1 Introduction and preliminaries

The (discrete) generalized Cesàro operators \(C_t\), for \(t\in [0,1]\), were first investigated by Rhaly [25, 26]. The action of \(C_t\) from the sequence space \(\omega := {{\mathbb {C}}}^{{{\mathbb {N}}}_0}\) into itself, with \({{\mathbb {N}}}_0:=\{0,1,2,\ldots \}\), is given by

$$\begin{aligned} C_tx:=\left( \frac{t^nx_0+t^{n-1}x_1+\cdots +x_n}{n+1}\right) _{n\in {{\mathbb {N}}}_0},\quad x=(x_n)_{n\in {{\mathbb {N}}}_0}\in \omega . \end{aligned}$$
(1.1)

For \(t=0\) and with \(\varphi :=(\frac{1}{n+1})_{n\in {{\mathbb {N}}}_0}\) note that \(C_0\) is the diagonal operator

$$\begin{aligned} D_\varphi x:= \left( \frac{x_n}{n+1}\right) _{n\in {{\mathbb {N}}}_0}, \quad x=(x_n)_{n\in {{\mathbb {N}}}_0}\in \omega , \end{aligned}$$
(1.2)

and, for \(t=1\), that \(C_1\) is the classical Cesàro averaging operator

$$\begin{aligned} C_1x:=\left( \frac{x_0+x_1+\cdots +x_n}{n+1}\right) _{n\in {{\mathbb {N}}}_0},\quad x=(x_n)_{n\in {{\mathbb {N}}}_0}\in \omega . \end{aligned}$$
(1.3)

The behaviour of \(C_t\) on various sequence spaces has been investigated by many authors. We refer the reader to [25,26,27], to the recent papers [28, 30, 31] and to the introduction of the papers [5, 13] and the references therein. The operator \(C_1\) was thoroughly investigated on weighted Banach spaces in [2]; see also [12]. Certain variants of the Cesàro operator \(C_1\) are considered in [9, 16].

Our aim is to investigate the operators \(C_t\), for \(t\in [0,1]\), when they are suitably interpreted to act on the space \(H({{\mathbb {D}}})\) of holomorphic functions on the open unit disc \({{\mathbb {D}}}:=\{z\in {{\mathbb {C}}}\,:\ |z|<1\}\), on the Banach space \(H^\infty \) of bounded analytic functions and on the weighted Banach spaces \(H_v^\infty \) and \(H_v^0\) with their sup-norms. The space \(H({{\mathbb {D}}})\) is equipped with the topology \(\tau _c\) of uniform convergence on the compact subsets of \({{\mathbb {D}}}\). According to [21, §27.3(3)] the space \(H({{\mathbb {D}}})\) is a Fréchet–Montel space. A family of norms generating \(\tau _c\) is given, for each \(0<r<1\), by

$$\begin{aligned} q_r(f):=\sup _{|z|\le r}|f(z)|,\quad f\in H({{\mathbb {D}}}). \end{aligned}$$
(1.4)

A weight v is a continuous, non-increasing function \(v:[0,1)\rightarrow (0,\infty )\). We extend v to \({{\mathbb {D}}}\) by setting \(v(z):=v(|z|)\), for \(z\in {{\mathbb {D}}}\). Note that \(v(z)\le v(0)\) for all \(z\in {{\mathbb {D}}}\). Given a weight v on [0, 1), we define the corresponding weighted Banach spaces of analytic functions on \({{\mathbb {D}}}\) by

$$\begin{aligned} H_v^\infty :=\{f\in H({{\mathbb {D}}}):\, \Vert f\Vert _{\infty ,v}:=\sup _{z\in {{\mathbb {D}}}}|f(z)|v(z)<\infty \}, \end{aligned}$$

and

$$\begin{aligned} H^0_v:=\{f\in H({{\mathbb {D}}}):\, \lim _{|z|\rightarrow 1^-}|f(z)|v(z)=0\}, \end{aligned}$$

both endowed with the norm \(\Vert \cdot \Vert _{\infty ,v}\). Since \(\Vert f\Vert _{\infty ,v}\le v(0)\Vert f\Vert _\infty \) whenever \(f\in H^\infty \), it is clear that \(H^\infty \subseteq H^\infty _v\) with a continuous inclusion. If \(v(z)=1\) for all \(z\in {{\mathbb {D}}}\), then \(H^\infty _v\) coincides with the space \(H^\infty \) of all bounded analytic functions on \({{\mathbb {D}}}\) with the sup-norm \(\Vert \cdot \Vert _\infty \) and \(H^0_v\) reduces to \(\{0\}\). Moreover, \(H^\infty _v\subseteq H({{\mathbb {D}}})\) continuously. Indeed, fix \(0<r<1\). Then \(\frac{1}{v(0)}\le \frac{1}{v(z)}\le \frac{1}{v(r)}\) for \(|z|\le r\) and so (1.4) implies that

$$\begin{aligned} q_r(f)=\sup _{|z|\le r}\frac{v(z)|f(z)|}{v(z)}\le \frac{1}{v(r)}\sup _{|z|\le r}v(z)|f(z)|\le \frac{1}{v(r)}\Vert f\Vert _{\infty , v},\quad f\in H^\infty _v. \end{aligned}$$

We refer the reader to [10] for a recent survey of such types of weighted Banach spaces and operators between them.

Whenever necessary we will identify a function \(f\in H({{\mathbb {D}}})\) with its sequence of Taylor coefficients \({\hat{f}}:=({\hat{f}}(n))_{n\in {{\mathbb {N}}}_0}\) (i.e., \({\hat{f}}(n):=\frac{f^{(n)}(0)}{n!}\), for \(n\in {{\mathbb {N}}}_0\)), so that \(f(z)=\sum _{n=0}^\infty {\hat{f}}(n)z^n\), for \(z\in {{\mathbb {D}}}\). The linear map \(\Phi :H({{\mathbb {D}}})\rightarrow \omega \) is defined by

$$\begin{aligned} \Phi \left( f=\sum _{n=0}^\infty {\hat{f}}(n)z^n \right) :={\hat{f}},\quad f\in H({{\mathbb {D}}}). \end{aligned}$$

It is injective (clearly) and continuous. Indeed, for each \(m\in {{\mathbb {N}}}_0\),

$$\begin{aligned} r_m(x):=\max _{0\le j\le m}|x_j|,\quad x=(x_j)_{j\in {{\mathbb {N}}}_0}\in \omega , \end{aligned}$$

is a continuous seminorm in \(\omega \). Fix \(0<r<1\), in which case

$$\begin{aligned} r_m(\Phi (f))&=\max _{0\le j\le m}|{\hat{f}}(j)|=\max _{0\le j\le m}\left| \frac{1}{2\pi i}\int _{|z|=r}\frac{f(z)}{z^{j+1}}\, dz\right| \le \max _{0\le j\le m}\sup _{|z|=r}\frac{|f(z)|}{|z|^j}\\ {}&=\max _{0\le j\le m}\frac{1}{r^j}q_r(f)= \frac{1}{r^m}q_r(f), \end{aligned}$$

for each \(f\in H({{\mathbb {D}}})\) because \(\frac{1}{r^j}\le \frac{1}{r^m}\) for all \(0\le j\le m\). Of course, the increasing sequence of seminorms \(\{r_m\,\ m\in {{\mathbb {N}}}_0\}\) generates the topology of \(\omega \).

We first provide an integral representation of the generalized Cesàro operators \(C_t\) defined on \(H({{\mathbb {D}}})\), for \(t\in [0,1)\). So, fix \(t\in [0,1)\) and define \(C_t:H({{\mathbb {D}}})\rightarrow H({{\mathbb {D}}})\) by \( C_tf(0):=f(0)\) and

$$\begin{aligned} C_tf(z):=\frac{1}{z}\int _0^z\frac{f(\xi )}{1-t\xi }\,d\xi ,\ z\in {{\mathbb {D}}}\setminus \{0\}, \end{aligned}$$
(1.5)

for every \(f\in H({{\mathbb {D}}})\). It turns out that \(C_t\) is continuous on \(H({{\mathbb {D}}})\); see Proposition 2.1. Moreover, the discrete Cesàro operator \(C_t:\omega \rightarrow \omega \), when restricted to the subspace \(\Phi (H({{\mathbb {D}}}))\subseteq \omega \) is transferred to \(H({{\mathbb {D}}})\) as follows. For a fixed \(f\in H({{\mathbb {D}}})\) we have \(f(\xi )=\sum _{n=0}^\infty a_n \xi ^n\), for \(\xi \in {{\mathbb {D}}}\), with \({\hat{f}}=(a_n)_{n\in {{\mathbb {N}}}_0}\) its sequence of Taylor coefficients. Since \(\frac{1}{1-t\xi }=\sum _{n=0}^\infty t^n\xi ^n\), for \(\xi \in {{\mathbb {D}}}\), we can form the Cauchy product of the two series, thereby obtaining

$$\begin{aligned} \frac{f(\xi )}{1-t\xi }=\sum _{n=0}^\infty (\sum _{k=0}^{n}t^{n-k}a_k)\xi ^n, \quad \xi \in {{\mathbb {D}}}. \end{aligned}$$

Then (1.5) yields

$$\begin{aligned} zC_tf(z)=\int _0^z\sum _{n=0}^\infty (\sum _{k=0}^{n}t^{n-k}a_k)\xi ^n\,d\xi =\sum _{n=0}^\infty \left( \frac{t^na_0+t^{n-1}a_1+\cdots +a_n}{n+1}\right) z^{n+1},\ z\in {{\mathbb {D}}}. \end{aligned}$$

The interchange of the infinite sum and the integral is permissible by uniform convergence of the series. This shows that \(C_tf\in H({{\mathbb {D}}})\) also has the series representation

$$\begin{aligned} C_tf(z)&=\sum _{n=0}^\infty \left( \frac{t^na_0+t^{n-1}a_1+\cdots +a_n}{n+1}\right) z^{n}\nonumber \\&=\sum _{n=0}^\infty \left( \frac{t^n{\hat{f}}(0)+t^{n-1}{\hat{f}}(1)+\cdots +{\hat{f}}(n)}{n+1}\right) z^{n}=\sum _{n=0}^\infty (C_t^\omega ({\hat{f}}))_nz^n, \end{aligned}$$
(1.6)

where the coefficients of the series are precisely as in (1.1). For the sake of clarity we will denote the discrete generalized Cesàro operator \(C_t:\omega \rightarrow \omega \) by \(C_t^\omega \) and reserve the notation \(C_t\) for the operator (1.5) acting in \(H({{\mathbb {D}}})\). Note that \(C_0^\omega =D_\varphi \) (see (1.2)). Moreover, \(C_0\) is given by \(C_0f(z)=\frac{1}{z}\int _0^z f(\xi )\,d\xi \) for \(z\not =0\) and \(C_0f(0)=f(0)\), which is the classical Hardy operator in \(H({{\mathbb {D}}})\).

The main results for \(C_t\) when acting in the Fréchet space \(H({{\mathbb {D}}})\) occur in Proposition 2.1 (continuity), Proposition 3.3 (non-compactness), Proposition 3.7 (spectra) and Proposition 3.8 (linear dynamics and mean ergodicity). For the analogous information concerning \(C_t\) when acting in the weighted Banach spaces \(H^\infty _v\) and \(H^0_v\) see Proposition 2.4 and Corollary 2.5 (continuity), Proposition 2.7 (compactness), Proposition 2.8 (spectra) and Proposition 3.2 (linear dynamics and mean ergodicity).

We end this section by recalling a few definitions and some notation concerning locally convex spaces and operators between them. For further details about functional analysis and operator theory relevant to this paper see, for example, [15, 18, 20,21,22, 29].

Given locally convex Haudorff spaces XY (briefly, lcHs) we denote by \({{\mathcal {L}}}(X,Y)\) the space of all continuous linear operators from X into Y. If \(X=Y\), then we simply write \({{\mathcal {L}}}(X)\) for \({{\mathcal {L}}}(X,X)\). Equipped with the topology of pointwise convergence on X (i.e., the strong operator topology) the lcHs \({{\mathcal {L}}}(X)\) is denoted by \({{\mathcal {L}}}_s(X)\). Equipped with the topology \(\tau _b\) of uniform convergence on the bounded subsets of X the lcHs \({{\mathcal {L}}}(X)\) is denoted by \({{\mathcal {L}}}_b(X)\).

Let X be a lcHs space. The identity operator on X is denoted by I. The transpose operator of \(T\in {{\mathcal {L}}}(X)\) is denoted by \(T'\); it acts from the topological dual space \(X':={{\mathcal {L}}}(X,{{\mathbb {C}}})\) of X into itself. Denote by \(X'_\sigma \) (resp., by \(X'_\beta \)) the topological dual \(X'\) equipped with the weak* topology \(\sigma (X',X)\) (resp., with the strong topology \(\beta (X',X)\)); see [21, §21.2] for the definition. It is known that \(T'\in {{\mathcal {L}}}(X'_\sigma )\) and \(T'\in {{\mathcal {L}}}(X_\beta ')\), [22, p. 134]. The bi-transpose operator \((T')'\) of T is simply denoted by \(T''\) and belongs to \({{\mathcal {L}}}((X'_\beta )'_\beta )\).

