A Nonperturbative Eliasson's Reducibility Theorem
Joaquim Puig
Departament de Matemàtica Aplicada I, Universitat Politècnica de Catalunya Av. Diagonal 647, 08028 Barcelona, Spain joaquim.puig@upc.edu
Abstract
This paper is concerned with discrete, one-dimensional Schrödinger operators with real analytic potentials and one Diophantine frequency. Using localization and duality we show that almost every point in the spectrum admits a quasi-periodic Bloch wave if the potential is smaller than a certain constant which does not depend on the precise Diophantine conditions. The associated first-order system, a quasi-periodic skew-product, is shown to be reducible for almost all values of the energy. This is a partial nonperturbative generalization of a reducibility theorem by Eliasson. We also extend nonperturbatively the genericity of Cantor spectrum for these Schrödinger operators. Finally we prove that in our setting, Cantor spectrum implies the existence of a
-set of energies whose Schrödinger cocycle is not reducible to constant coefficients.
Keywords: Quasi-periodic Schrödinger operators, Harper-like equations, reducibility, Floquet theory, quasi-periodic cocycles, skew-product, Cantor spectrum, localization, irreducibility, Bloch waves.
AMS Subject Classification: 47B39, (37C99s,37E10,37J40).
Contents
1 Introduction. Main results
Recently there has been substantial advance in the theory of quasi-periodic Schrödinger operators, both continuous and discrete, combining spectral and dynamical techniques. These operators arise naturally in many areas of physics and mathematics.
They appear in the study of electronic properties of solids [AA80, Jan92, OA01] , in the theory of KdV and related equations [JM82, Joh88, Chu89] or in Hamiltonian mechanics [Sin85] . Moreover, their eigenvalue equations are second order differential or difference linear equations with quasi-periodic coefficients like Hill's equation with quasi-periodic forcing [BPS03] or the Harper equation [KS97, DJKR04] which display a rich variety of dynamics ranging from quasi-periodicity to uniform and nonuniform hyperbolicity.
In this paper we pursue this fruitful combination of spectral and dynamical methods to study discrete, one-dimensional Schrödinger operators
|
(1)
|
where
is a real analytic function (the potential ),
is a phase and
a Diophantine frequency. The latter means that there exist positive constants
and
such that the bound
|
(2)
|
holds for any integer
This condition will be written as
In particular, this means that
is nonresonant,that is
unless
The best-studied example of quasi-periodic Schrödinger operator is the Almost Mathieu operator where
, being
a real coupling parameter. A vast amount of literature is devoted to the study of the spectral properties of this operator (see Simon [Sim82, Sim00] , Jitomirskaya [Jit95, Jit02] and Last [Las95] for surveys and references). In [Pui04a, Pui04b] (see also Avila & Jitomirskaya [AJ05] for the recent extension to the remaining frequencies) the Cantor structure of the spectrum of the Almost Mathieu operator was derived from the use of a localization result by Jitormirskaya [Jit99] and a dynamical analysis of its eigenvalue equation, the so-called Harper equation. This was a long-standing conjecture known as the “Ten Martini Problem”. The combined approach there is not limited to the Almost Mathieu operator as we plan to make evident in this paper.
The eigenvalue equation of a quasi-periodic Schrödinger operator
is the following Harper-like equation
|
(3)
|
where
is called the energy or spectral parameter. Since we want to study dynamical properties of this equation, it is better to transform it into a first-order system, obtaining the associated quasi-periodic skew-product on
,
|
(4)
|
which can be seen as an iteration of the corresponding Schrödinger cocycle,
on
. Here
denotes the matrix-valued function
|
(5)
|
The simplest class of Schrödinger cocycles occurs when
because in this case
does not depend on
(we will say that the corresponding cocycle is in constant coefficients ). In analogy with periodic differential equations we may try to reduce a quasi-periodic Schrödinger cocycle to constant coefficients.
Let us introduce first the notion of conjugation between cocycles, not necessarily of Schrödinger type. Here we restrict ourselves to the case of
-valued cocycles, although the notion of reducibility applies to more general cocycles (see [Pui04b] for an exposition). Two cocycles
and
on
are conjugated if there exists a continuous and nonsingular conjugation
such that the relation
holds for all
In this case the corresponding quasi-periodic skew-products
and
are conjugated through the change of variables
, so that they share the same dynamical properties.
A cocycle
is reducible to constant coefficients if it is conjugated to a cocycle
with
not depending on
(i.e. with constant coefficients). In this case,
is called a Floquet matrix. Sometimes it may be necessary to “halve the frequency” if we do not want to complexify the system (although this case will not be treated in this paper). In contrast with the situation in the periodic case (when the nonresonance condition fails), quasi-periodic cocycles need not to be reducible to constant coefficients (see Theorem 5 and the following Remark 7 ).
The reducibility of a Schrödinger cocycle and the eigenvalues of the reduced Floquet matrix have implications for the spectrum of the corresponding Schrödinger operator. The spectrum of
on
is a compact subset of the real line which we denote by
since it does not depend on
. It is known that an energy
lies in the spectrum of a Schrödinger operator if, and only if, the corresponding skew-product has an exponential dichotomy (it is uniformly hyperbolic), see Johnson [Joh82] . Under our assumptions,
real analytic and
Diophantine, this is equivalent to the reducibility of the Schrödinger cocycle to constant coefficients with a hyperbolic Floquet matrix (all its eigenvalues are outside the unit circle), see Johnson [Joh80] .
For energies in the spectrum, the situation is much more involved. However, when in addition to the present hypothesis, the potential
is small (in some complex neighbourhood around
which depends on
and
), then reducibility can be obtained by kam methods for a set of energies in the spectrum of large measure, see Dinaburg & Sinai [DS75] and Moser & Pöschel [MP84] .
