Hostname: page-component-5b777bbd6c-sbgtn Total loading time: 0 Render date: 2025-06-18T08:34:52.810Z Has data issue: false hasContentIssue false

Hereditarily frequently hypercyclic operators and disjoint frequent hypercyclicity

Published online by Cambridge University Press:  16 April 2025

FRÉDÉRIC BAYART
Affiliation:
Laboratoire de Mathématiques Blaise Pascal UMR 6620 CNRS, Université Clermont Auvergne, Campus universitaire des Cézeaux, 3 place Vasarely, 63178 Aubière Cedex, France (e-mail: frederic.bayart@uca.fr)
SOPHIE GRIVAUX
Affiliation:
Univ. Lille, CNRS, UMR 8524 - Laboratoire Paul Painlevé, F-59000 Lille, France (e-mail: sophie.grivaux@univ-lille.fr)
ÉTIENNE MATHERON
Affiliation:
Univ. Artois, UR 2462 - Laboratoire de Mathématiques de Lens (LML) F-62300 Lens, France (e-mail: etienne.matheron@univ-artois.fr)
QUENTIN MENET*
Affiliation:
Service d’Analyse fonctionnelle, Département de Mathématique, Université de Mons, Place du Parc 20, 7000 Mons, Belgium
Rights & Permissions [Opens in a new window]

Abstract

We introduce and study the notion of hereditary frequent hypercyclicity, which is a reinforcement of the well-known concept of frequent hypercyclicity. This notion is useful for the study of the dynamical properties of direct sums of operators; in particular, a basic observation is that the direct sum of a hereditarily frequently hypercyclic operator with any frequently hypercyclic operator is frequently hypercyclic. Among other results, we show that operators satisfying the frequent hypercyclicity criterion are hereditarily frequently hypercyclic, as well as a large class of operators whose unimodular eigenvectors are spanning with respect to the Lebesgue measure. However, we exhibit two frequently hypercyclic weighted shifts $B_w,B_{w'}$ on $c_0(\mathbb {Z}_+)$ whose direct sum ${B_w\oplus B_{w'}}$ is not $\mathcal {U}$-frequently hypercyclic (so that neither of them is hereditarily frequently hypercyclic), and we construct a C-type operator on $\ell _p(\mathbb {Z}_+)$, $1\le p<\infty $, which is frequently hypercyclic but not hereditarily frequently hypercyclic. We also solve several problems concerning disjoint frequent hypercyclicity: we show that for every $N\in \mathbb {N}$, any disjoint frequently hypercyclic N-tuple of operators $(T_1,\ldots ,T_N)$ can be extended to a disjoint frequently hypercyclic $(N+1)$-tuple $(T_1,\ldots ,T_N, T_{N+1})$ as soon as the underlying space supports a hereditarily frequently hypercyclic operator; we construct a disjoint frequently hypercyclic pair which is not densely disjoint hypercyclic; and we show that the pair $(D,\tau _a)$ is disjoint frequently hypercyclic, where D is the derivation operator acting on the space of entire functions and $\tau _a$ is the operator of translation by $a\in \mathbb {C}\setminus \{ 0\}$. Part of our results are in fact obtained in the general setting of Furstenberg families.

Type
Original Article
Copyright
© The Author(s), 2025. Published by Cambridge University Press

1. Introduction

This paper is devoted to two different topics, both pertaining to the study of the dynamics of linear operators. First, motivated by some questions regarding the behaviour of direct sums of operators, we introduce a new dynamical property of continuous linear operators on Banach or Fréchet spaces, which appears to be a very natural strengthening of the classical notion of frequent hypercyclicity; we call it hereditary frequent hypercyclicity. We believe that this is an interesting notion and we study it in some detail. Second, we address some questions concerning disjoint frequent hypercyclicity—also called diagonal frequent hypercyclicity. One notable connection between these two topics is that hereditarily frequently hypercyclic operators can be used to extend diagonally frequently hypercyclic tuples (see below).

In what follows, the letter X denotes an infinite-dimensional Polish topological vector space and $\mathfrak L(X)$ is the space of continuous linear operators on X. Recall that an operator $T\in \mathfrak L(X)$ is said to be hypercyclic if it has a dense orbit, that is, there exists $x\in X$ such that $\{T^n x:\ n\geq 0\}$ is dense in X; equivalently, for each non-empty open set $V\subset X$ , the ‘visit set’ $\mathcal N_T(x,V):=\{ n\in \mathbb {N};\; T^nx\in V\}$ is infinite. A much stronger property, introduced in [Reference Bayart and Grivaux5], is frequent hypercyclicity: the operator T is frequently hypercyclic if there exists $x\in X$ such that for each non-empty open set $V\subset X$ , the set $\mathcal N_T(x,V)$ has positive lower density. We refer the reader to [Reference Bayart and Matheron7, Reference Grosse-Erdmann and Peris33] for an in-depth presentation of various aspects of linear dynamics.

More recently, quantitative notions of hypercyclicity have begun to be studied in a very general framework [Reference Bès, Menet, Peris and Puig11, Reference Bonilla and Grosse-Erdmann17, Reference Ernst, Esser and Menet24]. Let $\mathcal F$ be a Furstenberg family, that is, a family of non-empty subsets of $\mathbb {N}$ which is hereditary upwards (if $A'\supset A\in \mathcal F$ , then $A'\in \mathcal F$ ). Following [Reference Bès, Menet, Peris and Puig11], we say that an operator $T\in \mathfrak L(X)$ is $\mathcal F$ -hypercyclic if there exists $x\in X$ , called an $\mathcal F$ -hypercyclic vector for T, such that for each non-empty open set $V\subset X$ , the set $\mathcal N_T(x,V)$ belongs to $\mathcal F$ . Thus, hypercyclicity corresponds to the family of all infinite subsets of $\mathbb {N}$ , and frequent hypercyclicity corresponds to the family of sets with positive lower density. The set of $\mathcal F$ -hypercyclic vectors for T will be denoted by ${\mathcal F}\text {-}\mathrm {HC}(T)$ . However, in accordance with a well-established notation, we write $\mathrm {HC}(T)$ in the hypercyclic case and $\mathrm {FHC}(T)$ in the frequently hypercyclic case. Also, when $\mathcal F$ is the family of all subsets of $\mathbb {N}$ with positive upper density, we say that T is $\mathcal U$ -frequently hypercyclic and we write $\text {UFHC}(T)$ .

The starting point of the paper is the following question.

Question 1.1. Let $T_1\in \mathfrak L(X_1)$ and $T_2\in \mathfrak L(X_2)$ be two frequently hypercyclic operators; is it true that $T_1\oplus T_2$ is frequently hypercyclic?

This question seems to have been considered for the first time in [Reference Grivaux, Matheron and Menet31, §8] and appears as a natural variant of the following well-known open problem in linear dynamics [Reference Bayart and Grivaux5]: if T is a frequently hypercyclic operator, is it true that $T\oplus T$ is frequently hypercyclic? Question 1.1 makes sense for $\mathcal F$ -hypercyclicity as well; and in the especially interesting case $T_1=T_2$ , the answer is known for the family of all infinite subsets of $\mathbb {N}$ and for the family of sets with positive upper density. Indeed, a famous example from [Reference De la Rosa and Read23] shows that hypercyclicity of T does not imply that of $T\oplus T$ , whereas it is proved in [Reference Ernst, Esser and Menet24] that $\mathcal U$ -frequent hypercyclicity of T does imply that of $T\oplus T$ .

Given $T_1\in \mathfrak L(X_1)$ and $T_2\in \mathfrak L(X_2)$ , two frequently hypercyclic operators, a natural way to show that $T_1\oplus T_2$ is frequently hypercyclic would be the following. Let $(V_i)_{i\in \mathbb {N}}$ be a countable basis of open sets for $X_1\times X_2$ and assume that each $V_i$ has the form ${V_i=V_{i,1}\times V_{i,2}}$ , where $V_{i,1}$ is open in $X_1$ and $V_{i,2}$ is open in $X_2$ . Pick a frequently hypercyclic vector $x_1\in X_1$ for $T_1$ . Then, for any $i\in \mathbb {N}$ , there exists a set $A_i\subset \mathbb {N}$ with positive lower density such that $T_1^{n}x\in V_{i,1}$ for all $n\in A_i$ . We would be done if we were able to find a vector $x_2\in X_2$ with the following property: for every $i\in \mathbb {N}$ , there exists a set $B_i$ with positive lower density and contained in $A_i$ (this is the important point, which cannot be guaranteed if $x_2$ is simply assumed to be frequently hypercyclic for $T_2$ ) such that $T_2^{n}x_2\in V_{i,2}$ for all $n\in B_i$ . This leads to the following definition.

Definition 1.2. Let $\mathcal F\subset 2^{\mathbb {N}}$ be a Furstenberg family. We say that an operator $T\in \mathfrak L(X)$ is hereditarily $\mathcal F$ -hypercyclic if, for any countable family $(V_i)_{i\in I} $ of non-empty open subsets of X and any family $(A_i)_{i\in I}\subset \mathcal F$ indexed by the same countable set I, there exists a vector $x\in X$ such that $\mathcal N_T(x,V_i)\cap A_i\in \mathcal F$ for every $i\in I$ ; in other words, for each $i\in I$ , there is a set $B_i\in \mathcal F$ such that $B_i\subset A_i$ and $T^nx\in V_i$ for all $n\in B_i$ .

When $\mathcal F$ is the family of sets with positive lower density, we say (of course) that the operator T is hereditarily frequently hypercyclic; and likewise for $\mathcal U$ -frequent hypercyclicity. By the above discussion, we get the following observation.

Observation 1.3. Let $\mathcal F\subset 2^{\mathbb {N}}$ be a Furstenberg family. If $T_1, T_2$ are two $\mathcal F$ -hypercyclic operators and at least one of them is hereditarily $\mathcal F$ -hypercyclic, then $T_1\oplus T_2$ is $\mathcal F$ -hypercyclic.

Note that when $\mathcal F$ is the family of all infinite subsets of $\mathbb {N}$ , hereditary $\mathcal F$ -hypercyclicity is equivalent to topological mixing; see §9.1 for the (easy) proof. So, Observation 1.3 implies in particular that the direct sum of a hypercyclic operator with a topologically mixing operator is hypercyclic; which is of course well known.

We also point out that—perhaps surprisingly—hereditary $\mathcal F$ -hypercyclicity automatically implies dense hereditary $\mathcal F$ -hypercyclicity: given $(A_i)$ and $(V_i)$ as in Definition 1.2 above, there is a dense set of vectors $x\in X$ satisfying the required property; see Proposition 6.4 below.

Having introduced a definition, we are immediately faced with some obvious questions.

• Are there any hereditarily frequently hypercyclic operators? We answer this in the affirmative by two different methods. Indeed, there are two ‘standard’ ways of proving that an operator is frequently hypercyclic: either by showing that it satisfies the so-called frequent hypercyclicity criterion (see [Reference Bayart and Matheron7] or [Reference Grosse-Erdmann and Peris33]) or by exhibiting a large supply of eigenvectors associated with unimodular eigenvalues (see for example [Reference Bayart and Grivaux5] or [Reference Bayart and Matheron8]). It turns out that in both cases, one gets in fact hereditary frequent hypercyclicity (Theorems 2.1, 3.1) or hereditary $\mathcal U$ -frequent hypercyclicity (Theorem 3.24).

• Is hereditary frequent hypercyclicity a new notion? In other words, are there any frequently hypercyclic operators which are not hereditarily frequently hypercyclic? The answer is ‘Yes’, and we prove this in two ways. On the one hand, we construct two frequently hypercyclic weighted shifts $B_w$ and $B_{w'}$ on $c_0(\mathbb Z_+)$ such that $B_w\oplus B_{w'}$ is not $\mathcal U$ -frequently hypercyclic (Theorem 4.2), so that neither of them can be hereditarily frequently hypercyclic by Observation 1.3. This also gives a strong negative answer to the $T_1\oplus T_2$ frequent hypercyclicity problem of Question 1.1. On the other hand, with the terminology of [Reference Grivaux, Matheron and Menet31], we construct a C-type operator on $\ell _p(\mathbb {Z}_+)$ , $1\leq p<\infty $ , which is frequently hypercyclic but not hereditarily frequently hypercyclic (Theorem 5.2).

• What are hereditarily frequently hypercyclic operators good for? We will use them in the context of ‘disjoint hypercyclicity’. The notion of disjointness in linear dynamics was introduced independently in [Reference Bernal-González10, Reference Bès and Peris14]. Let $N\geq 1$ and let $T_1,\ldots ,T_N\in \mathfrak L(X)$ . Following [Reference Bès and Peris14], we say that $T_1,\ldots ,T_N$ are disjoint or that the tuple $(T_1,\ldots ,T_N)$ is diagonally hypercyclic if there exists $x\in X$ such that the set $\{(T_1^n x,\ldots ,T_N^n x):\ n\geq 0\}$ is dense in $X^N$ ; in other words, the ‘diagonal’ vector $x\oplus \cdots \oplus x$ is hypercyclic for $T_1\oplus \cdots \oplus T_N$ . Such a vector x is said to be d-hypercyclic for the tuple $(T_1,\ldots ,T_N)$ and the set of d-hypercyclic vectors for $(T_1,\ldots ,T_N)$ will be denoted by $d{-}\mathrm {HC}(T_1,\ldots ,T_N)$ . Similarly, $(T_1,\ldots ,T_N)$ is said to be d-frequently hypercyclic if there exists $x\in X$ such that $x\oplus \cdots \oplus x$ is a frequently hypercyclic vector for $T_1\oplus \cdots \oplus T_N$ , and we denote by $d{-}\mathrm {FHC}(T_1,\ldots ,T_N)$ the set of d-frequently hypercyclic vectors for $(T_1,\ldots ,T_N)$ .

A natural problem regarding d-hypercyclicity is that of the extension of d-hypercyclic tuples. It was shown in [Reference Martin and Sanders39] that given any $N\geq 1,$ any Banach space X and any $T_1,\ldots ,T_N\in \mathfrak L(X)$ such that $(T_1,\ldots ,T_N)$ is d-hypercyclic, there exists $T_{N+1}\in \mathfrak L(X)$ such that $(T_1,\ldots ,T_{N+1})$ is also d-hypercyclic. As for d-frequent hypercyclicity, the situation is trickier since there exist Banach spaces which do not support any frequently hypercyclic operator [Reference Shkarin53]. The best one could hope for is that as soon as X supports a frequently hypercyclic operator, then one can extend d-frequently hypercyclic tuples. We are unable to prove this, but we show that one can indeed extend d-frequently hypercyclic tuples as soon as X supports a hereditarily frequently hypercyclic operator (Theorem 6.1).

The preceding discussion has outlined the content of §§26 of the paper and hopefully it is clear that §6 makes a transition between our two topics—hereditary frequent hypercyclicity and d-frequent hypercyclicity. The next two sections are exclusively devoted to d-frequent hypercyclicity. In §7, we show (in the spirit of [Reference Sanders and Shkarin51]) that there exists a d-frequently hypercyclic pair $(T_1,T_2)$ on some Banach space X which is not densely d-hypercyclic, that is, the set $d{-}\mathrm {HC}(T_1,T_2)$ is not dense in X (Theorem 7.2). In §8, we give a sufficient condition for d-frequent hypercyclicity of a tuple $(T_1,\ldots ,T_N)$ in terms of eigenvectors of the operators $T_i$ (Theorem 8.2); and this allows us for example to show that the pair $(D, \tau _a)$ is d-frequently hypercyclic, where D is the derivation operator on the space of entire functions $H(\mathbb {C})$ and $\tau _a$ is the operator of translation by $a\in \mathbb {C}\setminus \{ 0\}$ .

Finally, §9 contains a few additional remarks and a number of open questions originating in a rather natural way from our work.

2. The frequent hypercyclicity criterion

The frequent hypercyclicity criterion (FHCC) is a very efficient tool for showing that a given operator is frequently hypercyclic. Since [Reference Bayart and Grivaux5], there have been several versions of it in the literature; we choose here the most widely used [Reference Bonilla and Grosse-Erdmann16]: an operator $T\in \mathfrak L(X)$ satisfies the FHCC provided there exist a dense set $\mathcal D\subset X$ and a map $S:\mathcal D\to \mathcal D$ such that:

  • $TS=I$ on $\mathcal D$ ;

  • for any $x\in \mathcal D,$ the series $\sum T^n x$ and $\sum S^n x$ are unconditionally convergent.

In this section, we show that the FHCC implies, in fact, hereditary frequent hyper- cyclicity.

Theorem 2.1. If $T\in \mathfrak L(X)$ satisfies the FHCC, then T is hereditarily frequently hypercyclic.

The proof of this theorem will closely mimic the classical proof that an operator satisfying the FHCC is frequently hypercyclic. Recall that the latter depends on the construction of subsets of $\mathbb {N}$ with positive lower density which are ‘well separated’. To obtain hereditary frequent hypercyclicity, we need to control more precisely these subsets and, in particular, we have to be sure that one can find them inside some prescribed subsets of $\mathbb {N}$ (of positive lower density, of course). This is the content of the next lemma, which is useful in other situations as well (see [Reference Martin, Menet and Puig36] and the very recent [Reference Cardeccia and Muro18]). This lemma is actually contained in [Reference Martin, Menet and Puig36, Lemma 2.2], but we give a proof for completeness (and convenience of the reader).

Lemma 2.2. Let $(A_i)_{i\in I}$ be a countable family of subsets of $\mathbb {N}$ with positive lower density and let $(N_i)_{i\in I}$ be a family of positive integers indexed by the same countable set I. There exists a family $(B_i)_{i\in I}$ of pairwise disjoint subsets of $\mathbb {N}$ with positive lower density such that:

  1. (a) $B_i\subset A_i$ and $\min (B_i)\geq N_i$ for all $i\in I$ ;

  2. (b) for any $i, j\in I$ and any $(n,m)\in B_i\times B_j$ with $n\neq m, |n-m|\geq N_i+N_j.$

Proof. We may assume that $I=\mathbb {N}$ . For each $i\in \mathbb {N}$ , let $ M_i:= 2\, \max _{j\leq i} \, N_j.$

Enumerate each set $A_i$ as an increasing sequence $(n_i(k))_{k\in \mathbb N}$ and for $s\geq 1,$ define

$$ \begin{align*} A_{i, s}&:=\{n_{i}(sk):\ k\in\mathbb{N}\},\\ \widetilde{A}_{i, s}&:=(A_{i,s}+[-M_i,M_i])\cap \mathbb{N}. \end{align*} $$

Then (by subadditivity of upper density),

$$ \begin{align*} \overline{\textrm{dens}}(\widetilde{A}_{i,s}) \leq \frac{2M_i+1}{s}\cdot \end{align*} $$

Since all sets $A_{i,s}$ have positive lower density, it follows that one can construct, by induction, a sequence of positive integers $(s(i))_{i\in \mathbb {N}}$ such that for all $i\in \mathbb N, s(i)\geq M_i$ and

$$ \begin{align*}\overline{\textrm{dens}}(\widetilde{A}_{i, s(i)})\leq \min_{j<i}\frac{1}{4^{i-j}}\, \underline{\textrm{dens}}(A_{j, s(j)}).\end{align*} $$

We then set

$$ \begin{align*}B_i:=A_{i, s(i)}\backslash \bigcup_{j>i}\widetilde{A}_{j,s(j)},\end{align*} $$

so that

$$ \begin{align*} \underline{\textrm{dens}}(B_i)&\geq \underline{\textrm{dens}} (A_{i,s(i)})\bigg(1-\sum_{j>i}\frac 1{4^{j-i}}\bigg) >0. \end{align*} $$

Moreover, $B_i$ is clearly contained in $A_i$ , $\min (B_i)\geq s(i)\geq M_i$ , the sets $B_i$ are pairwise disjoint, and if $(n,m)\in B_i\times B_j$ with $n\neq m$ and $j\geq i$ , then:

  • either $j=i$ , in which case $|n-m|\geq s(i)\geq M_i\geq 2N_i=N_i+N_j$ ;

  • or $j> i$ , in which case $|n-m|\geq M_j\geq N_i+N_j$ since $n\notin B_j+[-M_j,M_j]$ .

We can now give the proof of Theorem 2.1.

Proof of Theorem 2.1

Let $(A_p)_{p\in \mathbb {N}}$ be a sequence of subsets of $\mathbb {N}$ with positive lower density and let $(V_p)_{p\in \mathbb {N}}$ be a sequence of non-empty open subsets of X. We have to find a vector $x\in X$ such that $\mathcal N_T(x,V_p)\cap A_p$ has positive lower density for all $p\in \mathbb {N}$ ; and for that, we follow the proof of [Reference Bayart and Matheron7, Theorem 6.18]. Let us fix an F-norm $\Vert \,\cdot \,\Vert $ defining the topology of X.

Let $\mathcal D$ be the dense set given by the FHCC. For each $p\geq 1$ , choose a vector ${x_p\in \mathcal D\cap V_p}$ and let $\alpha _p>0$ be such that $B(x_p,3\alpha _p)\subset V_p$ . Let also $(\varepsilon _p)_{p\geq 1}$ be a summable sequence of positive real numbers such that for all $p\geq 1,$

$$ \begin{align*}p\varepsilon_p+\sum_{q>p+1}\varepsilon_q<\alpha_p.\end{align*} $$

By unconditional convergence of the series involved in the FHCC, for each $p\geq 1,$ one can find a positive integer $N_p$ such that, for any set $F\subset \mathbb {N}\cap [N_p,\infty ),$

$$ \begin{align*} \bigg\|\!\sum_{n\in F}T^n x_i\bigg\|+\bigg\|\!\sum_{n\in F}S^nx_i\bigg\|\leq {\varepsilon_p}\quad\text{ for all } i\leq p.\end{align*} $$

Now, let $(B_p)_{p\in \mathbb {N}}$ be the sequence of subsets of $\mathbb {N}$ with positive lower density associated with $(A_p)$ and $(N_p)$ by Lemma 2.2. The vector x we are looking for is defined by

$$ \begin{align*}x:= \sum_{p=1}^{\infty}\sum_{n\in B_p}S^n x_p.\end{align*} $$

First, we note that x is well defined. Indeed, each series $\sum _{n\in B_p}S^n x_p$ is convergent and since $B_p\subset [N_p,\infty )$ for all p, we have

$$ \begin{align*}\sum_{p\geq 1}\bigg\|\!\sum_{n\in B_p}S^n x_p\bigg\|\leq\sum_{p\geq 1}\varepsilon_p<\infty.\end{align*} $$

Let us fix $p\geq 1$ and $n\in B_p$ : we show that $T^nx\in V_p$ . By definition of x, we have

$$ \begin{align*} \|T^n x-x_p\|&\leq \sum_{q=1}^{\infty}\bigg\|\!\sum_{\substack{m\in B_q\\ m>n}}S^{m-n} x_q \bigg\|+\sum_{q=1}^{\infty}\bigg\|\!\sum_{\substack{m\in B_q\\ m<n}}T^{n-m} x_q \bigg\|. \end{align*} $$

To estimate the first sum, we decompose it as

$$ \begin{align*}\sum_{q=1}^{p}\bigg\|\!\sum_{\substack{m\in B_q\\ m>n}}S^{m-n} x_q\bigg\|+\sum_{q=p+1}^{\infty}\bigg\|\!\sum_{\substack{m\in B_q\\ m>n}}S^{m-n} x_q\bigg\|.\end{align*} $$

Since $n\in B_p,$ we know that $m-n>\max (N_p,N_q)$ whenever $m\in B_q$ and $m>n$ . By the choice of the sequence $(N_p)$ , it follows that

$$ \begin{align*}\sum_{q=1}^{\infty}\bigg\|\!\sum_{\substack{m\in B_q\\ m>n}}S^{m-n} x_q\bigg\|\leq p\varepsilon_p+\sum_{q=p+1}^{\infty}\varepsilon_q<\alpha_p.\end{align*} $$

Estimating the second sum in the same way, we conclude that

$$ \begin{align*}\|T^n x-x_p\|< 3\alpha_p,\end{align*} $$

so that $T^n(x)\in V_p.$

As a direct consequence of Theorem 2.1 and the fact that every frequently hypercyclic weighted backward shift on $\ell _p(\mathbb Z_+)$ satisfies the FHCC [Reference Bayart and Ruzsa9], we obtain the following corollary.

Corollary 2.3. A weighted backward shift on $\ell _p(\mathbb Z_+), 1\leq p<\infty $ is frequently hypercyclic if and only if it is hereditarily frequently hypercyclic.

Remark 2.4. It should be clear from the proof of Theorem 2.1 that the FHCC implies hereditary $\mathcal F$ -hypercyclicity for any Furstenberg family $\mathcal F$ satisfying Lemma 2.2. For example, this holds true for the family of sets with positive upper density and the family of sets with positive Banach upper density, see [Reference Martin, Menet and Puig36, Lemma 2.2]; so any operator satisfying the FHCC is hereditarily $\mathcal U$ -frequently hypercyclic. It is not clear to us that this could be deduced from Theorem 2.1: indeed, if $\mathcal F$ and $\mathcal F'$ are two Furstenberg families, the inclusion $\mathcal F\subset \mathcal F'$ does not formally imply that hereditary $\mathcal F$ -hypercyclicity is a stronger property than hereditary $\mathcal F'$ -hypercyclicity.

3. Ergodic theory

3.1. Results and general strategy

It is well known that if $T\in \mathfrak L(X)$ and if one can find a T-invariant Borel probability measure $\mu $ on X with full support with respect to which T is an ergodic transformation, then T is frequently hypercyclic. Let us recall the argument. Let $(V_p)_{p\in \mathbb {N}}$ be a countable basis of open sets for X. Applying Birkhoff’s pointwise ergodic theorem to the characteristic functions $\mathbf 1_{V_p}$ , we obtain a sequence $(\Omega _p)$ of subsets of X with $\mu (\Omega _p)=1$ such that

$$ \begin{align*}\frac1N\, \# \mathcal (N_T(x,V_p)\cap[0,N-1])=\frac 1N\sum_{n=0}^{N-1}\mathbf 1_{V_p}(T^n x)\xrightarrow{N\to\infty}\mu(V_p)>0\quad\textrm{ for every }x\in \Omega_p.\end{align*} $$

Hence, any $x\in \bigcap _{p\in \mathbb {N}} \Omega _p$ is a frequently hypercyclic vector for T.

