Karl G. Grosse-Erdmann, Alfred Peris Manguillot Linear Chaos Universitext 2011
Karl G. Grosse-Erdmann, Alfred Peris Manguillot Linear Chaos Universitext 2011
Universitext
Series Editors:
Sheldon Axler
San Francisco State University
Vincenzo Capasso
Università degli Studi di Milano
Carles Casacuberta
Universitat de Barcelona
Angus J. MacIntyre
Queen Mary, University of London
Kenneth Ribet
University of California, Berkeley
Claude Sabbah
CNRS, École Polytechnique
Endre Süli
University of Oxford
Wojbor A. Woyczynski
Case Western Reserve University
Thus as research topics trickle down into graduate-level teaching, first textbooks
written for new, cutting-edge courses may make their way into Universitext.
Linear Chaos
Karl-G. Grosse-Erdmann Alfred Peris Manguillot
Institut de Mathématique Institut Universitari de Matemàtica
Université de Mons Pura i Aplicada
Le Pentagone, 7000 Mons Universitat Politècnica de València
Belgium Edifici 7A, 46022 València
[email protected] Spain
[email protected]
v
vi Preface
chaotic operators, and the Bourdon–Feldman theorem that any orbit that is
somewhere dense is (everywhere) dense.
It seems fair to say that while research in linear dynamics is still expanding
in both depth and breadth, the foundations have reached a certain stage of
maturity. At the same time the basic ideas as well as the applications of the
field have a broad appeal also to nonspecialists.
It is therefore our aim to make the theory of hypercyclic operators and
linear chaos accessible to a wider audience. The book is aimed at advanced
undergraduate or beginning graduate students, both as a basis for a lecture
course and for self-study. We have strived at a self-contained exposition. Each
chapter contains a large number of exercises and ends with a section that gives
references and directs the reader to further literature.
We have tried to keep the necessary prerequisites for reading this book
to a minimum. Since the concept of a hypercyclic operator requires both a
topological and a linear structure, the reader is supposed to be familiar with
metric spaces (up to the Baire category theorem) and with the basic theory of
Hilbert and Banach spaces, as it is often presented in advanced undergraduate
courses on analysis. Moreover, since many examples in the theory are given
by operators on spaces of holomorphic functions the reader is also expected
to have had an introductory course on complex analysis. Additional, more
advanced tools that are only needed occasionally will be provided in the two
appendices.
The book is divided into two parts. Part I presents an introduction to the
dynamics of linear operators. Its chapters form a unity and are best studied
in that order. In contrast, Part II covers selected topics from linear dynamics.
Its chapters are largely independent so that they can be read in an arbitrary
order. An occasional cross reference should pose no problem.
More specifically, Chapter 1 introduces the reader to the fundamental con-
cepts of the theory of (not necessarily linear) dynamical systems. Its high-
lights are the Birkhoff transitivity theorem, which is of fundamental impor-
tance for all that follows, and a close study of the various concepts of maps
with complicated behaviour, including chaotic maps. In Chapter 2, the no-
tions and results from the first chapter are revisited in the context of linear-
ity. Among other things it is proved that the operators of Birkhoff, MacLane
and Rolewicz are chaotic, and that linear dynamics can be as complicated as
nonlinear dynamics. We begin the chapter with an introduction to a straight-
forward generalization of Banach spaces, the so-called Fréchet spaces; they
provide the setting for some important chaotic operators. Chapter 3 presents
several criteria for hypercyclicity and chaos, in increasing order of sophisti-
cation. It culminates in the Hypercyclicity Criterion, which is discussed in
detail. In Chapter 4, some important classes of hypercyclic and chaotic oper-
ators are described: weighted shift operators on sequence spaces, differential
and composition operators on spaces of holomorphic functions, and adjoint
multiplication operators. In addition to the shift operators, which are studied
throughout the book, the reader may want to concentrate on one or two ad-
Preface vii
1 Topological dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.1 Dynamical systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Topologically transitive maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.3 Chaos . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.4 Mixing maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.5 Weakly mixing maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.6 Universality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
Sources and comments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
ix
x Contents
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 379
Part I
Introduction to linear dynamics
Chapter 1
Topological dynamics
Nn+1 = T (Nn ), n = 0, 1, 2, . . . ,
Nn = (T ◦ . . . ◦ T )(N0 ), n = 1, 2, . . .
xn+1 = T (xn ), n = 0, 1, 2, . . . .
Tn = T ◦ ... ◦ T (n times)
with
T 0 = I,
the identity on X.
Definition 1.2. Let T : X → X be a dynamical system. For x ∈ X we call
orb(x, T ) = {x, T x, T 2 x, . . .}
Nn+1 = (1 + γ)Nn ,
T : R+ → R+ , T x = (1 + γ)x.
Thus, the orbit tends to 0, x and ∞ for −1 < γ < 0, γ = 0 and γ > 0,
respectively.
1.1 Dynamical systems 5
Mn+1 = μMn (1 − Mn ), n ≥ 0.
Fig. 1.1 The logistic map L3 Fig. 1.2 The tent map T
For example, the property of having dense range is clearly preserved under
quasiconjugacy, while the property of being surjective is preserved under
conjugacy but not under quasiconjugacy.
One way of defining a new dynamical system from a given dynamical system
T is by restricting it to a subset. However, one has to ensure that T maps
this subset into itself.
Example 1.9. The interval [0, 1] is invariant under the logistic map Lμ for
0 ≤ μ ≤ 4.
and it has nonempty interior since it contains U . By (ii), B must have empty
interior, which implies that A is dense.
(iii)=⇒(ii). Suppose that X = A ∪ B, A ∩ B = ∅ and T (A) ⊂ A. Then
int(A) and int(B) are open sets with T n (int(A)) ∩ int(B) ⊂ A ∩ B = ∅ for all
n ≥ 0. By (iii) this can only be the case if either A or B has empty interior.
We clearly have that (iii)⇐⇒(iv). For (iii)⇐⇒(v) one need only note that
T n (U ) ∩ V = ∅ is equivalent to U ∩ T −n (V ) = ∅.
Example 1.12. (a) The tent map is topologically transitive. To see this, note
that T n is the piecewise linear map with T n ( 22kn ) = 0, k = 0, 1, . . . , 2n−1 ,
and T n ( 2k−1
2n ) = 1, k = 1, . . . , 2
n−1
; see Figure 1.4. Thus, let U ⊂ [0, 1] be
nonempty and open. Then U contains some interval J := [ 2mn , m+1 2n
]. But
since [0, 1] = T n (J) ⊂ T n (U ), T n (U ) in fact meets every nonempty set V .
(b) The doubling map on the circle, T : T → T, z → z 2 , is also topolog-
ically transitive. In fact, every nonempty open set U ⊂ T contains a closed
arc of angle 2π2n , for some n ≥ 1. Since the map T doubles angles, we have
1.2 Topologically transitive maps 9
The equivalence of conditions (iv) and (v) in Proposition 1.10 implies the
following.
Proof. (a) This follows easily from the fact that orb(x, T )\{x, T x, . . . , T n−1 x}
is contained in orb(T n x, T ) and that, in every metric space without isolated
points, a dense set remains dense even after removing finitely many points.
(b) Suppose that x ∈ X has dense orbit under T . Let U and V be nonempty
open sets in X. Then there is some n ≥ 0 such that T n x ∈ U . By (a), also
T n x has dense orbit, so that there is some m ≥ n such that T m x ∈ V . This
implies that T m−n (U ) ∩ V = ∅.
Proof. By the previous proposition, (ii) implies (i). For the converse, let T
be topologically transitive, and let D(T ) denote the set of points in X that
have dense orbit under T . Since X has a countable dense set {yj ; j ≥ 1},
the open balls of radius m1
around the yj , m, j ≥ 1, form a countable base
(Uk )k≥1 of the topology of X. Hence, x belongs to D(T ) if and only if, for
every k ≥ 1, there is some n ≥ 0 such that T n x ∈ Uk . In other words,
∞
∞
D(T ) = T −n (Uk ).
k=1 n=0
By continuity of T and Proposition 1.10, each set ∞
n=0 T
−n
(Uk ), k ≥ 0, is
open and dense. The Baire category theorem then implies that D(T ) is a
dense Gδ -set, and hence nonempty.
We note that the absence of isolated points was not needed for the proof
that (i) implies (ii).
Example 1.17. It follows from Example 1.12 that the tent map and the dou-
bling map have dense orbits. Irrational rotations (that is, with α ∈
/ πQ) have
the stronger property that each of their orbits is dense; see Exercise 1.2.9.
Example 1.18. Let X be the set of all points on the unit circle that are 2n th
roots of unity, for some n ≥ 1. By Example 1.12(b) the doubling map, re-
stricted to X, is topologically transitive, but clearly has no dense orbits.
Example 1.20. By Examples 1.6(c) and 1.17, the logistic map L4 on [0, 1] has
a dense orbit.
1.3 Chaos
What is chaos? Even when we restrict the meaning of this word to determin-
istic chaos, that is, chaotic behaviour of a dynamical system, mathematicians
have come up with different answers to this question. We will follow here the
definition that was suggested by Devaney in 1986. It has three ingredients,
which we discuss in turn.
The first ingredient tries to capture the idea of the so-called butterfly
effect: small changes in the initial state may lead, after some time, to large
discrepancies in the orbit. In order to be able to perturb points we consider
only spaces without isolated points.
Definition 1.21. Let (X, d) be a metric space without isolated points. Then
a dynamical system T : X → X is said to have sensitive dependence on
initial conditions if there exists some δ > 0 such that, for every x ∈ X and
ε > 0, there exists some y ∈ X with d(x, y) < ε such that, for some n ≥ 0,
d(T n x, T n y) > δ. The number δ is called a sensitivity constant for T .
We stress that the definition involves the metric of the space. In the fol-
lowing examples we will always work with the usual metric.
12 1 Topological dynamics
Example 1.22. (a) Using our knowledge of the iterates of the tent map (see
Example 1.12), we can easily show that it has sensitive dependence on initial
conditions with sensitivity constant 1/4, say. Indeed, if x ∈ [0, 1] and ε > 0
then there is some n ≥ 0 such that the open ball of radius ε around x contains
points y1 and y2 with T n y1 = 0 and T n y2 = 1; thus |T n x − T n yj | ≥ 1/2 for
some j ∈ {1, 2}.
(b) A similar argument, based on the fact that the doubling map doubles
angles, shows that it has sensitive dependence on initial conditions.
(c) No circle rotation has sensitive dependence on initial conditions because
we clearly have that |T n z1 − T n z2 | = |z1 − z2 | for any z1 , z2 ∈ T.
The second ingredient of chaos demands that the system is irreducible in
the sense that the map T connects any nontrivial parts of the space. We saw
in Section 1.2 that this idea is well captured by the notion of topological
transitivity of the system.
The third ingredient demands that the system has many orbits with a
regular behaviour; more precisely, there should be a dense set of points with
periodic orbit.
Definition 1.23. Let T : X → X be a dynamical system.
(a) A point x ∈ X is called a fixed point of T if T x = x.
(b) A point x ∈ X is called a periodic point of T if there is some n ≥ 1
such that T n x = x. The least such number n is called the period of x. The
set of periodic points is denoted by Per(T ).
A point is periodic if and only if it is a fixed point of some iterate T n ,
n ≥ 1. Thus, for real functions T , one easily detects them by searching for
the points where the graphs of T n and the identity function meet.
Example 1.24. (a) Considering the iterates of the tent map (see Example
1.12), we find that in every interval [ 2mn , m+1
2n ] there is a periodic point of
period n. Thus, the tent map has a dense set of periodic points.
(b) The periodic points of the doubling map on the circle are exactly the
(2n − 1)st roots of unity, n ≥ 1, so that also the doubling map has a dense
set of periodic points.
(c) For any rational rotation T there is some N ≥ 1 such that T N = I, so
that every point is periodic. In contrast, irrational rotations have no periodic
points at all.
Proposition 1.25. The property of having a dense set of periodic points is
preserved under quasiconjugacy.
Proof. Let T : X → X be quasiconjugate to S : Y → Y via φ : Y → X,
and let U ⊂ X be a nonempty open set. Then φ−1 (U ), being also open and
nonempty, contains a point y with S n y = y for some n ≥ 1. Hence φ(y) ∈ U
and T n φ(y) = φ(S n y) = φ(y).
Summarizing, we are led to Devaney’s definition of chaos.
1.3 Chaos 13
Example 1.27. By Examples 1.12, 1.22 and 1.24, the tent map and the dou-
bling map are chaotic, but no circle rotation is chaotic.
It then follows from the triangle inequality that, for any x ∈ X, either for
j = 1 or for j = 2 we have that d(x, T n pj ) ≥ η for all n ∈ N0 .
14 1 Topological dynamics
Let q have period N . As we have seen above there is also a periodic point p
such that
d(x, T n p) ≥ η = 4δ for n ∈ N0 . (1.2)
Since T is continuous there is some neighbourhood V of p such that
Definition 1.33. The dynamical system (Σ2 , σ) is called the shift on two
symbols.
Proof. Since the set of all finite 0-1-sequences is countable, one can construct
a 0-1-sequence that contains each finite 0-1-sequence as a block. Hence, by
Proposition 1.35(b), σ has a dense orbit, which shows that it is topologically
transitive because Σ2 has no isolated points.
Now, let x ∈ Σ2 and m ∈ N. Then the sequence y defined by yj(m+1)+k =
xk , 0 ≤ k ≤ m, j ≥ 0 is periodic, hence a periodic point for σ. Moreover, by
Lemma 1.34, we have that d(x, y) ≤ 21m . This shows that the periodic points
are dense in Σ2 .
Altogether, σ is chaotic.
It will be a recurrent theme in this book that shifts create chaos. And in
many cases maps are chaotic precisely because there is an underlying shift.
We illustrate this by the doubling map.
∞
xn
α= , xn ∈ {0, 1}.
n=0
2n+1
In other words, the doubling map acts as a shift on the binary representation
of the argument of z. In this form the dynamics of the doubling map become
much clearer. To put it formally, the map
∞
xn
φ : Σ2 → T, (xn )n → exp 2πi
n=0
2n+1
Example 1.39. As noted above, both the tent map and the doubling map are
mixing. An example of a topologically transitive system that is not mixing
will be given in Example 1.43.
(S × T )n = S n × T n .
the fact that topological transitivity and the mixing property can be tested
on a base of the topology, and, for (iv), the observation of Exercise 1.2.4.
to construct examples of weakly mixing maps that are not mixing. How-
ever, in the context of linear operators that will be discussed in the following
chapters, such examples will appear abundantly as consequences of our in-
vestigations there; see, for example, Remark 4.10.
Proposition 1.47. The weak mixing property is preserved under quasicon-
jugacy.
Proof. If φ : Y → X defines a quasiconjugacy from S : Y → Y to T : X → X,
then φ × φ defines a quasiconjugacy from S × S to T × T . Now the result
follows from Proposition 1.13.
As a consequence, we have the following as in Proposition 1.42.
Proposition 1.48. Let S : X → X and T : Y → Y be dynamical systems. If
S × T is weakly mixing then so are S and T .
In order to formulate the arguments involving weakly mixing maps more
succinctly we introduce the following useful concept.
Definition 1.49. Let T : X → X be a dynamical system. Then, for any sets
A, B ⊂ X, the return set from A to B is defined as
We usually drop the index T when this causes no ambiguity. In this nota-
tion, T is topologically transitive (or mixing) if and only if, for any pair U, V
of nonempty open subsets of X, the return set
N (U1 , V1 ) ∩ N (U2 , V2 ) = ∅.
Note also that if T is topologically transitive then the return sets N (U, V )
are even infinite for any nonempty open sets U, V ; see Exercise 1.2.4.
Incidentally, we observe that the larger the sets A and B are, the larger
also is the return set N (A, B).
It is our aim to give several characterizations of the weak mixing property.
To do this we will provide a useful lemma that, due to its form, we will call
the 4-set trick. We note that the return sets N (A, B) refer to T .
Lemma 1.50 (4-set trick). Let T : X → X be a dynamical system, and let
U1 , V1 , U2 , V2 ⊂ X be nonempty open sets.
(a) If there is a continuous map S : X → X commuting with T such that
then there exist nonempty open sets U1 ⊂ U1 , V1 ⊂ V1 such that
Proof. (a) Since S is continuous, it follows from the hypothesis that we can
find nonempty open sets U1 ⊂ U1 and V1 ⊂ V1 such that S(U1 ) ⊂ U2 and
S(V1 ) ⊂ V2 . If n ∈ N (U1 , V1 ), then there exists some x ∈ U1 with T n x ∈ V1 .
Therefore T n Sx = ST n x ∈ V2 and Sx ∈ U2 , which yields that n ∈ N (U2 , V2 ).
By symmetry we also obtain that N (V1 , U1 ) ⊂ N (V2 , U2 ). If, moreover, T is
topologically transitive then
n−1
N (Uk , Vk ) ∩ N (Un , Vn ) = ∅,
k=1
N (U, V1 ) ∩ N (U, V2 ) = ∅.
N (U, U ) ∩ N (U, V ) = ∅.
Proof. It suffices to show that the stated condition implies the condition of
Proposition 1.52. Thus, let U, V1 , V2 ⊂ X be arbitrary nonempty open sets.
By hypothesis there exists some n ∈ N0 such that U1 := U ∩ T −n (V1 ) is a
nonempty open set. Since topologically transitive maps have dense range, the
hypothesis also implies that T −n (V2 ) is nonempty and open, so that there
exists some m ∈ N (U1 , U1 ) ∩ N (U1 , T −n (V2 )). Therefore there are x, y ∈ U1
with T m x ∈ U1 and T n T m y ∈ V2 . We then have that T n T m x ∈ V1 , which
implies that n + m ∈ N (U, V1 ) ∩ N (U, V2 ), as desired.
T n (U ) ∩ T −k (V ) = ∅ for k = 1, . . . , m.
Condition (ii) in this result shows nicely how weak mixing sits between
topological transitivity and mixing.
1.6 Universality
In this case, Tn is the map that associates to f its Taylor polynomial of degree
n at 0.
The theory of universality will not be developed in any depth in this book.
We note that there is a difference in philosophy between universality and
topological dynamics: in the former one is interested in the universal elements
and their properties while in the latter the focus is rather on the map and
its properties.
24 1 Topological dynamics
Now, many of the results in this chapter extend, at least under suitable
assumptions, to general sequences. We will content ourselves here with some
examples.
Proof. Suppose that (ii) holds. If U and V are nonempty open sets of X and
Y , respectively, then there exists some x ∈ U with dense orbit under (Tn )n ,
so that there exists some n ≥ 0 with Tn x ∈ V . This implies (i).
The converse implication and the fact that the set of points with dense
orbit is a dense Gδ -set can be proved exactly as in the proof of the Birkhoff
transitivity theorem.
N (A, B) = {n ∈ N0 ; Tn (A) ∩ B = ∅}
then part (a) of the 4-set trick remains valid for sequences (Tn )n of self-maps
if the map S commutes with all Tn , n ≥ 0, as does part (b) for commuting
self-maps. As a consequence, Furstenberg’s theorem also holds for commuting
sequences (Tn )n , as does Proposition 1.52.
Exercises 25
Exercises
Exercise 1.1.1. Show that for μ = 2 the iterates of the logistic map L2 are given by
n
Ln
2x =
1
2 1 − (1 − 2x)2 , n ≥ 0.
Deduce from this the long-term behaviour of the orbits orb(x, L2 ) for x ∈ R.
Exercise 1.1.4. Show that the logistic map L4 , when restricted to the interval [0, 1], is
quasiconjugate to the doubling map on the interval via φ(x) = sin2 (πx). Also show that
the maps are not conjugate.
Exercise 1.2.1. Show that, in general, the implication (i)=⇒(ii) does not hold in Propo-
sition 1.10.
Exercise 1.2.3. Suppose that X has at least one isolated point. Prove that, if there is
any topologically transitive map T : X → X, then X is finite and X = orb(x, T ) for any
x ∈ X.
Exercise 1.2.4. Show that, if T : X → X is topologically transitive, then for any pair
U, V of nonempty open subsets of X, the return set N (U, V ) is infinite; see Definition
1.49. (Hint: For the trivial case in which X has isolated points apply Exercise 1.2.3. If
X has no isolated points, then given m ∈ N (U, V ) and W := U ∩ T −m (V ), observe that
N (W, W ) ∩ N = ∅ and m + N (W, W ) ⊂ N (U, V ).)
Exercise 1.2.9. Show that every orbit under an irrational rotation is dense. (Hint: Use
the pigeonhole principle to show that, for any ε > 0, some arc of angle ε must contain
two iterates of 1, T m 1 and T n 1, m > n. Then look at the iterates of T m−n .)
(a) Show that T has a dense orbit but that T 2 does not.
(b) Show that there are two points x, y ∈ [−1, 1] such that orb(x, T 2 ) ∪ orb(y, T 2 ) is
dense in [−1, 1] but neither of them has a dense orbit under T 2 .
(c) Show that there is a point x ∈ [−1, 1] such that orb(x, T 2 ) contains a nonempty
open set but x does not have a dense orbit under T 2 .
(Remark: We will prove in Chapter 6 that none of these properties can hold in a
linear setting.)
Exercise 1.3.1. Show that none of the three conditions in Definition 1.26 alone implies
chaos.
Exercise 1.3.2. Let X be a finite set, endowed with the discrete metric. Describe all
maps on X that are chaotic. Do the same for countably infinite sets under the discrete
metric.
Exercise 1.3.3. Suppose that (X, d) is a metric space without isolated points and T :
X → X is a contracting map, that is d(T x, T y) ≤ d(x, y) for all x, y ∈ X. Show that if
T has one dense orbit then T is minimal (see Exercise 1.2.10); in particular, it cannot
be chaotic.
Exercises 27
Exercise 1.3.4. Show that T : X → X is chaotic if and only if every finite family of
nonempty open sets shares a periodic orbit, in the following sense: for each finite family
Uj ⊂ X, j = 1, . . . , n, of nonempty open sets there is a periodic point x ∈ U1 such
that T kj x ∈ Uj for some kj ≥ 0, j = 2, . . . , n; see Figure 1.8. (Hint: One implication is
trivial; for the other one use continuity of T and an induction process.)
Exercise 1.3.5. Show that the space Σ2 is a complete metric space without isolated
points. Show also that the sequences with only finitely many nonzero entries form a
dense set.
Exercise 1.3.6. Why is the quasiconjugacy of Example 1.37 between the shift on two
symbols and the doubling map not a conjugacy? Use this quasiconjugacy to find a
representation of the periodic points and the points with dense orbit for the doubling
map.
Exercise 1.4.2. Prove that a dynamical system T is mixing if and only if, for any
strictly increasing sequence (nk )k of positive integers, the sequence (T nk )k is topologi-
cally transitive; see Definition 1.56 for the notion of topological transitivity for sequences
of maps.
Exercise 1.4.3. Let X be a complete metric space. Prove that a dynamical system
T : X → X is mixing if and only if, for every sequence (xn )n in X and for every strictly
increasing sequence (nk )k of positive integers for which {xnk ; k ∈ N} is relatively com-
pact there exists a dense Gδ -set of points y ∈ X such that lim inf k→∞ d(xnk , T nk y) = 0.
(Hint: Use the previous exercise; a subset A of a metric space X is relatively compact if
and only if every sequence in A has a subsequence that converges in X.)
Exercise 1.4.4. Let T : X → X be a dynamical system. For x ∈ X, the set JTmix (x) =
J mix (x) is defined as the set of all points y ∈ X for which there is a sequence (xn )n in
X such that xn → x and T n xn → y as n → ∞; see also Exercise 1.2.8.
(a) Show that J mix (x) is a closed T -invariant set.
(b) Show that J mix (x) = X if and only if, for any pair U, V of nonempty open subsets
of X with x ∈ U , there exists some N ≥ 0 such that T n (U ) ∩ V = ∅ for all n ≥ N .
(c) Show that the following assertions are equivalent:
(i) T is mixing;
(ii) for any x ∈ X, J mix (x) = X;
(iii) there is a dense set of points x ∈ X such that J mix (x) = X.
28 1 Topological dynamics
Exercise 1.4.5. Show that minimality and mixing are independent properties of a dy-
namical system.
Exercise 1.5.3. Prove that every chaotic and totally transitive dynamical system T :
X → X is weakly mixing. (Hint: Verify the hypothesis of Proposition 1.53 by finding a
periodic point in U , of period k say, and then by using topological transitivity of T k .)
Exercise 1.5.4. A dynamical system T : X → X is called flip transitive if, for any pair
U, V ⊂ X of nonempty open sets, N (U, V ) ∩ N (V, U ) = ∅. Show that the map T of
Exercise 1.2.11 is flip transitive but not weakly mixing.
Exercise 1.5.5. Show that T is weakly mixing if and only if it is flip transitive and T 2
is topologically transitive. (Hint: To prove the weak mixing property use condition (i)
in Exercise 1.5.1. To do this, given nonempty open sets U, V ⊂ X, find k ∈ N0 with
U := U ∩ T −2k (V ) = ∅ and then some m ∈ N (U , T −k (V )) ∩ N (T −k (V ), U ). Consider
m + k.)
Exercise 1.5.7. Let T : X → X be a dynamical system. Show that any of the following
conditions is equivalent to T being weakly mixing:
(i) for any nonempty open sets U, V ⊂ X and any m ∈ N there is some k with
k, k + m ∈ N (U, V );
Exercises 29
(ii) for any m ∈ N and for any increasing sequence (nk )k with nk+1 − nk ∈ {m, 2m},
k ∈ N, we have that (T nk )k is topologically transitive.
(Hint: See the proof of Theorem 1.54.)
k
V(U1 , . . . , Uk ) := K ∈ K(X) ; K ⊂ Uj and K ∩ Uj = ∅, j = 1, . . . , k ,
j=1
Exercise 1.6.2. Prove the Birkhoff transitivity theorem for commuting continuous maps
Tn : X → X, n ∈ N0 , of dense range on a separable complete metric space X, that is,
that the following assertions are equivalent:
(i) (Tn )n is topologically transitive;
(ii) there exists some x ∈ X such that orb(x, (Tn )) is dense in X.
If one of these conditions holds then the set of points in X with dense orbit is a dense
Gδ -set.
Exercise 1.6.3. Let S be a mixing map on a separable complete metric space X and T
a map on a metric space Y without isolated points that admits a dense orbit orb(y, T ),
y ∈ Y . Show that there exists some x ∈ X such that (x, y) has a dense orbit under the
map S × T . (Hint: Use the previous exercise.)
Exercise 1.6.5. Show that Theorem 1.54 does not hold for sequences (Tn )n even if the
maps Tn : X → X commute and have dense range. (Hint: A weakly mixing sequence
(Tn )n remains weakly mixing when adding arbitrary maps.)
Section 1.1. A standard reference for the theory of dynamical systems is Devaney [132].
For more recent textbooks we refer to Brin and Stuck [97] and Robinson [267], while
Gulick [188] provides an elementary introduction.
Section 1.2. Kolyada and Snoha [217] give an excellent survey on topological transi-
tivity, with many additional equivalent conditions. The original version of the Birkhoff
transitivity theorem can be found in [74, § 62].
Section 1.3. Chaos in the sense of Devaney was introduced in [132]. While there are
many other definitions of chaos (see for example Kolyada [216] or Forti [154]), Devaney’s
definition has become very popular. The theorem of Banks et al. was obtained in [31];
see also Silverman [294] and Glasner and Weiss [163].
Sections 1.4 and 1.5. For Furstenberg’s theorem see [157], which also contains the
4-set trick implicitly. Theorem 1.54 is due, independently, to Akin [4], Glasner [162],
and Peris and Saldivia [257]; in the context of linear operators on Banach spaces it was
also obtained by Grivaux [172]. The remaining characterizations of weak mixing in Sec-
tion 1.5, including Exercise 1.5.1, are due to Banks [29] and Akin [4]; more precisely,
Proposition 1.53 is called the “Furstenberg Intersection Lemma” in Akin’s book.
Section 1.6. The universal Taylor series mentioned in Section 1.6 is essentially due to
Fekete; see the discussion in [179, Section 3a]. The Universality Criterion was obtained
by Grosse-Erdmann [177].
Exercises. Exercises 1.2.8 and 1.4.4 are taken from Costakis and Manoussos [118, 119].
For Exercise 1.3.4 we refer to Touhey [299], for Exercise 1.4.3 to Moothathu [244], for
Exercise 1.5.3 to Bauer and Sigmund [32] and to Banks (attributed to Stacey) [28], for
Exercise 1.5.5 to Banks [29], and for Exercise 1.5.6 to Moothathu [245]. The two parts
of Exercise 1.5.7 are taken from Grivaux [172] and Peris and Saldivia [257], respectively;
Exercise 1.5.8 is from Grosse-Erdmann and Peris [187] (see also Grivaux [172] and Bayart
and Matheron [45]). The assertion of Exercise 1.5.9 is due to Bauer and Sigmund [32]
(one direction), and, independently, to Banks [30] and Peris [256] (the other direction).
Exercise 1.6.1 is from Godefroy and Shapiro [165], Exercise 1.6.4 from Bès and Peris [71].
Extensions. Let us add a word on the setting chosen in this chapter. Since the over-
whelming majority of linear dynamical systems studied in the literature acts on metric
spaces we have restricted our attention to such spaces. In general, however, a dynamical
system is given by a continuous map T : X → X on a topological space X. The def-
initions of topologically transitive, (weakly) mixing and chaotic maps extend verbatim
to such systems. The same applies to sequences (Tn )n of continuous maps Tn : X → Y
between arbitrary topological spaces X and Y .
Then, as the proofs show, all the results in this chapter on general dynamical systems
remain true in the setting of arbitrary topological spaces. To be more specific, this
concerns all results apart from Proposition 1.15 and the Theorems 1.16, 1.29 and 1.57.
Chapter 2
Hypercyclic and chaotic operators
D : f → f .
advanced results that are only used occasionally in this book will be covered
in Appendix A.
The following two examples will motivate the concept of a Fréchet space.
Example 2.1. The process of taking derivatives, that is, the operator D : f →
f , provides an interesting linear dynamical system. In order to have the
powerful tools of complex analysis at our disposal we regard D as acting on
the space of entire functions,
H(C) = {f : C → C ; f holomorphic}.
The natural concept of convergence for entire functions is that of local uni-
form convergence, that is, the uniform convergence on all compact sets. In
contrast to Banach spaces, convergence is described here by a countably in-
finite collection of conditions. More precisely, we have that fk → f in H(C)
if and only if, for all n ∈ N, pn (fk − f ) → 0 as k → ∞, where
pn (f ) := sup |f (z)|.
|z|≤n
Example 2.2. Many natural spaces of sequences are Banach spaces. But the
space of all (real or complex) sequences,
ω := KN = {(xn )n ; xn ∈ K, n ∈ N},
d(x, y) := x − y, x, y ∈ X,
2.1 Linear dynamical systems 33
and which is complete in that metric. If, moreover, the norm derives from an
inner product · , · via
x := x, x, x ∈ X,
The following result will be of constant use. We leave its proof as a useful
exercise to the reader; see Exercise 2.1.2.
∞
1
x := n
min(1, pn (x)), x ∈ X, (2.2)
n=1
2
The notion of an F-norm has the advantage that one can largely argue as
if one was working in a Banach space. One need only be aware of the fact
that the positive homogeneity of a norm is no longer available. In fact, in
many cases, this property is not needed at all or it can be replaced by the
following weaker property that follows directly from conditions (i) and (ii):
for all x ∈ X and λ ∈ K,
Having discussed Fréchet spaces and their topology we now turn to the
concept of operators on them.
pn x <ε
pn (x) + δ
and hence
pn (x) + δ
qm (T x) < .
ε
Since δ > 0 is arbitrary we obtain the result with M = 1/ε.
Ta f (z) = f (z + a), a ∈ C.
X ⊕ Y := {(x, y) ; x ∈ X, y ∈ Y }
With this we end our introduction to Fréchet spaces and their operators.
We will show in Chapter 12 that several important results in linear dynamics
2.2 Hypercyclic operators 37
The origin of this terminology is easily explained. For a long time, operator
theorists have been studying so-called cyclic vectors in connection with the
invariant subspace problem. Vectors with a more restrictive property were
then called supercyclic.
span {T n x ; n ≥ 0}
This suggested the name of hypercyclicity for the case when the orbit itself
is dense. Cyclic and supercyclic vectors will not be studied in detail in this
book.
38 2 Hypercyclic and chaotic operators
The invariant subspace problem, which is open to this day, asks whether
every Hilbert space operator possesses an invariant closed subspace other
than the trivial ones given by {0} and the whole space. Counterexamples do
exist for operators on non-reflexive spaces like 1 .
Obviously, the smallest closed T -invariant subspace of X that contains a
given point x coincides with the closure of the span of its orbit. Therefore, an
operator has no nontrivial invariant closed subspace precisely if every nonzero
vector is cyclic. By the same token we have a link between hypercyclicity and
the invariant subset problem: does every Hilbert space operator possess an
invariant closed subset other than the trivial ones given by {0} and the whole
space?
Observation 2.17. An operator has no nontrivial invariant closed subsets if
and only if every nonzero vector is hypercyclic.
Having explained the historical interest in hypercyclicity, our first question
has to be if hypercyclic operators exist. That is, does the additional require-
ment of linearity still allow us to find maps with dense orbits? Indeed, a very
simple operator on the Hilbert space 2 turns out to be hypercyclic.
Example 2.18. Let T : 2 → 2 be twice the backward shift, that is,
k−1 ∞
T nk x = 2nk −nj B nk −nj y (j) + y (k) + 2nk −nj F nj −nk y (j)
j=1 j=k+1
∞
= y (k) + 2nk −nj F nj −nk y (j) ,
j=k+1
2.2 Hypercyclic operators 39
we deduce that T nk x − y (k) ≤ 2−k . Since the y (k) form a dense set, x has
a dense orbit under T .
Directly or indirectly, the transitivity theorem will be our main tool for
proving the hypercyclicity of an operator.
The first examples of hypercyclic operators were found by G.D. Birkhoff
in 1929, G.R. MacLane in 1952 and S. Rolewicz in 1969. These operators will
accompany us throughout the book as they will serve as a testing ground
for any new concept in linear dynamics; indeed, Example 2.18 was already a
special Rolewicz operator.
Ta f (z) = f (z + a), a = 0.
sup |f (z) − p(z)| < ε and sup |g(z − na) − p(z)| < ε,
z∈K z∈K+na
N
k!|bk | k+n
sup |r(z) − p(z)| ≤ R →0
|z|≤R (k + n)!
k=0
T = λB : X → X, (x1 , x2 , x3 , . . .) → λ(x2 , x3 , x4 , . . .)
x = (x1 , x2 , . . . , xN , 0, 0, . . . ), y = (y1 , y2 , . . . , yN , 0, 0, . . . ),
There are various ways to derive the hypercyclicity of one operator from
that of another. The first result of this type follows from Proposition 1.14
by the Birkhoff transitivity theorem. Recall that by the inverse mapping
theorem (see Appendix A), any bijective operator has a continuous inverse
and is therefore.
2.2 Hypercyclic operators 41
We will see in Remark 4.17 that the converse fails in general. As an ap-
plication of Proposition 2.25 we obtain an interesting transference principle
that is specific to the linear setting. Let X be a real separable Fréchet space.
Then the complexification X of X is defined formally as
= {x + iy ; x, y ∈ X},
X
T(x + iy) = T x + iT y.
T x = T x1 + T x2 if x = x1 + x2 , x1 ∈ M1 , x2 ∈ M2 .
T = T |M1 ⊕ T |M2 ;
Remark 2.31. Devaney’s notion of chaos has been generally accepted in linear
dynamics. There are, however, also other definitions of chaos. We mention
here that a continuous map T : X → X on a metric space (X, d) is called
chaotic in the sense of Auslander and Yorke if it is topologically transitive and
it has sensitive dependence on initial conditions. By the previous proposition,
every hypercyclic operator is Auslander–Yorke chaotic.
In order to see that the set of periodic points is dense in X it suffices to ap-
proximate any finite sequence y = (y1 , . . . , yn , 0, . . .). By choosing a periodic
point whose∞ N ≥ n first coordinates coincide with those of y we see that
x − y ≤ j=1 |λ|−jN y → 0 as N → ∞. Therefore Rolewicz’s operators
are chaotic.
Let us observe that, for linear maps T on arbitrary vector spaces X, the
set of periodic points of T is a subspace of X. Indeed, let x, y ∈ X be periodic
points for T . Then we have that T n x = x and T m y = y for certain n, m ∈ N.
Thus T nm (ax + by) = a(T n )m x + b(T m )n y = ax + by, for any a, b ∈ K, so
that also ax + by is periodic.
There is, in fact, a nice and very useful description of the space of periodic
points in terms of eigenvectors to unimodular eigenvalues, that is, eigenvalues
of absolute value 1, provided that we have a complex space. The correspond-
ing result is of a purely algebraic nature.
z n − 1 = (z − λ1 )(z − λ2 ) · · · (z − λn ).
eλ (z) = eλz , z ∈ C.
Lemma 2.34. Let Λ ⊂ C be a set with an accumulation point. Then the set
span{eλ ; λ ∈ Λ}
is dense in H(C).
(λn − λ)2 z 2
eλn z = eλz e(λn −λ)z = eλz + eλz (λn − λ)z + eλz + ... (2.4)
2!
we see that
eλn z → eλz uniformly on compact sets,
which, incidentally, also follows directly. Therefore, eλ ∈ span{eλn ; n ≥ 1}.
But now (2.4) also shows that
2.3 Linear chaos 45
∞
∞
f (z) = eλz e−λz f (z) = eλz ak z k = ak z k eλz
k=0 k=0
The lemma allows us to show that Birkhoff’s and MacLane’s operators are
chaotic on H(C).
The restriction that Proposition 2.33 only holds for complex spaces can
sometimes be overcome by using complexifications: an operator on a real
space is chaotic if and only if its complexification is, as we will see in Corol-
lary 2.51 below.
46 2 Hypercyclic and chaotic operators
In this and the following section we study the mixing and weak mixing prop-
erties in the light of linearity.
Since we are in the setting of Fréchet spaces, the proofs could be formulated
in terms of their seminorms or their metric. However, the arguments become
particularly transparent when we use the topological language of open sets
and 0-neighbourhoods. For this we only need the following simple result.
As usual, we set A + B = {a + b ; a ∈ A, b ∈ B} for subsets A, B of a
vector space.
Proposition 2.37. An operator T is mixing if and only if, for any nonempty
open set U ⊂ X and any 0-neighbourhood W , the return sets
are cofinite.
We have also seen that these mixing operators are even chaotic. This is
not always the case, as the following example shows.
2.4 Mixing operators 47
which is a contradiction.
We reformulate part of a previous result, Proposition 1.42, for operators.
Proposition 2.40. Let S : X → X and T : Y → Y be hypercyclic operators.
If at least one of them is mixing then S ⊕ T is hypercyclic. Moreover, S ⊕ T
is mixing if and only if both S and T are.
For a later application (see Proposition 8.5), we also need to consider
direct sums of countably many operators on Banach spaces. Thus, let Tn be
operators on separable Banach spaces Xn , n ≥ 1. For 1 ≤ p < ∞, we define
the direct p -sum of these spaces as
∞
∞
Xn = (xn )n≥1 ; xn ∈ Xn , n ≥ 1, and xn p < ∞ ;
p
n=1 n=1
∞
endowed with the norm (xn )n = ( n=1 x n )
p 1/p
this space turns into a
∞
separable Banach space. The direct c0 -sum ( n=1 Xn )c0 is defined similarly.
Now suppose that supn∈N Tn < ∞. Then the direct sum of the operators
Tn , defined by
48 2 Hypercyclic and chaotic operators
∞
Tn (xn )n = (Tn xn )n ,
n=1
∞ ∞
is an operator on ( n=1 Xn )p and on ( n=1 Xn )c0 .
Proof. For the necessity ∞ part one need only note that each Tn , n ≥ 1, is
quasiconjugate to k=1 Tk via the map φ : (xk )k → xn .
Now suppose that each Tn , n ≥ 1, is mixing. Let U, V ⊂ ( ∞ n=1 Xn ) be
p
nonempty open sets. It follows from the definition of the norm on this space
that there are ε > 0, m ≥ 1, and points x := (x1 , . . . , xm , 0, 0, . . .) ∈ U and
y := (y1 , . . . , ym , 0, 0, . . .) ∈ V such that the open balls of radius ε around
these points belong to U and V , respectively. Since each Tk is mixing there is
(n)
some N ≥ 1 such that, for each 1 ≤ k ≤ m and n ≥ N , there are xk ∈ Xk
(n) (n)
such that xk − xk < ε/m1/p and Tkn xk − yk < ε/m1/p . Then, for all
(n) (n) ∞
n ≥ N , x(n) := ∞ (x1 , . . . , xm , 0, 0, . . .) ∈ U and ( k=1 Tk )n x(n) ∈ V , which
implies that k=1 Tk is mixing. The proof for direct c0 -sums is similar.
Refining the techniques of De la Rosa and Read, Bayart and Matheron have
shown that such operators even exist on any of the spaces p , 1 ≤ p < ∞, and
c0 , in particular on Hilbert spaces. The proof is, however, beyond the scope
of this book.
2.5 Weakly mixing operators 49
Lemma 2.44. Let T be a hypercyclic operator. Then, for any nonempty open
sets U and V in X and any 0-neighbourhood W , there is a nonempty open
set U1 ⊂ U and a 0-neighbourhood W1 ⊂ W such that
This proof copies our proof of the 4-set trick (see Lemma 1.50); a direct
application of that trick would not have given us that W1 contains 0.
Theorem 2.45. Let T be a hypercyclic operator. If, for any nonempty open
set U ⊂ X and any 0-neighbourhood W , there is a continuous map S : X →
X commuting with T such that
Proof. First, the 4-set trick and topological transitivity of T yield that, for
any nonempty open set U ⊂ X and for any 0-neighbourhood W ,
N (U, W ) ∩ N (W, U ) = ∅.
N (U, V ) ∩ N (V, U ) = ∅;
see also Exercises 1.5.4 and 1.5.5. By the Birkhoff transitivity theorem, such
an operator is hypercyclic. It follows from the previous result that even more
is true.
Theorem 2.47. An operator T is weakly mixing if and only if, for any
nonempty open sets U, V ⊂ X and any 0-neighbourhood W ,
N (U, W ) ∩ N (W, V ) = ∅.
X = HC(T ) + HC(T ),
that is, every vector x ∈ X can be written as the sum of two hypercyclic
vectors.
Proof. Let x ∈ X. Since both HC(T ) and x − HC(T ) are dense Gδ -sets,
their intersection must be nonempty by the Baire category theorem, which
implies that x ∈ HC(T ) + HC(T ).
Lemma 2.53. (a) Let T be a hypercyclic operator. Then its adjoint T ∗ has
no eigenvalues. Equivalently, every operator T − λI, λ ∈ K, has dense range.
(b) Let T be a hypercyclic operator on a real separable Fréchet space. Then
the adjoint T∗ of its complexification T has no eigenvalues. Equivalently,
every operator T − λI, λ ∈ C, has dense range.
T ∗ x∗ = λx∗
T n x, x∗ = x, (T ∗ )n x∗ = λn x, x∗ .
2.6 The set of hypercyclic vectors 53
∗ | = |Tn x, x
|T n x, x ∗ | = |x, (T∗ )n x∗ | = |λ|n |x, x
∗ |, (2.6)
∗ = x1 , x
for n ≥ 0. Since x1 + ix2 , x ∗ + ix2 , x
∗ for all x1 , x2 ∈ X, and
∗
= 0, there is some y ∈ X such that |y, x
since x ∗ | > 0. Hence | x∗ | can
take every positive value on X. By the hypercyclicity of x, the left-hand side
of (2.6) is dense in R+ , while the right-hand side clearly is not, which is a
contradiction. The remainder of the proof can be given as in (a).
N ≥ 1. For spaces X over the complex field the result follows immediately
from Lemma 2.53(a) and the fact that p can be written as a product of linear
factors, so that
p(T ) = aN (T − λ1 I) · · · (T − λN I)
with certain λk ∈ C, k = 1, . . . , N .
(Real case). If X is a real space we consider the complexification T of T .
With Lemma 2.53(b), it follows as in the complex case that, for any complex
polynomial p, p(T) has dense range on X. Now if p has real coefficients, then
Corollary 2.56. The set HC(T ) of hypercyclic vectors for a hypercyclic op-
erator T is a connected subset of X.
We have seen that chaotic linear operators exist. This is contrary to the com-
mon belief that (deterministic) chaos is necessarily connected to the nonlin-
earity of a system. In this section we want to explore the connection between
linear and nonlinear chaos.
The dynamics of linear operators on a finite-dimensional space X = KN
are easy to describe, thanks to the Jordan decomposition theorem. We assume
that KN is endowed with the Euclidean norm.
equivalent we can assume that this basis is the canonical basis of CN , and it
suffices to show the result for each operator given by a Jordan block
⎛ ⎞
λ 1 0 ... ...
⎜0 λ 1 0 ...⎟
⎜ ⎟
⎜ ⎟
T = ⎜ . . . . . . . . . ... ⎟ : CN → CN ,
⎜ ⎟
⎝ 0 λ 1 ⎠
0 λ
Of course, this also follows directly from Lemma 2.53 since every opera-
tor on CN has an eigenvalue. Further proofs of this result are suggested in
Exercises 2.7.1 and 2.7.2.
Since every finite-dimensional Fréchet space is isomorphic to some KN , N ≥
1 (see Appendix A), the theorem extends to such spaces.
Proposition 2.60. The orbit of any hypercyclic vector forms a linearly in-
dependent set.
56 2 Hypercyclic and chaotic operators
Proof. Let x be a hypercyclic vector for T and suppose that there are scalars
αk ∈ K, k = 0, . . . , N, such that
N
T N +1 x = αk T k x.
k=0
The proof that T has the desired properties will be split into three steps.
This can be made arbitrarily small by first choosing N sufficiently large and
then y sufficiently close to x. Thus φ : K → H is continuous.
Moreover, φ is injective; indeed, φ(x) = φ(y) implies that
1 1
k
d(yk , x) = k d(yk , y) for all k ≥ 0,
2 2
hence x = y by density of the yk .
As a continuous injection on a compact space, φ is a homeomorphism onto
a (compact) subset, L say, of H.
The proof is the same as the proof that Rolewicz’s operators are chaotic;
see Examples 2.22 and 2.32.
While this is not the common definition of the Julia set, which is more com-
plicated and involves the behaviour of the iterates of p near J (p), a result
by Fatou and Julia tells us that the two definitions are equivalent.
For instance, for the doubling map p(z) = z 2 on C (see Example 1.37), the
periodic points are given by e2πiα with α = 2nk−1 , n, k ∈ N, and all of them
are repelling, so that J (p) = T. We have also seen that p|T is chaotic.
This feature is shared by any complex polynomial of degree m ≥ 2. More
precisely one always has that J (p) is a compact p-invariant set such that
p|J (p) is chaotic.
The dynamical behaviour of p near the Julia set consists in spreading
points. Indeed, the following property, which can be regarded as a multi-
point approximation by the iterates of p on points near J (p), characterizes
the Julia set: a point z ∈ C belongs to J (p) if and only if
Now, one can deduce this property, for certain polynomials and at the
point 0, from the hypercyclic behaviour of Rolewicz’s operators.
mB
c0 −−−−→ c0
⏐ ⏐
⏐
φ
⏐φ
P
c0 −−−−→ c0 ,
where mB is the Rolewicz operator with λ = m on the complex space c0 ,
P (x1 , x2 , . . . ) = (p(x2 ), p(x3 ), . . . ), and φ(x1 , x2 , . . . ) = (ex1 − 1, ex2 − 1, . . . ).
It is easy to see that P and φ are continuous maps and that φ has dense
range. By Proposition 1.19, it follows from the hypercyclicity of Rolewicz’s
operators that P has a dense orbit.
In order to verify condition (2.7) at z = 0, we fix ε > 0 and arbitrary
zj ∈ C, j = 1, . . . k. Since P has a dense orbit we can find w ∈ c0 and n ≥ 0
such that
Exercises
Exercise 2.1.2. Prove Lemma 2.6. (Hint: Use the previous exercise.)
Exercise 2.1.3. Show that the spaces H(C) and ω are Fréchet spaces. Show that for
every separable Fréchet space X, X N is also a separable Fréchet space.
Exercise 2.1.5. Let X = C ∞ (R) be the space of infinitely differentiable (real or com-
plex) functions f : R → K, endowed with the seminorms
pn (f ) = max sup f (k) (x) .