A breakthrough in the kam approach came with Eliasson [Eli92] who proved, among other statements, that reducibility to constant coefficients holds for almost every energy provided the potential is small enough and
is Diophantine,
for some
and
(see Section 2.3 for a more precise formulation. Like the results in the previous paragraph, the smallness condition here depends on the precise Diophantine conditions on
. Eliasson's result, as well as the above kam results presented above, holds for real analytic or
potentials
depending on several frequencies.
Our main result states that in the presence of only one frequency,
in the notation above, the smallness condition in Eliasson's theorem does not depend on the constants
and
of the Diophantine condition as long as
is Diophantine (that is, it is “nonperturbative” in some sense). To be more precise, we consider real analytic potentials
having an analytic extension to
for some
such that
(the set of such potentials will be denoted by
) and a Diophantine frequency
for some positive constants
and
. Our extension of Eliasson's theorem reads as follows.
Theorem 1
Let
be a positive number. Then, there is a constant
such that, for any real analytic
with
the Schrödinger cocycle
is reducible to constant coefficients for every Diophantine frequency
and almost all
(with respect to Lebesgue measure).
The proof of this Theorem will be given in Section 3 .
Remarks 2
-
1.
Recently Avila & Krikorian [AK03] proved Theorem 1 with more restrictive hypothesis on
(although it is also a full measure condition). In fact, we will see that both results follow from a nonperturbative theorem on localization by Bourgain & Jitomirskaya [BJ02a] .
-
2.
When the potential is defined on a
-dimensional torus,
, Eliasson's theorem holds, but a nonperturbative version like Theorem 1 cannot be true, as Bourgain showed in [Bou02] . Indeed, he proved that, if
is a trigonometric polynomial with a nondegenerate maximum, there is a set of frequencies
, with positive Lebesgue measure, for which the operators
have some point spectrum. This point spectrum is incompatible with reducibility to constant coefficients. See Bourgain [Bou04b, Bou04a] for the differences between the cases of one and several frequencies.
-
3.
Let us stress that Theorem 1 is not a full nonperturbative version of Eliasson's theorem because the set of energies whose corresponding Schrödinger cocycle is reducible to constant coefficients is not explicitly characterized as it is in Eliasson's theorem (see Section 2.3 ).
-
4.
The smallness condition in Theorem 1 ,
is given by the localization result in [BJ02a] and can be explicitly given in terms of
,
,
and
. If
is kept fixed, then
as
Theorem 1 has some implications for the spectral properties of Schrödinger operators. The first one refers to the properties of solutions of the eigenvalue equation. An immediate application of Theorem 1 is the existence of analytic quasi-periodic Bloch waves for almost all
in the spectrum. An analytic quasi-periodic Bloch wave for a Harper-like equation ( 3 ) is a solution of the form
|
(6)
|
where
is called the Floquet exponent and
is a nontrivial analytic function. In Section 3 we will also prove the following about the existence of quasi-periodic Bloch waves.
Corollary 3
Let
,
and
be as in Theorem 1 . Then for (Lebesgue) almost all values of
in the spectrum
, the equation
has analytic quasi-periodic Bloch waves.
Our second application deals with the structure of the spectrum of quasi-periodic Schrödinger operators. The Cantor structure of the spectrum is not specific of the Almost Mathieu operator. Indeed, although the proof of the “Ten Martini Problem” [Pui04a] is restricted to this model, Cantor spectrum is “generic” in our setting. More precisely, if we consider the set
of real analytic functions furnished with the
norm, one has the following.
Theorem 4
Let
. Then there is a constant
such that for every Diophantine
there is a generic set of real analytic potentials
with
such that the spectrum of the Schrödinger operator
is a Cantor set.
Our last application is concerned with the existence of Schrödinger cocycles which are not reducible to constant coefficients. We will see that Cantor spectrum and nonreducibility are related concepts in our setting.
Theorem 5
Let
. Then there is a constant
such that if
is Diophantine and
, with
, is such that
is a Cantor set then for a
-dense subset of energies in the spectrum the corresponding Schrödinger cocycle is not reducible to constant coefficients (by a continuous transformation).
Remark 6
According to Theorem 1 , the
-set above has zero Lebesgue measure.
In particular, using Theorem 4 it is possible to give a nonperturbative version of a result in Eliasson [Eli92] , namely, that the existence of a
-subset of “nonreducible energies” is a generic property.
Remark 7
This kind of nonreducibility holds for a zero measure subset of energies and the corresponding Schrödinger cocycle has zero Lyapunov exponent.
Nonreducibility results can be obtained using, for instance, a result by Sorets & Spencer [
SS91]
, who prove that if the potential
is large enough then the Lyapunov exponent is positive for all energies in the spectrum and this prevents reducibility.
Let us finally outline the contents of this paper. In Section 2 we introduce some of the preliminaries needed for the proof of the main theorem. In Section 3 this is used to prove 1 and 3 using a similar technique to the one used for the Almost Mathieu operator. The applications are included in Section 4 .
2 Preliminaries
In this section we present some of the tools that will be needed in the proof of Theorem 1 . As said in the introduction, we plan to extend some of the ideas in the proof of the “Ten Martini Problem” given in [Pui04a] with the aid of a nonperturbative localization result for long-range potentials by Bourgain & Jitomirskaya [BJ02b] . In Section 2.1 we introduce a convenient version of Aubry duality, which will lead us to consider certain long-range operators which are not of Schrödinger type. In Section 2.2 we give the definition and some properties of the integrated density of states ids for these operators and its relation with Aubry duality. Finally, in Section 2.3 this ids is linked to the fibered rotation number of quasi-periodic cocycles to give a more precise version of Eliasson's result.
