Anderson Localization for Schrödinger Operators with Monotone Potentials over Circle Homeomorphisms
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 acting on , defined as
(1.1) |
where , is an irrational rotation on defined by , with , and is a potential function. Examples of such operators include for the almost Mathieu operator or 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 (and/or ) in many situations, as demonstrated in works such as [10, 5]. Since serves as a fundamental example of general circle homeomorphisms, a natural question arises: If is not a rotation but a more general circle homeomorphism with rotation number , can we still determine or get some information of the spectral type by the arithmetic properties of ?
As one can imagine, the answer may vary depending on properties of , , and . The study of (1.1) for general circle diffeomorphisms was initiated by [8] and [16]. In [8], the authors proved purely continuous spectrum for that is Hölder continuous, that is -smooth, and 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 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 with Diophantine and potential satisfying conditions (1) and (2) below. We work with the same conditions on but consider a more general orientation-preserving circle homeomorphism under the assumptions (1):
-
(1)
is one-periodic on and , .
-
(2)
is bi-Lipschitz monotone, i.e., there exist such that for all ,
-
(1)
Assume the invariant measure of is denoted by and that
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 to both weakly Liouville and Diophantine case. Specifically, we prove that
Theorem 1 (pure point spectrum).
Remark 1.
The theorem provides a meaningful statement for homeomorphisms with rotation number when is small or zero, which corresponds to weakly Liouville or Diophantine (see Sec 2). In fact, the smaller is, the more “irrational” is. For example, since we also proved positivity of Lyapunov exponent for all irrational in Corollary 3.3, when , this implies that , i.e. the spectrum of is pure point.
Note that condition (1) is equivalent to the existence of a bi-Lipschitz conjugacy between and , meaning that there exists a bi-Lipschitz function that is bounded from above and below such that . We acknowledge that if such a bi-Lipschitz conjugacy exists and is Diophantine, our results follow directly from [7] by a change of variables: letting , we obtain 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, ) conjugacy is currently only established for Diophantine .
As a byproduct, we also establish Lipschitz continuity of the integrated density of states (IDS, see (2.5)) for all in Lemma 3.3, as well as the continuity and positivity of the Lyapunov exponent for large in Corollary 3.3. Together with our key lemma 5.3, which is uniform in , and , these results allow us to achieve uniform localization of (see Definition 3) for sufficiently large and ”sufficiently irrational” , i.e. sufficiently small.
Theorem 2 (Uniform localization).
Let . If and if is weakly Liouville with , then has uniform localization for all .
We finally mention that a somewhat different proof was developed in [15] for unbounded lower-Lipschitz monotone and irrational rotation with Diophantine . 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 with weakly Liouville in similar ways to here. This will be explored in a future work.
Structure and key ideas. Under the assumption of (1) and allowing weakly Liouville , 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 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 in Sec 5. The main results follow immediately in Sec 6.
The extension of our results from to more general circle homeomorphisms is based on the observation that the behavior of box eigenvalues is closely related to the distribution of orbits of . While the orbits of 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 (1). Appendix A provides the key statements that enable us to carry out this argument. The extension to weakly Liouville , 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 , we use to denote the absolute value and to denote the closest distance between and integers.
Continued fraction expansion and weakly Liouville number. Any number can be written in the continued fraction expansion [20]:
with . Let denote the continued fraction approximants. They satisfy
(2.1) |
Definition 1 (weakly Liouville).
For , let
We call weakly Liouville if .
We mention that if is Diophantine111 is called Diophantine if there is and such that for all ., then .
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 is a generalized eigenfunction of an operator with respect to a generalized eigenvalue if is polynomially bounded, i.e. for some , and . Schnol’s theorem states that the spectral measure of an operator is supported on the set of its generalized eigenvalues.
According to Schnol’s theorem, to prove that has pure point spectrum, it is sufficient to show that all generalized eigenfunctions belong to . 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 and denote the restriction of to with Dirichlet and periodic boundary conditions, respectively. In particular, for the interval , we use the simplified notations and . More specifically,
Let denote the Green function, and let be its -entry. Denote , and let .
