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

Rigidity of stable Lyapunov exponents and integrability for Anosov maps

Jinpeng An,   Shaobo Gan,   Ruihao Gu,   Yi Shi
Abstract

Let ff be a non-invertible irreducible Anosov map on dd-torus. We show that if the stable bundle of ff is one-dimensional, then ff has the integrable unstable bundle, if and only if, every periodic point of ff admits the same Lyapunov exponent on the stable bundle with its linearization. For higher-dimensional stable bundle case, we get the same result on the assumption that ff is a C1C^{1}-perturbation of a linear Anosov map with real simple Lyapunov spectrum on the stable bundle. In both cases, this implies if ff is topologically conjugate to its linearization, then the conjugacy is smooth on the stable bundle.

1 Introduction

Let MM be a dd-dimensional smooth closed Riemannian manifold. A diffeomorphism f:MMf:M\to M is Anosov if there exists a continuous DfDf-invariant splitting TM=EsEuTM=E^{s}\oplus E^{u} such that DfDf is uniformly contracting in EsE^{s} and DfDf is uniformly expanding in EuE^{u}. The classical Stable Manifold Theorem (e.g. [32]) shows that both EsE^{s} and EuE^{u} are uniquely integrable. So there are ff-invariant stable and unstable foliations tangent to EsE^{s} and EuE^{u} respectively.

The most well-known example of Anosov diffeomorphisms is a linear automorphism AGLd()A\in{\rm GL}_{d}(\mathbb{Z}) with all eigenvalues whose absolute values are not equal to 11. The induced diffeomorphism A:𝕋d𝕋dA:\mathbb{T}^{d}\to\mathbb{T}^{d} is Anosov. All known Anosov diffeomorphisms are conjugate to affine automorphisms of infra-nilmanifolds. In particular, every Anosov diffeomorphism with dimEs=1{\rm dim}E^{s}=1 or dimEu=1{\rm dim}E^{u}=1 must be supported on 𝕋d\mathbb{T}^{d} [9], and every Anosov diffeomorphism f:𝕋d𝕋df:\mathbb{T}^{d}\to\mathbb{T}^{d} is topologically conjugate to its linearization f:π1(𝕋d)π1(𝕋d)f_{*}:\pi_{1}(\mathbb{T}^{d})\to\pi_{1}(\mathbb{T}^{d}) acting on 𝕋d\mathbb{T}^{d}[8, 26].

In 1974, Mañé and Pugh extended the concept of Anosov diffeomorphisms to non-invertible Anosov maps.

Definition 1.1 ([25]).

A C1C^{1} local diffeomorphism f:MMf:M\to M is called Anosov map, if there exists a DfDf-invariant continuous subbundle EsTME^{s}\subset TM such that it is uniformly DfDf-contracting and its quotient bundle TM/EsTM/E^{s} is unformly DfDf-expanding.

The set of Anosov maps on MM is C1C^{1}-open in the space Cr(M)C^{r}(M) which consists of all CrC^{r}-maps of MM with r1r\geq 1. All known Anosov maps are conjugate to affine endomorphisms of infra-nilmanifolds.

Differing from Anosov diffeomorphisms, there is a priori no DfDf-expanding subbundle EuTME^{u}\subset TM for a non-invertible Anosov map because the negative orbit for a point is not unique. For instance, Przytycki [33] constructed a class of Anosov maps on torus which has infinitely many expanding directions on certain points. In fact, the set of expanding directions on a certain point in Przytycki’s example contains a curve homeomorphic to interval in the (dimMdimEs)({\rm dim}M-{\rm dim}E^{s})-Grassman space [33, Theorem 2.15]. In the same paper [33], Przytycki defined Anosov maps in the way of orbit space (see Definition 2.11) which allowed us to define the unstable bundle along every orbit. However, these unstable bundles are not integrable in general when project on the manifold MM.

We say an Anosov map ff has an integrable unstable bundle, if there exists a continuous DfDf-invariant splitting TM=EsEuTM=E^{s}\oplus E^{u}, such that DfDf is uniformly contracting on EsE^{s} (stable bundle) and expanding on EuE^{u} (unstable bundle). Here EuE^{u} is uniquely integrable, see [33]. For example,

A0=[3111]:𝕋2𝕋2,A_{0}=\begin{bmatrix}3&1\\ 1&1\end{bmatrix}:\mathbb{T}^{2}\to\mathbb{T}^{2},

is an Anosov map on torus with integrable unstable bundle.

There are plenty of Anosov maps without integrable unstable bundles. Actually, Przytycki [33, Theorem 2.18] showed that any non-invertible Anosov map ff on any manifold MM with non-trivial stable bundle can be C1C^{1}-approximated by Anosov maps without integrable unstable bundles. Moreover, for every transitive Anosov map without integrable unstable bundle, it must have a residual set in the manifold in which every point has infinitely many expanding directions ([29]).

In this paper, we give an equivalent characterization for a class of Anosov maps on dd-torus 𝕋d\mathbb{T}^{d} which has integrable unstable bundle.

Let f:𝕋d𝕋df:\mathbb{T}^{d}\to\mathbb{T}^{d} be an Anosov map on torus, then ff is homotopic to a linear toral map A=f:π1(𝕋d)π1(𝕋d)A=f_{*}:\pi_{1}(\mathbb{T}^{d})\to\pi_{1}(\mathbb{T}^{d}). Here AA is also an Anosov map [1, Theorem 8.1.1] and is called the linearization of ff. A toral Anosov map ff is called irreducible, if its linearization AGLd()Md()A\in GL_{d}(\mathbb{R})\cap M_{d}(\mathbb{Z}) has irreducible characteristic polynomial over \mathbb{Q}.

Theorem 1.1.

Let f:𝕋d𝕋df:\mathbb{T}^{d}\to\mathbb{T}^{d} be a C1+αC^{1+\alpha} irreducible non-invertible Anosov map with one-dimensional stable bundle. Then ff has integrable unstable bundle, if and only if, every periodic point of ff admits the same Lyapunov exponent on the stable bundle.

Remark 1.2.

In both cases, the Lyapunov exponent of ff on the stable bundle is equal to its linearization ff_{*} on the stable bundle. Moreover, the necessity only need C1C^{1} regularity of ff: if a C1C^{1} irreducible non-invertible Anosov map f:𝕋d𝕋df:\mathbb{T}^{d}\to\mathbb{T}^{d} with dimEs=1{\rm dim}E^{s}=1 has integrable unstable bundle, then every periodic point of ff admits the same stable Lyapunov exponent to its linearization ff_{*}.

In fact, an Anosov map on torus is conjugate to its linearization if and only if it admits an integrable unstable bundle [30]. A direct corollary is the following, which is an interesting example of rigidity in smooth dynamics, in the sense of “weak equivalence”(topological conjugacy) implies “strong equivalence”(smooth conjugacy).

Corollary 1.3.

Let f:𝕋d𝕋df:\mathbb{T}^{d}\to\mathbb{T}^{d} be a C1+αC^{1+\alpha} irreducible non-invertible Anosov map with one-dimensional stable bundle. If ff is topologically conjugate to its linearization f:𝕋d𝕋df_{*}:\mathbb{T}^{d}\to\mathbb{T}^{d}, then the conjugacy is C1+αC^{1+\alpha}-smooth along stable foliation.

In particular, we have the following corollary on two torus 𝕋2\mathbb{T}^{2}.

Corollary 1.4.

Let f:𝕋2𝕋2f:\mathbb{T}^{2}\to\mathbb{T}^{2} be a non-invertible Anosov map, then the following are equivalent:

  • ff has integrable unstable bundle;

  • ff is topologically conjugate to its linearization f:𝕋2𝕋2f_{*}:\mathbb{T}^{2}\to\mathbb{T}^{2}.

Both of them imply the conjugacy between ff and ff_{*} is C1+αC^{1+\alpha}-smooth along the stable foliation.

Remark 1.5.

Recently, Micena [28, Theorem 1.10] shows that for a CC^{\infty} non-invertible Anosov map f:𝕋2𝕋2f:\mathbb{T}^{2}\to\mathbb{T}^{2} with integrable unstable bundle, if it admits periodic data on the stable and unstable bundle, then ff is CC^{\infty}-conjugate to ff_{*} (also see [27, Theorem C]). Our result shows that we only need to assume ff admits periodic data on the unstable bundle, then it is CC^{\infty}-conjugate to ff_{*}.

For higher-dimensional stable bundle case, we prove a local rigidity result for linear Anosov maps on 𝕋d\mathbb{T}^{d} with real simple spectrum. We say a hyperbolic matrix AMd()GLd()A\in M_{d}(\mathbb{Z})\cap GL_{d}(\mathbb{R}) has real simple spectrum on stable bundle, if all eigenvalues on the stable bundle are real and have mutually distinct moduli. Then AA admits a dominated splitting

T𝕋d=L1sLksLu,T\mathbb{T}^{d}=L^{s}_{1}\oplus\cdots\oplus L^{s}_{k}\oplus L^{u},

with dimLis=1{\rm dim}L^{s}_{i}=1 for i=1,,ki=1,\cdots,k.

If a map f:𝕋d𝕋df:\mathbb{T}^{d}\to\mathbb{T}^{d} is C1C^{1}-close to a hyperbolic AMd()GLd()A\in M_{d}(\mathbb{Z})\cap GL_{d}(\mathbb{R}) with real simple spectrum on stable bundle, then ff is Anosov and has kk-Lyapunov exponents on the stable bundle:

λ1s(p,f)<λ2s(p,f)<<λks(p,f)<0,pPer(f).\lambda^{s}_{1}(p,f)<\lambda^{s}_{2}(p,f)<\cdots\cdots<\lambda^{s}_{k}(p,f)<0,\qquad\forall p\in{\rm Per}(f).

We say ff has spectral rigidity on stable bundle if for every periodic point pPer(f)p\in{\rm Per}(f), it satisfies

λis(p,f)=log|μi|,i=1,,k.\lambda^{s}_{i}(p,f)=\log|\mu_{i}|,\qquad i=1,\cdots,k.

Here μi\mu_{i} is the eigenvalue of AA in the eigenspace LisL^{s}_{i}.

Theorem 1.2.

Let AMd()GLd()A\in M_{d}(\mathbb{Z})\cap GL_{d}(\mathbb{R}) be hyperbolic and irreducible with real simple spectrum on stable bundle. If A:𝕋d𝕋dA:\mathbb{T}^{d}\to\mathbb{T}^{d} is non-invertible, then for every fC1+α(𝕋d)f\in C^{1+\alpha}(\mathbb{T}^{d}) which is C1C^{1}-close to AA, ff has integrable unstable bundle, if and only if, it has spectral rigidity on stable bundle.

Since ff having integrable unstable bundle is equivalent to ff being topologically conjugate to AA, we have the following corollary.

Corollary 1.6.

Let AMd()GLd()A\in M_{d}(\mathbb{Z})\cap GL_{d}(\mathbb{R}) be hyperbolic and irreducible with real simple spectrum on stable bundle. Assume A:𝕋d𝕋dA:\mathbb{T}^{d}\to\mathbb{T}^{d} is non-invertible and fC1+α(𝕋d)f\in C^{1+\alpha}(\mathbb{T}^{d}) is C1C^{1}-close to AA. If ff is topologically conjugate to AA, then the conjugacy is C1+βC^{1+\beta}-smooth along stable foliation, for some 0<βα0<\beta\leq\alpha.

Remark 1.7.

Here we lose the regularity of conjugacy because the weak stable bundle may only be CβC^{\beta} continuous for some 0<βα0<\beta\leq\alpha.

We mention that the irreducible condition is necessary for our result. Indeed,

A1=[210110002]:𝕋3𝕋3,A_{1}=\begin{bmatrix}2&1&0\\ 1&1&0\\ 0&0&2\end{bmatrix}:\mathbb{T}^{3}\to\mathbb{T}^{3},

is a non-invertible Anosov map with one-dimensional stable bundle and integrable unstable bundle, and it can be treated as a product system on 𝕋2×S1\mathbb{T}^{2}\times S^{1}. Note that one of its factor systems

A2=[2111]:𝕋2𝕋2,A_{2}=\begin{bmatrix}2&1\\ 1&1\end{bmatrix}:\mathbb{T}^{2}\to\mathbb{T}^{2},

is an Anosov diffeomorphism. So, we can make a smooth perturbation f:𝕋2𝕋2f:\mathbb{T}^{2}\to\mathbb{T}^{2} of A2A_{2} which is not smooth conjugate to A2A_{2}. Then the product map

F=[f002]:𝕋3𝕋3,F=\begin{bmatrix}f&0\\ 0&2\end{bmatrix}:\mathbb{T}^{3}\to\mathbb{T}^{3},

is still an Anosov map with integrable unstable bundle, but it loses the rigidity of Lyapunov exponents on stable bundle.

We would like to give another view of Theorem 1.1 and Theorem 1.2. In [14], Gogolev and Guysinsky show that for an Anosov diffeomorphism g:𝕋3𝕋3g:\mathbb{T}^{3}\to\mathbb{T}^{3} with partially hyperbolic splitting T𝕋3=EssEwsEuT\mathbb{T}^{3}=E^{ss}\oplus E^{ws}\oplus E^{u}, if gg has spectral rigidity on the weak stable bundle EwsE^{ws}, then EssEuE^{ss}\oplus E^{u} is integrable. See [12, 15] for higher-dimensional Anosov diffeomorphisms on 𝕋d\mathbb{T}^{d}. Conversely, Gan and Shi [11] proved that if EssEuE^{ss}\oplus E^{u} is integrable, then gg has spectral rigidity on EwsE^{ws}. See [17] for higher-dimensional Anosov diffeomorphisms on 𝕋d\mathbb{T}^{d}.

For a non-invertible Anosov map f:𝕋d𝕋df:\mathbb{T}^{d}\to\mathbb{T}^{d} and every x𝕋dx\in\mathbb{T}^{d}, we can see the preimage set of xx:

Perimage(x)=k0Pk(x)wherePk(x)={z𝕋d|fk(z)=fk(x)}{\rm Perimage}(x)=\bigcup_{k\geq 0}{\rm P}_{k}(x)\qquad\text{where}\qquad{\rm P}_{k}(x)=\{z\in\mathbb{T}^{d}|f^{k}(z)=f^{k}(x)\}

as the strongest stable manifold of xx. So the stable bundle of ff is corresponding to the weak stable bundle EwsE^{ws} of the Anosov diffeomorphism gg. It is clear that ff has integrable unstable bundle implies for every x𝕋dx\in\mathbb{T}^{d}, the unstable bundle of xx is independent of the choice of negative orbits of xx. So we can see the unstable bundle is jointly integrable with the strongest stable bundle, which is corresponding to the case that EssEuE^{ss}\oplus E^{u} is integrable for the Anosov diffeomorphism gg. Thus we can expect the Anosov map has some rigidity on the stable bundle.

The regularity of conjugacy for Anosov diffeomorphisms under the assumption of rigidity for Lyapunov exponents of periodic points has been extensively studied by many researchers e.g. [24, 12, 15, 13]. Recently, there are elegant works about the smooth conjugacy for conservative Anosov diffeomorphisms under the assumption of rigidity for Lyapunov exponents with respect to Lebesgue measures e.g. [36, 16]. Our Corollary 1.3 and Corollary 1.6 show that we only need to assume spectral rigidity on the unstable bundle to get smooth conjugacy for non-invertible Anosov map with integrable unstable bundles, see Remark 1.5 and [28].

Finally, we would like to mention that Anosov maps on 𝕋2\mathbb{T}^{2} are a special class of partially hyperbolic maps on surfaces. A series of impressive results on SRB measures and statistical properties for partially hyperbolic maps on surfaces have been obtained, see [37, 6, 7]. Meanwhile, the classification of partially hyperbolic endomorphisms on 𝕋2\mathbb{T}^{2} up to leaf conjugacy has also been studied, see [20, 18, 19]. It will be interesting to classify all partially hyperbolic maps with integrable unstable bundle on surfaces.

Organization of this paper: In section 2, we recall some general properties of Anosov maps and give some useful properties on the assumptions of Theorem 1.1 and Theorem 1.2. In section 3, we prove the ”necessary” parts of Theorem 1.1 and Theorem 1.2 with C1C^{1} regularity, which state that the existence of integrable unstable bundle implies the spectral rigidity on stable bundle. In section 4, on the assumption of spectral rigidity on stable bundle, we endow an affine metric on each leaf of the lifting stable foliations, which will be useful for the proof of sufficient parts of our theorems. In section 5, we prove the ”sufficient” parts of Theorem 1.1 and Theorem 1.2, which state that the rigidity of periodic stable Lyapunov spectrums implies the existence of integrable unstable bundle and we also prove Corollary 1.3 and Corollary 1.6 in this section.

Acknowledgements: S. Gan was partially supported by NSFC (11831001, 12161141002) and National Key R&D Program of China (2020YFE0204200). Y. Shi was partially supported by National Key R&D Program of China (2021YFA1001900) and NSFC (12071007, 11831001, 12090015).

2 Preliminaries

For short, an Anosov map f:MMf:M\to M is called special, if it has the integrable unstable bundle.

2.1 Global properties

For studying an Anosov map, one can lift it to the universal cover. In fact, Man~e´\rm\tilde{n}\acute{e} and Pugh proved the following proposition which allows us to observe the dynamics on the universal cover.

Proposition 2.1 ([25]).

Let M~\tilde{M} be the universal cover of MM and F:M~M~F:\tilde{M}\to\tilde{M} be a lift of f:MMf:M\to M. Then ff is an Anosov map if and only if FF is an Anosov diffeomorphism.

As usual, we define the stable manifolds of the Anosov diffeomorphism FF, denoted by ~s(x)\tilde{\mathcal{F}}^{s}(x),

~s(x):={yM~|d(Fk(y),Fk(x))0ask+},\displaystyle\tilde{\mathcal{F}}^{s}(x):=\big{\{}y\in\tilde{M}\;\big{|}\;d(F^{k}(y),F^{k}(x))\to 0\;\;{\rm as}\;\;k\to+\infty\big{\}}, (2.1)

for all xM~x\in\tilde{M}, and the unstable manifolds ~u(x)\tilde{\mathcal{F}}^{u}(x) by iterating backward. And we define the local (un)stable manifolds with size δ\delta, denoted by ~s/u(x,δ)\tilde{\mathcal{F}}^{s/u}(x,\delta),

~s/u(x,δ):={y~s/u(x)|d~s/u(x,y)δ},\displaystyle\tilde{\mathcal{F}}^{s/u}(x,\delta):=\big{\{}y\in\tilde{\mathcal{F}}^{s/u}(x)\;\big{|}\;d_{\tilde{\mathcal{F}}^{s/u}}(x,y)\leq\delta\big{\}}, (2.2)

for all xM~x\in\tilde{M}, where d~s/u(,)d_{\tilde{\mathcal{F}}^{s/u}}(\cdot,\cdot) is induced by the metric on M~\tilde{M}.

In the rest of this paper, we restrict the manifolds MM to be a dd-torus 𝕋d\mathbb{T}^{d}. Let f:𝕋d𝕋df:\mathbb{T}^{d}\to\mathbb{T}^{d} be an Anosov map and A:𝕋d𝕋dA:\mathbb{T}^{d}\to\mathbb{T}^{d} its linearization. The following proposition exhibits the equivalent condtion of an Anosov map being conjugate with its linearization.

Proposition 2.2 ([30]).

Let ff be an Anosov map on torus, then ff is conjugate to its linearization if and only if ff is special.

On the other hand, from the observation of Proposition 2.1, we can expect that the liftings of ff and AA are conjugate. Let F:ddF:\mathbb{R}^{d}\to\mathbb{R}^{d} be a lift of ff and A¯:dd\overline{A}:\mathbb{R}^{d}\to\mathbb{R}^{d} be the lift of A:𝕋d𝕋dA:\mathbb{T}^{d}\to\mathbb{T}^{d} induced by the same projection π:d𝕋d\pi:\mathbb{R}^{d}\to\mathbb{T}^{d}. It means that πF=fπ\pi\circ F=f\circ\pi and πA¯=Aπ\pi\circ\overline{A}=A\circ\pi. For short, we denote A¯\overline{A} by AA if there is no confusion. The following proposition [1, Proposition 8.2.1 and Proposition 8.4.2] says that we do have a conjugacy between FF and AA.

Proposition 2.3 ([1]).

Let f:𝕋d𝕋df:\mathbb{T}^{d}\to\mathbb{T}^{d} be an Anosov map with lifting F:ddF:\mathbb{R}^{d}\to\mathbb{R}^{d} and AA be its linearization. There is a unique bijection H:ddH:\mathbb{R}^{d}\to\mathbb{R}^{d} such that

  1. 1.

    AH=HFA\circ H=H\circ F.

  2. 2.

    HH and H1H^{-1} are both uniformly continuous.

  3. 3.

    There exists C>0C>0 such that HId<C\|H-Id\|<C and H1Id<C\|H^{-1}-Id\|<C.

Remark 2.4.

Without losing generality, we can always assume that F(0)=0F(0)=0 and H(0)=0H(0)=0.

By proposition 2.2, it is clear that ff is special if and only if HH is commutative with d\mathbb{Z}^{d}-action, namely,

H(x+n)=H(x)+n,xdandnd.H(x+n)=H(x)+n,\quad\forall x\in\mathbb{R}^{d}\;\;\;{\rm and}\;\;\;\forall n\in\mathbb{Z}^{d}.

Although in general HH cannot be commutative with d\mathbb{Z}^{d}-action, we will see it can be commutative with d\mathbb{Z}^{d}-action as a stable leaf conjugacy.

Notation.

Denote the stable/unstable bundles and foliations of AA on 𝕋d\mathbb{T}^{d} by Ls/uL^{s/u}, s/u\mathcal{L}^{s/u} and on d\mathbb{R}^{d} by L~s/u\tilde{L}^{s/u}, ~s/u\tilde{\mathcal{L}}^{s/u} respectively.

It is clear that HH is a stable/unstable leaf conjugacy between FF and AA, namely,

H(~s/u)=~s/uandH(~s/u(Fx))=A(~s/u(Hx)).\displaystyle H(\tilde{\mathcal{F}}^{s/u})=\tilde{\mathcal{L}}^{s/u}\quad{\rm and}\quad H(\tilde{\mathcal{F}}^{s/u}(Fx))=A(\tilde{\mathcal{L}}^{s/u}(Hx)). (2.3)

Indeed, by the topological character of stable/unstable foliations (2.1) for AA and FF, one can get (2.3), directly. Especially, we mention that ~s\tilde{\mathcal{F}}^{s} and ~u\tilde{\mathcal{F}}^{u} admit the Global Product Structure, namely, any two leaves ~s(x)\tilde{\mathcal{F}}^{s}(x) and ~u(y)\tilde{\mathcal{F}}^{u}(y) intersect transversely at a unique point in d\mathbb{R}^{d}.

Proposition 2.5.

Assume that HH is given by Proposition 2.3. Then for every xdx\in\mathbb{R}^{d} and ndn\in\mathbb{Z}^{d},

H(x+n)n~s(H(x))andH1(x+n)n~s(H1(x)).H(x+n)-n\in\tilde{\mathcal{L}}^{s}\big{(}H(x)\big{)}\quad{\rm and}\quad H^{-1}(x+n)-n\in\tilde{\mathcal{F}}^{s}\big{(}H^{-1}(x)\big{)}.
Proof.

By Proposition 2.3, let H:ddH:\mathbb{R}^{d}\to\mathbb{R}^{d} be the unique conjugacy with H1Id<C0\|H^{-1}-Id\|<C_{0}. Now, we iterate these two points H1(x+n)nH^{-1}(x+n)-n and H1(x)H^{-1}(x) forward by FF. Note that since ff and AA are homotopic, we have Fk(x+n)=Fk(x)+AknF^{k}(x+n)=F^{k}(x)+A^{k}n, for all xdx\in\mathbb{R}^{d}, ndn\in\mathbb{Z}^{d} and kk\in\mathbb{N}. It follows that

d(FkH1(x),Fk(H1(x+n)n)),\displaystyle\quad d\left(F^{k}\circ H^{-1}(x)\;,\;F^{k}(H^{-1}(x+n)-n)\right),
=d(H1(Akx)+Akn,H1(Akx+Akn))),\displaystyle=d\left(H^{-1}(A^{k}x)+A^{k}n\;,\;H^{-1}(A^{k}x+A^{k}n))\right),
d(H1(Akx)+Akn,Akx+Akn)+d(Akx+Akn,H1(Akx+Akn))),\displaystyle\leq d\left(H^{-1}(A^{k}x)+A^{k}n\;,\;A^{k}x+A^{k}n\right)\;+\;d\left(A^{k}x+A^{k}n\;,\;H^{-1}(A^{k}x+A^{k}n))\right),
=d(H1(Akx),Akx)+d(Akx+Akn,H1(Akx+Akn)))2C0.\displaystyle=d\left(H^{-1}(A^{k}x)\;,\;A^{k}x\right)\;+\;d\left(A^{k}x+A^{k}n\;,\;H^{-1}(A^{k}x+A^{k}n))\right)\leq 2C_{0}.

Let kk tend to infinity, the fact that d(Fk(H1(x)+n),Fk(H1(x+n)))d\big{(}F^{k}(H^{-1}(x)+n),F^{k}(H^{-1}(x+n))\big{)} is always bounded by a uniform constant is sufficient to prove H1(x+n)n~s(H1(x))H^{-1}(x+n)-n\in\tilde{\mathcal{F}}^{s}\big{(}H^{-1}(x)\big{)}.

The proof for H(x+n)n~s(H(x))H(x+n)-n\in\tilde{\mathcal{L}}^{s}\big{(}H(x)\big{)} deduces from the fact that H(~s)=~sH(\tilde{\mathcal{F}}^{s})=\tilde{\mathcal{L}}^{s}. ∎

The following three propositions are all related to approching by ”special” d\mathbb{Z}^{d}-sequences which will be useful in Section 5.

Proposition 2.6.

Let f:𝕋d𝕋df:\mathbb{T}^{d}\to\mathbb{T}^{d} be an Anosov map with linearization AA, and HH be the conjugacy between its lifting FF and AA. There exist C>0C>0 and {εm}\{\varepsilon_{m}\} with εm0\varepsilon_{m}\to 0 as m+m\to+\infty, such that for every xdx\in\mathbb{R}^{d} and every nmdn_{m}\in\mathbb{Z}^{d} satisfying

Ainmd,1im,A^{-i}n_{m}\in\mathbb{Z}^{d},\qquad\forall 1\leq i\leq m,

the following two inequations hold

|H(x+nm)H(x)nm|<CA|Lsm,|H(x+n_{m})-H(x)-n_{m}|<C\cdot\|A|_{L^{s}}\|^{m},

and

|H1(x+nm)H1(x)nm|<εm.|H^{-1}(x+n_{m})-H^{-1}(x)-n_{m}|<\varepsilon_{m}.
Proof.

By Ainmd(1im)A^{-i}n_{m}\in\mathbb{Z}^{d}(1\leq i\leq m), we have Fi(x+nm)Fi(x)=AinmdF^{-i}(x+n_{m})-F^{-i}(x)=A^{-i}n_{m}\in\mathbb{Z}^{d}, for all 1im1\leq i\leq m. By |Hid|<C0|H-id|<C_{0}, we have

d(H(Fmx+Amnm),H(Fmx)+Amnm)2C0.d\left(H(F^{-m}x+A^{-m}n_{m})\;,\;H(F^{-m}x)+A^{-m}n_{m}\right)\leq 2C_{0}.

