1 Introduction

The open unit disc in the complex plane \({\mathbb {C}}\) will be denoted by \({\mathbb {D}}\) and \(\textrm{Hol}({\mathbb {D}})\) will stand for the space of all analytic functions in \({\mathbb {D}}\). Also, dA will denote the area measure on \({\mathbb {D}}\), normalized so that the area of \(\mathbb D\) is 1. Thus \(dA(z) = \frac{1}{\pi } dx dy = \frac{1}{\pi } r dr d\theta \).

For \(0\,\le \,r\,<\,1\), \(0<p\le \infty \), and f analytic in \({\mathbb {D}} \), the integral means \(M_p(r, f)\) of f are defined by

$$\begin{aligned} M_ p(r, f)= & {} \left( \frac{1}{2\pi }\int _ {0}^{2\pi }\left| f(re^{i\theta }) \right| ^pd\theta \right) ^{1/p}, \quad 0<p<\infty ,\\ M_ \infty (r, f)= & {} \max _ {\vert z\vert =r}\vert f(z)\vert . \end{aligned}$$

For \(0<p\le \infty \) the Hardy space \(H^p\) consists of those functions f, analytic in \({\mathbb {D}} \), for which

$$\begin{aligned} \left| \left| f \right| \right| _ {H^p}\,{\mathop {=}\limits ^{\text {def}}}\,\sup _ {0<r<1} M_ p(r, f)<\infty . \end{aligned}$$

We refer to [20] for the theory of Hardy spaces.

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

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

The unweighted Bergman space \(A^p_0\) is simply denoted by \(A^p\). We refer to [21, 31, 48] for the notation and results about Bergman spaces.

The space of Dirichlet type \({\mathcal {D}}^p_\alpha \) (\(0<p<\infty \), \(\alpha >-1\)) is the space of those \(f\in \textrm{Hol}({\mathbb {D}})\) such that \(f^\prime \in A^p_{\alpha }\). Thus, a function \(f\in \textrm{Hol}({\mathbb {D}})\) belongs to \({\mathcal {D}}^p_\alpha \) if and only if

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

In this paper we shall be mainly concerned with the Dirichlet space \({\mathcal {D}}={\mathcal {D}}^2_0\) which consists of those \(f\in \textrm{Hol}({\mathbb {D}})\) whose image Riemann surface has a finite area. We recall that if \(f\in {\mathcal {D}}\), \(f(z)=\sum _{n=0}^\infty a_nz^n\) (\(z\in {\mathcal {D}}\)), then

$$\begin{aligned} \Vert f\Vert _{{\mathcal {D}}}\,{\mathop {=}\limits ^{\text {def}}}\Vert f\Vert _{\mathcal D^2_0}=\vert f(0)\vert + \!\left( \int _{{\mathbb {D}}}\vert f^\prime (z)\vert ^2\,dA(z)\right) ^{1/2}=\,\vert a_0\vert \!+\!\left( \sum _{k=1}^\infty k\vert a_k\vert ^2\right) ^{1/2}.\qquad \quad \end{aligned}$$
(1.1)

Throughout the paper \(\mu \) will be a positive finite Borel measure on the radius [0, 1) and, for \(n=0, 1, 2, \dots \), we shall let \(\mu _n\) denote the moment of order n of \(\mu \), that is, \(\mu _n=\int _{[0, 1)}t^n\,d\mu (t)\). The matrices \({\mathcal {H}}_\mu \) and \({\mathcal {C}}_\mu \) are defined as follows

$$\begin{aligned}{\mathcal {H}}_\mu =\,\left( \begin{array}{ccccc} \mu _0 &{} \mu _1 &{} \mu _2 &{} . &{} . \\ \mu _1 &{} \mu _2 &{} \mu _3 &{} . &{} . \\ \mu _2 &{} \mu _3 &{} \mu _4 &{} . &{} . \\ . &{} . &{} . &{} . &{} . \\ . &{} . &{} . &{} . &{} . \\ \end{array} \right) ;&\,\,\,\,\,\, {\mathcal {C}}_{\mu }\,=\,\left( \begin{array}{cccccc} \mu _0 &{} 0 &{} 0 &{} 0 &{} . &{} . \\ \mu _1 &{} \mu _1 &{} 0 &{} 0 &{} . &{} . \\ \mu _2 &{} \mu _2 &{} \mu _2 &{} 0 &{} . &{} . \\ . &{} . &{} . &{} . &{} . &{} . \\ . &{} . &{} . &{} . &{} . &{} . \\ \end{array}\right) . \end{aligned}$$

As we shall see in Sects. 2 and  3, these matrices induce operators acting on spaces of analytic functions which are natural generalizations of the classical Hilbert and Cesàro operators. Recently a good amount of work has been devoted to study the action of these operators of Hilbert type and of Cesàro type on distinct subspaces of \(\textrm{Hol}({\mathbb {D}})\). Carleson-type measures play a basic role in this work.

Let us recall that if \(\mu \) is a positive finite Borel measure on [0, 1) then:

  • If \(s>0\), then \(\mu \) is said to be an s-Carleson measure if there exists a positive constant C such that

    $$\begin{aligned} \mu \left( [t, 1)\right) \,\le C(1-t)^{s},\quad 0\le t<1. \end{aligned}$$
  • If \(0\le \alpha <\infty \), and \(0<s<\infty \) we say that \(\mu \) is an \(\alpha \)-logarithmic s-Carleson measure if there exists a positive constant C such that

    $$\begin{aligned} \mu \left( [t, 1)\right) \le C(1-t)^s\left( \log \frac{2}{1-t}\right) ^{-\alpha }, \quad 0\le t<1. \end{aligned}$$

Let us close this section by saying that, as usual, we shall be using the convention that \(C=C(p, \alpha ,q,\beta , \dots )\) will denote a positive constant which depends only upon the displayed parameters \(p, \alpha , q, \beta \dots \) (which sometimes will be omitted) but not necessarily the same at different occurrences. Furthermore, for two real-valued functions \(K_1, K_2\) we write \(K_1\lesssim K_2\), or \(K_1\gtrsim K_2\), if there exists a positive constant C independent of the arguments such that \(K_1\le C K_2\), respectively \(K_1\ge C K_2\). If we have \(K_1\lesssim K_2\) and \(K_1\gtrsim K_2\) simultaneously, then we say that \(K_1\) and \(K_2\) are equivalent and we write \(K_1\asymp K_2\).

2 Hilbert-Type Operators

The matrix \({\mathcal {H}}_\mu \) induces formally an operator, which will be also called \({\mathcal {H}}_\mu \), on spaces of analytic functions by its action on the Taylor coefficients:

$$\begin{aligned} a_n\mapsto \sum _{k=0}^{\infty } \mu _{n+k}{a_k}, \quad n=0,1,2, \dots . \end{aligned}$$

To be precise, if  \(f(z)=\sum _{k=0}^\infty a_kz^k\in \textrm{Hol}({\mathbb {D}})\) we define

$$\begin{aligned} {\mathcal {H}}_\mu (f)(z)= \sum _{n=0}^{\infty }\left( \sum _{k=0}^{\infty } \mu _{n+k}{a_k}\right) z^n, \end{aligned}$$
(2.1)

whenever the right hand side makes sense and defines an analytic function in \({\mathbb {D}}\).

If \(\mu \) is the Lebesgue measure on [0, 1) the matrix \({\mathcal {H}}_\mu \) reduces to the classical Hilbert matrix   \({\mathcal {H}}= \left( {(n+k+1)^{-1}}\right) _{n,k\ge 0}\), which induces the classical Hilbert operator \({\mathcal {H}}\) which has extensively studied recently (see [1, 16, 17, 19, 32,33,34]).

The finite positive Borel measures \(\mu \) for which \({\mathcal {H}}_\mu \) is a bounded operator on distinct spaces of analytic functions in \({\mathbb {D}}\) have been characterized in a number of papers such as [9, 14, 25, 27,28,29, 35, 37, 38, 45]. Obtaining an integral representation of \({\mathcal {H}}_\mu \) plays a basic role in these works. If \(\mu \) is as above, we shall write throughout the paper

$$\begin{aligned} {\mathcal {I}}_\mu (f)(z)=\int _{[0,1)}\frac{f(t)}{1-tz}\,d\mu (t), \end{aligned}$$
(2.2)

whenever the right hand side makes sense and defines an analytic function in \({\mathbb {D}}\). It turns out that the operators \({\mathcal {H}}_\mu \) and \({\mathcal {I}}_\mu \) are very closely related.

