non-standard context of a general (non-lattice) subgroup is explained +Section 4.4.2 (see Definition 4.47). Sublinear neighbourhoods arise naturally in the presence of SBE maps: +the essential difference between a quasi-isometry and an SBE is that ‘far away in the space’, the ‘additive’ +error term of an SBE gets larger and larger. One is led to consider metric neighbourhoods that grow - +sublinearly, yet unboundedly - with the distance to some (arbitrary) fixed base point. +The hypothesis that u is sublinear is optimal in the sense that u could not be taken to be an arbitrary +linear function. Indeed, the geometric meaning of Γ ⊂ Nu(Λ) is that for every element γ ∈ Γ, the ball +BG +� +γ, u(|γ|) +� +‘of sublinear radius about γ’ must intersect Λ. +Observe that if f is the identity function +f(r) = r, then by definition f(|g|) = f +� +d(g, eG) +� += d(g, eG). In particular G lies in the f-neighbourhood of +the trivial subgroup which is, after all, not a lattice in G. For uniform lattices and for lattices with Kazhdan’s +property (T) one can however relax the assumption of sublinearity: +2 + +Theorem 1.7 (Theorems 3.3 and 5.2). Let G be a real centre-free semisimple Lie group without compact +factors, Γ ≤ G an irreducible lattice, Λ ≤ G a discrete subgroup. Fix ε > 0 and assume Γ ⊂ Nu(Λ) for the +function u(r) = ε · r. +1. If Γ is uniform and ε < 1, then Λ is a uniform lattice. +2. If Γ has Kazhdan’s property (T) then there is a constant ε(G) such that if ε < ε(G) then Λ is a lattice. +Theorem 1.6 generalizes the case where Γ lies in a bounded neighbourhood ND(Λ) for some D > 0, proved +by Eskin and Schwartz in a slightly modified version (see Section 4.3 below). In the bounded case, the result +is much stronger and states that Λ must be commensurable to Γ (except in groups locally isomorphic to +SL2(R), see [52]). In the sublinear setting I can only prove a limited commensurability result, which stems +from a reduction to the bounded setting: +Theorem 1.8. Let G, Γ and Λ be as in Theorem 1.6. If Γ is uniform then so is Λ. If Γ is of Q-rank 1 then: +1. If Γ ̸⊂ ND(Λ) for any D > 0, then also Λ is of Q-rank 1. . +2. If Γ ⊂ ND(Λ) for some D > 0 and in addition Γ sublinearly covers Λ, then Λ is commensurable to Γ. +I stress that the case where Γ and Λ each sublinearly covers the other arises naturally in the context of +SBE-completeness, see Theorem 6.4. +The most interesting case in the proof of Theorem 1.6 is when Γ is of Q-rank 1. In that case, the proof +is entirely geometric, relying on the following key proposition which might be of independent interest: +Proposition 1.9. In the setting of Theorem 1.6, assume that Γ is a Q-rank 1 lattice which does not lie +in any bounded neighbourhood of Λ. Then there exists a horosphere H based at the rational Tits building +associated to Γ such that +� +Λ ∩ StabG(H) +� +· x0 intersects H in a cocompact metric lattice. Moreover, the +bounded horoball HB does not intersect the orbit Λ · x0. +1.1 +Outline of Proof +My proofs rely and draw on the works on the quasi-isometric rigidity for non-uniform lattices, due to +Schwartz [52] in R-rank 1, and to Drut¸u [16] and Eskin [22] independently in groups of R-rank greater than +1, often called higher rank groups. The geometric proof of Theorem 1.6 for Q-rank 1 lattices is quite delicate +and involved. For this reason I give here a detailed sketch of the arguments and of the ideas one should have +in mind when reading the proof. What is written here is a good enough account if one wishes to understand +the main ideas while avoiding the technical details. I end the section with a brief sketch of the proofs for +SBE-completeness and for sublinear rigidity in the case of property (T) groups. +Strategy for Q-rank 1 Lattices. +Denote dγ := d(γ, Λ). The novel case is when {dγ}γ∈Γ is unbounded. +The rationale for the proof comes from a conjecture of Margulis, recently proved in full generality by Benoist +and Miquel ([5], see Theorem 4.4 below). Their result states that a discrete subgroup in a higher rank Lie +group is a lattice as soon as it intersects a horospherical subgroup in a lattice. This result could be seen +as an algebraic converse to the geometric structure of a Q-rank 1 lattice, whose orbit in X intersects some +parabolic horospheres in a cocompact (metric) lattice (see Section 2.2 for details). Proposition 1.9 is the +geometric analogue of the Benoist-Miquel criterion, and basically completes the proof in the higher R-rank +case (some non-trivial translation work is needed, see Section 4.4). Also, it easily follows from Proposition 1.9 +that every Γ-conical limit point is also Λ-conical (Corollary 4.27). The proof for R-rank 1 groups is then a +simple use of a criterion of Kapovich-Liu for geometrically finite groups ([30], see Theorem 4.6). I now give +a detailed description of the proof of Proposition 1.9. +3 + +The ABC of Sublinear Constraints. +Fix a point x0 ∈ X = G/K, identify Γ and Λ with Γ · x0 and +Λ · x0 respectively. Observe that by definition of dγ the interior of balls of the form B(γx0, dγ) does not +intersect Λ · x0. I call such balls (or general metric sets) Λ-free. Moreover, these balls intersect Λ · x0 (only) +in the bounding sphere: call such balls (sets) tangent to Λ. The Λ-free and, respectively, Γ-free regions in X +are the main objects of interest in this work. Since Theorem 1.6 is known in the case of bounded {dγ}γ∈Γ, +it makes sense to think about large Λ-free regions as ‘problematic’. The state of mind of the proof relies on +two easy observations that complete each other. +1. The sublinear constraint implies that dγn → ∞ forces |γn| → ∞, suggesting that ‘problematic’ Λ-free +regions should appear only ‘far away’ in the space. +2. On the other hand Λ is a group, and being Λ-free is a Λ-invariant property. In particular any metric +situation that can be described in terms of the Λ-orbit (e.g. B(γ, dγ) is a Λ-free ball tangent to Λ) can +be translated back to x0. This means that ‘problematic’ regions could actually be found near x0. +The moral of these observations can be formulated into a general principle that lies in the heart of the +argument. The sublinear constraint dγ ≤ u(|γ|) gives rise to many other constraints of ‘sublinear’ nature. +Each such constraint actually yields a uniform constraint inside any fixed bounded neighbourhood of x0. +Since Λ and Γ are groups, many of these uniform bounds which are produced ‘near’ x0 turn out to be global +bounds that depend only on the group and not on a specific orbit point. Put differently: the trick is to +describe metric situations in terms of the Γ and Λ orbits. One then uses the group invariance in order to +move these metric situations around the space to a place where the sublinear constraint can be exploited. +The Argument. +The above paragraph should more or less suffice the reader to produce a complete proof +for uniform lattices. For non-uniform lattices, denote by λγ the closest Λ-orbit point to the point γ ∈ Γ. For +H a cusp horosphere of Γ, let ΓH := {γ ∈ Γ | γx0 ∈ H}. The ultimate goal is to show that the metric lattice +ΓH · x0 yields a metric lattice that is more or less {λγx0}γ∈ΓH. One proceeds by the following steps. +1. Finding Λ-free horoballs (Section 4.2.1): The arbitrarily large Λ-free balls B(γn, u(|γn|)) are translated +to x0, and the compactness of the unit tangent space at x0 yields a converging direction which is the +base point at infinity of a Λ-free horoball. Translating by Λ, this yields Λ-free horoballs tangent to +every Λ-orbit point. +2. Controlling angles (Section 4.2.2): For every γ ∈ Γ with dγ uniformly large enough, one associates a +point ξ at X(∞) such that ξ is the base point of a Λ-free horoball tangent to λγx0. The angle between +the geodesics [λγx0, γx0] and [λγX0, ξ) is shown to be small as dγ grows large. This is used to show +that arbitrarily large dγ give rise to arbitrarily deep Γ-orbit points inside Λ-free horoballs. A key step +is Lemma 3.8, producing a Γ-free Euclidean cylinder between λγx0 and γx0. +3. Λ-cocompact horospheres (Section 4.2.3): One uses uniform bounds near x0 to prove that every Λ-free +horoball that is (almost) tangent to Λ must lie in a uniformly bounded neighbourhood of Λ · x0. If +dγ is large enough for some γ that lies on a horosphere HΓ of a cusp of Γ, then any γ′ ∈ ΓHΓ also +admits large dγ′. Since the bounds from the previous steps only depend on dγ′, all λγ′ are forced to +lie on the same horosphere HΛ parallel to HΓ. One concludes that Λ · x0 intersects HΛ on the nose in +a cocompact metric lattice. +Property (T). +Sublinear rigidity for groups with property (T) is established by the criterion that a discrete +subgroup there is a lattice if and only if it has the same exponential growth rate as a lattice (Leuzinger [36]). +It is quite straightforward that sublinear distortion cannot affect this growth rate (Corollary 5.10). +SBE Rigidity. +The general scheme for SBE-completeness is parallel to the quasi-isometry case: each +λ ∈ Λ naturally gives rise to an SBE X0 → X0 of the compact core of Γ. Each self SBE is close to an +isometry by Theorem 1.4, allowing to embed Λ as a discrete subgroup of isometries in G that sublinearly +4 + +covers Γ. The proof for Theorem 1.4 heavily relies on Drut¸u’s argument for quasi-isometries [16], which uses +the properties of the induced biLipschitz map on the asymptotic cone. As SBE also induce such a biLipschitz +map - indeed that was a main motivation for Cornulier to study SBE [13] - it is possible to generally follow +Drut¸u’s argument also in the SBE setting. +1.2 +Possible Improvements, Related and Future Work +The proof suggests three natural improvements to the statement of Theorem 1.6. One could probably relax +the assumption of trivial centre and allow finite centre, if the same relaxation is applicable in Leuzinger’s +work on property (T) groups [36] and in Prasad’s work [49]. In particular, the proofs for uniform lattices and +for Q-rank 1 lattices hold also for groups G with finite centre. The irreducibility of Λ may be derived directly +from the irreducibility of Γ. Lastly, in view of the geometric characterization of Q-rank (see Corollary D in +[35]), it is reasonable that the Q-rank of Λ should equal that of Γ. +There are problems that arise naturally from this work which seem to require new ideas: +Question. Let G be a real finite-centre semisimple Lie group without compact factors that admits a R- +rank 1 factor. Are the classes of uniform and non-uniform lattices of G SBE-complete? In particular, is this +true when G is of R-rank 1? +See Section 6.3.4 below for a discussion on the case of R-rank 1 factors. SBE of R-rank 1 symmetric spaces +is the main focus in Pallier’s work [44]. He investigated the sublinearly large scale geometry of hyperbolic +spaces, and proved that two R-rank 1 symmetric spaces that are SBE are homothetic, answering a question +of Drut¸u (see Remarks 1.16 and 1.17 in [13]). Also in this context Pallier and Qing [47] recently showed that +the sublinear Morse boundary is an SBE invariant. +Another problem is to find a non-trivial example of the setting of Theorem 1.6: +Question. Let G be a real finite-centre semisimple Lie group without compact factors, Γ ≤ G a lattice. +1. Does there exist a finitely generated group that is SBE to Γ but not quasi-isometric to it? Or, at least, +not known to be quasi-isometric to one? +2. Does there exist Λ ≤ G discrete that ε-linearly covers Γ but which does not sublinearly cover it? +On the side of the proof, it would be very interesting if the geometric ideas that prove Theorem 1.6 for +Q-rank 1 lattices could be applied to any Q-rank. While there are apparent places where the proof uses the +unique geometry of Q-rank 1 lattices, most of the geometric arguments leading to Proposition 1.9 seem to +be susceptible to the higher Q-rank setting. Such a generalization of the proof would definitely shed more +light on the mysterious lattice arising from growth considerations in property (T) groups, and in particular +on the question of commensurability of Λ and Γ in that case. It would also be interesting to see whether +one can push the geometric argument forward in order to establish a complete geometric analogue of the +Benoist-Miquel criterion. Namely, can one find a direct geometric proof that Λ admits finite co-volume +(perhaps similarly to Schwartz’s argument in the bounded case, see Section 4.3 below). +Lastly, one could possibly relate this work to the work of Fraczyk and Gelander [24], who proved that +a discrete subgroup (of a higher rank simple Lie group) is a lattice if and only if it has bounded injectivity +radius. While their result seem very much related to the condition Γ ⊂ Nu(Λ), the nature of their work does +not give explicit bounds on the injectivity radius. Specifically, given r > 0 one cannot tell directly from their +results how ‘far’ one must wander in X in order to find a point with injectivity radius r. Perhaps one could +use the sublinear results of this work to say something about the relation of |x|X and InjRad(x). +1.3 +Acknowledgments +This paper is based on my DPhil thesis, supervised by Cornelia Drut¸u. +I thank her for suggesting the +question of SBE-completeness and for guiding me in my first steps in the theory of Lie groups. I thank Uri +Bader, Tsachik Gelander and the Midrasha on groups at the Weizmann institute, where I learned the basics +5 + +of symmetric spaces. I thank my thesis examiners Emmanuel Breuillard and Yves Cornulier for their careful +inspection and numerous remarks. I thank Elon Lindenstrauss for telling me about Leuzinger’s result [36] +on property (T), and Or Landesberg, Omri Solan, Elyasheev Leibtag, Itamar Vigdorovich and Tal Cohen +for many discussions on different aspects of this paper. Finally I thank Gabriel Pallier for explaining his +examples of some unusual SBE in R-rank 1, and for his interest in this work. +2 +Preliminaries +For standard definitions and facts about fundamental domains, see Chapter 5.6.4 in [17]. The facts about +fundamental domains for Q-rank 1 lattices appear in Raghunathan’s book [50] and in Prasad’s work on +rigidity of Q-rank 1 lattices [49]. In notations and generalities I follow: Borel’s book on algebraic groups [6], +Helgason’s books on Lie groups and symmetric spaces [27, 28], and Eberline’s book on the geometry of +symmetric spaces of noncompact type [20]. +2.1 +Generalities on Semisimple Lie Groups and their Lattices +Let G be a real centre-free semisimple Lie group without compact factors. A discrete subgroup Γ ≤ G is +a lattice if Γ\G carries a finite volume G-invariant measure. Equivalently, Γ is a lattice if Γ\X is a finite +volume Riemannian manifold, where X = G/K is the symmetric space of noncompact type corresponding +to G. A lattice is irreducible if its projection to every simple factor of G is dense. The group G can be +viewed as an algebraic group via the adjoint representation. If G is of R-rank greater than 1, then by the +Margulis arithmeticity theorem every irreducible lattice of G is arithmetic. The Q-rank Γ is the Q-rank of +the Q-structure associated to (G, Γ) be the arithmeticity theorem. A result of Prasad [49] states that if G +admits a R-rank 1 factor, then a non-uniform irreducible lattice of G is of Q-rank 1 . +The group G has Kazhdan’s property (T) if and only if it does not admit an SO(n, 1) or an SU(n, 1) +factor, and an irreducible lattice Γ ≤ G has property (T) if and only if G has property (T). Together with +Prasad’s result, I may conclude: +Lemma 2.1. Let G be a real centre-free semisimple Lie group without compact factors, and Γ ≤ G an +irreducible lattice. Then at least one of the following occurs: (a) G has property (T) (b) Γ is a non-uniform +Q-rank 1 lattice (c) Γ is uniform. +Theorem 1.6 is therefore an immediate result of Theorem 3.1, Theorem 4.1 and Theorem 5.1. +2.2 +Cusps and the Rational Tits Building +The facts about symmetric spaces of noncompact type can be found in Eberline’s book [20]. +Since the +geometry of Q-rank 1 lattices resembles that of lattices in R-rank 1, the reader could for the most part +simply have the image of the hyperbolic plane in mind. If one wishes to see flats that are not geodesics, then +a product of two hyperbolic planes is enough. Even the product of the hyperbolic plane and R is helpful, +albeit this space has a Euclidean factor. +2.2.1 +Basic Geometry of Symmetric Spaces of Noncompact Type +Visual Boundary. +The visual boundary X(∞) of X is the set of equivalence classes of geodesic rays, +where two geodesic rays are equivalent if their Hausdorff distance is finite. For a ray η : [0, ∞) → X, η(∞) +denotes the equivalence class of η in X(∞). There are two natural topologies on X(∞) that will be of use. +The cone topology is the one given by viewing X(∞) as the set of all geodesic rays emanating from some +fixed base point x0, with topology induced by the unit tangent space at x0. There is a natural topology on +X := X ∪ X(∞) such that X is the compactification of X and the induced topology on X(∞) is the cone +topology. A well known fact about geodesic rays in nonpositively curved spaces, stating that two ‘close’ +geodesic rays fellow travel: +6 + +Lemma 2.2. Given time T and ε > 0, there is an angle α = α(T, ε) so that if η1, η2 are two geodesic rays +with η1(0) = η2(0) = x for some x ∈ X and ∡x(η1, η2) ≤ α then dX +� +η1(t), η2(t) +� +< ε for all t ≤ T . +The Tits metric on X(∞) is defined as follows. +Given a totally geodesic submanifold Y ⊂ X, let +Y (∞) ⊂ X(∞) be the subset of all points that admit a geodesic ray η lying inside Y (or, equivalently, those +points that admit a ray lying at bounded Hausdorff distance to Y ). For any ξ1, ξ2 ∈ X(∞) there exists a +flat F ⊂ X such that ξ1, ξ2 ∈ F(∞). Define dT (ξ1, ξ2) ∈ [0, π] to be the angle between two geodesic rays +η1, η2 ⊂ F emanating from some point x ∈ F and with η1(∞) = ξ1, η2(∞) = ξ2. This is a well defined metric +on X(∞), called the Tits metric. The pair +� +X(∞), dT +� +is a geodesic metric space, and isometries of X act +on it by isometries. I will use the following relation between the cone and the Tits topologies: +Proposition 2.3 (Section 3.1 in [20]). Let X be a symmetric space of noncompact type. The Tits metric on +X(∞) is semicontinuous with respect to the cone topology: for any ξ, ζ ∈ X(∞) and every ε > 0, there exists +neighbourhoods of the cone topology U, V ⊂ X(∞) of ξ and ζ respectively such that for all ξ′ ∈ U, ζ′ ∈ V one +has +∡(ξ′, ζ′) ≥ ∡(ξ, ζ) − ε +Moreover, for any flat F ⊂ X, the cone topology and the Tits topology coincide on F(∞). +Busemann Functions, Horoballs and Horospheres. +Horoballs and horospheres play a crucial role in +the proof, a role which stems from their role in the geometric description of the compact core of non-uniform +lattices (see Theorem 2.20 below). A Busemann function on X is any function of the form +fη(x) = lim +t→∞ d +� +x, η(t) +� +− t +for some geodesic ray η of X. A horoball HB ⊂ X is an open sublevel set of a Busemann function. A +horosphere H ⊂ X is a level set of a Busemann function. Two equivalent geodesic rays η1, η2 give rise to +Busemann function which differ by a constant, i.e. fη1 − fη2 = C for some C ∈ R. If HB is the sublevel set +of fη, then η(∞) is called the base point of the horoball HB (and respectively of the horosphere H = ∂HB). +The base point of a horoball is well defined, i.e. it depends only on η(∞) and not on η. For every choice of +x ∈ X, ξ ∈ X(∞) there is a unique horosphere H based at ξ with x ∈ H. I denote this horosphere by H(x, ξ) +and the bounded horoball HB(x, ξ). The following proposition collects some basic properties that will be of +use. +Proposition 2.4 (Proposition 1.10.5 in [20]). Let x ∈ X, ξ ∈ X(∞), and let H = H(x, ξ), HB the horoball +bounded by H and f the Busemann function based at ξ with f(x) = 0. +1. For any point y ∈ X, let η be the bi-infinite geodesic determined by the geodesic [y, ξ). Then PH(y) = +η ∩ H, where PH(y) is the unique point closest to y on H. +2. For any point y ∈ X, f(y) = ±d +� +y, PH(y) +� +. Moreover, f(y) is negative if and only if y ∈ HB. +3. If x′ ∈ X, then the horospheres H = H(x, ξ) and H′ = H(x′, ξ) are equidistant: if y ∈ H, y′ ∈ H′, then +d(y, H′) = d(y′, H). Such horospheres are called parallel. +A Busemann function fη thus naturally determines a filtration of X by the co-dimension 1 manifolds +{Ht}t ∈ R. By convention I usually assume that Ht := {x ∈ X | fη(x) = −t}. +Remark 2.5. The stabilizer of a point ξ ∈ X(∞) acts transitively on the set of horospheres based at ξ, so +every two such horospheres are isometric. Moreover, there is a close relation between the induced metrics on +horospheres with the same base point. Briefly, if dH denotes the induced distance on a horosphere H ⊂ X, +then dH +� +PH(x), PH(y) +� +for any two points x, y ∈ H′ can be bounded uniformly below and above as a function +of the distance dH′(x, y) and the curvature bounds on X. See Heintze-Im hof [26] for precise statements. +7 + +Using the above properties one can show that two horospheres are parallel if and only if they are based +at the same point. In particular for every x ∈ X, ξ ∈ X(∞) it holds that StabG +� +H(x, ξ) +� +⊂ StabG(ξ). In +addition, If H, H′ are two parallel horospheres based at the same ξ ∈ X(∞) and A ⊂ H is a cocompact +metric lattice in H, then πH′(A) is a cocompact metric lattice in H′. +For a point ξ ∈ X(∞) and a flat F with ξ ∈ F(∞), one readily observes that every horoball HB based +at ξ intersects F in a Euclidean half space. In particular for every ζ ∈ X(∞) with dT (ξ, ζ) < π +2 , for every +geodesic ray η with η(∞) = ζ and every horoball HB based at ξ there is some T for which η↾t>T ⊂ HB. +Parabolic and Horospherical Subgroups. +The isometries of X are classified into elliptic, hyperbolic, +and parabolic isometries. Most significant for this paper are the parabolic isometries, i.e. those g ∈ G whose +displacement function x �→ gx does not attain a minimum in X. Every such isometry fixes (at least) one +point in X(∞). A group P ≤ Isom(X) is called geometrically parabolic if it is of the form Gξ := StabG(ξ) for +some ξ ∈ X(∞). Such groups act transitively on X, and in particular act transitively on the set of geodesic +rays in the equivalence class of ξ. The same holds also for the identity component G◦ +ξ. An element g ∈ Gξ +acts by permutation on the set of horoballs based at ξ. This permutation is a translation with respect to +the filtration of the space X by horospheres based at ξ. Put differently, if {Ht}t∈R is a filtration of X by +horospheres based at ξ, then for every g ∈ Gξ there is l(g) ∈ R such that gHt = Ht+l(g). It is quite clear +from all of the above that for every horosphere based on ξ, the group GH := StabG(H) acts transitively on +H, and the same holds for G◦ +H. +I now present a fundamental structure theorem for geometrically parabolic groups. Denote g := Lie(G), +and let g = k ⊕ p be a Cartan decomposition defined using the maximal compact subgroup K ≤ G. Recall +that the Lie exponential map exp : g → G gives rise to a family of 1-parameter subgroups of the form +exp(tX) for each X ∈ p. +Proposition 2.6 (Proposition 2.17.3 in [20]). Let x ∈ X(∞), and let X ∈ p be the tangent vector of the +unit speed geodesic [x0, ξ). Let ht +ξ be the 1-parameter subgroup defined by t �→ exp(tX). Then an element +g ∈ G fixes ξ if and only if limt→∞ h−t +ξ ght +ξ exists. +Proposition 2.7 (Langlands Decomposition, Propositions 2.17.5 and 2.17.25 in [20]). Let ξ ∈ X(∞) and +ht +ξ as in Proposition 2.6. Let F be a flat containing [x0, ξ) and A ≤ G the maximal abelian subgroup such +that Ax0 = F. Denote Gξ := StabG(ξ), and define Tξ : Gξ → G