This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Global rigidity for some partially hyperbolic abelian actions with 1-dimensional center

Sven Sandfeldt [email protected]
(Date: December 2023)
Abstract.

We obtain a global rigidity result for abelian partially hyperbolic higher rank actions on certain 22-step nilmanifolds XΓX_{\Gamma}. We show that, under certain natural assumptions, all such actions are CC^{\infty}-conjugated to an affine model. As a consequence, we obtain a centralizer rigidity result, classifying all possible centralizers for any C1C^{1}-small perturbation of an irreducible, affine partially hyperbolic map on XΓX_{\Gamma}. Along the way, we prove two results of independent interest. We describe fibered partially hyperbolic diffeomorphisms on XΓX_{\Gamma} and we show that topological conjugacies between partially hyperbolic actions and higher rank affine actions are CC^{\infty}.

1. Introduction

Rigidity of k\mathbb{Z}^{k}-actions on tori (and nilmanifolds) with some hyperbolicity have been studied extensively. The general philosophy is: large abelian actions with some hyperbolicity should be globally rigid, i.e. smoothly conjugated to algebraic models. A big breakthrough result in this direction was obtained by Katok and Spatzier [44] where they prove that all perturbations of certain algebraic Anosov actions are smoothly conjugated back to the corresponding algebraic models. There they also outline a rigidity program for abelian actions with hyperbolicity. Since the paper by Katok and Spatzier, a lot of results have been obtained for large abelian actions close to some algebraic model with some hyperbolicity [16, 18, 14, 62, 63, 67, 68, 66, 23, 12]. Even earlier than the result by Katok and Spatzier, Katok and Lewis [43] proved a global rigidity statement for Anosov action on tori. In [43], the authors show that a certain class of Anosov d\mathbb{Z}^{d}-actions on the torus 𝕋d+1\mathbb{T}^{d+1} is, necessarily, smoothly conjugated to an algebraic model, even though the action might not be close to the algebraic model. A crucial assumption in [43] is that the d\mathbb{Z}^{d}-action contains many Anosov elements. The property of having many Anosov elements was removed by F. Rodriguez Hertz in [53], where Rodriguez Hertz only assumes that the action contains one Anosov element. Nilmanifolds are natural generalizations of tori. There has been a lot of work studying the global rigidity of higher rank111See Definition 2.3. Anosov actions on nilmanifolds since the paper by Katok and Lewis [38, 39, 40, 25]. The culmination of these works is the result by F. Rodriguez Hertz, Z. Wang and D. Fisher, B. Kalinin, R. Spatzier [55, 26] proving that abelian higher rank Anosov actions on (infra-)nilmanifolds are smoothly conjugated to algebraic models, completely resolving the question of global rigidity of abelian higher rank Anosov actions on these manifolds. Relaxing the Anosov assumption leads to the question:

When are partially hyperbolic k\mathbb{Z}^{k}-actions on nilmanifolds by affine maps?

These actions have been remarkably resistant to classification results. Even local rigidity for algebraic partially hyperbolic actions on (non-toral) nilmanifolds has been open for a long time until recent advances by Z. J. Wang [66]. In the global Anosov setting, any higher rank action is topologically conjugated to an affine action by the global topological rigidity for Anosov diffeomorphisms [27, 47]. Because of this, the key problem to solve is to upgrade the regularity of an already existing topological conjugacy. In contrast, for partially hyperbolic diffeomorphisms there is no global topological rigidity. In fact, there are nilmanifolds on which a generic partially hyperbolic diffeomorphism can not be conjugated to an affine model (for example, the manifolds considered in Theorem A). The reason for this is that affine partially hyperbolic diffeomorphisms always have isometric center whereas this is a very special property for general partially hyperbolic systems. As a consequence, when we study the global rigidity of partially hyperbolic actions, a significant problem to solve is the existence of a topological conjugacy. However, once the existence of a topological conjugacy is proved it turns out that upgrading to a smooth conjugacy follows similarly to the Anosov setting (see Theorem 1.2 and Section 8). In this paper, we produce an initial result towards answering the global rigidity question for partially hyperbolic actions.

In Theorem A we prove the first global rigidity result for higher rank abelian actions on nilmanifolds with one partially hyperbolic element. The nilmanifolds under consideration are products between tori and Heisenberg nilmanifolds. These manifolds are, in a sense, the simplest for studying partially hyperbolic diffeomorphisms (with 11-dimensional center). The reason for this is that the affine models of partially hyperbolic systems on these manifolds simultaneously satisfy three important properties, (i) they have 11-dimensional center direction, (ii) the center leaves are compact and (iii) these systems are accessible (see Section 2.1). This allows us to use significant portions of the partially hyperbolic theory when studying these systems. In contrast, when studying Anosov actions the simplest manifolds are tori. However, there are no accessible affine partially hyperbolic systems on tori, so studying global rigidity for partially hyperbolic actions on tori (that is not derived from Anosov) becomes very difficult because many of the tools from partially hyperbolic theory are not applicable.

The questions of local and global rigidity of higher rank actions can also be studied for different types of actions, either dropping the assumption that the action is abelian, or dropping the assumption that the action should have some hyperbolicity. Local rigidity has been obtained for large abelian parabolic actions, with no hyperbolicity [17, 15, 65, 13]. Removing the assumption that the action should be abelian, we can study the actions of higher rank lattices in semi-simple Lie groups, see for example [24] and the references therein. In fact, the rigidity result for Abelian actions in [43] was used to obtain local rigidity of SL(n,){\rm SL}(n,\mathbb{Z})-action on tori. When considering rigidity of Anosov lattice actions on nilmanifolds, there are also global results, see for example [10]. A key point in [10] is that any conjugacy between the hyperbolic lattice action and the algebraic model also conjugates the action of a large abelian subgroup to some algebraic abelian action. So, the results of [55, 26] can be applied to improve the regularity of the conjugacy. Considering the main results of this paper, see Theorem A, a natural question is: are higher rank partially hyperbolic lattice actions on nilmanifolds always by affine maps? Conjecturally this question should have an affirmative answer (see [30, Conjecture 5]). In [46] H. Lee and the author show, by using results from this paper, that the question has an affirmative answer when the manifold is a Heisenberg nilmanifold.

1.1. Global rigidity of partially hyperbolic actions

Let GG be a simply connected \ell-step nilpotent Lie group. That is, the lower central series G(1)=GG^{(1)}=G, G(j+1)=[G,G(j)]G^{(j+1)}=[G,G^{(j)}], terminate at +1\ell+1, G(+1)=eG^{(\ell+1)}=e. Given a lattice ΓG\Gamma\leq G we define the associated compact nilmanifold as the quotient XΓ=ΓGX_{\Gamma}=\Gamma\setminus G. Compact nilmanifolds have associated groups of automorphisms and affine maps

(1.1) Aut(XΓ)={LAut(G) : LΓ=Γ},\displaystyle{\rm Aut}(X_{\Gamma})=\{L\in{\rm Aut}(G)\text{ : }L\Gamma=\Gamma\},
(1.2) Aff(XΓ)={f0(x)=L(x)g1 : LAut(XΓ), gG}.\displaystyle{\rm Aff}(X_{\Gamma})=\{f_{0}(x)=L(x)g^{-1}\text{ : }L\in{\rm Aut}(X_{\Gamma}),\text{ }g\in G\}.

By automorphism rigidity of nilpotent lattices [51] we can, equivalently, define Aut(XΓ)=Aut(Γ)=Aut(π1XΓ){\rm Aut}(X_{\Gamma})={\rm Aut}(\Gamma)={\rm Aut}(\pi_{1}X_{\Gamma}). An automorphism ρ:kAut(XΓ)\rho:\mathbb{Z}^{k}\to{\rm Aut}(X_{\Gamma}) is said to have a rank1-1 factor if there is some quotient X^Γ\hat{X}_{\Gamma} of XΓX_{\Gamma} such that ρ\rho descends to X^Γ\hat{X}_{\Gamma} and the induced map ρ^:kAut(X^Γ)\hat{\rho}:\mathbb{Z}^{k}\to{\rm Aut}(\hat{X}_{\Gamma}) factor through a map Aut(X^Γ)\mathbb{Z}\to{\rm Aut}(\hat{X}_{\Gamma}). A homomorphism ρ:kAut(XΓ)\rho:\mathbb{Z}^{k}\to{\rm Aut}(X_{\Gamma}) is higher rank if it has no rank1-1 factor. More generally, given a smooth action α:k×XΓXΓ\alpha:\mathbb{Z}^{k}\times X_{\Gamma}\to X_{\Gamma} we have an induced map α:kAut(π1XΓ)\alpha_{*}:\mathbb{Z}^{k}\to{\rm Aut}(\pi_{1}X_{\Gamma}). We say that α\alpha is higher rank if the induced map α\alpha_{*} is higher rank.

A diffeomorphism f:XΓXΓf:X_{\Gamma}\to X_{\Gamma} is partially hyperbolic if there is a DfDf-invariant splitting TXΓ=EsEcEuTX_{\Gamma}=E^{s}\oplus E^{c}\oplus E^{u} such that EsE^{s} is exponentially contracted, EuE^{u} is exponentially expanded and the behaviour along EcE^{c} is dominated by the behaviour of DfDf along EsE^{s} and EuE^{u} (for a precise definition, see Section 2.1). We will need a technical assumption on the stable and unstable foliations WsW^{s} and WuW^{u}, tangent to EsE^{s} and EuE^{u}. We say that WsW^{s} and WuW^{u} are quasi-isometric in the universal cover if the metric along the leaves of the foliations, after we lift them to the universal cover, is comparable to the ambient metric (see Definition 2.1). If the center bundle EcE^{c} is the trivial bundle then ff is Anosov. The main result of this paper is an extension of the results of [55] to certain nilmanifolds by weakening the assumption that α\alpha is Anosov. Instead, we assume that the action α\alpha contains a partially hyperbolic element.

Theorem A.

Let GG be a 22-step nilpotent Lie group with dim[G,G]=1\dim[G,G]=1, ΓG\Gamma\leq G a lattice and XΓ=ΓGX_{\Gamma}=\Gamma\setminus G the associated nilmanifold. Let α:k×XΓXΓ\alpha:\mathbb{Z}^{k}\times X_{\Gamma}\to X_{\Gamma} be a smooth higher rank action with 𝐧0k\mathbf{n}_{0}\in\mathbb{Z}^{k} such that f=α𝐧0f=\alpha^{\mathbf{n}_{0}} is partially hyperbolic and satisfying

  1. (i)

    ff has 11-dimensional center,

  2. (ii)

    the stable and unstable foliations WsW^{s}, WuW^{u} are quasi-isometric in the universal cover.

Then α\alpha is CC^{\infty}-conjugated to some affine action α0:kAff(XΓ)\alpha_{0}:\mathbb{Z}^{k}\to{\rm Aff}(X_{\Gamma}).

Remark 1.

Condition (ii)(ii) in the theorem is a technical assumption that guarantees that ff is fibered (see Theorem 1.1). Motivated by results in [34, 33, 35], it seems plausible that (ii)(ii) is always satisfied. Moreover, in [46, Lemma A2] it is shown that if ff is topologically conjugated (or more generally, leaf conjugated) to some affine f0f_{0} then condition (ii)(ii) is satisfied. It follows that α\alpha satisfies the assumptions of the theorem if α\alpha satisfies the conclusion of the theorem, so the theorem is sharp in this sense. It also follows that if there are exotic partially hyperbolic systems that do not satisfy (ii)(ii), then there exists no global rigidity result for these systems.

Remark 2.

With GG as in Theorem A the group GG can be written as G=Hn×mG=H^{n}\times\mathbb{R}^{m}, n0n\neq 0, where HnH^{n} is a Heisenberg group. The nilmanifold XΓX_{\Gamma} is also a product of a Heisenberg nilmanifold and a torus. On these manifolds there exists no Anosov actions since the derived subgroup, [G,G][G,G], is isometric for any automorphism. So Theorem A is the only global rigidity result on these manifolds, since [55, 26] do not apply. In fact, to the author’s knowledge, Theorem A is the first global rigidity result for abelian actions assuming only one partially hyperbolic element.

Remark 3.

In principle the proof of Theorem A should work for \ell-step GG with >2\ell>2 as long as dimG()=1\dim G^{(\ell)}=1. In this case the quasi-isometric assumption, assumption (ii)(ii), would have to be changed. This is a work in progress.

Remark 4.

If α\alpha is topologically conjugated to some affine action, then the methods from [55] generalize to partially hyperbolic systems, see Theorem 1.2. So, the main novelty of Theorem A is that we produce a topological conjugacy from α\alpha to an affine action α0\alpha_{0}.

1.2. Applications to centralizer classification and centralizer rigidity

Given a diffeomorphism f:MMf:M\to M on a closed manifold we define its smooth centralizer as the group of diffeomorphisms that commute with ff. That is, we define

(1.3) Z(f)={gDiff(M) : fg=gf}.\displaystyle Z^{\infty}(f)=\{g\in{\rm Diff}^{\infty}(M)\text{ : }fg=gf\}.

We are interested in two questions about the group Z(f)Z^{\infty}(f):

  1. (i)

    What are possible groups that arise as Z(f)Z^{\infty}(f) for some fDiff(M)f\in{\rm Diff}^{\infty}(M)?

  2. (ii)

    If Z(f)Z^{\infty}(f) is large (compared to the conjecturally generic size \mathbb{Z}, [59, 60]) what can be said about ff?

In this level of generality, questions (i)(i) and (ii)(ii) are difficult (or possibly impossible) to answer. Instead, we fix f0Aff(XΓ)f_{0}\in{\rm Aff}(X_{\Gamma}) for some XΓ=ΓGX_{\Gamma}=\Gamma\setminus G, and consider questions (i)(i) and (ii)(ii) for those fDiff(XΓ)f\in{\rm Diff}^{\infty}(X_{\Gamma}) that are C1C^{1}-close to f0f_{0}. We call (i)(i) the question of local centralizer classification and (ii)(ii) the question of local centralizer rigidity around f0f_{0}. These questions were raised and addressed by Damjanović, Wilkinson and Xu in [20] where the authors study perturbations of timet0-t_{0} map of geodesic flows on negatively curved manifolds and trivial circle extensions of hyperbolic automorphisms. In [5] the authors study local centralizer rigidity of time1-1 maps of Anosov flows on 33-manifolds, generalizing results from [20] in the context of 33-manifolds. Another generalization of results from [20] was obtained by W. Wang in [69], where semi-simple Lie groups of higher rank were studied instead of rank1-1 simple groups. For ergodic toral automorphisms, Gan, Xu, Shi and Zhang studied partially hyperbolic diffeomorphisms on 𝕋3\mathbb{T}^{3} homotopic to an hyperbolic automorphism [28]. In [57] the author studies local centralizer classification and rigidity for some partially hyperbolic, irreducible222An automorphism AGL(d,)A\in{\rm GL}(d,\mathbb{Z}) is irreducible if the characteristic polynomial pA(t)p_{A}(t) is irreducible in [t]\mathbb{Q}[t]. toral automorphisms.

If ff is partially hyperbolic with (uniquely integrable) center foliation WcW^{c}, then we obtain a normal subgroup Zc(f)Z(f)Z_{c}^{\infty}(f)\subset Z^{\infty}(f), the center fixing centralizer:

(1.4) Zc(f):={gZ(f) : gxWc(x), xXΓ}.\displaystyle Z_{c}^{\infty}(f):=\{g\in Z^{\infty}(f)\text{ : }gx\in W^{c}(x),\text{ }x\in X_{\Gamma}\}.

From [21, Theorem 5], if Zc(f)Z_{c}^{\infty}(f) is sufficiently big and ff is fibered (see [21, Definition 1]) then ff is smoothly conjugated to an isometric extension of an Anosov map. Combining this with Theorem A we completely classify the centralizers of diffeomorphisms C1C^{1}-close to affine partially hyperbolic maps.

Let GG be the (d+1)(d+1)-dimensional Heisenberg group and XΓ=ΓGX_{\Gamma}=\Gamma\setminus G a compact Heisenberg nilmanifold. We have a natural fibration

(1.5) π:XΓ𝕋d,\displaystyle\pi:X_{\Gamma}\to\mathbb{T}^{d},

where we refer to 𝕋d\mathbb{T}^{d} as the base of XΓX_{\Gamma}. Any affine map on XΓX_{\Gamma} descends to an affine map on 𝕋d\mathbb{T}^{d} and from the group relations in GG any automorphism LAut(XΓ)L\in{\rm Aut}(X_{\Gamma}) induces an element of Sp(d,){\rm Sp}(d,\mathbb{Z}) on 𝕋d\mathbb{T}^{d}. Conversely, any LSp(d,)L\in{\rm Sp}(d,\mathbb{Z}) defines an element of Aut(XΓ){\rm Aut}(X_{\Gamma}), see Section 2.2. Given f0Aff(XΓ)f_{0}\in{\rm Aff}(X_{\Gamma}) we denote by LsuSp(d,)L_{su}\in{\rm Sp}(d,\mathbb{Z}) the induced automorphism on 𝕋d\mathbb{T}^{d}. Before stating the theorem we define for any f0Aff(XΓ)f_{0}\in{\rm Aff}(X_{\Gamma}), with irreducible, hyperbolic induced map on the base LsuSp(d,)L_{su}\in{\rm Sp}(d,\mathbb{Z}), the natural number

(1.6) r0(f0)=rank(Z(f0)Zc(f0))=rank(ZSp(d,)(Lsu)).\displaystyle r_{0}(f_{0})={\rm rank}\left(\frac{Z^{\infty}(f_{0})}{Z_{c}^{\infty}(f_{0})}\right)={\rm rank}\left(Z_{{\rm Sp}(d,\mathbb{Z})}(L_{su})\right).

Lemma A.3 explicitly calculates the number r0(f0)r_{0}(f_{0}), and if d6d\geq 6, then r0(f0)>1r_{0}(f_{0})>1.

Theorem B.

Let XΓX_{\Gamma} be a compact Heisenberg nilmanifold and let f0Aff(XΓ)f_{0}\in{\rm Aff}(X_{\Gamma}) be partially hyperbolic with 11-dimensional center and LsuL_{su} irreducible. If fDiff(XΓ)f\in{\rm Diff}^{\infty}(X_{\Gamma}) is C1C^{1}-close to f0f_{0} then one of the following holds

  1. (i)

    either Z(f)Z^{\infty}(f) is virtually trivial,

  2. (ii)

    or Z(f)Z^{\infty}(f) is virtually ×𝕋\mathbb{Z}\times\mathbb{T} in which case ff is an isometric extension of some Anosov diffeomorphism on 𝕋d\mathbb{T}^{d},

  3. (iii)

    or Z(f)Z^{\infty}(f) is virtually r0×𝕋\mathbb{Z}^{r_{0}}\times\mathbb{T} and if r0>1r_{0}>1 then ff is CC^{\infty}-conjugate to some (possibly different) affine map f~0Aff(XΓ)\tilde{f}_{0}\in{\rm Aff}(X_{\Gamma}).

Remark 5.

All cases (i)(i), (ii)(ii) and (iii)(iii) occur, so Theorem B completely classifies the centralizer of fDiff(XΓ)f\in{\rm Diff}^{\infty}(X_{\Gamma}) close to a partially hyperbolic f0Aff(XΓ)f_{0}\in{\rm Aff}(X_{\Gamma}). Case (i)(i) holds generically [6]. Case (ii)(ii) can be produced by fixing some irreducible, hyperbolic LSp(d,)L\in{\rm Sp}(d,\mathbb{Z}) and defining ff on Gd×G\cong\mathbb{R}^{d}\times\mathbb{R} by f(x,t)=(Lx,t+β(x))f(x,t)=(Lx,t+\beta(x)) where the second coordinate is identified with [G,G][G,G]\cong\mathbb{R} and β:𝕋d\beta:\mathbb{T}^{d}\to\mathbb{R} is a cocycle over LL that is not cohomologous to a constant. The last case holds when ff is CC^{\infty}-conjugate to some affine f~0\tilde{f}_{0}, so in particular when we take the trivial perturbation f=f0f=f_{0}.

Remark 6.

Similar results as Theorem B have been obtained independently by Damjanović, Wilkinson and Xu using different methods with additional assumptions [19].

1.3. Partially hyperbolic maps on nilmanifolds

When proving Theorems A and B, we use a description of partially hyperbolic diffeomorphisms on the nilmanifolds considered in Theorem A. The main property that we show is that, under the assumptions of Theorem A, the system ff is fibered in the terminology of [4]:

Theorem 1.1.

Let G=Hn×mG=H^{n}\times\mathbb{R}^{m} (where we allow n=0n=0), ΓG\Gamma\leq G a lattice with associated nilmanifold XΓ=ΓGX_{\Gamma}=\Gamma\setminus G. Let fDiff(XΓ)f\in{\rm Diff}^{\infty}(X_{\Gamma}) be partially hyperbolic and satisfy (i)(i), (ii)(ii) from Theorem A. If GG is abelian we assume, in addition, that the induced map f:H1(XΓ)H1(XΓ)f_{*}:H_{1}(X_{\Gamma})\to H_{1}(X_{\Gamma}) has at least one rational eigenvalue. The following holds

  1. (i)

    ff is dynamically coherent with global product structure,

  2. (ii)

    all foliations WσW^{\sigma}, σ=s,c,u,cs,cu\sigma=s,c,u,cs,cu, are uniquely integrable,

  3. (iii)

    the center foliation WcW^{c} have compact oriented circle leaves,

  4. (iv)

    ff is fibered over some hyperbolic LsuGL(d,)L_{su}\in{\rm GL}(d,\mathbb{Z}) in the sense that there is some Hölder Φ:XΓ𝕋d\Phi:X_{\Gamma}\to\mathbb{T}^{d} such that Φ(fx)=LsuΦ(x)\Phi(fx)=L_{su}\Phi(x), Wc(x)=Φ1(Φ(x))W^{c}(x)=\Phi^{-1}(\Phi(x)) and Φ\Phi is homotopic to the projection π:XΓ𝕋d\pi:X_{\Gamma}\to\mathbb{T}^{d},

  5. (v)

    there is a finite index subgroup Zfix(f)Z(f)Z_{\rm fix}^{\infty}(f)\leq Z^{\infty}(f) such that if gZfix(f)g\in Z_{\rm fix}^{\infty}(f) we have Φ(gx)=BΦ(x)\Phi(gx)=B\Phi(x) where BGL(d,)B\in{\rm GL}(d,\mathbb{Z}) is defined by BΦ=ΦgB\Phi_{*}=\Phi_{*}g_{*}, BB is the induced map on homology if GG is non-abelian,

moreover, if GG is not abelian then

  1. (vi)

    ff is accessible.

Remark 7.

The assumption that ff_{*} has at least one rational eigenvalue is to remove derived-from-Anosov examples since these examples are not fibered.

Remark 8.

Properties (i)(i) and (ii)(ii) follow from [7].

Remark 9.

This Theorem is similar to the classification of partially hyperbolic diffeomorphisms on 33-dimensional manifolds by Hammerlindl and Hammerlindl-Potrie [34, 33, 35]. In fact, in dimension 33, using [8], Theorem 1.1 essentially reduces to the main results of [34, 33] (in [33] we must make the extra assumption that the linearization LGL(3,)L\in{\rm GL}(3,\mathbb{Z}) has at least one rational eigenvalue).

1.4. Improved regularity of topological conjugacies between higher rank actions

The conjugacy in Theorems A and B, case (iii)(iii), is produced in two steps. First, we construct a topological conjugacy and second we show that the topological conjugacy is CC^{\infty}. The second step is the content of the following theorem, that may be of independent interest.

Theorem 1.2.

Let XΓX_{\Gamma} be a nilmanifold and α0:kAff(XΓ)\alpha_{0}:\mathbb{Z}^{k}\to{\rm Aff}(X_{\Gamma}) a homomorphism. Suppose that α0\alpha_{0} is higher rank. If α:kDiff(XΓ)\alpha:\mathbb{Z}^{k}\to{\rm Diff}^{\infty}(X_{\Gamma}) is bi-Hölder conjugate to α0\alpha_{0} by H:XΓXΓH:X_{\Gamma}\to X_{\Gamma} and there is some 𝐧0k\mathbf{n}_{0}\in\mathbb{Z}^{k} such that f=α𝐧0f=\alpha^{\mathbf{n}_{0}} is partially hyperbolic and accessible with center dim(Ec)=dim(Eα0𝐧0c)\dim(E^{c})=\dim(E_{\alpha_{0}^{\mathbf{n}_{0}}}^{c}), then HH is a CC^{\infty}-diffeomorphism.

Theorem 1.2 is a generalization of the global rigidity result by F. Rodriguez Hertz and Z. Wang [55] to some higher rank partially hyperbolic actions. In fact, large parts of the results in [55] generalize immediately to partially hyperbolic actions. One exception is that Rodriguez Hertz and Wang use a characterization of Anosov diffeomorphisms due to Mañé [48], to show that many elements of the action are Anosov. This characterization can not be applied in the partially hyperbolic setting. We also change some technical aspects of the proof, removing the use of Pesin theory.

1.5. Description of proofs

Let GG and α:k×XΓXΓ\alpha:\mathbb{Z}^{k}\times X_{\Gamma}\to X_{\Gamma} be as in Theorem A and f=α𝐧0f=\alpha^{\mathbf{n}_{0}} the partially hyperbolic element. By considering the Lie algebra of GG, 𝔤\mathfrak{g}, it is immediate that GG takes the form G=Hn×mG=H^{n}\times\mathbb{R}^{m}, n0n\neq 0, where HnH^{n} is the (2n+1)(2n+1)-dimensional Heisenberg group (for n=0n=0 we will consider HnH^{n} as the trivial group 11). Moreover, under the assumptions of Theorem B GG has to be a Heisenberg group HnH^{n} for some nn (this follows since any lattice Γ\Gamma in Hn×mH^{n}\times\mathbb{R}^{m} is, virtually, a product lattice so irreducibility of (f0)(f_{0})_{*} guarantee that either n=0n=0 or m=0m=0). The proof of Theorem A is divided into three steps, first we show that any ff as in Theorem A is fibered, then we show that any action α\alpha as in A is topologically conjugated to some affine model and finally we show that the topological conjugacy can be improved to a smooth conjugacy.

1.5.1. Step 1

The first part of the proof of Theorem A is to show that any element ff as in the theorem has to fiber over a hyperbolic automorphism of the torus and that ff is accessible. Both of these properties are contents of Theorem 1.1. First, we obtain a map Φ\Phi on GG, which is a contender for being the map in Theorem 1.1, this is done as in [27]. Second, we show that Φ\Phi is injective on the lifted stable and unstable leaves W^σ\hat{W}^{\sigma}, σ=s,u\sigma=s,u. By invariance of domain, this implies that the stable and unstable distributions of ff satisfy dim(Eσ)dim(E0σ)\dim(E^{\sigma})\leq\dim(E_{0}^{\sigma}), where E0σE_{0}^{\sigma} is the corresponding distribution for the linearization LL of ff. Since ff has center EcE^{c} of dimension 11 by assumption we conclude that dim(Eσ)=dim(E0σ)\dim(E^{\sigma})=\dim(E_{0}^{\sigma}) for σ=s,c,u\sigma=s,c,u. This shows, see Lemma 3.1, that Φ\Phi descends to a map Φ:XΓ𝕋d\Phi:X_{\Gamma}\to\mathbb{T}^{d}. Showing that Φ\Phi gives a fiber bundle structure as in Theorem 1.1 is then similar to [7].

Proving accessibility uses a topological argument. Since XΓX_{\Gamma} does not have a virtually abelian fundamental group, and since the kernel of the induced map

(1.7) Φ=π:π1XΓ=Γd\displaystyle\Phi_{*}=\pi_{*}:\pi_{1}X_{\Gamma}=\Gamma\to\mathbb{Z}^{d}

is the center of Γ\Gamma, there can not exist a connected compact set KXΓK\subset X_{\Gamma} such that Φ:XΓK𝕋d\Phi:X_{\Gamma}\supset K\to\mathbb{T}^{d} is a finite covering map. This implies, in particular, that ff can not have a compact susu-leaf. So, the proof of accessibility reduces to proving that if ff has a non-open accessibility class, then there is a compact susu-leaf. Obtaining a compact susu-leaf is done by studying the holonomies between center leaves in the universal cover induced by the fundamental group, as in [52, map TnT_{n} defined on page 71]. If KxGW^c(x)K_{x}\subset G\cap\hat{W}^{c}(x) is the closed set such that yKxy\in K_{x} does not have open accessibility class, then there is an action of Γ\Gamma on KxK_{x} defined by mapping yy to the unique intersection between the accessibility class of γy\gamma y and KxK_{x}. We show that the induced Γ\Gamma-action on the image of KxK_{x} in W^c(x)/[Γ,Γ]\hat{W}^{c}(x)/[\Gamma,\Gamma] has a fixed point if KxK_{x} is non-empty. This fixed point corresponds to a compact susu-leaf, which gives a contradiction so KxK_{x} must be empty.

1.5.2. Step 2

The remainder of the proof of Theorem A follows an idea by Spatzier and Vinhage [61]: instead of producing the conjugacy directly, we produce a homogeneous structure on XΓX_{\Gamma} that is compatible with α\alpha. The homogeneous structure on XΓX_{\Gamma} is obtained as the action of a certain quotient of the su-path group (see Section 4). We use the map Φ\Phi to define the susu-path group, 𝒫\mathcal{P}, and a natural action of 𝒫\mathcal{P} on XΓX_{\Gamma}. The most technical part of the paper is the following theorem from Section 6.

Theorem 1.3.

The system ff has a unique measure of maximal entropy μ\mu satisfying λμc=0\lambda_{\mu}^{c}=0.

Using Theorem 1.3 we show that there is a normal subgroup, 𝒩\mathcal{N}, of 𝒫\mathcal{P} such that N=𝒫/𝒩N=\mathcal{P}/\mathcal{N} is a Nilpotent Lie group that act transitively and freely on GG. Moreover, the action of NN is constructed such that α\alpha is compatible with the NN-action in the sense that the joint action of NkN\rtimes\mathbb{Z}^{k} is through a semi-direct product. We then use the NN-action to produce coordinates on XΓX_{\Gamma}, in which ff is affine. These coordinates gives a bi-Hölder conjugacy HH from α\alpha to some affine action α0\alpha_{0}.

1.5.3. Step 3

We finish the proof of Theorem A by proving Theorem 1.2, improving the regularity of HH from bi-Hölder to CC^{\infty}. The proof is similar to the proof in [55]. We begin the proof by using results from Wilkinson [70] to show that the conjugacy HH is smooth along the center WcW^{c}. The proof of Theorem 1.2 then follows [55] to show that the component of the conjugacy along some coarse exponent [χ][\chi] defining a chamber wall for the chamber that contains 𝐧0\mathbf{n}_{0} (see Section 2.2) is smooth. That is, we show that the [χ][\chi]-component of the bi-Hölder conjugacy HH restricted to Ws(x)W^{s}(x) and normalized by xex\mapsto e, denoted Hx[χ]:Ws(x)G[χ]H_{x}^{[\chi]}:W^{s}(x)\to G^{[\chi]} where G[χ]G^{[\chi]} is the coarse group with Lie algebra E0[χ]E_{0}^{[\chi]}, is uniformly CC^{\infty}. Once we know that Hx[χ]H_{x}^{[\chi]} is smooth, we study the map

(1.8) P:Gr(Es),P(x,V)=det(DxHx[χ]|V)\displaystyle P:{\rm Gr}^{\ell}(E^{s})\to\mathbb{R},\quad P(x,V)=\det\left(D_{x}H_{x}^{[\chi]}|_{V}\right)

where =dim(E0[χ])\ell=\dim(E_{0}^{[\chi]}) and the determinant is calculated with respect to some background Riemannian metric. The main observation is that P(x,V)P(x,V) can not vanish for all VGrx(Es)V\in{\rm Gr}_{x}^{\ell}(E^{s}) for any xXΓx\in X_{\Gamma} (see Lemma 8.7). This shows that Hx[χ]H_{x}^{[\chi]} is a submersion for every xx, so its fibers form a CC^{\infty}-foliation within WsW^{s}, denoted WssW^{ss}. Finally, we construct a α\alpha-invariant distribution E[χ]E^{[\chi]} transverse to the distribution Ess=TWssE^{ss}=TW^{ss} by using a graph transform argument. Existence of the distributions E[χ]E^{[\chi]} and EssE^{ss} allows us to produce new partially hyperbolic elements of the action α\alpha in a Weyl chamber adjacent to the Weyl chamber containing the first partially hyperbolic element 𝐧0\mathbf{n}_{0}. By induction we produce a partially hyperbolic element in every Weyl chamber. Using that α\alpha contains many partially hyperbolic elements it follows that Hx[χ]H_{x}^{[\chi]} is uniformly CC^{\infty} for every coarse exponent [χ][\chi], so HH is uniformly CC^{\infty} along WsW^{s} and WuW^{u}. Since HH is uniformly CC^{\infty} along WsW^{s}, WuW^{u}, and WcW^{c} we can apply Journé’s lemma twice to show that HH is CC^{\infty}.

Theorem B follows from Theorem A and results in [21].

1.6. Outline of paper

In Section 2 we go through some of the background results, and basic definitions from partially hyperbolic dynamics and higher rank actions on nilmanifolds. In Section 3 we prove Theorem 1.1. In Section 4 we introduce the susu-path group, one of the main objects in this paper, and show some of its basic properties. In Section 5 we recall the suspension construction of an abelian action and use it, combined with results from [3], to derive an invariance principle for higher rank actions on nilmanifolds. Section 6 is the most technical part of the paper, here we prove Theorem 1.3. In Section 7 we prove that the action α\alpha in Theorem A is topologically conjugated to some affine action. In Section 8 we prove Theorem 1.2, showing that the topological conjugacy is CC^{\infty}. Finally, in Section 9 we complete the proofs of Theorems A and B. We also include an appendix, Appendix A, proving some basic properties of higher rank, abelian algebraic actions on nilmanifolds.

1.7. Acknowledgements

The author thanks Danijela Damjanović, Homin Lee, Kurt Vinhage, Amie Wilkinson and Disheng Xu for useful discussion.

2. Background and definitions

2.1. Partially hyperbolic diffeomorphisms

Let MM be a smooth closed manifold and fDiff(M)f\in{\rm Diff}^{\infty}(M) a diffeomorphism. We fix a smooth metric gg on MM inducing a norm \left\lVert\cdot\right\rVert. We say that ff is (absolutely) partially hyperbolic if there is a continuous DfDf-invariant splitting

(2.1) TxM=Es(x)Ec(x)Eu(x)\displaystyle T_{x}M=E^{s}(x)\oplus E^{c}(x)\oplus E^{u}(x)

and constants ν,γ,γ^,ν^(0,1)\nu,\gamma,\widehat{\gamma},\widehat{\nu}\in(0,1), n0n_{0}\in\mathbb{N} such that for nn0n\geq n_{0}

(2.2) Dfn|Esνn<γn(Dfn)1|Ec1,\displaystyle\left\lVert Df^{n}|_{E^{s}}\right\rVert\leq\nu^{n}<\gamma^{n}\leq\left\lVert\left(Df^{n}\right)^{-1}|_{E^{c}}\right\rVert^{-1},
(2.3) Dfn|Ecγ^n<ν^n(Dfn|Eu)11.\displaystyle\left\lVert Df^{n}|_{E^{c}}\right\rVert\leq\widehat{\gamma}^{-n}<\widehat{\nu}^{-n}\leq\left\lVert\left(Df^{n}|_{E^{u}}\right)^{-1}\right\rVert^{-1}.

If we can choose the constants such that

(2.4) ν<γγ^r,ν^<γrγ^\displaystyle\nu<\gamma\widehat{\gamma}^{r},\quad\widehat{\nu}<\gamma^{r}\widehat{\gamma}

then we say that ff is rr-bunching. The distributions EsE^{s}, EcE^{c} and EuE^{u} are the stable, center and unstable distributions respectively.

Let fDiff(M)f\in{\rm Diff}^{\infty}(M) be partially hyperbolic. The stable and unstable distributions are always uniquely integrable to foliations WsW^{s} and WuW^{u} with uniformly CC^{\infty} leaves, but the center distribution may fail to be integrable. A sufficient condition for EcE^{c} being integrable is dynamical coherence. We say that ff is dynamically coherent if Ecs=EcEsE^{cs}=E^{c}\oplus E^{s} and Ecu=EcEuE^{cu}=E^{c}\oplus E^{u} are both integrable to foliations WcsW^{cs} and WcuW^{cu}. In this case we obtain a foliation tangent to EcE^{c} by intersecting Wc=WcsWcuW^{c}=W^{cs}\cap W^{cu}. We will denote the distance between two points p,qWσ(x)p,q\in W^{\sigma}(x), σ=s,c,u,cs,cu\sigma=s,c,u,cs,cu, in the leaf metric by dσ(p,q){\rm d}_{\sigma}(p,q). Denote the ball about xx of radius ε\varepsilon in dσ{\rm d}_{\sigma} by Wεσ(x)W_{\varepsilon}^{\sigma}(x). If ff is rr-bunching and dynamically coherent then WcsW^{cs}, WcuW^{cu} and WcW^{c} have uniformly CrC^{r} leaves [36] (or [20, Theorem 7]).

Let M^\hat{M} be the universal cover of MM, \mathcal{F} a foliation on MM, and ^\hat{\mathcal{F}} the lifted foliation to M^\hat{M}.

Definition 2.1.

We say that a continuous foliation with C1C^{1}-leaves \mathcal{F} of MM have quasi-isometric leaves in the universal cover if there is a constant Q1Q\geq 1 such that

(2.5) d(x,y)d(x,y)Qd(x,y),x,y^(p),\displaystyle{\rm d}(x,y)\leq{\rm d}_{\mathcal{F}}(x,y)\leq Q{\rm d}(x,y),\quad x,y\in\hat{\mathcal{F}}(p),

where d{\rm d}_{\mathcal{F}} is the metric along \mathcal{F}.

Remark 10.

The inequality d(x,y)d(x,y){\rm d}(x,y)\leq{\rm d}_{\mathcal{F}}(x,y) is immediate since any path connecting xx and yy along ^(p)\hat{\mathcal{F}}(p) also connect xx and yy in M^\hat{M}.

Remark 11.

We could have asked d(x,y)Ad(x,y)+B{\rm d}_{\mathcal{F}}(x,y)\leq A{\rm d}(x,y)+B in the definition, but this is equivalent to Definition 2.1 since d{\rm d}_{\mathcal{F}} and d{\rm d} are comparable on small balls in \mathcal{F} if \mathcal{F} have uniformly C1C^{1}-leaves.

In particular, if f:MMf:M\to M is partially hyperbolic (and dynamically coherent), we can apply Definition 2.1 to the lifted foliations W^σ\hat{W}^{\sigma}, σ=s,c,u,cs,cu\sigma=s,c,u,cs,cu. We can also lift the distributions EσE^{\sigma} on MM to distributions on M^\hat{M}, also denoted EσE^{\sigma}.

Assume now that f:MMf:M\to M is dynamically coherent. Since EsE^{s} is uniformly transverse to EcuE^{cu}, WsW^{s} and WcuW^{cu} have a local product structure. Similarly, WcW^{c} and WuW^{u} subfoliate WcuW^{cu} and EcE^{c} is transverse to EuE^{u}, so the foliations WcW^{c} and WuW^{u} have a local product structure in WcuW^{cu}.

Definition 2.2.

We say that ff have global product structure [32] if

(2.6) #W^cs(x)W^u(y)=1,x,yM^,\displaystyle\#\hat{W}^{cs}(x)\cap\hat{W}^{u}(y)=1,\quad x,y\in\hat{M},
(2.7) #W^s(x)W^cu(y)=1,x,yM^,\displaystyle\#\hat{W}^{s}(x)\cap\hat{W}^{cu}(y)=1,\quad x,y\in\hat{M},
(2.8) #W^s(x)W^c(y)=1,x,yW^cs(p),\displaystyle\#\hat{W}^{s}(x)\cap\hat{W}^{c}(y)=1,\quad x,y\in\hat{W}^{cs}(p),
(2.9) #W^c(x)W^u(y)=1,x,yW^cu(p).\displaystyle\#\hat{W}^{c}(x)\cap\hat{W}^{u}(y)=1,\quad x,y\in\hat{W}^{cu}(p).