Let us mention the following results.

Theorem A

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

  1. (i)

    The operator \({\mathcal {H}}_\mu \) is bounded from \(H^1\) into itself if and only if \(\mu \) is a 1-logarithmic 1-Carleson measure. In such a case \({\mathcal {H}}_\mu \) and \({\mathcal {I}}_\mu \) coincide on \(H^1\).

  2. (ii)

    If \(1<p<\infty \), then \({\mathcal {H}}_\mu \) is a bounded operator from \(H^p\) into itself if and only if \(\mu \) is a 1-Carleson measure. In such a case \({\mathcal {H}}_\mu \) and \({\mathcal {I}}_\mu \) coincide on \(H^p\).

  3. (iii)

    If \(p>1\) and \(-1<\alpha <\,p-2\) then the operator \(\mathcal H_\mu \) is well defined on \(A^p_\alpha \) and it is bounded from \(A^p_\alpha \) into itself if and only if \(\mu \) is a 1-Carleson measure. In such a case \({\mathcal {H}}_\mu \) and \({\mathcal {I}}_\mu \) coincide on \(A^p_\alpha \).

  4. (iv)

    If \(p>1\) and \(p-2<\alpha \le p-1\), then \(\mathcal H_\mu \) is well defined on \({\mathcal {D}}^p_\alpha \) and it is bounded from \({\mathcal {D}}^p_\alpha \) into itself if and only if \(\mu \) is a 1-Carleson measure. In such a case \({\mathcal {H}}_\mu \) and \({\mathcal {I}}_\mu \) coincide on \({\mathcal {D}}^p_\alpha \).

  5. (v)

    If \(0<\alpha <2\), \({\mathcal {H}}_\mu \) is a bounded operator from \({\mathcal {D}}^2_\alpha \) into itself if and only if \(\mu \) is a 1-Carleson measure. In such a case \({\mathcal {H}}_\mu \) and \({\mathcal {I}}_\mu \) coincide on \(\mathcal D^2_\alpha \).

The questions of characterizing those \(\mu \) for which \(\mathcal H_\mu \) is bounded on either the Dirichlet space \({\mathcal {D}}\) or on the Bergman space \(A^2\) are more delicate and remain open. Regarding the Dirichlet space, the following results are proved in [28].

Theorem B

  1. (i)

    Let \(\mu \) be a positive and finite Borel measure on [0, 1). If \(\gamma >1\) and \(\mu \) is a \(\gamma \)-logarithmic 1-Carleson measure, then \({\mathcal {H}}_\mu \) is bounded from \({\mathcal {D}}\) into itself.

  2. (ii)

    If \(0<\beta \le \frac{1}{2}\), then there exists a positive and finite Borel measure \(\mu \) on [0, 1) which is a \(\beta \)-logarithmic 1-Carleson measure but such that \({\mathcal {H}}_\mu (\mathcal D)\not \subset {\mathcal {D}}\).

We improve this result showing that being a 1-logarithmic 1-Carleson measure is enough to insure that \({\mathcal {H}}_\mu \) is bounded from \({\mathcal {D}}\) into itself and closing the gap between (i) and (ii). Indeed, we shall prove the following result.

Theorem 1

  1. (i)

    Let \(\mu \) be a positive and finite Borel measure on [0, 1). If \(\mu \) is a 1-logarithmic 1-Carleson measure, then \({\mathcal {H}}_\mu \) is bounded from \({\mathcal {D}}\) into itself.

  2. (ii)

    If \(0<\beta <1\), then there exists a positive and finite Borel measure \(\mu \) on [0, 1) which is a \(\beta \)-logarithmic 1-Carleson measure but such that \({\mathcal {H}}_\mu (\mathcal D)\not \subset {\mathcal {D}}\).

As a corollary of part (i) we obtain the following.

Corollary 2

  1. (a)

    Let \(\mu \) be a positive and finite Borel measure on [0, 1) and suppose that \(\mu \) is a 1-logarithmic 1-Carleson measure. Then there exists a positive constant C such that

    $$\begin{aligned} \int _{[0, 1)}\vert tf(t)f^\prime (t)\vert \,d\mu (t)\,\le C\Vert f\Vert _{{\mathcal {D}}}^2,\quad f\in \mathcal D.\end{aligned}$$
    (2.3)
  2. (b)

    There exists a positive constant C such that

    $$\begin{aligned} \int _0^1\vert tf(t)f^\prime (t)\vert \,\log \frac{2}{1-t}\,dt\,\le C\Vert f\Vert _{{\mathcal {D}}}^2,\quad f\in {\mathcal {D}}. \end{aligned}$$
    (2.4)

Regarding the Bergman space \(A^2\), Theorem 1.5 of [25] asserts the following.

Theorem C

Let \(\mu \) be a positive and finite Borel measure on [0, 1) and let \(h_\mu \) be defined by \(h_\mu (z)=\sum _{n=0}^\infty \mu _nz^n\)  (\(z\in {\mathbb {D}}.\)) If \(\mu \) satisfies the condition

$$\begin{aligned} \int _{[0, 1)} \frac{\mu \left( [t, 1)\right) }{(1-t)^2}\,d\mu (t)<\infty ,\end{aligned}$$
(2.5)

then \({\mathcal {H}}_\mu \) is bounded from \(A^2\) into itself if and only if the measure \(\vert h_\mu ^\prime (z)\vert ^2dA(z)\) is a Dirichlet-Carleson measure.

We recall that a finite positive Borel measure \(\nu \) on \({\mathbb {D}}\) is said to be a Dirichlet-Carleson messure if \({\mathcal {D}}\) is continuously embedded in \(L^2(d\nu )\). Stegenga [43] gave a characterization of these measures involving the logarithmic capacity of a finite union of intervals of \(\partial {\mathbb {D}}\). Shields [39] obtained a simpler characterization when dealing with measures supported on [0, 1). This result of Shields will be used below.

Using Theorem 1 we shall prove the following result.

Theorem 3

  1. (i)

    Let \(\mu \) be a positive and finite Borel measure on [0, 1). If \(\mu \) is a 1-logarithmic 1-Carleson measure, then \({\mathcal {H}}_\mu \) is bounded from \(A^2\) into itself.

  2. (ii)

    If \(0<\beta <1\), then there exists a positive and finite Borel measure \(\mu \) on [0, 1) which is a \(\beta \)-logarithmic 1-Carleson measure but such that \({\mathcal {H}}_\mu (A^2)\not \subset {\mathcal {A}}^2\).

In order to prove our results we start using the above mentioned result of Shields [39] to find a weak condition which insures that \({\mathcal {H}}_\mu \) and \({\mathcal {I}}_\mu \) are well defined in \({\mathcal {D}}\) and that \({\mathcal {H}}_\mu (f)=\mathcal I_\mu (f)\) for all \(f\in {\mathcal {D}}\).

Proposition 4

Let \(\mu \) be a positive and finite Borel measure on [0, 1). If there exists a positive constant C such that

$$\begin{aligned} \mu \left( [t, 1)\right) \le C\left( \log \frac{2}{1-t}\right) ^{-1},\quad 0<t<1, \end{aligned}$$
(2.6)

then \(\mathcal H_\mu \) and \({\mathcal {I}}_\mu \) are well defined in \({\mathcal {D}}\) and, furthermore, \({\mathcal {H}}_\mu (f)={\mathcal {I}}_\mu (f)\) for all \(f\in {\mathcal {D}}\).

Proof

Suppose that \(\mu \) satisfies (2.6). Shields proved in [39, Theorem 2] that this is equivalent to saying that there exists a positive constant A such that

$$\begin{aligned} \int _{[0, 1)}\vert f(t)\vert ^2 \,d\mu (t)\,\le \,A\Vert f\Vert _{{\mathcal {D}}}^2,\quad f\in \mathcal D. \end{aligned}$$
(2.7)

We can express (2.7) simply by saying that \(\mu \) is a radial Carleson-Dirichlet measure. Also, it is easy to see that (2.6) implies that there exists \(B>0\) such that

$$\begin{aligned} \mu _n\,\le \,\frac{B}{\log (n+2)},\quad n=0, 1, 2, \dots . \end{aligned}$$
(2.8)

Take \(f\in {\mathcal {D}}\), \(f(z)=\sum _{n=0}^\infty a_nz^n\) (\(z\in {\mathbb {D}}\)).

Let us prove that \({\mathcal {I}}_\mu (f)\) is well defined.

