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

Anderson Localization for Schrödinger Operators with Monotone Potentials over Circle Homeomorphisms

Jiranan Kerdboon University of California, Irvine [email protected]  and  Xiaowen Zhu University of Washington [email protected]
Abstract.

In this paper, we prove pure point spectrum for a large class of Schrödinger operators over circle maps with conditions on the rotation number going beyond the Diophantine. More specifically, we develop the scheme to obtain pure point spectrum for Schrödinger operators with monotone bi-Lipschitz potentials over orientation-preserving circle homeomorphisms with Diophantine or weakly Liouville rotation number. The localization is uniform when the coupling constant is large enough.

1. Introduction

The spectral theory of quasiperiodic Schrödinger operators has been the subject of extensive study over the past several decades due to its deep origins in physics and the richness of its unusual mathematical features. The general setup of a quasiperiodic operator is given by a family of operators Hλ,f,T,x{H_{\lambda,f,T,x}} acting on 2()\ell^{2}({\mathbb{Z}}), defined as

(Hλ,f,T,xψ)(n)=ψ(n+1)+ψ(n1)+λf(Tnx)ψ(n),({H_{\lambda,f,T,x}}\psi)(n)=\psi(n+1)+\psi(n-1)+\lambda f(T^{n}x)\psi(n), (1.1)

where x𝕋1x\in{\mathbb{T}}^{1}, TT is an irrational rotation on 𝕋1{\mathbb{T}}^{1} defined by Tx=Rαx=x+αTx=R_{\alpha}x=x+\alpha, with α\alpha\in{\mathbb{R}}\setminus{\mathbb{Q}}, and f:𝕋1f:{\mathbb{T}}^{1}\to{\mathbb{R}} is a potential function. Examples of such operators include f(x)=cos(x)f(x)=\cos(x) for the almost Mathieu operator or f(x)=tan(x)f(x)=\tan(x) for the Maryland model. One of the most interesting features of quasiperiodic operators is that their spectral type can often be fully characterized by the arithmetic properties of α\alpha (and/or xx) in many situations, as demonstrated in works such as [10, 5]. Since RαR_{\alpha} serves as a fundamental example of general circle homeomorphisms, a natural question arises: If TT is not a rotation but a more general circle homeomorphism with rotation number α\alpha, can we still determine or get some information of the spectral type by the arithmetic properties of α\alpha?

As one can imagine, the answer may vary depending on properties of ff, TT, and α\alpha. The study of (1.1) for general circle diffeomorphisms TT was initiated by [8] and [16]. In [8], the authors proved purely continuous spectrum for ff that is Hölder continuous, TT that is C1+BVC^{1+\operatorname{BV}}-smooth, and α\alpha that is super Liouville. [16] further explored similar phenomena for circle diffeomorphisms with a critical point or break. On the other hand, for quasiperiodic (1.1), a number of recent papers have proved the opposite, i.e., pure point spectrum, under arithmetic properties of α\alpha that go beyond the Diophantine condition, e.g., [19, 10, 1, 11, 12, 17, 5, 4, 18]. In this paper, we add to this list by extending the results in [7], where the authors worked with an irrational rotation T=RαT=R_{\alpha} with Diophantine α\alpha and potential ff satisfying conditions (\mathcal{F}1) and (\mathcal{F}2) below. We work with the same conditions on ff but consider a more general orientation-preserving circle homeomorphism TT under the assumptions (𝒯\mathcal{T}1):

  1. (\mathcal{F}1)

    ff is one-periodic on {\mathbb{R}} and f(0)=0f(0)=0, f(10):=limx1f(x)=1f(1-0):=\lim_{x\to 1^{-}}f(x)=1.

  2. (\mathcal{F}2)

    ff is bi-Lipschitz monotone, i.e., there exist γ,γ+>0\gamma_{-},\gamma_{+}>0 such that for all 0x<y<10\leq x<y<1,

    γ(yx)f(y)f(x)γ+(yx).\gamma_{-}(y-x)\leq f(y)-f(x)\leq\gamma_{+}(y-x).
  1. (𝒯\mathcal{T}1)

    Assume the invariant measure of TT is denoted by ν\nu and that

    Cν([x,y])|xy|C+ν([x,y]).C_{-}\nu([x,y])\leq|x-y|\leq C_{+}\nu([x,y]).

Under these conditions, we obtain results that are similar to the ones in [7]. In fact, in addition to extending to more general circle homeomorphisms, we also generalize the result by relaxing the Diophantine condition on α\alpha to both weakly Liouville and Diophantine case. Specifically, we prove that

Theorem 1 (pure point spectrum).

For ff satisfying (\mathcal{F}1), (\mathcal{F}2) and TT satisfying (𝒯\mathcal{T}1) with weakly Liouville or Diophatine rotation number α\alpha, or more specifically, 0β(α)<0\leq\beta(\alpha)<\infty, there is C0=C0(γ±,C±)=O(γCγ+C+)>0C_{0}=C_{0}(\gamma_{\pm},C_{\pm})=O\left(\frac{\gamma_{-}C_{-}}{\gamma_{+}C_{+}}\right)>0 such that for all λ>0\lambda>0, we have

σc(Hλ,f,T,x){E:β(α)<C0L(E;α)}=,x𝕋1,\sigma_{c}({H_{\lambda,f,T,x}})\cap\left\{E:\beta(\alpha)<C_{0}L(E;\alpha)\right\}=\emptyset,\quad\forall x\in{\mathbb{T}}^{1},

where β(α)\beta(\alpha) and the Lyapunov exponent L(E;α)L(E;\alpha) are defined in Sec 2.

Remark 1.

The theorem provides a meaningful statement for homeomorphisms with rotation number α\alpha when β(α)\beta(\alpha) is small or zero, which corresponds to weakly Liouville or Diophantine α\alpha (see Sec 2). In fact, the smaller β(α)\beta(\alpha) is, the more “irrational” α\alpha is. For example, since we also proved positivity of Lyapunov exponent L(E;α)>0L(E;\alpha)>0 for all irrational α\alpha in Corollary 3.3, when β(α)=0\beta(\alpha)=0, this implies that σc(Hf,T,x)=\sigma_{c}(H_{f,T,x})=\emptyset, i.e. the spectrum of Hf,T,xH_{f,T,x} is pure point.

Note that condition (𝒯\mathcal{T}1) is equivalent to the existence of a bi-Lipschitz conjugacy between TT and RαR_{\alpha}, meaning that there exists a bi-Lipschitz function ϕ\phi that is bounded from above and below such that ϕT=Rαϕ\phi\circ T=R_{\alpha}\circ\phi. We acknowledge that if such a bi-Lipschitz conjugacy exists and α\alpha is Diophantine, our results follow directly from [7] by a change of variables: letting y=ϕ(x)y=\phi(x), we obtain Hfϕ1,Rα,yH_{f\circ{\phi^{-1}},R_{\alpha},y} and can apply known localization results. However, we choose to present the proof in the more general setting using the invariant measure, since the existence of a bi-Lipschitz (in fact, C1+εC^{1+\varepsilon}) conjugacy is currently only established for Diophantine α\alpha.

As a byproduct, we also establish Lipschitz continuity of the integrated density of states (IDS, see (2.5)) for all λ\lambda in Lemma 3.3, as well as the continuity and positivity of the Lyapunov exponent for large λ\lambda in Corollary 3.3. Together with our key lemma 5.3, which is uniform in x,Ex,E, and α\alpha, these results allow us to achieve uniform localization of Hλ,f,T,x{H_{\lambda,f,T,x}} (see Definition 3) for sufficiently large λ\lambda and ”sufficiently irrational” α\alpha, i.e. β(α)\beta(\alpha) sufficiently small.

Theorem 2 (Uniform localization).

Let C0=C0(γ±,C±)>0C_{0}=C_{0}(\gamma_{\pm},C_{\pm})>0. If λ>4eγC\lambda>\frac{4e}{\gamma_{-}C_{-}} and if α\alpha is weakly Liouville with β(α)<C0ln(λγC4e)\beta(\alpha)<C_{0}\ln\left(\frac{\lambda\gamma_{-}C_{-}}{4e}\right), then Hλ,f,T,x{H_{\lambda,f,T,x}} has uniform localization for all xx.

We finally mention that a somewhat different proof was developed in [15] for unbounded lower-Lipschitz monotone ff and irrational rotation T=RαT=R_{\alpha} with Diophantine α\alpha. The key idea is, instead of controlling the change of eigenvalue functions horizontally (see Lemma 3.1), the author controls the change of counting function horizontally. We believe that the Lipschitz continuity of integrated density of states in our proof can be done through that of argument in [15] and the results there can be generalized to more general TT with weakly Liouville α\alpha in similar ways to here. This will be explored in a future work.

Structure and key ideas. Under the assumption of (𝒯\mathcal{T}1) and allowing weakly Liouville α\alpha, we re-develop the proof following the method in [7] in the key step: We use the non-perturbative proof of localization, first developed in [14], together with a detailed analysis of the behavior of box eigenvalues. We provide the latter for general TT in Sec 3, which helps with building the large deviation estimates in Sec 4. From the large deviation estimates, we get our key lemma on the uniform exponential decay of generalized eigenfunctions in x,α,Ex,\alpha,E in Sec 5. The main results follow immediately in Sec 6.

The extension of our results from RαR_{\alpha} to more general circle homeomorphisms TT is based on the observation that the behavior of box eigenvalues is closely related to the distribution of orbits of TT. While the orbits of TT are not evenly distributed with respect to distance, they are evenly distributed with respect to the invariant measure, allowing us to obtain quantitative estimates on their distribution under the comparability assumption (𝒯\mathcal{T}1). Appendix A provides the key statements that enable us to carry out this argument. The extension to weakly Liouville α\alpha, on the other hand, requires a more careful estimate of the decay of generalized eigenfunctions in Sec 5.

2. preliminaries

In this section, we will begin by discussing two fundamental concepts: continued fraction expansion and weakly Liouville numbers. Afterward, we will introduce several fundamental properties of discrete Schrödinger operators, including the generalized eigenvalue and Schnol’s theorem, the Green function and Poisson formula, transfer matrices and Lyapunov exponent, density of states measure, and the Thouless formula.

Notations. For xx\in{\mathbb{R}}, we use |x||x| to denote the absolute value and x=infn|xn|\|x\|=\inf\limits_{n\in{\mathbb{Z}}}|x-n| to denote the closest distance between xx\in{\mathbb{R}} and integers.

Continued fraction expansion and weakly Liouville number. Any number α[0,1)\alpha\in[0,1) can be written in the continued fraction expansion [20]:

α=1a1+1a2+1a3+:=[a1,a2,a3,].\displaystyle\alpha=\frac{1}{a_{1}+\frac{1}{a_{2}+\frac{1}{a_{3}+\ldots}}}:=[a_{1},a_{2},a_{3},\ldots].

with ak+a_{k}\in{\mathbb{N}}^{+}. Let pnqn=[a1,,an]\frac{p_{n}}{q_{n}}=[a_{1},\dots,a_{n}] denote the continued fraction approximants. They satisfy

pk=akpk1+pk2,p1=1,p0=0;qk=akqk1+qk2,q1=0,q0=1.\begin{split}p_{k}&=a_{k}p_{k-1}+p_{k-2},~{}p_{-1}=1,~{}p_{0}=0;\\ q_{k}&=a_{k}q_{k-1}+q_{k-2},~{}q_{-1}=0,~{}q_{0}=1.\end{split} (2.1)
Definition 1 (weakly Liouville).

