Skip to main content
Log in

Monotonic cocycles

  • Published:
Inventiones mathematicae Aims and scope

Abstract

We develop a “local theory” of multidimensional quasiperiodic \({\mathrm {SL}}(2,{\mathbb R})\) cocycles which are not homotopic to a constant. It describes a \(C^1\)-open neighborhood of cocycles of rotations and applies irrespective of arithmetic conditions on the frequency, being much more robust than the local theory of \({\mathrm {SL}}(2,{\mathbb R})\) cocycles homotopic to a constant. Our analysis is centered around the notion of monotonicity with respect to some dynamical variable. For such monotonic cocycles, we obtain a sharp rigidity result, minimality of the projective action, typical nonuniform hyperbolicity, and a surprising result of smoothness of the Lyapunov exponent (while no better than Hölder can be obtained in the case of cocycles homotopic to a constant, and only under arithmetic restrictions). Our work is based on complexification ideas, extended “à la Lyubich” to the smooth setting (through the use of asymptotically holomorphic extensions). We also develop a counterpart of this theory centered around the notion of monotonicity with respect to a parameter variable, which applies to the analysis of \({\mathrm {SL}}(2,{\mathbb R})\) cocycles over more general dynamical systems and generalizes key aspects of Kotani Theory. We conclude with a more detailed discussion of one-dimensional monotonic cocycles, for which results about rigidity and typical nonuniform hyperbolicity can be globalized using a new result about convergence of renormalization.

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.

Similar content being viewed by others

Notes

  1. It is somewhat delicate (in the ergodic case) to construct examples of cocycles not homotopic to a constant which are not premonotonic. A non-negligible (positive measure on parameters) set of such examples can be obtained by forcing a certain behavior of the “critical points” that arise in the approach of Lai-Sang Young [36] (of Benedicks–Carleson [9] flavor). In view of the results of this paper, the existence of non-premonotonic cocycles can also be deduced using the work of Wang–You (which is also related to [36]), see Remark 1.1.

  2. Here we follow the usual convention in dynamical system to reserve typical for a measure-theoretical notion of genericity. The most commonly used such notion is known as prevalence, see [24], which is in particular satisfied by sets whose complement have positive codimension, as is the case here.

  3. The set of premonotonic cocycles with zero Lyapunov exponent has infinite codimension when the number of frequencies \(d\) is at least \(2\), and finite codimension when \(d=1\). We will come back to this issue elsewhere.

  4. Particularly, the analysis of the spectrum of the critical almost Mathieu operator, where the Lyapunov exponent is still zero, does not reduce to the local situation.

  5. Since this result indeed assumes at least a Diophantine condition on \(\alpha \), and the cocycle is homotopic to a constant, the conclusion is in fact equivalent to the existence of a conjugacy to a constant cocycle of rotations.

  6. It is easy to see that \(\deg \) is just the \(\mu \)-average of \(\deg (x)\), where \(\deg (x)\) is the topological degree of \(\theta \mapsto A_{\theta ,s}(x)\) as a map \({\mathbb R}/{\mathbb Z}\rightarrow {\mathrm {SL}}(2,{\mathbb R})\) (in particular, if \(X\) is connected, \(\deg \) is an integer).

  7. A continuous harmonic function \(f\) on the disk is the Poisson integral of its restriction \(\rho \) to the boundary of the disk. If \(\rho \) is of bounded variation, the tangential derivative \(\partial _\sigma f\) is the Poisson integral of the measure \(d\rho \). Fatou theorem asserts that for a.e point on the boundary, the radial limit of \(\partial _\sigma f\) is \(d\rho /d\sigma \). The situation on the strip is easily reduced to the one on the disk.

  8. More generally, we can also let the dynamics vary, thus considering a sequence of \(\mu \)-preserving homeomorphisms \(f_k:X \rightarrow X\) converging to \(f\).

