1 Introduction and main results

Let \(\,{\mathbb {D}}=\{z\in {\mathbb {C}}:|z|<1\}\,\) denote the open unit disc in the complex plane \(\,{\mathbb {C}}\), \(\partial {\mathbb {D}}\,\) will be the unit circle. The space of all analytic functions in \(\,{\mathbb {D}}\,\) will be denoted by \(\,{\mathcal {H}}ol({\mathbb {D}})\). We also let \(\,H^p\,\) (\(0<p\le \infty \)) be the classical Hardy spaces. We refer to [11] for the notation and results regarding Hardy spaces.

For \(\,0<p<\infty \,\) and \(\,\alpha >-1\,\) the weighted Bergman space \(\,A^p_\alpha \,\) consists of those \(\,f\in {\mathcal {H}}ol({\mathbb {D}})\,\) such that

$$\begin{aligned} \Vert f\Vert _{A^p_\alpha }\,{\mathop {=}\limits ^{\text {def}}}\, \left( (\alpha +1)\int _{\mathbb {D}}(1-\vert z\vert ^2)^{\alpha }\vert f (z)\vert ^p\,dA(z)\right) ^{1/p}\,<\,\infty . \end{aligned}$$

Here, \(\,dA\,\) stands for the area measure on \(\,{\mathbb {D}}\), normalized so that the total area of \(\,{\mathbb {D}}\,\) is \(\,1\). Thus \(\,dA(z)\,=\,\frac{1}{\pi }\,dx\,dy\,=\,\frac{1}{\pi }\,r\,dr\,d\theta \). The unweighted Bergman space \(\,A^p_0\,\) is simply denoted by \(\,A^p\). We refer to [12, 18, 29] for the notation and results about Bergman spaces.

The space of Dirichlet type \(\,{{\mathcal {D}}^p_{\alpha }}\,\) (\(0<p<\infty \,\) and \(\,\alpha >-1\)) consists of those \(f\,\in {\mathcal {H}}ol({\mathbb {D}})\,\) such that \(\,f^\prime \in A^p_\alpha \). In other words, a function \(\,f\in {\mathcal {H}}ol({\mathbb {D}})\,\) belongs to \(\,{{\mathcal {D}}^p_{\alpha }}\,\) if and only if

$$\begin{aligned} \Vert f\Vert _{{{\mathcal {D}}^p_{\alpha }}}\,{\mathop {=}\limits ^{\text {def}}}\,\vert f(0)\vert \,+\,\left( (\alpha +1)\int _{{\mathbb {D}}}(1-\vert z\vert ^2)^{\alpha }\vert f^\prime (z)\vert ^p\,dA(z)\right) ^{1/p}\,<\,\infty . \end{aligned}$$

The Hilbert matrix is the infinite matrix \(\,{\mathcal {H}}=\left( \frac{1}{k+n+1}\right) _{k,n\ge 0}\). It induces formally an operator, called the Hilbert operator, on spaces of analytic functions as follows:

If \(\,f\in {\mathcal {H}}ol({\mathbb {D}})\), \(f(z)=\sum _{n=0}^\infty a_nz^n\), then we set

$$\begin{aligned} \mathcal {H}f(z)=\sum _{n=0}^\infty \left( \sum _{k=0}^\infty \frac{a_k}{n+k+1}\right) z^n,\quad z\in {\mathbb {D}}, \end{aligned}$$
(1)

whenever the right-hand side of (1) makes sense for all \(\,z\in {\mathbb {D}}\,\) and the resulting function is analytic in \(\,{\mathbb {D}}\). We define also

$$\begin{aligned} \mathcal {I}f(z)=\int _0^1\frac{f(t)}{1-tz}\,dt,\quad z\in {\mathbb {D}}, \end{aligned}$$
(2)

if the integrals in the right-hand side of (2) converge for all \(\,z\in {\mathbb {D}}\,\) and the resulting function \(\,\mathcal {I}f\,\) is analytic in \(\,{\mathbb {D}}\). It is clear that the correspondences \(\,f\mapsto \mathcal {H}f\,\) and \(\,f\mapsto \mathcal {I}f\,\) are linear.

If \(\,f\in H^1\), \(f(z)=\sum _{n=0}^\infty a_nz^z\), then by the Fejér-Riesz inequality [11, Theorem 3. 13, p. 46] and Hardy’s inequality [11, p. 48], we have

$$\begin{aligned} \int _0^1\vert f(t)\vert \,dt\le \pi \Vert f\Vert _{H^1}\quad \text {and}\quad \sum _{n=0}^\infty \frac{a_n}{n+1}\le \pi \Vert f\Vert _{H^1}. \end{aligned}$$

This immediately yields that if \(\,f\in H^1\,\) then \(\,\mathcal {H}f\,\) and \(\mathcal {I}f\,\) are well defined analytic functions in \(\,{\mathbb {D}}\,\) and that, furthermore, \(\mathcal {H}f=\mathcal {I}f\).

Diamantopoulos and Siskakis [9] proved that \(\,{\mathcal {H}}\,\) is a bounded operator from \(\,H^p\,\) into itself if \(\,1<p<\infty \), but this is not true for \(\,p=1\). In fact, they proved that \({\mathcal {H}}\left( H^1\right) \nsubseteq H^1\). Cima [6] has recently proved the following result.

Theorem A

  1. (i)

    The operator \(\,{\mathcal {H}} \,\) maps \(\,H^1\,\) into the space \(\,{\mathscr {C}}\,\) of Cauchy transforms of measures on the unit circle \(\,\partial {\mathbb {D}}\).

  2. (ii)

    \(\,{\mathcal {H}}:H^1\rightarrow {\mathscr {C}}\,\) is injective.

We recall that if \(\,\sigma \,\) is a finite complex Borel measure on \(\,\partial {\mathbb {D}}\), the Cauchy transform \(\,C\sigma \,\) is defined by

$$\begin{aligned} C\sigma (z)=\int _{\partial {\mathbb {D}}}\frac{d\sigma (\xi )}{1-\overline{\xi }\,z},\quad z\in {\mathbb {D}}. \end{aligned}$$

We let \(\mathscr {M}\) be the space of all finite complex Borel measure on \(\,\partial {\mathbb {D}}\). It is a Banach space with the total variation norm. The space of Cauchy transforms is \(\,{\mathscr {C}}=\{ C\sigma : \sigma \in \mathscr {M}\}\). It is a Banach space with the norm \(\,\Vert C\sigma \Vert {\mathop {=}\limits ^{\text {def}}}\inf \{ \Vert \tau \Vert : C\tau =C\sigma \}\). We mention [7] as an excellent reference for the main results about Cauchy transforms. We let \(\,\mathcal {A}\,\) denote the disc algebra, that is, the space of analytic functions in \(\,{\mathbb {D}}\,\) with a continuous extension to the closed unit disc, endowed with the \(\,\Vert \cdot \Vert _{H^\infty }\)-norm. It turns out [7, Chapter 4] that \(\,\mathcal {A}\,\) can be identified with the pre-dual of \(\,{\mathscr {C}}\,\) via the pairing

$$\begin{aligned} \langle g, C\sigma \rangle \, {\mathop {=}\limits ^{\text {def}}}\,\lim _{r\rightarrow 1}\frac{1}{2\pi }\int _0^{2\pi }g(re^{i\theta })\overline{C\sigma (re^{i\theta })}\,d\theta .\end{aligned}$$
(3)

This is the basic ingredient used by Cima to prove the inclusion \(\,{\mathcal {H}}(H^1)\subset {\mathscr {C}}\).

Now we turn to consider a class of operators which are natural generalizations of the operators \(\,{\mathcal {H}}\,\) and \(\,{\mathcal {I}}\). If \(\,\mu \,\) is a finite positive Borel measure on \(\,[0, 1)\,\) and \(\,n\, = 0, 1, 2, \ldots \), we let \(\,\mu _n\,\) denote the moment of order \(\,n\,\) of \(\,\mu \), that is, \(\mu _n=\int _{[0,1)}t^n\,d\mu (t),\) and we define \(\,{\mathcal {H}}_\mu \,\) to be the Hankel matrix \(\,(\mu _{n,k})_{n,k\ge 0}\,\) with entries \(\,\mu _{n,k}=\mu _{n+k}\). The measure \(\,\mu \,\) induces formally the operators \(\,{\mathcal {I}}_\mu \,\) and \(\,{\mathcal {H}}_\mu \,\) on spaces of analytic functions as follows:

$$\begin{aligned}{\mathcal {I}}_\mu f(z)=\int _{[0,1)}\frac{f(t)}{1-tz}\,d\mu (t),\quad {\mathcal {H}}_\mu f(z)=\sum _{n=0}^\infty \left( \sum _{k=0}^\infty a_k\mu _{n+k}\right) z^n,\quad z\in {\mathbb {D}},\end{aligned}$$

for \(\,f(z)=\sum _{n=0}^\infty a_nz^n\in {\mathcal {H}}ol({\mathbb {D}})\,\) being such that the terms on the right-hand sides make sense for all \(\,z\in {\mathbb {D}}\), and the resulting functions are analytic in \(\,{\mathbb {D}}\). If \(\,\mu \,\) is the Lebesgue measure on \(\,[0,1)\,\) the matrix \(\,{\mathcal {H}}_\mu \,\) reduces to the classical Hilbert matrix and the operators \(\,{\mathcal {H}}_\mu \,\) and \(\,{\mathcal {I}}_\mu \,\) are simply the operators \(\,{\mathcal {H}}\,\) and \(\,{\mathcal {I}}\).

If \(\,I\subset \partial {\mathbb {D}}\,\) is an interval, \(\,\vert I\vert \,\) will denote the length of \(\,I\). The Carleson square\(\,S(I)\,\) is defined as \(\,S(I)=\{re^{it}:\,e^{it}\in I,\quad 1-\frac{|I|}{2\pi }\le r <1\}\).

If \(\, s>0\,\) and \(\,\mu \,\) is a positive Borel measure on \(\,{\mathbb {D}}\), we shall say that \(\,\mu \,\) is an s-Carleson measure if there exists a positive constant \(\,C\,\) such that

$$\begin{aligned} \mu \left( S(I)\right) \le C{|I|^s}, \quad \hbox {for any interval}~ I\subset \partial {\mathbb {D}}. \end{aligned}$$

A 1-Carleson measure will be simply called a Carleson measure. We recall that Carleson [4] proved that \(\,H^p\,\subset \,L^p(d\mu )\,\) (\(0<p<\infty \)) if and only if \(\,\mu \,\) is a Carleson measure (see also [11, Chapter 9]).

