Skip to main content

Advertisement

Log in

Exponential decay of correlations for finite horizon Sinai billiard flows

  • Published:
Inventiones mathematicae Aims and scope

Abstract

We prove exponential decay of correlations for the billiard flow associated with a two-dimensional finite horizon Lorentz Gas (i.e., the Sinai billiard flow with finite horizon). Along the way, we describe the spectrum of the generator of the corresponding semi-group \(\mathcal {L}_t\) of transfer operators, i.e., the resonances of the Sinai billiard flow, on a suitable Banach space of anisotropic distributions.

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. Unfortunately, the spaces introduced e.g. in [6, 22, 24, 25] for Anosov diffeomorphisms or flows do not work well in the presence of discontinuities.

  2. It suffices to take \(\eta < L_0^5\), where \(L_0\) is from Lemma 3.8.

  3. Speed estimates can be found in [30].

  4. In view of defining \(\mathcal {C}^2\) stable curves which can touch the scatterers on their endpoints, \(\mathcal {C}^2\) test functions (when considering \(\mathcal {C}^\ell \) functions on \(\Omega _0\), we mean differentiable in the sense of Whitney, viewing \(\Omega _0\) as a subset of the three-torus).

  5. In particular strong continuity holds for the transfer operator \(\mathcal {L}_t\) associated to \(\Phi _{t}\) for \(t>0\), see Lemma 4.6.

  6. For integer \(\ell \), we set \(|f|_{\mathcal {C}^\ell (\Omega _0)}=\sum _{k=0}^\ell \sup _{x \in \Omega _0}\max _{|\mathbf {k}|=k}|\partial ^{\mathbf {k}} f(x)|\), where \(\partial ^{\mathbf {k}}\) is the partial derivative associated to \(\mathbf {k}=(k_1, k_2,k_3)\in \mathbb { Z}_+^3\) and \(|\mathbf {k}|=k_1+k_2+k_3\), using the natural chart on the torus.

  7. If \(\kappa \in (0,1)\), we set \(|f|_{\mathcal {C}^\kappa (\Omega _0)}=\sup _{x\in \Omega _0}|f(x)| + \sup _{x\ne y\in \Omega _0} \frac{|f(x)-f(y)|}{d(x,y)^{\kappa }}\), for the distance on \(\Omega _0\) induced by \(\mathbb {T}^3\).

  8. In the language of [5] the transversality condition is relative to the lateral sides of the flow boxes.

  9. Note however that the Sobolev nature of the norms in [5] made the corresponding estimates much easier than in the present work.

  10. Our curves W are diffeomorphic to a bounded open interval, in particular relatively open at endpoints. However, we require the \(\mathcal {C}^r\) (\(r={\alpha },\beta ,1,2\)) norms of functions supported on W to be bounded in the sense of Whitney on the closed interval.

  11. We use a bold upright symbol \({{\mathrm{~\mathbf {d}~}}}\) to denote the exterior derivative. This should not be confused with the d used for coordinates in a vector (dr,\(d\varphi \)), (\(d\eta ,d\xi ,d\omega \)), adopted from [15].

  12. This holds for example if \(W_1\cap W_2\ne \emptyset \). It does not hold if \(W_2=\Phi _t(W_1)\).

  13. This space \(\mathcal {C}^\beta (W)\) coincides with the “little Hölder space” noted \(b_{\infty ,\infty }^\beta \) in [39, Prop. 2.1.2, Def. 2.1.3.1] obtained by taking the closure of \(\mathcal {C}^\infty \) in the \(| \cdot |_{\mathcal {C}^\beta (W)}\) norm, it is strictly smaller than the set of all Hölder continuous functions with exponent \(\beta \), but contains all Hölder continuous functions with exponent \(\beta ' > \beta \).

  14. The value 1 / 3 is determined by the choice (2.7) of homogeneity layers, see Lemma 3.5. Replacing \(k^2\) by \(k^\chi \) for \(\chi >1\) replaces the bound \({\alpha }\le 1/3\) by \({\alpha }\le 1/(\chi +1)\).

  15. In view of (4.5), it will be natural to further restrict to those \(\beta \) which satisfy in addition \(\beta \le 1- {1/q}\).

  16. In previous works, \(d(\psi _1,\psi _2) =0\) was replaced by \(d_\beta (\psi _1,\psi _2) \le \varepsilon \) where \(\beta > 0\) and the distance used the \(\mathcal {C}^\beta \) instead of the \(\mathcal {C}^0\) norm. The two formulations are equivalent by the triangle inequality, using the strong stable norm, and since \(d(\psi _1,\psi _2)=0\) implies \(d_\beta (\psi _1,\psi _2)=0\).

  17. The requirement \(\Vert f\Vert _0<\infty \) essentially implies that \(f\) is in some kind of Sobolev space with positive exponent in the flow direction, thus continuous almost everywhere on the quotient \(\Omega \). We shall neither prove nor use this.

  18. Whenever we partition a curve into finitely or countably many subcurves, we drop the (at most countably many) division points.

  19. The factor 1 / 2 here is intrinsic and not related to the quadratic decay choice in (2.7).

  20. In Step 1 of Sect. 6, we shall apply a growth lemma [15, Theorem 5.52] directly.

  21. We write \(t_k\) for \(t_k+\tilde{\delta }(k,W_i^t)\), for simplicity.

  22. The bound \(<\sqrt{\epsilon }\) can perhaps be improved to \(<C \epsilon \) by making a special choice of \(W^u\).

  23. By construction of the family \(V_i\).

  24. We do not claim that \(I_\varepsilon \), \(N_\varepsilon \), or \(\Theta _\varepsilon ^{-1}\) are bounded uniformly in \(\varepsilon \).

  25. Defining the transfer operator on \(L^\infty \) leads to problems since an element of \(L^\infty \) is not in general well-defined on a stable curve W.

  26. Note for any curve W and any x in W, the point \(\Phi _s(x)\) only hits \(\partial \Omega _0\) at finitely many times s, recall also that \(f\in \mathcal {C}^0_\sim \).

  27. In particular, taking other values than the “Hilbert-space” choice \(q=2\) does not help here. However, for the proof of Proposition 9.1, it is crucial to let \(q\) tend to 1.

  28. The exponent \(-1/5\) comes from our choice of homogeneity layer (the length of the fiber, \(k_W^2 \rho \) must be less than the width of the strip \(k_W^{-3}\)); this choice will impact the exponents throughout this section, without further notice.

  29. The analogous time reversed counterpart from [15] would be \(|\ln J\mathbf {h}_V(x)| \le C (d(x,\tilde{x})^{1/3} + \phi _V(x,\tilde{x}))\). Here, we leverage the fact that we require the transverse foliation to survive for only n steps (not for all time) in addition to the fact that the surviving curves must be long in the length scale \(\rho \), to improve the exponent of \(d(x,\tilde{x})\) to 1, at the cost of having the constant depend on \(\rho \).

  30. We use implicitly here that \(T^{-j}(P^+(W))\) crosses an extended singularity line only if there exists \(\ell \) as constructed in Step 1 so that \(T^{n-j}(\ell )\) crosses a singularity line.

  31. Equation (6.23) below shows that we can get away with \(\varpi \le 1/15\), up to using a more complicated expression for the upper bound.

  32. The projection on the scatterers of Lebesgue measure is the T-invariant probability measure \(\mu _0=c \cos \varphi dr d\varphi \), where c is a normalization constant.

  33. Just like for (2.5), see the paragraph containing (8.6).

  34. In Sect. 8, we shall take \(\epsilon _0\) small enough as a function of \(n=\lceil c \ln |b| \rceil \).

  35. We do not assume here that \(f\) is supported in \(\Omega _0\), recalling (7.4).

  36. In such a way in the following we will never need to consider a curve in the the midst of a collision.

  37. When no confusion arises, to ease notation, we will identify \(\tilde{\kappa }_{x_{i,j}}(W_{A}^0)\) and \(W_{A}^0\).

  38. Some manifolds can be attributed either to a box or to an adjacent box and a choice can be made to resolve the ambiguity.

  39. The manifold \(W_B\) in the proof of Lemma 8.5 corresponds here to the central manifold of the box, which in the current coordinates reads \(\{(0,x^s,0)\}\), while \(M_A(x^s)\) corresponds to \(G(x^u_A,x^s)\).

  40. We cannot require that the leaf \((x^u,0,0)\) belongs to the foliation, but the proof of [5, Lemma  A.4] ensures that we can obtain the desired tangency.

  41. Remember that i runs from 1 to \(C_{\#}{\varepsilon }^{-3\theta }\), while j from 1 to \(C_{\#}{\varepsilon }^{\theta -1}\), so, all in all, the sum over ij consists of \(C_{\#}{\varepsilon }^{-1-2\theta }\) terms.

  42. Note that with the above choices if \(\alpha =\frac{1}{3}\), then we have the rather small value \(\gamma _{{ Do}}=\frac{1}{1222}\). Working more it is certainly possible to obtain a better estimate but it remains unclear what the optimal value is.

  43. In [24] another approach was used to address this issue, involving an auxiliary norm \(\Vert \cdot \Vert ^*\) based on exact stable leaves (bypassing the neutral norm, using the resolvent). See [24, Lemmas 7.4, 7.8, just after (7.4)].

  44. We use here that the constant C in (5.8) is uniform in a and b, also for \(\Vert \cdot \Vert _*\).

  45. The proof gives a constant \(C_{{ Do}}\) much larger than \(\gamma _{{ Do}}\), and we may safely assume that \(\gamma _{{ Do}}<1\).

