1. Introduction
Hypercyclic (that is, topologically transitive) and chaotic operators on separable Banach spaces have been studied for more than twenty years (the reader is referred to the work in [
1,
2] for good sources on linear dynamics). On the other hand, Toeplitz operators were introduced by Otto Toeplitz in [
3]. They are among the most studied families of operators on the Hardy–Hilbert space. On this space, the matrices of Toeplitz operators (with respect to the canonical basis) have constant diagonals.
A Toeplitz operator
on the Hardy–Hilbert space
with symbol
is defined by
,
, where
is the multiplication operator by
and
is the Riesz projection. Here, as usual,
are the open unit disc and its boundary, the unit circle, respectively. Actually,
so that it is naturally identified with the Hilbert sequence space
. The reader is referred to the work in [
4] for the basic theory of Toeplitz operators, and to the work in [
5,
6] for a detailed study of Hardy spaces.
It is known that analytic Toeplitz operators, that is, operators whose symbol is in
(the space of all the functions that are analytic and bounded on the open unit disk), cannot be hypercyclic, as their adjoints always have eigenvalues. However, Toeplitz operators with anti-analytic symbols, i.e., such that the symbol
satisfies
, provide many examples of hypercyclic operators, and they are the most studied Toeplitz operators in the topic of chaotic dynamics. Godefroy and Shapiro [
7] showed that a Toeplitz operator
with anti-analytic symbol
is chaotic if, and only if,
for
, a result that was extended by De Laubenfels and Emamirad in [
8] for
,
and
. Notice that, with the above identification of
with
, the anti-analytic Toeplitz operator
with
can be formally represented by
where
B is the backward shift
, so that an anti-analytic Toeplitz operator can be viewed as an upper triangular infinite matrix with constant diagonals. With this identification, Bourdon and Shapiro [
9] studied the dynamics of anti-analytic Toeplitz operators in the Bergman space, and Martínez [
10] in more general sequence spaces. The first example of an anti-analytic and hypercyclic Toeplitz operator was
for
[
11], which is represented by
. A special mention should be done to the exponential of the backward shift
that was shown to be hypercyclic in arbitrary “small” sequence spaces in [
12,
13].
Baranov and Lishanskii [
14], inspired by the work of Shkarin [
15], studied hypercyclicity of Toeplitz operators on
with symbols of the form
, where
p is a polynomial and
. They showed necessary conditions and sufficient conditions for hypercyclicity which almost coincide in the case the degree of
p is one. They characterized hypercyclicity in the tridiagonal case (i.e., when
p and
q have degree one) by refining a result of Shkarin [
15]. Based on these results, the chaotic behavior of certain non-local operators was studied in [
16]. Recently, some new classes of hypercyclic Toeplitz operators were also found in [
17], as a continuation of the work in [
14].
In this paper, a characterization of chaos in the tridiagonal case formulated in terms of the three symbol coefficients is obtained. Sufficient conditions for chaos in more general cases, also explicit on the symbol coefficients, are shown, and they easily provide us with examples of chaotic Toeplitz operators. The main contribution of the present work in comparison to the work in [
14,
15] is to offer conditions on the symbol coefficients of a Toeplitz operator for chaos, which are much easier to check than the previous ones. Moreover, the characterization of the tridiagonal case gives a full picture of the chaotic behavior in terms of the three coefficients. Finally, a rich variety of chaotic properties, in the topological and in the measure theoretical sense are provided, which must be compared with previous works on the dynamics of Toeplitz operators, dealing with hypercyclicity and/or Devaney chaos.
2. Preliminaries and Notation
Some definitions about hypercyclicity and chaos need to be recalled. From now on, unless otherwise specified, X will be assumed to be an infinite dimensional separable Banach space and a continuous and linear operator.