Using (2.7) and (2.8), we see that

$$\begin{aligned} \int _{[0, 1)}t^n\vert f(t)\vert \,d\mu (t)&\le \left( \int _{[0, 1)}t^{2n}\,d\mu (t)\right) ^{1/2}\left( \int _{[0, 1)}\vert f(t)\vert ^2\,d\mu (t)\right) ^{1/2}\\&\le A^{1/2}\mu _{2n}^{1/2}\Vert f\Vert _{\mathcal D}\\&\le \frac{A^{1/2}B^{1/2}\Vert f\Vert _{{\mathcal {D}}}}{\left( \log (2n+2)\right) ^{1/2}}, \end{aligned}$$

for all n. Then we have

$$\begin{aligned} \sum _{n=0}^\infty \left( \int _{[0, 1)}t^n\vert f(t)\vert \,d\mu (t)\right) \vert z\vert ^n\lesssim \sum _{n=0}^\infty \frac{\vert z\vert ^n}{\left( \log (2n+2)\right) ^{1/2}},\quad z\in {\mathbb {D}}. \end{aligned}$$

This implies that, for all \(z\in {\mathbb {D}}\), the integral

$$\begin{aligned} \int _{[0, 1)}\frac{f(t)}{1-tz}\,d\mu (t)\,=\,\int _{[0, 1)}f(t)\left( \sum _{n=0}^\infty t^nz^n\right) \,d\mu (t) \end{aligned}$$

converges and that

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

So \(\mathcal I_\mu (f) \) is a well defined analytic function in \({\mathbb {D}}\) and

$$\begin{aligned} {\mathcal {I}}_\mu (f)(z)=\sum _{n=0}^\infty \left( \int _{[0, 1)}t^nf(t)\,d\mu (t)\right) z^n,\quad z\in \mathbb D.\end{aligned}$$
(2.9)

Let us see now that \({\mathcal {H}}_\mu (f)\) is also well defined and that \({\mathcal {H}}_\mu (f)={\mathcal {I}}_\mu (f)\). Using (2.8), for all n, we have

$$\begin{aligned} \sum _{k=0}^\infty \vert \mu _{n+k}a_k\vert&\lesssim \mu _n\vert a_0\vert +\sum _{k=1}^\infty \frac{k^{1/2}\vert a_k\vert }{k^{1/2}\log (n+k+2)}\\ {}&\lesssim \mu _0\vert a_0\vert +\left( \sum _{k=1}^\infty k\vert a_k\vert ^2\right) ^{1/2}\left( \sum _{k=1}^\infty \frac{1}{k\left( \log (k+1)\right) ^2}\right) ^{1/2}\\&\lesssim \Vert f\Vert _{{\mathcal {D}}}. \end{aligned}$$

Clearly, this implies that \({\mathcal {H}}_\mu \) is a well defined analytic function in \({\mathbb {D}}\). Also,

$$\begin{aligned} \int _{[0, 1)}t^nf(t)\,d\mu (t)\,=\,\int _{[0, 1)}t^n\left( \sum _{k=0}^\infty a_kt^k\right) \,d\mu (t)\,=\,\sum _{k=0}^\infty \mu _{n+k}a_k \end{aligned}$$

for all k. Then (2.9) yields that \({\mathcal {H}}_\mu (f)={\mathcal {I}}_\mu (f)\). \(\square \)

Let us turn now to prove Theorem 1

Proof of Theorem 1 (i)

Suppose that \(\mu \) is a 1-logarithmic 1-Carleson measure. Take \(f\in {\mathcal {D}}\), \(f(z)=\sum _{k=0}^\infty a_kz^k\) (\(z\in {\mathbb {D}}\)). Proposition 4 implies that \({\mathcal {H}}_\mu (f)\) and \({\mathcal {I}}_\mu (f)\) are well defined and that \({\mathcal {H}}_\mu (f)={\mathcal {I}}_\mu (f)\). The above mentioned result of Shields yields that

$$\begin{aligned} \vert {\mathcal {H}}_{\mu }(f)(0)\vert \,=\,&\vert {\mathcal {I}}_{\mu }(f)(0)\vert \,=\,\left| \int _{[0, 1)}f(t)\,d\mu (t)\right| \nonumber \\ \,\lesssim \,&\left( \int _{[0, 1)}\vert f(t)\vert ^2\,d\mu (t)\right) ^{1/2} \,\lesssim \, \Vert f\Vert _{\mathcal D}. \end{aligned}$$
(2.10)

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

$$\begin{aligned} \mu _n={{\,\textrm{O}\,}}\left( \frac{1}{n\log (n+1)}\right) , \end{aligned}$$
(2.11)

(see e. g. [28, pp. 380-381]). Using (2.10) and (2.11), we obtain

$$\begin{aligned} \Vert {\mathcal {H}}_{\mu }(f)\Vert _{{\mathcal {D}}}^2\,\lesssim \,&\vert {\mathcal {H}}_{\mu }(f)(0)\vert ^2\,+\,\sum _{n=1}^\infty n\left( \sum _{k=0}^\infty \mu _{n+k}\vert a_k\vert \right) ^2 \\ \,\lesssim \,&\Vert f\Vert _{{\mathcal {D}}}^2\,+\,\sum _{n=1}^\infty n\left( \sum _{k=0}^\infty \frac{\vert a_k\vert }{(n+k)\log (n+k+1)}\right) ^2 \\ \,\lesssim \,&\Vert f\Vert _{{\mathcal {D}}}^2\,+\,I\,+\,II, \end{aligned}$$

where

$$\begin{aligned}I\,=\,&\sum _{n=1}^\infty n\left( \sum _{k=0}^n \frac{\vert a_k\vert }{(n+k)\log (n+k+1)}\right) ^2,\\II\,=\,&\sum _{n=1}^\infty n\left( \sum _{k=n+1}^\infty \frac{\vert a_k\vert }{(n+k)\log (n+k)}\right) ^2. \end{aligned}$$

Now, using a result of Holland and Walsh [30, Theorem 7] and simple estimates we deduce that

$$\begin{aligned} I\,=\,&\sum _{n=1}^\infty n\left( \sum _{k=0}^n \frac{\vert a_k\vert }{(n+k)\log (n+k+1)}\right) ^2\\ \,\le \,&\sum _{n=1}^\infty \frac{1}{n\left( \log (n+1)\right) ^2}\left( \sum _{k=0}^n\vert a_k\vert \right) ^2\,\lesssim \Vert f\Vert _{\mathcal D}^2. \end{aligned}$$

Also, since, for every n,

$$\begin{aligned} \sum _{k=n+1}^\infty \frac{\vert a_k\vert }{(n+k)\log (n+k)}\,\le \,&\frac{1}{\log (n+1)}\sum _{k=n+1}^\infty \frac{k^{1/2}\vert a_k\vert }{k^{1/2}(n+k)} \\ \,\le \,&\frac{1}{\log (n+1)}\left( \sum _{k=n+1}^\infty k\vert a_k\vert ^2\right) ^{1/2}\left( \sum _{k=n+1}^\infty \frac{1}{k(n+k)^2}\right) ^{1/2} \\ \,\le \,&\frac{\Vert f\Vert _{{\mathcal {D}}}}{\log (n+1)}\left( \sum _{k=n+1}^\infty \frac{1}{k(n+k)^2}\right) ^{1/2}\\ \,\le \,&\frac{\Vert f\Vert _{{\mathcal {D}}}}{n^{1/2}\log (n+1)}\left( \sum _{k=n+1}^\infty \frac{1}{(n+k)^2}\right) ^{1/2} \\ \,\lesssim \,&\frac{\Vert f\Vert _{{\mathcal {D}}}}{n\log (n+1)}, \end{aligned}$$

it follows that

$$\begin{aligned} II\,=\,&\sum _{n=1}^\infty n\left( \sum _{k=n+1}^\infty \frac{\vert a_k\vert }{(n+k)\log (n+k)}\right) ^2 \\ \,\lesssim \,&\Vert f\Vert _{{\mathcal {D}}}^2 \sum _{n=1}^\infty \frac{1}{n\left( \log (n+1)\right) ^2}\\\,\lesssim \,&\Vert f\Vert _{{\mathcal {D}}}^2. \end{aligned}$$

Putting everything together, we obtain \(\Vert {\mathcal {H}}_\mu (f)\Vert _{{\mathcal {D}}}^2\lesssim \Vert f\Vert _{{\mathcal {D}}}^2\). \(\square \)

Proof of Theorem 1 (ii)

Suppose that \(0<\beta <1\). Take \(\alpha \in {\mathbb {R}}\) with

$$\begin{aligned} \frac{1}{2}\,<\alpha \,<\min \left( 1, \frac{3-2\beta }{2}\right) . \end{aligned}$$

Let \(\mu \) be the Borel measure on [0, 1) defined by \(d\mu (t)=\left( \log \frac{2}{1-t}\right) ^{-\beta }\,dt\). Then (see [28, p. 392]) \(\mu \) is a \(\beta \)-logarithmic 1-Carleson measure and

$$\begin{aligned} \mu _n\,\asymp \,\frac{1}{n\left[ \log (n+1)\right] ^\beta }. \end{aligned}$$

Set \(a_n=\frac{1}{(n+1)\left[ \log (n+1)\right] ^\alpha }\) (\(n=1, 2, \dots \)) and \(g(z)=\sum _{n=1}^\infty a_nz^n\) (\(z\in {\mathbb {D}}\)).