For  \(0\le \alpha <\infty \,\) and \(\,0<s<\infty \,\) we say that a positive Borel measure \(\,\mu \,\) on \(\,{\mathbb {D}}\,\) is an \(\alpha \)-logarithmic s-Carleson measure if there exists a positive constant C such that

$$\begin{aligned}\frac{ \mu \left( S(I)\right) \left( \log \frac{2\pi }{\vert I\vert }\right) ^\alpha }{|I|^s}\le C, \quad \hbox {for any interval}~ I\subset \partial {\mathbb {D}}. \end{aligned}$$

A positive Borel measure \(\mu \) on [0, 1) can be seen as a Borel measure on \({\mathbb {D}}\) by identifying it with the measure \(\tilde{\mu } \) defined by

$$\begin{aligned} {\tilde{\mu }} (A)\,=\,\mu \left( A\cap [0,1)\right) ,\quad \text {for any Borel subset}~ A~ \text {of}~ {\mathbb {D}}. \end{aligned}$$

In this way a positive Borel measure \(\mu \) on [0, 1) is an s-Carleson measure if and only if there exists a positive constant C such that

$$\begin{aligned} \mu \left( [t,1)\right) \le C(1-t)^s, \quad 0\le t<1. \end{aligned}$$

We have a similar statement for \(\alpha \)-logarithmic s-Carleson measures.

The action of the operators \(\,{\mathcal {I}}_\mu \,\) and \(\,{\mathcal {H}}_\mu \,\) on distinct spaces of analytic functions have been studied in a number of articles (see, e. g., [2, 5, 14,15,16, 22, 25, 27]).

Combining results of [14] and of [16] we can state the following result.

Theorem B

Let \(\,\mu \,\) be a finite positive Borel measure on \(\,[0,1)\).

  1. (i)

    The operator \(\,{\mathcal {I}}_\mu \,\) is well defined on \(\,H^1\,\) if and only if \(\,\mu \,\) is a Carleson measure.

  2. (ii)

    If \(\,\mu \,\) is a Carleson measure, then the operator \(\,{\mathcal {H}}_\mu \,\) is also well defined on \(\,H^1\,\) and \(\,{\mathcal {I}}_\mu f={\mathcal {H}}_\mu f\,\) for all \(\,f\in H^1\).

  3. (iii)

    The operator \(\,{\mathcal {H}}_\mu \,\) is a bounded operator from \(\,H^1\,\) into itself if and only if \(\,\mu \,\) is a 1-logarithmic 1-Carleson measure.

Galanopoulos and Peláez [14, Theorem 2. 2] proved the following.

Theorem C

Let \(\,\mu \,\) be a positive Borel measure on \(\,[0,1)\). If \(\,\mu \,\) is a Carleson measure then \(\,{\mathcal {H}}_\mu (H^1)\subset {\mathscr {C}}\).

This result is stronger than Theorem A(i). In view of these results, the following question arises naturally.

Question 1

Suppose that \(\,\mu \,\) is a 1-logarithmic 1-Carleson measure on \(\,[0,1)\). What can we say about the image \(\,{\mathcal {H}}_\mu (H^1)\,\) of \(\,H^1\,\) under the action of the operator \(\,{\mathcal {H}}_\mu \)?

To answer Question 1, let us start noticing that it is known that, for \(0<p\le 2\), the space of Dirichlet type \(\,{{\mathcal {D}}^p_{p-1}}\) is continuously included in \(H^p\,\) (see [26, Lemma 1. 4]). In particular, the space \(\,{\mathcal {D}}^1_0\,\) is continuously included in \(\,H^1\). In fact, the estimates obtained by Vinogradov in the proof of his lemma easily yield the inequality

$$\begin{aligned} \Vert f\Vert _{H^1}\le 2\Vert f\Vert _{{\mathcal {D}}^1_0},\quad f\in {\mathcal {D}}^1_0. \end{aligned}$$

We shall prove that if \(\,\mu \,\) is a 1-logarithmic 1-Carleson measure on \(\,[0,1)\,\) then \(\,{\mathcal {H}}_\mu (H^1)\,\) is contained in the space \(\,{\mathcal {D}}^1_0\). Actually, we have the following stronger result.

Theorem 1

Let \(\,\mu \,\) be a positive Borel measure on \(\,[0,1)\). Then the following conditions are equivalent.

  1. (i)

    \(\mu \,\) is a 1-logarithmic 1-Carleson measure.

  2. (ii)

    \({\mathcal {H}}_\mu \,\) is a bounded operator from \(\,H^1\,\) into itself.

  3. (iii)

    \({\mathcal {H}}_\mu \,\) is a bounded operator from \(\,H^1\,\) into \(\,{\mathcal {D}}^1_0\).

  4. (iv)

    \({\mathcal {H}}_\mu \,\) is a bounded operator from \(\,{\mathcal {D}}^1_0\,\) into \(\,{\mathcal {D}}^1_0\).

There is a gap between Theorem C and Theorem 1 and so it is natural to discuss the range of \(\,H^1\,\) under the action of \(\,{\mathcal {H}}_\mu \,\) when \(\,\mu \,\) is an \(\,\alpha \)-logarithmic 1-Carleson measure with \(\,0<\alpha <1\). We shall prove the following result.

Theorem 2

Let \(\,\mu \,\) be a positive Borel measure on \(\,[0,1)\). Suppose that \(\,0<\alpha <1\,\) and that \(\,\mu \,\) is an \(\,\alpha \)-logarithmic 1-Carleson measure. Then \(\,{\mathcal {H}}_\mu \,\) maps \(\,H^1\,\) into the space \(\,{\mathcal {D}}^1(\log ^{\alpha -1})\,\) defined as follows:

$$\begin{aligned} {\mathcal {D}}^1(\log ^{\alpha -1})=\left\{ f\in {\mathcal {H}}ol({\mathbb {D}}) : \int _{{\mathbb {D}}} \vert f^\prime (z)\vert \left( \log \frac{2}{1-\vert z\vert }\right) ^{\alpha -1}\,dA(z)<\infty \right\} . \end{aligned}$$

These results will be proved in Sect. 2. Since the space of Dirichlet type \(\,{\mathcal {D}}^1_0\,\) has showed up in a natural way in our work, it seems natural to study the action of the operators \(\,{\mathcal {H}}_\mu \,\) and \(\,{\mathcal {I}}_\mu \,\) on the Bergman spaces \(\,A^p_{\alpha }\,\) and the Dirichlet spaces \(\,{{\mathcal {D}}^p_{\alpha }}\,\) for general values of the parameters \(\,p\,\) and \(\,\alpha \). This will be done in Sect. 3.

Throughout this paper the letter C denotes a positive constant that may change from one step to the next. Moreover, for two real-valued functions \(E_1, E_2\) we write \(E_1\lesssim E_2\), or \(E_1\gtrsim E_2\), if there exists a positive constant C independent of the arguments such that \(E_1\le C E_2\), respectively \(E_1\ge C E_2\). If we have \(E_1\lesssim E_2\) and \(E_1\gtrsim E_2\) simultaneously then we say that \(E_1\) and \(E_2\) are equivalent and we write \(E_1\asymp E_2\).

2 Proofs of the theorems 1 and  2

Proof of Theorem 1

We already know that (i) and (ii) are equivalent by Theorem B.

To prove that (i) implies (iii) we shall use some results about the Bloch space. We recall that a function \(\,f\in {\mathcal {H}}ol({\mathbb {D}})\,\) is said to be a Bloch function if

$$\begin{aligned} \Vert f\Vert _{\mathcal {B}}\,{\mathop {=}\limits ^{\text {def}}}\,\vert f(0)\vert \,+\,\sup _{z\in {\mathbb {D}}}(1-\vert z\vert ^2)\vert f^\prime (z)\vert \,<\,\infty . \end{aligned}$$

The space of all Bloch functions will be denoted by \(\,\mathcal {B}\). It is a non-separable Banach space with the norm \(\,\Vert \cdot \Vert _{\mathcal {B}}\,\) just defined. A classical source for the theory of Bloch functions is [1]. The closure of the polynomials in the Bloch norm is the little Bloch space\(\,\mathcal {B}_0\,\) which consists of those \(\,f\in {\mathcal {H}}ol({\mathbb {D}})\,\) with the property that

$$\begin{aligned}\lim _{\vert z\vert \rightarrow 1}(1-\vert z\vert ^2)\vert f^\prime (z)\vert =0.\end{aligned}$$

It is well known that (see [1, p. 13])

$$\begin{aligned} \vert f(z)\vert \lesssim \Vert f\Vert _{\mathcal {B}}\log \frac{2}{1-\vert z\vert }.\end{aligned}$$
(4)

The basic ingredient to prove that (i) implies (iii) is the fact that the dual \(\,\left( \mathcal {B}_0\right) ^*\,\) of the little Bloch space can be identified with the Bergman space \(\,A^1\) via the integral pairing

$$\begin{aligned} \langle h,f\rangle \,=\,\int _{{\mathbb {D}}}\,h(z)\,\overline{f(z)}\,dA(z),\quad h\in \mathcal {B}_0, \, f\in A^1. \end{aligned}$$
(5)

(See [29, Theorem 5. 15]).

