Skip to main content
Log in

Integrable measure equivalence and rigidity of hyperbolic lattices

  • Published:
Inventiones mathematicae Aims and scope

Abstract

We study rigidity properties of lattices in \(\operatorname {Isom}(\mathbf {H}^{n})\simeq \mathrm {SO}_{n,1}({\mathbb{R}})\), n≥3, and of surface groups in \(\operatorname {Isom}(\mathbf {H}^{2})\simeq \mathrm {SL}_{2}({\mathbb{R}})\) in the context of integrable measure equivalence. The results for lattices in \(\operatorname {Isom}(\mathbf {H}^{n})\), n≥3, are generalizations of Mostow rigidity; they include a cocycle version of strong rigidity and an integrable measure equivalence classification. Despite the lack of Mostow rigidity for n=2 we show that cocompact lattices in \(\operatorname {Isom}(\mathbf {H}^{2})\) allow a similar integrable measure equivalence classification.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Similar content being viewed by others

Notes

  1. Here ≃ means locally isomorphic.

  2. Any lcsc group containing a lattice is necessarily unimodular.

  3. If one only requires equivariance for almost all g 1,g 2G one can always modify Φ on a null set to get an everywhere equivariant map [63, Appendix B].

  4. For the formulation of Mostow rigidity above we have to assume that G has trivial center.

  5. Every connected lcsc group is compactly generated [57, Corollary 6.12 on p. 58].

  6. The formulation of the virtual isomorphism case in terms of induced actions is due to Kida [34].

  7. Any probability measure in the Lebesgue measure class would do it.

  8. \(\operatorname {vol}(B(0,r))\) is here the Lebesgue measure of the Euclidean ball of radius r around 0∈ℝn.

  9. The reader should note that in loc. cit. the Haar measure is normalized by \(\operatorname {vol}(\varGamma\backslash \mathbf {H}^{n})\) whereas we normalize it by 1.

  10. The target \(\mathcal {G}\) is assumed to be locally compact in this reference but the proof therein works the same for a Polish group \(\mathcal {G}\).

References

  1. Bader, U., Furman, A., Sauer, R.: Efficient subdivision in hyperbolic groups and applications. Groups Geom. Dyn. (2013, to appear). arXiv:1003.1562

  2. Bridson, M.R., Tweedale, M., Wilton, H.: Limit groups, positive-genus towers and measure-equivalence. Ergod. Theory Dyn. Syst. 27(3), 703–712 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  3. Bucher, M., Burger, M., Iozzi, A.: A dual interpretation of the Gromov–Thurston proof of Mostow rigidity and volume rigidity for representations of hyperbolic lattices (2012). arXiv:1205.1018

  4. Burger, M., Iozzi, A.: Boundary maps in bounded cohomology. Appendix to: “Continuous bounded cohomology and applications to rigidity theory” [Geom. Funct. Anal. 12(2), 219–280 (2002)] by Burger, and N. Monod. Geom. Funct. Anal. 12(2), 281–292 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  5. Burger, M., Iozzi, A.: A useful formula in bounded cohomology. In Seminaires et Congres, nr. 18, to appear. Available at http://www.math.ethz.ch/~iozzi/grenoble.ps

  6. Burger, M., Monod, N.: Continuous bounded cohomology and applications to rigidity theory. Geom. Funct. Anal. 12(2), 219–280 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  7. Burger, M., Mozes, S.: \({\rm CAT}\)(-1)-spaces, divergence groups and their commensurators. J. Am. Math. Soc. 9(1), 57–93 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  8. Casson, A., Jungreis, D.: Convergence groups and Seifert fibered 3-manifolds. Invent. Math. 118(3), 441–456 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  9. Corlette, K.: Archimedean superrigidity and hyperbolic geometry. Ann. Math. (2) 135(1), 165–182 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  10. Corlette, K., Zimmer, R.J.: Superrigidity for cocycles and hyperbolic geometry. Int. J. Math. 5(3), 273–290 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  11. Fell, J.M.G., Doran, R.S.: Representations of -Algebras, Locally Compact Groups, and Banach -Algebraic Bundles. Vol. 1. Pure and Applied Mathematics, vol. 125. Academic Press, San Diego (1988). Basic representation theory of groups and algebras

    MATH  Google Scholar 

  12. Fisher, D., Hitchman, T.: Cocycle superrigidity and harmonic maps with infinite-dimensional targets. Int. Math. Res. Not. (2006). doi:10.1155/IMRN/2006/72405

    MathSciNet  Google Scholar 

  13. Fisher, D., Morris, D.W., Witte, K.: Nonergodic actions, cocycles and superrigidity. N.Y. J. Math. 10, 249–269 (2004) (electronic)

    MATH  Google Scholar 

  14. Furman, A.: Orbit equivalence rigidity. Ann. Math. (2) 150(3), 1083–1108 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  15. Furman, A.: Gromov’s measure equivalence and rigidity of higher rank lattices. Ann. Math. (2) 150(3), 1059–1081 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  16. Furman, A.: Mostow-Margulis rigidity with locally compact targets. Geom. Funct. Anal. 11(1), 30–59 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  17. Furman, A.: A survey of measured group theory. In: Geometry, Rigidity, and Group Actions. Chicago Lectures in Math., pp. 296–374. Univ. Chicago Press, Chicago (2011)

    Google Scholar 

  18. Furstenberg, H.: Boundary theory and stochastic processes on homogeneous spaces. In: Harmonic Analysis on Homogeneous Spaces, Williams Coll., Williamstown, Mass., 1972. Proc. Sympos. Pure Math., vol. XXVI, pp. 193–229. Am. Math. Soc., Providence (1973)

    Chapter  Google Scholar 

  19. Furstenberg, H.: Rigidity and Cocycles for Ergodic Actions of Semisimple Lie Groups (after G.A. Margulis and R. Zimmer), Bourbaki Seminar, Vol. 1979/80. Lecture Notes in Math., vol. 842, pp. 273–292. Springer, Berlin (1981)

