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

Exact mobility edges for 1D quasiperiodic models

Yongjian Wang School of Mathematical Sciences, Laboratory of Mathematics and Complex Systems, MOE, Beijing Normal University, 100875 Beijing, China [email protected] Xu Xia Chern Institute of Mathematics and LPMC, Nankai University, Tianjin 300071, China [email protected] Jiangong You Chern Institute of Mathematics and LPMC, Nankai University, Tianjin 300071, China [email protected] ZuoHuan Zheng University of Chinese Academy of Sciences, Beijing 100049,China&Academy of Mathematics and Systems Science, Chinese Academy of Sciences, Beijing 100190, China&College of Mathematics and Statistics, Hainan Normal University, Haikou, Hainan 571158, China. [email protected]  and  Qi Zhou Chern Institute of Mathematics and LPMC, Nankai University, Tianjin 300071, China [email protected]
Abstract.

Mobility edges (ME), i.e. critical energies which separate absolutely continuous spectrum and purely point spectrum, is an important issue in quantum physics. So far there are two experimentally feasible 1D quasiperiodic models that have been discovered to have exact mobility edge. However, all the theoretical studies have remained at the numerical level. In this paper, we rigorously prove the existence and give the precise location of the MEs for these models.

1. Introduction

In his 1958 seminal article [3], Anderson argued that in one-dimensional or two-dimensional disordered systems, all states are localized at any disorder strengths. However, in a three-dimensional disordered system, a transition occurs at a finite disorder strength, i.e., there exists a critical energy EcE_{c} separating the localized states and the extended states. This kind of phenomenon became known as the Anderson metal–insulator transition, and the critical energy EcE_{c} was later termed the mobility edge (ME) by Mott. The idea of mobility edges would develop into one of the most studied concepts of condensed-matter physics. It has been the progenitor of many important problems in physics [35], and was one of the main reasons why Anderson and Mott shared the 1977 Nobel Prize in Physics.

The standard mathematical interpretation of Anderson transition is the following: the dd-dimensional (d3d\geq 3) random Schrödinger operator

H=Δ+V,H=-\Delta+V,

where V(n)V(n) is an independent identically distributed random variable with distribution uniformly in (λ,λ)(-\lambda,\lambda), has Anderson localization (pure point spectrum with exponentially decaying eigenfunctions) in the regime ±[Ec,2d+λ]\pm[E_{c},2d+\lambda], and absolutely continuous spectrum in the interval [Ec,Ec][-E_{c},E_{c}] for some EcE_{c}, if λ\lambda is small.

Over 40 years after Anderson-Mott’s Nobel Prize and 60 years after Anderson first proposed the theory, great progress has been made in understanding the corresponding physics, however experimental demonstration was notoriously difficult due to the problems in reliably controlling disorder in solid-state systems [19, 35]. On the other hand, the mathematical understanding of the whole picture is still unsatisfactory and one-sided: we know that if the coupling constant λ\lambda is large enough, the corresponding Schrödinger operator has Anderson localization [1, 37, 46]. But up to now, there are no rigorous results on the existence of the absolutely continuous spectrum for any random operators, not to mention the existence of ME. Indeed, this is such an important question that Simon [74] gave it as Problem 1 of a list of Schrödinger operator problems for the twenty-first century. One can consult [40] and the references therein for recent study on this subject.

The breakthrough came in the manipulation of ultra-cold atoms, which offer a completely new, well-controlled tool for directly observing ME [18, 71]. Consequently there is growing interest in exploring ME in 1D quasi-periodic models, especially exact ME to understand the extended-localized transition and to advance in-depth study of fundamental ME physics, e.g. to possibly eliminate the theoretical dispute on whether many-body MEs exist [72, 79]. However, finding experimentally realistic 1D quasi-periodic models with exact ME is difficult, and so far there are only two models in physics literature [41, 78]. In this paper, we rigorously prove ME for these two models.

Before introducing the models and our main results, let us first revisit the spectral results of the almost Mathieu operator (resp. Aubry-Andre model in physics literature):

(Hλ,α,θu)n=un+1+un1+2λcos2π(nα+θ)un,(H_{\lambda,\alpha,\theta}u)_{n}=u_{n+1}+u_{n-1}+2\lambda\cos 2\pi(n\alpha+\theta)u_{n},

where θ\theta\in\mathbb{R} is the phase, α\\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}} is the frequency, and λ\lambda\in{\mathbb{R}} is the coupling constant. The almost Mathieu operator (AMO) is the central quasi-periodic model, not only because of its importance in physics [16], but also as a fascinating mathematical object. It was first introduced by Peierls [70], as a model for an electron on a 2D lattice, acted on by a homogeneous magnetic field [47], and it plays a central role in the Thouless et al. theory of the integer quantum Hall effect [76]. We recall that α\alpha is Diophantine (denoted by DC(γ,σ)DC(\gamma,\sigma)), if there exist γ,σ>0\gamma,\sigma>0, such that

kα/γ|k|σk0.\|k\alpha\|_{{\mathbb{R}}/{\mathbb{Z}}}\geq\frac{\gamma}{|k|^{\sigma}}\quad\forall k\neq 0.

We also denote DC=γ>0,σ>0DC(γ,σ)DC=\cup_{\gamma>0,\sigma>0}DC(\gamma,\sigma). It is well known that if αDC\alpha\in DC, then λ=1\lambda=1 is the transition line from absolutely continuous spectrum to Anderson localization [7, 10, 53]. However, one should note that if α\alpha is not Diophantine, then there exists a second transition line from singular continuous spectrum to Anderson localization [12, 14, 58], which is neglected in the physics references. In any case, one has found that ME does not exist for AMO; nevertheless, based on a vast body of numerical work (one may consult [19, 41, 50, 73] and the references therein), the physical intuition is that if the symmetry of the almost Mathieu operator is broken in some controlled way, then the transition point λ=1\lambda=1 modifies into a ME. Different from random models, ME of quasiperiodic models could be any point in an interval due the the existence of gaps [49].

1.1. ME for the Generalized Aubry-Andre model

Our first result concerns the Generalized Aubry-Andre (GAA) model:

(1) (HV1,α,θu)n=un+1+un1+2λcos2π(θ+nα)1τcos2π(θ+nα)un,(H_{V_{1},\alpha,\theta}u)_{n}=u_{n+1}+u_{n-1}+2\lambda\frac{\cos 2\pi(\theta+n\alpha)}{1-\tau\cos 2\pi(\theta+n\alpha)}u_{n},

where τ(1,1)\tau\in(-1,1). If τ=0\tau=0, it is exactly AMO, and in the limiting case τ=1\tau=-1, it is the unbounded operator with potential tan2(πθ)\tan^{2}(\pi\theta). This model was first introduced by Ganeshan-Pixley-Das Sarma [41], where they not only give numerical evidence, but also introduce a generalized duality symmetry and show that

(2) sgn(λ)τE=2(1|λ|)sgn(\lambda)\tau E=2(1-|\lambda|)

is the ME. However, one should note that the generalized duality is mathematically rigorous. In this paper, without invoking generalized duality, we rigorously show that (2) really defines the exact ME.

It is well known that for any almost-periodic Schrödinger operator with potential VV, its spectrum Σ(V)\Sigma(V) is a perfect and compact set independent of the phase θ\theta. Denote

E¯(V)=minEΣ(V),E¯(V)=maxEΣ(V),\underline{E}(V)=\min_{E}\Sigma(V),\qquad\overline{E}(V)=\max_{E}\Sigma(V),

then our precise result can be formulated as follows:

Theorem 1.1.

For any αDC\alpha\in DC, |τ|<1|\tau|<1, λτ>0\lambda\tau>0, we have the following:

  1. (1)

    If |λ|<1|τ|2E¯(V1)|\lambda|<1-\frac{|\tau|}{2}\overline{E}(V_{1}), then HV1,α,θH_{V_{1},\alpha,\theta} has purely absolutely continuous spectrum for every θ.\theta.

  2. (2)

    If |λ|>1|τ|2E¯(V1)|\lambda|>1-\frac{|\tau|}{2}\underline{E}(V_{1}), then HV1,α,θH_{V_{1},\alpha,\theta} has Anderson localization for almost every θ.\theta.

  3. (3)

    If 1|τ|2E¯(V1)<|λ|<1|τ|2E¯(V1)1-\frac{|\tau|}{2}\overline{E}(V_{1})<|\lambda|<1-\frac{|\tau|}{2}\underline{E}(V_{1}), then sgn(λ)τE=2(1|λ|)sgn(\lambda)\tau E=2(1-|\lambda|) is the ME. More precisely,

    • HV1,α,θH_{V_{1},\alpha,\theta} has purely absolutely continuous spectrum for every θ\theta in the set {E:sgn(λ)τE<2(1|λ|)}\{E:sgn(\lambda)\tau E<2(1-|\lambda|)\}.

    • HV1,α,θH_{V_{1},\alpha,\theta} has Anderson localization for almost every θ\theta in the set {E:sgn(λ)τE>2(1|λ|)}\{E:sgn(\lambda)\tau E>2(1-|\lambda|)\}.

Remark 1.2.

One can consult Corollary 3.11 for the case λτ<0\lambda\tau<0.

Refer to caption
Figure 1. ME of the GAA model.

Figure 1 gives a numerical picture of ME of the GAA model, where orange color corresponds to extended states, and blue color corresponds to localized states. The physical mechanism of ME was explained by Anderson-Mott; we highlight that using synthetic lattices of laser-coupled atomic momentum modes [2], the GAA model can be experimentally realized to host the exact ME defined by (2). Theorem 1.1 gives the rigious proof of the existence of ME 222Theorem 1.1 covers partial result of our preprint [77], which is not intended for publication., and now the picture of ME from physics to mathematics is complete.

1.2. ME for quasi-periodic Mosaic model

Recently, the following quasi-periodic mosaic model was proposed in [78]

(3) (HV2,α,θu)n=un+1+un1+Vθ(n)un,(H_{V_{2},\alpha,\theta}u)_{n}=u_{n+1}+u_{n-1}+V_{\theta}(n)u_{n},

where

Vθ(n)={2λcos2πθ,nκ,0,else,λ>0.\quad V_{\theta}(n)=\left\{\begin{matrix}2\lambda\cos 2\pi\theta,&n\in\kappa\mathbb{Z},\\ 0,&else,\end{matrix}\right.\quad\lambda>0.

This model certainly defines a family of almost-periodic Schrödinger operators. If κ=1\kappa=1, then one can reduce it to AMO. As pointed out in [78], the model is experimentally realizable using an optical Raman lattice, thus a true physical model. We will show that, different from AMO, the mosaic model (3) with κ2\kappa\geq 2 do have MEs and we also give exact and complete description of all mobility edges for κ=2,3\kappa=2,3.

Theorem 1.3.

Let λ0\lambda\neq 0, αDC\alpha\in DC. If κ=2\kappa=2, then we have the following:

  1. (1)

    If |λ|E¯(V2)<1|\lambda|\overline{E}(V_{2})<1, then HV2,α,θH_{V_{2},\alpha,\theta} has purely absolutely continuous spectrum for every θ.\theta.

  2. (2)

    If |λ|E¯(V2)>1|\lambda|\overline{E}(V_{2})>1, then ±1λ\pm\frac{1}{\lambda} are MEs. More precisely,

    • HV2,α,θH_{V_{2},\alpha,\theta} has purely absolutely continuous spectrum in Σ(V2)(1λ,1λ)\Sigma(V_{2})\cap(-\frac{1}{\lambda},\frac{1}{\lambda}) for every θ\theta.

    • HV2,α,θH_{V_{2},\alpha,\theta} has Anderson localization in Σ(V2)[1λ,1λ]c\Sigma(V_{2})\cap[-\frac{1}{\lambda},\frac{1}{\lambda}]^{c} for almost every θ\theta.

Refer to caption
Figure 2. ME of the quasi-periodic mosaic model.

Figure 2 gives a numerical picture of ME of the quasi-periodic mosaic model. As is clear from the picture, the localization starts from the edges of the spectrum, and as the coupling constant λ\lambda is increased, then we have mobility edges, which move towards the center of the spectrum. This kind of behavior is similar to that of 3D disordered systems [66]. However, our results really demonstrate a new phenomenon, which does not even appear in previous physics literature. That is, no matter how large the coupling constant is, ME always occur. By contrast, in the random models or the quasi-periodic models (with smooth potential), all the states are believed to be localized when λ\lambda is large enough [1, 23, 24, 25, 26, 37, 46].

Also from Figure 2(a), if κ=2\kappa=2, it is clear (3) has two mobility edges. In general, one can anticipate arbitrary many even numbers of ME (Figure 2(b) for κ=3\kappa=3). In case κ=3\kappa=3, and denote

Ec1=1+1λ,Ec2=11λ,E_{c}^{1}=\sqrt{1+\frac{1}{\lambda}},\qquad E_{c}^{2}=\sqrt{1-\frac{1}{\lambda}},

then the complete picture is the following:

Theorem 1.4.

Let λ0\lambda\neq 0, αDC\alpha\in DC. If κ=3\kappa=3, then we have the following:

  1. (1)

    If |λ|(E¯(V2)21)<1|\lambda|(\overline{E}(V_{2})^{2}-1)<1, then HV2,α,θH_{V_{2},\alpha,\theta} has purely absolutely continuous spectrum for every θ.\theta.

  2. (2)

    If 1E¯(V2)21<|λ|<1\frac{1}{\overline{E}(V_{2})^{2}-1}<|\lambda|<1, then ±Ec1\pm E_{c}^{1} are MEs. More precisely,

    • HV2,α,θH_{V_{2},\alpha,\theta} has purely absolutely continuous spectrum in Σ(V2)(Ec1,Ec1)\Sigma(V_{2})\cap(-E_{c}^{1},E_{c}^{1}) for every θ\theta.

    • HV2,α,θH_{V_{2},\alpha,\theta} has Anderson localization in Σ(V2)[Ec1,Ec1]c\Sigma(V_{2})\cap[-E_{c}^{1},E_{c}^{1}]^{c} for almost every θ\theta.

  3. (3)

    If |λ|>1|\lambda|>1, then ±Ec1\pm E_{c}^{1}, ±Ec2\pm E_{c}^{2} are MEs. More precisely,

    • HV2,α,θH_{V_{2},\alpha,\theta} has purely absolutely continuous spectrum in Σ(V2)(Ec2,Ec1)\Sigma(V_{2})\cap(E_{c}^{2},E_{c}^{1}) and Σ(V2)(Ec1,Ec2)\Sigma(V_{2})\cap(-E_{c}^{1},E_{c}^{2}) for every θ\theta.

    • HV2,α,θH_{V_{2},\alpha,\theta} has Anderson localization in Σ(V2)[Ec1,Ec1]c\Sigma(V_{2})\cap[-E_{c}^{1},E_{c}^{1}]^{c} and Σ(V2)(Ec2,Ec2)\Sigma(V_{2})\cap(-E_{c}^{2},E_{c}^{2}) for almost every θ\theta.

Remark 1.5.

One can consult the result for general κ\kappa in Theorem 6.1.

1.3. Other models

The third model concerns the tight-binding model

(4) (H^V3,α,θx)n=jnep|nj|xj+λcos2π(nα+θ)xn,(\widehat{H}_{V_{3},\alpha,\theta}x)_{n}=\sum\limits_{j\neq n}e^{-p|n-j|}x_{j}+\lambda\cos 2\pi(n\alpha+\theta)x_{n},

with parameter p>0p>0. This is a quasi-periodic long-range operator acting on 2()\ell^{2}({\mathbb{Z}}). This quasi-periodic model was introduced by Biddle-Das Sarma in their groundbreaking work [19], where they predicted

E+1=cosh(p)|λ|E+1=\cosh(p)|\lambda|

is the exact energy dependent mobility edge, and this gives the first model which has exact ME in the physics literature. In this paper, we will actually show that the Aubry dual of (4) reduces to the GAA model, and as a consequence, we will rigorously show the following:

Corollary 1.6.

For any λ0\lambda\neq 0, αDC\alpha\in DC, the ME of H^V3,α,θ\widehat{H}_{V_{3},\alpha,\theta} takes place at

E+1=cosh(p)|λ|.E+1=\cosh(p)|\lambda|.
Remark 1.7.

One can consult the precise result at Corollary 3.13.

The final model is the Schrödinger operator with “Peaky” potential

(5) (HV4,α,θu)n=un+1+un1+λ1+4Ksin2π(θ+nα)un,K,λ>0,(H_{V_{4},\alpha,\theta}u)_{n}=u_{n+1}+u_{n-1}+\frac{\lambda}{1+4K\sin^{2}\pi(\theta+n\alpha)}u_{n},\quad K,\lambda>0,

which was first introduced by Bjerklöv and Krikorian [21]. Theorem B of [21] shows that for some sufficiently large KK and λ\lambda, there is a set 𝒜𝕋\mathcal{A}\subset{\mathbb{T}} of positive Lebesgue measure such that for any α𝒜\alpha\in\mathcal{A} the operator HV4,α,θH_{V_{4},\alpha,\theta} has both a.c. and p.p. components. In this paper, we will reveal exactly when this operator has ME and where is the ME.

Refer to caption
Figure 3. ME of Schrödinger operator with “Peaky” potential .
Corollary 1.8.

Let K>0K>0 and αDC\alpha\in DC. Then the following holds true:

  1. (1)

    If λK(2K+1)2<1KE¯(V4)2K+1\frac{\lambda K}{(2K+1)^{2}}<1-\frac{K\overline{E}(V_{4})}{2K+1}, then HV4,α,θH_{V_{4},\alpha,\theta} has purely absolutely continuous for every θ.\theta.

  2. (2)

    If λK(2K+1)2>1λE¯(V4)2K+1\frac{\lambda K}{(2K+1)^{2}}>1-\frac{\lambda\underline{E}(V_{4})}{2K+1}, then HV4,α,θH_{V_{4},\alpha,\theta} has Anderson localization for almost every θ.\theta.

  3. (3)

    If 1λE¯(V4)2K+1<λK(2K+1)2<1KE¯(V4)2K+11-\frac{\lambda\overline{E}(V_{4})}{2K+1}<\frac{\lambda K}{(2K+1)^{2}}<1-\frac{K\underline{E}(V_{4})}{2K+1}, then 2+1K2+\frac{1}{K} is the ME. More precisely,

    • HV4,α,θH_{V_{4},\alpha,\theta} has purely absolutely continuous spectrum in Σ(V4)[E¯(V4),2+1K)\Sigma(V_{4})\cap[\underline{E}(V_{4}),2+\frac{1}{K}) for every θ.\theta.

    • HV4,α,θH_{V_{4},\alpha,\theta} has Anderson localization in Σ(V4)(2+1K,E¯(V4)]\Sigma(V_{4})\cap(2+\frac{1}{K},\overline{E}(V_{4})] for almost every θ.\theta.

Remark 1.9.

We point out an interesting phenomenon, as is also clearly shown in Fig 3, that the ME of (5) doesn’t depend on the coupling constant λ\lambda. So one sees that KK gives the location of ME while λ\lambda determines whether ME will appear.

1.4. Coexistence of spectrums.

The coexistence of a.c. and p.p. spectrum is an active research subject which is related to and obviously weaker than exact ME. Bjerklöv [20] proved that if the potential is

V(θ)=exp(Kf(θ+α))+exp(Kf(θ)),V(\theta)=\exp(Kf(\theta+\alpha))+\exp(-Kf(\theta)),

where ff is assumed to be a non-constant real-analytic function with zero mean, then the Schrödinger operator has coexistence of regions of the spectrum with positive Lyapunov exponents and zero Lyapunov exponents, if KK is large enough. Elaborating on [20], Zhang [80] gives examples of the coexistence of a.c. and p.p. spectrum and coexistence of a.c. and s.c. spectrum. Bjerlöv and Krikorian [21] constructed a class of “peaky” potentials, such that the operator has coexistence of a.c. and p.p. spectrum. Avila [5] constructed examples of potentials which are real analytic perturbations of critical AMO, and for which the spectrum of the corresponding Schrödinger operator has both a.c. and p.p. components. For previous coexistence results on quasi-periodic potentials with two frequencies and almost periodic potential, one can consult [22, 36].

An effective method for proving coexistence of spectrum is studying the Lyapunov exponent of the Schrödinger cocycles333One can consult Section 2.2 for its definition associated to Schrödinger operators. This is a family of skew-products

(α,SEV):𝕋×2,(α,SEV)(θ,v)=(θ+α,SEV(θ)v)(\alpha,S_{E}^{V}):{\mathbb{T}}\times{\mathbb{R}}^{2}\circlearrowleft,\qquad(\alpha,S_{E}^{V})(\theta,v)=(\theta+\alpha,S_{E}^{V}(\theta)\cdot v)

where

SEV()=(EV()110).S_{E}^{V}(\cdot)=\begin{pmatrix}E-V(\cdot)&-1\\ 1&0\end{pmatrix}.

More precisely, coexistence of zero and positive Lyapunov exponents in the spectrum roughly implies the coexistence of a.c. and p.p. spectrum, since by the well-known Kotani’s theory [64], Σac(V)\Sigma_{ac}(V) is the essential support of the energies which have zero Lyapunov exponent, and it is a commonly used fact in physics literature that positive Lyapunov exponent implies localization.

1.5. Main ingredients of the proof

We stress that the above mentioned results [20, 21, 80] only give coexistence results, i.e. partial information on the spectrum, while ME requires complete information on the spectrum. For this purpose, we need to use the remarkable global theory of one-frequency analytic cocycles by Avila [5], where he establishes and gives classification of all SL(2,)SL(2,{\mathbb{C}}) cocycles. To be precise, cocycles that are not uniformly hyperbolic are classified in three regimes:

  1. (1)

    Subcritical, if there exists δ>0\delta>0 such that L(α,A(z))=0L(\alpha,A(z))=0 through some strip |z|δ|\Im z|\leq\delta,

  2. (2)

    Supercritical, or nonuniformly hyperbolic, if L(α,A)>0L(\alpha,A)>0,

  3. (3)

    Critical otherwise.

In the subcritical regime, the energy is related with extended states, while in the supercritical regime, the energy is related with localized states. To study ME, there are three key steps. The first is to find out the exact formula of the Lyapunov exponent L(α,SEV)L(\alpha,S_{E}^{V}) in the spectrum, which allows one to locate the zero Lyapunov exponent regime and positive Lyapunov exponent regime. Then one needs to prove a.c. spectrum in the subcritical regime, and prove localization in the supercritical regime.

Calculation of Lyapunov exponents. As we said, to obtain exact ME, the first step is to calculate the Lyapunov exponent. Based on the continuity of the Lyapunov exponent [28] and the Lyapunov exponent in the rational frequencies [65], Bourgain and Jitomirskaya [28] showed that if the energy belongs to the spectrum, then the Lyapunov exponent of AMO satisfies

(6) L(α,SE2λcos)=max{0,ln|λ|}.L(\alpha,S_{E}^{2\lambda\cos})=\max\{0,\ln|\lambda|\}.

However, this method can hardly be generalized. On the other hand, Avila’s global theory shows that, as a function of ϵ,\epsilon, the Lyapunov exponent L(α,SEV(+iϵ))L(\alpha,S_{E}^{V}(\cdot+i\epsilon)) is a convex, piecewise linear function, with integer slopes. Based on this fact, Avila [5] gives another proof of (6). In this paper, we will further generalize this argument, and calculate the Lyapunov exponent of the GAA model (Lemma 3.9) and quasi-periodic mosaic model (Lemma 3.1), and more importantly locate the subcritical and supercritical regime. Note that this method strongly depends on the fact that the acceleration of the Lyapunov exponent444Consult section 2.5 for its definition (the slope of the Lyapunov exponent) is not larger than 11, and this also explains why it is so difficult to find models with exact ME.

Absolutely continuous spectrum. Based on the KAM method, Dinaburg-Sinai [32] proved that if αDC\alpha\in DC, then Σac(λV)\Sigma_{ac}(\lambda V)\neq\varnothing in the perturbative small regime λ<λ0\lambda<\lambda_{0}. Here perturbative means that λ0\lambda_{0} depends on α\alpha through the Diophantine constants γ,σ\gamma,\sigma. Under the same assumption, Eliasson [33] showed that in fact the spectrum is purely absolutely continuous for any θ\theta. Specifically in the one-frequency case, one can even anticipate non-perturbative results. Making use of the specificity of one frequency, some new elaborate techniques have been developed to prove some sharp results. If αDC\alpha\in DC, based on non-perturbative Anderson localization results, Avila-Jitomirskaya [10] proved that there exists λ1\lambda_{1} which does not depend on α\alpha, such that Σ(λV)=Σac(λV)\Sigma(\lambda V)=\Sigma_{ac}(\lambda V) when λ<λ1\lambda<\lambda_{1}. Such a result was generalized by Avila to the weak Diophantine case [7]. Recently, Avila-Fayad-Krikorian [8] and Hou-You [51] independently developed non-standard KAM techniques, and showed that Σac(λV)\Sigma_{ac}(\lambda V)\neq\varnothing for λ<λ2(V)\lambda<\lambda_{2}(V) and for any irrational α\alpha. The breakthrough goes back to Avila, who established the deep relations between the existence of a.c. spectrum and the vanishing of the Lyapunov exponent. To be precise, his Almost Reducibility Conjecture (ARC) says that any subcritical cocycle is almost reducible, which furthermore supports a.c. spectrum. Our proof relies on the solution of ARC, as announced in [5], to appear in [4, 6]. ARC has many important dynamical and spectral consequences [4, 6, 11, 13, 14, 42, 67], indeed, it was already stated as Almost Reducibility Theorem (ART) in [11].

In our case, for the GAA model, one only needs to locate the subcritical regime, then one applies ARC directly to prove that the corresponding regime has pure a.c. spectrum. However, for the quasi-periodic mosaic model, the operator itself cannot induce a quasi-periodic Schrödinger cocycle. The observation here is that the iterates of the cocycle can be seen as a one-frequency analytic cocycle, thus one can locate the subcritical regime by Avila’s global theory, however ARC cannot apply directly, since an iterate of the cocycle does not define an operator any more. Here, we will develop a scheme to establish the link between absolutely continuous spectrum of almost periodic operators and almost reducibility of its iterated cocycle; the ideas first goes back to Avila [7], while the estimates are KAM based [29, 67]. One can found more discussions after Theorem 4.1 the difficulty and necessity for us to develop a general scheme. Indeed, such a scheme has already been used to study the purely a.c. spectrum of CMV matrices with small quasi-periodic Verblunsky coefficients [68].