0≤k≤n |x|≤n
Show that X is a separable Fréchet space. (Hint: Use the Weierstrass approximation
theorem to approximate f (n) on [−n, n] by a polynomial.)
Exercise 2.1.7. Let X be a Fréchet space with defining increasing sequence of semi-
norms (pn )n . Then a seminorm p : X → R is continuous if and only if there are n ≥ 1
and M > 0 such that
p(x) ≤ M pn (x), x ∈ X.
Show that this immediately implies the nontrivial part of Proposition 2.11.
Exercise 2.2.2. Let X be a Banach space. Show that there are no operators T on X
for which λT is hypercyclic for all λ = 0. (Hint: Consider |λ| ≤ 1/T .)
Exercise 2.2.3. Given any sequence of nonzero scalars (wn )n , we define the operator
Bw : ω → ω, (x1 , x2 , x3 , . . . ) → (w2 x2 , w3 x3 , w4 x4 , . . . ). Prove that Bw is hypercyclic.
Exercise 2.2.4. Let T be the translation operator on the space Lpv (R+ ) defined in
Exercise 2.1.6. We assume that there are constants M ≥ 1 and w ∈ R such that
Exercise 2.2.5. Let X = C ∞ (R) be the space of infinitely differentiable real functions
f : R → R; see Exercise 2.1.5. Show that the (real) differentiation operator D : X → X,
f → f is hypercyclic by defining a suitable quasiconjugacy φ : H(C) → C ∞ (R).
Exercise 2.2.7. Show that the complexification X of a real separable Fréchet space X
is a complex separable Fréchet space and that the complexification of a (real-linear)
operator on X is a (complex-linear) operator on X.
Exercise 2.3.1. Let T be an operator on a Banach space X. Show that the following
assertions are equivalent:
(i) T has sensitive dependence on initial conditions with respect to the usual metric;
(ii) supn≥0 T n = ∞;
(iii) T has an unbounded orbit.
(Hint: Use the Banach–Steinhaus theorem; see Appendix A.)
Exercise 2.3.4. Let H(Ω) be the Fréchet space of all holomorphic functions on a do-
main Ω in C; see Section 4.3. Then D : f → f is an operator on H(Ω). Show that the
following assertions are equivalent:
(i) D is chaotic on H(Ω);
(ii) D is hypercyclic on H(Ω);
(iii) Ω is simply connected.
(Hint: By Runge’s theorem, the polynomials are dense in H(Ω) if Ω is simply connected.
1
On the other hand, if Ω is not simply connected, approximate a suitable function z−a
by derivatives, and integrate both over closed curves in Ω.)
Exercise 2.4.1. Modify Example 2.39 to obtain a non-chaotic mixing operator on any
space p , 1 < p < ∞.
Exercise 2.4.2. Let T be the translation operator on Lpv (R+ ) with an admissible weight
function v; see Exercise 2.2.4. Show that T is mixing if and only if limx→∞ v(x) = 0.
Exercise 2.4.3. Using Proposition 2.37, show that λD is mixing on H(C) for any λ = 0.
(This should be contrasted with Exercise 2.2.2.)
Exercise 2.5.1. Let T be a weakly mixing (or mixing) operator and λ ∈ K, |λ| = 1.
Show that λT is also weakly mixing (or mixing, respectively). (Hint: Use Proposition
2.37 and Theorem 2.47.)
Exercise 2.5.2. Establish Theorem 2.45 without the use of Proposition 1.53, that is,
show the weak mixing property directly. (Hint: See Figure 2.2.)
Exercise 2.5.3. Let T be an operator such that, for any nonempty open sets U, V ⊂ X
and any 0-neighbourhood W , N (U, W ) is nonempty and N (W, V ) is syndetic. Then
prove that T is weakly mixing. Do likewise if the sets N (U, W ) are syndetic and the sets
N (W, V ) are nonempty. (Hint: Apply Theorem 2.47.)
Exercise 2.5.4. A subset A of N0 is called thickly syndetic if, for any n ∈ N, there is a
syndetic sequence (nk )k of positive integers such that {nk +j ; k ∈ N, j = 0, . . . , n} ⊂ A.
Show the following:
(i) the intersection of two thickly syndetic sets is thickly syndetic;
(ii) let T be an operator and U ⊂ X a nonempty open set; if, for any 0-neighbourhood
W , N (U, W ) is syndetic then these sets are all thickly syndetic; and similarly for
N (W, U ).
Deduce that, for any topologically ergodic operator T , N (U, V ) is thickly syndetic for
any nonempty open sets U, V ⊂ X; see Exercise 1.5.6 for the definition of topological
ergodicity.
Exercise 2.5.6. An operator T is called hereditarily ergodic if, for any pair U, V of
nonempty open subsets of X and any syndetic sequence (nk )k , there exists a subsequence
(nkj )j of (nk )k that is also syndetic such that T nkj (U ) ∩ V = ∅ for all j ≥ 1. Show
that an operator is hereditarily ergodic if and only if it is topologically ergodic. Deduce
that every power T p , p ≥ 1, of a topologically ergodic operator is topologically ergodic.
(Hint: Use Exercise 2.5.4.)
Exercises 63
Exercise 2.5.7. Prove that the direct sum S ⊕ T of two chaotic operators is chaotic.
(Hint: Use Exercises 1.5.6 and 2.5.5.)
Exercise 2.5.8. Show the following variant of Theorem 2.45: let T be an operator such
that, for any nonempty open sets U, V ⊂ X and any 0-neighbourhood W , there is a
continuous map S : X → X commuting with T such that S(U ) ∩ V = ∅, S(W ) ∩ W = ∅
and
N (U, W ) ∩ N (W, U ) = ∅.
Then T is weakly mixing. (Hint: Use the 4-set trick.)
is dense in X, then prove that T is weakly mixing. (Hint: Given U and W , pick u ∈ U
and p ∈ P such that p(T )u = 0. Since p(T ) has a dense range by Theorem 2.54, there is
some r > 0 with p(T )(rW ) ∩ U = ∅. Then define S = rp(T ) and apply Theorem 2.45.)
is weakly mixing. (Hint: Use the previous exercise and a well-known result from linear
algebra.)
Exercise 2.6.1. Show constructively, as in Example 2.18, that every vector x ∈ 2 is
the sum of two vectors that are hypercyclic for T = 2B.
Exercise 2.6.2. Let T be a hypercyclic operator on a complex space. Show that its
adjoint T ∗ has no finite-dimensional invariant subspace.
Exercise 2.6.3. Let S and T be commuting operators on X such that T has dense
range. Show that HC(S) is T -invariant.
Exercise 2.6.4. (a) Let us call an operator T (on a real or complex space) 2-hypercyclic
if there are vectors x, y ∈ X such that
{T n x + T m y ; n, m ≥ 0}
Exercise 2.7.5. Show that there are supercyclic operators on R and on R2 . (Hint: For
R2 , use a rotation.)
Section 2.1. For introductory texts on functional analysis that also cover Fréchet spaces
we refer to Rudin [271] and Meise and Vogt [237]. The notion of an F-norm can be found
in Kalton, Peck and Roberts [212].
Section 2.2. The term “hypercyclic” for vectors with a dense orbit was apparently first
used around 1986 (Beauzamy [46], [47], [48]) and then extended around 1988 to operators
with a dense orbit (Bourdon, Godefroy, Shapiro [94], [165]). Supercyclic vectors were
introduced by Hilden and Wallen [202] in 1973.
Beauzamy’s work was motivated by the invariant subspace problem. The negative
solution for non-Hilbert spaces is due to Enflo [142] and Read [265]. Subsequently, Read
[266] even constructed a counterexample to the invariant subset problem.
Theorem 2.64 (Read). There exists an operator on 1 all of whose nonzero vectors
are hypercyclic.
Section 2.3. Godefroy and Shapiro [165] suggested acceptance of Devaney’s definition of
chaos for linear operators, and they showed that the three classical operators are chaotic.
Chaos in the sense of Auslander and Yorke was introduced for continuous maps on metric
spaces in [18]. Proposition 2.30 is due to Godefroy and Shapiro [165]. Proposition 2.33
seems to be folklore, but see Herrero [195] and Bonet, Martínez and Peris [83].
We have taken the proof of Lemma 2.34 from Aron and Markose [15].
It does not seem to be easy to find a chaotic operator that is not mixing. The first
such operator was constructed by Badea and Grivaux [19].
Sections 2.5. In 1992, Herrero [195] posed the problem of whether every hypercyclic
operator (on a Hilbert space) is weakly mixing. De la Rosa and Read [126] constructed
a counterexample in a suitable Banach space, while Bayart and Matheron [43] showed
that counterexamples exist on many classical Banach spaces like p , 1 ≤ p < ∞, c0 ,
C[0, 1] and L1 [0, 1], as well as on the Fréchet space H(C).
The results of the section appear in Grosse-Erdmann and Peris [187]; see also Bayart
and Matheron [44], [45] and Moothathu [245]. Theorem 2.47 is due to Bernal and Grosse-
Erdmann [62] and León [219]; the characterizing condition first appeared in Godefroy
and Shapiro [165]. Unlike the proofs in [62] and [219], our argument avoids the Baire
category theorem, as does another proof by Yousefi and Rezaei [303].
Theorem 2.48 is due to Grivaux [172]. The fact that every chaotic operator is weakly
mixing, Corollary 2.49, is due Bès and Peris [71]; it also follows directly from a result
that is due to Bauer and Sigmund [32] and to Stacey (see [28]) using Ansari’s theorem
(see Section 6.1).
Section 2.6. Proposition 2.52 was observed, for example, by Grosse-Erdmann [177],
Godefroy (see [94]) and Kahane (see [249]). Lemma 2.53(a) is due to Kitai [215], while
part (b) can be found in Bonet and Peris [85]. Herzog and Lemmert [199] have char-
acterized the hypercyclic operators on ω = CN ; Conejero [108, p. 123] noted that their
condition can be rephrased as σp (T ∗ ) = ∅.
In the complex case, Theorem 2.54 is due to Bourdon [90]. The real case was added
by Bès [66]; our proof is due to Martínez [227]. Theorem 2.55 is due to Herrero [194]
and Bourdon [90].
A considerable improvement of Corollary 2.56 is due to Fathi [146] and Godefroy (see
[44, p. 16]). They showed that for any hypercyclic operator T on a Fréchet space X, the
set HC(T ) of hypercyclic vectors for T is homeomorphic to X.
Section 2.7. Kitai [215] and Rolewicz [268] observed that there are no hypercyclic op-
erators on finite-dimensional spaces. This prompted Rolewicz to pose the problem of
whether every infinite-dimensional separable Banach space admits a hypercyclic opera-
tor, to which we will turn in Chapter 8. Theorem 2.61 is due to Feldman [147].
The title of this section was inspired by Protopopescu [262], where one finds discus-
sions on the relationship between linear and nonlinear chaos; see also Protopopescu and
Azmy [263].
Section 2.8. In this section we follow Peris [255]. For an introduction to the dynamics
of complex polynomials we refer to Devaney [132] and Carleson and Gamelin [98].
Exercises. Exercise 2.2.1 can be found in Aron, Seoane and Weber [16]. The result of
Exercise 2.2.4 is essentially due to Desch, Schappacher and Webb [131]. Exercise 2.2.9
is taken from Feldman [148], Exercise 2.2.10 from Badea, Grivaux and Müller [21],
and Exercise 2.2.11 from Costakis and Manoussos [119]. Exercise 2.3.1 can be found
in Feldman [147], Exercise 2.3.2 in deLaubenfels and Emamirad [128], Exercise 2.3.4 in
Shapiro [280], and Exercise 2.4.2 in Bermúdez, Bonilla, Conejero, and Peris [49]. Ex-
ercises 2.5.1, 2.5.3 and 2.5.8 are extracted from Grosse-Erdmann and Peris [187]. The
fact that topologically ergodic operators are weakly mixing (part of Exercise 2.5.5) is
implicit in Grosse-Erdmann and Peris [185]. The main part of Exercise 2.5.5 is due to
Sources and comments 67
Desch and Schappacher [129], who introduce the concept of operators satisfying the
Recurrent Hypercyclicity Criterion, which is equivalent to topological ergodicity. Hered-
itarily ergodic operators (see Exercise 2.5.6), were introduced as hereditarily syndetic
operators by Badea and Grivaux [19]. Exercise 2.5.9 is taken from Badea, Grivaux and
Müller [20], Exercises 2.5.10 and 2.5.11 from Grivaux [172], Exercises 2.6.5 and 2.7.1
from Bès [66], and Exercise 2.6.6 from Bourdon [90].
Extensions. Let us add again some remarks on the setting chosen for this chapter (and
for most of the book). Typically, the basic results in linear dynamics either use a Baire
category argument, in which case they are often valid in all F-spaces (see below), or
they hold in all topological vector spaces (see Chapter 12). Only in more specialized
results do structural properties such as being locally convex, being normable or having
an inner product play a role. Since our aim has not been to always provide the best
possible result but to offer a widely accessible introduction to the main ideas of linear
dynamics we have chosen to restrict ourselves to the setting of Fréchet spaces.
The larger class of F-spaces consists of all vector spaces that are endowed with an
F-norm and that are complete under the induced metric. For example, the spaces p
with 0 < p < 1 are F-spaces. One can show that a vector space is an F-space if and only
if it carries a complete translation-invariant metric; see [212].
We finish with a citation that is representative of the common, and as we now know
erroneous, belief that chaos and nonlinearity go hand in hand.
Chaotic systems not only exhibit sensitive dependence, but two other properties
as well: they are deterministic, and they are nonlinear.
(L.A. Smith, Chaos: A very short introduction, [295, p. 1]; emphasis in the original.)
Chapter 3
The Hypercyclicity Criterion
The message of our first criterion is quite simple: a large supply of (appropri-
ate) eigenvectors is conducive to chaos. This extends the earlier observation
that linked periodic points with certain eigenvectors of modulus 1; see Propo-
sition 2.33.
The criterion, like others in this chapter, is named after its authors.
m
m
1
T nx = ak λnk xk → 0 and un := bk yk → 0
μnk
k=1 k=1
n ≥ N,
x + un ∈ U and T n (x + un ) = T n x + y ∈ V.
This shows that T is mixing and therefore hypercyclic.
According to Proposition 2.33, in the complex case, Z0 is precisely the set
of periodic points of T . Thus, if also Z0 is dense then T is even chaotic.
The Godefroy–Shapiro criterion provides a new proof for the chaotic be-
haviour of the three classical operators.
Example 3.2. (Rolewicz’s operators) Let T = μB, |μ| > 1, be the multiple
of the backward shift on any of the spaces X = p , 1 ≤ p < ∞, or X = c0 .
By Corollary 2.51 (see also Example 2.27), it suffices to consider the complex
case.
One easily determines the eigenvectors of B as being the nonzero multiples
of the sequences
eλ := (λ, λ2 , λ3 , . . .), |λ| < 1,
with corresponding eigenvalue λ; the condition |λ| < 1 ensures that eλ ∈ X.
Therefore, eλ is an eigenvector of T = μB to the eigenvalue μλ.
We claim that, for any subset Λ of the unit disk D = {λ ∈ C ; |λ| < 1}
that has an accumulation point inside the disk, the set
span{eλ ; λ ∈ Λ}
The essential point in the mixing part of the proof of Theorem 3.1 was
the fact that T n x → 0, for each x ∈ X0 , and that for each y ∈ Y0 we could
find a sequence (un )n in X such that un → 0 and T n un = y for n ≥ 0. The
second condition can be achieved, for example, via a map S : Y0 → Y0 such
that S n y → 0 and T Sy = y for any y ∈ Y0 ; one need only define un = S n y.
This leads us to another important criterion in linear dynamics.
Remark 3.5. We emphasize that the map S need not have any properties
other than satisfying (ii) and (iii). In particular, it need not be a linear or
continuous map. The same remark will apply to the forthcoming criteria.
72 3 The Hypercyclicity Criterion
Example 3.8. (Birkhoff’s operators) For simplicity we only show that the
translation operator T1 f (z) = f (z + 1) on H(C) is mixing. For X0 = Y0
we choose the set of all functions of the form fp,α,ν (z) = p(z)e−α(z−ν) ,
2
|λ|n 1
x = |x0 | 1 + + = ∞,
n+1 |λ|n (n + 1)
n∈N n∈N
T nk (xj + S nk yj ) = T nk xj + yj , j = 1, 2.
It follows from (i) and (ii) that, for sufficiently large k, xj + S nk yj ∈ Uj and
T nk xj + yj ∈ Vj for j = 1, 2. This shows that N (U1 , V1 ) ∩ N (U2 , V2 ) = ∅.
The requirements of this criterion are clearly weaker than those of Kitai’s
criterion. In fact we will provide an operator satisfying the Gethner–Shapiro
criterion but not Kitai’s criterion. This is also the first example that we
present of a weakly mixing operator that is not mixing; see also Remark
4.10.
(see also Section 4.1); note that the value of w1 is irrelevant. Since w is
bounded, T is continuous. Let (mk )k be the increasing sequence of all the
integers with wmk = 2−1 and wmk +1 = 2, k ≥ 1. Since
n+1
T nx = wν xn+1 , . . . , n ≥ 1,
ν=2
j
j
S nk ej = S nk wν S j−1 e1 = wν S j−1 (S nk e1 ) → 0.
ν=2 ν=2
(i) T nk x → 0,
(ii) Snk y → 0,
(iii) T nk Snk y → y,
then T is weakly mixing, and in particular hypercyclic.
Originally, Kitai had shown that the conditions in Theorem 3.4 imply
hypercyclicity. She did this by constructing, through a recursive procedure, a
hypercyclic vector. While the proofs we have presented here are shorter and
give better results, Kitai’s construction is useful in many other situations
where more abstract methods fail. We therefore give here a second proof of
the Hypercyclicity Criterion using Kitai’s approach.
x = x1 + Snk1 y1 + x2 + Snk2 y2 + x3 + . . .
Indeed, for j = 1 these inequalities are satisfied for x1 = 0 and for a suitable
k1 , using conditions (ii) and (iii). Now let j ≥ 2 and assume that x1 , . . . , xj−1
and k1 , . . . , kj−1 have been constructed. Then, by density of X0 , there is some
xj ∈ X such that (3.1) holds and
j−1
(xl + Snkl yl ) + xj ∈ X0 .
l=1
76 3 The Hypercyclicity Criterion
Choosing kj sufficiently large, conditions (i), (ii) and (iii) imply that one can
also achieve (3.2) and (3.3).
It then follows from the first inequalities in (3.1) and (3.2) that the series
defining x converges. The remaining inequalities imply that, for j ≥ 1,
j−1
nkj nkj
T x − yj = T (xl + Snkl yl ) + xj + T nkj Snkj yj − yj
l=1
∞
∞
+ T nkj xl + T nkj Snkl yl
l=j+1 l=j+1
∞
1 1 1 1
4
≤ j + j + + = j.
2 2 2l 2l 2
l=j+1
T nk (Uj ) ∩ Vj = ∅ for j = 1, . . . , k,
T mk (U ) ∩ V ⊃ T mk (Ul ) ∩ Vl = ∅
T nk y ∈ kUk , k ≥ 1.
T nk (T n x) = T n (T nk x) → T n 0 = 0, n ≥ 0,
which gives condition (i) of the Hypercyclicity Criterion. On the other hand,
let y ∈ Y0 . Then there is some n ≥ 0 such that y = T n x, and we define
Snk y = T n xk , k ≥ 1. Then we have that
Snk y = T n xk → 0, T nk Snk y = T n (T nk xk ) → T n x = y,
A careful look at the proof of Theorem 3.15 yields a more precise connec-
tion between hereditary hypercyclicity and the Hypercyclicity Criterion.
78 3 The Hypercyclicity Criterion
In view of the equivalence of the weak mixing property with the Hyper-
cyclicity Criterion, Theorem 2.48 and Corollary 2.49 provide us with a rich
source of operators that satisfy the Hypercyclicity Criterion.
Theorem 3.18. Every hypercyclic operator with a dense set of points with
bounded orbits satisfies the Hypercyclicity Criterion.
In particular, any of the following operators satisfy the Hypercyclicity Cri-
terion:
(i) chaotic operators;
(ii) hypercyclic operators that have a dense set of points for which the orbits
converge;
(iii) hypercyclic operators with dense generalized kernel.
In the last two sections we have discussed some popular criteria for hyper-
cyclicity. But many other criteria have been suggested, some formally stronger
than the Hypercyclicity Criterion, some formally weaker. So the problem
arises of determining whether they are indeed stronger or weaker than the
Hypercyclicity Criterion. We address this problem here for two particular
criteria, one of which we have already encountered. Additional results along
this line will be discussed in the exercises.
Our first result says that, in the Hypercyclicity Criterion, one may, without
loss of generality, assume that the sets X0 and Y0 coincide, are linear, and
that each Snk is linear.
{y ∈ X1 ; d1 (x, y) < ε} ⊂ U.
Set x1,1 = x. Since f1 has dense range, there are x1,2 ∈ X1 and x2,2 ∈ X2
such that f1 (x2,2 ) = x1,2 and d1 (x1,1 , x1,2 ) < ε/2. Proceeding by induction
we will find a countable family A = {xj,k ∈ Xj ; k ∈ N, 1 ≤ j ≤ k} satisfying
ε
fj (xj+1,k+1 ) = xj,k+1 and dj (xj,k , xj,k+1 ) < , k ∈ N, 1 ≤ j ≤ k.
2k
Indeed, suppose that we already have constructed the finite family {xj,k ∈
Xj ; 1 ≤ k ≤ n, 1 ≤ j ≤ k} with the above property. Since fn has dense
range and each fj , j < n, is continuous, we can find xn,n+1 ∈ fn (Xn+1 ) close
enough to xn,n so that, for xj,n+1 := fj (xj+1,n+1 ), j < n, we have
ε
dj (xj,n , xj,n+1 ) < , 1 ≤ j ≤ n.
2n
The induction process is completed if we pick xn+1,n+1 ∈ Xn+1 with
fn (xn+1,n+1 ) = xn,n+1 .
The defining property of the family A yields that for each j ∈ N the
sequence (xj,k )k≥j is a Cauchy sequence in Xj . We define xj = limk→∞ xj,k ∈
Xj , j ∈ N, to conclude that
k−1
k−1
ε
d1 (x, x1 ) = lim d1 (x1,1 , x1,k ) ≤ lim d1 (x1,l , x1,l+1 ) < lim = ε,
k→∞ k→∞ k→∞ 2l
l=1 l=1
so that x1 ∈ U , and
T (x1 , x2 , x3 , . . . ) := (T x1 , T x2 , T x3 , . . . ) = (T x1 , x1 , x2 , . . . ).
U = {(xn )n ∈ X ; xn ∈ Un , n = 1, . . . , N },
V = {(xn )n ∈ X ; xn ∈ Vn , n = 1, . . . , N }
with N ∈ N and nonempty open sets Un , Vn ⊂ X, n = 1, . . . , N .
Fix x = (xn )n ∈ U and y = (yn )n ∈ V. By continuity of T we can find
open neighbourhoods UN ⊂ UN , VN ⊂ VN of xN and yN , respectively, such
that T (UN ) ⊂ UN −j and T j (VN ) ⊂ VN −j , j = 1, . . . , N − 1.
j
T mk un = T mk (T N −n uN ) = T N −n (T mk uN ) ∈ T N −n (VN ) ⊂ Vn ,
82 3 The Hypercyclicity Criterion
and therefore T mk u ∈ V.
Since (yn )n ∈ X we also have that T Sy = y for all y ∈ Y0 . Finally, let (qk )k
be a subsequence of (mk )k such that Bqk y = (y1+qk , y2+qk , . . . ) → 0 in X .
Then
S qk yn = yn+qk → 0, n ∈ N.
Topological transitivity, mixing and weak mixing for (Tn )n are defined in
Definition 1.56. A notion of chaos can be defined, but seems less natural in
this context.
Now, the Birkhoff transitivity theorem extends to commuting sequences
of operators Tn : X → X with dense range; see Exercise 1.6.2. For general
sequences of operators one should instead turn to the Universality Criterion;
see Theorem 1.57.
The results of Sections 2.4 and 2.5 carry over to commuting sequences of
operators Tn : X → X (if, in Theorem 2.45, the map S : X → X commutes
with each Tn ).
The Hypercyclicity Criterion extends, with the previous proof, to arbitrary
sequences of operators.
84 3 The Hypercyclicity Criterion
The (simple) example of Exercise 1.6.1 shows that not every hypercyclic
sequence of operators satisfies the Hypercyclicity Criterion.
The Bès–Peris theorem extends likewise to commuting sequences. It is also
of interest to add the notion of hereditary transitivity; see Exercise 1.6.4.
The result breaks down for noncommuting operators; see Exercise 3.4.4.
Exercises
Exercise 3.1.1. Prove the following version of Kitai’s criterion. If there are dense sub-
sets X0 , Y0 ⊂ X such that
(i) T n x → 0 for any x ∈ X0 ,
(ii) for any y ∈ Y0 there is a sequence (un )n in X such that un → 0 and T n un → y,
then T is a mixing operator.
Exercise 3.1.3. Show that an operator T satisfies Kitai’s criterion if and only if there
is a syndetic increasing sequence (nk )k of positive integers such that T satisfies the
Gethner-Shapiro criterion with respect to (nk )k .
Exercise 3.1.4. Let C : Lp [0, 1] → Lp [0, 1], 1 < p < ∞, be the Cesàro operator
t
defined as Cf (t) = 1t 0 f (s) ds. Show that C is mixing and chaotic. (Hint: Use the
Hahn–Banach theorem to show the density of span{tα ; α ∈ A} in Lp [0, 1], for any set
A ⊂ H := {z ∈ C ; Re z > −1/p} with accumulation points in H. Then apply the
Godefroy–Shapiro criterion.)
Exercise 3.2.1. Give a new proof of Theorem 3.10 based on Theorem 2.47.
Exercises 85
Show that Bw is weakly mixing but not mixing. (Hint: Proceed as in Example 3.11;
note that this time S is not continuous, so calculate S nk ej directly.)
Exercise 3.2.5. Show that every weakly mixing operator T satisfies the Hypercyclicity
Criterion, without using the hereditary transitivity property of T . To do this, fix a
hypercyclic vector (x, y) for T ⊕ T and proceed as follows:
(i) for every k ≥ 1, the vector (x, T k y) is also hypercyclic for T ⊕ T ;
(ii) if Uk is the open ball of radius 1/k around 0, k ≥ 1, then there is an increasing
sequence (nk )k of positive integers and a sequence of vectors xk ∈ Uk ∩ orb(y, T )
such that T nk x ∈ Uk and T nk xk ∈ x + Uk , k ≥ 1;
(iii) the dense set X0 = Y0 := orb(x, T ), the sequences (nk )k and the maps Snk :
Y0 → X, Snk (T n x) := T n xk , k, n ≥ 1, satisfy the hypotheses of the Hypercyclicity
Criterion.
Exercise 3.2.6. Show that every hypercyclic operator on the space ω = CN of all com-
plex sequences satisfies the Hypercyclicity Criterion. (Hint: Use the Herzog–Lemmert
theorem, Theorem 2.65.)
Exercise 3.3.1. Prove that an operator T satisfies the Hypercyclicity Criterion if and
only if there exists an increasing sequence (nk )k of positive integers such that the fol-
lowing two conditions are satisfied:
(i) there exists a dense subset X0 ⊂ Xsuch that T nk x → 0 for any x ∈ X0 ;
∞
(ii) for any 0-neighbourhood W of X, k=1 T nk (W ) is dense in X.
∞particular case when X is a Banach space, the second condition can be simplified
In the
to k=1 T nk (BX ) being dense in X, where BX denotes the unit ball of X.
Exercise 3.3.2. A subset A of a metric space is precompact if for every ε > 0 the set A
can be covered by a finite union of open balls of radius ε in the space. If the metric space
is complete, then any sequence in a precompact set admits a convergent subsequence.
Let T be an operator satisfying the following criterion: there are dense subsets
X0 , Y0 ⊂ X, an increasing sequence (nk )k of positive integers, and maps Snk : Y0 → X,
k ∈ N, such that, for any x ∈ X0 , y ∈ Y0 ,
(i) the set {T nk x ; k ∈ N} is precompact,
(ii) Snk y → 0,
(iii) T nk Snk y → y.
Show that T satisfies the Hypercyclicity Criterion. (Hint: Prove that T is weakly mixing.)
86 3 The Hypercyclicity Criterion
Exercise 3.3.3. Consider the bilateral backward shift T = B on the weighted space
1 (Z, v) = {(xn )n∈Z ; x := n∈Z
|xn |vn < ∞}, where vn = n1 , n ≥ 1, and vn = 1,
n ≤ 0. Show that T is not hypercyclic. Deduce that in condition (i) of Exercise 3.3.2
one cannot replace precompactness by boundedness.
Exercise 3.3.4. Repeat Exercise 3.3.2 with conditions (i) and (ii) replaced by
(i) T nk x → 0 for any x ∈ X0 ,
(ii) the set {Snk y ; k ∈ N} is precompact for any y ∈ Y0 .
Exercise 3.3.7. Prove the following strengthened version of the Mittag-Leffler theorem.
Let (Xn )n be a sequence of complete metric spaces, let Un ⊂ Xn , n ∈ N, be dense open
sets and let fn : Xn+1 → Xn , n ∈ N, be continuous maps with dense range. Then,
for every nonempty open subset U ⊂ X1 , there exists a sequence (xn )n with xn ∈ Un ,
n ∈ N, such that x1 ∈ U and fn (xn+1 ) = xn , n ∈ N.
Deduce from this the Baire category theorem.
Exercise 3.4.1. Two operators T1 , T2 on a separable Fréchet space X are called disjoint
hypercyclic if there is some x ∈ X such that {(T1n x, T2n x) ; n ≥ 0} is dense in X ⊕ X.
(a) Confirm the following obvious facts. No operator T is disjoint hypercyclic with
itself. The operators T1 , T2 are disjoint hypercyclic if and only if some (x, x) ∈ X ⊕ X
is hypercyclic for T1 ⊕ T2 , and if and only if the sequence (Tn )n of operators Tn : X →
X ⊕ X, Tn x = (T1n x, T2n x) is hypercyclic.
(b) Derive a corresponding transitivity theorem. The operators T1 , T2 are disjoint
hypercyclic if, for any nonempty open subsets U, V1 , V2 of X, there exists some n ≥ 0
such that T1n (U ) ∩ V1
= ∅ and T2n (U ) ∩ V2
= ∅.
(c) Derive a corresponding Disjoint Hypercyclicity Criterion. If there are dense
subsets X0 , Y1 , Y2 ⊂ X, an increasing sequence (nk )k of positive integers, and maps
Sj,nk : Yj → X, k ≥ 1, j = 1, 2, such that, for any x ∈ X0 , y1 ∈ Y1 , y2 ∈ Y2 ,
(i) Tjnk x → 0, j = 1, 2,
(ii) Sj,nk yj → 0, j = 1, 2,
(iii) Tjnk Sj,nk yj → yj , j = 1, 2,
(iv) Tlnk Sj,nk yj → 0, j = 1, l = 2 or j = 2, l = 1,
then T1 , T2 are disjoint hypercyclic.
Exercise 3.4.2. Show that the following operators are disjoint hypercyclic:
(i) λB, μB 2 , where 1 < |λ| < |μ| and B is the backward shift on any of the spaces p ,
1 ≤ p < ∞, or c0 ;
(ii) Ta , Tb , where a, b ∈ C, a
= 0, b
= 0, a
= b, and Ta is the translation operator on
H(C) given by Ta f (z) = f (z + a).
Exercise 3.4.6. Let B be the backward shift operator on 2 and consider the operators
Tn : 2 → 2 , Tn x = 2n B n x − x, n ≥ 1. Show that (Tn )n is mixing and that the only
sequences x for which Tn x → 0 are multiples of the sequence ( 21n )n . Deduce that the
implication (ii) =⇒ (i) does not hold in the previous exercise.
Sections 3.1 and 3.2. The earliest forms of the Hypercyclicity Criterion were found
independently by Kitai [215] (who in addition demanded that X0 = Y0 ) and by Gethner
and Shapiro [161, Remark 2.3(b)]. In its general form it is due to Bès and Peris [71]. The
Godefroy–Shapiro criterion is contained implicitly in their paper [165] and was isolated
by Bernal [55].
Intended originally as simple tools for obtaining the hypercyclicity of an opera-
tor, these criteria have become tremendously important for the understanding of lin-
ear dynamics. While the Hypercyclicity Criterion, very unexpectedly, turned out to be
equivalent to the weak mixing property and therefore to be close to hypercyclicity, the
Godefroy–Shapiro criterion was further developed by Bayart and Grivaux in their study
of frequent hypercyclicity; see Chapter 9.
88 3 The Hypercyclicity Criterion
Grivaux [172] showed that there is a mixing operator T for which only x = 0 satisfies
that T n x → 0 as n → ∞. Thus, not every mixing operator satisfies the Hypercyclicity
Criterion for the full sequence (n), nor Kitai’s criterion; see Exercise 3.2.3.
We remark that the main idea of Example 3.8, that of considering the functions
z → p(z)e−αz , was taken from Birkhoff’s original proof in [75].
2
Theorem 3.15 is due Bès and Peris [71], as is Proposition 3.19. The notion of hered-
itary hypercyclicity with respect to the full sequence was introduced by Shapiro [279]
(who called it strong hypercyclicity) and by Ansari [9].
Section 3.3. Theorem 3.22 is due to Peris [253]; for the abstract Mittag-Leffler theorem
we refer to Arens [11, Theorem 2.4]; see also Esterle [145] for an interesting discussion.
It seems to be an open problem whether the strengthened conditions of the Gethner–
Shapiro criterion and of Proposition 3.20 can be combined, that is, if every operator
satisfying the Hypercyclicity Criterion admits a dense subspace X0 ⊂ X, an increasing
sequence (nk )k of positive integers, and a linear map S : X0 → X0 such that, for any
x ∈ X0 , T nk x → 0, S nk x → 0, and T Sx = x.
Section 3.4. Much closer studies of the Hypercyclicity Criterion for sequences of op-
erators were undertaken by Bernal and Grosse-Erdmann [62], Bermúdez, Bonilla, and
Peris [51] and León and Müller [223].
Exercises. For Exercise 3.1.3 we refer to Bermúdez, Bonilla, Conejero, and Peris [49].
Exercise 3.1.4 is taken from León, Piqueras and Seoane [224]; the continuity of the
Cesàro operators is due to Hardy’s inequality [193]. We will return to these operators
in Example 12.20. Exercise 3.2.5 follows the original argument in Bès and Peris [71].
Exercise 3.3.1 is taken from León and Müller [223]; see also Müller [247]. Exercises 3.3.2,
3.3.4 and 3.3.5 are due to Bermúdez, Bonilla, and Peris [51]. The notion of disjoint
hypercyclic operators is due to Bernal [58] and Bès and Peris [72], where one can also
find the results of Exercises 3.4.1 and 3.4.2. The example in Exercise 3.4.4 is a special
case of a class of operators studied by Bernal [56]. For Exercise 3.4.9 we refer to Aron,
Bès, León and Peris [14], for Exercise 3.4.10 to Bernal and Calderón [61]. Concerning
the latter exercise, it is interesting to note that by a remarkable result by Grivaux [169],
any countable family of hypercyclic operators on a Banach space has a common dense
subspace of hypercyclic vectors, except for 0.
Extensions. Repeating a remark we made in Chapter 2 we note that the results in this
chapter remain valid in separable F-spaces. For further extensions we refer to Chapter
12.
Chapter 4
Classes of hypercyclic and chaotic
operators
In order to distinguish this shift from the bilateral shift that we will discuss
later one also speaks of the unilateral backward shift.
Rolewicz has shown that, for any λ with |λ| > 1, the multiples of B,
λB(xn )n = (λxn+1 )n , are hypercyclic on the sequence space 2 . It is then a
small step to let the weights vary from coordinate to coordinate, which leads
to the (unilateral) weighted shift
where
w = (wn )n
is a sequence of nonzero scalars, called a weight sequence. Note that the value
of w1 is irrelevant.
We may also generalize these operators in a different direction. Rolewicz
had already replaced 2 by any of the spaces p , 1 ≤ p < ∞, and c0 . More
generally, one may take as the underlying space an arbitrary sequence space
X, that is, a linear space of sequences or, in other words, a subspace of
ω = KN . In addition, X should carry a topology that is compatible with
the sequence space structure of X. We interpret this as demanding that the
embedding X → ω is continuous, that is, convergence in X should imply co-
ordinatewise convergence. A Banach (Fréchet, . . . ) space of this kind is called
a Banach (Fréchet, . . . ) sequence space. The terms of a sequence x, y, z, . . .
will be denoted by xn , yn , zn , . . ., n ≥ 1.
By en , n ∈ N,
Bw en = wn en−1 , n ≥ 1, with e0 := 0.
X → K, x → xn , n ≥ 1,
which implies that each weighted shift has closed graph. From the closed
graph theorem (see Appendix A) we thus obtain that a weighted shift defines
an operator on a Fréchet sequence space X as soon as it maps X into itself.
Proposition 4.1. Let X be a Fréchet sequence space. Then every weighted
shift Bw : X → X is continuous.
We start by studying the (unweighted) backward shift B. Our results will
then extend immediately to all weighted shifts via a simple conjugacy.
The following technical result will help us simplify the condition charac-
terizing hypercyclicity of B.
Lemma 4.2. Let X be a metric space, vn ∈ X, n ≥ 1, and v ∈ X. Suppose
that there is a strictly increasing sequence (nk )k of positive integers such that
Proof. Let · stand for an F-norm that induces the topology of X; see
Section 2.1.
(i)=⇒(iii). Suppose that B is hypercyclic. Let N ∈ N and ε > 0. We show
that there exists some n ≥ N with en < ε.
It follows from the basis assumption that, for any x ∈ X, the sequence
(xn en )n converges to 0 in X. By the Banach–Steinhaus theorem (see Ap-
pendix A), applied to the operators x → xn en , n ≥ 1, there is some δ > 0
such that, for all x ∈ X,
x < δ =⇒ xn en < ε
2 for all n ≥ 1. (4.1)
xn en < ε
2 and |xn − 1| ≤ 12 ; (4.3)
enk −j = B j enk → 0 as k → ∞,
for all j ≥ 1. Since (nk )k must be strictly increasing, it follows from Lemma
4.2 that there is an increasing sequence (mk )k of positive integers such that
emk +j → 0 as k → ∞,
Smk y → 0
x ∈ p (v),
∞ 1/p
∞ 1/p
|xn+1 |p vn ≤M |xn |p vn ,
n=1 n=1
vn
which is equivalent to supn∈N vn+1 < ∞. Theorem 4.3 tells us that, under
this condition, the hypercyclicity of B is characterized by
inf vn = 0.
n∈N
where we have applied Parseval’s identity in L2 [0, 2π] for the orthonormal
basis ( √12π eint )n∈Z . As a consequence, A2 is isometrically isomorphic to the
1
weighted space 2 ( n+1 ) (with indices running from 0). By (a), the backward
shift is therefore hypercyclic on A2 . When acting on functions, B is the
operator
∞
1
Bf (z) = an+1 z n = (f (z) − f (0)) with Bf (0) = f (0).
n=0
z
(c) As in (b) we can consider the space H(C) of entire functions (see Ex-
ample
∞ 2.1)n as a sequence space by identifying the entire function f (z) =
n=0 an z with the sequence (an )n≥0 . By the formula for the radius of con-
vergence of Taylor series, this sequence space is given by
∞
(an )n≥0 ; lim |an |1/n = 0 = (an )n≥0 ; |an |mn < ∞, m ≥ 1 .
n→∞
n=0
Since |an+1 |1/n = (|an+1 |1/(n+1) )(n+1)/n → 0 if |an |1/n → 0, we have that
the backward shift B is an operator on H(C); see Proposition 4.1. Moreover,
the unit sequences en correspond to the monomials z → z n , n ≥ 0. It then
follows from Theorem 4.3 that B is not hypercyclic on H(C).
Using the same arguments as in the proof of Theorem 4.3, but employing
Kitai’s criterion instead of the Hypercyclicity Criterion, we obtain a charac-
terization of the mixing property for B.
Applying the backward shift N − 1 times and adding the results we obtain
(iii).
(iii)=⇒(ii) follows from our assumptions.
(ii)=⇒(i). First, by Theorem 4.3, condition (ii) implies that B is hyper-
cyclic.
Next, since (1, 1, 1, . . .) ∈ X, the unconditionality of the basis implies that
all the periodic 0-1-sequences belongs to X, and hence also all the periodic
sequences, which are exactly the periodic points for B. It remains to show
that these form a dense set in X.
To see this, let x = (xn )n ∈ X and ε > 0. Since (en )n is a basis, there is
some N ≥ 1 such that
N
x
:= xn e n
n=1
has distance less than ε/2 from x. The associated periodic sequence
∞
N
xn en+νN
ν=0 n=1
for any 0-1-sequence (εn )n ; see Theorem A.16. In particular we have that
∞ N ε
xn en+μmN < .
μ=1 n=1
2
and only if
lim vn = 0,
n→∞
in X as k → ∞.
(b) The following assertions are equivalent:
(i) Bw is mixing;
(ii) we have that
n −1
wν en → 0
ν=1
in X as n → ∞.
(c) Suppose that the basis (en )n is unconditional. Then the following asser-
tions are equivalent:
(i) Bw is chaotic;
(ii) the series
∞ n −1
wν en
n=1 ν=1
converges in X;
(iii) the sequence
n −1
wν
n
ν=1
belongs to X;
(iv) Bw has a nontrivial periodic point.
n
n ∞
1
sup |wν | = ∞, lim |wν | = ∞, n < ∞.
ν=1 |wν |
n→∞ p
n≥1 ν=1 ν=1 n=1
We remark that only the third condition depends on the parameter p. The
first condition also characterizes when Bw is hypercyclic on c0 , and the second
when it is mixing or, equivalently, chaotic on c0 .
nIn particular, for Rolewicz’s operator T = λB, |λ| > 1, we have that
ν=1 |wν | = λ , which implies once more that this operator is chaotic.
n
n
n
lim inf |wν | < lim sup |wν | = ∞
n→∞ n→∞
ν=1 ν=1
Fw (x1 , x2 , x3 , . . .) = (0, w1 x1 , w2 x2 , . . .)
with a weight sequence w = (wn )n . The first coordinate of every point in the
orbit of x is either x1 or 0. By the assumption that convergence in the space
implies coordinatewise convergence no orbit can be dense.
Some new and interesting phenomena arise when we study shifts on se-
quence spaces indexed over Z. The bilateral backward shift is given by
en = (δn,k )k∈Z , n ∈ Z,
Theorem 4.12. Let X be a Fréchet sequence space over Z in which (en )n∈Z
is a basis. Suppose that the bilateral shift B is an operator on X.
(a) The following assertions are equivalent:
(i) B is hypercyclic;
(ii) B is weakly mixing;
(iii) there is an increasing sequence (nk )k of positive integers such that,
for any j ∈ Z, ej−nk → 0 and ej+nk → 0 in X as k → ∞.
(b) The following assertions are equivalent:
(i) B is mixing;
(ii) e−n → 0 and en → 0 in X as n → ∞.
(c) Suppose that the basis (en )n is unconditional. Then the following asser-
tions are equivalent:
(i) B
∞ is chaotic;
(ii) n=−∞ en converges in X;
(iii) the constant sequences belong to X;
(iv) B has a nontrivial periodic point.
x < δ =⇒ xn en < ε
2 (n ∈ Z) and |xj | ≤ 1
2 (|j| ≤ N ). (4.5)
100 4 Classes of hypercyclic and chaotic operators
xn en < ε
2
(|n| > N ) and |xn+j − 1| ≤ 1
2
(|j| ≤ N ),
hence
|xj − 1| ≤ 1
2
(|j| ≤ N ) and xn+k ek < ε
2
(|k| > N ),
hence
(2xj )−1 ≤ 1 (|j| ≤ N ) and xj ej−n < ε
(|j| ≤ N ),
2
whence
ej−n = (2xj )−1 2xj ej−n < ε for |j| ≤ N .
(iii)=⇒(ii). One need only observe that for the forward shift
B nk ej = ej−nk → 0, F nk ej = ej+nk → 0,
B
Xv −−−−→ Xv
⏐ ⏐
⏐
φv
⏐φ
v
B
X −−−w−→ X,
where
Xv = {(xn )n∈Z ; (xn vn )n ∈ X}
and φv : Xv → X, (xn )n∈Z → (xn vn )n∈Z with
n −1
0
vn = wν for n ≥ 1, vn = wν for n ≤ −1, v0 = 1.
ν=1 ν=n+1
Theorem 4.13. Let X be a Fréchet sequence space over Z in which (en )n∈Z
is a basis. Suppose that the weighted shift Bw is an operator on X.
(a) The following assertions are equivalent:
(i) Bw is hypercyclic;
(ii) Bw is weakly mixing;
(iii) there is an increasing sequence (nk )k of positive integers such that,
for any j ∈ Z,
j j+n
k −1
wν ej−nk → 0 and wν ej+nk → 0
ν=j−nk +1 ν=j+1
in X as k → ∞.
(b) The following assertions are equivalent:
(i) Bw is mixing;
(ii) we have that
0
n −1
wν e−n → 0 and wν en → 0
ν=−n+1 ν=1
in X as n → ∞.
(c) Suppose that the basis (en )n∈Z is unconditional. Then the following as-
sertions are equivalent:
(i) Bw is chaotic;
(ii) the series
0
0 ∞
n −1
wν en + wν en
n=−∞ ν=n+1 n=1 ν=1
converges in X;
(iii) the sequence (xn )n∈Z with
102 4 Classes of hypercyclic and chaotic operators
0
n −1
xn = wν (n ≤ 0), xn = wν (n ≥ 1)
ν=n+1 ν=1
belongs to X;
(iv) Bw has a nontrivial periodic point.
Remark 4.14. In the bilateral case, forward shifts can be hypercyclic. A bi-
lateral weighted forward shift is given by an operator
j−1 −1 j+n
k −1
wν ej−nk → 0 and wν ej+nk → 0
ν=j−nk ν=j
j k
j+n
∃(nk )k ∀j ∈ Z : lim wν = 0 and lim |wν | = ∞;
k→∞ k→∞
ν=j−nk +1 ν=j+1
0
n
lim wν = 0 and lim |wν | = ∞;
n→∞ n→∞
ν=−n+1 ν=1
∞
0 ∞
1
|wν |p < ∞ and n < ∞.
ν=1 |wν |
p
n=0 ν=−n+1 n=1
In particular, a symmetric weight (that is, one with w−n = wn for all n ≥ 0)
never defines a hypercyclic weighted shift Bw on these spaces.
As a concrete example, the weight
w = . . . , 12 , 12 , 12 , 2, 2, 2, . . .
4.1 Weighted shifts 103
then Theorem 4.13 and Remark 4.14 tell us that Bw and F(wn+1 ) are weakly
mixing if and only if there are increasing sequences (nk )k and (mk )k of pos-
itive integers such that, for any j ∈ Z,
vj−nk → 0, vj+nk → 0,
vj−mk → ∞, vj+mk → ∞,
and the nk are the indices of the local minima, the mk the indices of the local
maxima of this sequence. Note that Bw is even invertible.
Remark 4.17. This proposition provides us with an example of two weakly
mixing, hence hypercyclic operators S, T : X → X whose direct sum S ⊕ T
is not hypercyclic.
We show more generally that for any operator T on a Banach space X
the operator T ⊕ T ∗ cannot be hypercyclic on X ⊕ X ∗ . Indeed, suppose
104 4 Classes of hypercyclic and chaotic operators
which is impossible since the left-hand side must be dense in K; note that
(x∗ , −x) cannot be the zero vector.
on the space H(C) of entire functions have little in common. But there is a
surprisingly simple connection. Since
∞ ∞
f (n) (z) n an Dn f
f (z + a) = a = (z)
n=0
n! n=0
n!
Ta = eaD .
In fact, this representation can be justified rigorously. We will need the fol-
lowing notion from complex analysis: an entire function ϕ is said to be of
exponential type if there are constants M, A > 0 such that
Setting ρ = n/A and using Stirling’s formula we get, with some C > 0,
√ n
M An n nA (2A)n
|an | ≤ n
e ≤ CM ≤ CM .
n n! n!
Conversely, if (4.8) holds then
∞
∞ n
(R|z|)
|ϕ(z)| ≤ |an z n | ≤ M = M eR|z| ,
n=0 n=0
n!
Proof. Let f ∈ H(C) and |z| ≤ m. By the Cauchy estimates and Lemma 4.18
there are M, R > 0 such that
n
an f (n) (z) ≤ |an | n! max |f (ζ)| ≤ M R max |f (ζ)|. (4.9)
mn |ζ|≤2m m |ζ|≤2m
∞
Therefore, if m > R then n=0 an f
(n)
(z) converges uniformly on |z| ≤ m.