It is also well known (see for example [Reference Bayart and Grivaux5], [Reference Bayart and Matheron7, Ch. 5], [Reference Bayart and Matheron8]) that if X is a complex Banach space and if an operator $T\in \mathfrak L(X)$ admits ‘sufficiently many’ $\mathbb {T}$ -eigenvectors (that is, eigenvectors whose associated eigenvalues have modulus $1$ ), then it is indeed possible to find an ergodic measure with full support for T. So, it may seem reasonable to expect that operators with sufficiently many $\mathbb {T}$ -eigenvectors are hereditarily frequently hypercyclic.

Now, if one wants to repeat the above argument to show that a given operator is hereditarily frequently hypercyclic, Birkhoff’s ergodic theorem is not enough: what is needed is a pointwise convergence result for averages of quantities of the form $\mathbf 1_V(T^nx)$ not only along the whole sequence of integers, but in fact along any sequence $(n_k)$ with positive lower density. Specifically, we will use a theorem of Conze [Reference Conze20], which we state in the next section (Theorem 3.3). The assumptions of Conze’s theorem are much stronger than merely asking that T is an ergodic transformation; so we will need to impose a rather strong condition on the $\mathbb {T}$ -eigenvectors to be able to conclude that our operator T is indeed hereditarily frequently hypercyclic.

Let us recall a few definitions. Assume that X is a complex Fréchet space. If $T\in \mathfrak L(X)$ , a $\mathbb {T}$ -eigenvector field for T is any bounded map $E:\mathbb {T}\to X$ such that $TE(z)=z E(z)$ for every $z\in \mathbb {T}$ . Given a positive Borel measure $\sigma $ on $\mathbb {T}$ , we say that a family of $\mathbb {T}$ -eigenvector fields $(E_i)_{i\in I}$ is $\sigma $ -spanning if $\overline {\mathrm {span}}\,( E_i(z)\,:\, z\in \mathbb {T}\setminus N\,,\, i\in I)=X$ for every Borel set $N\subset \mathbb {T}$ such that $\sigma (N)=0$ . Similarly, we say that the $\mathbb {T}$ -eigenvectors of T are $\sigma $ -spanning if for every Borel set $N\subset \mathbb {T}$ such that $\sigma (N)=0$ , the eigenvectors of T with eigenvalues in $\mathbb {T}\setminus N$ span a dense subspace of X. It follows from [Reference Bayart and Matheron7, Lemma 5.29] that the $\mathbb {T}$ -eigenvectors of T are $\sigma $ -spanning if and only if there exists a $\sigma $ -spanning countable family of Borel $\mathbb {T}$ -eigenvector fields for T. One of our main aims in this part of the paper will be to prove the following theorem. (We refer to any book on Banach space theory, for example [Reference Albiac and Kalton2], for the definition of type. Let us just recall that an $L_p$ -space has type $2$ if $2\leq p<\infty $ and type p if $1\leq p<2$ .)

Theorem 3.1. Let X be a separable complex Banach space and let $T\in \mathfrak L(X)$ . Assume that one can find a $\unicode{x3bb} $ -spanning, finite or countably infinite family $(E_i)_{i\in I}$ of $\mathbb T$ -eigenvector fields for T, where $\unicode{x3bb} $ is the Lebesgue measure on $\mathbb {T}$ . Moreover, assume that one of the following holds true:

  1. (a) X has type $2$ ;

  2. (b) X has type $p\in [1,2)$ and each eigenvector field $E_i$ is $\alpha _i$ -Hölderian for some ${\alpha _i>1/p-1/2}$ .

Then T is hereditarily frequently hypercyclic.

From this, we immediately get the following corollary.

Corollary 3.2. Let X be a separable complex Banach space and let $T\in \mathfrak L(X)$ . If the $\mathbb {T}$ -eigenvectors of T are spanning with respect to the Lebesgue measure and if, moreover, the Banach space X has type $2$ , then T is hereditarily frequently hypercyclic.

In view of the above discussion, our strategy for proving Theorem 3.1 should be clear: we will show that under the above assumptions, one can find a T-invariant (Borel probability) measure $\mu $ on X for which one can apply Conze’s Theorem 3.3 below to get hereditary frequent hypercyclicity. This will yield a more precise result, Theorem 3.9.

The proof of Theorem 3.1 will occupy us from §§3.2 to 3.6. After that, in §3.7, we will come back to the FHCC; in particular, we explain how the ergodic-theoretic point of view allows to give a completely different proof of Theorem 2.1, and we compare the assumptions in Theorems 2.1 and 3.1. Finally, in §3.8, we show that an assumption on the $\mathbb {T}$ -eigenvectors similar to that of Corollary 3.2 but much weaker—namely, the perfect spanning property—is sufficient to imply hereditary $\mathcal U$ -frequent hypercyclicity (without any additional assumption on the space X).

3.2. Conze’s theorem

Before stating Conze’s theorem, let us introduce some terminology.

In what follows, by a measure-preserving dynamical system $(X, \mathcal B, \mu , T)$ , we mean a pair consisting of a probability space $(X, \mathcal B, \mu )$ and a measurable map $T:X\to X$ such that $\mu \circ T^{-1}=\mu $ . Note that here, we are departing from our standing notation: X is an abstract space, not necessarily a topological vector space, and hence, T is not necessarily a linear operator. This ambiguity is in fact intentional and it should cause no confusion.

Given a measure-preserving dynamical system $(X, \mathcal B, \mu , T)$ , we denote by ${U_T:L^2(X, \mathcal B, \mu )\to L^2(X, \mathcal B, \mu )}$ the associated Koopman operator, which is defined by

$$ \begin{align*} U_Tf:= f\circ T.\end{align*} $$

This is an isometry of $L^2(X, \mathcal B, \mu )$ , and a unitary operator if T is bijective and bimeasurable, in which case we say that T is an automorphism of $(X, \mathcal B, \mu )$ or that the measure-preserving dynamical system $(X, \mathcal B, \mu , T)$ is invertible.

We stress a technical point: all probability spaces $(X, \mathcal B, \mu )$ under consideration will be assumed to be standard Borel, that is, the underlying measurable space $(X,\mathcal B)$ is isomorphic to $(Z, \mathcal B(Z))$ for some Polish space Z, where $\mathcal B(Z)$ is the Borel $\sigma $ -algebra of Z. Additionally, when X is already a Polish space, we assume that $\mathcal B$ is the Borel $\sigma $ -algebra of X.

A unitary operator $U:H\to H$ acting on a Hilbert space H is said to have Lebesgue spectrum if there exists a family of vectors $(f_i)_{i\in I}$ in H such that $\{U^n f_i:\ n\in \mathbb {Z},\ i\in I\}$ is an orthonormal basis of H. Observe that if H is separable, the family $(f_i)$ has to be finite or countably infinite. When there exists a countably infinite such family $(f_i)$ , we say that U has countable Lebesgue spectrum. Finally, we say that an invertible measure-preserving dynamical system $(X,\mathcal B,\mu , T)$ has (countable) Lebesgue spectrum if the restriction of the Koopman operator $U_T$ to $ L^2_0(X,\mathcal B,\mu ):=\{ f\in L^2(X,\mathcal B,\mu )\,:\, \int _{X} f\, d\mu =0\}$ has (countable) Lebesgue spectrum. We may also say that T itself has (countable) Lebesgue spectrum.

Conze’s theorem from [Reference Conze20] now reads as follows.

Theorem 3.3. Let $(X,\mathcal B,\mu , T)$ be an invertible measure-preserving dynamical system and assume that T has Lebesgue spectrum. If $(n_k)_{k\geq 0}$ is an increasing sequence of integers with positive lower density then, for any $f\in L^1(\mu ),$

$$ \begin{align*}\frac 1N\sum_{k=0}^{N-1} f(T^{n_k}x)\xrightarrow{N\to\infty}\int_X f\, d\mu\quad\text{ for } \mu\text{-almost every } x\in X.\end{align*} $$

Getting back to the case where X is a separable Banach space, it is easy to check that if $T\in \mathfrak L(X)$ and if one can find a T-invariant measure with full support $\mu $ such that $(X,\mathcal B,\mu , T)$ satisfies the assumptions of Conze’s theorem, then T is hereditarily frequently hypercyclic. However, Conze’s theorem is only stated for automorphisms and, in our context, it would be rather restrictive to confine ourselves to the case of automorphisms. We will get round this problem thanks to the notion of factor. Recall that a measure-preserving dynamical system $(X,\mathcal B,\mu ,T)$ is a factor of a measure-preserving dynamical system $(Y ,\mathcal C,\nu ,S)$ or that $(Y ,\mathcal C,\nu ,S)$ is an extension of $(X,\mathcal B,\mu ,T)$ , if (possibly after deleting two sets of measure $0$ in X and Y) there exists a measurable map $\pi :Y\to X$ such that $\pi \circ S=T\circ \pi $ and $\mu =\nu \circ \pi ^{-1}$ . Using suitable extensions, we will be able to apply Conze’s theorem to general operators T via the following corollary.

Corollary 3.4. Let $T\in \mathfrak L(X)$ and assume that there exists a T-invariant probability measure $\mu $ on X with full support such that $(X,\mathcal B,\mu ,T)$ is a factor of an invertible measure-preserving dynamical system $(Y,\mathcal C,\nu ,S)$ with Lebesgue spectrum. Then, T is hereditarily frequently hypercyclic. More precisely, given a countable family $(A_i)_{i\in I}$ of subsets of $\mathbb {N}$ with positive lower density and a family $(V_i)_{i\in I}$ of non-empty open sets, $\mu $ -almost every $x\in X$ is such that $\mathcal N_T(x,V_i)\cap A_i$ has positive lower density for every $i\in I$ .

Proof. It is enough to prove the result for a single pair $(A,V)$ , where $A\subset \mathbb {N}$ has positive lower density and V is a non-empty open set.

Let $(n_k)_{k\geq 1}$ be the increasing enumeration of A. Let also $W:= \pi ^{-1}(V)$ , where ${\pi :(Y ,\mathcal C,\nu ,S)\to (X,\mathcal B,\mu ,T)}$ is a factor map, so that $\nu (W)=\mu (V)>0$ , and let

$$ \begin{align*}\Omega:=\bigg\{y\in Y:\ \frac 1N\sum_{k=1}^N \mathbf 1_{{W}}(S^{n_{k}}y)\xrightarrow{N\to\infty}\mu(V)\bigg\}.\end{align*} $$

By Conze’s theorem, we know that $\nu (\Omega )=1$ . Since we are working with standard Borel spaces, it follows that the set $\pi (\Omega )$ is $\mu $ -measurable (being an analytic set, it is universally measurable) and that $\mu ( \pi (\Omega ))=1$ . So it is enough to show that every $x\in \pi (\Omega )$ is such that $\underline {\textrm {dens}}( \mathcal N_T(x,V)\cap A)>0$ .

Let $x\in \pi (\Omega )$ and write x as $x=\pi (y)$ for some $y\in \Omega $ . By definition of $\Omega $ , we know that the set $\{k\geq 1:\ S^{n_{k}}y\in {W}\}$ has positive lower density; enumerate it as an increasing sequence $(m_{k})_{k\geq 1}.$ Then, $B:=\{n_{m_{k}}:\ k\geq 1\}$ has positive lower density, it is contained in A and $S^n y\in \pi ^{-1}(V)$ for all $n\in B$ . This means that $T^n x\in V$ for all $n\in B,$ which concludes the proof.

3.3. Natural extensions

To apply Corollary 3.4, we need a simple way of extending a given measure-preserving dynamical system $(X,\mathcal B,\mu ,T)$ to a measure-preserving dynamical system $(\widetilde X,\widetilde {\mathcal B}, \widetilde \mu , \widetilde T)$ , where $\widetilde T$ is an automorphism of $(\widetilde X,\widetilde {\mathcal B},\widetilde \mu )$ . Fortunately, there is a canonical procedure doing precisely that. Consider the set

$$ \begin{align*}\widetilde X:=\{(x_k)_{k\geq 0}\in X^{\mathbb{Z}_+}:\ T(x_{k+1})=x_k\textrm{ for all }k\geq 0\},\end{align*} $$

and for $k\geq 0,$ let $\pi _k:\widetilde X\to X$ denote the projection onto the kth coordinate of $\widetilde X.$ Endow $\widetilde X$ with the smallest $\sigma $ -algebra $\widetilde {\mathcal B}$ which makes every projection $\pi _k$ measurable. The Rokhlin’s natural extension of T is the measurable transformation $\widetilde T:\widetilde X\to \widetilde X$ defined by

$$ \begin{align*}\widetilde T(x_0, x_1, \ldots):=(T(x_0),x_0,x_1,\ldots).\end{align*} $$

One can prove that there exists a unique probability measure $\widetilde \mu $ on $(\widetilde X,\widetilde {\mathcal B})$ such that ${\widetilde \mu \circ \pi _k^{-1}=\mu }$ for all $k\geq 0$ . (Here, the fact $(X,\mathcal B)$ is a standard Borel space is needed.) Then, it is a simple exercise to check that $\widetilde T:\widetilde X\to \widetilde X$ is an automorphism of $(\widetilde X,\widetilde B, \widetilde \mu )$ such that $\pi _k\circ \widetilde T=T\circ \pi _k$ for all $k\geq 0$ . In particular, $(X,\mathcal B,\mu ,T)$ is a factor of $(\widetilde { X},\widetilde { \mathcal B},\widetilde \mu ,\widetilde T)$ as witnessed by $\pi _0: \widetilde X\to X$ . Note also that if X is a Polish space, then $\widetilde X$ is a Borel subset of the Polish space $X^{\mathbb {Z}_+}$ endowed with the product topology, and $\widetilde {\mathcal B}$ is the Borel sigma-algebra of $\widetilde X$ . For more details, see for example [Reference Urbański, Roy and Munday56, §8.4] or [Reference Nicol and Petersen46].

We will need the following lemma. Recall that if X is a complex Fréchet space, a Borel probability measure $\mu $ on X is said to be Gaussian if every continuous linear functional on X has a complex symmetric Gaussian distribution when considered as a random variable on $(X,\mathcal B,\mu )$ .

Lemma 3.5. Assume that X is a complex Fréchet space, that $T\in \mathfrak L(X)$ and that the measure $\mu $ is Gaussian. Then, $\widetilde X$ is a Fréchet space when endowed with the induced product topology of $X^{{\mathbb {Z}}^{+}}$ , $\widetilde T$ is an invertible continuous linear operator on $\widetilde X$ and the measure $\widetilde \mu $ is Gaussian.

Proof. It is clear that $\widetilde X$ is a closed linear subspace of $X^{\mathbb {Z}_+}$ (and hence a Fréchet space) and that $\widetilde T$ is an invertible continuous linear operator on $\widetilde X$ .

Let $\phi $ be a continuous linear functional on $\widetilde X$ . By the Hahn–Banach theorem, one can extend $\phi $ to a continuous linear functional $\Phi $ on $X^{\mathbb {Z}_+}$ . Since $X^{\mathbb {Z}_+}$ is endowed with the product topology, this linear functional $\Phi $ has the form

$$ \begin{align*} \Phi=\sum_{k=0}^N x^*_k\circ \pi_k,\end{align*} $$

where $x_0^*,\ldots , x_N^*\in X^*$ . Moreover, by definition of $\widetilde X$ , we have $\pi _k=T^{N-k}\circ \pi _N$ on $\widetilde X$ for $k=0,\ldots ,N$ ; so we may write

$$ \begin{align*} \phi =x^*\circ \pi_N\quad \text{where } x^*=\sum_{k=0}^N x_k^*\circ T^{N-k}\in X^*.\end{align*} $$

Hence, $\widetilde \mu \circ \phi ^{-1}=( \widetilde \mu \circ \pi _N^{-1})\circ (x^*)^{-1}=\mu \circ (x^*)^{-1}$ . Since $\mu $ is a Gaussian measure, it follows that $\widetilde \mu $ is Gaussian as well.

3.4. Gaussian measures and countable Lebesgue spectrum

In this section, we state a general result about Gaussian linear measure-preserving dynamical systems. This result looks very much like [Reference Cornfeld, Fomin and Sinai21, Theorem 14.3.2], but it is not clear to us that it can be formally deduced from it; so we will give a more or less complete proof. However, this proof is not necessary for understanding the proof of Theorem 3.1, so it is postponed to §3.6 for the sake of better readability. This means that a reader who is ready to take Proposition 3.6 on faith can safely ignore that section.

Recall that if $(X,\mathcal B, \mu , T)$ is a measure-preserving dynamical system then, for any $f\in L^2(\mu )$ , there is a unique positive Borel measure $\sigma _{f}$ on $\mathbb {T}$ with non-negative Fourier coefficients

$$ \begin{align*} \widehat{\,\sigma_{f}}(n)=\langle U_T^n f,f\rangle_{L^2(\mu)},\quad n\geq 0.\end{align*} $$

This follows from Herglotz’s theorem and from the fact that the Koopman operator ${U_T:L^2(\mu )\to L^2(\mu )}$ is a contraction operator (so that the map $\mathbb {Z}_+\ni n\mapsto \langle U_T^n f,f\rangle _{L^2(\mu )}$ can be extended to a positive-definite function on $\mathbb {Z}$ , see for example [Reference Szőkefalvi-Nagy, Foias, Bercovici and Kérchy55, p. 28]). The measure $\sigma _f$ is the spectral measure of f with respect to $U_T$ .

Proposition 3.6. Let $(X,\mathcal B,\mu , T)$ be a measure-preserving dynamical system where X is a separable Fréchet space, T is an invertible linear operator and $\mu $ is a Gaussian measure on X. Assume that for any $x^*\in X^*\subset L^2(\mu )$ , the spectral measure $\sigma _{x^*}$ with respect to $U_T$ is absolutely continuous with respect to the Lebesgue measure. Then, $(X,\mathcal B,\mu , T)$ has countable Lebesgue spectrum.

Proposition 3.6 has the following consequence for not necessarily invertible Gaussian linear measure-preserving dynamical systems.

Corollary 3.7. Let $(X,\mathcal B,\mu , T)$ be a measure-preserving dynamical system, where X is a complex separable Fréchet space, T is a continuous linear operator and $\mu $ is a Gaussian measure on X. Assume that for any $x^*\in X^*\subset L^2(\mu )$ , the spectral measure $\sigma _{x^*}$ with respect to $U_T$ is absolutely continuous with respect to the Lebesgue measure. Then, the natural extension of $(X,\mathcal B,\mu , T)$ has countable Lebesgue spectrum.

Proof. Let $(\widetilde X,\widetilde {\mathcal B},\widetilde {\mu }, \widetilde {T})$ be the natural extension. By Lemma 3.5, we know that $\widetilde T$ is an invertible linear operator and $\widetilde \mu $ is a Gaussian measure. Hence, by Proposition 3.6, it is enough to show that for any continuous linear functional $\phi $ on $\widetilde X$ , the spectral measure $\sigma _{\phi }$ (with respect to $U_{\widetilde T}$ ) is absolutely continuous with respect to the Lebesgue measure.

By the proof of Lemma 3.5, one can find $N\in \mathbb {Z}_+$ and $x^*\in X^*$ such that $\phi =x^*\circ \pi _N$ on $\widetilde X$ , where $\pi _N:X^{\mathbb {Z}_+}\to X$ is the projection onto the Nth coordinate. Now, we apply the following purely formal fact, which holds true for any measure-preserving dynamical system $(X,\mathcal B,\mu , T)$ .

Fact 3.8. Let $f\in L^2(X,\mathcal B,\mu )$ and $F:=f\circ \pi _N\in L^2(\widetilde X, \widetilde B, \widetilde \mu )$ . Then, the spectral measure $\sigma _{F}$ with respect to $U_{\widetilde T}$ is equal to $\sigma _f$ , the spectral measure of f with respect to $U_T$ .

Proof of Fact 3.8

For all $n\geq 0$ ,

$$ \begin{align*} \langle U_{\widetilde T}^n F , F\rangle_{L^2(\widetilde\mu)}=\langle F \circ\widetilde T^n,F\rangle_{L^2(\widetilde \mu)}&=\langle f\circ\pi_N\circ\widetilde T^n,f\circ\pi_N\rangle_{L^2(\widetilde \mu)}\\ &=\langle f\circ T^n \circ\pi_N,f\circ\pi_N\rangle_{L^2(\widetilde \mu)}\\ &=\langle f\circ T^n ,f\rangle_{L^2(\mu)}\\&= \langle U_T^n f, f\rangle_{L^2(\mu)}.\\[-37pt] \end{align*} $$

By Fact 3.8, $\sigma _\phi =\sigma _{x^*}$ is absolutely continuous with respect to the Lebesgue measure, which concludes the proof of Corollary 3.7.

3.5. Proof of Theorem 3.1

We can now state and prove very easily the following more precise version of Theorem 3.1.

Theorem 3.9. Let X be a separable complex Banach space and let $T\in \mathfrak L(X)$ . Assume that one can find a $\unicode{x3bb} $ -spanning, finite or countably infinite family $(E_i)_{i\in I}$ of $\mathbb T$ -eigenvector fields for T, where $\unicode{x3bb} $ is the Lebesgue measure on $\mathbb {T}$ . Moreover, assume that one of the following holds true:

  1. (a) X has type $2$ ;

  2. (b) X has type $p\in [1,2)$ , and each $E_i$ is $\alpha _i$ -Hölderian for some $\alpha _i>1/p-1/2$ .

Then, there exists a T-invariant Gaussian measure $\mu $ on X with full support such that $(X,\mathcal B,\mu ,T)$ is a factor of a measure-preserving system which has countable Lebesgue spectrum. In particular, T is hereditarily frequently hypercyclic; and more precisely: given a countable family $(A_i)_{i\in I}$ of subsets of $\mathbb {N}$ with positive lower density and a family $(V_i)_{i\in I}$ of non-empty open sets, $\mu $ -almost every $x\in X$ is such that $\mathcal N_T(x,V_i)\cap A_i$ has positive lower density for every $i\in I$ .

Proof. Under the assumptions above, there exists a T-invariant Gaussian measure $\mu $ on X with full support such that, for all $x^*\in X^*,$ the measure $\sigma _{x^*}$ is absolutely continuous with respect to the Lebesgue measure; this is contained for instance in [Reference Bayart and Matheron7, Lemma 5.35, (4)]. So, the result follows immediately from Corollaries 3.7 and 3.4.

3.6. Proof of Proposition 3.6

In this section, we give the promised proof of Proposition 3.6.

3.6.1. Two facts concerning unitary operators

We will need several results on unitary operators. These results are certainly well known, but since they are rather difficult to locate in the literature, we provide some details.

Let U be a unitary operator acting on a complex separable Hilbert space $H.$ By Herglotz’s theorem, for any $f\in H$ , there exists a unique positive and finite Borel measure $\sigma _f$ on $\mathbb {T}$ , called the spectral measure of f with respect to U, such that

$$ \begin{align*} \text{ for all } n\in\mathbb{Z}\;:\; \langle U^n f,f\rangle=\widehat{\sigma_f}(n).\end{align*} $$

Note in particular that $\sigma _f$ is equal to the Lebesgue measure $\unicode{x3bb} $ if and only if the sequence $(U^nf)_{n\in \mathbb {Z}}$ is orthonormal. This explains the terminology ‘Lebesgue spectrum’.

Denote by $C(f)$ the cyclic subspace generated by $f,$ that is, $C(f)=\overline {\textrm {span}}(U^n f: n\in \mathbb Z)$ . With this notation, one form of the spectral theorem reads as follows (see for example [Reference Conway19, Ch. IX], [Reference Parry47, Appendix, §2] or [Reference Queffélec49]): there exists a finite or infinite sequence of vectors $f_i\in H, 0\leq i<m$ , where $m\in \mathbb N\cup \{\infty \}$ , such that

$$ \begin{align*}H=\bigoplus_{0\leq i< m} C(f_i)\quad\textrm{ and }\quad\sigma_{f_0}\gg\sigma_{f_1}\gg\cdots\gg\sigma_{f_i}\gg\cdots.\end{align*} $$

Moreover, these measures are essentially unique in the following sense: for any other sequence $(g_i)_{0\leq i<m'}$ satisfying $H=\bigoplus _{0\leq i<m'}C(g_i)$ and ${\sigma _{g_0}\gg \sigma _{g_1}\gg \cdots \gg \sigma _{g_i}\gg \cdots }$ , we have $m'=m$ and $\sigma _{f_i}\sim \sigma _{g_i}$ for all $i.$ The maximal spectral type of T is then defined as (the equivalence class of) the measure $\sigma _{f_0}$ .

Observe that for any $f\in H,$ the restriction of U to $C(f)$ is unitary equivalent to the multiplication operator $M_{\sigma _f}:L^2(\mathbb {T},\sigma _f)\to L^2(\mathbb {T},\sigma _f)$ defined by $M_{\sigma _f}u (z):= zu(z)$ , $z\in \mathbb T$ .

Suppose now that U has countable Lebesgue spectrum and let $(f_i)_{i\in \mathbb {N}}$ be a sequence in H such that $\{U^n f_i:\ n\in \mathbb Z,\ i\in \mathbb {N}\}$ is an orthonormal basis of $H.$ Then, $H=\bigoplus _{i\in \mathbb {N}}C(f_i)$ and $\sigma _{f_i}=\unicode{x3bb} $ for all $i\in I$ . Hence, U is unitarily equivalent to $M_{\unicode{x3bb} }^{(\infty )}:=\bigoplus _{i=1}^\infty M_\unicode{x3bb} $ acting on $\bigoplus _{i=1}^\infty L^2(\mathbb {T},\unicode{x3bb} )$ . Conversely, it is clear that if $U\cong M_{\unicode{x3bb} }^{(\infty )}$ , then U has countable Lebesgue spectrum.

By the uniqueness part of the spectral theorem, it follows in particular that the maximal spectral type of a unitary operator with countable Lebesgue spectrum is the Lebesgue measure. The next lemma is a kind of converse.

Lemma 3.10. Let U be a unitary operator on a complex separable Hilbert space H satisfying the following two conditions.

  1. (a) There exists a closed subspace $K\subset H$ such that $U(K)=K$ and $U_{| K}:K\to K$ has countable Lebesgue spectrum.

  2. (b) The maximal spectral type of U is the Lebesgue measure.

Then, $U:H\to H$ has countable Lebesgue spectrum.