For α[0,1)\alpha\in[0,1), let

β(α)=lim supklnqk+1qk.\beta(\alpha)=\limsup\limits_{k\to\infty}\frac{\ln q_{k+1}}{q_{k}}.

We call α\alpha weakly Liouville if 0<β(α)<0<\beta(\alpha)<\infty.

We mention that if α\alpha is Diophantine111α\alpha is called Diophantine if there is κ>0\kappa>0 and τ>0\tau>0 such that nα>κ|n|τ\|n\alpha\|>\frac{\kappa}{|n|^{\tau}} for all nn., then β(α)=0\beta(\alpha)=0.

For a detailed discussion on the next several definitions, please refer to [3, Ch.9,10] and [2, Ch.VII].

Generalized eigenfunction and Schnol’s theorem. We say ψ\psi is a generalized eigenfunction of an operator HH with respect to a generalized eigenvalue EE if ψ\psi is polynomially bounded, i.e. |ψ(n)|C(1+|n|)p|\psi(n)|\leq C(1+|n|)^{p} for some C>0C>0, pp\in{\mathbb{N}} and Hψ=EψH\psi=E\psi. Schnol’s theorem states that the spectral measure of an operator HH is supported on the set of its generalized eigenvalues.

According to Schnol’s theorem, to prove that HH has pure point spectrum, it is sufficient to show that all generalized eigenfunctions belong to 2\ell^{2}. This is because if all generalized eigenfunctions and eigenvalues become eigenfunctions and eigenvalues, respectively, then the spectrum is pure point.

Green function and Poisson formula. Let H[a,b](x){H_{[a,b]}(x)} and H~[a,b]\tilde{H}_{[a,b]} denote the restriction of Hλ,f,T,x{H_{\lambda,f,T,x}} to 2([a,b])\ell^{2}([a,b]) with Dirichlet and periodic boundary conditions, respectively. In particular, for the interval [a,b]=[0,n1][a,b]=[0,n-1], we use the simplified notations Hn(x)H_{n}(x) and H~n(x)\tilde{H}_{n}(x). More specifically,

Hn(x)=(λf(x)1111λf(Tn1x))n×n,H~n(x)=(λf(x)111111λf(Tn1x))n×n.\begin{split}&H_{n}(x)=\begin{pmatrix}\lambda f(x)&1&&\\ 1&\ddots&\ddots&\\ &\ddots&\ddots&1\\ &&1&\lambda f(T^{n-1}x)\end{pmatrix}_{n\times n,}\\ &\tilde{H}_{n}(x)=\begin{pmatrix}\lambda f(x)&1&&1\\ 1&\ddots&\ddots&\\ &\ddots&\ddots&1\\ 1&&1&\lambda f(T^{n-1}x)\end{pmatrix}_{n\times n.}\end{split}

Let Gx,E,[a,b]=(H[a,b](x)E)1{G_{x,E,[a,b]}}=({H_{[a,b]}(x)}-E)^{-1} denote the Green function, and let Gx,E,[a,b](m,n){G_{x,E,[a,b]}}(m,n) be its (m,n)(m,n)-entry. Denote Pn(x,E)=det(Hn(x)E)P_{n}(x,E)=\det(H_{n}(x)-E), and let P0(x,E)=1P_{0}(x,E)=1.

The Poisson formula provides a connection between the generalized eigenfunction and the Green function. Specifically, suppose ψ(n)\psi(n) is a generalized eigenfunction of Hλ,f,T,x{H_{\lambda,f,T,x}} with respect to generalized eigenvalue EE, then for nn in the interval [a,b][a,b], we have the following formula:

ψ(n)=Gx,E,[a,b](a,n)ψ(a1)Gx,E,[a,b](n,b)ψ(b+1).\psi(n)=-{G_{x,E,[a,b]}}(a,n)\psi(a-1)-{G_{x,E,[a,b]}}(n,b)\psi(b+1). (2.2)

Transfer matrix and Lyapunov exponent. Rewrite Hλ,f,T,xψ=Eψ{H_{\lambda,f,T,x}}\psi=E\psi into matrix form:

(ψnψn1)=An1(x,E)(ψn1ψn2)=An1(x,E)A0(x,E)(ψ0ψ1),\begin{pmatrix}\psi_{n}\\ \psi_{n-1}\end{pmatrix}=A_{n-1}(x,E)\begin{pmatrix}\psi_{n-1}\\ \psi_{n-2}\end{pmatrix}=A_{n-1}(x,E)\dots A_{0}(x,E)\begin{pmatrix}\psi_{0}\\ \psi_{-1}\end{pmatrix},

where

Ai(x,E):=(Eλf(Tix)110).A_{i}(x,E):=\begin{pmatrix}E-\lambda f(T^{i}x)&-1\\ 1&0\end{pmatrix}.

We define the nn-step transfer matrix by

Mn(x,E):=An1(x,E)A0(x,E).M_{n}(x,E):=A_{n-1}(x,E)\dots A_{0}(x,E).

One can verify by induction that

Mn(x,E):=(Pn(x,E)Pn1(Tx,E)Pn1(x,E)Pn2(Tx,E)).M_{n}(x,E):=\begin{pmatrix}P_{n}(x,E)&-P_{n-1}(Tx,E)\\ P_{n-1}(x,E)&-P_{n-2}(Tx,E)\end{pmatrix}. (2.3)

The Lyapunov exponent is defined to be

L(E):=limn1n01lnMn(x,E)dν(x).L(E):=\lim_{n\to\infty}\frac{1}{n}\int_{0}^{1}\ln\|M_{n}(x,E)\|\ d\nu(x). (2.4)

Integrated density of states (IDS) and the Thouless formula. Next, we introduce the density of states measure and the Thouless formula, which connects the Lyapunov exponent of EE with the density of states measure. The integrated density of states (IDS) is defined as follows:

N(E):=limn1n01Nn(x,E)𝑑ν(x),N(E):=\lim_{n\to\infty}\frac{1}{n}\int_{0}^{1}N_{n}(x,E)\ d\nu(x), (2.5)

where Nn(x,E):=#σ(Hn(x))(,E]N_{n}(x,E):=\#\sigma(H_{n}(x))\cap(-\infty,E].

Remark 2.

We can define P~n(x,E)\tilde{P}_{n}(x,E) and N~n(x,E)\tilde{N}_{n}(x,E), analogous to Pn(x,E)P_{n}(x,E) and Nn(x,E)N_{n}(x,E) for HnH_{n}, respectively, for H~n(x)\tilde{H}_{n}(x).

Remark 3.

Notice that H~n(x)\tilde{H}_{n}(x) is a rank-two perturbation of Hn(x)H_{n}(x). Thus we have |N~n(x,E)Nn(x,E)|2|\tilde{N}_{n}(x,E)-N_{n}(x,E)|\leq 2. Thus we can also define the IDS by

N(E)=limn1n01N~n(x,E)𝑑ν.N(E)=\lim_{n\to\infty}\frac{1}{n}\int_{0}^{1}\tilde{N}_{n}(x,E)\ d\nu. (2.6)

The function N(E)N(E) is right-continuous, non-decreasing, and approaches zero as EE approaches -\infty. Its derivative defines a unique probability measure, called the density of states measure N(dE)N(dE). The relation between the density of states measure N(dE)N(dE) and the Lyapunov exponents L(E)L(E) is known as the Thouless formula. We state it here without proof, but refer the interested reader to [3] for more details:

L(E)=ln|EE|N(dE).L(E)=\int_{\mathbb{R}}\ln|E^{\prime}-E|\,N(dE). (2.7)

3. Positive Lyapunov Exponent

In this section, we first establish some fundamental properties of box eigenvalue functions, which are the eigenvalues of H~n(x)\tilde{H}_{n}(x). We then derive estimates for the distance between these eigenvalue functions. Using these estimates, we obtain the Lipschitz continuity of the IDS N(E)N(E) with respect to EE and prove the positivity of the Lyapunov exponent L(E)L(E) for large λ\lambda.

Recall that H~n(x)\tilde{H}_{n}(x) is the periodic restriction of Hλ,f,T,x{H_{\lambda,f,T,x}} to [0,n1][0,n-1]. Let μ~m(x)\tilde{\mu}_{m}(x), 0mn10\leq m\leq n-1 be the eigenvalues of H~n(x)\tilde{H}_{n}(x) in increasing order. We refer to μ~m(x)\tilde{\mu}_{m}(x) as the box eigenvalue functions. Now we establish some of their basic properties:

Proposition 3.1.

μ~i(x)\tilde{\mu}_{i}(x) have the following properties:

  1. (1)

    μ~i(x)\tilde{\mu}_{i}(x) is 11-periodic, continuous on [0,1)[0,1) except at {Tj(0)}j=0n1\{T^{-j}(0)\}_{j=0}^{n-1}. By rearranging these discontinuity points in an increasing order, we denote them by {βl}l=0n1\{\beta_{l}\}_{l=0}^{n-1}. We also denote Il:=[βl,βl+1)I_{l}:=[\beta_{l},\beta_{l+1}).

  2. (2)

    μ~i(x)\tilde{\mu}_{i}(x) is bi-Lipschitz continuous with respect to the invariant measure, and strictly increasing on each IlI_{l}. In fact,

    λγCν([x,y])μi(y)μi(x)λγ+C+ν([x,y]).\lambda\gamma_{-}C_{-}\nu([x,y])\leq\mu_{i}(y)-\mu_{i}(x)\leq\lambda\gamma_{+}C_{+}\nu([x,y]). (3.1)
  3. (3)

    At each jump βl\beta_{l}, we have

    μ~i(βl0)μ~i+1(βl)μ~i+1(βl0),0in2.\tilde{\mu}_{i}(\beta_{l}-0)\leq\tilde{\mu}_{i+1}(\beta_{l})\leq\tilde{\mu}_{i+1}(\beta_{l}-0),\quad 0\leq i\leq n-2.
Remark 4.