2.1 Aubry Duality
Aubry Duality [AA80] was originally introduced for the study of the Almost Mathieu operator but the same idea (which is Fourier transform) works for other potentials. Let us give first the heuristic approach and then a more rigorous one.
Assume that
has an analytic quasi-periodic Bloch wave,
|
(7)
|
being
analytic and
the Floquet exponent. If
are the Fourier coefficients of
, a computation shows that they satisfy the following difference equation
where
are the Fourier coefficients of
,
This difference equation is the eigenvalue equation of the operator
which we call a dual operator of
. This is a self-adjoint and bounded operator on
(because
is real analytic) but it is not a Schrödinger operator unless
is exactly the cosine (this is what makes the Almost Mathieu operator so special). Such an operator will be called a long-range (quasi-periodic) operator even if it may be a finite-differences operator (if
is a trigonometric polynomial).
If
is nonresonant, the spectrum of the long-range operators
does not depend on the chosen
, so that one can write
This naive approach to Aubry duality shows that whenever
is a value in the spectrum
such that
is an analytic quasi-periodic Bloch wave with Floquet exponent
, then
is a point eigenvalue of the dual operator
whose eigenvector decays exponentially and, thus,
. The converse is also true: one can pass from exponentially decaying eigenvalues of
to quasi-periodic Bloch waves of
with Floquet exponent
.
The argument given above heavily relies on the existence of quasi-periodic Bloch waves or, equivalently, exponentially localized eigenvectors. Nevertheless both operators can be related without the assumption of such point eigenvalues.
This was done by Avron & Simon [AS83] . Here we will follow the idea by Gordon, Jitomirskaya, Last & Simon [GJLS97] (see also Chulaevsky & Delyon [CD89] ), who studied duality for the Almost Mathieu operator, although it can be extended to the general case, see Bourgain & Jitomirskaya [BJ02b] . The idea is to shift to more general spaces where the extensions of quasi-periodic Schrödinger operators and their duals are unitarily equivalent. Note that it is not true that the operators
and
are unitarily equivalent, since their spectral measures will, in general, be very different.
Let us consider the following Hilbert space,
which consists of functions
satisfying
The extensions of the Schrödinger operators
and their long-range duals
to
are given in terms of their direct integrals , which we now define. The direct integral of the Schrödinger operator
, is the operator
, defined as
and the direct integral of
, denoted as
, is
These two operators are bounded and self-adjoint in
. Let us now see that, for any fixed real analytic
and nonresonant frequency
, the direct integrals
and
are unitarily equivalent; i.e. there exists a unitary operator
on
such that the conjugation
holds. By analogy with the heuristic approach to Aubry duality in the beginning of this section, let
be the following operator on
,
where
is the Fourier transform. At a formal level this acts as
if we disregard the convergence of the sum in
The map
is unitary and satisfies
by construction of the dual long-range operators in terms of the Schrödinger operators. Therefore, the direct integrals
and
are unitarily equivalent and, in particular, their spectra are the same,
Hence, the spectrum of a quasi-periodic Schrödinger operator and its dual are the same. In the next section we will introduce the integrated density of states for Schrödinger operators (and their long-range duals) and we will see that this function is preserved by Aubry duality.
2.2 The integrated density of states and duality
The integrated density of states, ids for short, is a very useful object for the description of the spectrum of quasi-periodic Schrödinger operators and more general quasi-periodic self-adjoint operators. Here we want to introduce it both for quasi-periodic Schrödinger operators and their long-range duals. In order to give a unified approach, let us consider a more general class of operators.
If
are real analytic functions,
and
are their Fourier coefficients,
is a nonresonant frequency and
, let
be the following operator
acting on
which is bounded and self-adjoint. The operators in the previous section occur as particular cases,
Let us now define the ids for the operators
Take some integer
and consider
, the restriction of the operator
to the interval
with zero boundary conditions. Let
Then, due to the nonresonant character of
the limit
exists, it is independent of
and of the boundary conditions imposed above. It is called the integrated density of states of the operator
We will write this as
The map
|
(8)
|
is increasing and it is constant exactly at the open intervals in the resolvent set of the spectrum of
. It is the distribution function of a Borel measure
,
called is the density of states of the operator
which is supported on the spectrum of
. In the Schrödinger case we will use the notations
and
for their long-range duals.
The ids of the operators
can be seen as an average in
of the spectral measures of the operators (see Avron & Simon [AS83] ). By the spectral theorem we know that there is a Borel measure
such that
|
(9)
|
for every continuous function
, being
the delta function. The measures
are spectral measures in the sense that the spectral projection of
over a certain subset
of the spectrum is zero if, and only if,
. Avron & Simon prove that, for any continuous function
An approximation argument shows that, for any Borel subset of the spectrum,
In particular,
if
for Lebesgue almost every
.
Using this characterization of the ids one can prove the following adaption of the duality of the ids given in [GJLS97] .
Theorem 8 ([GJLS97] )
Let
and
be the integrated density of states of
and
respectively, for some real analytic
and nonresonant frequency
. Then
for all
.
Proof: Let
which belongs to
. Then
Moreover by ( 9 ) and the unitary equivalence between
and
we have that, for any continuous
Therefore, since
and
are the Borel measures such that
and
for every continuous
the two measures must coincide (and also their distribution functions,
and
).
Let us end this section summing up some facts useful in the sequel.
Proposition 9
Let
be real analytic,
nonresonant and
a spectral measure of
. Assume that there is a measurable set
such that
for almost every
. Then
and
2.3 The rotation number and Eliasson's theorem revisited
We have seen in the previous section that it is possible to assign an ids for quasi-periodic Schrödinger cocycles using its associated operator. Here we will see that it is possible to define an extension of this object, the fibered rotation number, for more general quasi-periodic cocycles. This object, introduced originally by Herman [Her83] in this discrete case (see also Johnson & Moser [JM82] , Delyon & Souillard [DS83b] ), allows us to give a version of Eliasson's theorem for these cocycles. Let us follow the presentation by Krikorian [Kri] .