Anderson localization. The above mentioned coexistence papers [5, 21, 80] all depend crucially on Bourgain-Goldstein’s result [25], where they prove that in the supercritical regime, for any fixed phase, HλV,α,θH_{\lambda V,\alpha,\theta} has AL for a.e.a.e. Diophantine frequency, i.e they have to remove a Hausdorff zero measure set of Diophantine frequencies. For the multi-frequency and multi-dimensional case, one can consult [23, 24, 26, 59, 52, 63] and the references therein. However, in physics applications, there is more interest in the case where α\alpha is a priori fixed as a Diophantine frequency. For localization results with fixed Diophantine frequency, if the potential is a cosine-like function, Fröhlich-Spencer-Wittwer [38] and Sinai [75] independently proved that for a.e. phase, HλV,α,θH_{\lambda V,\alpha,\theta} has AL for sufficiently large coupling constant. If the potential is analytic, Eliasson [34] proved that Hλ,α,θH_{\lambda,\alpha,\theta} has pure point spectrum for a.e. θ\theta and large enough λ.\lambda.

One can see that although these three localization results [34, 38, 75] hold for fixed Diophantine frequency, they are all perturbative, i.e. the coupling constant λ\lambda is assumed to be large enough. It is still open whether for non-constant analytic potentials and fixed Diophantine frequency, the operator HV,α,θH_{V,\alpha,\theta} has Anderson localization for a.e. θ\theta in the supercritical regime. To this stage, we should mention Jitomirskaya’s seminar paper [53], who not only proves Anderson localization result for the almost Mathieu operator, but also developed a non-perturbative localization approach which initiated other non-perturbative localization results (one may consult [10, 23, 24, 27, 25, 57, 58, 59, 60] and the reference therein). In this paper, we will further develop Jitomirskaya’s argument, and show that AL still holds for another family of analytic quasi-periodic Schrödinger operator in the whole supercritical regime.

2. Preliminaries

For a bounded analytic function ff defined on a strip {|z|<h}\{|\Im z|<h\}, let fh=sup|θ<h|f(θ)\mathop{||f||}_{h}=\sup_{|\Im\theta<h|}||f(\theta)|| and denote by Chω(𝕋,)C_{h}^{\omega}(\mathbb{T},*) the set of all these *-valued functions (* will usually denote \mathbb{R}, SL(2,)SL(2,\mathbb{R}), M(2,)M(2,\mathbb{C})). When θ\theta\in\mathbb{R}, we also set θ𝕋=infj|θj|||\theta||_{\mathbb{T}}=\inf_{j\in\mathbb{Z}}|\theta-j|.

2.1. Continued Fraction Expansion.

Let α(0,1)\alpha\in(0,1) be irrational, a0=0a_{0}=0 and b0=αb_{0}=\alpha. Inductively, for k1k\geq 1, we define

ak=bk11,bk=bk11ak,\quad a_{k}=\lfloor b_{k-1}^{-1}\rfloor,\ b_{k}=b_{k-1}^{-1}-a_{k},

Let p0=0p_{0}=0, p1=1p_{1}=1, q0=1q_{0}=1, q1=a1q_{1}=a_{1}. We define inductively pk=akpk1+pk1p_{k}=a_{k}p_{k-1}+p_{k-1}, qk=akqk1+qk2q_{k}=a_{k}q_{k-1}+q_{k-2}. Then (qn)n(q_{n})_{n} is the sequence of denominators of the best rational approximations of α\alpha, since we have kα𝕋qn1α𝕋||k\alpha||_{\mathbb{T}}\geq||q_{n-1}\alpha||_{\mathbb{T}}, 1k<qn\forall\ 1\leq k<q_{n}, and

12qn+1qnα𝕋1qn+1.\quad\frac{1}{2q_{n+1}}\leq||q_{n}\alpha||_{\mathbb{T}}\leq\frac{1}{q_{n+1}}.
Lemma 2.1.

[9] Let α\\alpha\in\mathbb{R}\backslash\mathbb{Q}, xx\in\mathbb{R} and 0l0qn10\leq l_{0}\leq q_{n}-1 be such that

|sinπ(x+l0α)|=inf0lqn1|sinπ(x+lα)|,|\sin\pi(x+l_{0}\alpha)|=\inf_{0\leq l\leq q_{n}-1}|\sin\pi(x+l\alpha)|,

then for some absolute constant C>0C>0,

Clnqn0lqn1,ll0ln|sinπ(x+lα)|+(qn1)ln2Clnqn.-C\ln q_{n}\leq\sum_{0\leq l\leq q_{n}-1,l\neq l_{0}}\ln|\sin\pi(x+l\alpha)|+(q_{n}-1)\ln 2\leq C\ln q_{n}.

2.2. Cocycle, Lyapunov exponent

Let XX be a compact metric space, (X,ν,T)(X,\nu,T) be ergodic. A cocycle (α,A)\×Cω(X,M(2,))(\alpha,A)\in{\mathbb{R}}\backslash{\mathbb{Q}}\times C^{\omega}(X,M(2,{\mathbb{R}})) is a linear skew product:

(T,A):\displaystyle(T,A): X×2X×2\displaystyle X\times{\mathbb{R}}^{2}\to X\times{\mathbb{R}}^{2}
(x,ϕ)(Tx,A(x)ϕ).\displaystyle(x,\phi)\mapsto(Tx,A(x)\cdot\phi).

For nn\in\mathbb{Z}, AnA_{n} is defined by (T,A)n=(Tn,An)(T,A)^{n}=(T^{n},A_{n}). Thus A0(x)=idA_{0}(x)=id,

An(x)=j=n10A(Tjx)=A(Tn1x)A(Tx)A(x),forn1,A_{n}(x)=\prod_{j=n-1}^{0}A(T^{j}x)=A(T^{n-1}x)\cdots A(Tx)A(x),\ for\ n\geq 1,

and An(x)=An(Tnx)1A_{-n}(x)=A_{n}(T^{-n}x)^{-1}. The Lyapunov exponent is defined as

L(T,A)=limn1nXlnAn(x)𝑑x.\quad L(T,A)=\lim_{n\rightarrow\infty}\frac{1}{n}\int_{X}\ln||A_{n}(x)||dx.

In this paper, we will consider the following two useful cocycles.

  • X=𝕋X=\mathbb{T} and T=RαT=R_{\alpha}, where Rαθ=θ+αR_{\alpha}\theta=\theta+\alpha, then (α,A):=(Rα,A)(\alpha,A):=(R_{\alpha},A) is a quasi-periodic cocycle.

  • X=𝕋×κX=\mathbb{T}\times\mathbb{Z}_{\kappa} and T=TαT=T_{\alpha}, where κ+\kappa\in{\mathbb{Z}}^{+}, Tα(θ,n)=(θ+α,n+1)T_{\alpha}(\theta,n)=(\theta+\alpha,n+1), then (Tα,A)(T_{\alpha},A) defines an almost-periodic cocycle.

These dynamical system (X,T)(X,T) is uniquely ergodic if α\alpha is irrational (Theorem 9.1 of [69]).

We say an SL(2,)SL(2,\mathbb{R}) cocycle (T,A)(T,A) is uniformly hyperbolic if, for every xXx\in X, there exists a continuous splitting 2=Es(x)Eu(x)\mathbb{R}^{2}=E_{s}(x)\oplus E_{u}(x) such that for every n0n\geq 0,

|An(x)v(x)|Cecn|v(x)|,v(x)Es(x),|An(x)v(x)|Cecn|v(x)|,v(x)Eu(x),\quad\begin{split}|A_{n}(x)v(x)|&\leq Ce^{-cn}|v(x)|,\ v(x)\in E_{s}(x),\\ |A_{-n}(x)v(x)|&\leq Ce^{-cn}|v(x)|,\ v(x)\in E_{u}(x),\end{split}

for some constans C,c>0C,c>0. Clearly, it holds that A(x)Es(x)=Es(Tx)A(x)E_{s}(x)=E_{s}(Tx) and A(x)Eu(x)=Eu(Tx)A(x)E_{u}(x)=E_{u}(Tx) for every xXx\in X, and if (T,A)(T,A) is uniformly hyperbolic, then L(T,A)>0L(T,A)>0.

2.3. Fibre rotation number.

Let 𝕊1\mathbb{S}^{1} be the set of unit vectors of 2\mathbb{R}^{2}, consider a projective cocycle FAF_{A} on X×𝕊1X\times\mathbb{S}^{1}:

(x,ϕ)(Tx,A(x)ϕA(x)ϕ).(x,\phi)\mapsto(Tx,\frac{A(x)\phi}{\|A(x)\phi\|}).

If AC0(𝕋,SL(2,))A\in C^{0}({\mathbb{T}},SL(2,{\mathbb{R}})) is homotopic to the identity, then there exists a lift F~A\tilde{F}_{A} of FAF_{A} to X×X\times\mathbb{R} such that F~A(x,ϕ)=(Tx,f~A(x,ϕ))\tilde{F}_{A}(x,\phi)=(Tx,\tilde{f}_{A}(x,\phi)) where f~A:X×\tilde{f}_{A}:X\times\mathbb{R}\rightarrow\mathbb{R} is a continuous lift such that

  • f~A(x,ϕ+1)=f~A(x,ϕ)+1;\tilde{f}_{A}(x,\phi+1)=\tilde{f}_{A}(x,\phi)+1;

  • for every xX,f~A(x,):RRx\in X,\tilde{f}_{A}(x,\cdot):R\rightarrow R is a strictly increasing homeomorphism;

  • if π2\pi_{2} is the projection map X×X×𝕊1:(x,ϕ)(x,e2πiϕ)X\times\mathbb{R}\rightarrow X\times\mathbb{S}^{1}:(x,\phi)\mapsto(x,e^{2\pi i\phi}), then FAπ2=π2F~AF_{A}\circ\pi_{2}=\pi_{2}\circ\tilde{F}_{A}.

If (X,ν,T)(X,\nu,T) is uniquely ergodic, then the number

ρ(T,A)=limnf~An(x,ϕ)ϕnmod\rho(T,A)=\lim_{n\rightarrow\infty}\frac{\tilde{f}_{A}^{n}(x,\phi)-\phi}{n}\mod{\mathbb{Z}}

is independent of (x,ϕ)X×/(x,\phi)\in X\times\mathbb{R/Z} and the lift of FAF_{A}, and is called the fibered rotation number of (T,A)(T,A), see [62, 48] for details.

If X=𝕋X=\mathbb{T} and T=RαT=R_{\alpha}, i.e. when we are dealing with quasi-periodic cocycles, we will simply denote its fiber rotation number as ρ(α,A)\rho(\alpha,A). The fibered rotation number is invariant under real conjugacies which are homotopic to the identity. In general, if the cocycles (α,A1)(\alpha,A_{1}) is conjugated to (α,A2)(\alpha,A_{2}):

B(θ+α)1A1(θ)B(θ)=A2(θ),B(\theta+\alpha)^{-1}A_{1}(\theta)B(\theta)=A_{2}(\theta),

and BC0(𝕋,B\in C^{0}({\mathbb{T}}, PSL(2,))PSL(2,{\mathbb{R}})) has degree n (that is, it is homotopic to θRnθ/2\theta\mapsto R_{n\theta/2}), where

Rϕ=(cos2πϕsin2πϕsin2πϕcos2πϕ),R_{\phi}=\begin{pmatrix}\cos 2\pi\phi&-\sin 2\pi\phi\\ \sin 2\pi\phi&\cos 2\pi\phi\end{pmatrix},

then we have

(7) ρ(α,A1)=ρ(α,A2)+12nαmod.\rho({\alpha,A_{1}})=\rho(\alpha,A_{2})+\frac{1}{2}n\alpha\mod\ \mathbb{Z}.

If furthermore BC0(𝕋,B\in C^{0}({\mathbb{T}}, SL(2,))SL(2,{\mathbb{R}})) with degB=n\deg B=n\in{\mathbb{Z}}, then we

ρ(α,A1)=ρ(α,A2)+nαmod.\rho({\alpha,A_{1}})=\rho(\alpha,A_{2})+n\alpha\mod{\mathbb{Z}}.

2.4. Dynamical defined Schrödinger operators.

Let XX be a compact metric space, (X,ν,T)(X,\nu,T) be ergodic, and V:XV:X\rightarrow{\mathbb{R}} is continuous. Then one can define the Schrödinger operator on 2()\ell^{2}({\mathbb{Z}}):

(HV,xu)n=un+1+un1+V(Tnx)un,xX.(H_{V,x}u)_{n}=u_{n+1}+u_{n-1}+V(T^{n}x)u_{n},\ \ \forall x\in X.

It is well known that the spectrum of HV,xH_{V,x} is a compact subset of \mathbb{R}, independent of xx if (X,T)(X,T) is minimal [30], we shall denote it by Σ(V)\Sigma(V). The integrated density of states (IDS) NV:[0,1]N_{V}:\mathbb{R}\rightarrow[0,1] of HV,xH_{V,x} is defined as

NV(E)=XμV,x(,E]𝑑ν,N_{V}(E)=\int_{X}\mu_{V,x}(-\infty,E]d\nu,

where μV,x\mu_{V,x} is the spectral measure of HV,xH_{V,x}. Note any formal solution u=(un)nu=(u_{n})_{n\in\mathbb{Z}} of HV,xu=EuH_{V,x}u=Eu can be rewritten as

(un+1un)=SEV(Tnx)(unun1),\begin{pmatrix}u_{n+1}\\ u_{n}\end{pmatrix}=S_{E}^{V}(T^{n}x)\begin{pmatrix}u_{n}\\ u_{n-1}\end{pmatrix},

where

SEV()=(EV()110),S_{E}^{V}(\cdot)=\begin{pmatrix}E-V(\cdot)&-1\\ 1&0\end{pmatrix},

and we call (T,SEV)(T,S_{E}^{V}) the Schrödinger cocycle. It is well-known that EΣ(V)E\notin\Sigma(V) if and only if (T,SEV)(T,S_{E}^{V}) is uniformly hyperbolic [61].

In this paper, we are interested in the case that (X,T)=(𝕋,Rα)(X,T)=({\mathbb{T}},R_{\alpha}) or (𝕋×κ,Tα)({\mathbb{T}}\times{\mathbb{Z}}_{\kappa},T_{\alpha}), where α\alpha is irrational, then the base dynamics is almost periodic (thus minimal and uniquely ergodic). For any fixed EE\in\mathbb{R}, the map θSEV(θ)\theta\mapsto S_{E}^{V}(\theta) is homotopic to the identity, hence the rotation number ρ(T,SEV)\rho(T,S_{E}^{V}) is well defined. Moreover, ρ(T,SEV)[0,12]\rho(T,S_{E}^{V})\in[0,\frac{1}{2}] relates to the integrated density of states NVN_{V} as follows:

NV(E)=12ρ(T,SEV).N_{V}(E)=1-2\rho(T,S_{E}^{V}).

By Thouless formula [15], we also have the following relation between the integrated density of states and the Lyapunov exponent:

L(T,SEV)=ln|EE|dNV(E).L(T,S_{E}^{V})=\int\ln|E-E^{\prime}|dN_{V}(E^{\prime}).

To get the existence of absolutely continuous spectrum, we need the following well-known result from subordinacy theory:

Theorem 2.2.

[44] Let \mathcal{B} be the set of EE\in{\mathbb{R}} such that the Schrödinger cocycle (T,SEV)(T,S_{E}^{V}) is bounded. Then μV,θ|\mu_{V,\theta}|\mathcal{B} is absolutely continuous for all θX\theta\in X.

Moreover, one can relate the growth of the cocycles to the spectral measure directly:

Lemma 2.3.

[7] There exists universal constant C>0C>0, such that μV,θ(Eϵ,E+ϵ)Cϵsup0sϵ1(SEV)s02\mu_{V,\theta}(E-\epsilon,E+\epsilon)\leq C\epsilon\sup_{0\leq s\leq\epsilon^{-1}}\|(S_{E}^{V})_{s}\|_{0}^{2}.

2.5. Global theory of one frequency quasiperiodic cocycle.

Let us make a short review of Avila’s global theory of one-frequency quasi-periodic cocycles [5]. Suppose that DCω(𝕋,M(2,))D\in C^{\omega}({\mathbb{T}},M(2,{\mathbb{C}})) admits a holomorphic extension to {|θ|<h}\{|\Im\theta|<h\}. Then for |ϵ|<h|\epsilon|<h, we define DϵCω(𝕋,M(2,))D_{\epsilon}\in C^{\omega}({\mathbb{T}},M(2,{\mathbb{C}})) by Dϵ()=A(+iϵ)D_{\epsilon}(\cdot)=A(\cdot+i\epsilon), and define the the acceleration of (α,Dε)(\alpha,D_{\varepsilon}) as follows

ω(α,Dε)=12πlimh0+L(α,Dε+h)L(α,Dε)h.\omega(\alpha,D_{\varepsilon})=\frac{1}{2\pi}\lim_{h\to 0+}\frac{L(\alpha,D_{\varepsilon+h})-L(\alpha,D_{\varepsilon})}{h}.

The acceleration was first introduced by Avila for analytic SL(2,)SL(2,{\mathbb{C}})-cocycles [5], and extended to analytic M(2,)M(2,{\mathbb{C}}) cocycles by Jitomirskaya-Marx [54]. It follows from the convexity and continuity of the Lyapunov exponent that the acceleration is an upper semicontinuous function in parameter ε\varepsilon. The key property of the acceleration is that it is quantized:

Theorem 2.4 (Quantization of acceleration[5, 54, 55]).

Suppose that (α,D)(\)×Cω(𝕋,M2())(\alpha,D)\in(\mathbb{R}\backslash\mathbb{Q})\times C^{\omega}(\mathbb{T},M_{2}(\mathbb{C})) with detD(θ)detD(\theta) bound away from 0 on the strip {|θ|<h}\{|\Im\theta|<h\}, then ω(α,Dε)12\omega(\alpha,D_{\varepsilon})\in\frac{1}{2}\mathbb{Z} in the strip. Morveover, if DCω(𝕋,SL(2,))D\in C^{\omega}(\mathbb{T},SL(2,{\mathbb{C}})), then ω(α,Dε)\omega(\alpha,D_{\varepsilon})\in\mathbb{Z}

If AA takes values in SL(2,)SL(2,\mathbb{R}), then εL(α,Aε)\varepsilon\mapsto L(\alpha,A_{\varepsilon}) is an even function. By convexity, ω(α,A)0\omega(\alpha,A)\geq 0. And if α\\alpha\in\mathbb{R}\backslash\mathbb{Q}, then (α,A)(\alpha,A) is uniformly hyperbolic if and only if L(α,A)>0,L(\alpha,A)>0, and ω(α,A)=0\omega(\alpha,A)=0. The cocycles in SL(2,)SL(2,\mathbb{R}) which are not uniformly hyperbolic are classified into three regimes: subcritical, critical, and supercritical. Especially, (α,A)(\alpha,A) is said to be subcritical if L(α,A)=0,L(\alpha,A)=0, ω(α,A)=0\omega(\alpha,A)=0; the cocycle (α,A)(\alpha,A) is said to be supercritical if L(α,A)>0,L(\alpha,A)>0, ω(α,A)>0\omega(\alpha,A)>0; otherwise (α,A)(\alpha,A) is critical.

The heart of Avila’s global theory is his “Almost Reducibility Conjecture” (ARC), which says that subcritical implies almost reducibility. Recall that a cocycle (α,A)(\alpha,A) is (analytically) reducible, if it can be CωC^{\omega} conjugated to a constant cocycle; (α,A)(\alpha,A) is (analytically) almost reducible if the closure of its analytic conjugates contains a constant. The full solution of ARC was recently given by Avila in [4, 6]:

Theorem 2.5.

[4, 6] Given α\\alpha\in\mathbb{R}\backslash\mathbb{Q}, and ACω(𝕋A\in C^{\omega}(\mathbb{T}, SL(2,))SL(2,\mathbb{R})), if (α,A)(\alpha,A) is subcritical, then it is almost reducible.

If we restrict ourself to the quasi-periodic Schrödinger cocycle (α,SEV)(\alpha,S_{E}^{V}), which comes from the quasi-periodic Schrödinger operator

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

Then we classify the energy EΣ(V)E\in\Sigma(V) by the dynamical behavior of (α,SEV)(\alpha,S_{E}^{V}). We denote EΣ(V)E\in\Sigma_{-}(V) if and only if (α,SEV)(\alpha,S_{E}^{V}) is subcritical, EΣc(V)E\in\Sigma_{c}(V) if and only if (α,SEV)(\alpha,S_{E}^{V}) is critical, and EΣ+(V)E\in\Sigma_{+}(V) if and only if (α,SEV)(\alpha,S_{E}^{V}) is supercritical.

3. Explicit formulas of Lyapunov exponent in the spectrum

It is known that the ac spectrum locates at the place where the Lyapunov exponent is zero, while pp spectrum locates at the place where the Lyapunov exponent is positive. Thus, the key for ME is the exact formula of the Lyapunov exponent. For this purpose, we consider the cocycle (α,A(+iϵ))(\alpha,A(\cdot+i\epsilon)) with ϵ>0\epsilon>0. For models considered in this paper, we can reduce the non-trival problem of computing L(α,A())L(\alpha,A(\cdot)) to the problem of computing limϵL(α,A(+iϵ))\lim_{\epsilon\to\infty}L(\alpha,A(\cdot+i\epsilon)). The later is much easier. This approach was based on Avila’s global theory of one-frequency quasi-periodic cocycles [5].

3.1. Lyapunov exponent for the mosaic model.

Note that for the mosaic model (3), let (abusing the notation a bit, we still denote it by V2V_{2})

V2(θ,n)={2λcos2πθ,nκ,0,else.V_{2}(\theta,n)=\left\{\begin{matrix}2\lambda\cos 2\pi\theta,&n\in\kappa\mathbb{Z},\\ 0,&else.\end{matrix}\right.

Then V2V_{2} is defined on 𝕋×κ\mathbb{T}\times\mathbb{Z}_{\kappa}, consequently (3) induces an almost-periodic Schrödinger cocycle (Tα,SEV2)(T_{\alpha},S_{E}^{V_{2}}) where Tα(θ,n)=(θ+α,n+1)T_{\alpha}(\theta,n)=(\theta+\alpha,n+1). Although (Tα,SEV2)(T_{\alpha},S_{E}^{V_{2}}) is not a quasi-periodic cocycle in the strict sense, its iterate

(κα,DEV2)=:(κα,SEV2(θ,κ1)SEV2(θ,0))(\kappa\alpha,D_{E}^{V_{2}})=:(\kappa\alpha,S_{E}^{V_{2}}(\theta,\kappa-1)\cdots S_{E}^{V_{2}}(\theta,0))

indeed defines an analytic quasi-periodic cocycle. By simple calculation,

DEV2(θ)=(E110)κ1(E2λcos2πθ110)=(aκaκ1aκ1aκ2)(E2λcos2πθ110)\begin{split}D_{E}^{V_{2}}(\theta)&=\begin{pmatrix}E&-1\\ 1&0\end{pmatrix}^{\kappa-1}\begin{pmatrix}E-2\lambda\cos 2\pi\theta&-1\\ 1&0\end{pmatrix}\\ &=\begin{pmatrix}a_{\kappa}&-a_{\kappa-1}\\ a_{\kappa-1}&-a_{\kappa-2}\end{pmatrix}\begin{pmatrix}E-2\lambda\cos 2\pi\theta&-1\\ 1&0\end{pmatrix}\\ &\end{split}

where

aκ=1E24((E+E242)κ(EE242)κ),a_{\kappa}=\frac{1}{\sqrt{E^{2}-4}}\left((\frac{E+\sqrt{E^{2}-4}}{2})^{\kappa}-(\frac{E-\sqrt{E^{2}-4}}{2})^{\kappa}\right),

and aκ(±2)=(1)κ1κa_{\kappa}(\pm 2)=(-1)^{\kappa-1}\kappa by continuity. It is easy to see that L(Tα,SEV2)=1κL(κα,DEV2)L(T_{\alpha},S_{E}^{V_{2}})=\frac{1}{\kappa}L(\kappa\alpha,D_{E}^{V_{2}}). The latter can be explicitly computed by Avila’s global theory, thus we have the following result:

Lemma 3.1.

Suppose that λ0\lambda\neq 0 and α\\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}}. Then for EΣ(V2)E\in\Sigma(V_{2}),

L(Tα,SEV2)=1κmax{ln|λaκ|,0}L(T_{\alpha},S_{E}^{V_{2}})=\frac{1}{\kappa}\max\{\ln|\lambda a_{\kappa}|,0\}

Moreover, if EΣ(V2)E\in\Sigma(V_{2}), then the cocycle (κα,DEV2)(\kappa\alpha,D_{E}^{V_{2}}) is:

  • supercritical, if and only if |λaκ|>1|\lambda a_{\kappa}|>1,

  • critical, if and only if |λaκ|=1|\lambda a_{\kappa}|=1,

  • subcritical, if and only if |λaκ|<1|\lambda a_{\kappa}|<1.

Proof.

It suffices to prove that for any EΣ(V2),E\in\Sigma(V_{2}), we have

L(κα,DEV2)=max{ln|λaκ|,0}.L(\kappa\alpha,D_{E}^{V_{2}})=\max\{\ln|\lambda a_{\kappa}|,0\}.