When ff has global product structure we define global holonomy maps in the universal cover M^\hat{M}. Given xM^x\in\hat{M} and yW^u(x)y\in\hat{W}^{u}(x) we define

(2.10) πx,yu:W^cs(x)W^cs(y),{πx,yu(z)}=W^u(z)W^cs(y).\displaystyle\pi_{x,y}^{u}:\hat{W}^{cs}(x)\to\hat{W}^{cs}(y),\quad\{\pi_{x,y}^{u}(z)\}=\hat{W}^{u}(z)\cap\hat{W}^{cs}(y).
Wu(q)W^{u}(q)Wc(x)W^{c}(x)Wc(y)W^{c}(y)Wu(p)W^{u}(p)Wu(r)W^{u}(r)ppqqrrπx,yu(p)\pi_{x,y}^{u}(p)πx,yu(r)\pi_{x,y}^{u}(r)πx,yu(q)\pi_{x,y}^{u}(q)
Figure 1. Unstable holonomy between Wc(x)W^{c}(x) and Wc(y)W^{c}(y), yWu(x)y\in W^{u}(x).

Since W^u\hat{W}^{u} and W^c\hat{W}^{c} subfoliate W^cu\hat{W}^{cu} the holonomy maps πx,yu\pi_{x,y}^{u} restricts to maps πx,yu:W^c(x)W^c(y)\pi_{x,y}^{u}:\hat{W}^{c}(x)\to\hat{W}^{c}(y). The holonomy maps between center leaves πx,yu|W^c(x)\pi_{x,y}^{u}|_{\hat{W}^{c}(x)} descend to holonomy maps between center leaves on MM (Figure 1). Similarly we define stable holonomies πx,ys:Wc(x)Wc(y)\pi_{x,y}^{s}:W^{c}(x)\to W^{c}(y) when yWs(x)y\in W^{s}(x). When considering holonomies between center leaves then the holonomy maps are C1+αC^{1+\alpha} [9], and if ff is rr-bunching then the holonomies πx,yσ:Wc(x)Wc(y)\pi_{x,y}^{\sigma}:W^{c}(x)\to W^{c}(y), σ=s,u\sigma=s,u, are CrC^{r} [50].

We say that a path γ:[0,1]M\gamma:[0,1]\to M is an susu-path if [0,1][0,1] has a subdivision 0=t0<t1<<tN1<tN=10=t_{0}<t_{1}<...<t_{N-1}<t_{N}=1 such that Im(γ|[tj,tj+1]){\rm Im}\left(\gamma|_{[t_{j},t_{j+1}]}\right) is entirely contained in either an WsW^{s}-leaf or a WuW^{u}-leaf. If any two points x,yMx,y\in M are connected by an susu-path, then we say that ff is accessible. A set EME\subset M is σ\sigma-saturated, σ=s,u\sigma=s,u, if xEx\in E implies Wσ(x)EW^{\sigma}(x)\subset E, and susu-saturated if it is ss and uu-saturated. Equivalently ff is accessible if the only susu-saturated sets are MM and \emptyset. For xMx\in M (or M^\hat{M}) we define the accessibility class of xx

(2.11) AC(x)={yM : there is an supath connecting x and y}.\displaystyle{\rm AC}(x)=\{y\in M\text{ : there is an }su-\text{path connecting }x\text{ and }y\}.

Define a closed set Λ(f)\Lambda(f) by xΛ(f)x\in\Lambda(f) if AC(x){\rm AC}(x) is not open. That is

(2.12) Λ(f)=[AC(x) is openAC(x)]c=AC(x) is openAC(x)c\displaystyle\Lambda(f)=\left[\bigcup_{{\rm AC}(x)\text{ is open}}{\rm AC}(x)\right]^{c}=\bigcap_{{\rm AC}(x)\text{ is open}}{\rm AC}(x)^{c}

If ff has 11-dimensional center direction, then Λ(f)\Lambda(f) is laminated by accessibility classes [54, Proposition A.3]. In particular, if ff has 11-dimensional center and Λ(f)=M\Lambda(f)=M then EsEuE^{s}\oplus E^{u} is jointly integrable to some continuous foliation WsuW^{su} with smooth leaves (in fact, the foliation WsuW^{su} will be a CrC^{r}-foliation if ff is rr-bunching). In the other extreme, ff is accessible if and only if Λ(f)=\Lambda(f)=\emptyset, in this case ff has a unique accessibility class.

2.2. Nilmanifolds and higher rank actions

Let GG be a (simply connected) Lie group with Lie algebra 𝔤\mathfrak{g}. We define the lower central series of 𝔤\mathfrak{g} inductively as

(2.13) 𝔤(1)=𝔤,𝔤(j+1)=[𝔤(j),𝔤].\displaystyle\mathfrak{g}^{(1)}=\mathfrak{g},\quad\mathfrak{g}^{(j+1)}=[\mathfrak{g}^{(j)},\mathfrak{g}].

If there is \ell such that 𝔤(+1)=0\mathfrak{g}^{(\ell+1)}=0 then we say that 𝔤\mathfrak{g} is nilpotent and the minimal \ell satisfying 𝔤(+1)=0\mathfrak{g}^{(\ell+1)}=0 is the step of 𝔤\mathfrak{g}. We say that GG is a \ell-step nilpotent Lie group if 𝔤\mathfrak{g} is \ell-step nilpotent. Given a discrete subgroup ΓG\Gamma\leq G, we say that Γ\Gamma is a lattice if the quotient space ΓG\Gamma\setminus G carries a finite Haar measure μΓ\mu_{\Gamma}. Equivalently, for nilpotent groups [11, Corollary 5.4.6], a discrete subgroup ΓG\Gamma\leq G is a lattice if the quotient ΓG\Gamma\setminus G is compact.

If GG is simply connected, nilpotent, and ΓG\Gamma\leq G is a lattice then we define the associated compact nilmanifold by

(2.14) XΓ=ΓG.\displaystyle X_{\Gamma}=\Gamma\setminus G.

Denote by μΓ\mu_{\Gamma} the normalized Haar measure on XΓX_{\Gamma} and write

(2.15) pΓ:GXΓ\displaystyle p_{\Gamma}:G\to X_{\Gamma}

for the natural projection map. If GG is \ell-step then there is a sequence of (normal) subgroups

(2.16) G=G(0)G(1)G(1)G()=e,G(j)=exp(𝔤(j)).\displaystyle G=G^{(0)}\unrhd G^{(1)}\unrhd...\unrhd G^{(\ell-1)}\unrhd G^{(\ell)}=e,\quad G^{(j)}=\exp\left(\mathfrak{g}^{(j)}\right).

The intersection Γ(j)=G(j)Γ\Gamma^{(j)}=G^{(j)}\cap\Gamma defines a lattice in G(j)G^{(j)} [11, Theorem 5.2.3], define

(2.17) XΓ(j)=XΓ/G(j),π(j):XΓXΓ(j),0j\displaystyle X_{\Gamma}^{(j)}=X_{\Gamma}/G^{(j)},\quad\pi^{(j)}:X_{\Gamma}\to X_{\Gamma}^{(j)},\quad 0\leq j\leq\ell

where XΓ(j)X_{\Gamma}^{(j)} is a compact nilmanifold and π(j):XΓXΓ(j)\pi^{(j)}:X_{\Gamma}\to X_{\Gamma}^{(j)} is a fiber bundle. In particular, if GG is 22-step then we get one (non-trivial) projection map π(2):XΓXΓ(2)\pi^{(2)}:X_{\Gamma}\to X_{\Gamma}^{(2)}. Since G(2)=[G,G]G^{(2)}=[G,G], G/G(2)=G/[G,G]G/G^{(2)}=G/[G,G] is abelian, so XΓ(2)X_{\Gamma}^{(2)} is a torus. In the case of 22-step nilmanifolds, write

(2.18) π:XΓ𝕋d\displaystyle\pi:X_{\Gamma}\to\mathbb{T}^{d}

for the projection, with 𝕋d\mathbb{T}^{d} the base of XΓX_{\Gamma}. The fibers of π\pi are G(2)/Γ(2)G^{(2)}/\Gamma^{(2)} which is also a torus. So π:XΓ𝕋d\pi:X_{\Gamma}\to\mathbb{T}^{d} is a fiber bundle with base and fibers both tori, but XΓX_{\Gamma} is not a torus.

We define the automorphism and affine group of XΓX_{\Gamma} by

(2.19) Aut(XΓ)={LAut(G) : LΓ=Γ},\displaystyle{\rm Aut}(X_{\Gamma})=\{L\in{\rm Aut(G)}\text{ : }L\Gamma=\Gamma\},
(2.20) Aff(XΓ)={f0(x)=L(x)g1 : LAut(XΓ), gG}.\displaystyle{\rm Aff}(X_{\Gamma})=\{f_{0}(x)=L(x)g^{-1}\text{ : }L\in{\rm Aut}(X_{\Gamma}),\text{ }g\in G\}.

There is a natural map Aff(XΓ)Aut(XΓ){\rm Aff}(X_{\Gamma})\to{\rm Aut}(X_{\Gamma}) defined by mapping f0(x)=L(x)g1f_{0}(x)=L(x)g^{-1} to the automorphism LL, and each projection π(j)\pi^{(j)} induce a map Aut(XΓ)Aut(XΓ(j)){\rm Aut}(X_{\Gamma})\to{\rm Aut}(X_{\Gamma}^{(j)}).

Fix a homomorphism

(2.21) ρ:kAut(XΓ),ρ𝐧:XΓXΓ, 𝐧k.\displaystyle\rho:\mathbb{Z}^{k}\to{\rm Aut}(X_{\Gamma}),\quad\rho^{\mathbf{n}}:X_{\Gamma}\to X_{\Gamma},\text{ }\mathbf{n}\in\mathbb{Z}^{k}.

We say that ρ\rho has a rank1-1 factor if there is a nilpotent group G^\hat{G}, of positive dimension, a surjective homomorphism ϕ:GG^\phi:G\to\hat{G} such that Γ^=ϕΓ\hat{\Gamma}=\phi\Gamma is a lattice in G^\hat{G} and an automorphism LAut(X^Γ^)L\in{\rm Aut}(\hat{X}_{\hat{\Gamma}}) such that for some finite index subgroup Λk\Lambda\subset\mathbb{Z}^{k} we have n:Λn:\Lambda\to\mathbb{Z} satisfying ϕρ𝐧=Ln(𝐧)ϕ\phi\rho^{\mathbf{n}}=L^{n(\mathbf{n})}\phi, 𝐧Λ\mathbf{n}\in\Lambda. That is, ρ\rho has a rank1-1 factor if there is a factor of XΓX_{\Gamma} where the projected action of k\mathbb{Z}^{k} is a 1\mathbb{Z}^{1}-action (up to finite index). More generally, if α0:kAff(XΓ)\alpha_{0}:\mathbb{Z}^{k}\to{\rm Aff}(X_{\Gamma}) is a homomorphism then α0\alpha_{0} has a rank1-1 factor if

(2.22) kα0Aff(XΓ)Aut(XΓ)\displaystyle\mathbb{Z}^{k}\xrightarrow{\alpha_{0}}{\rm Aff}(X_{\Gamma})\xrightarrow{}{\rm Aut}(X_{\Gamma})

has a rank1-1 factor.

Definition 2.3.

A homomorphism α0:kAff(XΓ)\alpha_{0}:\mathbb{Z}^{k}\to{\rm Aff}(X_{\Gamma}) is higher rank if it has no rank1-1 factor.

Let α0:kAff(XΓ)\alpha_{0}:\mathbb{Z}^{k}\to{\rm Aff}(X_{\Gamma}) be a homomorphism. We say that χ:k\chi:\mathbb{Z}^{k}\to\mathbb{R} is a Lyapunov exponent of α0\alpha_{0} if there is v𝔤0v\in\mathfrak{g}\setminus 0 such that

(2.23) χ(𝐧)=lim±1logα0𝐧(v).\displaystyle\chi(\mathbf{n})=\lim_{\ell\to\pm\infty}\frac{1}{\ell}\log\left\lVert\alpha_{0}^{\ell\mathbf{n}}(v)\right\rVert.

The Lyapunov exponents χ\chi are linear and extends uniquely to k\mathbb{R}^{k}, we will consider Lyapunov exponents as linear maps on k\mathbb{R}^{k}. The Lyapunov space associated to χ\chi, E0χ𝔤E_{0}^{\chi}\leq\mathfrak{g}, is the subspace where Equation 2.23 hold. For the 00-functional we write E0cE_{0}^{c}. Note that

(2.24) 𝔤=E0cχ0E0χ.\displaystyle\mathfrak{g}=E_{0}^{c}\oplus\bigoplus_{\chi\neq 0}E_{0}^{\chi}.

Denote the set of (non-zero) Lyapunov exponents for α0\alpha_{0} by

(2.25) Lyap(α0)={χ0 : χ is a Lyapunov exponent of α0}.\displaystyle{\rm Lyap}(\alpha_{0})=\{\chi\neq 0\text{ : }\chi\text{ is a Lyapunov exponent of }\alpha_{0}\}.

For χLyap(α0)\chi\in{\rm Lyap}(\alpha_{0}), define the associated coarse exponent and coarse space by

(2.26) [χ]={χLyap(α0) : χ=cχ, for some c>0},E0[χ]=χ[χ]E0χ.\displaystyle[\chi]=\{\chi^{\prime}\in{\rm Lyap}(\alpha_{0})\text{ : }\chi^{\prime}=c\chi,\text{ for some }c>0\},\quad E_{0}^{[\chi]}=\bigoplus_{\chi^{\prime}\in[\chi]}E_{0}^{\chi^{\prime}}.

If χ(𝐧)>0\chi(\mathbf{n})>0 (or χ(𝐧)<0\chi(\mathbf{n})<0) then χ(𝐧)>0\chi^{\prime}(\mathbf{n})>0 (or χ(𝐧)<0\chi(\mathbf{n})<0) for every χ[χ]\chi^{\prime}\in[\chi] so we define [χ](𝐧)[\chi](\mathbf{n}) as the sign of χ(𝐧)\chi(\mathbf{n}) (or as 0 if χ(𝐧)=0\chi(\mathbf{n})=0). We also define ker[χ]=kerχ\ker[\chi]=\ker\chi.

Definition 2.4.

Let {[χ1],,[χN]}\{[\chi_{1}],...,[\chi_{N}]\} be the coarse exponents of α0\alpha_{0} and

(2.27) U=(j=1Nker[χj])c.\displaystyle U=\left(\bigcup_{j=1}^{N}\ker[\chi_{j}]\right)^{c}.

Each connected component 𝒞\mathcal{C} of UU is a Weyl chamber of α0\alpha_{0}. The kernels ker[χ]\ker[\chi] are Weyl chamber walls. A wall ker[χ]\ker[\chi] is a wall of 𝒞\mathcal{C} if dim𝒞¯ker[χ]=k1\dim\overline{\mathcal{C}}\cap\ker[\chi]=k-1.

Two coarse exponents, [χ][\chi] and [η][\eta], are dependent if [χ](𝐧)=[η](𝐧)[\chi](\mathbf{n})=-[\eta](\mathbf{n}), otherwise the two exponents are independent. Given any two χ,χ′′[χ]\chi^{\prime},\chi^{\prime\prime}\in[\chi] it is immediate

(2.28) [E0χ,E0χ′′]E0χ+χ′′(with E0χ+χ′′=0 if χ+χ′′Lyap(α0))\displaystyle\left[E_{0}^{\chi^{\prime}},E_{0}^{\chi^{\prime\prime}}\right]\subset E_{0}^{\chi^{\prime}+\chi^{\prime\prime}}\quad\left(\text{with }E_{0}^{\chi^{\prime}+\chi^{\prime\prime}}=0\text{ if }\chi^{\prime}+\chi^{\prime\prime}\not\in{\rm Lyap}(\alpha_{0})\right)

so E0[χ]E_{0}^{[\chi]} is a subalgebra of 𝔤\mathfrak{g}. Define the associated group

(2.29) G[χ]G,G[χ]=exp(E0[χ]).\displaystyle G^{[\chi]}\leq G,\quad G^{[\chi]}=\exp\left(E_{0}^{[\chi]}\right).

In Section 8 we will use that every coarse group G[χ]G^{[\chi]} has a transverse normal subgroup in the stable group GsG^{s}. More precisely, if α0𝐧\alpha_{0}^{\mathbf{n}} has stable space E0sE_{0}^{s}, [χ](𝐧)<0[\chi](\mathbf{n})<0, and

(2.30) E0ss:=[η][χ][η](𝐧)<0E0[η]\displaystyle E_{0}^{ss}:=\bigoplus_{\begin{subarray}{c}[\eta]\neq[\chi]\\ [\eta](\mathbf{n})<0\end{subarray}}E_{0}^{[\eta]}

then E0ssE0sE_{0}^{ss}\leq E_{0}^{s} is an ideal in E0sE_{0}^{s} [55, Lemma 3.1]. Equivalently, the subgroup Gss=exp(E0ss)Gs=exp(E0s)G^{ss}=\exp(E_{0}^{ss})\leq G^{s}=\exp(E_{0}^{s}) is a normal subgroup.

The following two lemmas are well-known, we include proofs in Appendix A.

Lemma 2.1.

If α0:kAff(XΓ)\alpha_{0}:\mathbb{Z}^{k}\to{\rm Aff}(X_{\Gamma}) is higher rank then there are at least two independent coarse exponents.

Lemma 2.2.

If α0:kAff(XΓ)\alpha_{0}:\mathbb{Z}^{k}\to{\rm Aff}(X_{\Gamma}) is higher rank and [χ][\chi] is a coarse Lyapunov exponent then the space

V=[η]±[χ]E0[χ]\displaystyle V=\bigoplus_{[\eta]\neq\pm[\chi]}E_{0}^{[\chi]}

defines a minimal translation action on XΓX_{\Gamma} (the translation action by VV is the translation action of the exponential of the Lie algebra generated by VV).

Given a homomorphism α:kDiff(XΓ)\alpha:\mathbb{Z}^{k}\to{\rm Diff}^{\infty}(X_{\Gamma}), written α(𝐧)=α𝐧\alpha(\mathbf{n})=\alpha^{\mathbf{n}}, we obtain a linearization ρ:kAut(π1XΓ)Aut(XΓ)\rho:\mathbb{Z}^{k}\to{\rm Aut}(\pi_{1}X_{\Gamma})\cong{\rm Aut}(X_{\Gamma}).

Definition 2.5.

A smooth action α:kDiff(XΓ)\alpha:\mathbb{Z}^{k}\to{\rm Diff}^{\infty}(X_{\Gamma}) is higher rank if the linearization ρ:kAut(XΓ)\rho:\mathbb{Z}^{k}\to{\rm Aut}(X_{\Gamma}) is higher rank.

Fix n1n\geq 1, d=2nd=2n and define Hn:=n×n×H^{n}:=\mathbb{R}^{n}\times\mathbb{R}^{n}\times\mathbb{R}. With g=(q,p,z)Gg=(q,p,z)\in G and h=(q,p,z)Gh=(q^{\prime},p^{\prime},z^{\prime})\in G we define a multiplication

(2.31) gh=(q,p,z)(q,p,z)=(q+q,p+p,z+z+qp).\displaystyle gh=(q,p,z)(q^{\prime},p^{\prime},z^{\prime})=\left(q+q^{\prime},p+p^{\prime},z+z^{\prime}+q\cdot p^{\prime}\right).

This makes HnH^{n} into a group, the (d+1)(d+1)-dimensional Heisenberg group. Denote by ω\omega the symplectic form on d=nn\mathbb{R}^{d}=\mathbb{R}^{n}\oplus\mathbb{R}^{n}. The Lie bracket on 𝔤\mathfrak{g} is

(2.32) [(X,Z),(X,Z)]=(0,ω(X,X)),X,Xd Z,Z.\displaystyle\left[(X,Z),(X^{\prime},Z^{\prime})\right]=(0,\omega(X,X^{\prime})),\quad X,X^{\prime}\in\mathbb{R}^{d}\text{ }Z,Z^{\prime}\in\mathbb{R}.

Let ΓHn\Gamma\leq H^{n} be a lattice and XΓ=ΓHnX_{\Gamma}=\Gamma\setminus H^{n} the associated nilmanifold. For LAut(XΓ)L\in{\rm Aut}(X_{\Gamma}) we obtain a map LsuGL(d,)L_{su}\in{\rm GL}(d,\mathbb{Z}) by projecting onto the base, this element LsuL_{su} satisfies LsuSp(d,)L_{su}\in{\rm Sp}(d,\mathbb{Z}). In particular, if [χ][\chi] is a coarse exponent of α0:kAff(XΓ)\alpha_{0}:\mathbb{Z}^{k}\to{\rm Aff}(X_{\Gamma}) then [χ]-[\chi] is also a coarse exponent, so the coarse exponents come in negatively proportional pairs.

In the remainder, we will be interested in groups GG of the form G=×HnG=\mathbb{R}^{\ell}\times H^{n} for some 0\ell\geq 0 and n0n\geq 0. These groups constitute all abelian simply connected nilpotent groups and all 22-step, simply connected nilpotent Lie groups with dim[G,G]=1\dim[G,G]=1. Recall the Baker–Campbell–Hausdorff formula [11]

(2.33) eXeY=eX+Y+[X,Y]/2,X,Y𝔤.\displaystyle e^{X}e^{Y}=e^{X+Y+[X,Y]/2},\quad X,Y\in\mathfrak{g}.

Fix a left invariant metric, d{\rm d}, on GG. Using the Baker–Campbell–Hausdorff formula, it is immediate that for eZ=gc[G,G]e^{Z}=g_{c}\in[G,G], d(e,gc)4Z{\rm d}(e,g_{c})\leq 4\sqrt{\left\lVert Z\right\rVert}.

3. Some properties of partially hyperbolic diffeomorphisms with quasi isometric leaves in the universal cover

In this section, we prove Theorem 1.1. We begin by proving that Φ\Phi from Theorem 1.1 exists in Section 3.1. In section 3.2 we show that ff is accessible.

Let G=Hn×G=H^{n}\times\mathbb{R}^{\ell} be the product of some Heisenberg group and some abelian group, ΓG\Gamma\leq G a lattice and XΓX_{\Gamma} the associated compact nilmanifold. If n>0n>0 then we write d=2n+d=2n+\ell and let π:Gd\pi:G\to\mathbb{R}^{d} be the base projection. If n=0n=0 and we have an automorphism LAut(XΓ)L\in{\rm Aut}(X_{\Gamma}) with 11-dimensional center, then we let d=1d=\ell-1 and π:Gd\pi:G\to\mathbb{R}^{d} be the projection from GG onto G/E0cG/E_{0}^{c} (note that if LL has 11-dimensional center then the center direction E0cE_{0}^{c} is a rational line). Assume for the remainder of this section that fDiff(XΓ)f\in{\rm Diff}^{\infty}(X_{\Gamma}) satisfy all the assumptions of Theorem 1.1. Denote by LL the linearization of ff and LsuL_{su} the induced map on the base.

3.1. Existence of Franks-Manning coordinates

Write f:XΓXΓf:X_{\Gamma}\to X_{\Gamma} as

(3.1) fx=L(x)ev(x),\displaystyle fx=L(x)e^{-v(x)},

with v:XΓ𝔤v:X_{\Gamma}\to\mathfrak{g}. Fix a lift F:GGF:G\to G, Fx=L(x)ev(x)Fx=L(x)e^{-v(x)}. For xGx\in G let

(3.2) Fnx=:xn,n.\displaystyle F^{n}x=:x_{n},\quad n\in\mathbb{Z}.

Let E0sE_{0}^{s}, E0cE_{0}^{c} and E0uE_{0}^{u} be the stable, center and unstable direction of LL respectively. We decompose any v𝔤v\in\mathfrak{g} with respect to the splitting 𝔤=E0sE0cEu\mathfrak{g}=E_{0}^{s}\oplus E_{0}^{c}\oplus E^{u} as v=vs+vc+vuv=v_{s}+v_{c}+v_{u}. Denote by π~:GG/Gcd\tilde{\pi}:G\to G/G^{c}\cong\mathbb{R}^{d^{\prime}} the projection, where Gc=exp(E0c)G^{c}=\exp(E_{0}^{c}) is the center of LL (we do not know, a priori, that E0cE_{0}^{c} has dimension 11). Write A:ddA:\mathbb{R}^{d^{\prime}}\to\mathbb{R}^{d^{\prime}} for the map induced by LL, then AA is hyperbolic (if dimE0c1\dim E_{0}^{c}\neq 1 then ALsuA\neq L_{su}). Recall the following well-known lemma.

Lemma 3.1.

There exists a unique Hölder map Φ:GG/Gcd\Phi:G\to G/G^{c}\cong\mathbb{R}^{d^{\prime}}

(3.3) Φ(x)=π~(x)+φ(x),φ(γx)=φ(x), γΓ\displaystyle\Phi(x)=\tilde{\pi}(x)+\varphi(x),\quad\varphi(\gamma x)=\varphi(x),\text{ }\gamma\in\Gamma

such that Φ(Fx)=AΦ(x)\Phi(Fx)=A\Phi(x). If dim(E0c)=1\dim(E_{0}^{c})=1 then d=dd^{\prime}=d, Φ:Gd\Phi:G\to\mathbb{R}^{d} descends to a map Φ:XΓ𝕋d\Phi:X_{\Gamma}\to\mathbb{T}^{d} homotopic to π\pi and A=LsuA=L_{su}.

Proof.

The lemma follows from a calculation showing that φ\varphi satisfy vsu(x)=φ(fx)A(φ(x))v_{su}(x)=\varphi(fx)-A(\varphi(x)), which has a unique solution since AA is hyperbolic. If dim(E0c)=1\dim(E_{0}^{c})=1 then E0c=[𝔤,𝔤]E_{0}^{c}=[\mathfrak{g},\mathfrak{g}] if GG is non-abelian (since [𝔤,𝔤][\mathfrak{g},\mathfrak{g}] lie in the center of any automorphism) so G/GcG/G^{c} is the natural quotient by [G,G][G,G]. The lemma follows since φ\varphi is Γ\Gamma-invariant. If GG is abelian, then E0cE_{0}^{c} is some 11-dimensional rational line (since we assume that LsuL_{su} have at least one rational eigenvalue) and the last conclusion follows. ∎

Lemma 3.2.

If yW^σ(x)y\in\hat{W}^{\sigma}(x), σ=s,u\sigma=s,u, then Φ(y)=Φ(x)\Phi(y)=\Phi(x) if and only if x=yx=y. That is Φ:W^σ(x)Φ(x)+E0σ\Phi:\hat{W}^{\sigma}(x)\to\Phi(x)+E_{0}^{\sigma} is injective. Moreover, dimEσ=dimE0σ\dim E^{\sigma}=\dim E_{0}^{\sigma} for s,c,us,c,u, so Φ\Phi descends and A=LsuA=L_{su} is hyperbolic.

Remark 12.

We prove the lemma when GG is non-abelian. The proof when GG is abelian simplifies since all terms from brackets vanish.

Proof.

Assume that σ=u\sigma=u, for the other case we reverse time. Let yW^u(x)y\in\hat{W}^{u}(x). Write yn=xneγny_{n}=x_{n}e^{\gamma^{n}} where γn𝔤\gamma^{n}\in\mathfrak{g}. With respect to the decomposition 𝔤=E0sE0cE0u\mathfrak{g}=E_{0}^{s}\oplus E_{0}^{c}\oplus E_{0}^{u}, decompose γn=γsn+γcn+γun\gamma^{n}=\gamma_{s}^{n}+\gamma_{c}^{n}+\gamma_{u}^{n}. If Φ(x)=Φ(y)\Phi(x)=\Phi(y) then Φ(xn)=Φ(yn)\Phi(x_{n})=\Phi(y_{n}) for all n0n\geq 0, so π(xn)π(yn)2φC0=:C\left\lVert\pi(x_{n})-\pi(y_{n})\right\rVert\leq 2\left\lVert\varphi\right\rVert_{C^{0}}=:C independently of nn. On the other hand, π(xn)π(yn)2=γsn2+γun2\left\lVert\pi(x_{n})-\pi(y_{n})\right\rVert^{2}=\left\lVert\gamma_{s}^{n}\right\rVert^{2}+\left\lVert\gamma_{u}^{n}\right\rVert^{2}. So to show that Φ\Phi is injective on W^u(x)\hat{W}^{u}(x), it suffices to show that γun\gamma_{u}^{n}\to\infty as nn\to\infty if xyx\neq y.

Suppose for contradiction that xyx\neq y and γunK\left\lVert\gamma_{u}^{n}\right\rVert\leq K uniformly in nn. From our definitions

xn+1eγn+1=\displaystyle x_{n+1}e^{\gamma^{n+1}}= yn+1=F(yn)=L(yn)ev(yn)=L(xn)eL(γn)ev(yn)=\displaystyle y_{n+1}=F(y_{n})=L(y_{n})e^{-v(y_{n})}=L(x_{n})e^{L(\gamma^{n})}e^{-v(y_{n})}=
L(xn)eL(γn)v(yn)[L(γn),v(yn)]/2=\displaystyle L(x_{n})e^{L(\gamma^{n})-v(y_{n})-[L(\gamma^{n}),v(y_{n})]/2}=
F(xn)ev(xn)+L(γn)v(yn)([L(γn),v(yn)][v(xn),L(γn)]+[v(xn),v(yn)])/2\displaystyle F(x_{n})e^{v(x_{n})+L(\gamma^{n})-v(y_{n})-\left([L(\gamma^{n}),v(y_{n})]-[v(x_{n}),L(\gamma^{n})]+[v(x_{n}),v(y_{n})]\right)/2}

or if we take logarithms

(3.4) γn+1\displaystyle\gamma^{n+1} =v(xn)+L(γn)v(yn)\displaystyle=v(x_{n})+L(\gamma^{n})-v(y_{n})-
(3.5) [L(γn),v(yn)][v(xn),L(γn)]+[v(xn),v(yn)]2.\displaystyle\frac{[L(\gamma^{n}),v(y_{n})]-[v(x_{n}),L(\gamma^{n})]+[v(x_{n}),v(y_{n})]}{2}.

Using 3.5 we estimate

(3.6) γsn+1Lγsn+C,\displaystyle\left\lVert\gamma_{s}^{n+1}\right\rVert\leq\left\lVert L\gamma_{s}^{n}\right\rVert+C,
(3.7) γcn+1γcn+C+k(γsn+γun),\displaystyle\left\lVert\gamma_{c}^{n+1}\right\rVert\leq\left\lVert\gamma_{c}^{n}\right\rVert+C+k\left(\left\lVert\gamma_{s}^{n}\right\rVert+\left\lVert\gamma_{u}^{n}\right\rVert\right),

with constants C,K,kC,K,k that only depend on vC0\left\lVert v\right\rVert_{C^{0}}. Since LL is contracting on E0sE_{0}^{s} there is τ(0,1)\tau\in(0,1) such that γsn+1τγsn+C\left\lVert\gamma_{s}^{n+1}\right\rVert\leq\tau\left\lVert\gamma_{s}^{n}\right\rVert+C, or

(3.8) γsnC1τ\displaystyle\left\lVert\gamma_{s}^{n}\right\rVert\leq\frac{C}{1-\tau}

uniformly in nn. We have γunK\left\lVert\gamma_{u}^{n}\right\rVert\leq K by assumption, so for some possibly larger CC we obtain

(3.9) γcn+1γcn+C,γcnCn+γc0.\displaystyle\left\lVert\gamma_{c}^{n+1}\right\rVert\leq\left\lVert\gamma_{c}^{n}\right\rVert+C,\quad\left\lVert\gamma_{c}^{n}\right\rVert\leq Cn+\left\lVert\gamma_{c}^{0}\right\rVert.

After possibly enlarging CC again, we have

(3.10) d(xn,yn)=d(e,eγn)d(e,eγsn+γun)+d(e,eγcn)C(n+1).\displaystyle{\rm d}(x_{n},y_{n})={\rm d}(e,e^{\gamma^{n}})\leq{\rm d}(e,e^{\gamma_{s}^{n}+\gamma_{u}^{n}})+{\rm d}(e,e^{\gamma_{c}^{n}})\leq C(\sqrt{n}+1).

On the other hand, the assumption that W^u(xn)\hat{W}^{u}(x_{n}) is quasi-isometric implies that there is some λ>1\lambda>1 and Q1Q\geq 1 such that

(3.11) d(xn,yn)1Qdu(Fnx,Fny)1Qλndu(x,y).\displaystyle{\rm d}(x_{n},y_{n})\geq\frac{1}{Q}{\rm d}_{u}(F^{n}x,F^{n}y)\geq\frac{1}{Q}\lambda^{n}{\rm d}_{u}(x,y).

If du(x,y)0{\rm d}_{u}(x,y)\neq 0 then 3.10 and 3.11 gives a contradiction for nn sufficiently large, so x=yx=y.

Since Φ:W^σ(x)Φ(x)+E0σ\Phi:\hat{W}^{\sigma}(x)\to\Phi(x)+E_{0}^{\sigma}, σ=s,u\sigma=s,u, is injective it follows by invariance of domain that dim(Eσ)dim(E0σ)\dim(E^{\sigma})\leq\dim(E_{0}^{\sigma}). On the other hand, we have

(3.12) dim(Es)+dim(Eu)+1=d+1=dim(E0s)+dim(E0c)+dim(E0u)\displaystyle\dim(E^{s})+\dim(E^{u})+1=d+1=\dim(E_{0}^{s})+\dim(E_{0}^{c})+\dim(E_{0}^{u})

or 1dim(E0c)1\geq\dim(E_{0}^{c}), so dim(E0c)=1\dim(E_{0}^{c})=1. This implies dim(Es)+dim(Eu)=dim(E0u)+dim(E0s)\dim(E^{s})+\dim(E^{u})=\dim(E_{0}^{u})+\dim(E_{0}^{s}), which only hold if dim(Eσ)=dim(E0σ)\dim(E^{\sigma})=\dim(E_{0}^{\sigma}), σ=s,u\sigma=s,u, since dim(Eσ)dim(E0σ)\dim(E^{\sigma})\leq\dim(E_{0}^{\sigma}). That A=LsuA=L_{su} is hyperbolic follows by Lemma 3.1. ∎

Remark 13.

We will make no notational distinction between the map Φ:Gd\Phi:G\to\mathbb{R}^{d} and the induced map Φ:XΓ𝕋d\Phi:X_{\Gamma}\to\mathbb{T}^{d}.

Lemma 3.3.

For every xGx\in G the map Φσ,x:W^σ(x)E0σ\Phi_{\sigma,x}:\hat{W}^{\sigma}(x)\to E_{0}^{\sigma}, defined by

(3.13) Φσ,x(y)=Φ(y)Φ(x),\displaystyle\Phi_{\sigma,x}(y)=\Phi(y)-\Phi(x),

is a homeomorphism. For any yW^σ(x)y\in\hat{W}^{\sigma}(x), the map Φσ,x:W^Rσ(y)E0σ\Phi_{\sigma,x}:\hat{W}_{R}^{\sigma}(y)\to E_{0}^{\sigma} is uniformly bi-Hölder for fixed RR.

Proof.

By Lemma 3.2 Φ:W^σ(x)Φ(x)+E0σ\Phi:\hat{W}^{\sigma}(x)\to\Phi(x)+E_{0}^{\sigma} is injective so Φσ,x\Phi_{\sigma,x} is injective. Since W^σ(x)\hat{W}^{\sigma}(x) and E0σE_{0}^{\sigma} have the same dimension, it follows by invariance of domain that Φσ,x\Phi_{\sigma,x} has an open image and is homeomorphic onto its image. In particular, the image of Φσ,x\Phi_{\sigma,x} contain some ball Brxσ(0)B_{r_{x}}^{\sigma}(0) around 0 in E0σE_{0}^{\sigma} (to make rxr_{x} well-defined we take the maximal possible rxr_{x}). Given γΓ\gamma\in\Gamma we have

(3.14) Φσ,γx(γy)=Φ(γy)Φ(γx)=Φ(y)+π(γ)Φ(x)π(γ)=Φσ,x(y)\displaystyle\Phi_{\sigma,\gamma x}(\gamma y)=\Phi(\gamma y)-\Phi(\gamma x)=\Phi(y)+\pi(\gamma)-\Phi(x)-\pi(\gamma)=\Phi_{\sigma,x}(y)

from which it follows that xrxx\mapsto r_{x} is Γ\Gamma-invariant. Moreover, Φσ,x\Phi_{\sigma,x} and W^σ(x)\hat{W}^{\sigma}(x) vary continuously in xx, so rx>r0r_{x}>r_{0} is open. Combined with Γ\Gamma-invariance and the fact that XΓX_{\Gamma} is compact, we find r0>0r_{0}>0 such that rxr0r_{x}\geq r_{0} for all xGx\in G. Assume now σ=u\sigma=u, the other case follows by reversing time. We have

Φu,x(W^u(x))=\displaystyle\Phi_{u,x}(\hat{W}^{u}(x))= Φu,x(FnW^u(xn))=LnΦu,xn(W^u(xn))LnBr0u(0)\displaystyle\Phi_{u,x}(F^{n}\hat{W}^{u}(x_{-n}))=L^{n}\Phi_{u,x_{-n}}(\hat{W}^{u}(x_{-n}))\supset L^{n}B_{r_{0}}^{u}(0)

and letting nn\to\infty, using that LL expand E0uE_{0}^{u}, we obtain Φu,x(W^u(x))=E0u\Phi_{u,x}(\hat{W}^{u}(x))=E_{0}^{u}.

So Φu,x\Phi_{u,x} is a homeomorphism. Since Φ\Phi is Hölder, (x,y)Φu,x(y)(x,y)\mapsto\Phi_{u,x}(y) is Hölder in xx and yy. The set Φu,x1(B1u(0)¯)\Phi_{u,x}^{-1}\left(\overline{B_{1}^{u}(0)}\right) is compact in W^u(x)\hat{W}^{u}(x), so we define KxK_{x} as the minimal radius such that the closure of W^Kxu(x)\hat{W}_{K_{x}}^{u}(x) contain Φu,x1(B1u(0)¯)\Phi_{u,x}^{-1}\left(\overline{B_{1}^{u}(0)}\right). Since xΦu,xx\mapsto\Phi_{u,x} vary continuously the map xKxx\mapsto K_{x} also vary continuosly in xx. If K=supxKxK=\sup_{x}K_{x}, then KK is such that if z,wW^u(x)z,w\in\hat{W}^{u}(x) satisfy du(z,w)K{\rm d}_{u}(z,w)\geq K then Φ(z)Φ(w)1\left\lVert\Phi(z)-\Phi(w)\right\rVert\geq 1. Let μ=L|E0u>1\mu=\left\lVert L|_{E_{0}^{u}}\right\rVert>1 and λ>1\lambda>1 such that du(Fnw,Fnz)cλndu(z,w){\rm d}_{u}(F^{n}w,F^{n}z)\geq c\lambda^{n}{\rm d}_{u}(z,w). Let z,wW^u(x)z,w\in\hat{W}^{u}(x) satisfy du(z,w)1{\rm d}_{u}(z,w)\leq 1. There is n=n(z,w)0n=n(z,w)\geq 0 such that n(z,w)κlogdu(z,w)+Cn(z,w)\leq-\kappa\log{\rm d}_{u}(z,w)+C, du(zn,wn)K{\rm d}_{u}(z_{n},w_{n})\geq K and κ\kappa only depends on λ\lambda. We have

Φu,x(z)Φu,x(w)=\displaystyle\left\lVert\Phi_{u,x}(z)-\Phi_{u,x}(w)\right\rVert= Φ(z)Φ(w)=Ln(Φ(zn)Φ(wn))\displaystyle\left\lVert\Phi(z)-\Phi(w)\right\rVert=\left\lVert L^{-n}\left(\Phi(z_{n})-\Phi(w_{n})\right)\right\rVert\geq
μnΦ(zn)Φ(wn)μnμκlog(du(z,w))C=\displaystyle\mu^{-n}\left\lVert\Phi(z_{n})-\Phi(w_{n})\right\rVert\geq\mu^{-n}\geq\mu^{\kappa\log({\rm d}_{u}(z,w))-C}=
μCdu(z,w)κlog(μ).\displaystyle\mu^{-C}{\rm d}_{u}(z,w)^{\kappa\log(\mu)}.