  9. We apologize for this collision of definition with the notion of \(\epsilon \)-monotonic cocycle.

  10. To see this, define \(\delta _{v,w} \xi _n=\delta _\gamma \xi _n\) and \(\delta _{v,w} \zeta =\delta _\gamma \zeta \) where \(\gamma \) is any path homotopic to \(\gamma (t,x)=A(x+v+t w)\). Notice that \(\rho _{A^w_\theta }\) is given, up to additive constant, by \(\delta _{0,\theta w} \zeta \). Let us show that \(\delta _{v,w} \zeta =\langle l,w \rangle \). It is clear that \(\delta _{v,w} \xi _n(x,z_0,z_1)=\delta _{0,w}(x+v,z_0,z_1)\), so that \(\delta _{v,w} \zeta \) does not depend on \(v\). By Remark 2.2, \(\delta _{v,w} \zeta +\delta _{v+w,w'} \zeta =\delta _{v,w+w'} \zeta \). This shows that \(\delta _{v,w} \zeta \) is a linear function of \(w\). Moreover, for \(w \in {\mathbb Z}^d\) we have exactly \(\delta _{v,w} \xi _n=\langle l,w \rangle n\), so that \(\delta _{v,w} \zeta =\langle l,w \rangle \).

  11. Indeed we can construct a monotonic decreasing family \(\tilde{A}_t \in C^0({\mathbb R}^d/{\mathbb Z}^d,{\mathrm {SL}}(2,{\mathbb R}))\), \(t \in [0,1]\), such that \(\tilde{A}_0(\cdot )=A_{\theta _0}(\cdot -C h w))\), \(\tilde{A}_1(\cdot )=A_{\theta _0}(\cdot +C h w)\) and \(\tilde{A}_{1/2}=A_{\theta _0+h}\), and \(\tilde{A}_t\) is close to \(A_{\theta _0}\) for every \(t\) (so that we remain in a region where there is a continuous determination of \(\rho \)). Monotonicity of this family gives \(\rho _{\tilde{A}_1} \le \rho _{\tilde{A}_{1/2}} \le \rho _{\tilde{A}_0}\).

  12. Theorem 2.1 gives a \(C^1\) conjugacy under a slightly stronger \(C^{2+\epsilon }\) condition. Under \(C^{1+\epsilon }\), it still gives a continuous invariant section \(m:{\mathbb R}^d/{\mathbb Z}^d \rightarrow \overline{{\mathbb D}}\), which implies the \(C^0\) conjugacy, unless the invariant section is real, i.e., it lies in \(\partial {\mathbb D}\). However, this last possibility is impossible here, since \(A\) is non-homotopic to a constant so it can not admit invariant continuous sections in \(\partial {\mathbb D}\).

  13. It is easy to see that it actually coincides with \(u\), but this will not play a role here.

  14. Note that any \({\mathrm {SL}}(2,{\mathbb R})\) cocycle is homotopic to a \({\mathrm {SO}}(2,{\mathbb R})\)-valued cocycle (since \({\mathrm {SO}}(2,{\mathbb R})\) is a deformation retract of \({\mathrm {SL}}(2,{\mathbb R})\)), thus of the form \(R_{\phi (x)}\) for some continuous function \(\phi :{\mathbb R}^d/{\mathbb Z}^d \rightarrow {\mathbb R}/{\mathbb Z}\). Any such map \(\phi \) is of course homotopic to a unique map of the form \(x \mapsto \langle l,x \rangle \) for some \(l \in {\mathbb Z}^d\).

  15. This is most easily seen by working with \(C^r\) invariant sections (which arise from and give rise to a conjugacy to rotations in the usual way). If \(m \in C^r({\mathbb R}^d/{\mathbb Z}^d,{\mathbb D})\) satisfies \(\mathring{{A}}_n(x) \cdot m(x)=m(x+n\alpha )\), let \(m_j(x)=\mathring{{A}}_j(x-j\alpha ) \cdot m(x-j\alpha )\). Then \(m_{j+n}=m_j\) and \(\mathring{{A}}(x) \cdot m_j(x)=m_{j+1}(x)\). For each \(x \in {\mathbb R}^d/{\mathbb Z}^d\), let \(m_*(x)\) minimizes the sum of the squares of the hyperbolic distances (in \({\mathbb D}\)) to \((m_j(x))_{j=0}^{n-1}\): this is a well defined \(C^r\) function of \(x\) by strict convexity. Then \(\mathring{{A}}(x) \cdot m_*(x)=m_*(x+\alpha )\).