By Proposition 2.5, one has H(x)+n~s(H(x+n))H(x)+n\in\tilde{\mathcal{L}}^{s}\big{(}H(x+n)\big{)}, for every xdx\in\mathbb{R}^{d} and ndn\in\mathbb{Z}^{d}. Hence,

d(Am(H(Fmx+Amnm)),Am(H(Fmx)+Amnm))2C0A|Lsm.d\Big{(}A^{m}\left(H(F^{-m}x+A^{-m}n_{m})\right)\;,\;A^{m}\left(H(F^{-m}x)+A^{-m}n_{m}\right)\Big{)}\leq 2C_{0}\cdot\|A|_{L^{s}}\|^{m}.

That is

|H(x+nm)H(x)nm|<2C0A|Lsm.|H(x+n_{m})-H(x)-n_{m}|<2C_{0}\cdot\|A|_{L^{s}}\|^{m}.

On the other hand, by the uniform continuity of H1H^{-1}, there exists εm\varepsilon_{m} satisfying εm0\varepsilon_{m}\to 0 as m+m\to+\infty such that if d(x,y)<2C0A|Lsmd(x,y)<2C_{0}\cdot\|A|_{L^{s}}\|^{m}, then d(H1(x),H1(y))<εmd(H^{-1}(x),H^{-1}(y))<\varepsilon_{m}. Thus,

d(x+nm,H1(H(x)+nm))<εm.d\left(x+n_{m}\;,\;H^{-1}(H(x)+n_{m})\right)<\varepsilon_{m}.

Equivalently, for every xdx\in\mathbb{R}^{d}, we have

|H1(x+nm)H1(x)nm|<εm.\Big{|}H^{-1}(x+n_{m})-H^{-1}(x)-n_{m}\Big{|}<\varepsilon_{m}.

The following proposition is a corollary of Proposition 2.6. It says that although the FF-invariant foliation may not be commutative with d\mathbb{Z}^{d}-actions, it can ”almost” be commutative with nmn_{m}-actions if HH maps it to an AA-invariant linear foliation.

Proposition 2.7.

On the assumption of Proposition 2.6 and assume that ~\tilde{\mathcal{L}} is an AA-invariant linear foliation and the sequence {nm}md\{n_{m}\}_{m\in\mathbb{N}}\subset\mathbb{Z}^{d} satisfies nmAmdn_{m}\in A^{m}\mathbb{Z}^{d}. If each leaf of the FF-invariant foliation ~:=H1(~)\tilde{\mathcal{F}}:=H^{-1}(\tilde{\mathcal{L}}) is C1C^{1}-smooth, then for every xdx\in\mathbb{R}^{d} and every R>0R>0,

~(x+nm,R)nm~(x,R),\tilde{\mathcal{F}}(x+n_{m},R)-n_{m}\to\tilde{\mathcal{F}}(x,R),

as m+m\to+\infty, where the local manifold ~(x,R):={y~(x)|d~(x,y)R}\tilde{\mathcal{F}}(x,R):=\{y\in\tilde{\mathcal{F}}(x)\;|\;d_{\tilde{\mathcal{F}}}(x,y)\leq R\}.

Proof.

By Proposition 2.6, when m+m\to+\infty,

dH(H(~(x+nm,R)nm),H(~(x+nm,R))nm)0,\displaystyle d_{H}\Big{(}H\left(\tilde{\mathcal{F}}(x+n_{m},R)-n_{m}\right)\;\;,\;\;H\left(\tilde{\mathcal{F}}(x+n_{m},R)\right)-n_{m}\Big{)}\to 0, (2.4)

where dH(,)d_{H}(\cdot,\cdot) is Hausdorff distance. Note that 𝒯:=H(~(x+nm,R))\mathcal{T}:=H\left(\tilde{\mathcal{F}}(x+n_{m},R)\right) is a local leaf on ~(H(x+nm))\tilde{\mathcal{L}}\big{(}H(x+n_{m})\big{)}. Since ~\tilde{\mathcal{L}} is commutative with d\mathbb{Z}^{d}-actions, the set (𝒯nm)(\mathcal{T}-n_{m}) is a copy of 𝒯\mathcal{T} on ~(H(x+nm)nm)\tilde{\mathcal{L}}\big{(}H(x+n_{m})-n_{m}\big{)}. Again, by Proposition 2.6, as m+m\to+\infty,

dH(𝒯nm,𝒯H(x+nm)+H(x))0.\displaystyle d_{H}\big{(}\mathcal{T}-n_{m}\;\;,\;\;\mathcal{T}-H(x+n_{m})+H(x)\big{)}\to 0. (2.5)

Then, combining (2.4) and (2.5), one has

dH(H(~(x+nm,R)nm),𝒯H(x+nm)+H(x))0.d_{H}\Big{(}H\left(\tilde{\mathcal{F}}(x+n_{m},R)-n_{m}\right)\;\;,\;\;\mathcal{T}-H(x+n_{m})+H(x)\Big{)}\to 0.

By the uniform continuity of H1H^{-1}, it follows that

dH(~(x+nm,R)nm,H1(𝒯H(x+nm)+H(x)))0.d_{H}\Big{(}\tilde{\mathcal{F}}(x+n_{m},R)-n_{m}\;\;,\;\;H^{-1}\big{(}\mathcal{T}-H(x+n_{m})+H(x)\big{)}\Big{)}\to 0.

Since (𝒯H(x+nm)+H(x))\big{(}\mathcal{T}-H(x+n_{m})+H(x)\big{)} is a copy of 𝒯\mathcal{T} on ~(H(x))\tilde{\mathcal{L}}\big{(}H(x)\big{)}, the set H1(𝒯H(x+nm)+H(x))H^{-1}\big{(}\mathcal{T}-H(x+n_{m})+H(x)\big{)} is a local leaf on ~(x)\tilde{\mathcal{F}}(x). Moreover, its size tends to RR as m+m\to+\infty. ∎

Remark 2.8.

The foliation ~\tilde{\mathcal{F}} in Proposition 2.7 can be the unstable foliation or the center foliation of FF, where F:ddF:\mathbb{R}^{d}\to\mathbb{R}^{d} is a lifting of an Anosov map on torus.

A foliation \mathcal{F} on 𝕋d\mathbb{T}^{d} is called minimal, if its every leaf is dense. It is clear that if AA is irreducible, then every AA-invariant linear foliation on 𝕋d\mathbb{T}^{d} is minimal (a complete proof is available in [10]). The following proposition actually says that the projection of each leaf of FF-invariant foliation onto 𝕋d\mathbb{T}^{d} is dense if HH maps it to an AA-invariant linear foliation.

Proposition 2.9.

On the assumption of Proposition 2.6 and assume that AA is irreducible and ~\tilde{\mathcal{L}} is an AA-invariant linear foliation. Let ~=H1(~)\tilde{\mathcal{F}}=H^{-1}(\tilde{\mathcal{L}}) be a foliation on d\mathbb{R}^{d}. Then for any x,ydx,y\in\mathbb{R}^{d},

  1. 1.

    There exist xm~(x)x_{m}\in\tilde{\mathcal{L}}(x) and nmAmdn_{m}\in A^{m}\mathbb{Z}^{d} , such that (xm+nm)y(x_{m}+n_{m})\to y as m+m\to+\infty.

  2. 2.

    There exist zm~(x)z_{m}\in\tilde{\mathcal{F}}(x) and nmAmdn_{m}\in A^{m}\mathbb{Z}^{d} , such that (zm+nm)y(z_{m}+n_{m})\to y as m+m\to+\infty.

Proof.

Since AA is irreducible, the set {~(0)+d}\{\tilde{\mathcal{L}}(0)+\mathbb{Z}^{d}\} is dense in d\mathbb{R}^{d}. Fix mm\in\mathbb{N}, one has that the set Am{~(0)+d}A^{m}\{\tilde{\mathcal{L}}(0)+\mathbb{Z}^{d}\} is dense in d\mathbb{R}^{d}. It follows that {~(0)+Amd}\{\tilde{\mathcal{L}}(0)+A^{m}\mathbb{Z}^{d}\} is dense in d\mathbb{R}^{d}, since Am~(0)=~(0)A^{m}\tilde{\mathcal{L}}(0)=\tilde{\mathcal{L}}(0). This complete the proof for the first item.

Now, applying the first item for points H(x)H(x) and H(y)H(y), we can take nmAmdn_{m}\in A^{m}\mathbb{Z}^{d} and xm~(H(x))x_{m}^{\prime}\in\tilde{\mathcal{L}}\big{(}H(x)\big{)} with xm+nmH(y)x_{m}^{\prime}+n_{m}\to H(y). Let zm=H1(xm)~(x)z_{m}=H^{-1}(x_{m}^{\prime})\in\tilde{\mathcal{F}}(x). By Proposition 2.6, one has d(H(zm+nm),H(zm)+nm)0d\big{(}H(z_{m}+n_{m}),H(z_{m})+n_{m}\big{)}\to 0 as m+m\to+\infty. By the uniform continuity of H1H^{-1}, we get (zm+nm)y(z_{m}+n_{m})\to y. ∎

To end this subsection, we state a proposition about the density of preimage sets for an irreducible toral endomorphism (may not be Anosov) which will be used in Section 3.

Proposition 2.10.

Let AMd()GLA\in M_{d}(\mathbb{Z})\cap\rm{GL}()d{}_{d}(\mathbb{R}) be irreducible over \mathbb{Q}. It induces a torus endomorphism A:𝕋d𝕋dA:\mathbb{T}^{d}\to\mathbb{T}^{d}. Then there exists C>0C>0 such that for every k1k\geq 1 and every x0𝕋dx_{0}\in\mathbb{T}^{d}, the kk-preimage set of x0x_{0}

{x𝕋d|Ak(x)=x0}\displaystyle\{x\in\mathbb{T}^{d}|A^{k}(x)=x_{0}\}

is C|det(A)|k/d\;C|{\rm det}(A)|^{-k/d}-dense in 𝕋d\mathbb{T}^{d}.

Proof.

Consider the dd-dimensional real vector space

V=span{Id,A,,Ad1}V=\text{span}_{\mathbb{R}}\{I_{d},A,...,A^{d-1}\}

and the lattice V:=VMd()V_{\mathbb{Z}}:=V\cap M_{d}(\mathbb{Z}) in VV. We claim that nonzero matrices in VV_{\mathbb{Z}} are invertible in GLd()GL_{d}(\mathbb{R}). In fact, if MV{0}M\in V_{\mathbb{Z}}-\{0\}, then there is a nonzero rational polynomial m[x]m\in\mathbb{Q}[x] with deg(m)d1(m)\leq d-1 such that M=m(A)M=m(A). Note that mm and the characteristic polynomial χ\chi of AA are coprime over \mathbb{Q}, since AA is irreducible. So there exist g,h[x]g,h\in\mathbb{Q}[x] such that mg+χh=1mg+\chi h=1. It follows that m(A)g(A)=Idm(A)g(A)=I_{d}. So M=m(A)M=m(A) is invertible.

Fix 𝒦V\mathcal{K}\subset V be a compact covex symmetric subset with vol(𝒦)2dvol(V/V)(\mathcal{K})\geq 2^{d}\text{vol}(V/V_{\mathbb{Z}}), and let

C=dsupK𝒦,vd,v=1Kv.C={\sqrt{d}}\cdot\sup_{K\in\mathcal{K},v\in\mathbb{R}^{d},\|v\|=1}\|Kv\|.

We prove that for every k1k\geq 1, the kk-preimage set of every x0𝕋dx_{0}\in\mathbb{T}^{d} is C|det(A)|k/dC|\text{det}(A)|^{-k/d}-dense in 𝕋d\mathbb{T}^{d}. It suffices to prove that AkdA^{-k}\mathbb{Z}^{d} is C|det(A)|k/dC|\text{det}(A)|^{-k/d}-dense in d\mathbb{R}^{d}.

Let L:VVL:V\to V be the linear map defined be L(X)=AXL(X)=AX. The matrix of LL relative to the basis {Id,A,,Ad1}\{I_{d},A,...,A^{d-1}\} of VV is the companion matrix of χ\chi. So det(L)=det(A)\text{det}(L)=\text{det}(A). It follows that the compact convex symmetric set |det(A)|k/dLk(𝒦)|\text{det}(A)|^{-k/d}\cdot L^{k}(\mathcal{K}) has volume

vol(|det(A)|k/dLk(𝒦))\displaystyle{\rm vol}\left(|\text{det}(A)|^{-k/d}\cdot L^{k}(\mathcal{K})\right) =|det(A)|kvol(Lk(𝒦))\displaystyle=|\text{det}(A)|^{-k}\cdot\text{vol}\left(L^{k}(\mathcal{K})\right)
=|det(A)|k|det(L)|kvol(𝒦)\displaystyle=|\text{det}(A)|^{-k}|\text{det}(L)|^{k}\cdot\text{vol}(\mathcal{K})
=vol(𝒦)2dvol(V/V).\displaystyle=\text{vol}(\mathcal{K})\geq 2^{d}\text{vol}(V/V_{\mathbb{Z}}).

By Minkowski’s convex body theorem, it contains a nonzero matrix in VV_{\mathbb{Z}}. Namely, there exist K𝒦K\in\mathcal{K} and MV{0}M\in V_{\mathbb{Z}}-\{0\} such that |det(A)|k/dAkK=M|\text{det}(A)|^{-k/d}A^{k}K=M. Note that MM is invertible, so KK is also invertible. For r>0r>0, let B(r)dB(r)\subset\mathbb{R}^{d} denote the open ball with radius rr centered at the origin. Then

d\displaystyle\mathbb{R}^{d} =|det(A)|k/dKd=|det(A)|k/dK(d+B(d))\displaystyle=|\text{det}(A)|^{-k/d}K\mathbb{R}^{d}=|\text{det}(A)|^{-k/d}K\big{(}\mathbb{Z}^{d}+B(\sqrt{d})\big{)}
=AkMd+|det(A)|k/dKB(d)\displaystyle=A^{-k}M\mathbb{Z}^{d}+|\text{det}(A)|^{-k/d}KB(\sqrt{d})
Akd+B(C|det(A)|k/d).\displaystyle\subseteq A^{-k}\mathbb{Z}^{d}+B\big{(}C|\text{det}(A)|^{-k/d}\big{)}.

This means that AkdA^{-k}\mathbb{Z}^{d} is C|det(A)|k/dC|\text{det}(A)|^{-k/d}-dense in d\mathbb{R}^{d}. ∎

2.2 Dominated splitting on the inverse limit space

Now we introduce the dynamics on the inverse limit space. Note that the inverse limit space has compactness which the universal cover lacks.

Firstly, we clear the definition of inverse limit space. Let (M,d)(M,d) be a compact metric space and M:={(xi)|xiM,i}M^{\mathbb{Z}}:=\{(x_{i})\;|\;x_{i}\in M,\forall i\in\mathbb{Z}\} be the product topological space. MM^{\mathbb{Z}} is compact by Tychonoff theorem and it can be metrizable by the metric

d~((xi),(yi))=+d(xi,yi)2|i|.\tilde{d}((x_{i}),(y_{i}))=\sum_{-\infty}^{+\infty}\frac{d(x_{i},y_{i})}{2^{|i|}}.

Let σ:MM\sigma:M^{\mathbb{Z}}\to M^{\mathbb{Z}} be the (left) shift homeomorphism by (σ(xi))j=xj+1(\sigma(x_{i}))_{j}=x_{j+1}, for all jj\in\mathbb{Z}. For a continuous map f:MMf:M\to M, the inverse limit space of ff is

Mf:={(xi)|xiMandf(xi)=xi+1,i}.M_{f}:=\big{\{}(x_{i})\;|\;x_{i}\in M\;\;{\rm and}\;\;f(x_{i})=x_{i+1},\forall i\in\mathbb{Z}\big{\}}.

With the metric d~(,)\tilde{d}(\cdot,\cdot), the inverse limit space (Mf,d~)(M_{f},\tilde{d}) is a closed subset of (M,d~)(M^{\mathbb{Z}},\tilde{d}). So it is a compact metric space. It is clear that (Mf,d~)(M_{f},\tilde{d}) is σ\sigma-invariant.

Definition 2.11 ([33]).

A C1C^{1} local diffeomorphism f:MMf:M\to M is called Anosov map, if there exist constants C>0C>0 and 0<μ<10<\mu<1 such that, for every x~=(xi)Mf\tilde{x}=(x_{i})\in M_{f}, there exists a hyperbolic splitting

TxiM=Es(xi,x~)Eu(xi,x~),i,\displaystyle T_{x_{i}}M=E^{s}(x_{i},\tilde{x})\oplus E^{u}(x_{i},\tilde{x}),\quad\forall i\in\mathbb{Z},

which is DfDf-invariant

Dxif(Es(xi,x~))=Es(xi+1,x~)andDxif(Eu(xi,x~))=Eu(xi+1,x~),i,\displaystyle D_{x_{i}}f\left(E^{s}(x_{i},\tilde{x})\right)=E^{s}(x_{i+1},\tilde{x})\qquad{\rm and}\qquad D_{x_{i}}f\left(E^{u}(x_{i},\tilde{x})\right)=E^{u}(x_{i+1},\tilde{x}),\qquad\forall i\in\mathbb{Z},

and for all n>0n>0 the following estimates hold:

Dxifn(v)Cμnv,vEs(xi,x~),i,\displaystyle\|D_{x_{i}}f^{n}(v)\|\leq C\mu^{n}\|v\|,\qquad\quad\forall v\in E^{s}(x_{i},\tilde{x}),\;\;\forall i\in\mathbb{Z},
Dxifn(v)C1μnv,vEu(xi,x~),i.\displaystyle\|D_{x_{i}}f^{n}(v)\|\geq C^{-1}\mu^{-n}\|v\|,\;\;\quad\forall v\in E^{u}(x_{i},\tilde{x}),\;\;\forall i\in\mathbb{Z}.

We extend the hyperbolic splitting on the inverse limit space to the dominated splitting case. We say a local diffeomorphism g:MMg:M\to M admits a dominated splitting

TM=E1E2Em,TM=E_{1}\oplus E_{2}\oplus...\oplus E_{m},

if for every i=1,,mi=1,\cdots,m, each subbundle EiE_{i} is DgDg-invariant and there exist C>0C>0, 0<λ<10<\lambda<1 such that for any xMx\in M and any unit vectors uEi(x)u\in E_{i}(x) and vEi+1(x)v\in E_{i+1}(x),

DxgnuCλnDxgnv,n.\|D_{x}g^{n}u\|\leq C\lambda^{n}\|D_{x}g^{n}v\|,\quad\forall n\in\mathbb{N}.

Let A:𝕋d𝕋dA:\mathbb{T}^{d}\to\mathbb{T}^{d} be a linear Anosov map and admits the finest (on stable bundle) dominated splitting,

T𝕋d=L1sL2sLksLu,\displaystyle T\mathbb{T}^{d}=L^{s}_{1}\oplus L^{s}_{2}\oplus...\oplus L^{s}_{k}\oplus L^{u}, (2.6)

where dimLis=1{\rm dim}L^{s}_{i}=1, 1ik1\leq i\leq k, such that

0<|μ1s(A)|<|μ2s(A)|<<|μks(A)|<1<|μminu(A)|,\displaystyle 0<|\mu^{s}_{1}(A)|<|\mu^{s}_{2}(A)|<...<|\mu^{s}_{k}(A)|<1<|\mu_{\rm min}^{u}(A)|, (2.7)

where μis(A)\mu^{s}_{i}(A) is the eigenvalue with respect to LisL^{s}_{i} and μminu(A)=m(A|Lu)\mu_{\rm min}^{u}(A)=m(A|_{L^{u}}) the mininorm of AA restricted on LuL^{u}. Denote by μmaxu(A)=A\mu_{\rm max}^{u}(A)=\|A\|, the norm of AA. And denote by λis(A)=log|μis(A)|\lambda^{s}_{i}(A)={\rm log}|\mu^{s}_{i}(A)| the stable Lyapunov exponent of the subbundle LisL^{s}_{i}.

Notation.

We use the following notations to denote the joint bundles of AA which are anological for L~\tilde{L}, \mathcal{L}, ~\tilde{\mathcal{L}} and the bundles and foliations of ff and FF, if they are well defined.

  1. 1.

    L(i,j)s:=LisLi+1sLjsL^{s}_{(i,j)}:=L^{s}_{i}\oplus L^{s}_{i+1}\oplus...\oplus L^{s}_{j}, 1ijk1\leq i\leq j\leq k.

  2. 2.

    L(i,j)s,u:=LisLi+1sLjsLuL^{s,u}_{(i,j)}:=L^{s}_{i}\oplus L^{s}_{i+1}\oplus...\oplus L^{s}_{j}\oplus L^{u}, 1ijk1\leq i\leq j\leq k.

  3. 3.

    L(i,i)s=LisL^{s}_{(i,i)}=L^{s}_{i}, L(i,i)s,u=Lis,u=LisLuL^{s,u}_{(i,i)}=L^{s,u}_{i}=L^{s}_{i}\oplus L^{u}, 1ik1\leq i\leq k.

Proposition 2.12.

Let AMd()GLd()A\in M_{d}(\mathbb{Z})\cap GL_{d}(\mathbb{R}) induce a toral Anosov map with the finest (on stable bundle) dominated splitting satisfying (2.6) and (2.7). Then, there exists a C1C^{1} neighborhood 𝒰C1(𝕋d)\mathcal{U}\subset C^{1}(\mathbb{T}^{d}) of AA such that for all f𝒰f\in\mathcal{U} and all x~=(xn)𝕋fd\tilde{x}=(x_{n})\in\mathbb{T}^{d}_{f}, there exists a dominated splitting

Txn𝕋d=E1s(xn,x~)Eks(xn,x~)Eu(xn,x~),n,T_{x_{n}}\mathbb{T}^{d}=E^{s}_{1}(x_{n},\tilde{x})\oplus...\oplus E^{s}_{k}(x_{n},\tilde{x})\oplus E^{u}(x_{n},\tilde{x}),\quad\forall n\in\mathbb{Z},

where dimEis=1,1ikE^{s}_{i}=1,1\leq i\leq k. And the bundles Eis(xn,x~)E^{s}_{i}(x_{n},\tilde{x}) and Eu(xn,x~)E^{u}(x_{n},\tilde{x}) are continuous with x~\tilde{x}. Especially, there exist bundles E(1,i)s(1ik)E^{s}_{(1,i)}(1\leq i\leq k) defined on 𝕋d\mathbb{T}^{d} such that

E1sE(1,2)sE(1,k1)sE(1,k)s=Es.\displaystyle E^{s}_{1}\subset E^{s}_{(1,2)}\subset...\subset E^{s}_{(1,k-1)}\subset E^{s}_{(1,k)}=E^{s}. (2.8)

Moreover, for any α(0,π2)\alpha\in(0,\frac{\pi}{2}) and δ>0\delta>0, there exists 𝒰C1(𝕋d)\mathcal{U}\subset C^{1}(\mathbb{T}^{d}) such that for every f𝒰f\in\mathcal{U} and x~=(xn)𝕋fd\tilde{x}=(x_{n})\in\mathbb{T}^{d}_{f} and 1ik1\leq i\leq k,

(Eis(x0,x~),Lis(x0))αandDf|Eis(x0,x~)[μis(A)δ,μis(A)+δ].\displaystyle\angle\left(E^{s}_{i}(x_{0},\tilde{x}),L^{s}_{i}(x_{0})\right)\leq\alpha\quad{\rm and}\quad\|Df|_{E^{s}_{i}(x_{0},\tilde{x})}\|\in\big{[}\mu^{s}_{i}(A)-\delta,\mu^{s}_{i}(A)+\delta\big{]}.\qquad (2.9)

And,

(Eu(x0,x~),Lu(x0))αandDf|Eu(x0,x~)[μmaxu(A)δ,μmaxu(A)+δ].\displaystyle\angle\left(E^{u}(x_{0},\tilde{x}),L^{u}(x_{0})\right)\leq\alpha\quad{\rm and}\quad\|Df|_{E^{u}(x_{0},\tilde{x})}\|\in\big{[}\mu^{u}_{\rm max}(A)-\delta,\mu_{\rm max}^{u}(A)+\delta\big{]}. (2.10)
Proof.

We use invariant cone-fields to complete this proof. One may find more details in [2, Appendix B.1] and [5, Section 2.2 ].

Let SUS\oplus U be a dominated splitting on T𝕋dT\mathbb{T}^{d} of AA and 𝒞α\mathcal{C}_{\alpha} be an AA-invariant cone-fields contain UU with size α\alpha. That is

𝒞α(x)={ν=νS+νUTx𝕋d:νSανU}andDA(𝒞α(x))𝒞μα(x),\mathcal{C}_{\alpha}(x)=\big{\{}\nu=\nu_{S}+\nu_{U}\in T_{x}\mathbb{T}^{d}:\;\;\|\nu_{S}\|\leq\alpha\|\nu_{U}\|\big{\}}\quad{\rm and}\quad DA(\mathcal{C}_{\alpha}(x))\subset\mathcal{C}_{\mu\cdot\alpha}(x),

for some μ<1\mu<1. So, there exists a C1C^{1}-neighborhood 𝒰\mathcal{U} of AA such that for any f𝒰f\in\mathcal{U} and any x𝕋dx\in\mathbb{T}^{d}, one has Df(𝒞α(x))int(𝒞α(fx))Df(\mathcal{C}_{\alpha}(x))\subset{\rm int}\left(\mathcal{C}_{\alpha}(fx)\right). This implies the existence of dominated splitting (see [5, Section 2.2 ]). We mention that we can make a larger perturbation as long as 𝒞α\mathcal{C}_{\alpha} can be contracted into itself by finite iterations of DfDf. And let 𝒞α\mathcal{C}^{*}_{\alpha} be the closure of the complement of 𝒞α\mathcal{C}_{\alpha}. It is clear that 𝒞α\mathcal{C}^{*}_{\alpha} is also a cone-field which is contracted by Df1Df^{-1}, if the preimage is given. Fix an ff-orbit x~=(xn)𝕋fd\tilde{x}=(x_{n})\in\mathbb{T}^{d}_{f} and let

Sf(x0,x~):=n0Dfn(𝒞α(xn)),Uf(x0,x~):=n0Dfn(𝒞α(xn)).S_{f}(x_{0},\tilde{x}):=\bigcap_{n\leq 0}Df^{-n}\big{(}\mathcal{C}^{*}_{\alpha}(x_{n})\big{)},\quad U_{f}(x_{0},\tilde{x}):=\bigcap_{n\leq 0}Df^{n}\big{(}\mathcal{C}_{\alpha}(x_{-n})\big{)}.

Then SfUfS_{f}\oplus U_{f} give a dominated splitting on the inverse limits space 𝕋fd\mathbb{T}^{d}_{f}. Note that Sf(x0,x~)S_{f}(x_{0},\tilde{x}) is independent of the choice of orbits for x0x_{0}.

For finding out Eis(x0,x~)E_{i}^{s}(x_{0},\tilde{x}), we consider these two dominated splittings for AA

L(1,i)sL(i+1,k)s,uandL(1,i1)sL(i,k)s,u.L^{s}_{(1,i)}\oplus L^{s,u}_{(i+1,k)}\quad{\rm and}\quad L^{s}_{(1,i-1)}\oplus L^{s,u}_{(i,k)}.

So, we can get the following two dominated splittings under the C1C^{1}-perturbation,

E(1,i)s(x0,x~)E(i+1,k)s,u(x0,x~)andE(1,i1)s(x0,x~)E(i,k)s,u(x0,x~).E^{s}_{(1,i)}(x_{0},\tilde{x})\oplus E^{s,u}_{(i+1,k)}(x_{0},\tilde{x})\quad{\rm and}\quad E^{s}_{(1,i-1)}(x_{0},\tilde{x})\oplus E^{s,u}_{(i,k)}(x_{0},\tilde{x}).