Fix r>0r>0 such that Φu,x1(Bru(0))W^1u(x)\Phi_{u,x}^{-1}(B_{r}^{u}(0))\subset\hat{W}_{1}^{u}(x) for all xx. For v1,v2Bru(0)v_{1},v_{2}\in B_{r}^{u}(0) we obtain

(3.15) du(Φu,x1(v1),Φu,x1(v2))μC/κlog(μ)v1v21/κlog(μ)\displaystyle{\rm d}_{u}(\Phi_{u,x}^{-1}(v_{1}),\Phi_{u,x}^{-1}(v_{2}))\leq\mu^{C/\kappa\log(\mu)}\left\lVert v_{1}-v_{2}\right\rVert^{1/\kappa\log(\mu)}

so Φu,x1:Bru(0)W^u(x)\Phi_{u,x}^{-1}:B_{r}^{u}(0)\to\hat{W}^{u}(x) is uniformly Hölder. Given yW^u(x)y\in\hat{W}^{u}(x) and v0E0uv_{0}\in E_{0}^{u}

(3.16) Φu,y(Φu,x1(v+v0))=Φ(Φu,x1(v+v0))Φ(y)=v+v0+Φ(x)Φ(y)\displaystyle\Phi_{u,y}(\Phi_{u,x}^{-1}(v+v_{0}))=\Phi(\Phi_{u,x}^{-1}(v+v_{0}))-\Phi(y)=v+v_{0}+\Phi(x)-\Phi(y)

or if we choose Φ(y)Φ(x)=v0\Phi(y)-\Phi(x)=v_{0}, then

(3.17) Φu,x1(v+v0)=Φu,y1(v)\displaystyle\Phi_{u,x}^{-1}(v+v_{0})=\Phi_{u,y}^{-1}(v)

which shows that Φu,x1\Phi_{u,x}^{-1} is uniformly Hölder on any ball of radius rr. By covering any ball of radius RR with balls of radius rr, the lemma follows. ∎

Lemma 3.4.

If γ:[0,1]G\gamma:[0,1]\to G is a C1C^{1}-curve that is tangent to EcsE^{cs} (or EcuE^{cu}) then Φ(γ(1))Φ(γ(0))+E0s\Phi(\gamma(1))\in\Phi(\gamma(0))+E_{0}^{s} (or Φ(γ(1))Φ(γ(0))+E0u\Phi(\gamma(1))\in\Phi(\gamma(0))+E_{0}^{u}).

Proof.

Let γ\gamma be tangent to EcsE^{cs}. Write Hx=Φ1(Φ(x))H_{x}=\Phi^{-1}(\Phi(x)). For yHxy\in H_{x}

(3.18) π(yx)φ(y)φ(x)2φC0=C\displaystyle\left\lVert\pi(y-x)\right\rVert\leq\left\lVert\varphi(y)-\varphi(x)\right\rVert\leq 2\left\lVert\varphi\right\rVert_{C^{0}}=C

so Hxπ1BC(π(x))H_{x}\subset\pi^{-1}B_{C}(\pi(x)). In particular, if y=xeη(y)y=xe^{\eta(y)} for yHxy\in H_{x} then

(3.19) ηs(y)2+ηu(y)2C2.\displaystyle\left\lVert\eta_{s}(y)\right\rVert^{2}+\left\lVert\eta_{u}(y)\right\rVert^{2}\leq C^{2}.

Let ηn:W^u(x)𝔤\eta^{n}:\hat{W}^{u}(x)\to\mathfrak{g} be defined by yn=xneηn(y)y_{n}=x_{n}e^{\eta^{n}(y)}. Since

FnHx=FnΦ1(Φ(x))=Φ1(Φ(Fnx))=Hxn\displaystyle F^{n}H_{x}=F^{n}\Phi^{-1}(\Phi(x))=\Phi^{-1}(\Phi(F^{n}x))=H_{x_{n}}

we have ηsn(y),ηun(y)C\left\lVert\eta_{s}^{n}(y)\right\rVert,\left\lVert\eta_{u}^{n}(y)\right\rVert\leq C uniformly in nn\in\mathbb{Z}. As in the proof of Lemma 3.2 it follows that we also have ηcn(y)C|n|\left\lVert\eta_{c}^{n}(y)\right\rVert\leq C|n| for some (possibly larger) constant CC. Let γ:[0,1]G\gamma:[0,1]\to G be a C1C^{1}-curve tangent to EcsE^{cs} and denote the end-points of γ\gamma by x=γ(0)x=\gamma(0) and y=γ(1)y=\gamma(1). We find zW^s(y)z\in\hat{W}^{s}(y) and wW^u(z)w\in\hat{W}^{u}(z) such that wHxw\in H_{x} (this corresponds to choosing a two-legged susu-path from Φ(y)\Phi(y) to Φ(x)\Phi(x) in d\mathbb{R}^{d}). By the reverse triangle inequality

(3.20) d(xn,wn)d(wn,zn)d(xn,yn)d(yn,zn).\displaystyle{\rm d}(x_{n},w_{n})\geq{\rm d}(w_{n},z_{n})-{\rm d}(x_{n},y_{n})-{\rm d}(y_{n},z_{n}).

Since ww and zz lie in the same unstable leaf and W^u\hat{W}^{u} have quasi-isometric leaves, d(wn,zn)cλnd(w,z){\rm d}(w_{n},z_{n})\geq c\lambda^{n}{\rm d}(w,z) for some c>0c>0 and λ>1\lambda>1. But γ\gamma is a C1C^{1}-curve along EcsE^{cs} from xx to yy, so

(3.21) d(xn,yn)01Dγ(t)Fn(γ˙(t))dtCλ^n|γ|\displaystyle{\rm d}(x_{n},y_{n})\leq\int_{0}^{1}\left\lVert D_{\gamma(t)}F^{n}(\dot{\gamma}(t))\right\rVert{\rm d}t\leq C\hat{\lambda}^{n}\cdot|\gamma|

where |γ||\gamma| is the length of γ\gamma and λ^<λ\hat{\lambda}<\lambda. Finally, d(yn,zn)C{\rm d}(y_{n},z_{n})\leq C uniformly for n0n\geq 0 since yy and zz lie in the same stable leaf. Equation 3.20 implies

(3.22) cλnd(w,z)Cλ^n|γ|Cd(xn,wn)=d(e,eηn(w))Cn,n0\displaystyle c\lambda^{n}{\rm d}(w,z)-C\hat{\lambda}^{n}|\gamma|-C\leq{\rm d}(x_{n},w_{n})={\rm d}(e,e^{\eta^{n}(w)})\leq C\sqrt{n},\quad n\geq 0

for some constant CC. If d(w,z)>0{\rm d}(w,z)>0 then we obtain a contradiction for nn large enough. We conclude that d(w,z)=0{\rm d}(w,z)=0. That is, w=zw=z so W^s(y)Hx={w}\hat{W}^{s}(y)\cap H_{x}=\{w\}\neq\emptyset, which implies

(3.23) Φ(x)Φ(y)=Φ(w)Φ(y)=Φs,y(w)E0s\displaystyle\Phi(x)-\Phi(y)=\Phi(w)-\Phi(y)=\Phi_{s,y}(w)\in E_{0}^{s}

proving the lemma for scsc. The lemma follows for cucu by reversing time. ∎

We can now prove the first five points of Theorem 1.1

Proof of Theorem 1.1.

Dynamical coherence of ff follows from [7]. Any curve γ\gamma tangent to EcsE^{cs} satisfies Φ(γ(1))Φ(γ(0))+E0s\Phi(\gamma(1))\in\Phi(\gamma(0))+E_{0}^{s} by Lemma 3.4, so W^cs(x)Φ1(Φ(x)+E0s)\hat{W}^{cs}(x)\subset\Phi^{-1}(\Phi(x)+E_{0}^{s}) and Φ1(Φ(x)+E0s)\Phi^{-1}(\Phi(x)+E_{0}^{s}) is a union of W^cs\hat{W}^{cs}-leaves. We claim that Φ1(Φ(x)+E0s)\Phi^{-1}(\Phi(x)+E_{0}^{s}) is path-connected, which proves that Φ1(Φ(x)+E0s)=W^cs(x)\Phi^{-1}(\Phi(x)+E_{0}^{s})=\hat{W}^{cs}(x). Given any yGy\in G there is a unique intersection W^u(y)Φ1(Φ(x)+E0s)={w}\hat{W}^{u}(y)\cap\Phi^{-1}(\Phi(x)+E_{0}^{s})=\{w\} (since W^u(y)\hat{W}^{u}(y) maps homeomorphically onto Φ(y)+E0u\Phi(y)+E_{0}^{u} under Φ\Phi). Since Φ1(Φ(x)+E0s)\Phi^{-1}(\Phi(x)+E_{0}^{s}) is tangent to EcsE^{cs} and W^u\hat{W}^{u} is tangent to EuE^{u} the map

(3.24) GΦ1(Φ(x)+E0s),yW^u(y)Φ1(Φ(x)+E0s)\displaystyle G\to\Phi^{-1}(\Phi(x)+E_{0}^{s}),\quad y\mapsto\hat{W}^{u}(y)\cap\Phi^{-1}(\Phi(x)+E_{0}^{s})

is continuous. Since GΦ1(Φ(x)+E0s)G\to\Phi^{-1}(\Phi(x)+E_{0}^{s}) is surjective and GG is path-connected it follows that Φ1(Φ(x)+E0s)\Phi^{-1}(\Phi(x)+E_{0}^{s}) is path-connected. Properties (i)(i) and (ii)(ii) follows. Using

W^cs(x)=Φ1(Φ(x)+E0s),\displaystyle\hat{W}^{cs}(x)=\Phi^{-1}(\Phi(x)+E_{0}^{s}),
W^cu(x)=Φ1(Φ(x)+E0u), and\displaystyle\hat{W}^{cu}(x)=\Phi^{-1}(\Phi(x)+E_{0}^{u}),\text{ and}
W^c(x)=W^cs(x)W^cu(x)\displaystyle\hat{W}^{c}(x)=\hat{W}^{cs}(x)\cap\hat{W}^{cu}(x)

we obtain W^c(x)=Φ1(Φ(x))\hat{W}^{c}(x)=\Phi^{-1}(\Phi(x)). Since Φ\Phi descends to a map Φ:XΓ𝕋d\Phi:X_{\Gamma}\to\mathbb{T}^{d} the fibers Wc(x)=Φ1(Φ(x))W^{c}(x)=\Phi^{-1}(\Phi(x)) are compact. We define

(3.25) Γc={γΓ : Lγ=γ}\displaystyle\Gamma^{c}=\{\gamma\in\Gamma\text{ : }L\gamma=\gamma\}

where LAut(XΓ)L\in{\rm Aut}(X_{\Gamma}) is the linearization of ff. For γcΓc\gamma^{c}\in\Gamma^{c} we have

(3.26) γcW^c(x)=W^c(γcx)=Φ1(Φ(γcx))=Φ1(Φ(x))=W^c(x)\displaystyle\gamma^{c}\hat{W}^{c}(x)=\hat{W}^{c}(\gamma^{c}x)=\Phi^{-1}(\Phi(\gamma^{c}x))=\Phi^{-1}(\Phi(x))=\hat{W}^{c}(x)

where we have used Φ(γx)=π(γ)+Φ(x)\Phi(\gamma x)=\pi(\gamma)+\Phi(x) and kerπ=Γc\ker\pi=\Gamma^{c}. So, if we fix a generator γ0c\gamma_{0}^{c} of Γc\Gamma^{c} then we can orient W^c(x)\hat{W}^{c}(x) by letting γ0cx>x\gamma_{0}^{c}x>x. This is a well-defined orientation of W^c(x)\hat{W}^{c}(x) since xγ0cxx\mapsto\gamma_{0}^{c}x have no fixed points. That WcW^{c} are circles follows since they are compact 11-dimensional manifolds.

Let ε>0\varepsilon>0 be small and fix xXΓx\in X_{\Gamma}. We denote by D=Wεs(Wεu(x))D=W_{\varepsilon}^{s}(W_{\varepsilon}^{u}(x)), U=Wc(D)U=W^{c}(D) and D=Φ(D)=Φ(U)D^{\prime}=\Phi(D)=\Phi(U). With ε\varepsilon small enough we define

(3.27) πsu:UWc(x)\displaystyle\pi^{su}:U\to W^{c}(x)

by the unique holonomy first along WsW^{s} then WuW^{u} in UU. We obtain a map

(3.28) UD×Wc(x)y(Φ(y),πsu(y)),\displaystyle U\to D^{\prime}\times W^{c}(x)\quad y\mapsto(\Phi(y),\pi^{su}(y)),

this map is smooth along WcW^{c} since the holonomies are CrC^{r} (where rr depends on the bunching). That Φ\Phi semiconjugates (a finite index subgroup of) the centralizer Z(f)Z^{\infty}(f) onto its linearization is immediate since Φ\Phi is (essentially) unique homotopic to π\pi (note that the uniqueness in Lemma 3.1 implies that Φ:XΓ𝕋d\Phi:X_{\Gamma}\to\mathbb{T}^{d} is unique modulo the fact that we could change Φ(x)\Phi(x) to Φ(x)+p0\Phi(x)+p_{0} where Lsup0=p0L_{su}p_{0}=p_{0}). Indeed, for any gZ(g)g\in Z^{\infty}(g) we let BGL(d,)B\in{\rm GL}(d,\mathbb{Z}) be the induced map on 𝕋d\mathbb{T}^{d} by its linearization and let Φ~(x)=B1Φ(gx)\tilde{\Phi}(x)=B^{-1}\Phi(gx). Then Φ~\tilde{\Phi} is homotopic to π\pi and we still have

(3.29) Φ~(fx)=B1Φ(gfx)=B1Φ(fgx)=LsuB1Φ(gx)=LsuΦ~(x)\displaystyle\tilde{\Phi}(fx)=B^{-1}\Phi(gfx)=B^{-1}\Phi(fgx)=L_{su}B^{-1}\Phi(gx)=L_{su}\tilde{\Phi}(x)

so Φ~(x)=Φ(x)+p0\tilde{\Phi}(x)=\Phi(x)+p_{0} where p0p_{0} is a fixed point for LsuL_{su}. Define

(3.30) Zfix(f)={gZ(f) : gWc(x)=Wc(x) if fWc(x)=Wc(x)}.\displaystyle Z_{\rm fix}^{\infty}(f)=\{g\in Z^{\infty}(f)\text{ : }gW^{c}(x)=W^{c}(x)\text{ if }fW^{c}(x)=W^{c}(x)\}.

Since Φ(fx)=LsuΦ(x)\Phi(fx)=L_{su}\Phi(x) and LsuL_{su} has finitely many fixed points Zfix(f)Z_{\rm fix}^{\infty}(f) has finite index in Z(f)Z^{\infty}(f). For gZfix(f)g\in Z_{\rm fix}^{\infty}(f) and Φ(x0)=0\Phi(x_{0})=0 we have fWc(x0)=Wc(x0)fW^{c}(x_{0})=W^{c}(x_{0}) so gWc(x0)=Wc(x0)gW^{c}(x_{0})=W^{c}(x_{0}). It follows that

(3.31) p0=Φ~(x0)=B1Φ(gx0)=B1Φ(x0)=0\displaystyle p_{0}=\tilde{\Phi}(x_{0})=B^{-1}\Phi(gx_{0})=B^{-1}\Phi(x_{0})=0

or B1Φ(gx)=Φ(x)B^{-1}\Phi(gx)=\Phi(x). ∎

3.2. Proof of accessibility

In this section, we show the last point of Theorem 1.1. For γΓ\gamma\in\Gamma, define Tγ:GGT_{\gamma}:G\to G by

(3.32) Tγ(x)=W^u[W^s(γx)W^cu(x)]W^c(x)=W^u(W^s(γx))W^c(x)\displaystyle T_{\gamma}(x)=\hat{W}^{u}\left[\hat{W}^{s}(\gamma x)\cap\hat{W}^{cu}(x)\right]\cap\hat{W}^{c}(x)=\hat{W}^{u}(\hat{W}^{s}(\gamma x))\cap\hat{W}^{c}(x)

which is well-defined by point (i)(i) of Theorem 1.1. A calculation shows

FTγ(x)=TLγ(Fx).\displaystyle FT_{\gamma}(x)=T_{L\gamma}(Fx).

Recall that we denote by Λ(F)G\Lambda(F)\subset G and Λ(f)XΓ\Lambda(f)\subset X_{\Gamma} the complement of open accessibility classes. The set Λ(F)\Lambda(F) is closed and susu-saturated. Since EcE^{c} is 11-dimensional the set Λ(F)\Lambda(F) is laminated by accessibility classes, denoted AC(x)=W^su(x){\rm AC}(x)=\hat{W}^{su}(x), [54]. The union of open accessibility classes is Γ\Gamma-invariant, so Λ(F)\Lambda(F) is Γ\Gamma-invariant. Moreover, FF maps accessibility classes to accessibility classes so FΛ(F)=Λ(F)F\Lambda(F)=\Lambda(F). Given γ,γΓ\gamma,\gamma^{\prime}\in\Gamma and xΛ(F)x\in\Lambda(F) we have

Tγ(Tγ(x))=\displaystyle T_{\gamma}(T_{\gamma^{\prime}}(x))= W^su(γ[W^su(γx)W^c(x)]))W^c([W^su(γx)W^c(x)])=\displaystyle\hat{W}^{su}\left(\gamma\left[\hat{W}^{su}(\gamma^{\prime}x)\cap\hat{W}^{c}(x)\right]\right))\cap\hat{W}^{c}\left(\left[\hat{W}^{su}(\gamma^{\prime}x)\cap\hat{W}^{c}(x)\right]\right)=
W^su(W^su(γγx)W^c(γx))W^c(x)=\displaystyle\hat{W}^{su}(\hat{W}^{su}(\gamma\gamma^{\prime}x)\cap\hat{W}^{c}(\gamma x))\cap\hat{W}^{c}(x)=
W^su(γγx)W^c(x)=Tγγ(x)\displaystyle\hat{W}^{su}(\gamma\gamma^{\prime}x)\cap\hat{W}^{c}(x)=T_{\gamma\gamma^{\prime}}(x)

so restricted to Λ(F)\Lambda(F) the map (γ,x)Tγ(x)(\gamma,x)\mapsto T_{\gamma}(x) defines a group action of Γ\Gamma, see also [52, Lemma 6.1]. Before starting the proof we will need an elementary, but important, auxiliary lemma on k\mathbb{Z}^{k}-actions on the circle.

Lemma 3.5.

Let f1,,fk,gHomeo+(𝕋)f_{1},...,f_{k},g\in{\rm Homeo}_{+}(\mathbb{T}) be orientation preserving homeomorphisms on the circle and let K𝕋K\subset\mathbb{T} be a compact subset that is invariant by gg and each fjf_{j}. Moreover, assume that fifj=fjfif_{i}f_{j}=f_{j}f_{i} on KK and that there is some hyperbolic integer matrix (Aij)1i,jk(A_{i}^{j})_{1\leq i,j\leq k} such that

(3.33) gfi=f1Ai1fkAikg,on K.\displaystyle gf_{i}=f_{1}^{A_{i}^{1}}...f_{k}^{A_{i}^{k}}g,\quad{\rm on}\text{ }K.

Then the k\mathbb{Z}^{k}-action generated by f1,,fkf_{1},...,f_{k} on KK, β:k×KK\beta:\mathbb{Z}^{k}\times K\to K, has a periodic point. That is, there is a point pKp\in K and a finite index subgroup Λk\Lambda\leq\mathbb{Z}^{k} such that β𝐧p=p\beta^{\mathbf{n}}p=p for 𝐧Λ\mathbf{n}\in\Lambda.

Remark 14.

The condition in Equation 3.33 says that β:k×KK\beta:\mathbb{Z}^{k}\times K\to K joint with gg form an Abelian-by-Cyclic (AbC) action on KK.

In the proof, we will use the following two lemmas whose proofs are standard, but we include them for completeness.

Lemma 3.6.

If fHomeo+(𝕋)f\in{\rm Homeo}_{+}(\mathbb{T}) is an orientation preserving homeomorphism on the circle with zero rotation number, ω(f)=0\omega(f)=0, and K𝕋K\subset\mathbb{T} is a compact ff-invariant set, then ff have a fixed point in KK.

Proof.

If ff has zero rotation number then for any p𝕋p\in\mathbb{T} the sequence fnpf^{n}p converges to a fixed point of ff as nn\to\infty. For any xKx\in K the sequence fnxf^{n}x lie in KK since KK is ff-invariant. By compactness of KK any limit point of fnxf^{n}x also lies in KK, so ff has a fixed point in KK. ∎

Lemma 3.7.

Let f,gHomeo+(𝕋)f,g\in{\rm Homeo}_{+}(\mathbb{T}) be orientation preserving homeomorphisms of the circle and assume that there is a ff and gg-invariant probability measure ν\nu. Then the rotation numbers satisfies ω(fg)=ω(f)+ω(g)\omega(fg)=\omega(f)+\omega(g).

Proof.

Write f(x)=x+u(x)f(x)=x+u(x), g(x)=x+v(x)g(x)=x+v(x) where u,v:𝕋u,v:\mathbb{T}\to\mathbb{R}. Recall that if μ\mu is a ff-invariant measure then we obtain the rotation number of ff as

(3.34) ω(f)=𝕋u(x)dμ(x)+.\displaystyle\omega(f)=\int_{\mathbb{T}}u(x){\rm d}\mu(x)+\mathbb{Z}.

Similarly we obtain the rotation number of gg. The measure ν\nu is fgfg-invariant and f(gx)=x+u(gx)+v(x)f(gx)=x+u(gx)+v(x), so we can write the rotation number of fgfg as

ω(fg)=\displaystyle\omega(fg)= 𝕋[u(gx)+v(x)]dν(x)+=\displaystyle\int_{\mathbb{T}}\left[u(gx)+v(x)\right]{\rm d}\nu(x)+\mathbb{Z}=
𝕋u(gx)dν(x)+𝕋v(x)dν(x)+=\displaystyle\int_{\mathbb{T}}u(gx){\rm d}\nu(x)+\int_{\mathbb{T}}v(x){\rm d}\nu(x)+\mathbb{Z}=
𝕋u(x)dν(x)+𝕋v(x)dν(x)+=\displaystyle\int_{\mathbb{T}}u(x){\rm d}\nu(x)+\int_{\mathbb{T}}v(x){\rm d}\nu(x)+\mathbb{Z}=
ω(f)+ω(g)\displaystyle\omega(f)+\omega(g)

where the second to last equality use that ν\nu is gg-invariant. ∎

Proof of Lemma 3.5.

Since the β\beta-action is abelian on KK it has an invariant measure on KK. By Lemma 3.7 the rotation numbers satisfy

(3.35) ω(f1n1fknk)=n1ω(f1)++nkω(fk)\displaystyle\omega(f_{1}^{n_{1}}...f_{k}^{n_{k}})=n_{1}\omega(f_{1})+...+n_{k}\omega(f_{k})

for all integers n1,,nkn_{1},...,n_{k}\in\mathbb{Z}. Since gg preserves orientation, conjugacy invariance of rotation number and Equation 3.33 implies

(3.36) ω(fi)=ω(f1Ai1)++ω(fkAik).\displaystyle\omega(f_{i})=\omega(f_{1}^{A_{i}^{1}})+...+\omega(f_{k}^{A_{i}^{k}}).

Or if we denote the map f1n1fknkf_{1}^{n_{1}}...f_{k}^{n_{k}} by β𝐧\beta^{\mathbf{n}}, then we can write ω(β𝐧)=ω(βA𝐧)\omega(\beta^{\mathbf{n}})=\omega(\beta^{A\mathbf{n}}). Using Equation 3.35 we obtain

(3.37) ω(β(AI)𝐧)=0\displaystyle\omega\left(\beta^{(A-I)\mathbf{n}}\right)=0

for all 𝐧k\mathbf{n}\in\mathbb{Z}^{k}. Since AA is hyperbolic AIA-I is invertible over the rationals. So there is a finite index subgroup Λk\Lambda\subset\mathbb{Z}^{k} such that ω(β𝐧)=0\omega(\beta^{\mathbf{n}})=0 for all 𝐧Λ\mathbf{n}\in\Lambda. Let e1,,ekΛe_{1},...,e_{k}\in\Lambda be generators. By Lemma 3.6 the map βe1\beta^{e_{1}} has a fixed point in KK. Since βe2\beta^{e_{2}} commute with βe1\beta^{e_{1}} within KK it follows that βe2\beta^{e_{2}} preserve the compact set Fix(βe1)K{\rm Fix}(\beta^{e_{1}})\cap K\neq\emptyset. So, if we apply Lemma 3.6 once more we see that the set

(3.38) KFix(βe1)Fix(βe2)\displaystyle K\cap{\rm Fix}(\beta^{e_{1}})\cap{\rm Fix}(\beta^{e_{2}})

is non-empty. Proceeding by induction, we find a point pKp\in K that is fixed by βe1,,βek\beta^{e_{1}},...,\beta^{e_{k}}, and therefore by Λ\Lambda. Since Λ\Lambda has a finite index in k\mathbb{Z}^{k} the lemma follows. ∎

Recall that LAut(XΓ)L\in{\rm Aut}(X_{\Gamma}) is the linearization of ff and the LL-fixed part of Γ\Gamma is Γc\Gamma^{c} (Equation 3.25).

Lemma 3.8.

Assume that Λ(f)\Lambda(f)\neq\emptyset. There is a finite index subgroup ΓΓ\Gamma^{\prime}\leq\Gamma and xΛ(F)x\in\Lambda(F) such that Tγ(x)ΓcxT_{\gamma}(x)\in\Gamma^{c}x for γΓ\gamma\in\Gamma^{\prime}.

Proof.

Fix generators γ1,γ2,,γdΓ\gamma_{1},\gamma_{2},...,\gamma_{d}\in\Gamma. Let x0Gx_{0}\in G be such that Φ(x0)=e\Phi(x_{0})=e. Identify W^c(x0)/ΓcWc(pΓ(x0))\hat{W}^{c}(x_{0})/\Gamma^{c}\cong W^{c}(p_{\Gamma}(x_{0})) with 𝕋\mathbb{T}. Since Γc\Gamma^{c} is central in Γ\Gamma we can identify Tγj:W^c(x0)/ΓcW^c(x0)/ΓcT_{\gamma_{j}}:\hat{W}^{c}(x_{0})/\Gamma^{c}\to\hat{W}^{c}(x_{0})/\Gamma^{c} with circle diffeomorphisms. We also identify FF with a circle diffeomorphism (by our choice of x0x_{0} and the fact that Γc\Gamma^{c} is LL-fixed we have FW^c(x0)/Γc=W^c(x0)/ΓcF\hat{W}^{c}(x_{0})/\Gamma^{c}=\hat{W}^{c}(x_{0})/\Gamma^{c}). Let K=(Λ(F)W^c(x0))/ΓcK=(\Lambda(F)\cap\hat{W}^{c}(x_{0}))/\Gamma^{c}, which is compact, TγjT_{\gamma_{j}}-invariant, FF-invariant, and non-empty. Any γcΓc\gamma^{c}\in\Gamma^{c} act trivially on 𝕋\mathbb{T} (and therefore KK) under TγcT_{\gamma^{c}}, so the action TγT_{\gamma} on KK factor through Γ/Γcd\Gamma/\Gamma^{c}\cong\mathbb{Z}^{d}. Moreover, FF satisfy FTγ=TLγFFT_{\gamma}=T_{L\gamma}F so the assumptions of Lemma 3.5 are satisfied with Tγj=fjT_{\gamma_{j}}=f_{j} and g=Fg=F. Therefore there is a finite index subgroup of Γ/Γc\Gamma/\Gamma^{c} that admits a fixed point on 𝕋\mathbb{T} which implies the lemma. ∎

An immediate corollary of Lemma 3.8 is that there is a compact susu-leaf.

Lemma 3.9.

If Λ(f)\Lambda(f)\neq\emptyset, or equivalently if ff is not accessible, then there is a compact susu-leaf intersecting each center leaf q<q<\infty times.

Proof.

For any xΛ(F)x\in\Lambda(F) the map

(3.39) W^su(x)Φd\displaystyle\hat{W}^{su}(x)\xrightarrow{\Phi}\mathbb{R}^{d}

is a homeomorphism by points (i)(i) and (iv)(iv) in Theorem 1.1. Choose x0x_{0} as in Lemma 3.8 and let ΓΓ\Gamma^{\prime}\leq\Gamma be of finite index such that γW^su(x0)ΓcW^su(x0)\gamma\hat{W}^{su}(x_{0})\subset\Gamma^{c}\hat{W}^{su}(x_{0}) for all γΓ\gamma\in\Gamma^{\prime}. Note that ΦΓ=πΓ=Λd\Phi_{*}\Gamma^{\prime}=\pi_{*}\Gamma^{\prime}=\Lambda\subset\mathbb{Z}^{d} has finite index in d\mathbb{Z}^{d}. Define

(3.40) Λ={γΓ : γW^su(x0)=W^su(x0)}Γ.\displaystyle\Lambda_{*}=\{\gamma\in\Gamma\text{ : }\gamma\hat{W}^{su}(x_{0})=\hat{W}^{su}(x_{0})\}\subset\Gamma^{\prime}.

Since γW^su(x0)ΓcW^su(x0)\gamma\hat{W}^{su}(x_{0})\subset\Gamma^{c}\hat{W}^{su}(x_{0}) for all γΓ\gamma\in\Gamma^{\prime} there is for each γΓ\gamma\in\Gamma^{\prime} some γcΓc\gamma^{c}\in\Gamma^{c} such that γγcΛ\gamma\gamma^{c}\in\Lambda_{*}. In particular, ΦΛ=Λ\Phi_{*}\Lambda_{*}=\Lambda. Since Λ\Lambda is a lattice in d\mathbb{R}^{d} and we obtain a homeomorphism W^su(x0)/Λd/Λ\hat{W}^{su}(x_{0})/\Lambda_{*}\cong\mathbb{R}^{d}/\Lambda the set W^su(x0)/Λ\hat{W}^{su}(x_{0})/\Lambda_{*} is compact. Since pΓW^su(x0)=Wsu(pΓ(x0))p_{\Gamma}\hat{W}^{su}(x_{0})=W^{su}(p_{\Gamma}(x_{0})) is homeomorphic to W^su(x0)/Λ\hat{W}^{su}(x_{0})/\Lambda_{*} the lemma follows. ∎

We can now prove the last claim of Theorem 1.1

Proof of (vi)(vi) in Theorem 1.1.

If ff is not accessible then we construct ΛΓ\Lambda_{*}\subset\Gamma as in Equation 3.40. Since πΛ\pi_{*}\Lambda_{*} has finite index in d\mathbb{Z}^{d} the group Λ×kerπ\Lambda_{*}\times\ker\pi_{*} has finite index in Γ\Gamma, so Γ\Gamma is virtually abelian. This is a contradiction if GG is a non-abelian nilpotent Lie group. ∎

4. Action of the su-path group

In this section we introduce and prove basic properties of the susu-path group 𝒫\mathcal{P}. The susu-path group naturally acts on XΓX_{\Gamma} (Definition 4.1). The group 𝒫\mathcal{P}, its various subgroups, and its action on XΓX_{\Gamma} will be the key object in the following sections.

4.1. The su-path group

Let α:k×XΓXΓ\alpha:\mathbb{Z}^{k}\times X_{\Gamma}\to X_{\Gamma} be a smooth action satisfying the assumptions of Theorem A. Moreover, let ρ:kGL(d,)\rho:\mathbb{Z}^{k}\to{\rm GL}(d,\mathbb{Z}) the action defined by ρ𝐧Φ=Φα𝐧\rho^{\mathbf{n}}\Phi=\Phi\alpha^{\mathbf{n}} from Theorem 1.1, f=α(γ0)f=\alpha(\gamma_{0}) the partially hyperbolic element, and let LL be the linearization of ff. Denote by E0sE_{0}^{s} and E0uE_{0}^{u} the stable and unstable distributions for LL.

Definition 4.1.

We define the susu-path group 𝒫\mathcal{P} as the free product

(4.1) 𝒫=E0sE0u.\displaystyle\mathcal{P}=E_{0}^{s}*E_{0}^{u}.

If w𝒫w\in\mathcal{P} is a word in 𝒫\mathcal{P} then we define Π(w)=w1++wN\Pi(w)=w_{1}+...+w_{N} to be the sum of all factors in ww. We define the normal subgroup 𝒫c=Π1(0)\mathcal{P}^{c}=\Pi^{-1}(0).

Given any pair of negatively proportional course Lyapunov spaces, E0[χ]E_{0}^{-[\chi]}, E0[χ]E_{0}^{[\chi]}, of ρ\rho, we define the [χ][\chi]-path group

(4.2) 𝒫[χ]=E0[χ]E0[χ],𝒫[χ]c=𝒫[χ]Π1(0)\displaystyle\mathcal{P}_{[\chi]}=E_{0}^{-[\chi]}*E_{0}^{[\chi]},\quad\mathcal{P}_{[\chi]}^{c}=\mathcal{P}_{[\chi]}\cap\Pi^{-1}(0)

with E0[χ]=0E_{0}^{-[\chi]}=0 if [χ]-[\chi] is not a coarse exponent. We also define the complementary [χ][\chi]-path group

(4.3) 𝒬[χ]=([η]±[χ][η](𝐧0)>0E[η])([η]±[χ][η](𝐧0)>0E[η]),𝒬[χ]c=𝒬[χ]𝒫c.\displaystyle\mathcal{Q}_{[\chi]}=\left(\bigoplus_{\begin{subarray}{c}[\eta]\neq\pm[\chi]\\ [\eta](\mathbf{n}_{0})>0\end{subarray}}E^{-[\eta]}\right)*\left(\bigoplus_{\begin{subarray}{c}[\eta]\neq\pm[\chi]\\ [\eta](\mathbf{n}_{0})>0\end{subarray}}E^{[\eta]}\right),\quad\mathcal{Q}_{[\chi]}^{c}=\mathcal{Q}_{[\chi]}\cap\mathcal{P}^{c}.

It is immediate that 𝒫[χ],𝒬[χ]𝒫\mathcal{P}_{[\chi]},\mathcal{Q}_{[\chi]}\subset\mathcal{P}, 𝒫[χ]c,𝒬[χ]c𝒫c\mathcal{P}_{[\chi]}^{c},\mathcal{Q}_{[\chi]}^{c}\subset\mathcal{P}^{c}. The following well-known lemma on free products will be useful.

Lemma 4.1.

Let V,WV,W be vector spaces, 𝒢=VU\mathcal{G}=V*U and Π:GVW\Pi:G\to V\oplus W the map defined by

(4.4) Π(v1u1vnun)=v1++vn+u1++un.\displaystyle\Pi(v_{1}u_{1}...v_{n}u_{n})=v_{1}+...+v_{n}+u_{1}+...+u_{n}.

Then Π1(0)=kerΠ=[𝒢,𝒢]\Pi^{-1}(0)=\ker\Pi=[\mathcal{G},\mathcal{G}] and any w𝒢w\in\mathcal{G} can be written w=w~vuw=\tilde{w}vu with vVv\in V, uUu\in U and w~[𝒢,𝒢]\tilde{w}\in[\mathcal{G},\mathcal{G}].

Proof.

For w=v1u1vnun𝒢w=v_{1}u_{1}...v_{n}u_{n}\in\mathcal{G} we have

v1u1vnun=\displaystyle v_{1}u_{1}...v_{n}u_{n}= [v1u1(v1)(u1)]u1(v1+v2)u2vnun=\displaystyle\left[v_{1}u_{1}(-v_{1})(-u_{1})\right]u_{1}(v_{1}+v_{2})u_{2}...v_{n}u_{n}=
[v1u1(v1)(u1)][u1(v1+v2)(u1)(v1v2)]\displaystyle\left[v_{1}u_{1}(-v_{1})(-u_{1})\right]\cdot\left[u_{1}(v_{1}+v_{2})(-u_{1})(-v_{1}-v_{2})\right]\cdot
(v1+v2)(u1+u2)v3u3vnun.\displaystyle(v_{1}+v_{2})(u_{1}+u_{2})v_{3}u_{3}...v_{n}u_{n}.

Note that v1u1(v1)(u1),u1(v1+v2)(u1)(v1v2)[𝒢,𝒢]v_{1}u_{1}(-v_{1})(-u_{1}),u_{1}(v_{1}+v_{2})(-u_{1})(-v_{1}-v_{2})\in[\mathcal{G},\mathcal{G}] so (v1+v2)(u1+u2)v3u3vnun(v_{1}+v_{2})(u_{1}+u_{2})v_{3}u_{3}...v_{n}u_{n} consists of n1n-1 pairs. By induction we find w1,,w[𝒢,𝒢]w_{1},...,w_{\ell}\in[\mathcal{G},\mathcal{G}] such that

(4.5) w=w1w(v1++vn)(u1++un).\displaystyle w=w_{1}...w_{\ell}\cdot\left(v_{1}+...+v_{n}\right)\left(u_{1}+...+u_{n}\right).

This proves the last part of the lemma. Since Π(wj)=0\Pi(w_{j})=0 for each jj, we have Π(w)=v1++vn+u1++un\Pi(w)=v_{1}+...+v_{n}+u_{1}+...+u_{n}. In particular, if wkerΠw\in\ker\Pi then v1++vn=0v_{1}+...+v_{n}=0 and u1++un=0u_{1}+...+u_{n}=0, so w=w1w[𝒢,𝒢]w=w_{1}...w_{\ell}\in[\mathcal{G},\mathcal{G}]. ∎

Definition 4.2.

For tE0σt\in E_{0}^{\sigma} we define ησt:GG\eta_{\sigma}^{t}:G\to G (or ησt:XΓXΓ\eta_{\sigma}^{t}:X_{\Gamma}\to X_{\Gamma}) by ησtx=Φσ,x1(t)\eta_{\sigma}^{t}x=\Phi_{\sigma,x}^{-1}(t), where Φσ,x\Phi_{\sigma,x} is defined in Lemma 3.3.

Remark 15.

By Lemma 3.3 the map ησt\eta_{\sigma}^{t} is well-defined.

Lemma 4.2.

The map ησ:E0σ×GG\eta_{\sigma}:E_{0}^{\sigma}\times G\to G, (t,x)ησtx(t,x)\mapsto\eta_{\sigma}^{t}x, is a Hölder E0σE_{0}^{\sigma}-action that satisfies Fησtx=ησLsutFxF\eta_{\sigma}^{t}x=\eta_{\sigma}^{L_{su}t}Fx. The action ησt\eta_{\sigma}^{t} naturally descends to XΓX_{\Gamma} and if gZ(f)g\in Z^{\infty}(f), BB is the automorphism defined by Φ(gx)=BΦ(x)\Phi(gx)=B\Phi(x), then gησtx=ησBtgxg\eta_{\sigma}^{t}x=\eta_{\sigma}^{Bt}gx. Finally, Φ(ησtx)=t+Φ(x)\Phi(\eta_{\sigma}^{t}x)=t+\Phi(x), so Φ\Phi semi-conjugates ησt\eta_{\sigma}^{t} to the standard translation action along E0σE_{0}^{\sigma} on the base.

Proof.