    Google Scholar 

  20. Gabai, D.: Convergence groups are Fuchsian groups. Ann. Math. (2) 136(3), 447–510 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  21. Gaboriau, D.: Coût des relations d’équivalence et des groupes. Invent. Math. 139(1), 41–98 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  22. Gaboriau, D.: Invariants l 2 de relations d’équivalence et de groupes. Publ. Math. Inst. Hautes Études Sci. 95, 93–150 (2002) (French)

    Article  MathSciNet  MATH  Google Scholar 

  23. Gaboriau, D.: Examples of groups that are measure equivalent to the free group. Ergod. Theory Dyn. Syst. 25(6), 1809–1827 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  24. Ghys, É.: Groups acting on the circle. Enseign. Math. (2) 47(3–4), 329–407 (2001)

    MathSciNet  MATH  Google Scholar 

  25. Gromov, M.: Asymptotic invariants of infinite groups. In: Geometric Group Theory, Vol. 2, Sussex, 1991. London Math. Soc. Lecture Note Ser., vol. 182, pp. 1–295. Cambridge Univ. Press, Cambridge (1993)

    Google Scholar 

  26. Hjorth, G.: A converse to Dye’s theorem. Trans. Am. Math. Soc. 357(8), 3083–3103 (2005) (electronic)

    Article  MathSciNet  MATH  Google Scholar 

  27. Haagerup, U., Munkholm, H.J.: Simplices of maximal volume in hyperbolic n-space. Acta Math. 147(1–2), 1–11 (1981)

    Article  MathSciNet  MATH  Google Scholar 

  28. Ioana, A.: Cocycle superrigidity for profinite actions of property (T) groups. Duke Math. J. 157(2), 337–367 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  29. Ioana, A., Peterson, J., Popa, S.: Amalgamated free products of weakly rigid factors and calculation of their symmetry groups. Acta Math. 200(1), 85–153 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  30. Iozzi, A.: Bounded cohomology, boundary maps, and rigidity of representations into \({\rm Homeo}_{+}(S^{1})\) and \({\rm SU}(1,n)\). In: Rigidity in Dynamics and Geometry, Cambridge, 2000, pp. 237–260. Springer, Berlin (2002)

    Chapter  Google Scholar 

  31. Ivanov, N.V.: Foundations of the theory of bounded cohomology. Zap. Nauchn. Semin. Leningr. Otd. Mat. Inst. Steklova (LOMI) 143, 69–109 (1985), 177–178 (Russian, with English summary)

    MATH  Google Scholar 

  32. Kechris, A.S.: Classical Descriptive Set Theory. Graduate Texts in Mathematics, vol. 156. Springer, Berlin (1995)

    Book  MATH  Google Scholar 

  33. Kida, Y.: The mapping class group from the viewpoint of measure equivalence theory. Mem. Am. Math. Soc. 196(916), vii+190 (2008)

    MathSciNet  Google Scholar 

  34. Kida, Y.: Orbit equivalence rigidity for ergodic actions of the mapping class group. Geom. Dedic. 131, 99–109 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  35. Kida, Y.: Measure equivalence rigidity of the mapping class group. Ann. Math. (2) 171(3), 1851–1901 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  36. Kida, Y.: Rigidity of amalgamated free products in measure equivalence. J. Topol. 4(3), 687–735 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  37. Löh, C.: Measure homology and singular homology are isometrically isomorphic. Math. Z. 253(1), 197–218 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  38. Margulis, G.A.: Non-uniform lattices in semisimple algebraic groups. In: Lie Groups and Their Representations, Proc. Summer School on Group Representations of the Bolyai János Math. Soc., Budapest, 1971, pp. 371–553. Halsted, New York (1975)

    Google Scholar 

  39. Margulis, G.A.: Discrete groups of motions of manifolds of nonpositive curvature. In: Proceedings of the International Congress of Mathematicians, Vancouver, B.C., 1974, vol. 2, pp. 21–34. Canad. Math. Congress, Montreal (1975) (Russian)

    Google Scholar 

  40. Margulis, G.A.: Discrete Subgroups of Semisimple Lie Groups. Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)], vol. 17. Springer, Berlin (1991)

    Book  MATH  Google Scholar 

  41. Monod, N.: Continuous Bounded Cohomology of Locally Compact Groups. Lecture Notes in Mathematics, vol. 1758. Springer, Berlin (2001)

    Book  MATH  Google Scholar 

  42. Monod, N., Shalom, Y.: Cocycle superrigidity and bounded cohomology for negatively curved spaces. J. Differ. Geom. 67(3), 395–455 (2004)

    MathSciNet  MATH  Google Scholar 

  43. Monod, N., Shalom, Y.: Orbit equivalence rigidity and bounded cohomology. Ann. Math. (2) 164(3), 825–878 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  44. Mostow, G.D.: Strong Rigidity of Locally Symmetric Spaces. Annals of Mathematics Studies, vol. 78. Princeton University Press, Princeton (1973)

    MATH  Google Scholar 

  45. Pansu, P., Zimmer, R.J.: Rigidity of locally homogeneous metrics of negative curvature on the leaves of a foliation. Isr. J. Math. 68(1), 56–62 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  46. Popa, S.: On a class of type \({\rm II}\sb{1}\) factors with Betti numbers invariants. Ann. Math. (2) 163(3), 809–899 (2006)

    Article  MATH  Google Scholar 

  47. Popa, S.: Strong rigidity of \(\rm II\sb{1}\) factors arising from malleable actions of w-rigid groups. I. Invent. Math. 165(2), 369–408 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  48. Popa, S.: Strong rigidity of \(\rm II\sb{1}\) factors arising from malleable actions of w-rigid groups. II. Invent. Math. 165(2), 409–451 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  49. Popa, S.: Cocycle and orbit equivalence superrigidity for malleable actions of w-rigid groups. Invent. Math. 170(2), 243–295 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  50. Popa, S.: Deformation and Rigidity for Group Actions and von Neumann Algebras. International Congress of Mathematicians, vol. I, pp. 445–477. Eur. Math. Soc., Zürich (2007)