by g �→ limn→∞ h���n +ξ +ghn +ξ . Then Tξ is a +homomorphism, and there are subgroups Nξ, Aξ, Kξ ≤ Gξ such that: +1. Aξ = exp +� +Z(X) ∩ p +� +, where Z(X) is the centralizer of X in g. Moreover, every element a ∈ Aξ lies in +some conjugate Ag = gAg−1 with the property that [x0, ξ) ⊂ F g := Agx0. +2. Kξ ≤ K = StabG(x0) is the compact subgroup fixing the bi-infinite geodesic determined by [x0, ξ). +3. KξAξ = AξKξ. +4. Nξ = Ker(Tξ). It is a connected normal subgroup of Gξ. +5. Gξ = NξAξKξ, and the indicated decomposition of an element is unique. +6. G = NξAξK, and the indicated decomposition of an element is unique. In case ξ is a regular point at +X(∞), this decomposition is the Iwasawa decomposition. +7. Gξ has finitely many connected components, and G◦ +ξ = (KξAξ)◦Nξ. +8. Gξ is self normalizing. +Viewing G as an algebraic group, the geometrically parabolic subgroups are exactly the (algebraically) +non-trivial parabolic subgroups, i.e. proper subgroups of G that contain a normalizer of a maximal unipotent +subgroup. Proposition 2.7 is a geometric formulation of the algebraic Langlands decomposition of parabolic +groups. Recall that a horospherical subgroup is the unipotent radical of a non-trivial parabolic group, or +equivalently groups of the form Ug := {u ∈ G | limn→∞ g−nugn = idG}. The latter implies that Nξ a +horospherical subgroup of G. +8 + +Limit Set. +An important set associated to a discrete group ∆ ≤ G acting by isometries on X is the limit +set L∆. By definition L∆ := ∆ · x ∩ X(∞), i.e. it is the intersection with X(∞) of the closure, in the +compactification X = X ∪ X(∞), of an orbit ∆ · x. It is clear that L∆ does not depend on the choice of +x ∈ X. The limit set of any lattice is always the entire X(∞). In fact, much more is true: +Definition 2.8. Let ∆ ≤ G = Isom(X), and SX the unit tangent bundle. A vector v ∈ SX is ∆-periodic +if there is δ ∈ ∆ and s > 0 such that δη(t) = η(t + s) for all t ∈ R, where η is the bi-infinite geodesic +determined by the vector v. +For a flat F ⊂ X (including geodesics), denote ∆F := {δ | δF = F}. The flat F is called a ∆-periodic +flat if there exists a compact set C ⊂ F such that ∆F C = F. +Proposition 2.9 (Propositions 4.7.3., 4.7.5, 4.7.7 in [20], Lemma 8.3′ in [41]). If Γ ≤ G = Isom(X) is a +lattice, then: +1. The subset in SX of Γ-periodic vectors is dense. +2. Let F ⊂ X be any flat, η any bi-infinite geodesic in F, and denote v = ˙η(0) ∈ SX the initial velocity +vector. There is a sequence vn ∈ SX of regular vectors such that +(a) limn→∞ vn = v. +(b) The bi-infinite geodesics ηn determined by vn are all Γ-periodic. +(c) Denote by Fn the (unique) flat containing ηn. Each Fn is Γ-periodic. +Put differently, the set of Γ-periodic flats is dense in the set of flats of X. +Remark 2.10. See Definition 2.25 for the notion of regular tangent vectors. +Points in the limit set of a group are classified according to how the orbit approaches them. +Definition 2.11. Let ∆ ≤ G be a discrete subgroup, ξ ∈ L∆. +1. The point ξ is called conical if for some (hence any) x and some (hence every) geodesic ray η with +η(∞) = ξ there is a number D = D(x, η) such that for every T ∈ R>0 there is t > T for which +B +� +η(t), D +� +∩ ∆ · x ̸= ∅. Since ∆ is discrete, this is equivalent to ∆ · x ∩ ND(η) being infinite. +2. The point ξ is called horospherical if for every horoball HB based at ξ and every x ∈ X, ∆ · x ∩ HB is +non-empty. In particular, a conical limit point is horospherical. +3. The point ξ is non-horospherical if it is not horospherical. +As a corollary of Proposition 2.9, one has: +Corollary 2.12. If Γ ≤ Isom(X) is a lattice, then +1. The set of Γ-conical limit points is dense in X(∞) with the cone topology. +2. LΓ = X(∞). +I finish this section with some results on geometrically finite subgroups of isometries in R-rank 1. +Definition 2.13. Let X be a R-rank 1 symmetric space and ∆ ≤ Isom(X) a discrete subgroup. Denote by +Hull(∆) the closed convex hull in X = X ∪X(∞) of the limit set L∆, and Hull(∆) = X ∩Hull(∆). By virtue +of negative curvature, Hull(∆) is the union of all geodesics η such that η(∞), η(−∞) ∈ L∆. The convex core +of ∆ is defined to be ∆\Hull(∆) ⊂ ∆\X, i.e., the quotient of Hull(∆) by the ∆-action. +Definition 2.14 (Bowditch [9], see Theorem 1.4 in [30]). Let X be a R-rank 1 symmetric space, i.e. a +symmetric space of pinched negative curvature. A discrete group ∆ ≤ G = Isom(X) is geometrically finite if +for some δ > 0, the uniform δ-neighbourhood in ∆\X of the convex core Nδ +� +∆\Hull(∆) +� +, has finite volume +and there is a bound on the orders of finite subgroups of ∆. A group is geometrically infinite if it is not +geometrically finite. +9 + +Immediately from the definition of geometrical finiteness, one gets a simple criterion for a subgroup to +be a lattice: +Corollary 2.15. Let X be a R-rank 1 symmetric space. If ∆ ≤ Isom(X) is geometrically finite and admits +L∆ = X(∞), then ∆ is a lattice in Isom(X). +Sublinear distortion does not effect the limit set, as the following lemma shows. +Lemma 2.16. Let Γ, Λ ≤ G be discrete subgroups and u : R≥0 → R>0 a sublinear function. If Γ ⊂ Nu(Λ), +then LΓ ⊂ LΛ, i.e. every Γ-limit point is a Λ-limit point. In particular, if Γ is a lattice then LΛ = X(∞). +Proof. By definition, one has to show that given a point ξ ∈ X(∞) and a sequence γn ∈ Γ such that γnx0 → ξ +(in the cone topology on X), there is a corresponding sequence λn ∈ Λ with λnx0 → ξ. Define λn := λγn to +be the closest point to γn in Λ, and to ease notation denote xn := γnx0, x′ +n = λnx0. Let also ηn := [x0, xn] +and η′ +n := [x0, x′ +n] be unit speed geodesics. Finally, let Tn denote the time in which ηn terminates, i.e. +ηn(Tn) = xn. +Convergence in the cone topology xn → ξ is equivalent to the fact that the geodesics ηn converge to +η := [x0, ξ) uniformly on compact sets. This means in particular that Tn → ∞. Non-positive curvature +guarantees that the functions Fn(t) := d +� +η(t), ηn(t) +� +, F ′ +n(t) := d +� +η(t), η′ +n(t) +� +and Gn := d +� +ηn(t), η′ +n(t) +� +are +convex (F ′ +n is just a notation, completely unrelated to the derivative of Fn). Since Gn(0) = Fn(0) = F ′ +n(0) = 0 +any of these functions is either constant 0 or monotonically increasing, so proving uniform convergence of +η′ +n to η amounts to proving limn F ′ +n(T ) = 0 for every T ∈ R≥0. +Triangle inequality gives F ′ +n(T ) ≤ Fn(T ) + Gn(T ), and by assumption limn Fn(T ) = 0. Notice that +Gn(Tn) = d(xn, x′ +n) ≤ u(|γn|) = u(Tn). Writing T = +T +Tn · Tn, convexity of Gn implies +Gn(T ) ≤ (1 − T +Tn +)Gn(0) + T +Tn +Gn(Tn) ≤ 0 + T +Tn +u(Tn) = T · u(Tn) +Tn +As limn Tn = ∞ it follows from sublinearity that limn Gn(T ) = 0. I conclude that η′ +n converge to η +uniformly on compact sets, therefore ξ lies in the limit set of Λ. +Remark 2.17. In Section 4.2 I prove that in the setting of Proposition 4.12, the set of Λ-conical limit points +contains the set of Γ-conical limit points (Corollary 4.27). It holds that the conical limit points of Γ are +dense in X(∞) (in the cone topology, see Corollary 2.12) and therefore LΛ = LΓ = X(∞). In particular, +every Γ-limit point is a Λ-limit point. The strength of Lemma 2.16 is that it does not assume anything on Γ +other than that it is sublinearly covered by Λ. In particular, Lemma 2.16 does not require Γ to be a lattice. +2.2.2 +Cusps, Compact Core, and the Rational Tits Building +In this section I present some of the structure theory of non-compact quotients of X. The focus is on the +structure of ‘cusps’ in noncompact finite volume quotients of symmetric spaces, and the ‘location’ of cusps +on the visual boundary. +Cusps and Compact Core. +Consider V = Γ\X, for Γ ≤ G a non-uniform lattice. This is a locally +symmetric space of finite volume. The term ‘cusps’ is an informal name given to those areas in a locally +symmetric space through which one can ‘escape to infinity’. Another description is that cusps are the ends +of the complement of a large enough compact set in V. In strictly negative curvature, i.e. in R-rank 1 locally +symmetric spaces, these cusps have a precise description as submanifolds of the form C × R≥0 for a compact +manifold C, and metrically (C, t) gets narrower as t → ∞. There are finitely many cusps, each corresponding +to a point at X(∞) called a ‘parabolic point’. See e.g. Introduction in [3] or [19]. A fundamental feature +of the cusps is that one can ‘chop’ them out of the quotient manifold V and get a compact manifold. This +could be done in such a way so that: +1. The lifts of the chopped parts to the universal cover X are disjoint. +10 + +2. Each cusp is covered in X by the Γ-orbit of a horoball, that is, the lift of a cusp is the Γ-orbit of a +horoball. The respective base points are called parabolic points of Γ in X(∞). +3. Γ acts on X \ +� � +i∈I HBi +� +cocompactly, where {HBi}i∈I is the set of horoballs coming from the lifts +of cusps. +Since there are only finitely many cusps and Γ discrete, there are exactly countably many such horoballs. +See for example Section 12.6 in [17] where this is illustrated in the case of the real hyperbolic spaces Hn. +Formally, one has; +Theorem 2.18 (Theorem 3.1 in [19], see also Introduction therein). Assume X is of R-rank 1, and Γ ≤ G a +non-uniform lattice. The space V = Γ\X has only finitely many (topological) ends and each end is parabolic +and Riemannian collared. In particular, each cusp is a quotient of a horoball HB based at a parabolic limit +point ξ such that Γ ∩ Gξ acts cocompactly on H = ∂HB. +For symmetric spaces of higher rank, a similar construction is available (see [37]). By removing a countable +family of horoballs from X, one obtains a subspace on which Γ acts cocompactly. There are two main +differences from the situation in R-rank 1. One is that an orbit map γ �→ γx is a quasi-isometric embedding +of Γ (with the word metric) into X. +Theorem 2.19 (Lubozki-Mozes-Raghunathan, Theorem A in [38]). Let G be a semisimple Lie group of +higher R-rank, dG a left invariant metric induced from some Riemannian metric on G. Let Γ an irreducible +lattice, dΓ the corresponding word metric on Γ. Then dG↾Γ×Γ and dΓ are Lipschitz equivalent. +This result plays a significant preliminary role in the proofs of quasi-isometric rigidity for non-uniform +lattices in higher rank symmetric spaces in both [16] and [22]. The second difference in higher rank spaces +is that the horoballs could not in general be taken to be disjoint. However, in the special case of Q-rank 1 +lattices the horoballs can be taken to be disjoint. Recall that the Q-structure of (G, Γ) is a Q-structure on +G = G(R) in which Γ is an arithmetic lattice. The following theorem sums up the relevant properties for +Q-rank 1 lattices. +Theorem 2.20 (Theorem 4.2 and Proposition 2.1 in [34], see also Remarks 3 and 4 in [37], Section 13 +in [50], and Proposition 2.1 in [49]). Assume X is of higher rank, and Γ ≤ G an irreducible torsion-free +non-uniform lattice. On the locally symmetric space V = Γ\X there exists a continuous and piece-wise real +analytic exhaustion function h : V → [0, ∞) such that, for any s > 0, the sublevel set V(s) := {h < s} is a +compact submanifold with corners of V. Moreover the boundary of V(s), which is a level set of h, consists of +projections of subsets of horospheres in X. +The Γ-action on the above set of horospheres has finitely many orbits, and the following conditions are +equivalent: +1. The corresponding horoballs bounded by these horospheres can be taken to be disjoint. +2. For each such horosphere H the action of Γ ∩ StabG(H) on H is cocompact. +3. The Q-structure of (G, Γ) is of Q-rank 1. +4. The lattice Γ is of Q-rank 1. +Definition 2.21. In the setting of Theorem 2.18 and Theorem 2.20, the horoballs and horospheres that +appear in the statement are called (global) +horoballs (horospheres) of Γ. Base points of these are called +parabolic limit points of Γ. +The above geometric characterization of Q-rank 1 lattices is all that I use in order to prove the key +Proposition 4.12. The compact core of Γ is the complement in X of the horoballs of Γ. The group Γ acts on +it cocompactly. The following corollary describes the orbit of Q-rank 1 lattices in X, and especially some +finiteness properties which I will use. +11 + +Corollary 2.22. Let Γ ≤ G be a Q-rank 1 lattice, x ∈ X, and ξ a parabolic limit point. There is a unique +horosphere H based at ξ such that both following conditions hold: +1. Γ · x ∩ H is a cocompact metric lattice in H. +2. Γ · x ∩ HB = ∅, where HB is the horoball bounded by H. +Call H an x-horosphere of Γ at ξ, and the corresponding bounded horoball an x-horoball of Γ. Moreover, +one has: +1. For every C there exists a bound K = K(x, C) so that B(x, C) intersects at most K x-horospheres of +Γ. +2. There is D = D(x) > 0 such that H ⊂ ND(Γ · x ∩ H) for any x-horosphere H of Γ. The constant D is +called the compactness number of (Γ, x). +3. There is a number N = N(Γ) such that every point x ∈ X admits exactly N x-horospheres of Γ that +intersect x. These are called the horospheres of (Γ, x). +Proof. Let x ∈ X. Since Γ is a Q-rank 1 lattice, the stabilizer in Γ of a parabolic limit point ξ ∈ X(∞) acts +cocompactly on each horosphere based at ξ, and in particular on Hx := H(x, ξ). Let D(x, H) be such that +Hx ⊂ ND(Γ · x ∩ Hx). A priori D(x) depends on H, but the fact that Γ acts by isometries implies that for +every horosphere of the form H′ := γH = H(γx, γξ) for some γ ∈ Γ, one has H′ ⊂ ND(Γ · x ∩ H′). So D(x) +depends only on the Γ-orbit of H. Since Γ acts on the set of parabolic limit points with finitely many orbits +(that is to say Γ\X has finitely many cusps) one may take D = D(x) to be the maximum of the respective +bounds on each orbit. This gives the required compactness number. +Thinking of horoballs of γ as lifts of ends of the complement of some compact subset of V = Γ\X, one +sees that there is some horoball HB based at ξ so that Γ · x /∈ HB. Denote H = ∂HB, and y = PHB(x) ∈ H +be the projection on the closed convex set that is the closure of the horoball HB. Finally, Let η = [x, y] +and denote l = d(x, y). The existence of the required horosphere is equivalent to the fact that the following +non-empty set admits a maximum: +Hx,ξ := {t ∈ [0, l] | Γ · x ∩ H +� +η(t), ξ +� +̸= ∅} +Indeed 0 ∈ Hx,ξ and it is a bounded set, so it admits a supremum T . Moreover, this set is discrete. +If t were an accumulation point, then for any small ε > 0 the geodesic segment η↾(t−ε,t+ε) would intersect +D-cocompactly infinitely many horospheres of (Γ, x). Orbit points on different horospheres are in particular +different points, therefore the set B +� +η(t), D + ε +� +∩ Γ · x would be infinite, contradicting discreteness of Γ. I +conclude that T is a maximum, and that H +� +η(T ), ξ +� +is the unique desired horosphere. +The argument above generally shows that there cannot be an accumulation point in X of x-horospheres +of Γ. In particular, for every C, the ball B(x, C) intersects only finitely many x-horospheres of Γ, say K(C), +proving item 1 in the ‘moreover’ statement. +For the last statement, simply note that the horoballs of Γ are the Γ-translates of finitely many horoballs. +In the terminology of the statement, infinitely many horospheres of (Γ, x) imply that infinitely many of +them are in the same Γ-orbit. Suppose these are {Hn}n∈N, with base points ξn that are evidently pairwise +different. Finally let γn ∈ Γ for which γnH1 = Hn. Since Γ ∩ StabG(Hn) acts cocompactly on Hn, there +is γ′ +n ∈ Γ ∩ StabG(Hn) that maps γnx to B(x, D). Discreteness of Γ implies that the set γ′ +nγnx is finite, +hence infinitely many of these points are the same point. However γ′ +nγnξ1 = ξn and therefore γ′ +nγn ̸= IdX, +contradicting the fact that Γ is torsion-free. +Remark 2.23. For the most part, I am interested in a fixed base point x0 and the x0-horospheres and +horoballs. By a slight abuse of terminology I omit x0 and call these objects ‘horospheres of Γ’ and ‘horoballs +of Γ’, respectively, denoting the associated cocompactness number DΓ. +Corollary 2.22 allows to upgrade a metric lattice of H to a metric lattice coming from StabG(H) only. +12 + +Lemma 2.24. Assume that a torsion-free discrete group ∆ ≤ G = Isom(X) admits the following: +1. There is a bound N such that at each point x ∈ ∆ · x0 there are at most N horoballs that are tangent +to x and do not intersect ∆ · x0. +2. There is a horosphere H ⊂ X such that +(a) The set ∆ · x0 ∩ H is a cocompact metric lattice in H. +(b) The horoball HB bounded by H is ∆-free, i.e. ∆ · x0 ∩ HB ⊂ H. +Then +� +∆ ∩ StabG(H) +� +· x0 is also a cocompact metric lattice in H. +Proof. This is the Pigeonhole Principle. Fix x ∈ ∆ · x0 ∩ H, and let {HBi}i∈{1,...,N} be the finite set of +horoballs tangent to x that do not intersect ∆ · x0. Assume w.l.o.g that H is the bounding horosphere of +HB1. Fix a ∆-orbit point δ0x0 ∈ ∆ · x0 ∩ H. Up to translating by some element of ∆, I may assume x = x0. +For any other ∆-orbit point δx0 ∈ ∆ · x0 ∩ H let i(δ) ∈ {1, . . . , N} be the index of the horoball δ−1HB1. +Notice that from hypothesis 1 it indeed follows that δ−1HB1 ∈ {HBi}i∈{1,...,N}, because the action of ∆ is +by isometries. Define δi to be the element in ∆ for which: +1. i(δi) = i, i.e. δ−1 +i +(HB1) = HBi +2. d(δix0, x0) is minimal among all such δ ∈ ∆. +For some i ∈ {1, . . . , N} there is a δ ∈ ∆ with i = i(δ), while for others there might not be. I will only +care about those i for which there is such δ. Assume w.l.o.g that these are i ∈ {1, . . . M} for M ≤ N, and let +L := max1≤i≤M{d(x0, δix0)} < ∞. For such an index i0, the ∆-orbit points that share the same i(δ) = i0 +are in the same ∆ ∩ StabG(H) orbit, namely +{δx0 | i(δ) = i0} ⊂ +� +∆ ∩ StabG(H) +� +· δi0x0 +Indeed, if δ−1HB1 = HBi0 then by definition δδ−1 +i0 HB1 = δHBi0 = HB1, hence δδ−1 +i0 +∈ StabG(H) is +an element mapping δi0x0 to δx0. Let now δx0 ∈ H. Its distance from the orbit +� +∆ ∩ StabG(H) +� +· x0 is +� +∆ ∩ StabG(H) +� +-invariant, therefore +d +� +δx0, +� +∆ ∩ StabG(H) +� +x0 +� += d +� +δi(δ)x0, +� +∆ ∩ StabG(H) +� +x0 +� +≤ d(δi(δ), x0) +The right-hand side is uniformly bounded by L, proving that +(∆ · x0 ∩ H) ⊂ NL +�� +∆ ∩ StabG(H) +� +· x0 +� +The fact that ∆ · x0 ∩ H is a cocompact metric lattice in H renders (∆ ∩ StabG(H) +� +· x0 a cocompact +metric lattice in h as well, as claimed. +Real and Rational Tits Buildings. +The location of the parabolic points in X(∞) also plays an important +role in the geometry of X. In case Γ is an arithmetic lattice, the natural framework to consider these points +is the so called rational Tits building. +This is a building structure on the subset of parabolic points at +X(∞), sometimes referred to as ‘rational points’ in this case. They are exactly those points in X(∞) whose +stabilizers are Q-defined (algebraic) parabolic groups of G (see Section 4.4 for more details). I present this +object, denoted WQ(Γ), together with the more familiar real Tits building structure on X(∞) with the Tits +metric. The main goal is to present the results of Hattori [25], that give a good description of the rational +Tits building in terms of conical and horospherical limit points. In case G is of R-rank 1, by WQ(Γ) I mean +the (countable) set of parabolic limit points of Γ (so that X(∞) \ WQ(Γ) is comprised of conical limit points +only, see Theorem 2.29). +13 + +Definition 2.25. A geodesic η is said to be regular if it is contained in a unique maximal flat F ⊂ X. The +point η(∞) ∈ X(∞) is called a regular point of X(∞). A point ξ ∈ X(∞) is singular if it is not regular. +Regularity does not depend on the choice of representative geodesic ray η for ξ. +A Weyl chamber of X(∞), or an open spherical chamber, is any connected component in the Tits topology +of X(∞) \ S, where S ⊂ X(∞) is the subset of singular points at X(∞). +Proposition 2.26 (Propositions 2.2 and 3.2 in [29] and Section 8 in [3]). The Weyl chambers induce a +simplicial complex structure on X(∞) that is a spherical Tits building. The apartments of the building are +exactly the sets of the form F(∞) ⊂ X(∞) for all flats F ⊂ X, and the chambers are exactly the Weyl +chambers at X(∞). Moreover, the Tits metric completely determines the building structure, and vice versa, +and +� +X(∞), dT +� +is a metric realization of the Tits building at X(∞). +None of the rich theory of buildings is used directly in this paper. Given a non-uniform lattice of Γ ≤ G +the rational Tits building WQ(Γ) is a building structure on the subset of parabolic points. It is not in general +a sub-building of the real spherical building. Flats of X correspond to real maximal split tori in G. Since +G is an algebraic group defined over Q, one can consider the maximal Q-split tori. The rational flats of X +are then the G(Q)-orbits of maximal Q-split tori of G, and the rational boundary are all points ξ ∈ X(∞) +such that ξ ∈ FQ(∞) for some rational flat FQ(∞). One defines regular rational directions and rational Weyl +chambers in an analogous way to the real case, this time taking only rational flats into account. For further +details details see [29], and Section 2 in [25]. +Theorem 2.27 (Theorem A in [25]). Let X = G/K be a symmetric space of non-compact type and of higher +rank, and let Γ ≤ Isom(X) be an irreducible non-uniform lattice. Then WQ(Γ) does not include horospherical +limit points. The π +2 -neighbourhood +N π +2 +� +WQ(Γ) +� +:= {ξ ∈ X(∞) | dT +� +ξ, WQ(Γ) +� +< π +2 } +does not include conical limit points. +In Q-rank 1 , the converse statement also holds: +Theorem 2.28 (Theorem B in [25]). Let X = G/K be a symmetric space of non-compact type of higher +rank and Γ ≤ Isom(X) be an irreducible non-uniform lattice. Let +V = {ξ ∈ X(∞) | dT +� +ξ, WQ(Γ) +� +≥ π +2 } +Suppose that Γ is of Q-rank 1 . Then V consists of conical limit points only. +In groups of R-rank 1 one has the following well known fact: +Theorem 2.29 (Proposition 5.4.2 and Theorem 6.1 in [10], see also Theorem 12.29 in [17]). Let X be a +symmetric space of noncompact type and of R-rank 1, Γ ≤ Isom(X) a lattice. Then every ξ ∈ X(∞) is either +conical or non-horospherical. +Corollary 2.30. When Γ is of Q-rank 1 , the following holds: +1. WQ(Γ) = {ξ ∈ X(∞) | N π +2 (ξ) does not contain conical limit points} +2. Any two points ξ, ξ′ ∈ WQ(Γ) are at Tits distance = π. +Proof. In view of Theorem 2.29, for R-rank 1 both statements hold trivially. In higher rank, both follow +from the following observation: for any point ξ′ ∈ WQ(Γ) and any point ζ ∈ X(∞) with d(ζ, ξ′) = π +2 , ζ is +conical. To see this notice that ζ lies on the boundary of a horosphere based at ξ′: take a flat F ⊂ X with +ξ′, ζ ∈ F(∞). Any geodesic with limit ζ is contained in (a Euclidean) horosphere based at ξ′. The fact that +Γ is cocompact on the horospheres based at WQ(Γ) implies that ζ is conical. +14 + +The second item of the corollary follows: let ξ, ξ′ ∈ WQ(Γ) and c : [0, α] a Tits geodesic joining them. +There is a flat F ⊂ X containing both ξ, ξ′ as well as c ⊂ F(∞). Every point ζ that is at Tits distance +π +2 from either ξ or ξ′ is conical, and no point inside the π +2 neighbourhood of either ξ or ξ′ is conical. In +F(∞) the Tits metric is the same as the Tits metric on the Euclidean space of an equal rank. Therefore +one may prolong the geodesic c so that c(0) = ξ, c(α) = ξ′ and c(π) = ξ′′. If dT (ξ, ξ′) < π, then there is +a point along this prolonged geodesic that is at Tits distance exactly π +2 from ξ (so it is conical by the first +paragraph), but at Tits distance strictly less than π +2 from ξ′ (so it cannot be conical by Theorem 2.27). +Therefore dT (ξ, ξ′) = π. +For the first item, one containment is just Hattori’s Theorem 2.27. For the other containment, pretty +much