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

Anosov-Katok constructions for quasi-periodic SL(2,)\mathrm{SL}(2,{\mathbb{R}}) cocycles

Nikolaos Karaliolios Unaffiliated, Université de Lille. [email protected] Xu Xu Department of Mathematics, Nanjing University, Nanjing 210093, China [email protected]  and  Qi Zhou Chern Institute of Mathematics and LPMC, Nankai University, Tianjin 300071, China [email protected]
Abstract.

We prove that if the frequency of the quasi-periodic SL(2,)\mathrm{SL}(2,{\mathbb{R}}) cocycle is Diophantine, then the following properties are dense in the subcritical regime: for any 12<κ<1\frac{1}{2}<\kappa<1, the Lyapunov exponent is exactly κ\kappa-Hölder continuous; the extended eigenstates of the potential have optimal sub-linear growth; and the dual operator associated a subcritical potential has power-law decay eigenfunctions. The proof is based on fibered Anosov-Katok constructions for quasi-periodic SL(2,)\mathrm{SL}(2,{\mathbb{R}}) cocycles.

1. Introduction

In this paper, we are concerned with one-dimensional analytic quasi-periodic Schrödinger operators:

(1.1) (HV,α,θu)n=un+1+un1+V(θ+nα)un,(H_{V,\alpha,\theta}u)_{n}=u_{n+1}+u_{n-1}+V(\theta+n\alpha)u_{n},

where VCω(𝕋d,)V\in C^{\omega}({\mathbb{T}}^{d},{\mathbb{R}}) is the potential, α𝕋d\alpha\in{\mathbb{T}}^{d} is the frequency, assumed to be rationally independent, and θ𝕋d\theta\in{\mathbb{T}}^{d} is the phase. Due to the rich implications in quantum physics, quasi-periodic Schrödinger operators have been extensively studied [46]. Starting from the early 1980’s, there was already an almost periodic flu which swept the world, as pointed out by Simon [49]. In 2000’s, adopting a dynamical systems point of view (mainly analytic quasi-periodic cocycles) was found to be usefull in the study of such operators (1.1), and much progress has been made since then [2, 3, 12, 6].

We recall that an analytic quasi-periodic cocycle (α,A)𝕋d×Cω(𝕋d,SL(2,))(\alpha,A)\in{\mathbb{T}}^{d}\times C^{\omega}({\mathbb{T}}^{d},SL(2,{\mathbb{R}})) is a linear skew product system:

(α,A):{𝕋d×2𝕋d×2(θ,v)(x+α,A(θ)v)\displaystyle(\alpha,A):\left\{\begin{array}[]{c}{\mathbb{T}}^{d}\times{\mathbb{R}}^{2}\rightarrow{\mathbb{T}}^{d}\times{\mathbb{R}}^{2}\\ (\theta,v)\longmapsto(x+\alpha,A(\theta)\cdot v)\end{array}\right.

Among many important advances, we should highlight Avila’s global theory of one-frequency analytic quasi-periodic cocycles [2]: if d=1d=1, any (α,A)(\alpha,A) which is not uniformly hyperbolic, is either supercritical, subcritical or critical, properties that can be expressed in terms of the growth of the cocycle. More precisely, (α,A)(\alpha,A) is said to be

  1. (1)

    Supercritical, if supz𝕋A(z;n)\sup_{z\in{\mathbb{T}}}\|A(z;n)\| grows exponentially;

  2. (2)

    Subcritical, if there is a uniform subexponential bound on the growth of A(z;n)\|A(z;n)\| through some band |z|<δ|\Im z|<\delta;

  3. (3)

    Critical, otherwise.

Here we define the products of cocycles as (α,A)n=(nα,A(;n))(\alpha,A)^{n}=(n\alpha,A(\cdot;n)), where A(,0)=id,A(\cdot,0)=\operatorname{id},

A(;n)=A(+(n1)α)A(),forn1,A(\cdot;n)=A(\cdot+(n-1)\alpha)\cdots A(\cdot),\qquad{\rm for}\ n\geq 1,

and A(;n)=A(nα;n)1A(\cdot;-n)=A(\cdot-n\alpha;n)^{-1} for n1n\geq 1. We also recall that (α,A)(\alpha,A) is uniformly hyperbolic if there exists a continuous invariant splitting 2=Es(θ)Eu(θ){\mathbb{R}}^{2}=E^{s}(\theta)\oplus E^{u}(\theta) with the following property: there exist C>0C>0 and λ>0\lambda>0 such that A(θ;n)vCeλnv\|A(\theta;n)\cdot v\|\leq Ce^{-\lambda n}\|v\| holds for every n0n\geq 0 and every vEs(x)v\in E^{s}(x). Simultaneously A(θ;n)vCeλnv\|A(\theta;-n)\cdot v\|\leq Ce^{-\lambda n}\|v\| holds for every vEu(x)v\in E^{u}(x). We will denote the set of such cocycles as 𝒰α\mathcal{UH}_{\alpha} for short111In this paper, we will fix the frequency α\alpha, only vary the mapping AA in the fiber..

The fundamental observation, allowing for the dynamical systems point of view to be taken, is that (1.1) can be seen as a quasi-periodic SL(2,)\mathrm{SL}(2,{\mathbb{R}}) cocycle, since a sequence (un)n(u_{n})_{n\in{\mathbb{Z}}} is a formal solution of the eigenvalue equation HV,α,θu=EuH_{V,\alpha,\theta}u=Eu if and only if it satisfies

(1.2) (un+1un)=SEV(θ+nα)(unun1),\begin{pmatrix}u_{n+1}\\ u_{n}\end{pmatrix}=S_{E}^{V}(\theta+n\alpha)\cdot\begin{pmatrix}u_{n}\\ u_{n-1}\end{pmatrix},

where

SEV(θ)=(EV(θ)110)SL(2,).\displaystyle S_{E}^{V}(\theta)=\left(\begin{array}[]{ccc}E-V(\theta)&-1\cr 1&0\end{array}\right)\in SL(2,\mathbb{R}).

Thus, corresponding to (1.1), there is a family of naturally defined cocycles (α,SEV)(\alpha,S_{E}^{V}), called quasi-periodic Schrödinger cocycles. It is well-known that EΣV,αE\in\Sigma_{V,\alpha}, the spectrum of (1.1), if and only if (α,SEλ)𝒰α(\alpha,S_{E}^{\lambda})\notin\mathcal{UH}_{\alpha}. Therefore energies in the spectrum of HV,α,θH_{V,\alpha,\theta} can be characterized as supercritical, subcritical or critical in terms of its corresponding cocycle (α,SEV)(\alpha,S_{E}^{V}).

A cornerstone in Avila’s global theory is the “Almost Reducibility Theorem ”(ART) [2, 4, 5], which asserts that (α,A)(\alpha,A) is almost reducible (denoted by 𝒜α\mathcal{AR}_{\alpha} for any fixed α\alpha) if it is subcritical. We recall that a cocycle (α,A)(\alpha,A) is almost reducible, if the closure of its analytic conjugacy class contains a constant cocycle.

In this paper, we will focus on the dynamical and spectral behavior of (α,SEV)(\alpha,S_{E}^{V}) in the subcritical regime. We remark that our main results work also in the multifrequency case, d2d\geq 2, since they are proved in the regime 𝒜α\𝒰α\mathcal{AR}_{\alpha}\backslash\mathcal{UH}_{\alpha}. Indeed, by Avila’s global theory and ART [2, 4, 5], if α\\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}}, A𝒜α\𝒰αA\in\mathcal{AR}_{\alpha}\backslash\mathcal{UH}_{\alpha} if and only if (α,A)(\alpha,A) is subcritical.

1.1. Regularity of Lyapunov exponent

Our first result concerns the regularity of Lyapunov exponent (LE):

L(α,A):=limn1n𝕋dlnA(θ;n)dθL(\alpha,A):=\lim\limits_{n\to\infty}\frac{1}{n}\int_{{\mathbb{T}}^{d}}\ln\|A(\theta;n)\|d\theta

as a function of the mapping A()A(\cdot) defining the dynamics in the fibers. The LE is a central topic in the spectral theory of Schrödinger operators, since it relates with intergrated density of states through the Thouless formula. The LE also arise naturally in the study of smooth dynamics, and continuity of LE has been the object of important recent research, see [50] and the references therein.

In the analytic topology, the LE is always continuous with respect to both α{\alpha} and AA at any given cocycle (α,A)(\alpha,A), provided that α𝕋d\alpha\in{\mathbb{T}}^{d} is rationally independent, see [18, 21, 35]. This holds true even for cocycles in higher dimensional groups GL(d,)GL(d,{\mathbb{C}}), with α\\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}} [9]. If αDCd\alpha\in DC_{d}222We recall that α\alpha is Diophantine, denoted by αDCd(κ,τ)\alpha\in{\rm DC}_{d}(\kappa^{\prime},\tau), if there exist κ>0\kappa^{\prime}>0 and τ>d1\tau>d-1 such that (1.4) DCd(κ,τ):={α𝕋d:infj|n,αj|>κ|n|τ,nd\{0}}.{\rm DC}_{d}(\kappa^{\prime},\tau):=\left\{\alpha\in{\mathbb{T}}^{d}:\inf_{j\in{\mathbb{Z}}}\left|\langle n,\alpha\rangle-j\right|>\frac{\kappa^{\prime}}{|n|^{\tau}},\quad\forall\ n\in{\mathbb{Z}}^{d}\backslash\{0\}\right\}. The set of Diophantine numbers is denoted by DCd:=κ>0,τ>d1DCd(κ,τ){\rm DC}_{d}:=\bigcup_{\kappa^{\prime}>0,\,\tau>d-1}{\rm DC}_{d}(\kappa^{\prime},\tau)., and the potential VV is small enough (the smallness of VV depending on α\alpha), then L(α,SEV)L(\alpha,S_{E}^{V}) is 12\frac{1}{2} Hölder continuous w.r.t EE [15]. Subsequently, Avila-Jitomirskaya in [7] generalized the 12\frac{1}{2}-Hölder continuity to the non-perturbative regime, lifting the dependence of the smallness of VV on α\alpha. By the ART, see [2, 4], the LE is 12\frac{1}{2}-Hölder continuous in the whole subcritical regime [47]. This kind of 12\frac{1}{2}-Hölder continuity is sharp, since the LE is exactly 12\frac{1}{2} Hölder continuous at the end of spectral gaps by [48]. On the other hand, if (α,A)(\alpha,A) is conjugated to constant SO(2,)\mathrm{SO}(2,{\mathbb{R}}) cocycle or A𝒰αA\in\mathcal{UH}_{\alpha}, then LE is Lipschitz. Therefore a natural question is whether there exist subcritical cocycles such that the optimal Hölder exponent of L(α,A)L(\alpha,A) can be any fixed number between 12\frac{1}{2} and 11. In fact, we will show that this holds in a dense set in 𝒜α\𝒰α\mathcal{AR}_{\alpha}\backslash\mathcal{UH}_{\alpha} for each admissible Hölder exponent.

Theorem 1.1.

Let 12<κ<1\frac{1}{2}<\kappa<1, αDCd\alpha\in\mathrm{DC}_{d}. There exists a set 𝔖\mathfrak{S} which is dense in 𝒜α\𝒰α\mathcal{AR}_{\alpha}\backslash\mathcal{UH}_{\alpha} in the Cω(𝕋d,SL(2,))C^{\omega}({\mathbb{T}}^{d},\mathrm{SL}(2,{\mathbb{R}})) topology such that Lyapunov exponent is exactly κ\kappa-Hölder continuous at each point of 𝔖\mathfrak{S} in the sense that for any A𝔖A\in\mathfrak{S} we have for BC(𝕋d,SL(2,))B\in C({\mathbb{T}}^{d},\mathrm{SL}(2,{\mathbb{R}})),

lim infBA00log|L(α,A)L(α,B)|logBA0=κ.\liminf_{\|B-A\|_{0}\rightarrow 0}\frac{\log|L(\alpha,A)-L(\alpha,B)|}{\log\|B-A\|_{0}}=\kappa.

Some arithmetic condition on α\alpha is needed for the theorem to be true. As pointed out by Avila-Jitomirskaya [7], the Lyapunov Exponent is discontinuous at rational α\alpha, which implies then for generic α\\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}}, LE cannot be Hölder continuous at any order. Recently, Avila-Last-Shamis-Zhou [11] showed that if α\alpha is very Liouvillean, then LE can even be not logt\log^{t}-Hölder continuous.

In the positive Lyapunov Exponent regime, Goldstein and Schlag proved that the LE is Hölder continuous (for one-frequency cocycles) or weak Hölder continuous (for multi-frequency cocycles) for Schrödinger operators whose frequency satisfies a strong Diophantine condition [30, 31].

Continuity of the LE also depends sensitively on the smoothness of A()A(\cdot). In the C0C^{0} topology, any non uniformly hyperbolic SL(2,)\mathrm{SL}(2,{\mathbb{R}}) cocycle can be approximated by cocycles with zero LE, see [16, 17], and thus the LE is not continuous. In the CkC^{k} (kk\in{\mathbb{N}}) topology, results are quite different. Wang-You, [51], showed that the LE can be discontinuous. On the other hand, Xu-Ge-Wang, [53], recently constructed a class of C2+ϵC^{2+\epsilon} cos\cos-type potential, and for any 12κ<1\frac{1}{2}\leq\kappa<1 showed the existence of an energy, such that LE is exactly κ\kappa-Hölder continuous. For other results of regularity of LE in the smooth or in the Gervey topology, one can consult [22, 52, 45] and the references therein.

1.2. Optimal sub-linear growth of extended eigenstates

Next, we move to spectral applications. In contrast with 1D random Schrödinger operator, one of the most remarkable phenomena exhibited by quasiperiodic Schrödinger operators is the existence of absolutely continuous (ac) spectrum. If αDCd\alpha\in\mathrm{DC}_{d} and VV is small enough, then Dinaburg-Sinai in [23] proved that HV,α,θH_{V,\alpha,\theta} has ac spectrum. Eliasson in [24] proved the stronger result that HV,α,θH_{V,\alpha,\theta} has purely ac spectrum for every θ\theta. Avila’s ART [2, 4, 5] ensures that if the potential is subcritical, then the corresponding operator has purely ac spectrum. From the physicist’s point of view, ac spectrum corresponds to a phase where the material is a conductor, and the corresponding eigenstate is extended. Indeed, one can define the inverse participation ratio (IPR) [26] as

IPR(m)=n=1L|um(n)|4n=1L|um(n)|2,{\rm IPR}(m)=\frac{\sum_{n=1}^{L}|u_{m}(n)|^{4}}{\sum_{n=1}^{L}|u_{m}(n)|^{2}},

where umu_{m} is the mthm-th eigenstate. If Γ=limLln(IPR)lnL=1\Gamma=-\lim_{L\rightarrow\infty}\frac{\ln({\rm IPR})}{\ln L}=1, then the phase is extended. Thus, conductivity of the material is reflected in the growth of its eigenstate.

For αDCd\alpha\in\mathrm{DC}_{d} and VV small enough, Eliasson [24] proved that if the energy lies in the end of spectral gaps, then its extended eigenstates have linear growth. On the other hand, for any EΣV,αE\in\Sigma_{V,\alpha}, the spectrum of the operator, the extended eigenstates at most have sub linear growth, i.e. |uE(n)|o(n)|u_{E}(n)|\leq o(n). This kind of behavior can be generalized to the whole subcritical regime [2, 4]. Thus it is interesting to ask whether this kind of sub-linear growth was optimal. In this paper, we prove that for a dense subset of subcritical potentials, the extended eigenstates of the associated operator have optimal sub-linear growth.

Theorem 1.2.

Fix αDC\alpha\in\mathrm{DC}. For any non-increasing sequence {g(n)}n=1\{g(n)\}_{n=1} satisfying 0<g(n)<10<g(n)<1 and limnng(n)=\lim_{n\rightarrow\infty}n^{g(n)}=\infty, for any VChω(𝕋,)V\in C^{\omega}_{h}({\mathbb{T}},{\mathbb{R}}), if EΣV,αsubE\in\Sigma_{V,\alpha}^{sub}, then for any ε>0\varepsilon>0, there exist 0<h<h0<h^{\prime}<h, VChω(𝕋,)V\in C^{\omega}_{h^{\prime}}({\mathbb{T}},{\mathbb{R}}) with VVhε\|V-V^{\prime}\|_{h^{\prime}}\leq\varepsilon, such that EΣV,αsubE\in\Sigma_{V^{\prime},\alpha}^{sub}. Moreover, the eigenstate uE(n)u_{E}(n) of

(HV,α,θuE)(n)=EuE(n),(H_{V^{\prime},\alpha,\theta}u_{E})(n)=Eu_{E}(n),

has sub linear growth with rate {g(n)}n=1\{g(n)\}_{n=1}, i.e. there exist {nj}j=1\{n_{j}\}_{j=1}^{\infty} and c,Cc,C not depending on jj such that

c|nj|1g(nj)<|uE(nj)|<C|nj|112g(nj).c|n_{j}|^{1-g(n_{j})}<|u_{E}(n_{j})|<C|n_{j}|^{1-\frac{1}{2}g(n_{j})}.
Remark 1.1.

On the right side of the inequality, one can replace 112g(nj)1-\frac{1}{2}g(n_{j}) by 1(1δ)g(nj)1-(1-\delta)g(n_{j}) for any δ>0\delta>0.

Theorem 1.2 shows that there exists extended eigenfunctions with optimal sub-linear growth. We remark howevert, that the universal hierarchical structure of any extended quasi-periodic eigenfunctions for almost Mathieu operators was recently obtained in [33]

The study of growth of extended eigenstates is not only interesting from the physicist’s point of view, but also important from the mathematical point of view. By (1.2), the study of the growth of the extended eigenstates is equivalent to study the growth of the Schrödinger cocycle. On one hand, the growth of Schrödinger cocycle is crucial for proving existence of ac spectrum. In the subcritical regime, for almost every EΣV,αE\in\Sigma_{V,\alpha}, the cocycles are uniformly bounded, see [24], which is sufficient for establishing the existence of ac spectrum by subordinacy theory [29]. However, whether the spectral measure is purely ac or not really depends on the growth of SEV(θ;j)\|S_{E}^{V}(\theta;j)\| on a Lebesgue zero measure set of energies by [3]. On the other hand, the growth of Schrödinger cocycle is important in proving the regularity of the spectral measure, see [8]. Indeed, using Jitomiskaya-Last’s inequality, see [37], Avila-Jitomirskaya in [8] showed that

μ(Eϵ,E+ϵ)CϵPL,\mu(E-\epsilon,E+\epsilon)\leq C\epsilon\|P_{L}\|,

where PL=n=1L(SEV)(θ+α;2n1)SEV(θ+α;2n1)P_{L}=\sum_{n=1}^{L}(S_{E}^{V})^{*}(\theta+\alpha;2n-1)S_{E}^{V}(\theta+\alpha;2n-1) satisfies detPL=1/4ϵ2\det P_{L}=1/4\epsilon^{2}. Therefore, by Theorem 1.2, it seems interesting to study high order Hölder continuity of spectral measure for a dense set of subcritical energies.

1.3. Power-law decay eigenfunctions

Our third result concerns the localization property of the quasi-periodic long-range operator:

(1.5) (V,α,ϕu)n=kdVkunk+2cos2π(ϕ+n,α)un,(\mathcal{L}_{V,\alpha,\phi}u)_{n}=\sum_{k\in{\mathbb{Z}}^{d}}V_{k}u_{n-k}+2\cos 2\pi(\phi+\langle n,\alpha\rangle)u_{n},

where VkV_{k}\in{\mathbb{C}} is the Fourier coefficient of VCω(𝕋d,)V\in C^{\omega}({\mathbb{T}}^{d},{\mathbb{R}}). This operator has received a lot of attention, see [7, 13, 20] since it is the Aubry dual of the quasi-periodic Schrödinger operator defined in eq. (1.1). If V(θ)=2cosθV(\theta)=2\cos\theta, then it reduces to the extensively studied almost Mathieu operator:

(Hλ,α,θu)n=un+1+un1+2λcos(θ+nα)un.(H_{\lambda,\alpha,\theta}u)_{n}=u_{n+1}+u_{n-1}+2\lambda\cos(\theta+n\alpha)u_{n}.

For the almost Mathieu operator, there is a sharp phase transition line λ=eβ(α)\lambda=e^{\beta(\alpha)} 333 Here, β(α):=lim supnlnqn+1qn,\beta(\alpha):=\limsup_{n\rightarrow\infty}\frac{\ln q_{n+1}}{q_{n}}, where pnqn\frac{p_{n}}{q_{n}} is the continued fraction best rationnal approximants of α\\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}}. from singular continuous spectrum to Anderson localization (pure point spectrum with exponentially decaying eigenfunctions), see [13, 38]. In the transition line λ=eβ(α)\lambda=e^{\beta(\alpha)}, for frequencies in a dense set, Hλ,α,θH_{\lambda,\alpha,\theta} displays pure point spectrum, see [10], but the eigenfunction does not decay exponentially, see [38]. As pointed out in [10], the insights gained from the critical parameters often shed light on the creation, dissipation, and the mechanism behind the phases of non critical parameters as well. Thus it is interesting to ask whether there exists real power law decay eigenfunction, i.e. if the eigenfunctions can decay polynomially. In this paper, we establish the following result.

Theorem 1.3.