Hence, we get Eis(x0,x~)=E(1,i)s(x0,x~)E(i,k)s,u(x0,x~)E_{i}^{s}(x_{0},\tilde{x})=E^{s}_{(1,i)}(x_{0},\tilde{x})\cap E^{s,u}_{(i,k)}(x_{0},\tilde{x}). It is clear that this dominated splitting is continuous with respect to orbits. Meanwhile, E(1,i)s(x0,x~)E^{s}_{(1,i)}(x_{0},\tilde{x}) only depends on x0x_{0}. It follows that the bundle E(1,i)sE^{s}_{(1,i)} is well defined on 𝕋d\mathbb{T}^{d}. The previous proof also allows the control (2.9) and (2.10) for bundles of ff. ∎

2.3 Foliations on the universal cover

In this subsection, we always assume that AMd()GLd()A\in M_{d}(\mathbb{Z})\cap GL_{d}(\mathbb{R}) admits the finest (on stable bundle) dominated splitting (see (2.6) and (2.7)). It is clear that there exists a C1C^{1} neighborhood 𝒰\mathcal{U} of AA such that for every f𝒰f\in\mathcal{U}, its lifting F:ddF:\mathbb{R}^{d}\to\mathbb{R}^{d} admits dominated splitting

Td=E~1sE~2sE~ksE~u.T\mathbb{R}^{d}=\tilde{E}^{s}_{1}\oplus\tilde{E}^{s}_{2}\oplus...\oplus\tilde{E}^{s}_{k}\oplus\tilde{E}^{u}.

Let H:ddH:\mathbb{R}^{d}\to\mathbb{R}^{d} be the unique conjugacy between FF and AA guaranteed by Proposition 2.3.

Let π:d𝕋d\pi:\mathbb{R}^{d}\to\mathbb{T}^{d} be the natural projection such that πF=fπ\pi\circ F=f\circ\pi. Note that any ff-orbit x~𝕋fd\tilde{x}\in\mathbb{T}^{d}_{f} can be approached by FF-orbits. Hence for every 1ijk1\leq i\leq j\leq k, one can get

x~𝕋fd,x~0=x0E(i,j)s(x0,x~)=π(y)=x0Dyπ(E~(i,j)s(y))¯,x~=(xn)𝕋fd.\displaystyle\bigcup_{\tilde{x}\in\mathbb{T}^{d}_{f},\tilde{x}_{0}=x_{0}}E^{s}_{(i,j)}(x_{0},\tilde{x})=\overline{\bigcup_{\pi(y)=x_{0}}D_{y}\pi\left(\tilde{E}^{s}_{(i,j)}(y)\right)},\qquad\forall\tilde{x}=(x_{n})\in\mathbb{T}^{d}_{f}. (2.11)

It is similar to Eu(x0,x~)E^{u}(x_{0},\tilde{x}). We refer to [29, Proposition 2.5] for more details. This projection allows us to get properties of FF from ones of 𝕋fd\mathbb{T}^{d}_{f}. Especially, we have the following remarks.

Remark 2.13.

By Proposition 2.12, we can choose 𝒰C1(𝕋d)\mathcal{U}\subset C^{1}(\mathbb{T}^{d}) small enough such that

μis(A)ε03DF|E~is(x)μis(A)+ε03and(E~is(x),L~is(x))<α<π2,\displaystyle\mu^{s}_{i}(A)-\frac{\varepsilon_{0}}{3}\leq\|DF|_{\tilde{E}^{s}_{i}(x)}\|\leq\mu^{s}_{i}(A)+\frac{\varepsilon_{0}}{3}\quad{\rm and}\quad\angle\left(\tilde{E}^{s}_{i}(x),\tilde{L}^{s}_{i}(x)\right)<\alpha<\frac{\pi}{2}, (2.12)

for every xdx\in\mathbb{R}^{d} and 1ik1\leq i\leq k, where ε0=min{1μks(A),μi+1s(A)μis(A):1ik1}\varepsilon_{0}={\rm min}\left\{1-\mu_{k}^{s}(A),\mu^{s}_{i+1}(A)-\mu^{s}_{i}(A):1\leq i\leq k-1\right\}.

Remark 2.14.

The distributions E~(i,j)s(1ijk)\tilde{E}^{s}_{(i,j)}(1\leq i\leq j\leq k) and E~u\tilde{E}^{u} are all Ho¨\ddot{\rm o}lder continuous on TdT\mathbb{R}^{d}. One can get this from [32, Theorem 2.3] with the fact that the angle (E~1,E~2)\angle(\tilde{E}_{1},\tilde{E}_{2}) is uniformly away from 0, where E~1,E~2\tilde{E}_{1},\tilde{E}_{2} are two different bundles in {E~u,E~is: 1ik}\big{\{}\tilde{E}^{u},\;\tilde{E}^{s}_{i}:\;1\leq i\leq k\big{\}} (also see Remark 2.16).

We say ~\tilde{\mathcal{F}} is a quasi-isometric foliation on d\mathbb{R}^{d}, if there exist contants a,b>0a,b>0 such that

d~(x,y)ad(x,y)+b,xd,y~(x).\displaystyle d_{\tilde{\mathcal{F}}}(x,y)\leq a\cdot d(x,y)+b,\quad\forall x\in\mathbb{R}^{d},y\in\tilde{\mathcal{F}}(x). (2.13)

A foliation ~\tilde{\mathcal{F}} defined on d\mathbb{R}^{d} is called d\mathbb{Z}^{d}-periodic (equivalently, commutative with d\mathbb{Z}^{d}-actions), if

~(x+n)=~(x)+n,xd,nd.\tilde{\mathcal{F}}(x+n)=\tilde{\mathcal{F}}(x)+n,\qquad\forall x\in\mathbb{R}^{d},\quad\forall n\in\mathbb{Z}^{d}.

Similarly, a d\mathbb{Z}^{d}-periodic bundle E~\tilde{E} on TdT\mathbb{R}^{d} means that

E~(x+n)=DTn(E~(x)),xd,nd.\tilde{E}(x+n)=DT_{n}\left(\tilde{E}(x)\right),\qquad\forall x\in\mathbb{R}^{d},\quad\forall n\in\mathbb{Z}^{d}.

Here Tn:ddT_{n}:\mathbb{R}^{d}\to\mathbb{R}^{d} is the translation Tn(x)=x+nT_{n}(x)=x+n.

Proposition 2.15.

There exists a C1C^{1} neighborhood 𝒰C1(𝕋d)\mathcal{U}\subset C^{1}(\mathbb{T}^{d}) of AA such that for every f𝒰f\in\mathcal{U} and its lifting FF, we have the followings,

  1. 1.

    The FF-invariant bundles E~(i,j)s(1ijk)\tilde{E}^{s}_{(i,j)}(1\leq i\leq j\leq k) and E~u\tilde{E}^{u} are uniquely integrable. Denote the integral foliations by ~(i,j)s\tilde{\mathcal{F}}^{s}_{(i,j)}, ~u\tilde{\mathcal{F}}^{u} and ~is\tilde{\mathcal{F}}^{s}_{i} ( if j=ij=i). Moreover, the foliation ~(i0,j0)s\tilde{\mathcal{F}}^{s}_{(i_{0},j_{0})} is subfoliated by ~(i,j)s\tilde{\mathcal{F}}^{s}_{(i,j)}, for any 1i0ijj0k1\leq i_{0}\leq i\leq j\leq j_{0}\leq k.

  2. 2.

    The strong stable bundle E~(1,i)s\tilde{E}^{s}_{(1,i)} and strong stable foliation ~(1,i)s\tilde{\mathcal{F}}^{s}_{(1,i)} (1ik)(1\leq i\leq k) are both d\mathbb{Z}^{d}-periodic.

  3. 3.

    The foliation ~(i,j)s(1ijk)\tilde{\mathcal{F}}^{s}_{(i,j)}(1\leq i\leq j\leq k) is quasi-isometric. Especially, for every Anosov map on torus with one-dimensional stable bundle, the lifting of stable foliation is quasi-isometric.

  4. 4.

    HH preserves the weak stable foliations. It means that H(~(i,k)s)=~(i,k)sH(\tilde{\mathcal{F}}^{s}_{(i,k)})=\tilde{\mathcal{L}}^{s}_{(i,k)}, for all 1ik1\leq i\leq k.

Proof.

Let 𝒰\mathcal{U} be given by Remark 2.13. Consider the dominated splitting

E~(1,i1)sE~(i,k)sE~u,\tilde{E}^{s}_{(1,i-1)}\oplus\tilde{E}^{s}_{(i,k)}\oplus\tilde{E}^{u},

where 2ik2\leq i\leq k. By the Stable Manifold Theorem e.g. [32, Theorem 4.1 and Theorem 4.8], we have that E~(1,i1)s\tilde{E}^{s}_{(1,i-1)} and E~u\tilde{E}^{u} are always integrable. Although [32] prove it for diffeomorphism, in our case Remark 2.13 and Remark 2.14 provides the uniform continuity and domination of bundles to replace the compactness. Moreover, by Proposition 2.12 and (2.11), the bundle E~(1,i)s\tilde{E}^{s}_{(1,i)} and foliation ~(1,i)s\tilde{\mathcal{F}}^{s}_{(1,i)} (1ik)(1\leq i\leq k) are d\mathbb{Z}^{d}-periodic.

For the integrability of E~(i,k)s\tilde{E}^{s}_{(i,k)}, in the case of diffeomorphism, the linear Anosov system A:ddA:\mathbb{R}^{d}\to\mathbb{R}^{d} is robustly dynamically coherent (see [21, Theorem 7.6] also [34, Proposition 3.2]) and this also holds for our case by the same reason of integrability for strong stable bundles. Hence, we have that

~(i,j)s=~(1,j)s~(i,k)s\tilde{\mathcal{F}}^{s}_{(i,j)}=\tilde{\mathcal{F}}^{s}_{(1,j)}\cap\tilde{\mathcal{F}}^{s}_{(i,k)}

is a foliation tangent to E~(i,j)s\tilde{E}^{s}_{(i,j)}. So E~(i,j)s\tilde{E}^{s}_{(i,j)} is integrable for all 1ijk1\leq i\leq j\leq k. We refer to [12, Lemma 6.1] for uniquely integrable property which is proved on the universal cover and also holds for non-invertible Anosov maps. It is clear that ~(i0,j0)s\tilde{\mathcal{F}}^{s}_{(i_{0},j_{0})} is subfoliated by ~(i,j)s\tilde{\mathcal{F}}^{s}_{(i,j)} for any 1i0ijj0k1\leq i_{0}\leq i\leq j\leq j_{0}\leq k. Moreover, ~(i0,i)s\tilde{\mathcal{F}}^{s}_{(i_{0},i)} and ~(i+1,j0)s\tilde{\mathcal{F}}^{s}_{(i+1,j_{0})} admit the Global Product Structure on ~(i0,j0)s\tilde{\mathcal{F}}^{s}_{(i_{0},j_{0})}.

For a small perturbation, in particular from (2.12), we have that the foliation ~(i,j)s\tilde{\mathcal{F}}^{s}_{(i,j)} is uniformly transverse to ~(1,i1)s~(j+1,k)s~u\tilde{\mathcal{L}}^{s}_{(1,i-1)}\oplus\tilde{\mathcal{L}}^{s}_{(j+1,k)}\oplus\tilde{\mathcal{L}}^{u}. By [3, Proposition 4 ], we have the quasi-isometric property for ~(i,j)s\tilde{\mathcal{F}}^{s}_{(i,j)}. We mention that the proof of this actually need ~(i,j)s\tilde{\mathcal{F}}^{s}_{(i,j)} has a uniformly transverse plane and it is uniformly continuous (see Remark 2.16).

For the case of dimEs=1E^{s}=1, since |HId||H-Id| is bounded, the unstable foliation ~u\tilde{\mathcal{F}}^{u} is uniformly bounded by ~u\tilde{\mathcal{L}}^{u}. Namely, there exists C0>0C_{0}>0 such that ~u(x)\tilde{\mathcal{F}}^{u}(x) and ~u(x)\tilde{\mathcal{L}}^{u}(x) are contained in the C0C_{0}-neighborhoods of each other, for all xdx\in\mathbb{R}^{d}. Fix n0dn_{0}\in\mathbb{Z}^{d} such that d(x,x+n0)3C0d(x,x+n_{0})\geq 3C_{0}, for all xdx\in\mathbb{R}^{d}. It follows that the Hausdorff distance between ~u(x)\tilde{\mathcal{F}}^{u}(x) and ~u(x+n0)\tilde{\mathcal{F}}^{u}(x+n_{0}) is bigger than C0C_{0}. Since the stable foliation ~s\tilde{\mathcal{F}}^{s} and the unstable foliation ~u\tilde{\mathcal{F}}^{u} admit the Global Product Structure, there exists L0L_{0} such that for any xdx\in\mathbb{R}^{d} with L(x)L0L(x)\leq L_{0}, one has ~s(x,L(x))\tilde{\mathcal{F}}^{s}(x,L(x)) intersects ~u(x+n0)\tilde{\mathcal{F}}^{u}(x+n_{0}) exactly once and the distance between xx and the intersection is bigger than C0C_{0}. Thus the one -dimensional foliation ~s\tilde{\mathcal{F}}^{s} is always quasi-isometric, whether ff is a small perturbation or not. We refer readers to [4] for more details.

Finally, we prove that HH preserves the weak stable foliations. Let y~s(x)y\in\tilde{\mathcal{F}}^{s}(x) and we always have H(y)~s(H(x))H(y)\in\tilde{\mathcal{L}}^{s}\big{(}H(x)\big{)}. Note that H(y)~(i,k)s(H(x))H(y)\in\tilde{\mathcal{L}}^{s}_{(i,k)}\big{(}H(x)\big{)} if and only if

d(An(Hy),An(Hx))(μis(A))nd(Hy,Hx),n.d\left(A^{-n}(Hy),A^{-n}(Hx)\right)\leq\left(\mu^{s}_{i}(A)\right)^{-n}\cdot d\left(Hy,Hx\right),\quad\forall n\in\mathbb{N}.

By Proposition 2.3, let |Hid|<C0|H-id|<C_{0}. One has that H(y)~(i,k)s(H(x))H(y)\in\tilde{\mathcal{L}}^{s}_{(i,k)}\big{(}H(x)\big{)} if and only if

d(Fn(y),Fn(x))(μis(A))nd(Hy,Hx)+2C0,n.\displaystyle d\left(F^{-n}(y),F^{-n}(x)\right)\leq\left(\mu^{s}_{i}(A)\right)^{-n}\cdot d\left(Hy,Hx\right)+2C_{0},\quad\forall n\in\mathbb{N}. (2.14)

It implies that H(y)~(i,k)s(H(x))H(y)\in\tilde{\mathcal{L}}^{s}_{(i,k)}(H(x)) if and only if y~(i,k)s(x)y\in\tilde{\mathcal{F}}^{s}_{(i,k)}(x). Indeed, if y~(i,k)s(x)y\notin\tilde{\mathcal{F}}^{s}_{(i,k)}(x), then there exists the unique point z~(1,i1)s(x)~(i,k)s(y)z\in\tilde{\mathcal{F}}^{s}_{(1,i-1)}(x)\cap\tilde{\mathcal{F}}^{s}_{(i,k)}(y) with a:=d~(1,i1)s(x,z)>0a:=d_{\tilde{\mathcal{F}}^{s}_{(1,i-1)}}(x,z)>0. Let b=d~(i,k)s(z,y)b=d_{\tilde{\mathcal{F}}^{s}_{(i,k)}}(z,y), note that bb may be zero. For nn\in\mathbb{N} big enough, one has

d(Fny,Fnz)d~(i,k)s(Fny,Fnz)(μis(A)ε02)nb,\displaystyle d(F^{-n}y,F^{-n}z)\leq d_{\tilde{\mathcal{F}}^{s}_{(i,k)}}(F^{-n}y,F^{-n}z)\leq\left(\mu^{s}_{i}(A)-\frac{\varepsilon_{0}}{2}\right)^{-n}\cdot b, (2.15)

where ε0\varepsilon_{0} is given by (2.12) and

d~(1,i1)s(Fnx,Fnz)(μi1s(A)+ε02)na.d_{\tilde{\mathcal{F}}^{s}_{(1,i-1)}}(F^{-n}x,F^{-n}z)\geq\left(\mu^{s}_{i-1}(A)+\frac{\varepsilon_{0}}{2}\right)^{-n}\cdot a.

Since ~(1,i1)s\tilde{\mathcal{F}}^{s}_{(1,i-1)} is quasi-isometric, there exists 0<C1<10<C_{1}<1 such that

d(Fnx,Fnz)C1(μi1s(A)+ε02)na.\displaystyle d(F^{-n}x,F^{-n}z)\geq C_{1}\left(\mu^{s}_{i-1}(A)+\frac{\varepsilon_{0}}{2}\right)^{-n}\cdot a. (2.16)

Hence by (2.15) and (2.16), one has

d(Fnx,Fny)C1(μi1s(A)+ε02)na(μis(A)ε02)nb,d(F^{-n}x,F^{-n}y)\geq C_{1}\left(\mu^{s}_{i-1}(A)+\frac{\varepsilon_{0}}{2}\right)^{-n}\cdot a-\left(\mu^{s}_{i}(A)-\frac{\varepsilon_{0}}{2}\right)^{-n}\cdot b,

which contradicts with (2.14). ∎

Remark 2.16.

We state the uniform continuity of foliation as follow. For given ~is,(1ik)\tilde{\mathcal{F}}^{s}_{i},(1\leq i\leq k) and two constants C>0C>0, there exists δ>0\delta>0 such that for every xdx\in\mathbb{R}^{d} and y~is(x)y\in\tilde{\mathcal{F}}^{s}_{i}(x) with d~is(x,y)>Cd_{\tilde{\mathcal{F}}^{s}_{i}}(x,y)>C, we have d(x,y)>δd(x,y)>\delta. Just note that, by the choice of the neighborhood 𝒰\mathcal{U}, the angle (E~is,L~is)\angle(\tilde{E}^{s}_{i},\tilde{L}^{s}_{i}) is uniformly bounded by α\alpha. In fact, for any Anosov map (may not be a small perturbation) with a dominated splitting along orbit, Txn𝕋d=E1s(xn,x~)Eks(xn,x~)Eu(xn,x~)T_{x_{n}}\mathbb{T}^{d}=E^{s}_{1}(x_{n},\tilde{x})\oplus...\oplus E^{s}_{k}(x_{n},\tilde{x})\oplus E^{u}(x_{n},\tilde{x}) the angle between any two distinct subbundles is uniformly away from 0, where dimEisE^{s}_{i} may bigger than one.

The following proposition says that the same periodic Lyapunov exponent implies it coincides with one of the linearization on the assumption that HH preserves the corresponding foliation.

Proposition 2.17.

Let f𝒰f\in\mathcal{U} given by Proposition 2.15. Fix 1ik1\leq i\leq k and suppose that H(~is)=~isH(\tilde{\mathcal{F}}^{s}_{i})=\tilde{\mathcal{L}}^{s}_{i} and λis(p,f)=λis(q,f)\lambda^{s}_{i}(p,f)=\lambda^{s}_{i}(q,f) for every p,qPer(f)p,q\in{\rm Per}(f). Then λis(p,f)=λis(A)\lambda^{s}_{i}(p,f)=\lambda^{s}_{i}(A), for all pPer(f)p\in{\rm Per}(f). Especially, for every Anosov map ff on torus with dimEs=1E^{s}=1, if λs(p,f)=λs(q,f)\lambda^{s}(p,f)=\lambda^{s}(q,f) for every p,qPer(f)p,q\in{\rm Per}(f), then λis(p,f)=λis(A)\lambda^{s}_{i}(p,f)=\lambda^{s}_{i}(A), for all pPer(f)p\in{\rm Per}(f).

Proof of Proposition 2.17.

Since λis(p,f)=λis(q,f)\lambda^{s}_{i}(p,f)=\lambda^{s}_{i}(q,f) for all p,qPerp,q\in\rm{Per}(f)(f), μ:=exp(λis(f,p))\mu:={\rm exp}\big{(}\lambda^{s}_{i}(f,p)\big{)} is a constant. We claim that there exists an adapted metric on d\mathbb{R}^{d}.

Claim 2.18.

For any δ>0\delta>0, there exists a smooth adapted Riemannian metric on TdT\mathbb{R}^{d} such that

μ(1+δ)1<DF|E~is(x)<μ(1+δ),xd.\mu\cdot(1+\delta)^{-1}<\|DF|_{\tilde{E}_{i}^{s}(x)}\|<\mu\cdot(1+\delta),\quad\forall x\in\mathbb{R}^{d}.
Proof of Claim 2.18.

Since we have the Shadowing Lemma for Anosov maps (see [1]), it can be proved as the existence of adapted metrics for Anosov diffeomorphisms. For the convenience of readers, we prove it as follow.

Fix in advance a Riemannian metric on TdT\mathbb{R}^{d} which induces a norm |||\cdot|. Let

μ+:=supxd|DF|E~is(x)|andμ:=infxd|DF|E~is(x)|.\mu_{+}:=\sup_{x\in\mathbb{R}^{d}}|DF|_{\tilde{E}^{s}_{i}(x)}|\quad{\rm and}\quad\mu_{-}:=\inf_{x\in\mathbb{R}^{d}}|DF|_{\tilde{E}^{s}_{i}(x)}|.

By (2.11) and the compactness of 𝕋fd\mathbb{T}^{d}_{f}, one has μ+<+\mu_{+}<+\infty and μ>0\mu_{-}>0. Moreover, since Eis(x0,x~)E^{s}_{i}(x_{0},\tilde{x}) is continuous with respect to x~=(xi)𝕋d\tilde{x}=(x_{i})\in\mathbb{T}^{d}, for any δ>0\delta>0, there exists α>0\alpha>0 such that

(1+δ/2)1|DF|E~is(x)||DF|E~is(y)|1+δ/2,\displaystyle(1+\delta/2)^{-1}\leq\frac{|DF|_{\tilde{E}^{s}_{i}(x)}|}{|DF|_{\tilde{E}^{s}_{i}(y)}|}\leq 1+\delta/2, (2.17)

for any x,ydx,y\in\mathbb{R}^{d} with d(x,y)<αd(x,y)<\alpha.

By the Shadowing Lemma, for given α>0\alpha>0, there exists β>0\beta>0 such that each periodic β\beta-pseudo-orbit in 𝕋d\mathbb{T}^{d} can be α\alpha-shadowing by a periodic orbit. Let B1,,Bn(β)B_{1},...,B_{n(\beta)} be finite many open β\beta-balls cover 𝕋d\mathbb{T}^{d}. Since ff is transtive (also see Section 4), there exists N1N_{1}\in\mathbb{N} such that for any BiB_{i} and BjB_{j}, BiB_{i} can intersect BjB_{j} within N1N_{1}-times iteration by ff.

Let π:d𝕋d\pi:\mathbb{R}^{d}\to\mathbb{T}^{d} be the natural projection. For any xdx\in\mathbb{R}^{d} and N0N_{0}\in\mathbb{N} with x0:=π(x)Bi0x_{0}:=\pi(x)\in B_{i_{0}} and fN0(x0)Bi1f^{N_{0}}(x_{0})\in B_{i_{1}}, there exists y0Bi1y_{0}\in B_{i_{1}} and N2[0,N1]N_{2}\in[0,N_{1}] such that fN2y0Bi0f^{N_{2}}y_{0}\in B_{i_{0}}. It follows that

{x0,f(x0),,fN01(x0),y0,f(y0),,fN21y0}\big{\{}x_{0},f(x_{0}),...,f^{N_{0}-1}(x_{0}),y_{0},f(y_{0}),...,f^{N_{2}-1}y_{0}\big{\}}

is a periodic β\beta-pseudo-orbit and is α\alpha-shadowing by a periodic orbit p0𝕋dp_{0}\in\mathbb{T}^{d} with period N(p0):=N0+N2N(p_{0}):=N_{0}+N_{2}. Hence, there exists pπ1(p0)p\in\pi^{-1}(p_{0}) such that d(π(Fjx),π(Fjp))<αd\left(\pi(F^{j}x),\pi(F^{j}p)\right)<\alpha, for all j[0,N0]j\in[0,N_{0}].

Now, let N=N0+N1N=N_{0}+N_{1} and

νN:=n=0N1|DFn|E~is(x)ν|1N,νE~is(x).\|\nu\|_{N}:=\prod_{n=0}^{N-1}|DF^{n}|_{\tilde{E}^{s}_{i}(x)}\nu|^{\frac{1}{N}},\quad\forall\nu\in\tilde{E}^{s}_{i}(x).

Note that |DFN(p0)|E~is(p)|=μN(p0)|DF^{N(p_{0})}|_{\tilde{E}^{s}_{i}(p)}|=\mu^{N(p_{0})}, one has

|DFN0|E~is(p)|[μN0(μμ+)N2,μN0(μμ)N2].\displaystyle|DF^{N_{0}}|_{\tilde{E}^{s}_{i}(p)}|\in\left[\mu^{N_{0}}\cdot\left(\frac{\mu}{\mu_{+}}\right)^{N_{2}}\;,\;\mu^{N_{0}}\cdot\left(\frac{\mu}{\mu_{-}}\right)^{N_{2}}\right]. (2.18)

Taking νE~is(x){0}\nu\in\tilde{E}^{s}_{i}(x)-\{0\}, by (2.17) and (2.18), we calculate directly,

DxF(ν)NνN\displaystyle\frac{\|D_{x}F(\nu)\|_{N}}{\|\nu\|_{N}} =|DxFN(ν)|1/N|ν|1/N\displaystyle=\frac{|D_{x}F^{N}(\nu)|^{1/N}}{|\nu|^{1/N}}
=(|DFN0(x0)fN1Dx0FN0(ν)||ν|)1/N\displaystyle=\left(\frac{\left|D_{F^{N_{0}}(x_{0})}f^{N_{1}}\circ D_{x_{0}}F^{N_{0}}(\nu)\right|}{|\nu|}\right)^{1/N}
[μN1/N(μ1+δ/2)N0/N(μμ+)N2/N,μ+N1/N(μ(1+δ/2))N0/N(μμ)N2/N].\displaystyle\in\left[\mu_{-}^{N_{1}/N}\cdot\left(\frac{\mu}{1+\delta/2}\right)^{N_{0}/N}\cdot\left(\frac{\mu}{\mu_{+}}\right)^{N_{2}/N}\;,\;\mu_{+}^{N_{1}/N}\cdot\big{(}\mu\cdot(1+\delta/2)\big{)}^{N_{0}/N}\cdot\left(\frac{\mu}{\mu_{-}}\right)^{N_{2}/N}\right].

Hence, there exists N0N_{0} big enough such that

DxF(ν)NνN(μ(1+δ)1,μ(1+δ)).\frac{\|D_{x}F(\nu)\|_{N}}{\|\nu\|_{N}}\in\big{(}\mu\cdot(1+\delta)^{-1},\mu\cdot(1+\delta)\big{)}.

