Anosov-Katok constructions for quasi-periodic cocycles
Abstract.
We prove that if the frequency of the quasi-periodic cocycle is Diophantine, then the following properties are dense in the subcritical regime: for any , the Lyapunov exponent is exactly -Hölder continuous; the extended eigenstates of the potential have optimal sub-linear growth; and the dual operator associated a subcritical potential has power-law decay eigenfunctions. The proof is based on fibered Anosov-Katok constructions for quasi-periodic cocycles.
1. Introduction
In this paper, we are concerned with one-dimensional analytic quasi-periodic Schrödinger operators:
(1.1) |
where is the potential, is the frequency, assumed to be rationally independent, and is the phase. Due to the rich implications in quantum physics, quasi-periodic Schrödinger operators have been extensively studied [46]. Starting from the early 1980’s, there was already an almost periodic flu which swept the world, as pointed out by Simon [49]. In 2000’s, adopting a dynamical systems point of view (mainly analytic quasi-periodic cocycles) was found to be usefull in the study of such operators (1.1), and much progress has been made since then [2, 3, 12, 6].
We recall that an analytic quasi-periodic cocycle is a linear skew product system:
Among many important advances, we should highlight Avila’s global theory of one-frequency analytic quasi-periodic cocycles [2]: if , any which is not uniformly hyperbolic, is either supercritical, subcritical or critical, properties that can be expressed in terms of the growth of the cocycle. More precisely, is said to be
-
(1)
Supercritical, if grows exponentially;
-
(2)
Subcritical, if there is a uniform subexponential bound on the growth of through some band ;
-
(3)
Critical, otherwise.
Here we define the products of cocycles as , where
and for . We also recall that is uniformly hyperbolic if there exists a continuous invariant splitting with the following property: there exist and such that holds for every and every . Simultaneously holds for every . We will denote the set of such cocycles as for short111In this paper, we will fix the frequency , only vary the mapping in the fiber..
The fundamental observation, allowing for the dynamical systems point of view to be taken, is that (1.1) can be seen as a quasi-periodic cocycle, since a sequence is a formal solution of the eigenvalue equation if and only if it satisfies
(1.2) |
where
Thus, corresponding to (1.1), there is a family of naturally defined cocycles , called quasi-periodic Schrödinger cocycles. It is well-known that , the spectrum of (1.1), if and only if . Therefore energies in the spectrum of can be characterized as supercritical, subcritical or critical in terms of its corresponding cocycle .
A cornerstone in Avila’s global theory is the “Almost Reducibility Theorem ”(ART) [2, 4, 5], which asserts that is almost reducible (denoted by for any fixed ) if it is subcritical. We recall that a cocycle is almost reducible, if the closure of its analytic conjugacy class contains a constant cocycle.
In this paper, we will focus on the dynamical and spectral behavior of in the subcritical regime. We remark that our main results work also in the multifrequency case, , since they are proved in the regime . Indeed, by Avila’s global theory and ART [2, 4, 5], if , if and only if is subcritical.
1.1. Regularity of Lyapunov exponent
Our first result concerns the regularity of Lyapunov exponent (LE):
as a function of the mapping defining the dynamics in the fibers. The LE is a central topic in the spectral theory of Schrödinger operators, since it relates with intergrated density of states through the Thouless formula. The LE also arise naturally in the study of smooth dynamics, and continuity of LE has been the object of important recent research, see [50] and the references therein.
In the analytic topology, the LE is always continuous with respect to both and at any given cocycle , provided that is rationally independent, see [18, 21, 35]. This holds true even for cocycles in higher dimensional groups , with [9]. If 222We recall that is Diophantine, denoted by , if there exist and such that (1.4) The set of Diophantine numbers is denoted by ., and the potential is small enough (the smallness of depending on ), then is Hölder continuous w.r.t [15]. Subsequently, Avila-Jitomirskaya in [7] generalized the -Hölder continuity to the non-perturbative regime, lifting the dependence of the smallness of on . By the ART, see [2, 4], the LE is -Hölder continuous in the whole subcritical regime [47]. This kind of -Hölder continuity is sharp, since the LE is exactly Hölder continuous at the end of spectral gaps by [48]. On the other hand, if is conjugated to constant cocycle or , then LE is Lipschitz. Therefore a natural question is whether there exist subcritical cocycles such that the optimal Hölder exponent of can be any fixed number between and . In fact, we will show that this holds in a dense set in for each admissible Hölder exponent.
Theorem 1.1.
Let , . There exists a set which is dense in in the topology such that Lyapunov exponent is exactly -Hölder continuous at each point of in the sense that for any we have for ,
Some arithmetic condition on is needed for the theorem to be true. As pointed out by Avila-Jitomirskaya [7], the Lyapunov Exponent is discontinuous at rational , which implies then for generic , LE cannot be Hölder continuous at any order. Recently, Avila-Last-Shamis-Zhou [11] showed that if is very Liouvillean, then LE can even be not -Hölder continuous.
In the positive Lyapunov Exponent regime, Goldstein and Schlag proved that the LE is Hölder continuous (for one-frequency cocycles) or weak Hölder continuous (for multi-frequency cocycles) for Schrödinger operators whose frequency satisfies a strong Diophantine condition [30, 31].