Hence
∞
ϕ(D)f = an D n f
n=0
T = a0 I + a1 D + . . . + am Dm .
(ii)=⇒(iii). By the same token, using (4.10), we obtain for f ∈ H(C) that
∞ ∞ ∞
an n an an n
T Ta f = T D f= T Dn f = D (T f ) = Ta T f.
n=0
n! n=0
n! n=0
n!
(T en )(0)
an =
n!
and deduce that
Rn
|an | ≤ M .
n!
∞
It follows from Lemma 4.18 and Proposition 4.19 that ϕ(z) = n=0 an z n
∞
defines an entire function ϕ of exponential type and that ϕ(D) = n=0 an Dn
defines an operator on H(C). Then
By what we have shown above we also know that ϕ(D) commutes with each
Ta . Thus we get with (iii) for any z ∈ C and f ∈ H(C), using the definition
of Tz ,
so that T = ϕ(D).
With this in hand we can prove a remarkably general common extension
of the theorems of Birkhoff and MacLane.
Theorem 4.21 (Godefroy–Shapiro). Suppose that T : H(C) → H(C),
T = λI, is an operator that commutes with D, that is,
T D = DT.
eλ (z) = eλz , λ ∈ C,
we calculate that
∞
T eλ = ϕ(D)eλ = an λn eλ = ϕ(λ)eλ .
n=0
contains span{eλ ; |ϕ(λ)| < 1}, which is dense in H(C) by Lemma 2.34;
indeed, since any nonconstant entire function has dense range (see Appendix
A), {λ ∈ C ; |ϕ(λ)| < 1} is a nonempty open set and therefore has an
accumulation point. For the same reason, the eigenvectors of T to eigenvalues
μ with |μ| > 1 span a dense set in H(C). For the density of
it suffices to observe that also the set {λ ∈ C ; ϕ(λ) = eαπi for some α ∈ Q}
has an accumulation point. Indeed, since ϕ(C) is connected and dense, it
must intersect the unit circle. And since nonconstant holomorphic functions
are open mappings, infinitely many preimages under ϕ of roots of unity lie
in some bounded subset of C and therefore have an accumulation point.
Having established the hypercyclicity of every differential operator T =
ϕ(D) = λI we now want to focus our attention on properties of the cor-
responding hypercyclic functions. MacLane had already addressed such a
problem: he showed that there exists a D-hypercyclic entire function f of
exponential type 1, which means that for every ε > 0 there is some M > 0
such that
|f (z)| ≤ M e(1+ε)r for all z ∈ C.
Here we follow the usual convention of writing r = |z|. MacLane’s growth
condition can be improved, and one can even determine the least possible
rate of growth.
Theorem 4.22. (a) Let φ : ]0, ∞[ → [1, ∞[ be a function with φ(r) → ∞ as
r → ∞. Then there exists an entire function f that is hypercyclic for D and
that satisfies
er
|f (z)| ≤ M φ(r) √ for |z| = r > 0
r
with some M > 0.
(b) There is no entire function f that is hypercyclic for D and that satisfies
er
|f (z)| ≤ M √ for |z| = r > 0
r
Proving our assertion then amounts to showing that the sequence of operators
admits a dense orbit in the sense of Section 3.4; note that the Tn are indeed
operators because the inclusion map X → H(C) is obviously continuous.
To prove that (Tn )n is hypercyclic we apply the Hypercyclicity Criterion
for sequences of operators, Theorem 3.24.
It is an easy exercise to see that X is a Banach space. We would like
to take as X0 the set of polynomials, but we cannot guarantee that the
polynomials are dense in X. Thus we replace X by the closure X 0 of X0 in
X. Clearly Tn f → 0 for any f ∈ X0 . For Y0 we take the set of polynomials
4.2 Differential operators 109
e r n+1/2
n
Sn e0 = = sup r
.
n! r>0 n!φ(r)e
r n+1/2 (n + 1/2)n+1/2
sup = ,
r>0 n!er n!en+1/2
Choosing ρ = n we get
(n) n!
f (0) ≤ M en ,
nn+1/2
which is bounded by Stirling’s formula. Thus, f cannot be hypercyclic for D.
√
Exercise 4.2.5 explains how the critical rate of growth er / r is related to
the differentiation operator.
In contrast to the result for MacLane’s operator, entire functions that are
hypercyclic for Birkhoff’s operators can grow arbitrarily slowly. The proof
requires a different technique and will be provided in Chapter 8; see Exercise
8.1.3.
Theorem 4.23 (Duyos-Ruiz). Let a = 0. Let φ : ]0, ∞[ → [1, ∞[ be a
function so that, for any N ≥ 1, φ(r)/rN → ∞. Then there exists an entire
function f that is hypercyclic for Ta and that satisfies
By the method used in the proof of Theorem 4.21 one can derive certain
possible rates of growth for an arbitrary operator ϕ(D); see Exercise 4.2.4.
Ta f = f ◦ τ a .
ϕ : Ω → Ω;
pn (f ) = sup |f (z)|, n ∈ N.
z∈Kn
In this way H(Ω) turns into a Fréchet space; note that the topology is inde-
pendent of the chosen exhaustion. Moreover, by Runge’s theorem, H(Ω) is
separable; see Exercise 4.3.1.
Clearly, for any automorphism ϕ of Ω the composition operator Cϕ is con-
tinuous on H(Ω). Let us first note that conformal maps, that is, holomorphic
bijections between two domains, induce conjugacies between the correspond-
ing composition operators; the proof is immediate.
ϕn (K) ∩ K = ∅.
Example 4.28. (a) Let Ω = C. Then the automorphisms of C are the functions
ϕ(z) = az + b, a = 0, b ∈ C,
We first show that the run-away property is a necessary condition for the
hypercyclicity of the composition operator.
Proof. If (ϕn )n is not run-away then there exists a compact set K ⊂ Ω and
elements zn ∈ K such that
ϕn (zn ) ∈ K, n ∈ N. (4.11)
so that the functions (Cϕ )n f cannot approximate, for example, the constant
function M + 1 uniformly on K, a contradiction.
The following auxiliary result will be crucial for the proof of sufficiency in
Theorem 4.32. However, since its proof is rather technical we will postpone it
to the end of the section. For later use we formulate the lemma for arbitrary
sequences (ϕn )n of automorphisms.
114 4 Classes of hypercyclic and chaotic operators
hence
sup |f (z) − h(z)| < ε and sup |g(z) − h(ϕn (z))| < ε.
z∈L z∈L
Then
4.3 Composition operators I 115
Ω := D \ ϕn (A)
n∈Z
in particular, ϕm2 (K2 ) ∩ K2 = ∅; moreover, since ϕm2 (L) contains ϕm2 (K1 )
and L contains ϕk (K1 ) for k = 1, . . . , m1 , we must have that m2 > m1 .
Proceeding inductively we obtain a strictly increasing sequence (mn )n such
that ϕmn (Kn ) ∩ Kn = ∅, for any n ∈ N; as a consequence, (ϕmn )n and any
of its subsequences is run-away. The proofs in this section then show that
every subsequence of (Cϕmn )n admits a dense orbit. This tells us that Cϕ is
hereditarily hypercyclic, and hence weakly mixing by Theorem 3.15.
ϕ(z) = az + b, a = 0, b ∈ C,
1 − |a|2
1 − |w|2 = (1 − |z|2 ). (4.12)
|1 − az|2
Proof. The implications (vi)=⇒(iii) and (vi)=⇒(ii) follow from known prop-
erties of the Birkhoff operators, (iii)=⇒(i) and (ii)=⇒(i) hold for all operators
on H(Ω), and (i)⇐⇒(iv) was proved in Theorem 4.32.
(i)=⇒(v). If ϕ has a fixed point z0 ∈ Ω then, for any f ∈ H(Ω) and n ≥ 0,
((Cϕ )n f )(z0 ) = f (ϕn (z0 )) = f (z0 ), so that f cannot have a dense orbit.
It remains to prove that (v)=⇒(vi). In the case Ω = C the result was
shown in Example 4.28(a). In the case Ω = C we can assume by the discussion
leading up to Proposition 4.36 that Ω = D. The proof then requires certain
properties of linear fractional transformations. Since we will have occasion
to study them in Section 4.5 we will postpone the proof to the end of that
section.
Fig. 4.4 ϕmN +1 (KN ) ∪ KN is Ω-convex Fig. 4.5 Both ϕmN +1 (KN ) ∪ KN and
ϕmν (KN ) ∪ KN are Ω-convex
defines a holomorphic function on the complex unit disk D. The Hardy space
is then defined as the space of these functions, that is,
∞
∞
H 2 = f : D → C ; f (z) = an z n , z ∈ D, with |an |2 < ∞ .
n=0 n=0
In other words, the Hardy space is simply the sequence space 2 (N0 ), with
its elements written as holomorphic functions. It is then clear that H 2 is a
Banach space under the norm
∞ 1/2 ∞
f = |an |2 when f (z) = an z n ,
n=0 n=0
f (λ) = f, kλ .
f → f (λ)
Proof. It suffices to show that only the zero function can be orthogonal to
span{kλ ; λ ∈ Λ}. But that is immediate by the identity theorem for holo-
morphic functions: if, for f ∈ H 2 , f, kλ = f (λ) vanishes for all λ ∈ Λ, then
f = 0.
2π 1/2 2π 1/2
1 1
f = sup |f (reit )|2 dt = lim |f (reit )|2 dt
0≤r<1 2π 0 r
1 2π 0
and 2π
1
f, g = lim f (reit )g(reit ) dt.
r
1 2π 0
∞
Proof. Writing f (z) = n=0 an z n we obtain that
120 4 Classes of hypercyclic and chaotic operators
2π 2π ∞ 2
1 1
|f (reit )|2 dt = an (reit )n dt
2π 0 2π 0 n=0
2π ∞ 2 ∞
n 1 int
= an r √ e dt = |an |2 r2n ,
0 n=0
2π n=0
where we have used Parseval’s identity in L2 [0, 2π] for the orthonormal basis
( √12π eint )n∈Z ; see also Example 4.4(b). Since
∞
∞
∞
sup |an |2 r2n = lim |an |2 r 2n = |an |2 = f 2 ,
0≤r<1 n=0 r
1
n=0 n=0
the first part of the assertion follows. In the same way, one obtains the second
part by using Parseval’s identity for the inner product.
Now let ϕ be a bounded holomorphic function on D. Then, for any f ∈ H 2 ,
ϕf is holomorphic on D, and we have that
2π 2π
1 1
sup |(ϕf )(reit )|2 dt ≤ sup |ϕ(z)|2 sup |f (reit )|2 dt,
0≤r<1 2π 0 z∈D 0≤r<1 2π 0
Mϕ f = ϕf
the same would be true of the sequence (ϕ(0)n f (0))n≥0 in C, which is never
the case. Instead, we will consider the (Hilbert space) adjoint Mϕ∗ : H 2 → H 2
of Mϕ , called an adjoint multiplication operator or adjoint multiplier.
In fact, we already know that some operators Mϕ∗ are hypercyclic, as we will
2
see now. As usual, B and F denote the backward and forward shifts ∞on (Nn0 ),
respectively, which are operators of norm 1. Hence,
∞ if ϕ(z) = n=0 an z is
∞
holomorphic on some neighbourhood of D then n=0 an B n ≤ n=0 |an | <
∞, so that
∞
ϕ(B) = an B n
n=0
∞
defines an operator on 2 (N0 ), and the same is true for ϕ(F ) = n=0 an F n ;
see also Appendix B.
∞
Proposition 4.41. Let ϕ(z) = n=0 an z n be holomorphic on a neighbour-
∞
hood of D, and set ϕ∗ (z) = n=0 an z n . Then, via the identification of H 2
with 2 (N0 ):
4.4 Adjoint multipliers 121
∞
∞
∞
n
Mϕ f (z) = an z n bk z k = an−k bk z n .
n=0 k=0 n=0 k=0
In particular, the adjoint multipliers Mϕ∗ with ϕ(z) = λz, |λ| > 1, corre-
spond to the Rolewicz operators λB and are therefore hypercyclic.
The Godefroy–Shapiro criterion allows us to characterize the hypercyclic
adjoint multipliers. We can exclude constant multipliers because their adjoint
multiplication operators are multiples of the identity.
Proof. Suppose that condition (iv) holds. Considering the reproducing ker-
nels kλ , λ ∈ D, we find that, for all f ∈ H 2 ,
contains span{kλ ; |ϕ(λ)| < 1}, which is dense in H 2 by Lemma 4.39; indeed,
since nonconstant holomorphic functions are open mappings, condition (iv)
implies that {λ ∈ D ; |ϕ(λ)| < 1} is nonempty and open and therefore
contains an accumulation point in D. For the same reason the eigenvectors of
Mϕ∗ to eigenvalues of modulus greater than 1 span a dense set in H 2 . Finally,
the same is true for the eigenvectors of Mϕ∗ to eigenvalues that are roots
of unity. For this it suffices to show that {λ ∈ D ; ϕ(λ) is a root of unity}
has an accumulation point. But since ϕ is an open mapping, condition (iv)
implies that infinitely many preimages of roots of unity lie in some relatively
compact subset of D and therefore have an accumulation point in D. By the
Godefroy–Shapiro criterion, therefore, (iv) implies (ii) and (iii).
To finish the proof it suffices to show that (i) implies (iv). Let us suppose
that ϕ(D) does not intersect the unit circle. Since ϕ(D) is connected, it must
lie entirely inside or entirely outside D. If ϕ(D) ⊂ D then
(see Proposition A.8), and hence Mϕ∗ cannot be hypercyclic. On the other
hand, if ϕ(D) ⊂ C \ D then ψ := 1/ϕ is a bounded holomorphic function on
D with ψ(D) ⊂ D, which implies that Mψ∗ cannot be hypercyclic. But Mϕ
is the inverse of Mψ and therefore Mϕ∗ is the inverse of Mψ∗ ; see Proposition
A.8. By Proposition 2.23, Mϕ∗ cannot be hypercyclic.
This result can easily be extended to more general Hilbert spaces of holo-
morphic functions, for example the Bergman space (see Exercise 4.4.3); but
see also Exercise 4.4.4. An extension to some Banach spaces X of holomor-
phic functions is also possible, in which case, of course, Mϕ∗ is the Banach
space adjoint on the dual X ∗ ; see Exercise 4.4.5.
We pass on to another, partial generalization of Theorem 4.42 that is
motivated∞by Proposition 4.41. Under the assumptions of that proposition,
ϕ(B) = n=0 an B n defines an operator on each of the spaces p , 1 ≤ p < ∞,
and c0 .
span{eλ ; λ ∈ Λ}
I + λB and eλB , λ = 0
Cϕ f = f ◦ ϕ
Proof. By Proposition 4.36 there are a, b ∈ C with |a| < 1 and |b| = 1 such
that
a−z
ϕ(z) = b .
1 − az
124 4 Classes of hypercyclic and chaotic operators
we deduce that
1 + |a| 1 u0 +2π
1 + |a|
f ◦ ϕ ≤
2
· |f (eiu )|2 du = · f 2 . (4.14)
1 − |a| 2π u0 1 − |a|
Next, for Y0 we take the subspace of H 2 of all functions that are holomor-
phic on a neighbourhood of D and that vanish at z1 , and for S we take the
map S = Cϕ−1 . Since z1 is a fixed point of ϕ−1 , S maps Y0 into itself, and
clearly T S = I. It follows as above that Y0 is dense in H 2 and that S n f → 0
for all f ∈ Y0 .
Therefore the conditions of Kitai’s criterion are satisfied, so that Cϕ is
mixing.
Some concrete instances of this result are treated in Exercise 4.5.2.
We end this chapter by returning to Theorem 4.37 of Section 4.3. Propo-
sition 4.47 allows us to give the missing proof of the implication (v)=⇒(vi)
for Ω = D.
Conclusion of the proof of Theorem 4.37. Let ϕ be an automorphism of D
without fixed points in D. Again, ϕ is either parabolic or hyperbolic.
First, let ϕ be parabolic. By the discussion before Definition 4.46 there
is a linear fractional transformation σ that provides a conjugacy between ϕ
and ψ(z) = z + c, c = 0. Then ψ is an automorphism of σ(D), so that Cϕ is
conjugate to the operator Cψ on H(σ(D)) by Proposition 4.25. By Runge’s
theorem, the continuous restriction map H(C) → H(σ(D)), f → f |σ(D) has
dense range. Hence Cϕ is quasiconjugate to the Birkhoff operator Cψ on
H(C).
In the hyperbolic case, there is a linear fractional transformation σ such
that ϕ is conjugate to a dilation ψ(z) = λz, λ = 1 strictly positive, and ψ
is an automorphism of σ(D), which therefore must be a half-plane with 0 on
its boundary. After conjugation with a suitable rotation we can assume that
it is the right half-plane C+ , and ψ remains unchanged. Now, the principal
branch log of the logarithm is a conformal map from C+ to the strip S =
{z ∈ C ; |Im(z)| < π2 }, and conjugating ψ with log gives us the translation
τ (z) = z + log λ, log λ = 0, on S. We conclude as in the parabolic case that
Cϕ is quasiconjugate to the Birkhoff operator Cτ on H(C).
Exercises
Exercise 4.1.1. Show that a weighted shift Bw defines an operator on H(C) if and only
if supn≥1 |wn |1/n < ∞, and that an en → 0 in H(C) if and only if |an |1/n → 0.
Exercise 4.1.2. Let T := Bw be a weighted shift on p , 1 ≤ p < ∞.
(a) Given an increasing sequence (nk )k of positive integers, show that the sequence
of operators (T nk )k is hypercyclic if and only if, for each j ∈ N,
j+nk
sup |wν | = ∞.
k≥1
ν=1
(b) Show that T is hereditarily hypercyclic with respect to (nk )k if and only if, for
each j ∈ N,
128 4 Classes of hypercyclic and chaotic operators
j+nk
lim |wν | = ∞.
k→∞
ν=1
Exercise 4.1.3. Let X be the Banach space of all sequences (xn )n satisfying
∞
xn xn+1 xn
x = n − n + 1 < ∞ and n → 0 as n → ∞.
n=1
Show that the backward shift B is a hypercyclic operator on X and that conditions
(ii)–(iv) of Theorem 4.6 are satisfied, but that the only periodic points of B are the
constant sequences and that B is therefore not chaotic.
Exercise 4.1.4. Let X be a Fréchet sequence space over Z in which (en )n∈Z is a basis.
Suppose that the bilateral weighted shift Bw is an invertible operator on X. Show that
Bw is hypercyclic if and only if there is an increasing sequence (nk )k of positive integers
such that
0
nk −1
wν e−nk → 0 and wν enk → 0
ν=−nk +1 ν=1
Exercise 4.1.5. Find a (necessarily non-invertible) bilateral weighted shift that satisfies
the condition stated in the previous exercise but that is not hypercyclic. (Hint: See the
proof of Proposition 4.16, but choose nonsymmetric vn .)
Exercise 4.1.6. Prove the results stated in Remark 4.14; instead of using a conjugacy
one may also observe that a forward shift on the basis (en )n is a backward shift on the
basis (e−n )n .
j
j+n
1
|wν | < ε, |wν | > .
ε
ν=j−n+1 ν=j+1
Exercise 4.2.1. An entire function ϕ is of exponential type 0 if for any ε > 0 there is
some M > 0 such that
|ϕ(z)| ≤ M eε|z| for all z ∈ C.
For example, any polynomial but no exponential function z → eλz , λ
= 0, is of expo-
nential type 0.
For a domain Ω ⊂ C, let H(Ω) denote the Fréchet space of holomorphic functions
on Ω; see Section 4.3. Show the following:
∞
(i) an entire function ϕ(z) = a z n is of exponential type 0 if and only if, for
n=0 n n
any ε > 0, there is some M > 0 such that |an | ≤ M εn! ;
∞
(ii) for any domain Ω ⊂ C and any entire function ϕ(z) = a z n of exponential
∞ n
n=0 n
type 0, ϕ(D) = a D defines an operator on H(Ω);
n=0 n
(iii) for any simply connected domain Ω ⊂ C and any nonconstant entire function ϕ
of exponential type 0, ϕ(D) is chaotic on H(Ω). (Hint: Use the Godefroy–Shapiro
theorem and the fact, that, by Runge’s theorem, H(C) is dense in H(Ω).)
Exercises 129
Exercise 4.2.2. Let Ω be a domain and P a nonconstant polynomial. Show that the
following assertions are equivalent:
(i) P (D) is chaotic on H(Ω);
(ii) P (D) is hypercyclic on H(Ω);
(iii) Ω is simply connected.
there is a smooth Jordan curve Γ in Ω sur-
(Hint: If Ω is not simply connected then
rounding some a ∈ / Ω. Show that f → Γ f (ζ) dζ is an eigenvector of P (D)∗ , and use
Lemma 2.53.)
Exercise 4.2.3. Let X = CR∞ (R) be the space of infinitely differentiable real functions
f : R → R; see Exercise 2.1.5. Show that every (real) differential operator T : X → X,
N
Tf = a f (n) , T
= a0 I, is chaotic. (Hint: See Exercise 2.2.5.)
n=0 n
Exercise 4.2.4. Let ϕ be a nonconstant entire function of exponential type and A =
min{|z| ; z ∈ C, |ϕ(z)| = 1}. Show that, for any ε > 0, there is an entire function f that
is hypercyclic for ϕ(D) such that
Show that any f ∈ Eτ2 satisfies |f (z)| ≤ M eτ r ; use ideas from Example 3.2 to show
that for any Λ ⊂ Dτ with an accumulation point, span{eλ ; λ ∈ Λ} is dense in Eτ2 (see
Appendix A for the dual of Eτ2 ); show that ϕ(D) is an operator on any Eτ2 , and that
2
ϕ(D) is hypercyclic on EA+ε for any ε > 0.
Apply the result to MacLane’s and Birkhoff’s operators.
n4.2.5. Let
Exercise
∞ −1 n
Bw be a chaotic weighted shift on H(C); see Example 4.9(b). Then
n=0
( ν=1
wν ) z is an entire function, and its maximum term is defined by
rn
μw (r) = max n , r ≥ 0.
n≥0
ν=1
|wν |
(a) Let φ : ]0, ∞[ → [1, ∞[ be a function with φ(r) → ∞ as r → ∞. Show that there
exists an entire function f that is hypercyclic for Bw and that satisfies
f (λ) = f, kλ , f ∈ H.
Exercise 4.4.4. The Dirichlet space D is defined as the space of all holomorphic func-
tions f on D such that
1
f 2 := |f (0)|2 + |f (z)|2 dλ(z) < ∞,
π D
then ϕ(B) is a chaotic operator on X. (Hint: See Appendix A for the dual of X.)
Exercise 4.5.1. The aim of this exercise is to prove Littlewood’s subordination principle:
if ϕ : D → D is a holomorphic self-map then Cϕ : f → f ◦ ϕ defines an operator on H 2
with Cϕ ≤ ( 1−r ) , where r = |ϕ(0)|.
1+r 1/2
Exercise 4.5.2. For the following linear fractional transformations decide if they are
automorphisms of D; in the case of an automorphism, determine if the corresponding
composition operator Cϕ is hypercyclic on H 2 :
2z − 1
(i) ϕ(z) = ;
2−z
1 + (i − 1)z
(ii) ϕ(z) = ;
(i + 1) − z
4 − 5z
(iii) ϕ(z) = ;
5 − 4z
z+1
(iv) ϕ(z) = .
2
Exercise 4.5.3. Let α > −1. Then the weighted Bergman space A2α is defined as the
space of all holomorphic functions f on D such that
2 1 α
f := |f (z)|2 1 − |z|2 dλ(z) < ∞,
π D
n=0
|an |
2 1
(n+1)α+1 < ∞. (Hint: Stirling’s formula.)
(b) Let ϕ ∈ Aut(D). Show that Cϕ is an operator on A2α with Cϕ ≤ ( 1−r
1+r 1+α/2
) ,
−2
where r = |ϕ(0)|. (Hint: Show that if w = ϕ(z) then dλ(z) = |ϕ (z)| dλ(w), and use
(4.12).)
Exercise 4.5.4. Let α > −1. Then the weighted Dirichlet space Dα is defined as the
space of all holomorphic functions f on D such that
2
f 2 := |f (0)|2 +
1 f (z) 1 − |z|2 α dλ(z) < ∞,
π D
where λ denotes two-dimensional Lebesgue measure; see Exercise 4.4.4 and the previous
exercise. ∞
(a) Let f be a holomorphic function on D with f (z) = a z n , z ∈ D. Show that
∞ n=0 n
f ∈ Dα if and only if n=0
|an | (n + 1)
2 1−α
< ∞.
(b) Let ϕ ∈ Aut(D). Show that Cϕ is an operator on Dα .
(c) Let α > 0. Show that Theorem 4.48 remains true for Dα . (Hint: Proceed as in the
proof of that theorem; use the change of variables w = ϕn (z); note that 1−|ϕ−n (w)|2 →
0.)
Exercise 4.5.6. Let β = (βn )n≥0 be a sequence of strictly positive numbers such that
∞ −2 n
β r < ∞ whenever 0 ≤ r < 1. Then the weighted Hardy space H (β) is defined
n=0 n
2
∞
∞
n=0 n
|b |(n + 1)ν < ∞. Show that Mϕ is an operator on Sν .
(c) Let ϕ ∈ Aut(D). Deduce from Exercise 4.5.3 that Cϕ defines an operator on Sν
for ν < 0. Use parts (a) and (b) to conclude that Cϕ defines an operator on Sν for any
ν ∈ R. (Hint: (Cϕ f ) = Mϕ Cϕ f .) ∞
∞ ∞
(d) Let ϕ(z) = b z n with
n=0 n n=0 n
|b | ≤ 1 and |b |2 n = ∞ (existence?).
n=0 n
Show that ϕ is a holomorphic self-map of D for which Cϕ does not define an operator
on the Dirichlet space D.
Exercise 4.5.8. Let ν ∈ R and ϕ ∈ Aut(D). By Exercise 4.5.7, Cϕ is an operator on
Sν . Deduce the following from the previous exercises:
(i) if ν ≥ 12 then Cϕ is never hypercyclic on Sν ;
(ii) if ν < 12 then Cϕ is hypercyclic on Sν if and only if ϕ is not the identity and
non-elliptic.
Spell out these results for the (weighted) Bergman and Dirichlet spaces.
Section 4.1. Rolewicz’s multiples of the backward shift were the first Banach space op-
erators to be proved hypercyclic [268]. Due to its simple structure, the class of weighted
shifts is a favorite testing ground for operator-theorists (Salas [274]). Accordingly, when-
ever a new notion in linear dynamics is introduced it is usually first tested on weighted
shifts.
134 4 Classes of hypercyclic and chaotic operators
Salas [274] characterized hypercyclic and weakly mixing unilateral and bilateral
weighted shifts on 2 and 2 (Z), respectively. The characterizations for more general
sequence spaces and of chaos are due to Grosse-Erdmann [180], see also Martínez and
Peris [229] in the special case of Köthe sequence spaces, while mixing shifts on 2 and
2 (Z) were characterized by Costakis and Sambarino [124]. The approach chosen here of
first studying the unweighted shift and then using suitable conjugacies is due to Martínez
and Peris [229].
The first example of a hypercyclic operator whose adjoint is also hypercyclic (see
Proposition 4.16) was found by Salas [273]. He later showed [276] that every separa-
ble Banach space with separable dual supports such an operator. The observation that
T ⊕ T ∗ is never hypercyclic is due to Deddens; see [273].
Section 4.2. The investigation of hypercyclicity for differential operators ϕ(D) is due to
Godefroy and Shapiro [165]. Theorem 4.22 on the rate of growth of MacLane’s operator
was obtained independently by Grosse-Erdmann [178] and Shkarin [283]. The corre-
sponding result for Birkhoff’s operators was obtained by Duyos-Ruiz [137]; alternative
proofs can be found in Chan and Shapiro [106] and in Exercise 8.1.3.
Translation and differentiation operators have also been studied on spaces of har-
monic functions on RN , N ≥ 2. Hypercyclicity of these operators and corresponding
growth results have been obtained by Dzagnidze [138], Aldred and Armitage [6], [7],
[13], and Gómez, Martínez, Peris and Rodenas [166].
Section 4.3. This section draws heavily on the work of Bernal and Montes [64], who
also coined the term “run-away sequence”, and the work of Shapiro [281]. The material
up to Theorem 4.32 can be found in [64], while most of Theorem 4.37 is implicit in
[281]. We mention that Seidel and Walsh [278] were the first to study the analogue of
Birkhoff’s result in the unit disk.
For two different proofs of Proposition 4.31 we refer to Bernal and Montes [64] and to
Grosse-Erdmann and Mortini [184]. Example 4.34 is taken from Kim and Krantz [214];
see also Gorkin, León and Mortini [168].
Section 4.4. The study of the dynamical properties of adjoint multipliers was initiated
by Godefroy and Shapiro [165], who also obtained Theorem 4.42. Functions of the back-
ward shift on the spaces p and c0 were studied by deLaubenfels and Emamirad [128],
who also obtained Theorem 4.43. An interesting related investigation of functions of the
backward shift on the Bergman space is due to Bourdon and Shapiro [96].
For a more detailed introduction to Hardy spaces we refer to Duren [136] and Rudin
[270].
Section 4.5. In this section we closely follow the book of Shapiro [279]; see also Shapiro
[281]. Proposition 4.45 is a special case of the Littlewood subordination principle; see
Exercise 4.5.1. Theorem 4.48 is due to Bourdon and Shapiro [94], [95]. This result is
only the beginning of a fascinating story on the interplay between operator theory and
complex function theory. The extension of Theorem 4.48, first to non-automorphic linear
fractional transformations and then to more arbitrary holomorphic self-maps of D, can
be found in the cited work of Bourdon and Shapiro. The proofs, however, require a much
deeper understanding, for example, of the (nonlinear) dynamics of self-maps of D.
Hosokawa [204] proved that, for any automorphism ϕ of D, Cϕ is chaotic whenever
it is hypercyclic; see also Taniguchi [298]. Thus one can add chaos to the equivalent
conditions in Theorem 4.48.
We note that Gallardo and Montes [158] have obtained a complete characterization
of the cyclic, supercyclic and hypercyclic composition operators Cϕ for linear fractional
self-maps ϕ of D on any of the spaces Sν , ν ∈ R (see Exercises 4.5.7 and 4.5.8).
Sources and comments 135
Exercises. Exercise 4.1.2 is taken from Bès and Peris [71], Exercise 4.1.3 from Grosse-
Erdmann [180]. For Exercises 4.1.4 and 4.1.5 we refer to Feldman [150], Exercise 4.1.7
states the condition in the form found originally by Salas [274]. But note that the
weighted shifts considered by Feldman and Salas are forward shifts. For Exercise 4.2.1
we refer to Bernal [55] and Shapiro [280], for Exercise 4.2.2 to Shapiro [280], for Exer-
cise 4.2.4 to Chan and Shapiro [106] and to Bernal and Bonilla [60], for Exercise 4.2.5
to Grosse-Erdmann [181]. Exercises 4.3.2 and 4.3.3 follow Montes [241] and Grosse-
Erdmann and Mortini [184], while Exercise 4.3.4 is taken from Shapiro [280]. The mate-
rial for Exercises 4.4.1–4.4.4 can be found in Godefroy and Shapiro [165], with Exercise
4.4.4(vii) being taken from Chan and Seceleanu [104]; for Exercises 4.4.6–4.4.8 we refer
to deLaubenfels and Emamirad [128], for Exercise 4.4.9 to Martínez [228]. For Exercise
4.5.1 we have again followed Shapiro [279]; Exercises 4.5.4(c), 4.5.5 and 4.5.8 are taken
from Gallardo and Montes [158], Exercise 4.5.6 from Zorboska [304] and Exercise 4.5.7
from Hurst [205].
Chapter 5
Necessary conditions for hypercyclicity
and chaos
For mixing operators T one can even assert that (T ∗ )n x∗ → ∞ for any
nonzero element x∗ ∈ X ∗ ; see Exercise 5.1.1.
Next we study the spectrum σ(T ) of the operator T , which suggests con-
sideration of complex spaces.
Throughout the remainder of this section, T denotes an operator on a
complex Banach space X.
Proof. (i) Since σ(T ) is a compact set, the assumption and the spectral radius
formula imply that limn→∞ T n 1/n = r(T ) < r. Hence, if ε < r − r(T ) then
there is some M > 0 such that T n ≤ M (r − ε)n for all n ∈ N0 , and the
result follows.
(ii) The assumption implies that 0 ∈ / σ(T ), so that T is invertible. Since
σ(T −1 ) = σ(T )−1 (see Exercise 5.0.7), we obtain that σ(T −1 ) ⊂ {z ∈
C ; |z| < 1r }. By (i) there are then some η > 0 with η < 1r and M > 0
such that (T −1 )n y ≤ M ( 1r − η)n y for all y ∈ X and n ∈ N0 . Setting
y = T n x and defining ε by 1r − η = r+ε 1
we obtain the result.
Proposition 5.3. Let T be a hypercyclic operator. Then σ(T ) meets the unit
circle:
σ(T ) ∩ T = ∅.
Proof. Suppose, on the contrary, that σ(T ) does not meet the unit circle.
If σ(T ) ⊂ D then, by taking r = 1 in Lemma 5.2(i), we see that all orbits
of T tend to 0, which is impossible. Similarly, Lemma 5.2(ii) shows that
σ(T ) ⊂ C \ D is impossible. We therefore have that the sets
5.1 Spectral properties of hypercyclic and chaotic operators 139
form a partition of σ(T ) into nonempty closed sets. By the Riesz decomposi-
tion theorem there then exist nontrivial T -invariant closed subspaces M1 and
M2 such that X = M1 ⊕ M2 , σ(T |M1 ) = σ1 and σ(T |M2 ) = σ2 . By Propo-
sition 2.28, T |M1 would then be a hypercyclic operator with σ(T |M1 ) ⊂ D,
which is impossible, as we saw above.
so that
1
V n 1/n ≤ → 0.
n!1/n
140 5 Necessary conditions for hypercyclicity and chaos
The spectral radius formula implies that r(V ) = 0, so that σ(V ) = {0}. The
same result holds for any space X = Lp [0, 1], 1 ≤ p < ∞; see Exercise 5.1.2.
By Proposition 5.3, V cannot be hypercyclic on any of the spaces consid-
ered. Indeed, for the same reason, no multiple λV , λ ∈ C, can be hypercyclic.
Proof. By the preceding lemma we have to show that the compact set σ(T )∪T
is connected. If this is not the case then σ(T ) ∪ T can be partitioned into two
nonempty closed sets C1 and C2 . Since T is connected, it must lie entirely in
one of these two sets, C2 say. Then we define a partition of σ(T ) into closed
sets by setting
Proof. We first show that σp (T ) contains infinitely many roots of unity. Since
T has periodic points, Proposition 2.33 shows that σp (T ) contains at least
one root of unity. Now suppose that λ1 , . . . , λN , N ≥ 1, are the only roots of
unity in the point spectrum of T . Consider the polynomial
p(z) = (z − λ1 ) · · · (z − λN ).
p(T )x = (T − λ1 I) · · · (T − λN I)x = 0.
Again by Proposition 2.33 and by linearity of p(T ) we have that p(T )x = 0
for every periodic point of T . Since these points form a dense set in X we
obtain that p(T ) = 0, which contradicts Theorem 2.54.
Now suppose that λ is an isolated point of the spectrum of T . Then by
the Riesz decomposition theorem there are nontrivial T -invariant closed sub-
spaces M1 and M2 of X such that X = M1 ⊕ M2 and σ(T |M2 ) = {λ}. It
follows as in Proposition 2.28 (see also Exercise 2.2.8), that T |M2 is chaotic,
which is impossible since its spectrum is a singleton.
Based on necessary conditions for hypercyclicity we can now show that certain
classes of operators cannot contain hypercyclic operators.
Throughout this section, T denotes an operator on a (real or complex)
Banach space X.
is called a Volterra integral operator, and k is its kernel; note that the values
k(t, s), s > t, are irrelevant. It is obvious that Vk defines an operator on
t
C[0, 1] with Vk ≤ maxt∈[0,1] 0 |k(t, s)| ds < ∞. In the special case when k
is identically 1 we obtain the classical Volterra operator of Example 5.4.
A simple induction shows that every power of a Volterra integral operator
is a Volterra integral operator. More precisely, if kn is the kernel of (Vk )n ,
n ≥ 1, then
t
kn+1 (t, s) = k(t, u)kn (u, s) du, s ≤ t, n ≥ 1;
s
Mn
|kn (t, s)| ≤ (t − s)n−1 , s ≤ t, n ≥ 1,
(n − 1)!
and hence
Mn
(Vk )n ≤ ,
n!
which implies that limn→∞ (Vk )n 1/n = 0. We ask the reader to verify
these statements; see Exercise 5.2.1. Thus, each Volterra integral operator is
quasinilpotent and therefore not hypercyclic.
see Figure 5.1. Then x = 1, and for any y ∈ M we have that
x −y
0 0
x − y = − y
x0 − y0
1
= x0 − (y0 + yx0 − y0 )
x0 − y0
d
≥ ≥ 1 − ε,
x0 − y0
dn = inf{x ; (I − T )x = yn },
as n → ∞. This implies that the sequence (zn )n also converges. We call its
limit z. It then follows from (5.1) that z − T z = 0 and hence also that
xn − T xn = (I − T )xn = yn → y
as n → ∞. This implies that the sequence (xn )n itself converges. Calling its
limit x we obtain that
The next lemma tells us that the eigenvalues of a compact operator can
only have 0 as an accumulation point.
5.2 Classes of non-hypercyclic operators on Banach spaces 145
Lemma 5.14. Let T be a compact operator and ε > 0. Then T has only a
finite number of eigenvalues λ with |λ| ≥ ε.
Mn := span{x1 , . . . , xn }, n ≥ 1
Proof. If 0 is not in the spectrum then T is invertible. Thus the unit ball
of X, as the homeomorphic image (under T −1 ) of a relatively compact set,
is itself relatively compact. But this can only happen in finite-dimensional
spaces; see Appendix A.
We are now in a position to show that compact operators are never hy-
percyclic. Here we only give the proof in the case of complex Banach spaces.
In the real case the proof is outlined in Exercise 5.2.5; it uses the same ideas
as in the complex case but it is less straightforward.
146 5 Necessary conditions for hypercyclicity and chaos
Example 5.16. In Example 5.9 we showed that the Volterra integral operators
on C[0, 1] are quasinilpotent, and therefore not hypercyclic.
We want to show here that they are also compact, implying once more
that they are not hypercyclic. To this end let (fn )n be a sequence in C[0, 1]
with fn ≤ 1, n ≥ 1. We have to construct a subsequence (gν )ν of (fn )n
such that (T gν )ν converges.
We begin the proof by a familiar diagonal process. Let (tj )j be a dense
sequence in [0, 1]. Since T is continuous, the sequence (T fn )n is bounded in
C[0, 1], and hence the sequence ((T fn )(t1 ))n is bounded in K. We can there-
fore extract a subsequence (n1,ν )ν such that ((T fn1,ν )(t1 ))ν converges in K.
In the same way, since ((T fn1,ν )(t2 ))ν is bounded there exists a subsequence
(n2,ν )ν of (n1,ν )ν such that ((T fn2,ν )(t2 ))ν converges. Proceeding inductively,
we find subsequences (nj,ν )ν of (nj−1,ν )ν such that ((T fnj,ν )(tj ))ν converges,
for j ≥ 2. Now define mν = nν,ν . Since, for any j ≥ 1, (mν )ν≥j is a subse-
quence of (nj,ν )ν we have that ((T fmν )(tj ))ν converges.
We claim that gν = fmν , ν ≥ 1, defines the required subsequence of
(fn )n . To show that (T gν )ν converges in C[0, 1] it suffices to show that it is a
Cauchy sequence. Thus, let ε > 0. Since the kernel k of T is continuous, and
hence uniformly continuous, on [0, 1] × [0, 1] there is some δ > 0 such that, if
|τ1 − τ2 | < δ, τ1 , τ2 ∈ [0, 1], then maxs∈[0,1] |k(τ1 , s) − k(τ2 , s)| < ε. Moreover,
M := maxt,s∈[0,1] |k(t, s)| < ∞. By the density of (tj )j there is some J ∈ N
such that for any t ∈ [0, 1] there is some j ≤ J such that |t − tj | < min(δ, ε).
The definition of the sequence (gν )ν shows that there is some N ∈ N such
that, for any j ≤ J and ν, μ ≥ N ,
Now let t ∈ [0, 1]. Choose j ≤ J such that |t − tj | < min(δ, ε). Then, for
any ν ≥ 1,
tj
|(T gν )(t) − (T gν )(tj )| ≤ (k(t, s) − k(tj , s))gν (s) ds
0
t
+ k(t, s)gν (s) ds
tj
≤ ε + M ε,
5.2 Classes of non-hypercyclic operators on Banach spaces 147
where we have used that gν ≤ 1, ν ≥ 1. We therefore have that, for any
ν, μ ≥ N ,
The proof can be given as in the case of compact operators, taking account
of the fact that σ(T n ) = σ(T )n by the spectral mapping theorem. Alterna-
tively, the result is also an immediate consequence of the compact case and
of Ansari’s theorem that we will obtain in the next chapter; see Theorem 6.2.
By Ansari, T can only be hypercyclic if T n is, which is impossible if T n is
compact.
λI + F, λ ∈ K,
In fact, an even more general result can be established; we leave the proof
to the reader (see Exercise 5.2.8).
T = λI + K,
We restrict our attention again to the complex case and refer to Exercise
5.2.9 for the real case.
148 5 Necessary conditions for hypercyclicity and chaos
We will show later that there are indeed hypercyclic operators of the form
T = I + K with K compact; see Example 8.4. But such an operator cannot
be chaotic.
A = (ank )n,k≥1 .
Now suppose that in the nth row of the matrix A, all off-diagonal
∞ elements
are 0, that is, ank = 0 for all k = n. Then we have for x = k=1 xk ek ∈ X
that ∞
∗ ∗ ∗
x, T en = T x, en = ank xk = ann xn = ann x, e∗n .
k=1
(see Section 4.4), which are special Toeplitz matrices; chaos for ϕ(B) is char-
acterized in Theorem 4.43.
T = T ∗.
In order to explore these notions we need a preliminary result.
Lemma 5.24. Let T be an operator on a Hilbert space H.
(a) Suppose that H is a complex Hilbert space. If
T x, x ∈ R for all x ∈ H,
then T is self-adjoint. If
T x, x = 0 for all x ∈ H,
then T = 0.
(b) Suppose that H is a real Hilbert space. If T is self-adjoint and
T x, x = 0 for all x ∈ H,
then T = 0.
Proof. For any x, y ∈ H we have that
In the real case, this already implies (b). In fact, the left-hand side is zero by
assumption, and since T is self-adjoint and the inner product is symmetric,
we deduce that, for all x, y ∈ H,
Under the first assumption in (a), the left-hand sides in (5.2) and (5.3) are
both real-valued. Hence we deduce that
and therefore
T x, y = T y, x = x, T y
for all x, y ∈ H. This shows that T ∗ = T .
Under the second assumption in (a) we obtain from (5.2) and (5.3) that
T x, y + T y, x = 0 = −T x, y + T y, x
and hence
152 5 Necessary conditions for hypercyclicity and chaos
T x, y = 0
for all x, y ∈ H. This shows that T = 0.
(T ∗ T − T T ∗ )x, x = T ∗ T x, x − T T ∗ x, x = T x, T x − T ∗ x, T ∗ x
= T x2 − T ∗ x2 , (5.4)
T x, x ≥ 0.
Remark 5.26. Of course, one may just as well consider operators that satisfy
T T ∗ − T ∗ T ≥ 0.
That these are of less interest in the present context will become clear in
Exercise 5.3.5.
T x2 = T x, T x = T ∗ T x, x ≤ T ∗ T xx
for all x ∈ H. Proposition 5.27, applied to T x, then shows that for hyponor-
mal operators we have that
T x2 ≤ T 2 xx
self-adjoint T = T ∗
normal T ∗T = T T ∗ ∀ x ∈ H, T x = T ∗ x
hyponormal T ∗ T − T T ∗ ≥ 0 ∀ x ∈ H, T x ≥ T ∗ x
paranormal ∀ x ∈ H, T x2 ≤ T 2 xx
Table 5.1 Classes of Hilbert space operators
0 1
Remark 5.28. The operator on K2 that is given by the matrix A = −1 0
is normal, but not self-adjoint. For distinguishing the remaining classes of
operators we refer to Exercises 5.3.1 and 5.3.2.
Proof. We show that the orbits of paranormal operators are too well behaved
for the operator to be hypercyclic. Indeed, let T be a paranormal operator
on a Banach space X and x ∈ X. Suppose that, for some n ≥ 0, T n+1 x >
T n x, which implies, in particular, that T n+1 x = 0 and T n x = 0. Applying
the definition of paranormality to T n x we obtain that
T n+1 x2
T n+2 x ≥ > T n+1 x.
T n x
One may also arrive at the same conclusion from another point of view.
Considering T n x in the definition of paranormality and taking logarithms we
see that, if T is a paranormal operator, then, for any x ∈ X and n ≥ 0 with
T n+2 x = 0,
1
log T n+1 x ≤ (log T n+2 x + log T n x).
2
Thus, for any x ∈ X whose orbit does not end up in the origin, the sequence
(log T n x)n is convex. This leads to the behaviour of the sequence (T n x)n
found in the proof.
Corollary 5.31. No self-adjoint operator, no normal operator, no hyponor-
mal operator on a Hilbert space is hypercyclic. No positive operator on a
complex Hilbert space is hypercyclic.
We finish this section, and indeed the chapter, with two examples.
Example 5.32. We consider the bilateral weighted backward shift
on the Hilbert space 2 (Z), where w = (wn )n∈Z is a weight sequence; see
Section 4.1. An easy computation shows that
∗
Bw = Fv ,
the weighted forward shift with weight sequence v = (wn+1 )n∈Z ; see Remark
4.14. (Note that this is the Hilbert space adjoint of Bw ; its Banach space
adjoint was determined in the proof of Proposition 4.16.)
We calculate that, for x ∈ 2 (Z) and λ ≥ 0,
2 2
Fv Bw − 2λFv Bw + λ2 I x, x = |wn |2 |wn−1 |2 − 2λ|wn |2 + λ2 |xn |2 .
n∈Z
Exercises
Exercise 5.0.1. Consider the operator T given by T (x1 , x2 ) = (−x2 , x1 ) on the real
Banach space R2 . Show that σ(T ) = ∅ and σ(T 2 ) = {−1}. Thus, some results of
Appendix B do not hold in the real case.
Exercise 5.0.2. Let B be the backward shift operator and F the forward shift operator
on any of the sequence spaces X = p , 1 ≤ p < ∞, or X = c0 . Show that σp (B) = D
and σ(B) = D. Determine also σp (F ) and σ(F ). (Hint: Find F ∗ .)
Exercise 5.0.5. Prove the spectral mapping theorem. (Hint: Look at the proof of the
point spectral mapping theorem.)
Exercise 5.0.6. Give a direct proof of the spectral mapping theorem for polynomials,
without the use of functional calculus. (Hint: For λ ∈ C there are c, μ1 , . . . , μN ∈ C such
that λ − f (z) = c(μ1 − z) · · · (μN − z); use the corresponding factorization for λI − T .)
Exercise 5.0.7. If T is an invertible operator then 0 ∈/ σ(T ) and σ(T −1 ) = σ(T )−1 :=
{ z ; z ∈ σ(T )}. Give two proofs of the identity, one direct and one using the spectral
1
mapping theorem.
Exercise 5.0.9. Give a direct proof of the point spectral mapping theorem for polyno-
mials using the ideas of Exercise 5.0.6.
Exercise 5.1.1. Let T be a mixing operator on a (real or complex) Banach space. Then,
for every nonzero vector x∗ ∈ X ∗ , (T ∗ )n x∗ → ∞ as n → ∞. (Hint: Suppose that
(T ∗ )nk x∗ ≤ M and consider U = {x ; x < 1}, V = {x ∈ X ; | x, x∗
| > M }.)
Exercise 5.1.2. (a) Show that the Volterra operator also defines an operator on the
space L∞ [0, 1] of essentially bounded measurable functions on [0, 1]. Moreover, show
that V = 1 for V as an operator on L1 and on L∞ .
(b) Show that, on any space C[0, 1] or Lp [0, 1], 1 ≤ p ≤ ∞, the Volterra operator has
no eigenvalues.