The Poisson formula provides a connection between the generalized eigenfunction and the Green function. Specifically, suppose is a generalized eigenfunction of with respect to generalized eigenvalue , then for in the interval , we have the following formula:
(2.2) |
Transfer matrix and Lyapunov exponent. Rewrite into matrix form:
where
We define the -step transfer matrix by
One can verify by induction that
(2.3) |
The Lyapunov exponent is defined to be
(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 with the density of states measure. The integrated density of states (IDS) is defined as follows:
(2.5) |
where .
Remark 2.
We can define and , analogous to and for , respectively, for .
Remark 3.
Notice that is a rank-two perturbation of . Thus we have . Thus we can also define the IDS by
(2.6) |
The function is right-continuous, non-decreasing, and approaches zero as approaches . Its derivative defines a unique probability measure, called the density of states measure . The relation between the density of states measure and the Lyapunov exponents is known as the Thouless formula. We state it here without proof, but refer the interested reader to [3] for more details:
(2.7) |
3. Positive Lyapunov Exponent
In this section, we first establish some fundamental properties of box eigenvalue functions, which are the eigenvalues of . We then derive estimates for the distance between these eigenvalue functions. Using these estimates, we obtain the Lipschitz continuity of the IDS with respect to and prove the positivity of the Lyapunov exponent for large .
Recall that is the periodic restriction of to . Let , be the eigenvalues of in increasing order. We refer to as the box eigenvalue functions. Now we establish some of their basic properties:
Proposition 3.1.
have the following properties:
-
(1)
is -periodic, continuous on except at . By rearranging these discontinuity points in an increasing order, we denote them by . We also denote .
-
(2)
is bi-Lipschitz continuous with respect to the invariant measure, and strictly increasing on each . In fact,
(3.1) -
(3)
At each jump , we have
Remark 4.
Proof.
-
(1)
Note that the box eigenvalue functions are roots of the characteristic polynomial . Therefore, each is continuous with respect to the coefficients of , which are polynomials of . Since is only discontinuous at , is only potentially discontinuous at .
- (2)
-
(3)
Notice that
(3.2) where such that . This leads to the second inequality since is positive semi-definite. To derive the first inequality, by (3.2), let be the matrix obtained by deleting the row and column from or . Let be the eigenvalues of . By eigenvalue interlacing theorem,
(3.3) Therefore, , for all .
∎
Horizontal comparison. From now on, we fix , and consider 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 and horizontally. Note that the estimate is uniform in .
Lemma 3.1.
For any ,
Proof.
Define an unitary matrix where are standard unit vectors. Then
By Lemma A.2,
The result follows from Lindskii’s theorem. ∎
Corollary 3.2.
For any ,
Proof.
Vertical comparison. We now estimate the lower bound of vertical comparison between eigenvalue functions. Unfortunately, the vertical distance between two closest eigenvalue functions and is not always positive. However, we can show that at most eigenvalues can be very close to each other, others will be nicely seperated from them.
Lemma 3.2.
Given , and . For any , there is a , such that for any satisfying and , we have
(3.4) |
Proof.
First notice that given , there exists , such that each point in falls in precisely one interval among . Then for any ,
In particular, we can pick such that and are on the graph of the same , defined in Remark 4. Put such pairs of together and denote the set by . Then includes out of subintervals created by the partition on , where each intervals have the same invariant measure. Thus by pigeonhole principle,111In fact, we could bound , 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 in the vertical comparison of eigenvalue functions. It implies that vertical differences of eigenvalue functions at any is uniformly controlled by the largest vertical differences among all . This observation can be useful in dealing with singular where certain ’s are too small or the case when is flat at some ’s.
Thus
when .
∎
Lipschitz continuity of IDS. Recall that and
(3.5) |
Now we can derive Lipschitz continuity of and from vertical distance of :
Lemma 3.3.