Because of (2) and (3) above, it is natural to define

Λi(x):=l=0n1μi+l(x)χIl(x),0in1\Lambda_{i}(x):=\sum_{l=0}^{n-1}\mu_{i+l}(x)\chi_{I_{l}}(x),\quad 0\leq i\leq n-1

and extend it periodically from [0,1)[0,1) to {\mathbb{R}}. As a result, Λj(x)\Lambda_{j}(x) is monotone increasing on [βnj+1+N,βnj+1+N+1)[\beta_{n-j+1}+N,\beta_{n-j+1}+N+1) for any NN\in{\mathbb{Z}}, and it inherits the properties of μj(x)\mu_{j}(x) on each IlI_{l}. In particular, Λj(x)\Lambda_{j}(x) is also lower-Lipschitz with respect to invariant measure ν\nu on [βnj+1+N,βnj+1+N+1)[\beta_{n-j+1}+N,\beta_{n-j+1}+N+1):

Λi(y)Λi(x)λγCν([x,y]), for x<y and x,y[βnj+1+N,βnj+1+N+1).\Lambda_{i}(y)-\Lambda_{i}(x)\geq\lambda\gamma_{-}C_{-}\nu([x,y]),\quad\text{~{}for~{}}x<y\text{~{}and~{}}x,y\in[\beta_{n-j+1}+N,\beta_{n-j+1}+N+1).
Proof.
  1. (1)

    Note that the box eigenvalue functions μ~i(x)\tilde{\mu}_{i}(x) are roots of the characteristic polynomial P~n(x,E)\tilde{P}_{n}(x,E). Therefore, each μ~i(x)\tilde{\mu}_{i}(x) is continuous with respect to the coefficients of P~n(x,E)\tilde{P}_{n}(x,E), which are polynomials of {λf(Tjx)}j=0qk1\{\lambda f(T^{j}x)\}_{j=0}^{q_{k}-1}. Since f(Tjx)f(T^{j}x) is only discontinuous at Tj(x)T^{-j}(x), μ~i(x)\tilde{\mu}_{i}(x) is only potentially discontinuous at {Tj(x)}j=0qk1\{T^{-j}(x)\}_{j=0}^{q_{k}-1}.

  2. (2)

    Notice that for x<yx<y in the same IlI_{l}, H~(y)H~(x)\tilde{H}(y)-\tilde{H}(x) is a non-negative diagonal matrix. By Lidskii’s theorem (see [9, Theorem 2]), we have

    λminj(f(Tjy)f(Tjy))μj(y)μj(x)λmaxj(f(Tjy)f(Tjx)).\lambda\min\limits_{j}\left(f(T^{j}y)-f(T^{j}y)\right)\leq\mu_{j}(y)-\mu_{j}(x)\leq\lambda\max\limits_{j}\left(f(T^{j}y)-f(T^{j}x)\right).

    Notice that (\mathcal{F}2) and (𝒯\mathcal{T}1) implies that

    f(Tjy)f(Tjx)γ+C+ν([Tjx,Tjy]),f(T^{j}y)-f(T^{j}x)\leq\gamma_{+}C_{+}\nu([T^{j}x,T^{j}y]),

    and similarly, f(Tjy)f(Tjx)λγCν([x,y])f(T^{j}y)-f(T_{j}x)\geq\lambda\gamma_{-}C_{-}\nu([x,y]).

  3. (3)

    Notice that

    H~n(βl0)H~n(βl)=λejej\tilde{H}_{n}(\beta_{l}-0)-\tilde{H}_{n}(\beta_{l})=\lambda e_{j}\otimes e_{j} (3.2)

    where 0jqk10\leq j\leq q_{k}-1 such that T(j1)(x)=βlT^{-{(j-1)}}(x)=\beta_{l}. This leads to the second inequality since H~n(βl0)H~n(βl)\tilde{H}_{n}(\beta_{l}-0)-\tilde{H}_{n}(\beta_{l}) is positive semi-definite. To derive the first inequality, by (3.2), let DD be the matrix obtained by deleting the row jj and column jj from H~n(βl0)\tilde{H}_{n}(\beta_{l}-0) or H~n(βl)\tilde{H}_{n}(\beta_{l}). Let ω1ω2ωn1\omega_{1}\leq\omega_{2}\leq\ldots\leq\omega_{n-1} be the eigenvalues of DD. By eigenvalue interlacing theorem,

    μ~0(βl0)ω1μ~1(βl0)ω2ωn1μ~n1(βl0),μ~0(βl)ω1μ~1(βl)ω2ωn1μ~n1(βl).\begin{split}\ \ \ \ &\tilde{\mu}_{0}(\beta_{l}-0)\leq\omega_{1}\leq\tilde{\mu}_{1}(\beta_{l}-0)\leq\omega_{2}\leq\dots\leq\omega_{n-1}\leq\tilde{\mu}_{n-1}(\beta_{l}-0),\\ \ \ \ \ &\tilde{\mu}_{0}(\beta_{l})\leq\omega_{1}\leq\tilde{\mu}_{1}(\beta_{l})\leq\omega_{2}\leq\dots\leq\omega_{n-1}\leq\tilde{\mu}_{n-1}(\beta_{l}).\end{split} (3.3)

    Therefore, μ~m(βl0)ωm+1μ~m+1(βl)\tilde{\mu}_{m}(\beta_{l}-0)\leq\omega_{m+1}\leq\tilde{\mu}_{m+1}(\beta_{l}), for all 0mn20\leq m\leq n-2.

Horizontal comparison. From now on, we fix α\alpha, and consider n=qkn=q_{k} since we will use the dynamical properties of the irrational circle map to compare box eigenvalue functions horizontally and vertically. The following lemma provides an upper bound control if we compare the box eigenvalue functions μ~i(x)\tilde{\mu}_{i}(x) and μ~i(Trx)\tilde{\mu}_{i}(T^{r}x) horizontally. Note that the estimate is uniform in rr.

Lemma 3.1.

For any qk+1rqk1-q_{k}+1\leq r\leq q_{k}-1,

|μ~i(x)μ~i(Trx)|λγ+C+qk+1.\displaystyle\left|{\tilde{\mu}}_{i}(x)-{\tilde{\mu}}_{i}(T^{r}x)\right|\leq\frac{\lambda\gamma_{+}C_{+}}{q_{k+1}}.
Proof.

Define an qk×qkq_{k}\times q_{k} unitary matrix S=[eqk,e1,e2,,eqk1]S=[{e_{q_{k}}},{e_{1}},{e_{2}},\dots,{e_{{q_{k}}-1}}] where ejne_{j}\in{\mathbb{R}}^{n} are standard unit vectors. Then

SrH~qk(x)Sr=H~qkr(Trx)H~r(x),H~qk(Trx)=H~qkr(Trx)H~r(Tqkx).\begin{split}&S^{r}\tilde{H}_{q_{k}}(x)S^{-r}=\tilde{H}_{q_{k}-r}(T^{r}x)\oplus\tilde{H}_{r}(x),\\ &\tilde{H}_{q_{k}}(T^{r}x)=\tilde{H}_{q_{k}-r}(T^{r}x)\oplus\tilde{H}_{r}(T^{q_{k}}x).\end{split}

By Lemma A.2,

H~r(x)H~r(Tqkx)λmax0ir1|f(Trx)f(Tqk+rx)|λγ+C+ν([x,Tqkx])λγ+C+qk+1.\|\tilde{H}_{r}(x)-\tilde{H}_{r}(T^{q_{k}}x)\|\leq\lambda\max\limits_{0\leq i\leq r-1}|f(T^{r}x)-f(T^{q_{k}+r}x)|\leq\lambda\gamma_{+}C_{+}\nu([x,T^{q_{k}}x])\leq\frac{\lambda\gamma_{+}C_{+}}{q_{k+1}}.

The result follows from Lindskii’s theorem. ∎

Corollary 3.2.

For any x,y[0,1)x,y\in[0,1),

|μ~i(x)μ~i(y)|λγ+C+qk+1+2λγ+C+qk3λγ+C+qk.|\tilde{\mu}_{i}(x)-\tilde{\mu}_{i}(y)|\leq\frac{\lambda\gamma_{+}C_{+}}{q_{k+1}}+\frac{2\lambda\gamma_{+}C_{+}}{q_{k}}\leq\frac{3\lambda\gamma_{+}C_{+}}{q_{k}}.
Proof.

First notice that for given x[0,1)x\in[0,1), depending on which IkI_{k} it belongs to, there exists qk+1α0-q_{k}+1\leq\alpha\leq 0, such that each point in {Trx}r=αα+qk1\{T^{r}x\}_{r=\alpha}^{\alpha+q_{k}-1} precisely falls in one interval among {Il}l=0qk1\{I_{l}\}_{l=0}^{q_{k}-1}. Thus there is qk+1rqk1-q_{k}+1\leq r\leq q_{k}-1 such that TrxT^{r}x and yy are in the same IkI_{k}. Then

|μ~i(x)μ~i(y)||μ~i(x)μ~i(Trx)|+|μ~i(Trx)μi(y)|λγ+C+qk+1+λγ+C+ν([Trx,y])λγ+C+qk+1+λγ+C+(1qk+1qk+1)\begin{split}|\tilde{\mu}_{i}(x)-\tilde{\mu}_{i}(y)|&\leq|\tilde{\mu}_{i}(x)-\tilde{\mu}_{i}(T^{r}x)|+|\tilde{\mu}_{i}(T^{r}x)-\mu_{i}(y)|\\ &\leq\frac{\lambda\gamma_{+}C_{+}}{q_{k+1}}+\lambda\gamma_{+}C_{+}\nu([T^{r}x,y])\\ &\leq\frac{\lambda\gamma_{+}C_{+}}{q_{k+1}}+\lambda\gamma_{+}C_{+}\left(\frac{1}{q_{k}}+\frac{1}{q_{k+1}}\right)\end{split}

where the first inequality follows from Lemma 3.1 and the second follows from Lemma A.2. ∎

Vertical comparison. We now estimate the lower bound of vertical comparison between eigenvalue functions. Unfortunately, the vertical distance between two closest eigenvalue functions μ~i(x)\tilde{\mu}_{i}(x) and μ~i+1(x)\tilde{\mu}_{i+1}(x) is not always positive. However, we can show that at most MM eigenvalues can be very close to each other, others will be nicely seperated from them.

Lemma 3.2.

Given γ±\gamma_{\pm}, λ\lambda and C±C_{\pm}. For any ε>0\varepsilon>0, there is a j0=j0(ε)=2γ+C+εγCj_{0}=j_{0}(\varepsilon)=\frac{2\gamma_{+}C_{+}}{\varepsilon\gamma_{-}C_{-}}, such that for any i,j,qki,j,q_{k} satisfying jj0j\geq j_{0} and 0i<i+jqk10\leq i<i+j\leq q_{k}-1, we have

|μ~i+j(x)μ~i(x)|λγC(1ε)jqk=:d0(ε)jqk.|\tilde{\mu}_{i+j}(x)-\tilde{\mu}_{i}(x)|\geq\lambda\gamma_{-}C_{-}(1-\varepsilon)\frac{j}{q_{k}}=:d_{0}(\varepsilon)\frac{j}{q_{k}}. (3.4)
Proof.