That ησt\eta_{\sigma}^{t} defines an action is immediate from the definition. Indeed, for t,sE0σt,s\in E_{0}^{\sigma}

Φ(ησsησtx)=\displaystyle\Phi(\eta_{\sigma}^{s}\eta_{\sigma}^{t}x)= Φ(Φσ,ησtx1(s))=s+Φ(ησtx)=s+Φσ,x(ησtx)+Φ(x)=\displaystyle\Phi(\Phi_{\sigma,\eta_{\sigma}^{t}x}^{-1}(s))=s+\Phi(\eta_{\sigma}^{t}x)=s+\Phi_{\sigma,x}(\eta_{\sigma}^{t}x)+\Phi(x)=
s+t+Φ(x)\displaystyle s+t+\Phi(x)

so if we subtract Φ(x)\Phi(x) then Φσ,x(ησsησtx)=s+t\Phi_{\sigma,x}(\eta_{\sigma}^{s}\eta_{\sigma}^{t}x)=s+t. Applying Φσ,x1\Phi_{\sigma,x}^{-1} on both sides of the equality yields ησsησtx=ησs+tx\eta_{\sigma}^{s}\eta_{\sigma}^{t}x=\eta_{\sigma}^{s+t}x. This also shows that Φ(ηutx)=Φ(x)+t\Phi(\eta_{u}^{t}x)=\Phi(x)+t. Given γΓ\gamma\in\Gamma we have Φσ,γx(γy)=Φ(γy)Φ(γx)=Φ(y)Φ(x)=Φσ,x(y)\Phi_{\sigma,\gamma x}(\gamma y)=\Phi(\gamma y)-\Phi(\gamma x)=\Phi(y)-\Phi(x)=\Phi_{\sigma,x}(y) which implies ησt(γx)=γησtx\eta_{\sigma}^{t}(\gamma x)=\gamma\eta_{\sigma}^{t}x, so ησt\eta_{\sigma}^{t} descend to XΓX_{\Gamma}. For gZ(f)g\in Z^{\infty}(f) we have Φ(gx)=BΦ(x)+p0\Phi(gx)=B\Phi(x)+p_{0} for some p0𝕋dp_{0}\in\mathbb{T}^{d} (that is fixed by LsuL_{su}). It follows that

(4.6) Φσ,gx(gy)=Φ(gy)Φ(gx)=B(Φ(x)Φ(y))=BΦσ,x(y)\displaystyle\Phi_{\sigma,gx}(gy)=\Phi(gy)-\Phi(gx)=B\left(\Phi(x)-\Phi(y)\right)=B\Phi_{\sigma,x}(y)

or with Φσ,x(y)=t\Phi_{\sigma,x}(y)=t

(4.7) ησBt(gx)=gy=gησt(x).\displaystyle\eta_{\sigma}^{Bt}(gx)=gy=g\eta_{\sigma}^{t}(x).

Next, we show that ησ\eta_{\sigma} is Hölder. Let u=σu=\sigma, the other case is similar. Since Φu,x\Phi_{u,x} is a bi-Hölder homeomorphism (Lemma 3.3) it is immediate that (t,x)ηutx(t,x)\mapsto\eta_{u}^{t}x is Hölder in tt. The foliations W^u\hat{W}^{u} and W^cs\hat{W}^{cs} are uniformly transverse, so we find ε0>0\varepsilon_{0}>0 and KK such that for x,yXΓx,y\in X_{\Gamma} with d(x,y)ε0{\rm d}(x,y)\leq\varepsilon_{0} we have W^locu(x)W^loccs(y)={z}\hat{W}_{\rm loc}^{u}(x)\cap\hat{W}_{\rm loc}^{cs}(y)=\{z\},

(4.8) du(x,z)Kd(x,y),dcs(z,y)Kd(x,y).\displaystyle{\rm d}_{u}(x,z)\leq K{\rm d}(x,y),\quad{\rm d}_{cs}(z,y)\leq K{\rm d}(x,y).

If yW^cs(x)y\in\hat{W}^{cs}(x) then Lemma 3.4 shows that Φ(y)Φ(x)+E0s\Phi(y)\in\Phi(x)+E_{0}^{s}, so

Φ(ηuty)=\displaystyle\Phi(\eta_{u}^{t}y)= Φ(y)+t=Φ(x)+t+[Φ(y)Φ(x)]=\displaystyle\Phi(y)+t=\Phi(x)+t+[\Phi(y)-\Phi(x)]=
Φ(ηutx)+[Φ(y)Φ(x)]Φ(ηutx)+E0s,\displaystyle\Phi(\eta_{u}^{t}x)+[\Phi(y)-\Phi(x)]\in\Phi(\eta_{u}^{t}x)+E_{0}^{s},

or ηutyW^cs(Φ(ηutx))\eta_{u}^{t}y\in\hat{W}^{cs}(\Phi(\eta_{u}^{t}x)). That is, ηut\eta_{u}^{t} preserve the foliation W^cs\hat{W}^{cs}. Given tE0ut\in E_{0}^{u}

(4.9) d(ηutx,ηuty)du(ηutx,ηutz)+dcs(ηutz,ηuty)\displaystyle{\rm d}(\eta_{u}^{t}x,\eta_{u}^{t}y)\leq{\rm d}_{u}(\eta_{u}^{t}x,\eta_{u}^{t}z)+{\rm d}_{cs}(\eta_{u}^{t}z,\eta_{u}^{t}y)

with x,yx,y and zz as in Equation 4.8. So, it suffices to show that ηut\eta_{u}^{t} is Hölder along W^u\hat{W}^{u} and W^cs\hat{W}^{cs}. For yW^u(x)y\in\hat{W}^{u}(x)

Φu,x(ηuty)=\displaystyle\Phi_{u,x}(\eta_{u}^{t}y)= Φ(ηuty)Φ(x)=Φu,y(ηuty)+Φ(y)Φ(x)=\displaystyle\Phi(\eta_{u}^{t}y)-\Phi(x)=\Phi_{u,y}(\eta_{u}^{t}y)+\Phi(y)-\Phi(x)=
t+Φ(y)Φ(x)\displaystyle t+\Phi(y)-\Phi(x)

or ηuty=Φu,x1(t+Φ(y)Φ(x))\eta_{u}^{t}y=\Phi_{u,x}^{-1}(t+\Phi(y)-\Phi(x)). Since Φ\Phi is Hölder and Φu,x1\Phi_{u,x}^{-1} is uniformly Hölder

(4.10) d(ηutx,ηuty)CΦ(y)Φ(x)θCd(x,y)θ\displaystyle{\rm d}(\eta_{u}^{t}x,\eta_{u}^{t}y)\leq C\left\lVert\Phi(y)-\Phi(x)\right\rVert^{\theta}\leq C^{\prime}{\rm d}(x,y)^{\theta^{\prime}}

so ηut\eta_{u}^{t} is Hölder along W^u\hat{W}^{u}. Given yW^cs(x)y\in\hat{W}^{cs}(x) we have ηutyW^cs(ηutx)\eta_{u}^{t}y\in\hat{W}^{cs}(\eta_{u}^{t}x) since ηut\eta_{u}^{t} preserve W^cs\hat{W}^{cs}. On the other hand, ηutyW^u(y)\eta_{u}^{t}y\in\hat{W}^{u}(y) by the definition of Φu,y\Phi_{u,y}. So ηuty=πx,ηutxu(y)\eta_{u}^{t}y=\pi_{x,\eta_{u}^{t}x}^{u}(y). The unstable Holonomy is (uniformly) Hölder [50], so ηut\eta_{u}^{t} is Hölder along W^cs\hat{W}^{cs}. ∎

Definition 4.3.

We define an action of 𝒫\mathcal{P} (or 𝒫c,𝒫[χ],𝒬[χ],𝒫[χ]c,𝒬[χ]c\mathcal{P}^{c},\mathcal{P}_{[\chi]},\mathcal{Q}_{[\chi]},\mathcal{P}_{[\chi]}^{c},\mathcal{Q}_{[\chi]}^{c}) on XΓX_{\Gamma} (and on GG) by

(4.11) η𝒫wx=η𝒫w1sw1uwNswNux=ηsw1sηuw1uηswNsηuwNux\displaystyle\eta_{\mathcal{P}}^{w}x=\eta_{\mathcal{P}}^{w_{1}^{s}w_{1}^{u}...w_{N}^{s}w_{N}^{u}}x=\eta_{s}^{w_{1}^{s}}\eta_{u}^{w_{1}^{u}}...\eta_{s}^{w_{N}^{s}}\eta_{u}^{w_{N}^{u}}x

where ησ\eta_{\sigma} is defined in Definition 4.2 and shown to be an action in Lemma 4.2.

Remark 16.

We make no notational distinction between the action on XΓX_{\Gamma} and GG. It is clear that the action on GG covers the action on XΓX_{\Gamma} in the sense that the projection pΓ:GXΓp_{\Gamma}:G\to X_{\Gamma} intertwines the two actions.

Lemma 4.3.

We have Φ(η𝒫wx)=Φ(x)+Π(w)\Phi(\eta_{\mathcal{P}}^{w}x)=\Phi(x)+\Pi(w). That is Φ\Phi semiconjugates the 𝒫\mathcal{P}-action onto the translation action on 𝕋d\mathbb{T}^{d} (and d\mathbb{R}^{d}).

Proof.

By induction, it suffices to consider vE0σv\in E_{0}^{\sigma}. The Lemma follows from Lemma 4.2. ∎

Lemma 4.4.

For any w𝒫w\in\mathcal{P} the homeomorphism η𝒫w\eta_{\mathcal{P}}^{w} preserve the center foliation, WcW^{c}. Moreover, if w=vNsvNuv1sv1uw=v_{N}^{s}v_{N}^{u}...v_{1}^{s}v_{1}^{u} with vjσE0σv_{j}^{\sigma}\in E_{0}^{\sigma}, x0=xx_{0}=x and

x1u=ηuv1ux0,x1s=ηsv1sx1u,\displaystyle x_{1}^{u}=\eta_{u}^{v_{1}^{u}}x_{0},\quad x_{1}^{s}=\eta_{s}^{v_{1}^{s}}x_{1}^{u},
\displaystyle\vdots
xNu=ηuvNuxN1s,xNs=ηsvNsxNu\displaystyle x_{N}^{u}=\eta_{u}^{v_{N}^{u}}x_{N-1}^{s},\quad x_{N}^{s}=\eta_{s}^{v_{N}^{s}}x_{N}^{u}

then η𝒫wx=xNs\eta_{\mathcal{P}}^{w}x=x_{N}^{s} and the map η𝒫w:Wc(x)Wc(η𝒫wx)\eta_{\mathcal{P}}^{w}:W^{c}(x)\to W^{c}(\eta_{\mathcal{P}}^{w}x) coincide with the composition

(4.12) Wc(x0)πx,x1uuWc(x1u)πx1u,x1ssWc(x1s)πx1s,x2uuπxNu,xNssWc(xNs).\displaystyle W^{c}(x_{0})\xrightarrow{\pi_{x,x_{1}^{u}}^{u}}W^{c}(x_{1}^{u})\xrightarrow{\pi_{x_{1}^{u},x_{1}^{s}}^{s}}W^{c}(x_{1}^{s})\xrightarrow{\pi_{x_{1}^{s},x_{2}^{u}}^{u}}...\xrightarrow{\pi^{s}_{x_{N}^{u},x_{N}^{s}}}W^{c}(x_{N}^{s}).

That is, we have

(4.13) η𝒫w|Wc(x)=πxNu,xNssπxN1s,xNuuπxN1s,xN1usπx1u,x1ssπx0,x1uu.\displaystyle\eta_{\mathcal{P}}^{w}|_{W^{c}(x)}=\pi_{x_{N}^{u},x_{N}^{s}}^{s}\circ\pi_{x_{N-1}^{s},x_{N}^{u}}^{u}\circ\pi_{x_{N-1}^{s},x_{N-1}^{u}}^{s}\circ...\circ\pi_{x_{1}^{u},x_{1}^{s}}^{s}\circ\pi_{x_{0},x_{1}^{u}}^{u}.

If ff is rr-bunching then holonomies between center manifolds are CrC^{r}-smooth, so η𝒫w\eta_{\mathcal{P}}^{w} is CrC^{r} along WcW^{c} for all w𝒫w\in\mathcal{P}.

Proof.

By induction it suffices to consider w=vE0σw=v\in E_{0}^{\sigma} for σ=s,u\sigma=s,u. This was shown in the proof of Lemma 4.2. The regularity follows from [50]. ∎

Lemma 4.5.

We have w𝒫cw\in\mathcal{P}^{c} if and only if η𝒫wxW^c(x)\eta_{\mathcal{P}}^{w}x\in\hat{W}^{c}(x) for every xGx\in G.

Remark 17.

When w𝒫cw\in\mathcal{P}^{c} then Lemma 4.5 shows that η𝒫wWc(x)=Wc(x)\eta_{\mathcal{P}}^{w}W^{c}(x)=W^{c}(x) for every xx, so 𝒫c\mathcal{P}^{c} is the (homotopically trivial) center fixing part of 𝒫\mathcal{P}.

Proof.

The lemma is immediate from Φ(η𝒫wx)=Φ(x)+Π(w)\Phi(\eta_{\mathcal{P}}^{w}x)=\Phi(x)+\Pi(w) (Lemma 4.3). ∎

Lemma 4.6.

If ff is accessible, then the 𝒫c\mathcal{P}^{c}-action is transitive on W^c(x)\hat{W}^{c}(x) (and Wc(x)W^{c}(x)) for all xGx\in G (and xXΓx\in X_{\Gamma}).

Proof.

Since ff is accessible η𝒫𝒫xWc(x)\eta_{\mathcal{P}}^{\mathcal{P}}x\supset W^{c}(x) for every xXΓx\in X_{\Gamma}, and η𝒫wxWc(x)\eta_{\mathcal{P}}^{w}x\in W^{c}(x) if and only if Π(w)d\Pi(w)\in\mathbb{Z}^{d} (Lemma 4.3). Given 𝐧d\mathbf{n}\in\mathbb{Z}^{d} write 𝒫𝐧={w𝒫 : Π(w)=𝐧}\mathcal{P}_{\mathbf{n}}=\{w\in\mathcal{P}\text{ : }\Pi(w)=\mathbf{n}\}. Then 𝒫c=𝒫0\mathcal{P}^{c}=\mathcal{P}_{0}. It follows that

Wc(x)=𝐧dη𝒫𝒫𝐧x\displaystyle W^{c}(x)=\bigcup_{\mathbf{n}\in\mathbb{Z}^{d}}\eta_{\mathcal{P}}^{\mathcal{P}_{\mathbf{n}}}x

and since Wc(x)W^{c}(x) is uncountable and d\mathbb{Z}^{d} is countable there is at least one 𝐧0\mathbf{n}_{0} such that #η𝒫(𝒫𝐧0)x>1\#\eta_{\mathcal{P}}(\mathcal{P}_{\mathbf{n}_{0}})x>1. If we fix some w𝒫𝐧0w\in\mathcal{P}_{-\mathbf{n}_{0}} then

#η𝒫wη𝒫𝒫𝐧0x=#η𝒫w𝒫𝐧0x>1.\displaystyle\#\eta_{\mathcal{P}}^{w}\eta_{\mathcal{P}}^{\mathcal{P}_{\mathbf{n}_{0}}}x=\#\eta_{\mathcal{P}}^{w\mathcal{P}_{\mathbf{n}_{0}}}x>1.

For any w𝒫𝐧0w^{\prime}\in\mathcal{P}_{\mathbf{n}_{0}} we have Π(ww)=Π(w)+Π(w)=0\Pi(ww^{\prime})=\Pi(w)+\Pi(w^{\prime})=0 so w𝒫𝐧0𝒫cw\mathcal{P}_{\mathbf{n}_{0}}\subset\mathcal{P}^{c}, or #η𝒫𝒫cx>1\#\eta_{\mathcal{P}}^{\mathcal{P}^{c}}x>1. Given any w=w1sw1uwNswNu𝒫cw=w_{1}^{s}w_{1}^{u}...w_{N}^{s}w_{N}^{u}\in\mathcal{P}^{c} we define a path wt=(tw1s)(tw1u)(twNs)(twNu)𝒫cw_{t}=(tw_{1}^{s})(tw_{1}^{u})...(tw_{N}^{s})(tw_{N}^{u})\in\mathcal{P}^{c}, t[0,1]t\in[0,1], from 0 to ww, so η𝒫𝒫cx\eta_{\mathcal{P}}^{\mathcal{P}^{c}}x is path connected. The image I=η𝒫𝒫cxI=\eta_{\mathcal{P}}^{\mathcal{P}^{c}}x is an interval in Wc(x)W^{c}(x) since it contains at least 22 distinct points and is path connected. We claim that xx is an interior point in this interval. If xx is not an interior point, then I=[x,y),[x,y]I=[x,y),[x,y] or (y,x],[y,x](y,x],[y,x] for some yWc(x)y\in W^{c}(x). We will assume that one of the first two cases holds the other two cases are similar. Let xzIx\neq z\in I and let η𝒫wx=z\eta_{\mathcal{P}}^{w}x=z, then [x,z]I[x,z]\subset I. Since w1𝒫cw^{-1}\in\mathcal{P}^{c} and η𝒫w1\eta_{\mathcal{P}}^{w^{-1}} preserves orientation we have

(4.14) Iη𝒫w1[x,z]=[η𝒫w1x,x]\displaystyle I\supset\eta_{\mathcal{P}}^{w^{-1}}[x,z]=[\eta_{\mathcal{P}}^{w^{-1}}x,x]

which would imply η𝒫w1x=x\eta_{\mathcal{P}}^{w^{-1}}x=x if xx is an end point of II. After applying η𝒫w\eta_{\mathcal{P}}^{w} we obtain x=η𝒫wx=zx=\eta_{\mathcal{P}}^{w}x=z. This is a contradiction since we assumed zxz\neq x. The point xx is interior in η𝒫𝒫cx\eta_{\mathcal{P}}^{\mathcal{P}^{c}}x and xx was arbitrary, so the orbit of xx under η𝒫𝒫c\eta_{\mathcal{P}}^{\mathcal{P}_{c}} is open. This holds for every xx and Wc(x)W^{c}(x) is connected, so η𝒫𝒫cx=Wc(x)\eta_{\mathcal{P}}^{\mathcal{P}^{c}}x=W^{c}(x). The second part of the lemma follows.

The first claim follows from the second part. Indeed the second part implies that for any xGx\in G the 𝒫c\mathcal{P}^{c}-orbit of xx is open in W^c(x)\hat{W}^{c}(x). Connectedness of W^c(x)\hat{W}^{c}(x) implies the first part of the lemma. ∎

Recall that ρ:kGL(d,)\rho:\mathbb{Z}^{k}\to{\rm GL}(d,\mathbb{Z}) is defined by Φα𝐧=ρ𝐧Φ\Phi\alpha^{\mathbf{n}}=\rho^{\mathbf{n}}\Phi. Let ρ𝐧:𝒫𝒫\rho^{\mathbf{n}}:\mathcal{P}\to\mathcal{P} be the map

ρ𝐧(w1sw1uwNswNu)=(ρ𝐧w1s)(ρ𝐧w1u)(ρ𝐧wNs)(ρ𝐧wNu)\displaystyle\rho^{\mathbf{n}}(w_{1}^{s}w_{1}^{u}...w_{N}^{s}w_{N}^{u})=(\rho^{\mathbf{n}}w_{1}^{s})(\rho^{\mathbf{n}}w_{1}^{u})...(\rho^{\mathbf{n}}w_{N}^{s})(\rho^{\mathbf{n}}w_{N}^{u})

The following lemma is immediate from Lemma 4.2.

Lemma 4.7.

We have α𝐧η𝒫w=η𝒫ρ𝐧wα𝐧\alpha^{\mathbf{n}}\eta_{\mathcal{P}}^{w}=\eta_{\mathcal{P}}^{\rho^{\mathbf{n}}w}\alpha^{\mathbf{n}} and ρ𝐧\rho^{\mathbf{n}} preserve 𝒫c\mathcal{P}^{c}, 𝒫[χ]\mathcal{P}_{[\chi]}, 𝒬[χ]\mathcal{Q}_{[\chi]}, 𝒫[χ]c\mathcal{P}_{[\chi]}^{c}, and 𝒬[χ]c\mathcal{Q}_{[\chi]}^{c}.

5. An invariance principle for higher rank Anosov actions

Let XΓX_{\Gamma} be any nilmanifold and ρ:kAut(XΓ)\rho:\mathbb{Z}^{k}\to{\rm Aut}(X_{\Gamma}) a higher rank action. We will assume that ρ\rho is the restriction of some map Q:kAut(G)Q:\mathbb{R}^{k}\to{\rm Aut}(G). Let Φ𝒳:𝒳XΓ\Phi_{\mathcal{X}}:\mathcal{X}\to X_{\Gamma} be a Hölder fiber bundle over XΓX_{\Gamma} with fibers 𝒳x=Φ𝒳1(x)\mathcal{X}_{x}=\Phi_{\mathcal{X}}^{-1}(x) uniformly CrC^{r} for some r>1r>1 (we allow r(1,2)r\in(1,2)). We will assume throughout this section that 𝒳\mathcal{X} is compact, and therefore have compact fibers.

Definition 5.1.

We say that F:k×𝒳𝒳F:\mathbb{Z}^{k}\times\mathcal{X}\to\mathcal{X} is a cocycle over ρ\rho if FF is a k\mathbb{Z}^{k}-action covering ρ\rho. Moreover, FF is a CsC^{s}-cocycle if F(𝐧,):𝒳x𝒳ρ𝐧xF(\mathbf{n},\cdot):\mathcal{X}_{x}\to\mathcal{X}_{\rho^{\mathbf{n}}x} is uniformly CsC^{s}.

For a cocycle FF over ρ\rho we write F𝐧:𝒳x𝒳ρ𝐧xF^{\mathbf{n}}:\mathcal{X}_{x}\to\mathcal{X}_{\rho^{\mathbf{n}}x} for the map F(𝐧,)F(\mathbf{n},\cdot). For each coarse exponent [χ][\chi] of ρ\rho we have a translation action T[χ]:G[χ]×XΓXΓT_{[\chi]}:G^{[\chi]}\times X_{\Gamma}\to X_{\Gamma}, T[χ]gx=xg1T_{[\chi]}^{g}x=xg^{-1} for gG[χ]g\in G^{[\chi]}.

Definition 5.2.

A cocycle FF has [χ][\chi]-translations if there is a Hölder action η[χ]:G[χ]×𝒳𝒳\eta_{[\chi]}:G^{[\chi]}\times\mathcal{X}\to\mathcal{X} covering T[χ]T_{[\chi]} such that

(5.1) F𝐧η[χ]g=η[χ]ρ𝐧gF𝐧,\displaystyle F^{\mathbf{n}}\eta_{[\chi]}^{g}=\eta_{[\chi]}^{\rho^{\mathbf{n}}g}F^{\mathbf{n}},

for any 𝐧k\mathbf{n}\in\mathbb{Z}^{k}.

Our interest in cocycles over algebraic actions comes from the following lemma.

Lemma 5.1.

If α\alpha is as in Theorem A, with Φα𝐧=ρ𝐧Φ\Phi\alpha^{\mathbf{n}}=\rho^{\mathbf{n}}\Phi, then α\alpha is a C1+αC^{1+\alpha}-cocycle over ρ\rho that admit [χ][\chi]-translations for every coarse [χ][\chi] and η[χ]=η𝒫|E0[χ]\eta_{[\chi]}=\eta_{\mathcal{P}}|_{E_{0}^{[\chi]}}.

Proof.

The lemma is immediate from Theorem 1.1 and Lemma 4.3. ∎

The main result of this section is a sufficient condition for the translation action η[χ]\eta_{[\chi]} to preserve a FF-invariant measure.

Theorem 5.1.

Let FF be a CsC^{s}-cocycle, s>1s>1, over ρ\rho. Let ν\nu be a FF-invariant probability measure projecting onto μΓ\mu_{\Gamma}, (Φ𝒳)ν=μΓ(\Phi_{\mathcal{X}})_{*}\nu=\mu_{\Gamma}, and λF,ν1,,λF,νN:k\lambda_{F,\nu}^{1},...,\lambda_{F,\nu}^{N}:\mathbb{Z}^{k}\to\mathbb{R} the ν\nu-Lyapunov exponents of FF along the fibers of 𝒳\mathcal{X}. If

(5.2) i=1NkerλF,νiker[χ]\displaystyle\bigcap_{i=1}^{N}\ker\lambda_{F,\nu}^{i}\not\subset\ker[\chi]

then ν\nu is η[χ]\eta_{[\chi]}-invariant.

Remark 18.

In Theorem 5.1 we assume that ρ\rho is the restriction of some homomorphism Q:kAut(G)Q:\mathbb{R}^{k}\to{\rm Aut}(G). This is without loss of generality after possibly dropping to a finite index subgroup.

To apply results from [3] it will be convenient to reformulate η[χ]\eta_{[\chi]}-invariance of ν\nu into essential holonomy invariance. Let xXΓx\in X_{\Gamma}, gG[χ]g\in G^{[\chi]} and y=T[χ]gxy=T_{[\chi]}^{g}x. Since η[χ]g\eta_{[\chi]}^{g} cover T[χ]gT_{[\chi]}^{g}, we define

(5.3) hx,y[χ]:𝒳x𝒳y,hx,y[χ](ξ)=η[χ]g(ξ).\displaystyle h_{x,y}^{[\chi]}:\mathcal{X}_{x}\to\mathcal{X}_{y},\quad h_{x,y}^{[\chi]}(\xi)=\eta_{[\chi]}^{g}(\xi).

We say that hx,y[χ]h_{x,y}^{[\chi]} is the [χ][\chi]-holonomy between 𝒳x\mathcal{X}_{x} and 𝒳y\mathcal{X}_{y}.

Definition 5.3.

Let ν\nu be FF-invariant such that (Φ𝒳)ν=μΓ(\Phi_{\mathcal{X}})_{*}\nu=\mu_{\Gamma} and let {νx}xXΓ\{\nu_{x}\}_{x\in X_{\Gamma}} be the disintegration of ν\nu over Φ𝒳\Phi_{\mathcal{X}}. We say that ν\nu, or {νx}xXΓ\{\nu_{x}\}_{x\in X_{\Gamma}}, is essentially [χ][\chi]-holonomy invariant if there is a μΓ\mu_{\Gamma}-full measure set YXΓY\subset X_{\Gamma} such that (hx,y[χ])νx=νy(h_{x,y}^{[\chi]})_{*}\nu_{x}=\nu_{y} for x,yYx,y\in Y.

Lemma 5.2.

Let ν\nu be FF-invariant and projecting onto μΓ\mu_{\Gamma}. If ν\nu is essentially [χ][\chi]-holonomy invariant then ν\nu is η[χ]\eta_{[\chi]}-invariant.

Proof.

Let YXΓY\subset X_{\Gamma} be a full measure subset such that (hx,y[χ])νx=νy(h_{x,y}^{[\chi]})_{*}\nu_{x}=\nu_{y} for x,yYx,y\in Y. Let gG[χ]g\in G^{[\chi]} and Y~=YT[χ]g1Y\tilde{Y}=Y\cap T_{[\chi]}^{g^{-1}}Y so that x,T[χ]gxYx,T_{[\chi]}^{g}x\in Y for xY~x\in\tilde{Y}. If φC0(𝒳)\varphi\in C^{0}(\mathcal{X}) then

𝒳φ(η[χ]gξ)dν(ξ)=\displaystyle\int_{\mathcal{X}}\varphi(\eta_{[\chi]}^{g}\xi){\rm d}\nu(\xi)= Y~(𝒳xφ(η[χ]gξ)dνx(ξ))dμΓ(x)=\displaystyle\int_{\tilde{Y}}\left(\int_{\mathcal{X}_{x}}\varphi(\eta_{[\chi]}^{g}\xi){\rm d}\nu_{x}(\xi)\right){\rm d}\mu_{\Gamma}(x)=
Y~(η[χ]g𝒳xφ(ξ)d(hx,T[χ]gx[χ])νx(ξ))dμΓ(x)=\displaystyle\int_{\tilde{Y}}\left(\int_{\eta_{[\chi]}^{g}\mathcal{X}_{x}}\varphi(\xi){\rm d}(h_{x,T_{[\chi]}^{g}x}^{[\chi]})_{*}\nu_{x}(\xi)\right){\rm d}\mu_{\Gamma}(x)=
Y~(𝒳T[χ]gxφ(ξ)dνT[χ]gx(ξ))dμΓ(x)=\displaystyle\int_{\tilde{Y}}\left(\int_{\mathcal{X}_{T_{[\chi]}^{g}x}}\varphi(\xi){\rm d}\nu_{T_{[\chi]}^{g}x}(\xi)\right){\rm d}\mu_{\Gamma}(x)=
𝒳φ(ξ)dν(ξ),\displaystyle\int_{\mathcal{X}}\varphi(\xi){\rm d}\nu(\xi),

so ν\nu is η[χ]g\eta_{[\chi]}^{g}-invariant. ∎

5.1. The suspension construction

Fix a higher rank action ρ:kAut(XΓ)\rho:\mathbb{Z}^{k}\to{\rm Aut}(X_{\Gamma}), a cocycle F𝐧:𝒳𝒳F^{\mathbf{n}}:\mathcal{X}\to\mathcal{X} over ρ\rho, and a measure ν\nu as in Theorem 5.1. We recall the definition of the suspension of an action α:k×MM\alpha:\mathbb{Z}^{k}\times M\to M.

Definition 5.4.

Let τ:k×(M×k)M×k\tau:\mathbb{Z}^{k}\times(M\times\mathbb{R}^{k})\to M\times\mathbb{R}^{k} be defined by τ𝐧(x,𝐬)=(α𝐧x,𝐬𝐧)\tau^{\mathbf{n}}(x,\mathbf{s})=(\alpha^{\mathbf{n}}x,\mathbf{s}-\mathbf{n}). We define the suspension 𝒮\mathcal{S} of α\alpha as

𝒮:=(M×k)/τ.\displaystyle\mathcal{S}:=\left(M\times\mathbb{R}^{k}\right)/\tau.

Given (x,𝐬)M×k(x,\mathbf{s})\in M\times\mathbb{R}^{k} we denote by [x,𝐬][x,\mathbf{s}] the equivalence class of (x,𝐬)(x,\mathbf{s}) in 𝒮\mathcal{S}. We also define a natural action on 𝒮\mathcal{S} by α𝒮𝐭(x,𝐬)=(x,𝐬+𝐭)\alpha_{\mathcal{S}}^{\mathbf{t}}(x,\mathbf{s})=(x,\mathbf{s}+\mathbf{t}). Since

α𝒮𝐭τ𝐧(x,𝐬)=\displaystyle\alpha_{\mathcal{S}}^{\mathbf{t}}\tau^{\mathbf{n}}(x,\mathbf{s})= α𝒮𝐭(α𝐧x,𝐬𝐧)=(α𝐧x,𝐭+𝐬𝐧)=τ𝐧(x,𝐭+𝐬)=\displaystyle\alpha_{\mathcal{S}}^{\mathbf{t}}(\alpha^{\mathbf{n}}x,\mathbf{s}-\mathbf{n})=(\alpha^{\mathbf{n}}x,\mathbf{t}+\mathbf{s}-\mathbf{n})=\tau^{\mathbf{n}}(x,\mathbf{t}+\mathbf{s})=
τ𝐧α𝒮𝐭(x,𝐬)\displaystyle\tau^{\mathbf{n}}\alpha_{\mathcal{S}}^{\mathbf{t}}(x,\mathbf{s})

the action α𝒮\alpha_{\mathcal{S}} descends to an action on 𝒮\mathcal{S}. Moreover the map M×k(x,𝐬)𝐬+k𝕋kM\times\mathbb{R}^{k}\ni(x,\mathbf{s})\mapsto\mathbf{s}+\mathbb{Z}^{k}\in\mathbb{T}^{k} descends to a map π𝒮:𝒮𝕋k\pi_{\mathcal{S}}:\mathcal{S}\to\mathbb{T}^{k} with fibers MM. The map π𝒮\pi_{\mathcal{S}} semi-conjugates α𝒮\alpha_{\mathcal{S}} to the natural translation action on 𝕋k\mathbb{T}^{k}.

Given any α\alpha-invariant measure μ\mu on MM we define a measure μ𝒮\mu_{\mathcal{S}} on 𝒮\mathcal{S} as follows. For each x𝕋kx\in\mathbb{T}^{k} we choose some 𝐬k\mathbf{s}\in\mathbb{R}^{k} such that 𝐬+k=x\mathbf{s}+\mathbb{Z}^{k}=x. Let ι𝐬:Mπ𝒮1(x)𝒮\iota_{\mathbf{s}}:M\to\pi_{\mathcal{S}}^{-1}(x)\subset\mathcal{S} be defined by ι𝐬(y)=[y,𝐬]\iota_{\mathbf{s}}(y)=[y,\mathbf{s}]. Define a measure μx\mu_{x} on π𝒮1(x)\pi_{\mathcal{S}}^{-1}(x) by

(5.4) (ι𝐬)μ=μx.\displaystyle(\iota_{\mathbf{s}})_{*}\mu=\mu_{x}.

Given any 𝐧k\mathbf{n}\in\mathbb{Z}^{k} we have ι𝐬+𝐧(y)=[y,𝐬+𝐧]=[α𝐧y,𝐬]=ι𝐬α𝐧y\iota_{\mathbf{s}+\mathbf{n}}(y)=[y,\mathbf{s}+\mathbf{n}]=[\alpha^{\mathbf{n}}y,\mathbf{s}]=\iota_{\mathbf{s}}\alpha^{\mathbf{n}}y. So α\alpha-invariance of μ\mu implies

(ι𝐬+𝐧)μ=(ι𝐬)α𝐧μ=(ι𝐬)μ\displaystyle(\iota_{\mathbf{s}+\mathbf{n}})_{*}\mu=(\iota_{\mathbf{s}})_{*}\alpha^{\mathbf{n}}_{*}\mu=(\iota_{\mathbf{s}})_{*}\mu

showing that μx\mu_{x} is well-defined. Define a suspended measure μ𝒮\mu_{\mathcal{S}} by

(5.5) μ𝒮=𝕋kμxdvol𝕋k(x).\displaystyle\mu_{\mathcal{S}}=\int_{\mathbb{T}^{k}}\mu_{x}{\rm d}{\rm vol}_{\mathbb{T}^{k}}(x).

One checks that μ𝒮\mu_{\mathcal{S}} is α𝒮\alpha_{\mathcal{S}}-invariant.

In the remainder of this section we denote by 𝒮0\mathcal{S}_{0} the suspension of ρ\rho with action ρ𝒮0:k×𝒮0𝒮0\rho_{\mathcal{S}_{0}}:\mathbb{R}^{k}\times\mathcal{S}_{0}\to\mathcal{S}_{0} and by 𝒮\mathcal{S} the suspension of FF with action F𝒮:k×𝒮𝒮F_{\mathcal{S}}:\mathbb{R}^{k}\times\mathcal{S}\to\mathcal{S}. We also denote by μ𝒮0\mu_{\mathcal{S}_{0}} the suspension of μΓ\mu_{\Gamma} and ν𝒮\nu_{\mathcal{S}} the suspension of ν\nu. Note that μ𝒮0\mu_{\mathcal{S}_{0}} is a volume on 𝒮0\mathcal{S}_{0}. Let Q:kAut(G)Q:\mathbb{R}^{k}\to{\rm Aut}(G) be a homomorphism such that

(5.6) Q|k=ρ.\displaystyle Q|_{\mathbb{Z}^{k}}=\rho.

We suspend the actions T[χ]gT_{[\chi]}^{g} and η[χ]g\eta_{[\chi]}^{g} as

(5.7) T^[χ]g([x,𝐬])=[T[χ]Q𝐬gx,𝐬],η^[χ]g([ξ,𝐬])=[η[χ]Q𝐬gξ,𝐬].\displaystyle\hat{T}_{[\chi]}^{g}\left([x,\mathbf{s}]\right)=\left[T_{[\chi]}^{Q^{-\mathbf{s}}g}x,\mathbf{s}\right],\quad\hat{\eta}_{[\chi]}^{g}\left([\xi,\mathbf{s}]\right)=\left[\eta_{[\chi]}^{Q^{-\mathbf{s}}g}\xi,\mathbf{s}\right].

For any 𝐧k\mathbf{n}\in\mathbb{Z}^{k} we have

(5.8) T^[χ]g([ρ𝐧x,𝐬𝐧])=[T[χ]Q𝐧Q𝐬gρ𝐧x,𝐬𝐧]=[ρ𝐧T[χ]Q𝐬gx,𝐬𝐧]\displaystyle\hat{T}_{[\chi]}^{g}([\rho^{\mathbf{n}}x,\mathbf{s}-\mathbf{n}])=\left[T_{[\chi]}^{Q^{\mathbf{n}}Q^{-\mathbf{s}}g}\rho^{\mathbf{n}}x,\mathbf{s}-\mathbf{n}\right]=\left[\rho^{\mathbf{n}}T_{[\chi]}^{Q^{-\mathbf{s}}g}x,\mathbf{s}-\mathbf{n}\right]

so T^[χ]\hat{T}_{[\chi]} is a well-defined action on 𝒮0\mathcal{S}_{0} that acts in the fibers of π𝒮0:𝒮0𝕋k\pi_{\mathcal{S}_{0}}:\mathcal{S}_{0}\to\mathbb{T}^{k}. Similarly the action η^[χ]\hat{\eta}_{[\chi]} is well-defined on 𝒮\mathcal{S}. Define Φ𝒮:𝒮𝒮0\Phi_{\mathcal{S}}:\mathcal{S}\to\mathcal{S}_{0} by

(5.9) Φ𝒮([ξ,𝐬])=[Φ𝒳(ξ),𝐬].\displaystyle\Phi_{\mathcal{S}}([\xi,\mathbf{s}])=[\Phi_{\mathcal{X}}(\xi),\mathbf{s}].

Since Φ𝒳\Phi_{\mathcal{X}} semi-conjugates ρ\rho onto FF, Φ𝒮\Phi_{\mathcal{S}} is well-defined. The following lemma is immediate from our definitions.

Lemma 5.3.

Let Φ𝒮\Phi_{\mathcal{S}}, F𝒮F_{\mathcal{S}}, ρ𝒮0\rho_{\mathcal{S}_{0}}, T^[χ]\hat{T}_{[\chi]} and η^[χ]\hat{\eta}_{[\chi]} be as above. The following holds

  1. (i)

    the map Φ𝒮:𝒮𝒮0\Phi_{\mathcal{S}}:\mathcal{S}\to\mathcal{S}_{0} is a Hölder fiber bundle with uniformly CrC^{r} fibers. In fact, for any 𝐬𝕋k\mathbf{s}\in\mathbb{T}^{k} the restriction Φ𝒮|π𝒮1(𝐬)\Phi_{\mathcal{S}}|_{\pi_{\mathcal{S}}^{-1}(\mathbf{s})} coincides with Φ𝒳\Phi_{\mathcal{X}} using natural identifications of π𝒮1(𝐬)𝒳\pi_{\mathcal{S}}^{-1}(\mathbf{s})\cong\mathcal{X} and π𝒮01(𝐬)XΓ\pi_{\mathcal{S}_{0}}^{-1}(\mathbf{s})\cong X_{\Gamma},

  2. (ii)

    the map F𝒮F_{\mathcal{S}} is a CsC^{s} cocycle over ρ𝒮0\rho_{\mathcal{S}_{0}},

  3. (iii)

    the Lyapunov exponents for ρ𝒮0\rho_{\mathcal{S}_{0}} coincide with the Lyapunov exponents of ρ\rho,

  4. (iv)

    the ν𝒮\nu_{\mathcal{S}}-Lyapunov exponents along the fibers of Φ𝒮\Phi_{\mathcal{S}} for F𝒮F_{\mathcal{S}} coincide with the fiberwise Lyapunov exponents of FF,

  5. (v)

    the map Φ𝒮\Phi_{\mathcal{S}} semiconjugates η^[χ]\hat{\eta}_{[\chi]} to T^[χ]\hat{T}_{[\chi]}.

Proof.