An operator
is called
hypercyclic if there is some
whose orbit under
T is dense in
X. In such a case,
x is called a hypercyclic vector for
T. The operator
T is said to be
Devaney chaotic if it is hypercyclic and admits a dense set of periodic points. Actually the original definition of Devaney [
18] also included as an ingredient the sensitive dependence on initial conditions, but it was shown to be redundant (see, e.g., in [
19]).
The first notion of chaos coined in the mathematical literature appeared in the article of Li and Yorke [
20]. Let
be a metric space. A continuous map
is called
Li–Yorke chaotic if there exists an uncountable subset
such that for every pair
of distinct points one has
In this case,
is a
scrambled set and
a Li–Yorke pair.
A vector is said to be irregular for T if and .
A stronger notion of chaos was introduced by Schweizer and Smital [
21]: Let
be a metric space and let
be a continuous map. For any pair
and every
, the
distributional function is defined by
where card(
A) denotes the cardinality of the set
A. Define
The map f is called distributionally chaotic if there exist an uncountable subset and such that for every and each pair of distinct points , it happens that and . The set is a distributionally -scrambled set and the pair a distributionally chaotic pair. Moreover, f exhibits dense distributional chaos if the set may be chosen to be dense.
A subset
A of
is said to have
positive lower density if
Inspired by Birkhoff ergodic theorem, Bayart and Grivaux [
22,
23] considered a concept stronger than hypercyclicity: An operator
T on
X is said to be
frequently hypercyclic provided there exists a vector
x such that for every nonempty open subset
U of
X, the set of integers
n such that
belongs to
U has positive lower density. In this case,
x is called a frequently hypercyclic vector for
T.
Bowen [
24] introduced a very strong dynamical notion for maps on compact spaces that occurs when one can approximate distinct pieces of orbits by a single periodic orbit with a certain uniformity: A continuous map
on a compact metric space
has the
specification property (SP) if for any
there is a positive integer
such that for any integer
, any set
and any integers
satisfying
for
, there is a point
such that, for each positive integer
and any integer
i with
, the following conditions hold:
Bartoll et al. [
25] generalized this concept for operators: An operator
T on
X has the
operator specification property (OSP) if there exists an increasing sequence
of
T-invariant sets with
and
such that for each
the map
has the SP.
Some measure-theoretic concepts in dynamics ought to be recalled too. Let
be a probability space, where
X is a topological space and
denotes the
algebra of Borel subsets of
X. A Borel probability measure
is said to have
full support if
for each non-empty open set
. A measurable map
is called a
measure preserving transformation (or
is
T-invariant) if
for all
. The measure
is said to be
strongly mixing with respect to
T if
A recent work that, in particular, connects the OSP with the existence of strongly mixing measures is [
26].
A sufficient condition for frequent hypercyclicity was given by Bayart and Grivaux [
23], later refined by Bonilla and Grosse-Erdmann [
27] by replacing absolute convergence of series by unconditional convergence. This is what is known today as the Frequent Hypercyclicity Criterion. A series
in
X converges
unconditionally if it converges and, for any 0-neighborhood
U in
X, there exists some
such that
for every finite set
.
It was shown in [
28] that the Frequent Hypercyclicity Criterion implies the existence of mixing measures (see also in [
29] for more general results).
Theorem 1 ([
28]).
Let T be an operator on X. If there is a dense subset of X and a sequence of maps such that for each .
- (i)
converges unconditionally,
- (ii)
converges unconditionally, and
- (iii)
andif,
then there is a invariant strongly mixing Borel probability measure μ on X with full support.
A powerful tool in linear dynamics to obtain chaotic properties for operators is to have a wide source of eigenvectors associated to suitable eigenvalues. Certainly, the basis is found in the so-called Godefroy–Shapiro Criterion [
7]. Other sufficient conditions for hypercyclicity can be found in [
1,
2,
30,
31].
Theorem 2 (Godefroy–Shapiro Criterion).