Continuity of the LE also depends sensitively on the smoothness of . In the topology, any non uniformly hyperbolic cocycle can be approximated by cocycles with zero LE, see [16, 17], and thus the LE is not continuous. In the () topology, results are quite different. Wang-You, [51], showed that the LE can be discontinuous. On the other hand, Xu-Ge-Wang, [53], recently constructed a class of -type potential, and for any showed the existence of an energy, such that LE is exactly -Hölder continuous. For other results of regularity of LE in the smooth or in the Gervey topology, one can consult [22, 52, 45] and the references therein.
1.2. Optimal sub-linear growth of extended eigenstates
Next, we move to spectral applications. In contrast with 1D random Schrödinger operator, one of the most remarkable phenomena exhibited by quasiperiodic Schrödinger operators is the existence of absolutely continuous (ac) spectrum. If and is small enough, then Dinaburg-Sinai in [23] proved that has ac spectrum. Eliasson in [24] proved the stronger result that has purely ac spectrum for every . Avila’s ART [2, 4, 5] ensures that if the potential is subcritical, then the corresponding operator has purely ac spectrum. From the physicist’s point of view, ac spectrum corresponds to a phase where the material is a conductor, and the corresponding eigenstate is extended. Indeed, one can define the inverse participation ratio (IPR) [26] as
where is the eigenstate. If , then the phase is extended. Thus, conductivity of the material is reflected in the growth of its eigenstate.
For and small enough, Eliasson [24] proved that if the energy lies in the end of spectral gaps, then its extended eigenstates have linear growth. On the other hand, for any , the spectrum of the operator, the extended eigenstates at most have sub linear growth, i.e. . This kind of behavior can be generalized to the whole subcritical regime [2, 4]. Thus it is interesting to ask whether this kind of sub-linear growth was optimal. In this paper, we prove that for a dense subset of subcritical potentials, the extended eigenstates of the associated operator have optimal sub-linear growth.
Theorem 1.2.
Fix . For any non-increasing sequence satisfying and , for any , if , then for any , there exist , with , such that . Moreover, the eigenstate of
has sub linear growth with rate , i.e. there exist and not depending on such that
Remark 1.1.
On the right side of the inequality, one can replace by for any .
Theorem 1.2 shows that there exists extended eigenfunctions with optimal sub-linear growth. We remark howevert, that the universal hierarchical structure of any extended quasi-periodic eigenfunctions for almost Mathieu operators was recently obtained in [33]
The study of growth of extended eigenstates is not only interesting from the physicist’s point of view, but also important from the mathematical point of view. By (1.2), the study of the growth of the extended eigenstates is equivalent to study the growth of the Schrödinger cocycle. On one hand, the growth of Schrödinger cocycle is crucial for proving existence of ac spectrum. In the subcritical regime, for almost every , the cocycles are uniformly bounded, see [24], which is sufficient for establishing the existence of ac spectrum by subordinacy theory [29]. However, whether the spectral measure is purely ac or not really depends on the growth of on a Lebesgue zero measure set of energies by [3]. On the other hand, the growth of Schrödinger cocycle is important in proving the regularity of the spectral measure, see [8]. Indeed, using Jitomiskaya-Last’s inequality, see [37], Avila-Jitomirskaya in [8] showed that
where satisfies . Therefore, by Theorem 1.2, it seems interesting to study high order Hölder continuity of spectral measure for a dense set of subcritical energies.
1.3. Power-law decay eigenfunctions
Our third result concerns the localization property of the quasi-periodic long-range operator:
(1.5) |
where is the Fourier coefficient of . This operator has received a lot of attention, see [7, 13, 20] since it is the Aubry dual of the quasi-periodic Schrödinger operator defined in eq. (1.1). If , then it reduces to the extensively studied almost Mathieu operator:
For the almost Mathieu operator, there is a sharp phase transition line 333 Here, where is the continued fraction best rationnal approximants of . from singular continuous spectrum to Anderson localization (pure point spectrum with exponentially decaying eigenfunctions), see [13, 38]. In the transition line , for frequencies in a dense set, displays pure point spectrum, see [10], but the eigenfunction does not decay exponentially, see [38]. As pointed out in [10], the insights gained from the critical parameters often shed light on the creation, dissipation, and the mechanism behind the phases of non critical parameters as well. Thus it is interesting to ask whether there exists real power law decay eigenfunction, i.e. if the eigenfunctions can decay polynomially. In this paper, we establish the following result.
Theorem 1.3.
Let , and . Then, there exist and , such that for any , any
is accumulated by such that has point spectrum with eigenfunction for some , where we denote .
Just as Theorem 1.2, Theorem 1.3 also holds if the dual Schrödigner operator (1.1) lies in the subcritical regime. However, under the assumption that , the phenomenon exhibited by Theorem 1.3 was not expected for Schrödigner operators (1.1). Indeed, if is an even function, then for a dense set of , has no eigenvalues, see [40]. It is widely believed for a.e. , has Anderson localization in the positive LE regime. In fact, even exhibits a sharp transition in the phase between singular continuous spectrum and Anderson localization [39].
If , Bourgain-Jitomirskaya in [20] proved that for any fixed , has Anderson localization for sufficiently small and a.e. . If , not necessary Diophantine, then for dense set of small potentials and a.e. , has point spectrum with exponentially decaying eigenfunctions, see [56], while it is still open whether has Anderson localization. If , Jitomirskaya-Kachkovskiy in [36] proved that for fixed , has pure point spectrum for sufficiently small and . Recently, Ge-You-Zhou [32] further proved that under the same assumption, has exponentially dynamical localization.
1.4. Density coycles reducible in finite differentiability
Recall that a cocycle is reducible if there exists such that
then by Aubry duality, Theorem 1.3 is an immediately corollary of the following reducibility result.