Proof. We will use the following notation: if $\sigma $ is a positive finite Borel measure on $\mathbb {T}$ and $N\in \mathbb {N}\cup \{\infty \}$ , we denote by $M_{\sigma }^{(N)}$ the operator $\bigoplus _{0\leq j<N} M_\sigma $ acting on $\bigoplus _{0\leq j<N} L^2(\mathbb {T},\sigma )$ . Recall also that $\unicode{x3bb} $ is the Lebesgue measure on $\mathbb {T}$ .

By condition (a), we know that

$$ \begin{align*} U_{| K}\cong M_{\unicode{x3bb}}^{(\infty)}.\end{align*} $$

Consider now the operator $U_{| K^\perp }:K^\perp \to K^\perp $ . By condition (b), its maximal spectral type is absolutely continuous with respect to the Lebesgue measure $\unicode{x3bb} $ . By the ‘second formulation’ of the spectral theorem (see for example [Reference Queffélec49] or [Reference Conway19, Ch. IX]), there exist pairwise disjoint Borel sets $\Delta _\infty ,\Delta _1,\Delta _2,\ldots $ and positive finite measures $\mu _\infty ,\mu _1,\mu _2,\ldots $ on $\mathbb {T}$ supported on $\Delta _\infty ,\Delta _1,\Delta _2,\ldots $ and absolutely continuous with respect to $\unicode{x3bb} $ , such that $U_{|K^\perp }$ is unitarity equivalent to $M_{\mu _\infty }^{(\infty )}\oplus M_{\mu _1}^{(1)}\oplus M_{\mu _2}^{(2)}\oplus \cdots .$ Note that some of the measures $\mu _n$ may be $0$ . For $n\in \{ \infty \}\cup \mathbb {N}$ , we may write $\mu _n=f_n\, \unicode{x3bb} $ for some non-negative function $f_n\in L^1(\mathbb {T})$ and we may assume that $\Delta _n=\{ f_n>0\}$ . Hence, if we set ${\nu _n:=\mathbf {1}_{\Delta _n}\, \unicode{x3bb} ,}$ we get $M_{\mu _n}\cong M_{\nu _n}$ since the measures $\mu _n$ and $\nu _n$ are equivalent. Therefore,

$$ \begin{align*} U_{|K^\perp}\cong M_{\nu_\infty}^{(\infty)}\oplus M_{\nu_1}^{(1)}\oplus M_{\nu_2}^{(2)}\oplus\cdots.\end{align*} $$

Now, let $E:=\mathbb {T}\backslash (\Delta _\infty \cup \bigcup _{n\geq 1}\Delta _n)$ and $\nu :=\mathbf {1}_{E}\, \unicode{x3bb} .$ Then, $\unicode{x3bb} =\nu +\nu _\infty +\sum \limits _{n\in \mathbb {N}}\nu _n$ so that

$$ \begin{align*} M_\unicode{x3bb}\cong M_\nu\oplus M_{\nu_\infty}\oplus M_{\nu_1}\oplus \cdots.\end{align*} $$

Consequently, we obtain

$$ \begin{align*} U&\cong U_{|K}\oplus U_{| K^\perp}\\ &\cong M_\nu^{(\infty)}\oplus M_{\nu_\infty}^{(\infty)}\oplus M_{\nu_1}^{(\infty)}\oplus M_{\nu_2}^{(\infty)}\oplus \cdots\\ &\qquad\qquad \oplus M_{\nu_\infty}^{(\infty)}\oplus M_{\nu_1}^{(1)}\oplus M_{\nu_2}^{(2)}\oplus\cdots\\ &\cong M_\nu^{(\infty)} \oplus M_{\nu_\infty}^{(\infty)}\oplus M_{\nu_1}^{(\infty)}\oplus M_{\nu_2}^{(\infty)} \cdots\\ &\cong M_\unicode{x3bb}^{(\infty)}, \end{align*} $$

which means that U has countable Lebesgue spectrum.

The following observation is useful to check assumption (a) in Lemma 3.10.

Lemma 3.11. Let $(f_i)_{i\in I}$ be a countably infinite family of vectors in H. Assume that the cyclic subspaces $C(f_i)$ are pairwise orthogonal and let $K:=\oplus _{i\in I} C(f_i)$ . If $\sigma _{f_i}$ is equivalent to the Lebesgue measure for all $i\in I$ , then $U_{| K}$ has countable Lebesgue spectrum.

Proof. For each $i\in I$ , there exists $g_i\in C(f_i)$ such that $\sigma _{g_i}$ is exactly equal to the Lebesgue measure and for any such $g_i$ , we have $C(g_i)=C(f_i)$ ; see for example [Reference Parry47, pp. 93–94]. Hence, $\{U^ng_i\,:\, i\in I, n\in \mathbb {Z}\}$ is an orthonormal basis of K.

We will also need the following fact.

Lemma 3.12. Let U be a unitary operator acting on a separable Hilbert space H and let $\sigma $ be a finite, positive Borel measure on $\mathbb {T}$ . If there exists a dense set $\mathcal D\subset H$ such that $\sigma _f\ll \sigma $ for all $f\in \mathcal D$ , then $\sigma _f\ll \sigma $ for all $f\in H$ .

Proof. Denote by $M(\mathbb {T})$ the space of all complex Borel measures on $\mathbb {T}$ . By [Reference Queffélec49, Corollary 2.2], the map $f\mapsto \sigma _f$ is continuous from H into $M(\mathbb {T})$ endowed with the norm topology. The result follows immediately.

3.6.2. Proof of Proposition 3.6

Before really starting the proof, we need to recall some basic facts concerning the $L^2$ -space of a Gaussian measure. In what follows, $\mu $ is a Gaussian measure on the (separable) complex Fréchet space X.

Let $\mathcal G:=\overline {\textrm {span}}\,(\langle x^*,\cdot \rangle :\ x^*\in X^*)$ , where the closure is taken in $L^2(X,\mathcal B,\mu )$ . This is a Gaussian subspace of $L^2(X,\mathcal B,\mu )$ , in the sense that any function in $\mathcal G$ has symmetric complex Gaussian distribution. Moreover, $\mathcal G$ is clearly $U_T$ -invariant.

For $k\geq 0,$ let us denote by $\mathcal G^k$ the space of homogeneous polynomials of degree k in the elements of $\mathcal G$ , with $\mathcal G^0=\mathbb C.$ The subspaces $\mathcal G^k$ , $k\geq 0$ , are linearly independent (which is not obvious) and one can orthonormalize them thanks to the so-called Wick transform $f\mapsto \; \pmb :{f}\pmb :$ which is defined on $ \bigcup _{k\geq 0} \mathcal G^k$ as follows:

$$ \begin{align*} \pmb:{f}\pmb:\;=\begin{cases} f&\text{if } f \text{ is constant},\\ f- P_k f\quad&\text{if } f\in\mathcal G^k, k\geq 1, \end{cases} \end{align*} $$

where we denote by $ P_k$ the orthogonal projection onto ${\overline {\mathrm {span}}}\,({\mathcal {G}}^i:0\leq i\leq k-1).$

With this notation, we have the orthogonal decomposition

$$ \begin{align*} L^2(X,{\mathcal{B}},\mu)=\bigoplus_{k=0}^{\infty}\overline{\pmb:{{\mathcal{G}}^k}\pmb:}.\end{align*} $$

Moreover, for any $f_1,\ldots ,f_k, g_1,\ldots ,g_k\in \mathcal G$ , the scalar product $ \langle \pmb :{f_1\cdots f_k}\pmb :\; ,\, \pmb :{g_1\cdots g_k}\pmb :\rangle $ is given by

(3.1) $$ \begin{align} \langle \pmb:{f_1\cdots f_k}\pmb:\; ,\, \pmb:{g_1\cdots g_k}\pmb:\rangle=\sum_{\mathfrak s\in\mathfrak S_k}\langle f_{\mathfrak s(1)},g_1\rangle\cdots \langle f_{\mathfrak s(k)},g_k\rangle,\end{align} $$

where $\mathfrak S_k$ is the permutation group of $\{ 1,\ldots ,k\}$ .

For details on these general facts, we refer to [Reference Peller48, Ch. 8] or [Reference Janson34, Chs. 2 and 3]. We will also need the following lemma (see [Reference Bayart and Grivaux5, Lemma 3.28]).

Lemma 3.13. Let $T\in \mathfrak L(X)$ and let $\mu $ be a T-invariant Gaussian measure on X. For any $f_1,\ldots ,f_k\in \mathcal G$ , we have

$$ \begin{align*} U_T(\pmb:{f_1\cdots f_k}\pmb:)=\; \pmb:{(U_Tf_1)\cdots (U_Tf_k)}\pmb:.\end{align*} $$

In particular, each subspace $\pmb :{\mathcal G^k}\pmb :$ is $U_T$ -invariant.

We can now finally give the proof of Proposition 3.6.

Proof of Proposition 3.6

Let $(X,\mathcal B,\mu , T)$ be a measure-preserving dynamical system, where $T\in \mathcal (X)$ is invertible and the measure $\mu $ is Gaussian, such that for any $x^*\in X^*\subset L^2(\mu )$ , the spectral measure $\sigma _{x^*}$ with respect to $U_T$ is absolutely continuous with respect to the Lebesgue measure. We want to show that $(X,\mathcal B,\mu , T)$ has countable Lebesgue spectrum.

Let us first recall that for any $f,g\in L^2(\mu )$ , there is a unique complex Borel measure $\sigma _{f,g}$ on $\mathbb {T}$ with Fourier coefficients

$$ \begin{align*} \widehat{\sigma_{f,g}}(n)=\langle U_T^n f,g\rangle,\quad n\in\mathbb Z.\end{align*} $$

If $f=g$ , then $\sigma _{f,f}$ is the spectral measure $\sigma _f$ ; and the existence of $\sigma _{f,g}$ for arbitrary functions f and g follows from a polarization argument. Indeed, we have

$$ \begin{align*} \sigma_{f,g}= \frac14\sum_{k=0}^3 \mathbf i^k\sigma_{f+\mathbf i^k g}.\end{align*} $$

This formula and the assumption of the proposition show that $\sigma _{x^*, y^*}$ is absolutely continuous with respect to the Lebesgue measure for any $x^*, y^*\in X^*$ . Note also that the map $(f,g)\mapsto \sigma _{f,g}$ is obviously $\mathbb {R}$ -bilinear.

In what follows, we denote by $\star $ the convolution product for measures on $\mathbb {T}$ and we write the elements of $X^*$ as $f,g, \ldots $ rather than $x^*, y^*, \ldots $ to avoid the proliferation of stars.

Fact 3.14. If $f_1, \ldots ,f_k\in X^*$ , then $ \sigma _{\pmb :{f_1\cdots f_k}\pmb :}= \sum _{\mathfrak s\in \mathfrak S_k} \sigma _{f_{\mathfrak s(1)}, f_1}\star \cdots \star \sigma _{f_{\mathfrak s(k)},f_k}$ .

Proof of Fact 3.14

By Lemma 3.13 and equation (3.1), we have for all $n\in \mathbb {Z}$ :

$$ \begin{align*} \widehat{\sigma}_{\pmb:{f_1\cdots f_k}\pmb:}(n)&=\langle \pmb:{(U_T^nf_1)\cdots (U_T^nf_k)}\pmb:\, ,\, \pmb:{f_1\cdots f_k}\pmb:\rangle\\ &=\sum_{\mathfrak s\in\mathfrak S_k} \langle U_T^nf_{\mathfrak s(1)},f_1\rangle\cdots \langle U_T^nf_{\mathfrak s(k)},f_k\rangle\\ &= \sum_{\mathfrak s\in\mathfrak S_k} \widehat\sigma_{f_{\mathfrak s(1)}, f_1}(n)\cdots \widehat\sigma_{f_{\mathfrak s(k)}, f_k}(n).\\[-3.9pc] \end{align*} $$

Now, let us denote by $\mathcal D$ the set of all functions $f\in L^2(\mu )$ of the form

$$ \begin{align*} f=\sum_{k=1}^N \pmb:{\sum_{j=1}^{m_k}f_{j,1}\cdots f_{j,k}}\pmb:\quad\text{with } m_k\geq 1 \text{ and } f_{j,i}\in X^*.\end{align*} $$

This is a dense linear subspace of $L^2_0(\mu )$ .

Fact 3.15. If $f\in \mathcal D$ , then $\sigma _f$ is absolutely continuous with respect to the Lebesgue measure.

Proof of Fact 3.15

Let $f\in \mathcal D$ , so that $ f=\sum _{k=1}^N f_k$ , where

$$ \begin{align*} f_k= \sum_{j=1}^{m_k}\pmb:{f_{j,1}\cdots f_{j,k}}\pmb:\quad\text{with } f_{j,i}\in X^*, \quad\text{so that } f_k\in {\pmb:{\mathcal G^k}\pmb:}.\end{align*} $$

By orthogonality and $U_T$ -invariance of the subspaces $\pmb :{\mathcal G^k}\pmb :$ , we have for all $n\in \mathbb {Z}$ ,

$$ \begin{align*}\widehat{\sigma_f}(n)=\langle U_T^n f,f\rangle=\sum_{k=1}^N \langle U_T^n f_k,f_k\rangle=\sum_{k=1}^N \widehat{\,\sigma_{f_k}}(n),\end{align*} $$

so that $\sigma _f= \sigma _{f_1}+\cdots +\sigma _{f_N}$ . Hence, it is enough to check that each measure $\sigma _{f_k}$ is absolutely continuous with respect to the Lebesgue measure. However, this is clear by Fact 3.14 and bilinearity of the map $(f,g)\mapsto \sigma _{f,g}$ : indeed, we have

$$ \begin{align*} \sigma_{f_k}= \sum_{j,j'=1}^N\sigma_{ \pmb:{f_{j,1}\cdots f_{j,k}}\pmb:\,,\, \pmb:{f_{j',1}\cdots f_{j',k}}\pmb:}=\sum_{j,j'=1}^N \sum_{\mathfrak s\in\mathfrak S_k} \sigma_{f_{j,\mathfrak s(1)},f_{j',1}}\star\cdots\star \sigma_{f_{j,\mathfrak s(k)},f_{j',k}} \end{align*} $$

and the result follows since all measures $\sigma _{f_{j,\mathfrak s(i)},f_{j',i}}$ are absolutely continuous with respect to the Lebesgue measure.

We can now finish the proof of Proposition 3.6. By Fact 3.15 and Lemma 3.12, the maximal spectral type of $U_{T}$ is absolutely continuous with respect to the Lebesgue measure. Now, take any $f\in X^*\setminus \{ 0\}$ . Then, $\sigma _f$ is a non-zero measure which is absolutely continuous with respect to the Lebesgue measure. It follows that there exists $k_0\geq 1$ such that, for all $k\geq k_0,$ the measure $\sigma _{:f^k:}=k!\, \sigma _f\star \cdots \star \sigma _f$ (k times) is equivalent to the Lebesgue measure (see for example [Reference Newton and Parry45, Lemma 3.6], where the symmetry assumption on the measure is in fact not necessary). Let us set

$$ \begin{align*}K:=\overline{\textrm{span}}\,\{U_{T}^n (\pmb:{f^k}\pmb:) : \ n\in\mathbb Z,\ k\geq k_0\}.\end{align*} $$

By orthogonality and $U_T$ -invariance of the spaces $\pmb :{\mathcal G^k}\pmb :$ , and applying Lemma 3.11, we see that $(U_T)_{| K}$ has countable Lebesgue spectrum. Hence, by Lemma 3.10, $U_{T}$ has countable Lebesgue spectrum.

3.7. The frequent hypercyclicity criterion again

The tools introduced in the previous sections allow us to give another proof of Theorem 2.1. Arguably, this proof is much less elementary. Let us say that an operator $T\in \mathfrak L(X)$ has countable Lebesgue spectrum after extension if there exists a T-invariant Borel probability measure $\mu $ on X with full support such that $(X,\mathcal B,\mu ,T)$ is a factor of a measure-preserving dynamical system which has countable Lebesgue spectrum. By Corollary 3.4, any such operator T is hereditarily frequently hypercyclic.

Proposition 3.16. Let $T\in \mathfrak L(X)$ be an operator satisfying the FHCC. Then, T has countable Lebesgue spectrum after extension.

Proof. It is shown in [Reference Murillo-Arcila and Peris44] that there exists a T-invariant measure $\mu $ on X with full support such that $(X,\mathcal B,T,\mu )$ is a factor of a Bernoulli shift; and it is well known that Bernoulli shifts have countable Lebesgue spectrum (see for example [Reference Walters57, Theorems 4.30 and 4.33]).

One can also prove the following ‘probabilistic’ version of Proposition 3.16. Let us say that a sequence $(x_n)_{n\in \mathbb {Z}}$ is a bilateral backward orbit for an operator T if $Tx_n=x_{n-1}$ for all $n\in \mathbb {Z}$ .

Proposition 3.17. Let X be a complex Fréchet space and let $T\in \mathfrak L(X)$ . Assume that there exists a bilateral backward orbit $(x_n)_{n\in \mathbb {Z}}$ for T such that $\overline {\mathrm {span}}( x_n\,:\, n\in \mathbb {Z})=X$ and the series $\sum g_n x_n$ is almost surely convergent, where $(g_n)$ is a sequence of independent complex standard Gaussian variables. Then, T has countable Lebesgue spectrum after extension. More precisely, there exists a T-invariant Gaussian measure $\mu $ on X with full support such that $(X,\mathcal B, \mu , T)$ is a factor of a measure-preserving dynamical system with countable Lebesgue spectrum.

The (almost sure) convergence of the bilateral series $\sum g_n x_n$ means that both series

$$ \begin{align*}\sum_{n\geq 0} g_n x_n \quad \textrm{ and }\quad\sum_{n<0} g_n x_n\end{align*} $$

are (almost surely) convergent.

Proof. Let $\mu $ be the distribution of the random variable $\xi :=\sum _{n\in \mathbb {Z}} g_n x_n$ . This is a Gaussian measure, which has full support since $\overline {\mathrm {span}}( x_n\,:\, n\in \mathbb {Z})=X$ and which is T-invariant because $(x_n)$ is a bilateral backward orbit for T. By Corollary 3.7, it is enough to show that for any $x^*\in X^*\subset L^2(\mu )$ , the spectral measure $\sigma _{x^*}$ of $x^*$ with respect to $U_T$ is absolutely continuous with respect to the Lebesgue measure.

By orthogonality of the Gaussian variables $g_k$ , we have for all $n\geq 0$ ,

$$ \begin{align*} \widehat{\sigma_{x^*}}(n)=\langle U_T^n x^*, x^*\rangle&=\sum_{k\in\mathbb{Z}} \langle x^*, T^n x_k\rangle\,\overline{\langle x^*, x_k\rangle}\\ &=\sum_{k\in\mathbb{Z}} \langle x^*, x_{k-n}\rangle\,\overline{\langle x^*, x_k\rangle}. \end{align*} $$

The series is absolutely convergent since the almost sure convergence of the scalar Gaussian series $\sum \langle x^*, x_k\rangle \, g_k$ implies that $\sum _{k\in \mathbb {Z}} \vert \langle x^*, x_k\rangle \vert ^2<\infty $ .

Now, let $\varphi \in L^2(\mathbb {T})$ be the function with Fourier coefficients $\widehat \varphi (k):= \langle x^*, x_{-k}\rangle $ , $k\in \mathbb Z$ , that is,

$$ \begin{align*} \varphi(z)\sim \sum_{k\in\mathbb{Z}} \langle x^*, x_{-k}\rangle \,z^k;\end{align*} $$

and let $g:=\vert \varphi \vert ^2\in L^1(\mathbb {T})$ . By definition, we have for all $n\geq 0$ ,

$$ \begin{align*} \widehat g(n)=\sum_{k\in\mathbb{Z}} \widehat{\overline\varphi}(k)\,\widehat{\varphi}(n-k)=\sum_{k\in\mathbb{Z}} \overline{\langle x^*, x_k\rangle}\, \langle x^*, x_{k-n}\rangle.\end{align*} $$

Since two positive measures on $\mathbb {T}$ with the same non-negative Fourier coefficients must be equal, it follows that $\sigma _{x^*}= g(\unicode{x3bb} )\, d\unicode{x3bb} $ , which concludes the proof.

Remark 3.18. The above proof shows in particular that if $(x_n)_{n\in \mathbb {Z}}$ is a bilateral backward orbit for T such that the series $\sum g_n x_n$ is almost surely convergent, then the distribution of the random variable $\xi :=\sum _{n\in \mathbb {Z}} g_n x_n$ is a strongly mixing measure for T. This is not specific to Gaussian variables: as shown in [Reference Agneessens1], the same result holds true if $(g_n)$ is replaced by any sequence of independent, identically distributed random variables.

We mentioned above that Proposition 3.17 is a probabilistic version of Proposition 3.16; let us be a little bit more explicit. The following fact (which was observed independently by A. López-Martínez) can be extracted from the proof of [Reference Agneessens1, Theorem 4.9].

Fact 3.19. Let X be a Fréchet space. If $T\in \mathfrak L(X)$ satisfies the FHCC, then there exists a bilateral backward orbit $(x_n)$ for T such that $\overline {\mathrm {span}}( x_n\,:\, n\in \mathbb {Z})=X$ and the series $\sum x_n$ is unconditionally convergent.

In view of that, the next result is an improvement of Proposition 3.16 when X is a Banach space with non-trivial cotype.

Corollary 3.20. Let X be a Banach space with non-trivial cotype and let $T\in \mathfrak L(X)$ . Assume that there exists a bilateral backward orbit $(x_n)_{n\in \mathbb {Z}}$ for T such that $\overline {\mathrm {span}}( x_n\,:\, n\in \mathbb {Z})=X$ and the series $\sum \pm x_n$ is convergent for almost every choice of signs $\pm $ . Then, T has countable Lebesgue spectrum after extension.

Proof. By assumption on X, almost sure convergence of the Rademacher series $\sum \pm x_n$ is equivalent to almost sure convergence of the Gaussian series $\sum g_n x_n$ (this follows from [Reference Maurey and Pisier40, Corollaire 1.3, p. 67]; see also for example [Reference Ledoux and Talagrand35, Proposition 9.14]). So, the result follows immediately from Proposition 3.17.

Remark 3.21. In the same spirit, Proposition 3.17 can be used to show that if an operator T satisfies the ‘probabilistic frequent hypercyclicity criterion’ of [Reference Grivaux28, Theorem 1.2], then T is hereditarily frequently hypercyclic. Indeed, by [Reference Grivaux28, Lemma 3.2], there is a bilateral backward orbit $(x_n)_{n\in \mathbb {Z}}$ for T satisfying the assumption of Proposition 3.17.

We now show that, at least for operators on complex Hilbert spaces, Theorem 3.1 is ‘strictly stronger’ than Theorem 2.1. This answers a very natural question asked by the referee.

Proposition 3.22. Let X be a complex Fréchet space.

  1. (1) If $T\in \mathfrak L(X)$ satisfies the FHCC, then there exists a family $(E_i)_{i\in I}$ of continuous $\mathbb {T}$ -eigenvector fields for T such that $\overline {\mathrm {span}}\,\bigcup _{i\in I} E_i(\mathbb {T})=X$ . In particular, this family is spanning with respect to the Lebesgue measure $\unicode{x3bb} $ , so that if X is a Banach space with type $2$ , one can apply Theorem 3.1 to conclude that T is hereditarily frequently hypercyclic.

  2. (2) If X is a Hilbert space, there exists $T\in \mathfrak L(X)$ which admits a $\unicode{x3bb} $ -spanning $\mathbb {T}$ -eigenvector field (so that T is hereditarily frequently hypercyclic by Theorem 3.1) but does not satisfy the FHCC. In fact, one can even require that T has no periodic vector except $0$ .

Proof. (1) This is well known, see for example [Reference Bayart and Matheron8, §8.1].

(2) Let $\Lambda $ be any compact subset of $\mathbb {T}$ with empty interior, but such that all its non-empty (relative) open subsets have positive Lebesgue measure (to obtain such a set, take a compact set $\Lambda _0$ with empty interior and positive Lebesgue measure, and remove all open subsets of $\Lambda _0$ with Lebesgue measure $0$ ). Let $T_\Lambda :H_\Lambda \to H_\Lambda $ be the Kalish operator associated with $\Lambda $ (see for example [Reference Bayart and Matheron7, Ch. 5] for the definition). The operator $T_\Lambda $ has the following properties: its point spectrum is equal to $\Lambda $ and it admits a continuous $\mathbb {T}$ -eigenvector field $E:\Lambda \to H_\Lambda $ such that $\overline {\mathrm {span}}\, E(\Lambda )=H_\Lambda $ . By continuity and since $\Lambda \setminus N$ is dense in $\Lambda $ for every set $N\subset \mathbb {T}$ with Lebesgue measure $0$ , this eigenvector field is $\unicode{x3bb} $ -spanning. However, T does not satisfy the FHCC. Indeed, assume otherwise. The continuous $\mathbb {T}$ -eigenvector fields $E_i:\mathbb {T}\to H_\Lambda $ given by item (1) cannot vanish identically outside $\Lambda $ (unless they are identically $0$ on $\mathbb {T}$ ) since $\mathbb {T}\setminus \Lambda $ is dense in $\mathbb {T}$ ; so $T_\Lambda $ has eigenvalues outside $\Lambda $ , which is a contradiction since $\sigma _p(T_\Lambda )=\Lambda $ . Moreover, if $\Lambda $ contains no roots of unity, then $T_\Lambda $ has no periodic vector except $0$ since no root of unity can be an eigenvalue of $T_\Lambda $ . Recall that any operator (on a real or complex space) satisfying the FHCC is chaotic, see for example [Reference Bayart and Matheron7, Theorem 6.10].

Finally, the ergodic-theoretic point of view allows us to give an ‘extreme’ example of a hereditarily frequently hypercyclic operator T which does not satisfy the FHCC.

Example 3.23. There exists a separable complex Banach space X supporting a hereditarily frequently hypercyclic operator with no eigenvalues.

Proof. Let T be the Kalish operator acting on the space $X:=C_0([0,2\pi ])$ of continuous functions on $[0,2\pi ]$ vanishing at the point $0$ (see [Reference Bayart and Grivaux5, Example 4.2] or [Reference Bayart and Matheron7, §5.5.4]). The operator T has no eigenvalues. However, T admits a Gaussian invariant measure $\mu $ with full support such that all the spectral measures $\sigma _{x^*}$ , $x\in X^*$ relative to the Koopman operator $U_T:L^2(\mu )\to L^2(\mu )$ are absolutely continuous with respect to the Lebesgue measure; so T has countable Lebesgue spectrum after extension by Corollary 3.7 and hence it is hereditarily frequently hypercyclic.