There exists a norm \|\cdot\| whose restriction on subbundle E~is\tilde{E}^{s}_{i} is N\|\cdot\|_{N} gives the smooth adapted Riemannian metric we want. ∎

Assume that μ|μis(A)|=exp(λis(A))\mu\neq|\mu^{s}_{i}(A)|={\rm exp}\big{(}\lambda^{s}_{i}(A)\big{)}. Fix δ<min{|μμis(A)1|,|μis(A)μ1|}\delta<{\rm min}\big{\{}|\frac{\mu}{\mu^{s}_{i}(A)}-1|,|\frac{\mu^{s}_{i}(A)}{\mu}-1|\big{\}} and an adapted norm \|\cdot\| from Claim 2.18. Let H:ddH:\mathbb{R}^{d}\to\mathbb{R}^{d} be the conjugacy defined in Proposition 2.3 satisfying |HId|C0|H-Id|\leq C_{0}. We take two points x,ydx,y\in\mathbb{R}^{d} such that y~is(x)y\in\tilde{\mathcal{F}}^{s}_{i}(x). One has

μk(1+δ)kds(x,y)ds(Fkx,Fky)μk(1+δ)kds(x,y),\mu^{-k}(1+\delta)^{-k}\cdot d^{s}(x,y)\leq d^{s}(F^{-k}x,F^{-k}y)\leq\mu^{-k}(1+\delta)^{k}\cdot d^{s}(x,y),

further,

aμk(1+δ)kds(x,y)d(Fkx,Fky)μk(1+δ)kds(x,y),\displaystyle a\cdot\mu^{-k}(1+\delta)^{-k}\cdot d^{s}(x,y)\leq d(F^{-k}x,F^{-k}y)\leq\mu^{-k}(1+\delta)^{k}\cdot d^{s}(x,y), (2.19)

where aa is given by (2.13)\eqref{defquasi-isometric}, since ~is\tilde{\mathcal{F}}^{s}_{i} is quasi-isometric (by Proposition 2.15). Meanwhile, since H(~is)=~isH(\tilde{\mathcal{F}}^{s}_{i})=\tilde{\mathcal{L}}^{s}_{i} ( when dimEs=1E^{s}=1, H(~s)=~sH(\tilde{\mathcal{F}}^{s})=\tilde{\mathcal{L}}^{s} always holds),

d(H(Fkx),H(Fkx))=d(Ak(Hx),Ak(Hx))=(μis(A))kd(Hx,Hy).\displaystyle d\left(H(F^{-k}x),H(F^{-k}x)\right)=d\left(A^{-k}(Hx),A^{-k}(Hx)\right)=\left(\mu^{s}_{i}(A)\right)^{-k}\cdot d\left(Hx,Hy\right). (2.20)

The formulas (2.19) and (2.20) jointly contradict with the fact

|d(Fkx,Fky)d(H(Fkx),H(Fkx))|2C0.\Big{|}d(F^{-k}x,F^{-k}y)-d\left(H(F^{-k}x),H(F^{-k}x)\right)\Big{|}\leq 2C_{0}.

For the convenience of readers, we state Journe´\acute{\rm e} Lemma [22] as the following proposition which will be useful in Section 4 and Section 5.

Proposition 2.19 ( [22]).

Let Mi(i=1,2)M_{i}\;(i=1,2) be a smooth manifold and is,iu\mathcal{F}_{i}^{s},\mathcal{F}_{i}^{u} be continuous transverse foliations on MiM_{i} with uniformly Cr+αC^{r+\alpha}-smooth leaves (r1,0<α<1)(r\geq 1,0<\alpha<1). Assume that h:M1M2h:M_{1}\to M_{2} is a homeomorphism and maps 1σ\mathcal{F}_{1}^{\sigma} to 2σ(σ=s,u)\mathcal{F}^{\sigma}_{2}(\sigma=s,u). If hh restricted on leaves of both 1s\mathcal{F}_{1}^{s} and 1u\mathcal{F}_{1}^{u} is uniformly Cr+αC^{r+\alpha}, then hh is Cr+αC^{r+\alpha}-smooth.

3 Spectral rigidity on stable bundle

In this section, we prove the necessary parts of both Theorem 1.1 and Theorem 1.2. As mentioned, we can actually prove them under C1C^{1} assumption. For convenience, we restate these as follow.

Theorem 3.1.

Let A:𝕋d𝕋dA:\mathbb{T}^{d}\to\mathbb{T}^{d} be an irreducible linear non-invertible Anosov map. Assume that AA admits the finest (on stable bundle) dominated splitting,

T𝕋d=L1sL2sLksLu,T\mathbb{T}^{d}=L^{s}_{1}\oplus L^{s}_{2}\oplus...\oplus L^{s}_{k}\oplus L^{u},

where dimLis=1{\rm dim}L^{s}_{i}=1, 1ik1\leq i\leq k.

Then there exists a C1C^{1} neighborhood 𝒰C1(𝕋d)\mathcal{U}\subset C^{1}(\mathbb{T}^{d}) of AA such that for every f𝒰f\in\mathcal{U}, if ff is special, then λis(p,f)=λis(A)\lambda^{s}_{i}(p,f)=\lambda^{s}_{i}(A), for all pPer(f)p\in{\rm Per}(f) and all 1ik1\leq i\leq k. Moreover, ff admits the finest (on stable bundle) dominated splitting,

T𝕋d=E1sE2sEksEu,T\mathbb{T}^{d}=E^{s}_{1}\oplus E^{s}_{2}\oplus...\oplus E^{s}_{k}\oplus E^{u},

where dimEis=1E^{s}_{i}=1, for all 1ik1\leq i\leq k.

Moreover, when k=1k=1, for every C1C^{1}-smooth non-invertible Anosov map ff with irreducible linearization AA, if ff is special, then λs(p,f)=λs(A)\lambda^{s}(p,f)=\lambda^{s}(A), for all pPer(f)p\in{\rm Per}(f).

Now, we give the scheme of our proof. In this section, we always assume that A:𝕋d𝕋dA:\mathbb{T}^{d}\to\mathbb{T}^{d} satisfies the condition of Theorem 3.1. To get the spectral rigidity on stable bundle , we firstly prove that every periodic point pPer(f)p\in{\rm Per}(f) has the same stable Lyapunov spectrum {λis(p,f):i=1,2,,k}\big{\{}\lambda^{s}_{i}(p,f):i=1,2,...,k\big{\}}.

Proposition 3.1.

Let f:𝕋d𝕋df:\mathbb{T}^{d}\to\mathbb{T}^{d} be an irreducible non-invertible Anosov map with a DfDf-invariant one-dimensional subbundle EisEsE^{s}_{i}\subset E^{s}. If ff is special and there exists an ff-invariant foliation on 𝕋d\mathbb{T}^{d} tangent to EisE^{s}_{i}, then λis(p,f)=λis(q,f)\lambda^{s}_{i}(p,f)=\lambda^{s}_{i}(q,f), for all p,qPer(f)p,q\in{\rm Per}(f), where λis(p,f)\lambda^{s}_{i}(p,f) is the Lyapunov exponent of ff for pp corresponding the bundle EisE^{s}_{i}.

We emphasize here that in the proof of Proposition 3.1, ff need not be a small perturbation of AA. To get that every periodic point of ff has the same stable Lyapunov spectrum through Proposition 3.1, we need that ff admits the finest (on stable bundle) dominated splitting.

Proposition 3.2.

There exists a C1C^{1} neighborhood 𝒰C1(𝕋d)\mathcal{U}\subset C^{1}(\mathbb{T}^{d}) of AA such that for every f𝒰f\in\mathcal{U}, if it is special, then it admits the finest (on stable bundle) dominated splitting

T𝕋d=E1sE2sEksEu,T\mathbb{T}^{d}=E^{s}_{1}\oplus E^{s}_{2}\oplus...\oplus E^{s}_{k}\oplus E^{u},

where EisE^{s}_{i} is one-dimensional and integrable, for all 1ik1\leq i\leq k.

We leave the proofs of Proposition 3.1 and Proposition 3.2 in subsection 3.1.

To obtain the relationship between the periodic stable Lyapunov spectrum and one of its linearization, we can use Proposition 2.17. For a special f𝒰f\in\mathcal{U} given by Proposition 3.2, let hh be the conjugacy between ff and AA given by Proposition 2.2. We need to prove that the conjugacy hh is also a leaf conjugacy between is\mathcal{F}^{s}_{i} and is\mathcal{L}^{s}_{i}.

Proposition 3.3.

Let f𝒰f\in\mathcal{U} given by Proposition 3.2 be special. Then h(is)=ish(\mathcal{F}^{s}_{i})=\mathcal{L}^{s}_{i}, for every 1ik1\leq i\leq k.

We leave the proof of Proposition 3.3 in subsection 3.2. Now, we can prove Theorem 3.1.

Proof of Theorem 3.1.

For one-dimensional stable bundle case, the special Anosov map ff admits the dominated splitting T𝕋d=EsEuT\mathbb{T}^{d}=E^{s}\oplus E^{u} and H(~s)=~sH(\tilde{\mathcal{F}}^{s})=\tilde{\mathcal{L}}^{s} always holds whether ff is a small perturbation of its linearization or not. Hence by Proposition 2.17 and Proposition 3.1, we get λs(p,f)=λs(A)\lambda^{s}(p,f)=\lambda^{s}(A), for all pPer(f)p\in{\rm Per}(f), immediately.

For higher-dimensional stable bundle case, by Proposition 3.2, there exists a C1C^{1} neighborhood 𝒰\mathcal{U} of AA such that every special f𝒰f\in\mathcal{U} admits the finest (on stable bundle) dominated splitting. Thus by Proposition 3.1, every periodic point pPer(f)p\in{\rm Per}(f) has the same stable Lyapunov spectrum. Now, combining Proposition 2.17 and Proposition 3.3, we have that λis(p,f)=λis(A)\lambda^{s}_{i}(p,f)=\lambda^{s}_{i}(A), for every pPer(f)p\in{\rm Per}(f) and every 1ik1\leq i\leq k. ∎

3.1 Periodic stable Lyapunov spectrums coincide

In this subsection, we prove Proposition 3.1 and Proposition 3.2. Fix 1ik1\leq i\leq k, let f:𝕋d𝕋df:\mathbb{T}^{d}\to\mathbb{T}^{d} be a special irreducible non-invertible Anosov map with DfDf-invariant subbundle EisEsE^{s}_{i}\subset E^{s}. Let is\mathcal{F}^{s}_{i} be an ff-invariant integral foliation for EisE^{s}_{i}. For short, we denote μis(p,f):=exp(λis(p,f))\mu^{s}_{i}(p,f):={\rm exp}\left(\lambda^{s}_{i}(p,f)\right) by μis(p)\mu^{s}_{i}(p), for all pPer(f)p\in{\rm Per}(f).

Proof of Proposition 3.1.

We assume that there exist p,qPerp,q\in\rm{Per}(f)(f) such that μis(p)<μis(q)\mu^{s}_{i}(p)<\mu^{s}_{i}(q), then to get a contradiction. By the assumption of the existence of different periodic stable Lyapunov exponents, the infimum μ\mu_{-} and the supremum μ+\mu_{+} of the set {μis(p):pPer(f)}\big{\{}\mu_{i}^{s}(p):p\in{\rm Per}(f)\big{\}} satisfy 0<μ<μ+<10<\mu_{-}<\mu_{+}<1. Given δ>0\delta>0 arbitrarily small, we can choose two periodic points p,qp,q of ff such that

μis(p)μ(1+δ)andμis(q)μ+(1+δ)1,\mu^{s}_{i}(p)\leq\mu_{-}\cdot(1+\delta)\quad{\rm{and}}\quad\mu^{s}_{i}(q)\geq{\mu_{+}}\cdot(1+\delta)^{-1},

Moreover, as Claim 2.18, there exists a smooth adapted Riemannian metric such that

μ(1+δ)1<Df|Eis(x)<μ+(1+δ),x𝕋d.\mu_{-}\cdot(1+\delta)^{-1}<\|Df|_{E_{i}^{s}(x)}\|<\mu_{+}\cdot(1+\delta),\quad\forall x\in\mathbb{T}^{d}.

For convenience, we can assume that p,qp,q are both fixed points. Otherwise we can go through the rest of this proof by using fn0f^{n_{0}} instead of ff, where n0n_{0} is the minimal common period of pp and qq. Let η0>0\eta_{0}>0 small enough such that, for any x1,x2𝕋dx_{1},x_{2}\in\mathbb{T}^{d} with d(x1,x2)η0d(x_{1},x_{2})\leq\eta_{0}, we have

(1+δ)1Df|Eis(x1)Df|Eis(x2)1+δ.(1+\delta)^{-1}\leq\frac{\|Df|_{E^{s}_{i}(x_{1})}\|}{\|Df|_{E^{s}_{i}(x_{2})}\|}\leq 1+\delta.

Fix ε>0{\varepsilon}>0(εη0)({\varepsilon}\ll\eta_{0}), there exists xεx_{{\varepsilon}} in the ε{\varepsilon}-Ball Bε(q)B_{{\varepsilon}}(q) and kε=k(ε,xε)>0k_{{\varepsilon}}=k({\varepsilon},x_{{\varepsilon}})>0 such that fkε(xε)=pf^{k_{{\varepsilon}}}(x_{{\varepsilon}})=p. Indeed, since ff is special, the preimage set of pp for ff is dense by Proposition 2.2 and Proposition 2.10.

Shrinking ε{\varepsilon}, by the local product structure, there exist η1,η2>0\eta_{1},\eta_{2}>0 such that the local unstable leaf u(q,η1)Bη0(q)\mathcal{F}^{u}(q,\eta_{1})\subset B_{\eta_{0}}(q) intersects with the local stable leaf s(xε,η2)Bη0(q)\mathcal{F}^{s}(x_{{\varepsilon}},\eta_{2})\subset B_{\eta_{0}}(q) at the unique point yεy_{{\varepsilon}}, namely, yε=s(xε,η2)u(q,η1)y_{{\varepsilon}}=\mathcal{F}^{s}(x_{{\varepsilon}},\eta_{2})\cap\mathcal{F}^{u}(q,\eta_{1}). Note that one has du(yε,q)d(ε)d_{\mathcal{F}^{u}}(y_{{\varepsilon}},q)\leq d({\varepsilon}), where d(ε)d({\varepsilon}) tends to 0 as ε{\varepsilon} goes to 0.

Therefore, we can choose a point zεis(xε,η2)z_{{\varepsilon}}\in\mathcal{F}^{s}_{i}(x_{{\varepsilon}},\eta_{2}) such that dis(xε,zε)η2/3d_{\mathcal{F}^{s}_{i}}(x_{{\varepsilon}},z_{{\varepsilon}})\geq\eta_{2}/3. We denote by IεI_{{\varepsilon}} the curve in is(xε)\mathcal{F}^{s}_{i}(x_{{\varepsilon}}) from xεx_{{\varepsilon}} to zεz_{{\varepsilon}}. Since xεx_{{\varepsilon}} is a kεk_{{\varepsilon}}-preimage of the fixed point pp, we can find a curve JεJ_{{\varepsilon}} in is(p)\mathcal{F}^{s}_{i}(p) such that

fkε(Jε)=fkε(Iε).f^{k_{{\varepsilon}}}(J_{{\varepsilon}})=f^{k_{{\varepsilon}}}(I_{{\varepsilon}}).

Let NεN_{{\varepsilon}} be the maximal positive integer such that d(fj(w),fj(q))<η0d(f^{j}(w),f^{j}(q))<\eta_{0}, for all wIεw\in I_{{\varepsilon}} and j[0,Nε]j\in[0,N_{{\varepsilon}}]. Let kεk_{{\varepsilon}} be the minimal positive integer such that {x𝕋d|fkε=p}Bε(q)\big{\{}x\in\mathbb{T}^{d}|f^{k_{{\varepsilon}}}=p\big{\}}\cap B_{{\varepsilon}}(q)\neq\emptyset.

Claim 3.4.

There exist ε0>0{\varepsilon}_{0}>0 and C0>0C_{0}>0 such that

NεkεC0,εε0.\frac{N_{{\varepsilon}}}{k_{{\varepsilon}}}\geq C_{0},\quad\forall{\varepsilon}\leq{\varepsilon}_{0}.

We estimate the upper bound and the lower bound of kεk_{{\varepsilon}} and NεN_{{\varepsilon}}, respectively. Note that we can get the lower bound of NεN_{{\varepsilon}} by controlling the distance of fNε(yε)f^{N_{{\varepsilon}}}(y_{{\varepsilon}}) and fNε(q)f^{N_{{\varepsilon}}}(q) along unstable leaves directly. However, it is difficult to estimate the upper bound of kεk_{{\varepsilon}} under the dynamics of ff, while it is convient in linear systems (see Proposition 2.10). So, by Proposition 2.2, let ff conjugate to its linearization AA. We calculate the ”NεN_{{\varepsilon}}” and ”kεk_{{\varepsilon}}” of AA. A direct way to get Claim 3.4 is using the Ho¨\ddot{\rm o}lder continuity of hh, but here we prove it by only uniform continuity. See Figure 1.

Refer to caption
Figure 1: Iteration of local stable leaves in the case of one-dimensional stable bundle
Proof of Claim3.4.

Let h:𝕋d𝕋dh:\mathbb{T}^{d}\to\mathbb{T}^{d} be the conjugacy between ff and AA with hf=Ahh\circ f=A\circ h. For short, denote h(xε),h(yε),h(p)h(x_{{\varepsilon}}),h(y_{{\varepsilon}}),h(p) and h(q)h(q) by x~ε,y~ε,p~\tilde{x}_{{\varepsilon}},\tilde{y}_{{\varepsilon}},\tilde{p} and q~\tilde{q}, respectively. The homeomorphism hh maps Bη0(q)B_{\eta_{0}}(q) and Bε(q)B_{{\varepsilon}}(q) to be two neighborhoods of q~\tilde{q} such that we can choose two su-foliation boxes of AA, B~η~0(q~)h(Bη0(q))\tilde{B}_{\tilde{\eta}_{0}}(\tilde{q})\subset h(B_{\eta_{0}}(q)) and B~ε~(q~)h(Bε(q))\tilde{B}_{\tilde{{\varepsilon}}}(\tilde{q})\subset h(B_{{\varepsilon}}(q)). Here, η~0\tilde{\eta}_{0} is fixed by η0\eta_{0}, while ε~\tilde{{\varepsilon}} tends to 0 following ε{\varepsilon}. Thus we can shorten IεI_{{\varepsilon}} sligtly such that Iε~:=h(Iε)B~ε~(q~)s(x~ε)\tilde{I_{{\varepsilon}}}:=h(I_{{\varepsilon}})\subset\tilde{B}_{\tilde{{\varepsilon}}}(\tilde{q})\cap\mathcal{L}^{s}(\tilde{x}_{{\varepsilon}}) and the length |Iε~|=η~03|\tilde{I_{{\varepsilon}}}|=\frac{\tilde{\eta}_{0}}{3}.

Let Nε~\tilde{N_{{\varepsilon}}} be the maximal positive integer such that d(Aj(w~),Aj(q~))<η~0d(A^{j}(\tilde{w}),A^{j}(\tilde{q}))<\tilde{\eta}_{0}, for all w~Iε~\tilde{w}\in\tilde{I_{{\varepsilon}}} and j[0,Nε~]j\in[0,\tilde{N_{{\varepsilon}}}]. Let k~ε\tilde{k}_{{\varepsilon}} be the minimal positive integer such that {x𝕋d|Ak~ε=p~}B~ε~(q~)\big{\{}x\in\mathbb{T}^{d}|A^{\tilde{k}_{{\varepsilon}}}=\tilde{p}\big{\}}\cap\tilde{B}_{\tilde{{\varepsilon}}}(\tilde{q})\neq\emptyset, where p~:=h(p)\tilde{p}:=h(p). It is clear that NεNε~N_{{\varepsilon}}\geq\tilde{N_{{\varepsilon}}} and kεk~εk_{{\varepsilon}}\leq\tilde{k}_{{\varepsilon}}. So, we get Nε/kεNε~/k~εN_{{\varepsilon}}/k_{{\varepsilon}}\geq\tilde{N_{{\varepsilon}}}/\tilde{k}_{{\varepsilon}}.

For every w~Iε~\tilde{w}\in\tilde{I_{{\varepsilon}}} and j[0,Nε~]j\in[0,\tilde{N_{{\varepsilon}}}], we have

d(Aj(w~),Aj(q~))\displaystyle d\left(A^{j}(\tilde{w}),A^{j}(\tilde{q})\right) ds(Aj(w~),Aj(y~ε))+du(Aj(y~ε),Aj(q~))\displaystyle\leq d_{\mathcal{L}^{s}}\left(A^{j}(\tilde{w}),A^{j}(\tilde{y}_{{\varepsilon}})\right)+d_{\mathcal{L}^{u}}\left(A^{j}(\tilde{y}_{{\varepsilon}}),A^{j}(\tilde{q})\right)
η~03+d(x~ε,y~ε)+ε~(μmaxu(A))j.\displaystyle\leq\frac{\tilde{\eta}_{0}}{3}+d(\tilde{x}_{{\varepsilon}},\tilde{y}_{{\varepsilon}})+\tilde{{\varepsilon}}\cdot\left(\mu_{max}^{u}(A)\right)^{j}.

Note that as ε{\varepsilon} small enough, we have d(x~ε,y~ε)η~06d(\tilde{x}_{{\varepsilon}},\tilde{y}_{{\varepsilon}})\leq\frac{\tilde{\eta}_{0}}{6}. So, the maximal positive integer Nε~\tilde{N_{{\varepsilon}}} such that

η~03+d(x~ε,y~ε)+ε~(μmaxu(A))jη~02+ε~(μmaxu(A))jη~0,\frac{\tilde{\eta}_{0}}{3}+d(\tilde{x}_{{\varepsilon}},\tilde{y}_{{\varepsilon}})+\tilde{{\varepsilon}}\cdot\left(\mu_{max}^{u}(A)\right)^{j}\leq\frac{\tilde{\eta}_{0}}{2}+\tilde{{\varepsilon}}\cdot\left(\mu_{max}^{u}(A)\right)^{j}\leq\tilde{\eta}_{0},

holds should satisfy

Nε~lnη~0ln(2ε~)lnμmaxu(A).\displaystyle\tilde{N_{{\varepsilon}}}\geq\frac{{\rm ln}\tilde{\eta}_{0}-{\rm ln}(2\tilde{{\varepsilon}})}{{\rm ln}\mu_{\rm max}^{u}(A)}.

On the other hand, by Proposition 2.10, there exists C>0C>0 such that for every ε~>0\tilde{{\varepsilon}}>0,

k~εdlnClnε~ln|detA|.\displaystyle\tilde{k}_{{\varepsilon}}\leq d\cdot\frac{{\rm ln}C-{\rm ln}\tilde{{\varepsilon}}}{{\rm ln}|{\rm det}A|}.

Thus, as ε~\tilde{{\varepsilon}} tending to 0, there exists C0>0C_{0}>0 such that,

NεkεNε~k~εC0,\displaystyle\frac{N_{{\varepsilon}}}{k_{{\varepsilon}}}\geq\frac{\tilde{N_{{\varepsilon}}}}{\tilde{k}_{{\varepsilon}}}\geq C_{0},

where C0C_{0} could be close to 1dln|detA|lnμmaxu(A)\frac{1}{d}\cdot\frac{\rm{ln}|\rm{det}A|}{{\rm ln}\mu_{\rm max}^{u}(A)} arbitrarily. ∎

Now, using the uniform lower bound of the time ratio Nε/kεN_{{\varepsilon}}/k_{{\varepsilon}} of IεI_{{\varepsilon}} being around qq to reaching pp, we can get an exponential error between |fkε(Iε)||f^{k_{{\varepsilon}}}(I_{{\varepsilon}})| and |fkε(Jε)||f^{k_{{\varepsilon}}}(J_{{\varepsilon}})|.

Firstly, we claim that there exists C2>1C_{2}>1 such that

|Iε||Jε|[C21,C2],εη0.\displaystyle\frac{|I_{{\varepsilon}}|}{|J_{{\varepsilon}}|}\in\big{[}C_{2}^{-1},C_{2}\big{]},\quad\forall{\varepsilon}\ll\eta_{0}.

Indeed, by the construction of JεJ_{{\varepsilon}}, we have hfkε(Jε)=hfkε(Iε)h\circ f^{k_{{\varepsilon}}}(J_{{\varepsilon}})=h\circ f^{k_{{\varepsilon}}}(I_{{\varepsilon}}), equivalently, Akε(Jε~)=Akε(Iε~)A^{k_{{\varepsilon}}}(\tilde{J_{{\varepsilon}}})=A^{k_{{\varepsilon}}}(\tilde{I_{{\varepsilon}}}). This implies that Jε~\tilde{J_{{\varepsilon}}} is just a translation of Iε~\tilde{I_{{\varepsilon}}}, thus |Jε~|=|Iε~|η~02|\tilde{J_{{\varepsilon}}}|=|\tilde{I_{{\varepsilon}}}|\geq\frac{\tilde{\eta}_{0}}{2}. Since h1h^{-1} is uniformly continuous, by the uniform continuity for is\mathcal{F}^{s}_{i} (see Remark 2.16), we have that |Iε||Jε|\frac{|I_{{\varepsilon}}|}{|J_{{\varepsilon}}|} is uniformly bounded away from 0.

Note that we can assume JεBη0(p)J_{{\varepsilon}}\subset B_{\eta_{0}}(p). Otherwise, by the uniform continuity of hh again, we can shorten the length of IεI_{{\varepsilon}}, meanwhile ensure that the length is independent of ε{\varepsilon}.

Since we have assumed that p,qp,q are fixed points of ff, then fj(Jε)Bη0(p)f^{j}(J_{{\varepsilon}})\subset B_{\eta_{0}}(p), for every j0j\geq 0. So,

|fkε(Jε)|(μ(1+δ)2)kε|Jε|.\displaystyle\big{|}f^{k_{{\varepsilon}}}(J_{{\varepsilon}})\big{|}\leq\big{(}\mu_{-}\cdot(1+\delta)^{2}\big{)}^{k_{{\varepsilon}}}\big{|}J_{{\varepsilon}}\big{|}.

And,

|fkε(Iε)|\displaystyle|f^{k_{{\varepsilon}}}(I_{{\varepsilon}})| =|fkεNεfNε(Iε)|\displaystyle=|f^{k_{{\varepsilon}}-N_{{\varepsilon}}}\circ f^{N_{{\varepsilon}}}(I_{{\varepsilon}})|
(μ1+δ)kεNε(μ+(1+δ)2)Nε|Iε|.\displaystyle\geq\left(\frac{\mu_{-}}{1+\delta}\right)^{k_{{\varepsilon}}-N_{{\varepsilon}}}\cdot\left(\frac{\mu_{+}}{(1+\delta)^{2}}\right)^{N_{{\varepsilon}}}|I_{{\varepsilon}}|.

Consquently,

|fkε(Iε)||fkε(Jε)|\displaystyle\frac{|f^{k_{{\varepsilon}}}(I_{{\varepsilon}})|}{|f^{k_{{\varepsilon}}}(J_{{\varepsilon}})|} 1(1+δ)3kε+Nε(μ+μ)Nε|Iε||Jε|1(1+δ)4kε(μ+μ)Nε1C2\displaystyle\geq\frac{1}{(1+\delta)^{3k_{{\varepsilon}}+N_{{\varepsilon}}}}\cdot\left(\frac{\mu_{+}}{\mu_{-}}\right)^{N_{{\varepsilon}}}\cdot\frac{|I_{{\varepsilon}}|}{|J_{{\varepsilon}}|}\geq\frac{1}{(1+\delta)^{4k_{{\varepsilon}}}}\cdot\left(\frac{\mu_{+}}{\mu_{-}}\right)^{N_{{\varepsilon}}}\cdot\frac{1}{C_{2}}
1C2(1+δ)4kε(μ+μ)C0kε=1C2((1+δ)4(μ+μ)C0)kε.\displaystyle\geq\frac{1}{C_{2}\cdot(1+\delta)^{4k_{{\varepsilon}}}}\cdot\left(\frac{\mu_{+}}{\mu_{-}}\right)^{C_{0}\cdot k_{{\varepsilon}}}=\frac{1}{C_{2}}\cdot\Big{(}(1+\delta)^{4}\cdot\left(\frac{\mu_{+}}{\mu_{-}}\right)^{C_{0}}\Big{)}^{k_{{\varepsilon}}}.

We can assume that kεNεk_{{\varepsilon}}\geq N_{{\varepsilon}}, so that we have the second inequality. Otherwise, when kε<Nεk_{{\varepsilon}}<N_{{\varepsilon}}, the whole estimation of |fkε(Iε)||fkε(Jε)|\frac{|f^{k_{{\varepsilon}}}(I_{{\varepsilon}})|}{|f^{k_{{\varepsilon}}}(J_{{\varepsilon}})|} is trivial.