  46. We may take \(\gamma ={\alpha }-\beta \), which is smaller than (and close to) \(1/3<{1/q}\), given our other choices, so that \(\sigma \) is larger than (and close to) 3.

  47. Note that M exists and is finite by the first part of the lemma, (8.41).

  48. Remember that the \(J_j\) are all disjoint and \(\cup _j J_j =W\) and \(|W|\le L_0\).

References

  1. Araújo, V., Melbourne, I.: Exponential decay of correlations for nonuniformly hyperbolic flows with a \(C^{1+\alpha }\) stable foliation, including the classical Lorenz attractor. Ann. Henri Poincaré 17, 2975–3004 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  2. Avila, A., Gouëzel, S., Yoccoz, J.-C.: Exponential mixing for the Teichmüller flow. Pub. Math. IHÉS 104, 143–211 (2006)

    Article  MATH  Google Scholar 

  3. Arnold, V.: Mathematical methods of classical mechanics. In: Graduate Texts in Mathematics, vol. 60. Springer-Verlag, Berlin–New York (1978, 1989)

  4. Baladi, V., Gouëzel, S.: Banach spaces for piecewise cone hyperbolic maps. J. Mod. Dyn. 4, 91–137 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  5. Baladi, V., Liverani, C.: Exponential decay of correlations for piecewise cone hyperbolic contact flows. Comm. Math. Phys. 314, 689–773 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  6. Baladi, V., Tsujii, M.: Anisotropic Hölder and Sobolev spaces for hyperbolic diffeomorphisms. Ann. I. Fourier 57, 127–154 (2007)

    Article  MATH  Google Scholar 

  7. Baladi, V., Tsujii, M.: Dynamical determinants and spectrum for hyperbolic diffeomorphisms. In: Probabilistic and Geometric Structures in Dynamics, Contemporary Mathematics, vol. 469, pp. 29–68. American Mathematical Society (2008)

  8. Baladi, V., Vallée, B.: Exponential decay of correlations for surface semi-flows without finite Markov partitions. Proc. Am. Math. Soc. 133, 865–874 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  9. Bunimovich, L.A., Sinai, Y.G., Chernov, N.I., Statistical properties of two-dimensional hyperbolic billiards, Uspekhi Mat. Nauk 46, 43–92 (1991); translation in Russian Math. Surveys 46, 47–106 (1991)

  10. Burns, K., Matheus, C., Masur, H., Wilkinson, A.: Rates of mixing for the Weil–Petersson geodesic flow: Exponential mixing in exceptional moduli spaces. arXiv:1605.09037, to appear GAFA

  11. Butterley, O.: A note on operator semigroups associated to chaotic flows. Ergod. Theory Dyn. Syst. 36, 1396–1408 (2016). Corrigendum 36, 1409–1410 (2016)

  12. Butterley, O., Liverani, C.: Robustly invariant sets in fibre contracting bundle flows. J. Mod. Dyn. 7, 153–208 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  13. Chernov, N.: Decay of correlations in dispersing billiards. J. Stat. Phys. 94, 513–556 (1999)

    Article  MATH  Google Scholar 

  14. Chernov, N.: A stretched exponential bound on time correlations for billiard flows. J. Stat. Phys. 127, 21–50 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  15. Chernov, N., Markarian, R., Chaotic billiards. In: Mathematical Surveys and Monographs, vol. 127 (2006)

  16. Davies, E.B.: Linear operators and their spectra. In: Cambridge Studies in Advanced Mathematics, vol. 106 (2007)

  17. Demers, M.F., Liverani, C.: Stability of statistical properties in two-dimensional piecewise hyperbolic maps. Trans. American Math. Soc. 360, 4777–4814 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  18. Demers, M.F., Zhang, H.-K.: Spectral analysis for the transfer operator for the Lorentz gas. J. Mod. Dyn. 5, 665–709 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  19. Demers, M.F., Zhang, H.-K.: A functional analytic approach to perturbations of the Lorentz gas. Comm. Math. Phys. 324, 767–830 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  20. Demers, M.F., Zhang, H.-K.: Spectral analysis of hyperbolic systems with singularities. Nonlinearity 27, 379–433 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  21. Dolgopyat, D.: On decay of correlations in Anosov flows. Ann. Math. 147, 357–390 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  22. Faure, F., Roy, N., Sjöstrand, J.: Semi-classical approach for Anosov diffeomorphisms and Ruelle resonances. Open Math. J. 1, 35–81 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  23. Friedman, B., Martin, R.F.: Behavior of the velocity autocorrelation function for the periodic Lorentz gas. Phys. D 30, 219–227 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  24. Giulietti, P., Liverani, C., Pollicott, M.: Anosov flows and dynamical zeta functions. Ann. Math. 178, 687–773 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  25. Gouëzel, S., Liverani, C.: Banach spaces adapted to Anosov systems. Ergod. Theory Dyn. Syst. 26, 189–218 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  26. Hennion, H.: Sur un théorème spectral et son application aux noyaux lipchitziens. Proc. Am. Math. Soc. 188, 627–634 (1993)

    MathSciNet  MATH  Google Scholar 

  27. Iwata, Y.: A generalized local limit theorem for mixing semi-flows. Hokkaido Math. J. 37, 215–240 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  28. Kato, T.: Perturbation theory for linear operators. Reprint of the 1980 edition. In: Classics in Mathematics. Springer-Verlag, Berlin (1995)

  29. Katok, A., Strelcyn, J.-M., Ledrappier, F., Przytycki, F.: Invariant manifolds, entropy and billiards; smooth maps with singularities. In: Lecture Notes in Mathematics, vol. 1222. Springer-Verlag, Berlin (1986)

  30. Le Borgne, S., Pène, F.: Vitesse dans le théorème limite central pour certains systèmes dynamiques quasi-hyperboliques. Bull. Soc. Math. France 133, 395–417 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  31. Liverani, C.: Decay of correlations. Ann. Math. 142, 239–301 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  32. Liverani, C.: On contact Anosov flows. Ann. Math. 159, 1275–1312 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  33. Matsuoka, H., Martin, R.F.: Long-time tails of the velocity autocorrelation functions for the triangular periodic Lorentz gas. J. Stat. Phys. 88, 81–103 (1997)

    Article  MATH  Google Scholar 

  34. Melbourne, I.: Rapid decay of correlations for nonuniformly hyperbolic flows. Trans. Am. Math. Soc. 359, 2421–2441 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  35. Melbourne, I., Nicol, M.: Large deviations for nonuniformly hyperbolic systems. Trans. Am. Math. Soc. 360, 6661–6676 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  36. Melbourne, I., Török, A.: Statistical limit theorems for suspension flows. Isr. J. Math. 144, 191–209 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  37. Pène, F.: A Berry Esseen result for the billiard transformation. Preprint (2007). https://hal.archives-ouvertes.fr/hal-01101281

  38. Ruelle, D.: A measure associated to Axiom A attractors. Am. J. Math. 98, 619–654 (1976)

    Article  MathSciNet  MATH  Google Scholar 

  39. Runst, T., Sickel, W.: Sobolev Spaces of Fractional Order, Nemytskij Operators, and Nonlinear Partial Differential Equations. Walter de Gruyter & Co., Berlin (1996)

    Book  MATH  Google Scholar 

  40. Sinai, Y.G.: On the foundations of the ergodic hypothesis for a dynamical system of statistical mechanics. Dokl. Akad. Nauk SSSR 153, 1261–1264 (Russian); translated as Soviet Math. Dokl. 4, 1818–1822 (1963)

  41. Sinai, Y.G.: Dynamical systems with elastic reflections. Ergodic properties of dispersing billiards. Uspehi Mat. Nauk 25(152), 141–192 (1970) (Russian)

  42. Sinai, Y.G.: Gibbs measures in ergodic theory. Russ. Math. Surv. 166, 21–69 (1972)

    Article  MathSciNet  MATH  Google Scholar 

  43. Tsujii, M.: Quasi-compactness of transfer operators for contact Anosov flows. Nonlinearity 23, 1495–1545 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  44. Young, L.-S.: Statistical properties of dynamical systems with some hyperbolicity. Ann. Math. 2(147), 585–650 (1998)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Mark F. Demers.

Additional information

In memoriam Nikolai Chernov 1956–2014.

CL acknowledges the partial support of the European Advanced Grant Macroscopic Laws and Dynamical Systems (MALADY) (ERC AdG 246953). MD was partially supported by NSF Grants DMS 1101572 and 1362420. Part of this work was carried out by VB and MD at the Newton Institute in 2013 (Mathematics for the Fluid Earth), during visits of VB and MD to Roma in 2013–2015, and a visit of MD to Paris in 2014, using the Action incitative ENS Spectres en dynamique, géométrie et théorie des nombres. VB is grateful to Peter Bálint for his patient explanations on billiards and thanks Sébastien Gouëzel for many key ideas. We thank I.P. Toth for several useful remarks. We are grateful to the two anonymous referees for many constructive comments.

Appendices

Appendix A. Proofs of Lemmas 6.6 and 6.7 (approximate foliation holonomy)

This section contains the proofs of Lemmas 6.6 and 6.7 (Sects. 6.2 and 6.4) on the Hölder bounds of the Jacobians of the holonomy of the fake (approximate) foliation. These bounds are instrumental to get the four-point estimate (vii) for the fake unstable foliation.