Theorem 1.4.
Given , . There exists a set which is dense in in the topology, such that if , then is -reducible but not -reducible.
We point out that this result is interesting in itself, from the dynamical systems point of view. The study of the reducibility of cocycles is related with the linearization of circle diffeomorphisms. Arnol’d in [1] proved that if an analytic diffeomorphism is close to a rotation , where is the rotation number of , and , then is analytically linearizable. One of the great achievements of Herman-Yoccoz, see [34, 54], is the proof of the fact that the sharp arithmetic condition for linearizability of , without any smallness condition imposed a priori on , is that its rotation number be Diophantine. In the Liouvillean regime, Herman [34] (see also [44]) proved that for any and any Liouvillean number , the set of with rotation number which is but not linearizable is locally dense. For a more precise description of the theory, we refer the reader for the recent survey of Eliasson-Fayad-Krikorian [25] on circle diffeomorphisms.
Concerning the reducibility of cocycles, if the base frequency , and the cocycle is sufficiently close to constants (the closeness depends on ), Eliasson [24] proved that if the fibered rotation number is Diophantine w.r.t , then is analytically reducible. By global to local reduction, reducibility for a full measure set of frequencies holds in the non-perturbative regime, see [48], and in the subcritical regime [4]. Motivated by the results in circle diffeomorphism, then it is natural to study the density of but not reducible cocycles. Theorem 1.4 provides a positive answer to this question. The corresponding theorem in the case of smooth cocycles, which in some sense stands in midway between circle diffemorphisms and cocycles, was proved by the first author in [41].
1.5. Some comments on the proofs
The proof of these results are based on the fast approximation by conjugation method introduced by Anosov and Katok in [14], where they constructed mixing diffeomorphisms of the unit disc arbitrarily close to Liouvillean rotations. We refer the reader to §3 for a more detailed description of the method, but in a nutshell it consists of the following idea.
The transient dynamics of a Liouvillean rotation is, for all practical reasons, periodic: if is Liouville, then there exist sequences such that
which means that a diffeomorphism conjugate to the rotation by is practically periodic with period , in scales of iteration comparable with arbitrarily big powers of . It is tempting, therefore, to try to study diffeomorphisms with rotation number equal to by approximating them from the outside, i.e. with diffeomorphisms that have rational rotation numbers, and to boot, are conjugate to them.
Since the dynamics of a periodic rotation are determined by a finite number of iterations, they are easier to tamper with, and since , it is also possible to modify the dynamics at the scale in a way that preserves what was constructed in the previous scale .
The limit object, satisfying the mixing property was thus constructed as follows. A diffeomorphism is constructed such that
for some smooth conjugation , satisfying some finitary version of mixing at a scale of iteration . A -periodic conjugation is constructed, such that the diffeomorphism
satisfying some improved finitary version of mixing at a scale of iteration . This can be achieved by chosing very close to , so that the norm of can be allowed to explode. This will ensure that the limit diffeomorphism is smooth, but not measurably conjugate to . The Liouvillean character of is necessary in order to assure the convergence of despite the divergence of .
For more information on the method, results and references we point the reader to [28]. This approximation-by-conjugation method has been useful in producing examples of dynamics incompatible with quasi-periodicity, in the vicinity of quasi-periodic dynamics. It is in some sense the counterpart of the KAM method: KAM tends to prove rigidity in the Diophantine world, while Anosov-Katok is used in order to prove non-rigidity in the Liouvillean world. In the context of cocycles, the concept of reducibility, obtained notably via KAM, allows for studying the rigidity results [12] when the fibered rotation number is Diophantine with respect to the frequency, while Anosov-Katok’s construction will be an efficient method to study wild dynamics when the fibered rotation number is Liouville with respect to the frequency.
The fibered Anosov-Katok construction was introduced by the first author in [41]. In this context, the rotation in the basis is fixed, and the only freedom is in the choice of the mapping in the fibers. The rotation in the basis can be chosen to be Diophantine, and the role of periodic rotations in the classical constructions is taken by resonant cocycles, i.e. cocycles whose rotation number is , where . The rest of the construction remains the same.
From an almost reducibility point of view, the construction consists in engineering the parameters of the KAM normal form, introduced in [42], and further exploited in [41, 43] in order to study a variety of dynamical phenomena present in the almost reducibility regime for cocycles in . Using the KAM schemes of [22, 47], we further develop these techniques in order to adapt them to cocycles and in the context of the analytic category (instead of the smooth one, as in the [42]).
We point out that results obtainable by the fibered Anosov-Katok method can be obtained for cocycles over a fixed Liouvillean rotation, but the case of a Diophantine rotation is more difficult, and therefore more interesting.
2. A lemma from linear algebra
Given and calling
we have by direct calculation , where is the group of special unitary matrices preserving the scalar product of with signature i.e.
where . Since is an isometry between the upper half-plane and the disc models for the hyperbolic plane, we know that
The Lie algebra of , denoted by is formed by traceless Hermitian matrices,
Given two matrices in
their scalar product is defined by
so that the natural semi-Riemannian structure on be defined by
In we can therefore distinguish three regimes: the elliptic regime where , the parabolic regime where , and the hyperbolic regime where .
Parabolic matrices are not diagonalizable, and hyperbolic ones are anti-diagonalizable, but since in the present paper we focus on the elliptic regime, we prove the following lemma concerning the diagonalization of elliptic matrices in . The diagonalizing conjugugation given by the lemma is of optimal norm.
Lemma 2.1.