Let T be an operator on X. Suppose that the subspacesandare dense in X. Then, T is hypercyclic. If, moreover, the subspaceis dense in X, then T is Devaney chaotic. Suitable eigenvector fields will be very useful to obtain all the chaotic properties considered here, an idea that follows the work initiated by Bayart and Grivaux [
22]. Given an operator
on a complex Banach space
X, a collection of functions
,
, is called a
spanning eigenvector field associated to unimodular eigenvalues if
for any
,
, and
A map
defined on a non-empty open set
is said to be
weakly holomorphic on
U if
is holomorphic for any
.
The following result is essentially well known, but its proof is included for the sake of completeness.
Theorem 3. (Eigenfield Criterion) Given an operator on a complex Banach space X, if is a connected nonempty open set that intersects , , , is a collection of weakly holomorphic maps such that for any , , andthen - (i)
T is Devaney chaotic,
- (ii)
there exists a spanning eigenvector field associated to unimodular eigenvalues , , such that for any , , where is a non-trivial arc, and
- (iii)
T satisfies the Frequent Hypercyclicity Criterion.
Proof. To prove (i), consider
,
, and
It will suffice to show that, given
, and for any
, the equality
for every
implies
. Actually, as the holomorphic maps
annihilate on
for
, which are sets with accumulating points in
U, and
U is connected, then
for every
. The assumptions imply that
, and (i) is shown.
For (ii), let
be an open arc, and let
be a non-trivial closed sub-arc. It is possible to extend
from
I to
as a
function
such that
for all
for each
.
,
, is a spanning eigenvector field associated to unimodular eigenvalues as
which is dense in
X because
I has accumulating points in
U.
The fact that
T satisfies the Frequent Hypercyclicity Criterion is a consequence of, e.g., Remark 9.10 and Theorem 9.22 in [
2]. □
3. Tridiagonal Toeplitz Operators
The main purpose in this section is to reformulate the characterization of hypercyclic tridiagonal operators given by Shkarin [
15] and Baranov and Lishanskii [
14] to offer conditions expressed in terms of the three coefficients of the symbol.
More precisely, equivalent and sufficient conditions, expressed on the coefficients of the symbol, are provided in order to guarantee that a tridiagonal Toeplitz operator has a chaotic behavior. Tridiagonal Toeplitz operators were studied in [
32] (see also in [
33]) as generators of chaotic semigroups associated to birth-and-death processes.
Let
be an operator with symbol
, where
. If
is zero then
is an anti-analytic operator, and there are conditions for these operators to be hypercyclic [
2]. If
is zero then
is a analytic operator, and these operators are not hypercyclic, as was mentioned before. The goal of this section is to have conditions such that these operators are chaotic when
and
are not zero.
The previous eigenvalue criteria will be a key tool. To do this, one has to solve the equation
. It is known that, for
,
and
. Then
, which implies that
is equivalent to
Therefore,
If
one would have that
as
f is an analytic function on
this would imply that
is identically zero. Therefore,
and without loss of generality it is assumed that
.
As it is wanted that
belongs to
, then the polynomial
needs to have roots
and
with
and
. This is equivalent to the roots of the polynomial
being in
.
To find conditions on the coefficients such that both roots of
are in
, the following test (see, e.g., in [
34] for a proof of it) will play a key role.
Jury test: Consider the family of quadratic equations for where and are parameters. For a fixed r, let denote the set of all complex w such that the absolute value of each root is less then 1. If , then Actually, dealing with the case when demands to generalize the Jury test.
Lemma 1. (Generalized Jury Test).
The roots of
, with parameters
,
and
, belong to
if and only if
and
Proof. By applying the Jury test to , and taking into account that if and only if for , the result is obtained. □
The Generalized Jury test will be applied to the polynomial
Observe that, if its roots belong to
, then
. Therefore, from now on it is assumed that
.