First notice that given xx, there exists qk+1α0-q_{k}+1\leq\alpha\leq 0, such that each point in {Trx}r=αα+qk1\{T^{r}x\}_{r=\alpha}^{\alpha+q_{k}-1} falls in precisely one interval among {Il}l=0qk1\{I_{l}\}_{l=0}^{q_{k}-1}. Then for any αr,rα+qk1\alpha\leq r,r^{\prime}\leq\alpha+q_{k}-1,

|μ~i+j(x)μ~i(x)||μ~i+j(Trx)μ~i(Trx)||μ~i+j(x)μ~i+j(Trx)||μ~i(Trx)μ~i(x)||μ~i+j(Trx)μ~i(Trx)|2λγ+C+qk+1supr,r|μ~i+j(Trx)μ~i(Trx)|2λγ+C+qk+1.\begin{split}|\tilde{\mu}_{i+j}(x)-\tilde{\mu}_{i}(x)|\geq&|\tilde{\mu}_{i+j}(T^{r}x)-\tilde{\mu}_{i}(T^{r^{\prime}}x)|-|\tilde{\mu}_{i+j}(x)-\tilde{\mu}_{i+j}(T^{r}x)|-|\tilde{\mu}_{i}(T^{r^{\prime}}x)-\tilde{\mu}_{i}(x)|\\ \geq&|\tilde{\mu}_{i+j}(T^{r}x)-\tilde{\mu}_{i}(T^{r^{\prime}}x)|-\tfrac{2\lambda\gamma_{+}C_{+}}{q_{k+1}}\\ \geq&\sup\limits_{r,r^{\prime}}|\tilde{\mu}_{i+j}(T^{r}x)-\tilde{\mu}_{i}(T^{r^{\prime}}x)|-\tfrac{2\lambda\gamma_{+}C_{+}}{q_{k+1}}.\end{split}

In particular, we can pick r,rr,r^{\prime} such that (Trx,μ~i(Trx))(T^{r^{\prime}}x,\tilde{\mu}_{i}(T^{r^{\prime}}x)) and (Trx,μ~i+j(Trx))(T^{r}x,\tilde{\mu}_{i+j}(T^{r}x)) are on the graph of the same Λm\Lambda_{m}, defined in Remark 4. Put such pairs of (r,r)(r,r^{\prime}) together and denote the set by SjS_{j}. Then [Trx,Trx][T^{r}x,T^{r^{\prime}}x] includes jj out of qkq_{k} subintervals created by the partition {Tix}i=αα+qk1\{T^{i}x\}_{i=\alpha}^{\alpha+q_{k}-1} on [0,1)[0,1), where each intervals have the same invariant measure. Thus by pigeonhole principle,111In fact, we could bound |μ~i+j(Trx)μ~i(Trx)||\tilde{\mu}_{i+j}(T^{r}x)-\tilde{\mu}_{i}(T^{r^{\prime}}x)|, the distance between eigenvalue functions, directly by Lemma A.2 without taking the supremum or referring to the pigeonhole principle. However, the authors choose to prove it this way both because it is more interesting, and because it reveals the uniformity in xx in the vertical comparison of eigenvalue functions. It implies that vertical differences of eigenvalue functions at any xx is uniformly controlled by the largest vertical differences among all Λm\Lambda_{m}. This observation can be useful in dealing with singular ν\nu where certain IkI_{k}’s are too small or the case when ff is flat at some IkI_{k}’s.

supr,r|μ~i+j(Trx)μ~i(Trx)|λγCsupr,rSjν([Trx,Trx])λγCjqk.\sup\limits_{r,r^{\prime}}|\tilde{\mu}_{i+j}(T^{r}x)-\tilde{\mu}_{i}(T^{r^{\prime}}x)|\geq\lambda\gamma_{-}C_{-}\sup\limits_{r,r^{\prime}\in S_{j}}\nu([T^{r}x,T^{r^{\prime}}x])\geq\lambda\gamma_{-}C_{-}\frac{j}{q_{k}}.

Thus

|μ~i+j(x)μ~i(x)|λγCjqk2λγ+C+qk+1λγCjqk(12γ+C+γCqkqk+1j)λγC(1ε0)jqk|\tilde{\mu}_{i+j}(x)-\tilde{\mu}_{i}(x)|\geq\lambda\gamma_{-}C_{-}\frac{j}{q_{k}}-\frac{2\lambda\gamma_{+}C_{+}}{q_{k+1}}\geq\lambda\gamma_{-}C_{-}\frac{j}{q_{k}}\left(1-\tfrac{2\gamma_{+}C_{+}}{\gamma_{-}C_{-}}\tfrac{q_{k}}{q_{k+1}j}\right)\geq\lambda\gamma_{-}C_{-}(1-\varepsilon_{0})\frac{j}{q_{k}}

when jj0:=2γ+C+εγCj\geq j_{0}:=\frac{2\gamma_{+}C_{+}}{\varepsilon\gamma_{-}C_{-}}.

Lipschitz continuity of IDS. Recall that N~n(x,E)=#σ(H~n(x))(,E]\tilde{N}_{n}(x,E)=\#\sigma(\tilde{H}_{n}(x))\cap(-\infty,E] and

N(E)=limn1n01N~n(x,E)𝑑ν(x).N(E)=\lim_{n\to\infty}\frac{1}{n}\int_{0}^{1}\tilde{N}_{n}(x,E)\ d\nu(x). (3.5)

Now we can derive Lipschitz continuity of N~qk(x,E)\tilde{N}_{q_{k}}(x,E) and N(E)N(E) from vertical distance of μ~i(x)\tilde{\mu}_{i}(x):

Lemma 3.3.

Given λ,γ+,γ\lambda,\gamma_{+},\gamma_{-}, EE, EE^{\prime}\in{\mathbb{R}}, we have

|N(E)N(E)||EE|λγC.|N(E)-N(E^{\prime})|\leq\frac{|E-E^{\prime}|}{\lambda\gamma_{-}C_{-}}. (3.6)
Proof.

Fix EE and EE^{\prime}. For any ε>0\varepsilon>0, we see from Lemma 3.2 that any interval of length d0(ε)j0qkd_{0}(\varepsilon)\tfrac{j_{0}}{q_{k}} contains at most j0j_{0} eigenvalues for qkq_{k} large enough. This allows us to estimate the number of eigenvalues between EE and EE^{\prime}:

|N~qk(E,x)N~qk(E,x)|(qk|EE|d0(ε)j0+1)j0=qk|EE|d0(ε)(1+j0d0(ε)qk|EE|)\begin{split}|\tilde{N}_{q_{k}}(E,x)-\tilde{N}_{q_{k}}(E^{\prime},x)|&\leq\left(\frac{q_{k}|E-E^{\prime}|}{d_{0}(\varepsilon)j_{0}}+1\right)j_{0}\\ &=\frac{q_{k}|E-E’|}{d_{0}(\varepsilon)}\left(1+\frac{j_{0}d_{0}(\varepsilon)}{q_{k}|E-E^{\prime}|}\right)\end{split}

for any xx. Let kk\to\infty, we get

|N(E)N(E)|lim infk|EE|d0(ε)(1+j0d0(ε)qk|EE|)=|EE|d0(ε)=|EE|λγC(1ε).|N(E)-N(E^{\prime})|\leq\liminf\limits_{k\to\infty}\frac{|E-E^{\prime}|}{d_{0}(\varepsilon)}\left(1+\frac{j_{0}d_{0}(\varepsilon)}{q_{k}|E-E^{\prime}|}\right)=\frac{|E-E^{\prime}|}{d_{0}(\varepsilon)}=\frac{|E-E^{\prime}|}{\lambda\gamma_{-}C_{-}(1-\varepsilon)}.

Since this inequality is true for all ε\varepsilon, the result follows. ∎

Positivity of Lyapunov exponent. This is a corollary of Lemma 3.3 which is also useful in the later proof of uniform localization.

Corollary 3.3.

The Lyapunov exponent L(E)L(E) of Hλ,f,T,x{H_{\lambda,f,T,x}} is continuous in EE and L(E)L(E) admits a lower bound

L(E)max{0,ln(λγC2e)}.L(E)\geq\max\left\{0,\ln\left(\frac{\lambda\gamma_{-}C_{-}}{2e}\right)\right\}. (3.7)

Therefore, L(E)L(E) is uniformly positive if λ>2eγC\lambda>\frac{2e}{\gamma_{-}C_{-}}.

Proof.

By Lemma 3.3, dN(E)dN(E) is absolutely continuous with respect to dEdE and the Radon-Nikodym derivative dN(E)dE1λγC:=1d\frac{dN(E)}{dE}\leq\frac{1}{\lambda\gamma_{-}C_{-}}:=\frac{1}{d}, for a.e.Ea.e.E. Thus by Thouless formula,

L(E)=ln|EE|dN(E)=(ln|EE|)dN(E)dE𝑑EEd2E+d21dln|EE|dE=2d0d2ln|E|dE=lnd2e=lnλγC2e,\begin{split}L(E)&=\int_{\mathbb{R}}\ln|E^{\prime}-E|dN(E^{\prime})=\int_{\mathbb{R}}(\ln|E^{\prime}-E|)\frac{dN(E^{\prime})}{dE^{\prime}}dE^{\prime}\geq\int_{E-\frac{d}{2}}^{E+\frac{d}{2}}\frac{1}{d}\cdot\ln|E^{\prime}-E|\ dE^{\prime}\\ &=\frac{2}{d}\int_{0}^{\frac{d}{2}}\ln|E^{\prime}|\ dE^{\prime}=\ln\tfrac{d}{2e}=\ln\tfrac{\lambda\gamma_{-}C_{-}}{2e},\end{split}

where the first inequality follows from monotonicity of ln\ln function and boundedness of f(E)f(E^{\prime}). Finally notice that L(E)0L(E)\geq 0 follows from the definition. Thus we get (3.7). ∎

4. Large deviation theorem

In this section, we provide two essential ingredients for the proof of localization: Lemma 4.1 provides an upper bound of Pn(x,E)P_{n}(x,E) while Theorem 3 provides the large deviation estimate which is central of the non-perturbative proofs of localization, as introduced in [14]. The first is a result that can be directly adapted from [13, Lemma 3.5]. It holds for arbitrary α\alpha\in{\mathbb{R}}\setminus{\mathbb{Q}} and arbitrary piecewise potentials.

Lemma 4.1.

For any κ>0\kappa>0 and EE\in{\mathbb{R}}, there exists an NN\in{\mathbb{N}} such that for all n>Nn>N

|Pn(x,E)|en(L(E)+κ),x[0,1).\displaystyle|P_{n}(x,E)|\leq e^{n(L(E)+\kappa)},\quad\forall x\in[0,1).

Moreover, NN can be chosen to be uniform in EIE\in I as long as L(E)L(E) is continuous on interval II.

Proof.

This was proved in [13, Lemma 3.5] for irrational rotation T=RαT=R_{\alpha}. The same method applies to a general circle diffeomorphism under the assumption (𝒯\mathcal{T}1). ∎

Theorem 3 (Large deviation theorem).

Fix EE such that L(E)>0L(E)>0. There exists C0=C0(γ±,C±)>0C_{0}=C_{0}(\gamma_{\pm},C_{\pm})>0 such that for any 0<δ<L(E)0<\delta<L(E), there is k0k_{0} such that for any kk0k\geq k_{0},

ν{x[0,1):1qkln|Pqk(x,E)|<L(E)δ}<eC0δqk\nu\{x\in[0,1):\tfrac{1}{q_{k}}\ln|P_{q_{k}}(x,E)|<L(E)-\delta\}<e^{-C_{0}\delta q_{k}} (4.1)

where mm is the Lebesgue measure. Moreover, the set on the left-hand side is composed of at most qkq_{k} many intervals.