Let δ\delta is small enough such that (1+δ)4(μ+μ)C0>1(1+\delta)^{4}\cdot\left(\frac{\mu_{+}}{\mu_{-}}\right)^{C_{0}}>1, after that we can fix η0\eta_{0}. Let ε{\varepsilon} tend to zero. We have kε+k_{{\varepsilon}}\to+\infty, hence |fkε(Iε)||fkε(Jε)|+\frac{|f^{k_{{\varepsilon}}}(I_{{\varepsilon}})|}{|f^{k_{{\varepsilon}}}(J_{{\varepsilon}})|}\to+\infty. This contradics the fact that fkε(Iε)=fkε(Jε)f^{k_{{\varepsilon}}}(I_{{\varepsilon}})=f^{k_{{\varepsilon}}}(J_{{\varepsilon}}). ∎

Now, we show that there exists a C1C^{1} neighbohood 𝒰\mathcal{U} of AA in which every special ff admits the finest (on stable bundle) dominated splitting. Combining with Proposition 3.1, if f𝒰f\in\mathcal{U} is special, then λis(p,f)=λis(q,f)\lambda^{s}_{i}(p,f)=\lambda^{s}_{i}(q,f), for all p,qPer(f)p,q\in{\rm Per}(f) and 1ik1\leq i\leq k.

Proof of Proposition 3.2.

Since Eis=E(1,i)sE(i,k)sE^{s}_{i}=E^{s}_{(1,i)}\cap E^{s}_{(i,k)}, it suffices to prove the following lemma.

Lemma 3.5.

There exists a C1C^{1} neighborhood 𝒰C1(𝕋d)\mathcal{U}\subset C^{1}(\mathbb{T}^{d}) of AA such that, for every f𝒰f\in\mathcal{U} and 1ik11\leq i\leq k-1, if ff is special, then it admits the following dominated splitting

E(1,i)sE(i+1,k)sEu,E^{s}_{(1,i)}\oplus E^{s}_{(i+1,k)}\oplus E^{u},

and E(i+1,k)sE^{s}_{(i+1,k)} is integrable.

Proof of Lemma 3.5.

By Proposition 2.12, E(1,i)sE^{s}_{(1,i)} is well defined on 𝕋d\mathbb{T}^{d}. By the assumption that ff is special, the unstable bundle EuE^{u} is also well defined on 𝕋d\mathbb{T}^{d}.

Let FF be a lifting of ff and HH be the conjugacy between FF and AA. As the proof of Proposition 2.15, there exists a C1C^{1} neighborhood 𝒰C1(𝕋d)\mathcal{U}\subset C^{1}(\mathbb{T}^{d}) of AA such that for every f𝒰f\in\mathcal{U}, its lifting FF admits a dominated splitting

E~(1,i)s<E~(i+1,k)s<E~u,\tilde{E}^{s}_{(1,i)}\oplus_{<}\tilde{E}^{s}_{(i+1,k)}\oplus_{<}\tilde{E}^{u},

and E~(i+1,k)s\tilde{E}^{s}_{(i+1,k)} is integrable. Moreover, by the forth item of Proposition 2.15, H(y)~(i+1,k)s(H(x))H(y)\in\tilde{\mathcal{L}}^{s}_{(i+1,k)}(H(x)) if and only if y~(i+1,k)s(x)y\in\tilde{\mathcal{F}}^{s}_{(i+1,k)}(x). Note that, since ff is special, H(x+n)=H(x)+nH(x+n)=H(x)+n, for every xdx\in\mathbb{R}^{d} and ndn\in\mathbb{Z}^{d}. Thus, we have that

y+n~(i+1,k)s(x+n)\displaystyle y+n\in\tilde{\mathcal{F}}^{s}_{(i+1,k)}(x+n) H(y+n)~(i+1,k)s(H(x+n)),\displaystyle\Longleftrightarrow H(y+n)\in\tilde{\mathcal{L}}^{s}_{(i+1,k)}\big{(}H(x+n)\big{)},
H(y)+n~(i+1,k)s(H(x)+n),\displaystyle\Longleftrightarrow H(y)+n\in\tilde{\mathcal{L}}^{s}_{(i+1,k)}\big{(}H(x)+n\big{)},
H(y)~(i+1,k)s(H(x)),\displaystyle\Longleftrightarrow H(y)\in\tilde{\mathcal{L}}^{s}_{(i+1,k)}\big{(}H(x)\big{)},
y~(i+1,k)s(x).\displaystyle\Longleftrightarrow y\in\tilde{\mathcal{F}}^{s}_{(i+1,k)}(x).

It means that E~(i+1,k)s\tilde{E}^{s}_{(i+1,k)} is d\mathbb{Z}^{d}-periodic, hence it can descend to 𝕋d\mathbb{T}^{d} through (2.11). ∎

3.2 The conjugacy preserves strong stable foliations

Now we prove Proposition 3.3 that is h(is)=ish(\mathcal{F}^{s}_{i})=\mathcal{L}^{s}_{i}, for all 1ik1\leq i\leq k, where hh is the conjugacy between the special ff and its linearization AA.

Proof of Proposition 3.3.

By Proposition 2.15, we already have h((i,k)s)=(i,k)sh\left(\mathcal{F}^{s}_{(i,k)}\right)=\mathcal{L}^{s}_{(i,k)}, for every 1ik1\leq i\leq k. Since is=(1,i)s(i,k)s\mathcal{F}^{s}_{i}=\mathcal{F}^{s}_{(1,i)}\cap\mathcal{F}^{s}_{(i,k)}, it suffices to prove the following lemma.

Lemma 3.6.

Let f𝒰f\in\mathcal{U} given by Proposition 3.2 be special. Then for every 1ik1\leq i\leq k, h((1,i)s)=(1,i)sh(\mathcal{F}^{s}_{(1,i)})=\mathcal{L}^{s}_{(1,i)}.

Proof of Lemma 3.6.

Fix 1ik11\leq i\leq k-1. Firstly, we prove the joint integrability of the bundle E(1,i)sEuE^{s}_{(1,i)}\oplus E^{u}.

Claim 3.7.

The bundle E(1,i)s,u=E(1,i)sEuE^{s,u}_{(1,i)}=E^{s}_{(1,i)}\oplus E^{u} is jointly integrable.

Proof of Claim 3.7.

For any x𝕋dx\in\mathbb{T}^{d}, y(1,i)s(x,δ){x}y\in\mathcal{F}^{s}_{(1,i)}(x,\delta)\setminus\{x\} and xu(x){x}x^{\prime}\in\mathcal{F}^{u}(x)\setminus\{x\}, let y=Holx,xu(y)s(x)y^{\prime}={\rm Hol}^{u}_{x,x^{\prime}}(y)\in\mathcal{F}^{s}(x^{\prime}), where Hol:x,xus(x,δ)s(x,δ){}^{u}_{x,x^{\prime}}:\mathcal{F}^{s}(x,\delta)\to\mathcal{F}^{s}(x^{\prime},\delta) is the holonomy map along the unstable foliation u\mathcal{F}^{u}. It suffices to show that y(1,i)s(x)y^{\prime}\in\mathcal{F}^{s}_{(1,i)}(x^{\prime}).

Let II be any curve homeomorphic to [0,1][0,1] and laying on (1,i)s(x,δ)\mathcal{F}^{s}_{(1,i)}(x,\delta) with endpoints x,yx,y. Let J=Holx,xu(I)J={\rm Hol}^{u}_{x,x^{\prime}}(I). Since the conjugacy hh maps u\mathcal{F}^{u} to u\mathcal{L}^{u}, the curve h(J)h(J) is a translation of h(I)h(I). Now, by Proposition 2.10, there exists znh(x)z_{n}\to h(x^{\prime}) with Anzn=Anh(x)A^{n}z_{n}=A^{n}h(x). Moreover, we can pick curves InI_{n} with znInz_{n}\in I_{n} such that AnIn=Anh(I)A^{n}I_{n}=A^{n}h(I). Hence, the other endpoint wn(zn)w_{n}(\neq z_{n}) of InI_{n} also has Anwn=Anh(y)A^{n}w_{n}=A^{n}h(y). Since InI_{n} is also a translation of h(I)h(I), one has Inh(J)I_{n}\to h(J) and wnh(y)w_{n}\to h(y^{\prime}). See Figure 2.

Refer to caption
Figure 2: The image of holonomy map is approached by preimage sets.

Let xn=h1(zn)x_{n}=h^{-1}(z_{n}) and yn=h1(wn)y_{n}=h^{-1}(w_{n}). We have that xnxx_{n}\to x^{\prime} and ynyy_{n}\to y^{\prime}. Since AnIn=Anh(I)A^{n}I_{n}=A^{n}h(I), one has fn(h1In)=fn(I)f^{n}(h^{-1}I_{n})=f^{n}(I). It follows that h1(In)h^{-1}(I_{n}) is a curve laying on (1,i)s(xn)\mathcal{F}^{s}_{(1,i)}(x_{n}) with endpoints xnx_{n} and yny_{n}. Moreover, by the continuity of hh and h1h^{-1}, the curve h1(In)h^{-1}(I_{n}) is located in a small tubular neighborhood of JJ, when nn is big enough. Hence, we have that h1(In)Jh^{-1}(I_{n})\to J. By the continuity of the foliation (1,i)s\mathcal{F}^{s}_{(1,i)}, we get J(1,i)s(x)J\subset\mathcal{F}^{s}_{(1,i)}(x^{\prime}) and y(1,i)s(x)y^{\prime}\in\mathcal{F}^{s}_{(1,i)}(x^{\prime}). ∎

Now, by the fact that (1,i)s,u\mathcal{F}^{s,u}_{(1,i)} is subfoliated by the unstable foliation u\mathcal{F}^{u} which is minimal, we can get that h((1,i)s)h(\mathcal{F}^{s}_{(1,i)}) is a linear foliation.

Claim 3.8.

h((1,i)s)h(\mathcal{F}^{s}_{(1,i)}) is a linear foliation.

Proof of Claim 3.8.

For convenience, we prove it on the universal cover space d\mathbb{R}^{d}. Let 𝒲~=H(~(1,i)s,u)\tilde{\mathcal{W}}=H(\tilde{\mathcal{F}}^{s,u}_{(1,i)}). Note that

H(~s)=~s,H(~u)=~u,H(~(1,i)s)=𝒲~~s.H(\tilde{\mathcal{F}}^{s})=\tilde{\mathcal{L}}^{s},\quad H(\tilde{\mathcal{F}}^{u})=\tilde{\mathcal{L}}^{u},\quad H(\tilde{\mathcal{F}}^{s}_{(1,i)})=\tilde{\mathcal{W}}\cap\tilde{\mathcal{L}}^{s}.

We are going to prove that 𝒲~\tilde{\mathcal{W}} is additively closed, that is x+y𝒲~(0),x+y\in\tilde{\mathcal{W}}(0), for all x,y𝒲~(0)x,y\in\tilde{\mathcal{W}}(0). Combining with the fact that ~s\tilde{\mathcal{L}}^{s} is additively closed, we have that H(~(1,i)s)=~s𝒲~H(\tilde{\mathcal{F}}^{s}_{(1,i)})=\tilde{\mathcal{L}}^{s}\cap\tilde{\mathcal{W}} is additively closed. Hence it is a linear foliation. Note that

  1. 1.

    𝒲~(x+n)=𝒲~(x)+n\tilde{\mathcal{W}}(x+n)=\tilde{\mathcal{W}}(x)+n, for all xdx\in\mathbb{R}^{d} and ndn\in\mathbb{Z}^{d}.

  2. 2.

    𝒲~(x+v)=𝒲~(x)\tilde{\mathcal{W}}(x+v)=\tilde{\mathcal{W}}(x), for all xdx\in\mathbb{R}^{d} and v~u(0)v\in\tilde{\mathcal{L}}^{u}(0).

Indeed, we just need the fact that ~u\tilde{\mathcal{L}}^{u} is linear and ~u(x+n)=~u(x)+n\tilde{\mathcal{F}}^{u}(x+n)=\tilde{\mathcal{F}}^{u}(x)+n, for all xdx\in\mathbb{R}^{d} and ndn\in\mathbb{Z}^{d}.

Since the foliation u\mathcal{L}^{u} is minimal, there exist ndn\in\mathbb{Z}^{d} and vn~u(0)v_{n}\in\tilde{\mathcal{L}}^{u}(0) such that n+vnxn+v_{n}\to x. By y𝒲~(0)y\in\tilde{\mathcal{W}}(0), we have vn+y𝒲~(0)v_{n}+y\in\tilde{\mathcal{W}}(0). Hence, n+vn+y𝒲~(n)=𝒲~(n+vn)n+v_{n}+y\in\tilde{\mathcal{W}}(n)=\tilde{\mathcal{W}}(n+v_{n}). Let n+vnxn+v_{n}\to x, it follows that x+y𝒲~(x)=𝒲~(0)x+y\in\tilde{\mathcal{W}}(x)=\tilde{\mathcal{W}}(0). ∎

Now, we can finish our proof of Lemma 3.6. Note that the linear foliation H(~(1,i)s)~sH(\tilde{\mathcal{F}}^{s}_{(1,i)})\subset\tilde{\mathcal{L}}^{s} is AA-invariant. So, it must be a union foliation of ii subfoliations of ~s\tilde{\mathcal{L}}^{s}, that is H(~(1,i)s)=~j1s~jis.H(\tilde{\mathcal{F}}^{s}_{(1,i)})=\tilde{\mathcal{L}}^{s}_{j_{1}}\oplus...\oplus\tilde{\mathcal{L}}^{s}_{j_{i}}. On the other hand, by Proposition 2.15, we have H(~(i+1,k)s)=~(i+1,k)s.H(\tilde{\mathcal{F}}^{s}_{(i+1,k)})=\tilde{\mathcal{L}}^{s}_{(i+1,k)}. Since HH is a homeomophism, we get H(~(1,i)s)=~(1,i)sH(\tilde{\mathcal{F}}^{s}_{(1,i)})=\tilde{\mathcal{L}}^{s}_{(1,i)}. ∎

4 Affine metric

Let ff belong to 𝒰\mathcal{U} given by Proposition 2.15. Applying Livschitz Theorem for Anosov maps (Proposition 4.1), the spectral rigidity on the stable bundle implies that we can endow with a metric to each stable foliation ~is\tilde{\mathcal{F}}^{s}_{i} such that FF is affine restricted on it. Especially, for the case of dimEs=1E^{s}=1, the existence of the affine metric is just like the diffeomorphism case ([11, Lemma 3.1]) whether ff is a small perturbation or not. However, for the case of dimEs>1E^{s}>1, there are something quite different: we do not a priori have foliation is\mathcal{F}^{s}_{i} on 𝕋d\mathbb{T}^{d} for non-invertible Anosov maps. So, we will give this affine metric on the lifting ~is\tilde{\mathcal{F}}^{s}_{i}. Moreover, lack of the bundle Eis(2ik)E^{s}_{i}(2\leq i\leq k) on T𝕋dT\mathbb{T}^{d} prevents us from defining the contracting-rate function for EisE^{s}_{i}. We overcome this by using quotient dynamics.

Proposition 4.1 (Livschitz Theorem).

Let MM be a closed Riemannian manifold, f:MMf:M\to M be a C1+αC^{1+\alpha} transitive Anosov map and ϕ:M\phi:M\to\mathbb{R} be a Ho¨\ddot{o}lder continuous function. Suppose that, for every pPerp\in\rm{Per}(f)(f) with period π(p)\pi(p), i=0π(p)1ϕ(fi(p))=0\sum_{i=0}^{\pi(p)-1}\phi(f^{i}(p))=0 . Then there exists a continuous function ψ:M\psi:M\to\mathbb{R} such that ϕ=ψfψ\phi=\psi\circ f-\psi. Moreover ψ\psi is unique up to an additive constant and Ho¨\ddot{o}lder with the same exponent as ϕ\phi.

Note that the proof of this proposition is quite similar to the Livschitz Theorem for transitive Anosov diffeomorphisms whose complete proof can be found in [23, Corollary 6.4.17 and Theorem 19.2.1]. If one need, we also refer to [33, Theorem 2.1 and Proposition 2.3 ] for the (un)stable manifolds theorem and the local product structure for Anosov maps, which are useful to prove the Anosov (exponential) closing lemma (see [23, Corollary 6.4.17 ]) and the Livschitz Theorem for Anosov maps.

We also mention that every Anosov map on torus is transitive, so that the Livschitz Theorem always holds for toral Anosov maps. If one need, we refer to [1, Theorem 6.8.1] which says that σf:𝕋fd𝕋fd\sigma_{f}:\mathbb{T}^{d}_{f}\to\mathbb{T}^{d}_{f} the inverse limit system of ff is topologically conjugate to σA:𝕋Ad𝕋Ad\sigma_{A}:\mathbb{T}^{d}_{A}\to\mathbb{T}^{d}_{A} the inverse limit system of AA. Combining the fact that the toral Anosov map AA is transitive, we get the transitivity of ff. We also refer to [31, Proposition 1.2 ] to get a proof for transitivity of Anosov maps on infra-nilmanifolds.

Now, using quotient dynamics, we can apply Livschitz Theorem to F:ddF:\mathbb{R}^{d}\to\mathbb{R}^{d}.

Proposition 4.2.

Let f𝒰f\in\mathcal{U} given by Proposition 2.15. Fix 1ik1\leq i\leq k, if λis(p,f)=λis(A)\lambda^{s}_{i}(p,f)=\lambda^{s}_{i}(A), for every pPer(f)p\in{\rm Per}(f), then there exists a Ho¨\ddot{o}lder continuous function Ψ:d\Psi:\mathbb{R}^{d}\to\mathbb{R}, such that

logDF|E~is(x)=λis(A)+Ψ(x)Ψ(F(x)),{\rm log}\|DF|_{\tilde{E}^{s}_{i}(x)}\|=\lambda^{s}_{i}(A)+\Psi(x)-\Psi(F(x)),

for every xdx\in\mathbb{R}^{d}. Moreover, Ψ\Psi is bounded on d\mathbb{R}^{d}.

Proof.

For i=1i=1, since E1sE^{s}_{1} is well defined on T𝕋dT\mathbb{T}^{d} (see Proposition 2.12), we can use Livschitz Theorem (Proposition 4.1) for function Df|E1s(x)\|Df|_{E^{s}_{1}(x)}\|. By the assumption that λ1s(p,f)=λ1s(A)\lambda^{s}_{1}(p,f)=\lambda^{s}_{1}(A) for all pPer(f)p\in{\rm Per}(f), there exists a Ho¨\ddot{\rm o}lder continuous function ψ:𝕋d\psi:\mathbb{T}^{d}\to\mathbb{R} such that logDf|E1s(x)=λ1s(A)+ψ(x)ψ(f(x)){\rm log}\|Df|_{E^{s}_{1}(x)}\|=\lambda^{s}_{1}(A)+\psi(x)-\psi(f(x)) for all x𝕋dx\in\mathbb{T}^{d}. The lifting of ψ\psi, Ψ:d\Psi:\mathbb{R}^{d}\to\mathbb{R} is the function we need.

Fix 2ik2\leq i\leq k. Since the strong stable bundle E(1,j)sE^{s}_{(1,j)} (1jk)(1\leq j\leq k) is always well defined on T𝕋dT\mathbb{T}^{d}, we can define NT𝕋dN\subset T\mathbb{T}^{d} the normal bundle of E(1,i1)sE^{s}_{(1,i-1)} in E(1,i)sE^{s}_{(1,i)}. Let πN:E(1,i)sN\pi^{N}:E^{s}_{(1,i)}\to N be the natural projection and D¯f:E(1,i)sN\overline{D}f:E^{s}_{(1,i)}\to N defined as D¯f(x,v)=πNDxf(v)\overline{D}f(x,v)=\pi^{N}\circ D_{x}f(v), for all x𝕋dx\in\mathbb{T}^{d} and vE(1,i)s(x)v\in E^{s}_{(1,i)}(x).

Let μ(x):=D¯f|N(x)\mu(x):=\|\overline{D}f|_{N(x)}\| and Eis(p)E^{s}_{i}(p) be certained by the return map of the periodic point pp with period π(p)\pi(p), that is the eigenspace of the eigenvalue exp(λis(p,f))=μis(A){\rm exp}(\lambda^{s}_{i}(p,f))=\mu^{s}_{i}(A) for Dpfπ(p)D_{p}f^{\pi(p)}. For any unit vector vN(p)v\in N(p), let v=j=1ivjsv=\sum_{j=1}^{i}v^{s}_{j}, where vjsEjs(p)v^{s}_{j}\in E^{s}_{j}(p). One has

Dpfπ(p)(v)=j=1i(μjs(A))π(p)vjs.\displaystyle D_{p}f^{\pi(p)}(v)=\sum_{j=1}^{i}(\mu^{s}_{j}(A))^{\pi(p)}v^{s}_{j}. (4.1)

Note that

D¯f(x,w)=D¯fπN(x,w),wE(1,i)s(x).\displaystyle\overline{D}f(x,w)=\overline{D}f\circ\pi^{N}(x,w),\quad\forall w\in E^{s}_{(1,i)}(x). (4.2)

Indeed, let wE(1,i)s(x)w\in E^{s}_{(1,i)}(x) and w=w(1,i1)s+wNw=w^{s}_{(1,i-1)}+w^{N} be the decomposition in E(1,i1)sNE^{s}_{(1,i-1)}\oplus N, one has

πfxNDxf(w)\displaystyle\pi_{fx}^{N}\circ D_{x}f(w) =πfxNDxf(w(1,i1)s+wN)=πfxN(Dxfw(1,i1)s+DxfwN),\displaystyle=\pi_{fx}^{N}\circ D_{x}f\left(w^{s}_{(1,i-1)}+w^{N}\right)=\pi_{fx}^{N}\left(D_{x}fw^{s}_{(1,i-1)}+D_{x}fw^{N}\right),
=πfxNDxfw(1,i1)s+πfxNDxfwN=πfxNDxf(wN).\displaystyle=\pi_{fx}^{N}\circ D_{x}fw^{s}_{(1,i-1)}+\pi_{fx}^{N}\circ D_{x}fw^{N}=\pi_{fx}^{N}\circ D_{x}f(w^{N}).

Hence, by (4.1) and (4.2),

(D¯f)π(p)(x,v)=πpNDpfπ(p)(v)=(μis(A))π(p)πpN(vis)=(μis(A))π(p)v.(\overline{D}f)^{\pi(p)}(x,v)=\pi^{N}_{p}\circ D_{p}f^{\pi(p)}(v)=\left(\mu^{s}_{i}(A)\right)^{\pi(p)}\pi^{N}_{p}(v^{s}_{i})=\left(\mu^{s}_{i}(A)\right)^{\pi(p)}v.

It follows that i=0π(p)1logμ(fi(p))=π(p)μis(A)\sum_{i=0}^{\pi(p)-1}{\rm log}\mu(f^{i}(p))=\pi(p)\cdot\mu^{s}_{i}(A) for all pPer(f)p\in{\rm Per}(f). Now, using Livschitz Theorem for logμ(x)\mu(x), there exists a Ho¨\ddot{\rm o}lder continuous function ϕ:𝕋d\phi:\mathbb{T}^{d}\to\mathbb{R} such that

logμ(x)=λis(A)+ϕ(x)ϕ(f(x)).{\rm log}\mu(x)=\lambda^{s}_{i}(A)+\phi(x)-\phi(f(x)).

Let D¯F:E~(1,i)sN~\overline{D}F:\tilde{E}^{s}_{(1,i)}\to\tilde{N} be the lifting of D¯f:E(1,i)sN\overline{D}f:E^{s}_{(1,i)}\to N, where N~Td\tilde{N}\subset T\mathbb{R}^{d} is the lifting of NN. Let Φ:d\Phi:\mathbb{R}^{d}\to\mathbb{R} be the lifting of ϕ:𝕋d\phi:\mathbb{T}^{d}\to\mathbb{R}. Denote μ~(x):=D¯F|N~(x)\tilde{\mu}(x):=\|\overline{D}F|_{\tilde{N}(x)}\|. Thus, we have

logμ~(x)=λis(A)+Φ(x)Φ(F(x)).\displaystyle{\rm log}\tilde{\mu}(x)=\lambda^{s}_{i}(A)+\Phi(x)-\Phi(F(x)). (4.3)

Note that Φ\Phi is bounded and Ho¨\ddot{\rm o}lder continuous on d\mathbb{R}^{d}.

Let α(x):=logcos(N~(x),E~is(x))\alpha(x):={\rm log\;cos}\angle(\tilde{N}(x),\tilde{E}^{s}_{i}(x)), we claim that

logDF|E~is(x)=logμ~(x)+α(x)α(F(x)).\displaystyle{\rm log}\|DF|_{\tilde{E}^{s}_{i}(x)}\|={\rm log}\tilde{\mu}(x)+\alpha(x)-\alpha(F(x)). (4.4)

Indeed, it is just a linear algebraic calculation. Let vE~is(x)v\in\tilde{E}^{s}_{i}(x), we have

vN~=cos(N~(x),E~is(x))v=eα(x)v,\|v^{\tilde{N}}\|=\text{cos}\angle(\tilde{N}(x),\tilde{E}^{s}_{i}(x))\cdot\|v\|=e^{\alpha(x)}\cdot\|v\|,

and by (4.2) (this equation can also lift on TdT\mathbb{R}^{d}),

μ~(x)vN~=πFxN~DxF(v)=cos(N~(Fx),E~is(Fx))DxFv=eα(Fx)DxFv.\tilde{\mu}(x)\|v^{\tilde{N}}\|=\pi_{Fx}^{\tilde{N}}\circ D_{x}F(v)=\text{cos}\angle(\tilde{N}(Fx),\tilde{E}^{s}_{i}(Fx))\cdot\|D_{x}Fv\|=e^{\alpha(Fx)}\cdot\|D_{x}Fv\|.

Thus, we get (4.4).

Now, by (4.3) and (4.4), Ψ=Φ+α\Psi=\Phi+\alpha is a function satisfying

logDF|E~is(x)=λis(A)+Ψ(x)Ψ(F(x)).{\rm log}\|DF|_{\tilde{E}^{s}_{i}(x)}\|=\lambda^{s}_{i}(A)+\Psi(x)-\Psi(F(x)).

Moreover, since the angle (N~(x),E~is(x))\angle(\tilde{N}(x),\tilde{E}^{s}_{i}(x)) is uniformly away from π/2\pi/2 (see Proposition 2.12 and (2.12)), α(x)\alpha(x) is a bounded function defined on d\mathbb{R}^{d}. And, so is Ψ\Psi. The Ho¨\ddot{\rm o}lder continuity of both (N~(x),E~is(x))\angle(\tilde{N}(x),\tilde{E}^{s}_{i}(x)) (see Remark 2.14) and function Φ\Phi implies one of Ψ\Psi. ∎

Now, we can endow an affine metric to each leaf of ~is\tilde{\mathcal{F}}^{s}_{i} and it would be invariant under some certain holonomy maps. Let foliation ~\tilde{\mathcal{F}} be subfoliated by foliations ~1\tilde{\mathcal{F}}_{1} and ~2\tilde{\mathcal{F}}_{2} which admit the Global Product Structure on ~\tilde{\mathcal{F}}. We define the holonomy map of ~1\tilde{\mathcal{F}}_{1} along ~2\tilde{\mathcal{F}}_{2} restricted on ~\tilde{\mathcal{F}} as

Holx,x:~1(x)~1(x)withHolx,x(y)=~1(x)~2(y),\displaystyle{\rm Hol}_{x,x^{\prime}}:\tilde{\mathcal{F}}_{1}(x)\to\tilde{\mathcal{F}}_{1}(x^{\prime})\quad{\rm with}\quad{\rm Hol}_{x,x^{\prime}}(y)=\tilde{\mathcal{F}}_{1}(x^{\prime})\cap\tilde{\mathcal{F}}_{2}(y),

for every x,xdandy~1(x)x,x^{\prime}\in\mathbb{R}^{d}andy\in\tilde{\mathcal{F}}_{1}(x). Let dis(,)d^{s}_{i}(\cdot,\cdot) be a metric defined on each leaf of ~is\tilde{\mathcal{F}}^{s}_{i}, we say dis(,)d^{s}_{i}(\cdot,\cdot) is continuous, if for any ε>0{\varepsilon}>0, xdx\in\mathbb{R}^{d} and y~is(x)y\in\tilde{\mathcal{F}}^{s}_{i}(x), there exists δ>0\delta>0 such that

|dis(x,y)dis(x,y)|<ε,\left|d^{s}_{i}(x,y)-d^{s}_{i}(x^{\prime},y^{\prime})\right|<{\varepsilon},

for all xB(x,δ)x^{\prime}\in B(x,\delta) and yB(y,δ)y^{\prime}\in B(y,\delta) with y~is(x)y^{\prime}\in\tilde{\mathcal{F}}^{s}_{i}(x^{\prime}).