Let the matrix
satisfy . Then, calling , we have
where
Here and satisfies
(2.1) |
In addition we have
(2.2) |
Proof.
Firstly, the invariance of the determinant forces that
where the choice of the sign of is of course arbitrary and irrelevant.
Let, now, , so that can be seen as an element of parameterized by . Then means belongs to the cone . As was shown in Figure 1, we can rotate to which lies in the plane by the conjugation:
which is a rotation around the axis. We can thus restrict the problem to the diagonalization of matrices of the type
In the plane, conjugacy classes are hyperbola where is the square of the angle of the corresponding rotation. In the full space, the conjugacy classes are the hyperboloids obtained by revolving these hyperbola around the axis.
This is achieved by conjugation by
which is a hyperbolic rotation in plane. We now calculate the remaning free parameter, the hyperbolic angle .
Direct calculation shows that
and imposing that proves the lemma. Optimality of the conjugation follows since the path
is the shortest path in connecting
For any , its operator norm satisfies
where is the maximum eigenvalue of . Therefore, by the structure of , (2.2) follows by simple calculations. ∎

The following corollary is immediate.
Corollary 2.1.
Suppose that
with . Then the conjugation constructed in Lemma 2.2 has the following estimations:
-
(1)
,
-
(2)
.
3. The Fiberd Anosov-Katok construction
In this section, we give the Anosov-Katok constructions for quasi-periodic cocycles, the construction is initialized by fixing a minimal rotation and then inductively constructing the sequences , and satisfying :
-
(1)
, , and , which forces ,
-
(2)
,
-
(3)
.
Given these parameters, assume that at the -th step of a construction, we have a constant cocycle , where
Such a cocycle is said to be -resonant with respect to .
In the -th step of the construction, we perturb this cocycle to , where
and . This perturbation is spectrally supported in the resonant Fourier mode in the anti-diagonal direction, while the constant is diagonal. The goal of the construction is to exploit the non commutativity arising by this special type of perturbation.
In the strip , we have the following estimate for the perturbation
(3.1) |
Let then one has
By assumptions (2) and (3) and direct application of Lemma 2.1, there exists such that
(3.2) |
i.e. is -resonant. Note that, by Lemma 2.1, can be chosen in the form
with estimate
(3.3) |
the last inequality holds since by our assumption and .
Let . Then we have
(3.4) |
which means the cocycle is conjugated to the resonant cocycle , and the construction can be iterated.
Consequently, let , , then, starting with an arbitrary resonant cocycle, we can construct the desired cocycle sequences:
(3.5) |
Before introducing the application of this kind of Anosov-Katok construction, we first prove that, under some mild conditions on the the sequence of parameters, the cocycle converges, and the limit cocycle is almost reducible. The first assumption of the lemma is related to the fact that we work in the real analytic category, and therefore we need to impose an exponentially fast growth condition on the resonances. In the lemma we use the notation established in this paragraph.
Lemma 3.1.
Suppose that for some
(3.6) |
Then the cocycle converges to with and
Moreover, the cocycle is almost reducible:
where
(3.7) |
(3.8) |
Proof.
Notice that by our construction (3.4) and (3.5), we have
(3.9) |
then by (3.1) and (3.3), one has
if , one has
(3.10) |
therefore, by our assumption (3.6), the cocycle converges to some . Also note our construction (3.5) implies that , then
Furthermore, by (3.9), we have
then (3.7) follows immediately. By (3.6) we know , thus the cocycle is almost reducible. ∎
Let us now provide some motivation for the precise choice of the structure of the perturbations . Note firstly that is totally determined by the triple . After conjugation by , the cocycle becomes
Ths matrix is parameterized by , which shows that the construction of the matrix is determined by the choice of the perturbation . This calculation provides the only way to perturb an elliptic constant cocycle so that it becomes conjugate to a parabolic or a hyperbolic one.

If we consider the map from to :
we see that , and correspond to the elliptic, parabolic and hyperbolic matrices in , respectively. The phenomena we study can appear only in the elliptic, and not in the parabolic and hyperbolic regimes. This is the reason why the condition is imposed. A second restriction is related with the convergence of and almost reducibility of the limit cocycle, which is guaranteed by choosing sufficiently small with respect to . In other words, we impose that tend to in the elliptic region and sufficiently fast. Then, depending on the unexpected behavior that we want to produce, different restrictions concerning the relative size of with respect to are imposed.
In §4, we construct cocycles at which the Lyapunov exponend is exactly -Hölder continuous, with . The eigenvalues of a constant parabolic matrix are -Hölder continuous, while in the hyperbolic regime (regime of fig. 2) they depend smoothly on the matrix. Therefore, we first chose to be elliptic but close to parabolic, in the regime . We then perturb to which is now hyperbolc, but close to parabolic. By controlling the distance between and , we are able to obtain -Hölder continunity.
In §5, we construct cocycles with sublinear growth. Since , the -norm of the conjugations are determined by . Constant elliptic cocycles do not grow, while constant parabolic ones grow linearly. In order to obtain growth of elliptic cocycles, we construct cocycles that are conjugated arbitrarily close to parabolic ones, in regime of the figure 2, and in that case the growth of the cocycle is comparable to , as .
In §6, we construct cocycles that are reducible, but not reducible. This is obtained by restricting the cocycle to the regime of the figure, where , as , at a prescribed speed.
4. Optimal Hölder continuity of Lyapunov exponent
4.1. Local density results
We first prove that cocycles whose associated Lyapunov exponents are exactly -Hölder continuous, with , are locally dense.