A linear map \(T:X\rightarrow Y\), with XY lcHs’, is called compact if there exists a neighbourhood \({{\mathcal {U}}}\) of 0 in X such that \(T({{\mathcal {U}}})\) is a relatively compact set in Y. It is routine to show that necessarily \(T\in {{\mathcal {L}}}(X,Y)\). We recall the following well known result; see [20, Proposition 17.1.1], [22, §42.1(1)].

Lemma 1.1

Let X be a lcHs. The compact operators are a 2-sided ideal in \({{\mathcal {L}}}(X)\).

Given a lcHs X and \(T\in {{\mathcal {L}}}(X)\), the resolvent set \(\rho (T;X)\) of T consists of all \(\lambda \in {{\mathbb {C}}}\) such that \(R(\lambda ,T):=(\lambda I-T)^{-1}\) exists in \({{\mathcal {L}}}(X)\). The set \(\sigma (T;X):={{\mathbb {C}}}{\setminus } \rho (T;X)\) is called the spectrum of T. The point spectrum \(\sigma _{pt}(T;X)\) of T consists of all \(\lambda \in {{\mathbb {C}}}\) (also called an eigenvalue of T) such that \((\lambda I-T)\) is not injective. Some authors (eg. [29]) prefer the subset \(\rho ^*(T;X)\) of \(\rho (T;X)\) consisting of all \(\lambda \in {{\mathbb {C}}}\) for which there exists \(\delta >0\) such that the open disc \(B(\lambda ,\delta ):=\{z\in {{\mathbb {C}}}:\, |z-\lambda |<\delta \}\subseteq \rho (T;X)\) and \(\{R(\mu ,T):\, \mu \in B(\lambda ,\delta )\}\) is an equicontinuous subset of \({{\mathcal {L}}}(X)\). Define \(\sigma ^*(T;X):={{\mathbb {C}}}{\setminus } \rho ^*(T;X)\), which is a closed set with \(\sigma (T;X)\subseteq \sigma ^*(T;X)\). For the spectral theory of compact operators in lcHs’ we refer to [15, 18], for linear dynamics to [6, 17] and for mean ergodic operators to [23], for example.

2 Continuity, compactness and spectrum of \(C_t\)

In this section we establish, for \(t\in [0,1)\), the continuity of \(C_t:H({{\mathbb {D}}})\rightarrow H({{\mathbb {D}}})\) as well as the continuity of \(C_t\) from \(H^\infty \) (resp., \(H^\infty _v\)) into \(H^\infty \) (resp., \(H^\infty _v\)). The same is true for \(C_t:H^0_v\rightarrow H^0_v\) whenever \(\lim _{r\rightarrow 1^-}v(r)=0\). It is also shown that the bi-transpose \(C_t''\) of \(C_t\in {{\mathcal {L}}}(H^0_v)\) is the generalized Cesàro operator \(C_t\in {{\mathcal {L}}}(H^\infty _v)\), provided that \(\lim _{r\rightarrow 1^-}v(r)=0\). For such weights v it also turns out that both \(C_t\in {{\mathcal {L}}}(H^0_v)\) and \(C_t\in {{\mathcal {L}}}(H^\infty _v)\) are compact operators (cf. Proposition 2.7); their spectrum is identified in Proposition 2.8. Of particular interest are the standard weights \(v_\gamma (z):=(1-|z|)^\gamma \), for \(\gamma >0\) and \(z\in {{\mathbb {D}}}\).

Proposition 2.1

For every \(t\in [0,1)\) the operator \(C_t:H({{\mathbb {D}}})\rightarrow H({{\mathbb {D}}})\) is continuous. Moreover, the set \(\{C_t:\, t\in [0,1)\}\) is equicontinuous in \({{\mathcal {L}}}(H({{\mathbb {D}}}))\).

Proof

Fix \(f\in H({{\mathbb {D}}})\). Taking into account that \(C_tf(0)=f(0)\), for all \(t\in [0,1)\) and, for each \(r\in (0,1)\), that \(\sup _{|z|\le r}|C_tf(z)|=\sup _{|z|=r}|C_tf(z)|\), the formula (1.5) implies, for each \(z\in {{\mathbb {D}}}{\setminus }\{0\}\), that

$$\begin{aligned} |C_tf(z)|&=\frac{1}{|z|}\left| \int _0^z\frac{f(\xi )}{1-t\xi }d\xi \right| \le \frac{1}{|z|}|z|\max _{\xi \in [0,z]}\frac{|f(\xi )|}{|1-t\xi |}\\ {}&\le \frac{1}{1-|z|}\max _{|\xi |\le |z|}|f(\xi )|=\frac{1}{1-|z|}\max _{|\xi |=|z|}|f(\xi )|, \end{aligned}$$

because \(|1-t\xi |\ge 1-t|\xi |\ge 1-|\xi |\ge 1-|z|\), for all \(|\xi |\le |z|\). It follows from the previous inequality, for each \(r\in (0,1)\), that

$$\begin{aligned} q_r(C_tf)=\sup _{|z|\le r}|C_tf(z)|\le \frac{1}{1-r}\sup _{|\xi |\le r}|f(\xi )|=\frac{1}{1-r}q_r(f);. \end{aligned}$$

see (1.4). This implies the result. \(\square \)

The following example will prove to be useful in the sequel.

Example 2.2

Consider the constant function \(f_1(z):=1\), for every \(z\in {{\mathbb {D}}}\), in which case \(C_tf_1(0)=f_1(0)=1\) for every \(t\in [0,1]\). For \(t=0\), it was noted in Sect. 1 that \(C_0\) is

the Hardy operator. In particular, \(C_0f_1(z)=1\), for every \(z\in {{\mathbb {D}}}\). For \(t\in (0,1]\), note that \(C_tf_1(0)=1\) and

$$\begin{aligned} C_tf_1(z)=\frac{1}{z}\int _0^z \frac{d\xi }{1-t\xi }=-\frac{1}{tz}\log (1-tz),\quad z\in {{\mathbb {D}}}{\setminus }\{0\}. \end{aligned}$$

For \(t=1\) this shows, in particular, that \(C_1(H^\infty )\not \subset H^\infty \), which is well known. For an investigation of the operator \(C_1\) acting in \(H^\infty \) we refer to [14].

Concerning \(t\in (0,1)\), recall the Taylor series expansion

$$\begin{aligned} -\log (1-z)=z\sum _{n=0}^\infty \frac{z^n}{n+1},\quad z\in {{\mathbb {D}}}, \end{aligned}$$

from which it follows that

$$\begin{aligned} -\frac{\log (1-tz)}{tz}=\sum _{n=0}^\infty \frac{t^n}{n+1}z^n, \quad z\in {{\mathbb {D}}}{\setminus }\{0\}, \end{aligned}$$

with the series having radius of convergence \(\frac{1}{t}>1\). The claim is that \(\Vert C_tf_1\Vert _\infty =\sup _{|z|<1}|C_tf_1(z)|=-\frac{\log (1-t)}{t}\). Indeed, \(C_tf_1\) is clearly holomorhic in \(B(0,\frac{1}{t}):=\{\xi \in {{\mathbb {C}}}:\, |\xi |<\frac{1}{t}\}\) hence, continuous in \(B(0,\frac{1}{t})\), and satisfies \(C_tf_1(1)=-\frac{\log (1-t)}{t}\) with \(\lim _{r\rightarrow 1^-}C_tf_1(r)=C_tf_1(1)\). On the other hand, for every \(z\in {{\mathbb {D}}}{\setminus }\{0\}\) and \(t\in (0,1)\) we have that

$$\begin{aligned} |C_tf_1(z)|=\left| -\frac{\log (1-tz)}{tz}\right| \le \sum _{n=0}^\infty \frac{t^n}{n+1}|z|^n\le \sum _{n=0}^\infty \frac{t^n}{n+1}=-\frac{\log (1-t)}{t}. \end{aligned}$$

This completes the proof of the claim. Observe that \(\Vert C_tf_1\Vert _\infty >1\). Indeed, define \(\gamma (t)=-\log (1-t)-t\), for \(t\in [0,1)\). Then \(\gamma (0)=0\), \(\lim _{t\rightarrow 1^-}\gamma (t)=\infty \) and \(\gamma '(t)=\frac{1}{1-t}-1=\frac{t}{1-t}\), for \(t\in [0,1)\). Since \(\gamma '(t)>0\), for \(t\in (0,1)\), it follows that \(\gamma \) is strictly increasing and so \(\gamma (t)>0\) for all \(t\in (0,1)\). This implies that \(\Vert C_tf_1\Vert _\infty =-\frac{\log (1-t)}{t}>1\) for every \(t\in (0,1)\). On the other hand, for \(t\in (0,1)\), the inequality \(\sum _{n=0}^\infty t^n/(n+1)<\sum _{n=0}^\infty t^n\) implies that \(-\frac{\log (1-t)}{t}<\frac{1}{1-t}\). So, we have shown that \(\Vert C_0f_1\Vert _\infty =1\) and

$$\begin{aligned} 1<\Vert C_tf_1\Vert _\infty <\frac{1}{1-t},\quad t\in (0,1). \end{aligned}$$

We now turn to the action of \(C_t\) in various Banach spaces. For \(t=1\) it was noted above that \(C_1\) fails to act in \(H^\infty \).

Proposition 2.3

For \(t\in [0,1)\) the operator \(C_t:H^\infty \rightarrow H^\infty \) is continuous. Moreover, \(\Vert C_0\Vert _{H^\infty \rightarrow H^\infty }=1\) and

$$\begin{aligned} \Vert C_t\Vert _{H^\infty \rightarrow H^\infty }=-\frac{\log (1-t)}{t},\quad t\in (0,1). \end{aligned}$$

Proof

Let \(f\in H^\infty \) be fixed. Then

$$\begin{aligned} |C_0f(z)|=\left| \frac{1}{z}\int _0^zf(\xi )d\xi \right| \le \max _{|\xi |\le |z|} |f(\xi )|\le \Vert f\Vert _\infty . \end{aligned}$$

This implies that \(\Vert C_0\Vert _{H^\infty \rightarrow H^\infty }\le 1\). On the other hand, \(C_0f_1=f_1\) and so we can conclude that \(\Vert C_0\Vert _{H^\infty \rightarrow H^\infty }=1\).

Now let \(t\in (0,1)\). Then, for the parametrization \(\xi :=sz\), for \(s\in (0,1)\), it follows from \(|1-stz|\ge 1-|stz|\ge 1-st\) that

$$\begin{aligned} |C_tf(z)|&=\left| \frac{1}{z}\int _0^z\frac{f(\xi )}{1-t\xi }d\xi \right| =\left| \int _0^1\frac{f(sz)}{1-stz}ds\right| \le \max _{|\xi |\le |z|}|f(\xi )|\int _0^1\frac{ds}{1-st|z|}\\&\le \Vert f\Vert _\infty \int _0^1\frac{ds}{1-st}=-\frac{\log (1-t)}{t}\Vert f\Vert _\infty . \end{aligned}$$

So, \(C_t\in {{\mathcal {L}}}(H^\infty )\) with \(\Vert C_t\Vert _{H^\infty \rightarrow H^\infty }\le -\frac{\log (1-t)}{t}\). But, \(\Vert C_tf_1\Vert _\infty = -\frac{\log (1-t)}{t}\). Accordingly, \(\Vert C_t\Vert _{H^\infty \rightarrow H^\infty }= -\frac{\log (1-t)}{t}\). \(\square \)

Proposition 2.4

Let v be a weight function on [0, 1). For each \(t\in [0,1)\) the operator \(C_t:H_v^\infty \rightarrow H_v^\infty \) is continuous. Moreover, \(\Vert C_0\Vert _{H^\infty _v\rightarrow H_v^\infty }=1\) and

$$\begin{aligned} 1\le \Vert C_t\Vert _{H^\infty _v\rightarrow H^\infty _v}\le -\frac{\log (1-t)}{t},\quad t\in (0,1). \end{aligned}$$

Proof

Recall that \(C_tf(0):=f(0)\) for each \(f\in H({{\mathbb {D}}})\) and \(t\in [0,1]\). Fix \(t\in (0,1)\). Given \(f\in H^\infty _v\) and \(z\in {{\mathbb {D}}}{\setminus }\{0\}\), observe that

$$\begin{aligned} v(z)|C_tf(z)|&=\frac{v(z)}{|z|}\left| \int _0^z\frac{f(\xi )}{1-t\xi }d\xi \right| =v(z)\left| \int _0^1\frac{f(sz)}{1-stz}ds\right| \\&\le v(z)\int _0^1\frac{|f(sz)|}{|1-stz|}ds\le \int _0^1\frac{v(sz)|f(sz)|}{|1-stz|}ds\\ {}&\le \Vert f\Vert _{\infty ,v}\int _0^1\frac{ds}{|1-stz|}\le \Vert f\Vert _{\infty ,v}\int _0^1\frac{ds}{1-st|z|}\\&=-\frac{\log (1-t|z|)}{t|z|}\Vert f\Vert _{\infty ,v}, \end{aligned}$$

where we used that \(v(sz)=v(s|z|)\ge v(|z|)=v(z)\), for \(s\in (0,1)\), as v is non-increasing on (0, 1) and that \(|1-stz|\ge 1-st|z|\), for \(s\in (0,1)\). According to the calculations in Example 2.2 we can conclude that

$$\begin{aligned} \Vert C_tf\Vert _{\infty , v}=\sup _{z\in {{\mathbb {D}}}}|C_tf(z)|v(z)\le \Vert f\Vert _{\infty ,v}\sup _{z\in {{\mathbb {D}}}}\left[ -\frac{\log (1-t|z|)}{t|z|}\right] =-\frac{\log (1-t)}{t}\Vert f\Vert _{\infty ,v}. \end{aligned}$$

This implies that \(C_t\in {{\mathcal {L}}}(H^\infty _v)\) and \(\Vert C_t\Vert _{H^\infty _v\rightarrow H^\infty _v}\le -\frac{\log (1-t)}{t}\).