3.8. Perfect spanning and hereditary UFHC

The links between properties of unimodular eigenvectors of an operator T and frequent hypercyclicity of T have been very much studied since [Reference Bayart and Grivaux5]. The strongest available result may be the following (see [Reference Bayart and Matheron8, Reference Grivaux29]).

Let X be a separable complex Fréchet space and let $T\in \mathfrak L(X)$ . If the $\mathbb {T}$ -eigenvectors of T are perfectly spanning, then T is frequently hypercyclic and, in fact, there exists a Gaussian T-invariant measure $\mu $ with full support such that T is weakly mixing with respect to $\mu $ .

The perfect spanning assumption means that for any countable set $N\subset \mathbb {T}$ , the eigenvectors of T with eigenvalues in $\mathbb {T}\setminus N$ span a dense linear subspace of X; equivalently (see [Reference Grivaux and Matheron30, Proposition 6.1]), the $\mathbb {T}$ -eigenvectors of T are $\sigma $ -spanning for some continuous probability measure $\sigma $ on $\mathbb {T}$ . It is plausible that under this assumption, the operator T is, in fact, hereditarily frequently hypercyclic; but we are very far from being able to prove that. We would be already happy enough if we could weaken the assumptions of Theorem 3.1 and prove that for any complex Banach space X, an operator $T\in \mathfrak L(X)$ is hereditarily frequently hypercyclic as soon as the $\mathbb {T}$ -eigenvectors of T are spanning with respect to the Lebesgue measure—but again, this seems out of reach for the moment. However, we do have the following result.

Theorem 3.24. Let X be a complex Fréchet space and let $T\in \mathfrak L(X)$ . If the $\mathbb {T}$ -eigenvectors of T are perfectly spanning, then T is hereditarily $\mathcal U$ -frequently hypercyclic.

For the proof, we will need the following variant of [Reference Conze20, Lemme 5].

Lemma 3.25. Let $(X, \mathcal B,\mu , T)$ be a measure-preserving dynamical system and assume that T is weakly mixing with respect to $\mu $ . Let also $(n_k)_{k\geq 1}$ and $(k_i)_{i\geq 0}$ be two increasing sequence of integers. Assume that $n_{k_i}=O(k_i)$ as $i\rightarrow \infty $ . Then, for any measurable set $V\subset X$ ,

$$ \begin{align*} \frac1{k_i}\sum_{k=1}^{k_i} \mathbf 1_V\circ T^{n_k}\xrightarrow{L^2} \mu(V)\quad\text{as } i\to\infty.\end{align*} $$

Proof. The proof is similar to that of the classical Blum–Hanson theorem [Reference Blum and Hanson15]. We have

$$ \begin{align*} \bigg\Vert \frac1{k_i}\sum_{k=1}^{k_i} \mathbf 1_V\circ T^{n_k}-\mu(V)\bigg\Vert_2^2&= \frac1{k_i^2} \sum_{r,s=1}^{k_i} ( \mu( T^{-n_r}(V)\cap T^{-n_s}(V)) -\mu(V)^2)\\ &=\frac2{k_i^2} \sum_{1\leq r<s\leq k_i} ( \mu( V\cap T^{-(n_s-n_r)}(V)) -\mu(V)^2) +O\bigg(\frac1{k_i}\bigg). \end{align*} $$

So it is enough to check that

(3.2) $$ \begin{align} \sum_{1\leq r<s\leq k_i} \vert \mu( V\cap T^{-(n_s-n_r)}(V)) -\mu(V)^2\vert=o(k_i^2). \end{align} $$

In what follows, we put

$$ \begin{align*} \gamma_{s,r} :=\vert\mu( V\cap T^{-(n_s-n_r)}(V)) -\mu(V)^2\vert.\end{align*} $$

Since T is weakly mixing with respect to $\mu $ , there is a set $D\subseteq \mathbb {N}$ with $\mathrm {dens} (D)=1$ such that

(3.3) $$ \begin{align} \mu( V\cap T^{-d}(V))-\mu(V)^2\to 0\quad\text{as } d\to\infty, d\in D.\end{align} $$

Write

$$ \begin{align*} \sum_{1\leq r<s\leq k_i} \gamma_{s,r}=\sum_{r=1}^{k_i}\sum_{\substack{r<s\leq k_i \\ n_s-n_r\in D}} \gamma_{s,r}+ \sum_{r=1}^{k_i}\sum_{\substack{r<s\leq k_i \\ n_s-n_r\not\in D}} \gamma_{s,r}=:\alpha_i+\beta_i. \end{align*} $$

Using equation (3.3) and since $\gamma _{r,s}\leq 1$ for all $r,s$ , it is not hard to check that

$$ \begin{align*}\sum_{\substack{r<s\leq k_i \\ n_s-n_r\in D}} \gamma_{s,r}=o(k_i)\quad\text{ as } i\to\infty, \text{ uniformly in } r;\end{align*} $$

and it follows that $\alpha _i=o(k_i^2)$ . Moreover, since $\mathrm {dens}(D)=1$ , we see that

$$ \begin{align*} \beta_i\leq \sum_{r=1}^{k_i} \# \{ s\in (r,k_i] : n_s-n_r\not\in D\} \leq k_i\times \# ( (0,n_{k_i}]\setminus D) =k_i\times o(n_{k_i});\end{align*} $$

so $\beta _i=o(k_i^2)$ since we are assuming that $n_{k_i}=O(k_i)$ . This proves equation (3.2).

Corollary 3.26. Let $(X, \mathcal B,\mu , T)$ be a measure-preserving dynamical system and assume that T is weakly mixing with respect to $\mu $ . Let also $A\subset \mathbb {N}$ with $\overline {\mathrm {dens}}(A)>0$ . If $V\subset X$ is a measurable set such that $\mu (V)>0$ , then $\overline {\mathrm {dens}}(A\cap \mathcal N_T(x,V))>0$ for $\mu $ -almost every $x\in X$ .

Proof. Let $(n_k)_{k\geq 1}$ be the increasing enumeration of A. Since $\overline {\mathrm {dens}}(A)>0$ , one can find an increasing sequence of integers $(k_i)_{i\geq 0}$ such that $n_{k_i}=O(k_i)$ . By Lemma 3.25, one can find a subsequence $(k^{\prime }_i)$ of $(k_i)$ such that $(1/{k^{\prime }_i})\sum _{k=1}^{k^{\prime }_i} \mathbf 1_V\circ T^{n_k}\to \mu (V) \mu $ -almost everywhere. In other words: for $\mu $ -almost every $x\in X$ ,

$$ \begin{align*} \frac1{k^{\prime}_i}\, \#\{ k\in [1, k^{\prime}_i] : n_k\in \mathcal N_T(x,V)\}\to \mu(V).\end{align*} $$

Since $\#\{ k\in [1, k^{\prime }_i] : n_k\in \mathcal N_T(x,V)\}=\#( [1, n_{k^{\prime }_i}]\cap A\cap \mathcal N_T(x,V))$ and $n_{k^{\prime }_i}=O(k^{\prime }_i)$ , it follows that $\overline {\mathrm {dens}}(A\cap \mathcal N_T(x,V))>0$ for $\mu $ -almost every $x\in X$ .

Proof of Theorem 3.24

Assume that the $\mathbb {T}$ -eigenvectors of T are perfectly spanning. By [Reference Bayart and Matheron8], there exists a T-invariant Gaussian measure $\mu $ on X with full support such that T is weakly mixing with respect to $\mu $ . Let $(A_i)_{i\in I}$ be a countable family of subsets of $\mathbb {N}$ with positive upper density and let $(V_i)_{i\in I}$ be a family of non-empty open subsets of X. It follows immediately from Corollary 3.26 that one can find $x\in X$ (in fact, $\mu $ -almost every $x\in X$ will do) such that $\overline {\mathrm {dens}}(A_i\cap \mathcal N_T(x,V_i))>0$ for all $i\in I$ .

Let us point out one consequence of Theorem 3.24.

Corollary 3.27. If X is a complex Banach space admitting an unconditional Schauder decomposition, then X supports a hereditarily $\mathcal U$ -frequently hypercyclic operator.

Proof. It is shown in [Reference De la Rosa, Frerick, Grivaux and Peris22] that such a space X supports an operator with a perfectly spanning set of $\mathbb {T}$ -eigenvectors.

Remark 3.28. The proof of Theorem 3.24 shows the following: if $T\in \mathfrak L(X)$ and if there exists a T-invariant Borel measure with full support $\mu $ on X such that T is weakly mixing with respect to $\mu $ , then T is hereditarily $\mathcal U$ -frequently hypercyclic. It would be interesting to know if the weak mixing assumption can be replaced by ergodicity. Incidentally, we do not know any example of an operator T admitting an ergodic measure with full support but no weakly mixing measure with full support.

4. The $T_1\oplus T_2$ frequent hypercyclicity problem

One of the most intriguing open problems regarding frequent hypercyclicity is to decide whether $T\oplus T$ is frequently hypercyclic whenever T is frequently hypercyclic [Reference Bayart and Grivaux5]. Note that, by [Reference Ernst, Esser and Menet24], the corresponding question for $\mathcal U$ -frequent hypercyclicity has a positive answer. A related problem is Question 1.1, which asks whether $T_1\oplus T_2$ is frequently hypercyclic for every frequently hypercyclic operator $T_1$ and $T_2$ . This question is also open if we replace frequent hypercyclicity by $\mathcal {U}$ -frequent hypercyclicity. As observed in [Reference Grivaux, Matheron and Menet31], $T_1\oplus T_2$ is hypercyclic as soon as $T_1$ and $T_2$ are $\mathcal U$ -frequently hypercyclic. In the opposite direction, it seems that the best known result is [Reference Grivaux, Matheron and Menet31, Theorem 7.33] which deals with infinite sums: there exists a sequence $(T_n)_{n\geq 1}$ of frequently hypercyclic operators on $\ell _p(\mathbb {N})$ , $p>1$ , such that the operator $T=\bigoplus _{n\geq 1}T_n$ acting on the $\ell _p$ -sum $X=\bigoplus _{n\geq 1}\ell _p(\mathbb {N})$ is not $\mathcal U$ -frequently hypercyclic.

As mentioned in §1, things are much simpler if we consider hereditarily frequently hypercyclic operators. We now give the detailed proof for the convenience of the reader.

Proposition 4.1. Let $\mathcal F\subset 2^{\mathbb {N}}$ be a Furstenberg family and let $X_1,X_2$ be two Polish topological vector spaces. Let also $T_1\in \mathfrak L(X_1)$ and $T_2\in \mathfrak L(X_2)$ . If $T_1$ is $\mathcal F$ -hypercyclic and $T_2$ is hereditarily $\mathcal F$ -hypercyclic, then $T_1\oplus T_2$ is $\mathcal F$ -hypercyclic. If both $T_1$ and $T_2$ are hereditarily $\mathcal F$ -hypercyclic, then $T_1\oplus T_2$ is hereditarily $\mathcal F$ -hypercyclic.

Proof. Assume that $T_1$ is $\mathcal F$ -hypercyclic and $T_2$ is hereditarily $\mathcal F$ -hypercyclic. Let $(V_i)_{i\in I}$ be a countable basis of open sets for $X_1\times X_2$ . Without loss of generality, we may assume that each $V_i$ has the form $V_i=V_{i,1}\times V_{i,2}$ , where $V_{i,1}, V_{i,2}$ are open in $X_1, X_2$ . Let ${x_1\in X_1}$ be any $\mathcal F$ -hypercyclic vector for $T_1$ . Then, for each $i\in I$ , the set $A_i:=\mathcal N_{T_1}(x_1, V_{i,1})$ belongs to $\mathcal F$ . Since $T_2$ is hereditarily $\mathcal F$ -hypercyclic, it follows that one can find a vector $x_2\in X_2$ such that $B_i:=A_i\cap \, \mathcal N_{T_2}(x_2, V_{i,2})\in \mathcal F$ for all $i\in I$ . Then, $x:=(x_1,x_2)$ is a frequently hypercyclic vector for $T:=T_1\oplus T_2$ since $\mathcal N_T(x, V_i)\supset B_i$ for all $i\in I$ .

The proof of the second part of the proposition is essentially the same.

Using weighted backward shifts on $c_0(\mathbb Z_+)$ , we now find a counterexample to the ${T_1\oplus T_2}$ frequent hypercyclicity problem, and thus answer Question 1.1 in the negative. This counterexample also solves the $T_1\oplus T_2 \mathcal {U}$ -frequent hypercyclicity problem.

Theorem 4.2. There exist two frequently hypercyclic weighted shifts $B_w, B_{w'}$ on $c_0(\mathbb Z_+)$ such that $B_w\oplus B_{w'}$ is not $\mathcal U$ -frequently hypercyclic.

From this theorem and Proposition 4.1, we immediately deduce the following result, which is of course to be compared with Corollary 2.3.

Corollary 4.3. There exist weighted shifts on $c_0(\mathbb Z_+)$ which are frequently hypercyclic but not hereditarily frequently hypercyclic.

Let us also point out another consequence of Theorem 4.2 and Remark 3.28.

Corollary 4.4. There exist frequently hypercyclic weighted shifts on $c_0(\mathbb {Z}_+)$ which admit no weakly mixing invariant measure with full support, and hence no ergodic invariant Gaussian measure with full support.

This is not really a new result: the Gaussian part has been known since [Reference Bayart and Grivaux6] (with arguably a more complicated example than the one we are about to present here) and it was proved in [Reference Grivaux and Matheron30] that there exist frequently hypercyclic bilateral weighted shifts on $c_0(\mathbb {Z})$ which admit no ergodic invariant measure with full support.

In the proof of Theorem 4.2, we shall use the following lemma, which gives a simple characterization of frequent hypercyclicity for weighted shifts on $c_0(\mathbb Z_+)$ whose weight sequence is bounded below (see [Reference Bayart and Ruzsa9] or [Reference Bonilla and Grosse-Erdmann17, Corollary 34]).

Lemma 4.5. Let $w=(w_n)_{n\geq 1}$ be a bounded sequence of positive real numbers and assume that $\inf _{n\geq 1} w_n>0$ . Then, the associated weighted shift $B_w$ is frequently hypercyclic on $c_0(\mathbb {Z}_+)$ if and only if there exist a sequence $(M(p))_{p\geq 1}$ of positive real numbers tending to infinity and a sequence $(E_p)_{p\geq 1}$ of disjoint subsets of $\mathbb {N}$ with positive lower density such that:

  1. (a) $\lim _{n\to \infty ,\ n\in E_p}w_1\cdots w_n=\infty $ for all $p\geq 1$ ;

  2. (b) for all $p,q\geq 1,$ for all $m\in E_p$ and $n\in E_q$ with $m>n,$

    $$ \begin{align*}w_1\cdots w_{m-n}\geq \max(M(p),M(q)).\end{align*} $$

We will also need the following elementary lemma, which is almost the same as [Reference Bayart and Ruzsa9, Lemma 6.1]. For $\varepsilon>0$ , $a>1$ and $u\in \mathbb {N},$ we let

$$ \begin{align*} I_u^{a,\varepsilon}:=[(1-\varepsilon)a^u,(1+\varepsilon)a^u].\end{align*} $$

Lemma 4.6. There exist $\varepsilon>0$ and $a>1$ such that, for any integers $u>v\geq 1,$

$$ \begin{align*} I_u^{a,4\varepsilon}\cap I_v^{a,4\varepsilon}=\emptyset,\quad I_u^{a,2\varepsilon}-I_v^{a,2\varepsilon}\subset I_u^{a,4\varepsilon}\quad\mathrm{and}\quad I_v^{a,\varepsilon}+[-v,v]\subset I_v^{a,2\varepsilon}. \end{align*} $$

Proof. Provided that $\varepsilon \in (0,1/4),$ the first condition is equivalent to saying that

$$ \begin{align*}(1+4\varepsilon)a^u<(1-4\varepsilon)a^{u+1}\quad\text{for all } u\geq 1,\end{align*} $$

that is,

$$ \begin{align*} \frac{1+4\varepsilon}{(1-4\varepsilon)a}<1.\end{align*} $$

The second one is satisfied as soon as

$$ \begin{align*}(1-2\varepsilon)a^u-(1+2\varepsilon)a^{u-1}\geq (1-4\varepsilon)a^u\quad\text{for all } u\geq 2,\end{align*} $$

which is equivalent to

$$ \begin{align*} \frac{2\varepsilon a}{1+2\varepsilon}\geq 1.\end{align*} $$

The last condition is satisfied if $(1-\varepsilon )a^v-v\geq (1-2\varepsilon )a^v$ for all $v\geq 1,$ in other words,

$$ \begin{align*}\varepsilon a^v\geq v\quad\text{for all } v\geq 1.\end{align*} $$

Therefore, one can choose, for example, $\varepsilon :=1/8$ and then take a sufficiently large.

Proof of Theorem 4.2

Our construction is inspired by that of [Reference Bayart and Ruzsa9, §6]. In what follows, we fix once and for all $\varepsilon>0$ and $a>1$ satisfying the conclusions of Lemma 4.6.

For $k\geq 1,$ let

$$ \begin{align*} A_k:=2^{k-1}\mathbb N\backslash 2^k\mathbb{N}.\end{align*} $$

Note that each $A_k$ is a syndetic set, that is, it has bounded gaps, and the sets $A_k$ are pairwise disjoint. Moreover, since $2^{k-1}\geq k$ , we have $I_v^{a,\varepsilon }+[-k,k]\subset I_v^{a,2\varepsilon }$ for each $k\geq 1$ and all $v\in A_k$ .

We also fix an increasing sequence of positive integers $(b_p)_{p\geq 1}$ such that

$$ \begin{align*}\lim_{p\to\infty}\overline{\textrm{dens}} \bigg[\bigcup_{q\geq p} ( b_q\mathbb{N}+[-q,q])\bigg]=0.\end{align*} $$

Finally, we set for $p\geq 1,$

$$ \begin{align*} E_p:=&\bigcup_{u\in A_{2p}}(I_u^{a,\varepsilon}\cap b_{p}\mathbb{N}),\\ F_p:=&\bigcup_{u\in A_{2p+1}}(I_u^{a,\varepsilon}\cap b_{p}\mathbb{N}). \end{align*} $$

We note that since $A_{2p}$ and $A_{2p+1}$ are syndetic, we have

$$ \begin{align*}\underline{\textrm{dens}}(E_p)>0\quad\mathrm{and}\quad \underline{\textrm{dens}}(F_p)>0\quad\text{for all } p\geq 1.\end{align*} $$

Indeed, for all $p\geq 1,$ there exists $\delta _p>0$ such that, for all u sufficiently large,

$$ \begin{align*}\#(I_{u}^{a,\varepsilon}\cap b_p\mathbb{N})\geq \delta_p a^u.\end{align*} $$

Let $R_p$ be such that if u and v are two consecutive elements of $A_{2p}$ , then $v-u\leq R_p.$ If now n is very large and if we consider u and v as two consecutive elements of $A_{2p}$ such that

$$ \begin{align*}(1+\varepsilon)a^u<n \leq (1+\varepsilon)a^v,\end{align*} $$

then we see that

$$ \begin{align*}\frac{\#(E_p\cap [1,n])}{n}\geq \frac{\#(I_u^{a,\varepsilon}\cap b_p\mathbb{N})}{(1+\varepsilon)a^v}\geq\frac{\delta_p}{(1+\varepsilon)a^{R_p}}\cdot\end{align*} $$

Hence, $\underline {\textrm {dens}}(E_p)>0$ . A similar argument shows that $\underline {\textrm {dens}}(F_p)>0$ .

We now construct our weight sequences w and $w'$ .

For $p\geq 1,$ we first define a sequence $(w_n^p)_{n\geq 1}\subset \{1/2,1, 2\}$ such that, for all $n\geq 1,$

$$ \begin{align*}w_1^p\cdots w_n^p=\begin{cases} 1&\textrm{if }n\notin I_u^{a,2\varepsilon}, u\in A_{2p},\\ 2^u&\textrm{if }n\in I_u^{a,\varepsilon}, u\in A_{2p}. \end{cases}\end{align*} $$

This is possible since $I_u^{a,\varepsilon }+[-u,u]\subset I_u^{a,2\varepsilon }$ . These sequences will be used for handling condition (a) in Lemma 4.5.

For $p\geq 1,$ we also define a sequence $(\omega _n^p)_{n\geq 1}\subset \{1/2,1, 2\}$ such that, for all $n\geq 1,$

$$ \begin{align*}\omega_1^p\cdots \omega_n^p=\begin{cases} 1&\textrm{if }n\notin b_{p}\mathbb{N}+[-p,p],\\ 2^p&\textrm{if }n\in b_{p}\mathbb{N}. \end{cases}\end{align*} $$

These sequences will help us to verify condition (b) in Lemma 4.5 for $p=q$ .

Finally, for $u>v\geq 1$ with $u\in A_{2p}$ and $v\in A_{2q}$ for some $p,q\geq 1,$ we define a sequence $(w_n^{u,v})_{n\geq 1}\subset \{1/2,1, 2\}$ such that, for all $n\geq 1,$

$$ \begin{align*}w_1^{u,v}\cdots w_n^{u,v}=\begin{cases} 1&\textrm{if }n\notin I_u^{a,4\varepsilon},\\ \max(2^p,2^q)&\textrm{if }n\in I_u^{a,\varepsilon}-I_v^{a,\varepsilon}. \end{cases}\end{align*} $$

This is possible since

$$ \begin{align*} I_u^{a,\varepsilon}-I_v^{a,\varepsilon}+[-\max(p,q),\max(p,q)]&\subset (I_u^{a,\varepsilon}+[-p,p])-(I_v^{a,\varepsilon}+[-q,q])\\ &\subset I_u^{a,2\varepsilon}-I_v^{a,2\varepsilon}\\ & \subset I_u^{a,4\varepsilon}. \end{align*} $$

These sequences will be needed to check condition (b) in Lemma 4.5 for $p\neq q$ .

We finally define the weight sequence $w=(w_n)_{n\geq 1}$ as follows: for all $n\geq 1$ , we require that

$$ \begin{align*}w_1\cdots w_n=\max_{p,u>v}(w_1^p\cdots w_n^p,\omega_1^p\cdots \omega_n^p,w_1^{u,v}\cdots w_n^{u,v}).\end{align*} $$

It is not difficult to check that $w_n\in [1/2, 2]$ for all $n\geq 1$ . Indeed, assume for instance that $w_1\cdots w_n=w_1^p\cdots w_n^p.$ Then, $w_1\cdots w_{n-1}\geq w_1^p\cdots w_{n-1}^p$ and

$$ \begin{align*}w_n\leq w_n^p\leq 2.\end{align*} $$

The same argument works for the other cases; for the lower bound, assume for instance that $w_1\cdots w_{n-1}=w_1^p\cdots w_{n-1}^p.$ Then, $w_1\cdots w_n\geq w_1^p\cdots w_n^p$ so that

$$ \begin{align*}w_n\geq w_n^p\geq \tfrac 12.\end{align*} $$

We define in a similar way a weight sequence $w'=(w^{\prime }_n)_{n\geq 1}\subset [1/2,2]$ , replacing everywhere $A_{2p}$ by $A_{2p+1}$ and $A_{2q}$ by $A_{2q+1}$ .

Let us first show that w and $w'$ satisfy the conditions of Lemma 4.5, so that $B_w$ and $B_{w'}$ are frequently hypercyclic. It is clearly enough to do that for w.

  1. (a) If $n\in E_p$ , there is a unique $u=u(n)\in A_{2p}$ such that $n\in I_u^{a,\varepsilon }$ . Then, $w_1\cdots w_n\geq w_1^p\cdots w_n^p\geq 2^{u(n)}$ , which shows that $w_1\cdots w_n\to \infty $ as $n\to \infty $ , $n\in E_p$ .

  2. (b) Let $p,q\geq 1,$ , and let us fix $m\in E_p$ and $n\in E_q$ with $m>n$ . If $p=q$ , then $m-n\in b_p \mathbb {N}$ , so that $w_1\cdots w_{m-n}\geq \omega _1^p\cdots \omega _{m-n}^p\geq 2^p.$ If $p\neq q$ , there exist ${u>v\geq 1}$ such that $m\in I_{u}^{a,\varepsilon }$ and $n\in I_v^{a,\varepsilon }$ , and then $w_1\cdots w_{m-n}\geq w_1^{u,v}\cdots w_{m-n}^{u,v}\geq \max (2^p,2^q).$

Now, let us show that $B_w\oplus B_{w'}$ is not $\mathcal U$ -frequently hypercyclic. Denote by $(e_j)_{j\ge 0}$ the canonical basis of $c_0(\mathbb Z_+)$ . We show that for any vector $x\in c_0(\mathbb Z_+)\oplus c_0(\mathbb Z_+)$ , the set $E_x:=\{n\in \mathbb {N}:\ \|(B_w\oplus B_{w'})^nx-(e_0,e_0)\|<1/2\}$ has upper density equal to $0$ .

Towards a contradiction, assume that $\overline {\mathrm {dens}}(E_x)>0$ for some vector x. It is easy to check that

$$ \begin{align*}\lim_{n\to\infty,\ n\in E_x}w_1\cdots w_n =\lim_{n\to\infty,\ n\in E_x}w^{\prime}_1\cdots w^{\prime}_n=\infty.\end{align*} $$

It follows that if we set

$$ \begin{align*}G_p:=\{n\in\mathbb{N}:\ w_1\cdots w_n\geq 2^p\textrm{ and }w^{\prime}_1\cdots w_n'\geq 2^p\},\end{align*} $$

then $E_x\setminus G_p$ is finite and hence $ \overline {\textrm {dens}}(G_p)\geq \overline {\textrm {dens}}(E_x)>0$ for all $p\geq 1$ .