Proof of Lemma 6.6

Letting \(\mathbf {h}_{V^n_1, V^n_2}\) denote the holonomy map from \(V^n_1 = T^{-n}V_1\) to \(V^n_2 = T^{-n} V_2\), we begin by noting that

$$\begin{aligned} \begin{aligned}&\ln \frac{J\mathbf {h}_{12}( v_1(\bar{x}^s),\bar{x}^s)}{J\mathbf {h}_{12}( v_1(\bar{y}^s),\bar{y}^s)}\\&\qquad = \ln \frac{J^s_{V_1}T^{-n}( v_1(\bar{x}^s),\bar{x}^s)}{J^s_{V_2}T^{-n}(\mathbf {h}_{12}(v_1(\bar{x}^s),\bar{x}^s,))} - \ln \frac{J^s_{V_1}T^{-n}( v_1(\bar{y}^s),\bar{y}^s)}{J^s_{V_2}T^{-n}(\mathbf {h}_{12}(v_1(\bar{y}^s),\bar{y}^s))}\\&\quad + \ln \frac{J\mathbf {h}_{V^n_1, V^n_2}(x_n)}{J\mathbf {h}_{V^n_1, V^n_2}(y_n)} \\&\qquad = \sum _{j=0}^{n-1} \ln J^s_{T^{-j}V_1}T^{-1}(x_j) - \ln J^s_{T^{-j}V_2}T^{-1}(\tilde{x}_j)\\&\quad - \ln J^s_{T^{-j}V_1}T^{-1}(y_j) + \ln J^s_{T^{-j}V_2}T^{-1}(\tilde{y}_j) \\&\qquad \qquad \quad + \ln \frac{J\mathbf {h}_{V^n_1, V^n_2}(x_n)}{J\mathbf {h}_{V^n_1, V^n_2}(y_n)} , \end{aligned} \end{aligned}$$
(A.1)

where as in the proof of Lemma 6.4, \(x_j = T^{-j}(v_1(\bar{x}^s),\bar{x}^s)\), \(\tilde{x}_j = T^{-j}(\mathbf {h}_{12}(v_1(\bar{x}^s),\bar{x}^s))\), and similarly for \(y_j\) and \(\tilde{y}_j\).

We begin by estimating the difference of stable Jacobians in (A.1). The factors in each term involving the stable Jacobian are given by (6.4), and we can bound the differences by grouping together either the terms on the same stable curve (standard distortion bounds), or the terms on the same unstable curve (using Lemma 6.4). Assuming without loss of generality that \(d(x_0, \tilde{x}_0) \ge d(y_0, \tilde{y}_0)\), this yields the following bound on the sum,

$$\begin{aligned} C \sum _{j=0}^{n-1} \min \{ d(x_{j+1}, y_{j+1}) k_{j+1}^2, d(x_{j+1}, \tilde{x}_{j+1}) k_{j+1}^2 + d(x_j, \tilde{x}_j) + \phi (x_j, \tilde{x}_j) + \phi (x_{j+1}, \tilde{x}_{j+1}) \} , \end{aligned}$$
(A.2)

where we have used the uniform bound on \(J\mathbf {h}_{12}\) given by Lemma 6.4 and (6.7) to eliminate terms involving \(d(\tilde{x}_j, \tilde{y}_j)\) and \(d(y_j, \tilde{y}_j)\). We estimate each term using one of two cases.

Case 1. \(d(x_{j+1}, y_{j+1}) \le d(x_{j+1}, \tilde{x}_{j+1})\). We write,

$$\begin{aligned} d(x_{j+1},y_{j+1})&= d(x_{j+1}, y_{j+1})^{\varpi } d(x_{j+1}, y_{j+1})^{1-\varpi }\nonumber \\&\le C \left( J^s_{V_1}T^{-j-1}(x_0)\right) ^\varpi d(x_0, y_0)^\varpi d(x_{j+1}, \tilde{x}_{j+1})^{1-\varpi } . \end{aligned}$$
(A.3)

On the one hand, we have

$$\begin{aligned} d(x_{j+1} , \tilde{x}_{j+1}) \le C d(x_0, \tilde{x}_0) \left( J^u_{T^{n-j}\ell }T^{j+1}(x_j)\right) ^{-1} , \end{aligned}$$

where \(\ell \) is the curve so that \(x_0 \in T^n(\ell ) =: \bar{\gamma }\). On the other hand, by the uniform transversality of the curves \(T^{-j-1}(V_1)\) and \(T^{-j-1}(\bar{\gamma })\), we have

$$\begin{aligned} \begin{aligned}&J^s_{V_1}T^{-j-1}(x_0)J^u_{\bar{\gamma }}T^{-j-1}(x_0) = C^{\pm 1} JT^{-j-1}(x_0) = C^{\pm 1} \tfrac{\cos \varphi (x_0)}{\cos \varphi (x_{j+1})} \\&\quad \implies J^s_{V_1}T^{-j-1}(x_0) = C^{\pm 1} k_W^{-2} k_{j+1}^{2} J^u_{T^{-j-1}(\bar{\gamma })}T^{j+1}(x_{j+1}) , \end{aligned} \end{aligned}$$
(A.4)

where \(JT^{-j-1}\) is the full Jacobian of the map \(T^{-j-1}\). This estimate together with (6.3) implies

$$\begin{aligned} J^s_{V_1}T^{-j-1}(x_0) \le C k_W^{-10/3} \rho ^{-2/3} \left( J^u_{T^{-j-1}\bar{\gamma }}T^{j+1}(x_{j+1})\right) ^{5/3} . \end{aligned}$$
(A.5)

Now using this estimate together with (A.3) and again (6.3) yields,

$$\begin{aligned} d(x_{j+1},y_{j+1})k_{j+1}^2\le & {} C d(x_0,y_0)^\varpi d(x_0, \tilde{x}_0)^{1-\varpi } \frac{k_W^{-(4+10\varpi )/3}}{\rho ^{2(1+\varpi )/3}}\nonumber \\&\times \frac{1}{(J^u_{T^{-j-1}\bar{\gamma }}T^{j+1}(x_{j+1}))^{(1-8\varpi )/3}}, \end{aligned}$$
(A.6)

and this will be summable over j as long as \(\varpi < 1/8\).

Case 2. \(d(x_{j+1}, \tilde{x}_{j+1}) \le d(x_{j+1}, y_{j+1})\). In order to control the terms involving \(\phi (x_j, \tilde{x}_j)\), we use the expressions for the Jacobians given by (6.4) and (A.4) together with [15, eq. (5.27) and following] to write,

$$\begin{aligned} \begin{aligned} \phi (x_j, \tilde{x}_j)&\le C\sum _{m=1}^j \frac{d(x_{j-m}, \tilde{x}_{j-m}) }{J^s_{T^{m-j-1}V_1}T^{-m+1}(x_{j-m+1}) J^u_{T^{-j}\bar{\gamma }}T^{m-1}(x_{j}) }\\&\quad + C \frac{\phi (x_0, \tilde{x}_0)}{J^s_{V_1}T^{-j}(x_0) J^u_{T^{-j}\bar{\gamma }}T^j(x_j) } \\&\le C d(x_0, \tilde{x}_0) \sum _{m=1}^j \frac{\Lambda _0^{-m+1}}{J^u_{T^{m-j}\bar{\gamma }}T^{j-m}(x_{j-m}) J^u_{T^{-j}\bar{\gamma }}T^{m-1}(x_{j})} \\&\quad + C \frac{\Lambda _0^{-j} d(x_0, \tilde{x}_0)}{J^u_{T^{-j}\bar{\gamma }}T^j(x_j)} , \end{aligned} \end{aligned}$$
(A.7)

where in the second line we have used the assumption that \(\phi (x_0, \tilde{x}_0)\) is proportional to \(d(x_0, \tilde{x}_0)\) and that \(d(x_{j-m}, \tilde{x}_{j-m}) \le C d(x_0, \tilde{x}_0)/J^u_{T^{m-j}\bar{\gamma }}T^{j-m}(x_{j-m})\). Note that for terms with \(m \ge 2\), there is a gap between the expansion factors of the unstable Jacobians in the denominator of the sum: The Jacobian \(J^u_{T^{m-j-1}\bar{\gamma }}T(x_{j-m+1})\) is missing. We use the fact that this is proportional to \(k_{j-m}^2\) to estimate,

$$\begin{aligned} \begin{aligned}&\frac{d(x_0, \tilde{x}_0)^{9\varpi }}{J^u_{T^{m-j}\bar{\gamma }}T^{j-m}(x_{j-m}) J^u_{T^{-j}\bar{\gamma }}T^{m-1}(x_{j})}\\&\quad \le \frac{C d(x_j, \tilde{x}_j)^{9\varpi } k_{j-m}^{18\varpi }}{\left( J^u_{T^{m-j}\bar{\gamma }}T^{j-m}(x_{j-m}) J^u_{T^{-j}\bar{\gamma }}T^{m-1}(x_{j})\right) ^{1-9\varpi }} \\&\quad \le \frac{C d(x_j, \tilde{x}_j)^{9\varpi } k_W^{-12\varpi } \rho ^{-6\varpi }}{\left( J^u_{T^{m-j}\bar{\gamma }}T^{j-m}(x_{j-m})\right) ^{1-15\varpi } \left( J^u_{T^{-j}\bar{\gamma }}T^{m-1}(x_{j})\right) ^{1-9\varpi }} \\&\quad \le C d(x_j, \tilde{x}_j)^{9\varpi } k_W^{-12\varpi } \rho ^{-6\varpi } \Lambda _0^{-(j-1)(1-15\varpi )}, \end{aligned} \end{aligned}$$

where we have used (6.3) in the second line. Notice that this estimate holds for \(m=1\) as well. Putting this estimate together with (A.7) yields,

$$\begin{aligned} \phi (x_j , \tilde{x}_j) \le C d(x_0, \tilde{x}_0)^{1-9\varpi } d(x_j, \tilde{x}_j)^{9\varpi } k_W^{-12\varpi } \rho ^{-6\varpi } \Lambda _0^{-(j-1)(1-15\varpi )} . \end{aligned}$$
(A.8)

A similar estimate holds for \(\phi (x_{j+1}, \tilde{x}_{j+1})\) with \(d(x_{j+1}, \tilde{x}_{j+1})\) in place of \(d(x_j, \tilde{x}_j)\).