First we rewrite the matrix DEV2(θ)D_{E}^{V_{2}}(\theta) as

DEV2(θ)=(aκ(Eλ(ei2πθ+ei2πθ))aκ1aκaκ1(Eλ(ei2πθ+e2πθ))aκ2aκ1),D_{E}^{V_{2}}(\theta)=\begin{pmatrix}a_{\kappa}(E-\lambda(e^{i2\pi\theta}+e^{-i2\pi\theta}))-a_{\kappa-1}&-a_{\kappa}\\ a_{\kappa-1}(E-\lambda(e^{i2\pi\theta}+e^{-2\pi\theta}))-a_{\kappa-2}&-a_{\kappa-1}\end{pmatrix},

then we complexify the phase

(DEV2)ϵ=:DEV2(θ+iϵ)=(aκ(Eλ(ei2π(θ+iϵ)+ei2π(θ+iϵ)))aκ1aκaκ1(Eλ(ei2π(θ+iϵ)+e2π(θ+iϵ)))aκ2aκ1),\begin{split}(D_{E}^{V_{2}})_{\epsilon}&=:D_{E}^{V_{2}}(\theta+i\epsilon)\\ &=\begin{pmatrix}a_{\kappa}(E-\lambda(e^{i2\pi(\theta+i\epsilon)}+e^{-i2\pi(\theta+i\epsilon)}))-a_{\kappa-1}&-a_{\kappa}\\ a_{\kappa-1}(E-\lambda(e^{i2\pi(\theta+i\epsilon)}+e^{-2\pi(\theta+i\epsilon)}))-a_{\kappa-2}&-a_{\kappa-1}\end{pmatrix},\end{split}

thus for sufficiently large ϵ\epsilon

DEV2(θ+iϵ)=e2πϵei2πθ(λaκ0λaκ10)+o(1).D_{E}^{V_{2}}(\theta+i\epsilon)=e^{2\pi\epsilon}e^{-i2\pi\theta}\begin{pmatrix}-\lambda a_{\kappa}&0\\ -\lambda a_{\kappa-1}&0\end{pmatrix}+o(1).

Let A=(λaκ0λaκ10)A=\begin{pmatrix}-\lambda a_{\kappa}&0\\ -\lambda a_{\kappa-1}&0\end{pmatrix}. Then An=(λ)n(aκn0aκ1aκn10).A^{n}=(-\lambda)^{n}\begin{pmatrix}a_{\kappa}^{n}&0\\ a_{\kappa-1}a_{\kappa}^{n-1}&0\end{pmatrix}. It is obvious that

limnlnAnn=limnln|(λ)naκn|n=ln|λaκ|.\lim_{n\rightarrow\infty}\frac{\ln||A^{n}||}{n}=\lim_{n\rightarrow\infty}\frac{\ln|(-{\lambda})^{n}a_{\kappa}^{n}|}{n}=\ln|{\lambda a_{\kappa}}|.

By the continuity of Lyapunov exponent [28, 56], we have

L(κα,(DEV2)ϵ)=2πϵ+ln|λaκ|+o(1).L(\kappa\alpha,(D_{E}^{V_{2}})_{\epsilon})=2\pi\epsilon+\ln|{\lambda a_{\kappa}}|+o(1).

By Theorem 2.4, ω(κα,(DEV2)ϵ)=1\omega(\kappa\alpha,(D_{E}^{V_{2}})_{\epsilon})=1 and

L(κα,(DEV2)ϵ)=2πϵ+ln|λaκ|L(\kappa\alpha,(D_{E}^{V_{2}})_{\epsilon})=2\pi\epsilon+\ln|{\lambda a_{\kappa}}|

for sufficiently large ϵ\epsilon. By real-symmetry, ω(κα,(DEV2)ϵ)\omega(\kappa\alpha,(D_{E}^{V_{2}})_{\epsilon}) is either 0 or 1 for ϵ0\epsilon\geq 0. This implies that

(8) L(κα,(DEV2)ϵ)=max{ln|λaκ|+2πϵ,L(κα,DEV2)}.L(\kappa\alpha,(D_{E}^{V_{2}})_{\epsilon})=\max\{\ln|\lambda a_{\kappa}|+2\pi\epsilon,L(\kappa\alpha,D_{E}^{V_{2}})\}.

As a consequence, we have

L(κα,DEV2)max{ln|λaκ|,0}.L(\kappa\alpha,D_{E}^{V_{2}})\geq\max\{\ln|\lambda a_{\kappa}|,0\}.

If L(κα,DEV2)>max{ln|λaκ|,0}L(\kappa\alpha,D_{E}^{V_{2}})>\max\{\ln|\lambda a_{\kappa}|,0\}, then L(κα,DEV2)>0L(\kappa\alpha,D_{E}^{V_{2}})>0 and ω(κα,(DEV2)ϵ)=0\omega(\kappa\alpha,(D_{E}^{V_{2}})_{\epsilon})=0 for sufficient small and positive ϵ\epsilon, which implies that (κα,DEV2)(\kappa\alpha,D_{E}^{V_{2}}) is uniformly hyperbolic by Theorem 6 of [5], and thus (Tα,SEV2)(T_{\alpha},S_{E}^{V_{2}}) is uniformly hyperbolic. It contradicts with EΣ(V2)E\in\Sigma(V_{2}). Therefore

L(κα,DEV2)=max{ln|λaκ|,0}L(\kappa\alpha,D_{E}^{V_{2}})=\max\{\ln|\lambda a_{\kappa}|,0\}

for EΣ(V2).E\in\Sigma(V_{2}). Moreover, (8) implies that (κα,DEV2)(\kappa\alpha,D_{E}^{V_{2}}) is supercritical if and only if |λaκ|>1|\lambda a_{\kappa}|>1. The other cases follow similarly. ∎

To locate the spectrum, we need the following observations:

Lemma 3.2.

For any α\,κ+\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}},\kappa\in\mathbb{Z}^{+}, we have Σ(V2)=Σ(V2)\Sigma(V_{2})=-\Sigma(V_{2}).

Proof.

Suppose that EΣ(V2)E\in\Sigma(V_{2}) and HV2,α,θu=EuH_{V_{2},\alpha,\theta}u=Eu, define u~\widetilde{u} as

u~2n=u2n,u~2n+1=u2n+1.\widetilde{u}_{2n}=u_{2n},\ \widetilde{u}_{2n+1}=-u_{2n+1}.

Direct computation shows that HV2,α,θ+12u~=Eu~H_{V_{2},\alpha,\theta+\frac{1}{2}}\widetilde{u}=-E\widetilde{u}. Then EΣ(V2)-E\in\Sigma(V_{2}) since Σ(V2)\Sigma(V_{2}) is independence of the phase θ\theta. ∎

Lemma 3.2 says that Σ(V2)\Sigma(V_{2}) is symmetric with respect to 0, actually we will show that 0 always belongs to Σ(V2)\Sigma(V_{2}) (Lemma 3.3 and Lemma 3.4). Moreover, by direct calculation, we obtain that aκa_{\kappa} has κ1\kappa-1 roots (El=2cosπlκ,l=1,,κ1E_{l}=2\cos\frac{\pi l}{\kappa},l=1,\cdots,\kappa-1). Thus the set of EE satisfying |λaκ|<1|\lambda a_{\kappa}|<1 is a union of at most κ1\kappa-1 open intervals j=1mλ(bj,cj)\cup_{j=1}^{m_{\lambda}}(b_{j},c_{j}) such that |aκ(bj)|=|aκ(cj)|=1λ,|a_{\kappa}(b_{j})|=|a_{\kappa}(c_{j})|=\frac{1}{\lambda}, 1mλκ11\leq m_{\lambda}\leq\kappa-1, and each of these open intervals (bj,cj)(b_{j},c_{j}) has at least one root of aκa_{\kappa}. Obviously, the distance between these ElE_{l} is constant, thus mλ=κ1m_{\lambda}=\kappa-1 if λ\lambda is sufficiently large. Next we will show the roots El=2cosπlκE_{l}=2\cos\frac{\pi l}{\kappa} are always in the spectrum, indeed for any ElE_{l}, the corresponding cocycles (κα,DElV2)(\kappa\alpha,D_{E_{l}}^{V_{2}}) are always reducible:

Lemma 3.3.

For any α\,\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}}, λ0\lambda\neq 0, κ+\kappa\in\mathbb{Z}^{+}. Then there exists BCω(𝕋,SL(2,))B\in C^{\omega}({\mathbb{T}},SL(2,{\mathbb{R}})) such that

B(θ+2α)1DElV2(θ)B(θ)=(±100±1).B(\theta+2\alpha)^{-1}D_{E_{l}}^{V_{2}}(\theta)B(\theta)=\begin{pmatrix}\pm 1&0\\ 0&\pm 1\end{pmatrix}.

In particular, one has

(9) Σ(V2){E||λaκ|<1}.\Sigma(V_{2})\cap\{E\in{\mathbb{R}}\quad|\quad|\lambda a_{\kappa}|<1\}\neq\emptyset.
Proof.

Note that aκ=aκ1Eaκ2a_{\kappa}=a_{\kappa-1}E-a_{\kappa-2} and aκ(El)=0a_{\kappa}(E_{l})=0, direct computation shows that

DElV2(θ)=(aκ102λaκ1cos2πθaκ1)=aκ1(El)(102λcos2πθ1),\begin{split}D_{E_{l}}^{V_{2}}(\theta)=\begin{pmatrix}-a_{\kappa-1}&0\\ -2\lambda a_{\kappa-1}\cos 2\pi\theta&-a_{\kappa-1}\end{pmatrix}=-a_{\kappa-1}(E_{l})\begin{pmatrix}1&0\\ 2\lambda\cos 2\pi\theta&1\end{pmatrix},\end{split}

where aκ1(El)=±1a_{\kappa-1}(E_{l})=\pm 1 since detDElV2=1\det D_{E_{l}}^{V_{2}}=1. The equation

2λcos2πθ=h(θ+2α)h(θ)2\lambda\cos 2\pi\theta=h(\theta+2\alpha)-h(\theta)

always has a solution since 𝕋cos2πθdθ=0\int_{\mathbb{T}}\cos 2\pi\theta d\theta=0 and α\alpha is irrational. Let h(θ)Cω(𝕋,)h(\theta)\in C^{\omega}(\mathbb{T},\mathbb{R}) be its solution and denote B(θ)=(10h(θ)1)B(\theta)=\begin{pmatrix}1&0\\ h(\theta)&1\end{pmatrix}, then one can easily check that

B(θ+2α)1DElV2(θ)B(θ)=aκ1(El)(10h(θ+2α)1)(102λcos2πθ1)(10h(θ)1)=(±100±1),\begin{split}&B(\theta+2\alpha)^{-1}D_{E_{l}}^{V_{2}}(\theta)B(\theta)\\ &=-a_{\kappa-1}(E_{l})\begin{pmatrix}1&0\\ -h(\theta+2\alpha)&1\end{pmatrix}\begin{pmatrix}1&0\\ 2\lambda\cos 2\pi\theta&1\end{pmatrix}\begin{pmatrix}1&0\\ h(\theta)&1\end{pmatrix}\\ &=\begin{pmatrix}\pm 1&0\\ 0&\pm 1\end{pmatrix},\end{split}

This implies that L(Tα,SElV2)=1κL(κα,DElV2)=0L(T_{\alpha},S_{E_{l}}^{V_{2}})=\frac{1}{\kappa}L(\kappa\alpha,D_{E_{l}}^{V_{2}})=0, and ElΣ(V2)E_{l}\in\Sigma(V_{2}), then (9) follows directly. ∎

As a direct consequence, if κ\kappa is an even number, then 0Σ(V2)0\in\Sigma(V_{2}), if κ\kappa is an odd number, then we have the following observation:

Lemma 3.4.

For any α\,\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}}, λ0\lambda\neq 0. If κ2+1\kappa\in 2\mathbb{Z}+1, then 0Σ(V2)0\in\Sigma(V_{2}).

Proof.

If κ2+1\kappa\in 2\mathbb{Z}+1, and E=0E=0, by simple calculation,

(κα,D0V2(θ))=(κα,±(2λcos2πθ110)),(\kappa\alpha,D_{0}^{V_{2}}(\theta))=(\kappa\alpha,\pm\begin{pmatrix}-2\lambda\cos 2\pi\theta&-1\\ 1&0\end{pmatrix}),

which is just the almost Mathieu cocycle. Thus 0Σ(V2)0\in\Sigma(V_{2}) is equivalent to whether 0 belongs to the spectrum of almost Mathieu operator:

(Hλ,κα,θu)n=un+1+un1+2λcos2π(nκα+θ)un,(H_{\lambda,\kappa\alpha,\theta}u)_{n}=u_{n+1}+u_{n-1}+2\lambda\cos 2\pi(n\kappa\alpha+\theta)u_{n},

then the result follows from [17] directly. ∎

Now we summarize the above results in the case κ=2\kappa=2 and κ=3\kappa=3, which can be clarified very clearly:

Corollary 3.5.

Suppose that λ0\lambda\neq 0, α\\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}}, and κ=2\kappa=2. Then we have

L(Tα,SEV2)=12max{ln|λE|,0},EΣ(V2).L(T_{\alpha},S_{E}^{V_{2}})=\frac{1}{2}\max\{\ln|\lambda E|,0\},\ E\in\Sigma(V_{2}).

Moreover, the following holds true:

  1. (1)

    If λE¯(V2)<1\lambda\overline{E}(V_{2})<1, then for any EΣ(V2)E\in\Sigma(V_{2}), (2α,DEV2)(2\alpha,D_{E}^{V_{2}}) is subcritical.

  2. (2)

    If λE¯(V2)>1\lambda\overline{E}(V_{2})>1, then we have the following:

    • Σ(V2)(1λ,1λ)\Sigma(V_{2})\cap(-\frac{1}{\lambda},\frac{1}{\lambda})\neq\emptyset, furthermore (2α,DEV2)(2\alpha,D_{E}^{V_{2}}) is subcritical.

    • Σ(V2)[1λ,1λ]c\Sigma(V_{2})\cap[-\frac{1}{\lambda},\frac{1}{\lambda}]^{c}\neq\emptyset, furthermore (2α,DEV2)(2\alpha,D_{E}^{V_{2}}) is supercritical.

Proof.

Just note a2=Ea_{2}=E, then it is direct consequences of Lemma 3.1 and Lemma 3.3. ∎

Lemma 3.5 ensures Σ(V2)(1λ,1λ)\Sigma(V_{2})\cap(-\frac{1}{\lambda},\frac{1}{\lambda})\neq\emptyset for any λ.\lambda. Note if λ\lambda is small enough, then

Σ(V2)[22λ,2+2λ](1λ,1λ),\Sigma(V_{2})\subseteq[-2-2\lambda,2+2\lambda]\subset(-\frac{1}{\lambda},\frac{1}{\lambda}),

which means Corollary 3.5 (1)(1) holds for small λ\lambda and there is no ME. It follows that ME appears only when λ\lambda is relatively large. However, 1λ<E¯(V2)\frac{1}{\lambda}<\overline{E}(V_{2}) is not easy to be verified since E¯(V2)\overline{E}(V_{2}) depends implicitly on λ\lambda. Next result will tell us, Corollary 3.5 (2)(2) holds at least for λ>22:\lambda>\frac{\sqrt{2}}{2}:

Lemma 3.6.

If λ>22,\lambda>\frac{\sqrt{2}}{2}, κ=2\kappa=2, then Σ(V2)([22λ,1λ)(1λ,2+2λ])\Sigma(V_{2})\cap([-2-2\lambda,-\frac{1}{\lambda})\cup(\frac{1}{\lambda},2+2\lambda])\neq\emptyset

Proof.

We prove Σ(V2)(1λ,2+2λ]\Sigma(V_{2})\cap(\frac{1}{\lambda},2+2\lambda]\neq\emptyset. First by the spectral theorem, we have

2λcos2π(θ+2nα)=δ2n,HV2,α,θδ2n=E𝑑μV2,α,θ,δ2n.2\lambda\cos 2\pi(\theta+2n\alpha)=\left<\delta_{2n},H_{V_{2},\alpha,\theta}\delta_{2n}\right>=\int Ed\mu_{V_{2},\alpha,\theta,\delta_{2n}}.

One can select nn such that cos2π(2nα+θ)>12λ2\cos 2\pi(2n\alpha+\theta)>\frac{1}{2\lambda^{2}}, such nn exists since α\alpha is irrational and 12λ2<1\frac{1}{2\lambda^{2}}<1. On the other hand, suppose that Σ(V2)(1λ,2+2λ]=\Sigma(V_{2})\cap(\frac{1}{\lambda},2+2\lambda]=\emptyset, i.e. for any EΣ(V2)E\in\Sigma(V_{2}), we have |E|1λ|E|\leq\frac{1}{\lambda}. This immediately imply that

1λ<2λcos2π(2nα+θ)=E𝑑μV2,α,θ,δ2n1λ.\frac{1}{\lambda}<2\lambda\cos 2\pi(2n\alpha+\theta)=\int Ed\mu_{V_{2},\alpha,\theta,\delta_{2n}}\leq\frac{1}{\lambda}.

This is a contradiction. ∎

Remark 3.7.

This argument can be generalized to general κ\kappa almost without change. Consequently, combining Lemma 3.3, if |λ||\lambda| is relatively large, then ME always exists.

In the case κ=3\kappa=3, recall that

Ec1=1+1λ,Ec2=11λ,E_{c}^{1}=\sqrt{1+\frac{1}{\lambda}},\qquad E_{c}^{2}=\sqrt{1-\frac{1}{\lambda}},

then we have the following:

Corollary 3.8.

Suppose that λ0\lambda\neq 0, α\\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}}, and κ=3\kappa=3. Then we have

L(Tα,SEV2)=13max{ln|λ(E21)|,0},EΣ(V2).L(T_{\alpha},S_{E}^{V_{2}})=\frac{1}{3}\max\{\ln|\lambda(E^{2}-1)|,0\},\ E\in\Sigma(V_{2}).

Moreover, the following holds true:

  1. (1)

    If |λ|(E¯(V2)21)<1|\lambda|(\overline{E}(V_{2})^{2}-1)<1, then for any EΣ(V2)E\in\Sigma(V_{2}), (3α,DEV2)(3\alpha,D_{E}^{V_{2}}) is subcritical.

  2. (2)

    If 1E¯(V2)21<|λ|<1\frac{1}{\overline{E}(V_{2})^{2}-1}<|\lambda|<1, then we have the following:

    • Σ(V2)(Ec1,Ec1)\Sigma(V_{2})\cap(-E_{c}^{1},E_{c}^{1})\neq\emptyset, furthermore (3α,DEV2)(3\alpha,D_{E}^{V_{2}}) is subcritical.

    • Σ(V2)[Ec1,Ec1]c\Sigma(V_{2})\cap[-E_{c}^{1},E_{c}^{1}]^{c}\neq\emptyset, furthermore (3α,DEV2)(3\alpha,D_{E}^{V_{2}}) is supercritical.

  3. (3)

    If |λ|>1|\lambda|>1, then we have the following:

    • Σ(V2)(Ec2,Ec1)\Sigma(V_{2})\cap(E_{c}^{2},E_{c}^{1})\neq\emptyset and Σ(V2)(Ec1,Ec2)\Sigma(V_{2})\cap(-E_{c}^{1},E_{c}^{2})\neq\emptyset, furthermore (3α,DEV2)(3\alpha,D_{E}^{V_{2}}) is subcritical.

    • Σ(V2)[Ec1,Ec1]c\Sigma(V_{2})\cap[-E_{c}^{1},E_{c}^{1}]^{c}\neq\emptyset and Σ(V2)(Ec2,Ec2)\Sigma(V_{2})\cap(-E_{c}^{2},E_{c}^{2})\neq\emptyset, furthermore (3α,DEV2)(3\alpha,D_{E}^{V_{2}}) is supercritical.

Proof.

Just note a3=E21a_{3}=E^{2}-1, then it is direct consequences of Lemma 3.1, Lemma 3.3 and Lemma 3.4. ∎

3.2. Lyapunov exponent for the GAA model

Similar to the proof of Lemma 3.1, we can also calculate the Lyapunov exponent for the GAA model:

Lemma 3.9.

Suppose that λ\lambda\in{\mathbb{R}}, α\\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}} and |τ|1|\tau|\leq 1. Then, for any EΣ(V1)E\in\Sigma(V_{1}), we have

(10) L(α,SEV1)=max{ln||τE+2λ|+(τE+2λ)24τ22(1+1τ2)|,0}.L(\alpha,S_{E}^{V_{1}})=\max\{\ln|\frac{|\tau E+2\lambda|+\sqrt{(\tau E+2\lambda)^{2}-4\tau^{2}}}{2(1+\sqrt{1-\tau^{2}})}|,0\}.

Moreover, we have the following:

  • EΣ+(V1)E\in\Sigma_{+}(V_{1}), if and only if sgn(λ)τE>2(1|λ|)sgn(\lambda)\tau E>2(1-|\lambda|),

  • EΣc(V1)E\in\Sigma_{c}(V_{1}), if and only if sgn(λ)τE=2(1|λ|)sgn(\lambda)\tau E=2(1-|\lambda|),

  • EΣ(V1)E\in\Sigma_{-}(V_{1}), if and only if sgn(λ)τE<2(1|λ|)sgn(\lambda)\tau E<2(1-|\lambda|).

Proof.

We distinugish three cases: τ=0\tau=0, 0<|τ|<10<|\tau|<1 and |τ|=1|\tau|=1.

Case 1: If τ=0\tau=0, the result follows from [5], or Lemma 3.1 with κ=1\kappa=1.

Case 2: If 0<|τ|<10<|\tau|<1, the potential is bounded and analytic. Let D(θ):=2(1τcos2π(θ))SEV1(θ)D(\theta):=2(1-\tau\cos 2\pi(\theta))S_{E}^{V_{1}}(\theta), i.e.,

D(θ):=(2E(1τcos2πθ)4λcos2πθ2+2τcos2πθ22τcos2πθ0),\displaystyle D(\theta):=\begin{pmatrix}2E(1-\tau\cos 2\pi\theta)-4\lambda\cos 2\pi\theta&-2+2\tau\cos 2\pi\theta\\ &\\ 2-2\tau\cos 2\pi\theta&0\end{pmatrix},

then D(θ)D(\theta) admits a holomorphic extension to |θ|<|\Im\theta|<\infty. Note that

f(θ,ϵ)=2(1τcos2π(θ+ϵi))=2τe2πϵe2πθiτe2πϵe2πθi=τe2πϵ2πθi(e2πθi1+1τ2τe2πϵ)(e2πθi11τ2τe2πϵ).\begin{split}f(\theta,\epsilon)&=2(1-\tau\cos 2\pi(\theta+\epsilon i))\\ &=2-\tau e^{-2\pi\epsilon}e^{2\pi\theta i}-\tau e^{2\pi\epsilon}e^{-2\pi\theta i}\\ &=-\tau e^{-2\pi\epsilon-2\pi\theta i}(e^{2\pi\theta i}-\frac{1+\sqrt{1-\tau^{2}}}{\tau}e^{2\pi\epsilon})(e^{2\pi\theta i}-\frac{1-\sqrt{1-\tau^{2}}}{\tau}e^{2\pi\epsilon}).\end{split}

Thus SEV1Cω(𝕋,SL(2,))S_{E}^{V_{1}}\in C^{\omega}(\mathbb{T},SL(2,\mathbb{R})) admits a holomorphic extension to the strip |θ|<δ0|\Im\theta|<\delta_{0} with δ0=12πln1+1τ2|τ|\delta_{0}=\frac{1}{2\pi}\ln\frac{1+\sqrt{1-\tau^{2}}}{|\tau|}.

By Jensen’s formula, we have

𝕋ln|f(θ,ϵ)|dθ=ln(1+1τ2),|ϵ|<δ0.\int_{\mathbb{T}}\ln|f(\theta,\epsilon)|d\theta=\ln(1+\sqrt{1-\tau^{2}}),\quad\forall\,|\epsilon|<\delta_{0}.

Therefore we have

(11) L(α,Dϵ)=L(α,(SEV1)ϵ)+𝕋ln|f(θ,ϵ)|dθ=L(α,(SEV1)ϵ)+ln(1+1τ2),|ϵ|<δ0,\begin{split}L(\alpha,D_{\epsilon})&=L(\alpha,(S_{E}^{V_{1}})_{\epsilon})+\int_{\mathbb{T}}\ln|f(\theta,\epsilon)|d\theta\\ &=L(\alpha,(S_{E}^{V_{1}})_{\epsilon})+\ln(1+\sqrt{1-\tau^{2}}),\quad\forall\,|\epsilon|<\delta_{0},\end{split}

which implies that (α,(SEV1)ϵ)(\alpha,(S_{E}^{V_{1}})_{\epsilon}) and (α,Dϵ)(\alpha,D_{\epsilon}) have the same acceleration when |ϵ|<δ0|\epsilon|<\delta_{0}, i.e.

(12) ω(α,Dϵ)=ω(α,(SEV1)ϵ),|ϵ|<δ0.\omega(\alpha,D_{\epsilon})=\omega(\alpha,(S_{E}^{V_{1}})_{\epsilon}),\quad\forall|\epsilon|<\delta_{0}.

On the other hand, if we complexify the phase, and write

D(θ+ϵi)=(e2πθi+2πϵ+e2πθi2πϵ)((τE+2λ)ττ0)+(2E220).\begin{split}D(\theta+\epsilon i)=(e^{-2\pi\theta i+2\pi\epsilon}+e^{2\pi\theta i-2\pi\epsilon})\begin{pmatrix}-(\tau E+2\lambda)&\tau\\ -\tau&0\end{pmatrix}+\begin{pmatrix}2E&-2\\ 2&0\end{pmatrix}.\end{split}

Let ϵ\epsilon goes to infinity, then

D(θ+ϵi)=e2πθi+2πϵ((τE+2λ)+o(1)τ+o(1)τ+o(1)0).D(\theta+\epsilon i)=e^{-2\pi\theta i+2\pi\epsilon}\begin{pmatrix}-(\tau E+2\lambda)+o(1)&\tau+o(1)\\ -\tau+o(1)&0\end{pmatrix}.

By the continuity of Lyapunov exponent [28, 56], we have

L(α,Dϵ)=ln|h(E)|+2πϵ+o(1),L(\alpha,D_{\epsilon})=\ln|h(E)|+2\pi\epsilon+o(1),

where

h(E)=12(|τE+2λ|+(τE+2λ)24τ2).h(E)=\frac{1}{2}(|\tau E+2\lambda|+\sqrt{(\tau E+2\lambda)^{2}-4\tau^{2}}).