Now, the construction of the weight sequence w yields that if $n\in G_p$ , then either $n\in b_q\mathbb {N}+[-q,q]$ for some $q\geq p$ , or $n\in I_u^{a,4\varepsilon }$ for some $u\in \bigcup _{q\geq 1}A_{2q}$ . Similarly, by construction of the sequence $w'$ , we also know that if $n\in G_p$ , then either $n\in b_q\mathbb {N}+[-q,q]$ for some $q\geq p$ , or $n\in I_v^{a,4\varepsilon }$ for some $v\in \bigcup _{q\geq 1}A_{2q+1}$ . By disjointness of the sets $I_u^{a,4\varepsilon }$ and $I_v^{4,\varepsilon }$ for $u\neq v,$ it follows that

$$ \begin{align*}G_p\subset\bigcup_{q\geq p}( b_q\mathbb{N}+[-q,q]).\end{align*} $$

By our choice of the sequence $(b_p),$ we get a contradiction with $\overline {\textrm {dens}}(G_p)\geq \overline {\textrm {dens}}(E_x)>0$ .

Remark 4.7. The weighted shift $B_w$ cannot serve as a counterexample to the $T\oplus T$ frequent hypercyclicity problem. Indeed, it can be shown (see [Reference Grosse-Erdmann32, Theorem 18]) that any weighted shift on $c_0(\mathbb {Z}_+)$ satisfying the assumptions of Lemma 4.5 is such that any finite direct sum $B_w\oplus \cdots \oplus B_w$ is itself frequently hypercyclic.

5. FHC operators on $\ell _p({\mathbb Z}_+)$ which are not hereditarily FHC

5.1. The result

In this section, we use the machinery developed in [Reference Grivaux, Matheron and Menet31], following the construction in [Reference Menet41] of chaotic operators which are not frequently hypercyclic, to produce an operator on $\ell _p(\mathbb {Z}_+)$ , $1\leq p<\infty $ , which is frequently hypercyclic but not hereditarily frequently hypercyclic. We will, in fact, obtain a formally stronger result.

Definition 5.1. Let $A\subset \mathbb {N}$ be a set with $\underline {\text {dens}}(A)>0$ . We say that an operator $T\in \mathfrak L(X)$ is frequently hypercyclic along A if the sequence $(T^n)_{n\in A}$ is frequently hypercyclic: there exists $x\in X$ such that $\underline {\mathrm {dens}}\,( {A\cap \mathcal N_T(x,V)})>0$ for every non-empty open set $V\subset X$ .

Obviously, if an operator is hereditarily frequently hypercyclic, then it is frequently hypercyclic along any set $A\subset \mathbb {N}$ with positive lower density. Our aim is to prove the following theorem.

Theorem 5.2. Let $1\le p<\infty $ . There exist an operator T on $\ell _p(\mathbb {Z}_+)$ and a set $A\subset \mathbb {N}$ with $\underline {\text {dens}}(A)>0$ such that T is frequently hypercyclic and chaotic, but not frequently hypercyclic along A (and thus not hereditarily frequently hypercyclic).

With the terminology of [Reference Grivaux, Matheron and Menet31], the operator we are looking for will be a C $_{+,1}$ -type operator. So we will need to recall the definition of C $_{+,1}$ -type operators, and more generally of C-type and C $_{+}$ -type operators. However, before that, we will prove a general result allowing to check in a simple way that an operator is not frequently hypercyclic along some set with positive lower density.

5.2. How not to be hereditarily FHC

The next theorem gives simple conditions ensuring that an operator is not hereditarily frequently hypercyclic.

Theorem 5.3. Let X be a Banach space admitting a Schauder basis $(e_k)_{k\ge 0}$ and let ${T\in \mathfrak L(X)}$ . Denoting by $\pi _K$ , $K\geq 1$ the canonical projection onto $\mathrm {span}(e_k\,:\, 0\leq k\leq K-1)$ , assume that there exist increasing sequences of integers $(K_n)_{n\ge 1}$ and $(J_n)_{n\ge 1}$ such that for every n:

  1. (a) $T^{J_n} \pi _{K_n}=\pi _{K_n}$ ;

  2. (b) $\|\pi _{K_n}T^j (I-\pi _{K_n})x\|\le \|(I-\pi _{K_n})x\|$ for all $x\in X$ and $0\le j\le (n+1)^{2^{J_n}}J_n$ .

Then, there exists a set $A\subset \mathbb {N}$ with positive lower density such that T is not hereditarily frequently hypercyclic along A.

Proof. Extracting subsequences of $(K_n)$ and $(J_n)$ if necessary, we may assume without loss of generality that $J_{n+1}\ge 2(n+1)^{2^{J_n}}J_n$ for all n.

Let us denote by $\mathcal {F}_n\subset 2^{\mathbb {N}}$ the family of all finite sets $F\subset [0,J_n)$ such that $\#F\ge J_n/2$ . Let $C_n:=\#\mathcal {F}_n$ and let $(F_{n,j})_{0\le j< C_n}$ be an enumeration of $\mathcal {F}_n$ . We set $M_n:=(n+1)^{C_{n}}J_n$ and we remark that $M_n\le (n+1)^{2^{J_n}}J_n\le J_{n+1}/2$ .

We now construct the set A by induction. To start, let

$$ \begin{align*}A_1:=[0,J_1)\cup \bigcup_{0\le j< C_1}\bigcup_{0\le l<2^{j+1}-2^j}((2^j+l)J_1+F_{1,j}).\end{align*} $$

Given $k\ge 2$ , if $A_1,\ldots , A_{k-1}$ have been defined already, we define $A_k$ by setting

$$ \begin{align*} A_{k}:=[M_{k-1},J_{k})\; \cup \; \bigcup_{0\le j<C_{k}}\bigcup_{0\le l<(k+1)^{j+1}-(k+1)^j}((k+1)^j+l)J_{k}+F_{k,j}). \end{align*} $$

Observe that the sets $(k+1)^j+l)J_{k}+F_{k,j}$ involved in this definition are pairwise disjoint, since they are contained in successive intervals. More precisely,

$$ \begin{align*} A_k\cap [sJ_{k}, (s+1)J_{k})= sJ_{k}+F_{k,j}\end{align*} $$

for every $0\le j<C_k$ and every $(k+1)^{j}\le s<(k+1)^{j+1}$ , and $\max (A_k)< (k+1)^{C_k} J_k= M_k$ . Hence, $A_k\subseteq [M_{k-1}, M_k)$ . Finally, we let

$$ \begin{align*} A:=\bigcup_{k\ge 1} A_k. \end{align*} $$

Claim 5.4. We have ${\#(A\cap [0,N])}/({N+1})\ge 1/4$ for all $N\geq 0$ . In particular, $\underline {\textrm {dens}}(A)>0$ .

Proof of Claim 5.4

We will check by induction on $k\ge 1$ that

(5.1) $$ \begin{align} \frac{\#( A\cap [0,N])}{(N+1)}\ge \frac{1}{4} \quad\textrm{ for all } N<M_k. \end{align} $$

If $N<M_1$ , there exists $0\le s < 2^{C_1}$ such that $s J_1\le N<(s+1)J_1$ . Then, ${\#(A\cap [0,N])}/({N+1})\ge 1$ if $s=0$ and

$$ \begin{align*} \frac{\#(A\cap [0,N])}{N+1}\ge \frac{\#(A_1\cap [0,sJ_1])}{(s+1)J_1}\ge \frac{s}{2(s+1)}\ge \frac{1}{4}\quad \text{if } s\ge 1, \end{align*} $$

because the sets involved in the definition of $A_1$ are pairwise disjoint and $\#F_{1,j}\ge J_1/2$ for every $j<C_1$ .

Assume that the inequality (5.1) has been proved for $k-1$ . To get the result for k, it is enough to check that

$$ \begin{align*}\frac{\#(A\cap [0,N])}{N+1}\ge \frac{1}{4} \quad\textrm{ for every } M_{k-1}\le N<M_{k}.\end{align*} $$

If $M_{k-1}\le N< J_{k}$ then, since $[ M_{k-1}, N]\subset [M_{k-1},J_{k})\subset A_{k}$ , we get by the induction assumption that

$$ \begin{align*}\frac{\#(A\cap [0,N])}{N+1}\ge \frac{\#(A\cap [0,M_{k-1}))}{M_{k-1}}\geq \frac{1}{4}\cdot\end{align*} $$

Moreover, we even have ${\#(A\cap [0,J_{k}))}/{J_{k}}\ge \tfrac 12$ since $[M_{k-1},J_{k})\subset A_{k}$ and $M_{k-1}\le J_{k}/2$ . However, if $J_{k}\le N<M_{k}$ , there exists $1\le s < (k+1)^{C_{k}}$ such that $s J_{k}\le N<(s+1)J_{k}$ and we obtain in this case that

$$ \begin{align*}\frac{\#(A\cap [0,N])}{N+1}\ge \frac{\#(A\cap [0,J_{k}))}{(s+1)J_{k}}+\frac{\#(A_{k}\cap [J_{k},sJ_{k}])}{(s+1)J_{k}}\ge \frac{1}{2(s+1)}+\frac{s-1}{2(s+1)}\ge \frac{1}{4}.\end{align*} $$

This proves Claim 5.4.

Let us now get back to the proof of Theorem 5.3. Our aim is to show that under assumptions (a) and (b), T is not hereditarily frequently hypercyclic along the set A that we just constructed. To this end, we consider for any $c>0$ , the open sets

$$ \begin{align*} U_c:=\{y\in X\,:\, |\langle e^*_0, y\rangle|<c\}\quad\mathrm{and}\quad V_c=\{y\in X\,:\, |\langle e^*_0, y\rangle|>c\},\end{align*} $$

and we show that for every $x\in X$ , either $\mathcal N_T(x,U_{1/2})\cap A$ or $\mathcal N_T(x,V_{3/2})\cap A$ has a lower density equal to $0$ . Let $x\in X$ . Since $\vert \langle e_0^*, u\rangle \vert \leq C\, \Vert \pi _{K_n} u\Vert $ for some absolute constant $C>0$ , it follows from assumption (b) that for n sufficiently large, we have

$$ \begin{align*} \mathcal N_T(x,U_{1/2})\cap [0,(n+1)^{2^{J_n}}J_n)&\subset \mathcal N_T(\pi_{K_n}x,U_{3/4})\cap [0,(n+1)^{2^{J_n}}J_n)\\ &\subset \mathcal N_T(x,U_{1})\cap [0,(n+1)^{2^{J_n}}J_n) \end{align*} $$

and

$$ \begin{align*} \mathcal N_T(x,V_{3/2})\cap [0,(n+1)^{2^{J_n}}J_n)&\subset \mathcal N_T(\pi_{K_n}x,V_{4/3})\cap [0,(n+1)^{2^{J_n}}J_n)\\ &\subset \mathcal N_T(x,V_{1})\cap [0,(n+1)^{2^{J_n}}J_n). \end{align*} $$

Moreover, since $\mathcal N_T(x,U_1)\cap \, \mathcal N_T(x,V_1)=\emptyset $ , we have, for any $n\geq 1$ , that

$$ \begin{align*}\mathrm{either}\ \#(\mathcal N_T(x,U_{1})\cap [0,J_n))\le J_n/2 \quad\textrm{ or }\quad \#(\mathcal N_T(x,V_{1})\cap [0,J_n))\le J_n/2.\end{align*} $$

Hence, either $\#(\mathcal N_T(x,U_{1}){\kern-1pt}\cap{\kern-1pt} [0,J_n)){\kern-1pt}\le{\kern-1pt} J_n/2$ for infinitely many n values or $\#( \mathcal N_T(x,{\kern-1pt} V_{1}){\kern-1pt}\cap [0,J_n))\le J_n/2$ for infinitely many n values. Without loss of generality, we assume that $\#( \mathcal N_T(x,U_{1})\cap [0,J_n))\le J_n/2$ for infinitely many n values (the other case being similar). Hence, there exists an increasing sequence $(n_k)_{k\ge 1}$ of integers and a sequence $(j_k)_{k\ge 1}$ of integers such that

$$ \begin{align*}\mathcal N_T(x,U_{1})\cap [0,J_{n_k})\cap F_{n_k,j_k}=\emptyset \quad\textrm{ for every } k\ge 1.\end{align*} $$

Now, since $T^{J_{n_k}} \pi _{K_{n_k}}=\pi _{K_{n_k}}$ by assumption (a), we have for every $k\ge 1$ ,

$$ \begin{align*} \mathcal N_T(x,U_{1/2})\cap [&0,(n_k+1)^{2^{J_{n_k}}}J_{n_k})\\ &\quad\subset \;\mathcal N_T(\pi_{K_{n_k}}x,U_{3/4})\cap [0,(n_k+1)^{2^{J_{n_k}}}J_{n_k})\\ &\quad=\bigcup_{0\le l<(n_k+1)^{2^{J_{n_k}}}} (lJ_{n_k}+ \mathcal N_T(\pi_{K_{n_k}}x,U_{3/4})\cap[0,J_{n_k}) )\\ &\quad \subset \bigcup_{0\le l<(n_k+1)^{2^{J_{n_k}}}} (lJ_{n_k}+ \mathcal N_T(x,U_{1})\cap[0,J_{n_k}) ). \end{align*} $$

Intersecting with A and observing that $(n_k+1)^{j_k+1}\le (n_k+1)^{2^{J_{n_k}}}$ , we get that

$$ \begin{align*} (\mathcal N_T(x,U_{1/2})\cap A)\cap [0,(&n_k+1)^{j_k+1}J_{n_{k}})\\ &\quad\subset A\cap \bigcup_{0\le s<(n_k+1)^{j_k+1}} (sJ_{n_k}+\mathcal N_T(x,U_{1})\cap[0,J_{n_k})). \end{align*} $$

Now, by definition of A, we have

$$ \begin{align*}A\cap \bigcup_{(n_k+1)^{j_k}\le s<(n_k+1)^{j_k+1}}[sJ_{n_k}, (s+1)J_{n_k})= \bigcup_{(n_k+1)^{j_k}\le s<(n_k+1)^{j_k+1}}(sJ_{n_k}+F_{n_k,j_k}).\end{align*} $$

Since $\mathcal N_T(x,U_{1})\cap [0,J_{n_k})\cap F_{n_k,j_k}=\emptyset $ , it follows that

$$ \begin{align*} (\mathcal N_T(x,U_{1/2})\cap A)\cap [(n_k+1)^{j_k}J_{n_k},(n_k+1)^{j_k+1}J_{n_{k}}) =\emptyset, \end{align*} $$

so that

$$ \begin{align*} (\mathcal N_T(x,U_{1/2})\cap A)\cap [0,(n_k+1)^{j_k+1}J_{n_{k}}) \subset [0,(n_k+1)^{j_k}J_{n_k}). \end{align*} $$

So we see that

$$ \begin{align*} \frac{\# ((\mathcal N_T(x,U_{1/2})\cap A)\cap [0,(n_k+1)^{j_k+1}J_{n_{k}}))}{(n_k+1)^{j_k+1}J_{n_{k}}}\le \frac{(n_k+1)^{j_k}J_{n_k}}{(n_k+1)^{j_k+1}J_{n_{k}}}\cdot\end{align*} $$

The right-hand side of this inequality tends to $0$ as n tends to infinity, and this shows that $\underline {\textrm {dens}}(\mathcal N_T(x,U_{1/2})\cap A)=0$ .

Remark 5.5. Assumption (b) in Theorem 5.3 can be weakened: it is enough to assume that there exists a non-zero linear functional $x^*\in X^*$ such that $\vert \langle x^*, T^j (I-\pi _{K_n})x\rangle \vert \le \|(I-\pi _{K_n})x\|$ for all $x\in X$ and $j\le (n+1)^{2^{J_n}}J_n$ . This is apparent from the above proof.

5.3. C-type operators

We recall here very succinctly some basic facts concerning C-type operators, and we refer the reader to [Reference Grivaux, Matheron and Menet31, §§6 and 7] for more on this class of operators. In what follows, we denote by $(e_k)_{k\geq 0}$ the canonical basis of $\ell _p(\mathbb {Z}_+)$ , $1\le p<\infty $ .

Let us consider four ‘parameters’ v, w, $\varphi $ and b, where:

  • $v=(v_{n})_{n\ge 1}$ is a sequence of non-zero complex numbers such that $\sum _{n\ge 1}|v_{n}|<\infty $ ;

  • $w=(w_{j})_{j\geq 1}$ is a sequence of complex numbers such that $0<\inf _{k\ge 1} \vert w_k\vert \leq \sup _{k\ge 1}\vert w_k\vert <\infty $ ;

  • $\varphi $ is a map from $\mathbb {Z}_+$ into itself, such that $\varphi (0)=0$ , $\varphi (n)<n$ for every $n\ge 1$ , and the set $\varphi ^{-1}(l)=\{n\ge 0\,;\,\varphi (n)=l\}$ is infinite for every $l\ge 0$ ;

  • $b=(b_{n})_{n\ge 0}$ is a strictly increasing sequence of positive integers such that $b_{0}=0$ and $b_{n+1}-b_{n}$ is a multiple of $2(b_{\varphi (n)+1}-b_{\varphi (n)})$ for every $n\ge 1$ .

If w and b are such that

$$ \begin{align*} \inf_{n\geq 0} \prod_{b_n<j<b_{n+1}}\, \vert w_j\vert>0,\end{align*} $$

then, by [Reference Grivaux, Matheron and Menet31, Lemma 6.2], there is a unique bounded operator $T_{v,\,w,\, \varphi , \,b\,}$ on $\ell _p(\mathbb {Z}_+)$ such that

$$ \begin{align*} T_{v,\,w,\, \varphi, \,b\,}\ e_k= \begin{cases} w_{k+1}\,e_{k+1} & \textrm{if}\ k\in [b_{n},b_{n+1}-1),\; n\geq 0,\\ v_{n}\,e_{b_{\varphi(n)}}-\bigg(\displaystyle\prod_{j=b_{n}+1}^{b_{n+1}-1} w_{j}\bigg)^{ -1 } e_ { b_{n}} & \textrm{if}\ k=b_{n+1}-1,\ n\ge 1,\\ -\bigg(\displaystyle\prod_{j=b_0+1}^{b_{1}-1}w_j\bigg)^{-1}e_0& \textrm{if}\ k=b_1-1. \end{cases} \end{align*} $$

Any such operator $T_{v,\,w,\, \varphi , \,b\,}$ is called a C-type operator. A notable fact to be pointed out immediately is that C-type operators have lots of periodic points; indeed, we have the following fact, which is [Reference Grivaux, Matheron and Menet31, Lemma 6.4].

Fact 5.6. If $T=T_{v,\,w,\, \varphi , \,b\,}$ is a C-type operator, then

$$ \begin{align*}T^{2(b_{n+1}-b_n)}e_k=e_k\quad\text{if } k\in [b_n,b_{n+1}), \; n\geq 0.\end{align*} $$

It follows that every finitely supported vector is periodic for $T_{v,\,w,\, \varphi , \,b\,}$ ; in particular, a C-type operator is chaotic as soon as it is hypercyclic.

A C $_{+}$ -type operator is a C-type operator for which the parameters satisfy the following additional conditions: for every $k\geq 1$ :

  • $\varphi $ is increasing on each interval $[2^{k-1},2^{k})$ with $\varphi ([2^{k-1},2^{k}))=[0,2^{k-1})$ , that is,

    $$ \begin{align*}\varphi (n)=n-2^{k-1}\quad\text{for every } n\in[2^{k-1},2^{k});\end{align*} $$
  • the blocks $[b_{n},b_{n+1})$ , $n\in [2^{k-1},2^{k})$ all have the same size, which we denote by $\Delta ^{(k)}$ :

    $$ \begin{align*}b_{n+1}-b_{n}= \Delta ^{(k)}\quad\text{for every } n\in[2^{k-1},2^{k}); \end{align*} $$
  • the sequence v is constant on the interval $[2^{k-1}, 2^{k})$ : there exists $v^{(k)}$ such that

    $$ \begin{align*} v_{n}=v^{(k)}\quad\text{for every } n\in[2^{k-1},2^{k}); \end{align*} $$
  • the sequences of weights $(w_{b_{n}+i})_{1\le i <\Delta ^{(k)}}$ are independent of $n\in [2^{k-1},2^{k})$ : there exists a sequence $(w_{i}^{(k)})_{1\le i<\Delta ^{(k)}}$ such that

    $$ \begin{align*} w_{b_{n}+i}=w_{i}^{(k)} \quad\text{for every } n\in[2^{k-1},2^{k}) \text{ and } 1\le i<\Delta ^{(k)}. \end{align*} $$

Finally, a C $_{+,1}$ -type operator is a C $_{+}$ -type operator whose parameters are such that for all $k\geq 1$ ,

$$ \begin{align*} v^{(k)}=2^{-\tau ^{(k)}}\quad\textrm{and}\quad w_{i}^{(k)}= \begin{cases} 2&\textrm{if}\ \ 1\le i\le \delta ^{(k)},\\ 1&\textrm{if}\ \ \delta ^{(k)}<i<\Delta ^{(k)}, \end{cases} \end{align*} $$

where $(\tau ^{(k)})_{k\ge 1}$ and $(\delta ^{(k)})_{k\ge 1}$ are two increasing sequences of integers with $\delta ^{(k)}<\Delta ^{(k)}$ for each $k\ge 1$ .

These operators have been studied in detail in [Reference Grivaux, Matheron and Menet31, §7]. In particular, we have the following crucial fact [Reference Grivaux, Matheron and Menet31, Theorem 7.1].

Fact 5.7. A C $_{+,1}$ -type operator $T_{v,\,w,\, \varphi , \,b\,}$ is frequently hypercyclic as soon as

(5.2) $$ \begin{align} \limsup\limits_{k\to\infty }\, \dfrac{\delta ^{(k)} -\tau ^{(k)}}{\Delta ^{(k)}}>0. \end{align} $$

5.4. Proof of Theorem 5.2

Let $T=T_{v,\,w,\, \varphi , \,b\,}$ be an operator of C $_{+,1}$ -type on $\ell _{p}(\mathbb {Z}_+)$ , so that v and w are given by

$$ \begin{align*} v^{(k)}=2^{-\tau ^{(k)}}\quad\textrm{and}\quad w_{i}^{(k)}= \begin{cases} 2&\textrm{if}\ \ 1\le i\le \delta ^{(k)},\\ 1&\textrm{if}\ \ \delta ^{(k)}<i<\Delta ^{(k)}. \end{cases} \end{align*} $$

We assume that $\Delta ^{(k)}\in 8\mathbb {N}$ for all $k\ge 1$ , and we choose

$$ \begin{align*}\delta^{(k)}:=\tfrac{1}{4}\Delta^{(k)} \quad \text{and} \quad \tau^{(k)}:=\tfrac{1}{8}\Delta^{(k)}.\end{align*} $$

So the only ‘free’ parameter is now the sequence $(\Delta ^{(k)})_{k\geq 1}$ .

By Fact 5.7, the operator T is frequently hypercyclic (and hence chaotic since it is a C-type operator). So we just have to show that if the sequence $(\Delta ^{(k)})$ is suitably chosen, then T satisfies the assumptions of Theorem 5.3. We will in fact show that this holds as soon as the sequence $(\Delta ^{(k)})$ grows sufficiently rapidly. Let us set

$$ \begin{align*} K_n:=b_{2^{n}}\quad\mathrm{and}\quad J_n:=2\Delta^{(n)}\quad\textrm{ for every }n\ge 1.\end{align*} $$

With this choice of the sequences $(K_n)$ and $(J_n)$ , condition (a) in Theorem 5.3 is satisfied by Fact 5.6. So the only thing to check is condition (b).

Let $\gamma _k:=2^{\,\delta ^{(k-1)}-\tau ^{(k)}}(\Delta ^{(k)})^{1-1/{p}}$ . If $(\Delta ^{(k)})$ grows sufficiently rapidly, then the sequence $(\gamma _k)$ is decreasing and

$$ \begin{align*}2^n \sum_{k\ge n+1}2^{k-1}\gamma_k\le 1\quad\text{for all } n\geq 0.\end{align*} $$

Let us also define a sequence $(\beta _l)_{l\geq 1}$ as follows:

$$ \begin{align*}\beta _{l}:=4\,\gamma_k\quad\text{if } l\in [2^{k-1},2^{k}).\end{align*} $$

Finally, for any $l\geq 1$ , let $P_l$ be the projection of $\ell _p({\mathbb Z}_+)$ defined by

$$ \begin{align*} P_lx=\sum_{k=b_l}^{b_{l+1}-1}x_ke_k \quad\textrm{ for every } x\in\ell_p({\mathbb Z}_+). \end{align*} $$

As in the proof of [Reference Grivaux, Matheron and Menet31, Theorem 7.2], one can show that the following estimate holds for every $k\ge 0$ , every $l\in [2^{k-1},2^k[$ , every $0\le m<l$ and every $0\le j\le \Delta ^{(k)}-\delta ^{(k)}=\tfrac 34\Delta ^{(k)}$ :

$$ \begin{align*}\|P_mT^jP_lx\|\le \frac{\beta_l}{4}\bigg(\prod_{i=\Delta^{(k)}-j+1}^{\Delta^{(k)}-1}|w_i^{(k)}|\bigg) \|P_lx\|\le \frac{\beta_l}{4}\|P_lx\|.\end{align*} $$

Hence, we have for all n and $j\le \tfrac 34\Delta ^{(n+1)}$ ,

$$ \begin{align*} \|\pi_{K_n}T^j (I-\pi_{K_n})x\| &\le \sum_{m< 2^n}\sum_{l\ge 2^n}\|P_mT^j P_lx\|\\ &\le \sum_{m< 2^n}\sum_{l\ge 2^n} \frac{\beta_l}{4}\,\|P_lx\|\\ &\le 2^n \bigg(\sum_{l\ge 2^n}\frac{\beta_l}{4}\bigg)\,\|(I-\pi_{K_n})x\|\\ &\le 2^n \bigg(\sum_{k\ge n+1}2^{k-1}\gamma_k\bigg)\,\|(I-\pi_{K_n})x\|\le \|(I-\pi_{K_n})x\|. \end{align*} $$

So, if we take care to ensure that $\tfrac 34\Delta ^{(n+1)}\ge (n+1)^{2^{J_n}}J_n=(n+1)^{2^{2\Delta ^{(n)}}}2\Delta ^{(n)}$ for all $n\ge 1$ , then condition (b) in Theorem 5.3 is satisfied. This concludes the proof of Theorem 5.2.