    Google Scholar 

  51. Popa, S.: On the superrigidity of malleable actions with spectral gap. J. Am. Math. Soc. 21(4), 981–1000 (2008)

    Article  MATH  Google Scholar 

  52. Prasad, G.: Strong rigidity of \({\bf Q}\)-rank 1 lattices. Invent. Math. 21, 255–286 (1973)

    Article  MathSciNet  MATH  Google Scholar 

  53. Ratcliffe, J.G.: Foundations of Hyperbolic Manifolds. Graduate Texts in Mathematics, vol. 149. Springer, New York (1994)

    Book  MATH  Google Scholar 

  54. Ratner, M.: Interactions Between Ergodic Theory, Lie Groups, and Number Theory, Zürich, 1994, vol. 2, pp. 157–182. Birkhäuser, Basel (1995)

    Google Scholar 

  55. Shalom, Y.: Rigidity, unitary representations of semisimple groups, and fundamental groups of manifolds with rank one transformation group. Ann. Math. (2) 152(1), 113–182 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  56. Shalom, Y.: Measurable Group Theory. European Congress of Mathematics, pp. 391–423. Eur. Math. Soc., Zürich (2005)

    Google Scholar 

  57. Stroppel, M.: Locally Compact Groups. EMS Textbooks in Mathematics. Eur. Math. Soc., Zürich (2006)

    Book  MATH  Google Scholar 

  58. Tits, J.: Sur le groupe des automorphismes d’un arbre. In: Essays on Topology and Related Topics (Mémoires dédiés à Georges de Rham). Springer, Berlin (1970)

    Google Scholar 

  59. Thurston, W.P.: The geometry and topology of three-manifolds (1978). Available at http://www.msri.org/publications/books/gt3m

  60. Zastrow, A.: On the (non)-coincidence of Milnor-Thurston homology theory with singular homology theory. Pac. J. Math. 186(2), 369–396 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  61. Zimmer, R.J.: Strong rigidity for ergodic actions of semisimple Lie groups. Ann. Math. (2) 112(3), 511–529 (1980)

    Article  MathSciNet  MATH  Google Scholar 

  62. Zimmer, R.J.: On the Mostow rigidity theorem and measurable foliations by hyperbolic space. Isr. J. Math. 43(4), 281–290 (1982)

    Article  MathSciNet  MATH  Google Scholar 

  63. Zimmer, R.J.: Ergodic Theory and Semisimple Groups. Monographs in Mathematics, vol. 81. Birkhäuser, Basel (1984)

    Book  MATH  Google Scholar 

Download references

Acknowledgements

U. Bader and A. Furman were supported in part by the BSF grant 2008267. U. Bader was also supported in part by the ISF grant 704/08 and the ERC grant 306706. A. Furman was partly supported by the NSF grants DMS 0604611 and 0905977. R. Sauer acknowledges support from the Deutsche Forschungsgemeinschaft, made through grant SA 1661/1-2.

We thank the referee for his detailed and careful report, especially for his recommendations that led to a restructuring of Sect. 3 in a previous version.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Uri Bader.

Appendices

Appendix A: Measure equivalence

The appendix contains some general facts related to measure equivalence (Definition 1.1), the strong ICC property (Definition 2.2), and the notions of taut couplings and groups (Definition 1.3).

1.1 A.1 The category of couplings

Measure equivalence is an equivalence relation on unimodular lcsc groups. Let us describe explicitly the constructions which show reflexivity, symmetry and transitivity of measure equivalence.

1.1.1 A.1.1 Tautological coupling

The tautological coupling is the (G×G)-coupling (G,m G ) given by \((g_{1},g_{2}):g\mapsto g_{1} g g_{2}^{-1}\). It demonstrates reflexivity of measure equivalence.

1.1.2 A.1.2 Duality

Symmetry is implied by the following: Given a (G,H)-coupling (Ω,m) the dual (Ω ,m ) is the (H,G)-coupling Ω with the same underlying measure space (Ω,m) and the H×G-action (h,g):ω ↦(g,h)ω .

1.1.3 A.1.3 Composition of couplings

Compositions defined below shows that measure equivalence is a transitive relation. Let G 1,H,G 2 be unimodular lcsc groups, and (Ω i ,m i ) be a (G i ,H)-coupling for i∈{1,2}. We describe the (G 1,G 2)-coupling \(\varOmega_{1}\times_{H}{\varOmega}^{*}_{2}\) modeled on the space of H-orbits on (Ω 1×Ω 2,m 1×m 2) with respect to the diagonal H-action. Consider measure isomorphisms for (Ω i ,mi) as in (1.1): For i∈{1,2} there are finite measure spaces (X i ,μ i ) and (Y i ,ν i ), measure-preserving actions G i ↷(X i ,μ i ) and H↷(Y i ,ν i ), measurable cocycles α i :G i ×X i H and β i :H×Y i G i , and measure space isomorphisms G i ×Y i Ω i H×X i with respect to which the G i ×H-actions are given by

The space \(\varOmega_{1}\times_{H}{\varOmega}^{*}_{2}\) with its natural G 1×G 2-action is equivariantly isomorphic to (X 1×X 2×H,μ 1×μ 2×m H ) endowed with the G 1×G 2-action

$$(g_1,g_2): (x_1,x_2,h)\mapsto \bigl(g_1\mathbin{.}x_1,\, g_2\mathbin{.}x_2,\, \alpha_1(g_1,x_1) h \alpha_2(g_2,x_2)^{-1} \bigr). $$