Let
be a quasi-periodic cocycle on
which is homotopic to the identity. That is
is a continuous map (although we will later assume that it is real analytic) that is homotopic to the identity (for example a Schrödinger cocycle). The fibered rotation number, which we now introduce measures how solutions wind around the origin in
in average.
Let
be the set of unit vectors of
and let us denote by
the projection given by the exponential
, identifying
with
. Because of the linear character of the cocycle and the fact that it is homotopic to the identity, the continuous map
|
(10)
|
is also homotopic to the identity. Therefore, it admits a continuous lift
of the form:
such that
for all
and
. The map
is independent of the choice of
up to the addition of a constant
, with
. Since the iteration
is uniquely ergodic on
for all
, one has that the limit
exists modulus
and it is independent of
, see Herman [Her83] . This object is called the fibered rotation number of
, and it will be denoted by
.
The fibered rotation number of a Harper-like equation is defined as the fibered rotation number of the associated Schrödinger cocycle on
and will be denoted as
.
The rotation number of a Harper-like equation can be linked to its ids . Indeed, using a suspension argument (see Johnson [Joh83] ) it can be seen that
The rotation number is not invariant under conjugation, but one has the following.
Proposition 10 (cf. [Kri] )
Let
be nonresonant and
and
be two quasi-periodic cocycles on
homotopic to the identity. If they are conjugated for some continuous
then
where
is the degree of the map
. If the conjugation
is not defined on
but on
and it has degree
, then
Keeping in mind this result, we can define two classes of rotation numbers which are preserved under conjugation. An important class is that of resonant rotation numbers. A number of the form
for some
is called resonant with respect to
. We can also define the class of fibered rotation numbers which are Diophantine with respect to
. Its elements are the numbers
such that the bound
holds for all
and suitable fixed positive constants
and
. Both classes of rotation numbers are constant under conjugation.
With these definitions we can give a more precise version of Eliasson's reducibility theorem for general quasi-periodic cocycles on
homotopic to the identity. Again the result is valid for more than one frequency, but we restrict ourselves to this one-dimensional case.
Theorem 11 ([Eli92] )
Let
,
be Diophantine and
be a matrix in
. Then there is a constant
such that, if
is real analytic with
and the rotation number of the cocycle
is either resonant or Diophantine with respect to
, then
is reducible to constant coefficients of a quasi-periodic (perhaps with frequency
) and analytic transformation.
Remark 12
The proof of this theorem was originally given in [
Eli92]
in the continuous case and for Schrödinger operators (instead of cocycles), although it extends to the setting of Theorem 11 .
Applied to Schrödinger cocycles one obtains the perturbative version of Theorem 1 with the additional characterization of the set of reducible energies in terms of its rotation number. More precisely, the theorem above implies that the set of “reducible” rotation numbers is of full measure in
. To obtain a full-measure condition on the energies it is necessary to use some facts on the growth of the rotation number at these reducible points which will be also used in Section 3 and which are due to Deift & Simon [DS83a] .
3 Proof of Theorem 1
We are now ready to show that Theorem 1 is a direct consequence of the following result by Bourgain & Jitomirskaya [BJ02b] , which we restate in a convenient way:
Theorem 13 ([BJ02b] )
Let
be a positive number. Then there is a constant
such that, for any real analytic
with
and Diophantine
there is a set
, of zero (Lebesgue) measure such that, if
, the operator
has pure point spectrum with exponentially decaying eigenfunctions.
Remarks 14
-
1.
In [BJ02a] , the bound
depends on
and
. Since
belongs to
, all these previous norms can be controlled by
.
-
2.
The set
consists of those phases
for which the relation
|
(11)
|
holds for infinitely many values of
, where
. For any Diophantine
, this is a set of zero Lebesgue measure.
Our strategy to prove Theorem 1 will be, first of all, to show that Corollary 3 is a simple consequence of Theorem 13 and the duality of the ids . Then, in Section 3.1 it will be shown that Corollary 3 actually implies Theorem 1 .
Let
and
,
and
be as in the Theorem 13 . As a consequence of Proposition 9 , the set
where
is the set of point eigenvalues of
given by Theorem 13 satisfies that
. Indeed, according to Proposition 9 we only need to show that
for all
, where
are the spectral measures of the long-range operators
. This is a consequence of the fact that the spectral measures
, for
are supported on the set of point eigenvalues of the corresponding operator.
Therefore also
due to Proposition 9 . To prove Corollary 3 it only remains to show that also the Lebesgue measure of
is zero. To do so, one can invoke Deift & Simon [DS83a] . For almost periodic discrete Schrödinger operators they prove that for Lebesgue almost every
in the set where the Lyapunov exponent is zero, one has the inequality
|
(12)
|
Thus, under the additional assumption that the Lyapunov exponent vanishes in the spectrum, the inequality ( 12 ) implies that if
is a subset of
with
then also the Lebesgue measure of
is zero.
As a consequence of Bourgain & Jitormirskaya [BJ02a, BJ02b] , for any
in
(with
) the Lyapunov exponent is zero. Therefore, the set
has Lebesgue measure zero and for the values of
in its complement in the spectrum,
which is a total measure subset of
, the corresponding Harper-like equation
|
(13)
|
has an analytic quasi-periodic Bloch wave, using the argument of duality in the beginning of Section 2.1 . Indeed, we saw that if
is a point eigenvalue of the operator
whose eigenfunction decays exponentially then the Harper-like equation ( 13 ) has an analytic quasi-periodic Bloch wave with Floquet exponent
.