the same argument from above works. Assume for some ξ ∈ X(∞) that N π +2 (ξ) consists of non-conical +limit points. In particular ξ itself is not conical, and by Theorem 2.28 it holds that d(ξ, WQ(Γ)) < π +2 . Let +ξ′ ∈ WQ(Γ) be a point realizing this distance. As above, this gives rise to a flat F containing ξ, ξ′ and +another point ζ that is at Tits distance π +2 from ξ′ but at Tits distance strictly less than π +2 from ξ. The first +forces ζ to be conical, and the latter forces it to be non-conical, a contradiction. +Hattori’s characterization relies on a simple lemma which will also be of use in the sequel. It relates the +(linear) penetration rate of a geodesic into a horoball to the Tits distance. +Lemma 2.31 (Lemma 3.4 in [25]). Let X be a symmetric space of higher rank and of noncompact type. +Let η1, η2 : [0, ∞) → X be two geodesic rays, α := dT +� +η1(∞), η2(∞) +� +and b2 the Busemann function +corresponding to η2. Then there exists a positive constant C1, depending only on η1 and η2, such that: +1. If α > π +2 , then for all t ≥ 0 +b2 +� +η1(t) +� +≥ −t · cos α − C1 +2. If α = π +2 , then b2 +� +η1(t) +� +is monotone non-increasing in t and −C1 ≤ b2 +� +η1(t) +� +. +3. If α < π +2 , then for all t ≥ 0 +b2 +� +η1(t) +� +≤ −t · cos α − C1 +Remark 2.32. If X is a symmetric spaces of R-rank 1, maximal flats are geodesics. Therefore every two +points ξ, ζ ∈ X(∞) admit dT (ξ, ζ) = π, and it is clear from the strict negative curvature that Lemma 2.31 +is true also in this case. +3 +Uniform Lattices +In this section I prove: +Theorem 3.1. Let G be a semisimple Lie group without compact factors and with finite centre. Let Γ ≤ G +be a lattice, Λ ≤ G a discrete subgroup such that Γ ⊂ Nu(Λ) for some sublinear function u. If Γ is uniform, +then Λ is a uniform lattice. +The focus of this paper is on sublinear distortion, however for uniform lattices (and also for lattices that +have property (T), see Section 5), a slightly stronger result holds. I call this ε-linear rigidity. +Definition 3.2. Let f, g : R≥0 → R>0 be two monotonically increasing functions. Call f asymptotically +smaller than g if lim sup f +g ≤ 1. Denote this relation by f ⪯∞ g. +Theorem 3.3. In the setting of Theorem 3.1, the conclusion holds also under the relaxed assumption that +u(r) ⪯∞ εr for any 0 < ε < 1. +Clearly Theorem 3.3 implies Theorem 3.1. From now and until the end of this section, the standing +assumptions are those of Theorem 3.3. +15 + +Lattice Criterion. +A discrete group is a uniform lattice if and only if it admits a relatively compact +fundamental domain. The criterion I use is the immediate consequence that if Γ is uniform and u is bounded +(i.e. Γ ⊂ ND(Λ) for some D > 0), then Λ is a uniform lattice. +Outline of Proof and Use of ε-Linearity. +The goal is to show that the ε-linearity of u forces Γ ⊂ ND(Λ) +for some D > 0, i.e. that Γ actually lies inside a bounded neighbourhood of Λ. The proof is by way of +contradiction. If there is no such D > 0 then there are arbitrarily large balls that do not intersect Λ. The +proof goes by finding such large Λ-free balls that are all tangent to some fixed arbitrary point x ∈ X (see +Figure 1). The ε-linearity then gives rise to concentric Γ-free balls that are arbitrarily large, contradicting +the fact that Γ is a uniform lattice. +Remark 3.4. The main difference from the non-uniform case is that for a non-uniform lattice Γ, the space +X does admit arbitrarily large Γ-free balls. This situation requires different lattice criteria and much extra +work. Still the proof for the uniform case, though essentially no more than a few lines, lies the foundations +for and presents the logic of the much more involved case of Q-rank 1 lattices. +3.1 +Notations and Terminology +Definition 3.5. Let X be a metric space, Y, Z ⊂ X two closed subsets of X. The closest point projection +of Y to Z is the set theoretic map pZ : Y → Z defined by pZ(y) := zy, where zy ∈ Z is any point realizing +the distance d(y, Z) = d(y, z). If X is a proper metric space and Z discrete, then there are at most finitely +many such points. In any case of multiple points, pZ chooses one arbitrarily. +The particular case of interest from now on is where the metric space is the pointed symmetric space +(X, x0), the two subsets are the orbits Γ · x0 and Λ · x0, and the projection is pΛ·x0 : Γ · x0 → Λ · x0. To ease +notation I often denote this projection by pΛ (there is no risk of ambiguity since the subgroups Λ and Γ are +always considered in the context of their respective orbits in X and not in G). +The following definitions will be used repeatedly in both this section and in Section 4. It mainly fixes +terminology and notation of the geometric situation illustrated in Figure 1. +Definition 3.6. Let H ≤ G = Isom(X)◦. A set U ⊂ X is called H-free if H · x0 ∩ Int(U) = ∅, where Int(U) +is the topological interior of U. That is, U is called H-free if its interior does not intersect the H-orbit H ·x0. +Definition 3.7. Denote PΛ(γx0) = PΛ(γ) = λγx0. +1. dγ := d(γx0, λγx0). +2. Bγ := B(γx0, dγ). It is a Λ-free ball centred at γx0 and tangent to λγx0. +3. x′ +γ := λ−1 +γ γx0. Notice |x′ +γ| = dγ. +4. B′ +γ := λ−1 +γ Bγ = B(x′ +γ, dγ). It is Λ-free as a Λ-translate of the Λ-free ball Bγ, and is tangent to x0. +5. For s ∈ R>0 and a ball B = B(x, r), denote sB := B(x, sr), the rescaled ball with same centre and +radius sr. +6. For a sequence γn, denote by λn, dn, Bn, B′ +n, x′ +n the respective λγn, dγn, etc. +3.2 +Proof of Theorem 3.3 +Lemma 3.8. Let x ∈ X. There exists S = S(x, u) ∈ (0, 1) such that for every s ∈ (0, S) there is R = R(s, S) +such that if r > R and B = B(y, r) is a Λ-free ball tangent to x, then sB is Γ-free. +In particular, the existence of arbitrarily large Λ-free balls that are all tangent to a fixed point x ∈ X +implies the existence of arbitrarily large Γ-free balls. +16 + +x0 +x′ +γ += dγ += s · dγ +Γ − free +Λ − free +γx0 += dγ +Λ − free +λγx0 +Lλ−1 +γ +Figure 1: Basic Setting and Lemma 3.8. A Λ-free ball about γx0 of radius dγ, translated by λ−1 +γ +to a ball +tangent to x0. The linear ratio between |x′ +γ| = dγ and the Λ-free radius forces the red ball to be Γ-free. +There is a slightly stronger version of Lemma 3.8 if u is sublinear: +Lemma 3.9. Let G, Γ, Λ and u be as in Theorem 3.1 (in particular, u is a sublinear function and Γ ⊂ Nu(Γ)). +For every x ∈ X and every s ∈ (0, 1) there exists R = R(x, s) > 0 such that for every r > R, if B = B(y, r) +is a Λ-free ball tangent to x then sB is Γ-free. +I omit the proof of Lemma 3.9, which is a slightly simpler version of the proof of Lemma 3.8. +Proof of Lemma 3.8. The proof is more easily read if one assumes x = x0 and u(r) = εr so I begin with this +case. Assume B = B(y, R) is Λ-free for some y ∈ X. The assumption x = x0 means that |y| = d(y, x0) = r. +Assume that for a given s ∈ (0, 1), the ball sB intersects Γ · x0. This gives rise to an element γ ∈ Γ such +that: +1. |γ| = d(γx0, x0) ≤ d(y, x0) + d(γx0, y) = (1 + s)r (triangle inequality). In particular, d(γx0, Λ · x0) ≤ +ε(1 + s)r +2. B +� +γx0, (1 − s)r +� +⊂ B(y, r), so it is Λ-free. +I conclude that for s for which sB ∩ Γ · x0 ̸= ∅, one has the inequality (1 − s)r ≤ ε(1 + s)r, i.e. 1−s +1+s ≤ ε. +The number ε is fixed and smaller than 1, while 1−s +1+s limit to 1 monotonically from below as s > 0 tend to +0. I conclude that there is a segment (0, S) ⊂ (0, 1) such that for all s ∈ (0, S), sB is Γ-free. +Assume now that x ̸= x0 and u(r) ⪯∞ εr. As above, if γx0 ∈ B(y, sr) then it is the centre of a Λ-free +ball of radius (1 − s)r, and so (1 − s)r ≤ u(|γ|). I wish to use the ε-linear bound on u as I did before, +only this time u is only asymptotically smaller than εr. To circumvent this I just need to show that |γ| +is large enough. Indeed since B +� +γx0, (1 − s)r +� +is Λ-free it does not contain x0 ∈ Λ · x0 and in particular +(1 − s)r ≤ d(x0, γx0) = |γ|. For some R1(s) = R1(s, u) one therefore has for all r > R1(s) +(1 − s)r ≤ u(|γ|) ≤ ε|γ| +On the other hand |y| ≤ d(x, y) + d(x, x0) = r + |x|, and consequently |γ| ≤ (1 + s)r + |x|. For r > R1(s) +one has +17 + +(1 − s)r ≤ u(|γ|) ≤ ε|γ| ≤ ε +� +(1 + s)r + |x| +� +This means that s for which Γ · x0 ∩ B(y, sr) ̸= ∅ must admit, for all r > R1(s), +1 − s +1 + s + |x| +r += +(1 − s)r +(1 + s)r + |x| ≤ ε < 1 +(1) +The rest of the proof is just Calculus 1, and concerns with finding S = S(x, ε, u) ∈ (0, 1) so that for any +s ∈ (0, S) there is R(s) such that all r > R(s) satisfy +ε < +1 − S +1 + S + |x| +r +≤ +1 − s +1 + s + |x| +r +(2) +The lemma readily follows from inequalities 1,2. +Explicitly, fix ε′ > ε. As before, monotonic approach of +1−s +1+s to 1 allows to fix S ∈ (0, 1) for which +ε < ε′ < 1−s +1+s for all s ∈ (0, S). Next note that for any fixed s ∈ (0, S), limr→∞ +1−s +1+s+ |x| +r = 1−s +1+s, and that the +approach in monotonically increasing with r. Since ε < ε′, this limit implies that for some R2 > R1(S), all +r > R2 admit ε < +1−S +1+S+ |x| +r . Finally notice that for any fixed r the function +1−s +1+s+ |x| +r +is again monotonically +increasing as s tends to 0 from above. Therefore inequality 2 holds for every s ∈ (0, S) and all r > R2(S) +(capital S is intentional and important). +To conclude the proof, notice that if moreover r > R1(s) (again lowercase s is intentional and important) +then inequalities 1,2 both hold. This means that for any s ∈ (0, S) there is R(s) := max{R1(s), R2(S)} such +that r > R(s) ⇒ B(y, sr) is Γ-free. The constants R1(s), R2(S) have the desired dependencies, hence so +does R(s), proving the lemma. +Corollary 3.10 (Theorem 3.3). There is a uniform bound on {dγ}γ∈Γ, i.e., Γ ⊂ ND(Λ) for some D > 0. +In particular, Λ is a uniform lattice. +4 +Q-rank 1 Lattices +In this section I prove: +Theorem 4.1. Let G be a semisimple Lie group without compact factors and with finite centre, Γ ≤ G an +irreducible non-uniform Q-rank 1 lattice, Λ ≤ G a discrete irreducible subgroup. If Γ ⊂ Nu(Λ) for some +sublinear function u, then Λ is a lattice. Moreover, if Γ ̸⊂ ND(Λ) for any D > 0, then Λ is also of Q-rank 1. +If Γ ⊂ ND(Λ) for some D > 0, then Λ could be a uniform lattice. An obvious obstacle for that is if +Λ ⊂ Nu′(Γ) for some sublinear function u′. This condition turns out to be sufficient for commensurability. +Proposition 4.2. Let G be a semisimple Lie group without compact factors and with finite centre, Γ ≤ G an +irreducible non-uniform Q-rank 1 lattice, Λ ≤ G a discrete subgroup such that Γ ⊂ ND(Λ) for some D > 0, +and Λ ⊂ Nu(Γ) for some sublinear function u. Then Λ ⊂ ND′(Γ) for some D′. +As a result of Eskin’s and Schwartz’s arguments (see Section 4.3), I conclude: +Corollary 4.3. In the setting of Proposition 4.2 and unless G is locally isomorphic to SL2(R), Λ is com- +mensurable to Γ. +Theorem 1.8 is a result of Theorem 4.1 and Corollary 4.3 +18 + +4.1 +Strategy +Lattice Criteria. +I use three different lattice criteria, depending on the R-rank of G and on whether or +not Γ ⊂ ND(Λ) for some D > 0. My proof for Q-rank 1 lattices is motivated by a criterion of Benoist +and Miquel, resolving a conjecture of Margulis. It can be viewed as an algebraic converse to the geometric +structure of the compact core described in Theorem 2.20. +Theorem 4.4 (Theorem 2.16 in [5]). Let G be a semisimple real algebraic Lie group of real rank at least 2 +and U be a non-trivial horospherical subgroup of G. Let ∆ be a discrete Zariski dense subgroup of G that +contains an indecomposable lattice ∆U of U. Then ∆ is a non-uniform irreducible arithmetic lattice of G. +Remark 4.5. See Definition 4.47 for the precise meaning of an indecomposable horospherical lattice. +For R-rank 1 groups, one has the following theorem by Kapovich and Liu, stating that a group is +geometrically finite so long as ‘most’ of its limit points are conical. Recall L(∆) is the limit set of ∆ ≤ +Isom(X), and Lcon(∆) is the set of its conical limit points. +Theorem 4.6 (Theorem 1.5 in [30]). Let X be a R-rank 1 symmetric space. A discrete subgroup ∆ ≤ +Isom(X) is geometrically infinite if and only if the set L(∆) \ Lcon(∆) of non-conical limit points has the +cardinality of the continuum. +As a direct corollary I obtain the following criterion: +Corollary 4.7. Let X be a R-rank 1 symmetric space, Γ ≤ G = Isom(X) a non-uniform lattice and Λ ≤ G +a discrete subgroup. If L(Λ) = X(∞) and Lcon(Γ) ⊂ Lcon(Λ), then Λ is a lattice. +Proof. Since Γ is a lattice, L(Γ) = X(∞) and it is geometrically finite. Theorem 4.6 implies the cardinality +of X(∞) \ Lcon(Γ) is strictly smaller than the continuum. The assumption Lcon(Γ) ⊂ Lcon(Λ) implies the +same holds for Λ, and in particular that Λ is geometrically finite. The assumption that L(Λ) = X(∞) implies +that Λ is geometrically finite if and only if it is a lattice. +Corollary 4.8. Let X be a R-rank 1 symmetric space, Γ ≤ G = Isom(X) a non-uniform lattice and Λ ≤ G +a discrete subgroup. If Γ ⊂ ND(Λ) for some D > 0, then Λ is a lattice. +Proof. By definition of the limit set and of conical limit points, it is clear that every Γ-limit point is a Λ-limit +point, and every Γ-conical limit point is also Λ-conical limit point. I conclude from Corollary 4.7 that Λ is +a lattice. +Also in higher rank the inclusion Γ ⊂ ND(Λ) implies that Λ is a lattice. This result is due to Eskin. +Theorem 4.9 (Eskin, see Remark 4.11 below). Let G be a semisimple Lie group without compact factors and +of higher rank, Γ ≤ G an irreducible non-uniform lattice, Λ ≤ G a discrete subgroup such that Γ ⊂ ND(Λ) +for some D > 0. Then Λ is a lattice. +Theorem 4.9 was used in the proof of quasi-isometric rigidity for higher rank non-uniform lattices in [16], +[22]. In the (earlier) R-rank 1 case, Schwartz [52] used an analogous statement, which requires one extra +assumption. +Theorem 4.10 (Schwartz, see Section 10.4 in [52] and Remark 4.11 below). Let G be a real simple Lie group +of R-rank 1 and with finite centre, Γ ≤ G an irreducible non-uniform lattice, Λ ≤ G a discrete subgroup such +that both Γ ⊂ ND(Λ) and Λ ⊂ ND(Γ) for some D > 0. Then Λ is a lattice and commensurable to Γ. +Remark 4.11. Theorem 1.6 should be viewed as a generalization of the bounded case depicted in Theo- +rems 4.9 and 4.10, which were known to experts in the field in the late 1990’s. Complete proofs for these +statements were never given in print, and I take the opportunity to include them here. See Section 4.3, where +I also prove Proposition 4.2. I thank Rich Schwartz and Alex Eskin for supplying me with their arguments +and allowing me to include them in this paper. I also thank my thesis examiner Emmanuel Breuillard for +encouraging me to find and make these proofs public. +19 + +Outline of Proof and Use of Sublinearity. +Lattices of Q-rank 1 admit a concrete geometric structure +(see Section 2.2). This structure is manifested in the geometry of an orbit of such a lattice in the corresponding +symmetric space X = G/K. One important geometric property is the existence of a set of horoballs which +the orbit of the lattice intersects only in the bounding horospheres, and in each such horosphere the orbit +forms a (metric) cocompact lattice. +Corollary 4.8 and Theorem 4.9 reduce the proof to the case where Γ ̸⊂ ND(Λ) for any D > 0. In that +case, the essence lies in proving the existence of horospheres in X which a Λ-orbit intersects in a cocompact +lattice. +This is proved purely geometrically, using the geometric structure of Q-rank 1 lattices and the +sublinear constraint. Together with some control on the location of these horospheres, I prove two major +statements: +1. Λ · x0 intersects a horosphere H ⊂ X in a cocompact lattice (Proposition 4.12). +2. Every Γ-conical limit point is also a Λ-conical limit point (Corollary 4.27). +The R-rank 1 case of Theorem 4.1 follows directly from Corollary 4.7 using the second item above. The +higher rank case requires a bit more. namely it requires to deduce from the above items that Λ meets +the hypotheses of the Benoist-Miquel Theorem 4.4. To that end I use a well known geometric criterion +(Lemma 4.50) in order show that Λ is Zariski dense, and a lemma of Mostow (Lemma 4.36) to show that Λ +intersects a horospherical subgroup in a cocompact lattice. +Outline for Section 4. +Section 4.2 is the core of the original mathematics of this paper. It is devoted +to proving that Λ · x0 intersects some horospheres in a cocompact lattice. The proof is quite delicate and +somewhat involved, and I include a few figures and a detailed informal overview of the proof. The figures +are detailed and may take a few moments to comprehend, but I believe they are worth the effort. +Section 4.3 deals with the case where Γ ⊂ ND(Λ), and elaborates on Schwartz’s and Eskin’s proofs +of Theorem 4.9 and Theorem 4.10. Section 4.4 is devoted to the translation of the geometric results of +Section 4.2 to the algebraic language used in Theorem 4.4. +Though the work is indeed mainly one of +translation, some of it is non-trivial. Finally, in Section 4.5 I put everything together for a complete proof +of Theorem 4.1. +I highly recommend the reader to have a look at the uniform case in Section 3 before reading this one. +4.2 +A Λ-Cocompact Horosphere +Recall that dγ := d(γx0, λγx0). In this section I prove: +Proposition 4.12. If {dγ}γ∈Γ is unbounded, then there exists a horosphere H based at WQ(Γ) such that +� +Λ∩StabG(H) +� +·x0 intersects H in a cocompact metric lattice. Moreover, the bounded horoball HB is Λ-free. +Throughout Section 4.2 the standing assumptions are that {dγ}γ ∈ Γ is unbounded, and Γ is an irreducible +Q-rank 1 lattice. +The Argument. +The proof is by chasing down the geometric implications of unbounded dγ. +These +implications are delicate, but similar in spirit to the straight-forward proof for uniform lattices. The proof +consists of the following steps: +1. Unbounded dγ results in Λ-free horoballs HBΛ tangent to Λ-orbit points. Each such horoball is based +at WQ(Γ), giving rise to corresponding horoballs of Γ, denoted HBΓ. +2. If dγ is large, then γx0 must lie deep inside a unique Λ-free horoball tangent to λγx0. I use: +(a) A bound on the distance d(HΛ, HΓ). +(b) A bound on the angle ∠λγx0([λγx0, γx0], [λγx0, ξ)), where ξ ∈ X(∞) is the base point of a suitable +Λ-free horoball tangent to λγx0. +20 + +3. There exist horospheres of Γ, say HΓ, such that if γx0 ∈ HΓ then large dγ implies large Λ-free areas +along the bounding horosphere of some HBΛ. +4. If HBΛ is boundedly close to some Λ-orbit point, then HΛ is almost Λ-cocompact, that is HΛ ⊂ +ND(Λ·x0) for some universal D = D(Λ). Together with the previous step, this yields a uniform bound +on dγ along certain horospheres of Γ. +5. Finally I elevate the almost cocompactness to actual cocompactness and show HΛ ⊂ ND(Λ · x0 ∩ HΛ) +for some D > 0. This immediately elevates to HΛ ⊂ (Λ ∩ StabG(HΛ)) · x0, proving the proposition. +The Properties of Γ. +The geometric properties of Γ that are used in the proof are: +1. In higher rank, the characterization of WQ(Γ) using conical / non-horospherical limit points (Corol- +lary 2.30). In R-rank 1, the dichotomy of limit points being either non-horospherical or conical (The- +orem 2.29). +2. Γ-cocompactness along the horospheres of Γ. +3. For every point x ∈ X and C > 0 there is a bound K(C) on the number of horospheres of Γ that +intersect B(x, C) (Corollary 2.22). +4.2.1 +Λ-Free Horoballs +I retain the notations and objects defined in Section 3.1. +Lemma 4.13. There exists a Λ-free horoball tangent to x0. +Proof. Since {dγ}γ∈Γ is unbounded, there are γn ∈ Γ with dn = dγn = d(γn, λn) → ∞ monotonically, where +λn ∈ Λ is a Λ-orbit point closest to γn. Denote x′ +n = λ−1 +n γnx0, ηn := [x0, x′ +n], and vn ∈ Sx0X the initial +velocity vectors vn := +˙ηn(0). The tangent space Sx0X is compact, so up to a subsequence, vn converge +monotonically in angle to a direction v ∈ Sx0X. Let η be the unit speed geodesic ray emanating from x0 +with initial velocity ˙η(0) = v. Denote ξ := η(∞) the limit point of η in X(∞). +I claim that the horoball HB := ∪t>0B +� +η(t), t +� +, based at ξ and tangent to x0, is Λ-free. Let t > 0 and +consider η(t). For every ε > 0, there is some angle α = α(t, ε) such that any geodesic η′ with ∠x0(η, η′) < α +admits d +� +η(t), η′(t) +� +< ε/2. The convergence vn → v implies d +� +η(t), ηn(t) +� +< ε/2 for all but finitely many +n ∈ N. In particular, B +� +η(t), t +� +⊂ Nε +� +B +� +ηn(t), t +�� +for all such n ∈ N. +For a fixed t ≤ dn, it is clear from the definitions that B +� +ηn(t), t +� +⊂ B′ +n = B(x′ +n, dn). One has dn → ∞, +and so for a fixed t > 0 it holds that t < dn for all but finitely many n ∈ N. I conclude that for any fixed +t > 0 there is n large enough such that +B +� +η(t), t +� +⊂ Nε +� +B +� +ηn(t), t +�� +⊂ NεB′ +n +I conclude that for every ε > 0, HB ⊂ � +n Nε(B′ +n) = Nε +� � +n B′ +n +� +. This implies that any point in the +interior of HB is contained in the interior of one of the Λ-free balls B′ +n, proving HB is Λ-free. +Lemma 4.14. Suppose HB is a Λ-free horoball, based at some point ξ ∈ X(∞). Then ξ ∈ WQ(Γ). +Proof. For any geodesic η with limit ξ, the size d(x0, γx0) of the Γ-orbit points γx0 that lie boundedly close +to η grows linearly in the distance to any fixed horosphere based at ξ, and in particular to H = ∂HB. The +sublinear constraint d(γx0, λγx0) ≤ u(|γ|) together with the fact that HB is Λ-free imply that the size of +such γ is bounded. In R-rank 1 every limit point is either conical or in WQ(Γ), proving the lemma in this +case. For higher rank, the above argument actually shows more: it shows that a point ξ′ ∈ N π +2 (ξ) is not +conical, because every geodesic with limit ξ′ ∈ N π +2 (ξ) entres HB at a linear rate (Lemma 2.31). Hattori’s +characterization of WQ(Γ) (Corollary 2.30) implies ξ ∈ WQ(Γ). +21 + +Definition 4.15. Given a Λ-free horoball HBΛ, Lemma 4.14 gives rise to a horoball of Γ based at the same +point at X(∞). Call this the horoball corresponding to HBΛ, and denote it by HBΓ. The corresponding +horosphere is denoted HΓ. +Remark 4.16. In the course of my work I had had a few conversations with Omri Solan regarding the +penetration of geodesics into Λ-free horoballs. Assuming Λ ⊂ Nu(Γ) implies that Λ preserves WQ(Γ) (see +Lemma 4.32). This is the case in the motivating setting where Λ is an abstract finitely generated group that +is SBE to Γ, see Claim 6.4 in Chapter 6. In the case of SL2(R) Omri suggested to use the action of Λ on +the Bruhat-Tits tree of SL2(Qp) (for all primes p) and the classification of these elements into elliptic and +hyperbolic elements (separately for each p) in order to deduce that Λ actually lies in SL2(Z). We did not +pursue that path nor its possible generalization to the SLn case and general Bruhat-Tits buildings. +4.2.2 +A Γ-orbit point Lying Deep Inside a Λ-Free Horoball +I established the existence of Λ-free horoballs. It may seem odd that the first step in proving Λ · x0 is +‘almost everywhere’ is proving the existence of Λ-free regions. But this fits perfectly well with the algebraic +statement that non-uniform lattices must admit unipotent elements (see Proposition 5.3.1 in [39]). The goal +of this section is to obtain some control on the location of the Λ-free horoballs, in order to conclude that +some γx0 lies deep inside HBΛ. This results in yet more Λ-free regions, found on the bounding horosphere. +I need one property of sublinear functions. +I thank Panos Papazoglou for noticing a mistake in the +original formulation. +Lemma 4.17. Let u be a sublinear function, f, g : R≥0 −→ R>0 two positive monotone functions with +limx→∞ f(x) + g(x) = ∞. If for all large enough x it holds that f(x) ≤ u +� +f(x) + g(x) +� +, then for every 1 < s +and all large enough x it holds that f(x) ≤ u +� +s · g(x) +� +. In