Now since \(d(x_j, \tilde{x}_j) = C^{\pm 1} d(x_{j+1}, \tilde{x}_{j+1}) J^u_{T^{-j-1}\bar{\gamma }}T(x_{j+1})\), we have

$$\begin{aligned} d(x_j,\tilde{x}_j) \le C d(x_{j+1}, y_{j+1}) k_j^2 \end{aligned}$$

by the assumption of this case as well. Since \(d(x_{j+1}, \tilde{x}_{j+1}) \le C d(x_j, \tilde{x}_j)\), combining similar terms in (A.2), it remains to estimate the following expression in this case,

$$\begin{aligned} \begin{aligned}&d(x_{j+1}, \tilde{x}_{j+1}) k_{j+1}^2 + d(x_j, \tilde{x}_j)\\&\qquad + d(x_0, \tilde{x}_0)^{1-9\varpi } d(x_j, \tilde{x}_j)^{9\varpi } k_W^{-12\varpi } \rho ^{-6\varpi }\Lambda _0^{-(j-1)(1-15\varpi )} \\&\quad \le d(x_{j+1}, \tilde{x}_{j+1}) k_{j+1}^2 + d(x_{j+1}, \tilde{x}_{j+1}) k_j^2\\&\qquad + \frac{d(x_0, \tilde{x}_0)^{1-9\varpi } d(x_{j+1}, \tilde{x}_{j+1})^{9\varpi } k_j^{18\varpi }}{k_W^{12\varpi } \rho ^{6\varpi } \Lambda _0^{(j-1)(1-15\varpi )}} . \end{aligned} \end{aligned}$$
(A.9)

For the first term, we use the assumption of this case to write,

$$\begin{aligned} d(x_{j+1}, \tilde{x}_{j+1})&= d(x_{j+1}, \tilde{x}_{j+1})^\varpi d(x_{j+1}, \tilde{x}_{j+1})^{1-\varpi }\\&\le C \left( J^s_{V_1}T^{-j-1}(x_0)\right) ^\varpi d(x_0, y_0)^\varpi d(x_{j+1}, \tilde{x}_{j+1})^{1-\varpi } , \end{aligned}$$

which is the same expression as in (A.3). Thus the first term in (A.9) is bounded by (A.6).

For the second term in (A.9), we use (6.3) to bound \(k_j^2\) and again (A.5) to estimate,

$$\begin{aligned} \begin{aligned} d(x_{j+1}, \tilde{x}_{j+1}) k_j^2&\le C \left( J^s_{V_1}T^{-j-1}(x_0)\right) ^{\varpi } d(x_0, y_0)^\varpi \\&\quad \times \frac{d(x_0, \tilde{x}_0)^{1-\varpi } }{\left( J^u_{T^{-j-1}\bar{\gamma }}T^{j+1}(x_{j+1})\right) ^{1-\varpi }} \frac{\left( J^u_{T^{-j}\bar{\gamma }}T^j(x_j)\right) ^{2/3}}{k_W^{4/3} \rho ^{2/3}} \\&\le C d(x_0, y_0)^\varpi d(x_0, \tilde{x}_0)^{1-\varpi }\\&\quad \times \frac{k_W^{-(4+ 10\varpi )/3}}{\rho ^{2(1+\varpi )/3}} \frac{1}{\left( J^u_{T^{-j-1}\bar{\gamma }}T^{j+1}(x_{j+1})\right) ^{(1-8\varpi )/3}}, \end{aligned} \end{aligned}$$

where in the second line we have used the fact that \(J^u_{T^{-j}\bar{\gamma }}T^j(x_j) \le C J^u_{T^{-j-1}\bar{\gamma }}T^{j+1}(x_{j+1})\) as well as (A.4) to obtain a bound equivalent to (A.6).

Finally, we estimate the third term in (A.9) following a similar strategy, using again (A.5),

$$\begin{aligned} \begin{aligned}&d(x_{j+1}, \tilde{x}_{j+1})^{9\varpi } k_j^{18\varpi } \\&\quad \le C \left( J^s_{V_1}T^{-j-1}(x_0)\right) ^{\varpi } d(x_0, y_0)^\varpi \\&\qquad \times \frac{d(x_0, \tilde{x}_0)^{8\varpi } }{\left( J^u_{T^{-j-1}\bar{\gamma }}T^{j+1}(x_{j+1})\right) ^{8\varpi }} \frac{\left( J^u_{T^{-j}\bar{\gamma }}T^j(x_j)\right) ^{6\varpi }}{k_W^{12\varpi } \rho ^{6\varpi }} \\&\quad \le C d(x_0, y_0)^\varpi d(x_0, \tilde{x}_0)^{8\varpi } \frac{k_W^{-46\varpi /3}}{\rho ^{20\varpi /3}} \frac{1}{\left( J^u_{T^{-j-1}\bar{\gamma }}T^{j+1}(x_{j+1})\right) ^{\varpi /3}} . \end{aligned} \end{aligned}$$

Putting these three estimates together in (A.9) implies that the minimum factor from (A.2) in this case is bounded by

$$\begin{aligned}&Cd(x_0,y_0)^\varpi d(x_0, \tilde{x}_0)^{1-\varpi } \Lambda _0^{-(j-1)(1-44\varpi /3)} \\&\quad \cdot \max \left\{ k_W^{-82\varpi /3} \rho ^{-38\varpi /3}, k_W^{-4/3 - 10\varpi /3} \rho ^{-2(1+\varpi )/3} \right\} , \end{aligned}$$

which will be summable over j as long as \(\varpi \le 1/15\).

Finally, using Case 1 or Case 2 as appropriate for each term appearing in (A.2) and summing over j completes the required estimate on the difference of stable Jacobians appearing in (A.1).

It remains to estimate the term involving \(J\mathbf {h}_{V^n_1, V^n_2}\), which is the holonomy of the seeding foliation \(\{ \ell _z \}_{z \in \overline{W}_i}\). Since \(\{ \ell _z \}\) is uniformly \(\mathcal {C}^2\), we have on the one hand,

$$\begin{aligned} \ln \frac{J\mathbf {h}_{V^n_1, V^n_2}(x_n)}{J\mathbf {h}_{V^n_1, V^n_2}(y_n)} \le C(d(x_n, \tilde{x}_n) + \phi (x_n, \tilde{x}_n) + d(y_n, \tilde{y}_n) + \phi (y_n, \tilde{y}_n)) , \end{aligned}$$

while on the other hand, using the fact that the curvatures of \(V^n_1\) and \(V^n_2\) are uniformly bounded,

$$\begin{aligned} \ln \frac{J\mathbf {h}_{V^n_1, V^n_2}(x_n)}{J\mathbf {h}_{V^n_1, V^n_2}(y_n)} \le C ( d(x_n ,y_n) + d(\tilde{x}_n, \tilde{y}_n)) . \end{aligned}$$

Now \(\ln \frac{J\mathbf {h}_{V^n_1, V^n_2}(x_n)}{J\mathbf {h}_{V^n_1, V^n_2}(y_n)}\) is bounded by the minimum of these two estimates, which we recognize as a simplified version of the expression (A.2) with \(j=n\). Thus this quantity satisfies the same bounds as in Cases 1 and 2 above, completing the proof of Lemma 6.6. \(\square \)

Proof of Lemma 6.7

Let \(V_0\) be the subcurve of \(P^+(W)\) corresponding to the gap; in the coordinates \((\bar{x}^u, \bar{x}^s)\), \(V_0\) is the vertical line segment identified by the interval [ab]. The boundary curves of the gap from the surviving foliation on either side as described by \(\{ (\overline{G}(\bar{x}^u, a), a) : |\bar{x}^u| \le k_W^2 \rho \}\) and \(\{ (\overline{G}(\bar{x}^u, b), b) : |\bar{x}^u| \le k_W^2 \}\). Let \(x_a = (0, a)\) and \(y_b = (0,b)\). Fix \(\bar{x}^u\) with \(|\bar{x}^u| \le k_W^2 \rho \) and define \(\tilde{x}_a = (\bar{x}^u, \overline{G}(\bar{x}^u, a))\) and \(\tilde{y}_b = (\bar{x}^u, \overline{G}(\bar{x}^u, b))\). Let \(V_1\) denote the stable curve (vertical in the \((\bar{x}^u, \bar{x}^s)\) coordinates) connecting \(\tilde{x}_a\) and \(\tilde{y}_b\).

Using similar notation to Sect. 6.3, let \(j+1\) denote the least integer \(j' \ge 1\) such that \(T^{j'}(\mathcal {S}_0)\) intersects \(V_0\). This implies in particular, that the surviving parts of the foliation immediately on either side of the gap lie in the same homogeneity strip for the first j interates of \(T^{-1}\). Define \(x_i = T^{-i}(x_a)\), \(y_i = T^{-i}(y_b)\), \(\tilde{x}_i = T^{-i}(\tilde{x}_a)\) and \(\tilde{y}_i = T^{-i}(\tilde{y}_b)\), for \(i = 0, 1, \ldots j\).