Let αDCd\alpha\in\mathrm{DC}_{d}, 0<h<h0<h_{*}<h and s+s\in{\mathbb{N}}^{+}. Then, there exist ε0=ε0(α,h,h)\varepsilon_{0}=\varepsilon_{0}(\alpha,h,h_{*}) and 0<h<h0<h_{*}<h, such that for any ε<ε0\varepsilon<\varepsilon_{0}, any

Vh(ε):={VChω(𝕋,)|Vh<ε}V\in\mathcal{B}_{h}(\varepsilon):=\{\ V\in C^{\omega}_{h}({\mathbb{T}},{\mathbb{R}})\ \big{|}\ \|V\|_{h}<\varepsilon\}

is accumulated by Vkh(ε)V_{k}\in\mathcal{B}_{h_{*}}(\varepsilon) such that Vk,α,ϕ\mathcal{L}_{V_{k},\alpha,\phi} has point spectrum with eigenfunction uhs\hs+1u\in h^{s}\backslash h^{s+1} for some ϕ\phi\in{\mathbb{R}}, where we denote hs={ul2:k|k|s|uk|<}h^{s}=\{u\in l^{2}:\sum_{k}|k|^{s}|u_{k}|<\infty\}.

Just as Theorem 1.2, Theorem 1.3 also holds if the dual Schrödigner operator (1.1) lies in the subcritical regime. However, under the assumption that αDCd\alpha\in\mathrm{DC}_{d}, the phenomenon exhibited by Theorem 1.3 was not expected for Schrödigner operators (1.1). Indeed, if V(θ)V(\theta) is an even function, then for a GδG_{\delta} dense set of θ\theta, HV,α,θH_{V,\alpha,\theta} has no eigenvalues, see [40]. It is widely believed for a.e. θ\theta, HV,α,θH_{V,\alpha,\theta} has Anderson localization in the positive LE regime. In fact, Hλ,α,θH_{\lambda,\alpha,\theta} even exhibits a sharp transition in the phase θ\theta between singular continuous spectrum and Anderson localization [39].

If d=1d=1, Bourgain-Jitomirskaya in [20] proved that for any fixed αDC\alpha\in DC, λV,α,ϕ\mathcal{L}_{\lambda V,\alpha,\phi} has Anderson localization for sufficiently small λ\lambda and a.e. ϕ\phi. If α\\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}}, not necessary Diophantine, then for dense set of small potentials VV and a.e. ϕ\phi, V,α,ϕ\mathcal{L}_{V,\alpha,\phi} has point spectrum with exponentially decaying eigenfunctions, see [56], while it is still open whether V,α,ϕ\mathcal{L}_{V,\alpha,\phi} has Anderson localization. If d2d\geq 2, Jitomirskaya-Kachkovskiy in [36] proved that for fixed αDCd\alpha\in DC_{d}, LλV,α,θL_{\lambda V,\alpha,\theta} has pure point spectrum for sufficiently small λ\lambda and a.e.a.e. θ\theta. Recently, Ge-You-Zhou [32] further proved that under the same assumption, LλV,α,θL_{\lambda V,\alpha,\theta} has exponentially dynamical localization.

1.4. Density coycles reducible in finite differentiability

Recall that a cocycle (α,A)𝕋d×Cω(𝕋d,SL(2,))(\alpha,A)\in{\mathbb{T}}^{d}\times C^{\omega}({\mathbb{T}}^{d},\mathrm{SL}(2,{\mathbb{R}})) is CsC^{s} reducible if there exists BCs(2𝕋d,SL(2,))B\in C^{s}(2{\mathbb{T}}^{d},\mathrm{SL}(2,{\mathbb{R}})) such that

B1(+α)A()B()SL(2,),B^{-1}(\cdot+\alpha)A(\cdot)B(\cdot)\in\mathrm{SL}(2,{\mathbb{R}}),

then by Aubry duality, Theorem 1.3 is an immediately corollary of the following reducibility result.

Theorem 1.4.

Given s+s\in{\mathbb{N}}^{+}, αDCd\alpha\in\mathrm{DC}_{d}. There exists a set 𝔉\mathfrak{F} which is dense in 𝒜α\𝒰α\mathcal{AR}_{\alpha}\backslash\mathcal{UH}_{\alpha} in the Cω(𝕋d,SL(2,))C^{\omega}({\mathbb{T}}^{d},\mathrm{SL}(2,{\mathbb{R}})) topology, such that if A𝔉A\in\mathfrak{F}, then (α,A)(\alpha,A) is CsC^{s}-reducible but not Cs+1C^{s+1}-reducible.

We point out that this result is interesting in itself, from the dynamical systems point of view. The study of the reducibility of cocycles is related with the linearization of circle diffeomorphisms. Arnol’d in [1] proved that if an analytic diffeomorphism fDiffω(𝕋)f\in\mathrm{Diff}^{\omega}({\mathbb{T}}) is close to a rotation TρT_{\rho}, where ρ\rho is the rotation number of ff, and ρDC\rho\in DC, then ff is analytically linearizable. One of the great achievements of Herman-Yoccoz, see [34, 54], is the proof of the fact that the sharp arithmetic condition for CC^{\infty} linearizability of fDiff(𝕋)f\in\mathrm{Diff}^{\infty}({\mathbb{T}}), without any smallness condition imposed a priori on ff, is that its rotation number be Diophantine. In the Liouvillean regime, Herman [34] (see also [44]) proved that for any s+s\in{\mathbb{N}}^{+} and any Liouvillean number ρ\rho, the set of fDiff(𝕋)f\in\mathrm{Diff}^{\infty}({\mathbb{T}}) with rotation number ρ\rho which is CsC^{s} but not Cs+1C^{s+1} linearizable is locally dense. For a more precise description of the theory, we refer the reader for the recent survey of Eliasson-Fayad-Krikorian [25] on circle diffeomorphisms.

Concerning the reducibility of SL(2,)\mathrm{SL}(2,{\mathbb{R}}) cocycles, if the base frequency αDCd\alpha\in DC_{d}, and the cocycle (α,A)(\alpha,A) is sufficiently close to constants (the closeness depends on α\alpha), Eliasson [24] proved that if the fibered rotation number is Diophantine w.r.t α\alpha, then (α,A)(\alpha,A) is analytically reducible. By global to local reduction, reducibility for a full measure set of frequencies holds in the non-perturbative regime, see [48], and in the subcritical regime [4]. Motivated by the results in circle diffeomorphism, then it is natural to study the density of CsC^{s} but not Cs+1C^{s+1} reducible cocycles. Theorem 1.4 provides a positive answer to this question. The corresponding theorem in the case of CC^{\infty} smooth SU(2)SU(2) cocycles, which in some sense stands in midway between circle diffemorphisms and SL(2,)\mathrm{SL}(2,{\mathbb{R}}) cocycles, was proved by the first author in [41].

1.5. Some comments on the proofs

The proof of these results are based on the fast approximation by conjugation method introduced by Anosov and Katok in [14], where they constructed mixing diffeomorphisms of the unit disc arbitrarily close to Liouvillean rotations. We refer the reader to §3 for a more detailed description of the method, but in a nutshell it consists of the following idea.

The transient dynamics of a Liouvillean rotation is, for all practical reasons, periodic: if ρ\rho is Liouville, then there exist sequences pn,qnp_{n},q_{n}\in{\mathbb{N}}^{*} such that

ρpnqn𝕋<|qn|n\|\rho-\frac{p_{n}}{q_{n}}\|_{{\mathbb{T}}}<|q_{n}|^{-n}

which means that a diffeomorphism conjugate to the rotation by ρ\rho is practically periodic with period qnq_{n}, in scales of iteration comparable with arbitrarily big powers of qnq_{n}. It is tempting, therefore, to try to study diffeomorphisms with rotation number equal to ρ\rho by approximating them from the outside, i.e. with diffeomorphisms that have rational rotation numbers, and to boot, are conjugate to them.

Since the dynamics of a periodic rotation are determined by a finite number of iterations, they are easier to tamper with, and since |qn+1||qn||q_{n+1}|\gg|q_{n}|, it is also possible to modify the dynamics at the scale qn+1q_{n+1} in a way that preserves what was constructed in the previous scale qnq_{n}.

The limit object, satisfying the mixing property was thus constructed as follows. A diffeomorphism fnf_{n} is constructed such that

fn=HnTpn/qnHn1f_{n}=H_{n}\circ T_{p_{n}/q_{n}}\circ H^{-1}_{n}

for some smooth conjugation HnH_{n}, satisfying some finitary version of mixing at a scale of iteration qnq_{n}. A qnq_{n}-periodic conjugation hnh_{n} is constructed, such that the diffeomorphism

fn+1=HnhnTpn+1/qn+1hn1Hn1f_{n+1}=H_{n}\circ h_{n}\circ T_{p_{n+1}/q_{n+1}}\circ h_{n}^{-1}\circ H_{n}^{-1}

satisfying some improved finitary version of mixing at a scale of iteration qn+1qnq_{n+1}\gg q_{n}. This can be achieved by chosing pn+1/qn+1p_{n+1}/q_{n+1} very close to pn/qnp_{n}/q_{n}, so that the C0C^{0} norm of hnh_{n} can be allowed to explode. This will ensure that the limit diffeomorphism f=limfnf=\lim f_{n} is smooth, but not measurably conjugate to TρT_{\rho}. The Liouvillean character of ρ\rho is necessary in order to assure the convergence of fnf_{n} despite the divergence of HnH_{n}.

For more information on the method, results and references we point the reader to [28]. This approximation-by-conjugation method has been useful in producing examples of dynamics incompatible with quasi-periodicity, in the vicinity of quasi-periodic dynamics. It is in some sense the counterpart of the KAM method: KAM tends to prove rigidity in the Diophantine world, while Anosov-Katok is used in order to prove non-rigidity in the Liouvillean world. In the context of cocycles, the concept of reducibility, obtained notably via KAM, allows for studying the rigidity results [12] when the fibered rotation number is Diophantine with respect to the frequency, while Anosov-Katok’s construction will be an efficient method to study wild dynamics when the fibered rotation number is Liouville with respect to the frequency.

The fibered Anosov-Katok construction was introduced by the first author in [41]. In this context, the rotation in the basis is fixed, and the only freedom is in the choice of the mapping in the fibers. The rotation α\alpha in the basis can be chosen to be Diophantine, and the role of periodic rotations in the classical constructions is taken by resonant cocycles, i.e. cocycles whose rotation number is kαk\alpha, where kk\in{\mathbb{Z}}. The rest of the construction remains the same.

From an almost reducibility point of view, the construction consists in engineering the parameters of the KAM normal form, introduced in [42], and further exploited in [41, 43] in order to study a variety of dynamical phenomena present in the almost reducibility regime for cocycles in SU(2)SU(2). Using the KAM schemes of [22, 47], we further develop these techniques in order to adapt them to SL(2,)SL(2,{\mathbb{R}}) cocycles and in the context of the analytic category (instead of the smooth one, as in the [42]).

We point out that results obtainable by the fibered Anosov-Katok method can be obtained for cocycles over a fixed Liouvillean rotation, but the case of a Diophantine rotation is more difficult, and therefore more interesting.

2. A lemma from linear algebra

Given ASL(2,)A\in\mathrm{SL}(2,{\mathbb{R}}) and calling

M:=11+i(1i1i),M:=\frac{1}{1+i}\begin{pmatrix}1&-i\\ 1&i\end{pmatrix},

we have by direct calculation MAM1SU(1,1)MAM^{-1}\in\mathrm{SU}(1,1), where SU(1,1)\mathrm{SU}(1,1) is the group of special unitary 2×22\times 2 matrices preserving the scalar product of 2{\mathbb{C}}^{2} with signature (1,1)(1,-1) i.e.

BH(1001)B=(1001),B^{H}\begin{pmatrix}1&0\\ 0&-1\end{pmatrix}B=\begin{pmatrix}1&0\\ 0&-1\end{pmatrix},

where BSU(1,1)B\in\mathrm{SU}(1,1). Since MM is an isometry between the upper half-plane and the disc models for the hyperbolic plane, we know that

SU(1,1)={(abb¯a¯)||a|2|b|2=1anda,b}.\mathrm{SU}(1,1)=\{\begin{pmatrix}a&b\\ \bar{b}&\bar{a}\end{pmatrix}||a|^{2}-|b|^{2}=1\ \textit{and}\ a,b\in{\mathbb{C}}\}.

The Lie algebra of SU(1,1)\mathrm{SU}(1,1), denoted by su(1,1)\mathrm{su}(1,1) is formed by traceless Hermitian 2×22\times 2 matrices,

su(1,1)={(itzz¯it)|t,z}.\mathrm{su}(1,1)=\{\begin{pmatrix}it&z\\ \bar{z}&-it\end{pmatrix}|t\in{\mathbb{R}},z\in{\mathbb{C}}\}.

Given two matrices in su(1,1)\mathrm{su}(1,1)

A1=(it1z1z¯1it1) and A2=(it2z2z¯2it2)A_{1}=\begin{pmatrix}it_{1}&z_{1}\\ \bar{z}_{1}&-it_{1}\end{pmatrix}\text{ and }A_{2}=\begin{pmatrix}it_{2}&z_{2}\\ \bar{z}_{2}&-it_{2}\end{pmatrix}

their scalar product is defined by

{A1},{A2}=t1t2+(z1z¯2)=t1t2+z1.z2+z1.z2\langle\{A_{1}\},\{A_{2}\}\rangle=t_{1}t_{2}+\mathcal{R}(z_{1}\bar{z}_{2})=t_{1}t_{2}+\mathcal{R}z_{1}.\mathcal{R}z_{2}+\mathcal{I}z_{1}.\mathcal{I}z_{2}

so that the natural semi-Riemannian structure on su(1,1)\mathrm{su}(1,1) be defined by

A|t|2|z|2=det(A)A\rightarrow|t|^{2}-|z|^{2}=\det(A)

In su(1,1)\mathrm{su}(1,1) we can therefore distinguish three regimes: the elliptic regime where det(A)>0\det(A)>0, the parabolic regime where det(A)=0\det(A)=0, and the hyperbolic regime where det(A)<0\det(A)<0.

Parabolic matrices are not diagonalizable, and hyperbolic ones are anti-diagonalizable, but since in the present paper we focus on the elliptic regime, we prove the following lemma concerning the diagonalization of elliptic matrices in su(1,1)\mathrm{su}(1,1). The diagonalizing conjugugation given by the lemma is of optimal norm.

Lemma 2.1.

Let the matrix

A=(itzz¯it)su(1,1)A=\begin{pmatrix}it&z\\ \bar{z}&-it\end{pmatrix}\in\mathrm{su}(1,1)

satisfy detA>0\det A>0. Then, calling ρ=detA\rho=\sqrt{\det A}, we have

D1AD=(iρ00iρ),D^{-1}AD=\begin{pmatrix}i\rho&0\\ 0&-i\rho\end{pmatrix},

where

D=(cos2θ)12(eiϕ00eiϕ)(cosθsinθsinθcosθ)(eiϕ00eiϕ)=(cos2θ)12(cosθe2iϕsinθe2iϕsinθcosθ).\begin{array}[]{r@{}l}D&=(\cos 2\theta)^{-\frac{1}{2}}\begin{pmatrix}e^{i\phi}&0\\ 0&e^{-i\phi}\end{pmatrix}\begin{pmatrix}\cos\theta&\sin\theta\\ \sin\theta&\cos\theta\end{pmatrix}\begin{pmatrix}e^{-i\phi}&0\\ 0&e^{i\phi}\end{pmatrix}\\ &=(\cos 2\theta)^{-\frac{1}{2}}\begin{pmatrix}\cos\theta&e^{2i\phi}\sin\theta\\ e^{-2i\phi}\sin\theta&\cos\theta\end{pmatrix}.\end{array}

Here 2ϕ=argzπ22\phi=\arg z-\frac{\pi}{2} and θ(π2,π2)\theta\in(-\frac{\pi}{2},\frac{\pi}{2}) satisfies

(2.1) 2θ=arctan|z|t2|z|22\theta=-\arctan\frac{|z|}{\sqrt{t^{2}-|z|^{2}}}

In addition we have

(2.2) D2=(1tanθ)21tan2θ=|t|+|z|ρ\begin{array}[]{r@{}l}\|D\|^{2}=\frac{(1-\tan\theta)^{2}}{1-\tan^{2}\theta}=\frac{|t|+|z|}{\rho}\end{array}
Proof.

Firstly, the invariance of the determinant forces that

ρ=t2|z|2(0,)\rho=\sqrt{t^{2}-|z|^{2}}\in(0,\infty)

where the choice of the sign of ρ\rho is of course arbitrary and irrelevant.

Let, now, z=a+biz=a+bi, so that AA can be seen as an element of 3\mathbb{R}^{3} parameterized by (t,b,a)(t,b,a). Then detA>0\det A>0 means AA belongs to the cone {(t,b,a)|t2>b2+a2}\{(t,b,a)|t^{2}>b^{2}+a^{2}\}. As was shown in Figure 1, we can rotate AA to B=(t,|z|,0)B=(t,|z|,0) which lies in the (t,b)(t,b) plane by the conjugation:

(eiϕ00eiϕ).\begin{pmatrix}e^{i\phi}&0\\ 0&e^{-i\phi}\end{pmatrix}.

which is a rotation around the tt axis. We can thus restrict the problem to the diagonalization of matrices of the type

(iti|z|i|z|it)\begin{pmatrix}it&i|z|\\ -i|z|&-it\end{pmatrix}

In the (t,b)(t,b) plane, conjugacy classes are hyperbola t2b2=Ct^{2}-b^{2}=C where C>0C>0 is the square of the angle of the corresponding rotation. In the full R3R^{3} space, the conjugacy classes are the hyperboloids obtained by revolving these hyperbola around the tt axis.

This is achieved by conjugation by

R(θ)=(cos2θ)12(cosθsinθsinθcosθ)R(\theta)=(\cos 2\theta)^{-\frac{1}{2}}\begin{pmatrix}\cos\theta&\sin\theta\\ \sin\theta&\cos\theta\end{pmatrix}

which is a hyperbolic rotation in (t,b)(t,b) plane. We now calculate the remaning free parameter, the hyperbolic angle θ\theta.

Direct calculation shows that

R(θ)(iρ00iρ)R1(θ)=(iρcos12θiρtan2θiρtan2θiρ1cos2θ)R(\theta)\begin{pmatrix}i\rho&0\\ 0&-i\rho\end{pmatrix}R^{-1}(\theta)=\begin{pmatrix}i\rho\cos^{-1}2\theta&-i\rho\tan 2\theta\\ i\rho\tan 2\theta&-i\rho^{-1}\cos 2\theta\end{pmatrix}

and imposing that i|z|=iρtan2θi|z|=-i\rho\tan 2\theta proves the lemma. Optimality of the conjugation follows since the path

R(τθ)(iρ00iρ)R1(τθ),τ[0,1]R(\tau\theta)\begin{pmatrix}i\rho&0\\ 0&-i\rho\end{pmatrix}R^{-1}(\tau\theta),\tau\in[0,1]

is the shortest path in su(1,1)su(1,1) connecting

(iρ00iρ) with (iti|z|i|z|it)\begin{pmatrix}i\rho&0\\ 0&-i\rho\end{pmatrix}\text{ with }\begin{pmatrix}it&i|z|\\ -i|z|&-it\end{pmatrix}

For any BSU(1,1)B\in SU(1,1), its operator norm satisfies

B2=λmax(BTB),\|B\|^{2}=\lambda_{max}(B^{T}B),

where λmax(BTB)\lambda_{max}(B^{T}B) is the maximum eigenvalue of BTBB^{T}B. Therefore, by the structure of DD, (2.2) follows by simple calculations. ∎

Refer to caption
Figure 1. Structure of su(1,1)\mathrm{su}(1,1)

The following corollary is immediate.

Corollary 2.1.

Suppose that

A=i(tλλt)su(1,1)A=i\begin{pmatrix}t&\lambda\\ -\lambda&-t\end{pmatrix}\in\mathrm{su}(1,1)

with t>λ0t>\lambda\geq 0. Then the conjugation DD constructed in Lemma 2.2 has the following estimations:

  1. (1)

    |(cos2θ)12cosθ1|<(λt)21(λt)2|(\cos 2\theta)^{-\frac{1}{2}}\cos\theta-1|<\frac{(\frac{\lambda}{t})^{2}}{\sqrt{1-(\frac{\lambda}{t})^{2}}},

  2. (2)

    |(cos2θ)12sinθ|[λt2(1(λt)2)14,λt(1(λt)2)14]|(\cos 2\theta)^{-\frac{1}{2}}\sin\theta|\in[\frac{\frac{\lambda}{t}}{2(1-(\frac{\lambda}{t})^{2})^{\frac{1}{4}}},\frac{\frac{\lambda}{t}}{(1-(\frac{\lambda}{t})^{2})^{\frac{1}{4}}}].

Proof.

Direct calculation shows that

(cos2θ)12cosθ1=\displaystyle(\cos 2\theta)^{-\frac{1}{2}}\cos\theta-1= 11tan2θ1\displaystyle\sqrt{\frac{1}{1-\tan^{2}\theta}}-1
=\displaystyle= tan2θ1tan2θ11+1tan2θ,\displaystyle\frac{\tan^{2}\theta}{\sqrt{1-\tan^{2}\theta}}\cdot\frac{1}{1+\sqrt{1-\tan^{2}\theta}},

then estimate (1)(1) follows from

(2.3) λ2t<|tanθ|<λt<1.\frac{\lambda}{2t}<|\tan\theta|<\frac{\lambda}{t}<1.