particular f(x) ≤ u′� +g(x) +� +for some sublinear +function u′. +Proof. Assume as one may that u is non-decreasing. By definition of sublinearity limx→∞ +u +� +f(x)+g(x) +� +f(x)+g(x) += 0, +so by hypothesis limx→∞ +f(x) +f(x)+g(x) = 0. This means that for every ε > 0 one has f(x) ≤ ε · g(x) for all large +enough x, resulting in +f(x) ≤ u +� +f(x) + g(x) +� +≤ u +� +(1 + ε) · g(x) +� +Notice that for a fixed s > 0, the function u′(x) = u(sx) is sublinear, as +lim +x→∞ +u(sx) +x += lim +x→∞ s · u(sx) +sx += 0 +Lemma 4.18. Let C > 0. There is L = L(C) such that if HBΛ is any Λ-free horoball tangent to a point +x ∈ B(x0, C) then d(HΛ, HΓ) ≤ L. Moreover, there is a sublinear function u′ such that: +L(C) ≤ +� +u′(C) +if HBΓ ⊂ HBΛ +C +if HBΛ ⊂ HBΓ +Proof. If HBΛ ⊂ HBΓ, then clearly d(HΛ, HΓ) ≤ C, simply because HBΓ is Γ-free and in particular cannot +contain x0. Therefore HΓ must separate HΛ from x0 and in particular d(HΛ, HΓ) ≤ C. +Assume that HBΓ ⊂ HBΛ, and denote l = d(HΛ, HΓ). The horoball HBΓ is a horoball of Γ, hence Γ · x0 +is DΓ-cocompact along HΓ and there is an element γ ∈ Γ with |γ| ≤ C + l + DΓ and γx0 ∈ HΓ. Since HBΛ +is Λ-free one has +l ≤ d(γx0, λγx0) ≤ u(|γ|) ≤ u(C + l + DΓ) +C, DΓ are fixed, so this inequality can only occur for boundedly small l, say l < L′(C) (DΓ is a universal +constant and may be ignored). Consult figure 2 for a geometric visualization of this situation. +22 + +Rad = C +x0 +Λ − free HBΛ +ξ +Γ − free HBΓ +≤ L(C) +γx0 +PHΓ(x0) +≤ L(C) + DΓ +Figure 2: Lemma 4.18. A Λ-free horoball HBΛ intersects a ball of radius C about x0. The associated +Γ-free horoball HBΓ is boundedly close, essentially due to the uniform cocompactness of Γ · x0 along the Γ +horospheres. +It remains to show that L′(C) is indeed sublinear in C. First define L(C) to be the minimal L that +bounds the distance d(HΛ, HΓ) for all possible HBΛ tangent to a point x ∈ B(x0, C). This is indeed a +minimum, since by Corollary 2.22 there are only finitely many horoballs of Γ intersecting B +� +x0, C + L′(C) +� +. +For every C there is thus a horoball HBΓ +C and an element γ ∈ Γ such that γx0 ∈ HΓ, d(HΛ +C, HΓ +C) = L(C) +and |γ| ≤ C + L(C) + DΓ. The fact that HBΛ +C is Λ-free implies +L(C) = d(HΛ +C, HΓ +C) ≤ u(|γ|) ≤ u +� +C + L(C) + DΓ +� +Lemma 4.17 implies that L(C) ≤ u′(C) for some sublinear function u′. +The following is an immediate corollary, apparent already in the above proof. +Corollary 4.19. For every C > 0 there is a bound K = K(C) and a fixed set ξ1, ξ2, . . . ξK ∈ WQ(Γ) ⊂ X(∞) +so that every Λ-free horoball HBΛ which is tangent to some point x ∈ B(x0, C) is based at ξi for some +i ∈ {1, . . ., K}. In particular, for any specific point x ∈ NC(Λ · x0) there are at most K Λ-free horoballs +tangent to x. +Proof. Let HBΛ be a horoball tangent to a point x ∈ B(x0, C). +Lemma 4.18 bounds d(HBΛ, HBΓ) by +L(C), hence HBΓ is tangent to a point x′ ∈ B +� +x0, C + L(C) +� +. By Corollary 2.22 there are only finitely +many possibilities for such HBΓ. In particular there are finitely many base-points for these horoballs, say +ξ1, ξ2, . . . , ξK(C) ∈ WQ(Γ). Finally, recall that a horoball is determined by a base point and a point x ∈ X +tangent to it, so the last statement of the corollary holds for any x ∈ B(x0, C). But the property in question +is Λ-invariant so the same holds for any point x ∈ Λ · B(x0, C) = NC(Λ · x0). +The bound on d(HBΛ, HBΓ) given by Lemma 4.18 further strengthen the relation between HBΛ and HBΓ. +The ultimate goal is to show that the HBΛ-s play the role of the Γ-horoballs in the geometric structure of +Q-rank 1 lattices, namely to show that Λ · x0 is cocompact on the HΛ-s. This requires to actually find +Λ-orbit points somewhere in X, and not just Λ-free regions as was done up to now. As one might suspect, +these points arise as λγx0 corresponding to points γx0 ∈ HΓ, which exist in abundance since Γ · x0 ∩ HΓ is +a cocompact lattice in HΓ. +The hope is that a Λ-free horoball HBΛ tangent to λγx0 would correspond to a horoball of Γ tangent to +γx0. This would have forced all the λγ to actually lie on the same bounding horosphere, and {λγx0 | γx0 ∈ +HΓ} would then be a cocompact lattice in HΛ. This hope turns out to be more or less true, but it requires +23 + +some work. The goal of the rest of this section is to establish a relation between a Λ-free horoball HBΛ +tangent to λγx0 and γx0. I start with some notations. +Definition 4.20. In light of Corollary 4.19, there is a finite number N of Λ-free horoballs tangent to x0. +Denote: +1. {HBΛ +i }N +1=1 are the Λ-free horoballs tangent to x0. +2. ξi ∈ WQ(Γ) is the base point of HBΛ +i . +3. vi ∈ Sx0X is the unit tangent vector in the direction ξi. +4. ηi := [x0, ξi) is the unit speed geodesic ray emanating from x0 with limit ξi. In particular vi = d +dtηi(0). +5. HBΓ +i is the horoball of Γ that corresponds to HBΛ +i , based at ξi. +6. HBΛ +λ,i, ξi +λ, ηi +λ are the respective λ-translates of the objects above. For example, HBΛ +λ,i := λ · HBΛ +i . +7. H decorated by the proper indices denotes the horosphere bounding HB, the horoball with respective +indices, e.g. HΛ +i := ∂HBΛ +i . +8. For an angle α > 0 and a tangent vector v0 ∈ SxX, define +(a) The α-sector of v in SxX is the set {v ∈ SxX | v ∈ Nα(v0)}. Recall that the metric on SxX is +the angular metric. +(b) The α-sector of v in X are all points y ∈ X for which the tangent vector at 0 of the unit speed +geodesic [x, y] lies in the α-sector of v in SxX. +Lemma 4.21. For every angle α ∈ (0, π +2 ) there exists D = D(α) such that if dγ > D then for some +i ∈ {1, . . . , N}, γx0 lies inside the α-sector of vi +λγ at λγx0. Furthermore whenever α is uniformly small +enough, there is a unique such i = i(γ), independent of α. +Proof. Translation by the isometry λ−1 +γ +preserves angles and distances, so it is enough to prove that there +is an i for which x′ +γ := λ−1γx0 lies inside the α-sector of vi, and that this i is unique if α is uniformly small. +Assume towards contradiction that there is α ∈ (0, π +2 ) and a sequence γn ∈ Γ, λn := λγn ∈ Λ with dγn +unbounded, and x′ +n := λ−1 +n γnx0 not lying in the union of the α-sectors of vi. By perhaps taking smaller α +I may assume all the α-sectors of the vi in Sx0X are pairwise disjoint. This can be done because there are +only finitely many vi. +Compactness of Sx0X allows me to take a converging subsequence v′ +n := +˙ +[x0, x′n], with limit direction +v′. Denote by η′ the geodesic ray emanating from x0 with initial velocity v′. The exact same argument of +Lemma 4.13 proves that η′(∞) is the base point of a Λ-free horoball tangent to x0. But this means v′ = vi +for some i ∈ {1, . . . , N}, contradicting the fact that all v′ +n lie outside the α-sectors of the vi. This proves +that there is a bound D = D(α) such that if dγ > D then x′ +γ lies within the α-sector of some vi. +The proof clearly shows that whenever α is small enough so that the α-sectors of the vi are disjoint, x′ +γ +lies in the α-sector of a unique vi as soon as dγ > D(α). +Remark 4.22. In the proof of Lemma 4.13 I used compactness of Sx0X to induce a converging subsequence +of directions. Lemma 4.21 actually shows that the fact there are finitely many Λ-free horoballs tangent to +x0 implies a posteriori that there was not much choice in the process - all directions [x0, x′ +γ] must fall into +one of the finitely many directions vi. +Next, I want to control the actual location of certain points with respect to the horoballs of interest, and +not just the angles. This turns out to be a more difficult of a task than one might suspect, since control on +angles does not immediately give control on distances. +Recall that large Λ-free balls near x0 imply large concentric Γ-free balls. The precise quantities and +bounds are given by Lemma 3.8 (one can use Lemma 3.9 to obtain a slightly cleaner statement). +24 + +Proposition 4.23. Let S ∈ (0, 1) be the constant given by Lemma 3.8, and let s ∈ (0, S). There is a bound +D = D(s) such that dγ > D implies that γx0 lies sdγ deep in HBΛ +λγ. +Proof. The proof is a bit delicate but very similar to that of Lemma 3.8. In essence, I use the Γ-free balls +near x0 to produce a Γ-free cylinder, which would force a certain geodesic not to cross a horosphere of Γ, +i.e. force it to stay inside a Γ-free horoball. +As in Lemma 4.21 it is only required to show that x′ = λ−1 +γ γx0 is sdγ deep inside HBΛ +i(γ). I start with +proving that x′ ∈ HBΓ +i(γ). I learned the hard way that even this is not a triviality. Recall the notation +Bγ = B(γx0, dγ). The ball B′ +γ = λ−1 +γ Bγ is a Λ-free ball of radius dγ about x′ = λ−1 +γ γx0. Denote by x′ +t +the point at time t along the unit speed geodesic η′ := [x0, x′]. It holds that |x′ +t| = t and, for t ≤ dγ, x′ +t +is the centre of a Λ-free ball of radius t tangent to x0. The constant s is fixed and by Lemma 3.8 there is +T ′ = T ′(s) such that if t > T ′, the ball sB +� +x′ +t, t +� +is Γ-free. +The next goal is to show that x′ +T ∈ HBΓ for some adequate T . For any time T > 0, let α = α(ε, T ) be the +angle for which d +� +η(T ), ηi(γ)(T ) +� +< ε for every η in the α-sector of vi(γ). By perhaps taking smaller α I may +assume that α is uniformly small as stated in Lemma 4.21. Let D(α) be the bound given by Lemma 4.21 +guaranteeing +D(α) < dγ ⇒ d +� +x′ +T , ηi(γ)(T ) +� +< ε +For my needs in this lemma ε may as well be chosen to be 1. I now choose a specific time T for which +I want x′ +T and ηi(γ)(T ) to be close. There are only finitely many Λ-free horoballs {HBΛ +i }i∈{1,...,N} tangent +to x0, giving rise to a uniform bound L = maxi∈{1,...,N}{d(HΛ +i , HΓ +i )} on the distance d(HΛ +i(γ), HΓ +i(γ)). Fix +T to be any time in the open interval (T ′ + L + ε, dγ). The fact that L + ε < T implies that ηi(γ)(T ) lies +at least ε-deep inside HBΓ +i(γ), and therefore η′(T ) ∈ HBΓ +i(γ). Recall that any point on HΓ is DΓ-close to a +point γx0 ∈ HΓ. By perhaps enlarging T and shrinking α if necessary, I may assume that DΓ < sT . Thus +for all T < t ≤ dγ, x′ +t is the centre of a Γ-free ball of radius st > sT > DΓ, hence {x′ +t}T ≤t≤dγ does not cross +a horosphere of Γ. Since x′ +T ∈ HBΓ +i(γ), this implies that x′ +t stays in HBΓ +i(γ) for all T < t ≤ dγ. In particular +x′ +dγ = x′ ∈ HBΓ +i(γ). +To get the result of the proposition, recall that sB′ +γ = B(x′, sdγ) is Γ-free, so x′ must be at distance at +least sdγ − DΓ from any horosphere of Γ, and in particular from HΓ +i(γ). In terms of Busemann functions, this +means that bηi(γ)(x′) ≤ −sdγ + DΓ whenever one can find such T ′ + L + ε < T < dγ. Since HBΛ +i(γ) is tangent +to x0, the corresponding horoball HBΓ +i(γ) lies inside it, and so x′ lies (sdγ − DΓ)-deep inside HBΛ +i(γ). A close +look at the argument yields the desired bound D = D(s) such that the above holds whenever dγ > D(s). +To help the reader take this closer look, I reiterate the choice of constants and their dependencies as they +appear in the proof: +1. Fix ε = 1. +2. Let T ′ = T ′(s) the constant from Lemma 3.8 and L = maxi∈{1,...,N}{d(HBΛ +i , HBΓ +i )}. +3. Fix T > T ′ + L + 1. +4. Fix α = α(1, T ). +5. Fix D(s) = max{D(α), T + 1}. +I remark, for the reader worried about the DΓ which appears in the final bound but not in the statement, +that (a) DΓ is a fixed universal constant and may as well be ignored, and (b) the discrepancy can be formally +corrected by taking a slightly larger s < s′ to begin with and as a result perhaps enlarging the bound D for +dγ). Also note that L = L(Λ) is a universal constant. +25 + +4.2.3 +Intersection of Λ-Free Regions and the Existence of a Λ-Cocompact Horosphere +In this section I find Λ-orbit points that lie close to the bounding horosphere of a Λ-free horoball HBΛ. In +order to find such points I need to make sure HBΛ is not contained inside a much larger Λ-free horoball. I +introduce the following definition. +Definition 4.24. A Λ-free horoball HBΛ is called maximal if it is tangent to a point x = λx0 ∈ Λ · x0. It +is called ε-almost maximal if d(Λ · x0, HΛ) < ε. +Remark 4.25. It may happen that a discrete group admits free but not no maximally free horoballs - see +discussion in section 4 of [55]. In any case it is clear that any Λ-free horoball can be ‘blown-up’ to an ε-almost +maximal Λ-free horoball, for every ε > 0. Moreover, every two ε-almost maximal horoballs based at the +same point ξ ∈ WQ(Γ) lie at distance at most ε of one another. For my needs any fixed ε would suffice, and +I fix ε = 1. +Lemma 4.26. There is DΛ > 0 such that if HBΛ is 1-almost maximal Λ-free horoball then HΛ ⊂ NDΛ(Λ·x0), +i.e. d(x, Λ · x0) ≤ DΛ for all x ∈ HΛ. +Notice that Lemma 4.26 does not state Λ · x0 even intersects HΛ. +Proof. I start with a short sketch of the proof. Consider a 1-maximal horoball and a point x on its bounding +horosphere with d(x, Λ · x0) = D. One may translate this situation to x0, which results in a Λ-free horoball +HBΛ intersecting the (closed) D-ball about x0 at a point w with B(w, D) Λ-free. +The proof differs depending on whether HBΓ ⊂ HBΛ or the other way round, since I use the bounds +from 4.18: +1. If HBΓ ⊂ HBΛ, there is a sublinear bound on d(HBΛ, HBΓ), which readily yields a bound on D. +2. if HBΛ ⊂ HBΓ there is a bound on d(x0, HBΓ) that is independent of D. So there are only finitely +many possibilities for HBΓ, independent of D. Hence there are only finitely many possible base points +for HBΓ. These in turn correspond to possible base points for such HBΛ, and this finiteness yields +a bound on the distance d(HBΓ, HBΛ) < L that is independent of D. The rest of the proof is quite +routine. +Let HBΛ be a 1-almost maximal Λ-free horoball. By definition there is λ ∈ Λ and z ∈ HΛ such that +d(λx0, z) < 1. Fix D > 0. I show that if there is some z′ ∈ HΛ for which d(z′, Λ · x0) ≥ D, then D must be +uniformly small. Exactly how small will be set in the course of the proof. +Fix D > 1 and assume that there is z′ ∈ HΛ with d(z, Λ · x0) ≥ D. Up to sliding z′ along HΛ, the +continuity of the function x �→ d(x, Λ · x0) together with Intermediate Value Theorem allows to assume that +d(z′, Λ · x0) = D. Let λ′ ∈ Λ be the element for which d(z′, λ′x0) = D. Translating by λ′−1 yields +1. A Λ-free horoball HBΛ +0 := λ′−1HBΛ. +2. A point w := λ′−1z′ ∈ HΛ +0 for which |w| = d(w, x0) = d(w, Λ · x0) = D. +Assume first that HBΓ +0 ⊂ HBΛ +0 . By Lemma 4.18 there is a sublinear function u′ such that d(HΓ +0 , HΛ +0 ) ≤ +u′(D). This yields a point γx0 ∈ HΓ +0 for which d(w, γx0) ≤ u′(D) + DΓ. Thus |γx0| ≤ D + u′(D) + DΓ and +the reverse triangle inequality gives +D − +� +u′(D) + DΓ +� +≤ d(w, λγx0) − d(w, γx0) < d(γx0, λγx0) +Together with the bound d(γx0, λγx0) ≤ u(|γx0|) and rearranging, one obtains +D ≤ u +� +D + u′(D) + DΓ +� ++ u′(D) + DΓ +The right hand side is clearly a sublinear function in D, hence this inequality may hold only for boundedly +small D, say D < D1. I conclude that HBΓ +0 ⊂ HBΛ +0 may occur only when D < D1. Notice that D1 depends +only on u and u′, and not on HBΛ. +26 + +w +x0 +Λ − free +B(w, D) +τ(t0) +τ +Λ − free HBΛ +ξ +Γ − free HBΓ +γx0 +Figure 3: Lemma 4.26, case HBΛ ⊂ HBΓ. The red horosphere of Γ is trapped between x0 and HΛ, and is +at distance t0 from x0. A Γ-orbit point on the red horosphere close to x0 allows to use sublinearity to get a +bound on t0. +Assume next that HBΛ +0 ⊂ HBΓ +0 , and that the containment is strict. Since x0 ∈ Γ · x0, the geodesic +τ := [x0, w] is of length D and intersects HΓ +0 . Denote by t0 ∈ [0, D) the time in which τ intersects HΓ +0 , and +let w′ := ��(t0) ∈ HΓ +0 be the intersection point. In particular |w′| = t0. It is clear that B(w′, t0) is Λ-free, +as a subset of the ball B(w, D). Again there is γx0 ∈ B(w′, DΓ) ∩ HΓ +0 and so |γx0| ≤ t0 + DΓ. By reverse +triangle inequality +t0 − DΓ ≤ d(w′, λγx0) − d(w′, γx0) ≤ d(γx0, λγx0) +and the sublinear constraint gives t0 − DΓ ≤ u(t0 + DΓ). This can only happen for boundedly small t0, +say t0 < T . I conclude that if HBΛ +0 ⊂ HBΓ +0 , then HBΓ +0 is a horoball of Γ tangent to some point y ∈ B(x0, T ). +By Corollary 2.22 there are finitely many horoballs of Γ tangent to points in B(x0, T ). In particular +there is a finite set {ξ′ +1, . . . , ξ′ +K} ∈ WQ(Γ) of possible base points for HBΓ +0. This set depends only on T , and +since the choice of T was completely independent of D, the set of possible base points is independent of D +as well. Let � +HBΓ +i be the horoball of Γ based at ξ′ +i. +I can now bound the distance d(HBΓ +0 , HBΛ +0 ). Let 1 ≤ i ≤ K be an index for which there is a Λ-free +horoball based at ξ′ +i that is contained in � +HBΓ +i . There is thus some 1-almost-maximal Λ-free horoball based at +ξ′ +i. Fix an arbitrary such 1-almost-maximal Λ-free horoball � +HBΛ +i for each such i, and let Li := d(� +HBΛ +i , � +HBΓ +i ). +Finally, define L := max{Li} + 1 among such i. As stated in Remark 4.25, d(HBΛ +0 , � +HBΛ +i ) ≤ 1 for some i, +therefore d(HBΓ +0, HBΛ +0 ) ≤ L. +Recall |w| = D and B(w, D) is Λ-free. It holds that d(w, HΓ +0 ) ≤ L, and so there is γx0 ∈ HΓ +0 for which +d(w, γx0) ≤ L + DΓ. In particular |γx0| ≤ D + L + DΓ (in fact it is clear that |γx0| ≤ T + DΓ, but this +won’t be necessary). Reverse triangle inequality gives +D − (L + DΓ) ≤ d(w, λγx0) − d(w, γx0) ≤ d(γx0, λγx0) +and from the sublinear constraint I conclude D − (L + DΓ) ≤ u(D + L + DΓ). Since L, DΓ are fixed +constants independent of D, this can only hold for boundedly small D, say D < D2. In particular, one gets +a uniform bound DΛ := max{D1, D2} such that x ∈ HΛ ⇒ d(x, Λ · x0) < DΛ. +Corollary 4.27. Every Γ-conical limit point is a Λ-conical limit point. +27 + +Proof. Let ξ ∈ X(∞) be a Γ-conical limit point. Let η : R≥0 → X be a geodesic with η(∞) = ξ. By +definition there is a bound D > 0 and sequences tn → ∞, γn ∈ Γ such that d +� +γnx0, η(tn) +� +< D. Consider the +corresponding λn := λγn and λnx0. If dn is uniformly bounded, then ξ is Λ-conical by definition. Otherwise, +assume dn is monotonically increasing to ∞. For some fixed s ∈ (0, 1) it holds that for all but finitely many +n ∈ N, γnx0 is sdn deep inside HBΛ +n := HBΛ +λn,i(γn). I assume dn is large enough so that sdn > D, and in +particular η(tn) ∈ HBΛ +n. Let ξn ∈ WQ(Γ) be the respective base points of HBΛ +n. The point ξ is Γ-conical, +and by Theorem 2.28 π +2 ≤ d(ξ, WQ(Γ)) ≤ d(ξ, ξn). +The proof differs depending on whether the above inequality is strict or not for any n ∈ N. Assume +first that for some m ∈ N, d(ξ, ξm) = π +2 . By item 2 of Lemma 2.31, d +� +HΛ +m, η(t) +� +is uniformly bounded, i.e., +there is C > 0 such that for every t > 0 there is xt ∈ HΛ +m for which d +� +xt, η(t) +� +< C. By Lemma 4.26, +d(xt, Λ · x0) < DΛ, hence d +� +η(tn), xtn +� +≤ C + DΛ. This means that ξ is Λ-conical. +Otherwise, for all n ∈ N it holds that π +2 < d(ξ, ξn). The fact that η(tn) ∈ HBΛ +n together with Lemma 2.31 +implies that at some later time the geodesic ray η leaves HBΛ +n. Thus there is sn > tn for which η(sn) ∈ HΛ +n. +Since HBΛ +n are maximal Λ-free horoballs, Lemma 4.26 gives rise to points λnx0 such that d +� +λnx0, η(sn) +� +≤ +DΛ. This renders ξ as a Λ-conical limit point, as wanted. +I now prove Proposition 4.12. +Proof of Proposition 4.12. The strategy is as follows. For HBΛ = HBΛ +λγ,i(γ), one uses Proposition 4.23 to +get that HBΓ ⊂ HBΛ and that the distance d(HΛ, HΓ) is large with dγ. The horosphere HΓ admits a +Γ-cocompact metric lattice, and so the projections of these metric lattice points onto HΛ form a cocompact +metric lattice in HΛ. It remains to show that for each γ′x0 ∈ Γ · x0 ∩ HΓ, the corresponding λ′ = λγ′ indeed +lies on the same HΛ and boundedly close to the projection PHΛ(γ′x0). This is done by putting together all +the geometric facts obtained up to this point, specifically Lemma 4.26. One delicate fact that will be of use +is that two maximal Λ-free horoballs that are based at the same point must be equal, because none of them +can contain a Λ-orbit point while on the other hand both bounding horospheres intersect Λ · x0. +Fix s > 0 for which Proposition 4.23 yields a corresponding bound D(s), and let γ ∈ Γ such that +sdγ > 2 · +� +DΛ + D(s) +� +. +Consider the (maximal) Λ-free horoball HBΛ +λγ,i(γ) based at ξi(γ) +λ +. +I show that +Λ · x0 ∩ HΛ +λγ,i(γ) is a cocompact metric lattice in HΛ +λγ,i(γ). I keep the subscript notation because the proof is +a game between HBΛ +λγ,i(γ) and another Λ-free horoball. +Let HBΓ +λγ,i(γ) be the Γ-horoball corresponding to HBΛ +λγ,i(γ). I can conclude that HBΓ +λγ,i(γ) ⊂ HBΛ +λγ,i(γ), +because the choice of dγ > D(s) guarantees γx0 is sdγ deep inside HBΛ +λγ,i(γ). In particular HBΛ +λγ,i(γ) is not +Γ-free. Moreover, it holds that L := d(HΛ +λγ,i(γ), HΓ +λγ,i(γ)) ≥ sdγ. Let γ′ ∈ Γ be any element in the cocompact +metric lattice Γ · x0 ∩ HΓ +λγ,i(γ), and consider two associated points: (a) λ′x0 = λγ′x0 and (b) the projection +of γ′x0 on HΛ +λγ,i(γ), denoted p′ +γ := PHΛ +λγ ,i(γ)(γ′x0) ∈ HΛ +λγ,i(γ). The horoball HBΛ +λγ,i(γ) is a maximal Λ-free +horoball so it is also 1-almost maximal, hence d(p′ +γ, Λ · x0) ≤ DΛ and the following holds: +sdγ ≤ L ≤ dγ′ ≤ d(γ′x0, p′ +γ) + d(p′ +γ, Λ · x0) ≤ L + DΛ +(3) +Consider ξi(γ′) +λ′ +, and assume towards contradiction that ξi(γ′) +λγ′ +̸= ξi(γ) +λγ . Both points lie in WQ(Γ) and +therefore must be at Tits distance π of each other. Therefore the fact that γ′x0 lies in HBΛ +λγ,i(γ) implies that +the geodesic [γ′x0, ξi(γ′)] leaves HBΛ +λγ,i(γ) at some point z ∈ HΛ +λγ,i(γ). +The fact that D(s) ≤ sdγ ≤ dγ′ implies that γ′x0 lies s2dγ deep inside HBΛ +λγ′ ,i(γ′). Therefore the point z +also lies at least s2dγ deep inside HBΛ +λγ′ ,i(γ′), and therefore z is the centre of a Λ-free horoball of radius at +least s2dγ. +By choice of dγ the point z therefore admits a 2DΛ neighbourhood that is Λ-free. But z lies on HΛ +λγ,i(γ), +a maximal horosphere of Λ, contradicting Lemma 4.26. I conclude that ξi(γ′) +λγ′ += ξi(γ) +λγ , so HBΛ +λγ,i(γ) and +28 + +γx0 +dγ +γ′x0 +λγ′x0 +HBΛ +λγ′ ,i(γ′) +z′ := πHBΛ +λγ′ ,i(γ′)(γ′x0) +D1 +���� +1 +ξi(γ′) +λγ′ +ξi(γ) +λγ +≥ 1 +2dγ +z +≥ 1 +2 dγ′ +Λ − free ball +B +� +z, 1 +2dγ +� +HBΛ +λγ,i(γ) +HBΓ +λγ,i(γ) +λγx0 +Figure 4: Proposition 4.12. Assuming towards contradiction that ξi(γ) +λγ +̸= ξi(γ′) +λγ′ +results in a point z ∈ HBΛ +λγ,i(γ) +(blue coloured and bold faced in the bottom part of the figure) admitting a large Λ-free neighbourhood, +contradicting almost cocompactness. +HBΛ +λγ′ ,i(γ′) are two Λ-free horoballs that are tangent to a Λ · x0 point and based at the same point at +∞. This implies HBΛ +λγ,i(γ) = HBΛ +λγ′ ,i(γ′), and in particular λγ′x0 ∈ HΛ. Finally, it is clearly seen from +Inequality 3 that +d(λ′x0, p′ +γ) ≤ d(λ′x0, γ′x0) + d(γ′x0, p′ +γ) ≤ dγ′ + L ≤ L + DΛ + L +The element γ′x0 ∈ Γ · x0 ∩ HΓ +λγ,i(γ) was as arbitrary element, and the above argument shows that the +corresponding Λ-orbit points satisfy: +1. λ′x0 all lie on HΛ +λγ,i(γ). +2. Each p′ +γ is 2L + DΛ close to the point λ′x0. +This shows that the cocompact metric lattice {p′ +γ | γ′x0 ∈ HΓ +λγ,i(γ)} lies in a bounded neighbourhood of +the set of points Λ · x0 ∩ HΛ +λγ,i(γ), proving that Λ · x0 ∩ HΛ +λγ,i(γ) is a cocompact metric lattice in HΛ +λγ,i(γ). +Lemma 2.24 elevates this to +� +Λ ∩ StabG(HΛ +λγ,i(γ)) +� +· x0 ∩ HΛ +λγ,i(γ) +being a cocompact metric lattice in HΛ +λγ,i(γ), completing the proof. +4.3 +The Bounded Case +Proposition 4.12 is enough in order to prove Theorem 4.1 in the case Γ ̸⊂ ND(Λ) for any D > 0, i.e. in case +Γ does not lie in a bounded neighbourhood of Λ. The case where Γ and Λ lie at bounded Hausdorff distance, +i.e. where Γ ⊂ ND(Λ) and Λ ⊂ ND(Γ), arose naturally in the context of the quasi-isometric classification +of non-uniform lattices in the works of Schwartz [52] (R-rank 1), Drut¸u [16] and Eskin [22] (higher