Proposition 4.1.
Fix , and rationally independent. Then, for any there exists a cocycle with and
such that Lyapunov exponent is exactly -Hölder continuous at .
Proof.
Using the isomoprhism between and , we prove the theorem in the context of cocycles. Let us first introduce two auxiliary parameters . Since , there exists such that
(4.1) |
and satisfying:
(4.2) |
We will also use the auxiliary function
which satisfies , and and is monotonic increasing on .
We can now construct iteratively the sequence . Let , and choose satisfying:
(4.3) | |||
(4.4) |
Assuming we have constructed , we choose satisfying the following properties:
(4.5) | |||||
(4.6) |
We now call
(4.7) |
and, finally, let
We also remark that by (4.5)
which, combined with (4.7), gives
Thus by (4.1) and (4.2), we have
(4.8) |
With these parameters, we can construct and then construct by the Anosov-Katok method of §3. First we check the following equality:
(4.9) |
Note that by our selection of parameters and estimate (4.8), we have
(4.10) |
On the other hand, by (4.7) and (4.1), we have
(4.11) |
Moreover, by (4.6), we have . Thus we have
by(4.3), (4.11) and (4.10). Therefore by Lemma 3.1, the limit cocycle exists and satisfies
Moreover,
(4.12) |
by (4.10) and (4.11), we have estimate
(4.13) |
Similarly, by (3.8) of Lemma 3.1, and the following estimate holds true
(4.14) |
Turning to the size of the conjugations, by the construction and equations (3.3) and (4.10), we get the estimate
(4.15) |
We now prove that Lyapunov exponent is -Hölder continuous at .
Lemma 4.1.
Let and constructed as above. Then, for any satisfying there exists independent of such that
Proof.
For any , let . By (4.8), we have , and thus there exists such that . Consequently, we have
This implies that if is small enough, there exists such that
Recall that by construction, cf. eq. (3.2), we have
(4.16) |
which means can be diagonalized by large conjugacy . However, there always exists such that conjugates into Schur Form, i.e.
(4.17) |
where .
Let now
Then by equations (4.12), (4.16) and (4.17), we have
Let now
where . By construction, can be written as
By (4.15) and the parameter choice , we have
which implies that
(4.18) |
By (4.1), one has
Since , then by (4.8), we have
Consequently, by (4.13), (4.18), (4.1), we have
Since Lyapunov exponent is invariant under conjugacy, we immediately have
On the other hand, by (4.12), we have
by continuity of Lyapunov exponent [21], we then have , consequently,
∎
We now prove that Lyapunov exponent is exactly -Hölder continuous at , i.e. that the exponent is not higher than .
Lemma 4.2.
There exists a sequence where such that
where independent of and .
4.2. Proof of Theorem 1.1:
For any , we only need to conjugate the cocycle , such that the conjugated cocycle is close to the identity, then we can apply Proposition 4.1 to finish the proof. However, as we will see, some quantitative estimates are still needed. Therefore, we will first conjugate the global almost reducible cocycle to the local regime (for example, as defined in Proposition 7.1), then apply local KAM to get the desired results.
The following result is standard, but the basis of the proof. We will sketch the proof in the appendix, for the sake of completeness.
Proposition 4.2.
Let , , . There exist such if and then, for all , there exist , and , satisfying
and
Since , then there exist such that
(4.22) |
with
where as in Proposition 4.2. Now by Proposition 4.2, for any , there exist such that
(4.23) |
satisfying
(4.24) |
Let , then we have
We now separate three cases, following the regime to which belongs.
: is elliptic. Then there exists such that
where . Since is rationally independent, there exists such that
By Proposition 4.1, there exists such that Lyapunov exponent is exactly -Hölder continuous at and
Let . Set
Then, we have
Moreover, since the Lyapunov exponent is invariant under conjugation, one can easily check that
which means that the Lyapunov exponent is exactly -Hölder continuous at .
: is parabolic. In this case, without loss of generality, we assume the eigenvalues of are . Then there exists such that
Let
where , then is elliptic, and, moreover,
This situation has been transformed into , which ends the proof for this case.
: is hyperbolic. Let the eigenvalues of be with .
We first consider the case
(4.25) |
In view of Proposition 18 of [48], there exists , with
such that
Then
with , and (4.25) implies that
Consequently, is uniformly hyperbolic by the usual cone criterion [55], which contradicts our assumptions.
Therefore, Consequently, there exists an elliptic matrix , such that
Then by (4.24), we have
which again transforms this case into , which concludes the proof.
5. Sub-linear growth of extended eigenfunction
Proposition 5.1.
Let be rationally independent and fix , , and a non-increasing sequence satisfying and . Then there exists which has sub-linear growth with rate . Moreover, it satisfies
Proof.
We construct iteratively. Firstly we construct the sequence . Let . Assuming we have constructed . We choose satisfying the following:
(5.3) |
(5.4) |
The sequence always exists since is rationally independent and . We now call
(5.5) | ||||
where
(5.6) |
With these parameters, one can construct and then by the Anosov-Katok method as in §3. Convergence of follows from the following inequality
(5.7) |
To show this, notice that (5.3) directly implies that
then by (5.4) and the choice of , we have
(5.8) |
Therefore by (5.3), (5.4), (5.5), we have
which establishes eq. (5.7). We can now apply Lemma 3.1, and thus obtain such that and
Moreover,
(5.9) |
satisfying the estimate
(5.10) |
The estimates hold because of equations (5.3), (5.4) , (5.5) and the estimate (5.8). To estimate the growth of , we first estimate the growth of the approximating cocycle .