Let us proceed to prove the implication (i) \(\Rightarrow \) (iii). Assume that \(\,\mu \,\) is a 1-logarithmic 1-Carleson measure and take \(\,f\in H^1\). We have to show that \(\,{\mathcal {I}}_\mu f\in {\mathcal {D}}^1_0\,\) or, equivalently, that \(\,\left( {\mathcal {I}}_\mu f\right) ^\prime \in A^1\). Since \(\,\mathcal {B}_0\,\) is the closure of the polynomials in the Bloch norm, it suffices to show that

$$\begin{aligned} \left| \int _{{\mathbb {D}}}\,h(z)\,\overline{\left( {\mathcal {I}}_\mu f\right) ^\prime (z)}\,dA(z)\right| \,\lesssim \Vert h\Vert _{\mathcal {B}}\Vert f\Vert _{H^1},\quad \text {for any polynomial}~ h.\end{aligned}$$
(6)

So, let \(\,h\,\) be a polynomial. We have

$$\begin{aligned} \int _{{\mathbb {D}}}\,h(z)\,\overline{\left( {\mathcal {I}}_\mu f\right) ^\prime (z)}\,dA(z)\,=\,&\int _{{\mathbb {D}}}\,h(z)\,\overline{\left( \int _{[0,1)}\frac{t\,f(t)}{(1-tz)^2}\,d\mu (t)\right) }\,dA(z)\\ =\,&\int _{{\mathbb {D}}}\,h(z)\,\int _{[0,1)}\frac{t\,\overline{f(t)}}{(1-t\,\overline{z})^2}\,d\mu (t)\,dA(z)\\ =\,&\int _{[0,1)}\,t\,\overline{f(t)}\int _{{\mathbb {D}}}\,\frac{h(z)}{(1-t\,\overline{z})^2}\,dA(z)\,d\mu (t). \end{aligned}$$

Because of the reproducing property of the Bergman kernel [29, Proposition 4. 23], \(\,\int _{{\mathbb {D}}}\,\frac{h(z)}{(1-t\,\overline{z})^2}\,dA(z)\,=\,h(t)\). Then it follows that

$$\begin{aligned} \int _{{\mathbb {D}}}\,h(z)\,\overline{\left( {\mathcal {I}}_\mu f\right) ^\prime (z)}\,dA(z)\,=\,\int _{[0,1)}\,t\,\overline{f(t)}\,h(t)\,d\mu (t).\end{aligned}$$
(7)

Since \(\,\mu \,\) is a 1-logarithmic 1-Carleson measure, the measure \(\,\nu \,\) defined by

$$\begin{aligned} d\nu (t)\,=\,\log \frac{2}{1-t}\,d\mu (t) \end{aligned}$$

is a Carleson measure [15, Proposition 2. 5]. This implies that

$$\begin{aligned} \int _{[0,1)}\vert f(t)\vert \log \frac{2}{1-t}\,d\mu (t)\lesssim \Vert f\Vert _{H^1}. \end{aligned}$$

This and (4) yield

$$\begin{aligned} \int _{[0,1)}\left| t\,\overline{f(t)}\,h(t)\right| \,d\mu (t)\,\lesssim \,\Vert h\Vert _{\mathcal {B}}\Vert f\Vert _{H^1}. \end{aligned}$$

Using this and (7), (6) follows.

Since \(\,{\mathcal {D}}^1_0\subset H^1\,\), the implication (iii)  \(\Rightarrow \)  (iv) is trivial. To prove that (iv) implies (i) we shall use the following result of Pavlović [23, Theorem 3. 2].

Theorem D

Let \(\,f\in {\mathcal {H}}ol({\mathbb {D}})\), \(\,f(z)=\sum _{n=0}^\infty a_nz^n\), and suppose that the sequence \(\,\{ a_n\} \) is a decreasing sequence of non-negative real numbers. Then \(\,f\in {\mathcal {D}}^1_0\,\) if and only if \(\,\sum _{n=0}^\infty \frac{a_n}{n+1}<\infty \), and we have

$$\begin{aligned} \Vert f\Vert _{{\mathcal {D}}^1_0}\,\asymp \,\sum _{n=0}^\infty \frac{a_n}{n+1}. \end{aligned}$$

Now we turn to prove the implication (iv) \(\Rightarrow \) (i). Assume that \({\mathcal {H}}_\mu \) is a bounded operator from \(\,{\mathcal {D}}^1_0\,\) into \(\,{\mathcal {D}}^1_0\). We argue as in the proof of Theorem 1. 1 of [16]. For \(\,\frac{1}{2}<b<1\,\) set

$$\begin{aligned} f_b(z)=\frac{1-b^2}{(1-bz)^2},\quad z\in {\mathbb {D}}. \end{aligned}$$

We have \(\,f_b^\prime (z)=\frac{2b(1-b^2)}{(1-bz)^3}\,\) (\(z\in {\mathbb {D}}\)). Then, using Lemma 3. 10 of [29] with \(\,t=0\,\) and \(\,c=1\), we see that

$$\begin{aligned} \Vert f_b\Vert _{{\mathcal {D}}^1_0}\,\asymp \,\int _{{\mathbb {D}}}\frac{1-b^2}{\vert 1-bz\vert ^3}\,dA(z)\,\asymp 1. \end{aligned}$$

Since \(\,{\mathcal {H}}_\mu \,\) is bounded on \(\,{\mathcal {D}}^1_0\), this implies that

$$\begin{aligned} 1\gtrsim \Vert {\mathcal {H}}_\mu (f_b)\Vert _{{\mathcal {D}}^1_0}.\end{aligned}$$
(8)

We also have,

$$\begin{aligned} f_b(z)\,=\,\sum _{k=0}^\infty a_{k,b}z^k,\quad \text {with}~ a_{k,b}=(1-b^2)(k+1)b^k. \end{aligned}$$

Since the \(a_{k,b}\)’s are positive, it is clear that the sequence \(\{ \sum _{k=0}^\infty \mu _{n+k}a_{k,b}\} _{n=0}^\infty \) of the Taylor coefficients of \({\mathcal {H}}_\mu (f_b)\) is a decreasing sequence of non-negative real numbers. Using this, Theorem D, (8), and the definition of the \(a_{k,b}\)’s, we obtain

$$\begin{aligned} 1\,&\gtrsim \,\Vert {\mathcal {H}}_\mu (f_b)\Vert _{{\mathcal {D}}^1_0}\,\gtrsim \,\sum _{n=1}^\infty \frac{1}{n}\left( \sum _{k=0}^\infty \mu _{n+k}a_{k,b}\right) \\&= \sum _{n=1}^\infty \frac{1}{n}\left( \sum _{k=0}^\infty a_{k,b}\int _{[0,1)}t^{n+k}\,d\mu (t)\right) \\&\gtrsim \,(1-b^2)\sum _{n=1}^\infty \frac{1}{n}\left( \sum _{k=1}^\infty kb^k\int _{[b,1)}t^{n+k}\,d\mu (t)\right) \\&\gtrsim \,(1-b^2)\sum _{n=1}^\infty \frac{1}{n}\left( \sum _{k=1}^\infty kb^{n+2k}\,\mu \left( [b,1)\right) \right) \\&=\,(1-b^2)\mu \left( [b,1)\right) \sum _{n=1}^\infty \frac{b^n}{n}\left( \sum _{k=1}^\infty kb^{2k}\right) \\&= \,(1-b^2)\mu \left( [b,1)\right) \left( \log \frac{1}{1-b}\right) \frac{b^2}{(1-b^2)^2}. \end{aligned}$$

Then it follows that

$$\begin{aligned} \mu \left( [b,1)\right) \,=\,{{\,\mathrm{O}\,}}\left( \frac{1-b}{\log \frac{1}{1-b}}\right) ,\quad \text {as}~ b\rightarrow 1. \end{aligned}$$

Hence, \(\mu \) is a 1-logarithmic 1-Carleson measure. \(\square \)

Before embarking on the proof of Theorem 2 we have to introduce some notation and results. Following [24], for \(\,\alpha \in \mathbb {R}\,\) the weighted Bergman space \(\,A^1(\log ^\alpha )\,\) consists of those \(\,f\in {\mathcal {H}}ol({\mathbb {D}})\,\) such that

$$\begin{aligned} \Vert f\Vert _{A^1(\log ^\alpha )}\,{\mathop {=}\limits ^{\text {def}}}\,\int _{{\mathbb {D}}}\vert f(z)\vert \,\left( \log \frac{2}{1-\vert z\vert }\right) ^\alpha \,dA(z)<\infty . \end{aligned}$$

This is a Banach space with the norm \(\,\Vert \cdot \Vert _{A^1(\log ^\alpha )}\,\) just defined and the polynomials are dense in \(\,A^1(\log ^\alpha )\). Likewise, we define

$$\begin{aligned} {\mathcal {D}}^1(\log ^\alpha )=\{ f\in {\mathcal {H}}ol({\mathbb {D}}) : f^\prime \in A^1(\log ^\alpha )\} . \end{aligned}$$

We define also the Bloch-type space \(\,\mathcal {B}(\log ^\alpha )\,\) as the space of those \(\,f\in {\mathcal {H}}ol({\mathbb {D}})\,\) such that

$$\begin{aligned} \Vert f\Vert _{{\mathcal {B}}(\log ^\alpha )}{\mathop {=}\limits ^{\text {def}}}\vert f(0)\vert +\sup _{z\in {\mathbb {D}}}(1-\vert z\vert ^2)\left( \log \frac{2}{1-\vert z\vert }\right) ^{-\alpha }\vert f^\prime (z)\vert <\infty , \end{aligned}$$

and

$$\begin{aligned} {\mathcal {B}}_0(\log ^\alpha )=\left\{ f\in {\mathcal {H}}ol({\mathbb {D}}) : \vert f^\prime (z)\vert ={{\,\mathrm{o}\,}}\left( \frac{\left( \log \frac{2}{1-\vert z\vert }\right) ^\alpha }{1-\vert z\vert }\right) ,\,\,\,\text {as}~ \,\vert z\vert \rightarrow 1~ \right\} . \end{aligned}$$

The space \(\,\mathcal {B}(\log ^\alpha )\,\) is a Banach space and \(\,\mathcal {B}_0(\log ^\alpha )\,\) is the closure of the polynomials in \(\,\mathcal {B}(\log ^\alpha )\).

We remark that the spaces \(\,{\mathcal {D}}^1(\log ^\alpha )\,\), \(\,\mathcal {B}(\log ^\alpha )\,\), and \(\,\mathcal {B}_0(\log ^\alpha )\,\) were called \(\,\mathfrak {B}^1_{\log ^\alpha }\), \(\,\mathfrak {B}_{\log ^\alpha }\), and \(\,\mathfrak {b}_{\log ^\alpha }\,\) in [24]. Pavlović identified in [24, Theorem 2. 4] the dual of the space \(\,{\mathcal {B}}_0(\log ^\alpha )\).

Theorem E

Let \(\,\alpha \in \mathbb {R}\). Then the dual of \(\,{\mathcal {B}}_0(\log ^\alpha )\,\) is \(\,A^1(\log ^\alpha )\,\) via the pairing

$$\begin{aligned} \langle h, g\rangle \,=\,\int _{{\mathbb {D}}}\,f(z)\,\overline{g(z)}\,dA(z),\quad h\in {\mathcal {B}}_0(\log ^\alpha ),\,\,\,g \in A^1(\log ^\alpha ). \end{aligned}$$

Actually, Pavlović formulated the duality theorem in another way but it is a simple exercise to show that his formulation is equivalent to this one which is better suited to our work.