The following expressions for the quantities appearing in the slope will be useful,

$$\begin{aligned} \overline{G}(\bar{x}^u, b) - \overline{G}(\bar{x}^u, a)= & {} \int _{T^{-j}V_1} J^s_{T^{-j}V_1}T^j~dm_W, \\ b-a= & {} \int _{T^{-j}V_0} J^s_{T^{-j}V_0} T^j~dm_W. \end{aligned}$$

And similarly,

$$\begin{aligned} \partial _{\bar{x}^s} \overline{G}(\bar{x}^u, a) = \frac{J^s_{T^{-j}V_1}T^j(y_j)}{J^s_{T^{-j}V_0}T^j(x_j)} J \mathbf {h}_{-j}(x_j) , \end{aligned}$$

where \(\mathbf {h}_{-j}\) is the holonomy map from \(T^{-j}V_0\) to \(T^{-j}V_1\) along the surviving foliation on the sides of the gap. Using these expressions, we estimate,

$$\begin{aligned} \begin{aligned} \ln \left| \frac{\overline{G}(\bar{x}^u, b) - \overline{G}(\bar{x}^u, a)}{(b-a) \partial _{\bar{x}^s} \overline{G}(\bar{x}^u, a)} \right| =&\ln \frac{\frac{1}{|T^{-j}V_1|} \int _{T^{-j}V_1} J^s_{T^{-j}V_1}T^j~dm_W}{J^s_{T^{-j}V_1}T^j(y_j)} - \ln J \mathbf {h}_{-j}(x_j) \\&\; \; + \ln \frac{J^s_{T^{-j}V_0}T^j(x_j)}{\frac{1}{|T^{-j}V_0|} \int _{T^{-j}V_0} J^s_{T^{-j}V_0}T^j~dm_W} + \ln \frac{|T^{-j}V_0|}{|T^{-j}V_1|} . \end{aligned} \end{aligned}$$
(A.10)

Since \(J^s_{T^{-j}V_1}T^j\) is continuous on \(T^{-j}V_1\), the average value of the function is equal to the function evaluated at some point on \(T^{-j}V_1\). Thus by standard distortion bounds along stable curves, the first term on the right-hand side of (A.10) is bounded by \(C_d |a-b|^{1/3}\). A similar bound holds for the third term above.

In order to estimate \(J\mathbf {h}_{-j}(x_j)\), we note that this is part of the surviving foliation that originated at time \(-n\). Applying Lemma 6.4 from time \(-j\) to time \(-n\), there exists \(C>0\) such that

$$\begin{aligned} \ln J\mathbf {h}_{-j}(x_j) \le C \left( d(x_j, \tilde{x}_j) \rho ^{-2/3} k_{W}^{-4/3} + \phi (x_j, \tilde{x}_j)\right) , \end{aligned}$$

where \(\phi (x_j, \tilde{x}_j)\) represents the angle formed by the tangent vectors to \(T^{-j}V_0\) and \(T^{-j}V_1\) at \(x_j\) and \(\tilde{x}_j\), respectively.

For the first term above, \(d(x_j, \tilde{x}_j )\le C d(x_0 , \tilde{x}_0)/ J^u_{T^{-j}\bar{\gamma }}T^j(x_j)\), where \(\bar{\gamma }\) denotes the element of the surviving foliation containing \(x_a\). For the second term, we use (A.7) to write,

$$\begin{aligned} \phi (x_j, \tilde{x}_j)\le & {} C d(x_0, \tilde{x}_0) \sum _{m=1}^{j} \frac{\Lambda _0^{-m+1}}{J^u_{T^{m-j}\bar{\gamma }}T^{j-m}(x_{j-m}) J^u_{T^{-j}\bar{\gamma }}T^{m-1}(x_j)}\\&+ \,C \frac{\Lambda _0^{-j} d(x_0, \tilde{x}_0)}{J^u_{T^{-j}\bar{\gamma }}T^j(x_j)} , \end{aligned}$$

where we have used the fact that \(\phi (x_0, \tilde{x}_0) \le C d(x_0, \tilde{x}_0)\) since both \(V_0\) and \(V_1\) are vertical segments and the unstable curves in the surviving foliation have uniformly bounded curvature. As in the proof of Lemma 6.6, there is a gap in the product of unstable Jacobians; the factor \(J^u_{T^{m-j}\bar{\gamma }} T(x_{j-m+1})\) is missing. This is proportional to \(k^2_{j-m}\), but due to (6.3), we do not ask for a full power of this factor; rather, we estimate,

$$\begin{aligned} \begin{aligned}&\frac{1}{J^u_{T^{m-j}\bar{\gamma }}T^{j-m}(x_{j-m})J^u_{T^{-j}\bar{\gamma }}T^{m-1}(x_j)}\\&\quad \le \frac{C k_{j-m}^{6/5}}{J^u_{T^{m-j}\bar{\gamma }}T^{j-m}(x_{j-m}) \left( J^u_{T^{m-j}\bar{\gamma }} T(x_{j-m+1})\right) ^{3/5} J^u_{T^{-j}\bar{\gamma }}T^{m-1}(x_j)} \\&\quad \le C \frac{k_W^{4/5} \rho ^{-2/5}}{\left( J^u_{T^{-j}\bar{\gamma }} T^j(x_j)\right) ^{3/5}} . \end{aligned} \end{aligned}$$

This, together with the above estimates implies,

$$\begin{aligned} \ln J\mathbf {h}_{-j}(x_j) \le C \left( \frac{d(x_0, \tilde{x}_0)}{\rho ^{2/3} k_W^{4/3} J^u_{T^{-j} \bar{\gamma }}T^j(x_j)} + \frac{d(x_0, \tilde{x}_0) k_W^{4/5}}{\rho ^{2/5} \left( J^u_{T^{-j}\bar{\gamma }}T^j(x_j)\right) ^{3/5}} \right) . \end{aligned}$$
(A.11)

\(\square \)

Sublemma A.1

(Relating the gap size and the unstable Jacobian) Let ab define the gap in \(P^+(W)\) as in the statement of the lemma and let \(j, x_j\) be as defined above. There exists \(C>0\), depending only on T, and not on n, j or W, such that

$$\begin{aligned} \frac{1}{J^u_{T^{-j}\bar{\gamma }}T^j(x_j) } \le C k_W^{-2} \rho ^{-26/35} |a-b|^{9/35}. \end{aligned}$$

Proof

We consider the size of the gap created in a neighborhood of \(T^{-j-1}(x_a)\), depending on whether this gap is created by an intersection with homogeneity strips of high index or not.

Case 1. The connected component of \(T^{-j-1}(V_1)\) containing \(T^{-j-1}(x_a)\) intersects \(\mathcal {S}_0\). Then using (6.12) and (6.13), we have,

$$\begin{aligned} |a-b| \ge \frac{C^{\pm 1} k_{j+1}^{-4}}{J^s_{V_1}T^{-j}(x_a)} \ge \frac{C^{\pm 1} k_W^{8/3} \rho ^{4/3}}{\left( J^u_{T^{-j-1}\bar{\gamma }}T^{j+1}(x_{j+1})\right) ^{4/3} J^s_{V_1}T^{-j}(x_a)} , \end{aligned}$$

where we used (6.3) in the second inequality. Now \(J^u_{T^{-j-1}\bar{\gamma }}T^{j+1}(x_{j+1}) = C^{\pm 1} k_j^2 J^u_{T^{-j}\bar{\gamma }}T^j(x_j)\). Using this together with (A.4) to convert between stable and unstable Jacobians yields,

$$\begin{aligned} |a-b| \ge \frac{C^{\pm 1} k_W^{8/3} \rho ^{4/3}}{\left( J^u_{T^{-j}\bar{\gamma }}T^j(x_j)\right) ^{7/3} k_j^{14/3} k_W^{-2}} \ge \frac{C^{\pm 1} k_W^{70/9} \rho ^{26/9} }{\left( J^u_{T^{-j}\bar{\gamma }}T^j(x_j)\right) ^{35/9} } , \end{aligned}$$

where in the second inequality we have used (6.3). This proves the sublemma in this case.

Case 2. The component of \(T^{-j-1}(V_1)\) containing \(T^{-j-1}(x_a)\) does not intersect \(\mathcal {S}_0\). In this case, by the uniform transversality of \(\mathcal {S}_1\) with unstable curves, the gap is bounded below by

$$\begin{aligned} \begin{aligned} |a-b|&\ge \frac{C^{\pm 1} \rho k_W^2}{J^u_{T^{-j-1}\bar{\gamma }}T^{j+1}(x_{j+1}) J^s_{V_1}T^{-j-1}(x_a)} \ge \frac{C^{\pm 1} \rho k_W^4}{\left( J^u_{T^{-j-1}\bar{\gamma }}T^{j+1}(x_{j+1})\right) ^2 k_{j+1}^2 } \\&\ge \frac{C^{\pm 1} \rho k_W^4}{\left( J^u_{T^{-j}\bar{\gamma }}T^j(x_j)\right) ^2}, \end{aligned} \end{aligned}$$

where in the last step, we have used the fact that in Case 2, \(k_{j+1}\) is of order 1. This is clearly a greater lower bound on \(|a-b|\) than in Case 1, proving the sublemma. \(\square \)

With the sublemma proved, we return to our estimate (A.11),

$$\begin{aligned} \ln J\mathbf {h}_{-j}(x_j)\le & {} C \left( \frac{d(x_0, \tilde{x}_0)}{\rho ^{2/3} k_W^{4/3} J^u_{T^{-j} \bar{\gamma }}T^j(x_j)} + \frac{d(x_0, \tilde{x}_0) k_W^{4/5}}{\rho ^{2/5} \left( J^u_{T^{-j}\bar{\gamma }}T^j(x_j)\right) ^{3/5}} \right) \\\le & {} C d(x_0, \tilde{x}_0)\left( k_W^{-10/3} \rho ^{-148/105} |a-b|^{9/35} + k_W^{-2/5} \rho ^{-148/175} |a-b|^{27/175} \right) \\\le & {} C \left( k_W^{-4/3} \rho ^{-43/105} |a-b|^{9/35} + k_W^{8/5} \rho ^{27/175} |a-b|^{27/175} \right) \\\le & {} C \left( \rho ^{-43/105} |a-b|^{9/35} + \rho ^{-29/175} |a-b|^{27/175} \right) \le C \rho ^{-31/105} |a-b|^{1/7} , \end{aligned}$$

where in the third line we have used the fact that \(d(x_0, \tilde{x}_0) \le C \rho k_W^2\) and in the last line that \(k_W \le C \rho ^{-1/5}\). We have also opted to take simpler (slightly less than optimal) exponents, taking the power 1 / 7 rather than 27 / 175 in the second term of the last line and converting \(|a-b|^{4/35} \le C\rho ^{4/35}\) in the first term. Note that \(|a-b| \le C\rho \) follows from Sect. 6.1.