For \(t=0\) observe that

$$\begin{aligned} |C_0f(z)|\le \int _0^1|f(sz)|ds\le \max _{|\xi |\le |z|}|f(\xi )|=\frac{1}{v(z)}\max _{|\xi |=|z|}|f(\xi )|v(\xi )\le \frac{1}{v(z)}\Vert f\Vert _{\infty ,v}, \end{aligned}$$

as \(v(\xi )=v(z)\) whenever \(|\xi |=|z|\) with \(\xi \in {{\mathbb {D}}}\). This shows that \(\Vert C_0\Vert _{H^\infty _v\rightarrow H_v^\infty }\le 1\). Since \(C_0f_1=f_1\), it follows that actually \(\Vert C_0\Vert _{H^\infty _v\rightarrow H_v^\infty }=1\).

It remains to show that \(\Vert C_t\Vert _{H^\infty _v\rightarrow H^\infty _v}\ge 1\) for \(t\in (0,1)\). To this end, fix \(t\in (0,1)\) and consider the function \(g_0(z):=\frac{1}{1-tz}=\sum _{n=0}^\infty t^nz^n\), for \(z\in {{\mathbb {D}}}\). Then \(\Vert g_0\Vert _\infty =\frac{1}{1-t}\) and so \(g_0\in H^\infty \subseteq H^\infty _v\). Moreover, for every \(z\in {{\mathbb {D}}}{\setminus }\{0\}\), it is the case that

$$\begin{aligned} C_tg_0(z)=\frac{1}{z}\int _0^z\frac{d\xi }{(1-t\xi )^2}=\frac{1}{z} \left[ \frac{1}{t(1-t\xi )}\right] _0^z=\frac{1}{tz}\left[ \frac{1}{1-tz}-1\right] =\frac{1}{1-tz}=g_0(z). \end{aligned}$$

It follows that \(\Vert g_0\Vert _{\infty ,v}=\Vert C_tg_0\Vert _{\infty ,v}\le \Vert C_t\Vert _{H^\infty _v\rightarrow H^\infty _v}\Vert g_0\Vert _{\infty ,v}\) which implies that \(\Vert C_t\Vert _{H^\infty _v\rightarrow H^\infty _v}\ge 1\). \(\square \)

Corollary 2.5

Let v be a weight function on [0, 1) satisfying \(\lim _{r\rightarrow 1^-}v(r)=0\). For each \(t\in [0,1)\) the operator \(C_t:H^0_v\rightarrow H^0_v\) is continuous and satisfies \(\Vert C_t\Vert _{H^0_v\rightarrow H^0_v}=\Vert C_t\Vert _{H^\infty _v\rightarrow H^\infty _v}\).

Proof

By Proposition 2.4 and the fact that \(H^0_v\) is a closed subspace of \(H^\infty _v\), to obtain the result it suffices to establish that \(C_t(H_v^0)\subseteq H^0_v\). To this effect, observe that \(H^\infty \subseteq H^0_v\) and that \(H^\infty \) is dense in \(H^0_v\), as the space of polynomials is dense in \(H^0_v\); see Section 1 of [11] and also [7]. Proposition 2.3 implies that \(C_t(H^\infty )\subseteq H^\infty \subseteq H^0_v\). Since \(C_t\) acts continuously on \(H^\infty _v\), it follows that

$$\begin{aligned} C_t(H^0_v)=C_t(\overline{H^\infty })\subseteq \overline{C_t(H^\infty )}\subseteq H^0_v. \end{aligned}$$

Moreover, \(\lim _{r\rightarrow 1^-}v(r)=0\) implies that \(H^\infty _v\) is canonically isometric to the bidual of \(H^0_v\), [8, Example 2.1], and that the bi-transpose \(C_t'':H^\infty _v\rightarrow H_v^\infty \) of \(C_t:H^0_v\rightarrow H^0_v\) coincides with \(C_t:H^\infty _v\rightarrow H^\infty _v\) (see Lemma 2.6 below), from which the identity \(\Vert C_t\Vert _{H^0_v\rightarrow H^0_v}=\Vert C_t\Vert _{H^\infty _v\rightarrow H^\infty _v}\) follows. \(\square \)

Lemma 2.6

Let v be a weight function on [0, 1) satisfying \(\lim _{r\rightarrow 1^-}v(r)=0\). For each \(t\in [0,1)\), the bi-transpose \(C''_t:H^\infty _v\rightarrow H^\infty _v\) of \(C_t:H^0_v\rightarrow H_v^0\) coincides with \(C_t:H^\infty _v\rightarrow H^\infty _v\).

Proof

By Proposition 2.3 and Corollary 2.5, together with the fact that \(H^\infty _v\) is canonically isometric to the bidual of \(H^0_v\), both of the operators \(C_t''\) and \(C_t\) act continuously on \(H^\infty _v\).

To show that the bi-transpose \(C_t'':H^\infty _v\rightarrow H^\infty _v\) of \(C_t:H^0_v\rightarrow H^0_v\) coincides with \(C_t:H^\infty _v\rightarrow H^\infty _v\) we proceed via several steps.

First step Given \(f\in H({{\mathbb {D}}})\), its Taylor polynomials \(p_k(z)=\sum _{j=0}^k{\hat{f}}(j)z^j\), \(z\in {{\mathbb {D}}}\), for \(k\in {{\mathbb {N}}}_0\), converge to f uniformly on compact subsets of \({{\mathbb {D}}}\). That is, \(p_k\rightarrow f\) in \((H({{\mathbb {D}}}),\tau _c)\) as \(k\rightarrow \infty \). Accordingly, the averages of \((p_k)_{k\in {{\mathbb {N}}}_0}\), that is, the Cesàro means \(f_n(z):=\frac{1}{n+1}\sum _{j=0}^np_j(z)\), for \(z\in {{\mathbb {D}}}\) and \(n\in {{\mathbb {N}}}_0\), also converge to f in \((H({{\mathbb {D}}}),\tau _c)\) as \(n\rightarrow \infty \).

Second step Lemma 1.1 in [7] implies, for every \(f\in H^\infty _v\) and \(n\in {{\mathbb {N}}}_0\), that \(\Vert f_n\Vert _{\infty ,v}\le \Vert f\Vert _{\infty ,v}\), where \(f_n\) is the n-th Cesàro mean of f, as defined in the First step. Denote by \(U_v\) the closed unit ball of \((H^\infty _v,\Vert \cdot \Vert _{\infty ,v})\). Then, for any given \(f\in U_v\), its sequence of Cesàro means satisfies \((f_n)_{n\in {{\mathbb {N}}}_0}\subseteq U_v\) and \(f_n\rightarrow f\) in \((H({{\mathbb {D}}}),\tau _c)\) as \(n\rightarrow \infty \).

Third step With the topology of uniform convergence on the compact subsets of \(U_v\) denoted by \(\tau _c\), let \(X:=\{F\in (H_v^\infty )':\ F|_{U_v}\ \text { is } \tau _c-\text { continuous }\}\) be endowed with the norm \(\Vert F\Vert :=\sup \{|F(f)|:\ f\in U_v\}\). Then [8, Theorem 1.1(a)] ensures that \((X,\Vert \cdot \Vert )\) is a Banach space and that the evaluation map \(\Psi :H^\infty _v\rightarrow X'\) defined by \((\Psi (f))(F):=\langle f,F\rangle \), for \(F\in X\) and \(f\in H^\infty _v\), is an isometric isomorphism onto \(X'\) (where \(X'\) is the dual Banach space of \((X,\Vert \cdot \Vert )\)). Moreover, by [8, Theorem 1.1(b) and Example 2.1] the restriction map \(R:X\rightarrow (H^0_v)'\) given by \(F\mapsto F|_{H_v^0}\), is also a surjective isometric isomorphism. Therefore, the spaces \(H^\infty _v\) and \((H^0_v)''\) are isometrically isomorphic, that is, X and \((H^0_v)'\) are isometrically isomorphic and hence, also \(H^\infty _v\) and \((H^0_v)''\) are isometrically isomorphic.

It is easy to see, since the Banach space X above is the predual of \(H^\infty _v\), that the evaluation map \(\delta _z\in X\), for every \(z\in {{\mathbb {D}}}\), where \(\delta _z:f\mapsto f(z)\), for \(f\in H^\infty _v\), satisfies \(|\langle f,\delta _z\rangle |\le \Vert f\Vert _{\infty ,v}/v(z)\). In particular, the linear span L of the set \(\{\delta _z:\ z\in {{\mathbb {D}}}\}\) separates the points of \(H^\infty _v=X'\) and hence, L is dense in X. Therefore, the pointwise convergence topology \(\tau _p\) on \(H^\infty _v\) is Hausdorff and coarser than the \(w^*\)-topology \(\sigma (H^\infty _v, X)\).

Fourth step The closed unit ball \(U_v\) of \(H^\infty _v\) is a \(\tau _c\)-compact set by Montel’s theorem, as it is \(\tau _c\)-bounded and closed. On the other hand, \(U_v\) is also \(\sigma (H^\infty _v,X)\)-compact by the Alaoglu-Bourbaki theorem. Since \(\tau _p|_{U_v}\) is coarser than \(\tau _c|_{U_v}\) and Hausdorff, we can conclude that \(\tau _p|_{U_v}=\tau _c|_{U_v}\). In the same way, it follows that \(\tau _p|_{U_v}=\sigma (H^\infty _v,X)|_{U_v}\). Accordingly, \(\tau _p|_{U_v}=\tau _c|_{U_v}=\sigma (H^\infty _v,X)|_{U_v}\).

We are now ready to prove that \((C_t)''=C_t\). To show this, it suffices to establish that \((C_t)''f=C_tf\) for every \(f\in U_v\).

So, fix \(f\in U_v\). With \((f_n)_{n\in {{\mathbb {N}}}_0}\) as in the First step it follows from there that \(f_n\rightarrow f\) in \((H({{\mathbb {D}}}),\tau _c)\) as \(n\rightarrow \infty \) and, by the Second step, that \((f_n)_{n\in {{\mathbb {N}}}_0}\subseteq U_v\). This implies that \(C_tf_n\rightarrow C_tf\) in \((H({{\mathbb {D}}}),\tau _c)\) as \(n\rightarrow \infty \). Since \(C_t\in {{\mathcal {L}}}(H^\infty _v)\) and \(f\in U_v\), it is clear that \(C_tf\in H^\infty _v\). On the other hand, by the Fourth step the sequence \((f_n)_{n\in {{\mathbb {N}}}_0}\) also converges to f in \((H^\infty _v,\sigma (H^\infty _v,X))=(H^\infty _v,\sigma (H^\infty _v, (H^0_v)'))\). Since \((C_t)'':((H^0_v)'',\sigma ((H^0_v)'',(H^0_v)'))\rightarrow ((H^0_v)'',\sigma ((H^0_v)'',(H^0_v)')\) is continuous, [20, §8.6], that is, \((C_t)'':(H^\infty _v,\sigma (H^\infty _v, X))\rightarrow (H^\infty _v,\sigma (H^\infty _v, X))\) is continuous, it follows that \((C_t)''f_n\rightarrow (C_t)''f\) in \((H^\infty _v,\sigma (H^\infty _v, X))\) as \(n\rightarrow \infty \). Now, \((f_n)_{n\in {{\mathbb {N}}}_0}\subset H^\infty \subseteq H^0_v\), as each \(f_n\) is a polynomial, and \((C_t)''f_n=C_tf_n\) for every \(n\in {{\mathbb {N}}}_0\). Moreover, the sequence \(C_tf_n\rightarrow (C_t)''f\) in \((H({{\mathbb {D}}}),\tau _p)\) as \(n\rightarrow \infty \). Thus, \((C_t)''f=C_tf\) as desired. \(\square \)

Proposition 2.7

Let v be a weight function satisfying \(\lim _{r\rightarrow 1^-}v(r)=0\). For each \(t\in [0,1)\), both of the operators \(C_t:H^\infty _v\rightarrow H^\infty _v\) and \(C_t\rightarrow H^0_v\rightarrow H^0_v\) are compact.