The condition \(\alpha >\frac{1}{2}\) implies that \(g\in \mathcal D\). We are going to see that \({\mathcal {H}}_\mu (g)\notin \mathcal D\), this will finish the proof.

We have

$$\begin{aligned} \Vert {\mathcal {H}}_\mu (g)\Vert ^2_{{\mathcal {D}}}\,\gtrsim \,&\sum _{n=2}^\infty n\left( \sum _{k=2}^n\mu _{n+k}a_k\right) ^2\\ \,\asymp \,&\sum _{n=2}^\infty n\left( \sum _{k=2}^n\frac{1}{(n+k)\left[ \log (n+k)\right] ^\beta k\left[ \log k\right] ^\alpha }\right) ^2\\ \,\gtrsim \,&\sum _{n=2}^\infty \frac{n}{n^2\left[ \log n\right] ^{2\beta }}\left( \sum _{k=2}^n\frac{1}{k\left[ \log k\right] ^\alpha }\right) ^2 \\ \,=\,&\sum _{n=2}^\infty \frac{1}{n\left[ \log n\right] ^{2\beta }}\left( \sum _{k=2}^n\frac{1}{k\left[ \log k\right] ^\alpha }\right) ^2 \\ \,\gtrsim \,&\sum _{n=2}^\infty \frac{1}{n\left[ \log n\right] ^{2\beta +2\alpha -2}}. \end{aligned}$$

Since \(2\alpha +2\beta -2<1\), \(\sum _{n=2}^\infty \frac{1}{n\left[ \log n\right] ^{2\beta +2\alpha -2}}=\infty \) and, hence, \(\mathcal H_\mu (g)\notin {\mathcal {D}}\) as desired. \(\square \)

Proof of Corollary 2

The Dirichlet space is a Hilbert space with the inner product

$$\begin{aligned} <f, g>= f(0)\overline{g(0)}\,+\,\int _{{\mathbb {D}}}f^\prime (z)\overline{g^\prime (z)}\,dA(z),\quad f, g\in {\mathcal {D}}. \end{aligned}$$

Hence, \({\mathcal {D}}\) is identifiable with its dual with this pairing.

Assume that \(\mu \) is a finite Borel measure on [0, 1) which is a 1-logarithmic 1-Carleson measure. If \(f\in {\mathcal {D}}\), using Theorem 1, we see that \({\mathcal {H}}_\mu (f)\in {\mathcal {D}}\) and \(\Vert {\mathcal {H}}_\mu (f)\Vert _{{\mathcal {D}}}\lesssim \Vert f\Vert _{{\mathcal {D}}}\). Then \({\mathcal {H}}_\mu (f)\) induces a bounded linear functional on \({\mathcal {D}}\) with norm controlled by \(\Vert f\Vert _{\mathcal D}\). Thus

$$\begin{aligned} \left| \int _{\mathbb D}{\mathcal {H}}_\mu (f)^\prime (z)\overline{g^\prime (z)}\,dA(z)\right| \,\lesssim \Vert f\Vert _{{\mathcal {D}}}\Vert g\Vert _{{\mathcal {D}}},\quad f, g\in {\mathcal {D}}. \end{aligned}$$
(2.12)

Now, using the definitions, Fubini’s theorem, and the reproducing formula for the Bergman space \(A^2\), we have

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

Using (2.12), we obtain

$$\begin{aligned} \left| \int _{[0, 1)}tf(t)\overline{g^\prime (t)}\,d\mu (t)\right| \lesssim \Vert f\Vert _{{\mathcal {D}}}\Vert g\Vert _{{\mathcal {D}}},\quad f, g\in {\mathcal {D}}. \end{aligned}$$
(2.13)

Take \(f, g\in {\mathcal {D}}\), \(f(z)=\sum _{n=0}^\infty a_nz^n\), \(g(z)=\sum _{n=0}^\infty b_nz^n\) (\(z\in {\mathbb {D}}\)). Set

$$\begin{aligned} f_1(z)=\sum _{n=0}^\infty \vert a_n\vert z^n,\quad g_1(z)=\sum _{n=0}^\infty \vert b_n\vert z^n\quad (z\in {\mathbb {D}}). \end{aligned}$$

Then \(f_1, g_1\in {\mathcal {D}}\), \(\Vert f_1\Vert _{{\mathcal {D}}}=\Vert f\Vert _{{\mathcal {D}}}\), and \(\Vert g_1\Vert _{{\mathcal {D}}}=\Vert g\Vert _{{\mathcal {D}}}\). Using (2.13) with \(f_1\) and \(g_1\) in the places of f and g, we obtain

$$\begin{aligned} \int _{[0, 1)}\left| tf(t)\overline{g^\prime (t)}\right| \,d\mu (t)\,\le \,&\int _{[0, 1)}\left| tf_1(t)\overline{g_1^\prime (t)}\right| \,d\mu (t) \\ \,\lesssim&\Vert f_1\Vert _{{\mathcal {D}}}\Vert g_1\Vert _{{\mathcal {D}}}\\ \,=\,&\Vert f\Vert _{{\mathcal {D}}}\Vert g\Vert _{{\mathcal {D}}}. \end{aligned}$$

Taking \(f=g\), (2.3) follows.

Part (b) follows taking \(d\mu (t)=\log \frac{2}{1-t}\,dt\) in part (a). \(\square \)

Proof of Theorem 3

Our proof of Theorem 3 is based on the fact that the pairing

$$\begin{aligned} <f, g>\,=\,f(0)\overline{g(0)}\,+\,\int _{{\mathbb {D}}}f^\prime (z)\overline{\left( \frac{g(z)-g(0)}{z}\right) }\,dA(z),\quad f\in {\mathcal {D}},\,\,g\in A^2 \end{aligned}$$

is a “duality paring” between the Dirichlet space \({\mathcal {D}}\) and the Bergman space \(A^2\). Notice that if \(f(z)=\sum _{n=0}^\infty a_nz^n\) and \(g(z)=\sum _{n=0}^\infty b_nz^n\) (\(z\in {\mathbb {D}}\)), then

$$\begin{aligned} <f, g>\,=\,\sum _{n=0}^\infty a_n\overline{b_n}. \end{aligned}$$

It is a simple exercise to show that \(<{\mathcal {H}}_\mu (P), Q>=<P, {\mathcal {H}}_\mu (Q)>\) if P and Q are polynomials. Then it follows that if \({\mathcal {H}}_\mu \) is a bounded operator from \({\mathcal {D}}\) into itself then its adjoint (via this pairing) is \({\mathcal {H}}_\mu \), and then we see that \({\mathcal {H}}_\mu \) is a bounded operator from \(A^2\) into itself. Using this and Theorem 1 (i) we obtain part (a) of Theorem 3.