  16. In fact, as the proof will show the convergence holds uniformly on any compact subsets of larger and larger complex strips.

  17. Although we do not need this fact, it is easy to see that \({\mathcal B}\) is continuous in each \({\mathcal M}_K\), \(1 \le K<\infty \).

References

  1. Amor, S.H.: Hölder continuity of the rotation number for quasi-periodic co-cycles in \({\rm SL}(2,\mathbb{R})\). Commun. Math. Phys. 287, 565–588 (2009)

    Article  MATH  Google Scholar 

  2. Avila, A.: Local distribution of eigenvalues in the absolutely continuous spectrum of ergodic Schrödinger operators: a renormalization approach (in preparation)

  3. Avila, A.: Global theory of one-frequency operators II: acriticality and finiteness of phase transitions for typical potentials. http://www.impa.br/~avila/ (preprint)

  4. Avila, A., Bochi, J.: A formula with some applications to the theory of Lyapunov exponents. Isr. J. Math. 131, 125–137 (2002)

    Article  MATH  MathSciNet  Google Scholar 

  5. Avila, A., Fayad, B., Krikorian, R.: A KAM scheme for \({\rm SL}(2,\mathbb{R})\) cocycles with Liouvillean frequencies. Geom. Funct. Anal. 21, 1001–1019 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  6. Avila, A., Jitomirskaya, S.: Almost localization and almost reducibility. J. Eur. Math. Soc. 12, 93–131 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  7. Avila, A., Krikorian, R.: Reducibility and non-uniform hyperbolicity for one-dimensional quasiperiodic Schrödinger cocycles. Ann. Math. 164, 911–940 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  8. Avila, A., Last, Y., Simon, B.: Bulk universality and clock spacing of zeros for ergodic Jacobi matrices with ac spectrum. Anal. PDE 3, 81–118 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  9. Benedicks, M., Carleson, L.: The dynamics of the Hénon map. Ann. Math. (2) 133(1), 73–169 (1991)

    Article  MATH  MathSciNet  Google Scholar 

  10. Bjerklöv, K.: Positive Lyapunov exponent and minimality for a class of one-dimensional quasi-periodic Schrödinger equations. Ergodic Theory Dyn. Syst. 25(4), 1015–1045 (2005)

    Article  MATH  Google Scholar 

  11. Bjerklöv, K.: Dynamics of the quasi-periodic Schrödinger cocycle at the lowest energy in the spectrum. Commun. Math. Phys. 272, 397–442 (2007)

    Article  MATH  Google Scholar 

  12. Bjerklöv, K., Johnson, R.: Minimal subsets of projective flows. Discrete Contin. Dyn. Syst. Ser. B 9(3–4), 493–516 (2008)

    MATH  MathSciNet  Google Scholar 

  13. Bourgain, J.: Positivity and continuity of the Lyapounov exponent for shifts on \(\mathbb{T}^d\) with arbitrary frequency vector and real analytic potential. J. Anal. Math. 96, 313–355 (2005)

    Article  MATH  MathSciNet  Google Scholar 

  14. Bourgain, J., Jitomirskaya, S.: Continuity of the Lyapunov exponent for quasiperiodic operators with analytic potential. Dedicated to David Ruelle and Yasha Sinai on the occasion of their 65th birthdays. J. Stat. Phys. 108(5–6), 1203–1218 (2002)

    Article  MATH  MathSciNet  Google Scholar 

  15. Bourgain, J., Jitomirskaya, S.: Absolutely continuous spectrum for 1D quasiperiodic operators. Invent. Math. 148(3), 453–463 (2002)