Also by the fact that

(cos2θ)12sinθ=tanθ2tan2θ,\displaystyle(\cos 2\theta)^{-\frac{1}{2}}\sin\theta=\sqrt{\frac{\tan\theta}{2}\tan 2\theta},

then (2)(2) follows from (2.3) and (2.1). ∎

3. The Fiberd Anosov-Katok construction

In this section, we give the Anosov-Katok constructions for quasi-periodic SU(1,1)SU(1,1) cocycles, the construction is initialized by fixing a minimal rotation αd/d\alpha\in{\mathbb{R}}^{d}/{\mathbb{Z}}^{d} and then inductively constructing the sequences {kn}n=0\{k_{n}\}_{n=0}^{\infty}, {tn}n=0\{t_{n}\}_{n=0}^{\infty} and {λn}n=0\{\lambda_{n}\}_{n=0}^{\infty} satisfying n\forall n\in{\mathbb{N}}:

  1. (1)

    kndk_{n}\in{\mathbb{Z}}^{d}, k0=0k_{0}=0, and kn,α𝕋0\|\langle k_{n},\alpha\rangle\|_{{\mathbb{T}}}\rightarrow 0, which forces |kn||k_{n}|\rightarrow\infty,

  2. (2)

    tn>λn0t_{n}>\lambda_{n}\geq 0,

  3. (3)

    tn2λn2=kn+1,α𝕋0\sqrt{t_{n}^{2}-\lambda_{n}^{2}}=\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}\rightarrow 0.

Given these parameters, assume that at the nn-th step of a construction, we have a constant cocycle (α,A~n)(\alpha,\tilde{A}_{n}), where

A~n=(ekn(α)00ekn(α))\tilde{A}_{n}=\begin{pmatrix}e_{k_{n}}(\alpha)&0\\ 0&e_{-k_{n}}(\alpha)\end{pmatrix}

Such a cocycle is said to be knk_{n}-resonant with respect to α{\alpha}.

In the (n+1)(n+1)-th step of the construction, we perturb this cocycle to (α,A~neFn())(\alpha,\tilde{A}_{n}e^{F_{n}(\cdot)}), where

Fn()=2πi(tnλne2kn()λne2kn()tn),F_{n}(\cdot)=2\pi i\begin{pmatrix}t_{n}&\lambda_{n}e_{2k_{n}}(\cdot)\\ -\lambda_{n}e_{-2k_{n}}(\cdot)&-t_{n}\end{pmatrix},

and ek():=e2πik,e_{k}(\cdot):=e^{2\pi i\langle k,\cdot\rangle}. This perturbation is spectrally supported in the resonant Fourier mode in the anti-diagonal direction, while the constant A~n\tilde{A}_{n} is diagonal. The goal of the construction is to exploit the non commutativity arising by this special type of perturbation.

In the strip |x|<h|\Im x|<h, we have the following estimate for the perturbation

(3.1) Fnh2πtne4π|kn|h.\|F_{n}\|_{h}\leq 2\pi t_{n}e^{4\pi|k_{n}|h}.

Let Hn()=(ekn()00ekn()),H_{n}(\cdot)=\begin{pmatrix}e_{k_{n}}(\cdot)&0\\ 0&e_{-k_{n}}(\cdot)\end{pmatrix}, then one has

Hn1(+α)A~neFn()Hn()=exp(2πi(tnλnλntn))=A¯nH_{n}^{-1}(\cdot+\alpha)\tilde{A}_{n}e^{F_{n}(\cdot)}H_{n}(\cdot)=\exp\left(2\pi i\begin{pmatrix}t_{n}&\lambda_{n}\\ -\lambda_{n}&-t_{n}\end{pmatrix}\right)=\bar{A}_{n}\\

By assumptions (2) and (3) and direct application of Lemma 2.1, there exists DnSU(1,1)D_{n}\in SU(1,1) such that

(3.2) Dn1exp(2πi(tnλnλntn))Dn=(ekn+1(α)00ekn+1(α))=A~n+1D_{n}^{-1}\exp\left(2\pi i\begin{pmatrix}t_{n}&\lambda_{n}\\ -\lambda_{n}&-t_{n}\end{pmatrix}\right)D_{n}=\begin{pmatrix}e_{k_{n+1}}(\alpha)&0\\ 0&e_{-k_{n+1}}(\alpha)\end{pmatrix}=\tilde{A}_{n+1}

i.e. A~n+1\tilde{A}_{n+1} is kn+1k_{n+1}-resonant. Note that, by Lemma 2.1, DnD_{n} can be chosen in the form

Dn=(cos2θn)12(cosθnsinθnsinθncosθn),D_{n}=(\cos 2\theta_{n})^{-\frac{1}{2}}\begin{pmatrix}\cos\theta_{n}&\sin\theta_{n}\\ \sin\theta_{n}&\cos\theta_{n}\end{pmatrix},

with estimate

(3.3) Dn2=tn+λntn2λn2[tnkn+1,α𝕋,2tnkn+1,α𝕋],\|D_{n}\|^{2}=\frac{t_{n}+\lambda_{n}}{\sqrt{t_{n}^{2}-\lambda_{n}^{2}}}\in[\frac{t_{n}}{\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}},\frac{2t_{n}}{\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}}],

the last inequality holds since by our assumption (2)(2) and (3)(3).

Let Gn()=Hn()DnG_{n}(\cdot)=H_{n}(\cdot)D_{n}. Then we have

(3.4) Gn1(+α)A~neFn()Gn()=A~n+1,G_{n}^{-1}(\cdot+\alpha)\tilde{A}_{n}e^{F_{n}(\cdot)}G_{n}(\cdot)=\tilde{A}_{n+1},

which means the cocycle (α,A~neFn())(\alpha,\tilde{A}_{n}e^{F_{n}(\cdot)}) is conjugated to the resonant cocycle (α,A~n+1)(\alpha,\tilde{A}_{n+1}), and the construction can be iterated.

Consequently, let B0=IdB_{0}=\mathrm{Id}, Bn()=G0()Gn2()Gn1()B_{n}(\cdot)=G_{0}(\cdot)\cdots G_{n-2}(\cdot)G_{n-1}(\cdot), then, starting with an arbitrary resonant cocycle, we can construct the desired cocycle sequences:

(3.5) An()=Bn(+α)A~nBn1().A_{n}(\cdot)=B_{n}(\cdot+\alpha)\tilde{A}_{n}B_{n}^{-1}(\cdot).

Before introducing the application of this kind of Anosov-Katok construction, we first prove that, under some mild conditions on the the sequence of parameters, the cocycle (α,An())(\alpha,A_{n}(\cdot)) converges, and the limit cocycle is almost reducible. The first assumption of the lemma is related to the fact that we work in the real analytic category, and therefore we need to impose an exponentially fast growth condition on the resonances. In the lemma we use the notation established in this paragraph.

Lemma 3.1.

Suppose that for some ϵ>0\epsilon>0

(3.6) 4πt0+n=14πtne4πhΣj=1n|kj|j=0n12tjkj+1,α𝕋<ϵ.\displaystyle 4\pi t_{0}+\sum_{n=1}^{\infty}4\pi t_{n}e^{4\pi h\Sigma_{j=1}^{n}|k_{j}|}\prod_{j=0}^{n-1}\frac{2t_{j}}{\|\langle k_{j+1},\alpha\rangle\|_{{\mathbb{T}}}}<\epsilon.

Then the cocycle (α,An())(\alpha,A_{n}(\cdot)) converges to (α,A())(\alpha,A_{\infty}(\cdot)) with A()Chω(𝕋d,SU(1,1))A_{\infty}(\cdot)\in C_{h}^{\omega}({\mathbb{T}}^{d},SU(1,1)) and

AIdh<ϵ.\|A_{\infty}-\mathrm{Id}\|_{h}<\epsilon.

Moreover, the cocycle (α,A)(\alpha,A_{\infty}) is almost reducible:

Bn1(+α)A()Bn()=A~n+F~n(),B_{n}^{-1}(\cdot+\alpha)A_{\infty}(\cdot)B_{n}(\cdot)=\tilde{A}_{n}+\tilde{F}_{n}(\cdot),

where

(3.7) F~nh4πtne4π|kn|h+j=n+14πtje4πhΣi=nj|kj|i=nj12tiki+1,α𝕋,\|\tilde{F}_{n}\|_{h}\leq 4\pi t_{n}e^{4\pi|k_{n}|h}+\sum_{j=n+1}^{\infty}4\pi t_{j}e^{4\pi h\Sigma_{i=n}^{j}|k_{j}|}\prod_{i=n}^{j-1}\frac{2t_{i}}{\|\langle k_{i+1},\alpha\rangle\|_{{\mathbb{T}}}},
(3.8) AAnhj=n4πtje4πhΣi=1j|ki|i=0j12tiki+1,α𝕋.\|A_{\infty}-A_{n}\|_{h}\leq\sum_{j=n}^{\infty}4\pi t_{j}e^{4\pi h\Sigma_{i=1}^{j}|k_{i}|}\prod_{i=0}^{j-1}\frac{2t_{i}}{\|\langle k_{i+1},\alpha\rangle\|_{{\mathbb{T}}}}.
Proof.

Notice that by our construction (3.4) and (3.5), we have

(3.9) An+1An=Bn(+α)(A~neFnA~n)Bn1(),A_{n+1}-A_{n}=B_{n}(\cdot+\alpha)(\tilde{A}_{n}e^{F_{n}}-\tilde{A}_{n})B_{n}^{-1}(\cdot),

then by (3.1) and (3.3), one has

A1A0h2B0h2F0h4πt0\displaystyle\|A_{1}-A_{0}\|_{h}\leq 2\|B_{0}\|_{h}^{2}\|F_{0}\|_{h}\leq 4\pi t_{0}

if n1n\geq 1, one has

(3.10) AnAn+1h2Bnh2Fnh4πtne4πhΣj=1n|kj|j=0n12tjkj+1,α𝕋.\displaystyle\|A_{n}-A_{n+1}\|_{h}\leq 2\|B_{n}\|_{h}^{2}\|F_{n}\|_{h}\leq 4\pi t_{n}e^{4\pi h\Sigma_{j=1}^{n}|k_{j}|}\prod_{j=0}^{n-1}\frac{2t_{j}}{\|\langle k_{j+1},\alpha\rangle\|_{{\mathbb{T}}}}.

therefore, by our assumption (3.6), the cocycle (α,An())(\alpha,A_{n}(\cdot)) converges to some (α,A())(\alpha,A_{\infty}(\cdot)). Also note our construction (3.5) implies that A0=IdA_{0}=\mathrm{Id}, then

AIdh<ϵ.\|A_{\infty}-\mathrm{Id}\|_{h}<\epsilon.

Furthermore, by (3.9), we have

Bn1(+α)A()Bn()\displaystyle B_{n}^{-1}(\cdot+\alpha)A_{\infty}(\cdot)B_{n}(\cdot)
=\displaystyle= Bn1(+α)An()Bn()+j=nBn1(+α)(Aj+1()Aj())Bn()\displaystyle B_{n}^{-1}(\cdot+\alpha)A_{n}(\cdot)B_{n}(\cdot)+\sum_{j=n}^{\infty}B_{n}^{-1}(\cdot+\alpha)(A_{j+1}(\cdot)-A_{j}(\cdot))B_{n}(\cdot)
=\displaystyle= A~n+j=n(Bn1Bj)(+α)(A~jeFjA~j)()(Bn1Bj)1()\displaystyle\tilde{A}_{n}+\sum_{j=n}^{\infty}(B_{n}^{-1}B_{j})(\cdot+\alpha)(\tilde{A}_{j}e^{F_{j}}-\tilde{A}_{j})(\cdot)(B_{n}^{-1}B_{j})^{-1}(\cdot)
=\displaystyle= A~n+F~n(),\displaystyle\tilde{A}_{n}+\tilde{F}_{n}(\cdot),

then (3.7) follows immediately. By (3.6) we know F~nh0\|\tilde{F}_{n}\|_{h}\rightarrow 0, thus the cocycle (α,A())(\alpha,A_{\infty}(\cdot)) is almost reducible. ∎

Let us now provide some motivation for the precise choice of the structure of the perturbations FnF_{n}. Note firstly that FnF_{n} is totally determined by the triple {kn,tn,λn}\{k_{n},t_{n},\lambda_{n}\}. After conjugation by HnH_{n}, the cocycle (α,A~neFn)(\alpha,\tilde{A}_{n}e^{F_{n}}) becomes

exp2πi(tnλnλntn)=A¯n.\exp 2\pi i\begin{pmatrix}t_{n}&\lambda_{n}\\ -\lambda_{n}&-t_{n}\end{pmatrix}=\bar{A}_{n}.

Ths matrix A¯n\bar{A}_{n} is parameterized by {tn,λn}\{t_{n},\lambda_{n}\}, which shows that the construction of the matrix A¯n\bar{A}_{n} is determined by the choice of the perturbation FnF_{n}. This calculation provides the only way to perturb an elliptic constant cocycle so that it becomes conjugate to a parabolic or a hyperbolic one.

Refer to caption
Figure 2. algebra

If we consider the map from 2{\mathbb{R}}^{2} to GG:

(t,λ)exp2πi(tλλt),(t,\lambda)\rightarrow\exp 2\pi i\begin{pmatrix}t&\lambda\\ -\lambda&-t\end{pmatrix},

we see that {(t,λ)||λ|<|t|}\{(t,\lambda)||\lambda|<|t|\}, {(t,λ)||λ|=|t|}\{(t,\lambda)||\lambda|=|t|\} and {(t,λ)||λ|>|t|}\{(t,\lambda)||\lambda|>|t|\} correspond to the elliptic, parabolic and hyperbolic matrices in SU(1,1)\mathrm{SU}(1,1), respectively. The phenomena we study can appear only in the elliptic, and not in the parabolic and hyperbolic regimes. This is the reason why the condition tn>λn0t_{n}>\lambda_{n}\geq 0 is imposed. A second restriction is related with the convergence of AnA_{n} and almost reducibility of the limit cocycle, which is guaranteed by choosing tnt_{n} sufficiently small with respect to {kj}j=0n\{k_{j}\}_{j=0}^{n}. In other words, we impose that (tn,λn)(t_{n},\lambda_{n}) tend to OO in the elliptic region and sufficiently fast. Then, depending on the unexpected behavior that we want to produce, different restrictions concerning the relative size of λn\lambda_{n} with respect to tnt_{n} are imposed.

In §4, we construct cocycles at which the Lyapunov exponend is exactly κ\kappa-Hölder continuous, with 12<κ<1\frac{1}{2}<\kappa<1. The eigenvalues of a constant parabolic matrix are 12\frac{1}{2}-Hölder continuous, while in the hyperbolic regime (regime II of fig. 2) they depend smoothly on the matrix. Therefore, we first chose tnλnt_{n}\approx\lambda_{n} to be elliptic but close to parabolic, in the regime IIII. We then perturb A¯n\bar{A}_{n} to A¯n\bar{A}^{\prime}_{n} which is now hyperbolc, but close to parabolic. By controlling the distance between A¯n\bar{A}_{n} and A¯n\bar{A}^{\prime}_{n}, we are able to obtain κ\kappa-Hölder continunity.

In §5, we construct cocycles with sublinear growth. Since Hn0=1\|H_{n}\|_{0}=1, the C0C^{0}-norm of the conjugations BnB_{n} are determined by Dn\|D_{n}\|. Constant elliptic cocycles do not grow, while constant parabolic ones grow linearly. In order to obtain growth of elliptic cocycles, we construct cocycles that are conjugated arbitrarily close to parabolic ones, in regime IIII of the figure 2, and in that case the growth of the cocycle is comparable to Dn\|D_{n}\|\rightarrow\infty, as nn\rightarrow\infty.

In §6, we construct cocycles that are CsC^{s} reducible, but not Cs+1C^{s+1} reducible. This is obtained by restricting the cocycle to the regime IIIIII of the figure, where Dn1\|D_{n}\|\rightarrow 1, as nn\rightarrow\infty, at a prescribed speed.

4. Optimal Hölder continuity of Lyapunov exponent

4.1. Local density results

We first prove that cocycles whose associated Lyapunov exponents are exactly κ\kappa-Hölder continuous, with 12<κ<1\frac{1}{2}<\kappa<1, are locally dense.

Proposition 4.1.

Fix h>0h>0, 12<κ<1\frac{1}{2}<\kappa<1 and αd/d\alpha\in{\mathbb{R}}^{d}/{\mathbb{Z}}^{d} rationally independent. Then, for any ϵ>0\epsilon>0 there exists a cocycle (α,A())(\alpha,A(\cdot)) with A()Chω(𝕋d,SL(2,))A(\cdot)\in C_{h}^{\omega}({\mathbb{T}}^{d},\mathrm{SL}(2,{\mathbb{R}})) and

A()Idh<ϵ,\|A(\cdot)-\mathrm{Id}\|_{h}<\epsilon,

such that Lyapunov exponent is exactly κ\kappa-Hölder continuous at (α,A())(\alpha,A(\cdot)).

Proof.

Using the isomoprhism between SL(2,))\mathrm{SL}(2,{\mathbb{R}})) and SU(1,1)SU(1,1), we prove the theorem in the context of SU(1,1)SU(1,1) cocycles. Let us first introduce two auxiliary parameters q,δ~q,\tilde{\delta}. Since 12<κ<1\frac{1}{2}<\kappa<1, there exists qq\in{\mathbb{N}} such that

(4.1) q>101κ2κ1+10.\displaystyle q>10\frac{1-\kappa}{2\kappa-1}+10.

and 0<δ~<1q+10<\tilde{\delta}<\frac{1}{q+1} satisfying:

(4.2) (1δ~)1κ2κ1+11κ2κ1+1>qq+1.\displaystyle\frac{(1-\tilde{\delta})\frac{1-\kappa}{2\kappa-1}+1}{\frac{1-\kappa}{2\kappa-1}+1}>\frac{q}{q+1}.

We will also use the auxiliary function

f(x):=x|lnx|12f(x):=x^{|\ln x|^{-\frac{1}{2}}}

which satisfies limx0f(x)=0\lim_{x\rightarrow 0}f(x)=0, and limx1f(x)=1\lim_{x\rightarrow 1}f(x)=1 and is monotonic increasing on (0,1)(0,1).

We can now construct iteratively the sequence {kn}0\{k_{n}\}_{0}^{\infty}. Let k0=0k_{0}=0, and choose k1k_{1} satisfying:

(4.3) f(k1,α𝕋)18<ϵ32,\displaystyle f(\|\langle k_{1},\alpha\rangle\|_{{\mathbb{T}}})^{\frac{1}{8}}<\frac{\epsilon}{32},
(4.4) 8|lnk1,α𝕋|12<δ~.\displaystyle 8|\ln\|\langle k_{1},\alpha\rangle\|_{{\mathbb{T}}}|^{-\frac{1}{2}}<\tilde{\delta}.

Assuming we have constructed kj,jnk_{j},j\leq n, we choose kn+1dk_{n+1}\in{\mathbb{Z}}^{d} satisfying the following properties:

(4.5) f(kn+1,α𝕋)18\displaystyle f(\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}})^{\frac{1}{8}} <\displaystyle< ϵ4n+1e16π|kn|hkn,α𝕋q+1,\displaystyle\frac{\epsilon}{4^{n+1}}e^{-16\pi|k_{n}|h}\|\langle k_{n},\alpha\rangle\|_{{\mathbb{T}}}^{q+1},
(4.6) |kn+1|\displaystyle|k_{n+1}| >\displaystyle> e|kn|+10.\displaystyle e^{|k_{n}|}+10.

We now call

δn=18|lnkn+1,α𝕋|12,\delta_{n}=\frac{1}{8}|\ln\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}|^{-\frac{1}{2}},
(4.7) tn(14δn)1κ2κ1+1=kn+1,α𝕋,\displaystyle t_{n}^{(1-4\delta_{n})\frac{1-\kappa}{2\kappa-1}+1}=\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}},

and, finally, let

λn=tn2|kn+1,α𝕋2.\lambda_{n}=\sqrt{t_{n}^{2}-|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}^{2}}.

We also remark that by (4.5)

kn+1,α𝕋δn<kn,α𝕋q+1,\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}^{\delta_{n}}<\|\langle k_{n},\alpha\rangle\|_{{\mathbb{T}}}^{q+1},

which, combined with (4.7), gives

tnδn((14δn)1κ2κ1+1)<tn1(q+1)((14δn1)1κ2κ1+1).\displaystyle t_{n}^{\delta_{n}((1-4\delta_{n})\frac{1-\kappa}{2\kappa-1}+1)}<t_{n-1}^{(q+1)((1-4\delta_{n-1})\frac{1-\kappa}{2\kappa-1}+1)}.

Thus by (4.1) and (4.2), we have

(4.8) tn<tnδn<tn1(q+1)[(14δn1)1κ2κ1+1(14δn)1κ2κ1+1]<tn1q<tn1101κ2κ1+10.\displaystyle t_{n}<t_{n}^{\delta_{n}}<t_{n-1}^{(q+1)[\frac{(1-4\delta_{n-1})\frac{1-\kappa}{2\kappa-1}+1}{(1-4\delta_{n})\frac{1-\kappa}{2\kappa-1}+1}]}<t_{n-1}^{q}<t_{n-1}^{10\frac{1-\kappa}{2\kappa-1}+10}.