Similarly, if \({\mathcal {H}}_\mu \) is a bounded operator from \(A^2\) into itself, then \({\mathcal {H}}_\mu \) is also a bounded operator from \({\mathcal {D}}\) into itself and then part (b) of Theorem 3 follows using Theorem 1 (ii). \(\square \)

3 Cesàro-Type Operators

For \(\mu \) a finite positive Borel measure on [0, 1) as above, the matrix \({\mathcal {C}}_\mu \) induces a linear operator, also called \({\mathcal {C}}_\mu \), from \(\textrm{Hol}({\mathbb {D}})\) into itself as follows: If \(f\in \textrm{Hol}({\mathbb {D}})\), \(f(z)=\sum _{n=0}^\infty a_nz^n\) (\(z\in {\mathbb {D}}\)),

$$\begin{aligned} {\mathcal {C}}_\mu (f)(z)\,=\,\sum _{n=0}^\infty \left( \mu _n\sum _{k=0}^n a_k\right) z^n,\quad z\in {\mathbb {D}}. \end{aligned}$$

Let us remark that the operator \({\mathcal {C}}_\mu \) has the following integral representation: If \(f\in \textrm{Hol}({\mathbb {D}})\) then

$$\begin{aligned} {\mathcal {C}}_\mu (f)(z)=\int _{[0,1)}\frac{f(tz)}{1-tz}\,d\mu (t),\quad z\in {\mathbb {D}}. \end{aligned}$$
(3.1)

When \(\mu \) is the Lebesgue measure on [0, 1), the operator \({\mathcal {C}}_\mu \) reduces to the classical Cesàro operator \({\mathcal {C}}\).

The Cesàro operator \({\mathcal {C}}\) acting on distinct subspaces of \(\textrm{Hol}({\mathbb {D}})\) has been extensively studied in a good number of articles such as [2, 10, 12, 15, 23, 36, 40,41,42, 44]. Let us recall that it is bounded on \(H^p\) (\(0<p<\infty \)) and on \(A^p_\alpha \) (\(0<p<\infty \), \(\alpha >\,-1\)).

The operators \({\mathcal {C}}_\mu \) were introduced in [23] where, among other results, it was proved that the following conditions are equivalent:

  1. (i)

    \(\mu \) is a Carleson measure, that is, \(\mu ([t, 1))\le C(1-t)\) (\(0<t<1\)).

  2. (ii)

    \(\mu _n={{\,\textrm{O}\,}}\left( \frac{1}{n}\right) \).

  3. (iii)

    \(1\le p<\infty \) and \({\mathcal {C}}_\mu \) is bounded from \(H^p\) into itself.

  4. (iv)

    \(1<p<\infty \), \(\alpha >-1\), and \({\mathcal {C}}_\mu \) is bounded from \(A^p_\alpha \) into itself.

Blasco [12] has generalized the definition of the operators \({\mathcal {C}}_\mu \) by dealing with complex Borel measures on [0, 1) and he has extended results of [23] to this more general setting.

A further generalization has been given in [24] by working with the operators \({\mathcal {C}}_\mu \) associated to arbitrary complex Borel measures on \({\mathbb {D}}\), not necessarily supported on a radius. The complex Borel measures on \({\mathbb {D}}\) for which the operator \({\mathcal {C}}_\mu \) is bounded or Hilbert-Schmidt on \(H^2\) or on \(A^2_\alpha \) (\(\alpha >-1\)) are characterized in the mentioned paper [24].

We devote this section to study the operators \({\mathcal {C}}_\mu \) on the Dirichlet space, a question which has not been considered in the just mentioned papers. Our main results are contained in the following two theorems.

Theorem 5

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

  1. (i)

    If \(\mu \) is a 1-logarithmic 1-Carleson measure, then \({\mathcal {C}}_\mu \) is a bounded operator from the Dirichlet space \({\mathcal {D}}\) into itself.

  2. (ii)

    If \({\mathcal {C}}_\mu \) is a bounded operator from \({\mathcal {D}}\) into itself then \(\mu \) is a 1/2-logarithmic 1-Carleson measure.

Theorem 6

Suppose that \(\frac{1}{2}<\beta <1\). Then there exists a finite positive Borel measure \(\mu \) on [0, 1) which is \(\beta \)-logarithmic 1-Carleson measure for which \({\mathcal {C}}_\mu (\mathcal D)\not \subset {\mathcal {D}}\).

Proof of Theorem 5 (i)

Since \(\mu \) is a 1-logarithmic 1-Carleson measure, we have that

$$\begin{aligned} \mu _n={{\,\textrm{O}\,}}\left( \frac{1}{(n+1)\log (n+2)}\right) . \end{aligned}$$
(3.2)

Take \(f\in {\mathcal {D}}\), \(f(z)=\sum _{n=0}^\infty a_nz^n\) (\(z\in {\mathbb {D}}\)). Using (3.2) and Theorem 7 of [30], we obtain

$$\begin{aligned} \Vert {\mathcal {C}}_\mu (f)\Vert _{\mathcal D}^2\,\le \,&\sum _{n=0}^\infty (n+1)\mu _n^2\left( \sum _{k=0}^n\vert a_k\vert \right) ^2\\ \,\lesssim \,&\sum _{n=0}^\infty \frac{\left( \sum _{k=0}^n\vert a_k\vert \right) ^2}{(n+1)[\log (n+2)]^2} \\ \,\lesssim \,&\Vert f\Vert _{{\mathcal {D}}}^2. \end{aligned}$$

\(\square \)

Proof of Theorem 5 (ii)

Suppose that \(\mathcal C_\mu \) is a bounded operator from \({\mathcal {D}}\) into itself. For \(N\in {\mathbb {N}}\), set

$$\begin{aligned} f_N(z)\,=\,\sum _{n=1}^N\frac{z^n}{n},\quad z\in {\mathbb {D}}. \end{aligned}$$

Then,

$$\begin{aligned} \Vert f_N\Vert _{{\mathcal {D}}}^2\,=\,\sum _{n=1}^N\frac{1}{n}\,\asymp \,\log (N+1). \end{aligned}$$

Since \({\mathcal {C}}_\mu \) is bounded on \({\mathcal {D}}\), bearing in mind that the sequence of moments \(\{ \mu _n\} \) is decreasing, we have

$$\begin{aligned} \log (N+1)\,\asymp \,&\Vert f_N\Vert _{{\mathcal {D}}}^2\,\gtrsim \,\sum _{n=1}^\infty n\mu _n^2\left( \sum _{k=1}^n\frac{1}{k}\right) ^2 \\ \,\gtrsim \,&\mu _N^2\sum _{n=1}^Nn[\log (n+1)]^2 \,\asymp \,\mu _N^2N^2[\log (N+1)]^2. \end{aligned}$$

Then it follows that \(\mu _N={{\,\textrm{O}\,}}\left( \frac{1}{N[\log (N+1)]^{1/2}}\right) \). This implies that \(\mu \) is a 1/2-logarithmic 1-Carleson measure. \(\square \)

Proof of Theorem 6

Assume that \(1/2<\beta <1\). Let \(\mu \) be the Borel measure on [0, 1) defined by \(d\mu (t)=\left( \log \frac{2}{1-t}\right) ^{-\beta }\,dt\). Then, as mentioned before, \(\mu \) is a \(\beta \)-logarithmic 1-Carleson measure and \(\mu _n\asymp \frac{1}{n[\log (n+1)]^\beta }\).

Set \(\alpha =\beta -\frac{1}{2}\). Then \(0<\alpha <\frac{1}{2}\). Define

$$\begin{aligned} g(z)=\left( \log \frac{2}{1-z}\right) ^\alpha =\sum _{n=0}^\infty A_nz^n,\quad z\in {\mathbb {D}}. \end{aligned}$$

We have that

$$\begin{aligned} A_n\asymp \frac{1}{(n+1)[\log (n+2)]^{1-\alpha }}. \end{aligned}$$

Since \(\alpha <\frac{1}{2}\), we have that \(g\in {\mathcal {D}}\). Also

$$\begin{aligned} \Vert {\mathcal {C}}_\mu (g)\Vert _{{\mathcal {D}}}^2\,\ge \,&\sum _{n=2}^\infty n\mu _n^2\left( \sum _{k=2}^nA_k\right) ^2\,\gtrsim \,\sum _{n=2}^\infty \frac{n}{n^2[\log n]^{2\beta }[\log n]^{-2\alpha }}\\ \,=&\sum _{n=2}^\infty \frac{1}{n[\log n]^{2(\beta -\alpha )}} \,=\,\sum _{n=2}^\infty \frac{1}{n[\log n]}\,=\,\infty . \end{aligned}$$

\(\square \)

Danikas and Siskakis [15] proved that \({\mathcal {C}}(H^\infty )\not \subset H^\infty \) and that \(\mathcal C(H^\infty )\subset BMOA\). This was improved by Essén and Xiao who proved in [22] that \({\mathcal {C}}(H^\infty )\subset Q^p\) for \(0<p<\infty \). This result has been sharpened in [10].