Proof

Fix \(t\in [0,1)\). Since \(H^0_v\) is a closed subspace of \(H_v^\infty \) and \(C_t(H^0_v)\subseteq H_v^0\) (cf. Corollary 2.5), it suffices to show that \(C_t:H^\infty _v\rightarrow H_v^\infty \) is compact. First we establish the following Claim:

  1. (*)

    Let the sequence \((f_n)_{n\in {{\mathbb {N}}}}\subset H^\infty _v\) satisfy \(\Vert f_n\Vert _{\infty ,v}\le 1\) for every \(n\in {{\mathbb {N}}}\) and \(f_n\rightarrow 0\) in \((H({{\mathbb {D}}}),\tau _{c})\) for \(n\rightarrow \infty \). Then \(C_tf_n\rightarrow 0\) in \(H^\infty _v\).

To prove the Claim, let \((f_n)_{n\in {{\mathbb {N}}}}\subset H^\infty _v\) be a sequence as in (*). Fix \(\varepsilon >0\) and select \(\delta \in (0,\beta )\), where \(\beta :=\min \{1,\frac{\varepsilon (1-t)}{2},\frac{\varepsilon (1-t)}{2v(0)}\}\). Since \(\{\xi \in {{\mathbb {C}}}\ \ |\xi |\le (1-\delta )\}\) is a compact subset of \({{\mathbb {D}}}\), there exists \(n_0\in {{\mathbb {N}}}\) such that

$$\begin{aligned} \max _{|\xi |\le 1-\delta }|f_n(\xi )|<\delta , \quad n\ge n_0. \end{aligned}$$

Recall that \(C_tf_n(0)=f_n(0)\) for every \(n\in {{\mathbb {N}}}\). For \(z\in {{\mathbb {D}}}{\setminus }\{0\}\) we have seen previously that

$$\begin{aligned} v(z)|C_tf_n(z)|=v(z)\left| \int _0^1\frac{f_n(sz)}{1-stz}ds\right| \le v(z)\int _0^{1-\delta }\frac{|f_n(sz)|}{|1-stz|}ds+v(z)\int _{1-\delta }^1\frac{|f_n(sz)|}{|1-stz|}ds. \end{aligned}$$

Denote the first (resp., second) summand in the right-side of the previous inequality by \((A_n)\) (resp., by \((B_n)\)). Using the facts that \(|1-stz|\ge 1-st |z|\ge \max \{1-s,1-t,1-|z|\}\), for all \(s,t\in [0,1)\) and \(z\in {{\mathbb {D}}}\), and that v is non-increasing on [0, 1) it follows, for every \(n\ge n_0\), that \(\int _0^{1-\delta }|f_n(sz)|\,ds\le (1-\delta )\max _{|\xi |\le (1-\delta )}|f_n(\xi )|\) (as \(|sz|\le (1-\delta )\) for all \(s\in [0,1-\delta ]\)) and hence, that

$$\begin{aligned} (A_n)\le \frac{v(0)(1-\delta )}{1-t}\max _{|\xi |\le 1-\delta }|f_n(\xi )|<\frac{\varepsilon }{2}. \end{aligned}$$

On the other hand, for every \(n\ge n_0\), we have (as \(\Vert f_n\Vert _{\infty ,v}=\sup _{\xi \in {{\mathbb {D}}}}v(\xi )|f_n(\xi )|\le 1\)) that

$$\begin{aligned} (B_n)=\int _{1-\delta }^1 \frac{v(z)}{v(sz)}\frac{v(sz)|f_n(sz)|}{|1-stz|}\,ds\le \int _{1-\delta }^1\frac{\Vert f_n\Vert _{\infty ,v}}{1-t}\,ds\le \frac{\delta }{1-t}<\frac{\varepsilon }{2}. \end{aligned}$$

It follows that \(\Vert C_tf_n\Vert _{\infty ,v}<\varepsilon \) for every \(n\ge n_0\). That is, \(C_tf_n\rightarrow 0\) in \(H^\infty _v\) for \(n\rightarrow \infty \) and so (*) is proved.

The compactness of \(C_t\in {{\mathcal {L}}}(H^\infty _v)\) can be deduced from (*) as follows. Let \((f_n)_{n\in {{\mathbb {N}}}}\subset H^\infty _v\) be any bounded sequence. There is no loss of generality in assuming that \(\Vert f_n\Vert _{\infty ,v}\le 1\) for all \(n\in {{\mathbb {N}}}\). To establish the compactness of \(C_t\in {{\mathcal {L}}}(H^\infty _v)\) we need to show that \((C_tf_n)_{n\in {{\mathbb {N}}}}\) has a convergent subsequence in \(H^\infty _v\).

Since \(H^\infty _v\subseteq H({{\mathbb {D}}})\) continuously, the sequence \((f_n)_{n\in {{\mathbb {N}}}}\) is also bounded in the Fréchet–Montel space \(H({{\mathbb {D}}})\). Hence, there is a subsequence \(g_j:=f_{n_j}\), for \(j\in {{\mathbb {N}}}\), of \((f_n)_{n\in {{\mathbb {N}}}}\) and \(f\in H({{\mathbb {D}}})\) such that \(g_j\rightarrow f\) in \(H({{\mathbb {D}}})\) with respect to \(\tau _c\). In particular, \(g_j\rightarrow f\) pointwise on \({{\mathbb {D}}}\). Since \(v(z)|g_j(z)|=v(z)|f_{n_j}(z)|\le 1\) for all \(z\in {{\mathbb {D}}}\) and \(j\in {{\mathbb {N}}}\), letting \(j\rightarrow \infty \) it follows that \(v(z)|f(z)|\le 1\) for all \(z\in {{\mathbb {D}}}\), that is, \(f\in H^\infty _v\) with \(\Vert f\Vert _{\infty ,v}\le 1\). Let \(h_j:=\frac{1}{2}(g_j-f)\), for \(j\in {{\mathbb {N}}}\). Then \(\Vert h_j\Vert _{\infty , v}\le 1\), for \(j\in {{\mathbb {N}}}\), and \(h_j\rightarrow 0\) in \(H({{\mathbb {D}}})\) with respect to \(\tau _c\). Condition (*) implies that \(C_th_j\rightarrow 0\) in \(H^\infty _v\) from which it follows that \(C_tf_{n_j}=C_tg_j=C_t(g_j-f)+C_tf=2C_th_j+C_tf\rightarrow C_tf \) in \(H^\infty _v\), as desired. \(\square \)

Proposition 2.8

Let v be a weight function on [0, 1) satisfying \(\lim _{r\rightarrow 1^-}v(r)=0\). For each \(t\in [0,1)\) the spectra of \(C_t\in {{\mathcal {L}}}(H^\infty _v)\) and of \(C_t\in {{\mathcal {L}}}(H^0_v)\) are given by

$$\begin{aligned} \sigma _{pt}(C_t;H^\infty _v)=\sigma _{pt}(C_t;H^0_v)=\left\{ \frac{1}{m+1}\,: m\in {{\mathbb {N}}}_0\right\} ,\end{aligned}$$
(2.1)

and

$$\begin{aligned} \sigma (C_t;H^\infty _v)=\sigma (C_t;H^0_v)=\left\{ \frac{1}{m+1}\,: m\in {{\mathbb {N}}}_0\right\} \cup \{0\}.\end{aligned}$$
(2.2)

Proof

Let \(t\in [0,1)\) be fixed. By [13, Lemma 3.6] we know that the point spectrum of the operator \(C_t^\omega \in {{\mathcal {L}}}(\omega )\) is given by \(\sigma _{pt}(C_t^\omega ;\omega )=\{\frac{1}{m+1}:\, m\in {{\mathbb {N}}}_0\}\) and, for each \(m\in {{\mathbb {N}}}_0\), that the corresponding eigenspace \(\textrm{Ker}(\frac{1}{m+1}I-C_t^\omega )\) is 1-dimensional and is generated by an eigenvector \(x^{[m]}=(x_n^{[m]})_{n\in {{\mathbb {N}}}_0}\in \ell ^1\). Since \(H^0_v\subseteq H^\infty _v\subseteq H({{\mathbb {D}}})\) with continuous inclusions and \(\Phi :H({{\mathbb {D}}})\rightarrow \omega \) (cf. Sect. 1) is a continuous embedding, this implies that \(\sigma _{pt}(C_t;H^0_v)\subseteq \sigma _{pt}(C_t;H^\infty _v)\subseteq \{\frac{1}{m+1}\,:\, m\in {{\mathbb {N}}}_0\}\). Indeed, let \(f\in H({{\mathbb {D}}}){\setminus }\{0\}\) and \(\lambda \in {{\mathbb {C}}}\) satisfy \(C_tf=\lambda f\). Then \(\lambda f(z)=\sum _{n=0}^\infty \widehat{(\lambda f)}(n)z^n=\sum _{n=0}^\infty \lambda {\hat{f}}(n)z^n\) and, by (1.6), we have that \((C_tf)(z)=\sum _{n=0}^\infty (C_t^\omega {\hat{f}})_nz^n\). It follows that \(C_t^\omega {\hat{f}}=\lambda {\hat{f}}\) in \(\omega \) with \({\hat{f}}\not =0\) and so \(\lambda \in \sigma _{pt}(C_t^\omega ;\omega )=\{\frac{1}{m+1}\,\ m\in {{\mathbb {N}}}_0\}\).

To conclude the proof, it remains to show that \(\{\frac{1}{m+1}\,:\, m\in {{\mathbb {N}}}_0\}\subseteq \sigma _{pt}(C_t;H^0_v)\). To establish this recall, for each \(m\in {{\mathbb {N}}}_0\), that the eigenvector \(x^{[m]}\in \ell ^1\) and hence, the function \(g_m(z):=\sum _{n=0}^\infty (x^{[m]})_nz^n\) belongs to \(H^0_v\) because \(0\le v(z)|g_m(z)|\le v(z)\Vert x^{[m]}\Vert _{\ell ^1}\) for \(z\in {{\mathbb {D}}}\) and \(\lim _{r\rightarrow 1^-}v(r)=0\). Moreover, according to (1.5) and (1.6) we have, for each \(z\in {{\mathbb {D}}}\), that

$$\begin{aligned} C_tg_m(z)=\sum _{n=0}^{\infty }(C_t^\omega x^{[m]})_nz^n=\sum _{n=0}^{\infty }(\frac{1}{m+1} x^{[m]})_nz^n=\frac{1}{m+1}\sum _{n=0}^\infty (x^{[m]})_nz^n =\frac{1}{m+1}g_m(z). \end{aligned}$$

Thus \(g_m\) is an eigenvector of \(C_t\in {{\mathcal {L}}}(H^0_v)\) corresponding to the eigenvalue \(\frac{1}{m+1}\).

The validity of \(\sigma (C_t;H^0_v)=\sigma (C_t;H^\infty _v)=\{\frac{1}{m+1}\,:\, m\in {{\mathbb {N}}}_0\}\cup \{0\}\) follows from the fact that \(C_t\) is a compact operator on both spaces. \(\square \)

We now investigate the norm of \(C_t\) on \(H^\infty _v\) for the standard weights \(v_\gamma (z):=(1-|z|)^\gamma \), for \(\gamma >0\) and \(z\in {{\mathbb {D}}}\), which satisfy \(\lim _{r\rightarrow 1^-}v_\gamma (r)=0\).

Proposition 2.9

Let \(t\in (0,1)\) and \(\gamma >0\).

  1. (i)

    The operator norm \(\Vert C_t\Vert _{H^\infty _{v_\gamma }\rightarrow H^\infty _{v_\gamma }}=1\), for every \(\gamma \ge 1\).

  2. (ii)

    For each \(\gamma \in (0,1)\), the inequality \(\Vert C_t\Vert _{H^\infty _{v_\gamma }\rightarrow H^\infty _{v_\gamma }}\le \min \{-\frac{\log (1-t)}{t},\frac{1}{\gamma }\}\) is valid.

Proof

We adapt the arguments given for the Cesàro operator \(C_1\) in the proof of [2, Theorem 2.3].