This completes the proof of Corollary 3 .
3.1 From Bloch waves to reducibility
In this section we will see how Corollary 3 (which we proved in the previous section) implies our main result, Theorem 1 . By this corollary we know that if
,
and
are as in Theorem 13 then, for almost all
, the equation ( 13 ) has an analytic quasi-periodic Bloch wave with Floquet exponent
. Since we only want to prove a result for almost every
, it is sufficient to show that if
is such that
for all
and ( 13 ) has an analytic quasi-periodic Bloch wave with this Floquet exponent
, then the corresponding Schrödinger cocycle
is reducible to constant coefficients.
Remark 15
If
is resonant with respect to
,
for some integers
, then one can also prove reducibility [
Pui04a]
. In Section 4.3 we will consider the case of
which will be used for the Cantor structure of the spectrum.
The existence of a Bloch wave for Equation ( 13 ) implies that the Schrödinger cocycle has the following quasi-periodic solution
|
(15)
|
for all
. Moreover, writing
|
(16)
|
and
|
(17)
|
where the bar denotes complex conjugation, one always has the relation
|
(18)
|
where
Obviously,
will only define a conjugation between the cocycles
and
if it is nonsingular. Because of ( 18 ), the determinant of
is constant as a function of
and it is purely imaginary. In particular,
and
are linearly independent for all
if, and only if, they are independent for some
. In the case that
and
are linearly independent, it is not difficult to prove reducibility to constant coefficients of the cocycle.
Lemma 16
Let
be a real analytic map and
be nonresonant.
Assume that there is an analytic map
, with
and
linearly independent, such that
holds for all
, where
. Then the cocycle
is reducible to constant coefficients by means of a real analytic transformation. Moreover, the Floquet matrix can be chosen to be of the form
|
(19)
|
Proof: Let
as in ( 17 ),
and
be the determinant of
. Therefore
defines a conjugation between
and
because
and
are linearly independent,
is real analytic and, for every
,
is nonsingular.
Moreover, from the conjugacy ( 18 ) and the nonresonance of
,
is constant as a function of
. By the linearity of our system, we choose this constant value to be
(recall that, due to the form of
its determinant must be purely imaginary).
To obtain the real rotation consider the composition
where
is the constant matrix
Then
satisfies the desired conjugation
being
the rotation of angle
given by ( 19 ). Thanks to the construction
is real and with determinant one.
To complete the proof of Theorem 1 , it only remains to rule out the possibility that
satisfies ( 14 ) and
and
are linearly dependent at the same time. Recall that these two vectors are linearly independent for all
if, and only if, they are linearly independent for some
Note that both
and
are different from zero for all
by construction. Assume that
and
were linearly dependent for all
. Since these vectors depend analytically on
, there would exist an analytic
and an integer
such that
for all
. Using that
and
are quasi-periodic solutions of
, this would imply that
Therefore,
must satisfy the following small divisors equation
for all
where
is some fixed integer. Clearly, such analytic
cannot exist unless
which is a contradiction with the nonresonance condition ( 14 ). This ends the proof of Theorem 1 .
4 Applications
In this section we will prove several consequences of the main theorem which are summarized in theorems 4 and 5 . In Section 4.1 we will present the setting of this section. In 4.2 we adapt Moser-Pöschel perturbation arguments to the discrete case. This is applied in Section 4.3 to the proof of nonperturbative genericity of Cantor spectrum. Finally, in Section 4.4 we prove that, in our situation, Cantor spectrum implies nonreducibility for a
-set of energies.
4.1 Reducibility at gap edges
In previous sections, we discussed the reducibility of a quasi-periodic Schrödinger cocycle
when
is a point eigenvalue of the dual operator
and
satisfies a nonresonance condition of the form ( 14 ), which was enough to prove the main result. The “resonant” values of
:
|
(20)
|
are particularly important for the description of the spectrum of these operators because the corresponding point eigenvalues lie at endpoints of spectral gaps. Let us prove the reducibility at these endpoints. What follows mimics the proof of the “Ten Martini Problem” given in [Pui04a] .
Bourgain & Jitomirskaya [BJ02a] also prove that, provided
and
is Diophantine
has pure-point spectrum with exponentially localized eigenfunctions if
is of the form ( 20 ), see Remark 8.2 after Theorem 7 in [BJ02a] .
Taking into account the symmetries of the operators, the study reduces to the four cases
. For the sake of simplicity we consider here
. As a direct consequence of Aubry duality and Bourgain-Jitomirskaya result for the dual operators
, the set of pure-point eigenvalues
is a dense subset of
and any energy
has a quasi-periodic Bloch wave with
as a Floquet exponent for the dual eigenvalue equation. Let
be one of these eigenvalues. The condition above means that there is an analytic map
such that
is a nonzero solution of
. Clearly, due to the symmetry of the eigenvalue equation for
, the function
can be chosen real analytic. In terms of the cocycle we have that the relation
holds for all
. Instead of Lemma 16 we now have the following (see [Pui04a] for the proof, which is a simple triangularization and averaging argument).
Lemma 17
Let
be a real analytic map and
be Diophantine. Assume that there is a nonzero real analytic map
such that the relation
holds for all
Then the quasi-periodic cocycle
is reducible to constant coefficients by means of a quasi-periodic transformation which is analytic in
. Moreover the Floquet matrix can be chosen to be of the form
for some
.
In the Almost Mathieu case Ince's argument [Inc44, Pui04a] shows that
.
Otherwise the dual model (which is also a Schrödinger operator) would have a point eigenvalue with two linearly independent eigenvectors in
, and this is a contradiction with the limit-point character of Schrödinger operators (or just the preservation of the Wronskian in this discrete case). The fact that
is important for the description of the spectrum, because if a Schrödinger cocycle is reducible to a Floquet matrix of the form ( 21 ) with
then the corresponding energy is at the endpoint of an open gap in the spectrum, as it will be seen in the next section.