To see that it is a (G 1,G 2)-coupling, we identify this space with Z×G 1 equipped with the action

$$(g_1,g_2): \bigl(g',z\bigr)\mapsto \bigl(g_1 g' c(g_2,z)^{-1}, g_2\mathbin{.}z\bigr)\quad\bigl(g'\in G_1,\ z\in Z \bigr) $$

where Z=X 2×Y 1, while the action G 2Z and the cocycle c:G 2×ZG 1 are given by

$$ \begin{aligned} &g_2:(x,y) \mapsto\bigl(g_2\mathbin{.}x,\alpha_2(g_2,x)\mathbin{.}y\bigr), \\ &c\bigl(g_2,(x,y)\bigr)=\beta_1\bigl( \alpha_2(g_2,x), y\bigr). \end{aligned} $$
(A.1)

Similarly, \(\varOmega_{1}\times_{H}{\varOmega}^{*}_{2}\cong W\times G_{2}\), for WX 1×Y 2.

1.1.4 A.1.4 Morphisms

Let (Ω i ,m i ), i∈{1,2}, be two (G,H)-couplings. Let F:Ω 1Ω 2 be a measurable map such that for m 1-a.e. ωΩ 1 and every gG and every hH

$$F\bigl((g,h)\omega\bigr)=(g,h)F(\omega). $$

Such maps are called quotient maps or morphisms.

1.1.5 A.1.5 Compact kernels

Let (Ω,m) be a (G,H)-coupling, and let

$$\{1\}\to K\to G\to\bar{G}\to\{1\} $$

be a short exact sequence where K is compact. Then the natural quotient space \((\bar{\varOmega},\bar{m})=(\varOmega,m)/K\) is a \((\bar{G},H)\)-coupling, and the natural map \(F:\varOmega\to\bar {\varOmega}\), F:ω, is equivariant in the sense of \(F((g,h)\omega)=(\bar{g},h) F(\omega)\). This may be considered as an isomorphism of couplings up to compact kernel.

1.1.6 A.1.6 Passage to lattices

Let (Ω,m) be a (G,H)-coupling, and let Γ<G be a lattice. By restricting the G×H-action on (Ω,m) to Γ×H we obtain a (Γ,H)-coupling. Formally, this follows by considering (G,m G ) as a Γ×G-coupling and considering the composition G× G Ω as Ω with the Γ×H-action.

1.2 A.2 Lp-integrability conditions

Let G and H be compactly generated unimodular lcsc groups equipped with proper norms |⋅| G and |⋅| H . Let c:G×XH be a measurable cocycle, and fix some p∈[1,∞). For gG we define

$$\|g\|_{c,p}= \biggl(\int_X \bigl|c(g,x)\bigr|^p_H \,d\mu(x) \biggr)^{1/p}. $$

For p=∞ we use the essential supremum. Assume that ∥g c,p <∞ for a.e. gG. We claim that there are constants a,A>0 so that for every gG

$$ \|g\|_{c,p}\le A\cdot|g|_G+a. $$
(A.2)

Hence c is Lp-integrable in the sense of Definition 1.5. The key observation here is that ∥−∥ c,p is subadditive. Indeed, by the cocycle identity, subadditivity of the norm |−| H , and the Minkowski inequality, for any g 1,g 2G we get

For t>0 denote E t ={gG:∥g c,p <t}. We have E t E s E s+t for any t,s>0. Fix t large enough so that m G (E t )>0. By [11, Corollary 12.4 on p. 235], E 2t E t E t has a non-empty interior. Hence any compact subset of G can be covered by finitely many translates of E 2t . The subadditivity implies that ∥g c,p is bounded on compact sets. This gives (A.2).

Lemma A.1

Let G,H,L be compactly generated groups, G↷(X,μ), H↷(Y,ν) be finite measure-preserving actions, and α:G×XH and β:H×YL be Lp-integrable cocycles for some 1≤p≤∞. Consider Z=X×Y and GZ by g:(x,y)↦(g.x,α(g,x).y). Then the cocycle γ:G×ZL given by

$$\gamma\bigl(g,(x,y)\bigr)=\beta\bigl(\alpha(g,x),y\bigr). $$

is Lp-integrable.

Proof

For p=∞ the claim is obvious. Assume p<∞. Let A,a,B,b be constants such that ∥h β,p B⋅|h| H +b and ∥g α,p A⋅|g| G +a. Then

for appropriate constants c>0 and C>0. □

Lemma A.2

Let G 1,H,G 2 be compactly generated unimodular lcsc groups. For i∈{1,2} let (Ω i ,m i ) be an Lp-integrable (G i ,H)-coupling. Then \(\varOmega_{1}\times_{H} {\varOmega}^{*}_{2}\) is an Lp-integrable (G 1,G 2)-coupling.

Proof

This follows from Lemma A.1 using the explicit description (A.1) of the cocycles for \(\varOmega_{1}\times_{H} {\varOmega}^{*}_{2}\). □

We conclude that for each 1≤p≤∞, Lp-measure equivalence is an equivalence relation between compactly generated unimodular lcsc groups.