Let \(\gamma >0\) and \(t\in (0,1)\) be fixed. For \(f\in H^\infty _{v_\gamma }\) with \(\Vert f\Vert _{\infty ,v_\gamma }=1\) we have

$$\begin{aligned} |C_tf(z)|&=\frac{1}{|z|}\left| \int _0^1\frac{f(sz)}{1-stz}ds\right| \le \int _0^1\frac{|f(sz)|}{1-st|z|}ds\\&\le \int _0^1\frac{|f(sz)|}{1-s|z|}ds\le \int _0^1\frac{ds}{(1-s|z|)^{\gamma +1}}=\frac{1}{(1-|z|)^\gamma }\frac{1-(1-|z|)^\gamma }{\gamma |z|}, \end{aligned}$$

as \(z\in {{\mathbb {D}}}\) implies that \(1-st|z|\ge 1-s|z|\), for \(s\in (0,1)\). Accordingly,

$$\begin{aligned} v_\gamma (z)|C_tf(z)|=(1-|z|)^\gamma |C_tf(z)|\le \frac{1-(1-|z|)^\gamma }{\gamma |z|},\quad z\not =0, \end{aligned}$$

and hence,

$$\begin{aligned} \Vert C_tf\Vert _{\infty ,v_\gamma }\le \frac{1}{\gamma } \sup _{z\in {{\mathbb {D}}}}\frac{1-(1-|z|)^\gamma }{|z|}. \end{aligned}$$

Define \(\phi (s):=\frac{1-(1-s)^\gamma }{s}\) for \(s\in (0,1]\) and \(\phi (0)=\gamma \), in which case \(\phi \) is continuous. So, the previous inequality yields \(\Vert C_tf\Vert _{\infty ,v_\gamma }\le \frac{M_\gamma }{\gamma }\), for all \(\Vert f\Vert _{\infty ,v_\gamma }\le 1\), that is, \(\Vert C_t\Vert _{H^\infty _{v_\gamma }\rightarrow H^\infty _{v_\gamma }}\le \frac{M_\gamma }{\gamma }\), where \(M_\gamma :=\sup _{s\in [0,1]}\phi (s)\). Proposition 2.4 yields that \(1\le \Vert C_t\Vert _{H^\infty _{v_\gamma }\rightarrow H^\infty _{v_\gamma }}\le -\frac{\log (1-t)}{t}\) for \(t\in (0,1)\). On page 101 of [2] it is shown that \(\frac{M_\gamma }{\gamma }\le 1\) whenever \(\gamma \ge 1\) and that \(M_\gamma \le 1\) for all \(\gamma \in (0,1)\). The proof of both parts (i) and (ii) follows immediately. \(\square \)

Remark 2.10

For each \(\gamma >0\) let \(v_\gamma (z)=(1-|z|)^\gamma \), for \(z\in {{\mathbb {D}}}\). Proposition 2.9 implies that \(\sup _{0\le t<1}\Vert C_t\Vert _{H^\infty _{v_\gamma }\rightarrow H^\infty _{v_\gamma }}<\infty \). Moreover, if \(\gamma \ge 1\), then \(\Vert C_t^n\Vert _{H^\infty _{v_\gamma }\rightarrow H^\infty _{v_\gamma }}=1\) for every \(n\in {{\mathbb {N}}}\); see case (i) in the proof of [2, Theorem 2.3] together with the fact that \(1\in \sigma _{pt}(C_t, H^\infty _{v_\gamma })\) by Proposition 2.8.

Let \(n\in {{\mathbb {N}}}\) be fixed. Consider the weight \(v(z)=(\log \frac{e}{1-|z|})^{-n}\), for \(z\in {{\mathbb {D}}}\), which satisfies \(v(0)=1\) and \(\lim _{|z|\rightarrow 1^-}v(z)=0\).

The function \(f(z):=[\log (1-z)]^n\in H({{\mathbb {D}}})\) belongs to \(H^\infty _v\). Indeed, for each \(z\in {{\mathbb {D}}}\), we have that

$$\begin{aligned} |\log (1-z)|=\left| -\sum _{n=1}^\infty \frac{z^n}{n}\right| \le \sum _{n=1}^\infty \frac{|z|^n}{n}=-\log (1-|z|) \end{aligned}$$

and hence, that \(|f(z)|=|\log (1-z)|^n\le (-\log (1-|z|))^n\). Since v is given by \(v(z)=(1-\log (1-|z|))^{-n}\) and \(\lim _{|z|\rightarrow 1^-}\frac{-\log (1-|z|)}{1-\log (1-|z|)}=1\), it follows that \(\Vert f\Vert _{\infty ,v}<\infty \) and so \(f\in H^\infty _v\). On the other hand,

$$\begin{aligned} C_1f(z)=\frac{1}{z}\int _0^z\frac{(\log (1-\xi )^n)}{1-\xi }d\xi =-\frac{1}{(n+1)z}(\log (1-z))^{n+1},\quad z\in {{\mathbb {D}}}. \end{aligned}$$

Accordingly, \(C_1f\not \in H^\infty _v\) since

$$\begin{aligned} \lim _{s\rightarrow s^-}v(s)|(C_1)f(s)|&=\frac{1}{n+1}\lim _{s\rightarrow 1^-}\left| \frac{(\log (1-s))^{n+1}}{s(1-\log (1-s))^n}\right| \\&=\frac{1}{n+1}\lim _{s\rightarrow 1^-}\left| \left( \frac{\log (1-s)}{1-\log (1-s)}\right) ^n\frac{\log (1-s)}{s}\right| =\infty . \end{aligned}$$

This implies that the Cesàro operator \(C_1\) is not well-defined on \(H^\infty _v\), that is, \(C_1(H^\infty _v)\not \subseteq H^\infty _v\). But, by Proposition 2.4 the generalized Cesàro operator \(C_t\in {{\mathcal {L}}}(H^\infty _v)\) for every \(t\in [0,1)\). At this point, the following question arises: Is \(\sup _{t\in [0,1)}\Vert C_t\Vert _{H^\infty _v\rightarrow H^\infty _v}<\infty \) for this particular v? Our next two results show that the answer is negative for certain weights v, which includes \(v(z)=\left( \log \frac{e}{1-|z|}\right) ^{-n}\) for \(z\in {{\mathbb {D}}}\).

Proposition 2.11

Let v be a weight function on [0, 1) such that \(\sup _{t\in [0,1)}\Vert C_t\Vert _{H^\infty _v\rightarrow H^\infty _v}<\infty \). Then \(C_1\in {{\mathcal {L}}}(H^\infty _v)\).

Proof

Proposition 2.1 implies that \(\{C_t:\, t\in [0,1)\}\) is equicontinuous in \({{\mathcal {L}}}(H({{\mathbb {D}}}))\). The claim is that \(\lim _{t\rightarrow 1^-}C_tf(z)=C_1f(z)\), for every \(f\in H({{\mathbb {D}}})\) and \(z\in {{\mathbb {D}}}\).

To prove this claim fix \(f\in H({{\mathbb {D}}})\) and \(z\in {{\mathbb {D}}}{\setminus }\{0\}\). Recall, for \(t\in [0,1)\), that

$$\begin{aligned} C_tf(z)=\frac{1}{z}\int _0^z\frac{f(\xi )}{1-t\xi }d\xi =\int _0^1\frac{f(sz)}{1-stz}ds \end{aligned}$$

and

$$\begin{aligned} C_1f(z)=\frac{1}{z}\int _0^z\frac{f(\xi )}{1-\xi }d\xi =\int _0^1\frac{f(sz)}{1-sz}ds. \end{aligned}$$

Moreover, for each \(z\in {{\mathbb {D}}}{\setminus }\{0\}\), we have (as \(|1-stz|\ge (1-|z|)\)) that

$$\begin{aligned} \left| \frac{f(sz)}{1-st z}\right| \le \frac{|f(sz)|}{1-|z|}\le \frac{1}{1-|z|}\max _{|\xi |\le |z|}|f(\xi )|, \quad s\in [0,1], \end{aligned}$$

and that \(\lim _{t\rightarrow 1^-}\frac{f(sz)}{1-st z}=\frac{f(sz)}{1-s z}\) for every \(s\in [0,1]\). So, we can apply the dominated convergence theorem to conclude that \(\lim _{t\rightarrow 1^-}C_tf(z)=C_1f(z)\) for \(z\in {{\mathbb {D}}}{\setminus }\{0\}\). For \(z=0\) we have \(C_tf(0)=f(0)=C_1f(0)\) for each \(f\in H({{\mathbb {D}}})\) and \(t\in [0,1)\). So, for each \(f\in H({{\mathbb {D}}})\), we can conclude that \(C_tf\rightarrow C_1f\) pointwise on \({{\mathbb {D}}}\) for \(t\rightarrow 1^-\). The claim is thereby established.

We now show that \(C_tf\rightarrow C_1f\) in \(H({{\mathbb {D}}})\) as \(t\rightarrow 1^-\) for every \(f\in H^\infty _v\). The assumption \(\sup _{t\in [0,1)}\Vert C_t\Vert _{H^\infty _v\rightarrow H_v^\infty }<\infty \) implies that there exists \(M>0\) satisfying \(\Vert C_t\Vert _{H^\infty _v\rightarrow H^\infty _v}\le M\) for every \(t\in [0,1)\). Therefore,

$$\begin{aligned} \sup _{z\in {{\mathbb {D}}}}|C_tf(z)|v(z)\le M\Vert f\Vert _{\infty ,v}, \quad f\in H^\infty _v,\ t\in [0,1). \end{aligned}$$
(2.3)

Fix \(f\in H^\infty _v\). Then \(\{C_tf:\, t\in [0,1)\}\) is a bounded set in \(H({{\mathbb {D}}})\). Indeed, given \(r\in (0,1)\) and \(t\in [0,1)\) we have (as \(v(r)\le v(z)\) for all \(|z|\le r\)) that

$$\begin{aligned} q_r(C_tf)=\sup _{|z|\le r}|C_tf(z)|=\max _{|z|=r}|C_tf(z)|\le \frac{M}{v(r)}\Vert f\Vert _{\infty ,v}. \end{aligned}$$

So, the set \(\{C_tf:\, t\in [0,1)\}\) is bounded in the Fréchet–Montel space \(H({{\mathbb {D}}})\) and hence, it is relatively compact in \(H({{\mathbb {D}}})\). Since \(C_tf\rightarrow C_1f\) pointwise on \({{\mathbb {D}}}\) for \(t\rightarrow 1^-\), it follows that \(C_tf\rightarrow C_1f\) with respect to \(\tau _{c}\), that is, in the Fréchet space \(H({{\mathbb {D}}})\), for \(t\rightarrow 1^-\). In particular, \(C_1f\in H({{\mathbb {D}}})\).

Since \(H^\infty _v\subseteq H({{\mathbb {D}}})\) and \(C_th\rightarrow C_1h\) pointwise on \({{\mathbb {D}}}\) as \(t\rightarrow 1^-\), for every \(h\in H({{\mathbb {D}}})\), letting \(t\rightarrow 1^-\) in (2.3) it follows that

$$\begin{aligned} |C_1f(z)|v(z)\le M\Vert f\Vert _{\infty ,v},\quad z\in {{\mathbb {D}}}, \end{aligned}$$

that is, \(\Vert C_1f\Vert _{\infty ,v}\le M\Vert f\Vert _{\infty ,v}\). But, \(f\in H^\infty _v\) is arbitrary and so \(C_1\in {{\mathcal {L}}}(H^\infty _v)\). \(\square \)

Proposition 2.12

For each \(n\in {{\mathbb {N}}}\), let \(v(z)=(\log (\frac{e}{1-|z|}))^{-n}\) for \(z\in {{\mathbb {D}}}\). Then \(\sup _{t\in [0,1)}\Vert C_t\Vert _{H^\infty _v\rightarrow H^\infty _v}=\infty \).

Proof

Apply Proposition 2.11 and the discussion prior it. \(\square \)

3 Linear dynamics and mean ergodicity of \(C_t\)

The aim of this section is to investigate the mean ergodicity and the linear dynamics of the operators \(C_t\), for \(t\in [0,1)\), acting on \(H({{\mathbb {D}}})\), \(H^\infty _v\) and \(H^0_v\)

An operator \(T\in {{\mathcal {L}}}(X)\), with X a lcHs, is called power bounded if \(\{T^n:\ n\in {{\mathbb {N}}}_0\}\) is an equicontinuous subset of \({{\mathcal {L}}}(X)\). For a Banach space X, this means that \(\sup _{n\in {{\mathbb {N}}}_0}\Vert T^n\Vert _{X\rightarrow X}<\infty \). Given \(T\in {{\mathcal {L}}}(X)\), the averages

$$\begin{aligned} T_{[n]}:=\frac{1}{n}\sum _{m=1}^nT^m,\quad n\in {{\mathbb {N}}}, \end{aligned}$$

are usually called the Cesàro means of T. The operator T is said to be mean ergodic (resp., uniformly mean ergodic) if \((T_{[n]})_{n\in {{\mathbb {N}}}}\) is a convergent sequence in \({{\mathcal {L}}}_s(X)\) (resp., in \({{\mathcal {L}}}_b(X)\)). It is routine to check that \(\frac{T^n}{n}=T_{[n]}-\frac{n-1}{n}T_{[n-1]}\), for \(n\ge 2\), and hence, \(\tau _s\)-\(\lim _{n\rightarrow \infty }\frac{T^n}{n}=0\) whenever T is mean ergodic. Every power bounded operator on a Fréchet–Montel space X is necessarily uniformly mean ergodic, [1, Proposition 2.8]. Concerning the linear dynamics of \(T\in {{\mathcal {L}}}(X)\), with X a lcHs, the operator T is called supercyclic if, for some \(z\in X\), the projective orbit \(\{\lambda T^nz\,:\ \lambda \in {{\mathbb {C}}},\ n\in {{\mathbb {N}}}_0\}\) is dense in X. Since the closure of the linear span of a projective orbit is separable, if such a supercyclic operator \(T\in {{\mathcal {L}}}(X)\) exists, then X is necessarily separable.