Finally, for the fourth term on the right-hand side of (A.10), we note that the boundary curves for the gap containing \(T^{-j}V_0\) and \(T^{-j}V_1\) both lie in the unstable cone. Since \(T^{-j}V_0\) and \(T^{-j}V_1\) have bounded curvature, we have

$$\begin{aligned} \ln \frac{|T^{-j}V_0|}{|T^{-j}V_1|} \le C d(x_j, \tilde{x}_j) \le \frac{C d(x_0, \tilde{x}_0)}{J^u_{T^{-j} \bar{\gamma }}T^j(x_j)} . \end{aligned}$$

Using Sublemma A.1, this quantity is bounded by

$$\begin{aligned} C d(x_0, \tilde{x}_0) k_W^{-2} \rho ^{-26/35} |a-b|^{9/35} \le C \rho ^{9/35} |a-b|^{9/35}, \end{aligned}$$

where again we have used the fact that \(d(x_0, \tilde{x}_0) \le k_W^2 \rho \). Using the estimates for these four terms in (A.10) ends the proof of Lemma 6.7 since both the average slope and \(\partial _{\bar{x}^s} \overline{G}\) are uniformly bounded away from 0 and infinity. \(\square \)

Appendix B: Estimates for the Dolgopyat bound (Lemmas 8.7, 8.8, and 8.9)

This appendix contains several crucial, but technical, estimates used in the Dolgopyat type cancellation argument developed in Sect. 8.

Proof of Lemma 8.7

Recalling condition (ii) on the foliation from Sect. 6, the first identity in (8.35) together with (8.34) gives

$$\begin{aligned} \begin{aligned} x^s&= G_{i,j,\varkappa }(M_A(x^s),0)+\int _0^{\mathbf {h}^s_A(x^s)}\!\!\!\!\!\!\!\!ds\; \partial _{x^s} G_{i,j,\varkappa }(M_A(x^s),s)\\&= \int _0^{M_A(x^s)}\!\!\!\!\!\!\!\! du\; \partial _{x^u}G_{i,j,\varkappa }(u,0)+ \int _0^{\mathbf {h}^s_A(x^s)}\!\!\!\!\!\!\!\! ds\left[ 1+\int _0^{M_A(x^s)}\!\!\!\!\!\!\!\! du\; \partial _{x^u}\partial _{x^s}G_{i,j,\varkappa }(u,s)\right] \\&=\mathbf {h}^s_A(x^s)+\int _0^{M_A(x^s)}\!\!\!\!\!\!\!\!\!\! du\; \partial _{x^u}G_{i,j,\varkappa }(u,0)+ \int _0^{\mathbf {h}^s_A(x^s)}\!\!\!\!\!\!\!\! ds\int _0^{M_A(x^s)}\!\!\!\!\!\!\!\!\!\! du\; \partial _{x^u}\partial _{x^s}G_{i,j,\varkappa }(u,s) , \end{aligned} \end{aligned}$$
(B.1)

where we have used the fact that \(\partial _{x^s} G_{i,j,\varkappa }(0,s) =1\) for each s by property (ii) of the foliation. Thus, combining the bound (vi) from Sect. 6, that is \(|\partial _{x^u}\partial _{x^s} G_{i,j,\varkappa }|\le C\rho ^{-4/5}=C{\varepsilon }^{-4\varsigma /5}\), the bound \(|M'_A|\le C_{\#}{\varepsilon }^{1-\frac{4}{5}\theta }\) from (8.22) (which implies \(|M_A(x^s)|\le C_{\#}{\varepsilon }\)), and the condition \(|\mathbf {h}^s_A(x^s)|\le C_{\#}{\varepsilon }^\theta \) from (8.34), we get \(\mathbf {h}^s_A(x^s)=x^s(1+\mathcal {O}({\varepsilon }^{1-4\varsigma /5}))+\mathcal {O}({\varepsilon })\).

Next, differentiating the first identity in (8.35), we find

$$\begin{aligned} (\mathbf {h}^s_A)'(x^s)=\frac{1-\partial _{x^u} G_{i,j,\varkappa }(M_A(x^s),\mathbf {h}^s_A(x^s)) M'_A(x^s)}{\partial _{x^s} G_{i,j,\varkappa }(M_A(x^s),\mathbf {h}^s_A(x^s))}. \end{aligned}$$
(B.2)

We have

$$\begin{aligned} \begin{aligned} \partial _{x^s} G_{i,j,\varkappa }\left( M_A(x^s),\mathbf {h}^s_A(x^s)\right)&=\partial _{x^s} G_{i,j,\varkappa }\left( 0,\mathbf {h}^s_A(x^s)\right) \\&\quad + \int _0^{M_A(x^s)}\!\!\!\!\!\!\! \partial _{x^u}\partial _{x^s} G_{i,j,\varkappa }(u,\mathbf {h}^s_A(x^s)) du\\&=1+\mathcal {O}({\varepsilon }^{1-4\varsigma /5}), \end{aligned} \end{aligned}$$

while \(|\partial _{x^u} G_{i,j,\varkappa }|_\infty \le C_{\#}\), hence we have the second inequality of Lemma 8.7.

$$\begin{aligned} \left| 1-\left( \mathbf {h}^s_A\right) '\right| \le C_{\#}{\varepsilon }^{1-\frac{4}{5}\varsigma } , \end{aligned}$$
(B.3)

while the bound on \(\left| \left( \mathbf {h}^s_A\right) (x^s)-x^s\right| \) follows by integration. Finally, \(\mathbf {h}^s_A\) is invertible and the claimed bound on \(\left| \left( \mathbf {h}^s_A\right) ^{-1}(x^s)-x^s\right| \) holds using \(\left( (\mathbf {h}^s_A)^{-1}(x^s) - x^s\right) ' = \frac{1}{\left( \mathbf {h}^s_A\right) ' \circ \left( \mathbf {h}^s_A\right) ^{-1}} - 1\) and integrating as before. \(\square \)

Proof of Lemma 8.8

To start with, note that formula [5, (E.1)] is obtained by a purely geometric argument and uses only that the strong manifolds are in the kernel of the contact form and the weak manifolds in the kernel of its differential. Since the exact same situation holds here, we have, for all relevant manifolds \(W_A\), the formula

$$\begin{aligned} {\varvec{\omega }}_A(x^s)=\int _0^{x^s}ds\int _0^{M_A((\mathbf {h}_A^s)^{-1}(s))}du\;\partial _{x^s}G_{i,j\varkappa }(u,s). \end{aligned}$$
(B.4)

Note that this implies

$$\begin{aligned} |{\varvec{\omega }}_A|_{\infty }\le C_{\#}{\varepsilon }^{1+\theta } . \end{aligned}$$
(B.5)

To obtain the formula we are interested in for each pair AB, it suffices to differentiate. Remembering properties (ii) and (vi) of the foliation constructed in Sect. 6, we have

$$\begin{aligned} \begin{aligned} \partial _{x^s} {\varvec{\omega }}_B(x^s)-\partial _{x^s} {\varvec{\omega }}_A(x^s)&=\int _{M_A\left( \left( \mathbf {h}_A^s\right) ^{-1}(x^s)\right) }^{M_B\left( \left( \mathbf {h}_B^s\right) ^{-1}(x^s)\right) }\!du\;\partial _{x^s} G_{i,j,\varkappa }(u,x^s)\\&=\int _{M_A\left( \left( \mathbf {h}_A^s\right) ^{-1}(x^s)\right) }^{M_B\left( \left( \mathbf {h}_B^s\right) ^{-1}(x^s)\right) }\!du\left[ 1+\int _0^u du_1\; \partial _{x^u} \partial _{x^s}G_{i,j,\varkappa }(u_1,x^s)\right] \\&=\left[ M_B\left( \left( \mathbf {h}_B^s\right) ^{-1}(x^s)\right) {-}M_A\left( \left( \mathbf {h}_A^s\right) ^{-1}(x^s)\right) \right] \left( 1{+}\mathcal {O}\left( {\varepsilon }^{1-\frac{4}{5}\varsigma }\right) \right) . \end{aligned} \end{aligned}$$
(B.6)