For general potentials
, however, we cannot use Ince's argument and it may happen that some of these are collapsed. In fact, there are examples of quasi-periodic Schrödinger operators (with
small,
Diophantine) for which some
are zero [BPS03] or even do not display Cantor spectrum (see De Concini & Johnson [DCJ87] ).
Nevertheless, even if
can be zero, Moser & Pöschel [MP84] showed that, in this reducible setting, a closed gap can be opened by means of an arbitrarily small and generic real analytic perturbation of the potential. In the next section we give an adaption of their proof to the discrete case together with some extra properties which will be needed later.
4.2 Moser-Pöschel perturbation argument
In this section we prove the following adaption of Moser-Pöschel argument to the discrete case, which deals with cocycles which are perturbations of constant matrices of the form ( 21 ).
Proposition 18
Let
be real analytic,
Diophantine and
, for some
be a quasi-periodic Schrödinger cocycle. Assume that
is analytically reducible to the constant coefficients cocycle
with
for some
. Let
be real analytic and
real. If
is the real analytic reducing matrix and the conditions
|
(22)
|
or
|
(23)
|
are satisfied, then the quasi-periodic cocycle
has an exponential dichotomy provided
is small enough and
|
(24)
|
Moreover in the case
, the Lyapunov exponent of
and its rotation number,
satisfy
|
(25)
|
Remark 19
This type of perturbation arguments have been used in a variety of contexts, c.f. Moser & Pöschel [
Mos81,
MP84]
, Johnson [
Joh91]
, Nún͂ez [
Nn95]
, Broer, Puig & Simó [
BPS03]
and Puig & Simó [
PS04]
.
Proof: Since
is the reducing transformation of
to
, it also renders the perturbed cocycle
to
where
After one step of averaging this cocycle can be analytically conjugated to
where
denotes the average of a quasi-periodic function and
depends analytically on
and
in some open neighbourhoods of
and
. Moreover, a computation shows that
being
and
depending analytically on
and
. Let
whose determinant is
. Now let us distinguish between the cases
and
.
If ( 22 ) holds then the expression for the determinant becomes
so that it is negative if, in addition ( 24 ) holds. In this case, the matrix
which is well-defined, has determinant
and satisfies
where
Therefore the change of variables defined by
transforms the cocycle
|
(26)
|
into
|
(27)
|
where
which is
uniformly in
. Note that
so that if ( 24 ) holds and
is small enough the cocycle
has an exponential dichotomy and the Lyapunov exponent satisfies
see Coppel [Cop78] . To obtain the asymptotics of the rotation number, we can consider the transformation
defined for
. This, although complex, is a well-defined conjugation between
and
, which is now a complex rotation of angle
Therefore ( 26 ) is conjugated to ( 27 ), a perturbation of a complex rotation. Using the definition of the fibered rotation number given in Section 2.3 the result follows.
Let us now consider the situation when ( 23 ) holds. In this case, the matrix
becomes
Condition ( 23 ) is equivalent to the hyperbolicity of
whose determinant is
Therefore there is a change of variables
, independent of
and
which renders it to a diagonal form
with
and
as diagonal entries. This conjugation transforms the cocycle
into
where
Since,
the cocycle
has an exponential dichotomy when
is small enough (see, again Coppel [Cop78] ).
The perturbation argument in the previous proposition can applied to the reducible cocycles at endpoints of gaps as we do next.
Corollary 20
Let
,
and
as in Proposition 18 and assume that
.
Then
is at the endpoint of a noncollapsed spectral gap
of
(the right one if
and the left one if
). Moreover, the limits
|
(28)
|
exist and are different from zero.
Proof: Take
in Proposition 18 . Then, the cocycle
has an exponential dichotomy if
and
is small enough. This means that there is an open spectral gap besides
(to the left if
and to the right otherwise).
Moreover the asymptotics of formula ( 25 ) imply ( 28 ).
Finally we consider the variation of the rotation number in the case
in a more general setting which will be needed in the next section.
Proposition 21
Let
be continuous and
nonresonant. Assume that the Schrödinger cocycle
is reducible to the cocycle
with
a constant matrix. Then the map
is differentiable at
Proof: Let
be the angle of the rotation,
The cocycle
is conjugated to
where
and
. The cocycle
induces a lift
from
to itself of the form
where
is continuous and
-periodic in both
and
. Therefore,
as we wanted to show.
Remarks 22
-
1.
A computation shows that the derivative of the rotation number above is nonzero. In particular, when a Schrödinger cocycle is reducible to the identity, the corresponding energy lies at the endpoint of a collapsed gap.
-
2.
Similar results have been obtained when
are square integrable and
is measurable, compare with Moser [Mos81] and Deift & Simon [DS83a] .
4.3 Genericity of Cantor spectrum
In the previous section we have seen that if a Schrödinger cocycle is reducible to a matrix with trace
then the corresponding energy is at the endpoint of a spectral gap which is collapsed if the Floquet matrix is the identity. The next consequence of Proposition 18 is that when the Floquet matrix is the identity (a similar statement holds for minus the identity) one can “open up” the collapsed gap by means of a generic perturbation.
Corollary 23
Let
,
,
and
be as in Proposition 18 and assume that
. If
is a generic real analytic potential then for
small enough the spectrum
has an open spectral gap with ids
.
Proof: In Proposition 18 we proved that for a perturbation
satisfying ( 23 )
the cocycle
has an exponential dichotomy if
small enough.