With these parameters, we can construct (α,A~neFn())(\alpha,\tilde{A}_{n}e^{F_{n}(\cdot)}) and then construct (α,An())(\alpha,A_{n}(\cdot)) by the Anosov-Katok method of §3. First we check the following equality:

(4.9) 4πt0+n=14πtne4πhΣj=1n|kj|j=0n12tjkj+1,α𝕋<ϵ.\displaystyle 4\pi t_{0}+\sum_{n=1}^{\infty}4\pi t_{n}e^{4\pi h\Sigma_{j=1}^{n}|k_{j}|}\prod_{j=0}^{n-1}\frac{2t_{j}}{\|\langle k_{j+1},\alpha\rangle\|_{{\mathbb{T}}}}<\epsilon.

Note that by our selection of parameters and estimate (4.8), we have

j=0n12tjkj+1,α𝕋=j=0n12tj(14δj)1κ2κ1\displaystyle\prod_{j=0}^{n-1}\frac{2t_{j}}{\|\langle k_{j+1},\alpha\rangle\|_{{\mathbb{T}}}}=\prod_{j=0}^{n-1}2t_{j}^{-(1-4\delta_{j})\frac{1-\kappa}{2\kappa-1}}
\displaystyle\leq (j=0n122κ11κtj1)1κ2κ1(22κ11κtn11)1κ2κ1Σj=0n11qj\displaystyle(\prod_{j=0}^{n-1}2^{\frac{2\kappa-1}{1-\kappa}}t_{j}^{-1})^{\frac{1-\kappa}{2\kappa-1}}\leq(2^{\frac{2\kappa-1}{1-\kappa}}t_{n-1}^{-1})^{\frac{1-\kappa}{2\kappa-1}\Sigma_{j=0}^{n-1}\frac{1}{q^{j}}}
(4.10) \displaystyle\leq 2tnδn1κ2κ12tnδnq10,\displaystyle 2t_{n}^{-\delta_{n}\frac{1-\kappa}{2\kappa-1}}\leq 2t_{n}^{-\frac{\delta_{n}q}{10}},

On the other hand, by (4.7) and (4.1), we have

(4.11) tnδnq10tnδn(1κ2κ1+1)kn+1,α𝕋δn<ϵ4n+1e16π|kn|hkn,α𝕋q+1,\displaystyle t_{n}^{\frac{\delta_{n}q}{10}}\leq t_{n}^{\delta_{n}(\frac{1-\kappa}{2\kappa-1}+1)}\leq\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}^{\delta_{n}}<\frac{\epsilon}{4^{n+1}}e^{-16\pi|k_{n}|h}\|\langle k_{n},\alpha\rangle\|_{{\mathbb{T}}}^{q+1},

Moreover, by (4.6), we have Σj=1n|kj|<2|kn|\Sigma_{j=1}^{n}|k_{j}|<2|k_{n}|. Thus we have

4πt0+n=14πtne4πhΣj=1n|kj|j=0n12tjkj+1,α𝕋\displaystyle 4\pi t_{0}+\sum_{n=1}^{\infty}4\pi t_{n}e^{4\pi h\Sigma_{j=1}^{n}|k_{j}|}\prod_{j=0}^{n-1}\frac{2t_{j}}{\|\langle k_{j+1},\alpha\rangle\|_{{\mathbb{T}}}}
\displaystyle\leq 4πk1,α𝕋2κ1κ+n=14πtnϵe8π|kn|h4n+1tnδnq5\displaystyle 4\pi\|\langle k_{1},\alpha\rangle\|_{{\mathbb{T}}}^{\frac{2\kappa-1}{\kappa}}+\sum_{n=1}^{\infty}4\pi t_{n}\frac{\epsilon e^{-8\pi|k_{n}|h}}{4^{n+1}}t_{n}^{-\frac{\delta_{n}q}{5}}
\displaystyle\leq 4πk1,α𝕋δ0+4πt045ϵ8\displaystyle 4\pi\|\langle k_{1},\alpha\rangle\|_{{\mathbb{T}}}^{\delta_{0}}+\frac{4\pi t_{0}^{\frac{4}{5}}\epsilon}{8}
\displaystyle\leq 4πϵ32+ϵ32<ϵ.\displaystyle\frac{4\pi\epsilon}{32}+\frac{\epsilon}{32}<\epsilon.

by(4.3), (4.11) and (4.10). Therefore by Lemma 3.1, the limit cocycle A()Chω(𝕋d,SU(1,1))A(\cdot)\in C^{\omega}_{h}({\mathbb{T}}^{d},SU(1,1)) exists and satisfies

A()Idh=A()A0h<ϵ.\|A(\cdot)-\mathrm{Id}\|_{h}=\|A(\cdot)-A_{0}\|_{h}<\epsilon.

Moreover,

(4.12) Bn1(+α)A()Bn()=A~n+F~n(),B_{n}^{-1}(\cdot+\alpha)A(\cdot)B_{n}(\cdot)=\tilde{A}_{n}+\tilde{F}_{n}(\cdot),

by (4.10) and (4.11), we have estimate

F~nh\displaystyle\|\tilde{F}_{n}\|_{h} 4πtne4π|kn|h+j=n+14πtje4πhΣi=nj|kj|i=nj12tiki+1,α𝕋\displaystyle\leq 4\pi t_{n}e^{4\pi|k_{n}|h}+\sum_{j=n+1}^{\infty}4\pi t_{j}e^{4\pi h\Sigma_{i=n}^{j}|k_{j}|}\prod_{i=n}^{j-1}\frac{2t_{i}}{\|\langle k_{i+1},\alpha\rangle\|_{{\mathbb{T}}}}
j=n4πtj35ϵ4n+1e16πh|kj|kj,α𝕋2q+2\displaystyle\leq\sum_{j=n}^{\infty}4\pi t_{j}^{\frac{3}{5}}\frac{\epsilon}{4^{n+1}}e^{-16\pi h|k_{j}|}\|\langle k_{j},\alpha\rangle\|_{{\mathbb{T}}}^{2q+2}
(4.13) <2ϵ4n+1e8π|kn|hkn,α𝕋2q+2tn12,\displaystyle<\frac{2\epsilon}{4^{n+1}}e^{-8\pi|k_{n}|h}\|\langle k_{n},\alpha\rangle\|_{{\mathbb{T}}}^{2q+2}t_{n}^{\frac{1}{2}},

Similarly, by (3.8) of Lemma 3.1, and the following estimate holds true

A()An()h\displaystyle\|A(\cdot)-A_{n}(\cdot)\|_{h} j=n4πtje4πhΣi=1j|ki|i=0j12tiki+1,α𝕋\displaystyle\leq\sum_{j=n}^{\infty}4\pi t_{j}e^{4\pi h\Sigma_{i=1}^{j}|k_{i}|}\prod_{i=0}^{j-1}\frac{2t_{i}}{\|\langle k_{i+1},\alpha\rangle\|_{{\mathbb{T}}}}
(4.14) <2ϵ4n+1e8π|kn|hkn,α𝕋2q+2tn12.\displaystyle<\frac{2\epsilon}{4^{n+1}}e^{-8\pi|k_{n}|h}\|\langle k_{n},\alpha\rangle\|_{{\mathbb{T}}}^{2q+2}t_{n}^{\frac{1}{2}}.

Turning to the size of the conjugations, by the construction and equations (3.3) and (4.10), we get the estimate

Bn+102j=0nDj2=Dn2j=0n1Dj2\displaystyle\|B_{n+1}\|_{0}^{2}\leq\prod_{j=0}^{n}\|D_{j}\|^{2}=\|D_{n}\|^{2}\prod_{j=0}^{n-1}\|D_{j}\|^{2}
\displaystyle\leq 2tnkn+1,α𝕋j=0n12tjkj+1,α𝕋\displaystyle\frac{2t_{n}}{\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}}\prod_{j=0}^{n-1}\frac{2t_{j}}{\|\langle k_{j+1},\alpha\rangle\|_{{\mathbb{T}}}}
(4.15) \displaystyle\leq 2tn(14δn)1κ2κ12tnδn1κ2κ18tn(12δn)1κ2κ1.\displaystyle 2t_{n}^{-(1-4\delta_{n})\frac{1-\kappa}{2\kappa-1}}2t_{n}^{-\delta_{n}\frac{1-\kappa}{2\kappa-1}}\leq 8t_{n}^{-(1-2\delta_{n})\frac{1-\kappa}{2\kappa-1}}.

We now prove that Lyapunov exponent is κ\kappa-Hölder continuous at (α,A())(\alpha,A(\cdot)).

Lemma 4.1.

Let 12<κ<1\frac{1}{2}<\kappa<1 and (α,A())(\alpha,A(\cdot)) constructed as above. Then, for any A()C(𝕋d,SU(1,1))A^{\prime}(\cdot)\in C({\mathbb{T}}^{d},\mathrm{SU}(1,1)) satisfying AA01\|A^{\prime}-A\|_{0}\leq 1 there exists CC independent of AA^{\prime} such that

|L(α,A)L(α,A)|<CAA0κ.|L(\alpha,A^{\prime})-L(\alpha,A)|<C\|A-A^{\prime}\|_{0}^{\kappa}.
Proof.

For any A()C(𝕋d,SU(1,1))A^{\prime}(\cdot)\in C({\mathbb{T}}^{d},\mathrm{SU}(1,1)), let ε:=AA0\varepsilon:=\|A^{\prime}-A\|_{0}. By (4.8), we have limntnδn=0\lim_{n\rightarrow\infty}t_{n}^{\delta_{n}}=0, and thus there exists N2=N2(κ)N_{2}=N_{2}(\kappa) such that tN2δN2<(132)2κ11κt_{N_{2}}^{\delta_{N_{2}}}<(\frac{1}{32})^{\frac{2\kappa-1}{1-\kappa}}. Consequently, we have

(132)11κtn112δn12κ1>tn11δn12κ1,nN2.(\frac{1}{32})^{\frac{1}{1-\kappa}}t_{n-1}^{\frac{1-2\delta_{n-1}}{2\kappa-1}}>t_{n-1}^{\frac{1-\delta_{n-1}}{2\kappa-1}},\qquad n\geq N_{2}.

This implies that if ε\varepsilon is small enough, there exists nn such that

ε(tn1δn2κ1,(132)11κtn112δn12κ1).\varepsilon\in(t_{n}^{\frac{1-\delta_{n}}{2\kappa-1}},(\frac{1}{32})^{\frac{1}{1-\kappa}}t_{n-1}^{\frac{1-2\delta_{n-1}}{2\kappa-1}}).

Recall that by construction, cf. eq. (3.2), we have

(4.16) DnA~n+1Dn1=exp2πi(tnλnλntn):=A¯n,D_{n}\tilde{A}_{n+1}D_{n}^{-1}=\exp 2\pi i\begin{pmatrix}t_{n}&\lambda_{n}\\ -\lambda_{n}&-t_{n}\end{pmatrix}:=\bar{A}_{n},

which means A¯n\bar{A}_{n} can be diagonalized by large conjugacy DnD_{n}. However, there always exists UnU(2)U_{n}\in\mathrm{U}(2) such that UnU_{n} conjugates A¯n\bar{A}_{n} into Schur Form, i.e.

(4.17) Un1A¯nUn=(ekn+1(α)cn0ekn+1(α)),U_{n}^{-1}\bar{A}_{n}U_{n}=\begin{pmatrix}e_{k_{n+1}}(\alpha)&c_{n}\\ 0&e_{-k_{n+1}}(\alpha)\end{pmatrix},

where |cn|4πtn|c_{n}|\leq 4\pi t_{n}.

Let now

B~n():=Bn+1()Dn1Un.\tilde{B}_{n}(\cdot):=B_{n+1}(\cdot)D_{n}^{-1}U_{n}.

Then by equations (4.12), (4.16) and (4.17), we have

B~n1(+α)A()B~n()\displaystyle\tilde{B}_{n}^{-1}(\cdot+\alpha)A(\cdot)\tilde{B}_{n}(\cdot) =\displaystyle= Un1Dn(A~n+1+F~n+1())Dn1Un\displaystyle U_{n}^{-1}D_{n}(\tilde{A}_{n+1}+\tilde{F}_{n+1}(\cdot))D_{n}^{-1}U_{n}
=\displaystyle= (ekn+1(α)cn0ekn+1(α))+Un1DnF~n+1()Dn1Un.\displaystyle\begin{pmatrix}e_{k_{n+1}}(\alpha)&c_{n}\\ 0&e_{-k_{n+1}}(\alpha)\end{pmatrix}+U_{n}^{-1}D_{n}\tilde{F}_{n+1}(\cdot)D_{n}^{-1}U_{n}.

Let now

Bn():=B~n()(s00s1),B_{n}^{\prime}(\cdot):=\tilde{B}_{n}(\cdot)\begin{pmatrix}s&0\\ 0&s^{-1}\end{pmatrix},

where s1=B~n0ε1κ2s^{-1}=\|\tilde{B}_{n}\|_{0}\varepsilon^{\frac{1-\kappa}{2}}. By construction, B~n\tilde{B}_{n} can be written as

B~n()=Bn()Hn()Un.\tilde{B}_{n}(\cdot)=B_{n}(\cdot)H_{n}(\cdot)U_{n}.

By (4.15) and the parameter choice ε<(132)11κtn112δn12κ1\varepsilon<(\frac{1}{32})^{\frac{1}{1-\kappa}}t_{n-1}^{\frac{1-2\delta_{n-1}}{2\kappa-1}}, we have

B~n0=Bn022tn1(12δn1)1κ2(2κ1)<ε1κ2.\|\tilde{B}_{n}\|_{0}=\|B_{n}\|_{0}\leq 2\sqrt{2}t_{n-1}^{-(1-2\delta_{n-1})\frac{1-\kappa}{2(2\kappa-1)}}<\varepsilon^{-\frac{1-\kappa}{2}}.

which implies that

(4.18) Bn02ε1κ2.\displaystyle\|B_{n}^{\prime}\|_{0}\leq 2\varepsilon^{-\frac{1-\kappa}{2}}.

By (4.1), one has

Bn1(+α)A()Bn()\displaystyle B_{n}^{\prime-1}(\cdot+\alpha)A(\cdot)B_{n}^{\prime}(\cdot)
=\displaystyle= (ekn+1(α)cns20ekn+1(α))+(s100s)Un1DnF~n+1()Dn1Un(s00s1).\displaystyle\begin{pmatrix}e_{k_{n+1}}(\alpha)&c_{n}s^{-2}\\ 0&e_{-k_{n+1}}(\alpha)\end{pmatrix}+\begin{pmatrix}s^{-1}&0\\ 0&s\end{pmatrix}U_{n}^{-1}D_{n}\tilde{F}_{n+1}(\cdot)D_{n}^{-1}U_{n}\begin{pmatrix}s&0\\ 0&s^{-1}\end{pmatrix}.

Since ε>tn1δn2κ1\varepsilon>t_{n}^{\frac{1-\delta_{n}}{2\kappa-1}}, then by (4.8), we have

s2cnB~n02ε1κ4πtn\displaystyle s^{-2}c_{n}\leq\|\tilde{B}_{n}\|_{0}^{2}\varepsilon^{1-\kappa}4\pi t_{n}
\displaystyle\leq 32πεκε12κtntn1(12δn)1κ2κ1\displaystyle 32\pi\varepsilon^{\kappa}\varepsilon^{1-2\kappa}t_{n}t_{n-1}^{-(1-2\delta_{n})\frac{1-\kappa}{2\kappa-1}}
\displaystyle\leq εκ32πtnδn1tntn1q10<εκ\displaystyle\varepsilon^{\kappa}32\pi t_{n}^{\delta_{n}-1}t_{n}t_{n-1}^{-\frac{q}{10}}<\varepsilon^{\kappa}

Consequently, by (4.13), (4.18), (4.1), we have

Bn1(+α)A()Bn()A~n+10\displaystyle\|B_{n}^{\prime-1}(\cdot+\alpha)A^{\prime}(\cdot)B_{n}^{\prime}(\cdot)-\tilde{A}_{n+1}\|_{0}
\displaystyle\leq Bn1(+α)(A()A())Bn()0+Bn1(+α)A()Bn()A~n+10\displaystyle\|B_{n}^{\prime-1}(\cdot+\alpha)(A^{\prime}(\cdot)-A(\cdot))B_{n}^{\prime}(\cdot)\|_{0}+\|B_{n}^{\prime-1}(\cdot+\alpha)A(\cdot)B_{n}^{\prime}(\cdot)-\tilde{A}_{n+1}\|_{0}
\displaystyle\leq Bn02ε+s2|cn|+2s2Dn2F~n+10\displaystyle\|B_{n}^{\prime}\|_{0}^{2}\varepsilon+s^{-2}|c_{n}|+2s^{2}\|D_{n}\|^{2}\|\tilde{F}_{n+1}\|_{0}
\displaystyle\leq 4εκ+εκ+2ε(1κ)2tnkn+1,α𝕋kn+1,α𝕋2q+2tn+112\displaystyle 4\varepsilon^{\kappa}+\varepsilon^{\kappa}+2\varepsilon^{-(1-\kappa)}\frac{2t_{n}}{\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}}\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}^{2q+2}t_{n+1}^{\frac{1}{2}}
\displaystyle\leq 5εκ+εκ4tnδn1tntn+112\displaystyle 5\varepsilon^{\kappa}+\varepsilon^{\kappa}4t_{n}^{\delta_{n}-1}t_{n}t_{n+1}^{\frac{1}{2}}
\displaystyle\leq 8εκ.\displaystyle 8\varepsilon^{\kappa}.

Since Lyapunov exponent is invariant under conjugacy, we immediately have

L(α,A)=L(α,Bn1(+α)A()Bn())16εκ.L(\alpha,A^{\prime})=L(\alpha,B_{n}^{\prime-1}(\cdot+\alpha)A^{\prime}(\cdot)B_{n}^{\prime}(\cdot))\leq 16\varepsilon^{\kappa}.

On the other hand, by (4.12), we have

L(α,A)=L(α,A~n+F~n)2F~n,L(\alpha,A)=L(\alpha,\tilde{A}_{n}+\tilde{F}_{n})\leq 2\|\tilde{F}_{n}\|,

by continuity of Lyapunov exponent [21], we then have L(α,A)=0L(\alpha,A)=0, consequently,

|L(α,A)L(α,A)|=L(α,A)<CAA0κ.|L(\alpha,A^{\prime})-L(\alpha,A)|=L(\alpha,A^{\prime})<C\|A-A^{\prime}\|_{0}^{\kappa}.

We now prove that Lyapunov exponent is exactly κ\kappa-Hölder continuous at (α,A())(\alpha,A(\cdot)), i.e. that the exponent is not higher than κ\kappa.

Lemma 4.2.

There exists a sequence {An}n=1\{A_{n}^{\prime}\}_{n=1}^{\infty} where AnChω(𝕋d,SU(1,1))A_{n}^{\prime}\in C^{\omega}_{h}({\mathbb{T}}^{d},SU(1,1)) such that

|L(α,An)L(α,A)|>cAnAhκn,|L(\alpha,A_{n}^{\prime})-L(\alpha,A)|>c\|A_{n}^{\prime}-A\|_{h}^{\kappa_{n}},

where cc independent of nn and limnκn=κ\lim_{n\rightarrow\infty}\kappa_{n}=\kappa.

Proof.

Define for every nn\in{\mathbb{N}}

A¯n:=exp2πi(λn2tnλnλn2tnλn),\bar{A}^{\prime}_{n}:=\exp 2\pi i\begin{pmatrix}\lambda_{n}&2t_{n}-\lambda_{n}\\ \lambda_{n}-2t_{n}&-\lambda_{n}\end{pmatrix},

and

An+1():=(BnHn)(+α)A¯n(BnHn)1().A^{\prime}_{n+1}(\cdot):=(B_{n}H_{n})(\cdot+\alpha)\bar{A}^{\prime}_{n}(B_{n}H_{n})^{-1}(\cdot).

and notice that

A¯nA¯n4π(tnλn)\displaystyle\|\bar{A}^{\prime}_{n}-\bar{A}_{n}\|\leq 4\pi(t_{n}-\lambda_{n})
=\displaystyle= 4πtn2λn2tn+λn4πkn+1,α𝕋2tn\displaystyle 4\pi\frac{t_{n}^{2}-\lambda_{n}^{2}}{t_{n}+\lambda_{n}}\leq 4\pi\frac{\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}^{2}}{t_{n}}
(4.20) \displaystyle\leq 4πtn2(14δn)1κ2κ1+1.\displaystyle 4\pi t_{n}^{2(1-4\delta_{n})\frac{1-\kappa}{2\kappa-1}+1}.

Therefore, we have

An+1An+10\displaystyle\|A^{\prime}_{n+1}-A_{n+1}\|_{0} Bn02Hn02A¯nA¯n\displaystyle\leq\|B_{n}\|_{0}^{2}\|H_{n}\|_{0}^{2}\|\bar{A}^{\prime}_{n}-\bar{A}_{n}\|
4πtn1(12δn1)1κ2κ1tn2(14δn)1κ2κ1+1\displaystyle\leq 4\pi t_{n-1}^{-(1-2\delta_{n-1})\frac{1-\kappa}{2\kappa-1}}t_{n}^{2(1-4\delta_{n})\frac{1-\kappa}{2\kappa-1}+1}
tn(2(8+κ1κ)δn)1κ2κ1+1.\displaystyle\leq t_{n}^{(2-(8+\frac{\kappa}{1-\kappa})\delta_{n})\frac{1-\kappa}{2\kappa-1}+1}.