Point (i)(i) follows from the analogous properties of Φ𝒳\Phi_{\mathcal{X}} since Φ𝒮\Phi_{\mathcal{S}} is defined in the fibers of π𝒮:𝒮𝕋k\pi_{\mathcal{S}}:\mathcal{S}\to\mathbb{T}^{k}. That Φ𝒮\Phi_{\mathcal{S}} conjugates F𝒮F_{\mathcal{S}} to ρ𝒮0\rho_{\mathcal{S}_{0}} is immediate from its definition: Φ𝒮[ξ,𝐬+𝐭]=[Φ𝒳(ξ),𝐬+𝐭]\Phi_{\mathcal{S}}[\xi,\mathbf{s}+\mathbf{t}]=[\Phi_{\mathcal{X}}(\xi),\mathbf{s}+\mathbf{t}]. That F𝒮F_{\mathcal{S}} is CsC^{s} along the fibers of Φ𝒮\Phi_{\mathcal{S}} is immediate since FF is CsC^{s} along the fibers of Φ𝒳\Phi_{\mathcal{X}} (note that the identifications ι𝐬(ξ)=[ξ,𝐬]\iota_{\mathbf{s}}(\xi)=[\xi,\mathbf{s}], 𝒳π𝒮1(𝐬+k)\mathcal{X}\to\pi_{\mathcal{S}}^{-1}(\mathbf{s}+\mathbb{Z}^{k}), defines a smooth structure on the fibers of Φ𝒮\Phi_{\mathcal{S}} in which F𝒮F_{\mathcal{S}} is uniformly CsC^{s}). Points (iii)(iii) and (iv)(iv) holds for kk\mathbb{Z}^{k}\subset\mathbb{R}^{k}, and any functional is determined by its values on a lattice, proving (iii)(iii) and (iv)(iv). Point (v)(v) is immediate from the definitions and the fact that Φ𝒳\Phi_{\mathcal{X}} conjugate η[χ]\eta_{[\chi]} to T[χ]T_{[\chi]}. ∎

Define holonomies along the orbits of T^[χ]\hat{T}_{[\chi]} as in Equation 5.3. That is, if x𝒮0x\in\mathcal{S}_{0}, y=T^[χ]gxy=\hat{T}_{[\chi]}^{g}x then

(5.10) h^x,y[χ](ξ)=η^[χ]g(ξ).\displaystyle\hat{h}_{x,y}^{[\chi]}(\xi)=\hat{\eta}_{[\chi]}^{g}(\xi).

We say that ν𝒮\nu_{\mathcal{S}} (or the disintegration {ν𝒮,x}x𝒮0\{\nu_{\mathcal{S},x}\}_{x\in\mathcal{S}_{0}}) is essentially [χ][\chi]-holonomy invariant if there is a μ𝒮0\mu_{\mathcal{S}_{0}} full measure set Y𝒮0Y\subset\mathcal{S}_{0} such that

(5.11) (h^x,y[χ])ν𝒮,x=ν𝒮,y,x,yY.\displaystyle(\hat{h}_{x,y}^{[\chi]})_{*}\nu_{\mathcal{S},x}=\nu_{\mathcal{S},y},\quad x,y\in Y.

The key fact about the suspension, is that holonomy invariance of ν𝒮\nu_{\mathcal{S}} implies holonomy invariance of ν\nu. So Theorem 5.1 follows by [χ][\chi]-holonomy invariance of ν𝒮\nu_{\mathcal{S}} (by Lemma 5.2).

Lemma 5.4.

The measure ν\nu is essentially [χ][\chi]-holonomy invariant if and only if ν𝒮\nu_{\mathcal{S}} is essentially [χ][\chi]-holonomy invariant.

Lemma 5.4 is immediate from the following lemma.

Lemma 5.5.

We have (Φ𝒮)ν𝒮=μ𝒮0(\Phi_{\mathcal{S}})_{*}\nu_{\mathcal{S}}=\mu_{\mathcal{S}_{0}}. The disintegration of ν𝒮\nu_{\mathcal{S}} over μ𝒮0\mu_{\mathcal{S}_{0}} is given by

(5.12) ν𝒮,[x,𝐬]=(ι𝐬)νx,[x,𝐬]𝒮0\displaystyle\nu_{\mathcal{S},[x,\mathbf{s}]}=(\iota_{\mathbf{s}})_{*}\nu_{x},\quad[x,\mathbf{s}]\in\mathcal{S}_{0}

where ν\nu is the disintegration of ν\nu over μ\mu.

Proof.

By construction we have ν𝒮=(ι𝐬)νd𝐬\nu_{\mathcal{S}}=(\iota_{\mathbf{s}})_{*}\nu\otimes{\rm d}\mathbf{s}, so

(Φ𝒮)ν𝒮=\displaystyle(\Phi_{\mathcal{S}})_{*}\nu_{\mathcal{S}}= 𝕋k(Φ𝒮)(ι𝐬)νd𝐬=𝕋k(Φ𝒮ι𝐬)νd𝐬=𝕋k(ι𝐬Φ𝒳)νd𝐬=\displaystyle\int_{\mathbb{T}^{k}}(\Phi_{\mathcal{S}})_{*}(\iota_{\mathbf{s}})_{*}\nu{\rm d}\mathbf{s}=\int_{\mathbb{T}^{k}}(\Phi_{\mathcal{S}}\iota_{\mathbf{s}})_{*}\nu{\rm d}\mathbf{s}=\int_{\mathbb{T}^{k}}(\iota_{\mathbf{s}}\Phi_{\mathcal{X}})_{*}\nu{\rm d}\mathbf{s}=
𝕋k(ι𝐬)μd𝐬=μ𝒮0.\displaystyle\int_{\mathbb{T}^{k}}(\iota_{\mathbf{s}})_{*}\mu{\rm d}\mathbf{s}=\mu_{\mathcal{S}_{0}}.

If we define ν𝒮,[x,𝐬]:=(ι𝐬)νx\nu_{\mathcal{S},[x,\mathbf{s}]}:=(\iota_{\mathbf{s}})_{*}\nu_{x} then

(5.13) ν𝒮,[ρ𝐧x,𝐬𝐧]=(ι𝐬𝐧)νρ𝐧x=(ι𝐬F𝐧)νρ𝐧x=(ι𝐬)νx=ν𝒮,[x,𝐬]\displaystyle\nu_{\mathcal{S},[\rho^{\mathbf{n}}x,\mathbf{s}-\mathbf{n}]}=(\iota_{\mathbf{s}-\mathbf{n}})_{*}\nu_{\rho^{\mathbf{n}}x}=(\iota_{\mathbf{s}}F^{\mathbf{-n}})_{*}\nu_{\rho^{\mathbf{n}}x}=(\iota_{\mathbf{s}})_{*}\nu_{x}=\nu_{\mathcal{S},[x,\mathbf{s}]}

where we have used ι𝐬+𝐧(ξ)=[ξ,𝐬+𝐧]=[F𝐧F𝐧ξ,𝐬+𝐧]=[ρ𝐧ξ,𝐬]=ι𝐬(ρ𝐧ξ)\iota_{\mathbf{s}+\mathbf{n}}(\xi)=[\xi,\mathbf{s}+\mathbf{n}]=[F^{-\mathbf{n}}F^{\mathbf{n}}\xi,\mathbf{s}+\mathbf{n}]=[\rho^{\mathbf{n}}\xi,\mathbf{s}]=\iota_{\mathbf{s}}(\rho^{\mathbf{n}}\xi) for ξ𝒳\xi\in\mathcal{X}. So ν𝒮,[x,𝐬]\nu_{\mathcal{S},[x,\mathbf{s}]} is well-defined. We calculate

𝒮0ν𝒮,[x,𝐬]dμ𝒮0([x,𝐬])=\displaystyle\int_{\mathcal{S}_{0}}\nu_{\mathcal{S},[x,\mathbf{s}]}{\rm d}\mu_{\mathcal{S}_{0}}([x,\mathbf{s}])= 𝒮0(ι𝐬)νxdμ𝒮0([x,𝐬])=\displaystyle\int_{\mathcal{S}_{0}}(\iota_{\mathbf{s}})_{*}\nu_{x}{\rm d}\mu_{\mathcal{S}_{0}}([x,\mathbf{s}])=
𝕋k[ι𝐬XΓ(ι𝐬)νxd(ι𝐬)μΓ([x,𝐬])]d𝐬=\displaystyle\int_{\mathbb{T}^{k}}\left[\int_{\iota_{\mathbf{s}}X_{\Gamma}}(\iota_{\mathbf{s}})_{*}\nu_{x}{\rm d}(\iota_{\mathbf{s}})_{*}\mu_{\Gamma}([x,\mathbf{s}])\right]{\rm d}\mathbf{s}=
𝕋k[(ι𝐬)XΓνxdμΓ(x)]d𝐬=\displaystyle\int_{\mathbb{T}^{k}}\left[(\iota_{\mathbf{s}})_{*}\int_{X_{\Gamma}}\nu_{x}{\rm d}\mu_{\Gamma}(x)\right]{\rm d}\mathbf{s}=
𝕋k(ι𝐬)νd𝐬=ν𝒮\displaystyle\int_{\mathbb{T}^{k}}(\iota_{\mathbf{s}})_{*}\nu{\rm d}\mathbf{s}=\nu_{\mathcal{S}}

which proves that ν𝒮,[x,𝐬]\nu_{\mathcal{S},[x,\mathbf{s}]} is a disintegration of ν𝒮\nu_{\mathcal{S}} over Φ𝒮\Phi_{\mathcal{S}}. ∎

Proof of Lemma 5.4.

If ν\nu is essentially [χ][\chi]-holonomy invariant, then we find YXΓY\subset X_{\Gamma} such that (hx,y[χ])νx=νy(h_{x,y}^{[\chi]})_{*}\nu_{x}=\nu_{y} for x,yYx,y\in Y. Letting Y~𝒮0\tilde{Y}\subset\mathcal{S}_{0} be the image of Y×kY\times\mathbb{R}^{k} in 𝒮0\mathcal{S}_{0}, one direction in Lemma 5.4 follows from the formula in Lemma 5.5. For the converse direction, let Y𝒮0Y\subset\mathcal{S}_{0} be such that (h^x,y[χ])ν𝒮,x=ν𝒮,y(\hat{h}_{x,y}^{[\chi]})_{*}\nu_{\mathcal{S},x}=\nu_{\mathcal{S},y} for x,yYx,y\in Y. Since μ𝒮0(Y)=1\mu_{\mathcal{S}_{0}}(Y)=1 we have

μΓ(ι𝐬1(Yπ𝒮01(𝐬)))=1\displaystyle\mu_{\Gamma}(\iota_{\mathbf{s}}^{-1}(Y\cap\pi_{\mathcal{S}_{0}}^{-1}(\mathbf{s})))=1

for d𝐬{\rm d}\mathbf{s}-almost every 𝐬𝕋k\mathbf{s}\in\mathbb{T}^{k}. For any 𝐬k\mathbf{s}\in\mathbb{R}^{k}, gG[χ]g\in G^{[\chi]} and xXΓx\in X_{\Gamma} we have

ι𝐬T[χ]g(x)=[T[χ]g(x),𝐬]=[T[χ]Q𝐬Q𝐬g(x),𝐬]=T^[χ]Q𝐬g[x,𝐬]=T^[χ]Q𝐬gι𝐬(x).\displaystyle\iota_{\mathbf{s}}T_{[\chi]}^{g}(x)=\left[T_{[\chi]}^{g}(x),\mathbf{s}\right]=\left[T_{[\chi]}^{Q^{-\mathbf{s}}Q^{\mathbf{s}}g}(x),\mathbf{s}\right]=\hat{T}_{[\chi]}^{Q^{\mathbf{s}}g}\left[x,\mathbf{s}\right]=\hat{T}_{[\chi]}^{Q^{\mathbf{s}}g}\iota_{\mathbf{s}}(x).

Similarly, ι𝐬η[χ]g=η^[χ]Q𝐬gι𝐬\iota_{\mathbf{s}}\eta_{[\chi]}^{g}=\hat{\eta}_{[\chi]}^{Q^{\mathbf{s}}g}\iota_{\mathbf{s}}. Since ι𝐬\iota_{\mathbf{s}} maps fibers of Φ𝒳\Phi_{\mathcal{X}} to fibers of Φ𝒮\Phi_{\mathcal{S}}, it follows that ι𝐬1h^[x,𝐬],[y,𝐬][χ]ι𝐬=hx,y[χ]\iota_{\mathbf{s}}^{-1}\hat{h}_{[x,\mathbf{s}],[y,\mathbf{s}]}^{[\chi]}\iota_{\mathbf{s}}=h_{x,y}^{[\chi]}. With Y~=ι𝐬1(Yπ𝒮01(𝐬))\tilde{Y}=\iota_{\mathbf{s}}^{-1}(Y\cap\pi_{\mathcal{S}_{0}}^{-1}(\mathbf{s})) and Lemma 5.5

(hx,y[χ])νx=\displaystyle(h_{x,y}^{[\chi]})_{*}\nu_{x}= (ι𝐬1h^[x,𝐬],[y,𝐬][χ]ι𝐬)νx=(ι𝐬1h^[x,𝐬],[y,𝐬][χ])ν𝒮,[x,𝐬]=\displaystyle\left(\iota_{\mathbf{s}}^{-1}\hat{h}_{[x,\mathbf{s}],[y,\mathbf{s}]}^{[\chi]}\iota_{\mathbf{s}}\right)_{*}\nu_{x}=\left(\iota_{\mathbf{s}}^{-1}\hat{h}_{[x,\mathbf{s}],[y,\mathbf{s}]}^{[\chi]}\right)_{*}\nu_{\mathcal{S},[x,\mathbf{s}]}=
(ι𝐬1)ν𝒮,[y,𝐬]=νy,\displaystyle(\iota_{\mathbf{s}}^{-1})_{*}\nu_{\mathcal{S},[y,\mathbf{s}]}=\nu_{y},

for x,yY~x,y\in\tilde{Y}. Choosing 𝐬\mathbf{s} such that μΓ(Y~)=1\mu_{\Gamma}(\tilde{Y})=1, we see that ν\nu is essentially [χ][\chi]-holonomy invariant. ∎

5.2. Proof of Theorem 5.1

By Lemmas 5.2 and 5.4 it suffices to show that the disintegration of the suspension of ν\nu is essentially [χ][\chi]-holonomy invariant. We will use the following general criteria for obtaining holonomy invariance, proved in [3, Proposition 4.2] (or [3, Corollary 4.3]).

Theorem 5.2.

Let f:MMf:M\to M be a volume preserving diffeomorphism on a closed, smooth manifold with an invariant contracting smooth foliation WW. Let 𝒳M\mathcal{X}\to M be a Hölder fiber bundle with uniformly CrC^{r} fibers and F:𝒳𝒳F:\mathcal{X}\to\mathcal{X} a map covering ff such that FF is uniformly CsC^{s} along the fibers of 𝒳M\mathcal{X}\to M. Assume that WW admits holonomies in 𝒳\mathcal{X}, that is for every yW(x)y\in W(x) there is a map hx,yW:𝒳x𝒳yh_{x,y}^{W}:\mathcal{X}_{x}\to\mathcal{X}_{y} satisfying (sh1)(sh1), (sh2)(sh2) and (sh3)(sh3) in [3, Section 2.4]. Let ν\nu be an FF-invariant measure on 𝒳\mathcal{X} projecting onto volume. If the ν\nu-exponents of FF along the fibers of 𝒳M\mathcal{X}\to M are 0 then the disintegration of ν\nu is essentially WW-holonomy invariant. That is, there is a full volume set YMY\subset M such that (hx,yW)νx=νy(h_{x,y}^{W})_{*}\nu_{x}=\nu_{y} for x,yYx,y\in Y.

Proof.

The theorem would follow immediately from [3, Proposition 4.2] if the foliation WW coincided with the stable foliation of ff (in the sense of [3, Section 4.1]). However, following the proof, it suffices that WW is contracting. In fact, since WW is a contracting foliation it is standard to produce a measurable partition subordinate to WW (see for example [45]) which simplifies the proof. ∎

Proof of Theorem 5.1.

By assumption we find 𝐭0k0\mathbf{t}_{0}\in\mathbb{R}^{k}\setminus 0 such that F𝒮𝐭0F_{\mathcal{S}}^{\mathbf{t}_{0}} has zero exponents along the fibers of Φ𝒮\Phi_{\mathcal{S}} and [χ](𝐭0)<0[\chi](\mathbf{t}_{0})<0. Let f=ρ𝒮0𝐭0f=\rho_{\mathcal{S}_{0}}^{\mathbf{t}_{0}}, F=F𝒮𝐭0F=F_{\mathcal{S}}^{\mathbf{t}_{0}}, WW be the orbit foliation of T^[χ]\hat{T}_{[\chi]}, and apply Theorem 5.2 to conclude that ν𝒮\nu_{\mathcal{S}} is essentially [χ][\chi]-holonomy invariant. The theorem follows from Lemmas 5.2 and 5.4. ∎

6. Invariant structure in the center direction

Let α:k×XΓXΓ\alpha:\mathbb{Z}^{k}\times X_{\Gamma}\to X_{\Gamma} be a smooth action satisfying the assumptions of Theorem A. In this section we prove Theorem 1.3: ff, and α\alpha, have a unique measure of maximal entropy. Moreover, if μ\mu is the measure of maximal entropy then Φμ=vol\Phi_{*}\mu={\rm vol} and the disintegration of μ\mu is invariant under stable and unstable holonomy. Equivalently [3, 56] we show that the μ\mu-center exponent vanish λμc=0\lambda_{\mu}^{c}=0. The proof of Theorem 1.3 is by contradiction, so we assume that λμc(f)0\lambda_{\mu}^{c}(f)\neq 0. The proof splits into two cases. First, we have a generic case when the kernel of λμc:k\lambda_{\mu}^{c}:\mathbb{Z}^{k}\to\mathbb{R} does not coincide with the kernel of some exponent χ:k\chi:\mathbb{Z}^{k}\to\mathbb{R} of ρ\rho. Second, we have an exceptional case when kerλμc=kerχ\ker\lambda_{\mu}^{c}=\ker\chi for some exponent χ\chi of ρ\rho. The first case is dealt with by using Theorem 5.1 and Lemma 5.1. The second, more technical, case is dealt with by studying the circle dynamics induced by the holonomy maps on the center leaves. Suppose that λμc\lambda_{\mu}^{c} has the same kernel as [χ][\chi]. We begin by showing that the 𝒫[χ]\mathcal{P}_{[\chi]}-action commute with the 𝒬[χ]\mathcal{Q}_{[\chi]}-action. This implies that either 𝒫[χ]c\mathcal{P}_{[\chi]}^{c} or 𝒬[χ]c\mathcal{Q}_{[\chi]}^{c} act transitively on center leaves (see Lemma 6.2). If 𝒬[χ]c\mathcal{Q}_{[\chi]}^{c} act transitively on center leaves then Theorem 5.1 can be applied, as in the generic case. If 𝒫[χ]c\mathcal{P}_{[\chi]}^{c} acts transitively on center leaves, then we show that 𝒬[χ]\mathcal{Q}_{[\chi]} acts minimally on XΓX_{\Gamma}. We use the minimality of the 𝒬[χ]\mathcal{Q}_{[\chi]}-action, and the fact that 𝒬[χ]\mathcal{Q}_{[\chi]} commute with 𝒫[χ]c\mathcal{P}_{[\chi]}^{c} to produce a continuous 𝕋\mathbb{T}-action preserving WcW^{c} that commutes with α\alpha. This shows that the exponent λμc\lambda_{\mu}^{c} must vanish, a contradiction.

Denote by volf(XΓ)\mathcal{M}_{\rm vol}^{f}(X_{\Gamma}) the ff-invariant measures projecting to volume

(6.1) Φμ=vol,μf(XΓ).\displaystyle\Phi_{*}\mu={\rm vol},\quad\mu\in\mathcal{M}^{f}(X_{\Gamma}).

Equivalently the measures volf(XΓ)\mathcal{M}_{\rm vol}^{f}(X_{\Gamma}) are precisely the measures of maximal entropy for ff [56]. From [56] the set volf(XΓ)\mathcal{M}_{\rm vol}^{f}(X_{\Gamma}) is finite so we may assume, after possibly dropping to a finite index subgroup, that volf(XΓ)\mathcal{M}_{\rm vol}^{f}(X_{\Gamma}) consist of α\alpha-invariant measures. Assume for contradiction that λμc(f)0\lambda_{\mu}^{c}(f)\neq 0 for some μvolf(XΓ)\mu\in\mathcal{M}_{\rm vol}^{f}(X_{\Gamma}). This implies λνc(f)0\lambda_{\nu}^{c}(f)\neq 0 for all νvolf(XΓ)\nu\in\mathcal{M}_{\rm vol}^{f}(X_{\Gamma}) [56].

Lemma 6.1.

For any two μ,νvolf(XΓ)\mu,\nu\in\mathcal{M}_{\rm vol}^{f}(X_{\Gamma}) we have kerλμc=kerλνc\ker\lambda_{\mu}^{c}=\ker\lambda_{\nu}^{c} where λμc,λνc:k\lambda_{\mu}^{c},\lambda_{\nu}^{c}:\mathbb{Z}^{k}\to\mathbb{R}. Moreover, there is a vol{\rm vol}-full measure set Y𝕋dY\subset\mathbb{T}^{d} such that for any yYy\in Y we have x1,,xNΦ1(y)x_{1},...,x_{N}\in\Phi^{-1}(y) with x1<x2<<xNx_{1}<x_{2}<...<x_{N} in the orientation of Φ1(y)\Phi^{-1}(y) such that (xj1,xj+1)(x_{j-1},x_{j+1}) is the stable or unstable manifold for some νvol(XΓ)\nu^{\prime}\in\mathcal{M}_{\rm vol}(X_{\Gamma}) in Φ1(y)\Phi^{-1}(y).

Proof.

We sketch the construction of measures in [56]. Let μvolf(XΓ)\mu\in\mathcal{M}_{\rm vol}^{f}(X_{\Gamma}) and (μxc)xXΓ(\mu_{x}^{c})_{x\in X_{\Gamma}} be the disintegration of μ\mu with respect to the center foliation. The measures μxc\mu_{x}^{c} are atomic μ\mu-almost everywhere since λμc(f)0\lambda_{\mu}^{c}(f)\neq 0 and WcW^{c} has 11-dimensional leaves. Denote by μ\mathcal{F}_{\mu} the ff-invariant foliation (contracting or expanding) manifolds in WcW^{c} defined μ\mu-almost everywhere. Let Y𝕋dY\subset\mathbb{T}^{d} be such that μxc\mu_{x}^{c} exists and is atomic for each xΦ1(y)x\in\Phi^{-1}(y). For yYy\in Y let

(6.2) μxc=(δp1(x)++δpk(x))/k.\displaystyle\mu_{x}^{c}=(\delta_{p_{1}(x)}+...+\delta_{p_{k}(x)})/k.

Let qj(x)q_{j}(x) be the positively oriented endpoint of μ(pj(x))\mathcal{F}_{\mu}(p_{j}(x)). Define a new measure ν\nu by νy=δq1(x)++δqk(x)\nu_{y}=\delta_{q_{1}(x)}+...+\delta_{q_{k}(x)} and ν=νydvol(y)\nu=\nu_{y}\otimes{\rm d}{\rm vol}(y). Then Φν=vol\Phi_{*}\nu={\rm vol} and ν\nu is ergodic, so νvolf(XΓ)\nu\in\mathcal{M}_{\rm vol}^{f}(X_{\Gamma}). For any 𝐧k\mathbf{n}\in\mathbb{Z}^{k} such that λμc(𝐧)<0\lambda_{\mu}^{c}(\mathbf{n})<0, μ(pj(x))\mathcal{F}_{\mu}(p_{j}(x)) is a stable manifold for α𝐧\alpha^{\mathbf{n}}. If λνc(𝐧)<0\lambda_{\nu}^{c}(\mathbf{n})<0, then any point in (pj(x),qj(x))(p_{j}(x),q_{j}(x)) lie in the stable manifold for both pj(x)p_{j}(x) and qj(x)q_{j}(x), which is a contradiction. It follows that λμc(𝐧)<0\lambda_{\mu}^{c}(\mathbf{n})<0 implies λνc(𝐧)0\lambda_{\nu}^{c}(\mathbf{n})\geq 0, or λμc=cλνc\lambda_{\mu}^{c}=-c\lambda_{\nu}^{c} for some c>0c>0 (note that c0c\neq 0 since there are no measure νvolf(XΓ)\nu\in\mathcal{M}_{\rm vol}^{f}(X_{\Gamma}) with zero center exponent [56]). The first part of the lemma follows for measures constructed as above. Since #vol(XΓ)<\#\mathcal{M}_{\rm vol}(X_{\Gamma})<\infty [56] the construction of new invariant measures outlined above can only produce new measures finitely many times. This proves the last part of the lemma, since if the invariant manifolds did not cover the center leaves, then we could proceed the construction. This also proves the first part of the lemma since the measures constructed above have invariant manifolds covering the entire center leaves for vol{\rm vol}-almost every Φ1(y)\Phi^{-1}(y). ∎

6.1. Generic case of Theorem 1.3

Let μvolf(XΓ)\mu\in\mathcal{M}_{\rm vol}^{f}(X_{\Gamma}). If kerλμc\ker\lambda_{\mu}^{c} does not coincide with ker[χ]\ker[\chi] for any coarse exponent of ρ\rho, then η𝒫|E0[χ]\eta_{\mathcal{P}}|_{E_{0}^{[\chi]}} preserve μ\mu for all coarse [χ][\chi] by Theorem 5.1 and Lemma 5.1. So, μ\mu is η𝒫\eta_{\mathcal{P}}-invariant. Accessibility of ff implies that η𝒫\eta_{\mathcal{P}} acts transitively, which implies that the disintegration of μ\mu, μxc\mu_{x}^{c}, is not atomic. This is a contradiction since we assumed λμc(f)0\lambda_{\mu}^{c}(f)\neq 0.

6.2. Exceptional case of Theorem 1.3

Now we deal with the exceptional case of Theorem 1.3. Fix μvolf(XΓ)\mu\in\mathcal{M}_{\rm vol}^{f}(X_{\Gamma}) and assume that λμc=rχ\lambda_{\mu}^{c}=r\chi for some Lyapunov exponent χ\chi of ρ\rho. Denote by [χ][\chi] the corresponding coarse Lyapunov exponent. We will need two preliminary result on circle maps.

Lemma 6.2.

If G,HHomeo+(𝕋)G,H\subset{\rm Homeo}_{+}(\mathbb{T}) are two path-connected groups such that GH.x=HG.x=𝕋GH.x=HG.x=\mathbb{T} for all x𝕋x\in\mathbb{T} then either G.x=𝕋G.x=\mathbb{T} or H.x=𝕋H.x=\mathbb{T}.

Proof.

We have G.x=(a,b)G.x=(a,b), G.x=(a,b]G.x=(a,b], G.x=[a,b)G.x=[a,b) or G.x=[a,b]G.x=[a,b] since GG is path-connected. Write I=G.xI=G.x. If I𝕋I\neq\mathbb{T} then aa and bb are fixed by GG since gg is orientation preserving (and therefore fix the endpoints of the gg-invariant interval II). So, G.a=aG.a=a and G.b=bG.b=b. If H.xIH.x\subset I then 𝕋=GH.x=G.x=I𝕋\mathbb{T}=GH.x=G.x=I\neq\mathbb{T}, so H.xIH.x\not\subset I. Since HH is path-connected J=H.xJ=H.x is an interval that contain xx and JIJ\not\subset I, so aH.xa\in H.x or bH.xb\in H.x. Assume that aH.xa\in H.x. Since aa is fixed by GG we have 𝕋=HG.a=H.a=H.x\mathbb{T}=HG.a=H.a=H.x so H.x=𝕋H.x=\mathbb{T}. ∎

Lemma 6.3.

Let K𝕋K\subset\mathbb{T} be compact such that

  1. (i)

    there is a subgroup 𝒢Homeo+(𝕋)\mathcal{G}\subset{\rm Homeo}_{+}(\mathbb{T}), that acts transitively, with subgroup 𝒢K=StabK(𝒢)={g𝒢K : gK=K}\mathcal{G}_{K}={\rm Stab}_{K}(\mathcal{G})=\{g\in\mathcal{G}_{K}\text{ : }gK=K\},

  2. (ii)

    if xKx\in K and g𝒢g\in\mathcal{G} satisfy gxKgx\in K then g𝒢Kg\in\mathcal{G}_{K},

  3. (iii)

    if g𝒢Kg\in\mathcal{G}_{K} satisfy gx=xgx=x for xKx\in K then g|K=idKg|_{K}={\rm id}_{K},

  4. (iv)

    there is a compact subset 𝒢0𝒢\mathcal{G}_{0}\subset\mathcal{G} such that 𝒢0x=𝕋\mathcal{G}_{0}x=\mathbb{T} for every x𝕋x\in\mathbb{T}.

Then K=𝕋K=\mathbb{T} or KK is finite.

Proof.

The group 𝒢K\mathcal{G}_{K} act transitively and freely on KK by (i)(i), (ii)(ii) and (iii)(iii). Combining this with (iv)(iv) it follows that 𝒢K\mathcal{G}_{K} is a compact group, so 𝒢K\mathcal{G}_{K} preserves a measure ν\nu on 𝕋\mathbb{T}. From Lemma 3.7 it follows that the rotation number ω:𝒢K𝕋\omega:\mathcal{G}_{K}\to\mathbb{T} is a homomorphism. If ω(g)=0\omega(g)=0 for g𝒢Kg\in\mathcal{G}_{K}, then gg fix a point in KK by Lemma 3.6, so by (iii)(iii) we have g|K=idKg|_{K}={\rm id}_{K}. It follows that, if we view 𝒢KHomeo(K)\mathcal{G}_{K}\subset{\rm Homeo}(K), ω:𝒢K𝕋\omega:\mathcal{G}_{K}\to\mathbb{T} is injective. The image T=ω(𝒢K)T=\omega(\mathcal{G}_{K}) is compact, so either T=𝕋T=\mathbb{T} or #T<\#T<\infty. In the second case, since ω|𝒢K\omega|_{\mathcal{G}_{K}} is injective, we have #K=#𝒢K<\#K=\#\mathcal{G}_{K}<\infty. In the first case, ω:𝒢K𝕋\omega:\mathcal{G}_{K}\to\mathbb{T} is injective and surjective, so a homeomorphism. It follows that KK is homeomorphic to a circle, so K𝕋K\hookrightarrow\mathbb{T} is a local homeomorphism by invariance of domain. In particular, KK is both open and closed. Since 𝕋\mathbb{T} is connected it follows that K=𝕋K=\mathbb{T}. ∎

Lemma 6.4.

The action of 𝒫[χ]\mathcal{P}_{[\chi]} commute with the action of 𝒬[χ]\mathcal{Q}_{[\chi]}.

Proof.

By induction it suffices to consider vE0±[χ]v\in E_{0}^{\pm[\chi]} and wE0[η]w\in E_{0}^{[\eta]} for [η][\eta] independent of [χ][\chi]. Assume vE0[χ]v\in E_{0}^{[\chi]}, the other case is identical. Let g=η𝒫vη𝒫wη𝒫vη𝒫wg=\eta_{\mathcal{P}}^{v}\eta_{\mathcal{P}}^{w}\eta_{\mathcal{P}}^{-v}\eta_{\mathcal{P}}^{-w}. Fix

𝒞v,w={𝐧k : lim1logρ𝐧v,1logρ𝐧w<0}\displaystyle\mathcal{C}_{v,w}^{-}=\left\{\mathbf{n}\in\mathbb{Z}^{k}\text{ : }\lim_{\ell\to\infty}\frac{1}{\ell}\log\left\lVert\rho^{\ell\mathbf{n}}v\right\rVert,\frac{1}{\ell}\log\left\lVert\rho^{\ell\mathbf{n}}w\right\rVert<0\right\}

which is a non-empty cone since [χ][\chi] and [η][\eta] are independent. Let Y𝕋dY\subset\mathbb{T}^{d} be the full measure set from Lemma 6.1. Given yYy\in Y let xΦ1(y)x\in\Phi^{-1}(y). We have

α𝐧gx=\displaystyle\alpha^{\ell\mathbf{n}}gx= α𝐧η𝒫vη𝒫wη𝒫vη𝒫wx=\displaystyle\alpha^{\ell\mathbf{n}}\eta_{\mathcal{P}}^{v}\eta_{\mathcal{P}}^{w}\eta_{\mathcal{P}}^{-v}\eta_{\mathcal{P}}^{-w}x=
η𝒫ρ𝐧vη𝒫ρ𝐧wη𝒫ρ𝐧vη𝒫ρ𝐧wα𝐧x.\displaystyle\eta_{\mathcal{P}}^{\rho^{\ell\mathbf{n}}v}\eta_{\mathcal{P}}^{\rho^{\ell\mathbf{n}}w}\eta_{\mathcal{P}}^{-\rho^{\ell\mathbf{n}}v}\eta_{\mathcal{P}}^{-\rho^{\ell\mathbf{n}}w}\alpha^{\ell\mathbf{n}}x.

The action η𝒫\eta_{\mathcal{P}} is Hölder by Lemma 4.2, so from or choice of 𝒞v,w\mathcal{C}_{v,w}^{-} there is uniform κ>0\kappa>0 such that for any 𝐧𝒞v,w\mathbf{n}\in\mathcal{C}_{v,w}^{-}

lim sup1logdc(α𝐧x,α𝐧gx)\displaystyle\limsup_{\ell\to\infty}\frac{1}{\ell}\log{\rm d}_{c}\left(\alpha^{\ell\mathbf{n}}x,\alpha^{\ell\mathbf{n}}gx\right)\leq κmin[η~]=[η],[χ~]=[χ](χ~(𝐧)+η~(𝐧))\displaystyle\kappa\min_{[\tilde{\eta}]=[\eta],[\tilde{\chi}]=[\chi]}\left(\tilde{\chi}(\mathbf{n})+\tilde{\eta}(\mathbf{n})\right)\leq
κmin[η~]=[η]η~(𝐧).\displaystyle\kappa\min_{[\tilde{\eta}]=[\eta]}\tilde{\eta}(\mathbf{n}).

In particular, gxgx lie in the stable manifold of xx for any α𝐧\alpha^{\mathbf{n}} with 𝐧𝒞v,w\mathbf{n}\in\mathcal{C}_{v,w}^{-}. Assume for contradiction that gxxgx\neq x. Since gx,xWc(x)gx,x\in W^{c}(x) lie in the same stable manifold Lemma 6.1 implies

lim sup1logdc(α𝐧x,α𝐧gx)=λνc(𝐧)\displaystyle\limsup_{\ell\to\infty}\frac{1}{\ell}\log{\rm d}_{c}\left(\alpha^{\ell\mathbf{n}}x,\alpha^{\ell\mathbf{n}}gx\right)=\lambda_{\nu}^{c}(\mathbf{n})

for some νvolf(XΓ)\nu\in\mathcal{M}_{\rm vol}^{f}(X_{\Gamma}). So, λνc(𝐧)κmin[η~]=[η]η~(𝐧)\lambda_{\nu}^{c}(\mathbf{n})\leq\kappa\min_{[\tilde{\eta}]=[\eta]}\tilde{\eta}(\mathbf{n}) for any 𝐧𝒞v,w\mathbf{n}\in\mathcal{C}_{v,w}^{-}. Since [χ][\chi] is independent of [η][\eta] we find 𝐧j𝒞v,w\mathbf{n}_{j}\in\mathcal{C}_{v,w}^{-} such that χ~(𝐧j)0\tilde{\chi}(\mathbf{n}_{j})\to 0 for any [χ~]=[χ][\tilde{\chi}]=[\chi] and η~(𝐧j)\tilde{\eta}(\mathbf{n}_{j})\to-\infty for any [η~]=[η][\tilde{\eta}]=[\eta]. Since λνc\lambda_{\nu}^{c} has the same kernel as [χ][\chi] we also have λνc(𝐧j)0\lambda_{\nu}^{c}(\mathbf{n}_{j})\to 0. This is a contradiction since

0=limjλνc(𝐧j)limjκmin[χ~]=[χ]η~(𝐧j)=.\displaystyle 0=\lim_{j\to\infty}\lambda_{\nu}^{c}(\mathbf{n}_{j})\leq\lim_{j\to\infty}\kappa\min_{[\tilde{\chi}]=[\chi]}\tilde{\eta}(\mathbf{n}_{j})=-\infty.

We conclude that gx=xgx=x for xΦ1(Y)x\in\Phi^{-1}(Y) and since YY has full volume we conclude that gx=xgx=x holds on a dense set. The map gg is continuous so this implies g=idXΓg={\rm id}_{X_{\Gamma}}. ∎

By Lemmas 6.2 and 6.4 either 𝒫[χ]c\mathcal{P}_{[\chi]}^{c} or 𝒬[χ]c\mathcal{Q}_{[\chi]}^{c} acts transitively on Wc(x)W^{c}(x) for every xXΓx\in X_{\Gamma}. We will prove that neither 𝒫[χ]c\mathcal{P}_{[\chi]}^{c} or 𝒬[χ]c\mathcal{Q}_{[\chi]}^{c} can act transitively on Wc(x)W^{c}(x), which is a contradiction.

Lemma 6.5.

The group 𝒬[χ]c\mathcal{Q}_{[\chi]}^{c} can not act transitively on any Wc(x)W^{c}(x).

Proof.

Let YXΓY\subset X_{\Gamma} be the subset where 𝒬[χ]c\mathcal{Q}_{[\chi]}^{c} act transitively on Wc(y)W^{c}(y). Since 𝒬[χ]c\mathcal{Q}_{[\chi]}^{c} is normal in 𝒬[χ]\mathcal{Q}_{[\chi]} (it is the kernel of a homomorphism) YY is 𝒬[χ]\mathcal{Q}_{[\chi]}-invariant. But by Lemma 6.4 the set YY is also 𝒫[χ]\mathcal{P}_{[\chi]}-invariant. So YY is 𝒫\mathcal{P}-invariant (by Lemma 6.4). Since ff is accessible it follows that either Y=Y=\emptyset or Y=XΓY=X_{\Gamma}.

Assume for contradiction that YY\neq\emptyset, so Y=XΓY=X_{\Gamma}. The remainder of the proof is similar to the proof of the generic case of Theorem 1.3. By Theorem 5.1 and Lemma 5.1 the action 𝒬[χ]c\mathcal{Q}_{[\chi]}^{c} preserve μ\mu. Since 𝒬[χ]c\mathcal{Q}_{[\chi]}^{c} act transitively on every Wc(x)W^{c}(x) the disintegration of μ\mu, μxc\mu_{x}^{c}, is not atomic. This is a contradiction, so we conclude that Y=Y=\emptyset. ∎

Before proceeding we define the space

(6.3) V=Π𝒬[χ]=η±[χ]E0[η]\displaystyle V=\Pi\mathcal{Q}_{[\chi]}=\bigoplus_{\eta\neq\pm[\chi]}E_{0}^{[\eta]}

and the associated translation action on 𝕋d\mathbb{T}^{d} by Rv(x)=x+vR_{v}(x)=x+v. The action RvR_{v} is minimal (Lemma 2.2).

Lemma 6.6.

The group 𝒫[χ]c\mathcal{P}_{[\chi]}^{c} can not act transitively on any Wc(x0)W^{c}(x_{0}).

The proof of Lemma 6.6 is by contradiction. As in the proof of Lemma 6.5 we may assume that 𝒫[χ]c\mathcal{P}_{[\chi]}^{c} act transitively on every Wc(x)W^{c}(x), we will do this in the remainder. We split the proof of Lemma 6.6 into parts, beginning by proving that 𝒬[χ]\mathcal{Q}_{[\chi]} acts minimally.

For xXΓx\in X_{\Gamma} define

(6.4) Kx:={η𝒫wx : w𝒬[χ]}¯.\displaystyle K_{x}:=\overline{\{\eta_{\mathcal{P}}^{w}x\text{ : }w\in\mathcal{Q}_{[\chi]}\}}.

First, we show that KxK_{x} is a minimal set for the 𝒬[χ]\mathcal{Q}_{[\chi]}-action for every xXΓx\in X_{\Gamma}.

Lemma 6.7.

For any yKxy\in K_{x} we have

(6.5) Kx={η𝒫wy : w𝒬[χ]}¯,\displaystyle K_{x}=\overline{\{\eta_{\mathcal{P}}^{w}y\text{ : }w\in\mathcal{Q}_{[\chi]}\}},

that is KxK_{x} is a minimal set for the 𝒬[χ]\mathcal{Q}_{[\chi]}-action. In particular, the set {Kx}x\{K_{x}\}_{x} form a partition of XΓX_{\Gamma}.

Proof.