We recall that BMOA is the space of those functions \(f\in H^1\) whose boundary values have bounded mean oscillation. Alternatively, a function \(f\in \textrm{Hol}({\mathbb {D}})\) belongs to BMOA if and only if

$$\begin{aligned} \sup _{T\in \textrm{Aut}({\mathbb {D}})}\Vert f\circ T\,-\,f(T(0))\Vert _{H^2}<\infty , \end{aligned}$$

where \(\textrm{Aut}({\mathbb {D}})\) denotes the set of all Möbius transformations from \({\mathbb {D}}\) onto itself. We refer to [26] for the theory of BMOA-functions.

For \(0<s<\infty \) the space \(Q_s\) consists of those \(f\in \textrm{Hol}({\mathbb {D}})\) such that

$$\begin{aligned} \sup _{T\in \textrm{Aut}({\mathbb {D}})}\int _{{\mathbb {D}}}\vert f^\prime (z)\vert ^2(1-\vert T(z)\vert ^2)^s\,dA(z)<\infty . \end{aligned}$$

The spaces \(Q_s\) were introduced in [6] and [7]. We refer to [46] for the theory of \(Q_s\) spaces. Let us recall that

$$\begin{aligned} \mathcal D\subsetneq Q_{s_1}\subsetneq Q_{s_2}\subsetneq Q_1=BMOA,\quad 0<s_1<s_2<1. \end{aligned}$$

For \(s>1\) the space \(Q_s\) coincides with the Bloch space \({\mathcal {B}}\) of those functions \(f\in \textrm{Hol}({\mathbb {D}})\) for which

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

The paper [3] is an excellent reference for the theory of Bloch functions. Let us recall that \(BMOA\subsetneq {\mathcal {B}}\).

Blasco [12] has proved that

$$\begin{aligned} {\mathcal {C}}(H^\infty )\subset \bigcap _{1<p<\infty }\Lambda ^p_{1/p}. \end{aligned}$$
(3.3)

Here, for \(p\ge 1\), \(\Lambda ^p_{1/p}\) is the space of those functions \(f\in \textrm{Hol}({\mathbb {D}})\) having a non-tangential limit at almost every point of \(\partial {\mathbb {D}}\) and so that \(\omega _ p(\cdot , f)\), the integral modulus of continuity of order p of the boundary values \(f(e^{i\theta })\) of f, satisfies \(\omega _ p(\delta , f)={{\,\textrm{O}\,}}(\delta ^{1/p})\), as \(\delta \rightarrow 0\). Classical results of Hardy and Littlewood (see [13] and [20, Chapter 5]) show that \(\Lambda ^p_{1/p}\subset H^p\) and that

$$\begin{aligned} \Lambda ^p_{1/p}=\left\{ f \hbox { analytic in } {\mathbb {D}}: M_ p(r, f^\prime )= {{\,\textrm{O}\,}}\left( \frac{1}{(1-r)^{1-\frac{1}{p}}}\right) , \quad \hbox { as } r\rightarrow 1 \right\} . \end{aligned}$$

In particular, \(\Lambda ^1_1\) is the space of those \(f\in \textrm{Hol}({\mathbb {D}})\) such that \(f^\prime \in H^1\). The spaces \(\Lambda ^p_{1/p}\) increase with p and they are all contained in BMOA [13]. Since \(\Lambda ^2_{1/2}\subset Q_s\) for all \(s>0\) (see [5, p. 427]), (3.3) improves the mentioned result in [22].

Bao, Sun and Wulan [8, Theorem 3.1] have proved that for any given \(s>0\), \({\mathcal {C}}_\mu (H^\infty )\subset Q_s\) if and only if \(\mu \) is a Carleson measure.

It is natural to look for a result like (3.3) with \({\mathcal {D}}\) in the place of \(H^\infty \). It is easy to see that

$$\begin{aligned} {\mathcal {C}}({\mathcal {D}})\not \subset {\mathcal {B}}. \end{aligned}$$
(3.4)

Indeed, set \(a_n=\frac{1}{(n+1)\log (n+1)}\) (\(n\ge 1\)) and \(f(z)=\sum _{n=1}^\infty a_nz^n\) (\(z\in \mathbb D\)). Then \(f\in {\mathcal {D}}\) and, setting \(A_n=\sum _{k=1}^na_k\), we have, for \(0<r<1\),

$$\begin{aligned} (1-r){\mathcal {C}}(f)^\prime (r)\,=\,&(1-r)\sum _{n=1}^\infty \frac{n}{n+1}A_nr^{n-1}\,\ge \,\frac{1}{2}(1-r)\sum _{n=1}^\infty A_nr^{n-1}\\ \,=\,&\frac{1}{2}\left[ A_1\,+\,\sum _{n=2}^\infty (A_n\,-\,A_{n-1})r^{n-1}\right] \,=\,\frac{1}{2}\left[ A_1\,+\,\sum _{n=2}^\infty a_nr^{n-1}\right] \\ \,\asymp \,&\log \log \frac{2}{1-r}. \end{aligned}$$

Hence, \(C(f)\not \in {\mathcal {B}}\).

The next natural step is trying to characterize the measures \(\mu \) such that \({\mathcal {C}}_\mu ({\mathcal {D}})\subset \mathcal B\). We have the following result.

Theorem 7

Let X be a Banach space of analytic functions in \({\mathbb {D}}\) with \(\Lambda ^2_{1/2}\subset X\subset {\mathcal {B}}\) and let \(\mu \) be a positive finite Borel measure on [0, 1).

  1. (i)

    If \(\mu \) is a \(\frac{1}{2}\)-logarithmic 1-Carleson measure, then \({\mathcal {C}}_\mu \) is a bounded operator from \({\mathcal {D}}\) into X.

  2. (ii)

    If \(\mathcal C_\mu \) is a bounded operator from \({\mathcal {D}}\) into X and \(0<\beta <\frac{1}{2}\), then \(\mu \) is a \(\beta \)-logarithmic 1-Carleson measure.

Proof

Suppose that \(\mu \) is a \(\frac{1}{2}\)-logarithmic 1-Carleson measure. Then

$$\begin{aligned} \mu _n\lesssim \frac{1}{n[\log (n+1)]^{1/2}}. \end{aligned}$$
(3.5)

Take \(f\in {\mathcal {D}}\), \(f(z)=\sum _{n=0}^\infty a_nz^n\) (\(z\in {\mathbb {D}}\)). We have

$$\begin{aligned} {\mathcal {C}}_\mu (f)(z)=\sum _{n=0}^\infty \mu _n\left( \sum _{k=0}^na_k\right) z^n\,=\,\sum _{n=0}^\infty A_nz^n, \end{aligned}$$

where \(A_n=\mu _n\left( \sum _{k=0}^na_k\right) \). We have,

$$\begin{aligned} \left| \sum _{k=0}^na_k\right|&\le \vert a_0\vert \,+\,\sum _{k=1}^n\frac{k^{1/2}\vert a_k\vert }{k^{1/2}}\\&\le \vert a_0\vert \,+\,\left( \sum _{k=1}^nk\vert a_k\vert ^2\right) ^{1/2}\left( \sum _{k=1}^n\frac{1}{k}\right) ^{1/2}\,\lesssim \Vert f\Vert _{{\mathcal {D}}}[\log (n+1)]^{1/2}. \end{aligned}$$

This and (3.5) imply that \(\vert A_n\vert \lesssim \frac{\Vert f\Vert _{{\mathcal {D}}}}{n}\) a fact which easily yields that \({\mathcal {C}}_\mu (f)\in \Lambda ^2_{1/2}\). This finishes the proof of (i).

Let us turn to prove (ii). Assume that \(0<\beta <\frac{1}{2}\) and that \({\mathcal {C}}_\mu \) is a bounded operator from \({\mathcal {D}}\) into X.

Since \(X\subset {\mathcal {B}}\), \({\mathcal {C}}_\mu \) is a bounded operator from \({\mathcal {D}}\) into \({\mathcal {B}}\).

Set \(\alpha =1-\beta \), and \(f(z)=\sum _{n=0}^\infty \frac{z^n}{(n+1)[\log (n+2)]^\alpha }\) (\(z\in {\mathbb {D}}\)).