1.3 A.3 Tautening maps

Lemma A.3

Let G be a lcsc group, Γ a countable group and j 1,j 2:ΓG be homomorphisms with Γ i =j i (Γ) being lattices in G. Assume that G is taut (resp. p-taut and Γ i are Lp-integrable). Then there exists gG so that

$$j_2(\gamma)=g\,j_1(\gamma)\,g^{-1}\quad( \gamma\in\varGamma). $$

If \(\pi:G\to \mathcal {G}\) is a continuous homomorphism into a Polish group and G is taut relative to \(\pi:G\to \mathcal {G}\) (resp. G is p-taut relative to \(\pi:G\to \mathcal {G}\) and Γ i are Lp-integrable) then there exists \(y\in \mathcal {G}\) with

$$\pi\bigl(j_2(\gamma)\bigr)=y \pi\bigl(j_1(\gamma) \bigr)y^{-1}\quad(\gamma\in \varGamma). $$

Proof

We prove the more general second statement. The group

$$\varDelta =\bigl\{\bigl(j_1(\gamma),j_2(\gamma)\bigr)\in G \times G \mid\gamma\in \varGamma\bigr\} $$

is a closed discrete subgroup in G×G. The G×G-space Ω=(G×G)/Δ equipped with the G×G-invariant measure is easily seen to be a (G,G)-coupling. It will be Lp-integrable if Γ 1 and Γ 2 are Lp-integrable lattices. Let \(\varPhi:\varOmega\to \mathcal {G}\) be the tautening map. There are a,bG and \(x\in \mathcal {G}\) such that for all g 1,g 2G

$$\varPhi\bigl((g_1a,g_2b)\varDelta _f\bigr)= \pi(g_1) x \pi(g_2)^{-1}. $$

Since (a,b) and (j 1(γ)aa, j 2(γ)bb) are in the same Δ-coset, where g h=hgh −1, we get for all g 1,g 2G and every γΓ

$$\pi(g_1) x \pi(g_2)^{-1}=\pi(g_1) \pi\bigl(j_1(\gamma)^a\bigr) x\pi \bigl(j_2( \gamma)^b\bigr)^{-1} \pi(g_2)^{-1}. $$

This implies that j 1 and j 2 are conjugate homomorphisms. □

The following lemma relates tautening maps Φ:ΩG and cocycle rigidity for ME-cocycles.

Lemma A.4

Let G be a unimodular lcsc group, \(\mathcal {G}\) be a Polish group, \(\pi :G\to \mathcal {G}\) a continuous homomorphism. Let (Ω,m) be a (G,G)-coupling and α:G×XG, β:G×YG be the corresponding ME-cocycles. Then there is a tautening map \(\varOmega\to \mathcal {G}\) iff the \(\mathcal {G}\)-valued cocycle πα is conjugate to π, that is,

$$\pi\circ\alpha(g,x)=f(g\mathbin{.}x)^{-1} \pi(g) f(x). $$

Moreover, Ω is taut relative to π if such a measurable map \(f:X\to \mathcal {G}\) is unique. This is also equivalent to πβ being uniquely conjugate to \(\pi: G\to \mathcal {G}\).

Proof

Let α:G×XG be the ME-cocycle associated to a measure space isomorphism i:(G,m G )×(X,μ)→(Ω,m) as in (1.1). In particular,

$$(g_1,g_2): i(g,x)\mapsto i\bigl(g_2 g \alpha(g_1,x)^{-1},\,g_1\mathbin{.}x\bigr). $$

We shall now establish a 1-to-1 correspondence between Borel maps \(f:X\to\nobreak \mathcal {G}\) with

$$\pi\circ\alpha(g,x)=f(g\mathbin{.}x)^{-1}\pi(g) f(x) $$

and tautening maps \(\varPhi:\varOmega\to \mathcal {G}\). Given f as above one verifies that

$$\varPhi:\varOmega\to \mathcal {G},\qquad\varPhi\bigl(i(g,x)\bigr)=f(x) \pi(g)^{-1} $$

is G×G-equivariant.

For the converse direction, suppose Φ:ΩG is a tautening map. Thus,

$$g_1 \varPhi(g_0,x) g_2^{-1}= \varPhi\bigl((g_1,g_2) (g_0,x)\bigr)=\varPhi \bigl(g_2 g_0 \alpha(g_1,x)^{-1},\, g_1\mathbin{.}x\bigr). $$

For μ-a.e. xX and a.e. gG the value of Φ(g,x)g is constant f(x). If we substitute g 0=g 1=g and g 2=(g,x)g −1 in the above identity, then we obtain α(g,x)=f(g.x)−1 gf(x). □

Lemma A.5

Let G be a unimodular lcsc group, \(\mathcal {G}\) be a Polish group, \(\pi :G\to \mathcal {G}\) a continuous homomorphism. Then G is (p-)taut, that is every (p-integrable) (G,G)-coupling is taut relative to π, iff every ergodic (p-integrable) (G,G)-coupling is taut relative to π.

Proof

We give the proof in the p-integrable case, the case without the integrability condition being simpler. We assume that every ergodic p-integrable (G,G)-coupling is taut relative to π and let (Ω,m) be an arbitrary p-integrable coupling. We fix a fundamental domain (X,μ) for the second G action such that the associated cocycle α:G×XG is p-integrable.

Let μ=∫μ t (t) be the G-ergodic decomposition of (X,μ). By [15, Lemma 2.2] it corresponds to the (G×G)-ergodic decomposition of (Ω,m) into ergodic couplings (Ω,m t ), so that m=∫m t (t).

Let |_|:G→ℕ be the length function associated to some word-metric on G. The p-integrability of α means that

$$\int\int_X{ \bigl\vert\alpha(g,x) \bigr\vert}^p\,d \mu_t(x)\,d\eta (t)=\int_X{ \bigl\vert\alpha(g,x) \bigr\vert}^p\,d\mu(x)<\infty $$

for every gG. By Fubini’s theorem ∫ X |α(g,x)|p t (x)<∞ for η-a.e. t. Hence (Ω,m t ) is p-integrable for η-a.e. m t , and in particular it is taut relative to π, by our assumption. It follows by Lemma A.4 that the cocycle πα is conjugate to the constant cocycle π over η-a.e. (X,μ t ). Then by [13, Corollary 3.6]Footnote 10 πα is conjugate to π over (X,μ). Again, by Lemma A.4 we conclude that (Ω,m) is taut relative to π. □

1.4 A.4 Strong ICC property

Lemma A.6

Let G be a unimodular lcsc group, \(\mathcal {G}\) a Polish group, \(\pi:G\to \mathcal {G}\) a continuous homomorphism. Let Γ<G be a lattice. Then \(\mathcal {G}\) is strongly ICC relative to π(G) if and only if it is strongly ICC relative to π(Γ).