By (4.8) and (4.14), we have

An+1Ah\displaystyle\|A^{\prime}_{n+1}-A\|_{h} \displaystyle\leq An+1An+1h+An+1Ah\displaystyle\|A^{\prime}_{n+1}-A_{n+1}\|_{h}+\|A_{n+1}-A\|_{h}
\displaystyle\leq tn(2(8+κ1κ)δn)1κ2κ1+1+kn+1,α𝕋2q+2tn+112\displaystyle t_{n}^{(2-(8+\frac{\kappa}{1-\kappa})\delta_{n})\frac{1-\kappa}{2\kappa-1}+1}+\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}^{2q+2}t_{n+1}^{\frac{1}{2}}
\displaystyle\leq 2tn(2(8+κ1κ)δn)1κ2κ1+1.\displaystyle 2t_{n}^{(2-(8+\frac{\kappa}{1-\kappa})\delta_{n})\frac{1-\kappa}{2\kappa-1}+1}.

On the other hand, by (4.20), we have

L(α,An+1)=L(α,A¯n)=4πtn(tnλn)\displaystyle L(\alpha,A^{\prime}_{n+1})=L(\alpha,\bar{A}^{\prime}_{n})=4\pi\sqrt{t_{n}(t_{n}-\lambda_{n})}
=\displaystyle= 4πtn(tn2λn2)tn+λn4πtn2λn22\displaystyle 4\pi\sqrt{\frac{t_{n}(t_{n}^{2}-\lambda_{n}^{2})}{t_{n}+\lambda_{n}}}\geq 4\pi\sqrt{\frac{t_{n}^{2}-\lambda_{n}^{2}}{2}}
=\displaystyle= 22πtn(14δn)1κ2κ1+1An+1Ahκn.\displaystyle 2\sqrt{2}\pi t_{n}^{(1-4\delta_{n})\frac{1-\kappa}{2\kappa-1}+1}\geq\|A^{\prime}_{n+1}-A\|_{h}^{\kappa_{n}}.

where

(4.21) κn:=(14δn)1κ2κ1+1(2(8+κ1κ)δn)1κ2κ1+1=κ4δn(1κ)1(8+κ1κ)δn(1κ).\displaystyle\kappa_{n}:=\frac{(1-4\delta_{n})\frac{1-\kappa}{2\kappa-1}+1}{(2-(8+\frac{\kappa}{1-\kappa})\delta_{n})\frac{1-\kappa}{2\kappa-1}+1}=\frac{\kappa-4\delta_{n}(1-\kappa)}{1-(8+\frac{\kappa}{1-\kappa})\delta_{n}(1-\kappa)}.

Direct calculation shows that

κ<κn<κ+cδn,\kappa<\kappa_{n}<\kappa+c\delta_{n},

where c(κ)=8((2κ1)(1κ)+κ2)c(\kappa)=8((2\kappa-1)(1-\kappa)+\kappa^{2}), which ends the proof. ∎

The two lemmas imply the theorem. ∎

4.2. Proof of Theorem 1.1:

For any (α,A)𝒜α\𝒰α(\alpha,A)\in\mathcal{AR}_{\alpha}\backslash\mathcal{UH}_{\alpha}, we only need to conjugate the cocycle (α,A)(\alpha,A), such that the conjugated cocycle is close to the identity, then we can apply Proposition 4.1 to finish the proof. However, as we will see, some quantitative estimates are still needed. Therefore, we will first conjugate the global almost reducible cocycle to the local regime (for example, as defined in Proposition 7.1), then apply local KAM to get the desired results.

The following result is standard, but the basis of the proof. We will sketch the proof in the appendix, for the sake of completeness.

Proposition 4.2.

Let αDCd\alpha\in\mathrm{DC}_{d}, ASL(2,)A\in\mathrm{SL}(2,{\mathbb{R}}), h>h>0h>h^{\prime}>0. There exist ϵ=ϵ(A,α,h,h)\epsilon=\epsilon(\|A\|,\alpha,h,h^{\prime}) such if FChω(𝕋d,sl(2,))F\in C^{\omega}_{h}({\mathbb{T}}^{d},\mathrm{sl}(2,{\mathbb{R}})) and Fh<ϵ\|F\|_{h}<\epsilon then, for all n1n\geq 1, there exist FnChω(𝕋d,sl(2,))F_{n}\in C_{h^{\prime}}^{\omega}({\mathbb{T}}^{d},\mathrm{sl}(2,{\mathbb{R}})), BnChω(𝕋d,PSL(2,))B_{n}\in C_{h^{\prime}}^{\omega}({\mathbb{T}}^{d},\mathrm{PSL}(2,{\mathbb{R}})) and AnSL(2,)A_{n}\in\mathrm{SL}(2,{\mathbb{R}}), satisfying

Bn1(+α)AeF()Bn()=AneFn(),B_{n}^{-1}(\cdot+\alpha)Ae^{F(\cdot)}B_{n}(\cdot)=A_{n}e^{F_{n}(\cdot)},
AnA+2ϵ.\|A_{n}\|\leq\|A\|+2\epsilon.

and

limn+Bnh6Fnh=0.\lim_{n\rightarrow+\infty}\|B_{n}\|_{h^{\prime}}^{6}\|F_{n}\|_{h^{\prime}}=0.

Since (α,A)𝒜α(\alpha,A)\in\mathcal{AR}_{\alpha}, then there exist B~0Chω(𝕋d,PSL(2,))\tilde{B}_{0}\in C^{\omega}_{h}({\mathbb{T}}^{d},\mathrm{PSL}(2,{\mathbb{R}})) such that

(4.22) B~01(+α)A()B~0()=A0eF0().\tilde{B}_{0}^{-1}(\cdot+\alpha)A(\cdot)\tilde{B}_{0}(\cdot)=A_{0}e^{F_{0}(\cdot)}.

with

F0h1<ϵ4A0,\|F_{0}\|_{h_{1}}<\frac{\epsilon}{4\|A_{0}\|},

where ϵ=ϵ(A~1,α,h,h)\epsilon=\epsilon(\|\tilde{A}_{1}\|,\alpha,h,h^{\prime}) as in Proposition 4.2. Now by Proposition 4.2, for any ε>0\varepsilon>0, there exist BNChω(𝕋d,PSL(2,))B^{\prime}_{N}\in C_{h^{\prime}}^{\omega}({\mathbb{T}}^{d},\mathrm{PSL}(2,{\mathbb{R}})) such that

(4.23) BN1(+α)A0eF()BN()=ANeFN(),B_{N}^{{}^{\prime}-1}(\cdot+\alpha)A_{0}e^{F(\cdot)}B^{\prime}_{N}(\cdot)=A_{N}e^{F_{N}(\cdot)},

satisfying

(4.24) BNh6FNhε32B~0h2.\|B_{N}\|_{h^{\prime}}^{6}\|F_{N}\|_{h^{\prime}}\leq\frac{\varepsilon^{3}}{2\|\tilde{B}_{0}\|_{h}^{2}}.

Let BN:=B~0BNB_{N}:=\tilde{B}_{0}B^{\prime}_{N}, then we have

BN(+α)ANBN1()A()h\displaystyle\|B_{N}(\cdot+\alpha)A_{N}B_{N}^{-1}(\cdot)-A(\cdot)\|_{h^{\prime}}
\displaystyle\leq BNh2ANBN1(+α)A()BN()h\displaystyle\|B_{N}\|_{h^{\prime}}^{2}\|A_{N}-B_{N}^{-1}(\cdot+\alpha)A(\cdot)B_{N}(\cdot)\|_{h^{\prime}}
\displaystyle\leq B~0h2BNh2ANANeFNh\displaystyle\|\tilde{B}_{0}\|_{h}^{2}\|B^{\prime}_{N}\|_{h^{\prime}}^{2}\|A_{N}-A_{N}e^{F_{N}}\|_{h^{\prime}}
\displaystyle\leq 2(A0+2ϵ)B~0h2BNh2FNhε4.\displaystyle 2(\|A_{0}\|+2\epsilon)\|\tilde{B}_{0}\|_{h}^{2}\|B^{\prime}_{N}\|_{h^{\prime}}^{2}\|F_{N}\|_{h^{\prime}}\leq\frac{\varepsilon}{4}.

We now separate three cases, following the regime to which ANA_{N} belongs.

𝐂𝐚𝐬𝐞𝐈\mathbf{Case\ I}: ANA_{N} is elliptic. Then there exists PSL(2,)P\in\mathrm{SL}(2,{\mathbb{R}}) such that

P1ANP=Rθ,P^{-1}A_{N}P=R_{\theta},

where RθSO(2,)R_{\theta}\in SO(2,{\mathbb{R}}). Since αd/d\alpha\in{\mathbb{R}}^{d}/{\mathbb{Z}}^{d} is rationally independent, there exists kdk\in{\mathbb{Z}}^{d} such that

θk,α𝕋<ε16BNh2P2.\|\theta-\langle k,\alpha\rangle\|_{{\mathbb{T}}}<\frac{\varepsilon}{16\|B_{N}\|_{h^{\prime}}^{2}\|P\|^{2}}.

By Proposition 4.1, there exists A¯()Chω(𝕋d,SL(2,))\bar{A}(\cdot)\in C^{\omega}_{h^{\prime}}({\mathbb{T}}^{d},\mathrm{SL}(2,{\mathbb{R}})) such that Lyapunov exponent is exactly κ\kappa-Hölder continuous at (α,A¯())(\alpha,\bar{A}(\cdot)) and

A¯Idh<εe16π|k|h16BNh2P2.\|\bar{A}-\mathrm{Id}\|_{h^{\prime}}<\frac{\varepsilon e^{-16\pi|k|h^{\prime}}}{16\|B_{N}\|_{h^{\prime}}^{2}\|P\|^{2}}.

Let H:=M1(ek()00ek())MH:=M^{-1}\begin{pmatrix}e_{k}(\cdot)&0\\ 0&e_{-k}(\cdot)\end{pmatrix}M. Set

A()=BN(+α)PH(+α)A¯()H1()P1BN1().A^{\prime}(\cdot)=B_{N}(\cdot+\alpha)PH(\cdot+\alpha)\bar{A}(\cdot)H^{-1}(\cdot)P^{-1}B_{N}^{-1}(\cdot).

Then, we have

A()A()h\displaystyle\|A(\cdot)-A^{\prime}(\cdot)\|_{h^{\prime}} BNh2P2Hh2A¯Idh\displaystyle\leq\|B_{N}\|_{h^{\prime}}^{2}\|P\|^{2}\|H\|_{h^{\prime}}^{2}\|\bar{A}-\mathrm{Id}\|_{h^{\prime}}
+BNh2P2RθRk,α+BN(+α)ANBN1()A()h\displaystyle+\|B_{N}\|_{h^{\prime}}^{2}\|P\|^{2}\|R_{\theta}-R_{\langle k,\alpha\rangle}\|+\|B_{N}(\cdot+\alpha)A_{N}B_{N}^{-1}(\cdot)-A(\cdot)\|_{h^{\prime}}
ε.\displaystyle\leq\varepsilon.

Moreover, since the Lyapunov exponent is invariant under conjugation, one can easily check that

lim infBA00log|L(α,A)L(α,B)|logBA0=κ,\liminf_{\|B-A^{\prime}\|_{0}\rightarrow 0}\frac{\log|L(\alpha,A^{\prime})-L(\alpha,B)|}{\log\|B-A^{\prime}\|_{0}}=\kappa,

which means that the Lyapunov exponent is exactly κ\kappa-Hölder continuous at (α,A())(\alpha,A^{\prime}(\cdot)).

𝐂𝐚𝐬𝐞𝐈𝐈\mathbf{Case\ II}: ANA_{N} is parabolic. In this case, without loss of generality, we assume the eigenvalues of ANA_{N} are {1,1}\{1,1\}. Then there exists PSL(2,)P\in\mathrm{SL}(2,{\mathbb{R}}) such that

P1ANP=(1101).P^{-1}A_{N}P=\begin{pmatrix}1&1\\ 0&1\end{pmatrix}.

Let

AN:=P(1δ1δ1)P1.A_{N}^{\prime}:=P\begin{pmatrix}1-\delta&1\\ -\delta&1\end{pmatrix}P^{-1}.

where δ:=ε4BNh2P2\delta:=\frac{\varepsilon}{4\|B_{N}\|_{h}^{2}\|P\|^{2}}, then ANA_{N}^{\prime} is elliptic, and, moreover,

A()BN(+α)ANBN1()h\displaystyle\|A(\cdot)-B_{N}(\cdot+\alpha)A_{N}^{\prime}B_{N}^{-1}(\cdot)\|_{h^{\prime}}
\displaystyle\leq BN(+α)ANBN1()A()h+BNh2P2δ\displaystyle\|B_{N}(\cdot+\alpha)A_{N}B_{N}^{-1}(\cdot)-A(\cdot)\|_{h^{\prime}}+\|B_{N}\|_{h^{\prime}}^{2}\|P\|^{2}\delta
\displaystyle\leq ε.\displaystyle\varepsilon.

This situation has been transformed into 𝐂𝐚𝐬𝐞𝐈\mathbf{Case\ I}, which ends the proof for this case.

𝐂𝐚𝐬𝐞𝐈𝐈𝐈\mathbf{Case\ III}: ANA_{N} is hyperbolic. Let the eigenvalues of ANA_{N} be {λ,λ1}\{\lambda,\lambda^{-1}\} with λ>1\lambda>1.

We first consider the case

(4.25) |λ1|>2FN13.|\lambda-1|>2\|F_{N}\|^{\frac{1}{3}}.

In view of Proposition 18 of [48], there exists PSL(2,)P\in{\rm SL}(2,{\mathbb{R}}), with

P2(AN|λ1|)12\|P\|\leq 2\left(\frac{\|A_{N}\|}{|\lambda-1|}\right)^{\frac{1}{2}}

such that

P1ANP=(λ00λ1).P^{-1}A_{N}P=\begin{pmatrix}\lambda&0\\ 0&\lambda^{-1}\end{pmatrix}.

Then

P1BN1(+α)A()BN()P=(λ00λ1)eF~N(),P^{-1}B_{N}^{-1}(\cdot+\alpha)A(\cdot)B_{N}(\cdot)P=\begin{pmatrix}\lambda&0\\ 0&\lambda^{-1}\end{pmatrix}e^{\tilde{F}_{N}(\cdot)},

with F~Nh8|λ1|FNh\|\tilde{F}_{N}\|_{h^{\prime}}\leq\frac{8}{|\lambda-1|}\|F_{N}\|_{h^{\prime}}, and (4.25) implies that

|λ1|>F~Nh12|\lambda-1|>\|\tilde{F}_{N}\|_{h^{\prime}}^{\frac{1}{2}}

Consequently, (α,A())(\alpha,A(\cdot)) is uniformly hyperbolic by the usual cone criterion [55], which contradicts our assumptions.

Therefore, |λ1|<2FN13.|\lambda-1|<2\|F_{N}\|^{\frac{1}{3}}. Consequently, there exists an elliptic matrix ANA_{N}^{\prime}, such that

ANAN2|λ1|<4FN13.\|A_{N}-A_{N}^{\prime}\|\leq 2|\lambda-1|<4\|F_{N}\|^{\frac{1}{3}}.

Then by (4.24), we have

A()BN(+α)ANBN1()h\displaystyle\|A(\cdot)-B_{N}(\cdot+\alpha)A_{N}^{\prime}B_{N}^{-1}(\cdot)\|_{h^{\prime}}
\displaystyle\leq BN(+α)ANBN1()A()h+4BNh2FN13\displaystyle\|B_{N}(\cdot+\alpha)A_{N}B_{N}^{-1}(\cdot)-A(\cdot)\|_{h^{\prime}}+4\|B_{N}\|_{h^{\prime}}^{2}\|F_{N}\|^{\frac{1}{3}}
\displaystyle\leq ε,\displaystyle\varepsilon,

which again transforms this case into 𝐂𝐚𝐬𝐞𝐈\mathbf{Case\ I}, which concludes the proof.

5. Sub-linear growth of extended eigenfunction

Proposition 5.1.

Let α𝕋d\alpha\in{\mathbb{T}}^{d} be rationally independent and fix h>0h>0, ϵ>0\epsilon>0, and a non-increasing sequence {g(n)}n=1\{g(n)\}_{n=1} satisfying 0<g(n)<10<g(n)<1 and limnng(n)=\lim_{n\rightarrow\infty}n^{g(n)}=\infty. Then there exists (α,A())𝕋d×Chω(𝕋d,SL(2,))(\alpha,A(\cdot))\in{\mathbb{T}}^{d}\times C^{\omega}_{h}({\mathbb{T}}^{d},\mathrm{SL}(2,{\mathbb{R}})) which has sub-linear growth with rate {g(n)}n=1\{g(n)\}_{n=1}. Moreover, it satisfies

A()Idh<ϵ.\|A(\cdot)-\mathrm{Id}\|_{h}<\epsilon.
Proof.

We construct A()A(\cdot) iteratively. Firstly we construct the sequence {kn}n=0\{k_{n}\}_{n=0}^{\infty}. Let k0=0k_{0}=0. Assuming we have constructed kj,jnk_{j},j\leq n. We choose kn+1dk_{n+1}\in{\mathbb{Z}}^{d} satisfying the following:

(5.3) {(kn+1,α𝕋)18g([14kn+1,α𝕋])<ϵ32n=0,(kn+1,α𝕋)18g([14kn+1,α𝕋])<ϵ2e16π|kn|hkn,α𝕋44n+2n1.\displaystyle\left\{\begin{array}[]{lr}(\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}})^{\frac{1}{8}g([\frac{1}{4\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}}])}<\frac{\epsilon}{32}&n=0,\\ (\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}})^{\frac{1}{8}g([\frac{1}{4\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}}])}<\frac{\epsilon^{2}e^{-16\pi|k_{n}|h}\|\langle k_{n},\alpha\rangle\|_{{\mathbb{T}}}^{4}}{4^{n+2}}&n\geq 1.\end{array}\right.
(5.4) |kn+1|>e|kn|+10.|k_{n+1}|>e^{|k_{n}|}+10.

The sequence kn+1k_{n+1} always exists since α𝕋d\alpha\in{\mathbb{T}}^{d} is rationally independent and limnng(n)=\lim_{n\rightarrow\infty}n^{g(n)}=\infty. We now call

(5.5) tn\displaystyle t_{n} =kn+1,α𝕋34g(Nn),\displaystyle=\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}^{\frac{3}{4}g(N_{n})},
λn\displaystyle\lambda_{n} =tn2kn+1,α𝕋2.\displaystyle=\sqrt{t_{n}^{2}-\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}^{2}}.

where

(5.6) Nn=[14kn+1,α𝕋]N_{n}=[\frac{1}{4{\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}}}]

With these parameters, one can construct (α,A~neFn())(\alpha,\tilde{A}_{n}e^{F_{n}(\cdot)}) and then (α,An())(\alpha,A_{n}(\cdot)) by the Anosov-Katok method as in §3. Convergence of An()A_{n}(\cdot) follows from the following inequality

(5.7) 4πt0+n=14πtne4πhΣj=1n|kj|j=0n12tjkj+1,α𝕋<ϵ.\displaystyle 4\pi t_{0}+\sum_{n=1}^{\infty}4\pi t_{n}e^{4\pi h\Sigma_{j=1}^{n}|k_{j}|}\prod_{j=0}^{n-1}\frac{2t_{j}}{\|\langle k_{j+1},\alpha\rangle\|_{{\mathbb{T}}}}<\epsilon.

To show this, notice that (5.3) directly implies that

kn+1,α𝕋<kn+1,α𝕋18g(Nn)<kn,α𝕋4,\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}<\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}^{\frac{1}{8}g(N_{n})}<\|\langle k_{n},\alpha\rangle\|_{{\mathbb{T}}}^{4},

then by (5.4) and the choice of tnt_{n}, we have

j=0n12tjkj+1,α𝕋\displaystyle\prod_{j=0}^{n-1}\frac{2t_{j}}{\|\langle k_{j+1},\alpha\rangle\|_{{\mathbb{T}}}} (j=0n1kj+1,α𝕋)1kn,α𝕋(Σi=0n114i)\displaystyle\leq(\prod_{j=0}^{n-1}\|\langle k_{j+1},\alpha\rangle\|_{{\mathbb{T}}})^{-1}\leq\|\langle k_{n},\alpha\rangle\|_{{\mathbb{T}}}^{-(\Sigma_{i=0}^{n-1}\frac{1}{4^{i}})}
(5.8) kn,α𝕋2kn+1,α𝕋116g(Nn).\displaystyle\leq\|\langle k_{n},\alpha\rangle\|_{{\mathbb{T}}}^{-2}\leq\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}^{-\frac{1}{16}g(N_{n})}.

Therefore by (5.3), (5.4), (5.5), we have

4πt0+n=14πtne4πhΣj=1n|kj|j=0n12tjkj+1,α𝕋\displaystyle 4\pi t_{0}+\sum_{n=1}^{\infty}4\pi t_{n}e^{4\pi h\Sigma_{j=1}^{n}|k_{j}|}\prod_{j=0}^{n-1}\frac{2t_{j}}{\|\langle k_{j+1},\alpha\rangle\|_{{\mathbb{T}}}}
\displaystyle\leq 4πt0+n=14πtne8πh|kn|kn,α𝕋2\displaystyle 4\pi t_{0}+\sum_{n=1}^{\infty}4\pi t_{n}e^{8\pi h|k_{n}|}\|\langle k_{n},\alpha\rangle\|_{{\mathbb{T}}}^{-2}
\displaystyle\leq πϵ8+n=14πϵ2n+2kn+1,α𝕋1116g(Nn)ϵ.\displaystyle\frac{\pi\epsilon}{8}+\sum_{n=1}^{\infty}\frac{4\pi\epsilon}{2^{n+2}}\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}^{\frac{11}{16}g(N_{n})}\leq\epsilon.

which establishes eq. (5.7). We can now apply Lemma 3.1, and thus obtain A()Chω(𝕋d,SL(2,))A(\cdot)\in C^{\omega}_{h}({\mathbb{T}}^{d},\mathrm{SL}(2,{\mathbb{R}})) such that M1A()M=A()M^{-1}A_{\infty}(\cdot)M=A(\cdot) and

A()Idh=A()A0h<ϵ.\|A(\cdot)-\mathrm{Id}\|_{h}=\|A(\cdot)-A_{0}\|_{h}<\epsilon.