Lemma 5.1.
Letting
we have
Proof.
First by our construction (3.5), we have
Then for any we have
(5.11) |
Since , we have the Singular value decomposition of :
for some . Thus
(5.12) | |||||
where .
The key observation is that if , we have
and
By construction,
so that
Thus we have
which proves the lemma. ∎
Now, by eq. (3.3), we have
and, by eq. (5.8),
(5.13) |
Consequently, by Lemma 5.1, we have
(5.14) |
On the other hand, by (5.9) and (5.10), we have
(5.15) |
Combining (5.14) and (5.15) can finish the proof of the proposition. ∎
Corollary 5.1.
Given . For any non-increasing sequence satisfying and . There exists a set which is dense in in the topology such that has sub-linear growth with rate at each point of .
Proof.
Proof of Theorem 1.2
If , just note , if and only if is subcritical. In fact, by Avila’s almost reducibility theorem (ART) [2, 4, 5], is subcritical, then it is almost reducible and not uniformly hyperbolic. Conversely, if is almost reducible but not uniformly hyperbolic, then the Lyapunov exponent vanishes in a band [7], which ensures the cocycle is subcritical.
Therefore, if , then is subcritical, then by Corollary 5.1, one can perturb to , so that it has sub-linear growth. Then the result follows immediately from the following lemma.
Lemma 5.2 (Avila-Jitomirskaya [8]).
Let Let be rationally independent, for some , such that is almost reducible. There exists such that for any , one can find with , , and such that
Moreover, for every , there is such that if satisfies , then there exist with and such that and
Remark 5.1.
Avila-Jitomirskaya [8] state the result for . The proof, however, applies equally well to the multifrequency case.
∎
6. Power-law localized eigenfunction
Proposition 6.1.
Given , and which is rationally independent. Then for any , there exists which is -reducible but not -reducible. Moreover, it satisfies
We again construct by the Anosov-Katok method. The construction, however, will be a bit different from Section 3. Since the goal is to construct which is reducible and not merely almost reducible.
Proof.
First, we construct a sequence where . Let , . Suppose we have constructed , then we choose satisfying
(6.1) | |||||
(6.2) |
Then we construct
By direct calculation we have
(6.3) |
Once we have these parameters, we perturb the cocycle , to, where
Let
then we have
By (6.3) and Lemma 2.1, there exists such that
Let , then we have
which means the cocycle is conjugated to , which concludes one step of the iteration.
We point out the conjugacy used here is the main difference with respect to the construction in Section 3. Indeed, by Lemma 2.1, can be chosen in the form
therefore is the form as
By Corollary 2.1 and (6.3) we have
(6.4) |
and
(6.5) |
Then we can get the following estimation on :
(6.6) |
By (6.6) we know that
(6.7) |
Thus, there exists such that . Since , one can show that
(6.8) |
where , i.e. is -reducible.
Next we will prove is not -reducible. First, we have the following :
Claim 1.
(6.9) |
Proof.
Suppose that . Then we know for all . By (6.7) we know
(6.10) |
We now analyze the structure of . Let
and
Therefore, we have
By direct calculation we know if is even
otherwise
Now we need the following crucial observation: given two set , where . Then we have
(6.11) |
if and only if . This holds since by our construction (6.2), we have , if (6.11) satisfied then must happen. Iterating this step gives .
This observation implies that for any we have
In particular, for we have
By equations (6.4) and (6.5), for any , we have
(6.12) |
Because of (6.10) and (6.12), if which is large enough, we have
By the assumption we have
which contradicts (6.2). ∎
Now we finish the proof that the cocycle is not -reducible, proceeding by contradiction. Suppose that there exists such that
(6.13) |
where . Then, is diagonalizable. Otherwise can be conjugate to a Jordan block. Then the has linear growth on which contradicts equation (6.8). Combining equations (6.8) and (6.13) we have
(6.14) |
where . Define the linear operator on :
where . Applying the Fourier transform to eq. (6.14), we have that for every
Since there exists such that . Thus is an eigenvalue of . Therefore, the two eigenvalus of are ore . Without loss of generality, we assume the eigenvalues to be the former. Since is diagonalizable there exists such that
Let . We have
(6.15) |
Then there are two cases:
: For all where , we have .
In this case, combining (6.8) and we have
where . In the frequency domain, this implies that for every we have
However, the eigenvalues of the operator are , and since this implies that . Thus we have
Due to and , we have which contradicts eq. (6.9).
: There exists , , such that .
In this case, we just need to set . Then we have
This situation has been reduced , which concludes the proof. ∎
Proof of Theorem 1.4: The proof is as for Theorem 1.1, by replacing Proposition 4.1 by Proposition 6.1. ∎
Proof of Theorem 1.3:
By [24, 47], for any , , there exist , such that if , , then . By Theorem 1.4, for any , for any , there exists with such that
Moreover, while By Lemma 5.2, there exist with and such that
Let , then
and write then we have
(6.16) |
Applying the Fourier transformatiomation to eq. , we get
i.e., , moreover, since while then the eigenfunction .
7. Appendix: Proof of Proposition 4.2:
We follow the proof of the proposition as given in [47, 22]. An alternative proof can be obtained by use of the KAM normal form, following the method introduced in [41]. We need the following result, proved in [47, 22].
Proposition 7.1.
Let , . Suppose that , . Then for any , there exists constants and such that if
(7.1) |
then there exists , and such that
More precisely, letting , , we can distinguish two cases:
-
•
(Non-resonant case) if for any with , we have
then
Moreover, .