Proof of Theorem 2

Let \(\,\mu \,\) be a positive Borel measure on \(\,[0,1)\,\) and \(\,0<\alpha <1\). Suppose that \(\,\mu \,\) is an \(\,\alpha \)-logarithmic 1-Carleson measure. Take \(\,f\in H^1\). We have to show that \(\,{\mathcal {I}}_\mu f\in {\mathcal {D}}^1(\log ^{\alpha -1})\,\) or, equivalently, that \(\,\left( {\mathcal {I}}_\mu f\right) ^\prime \in A^1(\log ^{\alpha -1})\). Bearing in mind Theorem E and the fact that \(\,{\mathcal {B}}_0(\log ^{\alpha -1})\,\) is the closure of the polynomials in \(\,{\mathcal {B}}(\log ^{\alpha -1})\,\), it suffices to show that

$$\begin{aligned} \left| \int _{{\mathbb {D}}}\,h(z)\,\overline{\left( {\mathcal {I}}_\mu f\right) ^\prime (z)}\,dA(z)\right| \,\lesssim \Vert h\Vert _{{\mathcal {B}}(\log ^{\alpha -1})}\Vert f\Vert _{H^1},\quad \text {for any polynomial}~ h.\end{aligned}$$
(9)

So, let \(\,h\,\) be a polynomial. Arguing as in the proof of the implication (i) \(\Rightarrow \) (iii) in Theorem 1 we obtain

$$\begin{aligned} \int _{{\mathbb {D}}}h(z)\,\overline{\left( {\mathcal {I}}_\mu f\right) ^\prime (z)}\,dA(z)\,=\,\int _{[0,1)}t\,\overline{f(t)}\,h(t)\,d\mu (t).\end{aligned}$$
(10)

Now, it is clear that

$$\begin{aligned} \vert h(z)\vert \lesssim \Vert h\Vert _{{\mathcal {B}}(\log ^{\alpha -1})}\left( \log \frac{2}{1-\vert z\vert }\right) ^\alpha , \end{aligned}$$

and then it follows that

$$\begin{aligned} \int _{[0,1)}\left| t\,\overline{f(t)}\,h(t)\right| \,d\mu (t)\lesssim \Vert h\Vert _{{\mathcal {B}}(\log ^{\alpha -1})}\int _{[0,1)}\vert f(t)\vert \left( \log \frac{2}{1-t}\right) ^{\alpha }\,d\mu (t). \end{aligned}$$

Using the fact that the measure \(\,\left( \log \frac{2}{1-t}\right) ^{\alpha }\,d\mu (t)\,\) is a Carleson measure [15, Proposition 2. 5], this implies that

$$\begin{aligned} \int _{[0,1)}\left| t\,\overline{f(t)}\,h(t)\right| \,d\mu (t)\lesssim \Vert h\Vert _{{\mathcal {B}}(\log ^{\alpha -1})}\Vert f\Vert _{H^1}. \end{aligned}$$

This and (10) give (9). \(\square \)

3 The operators \({\mathcal {H}}_\mu \) acting on Bergman spaces and on Dirichlet spaces

Jevtić and Karapetrović [20] have recently proved the following result.

Theorem F

The Hilbert operator \(\,{\mathcal {H}}\,\) is a bounded operator from \(\,{{\mathcal {D}}^p_{\alpha }}\,\) into itself if and only if \(\,\max (-1, p-2)\,<\alpha \,<2p-2\).

Now, it is well known that \(\,A^p_\alpha \,=\,{\mathcal {D}}^p_{\alpha +p}\,\) (see [29, Theorem 4. 28]). Hence, regarding Bergman spaces Theorem F says the following.

Corollary G

The Hilbert operator \(\,{\mathcal {H}}\,\) is a bounded operator from \(\,A^p_\alpha \,\) into itself if and only if \(\,-1<\alpha <p-2\).

Let us recall that Diamantopoulos [8] had proved before that the Hilbert operator is bounded on \(\,A^p\,\) for \(\,p>2\), but not on \(\,A^2\). The situation on \(\,A^2\,\) is even worse. Dostanić, Jevtić, and Vukotić [10] proved that the Hilbert operator is not well defined on \(\,A^2\). Indeed, they considered the function \(\,f\,\) defined by

$$\begin{aligned} f(z)\,=\,\sum _{n=1}^\infty \frac{1}{\log (n+1)}\,z^n,\quad z\in {\mathbb {D}},\end{aligned}$$
(11)

which belongs to \(\,A^2.\) However, the series defining \(\,\mathcal {H}f(0)\,\) is \(\,\sum _{n=1}^\infty \frac{1}{(n+1)\log (n+1)}=\infty \,\) and the integral defining \(\,\mathcal {I}f(0)\,\) is \(\,\int _0^1\,f(t)\,dt\,=\,\infty \). Hence neither \(\,{\mathcal {H}}\,\) nor \(\,{\mathcal {I}}\,\) are defined on \(\,A^2\).

This result can be extended. We can assert that \(\,{\mathcal {H}}\,\) is not well defined on \(\,A^p_{p-2}\,\) for any \(\,p>1\). Indeed, let \(\,f\,\) be the function defined in (11). Notice that the sequence \(\,\{ \frac{1}{(n+1)\log (n+1)}\} \) is decreasing and that \(\,\sum _{n=1}^\infty \frac{1}{n\left( \log (n+1)\right) ^p}<\infty \). Then (see Proposition 1 below) it follows that \(\,f\in A^p_{p-2}\), and we have already seen that \(\,\mathcal {H}f\,\) and \(\,\mathcal {I}f\,\) are not defined. Since \(\,\alpha \ge p-2\,\,\Rightarrow \,\, A^p_{p-2} \subset A^p_\alpha \), it follows that the Hilbert operator \(\,{\mathcal {H}}\,\) is not defined on \(\,A^p_\alpha \,\) if \(\,\alpha \ge p-2\).

In this section we shall obtain extensions of the mentioned results of Jevtić and Karapetrović considering the generalized Hilbert operators \(\,{\mathcal {H}}_\mu \,\).

Theorem 3

Suppose that \(\,\max (-1, p-2)\,<\alpha \,<2p-2\,\) and let \(\,\mu \,\) be a finite positive Borel measure on \(\,[0,1)\). If \(\,\mu \,\) is a Carleson measure then the operators \(\,{\mathcal {H}}_\mu \,\) and \(\,{\mathcal {I}}_\mu \,\) are well defined on \(\,{{\mathcal {D}}^p_{\alpha }}\). Furthermore, \({\mathcal {I}}_\mu f={\mathcal {H}}_\mu f\), for all \(\,f\in {{\mathcal {D}}^p_{\alpha }}\).

When dealing with Bergman spaces Theorem 3 reduces to the following.

Corollary 1

Suppose that \(\,p>1\,\) and \(\,-1\,<\alpha \,<\,p-2\,\), and let \(\,\mu \,\) be a finite positive Borel measure on \(\,[0,1)\). If \(\,\mu \,\) is a Carleson measure then the operators \(\,{\mathcal {H}}_\mu \,\) and \(\,{\mathcal {I}}_\mu \,\) are well defined on \(\,A^p_\alpha \). Furthermore, \({\mathcal {I}}_\mu f={\mathcal {H}}_\mu f\), for all \(\,f\in A^p_\alpha \).

Proof of Theorem 3

Suppose that \(\,\mu \,\) is a Carleson measure and take \(\,f\in {{\mathcal {D}}^p_{\alpha }}\). Set \(\,\beta =\frac{2+\alpha }{p}-1\). Observe that \(\,0<\beta <1\). Using [29, Theorem 4. 14], we see that \(\,\vert f^\prime (z)\vert \,\lesssim \frac{1}{(1-\vert z\vert )^{(2+\alpha )/p}}\,\) and, hence, \(\,\vert f(z)\vert \,\lesssim \frac{1}{(1-\vert z\vert )^\beta }\). Then it follows that

$$\begin{aligned}\int _{[0,1)}\vert f(t)\vert \,d\mu (t)\,\lesssim \,\int _{[0,1)}\frac{d\mu (t)}{(1-t)^\beta }.\end{aligned}$$

Integrating by parts, using that \(\,\mu \,\) is a Carleson measure, and that \(\,0<\beta <1\), we obtain

$$\begin{aligned} \int _{[0,1)}\frac{d\mu (t)}{(1-t)^\beta }\,&=\,\mu ([0,1))\,-\,\lim _{t\rightarrow 1}\frac{\mu ([t,1))}{(1-t)^{\beta }}\,+\,\beta \int _0^1\,\frac{\mu ([t,1))}{(1-t)^{\beta +1}}\,dt \\&=\,\mu ([0,1))\,+\,\beta \int _0^1\,\frac{\mu ([t,1))}{(1-t)^{\beta +1}}\,dt \\&\lesssim \,\mu ([0,1))\,+\,\int _0^1\,\frac{1}{(1-t)^{\beta }}\,dt \\&<\,\infty . \end{aligned}$$

Consequently, we obtain that

$$\begin{aligned} \int _{[0,1)}\vert f(t)\vert \,d\mu (t)\,<\,\infty .\end{aligned}$$
(12)

Clearly, this implies that the integral

$$\begin{aligned} \int _{[0,1)}\frac{f(t)\,d\mu (t)}{1-tz}\quad \text {converges absolutely and uniformly on compact subsets of}~ {\mathbb {D}}.\end{aligned}$$
(13)

This gives that \(\,{\mathcal {I}}_\mu f\,\) is a well defined analytic function in \({\mathbb {D}}\) and that

$$\begin{aligned} {\mathcal {I}}_\mu f(z)=\sum _{n=0}^\infty \left( \int _{[0,1)}t^n\,f(t)\,d\mu (t)\right) z^n,\quad z\in {\mathbb {D}}.\end{aligned}$$
(14)

Using [19, Theorem 2. 1] (see also [20, Theorem 2. 1]) we see that for these values of \(\,p\,\) and \(\,\alpha \,\) we have that if \(\,f\in A^p_\alpha \), \(f(z)=\sum _{n=0}^\infty a_nz^n\), then \(\,\sum _{k=0}^\infty \frac{\vert a_k\vert }{k+1}<\infty \). Now, since \(\,\mu \,\) is a Carleson measure we have that \(\,\vert \mu _n\vert \lesssim \frac{1}{n+1}\) ([5, Proposition 1]). Then it follows that

$$\begin{aligned} \sum _{k=0}^\infty \vert \mu _{n+k}a_k\vert \,\lesssim \,\sum _{k=0}^\infty \frac{\vert a_k\vert }{k+n+1}\lesssim \sum _{k=0}^\infty \frac{\vert a_k\vert }{k+1},\quad \text {for all}~ \,n. \end{aligned}$$

Clearly, this implies that \(\,{\mathcal {H}}_\mu f\,\) is a well defined analytic function in \(\,{\mathbb {D}}\,\) and that \(\,\int _{[0,1)}t^n\,f(t)\,d\mu (t)\,=\,\sum _{k=0}^\infty \mu _{n+k}a_k\,\) for all \(\,n\). This and (13) give that \(\,{\mathcal {I}}_\mu f\,=\,{\mathcal {H}}_\mu f\). \(\square \)

Our next result is an extension of Corollary G.