    Article  MATH  MathSciNet  Google Scholar 

  16. Concini, D., Johnson, R.A.: The algebraic–geometric AKNS potentials. Ergodic Theory Dyn. Syst. 7(1), 1–24 (1987)

    Article  MATH  Google Scholar 

  17. Deift, P., Simon, B.: Almost periodic Schrödinger operators, III. The absolutely continuous spectrum in one dimension. Commun. Math. Phys. 90, 389–411 (1983)

    Article  MATH  MathSciNet  Google Scholar 

  18. Dinaburg, E.I., Sinai, Ja.G.: The one-dimensional Schrödinger equation with quasiperiodic potential. Funkcional. Anal. i Prilozen. 9(4), 8–21 (1975)

  19. Douady, A., Earle, C.J.: Conformally natural extension of homeomorphisms of the circle. Acta Math. 157(1–2), 23–48 (1986)

    Article  MATH  MathSciNet  Google Scholar 

  20. Eliasson, L.H.: Floquet solutions for the \(1\)-dimensional quasi-periodic Schrödinger equation. Commun. Math. Phys. 146(3), 447–482 (1992)

    Article  MATH  MathSciNet  Google Scholar 

  21. Herman, M.-R.: Une méthode pour minorer les exposants de Lyapounov et quelques exemples montrant le caractère local d’un théorème d’Arnold et de Moser sur le tore de dimension \(2\). Comment. Math. Helv. 58(3), 453–502 (1983)

    Article  MATH  MathSciNet  Google Scholar 

  22. Hirsch, M.W., Pugh, C.C., Shub, M.: Invariant manifolds. Lecture Notes in Mathematics, vol. 583. Springer, Berlin (1977)

  23. Hou, X., You, J.: Almost reducibility and non-perturbative reducibility of quasi-periodic linear systems. Invent. Math. 190(1), 209–260 (2012)

    Article  MATH  MathSciNet  Google Scholar 

  24. Hunt, B.R., Sauer, T., Yorke, J.A.: Prevalence: a translation-invariant “almost every” on infinite-dimensional spaces. Bull. Am. Math. Soc. (N.S.) 27(2), 217–238 (1992)

  25. Jitomirskaya, J., Kachkovskiy, I.: Anderson localization for discrete 1D quasiperiodic operators with piecewise monotonic sampling functions (in preparation)

  26. Johnson, R.A.: Ergodic theory and linear differential equations. J. Differ. Equ. 28(1), 23–34 (1978)

    Article  MATH  Google Scholar 

  27. Johnson, R.: Two-dimensional, almost periodic linear systems with proximal and recurrent behavior. Proc. Am. Math. Soc. 82, 417–422 (1981)

    Article  MATH  Google Scholar 

  28. Journé, J.-L.: A regularity lemma for functions of several variables. Rev. Mat. Iberoamericana 4, 187–193 (1988)

    Article  MathSciNet  Google Scholar 

  29. Khanin, K.M., Sinai, YaG: A new proof of M. Herman’s theorem. Commun. Math. Phys. 112(1), 89–101 (1987)

    Article  MATH  MathSciNet  Google Scholar 

  30. Kim, J.-W., Kim, S.-Y., Hunt, B., Ott, E.: Fractal properties of robust strange nonchaotic attractors in maps of two or more dimensions. Phys. Rev. E 67, 036211 (2003)

    Article  MathSciNet  Google Scholar 

  31. Kotani, S.: Ljapunov indices determine absolutely continuous spectra of stationary random one-dimensional Schrödinger operators. Stochastic analysis (Katata/Kyoto, 1982). North-Holland Math. Library, vol. 32, pp. 225–247. North-Holland, Amsterdam (1984)

  32. Krikorian, R.: Global density of reducible quasi-periodic cocycles on \(\mathbb{T}^1\times SU(2)\),-. Ann. Math. 154, 269–326 (2001)

    Article  MATH  MathSciNet  Google Scholar 

  33. Krikorian, R.: Reducibility, differentiable rigidity and Lyapunov exponents for quasi-periodic cocycles on \(\mathbb{T} \times {\rm SL}(2,\mathbb{R})\). http://www.arXiv.org (preprint)