156 5 Necessary conditions for hypercyclicity and chaos
Exercise 5.1.5. Let T be an ε-hypercyclic operator; see Exercise 2.2.10. Show that
every connected component of σ(T ) meets the unit circle.
Exercise 5.1.6. Let T be a J-class operator; see Exercise 2.2.11. Show that σ(T ) meets
the unit circle, but that not necessarily every connected component of σ(T ) does. (Hint:
Avoid the problem presented in Exercise 2.2.11(a) by modifying the proof of Proposition
5.3.)
Exercise 5.2.1. Prove the statements on the Volterra integral operator made in Exam-
ple 5.9.
Exercise 5.2.3. Let X be a Banach space and K(X) the space of compact operators
on X. Show the following:
(i) the identity operator on X is compact if and only if X is finite dimensional; (Hint:
Use a result from Appendix A.)
(ii) if K, L ∈ K(X), λ ∈ K and T ∈ L(X) then λK, K + L, T K and KT all belong to
K(X) (one says that K(X) is an ideal in L(X)).
Exercise 5.2.4. Let X be a Banach space and K(X) the space of all compact operators
on X. Show that K(X) is a closed subspace of L(X), the space of all operators on X,
endowed with the operator norm topology. Deduce that a limit, in the operator norm,
of finite-rank operators is compact. (Hint: Look at the argument contained in Example
5.16.)
Exercises 157
Exercise 5.2.5. Show that no compact operator on a real Banach space can be hyper-
cyclic by proceeding as follows. Suppose that T is a hypercyclic compact operator on an
infinite-dimensional real Banach space X, and let T : X →X denote its complexifica-
tion. By Exercise 2.6.4, T is 2-hypercyclic.
(a) Show that T is compact.
(b) Show that, for every ε > 0, σ(T) ∩ {z ∈ C ; |z| ≥ ε} is a finite set. (Hint: Use
Exercise 2.6.4.)
(c) By Lemma 5.15, 0 ∈ σ(T). If σ(T) = {0} deduce that all orbits under T tend to
0, contradicting the 2-hypercyclicity of T.
(d) If {0} σ(T), show that there are nontrivial T-invariant closed subspaces M1
and M2 with X = M1 ⊕ M2 such that σ(T|M1 ) ⊂ D. Prove that T|M1 is 2-hypercyclic
and deduce a contradiction as in (c).
Exercise 5.2.7. Let ϕ : [0, 1] → [0, 1] be a continuous function. Then the Volterra
composition operator Vϕ : C[0, 1] → C[0, 1] is defined as
ϕ(t)
Vϕ f (t) = f (s) ds, 0 ≤ t ≤ 1.
0
(a) Show that no Volterra composition operator is hypercyclic on C[0, 1]. (Hint:
Vϕ = Cϕ ◦ V .)
(b) Now let ϕ(x) = xα , α > 0, and consider Vϕ as an operator on C0 [0, 1[, the
space of continuous functions f on [0, 1[ with f (0) = 0; it is a Fréchet space under the
seminorms pn (f ) = supt∈[0,1−1/n] |f (t)|, n ≥ 1. Show the following:
(i) for α ≥ 1, Vϕ is not hypercyclic on C0 [0, 1[;
(ii) for α < 1, Vϕ is hypercyclic on C0 [0, 1[.
(Hint for (ii): Use Kitai’s criterion with X0 the set of polynomials in C0 [0, 1[ and Y0 =
span{xν ; ν > α/(1 − α)}; use the Weierstrass approximation theorem to show that Y0
is dense.)
Exercise 5.2.8. Show that no hypercyclic operator can commute with a finite-rank
operator. Deduce Proposition 5.18 from this. (Hint: Use a commutative diagram.)
Exercise 5.2.9. Prove Lemma 5.19 for real Banach spaces. (Hint: Use the ideas of
Exercise 5.2.5.)
Exercise 5.2.11. The aim of this and the next exercise is to get a better understanding
of the spectrum of compact operators. Show here that the adjoint T ∗ of a compact
operator is compact. (Hint: Let x∗n ≤ 1. For the unit ball BX of X, K := T (BX ) is
a compact metric space. Note that x∗n |K ∈ C(K), the space of continuous functions on
K, endowed with the sup-norm. Apply the Arzelà–Ascoli theorem (see Exercise 5.2.6),
which remains true for C(K).)
158 5 Necessary conditions for hypercyclicity and chaos
Exercise 5.3.3. An operator T on a Banach space X is called power-regular if, for any
x ∈ X, (T n x1/n )n converges. It is called normaloid if, for any n ≥ 1, T n = T n .
Show the following:
(i) every paranormal operator is power regular;
(ii) every paranormal operator is normaloid;
(iii) normaloid operators can be hypercyclic.
(Hint: In (i), study the behaviour of the sequence (T n+1 x/T n x)n≥0 ; in (ii), prove
inductively that T n x ≥ T xn , for all n ≥ 1 and all unit vectors x.)
Exercise 5.3.4. (a) Let p(λ) = aλ2 + bλ + c be a real polynomial with a = 0 and b ≤ 0.
Show that p(λ) ≥ 0 for all λ ≥ 0 if and only if a > 0 and b2 ≤ 4ac.
(b) Deduce from (a) that an operator T on a Hilbert space H is paranormal if and
only if, for all λ ≥ 0,
T ∗2 T 2 − 2λT ∗ T + λ2 I ≥ 0.
Exercise 5.3.6. Show that for bilateral weighted backward shifts on 2 (Z), hyponor-
mality and paranormality are equivalent properties.
Exercise 5.3.7. Show that the Volterra operator on L2 [0, 1] is not paranormal.
Sources and comments 159
Section 5.1. It seems fair to say that all the results in this chapter were either found,
or at least initiated, by C. Kitai in her PhD thesis [215]. Some of her results were also
found, independently, by Matache [233].
In particular, Theorem 5.6 is due to Kitai [215]. Lemma 5.5 is not as obvious as
it may at first seem. While one implication is standard (see for example [250, p. 74,
Corollary 2]), the other implication can be derived from [250, p. 83, Corollary 1].
Quite remarkably, Shkarin [287] has recently shown that Kitai’s necessary spectral
condition actually characterizes spectra of hypercyclic operators on Hilbert spaces.
Dilworth and Troitsky [135] have shown that Kitai’s theorem remains true for weakly
hypercyclic operators, that is, operators with a weakly dense orbit. Herrero [194] has
shown that for any supercyclic operator (on a complex Hilbert space) there is some
r ≥ 0 such that every connected component of its spectrum meets the circle |z| = r;
see also Feldman, Miller and Miller [151, Theorem 6.2]. Herrero also obtained a spec-
tral characterization of all operators on complex Hilbert spaces that are limits, in the
operator norm, of hypercyclic operators.
Concerning the other necessary conditions found in Section 5.1, Proposition 5.1, part
(i) is due to Kitai [215] while part (ii) was observed by Bourdon; see [106, p. 1446] and
[96, Lemma 3.1]. Proposition 5.7 is implicit in Bonet, Martínez and Peris [82].
Section 5.2. The main result of this section, Theorem 5.11, is once more due to Kitai
[215] (for complex Banach spaces); see also Matache [233]. The real case was provided
by Bonet and Peris [85]. Kitai’s proof is quite short but it draws heavily on the Riesz
spectral theory of compact operators by which the spectrum of any compact operator on
an infinite-dimensional complex Banach space is either finite, including 0, or it consists
of a sequence converging to 0, again including 0; see [116, Chapter VII, § 7]. In the
approach chosen here, Bourdon’s theorem allows us to avoid some parts of the Riesz
theory; but see Exercise 5.2.12.
Proposition 5.17, Proposition 5.18, Lemma 5.19 and Proposition 5.20 are due to Kitai
[215], Chan and Shapiro [106, p. 1445, p. 1446] and Martínez and Peris [229].
It is interesting to note that, by Shkarin [289], there are hypercyclic finite-rank per-
turbations of unitary operators on complex Hilbert spaces; by Grivaux [175], the per-
turbation may even be of rank 1.
Concerning specific operators, the Volterra operator (on Lp [0, 1], 1 ≤ p < ∞) and the
Cesàro operator (on p , 1 < p < ∞) are not even supercyclic; see Gallardo and Montes
[159] and León, Piqueras and Seoane [224]. As for Example 5.22, Hardy’s inequality can
be found in [193].
Section 5.3. We refer to Kubrusly [218] for a discussion of various types of normality
conditions for Hilbert space operators. It was once more Kitai [215] who showed that no
hyponormal operator can be hypercyclic. This was extended in two respects by Bourdon
[92, p. 352] who proved that no paranormal operator can even be supercyclic. Concern-
ing Example 5.33, it should be noted that some authors include self-adjointness in the
definition of positive operators on real Hilbert spaces. In that case, of course, no positive
operator is hypercyclic.
Exercises. We have taken Exercise 5.1.1 from Bonet [79] and Exercise 5.1.3 from Mat-
ache [234]. The notion of an irregular vector was introduced by Beauzamy [48, p. 41];
160 5 Necessary conditions for hypercyclicity and chaos
Exercises 5.1.4 and 5.1.5 are due to Prǎjiturǎ [261], while Exercise 5.1.6 is taken from
Costakis and Manoussos [119]. The result of Exercise 5.2.5 is due to Bonet and Peris
[85]. Exercise 5.2.7 is taken from Herzog and Weber [201], Exercise 5.2.8 from Shapiro
[281, 2.9]. The result of Exercise 5.2.11 is one half of Schauder’s theorem; see [116].
For the background to Exercises 5.3.1, 5.3.2 and 5.3.5 we refer to Kubrusly [218]. As
for Exercise 5.3.3, power-regular operators were introduced by Atzmon [17], who also
showed that operators with a countable spectrum are power regular. As a consequence,
power-regular operators can be hypercyclic; see Example 8.4. That every paranormal
operator is power regular and normaloid is due to Bourdon [92, p. 352] and Istrăţescu,
Saitô and Yoshino [206], respectively. Incidentally, it was in the latter article that the
class of paranormal operators was introduced (as operators of class (N)). The charac-
terization of paranormal operators in Exercise 5.3.4 is due to Ando [8].
Extensions. We have decided to restrict our attention in this chapter to Banach spaces.
But the main results in Section 5.2 remain true in arbitrary Fréchet spaces, and even in
all locally convex spaces; see Chapter 12. Indeed, no compact (and therefore no power-
compact) operator on a locally convex space can be hypercyclic (see Bonet and Peris
[85]), and no compact perturbation of a multiple of the identity can be chaotic; see
Martínez and Peris [229]. For an idea of the proofs we refer to Exercise 12.2.3. Proposi-
tions 5.10 and 5.18 also remain true; their proofs extend immediately.
Dj = orb(T j x, T p ), j = 0, . . . , p − 1.
p−1
p−1
D = orb(x, T ) = orb(T j x, T p ) = Dj
j=0 j=0
6.1 Ansari’s theorem 163
and
T (Dj ) ⊂ Dj+1(mod p) .
Let F ⊂ {0, . . . , p − 1} be a set of minimal cardinality such that
D= Dj .
j∈F
p−1
A := (Dj ∩ Dk )
l=0 j,k∈Fl
j=k
D = orb(y, T ) ⊂ A = A,
is a finite union of pairwise disjoint closed subsets. But this contradicts the
connectedness of D.
In conclusion, F = {j} is a singleton. Then D = Dj , and we obtain that
D = T p−j (Dj ) = D0 , which had to be shown.
The simple example T : {−1, 1} → {−1, 1}, T x = −x, shows that Ansari’s
theorem fails for nonlinear dynamical systems; see also Exercise 1.2.11 for an
example on a metric space without isolated points. Ansari’s theorem does ex-
tend to the nonlinear setting if the set of points with dense orbit is connected;
see Exercise 6.1.7.
164 6 Connectedness arguments in linear dynamics
The following properties can be easily deduced from the continuity of T and
the fact that X has no isolated points (see Exercise 6.2.1):
(i) if y ∈ D(x), then D(y) ⊂ D(x);
(ii) U (x) = U (T k x) for each k ∈ N;
(iii) if R : X → X is a continuous map that commutes with T , then
R(D(x)) ⊂ D(Rx).
We first need a generalization of Theorem 2.54. An easy adaptation of the
argument used there gives the result; see Exercise 6.2.2.
Lemma 6.3. If T admits a somewhere dense orbit and p is a nonzero poly-
nomial, then the operator p(T ) has dense range.
Before proving that a vector whose orbit is somewhere dense is necessarily
hypercyclic, we will show that it is cyclic, that is, the linear span of its orbit
is dense in X.
Lemma 6.4. If orb(x, T ) is somewhere dense, then the set {p(T )x ; p =
0 a polynomial} is connected and dense in X.
Proof. The set A := {p(T )x ; p = 0 a polynomial} is path connected. Indeed,
let p, q be nonzero polynomials. If q is not a multiple of p then the straight
path t → tp(T )x + (1 − t)q(T )x, t ∈ [0, 1], is contained in A. Otherwise we
select a third nonzero polynomial r that is not a multiple of p, and therefore
not of q, and we take the union of the straight paths connecting p(T )x and
q(T )x with r(T )x.
On the other hand, A is a subspace of X that contains orb(x, T ). It follows
from the hypothesis that there is some x0 ∈ X and a 0-neighbourhood W
such that x0 + W ⊂ A. Thus, for any y ∈ X, there is a scalar λ with y ∈ λW ;
hence y ∈ λ(x0 + W ) − λx0 ⊂ A. Consequently, A is dense in X.
Theorem 6.5 (Bourdon–Feldman). Let T be an operator on a Fréchet
space X and x ∈ X. If orb(x, T ) is somewhere dense in X, then it is dense
in X.
6.2 Somewhere dense orbits 165
Proof. We have to show that if U (x) = ∅ then D(x) = X. The proof will be
split into four steps.
p(T )xm ∈ p(T )(U (x)) = p(T )(U (T m+1 x)) ⊂ p(T )(D(T m+1 x)).
Moreover, since T p(T )xm ∈ U (x) ⊂ D(x), properties (iii) and (i) yield that
We have therefore shown that V ∩D(x) = ∅. Since V was arbitrary and D(x)
is closed, we deduce that y ∈ D(x) and hence T −1 (U (x)) ⊂ D(x). Continuity
of T implies that T −1 (U (x)) ⊂ U (x).
n
n
orb(xj , T ) = orb(xj , T ) = X.
j=1 j=1
Since a finite union of nowhere dense sets is nowhere dense, orb(xj , T ) must
be somewhere dense in X for some j ∈ {1, . . . , n}. By the Bourdon–Feldman
theorem, xj then has a dense orbit.
Ansari’s result can easily be derived from this theorem. Let x ∈ HC(T )
and p ∈ N. Since
6.4 Hypercyclic semigroup actions 167
p−1
orb(x, T ) = orb(T j x, T p )
j=0
HC(T ) = HC(λT ).
168 6 Connectedness arguments in linear dynamics
In Exercise 2.5.1 we saw that rotations of mixing (or weakly mixing) op-
erators are mixing (or weakly mixing, respectively). But that result did not
say anything about the sets of hypercyclic vectors.
Theorem 6.8 (Conejero–Müller–Peris). Let (Tt )t≥0 be a C0 -semigroup
on a Banach space X. If x ∈ X is hypercyclic for (Tt )t≥0 , then it is hypercyclic
for each operator Tt , t > 0.
It is particularly gratifying that the two problems can be treated within
a common framework, that of semigroup actions. We will also show that a
variant of the method leads to a new proof of Ansari’s theorem; see Exercise
6.4.5.
Throughout this section we will write
G = N0 × R+ ,
Ψ : G → L(X)
Ψ (n, t) = e2πti T n , n ∈ N0 , t ≥ 0.
Proof. We first note that it suffices to prove the claim for t = 1. Indeed, let
x be hypercyclic for Ψ , and let t > 0 be arbitrary. We distinguish the two
subcases of (α). If Ψ (1, 0) = I then
defines a semigroup action that satisfies (α) and (β). Since x is also hyper-
cyclic for Ψ we can conclude that x is hypercyclic for Ψ(1, 1) = Ψ (1, t). If
Ψ (0, 1) = I then we define
and we can conclude as before that x is hypercyclic for Ψ(1, 1) = Ψ (1, t+1) =
Ψ (1, t).
As usual, T = {z ∈ C ; |z| = 1} is the unit circle. For ease of notation we
introduce the map ρ : R+ → T, given by ρ(t) := e2πti . We then define, for
every pair u, v ∈ X, the subset Fu,v of T by
Fu,v := λ ∈ T ; ∃ ((nk , tk ))k ⊂ G with Ψ (nk , tk )u → v and ρ(tk ) → λ .
and
|λμ − ρ(t3 )| ≤ |λ| |μ − ρ(t1 )| + |ρ(t1 )| |λ − ρ(t2 )| < 2ε.
Hence λμ ∈ Fu,w .
We now fix x ∈ HC(Ψ ). Our aim is to show that x ∈ HC(Ψ (1, 1)). By Steps
1, 2 and 3, Fx,x is a nonempty closed subsemigroup of the multiplicative
group T.
Observe that, by property (α), Ψ (nk , jk )x ∈ orb(x, Ψ (1, 1)). Thus Ψ (0, 1)y ∈
orb(x, Ψ (1, 1)). Since Ψ (0, 1) has dense range by property (β), and y ∈ X is
arbitrary, x is hypercyclic for Ψ (1, 1).
For the rest of the proof we can now assume that Fx,x = T, and we will show
that this leads to a contradiction.
Step 5. There exists some m ∈ N such that, for each y ∈ HC(Ψ ), there is
λ ∈ T satisfying Fx,y = {λz ; z m = 1}.
We first note that Fx,x must be of the form Fx,x = {z ∈ T ; z m = 1} for
some m ∈ N. Indeed, if Fx,x contained points z = e2πti with t > 0 arbitrarily
small then, being a closed subsemigroup of T, Fx,x would be dense, and hence
coincide with T, which was excluded. Hence there is a minimal t0 ∈ ]0, 1] such
that z0 = e2πt0 i ∈ Fx,x . By the same argument, t0 cannot be irrational; see
6.4 Hypercyclic semigroup actions 171
Now let t ≥ 0. By property (β), Ψ (0, t) has dense range and therefore
Ψ (0, t)x ∈ HC(Ψ ). Since, by definition, ρ(t) = e2πti ∈ Fx,Ψ (0,t)x we conclude
that f (Ψ (0, t)x) = e2πmti .
Exercises
Exercise 6.1.1. In a metric space, show that a finite union of nowhere dense sets is
nowhere dense.
Exercise 6.1.3. Let T be an operator on a separable Fréchet space X that satisfies the
Hypercyclicity Criterion. Give two proofs of the fact that any T p , p ∈ N, also satisfies
the Hypercyclicity Criterion.
Exercise 6.1.4. Let T be a chaotic operator on a Fréchet space X. Without the use of
Theorem 6.2, show that any T p , p ∈ N, is also chaotic. This is not true for nonlinear
maps by the example of Exercise 1.2.11.
Exercise 6.1.8. An alternative proof of Exercise 6.1.7 (and thus of Ansari’s theorem)
is the following. With the notation of Theorem 6.2, let
Ak := (Dj1 ∩ · · · ∩ Djk ) ,
0≤j1 <···<jk ≤p−1
where k = 1, . . . , p. Prove
p−1the following assertions:
(i) A1 = D, Ap = j=0 Dj , and Ak+1 ⊂ Ak , k = 1, . . . , p − 1;
(ii) T (Ak ) ⊂ Ak , k = 1, . . . , p;
(iii) if Ak = D, then Ak+1 = D, k = 1, . . . , p − 1.
In particular, orb(x, T p ) is dense in X for every x ∈ D. (Hint: For (iii), observe that if
Ak+1 = D, then Ak+1 = ∅, since it is closed, T -invariant, and T : D → D is minimal;
hence Ak is a finite union of pairwise disjoint closed sets.)
Exercise 6.2.1. Prove assertions (i), (ii) and (iii) before Lemma 6.3. (Hint: See the
proof of Proposition 1.15.)
Exercise 6.2.2. Prove Lemma 6.3. (Hint: Follow the argument of Theorem 2.54.)
∞
orb(xj , T ) = X.
j=1
orb(xj , T ) = X.
j=1
Exercise 6.4.1. Let T = Bw be the weighted bilateral backward shift on 2 (Z) with
weights wn = n+1n if n ≥ 1 and w−n = n+1n+2 if n ≥ 0; see Section 4.1. Show that λT ,
λ ∈ C, is hypercyclic if and only if |λ| = 1. Discuss this result in the light of Kitai’s
theorem, showing first that σ(T ) ⊂ T. (Hint: For the first part note that λBw = Bλw ;
for the second part use the spectral radius formula for T and T −1 and Exercise 5.0.7.)
Exercise 6.4.4. Let X = C0 (R+ ), the space of continuous functions on R+ that van-
ish at ∞, endowed with the sup-norm. Consider the semigroup action Ψ (n, t)f (x) :=
2n−t f (x + t), n ∈ N0 , t ∈ R+ . Then Ψ is hypercyclic but the operator Ψ (1, 1) is not
hypercyclic. Which hypothesis of Theorem 6.10 is not satisfied?
Exercise 6.4.5. Give a new proof of Ansari’s theorem by proceeding as follows. Let T
be a hypercyclic operator on a Fréchet space X, x a hypercyclic vector for T and p ∈ N.
For u, v ∈ X define the subset Fu,v of T by
Fu,v = e2πji/p ; ∃ (nk )k ⊂ N0 with T nk p+j u → v, j = 0, . . . , p − 1 .
Fig. 6.3 Nonlinear dynamics if x has dense orbit under T but not under T p , m|p
Exercise 6.4.7. Show the following decomposition theorem. If x ∈ X has dense orbit
under T but not under T p , p > 1, then there is a divisor m > 1 of p and pairwise disjoint
open subsets S0 , S1 , . . . , Sm−1 of X with the following properties:
m−1
(i) S := j=0 Sj is dense in X;
(ii) T (Sj ) ⊂ Sj+1 , j = 0, . . . , m − 2, and T (Sm−1 ) ⊂ S0 ∪ (X \ S);
(iii) X \ S is invariant under T ;
(iv) for j = 0, . . . , m − 1, the orbit of T j x under T p is contained and dense in Sj ;
see Figure 6.3. m−2
(Hint: Consider the sets Dj of the previous exercise; set Sm−1 = X \ j=0 Dj , with
closure in X, and Sj = T −m+j+1 (Sm−1 ); show first that T −m (Sm−1 ) ⊂ Sm−1 , Dj ⊂ Sj ,
and that T n x ∈ Sj if and only if n = j(mod m).)
Exercise 6.4.8. Verify the results of the previous two exercises in the case of the map
of Exercise 1.2.11.
Sources and comments 177
The results in this chapter have in common that their proofs use connectedness argu-
ments. But their relationship runs deeper than that. As we have seen, Ansari’s theorem
is a consequence of the Costakis–Peris theorem, which in turn follows from the Bourdon–
Feldman theorem. Moreover, the theorems of León–Müller and Conejero–Müller–Peris
are proved by a common approach. Recently, Shkarin [284] was able to unify the latter
two theorems with Ansari’s theorem by deriving them as consequences of a single, quite
general result. An alternative common framework was developed by Bayart and Math-
eron [44], which was further generalized by Matheron [235] to include Shkarin’s result.
Section 6.1. Ansari [9] showed that powers of hypercyclic operators are hypercyclic. In-
dependently, Banks [28] proved a more general result: any power of a minimal map on a
connected topological space is also minimal (see Exercise 6.1.7). We combine ideas from
Banks [28] and Peris [254] for the proof of Theorem 6.2. Lemma 6.1 is from Peris [254].
Section 6.2. Theorem 6.5 is due to Bourdon and Feldman [93], answering a question
from Peris [254]. The corresponding result for semigroups of operators (see the next
chapter for this notion) is due to Costakis and Peris [121]. It is interesting to note that
for a weighted backward shift on p , 1 ≤ p < ∞, to be hypercyclic it already suffices to
have an orbit with a nonzero limit point, as was shown by Chan and Seceleanu [105];
such an orbit though, need not be dense.
Section 6.3. The fact that multi-hypercyclic operators are hypercyclic was indepen-
dently proved by Costakis [117] and Peris [254], answering a question raised by Her-
rero [195]. The original proofs motivated the question leading to the Bourdon–Feldman
theorem.
Section 6.4. Theorem 6.7 on rotations of hypercyclic operators is due to León and
Müller [222]. Bayart and Bermúdez [37] show that the corresponding result for chaos
fails. Badea, Grivaux and Müller [20] characterize the subsets of C that can appear as
{λ ∈ C ; λT hypercyclic} for invertible operators T on a complex Hilbert space.
Theorem 6.8 on discretizations of hypercyclic C0 -semigroups is due to Conejero,
Müller and Peris [110]. Exercise 6.4.4 shows that the result fails for semigroups indexed
over N0 × R+ ; see also Shkarin [284] and Exercise 7.3.1. Bayart [36] shows that it even
fails for holomorphic groups over C. And by Bayart and Bermúdez [37] there are chaotic
C0 -semigroups on a Hilbert space for which no individual operator is chaotic.
The unified proof of Theorem 6.10 essentially follows the argument of [110]. The re-
lated approach to Ansari’s theorem in Exercise 6.4.5 is due to Grosse-Erdmann, León
and Piqueras [183]. As mentioned above, Shkarin [284], Bayart and Matheron [44] and
Matheron [235] obtain much more general results that contain the theorems of Ansari,
León–Müller and Conejero–Müller–Peris as special cases. Shkarin and Matheron point
out that the main common idea in all these proofs can already be found in a paper by
Furstenberg [156].
Exercises. Exercise 6.1.6 is taken from Montes and Salas [243], Exercise 6.1.7 from
Banks [28]. Exercise 6.1.8 outlines essentially the original proof of Ansari [9]. Exercises
6.2.2 and 6.2.3 are taken from Bourdon and Feldman [93], while the result of Exercise
6.2.6 is due to Peris and Saldivia [257]. The result of Exercise 6.3.1 is due to Herrero
and Kitai [196]. The notion of a countably hypercyclic operator (see Exercise 6.3.3), as
well as the results of Exercises 6.3.4 and 6.3.5 are due to Feldman [149]. Exercise 6.4.1
is taken from León and Müller [222], Exercise 6.4.2 from Shkarin [284]. Exercises 6.4.6
and 6.4.7 are due to Grosse-Erdmann, León and Piqueras [183] (see also Marano and
178 6 Connectedness arguments in linear dynamics
Lemma 7.1. Let (Tt )t≥0 be a family of operators on X. Then the following
assertions are equivalent:
(i) lims→t Ts x = Tt x for all x ∈ X and t ≥ 0;
(ii) (Tt )t≥0 is locally equicontinuous and there is a dense subset X0 of X
such that lims→t Ts x = Tt x for all x ∈ X0 and t ≥ 0;
(iii) the map
R+ × X → X, (t, x) → Tt x
is continuous.
Proof. It suffices to show that (ii) implies (iii). This implication follows im-
mediately from the identity
for x, y ∈ X, x0 ∈ X0 and s, t ≥ 0.
Remark 7.2. Local equicontinuity implies, via (7.1), that Ttn xn → 0 whenever
(tn )n is bounded and xn → 0. This observation will be used repeatedly.
Moreover, one can establish an exponential bound for the operator norm
of the semigroup.
where ∞ 1/p
f = |f (x)|p v(x) dx .
0
The translation semigroup is then given by
Tt f (x) = f (x + t), t, x ≥ 0.
We claim that this defines a C0 -semigroup on Lpv (R+ ) if and only if there
exist M ≥ 1 and w ∈ R such that, for all t ≥ 0,
has Lebesgue measure λ(B) > 0. Let b > 0 be such that λ(B ∩ [0, b]) > 0,
and define f (x) = v(x)11/p if x ∈ t0 + (B ∩ [0, b]), and f (x) = 0 otherwise. It
is clear that f ∈ X and f > 0. On the other hand
184 7 Dynamics of semigroups, with applications to differential equations
p
Tt0 f = |f (x + t0 )|p v(x) dx
B∩[0,b]
p
≥ 2M e
p pwt0
|f (x + t0 )|p v(x + t0 ) dx = 2M p epwt0 f ,
B∩[0,b]
1
Ax := lim (Tt x − x)
t→0 t
Ax = λx =⇒ Tt x = eλt x (7.4)
for every t ≥ 0.
As an example, the generator (A, D(A)) of the translation semigroup on
X = Lpv (R+ ), 1 ≤ p < ∞, can be shown to be
7.2 Hypercyclic and chaotic C0 -semigroups 185
u(·, x) : R+ → X, u(t, x) := Tt x,
the function that describes the orbit of x under the semigroup. Whenever
x ∈ D(A), u(·, x) is the unique solution of the abstract Cauchy problem
dt u(t) = Au(t) for t ≥ 0,
d
(ACP)
u(0) = x.
Using the fact that X is separable, we easily deduce from the definitions
that hypercyclicity of a C0 -semigroup is equivalent to hypercyclicity of (Ttn )n
for some positive sequence (tn )n with tn → ∞; see Definition 3.23. Moreover,
hypercyclicity and topological transitivity for C0 -semigroups are equivalent
notions; see Exercise 7.2.1.
One should note that, for any C0 -semigroups (Tt )t≥0 and (St )t≥0 on Ba-
nach spaces X and Y , respectively, (Tt ⊕ St )t≥0 is a C0 -semigroup on X ⊕ Y .
As in the discrete case, the direct sum of a mixing semigroup with a hyper-
cyclic semigroup is hypercyclic.
Finally, we also have a concept of chaos.
on [0, 1], and 0 otherwise. Then we have for any f ∈ X with f < (2C)−1
and any t ≥ 0 that
1 1/p
Tt f − g ≥ |f (x + t) − g(x)|p v(x) dx
0
1 1/p 1 1/p
≥ |g(x)|p v(x) dx − |f (x + t)|p v(x) dx
0 0
1 1/p 1
≥1−M |f (x + t)|p v(x + t) dx ≥ 1 − Cf > .
0 2
2 wt0 v(t0 + t)
=M e f2 p < ε,
v(0)
We need only modify the arguments in (a) slightly. Thus, suppose that (ii)
fails, that is, there is some sequence (tn )n with tn → ∞ such that (v(tn ))n is
bounded away from zero. The ∞admissibility condition (7.3) implies that v is
bounded away from zero on n=1 [tn , tn + 1]. But then there is some C > 0
such that v(x) ≤ C p v(x + tn ) for all x ∈ [0, 1], n ≥ 1, and the argument in
(a) shows that the sequence of operators (Ttn )n cannot be hypercyclic, which
contradicts (i). The proof that (ii) implies (i) is exactly as in (a).
(c) Let X = C0 (R+ ) be the space of continuous functions f : R+ → K
with limx→∞ f (x) = 0, endowed with the sup-norm f = supx∈R+ |f (x)|.
Given a constant w > 0, we consider the family of operators defined by
(Tt f )(x) = ewt f (x + t), x ∈ R+ , which is easily seen to define a C0 -semigroup
on X.
We claim that this semigroup is mixing and chaotic. To this end, let X0
be the dense subspace of continuous functions with compact support, and let
f1 ∈ X0 , f2 ∈ X and ε > 0. For any t > t0 := max supp(f1 ) we define gt ∈ X
by ⎧
⎪f1 (x)
⎨ if x ≤ t0 ,
gt (x) = e−wt f2 (0) x−t 0
if t0 < x < t,
⎪
⎩ −wt
t−t0
e f2 (x − t) if x ≥ t.
Then we have that Tt gt = f2 and
for any t > t0 sufficiently large. As in (a), this yields the mixing property for
the semigroup.
In order to show that the semigroup is even chaotic, it suffices to prove
that, for any f ∈ X0 and ε > 0, there is a periodic point h ∈ X with
f − h < ε. To this end, for any t > t0 := max supp(f ), we define the
function ht ∈ X by
e−wkt f (x − kt) if kt ≤ x ≤ kt + t0 , k ∈ N0 ,
ht (x) = −w(k+1)t
e x−kt−t0
f (0) t−t0 if kt + t0 < x < (k + 1)t, k ∈ N0 .
Definition 7.11. Let (Tt )t≥0 and (St )t≥0 be C0 -semigroups on separable
Banach spaces X and Y , respectively.
7.2 Hypercyclic and chaotic C0 -semigroups 189
(a) Then (Tt )t≥0 is called quasiconjugate to (St )t≥0 if there exists a con-
tinuous map φ : Y → X with dense range such that Tt ◦ φ = φ ◦ St for all
t ≥ 0. If φ can be chosen to be a homeomorphism then (Tt )t≥0 and (St )t≥0
are called conjugate.
(b) A property P of C0 -semigroups is said to be preserved under (quasi)
conjugacy if any C0 -semigroup that is (quasi)conjugate to a C0 -semigroup
with property P also possesses property P.
Case 2: |λ| ≥ 1. On the one hand, let r > 0 be such that |Tr x, x∗ | > 1.
On the other hand, by equicontinuity of (Ts )s∈[0,t] , there is some ε > 0 such
that |Ts y, x∗ | < 1 if s ∈ [0, t] and y < ε. We can also find t > r such that
Tt x < ε. If we write t = mt − s + r for some m ∈ N and s ∈ [0, t], then,
putting all these facts together, we obtain that
which is a contradiction.
Next, let (A, D(A)) be the generator of (Tt )t≥0 , and suppose that A∗ has
an eigenvalue λ ∈ K. Let x∗ ∈ X ∗ \ {0} be such that Ax, x∗ = λx, x∗
for all x ∈ D(A). Given x ∈ D(A), we define h(t) = Tt x, x∗ , t ≥ 0. Then
Tt x ∈ D(A) and h (t) = ATt x, x∗ = λTt x, x∗ = λh(t) for all t ≥ 0; see
Exercise 7.1.7. Therefore,
We can also apply Theorem 7.16 to obtain a necessary condition for chaos
of a C0 -semigroup on complex Banach spaces; compare with Proposition 5.7.
σp (A) ∩ iR
N
2πnk i
Per((Tt )) ⊂ span ker I −A . (7.5)
tk
k=1
It follows from (7.5) and (7.6) that the operator T := (I − Tt1 ) · · · (I − TtN )
vanishes on Per((Tt )). Since this set is dense we have that T = 0. But this
contradicts Theorem 7.16 by which T has dense range.
(i) the
∞ translation semigroup is chaotic;
(ii) 0 v(x) dx < ∞.
We will see below that a semigroup on a real Banach space is chaotic if and
only if its complexification is; see Corollary 7.24. Therefore we may assume
that we are dealing with the complex space X. Thus, if the semigroup is
chaotic, Proposition 7.18 implies that its generator has an eigenvalue on the
imaginary axis. Since the generator is the derivative, we can find some s ∈ R
such that the function x → eisx , x ∈ R+ , belongs to X. This yields (ii).
Conversely, suppose that v is integrable on R+ . By Example 7.10(a), the
semigroup is then hypercyclic. Now let f be a continuous function of compact
support on R+ and ∞ ε > 0. Setting K = 1 + maxx∈R+ |f (x)|, we can find some
t > 0 such that t v(x) dx < (ε/2K)p . We define g(x + nt) = f (x), x ∈ [0, t[,
n ∈ N0 . The function g is bounded, thus g ∈ X by the integrability of the
weight. Moreover, Tt g = g and
∞ 1/p ∞ 1/p
f − g ≤ |f (x)|p v(x) dx + |g(x)|p v(x) dx
t
∞
t
1/p
≤ 2K v(x) dx < ε.
t
Since the continuous functions of compact support are dense in X, the semi-
group has a dense set of periodic points and is therefore chaotic.
Proof. The implications (i) =⇒ (ii) =⇒ (iii) =⇒ (iv) and (ii) =⇒ (v) =⇒
(vi) are trivial.
(iv) =⇒ (i). Suppose that (Tt )t≥0 is not mixing. Then there exists a pair
of nonempty open subsets U, V of X and a sequence (tn )n tending to ∞ such
that Ttn (U ) ∩ V = ∅ for all n ≥ 1. Then (Ttn )n cannot be topologically
transitive.
(vi) =⇒ (i). Let (Tnt0 )n be a mixing autonomous discretization of (Tt )t≥0 .
We fix arbitrary nonempty open sets U, V ⊂ X. By Lemma 2.36 there are
nonempty open sets U1 ⊂ U , V1 ⊂ V and a 0-neighbourhood W such that
U1 + W ⊂ U and V1 + W ⊂ V . By local equicontinuity of the semigroup
there is a 0-neighbourhood W1 such that Ts (W1 ) ⊂ W for all s ∈ [0, t0 ]. The
hypothesis then implies the existence of some N ∈ N such that
Tt u1 = Ts Tnt0 u1 ∈ Ts (W1 ) ⊂ W,
Tt w = T(n+1)t0 w1 ∈ V1 ,
But since its proof is highly nontrivial we will provide here a simpler proof of
a classical result that is somewhat weaker but still strong enough for many
applications.
Each Jk is an open subset of ]0, ∞[. We will see that it is also dense. Indeed,
if 0 < a < b < ∞, there is some n0 ∈ N such that n0 b > (n0 + 1)a. Then
]na, nb[ = ]n0 a, ∞[.
n≥n0
which is a dense Gδ -subset of ]0, ∞[; note that the Baire category theorem is
applicable since ]0, ∞[ is isomorphic to R. Let t ∈ J. Then, for every k ∈ N,
there exists some n ∈ N such that Tnt x ∈ Uk . That is, x is a hypercyclic
vector for Tt .
The theorem yields that, as in the discrete case, chaos implies weak mixing.
Proof. Suppose that the semigroup (Tt )t≥0 on a real Banach space X is
chaotic. Since its complexification (Tt )t≥0 can be identified with the direct
sum (Tt ⊕ Tt )t≥0 it is hypercyclic by the preceding result. Moreover, if U
and V are nonempty open subsets of X then, by topological transitivity and
continuity, there is a nonempty open subset U1 ⊂ U and some t ≥ 0 such
that Tt (U1 ) ⊂ V . We can then find some periodic point x ∈ U1 . Let Ts x = x,
s > 0. Then y := Tt x ∈ V and Ts y = Tt Ts x = y. Therefore, (x, y) ∈ U × V
is a periodic point for (Tt ⊕ Tt )t≥0 , which concludes the proof that (Tt )t≥0 is
chaotic. The converse implication was obtained in Corollary 7.13.
Proof. It suffices to show that (ii) implies (iii). In view of the Bès–Peris
theorem we have to show that if some Tt0 , t0 > 0, satisfies the Hypercyclicity
Criterion, then so does every Tt1 , t1 > 0. Since Tt1 = Ttp1 /p , p ∈ N, we may
assume that t1 < t0 ; see Exercise 6.1.3.
Now, Theorem 3.22 tells us that Tt0 even satisfies the Gethner–Shapiro
criterion; that is, there are dense subsets X0 , Y0 of X, an increasing sequence
(nk )k of positive integers and a mapping St0 : Y0 → Y0 such that, for any
x ∈ X0 and y ∈ Y0 , Ttn0k x → 0, Stn0k y → 0, and Tt0 St0 y = y. As the proof of
Theorem 3.22 shows, Y0 can be taken in the form Y0 = {yn ; n ∈ N}, where
the yn are pairwise distinct and St0 yn = yn+1 for all n ∈ N.
Since t1 < t0 there are mk ∈ N0 and sk ∈ [0, t0 [, k ∈ N, such that
mk t1 = nk t0 + sk , k ∈ N.
We claim that also Tt1 satisfies the Hypercyclicity Criterion, with respect to
X0 , Y0 , and the sequence (mk )k .
Indeed, for every x ∈ X0 we have that
Ttm
1
k
x = Tsk Ttn0k x → 0
as k → ∞, where we use the fact that the family (Ts )s∈[0,t0 ] is equicontinuous.
Next we define a family (St )t≥0 of maps on Y0 by setting St yn = Ts yn+k
if t = kt0 − s for some k ∈ N and 0 < s ≤ t0 ; note that this is consistent with
St0 . Then we have for every yn ∈ Y0 , n ∈ N, that
Smk t1 yn = S(nk +1)t0 −(t0 −sk ) yn = Tt0 −sk yn+nk +1 = Tt0 −sk Stn0k yn+1 → 0
196 7 Dynamics of semigroups, with applications to differential equations
as k → ∞, again by equicontinuity.
Finally, for every yn ∈ Y0 , n ∈ N, we obtain that
Ttm
1
k
Smk t1 yn = Tmk t1 Tt0 −sk Stn0k yn+1 = Ttn0k +1 Stn0k +1 yn = yn .
Proof. Suppose that (Tt )t≥0 is weakly mixing. Then, by Proposition 7.25, the
operator T1 is weakly mixing and therefore hereditarily hypercyclic, so that
the same holds for (Tt )t≥0 .
On the other hand, if (Tt )t≥0 is hereditarily hypercyclic then there is some
sequence (tn )n in R+ with tn → ∞ such that (Ttn )n is a hereditarily transitive
sequence of operators. By Theorem 3.25, (Tn )n satisfies the Hypercyclicity
Criterion for sequences of operators (Theorem 3.24), which implies that the
semigroup (Tt )t≥0 satisfies the Hypercyclicity Criterion.
If, in the Hypercyclicity Criterion, one has convergence along the whole
real line then we obtain a criterion for mixing.
Theorem 7.29. Let (Tt )t≥0 be a C0 -semigroup on X. If there are dense sub-
sets X0 , Y0 ⊂ X, and maps St : Y0 → X, t ≥ 0, such that, for any x ∈ X0 ,
y ∈ Y0 ,
(i) Tt x → 0,
(ii) St y → 0,
(iii) Tt St y → y,
then (Tt )t≥0 is mixing.
We will now consider complex Banach spaces. In the discrete case, the
Godefroy–Shapiro criterion tells us that a large supply of eigenvectors of an
operator T to eigenvalues in an open set intersecting the unit circle leads
to mixing and chaos of T . An analogous result is true when one studies
eigenvectors of the generator A to eigenvalues in an open set that intersects
the imaginary axis.
By a weakly holomorphic function f : U → X on an open set U ⊂ C we
understand an X-valued function such that, for every x∗ ∈ X ∗ , the complex-
valued function z → f (z), x∗ is holomorphic on U . In the sequel, J is a
nonempty index set.
Theorem 7.30. Let X be a complex separable Banach space, and (Tt )t≥0 a
C0 -semigroup on X with generator (A, D(A)). Assume that there exists an
open connected subset U and weakly holomorphic functions fj : U → X,
j ∈ J, such that
(i) U ∩ iR
= ∅,
(ii) fj (λ) ∈ ker(λI − A) for every λ ∈ U , j ∈ J,
(iii) for any x∗ ∈ X ∗ , if fj (λ), x∗ = 0 for all λ ∈ U and j ∈ J then x∗ = 0.
Then the semigroup (Tt )t≥0 is mixing and chaotic.
198 7 Dynamics of semigroups, with applications to differential equations
Proof. We will use the Hahn–Banach theorem to prove the density of certain
subspaces generated by eigenvectors, and then apply the Godefroy–Shapiro
criterion.
For a fixed t > 0, we consider the subspaces
We first show that these spaces are dense in X. To this end, suppose that
x∗ ∈ X ∗ is identically 0 on X0 . For any j ∈ J, the function λ → fj (λ), x∗
is holomorphic on U , and, in view of (i), it vanishes on some nonempty open
subset of U . Therefore it vanishes on U , for any j ∈ J. By (iii) we have that
x∗ = 0. The Hahn–Banach theorem then implies that X0 is dense in X. In
the same way it follows that also Y0 and Z0 are dense.
Now, by (7.4), we have that
Tt x = eλt x
for all x ∈ ker(λI −A). Therefore, by (ii), X0 and Y0 are contained in the span
of the eigenvectors of Tt to eigenvalues of modulus greater than 1 and smaller
than 1, respectively, and Z0 is contained in the span of the eigenvectors of
Tt to eigenvalues that are roots of unity. The Godefroy–Shapiro criterion
yields that Tt is mixing and chaotic, properties that are inherited by the
C0 -semigroup; see Proposition 7.21.
One can strengthen the above criterion in the sense that the selection map
f need not be defined on an open subset of C, and we can replace weak
holomorphy by continuity, as we will see next.
First of all, for any continuous vector-valued function f : [a, b] → X one
can define the Riemann integral
b
f (t) dt
a
in the usual way by its Riemann sums; see Appendix A for more details and
basic properties.
the step function that takes the value f (tj ) on ]tj−1 , tj ], j = 1, . . . , N . Then
we find that
b b b
eirt
f (t) dt ≤
e irt
(f (t) − g(t)) dt + eirt g(t) dt
a a a
N tj
≤ (b − a)ε + f (tj ) eirt dt.
j=1 tj−1
tj
Now, since tj−1
eirt dt → 0 as r → ±∞, the claim follows.
Theorem 7.32. Let X be a complex separable Banach space, and (Tt )t≥0 a
C0 -semigroup on X with generator (A, D(A)). Assume that there are a < b
and continuous functions fj : [a, b] → X, j ∈ J, such that
(i) fj (s) ∈ ker(isI − A) for every s ∈ [a, b], j ∈ J,
(ii) span{fj (s) ; s ∈ [a, b], j ∈ J} is dense in X.
Then the semigroup (Tt )t≥0 is mixing and chaotic.
Proof. We apply Theorem 7.29 to show that the semigroup is mixing. For
b
any r ∈ R, j ∈ J, we define xr,j = a eirs fj (s) ds, and we set
X0 = Y0 = span{xr,j ; r ∈ R, j ∈ J}.
To prove that this subspace is dense in X we will again use the Hahn–Banach
theorem. Thus let x∗ be a continuous linear functional on X such that, for
all r ∈ R, j ∈ J,
b
xr,j , x∗ = eirs fj (s), x∗ ds = 0.
a
The functions s → fj (s), x∗ are continuous on [a, b] and therefore belong to
1 2π
L2 [a, b]. Since ( √b−a exp ( b−a ikt))k∈Z is an orthonormal basis in this Hilbert
space we deduce that, in view of continuity,
so that
b b
irs
Tt xr,j = e Tt fj (s) ds = ei(t+r)s fj (s) ds = xt+r,j . (7.8)
a a
m
ut = ak xrk −t,jk .
k=1
By (7.8) we then have that Tt ut = y for all t ≥ 0 and, again by the Riemann–
Lebesgue lemma, that ut → 0 as t → ∞.
Thus, (Tt )t≥0 is a mixing semigroup.
Finally, by continuity of the functions fj , also
Example 7.33. Let (Tt )t≥0 be the translation semigroup on the complex Ba-
nach space X = Lpv (R+ ), 1 ≤ p < ∞, where v(x) = e−x . The dual of X
is given by the space Lqw (R+ ), where 1 < q ≤ ∞ satisfies p1 + 1q = 1, with
w = v −q/p for p > 1 and w = v−1 for p = 1 (see Example A.3(c)); note
that L∞w (R+ ) is the space of measurable functions g on R+ such that gw is
essentially bounded.
We define f : D → X by f (λ)(x) = eλx , λ ∈ D, x ∈ R+ . Then f is well
defined and weakly holomorphic on D. Clearly Af (λ) = dx d
f (λ) = λf (λ). If
∗
a continuous linear functional on X , represented by g ∈ Lqw (R+ ), satisfies
∞
f (λ), g = g(x)eλx dx = 0
0
for all λ ∈ D, then we obtain, by taking the nth derivative with respect to λ,
∞
xn g(x)eλx dx = 0, n ∈ N0 , λ ∈ D.
0
1 + (x + t)2
Tt ϕ(x) = ϕ(x + t), x, t ∈ R+ .
1 + x2
That is, the action of the semigroup is a translation, together with multipli-
2
cation by the function ht (x) = 1+(x+t)
1+x2
. The fact that ht (x) → ∞ as t → ∞
opens up the possibility of approximating arbitrary functions by small func-
tions under the action of Tt , and thus of obtaining the mixing property for
the semigroup. On the other hand, the fact that ht (x) → 1 as x → ∞ pre-
vents eigenvectors of Tt to eigenvalues of modulus greater than 1. Therefore
the eigenvalue criterion of Theorem 7.30 cannot be applied.
Proposition 7.34. The solution semigroup (Tt )t≥0 of (7.9) is mixing and
chaotic on L1 (R+ ).