Proposition 4.3.

Let f𝒰f\in\mathcal{U} given by Proposition 2.15. Fix 1ik1\leq i\leq k, if λis(p,f)=λis(A)\lambda^{s}_{i}(p,f)=\lambda^{s}_{i}(A), for every pPer(f)p\in{\rm Per}(f), then there exists a continuous metric dis(,)d^{s}_{i}(\cdot,\cdot) defined on each leaf of ~is\tilde{\mathcal{F}}^{s}_{i} satisfying,

  1. 1.

    There exists a constant K>1K>1, such that 1/Kd~is(x,y)<dis(x,y)<Kd~is(x,y)1/K\cdot d_{\tilde{\mathcal{F}}^{s}_{i}}(x,y)<d^{s}_{i}(x,y)<K\cdot d_{\tilde{\mathcal{F}}^{s}_{i}}(x,y), for every xdx\in\mathbb{R}^{d} and y~is(x)y\in\tilde{\mathcal{F}}^{s}_{i}(x).

  2. 2.

    dis(Fx,Fy)=exp(λis(A))dis(x,y)d^{s}_{i}(Fx,Fy)={\rm exp}(\lambda^{s}_{i}(A))\cdot d^{s}_{i}(x,y), for every xdx\in\mathbb{R}^{d} and y~is(x)y\in\tilde{\mathcal{F}}^{s}_{i}(x).

  3. 3.

    The holonomy maps of ~is\tilde{\mathcal{F}}^{s}_{i} along ~(1,i1)s(2ik)\tilde{\mathcal{F}}^{s}_{(1,i-1)}(2\leq i\leq k) restricted on ~(1,i)s\tilde{\mathcal{F}}^{s}_{(1,i)} are isometric under the metric dis(,)d^{s}_{i}(\cdot,\cdot).

  4. 4.

    If E~isE~u\tilde{E}^{s}_{i}\oplus\tilde{E}^{u} is integrable, then the holonomy maps of ~is\tilde{\mathcal{F}}^{s}_{i} along ~u\tilde{\mathcal{F}}^{u} restricted on ~is~u\tilde{\mathcal{F}}^{s}_{i}\oplus\tilde{\mathcal{F}}^{u} are isometric under the metric dis(,)d^{s}_{i}(\cdot,\cdot).

Especially, when dimEs=1E^{s}=1, if λs(p,f)=λs(A)\lambda^{s}(p,f)=\lambda^{s}(A), for every pPer(f)p\in{\rm Per}(f), there exists a continuous metric ds(,)d^{s}(\cdot,\cdot) defined on each leaf of ~s\tilde{\mathcal{F}}^{s} satisfying the first two items. Moreover, the holonomy maps of ~s\tilde{\mathcal{F}}^{s} along ~u\tilde{\mathcal{F}}^{u} are isometric under the metric ds(,)d^{s}(\cdot,\cdot).

Proof of Proposition 4.3.

For every xdx\in\mathbb{R}^{d} and y~(x)y\in\tilde{\mathcal{F}}(x), let γ:[0,1]~is(x)\gamma:[0,1]\to\tilde{\mathcal{F}}^{s}_{i}(x) be a C1C^{1}-parametrization with γ(0)=x\gamma(0)=x and γ(1)=y\gamma(1)=y. Using the same notations in Proposition 4.2, the following formula

dis(x,y):=01eΨγ(t)|γ(t)|𝑑t,d^{s}_{i}(x,y):=\int_{0}^{1}e^{\Psi\circ\gamma(t)}\cdot|\gamma^{\prime}(t)|dt,

defines the metric we need. It is clear that the metric dis(,)d^{s}_{i}(\cdot,\cdot) is continuous.

Since Ψ\Psi is a bounded function defined on d\mathbb{R}^{d}, say ΨC0logK\|\Psi\|_{C_{0}}\leq{\rm log}K, the metric dis(,)d^{s}_{i}(\cdot,\cdot) is KK-equivalent to d~is(,)d_{\tilde{\mathcal{F}}^{s}_{i}}(\cdot,\cdot). So, we get the first item. For the second one, we calculate directly. Let y~is(x)y\in\tilde{\mathcal{F}}^{s}_{i}(x),

dis(F(x),F(y))\displaystyle d^{s}_{i}(F(x),F(y)) =01eΨFγ(t)|(Fγ)(t)|𝑑t,\displaystyle=\int_{0}^{1}e^{\Psi\circ F\circ\gamma(t)}|(F\circ\gamma)^{\prime}(t)|dt,
=01μis(A)DF|E~is(γ(t))eΨγ(t)DF|E~is(γ(t))|γ(t)|dt,\displaystyle=\int_{0}^{1}\frac{\mu^{s}_{i}(A)}{\|DF|_{\tilde{E}^{s}_{i}(\gamma(t))}\|}\cdot e^{\Psi\circ\gamma(t)}\cdot\|DF|_{\tilde{E}^{s}_{i}(\gamma(t))}\|\cdot|\gamma^{\prime}(t)|dt,
=μis(A)dis(x,y).\displaystyle=\mu^{s}_{i}(A)d^{s}_{i}(x,y).

Denote the holonomy maps of ~is\tilde{\mathcal{F}}^{s}_{i} along ~(1,i1)s\tilde{\mathcal{F}}^{s}_{(1,i-1)} restricted on ~(1,i)s\tilde{\mathcal{F}}^{s}_{(1,i)} and the holonomy maps of ~is\tilde{\mathcal{F}}^{s}_{i} along ~u\tilde{\mathcal{F}}^{u} restricted on ~is,u\tilde{\mathcal{F}}^{s,u}_{i} by Hols and Holu, respectively.

Claim 4.4.

For any ε>0\varepsilon>0, there exists δ>0\delta>0, such that, for any two curves on ~is\tilde{\mathcal{F}}^{s}_{i}, γ1:[0,1]~is(x)\gamma_{1}:[0,1]\to\tilde{\mathcal{F}}^{s}_{i}(x) and γ2:[0,1]~is(x)\gamma_{2}:[0,1]\to\tilde{\mathcal{F}}^{s}_{i}(x^{\prime}). Then, each one of the follwing conditions,

  1. 1.

    γ2(t)=Holx,xsγ1(t)\gamma_{2}(t)={\rm Hol}_{x,x^{\prime}}^{s}\circ\gamma_{1}(t), and d~(1,i1)s(γ1(t),γ2(t))<δ,t[0,1]d_{\tilde{\mathcal{F}}^{s}_{(1,i-1)}}\big{(}\gamma_{1}(t),\gamma_{2}(t)\big{)}<\delta,\forall t\in[0,1].

  2. 2.

    γ2(t)=Holx,xuγ1(t)\gamma_{2}(t)={\rm Hol}_{x,x^{\prime}}^{u}\circ\gamma_{1}(t), and d~u(γ1(t),γ2(t))<δ,t[0,1]d_{\tilde{\mathcal{F}}^{u}}\big{(}\gamma_{1}(t),\gamma_{2}(t)\big{)}<\delta,\forall t\in[0,1].

implies,

dis(γ1(0),γ1(1))dis(γ2(0),γ2(1))(1ε,1+ε).\frac{d^{s}_{i}\big{(}\gamma_{1}(0),\gamma_{1}(1)\big{)}}{d^{s}_{i}\big{(}\gamma_{2}(0),\gamma_{2}(1)\big{)}}\in(1-\varepsilon,1+\varepsilon).
Proof of Claim 4.4.

Firstly, by the uniform continuity of Ψ\Psi (see Proposition 4.2), for any ε0>0\varepsilon_{0}>0, there exists δ>0\delta>0 such that Ψγ1Ψγ2C0ε0\|\Psi\circ\gamma_{1}-\Psi\circ\gamma_{2}\|_{C_{0}}\leq\varepsilon_{0}.

Then, we need control the deviation of holonomy maps. Applying [35, Theorem B ] for dominated splitting

E~(1,i1)sE~isE~(i+1,k)s,u,\tilde{E}^{s}_{(1,i-1)}\oplus\tilde{E}^{s}_{i}\oplus\tilde{E}^{s,u}_{(i+1,k)},

we get that Holsx,x{}_{x,x^{\prime}}^{s} is C1C^{1}-smooth. Hence, we can assume there exists δ>0\delta>0 such that |DHolx,xs(γ1(t))|(1ε0,1+ε0),t[0,1]|D{\rm Hol}_{x,x^{\prime}}^{s}(\gamma_{1}(t))|\in(1-\varepsilon_{0},1+\varepsilon_{0}),\forall t\in[0,1].

Finally, we compute by definition,

dis(γ1(0),γ1(1))dis(γ2(0),γ2(1))\displaystyle\frac{d^{s}_{i}(\gamma_{1}(0),\gamma_{1}(1))}{d^{s}_{i}(\gamma_{2}(0),\gamma_{2}(1))} =01eΨγ1(t)|γ1(t)|𝑑t01eΨHolx,xsγ1(t)|DHolx,xs(γ1(t))||γ1(t)|𝑑t,\displaystyle=\frac{\int_{0}^{1}e^{\Psi\circ\gamma_{1}(t)}\cdot|\gamma_{1}^{\prime}(t)|dt}{\int_{0}^{1}e^{\Psi\circ{\rm Hol}_{x,x^{\prime}}^{s}\circ\gamma_{1}(t)}\cdot|D{\rm Hol}_{x,x^{\prime}}^{s}(\gamma_{1}(t))|\cdot|\gamma_{1}^{\prime}(t)|dt}, (4.5)
[(1+ε0)1eε0,(1ε0)1eε0].\displaystyle\in\big{[}(1+\varepsilon_{0})^{-1}e^{-\varepsilon_{0}}\;,\;(1-\varepsilon_{0})^{-1}e^{\varepsilon_{0}}\big{]}. (4.6)

For the second case, note that by Journe´\acute{\rm e} Lemma [22](see Proposition 2.19), if E~isE~u\tilde{E}^{s}_{i}\oplus\tilde{E}^{u} is integrable, then each leaf of ~is,u\tilde{\mathcal{F}}^{s,u}_{i} is C1+αC^{1+\alpha}. Therefore, we can use the same way of [32, Theorem 7.1] to get the absolute continuity of Holux,x{}_{x,x^{\prime}}^{u}. In fact, we have the Radon-Nikodym derivative of Holx,xu{\rm Hol}_{x,x^{\prime}}^{u} as follow,

Jac(Holx,xu)(x)=n=0DF|E~is(Fnz)DF|E~is(Fnx),{\rm Jac}({\rm Hol}_{x,x^{\prime}}^{u})(x)=\prod_{n=0}^{\infty}\frac{\|DF|_{\tilde{E}^{s}_{i}(F^{-n}z)}\|}{\|DF|_{\tilde{E}^{s}_{i}(F^{-n}x)}\|},

where z=Holx,xu(x)z={\rm Hol}_{x,x^{\prime}}^{u}(x). Since the distribution E~is\tilde{E}^{s}_{i} is Ho¨\ddot{\rm o}lder continuous (Remark 2.14), by the standard distortion control techniques, there exists δ>0\delta>0 such that |Jac(Holx,xu)(γ1(t))|(1ε0,1+ε0),t[0,1]|{\rm Jac}({\rm Hol}_{x,x^{\prime}}^{u})(\gamma_{1}(t))|\in(1-\varepsilon_{0},1+\varepsilon_{0}),\forall t\in[0,1]. The rest of proof is similar to (4.6), since ~is\tilde{\mathcal{F}}^{s}_{i} is one-dimensional. Note that the case for Hols can also be proved by using only absolute continuity. ∎

Now, we prove that Holux,x{}_{x,x^{\prime}}^{u} is isometric under the metric dis(,)d^{s}_{i}(\cdot,\cdot) by iterating backward. An analogical way can prove one for Holsx,x{}_{x,x^{\prime}}^{s} by iterating forward.

If there exist y~is(x)y\in\tilde{\mathcal{F}}^{s}_{i}(x) and y~is(x)y^{\prime}\in\tilde{\mathcal{F}}^{s}_{i}(x^{\prime}) with Hol(x)x,xu=x{}_{x,x^{\prime}}^{u}(x)=x^{\prime} and Hol(y)x,xu=y{}_{x,x^{\prime}}^{u}(y)=y^{\prime} such that dis(x,y)dis(x,y)d^{s}_{i}(x,y)\neq d^{s}_{i}(x^{\prime},y^{\prime}). Iterating these points backward, we can assume that γ1(0)=Fn(x),γ1(1)=Fn(y)\gamma_{1}(0)=F^{-n}(x),\gamma_{1}(1)=F^{-n}(y) and γ2(0)=Fn(x),γ2(1)=Fn(y)\gamma_{2}(0)=F^{-n}(x^{\prime}),\gamma_{2}(1)=F^{-n}(y^{\prime}) satisfy the conditions of Claim 4.4, for a large nn\in\mathbb{N}. Since FF is affine along ~is\tilde{\mathcal{F}}^{s}_{i} under the metric dis(,)d^{s}_{i}(\cdot,\cdot), one has

dis(Fnx,Fny)dis(Fnx,Fny)=dis(x,y)dis(x,y).\displaystyle\frac{d^{s}_{i}(F^{-n}x,F^{-n}y)}{d^{s}_{i}(F^{-n}x^{\prime},F^{-n}y^{\prime})}=\frac{d^{s}_{i}(x,y)}{d^{s}_{i}(x^{\prime},y^{\prime})}. (4.7)

If we pick ε{\varepsilon} small enough in Claim 4.4, then it contradicts with (4.7). This complete the proof. ∎

5 Existence of integrable subbundles

In this section, we prove the sufficient parts of Theorem 1.1 and Theorem 1.2 under the assumption that every periodic point of ff has the same Lyapunov spectrum on the stable bundle ( a priori, need not equal to one of the linearization) and we restate as follow.

Theorem 5.1.

Let A:𝕋d𝕋dA:\mathbb{T}^{d}\to\mathbb{T}^{d} be an irreducible linear Anosov map. Assume that AA admits the finest (on stable bundle) dominated splitting,

T𝕋d=L1sL2sLksLu,T\mathbb{T}^{d}=L^{s}_{1}\oplus L^{s}_{2}\oplus...\oplus L^{s}_{k}\oplus L^{u},

where dimLis=1{\rm dim}L^{s}_{i}=1, 1ik1\leq i\leq k.

There exists a C1C^{1} neighborhood 𝒰C1(𝕋d)\mathcal{U}\subset C^{1}(\mathbb{T}^{d}) of AA such that for every C1+αC^{1+\alpha}-smooth f𝒰f\in\mathcal{U}, if λis(p,f)=λis(q,f)\lambda^{s}_{i}(p,f)=\lambda^{s}_{i}(q,f), for all p,qPer(f)p,q\in{\rm Per}(f) and all 1ik1\leq i\leq k, then ff admits the finest (on stable bundle) dominated splitting,

T𝕋d=E1sE2sEksEu,T\mathbb{T}^{d}=E^{s}_{1}\oplus E^{s}_{2}\oplus...\oplus E^{s}_{k}\oplus E^{u},

where dimEis=1{\rm dim}E^{s}_{i}=1, 1ik1\leq i\leq k. Especially, ff is special.

Moreover, when k=1k=1, for every C1+αC^{1+\alpha} Anosov map ff with linearization AA, if λs(p,f)=λs(q,f)\lambda^{s}(p,f)=\lambda^{s}(q,f), for every p,qPer(f)p,q\in{\rm Per}(f), then ff is special.

We mention that in Theorem 5.1, ff can be inverse since an Anosov diffeomorphism is always a special Anosov map.

It is convenient to give the scheme of our proof. In this section we always assume that AA is irreducible and has the finest (on stable bundle) dominated splitting. Let f𝒰f\in\mathcal{U} given by Proposition 2.15 and F:ddF:\mathbb{R}^{d}\to\mathbb{R}^{d} be a lifting of ff and H:ddH:\mathbb{R}^{d}\to\mathbb{R}^{d} be the conjugacy between FF and AA.

Firstly, we show that HH maps every one-dimensional stable foliation ~is(1ik)\tilde{\mathcal{F}}^{s}_{i}(1\leq i\leq k) to one of the linearization ~is\tilde{\mathcal{L}}^{s}_{i}. Moreover, it is an isometry along each leaf of ~is\tilde{\mathcal{F}}^{s}_{i}. As proved in Proposition 2.15, we already have that HH preserves weak stable foliations. For reducing to each single leaf, we need the following two propositions. The idea to reduce the leaf conjugacy originated from [12] (also see [15]). A main tool in [12] is the minimal property of the foliation is\mathcal{F}^{s}_{i}. However, in our case, there is a priori no is(i2)\mathcal{F}^{s}_{i}(i\geq 2) on 𝕋d\mathbb{T}^{d}. So, we cannot use the minimal property, directly. This obstruction can be overcomed by using a special d\mathbb{Z}^{d}-sequence described in Proposition 2.6, Proposition 2.7 and Proposition 2.9.

Again, by Proposition 2.15, we already have H(~ks)=~ksH(\tilde{\mathcal{F}}^{s}_{k})=\tilde{\mathcal{L}}^{s}_{k}. We show that it is an isometry restricted on each leaf of ~ks\tilde{\mathcal{F}}^{s}_{k} under the metric dks(,)d^{s}_{k}(\cdot,\cdot) given by Proposition 4.3. Generally, we have the following proposition.

Proposition 5.1.

Let f𝒰f\in\mathcal{U} given by Proposition 2.15. Fix 1ik1\leq i\leq k, assume that H(~is)=~isH(\tilde{\mathcal{F}}^{s}_{i})=\tilde{\mathcal{L}}^{s}_{i} and λis(p,f)=λis(A)\lambda^{s}_{i}(p,f)=\lambda^{s}_{i}(A) for all pPer(f)p\in{\rm Per}(f). Then HH is isometric restricted on each leaf of ~is\tilde{\mathcal{F}}^{s}_{i} under the metric dis(,)d^{s}_{i}(\cdot,\cdot) given by Proposition 4.3. Especially, for an irreducible Anosov map ff with dimEs=1E^{s}=1, if λs(p,f)=λis(A)\lambda^{s}(p,f)=\lambda^{s}_{i}(A) for all pPer(f)p\in{\rm Per}(f), then HH is isometric restricted on each leaf of ~s\tilde{\mathcal{F}}^{s} under the metric ds(,)d^{s}(\cdot,\cdot).

The following proposition allow us to reduce the leaf conjugacy by induction.

Proposition 5.2.

Let f𝒰f\in\mathcal{U} given by Proposition 2.15. Fix 1<ik1<i\leq k, assume that H(~(1,i)s)=~(1,i)sH(\tilde{\mathcal{F}}^{s}_{(1,i)})=\tilde{\mathcal{L}}^{s}_{(1,i)} and HH is isometric restricted on each leaf of ~is\tilde{\mathcal{F}}^{s}_{i} under the metric dis(,)d^{s}_{i}(\cdot,\cdot), then H(~(1,i1)s)=~(1,i1)sH(\tilde{\mathcal{F}}^{s}_{(1,i-1)})=\tilde{\mathcal{L}}^{s}_{(1,i-1)}.

We leave the proofs for Proposition 5.1 and Proposition 5.2 in subsection 5.1. Combining these two propositions, we can prove that HH preserves every one-dimensional stable foliation ~is\tilde{\mathcal{F}}^{s}_{i} and in fact is an isometry restricted on each leaf of ~is\tilde{\mathcal{F}}^{s}_{i}.

Corollary 5.3.

Let f𝒰f\in\mathcal{U} given by Proposition 2.15 with λis(p,f)=λis(q,f)\lambda^{s}_{i}(p,f)=\lambda^{s}_{i}(q,f), for all p,qPer(f)p,q\in{\rm Per}(f) and all 1ik1\leq i\leq k. Then, for every 1ik1\leq i\leq k, H(~is)=~isH(\tilde{\mathcal{F}}^{s}_{i})=\tilde{\mathcal{L}}^{s}_{i} and HH is isometric along each leaf of ~is\tilde{\mathcal{F}}^{s}_{i} under the metric dis(,)d^{s}_{i}(\cdot,\cdot).

Proof of Corollary 5.3.

We get the proof by induction.The beginning of the induction is H(~(i,k)s)=~(i,k)sH(\tilde{\mathcal{F}}^{s}_{(i,k)})=\tilde{\mathcal{L}}^{s}_{(i,k)} (see Proposition 2.15), especially, H(~ks)=~ksH(\tilde{\mathcal{F}}^{s}_{k})=\tilde{\mathcal{L}}^{s}_{k}. Then, by Proposition 2.17, the assumption that λks(p,f)=λks(q,f)\lambda^{s}_{k}(p,f)=\lambda^{s}_{k}(q,f), for every p,qPer(f)p,q\in{\rm Per}(f) implies λks(p,f)=λks(A)\lambda^{s}_{k}(p,f)=\lambda^{s}_{k}(A). It allows us to define an affine metric dks(,)d^{s}_{k}(\cdot,\cdot) by Proposition 4.3.

Now, by Proposition 5.1, H:~ks~ksH:\tilde{\mathcal{F}}^{s}_{k}\to\tilde{\mathcal{L}}^{s}_{k} is isometric. Thus, by Proposition 5.2, H(~(1,k)s)=~(1,k)sH(\tilde{\mathcal{F}}^{s}_{(1,k)})=\tilde{\mathcal{L}}^{s}_{(1,k)} implies H(~(1,k1)s)=~(1,k1)sH(\tilde{\mathcal{F}}^{s}_{(1,k-1)})=\tilde{\mathcal{L}}^{s}_{(1,k-1)}. Moreover, since HH preserves the weak stable foliation H(~(k1,k)s)=~(k1,k)sH(\tilde{\mathcal{F}}^{s}_{(k-1,k)})=\tilde{\mathcal{L}}^{s}_{(k-1,k)}, we have

H(~k1s)=H(~(1,k1)s~(k1,k)s)=~(1,k1)s~(k1,k)s=~k1s.H(\tilde{\mathcal{F}}^{s}_{k-1})=H\left(\tilde{\mathcal{F}}^{s}_{(1,k-1)}\cap\tilde{\mathcal{F}}^{s}_{(k-1,k)}\right)\ =\tilde{\mathcal{L}}^{s}_{(1,k-1)}\cap\tilde{\mathcal{L}}^{s}_{(k-1,k)}=\tilde{\mathcal{L}}^{s}_{k-1}.

Applying the preceding methods to ~k1s\tilde{\mathcal{F}}^{s}_{k-1} and ~(1,k1)s\tilde{\mathcal{F}}^{s}_{(1,k-1)} , we have H(~(1,k2)s)=~(1,k2)sH(\tilde{\mathcal{F}}^{s}_{(1,k-2)})=\tilde{\mathcal{L}}^{s}_{(1,k-2)}. Moreover, by intersecting with ~(k2,k)s\tilde{\mathcal{F}}^{s}_{(k-2,k)}, we have H(~k2s)=~k2sH(\tilde{\mathcal{F}}^{s}_{k-2})=\tilde{\mathcal{L}}^{s}_{k-2}.

Consequently, we can finish our proof by induction. ∎

Using the isometry HH along each leaf of single stable foliations, we also can show that all stable foliations ~is(1ik)\tilde{\mathcal{F}}^{s}_{i}\;(1\leq i\leq k) are d\mathbb{Z}^{d}-periodic. Moreover, there is no deviation between H1(x+n)H^{-1}(x+n) and H1(x)+nH^{-1}(x)+n along ~is\tilde{\mathcal{F}}^{s}_{i}, for all xdx\in\mathbb{R}^{d} and ndn\in\mathbb{Z}^{d}. More precisely, we have the following two propositions.

Proposition 5.4.

Let f𝒰f\in\mathcal{U} given by Proposition 2.15. Assume that for every 1jk1\leq j\leq k, H(~js)=~jsH(\tilde{\mathcal{F}}^{s}_{j})=\tilde{\mathcal{L}}^{s}_{j} and HH is isometric along each leaf of ~js\tilde{\mathcal{F}}^{s}_{j} under the metric djs(,)d^{s}_{j}(\cdot,\cdot). Fix 1i<k1\leq i<k, if ~(i,k)s\tilde{\mathcal{F}}^{s}_{(i,k)} is d\mathbb{Z}^{d}-periodic, then so is ~(i+1,k)s\tilde{\mathcal{F}}^{s}_{(i+1,k)}.

Remark 5.5.

Using Proposition 5.4 and by induction beginning with the fact that ~(1,k)s=~s\tilde{\mathcal{F}}^{s}_{(1,k)}=\tilde{\mathcal{F}}^{s} is d\mathbb{Z}^{d}-periodic, we have that ~(i,k)s(1ik)\tilde{\mathcal{F}}^{s}_{(i,k)}(1\leq i\leq k) is d\mathbb{Z}^{d}-periodic. Note that ~(1,i)s(1ik)\tilde{\mathcal{F}}^{s}_{(1,i)}(1\leq i\leq k) is always d\mathbb{Z}^{d}-periodic (see Proposition 2.15). Thus, ~is=~(1,i)s~(i,k)s\tilde{\mathcal{F}}^{s}_{i}=\tilde{\mathcal{F}}^{s}_{(1,i)}\cap\tilde{\mathcal{F}}^{s}_{(i,k)} is also d\mathbb{Z}^{d}-periodic, for all 1ik1\leq i\leq k. It follows that for every 1ik1\leq i\leq k, E~is\tilde{E}^{s}_{i} is d\mathbb{Z}^{d}-periodic. Hence, by (2.11), we get the bundle EisE^{s}_{i} defined well on 𝕋d\mathbb{T}^{d}.

By Corollary 5.3, E~(1,i1)sE~(i+1,k)s\tilde{E}^{s}_{(1,i-1)}\oplus\tilde{E}^{s}_{(i+1,k)} is interagble and denote the FF-invariant integral foliations by ~is,\tilde{\mathcal{F}}^{s,\perp}_{i}. We also denote ~(1,i1)s~(i+1,k)s\tilde{\mathcal{L}}^{s}_{(1,i-1)}\oplus\tilde{\mathcal{L}}^{s}_{(i+1,k)} by ~is,\tilde{\mathcal{L}}^{s,\perp}_{i}.

Proposition 5.6.

Let f𝒰f\in\mathcal{U} given by Proposition 2.15. Assume that for every 1ik1\leq i\leq k, H(~is)=~isH(\tilde{\mathcal{F}}^{s}_{i})=\tilde{\mathcal{L}}^{s}_{i} and HH is isometric along each leaf of ~is\tilde{\mathcal{F}}^{s}_{i} under the metric dis(,)d^{s}_{i}(\cdot,\cdot). Then

H1(x+n)n~is,(H1(x)),H^{-1}(x+n)-n\in\tilde{\mathcal{F}}^{s,\perp}_{i}\big{(}H^{-1}(x)\big{)},

for all 1ik1\leq i\leq k, xdx\in\mathbb{R}^{d} and ndn\in\mathbb{Z}^{d}. Especially, for every irreducible Anosov map ff on torus with one dimensional stable bundle, if HH is isometric along each leaf of ~s\tilde{\mathcal{F}}^{s} under the metric ds(,)d^{s}(\cdot,\cdot) then

H1(x+n)n=H1(x),H^{-1}(x+n)-n=H^{-1}(x),

for all xdx\in\mathbb{R}^{d} and ndn\in\mathbb{Z}^{d}.