Observe that the space \(H^\infty _v\) is never separable, [24, Theorem 1.1]. Therefore, every operator \(T\in {{\mathcal {L}}}(H^\infty _v)\) is clearly not supercyclic. However, the spaces \(H({{\mathbb {D}}})\), [21, Theorem 27.2.5], and \(H^0_v\), [24, Theorem 1.1], for every weight v are always separable. Hence, the problem of supercyclicity for non-zero operators \(T\in {{\mathcal {L}}}(H({{\mathbb {D}}}))\) and \(T\in {{\mathcal {L}}}(H^0_v)\) arises.

The following result, [5, Theorem 6.4], is stated here for Banach spaces.

Theorem 3.1

Let X be a Banach space and let \(T\in {{\mathcal {L}}}(X)\) be a compact operator such that \(1\in \sigma (T;X)\) with \(\sigma (T;X){\setminus } \{1\}\subseteq \overline{B(0,\delta )}\) for some \(\delta \in (0,1)\) and satisfying \(\textrm{Ker}(I-T)\cap \textrm{Im} (I-T)=\{0\}\). Then T is power bounded and uniformly mean ergodic.

A consequence of the previous theorem is the following result.

Proposition 3.2

Let v be a weight function on [0, 1) satisfying \(\lim _{r\rightarrow 1^-}v(r)=0\). For each \(t\in [0,1)\) both of the operators \(C_t\in {{\mathcal {L}}}(H^\infty _v)\) and \(C_t\in {{\mathcal {L}}}(H^0_v)\) are power bounded, uniformly mean ergodic and fail to be supercyclic.

Proof

Fix \(t\in [0,1)\). It was already noted that \(C_t\in {{\mathcal {L}}}(H^\infty _v)\) cannot be supercyclic. The operator \(C_t\) is a compact operator on both \(H^\infty _v\) and on \(H^0_v\) (cf. Proposition 2.7). Therefore, the compact transpose operators \(C_t'\in {{\mathcal {L}}}((H^\infty _v)')\) and \(C'_t\in {{\mathcal {L}}}((H^0_v)')\) have the same non-zero eigenvalues as \(C_t\) (see, e.g., [15, Theorem 9.10-2(2)]). In view of Proposition 2.8 it follows that \(\sigma _{pt}(C_t';(H^\infty _v)')=\sigma _{pt}(C_t';(H^0_v)')=\{\frac{1}{m+1}\,:\, m\in {{\mathbb {N}}}_0\}\). We can apply [6, Proposition 1.26] to conclude that \(C_t\) is not supercyclic on \(H^0_v\).

By Proposition 2.8 and its proof (as \(x^{[0]}=(t^n)_{n\in {{\mathbb {N}}}_0}\)) we have that \(\textrm{Ker}(I-C_t)=\textrm{span}\{g_0\}\), with \(g_0(z)=\sum _{n=0}^\infty t^nz^n\), for \(z\in {{\mathbb {D}}}\). On the other hand, \(\textrm{Im}(I-C_t)\) is a closed subspace of \(H^\infty _v\) (resp., of \(H^0_v\)), as \(C_t\) is compact in \(H^\infty _v\) (resp., in \(H^0_v\))), and \(\textrm{Im}(I-C_t)\subseteq \{g\in H^\infty _v\,:\, g(0)=0\}\) (resp., \(\subseteq \{g\in H^0_v\,:\, g(0)=0\}\)), because \(C_tf(0)=f(0)\) for each \(f\in H^\infty _v\) (resp., each \(f\in H^0_v\)). Moreover, [15, Theorem 9.10.1] implies that \(\mathrm{codim\,Im}(I-C_t)=\mathrm{dim\,Ker} (I-C_t)=1\). Accordingly, both \(\textrm{Im}(I-C_t)\) and \(\{g\in H^\infty _v\,:\, g(0)=0\}=\textrm{Ker}(\delta _0)\) are hyperplanes, where \(\delta _0\in (H^\infty _v)'\) is the linear evaluation functional \(f\mapsto f(0)\), for \(f\in H^\infty _v\). It follows that necessarily \(\textrm{Im}(I-C_t)=\{g\in H^\infty _v:\, g(0)=0\}\).

Let \(h\in \textrm{Im}(I-C_t)\cap \textrm{Ker}(I-C_t)\). Then \(h(0)=0\) and there exists \(\lambda \in {{\mathbb {C}}}\) such that \(h=\lambda g_0\). This yields that \(0=h(0)=\lambda g_0(0)=\lambda \). Hence, \(h=0\). So, \(\textrm{Im}(I-C_t)\cap \textrm{Ker}(I-C_t)=\{0\}\).

Proposition 2.8 implies that \(1\in \sigma (C_t; H^\infty _v)=\sigma (C_t; H^0_v)=\{\frac{1}{m+1}\,;\, m\in {{\mathbb {N}}}_0\}\cup \{0\}\). Consequently, for \(\delta =\frac{1}{2}\), all the assumptions of Theorem 3.1 are satisfied. So, we can conclude that \(C_t\) is power bounded and uniformly mean ergodic on both \(H^\infty _v\) and on \(H^0_v\). \(\square \)

In contrast to the compactness of \(C_t\) acting in the Banach spaces \(H^\infty _v\) and \(H^0_v\) (cf. Proposition 2.7) the situation for the Fréchet space \(H({{\mathbb {D}}})\) is different.

Proposition 3.3

For each \(t\in [0,1)\) the operator \(C_t:H({{\mathbb {D}}})\rightarrow H({{\mathbb {D}}})\) is an isomorphism and, hence, it is not compact.

Proof

Fix \(t\in [0,1)\). Consider the operator \(T_t:H({{\mathbb {D}}})\rightarrow H({{\mathbb {D}}})\), for \(f\in H({{\mathbb {D}}})\), given by

$$\begin{aligned} T_tf(z):= (1-tz)(zf(z))'=(1-tz)(f(z)+zf'(z)),\quad z\in {{\mathbb {D}}}. \end{aligned}$$

Then \(T_t\) is clearly well-defined. Moreover, its graph is closed. Indeed, for a given sequence \((f_n)_{n\in {{\mathbb {N}}}}\subset H({{\mathbb {D}}})\), suppose that \(f_n\rightarrow f\) in \(H({{\mathbb {D}}})\) and \(T_tf_n\rightarrow g\) in \(H({{\mathbb {D}}})\). Since multiplication operators (by elements from \(H({{\mathbb {D}}})\)) and the differentiation operator are continuous on \(H({{\mathbb {D}}})\) and the evaluation functionals at points of \({{\mathbb {D}}}\) belong to \(H({{\mathbb {D}}})'\), it follows that \(f'_n\rightarrow f'\) in \(H({{\mathbb {D}}})\) and hence, \(T_tf_n=(1-tz)(f_n+zf'_n)\rightarrow (1-tz)(f+zf')=T_tf\) in \(H({{\mathbb {D}}})\). Accordingly, \(g=T_tf\). Since \(H({{\mathbb {D}}})\) is a Fréchet space, the closed graph theorem, [20, Corollary 5.4.3], implies that \(T_t\in {{\mathcal {L}}}(H({{\mathbb {D}}}))\).

Finally, it is routine to verify that \(C_t\circ T_t=T_t\circ C_t=I\). So, the inverse operator \(C_t^{-1}=T_t\in {{\mathcal {L}}}(H({{\mathbb {D}}}))\) exists and hence, \(C_t\) is a bi-continuous isomorphism of \(H({{\mathbb {D}}})\) onto itself. In particular, \(C_t\) cannot be compact. \(\square \)

Let \(\Lambda :=\{\frac{1}{n+1}:\, n\in {{\mathbb {N}}}_0\}\) and \(\Lambda _0:=\Lambda \cup \{0\}\). We recall from [4, Lemma 2.7] the following lemma, which is an extension of a result of Rhoades [27].

Lemma 3.4

For every \(\mu \in {{\mathbb {C}}}{\setminus } \Lambda _0\) there exist \(\delta =\delta _\mu >0\) and constants \(d_\delta , D_\delta >0\) such that \(\overline{B(\mu ,\delta )}\cap \Lambda _0=\emptyset \) and

$$\begin{aligned} \frac{d_\delta }{n^{\alpha (\nu )}}\le \prod _{k=1}^n\left| 1-\frac{1}{k\nu }\right| \le \frac{D_\delta }{n^{\alpha (\nu )}},\quad \forall n\in {{\mathbb {N}}},\ \nu \in B(\mu ,\delta ), \end{aligned}$$
(3.1)

where \(\alpha (\nu ):=\textrm{Re}(\frac{1}{\nu })\).

Remark 3.5

As a direct application of Lemma 3.4 we obtain, for every \(\mu \in {{\mathbb {C}}}{\setminus } \Lambda _0\), that there exist \(\delta >0\) and \(d_\delta , D_\delta >0\) such that \(\overline{B(\mu , \delta )}\cap \Lambda _0=\emptyset \) and, for every \(\nu \in B(\mu ,\delta )\) and \(n\in {{\mathbb {N}}}_0\), we have that

$$\begin{aligned} d_\delta D_\delta ^{-1}\left( \frac{n-h}{n+1}\right) ^{\alpha (\nu )}\le \prod _{j=n-h+1}^{n+1}\left| 1-\frac{1}{j\nu }\right| \le D_\delta d_\delta ^{-1}\left( \frac{n-h}{n+1}\right) ^{\alpha (\nu )}, \end{aligned}$$
(3.2)

for all \(h\in \{1,\ldots , n-1\}\), where \(\alpha (\nu )=\textrm{Re}(\frac{1}{\nu })\).

For each \(k\in {{\mathbb {N}}}\) with \(k\ge 2\) define \(r_k:=(1-\frac{1}{k})\). Define the norms \(\Vert \cdot \Vert _k\) and \(|||\cdot |||_k\) on \(H({{\mathbb {D}}})\) by

$$\begin{aligned} \Vert f\Vert _k:=\sum _{n=0}^\infty |{\hat{f}}(n)|r_k^n,\quad f=\sum _{n=0}^\infty {\hat{f}}(n)z^n, \end{aligned}$$

and

$$\begin{aligned} |||f|||_k:=\sup _{n\in {{\mathbb {N}}}_0} |{\hat{f}}(n)|r_k^n\quad f=\sum _{n=0}^\infty {\hat{f}}(n)z^n. \end{aligned}$$

Lemma 3.6

Each of the sequences \(\{\Vert \cdot \Vert _k\}_{k\ge 2}\) and \(\{|||\cdot |||_k\}_{k\ge 2}\) is a fundamental system of norms for \((H({{\mathbb {D}}}),\tau _c)\).

Proof

Given \(r\in (0,1)\) choose any \(k\ge 2\) such that \(0<r<(1-\frac{1}{k})\). Then, for every \(f\in H({{\mathbb {D}}})\), we have

$$\begin{aligned} q_r(f)=\sup _{|z|=r}\left| \sum _{n=0}^\infty {\hat{f}}(n)z^n\right| \le \sum _{n=0}^\infty |{\hat{f}}(n)|r^n\le \sum _{n=0}^\infty |{\hat{f}}(n)|\left( 1-\frac{1}{k}\right) ^n=\Vert f\Vert _k. \end{aligned}$$

On the other hand, given \(k\ge 2\), let \(r_k:=(1-\frac{1}{k})<(1-\frac{1}{k+1}):=r_{k+1}\). By the Cauchy inequalities, for \(n\in {{\mathbb {N}}}_0\), we have

$$\begin{aligned} |{\hat{f}}(n)|\le \frac{1}{r_{k+1}^n}\max _{|z|=r_{k+1}}|f(z)|= \frac{1}{r_{k+1}^n} q_{r_{k+1}}(f),\quad f\in H({{\mathbb {D}}}), \end{aligned}$$

and hence,

$$\begin{aligned} \Vert f\Vert _{r_k}=\sum _{n=0}^\infty |{\hat{f}}(n)|r_k^n\le q_{r_{k+1}}(f)\sum _{n=0}^\infty \left( \frac{r_k}{r_{k+1}}\right) ^n= cq_{r_{k+1}}(f),\quad f\in H({{\mathbb {D}}}), \end{aligned}$$

with \(c=\frac{1}{1-\frac{r_k}{r_{k+1}}}=k^2>0\) as \(\frac{r_k}{r_{k+1}}<1\), which is independent of f.