Proof. We first apply the mixing criterion, Theorem 7.29. We take for X0
the dense subspace of X = L1 (R+ ) of functions with compact support. If the
202 7 Dynamics of semigroups, with applications to differential equations
1 + (x − t)2
St ϕ(x) = ϕ(x − t), for x ≥ t ≥ 0,
1 + x2
and St (x) = 0 for 0 ≤ x < t. Then Tt St ϕ = ϕ for every ϕ ∈ X. Moreover,
∞ ∞
1 + (x − t)2 1 + x2
St ϕ = 2
|ϕ(x − t)| dx = |ϕ(x)| dx → 0
t 1+x 0 1 + (x + t)2
endowed with the norm f = supn≥0 |an |, where c0 is the Banach space
of complex sequences tending to 0. Then Xρ is a Banach space of analytic
functions with a certain growth control. By its definition it is isometrically
isomorphic to c0 .
Since ⎛ ⎞
0 I
A := ⎝ α ∂ 2 1 ⎠
− I
τ ∂x2 τ
is easily seen to be an operator on X := Xρ ⊕ Xρ , we have that (etA )t≥0 is a
C0 -semigroup on X, which is the solution semigroup of (7.10) on X.
Proposition 7.35. Let ρ > 0 be such that ατ ρ2 > 2. Then the solution
semigroup (etA )t≥0 of (7.10) is mixing and chaotic on Xρ ⊕ Xρ .
∞ ∞
Rλn x2n Rλn x2n+1
ϕλ,z0 ,z1 (x) = z0 + z1 , x ∈ R.
n=0
(2n)! n=0
(2n + 1)!
Let U be the open disk of radius r = αρ2 /2τ > 0 centred at zero. If λ ∈ U
then
2
αρ2 !
|Rλ | τ αρ
2τ + 2τ 1 1 2
< = + < 1, (7.12)
ρ2 αρ2 2 2 ατ ρ2
so that ϕλ,z0 ,z1 ∈ Xρ . We now consider the functions fz0 ,z1 : U → X, z0 , z1 ∈
R, given by
ϕλ,z0 ,z1
fz0 ,z1 (λ) = .
λϕλ,z0 ,z1
Then one finds that
According to Theorem 7.30 it remains to prove that, for any x∗ ∈ Xρ∗ ⊕ Xρ∗ ,
the functions λ → fz0 ,z1 (λ), x∗ , z0 , z1 ∈ R, are holomorphic on U , and that
if they all vanish on U then x∗ = 0.
Thus, let x∗ ∈ Xρ∗ ⊕ Xρ∗ . By the isomorphism of Xρ with c0 , x∗ can be
represented in a canonical way by a pair ((bn )n≥0 , (cn )n≥0 ) ∈
1 ⊕
1 . We
then have that, for λ ∈ U ,
where (αn )n , (βn )n are bounded positive sequences and (an )n ∈
1 is a real
sequence.
In this case we consider the real Banach space X =
1 and the map A
given by
Af = (−αn fn + βn fn+1 )n for f = (fn )n .
Since A is an operator on
1 , it generates a C0 -semigroup (Tt )t≥0 , which is
then the solution semigroup of the abstract Cuachy problem (7.13).
Then the solution semigroup (Tt )t≥0 of (7.13) is mixing and chaotic on 1 .
Proof. Let α = supn≥1 αn , β = lim inf n→∞ βn and α/2 < r < β/2. We fix
U ⊂ C as the open disk of radius r centred at −α/2, which intersects the
imaginary axis iR. We can calculate the eigenvectors of A to the eigenvalues
λ ∈ U . Indeed, suppose that f = (fn )n satisfies Af = λf , λ ∈ U . Then
and therefore
"
n−1
λ + αk
fn = γn f1 with γn = , n ≥ 1,
βk
k=1
∞ n−1
" λ + αk
f (λ), x∗ = bn
n=1
βk
k=1
Exercises
Exercise 7.1.1. Let Δ = Δ(α) := {reiθ ; r ≥ 0, |θ| ≤ α}, 0 < α ≤ π/2, be a sector
in the complex plane, and v : Δ → R a strictly positive measurable function for which
206 7 Dynamics of semigroups, with applications to differential equations
there exist constants M ≥ 1 and w ∈ R such that v(z1 ) ≤ M ew|z2 | v(z1 + z2 ) for all
z1 , z2 ∈ Δ, called an admissible weight function on Δ. For 1 ≤ p < ∞ we define the
space
Lpv (Δ) = {f : Δ → K ; f is measurable and f < ∞}
with f = ( Δ |f (z)|p v(z) dλ(z))1/p , where λ denotes two-dimensional Lebesgue mea-
sure. Show that, for every z ∈ Δ, the translation semigroup (Ttz )t≥0 given by Ttz f (ζ) =
f (ζ + tz) is a C0 -semigroup on Lpv (Δ).
endowed with the norm f = supx≥0 |f (x)|v(x). Consider the translation semigroup
given by Tt f (x) = f (x + t), t, x ≥ 0. Show that this defines a C0 -semigroup on X if and
only if v is admissible in the sense of (7.3).
Exercise 7.1.3. For the operator A on R2 given by the matrix 01 −1 0 determine the
semigroup (etA )t≥0 , and verify that A is its infinitesimal
generator. Describe the orbits
{etA x ; t ≥ 0}, x ∈ R2 . Do the same for −1 1 1
−1 .
Exercise 7.1.4. Let ϕ be a bounded holomorphic function on the unit disk D. Then
Tt f = etϕ f , t ≥ 0, defines multiplication operators on the Hardy space H 2 ; see Section
4.4. Show that (Tt )t≥0 is a uniformly continuous semigroup on H 2 , and identify its
generator. Do the same for the family (Tt∗ )t≥0 of Hilbert space adjoints.
Exercise 7.1.5. Let T be an operator on a Banach space X such that T < 1. Then
∞ (−1)n+1 n
the series R := n=1 n T converges in the operator norm, so that R ∈ L(X).
Prove that eR = I + T .
Exercise 7.1.6. Let A be an operator on a Banach space X. Show that the operators
Tt = etA , t ≥ 0, define a C0 -semigroup on X. Moreover, show that lims→t Ts − Tt = 0
and Ax = limt→0 1t (Tt x − x) for all x ∈ X.
Exercise 7.1.7. Let (Tt )t≥0 be a C0 -semigroup on a Banach space X and (A, D(A))
its infinitesimal generator. Show the following for any t ≥ 0:
(i) for any x ∈ D(A), Tt x ∈ D(A) and ATt x = Tt Ax;
(ii) for any x ∈ D(A), lims→t Ts x−T tx
= ATt x;
t s−t
t
(iii) for any x ∈ X, Ts x ds ∈ D(A) and A Ts x ds = Tt x − x;
0 t 0
(iv) for any x ∈ D(A), 0 Ts Ax ds = Tt x − x,
where the integral is the vector-valued Riemann integral; see Appendix B. Deduce that
t+h t
D(A) is dense and that A has closed graph. (Hint for (ii): h Ts x ds − 0 Ts x ds =
t+h h
t
Ts x ds − 0
Ts x ds.)
Exercise 7.1.8. Prove Proposition 7.5. (Hint: For (i) =⇒ (ii), use Lemma 8.20 to show
t
that 1t 0 Ts ds is an invertible operator if t > 0 is sufficiently small, and use Exercise
7.1.7(ii). For (ii) =⇒ (iii), use the closed graph theorem and the fact that the infinitesimal
generator determines the C0 -semigroup.)
(b) Prove the Birkhoff transitivity theorem for C0 -semigroups, that is: (Tt )t≥0 is
hypercyclic if and only if it is topologically transitive. In that case, the set of hypercyclic
vectors for (Tt )t≥0 is a dense Gδ -set.
Exercise 7.2.3. Let (Tt )t≥0 be the translation semigroup on X = Lpv (R+ ) for an ad-
missible weight v. Establish the equivalence of the following assertions:
(i)
(Tt )t≥0 is chaotic;
∞
(ii) v(n) < ∞;
n=1 ∞
(iii) for all ε > 0 and t > 0, there is some s > 0 such that n=1
v(t + ns) < ε.
Exercise 7.2.4. Let (Tt )t≥0 be the translation semigroup on X = C0,v (R+ ) for an
admissible continuous weight function v; see Exercise 7.1.2. Show that (Tt )t≥0 is hyper-
cyclic if and only if lim inf x→∞ v(x)=0. And that it is mixing, if and only if it is chaotic,
and if and only if limx→∞ v(x)=0. (Hint: Use the fact that the generator A is given by
Af = f on the subspace of continuously differentiable functions f ∈ X with f ∈ X.)
Exercise 7.2.5. Let (Tt )t≥0 be the translation semigroup on the space X = Lpv (Δ); see
Exercise 7.1.1. Let Δm = {z ∈ Δ ; Im(z) ∈ [−m, m]}, m ∈ N. Demonstrate that the
following are equivalent:
translation semigroup is chaotic on X;
(i) the
(ii) Δ v(z) dλ(z) < ∞ for every m ∈ N.
m
Show that there is a mixing nonautonomous discretization (Tzn )n of (Tz )z∈Δ , but
that no C0 -subsemigroup (Ttz )t≥0 , z ∈ Δ, is hypercyclic. In particular, no autonomous
208 7 Dynamics of semigroups, with applications to differential equations
Exercise 7.3.4. Prove that a C0 -semigroup (Tt )t≥0 is weakly mixing if and only if, for
every pair U, V ⊂ X of nonempty open sets, R(U, V ) contains arbitrarily long intervals.
(Hint: For one implication take any autonomous discretization and apply Theorem 1.54;
the other implication is based on the local equicontinuity of the semigroup and, again,
on Theorem 1.54.)
Exercise 7.3.5. Prove that a C0 -semigroup (Tt )t≥0 is weakly mixing if and only if there
exists some ε > 0 such that, for any pair U, V ⊂ X of nonempty open sets, R(U, V )
contains some interval of length ε. (Hint: Exercise 7.3.4, and Exercise 1.5.8 for Tt0 with
t0 ∈ ]0, ε/2[.)
Exercise 7.3.6. A C0 -semigroup (Tt )t≥0 of operators is topologically ergodic if, for ev-
ery pair U, V ⊂ X of nonempty open sets, R(U, V ) is syndetic, that is, R+ \ R(U, V )
does not contain arbitrarily long intervals. Show that every autonomous discretization
of a topologically ergodic C0 -semigroup is topologically ergodic. (Hint: Given t > 0,
nonempty open sets U, V ⊂ X, and an arbitrary 0-neighbourhood W , use local equicon-
tinuity to prove that N (U, W ) = {n ∈ N0 ; nt ∈ R(U, W )} and N (W, V ) = {n ∈
N0 ; nt ∈ R(W, V )} are syndetic. Then show that N (U, W ) ∩ N (W, V ) is (nonempty
and) syndetic.)
Exercise 7.3.7. Prove that, if (Tt )t≥0 and (St )t≥0 are topologically ergodic C0 -semi-
groups, then (Tt ⊕ St )t≥0 is also topologically ergodic. (Hint: Exercises 7.3.6 and 2.5.5.)
Exercise 7.4.1. Show directly that under the hypotheses of Theorem 7.27 the semi-
group is topologically transitive. Prove Theorem 7.29 in the same way.
Exercise 7.4.2. By using a direct argument, prove that if a C0 -semigroup (Tt )t≥0 sat-
isfies the Hypercyclicity Criterion then the operator T1 satisfies the Hypercyclicity Cri-
terion. (Hint: See the proof of Proposition 7.25.)
Exercise 7.4.3. Let (Tt )t≥0 be a C0 -semigroup. Show that the following assertions are
equivalent:
Exercises 209
(i) (Tt )t≥0 satisfies the criterion for mixing of Theorem 7.29;
(ii) every discretization of (Tt )t≥0 satisfies the Hypercyclicity Criterion for sequences
of operators for the full sequence (n);
(iii) some autonomous discretization of (Tt )t≥0 satisfies the Hypercyclicity Criterion
for sequences of operators for the full sequence (n).
Exercise 7.4.4. Suppose that a C0 -semigroup (Tt )t≥0 admits a discretization (Ttn )n∈N
with supn≥1 (tn+1 − tn ) < ∞ satisfying the Hypercyclicity Criterion for sequences of
operators for the full sequence (n). Prove that (Tt )t≥0 satisfies the criterion for mixing
of Theorem 7.29.
Exercise 7.4.5. Show the following generalization of Theorem 7.32. Let X be a complex
separable Banach space, and (Tt )t≥0 a C0 -semigroup on X with generator (A, D(A)).
Assume that there are compact intervals Ij = [aj , bj ] with aj < bj and continuous
functions fj : Ij → X, j ∈ J, such that
(i) fj (s) ∈ ker(isI − A) for every s ∈ Ij , j ∈ J,
(ii) span{fj (s) ; s ∈ Ij , j ∈ J} is dense in X.
Then the semigroup (Tt )t≥0 is mixing and chaotic.
Exercise 7.5.1. Provide an alternative proof of Proposition 7.34 for complex functions
by applying Exercise 7.4.5 to fj : [−j, j] → X, fj (s)(x) = exp (isx)
1+x2 , j ∈ N. (Hint: For
condition (ii), use the Hahn–Banach theorem and the fact that if the Fourier transform
of an integrable function vanishes identically then so does the function itself.)
Show that, on Lp (R+ ), the semigroup is chaotic if and only if 0 exp −p 0 h(s) ds dx
x
<∞, mixing if and only if limx→∞ 0 h(s) ds = ∞, and hypercyclic if and only if
x
supx≥0 0 h(s) ds = ∞. Find the corresponding characterizations on C0 (R+ ). (Hint:
Determine an admissible weight function v so that the semigroup is conjugate to the
translation semigroup on Lpv (R+ ) or on C0,v (R+ ); see Exercise 7.2.4.)
∂t ∂x2
⎩
u(0, x) = ϕ(x), x ∈ R,
∂2
If we consider the space Xρ , ρ > 0, defined in (7.11), then D2 = ∂x2 is an operator on
2
Xρ and Tt = etD . Show that Tt is chaotic on Xρ for any t > 0.
Exercise 7.5.5. For the abstract Cauchy problem associated with the linear kinetic
model (7.13), let α1 = 3, αn = 1, n ≥ 2, and βn = 2, n ∈ N. Prove that the solution
semigroup is not hypercyclic on
1 . This shows that condition (7.14) cannot be replaced
by lim supn→∞ αn < lim inf n→∞ βn .
Section 7.1. For the theory of C0 -semigroups we refer to the books by Engel and Nagel
[143], [144]. All the results mentioned in this section can be found there.
Section 7.2. Rolewicz [268] was the first to observe the existence of a dense orbit under
a C0 -semigroup. A systematic study of the dynamical properties of semigroups, however,
was only started by Desch, Schappacher and Webb [131]. In particular, they introduced
the notions of hypercyclicity and chaos for semigroups. Several other results in this
chapter are due to them, for example the characterization of hypercyclic translation
semigroups on Lpv (R+ ). The characterization of the corresponding mixing property was
added by Bermúdez, Bonilla, Conejero, and Peris [49].
Lemma 7.14 on the necessity of empty point spectrum for the adjoint operators and
for the adjoint of the generator of a hypercyclic semigroup was obtained by Costakis
and Peris [121] and by Desch, Schappacher and Webb [131], respectively. Theorem 7.16
is due to Costakis and Peris [121].
Chaos for the translation semigroup was characterized by deLaubenfels and Emami-
rad [128]. Kalmes [209], [210] has undertaken a systematic study of hypercyclicity, weak
mixing, mixing and chaos of C0 -semigroups that are induced by solution semiflows of
differential equations on open subsets of Rd .
Section 7.3. Theorem 7.22 is due to Oxtoby and Ulam [251]. As for the discretization re-
sults, Proposition 7.20 is due to Conejero and Peris [113], Proposition 7.21 to Bermúdez,
Bonilla, Conejero, and Peris [49] and Conejero and Peris [113], and Proposition 7.25 to
Desch and Schappacher [129]; see also Conejero and Peris [113]. Theorem 7.23 can be
found in Bernal and Grosse-Erdmann [63]. As we already mentioned in Chapter 6, the
discretization results fail for chaotic semigroups; in fact, Bayart and Bermúdez [37] have
constructed a chaotic semigroup on a Hilbert space for which no autonomous discretiza-
tion is chaotic.
Section 7.4. The first criteria for hypercyclicity of C0 -semigroups were found by Desch,
Schappacher and Webb [131]. In the form given here, the Hypercyclicity Criterion, The-
Sources and comments 211
orem 7.27, is due to Conejero and Peris [111] and El Mourchid [139], while the criterion
for mixing, Theorem 7.29, is due to Bermúdez, Bonilla, Conejero, and Peris [49]. See
also Conejero and Peris [113].
As for the eigenvalue criteria, Theorem 7.30 is due to Desch, Schappacher and
Webb [131], while Theorem 7.32 is due to El Mourchid [140]; see also Conejero and
Mangino [109]. We mention that the latter result was motivated by a similar criterion
for frequent hypercyclicity; see Section 9.3.
Theorem 7.28, together with further equivalent forms of the Hypercyclicity Criterion,
can be found in Conejero and Peris [113]. Incidentally, Herrero’s problem if hypercyclic-
ity implies weak mixing (see Section 2.5), also has a negative answer for semigroups, as
was proved by Shkarin [292].
Section 7.5. The three applications treated in this section are, in order, from El
Mourchid [140], Conejero, Peris and Trujillo [114], and Banasiak and Lachowicz [24].
The first results on the hypercyclic behaviour of the heat equation were given by
Herzog [198]. Generalizations of Herzog’s results were accomplished by Betancor and
Bonilla [73]. Some recent developments on the heat dynamics are due to Ji and Weber
[207], [208].
The method presented here for the hyperbolic heat transfer equation is based on the
conversion of the second-order Cauchy problem to first order so that the theory of C0 -
semigroups can be applied. Alternatively, it is possible to keep the second-order problem
and solve it directly, by applying the theory of hypercyclic cosine operator functions; this
theory was initiated by Bonilla and Miana [89], and thoroughly developed by Kalmes
[211].
The chaotic dynamics associated with infinite systems of differential equations mod-
elling birth-and-death processes (evolution of cell population, etc.) have been systemat-
ically studied by Banasiak et al. in [24], [25], [22], [23], [26], [27].
We also mention an interesting recent application of the methods in this chapter to
the Black–Scholes formula. Emamirad, Goldstein and Goldstein [141] have shown that
the Black–Scholes semigroup is chaotic for certain choices of the parameters.
Exercises. The weight condition (iii) in Exercise 7.2.3 is due to Matsui, Yamada and
Takeo [236]. For Exercise 7.2.5 we refer to Conejero and Peris [112]. The results of
Exercise 7.2.4 are due to Desch, Schappacher and Webb [131] (hypercyclicity), Bermúdez,
Bonilla, Conejero and Peris [49] (mixing) and Matsui, Yamada and Takeo [236] (chaos).
Exercise 7.3.1 can be found in Conejero and Peris [113]. Exercises 7.3.3 and 7.3.4 are
taken from Bernal and Grosse-Erdmann [63]. Topologically ergodic semigroups were
considered by Desch and Schappacher in [129] under the name of semigroups satisfying
the Recurrent Hypercyclicity Criterion. Exercises 7.3.6, 7.3.7 and 7.3.8 are extracted
from their paper. Exercise 7.4.3 is taken from Conejero and Peris [113], Exercise 7.4.4
from Bermúdez, Bonilla, Conejero and Peris [49], and Exercise 7.4.5 from Conejero and
Mangino [109]. Exercise 7.5.1 is extracted from El Mourchid [140], Exercise 7.5.2 is
taken from, and improves on, Takeo [297], while Exercises 7.5.3 and 7.5.4 are taken from
Conejero, Peris and Trujillo [114] and Herzog [198], respectively.
Chapter 8
Existence of hypercyclic operators
In this section we prepare the ground for the proofs of the existence of hy-
percyclic operators and C0 -semigroups on arbitrary infinite-dimensional sep-
arable Fréchet and Banach spaces, respectively. What we will need are hy-
percyclic operators on 1 that are “small” perturbations of the identity or,
equivalently, perturbations of the identity on weighted 1 -spaces with rapidly
growing weights. Such operators are also interesting in their own right.
Let v = (vn )n be a positive weight sequence, that is, a sequence of strictly
positive numbers, and let X be one of the weighted spaces
∞
p (v) = (xn )n≥1 ; |xn |p vn < ∞ , 1 ≤ p < ∞,
n=1
c0 (v) = (xn )n≥1 ; lim |xn |vn = 0 .
n→∞
∞
1 k
eB = B
k!
k=0
2m
ε
N m−k vk < . (8.2)
eC
k=m+1
We will show that, for any n ≥ N , there is some z ∈ 1 (v) such that
8.1 Mixing perturbations of the identity 215
where
⎛ ⎞ ⎛ nm nm+1 n2m−1
⎞
nm−1 ···
1 n
1!
··· (m−1)! m! (m+1)! (2m−1)!
⎜ ⎟ ⎜ nm−1 nm n2m−2 ⎟
⎜0 1 ··· nm−2 ⎟ ⎜ ··· ⎟
A=⎜ (m−2)! ⎟, V =⎜
(m−1)! m! (2m−2)! ⎟.
⎜. .. . . .. ⎟ ⎜ .. .. .. .. ⎟
⎝ .. . . . ⎠ ⎝ . . . . ⎠
2
0 0 ··· nm
1 n
1!
n
2!
··· m!
where
⎛ ⎞ ⎛ ⎞
n−1 0 · · · 0 n−m+1 0 ··· 0
⎜ 0 n−2 · · · 0 ⎟ ⎜ 0 n−m+2
··· 0⎟
⎜ ⎟ ⎜ ⎟
D2−1 =⎜ . .. . . .. ⎟ , D1−1 =⎜ .. .. .. .. ⎟ .
⎝ .. . . . ⎠ ⎝ . . . .⎠
0 0 · · · n−m 0 0 ··· 1
216 8 Existence of hypercyclic operators
This shows that the sequence z with the stated properties exists. Moreover,
since the entries of D1−1 and of D1−1 A are bounded by 1, we have that
m
m
D1−1 [(y1 , . . . , ym )t − A(x1 , . . . , xm )t ]1 ≤ |yk | + m |xk |.
k=1 k=1
2m
2m
x − z = |zk |vk ≤ C nm−k vk < ε.
k=m+1 k=m+1
2m
≤ eC nm−k vk < ε.
k=m+1
Proof. Again, it suffices to prove the result for 1 (v). We claim that there is a
positive weight sequence w = (wn )n and an operator φ : 1 (w) → 1 (v) with
dense range such that the following diagram commutes:
eB
1 (w) −−−−→ 1 (w)
⏐ ⏐
φ
⏐ ⏐φ
I+B
1 (v) −−−−→ 1 (v).
n
n−1
(I + B)φ(en ) = ank (I + B)ek = ann en + (ank + an,k+1 )ek .
k=1 k=1
n
1
j
a n n
jk
= ajk ek = ek .
j=1
(n − j)! (n − j)!
k=1 k=1 j=k
Setting these expressions equal and comparing the coefficients of en−1 yields
that an,n−1 + ann = an−1,n−1 + an,n−1 for n ≥ 2, so that the elements on the
diagonal must coincide; comparing the remaining coefficients implies that,
for k = 1, . . . , n − 2,
n
ajk
ank + an,k+1 = ,
(n − j)!
j=k
hence
n−1
ajk
an,k+1 = ,
(n − j)!
j=k
We note that, conversely, Theorem 8.2 can also be derived from Theorem
8.1 (see Exercise 8.1.5), so that the two results are in fact equivalent. Another
equivalent statement will be discussed in Exercise 8.1.2.
Using a suitable conjugacy one can reformulate the previous two results in
terms of weighted shifts on unweighted spaces. We recall that, given a weight
sequence w = (wn )n , the weighted shift Bw is a well-defined operator on p ,
1 ≤ p < ∞, or on c0 if and only if supn∈N |wn | < ∞; see Example 4.9(a).
Proof. Once more it suffices to perform the proof for X = 1 . The operator
Dw is best understood by its action on the unit sequences en , n ≥ 1. If we
write n ≥ 1 as
then
T en = 0 if k = 1, and T en = wm2k−1 em2k−2 if k ≥ 2.
In this way it becomes obvious that Dw is conjugate to a countable direct
sum of weighted shifts; see the explanations before Proposition 2.41. More
precisely, if we define weights w (m) = (wm2k−1 )k≥1 for m ≥ 1 odd, and the
map
8.2 Existence of mixing operators and semigroups 219
φ: 1 → 1 , (xm,k )k≥1 m odd
→ (yn )n ,
1
m odd
is dense. If T is surjective then the condition simply says that T has dense
generalized kernel. Moreover, any weighted backward shift on a Banach se-
quence space in which the finite sequences are dense is an extended backward
shift.
Theorem 8.6. Let A be an extended backward shift on a Banach space X.
Then the operators eA and I + A are mixing.
Since this result will not be needed in the sequel we omit the proof; but
see Exercise 8.1.9.
In this section we will prove the existence of mixing, and therefore hypercyclic,
operators on arbitrary infinite-dimensional separable Fréchet spaces. For the
particular case of Banach spaces we even show the existence of mixing C0 -
semigroups.
We first need two auxiliary results for Fréchet spaces. The first one is
beyond the scope of this book and is therefore stated without proof.
220 8 Existence of hypercyclic operators
Lemma 8.7. Every separable Fréchet space that is not isomorphic to ω has
a dense subspace that admits a continuous norm.
Here, a continuous norm is exactly what it says: a norm functional that is
continuous on the underlying space.
The second lemma is the crucial tool for transferring the mixing operators
of the previous section to more general Fréchet spaces.
Lemma 8.8. Let X be an infinite-dimensional separable Fréchet space that
is not isomorphic to ω. Then there are sequences (xn )n in X and (x∗n )n in
X ∗ such that
(i) (xn )n converges to 0, and span{xn ; n ∈ N} is dense in X,
(ii) (x∗n )n is equicontinuous,
(iii)
xk , x∗n = 0 if k
= n, and 0 <
xn , x∗n ≤ 1, n ≥ 1.
Proof. By Lemma 8.7 there exists a dense subspace M of X that admits a
continuous norm | · |. We select a linearly independent sequence (zn )n in M
whose linear span is dense in M , and therefore in X. By the Hahn–Banach
theorem (see Appendix A) there are functionals yn∗ ∈ (M, | · |)∗ such that,
for all n ≥ 1,
zn , yn∗ = 1 and
zk , yn∗ = 0 for k < n. Following a Gram–
n−1
Schmidt procedure in setting y1 = z1 and yn = zn − k=1
zn , yk∗ yk for
n > 1, we obtain a sequence (yn )n in M such that span{yn ; n ∈ N} =
span{zn ; n ∈ N} is dense in X and
yk , yn∗ = δk,n , k, n ∈ N.
By continuity there are Kn ≥ 1, n ≥ 1, such that |
x, yn∗ | ≤ Kn |x| for
all x ∈ M . Since M is dense in X, there is a (unique) continuous seminorm p
on X whose restriction to M coincides with | · |, and each yn∗ has a unique
continuous linear extension to X, which we denote again by yn∗ . Clearly,
|
x, yn∗ | ≤ Kn p(x) for x ∈ X, n ≥ 1. This implies that {Kn−1 yn∗ ; n ∈ N}
is equicontinuous. Since X is metrizable, there are numbers αn ∈ ]0, 1] such
that (αn yn )n converges to 0 in X. Setting xn = αn yn and x∗n = Kn−1 yn∗ ,
n ≥ 1, the claim follows.
We are now ready for the main result of this section.
Theorem 8.9 (Ansari–Bernal). Every infinite-dimensional separable Fré-
chet space supports a mixing, and therefore hypercyclic, operator.
Proof. On X = ω, the backward shift is a mixing operator; see Example
4.9(c). We may therefore suppose that X is an infinite-dimensional separable
Fréchet space that is not isomorphic to ω. Applying Lemma 8.8 we find
sequences (xn )n in X and (x∗n )n in X ∗ with the specified properties. Let us
consider T : X → X,
∞
Tx = x + 2−n
x, x∗n+1 xn , x ∈ X.
n=1
The equicontinuity of (x∗n )n and the fact that (xn )n tends to 0 easily imply
that T is a well-defined operator on X.
8.2 Existence of mixing operators and semigroups 221
Remark 8.10. We note that in the case that X is a Banach space, the mixing
operator T constructed above is of the form
T = I + K,
where K is a compact operator (see Exercise 5.2.4), and one may make K
arbitrarily small; see also Exercise 8.2.2.
for semigroups (see Appendix B) that etσp (A) contains at most one element,
for every t > 0, where A is the infinitesimal generator of (Tt )t≥0 . This is
only possible if σp (A) itself contains at most one element, which contradicts
Proposition 7.18. In the case when K = R we consider the complexification
of the semigroup, which, by Corollary 7.24, is again a chaotic semigroup
consisting of compact perturbations of multiples of the identity, and we argue
as before.
For vectors we have the phenomenon that as soon as one hypercyclic vec-
tor (for a given operator) exists there is then automatically a large supply
of them. Could something similar be true for operators? That is, does the
existence of one hypercyclic operator on a given space imply that there must
then be many of them, in a certain sense?
The most natural interpretation of this question leads to a negative answer.
Indeed, no operator T on a Banach space X with T ≤ 1 can be hypercyclic,
because all of its orbits are bounded. Thus the set of hypercyclic operators
on X is not dense in the space L(X) of all operators, when we endow it with
the operator norm topology. One can even show that in this topology the set
of hypercyclic operators is nowhere dense.
But there is another well-known, weaker topology on L(X), the strong op-
erator topology, abbreviated SOT; it is the topology of pointwise convergence
of operators. We will show in this section that the set of hypercyclic operators
is indeed dense in this topology.
In order to prepare the ground, let X be any Fréchet space and L(X) the
space of operators on X. The strong operator topology is then defined in the
following way: for T ∈ L(X), a base of neighbourhoods is given by
is SOT-dense in L(X).
224 8 Existence of hypercyclic operators
x1 , . . . , xn , T x1 , . . . , T xn
We are now in a position to prove the announced density result for hyper-
cyclic operators.
Proof. We know from the Ansari–Bernal theorem that there exists a hyper-
cyclic operator T on X. By conjugacy, the similarity orbit S(T ) consists en-
tirely of hypercyclic operators. Thus the result follows from Proposition 8.16
once we know that, for any n ≥ 1, there are n vectors x1 , . . . , xn in X such
that x1 , . . . , xn , T x1 , . . . , T xn are linearly independent. But these are easily
found: the orbit of any hypercyclic vector x for T is linearly independent and
hence so is
x, T 2 x, T 4 x, . . . , T 2n−2 x, T x, T 3 x, T 5 x, . . . , T 2n−1 x;
The proof tells us much more: the same result will be true for any
nonempty class of hypercyclic operators that is invariant under conjugacies.
In particular, we obtain the following.
There is another sense in which one can measure the richness of the set of
hypercyclic operators. By Proposition 2.60 we know that a dense orbit is nec-
essarily a linearly independent set. Therefore one may wonder if, conversely,
every countable dense linearly independent set arises as an orbit of an oper-
ator; this operator is then automatically hypercyclic. We will show here that
the answer is positive, at least in the setting of Banach spaces. The idea of
the proof is not dissimilar to the approach in the last section: we construct
an invertible operator that transforms a known dense orbit into an arbitrary
countable dense linearly independent set.
We first need a property of perturbations of invertible operators.
A−1 2 A − B
A−1 − B −1 ≤ .
1 − A−1 A − B
226 8 Existence of hypercyclic operators
Proof.
∞ First, if T is an operator on X with T < 1 then the series C :=
n
n=0 T converges in L(X) under the operator norm, so that C defines an
operator on X. Moreover, since
(I − T )C = (I + T + T 2 + . . .) − (T + T 2 + . . .) = I,
Applying the operator norm and the sum formula of the geometric series
gives the desired estimate on A−1 − B −1 .
The next step is to prescribe values of an invertible operator.
Lemma 8.21. Let X be an infinite-dimensional Banach space and X0 and Y0
dense subsets of X. Let E and F be subspaces of X with dim(E) = dim(F ) <
∞ and A an invertible operator on X such that A(E) = F .
Then, for any x0 ∈/ E, y0 ∈
/ F and ε > 0, there exists an invertible operator
B on X such that B|E = A|E , Bx0 ∈ Y0 , B −1 y0 ∈ X0 and A − B < ε.
Proof. Since E is closed as a finite-dimensional subspace, the Hahn–Banach
theorem implies that there exists x∗ ∈ X ∗ such that
x, x∗ = 0 for all x ∈ E
and
x0 , x∗ = 1. By density of Y0 there is some y1 ∈ Y0 with x∗ y1 −
Ax0 < min(ε/2, 1/A−1 ); perturbing y1 if necessary we can assume that it
does not belong to F ⊕ span{y0 }. Then we define the operator C by
Cx = Ax + x, x∗ (y1 − Ax0 ), x ∈ X.
A more careful analysis of the proof shows that, for every ε > 0, the
operator T can be of the form T = I + K, where K is a compact operator
with K < ε; see Exercise 8.4.3.
Exercises
Exercise 8.1.3. Use the previous exercise to prove Theorem 4.23. (Hint: Note that
rn ≤ Mn φ(r); then find γn such that any f ∈ E 1 (γ) satisfies the required growth
condition, and apply the hypercyclic comparison principle.)
v
Exercise 8.1.4. Let v = (vn )n be a positive weight sequence with supn∈N vn+1 n
< ∞.
Show directly that I + B is mixing on p (v), 1 ≤ p < ∞, or on c0 (v). (Hint: Following
the proof of Theorem 8.1, use the same argument substituting the matrix W by a matrix
Wn such that limn Wn = W .)
Exercises 229
⏐ ⏐
⏐ ⏐
φ φ
T =ϕ(A)
X −−−−−−→ X
and apply Proposition 1.40.)
Exercise 8.1.9. Prove Theorem 8.6. (Hint: By Proposition 2.37 it suffices to show that
for any x ∈ X, ε > 0, and for any sufficiently large n there are z, z ∈ X (depending on
n) such that x − z !< ε and T n z < ε, z < ε and x − T n z < ε. It suffices to
∞
consider x = 0 from n=0 (ker T n ∩ ran T n ). Choose m ≥ 1 minimal and y ∈ X such
that T m x = 0 and x = T m y; then construct ek , k = 1, . . . , 2m, such that T e1 = 0,
T ek+1 = ek for k ≥ 1, and em = x. By minimality of m, the ek are linearly independent.
Now, for T = eA , use the argument in the proof of Theorem 8.1 to show that there
is some C > 0 such that, for any n ≥ 1, there is z ∈ X (depending on n) of the
2m 2m 2m nj−k
form z = em + z e that satisfies T n z =
k=m+1 k k k=m+1
( j=k (j−k)! zj )ek and
|zk | ≤ Cnm−k , k = m + 1, . . . , 2m. Deduce that, for large n, x − z < ε and T n z < ε.
Construct z in a similar way. For the proof of T = I +A replace W by Wn as in Exercise
8.1.4.)
Exercise 8.2.1. Show that Lemma 8.8 fails for X = ω.
Exercise 8.2.2. Let T be an operator on a Banach space X. For n ≥ 1 the nth approx-
imation number αn (T ) is defined as
Exercise 8.2.6. Give an alternative proof of Theorem 8.14 as follows. Let T be the
mixing operator on an infinite-dimensional separable Banach space X provided by the
proof of Theorem 8.9 of the form T = I + S, where we can even have that S < 1. Then
find an operator A such that eA = T to deduce that (etA )t≥0 is a mixing C0 -semigroup
on X. (Hint: Use Exercise 7.1.5.)
Exercise 8.3.1. Even though every hypercyclic operator on 2 has norm bigger than
one, Theorem 8.18 tells us that one can approximate the zero operator in the SOT-
topology by such operators. Show that, concretely, if B is the backward shift, then
n B , n ≥ 1, are hypercyclic operators that converge to 0 in SOT as n → ∞.
n+1 n
Exercise 8.3.2. Show that the class of all operators on an infinite-dimensional separable
Fréchet space that have no nontrivial closed invariant subset is either empty or SOT-
dense.
Exercise 8.3.3. Show that on any infinite-dimensional separable Fréchet space that is
not isomorphic to ω the set of hypercyclic non-chaotic operators is SOT-dense. (Hint:
Exercise 8.2.3.)
Exercise 8.3.6. Let X be a complex Banach space with the properties of Theorem
8.11. Show that if an operator T on X is the sum of two hypercyclic operators then it
is of the form T = λI + K, where |λ| ≤ 2 and K is a compact operator. In particular,
not every operator on X can be written as the sum of two hypercyclic operators. (Hint:
Lemma 5.19.)
Exercise 8.4.2. Show that every normed space of countably infinite dimension sup-
ports an operator without nontrivial closed invariant subsets. (Hint: Apply the previous
exercise to the completion of X.)
Exercise 8.4.3. Show that, for any ε > 0, the operator T in Theorem 8.24 can be of
the form T = I + K, where K is a compact operator with K < ε. (Hint: Remark 8.10
and Exercise 5.2.3.)
Exercise 8.4.4. Show directly that the operator C defined in the proof of Lemma 8.21
is invertible if and only if it is injective, and if and only if A−1 y1 − x0 , x∗ = −1. In
that case, calculate the inverse.
Exercise 8.4.5. Show that Theorem 8.24 is false in ω. Indeed, there exists a dense
linearly independent sequence in ω that cannot be the orbit of any operator on ω. To
see this, let X1 = {xn ; n ∈ N} be a dense linearly independent set in ω such that, for
n ≥ 1, xn = (xn,k )k is a finite sequence with xn,1 = 0. We consider X2 = {yn ; n ∈ N}
defined by yn,k = 0 if k ≤ n, and yn,k = nk if k > n. Then X0 = X1 ∪ X2 is dense
and linearly independent in ω, but there is no operator T on ω such that X0 coincides
with an orbit under T . (Hint: Otherwise there would be infinitely many elements in X2
whose image under T belonged to X1 ; derive that then the composition of T with the
projection onto the first coordinate would not be continuous.)
Exercise 8.4.6. Prove the following extension of Theorem 8.24 to C0 -semigroups. For
any dense linearly independent set {xn ; n ∈ N} in a Banach space X, there is a C0 -
semigroup (Tt )t≥0 on X such that orb(x1 , T1 ) = {xn ; n ∈ N}. (Hint: Combine the
proof of Theorem 8.24 with Exercises 8.2.6 and 8.4.3.)
Section 8.1. The main result of this section comes in three equivalent forms. Salas [274]
showed that any perturbation of the identity by a weighted backward shift is hypercyclic
on p or c0 ; see Corollary 8.3. Following his arguments, Desch, Schappacher and Webb
[131] showed that the exponential of the backward shift is hypercyclic on the correspond-
ing weighted spaces; see Theorem 8.1. Some years earlier, Chan and Shapiro [106] had
obtained the hypercyclicity of the translation operator Ta on Hilbert spaces of entire
functions of slow growth; see Exercise 8.1.2. The fact that the equivalence of these results
was only noticed much later is, perhaps, due to the fact that the authors worked in very
different contexts. While Salas regarded iterates of a function of the backward shift,
Desch, Schappacher and Webb were interested in C0 -semigroups generated by shifts,
and Chan and Shapiro studied translation operators on spaces of entire functions. In
232 8 Existence of hypercyclic operators
addition, the proof by Chan and Shapiro was very different from the one by Salas by
using, in a crucial way, tools from complex analysis. Incidentally, the quasiconjugacies
between I + B and eB were first observed in Martínez and Peris [232].
León and Montes [220] showed that Salas’ operators satisfy the Hypercyclicity Cri-
terion, while Grivaux [172] showed that they are even mixing.
Proposition 8.5 is due to Grivaux [171]. Theorem 8.6 was obtained by Grivaux and
Shkarin [176, 286]; see also Bayart and Matheron [44]. Hypercyclicity of functions ϕ(T )
of rather general operators T have been studied by Herzog and Schmoeger [200], Miller
and Miller [239], Bermúdez and Miller [53], Martínez and Peris [232], and Müller [248].
Section 8.3. Chan [100] showed that the set of hypercyclic operators on an infinite-
dimensional separable Hilbert space is SOT-dense in the space of all operators, which was
extended by Bès and Chan [68] to Fréchet spaces. The simplified proofs using Proposition
8.16 by Hadwin, Nordgren, Radjavi, and Rosenthal [189] are due to Bès and Chan [67]
and Prǎjiturǎ [260]. In the same vein many similar results are possible; see Bès and Chan
[67].
The fact that hypercyclic operators are nowhere dense under the operator norm
topology was observed by Wu [302], who even showed that the set of cyclic operators
has this property. Herrero [194] has obtained a spectral description of the operator norm
closure of the set of hypercyclic operators on an infinite-dimensional separable complex
Hilbert space; see also Müller [248]. Herrero then deduced in [195] that the chaotic
operators are norm dense in the set of hypercyclic operators.
As the set of hypercyclic operators is not dense in the operator norm topology one
might wonder if its linear hull is dense. First, Bès and Chan [67] proved even more: on an
infinite-dimensional separable complex Hilbert space the set of sums of two hypercyclic
Sources and comments 233
(or even chaotic) operators is dense in the operator norm. This motivated the following
deep result.
In fact, both operators can even be chosen to be chaotic. But Grivaux also showed
that, even for hypercyclicity, the result cannot be extended to all Banach spaces; see
Exercise 8.3.6.
Section 8.4. The problem of whether every linearly independent sequence in a separable
Banach space is contained in the orbit of a hypercyclic vector was posed by Halperin,
Kitai and Rosenthal [192], who had obtained the result for Hilbert spaces. Theorem
8.24, which solves the problem, is due to Grivaux [169]. Albanese [5] has extended the
theorem to Fréchet spaces with a continuous norm.
Exercises. Exercise 8.1.2 generalizes Exercise 4.2.4 for Birkhoff’s operators. It was
proved (with the exception of (iv)) by Chan and Shapiro [106] with completely different
techniques based on complex analysis. Exercise 8.1.3 was observed by Chan and Shapiro
[106]. Exercise 8.1.4 essentially asks for the original proof of Corollary 8.3 by Salas [274].
The notion of a generalized backward shift was introduced by Godefroy and Shapiro
[165], where one also finds Exercise 8.1.6. Exercises 8.1.7 and 8.1.8 are from Martínez
and Peris [229], and [230, 232], respectively. Exercise 8.2.1 is taken from Bonet and Peris
[85], Exercise 8.2.2 generalizes a result by Chan and Shapiro [106]. The result of Exercise
8.2.4 is due to Bès, Martin, Peris, Salas and Shkarin [70, 277, 291]. Exercise 8.2.6 is taken
from Bernal and Grosse-Erdmann [63], Exercise 8.3.1 from Chan [100], and Exercises
8.3.4 and 8.3.5 from Chan and Sanders [103]. Concerning Exercise 8.3.3, De la Rosa,
Frerick, Grivaux and Peris [127] have shown that also on ω there are hypercyclic non-
chaotic operators, so that the result of the exercise extends to all infinite-dimensional
separable Fréchet spaces. Exercise 8.3.6 is due to Grivaux [170]; the use of the Argyros–
Haydon spaces simplifies the proof. Exercises 8.4.1, 8.4.2, 8.4.3 are taken from Grivaux
[169], and Exercises 8.4.4, 8.4.5 and 8.4.6 from Albanese [5], Bonet, Frerick, Peris and
Wengenroth [81] and Bernal and Grosse-Erdmann [63], respectively.
Chapter 9
Frequently hypercyclic operators
The theory of linear dynamical systems has its roots in topological dynamics.
But there is also a parallel theory of measurable dynamics, which is better
known under the name of ergodic theory. In this chapter we show how con-
cepts and results from that theory lead to a deepened understanding of linear
dynamics. More specifically, we will see how the celebrated Birkhoff ergodic
theorem suggests an interesting and rather strong variant of hypercyclicity,
that of frequently hypercyclic operators. We point out that, while ergodic
theory has turned out to be a most powerful tool in linear dynamics, we will
use it here only for motivating the new concept.
Having introduced frequently hypercyclic operators, we then derive a Fre-
quent Hypercyclicity Criterion and an eigenvalue criterion that allow us to
show that, quite surprisingly, many of the hypercyclic operators met so far
are in fact frequently hypercyclic. In the final section we revisit several of the
structural properties of hypercyclicity within the new framework.
and μ(B) > 0, there is some n ∈ N0 such that μ(T −n (A) ∩ B) > 0. This
notion is not only formally similar to topological transitivity. Suppose that
the measure μ has the additional property that μ(U ) > 0 for any nonempty
open set U ; μ is then said to be of full (topological) support. Under this
assumption, ergodicity obviously implies topological transitivity.
But there is an added bonus in the form of the Birkhoff ergodic theorem.
It tells us that if T is ergodic with respect to μ then, for any μ-integrable
function f on X, its time average with respect to T coincides with its space
average; more precisely we have that
1
N
f (T n x) → f dμ, for μ-almost all x ∈ X, (9.1)
N + 1 n=0 X
1
N
card{0 ≤ n ≤ N ; T n x ∈ Uk }
1Uk (T n x) = ,
N + 1 n=0 N +1
while the right-hand side is simply X 1Uk dμ = μ(Uk ) > 0, where we have
assumed again that μ is of full support. Thus there are subsets Ak ⊂ X,
k ≥ 1, of full measure such that, for any x ∈ Ak ,
card{0 ≤ n ≤ N ; T n x ∈ Uk }
lim > 0.
N →∞ N +1
Since every nonempty open set contains some Uk and since k≥1 Ak has full
measure we obtain that, for μ-almost all x ∈ X and every nonempty open
subset U of X,
card{0 ≤ n ≤ N ; T n x ∈ U }
lim inf > 0.
N →∞ N +1
What we have found here is that, under the mentioned assumptions, the
operator T has a property that is much stronger than hypercyclicity. There
must even be an x ∈ X whose orbit meets every nonempty open set very
often, in the sense given above. Let us recall here the following.
Definition 9.1. The lower density of a subset A ⊂ N0 is defined as
card{0 ≤ n ≤ N ; n ∈ A}
dens(A) = lim inf .
N →∞ N +1
k card{0 ≤ n ≤ N ; n ∈ A} k
≤ ≤ ,
nk+1 N +1 nk
which implies that dens (A) = lim inf k→∞ nkk . Thus A has positive lower
density if and only if ( nkk )k is bounded; in other words, if nk = O(k).
Proposition 9.3. A vector x ∈ X is frequently hypercyclic for T if and only
if, for any nonempty open subset U of X, there is a strictly increasing se-
quence (nk )k of positive integers such that
Moreover, if yl
= yk then the sets A(l, ν) and A(k, μ) are disjoint if ν and μ
are big. In fact, in the sequel we will need the existence of sets A(l, ν) with
a stronger separation property.
|n − m| ≥ ν + μ if n
= m.
n = (0, . . . , 0, 1, . . . , 1, 0, ∗)
k−1
nk = 2 δ i + δk , k ≥ 1,
i=1
has the desired properties. First, these sets are pairwise disjoint. Moreover,
if nk ∈ A(l, ν) then nk ≥ δk = ν; and if nj ∈ A(l, ν), nm ∈ A(k, μ) with
nj
= nm , where we can assume that j > m, then
j−1
nj − nm = δm + 2 δi + δj ≥ μ + ν.
i=m+1
It remains to show that each set A(l, ν) has positive lower density. We
begin by proving that there is some M > 0 such that
nk ≤ M k, k ≥ 1. (9.2)
9.1 Frequently recurrent orbits 239
ν
N
2
n2 N ≤ 2 δi ≤ 2 2N +2−l−ν ν ≤ 8 2N ,
2l+ν
i=1 l+ν≤N +2 l,ν≥1
kj ≤ Kj, j ≥ 1.
nkj ≤ M kj ≤ M Kj, j ≥ 1.
sup |fk+1 (z) − fk (z)| < εk and sup |fk+1 (z) − gk+1 (z)| < εk .
|z|≤nk +rk z∈Bk+1
If ∞k=1 εk < ∞ then it follows from the first inequality and the fact that
nk → ∞ that
∞
f (z) := f1 (z) + (fk+1 (z) − fk (z)) = lim fk (z)
k→∞
k=1
that is,
∞
sup |f (z) − Pl (z − nk )| ≤ εj
|z−nk |≤ν/2 j=k−1
k ∈ A(l, ν). It is easy to see that we can choose the εj in such a way
for n
that ∞ j=k−1 εj < ν whenever nk ∈ A(l, ν). We therefore have that
1
1
sup |T1nk f (z) − Pl (z)| <
|z|≤ν/2 ν
for nk ∈ A(l, ν). Since the sets {g ∈ H(C) ; sup|z|≤ν/2 |g(z) − Pl (z)| < ν1 },
l, ν ≥ 1, form a basis of the topology of H(C) and since each set A(l, ν) has
positive lower density, it follows that T1 is frequently hypercyclic.
A − A = {n − m ; n, m ∈ A, n ≥ m};
9.1 Frequently recurrent orbits 241
T n x ∈ U0 for any n ∈ A.