Next, the distance between \(W_A^0\cap \{x^s=0\}\) and \(W_B^0\cap \{x^s=0\}\) is given by \(|M_A(0)-M_B(0)|=:d(W_A,W_B)\). Then, by the argument developed in (8.18) (proof of Lemma 8.5), we have

$$\begin{aligned} |~|M_A(0)-M_B(0)|-|M_A(x^s)-M_B(x^s)|~|\le C_{\#}d(W_A,W_B){\varepsilon }^{\theta /5} . \end{aligned}$$
(B.7)

On the other hand, by (B.1) and property (vi) of the foliation, we have

$$\begin{aligned} \begin{aligned} \left| \left( \mathbf {h}^s_B\right) ^{-1}(x^s) -\left( \mathbf {h}^s_A\right) ^{-1}(x^s)\right|&= \left| \int _{M_A\circ \left( \mathbf {h}^s_A\right) ^{-1}(x^s)}^{M_B\circ \left( \mathbf {h}^s_B\right) ^{-1}(x^s)} \!\!\!\!\!\!\! du \; \partial _{x^u} G_{i,j,\varkappa }(u,0) \right| \\&\quad + \left| \int _{M_A\circ \left( \mathbf {h}^s_A\right) ^{-1}(x^s)}^{M_B\circ \left( \mathbf {h}^s_B\right) ^{-1}(x^s)}\!\!\!\!\!\!\!du\int _0^{x^s}\! ds\; \partial _{x^u}\partial _{x^s}G_{i,j,\varkappa }(u,s)\right| \\&\le C_{\#}\left( 1+ {\varepsilon }^{\theta -\frac{4}{5} \varsigma }\right) \left| M_B\circ \left( \mathbf {h}^s_B\right) ^{-1}(x^s)-M_A\circ \left( \mathbf {h}^s_A\right) ^{-1}(x^s)\right| . \end{aligned} \end{aligned}$$
(B.8)

To conclude recall that we are working in coordinates in which \(|M'_{A}|\le C_{\#}{\varepsilon }^{1-\frac{4}{5}\theta }\), cf. the proof of Lemma 8.5, hence

$$\begin{aligned} \begin{aligned}&\left| M_B\circ \left( \mathbf {h}^s_B\right) ^{-1}-M_A\circ \left( \mathbf {h}^s_A\right) ^{-1}\right| \ge \left| M_B\circ \left( \mathbf {h}^s_B\right) ^{-1}-M_A\circ \left( \mathbf {h}^s_B\right) ^{-1}\right| \\&\qquad -C_{\#}{\varepsilon }^{1-\frac{4}{5}\theta }\left| \left( \mathbf {h}^s_B\right) ^{-1}-\left( \mathbf {h}^s_A\right) ^{-1}\right| \\&\quad \ge \left| M_B\circ \left( \mathbf {h}^s_B\right) ^{-1}-M_A\circ \left( \mathbf {h}^s_B\right) ^{-1}\right| \\&\qquad -C_{\#}\left( {\varepsilon }^{1-\frac{4}{5}\theta } + {\varepsilon }^{1-\frac{4}{5} \varsigma + \frac{1}{5} \theta }\right) \left| M_B\circ \left( \mathbf {h}^s_B\right) ^{-1}-M_A\circ \left( \mathbf {h}^s_A\right) ^{-1}\right| , \end{aligned} \end{aligned}$$

which, together with (B.6) and (B.7), proves (8.41).

To prove the second statement, let us introduce the shorthand notation \({\varvec{\omega }}_{A,B}(x^s)={\varvec{\omega }}_A(x^s)-{\varvec{\omega }}_B(x^s)\) and \({\varvec{\eta }}_A=\left( \mathbf {h}_A^s\right) ^{-1}\), \({\varvec{\eta }}_B=\left( \mathbf {h}_B^s\right) ^{-1}\). By (B.6) we have

$$\begin{aligned}&\left| \partial _{x^s}{\varvec{\omega }}_{A,B}(x^s)-\partial _{x^s}{\varvec{\omega }}_{A,B}(y^s)\right| \\&\quad =\left| \int _{M_A\left( {\varvec{\eta }}_A(x^s)\right) }^{M_B\left( {\varvec{\eta }}_B(x^s)\right) }du\;\partial _{x^s} G_{i,j,\varkappa }(u,x^s)-\int _{M_A\left( {\varvec{\eta }}_A(y^s)\right) }^{M_B\left( {\varvec{\eta }}_B(y^s)\right) }du\;\partial _{x^s} G_{i,j,\varkappa }(u,y^s)\right| \\&\quad \le \left| \int _{M_A({\varvec{\eta }}_A(x^s))}^{M_B({\varvec{\eta }}_B(x^s))}\!du\;\partial _{x^s} G_{i,j,\varkappa }(u,x^s)- \partial _{x^s}G_{i,j,\varkappa }(u,y^s)\right| \\&\qquad +\left| \int _{M_B({\varvec{\eta }}_B(y^s))}^{M_B({\varvec{\eta }}_B(x^s))}\!du\;\partial _{x^s} G_{i,j,\varkappa }(u,y^s)\right| \\&\qquad +\left| \int _{M_A({\varvec{\eta }}_A(x^s))}^{M_A({\varvec{\eta }}_B(y^s))}\!du\;\partial _{x^s} G_{i,j,\varkappa }(u,y^s)\right| . \end{aligned}$$

Notice that (ii) of the foliation implies

$$\begin{aligned} \begin{aligned}&\partial _{x^s}G_{i,j,\varkappa }(u,x^s)- \partial _{x^s}G_{i,j,\varkappa }(u,y^s)\\&\quad =\partial _{x^s}G_{i,j,\varkappa }(u,x^s)-\partial _{x^s} G_{i,j,\varkappa }(u,y^s)-\partial _{x^s} G_{i,j,\varkappa }(0,x^s)\\&\qquad +\partial _{x^s} G_{i,j,\varkappa }(0,y^s) . \end{aligned} \end{aligned}$$

Thus, the four-point property (vii) (applied to the points \((u,x^s)\) and \((0,y^s)\)) implies

$$\begin{aligned} |\partial _{x^s}G_{i,j,\varkappa }(u,x^s)- \partial _{x^s}G_{i,j,\varkappa }(u,y^s)|\le C_{\#}{\varepsilon }^{-\left( \frac{4}{5}+\frac{11{\varpi }}{15}\right) \varsigma +1-7{\varpi }}|x^s-y^s|^{\varpi }. \end{aligned}$$
(B.9)

On the other hand, Eq. (8.22) and Lemma 8.7 imply

$$\begin{aligned} |M_B({\varvec{\eta }}_B(x^s))-M_B({\varvec{\eta }}_B(y^s))|\le C_{\#}{\varepsilon }^{1-\frac{4}{5}\theta }|x^s-y^s|, \end{aligned}$$

and the same for A. Remembering that \(|x^s|, |y^s|\le c{\varepsilon }^\theta \), property (v) of the foliation, and our conditions on \({\varpi }, \theta ,\varsigma \) from (8.42), the above facts yield

$$\begin{aligned} \left| \partial _{x^s}{\varvec{\omega }}_{A,B}(x^s)-\partial _{x^s}{\varvec{\omega }}_{A,B}(y^s)\right|\le & {} C_{\#}{\varepsilon }|x^s-y^s|^{\varpi }+{\varepsilon }^{1-\frac{4}{5}\theta }|x^s-y^s|\\\le & {} C_{\#}{\varepsilon }|x^s-y^s|^{\varpi }\end{aligned}$$

which, together with (B.5), proves the second statement of Lemma 8.8.

To continue, let

$$\begin{aligned} {\varvec{L}}_{A,B}(x^s, x^0)=\frac{[(m-1)!]^2}{(\ell {\tau }_-)^{2m-2}} e^{-2a\ell {\tau }_-} \mathbf{G}^*_{\ell ,m,i,A}(x^s, x^0)\overline{\mathbf{G}^*_{\ell ,m,i,B}(x^s, x^0)} . \end{aligned}$$

Next we introduce a sequence \(\{w_j\}_{j=0}^M\subset {\mathbb R}\) such that \(w_0=-cr^\theta \) and \(\partial _{x^s}{\varvec{\omega }}_{A,B}(w_j)(w_{j+1}-w_j)=2\pi b^{-1}\) and let \(M\in {\mathbb N}\) be such that \(w_M\le c {\varepsilon }^\theta \) and \(w_{M+1}> c{\varepsilon }^\theta \).Footnote 47 Also, we set \({\varvec{\delta }}_j=w_{j+1}-w_j\). By Lemma 8.7 it follows that, for each \(x^s\in [w_j,w_{j+1}]\),

$$\begin{aligned} |{\varvec{\omega }}_{A,B}(x^s)-{\varvec{\omega }}_{A,B}(w_j)-\partial _{x^s}{\varvec{\omega }}_{A,B}(w_j)(x^s-w_j)|\le C_{\#}{\varepsilon }{\varvec{\delta }}_j^{1+{\varpi }}. \end{aligned}$$

In addition, the bounds in (8.39) imply

$$\begin{aligned} |{\varvec{L}}_{A,B}(x^s,x^0)-{\varvec{L}}_{A,B}(w_j,x^0)|\le C_{\#}{\varvec{\delta }}_j {\varepsilon }^{-3\theta } . \end{aligned}$$