6. Extending frequently d-hypercyclic tuples

Let us recall the definition of d- $\mathcal F$ -hypercyclicity, for a given Furstenberg family ${\mathcal F\subset 2^{\mathbb {N}}}$ : a tuple of operators $(T_1, \ldots ,T_N)$ is d- $\mathcal F$ -hypercyclic if there exists $x\in X$ such that $x\oplus \cdots \oplus x$ is $\mathcal F$ -hypercyclic for $T_1\oplus \cdots \oplus T_N$ .

In this section, our aim is to prove the following result, which is a natural analogue of [Reference Martin and Sanders39, Theorem 2.1] for d- $\mathcal F$ -hypercyclicity. Let us denote by SOT the strong operator topology on $\mathfrak L(X)$ , that is, the topology of pointwise convergence.

Theorem 6.1. Let $\mathcal F\subset 2^{\mathbb {N}}$ be a Furstenberg family and let X be a Banach space supporting a hereditarily $\mathcal F$ -hypercyclic operator. Let $T_1,\ldots ,T_N\in \mathfrak L(X)$ and assume that $(T_1,\ldots ,T_N)$ is d- $\mathcal F$ -hypercyclic. Then, for any countable and linearly independent set $Z\subset d\,\text {-}\mathcal F\text {-}\mathrm{HC}(T_1,\ldots ,T_N)$ , the set

$$ \begin{align*}\{T\in\mathfrak L(X):\ Z\subset d\text{-}\mathcal F\text{-}\mathrm{HC}(T_1,\ldots,T_N,T)\}\end{align*} $$

is SOT-dense in $\mathfrak L(X).$

Applying this result to $Z=\{x\}$ with $x\in d\,\text {-}\mathcal F\text {-HC}(T_1,\ldots ,T_N)$ , we get the following corollary.

Corollary 6.2. Let X be a Banach space supporting a hereditarily frequently hypercyclic operator. Let $T_1,\ldots ,T_N\in \mathfrak L(X)$ and assume that $(T_1,\ldots ,T_N)$ is d-frequently hypercyclic. Then, there exists $T_{N+1}\in \mathfrak L(X)$ such that $(T_1,\ldots ,T_{N+1})$ is d-frequently hypercyclic.

If $(T_1,\ldots ,T_N)$ is densely d- $\mathcal F$ -hypercyclic then, applying Theorem 6.1 with any dense linearly independent set $Z\subset X$ contained in $d\,\text {-}\mathcal F\text {-HC}(T_1,\ldots ,T_N)$ , we obtain the following corollary.

Corollary 6.3. Let $\mathcal F\subset 2^{\mathbb {N}}$ be a Furstenberg family and let X be a Banach space supporting a hereditarily $\mathcal F$ -hypercyclic operator. Let $T_1,\ldots ,T_N\in \mathfrak L(X)$ and assume that $(T_1,\ldots ,T_N)$ is densely d- $\mathcal F$ -hypercyclic. Then, the set

$$ \begin{align*}\{T\in\mathfrak L(X):\ (T_1,\ldots,T_N,T)\;\text{ is densely } d\text{-}\mathcal F\text{-hypercyclic}\}\end{align*} $$

is SOT-dense in $\mathfrak L(X).$

In the proof of Theorem 6.1, we will need the following fact (already mentioned in §1).

Proposition 6.4. If $T\in \mathfrak L(X)$ is hereditarily $\mathcal F$ -hypercyclic, then it is in fact densely hereditarily $\mathcal F$ -hypercyclic: given a countable family $(A_i)_{i\in I}\subset \mathcal F$ and a family $(V_i)_{i\in I}$ of non-empty open sets in X, there is a dense set of $x\in X$ such that $\mathcal N_T(x,V_i)\cap A_i\in \mathcal F$ for all $i\in I$ .

Proof. Let U be a non-empty open set in X: we want to find $x\in U$ such that $\mathcal N_T(x,V_i)\cap A_i\in \mathcal F$ for all $i\in I$ . For $(i,N)\in I\times \mathbb {N}$ , define $V_{i,N}:=T^{-N}(V_i)$ and $A_{i,N}:= A_i$ . Since T is hereditarily $\mathcal F$ -hypercyclic, one can find $x_0\in X$ and sets $B_{i,N}\in \mathcal F$ such that ${B_{i,N}\subset A_{i,N}=A_i}$ and $T^nx_0\in V_{i,N}$ for all $n\in B_{i,N}$ . Since we may assume that the family $(V_i)$ is a basis of open sets for X, the vector $x_0$ is in particular a hypercyclic vector for T. So one can find an integer $N_U$ such that $x:= T^{N_U}x_0\in U$ . Then, for all $n\in B_{i,N_{U}}$ , we see that $T^nx= T^{N_U}(T^nx_0)\in V_i$ .

Proof of Theorem 6.1

Let us denote by $\mathrm {GL}(X)$ the set of all invertible operators on X. The core of the proof is contained in the following fact.

Fact 6.5. Let $T_1,\ldots ,T_N\in \mathfrak L(X)$ and assume that $(T_1,\ldots ,T_N)$ is d- $\mathcal F$ -hypercyclic. Let $T\in \mathfrak L(X)$ be a hereditarily $\mathcal F$ -hypercyclic operator. For any countable and linearly independent set $Z\subset d\,\text {-}\mathcal F\text{-}\mathrm{HC}(T_1,\ldots ,T_N)$ , for any $S\in \mathrm {GL}(X)$ and any $\varepsilon>0$ , one can find $L\in \mathrm {GL}(X)$ such that $\Vert L-S\Vert <\varepsilon $ and $Z\subset d\,\text {-}\mathcal F\text{-}\mathrm{HC}(T_1,\ldots ,T_N, L^{-1}TL)$ .

Proof of Fact 6.5

Without loss of generality, we assume that Z is infinite; we enumerate Z as a sequence $(z_l)_{l\in \mathbb {N}}$ , without repetition. Let us fix $S\in \mathrm {GL}(X)$ and $\varepsilon>0$ . Since $\mathrm {GL}(X)$ is $\Vert \,\cdot \,\Vert $ -open in $\mathfrak L(X)$ , we may assume that any operator $L\in \mathfrak L(X)$ such that $\Vert L-S\Vert <\varepsilon $ is invertible.

Let $(W_p)_{p\in \mathbb {N}}$ be a countable basis of open sets for $X^{N+1}=X^N\times X$ , and assume that each set $W_p$ has the form $W_p=U_p\times V_p$ where $U_p$ is open in $X^N$ and $V_p$ is open in X. For each $(l,p)\in \mathbb {N}\times \mathbb {N}$ , we may fix a set $A_{l,p}\in \mathcal F$ such that

$$ \begin{align*} (T_1^n z_l,\ldots,T_N^n z_l)\in U_p\quad\text{ for all } n\in A_{l,p}.\end{align*} $$

We construct, by induction, a sequence $(L_l)_{l\geq 0}$ in $\mathfrak L(X)$ with $L_0=S$ , a sequence $(x_l)_{l\geq 1}$ of vectors of X and a family $(B_{l,p})_{l,p\geq 1}$ of sets in $\mathcal F$ , such that the following holds true for every $l,p\geq 1$ :

  1. (i) $B_{l,p}\subset A_{l,p}$ and $T^n x_l\in V_p$ for all $n\in B_{l,p}$ ;

  2. (ii) $L_l(z_s)=x_s$ for all $s\leq l$ ;

  3. (iii) $\Vert L_l-L_{l-1}\Vert < 4^{-l}\varepsilon $ .

The inductive step is as follows. Choose a linear functional $v_l^*\in X^*$ such that ${v_l^*(z_s)=0}$ for all $s<l$ and $v_l^*(z_l)=1$ , which is possible by linear independence of Z. Next, since T is densely hereditarily $\mathcal F$ -hypercyclic by Proposition 6.4, we can find a vector $x_l\in X$ and sets $B_{l,p}\subset A_{l,p}$ with $B_{l,p}\in \mathcal F$ for each $p\geq 1$ , such that

$$ \begin{align*} \|x_l-L_{l-1}(z_{l})\|< \frac \varepsilon{4^l\,\|v_l^*\|}\quad\mathrm{and}\quad T^{n}x_l\in V_p\text{ for all } n\in B_{l,p}.\end{align*} $$

Then, define $L_l:=L_{l-1}+v_l^*\otimes (x_l-L_{l-1}(z_l))\in \mathfrak L(X)$ , that is,

$$ \begin{align*}L_l(x)=L_{l-1}(x)+v_l^*(x)\,(x_l-L_{l-1}(z_l)).\end{align*} $$

Clearly, $L_l(z_s)=L_{l-1}(z_s)$ for all $s<l$ , so that $L_{l}(z_s)=x_s$ by the induction hypothesis, $L_l(z_l)=x_l$ and $\|L_l-L_{l-1}\|<4^{-l}\varepsilon .$

By properties (ii) and (iii), the sequence $(L_l)$ converges to some $L\in \mathfrak L(X)$ , which satisfies $L(z_l)=x_l$ for all $l\in \mathbb {N}$ and $\|L-S\|<\varepsilon $ ; in particular, L is invertible. Moreover, $T^n x_l\in V_p$ for all $l,p\geq 1$ and $n\in B_{l,p}$ . Since $B_{l,p}\subset A_{l,p}$ , it follows that

$$ \begin{align*} (T_1^nz_l, \ldots , T_N^n z_l, (L^{-1}TL)^nz_l)\in U_p\times L^{-1}(V_p)=:\widetilde{W_p}\quad\text{for all } n\in B_{p,l}.\end{align*} $$

Now, $( \widetilde {W_p})_{p\in \mathbb {N}}$ is a basis of the topology of $X^{N+1}= X^N\times X$ because $I\oplus L^{-1}$ is a homeomorphism of $X^{N+1}$ ; so we see that $z_l\in d\,\text {-}\mathcal F\text {-HC}(T_1,\ldots , T_N, L^{-1}TL)$ for each $l\geq 1$ .

To conclude the proof of Theorem 6.1, we observe that since the operator T is hypercyclic, the similarity orbit of T, that is, the set $\{S^{-1}TS:\ S\in \mathrm {GL}(X)\}$ , is SOT-dense in $\mathfrak L(X)$ ; see for example [Reference Bayart and Matheron7, Proposition 2.20]. By Fact 6.5, it follows that the set

$$ \begin{align*}\{L^{-1}TL:\ Z\subset d\,\text{-}\mathcal F\text{-HC}(T_1,\ldots,T_N,L^{-1}TL),\ L\in\mathrm{GL}(X)\}\end{align*} $$

is SOT-dense in $\mathfrak L(X)$ .

Remark 6.6. Our proof of Theorem 6.1 differs from that of [Reference Martin and Sanders39, Theorem 2.1] regarding d-hypercyclicity, where a Baire category argument was used; and it must be so at least for d-frequent hypercyclicity, since $\mathrm {FHC}(T)$ is always meagre in X, for any operator $T\in \mathfrak L(X)$ (see [Reference Moothathu43] or [Reference Bayart and Ruzsa9]). However, there may be a Baire category proof of Theorem 6.1 when $\mathcal F$ is the family of sets with positive upper density (or, more generally, an ‘upper’ Furstenberg family in the sense of [Reference Bonilla and Grosse-Erdmann17]).

To apply Theorem 6.1, in the frequently hypercyclic case, it would be nice to exhibit a class of Banach spaces as large as possible supporting hereditarily frequently hypercyclic operators. It is easy to see that any Banach space with a symmetric Schauder basis has this property: it suffices to take $T:=2B$ , where B is the backward shift associated with the basis. In view of Corollary 3.27, a natural (and much broader) class would be that of complex Banach spaces admitting an unconditional Schauder decomposition, but we are not able to prove that every such space has the required property. In any event, we can use the method of [Reference Shkarin54] to prove the existence of d-frequently hypercyclic tuples of arbitrary length for this class of spaces.

Proposition 6.7. Let X be a complex separable infinite-dimensional Banach space with an unconditional Schauder decomposition. For any $N\geq 1$ , there exist $T_1,\ldots ,T_N\in \mathfrak L(X)$ such that $(T_1,\ldots ,T_N)$ is d-frequently hypercyclic.

Proof. By [Reference De la Rosa, Frerick, Grivaux and Peris22], X supports an operator T with a perfectly spanning set of $\mathbb {T}$ -eigenvectors. Then, $T\oplus \cdots \oplus T$ has the same property and, in particular, it is frequently hypercyclic; let $x_1\oplus \cdots \oplus x_N$ be a frequently hypercyclic vector for $T\oplus \cdots \oplus T.$ Now, let $y\in X\setminus \{ 0\}$ be arbitrary. Since $\mathrm {GL}(X)$ acts transitively on X, we may choose $S_1,\ldots ,S_N\in \mathrm {GL}(X)$ such that $S_i(x_i)=y$ for $i=1,\ldots , N$ . Then, setting $T_i:=S_i T S_i^{-1}$ , we see that $y\in d\text {-}\mathrm {FHC}(T_1,\ldots ,T_N)$ .

7. Frequent d-hypercyclicity versus dense frequent d-hypercyclicity

Despite the similarity of the definitions, there are strong differences between hypercyclicity and d-hypercyclicity. For instance, if $T\in \mathfrak L(X)$ is hypercyclic, then $\mathrm {HC}(T)$ is always dense in X, and $\mathrm {HC}(T)\cup \{ 0\}$ contains a dense linear subspace of X. In contrast, for two operators $T_1$ and $T_2,$ the set $d\text{ - }\mathrm {HC}(T_1,T_2)\cup \{0\}$ may be equal to some finite-dimensional subspace (see [Reference Sanders and Shkarin51, Theorem 3.4]). In particular, d-hypercyclic tuples are not necessarily ‘densely d-hypercyclic’.

However, since frequent d-hypercyclicity is a strong form of d-hypercyclicity, it is natural to ask whether some properties that are not true for d-hypercyclic tuples might be true for d-frequently hypercyclic tuples. In this spirit, the following question was asked in [Reference Martin, Menet and Puig36Reference Martin and Sanders38].

Question 7.1. Let $(T_1,T_2)$ be a d-frequently hypercyclic pair of operators on a Banach space X. Is $(T_1,T_2)$ necessarily densely d-hypercyclic?

The next theorem provides a solution to this problem.

Theorem 7.2. There exist a Banach space X and $T_1,T_2\in \mathfrak L(X)$ such that $(T_1,T_2)$ is d-frequently hypercyclic but not densely d-hypercyclic.

Our proof is inspired by [Reference Sanders and Shkarin51], where the authors construct a d-hypercyclic pair which is not densely d-hypercyclic. A key role will be played by the similarity orbit of some well-chosen operator T. The next two lemmas point out the relationship between (frequently) hypercyclic vectors of T and $T\oplus T$ , and (frequently) d-hypercyclic vectors of $(T_1,T_2)$ when $T_1$ and $T_2$ belong to the similarity orbit of T.

Lemma 7.3. Let $T\in \mathfrak L(X)$ and $L_1,L_2\in \mathrm {GL}(X)$ , and set $T_m:=L_m^{-1}TL_m, m=1,2$ . Let also $x\in X$ . If $x\in d\text {-}\mathrm {HC}(T_1,T_2)$ , then $L_2x-L_1x\in \mathrm {HC}(T)$ .

Proof. This is [Reference Sanders and Shkarin51, Lemma 3.1].

Lemma 7.4. Let $T\in {\mathfrak {L}}(X)$ and $L_1,L_2\in {\mathrm {GL}}(X)$ , and set $T_m:=L_m^{-1}TL_m$ . Let also ${\mathcal {F}}\subset 2^{\mathbb {N}}$ be a Furstenberg family and let $x\in X$ . Then, $x\in d\,\text {-}{\mathcal { F}}\text {-HC}(T_1,T_2)$ if and only if $(L_1x,L_2x)\in {\mathcal {F}}\text {-}{\mathrm {HC}}(T\oplus T)$ .

Proof. It is identical to the proof of [Reference Sanders and Shkarin51, Lemma 2.1]. Just observe that for any positive integer n and any pair of non-empty open subsets $(U, V)$ in X,

$$ \begin{align*} (T^{n}L_1x, T^nL_2x)\in U\times V\iff ( T_1^nx, T_2^n x)\in L_1^{-1}(U)\times L_2^{-1}(V),\end{align*} $$

and apply the definition of $\mathcal F$ -hypercyclicity.

The operators $T_1$ and $T_2$ that we are going to construct will be such that $T_2=(cI+R)^{-1}T_1(cI+R)$ for some $R\in \mathfrak L(X)$ and $c>0$ . By Lemma 7.3 (with $T=T_1$ and $L_1=cI$ ), any $x\in d\text {-}\mathrm {HC}(T_1,T_2)$ is such that $Rx\in \mathrm {HC}(T_1)$ . We shall construct $T_1$ and R in such a way that this condition prevents $d\text {-}\mathrm {HC}(T_1,T_2)$ from being dense in X. The following result will be useful to prove that the pair $(T_1,T_2)$ is d-frequently hypercyclic.

Lemma 7.5. Let $T_1,R\in \mathfrak L(X)$ and $c>0$ be such that $L_2=cI+R$ is invertible, and let $T_2:=L_2^{-1}T_1L_2$ . Let also $\mathcal F\subset 2^{\mathbb {N}}$ be a Furstenberg family. If $x\in X$ is such that $(x,Rx)\in {\mathcal F}\text {-}\mathrm {HC}(T_1\oplus T_1)$ , then $x\in d\,\text {-}\mathcal F\text {-HC}(T_1,T_2)$ .

Proof. Let us set $L_1:=cI.$ By Lemma 7.4, it suffices to show that the condition ${(x,Rx)\in {\mathcal F}\text {-}\mathrm {HC}(T_1\oplus T_1)}$ implies $(L_1x,L_2x)\in {\mathcal F}\text {-}\mathrm {HC}(T_1\oplus T_1)$ . Now it is clear that $(cx,Rx)\in {\mathcal F}\text {-}\mathrm {HC}(T_1\oplus T_1)$ . Let $U, V$ be two non-empty open subsets of X, and let $U'\subset U$ and W be non-empty open sets such that $U'+W\subset V.$ There exists a set $A\in \mathcal F$ such that ${T_1^{n}(cx)\in U'}$ and $T_1^{n}(Rx)\in W$ for all $n\in A$ . Then, for all $n\in A$ , we have

$$ \begin{align*}T_1^n L_1x=T_1^n(cx)\in U\quad\mathrm{and}\quad T_1^{n}L_2x=T_1^{n}(cx+Rx)\in U'+ W\subset V.\\[-36pt]\end{align*} $$

We now go into the details of the construction. First, we define the Banach space X as

$$ \begin{align*} X:=\bigg( \bigoplus_{l\geq 1} X(l)\bigg)_{c_0}\quad \text{where } X(l)=\ell_1(\mathbb{Z}_+) \text{ for every } l\ge 1. \end{align*} $$

(Following a standard notation, the subscript ‘ $c_0$ ’ indicates that the direct sum is a $c_0$ -sum.)

Next, we introduce the following operator $T\in \mathfrak L(X)$ : denoting by B the canonical backward shift on $\ell _1(\mathbb {Z}_+)$ , let

$$ \begin{align*} T:=\bigoplus_{l\geq 1} T(l),\quad\text{where } T(l)=I+2^{-l}B\in\mathfrak L(X(l))\textrm{ for every }l\ge 1.\end{align*} $$

Lemma 7.6. The operator $T\oplus T$ is frequently hypercyclic.

Proof. It is enough to prove that T has a perfectly spanning set of $\mathbb {T}$ -eigenvectors. Indeed, $T\oplus T$ will have the same property and hence will be frequently hypercyclic.

We now define our first operator $T_1$ :

$$ \begin{align*} T_1:=\bigoplus_{l\geq 1} (I+B)\in\mathfrak L(X).\end{align*} $$

Note that the same proof as that of Lemma 7.6 shows that $T_1\oplus T_1$ is frequently hypercyclic. However, we will use the above operator T to produce a frequently hypercyclic vector for $T_1\oplus T_1$ with specific properties.

In what follows, we denote by $(e_k(l))_{k\geq 0}$ the canonical basis of the lth component $X(l)=\ell _1(\mathbb {Z}_+)$ of X and by $(e_k^*(l))_{k\geq 0}$ the associated sequence of coordinate functionals, which we will consider as linear functionals on X. A vector $x\in X$ will be written as $x=(x(l))_{l\geq 1}$ and we will use the notation $x_k(l)=\langle e_k^*(l), x\rangle $ for every $k\ge 0$ .

Lemma 7.7. There exists $(u,v)\in \mathrm {FHC}(T_1\oplus T_1)$ such that $u_0(1)\neq 0$ and $|v_k(l)|\leq 2^{-lk}$ for all $k\geq 0$ and all $l\geq 1$ .

Proof. For $l\geq 1,$ Let us consider the diagonal operator $D(l)$ on $X(l)$ defined by

$$ \begin{align*}D(l)(x(l)):=\sum_{k\geq 0}2^{-lk}x_k (l)e_k(l)\end{align*} $$

and set

$$ \begin{align*} D:=\bigoplus_{l\geq 1} D(l)\in\mathfrak L(X).\end{align*} $$

It is easy to check that $(I+B)D(l)=D(l)(I+2^{-l}B)$ for each $l\geq 1$ , so that ${T_1D=DT}$ . Moreover, the operator D has dense range. So $T_1$ is a quasi-factor of T with quasi-factoring map D and, hence, $T_1\oplus T_1$ is a quasi-factor of $T\oplus T$ with quasi-factoring map $D\oplus D$ . Let $(x,y)\in \mathrm {FHC}(T\oplus T)$ with $\|y\|\leq 1$ and $x_0(1)\neq 0$ , and let us set $(u,v):=(Dx,Dy)$ . Then, $(u,v)\in \mathrm {FHC}(T_1\oplus T_1)$ . Moreover, $u_0(1)=x_0(1)\neq 0$ and, for $k\geq 0$ and $l\geq 1,$

$$ \begin{align*}|v_k(l)|=|2^{-lk}y_k(l)|\leq 2^{-lk}.\\[-38pt]\end{align*} $$

We now give a result which provides a condition preventing a vector from being hypercyclic for $T_1$ .

Lemma 7.8. Let $x\in X$ . Assume that there exist $l\geq 1$ and $\unicode{x3bb} \neq 0$ such that, for all $k\geq 1$ sufficiently large, $\Re e\,\langle e_k^*(l), x/\unicode{x3bb} \rangle \geq 0.$ Then, $x\notin \mathrm {HC}(T_1)$ .

Proof. Since $x\in \mathrm {HC}(T_1)$ if and only if $x/\unicode{x3bb} \in \mathrm {HC}(T_1)$ , we may assume $\unicode{x3bb} =1$ . Now, if $\Re e (x_k(l))=\Re e\,\langle e_k^*(l), x\rangle \geq 0$ for all sufficiently large $k,$ then the arguments of [Reference Sanders and Shkarin51] (see the proof of Claim 1, p. 845) show that either $x(l)$ is finitely supported or

$$ \begin{align*}\Re e\, \Big\langle e_0^*(l), (I+B)^n(x(l))\Big\rangle\geq 0\quad\text{for all sufficiently large } n.\end{align*} $$

Therefore, $x(l)$ cannot be a hypercyclic vector for $I+B$ and, hence, $x\notin \mathrm {HC}(T_1)$ .

Let us fix a sequence of positive real numbers $(\varepsilon _l)_{l\geq 1}$ going to zero. Let also ${V:\ell _1(\mathbb {Z}_+)\to \ell _1(\mathbb {Z}_+)}$ be the (bounded) operator defined by

$$ \begin{align*} Vy:= \sum_{k\geq 1}\bigg(\sum_{j\geq 1}2^{-jk}y_j\bigg)e_k \quad\textrm{ for every }y\in \ell_1(\mathbb{Z}_+)\end{align*} $$

and let $R_0:X\to X$ be the (bounded) operator on X defined by

$$ \begin{align*}R_0(x):=(\varepsilon_1 V(x(1)),\ldots,\varepsilon_l V(x(1)),\ldots).\end{align*} $$

The operator $R_0$ satisfies the following crucial estimates.

Lemma 7.9. Let $x\in X$ and $\ m\geq 1$ be such that $\langle e_m^*(1),x\rangle \neq 0$ and $\langle e_j^*(1),x\rangle =0$ for $1\leq j<m.$ Then, there exists $\delta :\mathbb {Z}_+\longrightarrow \mathbb {R}_+$ such that $\delta (k)\longrightarrow 0$ as $k\rightarrow \infty $ and such that for all $l\geq 1$ and $k\ge 0$ , we can write

$$ \begin{align*}\langle e_k^*(l), R_0(x)\rangle=\varepsilon_l\, \langle e_m^*(1),x\rangle\,2^{-mk}(1+\delta(k)).\end{align*} $$

Proof. We just write

$$ \begin{align*} \langle e_k^*(l), R_0(x)\rangle&=\varepsilon_l \sum_{j=1}^{\infty}2^{-jk}\langle e_j^*(1),x\rangle\\ &=\varepsilon_l \sum_{j=m}^{\infty}2^{-jk}\langle e_j^*(1),x\rangle\\ &=\varepsilon_l 2^{-mk}\langle e_m^*(1),x\rangle\bigg(1+\sum_{j=1}^{\infty}2^{-jk}\frac{\langle e_{j+m}^*(1),x\rangle}{\langle e_m^*(1),x\rangle}\bigg); \end{align*} $$

and we conclude because

$$ \begin{align*} \bigg|\!\sum_{j=1}^{\infty}2^{-jk}\frac{\langle e_{j+m}^*(1),x\rangle}{\langle e_m^*(1),x\rangle}\bigg|\leq \frac{2^{-k}}{\langle e_m^*(1),x\rangle}\|x(1)\|_1\xrightarrow{k\to\infty}0.\\[-47pt] \end{align*} $$

We consider $(u,v)$ given by Lemma 7.7. We set, for $x\in X,$

$$ \begin{align*}R(x):=R_0(x)+\frac{x_0(1)}{u_0(1)}(v-R_0(u)).\end{align*} $$

This defines a bounded operator such that $R(u)=v\in \mathrm {HC}(T_1).$ It turns out that there are not so many vectors $x\in X$ such that $R(x)\in \mathrm {HC}(T_1)$ .

Lemma 7.10. Let $x\in X$ be such that $R(x)\in \mathrm {HC}(T_1)$ . Then, $x(1)$ is a scalar multiple of  $u(1)$ .