Define YXΓY\subset X_{\Gamma} as those xXΓx\in X_{\Gamma} such that KxK_{x} is minimal for the 𝒬[χ]\mathcal{Q}_{[\chi]}-action. A standard application of Zorn’s lemma shows that YY is non-empty. Given any w𝒬[χ]w\in\mathcal{Q}_{[\chi]} we have η𝒫wKx=Kx\eta_{\mathcal{P}}^{w}K_{x}=K_{x}, so YY is 𝒬[χ]\mathcal{Q}_{[\chi]}-invariant. By Lemma 6.4 the action of 𝒫[χ]\mathcal{P}_{[\chi]} commute with the action of 𝒬[χ]\mathcal{Q}_{[\chi]} so, given w𝒫[χ]w\in\mathcal{P}_{[\chi]}, the map η𝒫w:Kxη𝒫wKx\eta_{\mathcal{P}}^{w}:K_{x}\to\eta_{\mathcal{P}}^{w}K_{x} conjugates the 𝒬[χ]\mathcal{Q}_{[\chi]}-action on KxK_{x} to the 𝒬[χ]\mathcal{Q}_{[\chi]}-action on η𝒫wKx=Kη𝒫wx\eta_{\mathcal{P}}^{w}K_{x}=K_{\eta_{\mathcal{P}}^{w}x}. It follows that xYx\in Y if and only if η𝒫wxY\eta_{\mathcal{P}}^{w}x\in Y. Since YY is 𝒫[χ]\mathcal{P}_{[\chi]}-invariant and 𝒬[χ]\mathcal{Q}_{[\chi]}-invariant Lemma 6.4 implies that YY is susu-saturated. Accessibility and the fact that YY\neq\emptyset implies that Y=XΓY=X_{\Gamma}. ∎

Lemma 6.8.

The action of 𝒬[χ]\mathcal{Q}_{[\chi]} is minimal.

Proof.

Since KxK_{x} is 𝒬[χ]\mathcal{Q}_{[\chi]}-invariant, Lemma 4.3 implies

(6.6) Φ(Kx)=Φ(η𝒫wKx)=Φ(Kx)+Π(w),w𝒬[χ]\displaystyle\Phi(K_{x})=\Phi(\eta_{\mathcal{P}}^{w}K_{x})=\Phi(K_{x})+\Pi(w),\quad w\in\mathcal{Q}_{[\chi]}

so Φ(Kx)\Phi(K_{x}) is compact and invariant by RvR_{v} from Equation 6.3. The action RvR_{v} is minimal so Φ(Kx)=𝕋d\Phi(K_{x})=\mathbb{T}^{d}. It follows that KxWc(y)K_{x}\cap W^{c}(y)\neq\emptyset for every yXΓy\in X_{\Gamma}.

If 𝒢=η𝒫𝒫[χ]c\mathcal{G}=\eta_{\mathcal{P}}^{\mathcal{P}_{[\chi]}^{c}} and K=KxWc(y)K=K_{x}\cap W^{c}(y) then the assumptions in Lemma 6.3 are satisfied (after identifying Wc(y)𝕋W^{c}(y)\cong\mathbb{T}). Indeed, 𝒢\mathcal{G} act transitively by Lemma 4.6. Property (ii)(ii) and (iii)(iii) follows from Lemmas 6.4 and 6.7. To show property (iv)(iv) we follow [2, Section 8.3]. For NN\in\mathbb{N}, write

(6.7) 𝒫[χ],Nc={w=v1v2vN𝒫[χ]c : vjE0±[χ], vjN, nN}.\displaystyle\mathcal{P}_{[\chi],N}^{c}=\{w=v_{1}v_{2}...v_{N}\in\mathcal{P}_{[\chi]}^{c}\text{ : }v_{j}\in E_{0}^{\pm[\chi]},\text{ }\left\lVert v_{j}\right\rVert\leq N,\text{ }n\leq N\}.

That is 𝒫[χ],Nc\mathcal{P}_{[\chi],N}^{c} consist of those words in 𝒫[χ]c\mathcal{P}_{[\chi]}^{c} with at most NN letters, and each letter has length of at most NN. The following is immediate from the definition

(6.8) 𝒫[χ],N+1c𝒫[χ],Nc,𝒫[χ]c=N1𝒫[χ],Nc.\displaystyle\mathcal{P}_{[\chi],N+1}^{c}\supset\mathcal{P}_{[\chi],N}^{c},\quad\mathcal{P}_{[\chi]}^{c}=\bigcup_{N\geq 1}\mathcal{P}_{[\chi],N}^{c}.

Let 𝒦N={η𝒫w : w𝒫[χ],Nc}\mathcal{K}_{N}=\{\eta_{\mathcal{P}}^{w}\text{ : }w\in\mathcal{P}_{[\chi],N}^{c}\} and KN=𝒦Nx0={k(x0) : k𝒦N}K_{N}=\mathcal{K}_{N}x_{0}=\{k(x_{0})\text{ : }k\in\mathcal{K}_{N}\} for fixed x0Wc(y)x_{0}\in W^{c}(y). Then KNKN+1K_{N}\subset K_{N+1} is an ascending sequence of compact sets. If KN0=Wc(y)K_{N_{0}}=W^{c}(y) and xWc(x)x\in W^{c}(x) then x0𝒦N0xx_{0}\in\mathcal{K}_{N_{0}}x so 𝒦2N0x𝒦N0x0=Wc(y)\mathcal{K}_{2N_{0}}x\supset\mathcal{K}_{N_{0}}x_{0}=W^{c}(y). That is, property (iv)(iv) of Lemma 6.3 holds if KN0=Wc(y)K_{N_{0}}=W^{c}(y) for some N0N_{0}. Since

(6.9) N1KN={η𝒫wx0 : w𝒫[χ]c}=Wc(y),\displaystyle\bigcup_{N\geq 1}K_{N}=\{\eta_{\mathcal{P}}^{w}x_{0}\text{ : }w\in\mathcal{P}_{[\chi]}^{c}\}=W^{c}(y),

Baire’s category theorem implies that there is N1N_{1} such that int(KN){\rm int}(K_{N})\neq\emptyset for NN1N\geq N_{1} (where int(KN){\rm int}(K_{N}) is the interior in Wc(y)W^{c}(y)). Define UN=int(KN)U_{N}={\rm int}(K_{N}) and CN=UNcC_{N}=U_{N}^{c} (where the complement is in Wc(y)W^{c}(y)). If NN1N\geq N_{1} then x0𝒦NUNU2Nx_{0}\in\mathcal{K}_{N}U_{N}\subset U_{2N}, so KN𝒦NU2NU3NK_{N}\subset\mathcal{K}_{N}U_{2N}\subset U_{3N}. From Equation 6.9 we obtain

(6.10) Wc(y)=N1KN=NN1KNNN1U3N=N1UN\displaystyle W^{c}(y)=\bigcup_{N\geq 1}K_{N}=\bigcup_{N\geq N_{1}}K_{N}\subset\bigcup_{N\geq N_{1}}U_{3N}=\bigcup_{N\geq 1}U_{N}

or

(6.11) =N1CN.\displaystyle\emptyset=\bigcap_{N\geq 1}C_{N}.

Since CN+1CNC_{N+1}\subset C_{N} is a descending sequence of compact sets there is N0N_{0} such that CN0=C_{N_{0}}=\emptyset. Equivalently UN0=KN0=Wc(y)U_{N_{0}}=K_{N_{0}}=W^{c}(y) proving that property (iv)(iv) of Lemma 6.3 holds. Lemma 6.3 implies that #KxWc(y)<\#K_{x}\cap W^{c}(y)<\infty or KxWc(y)=Wc(y)K_{x}\cap W^{c}(y)=W^{c}(y) for every x,yXΓx,y\in X_{\Gamma}.

We claim that #KxWc(y)\#K_{x}\cap W^{c}(y) is independent of yy. Indeed, as in the proof of Lemma 6.7, for w𝒫w\in\mathcal{P} we have η𝒫wKx=Kη𝒫wx\eta_{\mathcal{P}}^{w}K_{x}=K_{\eta_{\mathcal{P}}^{w}x}. So given zKxWc(y)z\in K_{x}\cap W^{c}(y) we find, by accessibility, some w𝒫w\in\mathcal{P} that satisfy η𝒫wx=z\eta_{\mathcal{P}}^{w}x=z and in extension η𝒫wKx=Kz=Kx\eta_{\mathcal{P}}^{w}K_{x}=K_{z}=K_{x} (since zKxz\in K_{x}, and {Kx}xXΓ\{K_{x}\}_{x\in X_{\Gamma}} is a partition, Lemma 6.7). We also have η𝒫wWc(x)=Wc(y)\eta_{\mathcal{P}}^{w}W^{c}(x)=W^{c}(y) (since zWc(y)z\in W^{c}(y)) so η𝒫w(KxWc(x))=KxWc(y)\eta_{\mathcal{P}}^{w}(K_{x}\cap W^{c}(x))=K_{x}\cap W^{c}(y). It follows that

(6.12) #KxWc(y)=#KxWc(x)\displaystyle\#K_{x}\cap W^{c}(y)=\#K_{x}\cap W^{c}(x)

for all yXΓy\in X_{\Gamma}. If #KxWc(x)=q<\#K_{x}\cap W^{c}(x)=q<\infty, then the fibers of Φ\Phi intersect KxK_{x} precisely qq times. Given z0𝕋dz_{0}\in\mathbb{T}^{d} we define

ε0=infyy y,yΦ1(z0)Kxd(y,y)>0.\displaystyle\varepsilon_{0}=\inf_{y\neq y^{\prime}\text{ }y,y^{\prime}\in\Phi^{-1}(z_{0})\cap K_{x}}{\rm d}(y,y^{\prime})>0.

Let {y1,,yq}=KxΦ1(z0)\{y_{1},...,y_{q}\}=K_{x}\cap\Phi^{-1}(z_{0}) and δ>0\delta>0 small. For zBδ(z0)z\in B_{\delta}(z_{0}), we define yj(z)y_{j}(z) as the element in Φ1(z)Kx\Phi^{-1}(z)\cap K_{x} that minimize d(yj(z),yj(z0)){\rm d}(y_{j}(z),y_{j}(z_{0})) (if this does not define a unique point, choose an arbitrary minimizer). If znz0z_{n}\to z_{0} then any convergent subsequence of yj(zn)y_{j}(z_{n}) converges to an element of KxΦ1(z0)K_{x}\cap\Phi^{-1}(z_{0}). From our choice of yj(z)y_{j}(z) it is clear that any convergent subsequence converge to yj(z0)y_{j}(z_{0}). So, yj(zn)yj(z0)y_{j}(z_{n})\to y_{j}(z_{0}) and yjy_{j} is continuous at z0z_{0}. Choose δ>0\delta>0 small enough such that d(yj(z),yj(z0))<ε0/100{\rm d}(y_{j}(z),y_{j}(z_{0}))<\varepsilon_{0}/100 for all zBδ(z0)z\in B_{\delta}(z_{0}) and j=1,,qj=1,...,q. For z,zBδ(z0)z,z^{\prime}\in B_{\delta}(z_{0}) and iji\neq j the reverse triangle inequality implies

d(yi(z),yj(z))\displaystyle{\rm d}(y_{i}(z),y_{j}(z^{\prime}))\geq d(yi(z0),yj(z))d(yi(z0),yi(z))\displaystyle{\rm d}(y_{i}(z_{0}),y_{j}(z^{\prime}))-{\rm d}(y_{i}(z_{0}),y_{i}(z))\geq
d(yi(z0),yj(z0))d(yj(z0),yj(z))d(yi(z0),yi(z))>\displaystyle{\rm d}(y_{i}(z_{0}),y_{j}(z_{0}))-{\rm d}(y_{j}(z_{0}),y_{j}(z^{\prime}))-{\rm d}(y_{i}(z_{0}),y_{i}(z))>
ε0ε050>49ε050>0.\displaystyle\varepsilon_{0}-\frac{\varepsilon_{0}}{50}>\frac{49\varepsilon_{0}}{50}>0.

where the last inequality use the definition of ε0\varepsilon_{0}. In particular, yi(z)yj(z)y_{i}(z)\neq y_{j}(z) for every zBδ(z0)z\in B_{\delta}(z_{0}), so KxΦ1(z)={y1(z),,yq(z)}K_{x}\cap\Phi^{-1}(z)=\{y_{1}(z),...,y_{q}(z)\}. Moreover, we have

(6.13) d(yi(z),yi(z))d(yi(z),yi(z0))+d(yi(z0),yi(z))<ε050.\displaystyle{\rm d}(y_{i}(z),y_{i}(z^{\prime}))\leq{\rm d}(y_{i}(z),y_{i}(z_{0}))+{\rm d}(y_{i}(z_{0}),y_{i}(z^{\prime}))<\frac{\varepsilon_{0}}{50}.

Since ε0/50<49ε0/50\varepsilon_{0}/50<49\varepsilon_{0}/50, yi(z)y_{i}(z^{\prime}) is the element yKxΦ1(z)y\in K_{x}\cap\Phi^{-1}(z^{\prime}) that minimize d(yi(z),y){\rm d}(y_{i}(z),y). The same argument that showed continuity at z0z_{0}, now show that yjy_{j} is continuous at any zBδ(z0)z\in B_{\delta}(z_{0}). So the functions y1,,yq:Bδ(z0)Kxy_{1},...,y_{q}:B_{\delta}(z_{0})\to K_{x} are continuous. Note that Φ\Phi restricted to KxΦ1(Bδ(z0))Bε0(yj(z0))K_{x}\cap\Phi^{-1}(B_{\delta}(z_{0}))\cap B_{\varepsilon_{0}}(y_{j}(z_{0})) has inverse given by yjy_{j}. So, Φ:Kx𝕋d\Phi:K_{x}\to\mathbb{T}^{d} is a finite covering map. In particular, KxK_{x} is homeomorphic to 𝕋d\mathbb{T}^{d}. It follows that the map

(6.14) (Φ|Kx):π1𝕋ddΓπ1XΓ\displaystyle(\Phi|_{K_{x}})_{*}:\pi_{1}\mathbb{T}^{d}\cong\mathbb{Z}^{d}\to\Gamma\cong\pi_{1}X_{\Gamma}

is injective with image of finite index. The map

(6.15) dπ1KxiKxπ1XΓ=Γ\displaystyle\mathbb{Z}^{d}\cong\pi_{1}K_{x}\xrightarrow{i_{K_{x}}}\pi_{1}X_{\Gamma}=\Gamma

is injective since its injective after composition with Φ=π\Phi_{*}=\pi_{*}. The group Γ=Im(iKx)×kerΦ=Im(iKx)×[Γ,Γ]\Gamma^{\prime}={\rm Im}(i_{K_{x}})_{*}\times\ker\Phi_{*}={\rm Im}(i_{K_{x}})_{*}\times[\Gamma,\Gamma] has finite index in Γ\Gamma and is abelian, but Γ\Gamma is not virtually abelian so this is a contradiction. We conclude that KxWc(y)=Wc(y)K_{x}\cap W^{c}(y)=W^{c}(y) for every yy which implies Kx=XΓK_{x}=X_{\Gamma} for every xXΓx\in X_{\Gamma}. ∎

Lemma 6.9.

The map r:𝒫[χ]c𝕋r:\mathcal{P}_{[\chi]}^{c}\to\mathbb{T}, r(w)=ω(η𝒫w|Wc(x))r(w)=\omega(\eta_{\mathcal{P}}^{w}|_{W^{c}(x)}) is a well-defined surjective homomorphism. Moreover, r(w)=0r(w)=0 if and only if η𝒫w=idXΓ\eta_{\mathcal{P}}^{w}={\rm id}_{X_{\Gamma}}.

Proof.

Fix w𝒫[χ]cw\in\mathcal{P}_{[\chi]}^{c}, g=η𝒫wg=\eta_{\mathcal{P}}^{w}, and define φ:𝕋d𝕋\varphi:\mathbb{T}^{d}\to\mathbb{T} by

(6.16) φ(y):=ω(g|Φ1(y)).\displaystyle\varphi(y):=\omega(g|_{\Phi^{-1}(y)}).

The function φ\varphi is continuous since rotation numbers vary continuously in the C0C^{0}-topology [41, Proposition 11.1.6]. Given w𝒬[χ]w\in\mathcal{Q}_{[\chi]} we have gη𝒫w=η𝒫wgg\eta_{\mathcal{P}}^{w}=\eta_{\mathcal{P}}^{w}g (by Lemma 6.4) and Φη𝒫w=RΠ(w)Φ\Phi\eta_{\mathcal{P}}^{w}=R_{\Pi(w)}\Phi (by Lemma 4.3). Let VV be the space defined in Equation 6.3. Given vVv\in V let w𝒬[χ]w\in\mathcal{Q}_{[\chi]} be such that Π(w)=v\Pi(w)=v. We have

φ(Rv(x))=\displaystyle\varphi(R_{v}(x))= ω(g|Φ1(Rv(x)))=ω(g|η𝒫wΦ1(x))=\displaystyle\omega(g|_{\Phi^{-1}(R_{v}(x))})=\omega(g|_{\eta_{\mathcal{P}}^{w}\Phi^{-1}(x)})=
ω(η𝒫wgη𝒫w1|η𝒫wΦ1(x))=ω(g|Φ1(x))=φ(x)\displaystyle\omega(\eta_{\mathcal{P}}^{w}g\eta_{\mathcal{P}}^{w^{-1}}|_{\eta_{\mathcal{P}}^{w}\Phi^{-1}(x)})=\omega(g|_{\Phi^{-1}(x)})=\varphi(x)

where the second to last equality uses that the rotation number is conjugacy invariant [41, Proposition 11.1.3]. Since φ\varphi is invariant under RvR_{v} and RvR_{v} is minimal φ\varphi is constant. That is, g:Wc(x)Wc(x)g:W^{c}(x)\to W^{c}(x) has a rotation number independent of xXΓx\in X_{\Gamma}. This shows that r:𝒫[χ]c𝕋r:\mathcal{P}_{[\chi]}^{c}\to\mathbb{T} is well-defined. If r(w)=0r(w)=0 then η𝒫w\eta_{\mathcal{P}}^{w} fix some xx, so Lemmas 6.4 and 6.8 implies that η𝒫w=idXΓ\eta_{\mathcal{P}}^{w}={\rm id}_{X_{\Gamma}}.

Since H=η𝒫|𝒫[χ]c𝒫[χ]c/{w𝒫[χ]c : η𝒫w=idXΓ}H=\eta_{\mathcal{P}}|_{\mathcal{P}_{[\chi]}^{c}}\cong\mathcal{P}_{[\chi]}^{c}/\{w\in\mathcal{P}_{[\chi]}^{c}\text{ : }\eta_{\mathcal{P}}^{w}={\rm id}_{X_{\Gamma}}\} act transitively and freely on each Wc(x)W^{c}(x), it follows that HH is homeomorphic to Wc(x)𝕋W^{c}(x)\cong\mathbb{T}. So HH is a compact topological group. Since HH is compact the HH-action preserve a measure. That rr is a homomorphism follows from Lemma 3.7. The image r(H)𝕋r(H)\subset\mathbb{T} is a compact subgroup, so it is either finite or all of 𝕋\mathbb{T}. But r(H)r(H) can not be finite since rr is injective and HH acts transitively. ∎

We can now prove Lemma 6.6.

Proof of Lemma 6.6.

Suppose that 𝒫[χ]c\mathcal{P}_{[\chi]}^{c} act transitively on some Wc(x0)W^{c}(x_{0}). By Lemma 6.9 we can define an action ηc:𝕋×XΓXΓ\eta_{c}:\mathbb{T}\times X_{\Gamma}\to X_{\Gamma} by ηcsx=η𝒫wx\eta_{c}^{s}x=\eta_{\mathcal{P}}^{w}x with w𝒫cw\in\mathcal{P}^{c} chosen such that r(w)=sr(w)=s. The action ηc\eta_{c} is free and transitive on each center leaf Wc(x)W^{c}(x). Moreover, for 𝐧k\mathbf{n}\in\mathbb{Z}^{k} and w𝒫[χ]cw\in\mathcal{P}_{[\chi]}^{c} we have

(6.17) η𝒫ρ𝐧w=α𝐧η𝒫wα𝐧\displaystyle\eta_{\mathcal{P}}^{\rho^{\mathbf{n}}w}=\alpha^{\mathbf{n}}\eta_{\mathcal{P}}^{w}\alpha^{-\mathbf{n}}

so by conjugacy invariance of the rotation number r(ρ𝐧w)=r(w)=sr(\rho^{\mathbf{n}}w)=r(w)=s. It follows that

(6.18) ηcs=η𝒫ρ𝐧w=α𝐧η𝒫wα𝐧=α𝐧ηcsα𝐧\displaystyle\eta_{c}^{s}=\eta_{\mathcal{P}}^{\rho^{\mathbf{n}}w}=\alpha^{\mathbf{n}}\eta_{\mathcal{P}}^{w}\alpha^{-\mathbf{n}}=\alpha^{\mathbf{n}}\eta_{c}^{s}\alpha^{-\mathbf{n}}

which proves that α\alpha commute with ηc\eta_{c}. This implies that α\alpha have vanishing Lyapunov exponent along WcW^{c}, which is a contradiction. We conclude that 𝒫[χ]c\mathcal{P}_{[\chi]}^{c} can not act transitively on Wc(x0)W^{c}(x_{0}). ∎

We finish the proof of Theorem 1.3 in the exceptional case.

Proof of Theorem 1.3 in exceptional case.

By Lemma 6.4 the action of 𝒫[χ]c\mathcal{P}_{[\chi]}^{c} commute with the action of 𝒬[χ]c\mathcal{Q}_{[\chi]}^{c}. Since ff is accessible η𝒫𝒫[χ]cη𝒫𝒬[χ]cx=Wc(x)\eta_{\mathcal{P}}^{\mathcal{P}_{[\chi]}^{c}}\eta_{\mathcal{P}}^{\mathcal{Q}_{[\chi]}^{c}}x=W^{c}(x) for all xXΓx\in X_{\Gamma}. By Lemma 6.2 either 𝒫[χ]c\mathcal{P}_{[\chi]}^{c} or 𝒬[χ]c\mathcal{Q}_{[\chi]}^{c} act transitively on Wc(x)W^{c}(x). Lemma 6.6 shows that 𝒫[χ]c\mathcal{P}_{[\chi]}^{c} does not act transitively on Wc(x)W^{c}(x) so 𝒬[χ]c\mathcal{Q}_{[\chi]}^{c} must act transitively on Wc(x)W^{c}(x). This is a contradiction by Lemma 6.5. ∎

7. Compatible algebraic structure

In this section we produce a topological conjugacy between α\alpha and an affine action α0\alpha_{0} (Theorem 7.1). Let α:k×XΓXΓ\alpha:\mathbb{Z}^{k}\times X_{\Gamma}\to X_{\Gamma} and f:XΓXΓf:X_{\Gamma}\to X_{\Gamma} satisfy the assumptions of Theorem A. By Theorem 1.3 there is a unique α\alpha-invariant measure μ\mu, with center exponent λμc=0\lambda_{\mu}^{c}=0, that projects to volume.

We begin by using the measure μ\mu to construct a circle action ηc:𝕋×XΓXΓ\eta_{c}:\mathbb{T}\times X_{\Gamma}\to X_{\Gamma} that commutes with α\alpha and η𝒫\eta_{\mathcal{P}}, preserves center leaves and acts transitively and freely on each center leaf. Moreover, if w𝒫cw\in\mathcal{P}^{c} then η𝒫w=ηct\eta_{\mathcal{P}}^{w}=\eta_{c}^{t} for some t𝕋t\in\mathbb{T}. This shows that the action of 𝒫\mathcal{P} on XΓX_{\Gamma} factor through a nilpotent group with base E0uE0sE_{0}^{u}\oplus E_{0}^{s} and center isomorphic to 𝕋\mathbb{T}. Lifting this action to GG we obtain a transitive free action of a 22-step nilpotent group, NN, on GG.

Using that λμc=0\lambda_{\mu}^{c}=0, the following two lemmas are proved identically as Lemmas 6.4 and 6.9.

Lemma 7.1.

Let [χ][\chi] and [η][\eta] be independent. If w1𝒫[χ]w_{1}\in\mathcal{P}_{[\chi]} and w2𝒫[η]w_{2}\in\mathcal{P}_{[\eta]} then

(7.1) η𝒫w1η𝒫w2=η𝒫w2η𝒫w1.\displaystyle\eta_{\mathcal{P}}^{w_{1}}\eta_{\mathcal{P}}^{w_{2}}=\eta_{\mathcal{P}}^{w_{2}}\eta_{\mathcal{P}}^{w_{1}}.
Lemma 7.2.

For any w𝒫cw\in\mathcal{P}^{c} the rotation number of

(7.2) η𝒫w:Wc(x)Wc(x)\displaystyle\eta_{\mathcal{P}}^{w}:W^{c}(x)\to W^{c}(x)

is independent of xx. The map r(w)r(w) mapping w𝒫cw\in\mathcal{P}^{c} to the rotation number of η𝒫w\eta_{\mathcal{P}}^{w} is a homomorphism with kernel {w𝒫c : η𝒫w=idXΓ}\{w\in\mathcal{P}^{c}\text{ : }\eta_{\mathcal{P}}^{w}={\rm id}_{X_{\Gamma}}\}.

Using Lemmas 7.1 and 7.1 we produce an NN-action, with NN nilpotent, on GG.

Lemma 7.3.

The action of 𝒫\mathcal{P} on GG factor through a nilpotent Lie group NN that acts transitively and freely on GG, the action of NN descends to an action on XΓX_{\Gamma}.

Proof.

Denote by NN the image of 𝒫\mathcal{P} in Homeo(G){\rm Homeo}(G) and N~\tilde{N} the image of 𝒫\mathcal{P} in Homeo(XΓ){\rm Homeo}(X_{\Gamma}). We begin by showing that the 𝒫c\mathcal{P}^{c}-action on XΓX_{\Gamma} factor through an abelian group. As in the proof of Lemma 6.6 we define ηcs:XΓXΓ\eta_{c}^{s}:X_{\Gamma}\to X_{\Gamma} by ηcs=η𝒫w\eta_{c}^{s}=\eta_{\mathcal{P}}^{w} for any w𝒫cw\in\mathcal{P}^{c} that satisfies ω(η𝒫w|Wc(x))=s\omega(\eta_{\mathcal{P}}^{w}|_{W^{c}(x)})=s. By Lemma 7.2 ηc\eta_{c} is well-defined, continuous, and acts transitively and freely on each center leaf Wc(x)W^{c}(x).

As in the proof of Lemma 6.6, α\alpha commutes with ηc\eta_{c}. Let rr be from Lemma 7.2. Given w𝒫w\in\mathcal{P} and wc𝒫cw_{c}\in\mathcal{P}^{c} we have r(wwcw1)=r(wc)r(ww_{c}w^{-1})=r(w_{c}) since η𝒫wη𝒫wc(η𝒫w)1\eta_{\mathcal{P}}^{w}\eta_{\mathcal{P}}^{w_{c}}(\eta_{\mathcal{P}}^{w})^{-1} has the same rotation number as η𝒫wc\eta_{\mathcal{P}}^{w_{c}}. It follows that

(7.3) ηcr(wc)=η𝒫wη𝒫wc(η𝒫w)1=η𝒫wηcr(wc)(η𝒫w)1\displaystyle\eta_{c}^{r(w_{c})}=\eta_{\mathcal{P}}^{w}\eta_{\mathcal{P}}^{w_{c}}(\eta_{\mathcal{P}}^{w})^{-1}=\eta_{\mathcal{P}}^{w}\eta_{c}^{r(w_{c})}(\eta_{\mathcal{P}}^{w})^{-1}

which shows that ηc\eta_{c} commute with η𝒫\eta_{\mathcal{P}}. Let r~:𝒫c\tilde{r}:\mathcal{P}^{c}\to\mathbb{R} be defined by r~(w)=ω~(η𝒫w)\tilde{r}(w)=\tilde{\omega}(\eta_{\mathcal{P}}^{w}) where ω~(η𝒫w)\tilde{\omega}(\eta_{\mathcal{P}}^{w})\in\mathbb{R} is the rotation number of η𝒫w\eta_{\mathcal{P}}^{w} with respect to the natural lift to GG from Definition 4.3. Lift ηc\eta_{c} to GG, then η𝒫w=ηcr~(w)\eta_{\mathcal{P}}^{w}=\eta_{c}^{\tilde{r}(w)}. It is immediate from the analogous properties of ηc:𝕋×XΓXΓ\eta_{c}:\mathbb{T}\times X_{\Gamma}\to X_{\Gamma} that ηc:×GG\eta_{c}:\mathbb{R}\times G\to G commute with α\alpha and η𝒫\eta_{\mathcal{P}}.

Let NcNN^{c}\subset N be the image of 𝒫c\mathcal{P}^{c}. Since ηc\eta_{c} commute with η𝒫\eta_{\mathcal{P}} and since η𝒫w=η𝒫r~(w)\eta_{\mathcal{P}}^{w}=\eta_{\mathcal{P}}^{\tilde{r}(w)} for w𝒫cw\in\mathcal{P}^{c} the subgroup NcN^{c} is central in NN. Moreover, Lemma 4.1 implies that

(7.4) 𝒫c=[𝒫,𝒫]\displaystyle\mathcal{P}^{c}=\left[\mathcal{P},\mathcal{P}\right]

so Nc=[N,N]N^{c}=[N,N]. It follows that NN is 22-step nilpotent. By Lemma 4.1 we can write any w𝒫w\in\mathcal{P} as w=wcvsvuw=w_{c}v_{s}v_{u} with vσE0σv_{\sigma}\in E_{0}^{\sigma}, σ=s,u\sigma=s,u, and wc𝒫cw_{c}\in\mathcal{P}^{c}. It follows that η𝒫w=ηcr~(wc)ηsvsηuvu\eta_{\mathcal{P}}^{w}=\eta_{c}^{\tilde{r}(w_{c})}\eta_{s}^{v_{s}}\eta_{u}^{v_{u}}, so T:E0s×E0u×NT:E_{0}^{s}\times E_{0}^{u}\times\mathbb{R}\to N, T(vs,vu,t)=ηctηsvsηuvuT(v_{s},v_{u},t)=\eta_{c}^{t}\eta_{s}^{v_{s}}\eta_{u}^{v_{u}} is surjective. If ηctηsvsηuvu=ηctηsvsηuvu\eta_{c}^{t}\eta_{s}^{v_{s}}\eta_{u}^{v_{u}}=\eta_{c}^{t^{\prime}}\eta_{s}^{v_{s}^{\prime}}\eta_{u}^{v_{u}^{\prime}} then we apply Φ\Phi and obtain vs=vsv_{s}=v_{s}^{\prime}, vu=vuv_{u}=v_{u}^{\prime}, Lemma 4.3. If vs=vsv_{s}=v_{s}^{\prime}, vu=vuv_{u}=v_{u}^{\prime} then we can simplify and obtain ηct=ηct\eta_{c}^{t}=\eta_{c}^{t^{\prime}}. Since ηc\eta_{c} is a free action this implies t=tt=t^{\prime}. So T:E0s×E0u×NT:E_{0}^{s}\times E_{0}^{u}\times\mathbb{R}\to N is bijective. It is also clear that TT is continuous. Let (vs,vu,t)(v_{s},v_{u},t) and (vs,vu,t)(v_{s}^{\prime},v_{u}^{\prime},t^{\prime}) be such that T(vs,vu,t)T(v_{s},v_{u},t) is close to T(vs,vu,t)T(v_{s}^{\prime},v_{u}^{\prime},t^{\prime}). After applying Φ\Phi, it is immediate that vsv_{s} is close to vsv_{s}^{\prime} and vuv_{u} is close to vuv_{u}^{\prime}. Writing h=T(vs,vu,t)T(vs,vu,t)1h=T(v_{s},v_{u},t)T(v_{s}^{\prime},v_{u}^{\prime},t^{\prime})^{-1} we obtain

h=ηsvsηuvuvuηsvsηctt,ηctt=ηsvsηuvu+vuηsvsh\displaystyle h=\eta_{s}^{v_{s}}\eta_{u}^{v_{u}-v_{u}^{\prime}}\eta_{s}^{v_{s}^{\prime}}\eta_{c}^{t-t^{\prime}},\quad\eta_{c}^{t-t^{\prime}}=\eta_{s}^{-v_{s}^{\prime}}\eta_{u}^{-v_{u}+v_{u}^{\prime}}\eta_{s}^{-v_{s}}h

so ηctt\eta_{c}^{t-t^{\prime}} is close to idG{\rm id}_{G}. Since ηc\eta_{c} act freely ttt-t^{\prime} is close to 0, so TT has a continuous inverse and is therefore a homeomorphism. Since NN is a topological group homeomorphic to E0s×E0u×E_{0}^{s}\times E_{0}^{u}\times\mathbb{R}, NN has a unique structure as a Lie group [49, 29]. Finally, NN is 22-step nilpotent as an abstract group so NN is a 22-step nilpotent Lie group. ∎

Theorem 7.1.

The diffeomorphism f:XXf:X\to X is bi-Hölder conjugate to some affine map f0:XXf_{0}:X\to X where f0(x)=L(x)z0f_{0}(x)=L(x)z_{0} with z0Gcz_{0}\in G^{c}.

Proof.

Let NN be the group from Lemma 7.3 and FF a lift of ff. For w𝒫w\in\mathcal{P} we have Fη𝒫wF1=η𝒫LsuwF\eta_{\mathcal{P}}^{w}F^{-1}=\eta_{\mathcal{P}}^{L_{su}w} by Lemma 4.7. It follows that L~(n)=FnF1N\tilde{L}(n)=FnF^{-1}\in N for every nNn\in N. So, L~:NN\tilde{L}:N\to N is a continuous autormophism such that Fn=L~(n)FFn=\tilde{L}(n)F for nNn\in N.

Define ΛN\Lambda\subset N by λΛ\lambda\in\Lambda if λ(e)Γ\lambda(e)\in\Gamma for the identity element eGe\in G. Any two λ,λΛ\lambda,\lambda^{\prime}\in\Lambda are lifted from XΓX_{\Gamma} so there is γΓ\gamma\in\Gamma such that λ(λ(e))=λ(γ)=γλ(e)Γ\lambda^{\prime}(\lambda(e))=\lambda^{\prime}(\gamma)=\gamma\lambda^{\prime}(e)\in\Gamma. Similarly if λΛ\lambda\in\Lambda then λ1(λ(e))=λ(e)λ1(e)\lambda^{-1}(\lambda(e))=\lambda(e)\lambda^{-1}(e) so λ1(e)=(λ(e))1Γ\lambda^{-1}(e)=(\lambda(e))^{-1}\in\Gamma. It follows that Λ\Lambda is a subgroup of NN. Moreover, Γ\Gamma is closed in GG so ΛN\Lambda\leq N is closed.

The map q:Nhh(e)Gq:N\ni h\mapsto h(e)\in G is a homeomorphism by invariance of domain since qq is bijective. Let hNh\in N and λΛ\lambda\in\Lambda, then

q(hλ)=h(λ(e))=λ(e)h(e)=λ(e)q(h)=q(λ)q(h)\displaystyle q(h\lambda)=h(\lambda(e))=\lambda(e)h(e)=\lambda(e)q(h)=q(\lambda)q(h)

so qq descends to a map N/ΛΓG=XΓN/\Lambda\to\Gamma\setminus G=X_{\Gamma}. If q(h)=γq(h)q(h)=\gamma q(h^{\prime}) for h,hNh,h^{\prime}\in N and γΓ\gamma\in\Gamma then we find λΛ\lambda\in\Lambda such that γ=λ(e)\gamma=\lambda(e) (since the action of NN on GG is transitive). It follows that

q(h)=λ(e)q(h)=λ(e)h(e)=h(λ(e))=q(hλ)\displaystyle q(h)=\lambda(e)q(h^{\prime})=\lambda(e)h^{\prime}(e)=h^{\prime}(\lambda(e))=q(h^{\prime}\lambda)

or h(e)=h(λ(e))h(e)=h^{\prime}(\lambda(e)). The action of NN is free, so h=hλh=h^{\prime}\lambda. That is, we have a diagram

N{N}G{G}N/Λ{N/\Lambda}XΓ{X_{\Gamma}}q\scriptstyle{q}q\scriptstyle{q}

where both horizontal maps are homeomorphisms. The group NN is a nilpotent Lie group by Lemma 7.3, so N/ΛN/\Lambda is a nilmanifold. It follows that NN is isomorphic to GG as a Lie group and Λ\Lambda is isomorphic to Γ\Gamma under this map NGN\to G [51].

We claim that the induced map f~:N/ΛN/Λ\tilde{f}:N/\Lambda\to N/\Lambda, f~=q1fq\tilde{f}=q^{-1}\circ f\circ q is affine. Equivalently, F~:NN\tilde{F}:N\to N, F~=q1Fq\tilde{F}=q^{-1}\circ F\circ q is affine. From the relation Fn=(L~n)FFn=(\tilde{L}n)F, nNn\in N, we obtain

q(F~(n))=F(q(n))=F(n(e))=L~(n)(Fe)=L~(n)(n0(e))=q(L~(n)n0)\displaystyle q\left(\tilde{F}(n)\right)=F(q(n))=F(n(e))=\tilde{L}(n)(Fe)=\tilde{L}(n)\left(n_{0}(e)\right)=q\left(\tilde{L}(n)\cdot n_{0}\right)

where n0Nn_{0}\in N is chosen such that n0(e)=Fen_{0}(e)=Fe. It follows that F~\tilde{F}, and therefore also f~\tilde{f}, is affine.

Denote by H:XΓXΓH:X_{\Gamma}\to X_{\Gamma} the conjugacy such that H(fx)=f0(h(x))=L(H(x))z0H(fx)=f_{0}(h(x))=L(H(x))z_{0}. After conjugating with a translation we may assume that z0Gc=[G,G]z_{0}\in G^{c}=[G,G]. To finish the proof we show that HH is bi-Hölder. By uniqueness of Φ\Phi we have Φ(x)=π(H(x))=H(x)Gc\Phi(x)=\pi(H(x))=H(x)G^{c}. Since Φ:XΓ𝕋d\Phi:X_{\Gamma}\to\mathbb{T}^{d} is Hölder and the inverse of Φ\Phi restricted to stable and unstable leaves is Hölder (Lemma 3.3) HH is bi-Hölder along WsW^{s} and WuW^{u}. So it suffices to show that HH is Hölder along Wc(x)W^{c}(x) and H1H^{-1} is Hölder along W0c(x)W_{0}^{c}(x). Write

(7.5) H(x)=xeh(x)=xe(hs(x)+hc(x)+hu(x))\displaystyle H(x)=xe^{-h(x)}=xe^{-\left(h_{s}(x)+h_{c}(x)+h_{u}(x)\right)}

with h:XΓ𝔤h:X_{\Gamma}\to\mathfrak{g} and hσ:XΓE0σh_{\sigma}:X_{\Gamma}\to E_{0}^{\sigma}, σ=s,c,u\sigma=s,c,u. Since π(H(x))=Φ(x)\pi(H(x))=\Phi(x) and Φ\Phi is Hölder, both hsh_{s} and huh_{u} are Hölder. The functional equation for HH implies

H(fx)=\displaystyle H(fx)= (fx)eh(fx)=L(x)ev(x)eh(fx)=L(x)ev(x)h(fx)+[v(x),h(fx)]2=\displaystyle(fx)e^{-h(fx)}=L(x)e^{-v(x)}e^{-h(fx)}=L(x)e^{-v(x)-h(fx)+\frac{[v(x),h(fx)]}{2}}=
L(H(x))z0=L(x)eLh(x)+Z0\displaystyle L(H(x))z_{0}=L(x)e^{-Lh(x)+Z_{0}}

or

(7.6) h(fx)Lh(x)=[v(x),h(fx)]2v(x)+Z0.\displaystyle h(fx)-Lh(x)=\frac{[v(x),h(fx)]}{2}-v(x)+Z_{0}.