We leave the proofs for Proposition 5.4 and Proposition 5.6 in subsection 5.2. Now, by the previous propositions, we can prove Theorem 5.1.

Proof of Theorem 5.1.

In the case of dimEs=1E^{s}=1, since λs(p,f)=λs(q,f)\lambda^{s}(p,f)=\lambda^{s}(q,f) for all p,qPer(f)p,q\in{\rm Per}(f), combining Proposition 5.1 and Proposition 5.6, we have H1(x+n)=H1(x)+nH^{-1}(x+n)=H^{-1}(x)+n, for all xdx\in\mathbb{R}^{d} and ndn\in\mathbb{Z}^{d}. It means that HH can descend to 𝕋d\mathbb{T}^{d}. By Proposition 2.2, ff is special.

In the case of higher-dimensional stable bundle, by Corollary 5.3 and Proposition 5.6, we have

H1(x+n)ni=1k(~is,(H1x))={H1(x)},H^{-1}(x+n)-n\in\bigcap_{i=1}^{k}\left(\tilde{\mathcal{F}}^{s,\perp}_{i}(H^{-1}x)\right)=\big{\{}H^{-1}(x)\big{\}},

for all xdx\in\mathbb{R}^{d} and ndn\in\mathbb{Z}^{d}. Hence, ff is special. And by Remark 5.5, ff admits T𝕋d=E1s..EksEu.T\mathbb{T}^{d}=E^{s}_{1}\oplus..\oplus E^{s}_{k}\oplus E^{u}.

Combining Theorem 3.1 and Corollary 5.3, we can show that, if the conjugacy between the non-invertible Anosov maps ff and AA exists, then it must be smooth along the stable foliation (see Corollary 1.3 and Corollary 1.6). We mention that it can be proved without Proposition 5.4 and Proposition 5.6.

Proof of Corollary 1.3 and Corollary 1.6.

Assume that ff is C1+αC^{1+\alpha}-smooth. Let hh be a conjugacy between ff and AA. By Proposition 2.2, ff is special. Then, by Theorem 3.1, we have the spectral rigidity on stable bundle for ff which is exactly the condition stated in Theorem 5.1. Moreover, ff admits the finest (on stable bundle) dominated splitting and the conjugacy hh maps each stable foliation is\mathcal{F}^{s}_{i} to is\mathcal{L}^{s}_{i} (see Proposition 3.3). Since the bundle Eis(1ik)E^{s}_{i}(1\leq i\leq k) is Ho¨\ddot{\rm o}lder continuous (see Remark 2.14) with exponent β\beta, for some 0<βα0<\beta\leq\alpha, by Proposition 4.1 and the construction of dis(,)d^{s}_{i}(\cdot,\cdot) (see Proposition 4.3), the metric dis(,)d^{s}_{i}(\cdot,\cdot) is C1+βC^{1+\beta}-smooth along each leaf of ~is\tilde{\mathcal{F}}^{s}_{i}. So, Corollary 5.3 actually shows that hh is C1+βC^{1+\beta}-smooth along is\mathcal{F}^{s}_{i}, 1ik1\leq i\leq k. It follows that hh is C1+βC^{1+\beta}-smooth along the stable foliation s\mathcal{F}^{s}, by Journe´\acute{\rm e} Lemma [22]. In the case of dimEs=1E^{s}=1, since each leaf of the stable foliation s\mathcal{F}^{s} is C1+αC^{1+\alpha}-smooth, the metric ds(,)d^{s}(\cdot,\cdot) is C1+αC^{1+\alpha}-smooth along each leaf of the stable foliation s\mathcal{F}^{s} and so is hh. ∎

5.1 Induction of the leaf conjugacy

In this subsection, we prove Proposition 5.1 and Proposition 5.2. Let f𝒰f\in\mathcal{U} given by Proposition 2.15. Fix 1ik1\leq i\leq k, assume that H(~is)=~isH(\tilde{\mathcal{F}}^{s}_{i})=\tilde{\mathcal{L}}^{s}_{i} and λis(p,f)=λis(A)\lambda^{s}_{i}(p,f)=\lambda^{s}_{i}(A) for all pPer(f)p\in{\rm Per}(f). We show that HH is isometric restricted on ~is\tilde{\mathcal{F}}^{s}_{i}.

Proof of Proposition 5.1.

By the existence of affine metric dis(,)d^{s}_{i}(\cdot,\cdot) given by Proposition 4.3, we have that HH is bi-Lipschitz along ~is\tilde{\mathcal{F}}^{s}_{i}. Indeed, one can just iterate any two points x0,y0dx_{0},y_{0}\in\mathbb{R}^{d} such that they are away an almost fixed distance, say dis(Fn(x0),Fn(y0))[C1,C2]d^{s}_{i}\big{(}F^{n}(x_{0}),F^{n}(y_{0})\big{)}\in[C_{1},C_{2}]. By the uniform continuity of HH and ~is\tilde{\mathcal{F}}^{s}_{i} (see Remark 2.16), there exist C1,C2>0C_{1}^{\prime},C_{2}^{\prime}>0 such that for every x,ydx,y\in\mathbb{R}^{d}, if dis(x,y)[C1,C2]d^{s}_{i}(x,y)\in[C_{1},C_{2}] then d(H(x),H(y))[C1,C2]d\big{(}H(x),H(y)\big{)}\in[C_{1}^{\prime},C_{2}^{\prime}]. Hence, by Proposition 4.3 we have

d(H(x0),H(y0))dis(x0,y0)=d(HFn(x0),HFn(y0))dis(Fn(x0),Fn(y0))[C1C2,C2C1].\frac{d\big{(}H(x_{0}),H(y_{0})\big{)}}{d_{i}^{s}(x_{0},y_{0})}=\frac{d\big{(}H\circ F^{n}(x_{0}),H\circ F^{n}(y_{0})\big{)}}{d_{i}^{s}\big{(}F^{n}(x_{0}),F^{n}(y_{0})\big{)}}\in\left[\frac{C_{1}^{\prime}}{C_{2}},\frac{C_{2}^{\prime}}{C_{1}}\right].

Similarly, we have H1H^{-1} is Lipschitz along ~is\tilde{\mathcal{L}}^{s}_{i}.

It is convenient to prove that H1H^{-1} is isometric along ~is\tilde{\mathcal{L}}^{s}_{i}. Thus, so is HH along ~is\tilde{\mathcal{F}}^{s}_{i}.

By Lipschitz continuity, there exists a point x~is(0)x\in\tilde{\mathcal{L}}^{s}_{i}(0) differentiable and we assume (H1|~is)(x)=C0\big{(}H^{-1}|_{\tilde{\mathcal{L}}^{s}_{i}}\big{)}^{\prime}(x)=C_{0}. It means that, for any ε>0\varepsilon>0 small enough, there exists δ>0\delta>0 such that

|dis(H1(x),H1(w))d(x,w)C0|<ε2,\displaystyle\left|\frac{d^{s}_{i}\big{(}H^{-1}(x),H^{-1}(w)\big{)}}{d(x,w)}-C_{0}\right|<\frac{{\varepsilon}}{2}, (5.1)

for every w~is(x,δ)w\in\tilde{\mathcal{L}}^{s}_{i}(x,\delta). Fix ydy\in\mathbb{R}^{d}, for every z~is(y)z\in\tilde{\mathcal{L}}^{s}_{i}(y) and nmdn_{m}\in\mathbb{Z}^{d}, we denote

ym=~s(y+nm)~u(x),zm=~is(ym)~is,(z+nm)andxm=~is(x)~u(zm).y_{m}=\tilde{\mathcal{L}}^{s}(y+n_{m})\cap\tilde{\mathcal{L}}^{u}(x),\;z_{m}=\tilde{\mathcal{L}}^{s}_{i}(y_{m})\cap\tilde{\mathcal{L}}^{s,\perp}_{i}(z+n_{m})\;\;{\rm and}\;\;x_{m}=\tilde{\mathcal{L}}^{s}_{i}(x)\cap\tilde{\mathcal{L}}^{u}(z_{m}).
Claim 5.7.

There exists a sequence {nm}d\{n_{m}\}\subset\mathbb{Z}^{d} with nmAmdn_{m}\in A^{m}\mathbb{Z}^{d}, such that when m+m\to+\infty,

d(x,xm)d(y,z),\displaystyle d(x,x_{m})\to d(y,z), (5.2)

and

dis(H1(x),H1(xm))dis(H1(z),H1(y)).\displaystyle d^{s}_{i}\big{(}H^{-1}(x),H^{-1}(x_{m})\big{)}\to d^{s}_{i}\big{(}H^{-1}(z),H^{-1}(y)\big{)}. (5.3)
Proof of Claim 5.7.

Fix mm\in\mathbb{N}, we consider the point AmyA^{-m}y. Since AA is irreducible, the unstable foliation u\mathcal{L}^{u} is minimal on 𝕋d\mathbb{T}^{d}. Thus we can choose n~md\tilde{n}_{m}\in\mathbb{Z}^{d} such that d(Amy+n~m,~u(Amx))1d\big{(}A^{-m}y+\tilde{n}_{m},\tilde{\mathcal{L}}^{u}(A^{-m}x)\big{)}\leq 1. Let y(m)=~s(Amy+n~m)~u(Amx)y(m)=\tilde{\mathcal{L}}^{s}(A^{-m}y+\tilde{n}_{m})\cap\tilde{\mathcal{L}}^{u}(A^{-m}x) and nm=Amn~mn_{m}=A^{m}\tilde{n}_{m}. Note that ym=Amy(m)y_{m}=A^{m}y(m) and

d(ym,y+nm)A|Lsmd(y(m),Amy+n~m)0,d(y_{m},y+n_{m})\leq\|A|_{L^{s}}\|^{m}d(y(m),A^{-m}y+\tilde{n}_{m})\to 0,

as m+m\to+\infty. The sequence {nm}\{n_{m}\} is what we need. Let m+m\to+\infty, we have

d(ym,y+nm)0andd(z+nm,zm)0.\displaystyle d(y_{m},y+n_{m})\to 0\quad{\rm and}\quad d(z+n_{m},z_{m})\to 0. (5.4)

Moreover,

|d(y+nm,z+nm)d(ym,zm)|0.\left|d(y+n_{m},z+n_{m})-d(y_{m},z_{m})\right|\to 0.

Note that d(x,xm)=d(ym,zm)d(x,x_{m})=d(y_{m},z_{m}), then we get (5.2). See Figure 3.

Refer to caption
Figure 3: Approaching by nmn_{m} sequence.

Now, by the uniform continuity of H1H^{-1} and (5.4), one has

d(H1(zm),H1(z+nm))0andd(H1(ym),H1(y+nm))0,\displaystyle d\big{(}H^{-1}(z_{m}),H^{-1}(z+n_{m})\big{)}\to 0\quad{\rm and}\quad d\big{(}H^{-1}(y_{m}),H^{-1}(y+n_{m})\big{)}\to 0, (5.5)

as m+m\to+\infty. By H1(~is)=~isH^{-1}(\tilde{\mathcal{L}}^{s}_{i})=\tilde{\mathcal{F}}^{s}_{i} and H1(~u)=~uH^{-1}(\tilde{\mathcal{L}}^{u})=\tilde{\mathcal{F}}^{u}, we have E~isE~u\tilde{E}^{s}_{i}\oplus\tilde{E}^{u} is jointly integrable. Thus, by the forth item of Proposition 4.3, one has

dis(H1(x),H1(xm))=dis(H1(zm),H1(ym)).\displaystyle d^{s}_{i}\big{(}H^{-1}(x),H^{-1}(x_{m})\big{)}=d^{s}_{i}\big{(}H^{-1}(z_{m}),H^{-1}(y_{m})\big{)}. (5.6)

Combining (5.5) and (5.6), we get

|dis(H1(x),H1(xm))dis(H1(z+nm),H1(y+nm))|0.\displaystyle\Big{|}d^{s}_{i}\big{(}H^{-1}(x),H^{-1}(x_{m})\big{)}-d^{s}_{i}\big{(}H^{-1}(z+n_{m}),H^{-1}(y+n_{m})\big{)}\Big{|}\to 0. (5.7)

On the other hand, by Proposition 2.6, we have

(H1(z+nm)nm)H1(z)and(H1(y+nm)nm)H1(y),\displaystyle\big{(}H^{-1}(z+n_{m})-n_{m}\big{)}\to H^{-1}(z)\quad{\rm and}\quad\big{(}H^{-1}(y+n_{m})-n_{m}\big{)}\to H^{-1}(y), (5.8)

as m+m\to+\infty. Note that H1(z)+nmH^{-1}(z)+n_{m} may not belong to ~is(H1(y)+nm)\tilde{\mathcal{F}}^{s}_{i}(H^{-1}(y)+n_{m}). But, by Proposition 2.7,

d(H1(z)+nm,~is(H1(y)+nm))0.\displaystyle d\left(H^{-1}(z)+n_{m}\;,\;\tilde{\mathcal{F}}^{s}_{i}(H^{-1}(y)+n_{m})\right)\to 0. (5.9)

Hence, by (5.8) and (5.9), one has

dis(H1(z+nm),H1(y+nm))dis(H1(z),H1(y)).\displaystyle d^{s}_{i}\big{(}H^{-1}(z+n_{m}),H^{-1}(y+n_{m})\big{)}\to d^{s}_{i}\big{(}H^{-1}(z),H^{-1}(y)\big{)}. (5.10)

Consequently, according to (5.7) and (5.10), when m+m\to+\infty, we get (5.3).

Now, by Claim 5.7, for any ε>0{\varepsilon}>0 and z~is(y)z\in\tilde{\mathcal{L}}^{s}_{i}(y), there exists N0N_{0}\in\mathbb{N} such that when m>N0m>N_{0}, one has

|dis(H1(y),H1(z))d(y,z)dis(H1(x),H1(xm))d(x,xm)|<ε2.\displaystyle\left|\frac{d^{s}_{i}(H^{-1}(y),H^{-1}(z))}{d(y,z)}-\frac{d^{s}_{i}(H^{-1}(x),H^{-1}(x_{m}))}{d(x,x_{m})}\right|<\frac{{\varepsilon}}{2}. (5.11)

Moreover, let z~is(y,δ2)z\in\tilde{\mathcal{L}}^{s}_{i}(y,\frac{\delta}{2}), we have d(x,xm)<δd(x,x_{m})<\delta, for all m>N0m>N_{0}. Combining (5.1) and (5.11), for every ydy\in\mathbb{R}^{d} and every ε>0\varepsilon>0, there exists δ>0\delta>0 such that

|dis(H1(y),H1(z))d(y,z)C0|<ε,\left|\frac{d^{s}_{i}\big{(}H^{-1}(y),H^{-1}(z)\big{)}}{d(y,z)}-C_{0}\right|<\varepsilon,

for every z~is(y,δ2)z\in\tilde{\mathcal{L}}^{s}_{i}(y,\frac{\delta}{2}). It follows that HH is differentiable along ~is\tilde{\mathcal{L}}^{s}_{i} and the derivative is the constant C0C_{0}. Note that we can change the metric dis(,)d^{s}_{i}(\cdot,\cdot) by scaling such that C0=1C_{0}=1.

Now, fix 1<ik1<i\leq k, suppose that H(~(1,i)s)=~(1,i)sH(\tilde{\mathcal{F}}^{s}_{(1,i)})=\tilde{\mathcal{L}}^{s}_{(1,i)}. Note that since HH always preserves weak stable foliations, H(~(1,i)s)=~(1,i)sH(\tilde{\mathcal{F}}^{s}_{(1,i)})=\tilde{\mathcal{L}}^{s}_{(1,i)} implies H(~is)=H(~(1,i)s~(i,k)s)=~isH(\tilde{\mathcal{F}}^{s}_{i})=H(\tilde{\mathcal{F}}^{s}_{(1,i)}\cap\tilde{\mathcal{F}}^{s}_{(i,k)})=\tilde{\mathcal{L}}^{s}_{i}. Assume that HH is isometric restricted on each leaf of ~is\tilde{\mathcal{F}}^{s}_{i} under the metric dis(,)d^{s}_{i}(\cdot,\cdot). We show H(~(1,i1)s)=~(1,i1)sH(\tilde{\mathcal{F}}^{s}_{(1,i-1)})=\tilde{\mathcal{L}}^{s}_{(1,i-1)}.

Proof of Proposition 5.2.

By the assumption H(~(1,i)s)=~(1,i)sH(\tilde{\mathcal{F}}^{s}_{(1,i)})=\tilde{\mathcal{L}}^{s}_{(1,i)}, the foliation ~=H(~(1,i1)s)\tilde{\mathcal{F}}=H(\tilde{\mathcal{F}}^{s}_{(1,i-1)}) is a sub-foliation of ~(1,i)s\tilde{\mathcal{L}}^{s}_{(1,i)}. It is clear that ~\tilde{\mathcal{F}} and ~is\tilde{\mathcal{L}}^{s}_{i} give the Global Product Structure on ~(1,i)s\tilde{\mathcal{L}}^{s}_{(1,i)}. By Proposition 4.3, the holonomy maps of ~is\tilde{\mathcal{F}}^{s}_{i} along ~(1,i1)s\tilde{\mathcal{F}}^{s}_{(1,i-1)} restricted on ~(1,i)s\tilde{\mathcal{F}}^{s}_{(1,i)} are isometric under the metric dis(,)d^{s}_{i}(\cdot,\cdot). Combining with the assumption that H1:~is~isH^{-1}:\tilde{\mathcal{L}}^{s}_{i}\to\tilde{\mathcal{F}}^{s}_{i} is isometric, we have that the holonomy maps of ~is\tilde{\mathcal{L}}^{s}_{i} along ~\tilde{\mathcal{F}} restricted on ~(1,i)s\tilde{\mathcal{L}}^{s}_{(1,i)} are isometric.

Assume that H1(~(1,i1)s)~(1,i1)sH^{-1}(\tilde{\mathcal{L}}^{s}_{(1,i-1)})\neq\tilde{\mathcal{F}}^{s}_{(1,i-1)}. It follows that there exist points x0x_{0} and x1x_{1} such that x1~(x0)~(1,i1)s(x0)x_{1}\in\tilde{\mathcal{F}}(x_{0})\setminus\tilde{\mathcal{L}}^{s}_{(1,i-1)}(x_{0}).

Claim 5.8.

There exist bm~is(x0)b_{m}\in\tilde{\mathcal{L}}^{s}_{i}(x_{0}), kmAmdk_{m}\in A^{m}\mathbb{Z}^{d} and x2~(x0)x_{2}\in\tilde{\mathcal{F}}(x_{0}) such that when m+m\to+\infty,

  1. 1.

    (bm+km)x1(b_{m}+k_{m})\to x_{1}.

  2. 2.

    d(ym+km,~(bm+km))0d\left(y_{m}+k_{m},\tilde{\mathcal{F}}(b_{m}+k_{m})\right)\to 0, where ym=~(bm)~is(x1)y_{m}=\tilde{\mathcal{F}}(b_{m})\cap\tilde{\mathcal{L}}^{s}_{i}(x_{1}).

  3. 3.

    (ym+km)x2(y_{m}+k_{m})\to x_{2}.

Proof of Claim 5.8.

By the first item of Proposition 2.9, we can choose kmAmdk_{m}\in A^{m}\mathbb{Z}^{d} and bm~is(x0)b_{m}\in\tilde{\mathcal{L}}^{s}_{i}(x_{0}) such that when m+m\to+\infty, (bm+km)x1(b_{m}+k_{m})\to x_{1}.

Now, let ym=~(bm)~is(x1)y_{m}=\tilde{\mathcal{F}}(b_{m})\cap\tilde{\mathcal{L}}^{s}_{i}(x_{1}). Since |H1Id||H^{-1}-Id| is bounded, H1(~is(x1))H^{-1}\big{(}\tilde{\mathcal{L}}^{s}_{i}(x_{1})\big{)} is located in a neighborhood of H1(~is(x0))H^{-1}\big{(}\tilde{\mathcal{L}}^{s}_{i}(x_{0})\big{)}. Hence, there exists R>0R>0 such that d~(1,i1)s(H1(bm),H1(ym))<Rd_{\tilde{\mathcal{F}}^{s}_{(1,i-1)}}\big{(}H^{-1}(b_{m}),H^{-1}(y_{m})\big{)}<R, for all bm~is(x0)b_{m}\in\tilde{\mathcal{L}}^{s}_{i}(x_{0}). By Proposition 2.6, ym~(bm)y_{m}\in\tilde{\mathcal{F}}(b_{m}) implies

d(H1(ym),~(1,i1)s(H1(bm+km)km,R))0.d\left(H^{-1}(y_{m})\;,\;\tilde{\mathcal{F}}^{s}_{(1,i-1)}\left(H^{-1}(b_{m}+k_{m})-k_{m},R\right)\right)\to 0.

Since ~(1,i1)s\tilde{\mathcal{F}}^{s}_{(1,i-1)} is always d\mathbb{Z}^{d}-periodic (see Proposition 2.15), one has

d(H1(ym)+km,~(1,i1)s(H1(bm+km),R))0.d\left(H^{-1}(y_{m})+k_{m}\;,\;\tilde{\mathcal{F}}^{s}_{(1,i-1)}\left(H^{-1}(b_{m}+k_{m}),R\right)\right)\to 0.

Again, by Proposition 2.6,

d(H1(ym+km),~(1,i1)s(H1(bm+km),R))0.\displaystyle d\left(H^{-1}(y_{m}+k_{m})\;,\;\tilde{\mathcal{F}}^{s}_{(1,i-1)}\left(H^{-1}(b_{m}+k_{m}),R\right)\right)\to 0. (5.12)

It follows that

d(ym+km,~(bm+km))0.d(y_{m}+k_{m}\;,\;\tilde{\mathcal{F}}(b_{m}+k_{m}))\to 0.

So, we complete the second item of claim.

Note that since (bm+km)x1(b_{m}+k_{m})\to x_{1}, (5.12) also implies

d(H1(ym+km),~(1,i1)s(H1(x1),R))0.d\left(H^{-1}(y_{m}+k_{m})\;,\;\tilde{\mathcal{F}}^{s}_{(1,i-1)}\left(H^{-1}(x_{1}),R\right)\right)\to 0.

Hence, d(ym+km,~(x0))0d\left(y_{m}+k_{m},\tilde{\mathcal{F}}(x_{0})\right)\to 0 and we get the third item. ∎

Let z1=~is(x1)~(1,i1)s(x0)z_{1}=\tilde{\mathcal{L}}^{s}_{i}(x_{1})\cap\tilde{\mathcal{L}}^{s}_{(1,i-1)}(x_{0}) and z2=~is(x2)~(1,i1)s(x0)z_{2}=\tilde{\mathcal{L}}^{s}_{i}(x_{2})\cap\tilde{\mathcal{L}}^{s}_{(1,i-1)}(x_{0}). Since x1~(1,i1)s(x0)x_{1}\notin\tilde{\mathcal{L}}^{s}_{(1,i-1)}(x_{0}), we have d(z1,x1)>0d(z_{1},x_{1})>0. See Figure 4.

Claim 5.9.

d(z2,x2)=2d(z1,x1)d(z_{2},x_{2})=2\cdot d(z_{1},x_{1}) and d(x0,z2)=2d(x0,z1)d(x_{0},z_{2})=2\cdot d(x_{0},z_{1}).

Refer to caption
Figure 4: Deduction of HH preserving the strong stable foliations.
Proof of Claim 5.9.

Let ymdy_{m}\in\mathbb{Z}^{d} and bm~is(x0)b_{m}\in\tilde{\mathcal{L}}^{s}_{i}(x_{0}) given by Claim 5.8. Denote

cm=~(1,i1)s(bm)~is(x1),z2=~(1,i1)s(x1)~is(x2)andym=~(bm+km)~is(ym+km).c_{m}=\tilde{\mathcal{L}}^{s}_{(1,i-1)}(b_{m})\cap\tilde{\mathcal{L}}^{s}_{i}(x_{1}),\quad z_{2}^{\prime}=\tilde{\mathcal{L}}^{s}_{(1,i-1)}(x_{1})\cap\tilde{\mathcal{L}}^{s}_{i}(x_{2})\quad{\rm and}\quad y_{m}^{\prime}=\tilde{\mathcal{F}}(b_{m}+k_{m})\cap\tilde{\mathcal{L}}^{s}_{i}(y_{m}+k_{m}).

By the proof of Claim 5.8, we actually have d(ym,ym+km)0d(y_{m}^{\prime},y_{m}+k_{m})\to 0, ymx2y_{m}^{\prime}\to x_{2} and cm+kmz2c_{m}+k_{m}\to z_{2}^{\prime}. Since the holonomy maps of ~is\tilde{\mathcal{L}}^{s}_{i} along ~\tilde{\mathcal{F}} restricted on ~(1,i)s\tilde{\mathcal{L}}^{s}_{(1,i)} are isometric, we have d(cm,ym)=d(z1,x1)d(c_{m},y_{m})=d(z_{1},x_{1}). It follows that d(cm+km,ym+km)=d(z1,x1)d(c_{m}+k_{m},y_{m}+k_{m})=d(z_{1},x_{1}). Hence,

d(z2,x2)=limm+d(cm+km,ym)=limm+d(cm+km,ym+km)=d(z1,x1).d(z_{2}^{\prime},x_{2})=\lim_{m\to+\infty}d(c_{m}+k_{m},y_{m}^{\prime})=\lim_{m\to+\infty}d(c_{m}+k_{m},y_{m}+k_{m})=d(z_{1},x_{1}).

Since the foliation ~is\tilde{\mathcal{L}}^{s}_{i} is one-dimensional, d(z2,x2)=d(z2,z2)+d(z2,x2)=2d(z1,x1)d(z_{2},x_{2})=d(z_{2},z_{2}^{\prime})+d(z_{2}^{\prime},x_{2})=2d(z_{1},x_{1}).

For the other equation,

d(z1,z2)=d(x1,z2)=limm+d(bm+km,cm+km)=limm+d(bm,cm)=d(x0,z1).d(z_{1},z_{2})=d(x_{1},z_{2}^{\prime})=\lim_{m\to+\infty}d(b_{m}+k_{m},c_{m}+k_{m})=\lim_{m\to+\infty}d(b_{m},c_{m})=d(x_{0},z_{1}).