-
•
(Resonant case) if there exists with such that
then
Moreover, can be written as with .
The proof of proposition 4.2 follows by iteration of the proposition here above. Consider the initial cocycle , where , . Without loss of generality, assume that , as well as that
where is the constant defined in Proposition 7.1. Then we can define the sequence inductively. Let , , and assume that we are at the KAM step, i.e. we have already constructed such that
where has eigenvalues and
for some , and define
By our choice of , one can check that
(7.2) |
Indeed, on the left side of the inequality decays at least super-exponentially with , while on the right side decays exponentially with .
Note that implies that Proposition 7.1 can be applied iteratively, consequently one can construct
such that
More precisely, we can distinguish two cases:
Non-resonant case: If for any with , we have
then
Let , we have
with estimate
Resonant case: If there exists with such that
with estimate
Moreover, we can write with estimate
Let , then we have
with
The last inequality is possible since by our choise .
Acknowledgements
N. Karaliolios was partially supported by LABEX CEMPI (ANR-11-LABX-0007-01) while a post-doc at Université de Lille. He is grateful to his co-authors for their warm hospitality at the Chern Institute. Q. Zhou was partially supported by support by NSFC grant (11671192,11771077), The Science Fund for Distinguished Young Scholars of Tianjin (No. 19JCJQJC61300) and Nankai Zhide Foundation.
References
- [1] V. I. Arnold. Small Denominators I; on the Mapping of a Circle into itself Isvestijia Akad. Nauk, Math series. 25 1 (1961), 21-96.
- [2] A. Avila. Global theory of one-frequency Schródinger operators. Acta Math. 215 (2015), 1-54.
- [3] A. Avila. The absolutely continuous spectrum of the almost Mathieu operator. arXiv:0810.2965
- [4] A. Avila. KAM, Lyapunov exponents and the spectral dichotomy for one-frequency Schrödinger operators. In preparation.
- [5] A. Avila. Almost reducibility and absolute continuity, arXiv:1006.0704
- [6] A. Avila; S. Jitomirskaya, The Ten Martini Problem, Ann. Math.170, 303-342 (2009).
- [7] A. Avila; S. Jitomirskaya. Almost localization and almost reducibility. J. Eur. Math. Soc. 12 (2010), 93–131
- [8] A. Avila; S. Jitomirskaya. Hölder continuity of absolutely continuous spectral measures for one-frequency Schrödinger operators. Comm. Math. Phys. 301 (2011), no. 2, 563–581.
- [9] A. Avila; S. Jitomirskaya; C. Sadel. Complex one-frequency cocycles. J. Eur. Math. Soc. 16(9) (2013), 1915–1935.
- [10] A. Avila; S. Jitomirskaya; Q. Zhou. Second Phase transition line. Math. Ann. 370 (2016).
- [11] A.Avila; Y. Last; M. Shamis; Q. Zhou. On the abominable properties of the Almost Mathieu operator with well approximated frequencies. In preparation.
- [12] A. Avila; R. Krikorian. Reducibility or non-uniform hyperbolicity for quasi-periodic Schrödinger cocycles, Ann. Math. 164, 911-940 (2006).
- [13] A. Avila; J. You; Q. Zhou. Sharp phase transitions for the almost Mathieu operator. Duke Math. J. 166(14) (2017), 2697-2718.
- [14] D. V. Anosov; A. B. Katok. New examples in smooth ergodic theory. Ergodic diffeomorphisms. Trans. Moscow Math. Soc. 23 (1970), 1-35.
- [15] S. H. Amor. Hölder continuity of the rotation number for the quasi-periodic co-cycles in . Commun. Math. Phys. 287 (2009), 565–588
- [16] J. Bochi, Discontinuity of the Lyapunov exponent for non-hyperbolic cocycles. Unpublished. (1999)
- [17] J. Bochi, Genericity of zero Lyapunov exponents. Ergodic Theory Dyn. Syst. 22(6), 1667-1696 (2002)
- [18] J. Bourgain. Positivity and continuity of the Lyapunov exponent for shifts on with arbitrary frequency vector and real analytic potential. J. Anal. Math. 96, 313-355 (2005)
- [19] J. Bourgain. Green’s function estimates for lattice Schrödinger operators and applications. Annals of Mathematics Studies. Princeton University Press, Princeton, NJ. 158 (2005).
- [20] J. Bourgain; S. Jitomirskaya. Absolutely continuous spectrum for 1D quasi-periodic operators. Invent. Math. 148 (2002), 453-463.
- [21] J. Bourgain; S. Jitomirskaya. Continuity of the Lyapunov exponent for quasiperiodic operators with analytic potential. J. Stat. Phys. 108 (2002), 1203–1218
- [22] A. Cai; C. Chavaudret; J. You; Q. Zhou, Sharp Hölder continuity of the Lyapunov exponent of finitely differentiable quasi-periodic cocycles. Math. Z. 291, 931–958 (2019).
- [23] E. I. Dinaburg; Y. G. Sinai. The one-dimensional Schrödinger equation with quasiperiodic potential. Funk. Anal. i Prilozen. 9.4 (1975), 279-289.
- [24] L.H. Eliasson, Floquet solutions for the 1-dimensional quasi-periodic Schrödinger equation. Commun. Math. Phys. 146 (1992), 447–482
- [25] L.H. Eliasson; B. Fayad; R. Krikorian. Jean-Christophe Yoccoz and the theory of circle diffeomorphisms. arXiv:1810.07107.