Now, consider the following ellipse:
and its interior
Let
i.e., the interior of the outer parallel curve at distance one of the ellipse
E. If
, let
F be the inner parallel curve at distance one of
E. There is a connected component in the interior of
F that contains 0, and its closure is set as
.
Figure 1 illustrates an example of how
is defined.
All the conditions are now set to establish a key result that will allow us to obtain the desired characterization in terms of and .
Lemma 2. Let with . Set , , with , . Then, the following conditions are equivalent:
- (A)
There existssuch thathas its roots in.
- (B)
satisfies one of the following cases:
- 1.
Ifthen.
- 2.
Ifthen.
- 3.
Ifthen.
Proof. By applying the Generalized Jury test to the polynomial
it is known that its roots belong to
if, and only if, the following inequality is satisfied:
which is equivalent to
Let . The above inequality holds for some if, and only if, one can find such that . This is in turn equivalent to the existence of such that and . Indeed, one implication is obvious since is open, and the other one follows from an easy connectedness argument: Consider the function defined by . As is a connected set, the set is an interval in that has points greater than 1, and smaller than 1, so 1 is inside.
The equivalence with condition (B) will be shown now:
- 1.
.
As in this case , by the definition of the set it is clear that exist such that and .
- 2.
.
As in this case , the same reasoning as above holds for all , except for .
- 3.
.
As , by the definition of , there exists such that . Now, as , the definition of allows us to find a point such that .
□
Therefore, far it has been characterized the existence of
such that the
-eigenvector of
belongs to
. Actually, it will be shown that this defines an eigenvector field that satisfies the conditions of Theorem 3.
Theorem 4. Let be an open subset with non empty intersection with and suppose that for any , whereThen, the map , , is weakly holomorphic and Proof. For
open, set
. Let
g be a function in
. It will be shown that, if
for all
in
B, then
, which is equivalent to the fact that
is dense in
.
Consider
defined by
,
,
Suppose that
for all
. Thus, all the derivatives of
H vanish at certain
. That is,
,
and
One has that is an analytic function in an open disc . As it is known that , then is univalent on a neighborhood of , which is assumed to be U, and there exists , where is a simply connected open set. Then, , with , is an isomorphism. It is known that is dense in , thus is dense in , and then dense in . That is, is dense in which gives as , concluding that . □
The following theorem summarizes the previous results.
Theorem 5. Let be a Toeplitz operator with symbol the function , where , , with , and belong to . Set , and let , be the sets defined before Lemma 2. Then, the following affirmations are equivalent:
- (C1)
and satisfies one of the following conditions:
- (a)
If then .
- (b)
If then .
- (c)
If then .
- (C2)
T satisfies the Godefroy–Shapiro Criterion.
- (C3)
T satisfies the Eigenfield Criterion.
All the necessary ingredients are now given in order to establish the main result of this section.
Theorem 6. Let be a Toeplitz operator with symbol the function , where , , and belong to . Then, the following affirmations are equivalent:
- (1)
satisfies the Godefroy–Shapiro Criterion.
- (2)
, and .
- (3)
The coefficients , and satisfy the conditions (C1) of Theorem 5.
- (4)
satisfies the Eigenfield Criterion.
- (5)
is a distributionally chaotic operator.
- (6)
is a Li–Yorke chaotic operator.
- (7)
is a Devaney chaotic operator.
- (8)
admits an invariant strongly mixing Borel probability measure μ on with full support.
- (9)
has the .
- (10)
is a frequently hypercyclic operator.
- (11)
is a hypercyclic operator.