Proof.

Recall that

Pqk(x,E):=det(Hqk(x)E)=i=0qk1(μi(x)E).P_{q_{k}}(x,E):=\det(H_{q_{k}}(x)-E)=\prod_{i=0}^{q_{k}-1}\left(\mu_{i}(x)-E\right).

Denote for convenience

fqk(x):=1qkln|Pqk(x;E)|=1qki=0qk1ln|μi(x)E|.f_{q_{k}}(x):=\frac{1}{q_{k}}\ln|P_{q_{k}}(x;E)|=\frac{1}{q_{k}}\sum\limits_{i=0}^{q_{k}-1}\ln|\mu_{i}(x)-E|.

Notice that μi(x)\mu_{i}(x) is monotone and fqk(x)=f_{q_{k}}(x)=-\infty at {x:μi(x)=E for some i}\{x:\mu_{i}(x)=E\text{~{}for~{}some~{}}i\}. Thus “large deviation” happens near {x:μi(x)=E for some i}\{x:\mu_{i}(x)=E\text{~{}for~{}some~{}}i\}. The aim is to estimate how large this set can be without rising fqk(x)f_{q_{k}}(x) too high. The idea is since μi(x)\mu_{i}(x) are well-seperated, only the closest (to EE) several μi(x)\mu_{i}(x) contribute the most to the negativity of fqk(x)f_{q_{k}}(x), the rest are nicely controlled.

To do so, we split eigenvalues μ~j(x)\tilde{\mu}_{j}(x) into three clusters: 𝒦+\mathcal{K}^{+} above EE, 𝒦0\mathcal{K}^{0} around EE, and 𝒦\mathcal{K}^{-} below EE. Notice that by |Nqk(x;E)N~qk(x;E)|2|N_{q_{k}}(x;E)-\tilde{N}_{q_{k}}(x;E)|\leq 2 and Lemma 3.2, we can make sure that

  1. (1)

    The cluster of eigenvalues above EE, denoted by μi+(x)\mu_{i}^{+}(x), i=1,2,i=1,2,\cdots in an increasing order with μi+(x)E+id0qk\mu_{i}^{+}(x)\geq E+\frac{id_{0}}{q_{k}}.

  2. (2)

    The cluster of eigenvalues below EE, denoted by μi(x)\mu_{i}^{-}(x), i=1,2,i=1,2,\cdots in an decreasing order with μi+(x)Eid0qk\mu_{i}^{+}(x)\leq E-\frac{id_{0}}{q_{k}}.

  3. (3)

    The cluster of the rest of eigenvalues, denoted by μi0(x)\mu_{i}^{0}(x), with the number of eigenvalues in this cluster does not exceed some N0N_{0} uniform in EE.

For example, this can be achieved by considering the closest 2j0+42j_{0}+4 eigenvalues μi(x)\mu_{i}(x) to EE to be in the third cluster and every eigenvalue above/below them to be in the first/second cluster. Here j0j_{0} is to guarantee the lower and upper bound estimates above and 4=2×24=2\times 2 is due to |Nqk(x;E)N~qk(x;E)|2|N_{q_{k}}(x;E)-\tilde{N}_{q_{k}}(x;E)|\leq 2. In fact, we can do the same thing for μ~i(x)\tilde{\mu}_{i}(x), then we just need to pick the closest 2j02j_{0} eigenvalues instead of 2j0+42j_{0}+4.

Now decompose PqkP_{q_{k}}, P~qk\tilde{P}_{q_{k}} correspondingly,

Pqk(x;E)=Pqk+(x;E)Pqk0(x;E)Pqk(x;E),P~qk(x;E)=P~qk+(x;E)P~qk0(x;E)P~qk(x;E),\begin{split}&P_{q_{k}}(x;E)=P_{q_{k}}^{+}(x;E)P_{q_{k}}^{0}(x;E)P_{q_{k}}^{-}(x;E),\\ &\tilde{P}_{q_{k}}(x;E)=\tilde{P}_{q_{k}}^{+}(x;E)\tilde{P}_{q_{k}}^{0}(x;E)\tilde{P}_{q_{k}}^{-}(x;E),\end{split}

where Pqk(x;E)=μi𝒦μi(x)EP_{q_{k}}^{*}(x;E)=\prod\limits_{\mu_{i}^{*}\in\mathcal{K}^{*}}\mu_{i}^{*}(x)-E, where {+,,0}*\in\{+,-,0\}.

Claim 1.

Let a,b>0a,b>0,

j=1n[ln(aj+b)ln(aj)]j=1nln(1+baj)j=1nbajbaln(n+1).\sum\limits_{j=1}^{n}\left[\ln(aj+b)-\ln(aj)\right]\leq\sum_{j=1}^{n}\ln\left(1+\tfrac{b}{aj}\right)\leq\sum_{j=1}^{n}\frac{b}{aj}\leq\frac{b}{a}\ln(n+1).

By Corollary 3.2 and the claim, we have for any x,y[0,1)x,y\in[0,1),

|ln|P~qk±(x;E)|ln|P~qk±(y;E)||i=1qk[ln(jd0qk+3λγ+C+qk)ln(jd0qk)]Clnqk.\begin{split}\left|\ln|\tilde{P}^{\pm}_{q_{k}}(x;E)|-\ln|\tilde{P}_{q_{k}}^{\pm}(y;E)|\right|&\leq\sum\limits_{i=1}^{q_{k}}\left[\ln\left(\tfrac{jd_{0}}{q_{k}}+\tfrac{3\lambda\gamma_{+}C_{+}}{q_{k}}\right)-\ln\left(\tfrac{jd_{0}}{q_{k}}\right)\right]\leq C\ln q_{k}.\end{split}

Here we considered all maximum potential perturbation of all μi(x)\mu_{i}(x) at the maximum potential place {jd0qk}j=1qk\{\tfrac{jd_{0}}{q_{k}}\}_{j=1}^{q_{k}}. There might be extra terms of μi±(x)\mu_{i}^{\pm}(x) that does not pair to μi±(y)\mu_{i}^{\pm}(y) but since there are only finitely many terms and they are bounded, the result is still true with a modification of CC. For the same reason, the inequality holds for Pqk(x;E)P_{q_{k}}(x;E) as well.

Thus there is Lqk(E)L_{q_{k}}(E) such that

Lqk(E)1qkln|Pqk+(x;E)Pqk(x;E)|Lqk(E)+Clnqkqk,Lqk(E)1qkln|P~qk+(x;E)P~qk(x;E)|Lqk(E)+Clnqkqk,\begin{split}&L_{q_{k}}(E)\leq\frac{1}{q_{k}}\ln|P_{q_{k}}^{+}(x;E)P_{q_{k}}^{-}(x;E)|\leq L_{q_{k}}(E)+C\frac{\ln q_{k}}{q_{k}},\\ &L_{q_{k}}(E)\leq\frac{1}{q_{k}}\ln|\tilde{P}_{q_{k}}^{+}(x;E)\tilde{P}_{q_{k}}^{-}(x;E)|\leq L_{q_{k}}(E)+C\frac{\ln q_{k}}{q_{k}},\end{split} (4.2)
Claim 2.

There is C0=C0(j0)=C0(γ±,C±)C_{0}=C_{0}(j_{0})=C_{0}(\gamma_{\pm},C_{\pm}) such that for kk large enough, for δ>0\delta>0 small enough,

ν{x[0,1):1qkln|Pqk(x,E)|<Lqk(E)δ}<eC0δqk.\nu\{x\in[0,1):\tfrac{1}{q_{k}}\ln|P_{q_{k}}(x,E)|<L_{q_{k}}(E)-\delta\}<e^{-C_{0}\delta q_{k}}.
Proof.

If xx is such that 1qkln|Pqk(x;E)|Lqkδ\frac{1}{q_{k}}\ln|P_{q_{k}}(x;E)|\leq L_{q_{k}}-\delta, then 1qkln|Pqk0(x;E)|δ\frac{1}{q_{k}}\ln|P_{q_{k}}^{0}(x;E)|\leq-\delta. Since there are at most N0N_{0} eigenvalues in 𝒦0(x)\mathcal{K}^{0}(x), thus there is some ll such that

1qkln|μl(x)E|δ/N0|μl(x)E|eδqk/N0.\frac{1}{q_{k}}\ln|\mu_{l}(x)-E|\leq-\delta/N_{0}\quad\Rightarrow|\mu_{l}(x)-E|\leq e^{-\delta q_{k}/N_{0}}. (4.3)

Among all x[0,1)x\in[0,1), there are at most qkq_{k} intervals of xx such that some μl(x)E\mu_{l}(x)-E satisfies (4.3). In fact, there are at most qkq_{k} intersections of graph(Λj)¯\overline{\text{graph}(\Lambda_{j})} and [0,1)×{E}[0,1)\times\{E\}. Since each Λj\Lambda_{j} is monotone, (4.3) is only possible for xx near such intersections. And by Prop. 3.1 μi(x)\mu_{i}(x) are lower-Lipschitz with respect to invariant measure. Thus for qkq_{k} large enough,

ν{x: there is ls.t.|μl(x)E|eδqk/N0}qkeδqk/N0λγCeδqk2N0eC0δqk.\nu\{x:\text{~{}there~{}is~{}}l{~{}s.t.~{}}|\mu_{l}(x)-E|\leq e^{-\delta q_{k}/N_{0}}\}\leq q_{k}\frac{e^{-\delta q_{k}/N_{0}}}{\lambda\gamma_{-}C_{-}}\leq e^{-\frac{\delta q_{k}}{2N_{0}}}\leq e^{-C_{0}\delta q_{k}}.

Thus we have proved the result with Lqk(E)L_{q_{k}}(E) instead of L(E)L(E). Now we need the last component of the proof:

Claim 3.

For any ε>0\varepsilon>0, L(E)Lqk(E)+εL(E)\leq L_{q_{k}}(E)+\varepsilon uniform in EE when qkq_{k} is large enough.

Proof.

In fact, since the operator is bounded, ln|Pqk0(x;E)|N0C1\ln|P_{q_{k}}^{0}(x;E)|\leq N_{0}C_{1}. Together with (4.2), we get

1qkln|Pqk(x;E)|Lqk(E)+Clnqkqk+N0C1qkLqk(E)+ε,x[0,1)\frac{1}{q_{k}}\ln|P_{q_{k}}(x;E)|\leq L_{q_{k}}(E)+C\frac{\ln q_{k}}{q_{k}}+\frac{N_{0}C_{1}}{q_{k}}\leq L_{q_{k}}(E)+\varepsilon,\forall x\in[0,1) (4.4)

uniformly in EE when qkq_{k} is large enough (depending on λ,γ±,C±\lambda,\gamma_{\pm},C_{\pm}). The same holds for P~qk(x;E)\tilde{P}_{q_{k}}(x;E).