By quantization of acceleration(Theorem 2.4),

L(α,Dϵ)=ln|h(E)|+2πϵfor all ϵ sufficiently large,L(\alpha,D_{\epsilon})=\ln|h(E)|+2\pi\epsilon\quad\text{for all $\epsilon$ sufficiently large},

which also implies that ω(α,Dϵ)=1\omega(\alpha,D_{\epsilon})=1 for sufficiently large ϵ\epsilon.

By convexity, ω(α,Dϵ)1\omega(\alpha,D_{\epsilon})\leq 1 for every ϵ0\epsilon\geq 0. just note DSL(2,)D\notin SL(2,\mathbb{C}), in general one cann’t conclude ω(α,Dϵ)=0\omega(\alpha,D_{\epsilon})=0 or 11 for every ϵ0\epsilon\geq 0. Nevertheless, since SEV1Cω(𝕋,SL(2,))S_{E}^{V_{1}}\in C^{\omega}(\mathbb{T},SL(2,\mathbb{R})), again by Theorem 2.4, one has ω(α,(SEV1)ϵ)\omega(\alpha,(S_{E}^{V_{1}})_{\epsilon})\in{\mathbb{Z}} for any |ϵ|<δ0.|\epsilon|<\delta_{0}. Thus if EΣ+(V1)E\in\Sigma_{+}(V_{1}) or EΣc(V1)E\in\Sigma_{c}(V_{1}), then by (12), and the convexity of L(α,Dϵ)L(\alpha,D_{\epsilon}), we have ω(α,(SEV1)ϵ)=1\omega(\alpha,(S_{E}^{V_{1}})_{\epsilon})=1 for 0ϵ<δ00\leq\epsilon<\delta_{0} and ω(α,Dϵ)=1\omega(\alpha,D_{\epsilon})=1 for ϵ0\epsilon\geq 0. As a consequence, it holds that

L(α,Dϵ)=ln|h(E)|+2π|ϵ|,for allϵ.L(\alpha,D_{\epsilon})=\ln|h(E)|+2\pi|\epsilon|,\quad\text{for all}\ \epsilon.

where the case ϵ0\epsilon\leq 0 follows by real-symmetry. By (11) and the non-negativity of L(α,SEV1)L(\alpha,S_{E}^{V_{1}}), if EΣ+(V1)E\in\Sigma_{+}(V_{1}) or EΣc(V1)E\in\Sigma_{c}(V_{1}), we have

(13) L(α,(SEV1)ϵ)=ln|h(E)|1+1τ2+2π|ϵ|,|ϵ|<δ0.L(\alpha,(S_{E}^{V_{1}})_{\epsilon})=\ln\frac{|h(E)|}{1+\sqrt{1-\tau^{2}}}+2\pi|\epsilon|,\quad|\epsilon|<\delta_{0}.

If EΣ(V1)E\in\Sigma_{-}(V_{1}), then L(α,(SEV1)ϵ)=0L(\alpha,(S_{E}^{V_{1}})_{\epsilon})=0 for |ϵ|<δδ0|\epsilon|<\delta^{\prime}\leq\delta_{0}.

By Avila’s global theory [5], for any EΣ(V1)E\in\Sigma(V_{1}), the corresponding cocycle (α,SEV1)(\alpha,S_{E}^{V_{1}}) is either supercritical, critical, or subcritical, we thus only need to locate the energy which is supercritical or critical. Without losing generality, we assume λ<0\lambda<0, τ>0\tau>0. By (13)\eqref{equation_16}, EΣc(V1)E\in\Sigma_{c}(V_{1}), if and only if |h(E)|=1+1τ2|h(E)|=1+\sqrt{1-\tau^{2}}, which is equivalent to sgn(λ)τE=2(1|λ|)sgn(\lambda)\tau E=2(1-|\lambda|) by simple calculation. Meanwhile, EΣ+(V1)E\in\Sigma_{+}(V_{1}), if and only if |h(E)|>1+1τ2|h(E)|>1+\sqrt{1-\tau^{2}}, which is equivalent to |τE+2λ|>2|\tau E+2\lambda|>2. In our case λ<0\lambda<0 and τ>0\tau>0, we actually have τE+2λ<2\tau E+2\lambda<-2 since τE+2λ>2\tau E+2\lambda>2 is impossible. In fact if τE+2λ>2\tau E+2\lambda>2 then

E>22λτ>2+2λ1τ,E>\frac{2-2\lambda}{\tau}>2+\frac{2\lambda}{1-\tau},

which contradicts with the fact

Σ(V1)[22λ1+τ,2+2λ1τ].\Sigma(V_{1})\subseteq[-2-\frac{2\lambda}{1+\tau},2+\frac{2\lambda}{1-\tau}].

Case 3: In the limiting case |τ|=1|\tau|=1, the operator is unbounded. However, recall that (α,D)(\alpha,D) is a bounded and analytic cocycle, thus L(α,Dϵ)L(\alpha,D_{\epsilon}) is continuous in ϵ\epsilon. Moreover,

L(α,Dϵ)=L(α,(SEV1)ϵ)+𝕋ln|f(θ,ϵ)|dθ,L(\alpha,D_{\epsilon})=L(\alpha,(S_{E}^{V_{1}})_{\epsilon})+\int_{\mathbb{T}}\ln|f(\theta,\epsilon)|d\theta,

applying Jensen’s formula yields

𝕋ln|f(θ,ϵ)|dθ=2π|ϵ|,for allϵ.\int_{\mathbb{T}}\ln|f(\theta,\epsilon)|d\theta=2\pi|\epsilon|,\quad\text{for all}\ \epsilon.

Since the above equation explicitly implies the continuity of 𝕋ln|f(θ,ϵ)|dθ\int_{\mathbb{T}}\ln|f(\theta,\epsilon)|d\theta in ϵ\epsilon, the continuity of L(α,(SEV1)ϵ)L(\alpha,(S_{E}^{V_{1}})_{\epsilon}) follows.

Then, uniformly in θ𝕋\theta\in\mathbb{T}, one has

(SEV1)ϵ=B+o(1)(S_{E}^{V_{1}})_{\epsilon}=B_{\infty}+o(1)

as ϵ\epsilon goes to infinity, where

B=(E+2λτ110).B_{\infty}=\begin{pmatrix}E+2\lambda\tau&-1\\ 1&0\end{pmatrix}.

By continuity of the Lyapunov exponent [28, 56], we have

L(α,(SEV1)ϵ)=L(α,B)+o(1)L(\alpha,(S_{E}^{V_{1}})_{\epsilon})=L(\alpha,B_{\infty})+o(1)

as ϵ\epsilon goes to infinity.

The quantization of acceleration (Theorem 2.4) yields

L(α,(SEV1)ϵ)=L(α,B)for all ϵ sufficiently large.L(\alpha,(S_{E}^{V_{1}})_{\epsilon})=L(\alpha,B_{\infty})\quad\text{for all $\epsilon$ sufficiently large}.

In addition, the convexity, continuity, and symmetry of L(α,(SEV1)ϵ)L(\alpha,(S_{E}^{V_{1}})_{\epsilon}) with respect to ϵ\epsilon gives

L(α,(SEV1)ϵ)=L(α,B)for all ϵ.L(\alpha,(S_{E}^{V_{1}})_{\epsilon})=L(\alpha,B_{\infty})\quad\text{for all $\epsilon$}.

This actually implies that

L(α,SEV1)=max{|E+2λτ|+(E+2λτ)242,0},L(\alpha,S_{E}^{V_{1}})=\max\{\frac{|E+2\lambda\tau|+\sqrt{(E+2\lambda\tau)^{2}-4}}{2},0\},

which is exactly (10), since |τ|=1|\tau|=1. ∎

As a direct consequence of Lemma 3.9, we have

Corollary 3.10.

Suppose that λ\lambda\in{\mathbb{R}}, α\\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}} and |τ|<1|\tau|<1. Then if λτ>0\lambda\tau>0, the following holds true:

  1. (1)

    If |λ|<1|τ|2E¯(V1)|\lambda|<1-\frac{|\tau|}{2}\overline{E}(V_{1}), then we have

    Σ(V1)=Σ(V1).\Sigma(V_{1})=\Sigma_{-}(V_{1})\neq\emptyset.
  2. (2)

    If 1|τ|2E¯(V1)<|λ|<1|τ|2E¯(V1)1-\frac{|\tau|}{2}\overline{E}(V_{1})<|\lambda|<1-\frac{|\tau|}{2}\underline{E}(V_{1}), then we have

    (14) Σ(V1)[E¯(V1),2(1|λ|)|τ|)=Σ(V1),Σ(V1)(2(1|λ|)|τ|,E¯(V1)]=Σ+(V1).\begin{split}\Sigma(V_{1})\cap[\underline{E}(V_{1}),\frac{2(1-|\lambda|)}{|\tau|})&=\Sigma_{-}(V_{1})\neq\emptyset,\\ \Sigma(V_{1})\cap(\frac{2(1-|\lambda|)}{|\tau|},\overline{E}(V_{1})]&=\Sigma_{+}(V_{1})\neq\emptyset.\end{split}
  3. (3)

    If |λ|>1|τ|2E¯(V1)|\lambda|>1-\frac{|\tau|}{2}\underline{E}(V_{1}), then we have

    Σ(V1)=Σ+(V1).\Sigma(V_{1})=\Sigma_{+}(V_{1})\neq\emptyset.
Proof.

By Lemma 3.9, if λτ>0\lambda\tau>0, then EΣ+(V1)E\in\Sigma_{+}(V_{1}) if and only if |τ|E>2(1|λ|)|\tau|E>2(1-|\lambda|), and 2(1|λ|)|τ|\frac{2(1-|\lambda|)}{|\tau|} is the critical point. One can distinguish the following three cases:

  • if E¯(V1)>2(1|λ|)|τ|\underline{E}(V_{1})>\frac{2(1-|\lambda|)}{|\tau|}, then Σ(V1)=Σ+(V1)\Sigma(V_{1})=\Sigma_{+}(V_{1});

  • if E¯(V1)<2(1|λ|)|τ|\overline{E}(V_{1})<\frac{2(1-|\lambda|)}{|\tau|}, then Σ(V1)=Σ(V1)\Sigma(V_{1})=\Sigma_{-}(V_{1});

  • if E¯(V1)<2(1|λ|)|τ|<E¯(V1)\underline{E}(V_{1})<\frac{2(1-|\lambda|)}{|\tau|}<\overline{E}(V_{1}), then (14) holds.

We thus finish the whole proof. ∎

We have similar results for λτ<0\lambda\tau<0.

Corollary 3.11.

Suppose that λ\lambda\in{\mathbb{R}}, α\\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}} and |τ|<1|\tau|<1. Then if λτ<0\lambda\tau<0, the following holds true:

  1. (1)

    If |λ|<1+|τ|2E¯(V1)|\lambda|<1+\frac{|\tau|}{2}\underline{E}(V_{1}), then we have

    Σ(V1)=Σ(V1).\Sigma(V_{1})=\Sigma_{-}(V_{1})\neq\emptyset.
  2. (2)

    If 1+|τ|2E¯(V1)<|λ|<1+|τ|2E¯(V1)1+\frac{|\tau|}{2}\underline{E}(V_{1})<|\lambda|<1+\frac{|\tau|}{2}\overline{E}(V_{1}), then we have

    Σ(V1)(2(1|λ|)|τ|,E¯(V1)]=Σ(V1),\Sigma(V_{1})\cap(-\frac{2(1-|\lambda|)}{|\tau|},\overline{E}(V_{1})]=\Sigma_{-}(V_{1})\neq\emptyset,
    Σ+(V1)[E¯(V1),2(1|λ|)|τ|)=Σ+(V1).\Sigma_{+}(V_{1})\cap[\underline{E}(V_{1}),-\frac{2(1-|\lambda|)}{|\tau|})=\Sigma_{+}(V_{1})\neq\emptyset.
  3. (3)

    If |λ|>1+|τ|2E¯(V1)|\lambda|>1+\frac{|\tau|}{2}\overline{E}(V_{1}), then we have

    Σ(V1)=Σ+(V1).\Sigma(V_{1})=\Sigma_{+}(V_{1})\neq\emptyset.
Proof.

We omit the details, since the proof is the same as Corollary 3.10. ∎

In general, since E¯(V1)\underline{E}(V_{1}) and E¯(V1)\overline{E}(V_{1}) depend on the potential V1V_{1} implicitly, one does not exactly know when 1|τ|2E¯(V1)<|λ|<1|τ|2E¯(V1)1-\frac{|\tau|}{2}\overline{E}(V_{1})<|\lambda|<1-\frac{|\tau|}{2}\underline{E}(V_{1}) (or 1+|τ|2E¯(V1)<|λ|<1+|τ|2E¯(V1)1+\frac{|\tau|}{2}\underline{E}(V_{1})<|\lambda|<1+\frac{|\tau|}{2}\overline{E}(V_{1})) does happen, thus one does not exactly know when subcritical and supercritical energies coexist. However, we have the following:

Lemma 3.12.

Suppose that λ\lambda\in{\mathbb{R}}, α\\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}} and |τ|<1|\tau|<1. Then the following holds true:

  • If |λ|<(1|τ|)2|\lambda|<(1-|\tau|)^{2}, then we have

    Σ(V1)Σ(V1)=Σ(V1).\Sigma_{-}(V_{1})\cap\Sigma(V_{1})=\Sigma_{-}(V_{1})\neq\emptyset.
  • If 1|τ|<|λ|<1+|τ|1-|\tau|<|\lambda|<1+|\tau|, then we have

    (15) Σ(V1)[E¯(V1),2(1|λ|)|τ|)\displaystyle\Sigma(V_{1})\cap[\underline{E}(V_{1}),\frac{2(1-|\lambda|)}{|\tau|}) =\displaystyle= Σ(V1),\displaystyle\Sigma_{-}(V_{1})\neq\emptyset,
    (16) Σ(V1)(2(1|λ|)|τ|,E¯(V1)]\displaystyle\Sigma(V_{1})\cap(\frac{2(1-|\lambda|)}{|\tau|},\overline{E}(V_{1})] =\displaystyle= Σ+(V1).\displaystyle\Sigma_{+}(V_{1})\neq\emptyset.
  • If |λ|>(1+|τ|)2|\lambda|>(1+|\tau|)^{2}, then we have

    Σ+(V1)Σ(V1)=Σ+(V1).\Sigma_{+}(V_{1})\cap\Sigma(V_{1})=\Sigma_{+}(V_{1})\neq\emptyset.
Proof.

For simplicity, we only consider the case λ>0\lambda>0 and τ>0\tau>0, the other cases can be dealt with similarly. First note we have a trivial bound

22λ1+τE¯(V1)E¯(V1)2+2λ1τ,-2-\frac{2\lambda}{1+\tau}\leq\underline{E}(V_{1})\leq\overline{E}(V_{1})\leq 2+\frac{2\lambda}{1-\tau},

then the first statement and the third statement follows immediately from Corollary 3.10.

We are left to prove the second statement, and we only prove (15), since (16) can be proved similarly. First by the spectral theorem,

2λcos(2π(nα+θ))1τcos(2π(nα+θ))=δn,HV1,α,θδn=E𝑑μV1,α,θ,δn.\begin{split}\frac{2\lambda\cos(2\pi(n\alpha+\theta))}{1-\tau\cos(2\pi(n\alpha+\theta))}=\left<\delta_{n},H_{V_{1},\alpha,\theta}\delta_{n}\right>=\int Ed\mu_{V_{1},\alpha,\theta,\delta_{n}}.\end{split}

We argue by contradiction, assume that Σ(V1)Σ(V1)=\Sigma_{-}(V_{1})\cap\Sigma(V_{1})=\emptyset, then by Lemma 3.9, we have E2(1λ)τE\geq\frac{2(1-\lambda)}{\tau} for every EΣ(V1)E\in\Sigma(V_{1}). Select nn such that cos(2π(nα+θ))<1λτ\cos(2\pi(n\alpha+\theta))<\frac{1-\lambda}{\tau}, such nn exists since α\alpha is irrational and 1λτ>1\frac{1-\lambda}{\tau}>-1, as a consequence,

2(1λ)τ>2λcos(2π(nα+θ))1τcos(2π(nα+θ))=E𝑑μV1,α,θ,δn2(1λ)τ.\begin{split}\frac{2(1-\lambda)}{\tau}>\frac{2\lambda\cos(2\pi(n\alpha+\theta))}{1-\tau\cos(2\pi(n\alpha+\theta))}=\int Ed\mu_{V_{1},\alpha,\theta,\delta_{n}}\geq\frac{2(1-\lambda)}{\tau}.\end{split}

This is a contradiction. ∎

3.3. Application for the long-range tight-binding model.

Now we consider the long-range tight-binding model (4). By Aubry duality, the dual model of (4) can be written as

(HV3,α,θu)n=un+1+un1+V3(θ+nα)un,(H_{V_{3},\alpha,\theta}u)_{n}=u_{n+1}+u_{n-1}+V_{3}(\theta+n\alpha)u_{n},

where

V3(θ)=2λj0ep|j|cos(2πjθ)=2λ(j0ep|j|e2πjθi)=4λe2p+epcos2πθ1+e2p2epcos2πθ.\begin{split}V_{3}(\theta)&=\frac{2}{\lambda}\sum_{j\neq 0}e^{-p|j|}\cos(2\pi j\theta)=\frac{2}{\lambda}\Im(\sum_{j\neq 0}e^{-p|j|}e^{2\pi j\theta i})\\ &=\frac{4}{\lambda}\frac{-e^{-2p}+e^{-p}\cos 2\pi\theta}{1+e^{-2p}-2e^{-p}\cos 2\pi\theta}.\end{split}

Furthermore, by Aubry duality [45], we have

Σ(H^V3,α,θ)=λ2Σ(V3).\Sigma(\widehat{H}_{V_{3},\alpha,\theta})=\frac{\lambda}{2}\Sigma(V_{3}).

Then as a direct consequence of Lemma 3.9, one obtains:

Corollary 3.13.

Suppost that λ0\lambda\neq 0, α\\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}} and p>0p>0. Then we have

L(α,SEV3)=max{ln|ep2(|E+2λ|+(E+2λ)24)|,0},EΣ(V3).L(\alpha,S_{E}^{V_{3}})=\max\{\ln|\frac{e^{-p}}{2}(|E+\frac{2}{\lambda}|+\sqrt{(E+\frac{2}{\lambda})^{2}-4})|,0\},\ E\in\Sigma(V_{3}).

Moreover, we have the following:

  • EΣ+(V3)E\in\Sigma_{+}(V_{3}) if and only if sgn(λ)E>2coshp2|λ|sgn(\lambda)E>2\cosh p-\frac{2}{|\lambda|},

  • EΣc(V3)E\in\Sigma_{c}(V_{3}) if and only if sgn(λ)E=2coshp2|λ|sgn(\lambda)E=2\cosh p-\frac{2}{|\lambda|},

  • EΣ(V3)E\in\Sigma_{-}(V_{3}) if and only if sgn(λ)E<2coshp2|λ|sgn(\lambda)E<2\cosh p-\frac{2}{|\lambda|}.

Proof.

Note that

4λe2p+epcos(2πnα+θ)1+e2p2epcos(2πnα+θ)=E0+2λ~cos(2πnα+θ)1τcos(2πnα+θ),\frac{4}{\lambda}\frac{-e^{-2p}+e^{-p}\cos(2\pi n\alpha+\theta)}{1+e^{-2p}-2e^{-p}\cos(2\pi n\alpha+\theta)}=E_{0}+2\widetilde{\lambda}\frac{\cos(2\pi n\alpha+\theta)}{1-\tau\cos(2\pi n\alpha+\theta)},

thus we can rewrite HV3,α,θu=EuH_{V_{3},\alpha,\theta}u=Eu as

un+1+un1+2λ~cos(2πnα+θ)1τcos(2πnα+θ)un=(EE0)un,u_{n+1}+u_{n-1}+2\widetilde{\lambda}\frac{\cos(2\pi n\alpha+\theta)}{1-\tau\cos(2\pi n\alpha+\theta)}u_{n}=(E-E_{0})u_{n},

where

E0=2epλcoshp,λ~=tanhpλcoshp,τ=1coshp.E_{0}=-\frac{2e^{-p}}{\lambda\cosh p},\qquad\widetilde{\lambda}=\frac{\tanh p}{\lambda\cosh p},\qquad\tau=\frac{1}{\cosh p}.

Then the results follow from Lemma 3.9 directly. ∎

3.4. Application to the “Peaky” potential

Corollary 3.14.

Suppose that K0K\neq 0 and α\\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}}. Then we have

L(α,SEV4)=max{ln|K2K+1|E|+E241+1(2K2K+1)2|,0},EΣ(V4).L(\alpha,S_{E}^{V_{4}})=\max\{\ln|\frac{K}{2K+1}\frac{|E|+\sqrt{E^{2}-4}}{1+\sqrt{1-(\frac{2K}{2K+1})^{2}}}|,0\},\quad E\in\Sigma(V_{4}).

Moreover, the following holds true:

  1. (1)

    If λK(2K+1)2<1KE¯(V4)2K+1\frac{\lambda K}{(2K+1)^{2}}<1-\frac{K\overline{E}(V_{4})}{2K+1}, then we have

    Σ(V4)=Σ(V4).\Sigma(V_{4})=\Sigma_{-}(V_{4})\neq\emptyset.
  2. (2)

    If 1λE¯(V4)2K+1<λK(2K+1)2<1KE¯(V4)2K+11-\frac{\lambda\overline{E}(V_{4})}{2K+1}<\frac{\lambda K}{(2K+1)^{2}}<1-\frac{K\underline{E}(V_{4})}{2K+1}, then we have

    (17) Σ(V4)[E¯(V4),2+1K)=Σ(V4),Σ(V4)(2+1K,E¯(V4)]=Σ+(V4).\begin{split}\Sigma(V_{4})\cap[\underline{E}(V_{4}),2+\frac{1}{K})&=\Sigma_{-}(V_{4})\neq\emptyset,\\ \Sigma(V_{4})\cap(2+\frac{1}{K},\overline{E}(V_{4})]&=\Sigma_{+}(V_{4})\neq\emptyset.\end{split}
  3. (3)

    If λK(2K+1)2>1λE¯(V4)2K+1\frac{\lambda K}{(2K+1)^{2}}>1-\frac{\lambda\underline{E}(V_{4})}{2K+1}, then we have

    Σ(V4)=Σ+(V4).\Sigma(V_{4})=\Sigma_{+}(V_{4})\neq\emptyset.

In particular, if 1<λK(2K+1)<4K+11<\frac{\lambda K}{(2K+1)}<4K+1, then (17) holds.

Proof.

Note one can rewrite the “Peaky” potential as

V4(θ)=λ1+4Ksin2(πθ)=λ2K+1+2λK(2K+1)2cos2πθ12K2K+1cos2πθV_{4}(\theta)=\frac{\lambda}{1+4K\sin^{2}(\pi\theta)}=\frac{\lambda}{2K+1}+\frac{2\lambda K}{(2K+1)^{2}}\frac{\cos 2\pi\theta}{1-\frac{2K}{2K+1}\cos 2\pi\theta}

Thus the corresponding Schrödinger operator is in fact the GAA model:

un+1+un1+2λ~cos2π(nα+θ)1τcos2π(nα+θ)un=(Eλ2K+1)un,u_{n+1}+u_{n-1}+2\widetilde{\lambda}\frac{\cos 2\pi(n\alpha+\theta)}{1-\tau\cos 2\pi(n\alpha+\theta)}u_{n}=(E-\frac{\lambda}{2K+1})u_{n},

with λ~=λK(2K+1)2,τ=2K2K+1,\widetilde{\lambda}=\frac{\lambda K}{(2K+1)^{2}},\tau=\frac{2K}{2K+1}, and a shift of energy λ2K+1\frac{\lambda}{2K+1}. Then the results follows from Lemma 3.9, Corollary 3.10 and Lemma 3.12. ∎

4. Pure absolutely continuous spectrum

For any α\\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}}, κ+\kappa\in{\mathbb{Z}}^{+}, and for any VjChω(𝕋,)V^{j}\in C^{\omega}_{h}({\mathbb{T}},{\mathbb{R}}), j=0,1,κ1,j=0,1,\cdots\kappa-1, we consider the almost-periodic Schrödinger operator

(18) (HV,α,θu)n=un+1+un1+Vθ(n)un,(H_{V,\alpha,\theta}u)_{n}=u_{n+1}+u_{n-1}+V_{\theta}(n)u_{n},

where the potential takes the form

Vθ(n)=Vj(θ+nα)njmodκ.V_{\theta}(n)=V^{j}(\theta+n\alpha)\qquad n\equiv j\mod\kappa.

The case κ=1\kappa=1 is the one-frequency quasi-periodic Schrödinger operator, while the case κ2\kappa\geq 2, including the quasi-periodic mosaic model, is almost periodic with frequency modulo 𝕋×κ\mathbb{T}\times\mathbb{Z}_{\kappa}, consequently (18) induces an almost-periodic Schrödinger cocycle (Tα,SEV)(T_{\alpha},S_{E}^{V}), and associate with it, one can consider the quasi-periodic cocycle

(κα,DEV)\displaystyle(\kappa\alpha,D_{E}^{V}) :=\displaystyle:= (κα,SEV(θ,κ1)SEV(θ,0))\displaystyle(\kappa\alpha,S_{E}^{V}(\theta,\kappa-1)\cdots S_{E}^{V}(\theta,0))
=\displaystyle= (EVκ1(θ+(κ1)α)110)(EV0(θ)110).\displaystyle\begin{pmatrix}E-V^{\kappa-1}(\theta+(\kappa-1)\alpha)&-1\\ 1&0\end{pmatrix}\cdots\begin{pmatrix}E-V^{0}(\theta)&-1\\ 1&0\end{pmatrix}.

In general, for this kind of potentials, we have the following:

Theorem 4.1.

If αDC\alpha\in DC, κ+\kappa\in{\mathbb{Z}}^{+}, then for any θ\theta\in{\mathbb{R}}, HV,α,θH_{V,\alpha,\theta} is purely absolutely continuous in the set

𝒮={EΣ(V)|(κα,DEV)is almost reducible}.\mathcal{S}=\{E\in\Sigma(V)\quad|\quad(\kappa\alpha,D_{E}^{V})\quad\text{is almost reducible}\}.