Theorem 4

Suppose that \(\,-1\,<\alpha \,<\,p-2\,\) and let \(\,\mu \,\) be a finite positive Borel measure on \(\,[0,1)\).

The operator \(\,{\mathcal {H}}_\mu \,\) is well defined on \(\,A^p_\alpha \,\) and it is a bounded operator from \(\,A^p_\alpha \,\) to itself if and only if \(\,\mu \,\) is a Carleson measure.

A number of results will be needed to prove this theorem. We start with a characterization of the functions \(\,f\in {\mathcal {H}}ol({\mathbb {D}})\,\) whose sequence of Taylor coefficients is decreasing which belong to \(\,A^p_\alpha \).

Proposition 1

Let \(\,f\in {\mathcal {H}}ol({\mathbb {D}})\), \(f(z)=\sum _{n=0}^\infty a_n\,z^n\)\(\,(z\in {\mathbb {D}})\). Suppose that \(\,1<p<\infty \), \(\,\alpha >-1\), and that the sequence \(\{ a_n\} _{n=0}^\infty \) is a decreasing sequence of non-negative real numbers. Then

$$\begin{aligned} f\in A^p_\alpha \,\,\,\Leftrightarrow \,\,\,\sum _{n=1}^\infty n^{p-3-\alpha }a_n^p\,<\,\infty . \end{aligned}$$

Furthermore, \(\Vert f\Vert _{A^p_\alpha }^p\,\asymp \,\vert a_0\vert ^p\,+\,\sum _{n=1}^\infty n^{p-3-\alpha }a_n^p\,<\,\infty \).

This result can be proved with arguments similar to those used in the proofs of [15, Theorem 3.10] and [23, Theorem 3. 1] where the analogous results for the Besov spaces \(\,B^p={\mathcal {D}}^p_{p-2}\) (\(p>1\)) and for the spaces \(\,{{\mathcal {D}}^p_{p-1}}\,\) (\(p>1\)) were proved. The case \(\,\alpha =0\,\) is proved in [3, Proposition 2. 4]. Consequently, we omit the details.

The following lemma is a generalization of [13, Lemma 3 (ii)].

Lemma 1

Let \(\,\mu \,\) be a positive Borel measure on \(\,[0,1)\,\) which is a Carleson measure. Assume that \(\,0<p<\infty \,\) and \(\,\alpha >-1\). Then there exists a positive constant \(\,C=C(p,\alpha ,\mu )\,\) such that for any \(\,f\in A^p_\alpha \,\)

$$\begin{aligned} \int _{[0,1)}M_\infty ^p(r,f)(1-r)^{\alpha +1}\,d\mu (r)\,\le C\Vert f\Vert _{A^p_\alpha }^p. \end{aligned}$$

Of course, \(M_\infty (r,f)=\sup _{\vert z\vert =r}\vert f(z)\vert \).

Proof

Take \(\,f\in A^p_\alpha \,\) and set

$$\begin{aligned}&g(r)=M_\infty ^p(r,f)(1-r)^{\alpha +1},\,\, \\&\quad F(r)=\mu ([0,r))\,-\mu ([0,1))\,=\,-\mu ([r,1)),\,\, 0<r<1. \end{aligned}$$

Integrating by parts, we have

$$\begin{aligned}&\int _{[0,1)}M_\infty ^p(r,f)(1-r)^{\alpha +1}\,d\mu (r)\,=\,\int _{[0,1)}g(r)\,d\mu (r)\nonumber \\&\quad =\,\lim _{r\rightarrow 1}g(r)F(r)\,-\,g(0)F(0)\,-\int _0^1g^\prime (r)F(r)\,dr \nonumber \\&\quad =\,\vert f(0)\vert ^p\mu ([0,1))\,-\,\lim _{r\rightarrow 1}M_\infty ^p(r,f)(1-r)^{\alpha +1}\mu ([r,1))\,\nonumber \\&\qquad +\,\int _0^1g^\prime (r)\mu ([r, 1))\,dr. \end{aligned}$$
(15)

Since \(\,f\in A^p_\alpha \,\) we have that \(\,M_\infty ^p(r,f)={{\,\mathrm{o}\,}}\left( (1-r)^{-2-\alpha }\right) \), as \(\,r\rightarrow 1\) (see, e. g., [18, p. 54]). This and the fact that \(\,\mu \,\) is a Carleson measure imply that

$$\begin{aligned} \lim _{r\rightarrow 1}M_\infty ^p(r,f)(1-r)^{\alpha +1}\mu ([r,1))\,=\,0.\end{aligned}$$
(16)

Using again that \(\,\mu \,\) is a Carleson measure and integrating by parts we see that

$$\begin{aligned} \int _0^1g^\prime (r)\mu ([r, 1))\,dr\,&\lesssim \,\int _0^1g^\prime (r)(1-r)\,dr\\&=\,\lim _{r\rightarrow 1} g(r)(1-r)\,-g(0)\,+\,\int _0^1g(r)\,dr\\&\le \, \lim _{r\rightarrow 1}M_\infty ^p(r,f)(1-r)^{\alpha +2}\,+\,\int _0^1M_\infty ^p(r,f)(1-r)^{\alpha +1}\,dr \\&=\,\int _0^1M_\infty ^p(r,f)(1-r)^{\alpha +1}\,dr. \end{aligned}$$

Then, using [13, Lemma  3. (ii)], it follows that

$$\begin{aligned}\int _0^1g^\prime (r)\mu ([r, 1))\,dr\,\lesssim \,\Vert f\Vert _{A^p_\alpha }^p.\end{aligned}$$

Using this and (16) in (15) readily yields \(\,\int _{[0,1)}M_\infty ^p(r,f)(1-r)^{\alpha +1}\,d\mu (r)\,\lesssim \,\Vert f\Vert _{A^p_\alpha }^p\). \(\square \)

We shall also need the following characterization of the dual of the spaces \(\,A^q_\beta \) (\(q>1\)). It is a special case of [21, Theorem 2. 1].

Lemma 2

If \(1<q<\infty \) and \(\beta >-1\), then the dual of \(A^q_\beta \) can be identified with \(A^{p}_{\alpha }\) where \(\frac{1}{p}+\frac{1}{q}=1\) and \(\alpha \) is any number with \(\alpha >-1\), under the pairing

$$\begin{aligned} \langle h, f\rangle _{A_{q,\beta ,\alpha }}=\int _{\mathbb {D}}h(z)\overline{f(z)}(1-|z|^2)^{\frac{\beta }{q}+\frac{\alpha }{p}}\,dA(z),\quad h\in A^q_\beta ,\,\,\,\,f\in A^p_\alpha . \end{aligned}$$
(17)

Finally, we recall the following result from [13, (5. 2), p. 242] which is a version of the classical Hardy’s inequality [17, pp. 244–245].

Lemma 3

Suppose that \(\,k>0\), \(\,q>1\), and \(\,h\,\) is a non-negative function defined in \(\,(0,1)\), then

$$\begin{aligned} \int _0^1\left( \int _{1-r}^1 h(t)\, dt \right) ^q (1-r)^{k-1}\,dr \le \left( \frac{q}{k}\right) ^q \int _0^1 (h(1-r))^q (1-r)^{q+k-1}\,dr. \end{aligned}$$

Proof of Theorem 4

Suppose first that \(\,{\mathcal {H}}_\mu \,\) is a bounded operator from \(\,A^p_\alpha \,\) into itself. For \(0<b<1\), set

$$\begin{aligned} f_b(z)=\frac{(1-b^2)^{1-\frac{\alpha }{p}}}{(1-bz)^{\frac{2}{p}+1}},\quad z\in {\mathbb {D}}. \end{aligned}$$

Recall that \(\,p-\alpha >2\). Then using [29, Lemma 3. 10] with \(\,t=\alpha \,\) and \(\,c=p-\alpha \), we obtain

$$\begin{aligned} \Vert f_b\Vert _{A^p_\alpha }^p\,=\,(1-b^2)^{p-\alpha }\int _{{\mathbb {D}}}\frac{(1-\vert z\vert ^2)^\alpha }{\vert 1-bz\vert ^{2+p}}\,dA(z)\,\asymp \,1.\end{aligned}$$

Since \(\,{\mathcal {H}}_\mu \,\) is bounded on \(\,A^p_\alpha \), this implies

$$\begin{aligned} 1\,\gtrsim \Vert {\mathcal {H}}\mu (f_b)\Vert _{A^p_\alpha }.\end{aligned}$$
(18)

We also have

$$\begin{aligned} f_b(z)=\sum _{k=0}^\infty a_{k,b}z^k,\,\,\,(z\in {\mathbb {D}}),\quad \text {with}~ \,a_{k,b}\asymp (1-b^2)^{1-\frac{\alpha }{p}}k^{\frac{2}{p}}b^k. \end{aligned}$$

Since the \(a_{k,b}\)’s are positive, it is clear that the sequence \(\left\{ \sum _{k=0}^\infty \mu _{n+k}a_{k,b}\right\} _{n=0}^\infty \) of the Taylor coefficients of \({\mathcal {H}}_\mu (f_b)\) is a decreasing sequence of non-negative real numbers. Using this, Proposition 1, (18), and the definition of the \(a_{k,b}\)’s, we obtain

$$\begin{aligned} 1&\gtrsim \Vert {\mathcal {H}}_\mu (f_b)\Vert _{A^p_\alpha }^p\gtrsim \sum _{n=1}^\infty n^{p-\alpha -3}\left( \sum _{k=1}^\infty \mu _{n+k}a_{k,b}\right) ^p \\&= \sum _{n=1}^\infty n^{p-\alpha -3}\left( \sum _{k=1}^\infty a_{k,b}\int _{[0,1)} t^{n+k} d\mu (t)\right) ^p \\&\gtrsim (1-b^2)^{p-\alpha }\sum _{n=1}^\infty n^{p-\alpha -3}\left( \sum _{k=1}^\infty k^{\frac{2}{p}}b^{k}\int _{[b,1)} t^{n+k} d\mu (t)\right) ^p \\&\ge (1-b^2)^{p-\alpha }\sum _{n=1}^\infty n^{p-\alpha -3}\left( \sum _{k=1}^\infty k^{\frac{2}{p}}b^{n+2k}\mu ([b,1))\right) ^p \\&= (1-b^2)^{p-\alpha }\mu ([b,1))^p\sum _{n=1}^\infty n^{p-\alpha -3}b^{np}\left( \sum _{k=1}^\infty k^{\frac{2}{p}}b^{2k}\right) ^p \\&\asymp (1-b^2)^{p-\alpha }\mu ([b,1))^p\frac{1}{(1-b^2)^{2+p}}\sum _{n=1}^\infty n^{p-\alpha -3}b^{np} \\&\asymp (1-b^2)^{p-\alpha }\mu ([b,1))^p\frac{1}{(1-b^2)^{2+p}}\cdot \frac{1}{(1-b^2)^{p-\alpha -2}} \\&\asymp \mu ([b,1))^p\frac{1}{(1-b)^{p}}. \end{aligned}$$

Then it follows that

$$\begin{aligned} \mu \left( [b,1)\right) \,=\,{{\,\mathrm{O}\,}}\left( 1-b\right) ,\quad \text {as}~ b\rightarrow 1, \end{aligned}$$

and, hence, \(\mu \) is a Carleson measure.