We obtain an equation for hc(x)h_{c}(x)

hc(fx)hc(x)=[vs(x),hu(fx)]+[vu(x),hs(fx)]2vc(x)+Z0=w(x)\displaystyle h_{c}(fx)-h_{c}(x)=\frac{[v_{s}(x),h_{u}(fx)]+[v_{u}(x),h_{s}(fx)]}{2}-v_{c}(x)+Z_{0}=w(x)

with w:XΓE0cw:X_{\Gamma}\to E_{0}^{c} Hölder. It follows that hc(x)h_{c}(x) is Hölder [70], so HH is Hölder. Next we show that H1H^{-1} is Hölder along W0c(x)W_{0}^{c}(x). Fix XE0sX\in E_{0}^{s}, YE0uY\in E_{0}^{u} such that [X,Y]=Z0[X,Y]=Z\neq 0 with ZE0cZ\in E_{0}^{c}. Given t0t\geq 0 the Baker–Campbell–Hausdorff formula implies etXetYetXetY=et2[X,Y]e^{tX}e^{tY}e^{-tX}e^{-tY}=e^{t^{2}[X,Y]}, so

H1(xet2Z)H1(x)=\displaystyle H^{-1}\left(xe^{t^{2}Z}\right)-H^{-1}\left(x\right)= H1(xetXetYetXetY)H1(x)=\displaystyle H^{-1}\left(xe^{tX}e^{tY}e^{-tX}e^{-tY}\right)-H^{-1}\left(x\right)=
H1(xetXetYetXetY)H1(xetXetYetX)+\displaystyle H^{-1}\left(xe^{tX}e^{tY}e^{-tX}e^{-tY}\right)-H^{-1}\left(xe^{tX}e^{tY}e^{-tX}\right)+
H1(xetXetYetX)H1(xetXetY)+\displaystyle H^{-1}\left(xe^{tX}e^{tY}e^{-tX}\right)-H^{-1}\left(xe^{tX}e^{tY}\right)+
H1(xetXetY)H1(xetX)+\displaystyle H^{-1}\left(xe^{tX}e^{tY}\right)-H^{-1}\left(xe^{tX}\right)+
H1(xetX)H1(x).\displaystyle H^{-1}\left(xe^{tX}\right)-H^{-1}\left(x\right).

Letting tt be close to 0 and using that H1H^{-1} is Hölder along W0sW_{0}^{s} and W0uW_{0}^{u} we obtain

d(H1(xet2Z),H1(x))Ctθ.\displaystyle{\rm d}\left(H^{-1}\left(xe^{t^{2}Z}\right),H^{-1}\left(x\right)\right)\leq Ct^{\theta}.

Since d(xet2Z,x)ct2{\rm d}(xe^{t^{2}Z},x)\geq ct^{2} for small tt we have

d(H1(xet2Z),H1(x))Cd(x,xet2Z)θ/2,\displaystyle{\rm d}\left(H^{-1}\left(xe^{t^{2}Z}\right),H^{-1}\left(x\right)\right)\leq C{\rm d}\left(x,xe^{t^{2}Z}\right)^{\theta/2},

so for yW0c(x)y\in W_{0}^{c}(x) close. It follows that d(H1(x),H1(y))Cd(x,y)θ/2{\rm d}(H^{-1}(x),H^{-1}(y))\leq C{\rm d}(x,y)^{\theta/2}. ∎

8. Rigidity: Smoothness of the bi-Hölder conjugacy

In this section, we prove Theorem 1.2. Let α\alpha, α0\alpha_{0}, f=α𝐧0f=\alpha^{\mathbf{n}_{0}}, f0=α0𝐧0f_{0}=\alpha_{0}^{\mathbf{n}_{0}}, and HH be as in Theorem 1.2. We prove Theorem 1.2 under the assumption that ff is accessible, this is done for two reasons. First, the proof simplifies because we can apply [70] to obtain regularity of the conjugacy along the center direction. In particular, there is no loss of generality in assuming that the center of f0f_{0} coincides with the joint center of α0\alpha_{0} (see Remark 20). Second, if ff is accessible then α\alpha naturally preserves a volume form μ\mu and Hμ=μΓH_{*}\mu=\mu_{\Gamma} (see Lemma 8.2).

If α\alpha is assumed to preserve a volume form μ\mu such that Hμ=μΓH_{*}\mu=\mu_{\Gamma} then a result similar to Theorem 1.2 still holds, without accessibility. We give a brief sketch of the proof. Let H:GGH:G\to G be a lift of the conjugacy. By the argument below HH is uniformly CC^{\infty} along WσW^{\sigma} and DxH:Eσ(x)E0σD_{x}H:E^{\sigma}(x)\to E_{0}^{\sigma} is invertible at each xx for σ=s,u\sigma=s,u. Define

(8.1) H^:G𝐻GG/Gcs\displaystyle\hat{H}:G\xrightarrow{H}G\to G/G^{cs}

then the fibers of H^\hat{H} coincides with W^cs\hat{W}^{cs}. In particular, H^\hat{H} is uniformly CC^{\infty} along W^cs\hat{W}^{cs}. Moreover, HH is uniformly CC^{\infty} along WuW^{u} so H^\hat{H} is uniformly CC^{\infty} along WuW^{u}. By Journé’s lemma [37] H^\hat{H} is smooth. The map DxH^|Eu(x)D_{x}\hat{H}|_{E^{u}(x)} is invertible, so H^\hat{H} is a submersion. It follows that the leaves of H^\hat{H} form a CC^{\infty}-foliation, so WcsW^{cs} is a CC^{\infty}-foliation. Similarly, WcuW^{cu} is a CC^{\infty}-foliation. Once we know that WcW^{c} is a CC^{\infty}-foliation the arguments in [55, 26], to obtain regularity of HH along WsW^{s} and WuW^{u}, can also be used along WcW^{c} (note that, a priori, the assumptions of [55, Theorem A.1] are not satisfied for WcW^{c}).

8.1. Dynamical coherence and regularity of center leaves

We begin by proving that ff is dynamically coherent with WcsW^{cs}, WcuW^{cu} and WcW^{c} all uniquely integrable. We use this without mention in the remainder.

Lemma 8.1.

The map ff is dynamically coherent with WcsW^{cs}, WcuW^{cu} and WcW^{c} all uniquely integrable with uniformly CC^{\infty} leaves.

Proof.

Since ff is conjugated to f0=α0𝐧0f_{0}=\alpha_{0}^{\mathbf{n}_{0}}, and f0f_{0} is uniformly subexponetial along its center E0cE_{0}^{c}, ff is also uniformly subexponential along its center EcE^{c}. Indeed, for any ff-invariant measure ν\nu the stable Pesin manifold Wνs(x)W_{\nu}^{s}(x) maps into the stable manifold, W0s(H(x))W_{0}^{s}(H(x)), of f0f_{0} under HH. Since HH is invertible dimWνs(x)dimW0s(H(x))\dim W_{\nu}^{s}(x)\leq\dim W_{0}^{s}(H(x)). It follows that no ν\nu-exponents are negative along EcE^{c}. By exchanging ff for f1f^{-1}, no ν\nu-exponent is positive along EcE^{c}. By [58] ff is uniformly subexponential along EcE^{c}. Fix lifts FF, F0F_{0}, and HH to GG such that H(Fx)=F0(H(x))H(Fx)=F_{0}(H(x)). If γ:IG\gamma:I\to G is a C1C^{1}-curve tangent to EcsE^{cs} then the length |Fnγ||F^{n}\circ\gamma| satisfies

(8.2) |Fnγ|Cεenε,\displaystyle|F^{n}\circ\gamma|\leq C_{\varepsilon}e^{n\varepsilon},

for any ε>0\varepsilon>0. The conjugacy H:GGH:G\to G can be written H(x)=xh(x)1H(x)=xh(x)^{-1} with h:GGh:G\to G being Γ\Gamma-invariant. We estimate

d(H(x),H(y))=\displaystyle{\rm d}(H(x),H(y))= d(xh(x)1,yh(y)1)\displaystyle{\rm d}(xh(x)^{-1},yh(y)^{-1})\leq
\displaystyle\leq d(x,xh(x)1)+d(y,yh(y)1)+d(x,y)=\displaystyle{\rm d}(x,xh(x)^{-1})+{\rm d}(y,yh(y)^{-1})+{\rm d}(x,y)=
d(e,h(x)1)+d(e,h(y)1)+d(x,y)d(x,y)+C\displaystyle{\rm d}(e,h(x)^{-1})+{\rm d}(e,h(y)^{-1})+{\rm d}(x,y)\leq{\rm d}(x,y)+C

for some uniform CC. With x=H1(z)x=H^{-1}(z) and y=H1(w)y=H^{-1}(w) we obtain

d(H1(z),H1(w))d(z,w)C.\displaystyle{\rm d}(H^{-1}(z),H^{-1}(w))\geq{\rm d}(z,w)-C.

Now, d(Fn(γ(0)),Fn(γ(1)))Cεenε{\rm d}(F^{n}(\gamma(0)),F^{n}(\gamma(1)))\leq C_{\varepsilon}e^{n\varepsilon}, so

Cεenε\displaystyle C_{\varepsilon}e^{n\varepsilon}\geq d(Fn(γ(0)),Fn(γ(1)))=d(H1HFn(γ(0)),H1HFn(γ(1)))\displaystyle{\rm d}(F^{n}(\gamma(0)),F^{n}(\gamma(1)))={\rm d}(H^{-1}HF^{n}(\gamma(0)),H^{-1}HF^{n}(\gamma(1)))\geq
d(HFn(γ(0)),HFn(γ(1)))C=d(F0nH(γ(0)),F0n(H(γ(1))))C.\displaystyle{\rm d}(HF^{n}(\gamma(0)),HF^{n}(\gamma(1)))-C={\rm d}(F_{0}^{n}H(\gamma(0)),F_{0}^{n}(H(\gamma(1))))-C.

With ε\varepsilon sufficiently small it follows that H(γ(1))W0cs(H(γ(1)))H(\gamma(1))\in W_{0}^{cs}(H(\gamma(1))). We conclude that EcsE^{cs} is uniquely integrable with leaves given by

(8.3) Wcs(x)=H1(W0cs(H(x))).\displaystyle W^{cs}(x)=H^{-1}(W_{0}^{cs}(H(x))).

Similarly, EcuE^{cu} is uniquely integrable with Wcu(x)=H1(W0cu(H(x)))W^{cu}(x)=H^{-1}(W_{0}^{cu}(H(x))). Since ff is uniformly subexponential along EcE^{c} each foliation WcsW^{cs}, WcuW^{cu} and Wc=WcsWcuW^{c}=W^{cs}\cap W^{cu} have uniformly CC^{\infty} leaves, see [20, Theorem 7] (or [36]). ∎

8.2. Volume preservation and smoothness along the center direction

To apply arguments using exponential mixing we need to show that α\alpha preserves a smooth volume form. We assume that the action α\alpha is accessible so we show that HH is CC^{\infty} along WcW^{c} without using the exponential mixing argument from [26] and instead rely on results from [70].

Lemma 8.2.

Let α\alpha be as in Theorem 1.2, then α\alpha preserve a smooth volume form μ\mu. Moreover, the conjugacy H:XΓXΓH:X_{\Gamma}\to X_{\Gamma} is volume preserving in the sense that Hμ=μΓH_{*}\mu=\mu_{\Gamma} where μΓ\mu_{\Gamma} is the Haar measure on XΓX_{\Gamma}.

Remark 19.

By Moser’s trick, there is no loss of generality if we assume that μ=μΓ\mu=\mu_{\Gamma}.

Proof.

Existence of an invariant Hölder continuous volume form, μ\mu, follows from [31, Theorem 1.5]. We assume accessibility, so smoothness of μ\mu follows from [70, Theorem A, case (IV)]. As in [25, Proposition 2.4], HμH_{*}\mu is a measure of maximal entropy for the α0\alpha_{0}-action. From [64] it follows that Hμ=μΓH_{*}\mu=\mu_{\Gamma}. ∎

We show that HH is uniformly CC^{\infty} along WcW^{c} following [70].

Lemma 8.3.

The restriction H:Wc(x)W0c(H(x))H:W^{c}(x)\to W_{0}^{c}(H(x)) is uniformly CC^{\infty}.

Remark 20.

Using Lemma 8.3 we may assume, without loss of generality, that the center E0cE_{0}^{c} of f0f_{0} coincides with the joint center of α\alpha.

Proof.

Let M=Wc(x)×W0c(H(x))M=W^{c}(x)\times W_{0}^{c}(H(x)) and let NMN\subset M be the graph of HH:

(8.4) N={(y,H(y)) : yWc(x)}.\displaystyle N=\{(y,H(y))\text{ : }y\in W^{c}(x)\}.

Given any two z,wWc(x)z,w\in W^{c}(x) we fix a susu-path γ\gamma from zz to ww, such a path always exists by accessibility. Denote by hz,wγ:Wc(x)Wc(x)h_{z,w}^{\gamma}:W^{c}(x)\to W^{c}(x) the composition of holonomy maps along γ\gamma (z,wWc(x)z,w\in W^{c}(x) so Wc(z)=Wc(w)=Wc(x)W^{c}(z)=W^{c}(w)=W^{c}(x)). Since ff is uniformly subexponential along WcW^{c}, ff is \infty-bunching so hz,wγh_{z,w}^{\gamma} is CC^{\infty} [50]. Since HH map WσW^{\sigma}, σ=s,u\sigma=s,u, onto W0σW_{0}^{\sigma} we have Hhz,wγ=hH(z),H(w)Hγ,0HHh_{z,w}^{\gamma}=h_{H(z),H(w)}^{H\gamma,0}H where hH(z),H(w)Hγ,0:W0c(H(x))W0c(H(x))h_{H(z),H(w)}^{H\gamma,0}:W_{0}^{c}(H(x))\to W_{0}^{c}(H(x)) is the composition of holonomy maps along HγH\gamma. Define

(8.5) h^z,w:MM,h^z,w(p,q)=(hz,wγ(p),hH(z),H(w)Hγ,0(q)).\displaystyle\hat{h}_{z,w}:M\to M,\quad\hat{h}_{z,w}(p,q)=\left(h_{z,w}^{\gamma}(p),h_{H(z),H(w)}^{H\gamma,0}(q)\right).

Since Hhz,wγ=hH(z),H(w)Hγ,0HHh_{z,w}^{\gamma}=h_{H(z),H(w)}^{H\gamma,0}H we have h^z,w(N)=N\hat{h}_{z,w}(N)=N. Moreover, h^z,w(z,H(z))=(w,H(w))\hat{h}_{z,w}(z,H(z))=(w,H(w)), zz and ww were arbitrary, and h^z,w\hat{h}_{z,w} is smooth so NN is CC^{\infty}-homogeneous (see [70] for definitions). By [70, Corollary 1.3] NN is a CC^{\infty} submanifold. The graph of H:Wc(x)Wc(H(x))H:W^{c}(x)\to W^{c}(H(x)) is CC^{\infty} and it follows that H:Wc(x)W0c(H(x))H:W^{c}(x)\to W_{0}^{c}(H(x)) is also CC^{\infty}. Finally, HH is uniformly CC^{\infty} along WcW^{c} since it intertwines the holonomies of ff with the holonomies of f0f_{0}. ∎

8.3. Smoothness of coarse components along the stable foliation

Let 𝔤=E0sE0cE0u\mathfrak{g}=E_{0}^{s}\oplus E_{0}^{c}\oplus E_{0}^{u} be the splitting of 𝔤\mathfrak{g} with respect to f0f_{0}. Denote by GσG^{\sigma}, σ=s,c,u,cs,cu\sigma=s,c,u,cs,cu, the subgroup associated to E0σE_{0}^{\sigma}. We have a CC^{\infty}-diffeomorphism Gs×Gc×GuGG^{s}\times G^{c}\times G^{u}\to G defined by (gs,gc,gu)gsgcgu(g^{s},g^{c},g^{u})\mapsto g_{s}g_{c}g_{u}. Write H(x)=xh(x)1H(x)=xh(x)^{-1} with h:XΓGh:X_{\Gamma}\to G satisfying h(γx)=h(x)h(\gamma x)=h(x) for all γΓ\gamma\in\Gamma. We decompose h(x)h(x) with respect to the map Gs×Gc×GuGG^{s}\times G^{c}\times G^{u}\to G

(8.6) h(x)=hs(x)hc(x)hu(x).\displaystyle h(x)=h_{s}(x)h_{c}(x)h_{u}(x).

It is immediate that each hσh_{\sigma} is Hölder. Given a coarse exponent [χ][\chi] we denote by G[χ]G^{[\chi]} the subgroup associated to E0[χ]E_{0}^{[\chi]}.

Let [χ][\chi] be a coarse exponent along E0sE_{0}^{s}. We decompose hs(x)h_{s}(x) further as

(8.7) hs(x)=hss(x)h[χ](x)\displaystyle h_{s}(x)=h_{ss}(x)h_{[\chi]}(x)

where h[χ](x)h_{[\chi]}(x) is the component of hs(x)h_{s}(x) along G[χ]G^{[\chi]} and hss(x)h_{ss}(x) is the component of hs(x)h_{s}(x) along the complementary group GssG^{ss} (see [55, Lemma 3.1]). The following is proved in [55, Section 3].

Lemma 8.4.

Let α\alpha be as in Theorem 1.2 with α𝐧0=f\alpha^{\mathbf{n}_{0}}=f partially hyperbolic and ker[χ]\ker[\chi], [χ](𝐧0)<0[\chi](\mathbf{n}_{0})<0, a wall of the Weyl chamber that contains 𝐧0\mathbf{n}_{0}. The map h[χ](x)h_{[\chi]}(x) in the Equation 8.7 is CC^{\infty} along WsW^{s}, with all derivatives along WsW^{s} uniformly Hölder.

Proof.

The proof follows as in [55, Section 3] once we note that we do not need ff to be Anosov. Indeed, once we restrict to the foliation WsW^{s} the argument only requires the action α\alpha to be exponentially mixing with respect to volume. Exponential mixing with respect to volume follows from Lemma 8.2 and [31]. ∎

8.4. New partially hyperbolic elements: passing the chamber wall

Let 𝐧0\mathbf{n}_{0} be in a Weyl chamber 𝒞\mathcal{C} and let ker[χ]\ker[\chi] be a chamber wall for 𝒞\mathcal{C}. Now we start the work of passing the Weyl chamber wall ker[χ]\ker[\chi] by constructing a partially hyperbolic element in the chamber 𝒞\mathcal{C}^{\prime} adjacent to 𝒞\mathcal{C} through ker[χ]\ker[\chi]. We initially follow [55], but change the argument from Section 44 in [55], by not relying on smooth ergodic theory. If xXΓx\in X_{\Gamma} then the map H:Ws(x)W0s(H(x))=H(x)GsH:W^{s}(x)\to W_{0}^{s}(H(x))=H(x)G^{s} is a homeomorphism. Define

(8.8) Hs,x:Ws(x)Gs,H(y)=H(x)(Hs,x(y))1.\displaystyle H_{s,x}:W^{s}(x)\to G^{s},\quad H(y)=H(x)\left(H_{s,x}(y)\right)^{-1}.

If α0𝐧(x)=ρ𝐧(x)η𝐧1\alpha_{0}^{\mathbf{n}}(x)=\rho^{\mathbf{n}}(x)\eta_{\mathbf{n}}^{-1} with ρ𝐧Aut(XΓ)\rho^{\mathbf{n}}\in{\rm Aut}(X_{\Gamma}) and η𝐧G\eta_{\mathbf{n}}\in G then Hs,xH_{s,x} satisfy

Hs,α𝐧x(α𝐧y)=η𝐧ρ𝐧(Hs,x(y))η𝐧1.\displaystyle H_{s,\alpha^{\mathbf{n}}x}(\alpha^{\mathbf{n}}y)=\eta_{\mathbf{n}}\rho^{\mathbf{n}}\left(H_{s,x}(y)\right)\eta_{\mathbf{n}}^{-1}.

For yWs(x)y\in W^{s}(x) we write y=xgx(y)1y=xg_{x}(y)^{-1} where (x,y)gx(y)(x,y)\mapsto g_{x}(y) is chosen continuously and such that gx(x)=eg_{x}(x)=e. Then gx:Ws(x)Gsg_{x}:W^{s}(x)\to G^{s} is CC^{\infty}. With this notation

xgx(y)1h(y)1=yh(y)1=H(y)=H(x)Hs,x(y)1=xh(x)1Hs,x(y)1\displaystyle xg_{x}(y)^{-1}h(y)^{-1}=yh(y)^{-1}=H(y)=H(x)H_{s,x}(y)^{-1}=xh(x)^{-1}H_{s,x}(y)^{-1}

or

Hs,x(y)=\displaystyle H_{s,x}(y)= h(y)gx(y)h(x)1=\displaystyle h(y)g_{x}(y)h(x)^{-1}=
hs(y)hcu(y)[(h(x)gx(y)1)s(h(x)gx(y)1)cu]1=\displaystyle h_{s}(y)h_{cu}(y)\left[(h(x)g_{x}(y)^{-1})_{s}(h(x)g_{x}(y)^{-1})_{cu}\right]^{-1}=
hs(y)[hcu(y)(h(x)gx(y)1)cu1](h(x)gx(y)1)s1.\displaystyle h_{s}(y)\left[h_{cu}(y)\left(h(x)g_{x}(y)^{-1}\right)_{cu}^{-1}\right]\left(h(x)g_{x}(y)^{-1}\right)_{s}^{-1}.

If we multiply on the left by hs(y)1h_{s}(y)^{-1} and on the right by (h(x)gx(y))s(h(x)g_{x}(y))_{s}, then the right-hand side of the equality lie in GcuG^{cu}, but the left-hand side lies in GsG^{s}. It follows that both sides of the equality are identity, so for yWs(x)y\in W^{s}(x) we have

(8.9) hcu(y)=(h(x)gx(y)1)cu,\displaystyle h_{cu}(y)=\left(h(x)g_{x}(y)^{-1}\right)_{cu},
(8.10) Hs,x(y)=hs(y)(h(x)gx(y)1)s1.\displaystyle H_{s,x}(y)=h_{s}(y)\left(h(x)g_{x}(y)^{-1}\right)_{s}^{-1}.

Using the fact that the map aaσa\mapsto a_{\sigma}, σ=s,cu\sigma=s,cu, and the map gx:Ws(x)Gg_{x}:W^{s}(x)\to G are both smooth the following is immediate, see also [55, Corollary 3.14].

Lemma 8.5.

The map hcuh_{cu} is uniformly CC^{\infty} along WsW^{s}.

Define Hs,x(y)=(Hs,x(y))ssHs,x[χ](y)H_{s,x}(y)=(H_{s,x}(y))_{ss}H_{s,x}^{[\chi]}(y) with Hs,x[χ](y)G[χ]H_{s,x}^{[\chi]}(y)\in G^{[\chi]}. From our definitions it is immediate

{yWs(x) : Hs,x[χ](y)=e}=H1(W0ss(H(x))),W0ss(y)=yGss.\displaystyle\{y\in W^{s}(x)\text{ : }H_{s,x}^{[\chi]}(y)=e\}=H^{-1}(W_{0}^{ss}(H(x))),\quad W_{0}^{ss}(y)=yG^{ss}.

So if we prove that Hs,x[χ]H_{s,x}^{[\chi]} is a (local) CC^{\infty} submersion for every xx, then the fibers of Hs,x[χ]H_{s,x}^{[\chi]} defines a smooth foliation (within WsW^{s}). This shows that the, a priori only Hölder, foliation Wss(x)=H1(W0ss(H(x)))W^{ss}(x)=H^{-1}(W_{0}^{ss}(H(x))) is a Hölder foliation with uniformly smooth leaves.

Lemma 8.6.

The map Hs,x[χ]:Ws(x)G[χ]H_{s,x}^{[\chi]}:W^{s}(x)\to G^{[\chi]} is uniformly CC^{\infty}.

Proof.

Since GssG^{ss} is normal in GsG^{s} [55, Lemma 3.1] we have

Hs,x(y)=\displaystyle H_{s,x}(y)= hs(y)(h(x)gx(y)1)s1=\displaystyle h_{s}(y)\left(h(x)g_{x}(y)^{-1}\right)_{s}^{-1}=
hss(y)h[χ](y)[(h(x)gx(y)1)s1]ss[(h(x)gx(y)1)s1][χ]=\displaystyle h_{ss}(y)h_{[\chi]}(y)\left[\left(h(x)g_{x}(y)^{-1}\right)_{s}^{-1}\right]_{ss}\left[\left(h(x)g_{x}(y)^{-1}\right)_{s}^{-1}\right]_{[\chi]}=
assh[χ](y)[(h(x)gx(y)1)s1][χ]\displaystyle a_{ss}\cdot h_{[\chi]}(y)\left[\left(h(x)g_{x}(y)^{-1}\right)_{s}^{-1}\right]_{[\chi]}

for some assGssa_{ss}\in G^{ss}. That is, we obtain the formula

(8.11) Hs,x[χ](y)=h[χ](y)[(h(x)gx(y)1)s1][χ]\displaystyle H_{s,x}^{[\chi]}(y)=h_{[\chi]}(y)\left[\left(h(x)g_{x}(y)^{-1}\right)_{s}^{-1}\right]_{[\chi]}

so Hs,x[χ]H_{s,x}^{[\chi]} is uniformly CC^{\infty} along WsW^{s} since h[χ]h_{[\chi]} and gxg_{x} are (Lemma 8.4). ∎

Lemma 8.7.

The map DxHs,x[χ]:Es(x)E0[χ]D_{x}H_{s,x}^{[\chi]}:E^{s}(x)\to E_{0}^{[\chi]} is surjective at every xXΓx\in X_{\Gamma}. In particular, the foliation WssW^{ss} has uniformly CC^{\infty} leaves.

Proof.

Denote by K={DxHs,x[χ] : is not surjective}K=\{D_{x}H_{s,x}^{[\chi]}\text{ : is not surjective}\}, then KK is compact and α\alpha-invariant. Our goal is to show that K=K=\emptyset. Since H:Ws(x)W0s(H(x))H:W^{s}(x)\to W_{0}^{s}(H(x)) is surjective it follows by Sard’s theorem that KXΓK\neq X_{\Gamma}. We will prove that if KK is non-empty then KK contains a WssW^{ss}-leaf, which is a contradiction since every WssW^{ss}-leaf is dense (Lemma A.1).

In the remainder, fix some background metric ,\langle\cdot,\cdot\rangle and calculate determinants with respect to the top form induced by ,\langle\cdot,\cdot\rangle. Assume for contradiction that KK\neq\emptyset. Fix 𝐧\mathbf{n} close to the kernel ker[χ]\ker[\chi] (we specify later how close) such that α𝐧\alpha^{\mathbf{n}} contract Wss:=H1(H(x)Gss)W^{ss}:=H^{-1}(H(x)G^{ss}) and W[χ]W^{[\chi]}. Fix yWss(x)y\in W^{ss}(x) and some subspace V0Es(y)V_{0}\subset E^{s}(y) of dimension dim(E0[χ])\dim(E_{0}^{[\chi]}). We also define Vn=Dyαn𝐧V0V_{n}=D_{y}\alpha^{n\mathbf{n}}V_{0}. The relation Hs,α𝐦x[χ](α𝐦y)=η𝐦ρ𝐦(Hs,x[χ](y))η𝐦1H_{s,\alpha^{\mathbf{m}}x}^{[\chi]}(\alpha^{\mathbf{m}}y)=\eta_{\mathbf{m}}\rho^{\mathbf{m}}\left(H_{s,x}^{[\chi]}(y)\right)\eta_{\mathbf{m}}^{-1} implies

(8.12) det(DyHs,y[χ]|V0)=det(ρn𝐧|E0[χ])det(Dαn𝐧yHs,αn𝐧y[χ]|Vn)det(Dyαn𝐧|V0).\displaystyle\det(D_{y}H_{s,y}^{[\chi]}|_{V_{0}})=\det(\rho^{-n\mathbf{n}}|_{E_{0}^{[\chi]}})\det(D_{\alpha^{n\mathbf{n}}y}H_{s,\alpha^{n\mathbf{n}}y}^{[\chi]}|_{V_{n}})\det(D_{y}\alpha^{n\mathbf{n}}|_{V_{0}}).

If χ\chi is a representative of [χ][\chi] then det(Dρn𝐧)Cernχ(𝐧)\det(D\rho^{-n\mathbf{n}})\leq Ce^{-rn\chi(\mathbf{n})} for some uniform r>0r>0 depending on the dimension of E0[χ]E_{0}^{[\chi]}. If =dim(E0[χ])\ell=\dim(E_{0}^{[\chi]}) and Gr(Es){\rm Gr}_{\ell}(E^{s}) is the \ell-grassmannian bundle of EsE^{s} then Gr(Es)(V,x)det(DxHs,x[χ]|V){\rm Gr}_{\ell}(E^{s})\ni(V,x)\mapsto\det(D_{x}H_{s,x}^{[\chi]}|_{V}) is uniformly smooth along WsW^{s}. For xKx\in K we have det(DxHs,x[χ]|V)=0\det(D_{x}H_{s,x}^{[\chi]}|_{V})=0 since DxHs,x[χ]D_{x}H_{s,x}^{[\chi]} is not surjective. It follows that there is some constant CC such that

(8.13) |det(DyHs,y[χ]|V)|Cds(y,K)\displaystyle|\det(D_{y}H_{s,y}^{[\chi]}|_{V})|\leq C{\rm d}_{s}(y,K)

for yWs(K)y\in W^{s}(K). If yWss(x)y\in W^{ss}(x), then

(8.14) det(Dαn𝐧yHαn𝐧y[χ]|Vn)\displaystyle\det(D_{\alpha^{n\mathbf{n}}y}H_{\alpha^{n\mathbf{n}}y}^{[\chi]}|_{V_{n}})\leq Cd(αn𝐧y,K)Cd(αn𝐧y,αn𝐧x)Ceλn\displaystyle C{\rm d}(\alpha^{n\mathbf{n}}y,K)\leq C{\rm d}(\alpha^{n\mathbf{n}}y,\alpha^{n\mathbf{n}}x)\leq C^{\prime}e^{-\lambda n}

where λ>0\lambda>0 can be chosen independently of 𝐧\mathbf{n} since we let 𝐧\mathbf{n} be close to the kernel of [χ][\chi]. Equations 8.12 and 8.14 implies

|det(DyHs,y[χ]|V0)|Cen(λ+rχ(𝐧))|det(Dyαn𝐧|V0)|.\displaystyle|\det(D_{y}H_{s,y}^{[\chi]}|_{V_{0}})|\leq Ce^{-n(\lambda+r\chi(\mathbf{n}))}|\det(D_{y}\alpha^{n\mathbf{n}}|_{V_{0}})|.

Since α0𝐧\alpha_{0}^{\mathbf{n}} contract the foliations WssW^{ss} and W[χ]W^{[\chi]} we have a uniform bound |det(Dyαn𝐧|V0)|C|\det(D_{y}\alpha^{n\mathbf{n}}|_{V_{0}})|\leq C. We obtain an estimate

|det(DyHs,y[χ]|V0)|Cen(λ+rχ(𝐧)).\displaystyle|\det(D_{y}H_{s,y}^{[\chi]}|_{V_{0}})|\leq Ce^{-n(\lambda+r\chi(\mathbf{n}))}.

With 𝐧\mathbf{n} be sufficiently close to ker[χ]\ker[\chi] we have λ+rχ(𝐧)>0\lambda+r\chi(\mathbf{n})>0, so if nn\to\infty then

det(DyHs,y[χ]|V0)=0.\displaystyle\det(D_{y}H_{s,y}^{[\chi]}|_{V_{0}})=0.

The point yy was arbitrary so Wss(x)KW^{ss}(x)\subset K for xKx\in K. The foliation WssW^{ss} is minimal (Lemma A.1) and KK is closed so K=XΓK=X_{\Gamma}. This contradicts Sard’s theorem. ∎

To show that E[χ]E^{[\chi]} exists as a continuous bundle we will apply a linear graph transform argument. Denote by Ess(x)E^{ss}(x) the Dα𝐧D\alpha^{\mathbf{n}}-invariant, continuous subbundle tangent to Wss(x)W^{ss}(x). Let F(x)F(x) be any continuous bundle that is complementary to EssE^{ss} within EsE^{s}. Fix 𝐧k\mathbf{n}\in\mathbb{Z}^{k} such that [χ](𝐧)<0[\chi](\mathbf{n})<0 and α𝐧\alpha^{\mathbf{n}} expands WssW^{ss} (that is, we chose 𝐧\mathbf{n} such that 𝐧-\mathbf{n} have passed the chamber wall ker[χ]\ker[\chi] from the chamber that contains 𝐧0\mathbf{n}_{0}). Write g=α𝐧g=\alpha^{\mathbf{n}}. With respect to the splitting Es=FEssE^{s}=F\oplus E^{ss} we write

Dxg(u,v)=(A(x)u,B(x)v+C(x)u)\displaystyle D_{x}g(u,v)=\left(A(x)u,B(x)v+C(x)u\right)

where A(x)uA(x)u is Dxg(u)D_{x}g(u) projected onto F(gx)F(gx) and C(x)C(x) is Dxg(u)D_{x}g(u) projected onto Ess(gx)E^{ss}(gx).

Lemma 8.8.

We have AC0<1\left\lVert A\right\rVert_{C^{0}}<1.

Proof.

Note that DxHs,x[χ]Dxg=Deα0𝐧DxHs,x[χ]D_{x}H_{s,x}^{[\chi]}D_{x}g=D_{e}\alpha_{0}^{\mathbf{n}}D_{x}H_{s,x}^{[\chi]}. There is μ<1\mu<1, depending on our choice of 𝐧k\mathbf{n}\in\mathbb{Z}^{k}, such that for any uF(x)u\in F(x) we have

(8.15) DgxHs,gx[χ](A(x)u+B(x)u)=Deα0𝐧DxHs,x[χ](u)μDxHs,x[χ](u).\displaystyle\left\lVert D_{gx}H_{s,gx}^{[\chi]}\left(A(x)u+B(x)u\right)\right\rVert=\left\lVert D_{e}\alpha_{0}^{\mathbf{n}}D_{x}H_{s,x}^{[\chi]}(u)\right\rVert\leq\mu\left\lVert D_{x}H_{s,x}^{[\chi]}(u)\right\rVert.

Since B(x)uEss(gx)B(x)u\in E^{ss}(gx) and kerDxHs,x[χ]=Ess(x)\ker D_{x}H_{s,x}^{[\chi]}=E^{ss}(x) (the fibers of Hs,x[χ]H_{s,x}^{[\chi]} are the foliation Wss(x)W^{ss}(x)) it follows that DgxHs,gx[χ]B(x)=0D_{gx}H_{s,gx}^{[\chi]}B(x)=0. That is, Equation 8.15 simplifies

(8.16) DgxHs,gx[χ]A(x)uμDxHs,x[χ](u).\displaystyle\left\lVert D_{gx}H_{s,gx}^{[\chi]}A(x)u\right\rVert\leq\mu\left\lVert D_{x}H_{s,x}^{[\chi]}(u)\right\rVert.

Since DxHs,x[χ]:F(x)E0[χ]D_{x}H_{s,x}^{[\chi]}:F(x)\to E_{0}^{[\chi]} is an isomorphism (of vector bundles after identifying E0[χ]E_{0}^{[\chi]} with the trivial bundle XΓ×E0[χ]XΓX_{\Gamma}\times E_{0}^{[\chi]}\to X_{\Gamma}) the lemma follows (after either changing the norm used or exchanging 𝐧\mathbf{n} for N𝐧N\mathbf{n} with NN sufficiently large). ∎

Lemma 8.9.

The map T:Γ0(Hom(F,Ess))Γ0(Hom(F,Ess))T:\Gamma^{0}({\rm Hom}(F,E^{ss}))\to\Gamma^{0}({\rm Hom}(F,E^{ss})) defined by

(TP)(x)=B(x)1(P(gx)A(x)C(x))\displaystyle(TP)(x)=B(x)^{-1}\left(P(gx)A(x)-C(x)\right)

has a unique fixed point.

Proof.

Since AC0,B1C0<1\left\lVert A\right\rVert_{C^{0}},\left\lVert B^{-1}\right\rVert_{C^{0}}<1 the lemma follows from Banach’s fixed point theorem. ∎

Lemma 8.10.

There exists an α\alpha-invariant continuous subbundle E[χ]EsE^{[\chi]}\subset E^{s} such that Es=E[χ]EssE^{s}=E^{[\chi]}\oplus E^{ss}.

Proof.

Let PΓ0(Hom(F,Ess))P\in\Gamma^{0}({\rm Hom}(F,E^{ss})) be the unique TT-fixed point from Lemma 8.9. Define E[χ](x):=Graph(P(x))={(u,P(x)u) : uF(x)}E^{[\chi]}(x):={\rm Graph}(P(x))=\{(u,P(x)u)\text{ : }u\in F(x)\}. It is immediate that E[χ]Ess=EsE^{[\chi]}\oplus E^{ss}=E^{s}. Given uF(x)u\in F(x)

Dxg(u,P(x)u)=\displaystyle D_{x}g(u,P(x)u)= (A(x)u,B(x)P(x)u+C(x)u)=\displaystyle\left(A(x)u,B(x)P(x)u+C(x)u\right)=
(A(x)u,P(gx)A(x)uC(x)u+C(x)u)=\displaystyle\left(A(x)u,P(gx)A(x)u-C(x)u+C(x)u\right)=
(A(x)u,P(gx)A(x)u)Graph(P(gx))\displaystyle\left(A(x)u,P(gx)A(x)u\right)\in{\rm Graph}(P(gx))

so DxgE[χ](x)E[χ](gx)D_{x}gE^{[\chi]}(x)\subset E^{[\chi]}(gx). Or DxgE[χ](x)=E[χ](gx)D_{x}gE^{[\chi]}(x)=E^{[\chi]}(gx) since DxgD_{x}g is invertible. That E[χ](x)E^{[\chi]}(x) is α𝐦\alpha^{\mathbf{m}}-invariant for all 𝐦k\mathbf{m}\in\mathbb{Z}^{k} follows by applying the graph transform of α𝐦\alpha^{\mathbf{m}} on the element PΓ0(Hom(F,Ess))P\in\Gamma^{0}({\rm Hom}(F,E^{ss})). This defines a TT-fixed point, since gg commute with α𝐦\alpha^{\mathbf{m}}, and the TT-fixed point is unique (Lemma 8.9) so the α𝐦\alpha^{\mathbf{m}}-graph transform of PP is PP. Equivalently, E[χ]E^{[\chi]} is Dxα𝐦D_{x}\alpha^{\mathbf{m}}-invariant. ∎

Since DxHs,x[χ]:E[χ](x)E0[χ]D_{x}H_{s,x}^{[\chi]}:E^{[\chi]}(x)\to E_{0}^{[\chi]} conjugates Dxα𝐧D_{x}\alpha^{\mathbf{n}} to Ad(η𝐧)Dρ𝐧{\rm Ad}(\eta_{\mathbf{n}})D\rho^{\mathbf{n}} the following lemma follows by induction.

Lemma 8.11.

Every element 𝐧k0\mathbf{n}\in\mathbb{Z}^{k}\setminus 0 defines a partially hyperbolic α𝐧:XΓXΓ\alpha^{\mathbf{n}}:X_{\Gamma}\to X_{\Gamma} where the center of α𝐧\alpha^{\mathbf{n}} has the same dimension as the center of α0𝐧\alpha_{0}^{\mathbf{n}}.

Remark 21.

Lemma 8.11 lets us pass the Weyl chamber containing 𝐧0\mathbf{n}_{0} and produce new partially hyperbolic elements in adjacent chambers.

8.5. Finishing the proof of Theorem 1.2

We finish the proof of Theorem 1.2 by showing that HH is a CC^{\infty} diffeomorphism. We begin by proving that hσh_{\sigma}, σ=s,u\sigma=s,u, are CC^{\infty}.

Lemma 8.12.

The maps hσ:XΓGσh_{\sigma}:X_{\Gamma}\to G^{\sigma}, σ=s,u\sigma=s,u, are CC^{\infty}.

Proof.

It suffices to consider the case σ=s\sigma=s. By Lemma 8.5 and Journé’s lemma [37] it suffices to show that hsh_{s} is uniformly smooth along WsW^{s}. Number all coarse exponents along WsW^{s} by [χ1],,[χN][\chi_{1}],...,[\chi_{N}]. The function hsh_{s} can be decomposed with respect to the map G[χ1]××G[χN]GsG^{[\chi_{1}]}\times...\times G^{[\chi_{N}]}\to G^{s} (see [55, Lemma 3.2]) as

(8.17) hs(x)=h[χ1](x)h[χN](x).\displaystyle h_{s}(x)=h_{[\chi_{1}]}(x)...h_{[\chi_{N}]}(x).