Proof. Let us set $z:=x-({x_0(1)}/{u_0(1)})u$ , so that

$$ \begin{align*}R(x)=R_0(z)+\frac{x_0(1)}{u_0(1)}v.\end{align*} $$

Assume first that $z(1)\notin \mathrm {span}(e_0(1))$ . Then, there exists $m\geq 1$ such that $\langle e_m^*(1), z\rangle \neq 0,$ whereas $\langle e_j^*(1), z\rangle =0$ for $1\leq j<m$ . Let $l>m$ . By Lemma 7.9, it follows that

$$ \begin{align*}\langle e_k^*(l), R_0(z)\rangle=\varepsilon_l \langle e_m^*(1), z\rangle 2^{-mk}(1+\delta(k)),\end{align*} $$

so that

$$ \begin{align*} \langle e_k^*(l), R(x)\rangle&=\varepsilon_l \langle e_m^*(1), z\rangle 2^{-mk}(1+\delta(k))+\frac{x_0(1)}{u_0(1)}v_k(l)\\ &=\varepsilon_l \langle e_m^*(1)(z)2^{-mk}(1+\delta_l(k)), \end{align*} $$

where for all $l>m$ , $\delta _l(k)\rightarrow 0$ as $k\rightarrow \infty $ , since $|v_k(l)|\leq 2^{-lk}=o(2^{-mk})$ . By Lemma 7.8, $R(x)\notin \mathrm {HC}(T_1)$ , which is a contradiction.

Hence, there exists a complex number $\alpha (1)$ such that

$$ \begin{align*}x(1)-\frac{x_0(1)}{u_0(1)}u(1)=\alpha(1)e_0(1).\end{align*} $$

Applying the functional $e_0^*(1)$ to this equation, we easily get $\alpha (1)=0$ , which implies that $x(1)$ belongs to $\mathrm {span}(u(1))$ .

We can now give the proof.

Proof of Theorem 7.2

Let us fix $c>\|R\|$ . We set $L_2:=cI+R$ (which is invertible) and $T_2:=L_2^{-1}TL_2$ . We show that the pair $(T_1,T_2)$ is d-frequently hypercyclic but not densely d-hypercyclic.

By construction, $(u,Ru)=(u,v)\in \mathrm {FHC}(T_1\oplus T_1)$ . Hence, $u\in d\text {-}\mathrm {FHC}(T_1,T_2)$ by Lemma 7.5. Moreover, setting $T=T_1$ and $L_1=cI,$ Lemma 7.3 implies that if ${x\in d\text {-}\mathrm {HC}(T_1,T_2),}$ then $R(x)\in \mathrm {HC}(T_1)$ . By Lemma 7.10, it follows that $x(1)\in \mathrm {span}(u(1))$ for every $x\in d\text {-}\mathrm {HC}(T_1,T_2)$ . In particular, $d\text {-}\mathrm {HC}(T_1,T_2)$ cannot be dense in X.

8. Eigenvectors and d-frequent hypercyclicity

In this section, we give a criterion relying on properties of the eigenvectors for showing that a tuple of operators is (densely) d-frequently hypercyclic. The initial motivation was the following question asked by K. Grosse-Erdmann: let D be the derivation operator acting on the space of entire functions $H(\mathbb {C})$ , and for every $a\in \mathbb {C}\setminus \{ 0\}$ , denote by $\tau _a$ the operator of translation by a on $H(\mathbb {C})$ , defined by $\tau _a f(z):= f(z+a)$ . It is well known (see [Reference Bayart and Matheron7] or [Reference Grosse-Erdmann and Peris33]) that both D and $\tau _a$ are frequently hypercyclic. Now one can ask the following question.

Question 8.1. Do the operators D and $\tau _a$ have common frequently hypercyclic vectors?

It will follow from the next theorem that the answer to Question 8.1 is positive.

Theorem 8.2. Let $N\geq 2,$ let X be a complex Fréchet space and let $T_1, \ldots ,T_N\in \mathfrak L(X)$ . Assume that there exist a holomorphic vector field $E:O \to X$ , defined on some connected open set $O\subset \mathbb {C}$ , and non-constant holomorphic functions $\phi _1, \ldots ,\phi _N$ , defined on some connected open set containing O, such that:

  • $\overline {\mathrm {span}}\, E(O)=X$ ;

  • $T_i E(z)=\phi _i(z) E_i(z)$ for every $i=1,\ldots ,N$ and $z\in O$ ;

  • $O\cap \phi _i^{-1}(\mathbb {T})\cap \bigcap _{j\neq i} \phi _j^{-1}(\mathbb {D})\neq \emptyset $ for every $i=1,\ldots ,N$ .

Then, the N-tuple $(T_1,\ldots , T_N)$ is densely d-frequently hypercyclic.

Before proving this result, let us state two consequences and give some examples.

Corollary 8.3. Let D be the derivation operator on $X:=H(\mathbb {C})$ . If $\phi _1$ and $\phi _2$ are two entire functions of exponential type such that $\phi _1^{-1}(\mathbb {T})\cap \phi _2^{-1}(\mathbb {D})\neq \emptyset $ and $\phi _2^{-1}(\mathbb {T})\cap \phi _1^{-1}(\mathbb {D})\neq \emptyset $ , then the pair $(\phi _1(D), \phi _2(D))$ is densely d-frequently hypercyclic.

Proof. Let $E:\mathbb {C}\to X$ be the holomorphic vector field defined by $E(z):= e^{z\,\cdot \, }$ . We have $DE(z)=zE(z)$ for all $z\in \mathbb {C}$ and $\overline {\mathrm {span}}\, E(\mathbb {C})=X$ ; so we may apply Theorem 8.2 to the operators $T_i:=\phi _i(D)$ .

Since $\tau _a=\Phi _a(D)$ , where $\phi _a(z):=e^{az}$ , Corollary 8.3 applies to pairs of operators involving D and $\tau _a$ :

Example 8.4. Taking $\phi _1(z):= z$ and $\phi _2(z):=e^{az}$ , we see that for any $a\neq 0$ , the pair $(D,{\kern-1pt}\tau _a)$ is densely d-frequently hypercyclic (so that, in particular, ${\mathrm {FHC}(D){\kern-1pt}\cap{\kern-1pt} \mathrm {FHC}(\tau _a) {\kern-1pt}\neq{\kern-1pt} \emptyset} $ ). Indeed, any complex number z such that $\vert z\vert =1$ and $\Re e(az)<0$ belongs to ${\phi _1^{-1}(\mathbb {T})\cap \phi _2^{-1}(\mathbb {D})}$ , while any $z\in \ i\overline {a}\,\mathbb {R}$ such that $|z|<1$ belongs to $\phi _2^{-1}(\mathbb {T})\cap \phi _1^{-1}(\mathbb {D})$ . Similarly, if $a,b\neq 0$ and $a/b\not \in \mathbb {R}$ , then $(\tau _a,\tau _b)$ is densely d-frequently hypercyclic.

Corollary 8.5. Let B be the canonical backward shift acting on $X=\ell _p(\mathbb {Z}_+)$ or $c_0(\mathbb {Z}_+)$ . If $\phi _1$ and $\phi _2$ are two holomorphic functions defined in a neighbourhood of the closed unit disk $\overline {\mathbb {D}}$ such that $\mathbb {D}\cap \phi _1^{-1}(\mathbb {T})\cap \phi _2^{-1}(\mathbb {D})\neq \emptyset $ and $\mathbb {D}\cap \phi _2^{-1}(\mathbb {T})\cap \phi _1^{-1}(\mathbb {D})\neq \emptyset $ , then the pair $(\phi _1(B), \phi _2(B))$ is densely d-frequently hypercyclic.

Proof. The operators $T_i:=\phi _i(B)$ are well defined since $\sigma (B)=\overline {\,\mathbb {D}}$ . Let $E:\mathbb {D}\to X$ be the holomorphic vector field defined by $E(z):=\sum _{n=0}^\infty z^n e_n$ . We have $BE(z)=zE(z)$ for all $z\in \mathbb {D}$ and $\overline {\mathrm {span}}\, E(\mathbb {C})=X$ ; so Theorem 8.2 applies.

Example 8.6. If $\vert \unicode{x3bb} \vert>1$ and $0<\vert \alpha \vert <2\vert \unicode{x3bb} \vert $ , the pair $(\unicode{x3bb} B, I+\alpha B)$ is densely d-frequently hypercyclic. If $\vert \unicode{x3bb} \vert>1$ and $a\neq 0$ , the pair $(\unicode{x3bb} B, e^{aB})$ is densely d-frequently hypercyclic. However, Theorem 8.2 is completely inefficient to show, for example, that the pair $(aB, bB^2)$ is d-frequently hypercyclic if $1<a<b$ , which is nevertheless true by [Reference Martin, Menet and Puig36].

In the proof of Theorem 8.2, we will make use of the following straightforward observation.

Fact 8.7. Let $T_1,\ldots ,T_N\in \mathfrak L(X)$ and let $x_1,\ldots ,x_N\in X$ . Assume that $(x_1,\ldots ,x_N)\in \mathrm {FHC}(T_1\oplus \cdots \oplus T_N)$ and that $T_j^m x_i\to 0$ as $m\to \infty $ whenever $i\neq j$ . Then, $x:= x_1+\cdots +x_N$ is a d-frequently hypercyclic vector for $(T_1,\ldots ,T_N)$ .

Proof of Theorem 8.2

We first note that for $i=1,\ldots ,N$ , there is a non-empty open set $V_i\subset O$ and $r_i\in (0,1)$ such that $(\phi _i)_{| V_i}$ is a diffeomorphism (onto its range), $\phi _i(V_i)\cap \mathbb {T}\neq \emptyset $ and $\phi _j(V_i)\subset D(0,r_i)$ for $j\neq i$ . Indeed, let $a\in O$ be such that $\phi _i(a)\in \mathbb {T}$ and $\phi _{j}(a)\in \mathbb {D}$ for $j\neq i$ . Choose an open neighbourhood W of a and $r_i\in (0,1)$ such that ${\phi _{j}(W)\subset D(0,r_i)}$ for all $j\neq i$ . By the open mapping theorem, $\phi _i(W)$ is an open set intersecting $\mathbb {T}$ , so $\phi _i(W)\cap \mathbb {T}$ is uncountable. Hence, one can find $b\in W$ such that ${\phi _i(b)\in \mathbb {T}}$ and $\phi ^{\prime }_i(b)\neq 0$ ; and the claim follows from the inverse function theorem.

Taking the open set $V_i$ smaller if necessary, we may assume that $\Lambda _i:=\phi _i(V_i)\cap \mathbb {T}$ is a proper open arc of $\mathbb {T}$ . We choose a ‘cut-off’ function $\chi _i\in \mathcal C^\infty (\mathbb {T})$ such that ${\chi _i(\unicode{x3bb} )=0}$ outside $\Lambda _i$ and $\chi _i(\unicode{x3bb} )>0$ on $\Lambda _i$ , and (with the obvious abuse of notation) we define ${F_i:\mathbb {T}\to X}$ by setting

$$ \begin{align*} F_i(\unicode{x3bb}):=\chi_i(\unicode{x3bb}) E( \phi_{i}^{-1}(\unicode{x3bb}))\quad\textrm{ for every }\unicode{x3bb}\in\mathbb{T}.\end{align*} $$

Thus, $F_i$ is a $\mathcal C^\infty $ -smooth $\mathbb {T}$ -eigenvector field for $T_i$ , that is, $T_iF_i(\unicode{x3bb} )=\unicode{x3bb} F_i(\unicode{x3bb} )$ for every $\unicode{x3bb} \in \mathbb {T}$ . Let us denote by $\widehat {F_i}(n)$ the Fourier coefficients of $F_i$ :

$$ \begin{align*} \widehat{F_i}(n)=\int_{\mathbb{T}} F_i(\unicode{x3bb})\, \unicode{x3bb}^{-n}\, d\unicode{x3bb}, \quad n\in\mathbb{Z}.\end{align*} $$

Since $F_i$ is a $\mathbb {T}$ -eigenvector field for T, we have $T_i\widehat {F_i}(n)=\widehat {F_i}(n-1)$ for all $n\in \mathbb {Z}$ , that is, the sequence $(\widehat {F_i}(n))_{n\in \mathbb {Z}}$ is a bilateral backward orbit for $T_i$ . Moreover, since $\overline {\mathrm {span}}\,E(O)=X$ , it follows from the Hahn–Banach theorem, together with the identity principle for analytic functions, that $\overline {\mathrm {span}}\, ( \widehat {F_i}(n)\;:\; n\in \mathbb {Z})=X$ .

In the remainder of the proof, we fix a family $(g_{i,n})_{1\leq i\leq N, n\in \mathbb {Z}}$ of independent, standard complex Gaussian variables defined on some probability space $(\Omega ,\mathcal A,\mathbb P)$ .

Claim 8.8. For every $i=1,\ldots ,N$ , the series

$$ \begin{align*}\sum_{n\in\mathbb{Z}} g_{i, n} \widehat{F_i}(n)\end{align*} $$

is almost surely convergent and defines an X-valued random variable $\xi _i$ on $(\Omega ,\mathcal A,\mathbb P)$ , which is such that

$$ \begin{align*} \textit{for every } i,j=1,\ldots,N \textit{ with } j\neq i,\; T_j^m\xi_i\xrightarrow{m\to\infty} 0 \text{ almost surely}. \end{align*} $$

Proof of Claim 8.8

Since $F_i$ is $\mathcal C^2$ -smooth, two integrations by parts show that for any continuous semi-norm $\mathbf q$ on X, we have $\mathbf q( \widehat {F_i}(n))=O(1/n^2)$ as $\vert n\vert \to \infty $ , so that

$$ \begin{align*} \sum_{n=-\infty}^\infty \mathbb E(\mathbf q(g_{i, n} \widehat{F_i}(n)))<\infty.\end{align*} $$

This implies that the series $\sum _{n\in \mathbb {Z}} g_{i,n} \widehat {F_i}(n)$ is almost surely convergent.

Let us fix $j\neq i$ . By the definition of $F_i$ , we have $T_jF_i(\unicode{x3bb} )={\psi _{i,j}(\unicode{x3bb} )} F_i(\unicode{x3bb} )$ for every ${\unicode{x3bb} \in \Lambda _i}$ , where ${\psi _{i,j}(\unicode{x3bb} )}:=\phi _j(\phi _i^{-1}(\unicode{x3bb} ))$ ; and $T_jF_i(\unicode{x3bb} ) =0$ if $\unicode{x3bb} \not \in \Lambda _i$ . Hence, for almost every $\omega \in \Omega $ and every $m\in \mathbb {N}$ , we have

$$ \begin{align*} T_j^m (\xi_i(\omega))=\sum_{n\in\mathbb{Z}} g_n(\omega) \int_{\Lambda_i} {\psi_{i,j}(\unicode{x3bb})}^m F_i(\unicode{x3bb})\,\unicode{x3bb}^{-n}\, d\unicode{x3bb}.\end{align*} $$

Let $\mathbf q$ be a continuous semi-norm on X. Since $\vert {\psi _{i,j}(\unicode{x3bb} )}\vert <r_i$ for every $\unicode{x3bb} \in \Lambda _i$ by definition of $\psi $ , two integrations by parts show that there is a constant $C_{\mathbf q}$ such that

$$ \begin{align*} \mathbf{q}\bigg(\!\int_\Lambda {\psi_{i,j}(\unicode{x3bb})}^m F(\unicode{x3bb})\,\unicode{x3bb}^{-n}\, d\unicode{x3bb}\bigg)\leq C_{\mathbf q}\times \frac{m^2 r_i^m}{1+n^2}\quad\textrm{ for every }m\ge 0 \textrm{ and every }n\in\mathbb{Z}.\end{align*} $$

Moreover, it follows from the Borel–Cantelli lemma that for almost every $\omega \in \Omega $ , there exists an integer $N(\omega )$ such that

$$ \begin{align*} \text{ for all } \vert n\vert>N(\omega)\;:\; \vert g_n(\omega)\vert \leq \sqrt{n}.\end{align*} $$

Hence, given a continuous semi-norm $\mathbf q$ on X, one can find for almost every $\omega \in \Omega $ , some constant $M_{\mathbf q,\omega }$ such that

$$ \begin{align*} \text{ for all } m\in\mathbb{N}\;:\; \mathbf q( T_j^m\xi_i(\omega))\leq M_{q,\omega} \, m^2 r_i^m.\end{align*} $$

Hence, $\mathbf q( T^m_j\xi _i(\omega ))\to 0$ almost surely as $m\to \infty $ for any given continuous semi-norm $\mathbf q$ , that is, $T_j^m\xi _i\to 0$ almost surely.

We can now conclude the proof of Theorem 8.2. For $i=1,\ldots ,N$ , let us denote by $\mu _i$ the distribution of the random variable $\xi _i:\Omega \to X$ . By definition, $\mu _i$ is a $T_i$ -invariant Gaussian measure with full support; and by [Reference Agneessens1], $T_i$ is strongly mixing with respect to $\mu _i$ . Hence, the measure $\mu _1\otimes \cdots \otimes \mu _N$ is a $(T_1\oplus \cdots \oplus T_N)$ -invariant measure on $X^N$ with full support and $T_1\oplus \cdots \oplus T_N$ is mixing with respect to $\mu _1\otimes \cdots \otimes \mu _N$ . Since ${\mu _1\otimes \cdots \otimes \mu _N}$ is the distribution of the random vector $\xi :=(\xi _1,\ldots ,\xi _N)$ by independence of $\xi _1,\ldots ,\xi _N$ , it follows that the vector $\xi (\omega )$ is almost surely frequently hypercyclic for $T_1\oplus \cdots \oplus T_N$ . Moreover, by Claim 8.8, we see that $T_j^m \xi _i(\omega )\to 0$ almost surely as ${m\to \infty }$ , whenever $i\neq j$ . By Fact 8.7, it follows that the vector ${\xi _1(\omega )+\cdots +\xi _N(\omega )}$ is almost surely d-frequently hypercyclic for $(T_1,\ldots ,T_N)$ . In other words, $(\mu _1\star \cdots \star \mu _N)$ -almost every $x\in X$ is d-frequently hypercyclic for $(T_1,\ldots ,T_N)$ . Since the measure $\mu _1\star \cdots \star \mu _N$ has full support, this terminates the proof of Theorem 8.2.

9. Remarks and questions

9.1. Hereditary frequent hypercyclicity in a weak sense

Another natural definition for hereditary $\mathcal F$ -hypercyclicity could be the following: an operator $T\in \mathfrak L(X)$ is hereditarily $\mathcal F$ -hypercyclic in the weak sense if, for every $A\in \mathcal F,$ the sequence $(T^n)_{n\in A}$ is $\mathcal F$ -hypercyclic, that is, there exists $x\in X$ such that $\mathcal N_T(x,V)\cap A\in \mathcal F$ for all non-empty open sets $V\subset X.$ Equivalently, T is $\mathcal F_A$ -hypercyclic for every $A\in \mathcal F$ , where $\mathcal F_A:=\{ B\subset \mathbb {N}\,:\, B\cap A\in \mathcal F\}$ . Of course, hereditary $\mathcal F$ -hypercyclicity implies hereditary $\mathcal F$ -hypercyclicity in the weak sense. Note also that Theorem 5.2 says precisely that there exist frequently hypercyclic operators which are not hereditarily frequently hypercyclic in the weak sense.

When $\mathcal F$ is the family of all infinite subsets of $\mathbb {N}$ , an operator T is hereditarily $\mathcal F$ -hypercyclic in the weak sense if and only if it is ‘hereditarily hypercyclic with respect to the whole sequence of integers’ in the sense of [Reference Bès and Peris13]; and this means exactly that T is topologically mixing (see for example [Reference Grivaux27, Lemma 2.2]). The next result shows that this is also equivalent to hereditary $\mathcal F$ -hypercyclicity.

Proposition 9.1. Let $\mathcal F$ be a Furstenberg family with the following property: for any operator T and any $A\subset \mathbb {N}$ , the set $\mathcal F_A$ -HC $(T)$ is either empty or comeagre in the underlying space. Then, hereditary $\mathcal F$ -hypercyclicity and hereditary $\mathcal F$ -hypercyclicity in the weak sense are equivalent. In particular, when $\mathcal F$ is the family of all infinite subsets of $\mathbb {N}$ , an operator T is hereditarily $\mathcal F$ -hypercyclic if and only if it is topologically mixing.

Proof. Assume that T is hereditarily $\mathcal F$ -hypercyclic in the weak sense. Let $(A_i)_{i\in I}$ be a countable family of sets in $\mathcal F$ and let $(V_i)_{i\in I}$ be a family of non-empty open subsets of X. By assumption on $\mathcal F$ , for each $i\in I$ , the set $G_i$ of $\mathcal F$ -hypercyclic vectors for the sequence $(T^n)_{n\in A_i}$ is comeagre in X; so $G:=\bigcap _{i\in I} G_i$ is non-empty. Then, any $x\in G$ satisfies the required property: for every $i\in I$ , there is a set $B_i\in \mathcal F$ such that $B_i\subset A_i$ and $T^nx\in V_i$ for all $n\in B_i$ .

We can now ask the following question.

Question 9.2. For which Furstenberg families $\mathcal F$ do hereditary $\mathcal F$ -hypercyclicity and hereditary $\mathcal F$ -hypercyclicity in the weak sense coincide?

In view of the results of [Reference Bonilla and Grosse-Erdmann17], upper Furstenberg families may be good candidates. However, we are unable to handle even the case of sets with positive upper density. Proposition 9.1 leads naturally to the following question.

Question 9.3. Let us denote by $\overline {\mathcal D}$ the family of all sets $A\subset \mathbb {N}$ with positive upper density. Is it true that if $A\in \overline {\mathcal D}$ and $T\in \mathfrak L(X)$ is $\overline {\mathcal D}_A$ -hypercyclic, then $\overline {\mathcal D}_A$ -HC $(T)$ is comeagre in X?

However, it might seem more than plausible that the two notions are not equivalent in the case of frequent hypercyclicity, that is, when $\mathcal F$ is the family of sets with positive lower density. However again, we do not know how to prove this.

One may also think of ‘local’ versions of hereditary frequent hypercyclicity. For example, one could say that an operator $T\in \mathfrak L(X)$ is:

  • hereditarily $\mathcal F$ -hypercyclic with respect to some sequence $(\Lambda _i)_{i\in \mathbb {N}}\subset \mathcal F$ if, for any sequence $(A_i)\subset \mathcal F$ with $A_i\subset \Lambda _i$ and for any sequence of non-empty open sets $(V_i)$ in X, one can find a vector $x\in X$ such that $\mathcal N_T(x,V_i)\cap A_i\in \mathcal F$ for all $i\in \mathbb {N}$ ;

  • hereditarily $\mathcal F$ -hypercyclic with respect to some set $\Lambda \in \mathcal F$ if it is hereditarily $\mathcal F$ -HC with respect to the constant sequence $\Lambda _i=\Lambda $ ;

  • hereditarily $\mathcal F$ -hypercyclic in the weak sense with respect to some set $\Lambda \in \mathcal F$ if it is $\mathcal F_A$ -hypercyclic for any $A\in \mathcal F\cap 2^\Lambda $ .

When $\mathcal F$ is the family of all infinite subsets of $\mathbb {N}$ , hereditary $\mathcal F$ -hypercyclicity in the weak sense with respect to some set $\Lambda =\{ n_k\,:\, k\geq 0\}$ is the same as hereditary hypercyclicity with respect to the sequence $(n_k)$ in the sense of [Reference Bès and Peris13]; and hence, by [Reference Bès and Peris13, Theorem 2.3], an operator T is hereditarily $\mathcal F$ -hypercyclic in the weak sense with respect to some set $\Lambda $ if and only if it is topologically weakly mixing, that is, $T\oplus T$ is hypercyclic. Also, the proof of Proposition 4.1 makes it clear that if T is hereditarily $\mathcal F$ -hypercyclic with respect to some sequence $(\Lambda _i)$ , then $T\oplus T$ is $\mathcal F$ -hypercyclic. This leads to the following question.

Question 9.4. If $T\in \mathfrak L(X)$ is hereditarily $\mathcal F$ -hypercyclic in the weak sense with respect to some set $\Lambda \in \mathcal F$ , does it follow that $T\oplus T$ is $\mathcal F$ -hypercyclic? And conversely?

In the same spirit and with [Reference Ernst, Esser and Menet24] in mind, one may ask another question.

Question 9.5. Does $\mathcal U$ -frequent hypercyclicity imply some weak form of hereditary $\mathcal U$ -frequent hypercyclicity, yet strong enough to ‘explain’ why $T\oplus T$ is $\mathcal U$ -frequently hypercyclic as soon as T is?

9.2. $\mathcal F$ -transitivity and hereditary $\mathcal F$ -transitivity

In topological dynamics, there is a natural notion of ‘transitivity’ associated with a given Furstenberg family $\mathcal F$ (see for example [Reference Bès, Menet, Peris and Puig12, Reference Glasner26] in the linear setting): if X is a topological space, a continuous map ${T:X\to X}$ is said to be $\mathcal F$ -transitive if $\mathcal N_T(U,V)\in \mathcal F$ for every pair $(U,V)$ of non-empty open sets in X, where

$$ \begin{align*}\mathcal N_T(U,V):=\{ n\in\mathbb{N}\,:\, T^n(U)\cap V\neq \emptyset\}.\end{align*} $$

Following [Reference Bès, Menet, Peris and Puig12], one can consider a ‘hereditary’ version of $\mathcal F$ -transitivity: let us say that an operator $T\in \mathfrak L(X)$ is hereditarily $\mathcal F$ -transitive if $\mathcal N(U, V)\cap A\in \mathcal {F}$ for every $A\in \mathcal {F}$ and all non-empty open sets $U,V$ . There is an obvious link with hereditary $\mathcal F$ -hypercyclicity.

Remark 9.6. Hereditarily $\mathcal F$ -hypercyclic operators are hereditarily $\mathcal F$ -transitive.

Proof. By Proposition 6.4, we know that if T is hereditarily $\mathcal {F}$ -hypercyclic, then T is densely hereditarily $\mathcal {F}$ -hypercyclic. Let $U, V$ be non-empty open sets in X and let $A\in \mathcal {F}$ . By dense hereditary $\mathcal F$ -hypercyclicity, there exists $x\in U$ such that $N(x,V)\cap A\in \mathcal {F}$ . In particular, $N(U,V)\cap A\in \mathcal {F}$ , so T is hereditarily $\mathcal {F}$ -transitive.