Notice that \(\frac{1}{2}<\alpha <1\). This implies that \(f\in \mathcal D\) and, hence, \({\mathcal {C}}_\mu (f)\in {\mathcal {B}}\). Then, bearing in mind that the sequence \(\{ \mu _n\} \) is decreasing, we see that, for \(0<r<1\) and \(N\in {\mathbb {N}}\),

$$\begin{aligned} \frac{1}{1-r}&\gtrsim \,\sum _{n=1}^\infty n\mu _n\left( \sum _{k=1}^n\frac{1}{(k+1)[\log (k+2)]^\alpha }\right) r^{n-1} \\&\ge \sum _{n=1}^N n \mu _n\left( \sum _{k=1}^n\frac{1}{(k+1)[\log (k+2)]^\alpha }\right) r^n \\&\gtrsim \,\mu _N\sum _{n=1}^Nn[\log (n+2)]^{1-\alpha }r^n. \end{aligned}$$

Taking \(r=1-\frac{1}{N}\), we obtain

$$\begin{aligned} N\,\gtrsim \,\mu _NN^2[\log (N+2)]^{1-\alpha }\,=\,\mu _NN^2[\log (N+2)]^{\beta } \end{aligned}$$

and, hence, \(\mu _N\lesssim \frac{1}{N[\log (N+2)]^\beta }\). This implies that \(\mu \) is a \(\beta \)-logarithmic 1-Carleson measure. \(\square \)

4 Extensions to Besov Spaces

The Dirichlet space is one among the analytic Besov spaces \(B^p\). For \(1<p<\infty \), the analytic Besov space \(B^p\) is the space \({\mathcal {D}}^p_{p-2}\). Thus \(B^2\,=\,{\mathcal {D}}\).

The minimal Besov space \(B^1\) requires a special definition. It is the space of all \(f\in \textrm{Hol}({\mathbb {D}})\) such that \(f^{\prime \prime }\in A^1\). It is a Banach space with the norm \(\Vert \cdot \Vert _{B^1}\) defined by \(\Vert f\Vert _{B^1}\,=\,\vert f(0)\vert \,+\,\vert f^\prime (0)\vert \,+\,\Vert f^{\prime \prime }\Vert _{A^1}\).

The Besov spaces \(B^p\) form a nested scale of conformally invariant spaces and they are all contained in BMOA:

$$\begin{aligned} B^p\,\subsetneq B^q\,\subsetneq BMOA\subsetneq {\mathcal {B}},\quad 1\le p<q<\infty . \end{aligned}$$

Also \(B^p\,\subsetneq \,\Lambda ^p_{1/p}\) for all \(p\in [1,\infty )\). We mention [4, 11, 18, 30, 47, 48] for information on Besov spaces. Let us remark that, letting \(d\lambda \) be the Möbius invariant measure on \({\mathbb {D}}\) defined by \(d\lambda (z)=\frac{dA(z)}{(1-\vert z\vert ^2)^2} \), we have:

  1. (a)

    The Bergman projection P is a continuous linear operator from \(L^\infty ({\mathbb {D}})\) onto the Bloch space \({\mathcal {B}}\),

  2. (b)

    For \(1<p<\infty \), the Bergman projection P is a continuous linear operator from \(L^p(d\lambda )\) onto \(B^p\)

(see [48, Chapter 5]).

Our aim in this section is trying to extend to the spaces \(B^p\) some of the results obtained in the preceding ones for the Dirichlet space.

For the space \(B^1\) we have the following result.

Theorem 8

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

  1. (i)

    \(\int _{[0, 1)}\frac{d\mu (t)}{1-t}\,<\,\infty \).

  2. (ii)

    \(\sum _{n=0}^\infty \mu _n\,<\,\infty \).

  3. (iii)

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

  4. (iv)

    The operator \({\mathcal {C}}_\mu \) is a bounded operator from \(B^1\) into itself.

Proof

The equivalence (i)  \(\Leftrightarrow \)  (ii) is clear.

Suppose that (iii) holds. Let f be the constant function \(f(z)=1\), for all \(z\in {\mathbb {D}}\). Then \({\mathcal {H}}_\mu (f)={\mathcal {I}}_\mu (f)\in B^1\subset H^\infty \) and then

$$\begin{aligned} \int _{[0, 1)}\frac{d\mu (t)}{1-t}\,=\,\lim _{r\rightarrow 1^-}{\mathcal {I}}_\mu (f)(r)\,\le \,\Vert {\mathcal {I}}_\mu (f)\Vert _{H^\infty }\,<\infty . \end{aligned}$$

Thus (i) holds.

Conversely, suppose that (i) holds. Take \(f\in B^1\). We have

$$\begin{aligned} {\mathcal {H}}_\mu (f)^{\prime \prime }(z)\,=\,\int _{[0, 1)}\frac{2t^2f(t)}{(1-tz)^3}\,d\mu (t),\quad z\in {\mathbb {D}}. \end{aligned}$$

Then using Fubini’s theorem, [48, Lemma 3.10], and the fact that \(B^1\subset H^\infty \), we obtain

$$\begin{aligned} \int _{{\mathbb {D}}}\left| {\mathcal {H}}_\mu (f)^{\prime \prime }(z)\right| \,dA(z)\,&\lesssim \,\int _{{\mathbb {D}}}\int _{[0, 1)}\frac{\vert f(t)\vert }{\vert 1-tz\vert ^3}\,d\mu (t)\,dA(z)\\&= \int _{[0, 1)}\vert f(t)\vert \int _{{\mathbb {D}}}\frac{dA(z)}{\vert 1-tz\vert ^3} \,d\mu (t)\\&\lesssim \, \Vert f\Vert _{H^\infty }\int _{[0, 1)}\frac{d\mu (t)}{1-t}\\&\lesssim \,\Vert f\Vert _{B^1}\int _{[0, 1)}\frac{d\mu (t)}{1-t}. \end{aligned}$$

Thus, (iii) follows.

Let us prove next the equivalence (i)  \(\Leftrightarrow \)  (iv).

Suppose (i). Take \(f\in B^1\). Bearing in mind (3.1) and using Fubini’s theorem, we see that

$$\begin{aligned}&\int _{{\mathbb {D}}}\left| {\mathcal {C}}_\mu (f)^{\prime \prime }(z)\right| dA(z)\\&\quad \lesssim \int _{[0, 1)}\int _{{\mathbb {D}}}\frac{\vert f^{\prime \prime }(tz)\vert dA(z)}{\vert 1-tz\vert }d\mu (t)\,+\, \int _{[0, 1)}\int _{{\mathbb {D}}}\frac{\vert f^{\prime }(tz)\vert dA(z) }{\vert 1-tz\vert ^2}d\mu (t)\\&\qquad +\,\int _{[0,1)}\int _{\mathbb D}\frac{\vert f(tz)\vert dA(z)}{\vert 1-tz\vert ^3}d\mu (t). \end{aligned}$$

We now estimate each of the three terms in the last formula separately. For the first one we have

$$\begin{aligned} \int _{[0,1)} \int _{{\mathbb {D}}} \frac{\vert f^{\prime \prime }(tz)\vert }{\vert 1-tz\vert }\,dA(z)\,d\mu (t)&\le \int _{[0,1)}\frac{1}{1-t} \int _{{\mathbb {D}}} \vert f^{\prime \prime }(tz)\vert \,dA(z)\,d\mu (t) \\&\lesssim \Vert f\Vert _{B^1} \int _{[0,1)}\frac{d\mu (t)}{1-t}. \end{aligned}$$

For the second one, we use the fact that \(B^1\subset \Lambda ^1_1\) to obtain

$$\begin{aligned} \int _{[0,1)} \int _{{\mathbb {D}}} \frac{\vert f^{\prime }(tz)\vert }{\vert 1-tz\vert ^2} \,dA(z)\,d\mu (t)&\lesssim \int _{[0,1)} \int _0^1 \frac{M_1(tr,f^\prime )}{(1-tr)^2} \,dr\,d\mu (t) \\&\le \Vert f\Vert _{\Lambda ^1_1} \int _{[0,1)} \frac{d\mu (t)}{1-t} \lesssim \Vert f\Vert _{B^1} \int _{[0,1)}\frac{d\mu (t)}{1-t}. \end{aligned}$$

For the last integral, we use that \(B^1\subset H^\infty \) and Lemma 3.10 of [48] to see that

$$\begin{aligned} \int _{[0,1)} \int _{{\mathbb {D}}} \frac{\vert f(tz)\vert }{\vert 1-tz\vert ^3} \,dA(z)\,d\mu (t) \le \,&\Vert f\Vert _{H^\infty } \int _{[0,1)} \int _{{\mathbb {D}}} \frac{dA(z)}{\vert 1-tz\vert ^3} \,d\mu (t)\\ \, \lesssim \,&\Vert f\Vert _{B^1} \int _{[0,1)}\frac{d\mu (t)}{1-t}. \end{aligned}$$

Putting everything together we obtain (iv).