242 9 Frequently hypercyclic operators
T n0 +m−n (T n x) = T n0 (T m x) ∈ W.
n0 + (A − A) ⊂ N (U0 , W ) ⊂ N (U, W ).
T K−k (T k y) ∈ T K−k (W ) ∩ V.
1
T n yj
< l , (9.3)
l2
n∈F
1
S n yj
< l . (9.4)
l2
n∈F
and
z n = yl if n ∈ A(l, Nl ).
We then consider
x= S n zn . (9.5)
n∈A
∞
l ∞
S n zn = S n yj = S n yj + S n yj .
n∈A j=1 n∈A(j,Nj ) j=1 n∈A(j,Nj ) j=l+1 n∈A(j,Nj )
n∈F n∈F n∈F n∈F
1
S n yj
< l ;
l2
n∈A(j,Nj )
n∈F
1 1
S n yj
< j ≤ j .
j2 2
n∈A(j,Nj )
n∈F
l ∞
1 1 2
S n
z n
< l
+ j
= l.
j=1
l2 2 2
n∈A j=l+1
n∈F
244 9 Frequently hypercyclic operators
Since l was arbitrary we have proved that the series (9.5) converges uncon-
ditionally.
We now show that x is frequently hypercyclic for T . To this end, fix l ≥ 1.
Then, for n ∈ A(l, Nl ),
T n x − yl = T n S k zk + T n S k zk + T n S n zn − yl .
k∈A k∈A
k<n k>n
For the second sum we have, for any m ≥ n, using condition (iii),
l ∞
T n S k zk = S k−n yj + S k−n yj .
k∈A j=1 k∈A(j,Nj ) j=l+1 k∈A(j,Nj )
n<k≤m n<k≤m n<k≤m
hence
2
T n S k zk
≤ l .
2
k∈A
k>n
In the same way, but using (9.3) instead of (9.4), we obtain that also
2
T n S k zk
≤ l .
2
k∈A
k<n
T n S n zn = yl .
Remark 9.10. For a later application we note that the same proof works when
we replace conditions (ii) and (iii) by the following:
9.2 The Frequent Hypercyclicity Criterion 245
Let us also note here that the Frequent Hypercyclicity Criterion not only
implies frequent hypercyclicity but also two other strong forms of hypercyclic-
ity.
In the last section we saw that Birkhoff’s operators are frequently hyper-
cyclic. The Frequent Hypercyclicity Criterion allows us to show that also
the other two classical hypercyclic operators, the operators of MacLane and
Rolewicz, are in fact frequently hypercyclic; see also Exercise 9.2.2.
Example 9.16. (a) We follow Example 4.9(b) and consider weighted shifts
T = Bw on H(C), or rather its corresponding sequence space. Since the
sequences en , n ≥ 0, correspond
∞
to the monomials z n , Bw turns out to be
frequently hypercyclic if n=1 ( ν=1 wν )−1 z n converges unconditionally in
n
n −1
wν en converges.
n∈A ν=1
and since the projection onto the first coordinate is continuous on X, there
is a set B ⊂ N0 of positive lower density such that, for any n ∈ B,
hence
1
|xn+1 | > .
|w2 w3 · · · wn+1 |
∞
Together with the unconditional convergence of n=1 xn en this implies that
1
en+1
w1 w2 · · · wn+1
n∈B
converges; see Theorem A.16. This proves the claim for A = {n + 1 ; n ∈ B}.
We note that this condition is not, in general, a sufficient condition; see
Exercise 9.2.5.
While at the outset it was not even clear if frequently hypercyclic operators
exist, we have now actually seen that all the classical hypercyclic operators
and many others have this strong form of hypercyclicity. Although it is not to
be expected that hypercyclicity and frequent hypercyclicity coincide, we are
also in the position to give an example that differentiates the two concepts.
Example 9.18. On X = 2 we consider the weighted shift Bw with weights
wn = ( n+1
n )
1/2
. It follows from Example 4.9(a) that Bw is hypercyclic and
even mixing. However, if Bw were frequently hypercyclic then by the previous
proposition we could find a set A = {nk ; k ≥ 1} of positive lower density
∞
such that k=1 nk1+1 < ∞, which is impossible since nk = O(k); see the
discussion before Proposition 9.3. Note that Bw is conjugate to the shift
operator B on the Bergman space A2 ; see Example 4.4(b).
This example takes us back to the problem of comparing frequent hyper-
cyclicity with other forms of hypercyclicity. We saw in Theorem 9.8 that
every frequently hypercyclic operator is weakly mixing. Some other implica-
tions have turned out to be false. We have just seen that the mixing property
248 9 Frequently hypercyclic operators
does not imply frequent hypercyclicity. But there is also a frequently hyper-
cyclic operator on c0 that is neither mixing nor chaotic. In particular, by
Proposition 9.11, this operator does not satisfy the Frequent Hypercyclicity
Criterion. The construction of this example goes well beyond the scope of
this book.
The reader may have noticed that in frequent hypercyclicity so far we have
not made use of the Baire category theorem. This is unlike the situation in
hypercyclicity where the existence of a hypercyclic vector was deduced from
the fact that, in the sense of Baire category, there must be many of them;
see the proof of the Birkhoff transitivity theorem. In fact, this procedure is
ruled out in frequent hypercyclicity because, in general, the set F HC(T ) of
frequently hypercyclic vectors for an operator T is only of first Baire category.
We have that
E= Ek,M ,
k≥1 M ≥1
where
Ek,M = x ∈ X ; card{n ≤ N ; T n x ≥ δ} ≥ N +1
k .
N ≥M
is open, and it contains the dense set X0 . Hence each set Ek,M is nowhere
dense, so that E is of first Baire category.
Since there are uncountably many λ, one can find a finite subset, λ1 < λ2 <
. . . < λK say, such that
K
δλk > 2.
k=1
λ
Let ρ = min1≤k<K λk+1
k
,and choose M ∈ N such that ρM ≥ 3. We then have
for N sufficiently large that, for any k = 1, . . . , K,
The proof is similar to that of Theorem 7.32. We will need the Riemann
integral 2π
f (t) dt
0
for a continuous function f : [0, 2π] → X; see Appendix A for details and
basic properties.
Proof of Theorem 9.22. (a) Let (Ej )j∈J be the given eigenvector field of T .
Since each Ej : T → X is continuous the integrals
2π
xk,j := eikt Ej (eit ) dt ∈ X, k ∈ Z, j ∈ J,
0
X0 = Y0 = span{xk,j ; k ∈ Z, j ∈ J}.
We will use the Hahn–Banach theorem to show that this set is dense. Thus,
let x∗ be a continuous linear functional on X so that, for all k ∈ Z, j ∈ J,
2π
xk,j , x∗ = eikt Ej (eit ), x∗ dt = 0.
0
m
un = al xkl −n,jl , n ≥ 0.
l=1
is dense in X, and each vector in this span is a periodic point for T . Conse-
quently, T is chaotic.
(b) The proof follows the same lines, this time using Lemma 9.23(b) and
the Frequent Hypercyclicity Criterion in the form of Remark 9.10.
The eigenvalue criterion provides a new proof that the three classical hy-
percyclic operators are even frequently hypercyclic.
T D = DT.
Proof. Following the proof of Theorem 4.21 we can write T = ϕ(D) with a
nonconstant entire function ϕ of exponential type, which also implies that ev-
ery function eλ (z) = eλz , λ ∈ C, is an eigenvector of T to the eigenvalue ϕ(λ).
Since ϕ(C) is connected and dense (see Appendix A), there is a point z ∈ C
with w := ϕ(z) ∈ T; and since ϕ(C) is open and the zeros of ϕ are isolated
points we can also achieve that ϕ (z)
= 0. Thus ϕ maps a neighbourhood of
z conformally onto a neighbourhood U of w; let ψ be the inverse map, which
is holomorphic. Fix a nontrivial closed subarc γ ⊂ U of T containing w and a
C 2 -function f : T → C with f (w)
= 0 that vanishes outside γ. It follows that
E : T → H(C) with E(λ) = f (λ)eψ(λ) if λ ∈ γ and E(λ) = 0, else, defines an
9.3 An eigenvalue criterion for frequent hypercyclicity 253
As in the case of hypercyclicity one may ask how slowly a frequently hy-
percyclic entire function can grow at infinity. The eigenvalue criterion al-
lows us to deduce corresponding results for any operator T = ϕ(D); see
Exercise 9.3.3. Here we consider only the special case of Birkhoff’s operators
Ta f (z) = f (z + a), a
= 0. The theorem of Duyos-Ruiz tells us that corre-
sponding hypercyclic functions can grow arbitrarily slowly. This is no longer
true in the frequent context.
Proof. (a) This result follows from a general growth result for all operators
that commute with D (see Exercise 9.3.3) because Ta = eaD and eaz = 1 for
z = 0.
(b) We will assume that a = 1; see Example 4.26. Suppose, on the contrary,
that f is a frequently hypercyclic entire function with the stated growth
condition; by adding a constant, if necessary, we can assume that f (0) = 1.
Then there is a strictly increasing sequence (nk )k of positive integers with
nk = O(k) such that, for any k ≥ 1,
1 1
|f (z + nk ) − z| < for |z| ≤ .
2 2
Thus, by Rouché’s theorem (see Appendix A), f has a zero in |z − nk | < 12 . If
N (r) denotes the number of zeros of f in |z| < r, counting multiplicity, then
N (nk + 1) ≥ k, k ≥ 1.
On the other hand, it follows from Jensen’s formula (see Theorem A.23) and
the growth assumption on f that
254 9 Frequently hypercyclic operators
We saw in Section 6.3 how Ansari’s theorem follows from the fact that if
the union of the orbits of finitely many vectors is dense then one of these
orbits must already be dense. The corresponding result fails for frequent
hypercyclicity.
The proof of the Frequent Hypercyclicity Criterion then shows that the series
v := S n zn , w := S n zn
n∈B n∈C
M2k+1 1
≤
m2k+2 k
λ{t ∈ [0, T ] ; t ∈ A}
dens(A) := lim inf ,
T →∞ T
where λ denotes the Lebesgue measure.
Definition 9.29. A C0 -semigroup (Tt )t≥0 on a Banach space X is called
frequently hypercyclic if there is a vector x ∈ X such that, for any nonempty
open subset U of X,
dens {t ∈ R+ ; Tt x ∈ U } > 0.
1 N
dens(A) := lim inf λ{t ∈ [0, N ] ; (n, t) ∈ A},
N →∞ N (N + 1) n=0
Our main aim is to prove the converse statement. The following will be
crucial.
By Theorem 6.10, x is hypercyclic for Ψ (1, 1). It follows from property (β)
that, for any t ≥ 0, also Ψ (0, t)x is hypercyclic for Ψ (1, 1). Therefore there
are nj ∈ N0 , j = 1, . . . , k, such that
k )x ∈ U.
j−1
Ψ (nj , nj + k )x = Ψ (1, 1)nj Ψ (0, j−1
Ψ (nj , nj + j−1
k )(V ) ⊂ U, j = 1, . . . , k.
258 9 Frequently hypercyclic operators
1 N
λ{t ∈ [0, N ] ; Ψ (n, t)x ∈ V } ≥ δ.
N (N + 1) n=0
N
1
N
λ{t ∈ [0, N ] ; Ψ (n, t)x ∈ V, t ∈ Ij } ≥ λ{t ∈ [0, N ] ; Ψ (n, t)x ∈ V }.
n=0
k n=0
We conclude that
2N
λ{τ ∈ [0, 2N ] ; Ψ (ν, τ )x ∈ U, τ ∈ I1 }
ν=0
1
N
≥ λ{t ∈ [0, N ] ; Ψ (n, t)x ∈ V },
k n=0
so that
1
2N
δ
λ{τ ∈ [0, 2N ] ; Ψ (ν, τ )x ∈ U, τ ∈ I1 } ≥ .
2N (2N + 1) ν=0 4k
We can now prove the analogue of Theorem 6.10 for frequent hypercyclic-
ity.
N
∞
r := λ{t ∈ [0, N ] ; Ψ (n, t)x ∈ V, t ∈ m=1 [m − k1 , m[ ≥ N (N + 1)δ.
n=0
Now, if Ψ (n, t)x ∈ V and t ∈ [m− k1 , m[ then Ψ (n, m)x = Ψ (0, m−t)Ψ (n, t)x ∈
U . Thus, for
we have that p k1 ≥ r.
Next, let Ψ (1, 1)n x ∈ U . We distinguish the two cases described by (α). If
Ψ (1, 0) = I then Ψ (m, n)x = Ψ (n, n)x = Ψ (1, 1)n x ∈ U for any m ∈ Z; and
if Ψ (0, 1) = I then Ψ (n, m)x = Ψ (n, n)x ∈ U for any m ∈ Z. Thus, for
Proposition 9.31 and Theorem 9.33, applied to the semigroup action (9.7),
immediately imply a version of the León–Müller theorem for frequent hyper-
cyclicity.
Apart from being interesting in its own right, Theorem 9.35 has an impor-
tant application. We saw in Chapter 5 that the spectrum of a hypercyclic op-
erator on a complex Banach space has the property that each of its connected
components meets the unit circle; this is the content of Kitai’s theorem. In
particular, the spectrum cannot have isolated points outside the unit circle.
We will now show that the spectrum of frequently hypercyclic operators, just
like that of chaotic operators (see Proposition 5.7), cannot even have isolated
points on the unit circle.
We start with a crucial lemma whose proof uses complex analysis in a very
clever way.
Lemma 9.37. Let T be an operator on a real Fréchet space X. Let x ∈ X
and x∗ ∈ X ∗ with x, x∗
= 0 be such that
is nonempty.
We now consider the series
∞
z(z − 1) · · · (z − k + 1)
f (z) = (T − I)k x, x∗ , z ∈ C,
k!
k=0
|(T − I)n x, x∗ | ≤ M εn , n ≥ 0,
|f (z)| ≤ M eη|z| , z ∈ C.
Lemma 9.38 has another application. First, combining it with Lemma 5.19
yields the following.
Exercises
Exercise 9.1.1. Show that the Herrero–Bourdon theorem also holds for frequent hyper-
cyclicity. In particular, every frequently hypercyclic operator on a Fréchet space admits
a dense T -invariant subspace consisting, except for 0, of frequently hypercyclic vectors.
Exercise 9.1.2. Using Lemma 9.5, show that every weighted shift is frequently hyper-
cyclic on the space ω = KN .
Exercise 9.1.3. Show that every frequently hypercyclic operator on a Fréchet space is
topologically ergodic; see Exercise 1.5.6. Deduce that if T is a frequently hypercyclic
operator on a Banach space then its adjoint T ∗ cannot be frequently hypercyclic. (Hint:
Exercise 2.5.5, Remark 4.17.)
Exercise 9.1.4. Show that every entire function is the sum of two functions that are
frequently hypercyclic for the translation operator T1 f (z) = f (z + 1). (Hint: Use a
variant of the construction in Example 9.6.)
∞
Exercise 9.2.2. Use the Frequent Hypercyclicity Criterion to give a new proof that
Birkhoff’s operators are frequently hypercyclic. (Hint: Example 3.8.)
Exercise 9.2.3. Formulate and prove an analogue of Proposition 9.13 for weighted bi-
lateral shifts.
(Hint: Proceed as in the proof of Proposition 9.17 and consider the coordinates of index
n − m + 1 in Bwm
x − e1 .)
−2/j
as (Nj + j) for Nj + j ≤ n < Nj + 2j, as 1 for Nj + 2j ≤ n < Nj + 3j, and as
(Nj+1 )2/j for Nj + 3j ≤ n < Nj + 4j = Nj+1 . Show that Bw is a weighted shift on p ,
1 ≤ p < ∞, that satisfies the condition given in Proposition 9.17 but not the condition
in the previous exercise. Thus, the condition in Proposition 9.17 does not characterize
frequent hypercyclicity of weighted shifts on p . (Hint: Use the result by Erdős and
Sárközy.)
Exercise 9.2.6. The aim of this exercise is to show that not every vector x ∈ 1 is the
sum of two frequently hypercyclic vectors for the Rolewicz operator 2B. Suppose that
x = y+z with y and z frequently hypercyclic. Then there is an increasing sequence (mk )k
of positive integers such that T mk y < 1 for k ≥ 1; further let dens{n ∈ N0 ; T n z <
1} =: 2δ > 0. Deduce that there are positive integers nk such that T nk z < 1 and
δmk ≤ nk ≤ mk , k ≥ 1 sufficiently large, and hence that T mk x ≤ 1 + 2(1−δ)mk .
Finally find some x ∈ 1 that fails this inequality for any δ > 0 and any increasing
sequence (mk )k of positive integers.
Exercise 9.3.1. Let Lp (T), 1 ≤ p < ∞, be the space of all complex-valued functions
2π
f on T such that f p := ( 0 |f (eit )|p dt)1/p < ∞. Show that T f (λ) = λf (λ) −
(1,λ)
f (ζ) dζ defines a mixing and chaotic operator on Lp (T), where (λ1 , λ2 ) denotes the
positively oriented arc from λ1 to λ2 . (Hint: Consider the indicator functions f = 1(λ,1) .)
264 9 Frequently hypercyclic operators
Exercise 9.3.2. Let X be one of the complex spaces p , 1 ≤ p < ∞, or c0 . Show that
the map D → X, λ → (λn )n , is infinitely differentiable. Deduce that also the maps
D → H 2 , λ → kλ (see Proposition 4.38) and Dτ → Eτ2 , λ → eλ (see Exercise 4.2.4) are
infinitely differentiable.
with some M > 0. (Hint: Combine the ideas of Exercise 4.2.4 and the proof of Theorem
9.25.)
Exercise 9.3.4. Let D be the differentiation operator on H(C). Let φ : ]0, ∞[→ [1, ∞[
be a function with φ(r) → ∞ as r → ∞. Show that there exists an entire function f
that is frequently hypercyclic for D and that satisfies
with some M > 0. (Hint: Look at the proof of Theorem 4.22, using the Frequent Hy-
percyclicity Criterion in the version of Exercise 9.2.1.)
Exercise 9.4.1. In the proof of Theorem 9.27, show that dens(B) ≥ dens(A).
Exercise 9.4.5. Let T be an operator on a (real or complex) Banach space X. Show that
if there is some x∗ ∈ X ∗ , x∗ = 0, and some λ with |λ| = 1 such that (λI−T ∗ )n x∗ 1/n →
0 as n → ∞ then T is not frequently hypercyclic.
Section 9.1. Frequently hypercyclic operators were introduced by Bayart and Grivaux
[38], [40]. The idea of using ergodic theory to obtain the dynamical properties of linear
operators seems to be due to Rudnicki [272] and Flytzanis [152, 153]. Bayart and Grivaux
Sources and comments 265
[40] obtained Lemma 9.5 (see also Bonilla and Grosse-Erdmann [87]) as well as the
frequent hypercyclicity of the Birkhoff operators. The theorem of Erdős and Sárközy
can be found in [296]. Theorem 9.8 is due to Grosse-Erdmann and Peris [185]; Bayart
and Matheron [45] show that this result is essentially optimal.
For an introduction to ergodic theory we refer to Walters [300].
Section 9.2. The Frequent Hypercyclicity Criterion was obtained by Bayart and Gri-
vaux [38, 40]; the form given here is due to Bonilla and Grosse-Erdmann [87]. Grivaux
[173] also provided a probabilistic version of it. Proposition 9.11 is due to Bonilla and
Grosse-Erdmann [87]. The remaining results in this section can essentially be found in
Bayart and Grivaux [40]; see also Bonilla and Grosse-Erdmann [87]. The latter paper also
contains further conditions under which the set F HC(T ) of frequently hypercyclic op-
erators is of first Baire category, or when F HC(T ) + F HC(T ) does or does not coincide
with the full space.
Bayart and Grivaux [41] constructed a weighted shift on c0 that is frequently hy-
percyclic, but neither chaotic nor mixing; this also shows that not every frequently
hypercyclic operator satisfies the Frequent Hypercyclicity Criterion, and that Proposi-
tion 9.13 does not characterize frequently hypercyclic weighted shifts on c0 . Badea and
Grivaux [19] found operators on a Hilbert space that are frequently hypercyclic and
chaotic but not mixing.
It remains an open problem whether every chaotic operator is frequently hypercyclic,
and to find a characterization of frequently hypercyclic weighted shifts, even on 2 or on
c0 .
Section 9.3. The proof of Theorem 9.22 follows Bayart and Grivaux [38]; see also [39].
Theorems 9.25 and 9.26 are due to Blasco, Bonilla and Grosse-Erdmann [86, 76]; these
authors also show that the operators of differentiation and translation on the space of
harmonic functions on RN are frequently hypercyclic, and they obtain some related
growth results.
In order to keep the presentation simple we have imposed rather strong assumptions
on the eigenvector fields. A much deeper analysis leads to one of the most striking results
in linear dynamics.
To be more specific, an operator T on a complex separable Banach space X is said
to have a perfectly spanning set of eigenvectors associated to unimodular eigenvalues if
one of the following two equivalent conditions holds:
(i) there exists an atomless probability measure σ on T such that, for any measurable
set A ⊂ T with σ(A) = 1, span{ker(λI − T ) ; λ ∈ A} is dense in X;
(ii) for any countable set D ⊂ T, span{ker(λI − T ) ; λ ∈ T \ D} is dense in X.
These conditions were first introduced by Flytzanis [152, 153]. Their equivalence was
shown by Grivaux [174], who also obtained the following fundamental principle.
Theorem 9.42. Any operator on a complex separable Banach space with a perfectly
spanning set of eigenvectors associated to unimodular eigenvalues is frequently hyper-
cyclic.
When the underlying space is even a Hilbert space then one can show that there exists
a Borel probability measure of full support on X with respect to which T is ergodic (see
Bayart and Grivaux [40]); as explained in Section 9.1, this immediately implies that T
is frequently hypercyclic. The measure can even be a so-called Gaussian measure. A
similar result for nuclear Fréchet spaces is due to Grosse-Erdmann [182]. For surveys
on the application of ergodic theory to linear dynamics we refer to Godefroy [164] and
Grosse-Erdmann [182]. A detailed treatment can be found in Bayart and Matheron [44].
Bayart and Grivaux [40, 41] have applied their results to various operators. In par-
ticular they have shown that if ϕ is an automorphism of the unit disk D then the
266 9 Frequently hypercyclic operators
Section 9.4. Theorem 9.27 is due to Bayart and Grivaux [40], whose proof uses Ansari’s
theorem. The alternative proof given in Grosse-Erdmann and Peris [185] contains an
error; in fact, Example 9.28 contradicts Theorem 1.4 in that paper. The proof given
here is due to Grosse-Erdmann and Peris [186].
Theorem 9.33 provides a new common approach to Theorems 9.35 and Theorem 9.36
that were previously obtained by Bayart and Matheron [44] and by Conejero, Müller
and Peris [110], respectively. The remainder of the section, including Theorem 9.39
and Corollary 9.41, is due to Shkarin [287]. Grivaux [174] has recently shown that the
necessary spectral conditions of Theorems 5.6 and 9.39 actually characterize spectra of
frequently hypercyclic operators on Hilbert spaces.
Theorem 9.43. Let K ⊂ C be a nonempty compact set. There exists a frequently hy-
percyclic operator T on a complex Hilbert space such that σ(T ) = K if and only if K
has no isolated points and each of its connected components meets the unit circle.
Further interesting results on frequent hypercyclicity include the facts that every op-
erator on an infinite-dimensional complex separable Hilbert space is the sum of two
frequently hypercyclic operators (Bayart and Grivaux [40]) and that every infinite-
dimensional complex Fréchet space with an unconditional basis supports a frequently
hypercyclic and chaotic operator (De la Rosa, Frerick, Grivaux, and Peris [127]).
Many questions concerning frequently hypercyclic operators remain open. For exam-
ple (see Bayart and Grivaux [40]), whether the frequent hypercyclicity of an operator T
is inherited by its direct sum T ⊕ T ; and whether it is inherited by its inverse T −1 , if it
exists.
Exercises. Exercise 9.1.1 is taken from Bayart and Grivaux [40], Exercises 9.1.4, 9.2.1
and 9.2.6 from Bonilla and Grosse-Erdmann [87], and Exercises 9.2.4 and 9.2.5 from
Grosse-Erdmann and Peris [185]. For Exercise 9.3.1 we refer to Bayart and Grivaux
[39], for Exercise 9.3.3 to Bonilla and Grosse-Erdmann [86]. Exercise 9.3.4 is taken from
Blasco, Bonilla and Grosse-Erdmann [76] who also show that, in the converse direction,
given any function φ : R+ → R+ with limr→∞ φ(r) = 0 there r
is no entire function f
that is frequently hypercyclic for D such that |f (z)| ≤ φ(r) re1/4 for |z| = r sufficiently
large. Exercise 9.3.5 is taken from Bayart and Grivaux [40], Exercise 9.4.2 from Costakis
and Ruzsa [122], Exercise 9.4.4 from Grosse-Erdmann and Peris [185], and Exercise 9.4.5
from Shkarin [287].
Chapter 10
Hypercyclic subspaces
In the present context the following terminology has been generally accepted.
We have already met the notion of a continuous norm; see Lemma 8.7.
Example 10.3. Of course, the norm in any Banach space is continuous. The
Fréchet space H(C) of entire functions has a continuous norm. One may take,
for example, f = sup|z|≤1 |f (z)|, f ∈ H(C). By contrast, the space ω = KN
of all sequences (see Example 2.2) has no continuous norm: it would have to
be dominated by a multiple of some seminorm pn (x) = sup1≤k≤n |xk | (see
Exercise 2.1.7) and would thus assign the value 0 to some nonzero vector.
Remark 10.4. While Montes’ theorem is usually applied in the form stated
above, we will see that it remains true if condition (ii) is replaced by the
following weaker condition:
(ii ) there exists an infinite-dimensional closed subspace M0 of X such that
(T nk x)k converges for all x ∈ M0 .
We will give two proofs of this result. The first one is straightforward but
slightly technical, and it uses the notion of a basic sequence. The second
one, which we describe in the next section, provides a very interesting link
between the hypercyclicity of an operator T and the hypercyclicity of the
corresponding left-multiplication operator LT : S → T S.
Let us turn to the first proof. As is the case for several results in this
chapter, the proof of Montes’ theorem is considerably more transparent when
X is a Banach space. We will therefore restrict ourselves here to these spaces.
The proof in the general case will be given in Section 10.5.
e∗n : M → K, x → an
1
|y, fn∗ | = |an | = |x, e∗n | ≤ e∗n x ≤ e∗ y,
1−δ n
Since we will obtain a more general result later (see Lemma 10.39), we
omit the proof.
Proof of Theorem
10.7. First, let (εn )n be a sequence of positive numbers
such that ∞n=1 (1 + εn ) ≤ 2. Choose e1 ∈ X with e1 = 1. By Lemma 10.8
we can inductively construct vectors e2 , e3 , . . . of norm 1 such that, for all
n ≥ 1 and a1 , . . . , an+1 ∈ K,
n+1
n
ak ek ≤ (1 + εn ) ak ek .
k=1 k=1
k−1
k n
|ak | = ak ek ≤ aj ej + aj ej ≤ 4 aj ej . (10.5)
j=1 j=1 j=1
Nν
xν = aν,k ek
k=1
We are now in a position to prove Theorem 10.2 in a special case.
Proof of Theorem 10.2 for Banach spaces. The proof will be divided into
three steps. To simplify notation we perform the proof in the case when
(nk )k is the full sequence of positive integers. The general case follows in
exactly the same way.
Step 2. We show that (en )n can be perturbed into a basic sequence (fn )n of
hypercyclic vectors. To this end, let Kn = max(1, e∗n ). Further let X0 and
Y0 be the dense subsets of X appearing in the Hypercyclicity Criterion and
let (yn )n be a sequence in Y0 that is dense in X. We claim that there then
exist vectors xj,k ∈ X0 and positive integers n(j, k) such that (n(j, k))k≥1 is
increasing for each j ≥ 1 and such that, for all j, k, j , k ≥ 1,
272 10 Hypercyclic subspaces
1
xj,k ≤ , (10.6)
2j+k+1 Kj
1
T n(j,k) xj,k − yk ≤ k , (10.7)
2
n(j ,k ) 1
T xj,k ≤ j+k+k if (j , k ) = (j, k), (10.8)
2 Kj
1
T n(j,k) ej ≤ k . (10.9)
2
This is easily seen by induction with respect to the strict order < on N×N that
is defined by (1, 1) < (1, 2) < (2, 1) < (1, 3) < (2, 2) < (3, 1) < (1, 4) < . . ..
The existence of xj,k and n(j, k) satisfying (10.6), (10.7) and (10.9) follows
from the assumptions on Y0 in the Hypercyclicity Criterion and the fact that
ej ∈ M0 ; note that xj,k ∈ X0 can be achieved because X0 is dense in X.
Condition (10.8) can be rewritten as
1
,k ) 1
T n(j,k) xj ,k ≤ and T n(j xj,k ≤
2j +k +k Kj 2j+k+k Kj
if (j , k ) < (j, k); this condition can therefore also be ensured when we use
the fact that each xj ,k belongs to X0 and that T is continuous.
We now define, for any j ≥ 1,
∞
fj = ej + xj,k .
k=1
T n(m,k) z − yk
≤ T n(m,k) fm − yk + aj T n(m,k) (fj − ej ) + T n(m,k) w
j=m
3
≤ k + |aj |T n(m,k) (fj − ej ) + T n(m,k) w
2
j=m
3 1
≤ k + e∗j w j+k + T n(m,k) w
2 2 Kj
j=m
3 1
≤ + k w + T n(m,k) w → 0
2k 2
as k → ∞ because w ∈ M0 . Since the yk , k ≥ 1, are dense in X, z is
hypercyclic for T .
Remark 10.9. In order to see that Montes’ theorem remains true under con-
dition (ii ) of Remark 10.4 one need only weaken (10.9) to
1
T n(j,k) ej − vj ≤
2k
with certain vj ∈ X, j ≥ 1, so that (10.11) has to be replaced by
3
T n(j,k) fj − vj − yk ≤ .
2k
In addition, there is some v ∈ X such that T n w → v. We can then conclude
as before that T n(m,k) z − vm − v − yk → 0 as k → ∞. Since the vectors
yk + vm + v, k ≥ 1, form a dense set in X, we have again that z is hypercyclic.
Our first application of Montes’ theorem treats weighted shifts.
Example 10.10. Let w = (wn )n be a bounded weight sequence and Bw the
corresponding weighted backward shift on one of the spaces X = p , 1 ≤
p < ∞, or X = c0 ; see Section 4.1. By Example 4.9(a), Bw satisfies the
Hypercyclicity Criterion if (and only if)
n
sup |wν | = ∞.
n≥1 ν=1
274 10 Hypercyclic subspaces
subspace. In particular, the backward shift on the Bergman space does; see
Example 4.9(a).
In order to show the claim, let C denote the latter supremum, and suppose
that the Hypercyclicity Criterion is satisfied for the sequence (nk )k . Setting
m1 = n1 , we can find a subsequence (mk )k of (nk )k such that, for k ≥ 1,
mk
|wν+μ | ≤ C + 1 for μ ≥ mk+1 − mk .
ν=1
mj
|wν | ≤ C + 1. (10.13)
ν=mj −mk +1
M0 := ak emk ; (ak )k ∈ X
k=1
∞
is an infinite-dimensional closed subspace of X. For any x = k=1 ak emk ∈
M0 we have, using (10.13), that
∞
mj
T mk
x = aj T mk e mj = aj wν emj −mk
j=1 j>k ν=mj −mk +1
≤ (C + 1) aj emj −mk → 0
j>k
We remark that this result remains true for any λ with |λ| ≤ 1; see Exercise
10.1.3.
Now let M0 be the closed subspace of H(C) of all entire functions f of the
form
∞
f (z) = ak z nk −1 .
k=1
We claim that
Dnk f → 0 in H(C) as k → ∞.
Indeed, let R ≥ 1. Then we have that
276 10 Hypercyclic subspaces
∞
sup |Dnk f (z)| = sup aj Dnk z nj −1
|z|≤R |z|≤R j=k+1
∞
≤ |aj | (nj − 1) · · · (nj − nk )Rnj −nk −1
j=k+1
∞
≤ |aj | nnj k Rnj
j=k+1
∞
≤ |aj |(2R)nj → 0
j=k+1
LT : L(X) → L(X), LT S = T S,
shows that LT indeed defines an operator on L(X) when the latter is endowed
with the operator norm topology.
This new approach of studying the operator T is quite promising. For if
we can describe properties of an orbit
of some fixed operator S under LT then we can deduce properties of the orbit
{T n Sx ; n ≥ 0} ⊂ X (10.17)
10.2 Hypercyclic left-multiplication operators 277
For the proof we need a related result concerning the dual X ∗ of X. Note
that the dual of a separable Banach space need not be separable under the
usual operator norm; a simple example is provided by the sequence space
1 whose dual is ∞ . Again we obtain separability under a weaker topology.
Just like the strong operator topology on L(X), the weak-∗-topology on X ∗
is defined as the topology of pointwise convergence on X. A base of neigh-
bourhoods of an element x∗ ∈ X ∗ is given by
where e∗1 , . . . , e∗m are the coordinate functionals corresponding to the (finite)
basic sequence (ej )j=1,...,m (see Definition 10.5 and the subsequent discus-
sion); we assume that the e∗j are extended continuously to all of X, with
preservation of the norm.
We show that this countable set of functionals is weak-∗-dense in X ∗ . To
see this, let x∗ ∈ X ∗ , let x1 , . . . , xm , m ≥ 1, be linearly independent vectors
278 10 Hypercyclic subspaces
ε
m
< + |qj |e∗j ek − xk
2 j=1
ε ε m
≤ + 2M m |qj | ≤ ε,
2 4M ( j=1 |qj | + 1) j=1
as had to be shown.
The main result of this section provides us with a close link between the
hypercyclicity properties of the operator T and those of the operator LT .
K = KΦ = span{ · , y ∗ x ; y ∗ ∈ Φ, x ∈ X} ⊂ L(X),
where the closure is taken in the operator norm topology. Moreover, since
m
m
· , yj∗ xj ∈ U and · , yj∗ zj ∈ V ;
j=1 j=1
m
m
· , yj∗ xj ∈ U and · , yj∗ zj ∈ V.
j=1 j=1
m
m
S := · , yj∗ xj ∈ U and (LT )n S = T n S = · , yj∗ T n xj ∈ V,
j=1 j=1
Then exactly as in the proof of Theorem 10.20, using Theorems 3.24 and
3.25, we obtain the following generalization.
Second proof of Theorem 10.2 for Banach spaces. We suppose that T satis-
fies the Hypercyclicity Criterion for an increasing sequence (nk )k of positive
integers. Then also the sequence (T nk )k satisfies the Hypercyclicity Criterion;
see Theorem 3.24. By the preceding theorem, the sequence (LT nk )k admits
an SOT-hypercyclic vector S ∈ L(X). Since any nonzero multiple of an SOT-
hypercyclic vector is SOT-hypercyclic we may assume that S = 12 . Then,
for any x ∈ X,
1
(I + S)x ≥ x − Sx ≥ x.
2
Hence I + S is an injective operator, and it defines an isomorphism of X onto
its range ran(I + S).
Now let M0 be an infinite-dimensional closed subspace of X so that
T nk x → 0 for all x ∈ M0 . By the above, M := (I + S)M0 is an infinite-
dimensional closed subspace of X. We claim that every nonzero vector x ∈ M
282 10 Hypercyclic subspaces
T nk x = T nk y + T nk Sy, k ≥ 1,
For the deduction of Montes’ theorem in the Fréchet space setting we refer
to Section 10.5.
In this section we want to show that not all hypercyclic operators have hy-
percyclic subspaces. To this end we derive a useful necessary condition for
the existence of a hypercyclic subspace in which the behaviour of the iterates
of T on subspaces of finite codimension plays an essential role. We recall that
a subspace M of a vector space X is said to be of finite codimension if there
is a finite-dimensional subspace E of X such that X = M + E. We collect
here some rather obvious properties, whose proof we leave to the reader; see
Exercise 10.3.1.
Lemma 10.23. Let X be a vector space over K. Then we have the following:
(i) a subspace M is of finite codimension if and only if there is some n ≥ 1
and a linear map u : X → Kn such that M = ker u; n
(ii) if M1 , . . . , Mn are subspaces of finite codimension then so is k=1 Mk ;
(iii) if M is a subspace of finite codimension and T : X → X is a linear
map then T −1 (M ) is of finite codimension;
(iv) if M is an infinite-dimensional subspace and L is a subspace of finite
codimension then M ∩ L is infinite-dimensional.
We turn to the announced condition that prevents an operator from having
hypercyclic subspaces. We first state a version for Banach spaces; later on we
will present a more technical result for general Fréchet spaces.
Theorem 10.24. Let T be an operator on a Banach space X. Suppose that
there are subspaces Mn ⊂ X, n ≥ 1, of finite codimension and positive num-
bers Cn , n ≥ 1, with Cn → ∞ as n → ∞ such that
Proof of Theorem 10.24 for Hilbert spaces. Thus, let X be a Hilbert space.
We assume that T satisfies the hypotheses of the theorem and that M is
an infinite-dimensional closed subspace of X; we need to show then that M
contains a vector x = 0 that is not hypercyclic for T . First, it follows from
the assumption that there is an increasing sequence (kn )n of positive integers
such that, for n ≥ 2,
which we can assume to have norm n12 ; hence also (10.19) and (10.21) are
satisfied.
Having constructed the xn and Ln , n ≥ 1, we set
∞
x= xn ;
n=1
n−1 ∞
j j j
T x= T xν + T xn + T j xν .
ν=1 ν=n+1
T j xν ∈ Ln+1 .
284 10 Hypercyclic subspaces
∞
Since Ln+1 is closed, ν=n+1 T j xν belongs to Ln+1 and hence is orthogonal
n
to ν=1 T j xν by (10.20). By Pythagoras’ theorem we have for any orthogonal
vectors y, z ∈ X that y + z = (y2 + z2 )1/2 ≥ y. In our present
situation we therefore have that
n−1
j
T j x ≥ T xν + T j xn .
ν=1
n−1
Moreover, by (10.20) and (10.21) we have that ν=1 T j xν is orthogonal to
T j xn . Again by Pythagoras’ theorem this yields that
T j x ≥ T j xn ,
n3
T j x ≥ Cj xn ≥ = n.
n2
Thus T j x → ∞ as j → ∞, which shows that x ∈ M , x = 0, is not a
hypercyclic vector for T , as desired.
For Banach spaces this proof breaks down because we no longer have a
notion of orthogonality. In Section 10.5 we will prove the following general-
ization of Theorem 10.24, which then also contains the Banach space case.
We remark that the additional assumption that pN (x) > 0 for any hy-
percyclic vector is automatically satisfied if the defining seminorms can be
chosen to be norms, that is, if the space admits a continuous norm.
Our first application uses the criterion in the case of Banach spaces.
(Bw )n (xk )k = (w2 · · · wn+1 xn+1 , w3 · · · wn+2 xn+2 , w4 · · · wn+3 xn+3 , . . .);
For this we will apply Theorem 10.25. First, it is clear that every hyper-
cyclic sequence x satisfies p1 (x) > 0. For Mn we consider the subspaces of
finite codimension
p1 (T n x) = 2n pn (x).
We first show that operators with hypercyclic subspaces exist in a large class
of Fréchet spaces.
The usual techniques of Chapter 8 then also show that, in the setting of the
theorem, the mixing operators with a hypercyclic subspace are SOT-dense in
the space of operators on X; see Exercise 10.4.1.
Montes’ theorem provides us with a powerful tool for verifying that an
operator has hypercyclic subspaces. But it leaves us with the task of finding
an infinite-dimensional closed subspace on which (sub)orbits converge to 0. In
many cases, coming up with such a subspace is by no means easy or obvious.
In the remainder of this section we address this problem in two particular
cases. We will see that once again the notion of a basic sequence is crucial.
Our first result is in the spirit of Theorem 10.24.
Proof. For simplicity we will perform the proof in the case when (nk )k is the
full sequence; the general case follows in the same way.
By (i) there is a dense subset X0 of X such that, for all x ∈ X0 , T n x → 0 as
n → ∞. By (ii) there is some C > 0 such that T n |Mn ≤ C for n ≥ 1. And by
a slight strengthening of Mazur’s theorem (see Exercise 10.4.4), there exists
a basic sequence (en )n in X with en ∈ Mn for n ≥ 1; let Kn = max(1, e∗n ).
By density of X0 and continuity of T we can then find fn ∈ X0 , n ≥ 1,
such that
1
T n ej − T n fj < j , j ≥ 1, n = 0, 1, . . . , j. (10.22)
2 Kj
Since the fn belong to X0 one can construct inductively an increasing se-
quence (nk )k of positive integers such that, for all k ≥ 1,
1
T nk fj ≤ , j = 1, . . . , nk−1 , (10.23)
2j+k Kj
we have by Lemma 10.6 that (fnk )k is a basic sequence. Let M0 be its closed
linear span.
nk
∞the powers T of T tend pointwise to
We now claim that 0 on M0 . Indeed,
∞
let x ∈ M0 , x = k=1 ak fnk . By Lemma 10.6, also z := k=1 ak enk con-
verges. Moreover, since the subspaces Mk are closed and (Mk )k is decreasing,
we have that
∞
aj enj ∈ Mnk , k ≥ 1. (10.24)
j=k
k−1
∞ ∞
T nk x ≤ |aj |T nk fnj + T nk aj enj + |aj |T nk (fnj − enj )
j=1 j=k j=k
k−1
∞ ∞
1 1
≤ e∗nj z nj +k + C aj enj + e∗nj z
j=1
2 K nj 2nj Knj
j=k j=k
∞
1 2
≤ k z + C aj enj + n z → 0
2 2 k
j=k
As an application one obtains a more direct verification of Example 10.10;
see Exercise 10.4.2.
288 10 Hypercyclic subspaces
We use the theorem here to obtain a new, large class of operators with
hypercyclic subspaces. For this we need the following.
Lemma 10.30. Let K be a compact operator on a Banach space X. Then
there is a closed subspace M of finite codimension such that K|M ≤ 12 .
Proof. Since the image under K of the closed unit sphere is relatively compact
in X there are finitely many points xk ∈ X with xk = 1 for k = 1, . . . , N ,
such that, for any x ∈ X with x = 1 there is some k with Kxk −Kx ≤ 14 .
By the Hahn–Banach theorem (see Appendix A), there are continuous linear
functionals yk∗ , k = 1, . . . , N , on X such that yk∗ (Kxk ) = Kxk and yk∗ =
1. By Lemma 10.23,
N
M= ker(yk∗ ◦ K)
k=1
We can now present the announced class of operators with hypercyclic
subspaces.
Corollary 10.31. Let T be an operator on a separable Banach space of the
form
T = U + K,
where U ≤ 1 and K is compact. If T satisfies the Hypercyclicity Criterion
then it has a hypercyclic subspace.
Proof. For n ≥ 1 we have that
T n = (U + K)n = U n + Kn
with compact operators Kn ; see Exercise 5.2.3. It then follows from Lemma
10.30 that there are closed subspaces Mn of finite codimension such that
Kn |Mn ≤ 12 , hence
3
T n |Mn = (U n + Kn )|Mn ≤ .
2
Since X must be infinite dimensional, an application of Theorem 10.29, us-
ing the decreasing sequence (M1 ∩ . . . ∩ Mn )n of infinite-dimensional closed
subspaces, yields the result.
10.4 Further operators with hypercyclic subspaces 289
system in L2 (T) and are therefore a basic sequence there; the norms of the
corresponding coefficient functionals are 1.
By (10.25) we can apply Runge’s theorem to obtain, for any j ≥ 1, func-
tions fj,k ∈ H(Ω) such that
1 1
sup |ej (z) − fj,1 (z)| < , sup |fj,1 (z)| < (10.27)
z∈L1 2j+2 z∈ϕn1 (L 1)
2j+2
and, for k ≥ 2,
1 1
sup |fj,k−1 (z) − fj,k (z)| < , sup |fj,k (z)| < . (10.28)
z∈Lk 2j+k+1 z∈ϕnk (Lk ) 2j+k+1
Nν
gν = aν,j fj
j=1
that converge to g in H(Ω) and then also in L2 (T). Fix m ≥ 1; then we have
for k ≥ m that Lm ⊂ Lk and therefore, using (10.29),
10.5 The Fréchet space setting 291
Nν
sup |(Cϕ )nk g(z)| ≤ sup |(Cϕ )nk (g − gν )(z)| + |aν,j | sup |fj (ϕnk (z))|
z∈Lm z∈Lm j=1 z∈Lk
∞
1
≤ sup |(g − gν )(ϕnk (z))| + fj∗ gν 2
z∈Lm 2j+k
j=1
2
≤ sup |(g − gν )(w)| + gν 2 .
w∈ϕnk (Lm ) 2k
In this section we discuss the extensions of the main results of the previous
sections to Fréchet spaces.
Proof of Theorem 10.2. As in the Banach space case we assume that (nk )k
is the full sequence of positive integers.
Step 2. We will again perturb (en )n into a basic sequence (fn )n of hypercyclic
vectors. To do this, let Kn = max(1, |e∗n |), n ≥ 1, where |e∗n | is the norm
of the coefficient functionals e∗n on the space (X, | · |). Furthermore, let
· denote an F-norm defining the topology of X; see Section 2.1. As in
the Banach space case there is then a dense sequence (yn )n in X, vectors
xj,k ∈ X and positive integers n(j, k) such that (n(j, k))k≥1 is increasing for
each j ≥ 1 and such that, for all j, k, j , k ≥ 1,
292 10 Hypercyclic subspaces
1
max(xj,k , |xj,k |) ≤ , (10.30)
2j+k+1 Kj
1
T n(j,k) xj,k − yk ≤ k ,
2
n(j ,k ) 1
T xj,k ≤ j+k+k if (j , k ) = (j, k),
2 Kj
1
T n(j,k) ej ≤ k .
2
Defining
In (10.30) we have used the continuity of the inclusion of X into X.
∞
fj = ej + xj,k , j ≥ 1,
k=1
4 1
≤ + k |w| + T n(m,k) w. (10.35)
2k 2
Now let z ∈ M , z = 0. We want to show that z is hypercyclic for T . As z
we have a
also belongs to the closed linear span of the fn when taken in X,
representation
∞
z= a j fj
j=1
Nν
zν := aν,j fj
j=1
Since (aν,j )ν converges for all j ≥ 1 and since, by (2.3), (10.31) and (10.34),
with some C > 0, the dominated convergence theorem implies that (wν −zν )ν
is a Cauchy sequence in X; hence (wν )ν converges to some vector w, which
necessarily belongs to M0 .
By our previous argument, zν and wν satisfy (10.35) for any ν ≥ 1. Letting
ν → ∞ we have by continuity that, for any k ≥ 1,
4 1
T n(m,k) z − yk ≤ k
+ k |w| + T n(m,k) w.
2 2
Since w ∈ M0 we obtain that T n(m,k) z − yk → 0 as k → ∞, which implies
that z is hypercyclic.
Remark 10.34. Similar modifications to those in the Banach space case yield
that condition (ii ) of Remark 10.4 suffices in Montes’ theorem.
Note that the strong operator topology on L(X) is defined in Section 8.3,
and the definition of SOT-hypercyclicity is the same as in Definitions 10.16
and 10.21. Proposition 10.18 extends as well, with unchanged proof.
We endow the space L(X|·| , X) of operators X|·| → X with its natu-
ral operator topology. More precisely, if (pn )n is an increasing sequence of
seminorms defining the topology of X then we set
m
K = KΦ = span{ · , y ∗ x ; y ∗ ∈ Φ, x ∈ X},
Proof. (i) Suppose that · is an F-norm that defines the topology of X. Let
T ∈ L(X), let x1 , . . . , xm , m ≥ 1, be linearly independent vectors of X, and
ε > 0. The coordinate functionals x∗1 , . . . , x∗m corresponding to the basic se-
quence (xj )j=1,...,m are continuous with respect to | · | on span{x1 , . . . , xm }
(because all norms on a finite-dimensional space are equivalent) and hence
have continuous linear extensions to X|·| . One then concludes exactly as
in the proof of Proposition 10.14 that there is some S ∈ F such that
T xk − Sxk < ε for k = 1, . . . , m.
(ii) Since X is separable we can choose E to be countable. Then F is
countable, and the assertion follows from (i).
(iii) The proof of this assertion can be given as in Exercise 5.2.4.