Moreover,

(5.9) Bn1(+α)A()Bn()=A~n+F~n(),B_{n}^{-1}(\cdot+\alpha)A_{\infty}(\cdot)B_{n}(\cdot)=\tilde{A}_{n}+\tilde{F}_{n}(\cdot),

satisfying the estimate

F~nh\displaystyle\|\tilde{F}_{n}\|_{h} 4πtne4π|kn|h+j=n+14πtje4πhΣi=nj|kj|i=nj12tiki+1,α𝕋\displaystyle\leq 4\pi t_{n}e^{4\pi|k_{n}|h}+\sum_{j=n+1}^{\infty}4\pi t_{j}e^{4\pi h\Sigma_{i=n}^{j}|k_{j}|}\prod_{i=n}^{j-1}\frac{2t_{i}}{\|\langle k_{i+1},\alpha\rangle\|_{{\mathbb{T}}}}
<j=n4πtje8π|kj|hkj,α𝕋2j=n4πϵ2n+2kj+1,α𝕋1116g(Nj)\displaystyle<\sum_{j=n}^{\infty}4\pi t_{j}e^{8\pi|k_{j}|h}\|\langle k_{j},\alpha\rangle\|_{{\mathbb{T}}}^{-2}\leq\sum_{j=n}^{\infty}\frac{4\pi\epsilon}{2^{n+2}}\|\langle k_{j+1},\alpha\rangle\|_{{\mathbb{T}}}^{\frac{11}{16}g(N_{j})}
(5.10) kn+1,α𝕋1116g(Nn),\displaystyle\leq\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}^{\frac{11}{16}g(N_{n})},

The estimates hold because of equations (5.3), (5.4) , (5.5) and the estimate (5.8). To estimate the growth of (α,A())(\alpha,A(\cdot)), we first estimate the growth of the approximating cocycle (α,An+1())(\alpha,A_{n+1}(\cdot)).

Lemma 5.1.

Letting

Nn=[14kn+1,α𝕋],N_{n}=[\frac{1}{4{\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}}}],

we have

M1An+1(;Nn)M0[Dn22j=0n1Dj2,j=0nDj2].\|M^{-1}A_{n+1}(\cdot;N_{n})M\|_{0}\in[\frac{\|D_{n}\|^{2}}{2\prod_{j=0}^{n-1}\|D_{j}\|^{2}},\prod_{j=0}^{n}\|D_{j}\|^{2}].
Proof.

First by our construction (3.5), we have

An+1()\displaystyle A_{n+1}(\cdot) =\displaystyle= Bn+1(+α)A~n+1Bn+11()\displaystyle B_{n+1}(\cdot+\alpha)\tilde{A}_{n+1}B_{n+1}^{-1}(\cdot)
=\displaystyle= Bn+1(+α)Dn1(DnA~n+1Dn1)DnBn+11()\displaystyle B_{n+1}(\cdot+\alpha)D_{n}^{-1}(D_{n}\tilde{A}_{n+1}D_{n}^{-1})D_{n}B_{n+1}^{-1}(\cdot)
=\displaystyle= Bn+1(+α)Dn1(Dn(ekn+1(α)00ekn+1(α))Dn1)DnBn+11().\displaystyle B_{n+1}(\cdot+\alpha)D_{n}^{-1}(D_{n}\begin{pmatrix}e_{k_{n+1}}(\alpha)&0\\ 0&e_{-k_{n+1}}(\alpha)\end{pmatrix}D_{n}^{-1})D_{n}B_{n+1}^{-1}(\cdot).

Then for any ll\in{\mathbb{N}} we have

M1An+1(;l)M0\displaystyle\|M^{-1}A_{n+1}(\cdot;l)M\|_{0}
=\displaystyle= M1Bn+1(+lα)Dn1(DnA~n+1Dn1)lDnBn+11()M0\displaystyle\|M^{-1}B_{n+1}(\cdot+l\alpha)D_{n}^{-1}(D_{n}\tilde{A}_{n+1}D_{n}^{-1})^{l}D_{n}B_{n+1}^{-1}(\cdot)M\|_{0}
(5.11) \displaystyle\in [(DnA~n+1Dn1)l0Bn+1()Dn102,(DnA~n+1Dn1)l0Bn+1()Dn102].\displaystyle[\frac{\|(D_{n}\tilde{A}_{n+1}D_{n}^{-1})^{l}\|_{0}}{\|B_{n+1}(\cdot)D_{n}^{-1}\|_{0}^{2}},\|(D_{n}\tilde{A}_{n+1}D_{n}^{-1})^{l}\|_{0}\|B_{n+1}(\cdot)D_{n}^{-1}\|_{0}^{2}].

Since M1DnMSL(2,)M^{-1}D_{n}M\in\mathrm{SL}(2,{\mathbb{R}}), we have the Singular value decomposition of M1DnMM^{-1}D_{n}M:

M1DnM=Run(Dn00Dn1)Rsn,M^{-1}D_{n}M=R_{u_{n}}\begin{pmatrix}\|D_{n}\|&0\\ 0&\|D_{n}\|^{-1}\end{pmatrix}R_{s_{n}},

for some un,sn[0,2π)u_{n},s_{n}\in[0,2\pi). Thus

(5.12) M1(DnA~n+1Dn1)lM\displaystyle M^{-1}(D_{n}\tilde{A}_{n+1}D_{n}^{-1})^{l}M
=\displaystyle= Run(Dn00Dn1)RsnRϕRsn(Dn100Dn)Run\displaystyle R_{u_{n}}\begin{pmatrix}\|D_{n}\|&0\\ 0&\|D_{n}\|^{-1}\end{pmatrix}R_{s_{n}}R_{\phi}R_{-s_{n}}\begin{pmatrix}\|D_{n}\|^{-1}&0\\ 0&\|D_{n}\|\end{pmatrix}R_{-u_{n}}
=\displaystyle= Run(cosϕsinϕDn2sinϕDn2cosϕ)Run.\displaystyle R_{u_{n}}\begin{pmatrix}\cos\phi&-\sin\phi\|D_{n}\|^{2}\\ \sin\phi\|D_{n}\|^{-2}&\cos\phi\end{pmatrix}R_{-u_{n}}.

where ϕ=2πlkn+1,α𝕋\phi=2\pi l\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}.

The key observation is that if l=Nnl=N_{n}, we have

ϕ(π4,3π4),\phi\in(\frac{\pi}{4},\frac{3\pi}{4}),

and

(DnA~n+1Dn1)Nn0[Dn22,Dn2].\|(D_{n}\tilde{A}_{n+1}D_{n}^{-1})^{N_{n}}\|_{0}\in[\frac{\|D_{n}\|^{2}}{2},\|D_{n}\|^{2}].

By construction,

Bn+1()=Bn()Gn()=Bn()Hn()Dn,B_{n+1}(\cdot)=B_{n}(\cdot)G_{n}(\cdot)=B_{n}(\cdot)H_{n}(\cdot)D_{n},

so that

Bn+1()Dn102=Bn()02j=0n1Dj2.\displaystyle\|B_{n+1}(\cdot)D_{n}^{-1}\|_{0}^{2}=\|B_{n}(\cdot)\|_{0}^{2}\leq\prod_{j=0}^{n-1}\|D_{j}\|^{2}.

Thus we have

M1An+1(;Nn)M0[Dn22j=0n1Dj2,j=0nDj2]\displaystyle\|M^{-1}A_{n+1}(\cdot;N_{n})M\|_{0}\in[\frac{\|D_{n}\|^{2}}{2\prod_{j=0}^{n-1}\|D_{j}\|^{2}},\prod_{j=0}^{n}\|D_{j}\|^{2}]

which proves the lemma. ∎

Now, by eq. (3.3), we have

Dn2[12kn+1,α𝕋134g(Nn),4kn+1,α𝕋134g(Nn)],\|D_{n}\|^{2}\in[\frac{1}{2\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}^{1-\frac{3}{4}g(N_{n})}},\frac{4}{\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}^{1-\frac{3}{4}g(N_{n})}}],

and, by eq. (5.8),

(5.13) Bn()02j=0n1Dj2<j=0n12tjkj+1,α𝕋<kn+1,α𝕋116g(Nn)\|B_{n}(\cdot)\|_{0}^{2}\leq\prod_{j=0}^{n-1}\|D_{j}\|^{2}<\prod_{j=0}^{n-1}\frac{2t_{j}}{\|\langle k_{j+1},\alpha\rangle\|_{{\mathbb{T}}}}<\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}^{-\frac{1}{16}g(N_{n})}

Consequently, by Lemma 5.1, we have

(5.14) M1An+1(;Nn)M0[14kn+1,α𝕋178g(Nn),4kn+1,α𝕋158g(Nn)].\|M^{-1}A_{n+1}(\cdot;N_{n})M\|_{0}\in[\frac{1}{4\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}^{1-\frac{7}{8}g(N_{n})}},\frac{4}{\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}^{1-\frac{5}{8}g(N_{n})}}].

On the other hand, by (5.9) and (5.10), we have

A(;Nn)M1An+1(;Nn)M0\displaystyle\|A(\cdot;N_{n})-M^{-1}A_{n+1}(\cdot;N_{n})M\|_{0}
=\displaystyle= Bn+1(+Nnα)((A~n+1+F~n+1)(;Nn)A~n+1Nn)Bn+11()0\displaystyle\|B_{n+1}(\cdot+N_{n}\alpha)((\tilde{A}_{n+1}+\tilde{F}_{n+1})(\cdot;N_{n})-\tilde{A}_{n+1}^{N_{n}})B_{n+1}^{-1}(\cdot)\|_{0}
\displaystyle\leq 2NnBn+102F~n+1h\displaystyle 2N_{n}\|B_{n+1}\|_{0}^{2}\|\tilde{F}_{n+1}\|_{h}
\displaystyle\leq 2Nnkn+2,α𝕋116g(Nn+1)kn+2,α𝕋1116g(Nn+1)\displaystyle 2N_{n}\|\langle k_{n+2},\alpha\rangle\|_{{\mathbb{T}}}^{-\frac{1}{16}g(N_{n+1})}\|\langle k_{n+2},\alpha\rangle\|_{{\mathbb{T}}}^{\frac{11}{16}g(N_{n+1})}
(5.15) \displaystyle\leq kn+1,α𝕋4.\displaystyle\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}^{4}.

Combining (5.14) and (5.15) can finish the proof of the proposition. ∎

Corollary 5.1.

Given αDCd\alpha\in\mathrm{DC}_{d}. For any non-increasing sequence {g(n)}n=1\{g(n)\}_{n=1} satisfying 0<g(n)<10<g(n)<1 and limnng(n)=\lim_{n\rightarrow\infty}n^{g(n)}=\infty. There exists a set 𝔇\mathfrak{D} which is dense in 𝒜α\𝒰α\mathcal{AR}_{\alpha}\backslash\mathcal{UH}_{\alpha} in the Cω(𝕋d,SL(2,))C^{\omega}({\mathbb{T}}^{d},\mathrm{SL}(2,{\mathbb{R}})) topology such that A(;n)0\|A(\cdot;n)\|_{0} has sub-linear growth with rate {g(n)}n=1\{g(n)\}_{n=1} at each point of (α,A)𝔇(\alpha,A)\in\mathfrak{D}.

Proof.

The proof is same as Theorem 1.1, one only need to replace Proposition 4.1 by Proposition 4.1. ∎

Proof of Theorem 1.2

If d=1d=1, just note (α,A)𝒜α\𝒰α(\alpha,A)\in\mathcal{AR}_{\alpha}\backslash\mathcal{UH}_{\alpha}, if and only if (α,A)(\alpha,A) is subcritical. In fact, by Avila’s almost reducibility theorem (ART) [2, 4, 5], (α,A)(\alpha,A) is subcritical, then it is almost reducible and not uniformly hyperbolic. Conversely, if (α,A)(\alpha,A) is almost reducible but not uniformly hyperbolic, then the Lyapunov exponent vanishes in a band [7], which ensures the cocycle is subcritical.

Therefore, if EΣV,αsubE\in\Sigma_{V,\alpha}^{sub}, then (α,SEV)(\alpha,S_{E}^{V}) is subcritical, then by Corollary 5.1, one can perturb (α,SEV)(\alpha,S_{E}^{V}) to (α,A)(\alpha,A^{\prime}), so that it has sub-linear growth. Then the result follows immediately from the following lemma.

Lemma 5.2 (Avila-Jitomirskaya [8]).

Let Let α𝕋d\alpha\in{\mathbb{T}}^{d} be rationally independent, AChω(𝕋d,SL(2,))A\in C_{h_{*}}^{\omega}({\mathbb{T}}^{d},{\rm SL}(2,{\mathbb{R}})) for some h>0h_{*}>0, such that (α,A)(\alpha,A) is almost reducible. There exists h0(0,h)h_{0}\in(0,h_{*}) such that for any η>0\eta>0, one can find VCh0ω(𝕋d,)V\in C_{h_{0}}^{\omega}({\mathbb{T}}^{d},{\mathbb{R}}) with |V|h0<η|V|_{h_{0}}<\eta, EE\in{\mathbb{R}}, and ZCh0ω(𝕋d,PSL(2,))Z\in C_{h_{0}}^{\omega}({\mathbb{T}}^{d},{\rm PSL}(2,{\mathbb{R}})) such that

Z(+α)1A()Z()=SEV().Z(\cdot+\alpha)^{-1}A(\cdot)Z(\cdot)=S_{E}^{V}(\cdot).

Moreover, for every 0<hh00<h\leq h_{0}, there is δ>0\delta>0 such that if AChω(𝕋d,SL(2,))A^{\prime}\in C_{h}^{\omega}({\mathbb{T}}^{d},{\rm SL}(2,{\mathbb{R}})) satisfies |AA|h<δ|A-A^{\prime}|_{h}<\delta, then there exist VChω(𝕋d,)V^{\prime}\in C_{h}^{\omega}({\mathbb{T}}^{d},{\mathbb{R}}) with |V|h<η|V^{\prime}|_{h}<\eta and ZChω(𝕋d,PSL(2,))Z^{\prime}\in C_{h}^{\omega}({\mathbb{T}}^{d},{\rm PSL}(2,{\mathbb{R}})) such that |ZZ|h<η|Z-Z^{\prime}|_{h}<\eta and

Z(+α)1A()Z()=SEV().Z^{\prime}(\cdot+\alpha)^{-1}A^{\prime}(\cdot)Z^{\prime}(\cdot)=S_{E}^{V^{\prime}}(\cdot).
Remark 5.1.

Avila-Jitomirskaya [8] state the result for α\\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}}. The proof, however, applies equally well to the multifrequency case.

6. Power-law localized eigenfunction

Proposition 6.1.

Given s+s\in{\mathbb{Z}}^{+}, h>0h>0 and αd/d\alpha\in{\mathbb{R}}^{d}/{\mathbb{Z}}^{d} which is rationally independent. Then for any ϵ>0\epsilon>0, there exists (α,A())𝕋d×Chω(𝕋d,SL(2,))(\alpha,A(\cdot))\in{\mathbb{T}}^{d}\times C^{\omega}_{h}({\mathbb{T}}^{d},\mathrm{SL}(2,{\mathbb{R}})) which is CsC^{s}-reducible but not Cs+1C^{s+1}-reducible. Moreover, it satisfies

A()Idh<ϵ.\|A(\cdot)-\mathrm{Id}\|_{h}<\epsilon.

We again construct A()A(\cdot) by the Anosov-Katok method. The construction, however, will be a bit different from Section 3. Since the goal is to construct (α,A())(\alpha,A(\cdot)) which is reducible and not merely almost reducible.

Proof.

First, we construct a sequence {kn}n=0\{k_{n}\}_{n=0}^{\infty} where kndk_{n}\in{\mathbb{Z}}^{d}. Let k0=0k_{0}=0, kn=Σj=0nknk_{n}^{\prime}=\Sigma_{j=0}^{n}k_{n}. Suppose we have constructed ki,ink_{i},i\leq n, then we choose kn+1dk_{n+1}\in{\mathbb{Z}}^{d} satisfying

(6.1) kn+1,α𝕋\displaystyle\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}} <\displaystyle< ϵ4ne16π|kn|h,\displaystyle\frac{\epsilon}{4^{n}}e^{-16\pi|k_{n}^{\prime}|h},
(6.2) |kn+1|\displaystyle|k_{n+1}| >\displaystyle> e|kn|+10.\displaystyle e^{|k_{n}^{\prime}|}+10.

Then we construct

tn\displaystyle t_{n} =kn+1,α𝕋|kn|s(n+10)2|kn|2s(n+10)41,\displaystyle=\frac{\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}|k_{n}^{\prime}|^{s}(n+10)^{2}}{\sqrt{|k_{n}^{\prime}|^{2s}(n+10)^{4}-1}},
λn\displaystyle\lambda_{n} =kn+1,α𝕋|kn|2s(n+10)41.\displaystyle=\frac{\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}}{\sqrt{|k_{n}^{\prime}|^{2s}(n+10)^{4}-1}}.

By direct calculation we have

(6.3) tn2λn2=kn+1,α𝕋,λntn=1|kn|s(n+10)2.\displaystyle\sqrt{t_{n}^{2}-\lambda_{n}^{2}}=\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}},\quad\frac{\lambda_{n}}{t_{n}}=\frac{1}{|k_{n}^{\prime}|^{s}(n+10)^{2}}.

Once we have these parameters, we perturb the cocycle (α,A~n)(\alpha,\tilde{A}_{n}^{\prime}), to(α,A~neFn())(\alpha,\tilde{A}_{n}^{\prime}e^{F_{n}(\cdot)^{\prime}}), where

A~n=(ekn(α)00ekn(α)),Fn()=2πi(tnλne2kn()λne2kn()tn).\tilde{A}_{n}^{\prime}=\begin{pmatrix}e_{k_{n}^{\prime}}(\alpha)&0\\ 0&e_{-k_{n}^{\prime}}(\alpha)\end{pmatrix},F_{n}^{\prime}(\cdot)=2\pi i\begin{pmatrix}t_{n}&\lambda_{n}e_{2k_{n}^{\prime}}(\cdot)\\ -\lambda_{n}e_{-2k_{n}^{\prime}}(\cdot)&-t_{n}\end{pmatrix}.

Let

Hn():=(ekn()00ekn()),H_{n}^{\prime}(\cdot):=\begin{pmatrix}e_{k_{n}^{\prime}}(\cdot)&0\\ 0&e_{-k_{n}^{\prime}}(\cdot)\end{pmatrix},

then we have

Hn1(+α)A~neFn()Hn()=exp2πi(tnλnλntn).H_{n}^{\prime-1}(\cdot+\alpha)\tilde{A}_{n}^{\prime}e^{F_{n}^{\prime}(\cdot)}H_{n}^{\prime}(\cdot)=\exp{2\pi i\begin{pmatrix}t_{n}&\lambda_{n}\\ -\lambda_{n}&-t_{n}\end{pmatrix}}.

By (6.3) and Lemma 2.1, there exists DnSU(1,1)D_{n}\in SU(1,1) such that

Dn1exp2πi(tnλnλntn)Dn=(ekn+1(α)00ekn+1(α)).D_{n}^{-1}\exp 2\pi i\begin{pmatrix}t_{n}&\lambda_{n}\\ -\lambda_{n}&-t_{n}\end{pmatrix}D_{n}=\begin{pmatrix}e_{k_{n+1}}(\alpha)&0\\ 0&e_{-k_{n+1}}(\alpha)\end{pmatrix}.

Let Gn():=Hn()DnHn1()G_{n}^{\prime}(\cdot):=H_{n}^{\prime}(\cdot)D_{n}H_{n}^{\prime-1}(\cdot), then we have

Gn1(+α)A~neFnGn()=A~n+1,G_{n}^{\prime-1}(\cdot+\alpha)\tilde{A}_{n}^{\prime}e^{F_{n}^{\prime}}G_{n}^{\prime}(\cdot)=\tilde{A}_{n+1}^{\prime},

which means the cocycle (α,A~neFn())(\alpha,\tilde{A}_{n}^{\prime}e^{F_{n}(\cdot)^{\prime}}) is conjugated to (α,A~n+1)(\alpha,\tilde{A}_{n+1}^{\prime}), which concludes one step of the iteration.

Finally, let Bn:=G0Gn2Gn1B_{n}^{\prime}:=G_{0}^{\prime}\cdots G_{n-2}^{\prime}G_{n-1}^{\prime} and

An():=Bn(+α)A~nBn1(),A_{n}^{\prime}(\cdot):=B_{n}^{\prime}(\cdot+\alpha)\tilde{A}_{n}^{\prime}B_{n}^{\prime-1}(\cdot),

similarly as in Lemma 3.1, one can easily show that there exists A()Chω(𝕋d,SL(2,))A(\cdot)\in C_{h}^{\omega}({\mathbb{T}}^{d},\mathrm{SL}(2,{\mathbb{R}})) such that AnAh0\|A_{n}^{\prime}-A\|_{h}\rightarrow 0 and

A()Idh=A()A0<ϵ.\|A(\cdot)-\mathrm{Id}\|_{h}=\|A(\cdot)-A_{0}\|<\epsilon.