So, the systems \(\{q_r\}_{r\in (0,1)}\) and \(\{\Vert \cdot \Vert _k\}_{k\ge 2}\) are equivalent on \(H({{\mathbb {D}}})\).

Observe, for every \(k\ge 2\), that

$$\begin{aligned} |||f|||_k=\sup _{n\in {{\mathbb {N}}}_0}|{\hat{f}}(n)|r^n_k\le \sum _{n=0}^\infty |{\hat{f}}(n)|r_k^n=\Vert f\Vert _k,\quad f\in H({{\mathbb {D}}}), \end{aligned}$$

and that

$$\begin{aligned} \Vert f\Vert _k&=\sum _{n=0}^\infty |{\hat{f}}(n)|r_k^n=\sum _{n=0}^\infty |{\hat{f}}(n)|\left( \frac{r_k}{r_{k+1}}\right) ^nr^n_{k+1}\\&\le \sup _{n\in {{\mathbb {N}}}_0}|{\hat{f}}(n)|r^n_{k+1}\sum _{n=0}^\infty \left( \frac{r_k}{r_{k+1}}\right) ^n= k^2|||f|||_{k+1}, \end{aligned}$$

for \(f\in H({{\mathbb {D}}})\), where \(\sum _{n=0}^\infty \left( \frac{r_k}{r_{k+1}}\right) ^n=k^2\). Therefore, the systems \(\{\Vert \cdot \Vert _k\}_{k\ge 2}\) and \(\{|||\cdot |||_k\}_{k\ge 2}\) are equivalent. \(\square \)

Proposition 3.7

For each \(t\in [0,1)\) the spectra of the operator \(C_t\in {{\mathcal {L}}}(H({{\mathbb {D}}}))\) are given by

$$\begin{aligned} \sigma _{pt}(C_t;H({{\mathbb {D}}}))=\sigma (C_t;H({{\mathbb {D}}}))=\Lambda \end{aligned}$$
(3.3)

and

$$\begin{aligned} \sigma ^*(C_t;H({{\mathbb {D}}}))=\Lambda _0. \end{aligned}$$
(3.4)

Proof

Let \(t\in [0,1)\) be fixed. For any weight function v on [0, 1) satisfying \(\lim _{r\rightarrow 1^-}v(r)=0\), we have \(H^\infty _v\subseteq H({{\mathbb {D}}})\) continuously and \(\Phi :H({{\mathbb {D}}})\rightarrow \omega \) is a continuous imbedding. Accordingly, \(\sigma _{pt}(C_t;H_v^\infty )\subseteq \sigma _{pt}(C_t;H({{\mathbb {D}}}))\subseteq \Lambda \); see the proof of Proposition 2.8. Since \(\sigma _{pt}(C_t;H^\infty _v)=\Lambda \) (cf. Proposition 2.8) and \(\sigma _{pt}(C_t^\omega ;\omega )=\Lambda \) [5, Theorem 3.7], it follows that \(\sigma _{pt}(C_t;H({{\mathbb {D}}}))=\Lambda \). Moreover, in view of Proposition 2.8 above and Theorem 3.7 in [5], the eigenspace corresponding to each eigenvalue \(\frac{1}{n+1}\in \Lambda \) is 1-dimensional. By Proposition 3.3, the operator \(C_t:H({{\mathbb {D}}})\rightarrow H({{\mathbb {D}}})\) is a bi-continuous isomorphism and so \(0\not \in \sigma (C_t; H({{\mathbb {D}}}))\).

The claim is that \({{\mathbb {C}}}{\setminus } \Lambda _0\subseteq \rho (C_t;H({{\mathbb {D}}}))\). To establish this claim, fix \(\nu \in {{\mathbb {C}}}{\setminus }\Lambda _0\). Given \(g(z)=\sum _{n=0}^\infty c_nz^n\in H({{\mathbb {D}}})\), consider the identity

$$\begin{aligned} (C_t-\nu I)f(z)=g(z),\quad z\in {{\mathbb {D}}}, \end{aligned}$$
(3.5)

where \(f(z)=\sum _{n=0}^\infty a_n z^n\in H({{\mathbb {D}}})\) is to be determined. It follows from (1.6) that \(C_tf(z)=\sum _{n=0}^\infty (\frac{t^na_0+t^{n-1}a_1+\cdots +a_n}{n+1})z^n\) from which the identity \((C_t-\nu I)f(z)=\sum _{n=0}^\infty (\frac{t^na_0+t^{n-1}a_1+\cdots +a_n}{n+1}-\nu a_n)z^n\) is clear. So, (3.5) is satisfied if and only if

$$\begin{aligned} \sum _{n=0}^\infty \left( \frac{t^na_0+t^{n-1}a_1+\cdots +a_n}{n+1}-\nu a_n \right) z^n=\sum _{n=0}^\infty c_nz^n, \quad z\in {{\mathbb {D}}}, \end{aligned}$$

that is, if and only if

$$\begin{aligned} \frac{t^na_0+t^{n-1}a_1+\cdots +a_n}{n+1}-\nu a_n=c_n,\quad n\in {{\mathbb {N}}}_0. \end{aligned}$$

In view of this we can argue, as in the proof of [5, Lemma 3.6], to show that if a function \(f\in H({{\mathbb {D}}})\) exists which satisfies the identity (3.5), then the Taylor coefficients \((a_n)_{n\in {{\mathbb {N}}}_0}\) of f must verify the following equalities

$$\begin{aligned} a_0&=\frac{c_0}{1-\nu }\nonumber \\ a_n&=\frac{c_n}{(\frac{1}{n+1}-\nu )}+\sum _{h=1}^n(-1)^h\frac{\nu ^{h-1}t^hc_{n-h}}{(n+1) \prod _{j=n-h+1}^{n+1}(\frac{1}{j}-\nu )}\nonumber \\&=: A_n+B_n, \quad n\ge 1. \end{aligned}$$
(3.6)

Observe, for each \(n\ge 1\) and \(h\in \{1,\ldots ,n\}\), that

$$\begin{aligned} (-1)^h\prod _{j=n-h+1}^{n+1}\left( \frac{1}{j}-\nu \right) =-\prod _{j=n-h+1}^{n+1}\left( \nu -\frac{1}{j}\right) =-\nu ^{h+1}\prod _{j=n-h+1}^{n+1}\left( 1-\frac{1}{j\nu }\right) \end{aligned}$$

and so

$$\begin{aligned} B_n=-\sum _{h=1}^n\frac{\nu ^{h-1}t^hc_{n-h}}{\nu ^{h+1}(n+1) \prod _{j=n-h+1}^{n+1}(1-\frac{1}{j\nu })}=-\frac{1}{\nu ^2} \sum _{h=1}^n\frac{t^hc_{n-h}}{(n+1)\prod _{j=n-h+1}^{n+1}(1-\frac{1}{j\nu })}. \end{aligned}$$

Accordingly, to verify the claim we need to prove that the power series \(\sum _{n=0}^\infty a_nz^n\) is convergent in \({{\mathbb {D}}}\), with \((a_n)_{n\in {{\mathbb {N}}}_0}\) defined according to (3.6). First, observe that the series \(g(z)=\sum _{n=0}^\infty c_nz^n\) is convergent in \({{\mathbb {D}}}\) and satisfies

$$\begin{aligned} \limsup _{n\rightarrow \infty }\root n \of {|c_n|}=\limsup _{n\rightarrow \infty } \root n \of {\frac{|c_n|}{|\frac{1}{n+1}-\nu |}}=\limsup _{n\rightarrow \infty }\root n \of {|A_n|}. \end{aligned}$$

Therefore, the series \(\sum _{n=1}^\infty A_nz^n\) has the same radius of convergence as the series \(\sum _{n=0}^\infty c_nz^n\) and hence, it converges in \(H({{\mathbb {D}}})\). Accordingly, \(f_1(z):=\sum _{n=1}^\infty A_nz^n\), for \(z\in {{\mathbb {D}}}\), belongs to \(H({{\mathbb {D}}})\). On the other hand, the series

$$\begin{aligned} \sum _{n=1}^\infty B_nz^n&=-\frac{1}{\nu ^2}\sum _{n=1}^\infty \sum _{h=1}^n\frac{t^h c_{n-h}}{(n+1)\prod _{j=n-h+1}^{n+1}(1-\frac{1}{j\nu })}\\&=-\frac{1}{\nu ^2}\sum _{h=1}^\infty t^hz^h\sum _{n=h}^\infty \frac{c_{n-h}z^{n-h}}{(n+1)\prod _{j=n-h+1}^{n+1}(1-\frac{1}{j\nu })},\quad z\in {{\mathbb {D}}}. \end{aligned}$$

To establish the convergence of the series \(\sum _{n=1}^\infty B_nz^n\) in \(H({{\mathbb {D}}})\), fix \(z\in {{\mathbb {D}}}{\setminus }\{0\}\) and \(r\in (|z|,1)\). Recall, for every \(n\in {{\mathbb {N}}}_0\), that the Taylor coefficients of g satisfy (as \(\frac{1}{r}>1\))

$$\begin{aligned} |c_n|=\left| \frac{g^{(n)}(0)}{n!}\right| =\left| \frac{1}{2\pi i}\int _{|\xi |=r}\frac{g(\xi )}{\xi ^{n+1}}\,d\xi \right| \le \frac{1}{r^{n}}\max _{|\xi |=r}|g(\xi )|\le \frac{C}{r^{n+1}} \end{aligned}$$

where \(C:=\max _{|\xi |=r}|g(\xi )|\). Therefore, setting \(\alpha :=\alpha (\nu )=\textrm{Re}(\frac{1}{\nu })\) and \(d:=d_\delta \) and \(D:=D_\delta \) for a suitable \(\delta >0\) (cf. Remark 3.5), we obtain via (3.1) and (3.2) that

$$\begin{aligned}&\sum _{h=1}^\infty t^h|z|^h\sum _{n=h}^\infty \frac{|c_{n-h}|\,|z|^{n-h}}{(n+1)\prod _{j=n-h+1}^{n+1}|1-\frac{1}{j\nu }|}\\&\quad \le C\sum _{h=1}^\infty t^h|z|^{h-1}\left( \frac{|z|}{r}d^{-1}(h+1)^{-\alpha -1}+\sum _{n=h+1}^\infty \left( \frac{|z|}{r}\right) ^{n-h+1}Dd^{-1}\left( \frac{n+1}{n-h}\right) ^\alpha \right) \\&\quad =Cd^{-1}\sum _{h=1}^\infty t^h(h+1)^{-\alpha -1}|z|^{h}+ CDd^{-1}\sum _{h=1}^\infty t^h|z|^{h-1}\sum _{n=h+1}^\infty \left( \frac{|z|}{r}\right) ^{n-h+1}\left( \frac{n+1}{n-h}\right) ^\alpha \\&\quad \le Cd^{-1}\left( \sum _{h=1}^\infty t^h(h+1)^{-\alpha -1}|z|^{h}+D\sum _{h=1}^\infty t^h |z|^{h-1}\max \{1, (2+h)^\alpha \}\sum _{n=h+1}^\infty \left( \frac{|z|}{r}\right) ^{n-h+1}\right) , \end{aligned}$$

which is finite after observing that if \(\alpha \le 0\), then \(\left( \frac{n+1}{n-h}\right) ^\alpha =\left( \frac{n-h}{n+1}\right) ^{-\alpha }\le 1\) for every \(h\in {{\mathbb {N}}}\) and every \(n\ge h+1\), whereas if \(\alpha >0\), then \((\frac{n+1}{n-h})^\alpha =(1+\frac{h+1}{n-h})^\alpha \le (2+h)^\alpha \). This implies that the series \(\sum _{n=1}^\infty B^nz^n\) converges in \(H({{\mathbb {D}}})\). Accordingly, \(f_2(z):=\sum _{n=1}^\infty B_nz^n\), for \(z\in {{\mathbb {D}}}\), belongs to \(H({{\mathbb {D}}})\).

Set \(f(z):=\frac{c_0}{1-\nu }+f_1(z)+f_2(z)\), for \(z\in {{\mathbb {D}}}\). Then \(f\in H({{\mathbb {D}}})\). Moreover, the arguments above imply that f satisfies (3.5). The identities (3.6) imply that f is the unique solution of (3.5). Accordingly, the inverse operator \((C_t-\nu I)^{-1}:H({{\mathbb {D}}})\rightarrow H({{\mathbb {D}}})\) exists. In particular, \((C_t-\nu I)^{-1}\in {{\mathcal {L}}}(H({{\mathbb {D}}}))\) as it is the inverse of a continuous linear operator on a Fréchet space.

Since \(\nu \in {{\mathbb {C}}}{\setminus }\Lambda _0\) is arbitrary and \(0\in \rho (C_t;H({{\mathbb {D}}}))\), we can conclude that \(\sigma (C_t;H({{\mathbb {D}}}))=\Lambda \).