Given , , , we have
(3.6) |
Proof.
Fix and . For any , we see from Lemma 3.2 that any interval of length contains at most eigenvalues for large enough. This allows us to estimate the number of eigenvalues between and :
for any . Let , we get
Since this inequality is true for all , 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 of is continuous in and admits a lower bound
(3.7) |
Therefore, is uniformly positive if .
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 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 and arbitrary piecewise potentials.
Lemma 4.1.
For any and , there exists an such that for all
Moreover, can be chosen to be uniform in as long as is continuous on interval .
Proof.
Theorem 3 (Large deviation theorem).
Fix such that . There exists such that for any , there is such that for any ,
(4.1) |
where is the Lebesgue measure. Moreover, the set on the left-hand side is composed of at most many intervals.
Proof.
Recall that
Denote for convenience
Notice that is monotone and at . Thus “large deviation” happens near . The aim is to estimate how large this set can be without rising too high. The idea is since are well-seperated, only the closest (to ) several contribute the most to the negativity of , the rest are nicely controlled.
To do so, we split eigenvalues into three clusters: above , around , and below . Notice that by and Lemma 3.2, we can make sure that
-
(1)
The cluster of eigenvalues above , denoted by , in an increasing order with .
-
(2)
The cluster of eigenvalues below , denoted by , in an decreasing order with .
-
(3)
The cluster of the rest of eigenvalues, denoted by , with the number of eigenvalues in this cluster does not exceed some uniform in .
For example, this can be achieved by considering the closest eigenvalues to to be in the third cluster and every eigenvalue above/below them to be in the first/second cluster. Here is to guarantee the lower and upper bound estimates above and is due to . In fact, we can do the same thing for , then we just need to pick the closest eigenvalues instead of .
Now decompose , correspondingly,
where , where .
Claim 1.
Let ,
By Corollary 3.2 and the claim, we have for any ,
Here we considered all maximum potential perturbation of all at the maximum potential place . There might be extra terms of that does not pair to but since there are only finitely many terms and they are bounded, the result is still true with a modification of . For the same reason, the inequality holds for as well.
Thus there is such that
(4.2) |
Claim 2.
There is such that for large enough, for small enough,
Proof.
If is such that , then . Since there are at most eigenvalues in , thus there is some such that
(4.3) |
Among all , there are at most intervals of such that some satisfies (4.3). In fact, there are at most intersections of and . Since each is monotone, (4.3) is only possible for near such intersections. And by Prop. 3.1 are lower-Lipschitz with respect to invariant measure. Thus for large enough,
∎
Thus we have proved the result with instead of . Now we need the last component of the proof:
Claim 3.
For any , uniform in when is large enough.
Proof.
In fact, since the operator is bounded, . Together with (4.2), we get
(4.4) |
uniformly in when is large enough (depending on ). The same holds for .
On the other hand, by Lemma 4.1, for any , eventually. While by definition of Lyapunov exponent (2.4), is the limiting averaging of . But and are connected by (2.3). Thus we see that on a set of measure at least , the following is true for either , or :
(4.5) |
If , combining (4.5) and (4.4) gives us what we want. Otherwise, we first notice by row expansion of determinant, we have
(4.6) | |||
(4.7) |
Then when , by (4.6), we have either or satisfies (4.5) so we can combine it with (4.4) to derive the result. If , by (4.7), we have either or satisfies (4.5). For the former case, we get the result. For the latter, combining (4.5) and (4.4) with instead of . The claim follows. ∎
Now the result follows immediately: For any , apply Claim 3 to get eventually so that
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 . 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 is -regular if there is an interval with
(5.1) |
such that
Otherwise we say is -singular.
Lemma 5.1.
Fix such that . For large enough, for any , if is -singular, then for any ,
(5.2) |
Furthermore, let denote the number of such , then
Proof.