  34. Lyubich, M.: Teichmuller space of Fibonacci maps. Preprint IMS Stony Brook 1993/12

  35. Simon, B.: Kotani theory for one-dimensional stochastic Jacobi matrices. Commun. Math. Phys. 89(2), 227–234 (1983)

    Article  MATH  Google Scholar 

  36. Young, L.-S.: Lyapunov exponents for some quasi-periodic cocycles. Ergodic Theory Dyn. Syst. 17(2), 483–504 (1997)

    Article  MATH  Google Scholar 

  37. Wang, Y., You, J.: Examples of discontinuity of Lyapunov exponent in smooth quasi-periodic cocycles. Duke Math. J. (2014, to appear)

Download references

Acknowledgments

R.K would like to thank the hospitality of the IMPA and the support of the Institut Universitaire de France. This research was partially conducted during the period A.A. was a Clay Research Fellow. We are grateful to Zhenghe Zhang for detailed comments about a preliminary version of this paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Raphaël Krikorian.

Appendices

Appendix A: Conformal barycenter

Let \({\mathcal M}\) be the set of probability measures on \({\mathbb D}\), and for \(\mu \in {\mathcal M}\), let \(\Phi (\mu )=\int _{\mathbb D}\frac{1}{1-|z|^2} d\mu (z)\). For \(w \in {\mathbb D}\), let \(\Phi _w(\mu )=\Phi (\mu ')\) where \(\mu '\) is the pushforward of \(\mu \) by some Moebius transformation of \({\mathbb D}\) taking \(w\) to \(0\). Notice that if \(\Phi (\mu )<\infty \) then \(\Phi _w(\mu )<\infty \) for every \(w\). For every \(1 \le K<\infty \), let \({\mathcal M}_K=\{\mu \in {\mathcal M}, \Phi (\mu ) \le K\}\), and let \({\mathcal M}_\infty =\cup {\mathcal M}_K\). Notice that \({\mathcal M}_K\) is compact in the weak-* topology for every \(K<\infty \).

The next proposition can be proved using the conformal barycenter of Douady–Earle [19]. The construction is sufficiently simple for us to give the details here.

Proposition 5.1

There exists a Borelian function \({\mathcal B}:{\mathcal M}\rightarrow {\mathbb D}\), equivariant with respect to Möebius transformations of \({\mathbb D}\) and such that \(\Phi (\delta _{{\mathcal B}(\mu )}) \le \Phi (\mu )\).

Proof

Following an idea of Yoccoz, let us define a pairing \({\mathbb D}\times {\mathbb D}\rightarrow {\mathbb D}\) by setting \(z * w\) as the midpoint of the hyperbolic geodesic passing through \(z\) and \(w\) if \(z \ne w\), and \(z * z=z\). This pairing is continuous and equivariant, and we have

$$\begin{aligned} u_s(z,w) \equiv \Phi _s \left( \frac{\delta _z+\delta _w}{2} \right) -\Phi _s(\delta _{z*w})\ge 0, \end{aligned}$$
(5.1)

with equality if and only if \(z=w\). Notice that

$$\begin{aligned} u_s(z,s)=(2 \Phi _s(\delta _{z*s})-1) (\Phi _s(\delta _{z*s})-1) \ge \Phi _s(\delta _{z*s})-1. \end{aligned}$$
(5.2)

Extend the pairing \(*\) to \({\mathcal M}\times {\mathcal M}\rightarrow {\mathcal M}\) linearly. Thus

$$\begin{aligned} \mu * \nu =\int _{{\mathbb D}\times {\mathbb D}} \delta _{z * w} d\mu (z) d\nu (w). \end{aligned}$$
(5.3)