(iv) Since, for any y ∗ ∈ Φ and x ∈ X,
Lemma 10.37. Under the assumptions of Theorem 10.35 and Lemma 10.36,
the following assertions are equivalent:
(i) (Tn )n satisfies the Hypercyclicity Criterion;
(ii) (LTn )n is hypercyclic on K;
(iii) (LTn )n is SOT-hypercyclic.
Moreover, if S ∈ K is hypercyclic for (LTn )n on K, then it is SOT-hypercyclic
for (LTn )n on L(X).
Proof. The proof follows the same lines as that of Theorem 10.20, using
(10.37), Theorem 3.24 and Theorem 3.25.
296 10 Hypercyclic subspaces
Lemma 10.38. Under the assumptions of Theorem 10.35 and Lemma 10.36,
let S ∈ K be hypercyclic for (LTn )n on K. Then there is some λ = 0 such
that |λSx| ≤ 12 |x| for all x ∈ X. If M0 is an infinite-dimensional closed
subspace of X such that Tn x → 0 for all x ∈ M0 then (I + λS)M0 is a
hypercyclic subspace for (Tn )n .
Tn x = Tn y + λTn Sy, n ≥ 1,
The similarities between this proof and that of Lemma 10.30 are evident.
In Exercise 10.5.4 we ask the reader to prove a common generalization of the
two lemmas in the case of Banach spaces.
We can now prove the generalization of Theorem 10.24 to Fréchet spaces.
Proof of Theorem 10.25. Suppose that M is a hypercyclic subspace for T .
Then, exactly as in the proof of Theorem 10.24, there is an increasing se-
quence (kn )n of positive integers such that Cj ≥ n3 for kn−1 < j ≤ kn , n ≥ 2,
and with Lemma 10.39 we can construct points xn in X, finite-dimensional
subspaces En of X and corresponding closed subspaces Ln of finite codimen-
sion satisfying (10.39) for all x ∈ Ln , y ∈ En , with p = pN , such that, for
n ≥ 1 and 1 ≤ j ≤ kn ,
1
pn (xn ) = (n ≥ N ), xn ∈ M ∩ Mj , (10.40)
n2
T j x1 , . . . , T j xn−1 ∈ En (n ≥ 2),
T j xn ∈ L1 ∩ . . . ∩ Ln ,
it follows from (10.40) that the series converges in X, and we have that x ∈ M .
Once more following the proof of Theorem 10.24, replacing orthogonality by
condition (10.39), we find that, for n ≥ N and kn−1 < j ≤ kn ,
1 n pN (T j xn ) Cj pj (xn )
pN (T j x) ≥ pN T j xν ≥ ≥
1+ε ν=1
(1 + ε)(2 + ε) (1 + ε)(2 + ε)
n3 pn (xn ) n
≥ = →∞
(1 + ε)(2 + ε) (1 + ε)(2 + ε)
Exercises
Exercise 10.1.1. Show that the following operators have hypercyclic subspaces:
(i) the operator T on C0 (R+ ) given by T f (x) = λf (x + a), a > 0, λ > 1 (see Exercise
2.2.1);
(ii) the operator T on Lpv (R+ ), 1 ≤ p < ∞, given by T f (x) = f (x + 1), where v is an
admissible weight function with lim inf x→∞ v(x) = 0; see Exercise 2.2.4.
Exercises 299
Exercise 10.1.2. Show that a Fréchet space has a continuous norm if and only if its
topology can be defined by an increasing sequence (pn )n of norms.
Exercise 10.1.3. Let X be a separable Fréchet space with a continuous norm, and let T
be an operator on X that satisfies the Hypercyclicity Criterion. Show that if ker(λI − T )
is infinite-dimensional for some λ with |λ| ≤ 1, then T has a hypercyclic subspace. (Hint:
Use Remark 10.4 and the León–Müller theorem.)
Exercise 10.1.4. Let B wn be a weighted backward shift on H(C); see Example 4.9(b).
Show that if limn→∞ ( ν=1 |wν |)1/n = ∞ and limn→∞ |wn |1/n = 1 then Bw has a
hypercyclic subspace. See also Exercise 10.3.3.
Exercise 10.1.5. Let T be an operator on a separable Fréchet space. Show the follow-
ing:
(i) T and T n , n ≥ 1, have the same hypercyclic subspaces;
(ii) if T is surjective and M = X is a hypercyclic subspace then Mn := T −n (M ),
n ≥ 0, are hypercyclic subspaces of T ; moreover, Mn = Mk for n = k;
(iii) if T is bijective and M = X is a hypercyclic subspace then Mn := T −n (M ), n ∈ Z,
are hypercyclic subspaces of T ; moreover, Mn = Mk for n = k.
(Hint: Ansari’s theorem.)
Exercise 10.2.2. Let X be a Banach space with separable dual X ∗ . Show that X itself
is separable. (Hint: If {x∗n ; n ≥ 1} is dense in X ∗ , choose xn ∈ X with xn = 1,
|x∗n (xn )| ≥ 12 x∗n ; show that span{xn ; n ≥ 1} is dense in X.)
Exercise 10.2.5. A Banach space X is said to have the approximation property if, for
every Banach space Y , for every compact operator T : Y → X and for every ε > 0 there
is a finite-rank operator F : Y → X such that T − F < ε; see Exercise 5.2.2. Let X
be a Banach space with the approximation property and K(X) the space of compact
∗
operators on X. Let Φ ⊂ X mbe norm-dense and E ⊂ X be dense. Show that then
the operators of the form k=1
· , yk∗ xk , y1∗ , . . . , ym
∗
∈ Φ, x1 , . . . , xm ∈ E, m ≥ 1,
constitute a dense subset of K(X) under the operator norm topology. Deduce that if
X has the approximation property and X ∗ is separable then K(X) is separable. (Hint:
Exercise 10.2.2.)
300 10 Hypercyclic subspaces
Exercise 10.2.6. Let X be a Banach space with the approximation property such that
X ∗ is separable. By Exercises 5.2.4 and 10.2.5, the space K(X) of compact operators is
a separable Banach space under the operator norm topology. Show that LT : K(X) →
K(X), S → T S, is a well-defined operator on K(X). Show moreover that T satisfies the
Hypercyclicity Criterion if and only if LT is hypercyclic on K(X).
Exercise 10.2.7. Let X be a Banach space and T an operator on X. Then the right-
multiplication operator RT on L(X) is defined as RT S = ST for S ∈ L(X).
(a) Let xj ∈ X, yj ∈ X ∗ , j = 1, . . . , m. Show that RT ( j=1
· , yj∗ xj ) =
m
m
j=1
· , T ∗ yj∗ xj .
(b) Suppose that X and X ∗ are separable. Show that if T ∗ : X ∗ → X ∗ satisfies the
Hypercyclicity Criterion then RT is SOT-hypercyclic, that is, there is some S ∈ L(X)
such that {ST n ; n ≥ 0} is SOT-dense in L(X).
Exercise 10.2.9. Verify the contents of Remark 10.4 for Banach spaces using the ap-
proach of this section.
Exercise 10.3.2. Let Bw be a hypercyclic weighted backward shift on one of the spaces
X = p , 1 ≤ p < ∞, or c0 . Show the following:
(i) if lim inf n→∞ |wn | > 1, then Bw has no hypercyclic subspace;
(ii) if (|wn |)n is decreasing, then Bw has a hypercyclic subspace if and only if
limn→∞ |wn | = 1;
(iii) if (|wn |)n is increasing, then Bw is even mixing, but it has no hypercyclic subspace.
Exercise 10.3.3. Let Bw be a weighted backward shift on H(C); see Exercise 10.1.4.
n→∞
Show that if lim |wn |1/n = ∞ then Bw has no hypercyclic subspace. (Hint: The
∞ ∞
seminorms pn ( k=0 ak z k ) = k=0
|ak |nk , n ≥ 1, define the topology of H(C).)
Exercise 10.3.4. Use Example 10.27 to show that the existence of hypercyclic sub-
spaces is not preserved under quasiconjugacies.
Exercise 10.4.1. Using the methods of Section 8.3 show that in every Fréchet space X,
the set of operators having hypercyclic subspaces is either empty or SOT-dense in L(X).
Moreover, if X is an infinite-dimensional separable Fréchet space with a continuous norm
then the set of mixing operators having hypercyclic subspaces is SOT-dense in L(X).
Exercise 10.4.5. Show the following variant of Corollary 10.31. Let T be an operator
on a separable Banach space of the form T = S + K where ker(λI − S) is infinite
dimensional for some λ with |λ| ≤ 1 and K is compact. If T satisfies the Hypercyclicity
Criterion then it has a hypercyclic subspace.
Exercise 10.5.1. Let X be a separable Fréchet space with a continuous norm and T
an operator on X that satisfies the Hypercyclicity Criterion. Let K be the space defined
in Lemma 10.36.
(a) Show that the set of operators from K with dense range is a dense Gδ -set.
(b) Show that LT has an SOT-hypercyclic vector S that is a dense range operator.
(c) Deduce that there is a dense subspace of X consisting, except for zero, of hyper-
cyclic vectors for T . Compare with the Herrero–Bourdon theorem.
Exercise 10.5.2. Let Y be a separable normed space and X a separable Fréchet space.
The space L(Y, X) of operators from Y to X turns into a Fréchet space under seminorms
as in (10.36). For an operator T on X define the operator LT on L(Y, X) by LT S = T S.
The definitions of the strong operator topology on L(Y, X) and SOT-hypercyclicity of
LT are obvious. Show that the following assertions are equivalent when dim Y ≥ 2:
(i) T satisfies the Hypercyclicity Criterion;
(ii) LT is SOT-hypercyclic.
In that case, LT has an SOT-hypercyclic vector S that defines a compact operator
S : Y → X. Deduce Theorem 10.35 for iterates of an operator.
Exercise 10.5.3. Let T be the operator of Example 10.27. Show that it is even chaotic,
but that the left-multiplication operator LT is not SOT-hypercyclic. Hence, in Theorem
10.35, the assumption of existence of a continuous norm cannot be dropped, even for
iterates of an operator. (Hint: Let S ∈ L(X). If P denotes the canonical projection of
X onto 1 , then show that there exists some N ≥ 1 such that P Se−k = 0 for all k ≥ N
and deduce that Se−N is not hypercyclic for T .)
T x + T y − T y ≥ −ε y ,
In view of the fact that every hypercyclic operator possesses a dense subspace all of
whose nonzero vectors are hypercyclic, it seemed natural to reserve the term “hyper-
cyclic subspace” for the more interesting case of infinite-dimensional closed subspaces,
as was first suggested by Chan and Taylor [107]. An example of an operator for which
302 10 Hypercyclic subspaces
Section 10.1. For Banach spaces, Montes’ theorem is due to Montes [240]; see also
González, León and Montes [167]. Our proof is taken from Bonilla and Grosse-Erdmann
[88]; for a similar proof see León and Müller [223]. For Fréchet spaces, the result is due
independently to Bonet, Martínez and Peris [84] who use a tensor product technique,
and to Petersson [258] who adapts Montes’ technique. Bès and Conejero [69] have stud-
ied hypercyclic subspaces of operators on a Fréchet space without a continuous norm;
they have shown that every nonconstant polynomial p(B) of the backward shift has a
hypercyclic subspace on ω = KN .
For more on basic sequences in Banach spaces we refer to Diestel [133].
In the Banach space setting, González, León and Montes [167] (see also León and
Montes [221] and León and Müller [223]) have improved Montes’ theorem to a charac-
terization under the assumption that the operator satisfies the Hypercyclicity Criterion,
or equivalently, that it is weakly mixing.
It is important to note that the sequence (nk )k in (ii) need not be related to the
sequence in the Hypercyclicity Criterion. One way of defining the essential spectrum of
/ σe (T ) if and only if λI − T is a Fredholm operator, that is,
an operator T is that λ ∈
λI − T has finite-dimensional kernel and finite-codimensional closed range.
As an application, González, León and Montes [167], [221] identify the operators with
hypercyclic subspaces among various classes of operators. For example, a hypercyclic
weighted backward shift Bw on 2 has a hypercyclic subspace if and only if
n 1/n
lim inf |wν+k | ≤ 1,
n→∞ k≥0
ν=1
while any hypercyclic bilateral weighted backward shift on 2 (Z) has a hypercyclic sub-
space. A hypercyclic adjoint multiplier Mϕ∗ on the Hardy space H 2 , induced by an injec-
tive bounded holomorphic function ϕ on D (see Section 4.4), has a hypercyclic subspace
if and only if the boundary of ϕ(D) meets the closed unit disk. And for any automor-
phism ϕ of D the composition operator Cϕ on H 2 (see Section 4.5) has a hypercyclic
subspace whenever it is hypercyclic; see also Montes [240].
Remark 10.4 is due to Bernal [57]; in the case of Banach spaces it also follows from
the sufficiency of condition (iii) in Theorem 10.40 and the Banach–Steinhaus theorem.
Corollary 10.11 and Example 10.12 are due to Petersson [258]; hypercyclic subspaces for
Birkhoff’s operators were found earlier by Bernal and Montes [65]; see the remarks on
Proposition 10.33 below. The fact that MacLane’s operator has a hypercyclic subspace
was only recently proved by Shkarin [290]; he also notes that its essential spectrum is
empty, so that Theorem 10.40 breaks down for Fréchet spaces.
Frequently hypercyclic subspaces are introduced and studied in Bonilla and Grosse-
Erdmann [88].
notion of SOT-hypercyclicity and observed Proposition 10.18. Chan and Taylor [107]
and Montes and Romero [242] extended Chan’s investigation to Banach spaces. In all
three papers, the implication (i)=⇒(ii) in Theorem 10.20 is obtained by a construction.
Martínez and Peris [231] show this implication by a tensor product technique; they also
prove the converse implication as well as Theorem 10.22. The proof of Theorem 10.20
given here follows Martínez and Peris [231] but avoids the language of tensor products;
see also Aron, Bès, León and Peris [14]. Remark 10.19 was made in Montes and Romero
[242].
For more on the weak-∗-topology we refer to Diestel [133].
Section 10.3. Montes [240] not only proved his sufficient condition for the existence
of hypercyclic subspaces, he was also the first to come up with operators without such
a subspace, namely Rolewicz’s operators; see Example 10.26. Theorem 10.24 is due to
León and Müller [223]; its extension to Fréchet spaces, Theorem 10.25, seems to be new.
Example 10.27 provides a new proof of a result by Bonet, Martínez and Peris [84].
Section 10.4. Theorem 10.28 was obtained independently by Bernal [57] and Petersson
[258]; our proof follows that of Bernal. For Banach spaces the result was obtained earlier
by León and Montes [220]. The problem of whether Theorem 10.28 remains true for
all infinite-dimensional separable Fréchet spaces remains open; in the case of X = ω, a
positive answer was given by Bès and Conejero [69], as already mentioned.
Theorem 10.29 is due to León and Müller [223], but it is also implicitly contained in
León and Montes [220]. Corollary 10.31 is due to León and Montes [220] who deduced
Theorem 10.28 for Banach spaces from it. In a related result, Petersson [258] has shown
that any weakly mixing nuclear perturbation of the identity on a separable Fréchet space
with a continuous norm has a hypercyclic subspace; he used it to deduce Theorem 10.28.
The result with which we close this section, Proposition 10.33, is due to Bernal and
Montes [65]; historically, it was, in fact, the first result on hypercyclic subspaces after
Read’s theorem mentioned above.
Section 10.5. The proof of Montes’ theorem via basic sequences in a containing Banach
space, given here, is from Bonilla and Grosse-Erdmann [88].
Theorem 10.35 is due to Bonet, Martínez and Peris [84] who refine the tensor product
technique of Martínez and Peris [231]. The proof given here uses similar ideas without
relying on the theory of tensor products.
For Lemma 10.39 in the case of Banach spaces we refer to Müller [246].
Exercises. Exercises 10.1.3 and 10.1.5 are taken from Petersson [258], Exercise 10.2.3
from Chan [99], Exercise 10.2.4 from Martínez and Peris [231], Exercises 10.2.6 and 10.2.7
from Bonet, Martínez and Peris [84]. As for Exercise 10.2.8, the space constructed by
Argyros and Haydon [12] has the approximation property because it has a basis, and its
dual is separable because it is isomorphic to 1 . For Exercise 10.4.1 we refer to Bès and
Chan [100], [67], [68] and Bernal [57]. Exercises 10.4.5 and 10.5.1 are taken from León
and Müller [223], and Exercises 10.5.2 and 10.5.3 from Bonet, Martínez and Peris [84].
Chapter 11
Common hypercyclic vectors
In other words, the common hypercyclic vectors are exactly the elements
of
HC(Tλ ).
λ∈Λ
it serves to show that in many cases the set of common hypercyclic vectors
is a dense Gδ -set.
To this end we suppose that the index set Λ is a metric space. Then a
family (Tλ )λ∈Λ of operators on a Fréchet space X is called continuous if, for
any x ∈ X, the map
Λ → X, λ → Tλ x
is continuous. In the present setting, this implies joint continuity with respect
to λ and x.
Proposition 11.4. Let Λ be a metric space and (Tλ )λ∈Λ a continuous family
of operators on a Fréchet space X. Then the map (λ, x) → Tλ x is continuous
on Λ × X.
Since (Tλn z)n converges for any z ∈ X, the Banach–Steinhaus theorem (The-
orem A.10) implies that (Tλn )n is equicontinuous. It then follows easily that
Tλn xn → Tλ x as n → ∞.
Proof. (a) By separability, the topology of X has a countable base (Vk )k . Let
(Km )m be a sequence of compact sets whose union is Λ. We then have that
HC(Tλ ) = E(Km , Vk ),
λ∈Λ m≥1 k≥1
where we define
Example 11.8. The following are sufficient∞ conditions for uniform uncondi-
series n=1 xλ,n , λ ∈ Λ:
tional convergence of a family of
(i) there is a convergent series ∞n=1 cn of positive numbers such that
Let us also note that the criterion requires a one-dimensional parameter set,
a restriction that we will discuss later.
if 0≤μ−λ< δ
n then T n (λ)Sn (μ)x − x < ε;
Then the set of common hypercyclic vectors of the family (Tλ )λ∈Λ is a dense
Gδ -set, and in particular, nonempty.
In (i), we consider the finite sums as infinite series by adding 0 terms.
Proof. We will verify that the characterizing condition of Theorem 11.5 holds.
Thus let K ⊂ Λ be a compact set, which we can assume to be a subinterval
K = [a, b], and let U and V be nonempty open subsets of X. Then there are
points x0 , y0 ∈ X0 and some ε > 0 such that, whenever x − x0 < ε and
y − y0 < ε then x ∈ U and y ∈ V .
We can deduce from conditions (i), (ii) and (iv) that there is some N ≥ 0
such that, for any finite set F ⊂ {N, N + 1, N + 2, . . .}, we have that
ε
T m (λ)Sm−n (μn )y0 < for m ≥ 0, λ ≥ μ0 ≥ . . . ≥ μm , (11.2)
4
n∈F,n≤m
ε
T m (λ)Sm+n (μn )y0 < for m ≥ 0, λ ≤ μ0 ≤ μ1 ≤ . . . , (11.3)
4
n∈F
ε
T n (λ)x0 < for n ≥ N , λ ∈ K. (11.4)
4
ε
T n (λ)Sn (μ)y0 − y0 < for n ≥ 1, 0 ≤ μ − λ < δ
n. (11.5)
4
By the divergence of the harmonic series there is some J ≥ 1 such that
J−1
δ J
δ
a+ ≤b<a+ .
ν=1
2νN ν=1
2νN
We then set
j
δ
μ0 = a, μj = a + (0 < j < J), μJ = b,
ν=1
2νN
so that a = μ0 ≤ μ1 ≤ . . . ≤ μJ = b.
After these preparations we define
x − x0 = SN (μ1 )y0 + S2N (μ2 )y0 + . . . + SJN (μJ )y0 < ε,
j−1
T jN (λ)x − y0 = T jN (λ)x0 + T jN (λ)SνN (μν )y0
ν=1
J
+ T jN
(λ)SjN (μj )y0 − y0 + T jN (λ)SνN (μν )y0
ν=j+1
j−1
jN
=T (λ)x0 + T jN (λ)SjN −αN (μj−α )y0
α=1
J−j
+ T jN
(λ)SjN (μj )y0 − y0 + T jN (λ)SjN +αN (μj+α )y0
α=1
ε ε ε ε
< + + + = ε,
4 4 4 4
where we have applied, in turn, (11.4), (11.2), (11.5) and (11.3). This implies
that
T jN (λ)x ∈ V,
which proves the claim.
Remark 11.10. (a) In applications one often has that S0 (λ)x = x for all λ ∈ Λ
and x ∈ X0 . In such a case condition (iv) can be dropped because it follows
from condition (i) by considering the index n = m. Incidentally, condition
(iii) implies that T n (λ)Sn (λ)x = x for all n ≥ 1, λ ∈ K and x ∈ X0 .
(b) The conditions in the Common Hypercyclicity Criterion are rather
strong. They imply that any operator Tλ , λ ∈ Λ, is frequently hypercyclic
and mixing; see Exercise 9.2.1 and Remark 3.13(a). Moreover, if we can take,
for any λ ∈ Λ, Sn (λ) = S n (λ), n ≥ 0, with some map S(λ) : X0 → X0 , then
each operator Tλ , λ ∈ Λ, is also chaotic; see Proposition 9.11.
(c) Condition (i) in particular is quite restrictive; see the discussion before
Example 11.18.
(d) It is obvious from the proof that condition (iii) can be relaxed. Clearly,
∞onlyδ property of the sequence (δ/n)n that we needed was that the series
the
n=1 nN diverges for any N ≥ 1. Thus, condition (iii) can be weakened to
the following:
(iii ) for any ε
> 0 there is a decreasing sequence (δn )n of positive numbers
∞
such that n=1 δn diverges and such that, for all n ≥ 1, λ, μ ∈ K,
(iii) seems to create a serious problem. The property that we needed in the
proof was that, essentially, the balls of radius δ/n about the μn cover the
compact part K of the parameter set. To have something similar in a complex√
setting, say, would require choosing larger balls, for example of radius δ/ n;
see also the argument in Example 11.3. But in this weakened form condition
(iii) can usually no longer be satisfied.
Let us now see the Common Hypercyclicity Criterion at work.
Example 11.11. (Rolewicz’s operators) The multiples λB, |λ| > 1, of the
backward shift on any of the spaces p , 1 ≤ p < ∞, or c0 , share a common
hypercyclic vector, and the set of common hypercyclic vectors is a dense
Gδ -set.
Thanks to the León–Müller theorem, for any fixed λ > 1, the operators
zλB, |z| = 1, have the same hypercyclic vectors. Hence it suffices to prove
the claim for Λ = ]1, ∞[.
In order to apply the Common Hypercyclicity Criterion, let K = [a, b]
with 1 < a < b. As is usual for backward shifts, we set S(λ) = λ1 F , with F
the forward shift (see Example 3.6), and we let Sn (λ) = S n (λ), n ≥ 0. For
X0 we choose the set of finite sequences.
It then suffices to verify conditions (i)–(iii) for the canonical unit sequences
x = ek , k ≥ 1; see Remark 11.10(a). For any finite set G ⊂ {N, N + 1, . . .}
we have that
λm
T m (λ)Sm−n (μn )ek = n
m−n B ek ,
n∈G,n≤m n∈G,n≤m
μn
Therefore, if 0 ≤ μ − λ < δ
n and λ, μ ∈ K, then
λ n 1
λ δ δ
0≤1− =n tn−1 dt ≤ n 1 − < ≤ , (11.6)
μ λ/μ μ μ a
λT, λ ∈ K,
λ n
T n (λ)Sn (μ)x − x = − 1 x < ε,
μ
which implies condition (iii).
Finally, condition (iv) is trivial.
proof.
In Banach spaces the theorem leads to a very general result under rather
weak hypotheses.
11.2 Common hypercyclic vectors for multiples of an operator 315
Proof. By the open mapping theorem, the image of the open unit ball under
T contains the open ball of radius 2ε for some ε > 0. Hence for any y ∈ X
with y = ε there is some x ∈ X with x < 1, which we call Sy, such that
T x = T Sy = y. For general y ∈ X, y
= 0, we define
y
ε
Sy = S y ,
ε y
In situations where Theorem 11.13 can be applied, the task that remains
is that of determining a possibly small value of λ0 . We start with a general-
ization of Example 11.11; see Section 4.1 for the general framework of this
result.
J−1
δ J
δ
b+ ≤c<b+ .
ν=1
3νN ν=1
3νN
We set
j
δ
λ0 = b, λj = b + (0 < j < J), λJ = c
ν=1
3νN
and nj = 3jN for j = 1, . . . , J.
Applying Runge’s theorem to the union of the closed balls of radius N
around 0, n1 , . . . , nJ (see Figure 11.1), we obtain an entire function h such
that
sup |f (z) − h(z)| < ε
|z|≤N
and, for j = 1, . . . , J,
11.3 Further examples 317
−nj ε −nj
sup |λj g(z − nj ) − h(z)| < λ .
|z−nj |≤N 2 j
δ ε
sup |g(z) − (λT1 )nj h(z)| < M + = ε,
|z|≤N b 2
converges in X.
Then the set of common hypercyclic vectors for the family (Bw(λ) )λ∈Λ is a
dense Gδ -set, and in particular, nonempty.
It follows from assumption (iii) and the unconditionality of the basis that
∞
1
ek+n
n=1
wk+1 (a) · · · wk+n (a)
11.3 Further examples 319
Choose δ > 0 such that (et − 1)ek < ε whenever |t| < δ. By assumption
(ii) there is some L > 0 such that, for all ν ≥ 1 and λ, μ ∈ K,
w
k+n
k+1 (λ) · · · wk+n (λ)
log ≤ | log wν (λ) − log wν (μ)|
wk+1 (μ) · · · wk+n (μ)
ν=k+1
≤ nL(μ − λ) < δ,
so that Bw(λ)
n
Sn (μ)ek − ek < ε. Hence also condition (iii) of the Common
Hypercyclicity Criterion holds, which concludes the proof.
for λ > 0. Thus, assumption (iii) is satisfied for λ > 1/p for X = p , and
for all λ > 0 for X = c0 . Since the other assumptions hold as well, we have
common hypercyclic vectors for the respective families of operators.
(c) Consider the multiples λD, λ > 0, of the differentiation operator D on
the space H(C) of entire functions. By Example 4.9(b), D can be regarded
as a weighted shift with weights wn = n, n ≥ 1, on a suitable sequence space.
The proposition then implies easily that the operators λD, λ > 0, have a
common hypercyclic vector, confirming Example 11.17(a).
J−1
δ δ
J
≤ 2π < .
ν=1
3νN ν=1
3νN
We set
j
δ
θ0 = 0, θj = (0 < j < J), θJ = 2π
ν=1
3νN
and nj = 3jN for j = 1, . . . , J.
and, for j = 1, . . . , J,
sup g(z − nj eiθj ) − h(z) < ε .
|z−nj eiθj |≤ 5 N
2
4
iθ
nj e − nj eiθj ≤ nj |θ − θj | ≤ δ ≤ 1 N.
4
Now, if |z − nj eiθ | ≤ N then
z − nj eiθj ≤ z − nj eiθ + nj eiθ − nj eiθj ≤ N + δ ≤ 5 N.
4
Consequently we have that
sup g(z) − h z + nj eiθ ≤ sup g(z) − g z + nj eiθ − nj eiθj
|z|≤N |z|≤N
+ sup g z − nj eiθj − h(z)
|z−nj eiθ |≤N
In this section we combine the ideas of this chapter with those of Chapter 10.
Having established criteria for the existence of common hypercyclic vectors
one may wonder if a family of operators even has a common hypercyclic
subspace. Our first result might already come as a surprise: not even two
operators with hypercyclic subspaces need to share a hypercyclic subspace.
where L0 = ∅ and n0 = 0.
Denoting by en the functions en (z) = z n , n ≥ 1, and applying successively
Runge’s theorem we obtain, for any j ≥ 1, functions fj,k ∈ H(Ω), k ≥ 1,
such that
1 1
sup |ej (z) − fj,1 (z)| < , sup |fj,k (z) − fj,k+1 (z)| < ,
z∈L1 2j+2 z∈Lk+1 2j+k+2
1 1
sup |fj,2k−1 (z)| < , sup |fj,2k (z)| < .
z∈ϕn2k−1 (L2k−1 ) 2j+2k z∈ψn2k (L2k ) 2j+2k+1
l
S= · , yj∗ xj , yj∗ ∈ Φ, xj ∈ X0 ,
j=1
l
Mn (λ)S = · , yj∗ Sn (λ)xj ∈ K.
j=1
We then have that, for k ≥ 1, N ≥ 0 and any finite set F ⊂ {N, N + 1, . . .},
l
L m
Tλ M (μ
m−n n )S = Tλ
m
· , yj∗ Sm−n (μn )xj
k k
0≤n≤m 0≤n≤m j=1
n∈F n∈F
l
= sup pk x, yj∗ Tλm Sm−n (μn )xj
|x|≤1 j=1 0≤n≤m
n∈F
l
≤ sup |x, yj∗ | pk Tλm Sm−n (μn )xj ,
j=1 |x|≤1 0≤n≤m
n∈F
Combining this result with (the proof of) Theorem 11.13 we obtain the
following special case.
see also Proposition 8.5. Then the operators λT , |λ| > 1, have a common
hypercyclic subspace. To this end consider for X0 the set of finite sequences,
for Sn the maps Sn = S n with S(xn )n = (0, x1 , 0, x2 , 0, x3 , 0, . . .), x ∈ X, and
for M0 the subspace of all sequences x = (xn )n ∈ X whose even coordinates
vanish. Then the claim follows from Theorem 11.26 and Remark 11.27.
Exercises
Exercise 11.1.1. Consider the family of all weighted backward shifts Bw on 2 whose
weight sequence (wn )n only takes the values 1 and 2, with 2 appearing infinitely often.
Show that each of these operators is hypercyclic but that they do not possess a common
hypercyclic vector. (Hint: Construct, for any x ∈ 2 , a weight w such that w1 · · · wn |xn | ≤
1 for all sufficiently large n.)
326 11 Common hypercyclic vectors
Exercise 11.1.2. Let S and T be operators on a Banach space X and Λ ⊂ ]0, ∞[2 . Show
that if the family (λS ⊕ μT )(λ,μ)∈Λ of operators on X ⊕ X has a common hypercyclic
vector then Λ has two-dimensional Lebesgue measure 0. (Hint: First show as in Example
11.3 that the set {(log λ, log μ) ; (λ, μ) ∈ Λ} has Lebesgue measure 0.)
Exercise 11.1.4. Let B be the backward shift on one of the spaces p , 1 ≤ p < ∞, or c0 .
Show that, although the operators λB, λ > 1, have common hypercyclic vectors, no two
of them have the same hypercyclic vectors. More precisely, if λ, μ > 1 with λ = μ then
there is a vector that is λB-hypercyclic but not μB-hypercyclic. (Hint: Show that there
exists x ∈ HC(λB) with inf n≥1 μn |xn | > 0 (λ < μ) or supn≥1 μn |xn | < ∞ (λ > μ).)
Exercise 11.1.6. Let X be a separable Fréchet space, Λ a σ-compact metric space and
(Tλ )λ∈Λ a continuous family of operators on X. Show the following:
(i) if x ∈ X is a common hypercyclic vector then, for any compact subset K ⊂ Λ and
for any nonempty open subset V of X there are n1 , . . . , nJ ≥ 0 such that for any
n
λ ∈ K there is some j such that Tλ j x ∈ V ;
(ii) the set of common hypercyclic vectors is a dense Gδ -set if and only if, for any
compact set K ⊂ Λ and for any nonempty open subsets U and V of X, there are
x ∈ U and n1 , . . . , nJ ≥ 0 such that, for any λ ∈ K there is some j such that
n
Tλ j x ∈ V .
Moreover, show that the following condition is satisfied for any continuous family
(Tλ )λ∈Λ of hypercyclic operators and therefore does not characterize common hyper-
cyclicity:
(iii) for any compact set K ⊂ Λ and for any nonempty open subsets U and V of
X there are n1 , . . . , nJ ≥ 0 such that for any λ ∈ K there is some j such that
n
Tλ j (U ) ∩ V = ∅.
Show that the set of common hypercyclic vectors of the family (Tλ )λ∈Λ is a dense Gδ -
set, and in particular, nonempty. (Hint: Fix a partition K1 , . . . , Km of K of intervals
of length < δ; obtain U ⊃ U1 ⊃ . . . ⊃ Um and ν1 , . . . , νm such that T νj (Uj ) ⊂ V for
λ ∈ Kj .)
(b) Show that, for a singleton Λ, this criterion reduces to (a weak form of) the
Hypercyclicity Criterion.
(c) Show that the criterion cannot be applied in the canonical way to the family of
Rolewicz operators.
Exercise 11.2.1. Show that the family (λ(B ⊕ B))|λ|>1 = (λB ⊕ λB)|λ|>1 of operators
on p ⊕ p , 1 ≤ p < ∞, has a common hypercyclic vector. Compare the result with
Exercise 11.1.2.
Exercise 11.2.4. Let μ ∈ C with |μ| < 1. Consider the adjoint multiplier T = Mϕ∗ on
the Hardy space H 2 given by ϕ(z) = z − μ; see Section 4.4.
zn
n+1 , n ≥ 0. Show that X0 := span{gn ; n ≥ 0} is dense in H
2
(a) Let gn (z) = (1−μz)
and that T gn = gn−1 with g−1 = 0. (Hint: Show that
f, gn is a multiple of f (n) (μ).)
(b) Deduce that the operators λMϕ∗ , |λ| > 1−|μ|
1
, have common hypercyclic vectors.
(c) Deduce also that the operators λ(μI + B) on 2 , |λ| > 1
1−|μ| , have common
hypercyclic vectors, where B is the backward shift.
Exercise 11.2.5. Let ϕ be a nonconstant complex polynomial all of whose roots lie in D.
Show that the operators λMϕ∗ on H 2 , |λ| > max|z|=1 |ϕ(z)|
1
, have common hypercyclic
vectors. (Hint: Fix a root μ of ϕ; consider the set X0 of Exercise 11.2.4 and define
Sn : X0 → H 2 by Sn = Mψn , where ψn (z) = (1/ϕ∗ (1/z))n ; see Proposition 4.41. Use
Proposition 4.40.)
Exercise 11.2.8. The proof of the existence of entire functions that are hypercyclic for
all multiples λTa , λ > 0, of the Birkhoff operator Ta (see Example 11.18), can easily be
turned into a purely constructive proof. Explain.
Exercise 11.3.1. Show that Proposition 11.19 remains true when assumption (ii) is
weakened as follows: for any compact subinterval K ⊂ Λ, thefunctionsn log wn−1are
∞
Lipschitz continuous on K with Lipschitz constant Ln such that n=1
( L ) =
k=1 k
∞. Deduce that the weighted shifts Bw(λ) , λ > 0, with wn (λ) = nλ , n ≥ 1, have a
common hypercyclic vector on any of the spaces p , 1 ≤ p < ∞, or c0 . (Hint: Use
Remark 11.10(d).)
Exercise 11.3.4. Use the results of this section (and the León–Müller theorem) to
confirm the claim of Exercise 11.2.2.
Exercise 11.4.1. Show that the operators T1 and T2 of Example 11.22 have hypercyclic
subspaces.
Exercise 11.4.2. Explain how the construction in Example 11.24 can be modified in
order to show that any countable family of hypercyclic composition operators Cϕj ,
ϕj ∈ Aut(Ω), j ≥ 1, on a domain Ω ⊂ C possesses a common hypercyclic subspace.
n
n
Show that any countable family of operators in B has a common hypercyclic subspace.
Exercise 11.4.6. Show that, in the setting of Theorem 11.25 and its proof, (LTλ )λ∈Λ
is a continuous family on K. (Hint: Restrict the λ to a compact subset; show that the
corresponding Tλ are equicontinuous on X; use that F is dense in K).
Sources and comments 329
Exercise 11.4.7. Let X be one of the spaces p (Z), 1 ≤ p < ∞, or c0 (Z). Show that
the bilateral weighted backward shifts Bw(λ) , λ > 1, with weights
w(λ) = (. . . , λ1 , λ1 , λ1 , 2, 2, 2, . . .)
Exercise 11.4.9. Let a > 0, and let T be the operator on C0 (R+ ) given by T f (x) =
f (x + a); see Exercise 10.1.1. Show that the operators λT , λ > 1, have a common
hypercyclic subspace.
Exercise 11.4.10. Let Λ ⊂ R be an interval and (Tλ )λ∈Λ a continuous family of op-
erators on a separable Fréchet space X with a continuous norm. Show that if (Tλ )λ∈Λ
satisfies the Common Hypercyclicity Criterion then its set of common hypercyclic vec-
tors contains, except for 0, a dense subspace. (Hint: Use the ideas of Exercise 10.5.1.)
Section 11.1. The problem of the existence of common hypercyclic vectors for un-
countable families was first raised by Godefroy and Shapiro [165] when they asked if
the differential operators ϕ(D) on H(C) with nonconstant ϕ (see Section 4.2) share a
hypercyclic vector. However, this question was largely ignored, so that it was Salas [275]
who initiated the present study into common hypercyclicity by asking if the Rolewicz
operators λB, |λ| > 1, have a common hypercyclic vector. For real parameters, a positive
answer was given by Abakumov and Gordon [1] and, independently, by Peris [252]; the
general case is due to Costakis and Sambarino [123]; see Example 11.11.
Example 11.3 is due to Borichev (see [1, 42]). Theorem 11.5(a) is due to Saint Ray-
mond (see [1, 33]); the proof given here is taken from Costakis and Sambarino [123, p.
304]. Assertion (b) was obtained by Shkarin [288], while its equivalent formulation given
in Remark 11.6(b) is due to Chan and Sanders [102]. Remark 11.6(a) was observed by
Bayart [33].
The Common Hypercyclicity Criterion, under stronger assumptions, is due to Costakis
and Sambarino [123]. The present form was inspired by the Frequent Hypercyclicity Cri-
terion. An alternative criterion was found by Bayart and Matheron [42]. They use it to
show that, for the family of all automorphisms ϕ of D having 1 as an attractive fixed
point, the corresponding composition operators Cϕ on H 2 (see Section 4.5) share hyper-
cyclic vectors; see also Bayart and Grivaux [39]. By Bayart [33], this result breaks down
when the attractive fixed points are allowed to cover a set of positive Lebesgue measure
on T.
We have already commented on the difficulty of extending the criterion to more than
one-dimensional families of operators. Shkarin [288] confirmed that there is an intrinsic
obstruction to common hypercyclicity for higher-dimensional families by proving the
following remarkably general result.
For example, if D denotes the differentiation operator on H(C), then the operators
λI + μD, λ ∈ C, μ > 0, cannot have a common hypercyclic vector. This finally solved
the Godefroy–Shapiro problem in the negative.
Section 11.2. Proposition 11.12 is due to Bayart [33]. Theorem 11.13, Corollary 11.15
and Example 11.16 are due to Bayart and Matheron [42]; see also Bayart [33]. Remark
11.14 was made by Costakis and Mavroudis [120] and Bernal [59].
Common hypercyclic vectors for the multiples of MacLane’s operator (see Example
11.17) were found by Costakis and Sambarino [123]. This was extended to multiples
of certain operators ϕ(D) by Costakis and Mavroudis [120] and Bernal [59]. Finally,
Shkarin [288] proved that for any nonconstant entire function ϕ of exponential type the
differential operators λϕ(D), λ = 0, share hypercyclic vectors. This then also includes
Example 11.18.
Bayart [33] and Gallardo and Partington [160] have obtained common hypercyclic
vectors for multiples λMϕ∗ of certain adjoint multipliers on H 2 . Shkarin [288] proved
that, for any bounded holomorphic function ϕ, all hypercyclic multiples λMϕ∗ share
hypercyclic vectors.
The two mentioned results by Shkarin [288] are based on a powerful sufficient condi-
tion for common hypercyclicity for multiples of operators.
Section 11.3. Proposition 11.19 is a special case of a theorem by Bayart and Matheron
[42]. The first result in this direction is due to Costakis and Sambarino [123]; see Example
11.20(b). These authors also found Example 11.21; in fact, they gave a direct proof that
for parameters along rays starting from zero the sets of hypercyclic vectors are the same.
Chan and Sanders [102] showed that between any two hypercyclic weighted shifts
on 2 there is a continuous path of weighted shifts that shares hypercyclic vectors; and
there is another such path without any common hypercyclic vector.
Shkarin [288] studied common hypercyclicity for genuinely two- or higher-dimensional
families, that is, where one cannot reduce the number of dimensions by the León–
Müller theorem or the Conejero–Müller–Peris theorem. He was the first to obtain a
genuinely higher-dimensional family with common hypercyclic vectors: the multiples of
the Birkhoff operators f → λf (z + a), λ, a ∈ C \ {0}, on H(C) (note that one can reduce
the two complex dimensions to two real ones).
Section 11.4. Theorem 11.23 is due to Aron, Bès, León and Peris [14] where Example
11.22 can also be found. Theorem 11.25 improves on the corresponding result in Bayart
[35], Theorem 11.26 seems to be new. Common hypercyclic subspaces for some sequence
of operators on ω = KN were studied by Bès and Conejero [69].
As for the problem mentioned at the end of the section we recall that by a remarkable
result of Grivaux [169], any countable family of hypercyclic operators on a Banach space
has a common dense subspace of hypercyclic vectors, except for 0. See also Exercise
3.4.10.
Exercises. Exercise 11.1.1 is from Bayart and Matheron [42], Exercise 11.1.2 is due to
Borichev (see [42]), Exercise 11.1.5 is taken from Grivaux [169], and Exercises 11.1.6
and 11.1.7 from Chan and Sanders [102]. Exercise 11.2.1 is mentioned in Bayart and
Matheron [42], while Exercise 11.2.3 follows Costakis and Mavroudis [120] and Bernal
[59]. Exercises 11.3.1 and 11.3.2 are taken from Bayart and Matheron [42], Exercise
11.4.5 from Aron, Bès, León and Peris [14], and Exercise 11.4.10 partially improves a
result by Bayart [34].
Chapter 12
Linear dynamics in topological vector
spaces
+ : X × X → X, (x, y) → x + y,
· : K × X → X, (λ, x) → λx,
are continuous maps. We recall that a topology is Hausdorff if any two distinct
points in the space have disjoint neighbourhoods.
Many arguments in Banach and Fréchet spaces use the triangle inequality
of the norm or the seminorms. In general topological vector spaces, such
arguments are replaced by operations with 0-neighbourhoods.
A subset A of a vector space X is called balanced if λA ⊂ A whenever
λ ∈ K, |λ| ≤ 1.
Proposition 12.1. Let X be a topological vector space.
(a) A set U is a neighbourhood of a point x ∈ X if and only if there is a
0-neighbourhood W such that
x + W ⊂ U.
W 1 + W1 ⊂ W and W1 − W1 ⊂ W.
W1 ⊂ W ;
In view of Propositions 12.1(c) and 12.2(b), the set of all closed and
balanced 0-neighbourhoods of a topological vector space X is a base of 0-
neighbourhoods in X, which will be denoted by U0 (X).
There are some classes of topological vector spaces that deserve special
consideration.
To start with, any finite-dimensional topological vector space X is isomor-
phic to KN , for some N ≥ 0, where K is the scalar field of X; see Exer-
cise 12.1.2.
If a topological vector space X admits a countable base (Wn )n of 0-
neighbourhoods, then there is a translation-invariant metric d on X gen-
erating the topology of X. If, moreover, (X, d) is complete, then X is called
an F-space. Metrizable topological vector spaces are, thus, exactly the topo-
logical vector spaces admitting a countable base of 0-neighbourhoods, and
the completion X of a metrizable topological vector space is an F-space. See
also the related discussion in Section 2.1.
A topological vector space X whose topology is defined by a family of
seminorms is called a locally convex space; that is, X is locally convex if
there is a family (pα )α∈A of seminorms on X such that a subset W of X is
a 0-neighbourhood if and only if there are α1 , . . . , αn ∈ A, n ≥ 1, and ε > 0
such that
{x ∈ X ; pαk (x) < ε for k = 1, . . . , n} ⊂ W.
Fréchet spaces are, precisely, the locally convex F-spaces.
A subset A of a vector space X is called absolutely convex if, for any
x1 , x2 ∈ A and λ1 , λ2 ∈ K with |λ1 | + |λ2 | ≤ 1, the absolutely convex com-
bination λ1 x1 + λ2 x2 belongs to A. An easy observation nis that, if A ⊂ X is
absolutely convex, xk ∈ A, λk ∈ K, k = 1, . . . , n, with k=1 |λk | ≤ 1, then
n
λk xk ∈ A.
k=1
Example 12.3. In Sections 8.3 and 10.2 we considered the space L(X) of op-
erators on a Fréchet space X, endowed with the strong operator topology
(SOT). In this topology, a base of neighbourhoods of T ∈ L(X) is given by
The space ϕ has a natural locally convex topology, which is the strongest one
that can be defined on it; it is generated by the family of norms
∞
xv = |xn |vn , x ∈ ϕ,
n=1
1
vn = , n ∈ N,
min{|yn | ; y ∈ Fn }
Example 12.10. For a separable Banach space X, let us consider the space
L(X) of operators on X, endowed with the strong operator topology. In
Theorem 10.20 we proved that an operator T on X satisfies the Hypercyclicity
Criterion if, and only if, the left-multiplication operator LT : L(X) → L(X),
S → T S, is hypercyclic. This characterization provides a good collection of
hypercyclic operators on the non-metrizable locally convex space L(X).
We recall the useful result, Proposition 2.33, that the set Per(T ) of periodic
points for an operator T on a complex space X is given by
We next want to show that many fundamental results for hypercyclic op-
erators on Fréchet spaces extend to arbitrary topological vector spaces. For
this we need the notion of a quotient space.
Let X be a vector space and L ⊂ X a subspace. Defining x ∼ y if x−y ∈ L,
we obtain an equivalence relation on X. Let us denote by [x] = x + L the
equivalence class of x ∈ X, and by X/L the set of equivalence classes. Then
X/L inherits in a natural way a vector space structure, and we denote by
q : X → X/L, x → [x], the quotient map, which is linear and surjective.
If, now, X is a topological vector space and L ⊂ X is a closed subspace,
then X/L becomes a topological vector space, called the quotient space of X
modulo L, when endowed with the induced topology:
12.2 Hypercyclicity, topological transitivity, and linear chaos 337
⊂ X with q(U
U ⊂ X/L is open if and only if there is an open set U ) = U.
The quotient map q is then a continuous and open map. The requirement
that L be closed is necessary for the Hausdorff property of X/L; see Exer-
cise 12.2.5.
The following result, which generalizes Bourdon’s theorem, is the key to
the announced extensions.
Proof. We will only show the complex case. The real case can be deduced in
a similar way after some minor considerations; see Exercise 12.2.6.
As in the proof of Bourdon’s theorem it suffices to show that T − λI has
dense range for every λ ∈ C. Let L = (T − λI)(X), which is a closed subspace
of X, and suppose that L = X. We then consider the quotient space X/L,
which is nontrivial, and the quotient map q : X → X/L. Since, for any x ∈ X,
q((T − λI)x) = 0 we have that q(T x) = λq(x). Hence the operator S on X/L
given by S[x] = λ[x] is quasiconjugate to T via q and therefore inherits the
stated properties from T ; see the following section.
Under assumption (i), S is topologically transitive. On the other hand, let
[x] ∈ X/L, [x] = 0. By the Hausdorff property there is an open neighbourhood
U of [x] and a balanced 0-neighbourhood W such that U ∩ W = ∅. Now, if
|λ| ≥ 1, then W ⊂ λn W , and therefore S n (U ) ∩ W = ∅ for all n ∈ N0 . And
if |λ| < 1, then λn W ⊂ W , and therefore S n (W ) ∩ U = ∅ for all n ∈ N0 .
Thus, for any λ ∈ C, S is not topologically transitive, a contradiction.
Under assumption (ii), S has a somewhere dense orbit {λn [x] ; n ∈ N0 }.
Then span{[x]} = span{[x]} = X/L (see Exercises 12.1.1(v) and 12.1.2), so
that X/L is isomorphic to C. But every orbit {λn z ; n ∈ N0 }, z ∈ C, is
nowhere dense, a contradiction.
As a particular case, let 1 < p ≤ ∞, and consider the space p− := q<p q .
∞
Obviously, p− = n=1 pn for any strictly increasing sequence (pn )n in ]1, p[
tending to p. A natural topology on p− is the corresponding inductive limit
topology. If λ ∈ K is any scalar with |λ| > 1, then the multiple T = λB of the
backward shift satisfies the above requirements and is therefore hypercyclic
on p− .