Then, using the first part of the Lemma,

$$\begin{aligned} \begin{aligned}&\left| \int _{w_j}^{w_{j+1}} e^{-ib{\varvec{\omega }}_{A,B}(x^s)} {\varvec{L}}_{A,B}(x^s,x^0) d x^s\right| \\&\quad =\left| \int _{w_j}^{w_{j+1}} e^{-ib[\partial _{x^s}{\varvec{\omega }}_{A,B}(w_i)(x^s-w_j)+\mathcal {O}({\varepsilon }{\varvec{\delta }}_j^{1+{\varpi }})]} [{\varvec{L}}_{A,B}(w_j, x^0)+\mathcal {O}({\varepsilon }^{-3\theta }{\varvec{\delta }}_j)] dx^s\right| \\&\quad \le C_{\#}\left( b{\varvec{\delta }}_j^{1+{\varpi }}{\varepsilon }^{-2\theta +1}+{\varepsilon }^{-3\theta }{\varvec{\delta }}_j\right) {\varvec{\delta }}_j\\&\quad \le C_{\#}\left( \frac{{\varepsilon }^{-2\theta +1}}{d(W_A,W_B)^{1+{\varpi }} b^{\varpi }}+\frac{{\varepsilon }^{-3\theta }}{d(W_A,W_B) b}\right) {\varvec{\delta }}_j . \end{aligned} \end{aligned}$$

We may be left with the integral over the interval \([w_M,cr^\theta ]\) which is trivially bounded by \(C_{\#}{\varepsilon }^{-2\theta }{\varvec{\delta }}_M\le C_{\#}[{\varepsilon }^{2\theta }bd(W_A, W_B)]^{-1}\). The statement follows since the manifolds we are considering have length at most \(c{\varepsilon }^\theta \), hence \(\sum _{j=0}^{M-1}{\varvec{\delta }}_j \le c{\varepsilon }^\theta \). \(\square \)

Proof of Lemma 8.9

We start by introducing a function \(\bar{R}:W\rightarrow {\mathbb N}\) such that \(\bar{R}(\xi )\) is the first \(t\in {\mathbb N}\) at which \(\Phi _{-t{\tau }_-}\xi \) belongs to a component of \(\Phi _{-t{\tau }_-}W\) of size larger than \(\kappa _* L_{0}\), \(\kappa _*<1/3\), and distant more than \(L_0\) from \(\partial \Omega _0\). \(\square \)

Remark B.1

We choose \(\kappa _*\) such that, if \(\Phi _{-t{\tau }_-}W\) is a regular piece of size larger than \(L_{0}/3\) but in an \(L_0\) neighborhood of \(\partial \Omega _0\), then either \(\Phi _{-(t+k){\tau }_-}W\) or \(\Phi _{-(t-k){\tau }_-}W\) will satisfy our requirement for some \(C_{\#}>k{\tau }_->L_0\).

We define then \(R(\xi )=\min \{\bar{R}(\xi ),\ell \}\). Let \(\mathcal {P}=\{J_{i}\}\) be the coarser partition of W in intervals on which \(R\) is constant. Note that, for each \(W_{B,i}\), \(\Phi _{\ell {\tau }_-}(W_{B,i})\subset J_{j}\) for some \(J_{j}\in \mathcal {P}\).

Let

$$\begin{aligned} \Sigma _{\ell ,j}=\{(B,i)\;:\; i\in {\mathbb N},~ B\in E_{\ell ,i},~ d\left( W_{B,i}, W^0_{A,i}\right) \le \rho _{*},\; \Phi _{\ell {\tau }_-}W_{B,i}\subset J_{j}\}. \end{aligned}$$

Then, by Lemma 3.5, for each \((B,i)\in \Sigma _{\ell ,j}\)

$$\begin{aligned} Z_{\ell ,B,i} = \int _{W_{B,i}}J^s_{\ell {\tau }_-}\le C_{\#}\frac{|J_{j}|}{|\overline{W}_{j}|}\int _{W_{B,i}} J^s_{(\ell -R_j) {\tau }_-} \le C_{\#}\frac{|J_{j}|}{|\overline{W}_{j}|}|\Phi _{(\ell -R_{j}){\tau }_-}W_{B,i}|~, \end{aligned}$$
(B.10)

where \(R_{j}=R(J_{j})\) and \(\overline{W}_{j}=\Phi _{-R_{j}{\tau }_-}J_{j}\). Note that, by construction, either \(|\overline{W}_{j}|\ge \kappa _* L_{0}\) or \(R_j=\ell \) and \(\overline{W}_{j}=W_{B,i}\). Next, consider the local weak stable surfaces \(W_{B,i}^{0} := \cup _{t \in [-cr^\theta , cr^\theta ]} \Phi _t W_{B,i}\) and \(\overline{W}^{0}_j := \cup _{t \in [-cr^\theta , cr^\theta ]} \Phi _t \overline{W}_{j}\).

Let us analyze first the case in which \(R_j<\ell \). Let \(\rho \le C_{\#}L_0^{\frac{5}{3}}\). Then, by assumption, \(\overline{W}_{j}\) is a manifold with satisfies condition (a) at the beginning of Sect. 6. Indeed \(\overline{W}_{j}\) is too long to belong to \(\mathbb {H}_k\) with \(k\ge C\rho ^{-\frac{1}{5}}\), so that \(\overline{W}_j\) satisfies condition (b) of that section as well. We can then use the construction in Sect. 6 to define an approximate unstable foliation in a \(\rho \) neighborhood of the surface \(\overline{W}^{0}_j\). Let \(\Gamma _{B,i}\) be the set of leaves that intersect \(\Delta _{\ell -R_{j}}\cap \Phi _{(\ell -R_{j}){\tau }_-}W^{0}_{B,i}\). By the construction of the covering \(B_{c{\varepsilon }^\theta }(x_i)\), the \(\Gamma _{B,i}\) can have at most \(C_{\#}\) overlaps and since \(W_B\) completely crosses \(B_{c{\varepsilon }^\theta }(x_i)\), there can be no gaps between the curves in \(\Gamma _{B,i}\). Now using the uniform transversality between the stable, unstable and flow directions,

$$\begin{aligned} \sum _{(B,i)\in \Sigma _{\ell ,j}}m(\Gamma _{B,i}) \ge C_{\#}\!\!\sum _{(B,i)\in \Sigma _{\ell ,j}}|\Phi _{(\ell -R_{j}){\tau }_-}W_{B,i}|\rho {\varepsilon }^\theta . \end{aligned}$$

Accordingly, for each j such that \(R_j<\ell \), remembering (B.10),

$$\begin{aligned} \sum _{(B,i)\in \Sigma _{\ell ,j}}Z_{\ell ,B,i}\le C_{\#}|J_{j}| \rho ^{-1} {\varepsilon }^{-\theta } \sum _{(B,i)\in \Sigma _{\ell ,j}}m(\Phi _{-(\ell -R_{j}){\tau }_-}\Gamma _{B,i}) ~, \end{aligned}$$
(B.11)

where we have used the invariance of the volume. Remembering that the \(\Phi _{-(\ell -R_{j}){\tau }_-}\Gamma _{B,i}\) have a fixed maximal number of overlaps and since they must be all contained in a box containing \(\widetilde{O}\) of length \(C_{\#}{\varepsilon }^\theta \) in the flow direction, of length \(C_{\#}{\varepsilon }^\theta \) in the stable direction and of length \(C_{\#}(\rho _{*} + \Lambda ^{-(\ell -R_j){\tau }_-}\rho )\) in the unstable direction, we have,Footnote 48

$$\begin{aligned} \begin{aligned} \sum _j\sum _{(B,i)\in \Sigma _{\ell ,j}}Z_{B,i}\le&\sum _{\left\{ j\;:\; R_{j}\le \frac{\ell }{2}\right\} } C_{\#}|J_{j}|~\rho ^{-1} {\varepsilon }^\theta (\rho _{*}+\Lambda ^{-\frac{\ell }{2}{\tau }_-} \rho )\\&+\sum _{\left\{ j\;:\; R_{j}> \frac{\ell }{2}\right\} }\sum _{(B,i)\in \Sigma _{\ell ,j}}Z_{B,i}. \end{aligned} \end{aligned}$$
(B.12)

To estimate the sum on \(R_{j}> \frac{\ell }{2}\) we use the growth Lemma 3.8(a), with \({1/q_0}=0\), which, remembering Remark B.1, implies

$$\begin{aligned} \begin{aligned} \sum _{\left\{ j\;:\; R_{j}>\frac{\ell }{2}\right\} }\sum _{(B,i)\in \Sigma _{\ell ,j}}Z_{B,i}&\le \sum _{\left\{ j:\; R_{j}> \frac{\ell }{2}\right\} } \sum _{(B,i)\in \Sigma _{\ell ,j}}C_{\#}\int _{\Phi _{\ell {\tau }_-/2}W_{B,i}}J^s_{\Phi _{\ell {\tau }_-/2}W_{B,i}}\Phi _{\ell {\tau }_-/2}\\&\le \sum _{W_B\in \mathcal {I}_{\ell {\tau }_-/2}(W)} C_{\#}\int _{W_{B}}J^s_{W_{B}}\Phi _{\ell {\tau }_-/2} \\&\le \sum _{W_B\in \mathcal {I}_{\ell {\tau }_-/2}(W)} C_{\#}L_0|J^s_{W_{B}}\Phi _{\ell {\tau }_-/2}|_{\mathcal {C}^0}\le C_{\#}\lambda ^{\ell {\tau }_-/2}~. \end{aligned} \end{aligned}$$

Since \(\lambda > \Lambda ^{-1}\), the lemma follows by choosing \(\rho = \rho _{*}^{1/2}\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Baladi, V., Demers, M.F. & Liverani, C. Exponential decay of correlations for finite horizon Sinai billiard flows. Invent. math. 211, 39–177 (2018). https://doi.org/10.1007/s00222-017-0745-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00222-017-0745-1

Navigation