We point out the conjugacy Gn()G_{n}^{\prime}(\cdot) used here is the main difference with respect to the construction in Section 3. Indeed, by Lemma 2.1, DnD_{n} can be chosen in the form

Dn=(cos2θn)12(cosθnsinθnsinθncosθn),D_{n}=(\cos 2\theta_{n})^{-\frac{1}{2}}\begin{pmatrix}\cos\theta_{n}&\sin\theta_{n}\\ \sin\theta_{n}&\cos\theta_{n}\end{pmatrix},

therefore GnG_{n}^{\prime} is the form as

Gn=(cos2θn)12(cosθnsinθne2knsinθne2kncosθn).G_{n}^{\prime}=(\cos 2\theta_{n})^{-\frac{1}{2}}\begin{pmatrix}\cos\theta_{n}&\sin\theta_{n}e_{2k_{n}^{\prime}}\\ \sin\theta_{n}e_{-2k_{n}^{\prime}}&\cos\theta_{n}\end{pmatrix}.

By Corollary 2.1 and (6.3) we have

(6.4) |(cos2θn)12cosθn1|<1|kn|2s(n+10)41,\displaystyle|(\cos 2\theta_{n})^{-\frac{1}{2}}\cos\theta_{n}-1|<\frac{1}{|k_{n}^{\prime}|^{2s}(n+10)^{4}-1},

and

(6.5) |(cos2θn)12sinθn|(1|kn|s(n+10)2,2|kn|s(n+10)2).\displaystyle|(\cos 2\theta_{n})^{-\frac{1}{2}}\sin\theta_{n}|\in(\frac{1}{|k_{n}^{\prime}|^{s}(n+10)^{2}},\frac{2}{|k_{n}^{\prime}|^{s}(n+10)^{2}}).

Then we can get the following estimation on GnG_{n}^{\prime}:

(6.6) GnIds<2(n+10)2,\displaystyle\|G_{n}^{\prime}-\mathrm{Id}\|_{s}<\frac{2}{(n+10)^{2}},

By (6.6) we know that

(6.7) BnBn+1s=Bn(GnId)sC(s,d)(n+10)2.\displaystyle\|B_{n}^{\prime}-B_{n+1}^{\prime}\|_{s}=\|B_{n}^{\prime}(G_{n}^{\prime}-\mathrm{Id})\|_{s}\leq\frac{C(s,d)}{(n+10)^{2}}.

Thus, there exists B()Cs(𝕋d,SL(2,))B(\cdot)\in C^{s}({\mathbb{T}}^{d},\mathrm{SL}(2,{\mathbb{R}})) such that BnBs0\|B_{n}^{\prime}-B\|_{s}\rightarrow 0. Since A~nA~n+1<kn+1,α𝕋\|\tilde{A}_{n}^{\prime}-\tilde{A}_{n+1}^{\prime}\|<\|\langle k_{n+1},\alpha\rangle\|_{{\mathbb{T}}}, one can show that

(6.8) B(+α)1A()B()=A~=(e2πiρ00e2πiρ),\displaystyle B(\cdot+\alpha)^{-1}A(\cdot)B(\cdot)=\tilde{A}=\begin{pmatrix}e^{2\pi i\rho}&0\\ 0&e^{-2\pi i\rho}\end{pmatrix},

where e2πiρ=limnekn(α)e^{2\pi i\rho}=\lim_{n\rightarrow\infty}e_{k_{n}^{\prime}}(\alpha), i.e. (α,A())(\alpha,A(\cdot)) is CsC^{s}-reducible.

Next we will prove (α,A())(\alpha,A(\cdot)) is not Cs+1C^{s+1}-reducible. First, we have the following :

Claim 1.
(6.9) B()s+1=.\displaystyle\|B(\cdot)\|_{s+1}=\infty.
Proof.

Suppose that B()s+1=C<\|B(\cdot)\|_{s+1}=C<\infty. Then we know B^(k)C|k|(s+1)\|\widehat{B}(k)\|\leq C|k|^{-(s+1)} for all kdk\in{\mathbb{Z}}^{d}. By (6.7) we know

(6.10) B^(k)Bn^(k)C(s,d)|k|s(n+10)2.\displaystyle\|\widehat{B}(k)-\widehat{B_{n}^{\prime}}(k)\|\leq\frac{C(s,d)}{|k|^{s}(n+10)^{2}}.

We now analyze the structure of BnB_{n}. Let

un\displaystyle u_{n} :=(cos2θn)12cosθn1,\displaystyle:=(\cos 2\theta_{n})^{-\frac{1}{2}}\cos\theta_{n}-1,
vn\displaystyle v_{n} :=(cos2θn)12sinθn,\displaystyle:=(\cos 2\theta_{n})^{-\frac{1}{2}}\sin\theta_{n},

and

𝒥k:=(0ekek0),𝒢k:=(ek00ek).\mathcal{J}_{k}:=\begin{pmatrix}0&e_{k}\\ e_{-k}&0\end{pmatrix},\quad\mathcal{G}_{k}:=\begin{pmatrix}e_{k}&0\\ 0&e_{-k}\end{pmatrix}.

Therefore, we have

Bn\displaystyle B_{n}^{\prime} =j=0n1((1+uj)I+vj𝒥kj)\displaystyle=\prod_{j=0}^{n-1}((1+u_{j})I+v_{j}\mathcal{J}_{k_{j}^{\prime}})
=0j1,,jln1(jj1,,jl(1+uj))(m=1lvjm𝒥kjm).\displaystyle=\sum_{0\leq j_{1},\cdots,j_{l}\leq n-1}(\prod_{j\neq j_{1},\cdots,j_{l}}(1+u_{j}))(\prod_{m=1}^{l}v_{j_{m}}\mathcal{J}_{k_{j_{m}^{\prime}}}).

By direct calculation we know if ll is even

m=1l𝒥kjm=𝒢Σm=1l(1)m1kjm,\prod_{m=1}^{l}\mathcal{J}_{k_{j_{m}}^{\prime}}=\mathcal{G}_{\Sigma_{m=1}^{l}(-1)^{m-1}k_{j_{m}}^{\prime}},

otherwise

m=1l𝒥kjm=𝒥Σm=1l(1)m1kjm.\prod_{m=1}^{l}\mathcal{J}_{k_{j_{m}}^{\prime}}=\mathcal{J}_{\Sigma_{m=1}^{l}(-1)^{m-1}k_{j_{m}}^{\prime}}.

Now we need the following crucial observation: given two set Q:={i1,,ir}Q:=\{i_{1},\cdots,i_{r}\}, P:={j1,,jl}P:=\{j_{1},\cdots,j_{l}\} where in,jmi_{n},j_{m}\in{\mathbb{N}}. Then we have

(6.11) Σm=1l(1)m1kjm=Σn=1r(1)n1kin,\displaystyle\Sigma_{m=1}^{l}(-1)^{m-1}k_{j_{m}}^{\prime}=\Sigma_{n=1}^{r}(-1)^{n-1}k_{i_{n}}^{\prime},

if and only if Q=PQ=P. This holds since by our construction (6.2), we have |kn+1|Σj=0n|kj||k_{n+1}^{\prime}|\gg\Sigma_{j=0}^{n}|k_{j}^{\prime}|, if (6.11) satisfied then (1)l1kjl=(1)r1kir(-1)^{l-1}k_{j_{l}}^{\prime}=(-1)^{r-1}k_{i_{r}}^{\prime} must happen. Iterating this step gives P=QP=Q.

This observation implies that for any 0j1<<jln10\leq j_{1}<\cdots<j_{l}\leq n-1 we have

Bn^(Σm=1l(1)m1kjm)={jj1,,jl(1+uj)m=1lvjm(1000)leven,jj1,,jl(1+uj)m=1lvjm(0100)lodd.\widehat{B^{\prime}_{n}}(\Sigma_{m=1}^{l}(-1)^{m-1}k_{j_{m}}^{\prime})=\left\{\begin{array}[]{lr}\prod_{j\neq j_{1},\cdots,j_{l}}(1+u_{j})\prod_{m=1}^{l}v_{j_{m}}\begin{pmatrix}1&0\\ 0&0\end{pmatrix}&l\ \textit{even},\\ \prod_{j\neq j_{1},\cdots,j_{l}}(1+u_{j})\prod_{m=1}^{l}v_{j_{m}}\begin{pmatrix}0&1\\ 0&0\end{pmatrix}&l\ \textit{odd}.\end{array}\right.

In particular, for jn1j\leq n-1 we have

Bn^(kj)=vj(ij,0in1(1+ui))(0100).\widehat{B_{n}^{\prime}}(k_{j}^{\prime})=v_{j}(\prod_{i\neq j,0\leq i\leq n-1}(1+u_{i}))\begin{pmatrix}0&1\\ 0&0\end{pmatrix}.

By equations (6.4) and (6.5), for any n1n\geq 1, jn1j\leq n-1 we have

(6.12) Bn^(kj)>12|kj|s(j+10)2\displaystyle\|\widehat{B_{n}^{\prime}}(k_{j}^{\prime})\|>\frac{1}{2|k_{j}^{\prime}|^{s}(j+10)^{2}}

Because of (6.10) and (6.12), if nC(s,d)n\geq C(s,d) which is large enough, we have

B^(kj)Bn^(kj)B^(kj)Bn^(kj)>14|kj|s(j+10)2.\|\widehat{B}(k_{j}^{\prime})\|\geq\|\widehat{B_{n}^{\prime}}(k_{j}^{\prime})\|-\|\widehat{B}(k_{j}^{\prime})-\widehat{B_{n}^{\prime}}(k_{j}^{\prime})\|>\frac{1}{4|k_{j}^{\prime}|^{s}(j+10)^{2}}.

By the assumption Bs+1<C\|B\|_{s+1}<C we have

14|kj|s(j+10)2<C|kj|s+1,\frac{1}{4|k_{j}^{\prime}|^{s}(j+10)^{2}}<\frac{C}{|k_{j}^{\prime}|^{s+1}},

which contradicts (6.2). ∎

Now we finish the proof that the cocycle (α,A())(\alpha,A(\cdot)) is not Cs+1C^{s+1}-reducible, proceeding by contradiction. Suppose that there exists B1()Cs+1(2𝕋d,SL(2,))B_{1}(\cdot)\in C^{s+1}(2{\mathbb{T}}^{d},\mathrm{SL}(2,{\mathbb{R}})) such that

(6.13) B11(+α)A()B1()=B~1,\displaystyle B_{1}^{-1}(\cdot+\alpha)A(\cdot)B_{1}(\cdot)=\tilde{B}_{1},

where B~1SL(2,)\tilde{B}_{1}\in\mathrm{SL}(2,{\mathbb{R}}). Then, B~1\tilde{B}_{1} is diagonalizable. Otherwise B~1\tilde{B}_{1} can be conjugate to a Jordan block. Then the A(;n)A(\cdot;n) has linear growth on nn which contradicts equation (6.8). Combining equations (6.8) and (6.13) we have

(6.14) B2(+α)=A~B2()B~11,\displaystyle B_{2}(\cdot+\alpha)=\tilde{A}B_{2}(\cdot)\tilde{B}^{-1}_{1},

where B2=B1B1B_{2}=B^{-1}B_{1}. Define the linear operator on M(2,)\mathrm{M}(2,{\mathbb{C}}):

L(Y):=A~YB~11,L(Y):=\tilde{A}Y\tilde{B}^{-1}_{1},

where YM(2,)Y\in\mathrm{M}(2,{\mathbb{C}}). Applying the Fourier transform to eq. (6.14), we have that for every kdk\in{\mathbb{Z}}^{d}

ek(α2)B2^(k2)=L(B2^(k2)).e_{k}(\frac{\alpha}{2})\widehat{B_{2}}(\frac{k}{2})=L(\widehat{B_{2}}(\frac{k}{2})).

Since B2()0B_{2}(\cdot)\neq 0 there exists 2l0d2l_{0}\in{\mathbb{Z}}^{d} such that B2^(l0)0\widehat{B_{2}}(l_{0})\neq 0. Thus el0(α)e_{l_{0}}(\alpha) is an eigenvalue of LL. Therefore, the two eigenvalus of B~1\tilde{B}_{1} are {exp{2πi(l0,α+ρ)},exp{2πi(l0,α+ρ)}}\{\exp\{2\pi i(\langle l_{0},\alpha\rangle+\rho)\},\exp\{-2\pi i(\langle l_{0},\alpha\rangle+\rho)\}\} ore {exp{2πi(ρl0,α)},exp{2πi(l0,αρ)}}\{\exp\{2\pi i(\rho-\langle l_{0},\alpha\rangle)\},\exp\{2\pi i(\langle l_{0},\alpha\rangle-\rho)\}\}. Without loss of generality, we assume the eigenvalues to be the former. Since B~1\tilde{B}_{1} is diagonalizable there exists D~SL(2,)\tilde{D}\in\mathrm{SL}(2,{\mathbb{C}}) such that

D~1B~1D~=diag{exp{2πi(l0,α+ρ)},exp{2πi(l0,α+ρ)}}.\tilde{D}^{-1}\tilde{B}_{1}\tilde{D}=\mathrm{diag}\{\exp\{2\pi i(\langle l_{0},\alpha\rangle+\rho)\},\exp\{-2\pi i(\langle l_{0},\alpha\rangle+\rho)\}\}.

Let B3:=B1D~(el0()00el0())B_{3}:=B_{1}\tilde{D}\begin{pmatrix}e_{l_{0}}(\cdot)&0\\ 0&e_{-l_{0}}(\cdot)\end{pmatrix}. We have

(6.15) B31(+α)A()B3()=A~.\displaystyle B_{3}^{-1}(\cdot+\alpha)A(\cdot)B_{3}(\cdot)=\tilde{A}.

Then there are two cases:

𝐂𝐚𝐬𝐞𝐈\mathbf{Case\ I}: For all 2kd2k\in{\mathbb{Z}}^{d} where k0k\neq 0, we have 2ρk,αmod2\rho\neq\langle k,\alpha\rangle\operatorname{mod}{\mathbb{Z}}.

In this case, combining (6.8) and (6.15)\eqref{b4r} we have

B4(+α)=A~B4()A~1,B_{4}(\cdot+\alpha)=\tilde{A}B_{4}(\cdot)\tilde{A}^{-1},

where B4=B1B3B_{4}=B^{-1}B_{3}. In the frequency domain, this implies that for every 2kd2k\in{\mathbb{Z}}^{d} we have

ek(α)B4^(k)=A~B4^(k)A~1.e_{k}(\alpha)\widehat{B_{4}}(k)=\tilde{A}\widehat{B_{4}}(k)\tilde{A}^{-1}.

However, the eigenvalues of the operator L(Y):=A~YA~1L^{\prime}(Y):=\tilde{A}Y\tilde{A}^{-1} are {1,1,e4πiρ,e4πiρ}\{1,1,e^{4\pi i\rho},e^{-4\pi i\rho}\}, and since2ρk,α2\rho\neq\langle k,\alpha\rangle this implies that B4constantB_{4}\equiv\mathrm{constant}. Thus we have

B=B1D~(el0()00el0())B41.B=B_{1}\tilde{D}\begin{pmatrix}e_{l_{0}}(\cdot)&0\\ 0&e_{-l_{0}}(\cdot)\end{pmatrix}B_{4}^{-1}.

Due to D~,B4SL(2,)\tilde{D},B_{4}\in\mathrm{SL}(2,{\mathbb{C}}) and B1s+1<\|B_{1}\|_{s+1}<\infty, we have Bs+1<\|B\|_{s+1}<\infty which contradicts eq. (6.9).

𝐂𝐚𝐬𝐞𝐈𝐈\mathbf{Case\ II}: There exists 2l1d2l_{1}\in{\mathbb{Z}}^{d}, l10l_{1}\neq 0, such that 2ρ=l1,αmod2\rho=\langle l_{1},\alpha\rangle\operatorname{mod}{\mathbb{Z}}.

In this case, we just need to set B=B(el1(2)00el1(2))B^{\prime}=B\begin{pmatrix}e_{l_{1}}(\frac{\cdot}{2})&0\\ 0&e_{-l_{1}}(\frac{\cdot}{2})\end{pmatrix}. Then we have

B1(+α)A()B()=Id.B^{\prime-1}(\cdot+\alpha)A(\cdot)B^{\prime}(\cdot)=\mathrm{Id}.

This situation has been reduced 𝐂𝐚𝐬𝐞𝐈\mathbf{Case\ I}, which concludes the proof. ∎

Proof of Theorem 1.4: The proof is as for Theorem 1.1, by replacing Proposition 4.1 by Proposition 6.1. ∎

Proof of Theorem 1.3:

By [24, 47], for any αDCd\alpha\in\mathrm{DC}_{d}, 0<h<h0<h_{*}<h, there exist ε0=ε0(α,h,h)\varepsilon_{0}=\varepsilon_{0}(\alpha,h,h_{*}), such that if Vhε0\|V\|_{h}\leq\varepsilon_{0}, EΣV,αE\in\Sigma_{V,\alpha}, then (α,SEV)𝒜α\𝒰α(\alpha,S_{E}^{V})\in\mathcal{AR}_{\alpha}\backslash\mathcal{UH}_{\alpha}. By Theorem 1.4, for any s+s\in{\mathbb{Z}}^{+}, for any η>0\eta>0, there exists (α,A)𝒜α\𝒰α(\alpha,A^{\prime})\in\mathcal{AR}_{\alpha}\backslash\mathcal{UH}_{\alpha} with A()SEV()hη\|A^{\prime}(\cdot)-S_{E}^{V}(\cdot)\|_{h_{*}}\leq\eta such that

B(+α)1A()B()=(e2πiρ00e2πiρ),\displaystyle B(\cdot+\alpha)^{-1}A^{\prime}(\cdot)B(\cdot)=\begin{pmatrix}e^{2\pi i\rho}&0\\ 0&e^{-2\pi i\rho}\end{pmatrix},

Moreover, B()s<\|B(\cdot)\|_{s}<\infty while B()s+1=.\|B(\cdot)\|_{s+1}=\infty. By Lemma 5.2, there exist V~Chω(𝕋d,)\widetilde{V^{\prime}}\in C_{h_{*}}^{\omega}({\mathbb{T}}^{d},{\mathbb{R}}) with |V~|h<η|\widetilde{V^{\prime}}|_{h_{*}}<\eta and ZChω(𝕋d,PSL(2,))Z^{\prime}\in C_{h_{*}}^{\omega}({\mathbb{T}}^{d},{\rm PSL}(2,{\mathbb{R}})) such that

Z(+α)1A()Z()=SEV~().Z^{\prime}(\cdot+\alpha)^{-1}A^{\prime}(\cdot)Z^{\prime}(\cdot)=S_{E}^{\widetilde{V^{\prime}}}(\cdot).

Let B=ZBB^{\prime}=Z^{\prime}B, then

B(θ+α)1SEV~(θ)B(θ)=(e2πiρ00e2πiρ),B^{\prime}(\theta+\alpha)^{-1}S_{E}^{\widetilde{V^{\prime}}}(\theta)B^{\prime}(\theta)=\begin{pmatrix}e^{2\pi i\rho}&0\\ 0&e^{-2\pi i\rho}\end{pmatrix},

and write B(θ)=(z11(θ)z12(θ)z21(θ)z22(θ)),B^{\prime}(\theta)=\left(\begin{array}[]{ccc}z_{11}(\theta)&z_{12}(\theta)\cr z_{21}(\theta)&z_{22}(\theta)\end{array}\right), then we have

(6.16) (EV~k(θ))z11(θ)=z11(θα)e2πiρ+z11(θ+α)e2πiρ.\displaystyle(E-\widetilde{V}_{k}(\theta))z_{11}(\theta)=z_{11}(\theta-\alpha)e^{-2\pi i\rho}+z_{11}(\theta+\alpha)e^{2\pi i\rho}.

Applying the Fourier transformatiomation to eq. (6.16)(\ref{block-red}), we get

mdV~(m)z^11(nm)+2cos2π(ρ+nα)z^11(n)=Ez^11(n),\displaystyle\sum_{m\in{\mathbb{Z}}^{d}}\widetilde{V^{\prime}}(m)\widehat{z}_{11}(n-m)+2\cos 2\pi(\rho+n\alpha)\widehat{z}_{11}(n)=E\widehat{z}_{11}(n),

i.e., LV~,ρ,αz^11=Ez^11L_{\widetilde{V^{\prime}},\rho,\alpha}\widehat{z}_{11}=E\widehat{z}_{11}, moreover, since B()s<\|B^{\prime}(\cdot)\|_{s}<\infty while B()s+1=.\|B^{\prime}(\cdot)\|_{s+1}=\infty. then the eigenfunction {z^11(n)}nhs\hs+1\{\widehat{z}_{11}(n)\}_{n\in{\mathbb{Z}}}\in h^{s}\backslash h^{s+1}.

7. Appendix: Proof of Proposition 4.2:

We follow the proof of the proposition as given in [47, 22]. An alternative proof can be obtained by use of the KAM normal form, following the method introduced in [41]. We need the following result, proved in [47, 22].

Proposition 7.1.