rank). I +restate the theorems in the bounded setting. +29 + +Theorem 4.28 (Eskin [22], Theorem 4.9 above). Let G be a real centre-free semisimple Lie group without +compact factors and of higher rank, Γ ≤ G an irreducible non-uniform lattice, Λ ≤ G a discrete subgroup. If +Γ ⊂ ND(Λ) for some D > 0, then Λ is a lattice, and if moreover Λ ⊂ ND(Γ) then Λ and Γ are commensurable. +Theorem 4.29 (Schwartz [52], Theorem 4.10 above). Let G be a real simple Lie group of R-rank 1, Γ ≤ G +an irreducible non-uniform lattice, Λ ≤ G a discrete subgroup. If both Γ ⊂ ND(Λ) and Λ ⊂ ND(Γ) for some +D > 0, then Λ is a lattice, and if moreover G is not locally isomorphic to SL2(R), then Λ is commensurable +to Γ. +The notable difference between the two statements is that for higher rank groups, the inclusion Λ ⊂ ND(Γ) +is only required to prove commensurability. In view of Corollary 4.8, this allows me to omit that assumption +from Theorem 1.6. Notice also that for groups with property (T) the result easily follows from the (much +more recent) result by Leuzinger in Theorem 5.7. +In the context of commensurability in the sublinear setting, I can only prove a limited result, Namely +that Λ is commensurable to Γ if Γ is an irreducible Q-rank 1 lattice and both Γ ⊂ ND(Λ) and Λ ⊂ Nu(Γ) +for some constant D > 0 and a sublinear function u. This is done via a reduction to the bounded case. +Proposition 4.30. Let G be a real semisimple Lie group without compact factors and with finite centre, +Γ ≤ G an irreducible lattice of Q-rank 1, Λ ≤ G a discrete subgroup, and u a sublinear function. If Γ ⊂ ND(Λ) +for some D > 0 and Λ ⊂ Nu(Γ), then actually Λ ⊂ ND′(Γ) for some D′ > 0. Moreover, if G is of R-rank 1, +the conclusion holds under the relaxed assumption that u(r) ⪯∞ εr for some ε < 1. +Remark 4.31. While the setting of Proposition 4.2 is indeed rather limited, the situation that both Γ ⊂ +Nu(Λ) and Λ ⊂ Nu(Γ) arises naturally from the motivating example of SBE-rigidity in Theorem 6.9. Notice +however that Theorem 6.9 is not known for groups G that admit R-rank 1 factors, which is the only setting +for which I can prove Proposition 4.2. +4.3.1 +A Reduction +I start with the proof of Proposition 4.30. The first step is to establish the fact that Λ must preserve WQ(Γ). +Lemma 4.32. Let G be a real semisimple Lie group without compact factors and with finite centre, Γ ≤ G +an irreducible non-uniform lattice of Q-rank 1, Λ ≤ G a discrete subgroup. Assume that Γ ⊂ Nu(Λ) and +that Λ ⊂ Nu′(Γ) for sublinear functions u, u′. Then Λ · WQ(Γ) ⊂ WQ(Γ). Moreover, if G is of R-rank 1, the +conclusion holds under the relaxed assumption that u′(r) ⪯∞ εr for some ε < 1. +Proof. The proof is similar to the argument of Lemma 4.14, and uses the linear penetration rate of a geodesic +into a horoball. Let ξ ∈ WQ(Γ), and let HΓ be a horosphere bounding a Γ-free horoball HBΓ with HΓ ∩Γ·x0 +a metric lattice in HΓ. Assume first that u′ is sublinear. Since HBΓ is Γ-free and Λ ⊂ Nu′(Γ), I can conclude +that Λ · x0 ∩ HBΓ ⊂ Nu′(HΓ). Recall (Lemma 2.31) that every geodesic ray η with limit point ξ′ ∈ N π +2 (ξ) +penetrates HBΓ at linear rate. Therefore for every such geodesic ray η and every sublinear function v there +is R = R(η, v) > 0 for which Nv(η↾r>R) is Λ-free. +On the other hand, let λ ∈ Λ, and assume towards contradiction that λξ /∈ WQ(Γ). Then by Proposi- +tion 2.30 there is a Γ-conical limit point ξ′ ∈ N π +2 (λξ). The hypothesis that Γ ⊂ Nu(Λ) then implies that for +every R > 0, Nu(η↾r>R) ∩ Λ · x0 ̸= ∅. Translating by λ−1 yields a contradiction to the previous paragraph. +I conclude that λξ ∈ WQ(Γ). +I now modify the argument to include u′(r) ⪯∞ εr when G is of R-rank 1. In this case, the only point +ξ′ ∈ N π +2 (λξ) is λξ itself. Therefore by the same argument as above, the assumption that λξ /∈ WQ(Γ) +implies that the u-sublinear neighbourhood of every geodesic ray with limit point ξ intersects Λ · x0. I.e., for +every η with limit point ξ and every R > 0 it holds that Nu(η↾r>R) ∩ Λ · x0 ̸= ∅. On the other hand, every +such geodesic penetrates HBΓ at 1-linear rate. This amounts to the following fact: if v′(r) = εr for some +ε ∈ (0, 1), then for some R > 0, the set Nu(η↾r>R)∩Nv′(HΓ) = ∅. This is a contradiction to Λ ⊂ Nu′(Γ). +30 + +Proof of Proposition 4.30. Assume towards contradiction that there is a sequence λn such that d(λnx0, Γ · +x0) > n. Recall that Γ·x0 is a cocompact metric lattice in the compact core of Γ. This implies that there is a +number D′ > 0 such that any λ ∈ Λ for which λx0 /∈ ND′(Γ · x0) must lie at least 1 +2D′-deep inside a horoball +of Γ. I can assume that for all n ∈ N there are corresponding horoballs of Γ, which I denote HBΓ +n, for which +λn · x0 ∈ HBΓ +n. The fact that Γ ⊂ ND(Λ) then implies that ND(Λ · x0) covers a cocompact metric lattice in +HΓ +n, namely the metric lattice Γ · x0 ∩ HΓ +n. In the terminology of Section 4.2, HΓ +n is almost Λ-cocompact, or +D-almost Λ-cocompact. +I first prove that every horoball of Γ contains a Λ-free horoball (this is of course immediate if Λ ⊂ NC(Γ) +for some C > 0). Assume towards contradiction that there is a horoball HBΓ of Γ that does not contain a +Λ-free horoball. Denote HΓ := ∂HBΓ. In the notations of the previous paragraph, I can assume without loss +of generality that HBΓ = HBΓ +n for all n ∈ N. Denote by ξ the base point of HBΓ, fix some arbitrary x ∈ HΓ +and consider the geodesic ray η := [x, ξ). The constraint that Λ ⊂ Nu(Γ) implies that for every R > 0 there +is some L > 0 for which the ball B +� +η(L + t), R +� +is Λ-free, for all t ≥ 0. In particular, for all large enough +n ∈ N (depending on R), the horosphere H(ξ, λnx0) that is parallel to HΓ and that passes through λnx0 +contains a point that is the centre of Λ-free ball of radius R. This property is Λ-invariant, as well as the fact +that HBΓ is based at WQ(Γ). In particular, these two properties hold for the horoballs HBn := λ−1 +n +· HBΓ, +whose respective base points I denote ξn := λ−1 +n ξ ∈ WQ(Γ). +Fix R = D +2DΓ (recall that DΓ is such that every horosphere H of Γ admits H ⊂ NDΓ(Γ·x0 ∩H)). Let +L = L(D+2DΓ) be the corresponding bound from the previous paragraph. For every n > L the horoball HBn +has bounding horosphere Hn that admits a point zn ∈ Hn for which B(zn, D+2DΓ) is Λ-free. Moreover, the +same is true for every horosphere that is parallel to Hn which lies inside HBn. Since Γ ⊂ ND(Λ), this means +that every horosphere that lies inside HBn admits a point that is the centre of a Γ-free ball of radius 2DΓ. +I conclude that none of those horospheres could be the horosphere of Γ corresponding to the parabolic limit +point ξn ∈ WQ(Γ). Since x0 ∈ Hn it must therefore be that Hn is a horosphere of Γ. But this contradicts +the fact that zn ∈ Hn and B(zn, 2DΓ) is Γ-free. This shows that no horoball of Γ contains a sequence of +Λ-orbit points that lie deeper and deeper in that horoball. Put differently, it shows that every horoball of Γ +contains a Λ-free horoball. +I remark that the above argument shows something a bit stronger, which I will not use but which I find +illuminating. It proves that as soon as d(λx0, Γ · x0) is uniformly large enough, say more than M, then +λx0 must lie on a (D + 2DΓ)-almost Λ-cocompact horosphere parallel to HΓ, where HΓ is the bounding +horosphere of any horoball of Γ in which λx0 lies (recall that it must lie in at least one such horoball). On +the other hand if d(λx0, Γ · x0) < M, then since every point in the Γ-orbit lies on a horosphere of Γ one +concludes that λx0 lies on a horosphere H based at WQ(Γ) that is (M + D)-almost Λ-cocompact. +I can now assume that every HBΓ +n contains a Λ-free horoball. In particular it contains a 1-almost maximal +Λ-free horoball HBΛ +n (see Definition 4.24). By definition there is a point λ′ +nx0 that is at distance at most 1 +from HΛ +n = ∂HBΛ +n. Up to enlarging d(λnx0, Γ · x0) or decreasing it by at most 1, I can assume λn = λ′ +n to +begin with. Consider HBn := λ−1 +n +· HBΛ +n with Hn = ∂HBn. This is a sequence of horoballs, each of which +contains a Λ-free horoball at depth at most 1, based at corresponding parabolic limit points ξn ∈ WQ(Γ), +and tangent to points that are at distance at most 1 from x0, i.e., Hn ∩ B(x0, 1) ̸= ∅. +Since Γ ⊂ ND(Λ) I conclude that each of the HBn contain a horoball of depth at most D + 1 that +is Γ-free. Therefore the horoball of Γ that is based at ξn must have its bounding horosphere intersecting +B(x0, D + 2). By Corollary 2.22 there are only finitely such horoballs. I conclude that there are finitely +many points ξ′ +1, . . . , ξ′ +K ∈ WQ(Γ) such that for every n ∈ N there is i(n) ∈ {1, . . . , K} with ξn = ξ′ +i(n). From +the Pigeonhole Principle there is some ξ′ ∈ {ξ′ +1 . . . , ξ′ +K} for which ξn = ξ′ for infinitely many n ∈ N. Passing +to a subsequence I assume that this is the case for all n ∈ N. +To begin with the HBΓ +n are horoballs of Γ, and therefore as in the first case the bounding horospheres +HΓ +n are D + 2DΓ-almost Λ-cocompact. This is a Λ-invariant property and therefore the same holds for +the λ−1 +n +translate of it. These are the horospheres which are based at ξ′ and lie outside HBn at distance +d(λnx0, Γ · x0) > n − 1 from Hn. They form a sequence of outer and outer horospheres based at the same +point at WQ(Γ), all of which are D + 2DΓ-almost Λ-cocompact. This is a contradiction, since the union of +such horospheres intersect every horoball of X, contradicting the existence of Λ-free horoballs. Formally, +31 + +take some ζ ∈ WQ(Γ) different from ξ′. Since both ξ′ and ζ lie in WQ(Γ), they admit dT (ζ, ξ′) = π and +there is a geodesic η with η(−∞) = ξ′ and η(∞) = ζ. Let HBΓ +ζ be the horoball of Γ that is based at ζ. By +the first step of this proof, every such horoball must contain a Λ-free horoball HBΛ +ζ . Therefore there is some +T > 0 such that for all t > T the point η(t) lies 2(D + 2DΓ) deep in HBΛ +ζ . I conclude that for all t > T , +B +� +η(t), 2D +� +is Λ-free. On the other hand, for arbitrarily large t it holds that the horosphere based at ξ′ and +tangent to η(t) is D+2DΓ-almost Λ-cocompact, and in particular d +� +η(t), Λ·x0 +� +< D+2DΓ, a contradiction. +I conclude that Λ ⊂ ND′(Γ) for some D′ > 0, as claimed. +Corollary 4.33. In the setting of Proposition 4.30, Λ is a lattice commensurable to Γ. +4.3.2 +The Arguments of Schwartz and Eskin +The R-rank 1 case. +The statement of Theorem 4.10 is a slight modification of his original formulation. +His framework leads to a discrete subgroup ∆ ≤ G such that: +1. Every element of ∆ quasi-preserves the compact core of the lattice Γ. Namely, each element of ∆ is +an isometry of X that preserves WQ(Γ) and that maps every horosphere of Γ to within the D = D(∆) +neighbourhood of some other horosphere of Γ. +2. It holds that Γ ⊂ ND(∆). +From these two properties Schwartz is able to deduce that ∆ has finite covolume, i.e. that ∆ is a lattice +in G. Here is a sketch of his argument, which works whenever Γ is a Q-rank 1 lattice. +Theorem 4.34. In the setting described above, ∆ is a lattice in G. +Proof sketch. Consider X′ +0 := � +g∈∆ g · X0, where X0 is the compact core of Γ. +This space serves as a +‘compact core’ for ∆: the fact that ∆ quasi-preserves X0 implies that X′ +0 ⊂ ND(X0). It is a ∆-invariant +space, and therefore one gets an isometric action of ∆ on X′ +0. This action is cocompact: the reason is that +Γ acts cocompactly on X0, and Γ ⊂ ND(∆). Formally, every point in X′ +0 is D-close to a point in X0. Every +point in X0 is DΓ-close to a point in Γ · x0. Every point in Γ · x0 is D-close to a point in ∆ · x0. Therefore +the ball of radius 2D + DΓ contains a fundamental domain for the action of ∆ on X′ +0. +It remains to see that the action of ∆ on X \ X′ +0 is of finite covolume. As a result of the cocompact +action of ∆ on X′ +0, there is B := B(x0, R) so that X′ +0 ⊂ ∆·B. X′ +0 is the complement of a union of horoballs, +which one may call horoballs of ∆, with bounding horospheres of Λ. The fact that Γ is of Q-rank 1 means +that the horoballs of Γ are disjoint, and therefore those of Λ are almost disjoint: there is some C > 0 such +that for every horosphere H of Λ and every point x ∈ H, d(x, X′ +0) < C. Up to enlarging the radius of B by +C, I can assume that H ⊂ ∆ · B for every horosphere H of Λ. +Each horoball of Λ is based at WQ(Γ), and each lies uniformly boundedly close to the corresponding +horoballs of Γ. From Corollary 2.22 one therefore sees that there are finitely many horoballs of ∆ that inter- +sect B. Denote them by HB1, . . . , HBN, their bounding horospheres by Hi = ∂HBi, and their intersection +with B by Bi := B ∩ HBi. Let also ξi ∈ WQ(Γ) denote the base point of each HBi. Each Bi is pre-compact +and therefore the projection of each Bi on Hi is pre-compact as well (this is a consequence e.g. of the results +of Heintze-Im hof recalled in Remark 2.5). Let Di ⊂ Hi be a compact set that contains this projection, i.e. +PHi(Bi) ⊂ Di ⊂ Hi. In particular B ∩ Hi ⊂ Di. +Observe now that for every horoball HB of ∆, with bounding horosphere H = ∂HB, the ∆-orbit of every +point x ∈ H intersects some Di. First notice that for x ∈ H the choice of B implies that the ∆-orbit of x +must intersect B, say gx ∈ B. In particular gH ∩ B ̸= ∅, and since gHB is a horoball of ∆ then by definition +gH = Hi for some i ∈ {1, . . ., N}. One concludes that indeed gx ⊂ Di. +Moreover, let y ∈ X is any point that lies inside a horoball HB of ∆, and x = PH(y) its projection on +the bounding horosphere H = ∂HB. By the previous paragraph there is some g ∈ ∆ and i ∈ {1, . . . , N} for +which gx ∈ Di, and therefore it is clear that gy lies on a geodesic emanating from Di to ξi. +32 + +Finally, define Cone(Di) to be the set of all geodesic rays that emanate from Di and with limit point ξi. +The previous paragraph proves that �N +i=1 Cone(Di) contains a fundamental domain for the action of ∆ on +X \ X′ +0. Moreover, the fact that Di ⊂ Hi is compact readily implies that each Cone(Di) has finite volume, +and so this fundamental domain is of finite volume. +To conclude, B ∪ +� �N +i=1 Cone(Di) +� +is a set of finite volume and it contains a fundamental domain for the +∆-action on X, as claimed. The proof of commensurability of ∆ and Γ is given in full in [52]. +There is one essential difference between Theorem 4.10 and Theorem 4.34, namely the assumption that +Λ ⊂ ND(Γ) rather than quasi-preserving the compact core of Γ. In Schwartz’s work, the fact that ∆ ⊂ ND(Γ) +is not relevant (even though it easily follows from the construction of his embedding of ∆ in G). He only +uses the two properties described above, namely the quasi-preservation of X0 and Γ ⊂ ND(∆). +The assumption that Λ quasi-preserves the compact core of Γ does not feel appropriate in the context +of my thesis, while the metric condition Λ ⊂ ND(Γ) seems much more natural. It is a stronger condition +as I now show. By Lemma 4.32, Λ · WQ(Γ) ⊂ WQ(Γ). Let HΓ +1 be a horosphere of Γ, based at ξ ∈ WQ(Γ), +and let γx0 ∈ HΓ +1 be some point on the metric lattice of Γ · x0 on HΓ +1 . There is an element λ ∈ Λ such that +d(λx0, γx0) < D. Moreover, since Λ ⊂ ND(Γ) one knows that the parallel horoball that lies D-deep inside +HBΓ +1 is Λ-free. +Let λ′ ∈ Λ be an arbitrary element of Λ, and consider λ′ · HΓ +1 . +The last statement in the previous +paragraph is Λ-invariant, and so the horoball that lies D-deep inside λ′ · HBΓ +1 is Λ-free. +The fact that +Γ ⊂ ND(Λ) then implies that the parallel horoball that lies 2D deep inside λ′HBΓ +1 is Γ-free. Let HΓ +2 be the +horosphere of Γ that is based at λ′ξ. The last statement amounts to saying that HΓ +2 lies at most 2D-deep +inside λ′HBΓ +1 . On the other hand, one has d(λ′λx0, λ′HΓ +1 ) = d(λx0, HΓ +1 ) ≤ D, so there is a Λ-orbit point that +lies within D of λ′HΓ +1 . The parallel horoball that lies D-deep inside HBΓ +2 must also be Λ-free, so I conclude +that HΓ +2 must be contained in the parallel horoball to λ′HBΓ +1 which contains it and that is at distance D +from it. I conclude that d(λ′HΓ +1 , HΓ +2 ) ≤ 2D, and so that Λ quasi-preserves X0. +Remark 4.35. It is interesting to note that Schwartz’s arguments are similar in spirit to my arguments +in Section 4.2. In fact, one could also prove Theorem 4.10 using the same type of arguments that appear +repeatedly in section 4.2, namely by moving Λ-free horoballs around the space, specifically the proof of +Proposition 4.30. I do not present it here. +Higher rank. +Eskin’s proof is ergodic, and based on results of Mozes [42] and Shah [54]. I produce it here +without the necessary preliminaries, which are standard. +Proof of Theorem 4.9. To prove that Λ is a lattice amounts to finding a finite non-zero G-invariant measure +on Λ\G. By Theorem 2 in [42], if P ≤ G is a parabolic subgroup then every P-invariant measure on Λ\G is +automatically G-invariant. Fix a minimal parabolic subgroup P ≤ G and let µ0 be some fixed probability +measure on Λ\G. Since P is amenable it admits a tempered Følner sequence Fn ⊂ P, and one can average +µ0 along each Fn to get a sequence of probability measures µn. The weak* compactness of the unit ball +in the space of measures on Λ\G implies that there exists a weak* limit µ of the µn. The measure µ is +automatically a finite P-invariant measure. It remains to show that µ is not the zero measure. To see this +it is enough to show that for some compact set CΛ ⊂ Λ\G and some Λg = x ∈ Λ\G, one has +0 < lim inf +n +1 +|Fn| +� +Fn +1CΛ(xp−1)dp +(4) +Fix some compact neighbourhood CΓ ⊂ Γ\G of the trivial coset Γe. The hypothesis Γ ⊂ ND(Λ) implies +that there is a corresponding compact neighbourhood CΛ ⊂ Λ\G of the trivial coset Λe such that for any +p ∈ P, it holds that Γgp−1 ∈ CΓ ⇒ Λgp−1 ∈ CΛ (simply take CΛ to be the D + 1-blowup of CΓ). The action +of P on Γ\G is uniquely ergodic, therefore +0 < µΓ(CΓ) = lim +n +1 +|Fn| +� +Fn +1CΓ(Γp−1)dp +33 + +where µΓ denotes the natural G-invariant measure on Γ\G. The defining property of CΛ ensures that +Inequality (4) is satisfied, implying that µ is a non-zero P-invariant probability measure on Λ\G. I conclude +that µ is also G-invariant, and that Λ is a lattice. If moreover Λ ⊂ ND(Γ), one may use Shah’s Corollary +[54] to conclude that Λ is commensurable to Γ. +4.4 +Translating Geometry into Algebra +The goal of this section is to prove that the results of Section 4.2 imply that Λ satisfies the hypotheses +of the Benoist-Miquel criterion Theorem 4.4. +Namely, that Λ is Zariski dense, and that it intersects a +horospherical subgroup in a cocompact indecomposable lattice. +These are algebraic properties, and the +proof that Λ satisfies them is in essence just a translation of the geometric results of Section 4.2 to an +algebraic language. The geometric data given by Section 4.2 is that for some horosphere H bounding a +Λ-free horoball, Λ ∩ StabG(H) · x0 intersects H in a cocompact metric lattice (Proposition 4.12), and that +the set of Λ-conical limit points contains the set of Γ-conical limit points (Corollary 4.27). Note that since +K is compact the former implies that Λ ∩ StabG(H) is a uniform lattice in StabG(H). +4.4.1 +A Horospherical Lattice +I assume that Λ ∩ StabG(H) is a lattice in StabGH, and I want to show that Λ intersects a horospherical +subgroup U of G in a lattice. This step requires quite a bit of algebraic background, which I give below in full. +In short, the first goal is to show that StabG(H) admits a subgroup U ≤ StabG(H) that is a horospherical +subgroup of G. A lemma of Mostow (Lemma 4.36 below) allows to conclude that Λ intersects U in a lattice. +Lemma 4.36 (Lemma 3.9 in [40]). Let H be a Lie group having no compact connected normal semisimple +non-trivial Lie subgroups, and let N be the maximal connected nilpotent normal Lie subgroup of H. Let +Γ ≤ H be a lattice. Then N/N ∩ Γ is compact. +Remark 4.37. In the original statement Mostow uses the term ‘analytic group’, which I replaced here with +‘connected Lie subgroup’. This appears to be Mostow’s definition of an analytic group. See e.g. Section +10, Chapter 1 in [33]. In Chevalley’s Theory of Lie Groups, he defines a Lie group as a locally connected +topological group whose identity component is an analytic group (Definition 1, Section 8, Chapter 4 in [12]), +and proves (Theorem 1, Section 4, Chapter 4 therein) a 1-1 correspondence between analytic subgroups of +an analytic group and Lie subalgebras of the corresponding Lie algebra. +Lemma 4.36 lays the rationale for the rest of this section. Explicitly, I prove that StabG(H) admits +a subgroup that is a horospherical subgroup U of G (Corollary 4.39), and that U is maximal connected +nilpotent normal Lie subgroup of StabG(H) (Corollary 4.45). +In order to use Lemma 4.36, I show that the horospherical subgroup Nξ is a maximal normal nilpotent +connected Lie subgroup of StabG(H)◦, and that StabG(H)◦ admits no compact normal factors. This requires +to establish the structure of StabG(H)◦. +Definition 4.38. In the notation ht +ξ = exp(tX) and Aξ = exp +� +Z(X) ∩ p +� +of Proposition 2.7, define A⊥ +ξ to +be the codimension-1 submanifold of Aξ that is orthogonal to {hξ(t)}t∈R (with respect to the Killing form +in the Lie algebra). +Claim. Every element a ∈ A⊥ +ξ stabilizes H = H(x0, ξ). +Proof. An element in Aξ is an element that maps x0 to a point on a flat F ⊂ X that contains the geodesic +ray [x0, ξ). If a ∈ A⊥ +ξ , then the geodesic [x0, ax0] is orthogonal to [x0, ξ), and lies in F. From Euclidean +geometry and structure of horospheres in Euclidean spaces, it is clear that ax0 ∈ H(x,ξ). Since a ∈ Gξ, this +means aH = H(ax0, ξ) = H(x, ξ) = H. +Corollary 4.39. Let H be a horosphere based at ξ. Then StabG(H)◦ = (KξA⊥ +ξ )◦Nξ, and in particular it +contains a horospherical subgroup of G. Moreover, StabG(H)◦ is normal in StabG(ξ)◦ and acts transitively +on H. +34 + +Proof. Clearly (KξA⊥ +ξ )◦Nξ is a codimension-1 subgroup of StabG(ξ)◦. Since StabG(H) ̸= StabG(ξ) (e.g. +ht +ξ /∈ StabG(H) for t ̸= 0), it is enough to show that (KξA⊥ +ξ )◦Nξ ≤ StabG(H). Let kan ∈ (KξA⊥ +ξ )◦Nξ. It +fixes ξ, so it is enough to show that kanx0 ∈ H. Since k ∈ Kξ and kx0 = x0, it stabilizes H. From Claim 4.4.1 +a ∈ StabG(H). So it remains to check that Nξ stabilizes H, but this is more or less the definition: fixing +a base point x0, the horospheres based at ξ are parameterized by R. +Denote them by {Ht}t∈R, where +H = H0. In this parameterization, any element g ∈ Gξ acts on {Ht}t∈R by translation. I can thus define for +g ∈ StabG(ξ) the real number l(g) to be that number for which gHt = Ht+l(g). Clearly l +� +hξ(t) +� += t. The +element n fixes ξ, so one has +h−t +ξ nht +ξH0 = h−t +ξ Ht+l(n) = Ht+l(n)−t = Hl(n) +The fact that n ∈ Ker(Tξ), i.e. that limt→∞ h−t +ξ nht +ξ = eG readily implies that necessarily l(n) = 0. I +conclude that (KξA⊥ +ξ )◦Nξ = StabG(H)◦, as wanted. +Next recall that StabG(H)◦ acts transitively on X. Let x, y ∈ H, and consider g ∈ StabG(ξ)◦ with gx = y. +Writing an element g ∈ Gξ as kata⊥n ∈ Kξht +ξA⊥ +ξ Nξ, the argument above shows that kht +ξa⊥nH0 = H0 if +and only if t = 0, i.e., if and only if g ∈ StabG(H)◦. +Finally, let g ∈ StabG(ξ) and h ∈ StabG(H). By the discussion above h · Ht = Ht for all t ∈ R. Clearly +−l(g) = l(g−1), and therefore +ghg−1H0 = ghH−l(g) = g · H−l(g) = H0 +Therefore StabG(H) is normal in StabGξ, and the same is true for the respective identity components. +Corollary 4.40. StabG(H)◦ is a connected Lie group with no connected compact normal semisimple non- +trivial Lie subgroups. +Proof. Every compact subgroup of G fixes a point. Let H ≤ G be some closed subgroup. It is standard to +note that a normal N ≤ H that fixes a point x ∈ X must fix every point in the orbit H · x: hnh−1hx = hx. +Since H = StabG(H)◦ acts transitively on