Proof

Clearly if \(\mathcal {G}\) is strongly ICC relative to π(Γ) then it is also strongly ICC relative to π(G). The reverse implication follows by averaging a π(Γ) invariant measure over G/Γ. □

Lemma A.7

Let G be a unimodular lcsc group, \(\mathcal {G}\) a Polish group, \(\pi:G\to \mathcal {G}\) a continuous homomorphism. Suppose that \(\mathcal {G}\) is not strongly ICC relative to π(G). Then there is a (G,G)-coupling (Ω,m) with two distinct tautening maps to \(\mathcal {G}\).

Proof

Let μ be a Borel probability measure on \(\mathcal {G}\) invariant under conjugations by π(G). Consider \(\varOmega=G\times \mathcal {G}\) with the measure m=m G ×μ where m G denotes the Haar measure, and measure-preserving G×G-action

$$(g_1,g_2): (g,x)\mapsto\bigl(g_1 g g_2^{-1},\, \pi(g_2)\, x\, \pi(g_2)^{-1}\bigr). $$

This is clearly a (G,G)-coupling and the following measurable maps \(\varPhi_{i}:\varOmega\to \mathcal {G}\), i∈{1,2}, are G×G-equivariant: Φ 1(g,x)=π(g) and Φ 2(g,x)=π(g)⋅x. Note that Φ 1=Φ 2 on a conull set iff μ=δ e . □

Lemma A.8

Let G be a unimodular lcsc group and \(\mathcal {G}\) a Polish group. Assume that \(\mathcal {G}\) is strongly ICC relative to π(G). Let (Ω,m) be a (G,G)-coupling. Then:

  1. (1)

    There is at most one tautening map \(\varPhi\colon\varOmega\to \mathcal {G}\).

  2. (2)

    Let F:(Ω,m)→(Ω 0,m 0) be a morphism of (G,G)-couplings and suppose that there exists a tautening map \(\varPhi:\varOmega\to \mathcal {G}\). Then it descends to Ω 0, i.e., Φ=Φ 0F for a unique tautening map \(\varPhi_{0}:\varOmega_{0}\to \mathcal {G}\).

  3. (3)

    If Γ 1,Γ 2<G are lattices, then \(\varPhi:\varOmega \to \mathcal {G}\) is unique as a Γ 1×Γ 2-equivariant map.

  4. (4)

    If Γ 1,Γ 2<G are lattices, and (Ω,m) admits a Γ 1×Γ 2-equivariant map \(\varPhi:\varOmega \to \mathcal {G}\), then Φ is G×G-equivariant.

  5. (5)

    If \(\eta:\varOmega\to \operatorname {Prob}(\mathcal {G})\), ωη ω , is a measurable G×G-equivariant map to the space of Borel probability measures on \(\mathcal {G}\) endowed with the weak topology, then it takes values in Dirac measures: We have η ω =δ Φ(ω), where \(\varPhi:\varOmega\to \mathcal {G}\) is the unique tautening map.

Proof

We start from the last claim and deduce the other ones from it.

(5). Given an equivariant map \(\eta:\varOmega\to \operatorname {Prob}(\mathcal {G})\) consider the convolution

$$\nu_\omega=\check{\eta}_\omega* \eta_\omega, $$

namely the image of η ω ×η ω under the map (a,b)↦a −1b. Then

$$\nu_{(g,h)\omega}=\nu_\omega^{\pi(g)}\quad(g,h\in G), $$

where the latter denotes the push-forward of ν ω under the conjugation

$$a\mapsto a^{\pi(g)}=\pi(g)^{-1}a \pi(g). $$

In particular, the map ων ω is invariant under the action of the second G-factor. Therefore ν ω descends to a measurable map \(\tilde{\nu}:\varOmega/G\to \operatorname {Prob}(\mathcal {G})\), satisfying

$$\tilde{\nu}_{g.x}=\tilde{\nu}_x^{\pi(g)}\quad(x\in X= \varOmega/G,\ g\in G). $$

Here we identify Ω/G with a finite measure space (X,μ) as in (1.1). Consider the center of mass

$$\bar{\nu}=\frac{1}{\mu(X)}\int_{X}\tilde{\nu}_x \,d\mu(x). $$

It is a probability measure on \(\mathcal {G}\), which is invariant under conjugations. By the strong ICC property relative to π(G) we get \(\bar{\nu}=\delta_{e}\). Since δ e is an extremal point of \(\operatorname {Prob}(\mathcal {G})\), it follows that m-a.e. ν(ω)=δ e . This implies that η ω =δ Φ(ω) for some measurable \(\varPhi:\varOmega\to \mathcal {G}\). The latter is automatically G×G-equivariant.

(1). If \(\varPhi_{1},\varPhi_{2}:\varOmega\to \mathcal {G}\) are tautening maps, then \(\eta_{\omega}=\frac{1}{2}(\delta_{\varPhi_{1}(\omega)}+\delta_{\varPhi_{2}(\omega)})\) is an equivariant map \(\varOmega\to \operatorname {Prob}(\mathcal {G})\). By (5) it takes values in Dirac measures, which is equivalent to the m-a.e. equality Φ 1=Φ 2.

(2). Disintegration of m with respect to m 0 gives a G×G-equivariant measurable map \(\varOmega_{0}\to\mathcal{M}(\varOmega)\), \(\omega\mapsto m_{\omega_{0}}\), to the space of finite measures on Ω. Then the map \(\eta:\varOmega_{0}\to \operatorname {Prob}(\mathcal {G})\), given by

$$\eta_{\omega_0}=\|m_{\omega_0}\|^{-1}\cdot\varPhi_* (m_{\omega_0}) $$

is G×G-equivariant. Hence by (5), \(\eta_{\omega _{0}}=\delta_{\varPhi_{0}(\omega_{0})}\) for the unique tautening map \(\varPhi_{0}:\varOmega_{0}\to \mathcal {G}\). The relation Φ=Φ 0F follows from the fact that Dirac measures are extremal.