- [26] F. Evers; A. D. Mirlin, Anderson transitions. Rev. Mod. Phys. 80, 1355 (2008).
- [27] A. Fathi; M.R. Herman. Existence de difféomorphismes minimaux. Dynamical systems, Vol. I-Warsaw. Astérisque. 49 (1977), 37-59.
- [28] B, Fayad; A, Katok. Constructions in elliptic dynamics. Ergodic Theory Dyn. Syst. 24(5) (2005), 1477-1520.
- [29] D. J. Gilbert; D. B. Pearson. On subordinacy and analysis of the spectrum of one-dimensional Schrödinger operators, J. Math. Anal. Appl. 128 (1987).
- [30] M. Goldstein; W. Schlag. Hölder continuity of the integrated density of states for quasi-periodic Schrödinger equations and averages of shifts of subharmonic functions. Ann. Math. 154(1) (2001), 155–203
- [31] M. Goldstein; W. Schlag. Fine properties of the integrated density of states and a quantitative separation property of the Dirichlet eigenvalues. Geom. Funct. Anal. 18(3) (2008), 755–869
- [32] L. Ge; J. You; Q. Zhou. Exponential Dynamical Localization: Criterion and Applications. arXiv:1901.04258.
- [33] L. Ge; J. You; Q. Zhou. Universal hierarchical structure of extended quasi-periodic eigenfunctions. In preparation.
- [34] M. Herman. Sur la conjugaison différentiable des difféomorphismes du cercle à des rotations. Inst. Hautes Études Sci. Publ. Math. 49 (1979), 5-233
- [35] S. Jitomirskaya; D.A. Koslover; M.S. Schulteis. Continuity of the Lyapunov exponent for analytic quasiperiodic cocycles. Ergodic Theory Dyn. Syst. 29 (2009), 1881–1905
- [36] S.Jitomirskaya, I.Kachkovskiy. -reducibility and localization for quasi-periodic operators, Math Res Lett, 23, 431-444 (2016).
- [37] S. Jitomirskaya; Y. Last. Power-law subordinacy and singular spectra I. Half-line operators. Acta Math. 183 (1999), no. 2, 171-189.
- [38] S. Jitomirskaya; W. Liu Universal hierarchical structure of quasiperiodic eigenfunctions. Ann. Math. 187(3) (2016).
- [39] S. Jitormiskya; W. Liu. Universal reflective-hierarchical structure of quasi-periodic eigenfunctions and sharp spectral transition in phase. arXiv: 1802.00781v1.
- [40] S. Jitomirskaya; B. Simon, Operators with singular continuous spectrum. III. Almost periodic Schrödinger operators. Comm. Math. Phys. 165 no. 1, 201-205 (1994)
- [41] N. Karaliolios,Invariant distributions for quasiperiodic cocycles in , 2014. arXiv 1407.4763.
- [42] N. Karaliolios, Differentiable rigidity for quasiperiodic cocycles in compact Lie groups, J. Mod. Dyn. 11 (2017), 125–142.
- [43] N. Karaliolios, Continuous spectrum or measurable reducibility for quasiperiodic cocycles in , Comm. Math. Phys. 358 (2018), 741–766.
- [44] A. Katok; B. Hasselblatt. Introduction to the Modern Theory of Dynamical Systems. Encyclopedia of mathematics and its applications, 4 (1995), 27-29.
- [45] S. Klein. Localization for quasiperiodic Schrödinger operators with multivariable Gevrey potential functions. J. Spectr. Theory. 4, 1-53 (2014)
- [46] Y. Last. Spectral theory of Sturm-Liouville operators on infinite intervals: a review of recent developments. In Sturm-Liouville Theory. Birkhauser Basel. 99-120. (2005)
- [47] M. Leguil; J. You; Z. Zhao; Q. Zhou. Asymptotics of spectral gaps of quasi-periodic Schrödinger operators. arXiv:1712.04700.
- [48] Puig, J.; A nonperturbative Eliasson’s reducibility theorem, Nonlinearity, 19(2), 355–376 (2006).
- [49] B. Simon, Almost periodic Schrödinger operators: A review, Adv. Appl. Math., 3, 463-490 (1982).
- [50] M. Viana,Lectures on Lyapunov exponents. Cambridge Studies in Advanced Mathematics, Cambridge University Press. 145, (2014)
- [51] Y. Wang; J. You. Examples of discontinuity of Lyapunov exponent in smooth quasi-periodic cocycles. Duke Math. J. 162 (2013), 2363–2412
- [52] Y. Wang; Z. Zhang, Uniform positivity and continuity of Lyapunov exponents for a class of quasiperiodic Schrödinger cocycles. J. Func. Anal.268, 2525-2585 (2015)
- [53] J.Xu; L. Ge; Y. Wang. The Hölder continuity of Lyapunov exponents for a class of Cos-type quasi-periodic Schrödinger cocycles arXiv:2006.03381.
- [54] J.-C. Yoccoz. Conjugaison différentiable des difféomorphismes du cercle dont le nombre de rotation vérifie une condition diophantienne. Ann. Sci. École Norm. Sup. 17(4) (1984), 333-359
- [55] J.-C. Yoccoz. Some questions and remarks about cocycles. In Modern Dynamical Systems and Applications, 447-458. Cambridge University Press, Cambridge, 2004.
- [56] J. You; S. Zhang; Q. Zhou. Point spectrum for quasi-periodic long range operators. Journal of Spectral Theory, 4 (2015), pp.769-781.