H, it shows that a normal compact subgroup of StabG(H)◦ fixes +every point in H. An isometry fixing a horosphere pointwise while fixing its base point is clearly the identity, +proving the claim. +The following fact is well known but I could not find it in the literature. +Corollary 4.41. A horosphere in X is not convex. +Proof. Let H′ be some horosphere in X, with base point ζ ∈ X(∞), and assume towards contradiction that +it is convex. Fix x ∈ H′ and a′ +t the one parameter subgroup with η′(∞) = a′ +tx, and denote H′ +t = H(a′ +tx, ζ). +Let eG ̸= n ∈ Nζ (Nζ defined with respect to a′ +t in a corresponding Langlands decomposition), and consider +the curve η′ +n(t) := a′ +tnx. I claim that this is a geodesic. On the one hand, the fact that H′ is convex implies +that the geodesic segment [x, nx] is contained in H′. Therefore a′ +t[x, nx] = [a′ +tx, a′ +tnx] ⊂ H′ +t. More generally +it is clear that because a′ +tH′ +s = H′ +s+t it holds that H′ +t is convex for every t as soon as it is convex for some t. +On the other hand, for every point y ∈ [x, nx], d(y, H′ +t) = t, and more generally for any y ∈ [a′ +snx, a′ +sx] +it holds that d(y, H′ +t) = |s − t|. In particular this is true for η′ +n(t) = yt := a′ +tnx. I get that d +� +η′ +n(t), η′ +n(s) +� += +|s − t|. Therefore η′ +n is a geodesic (to be pedantic one has to show that η′ +n is a continuous curve, which is a +result of the fact that a′ +t is a one parameter subgroup of isometries). Clearly +d +� +η′ +n(t), η′(t) +� += d(a′ +tnx, a′ +tx) = d(nx, x) +and therefore η′ +n is at uniformly bounded distance to η′. This bounds d(ηn, η′ +n) as bi-infinite geodesics, +i.e. for all t ∈ R, not just as infinite rays. The Flat Strip Theorem (Theorem 2.13, Chapter 2.2 in [11]), then +implies that the geodesics ηn, η′ +n bound a flat strip: an isometric copy of R × [0, l] (where l = d(x, nx)). +Up to now I did not use the fact that n ∈ Nξ, only that the point nx lies on a geodesic that is contained +in H′ = H′ +0. Therefore the entire bi-infinite geodesic that is determined by [x, nx] lies on a 2-dimensional +flat F that contains η′. The two elements n, a′ +t therefore admit nx, a′ +tx ∈ F. It is a fact that two such +elements must commute. I can conclude therefore that [n, a′ +t] = eG, which contradicts the fact that that +n ∈ Nζ = Ker(Tζ). +35 + +Lemma 4.42 (Theorem 11.13 in [51]). Let N be a connected real Lie group. Then Lie(N) is a nilpotent Lie +algebra if and only if N is a nilpotent Lie group. +Proposition 4.43 (Proposition 13, Section 4, Chapter 1 in [7]). In the notation of Proposition 2.7, nξ = +Lie(Nξ) is a maximal nilpotent ideal in gξ = Lie(Gξ). +Remark 4.44. +1. The presentation of nξ in [8] is given by means of the root space decomposition of +StabG(ξ), that appears in Proposition 2.17.13 in [20]. +2. There are two main objects in the literature that are referred to as the nilpotent radical or the nilradical +of a Lie algebra. These are: (a) the maximal nilpotent ideal of the Lie algebra, and (b) the intersection +of the kernels of all irreducible finite-dimensional representations. +Proposition 13 in Section 4 of +Chapter 9 in [7] shows that in the case of Lie algebras of parabolic Lie groups, these notions coincide. +Corollary 4.45. Nξ is a maximal connected nilpotent normal Lie subgroup of the identity connected com- +ponent StabG(H)◦. +Proof. Lemma 4.42 implies Nξ is nilpotent. Since StabGH ⊳ StabG(ξ), every normal subgroup of +StabG(H) containing Nξ is in fact a normal subgroup of StabG(ξ), still containing Nξ. It remains to +prove maximality of Nξ among all connected nilpotent normal Lie subgroups of StabG(ξ). Any such +subgroup N ′ ⊳ StabG(ξ) gives rise to an ideal n′ of gξ = Lie +� +StabG(ξ) +� +, and by Lemma 4.42 it is a +nilpotent ideal. Therefore by Proposition 4.43 it is contained in nξ = Lie(Nξ), implying that N ′ ≤ Nξ. +Corollary 4.46. A lattice in StabG(H) intersects the horospherical subgroup Nξ in a lattice. +Proof. Corollaries 4.40 and 4.45 imply that the pair Nξ ⊳ StabG(H) satisfy the hypotheses of Mostow’s +Lemma 4.36. +4.4.2 +Indecomposable Horospherical Lattices +The Benoist and Miquel criterion requires the horospherical lattice to be indecomposable. It is shown in [5] +that if this lattice is contained in a Zariski dense discrete subgroup, then the indecomposability condition is +equivalent to irreducibility of the ambient group. The precise definitions and statements are as follows. +Definition 4.47 (Definition 2.14 in [5]). For a semisimple real algebraic Lie group G and U a horospherical +subgroup of G, let ∆U be a lattice in U. +1. ∆U is irreducible if for any proper normal subgroup N of G◦, one has ∆U ∩ N = {e}. +2. ∆U is indecomposable if one cannot write G◦ as a product G◦ = N ′N ′′ of two proper normal subgroups +N ′, N ′′ ⊳ G with finite intersection such that the group +∆′ +U := (∆U ∩ N ′)(∆U ∩ N ′′) +has finite index in ∆U. +Definition 4.48 (See Section 2.4.1 in [5]). Let G be a semisimple real algebraic Lie group. A discrete +subgroup Λ ≤ G is said to be irreducible if, for all proper normal subgroups N ⊳ G, the intersection Λ ∩ N +is finite. +Lemma 4.49 (Lemma 4.3 in [5]). Let G be a semisimple real algebraic Lie group, U ⊂ G a non-trivial +horospherical subgroup, and ∆U ≤ U a lattice of U which is contained in a discrete Zariski dense subgroup +∆ of G. Then the following are equivalent: +1. ∆ is irreducible. +2. ∆U is irreducible. +3. ∆U is indecomposable. +36 + +4.4.3 +Zariski Density +The last requirement is for Λ to be Zariski dense. I use a geometric criterion which is well known to experts. +Lemma 4.50 (Proposition 2 in [31]). Let X be a symmetric space of noncompact type, G = Isom(X)◦. A +subgroup ∆ ≤ G is Zariski dense if and only if: +1. ∆ does not globally fix a point in X(∞), i.e. ∆ ̸≤ StabG(ζ) for any ζ ∈ X(∞). +2. The identity component of the Zariski closure of ∆ does not leave invariant any proper totally geodesic +submanifold in X. +In the proof I use several facts - mostly algebraic, and two geometric. I warmly thank Elyasheev Leibtag +for his help and erudition in algebraic groups. The first property I need is very basic. +Lemma 4.51. Let ∆ ≤ G be a discrete subgroup, and let H ≤ G be the Zariski closure of ∆. Then ∆ ∩ H◦ +is of finite index in ∆. +Proof. H◦ is normal and of finite index in H. +The following fact is probably known to experts. It appears in a recent work by Bader and Leibtag[2]. +Lemma 4.52 (Lemma 3.9 in [2]). Let k be a field, G a connected k algebraic group, P ≤ G = G(R) a +parabolic subgroup. Then the centre of G contains the centre of P. +Still on the algebraic side, I need a Theorem of Dani, generalizing the Borel Density Theorem. +Theorem 4.53 (See [15]). Let S be a real solvable algebraic group. If S = S(R) is R-split, then every lattice +ΓS ≤ S is Zariski dense. +Remark 4.54. It is a fact (see Theorem 15.4 and Section 18 in [6]) that: +1. Every unipotent group over R is R-split. +2. For a field k of characteristic 0, a solvable linear algebraic k-group is k-split if and only if its maximal +torus is k-split. +Finally I need two geometric facts. The first is a characterization determining when does a unipotent +element belongs to Nζ for some ζ ∈ X(∞). +Proposition 4.55 (Proposition 4.1.8 in [20]). Let X be a symmetric space of noncompact type and of higher +rank, n ∈ G = Isom(X)◦ a unipotent element, and ζ ∈ X(∞). The following are equivalent: +1. For Nζ as in Proposition 2.7, n ∈ Nζ. +2. For some geodesic ray η with η(∞) = ζ it holds that limt→∞ d +� +nη(t), η(t) +� += 0. +3. For every geodesic ray η with η(∞) = ζ it holds that limt→∞ d +� +nη(t), η(t) +� += 0. +The last property I need is a characterization of the displacement function for unipotent elements. +Proposition 4.56 (See proof of Proposition 3.4 in [4]). Let X be a symmetric space of noncompact type, +ζ ∈ X(∞) some point and n ∈ Nζ an element of the unipotent radical of StabG(ζ). +The displacement +function x �→ d(nx, x) is constant on horospheres based at ζ, and for every ε > 0 there is a horoball HBε +based at ζ such that d(nx, x) < ε for every x ∈ HBε. +Corollary 4.57. Assume that: +1. +� +Λ ∩ StabG(H) +� +· x0 is a cocompact metric lattice in a horosphere H ⊂ X bounding a Λ-free horoball. +37 + +2. Every Γ-conical limit point is a Λ-conical limit point. +Then Λ is Zariski dense. +Proof. I show the criteria of Lemma 4.50 are met, starting with Λ ̸≤ StabG(ζ) for any ζ ∈ X(∞). To this +end, I first prove that Λ · x0 is not contained in any bounded neighbourhood of any horosphere H′. Let +ξ′ be the base point of H′. By Hattori’s Lemma 2.31 (and Remark 2.32), it is enough to find a Λ-conical +limit point ζ′ with dT (ξ′, ζ′) ̸= π +2 . Take some ζ′′ ∈ X(∞) at Tits distance π of ξ′, i.e. take a flat F on +which ξ′ lies and let ζ′′ be the antipodal point to ξ′ in F. +Fix ε = +π +4 . +By Proposition 2.3, there are +neighbourhoods of the cone topology U, V ⊂ X(∞) of ξ′, ζ′′ (respectively) so that every point ζ′ ∈ V admits +dT (ξ′, ζ′) ≥ dT (ξ′, ζ′′)− π +4 = 3 +4π. Recall that the set of Γ-conical limit points is dense (in the cone topology), +so the second hypothesis implies there is indeed a Λ-conical limit point in V and therefore at Tits distance +different (in this case larger) than π +2 from ξ′. I conclude that Λ · x0 is not contained in any bounded metric +neighbourhood of any horosphere of X. +Assume towards contradiction that Λ ≤ StabG(ζ). +I show that this forces Λ ∩ Nζ ̸= ∅. By Propo- +sition 4.55 it is enough to find a unipotent element λ ∈ Λ and a geodesic η with η(∞) = ζ such that +limt→∞ d +� +λη(t), η(t) +� += 0. Let F be a maximal flat with ξ, ζ ∈ F(∞), x ∈ F some point and X, Y ∈ a ≤ p +two vectors such that exp(tY ) = η(t) for the unit speed geodesic η = [x, ζ), and exp(tX) = η′(t) for the unit +speed geodesic η′ = [x, ξ) (where a ≤ p a maximal abelian subalgebra in a suitable Cartan decomposition +g = p ⊕ k). Let StabG(ξ) = KξAξNξ be the decomposition described in Proposition 2.7 with respect to Y +(notice that Nξ does not depend on choice of Y , see item 3 of Proposition 2.17.7 in [20]). +The assumption that Λ ≤ StabG(ζ) implies that for any λ ∈ Λ the distance d +� +λη(t), η(t) +� +either tends +to 0 as t → ∞ or is uniformly bounded for t ∈ R. In the latter case there is some constant c > 0 for which +d +� +λη(t), η(t) +� += c for all t ∈ R. As in the proof of Corollary 4.41, the Flat Strip Theorem (Theorem 2.13, +Chapter 2.2 in [11]) implies that λ and at := exp(tY ) commute. +From the first hypothesis of the statement and Mostow’s result (Corollary 4.46) I know that Λ ∩ Nξ is a +cocompact lattice in Nξ (attention to subscripts). Therefore Λ ∩ Nξ is Zariski dense in Nξ (Theorem 4.53). +Moreover, since commuting with an element is an algebraic property, an element g ∈ G that commutes with +Λ∩Nξ must also commute with its Zariski closure, namely with Nξ. This means that if at commutes with all +Λ∩Nξ then it commutes with Nξ, i.e. atn = nat for all t ∈ R and all n ∈ Nξ. I know that at ∈ Aξ commutes +with both Kξ and Aξ (Proposition 2.7) therefore if at also commutes with Nξ then at lies in the centre +of StabG(ξ). This means that at is central in G (Lemma 4.52). For a group G with compact centre this +cannot happen, so there is indeed some unipotent element λ ∈ Λ ∩ Nξ for which limt→∞ d +� +λη(t), η(t) +� += 0. +I conclude from Proposition 4.55 that Λ ∩ Nζ ̸= ∅. +The first paragraph of the proof implies in particular that Λ·x0 does not lie in any bounded neighbourhood +of a horosphere H′ based at ζ. +The assumption that Λ ⊂ StabG(ζ) implies that every λ ∈ Λ acts by +translation on the filtration {H′ +t}t∈R by horospheres based at ζ. Therefore as soon as Λ · x0 ̸⊂ Ht for some +t ∈ R one concludes that ζ is a horospherical limit point of Λ, i.e. that every horoball based at ζ intersects +the orbit Λ · x0. +By Proposition 4.56 it holds that for a unipotent element g ∈ Nζ the displacement function x �→ d(gx, x) +depends only on the horosphere H′ +t in which x lies and that, for xt ∈ H′ +t it holds that limt→∞ d(gxt, xt) = 0 +(up to reorienting the filtration t ∈ R so that η(t) ∈ H′ +t). +For a non-trivial element λζ ∈ Λ ∩ Nζ the +previous paragraph therefore yields a sequence of elements λn ∈ Λ such that limn→∞ d(λζλnx0, λnx0) = 0, +contradicting the discreteness of Λ. I conclude that Λ ̸≤ StabG(ζ) for every ζ ∈ X(∞). +Assume that H := +� +Λ +Z�◦, the identity connected component of the Zariski closure of Λ, stabilizes a totally +geodesic submanifold Y ⊂ X. By Lemma 4.51, Λ0 := Λ ∩ H is of finite index in Λ, therefore Λ0 ∩ StabG(H) +is also a cocompact lattice in StabG(H). The fact that +� +Λ0 ∩ StabG(H) +� +· x0 is a cocompact metric lattice +in H readily implies that +� +Λ0 ∩ StabG(H) +� +· y is a cocompact metric lattice in Hy = H(y, ξ). This goes to +show that there is no loss of generality in assuming x0 ∈ H ∩ Y . It follows that Λ0 ∩ StabG(H) · x0 ⊂ Y ∩ H, +and therefore H ⊂ ND(Y ) for some D > 0. A horosphere is a codimension-1 submanifold, implying that Y +is either all of X or of codimension-1. The latter forces Y = H, which is impossible since H is not totally +geodesic (H is not convex, see Corollary 4.41). I conclude that H does not stabilize any totally geodesic +38 + +proper submanifold, and hence that Λ is Zariski dense. +4.5 +Proof of Theorem 4.1 +I now complete the proof of the main sublinear rigidity theorem for Q-rank 1 lattices. +Proof of Theorem 4.1. If {dγ}γ∈Γ is bounded, then Λ is a lattice by Corollary 4.7 or Theorem 4.9, depending +on the R-rank of G. +If {dγ}γ∈Γ is unbounded, then Proposition 4.12 and Corollary 4.27 both hold. In R-rank 1 the proof +again follows immediately from Corollary 4.7 using Lemma 2.16 and Corollary 4.27. +In higher rank, Section 4.4 allows one to conclude that Λ is an irreducible, discrete, Zariski dense subgroup +that contains a horospherical lattice. By Theorem 4.4, this renders Λ a lattice. It is a Q-rank 1 lattice as a +result of Theorem 2.20. +Remark 4.58. The sublinear nature of the hypothesis in Theorem 1.6 induces coarse metric constraints. +A horospherical lattice on the other hand is a very precise object. It is not clear how to produce unipotent +elements in Λ, or even general elements that preserve some horosphere. The proof sketched above produces +a whole lattice of unipotent elements in Λ (this is Corollary 4.46); it is also the only proof that I know which +produces even a single unipotent (or parabolic) element in Λ. +5 +Lattices with Property (T) +Recall that a lattice in a locally compact group G has property (T) if and only if G has property (T). In +this section I prove: +Theorem 5.1. Let G be a real centre-free semisimple Lie group without compact factors, Γ ≤ G a lattice, +Λ ≤ G a discrete subgroup such that Γ ⊂ Nu(Λ) for some sublinear function u. If Γ has property (T), then +Λ is a lattice. +As in the case of uniform lattices, lattices with property (T) admit the stronger version of ε-linear rigidity, +for suitable ε = ε(G): +Theorem 5.2. Let G be a real centre-free semisimple Lie group without compact factors, Γ ≤ G a lattice +and Λ ≤ G a discrete subgroup. If Γ has property (T) then there exists ε = ε(G) > 0 depending only on G +such that if Γ ⊂ Nu(Λ) for some function u(r) ⪯∞ εr, then Λ is a lattice. +Clearly Theorem 5.2 implies Theorem 5.1. I thank Emmanuel Breuillard for suggesting this generaliza- +tion. From now and until the end of this section, the standing assumptions are those of Theorem 5.2. +Lattice Criterion. +For groups with property (T) I use a criterion by Leuzinger, stating that being a +lattice is determined by the exponential growth rate. The formulation requires a definition. +Definition 5.3. Given a pointed metric space (X, dX, x0), denote: +1. bX(r) = |B(x0, r)| +2. bu +X(r) = supx∈X |B(x, r)| +When a group ∆ acts on a pointed metric space X, the orbit ∆ · x0 together with the metric induced +from X is a pointed metric space (∆ · x0, dX↾∆·x0, x0). In this setting b∆·x0(r) = |BX(xo, r) ∩ ∆ · x0|. When +the action is by isometries, i.e. ∆ ≤ Isom(X), it is straightforward to observe that this quantity does not +depend on the centre of the ball, and so bu +∆·x0(r) = b∆·x0(r). The pointed metric spaces of interest in this +section are the Γ and Λ orbits in the symmetric space X = G/K. +39 + +Definition 5.4. Let X be a symmetric space and ∆ ≤ G = Isom(X)◦ a subgroup of isometries. The critical +exponent of ∆ is defined to be +δ(∆) := lim sup +r→∞ +log +� +b∆·x0(r) +� +r +Remark 5.5. Throughout this section there is no risk of ambiguity, and I allow myself to ease notation and +let b∆(r) = b∆·x0(r), bu +∆(r) = bu +∆·x0(r). +To a semisimple Lie group G one can associate a quantity ∥ρ∥, where ρ = ρ(G) is the half sum of positive +roots in the root system of (g, a) (see Section 2 in [36]). +Theorem 5.6 (Theorem 2 in [36]). Let G be a real centre-free semisimple Lie group without compact factors. +Let ∆ be a discrete, torsion-free subgroup of G that is not a lattice. If G has Kazhdan’s property (T ), then +there is a constant c∗(G) (depending on G but not on ∆) such that δ(∆) ≤ 2∥ρ∥ − c∗(G). +It is known that the critical exponent of a discrete subgroup ∆ ≤ G is bounded above by 2∥ρ∥ (see +Section 2.2 in [36]). Moreover, every lattice Γ ≤ G admits δ(Γ) = 2∥ρ∥ (Example 2.3.5 in [36], Theorem C +in [1]). Combining these facts with Theorem 5.6 yield: +Theorem 5.7 (Theorem B in [36]). Let G be a real centre-free semisimple Lie group without compact +factors. Let ∆ be a discrete, torsion-free subgroup of G. If G has Kazhdan’s property (T ), then ∆ is a lattice +iff δ(∆) = 2∥ρ∥. +Line of Proof and the Use of ε-Linearity. +The proof of Theorem 5.1 goes by showing that ε-linear +distortion cannot decrease the exponential growth rate by much. This fact is essentially manifested in a +proposition by Cornulier [13], stated here in Proposition 5.8. This is the only use I make of ε-linearity, and +the computations involved are straightforward. Theorem 5.2 is then an immediate consequence of Leuzinger’s +criterion Theorem 5.6. +5.1 +Proof of Theorem 5.1 +In his study on SBE maps, Cornulier proves the following growth discrepancy result. +Proposition 5.8 (Proposition 3.6 in [13]). Let X, Y be two pointed metric spaces. Let u be a non-decreasing +sublinear function and p : X → Y a map such that for some L, R0 > 0: +1. |p(x)| ≤ max(|x|, R0), i.e. p(|x|) ≤ |x| for all large enough x ∈ X. +2. dY +� +p(x), p(x′) +� +≥ 1 +LdX(x, x′) − u(max{|x|, |x′|}) +Then for all r > R0, bY (r) ≥ bX(r)/bu +X +� +L · u(r) +� +. +I need a slightly modified version of Proposition 5.8: +Proposition 5.9. Let X, Y be two pointed metric spaces. Let u be a non-decreasing function that admits +u(r) ⪯∞ εr for some ε < 1, and p : X → Y a map such that for some L, R0 > 0: +1. |p(x)| ≤ max(|x| + u(|x|), R0), i.e. |p(x)| ≤ |x| + u(|x|) for all large enough x ∈ X. +2. dY +� +p(x), p(x′) +� +≥ 1 +LdX(x, x′) − u(max{|x|, |x′|}) +Then for all r > R0, +bY (r) ≥ bX +� +r − u(r) +� +/bu +X +� +L · u +� +r − u(r) +�� +Proof. Repeat verbatim the proof for Proposition 3.6. in [13]. +40 + +Corollary 5.10. Let u(r) ⪯∞ ε · r for some ε < 1 +2. Assume that Γ ⊂ Nu(Λ). Then δ(Λ) ≥ (1 − 4ε) · δ(Γ). +Remark 5.11. Restricting to ε < 1 +2 stems from a 2 factor that appears in the proof and could possibly be +dropped using a slightly more sophisticated approach. For my needs this is more than enough, since in any +case I eventually restrict attention to a small interval around 0. +Corollary 5.10 can be formulated in a slightly more general fashion. +Using the notation δ(W) := +lim supr→∞ +log +� +bW (r) +� +r +for a general subset W in a general metric space X, the following general version +holds: +Corollary 5.12. Let (X, x0) be a pointed metric space, Y, Z ⊂ X two subsets, and u(r) ⪯∞ εr for some +ε < 1 +2. Assume that bu +Z(r) = bZ(r). If Z ⊂ Nu(Y ), then δ(Y ) ≥ (1 − 4ε) · δ(Z). Moreover, if u is sublinear, +then δ(Z) = δ(Y ). +In particular, Corollary 5.10 holds even when the group G does not have property (T ). Corollary 5.10 +follows from Corollary 5.12 because the fact that Γ is a group of isometries implies bu +Γ(r) = bΓ(r). +Proof (Corollary 5.12). Observe that the closest point projection pY : Z → Y defined by z �→ zy for some +point in the closed ball zy ∈ B +� +z, u(|z|) +� +admits: +1. |yz| ≤ |z| + u(|z|) +2. d +� +yz, yz′� +≥ d(z, z′) − 2u(max{|z|, |z′|}) +The first item follows from triangle inequality: |yz| ≤ d(yz, z) + d(z, x0). The second item follows from +the quadrilateral inequality, i.e., using triangle inequality twice along the quadrilateral [z, z′, yz′, yz]. +The above properties allow me to use Proposition 5.9 with constant L = 1 and function u′ = 2u to get +bY (r) ≥ bZ +� +r − u′(r) +� +/bu +Z +� +u′� +r − u′(r) +�� +Since I assume bu +Z = bZ, I can omit the superscript u in the last expression. Recalling the definition +δ(W) = lim supr→∞ +bW (r) +r +, it remains to prove: +lim sup +r→∞ +1 +r · log +� +bZ +� +r − u′(r) +� +/bZ +� +u′� +r − u′(r) +��� +≥ (1 − 4ε) · δ(Z) +The proof of this inequality involves nothing more than log rules and arithmetic of limits: +lim sup +r→∞ +1 +r · log +� +bZ +� +r − u′(r) +� +/bZ +� +u′� +r − u′(r) +��� += lim sup +r→∞ +1 +r · +� +log +� +bZ +� +r − u′(r) +�� +− log +� +bZ +� +u′� +r − u′(r) +���� +≥ lim sup +r→∞ +� +1 +r · log +� +bZ +� +r − u′(r) +�� +− lim sup +s→∞ +�1 +s log +� +bZ +� +u′� +s − u′(s) +����� += lim sup +r→∞ +1 +r · log +� +bZ +� +r − u′(r) +�� +− lim sup +s→∞ +1 +s log +� +bZ +� +u′� +s − u′(s) +��� +≥ lim sup +r→∞ +1 +r · log +� +bZ +� +r − 2εr +�� +− lim sup +s→∞ +1 +s log +� +bZ +� +2εs +�� += (1 − 2ε)δ(Z) − 2εδ(Z) = (1 − 4ε)δ(Z) +(5) +Below I justify the steps in the above inequalities: +41 + +1. First equality is by rules of log. +2. Second and third inequalities are by arithmetic of limits: let (an)n, (bn)n be two sequences of positive +numbers, and A = lim supn an, B = lim supn bn. Then lim sup(an − bn) ≥ lim supn(an − B) = A − B. +3. Fourth inequality: u′(r) < 2ε(r) for all large enough r. +4. Fifth equality: definition of δ. +This completes the proof in the general case, which is what is needed for the proof of Theorem 5.2. +For the more refined statement in the case u is sublinear, one has to show a bit more. From inequality 5 +(specifically from the fourth line of the inequality) it is clearly enough to prove: +1. lim supr→∞ +1 +r · log +� +bZ +� +r − u′(r) +�� += δ(Z). +2. lim sups→∞ +1 +s log +� +bZ +� +u′� +s − u′(s) +��� += 0. +Starting from the second item, indeed it holds that +1 +s log +� +bZ +� +u′� +s − u′(s) +��� += u′� +s − u′(s) +� +s +· +log +� +bZ +� +u′� +s − u′(s) +��� +u′� +s − u′(s) +� +Clearly lim sup of the right factor in the above product is bounded by δ(Z), and in particular it is +uniformly bounded. On the other hand sublinearity of u′ implies that the left factor tends to 0. I conclude +that this product tends to 0 as s tends to ∞. +It remains to prove lim supr→∞ +1 +r · log +� +bZ +� +r − u′(r) +�� += δ(Z). In a similar fashion, +1 +r log +� +bZ +� +r − u′(r) +�� += r − u′(r) +r +· +log +� +bZ +� +r − u′(r) +�� +r − u′(r) +The left factor limits to 1 by sublinearity of u′. The right factor is nearly the expression in the definition +of δ(Z), and I want to prove that indeed taking lim sup of it equals δ(Z). A priori {r − u′(r)}r∈R>0 is just a +subset of R>0, so changing variable and writing t := r−u′(r) requires a justification. But there is no harm in +assuming that u′ is a non-decreasing continuous function, hence R≥R ⊂ {r − u′(r)}r∈R>0 for some R ∈ R>0. +Therefore for any sequence rn → ∞ there is a sequence r′ +n with rn = r′ +n − u′(r′ +n) for all large enough n (note +that in particular r′ +n → ∞). In the other direction, for every sequence r′ +n → ∞ there is clearly a sequence +rn → ∞ for which