(3) follows from (4) and (1).

(4). The claim is equivalent to: For m-a.e. ωΩ the map \(F_{\omega}\colon G\times G\to \mathcal {G}\) with

$$F_\omega(g_1,g_2)=\pi(g_1)^{-1} \,\varPhi\bigl((g_1,g_2)\omega\bigr)\,\pi(g_2) $$

is m G ×m G -a.e. constant Φ 0(ω). Note that the family {F ω } has the following equivariance property: For g 1,g 2,h 1,h 2G we have

Since Φ is Γ 1×Γ 2-equivariant, for m-a.e. ωΩ the map F ω descends to G/Γ 1×G/Γ 2. Let \(\eta_{\omega}\in \operatorname {Prob}(\mathcal {G})\) denote the distribution of F ω (⋅,⋅) over the probability space G/Γ 1×G/Γ 2, that is, for a Borel subset \(E\subset \mathcal {G}\)

$$\eta_\omega(E)=m_{G/\varGamma_1}\times m_{G/\varGamma_2}\bigl\{ (g_1,g_2) \mid F_\omega(g_1,g_2) \in E\bigr\}. $$

Since this measure is invariant under translations by G×G, it follows that η ω is a G×G-equivariant maps \(\varOmega\to \operatorname {Prob}(\mathcal {G})\). By (5) one has η ω =δ f(ω) for some measurable G×G-equivariant map \(f:\varOmega\to \mathcal {G}\). Hence F ω (g 1,g 2)=f(ω) for a.e. g 1,g 2G; it follows that

$$ \varPhi\bigl((g_1,g_2)\omega\bigr)= \pi(g_1) \varPhi(\omega) \pi(g_2)^{-1} $$
(A.3)

holds for m G ×m G ×m-a.e. (g 1,g 2,ω). □

Corollary A.9

Let \(\pi:G\to \mathcal {G}\) be as above and assume that \(\mathcal {G}\) is strongly ICC relative to π(G). Then the collection of all (G,G)-couplings which are taut relative to \(\pi:G\to \mathcal {G}\) is closed under the operations of taking the dual, compositions, quotients and extensions.

Proof

The uniqueness of tautening maps follow from the relative strong ICC property (Lemma A.8.(1)). Hence we focus on the existence of such maps.

Let \(\varPhi:\varOmega\to \mathcal {G}\) be a tautening map. Then Ψ(ω )=Φ(ω)−1 is a tautening map \({\varOmega}^{*}\to \mathcal {G}\).

Let \(\varPhi_{i}:\varOmega_{i}\to \mathcal {G}\), i=1,2, be tautening maps. Then Ψ([ω 1,ω 2])=Φ(ω 1)⋅Φ(ω 2) is a tautening map \(\varOmega_{1}\times_{G}\varOmega_{2}\to \mathcal {G}\).

If F:(Ω 1,m 1)→(Ω 2,m 2) is a quotient map and \(\varPhi_{1}:\varOmega_{1}\to \mathcal {G}\) is a tautening map, then, by Lemma A.8.(2), Φ 1 factors as Φ 1=Φ 2F for a tautening map \(\varPhi_{2}:\varOmega_{2}\to \mathcal {G}\). On the other hand, given a tautening map \(\varPhi_{2}:\varOmega_{2}\to \mathcal {G}\), the map Φ 1=Φ 2F is tautening for Ω 1. □

Appendix B: Bounded cohomology

Our background references for bounded cohomology, especially for the functorial approach to it, are [6, 41]. We summarize what we need from Burger-Monod’s theory of bounded cohomology. Since we restrict to discrete groups, some results we quote from this theory already go back to Ivanov [31].

2.1 B.1 Banach modules

All Banach spaces are over the field ℝ of real numbers. By the dual of a Banach space we understand the normed topological dual. The dual of a Banach space E is denoted by E . Let Γ be a discrete and countable group. A Banach Γ-module is a Banach space E endowed with a group homomorphism π from Γ into the group of isometric linear automorphisms of E. We use the module notation γe=π(γ)(e) or just γe=π(γ)(e) for γΓ and eE whenever the action is clear from the context. The submodule of Γ-invariant elements is denoted by E Γ. Note that E ΓE is closed.

If E and F are Banach Γ-modules, a Γ-morphism EF is a Γ-equivariant continuous linear map. The space \(\mathcal {B}(E,F)\) of continuous, linear maps EF is endowed with a natural Banach Γ-module structure via

$$ (\gamma\cdot f) (e)=\gamma f\bigl( \gamma^{-1}e\bigr). $$
(B.1)

The contragredient Banach Γ-module structure E associated to E is by definition \(\mathcal {B}(E,{\mathbb{R}})=E^{\ast}\) with the Γ-action (B.1). A coefficient Γ-module is a Banach Γ-module E contragredient to some separable continuous Banach Γ-module denoted by E . The choice of E is part of the data. The specific choice of E defines a weak-∗ topology on E. The only examples that appear in this paper are E=L(X,μ) with E =L 1(X,μ) and E=E =ℝ.

For a coefficient Γ-module E let \(\mathrm{C}_{\mathrm {b}}^{k}(\varGamma,E)\) be the Banach Γ-module L(Γ k+1,E) consisting of bounded maps from Γ k+1 to E endowed with the supremum norm and the Γ-action:

$$ (\gamma\cdot f) (\gamma_0,\dots, \gamma_k)=\gamma\cdot f\bigl(\gamma^{-1} \gamma_0,\dots,\gamma^{-1}\gamma_k\bigr). $$
(B.2)

For a coefficient Γ-module E and a standard Borel Γ-space S with quasi-invariant measure let \(\mathrm {L}_{\mathrm{w}\ast}^{\infty}(S, E)\) be the space of weak-∗-measurable essentially bounded maps from S to E, where maps are identified if they only differ on a null set. The space \(\mathrm {L}_{\mathrm{w}\ast }^{\infty}(S, E)\) is endowed with the essential supremum norm and the Γ-action (B.2). For a measurable space X the Banach space \(\mathcal {B}^{\infty}(X, E)\) is the space of weak-∗-measurable bounded maps from X to E endowed with supremum norm [4, Sect. 2] and the Γ-action (B.2).