The second method is a kind of converse to the first technique. Of course,
if, for a given operator T , all operators S that are quasiconjugate to T are
hypercyclic then T itself must be hypercyclic; one may simply take S =
T . It is, however, remarkable that the result remains true when we only
admit operators S defined on F-spaces. In addition, the map φ defining the
quasiconjugacy may be required to be linear.
12.3 Dynamical transference principles 339
We turn to the third transference principle. For this we need a new concept.
A projective spectrum X of Fréchet spaces consists of a family (Xα )α∈I of
Fréchet spaces, where I is a directed index set, and operators α β : Xβ → X α
for α ≤ β, called the spectral maps, that satisfy β ◦ γ = γ and α
α β α
α = IXα ,
the identity on Xα , for any α ≤ β ≤ γ. The projective limit of X is defined
as
proj X = (xα )α∈I ∈ Xα ; αβ xβ = xα for all α ≤ β ,
α∈I
endowed with the topology inherited from the product topology on α∈I Xα ;
in this way, proj X is a locally convex space. We denote by α : proj X → Xα
the projection onto the component with index α. It is not difficult to see that
the sets (α )−1 (Wα ), Wα ∈ U0 (Xα ), α ∈ I, form a base of 0-neighbourhoods
for the topology of proj X . We say that X is strongly reduced if for each α
there is a larger β such that αβ (Xβ ) is contained in the closure of (proj X )
α
in Xα .
Now, a family (Tα )α∈I of operators Tα on Xα is called an endomorphism
of X if their elements commute with the spectral maps in the sense that, for
any α ≤ β, Tα ◦ α β = β ◦ Tβ . The projective limit of the endomorphism is
α
Proof. (a) We will only show the result for the mixing property. Let x, y ∈
X := proj X and W0 ∈ U0 (X) be given. Then there are α ∈ I and
W1 ∈ U0 (Xα ) with W0 ⊃ (α )−1 (W1 ), and there is some β ≥ α with
αβ Xβ ⊂ (X). For each W ∈ U0 (Xα ) we obtain that β (Xβ ) ⊂ (X) +
α α α
α −1
W , and thus Xβ ⊂ (X) + (β ) (W ). This means that the image of
β
Un := (β x + (α
β)
−1
(W1 )) ∩ (Tβ−n (β y + (α
β)
−1
(W1 ))) = ∅
for all n ≥ N . Since Un is open with respect to τ , there are zn ∈ X such that
β zn ∈ Un for all n ≥ N . Then we have that α (zn − x) = α β ( zn − x) ∈
β β
β ( (T zn )) = β (Tβ ( zn )) ∈ y + W1 ,
α (T n zn ) = α β n α n β α
(x + W0 ) ∩ (T −n (y + W0 )) = ∅,
Taking the inclusion maps as spectral maps, we see that (Xn )n is also
∞ and proj X is the space of constant sequences with
a projective spectrum,
entries from X = n=1 Xn . It is then clear that proj X is isomorphic to X.
Moreover, the projective spectrum is strongly reduced if, for example, X is
dense in each Xn , n ≥ 1.
Proof. This follows directly from Proposition 12.18 because, by the assump-
tions, (Tn )n is an endomorphism of proj X and the projective limit of (Tn )n
on proj X turns into T via the identification of proj X with X.
In this section we will convince ourselves that the central results of Sections
2.4 and 2.5 remain true in general topological vector spaces when we replace
the assumption of hypercyclicity there by topological transitivity. We will
omit the proofs when the arguments given in those sections translate directly
to the general situation.
Example 12.22. We pointed out in Example 12.17 that there are mixing op-
erators on ϕ. Another easy example of a mixing operator is T = λB, |λ| > 1,
12.4 Mixing and weakly mixing operators 343
on the inductive limit p− for 1 < p ≤ ∞; see Example 12.14. Also, if X is
an arbitrary topological vector space, then the product space X N , endowed
with the product topology, is a topological vector space, and the backward
shift B : X N → X N , (x1 , x2 , . . . ) → (x2 , x3 , . . . ) is a mixing operator. Thus,
if I is an infinite set then the space X I , endowed with the product topology,
is isomorphic to (X I )N and therefore admits a mixing operator.
Theorem 12.24 also implies that Theorem 2.47 extends to general topo-
logical vector spaces.
Proof. In order to adapt the proof of Theorem 2.48 one need only note that,
for any x ∈ X0 and any 0-neighbourhood W , there is some ε > 0 such that
εT n x ∈ W for all n ∈ N0 .
∞
Recall that the generalized kernel of an operator T is given by n=0 ker T n .
Corollary 12.28. Let T be a topologically transitive operator on a topological
vector space X. If one of the following conditions is satisfied:
(i) T is chaotic;
(ii) T has a dense set of points for which the orbits converge;
(iii) T has dense generalized kernel;
then T is weakly mixing.
As a final application of Theorem 12.24 we will characterize weakly mixing
operators by the behaviour of multiples of iterates of T . For the notion of
topological transitivity for sequences of operators we refer to Section 1.6.
Theorem 12.29. Let T be an operator on a topological vector space X and
λ, μ ∈ K \ {0} with λ = μ. Then the following assertions are equivalent:
(i) T is weakly mixing;
(ii) for any M > δ > 0 and for any (λn )n with δ ≤ |λn | ≤ M , n ∈ N0 , the
sequence (λn T n )n is topologically transitive;
(iii) for any (λn )n with {λn ; n ∈ N0 } ⊂ {λ, μ}, the sequence (λn T n )n is
topologically transitive.
Proof. (i) =⇒ (ii). Given (λn )n with δ ≤ |λn | ≤ M , n ∈ N0 , we let U, V ⊂ X
be nonempty open sets. By Exercise 12.1.1 there are nonempty open sets U1
and V1 and a 0-neighbourhood W such that U1 +W ⊂ U and V1 +W ⊂ V . By
Proposition 12.1, if L = max( 1δ , M ), then there is a 0-neighbourhood W1 such
that λW1 ⊂ W , for any λ ∈ K with |λ| ≤ L. Now, since T is weakly mixing,
there are n ∈ N0 , u ∈ U1 and w ∈ W1 such that T n u ∈ W1 and T n w ∈ V1 .
Thus u + λ−1 n −1
n w ∈ U1 + W ⊂ U and λn T (u + λn w) = λn T u + T w ∈
n n
M (λ − μ) λ μ
U1 ⊂ W, U1 − U1 ⊂ M −1 W, and U1 − U1 ⊂ U.
λ λ−μ λ−μ
λ
α−1 λu1 ∈ U1 ⊂ W, T n (α−1 λu1 ) ∈ U,
M (λ − μ)
and
Example 12.32. Let X = Lp [0, 1], 0 < p < 1, be the space of p-integrable
functions on [0, 1]; see Example 12.4. Let ϕ : [0, 1] → [0, 1] be the invertible
function given by ϕ(t) = t/2 if t ∈ [0, 1/2], and ϕ(t) = (3/2)t − 1/2 if
t ∈ ]1/2, 1]. We then consider the composition operator Cϕ : X → X, Cϕ f =
f ◦ ϕ. The set
346 12 Linear dynamics in topological vector spaces
Exercises
Exercise 12.1.1. Let X be a topological vector space. Prove the following assertions:
(i) if x ∈ X and W is a 0-neighbourhood, then there exists some M > 0 and a
neighbourhood U of x such that U ⊂ M W ;
(ii) if U is a nonempty open set, then there is a 0-neighbourhood W and a nonempty
open set U1 ⊂ U such that U1 + W ⊂ U ;
(iii) if U is a nonempty open set, λ, μ ∈ K with λ + μ = 0, and x ∈ U , then there is a
neighbourhood U1 ⊂ U of x such that λU1 + μU1 ⊂ (λ + μ)U ;
(iv) for any λ ∈ K \ {0} and y ∈ X, the operators Mλ : X → X, x → λx, and
Ty : X → X, x → x + y, are homeomorphisms;
(v) if A ⊂ X is somewhere dense in X then span A is dense in X.
Exercise 12.1.2. Show that every finite-dimensional topological vector space X over
the field K = R or C is isomorphic to KN , where N is the dimension of X. Deduce that
finite-dimensional subspaces of topological vector spaces are closed. Here, as usual, an
isomorphism between two topological vector spaces is, by definition, a linear homeomor-
phism.
Exercise 12.1.3. Given 0 < p < 1 and a < b, show that the vector space
b
X = Lp [a, b] = f : [a, b] → K ; f is measurable and |f (t)|p dt < ∞
a
Exercises 347
Exercise 12.1.4. (a) Let X be a vector space. Show that a subset A of X is absolutely
convex if and only if it is convex and balanced.
(b) Show that a topological vector space is locally convex if, and only if, it has a base
of 0-neighbourhoods consisting of convex sets.
is a seminorm on span A.
∞ such that each
Exercise 12.1.7. Let (Xn )n be an increasing sequence of Banach spaces
inclusion map in : Xn → Xn+1 , n ≥ 1, is continuous, and set X = n=1 Xn . Show that
the inductive limit topology defined in Example 12.5 is a locally convex topology on X.
Prove that it is not metrizable unless there is some m ∈ N such that Xn = Xm for all
n ≥ m.
Exercise 12.2.1. The weak topology on a Banach space X is the locally convex topology
on X defined by the seminorms x → |x, x∗ |, x∗ ∈ X ∗ ; that is, it is the topology of
pointwise convergence on X ∗ . An operator T on a Banach space X is called weakly
hypercyclic if it is hypercyclic on X endowed with the weak topology.
Let X = p , 1 ≤ p < ∞, or X = c0 . Show that a weighted backward shift Bw on X
is hypercyclic if and only if it is weakly hypercyclic.
Exercise 12.2.2. Construct a chaotic and mixing operator that is not hypercyclic.
(Hint: Enlarge the space in Example 12.9.)
Exercise 12.2.4. A subset B of a topological vector space X is called bounded if, for any
W ∈ U0 (X), there is some M > 0 such that B ⊂ M W . If, in addition, B is absolutely
convex then XB := span B is a normed space when endowed with the gauge of B.
An operator T on X is called bounded if there is some U ∈ U0 (X) such that T (U ) is a
bounded subset of X. Show that a bounded operator T with dense range is hypercyclic
(or mixing, weakly mixing or chaotic) if and only if there is a bounded absolutely convex
set B ⊂ X such that XB is a T -invariant dense subspace of X and the induced operator
TB : XB → XB is hypercyclic (or mixing, weakly mixing or chaotic, respectively).
Moreover, if T is a bounded operator and λ ∈ K, show that there is some M > 0
such that M (λI + T ) is not hypercyclic.
Exercise 12.2.5. Let X be a topological vector space and L a subspace of X. Show that
X/L is Hausdorff if and only if L is closed. (Hint: Use some properties of Exercise 12.1.1.)
348 12 Linear dynamics in topological vector spaces
Exercise 12.3.2. Consider the weighted Banach space of holomorphic functions on the
unit disk,
Hvn (D) := f ∈ H(D) ; f := sup|f (z)|vn (z) < ∞ ,
z∈D
where vn (z) := (1 − |z|)n , n ∈ N; the inclusions Hvn (D) → Hvn+1 (D), n ≥ 1, are
continuous. The Korenblum space A−∞ is defined as the inductive limit of (Hvn (D))n .
Show that the differentiation operator D is a well-defined operator on A−∞ , and prove
that any finite-order differential operator on A−∞ that is not a multiple of the identity
is chaotic. In contrast, observe that D(Hvn (D)) ⊂ Hvn (D) for any n ∈ N, so that the
argument in Example 12.14 cannot be applied.
Exercise 12.3.4. Let X be a topological sequence space, that is, a topological vector
space X such that X ⊂ ω = KN with continuous inclusion. Suppose that ϕ is contained
and dense in X. If the weighted backward shift Bw is a well-defined operator on X,
prove that T := I + Bw is mixing on X.
Exercise 12.3.5. For this exercise we will need Young’s inequality: given any x ∈ p (Z)
and y ∈ q (Z), 1 ≤ p, q < 2, the convolution product
x ∗ y := xk yn−k ,
n∈Z
k∈Z
Observe that T ( ) ⊂ p for any p ≥ 1, which shows that the above result cannot be
p
transferred from the Banach spaces defining the projective spectrum of 1+ .
Exercise 12.4.2. Inspired by Theorems 1.54 and 12.29, prove that an operator T on a
topological vector space X is weakly mixing if and only if, for any M > δ > 0, for any
(λn )n with δ ≤ |λn | ≤ M , n ∈ N0 , and for any syndetic sequence (nk )k , the sequence
(λk T nk )k is topologically transitive.
Exercise 12.5.1. Let ϕ : [0, 1] → [0, 1] be a continuous, surjective, and strictly increas-
ing function such that ϕ(t) = t for all t ∈ ]0, 1[. Show that the composition operator Cϕ
is mixing on Lp [0, 1] for any p > 0.
Exercise 12.5.3. Let X be a Banach space with separable dual X ∗ and T an operator
on X whose adjoint T ∗ : X ∗ → X ∗ satisfies the Hypercyclicity Criterion (or is chaotic).
Show that the right-multiplication operator RT : L(X) → L(X), S → ST , is hypercyclic
and weakly mixing (or is hypercyclic and chaotic, respectively) on L(X), endowed with
the strong operator topology; see also Exercise 10.2.7.
Section 12.1. All the basic results of this section can be found in the books by Meise and
Vogt [237] and Rudin [271]. For F-spaces we also refer to Kalton, Peck and Roberts [212].
Section 12.3. Proposition 12.15 is new, while Corollary 12.16 is due to Shkarin [284]
and Bayart and Matheron [44].
Example 12.17 and Proposition 12.18 are due to Bonet, Frerick, Peris and Wengen-
roth [81]. Since ϕ is thus a topological vector space without hypercyclic operators, but
that admits a mixing, and therefore topologically transitive, operator, one might wonder
if every infinite-dimensional topological vector space necessarily admits a topologically
transitive operator. This is not the case, as Bermúdez and Kalton [52] have shown: there
are (non-separable) Banach spaces, like ∞ , or L(2 ) with the operator norm, without
any topologically transitive operators. Further existence and nonexistence results con-
cerning hypercyclic or topologically transitive operators on locally convex spaces beyond
Fréchet spaces are due to Bonet and Peris [85], Bonet, Frerick, Peris and Wengenroth [81],
and Shkarin [286, 288, 293].
Example 12.20 solves a problem from León, Piqueras and Seoane [224].
Section 12.4. For the results of this section we refer to Grosse-Erdmann and Peris [187].
Similar investigations can be found in Bayart and Matheron [44], [45] and Moothathu
[245].
Exercises. Exercise 12.2.1 is taken from Chan and Sanders [101], where the authors
also show that there are weakly hypercyclic bilateral shifts that are not hypercyclic.
Exercise 12.2.2 is taken from Bonet [78], Exercise 12.2.3 from Bonet and Peris [85] and
Martínez and Peris [229]. The first part of Exercise 12.2.4 is from Bonet and Peris [85].
The second part is extracted from Bonet [80]; this paper studies the open problem of
the existence of non-normable Fréchet spaces X such that every operator on X is of the
form λI + T with T a bounded operator. It is also asked whether, for every infinite-
dimensional separable non-normable Fréchet space X, there exists an operator T on X
such that λT is hypercyclic for any λ = 0. Exercise 12.2.6 is taken from Wengenroth
[301]. For Exercise 12.3.2 we refer to Bonet [78], for Exercises 12.3.3 and 12.3.4 to Bonet,
Frerick, Peris and Wengenroth [81], and for Exercise 12.3.5 to Frerick and Peris [155].
Exercise 12.5.3 is taken from Bonet, Martínez and Peris [84].
Appendix A – Prerequisites
Throughout this book we suppose that the reader is familiar with metric
spaces, the basics of Hilbert and Banach space theory, and the fundamentals
of complex analysis. An introduction to the theory of Fréchet spaces and
their operators is given in Section 2.1. Some more advanced results of these
theories will be provided here, with suitable references.
Metric spaces
Banach spaces
We suppose that the reader has had a first introduction to Banach spaces,
as can be found, for example, in Chapter 5 of Rudin [270], Chapter III of
Conway [116], or Chapters 2–5 of Bollobás [77].
If X, Y are two normed spaces then L(X, Y ) denotes the space of (contin-
uous, linear) operators T : X → Y ; under the operator norm T → T this
space turns into a Banach space whenever Y is a Banach space.
As a special case, L(X) = L(X, X) is the space of operators on X. The
operator norm then has the property that
ST ≤ ST
x∗ (x) = x, x∗ , x ∈ X.
Only the last assertion is not immediate; for a proof see [116, Chapter VI,
Proposition 1.4].
Example A.3. (a) The spaces p , 1 ≤ p < ∞, c0 and Lp [a, b] are recalled in
Example 2.4. The dual of p is given by q , where p1 + 1q = 1, in the sense
that the continuous linear functionals x∗ on p are precisely the maps of the
form
∞
x∗ (x) = x, x∗ = xn y n (A.1)
n=1
with y = (yn )n ∈ q , and we have that x∗ = y. In the same way the dual
of c0 is given by 1 .
Prerequisites 353
Under the representation (A.1), for p > 1, the dual of p (v) is given by q (w),
−q/p
where p1 + 1q = 1 and wn = vn , n ≥ 1; for p = 1, it is given by the space
of sequences (yn )n with supn≥1 |yn vn−1 | < ∞. This follows from the fact that
1/p
(xn )n → (xn vn )n defines an isometric isomorphism from p (v) to p .
(c) The analogous representation, via (A.2), holds for the duals of the
spaces Lpv (R+ ) introduced in Example 7.4; see also Exercise 2.1.6.
Hilbert spaces
We suppose that the reader is familiar with the basic properties of Hilbert
spaces as can be found, for example, in Chapter 4 of Rudin [270], Chapter II
of Conway [116], or Chapter 9 of Bollobás [77].
Let H be a Hilbert space with inner product · , · . We formulate two
consequences of the projection theorem.
Fréchet spaces
Section 2.1 gives a short introduction to the theory of Fréchet spaces. For
more detailed accounts we refer to Rudin [271] and Meise and Vogt [237]. We
collect here some further definitions and results that are used in this book;
their proofs can be found in the two books mentioned or in the references
provided.
The dual X ∗ of a Fréchet space X and the adjoint T ∗ of an operator T
on X are defined as in the case of Banach spaces. However we will not (need
to) address the problem of topologizing X ∗ , which is more delicate than for
Banach spaces; as a consequence we will consider T ∗ only as a linear map.
Prerequisites 355
qm (Tj x) ≤ M pn (x), x ∈ X.
We will mostly apply the Hahn–Banach theorem through one of the fol-
lowing corollaries:
(i) if p is a seminorm on X and x0 ∈ X then there exists a linear functional
u on X such that u(x0 ) = p(x0 ) and |u(x)| ≤ p(x) for all x ∈ X;
if X is a Fréchet space then
(ii) every continuous linear functional on a subspace of X extends to a con-
tinuous linear functional on X (with preservation of norm, if X is a
Banach space);
(iii) if M is a closed subspace of X and x ∈/ M then there exists a continuous
linear functional x∗ on X that vanishes on M with x, x∗ = 0;
(iv) a subspace M of X is dense in X if and only if every continuous linear
functional that vanishes on M also vanishes on X;
(v) for any x ∈ X, if x, x∗ = 0 for all x∗ ∈ X ∗ then x = 0.
A sequence (en )n in a Fréchet space X is called a basis if every x ∈ X has a
unique representation
∞
x= an e n
n=1
e∗n : X → K, x → an ,
converges.
Theorem A.16. Let X be a Fréchet space. Then the following assertions are
equivalent:
∞
(i) n=1 xn is unconditionally convergent;
∞
(ii) for any 0-1-sequence (εn )n , n=1 εn xn converges;
(iii) for any bounded sequence (αn )n of scalars, ∞ n=1 αn xn converges;
(iv) for any ε > 0 there is some N ∈ N such that for any finite set F ⊂
{N, N + 1, N + 2, . . .} we have that
xn < ε;
n∈F
Prerequisites 357
∞any ε > 0 there is some N ∈ N such that for any 0-1-sequence (εn )n ,
(v) for
n=1 εn xn converges and
εn xn < ε;
n≥N
converges unconditionally.
In connection with eigenvalue criteria for hypercyclicity, but also for Ap-
pendix B on spectral theory, we need the Riemann integral for Fréchet space-
valued continuous functions. Its definition poses no particular problems and
follows exactly the same lines as in the real-valued case. Thus, if f : [a, b] → X
is a continuous function with values in a Fréchet space X, then its Riemann
integral is defined as
b
N −1
f (t) dt = lim f (tk )(tk+1 − tk ),
a k=0
Complex analysis
Then f and g have the same number of zeros, counting multiplicity, for
|z − a| < r.
Let now D = {z ∈ C ; |z| < 1} be the open unit disk.
Theorem A.21 (Schwarz lemma). Let f : D → D be a holomorphic func-
tion with f (0) = 0. Then |f (z)| ≤ |z| for any z ∈ D. Moreover, if there is
some z ∈ D, z = 0, with |f (z)| = |z| then there is some a ∈ C with |a| = 1
such that f (z) = az, z ∈ D.
Theorem A.22 (Big Picard theorem). Let f be an entire function that is
not a polynomial. Then f takes every value in C, with at most one exception,
infinitely often.
The following is a consequence of Jensen’s formula; see Rudin [270, 15.20] or
Conway [115, Chapter XI, § 1].
Theorem A.23. Let f be an entire function, and let N (r) denote the number
of zeros of f in |z| < r, counting multiplicity. If f (0) = 1 and M (r) =
max{|f (z)| ; |z| = r} then
The Runge approximation theorem is a crucial tool for several of our exam-
we denote the extended complex plane C ∪ {∞}.
ples. By C
Theorem A.24 (Runge’s theorem). Let K ⊂ C be a compact set and A a
\ K.
set that contains at least one point from each connected component of C
Let f be a function that is holomorphic on some neighbourhood of K, and
let ε > 0. Then there exists a rational function h with poles only at points
from A such that
sup |f (z) − h(z)| < ε.
z∈K
We note that the complement of a compact set is open and therefore has
at most a countable number of connected components. Thus, A may always
be chosen to be at most countable.
\ K is connected then A may be taken
Moreover, if K ⊂ Ω is such that C
to be {∞}, and the function h will a polynomial.
Corollary A.25. Let Ω ⊂ C be a domain.
(a) Let A ⊂ C contain at least one point from each bounded component of
C \ Ω. Then the rational functions with poles only in A form a dense set in
H(Ω).
360 Appendix A
σ(λI + μT ) = λ + μσ(T )
for any λ, μ ∈ C, which is a special case of the spectral mapping theorem; see
below.
Proposition B.2. The spectrum σ(T ) is a nonempty compact set. Moreover,
|λ| ≤ T for any λ ∈ σ(T ).
Theorem B.3 (Spectral radius formula). For the spectral radius we have
that
r(T ) = lim T n 1/n .
n→∞
The existence of the limit is part of the result. In particular, we have that
σ(T ) = {0} if and only if limn→∞ T n 1/n = 0.
For the adjoint T ∗ : X ∗ → X ∗ of T we refer to Appendix A.
N N
(vi) if p(z) = n=0 an z n is a polynomial then p(T ) = n=0 an T n ;
(vii) if f (z) = ∞ n
n=0 an z is holomorphic on {z ∈ C ; |z| < r} for some
∞
r > T then f (T ) = n=0 an T n .
The following is a central result of spectral theory.
Theorem B.6 (Spectral mapping theorem). Let f be a holomorphic
function on a neighbourhood of σ(T ). Then
There is also a version for the point spectrum. Since the result is more
difficult to find in the literature we give a proof here.
Theorem B.7 (Point spectral mapping theorem). Let f be a holomor-
phic function on an open neighbourhood O of σ(T ) that is not constant on
any connected component of O. Then
σp (f (T )) = f (σp (T )).
Proof. First suppose that λ ∈ σp (T ). Since the function z → f (λ) − f (z) has
a zero at λ, it factorizes as f (λ) − f (z) = (λ − z)g(z), where g is holomorphic
on some neighbourhood of σ(T ). Hence
n
μI − f (T ) = (λk I − T )Nk g(T ).
k=1
σ(T ) = σ1 ∪ σ2 .
Proof. Since σ(T ) is compact, the sets σ1 and σ2 are disjoint compact subsets
of C, so that there exist disjoint open neighbourhoods O1 of σ1 and O2 of σ2 .
The function f1 that takes the value 1 on O1 and the value 0 on O2 is then
holomorphic on a neighbourhood of σ(T ), as is the function f2 that takes the
value 0 on O1 and the value 1 on O2 . We can therefore define two operators
by
P1 = f1 (T ) and P2 = f2 (T ).
Since the functions f1 and f2 have the properties
f1 + f2 = 1, f1 f2 = f2 f1 = 0, f12 = f1 , f22 = f2 ,
We define
M1 = ran P1 and M2 = ran P2
and claim that these are the desired subspaces.
First, (B.1) implies that
Moreover, let λ ∈
/ σ1 . Then the function
For the Riesz decomposition theorem we have followed Radjavi and Rosen-
thal [264].
References
75. G.D. Birkhoff, Démonstration d’un théorème élémentaire sur les fonctions entières,
C. R. Acad. Sci. Paris 189 (1929), 473–475.
76. O. Blasco, A. Bonilla and K.-G. Grosse-Erdmann, Rate of growth of frequently
hypercyclic functions, Proc. Edinb. Math. Soc. (2) 53 (2010), 39–59.
77. B. Bollobás, Linear analysis, second edition, Cambridge University Press, Cam-
bridge, 1999.
78. J. Bonet, Hypercyclic and chaotic convolution operators, J. London Math. Soc. (2)
62 (2000), 253–262.
79. J. Bonet, Dynamics of the differentiation operator on weighted spaces of entire
functions, Math. Z. 261 (2009), 649–657.
80. J. Bonet, A problem on the structure of Fréchet spaces, Rev. R. Acad. Cienc.
Exactas Fís. Nat. Ser. A Mat. RACSAM 104 (2010), 427–434.
81. J. Bonet, L. Frerick, A. Peris and J. Wengenroth, Transitive and hypercyclic op-
erators on locally convex spaces, Bull. London Math. Soc. 37 (2005), 254–264.
82. J. Bonet, F. Martínez-Giménez and A. Peris, A Banach space which admits no
chaotic operator, Bull. London Math. Soc. 33 (2001), 196–198.
83. J. Bonet, F. Martínez-Giménez and A. Peris, Linear chaos on Fréchet spaces, In-
ternat. J. Bifur. Chaos Appl. Sci. Engrg. 13 (2003), 1649–1655.
84. J. Bonet, F. Martínez-Giménez and A. Peris, Universal and chaotic multipliers on
spaces of operators, J. Math. Anal. Appl. 297 (2004), 599–611.
85. J. Bonet and A. Peris, Hypercyclic operators on non-normable Fréchet spaces, J.
Funct. Anal. 159 (1998), 587–595.
86. A. Bonilla and K.-G. Grosse-Erdmann, On a theorem of Godefroy and Shapiro,
Integral Equations Operator Theory 56 (2006), 151–162.
87. A. Bonilla and K.-G. Grosse-Erdmann, Frequently hypercyclic operators and vec-
tors, Ergodic Theory Dynam. Systems 27 (2007), 383–404. Erratum: Ergodic Theory
Dynam. Systems 29 (2009), 1993–1994.
88. A. Bonilla and K.-G. Grosse-Erdmann, Frequently hypercyclic subspaces, preprint.
89. A. Bonilla and P. Miana, Hypercyclic and topologically mixing cosine functions on
Banach spaces, Proc. Amer. Math. Soc. 136 (2008), 519–528.
90. P.S. Bourdon, Invariant manifolds of hypercyclic vectors, Proc. Amer. Math. Soc.
118 (1993), 845–847.
91. P.S. Bourdon, The second iterate of a map with dense orbit, Proc. Amer. Math.
Soc. 124 (1996), 1577–1581.
92. P.S. Bourdon, Orbits of hyponormal operators, Michigan Math. J. 44 (1997), 345–
353.
93. P.S. Bourdon and N.S. Feldman, Somewhere dense orbits are everywhere dense,
Indiana Univ. Math. J. 52 (2003), 811–819.
94. P.S. Bourdon and J.H. Shapiro, Cyclic composition operators on H 2 , Operator
Theory: Operator Algebras and Applications, Part 2 (Proc. Summer Res. Inst.,
Durham, NH, 1988), 43–53, Amer. Math. Soc., Providence, RI, 1990.
95. P.S. Bourdon and J.H. Shapiro, Cyclic phenomena for composition operators, Mem.
Amer. Math. Soc. 125 (1997), no. 596.
96. P.S. Bourdon and J.H. Shapiro, Hypercyclic operators that commute with the
Bergman backward shift, Trans. Amer. Math. Soc. 352 (2000), 5293–5316.
97. M. Brin and G. Stuck, Introduction to dynamical systems, Cambridge University
Press, Cambridge, 2002.
98. L. Carleson and T.W. Gamelin, Complex dynamics, Springer, New York, 1993.
99. K.C. Chan, Hypercyclicity of the operator algebra for a separable Hilbert space,
J. Operator Theory 42 (1999), 231–244.
100. K.C. Chan, The density of hypercyclic operators on a Hilbert space, J. Operator
Theory 47 (2002), 131–143.
101. K.C. Chan and R. Sanders, A weakly hypercyclic operator that is not norm hy-
percyclic, J. Operator Theory 52 (2004), 39–59.
References 371
102. K.C. Chan and R. Sanders, Two criteria for a path of operators to have common
hypercyclic vectors, J. Operator Theory 61 (2009), 191–223.
103. K.C. Chan and R. Sanders, Common hypercyclic vectors for the conjugate class of
a hypercyclic operator, J. Math. Anal. Appl. 375 (2011), 139–148.
104. K.C. Chan and I. Seceleanu, Orbital limit points and hypercyclicity of operators
on analytic function spaces, Math. Proc. R. Ir. Acad. 110A (2010), 99–109.
105. K.C. Chan and I. Seceleanu, Hypercyclicity of shifts as a zero-one law of orbital
limit points, J. Operator Theory, to appear.
106. K.C. Chan and J.H. Shapiro, The cyclic behavior of translation operators on
Hilbert spaces of entire functions, Indiana Univ. Math. J. 40 (1991), 1421–1449.
107. K.C. Chan and R.D. Taylor, Jr., Hypercyclic subspaces of a Banach space, Integral
Equations Operator Theory 41 (2001), 381–388.
108. J.A. Conejero, Operadores y semigrupos de operadores en espacios de Fréchet y es-
pacios localmente convexos, Thesis, Univ. Politécnica de Valencia, Valencia, 2004.
109. J.A. Conejero and E.M. Mangino, Hypercyclic semigroups generated by Ornstein–
Uhlenbeck operators, Mediterr. J. Math. 7 (2010), 101–109.
110. J.A. Conejero, V. Müller and A. Peris, Hypercyclic behaviour of operators in a
hypercyclic C0 -semigroup, J. Funct. Anal. 244 (2007), 342–348.
111. J.A. Conejero and A. Peris, Linear transitivity criteria, Topology Appl. 153 (2005),
767–773.
112. J.A. Conejero and A. Peris, Chaotic translation semigroups, Discrete Contin. Dyn.
Syst. 2007, Dynamical Systems and Differential Equations (Proc. Conf., Poitiers,
2006), suppl., 269–276.
113. J.A. Conejero and A. Peris, Hypercyclic translation C0 -semigroups on complex
sectors, Discrete Contin. Dyn. Syst. 25 (2009), 1195–1208.
114. J.A. Conejero, A. Peris and M. Trujillo, Chaotic asymptotic behaviour of the hyper-
bolic heat transfer equation solutions, Internat. J. Bifur. Chaos Appl. Sci. Engrg.
20 (2010), 2943–2947.
115. J.B. Conway, Functions of one complex variable, second edition, Springer, New
York-Berlin, 1978.
116. J.B. Conway, A course in functional analysis, second edition, Springer, New York-
Berlin, 1990.
117. G. Costakis, On a conjecture of D. Herrero concerning hypercyclic operators, C.
R. Acad. Sci. Paris Sér. I Math. 330 (2000), 179–182.
118. G. Costakis and A. Manoussos, J-class weighted shifts on the space of bounded
sequences of complex numbers, Integral Equations Operator Theory 62 (2008), 149–
158.
119. G. Costakis and A. Manoussos, J-class operators and hypercyclicity, J. Operator
Theory, to appear.
120. G. Costakis and P. Mavroudis, Common hypercyclic entire functions for multiples
of differential operators, Colloq. Math. 111 (2008), 199–203.
121. G. Costakis and A. Peris, Hypercyclic semigroups and somewhere dense orbits, C.
R. Math. Acad. Sci. Paris 335 (2002), 895–898.
122. G. Costakis and I.Z. Ruzsa, Frequently Cesàro hypercylic operators are hypercyclic,
preprint.
123. G. Costakis and M. Sambarino, Genericity of wild holomorphic functions and com-
mon hypercyclic vectors, Adv. Math. 182 (2004), 278–306.
124. G. Costakis and M. Sambarino, Topologically mixing hypercyclic operators, Proc.
Amer. Math. Soc. 132 (2004), 385–389.
125. C.C. Cowen and B.D. MacCluer, Composition operators on spaces of analytic func-
tions, CRC Press, Boca Raton, FL, 1995.
126. M. De la Rosa and C. Read, A hypercyclic operator whose direct sum T ⊕ T is not
hypercyclic, J. Operator Th. 61 (2009), 369–380.
127. M. De la Rosa, L. Frerick, S. Grivaux and A. Peris, Frequent hypercyclicity, chaos,
and unconditional Schauder decompositions, Israel J. Math., to appear.
372 References
128. R. deLaubenfels and H. Emamirad, Chaos for functions of discrete and continuous
weighted shift operators, Ergodic Theory Dynam. Systems 21 (2001), 1411–1427.
129. W. Desch and W. Schappacher, On products of hypercyclic semigroups, Semigroup
Forum 71 (2005), 301–311.
130. W. Desch and W. Schappacher, Spectral characterization of weak topological tran-
sitivity, Rev. R. Acad. Cienc. Exactas Fís. Nat. Ser. A Mat. RACSAM, to appear.
131. W. Desch, W. Schappacher and G.F. Webb, Hypercyclic and chaotic semigroups
of linear operators, Ergodic Theory Dynam. Systems 17 (1997), 793–819.
132. L.R. Devaney, An introduction to chaotic dynamical systems, Benjamin/Cum-
mings, Menlo Park, CA, 1986; second edition, Addison-Wesley, Redwood City,
CA, 1989.
133. J. Diestel, Sequences and series in Banach spaces, Springer, New York, 1984.
134. J. Diestel, H. Jarchow and A. Tonge, Absolutely summing operators, Cambridge
University Press, Cambridge, 1995.
135. S.J. Dilworth and V.G. Troitsky, Spectrum of a weakly hypercyclic operator meets
the unit circle, Trends in Banach spaces and operator theory (Proc. Conf., Mem-
phis, TN, 2001), 67–69, Amer. Math. Soc., Providence, RI, 2003.
136. P.L. Duren, Theory of Hp spaces, Academic Press, New York-London, 1970.
137. S.M. Duyos-Ruiz, On the existence of universal functions, Soviet Math. Dokl. 27
(1983), 9–13.
138. O.P. Dzagnidze, The universal harmonic function in the space En (Russian), Soob-
shch. Akad. Nauk Gruzin. SSR 55 (1969), 41–44.
139. S. El Mourchid, On a hypercyclicity criterion for strongly continuous semigroups,
Discrete Contin. Dyn. Syst. 13 (2005), 271–275.
140. S. El Mourchid, The imaginary point spectrum and hypercyclicity, Semigroup Fo-
rum 73 (2006), 313–316.
141. H. Emamirad, G.R. Goldstein and J.A. Goldstein, Chaotic solution for the Black–
Scholes equation, preprint.
142. P. Enflo, On the invariant subspace problem for Banach spaces, Acta Math. 158
(1987), 213–313.
143. K.-J. Engel and R. Nagel, One-parameter semigroups for linear evolution equations,
Springer, New York-Berlin, 2000.
144. K.-J. Engel and R. Nagel, A short course on operator semigroups, Springer, New
York, 2006.
145. J. Esterle, Mittag-Leffler methods in the theory of Banach algebras and a new
approach to Michael’s problem, Banach algebras and several complex variables
(Proc. Conf., New Haven, CT, 1983), 107–129, American Mathematical Society,
Providence, RI, 1984.
146. A. Fathi, Existence de systèmes dynamiques minimaux sur l’espace de Hilbert
séparable, Topology 22 (1983), 165–167.
147. N.S. Feldman, Linear chaos?, <http://home.wlu.edu/~feldmann/research.html>,
2001.
148. N.S. Feldman, Perturbations of hypercyclic vectors, J. Math. Anal. Appl. 273
(2002), 67–74.
149. N.S. Feldman, Countably hypercyclic operators, J. Operator Theory 50 (2003),
107–117.
150. N.S. Feldman, Hypercyclicity and supercyclicity for invertible bilateral weighted
shifts, Proc. Amer. Math. Soc. 131 (2003), 479–485.
151. N.S. Feldman, V.G. Miller and T.L. Miller, Hypercyclic and supercyclic cohyponor-
mal operators, Acta Sci. Math. (Szeged) 68 (2002), 965–990.
152. E. Flytzanis, Mixing properties of linear operators in Hilbert spaces, Séminaire
d’Initiation à l’Analyse 34ème année (1994/1995), Exposé no. 6.
153. E. Flytzanis, Unimodular eigenvalues and linear chaos in Hilbert spaces, Geom.
Funct. Anal. 5 (1995), 1–13.
References 373
154. G.L. Forti, Various notions of chaos for discrete dynamical systems. A brief survey,
Aequationes Math. 70 (2005), 1–13.
155. L. Frerick and A. Peris, Hypercyclic operators for which every non-zero multiple
is also hypercyclic, preprint.
156. H. Furstenberg, The structure of distal flows, Amer. J. Math. 85 (1963) 477–515.
157. H. Furstenberg, Disjointness in ergodic theory, minimal sets, and a problem in
Diophantine approximation, Math. Systems Theory 1 (1967), 1–49.
158. E.A. Gallardo-Gutiérrez and A. Montes-Rodríguez, The role of the spectrum in
the cyclic behavior of composition operators, Mem. Amer. Math. Soc. 167 (2004),
no. 791.
159. E.A. Gallardo-Gutiérrez and A. Montes-Rodríguez, The Volterra operator is not
supercyclic, Integral Equations Operator Theory 50 (2004), 211–216.
160. E.A. Gallardo-Gutiérrez and J.R. Partington, Common hypercyclic vectors for fam-
ilies of operators, Proc. Amer. Math. Soc. 136 (2008), 119–126.
161. R.M. Gethner and J.H. Shapiro, Universal vectors for operators on spaces of holo-
morphic functions, Proc. Amer. Math. Soc. 100 (1987), 281–288.
162. E. Glasner, Ergodic theory via joinings, American Mathematical Society, Provi-
dence, RI, 2003.
163. E. Glasner and B. Weiss, Sensitive dependence on initial conditions, Nonlinearity
6 (1993), 1067–1075.
164. G. Godefroy, Linear dynamics, Advanced courses of mathematical analysis. II
(Proc. 2nd Int. School, Granada, 2004), 57–75, World Sci. Publ., Hackensack, NJ,
2007.
165. G. Godefroy and J.H. Shapiro, Operators with dense, invariant, cyclic vector man-
ifolds, J. Funct. Anal. 98 (1991), 229–269.
166. M.C. Gómez-Collado, F. Martínez-Giménez, A. Peris and F. Rodenas, Slow growth
for universal harmonic functions, J. Inequal. Appl. 2010, Art. ID 253690, 6 pp.
167. M. González, F. León-Saavedra and A. Montes-Rodríguez, Semi-Fredholm theory:
hypercyclic and supercyclic subspaces, Proc. London Math. Soc. (3) 81 (2000),
169–189.
168. P. Gorkin, F. León-Saavedra and R. Mortini, Bounded universal functions in one
and several complex variables, Math. Z. 258 (2008), 745–762.
169. S. Grivaux, Construction of operators with prescribed behaviour, Arch. Math.
(Basel) 81 (2003), 291–299.
170. S. Grivaux, Sums of hypercyclic operators, J. Funct. Anal. 202 (2003), 486–503.
171. S. Grivaux, Hypercyclic operators with an infinite dimensional closed subspace of
periodic points, Rev. Mat. Complut. 16 (2003), 383–390.
172. S. Grivaux, Hypercyclic operators, mixing operators, and the bounded steps prob-
lem, J. Operator Theory 54 (2005), 147–168.
173. S. Grivaux, A probabilistic version of the frequent hypercyclicity criterion, Studia
Math. 176 (2006), 279–290.
174. S. Grivaux, A new class of frequently hypercyclic operators, with applications,
Indiana Univ. Math. J., to appear.
175. S. Grivaux, A hypercyclic rank one perturbation of a unitary operator, preprint.
176. S. Grivaux and S. Shkarin, Non-mixing hypercyclic operators, unpublished (2007).
177. K.-G. Grosse-Erdmann, Holomorphe Monster und universelle Funktionen, Mitt.
Math. Sem. Giessen 176 (1987).
178. K.-G. Grosse-Erdmann, On the universal functions of G.R. MacLane, Complex
Variables Theory Appl. 15 (1990), 193–196.
179. K.-G. Grosse-Erdmann, Universal families and hypercyclic operators, Bull. Amer.
Math. Soc. (N.S.) 36 (1999), 345–381.
180. K.-G. Grosse-Erdmann, Hypercyclic and chaotic weighted shifts, Studia Math. 139
(2000), 47–68.
181. K.-G. Grosse-Erdmann, Rate of growth of hypercyclic entire functions, Indag.
Math. (N.S.) 11 (2000), 561–571.
374 References
263. V. Protopopescu and Y.Y. Azmy, Topological chaos for a class of linear models,
Math. Models Methods Appl. Sci. 2 (1992), 79–90.
264. H. Radjavi and P. Rosenthal, Invariant subspaces, second edition, Dover Publica-
tions, Mineola, NY, 2003.
265. C.J. Read, A solution to the invariant subspace problem, Bull. London Math. Soc.
16 (1984), 337–401.
266. C.J. Read, The invariant subspace problem for a class of Banach spaces. II. Hy-
percyclic operators, Israel J. Math. 63 (1988), 1–40.
267. C. Robinson, Dynamical systems. Stability, symbolic dynamics, and chaos, second
edition, CRC Press, Boca Raton, FL, 1999.
268. S. Rolewicz, On orbits of elements, Studia Math. 32 (1969), 17–22.
269. S. Rolewicz, Metric linear spaces, second edition, D. Reidel Publishing Co., Dor-
drecht, 1985.
270. W. Rudin, Real and complex analysis, third edition, McGraw-Hill, New York, 1987.
271. W. Rudin, Functional analysis, second edition, McGraw-Hill, New York, 1991.
272. R. Rudnicki, Gaussian measure-preserving linear transformations, Univ. Iagel. Acta
Math. 30 (1993), 105–112.
273. H. Salas, A hypercyclic operator whose adjoint is also hypercyclic, Proc. Amer.
Math. Soc. 112 (1991), 765–770.
274. H.N. Salas, Hypercyclic weighted shifts, Trans. Amer. Math. Soc. 347 (1995), 993–
1004.
275. H.N. Salas, Supercyclicity and weighted shifts, Studia Math. 135 (1999), 55–74.
276. H.N. Salas, Banach spaces with separable duals support dual hypercyclic operators,
Glasg. Math. J. 49 (2007), 281–290.
277. H.N. Salas, Dual disjoint hypercyclic operators, J. Math. Anal. Appl. 374 (2011),
106–117.
278. W. Seidel and J.L. Walsh, On approximation by Euclidean and non-Euclidean
translations of an analytic function, Bull. Amer. Math. Soc. 47 (1941), 916–920.
279. J.H. Shapiro, Composition operators and classical function theory, Springer, New
York, 1993.
280. J.H. Shapiro, Simple connectivity and linear chaos, Rend. Circ. Mat. Palermo (2)
Suppl. (1998), no. 56, 27–48.
281. J.H. Shapiro, Notes on the dynamics of linear operators,
<http://www.mth.msu.edu/~shapiro>, 2001.
282. S. Shirali and H.L. Vasudeva, Metric spaces, Springer, London, 2006.
283. S.A. Shkarin, On the growth of D-universal functions, Moscow Univ. Math. Bull.
48 (1993), no. 6, 49–51.
284. S. Shkarin, Universal elements for non-linear operators and their applications, J.
Math. Anal. Appl. 348 (2008), 193–210.
285. S. Shkarin, The Kitai criterion and backward shifts, Proc. Amer. Math. Soc. 136
(2008), 1659–1670.
286. S. Shkarin, Existence theorems in linear chaos, arXiv:0810.1192v2 (2008).
287. S. Shkarin, On the spectrum of frequently hypercyclic operators, Proc. Amer. Math.
Soc. 137 (2009), 123–134.
288. S. Shkarin, Remarks on common hypercyclic vectors, J. Funct. Anal. 258 (2010),
132–160.
289. S. Shkarin, A hypercyclic finite rank perturbation of a unitary operator, Math.
Ann. 348 (2010), 379–393.
290. S. Shkarin, On the set of hypercyclic vectors for the differentiation operator, Israel
J. Math. 180 (2010), 271–283.
291. S. Shkarin, A short proof of existence of disjoint hypercyclic operators, J. Math.
Anal. Appl. 367 (2010), 713–715.
292. S. Shkarin, Chaotic Banach algebras, preprint.
293. S. Shkarin, Hypercyclic operators on topological vector spaces, preprint.
378 References
294. S. Silverman, On maps with dense orbits and the definition of chaos, Rocky Moun-
tain J. Math. 22 (1992), 353–375.
295. L.A. Smith, Chaos: A very short introduction, Oxford University Press, Oxford,
2007.
296. C.L. Stewart and R. Tijdeman, On infinite-difference sets, Canad. J. Math. 31
(1979), 897–910.
297. F. Takeo, Chaos and hypercyclicity for solution semigroups to some partial differ-
ential equations, Nonlinear Anal. 63 (2005), e1943–e1953.
298. M. Taniguchi, Chaotic composition operators on the classical holomorphic spaces,
Complex Var. Theory Appl. 49 (2004), 529–538.
299. P. Touhey, Yet another definition of chaos, Amer. Math. Monthly 104 (1997), 411–
414.
300. P. Walters, An introduction to ergodic theory, Springer, New York-Berlin, 1982.
301. J. Wengenroth, Hypercyclic operators on non-locally convex spaces, Proc. Amer.
Math. Soc. 131 (2003), 1759–1761.
302. P.Y. Wu, Sums and products of cyclic operators, Proc. Amer. Math. Soc. 122
(1994), 1053–1063.
303. B. Yousefi and H. Rezaei, Conditions for hypercyclicity criterion, Int. J. Contemp.
Math. Sci. 1 (2006), 99–107.
304. N. Zorboska, Cyclic composition operators on smooth weighted Hardy spaces,
Rocky Mountain J. Math. 29 (1999), 725–740.
Index
A2 , 93 Mϕ , 120
A2α , 132 Mϕ∗ , 120
B, 36, 98 N (A, B), 19, 24
Bw , 89, 99 Per((Tt )), 186
C∗ , 112 Per(T ), 12
, 359
C Sν , 133
C[a, b], 33 S ⊕ T , 36
C ∞ (R), 60 S × T , 17
C0 (R+ ), 60, 188 T, 5
Cϕ , 123 T, 41
C0,v (R+ ), 206 T ∗ , 352, 354
D, 70 Ta , 36
D, 130 U0 (X), 333
Dα , 132 V , 139
D, 36 41
X,
Dw , 218 X ⊕ Y , 36
Eτ2 , 129 X × Y , 17
E p (γ), 228 X ∗
, 352
F , 92, 100 ∞
( n=1 Xn )p , 47
F HC(T ), 237 ∞
Fw , 98, 102 ( n=1 Xn )c0 , 47
H(C), 32 · , · , 352, 353
HC(T ), 37 σ(T ), 361
HC(Ψ ), 168 σp (T ), 361
H 2 , 118 ϕ, 335
H 2 (β), 133 ϕ(D), 105
I, 4, 352 ω, 32
J(x), 26 ω(Z), 99
J mix (x), 27 c0 , 33
K, 32 c0 (v), 93
K(X), 352 dens(A), 236
K(X, Y ), 352 ∞ , 33
L(X), 352 p , 33
L(X, Y ), 352 p (Z), 33
Lp (T), 263 p (v), 353
Lp [a, b], 33, 334 orb(x, (Tn )), 23
Lpv (R+ ), 183 orb(x, (Tt )), 186