On the other hand, by Lemma 4.1, for any ε>0\varepsilon>0, 1nln|Pn(x;E)|L(E)+ε\frac{1}{n}\ln|P_{n}(x;E)|\leq L(E)+\varepsilon eventually. While by definition of Lyapunov exponent (2.4), L(E)L(E) is the limiting averaging of 1nlnMn(x;E)\frac{1}{n}\ln\|M_{n}(x;E)\|. But MnM_{n} and PnP_{n} are connected by (2.3). Thus we see that on a set of measure at least 1/41/4, the following is true for either n=qkn=q_{k}, qk1q_{k}-1 or qk2q_{k}-2:

1nln|Pn(x;E)|L(E)ε.\frac{1}{n}\ln|P_{n}(x;E)|\geq L(E)-\varepsilon. (4.5)

If n=qkn=q_{k}, combining (4.5) and (4.4) gives us what we want. Otherwise, we first notice by row expansion of determinant, we have

Pn(x;E)+Pn2(x;E)=(λf(Tn1x)E)Pn1(x;E)\displaystyle P_{n}(x;E)+P_{n-2}(x;E)=(\lambda f(T^{n-1}x)-E)P_{n-1}(x;E) (4.6)
P~n(x;E)+2(1)n=Pn(x;E)Pn2(Tx;E).\displaystyle\tilde{P}_{n}(x;E)+2(-1)^{n}=P_{n}(x;E)-P_{n-2}(Tx;E). (4.7)

Then when n=qk1n=q_{k}-1, by (4.6), we have either PqkP_{q_{k}} or Pqk2P_{q_{k}-2} satisfies (4.5) so we can combine it with (4.4) to derive the result. If n=qk2n=q_{k}-2, by (4.7), we have either PqkP_{q_{k}} or P~qk\tilde{P}_{q_{k}} satisfies (4.5). For the former case, we get the result. For the latter, combining (4.5) and (4.4) with P~qk\tilde{P}_{q_{k}} instead of Pqk(x;E)P_{q_{k}}(x;E). The claim follows. ∎

Now the result follows immediately: For any δ>0\delta>0, apply Claim 3 to get L(E)Lqk(E)+δ/2L(E)\leq L_{q_{k}}(E)+\delta/2 eventually so that

{x[0,1):1nln|Pn(x;E)|L(E)δ}{x[0,1):1nln|Pn(x;E)|Lqk(E)δ/2}.\{x\in[0,1):\frac{1}{n}\ln|P_{n}(x;E)|\leq L(E)-\delta\}\subset\{x\in[0,1):\frac{1}{n}\ln|P_{n}(x;E)|\leq L_{q_{k}}(E)-\delta/2\}.

Then the result follows from Claim 2.

5. Exponential decay of eigenfunctions

We prove our key lemma 5.3, which provides uniform exponential decay of generalized eigenfunction in x,E,αx,E,\alpha. To do so, we introduce some definitions and prove a typical “either or” argument in the proof of localization in Lemma 5.2.

Definition 2 (Regular point).

We say a point nn\in{\mathbb{Z}} is (x,c,qk)(x,c,q_{k})-regular if there is an interval [a,b][a,b] with

n[a,b],b=a+qk1,|an|qk5,|nb|qk5,n\in[a,b],\ b=a+q_{k}-1,\ |a-n|\geq\frac{q_{k}}{5},\ |n-b|\geq\frac{q_{k}}{5}, (5.1)

such that

|Gx,E,[a,b](a,n)|ec|na|, and |Gx,E,[a,b](n,b)|ec|nb|.|{G_{x,E,[a,b]}}(a,n)|\leq e^{-c|n-a|},\text{~{}and~{}}|{G_{x,E,[a,b]}}(n,b)|\leq e^{-c|n-b|}.

Otherwise we say nn is (x,c,qk)(x,c,q_{k})-singular.

Lemma 5.1.

Fix δ,E\delta,E such that 0<δ<L(E)0<\delta<L(E). For qkq_{k} large enough, for any xx, if nn is (x,L(E)δ,qk)(x,L(E)-\delta,q_{k})-singular, then for any a[n3qk4,nqk4]a\in[n-\lfloor\frac{3q_{k}}{4}\rfloor,n-\lfloor\frac{q_{k}}{4}\rfloor],

|Pqk(Tax)|eqk(L(E)δ/10).|P_{q_{k}}(T^{a}x)|\leq e^{q_{k}(L(E)-\delta/10)}. (5.2)

Furthermore, let Nk=3qk4qk4+1N_{k}=\lfloor\frac{3q_{k}}{4}\rfloor-\lfloor\frac{q_{k}}{4}\rfloor+1 denote the number of such aa, then

qk+12Nkqk+32.\frac{q_{k}+1}{2}\leq N_{k}\leq\frac{q_{k}+3}{2}.
Proof.

Since nn is (x,L(E)δ,qk)(x,L(E)-\delta,q_{k})-singular, for any [a,b][a,b] satisfying (5.1), in particular, for any a[n3qk4,nqk4]a\in[n-\lfloor\frac{3q_{k}}{4}\rfloor,n-\lfloor\frac{q_{k}}{4}\rfloor], b=a+qk1b=a+q_{k}-1, we have

{either|Gx,E,[a,b](a,m)|e(L(E)δ)(ma), or |Gx,E,[a,b](m,b)|e(L(E)δ)(bm).\begin{cases}\text{either}&|{G_{x,E,[a,b]}}(a,m)|\geq e^{-(L(E)-\delta)(m-a)},\\ \text{~{}or~{}}&|{G_{x,E,[a,b]}}(m,b)|\geq e^{-(L(E)-\delta)(b-m)}.\end{cases} (5.3)

Notice that

{|Gx,E,[a,b](a,m)|=|Pbm(Tm+1x)||Pqk(Tax)|,|Gx,E,[a,b](m,b)|=|Pma(Tax)||Pqk(Tax)|.\begin{cases}|{G_{x,E,[a,b]}}(a,m)|=\frac{|P_{b-m}(T^{m+1}x)|}{|P_{q_{k}}(T^{a}x)|},\\ |{G_{x,E,[a,b]}}(m,b)|=\frac{|P_{m-a}(T^{a}x)|}{|P_{q_{k}}(T^{a}x)|}.\end{cases} (5.4)

Now we consider the first case in (5.3) for simplicity. The other case is similar. By Lemma 4.1, we have when qkq_{k} is large enough

|Pbm(Tm+1x)|e(L(E)+δ/10)(bm).|P_{b-m}(T^{m+1}x)|\leq e^{(L(E)+\delta/10)(b-m)}. (5.5)

By (5.3),(5.4) and (5.5), we see that

|Pqk(Tax)|e(L(E)+δ10)(bm)+(L(E)δ)(ma)eL(E)(ba)+δ10(bm)δ(ma)eL(E)qk+δ10qkδqk5e(L(E)δ10)qk,\begin{split}|P_{q_{k}}(T^{a}x)|&\leq e^{(L(E)+\frac{\delta}{10})(b-m)+(L(E)-\delta)(m-a)}\\ &\leq e^{L(E)(b-a)+\frac{\delta}{10}(b-m)-\delta(m-a)}\\ &\leq e^{L(E)q_{k}+\frac{\delta}{10}q_{k}-\delta\frac{q_{k}}{5}}\leq e^{(L(E)-\frac{\delta}{10})q_{k}},\end{split}

Thus we proved (5.2). The bound of NkN_{k} follows from direct computation when qk0,1,2,3(mod 4)q_{k}\equiv 0,1,2,3~{}(\text{mod~{}}4). ∎

In other words, there are many “large deviation points” near each singular point. This fact, together with the large deviation estimates in theorem 3 and appropriate weakly Liouville assumption (Definition 1), leads to the repelling of two singular points. In fact, we prove below that two (x,L(E)δ,qk)(x,L(E)-\delta,q_{k}) singular points are at least “qk+1qk/2q_{k+1}-q_{k}/2” away from each other:

Lemma 5.2 (Either or argument).

Let C0C_{0} be as in Theorem 3. Assume α\alpha and EE satisfy β(α)<C0L(E)\beta(\alpha)<C_{0}L(E). For any β(α)C0<δ<L(E)\frac{\beta(\alpha)}{C_{0}}<\delta<L(E), we have that for qkq_{k} large enough, and for any qk+12<|nm|qk+11qk+12\frac{q_{k}+1}{2}<|n-m|\leq q_{k+1}-1-\frac{q_{k}+1}{2}, either mm or nn is (x,L(E)δ,qk)(x,L(E)-\delta,q_{k})-regular for any xx.

Proof.

WLOG assume n>mn>m. For any δ<L(E)\delta<L(E), assume both mm and nn are (x,L(E)δ,qk)(x,L(E)-\delta,q_{k})-singular. By Lemma 5.1, we have

|Pqk(Tax)|e(L(E)δ/10)qk|P_{q_{k}}(T^{a}x)|\leq e^{(L(E)-\delta/10)q_{k}}

for any a[m3qk4,mqk4][n3qk4,nqk4]a\in[m-\lfloor\frac{3q_{k}}{4}\rfloor,m-\lfloor\frac{q_{k}}{4}\rfloor]\cup[n-\lfloor\frac{3q_{k}}{4}\rfloor,n-\lfloor\frac{q_{k}}{4}\rfloor]. Notice further that

n3qk4(mqk4)=nmNk+1>qk+12qk+32+1=0.n-\lfloor\frac{3q_{k}}{4}\rfloor-(m-\lfloor\frac{q_{k}}{4}\rfloor)=n-m-N_{k}+1>\frac{q_{k}+1}{2}-\frac{q_{k}+3}{2}+1=0.

Thus the two intervals of aa have no intersection. Overall there are 2Nkqk+12N_{k}\geq q_{k}+1 many possible aa such that |Pqk(Tax)|e(L(E)δ/10)qk|P_{q_{k}}(T^{a}x)|\leq e^{(L(E)-\delta/10)q_{k}}. By Theorem 3 and pigeonhole principle, there are i,j[m3qk4,mqk4][n3qk4,nqk4]i,j\in[m-\lfloor\frac{3q_{k}}{4}\rfloor,m-\lfloor\frac{q_{k}}{4}\rfloor]\cup[n-\lfloor\frac{3q_{k}}{4}\rfloor,n-\lfloor\frac{q_{k}}{4}\rfloor] such that

ν([Tix,Tjx])eC0δqk.\nu([T^{i}x,T^{j}x])\leq e^{-C_{0}\delta q_{k}}. (5.6)

Notice that |ij|nqk4(m3qk4)=nm+Nk1qk+11.|i-j|\leq n-\lfloor\frac{q_{k}}{4}\rfloor-(m-\lfloor\frac{3q_{k}}{4}\rfloor)=n-m+N_{k}-1\leq q_{k+1}-1. By Lemma A.1 and (A.2), we have

eC0δqkν([Tix,Tjx])ν([x,Tqkx])1qk+1.e^{-C_{0}\delta q_{k}}\geq\nu([T^{i}x,T^{j}x])\geq\nu([x,T^{q_{k}}x])\geq\frac{1}{q_{k+1}}.