This result establish the link between absolutely continuous spectrum of almost periodic operators and almost reducibility of its iterated cocycle. If κ=1\kappa=1, then this is well-known result of Avila [6]. However, the real challenge is the case κ2\kappa\geq 2, and so far none of existence approaches can be applied to this situation. As a matter of fact, there are basically two existing approach in proving pure ac spectrum based on almost reducibility, which are developed by Eliasson [33] and Avila [7] separately. Eliasson’s proof [33] based on parameterized KAM, and to study almost reducibility of (α,A(EE¯)eFE(θ))(\alpha,A(E-\bar{E})e^{F_{E}(\theta)}), his approach strongly depends on the fact the constant part A(E)A(E) is non-degenerated, i.e.,

(19) ddEρ(α,A(E))|E=E¯0,\frac{d}{dE}\rho(\alpha,A(E))|_{E=\bar{E}}\neq 0,

as also explored by Bjerklöv and Krikorian (Theorem 2.2 of [21]), however neither in the case λ\lambda is large (one may first apply Lemma 3.3 to reduce it to local situation) nor in the case λ\lambda is small enough, (19) is satisfied. Avila’s approach [7] is duality based where the desired almost reducibility estimates are provided by almost localization of the dual operator. However, it is obvious that we do not have the duality approach for κ2\kappa\geq 2 case.

Now we give the proof of Theorem 4.1. For given α,κ,V\alpha,\kappa,V, let

𝒜={E|(κα,DEV)is almost reducible},\mathcal{AR}=\{E\in{\mathbb{R}}\quad|\quad(\kappa\alpha,D_{E}^{V})\quad\text{is almost reducible}\},

which is an open set since almost reducibility cocycles is an open set in (\)×Cω(𝕋,SL(2,))({\mathbb{R}}\backslash{\mathbb{Q}})\times C^{\omega}({\mathbb{T}},SL(2,{\mathbb{R}})) (Corollary 1.3 of [4]), i.e.,

𝒜=j=1JIj=j=1J(aj,bj),\mathcal{AR}=\cup_{j=1}^{J}I_{j}=\cup_{j=1}^{J}(a_{j},b_{j}),

where JJ\in{\mathbb{N}} or J=J=\infty. Note that the operators we consider are bounded, thus we only need to prove purely absolutely continuous spectrum in j=1J\cup_{j=1}^{J} (aj,bj)(M,M)(a_{j},b_{j})\cap(-M,M) for some M>0M>0. Take any interval in j=1J(aj,bj)(M,M),\cup_{j=1}^{J}(a_{j},b_{j})\cap(-M,M), omiting jj and denote it by (a,b)(a,b). To prove HV,α,θH_{V,\alpha,\theta} has purely absolutely continuous spectrum in the bounded interval (a,b)(a,b), one only need to prove that HV,α,θH_{V,\alpha,\theta} has purely absolutely continuous spectrum in

𝒮(δ0)=Σ(V)[a+δ0,bδ0]\mathcal{S}(\delta_{0})=\Sigma(V)\cap[a+\delta_{0},b-\delta_{0}]

for any sufficiently small δ0>0\delta_{0}>0.

Now we give the full proof. First we need the following result, which states that for any E𝒮(δ0)E\in\mathcal{S}(\delta_{0}), then after a finite number (that is uniform with respect to E𝒮(δ0)E\in\mathcal{S}(\delta_{0})) of conjugation steps, one can reduce the cocycle to the perturbative regime.

Lemma 4.2.

For any ϵ0>0,\epsilon_{0}>0, α\,\alpha\in{\mathbb{R}}\backslash{\mathbb{Q}}, there exist h¯=h¯(α)>0\bar{h}=\bar{h}(\alpha)>0 and Γ=Γ(α,ϵ0)>0\Gamma=\Gamma(\alpha,\epsilon_{0})>0 such that for any E𝒮(δ0)E\in\mathcal{S}(\delta_{0}), there exist ΦECh¯ω(𝕋,PSL(2,))\Phi_{E}\in C^{\omega}_{\bar{h}}(\mathbb{T},PSL(2,\mathbb{R})) with ΦEh¯<Γ\|\Phi_{E}\|_{\bar{h}}<\Gamma such that

ΦE(θ+κα)1DEV(θ)ΦE(θ)=RΦEefE(θ)\Phi_{E}(\theta+\kappa\alpha)^{-1}D^{V}_{E}(\theta)\Phi_{E}(\theta)=R_{\Phi_{E}}e^{f_{E}(\theta)}

with fEh¯<ϵ0,\left\|f_{E}\right\|_{\bar{h}}<\epsilon_{0}, |degΦE|C|lnΓ|\left|\operatorname{deg}\Phi_{E}\right|\leq C|\ln\Gamma| for some constant C=C= C(V,α)>0.C(V,\alpha)>0.

Proof.

Similar proof first appeared in Proposition 5.2 of [67], we give the proof just for completeness. The crucial fact for this proposition is that we can choose h¯(α)\bar{h}(\alpha) to be independent of EE and ϵ0,\epsilon_{0}, and choose ϵ0\epsilon_{0} to be independent of EE.

For any E𝒮(δ0),E\in\mathcal{S}(\delta_{0}), there exists h0=h0(E,α)>0,h_{0}=h_{0}(E,\alpha)>0, such that

ΦE(+κα)1DEV()ΦE()=Rϕ(E)+FE(),\Phi_{E}(\cdot+\kappa\alpha)^{-1}D_{E}^{V}(\cdot)\Phi_{E}(\cdot)=R_{\phi(E)}+F_{E}(\cdot),

with FEh0<ϵ0/2\left\|F_{E}\right\|_{h_{0}}<\epsilon_{0}/2 and ΦEh0<Γ~\left\|\Phi_{E}\right\|_{h_{0}}<\tilde{\Gamma} for some Γ~=Γ(α,ϵ0,E)>0.\tilde{\Gamma}=\Gamma(\alpha,\epsilon_{0},E)>0. Note that for any E,E𝒮(δ0)E,E^{\prime}\in\mathcal{S}(\delta_{0}),

DEVDEVh0<C(V,h0)|EE|.||D_{E}^{V}-D_{E^{\prime}}^{V}||_{h_{0}}<C(V,h_{0})|E-E^{\prime}|.

Thus for any E,E^{\prime}\in\mathbb{R}, one has

ΦE(+κα)1DEV()ΦE()Rϕ(E)h0<ϵ02+C|EE|ΦEh02.\left\|\Phi_{E}(\cdot+\kappa\alpha)^{-1}D_{E^{\prime}}^{V}(\cdot)\Phi_{E}(\cdot)-R_{\phi(E)}\right\|_{h_{0}}<\frac{\epsilon_{0}}{2}+C\left|E-E^{\prime}\right|\left\|\Phi_{E}\right\|_{h_{0}}^{2}.

It follows that with the same ΦE,\Phi_{E}, we have

ΦE(+κα)1DEV()ΦE()Rϕ(E)h0<ϵ0\left\|\Phi_{E}(\cdot+\kappa\alpha)^{-1}D_{E^{\prime}}^{V}(\cdot)\Phi_{E}(\cdot)-R_{\phi(E)}\right\|_{h_{0}}<\epsilon_{0}

for any energy EE^{\prime} in a neighborhood 𝒰(E)\mathcal{U}(E) of EE. Since 𝒮(δ0)\mathcal{S}(\delta_{0}) is compact, by compactness argument, we can select h0(E,α),Γ(α,ϵ0,E)>0h_{0}(E,\alpha),\Gamma(\alpha,\epsilon_{0},E)>0 to be independent of the energy EE. ∎

Once having Lemma 4.2, one can apply the KAM scheme (Proposition A.1) to get precise control of the growth of the cocycles in the resonant sets. We inductively give the parameters, for any h¯>h~>0\bar{h}>\tilde{h}>0, γ>0,σ>0\gamma>0,\sigma>0, define

h0=h¯,ϵ0D0(γκσ,σ)(h¯h~8)C0σ,\displaystyle h_{0}=\bar{h},\qquad\epsilon_{0}\leq D_{0}(\frac{\gamma}{\kappa^{\sigma}},\sigma)(\frac{\bar{h}-\tilde{h}}{8})^{C_{0}\sigma},

where D0=D0(γ,σ)D_{0}=D_{0}(\gamma,\sigma) and C0C_{0} are constant defined in Proposition A.1, and define

ϵj=ϵ02j,hjhj+1=h¯h¯+h~24j+1,Nj=2|lnϵj|hjhj+1.\epsilon_{j}=\epsilon_{0}^{2^{j}},\quad h_{j}-h_{j+1}=\frac{\bar{h}-\frac{\bar{h}+\tilde{h}}{2}}{4^{j+1}},\quad N_{j}=\frac{2|\ln\epsilon_{j}|}{h_{j}-h_{j+1}}.

Then we have the following:

Proposition 4.3.

Let αDC(γ,σ)\alpha\in DC(\gamma,\sigma). Then there exists BjChjω(𝕋,PSL(2,))B_{j}\in C_{h_{j}}^{\omega}\left(\mathbb{T},PSL(2,\mathbb{R})\right) with |degBj|2Nj1|\deg B_{j}|\leq 2N_{j-1}, such that

Bj1(θ+κα)RΦEefE(θ)Bj(θ)=Aj(E)efj(θ),B_{j}^{-1}(\theta+\kappa\alpha)R_{\Phi_{E}}e^{f_{E}(\theta)}B_{j}(\theta)=A_{j}(E)e^{f_{j}(\theta)},

with estimates Bj0|lnϵj1|4σ\left\|B_{j}\right\|_{0}\leq|\ln\epsilon_{j-1}|^{4\sigma}, fjhjϵj.\left\|f_{j}\right\|_{h_{j}}\leq\epsilon_{j}. Moreover, for any 0<|n|Nj10<|n|\leq N_{j-1}, denote

Λj(n)={E𝒮(δ0):2ρ(κα,Aj1(E))n,κα𝕋<ϵj1115}.\begin{split}\Lambda_{j}(n)=\left\{E\in\mathcal{S}(\delta_{0}):\|2\rho(\kappa\alpha,A_{j-1}(E))-\langle n,\kappa\alpha\rangle\|_{{\mathbb{T}}}<\epsilon_{j-1}^{\frac{1}{15}}\right\}.\end{split}

If EKj:=|n|=1Nj1Λj(n)E\in K_{j}:=\cup_{|n|=1}^{N_{j-1}}\Lambda_{j}(n), then Aj(E)A_{j}(E) can be written as

Aj(E)=M1exp(itjvjv¯jitj)M,A_{j}(E)=M^{-1}\exp\left(\begin{array}[]{cc}{it_{j}}&{v_{j}}\\ {\bar{v}_{j}}&{-it_{j}}\end{array}\right)M,

where

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

with estimates

|tj|<ϵj1116,|vj|<ϵj11516.|t_{j}|<\epsilon^{\frac{1}{16}}_{j-1},\ |v_{j}|<\epsilon_{j-1}^{\frac{15}{16}}.
Proof.

We prove Proposition 4.3 by iteration. In the proof, we will omit its dependence on the energy EE for simplicity. Assume that we have completed the jj-th step and are at the (j+1)(j+1)-th KAM step, i.e. we already construct BjChjω(𝕋,PSL(2,))B_{j}\in C_{h_{j}}^{\omega}\left(\mathbb{T},PSL(2,\mathbb{R})\right) such that

Bj1(θ+κα)RΦEefE(θ)Bj(θ)=Ajefj(θ),B_{j}^{-1}(\theta+\kappa\alpha)R_{\Phi_{E}}e^{f_{E}(\theta)}B_{j}(\theta)=A_{j}e^{f_{j}(\theta)},

with estimates

fjhjϵj,|degBj|2Nj1.||f_{j}||_{h_{j}}\leq\epsilon_{j},\ |\deg B_{j}|\leq 2N_{j-1}.

Note αDC(γ,σ)\alpha\in DC(\gamma,\sigma) implies καDC(γκσ,σ)\kappa\alpha\in DC(\frac{\gamma}{\kappa^{\sigma}},\sigma). Then by our selection of ϵ0\epsilon_{0} (see also Remark A.2), one can check that

ϵjD0AjC0(hjhj+1)C0σ.\epsilon_{j}\leq\frac{D_{0}}{\left\|A_{j}\right\|^{C_{0}}}\left(h_{j}-h_{j+1}\right)^{C_{0}\sigma}.

Indeed, ϵj\epsilon_{j} on the left side of the inequality decays super-exponentially with j,j, while (hjhj+1)C0σ\left(h_{j}-h_{j+1}\right)^{C_{0}\sigma} on the right side decays exponentially with jj. Thus, Proposition A.1 can be applied iteratively, consequently one can construct

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

such that

B¯j+11(θ+κα)Ajefj(θ)B¯j+1(θ)=Aj+1efj+1(θ).\bar{B}_{j+1}^{-1}(\theta+\kappa\alpha)A_{j}e^{f_{j}(\theta)}\bar{B}_{j+1}(\theta)=A_{j+1}e^{f_{j+1}(\theta)}.

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

Bj+11(θ+κα)RΦEefE(θ)Bj+1(θ)=Aj+1efj+1(θ).B_{j+1}^{-1}(\theta+\kappa\alpha)R_{\Phi_{E}}e^{f_{E}(\theta)}B_{j+1}(\theta)=A_{j+1}e^{f_{j+1}(\theta)}.

Moreover, according to the resonance relation, we can distinguish the following two cases:

Non-resonant case: If E|n|=1NjΛj+1(n)E\notin\cup_{|n|=1}^{N_{j}}\Lambda_{j+1}(n), i.e. for any nn\in\mathbb{Z} with 0<|n|Nj,0<|n|\leqslant N_{j}, we have

2ρ(κα,Aj)n,κα/ϵj115,\left\|2\rho(\kappa\alpha,A_{j})-\left<n,\kappa\alpha\right>\right\|_{\mathbb{R}/\mathbb{Z}}\geqslant\epsilon_{j}^{\frac{1}{15}},

then by Proposition A.1, we have

(20) Aj+1Ajϵj,B¯j+1idhj+1ϵj12,fj+1hj+1ϵj+1,\|A_{j+1}-A_{j}\|\leq\epsilon_{j},\quad\left\|\bar{B}_{j+1}-id\right\|_{h_{j+1}}\leq\epsilon_{j}^{\frac{1}{2}},\quad\left\|f_{j+1}\right\|_{h_{j+1}}\leq\epsilon_{j+1},

which implies that

|degBj+1|=|degBj|2Nj12Nj.|\deg B_{j+1}|=|\deg B_{j}|\leq 2N_{j-1}\leq 2N_{j}.

Resonant case: If E|n|=1NjΛj+1(n)E\in\cup_{|n|=1}^{N_{j}}\Lambda_{j+1}(n), i.e. there exists nn with 0<|n|Nj0<\left|n\right|\leq N_{j} such that

2ρ(κα,Aj)n,κα/<ϵj115.\|2\rho(\kappa\alpha,A_{j})-\langle n,\kappa\alpha\rangle\|_{\mathbb{R}/\mathbb{Z}}<\epsilon_{j}^{\frac{1}{15}}.

By Proposition A.1, we have

fj+1hj+1ϵjehj+1ϵj118σϵj+1,\left\|f_{j+1}\right\|_{h_{j+1}}\leq\epsilon_{j}e^{-h_{j+1}\epsilon_{j}^{-\frac{1}{18\sigma}}}\leq\epsilon_{j+1},

and the conjugacy satisfy

(21) B¯j+10<κσγ|n|σ,degB¯j=n,\left\|\bar{B}_{j+1}\right\|_{0}<\frac{\kappa^{\sigma}}{\gamma}|n|^{\sigma},\quad\deg\bar{B}_{j}=n,

which implies that

|degBj+1|=|degBj+degB¯j|2Nj1+Nj2Nj.|\deg B_{j+1}|=|\deg B_{j}+\deg\bar{B}_{j}|\leq 2N_{j-1}+N_{j}\leq 2N_{j}.

Moreover, we can write

Aj+1=M1exp(itj+1vj+1v¯j+1itj+1)MA_{j+1}=M^{-1}\exp\left(\begin{array}[]{cc}{it_{j+1}}&{v_{j+1}}\\ {\bar{v}_{j+1}}&{-it_{j+1}}\end{array}\right)M

with estimates

|tj+1|<ϵj116,|vj+1|<ϵj1516.|t_{j+1}|<\epsilon_{j}^{\frac{1}{16}},\ |v_{j+1}|<\epsilon_{j}^{\frac{15}{16}}.

Finally, we are left to prove

Bj+10|lnϵj|4σ.||B_{j+1}||_{0}\leq|\ln\epsilon_{j}|^{4\sigma}.

To estimate this, we need more detailed analysis on the resonances. Assume that there are at least two resonant steps, say the (mi+1)th(m_{i}+1)^{th} and (mi+1+1)th(m_{i+1}+1)^{th}. At the (mi+1+1)th(m_{i+1}+1)^{th}-step, the resonance condition implies

2ρ(κα,Ami+1)nmi+1,κα/<ϵmi+1115,\|2\rho(\kappa\alpha,A_{m_{i+1}})-\langle n_{m_{i+1}},\kappa\alpha\rangle\|_{\mathbb{R}/\mathbb{Z}}<\epsilon_{m_{i+1}}^{\frac{1}{15}},

hence by αDC(γ,σ)\alpha\in DC(\gamma,\sigma), we have

(22) |ρ(κα,Ami+1(E))|γ2|κnmi+1|σϵmi+1115γ3|κnmi+1|σ.|\rho(\kappa\alpha,A_{m_{i+1}}(E))|\geq\frac{\gamma}{2|\kappa n_{m_{i+1}}|^{\sigma}}-\epsilon_{m_{i+1}}^{\frac{1}{15}}\geq\frac{\gamma}{3|\kappa n_{m_{i+1}}|^{\sigma}}.

On the other hand, according to Proposition A.1, after the (mi+1)th(m_{i}+1)^{th}-step, |ρ(κα,Ami(E))|ϵmi116|\rho(\kappa\alpha,A_{m_{i}}(E))|\leq\epsilon_{m_{i}}^{\frac{1}{16}}. Then (20) implies that |ρ(κα,Ami+1(E))|2ϵmi116,|\rho(\kappa\alpha,A_{m_{i+1}}(E))|\leq 2\epsilon_{m_{i}}^{\frac{1}{16}}, since by our selection, between (mi+1)th(m_{i}+1)^{th} and (mi+1+1)th(m_{i+1}+1)^{th} step, there are no resonant steps. Thus by (22) and |nmi|Nmi=2|lnϵmi|hmihmi+1|n_{m_{i}}|\leq N_{m_{i}}=\frac{2|\ln\epsilon_{m_{i}}|}{h_{m_{i}}-h_{m_{i}+1}}, we have

(23) |nmi+1|12κγ1σϵmi116σ>|nmi|2.|n_{m_{i+1}}|\geq\frac{1}{2\kappa}\gamma^{\frac{1}{\sigma}}\epsilon_{m_{i}}^{-\frac{1}{16\sigma}}>|n_{m_{i}}|^{2}.

Assuming that there are ss resonant steps, associated with integers vectors

nm1,,nms, 0<|nmi|Nmi,i=1,,s,n_{m_{1}},...,n_{m_{s}}\in{\mathbb{Z}},\ \ 0<|n_{m_{i}}|\leq N_{m_{i}},\ \ i=1,...,s,

By (20), (21) and (23), we have

Bj+102i=1sB¯mi02i=1sκσγ|nmi|σ<2(κσγ)s|nms|σ(1+12+12s)<2(κσγ)s|nms|2σ<|lnϵms|4σ<|lnϵj|4σ.\begin{split}||B_{j+1}||_{0}&\leq 2\prod_{i=1}^{s}||\bar{B}_{m_{i}}||_{0}\leq 2\prod_{i=1}^{s}\frac{\kappa^{\sigma}}{\gamma}|n_{m_{i}}|^{\sigma}\\ &<2(\frac{\kappa^{\sigma}}{\gamma})^{s}|n_{m_{s}}|^{\sigma(1+\frac{1}{2}+\cdots\frac{1}{2^{s}})}\\ &<2(\frac{\kappa^{\sigma}}{\gamma})^{s}|n_{m_{s}}|^{2\sigma}<|\ln\epsilon_{m_{s}}|^{4\sigma}<|\ln\epsilon_{j}|^{4\sigma}.\end{split}

We thus finish the proof.

Remark 4.4.

As we noted in Proposition A.1, in the resonant case, the new perturbation can be chosen as ϵeh+ϵ118τ\epsilon e^{-h_{+}\epsilon^{-\frac{1}{18\tau}}}, which is much smaller than ϵ2\epsilon^{2}. However, here we just choose ϵj+1=ϵj2\epsilon_{j+1}=\epsilon_{j}^{2}, otherwise if the perturbation ϵj(E)\epsilon_{j}(E) depends on EE (due to the fact the resonances depend on EE), one cann’t give a good stratification of the energies in the spectrum.

In the construction, KjK_{j} just means the cocycle (κα,RΦEefE(θ))(\kappa\alpha,R_{\Phi_{E}}e^{f_{E}(\theta)}) is resonant in the jj-th KAM step. If EKjE\in K_{j}, then we have the following characterization of its IDS and the growth of the cocycles in the resonant sets:

Lemma 4.5.

Assume that αDC(γ,σ)\alpha\in DC(\gamma,\sigma), EKj,E\in K_{j}, then there exists n~j\tilde{n}_{j}\in\mathbb{Z} with 0<|n~j|<2Nj10<|\tilde{n}_{j}|<2N_{j-1} such that

(24) κNV(E)+n~j,κα/2ϵj1115.\|\kappa N_{V}(E)+\langle\tilde{n}_{j},\kappa\alpha\rangle\|_{\mathbb{R}/\mathbb{Z}}\leqslant 2\epsilon_{j-1}^{\frac{1}{15}}.

Moreover, we have

sup0sϵj118(DEV)s04Γ2|lnϵj1|8σ.\sup_{0\leq s\leq\epsilon_{j-1}^{-\frac{1}{8}}}\|(D_{E}^{V})_{s}\|_{0}\leq 4\Gamma^{2}|\ln\epsilon_{j-1}|^{8\sigma}.
Proof.

First by Lemma 4.2 there exist ΦECω(𝕋,PSL(2,))\Phi_{E}\in C^{\omega}(\mathbb{T},PSL(2,\mathbb{R})) with |degΦE|C|lnΓ|\left|\deg\Phi_{E}\right|\leq C|\ln\Gamma| such that

ΦE(θ+κα)1DEV(θ)ΦE(θ)=RΦEefE(θ)\Phi_{E}(\theta+\kappa\alpha)^{-1}D^{V}_{E}(\theta)\Phi_{E}(\theta)=R_{\Phi_{E}}e^{f_{E}(\theta)}

Furthermore by Proposition 4.3, there exist Bj1Chj1ω(𝕋,PSL(2,))B_{j-1}\in C_{h_{j-1}}^{\omega}\left(\mathbb{T},PSL(2,\mathbb{R})\right) with |degBj1|2Nj2|\deg B_{j-1}|\leq 2N_{j-2} such that

Bj1(θ+κα)1ΦE(θ+κα)1DEV(θ)ΦE(θ)Bj1(θ)=Aj1efj1(θ),B_{j-1}(\theta+\kappa\alpha)^{-1}\Phi_{E}(\theta+\kappa\alpha)^{-1}D^{V}_{E}(\theta)\Phi_{E}(\theta)B_{j-1}(\theta)=A_{j-1}e^{f_{j-1}(\theta)},

and for any EKjE\in K_{j}, we have

(25) 2ρ(κα,Aj1)n,κα/ϵj1115,||2\rho\left(\kappa\alpha,A_{{j-1}}\right)-\left\langle n^{\prime},\kappa\alpha\right\rangle||_{\mathbb{R}/\mathbb{Z}}\leqslant\epsilon_{j-1}^{\frac{1}{15}},

for some nn^{\prime}\in\mathbb{Z} with 0<|n|Nj10<|n^{\prime}|\leq N_{j-1}.

Thus, we deduce that

(26) 2ρ(κα,DEV)=degBj1+degΦE,κα+2ρ(κα,Aj1efj1(θ)).2\rho(\kappa\alpha,D_{E}^{V})=\left<\deg B_{j-1}+\deg\Phi_{E},\kappa\alpha\right>+2\rho(\kappa\alpha,A_{j-1}e^{f_{j-1}(\theta)}).

Note that fj10ϵj1||f_{j-1}||_{0}\leq\epsilon_{j-1}, we have

(27) |ρ(κα,Aj1efj1(θ))ρ(κα,Aj1)|cϵj112.|\rho(\kappa\alpha,A_{j-1}e^{f_{j-1}(\theta)})-\rho(\kappa\alpha,A_{j-1})|\leq c\epsilon_{j-1}^{\frac{1}{2}}.

Combining (25), (26) and (27) suggest that

2ρ(κα,DEV)degBj1+degΦE,καn,κα/2ϵj1115.||2\rho(\kappa\alpha,D_{E}^{V})-\left<\deg B_{j-1}+\deg\Phi_{E},\kappa\alpha\right>-\left<n^{\prime},\kappa\alpha\right>||_{\mathbb{R}/\mathbb{Z}}\leq 2\epsilon_{j-1}^{\frac{1}{15}}.

Let n~j=degBj1+n+degΦE\tilde{n}_{j}=\deg B_{j-1}+n^{\prime}+\deg\Phi_{E}, then

2ρ(κα,DEV)n~j,κα/2ϵj1115,\|2\rho(\kappa\alpha,D_{E}^{V})-\langle\tilde{n}_{j},\kappa\alpha\rangle\|_{\mathbb{R}/\mathbb{Z}}\leqslant 2\epsilon_{j-1}^{\frac{1}{15}},

with estimate

|n~j|2Nj2+Nj1+C|lnΓ|2Nj1.|\tilde{n}_{j}|\leq 2N_{j-2}+N_{j-1}+C|\ln\Gamma|\leq 2N_{j-1}.

Next we observe that

κρ(Tα,SEV)=ρ(κα,DEV)mod,\kappa\rho(T_{\alpha},S_{E}^{V})=\rho(\kappa\alpha,D_{E}^{V})\ mod\ \mathbb{Z},

consequently by the fact that NV(E)=12ρ(Tα,SEV),N_{V}(E)=1-2\rho(T_{\alpha},S_{E}^{V}), we obtain (24).