This means that
lies in a spectral gap of
which, by continuity must satisfy that
for
small enough. Let us now show that if
is a generic potential, then for every
small enough there is a value of
such that
lies in a spectral gap of
with
Note that the condition ( 23 ) can be rewritten as
where
Let
be such that
(this determines
since
). Then the shifted perturbation
satisfies condition ( 23 ) unless
which is clearly a generic condition. Then, if
is small enough, the spectrum
has an open gap with
as we wanted to show.
Remark 24
As Moser & Pöschel show, when
it is always possible to choose the reducing transformation such that
and
so that
,
,
and the generic
must satisfy
Let us now summarize the situation. Using the two past sections we have seen that if
is a real analytic potential on
, with
and
is Diophantine, there is a countable dense subset of energies in the spectrum where the system is reducible to a Floquet matrix with trace
. These lie at endpoints of gaps. Although these can be collapsed, a generic and arbitrarily small perturbation opens them as Corollary 23 says. Since there is a countable number of gaps Theorem 4 follows.
4.4 Cantor spectrum implies nonreducibility
In [Eli92] it was seen that, for a generic real analytic Schrödinger cocycle (with Diophantine frequencies) besides the almost everywhere reducibility there was a set of zero measure of energies for which the cocycle was not reducible to constant coefficients. The proof relies on the kam procedure developed there, but the Cantor structure of the spectrum is seen to play a key role. In this section we prove irreducibility for a
-set of energies assuming only Cantor structure of the spectrum and Theorem 1 . This argument is reminiscent of some techniques in circle maps, see Arnol'd [Arn61] . We state here a slightly more general version than that of Theorem 5 . More applications will be given elsewhere.
Theorem 25
Let
. There is a constant
such that if
is real analytic with
,
is Diophantine and
is an open interval such that
is a nonvoid Cantor set, then there is a
-dense set of energies in
for which the corresponding Schrödinger cocycle is not reducible to constant coefficients by means of a continuous transformation.
Proof: Consider, for any
with
,
Now, for any
we can define
which is either a positive real number or
If
is reducible to constant coefficients then we have two situations.
Either the Floquet matrix
has trace
, in which case
(see Corollary 20 ) or
and then
(see Proposition 21 ). Due to the fact that
is Diophantine and the Cantor structure of the spectrum there is a dense set of endpoints of gaps,
, where the system is reducible to constant coefficients because of Eliasson's Theorem 11 .
We will show that the set where
is a
-dense subset of
.
Excluding the endpoints of gaps where there is reducibility to a Floquet matrix with trace
(which are at most countable) we will still have a
-dense subset of energies in
whose corresponding cocycle cannot be reducible to constant coefficients.
Let, for any
and
,
and
The sets
are open in
because of the continuity of the rotation number.
Moreover they are dense in
because they contain
, which is dense in
.
Therefore
is a
-dense subset of
. If we exclude the endpoints of open gaps the remaining energies, which still form a
-dense subset of
, cannot be reducible to constant coefficients by means of a continuous transformation. This proves 25 and also 5 .
Acknowledgements The author wishes to thank H. Eliasson, S. Jitomirskaya, C. Simó and B. Simon for many ideas which have been decisive in this paper. The paper arose in a stay at Caltech Mathematics Department and completed mostly in the Departament de Matemàtica Aplicada i Anàlisi at the Universitat de Barcelona as part of his PhD thesis. I am grateful to both institutions. This work has been supported by grants DGICYT BFM2003-09504-C02-01 (Spain) and CIRIT 2001 SGR-70 (Catalonia).
References
-
S. Aubry and G. André. Analyticity breaking and Anderson localization in incommensurate lattices. In Group theoretical methods in physics (Proc. Eighth Internat. Colloq., Kiryat Anavim, 1979), pages 133–164. Hilger, Bristol, 1980.
-
A. Avila and S. Jitomirskaya. The Ten Martini Problem. In preparation, 2005.
-
A. Avila and R. Krikorian. Reducibility or non-uniform hyperbolicity for quasiperiodic Schrödinger cocycles. Preprint, 2003.
-
V. I. Arnol
d. Small denominators. I. Mapping the circle onto itself. Izv. Akad. Nauk SSSR Ser. Mat., 25:21–86, 1961.
-
J. Avron and B. Simon. Almost periodic Schrödinger operators II. The integrated density of states. Duke Math. J., 50:369–391, 1983.
-
J. Bourgain and S. Jitomirskaya. Absolutely continuous spectrum for 1D quasiperiodic operators. Invent. Math., 148(3):453–463, 2002.
-
J. Bourgain and S. Jitomirskaya. Continuity of the Lyapunov exponent for quasiperiodic operators with analytic potential. J. Statist. Phys., 108(5-6):1203–1218, 2002. Dedicated to David Ruelle and Yasha Sinai on the occasion of their 65th birthdays.
-
J. Bourgain. On the spectrum of lattice Schrödinger operators with deterministic potential. II. J. Anal. Math., 88:221–254, 2002. Dedicated to the memory of Tom Wolff.
-
J. Bourgain. Green's function estimates for lattice Schrdinger operators and applications. Princeton University Press, Princeton, N. J., 2004. Annals of Mathematics Studies, to appear.
-
J. Bourgain. On quasi-periodic lattice Schrödinger operators. Discrete Contin. Dyn. Syst., 10(1-2):75–88, 2004.
-
H. W. Broer, J. Puig, and C. Simó. Resonance tongues and instability pockets in the quasi-periodic Hill-Schrödinger equation. Comm. Math. Phys, 241(2–3):467–503, 2003.
-
V. Chulaevsky and F. Delyon. Purely absolutely continuous spectrum for almost Mathieu operators. J. Statist. Phys., 55(5-6):1279–1284, 1989.
-
V. A. Chulaevsky. Almost periodic operators and related nonlinear integrable systems. Nonlinear Science: Theory and Applications. Manchester University Press, Manchester, 1989. With a foreword by Ya. G. Sinaĭ, Translated from the Russian.