2.2 B.2 Injective resolutions

Let Γ be a discrete group and E be a Banach Γ-module. The sequence of Banach Γ-modules \(\mathrm{C}_{\mathrm {b}}^{k}(\varGamma, E)\), k≥0, becomes a chain complex of Banach Γ-modules via the standard homogeneous coboundary operator

$$ d(f) (\gamma_0,\dots, \gamma_k)=\sum_{i\ge0}^k(-1)^i f(\gamma_0,\dots, \hat{\gamma_i},\dots, \gamma_k). $$
(B.3)

The bounded cohomology \(\mathrm{H}_{\mathrm{b}}^{\bullet}(\varGamma,E)\) of Γ with coefficients E is the cohomology of the complex of Γ-invariants \(\mathrm{C}_{\mathrm{b}}^{\bullet}(\varGamma, E)^{\varGamma}\). The bounded cohomology \(\mathrm{H}_{\mathrm{b}}^{\bullet}(\varGamma,E)\) inherits a semi-norm from \(\mathrm{C}_{\mathrm{b}}^{\bullet}(\varGamma, E)\): The (semi-)norm of an element \(x\in\mathrm{H}_{\mathrm {b}}^{k}(\varGamma, E)\) is the infimum of the norms of all cocycles in the cohomology class x.

Next we briefly recall the functorial approach to bounded cohomology as introduced by Ivanov [31] for discrete groups and further developed by Burger-Monod [6, 41]. We refer for the definition of relative injectivity of a Banach Γ-module to [41, Definition 4.1.2 on p. 32]. A strong resolution E of E is a resolution, i.e. an acyclic complex,

$$0\to E\to E^0\to E^1\to E^2\to\cdots $$

of Banach Γ-modules that has a chain contraction which is contracting with respect to the Banach norms. The key to the functorial definition of bounded cohomology are the following two theorems:

Theorem B.1

([6, Proposition 1.5.2])

Let E and F be Banach Γ-modules. Let E be a strong resolution of E. Let F be a resolution F by relatively injective Banach Γ-modules. Then any Γ-morphism EF extends to a Γ-morphism of resolutions E F which is unique up to Γ-homotopy. Hence EF induces functorially continuous linear maps H(E Γ)→H(F Γ).

Theorem B.2

([41, Corollary 7.4.7 on p. 80])

Let E be a Banach Γ-module. The complex \(E\to\mathrm{C}_{\mathrm{b}}^{\bullet}(\varGamma,E)\) with \(E\to \mathrm{C}_{\mathrm{b}}^{0}(\varGamma,E)\) being the inclusion of constant functions is a strong, relatively injective resolution.

For a coefficient Γ-module, a measurable space X with measurable Γ-action, and a standard Borel Γ-space S with quasi-invariant measure we obtain chain complexes \(\mathcal {B}^{\infty}(X^{\bullet+1},E)\) and Lw∗(S •+1,E) of Banach Γ-modules via the standard homogeneous coboundary operators (similar as in (B.3)).

The following result is important for expressing induced maps in bounded cohomology in terms of boundary maps [4].

Proposition B.3

([4, Proposition 2.1])

Let E be a coefficient Γ-module. Let X be a measurable space with measurable Γ-action. The complex \(E\to \mathcal {B}^{\infty}(X^{\bullet+1},E)\) with \(E\to \mathcal {B}^{\infty}(X,E)\) being the inclusion of constant functions is a strong resolution of E.

The next theorem is one of the main results of the functorial approach to bounded cohomology by Burger-Monod:

Theorem B.4

([6, Corollary 2.3.2, 41, Theorem 7.5.3 on p. 83])

Let S be a regular Γ-space and be E a coefficient Γ-module. Then E→Lw∗(S •+1,E) with E→Lw∗(S •+1,E) being the inclusion of constant functions is a strong resolution. If, in addition, S is amenable in the sense of Zimmer [41, Definition 5.3.1], then each Lw∗(S k+1,E) is relatively injective, and according to Theorem B.1 the cohomology groups H(Lw∗(S •+1,E)Γ) are canonically isomorphic to \(\mathrm{H}_{\mathrm{b}}^{\bullet}(\varGamma,E)\).

Definition B.5

Let S be a standard Borel Γ-space with a quasi-invariant probability measure μ. Let E be a coefficient Γ-module. The Poisson transform \(\mathrm {PT}^{\bullet}: \mathrm {L}_{\mathrm{w}\ast}(S^{\bullet+1},E)\to\mathrm{C}_{\mathrm {b}}^{\bullet}(\varGamma , E)\) is the Γ-morphism of chain complexes defined by

$$\mathrm {PT}^k(f) (\gamma_0,\dots,\gamma_k)= \int _{S^{k+1}} f(\gamma_0 s_0,\dots, \gamma_k s_k)\,d\mu(s_0)\cdots d \mu(s_k). $$

If S is amenable, then the Poisson transform induces a canonical isomorphism in cohomology (Theorem B.4). By the same theorem this isomorphism does not depend on the choice of μ within a given measure class.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Bader, U., Furman, A. & Sauer, R. Integrable measure equivalence and rigidity of hyperbolic lattices. Invent. math. 194, 313–379 (2013). https://doi.org/10.1007/s00222-012-0445-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00222-012-0445-9

Keywords

Navigation