On the other hand, note if B,DB,D are small sl(2,)sl(2,\mathbb{R}) matrices, then there exists Esl(2,)E\in sl(2,\mathbb{R}) such that

eBeD=eB+D+E,e^{B}e^{D}=e^{B+D+E},

where EE is a sum of terms at least 2 orders in B,D.B,D. Thus by Proposition 4.3, there exist BjChjω(𝕋,PSL(2,))B_{j}\in C_{h_{j}}^{\omega}\left(\mathbb{T},PSL(2,\mathbb{R})\right) with Bj0|lnϵj1|4σ\|B_{j}\|_{0}\leq|\ln\epsilon_{j-1}|^{4\sigma} such that

Bj(θ+κα)1ΦE(θ+κα)1DEV(θ)ΦE(θ)Bj(θ)=Rtjef~j(θ)B_{j}(\theta+\kappa\alpha)^{-1}\Phi_{E}(\theta+\kappa\alpha)^{-1}D^{V}_{E}(\theta)\Phi_{E}(\theta)B_{j}(\theta)=R_{t_{j}}e^{\tilde{f}_{j}(\theta)}

with estimates |tj|ϵj1116|t_{j}|\leq\epsilon_{j-1}^{\frac{1}{16}}, f~j0ϵj118\|\tilde{f}_{j}\|_{0}\leq\epsilon_{j-1}^{\frac{1}{8}}. This imply that

sup0sϵj118(DEV)s04ΦEh¯2Bj024Γ2|lnϵj1|8σ.\sup_{0\leq s\leq\epsilon_{j-1}^{-\frac{1}{8}}}||(D_{E}^{V})_{s}||_{0}\leq 4\|\Phi_{E}\|^{2}_{\bar{h}}\|B_{j}\|_{0}^{2}\leq 4\Gamma^{2}|\ln\epsilon_{j-1}|^{8\sigma}.

Next we study the regularity of NVN_{V}. The observation here is that while (Tα,SEV)(T_{\alpha},S_{E}^{V}) is not an analytic quasi-periodic cocycle, its iterate (κα,DEV)(\kappa\alpha,D_{E}^{V}) indeed defines an analytic quasi-periodic cocycle, then we one can pass the 12\frac{1}{2}-Hölder continuity of (κα,DEV)(\kappa\alpha,D_{E}^{V}) to (Tα,SEV)(T_{\alpha},S_{E}^{V}), the rest proof is standard.

Lemma 4.6.

Assume that αDC(γ,σ)\alpha\in DC(\gamma,\sigma), then the integrated density of states NVN_{V} is 12\frac{1}{2}-Hölder continuous on 𝒜\mathcal{AR}.

Proof.

The assumption αDC(γ,σ)\alpha\in DC(\gamma,\sigma) implies καDC(γκσ,σ)\kappa\alpha\in DC(\frac{\gamma}{\kappa^{\sigma}},\sigma). Thus for any E𝒮(δ0)E\in\mathcal{S}(\delta_{0}), we apply the following result:

Lemma 4.7.

Let α~DC\tilde{\alpha}\in DC, if (α~,A)(\tilde{\alpha},A) is analytically almost reducible, then for any continuous map B:𝕋SL(2,),B:\mathbb{T}\rightarrow SL(2,\mathbb{C}), we have

|L(α~,A)L(α~,B)|C0BA012|L(\tilde{\alpha},A)-L(\tilde{\alpha},B)|\leqslant C_{0}\|B-A\|_{0}^{\frac{1}{2}}

where C0C_{0} is a constant depending on α\alpha.

Proof.

This is essentially contained in the Corollary 4.6 of [10], see also Proposition 7.1 of [67]. ∎

Consequently, by Lemma 4.7, we have |L(κα,DE+iϵV)L(κα,DEV)|Cϵ12,|L(\kappa\alpha,D_{E+i\epsilon}^{V})-L(\kappa\alpha,D_{E}^{V})|\leq C\epsilon^{\frac{1}{2}}, which implies that |L(Tα,SE+iϵV)L(Tα,SEV)|Cκϵ12.|L(T_{\alpha},S_{E+i\epsilon}^{V})-L(T_{\alpha},S_{E}^{V})|\leq\frac{C}{\kappa}\epsilon^{\frac{1}{2}}. On the other hand, Thouless formula state that

L(α,SEV)=ln|EE|dNV(E),L(\alpha,S_{E}^{V})=\int\ln\left|E-E^{\prime}\right|dN_{V}\left(E^{\prime}\right),

then for every ϵ>0\epsilon>0, we have

|L(Tα,SE+iϵV)L(Tα,SEV)|=12ln(1+ϵ2(EE)2)𝑑NV(E)12EϵE+ϵln(1+ϵ2(EE)2)𝑑NV(E)ln22(NV(E+ϵ)NV(Eϵ)),\begin{split}|L(T_{\alpha},S_{E+i\epsilon}^{V})-L(T_{\alpha},S_{E}^{V})|&=\frac{1}{2}\int\ln\left(1+\frac{\epsilon^{2}}{\left(E-E^{\prime}\right)^{2}}\right)dN_{V}\left(E^{\prime}\right)\\ &\geq\frac{1}{2}\int_{E-\epsilon}^{E+\epsilon}\ln\left(1+\frac{\epsilon^{2}}{\left(E-E^{\prime}\right)^{2}}\right)dN_{V}\left(E^{\prime}\right)\\ &\geq\frac{\ln 2}{2}\left(N_{V}(E+\epsilon)-N_{V}(E-\epsilon)\right),\end{split}

which gives

NV(E+ϵ)NV(Eϵ)Cln2ϵ12.N_{V}(E+\epsilon)-N_{V}(E-\epsilon)\leq\frac{C}{\ln 2}\epsilon^{\frac{1}{2}}.

Since NVN_{V} is locally constant in the complement of Σ(V)\Sigma(V), this means precisely that NVN_{V} is 12\frac{1}{2}-Hölder continuous. ∎

As a consequence of Lemma 4.6, we can show NVN_{V} has a lower bound estimate:

Lemma 4.8.

For any δ0>0\delta_{0}>0 which is small enough, if E𝒮(δ0)E\in\mathcal{S}(\delta_{0}), then for sufficiently small ϵ>0\epsilon>0,

NV(E+ϵ)NV(Eϵ)c(δ0)ϵ32,N_{V}(E+\epsilon)-N_{V}(E-\epsilon)\geq c(\delta_{0})\epsilon^{\frac{3}{2}},

where c(δ0)>0c(\delta_{0})>0 is a small universal constant.

Proof.

The proof is first developed by Avila [4] in the quasi-periodic case, which can be generalized to the general case almost without change, we sketch the proof here just for completeness. Let δ=cϵ3/2\delta=c\epsilon^{3/2}. For any E𝒮(δ0),E\in\mathcal{S}(\delta_{0}), we have L(Tα,SEV)=0,L(T_{\alpha},S_{E}^{V})=0, then by Thouless formula we have

L(Tα,SE+iδV)=12ln(1+δ2|EE|2)𝑑NV(E).L(T_{\alpha},S_{E+i\delta}^{V})=\int\frac{1}{2}\ln(1+\frac{\delta^{2}}{\left|E-E^{\prime}\right|^{2}})\ dN_{V}\left(E^{\prime}\right).

We split the integral into four parts: I1=|EE|δ02I_{1}=\int_{\left|E-E^{\prime}\right|\geq\frac{\delta_{0}}{2}}, I2=ϵ|EE|<δ02I_{2}=\int_{\epsilon\leq\left|E-E^{\prime}\right|<\frac{\delta_{0}}{2}}, I3=ϵ4|EE|<ϵI_{3}=\int_{\epsilon^{4}\leq\left|E-E^{\prime}\right|<\epsilon} and I4=|EE|<ϵ4I_{4}=\int_{\left|E-E^{\prime}\right|<\epsilon^{4}}.

For sufficiently small ϵ>0\epsilon>0, by Lemma 4.6 we have I1<2c2ϵ3δ02,I_{1}<\frac{2c^{2}\epsilon^{3}}{\delta_{0}^{2}}, and

I4=k4ϵk>|EE|ϵk+112ln(1+δ2|EE|2)𝑑NV(E)12k4ϵk2ln(1+c2ϵ12k)ϵ74.\begin{split}I_{4}&=\sum_{k\geq 4}\int_{\epsilon^{k}>\left|E-E^{\prime}\right|\geq\epsilon^{k+1}}\frac{1}{2}\ln(1+\frac{\delta^{2}}{\left|E-E^{\prime}\right|^{2}})\ dN_{V}\left(E^{\prime}\right)\\ &\leq\frac{1}{2}\sum_{k\geq 4}\epsilon^{\frac{k}{2}}\ln(1+c^{2}\epsilon^{1-2k})\leq\epsilon^{\frac{7}{4}}.\end{split}

We also have the estimate

I2\displaystyle I_{2} k=0mek1|EE|<ek12ln(1+δ2|EE|2)𝑑NV(E)\displaystyle\leq\sum_{k=0}^{m}\int_{e^{-k-1}\leq|E-E^{\prime}|<e^{-k}}\frac{1}{2}\ln(1+\frac{\delta^{2}}{\left|E-E^{\prime}\right|^{2}})\ dN_{V}\left(E^{\prime}\right)
k=0m12ek2δ2e2k+2Cc2δ,\displaystyle\leq\sum_{k=0}^{m}\frac{1}{2}e^{-\frac{k}{2}}\delta^{2}e^{2k+2}\leq Cc^{2}\delta,

with m=[lnϵ]m=[-\ln\epsilon]. It follows that I3L(Tα,SE+iδV)140δCc2δ.I_{3}\geq L(T_{\alpha},S_{E+i\delta}^{V})-\frac{1}{40}\delta-Cc^{2}\delta. It is well known that L(Tα,SE+iδV)δ/10L(T_{\alpha},S_{E+i\delta}^{V})\geq\delta/10, for 0<δ<10<\delta<1 [31]. Since the constant cc above is consistent with our choice of δ\delta, we can shrink it such that I3120δI_{3}\geq\frac{1}{20}\delta. Since I3C(NV(E+ϵ)NV(Eϵ))lnϵ1,I_{3}\leq C(N_{V}(E+\epsilon)-N_{V}(E-\epsilon))\ln\epsilon^{-1}, the result follows. ∎

Proof of Theorem 4.1

Let \mathcal{B} be the set of E𝒮(δ0)E\in\mathcal{S}(\delta_{0}) such that the Schrödinger cocycle (Tα,SEV)(T_{\alpha},S_{E}^{V}) is bounded, which equals to the set (κα,DEV)(\kappa\alpha,D_{E}^{V}) is bounded. By Theorem 2.2, it is enough to prove that μV,θ(𝒮(δ0)\)=0\mu_{V,\theta}(\mathcal{S}(\delta_{0})\backslash\mathcal{B})=0 for any θ\theta\in{\mathbb{R}}.

Let \mathcal{R} be the set of E𝒮(δ0)E\in\mathcal{S}(\delta_{0}) such that (κα,DEV)(\kappa\alpha,D_{E}^{V}) is reducible, then \\mathcal{R}\backslash\mathcal{B} only contains EE for which (κα,DEV)(\kappa\alpha,D_{E}^{V}) is analytically reducible to a constant parabolic cocycle. Recall that for any E𝒮(δ0)E\in\mathcal{S}(\delta_{0}), by well-known result of Eliasson [33], if ρ(κα,DEV)\rho(\kappa\alpha,D_{E}^{V}) is rational or Diophantine w.r.t κα\kappa\alpha, then (κα,DEV)(\kappa\alpha,D_{E}^{V}) is reducible. It follows that \\mathcal{R}\backslash\mathcal{B} is countable: indeed for any such EE there exists mm\in\mathbb{Z} such that 2ρ(κα,DEV)=m,αmod.2\rho(\kappa\alpha,D_{E}^{V})=\left<m,\alpha\right>\mod\mathbb{Z}. Moreover, if EE\in\mathcal{R}, then any non-zero solution of HV,α,θu=EuH_{V,\alpha,\theta}u=Eu, satisfies infn|uκn|2+|uκn+1|2>0\inf_{n\in\mathbb{Z}}|u_{\kappa n}|^{2}+|u_{\kappa n+1}|^{2}>0, so there are no eigenvalues in \mathcal{R} and μV,θ(\)=0\mu_{V,\theta}(\mathcal{R}\backslash\mathcal{B})=0. Therefore, it is enough to show that for sufficiently small δ0>0\delta_{0}>0, μV,θ(𝒮(δ0)\)=0.\mu_{V,\theta}(\mathcal{S}(\delta_{0})\backslash\mathcal{R})=0. Note that 𝒮(δ0)\lim supKm\mathcal{S}(\delta_{0})\backslash\mathcal{R}\subset\limsup K_{m}, by Borel-Cantelli Lemma, we only need to prove mμV,θ(K¯m)<\sum_{m}\mu_{V,\theta}(\overline{K}_{m})<\infty.

Let Jm(E)J_{m}(E) be an open ϵm1245\epsilon_{m-1}^{\frac{2}{45}} neighborhood of EKmE\in K_{m}. By Lemma 2.3 and Lemma 4.5, we have

μV,θ(Jm(E))sup0sϵm1245(DEV)s02|Jm(E)|sup0sϵm118(DEV)s02|Jm(E)|C|lnϵm1|16σϵm1245.\begin{split}\mu_{V,\theta}(J_{m}(E))&\leq\sup_{0\leq s\leq\epsilon_{m-1}^{-\frac{2}{45}}}||(D_{E}^{V})_{s}||_{0}^{2}|J_{m}(E)|\\ &\leq\sup_{0\leq s\leq\epsilon_{m-1}^{-\frac{1}{8}}}||(D_{E}^{V})_{s}||_{0}^{2}|J_{m}(E)|\leq C|\ln\epsilon_{m-1}|^{16\sigma}\epsilon_{m-1}^{\frac{2}{45}}.\end{split}

Let l=0rJm(El)\cup_{l=0}^{r}J_{m}(E_{l}) be a finite subcover of K¯m\overline{K}_{m}. By refining this subcover, we can assume that every EE\in\mathbb{R} is contained in at most two different Jm(El)J_{m}(E_{l}).

On the other hand, by Lemma 4.5, if EKmE\in K_{m}, then

κNV(E)+n,κα/2ϵm1115,||\kappa N_{V}(E)+\left<n,\kappa\alpha\right>||_{\mathbb{R/Z}}\leq 2\epsilon_{m-1}^{\frac{1}{15}},

for some |n|<2Nm1|n|<2N_{m-1}. This shows that κNV(Km)\kappa N_{V}({K}_{m}) can be covered by 2Nm12N_{m-1} intervals IsI_{s} of length 2ϵm11152\epsilon_{m-1}^{\frac{1}{15}}. By Lemma 4.8,

κNV(Jm(E))c|Jm(E)|32,\kappa N_{V}(J_{m}(E))\geq c|J_{m}(E)|^{\frac{3}{2}},

thus by our selection |Is|1c|κNV(Jm(E))||I_{s}|\leq\frac{1}{c}|\kappa N_{V}(J_{m}(E))| for any ss and EKmE\in{K}_{m}, there are at most 2([1c]+1)+42([\frac{1}{c}]+1)+4 intervals Jm(El)J_{m}(E_{l}) such that κNV(Jm(El))\kappa N_{V}(J_{m}(E_{l})) intersects IsI_{s}. We conclude that there are at most 2(2([1c]+1)+4)Nm12(2([\frac{1}{c}]+1)+4)N_{m-1} intervals Jm(El)J_{m}(E_{l}) to cover KmK_{m}. Then

μV,θ(K¯m)j=0rμV(Jm(Ej))CNm1|lnϵm1|16σϵm1245<ϵm1145,\mu_{V,\theta}(\overline{K}_{m})\leq\sum_{j=0}^{r}\mu_{V}(J_{m}(E_{j}))\leq CN_{m-1}|\ln\epsilon_{m-1}|^{16\sigma}\epsilon_{m-1}^{\frac{2}{45}}<\epsilon_{m-1}^{\frac{1}{45}},

which gives mμV,θ(K¯m)<\sum_{m}\mu_{V,\theta}(\overline{K}_{m})<\infty.

5. Anderson Localization

In this section, we will prove Anderson localization for GAA model and quasi-periodic mosaic model. We will fix V=V1V=V_{1} (the GAA model) or V2V_{2} (quasi-periodic mosaic model) in this section, and for the quasi-periodic mosaic model, we will just consider κ=2\kappa=2, since the general case follows similarly.

Theorem 5.1.

Suppose that αDC(γ,σ)\alpha\in DC(\gamma,\sigma). Then HV,α,θH_{V,\alpha,\theta} has Anderson Localization in the set

𝒫:=Σ(V){E|L(α,SEV)>0}\mathcal{P}:=\Sigma(V)\cap\{E\in{\mathbb{R}}\quad|\quad L(\alpha,S_{E}^{V})>0\}

for every θΘ\theta\in\Theta, where

Θ=η>0Θ(η):=η>0{θ|2θkαη|k|σk0}.\Theta=\cup_{\eta>0}\Theta(\eta):=\cup_{\eta>0}\{\theta\in{\mathbb{R}}|\quad\|2\theta-k\alpha\|\geq\frac{\eta}{|k|^{\sigma}}\quad\forall k\neq 0\}.

We start with the basic setup going back to [53]. We will use the notation G[n1,n2](n,m)G_{[n_{1},n_{2}]}(n,m) for the Green’s function (HV,α,θE)1(n,m)(H_{V,\alpha,\theta}-E)^{-1}(n,m) of the operator HV,α,θH_{V,\alpha,\theta} restricted to the interval [n1,n2][n_{1},n_{2}] with zero boundary conditions at n11n_{1}-1 and n2+1n_{2}+1. To simplify the notations, we replace L(α,SEV)L(\alpha,S_{E}^{V}) by L(E)L(E), the V,αV,\alpha-dependence of various quantities will be omitted in some cases.

Denote by Mk(θ)M_{k}(\theta) the kk-step transfer-matrix of HV,α,θu=EuH_{V,\alpha,\theta}u=Eu, and denote

Pk(θ)=det[(HV,α,θE)|[0,k1]],Qk(θ)=det[(HV,α,θE)|[1,k]],P_{k}(\theta)=\det[(H_{V,\alpha,\theta}-E)|_{[0,k-1]}],\quad Q_{k}(\theta)=\det[(H_{V,\alpha,\theta}-E)|_{[1,k]}],

then the kk-step transfer-matrix can be written as

Mk(θ)=(1)k(Pk(θ)Qk1(θ)Pk1(θ)Qk2(θ)).M_{k}(\theta)=(-1)^{k}\begin{pmatrix}P_{k}(\theta)&Q_{k-1}(\theta)\\ -P_{k-1}(\theta)&-Q_{k-2}(\theta)\end{pmatrix}.

By Kingman’s Subadditive Ergodic Theorem, Lyapunov exponent satisfies

(28) L(E)=infn11n𝕋lnMn(θ)dθ=limn1nlnMn(θ),L(E)=\inf_{n\geq 1}\frac{1}{n}\int_{\mathbb{T}}\ln\|M_{n}(\theta)\|d\theta=\lim_{n\to\infty}\frac{1}{n}\ln\|M_{n}(\theta)\|,

for almost every θ𝕋\theta\in\mathbb{T}. Moreover, if we recall the following Furman’s result:

Theorem 5.2.

[39] Suppose (X,T)(X,T) is uniquely ergodic. If fn:Xf_{n}:X\to\mathbb{R} are continuous and satisfy fn+m(θ)fn(θ)+fm(Tnθ)f_{n+m}(\theta)\leq f_{n}(\theta)+f_{m}(T^{n}\theta), then

lim supn1nfn(θ)infn11n𝔼(fn)\limsup_{n\to\infty}\frac{1}{n}f_{n}(\theta)\leq\inf_{n\geq 1}\frac{1}{n}\mathbb{E}(f_{n})

for every θX\theta\in X and uniformly on XX.

Then we have uniform growth of the transfer matrix:

Lemma 5.3.

For every EE\in\mathbb{R} and ϵ>0\epsilon>0, there exists k1(E,ϵ)k_{1}(E,\epsilon) such that

Mk(θ)<e(L(E)+ϵ)k\|M_{k}(\theta)\|<e^{(L(E)+\epsilon)k}

for every k>k1(E,ϵ)k>k_{1}(E,\epsilon) and every θ𝕋\theta\in\mathbb{T}.

Proof.

If V=V1,V=V_{1}, then the base dynamics is (𝕋,Rα)(\mathbb{T},R_{\alpha}) which is uniquely ergodic. If V=V2V=V_{2}, the base dynamics is (𝕋×κ,Tα)(\mathbb{T}\times\mathbb{Z}_{\kappa},T_{\alpha}), which is also uniquely ergodic (Theorem 9.1 of [69]). Apply Theorem 5.2 to fn(θ)=lnMn(θ)f_{n}(\theta)=\ln\|M_{n}(\theta)\|, we get desired results. ∎

When the Lyapunov exponent is positive, Lemma 5.3 implies that some of the entries must be exponentially large. These entries in turn appear in a description of the Green’s function of the operator restricted to a finite interval. Namely, by Cramer’s Rule, if we denote

Δm,n(θ)=det[(HV,α,θE)|[m,n]],\Delta_{m,n}(\theta)=\det[(H_{V,\alpha,\theta}-E)|_{[m,n]}],

then for n1,n2=n1+k1n_{1},n_{2}=n_{1}+k-1, and n[n1,n2]n\in[n_{1},n_{2}],

(29) |G[n1,n2](n1,n)|=|Δn+1,n2(θ)Δn1,n2(θ)|,|G[n1,n2](n,n2)|=|Δn1,n1(θ)Δn1,n2(θ)|.\begin{split}|G_{[n_{1},n_{2}]}(n_{1},n)|&=|\frac{\Delta_{n+1,n_{2}}(\theta)}{\Delta_{n_{1},n_{2}}(\theta)}|,\\ |G_{[n_{1},n_{2}]}(n,n_{2})|&=|\frac{\Delta_{n_{1},n-1}(\theta)}{\Delta_{n_{1},n_{2}}(\theta)}|.\end{split}

A useful definition about Green’s function is the following:

Definition 5.4.

[53] Fix EE\in\mathbb{R} and ξ\xi\in\mathbb{R}. A point nn\in\mathbb{Z} will be called (ξ,k)(\xi,k)-regular if there exists an interval [n1,n2][n_{1},n_{2}], n2=n1+k1n_{2}=n_{1}+k-1, containing nn, such that

|G[n1,n2](n,ni)|<eξ|nni|,and|nni|17k;i=1,2.|G_{[n_{1},n_{2}]}(n,n_{i})|<e^{-\xi|n-n_{i}|},\quad and\ |n-n_{i}|\geq\frac{1}{7}k;\ i=1,2.

Otherwise, nn will be called (ξ,k)(\xi,k)-singular.

It is well known that any formal solution uu of the HV,α,θ=EuH_{V,\alpha,\theta}=Eu at a point n[n1,n2]n\in[n_{1},n_{2}] can be reconstructed from the boundary values via

(30) u(n)=G[n1,n2](n,n1)u(n11)G[n1,n2](n,n2)u(n2+1).u(n)=-G_{[n_{1},n_{2}]}(n,n_{1})u(n_{1}-1)-G_{[n_{1},n_{2}]}(n,n_{2})u(n_{2}+1).

This implies that if uEu_{E} is a generalized eigenfunction, then every point nn\in\mathbb{Z} with uE(n)0u_{E}(n)\neq 0 is (ξ,k)(\xi,k)-singular for kk sufficiently large: k>k2(E,ξ,θ,n)k>k_{2}(E,\xi,\theta,n). In the following, we just just assume uE(0)0u_{E}(0)\neq 0 (otherwise replace uE(0)u_{E}(0) by uE(1)u_{E}(1)). Then Theorem 5.1 will follow from the next result:

Proposition 5.5.

Assume that αDC\alpha\in DC, θΘ\theta\in\Theta, L(E)>0L(E)>0. Then for every ϵ>0\epsilon>0, for any |y|>y(α,θ,E,ϵ)|y|>y(\alpha,\theta,E,\epsilon) sufficiently large, there exists k>516|y|,k>\frac{5}{16}|y|, such that yy is (L(E)ϵ,k)(L(E)-\epsilon,k)-regular.

Proof of Theorem 5.1.

It is well known that if every generalized eigenfunction of HV,α,θH_{V,\alpha,\theta} decays exponentially, then the operator HV,α,θH_{V,\alpha,\theta} displays Anderson localization. Let E𝒫E\in\mathcal{P} be a generalized eigenvalue of HV,α,θH_{V,\alpha,\theta}, and denote the corresponding generalized eigenfunction by uEu_{E}. Let ϵ\epsilon small enough, by (30) and Proposition 5.5, if |y|>y(α,θ,E,ϵ)|y|>y(\alpha,\theta,E,\epsilon) the point yy is (L(E)ϵ,k)(L(E)-\epsilon,k)-regular for some k>516|y|k>\frac{5}{16}|y|. Thus, there exists an interval [n1,n2][n_{1},n_{2}] of length kk containing yy such that 17k|yni|67k,\frac{1}{7}k\leq|y-n_{i}|\leq\frac{6}{7}k, and

|G[n1,n2](y,ni)|<e(L(E)ϵ)|yni|,i=1,2.|G_{[n_{1},n_{2}]}(y,n_{i})|<e^{-(L(E)-\epsilon)|y-n_{i}|},\quad i=1,2.

Using (30), we obtain that

|uE(y)|2C(E)(2|y|+1)eL(E)ϵ7keL(E)ϵ24|y|.\begin{split}|u_{E}(y)|\leq 2C(E)(2|y|+1)e^{-\frac{L(E)-\epsilon}{7}k}\leq e^{-\frac{L(E)-\epsilon}{24}|y|}.\end{split}

This implies exponential decay of the eigenfunction if ϵ\epsilon is chosen small enough. ∎

5.1. Anderson localization for the GAA model.

For the GAA model, the basic observation is that Qk(θ)=Pk(θ+α)Q_{k}(\theta)=P_{k}(\theta+\alpha), then all the elements of Mk(θ)M_{k}(\theta) can be expressed by Qk(θ)Q_{k}(\theta). The key observation is the following:

Lemma 5.6.