If \(\mu ,\nu \in {\mathcal M}_\infty \) then

$$\begin{aligned} u_s(\mu ,\nu ) \equiv \Phi _s\left( \frac{1}{2}(\mu +\nu )\right) \!-\!\Phi _s(\mu *\nu )\!=\!\int _{{\mathbb D}\times {\mathbb D}} u_s(z,w) d\mu (z)d\nu (w) \ge 0,\nonumber \\ \end{aligned}$$
(5.4)

with equality if and only if \(\mu =\nu \) is a Dirac mass. Notice that \(u_s:{\mathcal M}\times {\mathcal M}\rightarrow [0,\infty ]\) is lower semicontinuous, so if \(\mu _k \rightarrow \mu \) and \(u_s(\mu _k,\mu _k) \rightarrow 0\) then \(\mu \) is a Dirac mass. If \(\mu _k \rightarrow \delta _s\) we have

$$\begin{aligned} \limsup _{k \rightarrow \infty } u_s(\mu _k,\mu _k) \ge \limsup _{k \rightarrow \infty } u_s(\mu _k,\delta _s) \ge \limsup _{k \rightarrow \infty } \int _{\mathbb D}\Phi _s(\delta _{z*s})-1 d\mu _k(z),\nonumber \\ \end{aligned}$$
(5.5)

and in particular if additionally \(\lim u_s(\mu _k,\mu _k)=0\) then \(\lim \Phi _s(\mu _k*\delta _s)=1\).

Given \(\mu \in {\mathcal M}\), define \(\mu ^{(k)}\) inductively by \(\mu ^{(0)}=\mu \) and \(\mu ^k=\mu ^{(k-1)}*\mu ^{(k-1)}\). If \(\mu \in {\mathcal M}_\infty \) then \(\mu ^{(k)} \in {\mathcal M}_\infty \) and we have \(\Phi (\mu ^{(k+1)})=\Phi (\mu ^{(k)})-u_{s}(\mu ^{(k)},\mu ^{(k)})\). Thus \(u_s(\mu ^{(k)},\mu ^{(k)}) \rightarrow 0\), and any limit of \(\mu ^{(k)}\) (which exists by compactness) must be a Dirac mass. Moreover, if \(\mu ^{(n_k)} \rightarrow \delta _s\) then \(\Phi _s(\mu ^{(n_k)}*\delta _s) \rightarrow 1\), so \(\Phi _s(\mu ^{(n)}*\delta _s) \rightarrow 1\) as well and \(\delta _s\) must be the unique limit of \(\mu ^{(n)}\). Now we can set \({\mathcal B}(\mu )=s\), which is clearly Borelian.Footnote 17 \(\square \)

The estimates above allow us to obtain compactness result for invariant sections of cocycles. For instance, we have the following.

Proposition 5.2

Let \(f_k:X \rightarrow X\) be a sequence of homeomorphisms of \(X\) preserving a probability measure \(\mu \) and converging uniformly to a homeomorphism \(f:X \rightarrow X\). Let \(A_k \in C^0(X,\Upsilon )\) be a sequence converging to \(A \in C^0(X,\Upsilon )\). Assume there exists measurable \(m_k:X \rightarrow {\mathbb D}\) satisfying \(A_k(x) \cdot m_k(x)=m_k(f_k(x))\), such that

$$\begin{aligned} H \equiv \liminf _{K,k \rightarrow \infty } \int _X \min \left\{ K,\frac{1}{1-|m_k(x)|^2} \right\} d\mu (x)<\infty . \end{aligned}$$
(5.6)

Then there exists a measurable \(m:X \rightarrow {\mathbb D}\) such that \(A(x) \cdot m(x)=m(f(x))\) and \(\int _X \frac{1}{1-|m(x)|^2} d\mu (x) \le H\).