We turn to prove the other implication. So, suppose that \(\,\mu \,\) is a Carleson measure and take \(\,f\in A^p_\alpha \). Let \(\,q\,\) be defined by the relation \(\,\frac{1}{p}+\frac{1}{q}=1\,\) and take \(\,\beta =\frac{-\alpha q}{p}=\frac{-\alpha }{p-1}\). Observe that \(\,\beta >-1\,\) and that with this election of \(\,\beta \,\) the weight in the pairing (17) is identically equal to 1. We have to show that \(\,{\mathcal {H}}_\mu f\in A^p_\alpha \,\) which is equal to \(\,\left( A^q_\beta \right) ^*\,\) under the pairing \(\,\langle \cdot ,\cdot \rangle _{q,\beta ,\alpha }\). So take \(\,h\in A^q_\beta \).

$$\begin{aligned} \langle h,{\mathcal {H}}_\mu f \rangle _{q,\beta ,\alpha }\,=&\,\int _{{\mathbb {D}}}h(z)\,\overline{{\mathcal {H}}_\mu f(z)}\,dA(z)\\ =&\, \int _{[0,1)}\overline{f(t)}\left( \int _{{\mathbb {D}}}\frac{h(z)}{1-t\,\overline{z}}\,dA(z)\right) \,d\mu (t)\\ =&\, \int _{[0,1)}\overline{f(t)}\left( \int _0^1\,\frac{r}{\pi }\,\int _0^{2\pi } \frac{h(re^{i\theta })}{1-tre^{-i\theta }}\,d\theta \,dr\right) \,d\mu (t) \\ =&\, \int _{[0,1)}\overline{f(t)}\left( \int _0^1\,\left( \frac{r}{\pi i}\,\int _{\vert \xi \vert =1} \frac{h(r\xi )}{\xi -tr}\,d\xi \right) \,dr\right) \,d\mu (t) \\ =&\, 2\int _{[0,1)}\overline{f(t)}\left( \int _0^1\,rh(r^2t)\,dr\right) \,d\mu (t). \end{aligned}$$

Thus,

$$\begin{aligned}\left| \langle h,{\mathcal {H}}_\mu f \rangle _{q,\beta ,\alpha }\right| \le 2\int _0^1\vert f(t)\vert G(t)\,d\mu (t),\end{aligned}$$

where \(G(t)=\int _0^1 r\vert h(r^2t)\vert \,dr\). Using Hölder’s inequality we obtain,

$$\begin{aligned}&\int _{[0,1)} f(t)G(t)\,d\mu (t)=\int _{[0,1)} |f(t)|(1-t)^{\frac{\alpha +1}{p}} G(t)(1-t)^{-\frac{\alpha +1}{p}}\,d\mu (t) \\&\quad \le \left( \int _{[0,1)} |f(t)|^p(1-t)^{\alpha +1}\,d\mu (t) \right) ^{1/p} \cdot \left( \int _{[0,1)} G(t)^q(1-t)^{-\frac{q(\alpha +1)}{p}}\,d\mu (t)\right) ^{1/q}. \end{aligned}$$

Lemma 1 implies that

$$\begin{aligned} \left( \int _{[0,1)} |f(t)|^p(1-t)^{\alpha +1}\,d\mu (t) \right) ^{1/p}\,\lesssim \,\Vert f\Vert _{A^p_\alpha }. \end{aligned}$$

Next we will show that

$$\begin{aligned} \int _{[0,1)} G(t)^q(1-t)^{-\frac{q(\alpha +1)}{p}}\,d\mu (t)\lesssim \Vert h\Vert ^q_{A^q_\beta }.\end{aligned}$$
(19)

This will give that

$$\begin{aligned} \left| \langle h,{\mathcal {H}}_\mu f \rangle _{q,\beta ,\alpha }\right| \,\lesssim \,\Vert f\Vert _{A^p_\alpha }\cdot \Vert h\Vert ^q_{A^q_\beta }. \end{aligned}$$

By the duality theorem, this implies that \(\,{\mathcal {H}}_\mu f\in A^p_\alpha \).

Let us prove (19). Observe first that if \(0<t<1/2\) then \(|h(r^2t)|\le M_\infty (\frac{1}{2},h)\) for each \(r\in (0,1)\), thus

$$\begin{aligned} G(t)=\int _0^1 |h(r^2t)|r\,dr\,\le M_\infty \left( \frac{1}{2},h\right) , \quad 0<t<1/2. \end{aligned}$$

Clearly, this implies

$$\begin{aligned} \int _{[0,1/2)} G(t)^q(1-t)^{-\frac{q(\alpha +1)}{p}}\,d\mu (t)\,\lesssim M_\infty ^q\left( \frac{1}{2},h\right) \,\lesssim \,\Vert h\Vert ^q_{A^q_\beta }. \end{aligned}$$
(20)

Notice that \(\,-\frac{q(\alpha +1)}{p}\,=\,\frac{p-2-\alpha }{p-1}-1\,>-1\). Making the change of variables \(\,r^2t=s\), we obtain \(\int _0^1 r|h(r^2t)|\,dr=\frac{1}{2t}\int _0^t |h(s)|\,ds\) and, hence,

$$\begin{aligned}&\int _{[1/2,1)} G(t)^q(1-t)^{-\frac{q(\alpha +1)}{p}}\,d\mu (t)\nonumber \\&\quad = \int _{[1/2,1)} \left( \int _0^1 |h(r^2t)|r\,dr\right) ^q(1-t)^{-\frac{q(\alpha +1)}{p}}\,d\mu (t)\nonumber \\&\quad =\int _{[1/2,1)} \frac{1}{(2t)^q} \left( \int _0^t |h(s)|\,ds\right) ^q(1-t)^{-\frac{q(\alpha +1)}{p}}\,d\mu (t)\nonumber \\&\quad \le \int _{[1/2,1)} \left( \int _0^t M_\infty (s,h)\,ds\right) ^q(1-t)^{-\frac{q(\alpha +1)}{p}}\,d\mu (t)\nonumber \\&\quad \le \int _{[0,1)} \left( \int _{1-t}^1 M_\infty (1-s,h)\,ds\right) ^q(1-t)^{-\frac{q(\alpha +1)}{p}}\,d\mu (t) \end{aligned}$$
(21)

Let us call \(H(t)= \left( \int _{1-t}^1 M_\infty (1-s,h)\,ds\right) ^q(1-t)^{-\frac{q(\alpha +1)}{p}}\) for \(0\le t <1\). Integrating by parts we obtain the following

$$\begin{aligned} \int _{[0,1)} H(t)\,d\mu (t)=H(0)\mu ([0,1))-\lim _{t\rightarrow 1^-} H(t)\mu ([t,1))+ \int _0^1 \mu ([t,1))H^\prime (t)\,dt. \end{aligned}$$
(22)

The first term is equal to 0. Using the fact that \(\mu \) is a Carleson measure we have that

$$\begin{aligned} H(t)\mu ([t,1))&\lesssim (1-t)H(t) \\&= \left( \int _{1-t}^1 M_\infty (1-s,h)\,ds\right) ^q(1-t)^{1-\frac{q(\alpha +1)}{p}} \\&=\left( \int _{0}^t M_\infty (s,h)\,ds\right) ^q(1-t)^{1-\frac{q(\alpha +1)}{p}} . \end{aligned}$$

Since \(\,h\in A^q_\beta \,\) we have \(\,M_\infty (t,h)=o\left( (1-t)^{-\frac{\beta +2}{q}}\right) \), as \(\,t\rightarrow 1\). Then, bearing in mind that \(\,\frac{\beta +2}{q}>1\), it follows that

$$\begin{aligned} H(t)\mu ([t,1))\,=\,{{\,\mathrm{o}\,}}\left( (1-t)^{-\beta -2+q}\cdot (1-t)^{1-\frac{q(\alpha +1)}{p}}\right) ={{\,\mathrm{o}\,}}(1),\quad \text {as}~ \,t\rightarrow 1.\end{aligned}$$
(23)

Actually, we have also proved that

$$\begin{aligned} (1-t)H(t)\,=\,{{\,\mathrm{o}\,}}(1),\quad \text {as}~ \,t\rightarrow 1.\end{aligned}$$
(24)

Using that \(\mu \) is a Carleson measure, integrating by parts, and using the definition of \(\,H\,\) and (24), we obtain

$$\begin{aligned} \int _0^1 \mu ([t,1))H^\prime (t)\,dt&\lesssim \int _0^1 (1-t)H^\prime (t)\,dt\nonumber \\&= \lim _{t\rightarrow 1}(1-t)H(t)\,-\,H(0)\,+\,\int _0^1 H(t)\,dt \nonumber \\&= \int _0^1 \left( \int _{1-t}^1 M_\infty (1-s,h)\,ds\right) ^q(1-t)^{-\frac{q(\alpha +1)}{p}}\,dt. \end{aligned}$$
(25)

Now, using Lemma 3 and [13, Lemma 3], we see that

$$\begin{aligned}&\int _0^1 \left( \int _{1-t}^1 M_\infty (1-s,h)\,ds\right) ^q(1-t)^{-\frac{q(\alpha +1)}{p}}\,dt\\&\quad \lesssim \int _0^1 M^q_\infty (t,h)(1-t)^{\alpha +1}\,dt \lesssim \Vert h\Vert ^q_{A^q_\beta }. \end{aligned}$$

Using this, (25), (23), (22), and (21), it follows that

$$\begin{aligned} \int _{[1/2,1)} G(t)^q(1-t)^{-\frac{q(\alpha +1)}{p}}\,d\mu (t)\,\lesssim \Vert h\Vert ^q_{A^q_\beta }. \end{aligned}$$

This and (20) yield (19). \(\square \)

Our final aim in this article is to find the analogue of Theorem 4 for Dirichlet spaces. In other words, we wish give an answer to the following question.