This implies that

C0δ<lnqk+1qkC0δC0δlim suplnqk+1qk=β(α).\displaystyle C_{0}\delta<\frac{\ln q_{k+1}}{q_{k}}\Rightarrow C_{0}\delta\Rightarrow C_{0}\delta\leq\limsup\frac{\ln q_{k+1}}{q_{k}}=\beta(\alpha).

which leads to a contradiction with the assumption. ∎

Lemma 5.3.

Let C0C_{0} be as in Theorem 3. If (x,E,α)(x,E,\alpha) satisfy

  1. (\mathcal{E}1)

    EE is a generalized eigenvalue of Hλ,f,T,x{H_{\lambda,f,T,x}},

  2. (\mathcal{E}2)

    β(α)<C0L(E)\beta(\alpha)<C_{0}L(E),

then EE is an eigenvalue with exponentially decaying eigenfunction. Denote the normalized eigenfunction by ψ\psi with ψ=1\|\psi\|_{\infty}=1.

Furthermore, for any ε>0\varepsilon>0, there is a C=C(ε)C=C(\varepsilon), uniform in all x,E,αx,E,\alpha which satisfy (\mathcal{E}1), (\mathcal{E}2) above, such that for any pair of eigenvalue EE and normalized eigenvector ψ\psi, there is n0=n0(E)n_{0}=n_{0}(E) such that

|ψ(n)|C(ε)e110(L(E)β(α)C0ε)|nn0|.|\psi(n)|\leq C(\varepsilon)e^{-\frac{1}{10}\left(L(E)-\frac{\beta(\alpha)}{C_{0}}-\varepsilon\right)|n-n_{0}|}. (5.7)
Proof.

Take any β(α)C0<δ<L(E)\frac{\beta(\alpha)}{C_{0}}<\delta<L(E). Let ψ\psi be a generalized eigenfunction of Hλ,f,T,x{H_{\lambda,f,T,x}} with respect to EE. Thus ψ(n)C1(1+|n|)p\psi(n)\leq C_{1}(1+|n|)^{p} where C1=C1(E,x,α)C_{1}=C_{1}(E,x,\alpha). We first prove ψ\psi decay exponentially so that EE is an eigenvalue, then we prove the decay is uniform in the sense of (5.7).

WLOG assume ψ(0)0\psi(0)\neq 0. By (2.2), 0 is eventually (x,L(E)δ,qk)(x,L(E)-\delta,q_{k})-singular. By Lemma 5.2, we have for qkq_{k} large enough, any n(qk+12,qk+11qk+12]:=(Ak,Bk]n\in(\frac{q_{k}+1}{2},q_{k+1}-1-\frac{q_{k}+1}{2}]:=(A_{k},B_{k}] is (L(E)δ,qk)(L(E)-\delta,q_{k})-regular. Notice further that Ak+1BkA_{k+1}\leq B_{k} since qk+1qk+4q_{k+1}\geq q_{k}+4 for k4k\geq 4. Thus eventually for any nn, there is kk such that n(Ak,Ak+1]n\in(A_{k},A_{k+1}]. We derive exponential decay by considering two cases seperately:

  1. (1)

    If n(Ak,qk]n\in(A_{k},q_{k}], nn is (L(E)δ,qk)(L(E)-\delta,q_{k})-regular, by (2.2), we have for arbitrarily small ε>0\varepsilon>0, eventually

    |ψ(n)|C1e(L(E)δ)qk/5(1+3n)pe(L(E)δε)n/5.|\psi(n)|\leq C_{1}e^{-(L(E)-\delta)q_{k}/5}(1+3n)^{p}\leq e^{-(L(E)-\delta-\varepsilon)n/5}. (5.8)
  2. (2)

    If n[qk+1,Ak+1]n\in[q_{k}+1,A_{k+1}], then it is easy to check that |nBk||nAk|n/2|n-B_{k}|\geq|n-A_{k}|\geq n/2. By (2.2), we have

    |ψ(n)|2e(L(E)δ)qk/5|ψ(n1)||\psi(n)|\leq 2e^{-(L(E)-\delta)q_{k}/5}|\psi(n_{1})|

    where n1=a1n_{1}=a-1 or b+1b+1 for suitable [a,b][a,b] satifying (5.1). As long as n1(Ak,Bk]n_{1}\in(A_{k},B_{k}], where n1n_{1} would be (x,L(E)δ,qk)(x,L(E)-\delta,q_{k}) regular, then we can apply (2.2) again to ψ(n1)\psi(n_{1}). We can repeat this process to get ψ(n2)\psi(n_{2}), ψ(n3),\psi(n_{3}),\dots, as long as nin_{i} stays in (Ak,Bk](A_{k},B_{k}]. Since |nBk||nAk|n/2|n-B_{k}|\geq|n-A_{k}|\geq n/2 while |nini+1|qk|n_{i}-n_{i+1}|\leq q_{k}, thus we can at least do

    J|nAk|qkn2qkJ\geq\frac{|n-A_{k}|}{q_{k}}\geq\frac{n}{2q_{k}}

    many times. Then we get

    |ψ(n)|2Je(L(E)δ)qkJ/5|ψ(nJ)|e(L(E)δ5qk)n10|ψ(nJ)|C1e(L(E)δ5qk)n/10(1+3n)pC1e(L(E)δε)n/10.\begin{split}|\psi(n)|&\leq 2^{J}e^{-(L(E)-\delta)q_{k}J/5}|\psi(n_{J})|\leq e^{-\left(L(E)-\delta-\frac{5}{q_{k}}\right)\frac{n}{10}}|\psi(n_{J})|\\ &\leq C_{1}e^{-(L(E)-\delta-\frac{5}{q_{k}})n/10}(1+3n)^{p}\leq C_{1}e^{-(L(E)-\delta-\varepsilon)n/10}.\end{split} (5.9)

Combining (5.8) and (5.9) gives us the first half of the theorem. Now since ψ2\psi\in\ell^{2}, we can normalize it so that ψ=1\|\psi\|_{\infty}=1.

The key point of the second half is the uniformity in x,E,αx,E,\alpha. Take n0=min{n:ψ(n)=1}>n_{0}=\min\{n:\psi(n)=1\}>-\infty to be the leftmost maximum point of ψ\psi. By (2.2), we see that maximum point n0n_{0} is always (x,L(E)δ,qk)(x,L(E)-\delta,q_{k})-singular for all qkq_{k} . Thus nn is (x,L(E)δ,qk)(x,L(E)-\delta,q_{k})-regular if Ak<|nn0|BkA_{k}<|n-n_{0}|\leq B_{k}. We can now repeat the estimates (5.8) and (5.9) above with the new, uniform (in x,E,αx,E,\alpha) improvement that |ψ(ni)|1|\psi(n_{i})|\leq 1 instead of |ψ(ni)|C1(x,E,α)(1+ni)p|\psi(n_{i})|\leq C_{1}(x,E,\alpha)(1+n_{i})^{p}, where we get

{|ψ(n)|e(L(E)δ)|nn0|5,n(Ak,qk],|ψ(n)|e(L(E)δ5qk)|nn0|10,n(qk,Ak+1].\begin{cases}|\psi(n)|\leq e^{-(L(E)-\delta)\frac{|n-n_{0}|}{5}},&n\in(A_{k},q_{k}],\\ |\psi(n)|\leq e^{-(L(E)-\delta-\frac{5}{q_{k}})\frac{|n-n_{0}|}{10}},&n\in(q_{k},A_{k+1}].\end{cases}

Since β(αC0<δ<L(E)\frac{\beta(\alpha}{C_{0}}<\delta<L(E) is arbitrary and 5qk\frac{5}{q_{k}} is arbitrarily small once qkq_{k} is large enough uniformly in x,E,αx,E,\alpha. Thus (5.7) follows. ∎

6. Localization results

Now we prove our main results. Both of them follow directly from Lemma 5.3:

Proof of Theorem 1.

Recall that by the Schnol’s theorem, spectral measure is supported on the set of generalized eigenvalues (see [2, Ch. VII]. Fix λ\lambda and xx, the theorem follows directly from Lemma 5.3. ∎

Definition 3 (Uniform localization).

An operator HH exhibits uniform localization if there exixts C,cC,c such that for any pair of eigenvalue and eigenfunction EE, ψ\psi, there exists n0=n0(E)n_{0}=n_{0}(E) such that

|ψ(n)|Cec|nn0|.\displaystyle|\psi(n)|\leq Ce^{-c|n-n_{0}|}.
Proof of Theorem 2.

By Corollary 3.3, 0<ln(λγC4e)L(E)0<\ln\left(\frac{\lambda\gamma_{-}C_{-}}{4e}\right)\leq L(E) for all EE. It follows that β(α)<C0L(E)\beta(\alpha)<C_{0}L(E). Thus Lemma 5.3 applies to all xx, all EE and those α\alpha which satisfy our assumption. By taking ε=12ln(λC^γ4ηe)\varepsilon=\frac{1}{2}\ln\left(\frac{\lambda\hat{C}\gamma_{-}}{4\eta e}\right) in Lemma 5.3, we get uniform localization. ∎

Appendix A Orbital analysis

It is well-known that irrational rotation on 1d-torus, Rα(x)=x+αR_{\alpha}(x)=x+\alpha, has best-approximation property, c.f. [20],

qkαnα,1n<qk+1α\|q_{k}\alpha\|\leq\|n\alpha\|,\quad\forall 1\leq n<q_{k+1}\alpha (A.1)

with estimates

12qk+1qkα1qk+1,\frac{1}{2q_{k+1}}\leq\|q_{k}\alpha\|\leq\frac{1}{q_{k+1}}, (A.2)

where qkq_{k} is defined in (2.1). Furthermore, the orbits of RαR_{\alpha} is also well-understood, we cite [7, Proposition 4.1, 4.2] here:

Proposition A.1.

Let k1k\geq 1. The points {jα}\{j\alpha\}, j=0,1,2,,qk1j=0,1,2,\dots,q_{k}-1 splits [0,1)[0,1) into qk1q_{k-1} “large” gaps of length (qkqk1)α\|(q_{k}-q_{k-1})\alpha\| and qkqk1q_{k}-q_{k-1} “small” gaps of length qk1α\|q_{k-1}\alpha\|. Furthermore, we have the estimates

1qkqk1qkqk+1qk1α1qk,1qk||(qkqk1)α)||1qk+1qk+1.\begin{split}&\tfrac{1}{q_{k}}-\tfrac{q_{k-1}}{q_{k}q_{k+1}}\leq||q_{k-1}\alpha||\leq\tfrac{1}{q_{k}},\\ &\tfrac{1}{q_{k}}\leq||(q_{k}-q_{k-1})\alpha)||\leq\tfrac{1}{q_{k}}+\tfrac{1}{q_{k+1}}.\end{split}

For a general measure-preserving circle homeomorphism TT with rotation number α\alpha, such kind of best approximate properties and orbital analysis can be done in a similar way with invariant measure ν\nu instead of distance ||||||\cdot||. In fact,

Lemma A.1 (Best approximation).