-
W. A. Coppel. Dichotomies in stability theory. Springer-Verlag, Berlin, 1978. Lecture Notes in Mathematics, Vol. 629.
-
C. De Concini and R. A. Johnson. The algebraic-geometric AKNS potentials. Ergodic Theory Dynam. Systems, 7(1):1–24, 1987.
-
S. Datta, T. Jäger, G. Keller, and R. Ramaswamy. On the dynamics of the critical harper map. Nonlinearity, 17(6):2315–2323, 2004.
-
E. I. Dinaburg and Y.G. Sinai. The one-dimensional Schrödinger equation with quasi-periodic potential. Funkt. Anal. i. Priloz., 9:8–21, 1975.
-
P. Deift and B. Simon. Almost periodic Schrödinger operators III. The absolute continuous spectrum. Comm. Math. Phys., 90:389–341, 1983.
-
F. Delyon and B. Souillard. The rotation number for finite difference operators and its properties. Comm. Math. Phys., 89(3):415–426, 1983.
-
L. H. Eliasson. Floquet solutions for the one-dimensional quasi-periodic Schrödinger equation. Comm. Math. Phys., 146:447–482, 1992.
-
A. Y. Gordon, S. Jitomirskaya, Y. Last, and B. Simon. Duality and singular continuous spectrum in the almost Mathieu equation. Acta Math., 178(2):169–183, 1997.
-
M. R. Herman. Une méthode pour minorer les exposants de Lyapunov et quelques exemples montrant le caractère local d'un théorème d'Arnold et de Moser sur le tore de dimension 2. Comment. Math. Helvetici, 58(3), 1983.
-
E. L. Ince. Ordinary Differential Equations. Dover Publications, New York, 1944.
-
C. Janot. Quasicrystals : a primer. Clarendon Press, 1992.
-
S. Ya. Jitomirskaya. Almost everything about the almost Mathieu operator. II. In XIth International Congress of Mathematical Physics (Paris, 1994), pages 373–382. Internat. Press, Cambridge, MA, 1995.
-
S. Jitomirskaya. Metal-insulator transition for the almost Mathieu operator. Ann. of Math. (2), 150(3):1159–1175, 1999.
-
S. Jitomirskaya. Nonperturbative localization. In Proceedings of the International Congress of Mathematicians, Vol. III (Beijing, 2002), pages 445–455, Beijing, 2002. Higher Ed. Press.
-
R. Johnson and J. Moser. The rotation number for almost periodic potentials. Comm. Math. Phys., 84:403–438, 1982.
-
R. Johnson. Analyticity of spectral subbundles. J. Diff. Eq., 35:366–387, 1980.
-
R. Johnson. The recurrent Hill's equation. J. Diff. Eq., 46:165–193, 1982.
-
R. Johnson. A review of recent work on almost periodic differential and difference operators. Acta Appl. Math., 1(3):241–261, 1983.
-
R. A. Johnson. On the Sato-Segal-Wilson solutions of the K-dV equation. Pacific J. Math., 132(2):343–355, 1988.
-
R. Johnson. Cantor spectrum for the quasi-periodic Schrödinger equation. J. Diff. Eq., 91:88–110, 1991.
-
R. Krikorian. Reducibility, differentiable rigidity and Lyapunov exponents for quasi-periodic cocycles on
. Preprint.
-
J. A. Ketoja and I. I. Satija. Harper equation, the dissipative standard map and strange nonchaotic attractors: relationship between an eigenvalue problem and iterated maps. Phys. D, 109(1-2):70–80, 1997. Physics and dynamics between chaos, order, and noise (Berlin, 1996).
-
Y. Last. Almost everything about the almost Mathieu operator. I. In XIth International Congress of Mathematical Physics (Paris, 1994), pages 366–372. Internat. Press, Cambridge, MA, 1995.
-
J. Moser. An example of Schrödinger equation with almost periodic potential and nowhere dense spectrum. Comment. Math. Helvetici, 56:198–224, 1981.
-
J. Moser and J. Pöschel. An extension of a result by Dinaburg and Sinai on quasi-periodic potentials. Comment. Math. Helvetici, 59:39–85, 1984.
-
C. Nún͂ez. Extension of a Moser-Pöschel theorem for the Schrödinger equation with ergodic potential. In XIV CEDYA/IV Congress of Applied Mathematics (Spanish)(Vic, 1995), page 10 pp. (electronic). Univ. Barcelona, Barcelona, 1995.
-
D. Osadchy and J. Avron. Hofstadter butterfly as quantum phase diagram. J. Math. Phys., 42(12):5665–5671, 2001.
-
J. Puig and C. Simó. Analytic families of reducible linear quasi-periodic equations. Preprint, 2004.
-
J. Puig. Cantor spectrum for the Almost Mathieu operator. Comm. Math. Phys, 244(2):297 – 309, 2004.
-
J. Puig. Reducibility of Quasi-Periodic Skew-Products and the Spectrum of Schrödinger Operators (PhD Thesis). University of Barcelona, 2004.
-
B. Simon. Almost periodic Schrödinger operators: a review. Adv. in Appl. Math., 3(4):463–490, 1982.
-
B. Simon. Schrödinger operators in the twentieth century. J. Math. Phys., 41(6):3523–3555, 2000.
-
Ya. G. Sinaĭ. Structure of the spectrum of a Schrödinger difference operator with almost periodic potential near the left boundary. Funktsional. Anal. i Prilozhen., 19(1):42–48, 96, 1985.
-
E. Sorets and T. Spencer. Positive Lyapunov exponents for Schrödinger operators with quasi-periodic potentials. Comm. Math. Phys., 142(3):543–566, 1991.