Question 2

If \(\,\max (-1, p-2)<\alpha <2p-2\), is it true that \(\,{\mathcal {H}}_\mu \,\) is a bounded operator from \(\,{{\mathcal {D}}^p_{\alpha }}\,\) into itself if and only if \(\,\mu \,\) is a Carleson measure?

Since \(\,p-1<\alpha <2p-2\,\) implies that \(\,{{\mathcal {D}}^p_{\alpha }}\,=\,A^p_{\alpha -p},\)    Theorem 4 answers the question affirmatively for these values of \(\,p\,\) and \(\,\alpha \). It remains to consider the case \(\,\max (-1, p-2)<\alpha \le p-1\). We shall prove the following result which gives a positive answer to Question 2 in the case \(\,p>1\).

Theorem 5

Suppose that \(\,p>1\,\) and \(\,p-2<\alpha \le p-1\), and let \(\,\mu \,\) be a finite positive Borel measure on \(\,[0,1)\).

The operator \(\,{\mathcal {H}}_\mu \,\) is well defined on \(\,{{\mathcal {D}}^p_{\alpha }}\,\) and it is a bounded operator from \(\,{{\mathcal {D}}^p_{\alpha }}\,\) into itself if and only if \(\,\mu \,\) is a Carleson measure.

The following two lemmas will be needed in the proof of Theorem 5. The first one follows trivially from Proposition 1.

Lemma 4

Let \(\,f\in {\mathcal {H}}ol({\mathbb {D}})\), \(f(z)=\sum _{n=0}^\infty a_nz^n\,\)\(\,(z\in {\mathbb {D}})\). Suppose that \(1\,<p<\infty \,\) and \(\,p-2<\alpha \le p-1\), and that the sequence \(\,\{ a_n\} _{n=0}^\infty \,\) is a decreasing sequence of non-negative real numbers. Then

$$\begin{aligned} f\in {{\mathcal {D}}^p_{\alpha }}\,\,\,\Leftrightarrow \,\,\,\sum _{n=0}^\infty (n+1)^{2p-\alpha -3}a_n^p<\infty . \end{aligned}$$

The following lemma is a generalization of [13, Lemma 4].

Lemma 5

Let \(\,\mu \,\) be a positive Borel measure on \(\,[0,1)\,\) which is a Carleson measure. Assume that \(\,0<p<\infty \,\) and \(\,\alpha >-1\). Then there exists a positive constant \(\,C=C(p,\alpha ,\mu )\,\) such that for any \(\,f\in {{\mathcal {D}}^p_{\alpha }}\,\)

$$\begin{aligned} \int _{[0,1)}M_\infty ^p(r,f)(1-r)^{\alpha -p+1}\,d\mu (r)\,\le C\Vert f\Vert _{{{\mathcal {D}}^p_{\alpha }}}^p. \end{aligned}$$

Proof

We argue as in the proof of Lemma 1. Take \(\,f\in {{\mathcal {D}}^p_{\alpha }}\,\) and set

$$\begin{aligned}&g(r)=M_\infty ^p(r,f)(1-r)^{\alpha -p+1},\,\, \\&\quad F(r)=\mu ([0,r))\,-\mu ([0,1))\,=\,-\mu ([r,1)),\,\, 0<r<1. \end{aligned}$$

Integrating by parts, we have

$$\begin{aligned}&\int _{[0,1)}M_\infty ^p(r,f)(1-r)^{\alpha -p+1}\,d\mu (r)\,=\,\int _{[0,1)}g(r)\,d\mu (r)\nonumber \\&\quad =\,\lim _{r\rightarrow 1}g(r)F(r)\,-\,g(0)F(0)\,-\int _0^1g^\prime (r)F(r)\,dr \nonumber \\&\quad =\,\vert f(0)\vert ^p\mu ([0,1))\,-\,\lim _{r\rightarrow 1}M_\infty ^p(r,f)(1-r)^{\alpha -p+1}\mu ([r,1))\,\nonumber \\&\qquad +\,\int _0^1g^\prime (r)\mu ([r, 1))\,dr. \end{aligned}$$
(26)

Since \(\,f\in {{\mathcal {D}}^p_{\alpha }}\,\) we have that \(\,M_\infty ^p(r,f^\prime )={{\,\mathrm{o}\,}}\left( (1-r)^{-2-\alpha }\right) \), as \(\,r\rightarrow 1\). Hence, \(\,M_\infty ^p(r,f)={{\,\mathrm{o}\,}}\left( (1-r)^{-2-\alpha +p}\right) \), as \(\,r\rightarrow 1\). This and the fact that \(\,\mu \,\) is a Carleson measure imply that

$$\begin{aligned} \lim _{r\rightarrow 1}M_\infty ^p(r,f)(1-r)^{\alpha -p+1}\mu ([r,1))\,=\,0.\end{aligned}$$
(27)

Using again that \(\,\mu \,\) is a Carleson measure and integrating by parts we see that

$$\begin{aligned} \int _0^1g^\prime (r)\mu ([r, 1))\,dr\,\lesssim&\int _0^1g^\prime (r)(1-r)\,dr\\ =&\lim _{r\rightarrow 1}g(r)(1-r)\,-g(0)\,+\,\int _0^1g(r)\,dr\\ \,\le \,&\lim _{r\rightarrow 1}M_\infty ^p(r,f)(1-r)^{\alpha -p +2}\,\\&+\int _0^1M_\infty ^p(r,f)(1-r)^{\alpha -p+1}\,dr \\ =&\int _0^1M_\infty ^p(r,f)(1-r)^{\alpha -p+1}\,dr. \end{aligned}$$

Then, using [13, Lemma  3], it follows that

$$\begin{aligned}\int _0^1g^\prime (r)\mu ([r, 1))\,dr\,\lesssim \,\Vert f\Vert _{{{\mathcal {D}}^p_{\alpha }}}^p.\end{aligned}$$

Using this and (27) in (26) readily yields \(\,\int _{[0,1)}M_\infty ^p(r,f)(1-r)^{\alpha -p+1}\,d\mu (r)\,\lesssim \,\Vert f\Vert _{{{\mathcal {D}}^p_{\alpha }}}^p\). \(\square \)

Proof of Theorem 5

Suppose first that \(\,{\mathcal {H}}_\mu \,\) is a bounded operator from \(\,{{\mathcal {D}}^p_{\alpha }}\,\) into itself. For \(\,1/2<b<1\,\) we set

$$\begin{aligned} f_b(z)=\frac{(1-b^2)^{1-\frac{\alpha }{p}}}{(1-bz)^{2/p}},\quad z\in {\mathbb {D}}. \end{aligned}$$

We have \(\,\Vert f_b\Vert _{{{\mathcal {D}}^p_{\alpha }}}\asymp 1\). Then arguing as in the proof of the correspondent implication in Theorem 4 we obtain that \(\,\mu \,\) is a Carleson measure. We omit the details.

To prove the other implication, suppose that \(\,\mu \,\) is a Carleson measure and take \(\,f\in {{\mathcal {D}}^p_{\alpha }}\). Since \(\,{\mathcal {H}}_ \mu \,\) and \(\,{\mathcal {I}}_\mu \,\) coincide on \(\,{{\mathcal {D}}^p_{\alpha }}\), we have to prove that \(\,{\mathcal {I}}_\mu f\in {{\mathcal {D}}^p_{\alpha }}\,\) and that \(\,\Vert {\mathcal {I}}_\mu f\Vert _ {{\mathcal {D}}^p_{\alpha }}\lesssim \Vert f\Vert _ {{\mathcal {D}}^p_{\alpha }}\,\) or, equivalently, that \(\,\left( {\mathcal {I}}_\mu f\right) ^\prime \in A^p_\alpha \,\) and

$$\begin{aligned} \,\Vert \left( {\mathcal {I}}_\mu f\right) ^\prime \Vert _ {A^p_\alpha }\lesssim \Vert f\Vert _ {A^p_\alpha }.\end{aligned}$$
(28)

We shall distinguish two cases.

First case: \({\varvec{\alpha <p-1}}\). Let \(\,q\,\) be defined by the relation \(\,\frac{1}{p}+\frac{1}{q}=1\,\) and take \(\,\beta =\frac{-\alpha q}{p}\). In view of Lemma 2, (28) is equivalent to

$$\begin{aligned} \left| \int _{{\mathbb {D}}}\,h(z)\,\overline{\left( {\mathcal {I}}_\mu f\right) ^\prime (z)}\,dA(z)\right| \,\lesssim \Vert f\Vert _{{{\mathcal {D}}^p_{\alpha }}}\Vert h\Vert _{A^q_\beta },\quad h\in A^q_\beta .\end{aligned}$$
(29)

So, take \(\,h\in A^q_\beta \). Just as in the proof of Theorem 1, we have

$$\begin{aligned} \int _{{\mathbb {D}}}\,h(z)\,\overline{\left( {\mathcal {I}}_\mu f\right) ^\prime (z)}\,dA(z)\,=\,\int _{[0,1)}\,t\,\overline{f(t)}\,h(t)\,d\mu (t).\end{aligned}$$
(30)

Set \(\,s=-1 + \frac{\alpha +1}{p}\). Observe that \(\,ps=\alpha -p+1\,\) and \(\,-qs=\beta +1\). Then, using (30), Hölder’s inequality, Lemma 1, and Lemma 5, we obtain

$$\begin{aligned}&\left| \int _{{\mathbb {D}}}\,h(z)\,\overline{\left( {\mathcal {I}}_\mu f\right) ^\prime (z)}\,dA(z)\right| \,\le \, \int _{[0,1)}\vert f(t)\vert (1-t)^s\,\vert h(t)\vert (1-t)^{-s}\,d\mu (t)\\&\quad \le \,\left( \int _{{\mathbb {D}}}\vert f(t)\vert ^p(1-t)^{\alpha -p+1}\,d\mu (t)\right) ^{1/p}\left( \int _{[0,1)}\vert h(t)\vert ^q(1-t)^{\beta +1}\,d\mu (t)\right) ^{1/q} \\&\quad \le \,\left( \int _{{\mathbb {D}}}M_\infty ^p(t,f)(1-t)^{\alpha -p+1}\,d\mu (t)\right) ^{1/p}\\&\qquad \times \left( \int _{[0,1)}M_\infty ^q(t,h)(1-t)^{\beta +1}\,d\mu (t)\right) ^{1/q} \\&\quad \le \,\Vert f\Vert _{{{\mathcal {D}}^p_{\alpha }}}\Vert h\Vert _{A^q_\beta }. \end{aligned}$$

Thus, (29) holds.