Since is -singular, for any satisfying (5.1), in particular, for any , , we have
(5.3) |
Notice that
(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 is large enough
(5.5) |
By (5.3),(5.4) and (5.5), we see that
Thus we proved (5.2). The bound of follows from direct computation when . ∎
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 singular points are at least “” away from each other:
Lemma 5.2 (Either or argument).
Let be as in Theorem 3. Assume and satisfy . For any , we have that for large enough, and for any , either or is -regular for any .
Proof.
WLOG assume . For any , assume both and are -singular. By Lemma 5.1, we have
for any . Notice further that
Thus the two intervals of have no intersection. Overall there are many possible such that . By Theorem 3 and pigeonhole principle, there are such that
(5.6) |
Notice that By Lemma A.1 and (A.2), we have
This implies that
which leads to a contradiction with the assumption. ∎
Lemma 5.3.
Let be as in Theorem 3. If satisfy
-
(1)
is a generalized eigenvalue of ,
-
(2)
,
then is an eigenvalue with exponentially decaying eigenfunction. Denote the normalized eigenfunction by with .
Proof.
Take any . Let be a generalized eigenfunction of with respect to . Thus where . We first prove decay exponentially so that is an eigenvalue, then we prove the decay is uniform in the sense of (5.7).
WLOG assume . By (2.2), is eventually -singular. By Lemma 5.2, we have for large enough, any is -regular. Notice further that since for . Thus eventually for any , there is such that . We derive exponential decay by considering two cases seperately:
-
(1)
If , is -regular, by (2.2), we have for arbitrarily small , eventually
(5.8) - (2)
Combining (5.8) and (5.9) gives us the first half of the theorem. Now since , we can normalize it so that .
The key point of the second half is the uniformity in . Take to be the leftmost maximum point of . By (2.2), we see that maximum point is always -singular for all . Thus is -regular if . We can now repeat the estimates (5.8) and (5.9) above with the new, uniform (in ) improvement that instead of , where we get
Since is arbitrary and is arbitrarily small once is large enough uniformly in . 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.
Definition 3 (Uniform localization).
An operator exhibits uniform localization if there exixts such that for any pair of eigenvalue and eigenfunction , , there exists such that
Appendix A Orbital analysis
It is well-known that irrational rotation on 1d-torus, , has best-approximation property, c.f. [20],
(A.1) |
with estimates
(A.2) |
where is defined in (2.1). Furthermore, the orbits of is also well-understood, we cite [7, Proposition 4.1, 4.2] here:
Proposition A.1.
Let . The points , splits into “large” gaps of length and “small” gaps of length . Furthermore, we have the estimates
For a general measure-preserving circle homeomorphism with rotation number , such kind of best approximate properties and orbital analysis can be done in a similar way with invariant measure instead of distance . In fact,
Lemma A.1 (Best approximation).
For any and ,
where .
Proof.
Note that Lemma A.1 holds when the invariant measure is the Lebesgue measure - in other words, when the map 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 with a rotation , and is also the distribution function for the unique invariant measure . Hence, for any and , we have
∎
Lemma A.2.
Fix , Let . The points split into “large” gaps of invariant measure and “small” gaps of invariant measure . Furthermore, we have the estimates
Proof.
To prove the theorem, let us first introduce the dynamical partition on the circle by following the convention using in [8]. For each , let be the interval between and . It can be verified by induction in that the following collection of intervals forms a dynamical partition of
That is, they are disjoint except for the endpoints, and the union cover the whole circle. Notice that intervals in all have smaller invariant measure than intervals in , thus we call them “short” and “long” intervals correspondingly. One can check by induction on that, each “long” interval in dynamical partition is divided into “long” intervals and one “short” interval in dynamical partition. More specifically,
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 is the invariant measure of , for dynamical partition , we have
(A.3) |
By (A.3), we get
Moreover, since (A.3) holds for any , we also get . So,
(A.4) |
The last inequality follows from the recurrence relation (2.1) and :
(A.5) |
By the comparability between and the Lebesgue measure on a circle (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.