Using Lemmas 8.11 and 8.4 it is immediate that h[χN](x)h_{[\chi_{N}]}(x) is CC^{\infty}. Moreover, we use Lemmas 8.11 and 8.4 to show that h[χj](2)(x)=h[χj](x)/G(2)h_{[\chi_{j}]}^{(2)}(x)=h_{[\chi_{j}]}(x)/G^{(2)} is smooth for all j=1,,Nj=1,...,N. If we define h[χj](i)(x)=h[χj](x)/G(i)h_{[\chi_{j}]}^{(i)}(x)=h_{[\chi_{j}]}(x)/G^{(i)} and assume that h[χj](i1)(x)h_{[\chi_{j}]}^{(i-1)}(x) is smooth, then we can change the order of products in 8.17 modulo a polynomial in h[χj](i1)h_{[\chi_{j}]}^{(i-1)}. Applying Lemmas 8.11 and 8.4 once more, h[χj](i)h_{[\chi_{j}]}^{(i)} is CC^{\infty} for each j=1,,Nj=1,...,N. For ii large enough, using that GG is nilpotent, it follows that h[χj]h_{[\chi_{j}]} is CC^{\infty} along WsW^{s}. ∎

We are now ready to prove Theorem 1.2.

Proof of Theorem 1.2.

If we show that HH is CC^{\infty} then HH is automatically a diffeomorphism since the Jacobian can never vanish, this would contradict volume preservation of α\alpha (Lemma 8.2).

By Lemma 8.3 the map HH is uniformly CC^{\infty} along WcW^{c}. From equation 8.9 we have Hs,x(y)=hs(y)(h(x)gx(y)1)s1H_{s,x}(y)=h_{s}(y)\left(h(x)g_{x}(y)^{-1}\right)_{s}^{-1}. Since gx(y)g_{x}(y) and hs(y)h_{s}(y) are both uniformly CC^{\infty} along WsW^{s} byLemma 8.12, it follows that Hs,xH_{s,x} is uniformly CC^{\infty} along WsW^{s}, so HH is uniformly CC^{\infty} along WsW^{s}. Similarly HH is uniformly CC^{\infty} along WuW^{u}. By using Journé’s lemma along WcW^{c} and WsW^{s} it follows that HH is uniformly CC^{\infty} along WcsW^{cs}. Using Journé’s lemma once more along WuW^{u} and WcsW^{cs} it follows that HH is CC^{\infty}. ∎

9. Proof of main theorems

We are now ready to complete the proofs of Theorems A and B.

Proof of Theorem A.

By Theorem 7.1 the action α\alpha is bi-Hölder conjugated to some affine action α0\alpha_{0}. We produce the conjugacy HH for the special element f=α𝐧0f=\alpha^{\mathbf{n}_{0}}, but it conjugates the full action into an affine action. This is immediate from the construction of HH, but also follows from an argument as in [1]. Theorem A now follows from Theorem 1.2. ∎

Proof of Theorem B.

Let f0Aff(XΓ)f_{0}\in{\rm Aff}(X_{\Gamma}) and fDiff(XΓ)f\in{\rm Diff}^{\infty}(X_{\Gamma}) be C1C^{1}-close to f0f_{0}. Write Lsu:𝕋d𝕋dL_{su}:\mathbb{T}^{d}\to\mathbb{T}^{d} for the induced hyperbolic automorphism. Note that ff satisfy assumptions (i)(i) and (ii)(ii) of Theorem A since it is close to f0f_{0} (see also [46, Lemma A.2]). In particular, we obtain Φ\Phi from Theorem 1.1. Denote by Z(f)Z^{\infty}(f) the CC^{\infty}-centralizer of ff and let Zc(f)Z(f)Z_{c}^{\infty}(f)\subset Z^{\infty}(f) be the center fixing, normal subgroup of Z(f)Z^{\infty}(f) from Equation 1.4. Define the quotient

(9.1) Zsu(f)=Z(f)/Zc(f).\displaystyle Z_{su}^{\infty}(f)=Z^{\infty}(f)/Z_{c}^{\infty}(f).

If gZc(f)g\in Z_{c}^{\infty}(f) then Φ(gx)=Φ(x)\Phi(gx)=\Phi(x) so the induced map on H1XΓH_{1}X_{\Gamma} satisfy Φg=Φ\Phi_{*}g_{*}=\Phi_{*}. Since Φ:H1XΓH1𝕋d\Phi_{*}:H_{1}X_{\Gamma}\to H_{1}\mathbb{T}^{d} is an isomorphism it follows that g=idg_{*}={\rm id}. Conversely, if g=idg_{*}={\rm id} then Φ(gx)=Φ(x)\Phi(gx)=\Phi(x) (Theorem 1.1) so gZc(f)g\in Z_{c}^{\infty}(f). It follows that each non-trivial gZsu(f)g\in Z_{su}^{\infty}(f) represent an element of Z(f)Z^{\infty}(f) that project onto a non-trivial automorphism on 𝕋d\mathbb{T}^{d}. In particular, if rank(Zsu(f))>1{\rm rank}(Z_{su}^{\infty}(f))>1 then the image of Z(f)Z^{\infty}(f) in ZAut(Lsu)Z_{\rm Aut}(L_{su}) contains a subgroup isomorphic to 2\mathbb{Z}^{2}. Irreducibility of LsuL_{su} implies that this 2\mathbb{Z}^{2}-subgroup in Z(f)Z^{\infty}(f) is higher rank, so the action of Z(f)Z^{\infty}(f) is CC^{\infty}-conjugate to some affine action by Theorem A. If rank(Zsu(f))=1{\rm rank}(Z_{su}^{\infty}(f))=1 and #Zc(f)=\#Z_{c}^{\infty}(f)=\infty then [21, Corollary 18] implies case (ii)(ii) of Theorem B. Finally, if rank(Zsu(f))=1{\rm rank}(Z_{su}^{\infty}(f))=1 and #Zc(f)<\#Z_{c}^{\infty}(f)<\infty then Z(f)Z^{\infty}(f) is virtually \mathbb{Z} so case (i)(i) of Theorem B holds. ∎

Appendix A Some algebraic lemmas

In this appendix, we show some basic properties of higher rank algebraic actions on nilmanifolds. The first two lemmas, 2.1 and 2.2, are stated in Section 2.2.

Proof of Lemma 2.1.

If the conclusion does not hold then there is a decomposition 𝔤=E0sE0cE0u\mathfrak{g}=E_{0}^{s}\oplus E_{0}^{c}\oplus E_{0}^{u} so that every α0𝐧\alpha_{0}^{\mathbf{n}} is subexponential along E0cE_{0}^{c} and either contract or expand E0sE_{0}^{s} and E0uE_{0}^{u}. If the projected action on the base is rank1-1, then the whole action is rank1-1 (and has a rank1-1 factor). For any nn\in\mathbb{N} there is εn>0\varepsilon_{n}>0 such that if LGL(n,)L\in{\rm GL}(n,\mathbb{Z}) satisfies that the eigenvalues of LL with modulus larger than one have a product bounded by 1+εn1+\varepsilon_{n}, then LL have no eigenvalues of modulus larger than 11.

Let WkW\subset\mathbb{R}^{k} be the kernel of the unique pair of negatively proportional exponents of α0\alpha_{0}. If 𝐧k\mathbf{n}\in\mathbb{Z}^{k} is sufficiently close to WW then all eigenvalues of α0𝐧GL(n,)\alpha_{0}^{\mathbf{n}}\in{\rm GL}(n,\mathbb{Z}) will be close to the unit circle, which implies that all eigenvalues of α0𝐧\alpha_{0}^{\mathbf{n}} lie on the unit circle. It follows that any 𝐧k\mathbf{n}\in\mathbb{Z}^{k} sufficiently close to WW lies in WW. In particular

(A.1) rank(kW)=dim(W)=k1.\displaystyle{\rm rank}(\mathbb{Z}^{k}\cap W)=\dim(W)=k-1.

Elements in WW have all eigenvalues on the unit circle so after dropping to a finite index subgroup of k\mathbb{Z}^{k}, we may assume that all eigenvalues are 11. After taking a quotient to remove Jordan blocks the action of WW is trivial. Since dim(W)=k1\dim(W)=k-1 the action α0\alpha_{0} factor through a \mathbb{Z}-action. ∎

Proof of Lemma 2.2.

The translation action of VV is minimal if and only if the induced translation action on the base is minimal, so we assume without loss of generality that XΓX_{\Gamma} is a torus. Let WW be the rational span of VV, then VV acts minimally if and only if W=dW=\mathbb{R}^{d}. If WdW\neq\mathbb{R}^{d} then 𝕋𝕋d/W\mathbb{T}^{\ell}\cong\mathbb{T}^{d}/W with 1\ell\geq 1 and α0\alpha_{0} descend to 𝕋\mathbb{T}^{\ell}. We have quotiened out all coarse exponents except for one negatively proportional pair, so the factor 𝕋\mathbb{T}^{\ell} has only one pair of negatively proportional exponents. By Lemma 2.1 𝕋\mathbb{T}^{\ell} is a rank1-1 factor, which is a contradiction. We conclude that W=dW=\mathbb{R}^{d} so VV act minimally. ∎

We will need a lemma like Lemma 2.2, but only considering coarse directions that lie in the same stable direction for some element of the action. The following lemma is a consequence of Lemma 2.1.

Lemma A.1.

Let α0:kAff(XΓ)\alpha_{0}:\mathbb{Z}^{k}\to{\rm Aff}(X_{\Gamma}) be higher rank. We say that 𝐧k\mathbf{n}\in\mathbb{Z}^{k} is regular if the center of α0𝐧\alpha_{0}^{\mathbf{n}} coincide with the joint center of α0\alpha_{0}. Let 𝐧0\mathbf{n}_{0} be regular and E0sE_{0}^{s} be the stable space associated to α0𝐧0\alpha_{0}^{\mathbf{n}_{0}}. Let [χ][\chi] be a coarse exponent such that ker[χ]\ker[\chi] is a wall for the Weyl chamber that contains 𝐧0\mathbf{n}_{0} and E0[χ]E0sE_{0}^{[\chi]}\subset E_{0}^{s}. Either E0s=E0[χ]E_{0}^{s}=E_{0}^{[\chi]} or the complementary subspace E0ssE_{0}^{ss}, E0ssE0[χ]=E0sE_{0}^{ss}\oplus E_{0}^{[\chi]}=E_{0}^{s}, defines a minimal foliation in XΓX_{\Gamma}.

Proof.

After projecting to the base, we assume without loss of generality that XΓX_{\Gamma} is a torus 𝕋d\mathbb{T}^{d}. If E0[χ]=E0sE_{0}^{[\chi]}=E_{0}^{s} for [χ](𝐧0)<0[\chi](\mathbf{n}_{0})<0 then there is nothing to prove, so assume that there is at least one coarse exponent [η][\eta] satisfying [η][χ][\eta]\neq[\chi] and [η](𝐧0)<0[\eta](\mathbf{n}_{0})<0. Write E0ssE_{0}^{ss} for the complementary subspace defined by

(A.2) E0ss=[η](𝐧0)<0[η][χ]E0[η].\displaystyle E_{0}^{ss}=\bigoplus_{\begin{subarray}{c}[\eta](\mathbf{n}_{0})<0\\ [\eta]\neq[\chi]\end{subarray}}E_{0}^{[\eta]}.

Let WW be the rational closure of E0ssE_{0}^{ss}, then WW is α0\alpha_{0}-invariant and rational. Each α0𝐧|W\alpha_{0}^{\mathbf{n}}|_{W} preserves the lattice WdW\cap\mathbb{Z}^{d} so det(α0𝐧|W)=±1\det(\alpha_{0}^{\mathbf{n}}|_{W})=\pm 1. Given a coarse exponent [η][\eta], define η=r(η)η\eta^{\prime}=r(\eta^{\prime})\eta for [η]=[η][\eta^{\prime}]=[\eta] and d[η]W[0,1]d_{[\eta]}^{W}\in[0,1] by

(A.3) d[η]W=η[η]r(η)dim(E0ηW)η[η]r(η)dim(E0η).\displaystyle d_{[\eta]}^{W}=\frac{\sum_{\eta^{\prime}\in[\eta]}r(\eta^{\prime})\dim(E_{0}^{\eta^{\prime}}\cap W)}{\sum_{\eta^{\prime}\in[\eta]}r(\eta^{\prime})\dim(E_{0}^{\eta^{\prime}})}.

From this definition it is immediate that

(A.4) |det(α0𝐧|E0[η]W)|=|det(α0𝐧|E0[η])|d[η]W,𝐧k.\displaystyle\left|\det(\alpha_{0}^{\mathbf{n}}|_{E_{0}^{[\eta]}\cap W})\right|=\left|\det(\alpha_{0}^{\mathbf{n}}|_{E_{0}^{[\eta]}})\right|^{d_{[\eta]}^{W}},\quad\mathbf{n}\in\mathbb{Z}^{k}.

Our choice of WW implies that d[η]W=1d_{[\eta]}^{W}=1 if [η](𝐧0)<0[\eta](\mathbf{n}_{0})<0 and [η][χ][\eta]\neq[\chi]. Rewrite

1=\displaystyle 1= |det(α0𝐧|W)|=[η]|det(α0𝐧|E0[η])|d[η]W=\displaystyle\left|\det(\alpha_{0}^{\mathbf{n}}|_{W})\right|=\prod_{[\eta]}\left|\det(\alpha_{0}^{\mathbf{n}}|_{E_{0}^{[\eta]}})\right|^{d_{[\eta]}^{W}}=
([η](𝐧0)<0[η][χ]|det(α0𝐧|E0[η])|)|det(α0𝐧|E0[χ])|d[χ]W([η](𝐧0)>0|det(α0𝐧|E0[η])|d[η]W).\displaystyle\left(\prod_{\begin{subarray}{c}[\eta](\mathbf{n}_{0})<0\\ [\eta]\neq[\chi]\end{subarray}}|\det(\alpha_{0}^{\mathbf{n}}|_{E_{0}^{[\eta]}})|\right)\cdot\left|\det(\alpha_{0}^{\mathbf{n}}|_{E_{0}^{[\chi]}})\right|^{d_{[\chi]}^{W}}\cdot\left(\prod_{[\eta](\mathbf{n}_{0})>0}|\det(\alpha_{0}^{\mathbf{n}}|_{E_{0}^{[\eta]}})|^{d_{[\eta]}^{W}}\right).

Using |det(α0𝐧)|=1|\det(\alpha_{0}^{\mathbf{n}})|=1

[η](𝐧0)<0[η][χ]|det(α0𝐧|E0[η])|=1|det(α0𝐧|E0[χ])|[η](𝐧0)>01|det(α0𝐧|E0[η])|\displaystyle\prod_{\begin{subarray}{c}[\eta](\mathbf{n}_{0})<0\\ [\eta]\neq[\chi]\end{subarray}}|\det(\alpha_{0}^{\mathbf{n}}|_{E_{0}^{[\eta]}})|=\frac{1}{\left|\det(\alpha_{0}^{\mathbf{n}}|_{E_{0}^{[\chi]}})\right|}\cdot\prod_{[\eta](\mathbf{n}_{0})>0}\frac{1}{\left|\det(\alpha_{0}^{\mathbf{n}}|_{E_{0}^{[\eta]}})\right|}

and combining estimates

(A.5) 1=|det(α0𝐧|E0[χ])|d[χ]W1[η](𝐧0)>0|det(α0𝐧|E0[η])|d[η]W1.\displaystyle 1=\left|\det(\alpha_{0}^{\mathbf{n}}|_{E_{0}^{[\chi]}})\right|^{d_{[\chi]}^{W}-1}\cdot\prod_{[\eta](\mathbf{n}_{0})>0}|\det(\alpha_{0}^{\mathbf{n}}|_{E_{0}^{[\eta]}})|^{d_{[\eta]}^{W}-1}.

If 𝐧jk\mathbf{n}_{j}\in\mathbb{Z}^{k} is a sequence in the same Weyl chamber as 𝐧0\mathbf{n}_{0} such that 𝐧jker[χ]\mathbf{n}_{j}\to\ker[\chi], then

|det(α0𝐧j|E0[χ])|d[χ]W11.\displaystyle\left|\det(\alpha_{0}^{\mathbf{n}_{j}}|_{E_{0}^{[\chi]}})\right|^{d_{[\chi]}^{W}-1}\to 1.

Each [η]±[χ][\eta]\neq\pm[\chi] satisfies |det(α0𝐧j|E0[η])|μ>1|\det(\alpha_{0}^{\mathbf{n}_{j}}|_{E_{0}^{[\eta]}})|\geq\mu>1 for some uniform μ\mu. Letting jj\to\infty, d[η]W=1d_{[\eta]}^{W}=1 for [η](𝐧0)>0[\eta](\mathbf{n}_{0})>0, [η][χ][\eta]\neq-[\chi]. It follows that E0[η]WE_{0}^{[\eta]}\subset W for [η]±[χ][\eta]\neq\pm[\chi]. If 𝕋d/W\mathbb{T}^{d}/W is non-trivial then the projected action on 𝕋d/W\mathbb{T}^{d}/W does not have two independent coarse Lyapunov exponents. By Lemma 2.1 𝕋d/W\mathbb{T}^{d}/W is a rank1-1 factor of α0\alpha_{0}. This contradicts the assumption that α0\alpha_{0} is higher rank so W=dW=\mathbb{R}^{d} which proves the lemma. ∎

Lemma A.2.

Let AGL(n,)A\subset{\rm GL}(n,\mathbb{Z}) be a free abelian subgroup with Lyapunov exponents Lyap(A){\rm Lyap}(A). Let NN be the maximal number of linearly independent Lyapunov exponents (so N=dim(span(Lyap(A)))N=\dim({\rm span}({\rm Lyap}(A)))). If the intersection of the kernels of all χLyap(A)\chi\in{\rm Lyap}(A) is trivial in AA then rank(A)N{\rm rank}(A)\leq N.

Proof.

Let p(t)[t]p(t)\in\mathbb{Z}[t] be of degree dd, monic and with constant term ±1\pm 1. Let λ1,,λd\lambda_{1},...,\lambda_{d} be the roots of p(t)p(t) (possibly with multiplicity). There is a constant μd>1\mu_{d}>1 such that either p(t)p(t) has only roots on the unit circle or

(A.6) M(p(t))=j=1dmax(1,|λj|)μd\displaystyle M(p(t))=\prod_{j=1}^{d}\max(1,|\lambda_{j}|)\geq\mu_{d}

see for example [22]. We number {χ1,,χn}=Lyap(A)\{\chi_{1},...,\chi_{n}\}={\rm Lyap}(A). Let djd_{j} be the dimension of the Lyapunov space associated to χj\chi_{j}. Given aAa\in A let pa(t)p_{a}(t) be the corresponding characteristic polynomial. We obtain

(A.7) logM(pa(t))=χj(a)>0djχj(a).\displaystyle\log M(p_{a}(t))=\sum_{\chi_{j}(a)>0}d_{j}\chi_{j}(a).

Suppose that rank(A)>N{\rm rank}(A)>N for contradiction. Let χ1,,χN\chi_{1},...,\chi_{N} be chosen such that every χj\chi_{j} lie in span(χ1,,χN){\rm span}(\chi_{1},...,\chi_{N}). If anAa_{n}\in A is such that χ1(an),,χN(an)0\chi_{1}(a_{n}),...,\chi_{N}(a_{n})\to 0 then χj(an)0\chi_{j}(a_{n})\to 0 for all j=1,,nj=1,...,n. The intersection

(A.8) V:=j=1NkerχjA\displaystyle V:=\bigcap_{j=1}^{N}\ker\chi_{j}\subset A\otimes\mathbb{R}

is non-trivial since rank(A)>N{\rm rank}(A)>N. The set AA is a lattice in AA\otimes\mathbb{R} so we find a sequence anAa_{n}\in A such that d(an,V)0{\rm d}(a_{n},V)\to 0 as nn\to\infty but an↛ea_{n}\not\to e (where ee is the identity in AGL(n,)A\subset{\rm GL}(n,\mathbb{Z})). It follows that

(A.9) logM(pan(t))=χj(a)>0djχj(an)\displaystyle\log M(p_{a_{n}}(t))=\sum_{\chi_{j}(a)>0}d_{j}\chi_{j}(a_{n})

tends to 0 as nn\to\infty. With nn big enough logM(pan(t))<logμd\log M(p_{a_{n}}(t))<\log\mu_{d} which implies that panp_{a_{n}} has all roots on the unit circle. It follows that ana_{n} lie in the kernel of all χj\chi_{j}, so an=ea_{n}=e by assumption. This implies that anea_{n}\to e which is a contradiction. ∎

Lemma A.3.

Let ASp(d,)A\in{\rm Sp}(d,\mathbb{Z}) be hyperbolic with irreducible characteristic polynomial. If r1(A)r_{1}(A) denotes the number of real eigenvalues of AA and r2(A)r_{2}(A) the number of pairs of complex eigenvalues of AA then rank(ZSp(d,)(A))=r1(A)/2+r2(A)/2{\rm rank}(Z_{{\rm Sp}(d,\mathbb{Z})}(A))=r_{1}(A)/2+r_{2}(A)/2.

Proof.

Fix ASp(d,)A\in{\rm Sp}(d,\mathbb{Z}) with irreducible characteristic polynomial. Let

(A.10) Λ2(d)=dd\displaystyle\Lambda_{2}(\mathbb{R}^{d})=\mathbb{R}^{d}\wedge\mathbb{R}^{d}

be the vector space of 22-vectors. Let AA:Λ2(d)Λ2(d)A\wedge A:\Lambda_{2}(\mathbb{R}^{d})\to\Lambda_{2}(\mathbb{R}^{d}) be the induced map on 22-vectors. Note that Λ2(d)\Lambda_{2}(\mathbb{Z}^{d}) is a AAA\wedge A-invariant lattice in Λ2(d)\Lambda_{2}(\mathbb{R}^{d}). Denote by WΛ2(d)W\leq\Lambda_{2}(\mathbb{R}^{d}) the eigenspace of 11 for AAA\wedge A. Write Γ:=WΛ2(d)\Gamma:=W\cap\Lambda_{2}(\mathbb{Z}^{d}), since WW is a rational subspace of Λ2(d)\Lambda_{2}(\mathbb{R}^{d}) the subgroup ΓW\Gamma\leq W is a lattice.

Given BZGL(d,)(A)B\in Z_{{\rm GL}(d,\mathbb{Z})}(A) the wedge BBB\wedge B preserves WW and stabilize Γ\Gamma in WW. So, after identifying Wd/2W\cong\mathbb{R}^{d/2} and Γd/2\Gamma\cong\mathbb{Z}^{d/2} we obtain a map Ψ:ZGL(d,)(A)GL(d/2,)\Psi:Z_{{\rm GL}(d,\mathbb{Z})}(A)\to{\rm GL}(d/2,\mathbb{Z}) defined by Ψ(B):=(BB)|W\Psi(B):=(B\wedge B)|_{W}. It is immediate that Ψ\Psi is a homomorphism. Fix eigenvectors e1,,ed/2,e~1,,e~d/2de_{1},...,e_{d/2},\tilde{e}_{1},...,\tilde{e}_{d/2}\in\mathbb{C}^{d} such that Aej=λjejAe_{j}=\lambda_{j}e_{j} and Ae~j=e~j/λjA\tilde{e}_{j}=\tilde{e}_{j}/\lambda_{j}. We identify (the complexification of) WW by W=span(e1e~1,,ed/2e~d/2)W={\rm span}(e_{1}\wedge\tilde{e}_{1},...,e_{d/2}\wedge\tilde{e}_{d/2}) so for BZGL(d,)(A)B\in Z_{{\rm GL}(d,\mathbb{Z})}(A) we have

(A.11) BB(eje~j)=μj(B)μ~j(B)eje~j\displaystyle B\wedge B(e_{j}\wedge\tilde{e}_{j})=\mu_{j}(B)\tilde{\mu}_{j}(B)e_{j}\wedge\tilde{e}_{j}

where Bej=μj(B)ejBe_{j}=\mu_{j}(B)e_{j} and Be~j=μ~j(B)e~jB\tilde{e}_{j}=\tilde{\mu}_{j}(B)\tilde{e}_{j}. If Ψ(B)=e\Psi(B)=e then μj(B)μ~j(B)=1\mu_{j}(B)\tilde{\mu}_{j}(B)=1 which implies that BB preserve the symplectic form that AA preserve (note that e1,,ed/2,e~1,,e~d/2e_{1},...,e_{d/2},\tilde{e}_{1},...,\tilde{e}_{d/2} can be chosen such that the symplectic form can be written e1e~1++ed/2e~d/2e^{1}\wedge\tilde{e}^{1}+...+e^{d/2}\wedge\tilde{e}^{d/2}). Conversely, if BZSp(d,)(A)B\in Z_{{\rm Sp}(d,\mathbb{Z})}(A) then Ψ(B)=e\Psi(B)=e. Equation A.11 also implies that the Lyapunov exponents of Im(Ψ){\rm Im}(\Psi) are given by

(A.12) χjΨ(Ψ(B))=log|μj(B)|+log|μ~j(B)|.\displaystyle\chi_{j}^{\Psi}(\Psi(B))=\log|\mu_{j}(B)|+\log|\tilde{\mu}_{j}(B)|.

It follows that Im(Ψ){\rm Im}(\Psi) has r1(A)/2+r2(A)/2r_{1}(A)/2+r_{2}(A)/2 Lyapunov exponents. Indeed, if μj(B),μ~j(B)\mu_{j}(B),\tilde{\mu}_{j}(B) takes values in \mathbb{R} then log|μj(B)|+log|μ~j(B)|\log|\mu_{j}(B)|+\log|\tilde{\mu}_{j}(B)| defines one Lyapunov exponent. If μj(B),μ~j(B)\mu_{j}(B),\tilde{\mu}_{j}(B) takes values in \mathbb{C}\setminus\mathbb{R} then μj(B)¯=μj(B)\overline{\mu_{j}(B)}=\mu_{j^{\prime}}(B) and μ~j(B)¯=μ~j(B)\overline{\tilde{\mu}_{j}(B)}=\tilde{\mu}_{j^{\prime}}(B) for some jjj\neq j^{\prime}, so

(A.13) χjΨ(B)=χjΨ(B).\displaystyle\chi_{j}^{\Psi}(B)=\chi_{j^{\prime}}^{\Psi}(B).

By Lemma A.2 rank(Im(Ψ))r1(A)/2+r2(A)/21{\rm rank}({\rm Im}(\Psi))\leq r_{1}(A)/2+r_{2}(A)/2-1 since |det(Ψ(B))|=1|\det(\Psi(B))|=1 implies that, at least, one Lyapunov exponent of Im(Ψ){\rm Im}(\Psi) can be written as a combination of the other exponents. By [42] the rank of ZGL(d,)(A)Z_{{\rm GL}(d,\mathbb{Z})}(A) is r1(A)+r2(A)1r_{1}(A)+r_{2}(A)-1, so

rank(ker(Ψ))=\displaystyle{\rm rank}(\ker(\Psi))= rank(ZGL(d,)(A))rank(Im(Ψ))\displaystyle{\rm rank}(Z_{{\rm GL}(d,\mathbb{Z})}(A))-{\rm rank}({\rm Im}(\Psi))\geq
r1(A)+r2(A)1[r1(A)+r2(A)21]=\displaystyle r_{1}(A)+r_{2}(A)-1-\left[\frac{r_{1}(A)+r_{2}(A)}{2}-1\right]=
r1(A)+r2(A)2\displaystyle\frac{r_{1}(A)+r_{2}(A)}{2}

or since ker(Ψ)=ZSp(d,)(A)\ker(\Psi)=Z_{{\rm Sp}(d,\mathbb{Z})}(A), rank(ZSp(d,)(A))(r1(A)+r2(A))/2{\rm rank}(Z_{{\rm Sp}(d,\mathbb{Z})}(A))\geq(r_{1}(A)+r_{2}(A))/2. The converse inequality is clear from Lemma A.2 since if χ\chi is a Lyapunov exponent of ZSp(d,)(A)Z_{\rm Sp(d,\mathbb{Z})}(A) then so is χ-\chi. ∎

References

  • [1] R. L. Adler and R. Palais. Homeomorphic conjugacy of automorphisms on the torus. Proceedings of the American Mathematical Society, 16(6):1222–25, 1965.
  • [2] A. Avila, J. Santamaria, and M. Viana. Cocycles over partially hyperbolic maps. Asterisque, 01 2013.
  • [3] A. Avila and M. Viana. Extremal lyapunov exponents: an invariance principle and applications. Inventiones mathematicae, 181:115–178, 2010.
  • [4] A. Avila, M. Viana, and A. Wilkinson. Absolute continuity, lyapunov exponents, and rigidity ii: systems with compact center leaves. Ergodic Theory and Dynamical Systems, 42(2):437–490, 2022.
  • [5] T. Barthelmé and A. Gogolev. Centralizers of partially hyperbolic diffeomorphisms in dimension 3. Discrete and Continuous Dynamical Systems, 41(9):4477–4484, 2021.
  • [6] C. Bonatti, S. Crovisier, and A. Wilkinson. The C1C^{1} generic diffeomorphism has trivial centralizer. Publications mathématiques, 109:185–244, 2009.
  • [7] M. Brin. On dynamical coherence. Ergodic Theory and Dynamical Systems, 23(2):395–401, 2003.
  • [8] M. Brin, D. Burago, and S. Ivanov. Dynamical coherence of partially hyperbolic diffeomorphisms of the 3-torus. Journal of Modern Dynamics, 3(1):1–11, 2009.
  • [9] A. Brown. Smoothness of stable holonomies inside center-stable manifolds. Ergodic Theory and Dynamical Systems, 42(12):3593–3618, 2022.
  • [10] A. Brown, F. Rodriguez Hertz, and Z. Wang. Global smooth and topological rigidity of hyperbolic lattice actions. Annals of Mathematics, 186(3):913–72, 2017.
  • [11] L. Corwin and F. P. Greenleaf. Representations of Nilpotent Lie Groups and Their Applications: Volume 1, Part 1, Basic Theory and Examples. Cambridge University Press, 1990.
  • [12] D. Damjanović and Q. Chen. Rigidity properties for some isometric extensions of partially hyperbolic actions on the torus. Transactions of the American Mathematical Society, 376(6), 2022.
  • [13] D. Damjanović, Q. Chen, and B. Petković. On simultaneous linearization of certain commuting nearly integrable diffeomorphisms of the cylinder. Transactions of the American Mathematical Society, 301:1881–1912, 2022.
  • [14] D. Damjanović and B. Fayad. On local rigidity of partially hyperbolic affine k\mathbb{Z}^{k} actions. Reine Angew. Math., 751:1–26, 2019.
  • [15] D. Damjanović, B. Fayad, and M. Saprykina. Kam-rigidity for parabolic affine abelian actions. arXiv:2303.04367, 2023.
  • [16] D. Damjanović and A. Katok. Local rigidity of partially hyperbolic actions i. kam method and k\mathbb{Z}^{k} actions on the torus. Annals of Mathematics, 172(3):1805–58, 2010.
  • [17] D. Damjanović and A. Katok. Local rigidity of homogeneous parabolic actions: I. a model case. Journal of Modern Dynamics, 5(2):203–235, 2011.
  • [18] D. Damjanović and A. Katok. Local rigidity of partially hyperbolic actions. ii: The geometric method and restrictions of weyl chamber flows on sl(n,r)/γ{\rm sl}(n,r)/\gamma. International Mathematics Research Notices, 2011(19):4405–4430, 2011.
  • [19] D. Damjanović, A. Wilkinson, and D. Xu. Private communication.
  • [20] D. Damjanović, A. Wilkinsson, and D. Xu. Pathology and asymmetry: Centralizer rigidity for partially hyperbolic diffeomorphisms. Duke mathematical journal, 170(17):3815–3890, 2021.
  • [21] D. Damjanović, A. Wilkinsson, and D. Xu. Transitive centralizers and fibered partially hyperbolic systems. arXiv:2303.17739, 2023.
  • [22] E. Dobrowolski. On a question of lehmer and the number of irreducible factors of a polynomial. Acta arithmetica, 34(4):391–401, 1979.
  • [23] M. Einsiedler and T. Fisher. Differentiable rigidity for hyperbolic toral actions. Israel J. Math, 157:347–377, 2007.
  • [24] D. Fisher. Recent progress in the zimmer program. In Ergodic Theory, Geometry, and Topology. University of Chicago Press, Chicago, 2019.
  • [25] D. Fisher, B. Kalinin, and R. Spatzier. Totally nonsymplectic anosov actions on tori and nilmanifolds. Geometry & Topology, 15(1):191–216, 2011.
  • [26] D. Fisher, B. Kalinin, R. J. Spatzier, and with an appendix by J. Davis. Global rigidity of higher rank anosov actions on tori and nilmanifolds. J. Amer. Math. Soc., 26(1):167–198, 2013.
  • [27] J. Franks. Anosov diffeomorphisms on tori. Transactions of the American Mathematical Society, 145:117–24, 1969.
  • [28] S. Gan, Y. Shi, D. Xu, and J. Zhang. Centralizers of derived-from-anosov systems on 𝕋3\mathbb{T}^{3}: rigidity versus triviality. Ergodic Theory and Dynamical Systems, 42(9):2841–2865, 2022.
  • [29] A. M. Gleason. Groups without small subgroups. Annals of Mathematics, 56(2):193–212, 1952.
  • [30] A. Gorodnik. Open problems in dynamics and related fields. Journal of Modern Dynamics, 1:1–35, 2007.
  • [31] A. Gorodnik and R. J. Spatzier. Mixing properties of commuting nilmanifold automorphisms. Acta Math., 215(1):127–159, 2015.
  • [32] A. Hammerlindl. Dynamics of quasi-isometric foliations. Nonlinearity, 25(6):1585–1599, 2012.
  • [33] A. Hammerlindl. Leaf conjugacies on the torus. Ergodic Theory and Dynamical Systems, 33(3):896–933, 2013.
  • [34] A. Hammerlindl. Partial hyperbolicity on 3-dimensional nilmanifolds. Discrete and Continuous Dynamical Systems, 33(8):3641–3669, 2013.
  • [35] A. Hammerlindl and R. Potrie. Classification of partially hyperbolic diffeomorphisms in 3-manifolds with solvable fundamental group. Journal of Topology, 8:842–870, 2015.
  • [36] M. W. Hirsch, C. C. Pugh, and M. Shub. Invariant Manifolds. Springer, 1977.
  • [37] J. L. Journé. A regularity lemma for functions of several variables. Revista Matematica Iberoamericana, 4:187–193, 1988.
  • [38] B. Kalinin and V. Sadovskaya. Global rigidity for totally nonsymplectic Anosov k\mathbb{Z}^{k} actions. Geometry & Topology, 10(2):929 – 954, 2006.
  • [39] B. Kalinin and V. Sadovskaya. On the classification of resonance-free anosov k\mathbb{Z}^{k} actions. Michigan Math. J, 55(3):651–67, 2007.
  • [40] B. Kalinin and R. Spatzier. On the classification of cartan actions. Geom. Funct. Anal, 17(2):468–490, 2007.
  • [41] A. Katok and B. Hasselblatt. Introduction to the Modern Theory of Dynamical Systems. Cambridge University Press, 1995.
  • [42] A. Katok, S. Katok, and K. Schmidt. Rigidity of measurable structure for k\mathbb{Z}^{k}-actions by automorphisms of a torus. Comment. Math. Helv., 77:718–745, 2002.
  • [43] A. Katok and J. Lewis. Local rigidity for certain groups of toral automorphisms. Israel J. Math, 75:203–241, 1991.
  • [44] A. Katok and R. Spatzier. First cohomology of anosov actions of higher rank abelian groups and applications to rigidity. Inst. Hautes Etudes Sei. Puhl Math, pages 131–156, 1994.
  • [45] F. Ledrappier and J.-M. Strelcyn. A proof of the estimation from below in pesin’s entropy formula. Ergodic Theory and Dynamical Systems, 2(2):203–219, 1982.
  • [46] H. Lee and S. Sandfeldt. Partially hyperbolic lattice actions on 2-step nilmanifolds. arXiv:2410.00784, 2024.
  • [47] A. Manning. There are no new anosov diffeomorphisms on tori. American Journal of Mathematics, 96(3):422–29, 1974.
  • [48] R. Mañé. Quasi-anosov diffeomorphisms and hyperbolic manifolds. Transactions of the American Mathematical Society, 229:351–370, 1977.
  • [49] D. Montgomery and L. Zippin. Small subgroups of finite-dimensional groups. Annals of Mathematics, 56(2):213–241, 1952.
  • [50] C. Pugh, M. Shub, and A. Wilkinson. Hölder foliations. Duke Mathematical Journal, 86(3):517 – 546, 1997.
  • [51] M. S. Raghunathan. Discrete Subgroups of Lie Groups. Springer Berlin, Heidelberg, 1972.
  • [52] F. Rodriguez Hertz. Stable ergodicity of certain linear automorphisms of the torus. Annals of Mathematics, 162:65–107, 2005.
  • [53] F. Rodriguez Hertz. Global rigidity of certain abelian actions by toral automorphisms. Journal of Modern Dynamics, 1(3):425–442, 2007.
  • [54] F. Rodriguez Hertz, M. Rodriguez Hertz, and R. Ures. Accessibility and stable ergodicity for partially hyperbolic diffeomorphisms with 1d-center bundle. Inventiones mathematicae, 172:353–381, 2006.
  • [55] F. Rodriguez Hertz and Z. Wang. Global rigidity of higher rank abelian anosov algebraic actions. Invent. math., 198:165–209, 2014.
  • [56] F. Rodríguez Hertz, M. Rodríguez Hertz, A. Thazibi, and R. Ures. Maximizing measures for partially hyperbolic systems with compact center leaves. Ergodic Theory and Dynamical Systems, 32(2):825–839, 2012.
  • [57] S. Sandfeldt. Centralizer classification and rigidity for some partially hyperbolic toral automorphisms. Journal of Modern Dynamics, 20:479–523, 2024.
  • [58] S. J. Schreiber. On growth rates of subadditive functions for semiflows. Journal of Differential Equations, 148(2):334–350, 1998.
  • [59] S. Smale. Dynamics retrospective: great problems, attempts that failed. Physica D: Nonlinear Phenomena, 51(1–3):267–273, 1991.
  • [60] S. Smale. Mathematical problems for the next century. The Mathematical Intelligencer, 20:7–15, 1998.
  • [61] R. Spatzier and K. Vinhage. Cartan actions of higher rank abelian groups and their classification. Journal of the American Mathematical Society, 01 2019.
  • [62] K. Vinhage. On the rigidity of weyl chamber flows and schur multipliers as topological groups. Journal of Modern Dynamics, 9(1):25–49, 2015.
  • [63] K. Vinhage and Z. Wang. Local rigidity of higher rank homogeneous abelian actions: a complete solution via the geometric method. Geom Dedicata, 200:385–439, 2019.
  • [64] Z. Wang. Multi-invariant measures and subsets on nilmanifolds. Journal d’Analyse Mathématique, 135:123–183, 2018.
  • [65] Z. Wang. Local rigidity of parabolic algebraic actions. arXiv:1908.10496, 2019.
  • [66] Z. Wang. Local rigidity of higher rank partially hyperbolic algebraic actions. arXiv:1103.3077, 2022.
  • [67] Z. J. Wang. Local rigidity of partially hyperbolic actions. Journal of Modern Dynamics, 4(2):271– 327, 2010.
  • [68] Z. J. Wang. New cases of differentiable rigidity for partially hyperbolic actions: symplectic groups and resonance directions. Journal of Modern Dynamics, 4(4):585–608, 2010.
  • [69] Z. W. Wang. Centralizer rigidity near elements of the weyl chamber flow. arXiv:2309.07282, 2023.
  • [70] A. Wilkinson. The cohomological equation for partially hyperbolic diffeomorphisms. Astérisque, (358):75–165, 2013.