Note that the dimension of ~(1,i1)s\tilde{\mathcal{L}}^{s}_{(1,i-1)} could be more than one, but the line x0z1¯\overline{x_{0}z_{1}} is parallel to the line bmcm¯\overline{b_{m}c_{m}} and also the line (bm+km)(cm+km)¯\overline{(b_{m}+k_{m})(c_{m}+k_{m})}. Hence, x0z1¯\overline{x_{0}z_{1}} is parallel to x1z2¯\overline{x_{1}z_{2}^{\prime}} and also z1z2¯\overline{z_{1}z_{2}}. Thus, we have d(x0,z2)=d(x0,z1)+d(z1,z2)=2d(x0,z1)d(x_{0},z_{2})=d(x_{0},z_{1})+d(z_{1},z_{2})=2d(x_{0},z_{1}). ∎

Repeating the construction in Claim 5.8 and Claim 5.9, there exists a sequence {xl}~(x0),l\{x_{l}\}\subset\tilde{\mathcal{F}}(x_{0}),l\in\mathbb{N} with zl=~is(xl)~(1,i1)s(x0)z_{l}=\tilde{\mathcal{L}}^{s}_{i}(x_{l})\cap\tilde{\mathcal{L}}^{s}_{(1,i-1)}(x_{0}) such that d(zl,xl)=ld(z1,x1)d(z_{l},x_{l})=l\cdot d(z_{1},x_{1}) and d(x0,zl)=ld(x0,z1)d(x_{0},z_{l})=l\cdot d(x_{0},z_{1}). Fix δ>0\delta>0, let Nl>0N_{l}>0 be the minimal number such that ANlzl~(1,i1)s(ANlx0,δ)A^{N_{l}}z_{l}\in\tilde{\mathcal{L}}^{s}_{(1,i-1)}(A^{N_{l}}x_{0},\delta). Since μis(A)>μi1s(A)\mu^{s}_{i}(A)>\mu^{s}_{i-1}(A), we have d(ANlzl,ANlxl)+d(A^{N_{l}}z_{l},A^{N_{l}}x_{l})\to+\infty. It contradicts with the fact that d(y,y)d(y,y^{\prime}) is bounded, for every y~(1,i1)s(x0,δ)y\in\tilde{\mathcal{L}}^{s}_{(1,i-1)}(x_{0},\delta) and y=~is(y)~(x0)y^{\prime}=\tilde{\mathcal{L}}^{s}_{i}(y)\cap\tilde{\mathcal{F}}(x_{0}). ∎

5.2 d\mathbb{Z}^{d}-periodic foliations

Let f𝒰f\in\mathcal{U} given by Proposition 2.15 with λis(p,f)=λis(q,f)\lambda^{s}_{i}(p,f)=\lambda^{s}_{i}(q,f), for all p,qPer(f)p,q\in{\rm Per}(f) and 1ik1\leq i\leq k. Now, we already have gotten Corollay 5.3 that is for each 1ik1\leq i\leq k, H(~is)=~isH(\tilde{\mathcal{F}}^{s}_{i})=\tilde{\mathcal{L}}^{s}_{i} and HH is isometric along ~is\tilde{\mathcal{F}}^{s}_{i} under the metric dis(,)d^{s}_{i}(\cdot,\cdot). It follows that the holonomy maps of ~is\tilde{\mathcal{F}}^{s}_{i} along ~j1s~jls\tilde{\mathcal{F}}^{s}_{j_{1}}\oplus\dots\oplus\tilde{\mathcal{F}}^{s}_{j_{l}} restricted on (~j1s~jls~is)\big{(}\tilde{\mathcal{F}}^{s}_{j_{1}}\oplus\dots\oplus\tilde{\mathcal{F}}^{s}_{j_{l}}\oplus\tilde{\mathcal{F}}^{s}_{i}\big{)} are isometric under dis(,)d^{s}_{i}(\cdot,\cdot), where 1j1<j2<<jlk1\leq j_{1}<j_{2}<...<j_{l}\leq k.

Fix 1i<k1\leq i<k, assume that ~(i,k)s\tilde{\mathcal{F}}^{s}_{(i,k)} is d\mathbb{Z}^{d}-periodic, we show that ~(i+1,k)s\tilde{\mathcal{F}}^{s}_{(i+1,k)} is also d\mathbb{Z}^{d}-periodic. Note that the assumption d\mathbb{Z}^{d}-periodic property for foliation ~(i,k)s\tilde{\mathcal{F}}^{s}_{(i,k)} implies one for ~is=~(1,i)s~(i,k)s\tilde{\mathcal{F}}^{s}_{i}=\tilde{\mathcal{F}}^{s}_{(1,i)}\cap\tilde{\mathcal{F}}^{s}_{(i,k)}.

Proof of Proposition 5.4.

Assume that there exist x0d,ndx_{0}\in\mathbb{R}^{d},n\in\mathbb{Z}^{d} and x1~(i+1,k)s(x0+n)nx_{1}\in\tilde{\mathcal{F}}^{s}_{(i+1,k)}(x_{0}+n)-n, but x1~(i+1,k)s(x0)x_{1}\notin\tilde{\mathcal{F}}^{s}_{(i+1,k)}(x_{0}). Note that, for every xdx\in\mathbb{R}^{d} and ndn\in\mathbb{Z}^{d}, ~(i+1,k)s(x+n)n~(i,k)s(x+n)n=~(i,k)s(x)\tilde{\mathcal{F}}^{s}_{(i+1,k)}(x+n)-n\subset\tilde{\mathcal{F}}^{s}_{(i,k)}(x+n)-n=\tilde{\mathcal{F}}^{s}_{(i,k)}(x). Thus we have x1~(i,k)s(x0)x_{1}\in\tilde{\mathcal{F}}^{s}_{(i,k)}(x_{0}). Let z1=~is(x1)~(i+1,k)s(x0)z_{1}=\tilde{\mathcal{F}}^{s}_{i}(x_{1})\cap\tilde{\mathcal{F}}^{s}_{(i+1,k)}(x_{0}). By the assumption, a:=dis(x1,z1)>0a:=d^{s}_{i}(x_{1},z_{1})>0.

For every bm~is(x0)b_{m}\in\tilde{\mathcal{F}}^{s}_{i}(x_{0}), we denote

cm=~(i+1,k)s(bm)~is(x1)andym=(~(i+1,k)s(n+bm)n)~is(x1).c_{m}=\tilde{\mathcal{F}}^{s}_{(i+1,k)}(b_{m})\cap\tilde{\mathcal{F}}^{s}_{i}(x_{1})\quad{\rm and}\quad y_{m}=\big{(}\tilde{\mathcal{F}}^{s}_{(i+1,k)}(n+b_{m})-n\big{)}\cap\tilde{\mathcal{F}}^{s}_{i}(x_{1}).

We claim dis(ym,cm)=ad^{s}_{i}(y_{m},c_{m})=a. Indeed, since dis(,)d^{s}_{i}(\cdot,\cdot) is holonomy invariant, we have

dis(x0,bm)=dis(z1,cm)anddis(x0+n,bm+n)=dis(x1+n,ym+n).d^{s}_{i}(x_{0},b_{m})=d^{s}_{i}(z_{1},c_{m})\quad{\rm and}\quad d^{s}_{i}(x_{0}+n,b_{m}+n)=d^{s}_{i}(x_{1}+n,y_{m}+n).

Again, ~(i,k)s\tilde{\mathcal{F}}^{s}_{(i,k)} is d\mathbb{Z}^{d}-periodic implies that for ~is\tilde{\mathcal{F}}^{s}_{i}. It follows that

dis(x0,bm)=dis(x0+n,bm+n)anddis(x1+n,ym+n)=dis(x1,ym).d^{s}_{i}(x_{0},b_{m})=d^{s}_{i}(x_{0}+n,b_{m}+n)\quad{\rm and}\quad d^{s}_{i}(x_{1}+n,y_{m}+n)=d^{s}_{i}(x_{1},y_{m}).

Hence, dis(ym,cm)=dis(x1,z1)=ad^{s}_{i}(y_{m},c_{m})=d^{s}_{i}(x_{1},z_{1})=a. See Figure 5.

Refer to caption
Figure 5: Deduction of d\mathbb{Z}^{d}-periodic foliations.

Since we already have H(~is)=~isH(\tilde{\mathcal{F}}^{s}_{i})=\tilde{\mathcal{L}}^{s}_{i}, by the second item of Proposition 2.9, we can choose bm~is(x0)b_{m}\in\tilde{\mathcal{F}}^{s}_{i}(x_{0}) and kmAmdk_{m}\in A^{m}\mathbb{Z}^{d} such that (bm+km)x1(b_{m}+k_{m})\to x_{1}. By Proposition 2.7, for a fixed size R>0R>0, one has

dH(~(i+1,k)s(bm+n,R)+km,~(i+1,k)s(bm+n+km,R))0,d_{H}\left(\tilde{\mathcal{F}}^{s}_{(i+1,k)}(b_{m}+n,R)+k_{m}\;\;,\;\;\tilde{\mathcal{F}}^{s}_{(i+1,k)}(b_{m}+n+k_{m},R)\right)\to 0,

as m+m\to+\infty, where dH(,)d_{H}(\cdot,\cdot) is Hausdorff distance. Since (bm+km)x1(b_{m}+k_{m})\to x_{1} and (x1+n)~(i+1,k)s(x0+n)(x_{1}+n)\in\tilde{\mathcal{F}}^{s}_{(i+1,k)}(x_{0}+n), we have that

dH(~(i+1,k)s(bm+n,R)+kmn,~(i+1,k)s(x1+n,R)n)0.d_{H}\left(\tilde{\mathcal{F}}^{s}_{(i+1,k)}(b_{m}+n,R)+k_{m}-n\;\;,\;\;\tilde{\mathcal{F}}^{s}_{(i+1,k)}(x_{1}+n,R)-n\right)\to 0.

It means that the sequence (ym+km)~(i+1,k)s(bm+n)+kmn(y_{m}+k_{m})\in\tilde{\mathcal{F}}^{s}_{(i+1,k)}(b_{m}+n)+k_{m}-n converges to x2~(i+1,k)s(x0+n)nx_{2}\in\tilde{\mathcal{F}}^{s}_{(i+1,k)}(x_{0}+n)-n.

Similarly, we can get

dH(~(i+1,k)s(bm,R)+km,~(i+1,k)s(x1,R))0.d_{H}\left(\tilde{\mathcal{F}}^{s}_{(i+1,k)}(b_{m},R)+k_{m}\;\;,\;\;\tilde{\mathcal{F}}^{s}_{(i+1,k)}(x_{1},R)\right)\to 0.

It follows that (cm+km)~(i+1,k)s(bm)+km(c_{m}+k_{m})\in\tilde{\mathcal{F}}^{s}_{(i+1,k)}(b_{m})+k_{m} converges to z2~(i+1,k)s(x1)z_{2}^{\prime}\in\tilde{\mathcal{F}}^{s}_{(i+1,k)}(x_{1}). Moreover, since ~is\tilde{\mathcal{F}}^{s}_{i} is d\mathbb{Z}^{d}-periodic, we have

dis(x2,z2)=limm+dis(ym+km,cm+km)=limm+dis(ym,cm)=a.d^{s}_{i}(x_{2},z_{2}^{\prime})=\lim_{m\to+\infty}d^{s}_{i}(y_{m}+k_{m},c_{m}+k_{m})=\lim_{m\to+\infty}d^{s}_{i}(y_{m},c_{m})=a.

Let z2=~is(x2)~(i+1,k)s(x0)z_{2}=\tilde{\mathcal{F}}^{s}_{i}(x_{2})\cap\tilde{\mathcal{F}}^{s}_{(i+1,k)}(x_{0}), by the holonomy-invariant metric, we have dis(x2,z2)=2a.d^{s}_{i}(x_{2},z_{2})=2a.

Repeating the preceding method, we can pick xl~(i+1,k)s(x0+n)nx_{l}\in\tilde{\mathcal{F}}^{s}_{(i+1,k)}(x_{0}+n)-n and zl=~is(xl)~(i+1,k)s(x0)z_{l}=\tilde{\mathcal{F}}^{s}_{i}(x_{l})\cap\tilde{\mathcal{F}}^{s}_{(i+1,k)}(x_{0}) such that

dis(xl,zl)=la+,d^{s}_{i}(x_{l},z_{l})=la\to+\infty,

as l+l\to+\infty. Since ~is\tilde{\mathcal{F}}^{s}_{i} is quasi-isometric (Proposition 2.15), it follows that

d(xl,zl)+,asl+.\displaystyle d(x_{l},z_{l})\to+\infty,\quad{\rm as}\;\;l\to+\infty. (5.13)

On the other hand, since there exists C0>0C_{0}>0 such that |HId|<C0|H-Id|<C_{0}, one has

~(i+1,k)s(x0+n)\displaystyle\tilde{\mathcal{F}}^{s}_{(i+1,k)}(x_{0}+n) BC0(~(i+1,k)s(H(x0+n))),\displaystyle\subset B_{C_{0}}\left(\tilde{\mathcal{L}}^{s}_{(i+1,k)}\big{(}H(x_{0}+n)\big{)}\right),
B3C0(~(i+1,k)s(H(x0)+n)),\displaystyle\subset B_{3C_{0}}\left(\tilde{\mathcal{L}}^{s}_{(i+1,k)}\big{(}H(x_{0})+n\big{)}\right),
=B3C0(~(i+1,k)s(H(x0)))+nB4C0(~(i+1,k)s(x0))+n.\displaystyle=B_{3C_{0}}\left(\tilde{\mathcal{L}}^{s}_{(i+1,k)}\big{(}H(x_{0})\big{)}\right)+n\subset B_{4C_{0}}\left(\tilde{\mathcal{F}}^{s}_{(i+1,k)}(x_{0})\right)+n.

It follows that there exists C4C0C\geq 4C_{0} such that d(xl,zl)Cd(x_{l},z_{l})\leq C, for all ll\in\mathbb{N}. This contradicts with (5.13). ∎

By Remark 5.5, we have that ~is(1ik)\tilde{\mathcal{F}}^{s}_{i}(1\leq i\leq k) is d\mathbb{Z}^{d}-periodic. By Corollary 5.3, H:~is~isH:\tilde{\mathcal{F}}^{s}_{i}\to\tilde{\mathcal{L}}^{s}_{i} is isomtric along each leaf of ~is\tilde{\mathcal{F}}^{s}_{i}. Now, we can show that there is no deviation between H1(x+n)H^{-1}(x+n) and H1(x)+nH^{-1}(x)+n along ~is\tilde{\mathcal{F}}^{s}_{i}. We mention that the following proof can also apply for the case of dimEs=1E^{s}=1 without small perturbation, since H:~s~sH:\tilde{\mathcal{F}}^{s}\to\tilde{\mathcal{L}}^{s} is also isometric by Proposition 5.1.

Proof of Proposition 5.6.

Recall that H1(x+n)n~s(H1(x))H^{-1}(x+n)-n\in\tilde{\mathcal{F}}^{s}(H^{-1}(x)) (see Proposition 2.5). Hence, we just need focus on ~s\tilde{\mathcal{F}}^{s}. For any xdx\in\mathbb{R}^{d} and y~s(x)y\in\tilde{\mathcal{F}}^{s}(x), let

d~is(x,y):=dis(x,y),\tilde{d}^{s}_{i}(x,y):=d^{s}_{i}(x,y^{\prime}),

where y=~is,(y)~is(x)y^{\prime}=\tilde{\mathcal{F}}^{s,\perp}_{i}(y)\cap\tilde{\mathcal{F}}^{s}_{i}(x). Note that the metric d~is(,)\tilde{d}^{s}_{i}(\cdot,\cdot) is well defined, since the holonomy maps of ~is\tilde{\mathcal{F}}^{s}_{i} along ~is,\tilde{\mathcal{F}}^{s,\perp}_{i} restricted on ~s\tilde{\mathcal{F}}^{s} are isometric under dis(,)d^{s}_{i}(\cdot,\cdot). For proving Proposition 5.6, it suffices to show that d~is(H1(x),H1(x+n)n)=0\tilde{d}^{s}_{i}\left(H^{-1}(x)\;,\;H^{-1}(x+n)-n\right)=0, for all xdx\in\mathbb{R}^{d} and ndn\in\mathbb{Z}^{d}.

Assume that there exist xdx\in\mathbb{R}^{d} and ndn\in\mathbb{Z}^{d} such that

d~is(H1(x),H1(x+n)n)=α0.\tilde{d}^{s}_{i}\left(H^{-1}(x)\;,\;H^{-1}(x+n)-n\right)=\alpha\neq 0.
Claim 5.10.

For every ydy\in\mathbb{R}^{d}, d~is(H1(y),H1(y+n)n)=α\tilde{d}^{s}_{i}\left(H^{-1}(y)\;,\;H^{-1}(y+n)-n\right)=\alpha.

Proof of Claim 5.10 .

By Proposition 2.9, there exist zm~is(x)z_{m}\in\tilde{\mathcal{L}}^{s}_{i}(x) and kmAmdk_{m}\in A^{m}\mathbb{Z}^{d} such that (zm+km)y(z_{m}+k_{m})\to y. By Proposition 2.6 and the uniform continuity of HH, as m+m\to+\infty,

d(H1(zm+km),H1(zm)+km)0\displaystyle d\left(H^{-1}(z_{m}+k_{m})\;,\;H^{-1}(z_{m})+k_{m}\right)\to 0\quad andd(H1(zm+km),H1(y))0,\displaystyle{\rm and}\quad d\left(H^{-1}(z_{m}+k_{m})\;,\;H^{-1}(y)\right)\to 0,
d(H1(zm+n)+km,H1(zm+n+km))0\displaystyle d\left(H^{-1}(z_{m}+n)+k_{m}\;,\;H^{-1}(z_{m}+n+k_{m})\right)\to 0\quad andd(H1(zm+km+n),H1(y+n))0.\displaystyle{\rm and}\quad d\left(H^{-1}(z_{m}+k_{m}+n)\;,\;H^{-1}(y+n)\right)\to 0.

Hence,

d(H1(zm)+km,H1(zm+n)n+km)d(H1(y),H1(y+n)n).\displaystyle d\left(H^{-1}(z_{m})+k_{m}\;,\;H^{-1}(z_{m}+n)-n+k_{m}\right)\to d\left(H^{-1}(y)\;,\;H^{-1}(y+n)-n\right). (5.14)

On the other hand, since foliations ~is\tilde{\mathcal{F}}^{s}_{i} and ~is,\tilde{\mathcal{F}}^{s,\perp}_{i} are both d\mathbb{Z}^{d}-periodic,

d~is(H1(zm)+km,H1(zm+n)n+km)=d~is(H1(zm),H1(zm+n)n).\displaystyle\tilde{d}^{s}_{i}\left(H^{-1}(z_{m})+k_{m}\;,\;H^{-1}(z_{m}+n)-n+k_{m}\right)=\tilde{d}^{s}_{i}\left(H^{-1}(z_{m})\;,\;H^{-1}(z_{m}+n)-n\right). (5.15)

Since H1H^{-1} is isometric along ~is\tilde{\mathcal{L}}^{s}_{i}, we have

dis(H1(zm),H1(x))\displaystyle d^{s}_{i}\left(H^{-1}(z_{m})\;,\;H^{-1}(x)\right) =dis(H1(zm+n),H1(x+n)),\displaystyle=d^{s}_{i}\left(H^{-1}(z_{m}+n)\;,\;H^{-1}(x+n)\right),
=dis(H1(zm+n)n,H1(x+n)n).\displaystyle=d^{s}_{i}\left(H^{-1}(z_{m}+n)-n\;,\;H^{-1}(x+n)-n\right).

It follows that

d~is(H1(zm),H1(zm+n)n)=d~is(H1(x),H1(x+n)n)=α.\displaystyle\tilde{d}^{s}_{i}\left(H^{-1}(z_{m})\;,\;H^{-1}(z_{m}+n)-n\right)=\tilde{d}^{s}_{i}\left(H^{-1}(x)\;,\;H^{-1}(x+n)-n\right)=\alpha. (5.16)

Now, combining (5.14) with (5.15) and (5.16), we get

d~is(H1(y),H1(y+n)n)=limm+d~is(H1(zm)+km,H1(zm+n)n+km)=α.\tilde{d}^{s}_{i}\left(H^{-1}(y)\;,\;H^{-1}(y+n)-n\right)=\lim_{m\to+\infty}\tilde{d}^{s}_{i}\left(H^{-1}(z_{m})+k_{m}\;,\;H^{-1}(z_{m}+n)-n+k_{m}\right)=\alpha.

For every ll\in\mathbb{N}, applying the previous claim for H1y=H1(x+(l1)n)(l1)nH^{-1}y=H^{-1}\big{(}x+(l-1)n\big{)}-(l-1)n, one has

d~is(H1(x+ln)ln,H1(x))\displaystyle\tilde{d}^{s}_{i}\left(H^{-1}(x+ln)-ln\;,\;H^{-1}(x)\right) =d~is(H1y,H1(x))+d~is(H1y,H1(x+ln)ln),\displaystyle=\tilde{d}^{s}_{i}\left(H^{-1}y\;,\;H^{-1}(x)\right)+\tilde{d}^{s}_{i}\left(H^{-1}y\;,\;H^{-1}(x+ln)-ln\right),
=(l1)α+α=lα+,\displaystyle=(l-1)\cdot\alpha+\alpha=l\cdot\alpha\to+\infty,

as l+l\to+\infty. Since ~is\tilde{\mathcal{F}}^{s}_{i} is quasi-isometric (Proposition 2.15), it contradicts with the fact that for all xdx\in\mathbb{R}^{d} and all mdm\in\mathbb{Z}^{d}, d(H1(x+m)m,H1(x))d\left(H^{-1}(x+m)-m\;,\;H^{-1}(x)\right) is bounded.

References

  • [1] N. Aoki, K. Hiraide, Topological theory of dynamical systems, Recent advances, North-Holland Mathematical Library, 52, North-Holland Publishing Co., Amsterdam, 1994, viii+416 pp.
  • [2] C. Bonatti, L.J. Díaz, and M. Viana, Dynamics beyond uniform hyperbolicity, A global geometric and probabilistic perspective, Encyclopaedia of Mathematical Sciences, 102, Mathematical Physics, III Springer-Verlag, Berlin, 2005, xviii+384 pp.
  • [3] M. Brin, On dynamical coherence, Ergodic Theory Dynam. Systems, 23 (2003), no. 2, 395–401.
  • [4] M. Brin, D. Burago, and S. Ivanov, Dynamical coherence of partially hyperbolic diffeomorphisms of the 3-torus J. Mod. Dyn., 3 (2009), no. 1, 1–11.
  • [5] S. Crovisier, R. Potrie Introduction to partially hyperbolic dynamics, Unpublished course notes, School on Dynamical Systems, ICTP, Trieste, 2015.
  • [6] J. De Simoi, C. Liverani, Statistical properties of mostly contracting fast-slow partially hyperbolic systems, Invent. Math., 206 (2016), no. 1, 147–227.
  • [7] J. De Simoi, C. Liverani, Limit theorems for fast-slow partially hyperbolic systems, Invent. Math., 213 (2018), no. 3, 811–1016.
  • [8] J. Franks, Anosov diffeomorphisms on tori, Trans. Amer. Math. Soc., 145 (1969), 117–124.
  • [9] J. Franks, Anosov diffeomorphisms, 1970 Global Analysis (Proc. Sympos. Pure Math., Vol. XIV, Berkeley, Calif., 1968) pp. 61–93, Amer. Math. Soc., Providence, R.I..
  • [10] S. Gan, Y. Ren, and P. Zhang, Accessibility and homology bounded strong unstable foliation for Anosov diffeomorphisms on 3-torus, Acta Math. Sin. (Engl. Ser.), 33 (2017), no. 1, 71–76.
  • [11] S. Gan, Y. Shi, Rigidity of center Lyapunov exponents and susu-integrability, Comment. Math. Helv., 95 (2020), no. 3, 569–592.
  • [12] A. Gogolev, Smooth conjugacy of Anosov diffeomorphisms on higher-dimensional tori, J. Mod. Dyn., 2 (2008), no. 4, 645–700.
  • [13] A. Gogolev, Bootstrap for local rigidity of Anosov automorphisms on the 3-torus, Comm. Math. Phys., 352 (2017), no. 2, 439–455.
  • [14] A. Gogolev, M. Guysinsky, C1C^{1}-differentiable conjugacy of Anosov diffeomorphisms, on three dimensional torus, Discrete Contin. Dyn. Syst., 22 (2008), no. 1–2, 183–200.
  • [15] A. Gogolev, B. Kalinin and V. Sadovskaya, Local rigidity for Anosov automorphisms, Math. Res. Lett., 18 (2011), no. 5, 843–858.
  • [16] A. Gogolev, B. Kalinin, and V. Sadovskaya, Local rigidity of Lyapunov spectrum for toral automorphisms, Israel J. Math., 238 (2020), no. 1, 389–403.
  • [17] A. Gogolev, Y. Shi, Spectrum rigidity and jointly integrability for Anosov diffeomorphisms, Preprint.
  • [18] L. Hall, A. Hammerlindl, Partially hyperbolic surface endomorphisms, Theory Dynam. Systems, 41 (2021), no. 1, 272–282.
  • [19] L. Hall, A. Hammerlindl, Classification of partially hyperbolic surface endomorphisms, Geom. Dedicata, 216 (2022), no. 3, Paper No. 29.
  • [20] B. He, Y. Shi, and X. Wang, Dynamical coherence of specially absolutely partially hyperbolic endomorphisms on 𝕋2\mathbb{T}^{2}, Nonlinearity, 32 (2019), no. 5, 1695–1704.
  • [21] M. Hirsch, C. Pugh, and M. Shub, Invariant manifolds, Lecture Notes in Mathematics, Vol. 583, Springer-Verlag, Berlin-New York, 1977, ii+149 pp.
  • [22] J.-L. Journé, A regularity lemma for functions of several variables, Rev. Mat. Iberoamericana, 4 (1988), no. 2, 187–193.
  • [23] A. Katok, B. Hasselblatt, Introduction to the modern theory of dynamical systems, With a supplementary chapter by Katok and Leonardo Mendoza, Encyclopedia of Mathematics and its Applications, 102, Cambridge University Press, Cambridge, 1995, xviii+802 pp.
  • [24] R. de la Llave, Smooth conjugacy and S-R-B measures for uniformly and non-uniformly hyperbolic systems, Comm. Math. Phys., 150 (1992), no. 2, 289–320.
  • [25] R. Mañé, C. Pugh, Stability of endomorphisms, Dynamical systems—Warwick 1974 (Proc. Sympos. Appl. Topology and Dynamical Systems, Univ. Warwick, Coventry, 1973/1974; presented to E. C. Zeeman on his fiftieth birthday), pp 175–184, Lecture Notes in Math., Vol. 468, Springer, Berlin, 1975.
  • [26] A. Manning, There are no new Anosov diffeomorphisms on tori, Amer. J. Math., 96 (1974), 422–429.
  • [27] C. Marisa, V. Régis, Anosov Endomorphisms on the 2-torus: Regularity of foliations and rigidity, Preprint, arXiv: 2104.01693, (2021).
  • [28] F. Micena, Rigidity for some cases of Anosov endomorphisms of torus, Preprint, arXiv: 2006.00 407, (2022).
  • [29] F. Micena, A. Tahzibi, On the unstable directions and Lyapunov exponents of Anosov endomorphisms, Fund. Math., 235 (2016), no. 1, 37–48.
  • [30] S. Moosavi, K. Tajbakhsh, Classification of special Anosov endomorphisms of nil-manifolds, Acta Math. Sin. (Engl. Ser.), 35 (2019), no. 12, 1871–1890.
  • [31] S. Moosavi, K. Tajbakhsh, Robust special Anosov endomorphisms, Commun. Korean Math. Soc., 34 (2019), no. 3, 897–910.
  • [32] Ya. Pesin, Lectures on partial hyperbolicity and stable ergodicity, Zurich Lectures in Advanced Mathematics, European Mathematical Society (EMS), Zürich, 2004, vi+122 pp.
  • [33] F. Przytycki, Anosov endomorphisms, Studia Math., 58 (1976), no. 3, 249–285.
  • [34] C. Pugh, M. Shub, Stably ergodic dynamical systems and partial hyperbolicity, J. Complexity, 13 (1997), no. 1, 125–179.
  • [35] C. Pugh, M. Shub, and A. Wilkinson, Hölder foliations, Duke Math. J., 86 (1997), no. 3, 517–546.
  • [36] R. Saghin, J. Yang, Lyapunov exponents and rigidity of Anosov automorphisms and skew products, Adv. Math., 355 (2019), 106764, 45 pp.
  • [37] M. Tsujii, Physical measures for partially hyperbolic surface endomorphisms, Acta Math., 194 (2005), no. 1, 37–132.

Jinpeng An

School of Mathematical Sciences, Peking University, Beijing, 100871, China

E-mail: [email protected]