Proof

Let \(X_{K,k}=\{x \in X,\, \frac{1}{1-|m_k(x)|^2}<K\}\), and let

\(\nu _{K,k}=\int _{X_{K,k}} \delta _{m_k(x)} d\mu (x)\). Let \(\nu \) be any limit of \(\nu _{K,k}\) along a sequence \(K_i \rightarrow \infty \), \(k_i \rightarrow \infty \) attaining the \(\liminf \) in (5.6). Then \(\nu \) is a probability measure which projects onto \(\mu \) and satisfies \(\int _{X \times {\mathbb D}} \frac{1}{1-|z|^2} d\nu (x,z) \le H\). Let \(\nu _x\), \(x \in {\mathbb R}/{\mathbb Z}\) be a disintegration of \(\nu \): \(\int _{X \times {\mathbb D}} \phi (x,z) d\nu (x,z)=\int _X (\int _{\mathbb D}\phi (x,z) d\nu _x(z)) d\mu (x)\). Then \(\nu _{f(x)}\) is the pushforward of \(\nu _x\) by \(w \mapsto \mathring{{A}}(x) \cdot w\), and \(\nu _x \in {\mathcal M}_\infty \) for \(\mu \)-almost every \(x\). Let \(m(x)={\mathcal B}(\nu _x)\). Then \(m(f(x))=\mathring{{A}}(x) \cdot m(x)\) and we have \(\int \frac{1}{1-|m(x)|^2} d\mu (x) \le \int \int \frac{1}{1-|z|^2} d\nu _x(z) \ d\mu (x) \le H\). \(\square \)

Appendix B: Transitivity of the projective action

We follow the notation of Sect. 3.3. Our goal is to show the transitivity of the projective action of multidimensional quasiperiodic cocycles which are not homotopic to the identity. The one-dimensional case was considered in [30], and in fact the topological ideas that make the one-dimensional argument work are easily implemented in the multidimensional case as well. Let \(f_\alpha :x \mapsto x+\alpha \) be an ergodic translation in \({\mathbb R}^d/{\mathbb Z}^d\).

Proposition 6.1

Let \(A \in C^0({\mathbb R}^d/{\mathbb Z}^d,{\mathrm {SL}}(2,{\mathbb R}))\) be non-homotopic to the identity. If \(f_\alpha :x \mapsto x+\alpha \) is an ergodic translation in \({\mathbb R}^d/{\mathbb Z}^d\) then \((f_\alpha ,A)\) is transitive on \({\mathbb R}^d/{\mathbb Z}^d \times \partial {\mathbb D}\).

Proof

Up to change of coordinate, we may assume that \(x_1 \mapsto A(x_1,\dots ,x_d)\) has positive degree \(\deg \ge 1\). To prove transitivity, it is enough to show that for any open set \(U \subset {\mathbb R}^d/{\mathbb Z}^d \times \partial {\mathbb D}\), the set \(\cup _{k \ge 0} (f_\alpha ,A)^k(U)\) is dense in \({\mathbb R}^d/{\mathbb Z}^d \times \partial {\mathbb D}\).

We will actually show a stronger statement. Let \(\Pi _1:{\mathbb R}^d/{\mathbb Z}^d \times \partial {\mathbb D}\rightarrow {\mathbb R}/{\mathbb Z}\), \(\Pi _2:{\mathbb R}^d/{\mathbb Z}^d \times \partial {\mathbb D}\rightarrow {\mathbb R}^{d-1}/{\mathbb Z}^{d-1}\) and \(\Pi _3:{\mathbb R}^d/{\mathbb Z}^d \times \partial {\mathbb D}\rightarrow \partial {\mathbb D}\) be given by \(\Pi _1(x_1,\dots ,x_d,z)=x_1\), \(\Pi _2(x_1,\dots ,x_d,z)=(x_2,\dots ,x_d)\) and \(\Pi _3(x_1,\dots ,x_d,z)=z\).

Let \(0<\epsilon <1/10\). Let us say that a point \(x \in {\mathbb R}^d/{\mathbb Z}^d\) is \(\epsilon \)-short if there exists a sequence of paths \(\gamma _n:[0,1] \rightarrow {\mathbb R}^d/{\mathbb Z}^d \times \partial {\mathbb D}\) such that \(\Pi _1 \circ \gamma _n(t)\) converges uniformly to \(\Pi _1(x)+\epsilon t\), \(\Pi _2 \circ \gamma _n(t)\) converges uniformly to \(\Pi _2(x)\), and \(\Pi _3((f_\alpha ,A)^k(\gamma _n(t)))\) has (algebraic) length at most \(2 \pi -1/10\) for every \(k \ge 0\). It is clear that the set of \(\epsilon \)-short \(x\) is forward invariant and closed, so for each \(\epsilon \), either every point is \(\epsilon \)-short or no point is \(\epsilon \)-short.