rn = r′ +n − u′(r′ +n). I conclude +lim sup +r→∞ +log +� +bZ +� +r − u′(r) +�� +r − u′(r) += lim sup +r→∞ +1 +r · log +� +bZ(r) +� += δ(Z) +This completes the proof. +Proof of Theorem 5.2. Define ε(G) = c∗(G) +4·2∥ρ∥, and assume u(r) ⪯∞ ε(G) · r. Notice that ε(G) < 1 +2, and since +δ(Γ) = 2∥ρ∥ Corollary 5.10 gives +δ(Λ) ≥ +� +1 − 4ε(G) +� +· 2∥ρ∥ = 2∥ρ∥ − 4ε(G) · 2∥ρ∥ ≥ 2∥ρ∥ − c∗(G) +By Theorem 5.6, Λ is a lattice. +42 + +Remark 5.13. The question of existence of interesting groups that coarsely cover a lattice is a key question +that arises naturally in the context of this paper. The first question that comes to mind is whether there +exist groups that are not commensurable to a lattice but that sublinearly, or even ε-linearly, cover one. +Perhaps the growth rate point of view could be used to rule out groups that cover a lattice ε-linearly but +not sublinearly. +6 +SBE Rigidity for Lattices of Higher Q-Rank +6.1 +Sublinear Distortion and SBE Maps +Denote a ∨ b := max(a, b) for a, b ∈ R. For a pointed metric space (X, x0, dX) and x, x1, x2 ∈ X, denote +|x|X := dX(x, x0) and |x1 −x2|X := dX(x1, x2) (or simply |x| and |x1 −x2| when there is no ambiguity about +the space X). +Following Cornulier [14], Pallier [44] makes the following definition: +Definition 6.1. A function u : R≥0 → R is admissible if it satisfies the following conditions: +• u is non-decreasing +• u grows sublinearly: lim sup +r�→∞ +u(r) +r += 0. +• u is doubling: +u(tr) +u(r) is bounded above for all t > 0. +• u ≥ 1. +The focus in this paper is condition 2, namely that the function u is strictly sublinear. I moreover require +it to be subadditive, resulting in the following terminology which I use from now on. +Definition 6.2. A function u : R≥0 → R is sublinear if it is admissible and subadditive, i.e. u(t + s) ≤ +u(t) + u(s) for all t, s > 0. +From now on by an SBE I mean an (L, u)-SBE where u is sublinear in the sense of Definition 6.2. +6.2 +SBE Rigidity +Two finitely generated groups Γ and Λ are said to be SBE if they are SBE when viewed as metric spaces with +some word metrics and base points eΓ, eΛ. Observe that every quasi-isometry is an SBE, and in particular +the word metric is an SBE-invariant. An SBE admits an SBE inverse, defined as quasi-inverse maps are +defined for quasi-isometries. +Definition 6.3. A class of groups A is said to be SBE complete if, for every finitely generated group Λ that +is SBE with some group Γ ∈ A, there is a short exact sequence +1 → F → Λ → Λ1 → 1 +for a finite group F ≤ Λ and some Λ1 ∈ A. +In this chapter I prove: +Theorem 6.4. Let G be a real centre-free semisimple Lie group without compact or R-rank 1 factors. +1. The class of uniform lattices of G is SBE complete. +2. The class of non-uniform lattices of G is SBE complete. +43 + +Remark 6.5. The proof I present is quite indifferent to whether the lattice Γ is uniform or not. In order +to have a unified proof and to ease notation, I fix the convention that for both uniform and non-uniform +lattices, X0 denotes the compact core of the lattice. This just means that X = X0 in case Γ is uniform. +My works heavily relies on that of Drut¸u in [16], where the theorems are stated for non-uniform lattices. +Nonetheless one readily sees that her proofs work perfectly well for uniform lattices. Indeed the arguments +of [16] are only much simpler in the uniform case. +6.2.1 +The Quasi-Isometry Case +The outline of the proof I present for Theorem 6.4 is identical to that of quasi-isometric rigidity, which I +now describe briefly. The main step is that for any quasi-isometry f : X0 → X0 of the compact core of Γ, +there exists an isometry g : X → X such that f, g are boundedly close, i.e. there is some D > 0 for which +d +� +f(x), g(x) +� +< D for all x ∈ X0. +Let Λ be an abstract group with a quasi-isometry q : Λ → Γ. +Using Lubotzki-Mozes-Raghunathan +([38], see Theorem 2.19 above), Γ is quasi-isometrically embedded in X as the compact core X0. One can +thus extend q to a quasi-isometry q0 : Λ → X0. A conjugation trick allows to associate to each λ ∈ Λ a +quasi-isometry fλ : X0 → X0 defined by fλ := q ◦ Lλ ◦ q−1 (Lλ : Λ → Λ is the left multiplication by λ in Λ). +By the first paragraph, there exists gλ ∈ Isom(X) that is boundedly close to fλ. Moreover, the proof also +shows that the bound D depends only on the quasi-isometry constants of fλ. These could be seen to depend +only on q and not on any specific λ. From this one concludes that the map λ �→ gλ is a group homomorphism +Φ : Λ → G. It is then straightforward to show that Φ has finite kernel and that Γ ⊂ ND +� +Im(Φ) +� +. One then +uses Theorem 4.3.2 (for higher rank groups, see Section 4.3) to deduce that Im(Φ) is a non-uniform lattice +in G that is commensurable to Γ. +6.2.2 +The SBE Case +Moving to SBE rigidity, one starts with an SBE q : Λ → Γ. The first step is to find an isometry of X that is +close to an SBE of X0. Drut¸u’s proof is preformed in the asymptotic cone of X, which allows for a smooth +transition to the SBE setting. Indeed, given an SBE f : X0 → X0, one can find an isometry g : X → X that +is close to it. The difference is that in the SBE setting, these maps are only sublinearly close. +Definition 6.6. Let (X, x0) be a pointed metric space, and f, g : X → X be two maps. Maps f, g are said +to be sublinearly close maps on X if there is a sublinear function u such that d +� +f(x), g(x) +� +≤ u(|x|). +Theorem 6.7 (Theorem 6.10). Let f : X0 → X0 be an SBE. Then there exists a unique isometry g ∈ +Isom(X) that is sublinearly close to f (in X0). +From this point on, one would like to continue as in the quasi-isometry case: define the map Φ : Λ → G +in a similar fashion and show that Γ ⊂ Nu +� +Im(Φ) +� +for some sublinear function u. That Im(Φ) is a lattice is +then a result of Theorem 1.6, proving Theorem 6.4. +There is however one additional obstacle that is unique to the SBE setting. Namely the SBE constants +of fλ do depend on λ, and the resulting sublinear bound on d +� +fλ(x), gλ(x) +� +in Theorem 6.7 is not enough +to define Φ properly. As far as I can see, one needs to get some uniform control on that bound in terms of +the SBE constants. The following statement is enough: +Lemma 6.8 (Lemma 6.11). Let {fr}r∈R>0 be a family of (L′, vr)-SBE maps fr : X0 → X0, where vr(s) = +L′·v(s)+v(r) for some sublinear function v ∈ O(u) and a constant L′. Let gr be the associated isometry given +by Theorem 6.10. Then for any x ∈ X0, there is a sublinear function ux ∈ O(u) such that d +� +fr(x), gr(x) +� +≤ +ux(r). +This type of uniform control is often needed when working with SBE maps, see e.g. Section I.3 in Pallier’s +thesis [43]. Using it, I am able to complete the argument as in the quasi-isometry case and prove: +Theorem 6.9. Let G be as in Theorem 6.4. In the notations described above, the map Φ : Λ → G is a group +homomorphism with Ker(Φ) finite, and there is a sublinear function u such that for Λ1 := Im(Φ) it holds +that Γ ⊂ Nu(Λ1) and Λ1 ⊂ Nu(Γ). +44 + +Theorem 6.4 is an immediate corollary of Theorem 6.9 and Theorem 1.6. +6.2.3 +Outline +This section is divided into two parts that correspond to the steps of the proof. Section 6.3 deals with the +task of finding an isometry that is sublinearly close to an SBE, and Section 6.4 establishes the properties of +the map Φ : Λ → G. I keep Section 6.3 slim and concise. The main reason for this choice is that the proof +of Theorem 6.10 is merely a mimic of Drut¸u’s argument in [16], or an adaptation of it to the SBE setting. +While these adaptations are somewhat delicate, giving a complete detailed proof would require reproducing +Drut¸u’s argument more or less in full. I felt that this is not desirable, and instead I only indicate the required +adaptations. I believe that a reader who is familiar with Drut¸u’s argument and with asymptotic cones could +easily produce a complete proof using these indications. In particular, there is no preliminary section. I do +not present buildings or dynamical results that go into Drut¸u’s argument. I only shortly present asymptotic +cones and some ideas from Drut¸u’s proof of the quasi-isometry version of Theorem 6.10. Section 6.4 is +elementary. +6.3 +SBE Maps are Close to Isometries +In this section I indicate how to adapt Drut¸u’s arguments in [16] in order to prove: +Theorem 6.10. There is a sublinear function v = v(L, u) such that for every +� +L, u +� +-SBE f : X0 → X0, +there exists a unique isometry g = g(f) ∈ Isom(X) such that d +� +f(x), g(x) +� +≤ v(|x|). +The proof of Theorem 6.4 requires some control on the sublinear distance between f and g, in terms of +the sublinear constants of f. This is the meaning of the following lemma. +Lemma 6.11. Let {fr}r∈R>0 be a family of (L′, vr)-SBE maps fr : X0 → X0, where vr = L′ · v + v(r) for +some sublinear function v ∈ O(u) and a constant L′. Let gr be the associated isometry given by Theorem 6.10. +Then for any x ∈ X0, there is a sublinear function ux ∈ O(u) such that d +� +fr(x), gr(x) +� +≤ ux(r). +Remark 6.12. Combined with Theorem 6.10, a different way to phrase the above statement is to say that +the function D : Λ× X0 → R≥0 defined by D(λ, x) = d +� +fλ(x), gλ(x) +� +is sublinear in each variable. I.e., there +is a function u : R≥0 × R≥0 → R≥1 such that u is sublinear in each variable and D(λ, x) ≤ u(|λ|, |x|). +Outline. +I begin with a short presentation of asymptotic cones. I then give an account of the original +proof of Theorem 6.10 when f : X0 → X0 is a quasi-isometry. I present the routines required to modify the +proof for the SBE setting. I exemplify the modification procedure in a specific representative example, and +finish with a road map for proving Theorem 6.10 and Lemma 6.11 using the aforementioned routines. +6.3.1 +Asymptotic Cones +Definition 6.13. Let (X, d) be a metric space. Fix an ultrafilter ω, a sequence of points xn ∈ X and a +sequence of scaling factors ın −→ +ω 0. The asymptotic cone of X w.r.t. xn, ın, denoted C(X), is the metric +ω-ultralimit of the sequence of pointed metric spaces (X, 1 +ın · d, xn). The metric on C(X) is denoted dω. +See Section 2.4 in [16] for an elaborate account, including the definitions of ultrafilters and ultralimits. +The strength of SBE maps is that they induce bi-Lipschitz maps between the respective asymptotic cones. +Lemma 6.14 (See e.g. Cornulier [13]). Let f : X → Y be an (L, u)-SBE. Then f induces an L-bi-Lipschitz +map C(f) : C(X) → C(Y ) between the corresponding asymptotic cones with the same scaling factors C(X) = +(X, 1 +ın dX, x0 +n) and C(Y ) = (Y, 1 +ın dY , y0 +n). +45 + +6.3.2 +The Argument +A High-Level Description. +The core of the argument lies in elevating an SBE f0 : X0 → X0 to an +isometry g0 ∈ G = Isom(X). There are two gaps to fill: first, Γ is non-uniform and so f0 is not even defined +on the whole space X. And obviously, f0 is just an SBE. +Assume for a moment that Γ is uniform and that f is defined on the whole space X. Elevating f : X → X +to an isometry is done by considering the map C(f) : C(X) → C(X) that f induces on an asymptotic cone +C(X). This map is bi-Lipschitz, and the work of Kleiner and Leeb [32] allows one to conclude that C(f) +is, up to a scalar, an isometry. In turn, this isometry induces an isometry ∂g on the spherical building +structure of X(∞). This is done by the relation between the Euclidean building structure of C(X) and the +spherical building structure of ∂∞X. A theorem of Tits [56] associates to ∂q a unique isometry g ∈ Isom(X) +that induces ∂f as its boundary map. By construction, it is then not too difficult to see that g and f are +‘close’. In case Γ is non-uniform, an SBE f : Γ → Γ does not readily yield a cone map on C(X), but only on +C(X0). Overcoming this difficulty requires substantial work and is the heart of Drut¸u’s proof. In short, she +uses dynamical results stating that the vast majority of flats in X are close enough to X0. As mentioned +in Section 2.2, the building structure on X(∞) is determined by the boundaries of flats F ⊂ X. The same +holds for the (Euclidean) building structure of C(X). Therefore the fact that the majority of flats in X are +‘close enough’ to X0 results in the fact that C(X0) composes the majority of C(X). This is a very rough +sketch of the logic behind Drut¸u’s argument. +The procedure described above results in an isometry g ∈ Isom(X) associated to f0. To complete the +argument one needs to verify that the map f0 �→ g is a group homomorphism between SBE(Γ) = SBE(X0) +and Isom(X). +Composing this map with a representation of Λ into SBE(Γ) yields a map Λ → G by +λ �→ fλ �→ gλ := gfλ. A computation then shows that this map has finite kernel and that Γ lies in a sublinear +neighbourhood of the image. +Flat Rigidity. +The adaptations that are required for the SBE setting lie mainly in the part of Drut¸u’s +work that concerns flat rigidity. That is, the proof that the quasi-isometry q0 maps a flat F ⊂ X0 to within +a uniformly bounded neighbourhood of another flat F ′ ⊂ X0. This is proved by passing to the cone map, +using an analogous result for bi-Lipschitz maps between Euclidean buildings, which translates back down to +the space X0. +Remark 6.15. Drut¸u’s argument requires many geometric and combinatorial definitions - some classical +and widely known (e.g., Weyl chambers of a symmetric space X) some less known (e.g. an asymptotic cone +with respect to an ultrafilter ω) and some new (e.g., the horizon of a set A ⊂ X). I use her definitions, +terminology and notations freely without giving the proper preliminaries or even the definitions. I assume +most readers who are interested in the question of SBE rigidity are familiar to some extent with most of these +objects. For the new definitions, I try to say as little as needed to allow the reader to follow the argument. +The proof consists of 6 steps: +1. The horizon of an image of a Weyl chamber is contained in the horizon of a finite union of Weyl +chambers, and the number of chambers in this union depend only on the Lipschitz constant. (Lemmas +3.3.5, 3.3.6 in [16], consult Remark 6.16 below for a sketchy definition of horizon). +2. The horizon of an image of a flat coincides with the horizon of a finite union of Weyl chambers, and +the number of chambers depends only on the Lipschitz constant of the quasi-isometry. +3. The union of Weyl chambers in the previous step limits to an apartment in the Tits building at X(∞). +Such a union is called a fan over an apartment. +4. For each Weyl chamber W there corresponds a unique chamber W ′ such that q0(W) and W ′ have the +same horizon. This amounts to an induced map on the Weyl chambers of the Tits building at X(∞) +(Lemma 4.2.1 in [16]). +46 + +5. Given a flat F through a point x, the unique flat F ′ asymptotic to the union of Weyl chambers obtained +in step 3 is at uniform bounded distance from f0(x). The bound depends only on the quasi-isometry +constants. The flat F ′ is called the flat associated to F. +6. If F1 and F2 are two flats through x which intersect along a hyperplane H, then the boundaries at +X(∞) of the associated flats F ′ +1 and F ′ +2 intersect along a hyperplane of the same codimension as H. +Remark 6.16. For a precise definition of horizon see [16], section 3. For now, it suffices to say the following. +The horizon of a set A ⊂ X is contained in the horizon of a set B ⊂ X if, looking far away at A from some +point x ∈ X, A appears to be contained in an ε-neighbourhood of B. This intuition is made precise by +considering the angle at x that a point a ∈ A makes with the set B. Two sets have the same horizon if each +set’s horizon is contained in the other. In the case A and B have the same horizon, an important aspect is +the distance R starting from which A and B seem to be ε-contained in one another. Call this distance the +horizon radius. It depends on x and ε. +The proofs for most of these steps have similar flavour: in any asymptotic cone C(X), f0 induces a +bi-Lipschitz map. Kleiner and Leeb [32] proved many results about such maps between cones of higher rank +symmetric spaces. One assumes towards contradiction that some assertion fails (say, in step 5, assume that +there is no bound on the distance between f0(xn) and the associated flat F ′ +n). This gives an unbounded +sequence of scalars (say, d +� +f0(xn), F ′ +n +� += ın → ∞). These scalars are used to define a cone in which one +obtains a contradiction to some fact about bi-Lipschitz maps (say, that the point [q0(xn)]ω is at dω-distance +1 from [F ′ +n]ω, while it should lie in [F ′ +n]ω). +Typically, the bounds obtained this way depend on the quasi-isometry constants. A priori, they also +depend on the specific point x ∈ X or flat F ⊂ X in which you work (e.g. the horizon radius for the chambers +in step 3 or the bound on d +� +q(x), F ′� +in step 5). However, it is easy to see that in the quasi-isometry setting, +the bounds are actually independent of the choice of point/flat/chamber. This independence stems from +the fact that one can pre-compose f0 with an isometry translating any given point/flat/chamber to a fixed +point/flat/chamber (resp.), without changing the quasi-isometry constants (see e.g. Remark 3.3.11 in [16]). +The fact that these bounds depend only on the quasi-isometry constants is essential for the proof that the +map Λ → Isom(X) has the desired properties. +Moving to the SBE setting, the essential difference is exactly that the bounds one obtains depend on +the specific point, Weyl chamber or flat. Indeed it is clear that these bounds should depend on the size |x|, +as they depend on the additive constant in the quasi-isometry case. It is sensible to guess though that the +bounds only grow sublinearly in |x|, which is enough in order to push the argument forward. In the next +section I show how to elevate a typical cone argument from the quasi-isometry setting to the SBE setting. I +focus on showing that the bound one obtains depend only on the SBE constants (L, u) and sublinearly |x|. +6.3.3 +Generalization to SBE: Adapting Cone Arguments +To adapt for the SBE setting, split each step into three sub-steps: the first two amount to proving Theo- +rem 6.10, and the third step amounts to proving Lemma 6.11. +Sub-Step 1. +Repeat the argument of the quasi-isometry setting verbatim, to obtain a bound c = c(x) +which depend on the point x. +Sub-Step 2. +Assume towards contradiction that there is a sequence of points xn for which limω +c(xn) +|xn| ̸= 0. +This means limω +|xn| +c(xn) ̸= ∞, and so the point (xn)n lies in the cone C(X) = Cone +� +X, x0, c(xn) +� +, and one +may proceed as in the corresponding quasi-isometry setting to obtain a contradiction. +Sub-step 3. +Fix x ∈ X and a sequence of SBE maps as in Lemma 6.11, i.e. {fn}n∈N with the same Lipschitz +constant and with sublinear constants vn(s) = v(s) + u(n), for some sublinear functions u, v. Denote by +cn(x) the constants that were achieved in the previous steps for x and the SBE map fn, and assume towards +47 + +contradiction that |cn(x)| is not bounded above by any function sublinear in n. This means in particular +that limω +u(n) +cn(x) = 0. One concludes that the cone map C(frn) is bi-Lipschitz, and gets a contradiction in the +same manner as in the first step. +Example 6.17. In order to give the reader a sense of what is actually required, I now demonstrate this +procedure in full in a specific claim. I chose to do this for proposition 4.2.7 of [16], which is complicated +enough to require some attention to details, but not too much. The statement is as follows: +Proposition 6.18 (SBE version of Proposition 4.2.7 in [16]). Let f : X → X be an (L, u)-SBE, and F ⊂ X +a flat through x to which f associates a fan over an apartment, ∪p +i=0Wi. If F ′ is the maximal flat asymptotic +to the fan, then d +� +f(x), F ′� +≤ c(x) where c(x) = c(|x|) is sublinear. +Moreover, let fn : X → X be a sequence of (L, vn) SBE maps for vn = v + u′(n) for v, u′ some sublinear +functions. The constant cn(x) associated to x and the SBE fn achieved in the first part of the proposition +admits cn(x) ≤ ux(n) for some sublinear function ux. +F ′ is then said to be the associated flat to F by f. +Proof. Proceed in two (sub-)steps. +Step 1. +I show that for a given x, there exists such a constant c(x) independent of the flat F. This is done +exactly as in [16], but I repeat the proof here because it contains the terminology and necessary preparation +for the second step. +Fix x ∈ X and assume towards contradiction that there exists a sequence Fn of flats through x and a +sequence fn : X → X of (L, u)-SBE maps such that cn := d +� +fn(x), F ′ +n +� +→ ∞. In Cone(X, x, c−1 +n ) one can +show that [∪p +i=0W n +i ], the union of Weyl chambers associated to fn(Fn), is a maximal flat (see Proposition +4.2.6 in [16]). Denote Fω := [∪p +i=0W n +i ]. Furthermore, since the bi-Lipschitz flat [fn(Fn)] ⊂ Cone(X, x, c−1 +n ) +is contained in it, it coincides with it: [fn(Fn)] = Fω. On the other hand, since the Hausdorff distance +between ∪p +i=0W n +i and F ′ +n is by assumption cn = d(x, F ′ +n), in the cone the maximal flats Fω and F ′ +ω := [F ′ +n] +are at Hausdorff distance 1. +This implies that Fω = F ′ +ω (see Corollary 4.6.4 in [32]). But since d +� +q(x), F ′ +n +� += cn the limit point +yω := Q(xω) = [q(x)] , which is contained in Fω, is at distance 1 from F ′ +ω - a contradiction. +Step 2. +Assume c(x) is taken to be the smallest possible for each x, and then modify the function c +so that c(x) = maxy:|y|=|x| c(y). +The function c : X → R now only depends on |x|. +I wish to show +that c(|x|) = O(u(|x|). +Assume towards contradiction that there exists a sequence xn with |xn| → ∞ +such that limω +c(xn) +u(|xn|) = ∞. Denote cn = c(xn) and consider the cone Cone(X, xn, c−1 +n ). The assumption +limω +c(xn) +u(|xn|) = ∞ implies (x0)ω = (xn)ω hence Cone(X, xn, cn) = Cone(X, x0, cn). By the definition of c(x) +this means that there is a sequence of flats Fn through xn such that d +� +q(xn), F ′ +n +� += cn, so one may proceed +as in step 1 for a cone with a fixed base point (x0)ω. In this cone the flat F ′ +ω := [F ′ +n] is at distance 1 +from the point [q(xn)], which lies on the maximal flat [∪p +i=0W n +i ]. The latter flat is, on the one hand, at +Hausdorff distance 1 from F ′ +ω (by the definition of the scaling factors cn), so they actually coincide. On +the other hand, [∪p +i=0W n +i ] coincides with Fω := Q(Fω) = [q(Fn)], so Fω = F ′ +ω, contradicting the fact that +dω([q(xn)], F ′ +ω) = 1. Thus c(|x|) = O(u(|x|)), as wanted. +Step 3. +For the moreover part (uniform control on the growth of c(x) as a function of the sublinear +constants), the proof is identical to Step 1. This time, consider a sequence fn as in the statement, and +denote by cn = cn(x) the constant obtained in step 1 w.r.t. the SBE constants (L, vn). Assume towards +contradiction that limω +u(n) +cn(x) = 0. The proof goes exactly as in step 1, with the sole difference that now one +might need convincing in the fact that in C(X), the cone with +1 +cn as scaling factors, the cone map C(fn) is +bi-Lipschitz. But indeed for any two cone points (xn), (yn) ∈ C(X) it holds: +dω +� +C(fn)(xn), C(fn)(yn) +� += lim +ω +1 +cn +d +� +fn(xn), fn(yn) +� +≤ lim +ω +1 +cn +L · d(xn, yn) + vn(|xn| ∨ |yn|) +48 + +By definition of the cone metric, limω +1 +cn L · d(xn, yn) = L · dω +� +(xn), (yn) +� +. +It thus remains to show +limω +1 +cn vn(|xn| ∨ |yn|) = 0. By definition of vn it amounts to proving limω +1 +cn v(|xn|) = 0 = limω +1 +cn v(|yn|) +and limω +1 +cn u(n) = 0. The former follows from the fact that (xn), (yn) ∈ C(X) and therefore both limω +1 +cn |xn| +and limω +1 +cn |xn| are finite. The latter follows from the assumption on the cn. One obtains a