Suppose now that (iv) holds. Let f be the constant function given by \(f(z)=1\), for all \(z\in {\mathbb {D}}\). Then \({\mathcal {C}}_\mu (f)\in B^1\subset H^\infty \). Using the integral representation of \({\mathcal {C}}_\mu \) we see that

$$\begin{aligned} \int _{[0, 1)}\frac{d\mu (t)}{1-t}\,=\,\lim _{r\rightarrow 1^-}{\mathcal {C}}_\mu (f)(r)\,\le \,\Vert {\mathcal {C}}_\mu (f)\Vert _{H^\infty }. \end{aligned}$$

Thus, \(\int _{[0, 1)}\frac{d\mu (t)}{1-t}<\infty \). This is (i). \(\square \)

Let us turn now to deal with the possible extensions in the range \(1<p<\infty \). The following result comes from [28, Theorem 2.4] and [23, Theorem 7].

Theorem D

Let \(\mu \) be a positive finite Borel measure on [0, 1). If \(\mu \) is a 1-logarithmic 1-Carleson measure then the operators \({\mathcal {H}}_\mu \) and \({\mathcal {C}}_\mu \) are bounded from the Bloch space \({\mathcal {B}}\) into itself.

Using this result and those obtained in Sects. 2 and  3 we will prove the following.

Theorem 9

Suppose that \(2<p<\infty \) and let \(\mu \) be a positive finite Borel measure on [0, 1). If \(\mu \) is a 1-logarithmic 1-Carleson measure then the operators \({\mathcal {H}}_\mu \) and \({\mathcal {C}}_\mu \) are bounded from the Besov space \(B^p\) into itself.

Proof

We shall use complex interpolation in the proof. Let us refer to [48, Chapter 2] for the terminology and basic results concerning complex interpolation.

If \(X_0\) and \(X_1\) are two compatible Banach spaces then, for \(0<\theta <1\), \([X_0, X_1]_{\theta }\) stands for the space obtained by the complex method of interpolation of Calderón. As a consequence of the above mentioned results characterizing the spaces \(B^p\) as the image of \(L^p(d\lambda )\) under the Bergman projection and the Bloch space as the image of \(L^\infty (d\lambda )\) under the Bergman projection, Zhu proves in [48, Theorem 5.25] that if \(1<p_0<\infty \), \(0<\theta <1\), and \(1/p=(1-\theta )/p_0\), then

$$\begin{aligned}{}[B^{p_0}, {\mathcal {B}}]_{\theta }=B^p. \end{aligned}$$
(4.1)

In particular,

$$\begin{aligned} B^p=[{\mathcal {D}}, {\mathcal {B}}]_\theta , \quad \text {if } 2<p<\infty \text { and } \theta =1-\frac{2}{p}. \end{aligned}$$
(4.2)

Theorem 9 follows using (4.2), Theorem 1 (i), Theorem 5 (i), and the interpolation theorem of operators [48, Theorem 2.4]. \(\square \)

Regarding the sharpness of Theorem 9, we have the following result.

Theorem 10

Suppose that \(0<\beta <1\).

  1. (i)

    If \(1<p<\infty \) then there exists a positive Borel measure \(\mu \) on [0, 1) which is a \(\beta \)-logarithmic 1-Carleson measure with the property that \({\mathcal {H}}_\mu (B^p)\not \subset B^p\).

  2. (ii)

    If \(1<p\le 2\) then there exists a positive Borel measure \(\mu \) on [0, 1) which is a \(\beta \)-logarithmic 1-Carleson measure with the property that \({\mathcal {C}}_\mu (B^p)\not \subset B^p\).

Proof

Assume that \(1<p<\infty \) and \(0<\beta <1\). Take \(\alpha \in {\mathbb {R}}\) with

$$\begin{aligned} \frac{1}{p}\,<\alpha \,<\,\min \left( 1, 1+\frac{1}{p}-\beta \right) . \end{aligned}$$

Let \(\mu \) be the Borel measure on [0, 1) defined by \(d\mu (t)=\left( \log \frac{2}{1-t}\right) ^{-\beta } \,dt\). We know that \(\mu \) is a \(\beta \)-logarithmic 1-Carleson measure and that \(\mu _n\asymp \frac{1}{(n+1)[\log (n+2)]^\beta }.\)

For \(n\ge 1\), set \(a_n=\frac{1}{n[\log (n+1)]^\alpha }\) and \(g(z)=\sum _{n=1}^\infty a_nz^n\) (\(z\in {\mathbb {D}}\)).

Since the sequence \(\{ a_n\} \) is decreasing and \(\sum _{n=1}^\infty n^{p-1}\vert a_n\vert ^p<\infty \), using [28, Theorem 3.10] we see that \(g\in B^p\).

We have that \({\mathcal {H}}_\mu (g)(z)=\sum _{n=0}^\infty \left( \sum _{k=0}^\infty \mu _{n+k}a_k\right) z^n\) (\(z\in {\mathbb {D}}\)). Since the \(a_k\)’s are positive and the sequence of moments \(\{ \mu _n\} \) is decreasing, it follows that the sequence \(\left\{ \sum _{k=0}^\infty \mu _{n+k}a_k\right\} \) is also decreasing. Then using again [28, Theorem 3.10] we see that

$$\begin{aligned} H_\mu (g)\in B^p\,\,\,\,\Leftrightarrow \,\,\,\,\,\sum _{n=1}^\infty n^{p-1}\left( \sum _{k=0}^\infty \mu _{n+k}a_k\right) ^p<\infty . \end{aligned}$$
(4.3)

Now,

$$\begin{aligned}\sum _{n=1}^\infty n^{p-1}\left( \sum _{k=0}^\infty \mu _{n+k}a_k\right) ^p\,\gtrsim \,&\sum _{n=2}^\infty n^{p-1}\left( \sum _{k=2}^\infty \frac{1}{(n+k)[\log (n+k)]^\beta k(\log k)^\alpha }\right) ^p\\ \,\ge \,&\sum _{n=2}^\infty n^{p-1}\left( \sum _{k=2}^n \frac{1}{(n+k)[\log (n+k)]^\beta k(\log k)^\alpha }\right) ^p \\ \,\gtrsim \,&\sum _{n=2}^\infty \frac{n^{p-1}}{n^p(\log n)^{p\beta }}\left( \sum _{k=2}^n\frac{1}{k(\log k)^\alpha }\right) ^p \\ \,\asymp \,&\sum _{n=2}^\infty \frac{1}{n(\log n)^{p\beta }(\log n)^{p(\alpha -1)}} \\ \,= \,&\sum _{n=2}^\infty \frac{1}{n(\log n)^{p(\beta +\alpha -1)}}. \end{aligned}$$

Since \(p(\beta +\alpha -1)<1\), it follows that \(\sum _{n=1}^\infty n^{p-1}\left( \sum _{k=0}^\infty \mu _{n+k}a_k\right) ^p=\infty \) and then (4.3) gives that \(H_\mu (g)\not \in B^p\).

Assume now that \(1<p\le 2\). We have

$$\begin{aligned} {\mathcal {C}}_\mu (g)(z)\,=\,\sum _{n=0}^\infty \mu _n\left( \sum _{k=0}^na_k\right) z^n. \end{aligned}$$

Using the fact that \(1<p\le 2\) and [20, Theorem 6.s2] it readily follows that

$$\begin{aligned} {\mathcal {C}}_\mu (g)\in B^p\,\,\,\,\Rightarrow \,\,\,\,\,\sum _{n=1}^\infty n^{p-1}\mu _n^p\left( \sum _{k=1}^na_k\right) ^p<\infty .\end{aligned}$$
(4.4)

But,

$$\begin{aligned}\sum _{n=1}^\infty n^{p-1}\mu _n^p\left( \sum _{k=1}^na_k\right) ^p\,\gtrsim \,&\sum _{n=1}^\infty \frac{1}{n[\log (n+1)]^{\beta p}}\left( \sum _{k=2}^n\frac{1}{k(\log k)^\alpha }\right) ^p \\ \,\gtrsim \,&\sum _{n=1}^\infty \frac{1}{n[\log (n+1)]^{p(\beta +\alpha -1)}}\\ \,=\,&\infty . \end{aligned}$$

Using (4.4) we obtain that \({\mathcal {C}}_\mu (g)\not \in B^p\). \(\square \)