Proof. The equivalence of (1) and (2) follows from the works in [
14,
15]. The equivalence of (1), (3), and (4) is given in Theorem 5. Equivalent conditions (1)–(4) imply that
satisfies the Frequent Hypercyclicity Criterion by Theorems 3 and 4, so any of the remaining conditions by [
35] ((5) and (6)), [
7] ((7) and (11)), [
28] (8), [
25] (9), and [
22] (10). Furthermore, condition (6) is the weakest one among (5)–(11), and one just needs to show that (6) implies (3) in order to conclude all the equivalences. Actually, a direct computation shows
, where
. As
, then
is hyponormal if
, and by [
36]
does not have Li–Yorke pairs. If
then
and
is a contraction. Therefore
does not have Li–Yorke pairs. If
, by an argument from Proposition 4.1 in [
14] one has that
for all
x, then
does not have irregular vectors. In any case
can not be Li–Yorke chaotic. □
Although the conditions of Theorem 5 offer a complete characterization in terms of the 3 coefficients , and , it is useful to have a more “handy” sufficient condition which can be expressed by a single inequality. This is the purpose of the final result in this section. To simplify, once the equivalences of Theorem 6 are known, from now on the term “chaotic” will refer to any of properties (5)–(11) given there.
Corollary 1. Let . If it satisfies , then the Toeplitz operator with symbol the function is a chaotic operator.
Proof. Recall that , where is the interior of a ellipse with semiaxis and . The first observation is that, as , then for all .
In case that , the conditions of Theorem 5 are satisfied.
If , then implies that , and the result follows.
Finally, if , then the fact that yields that for all , concluding the result. □
4. Toeplitz Operators with General Analytic Part
In this section, for more general
, some conditions are given for
to satisfy the Eigenvalue Criterion. Precisely,
will be assumed to be a Toeplitz operator with symbol
that is, with general analytic part
and a polynomial
q in the anti-analytic part. It will be imposed that
, that is,
is analytic in
.
If
,
, then
is an anti-analytic Toeplitz operator, and conditions for the hypercyclicity of such operators are well known [
7].
First, one has to solve the equation
, where
, with
. It is known that
By definition,
where
for
.
By the above remarks,
The equality
yields
and therefore
As
one has that
Half of the following result has been shown.
Lemma 3. Let be a Toeplitz operator with symbol , where and . If is an eigenvalue of , then the solutions of the equation form an m-dimensional vector space and are of the formwith a polynomial of degree at most . Proof. If
satisfies the equation
then
f has the form
Observe that the numerator of
f is a polynomial of degree
and it has the form
Therefore, as , the numerator of is nonzero if at least one of the coefficients , for , is nonzero. Clearly, every polynomial of degree at most can be obtained by choosing the coefficients , for , appropriately. Therefore, the space of solutions is m-dimensional. □
The following result is implicit in the proof of Statement 2 of Theorem 1.2 in [
14].
Proposition 1. Let , letwith and . Let U be a nonempty open set in such that , and let with accumulation points in U. If for any the equation has exactly m different solutions in , then is dense in , where , for . This section concludes by putting together the above results to obtain the next result.
Theorem 7. Let be a Toeplitz operator with symbol , with and . Assume that that the following conditions are satisfied:
- 1.
- 2.
for any the equation has exactly m solutions in .
Then, is chaotic.
Proof. For
, define
One needs that the roots of the denominator of
are not in
in order to have
and, as shown before,
. As it is wanted this to happen for certain
, where
is an open set that intersects
, select
(
if
) and, by the first hypothesis,
for any
, as
. For a sufficiently small neighborhood
U of
one gets
for all
. In addition, by construction,
is a
-eigenvector of
,
,
.
Observe that
. Otherwise,
with
, that implies
which is a contradiction. W.l.o.g. it is assumed that the neighborhood
U of
is small enough so that
U does not intersect
.
Let , , and for some . It is clear that , , has accumulation points in U.
By the second hypothesis and Proposition 1 applied to U and , , the sets , , are dense in , so , , is an eigenfield satisfying the hypothesis of Theorem 3, and the result is concluded. □
Remark 1. In the tridiagonal case, that is when , observe that the condition of Corollary 1 coincides with the hypothesis of Theorem 7. Actually, the firs part is obvious, and the second hypothesis is a consequence as for with yields , which implies , a contradiction.