It remains to show that \(\sigma ^*(C_t;H({{\mathbb {D}}}))=\Lambda _0\). To establish this, fix \(\mu \in {{\mathbb {C}}}{\setminus }\Lambda _0\) and observe, by Lemma 3.4, that there exist \(\delta >0\) and constants \(d_\delta , D_\delta >0\) such that \(\overline{B(\mu ,\delta )}\cap \Lambda _0=\emptyset \) and the inequalities (3.1) and (3.2) are satisfied. We will show that \(B(\mu ,\delta )\subset \rho (C_t;H({{\mathbb {D}}}))\) and that the set \(\{(C_t-\nu I)^{-1}:\, \nu \in B(\mu ,\delta )\}\) is equicontinuous in \({{\mathcal {L}}}(H({{\mathbb {D}}}))\). To see this, first observe that the function \(\nu \in \overline{B(\mu ,\delta )} \mapsto \textrm{Re}(\frac{1}{\nu })\in {{\mathbb {R}}}\) is continuous and hence, \(\alpha _0:=\max _{\nu \in \overline{B(\mu ,\delta )} }\{\textrm{Re}(\frac{1}{\nu })\}\) exists. For the sake of simplicity of notation set \(d:=d_\delta \) and \(D:=D_\delta \).

Let \(\nu \in B(\mu ,r)\), where \(r:=\frac{1}{2}d(\Lambda _0, \overline{B(\mu ,\delta )})>0\) has the property that \(|\nu -\frac{1}{j}|>r\) for all \(j\in {{\mathbb {N}}}\). It was proved above, for any fixed \(g(z)=\sum _{n=0}^\infty c_nz^n\in H({{\mathbb {D}}})\), that

$$\begin{aligned} (C_t-\nu I)^{-1}g(z)=\frac{c_0}{1-\nu }+\sum _{n=1}^\infty \left( \frac{c_n}{\frac{1}{n+1}-\nu }-\frac{1}{\nu ^2} \sum _{h=1}^n\frac{(-1)^ht^hc_{n-h}}{(n+1)\prod _{j=n-h+1}^{n+1}(1-\frac{1}{j\nu })}\right) z^n, \end{aligned}$$

for each \(z\in {{\mathbb {D}}}\). So, for \(k\ge 2\) fixed, consider the norm \(\Vert \cdot \Vert _k\) in \(H({{\mathbb {D}}})\). Then we have, via (3.6), that

$$\begin{aligned}&\Vert (C_t-\nu I)^{-1}g\Vert _k\\&\quad \le \frac{|c_0|}{|1-\nu |}+\sum _{n=1}^\infty \left| \frac{c_n}{\frac{1}{n+1}-\nu }- \frac{1}{\nu ^2}\sum _{h=1}^n\frac{(-1)^ht^hc_{n-h}}{(n+1) \prod _{j=n-h+1}^{n+1}(1-\frac{1}{j\nu })}\right| \left( 1-\frac{1}{k}\right) ^n\\&\quad \le \left( \frac{1}{r}\sum _{n=0}^\infty |c_n|\left( 1-\frac{1}{k}\right) ^n\right) +\frac{1}{|\nu |^2}\sum _{n=1}^\infty \sum _{h=1}^n\frac{t^h|c_{n-h}|}{(n+1) \prod _{j=n-h+1}^{n+1}|1-\frac{1}{j\nu }|}\left( 1-\frac{1}{k}\right) ^n\\&\quad =\frac{1}{r}\Vert g\Vert _k+\frac{1}{|\nu |^2}\sum _{h=1}^\infty t^h\left( 1-\frac{1}{k}\right) ^h\sum _{n=h}^\infty \frac{|c_{n-h}|}{(n+1) \prod _{j=n-h+1}^{n+1}|1-\frac{1}{j\nu }|}\left( 1-\frac{1}{k}\right) ^{n-h}. \end{aligned}$$

Moreover, (3.1) and (3.2) with \(\alpha (\nu )=\textrm{Re}(\frac{1}{\nu })\le \alpha _0\) imply, for each \(h\in {{\mathbb {N}}}\), that

$$\begin{aligned}&\sum _{n=h}^\infty \frac{|c_{n-h}|}{(n+1)\prod _{j=n-h+1}^{n+1}|1-\frac{1}{j\nu }|}\left( 1-\frac{1}{k}\right) ^{n-h}=\sum _{l=0}^\infty \frac{|c_{l}|}{(l+h+1)\prod _{j=l+1}^{l+h+1}|1-\frac{1}{j\nu }|}\left( 1-\frac{1}{k}\right) ^{l}\\&\quad = \frac{|c_0|}{(h+1)\prod _{j=1}^{h+1}|1-\frac{1}{j\nu }|}+\sum _{l=1}^\infty \frac{|c_{l}|}{(l+h+1)\prod _{j=l+1}^{l+h+1}|1-\frac{1}{j\nu }|}\left( 1-\frac{1}{k}\right) ^{l}\\&\quad \le d^{-1}|c_0|(h+1)^{\alpha (\nu ) -1}+d^{-1}D \sum _{l=1}^\infty \frac{|c_l|}{l+h+1}\left( \frac{l+h+1}{l}\right) ^{\alpha (\nu )}\left( 1-\frac{1}{k}\right) ^l\\&\quad \le d^{-1}|c_0|(h+1)^{\alpha _0 -1}+d^{-1}D \sum _{l=1}^\infty \frac{|c_l|}{l+h+1}\left( \frac{l+h+1}{l}\right) ^{\alpha _0}\left( 1-\frac{1}{k}\right) ^l\\&\quad \le \max \{d^{-1},d^{-1}D\}(2+h)^{\alpha _0}\sum _{l=0}^\infty |c_l|\left( 1-\frac{1}{k}\right) ^l=K(2+h)^{\alpha _0}\Vert g\Vert _k, \end{aligned}$$

with \(K:=\max \{d^{-1},d^{-1}D\}\), and hence, since \(|\nu |>r\) for all \(\nu \in B(\mu ,\delta )\), that

$$\begin{aligned}&\frac{1}{|\nu |^2}\sum _{h=1}^\infty t^h\left( 1-\frac{1}{k}\right) ^h\sum _{n=h}^\infty \frac{|c_{n-h}|}{(n+1) \prod _{j=n-h+1}^{n+1}|1-\frac{1}{j\nu }|}\left( 1-\frac{1}{k}\right) ^{n-h}\\&\quad \le \frac{K}{r^2}\Vert g\Vert _k\sum _{h=1}^\infty t^h\left( 1-\frac{1}{k}\right) ^h(2+h)^{\alpha _0}=K'\Vert g\Vert _k, \end{aligned}$$

with \(K'=\frac{K}{r^2}\sum _{h=1}^\infty t^h\left( 1-\frac{1}{k}\right) ^h(2+h)^{\alpha _0}<\infty \), by the ratio test, for instance.

We have established, for every \(\nu \in B(\mu ,\delta )\), that

$$\begin{aligned} \Vert (C_t-\nu I)^{-1}g\Vert _k\le (\frac{1}{r}+K')\Vert g\Vert _k. \end{aligned}$$

Since \(g\in H({{\mathbb {D}}})\) and \(k\ge 2\) are arbitrary, this shows that the set \(\{(C_t-\nu I)^{-1}:\, \nu \in B(\mu ,\delta )\}\) is equicontinuous. Hence, \(\sigma ^*(C_t;H({{\mathbb {D}}}))=\Lambda _0\).\(\square \)

Proposition 3.8

For each \(t\in [0,1)\) the operator \(C_t:H({{\mathbb {D}}})\rightarrow H({{\mathbb {D}}})\) is power bounded, uniformly mean ergodic but, it fails to be supercyclic. Moreover, \( (I-C_t)(H({{\mathbb {D}}}))\) is the closed subspace of \(H({{\mathbb {D}}})\) given by

$$\begin{aligned} (I-C_t)(H({{\mathbb {D}}}))=\{g\in H({{\mathbb {D}}})\, :\, g(0)=0\} \end{aligned}$$
(3.7)

and we have the decomposition

$$\begin{aligned} H({{\mathbb {D}}})=\textrm{Ker}(I-C_t)\oplus (I-C_t)(H({{\mathbb {D}}})). \end{aligned}$$
(3.8)

Proof

Fix \(t\in [0,1)\). We first prove that \(C_t\) is power bounded. Once this is established, \(C_t\) is necessarily uniform mean ergodic because \(H({{\mathbb {D}}})\) is a Fréchet- Montel space (see [1, Proposition 2.8]).

Given \(k\ge 2\) we have, for every \(f\in H({{\mathbb {D}}})\) and with \(r_k:=(1-\frac{1}{k})\), that

$$\begin{aligned} |||C_tf|||_k&=\sup _{n\in {{\mathbb {N}}}_0}\left| \frac{1}{n+1}\sum _{j=0}^n t^{n-j}{\hat{f}}(j)\right| r_k\le \sup _{n\in {{\mathbb {N}}}_0}\frac{1}{n+1}\sum _{j=0}^n |{\hat{f}}(j)|r_k^n\\&\le \sup _{n\in {{\mathbb {N}}}_0}\frac{1}{n+1}\sum _{j=0}^n |{\hat{f}}(j)|r_k^j\le \sup _{j\in {{\mathbb {N}}}_0}|{\hat{f}}(j)|r_k^j =|||f|||_k, \end{aligned}$$

because \(r_k^n\le r_k^j\) for all \(j\in \{0,1,\ldots ,n\}\). It follows, for every \(n\in {{\mathbb {N}}}\), that

$$\begin{aligned} |||C_t^nf|||_k\le |||f|||_k,\quad f\in H({{\mathbb {D}}}). \end{aligned}$$

Since \(k\ge 2\) is arbitrary, the operator \(C_t\in {{\mathcal {L}}}(H({{\mathbb {D}}}))\) is indeed power bounded.

To establish that \(C_t:H({{\mathbb {D}}})\rightarrow H({{\mathbb {D}}})\) is not supercyclic, note that the continuous embedding \(\Phi :H({{\mathbb {D}}})\rightarrow \omega \) has dense range. The operator \(C_t^\omega \in {{\mathcal {L}}}(\omega )\) satisfies \(\Phi \circ C_t=C_t^\omega \circ \Phi \) as an identity in \({{\mathcal {L}}}(H({{\mathbb {D}}}),\omega )\), which implies if \(C_t:H({{\mathbb {D}}})\rightarrow H({{\mathbb {D}}})\) is supercyclic, then also \(C_t^\omega :\omega \rightarrow \omega \) must be supercyclic as \(\Phi \circ C_t^n=\Phi \circ C_t\circ C_t^{n-1}=C_t^\omega \circ \Phi \circ C_t^{n-1}=\cdots =(C_t^\omega )^n\circ \Phi \), for all \(n\in {{\mathbb {N}}}\), and \(\Phi (H({{\mathbb {D}}}))\) is dense in \(\omega \). A contradition with [5, Theorem 6.1].

To establish (3.7) note that \((I-C_t)(H({{\mathbb {D}}}))\subseteq \{g\in H({{\mathbb {D}}})\,:\ g(0)=0\}\) because \(C_tf(0)=f(0)\) for every \(f\in H({{\mathbb {D}}})\). To show the reverse inclusion, let \(g\in H({{\mathbb {D}}})\) satisfy \(g(0)=0\). Then \(h(z):=zg'(z)+g(z)\), for \(z\in {{\mathbb {D}}}\), is holomorphic and \(h(0)=0\). Accordingly, also \(z\mapsto \frac{h(z)}{z}\), for \(z\in {{\mathbb {D}}}{\setminus }\{0\}\), and taking the value \(h'(0)\) at \(z=0\) is holomorphic in \({{\mathbb {D}}}\). Define \(f\in H({{\mathbb {D}}})\) by

$$\begin{aligned} f(z):=\frac{1}{tz-1}\int _0^z(1-t\xi )\frac{h(\xi )}{\xi }\,d\xi ,\quad z\in {{\mathbb {D}}}, \end{aligned}$$

and note that \(f(0)=0\). Direct calculation reveals that

$$\begin{aligned} \frac{f(z)}{1-tz}-(zf(z))'=h(z)=(zg(z))',\quad z\in {{\mathbb {D}}}, \end{aligned}$$

from which it follows that

$$\begin{aligned} \int _0^z\frac{f(\xi )}{1-t\xi }\,d\xi -zf(z)=zg(z),\quad z\in {{\mathbb {D}}}. \end{aligned}$$

Since \(f(0)=0\), we can conclude that

$$\begin{aligned} \frac{1}{z}\int _0^z\frac{f(\xi )}{1-t\xi }\,d\xi -f(z)=g(z),\quad z\in {{\mathbb {D}}}, \end{aligned}$$

that is, \((C_t-I)f=g\) and so \(g\in (I-C_t)(H({{\mathbb {D}}}))\). Hence, (3.7) is valid.

To show the validity of (3.8) it suffices to repeat the argument given in the proof of Proposition 3.2. \(\square \)