We have

Qk(θ)=Rk(cos2π(θ+k+12α))j=1k(1τcos2π(θ+jα)),Q_{k}(\theta)=\frac{R_{k}(\cos 2\pi(\theta+\frac{k+1}{2}\alpha))}{\prod\limits_{j=1}^{k}(1-\tau\cos 2\pi(\theta+j\alpha))},

where Rk()R_{k}(\cdot) is a polynomial of degree k.

Proof.

First notice that

gk(θ):=Qk(θ)j=1k(1τcos2π(θ+jα))=det(c1d2d1dkdk1ck)g_{k}(\theta):=Q_{k}(\theta)\prod\limits_{j=1}^{k}(1-\tau\cos 2\pi(\theta+j\alpha))=\det\begin{pmatrix}c_{1}&d_{2}&&\\ d_{1}&\ddots&\ddots&\\ &\ddots&\ddots&d_{k}\\ &&d_{k-1}&c_{k}\end{pmatrix}

where cn=(τE+2λ)cos2π(θ+nα)Ec_{n}=(\tau E+2\lambda)\cos 2\pi(\theta+n\alpha)-E, dn=1τcos2π(θ+nα)d_{n}=1-\tau\cos 2\pi(\theta+n\alpha), then it is a trigonometric polynomial with degree less than kk.

On the other hand, since cos2πθ\cos 2\pi\theta is an even function, denote UU the change of basis δjδk+1j\delta_{j}\mapsto\delta_{k+1-j}, then

U1HV1,α,θk+12α|[1,k]U=HV1,α,θk+12α|[1,k],U^{-1}H_{V_{1},\alpha,\theta-\frac{k+1}{2}\alpha}|_{[1,k]}U=H_{V_{1},\alpha,-\theta-\frac{k+1}{2}\alpha}|_{[1,k]},

which implies that

Qk(θk+12α)=Qk(θk+12α).Q_{k}(\theta-\frac{k+1}{2}\alpha)=Q_{k}(-\theta-\frac{k+1}{2}\alpha).

Due to the fact

j=1k(1τcos2π(θk+12α+jα))=j=1k(1τcos2π(θk+12α+jα)),\prod\limits_{j=1}^{k}(1-\tau\cos 2\pi(\theta-\frac{k+1}{2}\alpha+j\alpha))=\prod\limits_{j=1}^{k}(1-\tau\cos 2\pi(-\theta-\frac{k+1}{2}\alpha+j\alpha)),

then we have

gk(θk+12α)=gk(θk+12α).g_{k}(\theta-\frac{k+1}{2}\alpha)=g_{k}(-\theta-\frac{k+1}{2}\alpha).

Therefore, we obtain

gk(θ)=j=0ka~jcos(2πj(θ+k+12α)),g_{k}(\theta)=\sum\limits_{j=0}^{k}\widetilde{a}_{j}\cos(2\pi j(\theta+\frac{k+1}{2}\alpha)),

since the linear span of {1,cos(2πx),\{1,\cos(2\pi x), cos(2π2x),\cos(2\pi 2x), ,cos(2πkx)}\dots,\cos(2\pi kx)\} is equal to that of {1,cos(2πx),\{1,\cos(2\pi x), cos2(2πx),\cos^{2}(2\pi x), ,cosk(2πx)}\dots,\cos^{k}(2\pi x)\}, consequently we have

Qk(θ)=j=0ka~jcos2πj(θ+k+12α)j=1k(1τcos2π(θ+jα))=j=0kaj(cos2π(θ+k+12α))jj=1k(1τcos2π(θ+jα)).Q_{k}(\theta)=\frac{\sum\limits_{j=0}^{k}\widetilde{a}_{j}\cos 2\pi j(\theta+\frac{k+1}{2}\alpha)}{\prod\limits_{j=1}^{k}(1-\tau\cos 2\pi(\theta+j\alpha))}\\ =\frac{\sum\limits_{j=0}^{k}a_{j}(\cos 2\pi(\theta+\frac{k+1}{2}\alpha))^{j}}{\prod\limits_{j=1}^{k}(1-\tau\cos 2\pi(\theta+j\alpha))}.

By Lemma 5.6, if we denote

Mk,r={x𝕋:|Rk(cos2πx)|e(k+1)(r+ln1+1τ22)},M_{k,r}=\{x\in\mathbb{T}:|R_{k}(\cos 2\pi x)|\leq e^{(k+1)(r+\ln\frac{1+\sqrt{1-\tau^{2}}}{2})}\},

then we have the following:

Lemma 5.7.

Suppose yy\in\mathbb{Z} is (L(E)ϵ,k)(L(E)-\epsilon,k)-singular. Then for every jj satisfying y56kjy16ky-\frac{5}{6}k\leq j\leq y-\frac{1}{6}k, we have θ+(j+k+12)α\theta+(j+\frac{k+1}{2})\alpha belongs to Mk,L(E)ϵ8M_{k,L(E)-\frac{\epsilon}{8}} for k>k3(E,ϵ)k>k_{3}(E,\epsilon).

Proof.

Since yy is (L(E)ϵ,k)(L(E)-\epsilon,k)-singular, without loss of generality, assume that for every interval [j+1,j+k][j+1,j+k] of length kk containing yy with y56kjy16ky-\frac{5}{6}k\leq j\leq y-\frac{1}{6}k, then |yj1|>17k|y-j-1|>\frac{1}{7}k and |j+ky|>17k|j+k-y|>\frac{1}{7}k, we have that

|G[j+1,j+k](y,j+k)|e(L(E)ϵ)|yjk|.|G_{[j+1,j+k]}(y,j+k)|\geq e^{-(L(E)-\epsilon)|y-j-k|}.

Using (29), we have

|G[j+1,j+k](y,j+k)|=|Qyj1(θ+jα)Qk(θ+jα)|e(L(E)ϵ)|yjk|,\begin{split}|G_{[j+1,j+k]}(y,j+k)|=|\frac{Q_{y-j-1}(\theta+j\alpha)}{Q_{k}(\theta+j\alpha)}|\geq e^{-(L(E)-\epsilon)|y-j-k|},\end{split}

by Lemma 5.3, we obtain

|Qyj1(θ+jα)|e|yj1|(L(E)+ϵ90),fork>k1(E,ϵ90).|Q_{y-j-1}(\theta+j\alpha)|\leq e^{|y-j-1|(L(E)+\frac{\epsilon}{90})},\quad\text{for}\quad k>k_{1}(E,\frac{\epsilon}{90}).

which implies that

|Qk(θ+jα)|ekL(E)+(67k×19017k)ϵ=ek(L(E)2ϵ15).\begin{split}|Q_{k}(\theta+j\alpha)|&\leq e^{kL(E)+(\frac{6}{7}k\times\frac{1}{90}-\frac{1}{7}k)\epsilon}=e^{k(L(E)-\frac{2\epsilon}{15})}.\end{split}

On the other hand, by Jensen’s formula and uniquely ergodicity,

limk1kj=1kln(1τcos(2π(θ+jα)))=ln1+1τ22,θ,\displaystyle\lim_{k\to\infty}\frac{1}{k}\sum\limits_{j=1}^{k}\ln(1-\tau\cos(2\pi(\theta+j\alpha)))=\ln\frac{1+\sqrt{1-\tau^{2}}}{2},\quad\forall\theta\in{\mathbb{R}},

thus there exists k3(E,ϵ)>k1(E,ϵ90)k_{3}(E,\epsilon)>k_{1}(E,\frac{\epsilon}{90}) such that if k>k3(E,ϵ)k>k_{3}(E,\epsilon), then

j=1k(1τcos2π(θ+jα))<ek(ln1+1τ22+ϵ120),θ\prod\limits_{j=1}^{k}(1-\tau\cos 2\pi(\theta+j\alpha))<e^{k(\ln\frac{1+\sqrt{1-\tau^{2}}}{2}+\frac{\epsilon}{120})},\quad\forall\theta\in{\mathbb{R}}

Consequently, by Lemma 5.6, we have

Rk(cos2π(θ+(j+k+12)α))ek(L(E)2ϵ15+ln1+1τ22+ϵ120)e(k+1)(L(E)+ln1+1τ22ϵ8).\begin{split}R_{k}(\cos 2\pi(\theta+(j+\frac{k+1}{2})\alpha))&\leq e^{k(L(E)-\frac{2\epsilon}{15}+\ln\frac{1+\sqrt{1-\tau^{2}}}{2}+\frac{\epsilon}{120})}\\ &\leq e^{(k+1)(L(E)+\ln\frac{1+\sqrt{1-\tau^{2}}}{2}-\frac{\epsilon}{8})}.\end{split}

which just means θ+(j+k+12)αMk,L(E)ϵ8\theta+(j+\frac{k+1}{2})\alpha\in M_{k,L(E)-\frac{\epsilon}{8}}. ∎

On the other hand, we may write the polynomial Rk(x)R_{k}(x) in Lagrange interpolation form

(31) |Rk(x)|=|j=0kRk(cos2πθj)lj(xcos2πθl)lj(cos2πθjcos2πθl)||R_{k}(x)|=|\sum\limits_{j=0}^{k}R_{k}(\cos 2\pi\theta_{j})\frac{\prod\nolimits_{l\neq j}(x-\cos 2\pi\theta_{l})}{\prod\nolimits_{l\neq j}(\cos 2\pi\theta_{j}-\cos 2\pi\theta_{l})}|

and introduce the following useful definition:

Definition 5.8.

We say that the set {θ0,,θk}\{\theta_{0},\cdots,\theta_{k}\} is ϵ\epsilon-uniform if

maxx[1,1]maxi=0,,kj=0,jik|xcos2πθj||cos2πθicos2πθj|<ekϵ.\max_{x\in[-1,1]}\max_{i=0,\cdots,k}\prod_{j=0,j\neq i}^{k}\frac{|x-\cos 2\pi\theta_{j}|}{|\cos 2\pi\theta_{i}-\cos 2\pi\theta_{j}|}<e^{k\epsilon}.
Lemma 5.9.

Let 0<ϵ<ϵ0<\epsilon^{\prime}<\epsilon, L(E)>0L(E)>0. If θ0,,θkMk,L(E)ϵ\theta_{0},\cdots,\theta_{k}\in M_{k,L(E)-\epsilon}, then {θ0,,θk}\{\theta_{0},\cdots,\theta_{k}\} is not ϵ\epsilon^{\prime}-uniform for k>k4(ϵ,ϵ)k>k_{4}(\epsilon,\epsilon^{\prime}).

Proof.

Otherwise, using (31) we get

|Rk(x)|(k+1)e(k+1)(L(E)ϵ+ln1+1τ22+ϵ),|R_{k}(x)|\leq(k+1)e^{(k+1)(L(E)-\epsilon+\ln\frac{1+\sqrt{1-\tau^{2}}}{2}+\epsilon^{\prime})},

for any x[1,1]x\in[-1,1]. On the other hand, if kk4(ϵ,ϵ)0k\geq k_{4}(\epsilon,\epsilon^{\prime})0 is large enough, then

j=1k(1τcos2π(θ+jα))>ek(ln1+1τ22ϵϵ2),\prod\limits_{j=1}^{k}(1-\tau\cos 2\pi(\theta+j\alpha))>e^{k(\ln\frac{1+\sqrt{1-\tau^{2}}}{2}-\frac{\epsilon-\epsilon^{\prime}}{2})},

which implies that |Qk(θ)|ek(L(E)ϵϵ3)|Q_{k}(\theta)|\leq e^{k(L(E)-\frac{\epsilon-\epsilon^{\prime}}{3})} for all θ𝕋\theta\in\mathbb{T}. However, by Herman’s subharmonic function argument, /ln|Qk(x)|dxkL(E)\int_{\mathbb{R/Z}}\ln|Q_{k}(x)|dx\geq kL(E), this is a contradiction. ∎

We consider two points 0 and yy, without loss of generality, assume y>0y>0. let k=238y+1k=2\lfloor\frac{3}{8}y\rfloor+1, n1=34k,n2=y34kn_{1}=-\lfloor\frac{3}{4}k\rfloor,n_{2}=y-\lfloor\frac{3}{4}k\rfloor. Then we can construct the following sequence:

θj={θ+(n1+k12+j)α,j=0,1,,k+121,θ+(n2+k12+jk+12)α,j=k+12,k+12+1,,k.\theta_{j}=\left\{\begin{matrix}\theta+(n_{1}+\frac{k-1}{2}+j)\alpha,&j=0,1,\dots,\lfloor\frac{k+1}{2}\rfloor-1,\\ \theta+(n_{2}+\frac{k-1}{2}+j-\lfloor\frac{k+1}{2}\rfloor)\alpha,&j=\lfloor\frac{k+1}{2}\rfloor,\lfloor\frac{k+1}{2}\rfloor+1,\dots,k.\end{matrix}\right.

These points θ0,θ1,,θk\theta_{0},\theta_{1},\dots,\theta_{k} are distinct and satisfy the following:

Lemma 5.10.

Suppose that αDC\alpha\in DC, θΘ\theta\in\Theta. Then for any ϵ>0\epsilon>0, there exists k5(α,θ,ϵ)>0k_{5}(\alpha,\theta,\epsilon)>0, such that for k>k5(α,θ,ϵ)k>k_{5}(\alpha,\theta,\epsilon), the above constructed sequence {θj}j=0k\{\theta_{j}\}_{j=0}^{k} is ϵ\epsilon-uniform.

Proof.

This is essentially Lemma 7 of [53]. ∎

Proof of Proposition 5.5: GAA case.

By Lemma 5.9 and Lemma 5.10, we know that {θj}j=0k\{\theta_{j}\}_{j=0}^{k} can not be inside Mk,L(E)ϵ8M_{k,L(E)-\frac{\epsilon}{8}} at the same time for sufficiently large kk. Since uEu_{E} is a generalized function satisfying uE(0)0u_{E}(0)\neq 0, 0 is (L(E)ϵ,k)(L(E)-\epsilon,k)-singular for sufficiently large kk. Applying Lemma 5.7, one obtains

{θj}j=0k+121Mk,L(E)ϵ8.\{\theta_{j}\}_{j=0}^{\lfloor\frac{k+1}{2}\rfloor-1}\subset M_{k,L(E)-\frac{\epsilon}{8}}.

Assume yy is (L(E)ϵ,k)(L(E)-\epsilon,k)-singular, then we also have

{θj}j=k+12kMk,L(E)ϵ8.\{\theta_{j}\}_{j=\lfloor\frac{k+1}{2}\rfloor}^{k}\subset M_{k,L(E)-\frac{\epsilon}{8}}.

Thus {θj}j=0kMk,L(E)ϵ8\{\theta_{j}\}_{j=0}^{k}\subset M_{k,L(E)-\frac{\epsilon}{8}}, this contradiction means yy must be (L(E)ϵ,k)(L(E)-\epsilon,k)-regular for y>y(α,θ,E,ϵ)y>y(\alpha,\theta,E,\epsilon). Notice that k=238|y|+1>516|y|k=2\lfloor\frac{3}{8}|y|\rfloor+1>\frac{5}{16}|y|, we thus finish the proof of GAA case.

5.2. Anderson localization for the mosaic model.

Note in the GAA case, one of the basic observation is that the elements of Mk(θ)M_{k}(\theta) can be expressed by Qk(θ)Q_{k}(\theta). In the quasi-periodic mosaic case, the transfer matrix reads as

M2k(θ)=(P2k(θ)Q2k1(θ)P2k1(θ)Q2k2(θ)),M_{2k}(\theta)=\begin{pmatrix}P_{2k}(\theta)&Q_{2k-1}(\theta)\\ -P_{2k-1}(\theta)&-Q_{2k-2}(\theta)\end{pmatrix},

the key observation is that elements of M2k(θ)M_{2k}(\theta) can be written as linear combination of Q2k1(θ)Q_{2k-1}(\theta) (possibly with different kk and different θ\theta):

Lemma 5.11.

We have

EQ2k2(θ)=Q2k1(θ)Q2k3(θ),EP2k(θ)=Q2k+1(θ2α)Q2k1(θ),E2P2k1(θ)=Q2k+1(θ2α)+Q2k1(θ)+Q2k1(θ2α)+Q2k3(θ).\begin{split}EQ_{2k-2}(\theta)&=-Q_{2k-1}(\theta)-Q_{2k-3}(\theta),\\ EP_{2k}(\theta)&=-Q_{2k+1}(\theta-2\alpha)-Q_{2k-1}(\theta),\\ E^{2}P_{2k-1}(\theta)&=Q_{2k+1}(\theta-2\alpha)+Q_{2k-1}(\theta)+Q_{2k-1}(\theta-2\alpha)+Q_{2k-3}(\theta).\end{split}
Proof.

Note that V2(θ,2n+1)=0V_{2}(\theta,2n+1)=0 and V2(θ,n+2)=V2(θ+2α,n)V_{2}(\theta,n+2)=V_{2}(\theta+2\alpha,n). Then if we expand the determinant det[(HV,α,θE)|[0,2k]]\det[(H_{V,\alpha,\theta}-E)|_{[0,2k]}] by the last column, we have

P2k(θ)=EP2k1(θ)P2k2(θ),EQ2k2(θ)=Q2k1(θ)Q2k3(θ).\begin{split}P_{2k}(\theta)&=-EP_{2k-1}(\theta)-P_{2k-2}(\theta),\\ EQ_{2k-2}(\theta)&=-Q_{2k-1}(\theta)-Q_{2k-3}(\theta).\end{split}

Meanwhile, if we expand the determinant det[(HV,α,θE)|[0,2k]\det[(H_{V,\alpha,\theta}-E)|_{[0,2k]} by the first column, we have

Q2k1(θ)=EP2k2(θ+2α)Q2k3(θ+2α).Q_{2k-1}(\theta)=-EP_{2k-2}(\theta+2\alpha)-Q_{2k-3}(\theta+2\alpha).

which implies that

EP2k(θ)\displaystyle EP_{2k}(\theta) =\displaystyle= Q2k+1(θ2α)Q2k1(θ)\displaystyle-Q_{2k+1}(\theta-2\alpha)-Q_{2k-1}(\theta)
E2P2k1(θ)\displaystyle E^{2}P_{2k-1}(\theta) =\displaystyle= Q2k+1(θ2α)+Q2k1(θ)+Q2k1(θ2α)+Q2k3(θ).\displaystyle Q_{2k+1}(\theta-2\alpha)+Q_{2k-1}(\theta)+Q_{2k-1}(\theta-2\alpha)+Q_{2k-3}(\theta).

We thus finish the proof. ∎

Similar to Lemma 5.6, we have the following:

Lemma 5.12.

For every k2+1k\in 2\mathbb{N}+1, there exists a polynomial R~k12\widetilde{R}_{\frac{k-1}{2}} of degree k12\frac{k-1}{2} such that

Qk(θ)=R~k12(cos2π(θ+k+12α).Q_{k}(\theta)=\widetilde{R}_{\frac{k-1}{2}}(\cos 2\pi(\theta+\frac{k+1}{2}\alpha).
Proof.

Since cos2πθ\cos 2\pi\theta is an even function, it follow that the changes of basis δjδ2k+2j\delta_{j}\mapsto\delta_{2k+2-j} transforms

HV2,α,θ(k+1)α|1,2k+1intoHV2,α,θ(k+1)α|1,2k+1.H_{V_{2},\alpha,\theta-(k+1)\alpha}|_{1,2k+1}\qquad\text{into}\qquad H_{V_{2},\alpha,-\theta-(k+1)\alpha}|_{1,2k+1}.

which implies that

Q2k+1(θ(k+1)α)=Q2k+1(θ(k+1)α).Q_{2k+1}(\theta-(k+1)\alpha)=Q_{2k+1}(-\theta-(k+1)\alpha).

The rest proof is similar to Lemma 5.6, we thus omit the details. ∎

By Lemma 5.12, if we denote the set

M~k,r={x𝕋:|R~k12(cos2πx)|e(k+1)r},\widetilde{M}_{k,r}=\{x\in\mathbb{T}:|\widetilde{R}_{\frac{k-1}{2}}(\cos 2\pi x)|\leq e^{(k+1)r}\},

then we have the following:

Lemma 5.13.

Suppose yy\in\mathbb{Z} is (L(E)ϵ,k)(L(E)-\epsilon,k)-singular, k2+1k\in 2\mathbb{N}+1. Then for every jj\in\mathbb{Z} satisfying y56k+k+122jy16k+k+12y-\frac{5}{6}k+\frac{k+1}{2}\leq 2j\leq y-\frac{1}{6}k+\frac{k+1}{2}, we have θ+2jα\theta+2j\alpha belongs to M~k,L(E)ϵ8\widetilde{M}_{k,L(E)-\frac{\epsilon}{8}} for k>k1(E,ϵ48)k>k_{1}(E,\frac{\epsilon}{48}).

Proof.

The proof is similar to Lemma 5.7, we omit the details. ∎

Lemma 5.14.

Let 0<ϵ<ϵ0<\epsilon^{\prime}<\epsilon, k2+1k\in 2\mathbb{N}+1, L(E)>0L(E)>0. If θ0,,θk12M~k,L(E)ϵ\theta_{0},\cdots,\theta_{\frac{k-1}{2}}\in\widetilde{M}_{k,L(E)-\epsilon}, then {θ0,,\{\theta_{0},\cdots, θk12}\theta_{\frac{k-1}{2}}\} is not ϵ\epsilon^{\prime}-uniform for k>k6(ϵ,ϵ)k>k_{6}(\epsilon,\epsilon^{\prime}).

Proof.

Otherwise, using Lagrange interpolation form (31), we get |R~k12(x)|<ek(L(E)ϵϵ2)|\widetilde{R}_{\frac{k-1}{2}}(x)|<e^{k(L(E)-\frac{\epsilon-\epsilon^{\prime}}{2})} for all x[1,1]x\in[-1,1] when k>k6(ϵ,ϵ)k>k_{6}(\epsilon,\epsilon^{\prime}), which implies

|Qk(θ)|<ek(L(E)ϵϵ2),θ.|Q_{k}(\theta)|<e^{k(L(E)-\frac{\epsilon-\epsilon^{\prime}}{2})},\qquad\forall\theta\in{\mathbb{R}}.

On the other hand, Lemma 5.11 imply that

||M2n(θ)Cmax{|Q2n+1(θ2α)|,|Q2n1(θ)|,|Q2n1(θ2α)|,|Q2n3(θ)|},||M_{2n}(\theta)\|\leq C\max\{|Q_{2n+1}(\theta-2\alpha)|,|Q_{2n-1}(\theta)|,|Q_{2n-1}(\theta-2\alpha)|,|Q_{2n-3}(\theta)|\},

for some constant C=C(λ)C=C(\lambda), since by Corollary 3.5, we have 2+λ|E|>1λ2+\lambda\geq|E|>\frac{1}{\lambda}. However, this contradicts to (28) for sufficiently large nn. We thus finish the proof. ∎

Assume that (qn)n(q_{n})_{n} is the sequence of denominators of the best rational approximations of 2α2\alpha. Select nn such that qny8<qn+1q_{n}\leq\frac{y}{8}<q_{n+1} and let ss be the largest positive integer satisfying sqny8sq_{n}\leq\frac{y}{8}. Set I1,I2I_{1},I_{2}\subset\mathbb{Z} as follows

I1=[0,sqn1] and I2=[1+y2sqn,y2+sqn].I_{1}=[0,sq_{n}-1]\text{ and }I_{2}=[1+\lfloor\frac{y}{2}\rfloor-sq_{n},\lfloor\frac{y}{2}\rfloor+sq_{n}].
Lemma 5.15.

Let θj=θ+2jα\theta_{j}=\theta+2j\alpha, then for any ϵ>0\epsilon>0, the set {θj}jI1I2\{\theta_{j}\}_{j\in I_{1}\cup I_{2}} is ϵ\epsilon-uniform if y>y(α,θ,ϵ)y>y(\alpha,\theta,\epsilon).

Proof.

Take x=cos2πax=\cos 2\pi a. Now it suffices to estimate

jI1I2,ji(ln|cos2πacos2πθj|ln|cos2πθicos2πθj|)=12.\sum_{j\in I_{1}\cup I_{2},j\neq i}(\ln|\cos 2\pi a-\cos 2\pi\theta_{j}|-\ln|\cos 2\pi\theta_{i}-\cos 2\pi\theta_{j}|)=\sum_{1}-\sum_{2}.

Then Lemma 2.1 reduces this problem to estimating the minimal terms.

First we estimate 1\sum_{1}:

1=jI1I2,jiln|sinπ(a+θj)|+jI1I2,jiln|sinπ(aθj)|+(3sqn1)ln2=1,++1,+(3sqn1)ln2,\begin{split}\sum_{1}&=\sum_{j\in I_{1}\cup I_{2},j\neq i}\ln|\sin\pi(a+\theta_{j})|+\sum_{j\in I_{1}\cup I_{2},j\neq i}\ln|\sin\pi(a-\theta_{j})|+(3sq_{n}-1)\ln 2\\ &=\sum_{1,+}+\sum_{1,-}+(3sq_{n}-1)\ln 2,\end{split}

we cut 1,+\sum_{1,+} or 1,\sum_{1,-} into 3s3s sums and then apply Lemma 2.1, we get that for some absolute constant C1C_{1}:

13sqnln2+C1slnqn.\sum_{1}\leq-3sq_{n}\ln 2+C_{1}s\ln q_{n}.

Next, we estimate 2\sum_{2} as follows:

2=jI1I2,jiln|sinπ(2θ+(i+j)2α)|+jI1I2,jiln|sinπ(ij)2α|+(3sqn1)ln2=2,++2,+(3sqn1)ln2.\begin{split}\sum_{2}&=\sum_{j\in I_{1}\cup I_{2},j\neq i}\ln|\sin\pi(2\theta+(i+j)2\alpha)|\\ &\quad+\sum_{j\in I_{1}\cup I_{2},j\neq i}\ln|\sin\pi(i-j)2\alpha|+(3sq_{n}-1)\ln 2\\ &=\sum_{2,+}+\sum_{2,-}+(3sq_{n}-1)\ln 2.\end{split}

For any 0<|j|<qn+10<|j|<q_{n+1}, since αDC(γ,τ)\alpha\in DC(\gamma,\tau) we have

j2α/qn2α/γ(2qn)σ.\|j2\alpha\|_{\mathbb{R/Z}}\geq\|q_{n}2\alpha\|_{\mathbb{R/Z}}\geq\frac{\gamma}{(2q_{n})^{\sigma}}.