If for every \(\epsilon >0\) there is no point which is \(\epsilon \)-short, then for any \((x,z)=(x_1,\dots ,x_d,z)\) and for every \(\delta >0\), there exists \(k \ge 0\) such that, letting \(J_\delta (x,z)=[x_1,x_1+\delta ] \times \{(x_2,\dots ,x_d,z)\}\), we have \(|\Pi _3(f_\alpha ,A)^k(J_\delta (x,z))|>2 \pi -\delta \). It follows that for every \(\delta _0\), the closure of \(\cup _{k \ge 0} (f_\alpha ,A)^k(J_{\delta _0}(x,z))\) contains some circle \(\{y\} \times \partial {\mathbb D}\). Since this set is also forward invariant, it must contain also the circles \(\{y+l \alpha \} \times \partial {\mathbb D}\) for every \(l \ge 0\), and hence, by minimality of \(x \mapsto x+\alpha \), the whole \({\mathbb R}^d/{\mathbb Z}^d \times \partial {\mathbb D}\). Since \((x,z)\) and \(\delta _0>0\) are arbitrary, transitivity follows.

Assume now that there exists \(\epsilon >0\) such that every point is \(\epsilon \)-short. We may assume that \(\epsilon =\frac{1}{k}\) for some \(k \ge 2\). Let \(\gamma _{n,i}\), \(1 \le i \le k\), \(n \ge 1\), be the sequences of paths associated to \((\frac{i-1}{k},0,\dots ,0)\). We define a sequence of paths \(\tilde{\gamma }_n:{\mathbb R}/{\mathbb Z}\rightarrow {\mathbb R}^d/{\mathbb Z}^d \times \partial {\mathbb D}\) so that

  1. (1)

    \(\tilde{\gamma }_n|[(i-1)/k,(3i-2)/3k]\) is given by \(\tilde{\gamma }_n(t)=\gamma _{n,i}(3kt-3i+3)\),

  2. (2)

    \(\tilde{\gamma }_n|[(3i-2)/3k,(3i-1)/3k]\) is such that \(\Pi _1 \circ \tilde{\gamma }_n\) and \(\Pi _2 \circ \tilde{\gamma }_n\) are constant and \(\Pi _3 \circ \tilde{\gamma }_n\) is a homeomorphism,

  3. (3)

    the diameter of the image of \(\tilde{\gamma }_n|[(3i-1)/3k,i/k]\) converges to \(0\).

One readily checks that these properties imply that for every \(l \ge 0\), if \(n\) is sufficiently large, then \(\Pi _1 \circ (f_\alpha ,A)^l \circ \tilde{\gamma }_n\) has topological degree \(1\), \(\Pi _2 \circ (f_\alpha ,A)^l \circ \tilde{\gamma }_n\) is homotopic to a constant and \(\Pi _3 \circ (f_\alpha ,A)^l \circ \tilde{\gamma }_n\) has topological degree \(\deg _{l,n}\) satisfying \(|\deg _{l,n}| \le 2k-1\). But \(\deg _{l+1,n}=\deg _{l,n}+\deg \ge \deg _{l,n}+1\) for every \(l\) and \(n\), since \(A\) is not homotopic to the identity. Thus for large \(n\) we have both \(\deg _{4k,n}-\deg _{0,n} \ge 4k\) and \(|\deg _{4k,n}|,|\deg _{0,n}| \le 2k-1\), a contradiction. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Avila, A., Krikorian, R. Monotonic cocycles. Invent. math. 202, 271–331 (2015). https://doi.org/10.1007/s00222-014-0572-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00222-014-0572-6

Navigation