The converse is definitely not true in general, for the following reason: there exist topologically mixing operators that are not frequently hypercyclic. In particular, any such operator is hereditarily $\underline {\mathcal {D}}$ -transitive, where $\underline {\mathcal {D}}$ is the family of sets with positive lower density, but not frequently hypercyclic (that is, not $\underline {\mathcal {D}}$ -hypercyclic). This leads to the following questions (the third one was suggested by the referee).

Question 9.7. Are there operators which are frequently hypercyclic and topologically mixing, but not hereditarily frequently hypercyclic?

Question 9.8. Are there at least operators which are hereditarily $\underline {\mathcal {D}}$ -transitive and frequently hypercyclic, but not hereditarily frequently hypercyclic?

Question 9.9. Assume that $T\in \mathfrak L(X)$ admits an invariant measure with full support with respect to which it is a strongly mixing transformation. Does it follow that T is hereditarily frequently hypercyclic? What if, additionally, the measure is Gaussian?

Given a Furstenberg family $\mathcal {F}$ , one can define the dual family $\mathcal {F}^*$ as the collection of all subsets A of $\mathbb {N}$ such that $A\cap B\neq \emptyset $ for every $B\in \mathcal {F}$ . It is clear by definition that every hereditarily $\mathcal {F}$ -transitive operator is $\mathcal {F}^*$ -transitive; and it is also clear that ${(\underline {\mathcal {D}})^*=\overline {\mathcal {D}}_1}$ , the family of sets with upper density equal to $1$ . Hence, every hereditarily frequently hypercyclic operator is $\overline {\mathcal {D}}_1$ -transitive. It is natural to wonder if every frequently hypercyclic operator is $\overline {\mathcal {D}}_1$ -transitive too. The next proposition shows that this is not the case. This is an improvement of [Reference Bès, Menet, Peris and Puig12, Proposition 5.1], where it is shown that reiterative hypercyclicity does not imply $\overline {\mathcal {D}}_1$ -transitivity. Moreover, the example we give is any of the weighted shifts introduced in the proof of Theorem 4.2; so this provides another proof that these shifts are not hereditarily frequently hypercyclic.

Proposition 9.10. There exists a frequently hypercyclic weighted shift $B_w$ on $c_0(\mathbb {Z}_+)$ which is not $\overline {\mathcal {D}}_1$ -transitive.

Proof. Let $B_w$ be one of the weighted shifts introduced in the proof of Theorem 4.2. By [Reference Bès, Menet, Peris and Puig12, Proposition 3.3], to show that $B_w$ is not $\overline {\mathcal {D}}_1$ -transitive, it is enough to find $M>0$ such that

$$ \begin{align*}{C}_M:= \{n\in\mathbb{N} : |w_1\cdots w_n|> M\}\not\in \overline{\mathcal{D}}_1.\end{align*} $$

With the notation of the proof of Theorem 4.2, we know that for every $p\geq 1$ ,

$$ \begin{align*}{C}_{2^p}\subset \bigcup_{q \ge p} (b_q\mathbb{N}+[-q,q])\cup \bigcup_{q\ge 1}\bigcup_{u\in {A}_{2q}}I_u^{a,4\varepsilon}.\end{align*} $$

Moreover, by assumption on $(b_q)$ , we have

$$ \begin{align*}\lim_{p\to\infty} \overline{\text{dens}}\bigg(\bigcup_{q \ge p} (b_q\mathbb{N}+[-q,q])\bigg)=0;\end{align*} $$

and we also have

$$ \begin{align*}\bigcup_{q\ge 1}\bigcup_{u\in {A}_{2q}}I_u^{a,4\varepsilon}\subset \bigcup_{u\ge 1}I_u^{a,4\varepsilon}.\end{align*} $$

Since

$$ \begin{align*}\overline{\text{dens}}\bigg(\bigcup_{u\ge 1}I_u^{a,4\varepsilon}\bigg) \le \lim_{u\to\infty} \frac{\sum_{k=1}^u 8\varepsilon a^k}{(1+4\varepsilon)a^u}=\frac{8\varepsilon}{1+4\varepsilon}\sum_{k=0}^{\infty} a^{-k}<1\quad \text{if}\ a\ \text{is sufficiently big},\end{align*} $$

it follows that if p is sufficiently big, then

$$ \begin{align*}\overline{\text{dens}}\ {C}_{2^p}<1.\end{align*} $$

This concludes the proof of Proposition 9.10.

9.3. About disjointness

The original definition of disjointness in topological dynamics goes back to Furstenberg’s seminal paper [Reference Furstenberg25]. The setting is that of compact dynamical systems $(X,T)$ , that is, X is a compact metric space and $T:X\to X$ is a continuous map. Two compact dynamical systems $(X_1, T_1)$ and $(X_2, T_2)$ are said to be disjoint if the only closed, $(T_1\times T_2)$ -invariant set $\Gamma \subset X_1\times X_2$ such that $\pi _{X_1}(\Gamma )=X_1$ and $\pi _{X_2}(\Gamma )=X_2$ is $\Gamma = X_1\times X_2$ . Note that since the spaces are compact, one could replace $\pi _{X_i}(\Gamma )$ by $\overline {\pi _{X_i}(\Gamma )}$ in the definition. For dynamical systems $(X,T)$ whose underlying space is not necessarily compact, both definitions make sense and lead to a priori different notions of disjointness (the one ‘with closure’ being stronger than the one ‘without closure’). In particular, one could consider these notions in the linear setting. However, there are no disjoint pairs of linear dynamical systems in this sense, even ‘without closures’. Indeed, if $T_1\in \mathfrak L(X_1)$ and $X_2\in \mathfrak L(X_2)$ , then $\Gamma :=( \{ 0\}\times X_2)\, \cup \,( X_1\times \{ 0\})$ shows that disjointness cannot be met. One can get round this difficulty by changing the definitions a little bit as follows: instead of $\pi _{X_i}(\Gamma )=X_i$ , require that ${\pi _{X_i}( \Gamma \cap (X_1\setminus \{ 0\})\times (X_2\setminus \{ 0\}))=X_i\setminus \{ 0\}}$ ; and likewise for the definition ‘with closures’.

Even though these definitions of disjointness are likely to be artificial, one can try to play with them a little. For example, copying out the relevant parts of [Reference Furstenberg25]—namely, the proofs of Theorems II.1 and II.2—one gets the following results. Let us say that a linear dynamical system $(X,T)$ is minimal apart from $0$ if every non-zero vector $x\in X$ is hypercyclic for T; equivalently, if the only closed T-invariant subsets of X are $\{ 0\}$ and X. Famous examples of Read [Reference Read50] show that this can indeed happen.

Proposition 9.11. Let $(X_1, T_1)$ and $(X_2, T_2)$ be two linear dynamical systems. If $(X_1, T_1)$ and $(X_2, T_2)$ are disjoint ‘without closures’, then at least one of them is minimal apart from $0$ .

Proof. Assume that $(X_1, T_1)$ and $(X_2, T_2)$ are not minimal apart from $0$ . Then, for $i=1,2$ , one can find a closed $T_i$ -invariant set $C_i\subset X_i$ such that $C_i\neq X_i$ and ${C_i\cap (X_i\setminus \{ 0\})\neq \emptyset }$ ; and $\Gamma := (C_1\times X_2) \cup (X_1\times C_2)$ shows that $(X_1, T_1)$ and $(X_2, T_2)$ are not disjoint ‘without closures’.

Proposition 9.12. Let $(X_1, T_1)$ and $(X_2, T_2)$ be two linear dynamical systems. Assume that the periodic points of $T_1$ are dense in $X_1$ and that $(X_2, T_2)$ is minimal apart from $0$ . Then, $(X_1, T_1)$ and $(X_2, T_2)$ are disjoint ‘without closures’.

Proof. Let $\Gamma \subset X_1\times X_2$ be a closed, $(T_1\times T_2)$ -invariant set such that $\pi _{X_i}( \Gamma \cap (X_1\setminus \{ 0\})\times (X_2\setminus \{ 0\}))=X_i\setminus \{ 0\}$ for $i=1,2$ . We have to show that $\Gamma =X_1\times X_2$ ; and since the periodic points of $T_1$ are dense in $X_1$ , it is enough to show that $(\mathrm { Per}(T_1)\setminus \{ 0\})\times {X_2\subset \Gamma }$ .

Let $u\in X_1$ be any non-zero periodic point of $T_1$ and choose $d\in \mathbb {N}$ such that $T^du=u$ . By assumption on $\Gamma $ , one can find $v\in X_2\setminus \{ 0\}$ such that $(u,v)\in \Gamma $ . Then, ${(u, T_2^{dn}v)\in \Gamma }$ for all $n\in \mathbb {N}$ by $(T_1\times T_2)$ -invariance of $\Gamma $ . Moreover, $v\in \mathrm {HC}(T_2)$ by assumption on $(X_2,T_2)$ . Hence, by Ansari’s theorem [Reference Ansari3], v is also a hypercyclic vector for $T_2^d$ ; and since $\Gamma $ is closed in $X_1\times X_2$ , it follows that $\{ u\}\times X_2\subset \Gamma $ .

Proposition 9.12 implies in particular that linear dynamical systems which are disjoint ‘without closure’ do exist. We do not know if this is also true ‘with closure’; so one could think of possible weakenings of the definition of disjointness ‘with closures’. For hypercyclic operators, one possible such weakening could be the following: one could say that two hypercyclic operators $T_1\in \mathfrak L(X_1)$ and $T_2\in \mathfrak L(X_2)$ are pseudo-disjoint (just to give a name) if, whenever $x_1$ is a hypercyclic vector for $T_1$ and $x_2$ is a hypercyclic vector for $T_2$ , it follows that $(x_1, x_2)$ is hypercyclic for $T_1\times T_2$ . This is indeed weaker than the definition of disjointness ‘with closures’ (consider $\Gamma :=\overline {\mathrm {Orb}((x_1, x_2), T_1\times T_2)}\,$ ), yet formally much stronger than the disjointness notion introduced in [Reference Bernal-González10, Reference Bès and Peris14], that is, diagonal hypercyclicity. We are not much further ahead since we do not know if there are any pseudo-disjoint pairs of linear operators (whereas there are lots of interesting examples for diagonal hypercyclicity). So we ask the following question.

Question 9.13. Are there any pseudo-disjoint pairs of operators, that is, pairs of hypercyclic operators $(T_1,T_2)$ such that $\mathrm {HC}(T_1)\times \mathrm {HC}(T_2)\subset \mathrm {HC}(T_1\times T_2)$ ?

Regarding this question, one may observe that two linear operators $T_1$ and $T_2$ are trivially pseudo-disjoint if it happens that every vector $x\in X_1\times X_2$ with non-zero coordinates is hypercyclic for $T_1\times T_2$ . This leads to the following ‘strong’ form of Question 9.13.

Question 9.14. Are there pairs of operators $(T_1,T_2)$ such that every $x\in (X_1\setminus \{ 0\})\times (X_2\setminus \{ 0\})$ is hypercyclic for $T_1\times T_2$ ?

Let us point out the following amusing fact: if such a pair $(T_1,T_2)$ can be found, then the operator $T=T_1\times T_2$ acting on $X=X_1\times X_2$ is a hypercyclic operator such that $HC(T)$ is an open set but $HC(T)\neq X\setminus \{ 0\}$ . We do not know of any example of operators with that property.

Finally, we note that if one extends the definition of pseudo-disjointness to possibly nonlinear systems in the obvious way, it follows from the main result of [Reference Shkarin52] that any irrational rotation of the circle is pseudo-disjoint from any hypercyclic operator. In view of that, one may consider the following variant of Question 9.13.

Question 9.15. Are there natural classes of hypercyclic operators $\mathfrak C_1$ , $\mathfrak C_2$ such that any $T_1\in \mathfrak C_1$ is pseudo-disjoint from any $T_2\in \mathfrak C_2$ ?

9.4. Other questions

We conclude the paper by adding some other possibly interesting questions motivated by the results obtained in the paper.

The first question asks for a converse to Observation 1.3.

Question 9.16. Let $\mathcal F$ be a Furstenberg family and let $T\in \mathfrak L(X)$ . Assume that $S\oplus T$ is $\mathcal F$ -hypercyclic for every $\mathcal F$ -hypercyclic operator S. Does it follow that T is hereditarily $\mathcal F$ -hypercyclic?

The next two questions are related to an already mentioned ‘trap’ into which it is easy to fall: if $\mathcal F$ and $\mathcal F'$ are two Furstenberg families, the fact that $\mathcal F\subset \mathcal F'$ does not formally imply that hereditary $\mathcal F$ -hypercyclicity is a stronger property than hereditary $\mathcal F'$ -hypercyclicity.

Question 9.17. Are there hereditarily frequently hypercyclic operators which are not topologically mixing?

Question 9.18. Does hereditary frequent hypercyclicity imply hereditary $\mathcal U$ -frequent hypercyclicity?

In the theory of frequently hypercyclic operators, there are non-trivial counterexamples to some tempting ‘conjectures’. It is natural to ask if these examples are in fact hereditarily frequently hypercyclic or if it is possible to modify them to get hereditarily frequently hypercyclic examples. In particular, with [Reference Badea and Grivaux4, Reference Menet42] in mind, this leads to the following questions (the second one is a strengthening of Question 9.17).

Question 9.19. Are there invertible hereditarily frequently hypercyclic operators whose inverse is not frequently hypercyclic?

Question 9.20. Are there operators which are both hereditarily frequently hypercyclic and chaotic but not topologically mixing?

The next two questions are related to the sufficient conditions we found for hereditary frequent hypercyclicity. Observe first that in addition to the FHCC and the unimodular eigenvectors machinery, there are other criteria to prove frequent hypercyclicity (see [Reference Bès, Menet, Peris and Puig11] or [Reference Grivaux, Matheron and Menet31, Theorem 5.35]). They do not imply hereditarily frequent hypercyclicity. Indeed, the criterion of [Reference Bès, Menet, Peris and Puig11] is equivalent to frequent hypercyclicity for weighted shifts on $c_0$ , whereas the C-type operator of Theorem 5.2 satisfies [Reference Grivaux, Matheron and Menet31, Theorem 5.35].

Question 9.21. If $T\in \mathfrak L(X)$ is such that the $\mathbb T$ -eigenvectors of T are spanning with respect to the Lebesgue measure, does it follow that one can find a T-invariant measure $\mu $ with full support such that $(X,\mathcal B,\mu , T)$ is a factor of a dynamical system with countable Lebesgue spectrum? Does it follow at least that T is hereditarily frequently hypercyclic?

Question 9.22. Let $T\in \mathfrak L(X)$ . Is it true that if the $\mathbb {T}$ -eigenvectors of T are perfectly spanning, then T is hereditarily frequently hypercyclic?

Concerning the invariant measure business, the next two questions seem natural. The first one is motivated by Theorem 3.24.

Question 9.23. Does there exist an operator T which is not hereditarily $\mathcal U$ -frequently hypercyclic but admits an ergodic measure with full support?

Question 9.24. If X is a reflexive Banach space, then any frequently hypercyclic operator T on X admits a continuous invariant probability measure with full support (see [Reference Grivaux and Matheron30]). Is it possible to improve this result if T is assumed to be hereditarily frequently hypercyclic?

The next question is, of course, strongly reminiscent of the Bès–Peris theorem [Reference Bès and Peris13], according to which the hypercyclicity criterion characterizes topological weak mixing.

Question 9.25. Given a Furstenberg family $\mathcal F$ , is there some ‘ $\mathcal F$ -hypercyclicity criterion’ characterizing the operators T such that $T\oplus T$ is $\mathcal F$ -hypercyclic?

Finally, our last three questions concern the links between (hereditary) frequent hypercyclicity and the geometry of the underlying space X.

Question 9.26. On which spaces X is it possible to find hereditarily frequently hypercyclic operators? Is it possible at least on any complex Banach space admitting an unconditional Schauder decomposition?

Question 9.27. Are there spaces X which support frequently hypercyclic operators, but no hereditarily frequently hypercyclic operator?

Question 9.28. On which Banach spaces X is it possible to construct d-frequently hypercyclic pairs $(T_1,T_2)$ which are not densely d-frequently hypercyclic?

Acknowledgements

We gratefully acknowledge the support of the Mathematisches Forschungsinstitut Oberwolfach, where part of this work was carried out. The second author was also supported in part by the Labex CEMPI (ANR-11-LABX-0007-01). The fourth author is a Research Associate of the Fonds de la Recherche Scientifique - FNRS. Finally, we would like to thank the anonymous referee for suggesting two questions and helping us to improve the presentation in §3.

References

Agneessens, K.. Frequently hypercyclic random vectors. Proc. Amer. Math. Soc. 151 (2023), 11031117.CrossRefGoogle Scholar
Albiac, F. and Kalton, N. J.. Topics in Banach Space Theory. Springer, New York, 2006.Google Scholar
Ansari, S. I.. Hypercyclic and cyclic vectors. J. Funct. Anal. 128 (1995), 374383.Google Scholar
Badea, C. and Grivaux, S.. Unimodular eigenvalues, uniformly distributed sequences and linear dynamics. Adv. Math. 211 (2007), 766793.CrossRefGoogle Scholar
Bayart, F. and Grivaux, S.. Frequently hypercyclic operators. Trans. Amer. Math. Soc. 358 (2006), 50835117.Google Scholar
Bayart, F. and Grivaux, S.. Invariant Gaussian measures for operators on Banach spaces and linear dynamics. Proc. Lond. Math. Soc. (3) 94 (2007), 181210.CrossRefGoogle Scholar
Bayart, F. and Matheron, É.. Dynamics of Linear Operators (Cambridge Tracts in Mathematics, 179). Cambridge University Press, Cambridge, 2009.Google Scholar
Bayart, F. and Matheron, É.. Mixing operators and small subsets of the circle. J. Reine Angew. Math. 715 (2016), 75123.Google Scholar
Bayart, F. and Ruzsa, I.. Difference sets and frequently hypercyclic weighted shifts. Ergod. Th. & Dynam. Sys. 35 (2015), 691709.CrossRefGoogle Scholar
Bernal-González, L.. Disjoint hypercyclic operators. Studia Math. 182(2) (2007), 113131.CrossRefGoogle Scholar
Bès, J. P., Menet, Q., Peris, A. and Puig, Y.. Recurrence properties of hypercyclic operators. Math. Ann. 366 (2016), 545572.Google Scholar
Bès, J. P., Menet, Q., Peris, A. and Puig, Y.. Strong transitivity properties for operators. J. Differential Equations 266 (2019), 13131337.Google Scholar
Bès, J. P. and Peris, A.. Hereditarily hypercyclic operators. J. Funct. Anal. 167 (1999), 94112.CrossRefGoogle Scholar
Bès, J. P. and Peris, A.. Disjointness in hypercyclicity. J. Math. Anal. Appl. 336 (2007). 297315.Google Scholar
Blum, J. R. and Hanson, D. L.. On the mean ergodic theorem for subsequences. Bull. Amer. Math. Soc. 66 (1960), 308311.CrossRefGoogle Scholar
Bonilla, A. and Grosse-Erdmann, K.-G.. Frequently hypercyclic operators and vectors. Ergod. Th. & Dynam. Sys. 27 (2007), 383404.CrossRefGoogle Scholar
Bonilla, A. and Grosse-Erdmann, K.-G.. Upper frequent hypercyclicity and related notions. Rev. Mat. Complut. 31 (2018), 673711.CrossRefGoogle Scholar
Cardeccia, R. and Muro, S.. Frequently recurrent backward shifts. Preprint, 2024, arXiv:2407.11799.Google Scholar
Conway, J. B.. A Course in Functional Analysis (Graduate Texts in Mathematics, 96). Springer, New York, 1985.Google Scholar
Conze, J.-P.. Convergence des moyennes ergodiques pour des sous-suites. Mém. Soc. Math. Fr. (N.S.) 35 (1973), 715.Google Scholar
Cornfeld, I. P., Fomin, S. V. and Sinai, Ya. G.. Ergodic Theory (Grundlehren der mathematischen Wissenschaften, 245). Springer, New York, 1982.CrossRefGoogle Scholar
De la Rosa, M., Frerick, L., Grivaux, S. and Peris, A.. Frequent hypercyclicity, chaos, and unconditional Schauder decompositions. Israel J. Math. 190 (2012), 389399.Google Scholar
De la Rosa, M. and Read, C.. A hypercyclic operator whose direct sum ${T}\oplus {T}$ is not hypercyclic. J. Operator Theory 61 (2009), 369380.Google Scholar
Ernst, R., Esser, C. and Menet, Q.. $\mathcal{U}$ -frequent hypercyclicity notions and related weighted densities. Israel J. Math. 241 (2021), 817848.CrossRefGoogle Scholar
Furstenberg, H., Disjointness in ergodic theory, minimal sets, and a problem in Diophantine approximation. Math. Systems Theory 1 (1967), 149.Google Scholar
Glasner, E.. Classifying dynamical systems by their recurrence properties. Topol. Methods Nonlinear Anal. 24 (2004), 2140.Google Scholar
Grivaux, S.. Hypercyclic operators, mixing operators and the bounded steps problem. J. Operator Theory 54 (2005), 147168.Google Scholar
Grivaux, S.. A probabilistic version of the frequent hypercyclicity criterion. Studia Math. 176 (2006), 279290.CrossRefGoogle Scholar
Grivaux, S.. A new class of frequently hypercyclic operators. Indiana Univ. Math. Math. J. 60 (2011), 11771201.Google Scholar
Grivaux, S. and Matheron, É.. Invariant measures for frequently hypercyclic operators. Adv. Math. 265 (2014) 371427 Google Scholar
Grivaux, S., Matheron, É. and Menet, Q.. Linear dynamical systems on Hilbert spaces: typical properties and explicit examples. Mem. Amer. Math. Soc. 269 (2021), 147 pp.Google Scholar
Grosse-Erdmann, K.. Frequently hypercyclic bilateral weighted shifts. Glasg. Math. J. 61 (2019), 271286.CrossRefGoogle Scholar
Grosse-Erdmann, K. and Peris, A., Linear Chaos (Universitext). Springer, London, 2011.CrossRefGoogle Scholar
Janson, S., Gaussian Hilbert Spaces (Cambridge Tracts in Mathematics, 129). Cambridge University Press, Cambridge, 1997.CrossRefGoogle Scholar
Ledoux, M. and Talagrand, M.. Probability in Banach Spaces (Ergebnisse der Mathematik und ihrer Grenzgebiete, 23). Springer, Berlin, 1991.Google Scholar
Martin, Ö., Menet, Q. and Puig, Y.. Disjoint frequently hypercyclic pseudo-shifts. J. Funct. Anal. 283 (2022), 109474.CrossRefGoogle Scholar
Martin, Ö. and Puig, Y.. Existence of disjoint frequently hypercyclic operators which fail to be disjoint weakly mixing. J. Math. Anal. Appl. 500 (2021), 125106.CrossRefGoogle Scholar
Martin, Ö. and Sanders, R.. Disjoint supercyclic weighted shifts. Integral Equ. Oper. Theory 85 (2016), 191220.CrossRefGoogle Scholar
Martin, Ö. and Sanders, R., Extending families of disjoint hypercyclic operators. J. Math. Anal. Appl. 539 (2024), 128539 Google Scholar
Maurey, B. and Pisier, G.. Séries de variables aléatoires vectorielles indépendantes et propriétés géométriques des espaces de Banach. Studia Math. 58 (1976), 4590.Google Scholar
Menet, Q.. Linear chaos and frequent hypercyclicity. Trans. Amer. Math. Soc. 369 (2017), 49774994.CrossRefGoogle Scholar
Menet, Q.. Inverse of frequently hypercyclic operators. J. Inst. Math. Jussieu 21 (2022), 18671886.CrossRefGoogle Scholar
Moothathu, T. K. S.. Two remarks on frequent hypercyclicity. J. Math. Anal. Appl. 408 (2013), 843845.CrossRefGoogle Scholar
Murillo-Arcila, M. and Peris, A.. Strong mixing measures for linear operators and frequent hypercyclicity. J. Math. Anal. Appl. 398 (2013), 462465.Google Scholar
Newton, D. and Parry, W.. On a factor automorphism of a normal dynamical system. Ann. Math. Statist. 37 (1966), 15281533.CrossRefGoogle Scholar
Nicol, M. and Petersen, K.. Ergodic theory: basic examples and constructions. Mathematics of Complexity and Dynamical Systems. Ed. R. A. Meyers. Springer, New York, 2012, pp. 264288.CrossRefGoogle Scholar
Parry, W.. Topics in Ergodic Theory. Cambridge University Press, Cambridge, 1981.Google Scholar
Peller, V.. Hankel Operators and Their Applications (Springer Monographs in Mathematics). Springer, New York, 2003.CrossRefGoogle Scholar
Queffélec, M., Substitution Dynamical Systems Spectral Analysis (Lecture Notes in Mathematics, 1294). 2nd edn. Springer, 2010.Google Scholar
Read, C. J.. The invariant subspace problem for a class of Banach spaces. II. Hypercyclic operators. Israel J. Math. 63 (1988), 140.CrossRefGoogle Scholar
Sanders, R. and Shkarin, S.. Existence of disjoint weakly mixing operators that fail to satisfy the Disjoint Hypercyclicity Criterion. J. Math. Anal. Appl. 417 (2014), 834855.Google Scholar
Shkarin, S.. Universal elements for nonlinear operators and their applications. J. Math. Anal. Appl. 348 (2008), 193210.Google Scholar
Shkarin, S.. On the spectrum of frequently hypercyclic operators. Proc. Amer. Math. Soc. 137 (2009), 123134.CrossRefGoogle Scholar
Shkarin, S.. A short proof of existence of disjoint hypercyclic operators. J. Math. Anal. Appl. 367 (2010), 713715.CrossRefGoogle Scholar
Szőkefalvi-Nagy, B., Foias, C., Bercovici, H. and Kérchy, L.. Harmonic Analysis of Operators on Hilbert Space. Springer, New York, 2010.CrossRefGoogle Scholar
Urbański, M., Roy, M. and Munday, S.. Non-invertible Dynamical Systems. Vol. 1. De Gruyter, Berlin, 2022.Google Scholar
Walters, P.. An Introduction to Ergodic Theory (Graduate Texts in Mathematics, 79). Springer, Berlin, 1982.CrossRefGoogle Scholar