contradiction +identical to the one in Step 1. +Following the claims of Sections 3, 4, 5 in [16] carefully, and making the SBE adaptations as depicted +in the above example, one obtains flat rigidity in the SBE setting, that is Theorem 6.10 together with the +uniform control described in Lemma 6.11. Here is the complete list of claims involving cone arguments in +[16] that should be modified. +• Section 3. All claims starting from Lemma 3.3.5 through Corollary 3.3.10. All statements should +consider, instead of a quasi-isometry, a general (L, u)-SBE f and, when relevant, a general point x ∈ X +with f(x) = y (i.e., f(x) does not necessarily equal x). Also when relevant one should consider a family +fn of (L, vn) SBE maps as depicted in Lemma 6.11 above. +• Section 4. Propositions 4.2.6, 4.2.7, 4.2.9. When relevant, statements should be modified so that the +distance between f(x) and an associated flat of it should be uniformly sublinear in |x|. Also when +relevant one should consider a family fn of (L, vn) SBE maps as above. +• Section 5. Lemma 5.4.1 (D = D(x) should be uniformly linear in c = c(x)). Proposition 5.4.2 (the +constant D = D(x) should be replaced be a sublinear function D(|x|)). When relevant one should +consider a family fn of (L, vn) SBE maps as above. +This yields a proof for uniform flat rigidity in the SBE setting. The other major part of Drut¸u’s argument +concerns the fact that f0 is defined only on X0 and not on X. +The considerations for this aspect are +intertwined in the proof, but they all involve only Γ and the quasi-isometry between Γ and X0. For this +reason, the fact that f0 is an SBE to begin with does not effect any of these arguments. Moreover, this +argument is indifferent to whether or not Γ is uniform or not. If Γ is uniform all that changes is that that +part of Drut¸u’s argument dealing with extending the cone map from C(X0) to C(X) is not necessary since +X = X0. Her proof still works perfectly well also for uniform lattices. Therefore the argument above proves +Theorem 6.10 and Lemma 6.11. +6.3.4 +Some Remarks On R-rank 1 Factors +Quasi-isometric rigidity holds for groups of R-rank 1. It is worth mentioning that Schwartz’s proof also +relies on ‘flat rigidity’ - but in this case the flats are the horospheres of Γ. While these are not isometrically +embedded flats, the induced metric on horospheres is flat and Schwartz uses that in order to construct the +boundary map and find the associated isometry. +The same phenomena occurs in the SBE setting. Considering the compact core X0 ⊂ X of Γ, one can +use Proposition 5.6 in Drut¸u-Sapir [18], in order to show that horospheres are mapped boundedly close to +horospheres. In general, this work characterizes and explores a certain class of spaces they call asymptotically +tree graded, a class that is very suitable for the setting of the compact core of a non-uniform lattice in R- +rank 1. +A key ingredient in the proof is the fact that the boundary map ∂q induced by the quasi-isometry is +quasi-conformal. This in particular implies that it is almost everywhere differentiable. I spent some time +trying to generalize the proof of Schwartz to the SBE setting. One obstacle is that it is not clear that the +boundary map is going to be differentiable almost everywhere. Gabriel Pallier found that there are SBE +maps of the hyperbolic space whose boundary maps are not quasi-conformal (see Appendix A in [45]). For +this reason Pallier develops the notion of quasi-conformality [46]. While these maps may be differentiable, +he told me of examples he constructed where the differential is almost everywhere 0 - a property which also +nullifies Schwartz’s argument. +In the context of SBE rigidity, the maps I consider seem to indeed have ‘flat rigidity’, i.e. to map a +horosphere to within bounded distance of a unique horosphere. As in the higher rank flat rigidity, this +49 + +bound is not uniform but rather grows sublinearly with the distance of the horosphere to a fixed base point. +These are very specific maps, that coarsely preserve the compact core of that lattice X0 ⊂ X and basically +map horospheres to horospheres. This means there might still be hope for these specific maps to induce +boundary maps that admit the required analytic properties. +In a subsequent paper [53], Schwartz proves quasi-isometric rigidity for lattices in products of R-rank 1 +groups, i.e. in Hilbert modular groups. His proof there is different, but it also makes use of the fact that +horospheres are mapped to within uniformly bounded distance of horospheres. The fact that in the SBE +setting this bound is not uniform seems like a real obstruction to any attempt of generalizing his proof in +that case. +6.4 +From SBE to Sublinearly Close Groups +In this section I prove Theorem 6.9. I restate it here for convenience +Theorem 6.19. Let G be a real centre-free semisimple Lie group without compact or R-rank 1 factors. Let +Γ ≤ G be an irreducible lattice, and Λ an abstract finitely generated group that is SBE to Γ. Then there is +a group homomorphism Φ : Λ → G with finite kernel such that Γ ⊂ Nu +� +Φ(Λ) +� +and Φ(Λ) ⊂ Nu(Γ) for a +sublinear function u. +The proof is a sublinear adaptation of the classical arguments by Schwartz. The only difference is that +some calculations are in order, but there is no essential difference from Section 10.4 in [52]. +Before I start, I need one well known preliminary fact, namely that sublinearly close isometries are equal. +Lemma 6.20. Let X be a symmetric space of noncompact type and with no R-rank 1 factors. Let Γ ≤ +Isom(X) be a non-uniform lattice, X0 its compact core with respect to x0 ∈ X. Let g, h ∈ G = Isom(X) and +u a sublinear function such that for every x ∈ X0, d +� +g(x), h(x) +� +≤ u(|x|). Then g = h. +Proof. The proof is essentially just the fact that a sublinearly bounded convex function is uniformly bounded. +Up to multiplying by h−1, one may assume h = idX. First I show that the continuous map ∂g : X(∞) → +X(∞) is the identity map. Recall that the space X(∞) can be represented by all geodesics emanating from +the fixed point x0. Let η : [x0, ξ) be a Γ-periodic geodesic. By definition, there is some T > 0 and a sequence +γn ∈ G for which η(nT ) = γnx0. In particular, xn := γnx0 ∈ X0 hence d +� +g(xn), xn +� +≤ u(|xn|) = u(nT ). +On the other hand, the distance function d +� +η(t), g·η(t) +� +is convex. A convex sublinear function is bounded, +and so by definition in X(∞) one has [η] = [g · η], for all Γ-periodic geodesics η. The manifold X is of non- +positive curvature, hence ∂g is a homeomorphism of X(∞), and the density of Γ-periodic geodesics implies +∂g = idX(∞). This implies that g = idX (see Section 3.10 in [20] for a proof of this last implication). +Remark 6.21. The proof for the fact that ∂g = idX(∞) implies g = idX appears in [20] (section 3.10) as +part of the proof of the following important theorem of Tits: +Theorem 6.22 ([56], see Theorem 3.10.1 in [20]). Let X, X′ be symmetric spaces of noncompact type and +of higher R-rank. Assume X has no R-rank 1 factors, and let φ : X(∞) → X′(∞) be a bijection that is a +homeomorphism with respect to the cone topology and an isometry with respect to the Tits metric. Then, +after multiplying the metric of X by positive constants on de Rham factors, there exists a unique isometry +g : X → X′ such that φ = ∂g. +This theorem is actually a key ingredient in Drut¸u’s argument. Much of her work is directed towards +showing that the cone map C(q) will correspond to a map on X(∞) satisfying the above hypothesis. The +restriction to X with no R-rank 1 factors in Theorem 6.4 comes from this restriction in Tits’ Theorem 6.22 +Remark 6.23. The proof that ∂g = idX(∞) ⇒ g = idX only uses the fact that X has no Euclidean de +Rham factors (see pg. 251 in [20]). Since I only use it in the setting of no R-rank 1 factors, I added that +assumption to Lemma 6.20. +50 + +The Map Φ : Λ → G. +The orbit map q0 : Γ → X0 defined by γ �→ γx0 is a quasi-isometric embedding: +this is ˘Svarc-Milnor in case Γ is uniform and X0 = X, and Lubotzki-Mozes-Raghunathan (Theorem 2.19 +above) if Γ is non-uniform. An SBE f : Λ → Γ thus gives rise to an SBE Λ → X0, which I also denote by f. +For each λ ∈ Λ let fλ := f ◦ Lλ ◦ f −1 : X0 → X0, where Lλ is the left multiplication by λ. The left +translation Lλ is an isometry, hence fλ is a self SBE of X0. By Theorem 6.10, there exists a unique isometry +gλ ∈ Isom(X) that is sublinearly close to fλ. Define the map Φ : Λ → G by λ �→ gλ. The goal in this section +is to prove Φ is a homomorphism with finite kernel, and that Γ and Φ(Λ) are each contained in a sublinear +neighbourhood of the other. +I begin by controlling the SBE constants of the fλ. +Lemma 6.24. For each λ ∈ Λ, fλ is an (L2, vλ)-SBE, for +vλ(|x|) := (L + 1)u(|x|) + u(|λ|) +In particular vλ ∈ O(u). +Before the proof I state a corollary which follows immediately by combining Lemma 6.24 with Lemma 6.11. +Corollary 6.25. Assume Lemma 6.11 holds. Then for any x ∈ X there is a sublinear function ux such that +d +� +fλ(x), gλ(x) +� +≤ ux(|λ|) +. +Proof of Lemma 6.24. The proof is a straightforward computation. Up to an additive constant I may assume +f −1 is an (L, u)-SBE with f −1(eΓ) = eΛ. Let x1, x2 ∈ X0, and assume w.l.o.g |x2| ≤ |x1|. By the properties +of an SBE, this also means that for i ∈ {1, 2}: +|f −1(xi)| ≤ L|xi − x0| + u(|xi|) ≤ L|x1| + u(|x1|) +(6) +Notice that fλ(x) = f +� +λ · f −1(x) +� +, and f is an (L, u)-SBE. The following inequalities, justified below, +give the required upper bounds: +��fλ(x1) − fλ(x2) +�� ≤ L · +��λf −1(x1) − λf −1(x2) +�� + u +� +|λf −1(x1)| ∨ |λf −1(x2)| +� +≤ L2|x1 − x2| + Lu +� +|x1| +� ++ u +� +|λ|) + L|x1| + u(|x1|) +� +≤ L2|x1 − x2| + (L + 1)u +� +|x1| +� ++ u(|λ|). +(7) +From the first line to the second line I used: +1. For the first term: left multiplication in Λ is an isometry, and f −1 is an (L, u)-SBE. +2. For the second term: triangle inequality, left multiplication in Λ is an isometry, and Equation 6. +From the second to the third line I used the properties of u as an admissible function, namely that it is +sub-additive and doubling, so u +� +(L + 1)|x1| +� +≤ (L + 1)u(|x1|) for all large enough x1. +Remark 6.26. I remark that the proof of Lemma 6.24 is the only place where I use the properties of an +admissible function and not just the sublinearity of u. +Claim. Φ : Λ → G is a group homomorphism. +Proof. Let λ1, λ2 ∈ Λ. I begin with some notations: +1. f1 = fλ1, f2 = fλ2, f12 = fλ1λ2. By Lemma 6.24, these are all O(u) SBE maps with the same Lipschitz +constant L′ := L2 and sublinear constants v1, v2, v12 ∈ O(u). +51 + +2. g1 = Φ(λ1), g2 = Φ(λ2), g12 = Φ(λ1λ2) +3. u1, u2, u12 the sublinear functions that bound the respective distances between any g and f, e.g. +|g1(x) − f1(x)| ≤ u1(|x|). +One has to prove that gλ2 ◦ gλ1 = gλ1λ2. In view of Lemma 6.20, it is enough to find a sublinear function +v such that for all x ∈ X0 |g1g2(x) − g12(x)| ≤ v(|x|). By triangle inequality and the above definitions and +notation, it is enough to show that each of the following four terms are bounded by a function sublinear in +x: +1. |g1g2(x) − g1f2(x)| = |g2(x) − f2(x)| ≤ u2(|x|) (g1 is an isometry). +2. |g1f2(x) − f1f2(x)| ≤ u1 +� +|f2(x)| +� +≤ u1 +� +L2|x| + v2(|x|) +� +3. |f1f2(x) − f12(x)| +4. |f12(x) − g12(x)| ≤ u12(|x|) +Clearly items 1, 2, 4 are bounded by a sublinear function in |x|. It remains to bound |f1f2(x) − f12(x)|. +The map Λ → Aut(Λ) given by λ �→ Lλ is a group homomorphism, i.e. Lλ1λ2 = Lλ1Lλ2, so it remains to +bound: +|f1f2(x) − f12(x)| = |fLλ1f −1fLλ2f −1(x) − fLλ1Lλ2f −1(x)| +f ◦ Lλ is a composition of an isometry with an SBE, so it is still an SBE. Denote the SBE constants of +fLλ1 by L′, v (clearly one can take L′ = L and v ∈ O(u), but this is not needed). Writing y := Lλ2f −1(x), +this shows +|fLλ1f −1f(y) − fLλ1(y)| ≤ L|f −1fy − y| + v(|f −1fy| ∨ |y|) +By definition of an SBE inverse it holds that |f −1f(y) − y| ≤ u(|y|) and in particular also |f −1f(y)| ≤ +|y| + u(|y|). I conclude that +|f1f2(x) − f12(x)| ≤ L · u(|y|) + v +� +|y| + u(|y|) +� +The right-hand side is a sublinear function in |y|, hence it only remains to show that |y| is bounded by a +linear function in x. Indeed +|y| = |Lλ2f −1(x)| ≤ |λ2| + |f −1(x)| ≤ |λ2| + L|x| + u(|x|) +This completes the proof, rendering Φ a group homomorphism. +Claim. Φ has discrete image and finite kernel. +Proof. I show that for any radius R > 0, there are finitely many λ ∈ Λ for which gλx0 ∈ B(x0, R). I.e., that +there is a finite number of Φ(Λ)-orbit points, with multiplicities, inside an R ball in X. In particular the +set {λ ∈ Λ | gλx0 = x0} is finite, and clearly contains Ker(Φ). In addition, the actual number of Φ(Λ)-orbit +points inside that R ball is finite, so Φ(Λ) is discrete. +Let R > 0, and λ ∈ Λ. By the defining property of gλ and the definition of fλ, reverse triangle inequality +gives +d +� +x0, gλ(x0) +� +≥ d +� +x0, fλ(x0) +� +− d +� +gλ(x0), fλ(x0) +� +≥ |f(λ)| − uλ(|x0|) +Corollary 6.25 gives d +� +gλ(x0), fλ(x0) +� +≤ ux0(|λ|) for some sublinear function ux0 ∈ O(u). On the other +hand f is an SBE, and so |f(λ)| grows close to linearly in λ. Formally, +|f(λ)| = d +� +f(λ), x0 +� += d +� +f(λ), f(eΛ) +� +≥ 1 +Ld(λ, x0) − u(|λ| ∨ |eΛ|) ≥ 1 +L|λ| − u(|λ|) +52 + +To conclude, one has +d +� +x0, gλ(x0) +� +≥ 1 +L|λ| − u(|λ|) − ux0(|λ|) +and both u, ux0 are sublinear in |λ|. Therefore there is a bound S ∈ R>0 such that |λ| > S ⇒ 1 +L|λ| − +u(|λ|) − ux0(|λ|) > R. The group Λ is finitely generated and so only finitely many λ ∈ Λ admit |λ| ≤ S, +hence gλ(x0) ∈ B(x0, R) only for finitely many λ ∈ Λ. +Claim. There exists a sublinear function u′ : R≥0 → R≥1 such that +Γ · x0 ⊂ Nu′� +Φ(Λ) · x0 +� +Proof. I claim that there is a sublinear function u0, depending only on f and q0, such that for all γ ∈ G, +d +� +gλ(x0), γ(x0) +� +≤ u0(|γ|). +As before, I only have control on gλ via fλ, and so I use triangle inequality to get: +d +� +gλ(x0), γ(x0) +� +≤ d +� +gλ(x0), fλ(x0) +� ++ d +� +fλ(x0), γ(x0) +� +By Corollary 6.25, d +� +gλ(x0), fλ(x0) +� +≤ ux +0(|λ|) for a sublinear function ux0. +It is beneficial to distinguish between the SBE fΓ : Λ → Γ and the same SBE composed with the orbit +quasi-isometry q0 : Γ → X0. From now on I keep the notation fΓ : Λ → Γ for the SBE of the groups and f0 +for the same SBE composed with the orbit quasi-isometry so f0 = q0 ◦ fΓ. +Define λγ := f −1 +Γ (γ). I show that d +� +fλγ(x0), γ(x0) +� +is bounded by a function sublinear in γ. Indeed, +recall that I assumed without loss of generality q0 ◦f −1 +Γ (x0) = eΛ. Moreover, Γ is assumed to be torsion-free, +and so there is no ambiguity or trouble in defining the restriction of the map q−1 to the orbit Γ · x0 to be of +the form q−1(γx0) = γ. All together, this gives +d +� +fλγ(x0), γ(x0) +� += d +� +f0(λγ), γ(x0) +� += d +� +q0 ◦ fΓf −1 +Γ (γ), q0(γ) +� +Since fΓ is an SBE d +� +fΓf −1 +Γ (γ), γ +� +≤ u(γ). The fact that q0 is an (L′, C)-quasi-isometry implies that +d +� +q0 ◦ fΓf −1 +Γ (γ), q(γ) +� +≤ Ld +� +fΓf −1 +Γ (γ), γ +� ++ C ≤ L′u(|γ|) + C +Combining everything, one has +d +� +gλγ(x0), γ(x0) +� +≤ ux0(|λγ|) + L′u(|γ|) + C +As before, |λγ| = |f −1(γ)| ≤ |γ| + u(|γ|) ≤ 2|γ|, where the last inequality holds for all large enough γ. +What matters is that |λγ| is linear in |γ|. I conclude that indeed Γ · x0 ⊂ Nu′� +Φ(Λ) · x0 +� +for the sublinear +function u′ = ux0 + L′u + C ∈ O(u), as wanted. (To be pedantic, u′ = ux0 ◦ 2 + L′u + C ∈ O(u) where 2 is +the ‘multiplication by 2’ function, r �→ 2r). +Claim. There exists a sublinear function u′ : R≥0 → R≥1 such that +Φ(Λ) · x0 ⊂ Nu′(Γ · x0) +Proof. Let λ ∈ Λ. +Let γ = f(λ) and consider the distance d(gλx0, γx0). +From triangle inequality and +Corollary 6.25 one has +d(gλx0, γx0) ≤ d(gλx0, fλx0) + d(fλx0, γx0) ≤ u′(|λ|) + d(fλx0, γx0) +By definition of fλ it holds that fλ(x0) = f(λ) · x0 = γ · x0 and so d +� +fλ(x0), γx0 +� += 0. +Claims 6.4, 6.4, 6.4 and 6.4 result in Theorem 6.9. Theorem 6.4 then follows from Theorem 1.6. +53 + +References +[1] Paul Albuquerque. Patterson-Sullivan theory in higher rank symmetric spaces. GAFA Geom. Funct. +Anal., 9(1):1–28, March 1999. +[2] U. Bader and E. Leibtag. Homomorhic images of algebraic groups. arXiv preprint arXiv:2212.03055, +2022. +[3] W. Ballmann, M. Gromov, and V. Schroeder. Manifolds of Nonpositive Curvature. Birkh¨auser Boston, +1985. +[4] Werner Ballmann. Lectures on spaces of nonpositive curvature, volume 25. Springer Science & Business +Media, 1995. +[5] Yves Benoist and S´ebastien Miquel. Arithmeticity of discrete subgroups containing horospherical lat- +tices. Duke Mathematical Journal, 169(8):1485 – 1539, 2020. +[6] Armand Borel. Linear Algebraic Groups. Springer New York, 1991. +[7] N. Bourbaki. Lie Groups and Lie Algebras: Chapters 1-3. Bourbaki, Nicolas: Elements of mathematics. +Springer-Verlag, 1989. +[8] N. Bourbaki. Lie Groups and Lie Algebras: Chapters 7-9. Number 7-9 in Elements of mathematics. +Springer Berlin Heidelberg, 2004. +[9] Brian H. Bowditch. +Geometrical finiteness for hyperbolic groups. +Journal of Functional Analysis, +113(2):245–317, 1993. +[10] Brian H Bowditch. Geometrical finiteness with variable negative curvature. Duke Mathematical Journal, +77(1):229–274, 1995. +[11] Martin R Bridson and Andr´e Haefliger. Metric spaces of non-positive curvature, volume 319. Springer +Science & Business Media, 2013. +[12] Claude Chevalley. Theory of Lie Groups. Princeton University Press, 1946. +[13] Yves Cornulier. On sublinear bilipschitz equivalence of groups. Annales scientifiques de l’´Ecole normale +sup´erieure, 52, 02 2017. +[14] Yves Cornulier. On sublinear bilipschitz equivalence of groups. Ann. ENS, 52:1201–1242, 2019. +[15] SG Dani. On ergodic quasi-invariant measures of group automorphism. Israel Journal of Mathematics, +43(1):62–74, 1982. +[16] Cornelia Drut¸u. Quasi-isometric classification of non-uniform lattices in semisimple groups of higher +rank. Geometric and Functional Analysis - GAFA, 10, 06 2000. +[17] Cornelia Drut¸u and Michael Kapovich. Geometric group theory. American Mathematical Society, 2017. +[18] Cornelia Drut¸u and Mark Sapir. +Tree-graded spaces and asymptotic cones of groups. +Topology, +44(5):959–1058, 2005. +[19] Patrick Eberlein. Lattices in spaces of nonpositive curvature. Annals of Mathematics, 111(3):435–476, +1980. +[20] Patrick Eberlein. +Geometry of Nonpositively Curved Manifolds. +Chicago Lectures in Mathematics. +University of Chicago Press, 1996. +54 + +[21] A. Eskin and B. Farb. Quasi-flats and rigidity in higher rank symmetric spaces. Journal of the American +Mathematical Society, 10:653–692, 1997. +[22] Alex Eskin. Quasi-isometric rigidity of nonuniform lattices in higher rank symmetric spaces. Journal +of the American Mathematical Society, 11(2):321–361, 1998. +[23] Benson Farb. +The quasi-isometry classification of lattices in semisimple lie groups. +Mathematical +Research Letters, 4:705–717, 1997. +[24] Mikolaj Fraczyk and Tsachik Gelander. +Infinite volume and infinite injectivity radius. +Annals of +Mathematics, 197(1):389–421, 2023. +[25] Toshiaki Hattori. Geometric Limit Sets of Higher Rank Lattices. Proceedings of the London Mathemat- +ical Society, 90(3):689–710, 05 2005. +[26] Ernst Heintze and Hans-Christoph Im Hof. Geometry of horospheres. Journal of Differential Geometry, +12(4):481 – 491, 1977. +[27] Sigurdur Helgason. Differential Geometry, Lie Groups, and Symmetric Spaces. ISSN. Elsevier Science, +1979. +[28] Sigurdur Helgason. Differential geometry and symmetric spaces, volume 341. American Mathematical +Soc., 2001. +[29] Lizhen Ji. From symmetric spaces to buildings, curve complexes and outer spaces. Innovations in +Incidence Geometry, 10(none):33 – 80, 2009. +[30] M. Kapovich and B. Liu. Geometric finiteness in negatively pinched hadamard manifolds. Annales +Academiae Scientiarum Fennicae Mathematica, 2019. +[31] Inkang Kim. Rigidity on symmetric spaces. Topology, 43(2):393–405, 2004. +[32] B. Kleiner and B. Leeb. Rigidity of quasi-isometries for symmetric spaces and euclidean buildings. Publ. +Math. IHES, 86:115–197, 1997. +[33] Anthony W. Knapp. Lie Groups Beyond an Introduction. Progress in Mathematics. Birkh¨auser Boston, +2002. +[34] Enrico Leuzinger. An exhaustion of locally symmetric spaces by compact submanifolds with corners. +Inventiones Mathematicae, 121(1):389–410, December 1995. +[35] Enrico Leuzinger. Tits geometry, arithmetic groups and the proof of a conjecture of siegel. Journal of +Lie Theory, 14, 2002. +[36] Enrico Leuzinger. Critical exponents of discrete groups and ²–spectrum. Proceedings of the American +Mathematical Society, 132(3):919–927, 2004. +[37] Enrico Leuzinger. On polyhedral retracts and compactifications of locally symmetric spaces. Differential +Geometry and its Applications, 20(3):293–318, 2004. +[38] A. Lubotzky, S. Mozes, and M. S. Raghunathan. +The word and riemannian metrics on lattices of +semisimple groups. Publ. Math. IHES, 91(1):5–53, 2000. +[39] Dave Witte Morris. Introduction to arithmetic groups. arXiv preprint math/0106063, 2001. +[40] George D. Mostow. Arithmetic subgroups of groups with radical. Annals of Mathematics, 93(3):409–438, +1971. +55 + +[41] George D. Mostow. +Strong Rigidity of Locally Symmetric Spaces. +Annals of Mathematics Studies. +Princeton University Press, 1973. +[42] Shahar Mozes. Epimorphic subgroups and invariant measures. Ergodic Theory and Dynamical Systems, +15(6):1207–1210, 1995. +[43] Gabriel Pallier. G´eom´etrie asymptotique sous-lin´eaire : hyperbolicit´e, autosimilarit´e, invariants. PhD +thesis, Universit´e Paris-Saclay, 2019. +Th`ese de doctorat dirig´ee par Pansu, Pierre Math´ematiques +fondamentales Universit´e Paris-Saclay (ComUE) 2019. +[44] Gabriel Pallier. Large-scale sublinearly lipschitz geometry of hyperbolic spaces. Journal of the Institute +of Mathematics of Jussieu, 19(6):1831–1876, 2020. +[45] Gabriel Pallier. Sublinear quasiconformality and the large-scale geometry of Heintze groups. Conform. +Geom. Dyn., 24:131–163, 2020. +[46] Gabriel Pallier. Sublinear quasiconformality and the large-scale geometry of Heintze groups. Conform. +Geom. Dyn., 24:131–163, 2020. +[47] Gabriel Pallier and Yulan Qing. Sublinear bilipschitz equivalence and sublinearly morse boundaries, +2022. Preprint. +[48] Pierre Pansu. Metriques de Carnot-Caratheodory et quasiisometries des espaces symetriques de rang +un. Annals of Mathematics, 129(1):1–60, 1989. +[49] Gopal Prasad. Strong rigidity of q-rank 1 lattices. Inventiones mathematicae, 21:255–286, 1973. +[50] M.S. Raghunathan. Discrete Subgroups of Lie Groups. Ergebnisse der Mathematik und ihrer Grenzge- +biete. Springer-Verlag, 1972. +[51] A.A. Sagle, R.E. Walde, and R. Walde. Introduction to Lie Groups and Lie Algebras. Pure and Applied +Mathematics; A Series of Monographs and Tex. Academic Press, 1973. +[52] Richard Evan Schwartz. +The quasi-isometry classification of rank one lattices. +Publ. Math. IHES, +82(1):133–168, 1995. +[53] Richard Evan Schwartz. Quasi-isometric rigidity and diophantine approximation. Acta Mathematica, +177(1):75–112, 1996. +[54] Nimish Shah. Invariant measures and orbit closures on homogeneous spaces. In S.G. Dani and Tata Insti- +tute of Fundamental Research, editors, Proceedings of the International Colloquium on Lie Groups and +Ergodic Theory, Mumbai, 1996, Studies in mathematics, pages 229–272. Tata Institute of Fundamental +Research, 1996. +[55] Dennis Sullivan. +On The Ergodic Theory at Infinity of an Arbitrary Discrete Group of Hyperbolic +Motions, pages 465–496. Princeton University Press, 2016. +[56] Jacques Tits. Buildings of spherical type and finite BN-pairs. Springer-Verlag, Berlin,New York, 1974. +56 +