Let αDCd(γ,τ)\alpha\in DC_{d}(\gamma,\tau), σ>0\sigma>0. Suppose that ASL(2,)A\in SL(2,{\mathbb{R}}), fChω(𝕋d,sl(2,))f\in C^{\omega}_{h}({\mathbb{T}}^{d},{\mathrm{sl}}(2,{\mathbb{R}})). Then for any h+<hh_{+}<h, there exists constants C0C_{0} and D0=D0(κ,τ,d)D_{0}=D_{0}(\kappa^{\prime},\tau,d) such that if

(7.1) fhϵD0AC0(min{1,1h}(hh+))C0τ,\displaystyle\|f\|_{h}\leq\epsilon\leq\frac{D_{0}}{\|A\|^{C_{0}}}(\min\{1,\frac{1}{h}\}(h-h_{+}))^{C_{0}\tau},

then there exists BCh+ω(2𝕋d,SL(2,))B\in C_{h_{+}}^{\omega}(2{\mathbb{T}}^{d},SL(2,{\mathbb{R}})), A+SL(2,)A_{+}\in SL(2,{\mathbb{R}}) and f+Ch+ω(𝕋d,sl(2,))f_{+}\in C_{h_{+}}^{\omega}({\mathbb{T}}^{d},{\mathrm{sl}}(2,\\ {\mathbb{R}})) such that

B1(θ+α)Aef(θ)B(θ)=A+ef+(θ).B^{-1}(\theta+\alpha)Ae^{f(\theta)}B(\theta)=A_{+}e^{f_{+}(\theta)}.

More precisely, letting spec(A)={e2πiξ,e2πiξ}spec(A)=\{e^{2\pi i\xi},e^{-2\pi i\xi}\}, N=2hh+|lnϵ|N=\frac{2}{h-h_{+}}|\ln\epsilon|, we can distinguish two cases:

  • (Non-resonant case) if for any ndn\in{\mathbb{Z}}^{d} with 0<|n|N0<|n|\leq N, we have

    2ξ<n,α>/ϵ115,\|2\xi-<n,\alpha>\|_{{\mathbb{R}}/{\mathbb{Z}}}\geq\epsilon^{\frac{1}{15}},

    then

    Bidh+ϵ12,f+h+ϵ2.\|B-id\|_{h_{+}}\leq\epsilon^{\frac{1}{2}},\quad\|f_{+}\|_{h_{+}}\leq\epsilon^{2}.

    Moreover, A+A<2ϵ\|A_{+}-A\|<2\epsilon.

  • (Resonant case) if there exists nn_{\ast} with 0<|n|N0<|n_{\ast}|\leq N such that

    2ξ<n,α>/<ϵ115,\|2\xi-<n_{\ast},\alpha>\|_{{\mathbb{R}}/{\mathbb{Z}}}<\epsilon^{\frac{1}{15}},

    then

    Bh+|lnϵ|τϵh+hh+,f+h+<ϵeh+ϵ118τ.\|B\|_{h_{+}}\leq|\ln\epsilon|^{\tau}\epsilon^{-\frac{h_{+}}{h-h_{+}}},\ \|f_{+}\|_{h_{+}}<\epsilon e^{-h_{+}\epsilon^{-\frac{1}{18\tau}}}.

    Moreover, A+A_{+} can be written as A+=eA~+A_{+}=e^{\tilde{A}_{+}} with |A~+|ϵ116|\tilde{A}_{+}|\leq\epsilon^{\frac{1}{16}}.

The proof of proposition 4.2 follows by iteration of the proposition here above. Consider the initial cocycle (α,A0ef0(θ))(\alpha,A_{0}e^{f_{0}(\theta)}), where A0SL(2,)A_{0}\in SL(2,{\mathbb{R}}), f0Chω(𝕋d,sl(2,))f_{0}\in C_{h}^{\omega}({\mathbb{T}}^{d},{\mathrm{sl}}(2,{\mathbb{R}})). Without loss of generality, assume that h<1h<1, as well as that

f0hϵD0A0C0(hh~8)C0τ,\displaystyle\|f_{0}\|_{h}\leq\epsilon_{*}\leq\frac{D_{0}}{\|A_{0}\|^{C_{0}}}(\frac{h-\tilde{h}}{8})^{C_{0}\tau},

where D0=D0(κ,τ,d)D_{0}=D_{0}(\kappa^{\prime},\tau,d) is the constant defined in Proposition 7.1. Then we can define the sequence inductively. Let ϵ0=ϵ\epsilon_{0}=\epsilon_{*}, h0=hh_{0}=h, and assume that we are at the (j+1)th(j+1)^{th} KAM step, i.e. we have already constructed BjChjω(𝕋d,PSL(2,))B_{j}\in C^{\omega}_{h_{j}}({\mathbb{T}}^{d},PSL(2,{\mathbb{R}})) such that

Bj1(θ+α)A0ef0(θ)Bj(θ)=Ajefj(θ),B_{j}^{-1}(\theta+\alpha)A_{0}e^{f_{0}(\theta)}B_{j}(\theta)=A_{j}e^{f_{j}(\theta)},

where AjSL(2,)A_{j}\in SL(2,{\mathbb{R}}) has eigenvalues e±iξje^{\pm i\xi_{j}} and

Bjhjϵjhh~4h~,fjhjϵj\|B_{j}\|_{h_{j}}\leq\epsilon_{j}^{-\frac{h-\tilde{h}}{4\tilde{h}}},\qquad\|f_{j}\|_{h_{j}}\leq\epsilon_{j}

for some ϵjϵ02j\epsilon_{j}\leq\epsilon_{0}^{2^{j}}, and define

hjhj+1=hh+h~24j+1,Nj=2|lnϵj|hjhj+1.h_{j}-h_{j+1}=\frac{h-\frac{h+\tilde{h}}{2}}{4^{j+1}},\ \ N_{j}=\frac{2|\ln\epsilon_{j}|}{h_{j}-h_{j+1}}.

By our choice of ϵ0\epsilon_{0}, one can check that

(7.2) ϵjD0AjC0(hjhj+1)C0τ.\epsilon_{j}\leq\frac{D_{0}}{\|A_{j}\|^{C_{0}}}(h_{j}-h_{j+1})^{C_{0}\tau}.

Indeed, ϵj\epsilon_{j} on the left side of the inequality decays at least super-exponentially with jj, while (hjhj+1)C0τ(h_{j}-h_{j+1})^{C_{0}\tau} on the right side decays exponentially with jj.

Note that (7.2)(\ref{iter}) implies that Proposition 7.1 can be applied iteratively, consequently one can construct

B¯jChj+1ω(𝕋d,PSL(2,)),Aj+1SL(2,),fj+1Chj+1(𝕋d,sl(2,))\bar{B}_{j}\in C^{\omega}_{h_{j+1}}({\mathbb{T}}^{d},PSL(2,{\mathbb{R}})),\ \ A_{j+1}\in SL(2,{\mathbb{R}}),\ \ f_{j+1}\in C_{h_{j+1}}({\mathbb{T}}^{d},sl(2,{\mathbb{R}}))

such that

B¯j1(θ+α)Ajefj(θ)B¯j(θ)=Aj+1efj+1(θ).\bar{B}_{j}^{-1}(\theta+\alpha)A_{j}e^{f_{j}(\theta)}\bar{B}_{j}(\theta)=A_{j+1}e^{f_{j+1}(\theta)}.

More precisely, we can distinguish two cases:

Non-resonant case: If for any ndn\in{\mathbb{Z}}^{d} with 0<|n|Nj0<|n|\leq N_{j}, we have

2ξj<n,α>/ϵj115,\|2\xi_{j}-<n,\alpha>\|_{{\mathbb{R}}/{\mathbb{Z}}}\geq\epsilon_{j}^{\frac{1}{15}},

then

B¯jidhj+1ϵj12,fj+1hj+1ϵj2:=ϵj+1,Aj+1Aj2ϵj.\|\bar{B}_{j}-id\|_{h_{j+1}}\leq\epsilon_{j}^{\frac{1}{2}},\ \ \|f_{j+1}\|_{h_{j+1}}\leq\epsilon_{j}^{2}:=\epsilon_{j+1},\ \ \|A_{j+1}-A_{j}\|\leq 2\epsilon_{j}.

Let Bj+1=Bj(θ)B¯j(θ)B_{j+1}=B_{j}(\theta)\bar{B}_{j}(\theta), we have

Bj+11(θ+α)A0ef0(θ)Bj+1(θ)=Aj+1efj+1(θ),B_{j+1}^{-1}(\theta+\alpha)A_{0}e^{f_{0}(\theta)}B_{j+1}(\theta)=A_{j+1}e^{f_{j+1}(\theta)},

with estimate

Bj+1hj+12ϵjhh~4h~ϵj+1hh~4h~.\|B_{j+1}\|_{h_{j+1}}\leq 2\epsilon_{j}^{-\frac{h-\tilde{h}}{4\tilde{h}}}\leq\epsilon_{j+1}^{-\frac{h-\tilde{h}}{4\tilde{h}}}.

Resonant case: If there exists njn_{j} with 0<|nj|Nj0<|n_{j}|\leq N_{j} such that

2ξj<nj,α>/<ϵj115,\|2\xi_{j}-<n_{j},\alpha>\|_{{\mathbb{R}}/{\mathbb{Z}}}<\epsilon_{j}^{\frac{1}{15}},

with estimate

B¯jhj+1|lnϵj|τϵjhj+1hjhj+1,fj+1hj+1ϵjehj+1εj118τ:=ϵj+1.\|\bar{B}_{j}\|_{h_{j+1}}\leq|\ln\epsilon_{j}|^{\tau}\epsilon_{j}^{-\frac{h_{j+1}}{h_{j}-h_{j+1}}},\ \ \|f_{j+1}\|_{h_{j+1}}\leq\epsilon_{j}e^{-h_{j+1}\varepsilon_{j}^{-\frac{1}{18\tau}}}:=\epsilon_{j+1}.

Moreover, we can write Aj+1=eA~jA_{j+1}=e^{\tilde{A}_{j}} with estimate

|A~j|<2ϵj115.|\tilde{A}_{j}|<2\epsilon_{j}^{\frac{1}{15}}.

Let Bj+1(θ)=Bj(θ)B¯j(θ)B_{j+1}(\theta)=B_{j}(\theta)\bar{B}_{j}(\theta), then we have

Bj+11(θ+α)A0ef0(θ)Bj+1(θ)=Aj+1efj+1(θ),B_{j+1}^{-1}(\theta+\alpha)A_{0}e^{f_{0}(\theta)}B_{j+1}(\theta)=A_{j+1}e^{f_{j+1}(\theta)},

with

Bj+1hj+1\displaystyle\|B_{j+1}\|_{h_{j+1}} ϵjhh~4h~|lnϵj|τϵjhj+1hjhj+1ϵj+1hh~4h~.\displaystyle\leq\epsilon_{j}^{-\frac{h-\tilde{h}}{4\tilde{h}}}|\ln\epsilon_{j}|^{\tau}\epsilon_{j}^{-\frac{h_{j+1}}{h_{j}-h_{j+1}}}\leq\epsilon_{j+1}^{-\frac{h-\tilde{h}}{4\tilde{h}}}.

The last inequality is possible since by our choise ϵj+1=ϵjehj+1εj118τ\epsilon_{j+1}=\epsilon_{j}e^{-h_{j+1}\varepsilon_{j}^{-\frac{1}{18\tau}}}.

Acknowledgements

N. Karaliolios was partially supported by LABEX CEMPI (ANR-11-LABX-0007-01) while a post-doc at Université de Lille. He is grateful to his co-authors for their warm hospitality at the Chern Institute. Q. Zhou was partially supported by support by NSFC grant (11671192,11771077), The Science Fund for Distinguished Young Scholars of Tianjin (No. 19JCJQJC61300) and Nankai Zhide Foundation.

References

  • [1] V. I. Arnold. Small Denominators I; on the Mapping of a Circle into itself Isvestijia Akad. Nauk, Math series. 25 1 (1961), 21-96.
  • [2] A. Avila. Global theory of one-frequency Schródinger operators. Acta Math. 215 (2015), 1-54.
  • [3] A. Avila. The absolutely continuous spectrum of the almost Mathieu operator. arXiv:0810.2965
  • [4] A. Avila. KAM, Lyapunov exponents and the spectral dichotomy for one-frequency Schrödinger operators. In preparation.
  • [5] A. Avila. Almost reducibility and absolute continuity, arXiv:1006.0704
  • [6] A. Avila; S. Jitomirskaya, The Ten Martini Problem, Ann. Math.170, 303-342 (2009).
  • [7] A. Avila; S. Jitomirskaya. Almost localization and almost reducibility. J. Eur. Math. Soc. 12 (2010), 93–131
  • [8] A. Avila; S. Jitomirskaya. Hölder continuity of absolutely continuous spectral measures for one-frequency Schrödinger operators. Comm. Math. Phys. 301 (2011), no. 2, 563–581.
  • [9] A. Avila; S. Jitomirskaya; C. Sadel. Complex one-frequency cocycles. J. Eur. Math. Soc. 16(9) (2013), 1915–1935.
  • [10] A. Avila; S. Jitomirskaya; Q. Zhou. Second Phase transition line. Math. Ann. 370 (2016).
  • [11] A.Avila; Y. Last; M. Shamis; Q. Zhou. On the abominable properties of the Almost Mathieu operator with well approximated frequencies. In preparation.
  • [12] A. Avila; R. Krikorian. Reducibility or non-uniform hyperbolicity for quasi-periodic Schrödinger cocycles, Ann. Math. 164, 911-940 (2006).
  • [13] A. Avila; J. You; Q. Zhou. Sharp phase transitions for the almost Mathieu operator. Duke Math. J. 166(14) (2017), 2697-2718.
  • [14] D. V. Anosov; A. B. Katok. New examples in smooth ergodic theory. Ergodic diffeomorphisms. Trans. Moscow Math. Soc. 23 (1970), 1-35.
  • [15] S. H. Amor. Hölder continuity of the rotation number for the quasi-periodic co-cycles in SL(2,)\mathrm{SL}(2,{\mathbb{R}}). Commun. Math. Phys. 287 (2009), 565–588
  • [16] J. Bochi, Discontinuity of the Lyapunov exponent for non-hyperbolic cocycles. Unpublished. (1999)
  • [17] J. Bochi, Genericity of zero Lyapunov exponents. Ergodic Theory Dyn. Syst. 22(6), 1667-1696 (2002)
  • [18] J. Bourgain. Positivity and continuity of the Lyapunov exponent for shifts on 𝕋d{\mathbb{T}}^{d} with arbitrary frequency vector and real analytic potential. J. Anal. Math. 96, 313-355 (2005)
  • [19] J. Bourgain. Green’s function estimates for lattice Schrödinger operators and applications. Annals of Mathematics Studies. Princeton University Press, Princeton, NJ. 158 (2005).
  • [20] J. Bourgain; S. Jitomirskaya. Absolutely continuous spectrum for 1D quasi-periodic operators. Invent. Math. 148 (2002), 453-463.
  • [21] J. Bourgain; S. Jitomirskaya. Continuity of the Lyapunov exponent for quasiperiodic operators with analytic potential. J. Stat. Phys. 108 (2002), 1203–1218
  • [22] A. Cai; C. Chavaudret; J. You; Q. Zhou, Sharp Hölder continuity of the Lyapunov exponent of finitely differentiable quasi-periodic cocycles. Math. Z. 291, 931–958 (2019).
  • [23] E. I. Dinaburg; Y. G. Sinai. The one-dimensional Schrödinger equation with quasiperiodic potential. Funk. Anal. i Prilozen. 9.4 (1975), 279-289.
  • [24] L.H. Eliasson, Floquet solutions for the 1-dimensional quasi-periodic Schrödinger equation. Commun. Math. Phys. 146 (1992), 447–482
  • [25] L.H. Eliasson; B. Fayad; R. Krikorian. Jean-Christophe Yoccoz and the theory of circle diffeomorphisms. arXiv:1810.07107.
  • [26] F. Evers; A. D. Mirlin, Anderson transitions. Rev. Mod. Phys. 80, 1355 (2008).
  • [27] A. Fathi; M.R. Herman. Existence de difféomorphismes minimaux. Dynamical systems, Vol. I-Warsaw. Astérisque. 49 (1977), 37-59.
  • [28] B, Fayad; A, Katok. Constructions in elliptic dynamics. Ergodic Theory Dyn. Syst. 24(5) (2005), 1477-1520.
  • [29] D. J. Gilbert; D. B. Pearson. On subordinacy and analysis of the spectrum of one-dimensional Schrödinger operators, J. Math. Anal. Appl. 128 (1987).
  • [30] M. Goldstein; W. Schlag. Hölder continuity of the integrated density of states for quasi-periodic Schrödinger equations and averages of shifts of subharmonic functions. Ann. Math. 154(1) (2001), 155–203
  • [31] M. Goldstein; W. Schlag. Fine properties of the integrated density of states and a quantitative separation property of the Dirichlet eigenvalues. Geom. Funct. Anal. 18(3) (2008), 755–869
  • [32] L. Ge; J. You; Q. Zhou. Exponential Dynamical Localization: Criterion and Applications. arXiv:1901.04258.
  • [33] L. Ge; J. You; Q. Zhou. Universal hierarchical structure of extended quasi-periodic eigenfunctions. In preparation.
  • [34] M. Herman. Sur la conjugaison différentiable des difféomorphismes du cercle à des rotations. Inst. Hautes Études Sci. Publ. Math. 49 (1979), 5-233
  • [35] S. Jitomirskaya; D.A. Koslover; M.S. Schulteis. Continuity of the Lyapunov exponent for analytic quasiperiodic cocycles. Ergodic Theory Dyn. Syst. 29 (2009), 1881–1905
  • [36] S.Jitomirskaya, I.Kachkovskiy. L2L^{2}-reducibility and localization for quasi-periodic operators, Math Res Lett, 23, 431-444 (2016).
  • [37] S. Jitomirskaya; Y. Last. Power-law subordinacy and singular spectra I. Half-line operators. Acta Math. 183 (1999), no. 2, 171-189.
  • [38] S. Jitomirskaya; W. Liu Universal hierarchical structure of quasiperiodic eigenfunctions. Ann. Math. 187(3) (2016).
  • [39] S. Jitormiskya; W. Liu. Universal reflective-hierarchical structure of quasi-periodic eigenfunctions and sharp spectral transition in phase. arXiv: 1802.00781v1.
  • [40] S. Jitomirskaya; B. Simon, Operators with singular continuous spectrum. III. Almost periodic Schrödinger operators. Comm. Math. Phys. 165 no. 1, 201-205 (1994)
  • [41] N. Karaliolios,Invariant distributions for quasiperiodic cocycles in 𝕋d×SU(2){\mathbb{T}}^{d}\times SU(2), 2014. arXiv 1407.4763.
  • [42] N. Karaliolios, Differentiable rigidity for quasiperiodic cocycles in compact Lie groups, J. Mod. Dyn. 11 (2017), 125–142.
  • [43] N. Karaliolios, Continuous spectrum or measurable reducibility for quasiperiodic cocycles in 𝕋d×SU(2){\mathbb{T}}^{d}\times SU(2), Comm. Math. Phys. 358 (2018), 741–766.
  • [44] A. Katok; B. Hasselblatt. Introduction to the Modern Theory of Dynamical Systems. Encyclopedia of mathematics and its applications, 4 (1995), 27-29.
  • [45] S. Klein. Localization for quasiperiodic Schrödinger operators with multivariable Gevrey potential functions. J. Spectr. Theory. 4, 1-53 (2014)
  • [46] Y. Last. Spectral theory of Sturm-Liouville operators on infinite intervals: a review of recent developments. In Sturm-Liouville Theory. Birkhauser Basel. 99-120. (2005)
  • [47] M. Leguil; J. You; Z. Zhao; Q. Zhou. Asymptotics of spectral gaps of quasi-periodic Schrödinger operators. arXiv:1712.04700.
  • [48] Puig, J.; A nonperturbative Eliasson’s reducibility theorem, Nonlinearity, 19(2), 355–376 (2006).
  • [49] B. Simon, Almost periodic Schrödinger operators: A review, Adv. Appl. Math., 3, 463-490 (1982).
  • [50] M. Viana,Lectures on Lyapunov exponents. Cambridge Studies in Advanced Mathematics, Cambridge University Press. 145, (2014)
  • [51] Y. Wang; J. You. Examples of discontinuity of Lyapunov exponent in smooth quasi-periodic cocycles. Duke Math. J. 162 (2013), 2363–2412
  • [52] Y. Wang; Z. Zhang, Uniform positivity and continuity of Lyapunov exponents for a class of C2C^{2} quasiperiodic Schrödinger cocycles. J. Func. Anal.268, 2525-2585 (2015)
  • [53] J.Xu; L. Ge; Y. Wang. The Hölder continuity of Lyapunov exponents for a class of Cos-type quasi-periodic Schrödinger cocycles arXiv:2006.03381.
  • [54] J.-C. Yoccoz. Conjugaison différentiable des difféomorphismes du cercle dont le nombre de rotation vérifie une condition diophantienne. Ann. Sci. École Norm. Sup. 17(4) (1984), 333-359
  • [55] J.-C. Yoccoz. Some questions and remarks about SL(2,)SL(2,{\mathbb{R}}) cocycles. In Modern Dynamical Systems and Applications, 447-458. Cambridge University Press, Cambridge, 2004.
  • [56] J. You; S. Zhang; Q. Zhou. Point spectrum for quasi-periodic long range operators. Journal of Spectral Theory, 4 (2015), pp.769-781.