For any x𝕋1x\in{\mathbb{T}}^{1} and kk\in{\mathbb{N}},

ν([x,Tix])ν([x,Tqkx]),\displaystyle\nu([x,T^{i}x])\geq\nu([x,T^{q_{k}}x]),

where 0i<qk+10\leq i<q_{k+1}.

Proof.

Note that Lemma A.1 holds when the invariant measure is the Lebesgue measure - in other words, when the map TT is the irrational rotation. For a general measure-preserving circle homeomorphism, this inequality holds since it is equivalent to the irrational rotation case. In fact, the Poincaré classification theorem [6, Theorem 4.3.20] guarantees the existence of the topological conjucacy hh with a rotation RαR_{\alpha}, and hh is also the distribution function for the unique invariant measure ν\nu. Hence, for any x𝕋1x\in{\mathbb{T}}^{1} and ii\in{\mathbb{N}}, we have

ν([x,Tix])=|h(Tix)h(x)|=|Rαi(h(x))h(x)|=iα.\displaystyle\nu([x,T^{i}x])=|h(T^{i}x)-h(x)|=|R^{i}_{\alpha}(h(x))-h(x)|=\|i\alpha\|.

Lemma A.2.

Fix xx, Let k1k\geq 1. The points {Tjx}j=0qk1\{T^{j}x\}_{j=0}^{q_{k}-1} split [0,1)[0,1) into qk1q_{k-1} “large” gaps of invariant measure ν([Tqkx,Tqk1x])=ν([x,Tqkqk1x])\nu\left([T^{q_{k}}x,T^{q_{k-1}}x]\right)=\nu\left([x,T^{q_{k}-q_{k-1}}x]\right) and qkqk1q_{k}-q_{k-1} “small” gaps of invariant measure ν([x,Tqk1x])\nu\left([x,T^{q_{k-1}}x]\right). Furthermore, we have the estimates

1qkqk1qkqk+1ν([x,Tqk1x])1qk,1qkν([x,Tqkqk1x])1qk+1qk+1.\begin{split}&\tfrac{1}{q_{k}}-\tfrac{q_{k-1}}{q_{k}q_{k+1}}\leq\nu\left([x,T^{q_{k-1}}x]\right)\leq\tfrac{1}{q_{k}},\\ &\tfrac{1}{q_{k}}\leq\nu\left([x,T^{q_{k}-q_{k-1}}x]\right)\leq\tfrac{1}{q_{k}}+\tfrac{1}{q_{k+1}}.\end{split}
Proof.

To prove the theorem, let us first introduce the dynamical partition on the circle by following the convention using in [8]. For each kk\in{\mathbb{N}}, let IkI_{k} be the interval between xx and TqkxT^{q_{k}}x. It can be verified by induction in kk that the following collection of intervals forms a kthk^{th} dynamical partition of 𝕋1{\mathbb{T}}^{1}

𝒫k(z):={Ik,T(Ik),,Tqk11(Ik)}{Ik1,T(Ik1),,Tqk1(Ik1)}:=𝒮kk.\mathcal{P}_{k}(z):=\{I_{k},T(I_{k}),\dots,T^{q_{k-1}-1}(I_{k})\}\cup\{I_{k-1},T(I_{k-1}),\dots,T^{q_{k}-1}(I_{k-1})\}:=\mathcal{S}_{k}\cup\mathcal{L}_{k}.

That is, they are disjoint except for the endpoints, and the union cover the whole circle. Notice that intervals in 𝒮k\mathcal{S}_{k} all have smaller invariant measure ν(Ik)<ν(Ik1)\nu(I_{k})<\nu(I_{k-1}) than intervals in k\mathcal{L}_{k}, thus we call them “short” and “long” intervals correspondingly. One can check by induction on kk that, each “long” interval Tj(Ik1)T^{j}(I_{k-1}) in kthk^{th} dynamical partition is divided into ak+1a_{k+1} “long” intervals and one “short” interval in k+1thk+1^{th} dynamical partition. More specifically,

Tj(Ik1)k{Tj+qk1(Ik),Tj+qk1+qk(Ik),,Tj+qk1+(ak+11)qk(Ik)k+1,Tj(Ik+1)𝒮k+1.T^{j}(I_{k-1})\in\mathcal{L}_{k}\Rightarrow\begin{cases}T^{j+{q_{k-1}}}(I_{k}),T^{j+{q_{k-1}}+{q_{k}}}(I_{k}),\dots,T^{j+{q_{k-1}}+(a_{k+1}-1){q_{k}}}(I_{k})\in\mathcal{L}_{k+1},\\ T^{j}(I_{k+1})\in\mathcal{S}_{k+1}.\end{cases}

This allows us to estimates the “large” and “small” gaps 222Notice that the partition in Lemma A.2 is different from dynamical partition, “long” and “short” intervals are also different concepts from “large” and “small” gaps. in Lemma A.2 now.

Proof of Theorem A.2.

Since μ\mu is the invariant measure of TT, for dynamical partition 𝒫k+1(z)\mathcal{P}_{k+1}(z), we have

1=i=0qk+11ν(Ti(Ik))+j=0qk1ν(Tj(Ik+1))=qk+1ν(Ik)+qkν(Ik+1).1=\sum_{i=0}^{q_{k+1}-1}\nu(T^{i}(I_{k}))+\sum_{j=0}^{q_{k}-1}\nu(T^{j}(I_{k+1}))=q_{k+1}\nu(I_{k})+q_{k}\nu(I_{k+1}). (A.3)

By (A.3), we get

ν(Ik)=1qkν(Ik+1)qk+11qk+1.\displaystyle\nu(I_{k})=\frac{1-q_{k}\nu(I_{k+1})}{q_{k+1}}\leq\frac{1}{q_{k+1}}.

Moreover, since (A.3) holds for any kk, we also get ν(Ik+1)1qk+2\nu(I_{k+1})\leq\frac{1}{q_{k+2}}. So,

ν(Ik)1qk+1qkqk+1qk+212qk+1.\nu(I_{k})\geq\frac{1}{q_{k+1}}-\frac{q_{k}}{q_{k+1}q_{k+2}}\geq\frac{1}{2q_{k+1}}. (A.4)

The last inequality follows from the recurrence relation (2.1) and ak1a_{k}\geq 1:

qk+2=ak+2qk+1+qk2qk.q_{k+2}=a_{k+2}q_{k+1}+q_{k}\geq 2q_{k}. (A.5)

By the comparability between ν\nu and the Lebesgue measure on a circle (𝒯\mathcal{T}1), the claim follows. ∎

Acknowledgement

We would like to thank Svetlana Jitomirskaya for suggesting this problem, and helpful discussion and comments; Ilya Kachkovskiy for the helpful discussions on potential sharp conditions and improvements; Saša Kocić for his comments on the conjugacies. J.K. would also like to thank UCI for their wonderful hospitality. X.Z. was partially supported by Simons 681675, NSF DMS-2052899, DMS-2155211, DMS-2054589. J.K. was partially supported by the National Science Foundation EPSCoR RII Track-4 Award No. 1738834 and she appreciates Saša Kocić’s generous support on the visit to UCI.

References

  • [1] Artur Avila, Jiangong You, and Qi Zhou. Sharp phase transitions for the almost mathieu operator. Duke Mathematical Journal, 166(14):2697–2718, 2017.
  • [2] I.M. Berezanskii. Expansions in eigenfunctions of selfadjoint operators, volume 17. American Mathematical Soc., 1968.
  • [3] Hans L Cycon, Richard G Froese, Werner Kirsch, and Barry Simon. Schrödinger operators: With application to quantum mechanics and global geometry. Springer, 2009.
  • [4] Lingrui Ge, Jiangong You, and Xin Zhao. The arithmetic version of the frequency transition conjecture: New proof and generalization. Peking Mathematical Journal, pages 1–16, 2021.
  • [5] Rui Han, Svetlana Jitomirskaya, and Fan Yang. Anti-resonances and sharp analysis of maryland localization for all parameters. arXiv preprint arXiv:2205.04021, 2022.
  • [6] Boris Hasselblatt and Anatole Katok. A first course in dynamics: with a panorama of recent developments. Cambridge University Press, 2003.
  • [7] Svetlana Jitomirskaya and Ilya Kachkovskiy. All couplings localization for quasiperiodic operators with monotone potentials. Journal of the European Mathematical Society, 21(3):777–795, 2018.
  • [8] Svetlana Jitomirskaya and Saša Kocić. Spectral theory of Schrödinger operators over circle diffeomorphisms. International Mathematics Research Notices, 2022(13):9810––9829, 2021.
  • [9] Svetlana Jitomirskaya, Lyuben Konstantinov, and Igor Krasovsky. On the spectrum of critical almost mathieu operators in the rational case. arXiv preprint arXiv:2007.01005, 2020.
  • [10] Svetlana Jitomirskaya and Wencai Liu. Arithmetic spectral transitions for the maryland model. Communications on Pure and Applied Mathematics, 70(6):1025–1051, 2017.
  • [11] Svetlana Jitomirskaya and Wencai Liu. Universal hierarchical structure of quasiperiodic eigenfunctions. Annals of Mathematics, 187(3):721–776, 2018.
  • [12] Svetlana Jitomirskaya, Wencai Liu, and Shiwen Zhang. Arithmetic spectral transitions: a competition between hyperbolicity and the arithmetics of small denominators. Harmonic Analysis and Applications, 27:35, 2020.
  • [13] Svetlana Jitomirskaya and Rajinder Mavi. Dynamical bounds for quasiperiodic Schrödinger operators with rough potentials. International Mathematics Research Notices, 2017(1):96–120, 2016.
  • [14] Svetlana Ya Jitomirskaya. Metal-insulator transition for the almost mathieu operator. Annals of Mathematics, pages 1159–1175, 1999.
  • [15] Ilya Kachkovskiy. Localization for quasiperiodic operators with unbounded monotone potentials. Journal of Functional Analysis, 277(10):3467–3490, 2019.
  • [16] Saša Kocić. Singular continuous phase for Schrödinger operators over circle diffeomorphisms with a singularity. preprint on webpage at https://web.ma.utexas.edu/mp_arc/index-20.html, October 2020.
  • [17] Wencai Liu. Almost mathieu operators with completely resonant phases. Ergodic Theory and Dynamical Systems, 40(7):1875–1893, 2020.
  • [18] Wencai Liu. Distributions of resonances of supercritical quasi-periodic operators. arXiv preprint arXiv:2208.06944, 2022.
  • [19] Wencai Liu and Xiaoping Yuan. Anderson localization for the completely resonant phases. Journal of Functional Analysis, 268(3):732–747, 2015.
  • [20] Alfred Jacobus Van der Poorten. Notes on continued fractions and recurrence sequences. In Number Theory and Cryptography, volume 154, pages 86––97. Cambridge University Press, 1990.