Second case: \({\varvec{\alpha =p-1}}\). We let again \(\,q\,\) be defined by the relation \(\,\frac{1}{p}+\frac{1}{q}=1\,\) and take \(\,\beta =q-1\). Using Lemma 2 and arguing as in the preceding case, we have to show that

$$\begin{aligned} \left| \int _{{\mathbb {D}}}\, (1-\vert z\vert ^2)\,h(z)\,\overline{\left( {\mathcal {I}}_\mu f\right) ^\prime (z)}\,dA(z)\right| \,\lesssim \,\Vert f\Vert _{{{\mathcal {D}}^p_{p-1}}}\Vert h\Vert _{A^q_{q-1}},\quad h\in A^q_{q-1}.\end{aligned}$$
(31)

We have

$$\begin{aligned} \int _{{\mathbb {D}}}\, (1-\vert z\vert ^2)\,h(z)\,\overline{\left( {\mathcal {I}}_\mu f\right) ^\prime (z)}\,dA(z)\,=\,\int _{[0,1)}\,t\,\overline{f(t)}\,\int _{{\mathbb {D}}}\frac{(1-\vert z\vert ^2)h(z)}{(1-t\,\overline{z})^2}\,dA(z)\,d\mu (t).\end{aligned}$$
(32)

Now, \(\int _{{\mathbb {D}}}\frac{h(z)}{(1-t\,\overline{z})^2}\,dA(z)\,=\,h(t)\) and

$$\begin{aligned}&\int _{{\mathbb {D}}}\frac{\vert z\vert ^2\,h(z)}{(1-t\,\overline{z})^2}\,dA(z)\,=\,\int _0^1\,\frac{r^3}{\pi }\int _0^{2\pi }\frac{h(re^{i\theta })\,d\theta }{(1-tre^{-i\theta })^2}\,dr\\&\quad =\, \int _0^1\,\frac{2r^3}{2\pi i}\int _0^{2\pi }\frac{e^{i\theta }h(re^{i\theta })ie^{i\theta }\,d\theta }{(e^{i\theta }-tr)^2}\,dr\,=\,\int _0^1\,\frac{2r^3}{2\pi i}\int _{\vert z\vert =1}\frac{zh(rz)}{(z-tr)^2}\,dz\,dr\\&\quad =\,\int _0^12r^3\left[ h(r^2t)\,+\,r^2th^\prime (r^2t)\right] \,dr. \end{aligned}$$

Then it is clear that \(\left| \int _{{\mathbb {D}}}\frac{(1-\vert z\vert ^2)\,h(z)}{(1-t\,\overline{z})^2}\,dA(z)\right| \,\lesssim M_\infty (t,h)\). Using this, (32), Hölder’s inequality, Lemma 1, and Lemma 5, we obtain

$$\begin{aligned}&\left| \int _{{\mathbb {D}}}\, (1-\vert z\vert ^2)\,h(z)\,\overline{\left( {\mathcal {I}}_\mu f\right) ^\prime (z)}\,dA(z)\right| \,\lesssim \,\int _{[0,1)}M_\infty (t,f)\,M_\infty (t,h)\,d\mu (t)\\&\quad \le \,\left( \int _{[0,1)}M_\infty ^p(t,f)\,d\mu (t)\right) ^{1/p} \,\left( \int _{[0,1)}M_\infty ^q(t,h)\,d\mu (t)\right) ^{1/q} \,\\&\quad \le \,\Vert f\Vert _{{{\mathcal {D}}^p_{p-1}}}\Vert h\Vert _{A^q_{q-1}}. \end{aligned}$$

This is (31). \(\square \)

We shall close the article with some comments about the case \(\,p=1\,\) in Question 2. We have the following result.

Theorem 6

Let \(\,\mu \,\) be a finite positive Borel measure on \(\,[0,1)\,\) and \(\,-1<\alpha <0\). If \(\,\mu \,\) is a Carleson measure then the operator \(\,{\mathcal {H}}_\mu \,\) is a bounded operator form \(\,{\mathcal {D}}^1_\alpha \,\) to itself.

Proof

Using [29, Theorem 5. 15, p. 113], we see that \(\,A^1_\alpha \,\) can be identified as the dual of the little Bloch space under the pairing

$$\begin{aligned} \langle h, g\rangle \,=\,\int _{{\mathbb {D}}}(1-\vert z\vert ^2)^\alpha \,h(z)\,\overline{g(z)}\,dA(z),\quad h\in {\mathcal {B}}_0,\,\,\,g\in A^1_\alpha .\end{aligned}$$
(33)

Suppose that \(\,\mu \,\) is a Carleson measure. Using this duality relation and the fact that \(\,{\mathcal {H}}_\mu ={\mathcal {I}}_\mu \,\) on \(\,{\mathcal {D}}^1_\alpha \), showing that \(\,{\mathcal {H}}_\mu \,\) is a bounded operator from \(\,{\mathcal {D}}^1_\alpha \,\) to itself is equivalent to showing that

$$\begin{aligned} \left| \int _{{\mathbb {D}}}(1-\vert z\vert ^2)^{\alpha }\,h(z)\,\overline{\left( {\mathcal {I}}_\mu f\right) ^\prime (z)}\,dA(z)\right| \,\lesssim \Vert h\Vert _{{\mathcal {B}}}\cdot \Vert f\Vert _{{\mathcal {D}}^1_\alpha },\quad h\in {\mathcal {B}}_0,\,\,\,f\in {\mathcal {D}}^1_\alpha .\end{aligned}$$
(34)

Let us prove (34). Take \(\,h\in {\mathcal {B}}_0\,\) and \(\,f\in {\mathcal {D}}^1_\alpha \). We have

$$\begin{aligned}&\int _{{\mathbb {D}}}(1-\vert z\vert ^2)^{\alpha }\,h(z)\,\overline{\left( {\mathcal {I}}_\mu f\right) ^\prime (z)}\,dA(z)\,\nonumber \\&\quad =\,\int _{[0,1)}t\,\overline{f(t)}\,\int _{{\mathbb {D}}}\frac{(1-\vert z\vert ^2)^\alpha h(z)}{(1-t\,\overline{z})^2}\,dA(z)\,d\mu (t). \end{aligned}$$
(35)

Using [29, Lemma 5. 14, pp. 113-114] we have that the operator \(\,T\,\) defined by

$$\begin{aligned} T\phi (\xi )\,=\,(1-\vert \xi \vert ^2)^{-\alpha }\int _{{\mathbb {D}}}\frac{(1-\vert z\vert ^2)^\alpha \phi (z)}{(1-\xi \,\overline{z})^2}\,dA(z) \end{aligned}$$

is a bounded operator from \(\,{\mathcal {B}}\,\) into \(\,L^\infty ({\mathbb {D}})\). Then it follows that

$$\begin{aligned} \left| \int _{{\mathbb {D}}}\frac{(1-\vert z\vert ^2)^\alpha h(z)}{(1-t\,\overline{z})^2}\,dA(z)\right| \,\lesssim \,\Vert h\Vert _{{\mathcal {B}}}(1-t^2)^\alpha ,\quad t\in [0,1). \end{aligned}$$

Using this in (35), we obtain

$$\begin{aligned} \left| \int _{{\mathbb {D}}}(1-\vert z\vert ^2)^{\alpha }\,h(z)\,\overline{\left( {\mathcal {I}}_\mu f\right) ^\prime (z)}\,dA(z)\right| \,\lesssim \Vert h\Vert _{{\mathcal {B}}}\int _{{\mathbb {D}}}(1-t)^{\alpha }\vert f(t)\vert \,d\mu (t).\end{aligned}$$
(36)

The fact that \(\,\mu \,\) is a Carleson measure readily implies that the measure \(\,\nu \,\) defined by \(\,d\nu (t)=(1-t)^\alpha \,d\mu (t)\,\) is a \(\,(1-\alpha )\)-Carleson measure. Using Theorem 1 of [28] we see that then \(\,\nu \,\) is a Carleson measure for \(\,{\mathcal {D}}^1_\alpha \), that is,

$$\begin{aligned} \int _{[0,1)}(1-t)^\alpha \vert g(t)\vert \,d\mu (t)\,\lesssim \,\Vert g\Vert _{{\mathcal {D}}^1_\alpha },\quad g\in {\mathcal {D}}^1_\alpha . \end{aligned}$$

Using this in (36), (34) follows. \(\square \)

We do not know whether the converse of Theorem 6 is true. This is due to the fact that we do not know whether Lemma 4 remains true for \(\,p=1\). The inequality

$$\begin{aligned} \sum _{n=0}^\infty \vert a_n\vert (n+1)^{-(1+\alpha )}\lesssim \Vert f\Vert _{{\mathcal {D}}^1_\alpha }.\end{aligned}$$
(37)

is certainly true with no assumption on the sequence \(\,\{ a_n\} \). Indeed, by Hardy’s inequality [11, p. 48], \(\sum _{n=1}^\infty \vert a_n\vert r^{n-1}\lesssim \int _0^{2\pi }\vert f^\prime (re^{i\theta })\vert d\theta \). Hence

$$\begin{aligned}&\Vert f\Vert _{{\mathcal {D}}^1_\alpha }\asymp \int _0^1(1-r)^\alpha \int _0^{2\pi }\vert f^\prime (re^{i\theta })\vert d\theta dr \\&\quad \gtrsim \sum _{n=1}^\infty \vert a_n\vert \int _0^1(1-r)^\alpha r^{n-1}\,dr = \sum _{n=1}^\infty \vert a_n\vert B(\alpha +1, n), \end{aligned}$$

where \(\,B(\cdot ,\cdot )\,\) is the Beta function. Stirling’s formula gives \(\,B(\alpha +1, n)\asymp n^{-(\alpha +1)}\) and then (37) follows.

However, the proof of Theorem D in [23] does not seen to work to prove the opposite inequality when \(\{a_n\} \) is decreasing.