Therefore we obtain

max{ln|sinx|,ln|sin(x+πj2α)|}2lnγ2σln2qnfory>y1(α).\max\{\ln|\sin x|,\ln|\sin(x+\pi j2\alpha)|\}\geq 2\ln\gamma-2\sigma\ln 2q_{n}\quad\text{for}\ y>y_{1}(\alpha).

This means in any interval of length sqnsq_{n}, there can be at most one term which is less than 2lnγ2σln2qn2\ln\gamma-2\sigma\ln 2q_{n}. Then there can be at most 3 such terms in total.

For the part 2,\sum_{2,-}, since

(ij)2α/γ2σ|ij|σγ(18sqn)σ,\|(i-j)2\alpha\|_{\mathbb{R/Z}}\geq\frac{\gamma}{2^{\sigma}|i-j|^{\sigma}}\geq\frac{\gamma}{(18sq_{n})^{\sigma}},

these 3 smallest terms must be bounded by lnγσln18sqn\ln\gamma-\sigma\ln 18sq_{n} from below. Hence by Lemma 2.1, we have

(32) 2,3sqnln2+3lnγ3σln18sqnC2slnqn,\sum_{2,-}\geq-3sq_{n}\ln 2+3\ln\gamma-3\sigma\ln 18sq_{n}-C_{2}s\ln q_{n},

for y>y2(α)y>y_{2}(\alpha) and some absolute constant C2C_{2}. For the part 2,+\sum_{2,+}, since θΘ\theta\in\Theta, then

2θ+(i+j)2α/η|i+j|ση(18sqn)σ,\|2\theta+(i+j)2\alpha\|_{\mathbb{R/Z}}\geq\frac{\eta}{|i+j|^{\sigma}}\geq\frac{\eta}{(18sq_{n})^{\sigma}},

these 3 smallest terms must be greater than lnησln18sqn\ln\eta-\sigma\ln 18sq_{n}. Therefore combining with (32), we have

23sqnln2+3lnγ3σln18sqn+3lnη3σln18sqn(C2+C3)slnqn,\sum_{2}\geq-3sq_{n}\ln 2+3\ln\gamma-3\sigma\ln 18sq_{n}+3\ln\eta-3\sigma\ln 18sq_{n}-(C_{2}+C_{3})s\ln q_{n},

consequently, for any ϵ>0\epsilon>0 if y>y(α,θ,ϵ)y>y(\alpha,\theta,\epsilon), 126ϵsqn,\sum_{1}-\sum_{2}\leq 6\epsilon sq_{n}, i.e. the set {θj}jI1I2\{\theta_{j}\}_{j\in I_{1}\cup I_{2}} is ϵ\epsilon-uniform. ∎

Proof of Proposition 5.5: Quasi-periodic mosaic case:

Combining Lemma 5.14 and Lemma 5.15, we know that when yy is sufficiently large, {θj}jI1I2\{\theta_{j}\}_{j\in I_{1}\cup I_{2}} can not be inside the set M~6sqn1,L(E)ϵ8\widetilde{M}_{6sq_{n}-1,L(E)-\frac{\epsilon}{8}} at the same time. Therefore 0 and yy can not be (L(E)ϵ,6sqn1)(L(E)-\epsilon,6sq_{n}-1)-singular at the same time by Lemma 5.13. However 0 is (L(E)ϵ,6sqn1)(L(E)-\epsilon,6sq_{n}-1)-singular given yy large enough. Therefore

{θj}jI1M~6sqn1,L(E)ϵ8.\{\theta_{j}\}_{j\in I_{1}}\subset\widetilde{M}_{6sq_{n}-1,L(E)-\frac{\epsilon}{8}}.

Thus yy must be (L(E)ϵ,6sqn1)(L(E)-\epsilon,6sq_{n}-1)-regular for y>y(α,θ,E,ϵ)y>y(\alpha,\theta,E,\epsilon). Notice that 6sqn1>6/16y1>516y6sq_{n}-1>6/16y-1>\frac{5}{16}y, thus we complete the proof.

6. Proof of Main results

Proof of Theorem 1.1: By Corollary 3.10, Theorem 1.1 (1) and the first statement of Theorem 1.1 (3) follow from Theorem 2.5 and Theorem 4.1. Theorem 1.1 (2) and the second statement of Theorem 1.1 (3) follow from Theorem 5.1.

Proof of Theorem 1.3: The proof is same as Theorem 1.1, one only needs to replace Corollary 3.10 by Corollary 3.5.

Proof of Theorem 1.4: The proof is same as Theorem 1.1, one only needs to replace Corollary 3.10 by Corollary 3.8.

Theorem 1.3 and Theorem 1.4 covers the quasi-periodic mosaic model κ=2\kappa=2 and κ=3\kappa=3, for the general κ\kappa, recall that

aκ(E)=1E24((E+E242)κ(EE242)κ),\displaystyle a_{\kappa}(E)=\frac{1}{\sqrt{E^{2}-4}}\left((\frac{E+\sqrt{E^{2}-4}}{2})^{\kappa}-(\frac{E-\sqrt{E^{2}-4}}{2})^{\kappa}\right),\ \

and we have the following

Theorem 6.1.

For any λ0\lambda\neq 0, αDC\alpha\in DC, κ+\kappa\in{\mathbb{Z}}^{+}, then |λaκ(E)|=1|\lambda a_{\kappa}(E)|=1 are the MEs. More precisely,

  1. (1)

    HV2,α,θH_{V_{2},\alpha,\theta} has purely absolutely continuous spectrum for every θ\theta in

    (33) Σ(V2){E||λaκ(E)|<1}.\Sigma(V_{2})\cap\{E\in{\mathbb{R}}||\lambda a_{\kappa}(E)|<1\}.
  2. (2)

    If

    Σ(V2){E||λaκ(E)|>1},\Sigma(V_{2})\cap\{E\in{\mathbb{R}}||\lambda a_{\kappa}(E)|>1\}\neq\emptyset,

    then HV2,α,θH_{V_{2},\alpha,\theta} has Anderson localization in this set for almost every θ\theta.

Proof.

The proof is same as Theorem 1.3, one only needs to replace Corollary 3.5 by Lemma 3.1 and Lemma 3.3.∎

Proof of Corollary 1.6: By Aubry duality, we only to need consider its dual operator HV3,α,θH_{V_{3},\alpha,\theta}. By Corollary 3.13, Theorem 2.5 and Theorem 4.1, HV3,α,θH_{V_{3},\alpha,\theta} has purely absolutely continuous spectrum in sgn(λ)E<2coshp2|λ|sgn(\lambda)E<2\cosh p-\frac{2}{|\lambda|} for every θ\theta. By Corollary 3.13 and Theorem 5.1, HV3,α,θH_{V_{3},\alpha,\theta} has Anderson localization in sgn(λ)E>2coshp2|λ|sgn(\lambda)E>2\cosh p-\frac{2}{|\lambda|} for a.e. θ\theta. By Aubry duality [27, 45], and the fact that Σ(H^V3,α,θ)=λ2Σ(V3),\Sigma(\widehat{H}_{V_{3},\alpha,\theta})=\frac{\lambda}{2}\Sigma(V_{3}), ME of (4) has the form sgn(λ)E=2coshp2|λ|sgn(\lambda)E=2\cosh p-\frac{2}{|\lambda|}, which is just E+1=2|λ|coshpE+1=2|\lambda|\cosh p.

Proof of Corollary 1.8: By Corollary 3.14, then Corollary 1.8 follows from Theorem 1.1.

Appendix A A quantitative almost reducibility result

The following quantitative almost reducibility result from [29, 67] is the basis of our proof.

Proposition A.1.

Let αDC(γ,σ)\alpha\in DC(\gamma,\sigma). Suppose that ASL(2,)A\in SL(2,{\mathbb{R}}), fChω(𝕋,sl(2,)).f\in C_{h}^{\omega}\left(\mathbb{T},sl(2,{\mathbb{R}})\right). Then for any h+<h,h_{+}<h, there exists numerical constant C0,C_{0}, and constant D0=D0(γ,σ)D_{0}=D_{0}\left(\gamma,\sigma\right) such that if

fhϵD0AC0(min{1,1h}(hh+))C0σ,\|f\|_{h}\leq\epsilon\leq\frac{D_{0}}{\|A\|^{C_{0}}}\left(\min\left\{1,\frac{1}{h}\right\}\left(h-h_{+}\right)\right)^{C_{0}\sigma},

then there exist BCh+ω(2𝕋,PSL(2,)),B\in C_{h_{+}}^{\omega}\left(2{\mathbb{T}},PSL(2,{\mathbb{R}})\right), 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, let spec(A)={e2πiξ,e2πiξ},N=2hh+|lnϵ|spec(A)=\left\{e^{2\pi i\xi},e^{-2\pi i\xi}\right\},N=\frac{2}{h-h_{+}}|\ln\epsilon|, then we can distinguish two cases:

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

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

    then

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

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

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

    2ξn,α/<ϵ115\left\|2\xi-\left<n_{*},\alpha\right>\right\|_{\mathbb{R}/\mathbb{Z}}<\epsilon^{\frac{1}{15}}

    then we have

    Bh+1γ|n|τ2ϵh+hh+,B01γ|n|τ2,f+h+<ϵeh+ϵ118τ.\|B\|_{h_{+}}\leq\frac{1}{\gamma}|n_{*}|^{\frac{\tau}{2}}\epsilon^{-\frac{h_{+}}{h-h_{+}}},\quad\left\|B\right\|_{0}\leq\frac{1}{\gamma}|n_{*}|^{\frac{\tau}{2}},\quad\|f_{+}\|_{h_{+}}<\epsilon e^{-h_{+}\epsilon^{-\frac{1}{18\tau}}}.

    Moreover, degB=n\deg B=n_{*}, letting M=11+i(1i1i)M=\frac{1}{1+i}\begin{pmatrix}1&-i\\ 1&i\end{pmatrix}, then the constant A+A_{+} can be written as

    A+=M1exp(it+ν+ν¯+it+)MA_{+}=M^{-1}\exp\left(\begin{array}[]{cc}{it_{+}}&{\nu_{+}}\\ {\bar{\nu}_{+}}&{-it_{+}}\end{array}\right)M

    with estimates |t+|ϵ116|t_{+}|\leq\epsilon^{\frac{1}{16}}, |ν+|ϵ1516e2π|n|h\left|\nu_{+}\right|\leq\epsilon^{\frac{15}{16}}e^{-2\pi\left|n_{*}\right|h}.

Remark A.2.

Assume that AA varies in some compact subset of SL(2,)SL(2,{\mathbb{R}}). Then ϵ\epsilon can be taken uniform with respect to AA.

Acknowledgements

The authors would like to thank D. Damanik, R. Krikorian and S. Jitomirskaya for useful discussions. X. Xia, J. You and Q.Zhou were partially supported by National Key R&D Program of China (2020YFA0713300) and Nankai Zhide Foundation. Y. Wang is supported by the NSFC grant (12061031). J. You was also partially supported by NSFC grant (11871286). Z. Zheng acknowledges financial supports of NSFC grant (12031020, 11671382), CAS Key Project of Frontier Sciences (No. QYZDJ-SSW-JSC003), the Key Lab. of Random Complex Structures and Data Sciences CAS and National Center for Mathematics and Interdisciplinary Sciences CAS. Q.Zhou was supported by NSFC grant (12071232), the Science Fund for Distinguished Young Scholars of Tianjin (No. 19JCJQJC61300).

References

  • [1] M. Aizenman and S. Molchanov, Localization at large disorder and at extreme energies: An elementary derivation, Commun. Math. Phys. 157.2 (1993): 245-278.
  • [2] F. A. An, K. Padavić, E. J. Meier, S. Hegde, S. Ganeshan, J. H. Pixley, S. Vishveshwara, and B. Gadway, Interactions and mobility edges: Observing the generalized Aubry-Andre model, Phys. Rev. Lett. 126.4 (2021): 040603.
  • [3] P. W. Anderson, Absence of diffusion in certain random lattices, Phys. Rev. 109.5 (1958): 1492-1505.
  • [4] A. Avila, Almost reducibility and absolute continuity I, arXiv preprint arXiv:1006.0704, 2010.
  • [5] A. Avila, Global theory of one-frequency Schrödinger operators, Acta Math. 215.1 (2015): 1-54.
  • [6] A Avila. Lyapunov exponents, KAM and the spectral dichotomy for one-frequency schrödinger operators.
  • [7] A. Avila, The absolutely continuous spectrum of the almost Mathieu operator. https://webusers.imj-prg.fr/ artur.avila/papers.html.
  • [8] A. Avila, B. Fayad and R. Krikorian, A KAM scheme for SL(2,)SL(2,{\mathbb{R}}) cocycles with Liouvillean frequencies, Geom. Funct. Anal. 21.5 (2011): 1001-1019.
  • [9] A.Avila and S. Jitomirskaya, The ten Martini problem, Ann.of Math. 170.1 (2009): 303-342.
  • [10] A. Avila and S. Jitomirskaya, Almost localization and almost reducibility, J. Eur. Math. Soc. 12.1 (2010): 93-131.
  • [11] A. Avila, S. Jitomirskaya and C.A. Marx, Spectral theory of extended Harper’s model and a question by Erdös and Szekeres, Invent. math. 210.1 (2017): 283-339.
  • [12] A. Avila, S. Jitomirskaya and Q. Zhou, Second Phase transition line, Math. Ann. 370.1 (2018): 271-285.
  • [13] A. Avila, K. Khanin and M. Leguil, Invariant graphs and spectral type of Schrödinger operators, Pure Appl. Funct. Anal. 5 (2020): 1257-1277.
  • [14] A. Avila, J. You, and Q. Zhou, Sharp phase transitions for the almost Mathieu operator, Duke Math. J. 166.14 (2017): 2697-2718.
  • [15] J. Avron and B. Simon, Almost periodic Schrödinger operators. II. The integrates density of states, Duke Math. J. 50.1 (1983): 369-391.
  • [16] J. Avron, D. Osadchy and R. Seiler, A topological look at the quantum Hall effect, Physics today 56.8 (2003): 38-42.
  • [17] R. Balasubramanian, S. H. Kulkarni and R. Radha, Non-invertibility of certain almost Mathieu operators, Proc AMS 129, 2017 - 2018 (2001)
  • [18] J. Billy et al. Nature (London) 453, 891(2008).
  • [19] J. Biddle and S. D. Sarma, Predicted mobility edges in one-dimensional incommensurate optical lattices: an exactly solvable model of Anderson localization, Phys. Rev. Lett. 104.7 (2010): 070601.
  • [20] K. Bjerklöv, Explicit examples of arbitrarily large analytic ergodic potentials with zero Lyapunov exponent, Geom. Funct. Anal. 16.6 (2006): 1183-1200.
  • [21] K. Bjerklöv and R. Krikorian, Coexistence of ac and pp spectrum for kicked quasi-periodic potentials, J. Spectr. Theory, DOI: 10.4171/JST/370
  • [22] J. Bourgain, On the spectrum of lattice Schrödinger operators with deterministic potential, J. Anal. Math. 87.1 (2002): 37-75.
  • [23] J. Bourgain, Green’s function estimates for lattice Schrödinger operators and applications. Annals of Mathematics Studies. Princeton University Press, Princeton, NJ. 158 (2005).
  • [24] J. Bourgain, Anderson localization for quasi-periodic lattice Schrödinger operators on d{\mathbb{Z}}^{d}, dd arbitrary, Geom. Funct. Anal. 17.3 (2007): 682-706.
  • [25] J. Bourgain and M. Goldstein, On nonperturbative localization with quasi-periodic potential, Ann. of Math. 152.3 (2000), 835-879.
  • [26] J. Bourgain, M. Goldstein and W. Schlag, Anderson localization for Schrödinger operators on 2{\mathbb{Z}}^{2} with quasi-periodic potential, Acta Math. 188.1 (2002): 41-86.
  • [27] J. Bourgain and S. Jitomirskaya, Absolutely continuous spectrum for 1D quasiperiodic operators, Invent. Math. 148(3), 453-463 (2002).
  • [28] J. Bourgain and S. Jitomirskaya, Continuity of the Lyapunov exponent for quasiperiodic operators with analytic potential, J. Statist. Phys. 108.5 (2002): 1203-1218.
  • [29] A. Cai, C. Chavaudret, J. You and Q. Zhou, Sharp Hölder continuity of the Lyapunov exponent of finitely differentiable quasi-periodic cocycles, Math. Z. 291.3 (2019): 931–958.
  • [30] D. Damanik, Schrödinger operators with dynamically defined potentials, Ergodic. Theory. Dynam. Systems, 37.6 (2017): 1681-1764.
  • [31] P. Deift and B. Simon, Almost periodic Schrödinger operators. III. The absolutely continuous spectrum in one dimension. Comm. Math. Phys. 90.3 (1983): 389-411.
  • [32] E. I. Dinaburg and Y. G. Sinai, The one-dimensional Schrödinger equation with quasiperiodic potential, Funk. Anal. i Prilozen. 9.4 (1975): 279-289.
  • [33] H. Eliasson, Floquet solutions for the 1-dimensional quasi-periodic Schrödinger equation, Comm. Math. Phys. 146.3 (1992): 447-482.
  • [34] H. Eliasson, Discrete one-dimensional quasi-periodic Schrödinger operators with pure point spectrum, Acta Math. 179.2 (1997): 153-196.
  • [35] F. Evers and A. D. Mirlin, Anderson transitions, Rev. Mod. Phys. 80.4 (2008): 1355.
  • [36] A. Fedotov and F. Klopp, Coexistence of different spectral types for almost periodic Schrödinger equations in dimension one, Mathematical results in quantum mechanics. Birkhäuser, Basel, 1999: 243-251.
  • [37] J. Fröhlich and T. Spencer, Absence of diffusion in the Anderson tight binding model for large disorder or low energy, Commun. Math. Phys. 88.2 (1983): 151-184.
  • [38] J. Fröhlich, T. Spencer and P. Wittwer, Localization for a class of one dimensional quasi-periodic Schrödinger operators, Commun. Math. Phys. 132.1 (1990): 5-25.
  • [39] A. Furman, On the multiplicative ergodic theorem for the uniquely ergodic systems. Ann. Inst. Henri Poincaré. 33(1997), 797-815.
  • [40] G. Francois, and A. Klein, A characterization of the Anderson metal-insulator transport transition, Duke Math. J. 124.2 (2004): 309-350
  • [41] G. Sriram, J. H. Pixley and S. D. Sarma, Nearest neighbor tight binding models with an exact mobility edge in one dimension, Phys. Rev. Lett. 144.14 (2015): 146601.
  • [42] L. Ge and J. You, Arithmetic version of Anderson localization via reducibility, Geom. Funct. Anal. 30.5 (2020): 1370-1401.
  • [43] L. Ge, J. You and Q. Zhou, Exponential Dynamical Localization: Criterion and Applications. To appear in Ann. Sci. Ec. Norm. Super
  • [44] D. J. Gilbert and D. B. Pearson, On subordinacy and analysis of the spectrum of one-dimensional Schrödinger operators, J. Math. Anal. Appl. 128.1 (1987): 30-56.
  • [45] A. Gordon, S. Jitomirskaya, Y. Last, and B. Simon, Duality and singular continuous spectrum in the almost Mathieu equation. Acta Math. 178 (1997), 169-183.
  • [46] I. Goldsheid, S. Molchanov and L. Pastur, A pure point spectrum of the stochastic one-dimensional Schrödinger operator, Funct. Anal. Appl. 11.1 (1977): 1-10.
  • [47] P. G. Harper, Single band motion of conduction electrons in a uniform magnetic field, Proc. Phys. Soc. London A. 68 (1955): 874-892.
  • [48] M. R. Herman, Une méthode pour minorer les exposants de Lyapounov et quelques exemples montrant le caractère local d’un théorème d’Arnold et de Moser sur le tore de dimension 22. Comment. Math. Helv. 58.3 (1983): 453-502.
  • [49] H. Hiramoto and M. Kohmoto, Scaling analysis of quasiperiodic systems: Generalized Harper model, Phys. Rev. B 40.12 (1989): 8225.
  • [50] H. Hiramoto, M. Kohmoto, New localization in a quasiperiodic systems. Phys. Rev. Lett. 62.23 (1989): 2714-2717.
  • [51] X. Hou and J. You, Almost reducibility and non-perturbative reducibility of quasi-periodic linear systems, Invent. Math. 190.1 (2012): 209-260.
  • [52] W. Jian, Y. Shi and X. Yuan, Anderson localization for one-frequency quasi-periodic block operators with long-range interactions, J. Math. Phys. 60.6 (2019): 063504.
  • [53] S. Jitomirskaya, Metal-insulator transition for the almost Mathieu operator, Ann. of Math. (2) 150.3 (1999): 1159-1175.
  • [54] S. Jitomirskaya and C. A. Marx, Analytic quasi-perodic cocycles with singularities and the Lyapunov exponent of extended Harper’s model, Commun. Math. Phys. 316.1 (2012): 237-267.
  • [55] S. Jitomirskaya and C. A. Marx, Erratum to: Analytic quasi-perodic cocycles with singularities and the Lyapunov exponent of extended Harper’s model, Commun. Math. Phys 317 (2013): 269-271.
  • [56] S. Jitomirskaya and D. Koslover, Continuity of the Lyapunov exponent for analytic quasiperiodic cocycles, Ergodic Theory Dynam. Systems 29.6 (2009): 1881-1905.
  • [57] S. Jitomirskaya, D. A. Koslover and M. S. Schulteis. Localization for a family of one-dimensional quasiperiodic operators of magnetic origin. Ann. Henri Poincaré 6(1) (2005), 103–124
  • [58] S. Jitormiskya and W. Liu, Universal hierarchical structure of quasi-periodic eigenfuctions, Ann. of Math. 187.3 (2018): 721-776.
  • [59] S. Jitormiskya, W. Liu and Y. Shi, Anderson localization for multi-frequency quasiperiodic operators on d{\mathbb{Z}}^{d}, Geom. Funct. Anal. 30.2 (2020): 457-481.
  • [60] S. Jitormiskya and F. Yang, Pure point spectrum for the Maryland model: A constructive proof. Ergodic Theory and Dynamical Systems, 41(1), 283-294. doi:10.1017/etds.2019.50
  • [61] R. A. Jonhnson, Exponential dichotomy, rotation number, and linear differential operators with bounded coefficients, J. Differential Equations 61.1 (1986): 54-78.
  • [62] R. Johnson and J. Moser, The rotation number for almost periodic potentials. Commun. Math. Phys. 84 (1982): 403-438.
  • [63] S. Klein, Anderson localization for the discrete one-dimensional quasi-periodic Schrödinger operator with potential defined by a Gevery-class function, J. Funct. Anal. 218.2 (2005): 255-292.
  • [64] S. Kotani, Lyapunov exponents and spectra for one-dimensional random Schrödinger operators, Contemp. Math. 50 (1986): 277-286.
  • [65] I. V. Krasovsky, Bloch electron in a magnetic field and the Ising model, Phys. Rev. Lett. 85.23 (2000): 4920-4923.
  • [66] P. A. Lee and T. V. Ramakrishnan, Disordered electronic systems. Rev. Mod. Phys. 57.2 (1985): 287.
  • [67] M. Leguil, J. You, Z. Zhao and Q. Zhou, Asymptotics of spectral gaps of quasi-periodic Schrödinger operators. arxiv.org:1712.04700
  • [68] L. Li, D. Damanik and Q. Zhou, Absolutely continuous spectrum for CMV operators with small quasi-periodic Verblunsky coefficients, arXiv:2102.00586
  • [69] Ricardo Mañé, Ergodic theory and differentiable dynamics, volume 8 of Ergebnisse der Mathematik und ihrer Grenzgebiete (3). Springer-Verlag, Berlin, 1987. Translated from the Portuguese by Silvio Levy.
  • [70] R. Peierls, Zur theorie des diamagnetismus von leitungselektronen, Z. Phys. 80.11-12 (1933): 763-791.
  • [71] G. Roati et al, Nature (London) 453, 895 (2008).
  • [72] W. D. Roeck, F. Huveneers, M. Müller and M. Schiulaz, Absence of many-body mobility edges, Phys. Rev. B 93.1 (2016): 014203.
  • [73] S. D. Sarma, S. He and X. C. Xie, Mobility edge in a model one-dimensional potential, Phys. Rev. Lett. 61.18 (1988): 2144.
  • [74] B. Simon, Schrödinger operators in the twenty-first century. Mathematical physics 2000, 283–288, Imp. Coll. Press, London, 2000.
  • [75] Y. Sinai, Anderson localization for one-dimensional difference Schrödinger operator with quasi-periodic potential, J. Stat. Phys. 46.5-6 (1987): 861-909.
  • [76] D. J. Thouless, M. Kohmoto, M. P. Nightingale and M. Den Nijs, Quantized Hall conductance in a two dimensional periodic potential, Phys. Rev. Lett. 49.6 (1982): 405-408.
  • [77] Y. Wang and Z. Zheng, Coexistence of zero Lyapunov exponent and positive Lyapunov exponent for new quasi-periodic Schrödinger operator. Not intended for publication. arXiv:2009.06189
  • [78] Y. Wang, X. Xia, L. Zhang, H. Yao, S. Chen, J. You, Q. Zhou and X. Liu, One-Dimensional Quasiperiodic Mosaic Lattice with Exact Mobility Edges, Phys. Rev. Lett. 125.19 (2020): 196604.
  • [79] X. B. Wei, C. Cheng, X. Gao and R. Mondaini, Investigating many-body mobility edges in isolated quantum systems, Phys. Rev. B 99.16 (2019): 165137.
  • [80] S. Zhang, Mixed spectral types for the one-frequency discrete quasi-periodic Schrödinger operator, Proc. Amer. Math. Soc. 144.6 (2016): 2603-2609.